Abstract
Interfacing magnetism with superconducting condensates is rapidly emerging as a viable route for the development of innovative quantum technologies. In this context, the development of rational design strategies to controllably tune the interaction between magnetic moments is crucial. Here we address this problem demonstrating the possibility of tuning the interaction between local spins coupled through a superconducting condensate with atomic scale precision. By using Cr atoms coupled to superconducting Nb, we use atomic manipulation techniques to precisely control the relative distance between local spins along distinct crystallographic directions while simultaneously sensing their coupling by scanning tunneling spectroscopy. Our results reveal the existence of highly anisotropic interactions, lasting up to very long distances, demonstrating the possibility of crossing a quantum phase transition by acting on the direction and interatomic distance between spins. The high tunability provides novel opportunities for the realization of topological superconductivity and the rational design of magneto-superconducting interfaces.
Subject terms: Spintronics; Superconducting properties and materials; Surfaces, interfaces and thin films
The possibility of using superconductivity to control indirect interactions between magnetic moments remains largely unexplored. Here, the authors report tuning the interaction between local spins of Cr atoms coupled through a superconducting condensate in Nb.
Introduction
Interfacing materials with radically different properties is a viable route to discover new physical phenomena and create new functionalities going beyond those hosted by each one of the building blocks1–3. In this context, the interface between a superconductor and a magnet is one of the most remarkable examples4. These ordered phases have been for a very long time thought to be mutually exclusive5–7. However, over the last decade, it has been experimentally demonstrated that under the right conditions, at a superconductor–magnetic interface both superconductivity and spin polarization can coexist creating new states such as Majorana modes8,9 and spin-triplet Cooper pairs10,11, both offering tantalizing opportunities for radically new technologies. Accelerating the progress in the field crucially relies on identifying, understanding, and controlling the interaction between superconducting condensates and magnetic order. In this context, indirect magnetic coupling, i.e., the coupling of magnetic moments mediated via the superconducting condensate, occupies a central role. In conventional metals, the polarized spins of conduction electrons can create such links called Ruderman–Kittel–Kasuya–Yosida (RKKY) interactions12–14. Their study proved to be essential to the development of spin-valve and tunneling magnetoresistance devices, through spin-engineered and especially synthetic antiferromagnet heterostructures that have had a tremendous impact on the exponential increase in magnetic storage data capacity15–17. With the advancements in scanning tunneling microscopy (STM) and spectroscopy (STS) techniques, it became possible to scrutinize RKKY interactions on the nanometer scale by manipulation of magnetic adatoms on a metal substrate18, thereby creating atomically precise spatial maps of the coupling strength up to 1.5 nm between the atoms19–21.
In superconducting condensates, it was theoretically predicted that besides the RKKY interactions, the additional longer-range spin coupling can be enabled by Cooper pairs with an exponential decay over the superconducting coherence length and which can dominate over conventional RKKY even at distances significantly smaller than the coherence length22,23. Despite their relevance for both fundamental aspects as well for their far-reaching implications for the design and creation of nontrivial magneto-superconducting states of matter, the indirect coupling between local magnetic moments mediated by superconducting condensates remains largely unexplored.
Here, we combine atomic manipulation techniques with high-resolution spectroscopy to investigate the emergence of superconducting condensate-mediated coupling between magnetic moments at the atomic scale. By systematically sensing the interaction between localized spins placed at different distances and along distinct crystallographic directions, we demonstrate that: (i) local spins can talk at large distances, up to several lattice constants, (ii) the use of a superconductor hosting a highly anisotropic Fermi surface allows to precisely tune the interaction depending on the crystallographic direction, and (iii) the possibility to tailor not only the strength, but also the sign of the energy shift induced by the interaction, and to use it to controllably cross the quantum phase transition (QPT). The latter is obtained in two distinct ways: by changing the crystallographic direction and/or the distance between magnetic impurities. A direct comparison with interactions mediated by conventional conduction electrons confirms the high anisotropy of the coupling and the existence of longer-range effects in the superconducting regime. Our findings project a plethora of opportunities for highly controllable magneto-superconducting interactions that can be extended to more complex systems.
Results
Experimental lineup. Figure 1a schematically illustrates how local magnetic moments can couple through a superconducting condensate. An isolated magnetic adatom (upper panel) induces the so-called Yu–Shiba–Rusinov (YSR) quasi-particle state24–27 characterized by a spatially decaying wave function28–32. When a second magnetic atom is brought close enough (middle panel), the local spins start coupling to each other, with the interaction being mediated by the substrate. By progressively reducing their distance (bottom panel), the direct magnetic exchange becomes stronger and finally prevails over substrate-mediated interactions.
Overall, this is expected to result in a rich-phase diagram where both the sign and the strength of the interactions can be tuned on demand by controllably acting on the distance between adatoms and crystallographic directions, as proposed in refs. 33,34.
As a prototypical system to scrutinize these aspects, we focus on Cr adatoms deposited on the (110) surface of Nb (see “Methods” for details on sample preparation). Niobium offers several advantages35–37. Its wide superconducting gap (2∆ = 3.05 meV with Tc = 9.2 K) allows to spectroscopically disentangle low-energy magnetic signatures from thermal broadening effects. Moreover, contrary to other superconductors such as Pb, its high cohesive energy makes it suitable for single-atom manipulation techniques35,36,38. Finally, its highly anisotropic Fermi surface39 (see also Fig. 2c) allows us to investigate the emergence of directional-dependent coupling between magnetic moments, which is expected to enhance the tunability.
Figure 1b shows a clean Nb area with the inset corresponding to the atomically resolved lattice. Only two Cr adatoms are present in the area, referred to as atoms A and B. As described in the following, this allows to detect their interaction ruling out spurious effects such as impurities or the influence of other adatoms in the nearby.
Figure 1c reports the STS spectrum acquired by positioning the tip on top of an isolated Cr atom (green line) and over the bare superconductor (black line). A Nb terminated tip has been used to enhance the energy resolution, resulting in the typical convoluted spectrum of tip and sample superconducting energy gap, ∆tip and ∆sample (see Supplementary Note 1). A rich spectroscopic scenario is visible for the Cr adatom, with several peaks emerging within the superconducting gap. These peaks are direct fingerprints of magnetic impurity–superconductor interactions, with magnetic moments locally breaking Cooper pairs40 and inducing the YSR quasi-particle resonances residing inside the superconducting energy gap24–27. YSR states always appear in pairs that are energetically particle–hole symmetric with respect to the Fermi level, with the energy position EYSR being determined by the coupling strength between the magnetic impurity and the superconducting condensate. Their orbital character can be visualized by spatially mapping their wave function distribution as shown in Supplementary Note 241,42.
The YSR pairs are generally characterized by a strong spectral weight asymmetry, the higher intensity in the occupied or unoccupied states being directly linked to a spin-screened and free-spin ground state, respectively. The crossing of YSR pairs through zero marks a first-order quantum phase transition (QPT) between the two different regimes43–46. In the present case, the -derived YSR state dominates the scene, showing a significant spectral weight asymmetry δ with higher intensity below the Fermi level (see Fig. 1). Note that the dz2-derived YSR pair is very close to zero energy (±∆tip in our case because of the use of superconducting tips). For strong interactions, occurring for adatoms placed at short distances, electron and hole components of the YSR pair can significantly shift becoming well-separated. The resulting energy shift ∆EYSR can be directly evaluated based on the shift of their peak position with respect to the single adatom case. However, when the distance between the adatoms is increased, their interaction and consequently ∆EYSR become progressively weaker. In such a regime, any small energy shift ∆EYSR results in a substantial transfer of spectral weight between the positive (E ≃ + ∆tip) and the negative (E ≃ − ∆tip) peak. The change in peak height ∆δ(r) as function of the interatomic distance r can thus be directly linked to the energy shift of the YSR pairs, i.e., ∆δ(r) ∼∆EYSR (see Supplementary Note 4), offering a very sensitive measurement protocol to sense both the strength as well as the sign of the shift induced by the interaction between magnetic adatoms once YSR pairs are close to the QPT. These two different regimes are analyzed in Figs. 2 and 3, respectively.
Directional dependence of short distant coupled spins
Figure 2 compares Cr dimers aligned along five different directions while keeping them at a small distance, i.e., below 1 nm. Note that none of the cases corresponds to the first neighbor configuration. The corresponding dI/dU signals are shown in Fig. 2a. The spectrum obtained on an isolated Cr adatom is provided as a reference (gray line). While the [001] direction (90°) shows a negative energy shift compared to the single adatom case, all other directions are characterized by a positive shift, in agreement with our theoretical calculations (see below). Coupled YSR states are also characterized by a distinct spatial distribution compared to the isolated adatom case, as illustrated in Supplementary Note 3.
Figure 2b allows us to analyze these aspects more into details by showing the evolution of spectral intensity for the YSR pair dominating the scene within the superconducting gap (see Fig. 1b). A significant transfer of spectral weight from hole- (red line) to electron-like (blue line) YSR states is visible by progressively increasing the angle between substrate and magnetic dimers. At specific angles, the lines are crossing each other, revealing the occurrence of a quantum phase transition from a free-spin (0°, direction ) into a screen-spin regime (90°, direction [001]). These results demonstrate the existence of highly anisotropic coupling between the magnetic impurities. The signature of the anisotropy of the substrate electronic structure can be grasped from the Fermi surface as calculated from ab-initio (see “Methods”) and illustrated in Fig. 2c. These effects are found to be remarkably strong: they allow to tune not only the strength but also the sign of energy shift induced by the adatom-adatom interaction by properly acting on the crystallographic direction.
Distance and directional dependence of indirectly coupled spins. To investigate the evolution of the interaction over larger distances, we performed a distance-dependence study of the coupling between local spins along three distinct crystallographic directions, i.e., [001], and . The distance between two atoms has been progressively increased step-by-step, with each step being defined by the smallest possible vector compatible with the underlying lattice periodicity, i.e., the vector connecting two subsequent adsorption sites.
Figure 3a–c reports the spectra obtained for each configuration of the adatoms subtracted by the spectrum obtained on the isolated atom (reference spectra are reported in Supplementary Note 5). This procedure allows to detect and quantitatively analyze the effects induced by the coupling (see Supplementary Note 4). In line with expectations, our data reveal that the coupling becomes progressively stronger by reducing the distance between the adatoms. However, as described in the following, the coupling is found to persist at distances significantly larger than those observed in metallic systems19–21. Interestingly, the data obtained at larger distances reveal also that the decay length of the interaction is highly anisotropic. While along the [001] and direction a clear coupling between the magnetic adatoms can be detected up to distances of ~4 nm, the interaction decays much faster along the direction, where the interaction is strongly suppressed already at ~1.4 nm. The different crystallographic direction behavior can be traced back to the anisotropy of the Fermi surface. Good nesting vectors between flat parts of the Fermi surface can have a focusing effect on the wave function of quasi particles, significantly enhancing the coupling strength over larger distances47–49. Although the Fermi surface is rather complex in the present case, good nesting conditions are expected to manifest along the [001] direction, in line with our experimental observations (see Fig. 2c).
A careful inspection of Fig. 3 additionally reveals the existence of an oscillatory behavior of the coupling strength which is superimposed to the decay. This is clearly visible along the [001] and directions, while it becomes more subtle along the direction.
These oscillations are further highlighted in Fig. 3d, e, where the in-gap peak intensities are plotted separately for negative and positive energies. The comparison between electron- and hole-like states clearly reveal that their distance-dependence trends are anticorrelated. This observation confirms the transfer of spectral weight within the YSR pair. Note that a distance-dependent quantum phase transition from a free-spin to a screened-spin regime takes place along all directions, highlighted by a red cloud in Fig. 3d, e.
Figure 3g directly compares the change of the peaks asymmetry with respect to the isolated adatom ∆δ as a function of the atom separation r. Positive and negative values correspond to a shift ∆EYSR to positive and negative energies, respectively. Although the spatial resolution imposed by the lattice impedes a high sampling rate, the curves suggest a similar oscillation period for the different directions (see Supplementary Notes 6 and 7). To explore the origin of the observed behavior, we calculated the change of the YSR energy of the z2-derived orbital of the dimer atoms with respect to the YSR energy of the isolated adatom, ∆EYSR, using first principles and an effective Hamiltonian construction (see “Methods” and Supplementary Notes 8 and 9), which is shown in Fig. 3h using a spatial map for all discussed dimer configurations. In good agreement with the experimental observations, the energy shift is strongest for the next-nearest-neighbor dimers along the two high-symmetry directions [001] and and significantly weaker for the directions in between. The simulations also shed light on the fundamental physical origins determining the strength of the YSR energy shifts: (i) the nature of the magnetic coupling begin ferromagnetic or anti-ferromagnetic, (ii) the strength of the direct electronic coupling between the dimer atoms, and (iii) the strength of the indirect superconductor-mediated electronic and magnetic coupling between the dimer atoms. Analytic model calculations based on a simplified Alexander–Anderson model50 (see Supplementary Note 9) predict distinct shifts of the YSR energies depending on ferromagnetic or anti-ferromagnetic interactions. These results from the influence of the magnetic coupling on the formation of the electronic bonding and anti-bonding states51,52. However, spin–orbit coupling can lift this degeneracy resulting in split YSR states also for AFM coupled spins, as shown in ref. 53 and Supplementary Note 3. The more advanced theoretical model based on ab-initio parameters (leading to Fig. 3h) demonstrates that the sign of the energy shift and the occurrence of a QPT cannot be trivially linked to the sign of the magnetic exchange interaction. In agreement with our theoretical calculations reported in Fig. 3h, the experimentally scrutinized magnetic structures reported in Supplementary Note 10 are indicative of ferromagnetic order for adatoms coupled along the direction at the next-nearest-neighbor distance, while those oriented along [001] and both show antiferromagnetically coupled adatoms, despite being characterized by opposite energy shift of the YSR pairs, as correctly captured in our calculations reported in Fig. 3h.
Indirect coupling between spins in the metallic regime
To corroborate our findings, we compare the coupling between adatoms in the superconducting regime with the one observed when the interaction is mediated by conventional conduction electrons. To this end, the substrate has been driven into the metallic regime by applying a magnetic field of 1 T perpendicular to the sample surface. As described in the following, this allows to confirm the existence of highly anisotropic indirect interactions as well as to highlight to the existence of a distinct contribution to the magnetic coupling which originates from the superconducting condensate.
Figure 4a–c reports the spectra acquired over the Cr adatoms by progressively changing their relative distance along each one of the three distinct directions scrutinized in the superconducting regime, i.e., [001], , and . All spectra are overlapped to a reference spectrum obtained for an isolated Cr atom (gray line). The spectrum of the Cr single adatom is characterized by a sharp step-like signature at the Fermi level which has recently been demonstrated being directly linked to its magnetic ground state37. Changes in this spectral shape can thus be used to reveal the existence of magnetic interactions between the adatoms.
The direct comparison of the spectra acquired along the three distinct directions when adatoms are at the closest distance (step 1 in Fig. 4a–c) confirms the existence of a highly anisotropic indirect interaction. This effect can be effectively visualized by comparing the gray areas in panels (a–c), which highlight the crystallographic direction-dependent changes with respect to the single adatom case. The largest spectral change is observed along the [001] direction. Along this direction, the step-like feature characteristic of the single adatom is significantly suppressed and broadened, signaling a strong magnetic interaction between the adatoms. The spectral change is weaker along the direction and it becomes almost negligible along the direction. Note that the distance between the adatoms is smaller along the direction (r = 0.57 nm) compared to the [001] direction (r = 0.66 nm). If the effect would be distance-dependent, the spectral change along the direction should be stronger than the one along the [001] direction. Our data clearly reveal the opposite. This allows to identify the highly anisotropic Fermi surface as the origin of the anisotropic coupling, supporting the conclusions drawn from the analysis of the superconducting regime.
Finally, we investigate the distance-dependence behavior of conduction electrons-mediated interactions. Figure 4a–c evidence that along the [001] direction, where the interaction at short distances is the strongest, the spectrum is perfectly overlapping the isolated adatom case already at r = 1.32 nm. A slower decay is observed along the direction, where the difference with respect to the isolated adatom case shows an oscillatory on-off behavior depending on the odd or even number of steps between the adatoms, respectively. An oscillatory behavior along this crystallographic direction was also revealed in the superconducting regime (see Fig. 3e, g). Finally, along the direction, the difference with respect to the single adatom vanishes already at step 2, an observation in line with the very rapid spatial decay of the interaction between magnetic moments observed also in the superconducting regime. The decay length in the metallic regime is in line with earlier STM studies on other metallic systems19–21 and significantly faster than the one observed in the superconducting regime, where interactions between adatoms could be detected up to 4 nm (see Supplementary Notes 6 and 7). All together, these observations could be originating from the decisive role of the superconducting condensate in mediating interactions over long distances, an effect predicted decades ago but which always escaped experimental verification.
Discussion
In conclusion, we demonstrated the emergence of long-range, crystallographic direction-dependent, and highly tunable superconducting condensate-mediated coupling between local spins. We could directly link the strong crystallographic direction dependence of the coupling to the highly anisotropic Fermi surface, which can be used to controllably cross the QPT between screen-spin and free-spin regime by changing the alignment of the atom pair with respect to the underlying lattice with atomic precision. Longer-range coupling effects between magnetic impurities can be observed along the [001] and directions compared to , an enhancement which is potentially enabled by the optimal Fermi surface nesting conditions significantly strengthening substrate-mediated magnetic interactions. Our results unveil, in an extremely simple material platform, the existence of a very rich-phase diagram which can be used for tuning magnetic interactions mediated by superconducting condensates on demand. The simultaneous fulfilling of the following requirements: (i) magnetic impurities hosting deep YSR states, (ii) indirect coupling between adatoms taking place at large distances, and (iii) the possibility to cross the QPT upon hybridization54, create ideal conditions for experimentally scrutinizing the emergence of topologically superconductivity in dilute spin systems55 going beyond the nearest neighbors and ferromagnetically coupled spin chains explored so far8,9,56. We envision our results to be applicable to more complex systems and phenomena, creating precise guidelines for the rational design of exotic phenomena emerging at magnetic-superconducting interfaces57,58.
Methods
Sample and tip preparation
The bulk single-crystal Nb(110) substrate was cleaned in UHV by repeatedly heating the surface to 2300 K in 12 s38. Cr adatoms were deposited in-situ and with the substrate below a temperature of 15 K. All atoms were found to be adsorbed in the hollow site of the Nb(110) surface37. Measurements were taken in UHV at 600 mK using the Tribus STM head (Scienta Omicron). Cr adatoms could be moved with atomic precision by approaching them with the STM tip in constant current mode and a setpoint of −5 mV; 70 nA. dI /dU spectra were taken with the standard lock-in technique. For measurements in the normal metallic regime, we used a tungsten tip characterized on Ag(111) and applied a magnetic field of >1 T. For measurements in the superconducting regime, we used a superconducting Nb microtip obtained by deep indentations into the Nb single crystal. As described in Supplementary Note 1, this allows to improve the energy resolution.
Measurement protocol
Throughout the distance-dependence experiments, atom A was not moved from its initial position while atom B was subsequently positioned in a controlled way to specific hollow site positions, so that all changes seen on atom A are exclusively triggered by movements of atom B, therefore all spectra shown were taken on atom A. Both at the start and the end of the experiment, the very same configuration shown in Fig. 1a was arranged (atom distance |rref| = 5.6 nm) and measured to rule out any tip changes during the multiple atom manipulations required. An exactly identical dI/dU signal could be acquired before and after and was then used as the reference for all other spectra.
First-principles calculations
Our first-principles approach is implemented in the framework of the scalar-relativistic full-electron Korringa–Kohn–Rostoker (KKR) Green function augmented self-consistently with spin–orbit interaction59,60. The method is based on multiple-scattering theory allowing an embedding scheme, which is versatile for the treatment of nanostructures in real space. The full charge density is computed within the atomic-sphere approximation (ASA) and local spin density approximation (LSDA) is employed for the evaluation of the exchange-correlation potential61. We assume an angular momentum cutoff at lmax = 3 for the orbital expansion of the Green function and when extracting the local density of states a k-mesh of 150 × 150 is considered. The Nb(110) surface is modeled by a slab containing 22 layers enclosed by two vacuum regions with a thickness of 9.33 Å° each. The Cr adatoms are placed on the hollow stacking site relaxed toward the surface by 20% of the inter-layer distance of the underlying Nb(110) surface, which was shown to be the energetically favored stacking in ref. 37. The magnetic exchange interactions in the dimers were obtained using the magnetic force theorem in the frozen-potential approximation and the infinitesimal rotation method62,63. The on-site magnetic anisotropy of the Cr atoms in all dimers is estimated by one of the isolated Cr adatom, which is obtained from the method of constraining fields64. Further information can be found in Supplementary Note 7. The change of the YSR energies (∆EYSR) is estimated using a tight-binding model with parameters obtained from density functional theory and an effective Hamiltonian construction (see ref. 65 and Supplementary Note 8).
Supplementary information
Author contributions
F.K. performed the STM measurements and analyzed the data. S.B. and S.L. conceived the theoretical framework describing the YSR states and analyzed the theoretical data. S.L. and P.S. wrote the initial version of the manuscript to which all authors contributed. S.L., S.S.P.P., and P.S. supervised the project.
Funding
Open Access funding enabled and organized by Projekt DEAL.
Data availability
All data needed to evaluate the conclusions in the paper are present in the paper and/or the supplementary materials. Additional data related to this paper may be requested from the authors.
Code availability
The KKR Green function code that supports the findings of this study is available from the corresponding author on reasonable request.
Competing interests
The authors declare no competing interests.
Footnotes
Peer review information Nature Communications thanks Saw Hla and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Contributor Information
Samir Lounis, Email: s.lounis@fz-juelich.de.
Stuart S. P. Parkin, Email: stuart.parkin@mpi-halle.mpg.de
Paolo Sessi, Email: paolo.sessi@mpi-halle.mpg.de.
Supplementary information
The online version contains supplementary material available at 10.1038/s41467-021-26802-x.
References
- 1.Ohtomo A, Hwang HY. A high-mobility electron gas at the LaAlO3/SrTiO3 heterointerface. Nature. 2004;427:423–426. doi: 10.1038/nature02308. [DOI] [PubMed] [Google Scholar]
- 2.Wang Q-Y, et al. Interface-induced high-temperature superconductivity in single unit-cell FeSe films on SrTiO3. Chin. Phys. Lett. 2012;29:037402. [Google Scholar]
- 3.Zhong, D. et al. Van der Waals engineering of ferromagnetic semiconductor heterostructures for spin and valleytronics. Sci. Adv.3, e1603113 (2017). [DOI] [PMC free article] [PubMed]
- 4.Linder J, Robinson JWA. Superconducting spintronics. Nat. Phys. 2015;11:307–315. [Google Scholar]
- 5.Suhl H, Matthias BT. Impurity scattering in superconductors. Phys. Rev. 1959;114:977–988. [Google Scholar]
- 6.Matthias BT, Suhl H, Corenzwit E. Spin exchange in superconductors. Phys. Rev. Lett. 1958;1:92–94. [Google Scholar]
- 7.Reif F, Woolf MA. Energy gap in superconductors containing paramagnetic impurities. Phys. Rev. Lett. 1962;9:315–316. [Google Scholar]
- 8.Nadj-Perge S, et al. Observation of Majorana fermions in ferromagnetic atomic chains on a superconductor. Science. 2014;346:602–607. doi: 10.1126/science.1259327. [DOI] [PubMed] [Google Scholar]
- 9.Ruby M, et al. End states and subgap structure in proximity-coupled chains of magnetic adatoms. Phys. Rev. Lett. 2015;115:197204. doi: 10.1103/PhysRevLett.115.197204. [DOI] [PubMed] [Google Scholar]
- 10.Robinson JWA, Witt JDS, Blamire MG. Controlled injection of spin-triplet supercurrents into a strong ferromagnet. Science. 2010;329:59–61. doi: 10.1126/science.1189246. [DOI] [PubMed] [Google Scholar]
- 11.Di Bernardo A, et al. Signature of magnetic-dependent gapless odd frequency states at superconductor/ferromagnet interfaces. Nat. Commun. 2015;6:8053. doi: 10.1038/ncomms9053. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Ruderman MA, Kittel C. Indirect exchange coupling of nuclear magnetic moments by conduction electrons. Phys. Rev. 1954;96:99–102. [Google Scholar]
- 13.Kasuya T. A theory of metallic ferro- and antiferromagnetism on Zener’s Model. Prog. Theor. Phys. 1956;16:45–57. [Google Scholar]
- 14.Yosida K. Magnetic properties of Cu-Mn alloys. Phys. Rev. 1957;106:893–898. [Google Scholar]
- 15.Parkin M, Roche Oscillations in exchange coupling and magnetoresistance in metallic superlattice structures: Co/Ru, Co/Cr, and Fe/Cr. Phys. Rev. Lett. 1990;64:2304–2307. doi: 10.1103/PhysRevLett.64.2304. [DOI] [PubMed] [Google Scholar]
- 16.Parkin SSP. Systematic variation of the strength and oscillation period of indirect magnetic exchange coupling through the 3d, 4d, and 5d transition metals. Phys. Rev. Lett. 1991;67:3598–3601. doi: 10.1103/PhysRevLett.67.3598. [DOI] [PubMed] [Google Scholar]
- 17.Parkin SSP, Mauri D. Spin engineering: direct determination of the Ruderman-Kittel-Kasuya-Yosida far-field range function in ruthenium. Phys. Rev. B. 1991;44:7131–7134. doi: 10.1103/physrevb.44.7131. [DOI] [PubMed] [Google Scholar]
- 18.Eigler DM, Schweizer EK. Positioning single atoms with a scanning tunnelling microscope. Nature. 1990;344:524–526. [Google Scholar]
- 19.Zhou L, et al. Strength and directionality of surface Ruderman–Kittel–Kasuya–Yosida interaction mapped on the atomic scale. Nat. Phys. 2010;6:187–191. [Google Scholar]
- 20.Khajetoorians AA, et al. Atom-by-atom engineering and magnetometry of tailored nanomagnets. Nat. Phys. 2012;8:497–503. [Google Scholar]
- 21.Khajetoorians AA, et al. Tailoring the chiral magnetic interaction between two individual atoms. Nat. Commun. 2016;7:10620. doi: 10.1038/ncomms10620. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Kochelaev B, Tagirov L, Khusainov M. Spatial dispersion of the spin susceptibility of conductivity electrons in superconductors. Zh. Eksp. Teor. Fiz. 1979;76:578–587. [Google Scholar]
- 23.Yao NY, Glazman LI, Demler EA, Lukin MD, Sau JD. Enhanced antiferromagnetic exchange between magnetic impurities in a superconducting host. Phys. Rev. Lett. 2014;113:087202. doi: 10.1103/PhysRevLett.113.087202. [DOI] [PubMed] [Google Scholar]
- 24.Yu L. Bound state in superconductors with paramagnetic impurities. Acta Phys. Sin. 1965;21:75–98. [Google Scholar]
- 25.Shiba H. Classical spins in superconductors. Prog. Theor. Phys. 1968;40:435–451. [Google Scholar]
- 26.Rusinov AI. On the theory of gapless superconductivity in alloys containing paramagnetic impurities. JETP. 1969;29:1101–1106. [Google Scholar]
- 27.Yazdani J, Lutz C, Eigler Probing the local effects of magnetic impurities on superconductivity. Science. 1997;275:1767–1770. doi: 10.1126/science.275.5307.1767. [DOI] [PubMed] [Google Scholar]
- 28.Ménard GC, et al. Coherent long-range magnetic bound states in a superconductor. Nat. Phys. 2015;11:1013–1016. [Google Scholar]
- 29.Ruby M, Heinrich BW, Peng Y, von Oppen F, Franke KJ. Wave-function hybridization in Yu-Shiba-Rusinov dimers. Phys. Rev. Lett. 2018;120:156803. doi: 10.1103/PhysRevLett.120.156803. [DOI] [PubMed] [Google Scholar]
- 30.Choi D-J, et al. Influence of magnetic ordering between Cr adatoms on the Yu-Shiba-Rusinov states of the β−Bi2Pd superconductor. Phys. Rev. Lett. 2018;120:167001. doi: 10.1103/PhysRevLett.120.167001. [DOI] [PubMed] [Google Scholar]
- 31.Kezilebieke S, Dvorak M, Ojanen T, Liljeroth P. Coupled Yu–Shiba–Rusinov states in molecular dimers on NbSe2. Nano Lett. 2018;18:2311–2315. doi: 10.1021/acs.nanolett.7b05050. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Kim H, Rozsa L, Schreyer D, Simon E, Wiesendanger R. Long-range focusing of magnetic bound states in superconducting lanthanum. Nat. Commun. 2020;11:4573. doi: 10.1038/s41467-020-18406-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Morr DK, Yoon J. Impurities, quantum interference, and quantum phase transitions in s-wave superconductors. Phys. Rev. B. 2006;73:224511. [Google Scholar]
- 34.Morr DK, Stavropoulos NA. Quantum interference between impurities: creating novel many-body states in s-wave superconductors. Phys. Rev. B. 2003;67:020502. [Google Scholar]
- 35.Odobesko A, et al. Observation of tunable single-atom Yu-Shiba-Rusinov states. Phys. Rev. B. 2020;102:174504. [Google Scholar]
- 36.Schneider, L., Beck, P., Wiebe, J. & Wiesendanger, R. Atomic-scale spin-polarization maps using functionalized superconducting probes. Sci. Adv.7, https://advances.sciencemag.org/content/7/4/eabd7302 (2021). [DOI] [PMC free article] [PubMed]
- 37.Küster F, et al. Correlating Josephson supercurrents and Shiba states in quantum spins un- conventionally coupled to superconductors. Nat. Commun. 2021;12:1108. doi: 10.1038/s41467-021-21347-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Odobesko, A. B. et al. Preparation and electronic properties of clean superconducting nb(110) surfaces. Phys. Rev. B99, 115437 (2019).
- 39.Odobesko A, et al. Anisotropic vortices on superconducting nb(110) Phys. Rev. B. 2020;102:174502. [Google Scholar]
- 40.Bardeen J, Cooper LN, Schrieffer JR. Theory of superconductivity. Phys. Rev. 1957;108:1175–1204. [Google Scholar]
- 41.Ruby, M., Peng, Y., von Oppen, F., Heinrich, B. W. & Franke, K. J. Orbital picture of Yu-Shiba-Rusinov multiplets. Phys. Rev. Lett.117, 186801 (2016). [DOI] [PubMed]
- 42.Choi D-J, et al. Mapping the orbital structure of impurity bound states in a superconductor. Nat. Commun. 2017;8:15175. doi: 10.1038/ncomms15175. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Franke KJ, Schulze G, Pascual JI. Competition of superconducting phenomena and kondo screening at the nanoscale. Science. 2011;332:940–944. doi: 10.1126/science.1202204. [DOI] [PubMed] [Google Scholar]
- 44.Farinacci L, et al. Tuning the coupling of an individual magnetic impurity to a superconductor: quantum phase transition and transport. Phys. Rev. Lett. 2018;121:196803. doi: 10.1103/PhysRevLett.121.196803. [DOI] [PubMed] [Google Scholar]
- 45.Huang, H. et al. Quantum phase transitions and the role of impurity-substrate hybridization in Yu-Shiba-Rusinov states. Commun. Phys.3. https://www.nature.com/articles/s42005-020-00469-0 (2020).
- 46.Salkola MI, Balatsky AV, Schrieffer JR. Spectral properties of quasiparticle excitations induced by magnetic moments in superconductors. Phys. Rev. B. 1997;55:12648–12661. [Google Scholar]
- 47.Weismann A, et al. Seeing the Fermi surface in real space by nanoscale electron focusing. Science. 2009;323:1190–1193. doi: 10.1126/science.1168738. [DOI] [PubMed] [Google Scholar]
- 48.Lounis S, et al. Theory of real space imaging of Fermi surface parts. Phys. Rev. B. 2011;83:035427. [Google Scholar]
- 49.Bouhassoune M, et al. Quantum well states and amplified spin-dependent Friedel oscillations in thin films. Nat. Commun. 2014;5:5558. doi: 10.1038/ncomms6558. [DOI] [PubMed] [Google Scholar]
- 50.Oswald A, Zeller R, Braspenning PJ, Dederichs PH. Interaction of magnetic impurities in Cu and Ag. J. Phys. F: Met. Phys. 1985;15:193–212. [Google Scholar]
- 51.Flatte ME, Reynolds DE. Local spectrum of a superconductor as a probe of interactions between magnetic impurities. Phys. Rev. B. 2000;61:14810–14814. [Google Scholar]
- 52.Yao NY, Glazman LI, Demler EA, Lukin MD, Sau JD. Enhanced antiferromagnetic exchange between magnetic impurities in a superconducting host. Phys. Rev. Lett. 2014;113:087202. doi: 10.1103/PhysRevLett.113.087202. [DOI] [PubMed] [Google Scholar]
- 53.Beck P, et al. Spin-orbit coupling induced splitting of Yu-Shiba-Rusinov states in antiferromagnetic dimers. Nat. Commun. 2021;12:2040. doi: 10.1038/s41467-021-22261-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Hoffman S, Klinovaja J, Meng T, Loss D. Impurity-induced quantum phase transitions and magnetic order in conventional superconductors: competition between bound and quasiparticle states. Phys. Rev. B. 2015;92:125422. [Google Scholar]
- 55.Pientka F, Glazman LI, von Oppen F. Topological superconducting phase in helical Shiba chains. Phys. Rev. B. 2013;88:155420. [Google Scholar]
- 56.Schneider, L. et al. Topological Shiba bands in artificial spin chains on superconductors. Nat. Phys. (2021).
- 57.Steiner, J. F., Mora, C., Franke, K. J. & von Oppen, F. Quantum magnetism and topological superconductivity in Yu-Shiba-Rusinov chains. Preprint at https://arxiv.org/abs/2107.00031 (2021). [DOI] [PubMed]
- 58.Mishra, A., Simon, P., Hyart, T. & Trif, M. A Yu-Shiba-Rusinov qubit. Preprint at https://arxiv.org/abs/2106.01188 (2021).
- 59.Papanikolaou N, Zeller R, Dederichs PH. Conceptual improvements of the KKR method. J. Phys.: Condens. Matter. 2002;14:2799. [Google Scholar]
- 60.Bauer, D. S. G. Development of a relativistic full-potential first-principles multiple scattering Green function method applied to complex magnetic textures of nanostructures at surfaces. Forschungszentrum Ju ¨lich PhD thesis https://www.publications.rwth-aachen.de/record/229375 (2014).
- 61.Vosko SH, Wilk L, Nusair M. Accurate spin-dependent electron liquid correlation energies for local spin density calculations: a critical analysis. Can. J. Phys. 1980;58:1200–1211. [Google Scholar]
- 62.Liechtenstein AI, Katsnelson MI, Antropov VP, Gubanov VA. Local spin density functional approach to the theory of exchange interactions in ferromagnetic metals and alloys. J. Magn. Magn. Mater. 1987;67:65–74. [Google Scholar]
- 63.Ebert H, Mankovsky S. Anisotropic exchange coupling in diluted magnetic semiconductors: ab initio spin-density functional theory. Phys. Rev. B. 2009;79:045209. [Google Scholar]
- 64.Brinker S, Dias MDS, Lounis S. The chiral biquadratic pair interaction. N. J. Phys. 2019;21:083015. [Google Scholar]
- 65.Schneider L, et al. Controlling in-gap end states by linking nonmagnetic atoms and artificially-constructed spin chains on superconductors. Nat. Commun. 2020;11:4707. doi: 10.1038/s41467-020-18540-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Data Availability Statement
All data needed to evaluate the conclusions in the paper are present in the paper and/or the supplementary materials. Additional data related to this paper may be requested from the authors.
The KKR Green function code that supports the findings of this study is available from the corresponding author on reasonable request.