Citation PHYSICAL REVIEW LETTERS (2000), 85( RightCopyright 2000 American Physical So

Similar documents
arxiv:cond-mat/ v3 [cond-mat.supr-con] 23 May 2000

doi: /PhysRevLett

SUPPLEMENTARY INFORMATION

Tunneling Spectra of Hole-Doped YBa 2 Cu 3 O 6+δ

What's so unusual about high temperature superconductors? UBC 2005

arxiv:cond-mat/ v1 [cond-mat.supr-con] 14 May 1999

SUM RULES and ENERGY SCALES in BiSrCaCuO

Hole-concentration dependence of band structure in (Bi,Pb) 2 (Sr,La) 2 CuO 6+δ determined by the angle-resolved photoemission spectroscopy

Absence of Andreev reflections and Andreev bound states above the critical temperature

Crossover from phase fluctuation to amplitudedominated superconductivity: A model system

Dynamics of fluctuations in high temperature superconductors far from equilibrium. L. Perfetti, Laboratoire des Solides Irradiés, Ecole Polytechnique

The pseudogap in high-temperature superconductors: an experimental survey

Can superconductivity emerge out of a non Fermi liquid.

arxiv:cond-mat/ v1 [cond-mat.supr-con] 9 Jul 2001

Dependence of the gap parameter on the number of CuO 2 layers in a unit cell of optimally doped BSCCO, TBCCO, HBCCO and HSCCO

BSCCO Superconductors: Hole-Like Fermi Surface and Doping Dependence of the Gap Function

ARPES studies of cuprates. Inna Vishik Physics 250 (Special topics: spectroscopies of quantum materials) UC Davis, Fall 2016

Universal scaling relation in high-temperature superconductors

The Chemical Control of Superconductivity in Bi 2 Sr 2 (Ca 1 x Y x )Cu 2 O 8+±

Supplementary Figures

Isotope Effect in High-T C Superconductors

The Hubbard model in cold atoms and in the high-tc cuprates

Key words: High Temperature Superconductors, ARPES, coherent quasiparticle, incoherent quasiparticle, isotope substitution.

Fine Details of the Nodal Electronic Excitations in Bi 2 Sr 2 CaCu 2 O 8+δ

C. C. Tsuei IBM T.J. Watson Research Center Yorktown Heights, NY 10598

μsr Studies on Magnetism and Superconductivity

High temperature superconductivity

High-T c superconductors

arxiv:cond-mat/ v1 [cond-mat.supr-con] 23 Feb 1999

arxiv: v2 [cond-mat.supr-con] 24 Aug 2012

What is strange about high-temperature superconductivity in cuprates?

A momentum-dependent perspective on quasiparticle interference in Bi 2 Sr 2 CaCu 2 O 8+δ

Large-scale Simulation for a Terahertz Resonance Superconductor Device

Temperature and Spatial Dependence of the Superconducting

Aspects of of Intrinsic Josephson Tunneling

Tunneling Spectroscopy of PCCO

arxiv:cond-mat/ v1 8 Mar 1995

Real Space Bogoliubov de Gennes Equations Study of the Boson Fermion Model

Pressure Effect in Bi-2212 and Bi-2223 Cuprate Superconductor

Pseudogap and Conduction Dimensionalities in High-T c. Superconductors arxiv:cond-mat/ v1 [cond-mat.supr-con] 7 Mar 2002.

High-Temperature Superconductors: Playgrounds for Broken Symmetries

Rebuttal to Comment by V.M. Krasnov on Counterintuitive consequence. of heating in strongly-driven intrinsic junctions of Bi 2 Sr 2 CaCu 2 O 8+d Mesas

Spatial Symmetry of Superconducting Gap in YBa 2 Cu 3 O 7-δ Obtained from Femtosecond Spectroscopy

Superconducting Single-photon Detectors

Quantum Melting of Stripes

Tuning order in cuprate superconductors

Origin of the anomalous low temperature upturn in resistivity in the electron-doped cuprates.

Intrinsic Josephson junctions: recent developments

arxiv:cond-mat/ v1 8 May 1997

arxiv: v1 [cond-mat.supr-con] 29 Oct 2010

Electronic Noise Due to Thermal Stripe Switching

Periodic Oscillations of Josephson-Vortex Flow Resistance in Oxygen-Deficient Y 1 Ba 2 Cu 3 O x

epl draft M. L. Teague 1,A.D.Beyer 1,M.S.Grinolds 1,S.I.Lee 2 and N.-C. YEH 1 Korea

requires going beyond BCS theory to include inelastic scattering In conventional superconductors we use Eliashberg theory to include the electron-

Pseudogap and the chemical potential position topology in multiband superconductors

A BCS Bose-Einstein crossover theory and its application to the cuprates

arxiv: v1 [cond-mat.supr-con] 7 May 2011 Edge states of Sr 2 RuO 4 detected by in-plane tunneling spectroscopy

Supplementary figures

YBCO. CuO 2. the CuO 2. planes is controlled. from deviation from. neutron. , blue star for. Hg12011 (this work) for T c = 72

Temperature dependence of spin diffusion length in silicon by Hanle-type spin. precession

2nd Annual International Conference on Advanced Material Engineering (AME 2016)

An unusual isotope effect in a high-transition-temperature superconductor

Vortex Checkerboard. Chapter Low-T c and Cuprate Vortex Phenomenology

arxiv:cond-mat/ v2 [cond-mat.supr-con] 26 Jul 2006 Doping dependent optical properties of Bi 2 Sr 2 CaCu 2 O 8+δ

Effect of the magnetic resonance on the electronic spectra of high-t c superconductors

F. Rullier-Albenque 1, H. Alloul 2 1

Conference on Superconductor-Insulator Transitions May 2009

Recent Advances in High-Temperature Superconductivity

Superconductivity and Electron Correlations in Ruthenates

LB [1] have calculated T c for a 2D conductor having logarithmic singularities near the FL, a simple

Low energy excitations in cuprates: an ARPES perspective. Inna Vishik Beyond (Landau) Quasiparticles: New Paradigms for Quantum Fluids Jan.

Visualizing the evolution from the Mott insulator to a charge-ordered insulator in lightly doped cuprates

Superconductivity. Superconductivity. Superconductivity was first observed by HK Onnes in 1911 in mercury at T ~ 4.2 K (Fig. 1).

arxiv: v1 [cond-mat.supr-con] 4 Jul 2016

Universal valence-band picture of. the ferromagnetic semiconductor GaMnAs

Fermi Surface Reconstruction and the Origin of High Temperature Superconductivity

Scanning Tunnelling Microscopy Observations of Superconductivity

SUPPLEMENTARY INFORMATION

ARPES studies of c-axis intracell coupling in Bi 2 Sr 2 CaCu 2 O 8þd

Twenty years have passed since the discovery of the first copper-oxide high-temperature superconductor

The High T c Superconductors: BCS or Not BCS?

Origin of time reversal symmetry breaking in Y 1-y Ca y Ba 2 Cu 3 O 7-x.

Magnetic penetration depth in superconducting La 2 x Sr x CuO 4 films

An infrared study of the Josephson vortex state in high-t c cuprates

Splitting of a Cooper pair by a pair of Majorana bound states

Electron Doped Cuprates

b imaging by a double tip potential

Strongly Correlated Systems:

Using Disorder to Detect Order: Hysteresis and Noise of Nematic Stripe Domains in High Temperature Superconductors

Anisotropy in the ab-plane optical properties of Bi 2 Sr 2 CaCu 2 O 8 single-domain crystals

Coexistence of Fermi Arcs and Fermi Pockets in High Temperature Cuprate Superconductors

Principles of Electron Tunneling Spectroscopy

arxiv:cond-mat/ v2 [cond-mat.str-el] 25 Jun 2003

The change in conductivity anisotropy due to 1 superconductivity onset in the form of rare isolated islands: the theory and its application to FeSe

SESSION 3. Lattice fluctuations and stripes - I. (September 26, 2000) S3-I S3-II S3-III S3-IV S3-V. E. Kaldis New aspects of Ca doping in 123

High-T c superconductors. Parent insulators Carrier doping Band structure and Fermi surface Pseudogap and superconducting gap Transport properties

The pseudogap state in high-t c superconductors: an infrared study

I.3 MODELLING CUPRATE GAPS IN A COMPOSITE TWO-BAND MODEL 1. INTRODUCTION

Talk online at

Hall effect in the normal state of high T c cuprates

Transcription:

Title Discriminating the superconducting Bi2Sr2CaCu2O8+delta by interlayer t Author(s) Suzuki, M; Watanabe, T Citation PHYSICAL REVIEW LETTERS (2), 85( Issue Date 2-11-27 URL http://hdl.handle.net/2433/39919 RightCopyright 2 American Physical So Type Journal Article Textversion publisher; none Kyoto University

VOLUME 85, NUMBER 22 P H Y S I C A L R E V I E W L E T T E R S 27 NOVEMBER 2 Discriminating the Superconducting Gap from the Pseudogap in Bi 2 Sr 2 CaCu 2 O 81d by Interlayer Tunneling Spectroscopy Minoru Suzuki 1 and Takao Watanabe 2 1 Department of Electronic Science and Engineering, Kyoto University, Kyoto 66-851, Japan 2 NTT Basic Research Laboratories, Nippon Telegraph and Telephone Corporation, 3-1 Morinosato, Wakamiya, Atsugi, Kanagawa 243-198, Japan (Received 1 March 2) Tunneling spectroscopy using a very thin stack of intrinsic Josephson junctions has revealed that the superconducting gap is definitely different from the pseudogap in the Bi 2 Sr 2 CaCu 2 O 81d system. In the underdoped region, the conductance peak arising from the superconducting gap is independently observed in the di dv-v curve and its position is much lower than that of the pseudogap. Near the optimum doping level and in the overdoped region, both peaks are located in close proximity. These findings are in conflict with a previous understanding of the pseudogap. PACS numbers: 74.72.Hs, 74.5. +r, 74.25.Jb It is now widely accepted that the electronic density of states in high-t c superconductors undergoes a decremental change in its spectrum around the Fermi level below a certain temperature T, which is much higher than T c. This change, called the pseudogap (PG) evolution [1], is observed almost commonly in high-t c superconductors. Since it is believed that PG is closely related to the pairing mechanism of high-t c superconductivity, the origin of PG and its relation to the superconducting gap (SG) have been attracting wide and continued interest both theoretically and experimentally. Angle resolved photoemission spectroscopy (ARPES) experiments [2] revealed that PG has the same d-wave symmetry as SG. Scanning tunneling spectroscopy (STS) experiments [3] suggested that SG smoothly connects with PG at T c with a sizable magnitude. These results have postulated a picture that the order parameter amplitude persists up to T high above T c, while the macroscopic phase coherence sets in only below T c [4,5]. While this picture appears persuasive, the T dependence of SG [6], penetration depth (l 22 ) [7], and the maximum Josephson current [8 1] also seem to indicate the disappearance of the order parameter amplitude at T c. If SG starts to evolve at T c, PG must be interpreted differently such as a spin excitation of a certain kind. Thus the elucidation of detailed behavior of SG and PG near T c is crucially important. In this Letter, we report the results of tunneling spectroscopy (TS) intended to discriminate SG from PG. In order to obtain a sufficient energy resolution and a clear energy structure in TS measurements, we employed superconductor-insulator-superconductor-type intrinsic Josephson junctions (IJJ) of Bi 2 Sr 2 CaCu 2 O 81d (BSCCO) [11]. The most important advantage of the use of IJJ for TS is that we can ascertain the T dependence of the c-axis resistivity r c of the very portion to be probed. With this means, we can estimate the doping level almost exactly [12]. We have measured more than 2 specimens with different doping levels. From the results, we have deduced the systematic behavior of SG and PG. The major consequence is that SG (order parameter magnitude) disappears at T c and is definitely distinct from PG. Specimens used for the interlayer TS are very thin IJJ stacks made of BSCCO crystals with different doping levels. They were fabricated by forming a 1 or 2 mm square 15 to 2 nm thick mesa (approximately ten junctions connected in a series) on a cleaved surface of a BSCCO crystal grown by the traveling-solvent-floating-zone method [12] or by the self-flux method. Before the photolithograph process, a 25 nm thick Ag thin film and a 5 nm thick Au thin film were evaporated on the cleaved surface and then annealed at 43 45 ± C for 1 to 1.5 h in oxygen atmosphere or in vacuum, depending on the carrier doping level. The other fabrication processes were detailed in previous publications [8,13]. Current-voltage (I-V) characteristics were measured by the short pulse method [14] with 1 ms wide current pulses at a duty of.5%. The influence of heating due to selfinjection of current is less than 3% on the voltage scale for 2 mm square overdoped specimens, as detailed elsewhere [14]. Since smaller junction size reduces the heating, the influence of heating is expected to be much less than 3% for 1 mm square junctions and for underdoped specimens, in particular. In this method, the maximum applied voltage was approximately 1.5 V. When the pulse voltage was higher, the specimens were destroyed by the electric field surrender during measurements. This limited the voltage range to less than approximately 12 mv for a single junction. Figure 1 shows r c -T characteristics for four specimens with different doping levels. It is known that in the BSCCO system there is a clear and nearly unique relationship [12] between the doping level d and the ratio r rc max rc 3 K, where rc max is the maximum r c just above T c and rc 3 K is r c at 3 K. The relationship is expressed as d.174 1.321 r 1 1.932 with an error of Dd 6.5. Thus, the doping levels of the specimens in Figs. 1(a) 1(d) were estimated as d.22 (underdoped),.24,.25 (near optimum), and.28 (overdoped), 31-97 85(22) 4787(4)$15. 2 The American Physical Society 4787

VOLUME 85, NUMBER 22 P H Y S I C A L R E V I E W L E T T E R S 27 NOVEMBER 2 ρ c (Ω cm) 3 2 1 6 5 4 3 2 1 δ =.22 δ=.24 N=12 δ=.25 (a) (c) N=13 1 2 T (K) N=7 δ=.28 (b) (d) N=14 1 8 6 4 2 1 2 3 5 4 3 2 1 ρ c ( Ωcm ) FIG. 1. Temperature dependence of the c-axis resistivity calculated from the stack resistance R c for four specimens having different carrier doping levels: (a) d.22 (underdoped), (b) d.24, (c) d.25 (near optimum), (d) d.28 (overdoped). The dashed line in (c) is the contact resistance inferred from the oscilloscope I-V measurements. respectively, as indicated in Fig. 1 together with a value for T c. The inset of Fig. 2 displays an oscilloscope image of the I-V characteristics for the specimen of Fig. 1(c), whose doping level is near the optimum. The characteristics exhibit multiple resistive branches, from which we determined the value of N 13 for the number of junctions. The maximum Josephson current is rather homogeneous except for one junction which is probably located in contact with the Ag Au electrode on the top. Since the I-V characteristics are measured for all the junctions in series, the influence of the outermost junction is negligible. The I (ma) 2 1-1 -2 δ=.25 7 K 1 K 85 K T c =85 K T=17 K 11 K -1-5 5 1 V (mv) I (ma) 2-2 -4 -.8 N=13 -.4 V ( V ).4 6. K FIG. 2. I-V characteristics at different temperatures for the specimens in Fig. 1(c). The inset displays the oscilloscope I-V characteristics measured at 6 K, in which the lower branch is used for the tunneling spectroscopy. -.8 numbers of junctions determined in this way for the specimens in Figs. 1(a) 1(d) are 12, 7, 13, and 14, respectively. The main panel of Fig. 2 shows the I-V characteristics measured by the short pulse method for the same specimen. In this I-V curve, the gap position is not necessarily clear unlike the case of overdoped specimens [6]. Furthermore, the gap structure in the I-V curve is much less discernible for underdoped specimens. Figures 3 show four sets of di dv-v characteristics obtained numerically from the short pulse I-V characteristics for the same specimens shown in Fig. 1 at various temperatures from 1 to 18 K. The values for V are for a single junction. The thick curves indicate the results obtained at a temperature very close to T c. Each curve is shifted vertically for an appropriate amount for convenience. For the specimen in Fig. 3(a), the excess conductance at V V is approximately 4.8% of the di dv value at 18 K and V 1 mv. Those for specimens in Figs. 3(b) to 3(d) are 8.4%, 3.4%, and 12%, respectively. Thus, the influence of the excess conductance on the electronic density spectra seen in Fig. 3 is regarded as very small. When we turn to di dv-v curves for T. T c in Fig. 3, it is clearly seen that they exhibit a significant gap structure even above T c. This strongly T dependent structure is presumed to be due to a PG observed for the BSCCO system extensively by various methods, particularly by ARPES [2,5] and STS [3,15]. Then, the di dv-v curves indicate that PG starts to evolve at a higher than 2 K temperature for specimens in Figs. 3(a) 3(c) (underdoped and optimum), while it starts to evolve at 14 K for the specimen in Fig. 3(d) (overdoped). The PG magnitude is basically reflected by the depth and the width of the di dv-v curve, e.g., at T c (thick line). It is evident from Fig. 3 that the PG magnitude systematically decreases as the doping level increases from Figs. 3(a) to 3(d). This tendency is consistent with the previous observations by ARPES and STS. In the present di dv-v characteristics of Figs. 3(a) and 3(b), the PG peak is located outside the maximum voltage due to the experimental limitation. Therefore, we approximately estimated the peak position by fitting the BCS model [15] with a finite quasiparticle relaxation time to the di dv-v curve near T c (thick line). The estimates of the peak position are 15 6 2 mv, 13 6 2 mv, 8 6 1 mv, and 3 6 5 mv, for Figs. 3(a) to 3(d), respectively. The PG magnitude decreases with the doping level d but still persists in the overdoped region. This is in accordance with the existence of r c upturn, which is clearly present in r c -T curves in the overdoped region. In Fig. 3(a), the di dv-v curve at 1 K exhibits a small peak at around V 8 mv, which changes from a small peak to a cusp as T increases and disappears at T c. This peak is superposed on the shoulder of the PG conductance peak, so that the T dependence of the peak position is not very clear. However, a close inspection reveals that it tends to shift toward lower energies as T increases. We observed a similar peak and behavior for all the specimens 4788

VOLUME 85, NUMBER 22 P H Y S I C A L R E V I E W L E T T E R S 27 NOVEMBER 2 FIG. 3. di dv-v curves at various temperatures for the specimens in Fig. 1. Each curve is shifted vertically for an appropriate value. The thick lines indicate the curve very close to T c. The normal tunneling resistance for the specimen in (d) exhibited a characteristic T dependence very similar to those in Ref. [6]. that were measured with a similar d value. Based on these observations, we can conclude that the peak near V 8 mv corresponds to SG. The SG magnitude 2D pp defined as half the peak separation at 1 K is 79 mev, which is also reasonably compared with the STS results [3,15,16]. The present result implies that SG disappears at T c and PG which exists above T c is distinct from SG. Figures 3(a) and 3(b) also indicate that the PG magnitude is much greater than SG in the underdoped region. The relationship between SG and PG is rather different from the ARPES and STS results in which SG connects smoothly at T c with PG. It appears that the smooth connection of both gaps reflects the behavior only to be observed near the optimum doping, as seen in Fig. 3(c). The present results also imply that the quasiparticle excitation spectrum has two energy scales. This is also argued in the underdoped YBa 2 Cu 3 O 72x [17 19] and La 22x Sr x CuO 4 [2,21] through different experimental probes. Figures 3(a) 3(d) show that, as the doping level increases, the superconducting peak becomes increasingly pronounced, the height becomes greater, and the peak sharper. The behavior reflects that the superfluid density increases significantly as the doping level increases from the underdoped to overdoped region. With these changes in the peak profile, 2D remains almost unchanged from 2D pp 79 mev at d.22 in the underdoped region to 4789

VOLUME 85, NUMBER 22 P H Y S I C A L R E V I E W L E T T E R S 27 NOVEMBER 2 2 pp (mev) 12 1 8 6 4 2 δ=.22 δ=.24 δ=.25 δ=.28 2 4 6 8 1 12 14 T (K) δ=.25 δ=.28 FIG. 4. T dependence of 2D pp for SG and, partly, for PG for the same specimens in Fig. 3. Open squares and triangles indicate PG values and the other symbols indicate SG peak-topeak values. 2D pp 77 mev near the optimum level of d.25. It then decreases to 2D pp 45 mev at d.28 in the overdoped region. In the overdoped region, the SG peak is much sharper than in the underdoped region and the T dependence of 2D pp is clearly visible. Figure 4 shows the T dependence of 2D pp for the four specimens together with values for the PG peaks that are visible. It is noted that the PG values for d.28 obtained by fitting are much smaller than the 2D pp values. It should be noted that the PG peak position also shifts to lower energies with the increasing doping level at a faster rate than 2D pp. Near the optimum doping level, SG and PG are located in rather close proximity. In this situation, both peaks are merged to form a single peak as Fig. 3(c) reflects. It can then occur that SG appears to connect smoothly with PG at T c near the optimum doping level. Indeed, in that doping range, the characteristics are very similar to the STS results [3,15]. In the overdoped region, the PG peak is less pronounced and the SG peak is dominant so that we observe rather conventional behavior for the evolution of SG. The T dependence of the peak position is not necessarily the same as, but comparable to the BCS T dependence of the order parameter which disappears at T c. In spite of many STS measurements of the BSCCO system, reports on the di dv-v curve similar to Figs. 3(a) or 3(b) are astonishingly rare. Oda et al. [16] observed a similar di dv-v curve for an underdoped BSCCO crystal, although they attributed the shoulder of the pseudogap to a large excess conductance which happened to accompany their characteristics. The rareness of such data might be due to an unknown shift of the oxygen doping level towards the optimum level, which might occur in the doping dependence measurements by surface spectroscopy. Concerning the origin of PG, several models were proposed [4,22,23], among which the fluctuation model might be in conflict with the present result, since SG evolves independently from PG with a different gap energy. This is at variance with the argument that the tunneling probes the single particle excitation and not the phase coherence and that PG is the order parameter amplitude. The present result is also at variance with a model which assumes the establishment of the macroscopic phase coherence at T c for bosons that are formed below T, which is much higher than T c. It is also in conflict with a model which assumes a smooth connection of SG and PG. After the submission of this manuscript, we noted a paper by Krasnov et al. [24] with the same conclusion. The authors thank Dr. Azusa Matsuda for very valuable discussions. [1] For a review, see T. Timusk and B. Statt, Rep. Prog. Phys. 62, 61 (1999). [2] H. Ding et al., Nature (London) 382, 51 (1996). [3] Ch. Renner et al., Phys. Rev. Lett. 8, 149 (1998). [4] V. J. Emery and S. A. Kivelson, Nature (London) 374, 434 (1995). [5] M. R. Norman et al., Nature (London) 392, 157 (1998). [6] M. Suzuki et al., Phys. Rev. Lett. 82, 5361 (1999). [7] A. Hosseini et al., Phys. Rev. Lett. 81, 1298 (1999). [8] M. Suzuki et al., Phys. Rev. Lett. 81, 4248 (1998). [9] M. Suzuki, T. Watanabe, and A. Matsuda, IEEE Trans. Appl. Supercond. 9, 4511 (1999). [1] M. B. Gaifullin et al., Phys. Rev. Lett. 83, 3928 (1999). [11] R. Kleiner et al., Phys. Rev. Lett. 68, 2394 (1991). [12] T. Watanabe et al., Phys. Rev. Lett. 79, 2113 (1997). [13] M. Suzuki et al., IEEE Trans. Appl. Supercond. 7, 2956 (1997). [14] M. Suzuki, T. Watanabe, and A. Matsuda, IEEE Trans. Appl. Supercond. 9, 457 (1999). [15] A. Matsuda et al., Phys. Rev. B 6, 1377 (1999). [16] M. Oda et al., Physica (Amsterdam) 281C, 135 (1997). [17] G. Deutscher, Nature (London) 397, 41 (1999). [18] D. Mihailovic et al., Phys. Rev. B 6, R6995 (1999). [19] J. Demsar et al., Phys. Rev. Lett. 82, 4918 (1999). [2] T. Sato et al., Phys. Rev. Lett. 83, 2254 (1999). [21] A. Ino et al., Phys. Rev. Lett. 81, 2124 (1998). [22] V. J. Emery et al., Phys. Rev. B 56, 612 (1997). [23] E. Delmer and S.-C. Zhang, Nature (London) 396, 733 (1998). [24] V. M. Krasnov et al., Phys. Rev. Lett. 84, 586 (2). 479