cond-mat0112364/ges.tex
1: 
2: %\documentclass[aps,prb,preprint,groupedaddress,byrevtex]{revtex4}
3: \documentclass[aps,prb,twocolumn,groupedaddress]{revtex4}
4: %\documentclass[aps,prl,preprint,superscriptaddress]{revtex4}
5: %\documentclass[aps,prl,twocolumn,groupedaddress]{revtex4}
6: %\documentstyle[aps,prl,multicol]{revtex}
7: 
8: \usepackage{amsfonts}
9: \usepackage{graphicx}
10: %\input{psfig}
11: \bibliographystyle{apsrev}
12: \begin{document}
13: 
14: \title{Low Frequency Transport Measurements in GdSr$_2$RuCu$_2$O$_8$}
15: 
16: \author{A. Vecchione}
17: \affiliation{Departement of Physics \lq\lq E. R. Caianiello\rq\rq
18: , and INFM Research Unit, University of Salerno, via
19: S. Allende, I-84081 Baronissi, Italy. }
20: \author{D. Zola}
21: \affiliation{Departement of Physics \lq\lq E. R. Caianiello\rq\rq
22: , and INFM Research Unit, University of Salerno, via
23: S. Allende, I-84081 Baronissi, Italy. }
24: 
25: \author{G. Carapella}
26: \thanks{Corresponding author}
27: \email[\newline e-mail: ]{giocar@sa.infn.it}
28: \thanks{\newline FAX: +3908965390 \\}
29: %\altaffiliation{}
30: \affiliation{Departement of Physics \lq\lq E. R. Caianiello\rq\rq
31: , and INFM Research Unit, University of Salerno, via
32: S. Allende, I-84081 Baronissi, Italy. }
33: 
34: \author{M. Gombos}
35: \affiliation{Departement of Physics \lq\lq E. R. Caianiello\rq\rq
36: , and INFM Research Unit, University of Salerno, via
37: S. Allende, I-84081 Baronissi, Italy. }
38: 
39: \author{S. Pace}
40: \affiliation{Departement of Physics \lq\lq E. R. Caianiello\rq\rq
41: , and INFM Research Unit, University of Salerno, via
42: S. Allende, I-84081 Baronissi, Italy. }
43: 
44: \author{G. Costabile}
45: \affiliation{Departement of Physics \lq\lq E. R. Caianiello\rq\rq
46: , and INFM Research Unit, University of Salerno, via
47: S. Allende, I-84081 Baronissi, Italy. }
48: 
49: \author{C. Noce}
50: \affiliation{Departement of Physics \lq\lq E. R. Caianiello\rq\rq
51: , and INFM Research Unit, University of Salerno, via
52: S. Allende, I-84081 Baronissi, Italy. }
53: 
54: 
55: \date{\today}
56: 
57: \begin{abstract}
58: Low frequency transport measurements are performed on
59: GdSr$_2$RuCu$_2$O$_8$ pellets. The observed current-voltage curves
60: are qualitatively explained in the framework of a simple
61: phenomenological model accounting for  coexistence of
62: ferromagnetism and superconductivity in the sample. A Curie
63: temperature $T_{cM}$=133~K and a superconducting critical
64: temperature $T_{cS}$=18~K, with an onset temperature
65: $T_{cO}$=33~K, are extracted from the analysis of the
66: current-voltage curves.
67: \end{abstract}
68: 
69: 
70: \pacs{74.50.+r}
71: 
72: \maketitle
73: 
74: 
75: %\begin{multicols}{2}
76: %\narrowtext
77: 
78: % body of paper here - Use proper section commands
79: % References should be done using the \cite, \ref, and \label commands
80: \section{Introduction}
81: 
82: The interplay of magnetism and superconductivity is a fundamental
83: problem in condensed-matter physics and it has been studied
84: experimentally and theoretically for almost four decades. These
85: two cooperative phenomena are mutually antagonists. Indeed, the
86: superconductivity is associated with the pairing of electrons
87: states related to time reversal while in the magnetic states the
88: time-reversal symmetry is lost and therefore there is a strong
89: competition with superconductivity\cite{maple}. However, Schlabitz
90: {\it et al.}\cite{schlabitz} showed that surprisingly magnetism
91: and superconductivity could coexist in the heavy fermion compound
92: URu$_2$Si$_2$. Other heavy fermion superconductors have also been
93: shown to exhibit magnetic moments in their superconducting phase
94: \cite{review}. All these compounds contain rare-earth or actinide
95: ions with very localized 4f or 5f orbitals, strongly interacting
96: with the conduction band electrons. This is in contrast to the
97: Chevrel phases where magnetism and superconductivity coexist
98: because the magnetic moments responsible for magnetism are only
99: very weakly coupled with the electrons that form the condensate
100: \cite{fisher}.
