You are on page 1of 7

Physica C 470 (2010) 421–427

Contents lists available at ScienceDirect

Physica C
journal homepage: www.elsevier.com/locate/physc

Doping dependence of irreversibility line in Y1xCaxBa2Cu3O7d


E. Nazarova a,*, A. Zaleski b, K. Buchkov a
a
Georgy Nadjakov Institute of Solid State Physics, Bulgarian Academy of Sciences, 72 Tzarigradsko Chaussee Blvd., 1784 Sofa, Bulgaria
b
Institute of Low Temperature and Structure Research, Polish Academy of Sciences, 2 Okolna Str., 50-950 Wroclaw, Poland

a r t i c l e i n f o a b s t r a c t

Article history: The irreversibility lines (IL) for series of Y1xCaxBa2Cu3O7d (x = 0.025; 0.10 and 0.20) polycrystalline sam-
Received 10 August 2009 ples with different overdoping were investigated. The irreversibility fields were determined from mea-
Received in revised form 27 February 2010 surements of third harmonics AC susceptibility as a function of DC field at constant temperature. For
Accepted 7 March 2010
the weakly overdoped sample (with x = 0.025) Hirr(77 K) is about 7 T, which is higher than the previously
Available online 10 March 2010
reported for the non-substituted one. The irreversibility line behavior is typical for glass–liquid phase
transition and this is confirmed by transport measurements. On increasing the overdoping the irrevers-
Keywords:
ibility fields were shifted towards lower temperatures. The behavior of Hirr(T) for the highly overdoped
Irreversibility line
Third harmonics AC susceptibility
sample (with x = 0.20) is influenced by the surface barrier effect. It is supposed that in highly overdoped
Overdoped Y(Ca)BCO specimen the process of phase separation is enhanced and the Fermi clusters grow in size. This leads to a
Surface barrier suppression of the bulk pinning and to a domination of the surface barrier effects and flux creep as well.
Phase separation As a confirmation, the obtained quadratic Jc(T) dependences were presented demonstrating the existence
of S–N–S type inter-grain joints in the highly overdoped samples.
Ó 2010 Elsevier B.V. All rights reserved.

1. Introduction ics signal which is very sensitive method for their separation [7–9].
The AC susceptibility signal could be investigated as a function of
According to the Bean critical state model the hysteretic non- temperature, DC magnetic field intensity, AC magnetic field ampli-
linear relationship between the magnetization and the external tude and frequency. By variation of these parameters different dy-
field in type II superconductors emerges as a result of flux pinning. namic regimes could be studied. For their proper identification the
This brings about higher harmonics generation of AC magnetic sus- experimental results are compared to susceptibility curves simu-
ceptibility. However, the AC magnetic response is frequency lated by numerical calculations of the non-linear diffusion equa-
dependent, which is not accounted for in the model. In its extended tion for the magnetic field [10]. The observations of high
version, dynamic losses related to the thermally activated vortex harmonics signals (vn) imply the non-linear response of the flux
motion are included [1,2]. The vortex motion generates a resistive system. Non-linearity is not always a precondition to irreversibility
state described by Ohmic dependence E = qJ, where q(B, T) is inde- (for example, non-linear flux flow). The irreversibility in the mag-
pendent of J. This is the so-called linear response of type II super- netization behavior can be a result of bulk pinning, surface or edge
conductors to an AC magnetic field. It is connected to the thermally barrier effects and contributes to the third harmonics (v3) signal
assisted flux flow (TAFF) regime, characterized by thermally acti- appearance. At certain conditions (superposition of DC and AC
vated vortex hoping and flux flow (FF) regime – dominated by vis- magnetic fields when Hac  Hdc) the flux creep is non-linear and
cous motion of the vortex liquid. In fact, FF regime could also be the irreversible dynamical regime determines the v3 signal [7,9].
non-linear at high vortex velocities due to the non-equilibrium dis- In this case the third harmonics signal is closely related to the flux
tribution of quasiparticles in the vortex core and to the decrease of depinning (irreversibility) line and is often used for its determina-
viscosity [3–6]. The sample’s resistivity is expressed in J-dependent tion [11–13].
form when the vortex motion is governed by the thermally acti- The conventional superconductors have two well-known
vated flux creep (FC). The E–J characteristic has a non-linear behav- phases in the H–T plane: mixed and Meissner phase. In high tem-
ior and the higher harmonics generation is observed. In general, perature superconductors (HTSC) many different phases have been
the AC susceptibility technique is a useful tool for the investigation predicted and established to exist [14–17]. The IL distinguishes a
of different flux-dynamic regimes and especially the third harmon- pinned vortex solid where the flux creep regime occurs from un-
pinned vortex liquid characterized by TAFF regime [18]. The IL
was established for the first discovered HTSC (Ba–La–Cu–O) [19]
* Corresponding author. Tel.: +359 2 7144211; fax: +359 2 9753632. and was extensively studied for all other families of HTSC
E-mail address: nazarova@issp.bas.bg (E. Nazarova). compounds [20–22].

