1: \documentclass[prb,twocolumn,showpacs]{revtex4}
2: \begin{document}
3: \title{High Temperature Mixed State $c-$Axis Dissipation in Low
4: Carrier Density
5: $Y_{0.54}Pr_{0.46}Ba_{2}Cu_{3}O_{7-\delta}$}
6: \author{T. Katuwal}
7: \author{V. Sandu\thanks}
8: \thanks{Permanent Address: National
9: Institute of Materials Physics, 077125
10: Bucharest-Magurele, Romania}
11: \author{C. C. Almasan }
12: \affiliation{Kent State University, Kent, OH-44242 }
13: \author{B. J. Taylor and M. B. Maple}
14: \affiliation{University of California at San Diego, La Jolla, CA-92093}
15: \author{}
16: \affiliation{ }
17: \date{\today}
18: \begin{abstract}
19: The nature of the out-of-plane dissipation was investigated in
20: underdoped $Y_{0.54}Pr_{0.46}Ba_{2}Cu_{3}O_{7-\delta}$ single
21: crystals at temperatures close to the critical temperature. For this
22: goal, temperature and angle dependent
23: out-of-plane resistivity measurements were carried out both below and above
24: the critical temperature. We found that the Ambegaokar-Halperin
25: relationship [V. Ambegaokar, and B. I. Halperin, Phys. Rev.
26: Lett. \textbf{22}, 1364 (1969)] depicts very well the angular
27: magnetoresistivity in
28: the investigated range of field and temperature. The main finding is
29: that the in-plane phase fluctuations decouple the layers above the
30: critical temperature and the charge transport is governed only by the
31: quasiparticles.
32: We also have calculated the interlayer Josephson critical current
33: density, which was
34: found to be much smaller than the one predicted by the theory of
35: layered superconductors.
36: This discrepancy could be a result of the $d$-wave symmetry of the order parameter and/or of the non
37: BCS temperature dependence of the $c$-axis penetration length.
38:
39:
40: \end{abstract}
41: \pacs{72.20.My, 75.30.Vn }
42: \maketitle
43:
44:
45: Since the discovery of high temperature superconductivity in cuprates, the
46: contrasting temperature $T$ dependence of in-plane $\rho_{ab}$ and
47: out-of-plane $\rho_{c}$ resistivities have been an issue of debate. This
48: topic is even more complex in the
49: underdoped systems where the density of states DOS of quasiparticles as well
50: as the
51: transfer integrals are momentum and temperature dependent. Additionally,
52: strong fluctuations, which are expected at low carrier density,
53: have a major contribution to dissipation.
54:
55: The most debated issue is the
56: interlayer dissipation and its field and angle dependence.
57: In the most general way, the $c$-axis conductivity is presumably
58: controlled by the
59: tunneling of the Cooper pairs and quasiparticles, with conductivities
60: $\sigma_{J}$ and $\sigma_{c,qp}$, respectively; i.e., the $c$-axis
61: conductivity $\sigma_c =
62: \sigma_J+\sigma_{c,qp}$. This tunneling is the main consequence of the layered
63: structure of the cuprates, which can be depicted as stacks of
64: Josephson junctions made out of superconducting CuO$_2$
65: "electrodes"
66: and intermediate blocking layers. Actually, this Josephson coupling
67: of the layers distinguishes the layered cuprates from the ordinary anisotropic
68: superconductors.
69: The existence of this coupling was unambiguously demonstrated for
70: cuprates with large anisotropy
71: $\gamma>100$, where $\gamma$ is the ratio of the $c$-axis and
72: in-plane effective masses,
73: $\gamma\equiv m_{c}/m_{ab}$, either directly by $I-V$ characteristics
74: on small single crystals or messa
75: structures \cite{Kleiner1, Kleiner2, Sakai, Irie, Yurgens, Latyshev}
76: or by Josephson plasma resonance experiments.
77: \cite{Matsuda,Tsui, Bulaevski1, Matsuda2, Hanaguri, Gaifulin} In the
78: case of systems with lower anisotropy, e.g.,
79: YBa$_2$Cu$_3$O$_{7-\delta}$ ($\gamma\approx 5-9$), the existence of a
80: Josephson coupling in the $c$-axis direction was largely debated.
81: Direct measurements
82: were reported only in underdoped YBa$_2$Cu$_3$O$_{6+x}$.\cite{Rapp}
83: while optical conductivity measurements
84: \cite{ Basov, Homes} showed evidence of Josephson coupling in these
85: low anisotropic cuprates.
86:
87: As in the case of the
88: in-plane transport, the interlayer dissipation is strongly influenced by phase
89: fluctuations. A magnetic field applied perpendicular to the layers
90: penetrates as pancake
91: vortices with a particular phase distribution around each core. It is
92: known that position fluctuations of the pancake
93: vortices due to pinning and/or
94: thermal diffusion of the vortex core, reduce phase correlations,
95: hence, the Josephson
96: coupling. Nevertheless, as long as phase correlations exist, they could provide
97: a
98: Josephson-type contribution
99: to the out-of-plane transport. In fact, such phase correlations,
100: though weak, were
101: identified experimentally even in the liquid state of the vortex system.
102: \cite{Cubitt, Tsui, Matsuda3}
103: Generally, these phase (vortex) fluctuations have important consequences both
104: below and above the critical temperature $T_{c0}$. For $T<T_{c0}$,
105: they drive the vortex
106: system in a liquid state, whereas above $T_{c0}$, they allow vorticity to
107: survive and to contribute to the in-plane
108: dissipation.\cite{Sandu, Katuwal}
109: To be specific, in the latter case, there is a crossover in the
110: in-plane dissipation from a regime of pure flux-flow to a regime
111: entirely due to
112: quasiparticles, which occurs at a particular temperature {\it higher} than
113: $T_{c0}$. The importance of the fluctuations increases when the
114: density of charge carriers is reduced,
115: i.e., in the case of underdoped cuprates.
