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

Similar documents
Dirac-Fermion-Induced Parity Mixing in Superconducting Topological Insulators. Nagoya University Masatoshi Sato

Specific Heat and Electrical Transport Properties of Sn 0.8 Ag 0.2 Te Superconductor

Superconductivity in Cu x Bi 2 Se 3 and its Implications for Pairing in the Undoped Topological Insulator

Surface Majorana Fermions in Topological Superconductors. ISSP, Univ. of Tokyo. Nagoya University Masatoshi Sato

arxiv: v1 [cond-mat.supr-con] 27 Jun 2016

Exotic Phenomena in Topological Insulators and Superconductors

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

Superconductivity and non-metallicity induced by doping the. topological insulators Bi 2 Se 3 and Bi 2 Te 3

SUPPLEMENTARY INFORMATION

Superconductivity with two fold symmetry in topological superconductor Sr x Bi 2 Se 3

arxiv: v1 [cond-mat.mes-hall] 29 Jul 2010

Transport through Andreev Bound States in a Superconductor-Quantum Dot-Graphene System

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

Supplementary figures

SUPPLEMENTARY INFORMATION

Topological Surface States Protected From Backscattering by Chiral Spin Texture

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

Massive Dirac Fermion on the Surface of a magnetically doped Topological Insulator

Visualizing Electronic Structures of Quantum Materials By Angle Resolved Photoemission Spectroscopy (ARPES)

Tunneling Spectroscopy of PCCO

Topological Insulators

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

A combined method for synthesis of superconducting Cu doped Bi 2 Se 3

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

High-Temperature Superconductors: Playgrounds for Broken Symmetries

Theory of Lifetime Effects in Point-Contacts: Application to Cd 2 Re 2 O 7

Superconductivity Induced Transparency

Topological Insulators

arxiv: v1 [cond-mat.mes-hall] 23 Sep 2011

Supplementary Figure S1: Number of Fermi surfaces. Electronic dispersion around Γ a = 0 and Γ b = π/a. In (a) the number of Fermi surfaces is even,

Topological Kondo Insulators!

Supplementary Figures

Supplementary Figures

Crystalline Symmetry and Topology. YITP, Kyoto University Masatoshi Sato

arxiv: v1 [cond-mat.supr-con] 12 Jul 2017

This article is available at IRis:

Out-of-equilibrium electron dynamics in photoexcited topological insulators studied by TR-ARPES

arxiv: v2 [cond-mat.str-el] 24 Nov 2009

arxiv: v1 [cond-mat.supr-con] 9 Aug 2011

arxiv: v1 [cond-mat.supr-con] 17 Dec 2009

Observation of Fermi-energy dependent unitary impurity resonances in a strong topological insulator Bi 2 Se 3 with scanning tunneling spectroscopy

Observation of topological surface state quantum Hall effect in an intrinsic three-dimensional topological insulator

SUPPLEMENTARY INFORMATION

Topological Insulators and Ferromagnets: appearance of flat surface bands

Superconducting fluctuations, interactions and disorder : a subtle alchemy

ARPES experiments on 3D topological insulators. Inna Vishik Physics 250 (Special topics: spectroscopies of quantum materials) UC Davis, Fall 2016

Topological Defects inside a Topological Band Insulator

Graphite, graphene and relativistic electrons

Graphene and Carbon Nanotubes

Study of the Potential Dirac Material Pb 0.9 Sn 0.1 Se by Soft Point-Contact Spectroscopy

arxiv: v1 [cond-mat.supr-con] 27 Feb 2014

InAs/GaSb A New Quantum Spin Hall Insulator

Observation of a robust zero-energy bound state in iron-based superconductor Fe(Te,Se)

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

Topological insulator (TI)

SUPPLEMENTARY INFORMATION

Scientific Reports 3: 2016(2013)

Superconductivity at nanoscale

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

arxiv: v1 [cond-mat.supr-con] 18 Jul 2016

Physics in two dimensions in the lab

Two-Fold-Symmetric Magnetoresistance in Single Crystals of Tetragonal BiCh2- Based Superconductor LaO 0.5F 0.5BiSSe

Influence of tetragonal distortion on the topological electronic structure. of the half-heusler compound LaPtBi from first principles

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

PC-10-INV Superconducting States in Doped Topological Materials

Scanning Tunneling Microscopy & Spectroscopy: A tool for probing electronic inhomogeneities in correlated systems

doi: /PhysRevLett

Scanning Tunneling Microscopy Studies of Topological Insulators Grown by Molecular Beam Epitaxy

Quantum Condensed Matter Physics Lecture 12

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

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

FIG. 1: (Supplementary Figure 1: Large-field Hall data) (a) AHE (blue) and longitudinal

Scanning Tunneling Microscopy/Spectroscopy

Electron transport through Shiba states induced by magnetic adsorbates on a superconductor

InAs/GaSb A New 2D Topological Insulator

Multichannel Majorana Wires

Topological protection, disorder, and interactions: Life and death at the surface of a topological superconductor

Scanning Tunnelling Microscopy Observations of Superconductivity

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

STM studies of impurity and defect states on the surface of the Topological-

SUPPLEMENTARY INFORMATION

New Quantum Transport Results in Type-II InAs/GaSb Quantum Wells

Gate-tuned normal and superconducting transport at the surface of a topological insulator

arxiv: v1 [cond-mat.str-el] 16 Jun 2013

File name: Supplementary Information Description: Supplementary Figures and Supplementary References. File name: Peer Review File Description:

Crossover from phase fluctuation to amplitudedominated superconductivity: A model system

arxiv: v2 [cond-mat.supr-con] 3 Aug 2013

Phonon-induced topological insulating phases in. group IV-VI semiconductors

Topological nonsymmorphic crystalline superconductors

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

Supplementary Materials for

Time Reversal Invariant Ζ 2 Topological Insulator

Introductory lecture on topological insulators. Reza Asgari

Time - domain THz spectroscopy on the topological insulator Bi2Se3 (and its superconducting bilayers)

Electron Doped Cuprates

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

Journal of the Korean Magnetic Resonance Society 2003, 7, Kwangju, , KOREA Received September 29, 2003

Topological Kondo Insulator SmB 6. Tetsuya Takimoto

Observation of Unconventional Quantum Spin Textures in Topologically Ordered Materials

A Short Introduction to Topological Superconductors

Transcription:

Point contact spectroscopy of Cu 0.2 Bi 2 Se 3 single crystals arxiv:1111.5805v2 [cond-mat.supr-con] 24 Aug 2012 T. Kirzhner, 1 E. Lahoud, 1 K.B. Chaska, 1 Z. Salman, 2 and A. Kanigel 1 1 Physics Department, Technion-Israel Institute of Technology, Haifa 32000, Israel 2 Paul Scherrer Institute, CH-5232 Villigen PSI, Switzerland (Dated: November 7, 2018) We report point contact measurements in high quality single crystals of Cu 0.2Bi 2Se 3. We observe three different kinds of spectra: (1) Andreev-reflection spectra, from which we infer a superconducting gap size of 0.6mV; (2) spectra with a large gap which closes above T c at about 10K; and (3) tunneling-like spectra with zero-bias conductance peaks. These tunneling spectra show a very large gap of 2meV (2 /K b T c 14). INTRODUCTION One of the most promising directions in the fastgrowing field of topological phases of matter is the subject of Topological Superconductors [1, 2]. Closely related to topological insulators, topological superconductors have non-trivial topological properties and in addition are bulk superconductors [3]. The topological superconductors caused a lot of excitement since they are predicted to host Majorana fermions which could be used for quantum computing [4]. The recently discovered Cu x Bi 2 Se 3 superconductor is a prime candidate for being a topological superconductor [5]. Superconductivity in this compound is achieved by intercalating Cu between the layers of Bi 2 Se 3, which is an archetypical topological insulator. The intercalated Cu acts as an electron dopant leading to superconductivity below 3.5K. In this paper, we study the pairing symmetry and the topological features of the Cu x Bi 2 Se 3 crystals using point-contact spectroscopy. Differential conductance measurement of point contacts is a common and effective tool to determine the pairing symmetry of superconductors. Our point-contact measurements showed signs of unconventional superconductivity and a possible realization of topological superconductivity. EXPERIMENTAL DETAILS High-quality single crystals of Cu 0.2 Bi 2 Se 3 were prepared using the modified Bridgman method. High-purity Cu, Bi and Se, were mixed and sealed in an evacuated fused silica ampoule. The ampoule is warmed up to 850 C in an inclined tube furnace and then cooled down at a rate of 2.5 C/h to 560 C. The ampoule is held at this temperature for 24 hours and quenched to room temperature. This process is repeated twice with an intermediate grinding. The result is a boule which is easily cleaved normal to the (001) direction of the crystal. Cu 0.2 Bi 2 Se 3 is notorious for being a tricky sample to prepare. This is probably related to the am-bipolar nature of the Cu as a dopant [6]. We found that the sample R(W) 0.003 0.002 0.001 Binding energy(ev) 0.00 0 (a) (c) 0.0-0.3-0.4-0.5 1 2 3 4 T (K) o k x (A) -1 Volume Fraction(%) 0 5 10 15 20 25 30 5 6 0 (b) DeF 1 2 3 T (K) Bi 2Se3 (d) Cu0.2Bi 2Se3 0.0 0.1 o k x (A) -1 4 0.0 0.1 0.2 FIG. 1: (a) Resistivity data showing a zero resistance at T c = 3.3K. (b) SQUID magnetization data showing T c=3.3k and a SC volume fraction of 17% at 1.8K. (c),(d) ARPES data taken along the Γ K direction at 25K with photon energy of 22eV. The data was measured in the PGM beam-line at the SRC Madison, WI. Panel (c) shows the data for a pristine Bi 2Se 3 sample and panel (d) for the same Cu 0.2Bi 2Se 3 sample used in the point-contact experiments. A large change in the chemical potential can be seen. is very sensitive to the temperature gradient in the furnace. Resistivity measurement shows a zero resistance at T c = 3.3K, as seen in Fig 1(a). Superconducting quantum interference device (SQUID) magnetometry shows these crystals to be superconducting below T c = 3.3K. The shielding volume fraction is about 17% at 2K, as can

2 d I/d V (1 /Ω) 1.6 1.5 1.4 1.3 1.2 1.1 1.0 0.9 0.8 0.0 1 T 1 T 2 T 0.7-5 -4-3 -2-1 0 1 2 3 4 5 FIG. 2: Differential conductance versus the voltage bias for the Au point contact at temperature of 1.8K at different magnetic fields. be seen in Fig 1(b). In order to study the Cu intercalation effect on the electronic properties of our sample we performed angleresolved photoemission spectroscopy (ARPES) measurements. A comparison between the ARPES spectra of a non-intercalated Bi 2 Se 3 sample grown using the same method and of the superconducting sample is shown in panels (c) and (d) of Fig 1. We show the dispersion along the K Γ K direction taken with 22eV photons. The n-type doping effect of the Cu intercalation is clearly visible in the figure. The Dirac point sunk down by about 180meV, indicating a large change in the chemical-potential of the sample. We can estimate the carrier concentration in the superconducting (SC) sample to be 2.2 10 20 cm 3. Nevertheless, the surface states are well preserved; we do not see mixing with the conduction band. One can see that the momentum width of the surface states of the two samples is similar, indicating that the topological protection of the surface state is still intact in the superconducting sample. We now turn to the main part of this work, the use of point contact spectroscopy to probe the order parameter of the Cu 0.2 Bi 2 Se 3 single crystals. A Au tip was used as the point contact. The tip was pressed against the sample surface, and the contact pressure was controlled by mechanical means. The conductance measurements were made using a four-probe technique, and a standard lock-in amplifier technique was used to measure the differential conductance. The quality of our point contacts was verified by measuring Andreev gaps of Nb films using the same system. The surface of the single crystals was imaged using an atomic force microscope after a point-contact measurement, and the contact size was estimated to be a few square microns. The large size of the point contacts increased our chances of contacting superconducting areas. The Cu 0.2 Bi 2 Se 3 single crystals were cleaved using a sticky tape, creating a fresh (001) surface. Immediately after the cleaving, the sample was cooled down to 1.8K and a differential conductance measurement of the point contacts was made. RESULTS AND DISCUSSIONS The resistance of the point contacts was in the range of 1 to 20Ω. About 15% of the point contacts showed enhancement of the conductance. The zero bias enhancement in all cases appeared below the critical temperature of the bulk superconductivity and the amplitude enhancement ranged between a few percent to about twice of the normal conductance. Application of a 2T field perpendicular to the (001) plane suppressed the zero-bias enhancement in all the point contacts. This result is in line with the reported value for the critical magnetic field in Cu 0.2 Bi 2 Se 3 [7]. The small success ratio of the point contacts can be explained by the low volume fraction of the superconductor. Figure 2 shows a differential conductance measurement of a point contact at a temperature of 1.8K. The zero-field curve shows a zero bias peak which starts around 1mV. The amplitude of the peak is about 1.9 times the conductance at higher bias. The Andreevreflection process is known to enhance the conductance of a superconductor-normal metal junction at energies below the energy gap of the superconductor to a value which is twice the conductance at energies above the gap. Our observation of the zero-bias enhancement seems to fit to a scenario of an Andreev-reflection process. High magnetic field suppresses the zero bias enhancement and at about 2T it disappears. Zero bias enhancement at point contacts can be also explained by heating effects [8] or by magnetic impurities [9], which both can be ruled out by application of low magnetic fields. As seen in Fig. 2 application of 100G did not have much effect on the zerobias width and height, reinforcing our conclusion that the origin of our zero bias peak is an Andreev-reflection process. Point contacts with different contact resistivity which showed zero bias enhancement, had a common energy scale of 1mV, suggesting that the superconducting energy gap size is around 1mV. Figure 3(a) shows a differential conductance of point contact with a resistance of 2.5Ω. In this spectra, the zero bias amplitude is about 1%, which is much lower than in Fig. 2. The low amplitude can be explained by the large size of our point contacts and the low volume fraction of the superconductor. The large point contact consists of many conduction channels, where many of them are just plain metal-metal junctions and only a small part is a metal-superconductor junction, and this may explain the low amplitude of the Andreev peaks. The zero field curve shows an Andreev-reflection like feature, which sits

3 d I/d V (1 /Ω) 0.4 0 8 0.4 0 4 0.4 0 0 (a ) d I/d V (1 /O h m ) 0.2 4 0.2 3 0.2 2-1 5-1 0-5 0 5 1 0 1 5 m V 1.0 2 0 1.0 1 6 1.0 1 2 1.0 0 8 1.0 0 4 (b ) E x p e rim e n t s -w a v e B T K fit 0.7 0.6 0.5 0.4 0.3 0.2 0.1 E x p e rim e n t B C S F it 0.0 0.0 0.5 1.0 1.5 2.0 2.5 3.0 T (K ) 0.3 9 6-5 -4-3 -2-1 0 1 2 3 4 5 FIG. 3: (a) Differential conductance versus the voltage bias for the Au point contact at a temperature of 1.8K at different magnetic fields. Inset: Differential conductance of a point contact on top of a non superconducting spot on the sample. (b) Differential conductance of a point contact with a subtracted background and a fit to an s-wave BTK model with the following parameters: = 0.5 mv, and Z = 0 with a quasiparticle broadening parameter of Γ = 0.23 mv. Inset: The size of the superconducting gap, extracted from the BTK fit as a function of temperature and a fit to the data using a BCS model with a gap size of 0.625mV and T c of 2.65K. inside an asymmetric V-shaped background. An example spectra of point contact in a non superconducting region with a V-shaped background is shown in the inset of Fig. 3(a). This background persists well above the critical temperature of the superconductor. As seen in Fig. 3(a), application of a magnetic field of 0.5T perpendicular to the sample surface, has very little effect on the height and width of the peak, unlike the magnetic field dependence of the point contact in Fig. 2. The variation of the magnetic field dependence in different point contacts is not clear. In Fig. 3(b) we show a differential conductance curve of the same point contact after a background subtraction. We fit the data using an s-wave Blonder-Tinkham- Klapwijk (BTK) [10] model. The results give an energy gap of about 0.6 mv. The point contacts were measured at a temperature of 1.8K which is high compared to the size of the energy gap, causing broadening of the Andreev peak. We used a quasiparticle broadening parameter of 0.23mV in our fits. This broadening effect makes it impossible to infer from the Andreev data whether a fit to other BTK models with different order parameters, such as d-wave or p-wave, would give better results. Nevertheless, using a simple s-wave model we can get a reasonable estimate of the size of the superconductor energy gap. In the inset to Fig. 3(b) we show the SC gap extracted from various curves measured at different temperatures, including a fit to the data using the BCS model. From the fit we find a gap size of 0.625mV at zero temperature and a critical temperature of 2.6K. The gap value we find is consistent with the results of heat capacity measurements [7]. In Fig. 4(a) we show data of a different point contact, this contact had a higher resistance of 20Ω. In this spectra we find the characteristic depletion of density of states that appears in junctions with a non transparent barrier (high Z in the BTK model), but without coherence peaks at the gap edges. Within the gap we find a zero bias conductance enhancement. One possible interpretation of such conductance structure is that the observed zerobias conductance peak (ZBCP) is a manifestation of an Andreev bound state originating from the surface of an unconventional superconductor. The peak is not as narrow as in the case of a d-wave superconductor [11]. The shape of the spectrum with the peak and the dips around it, is similar to the expected spectrum in a chiral p-wave superconductor [12], where the Andreev bound state is dispersive and wide. Fu and Berg showed in Ref [3], that the band structure of Cu x Bi 2 Se 3 may favor an unconventional odd-parity symmetry leading to a time-reversal odd-parity topological superconductor. The proposed superconductor is fully gapped in the bulk, but the surface has gapless surface Andreev bound states. Recently, Hseih and Fu calculated the density of states of the surface states and showed that a dispersive Andreev bound state should appear on the surface of the superconductor, which is consistent with the bound states we see in our point contact data [13]. They also claimed that the zero bound state seen in Cu x Bi 2 Se 3 crystals host Majorana fermions. Fig. 4(b) shows the temperature dependance of the differential conductance of the same point contact. The position of the dips did not change as the temperature was increased and the peak disappeared at about 2.1K. The

4 1.0 2 5 1.0 2 0 1.0 1 5 1.0 1 0 1.0 0 5 (a ) 0.2 T 1 T 2 T 1.0 1 5 1.0 1 0 1.0 0 5 (b ) T (K ) = 1.8 2 1.8 4 1.8 7 1.9 2 1.9 5 2.1 8 FIG. 4: (a) Differential conductance of a 20 Ω point contact at different magnetic fields. (b) Temperature dependence of the differential conductance. temperature dependence of the gap-size we find here is not BCS-like and can be an indication for an unconventional superconductivity. The height of the ZBCP was suppressed by a magnetic field as seen in Fig. 4(a), and it was suppressed completely at 2T. A strong magnetic field did not change the position of the dips, ruling out heating effects [8]. The dips in the measured spectra start at 2mV, suggesting a very large gap compared to the results inferred from the Andreev-reflection spectra. This gap size is very large compared to the T c of the crystal, giving a gap to critical temperature ratio of 2 /K b T c 14 (taking T c = 3.3K). Such a large gap size makes the scenario of a ZBCP coming from unconventional superconductivity less likely. The observed ZBCP could, in principle, be the result of two parallel channels, one with a transparent barrier with an Andreev peak and one channel with a larger gap. The width of the ZBCP in Fig. 4 is very close to the width of the Andreev peaks we have shown in previous figures, strengthening the proposed scenario. We had observed many spectra with large gaps in many point contacts, with size ranging from 2mv to 5mv. In Fig. 5 we see an example of such large gap with a size of 5mV, which closes above 10K. In Fig. 4 we can observe that at high magnetic fields and at high temperature the ZBCP vanishes, revealing a large gap of 2mV. The sharpness of the revealed gap is less pronounced than in the low temperature regime, not ruling out completely the existence of a ZBCP with dips. Such large gaps can also be seen in the background of Fig. 3(a). Our results partly agree with the recent results of Sasaki et al. [14], but there are few differences. First, they measured point contacts of Cu x Bi 2 Se 3 crystals which were made using an electrochemical intercalation method. Second, they observed a ZBCP which vanished at a lower temperature of 1K and the gap size was 1.1 0 1.0 8 1.0 6 1.0 4 1.0 2 1.0 0-1 5-1 0-5 0 5 1 0 1 5 T (K ) = 2.5 6 3.0 4 4.1 8 5.4 3 7.7 1 8.7 2 1 1.7 2 1 8.1 8 2 5.3 3 FIG. 5: Differential conductance measurements made at high temperatures, showing a large gap. smaller ( 0.6mV ) compared to the gap size we have observed in our ZBCP spectra. The inferred gap in their measurements agrees with our Andreev spectra. CONCLUSIONS To summarize, we measured the differential conductance of single crystals of Cu 0.2 Bi 2 Se 3 using a point contact. We find three different kinds of spectra. The first kind consists of an Andreev-like peak, a fit to a BTK model givne a gap size of 0.6meV, which is in agreement with the observations of other groups. The second kind are spectra showing a large gap, 2-5meV, which persist above T c. The third kind, which is the most interesting, are tunneling-like spectra with a ZBCP. The position of

5 the dips in the tunneling spectra implies a very large gap size. These spectra can be the result of an Andreev bound state in an unconventional superconductor, but can also be a result of two parallel conductance channels: an Andreev-like one and one having a large gap. Using a scanning tunneling microscope, it will probably be possible to distinguish between these two scenarios. We are grateful to G. Koren for helpful discussions and for the critical reading of this manuscript. This work was supported by the Israeli Science Foundation. The Synchrotron Radiation Center is supported by NSF Grant No. DMR-0084402. [1] A. P. Schnyder, S. Ryu, A. Furusaki, and A. W. W. Ludwig, Phys. Rev. B 78, 195125 (2008). [2] A. Kitaev, AIP Conf. Proc.No. 1134 (2009). [3] L. Fu and E. Berg, Phys. Rev. Lett. 105, 097001 (2010). [4] L. Fu and C. L. Kane, Phys. Rev. Lett. 100, 096407 (2008). [5] Y. S. Hor, A. J. Williams, J. G. Checkelsky, P. Roushan, J. Seo, Q. Xu, H. W. Zandbergen, A. Yazdani, N. P. Ong, and R. J. Cava, Phys. Rev. Lett. 104, 057001 (2010). [6] J. H. A. Vasko, L. Tichy and J. Weissenstein, App. Phys. A 5, 217 (1974). [7] M. Kriener, K. Segawa, Z. Ren, S. Sasaki, and Y. Ando, Phys. Rev. Lett. 106, 127004 (2011). [8] P. Xiong, G. Xiao, and R. B. Laibowitz, Phys. Rev. Lett. 71, 1907 (1993). [9] J. A. Appelbaum, Phys. Rev. 154, 633 (1967). [10] G. E. Blonder, M. Tinkham, and T. M. Klapwijk, Phys. Rev. B 25, 4515 (1982). [11] S. Kashiwaya and Y. Tanaka, Reports on Progress in Physics 63, 1641 (2000). [12] M. Yamashiro, Y. Tanaka, and S. Kashiwaya, Phys. Rev. B 56, 7847 (1997). [13] T. H. Hsieh and L. Fu, Phys. Rev. Lett. 108, 107005 (2012). [14] S. Sasaki, M. Kriener, K. Segawa, K. Yada, Y. Tanaka, M. Sato, and Y. Ando, Phys. Rev. Lett. 107 (2011).