0921-4534/$ - see front matter Ó 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.physc.2010.03.002
422 E. Nazarova et al. / Physica C 470 (2010) 421–427

The irreversibility line, for underdoped YBCO samples with re- 3.8689–3.8769 Å; c = 11.6421–11.6468 Å. Only the sample with
duced oxygen content, was investigated by many authors [23– x = 0.025 has larger c = 11.6519 Å. Very close values for c lattice
26]. The observed decrease of irreversibility field (Hirr) in oxygen parameter presume similar oxygen content. Using the correlation
deficient samples is explained by the reduction of condensation en- between the c lattice parameter and oxygen content [31–33], we
ergy and the presence of a pseudo gap. Few articles are published estimated that the latter is higher than 6.8 for Y0.8Ca0.2Ba2Cu3O7d
where the irreversibility lines for overdoped and optimally doped and about 6.9 for Y0.9Ca0.1Ba2Cu3O7d. There are no data available
YBCO system are investigated. It has been established that a strong for the sample Y0.975Ca0.025Ba2Cu3O7d, but comparing it with the
correlation exists between irreversibility line and interlayer cou- non-substituted YBCO its oxygen content might be close to 7.
pling and anisotropy [27,28], oxidation state of CuO chains The lattice parameters of the specimen treated in nitrogen atmo-
[28,29], electronic and structural characteristics [21,28,30]. In sin- sphere are a = 3.8210 Å; b = 3.8742 Å and c = 11.6708 Å, respec-
gle crystals and polycrystalline samples Hirr is found to decrease tively. Its oxygen content is estimated to be 6.6 [34]. On
when a significant amount of Ca substitution is present. accepting Tcmax = 93.5 K for YBa2Cu3O7d and a Tc reduction by
In the present study we used the third harmonics magnetic sus- 0.35 K/at.% Ca, we evaluate the Tcmax values for the substituted
ceptibility measurements for IL investigations of a series of samples. The coincidence of Tcmax for the sample with x = 0.20 with
Y1xCaxBa2Cu3O7d (x = 0.025; 0.10 and 0.20) samples with differ- the experimentally observed data [35] confirms that the obtained
ent extent of overdoping. Tcmax values are realistic. The values of the Tconset are determined
from v01 ðTÞ dependences (Fig. 1) and presented in Table 1.
The carrier concentration p in the CuO2 plane was estimated
2. Experimental details
from the relation [36].
We investigated a series of polycrystalline samples of type T c =T cmax ¼ 1  82:6ðp  0:16Þ2 ð1Þ
Y1xCaxBa2Cu3O7d (x = 0.025; 0.10; 0.20) oxygenated for a long
time (48 h) at 450 °C in order to make them overdoped. One of and the obtained data are listed in Table 1. The results show that the
these samples (with x = 0.2) was additionally heat treated at nitro- overdoping increases by the Ca content: it is lowest in the sample
gen atmosphere in order to prepare the underdoped sample for with x = 0.025 and highest in sample with x = 0.20. The nitrogen
comparison. To this end a non-substituted YBCO sample and treated specimen is underdoped with p = 0.135.
Y0.8Ca0.2Ba2Cu3O7d oxygenated for 100 h was also prepared. Stan-
dard X-ray powder diffraction analysis with Cu ka radiation was 3.2. AC susceptibility measurements
used for the examination of the purity and the material structure.
Magnetic measurements were performed on the commercial v01 ðTÞ dependences presented in Fig. 1 show double step transi-
MagLab-Oxford 7000 susceptometer. AC magnetic susceptibility tion for all investigated samples. This is characteristic for the gran-
measurements were performed in parallel configuration with AC ular high-temperature superconducting materials. The first step is
magnetic field amplitude 0.1 Oe and frequency 1000 Hz and DC connected with the screening of the grain itself, while the second
magnetic field in the range 0–9 T. In order to obtain the proper sep- transition at lower temperatures is due to the establishing of in-
aration of real and imaginary parts of all harmonics, a special care ter-grain connections and diamagnetism in the whole sample.
was taken to properly adjust the phase for the measurements. It
was chosen in this way so that the imaginary part of the first order
susceptibility component was zeroed for the lowest obtained tem-
perature before every measurement. In the experiments carried
out for the Jc (T) determination the AC field amplitude was varied
from 1 Oe up to 15 Oe at frequency 1000 Hz. The samples’ dimen-
sions used for these measurements were 9.5  2.4  1.6 mm3 for
the Y0.8Ca0.2Ba2Cu3O7d oxygenated for 48 h, and 10.7  2.7 
1.9 mm3 for the Y0.8Ca0.2Ba2Cu3O7d oxygenated for 100 h.
The M(H) measurements were carried out by the extraction
method. To increase the sensitivity of the measurements 10 repe-
titions were used for rather slow sample movement.
The transport measurements were performed on Quantum De-
sign PPMS. In order to prevent the sample from Joule heating effect
the DC current was applied for a very short time 0.002 s. The
thick current leads have been used, soldered to the sample’s sur-
face on the big spots. The boundary values for I, V and power have
been specified and the measurement was impossible if one of these
values was exceeded. The applied magnetic field was perpendicu-
lar to the current direction. Sample dimensions were S = 0.116  Fig. 1. Temperature dependence of the normalized real part of fundamental AC
0.211 cm2, L = 0.420 cm for sample with x = 0.025 and S = magnetic susceptibility for all investigated samples at Hac = 0.1 Oe and f = 1000 Hz.
0.184  0.283 cm2, L = 0.340 cm for sample with x = 0.20, where L
is the distance between the voltage leads.
Table 1
Onset critical temperature Tconset, critical transition temperature Tcmax, carrier
concentration p and estimated ‘n’ value for all samples.
3. Experimental results
Sample Y1xCaxBa2Cu3O7d Tconset [K] Tcmax [K] p n
3.1. Samples’ characterization (x value)
0.025 90.6 92.6 0.176 2.23 ± 0.08
The X-ray analysis showed that all studied samples were single 0.10 85.3 90.0 0.185 3.96 ± 0.05
0.20 – 48 h 81.1 86.5 0.188 6.36 ± 0.15
phase. The oxygenated specimens have similar unit cell lattice
0.20 (N2 atmosphere) 82.1 85.5 0.135 2.15 ± 0.07
parameters in the following limits: a = 3.8126–3.8130 Å; b =
E. Nazarova et al. / Physica C 470 (2010) 421–427 423