101: 
102: Nevertheless, there are been a number of recent studies reporting
103: the coexistence of superconductivity and magnetic order in
104: R$_{1.4}$Ce$_{0.6}$Sr$_2$RuCu$_2$O$_{10-\delta}$ \cite{felner1}
105: and RSr$_2$RuCu$_2$O$_8$
106: \cite{tallon,tallon2,williams} where R=Gd, Eu.
107: These latter compounds were originally synthesized by Bauernfeind
108: {\it et al.} \cite{bauer} and Felner and co-workers \cite{felner}.
109: Most recent reports have focused on GdSr$_2$RuCu$_2$O$_8$, which
110: has a unit cell similar to that of the YBa$_2$Cu$_3$O$_7$ high
111: temperature cuprate, where there are two CuO$_2$ layers and one
112: RuO$_2$ layer with the CuO$_2$ and RuO$_2$ layers being separated
113: by insulating layers. Magnetization and muon spin rotation
114: studies \cite{tallon2} have shown that there exists a magnetic
115: ordering transition at temperature much greater than the
116: superconducting transition temperature. Some studies have been
117: interpreted in terms of ferromagnetic order arising from the Ru
118: moment in the RuO$_2$ layers. This idea has generated considerable
119: interest because ferromagnetic order and superconductivity are
120: mutually competing processes and could only coexist via some
121: accommodation of the respective order parameters by a spatial
122: modulation\cite{chu} or via the formation of a spontaneous vortex
123: phase \cite{vortex}. However, powder neutron diffraction study
124: \cite{neutron} has shown that while there is a small ferromagnetic
125: component, the low-field magnetic order is predominantly
126: antiferromagnetic. These contrasting reports cast some doubt about
127: the magnetic nature of this compound and at the present the
128: situation has not been completely clarified. The aim of this paper
129: is to give a contribution to this discussion. Indeed, we have
130: found that transport measurements performed on
131: GdSr$_2$RuCu$_2$O$_8$ sample are in agreement with predictions of
132: a simple phenomenological model where {\it ferromagnetism} and
133: superconductivity coexist. From the experimental results a Curie
134: temperature $T_{cM}$=133K and a superconducting critical
135: temperature $T_{cS}$=18K, with an onset temperature $T_{cO}$=33K,
136: have been inferred.
137: 
138: \noindent
139: The paper is organized as follows. In Section II the
140: sample preparation is discussed. A phenomenological model for
141: expected current-voltage curves is then given in Section III.
142: Experimental results are presented and discussed in connection
143: with the theoretical prediction of the proposed  model in Section
144: IV.  Some conclusions are finally given in the last
145:  Section.
146: 
147: \section{Sample preparation and characterization}
148: \begin{figure}[!h]
149: \includegraphics[width=7cm]{figure1.eps}
150: \caption{(a) X-ray diffraction patterns of the
151: GdSr$_2$RuCu$_2$O$_8$.} \label{fig1}
152: \end{figure}
153: Precursors powders have been synthesized starting from the pure
154: binary oxide and carbonate powders, Gd$_2$O$_3$, SrCO$_3$, CuO,
155: and RuO$_2$, mixed together in the proper amount and solid state
156: reacted. The mixed powder was calcinated in air at 960 $^\circ$C
157: for 10~h. Annealing in flowing pure nitrogen at 1000~$^\circ$C
158: during 10~h was performed to reduce the formation of undesired
159: phases such as SrRuO$_3$ \cite{bauer}.
160: \begin{figure}[tbp]
161: \includegraphics[width=7cm]{figure2.eps}
162: \caption{(a) Sketch of wire connections used to measure the
163: GdSr$_{2}$RuCu$_{2}$O$_{8}$ pellet. (b) Measured voltage is
164: assumed to be the sum of two contributions. $V_{M}$ is associated
165: to the magnetic phase and $V_{S}$ to the superconductive (or
166: normal) phase.} \label{fig2}
167: \end{figure}
168: Additional two steps of annealing of 10~h in pure flowing argon at
169: 1020~$^\circ$C also contributed to suppress the SrRuO$_3$ phase.