116:
117: Even though the interlayer coupling was investigated on a large extent
118: in large-$\gamma$ superconductors, the
119: scarcity of data is evident
120: in low and medium-anisotropic superconductors. Therefore, in the
121: present study, using temperature, field,
122: and angle dependence of the out-of-plane resistivity, we investigate the
123: features of the
124: $c$-axis dissipation in a medium-$\gamma$ superconductor in an
125: attempt to find the extent of Josephson response close to
126: $T_{c0}$, i.e., in a temperature range where the phase fluctuations are
127: important enough to
128: reduce and/or suppress the Josephson coupling. As in the case of
129: in-plane resistivity, \cite {Sandu, Katuwal} we
130: take advantage of the different angular dependences of the different
131: contributions to resistivity
132: to obtain the desired information. This investigation is
133: performed on Y$_{0.54}$Pr$_{0.46}$Ba$_{2}$Cu$_{3}$O$_{7-\delta}$ single
134: crystals, where we used the antidoping effect of praseodymium to reduce the
135: charge carrier density and increase the electronic anisotropy,
136: hence, to drive the system in the strong fluctuation regime.
137: In this way the electromagnetic and Josephson coupling of the pancake
138: vortices
139: and, subsequently, the interlayer coherence are weakened compared with the
140: thermal fluctuations.
141: Our main finding is that the $c$-axis dissipation scales with
142: $Hcos\theta$ below $T_{c0}$, in a temperature range that ends at the critical
143: temperature. We relate the failing of the scaling above $T_{c0}$ to the
144: suppression of the interlayer Josephson coupling by the
145: in-plane phase fluctuations. So, even though in-plane strong
146: superconducting fluctuations, hence, a large amount of condensate,
147: persist above $T_{c0}$,\cite{Sandu, Katuwal} there is no out-of-plane
148: Josephson coupling
149: which would provide an enhanced conductivity above $T_{c0}$.
150:
151: Y$_{0.54}$Pr$_{0.46}$Ba$_{2}$Cu$_{3}$O$_{7-\delta}$ single crystals
152: with typical
153: dimensions of 1.0
154: $\times$ 0.5 $\times$ 0.02 mm$^{3}$ were grown using a standard procedure
155: described elsewhere. \cite{Paulius} The dimensions the the crystal
156: for which the data are shown
157: here are 0.6
158: $\times$ 0.65 $\times$ 0.017 mm$^{3}$. We attached four gold wires
159: (0.025 mm in diameter) with
160: silver epoxy onto each of the two large faces of the single crystal
161: (see top Inset to Fig. 1). The two
162: outer (inner)
163: contacts on the same face were used as current (voltage) terminals. The
164: contact resistance is 2 $\Omega$ at room temperature. First, a
165: magnetic field H up to 14 T was applied along the $c-$direction of
166: the sample, a
167: constant current
168: $I\leq 1$ mA was fed through pads of size 0.05 $\times$ 0.5 mm$^{2}$,
169: alternately
170: on both faces, and the voltage on each face of the single crystal
171: was measured at set temperatures between
172: 0 and 300 K. Next, we measured in the same way the voltages at
173: different constant temperatures but this time
174: the single crystal was rotated in the applied magnetic field with the
175: angle $\theta$ between
176: $H$ and the $c$-axis varying between 0 and 360$^\circ$. The
177: out-of-plane $\rho_{c}$ and in-plane $\rho_{ab}$ resistivities were
178: calculated using an algorithm described elsewhere.\cite{Levin} The critical
179: temperature $T_{c0}$ was taken at
180: the midpoint of the normal-superconductor transition.
181:
182: Figure 1 shows the temperature $T$ dependence of $\rho_c$ of an
183: Y$_{0.54}$Pr$_{0.46}$Ba$_{2}$Cu$_{3}$O$_{7-\delta}$ single crystal measured at
184: different applied magnetic fields, while its bottom
185: Inset shows the zero field $\rho_{c}(T)$ over the whole measured
186: temperature range. Even though this is a
187: strongly underdoped sample, with a zero-field superconducting transition
188: temperature $T_{c0}=38$ K, it has a sharp transition,
189: attesting to the good quality of this single crystal. The
190: magnetoresistivity in the normal
191: state is very small and positive. The normal state is metallic at
192: high temperatures and becomes nonmetallic for
193: temperatures lower than 133 K. The upturn in $\rho_{c}(T)$ observed below
194: this temperature is the result of a complex process. First, there is a
195: reduction in the planar density of states DOS due to the opening of
196: the pseudogap at
197: $\mathbf{k}=(\pi/2a,0)$ (nodes) with decreasing $T$;\cite{Damascelli}
198: second, the transfer
199: integral of the coherent contribution is angle dependent with maxima
200: at the nodal points.
201:
202: The angular dependence of the normalized $c-$axis resistivity
203: $\rho_{c}(\theta)/\rho_{c}(\theta = 0^\circ)$ of
204: $Y_{0.54}Pr_{0.46}Ba_{2}Cu_{3}O_{7-\delta}$ measured at 30, 35 and 40
205: K in a magnetic field of 14 T
206: is shown in Fig. 2. Both below and above $T_{c0}$, $\rho_c(\theta)$
207: displays a minimum at $\theta=90^{\circ}$ (i. e. for
208: $H\parallel ab-$plane) and a maximum value at
209: $\theta = 0^\circ$ (i. e. for $H\parallel c-$axis). The former
210: (latter) value of the
211: angle corresponds to maximum (zero) transverse Lorentz force on the
212: flux vortices.