Fig. 2. Temperature dependences of the imaginary part of fundamental AC Fig. 3. Estimated critical current density, Jc, versus reduced temperature T/Tc for
magnetic susceptibility for a sample with x = 0.20 (48 h oxygenated) at different overdoped samples with x = 0.20 (48 h and 100 h oxygenated). The best quadratic
Hac amplitudes and f = 1000 Hz. The inset shows also the v01 (for Hac = 1 Oe and fits to the data are indications for the presence of S–N–S type junctions in these
10 Oe) in order to identify that the v001 peaks are of intergranular origin. samples.

The superconducting transition is shifted to the lower temperature Thus for a given magnetic field the position of IL for the highly
on increasing the overdoping. This is in accord with the overdoped overdoped sample (x = 0.20) is by about 20 K lower than
side of the T(p) phase diagram, where the increasing of the carrier that for the sample with x = 0.025. The irreversibility field for
concentration leads to a suppression of Tc [36]. Y0.975Ca0.025Ba2Cu3O7d at 77 K is estimated to be about 7 T. This
At small AC field amplitudes (Hac) there is no, flux penetration value is higher than the reported for YBCO polycrystalline sample,
and, correspondingly, no AC losses inside the grains. Only one peak which is about 5 T [40]. Usually, IL is fitted by the dependence:
is observed in the v001 ðTÞ dependence and it is attributed to the
Birr ¼ Birr ð0Þð1  T=T c Þn ð3Þ
losses generated in the intergranular region. In Fig. 2 series of
v001 ðTÞ dependences for highly overdoped sample (with x = 0.20; predicted by the collective pinning theory [41], where ‘n’ is a model
48 h oxygenated) are presented at different AC field amplitudes dependent parameter. The assumption for a 3D–2D flux-line transi-
ranging from 1 to 15 Oe and frequency f = 1000 Hz. The inset tion due to the decoupling of conducting layers leads to n = 3/2 [42],
shows that the peak position corresponds to the intergranular while flux-line melting model using a Lindemann type melting cri-
transition. The maximum of v001 ðTÞ is reached when the applied terion predicts n = 2 [43]. The double logarithmic plots of the irre-
magnetic field penetrates to the center of the sample. The increas- versibility field versus [1  (T/Tc)] are presented for all samples in
ing of the magnetic field amplitude shifts the maximum position Fig. 6. The slope of the lines, which gives the values of ‘n’, grows
towards lower temperatures in conformity with [37]. This is used on increasing the overdoping. The estimated values of ‘n’ and their
for the determination of the Jc (T) dependence when the sample’s error range are given in Table 1 for all samples. The values of ‘n’ for
shape is known. The sample has been approximated with a long Y0.975Ca0.025Ba2Cu3O7d and nitrogen treated samples are close to 2
cylinder (l = 10R, where R is the radius and l is the length of the cyl- which might indicate the existence of glass–liquid phase transition.
inder). The exact sample’s volume V = 10.7  2.7  1.9 mm3 is used The ‘n’ value increases when the overdoping grows and the values
for the determination of R according to the formula V = pR2(10R). for the other two samples (with x = 0.10 and x = 0.20) are large
The critical current density at Tp has been calculated by using the within such an interpretation. In fact, using only the ‘n’ value, it is
relation [38]: impossible to identify what kind of irreversibility is determinant
in different samples. Additional investigations have been performed
J c ðTÞ ¼ Hac =R ð2Þ
for that purpose.
The obtained in this manner Jc (T) dependences for two highly Temperature dependences of first harmonics susceptibility are
overdoped specimens are presented in Fig. 3 and well approxi- similar for all investigated samples as it is seen from Fig. 1. On
mated (with small mean square deviation) with quadratic fit. Sim- the contrary, a significant difference is observed in the temperature
ilar behavior was already observed from direct Jc(T) measurements dependences of third harmonics signals when the overdoping in-
in series of Ca substituted GdBa2Cu3Oz samples [39]. creases. In Fig. 7a–c v3(T) curves are presented for the overdoped
For the determination of IL we used the module of third har- samples (with x = 0.025, x = 0.10 and x = 0.20, respectively). For
monics signal (|v3|) as a function of a DC magnetic field (Hdc) at Y0.975Ca0.025Ba2Cu3O7d sample a positive peak (at temperatures
fixed temperature and small AC magnetic field amplitude near Tc) followed by a negative one (at lower temperatures) is ob-
(Hac < Hdc). In Fig. 4a–d |v3| versus Hdc dependences are presented served for the v03 ðTÞ, while the v003 ðTÞ shows a positive peak that
for all investigated samples (with x = 0.025, 0.10, 0.20 and the goes monotonically to zero when the temperature decreases. For
nitrogen treated one). The criterion used to determine the Hirr va- sample Y0.9Ca0.1Ba2Cu3O7d the positive peak in v03 ðTÞ is sup-
lue is the deviation of |v3| from the noise signal at fixed tempera- pressed and the negative depression becomes deeper, while v003 ðTÞ
ture. The signal which is two orders of magnitude smaller than that shows a positive peak again. For Y0.8Ca0.2Ba2Cu3O7d specimen –
in the peak position is referred to as noise signal (see inset of deep minimum is detected close to Tc, followed by a small positive
Fig. 4b). In this manner the value where the |v3| signal starts to in- peak at lower temperatures in v03 ðTÞ. v003 ðTÞ shows small positive
crease is determined within 1% error. This responds to no more peak close to Tc and a deep minimum at lower temperatures. In or-
than 0.1 T error when the Hirr is determined. der to verify the last observations the measurements were re-
The obtained irreversibility lines are presented in Fig. 5. On peated on a Quantum Design PPMS and the result was the same.
increasing the overdoping they are shifted to lower temperatures. In Fig. 7c inset the v03 ðTÞ and v003 ðTÞ dependences are shown ob-
424 E. Nazarova et al. / Physica C 470 (2010) 421–427

a b

c d

Fig. 4. Module of third harmonics AC magnetic susceptibility as a function of DC field at different temperatures (Hac = 0.1 Oe and f = 1000 Hz) for the sample with x = 0.025
(a), for the sample with x = 0.10 (b), for the sample with x = 0.20 (c), for the sample with x = 0.20 treated in nitrogen atmosphere (d). The inset at (b) shows the noise level two
orders of magnitude smaller than the signal at peak position.