170: Subsequently, the powders were oxygenated. Seven oxygenation
171: cycles of a mean duration of 10~h, were performed at
172: 1060~$^\circ$C in flowing pure oxygen. These fully oxygenated
173: powders were pressed in pellets by means of an hydrostatic press.
174: Five 10~h cycles in pure oxygen flux, at temperatures of
175: 1050~$^\circ$C, 1055~$^\circ$C, 1060~$^\circ$C, 1065~$^\circ$C,
176: and 1070~$^\circ$C, with intermediate grinding and mixing, have
177: been performed on the pellets. Then, a last 90~h long cycle at
178: 1070~$^\circ$C and a refining one of 10~h at 1065~$^\circ$C
179: assured the complete oxygenation of the pellets. The crystal
180: structure of the GdSr$_2$RuCu$_2$O$_8$ pellets was analyzed by
181: X-ray powder diffraction method. The data were collected with a
182: Philips PW-1700 powder diffractometer using Ni-filtered Cu
183: K$\alpha$ radiation. The X-ray spectrum of a typical fully
184: oxygenated pellet is shown in Fig.~\ref{fig1}.
185:  The scan pattern confirms that the sample is
186: GdSr$_2$RuCu$_2$O$_8$ single phased.
187: 
188: 
189: \section{Expected ac current-voltage curves}
190: 
191: If a magnetic phase is present in the
192: GdSr$_2$RuCu$_2$O$_8$, an hysteretic current-voltage ($I-V$) curve
193: should be expected when the current is swept with a frequency $\omega $.
194: \begin{figure}[tbp]
195: \includegraphics[width=7cm]{figure3.eps}
196: \caption{(a) Low magnetic field approximation of $B(H)$ for the
197: ferromagnetic phase ($T<T_{cM}$) and the paramagnetic phase
198: ($T>T_{cM}$). (b)$I-V_{M}$ curves of the ferromagnetic phase for
199: two different pulsations of a sinusoidal current forcing. (c) The
200: shape of ac $I-V_{M}$ curves is truly elliptical for the
201: paramagnetic phase and becomes a distorted ellipse in the
202: ferromagnetic phase. (d) Typical $I-V_{S}$ curves of the
203: superconductive ($T<T_{cS}$) or resistive ($T>T_{cS}$) phases.}
204: \label{fig3}
205: \end{figure}
206: In transport measurements, the four contact wire connection
207: sketched in Fig.~\ref{fig2}(a) is typically used. Here, the
208: forcing current $I(t)$ generates a magnetic field {\bf H}$(t)${\bf
209: =H}[$I(t)$] with an associated
210: magnetic induction {\bf %
211: B}$(t)${\bf =B[H}${\bf (}t{\bf )}${\bf ].} To the first order, the
212: magnetic field depends linearly on the forcing current,
213: $H(t)\propto I(t),$ so that $B(t)=B[I(t)]$ is too. From Maxwell
214: equations, we expect a voltage drop contribution due to the temporal
215: derivative of the magnetic flux linked to the voltage wires.
216: However, such a contribution is quite relevant only if the
217: magnetic induction field is quite high, i.e., if magnetic phases
218: are involved. The GdSr$_2$RuCu$_2$O$_8$ can be phenomenologically
219: seen as a series connection of superconducting and magnetic
220: phases. Hence, we expect the measured total voltage to be the sum
221: of a superconducting contribution $V_{S}$ and a magnetic
222: contribution $V_{M}$ [see Fig.~\ref{fig2}(b)]:
223: \[
224: \begin{array}{l}
225: V=V_{S}+V_{M} \\
226: V_{M}=\frac{d\Phi [B]}{dt}\propto
227: \frac{dB(t)}{dt}=\frac{dB[I(t)]}{dt}
228: \end{array}
229: \]
230: where we assumed that the relevant inductive voltage is
231: essentially due to the magnetic component. Some qualitative
232: predictions of the $I-V_{M}$ characteristic are possible from an
233: analysis of the expected $B[I(t)]$.