213: This fact rules out the possibility that flux motion contributes to
214: the measured
215: $c-$axis dissipation, since the measured dissipation is maximum for
216: angles for which
217: the Lorentz force is zero. Therefore, we assume that the
218: $c$-axis transport in the mixed state involves Josephson and quasiparticle
219: contributions which depend on $T$, $H$, and $\theta$.
220:
221: In the geometry we have used, the $c$-axis component of the current
222: density flows mainly
223: along the crystal edges in a wall of width almost equal to the
224: pad width $\Delta=50$ $\mu$m. We call this region the
225: active area. The variation of the current
226: density over this width is $[J_{z}(z, L/2)-J_{z}(z,
227: L/2-\Delta)]/J_{z}(z, L/2) <14$ \%, while, outside of this
228: width, the current density decreases
229: fast toward the center of the crystal.
230: Therefore, we conclude that only these edge walls of the single crystal
231: are involved in the current transport along c-axis.
232:
233: The characteristic
234: lengths for a stack of Josephson
235: junctions are the Josephson penetration length $\lambda_{j}$, which
236: accounts for the field
237: penetration within the nonsuperconducting interlayer space and the
238: $c$-axis London penetration
239: length $\lambda_{c}$, which accounts for the magnetic screening. For this single crystal,
240: $\lambda_{j}
241: \approx 0.03$ $\mu$m, much smaller than $\Delta$. Therefore, it is inappropriate to
242: consider the single crystal as a stack of short junctions. For a stack
243: of long junctions, Josephson vortices might be generated even in the
244: absence of in-plane external fields if the in-plane
245: currents are strong enough to create the required phase gradient. This phase gradient is generated
246: if the junction length is larger than $\lambda_{c}$.
247: Hence, if the width of the active area $\Delta$ is of the order of the $c$-axis magnetic penetration
248: length
249: $\lambda_{c}$ one can assume that the
250: in-plane currents are too small to generate Josephson vortices in
251: the active area. There are no available data concerning $\lambda_{c}$ for
252: strongly underdoped Y$_{1-x}$Pr$_{x}$Ba$_{2}$Cu$_{3}$O$_{7-\delta}$.
253: An estimate
254: of
255: $\lambda_{c}$ based on a similar underdoped YBa$_2$Cu$_3$O$_{x}$ is of the
256: order of 30 $\mu$m.
257: \cite{Hosseini} Therefore, this Josephson system satisfies the condition
258: $\lambda_{j}
259: \ll\Delta \sim\lambda_{c}$. Hence, we assume that in this case the
260: Josephson vortices are most likely
261: absent and we can use the Ambegaokar-Halperin relationship.
262: Additionally, the current we use
263: is very low with a total current density $J=4$ Acm$^{-2} \ll J_{c}$.
264: Under these assumptions, we proceed to obtain
265: an analytical relationship for the $c$-axis resistivity
266: as a function of field, temperature, and angle.
267:
268: Generally, the total conductivity can be derived from Kubo's
269: relationship for both Josephson
270: and quasiparticle current.\cite{Koshelev2} However, this
271: relationship is difficult to handle in the absence of an analytical
272: dependence of the in-plane diffusion coefficient on field
273: and temperature. Several experimental
274: reports\cite{Briceno,Gray,Hettinger,Yoo} have
275: shown that the Ambegaokar-Halperin (AH) expression, \cite{Ambegaokar}
276: which is valid
277: for a single Josephson junction, can be successfully used to fit the
278: $c$-axis resistivity
279: data when making specific assumptions on the expression of the
280: critical current.
281: The Ambegaokar-Halperin expression for $c$-axis resistivity
282: is given by:
283:
284: \begin{equation}\rho_{c}(T) = \rho_{n} \left[\mathcal{I}_{0}\left(\frac
285: {\Phi_0
286: I_{c}(T)}{2\pi k_{B} T}\right)\right]^{-2},
287: \end {equation}
288: where $\rho_n$ is the intrinsic normal-state resistivity of the
289: junction, $\mathcal{I}_{0}$ is
290: the modified Bessel function, $\Phi_0$ is the flux quantum,
291: $I_{c}$ is the critical current at a temperature T, and
292: $k_{B}$ is Boltzmann's constant.
293: Note that the AH relationship accounts also for the contribution of the
294: quasiparticles through $\rho_n$.
295:
296: Because of the high energy of the
297: Josephson coupling $\Phi_0 I_{c}/2\pi$ relative to the thermal
298: energy, one can use the asymptotic expansion
299: $\mathcal{I}_0(x)\approx
300: \exp(x)/\sqrt{2\pi x}$ for $x > 1$. This approximation gives, for example, for
301: $x=3$ a 5\% error compared
302: with the exact Bessel function. An estimate of the zero-field conductivity of
303: our samples gives $x(B=0)=5.5$ at 35 K and an error of approximately 2\%.
304: With this approximation, Eq. (1) gives the the following expression
305: for the $c-$axis
306: resistivity at high temperatures:
307:
308: \begin{equation}
309: \rho_{c}(T,H)\approx
310: \rho_{c,qp}\frac{\Phi_0I_{c}(T,H)}{k_{B}T}{\exp\left(-\frac{\Phi_0I_{c}(T,H)}{\pi
311: k_{B}T}\right)}.
312: \end{equation}
313:
314: Next, we discuss the $T$ and $H$ dependence of $I_c$. The temperature
315: dependence of the
316: AH relationship
317: is limited only to the spin-wave type fluctuation of the order parameter.
318: Therefore, we have to include in the above critical current term the
319: contribution
320: accounting for the presence of vortices. The maximum Josephson current
321: $I_{c}$, which is related to the interlayer phase difference, is strongly
322: influenced by the level of the fluctuations of the phases in each layer.
323: To be specific, the Josephson current density $J_{c}$ decreases as a result of
324: both the thermally-induced misalignment of the planar
325: vortices, which creates a gauge invariant phase difference
326: $\varphi_{n,n+1}$ from layer to layer, and the thermally induced phase
327: slippage (spin-wave type phase fluctuations).
328: Regarding the thermal motion of the pancake vortices, there is a
329: complex process of renormalization of $J_{c}$,
330: which suppresses $J_{c}$.\cite {Daemen} A suppressed $J_{c}$ increases
331: in turn the penetration depth
332: $\lambda_{c}$, hence, reduces the elastic constants, which in turn
333: further suppresses $J_{c}$. This process is present both
334: in the solid and liquid phases of the vortex system because
335: the only difference between these two phases is the vanishing of the shear
336: constant $c_{66}$ in the latter
337: phase. Additionally, the phase difference depends on the pancake
338: position within each plane. Therefore, to obtain the interlayer critical
339: current, one has to go beyond the ensemble average
340: \cite{Koshelev4, Koshelev3} and to also use a space average of the
341: critical current. Hence, \cite{Fistul,Logvenov}
342: \begin{equation}
343: I_{c}^2=J_{c0}^2\!\int\!\!d\mathbf{r}_1\!\!\int\!\!d\mathbf{r}_2
344: \exp\left
345: \{i\left[\phi_{n,n+1}\left(\mathbf{r}_1\right)-\phi_{n,n+1}\left(\mathbf{r}_2\right)\right]\right\},
346: \end{equation}
347: with $J_{c0}$ the local intrinsic (bare) Josephson critical current
348: density. At high temperatures, Eq. (3) becomes:
349: \begin{displaymath} I_{c}^2 (T, H)=J_{c0}^2(T)AS(H),
350: \end{displaymath} where $A$ is the
351: area of the junction and $S(H)$, given by
352:
353: \begin{equation} S(H)=\int d\mathbf{r}\langle
354: cos\left[\phi_{n,n+1}\left(\mathbf{r}\right)\right]-cos\left[\phi_{n,n+1}(0)\right]\rangle,
355: \end{equation}
356: is the correlation area, which needs to be evaluated. Following
357: Koshelev, Bulaevski, and Maley, \cite{Koshelev3} we
358: make the approximation
359: \begin{equation}
360: S(H)\approx f\gamma^2s^2\left(H_{j}/H_{z}\right)^\alpha.
361: \end{equation}
362: Here,
363: $f(H,T)$ is a function of order unity with a weak $T$ dependence,
364: $s$ is the interlayer spacing, $H_{j}=\Phi_0/\gamma^2
365: s^2$ is a characteristic field, and $H_{z}=Hcos\theta$ is the
366: magnetic field in the
367: $c$-axis direction. The deviation of the exponent
368: $\alpha=1-k_{B}T/2\pi E_0(T)$ from unity
369: is a result of the spin-wave type phase fluctuations and increases
370: strongly close to the critical
371: temperature. [$E_0=s\Phi_{0}^2/(4\pi\mu_0\lambda_{ab}^2)$ is the
372: Josephson energy of the area $\gamma^2 s^2$].\cite{Koshelev1}
373: The above approximation [Eq. (5)] is
374: valid for applied magnetic fields larger than the
375: characteristic field $H_{j}$.
376: For the single crystals
377: with a Pr doping $x$ = 0.46, for
378: which
379: $\gamma=26$ and $s=11.7 \AA$, the characteristic field $H_{j}\approx 2.3$ T.
380: Hence, Eq. (5) is valid for $H>2.3 $ T. With this
381: approximation,
382: the Josephson critical current becomes
383:
384: \begin{equation}
385: I_{c}(H,T)=J_{c0}(T)\gamma s f^{1/2}A^{1/2}(H_j/H_{z})^{\alpha/2}.
386: \end{equation}
387: Equation (4), hence Eq. (6), was derived in the approximation of a completely
388: decoupled pancake vortex liquid, when the correlation of the
389: $cos\varphi_{n,n+1}$ terms drops on a length scale of the order of the
390: intervortex
391: spacing.
392:
393: Equations
394: (2) and (6) give the following expression for the out-of-plane
395: resistivity:
396: \begin{displaymath} \rho_{c}(T,H,\theta)\approx
397: \rho_{c,qp}\frac{\Phi_0J_{c0}(T)\gamma s
398: f^{1/2}A^{1/2}}{k_{B}T}\left(\frac{H_{j}}{H|cos\theta|}
399: \right)^{\nu}\times\end{displaymath}
400: \begin{equation}\;\;\; \exp\left [-\frac{\Phi_0J_{c0}(T)\gamma s
401: f^{1/2}A^{1/2}}{\pi k_{B}T}\left(\frac{H_{j}}{H|cos\theta|}
402: \right)^{\nu}\right], \end{equation} with $\nu=\alpha/2$. An
403: important result of Eq.
404: (7) is that the out-of-plane resistivity in the mixed state where Josephson
405: tunneling dominates the conduction should scale with
406: $H\cos\theta$.
407: Figure 3(a) is a plot of $\rho_c$ vs $Hcos\theta$, measured at several
408: temperatures below
409: $T_{c0}$ and in applied magnetic fields of 6, 8, 10, 12, and 14 T. Note
410: that the data, indeed, follow the $H\cos\theta$ scaling.
411:
412: Above the critical temperature, $\rho_{c}$ vs $H
413: cos\theta$ plots
414: do not map anymore onto a single curve [see Inset to Fig. 3(a) for $T$ = 40 K].
415: This fact hints to the complete vanishing of the interplane phase
416: correlations, hence, of the Josephson contribution to dissipation
417: above $T_{c0}$. Thus, the phase fluctuations suppress the mechanism
418: responsible for bulk superconductivity at the critical
419: temperature, even though the in-plane dissipative processes still carry
420: the hallmark of superconducting phase fluctuations
421: up to temperatures well above $T_{c0}$.\cite{Sandu, Katuwal}
422:
423: The term $\rho_{c,qp}$ in Eq. (7) ensues from the quasiparticle
424: current driven by the time variation of the gauge invariant phase
425: difference. In the case of d-wave superconductors, the quasiparticle
426: concentration does not vanish with decreasing temperature. At any
427: temperature $T$, there is always a $\mathbf {k}$ range so
428: that $\Delta (\mathbf{k})< k_{B}T$, which facilitates the quasiparticle
429: excitation near the gap nodes. Microscopic models
430: have shown that in the case of constant DOS, the quasiparticle out-of-plane
431: resistivity below the critical temperature depends on
432: temperature as $\rho_{c,qp} = \rho_{n} (3 \Delta_{0}^{2}/\pi T^2)$ if the
433: tunneling is coherent and is $T$-independent if the
434: tunneling is completely incoherent.\cite{Artemenko} A real material
435: displays both contributions, hence, it follows a power law
436: temperature dependence.
437: Additionally, the DOS decreases with decreasing $T$ in underdoped cuprates. The
438: absence of an analytical expression for the temperature dependence of
439: the DOS makes impossible the
440: determination of the
441: temperature dependence of the quasiparticle contribution to
442: resistivity.
443: In the presence of a magnetic field, there is a small
444: change in conductivity due to the Doppler
445: shift in the quasiparticle spectrum.\cite{Vekhter}
446: However, a
447: sensitive change requires extremely high magnetic
448: fields,\cite{Morozov} so that in
449: the present measurements ($H\leq 14 $ T)
450: $\rho_{c, qp}$ is practically field independent. Indeed, the
451: $c$-axis resistivity data show that the normal-state magnetoresistivity is very
452: small, which implies an almost $H$ independent quasiparticle contribution
453: to the
454: out-of-plane conduction. Therefore, we assume that $\rho_c(T)$ measured in 14 T
455: in the normal state and its extrapolation at lower temperatures in
456: the mixed state is the
457: out-of-plane quasiparticle resistivity $\rho_{c,qp}(T)$.
458:
459: With $\rho_{c,qp}(T)$ determined as just discussed above, we fit the
460: $\rho_c(T,H,\theta)$ data with Eq. (7) with two fitting parameters:
461: the exponent
462: $\nu(T)$ and
463:
464: \begin{displaymath}C(T)=\frac{\Phi_0 J_{c0}(T)\gamma s
465: A^{1/2}H_{j}^{\nu(T)}}{\pi k_{B}T},\end{displaymath} where we take
466: $f\approx 1$. The results of the fitting of the data measured at several
467: temperatures, and in 14 T and 10 T are shown in Fig. 3(b) and its Inset,
468: respectively. The excellent fit of the out-of-plane resistivity data
469: with Eq. (7)
470: confirms the validity of our approach and shows that the $T$,
471: $H$, and
472: $\theta$ dependence of the measured out-of-plane
473: resistivity in the mixed state is dominated by the Josephson tunneling of the
474: Cooper pairs and the quasiparticle tunneling.
475:
476: The picture that immerges from these results is as follows. As in the case of
477: the in-plane dissipation, the fluctuations have a significant
478: effect on the nature of $\rho_{c}$ at high temperatures in low charge
479: carrier density cuprates such as
480: Y$_{0.54}$Pr$_{0.46}$Ba$_{2}$Cu$_{3}$O$_{7-\delta}$. Nevertheless,
481: although the dissipations along the two directions have the same
482: origin, they are governed by different mechanisms, hence, display two
483: temperature scales.
484: The out-of-plane dissipation is
485: governed by the Josephson tunneling of the Cooper pairs and the
486: quasiparticle tunneling.
487: With increasing $T$, the in-plane phase fluctuations give rise to a
488: rapid suppression
489: of the Josephson coupling, the unique process which makes
490: superconductivity a bulk phenomenon, and of the corresponding
491: interlayer supercurrent density.
492: Small interlayer correlations survive above the
493: irreversibility temperature up to 37 K above which both the Josephson
494: coupling and the corresponding interlayer supercurrent vanish. At $T>T_{c0}$,
495: even though in-plane vorticity has been shown to exist in a strong
496: fluctuating regime, \cite{Sandu} these phase fluctuations are too
497: fast to allow any phase
498: correlation along the $c$-axis. Hence, $\sigma_{J}=0$ and the
499: out-of-plane conductivity is only a result of quasiparticle
500: tunneling. This is not
501: the case of the in-plane
502: dissipation, which displays a
503: contribution from phase fluctuations (pancake vortices) arising from
504: their motion driven by the
505: transport current up to a charge carrier density dependent temperature
506: $T_{\varphi}>>T_{c0}$.\cite{Sandu, Katuwal} Therefore, the temperature
507: scale is $T_{c0}$ ($T_{\varphi}$) for
508: the contribution of the superconducting dissipation to the total out-of-plane
509: (in-plane) resistivity.
510: These two temperatures merge as the density of charge
511: carriers increases toward the optimal doping.
512:
513: Additional improvements to the model used in this study would require
514: the incorporation
515: of nonequilibrium effects due to the nodal quasiparticles,
516: mainly in the high temperature range.\cite{Artemenko2}
517: This simple AH approach, which constitutes the starting point of our
518: analysis, though fruitful, has been the subject of criticisms regarding
519: the omission of the interaction
520: between adjacent junctions, \cite{Goldobin} and the assumption of a Fermi
521: liquid behavior in underdoped cuprates.
522: However, the simplicity of the AH relationship and
523: its reported success
524: at low temperatures make it attractive, with appropriate assumptions,
525: in the high
526: temperature regime. Our present results confirm, indeed, its
527: applicability close to
528: $T_{c0}$.
529:
530: From the angular
531: magnetoresistivity data, we also extracted the temperature dependence of the
532: bare interlayer (Josephson) critical current density at high
533: temperatures, close to
534: $T_{c0}$, from
535: \begin{equation} J_{c0}(T) = \frac{\pi
536: k_{B}TC(T)}{\Phi_0\gamma s A^{1/2}H_{j}^{\nu(T)}},
537: \end{equation}
538: using the fitting parameter $C(T)$ as obtained from the
539: fit of the angular magnetoresistivity data at high temperatures with Eq. (7). A
540: plot of $J_{c0}$ vs $T$ is shown in Fig. 4.
541: These values of $J_{c0}$ are
542: smaller than the values smaller than the values predicted by a simple model of layered
543: superconductors,
544: which gives
545: $J_{c0}=\Phi_0/[2\pi \mu_0\lambda_{c}^2(T)]$.
546: The temperature dependence of the
547: data follow the power law
548: $J_{c0}(T)\approx 7.6 (T/T_{c0})^{-1.73}$. Such a $T$ dependence could
549: be the result of the complexity of the interlayer Cooper pair
550: transport in cuprates containing conducting CuO chains combined with
551: the $d-$wave symmetry of the superconducting order
552: parameter. Therefore, the temperature dependence of $J_{c0}$ is
553: provided not only by $\lambda_{c}^{-2}(T)$, but also by the
554: $T$-dependence of the density of states of the localized resonant centers.\cite{Abrikosov}
555: The $\lambda_{c}^{-2}(T)$ itself changes its convexity at high
556: temperatures, \cite{Hosseini} most probably due to the excitation of
557: the quasiparticles out of the condensate at gap nodes.
558:
559: The temperature dependence of the exponent $\nu$ is shown in the Inset to
560: Fig. 4. A fit of the data gives a linear $T$ dependence, i.e.,
561: $\nu(T)=1.7(1-T/T^{*})$, where $T^{*}=39.5$ K is slightly higher than
562: $T_{c0}$ defined
563: as the mid point of the transition curve. Theoretically, the exponent
564: $\nu$ should be linear in
565: $T\lambda_{ab}^2(T)$. A plot of the theoretical $\nu(T)$ curve using the BCS
566: dependence of
567: $\lambda_{ab}^2(T)$ is also shown in Fig. 4. The experimental and
568: theoretical values are close to each other for $T\geq 30$ K. At lower
569: temperatures, the data obtained from fitting are almost twice as high as the
570: theoretical values. Actually, the expression used for the field
571: dependence of $I_{c}$ [Eq. (6)] is
572: not valid at low temperatures. This could be one reason for the above
573: discrepancy. It is interesting to note, however, that the
574: scaling of
575: $\rho_{c}(H|cos\theta|)$ still works down to 20 K. Another reason for
576: the discrepancy between the experimental and
577: theoretical values of $\nu$ could be that $\lambda_{ab}^{-2}(T)$ has a non BCS
578: temperature dependence due to the nodal quasiparticles.
579:
580: In summary, we analyzed the out-of-plane dissipation in a medium
581: anisotropic underdoped cuprate at temperatures around $T_{c0}$. We
582: performed these measurements in order to investigate the origin of
583: the large $\rho_{c}$ and its $T$, $H$, and angle dependence in
584: this material. The data are well fitted by the Ambegaokar-Halperin
585: expression for
586: temperatures up to the critical temperature and applied magnetic fields as
587: high as 14 T. We found that the interlayer resistivity follows a
588: simple scaling law as a function of magnetic field and angle; i.e.,
589: $\rho_{c}(H,\theta)=
590: \rho_{c}(|Hcos\theta|)$. The existence of the scaling close to the critical
591: temperature proves the persistence of
592: interlayer correlations above the irreversibility temperature.
593: Nevertheless, the scaling fails above the mid point critical
594: temperature, above which the $c$-axis charge transport is governed by
595: quasiparticles only. This is different from the in-plane dissipation,
596: in which the contribution of the superconducting fluctuations can be
597: discerned up to temperatures as high as
598: $1.5\times T_{c0}$. We also have determined the interlayer critical current
599: density. It was found to be lower than predicted by simple
600: models of Josephson coupled superconductors.
601:
602:
603: \textbf{Acknowledgments}
604:
605: This research was supported by the National Science
606: Foundation under Grant
607: No. DMR-0406471 at KSU and the US Department of Energy under Grant No.
608: DE-FG02-04ER46105 at UCSD. \label{}
609:
610:
611:
612: \begin{thebibliography}{99}
613: \bibitem{Kleiner1} R. Kleiner, F. Steinmeyer, G. Kunkel, and P.
614: M\"{u}ller, Phys. Rev. Lett. \textbf{68}, 2394 (1992).
615: \bibitem{Kleiner2} R. Kleiner and P. M\"{u}ller, Phys. Rev. B
616: \textbf{49}, 1327 (1994).
617: \bibitem{Sakai} M. Sakai, A. Odagawa, H. Adachi, and K. Setsune,
618: Physica C \textbf{299}, 31 (1998).
619: \bibitem{Irie} A. Irie, Y. Hirai, and G. Oya, Appl. Phys. Lett. \textbf{72},
620: 2159 (1998); {\it ibid} \textbf{57}, 13399 (1998).
621: \bibitem{Yurgens} A. Yurgens, D. Winkler, T. Claeson, G. Yang, I. F.
622: G. Parker, and C. E. Gough, Phys Rev B \textbf{59},
623: 7196 (1999).
624: \bibitem{Latyshev} Y. I. Latyshev, T. Yamashita, L. N. Bulaevskii, M.
625: J. Graf, A. V. Balatsky, and M. P. Maley, Phys. Rev. Lett.
626: \textbf{82}, 5345 (1999).
627: \bibitem{Matsuda} Y. Matsuda, M. B. Gaifullin, K. Kumagai, K. Kadowaki, and
628: T. Mochiku, Phys. Rev. Lett. \textbf{75}, 4512 (1995).
629: \bibitem{Tsui} O. K. C. Tsui, N. P. Ong, and J. B. Peterson, Phys.
630: Rev. Lett \textbf{76}, 819 (1996).
631: \bibitem{Bulaevski1} L. N. Bulaevskii, D. Dominguez, M. P. Maley, A. R.
632: Bishop, O. K. C. Tsui, and N. P. Ong, Phys. Rev. B \textbf{54}, 7521 (1996).
633: \bibitem{Matsuda2} Y. Matsuda, M. B. Gaifullin, K. Kumagai, M. Kosugi, and
634: K. Hirata, Phys. Rev. Lett. \textbf{78}, 1972 (1997).
635: \bibitem{Hanaguri} T. Hanaguri, Y. Tsuchiya, and A. Maeda, Phys. Rev.
636: B \textbf{58},
637: R8929 (1998).
638: \bibitem{Gaifulin} M. B. Gaifullin, Y. Matsuda, N. Chikumoto, J. Shimoyama,
639: K. Kishio, and R. Yoshizaki, Phys. Rev. Lett. \textbf{83}, 3928 (1999).
640: \bibitem{Rapp} M. Rapp, A. Murk, R. Semerad, and W. Prusseit, Phys.
641: Rev. Lett. \textbf{77}, 928 (1996).
642: \bibitem{Basov} D. N. Basov, T. Timusk, B. Dabrowski, and J. D.
643: Jorgensen, Phys. Rev. B \textbf{50}, 3511 (1994).
644: \bibitem{Homes} C. C. Homes, S. V. Dordevic, D. A. Bonn, R. Liang, W.
645: A. Hardy, and T. Timusk, Phys. Rev. B
646: \textbf{71}, 184515 (2005).
647: \bibitem{Cubitt} R. Cubitt and E. M. Forgan, Nature (London)
648: \textbf{365}, 407 (1993).
649: \bibitem{Matsuda3} Y. Matsuda, M. B. Gaifullin, K. Kumagai, K.
650: Kadowaki, T. Mochiku, and K. Hirata,
651: Phys. Rev. B \textbf{55}, R8685 (1997).
652: \bibitem{Sandu} V. Sandu, E. Cimpoiasu, T. Katuwal, Shi Li, M. B. Maple,
653: and C. C. Almasan, Phys. Rev. Lett. \textbf{93}, 177005 (2004).
654: \bibitem{Katuwal} T. Katuwal, V. Sandu, E. Shi Li, M. B. Maple,
655: and C. C. Almasan, Phys. Rev. B \textbf{72}, 174501 (2005).
656: \bibitem{Paulius} L. M. Paulius, B. W. Lee, M. B. Maple, and P. K. Tsai,
657: Physica C \textbf{230}, 255(1994).
658: \bibitem{Levin} G. A. Levin, T. Stein, C. N. Jiang, C. C. Almasan, D. A.
659: Gajewski,S. H. Han, and M. B. Maple, Physica C \textbf{282-287}, 1147 (1997).
660: \bibitem{Damascelli} A. Damascelli, Z. Hussain, and Z.-H. Shen, Rev.
661: Mod. Phys. \textbf{75}, 473 (2003); D. N. basov and T.
662: Timusk, Rev. Mod. Phys. \textbf{77}, 721 (2005).
663: \bibitem{Hosseini} A. Hosseini, D. M. Broun, D. E. Sheehy, T. P.
664: Davis, M. Franz, W. N. Hardy, R. Liang, and D. A. Bonn,
665: Phys. Rev. Lett. \textbf{93}, 107003 (2004).
666: \bibitem{Koshelev2} A. E. Koshelev, Phys. Rev. Lett. \textbf{76}, 1340 (1996).
667: \bibitem{Briceno} G. Brice\~{n}o, M. F. Crommie, and A. Zettl, Phys. Rev.
668: Lett. \textbf{66}, 2164 (1991).
669: \bibitem{Gray} K. E. Gray and D. H. Kim, Phys. Rev. Lett.
670: \textbf{70}, 1693(1993).
671: \bibitem{Hettinger} J. D. Hettinger, K. E. Gray, B. W. Veal, A. P.
672: Paulikas, P. Kostic, B. R. Washburn, W. C. Tonjes, and A. C. Flewelling,
673: Phys. Rev. Lett. \textbf{74}, 4726(1995).
674: \bibitem{Yoo} K.-H. Yoo, D. H. Ha, Y. K. Park, and J. C. Park, Phys.
675: Rev. B \textbf{49}, 4399 (1994).
676: \bibitem{Ambegaokar} V. Ambegaokar, and B. I. Halperin, Phys. Rev.
677: Lett. \textbf{22},
678: 1364(1969).
679: \bibitem{Daemen} L. L. Daemen, L. N. Bulaevskii, M. P. Maley, and J.
680: Y. Coulter, Phys. Rev. Lett.
681: \textbf{70}, 1167 (1993).
682: \bibitem{Koshelev3} A. E. Koshelev, L. N. Bulaevskii, and M. P. Maley, Phys.
683: Rev. Lett. \textbf{81}, 902(1998).
684: \bibitem{Koshelev4} A. E. Koshelev, Phys. Rev. Lett. \textbf{77}, 3901(1996).
685: \bibitem{Fistul} M. V. Fistul and G. F. Giuliani, Physica C
686: \textbf{289}, 291 (1997).
687: \bibitem{Logvenov} G. Yu. Logvenov, M. V. Fistul, and P. M\"{u}ller,
688: Phys. Rev. B \textbf{59}, 4524 (1999).
689: \bibitem{Koshelev1} A. E. Koshelev, L. N. Bulaevskii, and M. P. Maley,
690: Phys. Rev. B \textbf{62}, 14403 (2000).
691: \bibitem{Artemenko} S. N. Artemenko, L. N. Bulaevskii, M. P. Maley, V. M.
692: Vinokur, Phys. Rev. B \textbf{59}, 11587 (1999).
693: \bibitem{Vekhter} I. Vekhter, L. N. Bulaevskii, A. E. Koshelev, and M. P.
694: Maley, Phys, Rev. Lett. \textbf{84}, 1296 (2000).
695: \bibitem{Morozov} N. Morozov, L. Krusin-Elbaum, T. Shibauchi, L. N.
696: Bulaevskii, M. P. Maley, Yu. I. Latyshev, and T. Yamashita,
697: Phys. Rev. Lett. \textbf{84}, 1784(2000).
698: \bibitem{Artemenko2} S. N. Artemenko and A. G. Kobelkov, Phys. Rev.
699: Lett. \textbf{78}, 3551 (1997).
700: \bibitem{Goldobin} E. Goldobin and A. V. Ustinov, Phys. Rev. B
701: \textbf{59}, 11532 (1999).
702: \bibitem{Abrikosov} A. A. Abrikosov, Phys. Rev B \textbf{57}, 7488 (1998).
703:
704:
705: \end{thebibliography}
706:
707:
708: \newpage
709: \begin{figure}[tbp]
710: \caption{Temperature $T$ dependence of out-of-plane resistivity $\rho_{c}$
711: of an Y$_{0.54}$Pr$_{0.46}$Ba$_{2}$Cu$_{3}$O$_{7-\delta}$ single
712: crystal, measured in an
713: applied magnetic field of 0, 6, 8, 10, 12, and 14 T and for $T\leq
714: 80$ K. Insets: (top) Sketch of sample geometry and leads configuration.
715: (bottom) Zero field
716: $\rho_c(T)$ shown over the whole measured $T$ range. The solid lines
717: are guides to
718: the eye.}
719: \end{figure}
720:
721: \begin{figure}[tbp]
722: \caption{Angular $\theta$ dependence of normalized out-of-plane
723: resistivity $\rho(\theta)/\rho(0)$ of an
724: Y$_{0.54}$Pr$_{0.46}$Ba$_{2}$Cu$_{3}$O$_{7-\delta}$ single crystal,
725: measured at 30, 35,
726: and 40 K and 14 T.}
727: \end{figure}
728:
729: \begin{figure}[tbp]
730: \caption{(a) Plot of the out-of-plane resistivity $\rho_c$ vs
731: $H|cos\theta|$ of an Y$_{0.54}$Pr$_{0.46}$Ba$_{2}$Cu$_{3}$O$_{7-\delta}$
732: single crystal,
733: measured at 25, 30, 35, and 36 K and 6 T (open circles), 8 T (diamonds), 10 T (triangles), 12 T
734: (inverted triangles), and 14 T (open squares). Inset: Plot of
735: $\rho_c(H|\cos\theta|)$ measured at 40 K. (b) and its Inset: Same
736: plot of the data measured at 14 T
737: and 10 T, respectively. The solid lines are fits of the data with Eq. (7).}
738: \end{figure}
739:
740:
741: \begin{figure} [tbp]
742: \caption{Josephson critical current density $J_{c0}$, calculated with the
743: fitting parameters
744: obtained by fitting the angular
745: magnetoresistivity, vs reduced
746: temperature $T/T_{c0}$. The solid line is a power-law fit. Inset:
747: Temperature $T$ dependence of exponent $\nu$ (empty circles)
748: obtained by fitting the magnetoresistivity data. The theoretical $\nu(T)$
749: dependence is calculated with the $T$ dependence of the penetration depth
750: $\lambda_{ab}(T)$ given by the BCS theory and taking as the critical
751: temperature $T_{c0}=38$ K.}
752: \end{figure}
753: \end{document}
754:
755:
756:
757:
758:
759:
760:
761:
762:
763: