Next Article in Journal / Special Issue
Simple Model for Tc and Pairing Symmetry Changes in Sr2RuO4 Under (100) Uniaxial Strain
Previous Article in Journal
Two-Band Electronic Reconstruction Induced via Correlation and CDW Order Effects
Previous Article in Special Issue
Josephson Critical Currents and Related Effects in Ultracold Atomic Superfluid Sytems
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Kondo Versus Fano in Superconducting Artificial High-Tc Heterostructures

1
Institute of Crystallography, Italian National Research Council, IC-CNR, Via Salaria Km 29.300, 00015 Rome, Italy
2
Rome International Center for Materials Science Superstripes RICMASS, Via dei Sabelli 119A, 00185 Rome, Italy
3
Max Planck Institute for Solid State Research, Heisenbergstraße 1, 70569 Stuttgart, Germany
4
Department of Physics, Sapienza Università di Roma, Piazzale Aldo Moro, 5, 00185 Rome, Italy
5
Institute for Microelectronics and Microsystems, Italian National Research Council IMM-CNR, Via del Fosso del Cavaliere, 100, 00133 Rome, Italy
*
Authors to whom correspondence should be addressed.
Condens. Matter 2024, 9(4), 43; https://doi.org/10.3390/condmat9040043
Submission received: 14 October 2024 / Revised: 29 October 2024 / Accepted: 30 October 2024 / Published: 31 October 2024
(This article belongs to the Special Issue Superstripes Physics, 3rd Edition)

Abstract

:
Recently, the quest for high-Tc superconductors has evolved from the trial-and-error methodology to the growth of nanostructured artificial high-Tc superlattices (AHTSs) with tailor-made superconducting functional properties by quantum design. Here, we report the growth by molecular beam epitaxy (MBE) of a superlattice of Mott insulator metal interfaces (MIMIs) made of nanoscale superconducting layers of quantum confined-space charge in the Mott insulator La2CuO4 (LCO), with thickness L intercalated by normal metal La1.55Sr0.45CuO4 (LSCO) with period d. The critical temperature shows the superconducting dome with Tc as a function of the geometrical parameter L/d showing the maximum at the magic ratio L/d = 2/3 where the Fano–Feshbach resonance enhances the superconducting critical temperature. The normal state transport data of the samples at the top of the superconducting dome exhibit Planckian T-linear resistivity. For L/d > 2/3 and L/d < 2/3, the heterostructures show a resistance following Kondo universal scaling predicted by the numerical renormalization group theory for MIMI nanoscale heterostructures. We show that the Kondo temperature, TK, and the Kondo scattering amplitude, R0K, vanish at L/d = 2/3, while TK and R0K increase at both sides of the superconducting dome, indicating that the T-linear resistance regime competes with the Kondo proximity effect in the normal phase of MIMIs.

1. Introduction

It is known that high-Tc cuprate superconductors can now be engineered by relying on the quantum design of nanoscale artificial high-Tc superlattices (AHTSs) of quantum wells [1,2,3,4] that capture the key features of cuprates. The growth of these novel nanoscale non-conventional heterostructures has been guided by the quantum material design of an artificially modulation-doped [5] correlated electron gas. This design relies on the predictions of the Bianconi–Perali–Valletta (BPV) theory [6,7,8] describing the amplification of the critical temperature driven by Fano–Feshbach shape resonance [9,10,11,12] in two-gap superconductors. This occurs in the presence of spin–orbit coupling (SOC) at the interfaces of nanostructured materials (NsMs) exhibiting quantum-size effects. Recently, this approach has been applied to the material design of the AHTS samples studied in this work [1,2]. These three-dimensional AHTS heterostructures are formed by superconductor layers (S) of a stoichiometric modulation-doped Mott insulator La2CuO4 (LCO) of thickness L, intercalated with potential barriers of a normal metal (N) made of chemically overdoped non-superconducting La1.55Sr0.45CuO4 (LSCO) with ten repeats of period d variable within the 2.97 < d < 5.28 nm nanoscale range. The normal N units play the role of charge reservoirs and transfer the interface space charge into the superconducting Mott insulator units, forming a 3D superlattice of a 2D Mott insulator–metal interface (MIMI) of period d. The internal interface electric field at the SNS junctions induces Rashba SOC in the superconducting interface space charge in the S layers, which is split into two electronic components by quantum-size effects, where the lowest sub-band shows a large cylindrical Fermi surface with high Fermi velocity, while the upper sub-band exhibits a low Fermi velocity and an unconventional extended van Hove singularity generated by the SOC at the interface [1,2,3,4]. It has been established that these artificial samples exhibit the superconducting dome predicted by the BPV theory [1,2]. According to this theory, the critical temperature, TC, is a function of the geometrical parameter, L/d, with a maximum at the magic ratio, L/d = 2/3. At this ratio, the theory predicts that the Fano–Feshbach resonance enhances the superconducting critical temperature.
The key result of this work is the compelling evidence that the temperature-dependent sheet resistance of various AHTSs in the normal phase above the top of the superconducting dome, i.e., within the 50 K < T < 270 K range, evolves as the geometric L/d ratio changes. At the top of superconducting dome for the magic ratio L/d = 2/3 for optimum Tc [1,2], the normal phase of the AHTS exhibits T-linear resistivity, similar to what is observed in slightly overdoped cuprate perovskites [13,14]. On either side of the superconducting dome for AHTS samples with L/d < 2/3 and L/d > 2/3, the temperature-dependent normal electrical resistance is predicted and experimentally probed for the Kondo proximity effect in Mott insulator–metal interfaces with the single variable T/TK, where TK is the Kondo temperature. While other factors such as weak localization or electron–electron interactions may also increase resistance at low temperatures, the proximity Kondo effect is particularly notable for its universal scaling behavior, as described by the numerical renormalization group (NRG) theory expressed by the universal function. This Kondo effect has been found in heterostructures made of metal–Mott insulator interfaces, where spin–orbit coupling driven by a potential drop at the interface and the coexistence of itinerant and localized electronic components generate a unique proximity Kondo effect. This effect appears in the nanoscale heterostructures studied in this work, which are made of Mott insulator–metal interfaces [15,16,17,18,19]. While the large charge correlation gap due to the local Coulomb repulsion U prohibits the tunneling of electrons from the metal into a Mott insulator, the resonant spin-flip scattering opens a new channel for tunneling. In this manner, the metal ‘eats’ itself layer by layer into the Mott insulator. The peculiar temperature dependence of the proximity Kondo in MIMI resistivity was observed in two-component electronic systems like, e.g., SrTiO3/LaTiO3/SrTiO3 heterostructures [20].
Evidence of the Kondo scattering of itinerant electrons by localized electrons has been reported in systems with logarithmic van Hove singularities [21], kagome local-moment metals [22], twisted bilayer graphene [23], and cuprates [24,25], including Rashba SOC [26]. In our MIMI superlattices, the superconducting transition temperature was amplified by the Fano–Feshbach shape resonance between two superconducting gaps tuned by quantum-size effects in the doped Mott insulator nanoscale layers of thickness L due to resonant scattering between the closed and open scattering channels. The Fano–Feshbach shape resonance in nanoscale heterostructures has been shown to compete [27] and coexist [28,29] with the Kondo effect.
The competition between the proximity Kondo scattering in the normal phase and the Fano–Feshbach shape resonance in MIMI superlattices can be tuned by changing the chemical potential in superlattices with variable L/d. While the electronic structure of the 2D electron gas at cuprate oxide interfaces with the associated interlayer phase separation has attracted high interest [30,31], we focused on quantum-size effects due to the nanoscale confinement of the interphase-correlated electron gas [1]. We first synthesized MIMI superlattices made of quantum wells with a period in the range of 1.98 < d < 5.28 nm, tuned at the resonant geometry where the pure Mott insulators LCO layers of thickness L, free of chemical substitutions, are intercalated with the metallic LSCO layers of thickness W, so that L/d = 2/3, where d = L + W, which form a superlattice of Mott insulator–metal interfaces (MIMIs). These superlattices exhibit T-linear resistivity in the metallic phase from 50 K to 270 K. By changing the conformational geometry parameter, L/d, away from the shape resonance, in the range 0.3 < L/d < 0.9, we observed a Kondo-like temperature dependence on resistivity, with the Kondo temperature, TK, being minimal at L/d = 2/3, where the superconducting critical temperature is maximal and the amplitude of the Kondo effect in the resistivity, R0K, is minimal.

2. Results

The nanoscale LSCO/LCO superlattices of alternating overdoped LSCO layers and undoped LCO layers were on a LaSrAlO4 substrate using molecular beam epitaxy (MBE). These superlattices formed a 2D electron gas (2DEG) at the interface space charge and exhibited 2D high-Tc superconductivity. The overall superlattice structure has a periodicity represented by parameter d, as shown in Figure 1a.
Our experimental approach involves the manipulation of the 2DEG interface space charge layer, which extends approximately 2.6 nm into the LCO layer from the LSCO–LCO interface. This layer experiences quantum confinement between the two LSCO potential barriers, of which the width is denoted as W within the superlattice. As a result, two artificial sub-bands emerge, allowing us to modulate their energy splitting or the transparency of the potential barrier by adjusting the thickness ratio, L/d. Figure 1a depicts a typical LSCO–LCO superlattice with a period of d = 3.96 nm. In this arrangement, the heterostructure consists of four LCO layers (MLs), with thickness L = 2.64 nm, while the LSCO layer has a thickness of two MLs, W = 1.32 nm. The superconducting critical temperature, Tc, as a function of L/d assumes a dome shape where, at L/d = 2/3, the TC reaches its maximum values [8] (see Figure 1b).
We measured the temperature dependence of the resistance in eighteen LSCO–LCO superlattices with varying values of d, L, and W. The changes in resistivity with temperature, as shown in Figure 1c, display curves that both approach and deviate from a linear trend (thick black line). The resistances that most closely follow a linear pattern are reported in Figure 1d, where we present the resistance of two significant samples. These samples have the same L/d parameter (L/d = 2/3) but exhibit different d-periodicities. At this L/d value, where the superlattices approach the maximum TC, in accordance with [8], the Fano–Fashback shape resonance drives up the critical temperature; we clearly observe the compelling experimental evidence that normal resistivity in the normal phase shows the T-linear behavior in our artificial MIMI superlattices in this figure.
To investigate the interplay of Fano resonance and Kondo scattering, away from the top of the superconducting dome, we considered the sheet resistance as a function of temperature and L/d in the measured eighteen LCO–LSCO superlattices listed in Figure 2. These samples were classified by different d-values, corresponding to different colors and different L/d values. The sheet resistance of all eighteen superlattices, as a function of temperature, is shown in Figure 2a, while Figure 2c depicts the normalized sheet resistance in units of R(T = 150 K) on a semilogarithmic scale.
The R(T) curves of all samples were fitted by a coexisting contribution of the Planckian T-linear resistance and the universal resistance function for the Kondo proximity effect at Mott insulator–metal interfaces [16,17,19,20]:
R T R T = 150   K =   r 0 + T 150 + A T 2 + B T 5 + R 0 K 1 + 2 1 s 1 T T K 2 s
for experimental R(T) values measured at T > 50 K. The Kondo temperature, TK, represents a characteristic temperature below which the coupling between the impurity and conduction electrons leads to the screening of the impurity spin. R0K is the amplitude of the Kondo-like contribution, while r0 is the residual contribution at T = 0 that represents the baseline resistivity of the material in the absence of Kondo physics, including contributions such as impurities and lattice defects. The term AT2 accounts for electron–electron scattering, while BT5 represents phonon scattering. Both AT2 and BT5 terms are non-Kondo contributions. The best-fitted curves are represented by continuous lines. The three different representations (Figure 2, panels a, and b) of the sheet resistance well visualize the fact that resistivity decreases as the temperature is lowered, reaching a minimum. At even lower temperatures, the electrical resistivity of the system increases (logarithmically, in the original perturbative Kondo treatment, or as a power law, beyond the perturbation theory). We observed how the behavior of the resistance deviates from linearity as the L/d value moves away from 2/3, where Fano resonance prevails. A phase diagram of normalized sheet resistance as a function of both temperature and the geometrical parameter L/d is given by the 3D color map in Figure 2c. In Figure 2d, we visualize a projection of this phase diagram on a T vs. L/d plane, highlighting the correspondence between L/d and the charge <δ> = 0.45 × (1 − L/d).
We used a least-squares fitting algorithm, extracting parameters A, B, R0K, and TK. The evolution of these parameters as a function of L/d is shown in Figure 3. The exponent, s, is a material-dependent parameter depending on the metal–Mott insulator interfaces, and, here, was found to be in the [0.10–0.18] range. In the sheet resistance measured at 150 K in all 18 measured samples with different L/d values, we observed an exponential increase for L/d > 2/3, indicating a metal insulator transition triggered by the geometrical conformational parameter. The evolutions of the fitting parameters A, B, r0 + R0K, and TK in Equation (1) are shown in Figure 3b–e. All parameters show a minimum at L/d = 2/3. In the scatter plot of TK vs. R0K, the positive correlation between these two parameters is made evident in Figure 3f.
Finally, in Figure 4, we show the anticorrelation between TC and TK. We report the theoretical calculations of superconducting critical temperature as a function of the geometry conformational parameter L/d, using the BPV theory in the case of superlattices with period 3 and 4 nm. This clearly shows the agreement between the theory and experiments, where the maximum TC can be controlled by the FFSR amplification [8].

3. Discussion

Many authors have observed that cuprate superconductors exhibiting high-Tc superconductivity display unconventional transport properties in the normal phase above the critical temperature. These properties are associated with quantum tunneling characterized by T-linear resistivity, the Planckian limit of the scattering rate assigned to quantum criticality, strong electronic correlation, and the coexistence of localized and itinerant states [32,33,34,35,36,37,38,39,40,41,42,43] in a homogeneous strange metal.
In this work, we provided further experimental evidence that artificial high Tc superlattices, made of modulation-doped quantum wells formed by the superconducting (S) layers of a chemically stoichiometric Mott insulator, La2CuO4, of thickness within the range of 2–3 nanometers intercalated by a normal metal (M), with the role of charge reservoir and a potential barrier of 500 meV [1], encompass the key feature of natural chemical-doped La2-xSrxCuO4. We confirm that AHTSs exhibit first (i) the superconducting dome of the critical temperature versus modulation doping, and second (ii), the T-linear Planckian sheet resistance around optimum modulation doping reported in reference [1]. The present results support the physical paradigm describing unconventionally high Tc cuprate perovskites characterized by intrinsic functional-arrested chemical nanoscale-phase separation observed by local and fast experimental probes [44,45,46,47,48,49,50], called the “superstripes” [46] or “swiss cheese” [48] scenario near the BEC–BCS crossover.

4. Materials and Methods

The MBE synthesis of artificial high-Tc superlattices: AHTS based on normal metal LSCO alternating with superconducting space-charge layers in LCO thin layers were synthesized via an ozone-assisted MBE method (DCA Instruments Oy, Turku, Finland) on LaSrAlO4 (001) substrates (compressive strain for La2CuO4 on LaSrAlO4 is +1.4%). The superlattice growth was controlled by the in situ reflection high-energy electron diffraction (RHEED). This method is characterized by the sequence of deposition of single atomic layers and minimal kinetic energy of impinging atoms (about 0.1 eV). The substrate temperature, Ts, according to a radiation pyrometer reading, was 650 °C, and the chamber pressure p ≈ 1.5 × 10−5 Torr of mixed ozone, and atomic and molecular oxygen. At the end of the procedure, the samples were cooled down to Ts.
Resistance measurements: The temperature dependence of resistance was determined in a four-point van der Pauw configuration with alternative DC currents ± 10 μA in a temperature range from room temperature to 4.2 K (liquid helium). The temperature-dependent resistance was measured by using a motorized custom-made dipstick in a transport helium dewar with temperature rate < 0.1 K/s.

5. Conclusions

The key remarkable result, shedding further light on the mechanism of high-Tc superconductivity, is that in the T-linear regime of sheet resistance in the normal phase of the AHTS made of MIMI at the magic ratio L/d = 2/3 and proximity Kondo scattering at a low temperature, vanishes, together with electron–electron Fermi liquid scattering. The Fano–Feshbach resonance between the two superconducting gaps, determined by quantum-size effects of nanoscale units, were optimized by quantum design at the geometrical conformational parameter of the superlattice, L/d = 2/3, at the top of the superconducting dome. Here, the Kondo temperature, TK, reached a minimum of the order of 4 Kelvin and the Kondo amplitude R0K vanished, as shown in Figure 3. On both sides of the superconducting dome where the average doping was in the L/d < 2/3 or L/d > 2/3 regime, the Kondo temperature, TK, increased and became higher than the superconducting critical temperature. These results are supported by the clear experimental anticorrelation between TK and RK. Finally, it is remarkable that where the resonant quantum tunnelling is tuned at the Fano–Feshbach for optimum resonant superconductivity [1,2], i.e., the residual T = 0 K resistance, the amplitude of the electron–electron Fermi scattering (T2 term) and the electron–phonon (T5 term) vanished in the normal phase.
The observed competition between the Fano–Feshbach resonance in multi-gap AHTS superconductors [6,7,8,51] can be expected since the Kondo effect [52] is based on the Anderson impurity model [53] and the atomic Fano resonance [54]. In fact, the Anderson model used in Kondo physics is similar to the Fano–Feshbach resonance near the quasi-flat band, although their band-mixing terms are different (i.e., one-particle and two-particle scatterings) [55,56,57,58,59].

Author Contributions

Conceptualization, A.B., G.L. and G.C.; methodology, G.L. and A.B.; software, A.V., G.C. and A.B.; validation, G.L. and A.B.; formal analysis, G.L., G.C. and A.B.; investigation, G.L., G.C., A.V. and A.B.; sample synthesis, G.L.; data curation, G.L., G.C. and A.B.; writing original draft preparation A.B., G.C., G.L. and S.C.; writing review and editing, A.B., G.C., S.C. and G.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research received funding from Superstripes onlus. SC acknowledges financial support from the Ateneo Research Projects of the University of Rome Sapienza: Competing phases and non-equilibrium phenomena in low-dimensional systems with microscopic disorder and nanoscale inhomogeneities (n. RM12117A4A7FD11B), Models and theories from anomalous diffusion to strange-metal behavior (n. RM12218162CF9D05), Non-conventional aspects for transport phenomena and non-equilibrium statistical mechanics (n. 234 RM123188E830D258).

Data Availability Statement

The data that support the findings of this study are available from the corresponding authors (A.B., G.C.) upon reasonable request.

Acknowledgments

We thank A. Perali for useful discussions, H. Tajima for comments on the manuscript and an unknown referee for very useful remarks for improving the paper.

Conflicts of Interest

The authors declare no conflicts of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Logvenov, G.; Bonmassar, N.; Christiani, G.; Campi, G.; Valletta, A.; Bianconi, A. The superconducting dome in artificial high-Tc superlattices tuned at the Fano–Feshbach resonance by quantum design. Condens. Matter 2023, 8, 78. [Google Scholar] [CrossRef]
  2. Valletta, A.; Bianconi, A.; Perali, A.; Logvenov, G.; Campi, G. High-Tc superconducting dome in artificial heterostructures made of nanoscale quantum building blocks. arXiv 2024, arXiv:2407.19069. [Google Scholar]
  3. Mazziotti, M.V.; Bianconi, A.; Raimondi, R.; Campi, G.; Valletta, A. Spin–orbit coupling controlling the superconducting dome of artificial superlattices of quantum wells. J. Appl. Phys. 2022, 132, 193908. [Google Scholar] [CrossRef]
  4. Mazziotti, M.V.; Valletta, A.; Raimondi, R.; Bianconi, A. Multigap superconductivity at an unconventional Lifshitz transition in a three-dimensional Rashba heterostructure at the atomic limit. Phys. Rev. B 2021, 103, 024523. [Google Scholar] [CrossRef]
  5. Mondal, D.; Mahapatra, S.R.; Derrico, A.M.; Rai, R.K.; Paudel, J.R.; Schlueter, C.; Gloskovskii, A.; Banerjee, R.; Hariki, A.; DeGroot, F.M.F.; et al. Modulation-doping a correlated electron insulator. Nat. Commun. 2023, 14, 6210. [Google Scholar] [CrossRef]
  6. Bianconi, A. On the possibility of new high Tc superconductors by producing metal heterostructures as in the cuprate perovskites. Solid State Commun. 1994, 89, 933–936. [Google Scholar] [CrossRef]
  7. Bianconi, A.; Valletta, A.; Perali, A.; Saini, N.L. Superconductivity of a striped phase at the atomic limit. Phys. C Supercond. 1998, 296, 269–280. [Google Scholar] [CrossRef]
  8. Bianconi, A. Feshbach shape resonance in multiband superconductivity in heterostructures. J. Supercond. 2005, 18, 625–636. [Google Scholar] [CrossRef]
  9. Cariglia, M.; Vargas-Paredes, A.; Doria, M.M.; Bianconi, A.; Miloševic, M.V.; Perali, A. Shape-resonant superconductivity in nanofilms: From weak to strong coupling. J. Supercond. Nov. Magn. 2016, 29, 3081–3086. [Google Scholar] [CrossRef]
  10. Salasnich, L.; Shanenko, A.A.; Vagov, A.; Aguiar, J.A.; Perali, A. Screening of pair fluctuations in superconductors with coupled shallow and deep bands: A route to higher-temperature superconductivity. Phys. Rev. B 2019, 100, 064510. [Google Scholar] [CrossRef]
  11. Ochi, K.; Tajima, H.; Iida, K.; Aoki, H. Resonant pair-exchange scattering and BCS-BEC crossover in a system composed of dispersive and heavy incipient bands: A Feshbach analogy. Phys. Rev. Res. 2022, 4, 013032.e26. [Google Scholar] [CrossRef]
  12. Valentinis, D.; Gariglio, S.; Fête, A.; Triscone, J.M.; Berthod, C.; Van Der Marel, D. Modulation of the superconducting critical temperature due to quantum confinement at the LaAlO3/SrTiO3 interface. Phys. Rev. B 2017, 96, 094518. [Google Scholar] [CrossRef]
  13. Barišić, N.; Chan, M.K.; Li, Y.; Yu, G.; Zhao, X.; Dressel, M.; Smontara, A.; Greven, M. Universal sheet resistance and revised phase diagram of the cuprate high-temperature superconductors. Proc. Natl. Acad. Sci. USA 2013, 110, 12235. [Google Scholar] [CrossRef]
  14. Pelc, D.; Veit, M.J.; Dorow, C.J.; Ge, Y.; Barišić, N.; Greven, M. Resistivity phase diagram of cuprates revisited. Phys. Rev. B 2020, 102, 075114. [Google Scholar] [CrossRef]
  15. Helmes, R.W.; Costi, T.A.; Rosch, A. Kondo proximity effect: How does a metal penetrate into a Mott insulator? Phys. Rev. Lett. 2008, 101, 066802. [Google Scholar] [CrossRef]
  16. Costi, T.A.; Hewson, A.C.; Zlatic, V. Transport coefficients of the Anderson model via the numerical renormalization group. J. Phys. Condens. Matter 1994, 6, 2519–2558. [Google Scholar] [CrossRef]
  17. Goldhaber-Gordon, D.; Göres, J.; Kastner, M.A.; Shtrikman, H.; Mahalu, D.; Meirav, U. From the Kondo regime to the mixed-valence regime in a single-electron transistor. Phys. Rev. Lett. 1998, 81, 5225. [Google Scholar] [CrossRef]
  18. Lee, M.; Williams, J.R.; Zhang, S.; Frisbie, C.D.; Goldhaber-Gordon, D. Electrolyte gate-controlled Kondo effect in SrTiO3. Phys. Rev. Lett. 2011, 107, 256601. [Google Scholar] [CrossRef]
  19. Mozaffari, S.; Guchhait, S.; Markert, J.T. Spin–orbit interaction ad Kondo scattering at the PrAlO3/SrTiO3 interface: Effects of oxygen content. J. Phys. Condens. Matter 2017, 29, 395002. [Google Scholar] [CrossRef]
  20. Yang, F.; Wang, Z.; Liu, Y.; Yang, S.; Yu, Z.; An, Q.; Ding, Z.; Meng, F.; Cao, Y.; Zhang, Q.; et al. Engineered Kondo screening and nonzero Berry phase in SrTiO3/LaTiO3/SrTiO3 heterostructures. Phys. Rev. B 2022, 106, 165421. [Google Scholar] [CrossRef]
  21. Zhuravlev, A.K.; Anokhin, A.O.; Irkhin, V.Y. One- and two-channel Kondo model with logarithmic Van Hove singularity: A numerical renormalization group solution. Phys. Lett. A 2018, 382, 528–533. [Google Scholar] [CrossRef]
  22. Kourris, C.; Vojta, M. Kondo screening and coherence in kagome local-moment metals: Energy scales of in the presence of flat bands. Phys. Rev. B 2023, 108, 235106. [Google Scholar] [CrossRef]
  23. Shankar, A.S.; Oriekhov, D.O.; Mitchell, A.K.; Fritz, L. Kondo effect in twisted bilayer graphene. Phys. Rev. B 2023, 107, 245102. [Google Scholar] [CrossRef]
  24. Miura, N.; Nakagawa, H.; Sekitani, T.; Naito, M.; Sato, H.; Enomoto, Y. High-magnetic-field study of high-Tc cuprates. Phys. B Condens. Matter 2002, 319, 310–320. [Google Scholar] [CrossRef]
  25. Sekitani, T.; Naito, M.; Miura, N. Kondo effect in underdoped n-type superconductors. Phys. Rev. B 2003, 67, 174503. [Google Scholar] [CrossRef]
  26. Yin, H.T.; Liu, X.J.; Feng, L.F.; Lü, T.Q.; Li, H. Spin-dependent Kondo effect induced by Rashba spin–orbit interaction in parallel coupled double quantum dots. Phys. Lett. A 2010, 374, 2865–2873. [Google Scholar] [CrossRef]
  27. Bułka, B.R.; Stefański, P.; Tagliacozzo, A. Interplay of Kondo and Fano resonance in electronic transport in nanostructures. Acta Phys. Pol. A 2005, 108, 555–569. [Google Scholar] [CrossRef]
  28. Fang, T.F.; Luo, H.G. Tuning the Kondo and Fano effects in double quantum dots. Phys. Rev. B 2010, 81, 113402. [Google Scholar] [CrossRef]
  29. Stefański, P. Interplay between quantum interference and electron interactions in a Rashba system. J. Phys. Condens. Matter 2010, 22, 505303. [Google Scholar] [CrossRef]
  30. Misawa, T.; Nomura, Y.; Biermann, S.; Imada, M. Self-optimized superconductivity attainable by interlayer phase separation at cuprate interfaces. Sci. Adv. 2016, 2, e1600664. [Google Scholar] [CrossRef]
  31. Tadano, T.; Nomura, Y.; Imada, M. Ab initio derivation of an effective Hamiltonian for the La2CuO4/La1.55Sr0.45CuO4 heterostructure. Phys. Rev. B 2019, 99, 155148. [Google Scholar] [CrossRef]
  32. Phillips, J.C. Reconciliation of normal-state and superconductive specific-heat, optical, tunneling, and transport data on Y-Ba-Cu-O. Phys. Rev. B 1989, 40, 7348–7349. [Google Scholar] [CrossRef]
  33. Varma, C.M.; Littlewood, P.B.; Schmitt-Rink, S.; Abrahams, E.; Ruckenstein, A.E. Phenomenology of the normal state of Cu-O high-temperature superconductors. Phys. Rev. Lett. 1989, 63, 1996. [Google Scholar] [CrossRef]
  34. Martin, S.; Fiory, A.T.; Fleming, R.M.; Schneemeyer, L.F.; Waszczak, J.V. Normal-state transport properties of Bi2+xSr2−yCuO6+δ crystals. Phys. Rev. B 1990, 41, 846–849. [Google Scholar] [CrossRef]
  35. Zaanen, J. Why the temperature is high. Nature 2004, 430, 512–513. [Google Scholar] [CrossRef] [PubMed]
  36. Cooper, R.A.; Wang, Y.; Vignolle, B.; Lipscombe, O.J.; Hayden, S.M.; Tanabe, Y.; Adachi, T.; Koike, Y.; Nohara, M.; Takagi, H.; et al. Anomalous criticality in the electrical resistivity of La2–xSrxCuO4. Science 2009, 323, 603–607. [Google Scholar] [CrossRef]
  37. Bruin, J.A.N.; Sakai, H.; Perry, R.S.; Mackenzie, A.P. Similarity of scattering rates in metals showing T-linear resistivity. Science 2013, 339, 804–807. [Google Scholar] [CrossRef] [PubMed]
  38. Haldane, F.D.M. Fermi-surface geometry and “Planckian dissipation”. arXiv 2018, arXiv:1811.12120. [Google Scholar]
  39. Shaginyan, V.R.; Amusia, M.Y.; Msezane, A.Z.; Stephanovich, V.A.; Japaridze, G.S.; Artamonov, S.A. Fermion condensation, T-linear resistivity, and Planckian limit. JETP Lett. 2019, 110, 290–295. [Google Scholar] [CrossRef]
  40. Amusia, M.; Shaginyan, V. Quantum Criticality, T-linear Resistivity, and Planckian Limit. In Strongly Correlated Fermi Systems; Springer Tracts in Modern Physics; Springer: Cham, Switzerland, 2020; Volume 283. [Google Scholar] [CrossRef]
  41. Patel, A.A.; Sachdev, S. Theory of a Planckian metal. Phys. Rev. Lett. 2019, 123, 066601. [Google Scholar] [CrossRef]
  42. Legros, A.; Benhabib, S.; Tabis, W.; Laliberté, F.; Dion, M.; Lizaire, M.; Vignolle, B.; Vignolles, D.; Raffy, H.; Li, Z.Z.; et al. Universal T-linear resistivity and Planckian dissipation in overdoped cuprates. Nat. Phys. 2019, 15, 142–147. [Google Scholar] [CrossRef]
  43. Balm, F.; Chagnet, N.; Arend, S.; Aretz, J.; Grosvenor, K.; Janse, M.; Zaanen, J. T-linear resistivity, optical conductivity, and Planckian transport for a holographic local quantum critical metal in a periodic potential. Phys. Rev. B 2023, 108, 125145. [Google Scholar] [CrossRef]
  44. Fratini, M.; Poccia, N.; Ricci, A.; Campi, G.; Burghammer, M.; Aeppli, G.; Bianconi, A. Scale-free structural organization of oxygen interstitials in La2CuO4+y. Nature 2010, 466, 841–844. [Google Scholar] [CrossRef] [PubMed]
  45. Campi, G.; Bianconi, A.; Poccia, N.; Bianconi, G.; Barba, L.; Arrighetti, G.; Innocenti, D.; Karpinski, J.; Zhigadlo, N.; Kazakov, S.M.; et al. Inhomogeneity of charge-density-wave order and quenched disorder in a high-Tc superconductor. Nature 2015, 525, 359–362. [Google Scholar] [CrossRef]
  46. Bianconi, A. Shape resonances in superstripes. Nat. Phys. 2013, 9, 536–537. [Google Scholar] [CrossRef]
  47. Jarlborg, T.; Bianconi, A. Fermi surface reconstruction of superoxygenated La2CuO4 superconductors with ordered oxygen interstitials. Phys. Rev. B—Condens. Matter Mater. Phys. 2013, 87, 054514. [Google Scholar] [CrossRef]
  48. Uemura, Y.J. Bose-Einstein to BCS crossover picture for high-Tc cuprates. Phys. C Supercond. 1997, 282, 194–197. [Google Scholar] [CrossRef]
  49. Li, Y.; Sapkota, A.; Lozano, P.M.; Du, Z.; Li, H.; Wu, Z.; Kundu, A.K.; Koch, R.J.; Wu, L.; Winn, B.L.; et al. Strongly overdoped La2−xSrxCuO4: Evidence for Josephson-coupled grains of strongly correlated superconductor. Phys. Rev. B 2022, 106, 224515. [Google Scholar] [CrossRef]
  50. Giraldo-Gallo, P.; Zhang, Y.; Parra, C.; Manoharan, H.C.; Beasley, M.R.; Geballe, T.; Kramer, M.; Fisher, I.R. Stripe-like nanoscale structural phase separation in superconducting BaPb1−x BixO3. Nat. Commun. 2015, 6, 8231. [Google Scholar] [CrossRef]
  51. Kondo, J. Superconductivity in transition metals. Prog. Theor. Phys. 1963, 29, 1–9. [Google Scholar] [CrossRef]
  52. Kondo, J. Resistance minimum in dilute magnetic alloys. Prog. Theor. Phys. 1964, 32, 37–49. [Google Scholar] [CrossRef]
  53. Anderson, P.W. Localized magnetic states in metals. Phys. Rev. 1961, 124, 41. [Google Scholar] [CrossRef]
  54. Fano, U. Effects of configuration interaction on intensities and phase shifts. Phys. Rev. 1961, 124, 1866. [Google Scholar] [CrossRef]
  55. Tajima, H.; Aoki, H.; Perali, A.; Bianconi, A. Emergent Fano-Feshbach resonance in two-band superconductors with an incipient quasiflat band: Enhanced critical temperature evading particle-hole fluctuations. Phys. Rev. B 2024, 109, L140504. [Google Scholar] [CrossRef]
  56. Tajima, H.; Perali, A.; Pieri, P. BCS-BEC crossover and pairing fluctuations in a two band superfluid/superconductor: AT matrix approach. Condens. Matter 2020, 5, 10. [Google Scholar] [CrossRef]
  57. Tajima, H.; Yerin, Y.; Perali, A.; Pieri, P. Enhanced critical temperature, pairing fluctuation effects, and BCS-BEC crossover in a two-band Fermi gas. Phys. Rev. B 2019, 99, 180503. [Google Scholar] [CrossRef]
  58. Paramasivam, S.K.; Gangadharan, S.P.; Milošević, M.V.; Perali, A. High-Tc Berezinskii-Kosterlitz-Thouless transition in two-dimensional superconducting systems with coupled deep and quasiflat electronic bands with Van Hove singularities. Phys. Rev. B 2024, 110, 024507. [Google Scholar] [CrossRef]
  59. Tajima, H.; Yerin, Y.; Pieri, P.; Perali, A. Mechanisms of screening or enhancing the pseudogap throughout the two-band Bardeen-Cooper-Schrieffer to Bose-Einstein condensate crossover. Phys. Rev. B 2020, 102, 220504. [Google Scholar] [CrossRef]
Figure 1. (a) Pictorial view of a practical realization of the nanoscale AHTS superlattice of quantum wells made of four monolayers (L = 2.64 nm) of undoped La2CuO4 (LCO), electronically doped by the interface space charge, which are intercalated by normal metal units made of two monolayers (W = 1.32 nm) of La1.55Sr0.45CuO4 (LSCO), forming a superlattice with a period of d = L + W = 3.96 nm and geometric parameter L/d = 2/3, giving the optimum critical temperature. (b) The superconductivity dome of the critical temperature, TC, as a function of L/d in the range of 0.25 < L/d < 0.9. The dashed line represents the predictions by the BPV theory at a Fano–Feshbach shape resonance for the superlattice of the quantum wells. We show the Tc determined as maximum of a derivative of the sheet resistance (blue circles) and maximum of the imaginary part of the mutual inductance (red squares). (c) Sheet resistance as a function of temperature of several LSCO/LCO superlattices where it is normalized at Rs 150K, which is the resistance measured at 150 K. The samples show the maximum Tc ≈ 43 K around L/d = 2/3, corresponding to a critical temperature in the optimum-doped LSCO. We outline a linear behavior (tick black line), observing how the samples with a maximum Tc, with L/d = 2/3, approach this linear resistivity regime. (d) The normalized sheet resistance as a function of temperature of two LSCO/LCO superlattices with L/d = 2/3 and different periods d = 3.96 nm and 2.97 nm, showing T-linear resistivity in the temperature range 50 K < T < 270 K.
Figure 1. (a) Pictorial view of a practical realization of the nanoscale AHTS superlattice of quantum wells made of four monolayers (L = 2.64 nm) of undoped La2CuO4 (LCO), electronically doped by the interface space charge, which are intercalated by normal metal units made of two monolayers (W = 1.32 nm) of La1.55Sr0.45CuO4 (LSCO), forming a superlattice with a period of d = L + W = 3.96 nm and geometric parameter L/d = 2/3, giving the optimum critical temperature. (b) The superconductivity dome of the critical temperature, TC, as a function of L/d in the range of 0.25 < L/d < 0.9. The dashed line represents the predictions by the BPV theory at a Fano–Feshbach shape resonance for the superlattice of the quantum wells. We show the Tc determined as maximum of a derivative of the sheet resistance (blue circles) and maximum of the imaginary part of the mutual inductance (red squares). (c) Sheet resistance as a function of temperature of several LSCO/LCO superlattices where it is normalized at Rs 150K, which is the resistance measured at 150 K. The samples show the maximum Tc ≈ 43 K around L/d = 2/3, corresponding to a critical temperature in the optimum-doped LSCO. We outline a linear behavior (tick black line), observing how the samples with a maximum Tc, with L/d = 2/3, approach this linear resistivity regime. (d) The normalized sheet resistance as a function of temperature of two LSCO/LCO superlattices with L/d = 2/3 and different periods d = 3.96 nm and 2.97 nm, showing T-linear resistivity in the temperature range 50 K < T < 270 K.
Condensedmatter 09 00043 g001
Figure 2. (a) (Open circles) Sheet resistance as a function of temperature of 18 superlattice samples with different d and L values indicated in the list on the left. The different d-values correspond to the different colors indicated. (b) Normalized sheet resistance as a function of temperature alongside the modeled lines through Equation (1). We show both the (left panel) linear and the (right panel) logarithmic evolution with temperature. (c) Color maps of normalized sheet resistance as a function of temperature, the geometrical parameter L/d. In panel (d), we show the correspondence between L/d and <δ> = 0.45 × (1 − L/d). The white dashed line indicates the L/d = 2/3 value.
Figure 2. (a) (Open circles) Sheet resistance as a function of temperature of 18 superlattice samples with different d and L values indicated in the list on the left. The different d-values correspond to the different colors indicated. (b) Normalized sheet resistance as a function of temperature alongside the modeled lines through Equation (1). We show both the (left panel) linear and the (right panel) logarithmic evolution with temperature. (c) Color maps of normalized sheet resistance as a function of temperature, the geometrical parameter L/d. In panel (d), we show the correspondence between L/d and <δ> = 0.45 × (1 − L/d). The white dashed line indicates the L/d = 2/3 value.
Condensedmatter 09 00043 g002
Figure 3. (a) Sheet resistance measured at 150 K in all 18 measured samples with different L/d values. We observed an exponential increase for L/d > 2/3. Evolution of the fitting parameters (b) A, (c) B, (d) r0 + R0k, and (e) Tk in Equation (1). (f) Scatter plot of Tk versus R0k showing the positive correlation between these two parameters.
Figure 3. (a) Sheet resistance measured at 150 K in all 18 measured samples with different L/d values. We observed an exponential increase for L/d > 2/3. Evolution of the fitting parameters (b) A, (c) B, (d) r0 + R0k, and (e) Tk in Equation (1). (f) Scatter plot of Tk versus R0k showing the positive correlation between these two parameters.
Condensedmatter 09 00043 g003
Figure 4. TK (left y-axis) and TC (right y-axis) as a function of L/d. The three red lines represent the theoretical TC in the superlattice with period 2.97 (red solid line) and 3:96 nm (dashed red line) as a function of L = d calculated in [8].
Figure 4. TK (left y-axis) and TC (right y-axis) as a function of L/d. The three red lines represent the theoretical TC in the superlattice with period 2.97 (red solid line) and 3:96 nm (dashed red line) as a function of L = d calculated in [8].
Condensedmatter 09 00043 g004
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Campi, G.; Logvenov, G.; Caprara, S.; Valletta, A.; Bianconi, A. Kondo Versus Fano in Superconducting Artificial High-Tc Heterostructures. Condens. Matter 2024, 9, 43. https://doi.org/10.3390/condmat9040043

AMA Style

Campi G, Logvenov G, Caprara S, Valletta A, Bianconi A. Kondo Versus Fano in Superconducting Artificial High-Tc Heterostructures. Condensed Matter. 2024; 9(4):43. https://doi.org/10.3390/condmat9040043

Chicago/Turabian Style

Campi, Gaetano, Gennady Logvenov, Sergio Caprara, Antonio Valletta, and Antonio Bianconi. 2024. "Kondo Versus Fano in Superconducting Artificial High-Tc Heterostructures" Condensed Matter 9, no. 4: 43. https://doi.org/10.3390/condmat9040043

APA Style

Campi, G., Logvenov, G., Caprara, S., Valletta, A., & Bianconi, A. (2024). Kondo Versus Fano in Superconducting Artificial High-Tc Heterostructures. Condensed Matter, 9(4), 43. https://doi.org/10.3390/condmat9040043

Article Metrics

Back to TopTop