234: 
235: \begin{figure}[tbp]
236: \includegraphics[width=7cm]{figure4.eps}
237: \caption{ Total ac  current-voltage curve predicted for
238: GdSr$_2$RuCu$_2$O$_8$ pellet at three relevant
239: temperatures.} \label{fig4}
240: \end{figure}
241: 
242: 
243: 
244: Generally, {\bf B} is a nonlinear function of {\bf H } when a
245: material is in a magnetic phase. In the following we are concerned
246: with low magnetic field (forcing current) amplitudes. In such a
247: case, a linear relationship between {\bf B} and {\bf H} can be
248: assumed for the paramagnetic phase. Due to the vanishingly small
249: net magnetization, for an antiferromagnetic phase an
250: approximatively linear $B(H)$ relation could be again inferred,
251: while a nonlinear relation should be expected for a strongly
252: ordered phase as the ferromagnetic one. The last case can be
253: qualitatively discussed as follows. At a given low amplitude
254: magnetic field, a linear relation between {\bf B} and {\bf H} [see
255: Fig.~\ref{fig3}(a)] can be expected for temperatures above the
256: ferromagnetic transition temperature $T_{cM}$ (i.e., when the
257: material is in the paramagnetic phase) while a strongly nonlinear
258: relation between {\bf B} and {\bf H} should be  expected for
259: temperatures below $T_{cM}$ (i.e., when the material is in the
260: ferromagnetic phase). Due to the very low magnetic fields we can
261: generate with the normally used forcing currents (of the order of
262: some mA) we can assume that the saturation field will never
263: reached when the material is in the ferromagnetic phase. In other
264: words, for the used currents only the virgin curve of the
265: hysteresis loop will be normally swept, so that a single-valued
266: functional form $B(t)=B[I(t)]$ similar to the one shown in
267: Fig.~\ref{fig3}(a) can be expected to approximately describe the
268: material in the ferromagnetic phase. In such a limit, for a
269: sinusoidal forcing current of amplitude $I_{0}$ and pulsation
270: $\omega $  the $I-V_{M}$ curves shown in Fig.~\ref{fig3}(b) should
271: be observed for the ferromagnetic phase. Moreover, the distorted
272: ellipse typical of the ferromagnetic phase (at $T<T_{cM}$) should
273: become a pure ellipse in the paramagnetic phase (at $T>T_{cM}$),
274: as shown in Fig.~\ref{fig3}(c).
275: 
276: \begin{figure}[!h]
277: \includegraphics[width=7cm]{figure5a.eps}
278: \includegraphics[width=7cm]{figure5b.eps}
279:  \caption{\label{fig5} (a) Experimental $I-V$ curves measured at three different temperatures.
280:  The frequency of current supply was 20 Hz.
281:   (b) $I-V$ curves measured at $T$=28.4~K at different frequencies}
282:  \end{figure}
283: \begin{figure}[!ht]
284: \includegraphics[width=7cm,clip]{figure6a.eps}
285: \includegraphics[width=7cm,clip]{figure6b.eps}
286:  \caption{\label{fig6} Irreversible (a) and reversible (b) component
287:  extracted by experimental $I-V$ curves.}
288:  \end{figure}
289: 
290: Referring to the superconducting phase, the standard ac $I-V_{S}$
291: curves schematically plotted in Fig.~\ref{fig3}(d) are expected
292: for temperatures below or above the superconducting transition temperature $T_{cS}$. For $%
293: T<T_{cS}$ is $V_{S}=0$ for amplitude of the forcing current lower
294: than a critical current value $I_{c},$ while a truly resistive
295: curve is observed for $T>T_{cS}$
296: 
297: As stated above, the measured voltage of GdSr$_2$RuCu$_2$O$_8$
298: pellet is $V=V_{M}+V_{S}$. Hence, from information in
299: Figs.~\ref{fig3}(c) and (d), the expected ac $I-V$ curves should
300: look similar to the ones we plotted in Fig.~\ref{fig4} for three
301: relevant temperatures.
302: 
303: We should remark that, if observed, the peculiar outward cusp-like
304: distortion of the $I-V$ curve around $I=0$ is a signature of a
305: ferromagnetic order originating from the strong nonlinear increase
306: of the magnetic susceptibility below the Curie temperature.