Fig. 5. Plot of the irreversibility fields against the temperature (irreversibility line) Fig. 6. The double logarithmic plots of irreversibility field versus [1  (T/Tc)] for all
for all samples. The 77 K boundary is explicitly indicated. samples. The slope of the line gives the ‘n’ value for each sample.

tained from these measurements. One possible explanation of the terms of the Bean–Livingston surface barrier theory [44]. The
observed behavior of third harmonics signal versus temperature established sign reversal for third harmonics AC susceptibility is
for the highly overdoped sample (x = 0.20) can be offered in the predicted by Maksimova and Vodolazov [45] when the irreversibil-
E. Nazarova et al. / Physica C 470 (2010) 421–427 425

3.3. DC magnetic measurements


a
In Fig. 8 M(H) dependences, measured at T = 4.2 K, are pre-
sented. For the purpose of clarity the M(H) curves only for
Y0.975Ca0.025Ba2Cu3O7d and Y0.8Ca0.2Ba2Cu3O7d samples are pre-
sented. In order to make a comparison the hysteresis curve for
polycrystalline non-substituted YBCO sample is also presented.
The lightly overdoped sample (x = 0.025) exhibits a broad, almost
symmetric with respect to M = 0 curve. The highly overdoped sam-
ple (x = 0.20, oxygenated for 100 h) exhibits a significant reduction
of the hysteresis loop width as compared to the substituted
(x = 0.025) and non-substituted samples. The other important
observation is the almost zero magnetization of the descending
branch of M(H) curve for the Y0.8Ca0.2Ba2Cu3O7d sample. This is
a telling demonstration of the presence of Bean–Livingston surface
barrier [47,48]. As a result of the competition between the attrac-
tive force acting among the vortex and its mirrored antivortex and
b repulsive force between the vortex at the surface and the shielding
currents – a surface barrier appears. It prevents the vortex penetra-
tion in the sample up to field Hp (identical to the thermodynamic
critical field Hc for the smooth surface) and does not prevent their
exit. Surface imperfections allow partial penetration thus reducing
the penetration field down to Hc1 [49]. In our case the highly overd-
oped sample Y0.8Ca0.2Ba2Cu3O7d (oxygenated for 100 h) shows
M(H) curve determined by surface barrier effects. It should be
mentioned that v3(T) measurements are more sensitive than the
magnetization, indicating surface barrier effect in less overdoped
Y0.8Ca0.2Ba2Cu3O7d sample, which was oxygenated for 48 h.
In order to determine Hp(T) for the overdoped sample Y0.8Ca0.2-
Ba2Cu3O7d (oxygenated for 48 h) the offset of M(H) curves at dif-
ferent temperatures (4 K, 10 K, 20 K, 30 K, 40 K, 50 K and 60 K)
were recorded. It has been established that penetration field
monotonically decreases when the temperature increases
c (600 Oe, 535 Oe, 398 Oe, 270 Oe, 188 Oe, 155 Oe and 70 Oe respec-
tively). According to Reissner [49].

Hp ðTÞ ¼ Hc ð0Þ½ð1  t 2 Þð1 þ t2 Þ1=2  expð2T=T 0 Þ ð4Þ

where Hc(0) is thermodynamic critical field, t = T/Tc and T0 is a char-


acteristic temperature. By plotting ln[Hp/(1  t2)(1 + t2)1/2] versus T
both Hc(0) and T0 were determined to be 810 Oe and 10.44 K
respectively. The irreversibility field is determined by

Hirr  Hc2 ðT 0 =2TÞ expð2T=T 0 Þ ð5Þ

in the case of vortex penetration through a surface barrier at T > T0


[50]. The experimental data for Hirr were collected at T > T0. In Fig. 9
these data are presented and fitted using the above-mentioned

Fig. 7. (a) Temperature dependence of the real and imaginary parts of third
harmonics susceptibilities at Hac = 0.1 Oe, and f = 1000 Hz for sample with x = 0.025
(a), for sample with x = 0.10 (b) and for the sample with x = 0.20 (c). Inset are shown
the same dependences at the same conditions measured on PPMS-9 (Quantum
Design).

ity mechanism is changed from bulk pinning to edge barrier in thin


film. The calculated third harmonics susceptibility in a flat type II
superconductor for various barrier heights is presented by Zhang
et all. [46]. It was shown that increasing of the barrier suppressed
the positive peaks (in v03 ðTÞ and v003 ðTÞ) while the negative ones (in
v03 ðTÞ) became deeper. In the case when the bulk pinning decreases
or completely vanishes the irreversible behavior is dominated by
surface barrier effects. Similar behavior is observed in the highly
overdoped sample (Fig. 5c). Obviously, the bulk pinning is sup-
Fig. 8. Magnetization hysteresis loops for the samples with x = (0; 0.025 and 0.20 –
pressed in this specimen and this was confirmed by M(H) measure- 100 h oxygenated) at 4.2 K. In the inset is shown linear fit to the initial curve slope
ments too. for the sample with x = 0.
426 E. Nazarova et al. / Physica C 470 (2010) 421–427

Fig. 9. Plot of the irreversibility fields versus temperature for the sample with
x = 0.2 (48 h oxygenated) fitted by power law Hirr  Hc2(T0/2T)exp(2T/T0), where Fig. 11. I–V curves at given constant temperatures for sample with x = 0.20 and
T0 = 10.44 K and Hc2 is fitting parameter. H = 1.25 T. The field value is determined from the first point of IL (1.25 T; 65 K) for
this sample.
dependence. This is another consideration supporting the assump-
tion that the experimental results for the overdoped Y0.8Ca0.2Ba2-
Cu3O7d sample are influenced by the surface barrier effects. As For the Y0.8Ca0.2Ba2Cu3O7d specimen we found that the best linear
the specimen surfaces have similar quality, the appearance of the fit with the smallest square deviation is reached for the I–V depen-
surface barrier effect is associated with the overdoping of the dence at T = 64 K. For higher temperatures the vortex liquid phase
sample. appeared. This result coincides very well with the previously deter-
mined IL and suggests that the flux creep contribution cannot be
excluded from the |v3(Hdc)| signal. Probably the decisive contribu-
3.4. I–V characteristics tion comes from the surface barrier effects in accordance with DC
measurements.
The evidence for a presence of vortex-glass phase in HTSC
comes from transport experiments [51,52]. It has been shown
[51] that at given temperature T = Tg a power-law described I–V 4. Discussion
curve. At T < Tg a negative curvature on log I versus log V plot ap-
peared indicating the presence of vortex-glass state. On increasing From detailed examinations of the hysteresis curves (Fig. 8),
the temperature at T > Tg the typical for vortex creep phenomena several important facts have to be outlined. Sample Y0.975Ca0.025
positive curvature is observed. Ba2Cu3Oz has the largest M(H) curve and the highest trapped mag-
In order to verify the existence of vortex-glass state in investi- netic field, as a result of the strong bulk pinning (A). A linear M(H)
gated samples the I–V characteristics have been measured at fixed dependence is observed for substituted specimens with x = 0.025
magnetic field and at different temperatures. For these measure- and for non-substituted one as well (B). For the highly overdoped
ments it is very important to keep constant temperature and to sample Y1xCaxBa2Cu3Oz – (x = 0.2), the usual linear slope is miss-
protect sample from Joule heating. In Figs. 10 and 11 log I–log V ing, thus indicating that Hc1 is almost zero (C). For Y0.8Ca0.2Ba2-
curves are presented for the Y0.975Ca0.025Ba2Cu3Oz and Y0.8Ca0.2Ba2- Cu3Oz – (100 h), the penetration peak is the smallest and
Cu3O7d samples at H = 0.1 T and 1.25 T, respectively, and different coincides with the central peak, as seen in the inset (D). The last
temperatures. It is seen from these figures that the negative curva- two facts suggest that the smallest critical current density in the
tures, characteristic for vortex-glass state, exist in both samples. Y0.8Ca0.2Ba2Cu3Oz sample might be a result of percolative super-
conductivity. This could be the result of an inhomogeneous charge
distribution on the nanoscale (clusters) and sub-nanoscale
(stripes) levels [53]. In the overdoped region (p > 0.19) new doped
holes assemble between stripes forming a sea of fermions. In HTSC,
due to the short coherence length, the energy cost for a creation of
phase boundaries is small, which facilitates phase separation [54].
When p increases the Fermi clusters grow in size. We can speculate
that in the case of slight overdoping, the normal regions are small
and their dimensions probably are comparable to the coherence
length n. As a result, a broad M(H) curve is observed due to the
effective pinning. The volume of Fermi regions increases in size
when the doping considerably exceeds the optimal concentration.
They are not able to pin the flux vortices any more. The intragran-
ular critical current is significantly suppressed because of the car-
rier and pinning reduction. In fact, the critical current is a
parameter highly dependent upon the material processing. How-
ever, the experimental results show that the quality of the super-
conducting condensate is also strongly influenced by overdoping.
Fig. 10. I–V curves at given constant temperatures for sample with x = 0.025 and The Jc versus T/Tc dependences presented in Fig. 3 for highly
H = 0.1 T. overdoped samples (with x = 0.20, 48 h and 100 h oxygenated)
E. Nazarova et al. / Physica C 470 (2010) 421–427 427

could be one confirmation of the above assumptions. These depen- [11] D. Di Gioacchino, A. Mrcelli, S. Zhang, M. Fratini, N. Poccia, A. Ricci, A. Bianconi,
cod-math-0905-1633.
dences are well approximated with quadratic fit, which is an indi-
[12] J. Deak, M. McElfresh, J.R. Clem, Z. Hao, M. Konczykowski, R. Muenchausen, S.
cation of the existence of S–N–S type joints between the grains in Foltyn, R. Dye, Phys. Rev. B 49 (1994) 6270.
these polycrystalline samples. Similar behavior is well-known for [13] K. Waki, T. Higuchi, S.I. Yoo, M. Murakami, Appl. Supercond. 5 (1997) 133.
superconducting samples with various metallic additions (Ag, Pt) [14] G. Blatter, M.V. Feigel’man, V.B. Geshkenbein, A.I. Larkin, V.M. Vinokur, Rev.
Mod. Phys. 66 (1994) 1125.
[55]. Using this analogy we conclude that normal carriers and/or [15] D.S. Fisher, M.P.A. Fisher, D.A. Huse, Phys. Rev. B 43 (1991) 130.
small regions of normal state are present at grain boundaries, con- [16] T. Klein, I. Joumard, S. Blanchard, J. Marcus, R. Cubitt, T. Giamarchi, P. Le
firming the idea of phase separation in overdoped samples. Doussal, Nature 413 (2001) 404.
[17] U. Divakar, A.J. Drew, S.L. Lee, R. Gilardi, J. Mesot, F.Y. Ogrin, D. Charalambous,
E.M. Forgan, G.I. Menon, N. Momono, M. Oda, C.D. Dewhurst, C. Baines, Phys.
5. Conclusions Rev. Lett. 92 (2004) 237004.
[18] M. Nikolo, Supercond. Sci. Technol. 6 (1993) 618.
[19] K.A. Muller, M. Takashige, J.G. Bednorz, Phys. Rev. Lett. 58 (1987) 1143.
In conclusion, we have demonstrated that 2.5% Ca substitution [20] A. Schilling, H.R. Ott, Th. Wolf, Phys. Rev. B 46 (1992) 14253.
in polycrystalline YBCO sample enhances the irreversibility field up [21] G.V.M. Williams, J.L. Tallon, D.M. Pooke, Phys. Rev. B 62 (2000) 9132.
[22] P. Chowdhury, Heon-Jung Kim, In-Sun Jo, Sung-Ik Lee, Physica C 384 (2003)
to 7 T at 77 K and makes the hysteresis loop (at 4 K) broader than
411.
that for non-substituted YBCO. Further increasing of Ca content [23] W. Bauhofer, W. Biberacher, B. Gegenheimer, W. Joss, R.K. Kremer, Hj.
deteriorates bulk pinning, and observed irreversibility is domi- Mattaussh, A. Muller, A. Simon, Phys. Rev. Lett. 63 (1989) 2520.
nated by surface barrier effects. As a result irreversibility lines [24] J.G. Ossandon, J.R. Tompson, D.K. Christen, B.C. Sales, H.R. Kerchner, J.O.
Thomson, Y.R. Sun, K.W. Lay, J.E. Tkaczyk, Phys. Rev. B 45 (1992) 12534.
are shifted to lower temperatures. We suppose that on increasing [25] H.B. Sun, K.N.R. Taylor, G.J. Russell, Physica C 227 (1994) 55.
the overdoping phase separation takes place. This assumption is [26] D. Babic, J.R. Cooper, J.W. Hodby, Ch. Changkang, Phys. Rev. B 60 (1999) 698.
supported by the Jc(T) dependence deduced from the AC suscepti- [27] J.L. Tallon, G.V.M. Williams, C. Bernhard, D.M. Pooke, M.P. Staines, J.D. Johnson,
R.H. Meinhold, Phys. Rev. B 53 (1996) R11972.
bility measurements as well as from previous direct measurements [28] T. Nakane, M. Karppinen, H. Yamauchi, Supercond. Sci. Technol. 13 (2000)
[39]. 1436.
[29] T. Masui, E. Ohmichi, S. Tajima, T. Osada, Physica C 426–431 (2005) 335.
[30] A. Semwal, N.M. Strickland, A. Bubendorfer, S.H. Naqib, S.K. Goh, G.V.M.
Acknowledgements Williams, Supercond. Sci. Technol. 17 (2004) 8506.
[31] A. Sedky, A. Gupta, V.P.S. Awana, A.V. Narlikar, Phys. Rev. B 58 (1998) 12495.
The authors are very grateful to K. Nenkov (IFW, Leibniz Insti- [32] V.P.S. Awana, A.V. Narlikar, Phys. Rev. B 49 (1994) 6353.
[33] A. Manthiram, S.-J. Lee, J.B. Goodenough, J. Solid State Chem. 73 (1988) 278.
tute for Solid State and Materials Research, Dresden, Germany)
[34] J. Hejtmanek, Z. Jirak, K. Knizek, M. Dlouha, S. Vratislav, Phys. Rev. B 54 (1996)
for the measurement of V–I characteristics and for useful 16226.
discussions. [35] K. Hatada, H. Shimizu, Physica C 304 (1998) 89.
[36] J.L. Tallon, C. Bernhard, H. Shaked, R.L. Hitterman, J.D. Jorgensen, Phys. Rev. B
This work was supported under the Euratom Project FU07-CT-
51 (1995) 12911.
2007-00059. One of the authors (E.N.) is grateful to the Institute [37] K.H. Muller, Physica C 168 (1990) 585.
of Low Temperature and Structure Research, Polish Academy of [38] I. Dhingra, B.K. Das, Supercond. Sci. Technol. 6 (1993) 765.
Sciences, Wroclaw, for the hospitality and useful cooperation. [39] E. Nazarova, K. Nenkov, G. Fuchs, K.-H. Muller, Physica C 436 (2006) 25.
[40] T. Nakane, M. Karppinen, H. Yamauchi, Physica C 375–360 (2001) 226.
[41] A.I. Larkin, Yu.N. Ovchinnikov, J. Low Temp. Phys. 34 (1979) 409.
References [42] Y. Yeshurun, A.P. Malozemoff, Phys. Rev. Lett. 60 (1988) 2202.
[43] A. Houghton, R.A. Pelcovits, A. Sudbo, Phys. Rev. B 40 (1989) 6763.
[1] P. Fabbricatore, S. Farinon, G. Gemme, R. Musenich, R. Parodi, B. Zhang, Phys. [44] C.P. Bean, J.D. Livingston, Phys. Rev. Lett. 12 (1964) 14.
Rev. B 50 (1994) 3189. [45] G.M. Maksimova, D.Yu. Vodolazov, Physica C 356 (2001) 67.
[2] Y.J. Zhang, M.J. Qin, C.K. Ong, Physica C 351 (2001) 395. [46] Y.J. Zhang, C.K. Ong, M.J. Qin, X.F. Sun, Physica C 350 (2001) 97.
[3] D.E. Farrell, I. Dinewitz, B.S. Chandrasekhar, Phys. Rev. Lett. 16 (1966) 91. [47] A.M. Campell, J.E. Evetts, Critical Currents in Superconductors, Taylor &
[4] F. Lefloch, C. Hoffmann, O. Demolliens, Physica C 319 (1999) 258. Francis, London, 1972, p.142.
[5] S. Pace, G. Filatrella, G. Grimadi, A. Nigro, M.D. Adesso, AIP Conf. Proc. 850 [48] M. Konczykowski, L.I. Burlachkov, Y. Yeshurun, F. Holtzberg, Phys. Rev. B 43
(2006) 873. (1991) 13707.
[6] R.P. Huebener, J. Phys. Condens. Matter 21 (2009) 254208. [49] M. Reissner, Physica C 290 (1997) 173.
[7] C. Senatore, N. Clayton, P. Lezza, M. Polichetti, S. Pace, R. Flukiger, IEEE Trans. [50] L. Burlachkov, V.B. Geshkenbein, A.E. Koshelev, A.I. Larkin, V.M. Vinokur, Phys.
Appl. Supercond. 15 (2005) 3329. Rev. B 50 (1994) 16770.
[8] C. Senatore, M. Polichetti, N. Clayton, R. Flukiger, S. Pace, Physica C 401 (2004) [51] R.H. Koch, V. Foglietti, Phys. Rev. Lett. 63 (1989) 511.
82. [52] D. Huse, M. Fisher, D. Fisher, Nature 358 (1992) 553.
[9] M. Polichetti, M.G. Adesso, S. Pace, Eur. Phys. J. B – Condens. Matter 36 (2003) [53] Andrei Mourachkine, Room-Temperature Superconductivity, Cambridge
27. International Science Publishing, 2004, p. 175.
[10] D.Di. Gioacchino, F. Celani, P. Tripodi, A.M. Testa, S. Pace, Phys. Rev. B 59 (1999) [54] Y.J. Uemura, Physica C 282–287 (1997) 194.
11539. [55] E. Nazarova, A. Zahariev, A. Angelow, K. Nenkov, J. Supercond. 13 (2000) 329.

You might also like