Uploaded by k.habanjar

1-s2.0-S0011227521001119-main-2

advertisement
Cryogenics 119 (2021) 103353
Contents lists available at ScienceDirect
Cryogenics
journal homepage: www.elsevier.com/locate/cryogenics
Structural and superconducting properties of low-density Bi(Pb)-2223
superconductor: Effect of Eu2O3 nanoparticles addition
E.S. Nurbaisyatul a, H. Azhan b, *, N. Ibrahim a, S.F. Saipuddin a, b
a
b
Faculty of Applied Sciences, Universiti Teknologi MARA Shah Alam, 40450 Shah Alam, Selangor, Malaysia
Faculty of Applied Sciences, Universiti Teknologi MARA Pahang, 26400 Bandar Tun Abdul Razak Jengka, Pahang, Malaysia
A R T I C L E I N F O
A B S T R A C T
Keywords:
Bi(Pb)-2223
Rare earth nanoparticles
Porous
Addition
Grain size
This paper presented the result of the structural analysis of Europium, Eu nanoparticles addition in low-density Bi
(Pb)-2223 superconductor. Polycrystalline bulks with nominal composition Bi1.6Pb0.4Sr2Ca2Cu3Oy + xEu2O3
samples where x = 0.0, 0.2, 0.4, 0.6 and 0.8 wt% were prepared via solid-state reaction method. Low-density
samples were created by adding crystalline sucrose into mixed powders and burn at 400 ◦ C for two hours.
The samples were characterized using X-Ray Diffraction (XRD) and Scanning Electron Microscopy (SEM) for
structural characterization while resistivity measurement was determined by using four-point probe method.
Phase analysis by XRD indicated that both Bi-2212 and Bi-2223 phases coexist within the samples having a
tetragonal crystal structure. As Eu concentration increases, the amount of Bi-2223 phase slightly decreased which
signifies that Eu nanoparticles favour the growth of Bi-2212 phase. It was observed that the highest Tc and Jc
values for the addition sample found at x = 0.2 wt%.
1. Introduction
Nowadays, researchers have conducted numerous studies to discover
the best materials that can maintain superconductivity at room tem­
perature [1,2]. An extensive search for novel superconductors has been
performed to find the best materials with higher critical temperature, Tc.
Superconducting materials can be classified into three groups which are
metal-based system, copper oxides (cuprates) and iron-based super­
conductors [3]. Among those categories, cuprates exhibit admirable
characteristics due to their strong magnetic field performance, zero
energy losses and current-carrying capacity [4]. BSCCO system is one of
the promising materials that attracted researchers’ attention due to their
great capability to carry electrical current without any resistance. This
system mainly consists of three phases with general formula
Bi2Sr2CanCunO2n+4+y where n = 1, 2 and 3 considering the number of
CuO2 layers in the sub-unit cell respectively. Notably, Bi-2223 phases
more preferable due to the highest critical temperature, Tc (~110 K)
compared to Bi-2201 (~20 K) and Bi-2212 (~90 K) [5]. Different
method of preparation, structural and superconducting properties has
been reported through the experience of past research [6–11]. There are
still lacking in terms of this behaviour although numerous attempts have
been made to further understand the phenomenon of a superconductor.
Besides BSCCO system, there are other classes of oxide hightemperature superconductor materials such as Y-Ba-Cu-O and Tl-based
family (Tl-Ba-Ca-Cu-O) that are also promising for practical uses. The
YBCO family has unique characteristic that makes them promising for
electronic and magnetic applications, such as particle accelerators,
electronic motors and power transmission and magnetic levitator de­
vices. Meanwhile, Tl-based family consist of different phases which are
Tl-2212, Tl-1223 and Tl-2223 with critical temperature of more than
120 K but have inability in carrying current in wires and tapes due to
incompetence in controlling the crystallographic orientation, grain
boundaries and microcracking [12].
One of the possible ways to increase Jc value is through a low-density
high-temperature superconductor. Low-density samples are also known
as porous samples which have an intermediate medium. According to
previous research, sucrose has been used as an additive to produce
porous surfaces, as it cannot react with the alumina matrix while
burning [13]. Study by Wu, I.J. in 2001 discovered that the Jc values of
the porous YBCO films were higher by at least fifty percent rather than
the free-pores YBCO films even for thick films [14]. One of the bene­
fits of the porous structure is that it enables effective contact to the
sample by applying a silver-conductive paint which then immerses into
the body of the sample, making a deep contact instead of applying just
* Corresponding author.
E-mail address: dazhan@uitm.edu.my (H. Azhan).
https://doi.org/10.1016/j.cryogenics.2021.103353
Received 21 January 2021; Received in revised form 24 August 2021; Accepted 29 August 2021
Available online 2 September 2021
0011-2275/© 2021 Elsevier Ltd. All rights reserved.
E.S. Nurbaisyatul et al.
Cryogenics 119 (2021) 103353
on the surface.
Over the past few years, much effort has been taken to prepare the
high Tc Bi-2223 phase superconductors. Several synthesizing techniques
such as the conventional solid state method [15–18], sol–gel method
[19,20] and co-precipitation (COP) method [21,22] have been devel­
oped so that superconductors with desired properties can be obtained.
Unfortunately, a major problem rises from previous study which turned
out to be intergrain weak links leading to a weak critical current density,
Jc. Note that the amount of dopant and additive elements need to be
controlled to improve the Jc values. However, the type of additive used
must be taken into consideration as it either acts as an impurity or
destruct the characteristics of the crystal structure [23]. Normally, the
preparation of BSCCO system always leads to the coexistence of Bi-2212
and Bi-2223 phases despite many attempts to obtain Bi-2223 single
phase.
Subsequently, it is of interest to introduce nano-sized rare earth el­
ements into the porous structure of Bi(Pb)-2223 to increase the con­
nectivity between the grains. In recent years, there is an increasing
interest among researchers to use nanomaterials to improve the super­
conductivity characteristic of the BSCCO system. Numerous studies have
investigated the effect of substitution or addition of nanomaterials into
high-temperature superconductors as demonstrated in [17] and
[24–29]. The latest study by Loudhaief et al. investigated the impact of
Cadmium Sulfide (CdS) nanoparticles on the superconducting properties
and flux pinning ability in Bi(Pb)-2223 superconductor [30]. Based on
the result obtained, they found that small addition of Cadmium Sulfide
(CdS) nanoparticles (≤0.3 wt%) exhibit the higher critical current
densities, Jc and produced strong pinning centres in the Bi(Pb)-2223
system compared to CdS-free sample. The effect of TiO2 nanoparticles
addition into YBCO superconductor has been recently studied by Han­
nachi et al. which revealed that adding 0.1 wt% of TiO2 nanoparticles
allowed a significant improvement in the connectivity between YBCO
grains [23]. Hence, this paper aims to study the influence of the nano­
sized Eu2O3 addition in porous BSCCO compound on the structural and
superconducting properties synthesized by solid-state reaction method.
In this research, we added Eu2O3 nanoparticles to the BSCCO system to
improve the connectivity between the grains owing to the metallic
behaviour of Europium. Different parameters such as lattice parameters,
hole concentrations, critical temperature (Tc) and critical current den­
sity (Jc) were evaluated and compared.
Fig. 1. Crystal structure of Bi-2223 compound according to the results of
VESTA software.
SrCaCuO superconductor. The low-density pellets were crushed into
powder and Eu2O3 nanopowder were added according to different wt%
(x = 0.0, 0.2, 0.4, 0.6 and 0.8 wt%) and ground again for one hour.
Under the same amount of pressure, the powders were pressed and then
sintered at 850 ◦ C for 48 h.
Analysis of the structure and phase formation was performed at room
temperature using X’Pert PRO PANalytical X-ray Diffractometer (XRD)
in a 2θ range of 5◦ to 90◦ . Meanwhile, the fractured pellet was examined
using Scanning Electron Microscopy (SEM) and Energy Dispersive X-ray
(EDX) to investigate the morphology and elemental composition of the
sample. The electrical resistance was measured with a current of 20 mA
using the standard four-point probe method over a temperature range of
30 to 300 K controlled using a 12 K closed-cycle He-cryostat system,
Lake Shore 9700 Temperature Controller, Keithley 2410 SMU pro­
grammable current source and Keithley 2182A nanovoltmeter. Four
wire probes were attached on top and side of the surface sample using
Ag-paste where the outer probes carry current, I across the surface while
the two inner probes are the voltages probes which measure the voltage,
V in between the outer probes. The temperature was measured from 30
2. Methodology
In this research, all samples were prepared by the conventional solidstate reactions method. The required proportions of raw materials are
weighed of the total weight of 20 g according to BSCCO stoichiometry
which is Bi1.6Pb0.4Sr2Ca2Cu3Oy + xEu2O3 where x = 0.0, 0.2, 0.4, 0.6
and 0.8 wt% by using a sensitive weighing balancing model Setra EL200S (±0.0001 g). The solutions of the samples were prepared by
mixing Bi2O3, PbO, SrCO3, CaCO3 and CuO powders according to sto­
chiometric weights with an absolute ethanol (99.5% purity) in milling
jar together with milling balls. The solutions were wet-milled for 24 h to
obtain a well-mixed dark grey slurry. The slurry obtained was dried out
in the oven for six hours at 120 ◦ C. Once it was dried, the grey powder
was scraped off from the beaker and ground for 30 min in a mortar.
There were two calcination processes involved in the making of Bi-2223
sample, both were done in an ambient atmosphere with intermediate
grinding. The grey precursor obtained was placed in an alumina boat
and undergoes pre-calcination process at 800 ◦ C for 15 h. The second
calcination process also subjected to 830 ◦ C for 15 h. The powder was
then weighed into 2 g samples and each were pressed under a pressure of
30 Mpa. For porous samples, the powder was weighed together with
polycrystalline sucrose, C12H22O11 (weight ratio of mixed powder to
sucrose is 1.9500 ± 0.0001 g : 0.0500 ± 0.0001 g). The bulk obtained
were heated at 200 ◦ C for two hours to eliminate the sucrose. The porous
pellets were sintered at 850 ◦ C for 48 h to produce a single-phase Bi(Pb)
Fig. 2. X-ray diffraction patterns of samples with different addition of Eu
nanoparticles.
2
E.S. Nurbaisyatul et al.
Cryogenics 119 (2021) 103353
Table 1
Lattice parameter and relative volume fraction of different concentration of Eu2O3 nanoparticles addition on low density Bi-2223 superconductor.
Samples, x (wt%)
0.0
0.2
0.4
0.6
0.8
Volume [Å 3]
Lattice parameter [Å]
a
b
c
5.394
5.406
5.409
5.404
5.409
5.399
5.405
5.403
5.401
5.405
37.149
37.169
37.214
37.041
37.062
1082.1
1086.3
1087.7
1081.2
1083.6
Volume fraction [%]
Williamson-Hall
Bi-2223
Bi-2212
Crystallite size (nm)
Lattice strain, Cε (x10-3)
88.3
83.1
82.0
70.6
71.1
11.7
16.9
18.0
29.4
28.9
32.50
47.10
43.37
42.60
43.66
2.2
4.3
3.2
2.5
3.3
Fig. 3. Volume fraction of Bi-2212 and Bi-2223 phases for different concen­
tration of Eu nanoparticles addition.
to 300 K where the critical temperature, Tc zero was defined as the
temperature where the resistance starts to diverge from the lowtemperature horizontal line. The XRD peaks intensities that was
observed can be used to calculate the relative volume fractions by using
the following relations:
Bi − 2223 =
ΣI(2223)
× 100%
ΣI(2223) + ΣI(2212)
(1)
Bi − 2212 =
ΣI(2212)
× 100%
ΣI(2223) + ΣI(2212)
(2)
Fig. 4. Williamson-Hall (W-H) plots of Bcosθ versus 4sinθ for low density
Bi1.6Pb0.4Sr2Ca2Cu3Oy + xEu2O3 where x = 0.0, 0.2, 0.4, 0.6 and 0.8 wt%.
along with their relative amounts of Bi-2212 and Bi-2223 phases. Using
the least square method, the lattice parameters are calculated through
the distances between atomic layers in a crystal, d and (hkl) planes for a
tetragonal unit cell structure. Meanwhile, the volume fraction of Bi2212 and Bi-2223 phases for different concentration of Eu nano­
particles addition was calculated using Eqs. (1) and (2) and depicted in
Fig. 3. Table 1 indicates that both Bi-2212 and Bi-2223 phases were
affaected by the addition of Eu nanoparticles, indicating a reduction in
the volume fraction of the Bi-2223 phase. As can be seen, Bi-2223 peaks
were most prominent in the Eu-free sample. Previous study by Mahtali,
M. and Chamekh, S. mentioned that annealing below 875 ◦ C led to the
formation of the secondary phase with single and multilayer Bi com­
pounds [31]. Besides, the effect of Eu substitution on Ca site in Bi(Pb)2223 superconductor prepared via co-precipitation method have been
studied, indicating that Eu-doped sample which is sintered at 850 ◦ C for
48 h produce more Bi-2212 phase as Eu concentration increases [9].
Based on the calculated lattice parameter, all sample exhibit a
tetragonal structure with different values of lattice parameters. The
variation of a and c values for Eu-free and Eu-addition samples has
slightly different as Eu addition increases but the crystal structure re­
mains unchanged. This result shows that Eu particles did not engage in
the host system. It is consistent with previous research where the crystal
where I(2223) and I(2212) refer to the intensities of the Bi-2223 and Bi2212 peak, respectively.
3. Result and discussions
Fig. 1 shows the crystal structure of Bi-2223 high-temperature su­
perconductor as presented using the VESTA software. X-ray diffraction
patterns for all samples sintered at 850 ◦ C for 48 h were plotted as shown
in Fig. 2. The formation of Bi-2212 and Bi-2223 phases were determined
through peak matching using Xpert Highscore software. Based on Fig. 2,
the Eu-free sample showed a few peaks of low intensity which belongs to
Bi-2212 phase and major peaks belonging to the Bi-2223 phase. Mean­
while, the consistency of peaks can be seen for x = 0.2, 0.4, 0.6 and 0.8
wt% samples as the mixed phases of Bi-2212 and Bi-2223 were detected.
The peaks for the phases of Bi-2212 and Bi-2223 was indicated by (•)
and (*), respectively. XRD patterns clearly demonstrate that the in­
tensity of Bi-2223 phase increased compared to the Eu-free sample
apparent at peaks (0111), (2012), (2 2 2) at 2θ = 31.02◦ , 44.65◦ and
47.66◦ respectively. The peak of Bi-2212 phase around 2θ = 24.89◦ and
33.64◦ corresponding to (1 1 3) and (0 2 2) respectively increased as Eu
nanoparticles concentration increased.
The lattice parameters of the prepared sample are shown in Table 1,
3
E.S. Nurbaisyatul et al.
Cryogenics 119 (2021) 103353
Fig. 5. SEM surface micrographs of low-density Bi1.6Pb0.4Sr2Ca2Cu3Oy + xEu2O3 samples (a) x = 0.0, (b) x = 0.2, (c) x = 0.4, (d) x = 0.6 and (e) × = 0.8 wt%.
structure of Bi-2223 system remains unchanged when adding with Eu
nanoparticles [9]. Furthermore, the decreases in lattice parameters are
possibly due to the change in oxygen contents related to the substitution
of Eu3+ with higher valence cations [32,33]. It is a very challenging task
to understand the effect of substitution on parameters a and b, since
these parameters are controlled by the length of the Cu-O bond in the
plane [34].
The crystallite size, D of the sample was estimated by applying
4
E.S. Nurbaisyatul et al.
Cryogenics 119 (2021) 103353
Fig. 6. Elemental mapping for Bi-2223 sample adding with 0.2 wt% Eu nanoparticles.
Scherer equations [35]:
D =
kλ
Bcosθ
thus preventing grain growth. Hence, it is noticeable to conclude that
the reduce in grain size degrade the connectivity between the super­
conducting grains [38]. Alaghbari et al. reported that when grain size
decreases, the weak link increases and ultimately the current can no
longer flow between the grains, resulting in a lower Jc [39]. Meanwhile,
Kir et al. concludes that some randomly distributed large grains reflect
the presence of Bi-2212 phase thus lower the Tc values due to the weak
connectivity between grains [33].
Fig. 6 presents the elemental mapping of Bi-2223 adding with x =
0.2 wt% of Eu indicating that this Bi-2223 addition superconductor
grains consist of Bi, Pb, Sr, Ca, Cu, O and Eu elements. The summary of
composition elements of low-density Bi-2223 added with Er2O3 nano­
particles are tabulated in Table 2. All the elements such as Bismuth,
Lead, Strontium, Calcium, Copper and Europium which have been taken
initially are present in the atomic% with their appropriate amount. From
the results obtained, all the elements are present in the prepared sam­
ples. The presence of Europium nanoparticles in the superconductor
system is confirmed with the increment of the composition of Neo­
dymium from 0 up to 0.26 atomic%, respectively. These results defi­
nitely indicate that the Europium nanoparticles have been successfully
incorporated into the structure of the low-density Bi-2223.
Fig. 7 displays the electrical resistivity versus temperature from 30 K
to 300 K for all the samples. As can be seen, a smooth resistivity curve is
shown for all samples that correspond to Bi-2223 phase one-step tran­
sition. High-temperature superconductors normally exhibit metallic
behaviour in their resistivity up to their onset temperatures, suggesting
that the resistance behaviour in this range is determined by the degree of
electron–phonon interaction [33]. It is known that the Tc onset values
represent the points at which the relationship deviates from the linear
trend, while the Tc offset values is considered the point where the elec­
trical resistance approaches zero [40]. The critical onset temperature, Tc
onset of the samples were 110, 105, 97, 106 and 107 K for x = 0.0, 0.2,
0.4, 0.6 and 0.8 wt% of Eu nanoparticles addition respectively. The
resistivity values corresponding to the Tc onset of the samples are 0.385,
0.550, 0.542, 0.621 and 0.636 Ω cm respectively. From Fig. 7, it is
obvious that the sample with x = 0.8 wt% has the highest Tc onset value
implying the presence of the many high-sized Bi-2212 grains as stated in
XRD analysis. Finding by Kir et al concludes that lower Tc offset values
indicate that there might be a large amount of porosity due to the
randomly oriented larger grains [33].
Table 3 summarizes the data of Tc onset and Tc offset derived from the
resistance versus temperature, R(T) graph. The highest value of Tc offset
can be seen at 99 K for the Eu-free sample and 89 K was measured in the
sample adding with 0.2 wt% of Eu nanoparticles. However, no conclu­
sion can be drawn as to whether intergrain connectivity has improved or
(3)
where B is the FWHM of the X-ray peaks, k is a shape factor without
dimension with the value of 0.9, θ is the Bragg angle and λ is the
wavelength of X-ray (1.5406 Å). Whereas, Williamson-Hall (W-H) the­
ory has been used to calculate the lattice strain, Cε for each sample
through equations [36]:
Bcosθ = 4Cε sinθ +
kλ
D
(4)
It is possible to determine the value of Cε by plotting the graph Bcosθ
versus 4sinθ as shown in Fig. 4. The value of Cε was deduced based on the
slope of linear fitting of the scattered graph. All the obtained D and Cε
values are recorded in Table 1. The crystallite size is slightly increased
for sample with 0.2 wt% compared to the Eu-free sample, suggesting
that Eu2O3 nanoparticles stimulated crystal growth by filling up the
voids within the grain boundary networks in Bi-2223. The development
of lattice strain results from the displacement of the atoms with respect
to their reference positions in the lattice. Clearly, there is no significant
variation found in the lattice strain as the Eu2O3 nanoparticles increases.
The highest Cε value was found at 0.2 wt% sample. These strains might
be resulting from the lattice shrinkage observed in the lattice parameters
calculation [37].
Fig. 5 displays the SEM micrograph of the fractured bulk samples for
Eu-free sample and Eu-added samples at a magnification of 10000 X
with an energy range of 5.00 kV to 10.00 kV. It is observed that the
microstructure of Bi-2223 superconductor exhibit a common flaky
layers of plate-like grains. The SEM images of Eu-free sample shows the
presence of some pores. These pores lead to poor grains connectivity
thus lead to lower Jc values. Furthermore, orientation and grains con­
nectivity for x = 0.8 wt% sample are slightly worsened as the size of
flaky layers start to reduce and form smaller plate-like grains and
distributed randomly without specific alignment. In addition, calcula­
tion of grain size was extracted from this research using Image J soft­
ware by measuring the average length of both sides of the grain. Noted
also the image of Fig. 5 taken at different spot of fractured samples thus
yielded different value of grain size. The findings display that the
maximum grain size for Eu2O3 nanoparticles addition sample is
observed at 0.2 wt%. However, the value of grain size decrease
compared to Eu-free sample might be due to the increasing of grain
boundaries. As grain growth has always been caused by grain boundary
energy, it makes sense that when Eu2O3 nanoparticles were added into
Bi-2223 system, they settle within the pores and the edges of the grains
5
E.S. Nurbaisyatul et al.
Cryogenics 119 (2021) 103353
Fig. 6. (continued).
destructed in Tc offset values for 0.6 wt% and 0.8 wt% samples [41]. The
difference between Tc onset and Tc offset is known as transition width
difference, ΔTc = Tc onset - Tc offset. From the result obtained, ΔTc
occurred for the sample x = 0.6 wt% due to the phase inhomogeneities
between grains boundaries [42] and the presence of impurities within
the structure of the samples. Meanwhile, the increment in the transition
6
E.S. Nurbaisyatul et al.
Cryogenics 119 (2021) 103353
Fig. 6. (continued).
width difference, ΔTc values as wt% of Eu nanoparticles increase indi­
cating that more than one phase is present and the weak links between
superconducting grains is enhanced [9].
The theoretical density of the pure BSCCO system is about 6.302 g/
cm3 [38]. We determined the densities of samples according to the
Archimedes principle by weighing in distilled water and air. The
calculated densities are listed in Table 3. The table clearly shows that the
density values obtained for addition samples are higher than that of the
Eu-free sample. While the maximum density value is attributed to x =
0.2 wt%, the smallest density value belongs to the x = 0.4 wt% for the
addition sample. The density values obtained are in the range of 78 to
85% compared to the theoretical values.
The variations of the hole-carrier concentration versus Eu content is
Table 2
Summary of composition elements in low density Bi-2223 superconductor added
with different composition of Eu2O3 nanoparticles.
Sample, x (wt%)
0.0
0.2
0.4
0.6
0.8
Element (atomic %)
Bi
Pb
Sr
Ca
Cu
O
Eu
11.27
8.61
8.86
8.31
8.76
2.60
1.82
2.83
2.26
2.22
10.22
9.42
9.37
9.02
9.81
13.09
8.94
9.47
9.83
9.27
33.08
16.51
15.42
13.45
16.67
29.74
54.62
53.93
56.95
53.01
0.00
0.08
0.12
0.18
0.26
Fig. 7. Normalized resistance at room temperature for low density
Bi1.6Pb0.4Sr2Ca2Cu3Oy + xEu2O3 where x = 0.0, 0.2, 0.4, 0.6 and 0.8 wt%.
Fig. 8. Superconductivity transition temperature versus hole concentration.
Table 3
The critical temperatures Tc,onset, Tc,zero, ΔTc and hole concentration of low density Bi1.6Pb0.4Sr2Ca2Cu3Oy + xEu2O3 samples (x = 0.0, 0.2, 0.4, 0.6 and 0.8 wt%).
Samples, x (wt%)
0.0
0.2
0.4
0.6
0.8
Tc
onset
110
105
97
106
107
(K)
Tc
99
89
74
59
62
zero
(K)
ΔTC (K)
Hole concentration, p
Grain size (µm)
Density, ρ (±0.01 g/cm3)
Porosity, P (%)
11
15
23
47
45
0.125
0.111
0.097
0.085
0.087
3.678
1.746
1.634
1.552
1.539
4.88
5.38
4.92
5.04
5.25
22.56
14.63
21.93
20.03
16.69
7
E.S. Nurbaisyatul et al.
Cryogenics 119 (2021) 103353
nanoparticles into BSCCO system enhance the Jc values due to the
improvement of the grains connectivity within Eu nanoparticles addi­
tion. Research by Zelati et al. also pointed out the addition of 0.5 wt% of
Dy2O3 nanoparticles into Bi-2223 superconductor increases the Jc values
where nanoparticles settle between the grains thus improve the grains
connectivity [46]. The latest study by Fallah-Arani et al. revealed that
the optimum Jc value for Bi-2223 superconductor adding with Titanium
Dioxide nanorods (TiO2-NR) found at a sample with 0.2 wt% TiO2-NR
content [4].
4. Conclusion
Low-density Bi(Pb)-2223 samples added with different wt% of Eu
nanoparticles were prepared by the standard solid-state reaction
method. The transition temperature changed with the increase of the wt
% of Eu addition. The optimum value of Tc zero for the addition sample
was found at sample x = 0.2 wt%. X-ray diffraction analysis showed both
Bi-2223 and Bi-2212 phases coexisted in the samples with tetragonal
structure. As the concentration of Eu nanoparticles increased, the vol­
ume fraction calculated from the XRD peaks intensities decreased. All
the elements in the prepared samples were present in the Bi(Pb)-2223
system and was comfirmed by EDX analysis. Based on resistivity mea­
surements, the highest value of Tc zero and Jc for the Eu-adding sample
corresponded at 0.2 wt% sample. Hole concentrations of the samples
were determined from the equation showed decrement as Eu concen­
tration increases. Hence, the best sample in this study was found at x =
0.2 wt% of Eu nanoparticles as it gives the best critical current density
values compared to the Eu-free sample.
Fig. 9. Variation of critical current density versus temperature.
Table 4
Variation of critical current density, Jc at different temperature of
Bi1.6Pb0.4Sr2Ca2Cu3Oy + xEu2O3 samples (x = 0.0 , 0.2, 0.4, 0.6 and 0.8 wt%).
Samples, x (wt%)
0.0
0.2
0.4
0.6
0.8
Critical current density, JC [A/cm2]
30 K
40 K
50 K
60 K
70 K
0.75
7.29
3.18
3.10
2.40
0.46
6.82
3.07
2.41
2.03
0.34
6.20
2.61
1.10
1.01
0.23
5.27
2.16
0.09
–
0.11
4.19
1.36
–
–
Declaration of Competing Interest
depicted in Fig. 8. The hole concentrations shows parabolic relationship
dependence on the critical temperature, Tc. The hole concentrations
were determined based on the equations given below:
Tc /Tc max = 1 − − 82.6 × (p − − 0.16)2
The authors declare that they have no known competing financial
interests or personal relationships that could have appeared to influence
the work reported in this paper.
(5)
Acknowledgements
where p indicates hole concentrations and Tc max for Bi-2223 samples are
considered as 110 K. Previous calculations for the Eu-free Bi-2223
samples had shown that the value of p ranged from 0.116 up to 0.160
[43,44]. Study by Tallon et al. revealed that at p = 0.19 value is where
the superconductivity is most sturdy [45]. Despite the fact that this
research shows a parabolic relationship, the hole concentration values
decrease from 0.125 to 0.087 as Eu addition increases. As suggested by
Presland et al., the holes transferred from the Bi2O2 bilayer was
distributed uniformly onto the three superconducting CuO2 layers in Bi2223 phase [43]. Hence, it can be said that the Eu addition reduces the
grain connectivity of the samples thus decreases the Tc offset values.
The critical current density, Jc of the samples was measured in zero
magnetic fields at temperatures ranging from 30 K to 70 K as shown in
Fig. 9 while Jc values are given in Table 4. The critical current density, Jc
values were calculated using the equation:
Jc =
Ic
A
The authors wish to thank the Ministry of Higher Education Malaysia
(MOHE) for financial support in carrying out this research through
Fundamental Research Grant Scheme (FRGS) no. 600- RMI/FRGS 5/3
(74/2016).
References
[1] Snider E, Dasenbrock-Gammon N, McBride R, Debessai M, Vindana H,
Vencatasamy K, et al. Room-temperature superconductivity in a carbonaceous
sulfur hydride. Nature 2020;586:373–7. https://doi.org/10.1038/s41586-0202801-z.
[2] Pickett WE. Design for a room temperature superconductor 2018;95616.
[3] Hosono H, Yamamoto A, Hiramatsu H, Ma Y. Recent advances in iron-based
superconductors toward applications. Mater Today 2018;21:278–302. https://doi.
org/10.1016/j.mattod.2017.09.006.
[4] Fallah-Arani H, Baghshahi S, Sedghi A. Impact of functionalized SiC nano-whisker
on the flux pinning ability and superconductor features of Bi-2223 ceramics. Ceram
Int 2021;47:3706–12. https://doi.org/10.1016/j.ceramint.2020.09.226.
[5] Anis-ur-Rehman M, Mubeen M. Synthesis and enhancement of current density in
cerium doped Bi(Pb)Sr(Ba)-2223 high Tc superconductor. Synth Met 2012;162:
1769–74. https://doi.org/10.1016/j.synthmet.2012.03.006.
[6] Hamadneh I, Halim SA, Lee CK. Characterization of Bi1.6Pb0.4Sr2Ca 2Cu3O y
ceramic superconductor prepared via coprecipitation method at different sintering
time. J Mater Sci 2006;41:5526–30. https://doi.org/10.1007/s10853-006-0277-3.
[7] Hamadneh I, Agil A, Yahya AK, Halim SA. Superconducting properties of bulk
Bi1.6Pb0.4Sr2Ca2- xCdxCu3O10 system prepared via conventional solid state and
coprecipitation methods. Phys C Supercond Its Appl 2007;463–465:207–10.
https://doi.org/10.1016/j.physc.2007.03.445.
[8] Hawa JS, Azhan H, Yahya SYS, Azman K, Hidayah HN, Norazidah AW. Properties
of rare-earth substitution in Bi(Pb)-2223 superconductor prepared by
coprecipitation method. Adv Mater Res 2014;895:83–6. https://doi.org/10.4028/
www.scientific.net/AMR.895.83.
[9] Hawa JS, Azhan H, Yahya SY, Azman K, Hidayah HN, Norazidah AW. The effect of
Eu substitution onto Ca site in Bi(Pb)-2223 superconductor via co-precipitation
(6)
where Ic is critical current (A) and A refer to the cross-sectional area of
the sample (cm2). Based on the result found in Table 4, we can see that Jc
values of the addition sample with x = 0.2 wt% are higher than those of
the other samples. The enhancement in Jc can also be possible due to the
improvement in the intergrain connectivity thus increases the effective
area of the current flow. This result agrees with the SEM images as
shown in Fig. 4. Note that improving of Jc values can also be explained
by increasing of the conductivity between the grains thus introducing
the new effective pinning centers. On the other hand, a similar expla­
nation is discussed by Zelati et al [41] that stated adding Eu2O3
8
E.S. Nurbaisyatul et al.
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]
[28]
Cryogenics 119 (2021) 103353
method. J Supercond Nov Magn 2013;26:979–83. https://doi.org/10.1007/
s10948-012-2043-x.
Mohammed NH. Optimizing the preparation conditions of Bi-2223
superconducting phase using PbO and PbO2. Mater Sci Appl 2012;03:224–33.
https://doi.org/10.4236/msa.2012.34033.
Chu S, McHenry ME. Critical current density in high-Tc Bi-2223 single crystals
using AC and DC magnetic measurements. Phys C Supercond Its Appl 2000;337:
229–33. https://doi.org/10.1016/S0921-4534(00)00107-6.
Siegal MP, Venturini EL, Marosin B, Aselage TL. Synthesis and properties of Tl-BaCa-Cu-O superconductors. J Mater Res 1997;12:2825–54. https://doi.org/
10.1557/JMR.1997.0378.
Silveira ZC, Nicoletti R, Fortulan CA, Purquerio BM, Carlos S. Ceramic matrices
applied to aerostatic porous journal bearings : material characterization and
bearing modeling (Matrizes cerâmicas com aplicação em mancais aerostáticos
cerâmicos porosos : caracterização do material e modelagem do mancal) 2010;56:
201–11.
Wu IJ, Us KS, Us OH. (12) United States Patent 2011;2.
Amira A, Boudjadja Y, Saoudel A, Varilci A, Akdogan M, Terzioglu C, et al. Effect of
doping by low content of yttrium at Ca and Sr sites of Bi (Pb) -2212
superconducting ceramics. Phys B Phys Condens Matter 2011;406:1022–7. https://
doi.org/10.1016/j.physb.2010.12.052.
Bal S, Dogruer M, Yildirim G, Varilci A, Terzioglu C, Zalaoglu Y. Role of cerium
addition on structural and superconducting properties of Bi-2212 system.
J Supercond Nov Magn 2012;25:847–56. https://doi.org/10.1007/s10948-0111360-9.
Marhaba MRS, Kork RAM, Mawassi IHR. Effect of Fe 2 O 3 Nano-Oxide Addition on
the Superconducting Properties of the (Bi , Pb) -2223 Phase 2014:143–53. https://
doi.org/10.1007/s10948-013-2288-z.
Kothawale RR, Dole BN, Shah SS. Effect of substitution of Ce on superconducting
properties of Bi1. 7Pb0. 3Sr2Ca2− x Cex Cu3O10+ δ system. Pramana-J Phys 2002;
58:871–5.
Arshad M, Qureshi AH. Time and temperature-based study for the production of
high TC phase by sol-gel technique in Pb-BSCCO system. J Therm Anal Calorim
2006;83:415–9. https://doi.org/10.1007/s10973-005-7233-3.
Qureshi AH, Arshad M, Durrani SK, Waqas H. Impact of Pb substitution on the
formation of high Tc superconducting phase in BSCCO system derived through solgel process. J Therm Anal Calorim 2008;94:175–80. https://doi.org/10.1007/
s10973-007-8915-9.
Salleh FM, Yahya AK, Imad H, Jumali MH. Synthesis and formation of TlSr1212
superconductors from coprecipitated oxalate precursors. Phys. C Supercond. its
Appl. 2005;426–431:319–24. https://doi.org/10.1016/j.physc.2005.02.043.
Hamadneh MFNI, Zainal WTTZ. The effect of sintering temperature variation on
the superconducting properties of ErBa 2 Cu 3 O 7 − δ superconductor prepared via
coprecipitation method 2011:1745–50. https://doi.org/10.1007/s10948-0101118-9.
Hannachi E, Slimani Y, Ben Azzouz F, Ekicibil A. Higher intra-granular and intergranular performances of YBCO superconductor with TiO2 nano-sized particles
addition. Ceram Int 2018;44:18836–43. https://doi.org/10.1016/j.
ceramint.2018.07.118.
Xu S, Gu Y, Wu X. Effects of Yb-doping on flux pinning properties in YBa2Cu3O7-δ.
J Rare Earths 2010;28:431–3. https://doi.org/10.1016/S1002-0721(10)60347-X.
Moutalbi N, Ouerghi A, Djurado E, Noudem JG, M’Chirgui A. Vortex pinning in
bulk-processed Y-Ba-Cu-O with ZrO2nano-particles: Optimum pinning center size.
Phys C Supercond Its Appl 2011;471:97–103. https://doi.org/10.1016/j.
physc.2010.12.012.
Farbod M, Rafati Z, Shoushtari MZ. Optimization of parameters for the synthesis of
Y2Cu2O5nanoparticles by Taguchi method and comparison of their magnetic and
optical properties with their bulk counterpart. J Magn Magn Mater 2016;407:
266–71. https://doi.org/10.1016/j.jmmm.2016.01.069.
Dadras S, Gharehgazloo Z. Effect of Au nano-particles doping on polycrystalline
YBCO high temperature superconductor. Phys B Condens Matter 2016;492:45–9.
https://doi.org/10.1016/j.physb.2016.04.005.
Slimani Y, Almessiere MA, Hannachi E, Mumtaz M, Manikandan A, Baykal A, et al.
Improvement of flux pinning ability by tungsten oxide nanoparticles added in YBa
[29]
[30]
[31]
[32]
[33]
[34]
[35]
[36]
[37]
[38]
[39]
[40]
[41]
[42]
[43]
[44]
[45]
[46]
9
2 Cu 3 O y superconductor. Ceram Int 2019;45:6828–35. https://doi.org/10.1016/
j.ceramint.2018.12.176.
Almessiere MA, Slimani Y, Hannachi E, Algarni R, Ben AF. Impact of Dy2O3
nanoparticles additions on the properties of porous YBCO ceramics. J Mater Sci
Mater Electron 2019;30:17572–82. https://doi.org/10.1007/s10854-019-02106-1.
Loudhaief N, Labiadh H, Hannachi E, Zouaoui M, Ben SM. Synthesis of CdS
nanoparticles by hydrothermal method and their effects on the electrical properties
of Bi-based superconductors. J Supercond Nov Magn 2018;31:2305–12. https://
doi.org/10.1007/s10948-017-4496-4.
Mahtali M, Chamekh S. Superconductive and magnetic properties of Bi 2Sr 2Ca
2Cu 3O 10+δ ceramics doped by Pb. J Supercond Nov Magn 2011;24:351–5.
https://doi.org/10.1007/s10948-010-1008-1.
Ozturk O, Akdogan M, Aydin H, Yilmazlar M, Terzioglu C, Belenli I. Substitution of
Sm at Ca site in Bi1.6 Pb0.4 Sr2 Ca2 - x Smx Cu3 Oy superconductors. Phys B
Condens Matter 2007;399:94–100. https://doi.org/10.1016/j.physb.2007.05.028.
Kir ME, Özkurt B, Aytekin ME. The effect of K-na co-doping on the formation and
particle size of Bi-2212 phase. Phys B Condens Matter 2016;490:79–85. https://
doi.org/10.1016/j.physb.2016.03.016.
Lee MS, Song KY. Effect of Nd substitution for the Ca site in the 110 K phase of (Bi,
Pb)-Sr-Ca-Cu-O superconductors. Supercond Sci Technol 2002;15:851–4. https://
doi.org/10.1088/0953-2048/15/6/301.
He K, Chen N, Wang C, Wei L, Chen J. Method for determining crystal grain size by
X-ray diffraction. Cryst Res Technol 2018;53:1–6. https://doi.org/10.1002/
crat.201700157.
Slimani Y, Almessiere MA, Hannachi E, Baykal A, Manikandan A, Mumtaz M, et al.
Influence of WO3 nanowires on structural, morphological and flux pinning ability
of YBa2Cu3Oy superconductor. Ceram Int 2019;45:2621–8. https://doi.org/
10.1016/j.ceramint.2018.10.201.
Bindu P, Thomas S. Estimation of lattice strain in ZnO nanoparticles: X-ray peak
profile analysis. J Theor Appl Phys 2014;8:123–34. https://doi.org/10.1007/
s40094-014-0141-9.
Yildirim G. Beginning point of metal to insulator transition for Bi-2223
superconducting matrix doped with Eu nanoparticles. J Alloys Compd 2013;578:
526–35. https://doi.org/10.1016/j.jallcom.2013.07.016.
Alaghbari ES, Zainal Z, Hamadneh I, Bin HMZ, Bin WMH. Effect of doping of Zn
and Ca into ErBa 2Cu 3O 7-δ superconductor prepared via Co-precipitation
method. J Supercond Nov Magn 2012;25:255–60. https://doi.org/10.1007/
s10948-011-1293-3.
Safran S, Ozturk H, Bulut F, Ozturk O. Experimental and theoretical approaches for
electrical, magnetic, micromechanical, and structural characterization of BSCCO
ceramic superconductors. Ceram Int 2018;44:11674–81. https://doi.org/10.1016/
j.ceramint.2018.03.243.
Zelati A, Amirabadizadeh A, Kompany A, Salamati H, Sonier J. Effect of
Eu2O3nanoparticles addition on structural and superconducting properties of
BSCCO. J Supercond Nov Magn 2014;27:1369–79. https://doi.org/10.1007/
s10948-013-2475-y.
Biju A, Vinod K, Aloysius RP, Syamaprasad U. Improved superconducting
properties by La addition in (Bi, Pb)-2212 bulk superconductor. J Alloys Compd
2007;431:49–55. https://doi.org/10.1016/j.jallcom.2006.05.080.
Presland MR, Tallon JL, Buckley RG, Liu RS, Flower NE. General trends in oxygen
stoichiometry effects on Tc in Bi and Tl superconductors. Phys C Supercond Its
Appl 1991;176:95–105. https://doi.org/10.1016/0921-4534(91)90700-9.
O. Bilgili, Y. Selamet KK. Effects of Li substitution in Bi-2223 superconductors. J
Supercond Nov Magn 2008;21:439–49. https://doi.org/10.1007/s10948-0080374-4.
Tallon JL, Loram JW, Williams GVM, Cooper JR, Fisher IR, Johnson JD, et al.
Critical doping in overdoped high-Tc superconductors: a quantum critical point?
Phys Status Solidi 1999;215:531–40. https://doi.org/10.1002/(sici)1521-3951
(199909)215:1<531::aid-pssb531>3.3.co;2-n.
Zelati A, Amirabadizadeh A, Kompany A, Salamati H, Sonier J. Effects of Dy 2 O 3
nanoparticle addition on structural and superconducting properties of BSCCO
2014;7:123–34.
Download