307: Conversely, for an antiferromagnetic order, a smoother distortion
308: of the $I-V$ curve should be expected, and the area of the ellipse
309: of the paramagnetic phase should decrease for temperatures below
310: the Neel temperature due to the decrease of the
311:  magnetic susceptibility of the antiferromagnetic phase when
312:  the temperature is lowered.
313: 
314: 
315: 
316: \section{Transport measurements and discussion}
317: 
318: Measurements of $I-V$ curves were performed on a slice of
319: GdSr$_2$RuCu$_2$O$_8$ using the four contact technique shown in
320: Fig.~\ref{fig2}(a). The sizes of the slice were L=5~mm, W=2~mm,
321: and d=0.7~mm. A sinusoidal forcing current ranging from 4~mA to
322: 6~mA and frequency values of 10, 20 and 40~Hz were used. In order
323: to reduce external electromagnetic interference, measurements were
324: performed in a shielded room. The sample was also enclosed in a
325: cryoperm shield to minimize external spurious magnetic field.
326: 
327: 
328: In Fig.~\ref{fig5}(a),
329:  $I-V$ curves recorded at three different
330: temperatures are shown. At first sight, the curves are in
331: qualitative agreement with the calculated ones [see
332: Fig.~\ref{fig4}], resulting from the phenomenological model
333: reported in the previous Section. The $I-V$ curves  show an
334: hysteretic behavior at each temperature measured. Below a certain
335: temperature, an outward cusp-like distortion of the elliptical
336: shape at T=256~K is evident in the curves. Moreover,  the loop
337: area always increases when temperature is lowered. As stated in
338: the previous section this means that a {\it ferromagnetic} phase
339: is involved in the material.
340: 
341: 
342:  Figure ~\ref{fig5}(b)
343: shows that the loop area increases with the frequency of the
344: current sweep, as expected for an inductive (magnetic)
345: contribution $V_{M}\propto dB/dt$ to the total voltage drop. By
346: comparison of  Fig.~\ref{fig5}(b) and Fig.~\ref{fig3}(b), a
347: voltage contribution from a ferromagnetic phase is achieved. In
348: the previous section, we have assumed that the electrical response
349: of the material can be described as the series connection of a
350: normal (superconducting/resistive) phase and a magnetic
351: (ferromagnetic/paramagnetic) phase. When the material is a.c.
352: supplied, the resistive (normal) component  gives a reversible
353: voltage signal, whereas the inductive (magnetic) one gives rise to
354: an irreversible response accounting for the hysteretic shape of
355: the voltage-current curves in the $I-V$ plane. In order to study
356: separately the resistive and the inductive components of the
357: measured $I-V$ curves, the reversible ($V_S$) and the irreversible
358: ($V_M$) voltage were extracted  in each curve. The reversible
359: component in the total voltage signal, was calculated by using the
360: simple formula
361: \begin{equation}\label{eq1}
362:   V_S(I)=\frac{V_{up}(I)+V_{dw}(I)}{2}
363: \end{equation}
364: where $V_{up}$ and $V_{dw}$ are respectively the voltage values
365: measured during the increasing  and the decreasing branch of the
366: sinusoidal forcing current.  Then, the irreversible component was
367: extracted according to
368: \begin{equation}\label{eq2}
369:    V_M(I)=V(I)-V_S(I)
370: \end{equation}
371: 
372: The irreversible part extracted from  the total signal measured is
373: shown in Fig.~\ref{fig6}(a). Again, a qualitative agreement with
374: the computed curves [see Fig.~\ref{fig3}(b)] is recognized. For
375: temperatures ranging from 4.2~K up to about 70~K, the loop area
376: diminishes very slowly. Then the area decreases quickly and
377: smoothly changes shape becoming elliptical around $T_{cM}$=133~K.
378: From analysis of the previous Section we identify $T_{cM}$=133~K
379: as the Curie transition temperature of the magnetic phase in the
380: sample. In Fig.~\ref{fig6}(b) the reversible curves, ascribed to
381: the resistive share in the total voltage signal, are shown for
382: different temperatures. The typical non linear $I-V$ for
383:  a superconductor ($V_{S}=0$ for $-Ic<I<Ic$) can be recognized for
384: temperatures below $T_{cS}$=18~K while linear behavior is
385: recovered above this temperature.
386: 
387: From  data of the reversible curve, we extracted the resistance as
388: a function of the temperature shown in Fig.~\ref{fig7}. The
389: temperature $T_{cS}$, corresponding to a full superconducting
390: phase in the sample ($V_{S}=0$) and the onset temperature $T_{cO}$
391: were estimated 18~K and 33~K, respectively. In our measurements
392: the non-linear behavior in reversible $I-V$ curves, can be
393: recognized up to 18~K. Increasing the temperature from T$_{cS}$ up
394: to T$_{cO}$ the reversible $I-V$ curves are linear with a quite
395: fast increase of the resistivity. Above $T_{cO}$, the measured
396: resistance get lower and for $T_{cM}$=133~K the resistance shows a
397: peak. For temperatures above the magnetic transition temperature
398: the resistance diminishes again.
399: \begin{figure}[!htb]
400: \includegraphics[width=7cm]{figure7.eps}
401: \caption{\label{fig7} Resistance versus temperature calculated by
402: using the reversible $I-V$ curves.}
403: \end{figure}
404: 
405: 
406: Two main results can be drawn from the data above presented.
407: 
408: Firstly, we find a clear evidence of changes near 130K of
409: irreversible and reversible components of the $I-V$ curves and
410: therefore we infer that they could be ascribed to a
411: ferromagnetic/paramagnetic-like transition. This speculation
412: agrees well with the results reported in literature that find a
413: magnetic ordering temperature at around 130K \cite{tallon}. The
414: appearance of a spontaneous magnetic moment below this temperature
415: at a very low field suggests that the transition at T$_{cM}$ must
416: have a significant ferromagnetic component. The experimental
417: results also suggest that the ferromagnetic component persists to
418: the lowest measured temperature attained in the experiments and
419: does not appear to weaken when the superconductivity comes in at
420: T$_{cO}$=33K. The existence of a ferromagnetic component in the
421: superconducting state of this sample suggested by the low
422: frequency data here presented, is also supported by magnetic
423: measurements performed on the same sample and reported elsewhere
424: \cite{eucas}. Moreover, because no impurity lines were detected in
425: the X-ray diffraction pattern within the experimental resolution
426: we may argue that no extra phases are responsible for
427: ferromagnetism implying that this ordering is due to an intrinsic
428: phase and in this respect we could infer that the coexistence of
429: superconductivity and ferromagnetism is realized within a
430: microscopic scale. This hypothesis is corroborated by
431: magneto-optical-imaging measurements where ferromagnetism and
432: superconductivity are directly observed to coexist in the same
433: space within the experimental resolution\cite{chu2}.
434: 
435: Second, we speculate briefly  on the significance of the
436: phenomenological model previously introduced. Within our model, we
437: assume that the measured total voltage is the sum of two
438: contributions: one coming from the superconducting channel and the
439: other  due to the  ferromagnetic ordering. Although the crudeness
440: of the assumptions, we have been able to reproduce fairly the
441: shape of the $I-V$ curves and more importantly we clearly identify
442: the superconducting contribution only when the ferromagnetic one
443: is subtracted. This contribution is of the standard form for a
444: generic superconductor and this in turn further supports the
445: correctness of our assumptions.
446: \section{Conclusion}
447: 
448: In conclusion, we performed measurements on
449: GdSr$_{2}$RuCu$_{2}$O$_{8}$ ruthenate-cuprate
450: with the aim to address the question of the
451: nature of the magnetic order in the superconducting phase and
452: trying to improve the understanding of the physics of
453: ruthenate-cuprate materials. We used a relatively
454: inexplored approach, based on the analysis of low
455: frequency electrical transport measurements.
456: The observed current-voltage curves
457: have been found in quite good qualitative agreement with
458: the predictions of a
459: phenomenological model
460: accounting for coexistence
461: of both magnetic and superconducting phases in the sample.
462: Our experimental results suggest that
463: GdSr$_{2}$RuCu$_{2}$O$_{8}$ is paramagnetic above
464: $T_{cM}$=133~K, ferromagnetic between
465: $T_{cS}$=18~K and $T_{cM}$=133~K, and
466: both ferromagnetic and
467: superconducting
468:  below $T_{cS}$=18~K.
469: 
470: 
471: \section{Acknowledgements}
472: We gratefully acknowledge the contribution of D. Sisti in the
473: sample preparation.
474: 
475: %\bibliography{lowfreq01}
476: 
477: \begin{thebibliography}{99}
478: \bibitem{maple} M.B. Maple, Physica B {\bf 215}, 110, (1995).
