Vous êtes sur la page 1sur 19

CRC_2308_Ch015.

qxd 11/30/2005 3:56 PM Page 455

15 Nanostructured
Superconductors
Oxide

Pavel E. Kazin and Yuri D. Tretyakov


Moscow State University, Moscow, Russia

CONTENTS

Abstract
15.1 Introduction
15.2 Superconductor Parameters and Magnetic Flux Pinning
15.3 Design of Flux Pinning Centers
15.4 Superconductor Materials with Crystalline Defects
15.4.1 Point Defects
15.4.2 Dislocations and Grain Boundaries
15.4.3 Irradiation Defects
15.5 Superconductor Matrix-Based Composites
15.5.1 Formation of Composites
15.5.2 Composites: Superconductor Matrix — Secondary-Phase Inclusions
15.5.3 Composites: Superconductor Matrix — Foreign-Phase Inclusions
15.5.4 Composites Obtained by Superconductor Phase Decomposition
15.6 Conclusions
Acknowledgments
References

ABSTRACT
Oxide superconductors are considered materials structured on micro- and nanoscale level, a neces-
sary condition for obtaining a superconducting material with high critical current density (Jc) val-
ues. The fundamentals of superconductivity in relation to magnetic flux pinning and Jc are briefly
discussed. Design of superconductors with high density of flux pinning centers is envisaged. The
nanostructured materials are classified by the nature of the pinning centers and the methods of their
formation. Superconductor materials with advanced properties are shown to correspond to either
single-phase materials with crystalline defects or to nanocomposites consisting of superconductor
matrix with ultrafine nonsuperconducting inclusions. Preparation routes and properties of such
materials are reviewed.

Copyright 2006 by Taylor & Francis Group, LLC


CRC_2308_Ch015.qxd 11/30/2005 3:56 PM Page 456

456 Nanomaterials Handbook

15.1 INTRODUCTION
Since the discovery of high Tc oxide superconductors (HTSC) in 1986 by Müller and Bednorz,1
these materials have been attracting considerable attention due to their high superconducting tran-
sition temperature, which exceeds the boiling temperature of liquid nitrogen. It determines their
high potential for application in electronics, for electric power transmission, storage, and utilization,
and for producing strong magnetic fields and magnetic levitation. Till now Tc values up to 135 K
have been achieved in Hg-based cuprates of the HgBa2CanCun1Oz homologue series.2,3 Most prom-
ising for application rare earth- and bismuth-based superconductors exhibit Tc in the range 90 to 110
K.4,5 A very important technical parameter of a superconductor is the critical current density Jc,
which determines maximum electric current that can flow in the material without energy dissipa-
tion. The Jc value has to be high enough and for most applications spans the range of 102 to 107
A/cm2. The first HTSC materials prepared possessed low Jc values that further decreased sharply
with increasing temperature and applied magnetic field. The main reasons for this were weak inter-
grain junctions and low magnetic flux pinning. Elaboration of the preparation technique made it
possible to overcome these problems to a certain extent and to get high-performance HTSC in the
form of epitaxial films, wires with textured superconductor core, and bulk items.6–8
High values of Jc can be achieved in inhomogeneous material only, as the magnetic flux lines pin
to small areas in the superconductor matrix with different electromagnetic properties. As we show in
the next section, advanced superconductors have the necessity to be nanostructured materials.

15.2 SUPERCONDUCTOR PARAMETERS AND MAGNETIC FLUX PINNING


Superconducting state is characterized by zero resistivity and zero magnetic field induction inside
the sample. Below Tc the superconducting state (phase) is stable, and consequently a certain amount
of energy is required to transform the superconductor to normal metal. That occurs, for example,
when high enough magnetic field is applied. Important parameters of the superconducting state are
coherence length ξ, determining the interaction distance of the electron pairs and London penetra-
tion depth λ determining the magnetic field spreading in the superconducting volume.9 For a super-
conductor with λ/ξ1/ 2 , the magnetic field starts to penetrate the material above lower critical
field Hc1 forming quantum vortices comprising of thin channels with suppressed superconductivity
(core) surrounded by circulating current (Figure 15.1), and the main volume of the material remains
superconducting. The vortex core diameter is approximately 2ξ, while the superconducting current
(and bound magnetic field) extends to a distance of λ from the vortex center. When a vortex appears,
some energy (electron pair condensation energy) is consumed to suppress superconductivity in its
core, which is compensated by the energy gain due to the magnetic field penetration in the area sur-
rounding the vortex. Repulsion forces acting between the vortices arrange them in a triangular array.
On increasing the field further, the number of vortices increases and at upper critical field Hc2 the
vortices touch their cores so that the volume superconductivity disappears.
Oxide superconductors are characterized by ξab  2 to 4 nm, λab  100 nm (in ab plane), Hc1 
10−1 T, and Hc2  102 T.10 Between Hc1 and Hc2 the electrical current applied through the material
induces the flow of the vortices under the Lorenz force. The flow is accompanied by energy dissi-
pation, and hence the material cannot be considered as a superconducting one. Fortunately, a vor-
tex can pin when its core intersects, for example, a nonsuperconducting inclusion (see Figure 15.1).
In this case, no energy is required for the field to penetrate the volume of the inclusion, and the vor-
tex energy drops by the value corresponding to the condensation energy in the superconductor vol-
ume equal to the volume of the intersection of the vortex core with the inclusion.8 Thus, the vortex
gets into an energy (pinning) well and is trapped by the inclusion. The well is characterized by pin-
ning energy Ep and pinning force fp  dEp/dx. The former characterizes stability of the pinning cen-
ter toward thermal excitation. The latter determines minimum Lorenz force which must be applied
to move the vortex and hence provides nonzero Jc, which is proportional to volume pinning force

Copyright 2006 by Taylor & Francis Group, LLC


CRC_2308_Ch015.qxd 11/30/2005 3:56 PM Page 457

Nanostructured Oxide Superconductors 457

I ∼

∼ 2

Energy E

Ep
fp = dE/ dx

Distance x

FIGURE 15.1 Magnetic vortices (shown as thin cylinders) in the superconductor. One vortex intersects a
nonsuperconducting inclusion (thick cylinder). Such a position of the vortex corresponds to the energy profile
depicted below (schematically). The energy drop Ep is equal to the energy required to convert superconductor
into metal in the volume of the intersection. The energy well formed causes the vortex to pin to the inclusion.

Fp  fp/V. In general, higher concentration of efficient pinning centers provides higher Jc values
and weaker suppression of Jc with temperature and magnetic field.
For the best performance, the inclusion diameter should be close to the vortex core diameter,
i.e., 4 to 8 nm. Smaller inclusions will give lower Ep and fp, while bigger ones will have consider-
able part of their volume not involved in the interaction.
One can imagine that the ideal structure of the material should represent continuous supercon-
ductor matrix penetrated by a regular array of nonsuperconducting columns a few nanometers in
diameter, the columns being parallel to the magnetic field and perpendicular to the superconduct-
ing current flow. It is a challenging task to design such a composite, and the problem is still not fully
solved. Less appropriate, but still very efficient is a random distribution of aligned columnar or even
equiaxed nanoinclusions or crystalline defects. Calculations show that Jc will grow up to ca. 40
vol% of the included nanophase.11 As a matter of fact large inclusions also pin the vortices; in fact
they can pin several vortices simultaneously, and theoretically Jc can achieve 104 to 105 A/cm2 for
micron-sized inclusions.8 However, because of higher concentration at the same volume fraction,
nanosized inclusions may provide one to two orders of magnitude higher Jc.
The next important issue is the nature of a pinning center. The steepness in the wall of the pinning
well (dEp/dx) depends on the properties of the inclusion material: a metallic inclusion gives low-
gradient dEp/dx due to the delocalization of the metal electrons, an isolator provides “normal” gradi-
ent, while a ferromagnetic particle gives an increased gradient because of fast magnetic suppression

Copyright 2006 by Taylor & Francis Group, LLC


CRC_2308_Ch015.qxd 11/30/2005 3:56 PM Page 458

458 Nanomaterials Handbook

Atomic layers O O O
O Cu O Cu O Cu O
Isolating layers BiO
SrO O O O
Conducting layers CuO2 O Cu O Cu O Cu O
Ca O O O
CuO2
SrO O Cu O Cu O Cu O
c
BiO a b O O O

FIGURE 15.2 Crystal structure features of cuprate superconductors exemplified by Bi2Sr2CaCu2O8x.


Atomic layer sequence along the c-axis (left) and structure of CuO2 plane running along the ab plane (right)
are shown.

Jc

Jc(H ) Jc(T )

H H irr(T ) T

FIGURE 15.3 Schematic diagram of Jc vs. temperature and magnetic field for a superconductor with strong
(full curves) and weak (broken curves) magnetic-flux pinning.

of the superconductor order parameter through the interface. An area with weakened (and even
strengthened) superconductivity also pins a vortex, but the pinning well has lower energy, proportional
to the difference in the condensation energies of the matrix and the inclusion. The wall gradient also
depends on the sharpness of the interface matrix or inclusion. Gradual transition from superconduct-
ing to nonsuperconducting volume causes smoothing of the wall of the pinning well.
A specific feature of HTSCs is their layered crystal structure (Figure 15.2) representing super-
conducting CuO2 planes separated by metal oxide layers with electric isolator character.12 Thus, a
large superconducting current can easily flow in the ab plane while along the c-axis its value is sub-
stantially lower. Therefore, it is important to make textured material and apply the superconducting
current flow in the ab-direction. Such an electromagnetic anisotropy is also manifested at lower val-
ues of ξc (of the order of interatomic distance) as compared with ξab. As a result, the initial c-axis
columnar vortex can split into “pancake” vortices localized within the superconducting ab planes,
which diminishes the effect of pinning. In the case of strong anisotropy, columnar inclusions
become less efficient and with regard to the pinning effect, approach the equiaxed ones. The criti-
cal current density becomes stronger suppressed by the magnetic field and decreases to zero at a
field Hirr (irreversibility field) much smaller than Hc2.
A schematic diagram showing the Jc dependence on H and T for the cases of weak and strong
pinning is presented in Figure 15.3.

15.3 DESIGN OF FLUX PINNING CENTERS


In principle, any inhomogeneity in the superconducting state can be a candidate for flux pinning.
The potential pinning sites can be conventionally divided into two groups: crystal structure defects
and inclusions. In the first case, one considers the superconductor as single-phase material, and in

Copyright 2006 by Taylor & Francis Group, LLC


CRC_2308_Ch015.qxd 11/30/2005 3:56 PM Page 459

Nanostructured Oxide Superconductors 459

Superconductor materials with crystal defects

Grain boudaries, Radiation damage


Point defects dislocations defects

1
YBa2Cu2.999Zn0.001O 7−x YBa2Cu3O 7−x film n0, 1p1 Heavy ions

O O O
O Cu O Cu O Cu O
O O O
235U-doped
O Cu O Zn O Cu O superconductor, 1n0
O O O
U-containing
O Cu O Cu O Cu O precipitates
O O O

Fission tracks

FIGURE 15.4 Crystalline defects in single-phase superconductor, which act as effective pinning centers.

the latter as composite. The defects include point defects such as atomic vacancies, substituted and
interstitial atoms (ions), extended defects such as dislocations and grain boundaries, irradiation
damage defects representing small areas of amorphized material. Formation of such defects in the
superconductor is illustrated in Figure 15.4.
As the included phases the composites may contain secondary phases composed of the same
components as the superconductor phase, or foreign phases containing additional chemical ele-
ments absent in the superconductor phase. In a special case, the suppression of superconductivity
can take place in crystal areas with shifted chemical composition. The nature of the inhomo-
geneities determines approaches to material formation. We will consider particular materials fol-
lowing that classification in the following section.

15.4 SUPERCONDUCTOR MATERIALS WITH CRYSTALLINE DEFECTS


15.4.1 POINT DEFECTS
Often oxide superconductors prepared as homogeneous material (e.g., single crystal) exhibit very
high Jc (105 A/cm2) and Hirr (10 T) at 4.2 K, while at higher temperatures, Jc and Hirr decrease
quickly and at 77 K the material has poor properties. High parameters at low temperature are usu-
ally associated with the presence of large number of pinning centers with low energy, presumably
oxygen vacancies. With the increase in temperature, the raised thermal energy becomes compara-
ble with the pinning energy so that these defects cannot coerce the vortices. It especially concerns
highly anisotropic bismuth-based cuprates Bi2Sr2CaCu2O8x (Bi-2212) and (Bi,Pb)2Sr2Ca2Cu3O10
(Bi-2223).
To increase pinning these superconductors were doped with many metal oxides (usually in
amount of several mol%), and sometimes this led to moderate pinning enhancement, as in the case
of Pb and Pr substitution.13,14 However, the nature of the increase in pinning was not clear, as the
dopants could impact the material in various ways: e.g., cation substitution could alter the charge

Copyright 2006 by Taylor & Francis Group, LLC


CRC_2308_Ch015.qxd 11/30/2005 3:56 PM Page 460

460 Nanomaterials Handbook

carrier density and electromagnetic anisotropy, possible secondary-phase impurities formed could
act as pinning centers as well.
Later on, stronger evidence was obtained on point defect pinning.15,16 Melt-textured bulk super-
conductor YBa2Cu3O7 (Y-123) was doped with “homeopathic” amounts of ZnO. Tc of the sample
decreased a little (ca. 1 K), while the critical current density increased twice at 77 K in zero field in
comparison with the undoped sample. Besides, the increase in Jc became more pronounced in a
magnetic field, sometimes showing “peak effect” (a maximum on the Jc(H) curve).
The Zn effect can be illustrated as follows (see Figure 15.4, left). It is well known that Zn
replaces Cu in CuO2 planes and causes a decrease in Tc.17 Thus, a single Zn ion produces an area
with suppressed superconductivity. The diameter of the area in the ab plane is expected to be of the
order of 2ξab. To allow the superconducting current flow, the areas formed around different Zn ions
should not overlap, which gives an estimation of upper doping limit x in YBa2(Cu1−xZnx)3O7 of the
order of 103. Indeed, the highest Jc achieved was for x  0.0003, and decreased to the Jc value for
undoped Y-123 at x  0.001.16
Theoretically, to make an efficient pinning center the dopant ion should completely destroy
superconductivity at the distance comparable with ξab, i.e., much larger than the interatomic dis-
tance. Hence, one can expect that a foreign metal ion such as Zn causes pair breaking also at the
many neighboring Cu ions. The real mechanism of the doping effect is not yet clear. It cannot be
excluded that the foreign ions could form associates in the lattice, which would be more efficient
for the pinning, or induce some extended defects. For example, in single crystals of Bi-2212 doped
with Ti ions it was observed that substitution of Cu by Ti led to formation of pairs of antiphase
boundaries parallel to c-axis so that the CuO2 planes were broken.19
An important feature of point defect pinning is the very small fraction of dopant required.
Enhancement in Jc and Hirr was observed for Cr-doped (Bi,Pb)2.2Sr2Ca2.2Cu3−xCrxOy silver-sheathed
tapes,18 with the maximum effect found for x0.001. A local maximum on Jc vs. dopant concentra-
tion was reported for Bi-2223 ceramics with contents of TaC, NbC, and HfN additives to be of the
order of 0.1 mass%.20,21 Small amounts of Ti, Zr, and Hf diffused from the silver sheath to Bi-2223
was accounted for the Jc increase observed.22,23 In all these cases, substantial contribution of point
defect pinning to Jc enhancement is quite probable.

15.4.2 DISLOCATIONS AND GRAIN BOUNDARIES


It is well known that Jc in Nb3Sn grows with decrease in the grain size of the polycrystalline mate-
rial.24 In that case, grain boundaries play a role of efficient pinning sites, still remaining transparent
for superconducting current flow. In nontextured oxide ceramics, grain boundaries usually impede
current flow due to low ξab value and are responsible for “weak link” behavior (strong suppression
of intergrain Jc even in low field). Mechanical treatment of oxide superconductors introduces dis-
locations, which contribute strongly to the intragrain Jc. At the same time, the intergrain junctions
can be destroyed, which leads to poor Jc transport properties. To prevent the latter, hot deformation
procedures were applied providing some pinning improvement.25,26
The highest Jc is observed in epitaxial thin films. They are usually prepared by vapor deposi-
tion (laser ablation, magnetron sputtering, and chemical vapor deposition) on a heated single crys-
talline substrate (SrTiO3, NdGaO3, MgO, ZrO2, etc.) having low lattice mismatch with the
superconductor. The films have thickness of 1 µm with their ab plane parallel to the film surface
and exhibit Jc exceeding 106 A/cm2 for R-123 and reaching 105 A/cm2 for Bi-2212 at 77 K.27
Microstructurally they consist of biaxially oriented submicron grains with low-angle boundaries
(see Figure 15.4, center). Each boundary comprises an array of dislocations running perpendicular
to the film surface (along the c-axis). It was proposed that such dislocations were efficient pinning
centers.28 Direct observations of the vortices supported this.29,30 The flux lines are captured along
the dislocations perpendicular to the current flow in the most advantageous way. The defects exhibit
short-range order that leads to a decrease in the inter-vortex repulsion providing stronger pinning.

Copyright 2006 by Taylor & Francis Group, LLC


CRC_2308_Ch015.qxd 11/30/2005 3:56 PM Page 461

Nanostructured Oxide Superconductors 461

The boundary (between dislocations), representing only slightly disturbed crystal lattice, remains a
good conductor.
Other crystalline defects and nanoinclusions in epitaxial films can also contribute to the pinning
to some extent.31

15.4.3 IRRADIATION DEFECTS


Irradiation damage can be produced by exposition of the sample to the beams of particles, or in situ,
by introducing additives containing atoms subjected to nuclear fission (see Figure 15.4, right).
Beam irradiation is extensively used as the defects concentration can be easily controlled by the
quantity of radiation. Irradiation with neutrons and protons gives local defects a few nanometers in
diameter. Heavy ions with high enough energy (Au, Pb, etc., 1 to 20 GeV) produce long columns
of amorphized material with a diameter of the order of 10 nm.32 Such extended defects impact the
flux pinning most strongly, and hence heavy ion irradiation has been studied most extensively.
Single crystals of HTSC having initially low Jc at enhanced temperatures exhibit very pronounced
increase in Jc (by orders of magnitude) and in Hirr after the irradiation. The pinning first grows with
the irradiation dose, then decreases when the fraction of the destroyed superconductor becomes sub-
stantial. The pinning effect also depends on the direction of the channels. In Y-123 with low
anisotropy, substantially stronger effect is observed for the channels running along the c-axis (while
the current is applied in the ab plane). In highly anisotropic Bi-2212, the direction of the channels
is not so important which is accounted for by the pancake vortex pinning.33 Heavy ion irradiation
allows to increase Jc considerably even in high-Jc epitaxial films.34
It should be noted that irradiation with fast protons causes atomic fission resulting in randomly
oriented amorphous tracks in oxide superconductors, especially those containing Hg, Tl, Pb, and
Bi.35 These tracks are similar to columnar defects produced by heavy ion irradiation both in struc-
ture and pinning effect.
To employ nuclear fission in situ the initial precursors were doped with 235U containing com-
pounds. In the doped Y-123 bulk material after the melt processing, the submicrometer-sized pre-
cipitates of Ba(U0.6Y0.4)O3 or Ba2YU0.6Pt0.4O6 (in the presence of platinum) were formed.36,37 The
precipitates themselves acting as pinning centers caused moderate increase in Jc. Then the samples
were irradiated by neutrons, which resulted in the appearance of numerous tracks in the supercon-
ductor matrix. These new pinning sites provided manifold increase in Jc. This technique allowed the
preparation of bulk material with the record Jc value of 3  105 A/cm2 at 77 K.37
Silver-sheathed tapes of Bi-2223 were also doped using U3O8, which resulted in rather uniform
dopant distribution in the material mostly found in fine precipitates of (Sr,Ca)3UO6.38 Neutron irra-
diation led to an increase in Jc in the effective pinning energy and to the shift of the irreversibility
line to higher fields.39

15.5 SUPERCONDUCTOR MATRIX-BASED COMPOSITES


15.5.1 FORMATION OF COMPOSITES
Considering the oxide superconductor material as composite containing two or more phases proved
to be a fruitful way to design material with advanced functional properties.40 Matrix phase of such a
composite is a superconducting one that allows smooth current flow without dissipation on grain
boundaries. It can be dense highly textured polycrystalline ceramics with clean grain boundaries
(directionally solidified ceramics and metal clad wires and tapes) or (quasi) single crystalline mate-
rials (epitaxial films and melt-textured bulks). The dispersed phase (pinning additive) should be uni-
formly distributed in the material to form ultrafine inclusions. The phases have to be chemically inert
toward each other both at preparation and application conditions in order to prevent the suppression
of superconductivity. Besides, the dispersed phase has to be stable toward its particles coarsening

Copyright 2006 by Taylor & Francis Group, LLC


CRC_2308_Ch015.qxd 11/30/2005 3:56 PM Page 462

462 Nanomaterials Handbook

(Ostvald ripening) and segregation. Chemical stability is best provided if both components of the
composite are thermodynamically compatible, i.e., within the range of the material formation condi-
tions (temperature and oxygen partial pressure) they co-exist in chemical and phase equilibrium. The
compatibility gains a special importance since the superconductor phase is often obtained either from
the melt or solid at temperatures close to the melting. Under such conditions most of the substances
regarded as “chemically inert” interact with the superconductor phase or its melt. The interaction has
to be much faster for the pinning additive as it represents highly dispersed phase.
Considering the addition of new components to the superconductor system, it should be ensured
that superconductive thermodynamic parameters are sustained at high level. As the best chemical
compositions for superconductors had mostly been determined in early studies, it is presumable for
new components not to be dissolved in the superconductor phase.
There are several pathways to form a composite from precursor materials, which are schemat-
ically illustrated in Figure 15.5 with several examples. The most straightforward approach is to
choose the components ratio in the precursor, which corresponds to two-phase region on the
phase diagram (figurative points 1, 1): superconductor — dispersed phase, and perform a corre-
sponding thermal treatment. For the preparation of the Bi-2223 core tapes usually a solid-state
thermal treatment is applied, though sometimes very short “shock” melting is introduced as an
intermediate step. For the R-123 bulk as well as for the Bi-2212 bulk and tapes, the most appro-
priate is slow solidification from peritectic melt. The process corresponds to the movement from
point 1 to point 1. In order to get very fine inclusions, the precursors have to be homogeneous
on a submicron level.
Another approach is to prepare single-phase material with certain composition lying within the
superconductor solid solution (figurative points 2 and 3), then change the temperature such as to get
in the two-phase region superconductor — secondary phase (2 → 2, 2 → 2 ). A special type of the
decomposition takes place when on cooling an immiscibility gap appears in the superconductor
solid solution (3 → 3). In this case, the phase separation by spinodal mechanism can occur, lead-
ing to nanostructuring of the material. One of the phases formed with weaker superconductivity will
play a role of a pinning phase.
Similarly, transition to a certain phase region is possible by changing partial pressure of oxygen.

15.5.2 COMPOSITES: SUPERCONDUCTOR MATRIX — SECONDARY-PHASE INCLUSIONS


The multicomponent chemical systems to which oxide superconductors belong are quite complex
and there are many adjacent phases co-existing with the superconductor phase on the phase dia-
grams.4,8,40,41 Quite naturally, these phases were first considered as potential pinning additives.
Widely applied is melt processing of RBa2Cu3O7−x (R-123)–R2BaCuO5 (R-211) bulk compos-
ites, where R  Y, Nd, Eu, Sm, Gd, and their mixtures.8 The R-123 phase melts incongruently
around 1000°C according to the reaction

R-123 → R-211  liquid

On solidification, the reverse reaction does not proceed completely as the solid particles of R-211
dissolve slowly and are trapped by the growing R-123 grains. An addition of the R-211 surplus (20 to
40 mol%) aids to complete the reaction and leads to the formation of two-phase composite
R-123–R-211 (see Figure 15.5). The material is usually prepared by very slow cooling ( 1°C/h) of
the peritectic melt combined with the initialization of the crystallization with small single crystal seed.
The R-123 superconductor forms a matrix composed of quasi-single crystal blocks up to several cen-
timeters in size, while the R-211 phase forms small inclusions. First experiments gave inclusions a few
micrometers in diameter.8 Nevertheless, high Jc exceeding 104 A/cm2 at 77 K was obtained. By tun-
ing the material composition and preparation conditions, the R-211 particles were refined to a submi-
cron size, which allowed to reach Jc above 105 A/cm2 at 77 K.42–45 It was theoretically estimated that

Copyright 2006 by Taylor & Francis Group, LLC


CRC_2308_Ch015.qxd 11/30/2005 3:56 PM Page 463

Nanostructured Oxide Superconductors 463

Liquid (L)

Temperature
L+SC
L+X

2″ 1
3
2
SC
SC+X

2′
1′
3″

Superconductor Inclusion
phase (SC) phase (X)

Composite: SC - secondary phase


1 1′

Melting Slow cooling

Oxide mixture Y2BaCuO5 YBa2Cu3O7−x


Melt
Y−Ba−Cu−O paricles matrix

Composite: SC - foreign phase


1 1′

Melting Slow cooling

Oxide mixture Melt Peritectic phase SrZrO3 Bi-2212 Grain


Bi−Sr−Ca−Cu−Zr−O particles particles matrix boundaries

SC phase decomposition
2 2′ 3 3′

Low T
annealing Cooling

Bi-2212 Secondary phase Bi,Pb-2212 Pb-enriched


ceramics precipitates crystal Pb-poor

FIGURE 15.5 Pathways of superconductor composite formation. Illustrated by the figurative point move-
ment on the schematic phase diagram and synthesis schemes of particular materials.

flux pinning by the inclusions alone could explain the Jc values observed, though other defects, includ-
ing those arising from the inclusions (e.g., dislocations), could also contribute.8 Studies of the bulk
pinning force dependence on the magnetic field supported a dominant role of the inclusions (or inclu-
sion–matrix interface).46,47
Further refinement of R-211 inclusions is expected to give even higher Jc. The obstacle could
be an increasing tendency for the particles to be pushed by the growing R-123 crystal. The theory

Copyright 2006 by Taylor & Francis Group, LLC


CRC_2308_Ch015.qxd 11/30/2005 3:56 PM Page 464

464 Nanomaterials Handbook

of the interaction of a growing crystal with an inclusion introduces critical crystal growth rate vc
below which the trapping rate of the foreign particles dramatically decreases.48,49

νc  ∆σ/ηr, ∆σ  σ1  σ2 σ3 (15.1)

where σ1, σ2, and σ3 are the energies of interfaces particle — crystal, particle — melt and crystal —
melt, respectively, η the melt viscosity, and r the particle radius. For very small particles vc becomes
high and it is not possible to incorporate large amounts of ultrafine inclusions in the matrix using the
low crystallization rate required for the formation of high-quality superconductor. As was shown in
[50], ultrafine Y-211 grains indeed were pushed out of the crystal. Besides, some dopants increased a
pushing effect (e.g., Pt) and for Ru doping the Y-211 particles were fully expelled from the Y-123
bulk.51 The dopant likely changes the interface energies resulting in the increase in ∆σ. The right way
out seems to be to diminish ∆σ or even make it negative, and probably some other doping will do it.
Similar to R-123 materials the (Bi-2212 — secondary phase) composites can be obtained. By slow
solidification of the oxide melt with surplus of CaO and CuO the Bi(Pb)-2212 material was prepared,
which contained high concentration of Ca2CuO3 inclusions.52,53 Using a ferromagnetic particles dec-
oration technique, it was shown that vortices pinned to the particle–matrix interface.54 Single crystals
of Bi-2212 with elongated CuO particles were obtained by crystallization from the KCl flux.55
In melt-textured Bi-2212, the secondary-phase particles formed are usually several microns in
size, and it seems problematic to make them much smaller. This is related to the fact that the peri-
tectic melt contains all the components of the system in comparable quantity and the secondary-
phase particle coarsening occurs during the processing.

15.5.3 COMPOSITES: SUPERCONDUCTOR MATRIX — FOREIGN-PHASE INCLUSIONS


Usage of foreign-phase inclusions extends the possibilities to vary microstructure and properties of
the composites.
In melt-processed R-123, at first additional components were introduced to impact on the dis-
tribution of R-211 phase in the composite. The addition of small amounts of Pt, CeO2, and ZrO2 led
to the R-211 particles refinement. The additives themselves often formed particles of foreign
phases. A range of satellite phases such as phases of Y–Ba–Cu–Pt–O system,56,57 BaSnO3,58
YBa2SnO5.5, and its solid solutions were identified in the R-123 matrix.59 All of them formed sub-
micrometer grains, which were partially pushed from the R-123 crystallites. These phases were
often considered as undesired impurities. The situation changed when detailed investigation of
the uranium-oxide-doped Y-123 revealed highly dispersed inclusions of Ba(U0.6Y0.4)O3,
Ba2Y(U0.6Pt0.4)O8.36,37 Their presence was related to increased Jc observed in the material prior to
neutron irradiation. This encouraged the search for other possible pinning additives, as a result of
which stable phases containing Mo and W oxides were found.60,61 In particular, the presence of
Ba2Y(W0.5Pt0.5)O6 inclusions of 200 to 300 nm in size provided ca. 60% increase in Jc.61
Bi-2212 superconductor looks promising for application as it can be easily melt processed to
form highly textured material by that providing uniform superconducting current distribution along
conductors such as silver-sheathed tapes, wires, and long rods. Yet, extremely high anisotropy
causes low pinning ability at increased temperatures and artificial high-energy pinning centers
would help to overcome the problem. As was mentioned above the effect of second phases of the
same system was restricted if the melt processing was applied, and therefore the application of
foreign-phase additives acquires a special importance.
An extended study on introduction of new components to the Bi–(Pb)–Sr–Ca–Cu–O system
showed that a range of metal oxides could be added to the Bi-2212- and Bi-2223-based materials with-
out significant incorporation of their ions into the superconductor crystal lattice thus conserving high
Tc values.41,62,63 The additives form separate phases chemically compatible with Bi-2212 and Bi-2223.
In most cases, the phases represent complex oxides, so if a foreign binary oxide is added it
reacts with the superconducting phase consuming certain oxides of the system and thus partially
Copyright 2006 by Taylor & Francis Group, LLC
CRC_2308_Ch015.qxd 11/30/2005 3:56 PM Page 465

Nanostructured Oxide Superconductors 465

destroying the material. Hence, it is important that the additive (or surplus to the precursor) should
have quite certain elements ratio. The only binary oxide found to be partially compatible with the
superconductors is magnesium oxide.64,65 On annealing in contact with Bi-2212, being in highly
dispersed state, pure MgO acquires copper oxide, so that the equilibrium composition is
Mg1xCuxO, where x  0.08–0.15. As long as the free energy of MgO reaction with CuO is small
enough, single-crystalline MgO does not interact with Bi-2212 and its melt and proved to be an
excellent substrate for the melt processing of the Bi-2212 material.
Using certain melt processing of a precursor, fine inclusions of the compatible phases can be
created in the Bi-2212 superconductor matrix. Different routes of the precursor preparation can be
applied (they are further consequently numbered). Route I considers preparation of single highly
homogeneous precursor containing all chemical components of the target composite. Co-precipita-
tion and sol–gel methods make it possible to get component mixing on an atomic scale. The same
is applicable to the homogeneous oxide glass prepared by rapid quenching of the melt. Route II
involves mixing of the final components of the composite: superconductor and nanodispersed for-
eign phase. In Route III, two precursors are mixed which on subsequent chemical reaction give an
equilibrium assemblage: superconductor — foreign phase.
As-obtained precursor is subjected to partial melting followed by slow crystallization of Bi-2212
from the melt (see Figure 15.5). The foreign phase forms small solid particles suspended in the melt
which are trapped by the growing Bi-2212 grains. The composite ceramics contain large Bi-2212
plate-like grains (of the order of 100 µm) and foreign-phase particles distributed inside and between
the Bi-2212 grains. The particle size, shape, and spatial distribution in the composite strongly depend
on the nature of the foreign phase as well as on preparation conditions. In Figure 15.6, electron
micrographs of some composites with different inclusion phases are shown.

(a) (b)

5 µm 2 µm

(c) (d)

10 µm 5 µm

FIGURE 15.6 Scanning electron microscopy images (polished cross-sections, back-scattered electrons) of
the melt-processed Bi-2212 superconductor with foreign-phase inclusions. Dark-gray and black rounded and
elongated particles, inclusions of foreign phase; light-gray matrix, Bi-2212. Foreign phases: (a) Mg1−xCuxO,
(b) SrZrO3, (c) (Sr,Ca)In2O4, and (d) Sr5(PO4)3CuxOHy.
Copyright 2006 by Taylor & Francis Group, LLC
CRC_2308_Ch015.qxd 11/30/2005 3:56 PM Page 466

466 Nanomaterials Handbook

Particles of the majority of foreign phases tend to segregate partially on the Bi-2212 grain
boundaries. The exception is Mg1−xCuxO, which is always found to be uniformly distributed in the
matrix (Figure 15.6a).
By melt processing the mixture of Bi-2212 and 10-nm-sized MgO powder, material with inclusions
in the range of 200 to 300 nm is obtained.65 During the processing, apparently foreign phase grain
growth takes place. Application of sol–gel precursor (route II) gives even larger grains of Mg1xCuxO.
That outlines the problem of keeping the included particle size small enough. The composites exhibit
enhanced Jc both at 5 and 77 K, with the increase more pronounced at high temperature.
Finely dispersed magnesium oxide was also used as pinning additive to Bi-2212 tapes, bulk,
and single crystals.66–68
Doping of Bi-2212 with Al2O3 has revealed two thermodynamically compatible foreign phases
BiSr1.5Ca0.5Al2Oz and (Sr,Ca)3Al2O6, forming micron-sized grains in the material.69 The former
phase particles are totally pushed from the Bi-2212 crystallites, while the latter phase particles are
trapped. Using an oxide glass precursor the (Sr,Ca)3Al2O6 particle size was diminished by the order
of a magnitude,70 yet the particles started to aggregate. Considering Equation (15.1) for a fixed Bi-
2212 crystal growth rate vc, the particle size r becomes too small for efficient trapping. In contrast
the sub-micron Mg1xCuxO particles are totally trapped, which allows very low or even negative
value for ∆σ.
Perovskites SrAO3, where AZr, Hf, Sn, under similar conditions form in the Bi-2212 matrix
finer particles of 100 to 300 nm in size (see Figure 15.6b).62,71,72 However, even application of nano-
sized SrAO3 powders (route II) does not give the resulting particles grain size lower than 100 nm.
The particles are partially incorporated into Bi-2212 grains and partially agglomerated, forming thin
closed areas between the Bi-2212 lamellas so that many direct junctions between the lamellas are
sustained. That allows to obtain high transport Jc in spite of the inhomogeneous foreign-phase dis-
tribution. For example, in the silver-sheathed tapes the SrZrO3 inclusions (prepared using route I
from co-precipitated oxalates) provided considerably better superconducting parameters.71 In
Particular, in comparison with Bi-2212 undoped tapes at 5 K the difference in Jc was negligible,
while at 60 K in zero external field Jc increased twice and in the field 50 mT, the increase exceeded
an order of magnitude. By magnetization decay measurements, it was shown that SrZrO3 doping led
to appearance of additional number of pinning centers with high energy.71 It should be mentioned
that positive effects of other phase inclusions is also mostly observed at enhanced temperatures.
Unlike foreign phases discussed above, (Sr,Ca)In2O4 forms anisotropic rod-like particles in the
Bi-2212 material (Figure 15.6c).73 By varying the preparation route, it is possible to get submicron-
sized particles with different aspect ratio. Particles with higher aspect ratio are better trapped by the
Bi-2212 grains. The most anisotropic particles were prepared by melt processing of the homoge-
neous oxide glass precursor.74
An apatite type Sr5(PO4)3CuxOHy phase75 yields submicron needles in the Bi-2212 matrix
(Figure 15.6d). Elongated inclusions are expected to be more favorable for flux pinning in compar-
ison with equiaxed ones if the magnetic field has a direction parallel to the long dimension of the
particle. Therefore, the next step would be to create aligned rods of such foreign phase in the super-
conductor matrix.
Route III utilizes topochemical reaction of fine grains of the compound containing foreign ele-
ment with the molten precursor of the Bi–Sr–Ca–Cu–O system. This way complex-shaped inclu-
sions can be formed.76,77 Thus, micron-sized grains of Sr2SnO4 transform into perforated shells of
SrSnO3 (Figure 15.7), while Sr3Sn2O7 give serrated particles of SrSnO3. Similar reaction with
Sr2TiO4 results in SrTiO3 shell-like crystals. Ca2SnO4 grains with the size of 1 µm are converted
into submicron grains of SrSnO3, so that dispersed-phase refinement takes place.
A substantial advance in nanocomposite design was the synthesis of Bi-2212 thick films with
aligned uniaxial inclusions.78 On a single crystal of MgO with specially prepared surface, a “forest”
of MgO nanorods was grown using Mg vapor oxidation (Figure 15.8). The nanorods grew perpen-
dicular to the substrate surface, had mean diameter of 25 nm, length of
2 µm, and areal density
of 3109 cm2. Then, the amorphous BSCCO layer of 1- to 2-µm thickness was deposited by laser
Copyright 2006 by Taylor & Francis Group, LLC
CRC_2308_Ch015.qxd 11/30/2005 3:56 PM Page 467

Nanostructured Oxide Superconductors 467

5 µm

FIGURE 15.7 Scanning electron microscopy image (polished cross-section, back scattered electrons) of the
Bi-2212–SrSnO3 composite obtained by melt processing of the mixture of the Bi–Sr–Ca–Cu–O oxide precur-
sor with the Sr2SnO4 micron-sized powder. Dark-gray micron-sized broken circles, shell-like inclusions of
SrSnO3; light-gray matrix, Bi-2212.

MgO Whiskers

Bi2Sr2CaCu2O8+x
Layer

MgO Substrate
(single crystal)

FIGURE 15.8 Thick film of Bi-2212 with MgO nanowhiskers aligned perpendicular to the film surface on
the MgO substrate.

sputtering and was subjected to a conventional melt texturing processing. The MgO nanorods pen-
etrated the Bi-2212 matrix perpendicular to the film surface. As mentioned above, such columnar
defects had to be most efficient for flux pinning. The introduction of nanorods resulted in the
several-fold increase in Jc, which was more pronounced at elevated temperatures. The irreversibil-
ity field values increased roughly twice.
The Bi-2223 superconductor was doped also with magnesium oxide.79–81 It was introduced in
the Bi-2223 silver-sheathed tapes in the form of nanoparticles or nanowhiskers and some increase
in Jc was observed.
Bi-2223 superconductor is usually obtained by a solid-state reaction. One can expect that the
foreign-phase particles coarsening will not be so pronounced as for the melt processing: however
the trapping ability of the particles is under question, as the superconductor grains grow much
slower. As of now, there is not enough data on composites microstructure to consider these issues.

15.5.4 COMPOSITES OBTAINED BY SUPERCONDUCTOR PHASE DECOMPOSITION


The Bi-2212 and Bi-2223 phases exhibit extended ranges of cation solid solutions, with the compo-
sitional boundaries depending on temperature and oxygen partial pressure. That makes them con-
venient to perform partial decomposition of the initially single-phase material by changing the
temperature and oxygen partial pressure. This approach was used to synthesize the Bi-2212 ceram-
ics containing sub-micrometer-sized inclusions of Bi-2201, Sr3Bi2O6, Sr14Cu24O41, and Ca2CuO382,83
Copyright 2006 by Taylor & Francis Group, LLC
CRC_2308_Ch015.qxd 11/30/2005 3:56 PM Page 468

468 Nanomaterials Handbook

and Bi-2223 ceramics containing inclusions of Bi-2212, Ca2CuO3, and Ca2CuO3.84 For example, at
870°C, pure Bi-2212 was prepared enriched with Sr for account of Ca. Then the sample was heat
treated at 700°C for a restricted time, which caused precipitation of secondary-phase inclusions of
100 nm in size, while Bi-2212 adopted composition with lower Sr to Ca ratio (see Figure 15.5). After
this procedure, the intra-grain Jc at 30 K increased 3 to 6 times. Longer heat treatment resulted in the
secondary-phase particles coarsening and decrease in Jc. Annealing of Bi-2212 at 550°C led to pre-
cipitation of smaller secondary-phase particles (20 to 100 nm).85
In Refs. 86, 87 the low temperature oxidation of rare earth for alkaline earth substituted Bi-2212
ceramics was investigated. Precipitation of second phases was observed for R  La, Pr, Nd. The
fastest precipitation occurred for Pr-substituted samples, which was explained by oxidation of Pr3
to Pr4 followed by its segregation in SrPrO3.
Solid solution separation for composites formation was applied both for Bi-2212 and R-123
superconductors. Phase separation is observed in the lead-doped Bi-2212 with the composition
Bi2−xPbxSr2CaCu2O8z when x exceeds 0.4.88,89 In single crystals as well in the large-grained ceram-
ics on cooling, sub-micrometer-sized domains arise, which form layered pattern or appear as
nanoscale rounded areas in the matrix (see also Figure 15.5). The crystal structure remains the
same, but the chemical composition of the domains is different. The domains are enriched with lead
while the matrix is depleted of lead. It is assumed that enriched with lead small areas have lower Tc
values and play the role of efficient pinning centers. This explains the fact that the materials exhibit
considerably increased Jc and Hirr values in comparison with the samples with x  0.4.
The R-123 compound forms an extended solid solution range R1xBa2xCu3O7z when rare
earth cation has large radius (La, Nd, Pr). With increasing x, the Tc value decreases quickly. The
Nd-containing compound has been studied most extensively.47,90,91 Samples of Nd1xBa2xCu3O7z
with x  0 prepared by melt texturing undergo a solid solution separation under particular condi-
tions. The phase with larger x forms nanometer-sized inclusions in the matrix with lower x. The
material shows better Jc values at enhanced magnetic field and its M(H) hysteresis curve is often
characterized by a peak effect (at an intermediate field Jc first increases with field, reaching a max-
imum and then decreases back). The effect is associated with the destruction of the superconduc-
tivity under the magnetic field in weakly superconducting regions that makes them effective pinning
centers. The dependence of a pinning force upon the reduced magnetic field H/Hirr passes through
a maximum at H/Hirr  0.4–0.5, which indicates a pinning of magnetic vortices to small regions
with weakened superconductivity.47,90,91 In contrast to that, pinning to nonsuperconducting inclu-
sions is characterized by the pinning force maximum at a twice lower H/Hirr. Similar nanoscale
weakly superconducting regions are believed to be responsible for pronounced peak effects in
(Nd,Eu,Gd)-123 and (Nd,Sm,Gd)-123.92,93

15.6 CONCLUSIONS
Nanostructuring of a superconductor is a key condition to obtain material with higher critical current
density values sustained at increased temperatures and magnetic fields. Considerable enhancement in
Jc has been achieved by the introduction of high concentration of pinning centers such as crystalline
defects or ultrafine nonsuperconducting phase inclusions. General trends in the material development
can be defined as moving from micro- to nanoscale pinning centers, constructing the pinning sites
with a certain shape (columns), creating a correlated array of the pinning sites. It should be recognized
that in a complex material, the advanced parameters obtained can be associated frequently with many
factors impacting simultaneously, such as different kinds of pinning centers, quality of inter-grain
junctions, and modified matrix superconductor characteristics (lower anisotropy and higher charge
carrier density). Further progress in high-performance oxide superconductor design is connected with
closer understanding of a particular microstructure and its relation to the superconducting character-
istics, with thorough selection of the material chemical composition in multicomponent systems and
development of processing to control efficiently the materials structure on a nanoscale level.

Copyright 2006 by Taylor & Francis Group, LLC


CRC_2308_Ch015.qxd 11/30/2005 3:56 PM Page 469

Nanostructured Oxide Superconductors 469

ACKNOWLEDGMENTS
The work was partially supported by the Russian Foundation for Basic Research and the Russian
Ministry of Education under a Universities of Russia program.

REFERENCES
1. K.A. Müller and J.G. Bednorz, The discovery of a class of high temperature superconductors, Science
237, 1133–1139 (1987).
2. S.N. Putilin, E.V. Antipov, O. Chmaissem, and M. Marezio, Superconductivity at 94 K in
HgBa2CuO4y, Nature 362, 226–228 (1993).
3. E.V. Antipov, A.M. Abakumov, and S.N. Putilin, Chemistry and structure of Hg-based superconduct-
ing Cu mixed oxides, Supercond. Sci. Technol. 15, 31–49 (2002).
4. P. Majewski, Phase diagram studies in the system Bi–Pb–Sr–Ca–Cu–O–Ag, Supercond. Sci. Technol.
10, 453–467 (1997).
5. R.J. Cava, Oxide superconductors, J. Am. Ceram. Soc. 83, 5–28 (2000).
6. J.L. MacManus-Driscoll, Recent developments in conductor processing of high irreversibility field
superconductors, Ann. Rev. Mater. Sci. 28, 421–462 (1998).
7. P. Vase, R. Flükiger, M. Leghissa, and B. Glowacki, Current status of high-Tc wire, Supercond. Sci.
Technol. 13, 71–84 (2000).
8. M. Murakami (Ed.), Melt Processed High-Temperature Superconductors, World Scientific, Singapore,
1992.
9. E.H. Brandt, Vortices in superconductors, Physica C 369, 10–20 (2002).
10. E.Z. Meilikhov and V.G. Shapiro, Critical fields in high-temperature superconductors (review),
Sverkhprovodimost’: fizika, khimia, tekhnika 4, 1437–1492 (1991).
11. L.D. Cooley and L.R. Motowidlo, Advances in high-field superconducting composites by addition of
artificial pinning centres to niobium–titanium, Supercond. Sci. Technol. 12, R135–R151 (1999).
12. A.P. Malozemoff, L. Krusin-Elbaum, and J.R. Clem, Interlayer coupling in high temperature super-
conductors, Physica C 162–164, 353–354 (1989).
13. M. Wakata, S. Takano, F. Munakata, and H. Yamauchi, Effects of cation substitution on flux pinning
in Bi 2212 supercoductors, Cryogenics 32, 1046–1051 (1992).
14. P.L. Paulose, S. Patil, H. Franck, and G. Guntherodt, Influence of Pb and Nb substitution on pinning
and irreversibility behavior of Bi 2212, Physica C 208, 11–17 (1993).
15. G. Krabbes, G. Fuchs, P. Shaetze, S. Gruss, J.W. Park, F. Hardinghaus, G. Stoever, R. Hayn, S.-L.
Drechsler, and T. Fahr, Zn doping of YBa2Cu3O7 in melt textured materials: peak effect and high
trapped fields, Physica C 330, 181–190 (2000).
16. M.T. González, N. Hari-Babu, and D.A. Cardwell, Enhancement of Jc under magnetic field by Zn dop-
ing in melt-textured Y–Ba–Cu–O superconductors, Supercond. Sci. Technol. 15, 1372–1376 (2002).
17. N. Kawaji, K. Muranaka, Y. Oda, and K. Asayama, Tc suppression and hole concentration in Cu sub-
stituted YBa2[Cu1xMx]3O7−δ with Fe, Co, Ni, Zn, Physica B 165–166, 1543–1544 (1990).
18. M.H. Pu, Y. Feng, P.X. Zhang, L. Zhou, J.X. Wang, Y.P. Sun, and J.J. Du, Enhanced the flux pinning
in Bi-2223/Ag by induced Cr-ion defects, Physica C 386, 41–46 (2003).
19. T.W. Li, R.J. Drost, P.H. Kes, C. Traeholt, H.W. Zandbergen, N.T. Hien, A.A. Menovsky, and J.J.M.
France, TEM analysis of planar defects induced by Ti doping in Bi-2212 single crystals, Physica C
290, 239–251 (1997).
20. B.P. Mikhailov, P.E. Kazin, V.V. Lennikov, S.V. Shavkin, G.V. Laskova, and A.A. Titov, Influence of
finely dispersed additions of niobium carbide on the structure and superconducting properties of
(Bi,Pb)2Sr2Ca2Cu3O10x ceramics, Neorg. Mater. 37, 753–757 (2001).
21. B.P. Mikhailov, G.S. Burkhanov, P.E. Kazin, V.V. Lennikov, M.V. Makarova, I.A. Rudnev, A.E.
Khodot, A.V. Eremin, and A.A. Titov, Effect of HfN microadditions on the microstructure and super-
conducting properties of (Bi,Pb)2Sr2Ca2Cu3O10x ceramics, Neorg. Mater. 39, 462–468 (2003).
22. M. Ishizuka, Y. Tanaka, and H. Maeda, Superconducting properties and microstructures of Bi-2223
Ag–Cu alloy sheathed tapes doped with Ti, Zr, or Hf, Physica C 252, 339–347 (1995).
23. M. Ishizuka, Y. Tanaka, T. Hashimoto, and H. Maeda, Influence of Cu composition and sintering condi-
tion in Bi-2223 tapes using Ag–Cu alloy sheath doped with Ti, Zr or Hf, Physica C 290, 265–274 (1997).

Copyright 2006 by Taylor & Francis Group, LLC


CRC_2308_Ch015.qxd 11/30/2005 3:56 PM Page 470

470 Nanomaterials Handbook

24. C.S. Pande, Microstructural aspects of high and low Tc superconductors, Mater. Phys. Mech. 2, 1–9
(2000).
25. S.X. Dou, H.K. Liu, M.H. Apperley, K.H. Song, C.C. Sorrell, K.E. Easterling, J. Niska, and S.J. Guo,
Improvement of critical current density in the Bi–Pb–Sr–Ca–Cu–O system through hot isostatic press-
ing, Physica C 167, 525–528 (1990).
26. R.R. Daminov, M.F. Imayev, M. Reissner, W. Steiner, M.V. Makarova, and P.E. Kazin, Improvement
of pinning in Bi2212 ceramics by hot plastic deformation, Physica C 408–410, 46–47 (2004).
27. O. Yu Gorbenko, V.N. Fuflygin, Yu.Yu. Erokhin, I.E. Graboy, A.R. Kaul, Yu.D. Tretyakov, G. Wahl,
and L. Klippe, YBCO and BSCCO thin films prepared by wet MOCVD, J. Mater. Chem. 4,
1585–1589 (1994).
28. V.M. Pan, A.L. Kasatkin, V.L. Svetchnikov, and H.W. Zandbergen, Dislocation model of supercon-
ducting transport properties of YBCO thin films and single crystals, Cryogenics 33, 21–27 (1993).
29. B. Dam, J.M. Huijbregtse, F.C. Klaassen, R.C.F. Van der Geest, G. Doornbos, J.H. Rector, A.M. Testa,
S. Freisem, J.C. Martinez, B. Stäuble-Pümpin, and R. Griessen, Origin of high critical currents in
YBa2Cu3O7−δ superconducting thin films, Nature 399, 439–442 (1999).
30. A. Volodin, K. Temst, Y. Bruynseraede, C.V. Haesendonck, M.I. Montero, I.K. Schuller, B. Dam, J.M.
Huijbregtse, and R. Griessen, Magnetic force microscopy of vortex pinning at grain boundaries in
superconducting thin films, Physica C 369, 165–170 (2002).
31. H. Yamada, H. Yamasaki, K. Develos-Bagarinao, Y. Nakagawa, Y. Mawatari, J.C. Nie, H. Obara, and
S. Kosaka, Flux pinning centres correlated along the c-axis in PLD-YBCO films, Supercond. Sci.
Technol. 17, 58–64 (2004).
32. J. Wiesner, C. Traeholt, J.-G. Wen, H.-W. Zandbergen, G. Wirth, and H. Fuess, High resolution elec-
tron microscopy of heavy-ion induced defects in superconducting Bi-2212 thin films in relation to
their effect on Jc, Physica C 268, 161–172 (1996).
33. W. Gerhauser, G. Ries, H.W. Neumuller, W. Schmidt, O. Eibl, G. Saemann-Ischenko, and S.
Klaumunzer, Flux line pinning in Bi2Sr2Ca1Cu2Ox crystals: interplay of intrinsic 2D behavior and irra-
diation induced columnar defects, Phys. Rev. Lett. 68, 879–882 (1992).
34 B. Roas, L. Schultz, B. Hensel, G. Saemann-Ischenko, and G. Endres, Superconducting properties of
irradiation induced effects in epitaxial Y–Ba–Cu–O thin films, Thin Solid Films 174, 179–183
(1989).
35. J.R. Thompson, J.G. Ossandon, L. Krusin-Elbaum, H.J. Kim, K.J. Song, D.K. Christen, and J.L.
Ullmann, Vortex pinning in high-Tc materials via randomly oriented columnar defects, created by GeV
proton-induced fission fragments, Physica C 378–381, 409–415 (2002).
36. R. Weinstein, R. Sawh, Y. Ren, M. Eisterer, and H.W. Weber, The role of uranium chemistry and ura-
nium fission in obtaining ultra-high Jc in textured Y123, Supercond. Sci. Technol. 11, 959–962 (1998)
37. R. Weinstein, R. Sawh, Y. Ren, and D. Parks, The role of uranium with and without irradiation, in the
achievement of Jc ≈ 300000 A cm−1 at 77 K in large grain melt-textured Y123, Mater. Sci. Eng. B 53,
38–44 (1998).
38. D. Milliken and S.X. Dou, Chemistry of uranium compound doping in (Bi,Pb)2Sr2Ca2Cu3Ox/silver
superconducting tapes, Physica C 341–348, 1411–1414 (2000).
39. D. Marinaro, S.X. Dou, J. Horvat, Y.C. Guo, J. Boldeman, A. Gandini, R. Weinstein, R. Sawh, and
Y. Ren, The effects of uranium doping and thermal neutron irradiation on the pinning properties of
Ag/Bi-2223 Tapes, Physica C 341–348, 1119–1120 (2000).
40. P.E. Kazin and Yu.D. Tretyakov, Microcomposites based on superconducting cuprates, Russian Chem.
Rev. 72, 849–865 (2003).
41. Yu.D. Tretyakov and E.A. Goodilin, Chemical principles of preparation of metal-oxide superconduc-
tors, Russian Chem. Rev. 69, 1–34 (2000).
42. L. Zhou, S.K. Chen, K.G. Wang, X.Z. Wu, P.X. Zhang, and Y. Feng, Synthesis of ultrafine Y2BaCuO5
powder and its incorporation into YBCO bulk by powder melting process, Physica C 363, 99–106
(2001).
43. S. Nariki, M. Matsui, N. Sakai, and M. Murakami, Refinement of RE211 particles in melt-textured
RE–Ba–Cu–O bulk superconductors, Supercond. Sci. Technol. 15, 679–682 (2002).
44. Y. Feng, A.K. Pradhan, Y. Zhao, S.K. Chen, Y. Wu, C.P. Zhang, G. Yan, J.K.F. Yau, L. Zhou, and
N. Koshizuka, Improved flux pinning in PMP YBa2Cu3Oy superconductors by submicron Y2BaCuO5
addition, Physica C 385, 363–367 (2003).

Copyright 2006 by Taylor & Francis Group, LLC


CRC_2308_Ch015.qxd 11/30/2005 3:56 PM Page 471

Nanostructured Oxide Superconductors 471

45. L. Zhou, S.K. Chen, K.G. Wang, X.Z. Wu, P.X. Zhang, Y. Feng, H.H. Wen, and S.L. Li, Preparation
of enhanced Jc YBCO bulks by powder melting process with a combination of submicron 211 precur-
sor and Pt addition, Physica C 371, 62–68 (2002).
46. T. Mochida, N. Sakai, S.-I. Yoo, and M. Murakami, Pinning properties in melt-processed
YbBa2Cu3O7−δ with finely dispersed Yb2BaCuO5 inclusions, Physica C 366, 229–237 (2002).
47. T. Mochida, N. Chikumoto, and M. Murakami, Flux pinning by Nd4Ba2Cu2O10 inclusions in
NdBa2Cu3O7−δ superconductors: a combined effect of point, interface, and pinning at elevated tem-
peratures, Phys. Rev. B 62, 1350–1360 (2000).
48. M.P. Delamare, H. Walter, B. Bringmann, A. Leenders, and H.C. Freyhardt, Macrosegregation of
Y2BaCuO5 particles in top-seeded melt textured monoliths, Physica C 323, 107–114 (1999).
49. J. Pötschke and V. Rogge, On the behaviour of foreign particles at an advancing solid–liquid interface,
J. Cryst. Growth 94, 726–738 (1989).
50. S. Nariki, N. Sakai, M. Murakami, and I. Hirabayashi, High critical current density in Y–Ba–Cu–O
bulk superconductors with very fine Y211 particles, Supercond. Sci. Technol. 17, 30–35 (2004).
51. P. Diko, L. Shlyk, and G. Krabbes, Ru assisted total 211 particle pushing in melt-grown Y123 super-
conductor, Physica C 390, 143–150 (2003).
52. T. Umemura, K. Egawa, S. Kinouchi, S. Utsunomiya, and M. Nojiri, Synthesis and superconducting
properties of BSCCO including precipitates with high density, Phase Transit. 42, 47–51 (1993).
53. T. Matsushita, T. Nakatani, E.S. Otabe, B. Ni, T. Umemura, K. Egava, S. Kinouchi, A. Nozaki, and
S. Utsunomiya, Irreversibility line superconducting Bi-2212 single grain with fine normal particles,
Cryogenics 33, 251–255 (1993).
54. M.R. Koblischka, S.G. Huang, K. Fossheim, T.H. Johansen, and H. Bratsberg, Evidence for pinning by
(Sr,Ca)2CuOy partial-melting processed bulk Bi2Sr2CaCu2O8δ ceramics, Physica C 300, 207–211 (1998).
55. X.L. Wang, J. Horvat, H.K. Liu, and S.X. Dou, Enhanced flux pinning from CuO inclusions in
Bi2Sr2CaCu2Oy crystals, J. Appl. Phys. 81, 533–535 (1997).
56. J.C.L. Chow, H.T. Leung, W. Lo, and D.A. Cardwell, Effects of Pt doping on the size distribution and
uniformity of Y2BaCuO5 particles in large-grain YBCO, Supercond. Sci. Technol. 11, 369–374 (1998).
57. M. Yoshida, N. Ogava, I. Hirabayashi, and S. Tanaka, Effects of the platinum group element addition
on preparation of Y–Ba–Cu–O superconductor by melt growth method, Physica C 185–189,
2409–2410 (1991).
58. K. Osamura, N. Matsukura, Y. Kusumoto, S. Ochiai, B. Ni, and T. Matsushita, Improvement of criti-
cal current density in YBa2Cu3O6x superconductor by Sn addition, Jpn. J. Appl. Phys. 29, 1621–1623
(1990).
59. K.V. Paulose, P. Murugaraj, J. Koshy, and A.D. Damodaran, YBa2SnO5.5: a new phase in
YBa2Cu3O7–SnO2 system, Jpn. J. Appl. Phys. 31, 1323–1325 (1992).
60. R. Weinstein and R.-P. Sawh, A class of chemical pinning centers including two elements foreign to
HTS, Physica C 383, 438–444 (2003).
61. R.-P. Sawh, R. Weinstein, D. Parks, A. Gandini, Y. Ren, and I. Rusakova, Tungsten and molybdenum
double perovskites as pinning centers in melt-textured Y123, Physica C 383, 411–416 (2003).
62. P.E. Kazin, V.V. Poltavets, V.V. Lennikov, R.A. Shuba, E.A. Eremina, Yu.D. Tretyakov, M. Jansen,
B. Freitag, G.F. de la Fuente, and A. Larrea, Formation of stable phase inclusions in Bi-2212 and
Bi(Pb)-2223 materials, in High-Temperature Superconductors and Novel Inorganic Materials. G. Van
Tendeloo, E.V. Antipov and S.N. Putilin, Eds., Kluwer Academic Publishers, Dordrecht, 1999, pp.
69–74.
63. P.E. Kazin, Superconducting phases Bi2Sr2CaCu2O8δ and Bi(Pb)2Sr2Ca2Cu3O10δ in multi-compo-
nent oxide systems, Zh. Neorg. Khim. 47, 703–711 (2002).
64. V.V. Lennikov, P.E. Kazin, V.I. Putlyaev, Yu.D. Tretyakov, and M. Jansen, Influence of magnesium
oxide on the properties of high temperature superconductor Bi2Sr2CaCu2O8x obtained by melt tech-
niques, Zh. Neorg. Khim. 41, 911–915 (1996).
65. P.E. Kazin, Y.D. Tretyakov, V.V. Lennikov, and M. Jansen, Formation of the Bi2Sr2CaCu2O8x super-
conductor with Mg1−xCuxO inclusions: the phases compatibility and the effect of the preparation route
on the material microstructure and properties, J. Mater. Chem. 11, 168–172 (2001).
66. W. Wei, Y. Sun, Y. Schwartz, K. Goretta, U. Balachandran, and A. Bhargava, Preparation and proper-
ties of nanosize TiO2 and MgO-doped Bi2Sr2CaCu2Ox tapes, IEEE Trans. Appl. Supercond. 7,
1556–1559 (1997).

Copyright 2006 by Taylor & Francis Group, LLC


CRC_2308_Ch015.qxd 11/30/2005 3:56 PM Page 472

472 Nanomaterials Handbook

67. W. Wei, J. Schwartz, K.C. Goretta, U. Balachandran, and A. Bhargava, Effects of nanosize MgO addi-
tions to bulk Bi2.1Sr1.7CaCu2Ox, Physica C 298, 279–288 (1998).
68. B. Zhao, W.H. Song, X.C. Wu, W.D. Huang, M.H. Pu, J.J. Du, Y.P. Sun, and H.C. Ku, Enhanced flux
pinning in Bi2212 single crystal embedded with MgO particles, Supercond. Sci. Technol. 13, 165–168
(2000).
69. P.E. Kazin, V.V. Poltavets, Y.D. Tretyakov, M. Jansen, B. Freitag, and W. Mader, Study on the super-
conducting composite material formation in the system Bi2Sr2CaCu2O8x/Al-containing phases,
Physica C 280, 253–266 (1997).
70. P.E. Kazin, V.V. Poltavets, Yu.D. Tretyakov, M. Jansen, B. Freitag, and W. Mader, Phase and
microstructure evolution in the process of the composite Bi-2212 — (Sr,Ca)3Al2O6 glass-ceramics for-
mation, Supercond. Sci. Technol. 12, 475–480 (1999).
71. P.E. Kazin, M. Jansen, A. Larrea, G.F. de la Fuente, and Yu.D. Tretyakov, Flux pinning improvement
in Bi-2212 silver sheathed tapes with submicron SrZrO3 inclusions, Physica C 253, 391–400 (1995).
72. P.E. Kazin, R.A. Shuba, Yu.D. Tretyakov, A.V. Knotko, M. Jansen, and B. Freitag, Formation of
Bi-2212-based composites with submicrometer-grained (Sr,Ca)SnO3, Supercond. Sci. Technol. 13,
134–139 (2000).
73. P.E. Kazin, V.V. Poltavets, M.S. Kuznetsov, D.D. Zaytsev, Yu.D. Tretyakov, M. Jansen, and M.
Schreyer, Phase compatibility and preparation of Bi-2212-Sr1−xCaxIn2O4 composite, Supercond. Sci.
Technol. 11, 880–886 (1998).
74. P.E. Kazin, A.A. Kovalevski, V.V. Poltavets, Yu.D. Tretyakov, and M. Jansen, Preparation of the
Bi2Sr2CaCu2O8x-matrix composites containing fine strontium calcium indate inclusions via glass
crystallization, Inorg. Mater. 37, 1183–1187 (2001).
75. P.E. Kazin, A.S. Karpov, M. Jansen, J. Nuss, and Yu.D. Tretyakov, Crystal structure and properties of
strontium phosphate apatite with oxocuprate ions in hexagonal channels, Z. Anorg. Allgem. Chem.
629, 344–352 (2003).
76. P.E. Kazin, A.S. Karpov, Yu.D. Tretyakov, and M. Jansen, Formation of SrSnO3 shell-like inclusions
in the Bi2Sr2CaCu2O8x superconductor via chemical reaction, Solid State Sci. 3, 285–290 (2001).
77. P.E. Kazin, A.S. Karpov, M.V. Makarova, A.A. Kovalevski, Yu.D. Tretyakov, and M. Jansen,
Formation of fine precipitates with different shapes in Bi-2212 material. VI Proceedings of the
International Workshop on High Temperature Superconductors and Novel Inorganic Materials
Engineering, June 24–30, 2001, Moscow - St.Petersburg, Book of Abstracts, PII-29.
78. P. Yang and C.M. Lieber, Nanorod-superconductor composites: a pathway to materials with high crit-
ical current densities, Science 273, 1936–1940 (1996).
79. L. Hua, G. Qiao, J. Yoo, J. Ko, H. Kim, and H. Chung, Microstructure and phase evolution of ultra-
fine MgO doped Bi-2223/Ag tapes, Physica C 291, 149–154 (1997).
80. X. Wan, Y. Sun, W. Song, L. Jiang, K. Wang, and J. Du, Enhancement flux pinning of silver-sheathed
(Bi,Pb)2Sr2Ca2Cu3Ox tapes with nano-MgO particle addition, Supercond. Sci. Technol. 11, 1079–1081
(1998).
81. W.D. Huang, W.H. Song, Z. Cui, B. Zhao, M.H. Pu, X.C. Wu, T. Hu, Y.P. Sun, and J.J. Du,
Enhancement of flux pinning in (Bi,Pb)-2223/Ag tapes doped with MgO nanorods, Supercond. Sci.
Technol. 13, 1499–1504 (2000).
82. P. Majewski, The use of phase diagrams for the engineering of flux pinning centres in Bi2Sr2CaCu2O8
ceramics, Appl. Supercond. 3, 289–301 (1995).
83. P. Majewski, F. Aldinger, and S. Elschner, Enhanced pinning by second-phase precipitates in Sr rich
“Bi2Sr2CaCu2O8” ceramics, Physica C 249, 234–240 (1995).
84. P. Majewski, S. Kaesche, F. Aldinger, S. Elschner, B. Hettich, and C. Lang, The increase of pinning in
(Bi,Pb)2Sr2Ca2Cu3O10d bulk ceramics, Supercond. Sci. Technol. 7, 514–517 (1994).
85. V. Putlayev, S. Sokolov, P. Kazin, A. Veresov, and Yu. Tretyakov, On the phase decomposition of
Bi2Sr2CaCu2O8, Solid State Ionics 101–103, 1075–1078 (1997).
86. A.V. Knotko, A.V. Garshev, M.N. Pulkin, and V.I. Putlyaev, Solid solution homogeneity area on the
base of Bi2Sr2CaCu2O8 with alkaline-earth elements by Nd and La substitution, Zh. Neorg. Khim. 46,
1364–1367 (2001).
87. A.V. Knotko, A.V. Garshev, M.N. Pulkin, V.I. Putlyaev, and S.I. Morozov, Interconnection between
oxygen atoms and the kinetics of solid solutions oxidation on the base of Bi2Sr2CaCu2O8, Fizika
Tverdogo Tela 46, 414–418 (2004).

Copyright 2006 by Taylor & Francis Group, LLC


CRC_2308_Ch015.qxd 11/30/2005 3:56 PM Page 473

Nanostructured Oxide Superconductors 473

88. J. Shimoyama, Y. Nakayama, K. Kitazawa, K. Kishio, Z. Hiroi, I. Chong, and M. Takano, Strong flux
pinning up to liquid nitrogen temperature discovered in heavily Pb-doped and oxygen controlled
Bi2212 single crystals, Physica C 281, 69–75 (1997).
89. I. Chong, Z. Hiroi, M. Izumi, J. Shimoyama, Y. Nakayama, K. Kishio, T. Terashima, Y. Bando, and M.
Takano, High critical-current density in the heavily Pb-doped Bi2Sr2CaCu2O8δ superconductor: gen-
eration of efficient pinning centers, Science 276, 770–773 (1997).
90. M.R. Koblischka, A.J.J. van Dalen, T. Higuchi, S.I. Yoo, and M. Murakami, Analysis of pinning in
NdBa2Cu3O7−δ superconductors, Phys. Rev. B 58, 2863–2867 (1998).
91. T. Higuchi, S.I. Yoo, and M. Murakami, Comparative study of critical current densities and flux pin-
ning among a flux-grown NdBa2Cu3Oy single crystal, melt–textured Nd–Ba–Cu–O, and Y–Ba–Cu–O
bulks, Phys. Rev. B 59, 1514–1527 (1999).
92. M. Muralidhar, M. Jirsa, N. Sakai, and M. Murakami, Matrix chemical ratio and its optimization for
highest flux pinning in ternary (Nd–Eu–Gd)Ba2Cu3Oy, Supercond. Sci. Technol. 15, 688–693 (2002).
93. M. Muralidhar, M. Jirsa, N. Sakai, and M. Murakami, Progress in melt-processed (Nd–Sm–Gd)Ba2Cu3Oy
superconductors, Supercond. Sci. Technol. 16, R1–R16 (2003).

Copyright 2006 by Taylor & Francis Group, LLC

Vous aimerez peut-être aussi