479: \bibitem{schlabitz} W. Schlabitz {\it et al}, Z. Phys. B: Condens.
480: Matter {\bf 62}, 171 (1986).
481: \bibitem{review}A. Amato, Rev. Mod. Phys. {\bf 69}, 1119, (1997).
482: \bibitem{fisher}O. Fischer, Appl. Phys. {\bf 16}, 1 (1978).
483: \bibitem{felner1}I. Felner, U. Asaf and O. Millo, Phys. Rev. B {\bf
484: 55}, R3374 (1997).
485: \bibitem{tallon}J. Tallon {\it et al.}, IEEE Trans. Appl. Supercond.
486: \textbf{9}, 1051 (1999).
487: \bibitem{tallon2} C. Bernhard {\it et al.}, Phys. Rev. B \textbf{59}, 14099 (1999).
488: \bibitem{williams} G. V. M. Williams, S. Kr\"amer Phys. Rev. B, \textbf{62},
489: 4132(2000)).
490: \bibitem{bauer} L. Bauernfeind, W. Widder, and H. F. Braun, Physica C \textbf{254}, 151
491: (1995); L. Bauernfeind, W. Widder, and H. F. Braun, Journ. of Low
492: Temp. Phys. \textbf{105}, 1605 (1996).
493: \bibitem{felner}I. Felner, U. Asaf, S. Reich, and Y. Tsabba, Physica C
494: \textbf{311}, 163 (1999).
495: \bibitem{chu}C. W. Chu {\it et al}, Physica C {\bf 335},231 (2000).
496: \bibitem{vortex} E. B. Sonin and I. Felner, Phys. Rev. B \textbf{57}, 14000
497: (1998);  I. Felner, U. Asaf, Y. Levi, and O. Millo, Physica C
498: \textbf{334}, 141 (2000).
499: \bibitem{neutron} J.W. Lynn {\it et al.}, Phys Rev B \textbf{ 61}, 14 964 (2000).
500: \bibitem{eucas}A. Vecchione {\it et al.}, cond-mat/0110482.
501: \bibitem{chu2}C.W. Chu {\it et al.}, cond-mat/9910056.
502: 
503: %\bibitem{tallon2} A. C. McLaughlin {\it et al.}, Phys. Rev. B \textbf{60}, 7512 (1999).
504: %\bibitem{tallon3} A. Fainstein {\it et al.}, Phys. Rev. B \textbf{60}, 12597 (1999);
505: %A. C. McLaughlin and J. P. Attfield, Phys. Rev. B \textbf{60},
506: %14605 (1999).
507: 
508: 
509: %\bibitem{tang}L.B. Tang {\it et al.},
510: %Physica C \textbf{282-287}, 947 (1997).
511: %\bibitem{Cox:1991} D.E. Cox, {Handbook of Synchrotron Radiation}, Vol.3, Ch. 5, Powder
512: %Diffraction (eds. Brown, G. and Moncton, D.E.), Elsevier,
513: %Amsterdam.
514: %\bibitem{GSAS} A.C. Larson, R.B. Von Dreele,
515: %GSAS - General Structure Analysis System, Los Alamos National
516: %Laboratory, Los Alamos (USA), LANL Report LAUR 86-748.
517: %\bibitem{pvoigt} P. Thompson, D.E. Cox, J.B. Hastings,
518: % J. Appl. Cryst. \textbf{20}, 79 (1987).
519: %\bibitem{axdiv} L.W. Finger, D.E. Cox,  A.P. Jephcoat,
520: % J. Appl. Cryst. \textbf{27}, 892 (1994).
521: %
522: %
523: %\bibitem{chma} O. Chmaissem {\it et al.}, Phys. Rev. B \textbf{61}, 6401
524: %(2000).
525: %\bibitem{ndbacuo} E.A. Goodilin et al., Physica C \textbf{272}, 65 (1996); S.I.Yoo and R.W.
526: %McCallum, Physica C \textbf{210}, 147 (1993).
527: %
528: %\bibitem{cant} J. D. Jorgensen, O. Chmaissem, H. Shaked, S. Short, P. W. Klamut,
529: %B. Dabrowski, J. L. Tallon Phys. Rev. B 63, 054440 (2001).
530: 
531: \end{thebibliography}
532: 
533: 
534: 
535: \end{document}
536: