Abstract
Vacancies in solid-state physics are underexplored in materials with strong electron-electron correlations. Recent research on the Ir-Sb binary system revealed an extended buckled-honeycomb vacancy (BHV) order. Superconductivity arises by suppressing BHV ordering through high-pressure growth with excess Ir atoms or Rh substitution, yet the superconducting pairing nature remains unknown. To explore this, we conducted muon spin rotation experiments on Ir1−δSb (synthesized at 5.5 GPa, Tc = 4.2 K) and ambient pressure synthesized optimally Rh-doped Ir1−xRhxSb (x=0.3, Tc = 2.7 K). The exponential temperature dependence of the superfluid density suggests a fully gapped superconducting state exists in both samples. The ratio of Tc to the superfluid density resembles that of unconventional superconductors. A significant increase in the superfluid density in the high-pressure synthesized sample correlates with Tc, indicating that unconventional superconductivity is intrinsic to the Ir-Sb binary system. These findings, along with the dome-shaped phase diagram, highlight IrSb as the first unconventional superconducting parent phase with ordered vacancies, requiring further theoretical investigations.
Subject terms: Superconducting properties and materials, Superconducting properties and materials
Vacancies or defects are structural features of the crystal lattice that can be used to engineer the physical properties of a solid-state system, and have played an important role in the investigation of quantum materials. Here, the authors apply muon spin rotation to explore the suppression of vacancy ordering in Rh-doped Ir1−xRhxSb and discuss the potential presence of unconventional superconductivity in the system.
Introduction
Vacancies and defects play crucial roles in the properties of materials, and their significance is evident in various classes of materials, including transition metal dichalcogenides (TMDs)1–5, Fe-based superconductors6–9 and cuprate high-temperature superconductors10–12. In transition metal dichalcogenides like MoS2 or WSe2, vacancies and defects can significantly influence electronic and optical properties1. For instance, point defects such as sulfur or selenium vacancies can introduce localized states into the band gap, affecting the materials conductivity and optical absorption. Defects can serve as active sites for chemical reactions and play a role in catalysis2. Defects in TMDs can also induce magnetism and lead to interesting magnetic properties3–5.
Several studies have investigated the impact of vacancy ordering in different iron-based superconductors6,7. For example, in iron chalcogenide superconductors such as FeSe, the ordering of selenium vacancies has been observed to influence the electronic structure and can lead to novel phenomena, including the emergence of superconductivity8. Understanding and controlling the role that vacancies and defects play in materials is therefore essential for tailoring their properties for specific applications. Researchers often explore the effects of defects to harness their potential benefits or mitigate undesirable consequences in various materials systems, ranging from electronics and catalysis to energy storage and superconductivity.
In this context, scientists have uncovered a distinct type of vacancy ordering in the Ir-Sb binary system Ir16Sb18, manifesting as an extended buckled-honeycomb vacancy (BHV) order13,14. The system Ir16Sb18 has been identified as the first superconducting parent phase known to exhibit ordered vacancies. The emergence of superconductivity in Ir-Sb is closely linked to the suppression of the BHV ordering. This suppression is achieved through two distinct methods: high-pressure growth of Ir1−δSb involving the squeezing of additional Ir atoms into the vacancies, and isovalent Rh substitution Ir1−xRhxSb. These interventions disrupt the ordered vacancy structure, paving the way for superconductivity. However, while the connection between vacancy ordering and superconductivity is established, the exact nature of the superconducting pairing in this system remains an intriguing aspect which has not yet been fully explored.
The comprehensive exploration of superconductivity at the microscopic level in the bulk of Ir1−δSb and optimally doped Ir1−xRhxSb (x = 0.3) is therefore essential, requiring both experimental and theoretical investigations. In this context, our focus is on muon spin rotation/relaxation/resonance (μSR) measurements of the magnetic penetration depth, λ in these superconductors15,16. This parameter is fundamental to understanding superconductivity, as it is directly related to the superfluid density, ns through the expression, 1/λ2 = μ0e2ns/m* (where m* is the effective mass). The temperature dependence of λ is particularly sensitive to the structure of the superconducting gap15,17. Moreover, zero-field μSR proves to be a powerful tool for detecting a spontaneous magnetic field arising from time-reversal symmetry (TRS) breaking in exotic superconductors18–21. This is particularly noteworthy as internal magnetic fields as small as 0.1 G can be detected in measurements without the application of external magnetic fields. These investigations aim to unveil the intricate details of the superconducting state in Ir1−δSb and Ir1−xRhxSb, and contribute to our broader understanding of unconventional superconductivity in these materials.
We report on the fully gapped and time-reversal invariant superconducting state in the bulk of Ir1−δSb and Ir0.7Rh0.3Sb. The fully gapped nature suggests a well-defined energy structure in the superconducting state, while time-reversal invariance emphasizes the preservation of fundamental symmetries in the superconducting order parameter. The zero-temperature limit of the penetration depth was evaluated to be approximately 319(7) nm and 639(10) nm for Ir1−δSb and Ir0.7Rh0.3Sb, respectively. The Tc/ ratio was found to be comparable to that of unconventional superconductors. The relatively high critical temperature (Tc) despite a small carrier density raises the intriguing possibility of an unconventional pairing mechanism in Ir-Sb binary superconductors. This observation opens avenues for further investigation into the underlying physics of superconductivity in these materials, potentially revealing mechanisms that contribute to their unique superconducting properties.
Results and discussion
The investigation into the possible magnetism, both static and fluctuating, in Ir1−δSb and Ir1−xRhxSb with x = 0.3 involved zero-field muon spin relaxation (ZF-μSR) experiments conducted both above and below the critical temperature, Tc. Figure 1a and b illustrate that, down to 1.7 K, no evidence of either static or fluctuating magnetism was detected in the ZF time spectra. The ZF-μSR spectra can be well described by the Gaussian Kubo-Toyabe (GKT) depolarization function22:
1 |
Here, Δ/γμ is the width of the local field distribution due to the presence of the dense system of nuclear moments. γμ = 0.085 μs−1G−1 is the muon gyromagnetic ratio. The inset of Fig. 1a shows the temperature evolution of Δ, which shows no enhancement across Tc. The initial asymmetry was also found to be temperature independent and remained constant throughout the analysis. The absence of any change in the ZF-μSR relaxation rate across Tc suggests the lack of spontaneous magnetic fields associated with a time-reversal symmetry (TRS) breaking pairing state in Ir1−δSb and Ir1−xRhxSb with x = 0.3.
Figure 2a, b depicts the TF-μSR time spectra for Ir1−δSb and Ir0.7Rh0.3Sb, respectively. These measurements were conducted in an applied magnetic field of 30 mT, both above (4 K) and below (0.08 K) the superconducting transition temperature, Tc. The measurements for Ir0.7Rh0.3Sb were also conducted under an applied magnetic field of 10 mT. The data were collected after field-cooling the sample from above TC. The applied field was selected based on the criterion Bc1<B ≪ Bc2. Under this condition, a well-ordered Abrikosov vortex lattice is formed. The zero-temperature limits of Bc1 and Bc2 of Ir1−δSb synthesized at 5.5 GPa are determined to be 28 mT and 1.19 T, respectively13. Consequently, the calculated Ginzburg-Landau parameter κ is estimated to be κ ≃ 6.2. For Ir0.7Rh0.3Sb, the zero-temperature limits of Bc1 and Bc2 are determined to be 8 mT and 0.9 T, respectively13. These values confirm the suitability of our applied magnetic field. Above Tc the oscillations show a small relaxation due to the random local fields from the nuclear magnetic moments. At 0.08 K, the relaxation rate increases due to the formation of a flux-line lattice (FLL) in the superconducting state, resulting in a nonuniform local field distribution. It is noteworthy that the rise in relaxation rate in the superconducting state is more pronounced in Ir1−δSb compared to Ir0.7Rh0.3Sb. This distinction is further evident in the Fourier transforms (see Fig. 2c, d) of the μSR time spectra, highlighting a significant broadening of the signal in the superconducting state for Ir1−δSb, whilst the spectra are almost identical above and below Tc in Ir0.7Rh0.3Sb.
As denoted by the solid lines in Fig. 2a, c, TF-μSR data were analyzed using the following functional form23:
2 |
The field distribution in the superconducting state of the sample Ir1−δSb synthesized at 5.5 GPa exhibits a noticeable broadening with an asymmetric line shape, as anticipated in the vortex state (see Fig. 2c). As a result, a two-component expression was utilized to accurately represent this broad and asymmetric field distribution. Conversely, the field distribution for the sample Ir0.7Rh0.3Sb (see Fig. 2d) in the superconducting state displays a significantly narrower width compared to Ir1−δSb and can be adequately represented by a single-component Gaussian. In Eq. (2), As,i, Bint,s,i and σsc,i are the the initial asymmetry, the internal magnetic field at the muon site and the superconducting relaxation rates of the i-th component. σnm characterizes the damping due to the nuclear magnetic dipolar contribution. Since the relaxation rate remains nearly unchanged above Tc upon the application of a magnetic field, this indicates the absence of field-induced magnetism, with only nuclear magnetic moments contributing to the normal state muon depolarization rate. During the analysis σnm was assumed to be constant over the entire temperature range and was fixed to the value obtained above Tc. In order to extract the superconducting muon spin depolarization rate σsc (the second moment of the field distribution) and Bint,s (the first moment of the field distribution) from the two-component fit we used the same procedure as described in ref. 24.
In Fig. 3a, σsc is plotted against temperature for both Ir1−δSb, synthesized at 5.5 GPa (at μ0H = 0.03 T), and for Ir0.7Rh0.3Sb (at μ0H = 0.01 T and 0.03 T). Below Tc, the relaxation rate σsc begins to increase from zero due to the formation of the flux-line lattice (FLL) and saturates at low temperatures. The temperature dependence of σsc reflects the topology of the superconducting gap and is consistent with the presence of a single superconducting gap on the Fermi surface of these materials, as we show below. The absolute value of σsc is five times smaller for Ir0.7Rh0.3Sb compared to Ir1−δSb, indicating a lower superfluid density for the Rh-doped sample. Below Tc, a large diamagnetic shift of Bint,s experienced by the muons is observed in both samples. In Fig. 3b, the temperature dependence of the diamagnetic shift ΔBdia = Bint,s,SC - Bint,s,NS is plotted, where Bint,s,SC represents the internal field measured in the superconducting state, and Bint,s,NS is the internal field measured in the normal state at 5 K. This diamagnetic shift indicates the bulk nature of superconductivity and rules out the possibility of field-induced magnetism in these superconductors.
To perform a quantitative analysis, it is important to note that the London magnetic penetration depth, λ(T) is directly related to the measured relaxation rate in the superconducting state, σsc. The triangular FLL relationship is described by the equation25:
3 |
where Φ0 = 2.068 × 10−15 Wb is the magnetic-flux quantum. Equation (3) is applicable only when the separation between vortices is smaller than λ. In this particular scenario, as per the London model, σsc becomes field-independent25.
To explore the superconducting gap structure of Ir1−δSb and Ir0.7Rh0.3Sb, we conducted an analysis of the temperature dependence of the magnetic penetration depth, λ(T), directly linked to the superconducting gap. The behavior of λ(T) can be characterized within the local (London) approximation (λ ≫ ξ) using the following expression23,26:
4 |
where is the Fermi function, φ is the angle along the Fermi surface, and Δi(T, φ) = Δ0,iΓ(T/Tc)g(φ) (Δ0,i is the maximum gap value at T = 0). The temperature dependence of the gap is approximated by the expression 27. while g(φ) describes the angular dependence of the gap and it is replaced by 1 for an s-wave gap, [1+acos(4φ)/(1+a)] for an anisotropic s-wave gap and for a d-wave gap28.
In Fig. 3a, the experimentally obtained (T) dependence is most accurately described by a momentum-independent s-wave model with a gap value of Δ = 0.6(1) meV and Tc = 4.1(1) K for Ir1−δSb, and a gap value of Δ = 0.4(1) meV and Tc = 2.7(2)K for Ir0.7Rh0.3Sb. The d-wave and p-wave gap symmetries were also considered but were found to be inconsistent with the data (illustrated by the dashed line in Fig. 3a). Particularly, these models struggle to account for the very weak temperature dependence of λ(T) at low temperatures. We note that the (px + ipy) pairing symmetry is also characterized by the full gap in 2D systems and should also saturate at low temperatures. However, the possibility of px + ipy pairing is excluded by the absence of a TRS breaking state. It should also be noted, though, that saturation of the muon depolarization rate at low temperatures can arise even in a nodal d-wave29–34 superconductor, but in the dirty limit. In this case a power law temperature dependence was proposed theoretically35. We tested this function but deemed it inconsistent with the data (see Fig. 3a). Therefore, our analysis shows that a nodeless or fully gapped state is the most plausible bulk superconducting pairing state for Ir1−δSb and Ir0.7Rh0.3Sb. Our specific heat results down to 300 mK for Ir1−δSb, synthesized at 5.5 GPa (see Fig. 4), measured on the same sample used for muon spin rotation experiments, show saturation at low temperatures and a large jump at Tc, providing the confirmation for bulk nodeless superconductivity.
The estimated ratio of the superconducting gap to Tc, (2Δ/kBTc), is approximately 3.4, aligning with the BCS (Bardeen-Cooper-Schrieffer) expectation15. However, it’s crucial to acknowledge that a similar ratio can also be anticipated within a Bose-Einstein Condensation (BEC)-like framework. Importantly, the ratio 2Δ/kBTc, on its own, does not effectively distinguish between BCS or BEC condensation scenarios. Further insights are required to differentiate between these two possibilities and elucidate the nature of the superconducting state in the studied materials. What distinguishes between BCS and BEC superconductivity is a key parameter: the ratio of the superconducting critical temperature to the superfluid density. This ratio, Tc/ns, plays a crucial role in characterizing the nature of the superconducting state in different materials. In a simplified interpretation of the BEC to BCS crossover, the Tc/ns ratio serves as a critical parameter. Systems characterized by a small Tc/ns (with a large superfluid density, ns) are often considered to reside on the “BCS” side of the crossover. Conversely, systems with a large Tc/ns (exhibiting a small superfluid density, ns) are expected to be on the BEC side. Moreover, the correlation between Tc and the superfluid density is anticipated to be significant primarily on the BEC side of the crossover. As one moves from the BCS limit to the BEC limit in the crossover, the nature of the pairing mechanism evolves, transitioning from Cooper pairs formed through electron-phonon interactions (BCS) to a more Bose-Einstein condensation-like scenario involving preformed pairs.
The observation of a correlation between Tc and the superfluid density () was first noted in hole-doped cuprates back in 1988–8936,37, extending later to include electron-doped cuprates38. This intriguing relationship has been investigated across various superconducting systems. Guguchia and collaborators demonstrated that this linear correlation is an intrinsic feature in superconductors such as transition metal dichalcogenides15,39 and kagome-lattice superconductors20,40,41. The ratio Tc/ in these systems tends to be lower than that observed in hole-doped cuprates (see Fig. 5). To contextualize the superconductors Ir1−δSb and Ir0.7Rh0.3Sb within this framework, Fig. 5 illustrates the critical temperature plotted against the superfluid density. For Ir0.7Rh0.3Sb, the estimated ratio Tc/ is approximately 1 (K/μm−2), closely resembling electron-doped cuprates known for their correlated superconductivity. In the case of Ir1−δSb, the ratio is reduced to 0.4 (K/μm−2) but remains notably distant from conventional BCS superconductors. Intriguingly, it aligns nearly perfectly with the trend line occupied by charge density wave superconductors like 2H-NbSe2, 4H-NbSe2, LaRu3Si2, as well as the Weyl-superconductor Td-MoTe215. This finding strongly suggests an unconventional pairing mechanism in Ir1−δSb and Ir0.7Rh0.3Sb, characterized by a low density of Cooper pairs. It is beyond the scope of this work to comment on the microscopic origin of the similarity between Ir-Sb and other superconductors.
Another unconventional feature in the superconducting phase diagram of Ir1−xRhxSb is a dome-shaped dependence of Tc (see Fig. 6a). This pattern is characterized by an optimal Tc value occurring at x = 0.3, followed by a reduction as the Rh concentration deviates from this optimal point. The isovalent Rh substitution on the Ir site in IrSb, without introducing additional holes or electrons, creates a condition often termed “chemical pressure”. Typically, chemical substitution introduces disorder effects, potentially influencing Tc. To discern the intrinsic nature of this dome shape, a cleaner external parameter is essential. For example, hydrostatic pressure introduces fewer disorder effects compared to isovalent chemical substitutions ("chemical pressure”). For this reason, we explored the impact on Tc in the optimally Rh-doped, Ir0.7Rh0.3Sb with hydrostatic pressure, spanning a range up to p = 2.2 GPa. As shown in Fig. 6b, c, the observed trend revealed a linear decrease in Tc with increasing pressure. This behavior qualitatively mirrors the effect of Rh doping. The figure is intended to illustrate the linear suppression of Tc by both hydrostatic pressure and Rh content. In the manuscript, we do not provide a quantitative relationship between hydrostatic pressure and the chemical pressure induced by Rh content. Our main point is that this finding suggests the impact of both external hydrostatic pressure and chemical modifications is consistent in the Ir0.7Rh0.3Sb system. This underscores that the reduction in Tc beyond x = 0.3, resulting in a dome-shaped dependence, is primarily due to chemical pressure rather than disorder.
Conclusions
In summary, our study provides a microscopic exploration of superconductivity in Ir1−δSb (synthesized at 5.5 GPa with Tc = 4.2 K) and optimally Rh-doped Ir0.7Rh0.3Sb (Tc = 2.7 K) in close proximity to vacancy ordering, employing a bulk sensitive local probe. Specifically, we investigated the zero-temperature magnetic penetration depth and the temperature dependence of through μSR experiments. The superfluid density in both systems aligns with a scenario of a complete gap. Intriguingly, the Tc/ ratio is comparable to that of high-temperature unconventional superconductors, suggesting the unconventional nature of superconductivity in Ir-Sb binary superconductors. Additionally, the μSR experiments, serving as an extremely sensitive magnetic probe, do not exhibit evidence of spontaneous magnetic fields, which would be expected for a time-reversal-symmetry-breaking state in the bulk of the superconductor. Consequently, our results categorize Ir1−δSb and Ir0.7Rh0.3Sb as unconventional, time-reversal-invariant, and fully gapped bulk superconductors. We further demonstrate the similarity between the effects of chemical pressure, induced by isovalent Rh substitution, and hydrostatic pressure on the superconducting critical temperature in Ir0.7Rh0.3Sb. This highlights that the observed dome-shaped dependence of Tc is not merely a consequence of disorder effects introduced by chemical substitution but is rooted in the intrinsic properties of the material. These results offer valuable insights into the underlying mechanisms governing the materials behavior. A more comprehensive analysis requires consideration of various factors, including the specific pairing mechanisms and the role of interactions in the superconducting state.
Methods
Sample growth and characterization
The details of the synthesis of the polycrystalline samples of Ir1−δSb and Ir0.7Rh0.3Sb are reported elsewhere13,14.
Sample characterization
A comprehensive characterization of the δ parameter was conducted in our previous publication13. Using an electron probe microanalyzer (EPMA), we determined the Ir deficiency in the Ir1−δSb series, observing that the filling rate increased gradually with higher synthesis pressure. At 5.5 GPa, the δ was approximately 4%. The characterization of the basic physical properties (magnetic susceptibility, electrical resistivity, and heat capacity) for Ir1−δSb at various synthesized external pressures and Ir1−xRhxSb samples is also detailed in the same publication13. Samples from the same batch were used in the current investigation.
Experimental details
Zero field (ZF) and transverse field (TF) μSR experiments were performed on the GPS (πM3 beamline)42, and high-field HAL-9500 instruments (πE3 beamline)43, equipped with BlueFors vacuum-loaded cryogen-free dilution refrigerator (DR), at the Swiss Muon Source (SμS) at the Paul Scherrer Institute, in Villigen, Switzerland. Zero field is dynamically obtained (compensation better than 30 mG) by a newly installed automatic compensation device42. When performing measurements in zero-field the geomagnetic field or any stray fields are tabulated and automatically compensated by the automatic compensation device. The μSR time spectra were analyzed using the free software package MUSRFIT23. AC susceptibility experiments under pressure were carried out using double wall piston-cylinder type of cell made of MP35N/MP35N material generate pressures up to 2.3 GPa40,44,45. A small indium plate was placed together with the sample in the pressure cell filled with the Daphne oil. The pressure was estimated by tracking the SC transition of a indium plate by AC susceptibility.
Acknowledgements
The μSR experiments were carried out at the Swiss Muon Source (SμS) Paul Scherrer Insitute, Villigen, Switzerland. Z.G. acknowledges support from the Swiss National Science Foundation (SNSF) through SNSF Starting Grant (No. TMSGI2_211750). Z.G. acknowledges the useful discussions with Dr. Robert Scheuermann. T.Y. would like to acknowledge the National Natural Science Foundation of China (Grant No. 52250308).
Author contributions
Z.G. conceived and supervised the project. Sample Growth: Y.T. and H.H. μSR experiments, data analysis and corresponding discussions: V.S., J.N.G., D.D., C.M.III., S.S.I., R.K., H.L., and Z.G. Specific heat experiments and corresponding discussions: V.S., S.S., M.M., M.B., and Z.G. Figure development and writing of the paper: Z.G., and V.S., with contributions from all authors. All authors discussed the results, interpretation and conclusion.
Peer review
Peer review information
Communications Physics thanks James Annett and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.
Data availability
All relevant data are available from the authors. Alternatively, the data can be accessed through the data base at the following link http://musruser.psi.ch/cgi-bin/SearchDB.cgiusing the following details: 1. Area: HAL. Year: 2020. Run from 0107 to 0198 and from from 0294 to 0356. 2. Area: GPS. Year: 2020. Run from 0014 to 0037. Year: 2024. Run from 2004 to 2008.
Competing interests
The authors declare no competing interests.
Footnotes
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
References
- 1.Ataca, C. et al. Mechanical and Electronic Properties of MoS2 Nanoribbons and Their Defects. J. Phys. Chem. C.115, 3934–3941 (2011). [Google Scholar]
- 2.Luo, Y. et al. Defect Engineering of Nanomaterials for Catalysis. Nanomaterials13, 1116 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Guguchia, Z. et al. Magnetism in Semiconducting Molybdenum Dichalcogenides. Sci. Adv.4, eaat3672 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Tongay, S. et al. Magnetic properties of MoS2: Existence of ferromagnetism. Appl. Phys. Lett.101, 123105 (2012). [Google Scholar]
- 5.Wang, Q. H. et al. The Magnetic Genome of Two-Dimensional van der Waals Materials. ACS Nano16, 6960–7079 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Fang, Y. et al. Fe-vacancy ordering in superconducting K1−xFe2−ySe2: first-principles calculations and Monte Carlo simulations. Supercond. Sci. Technol.28, 095004 (2015). [Google Scholar]
- 7.Zhang, A. M. et al. Vacancy ordering and phonon spectrum of the iron-based superconductor K0.8Fe1.6Se2. Phys. Rev. B85, 024518 (2012). [Google Scholar]
- 8.Chen, T. K. et al. Fe-vacancy order and superconductivity in tetragonal β-Fe1−xSe. PNAS111, 63–68 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Kazakov, S. M. et al. Uniform patterns of Fe-vacancy ordering in the Kx(Fe,Co)2−ySe2 superconductors. Chem. Mater.23, 4311–4316 (2011). [Google Scholar]
- 10.Ourmazd, A. & Spence, J. C. H. Detection of oxygen ordering in superconducting cuprates. Nature329, 425–427 (1987). [Google Scholar]
- 11.Zimmermann, M. V. et al. Oxygen-ordering superstructures in underdoped YBa2Cu3O6+x studied by hard x-ray diffraction. Phys. Rev. B68, 104515 (2003). [Google Scholar]
- 12.Li, Y., Du, S., Weng, Z.-Y. & Liu, Z. In-plane ordering of oxygen vacancies in a high-Tc cuprate superconductor with compressed Cu-O octahedrons: An automated cluster expansion study. Phys. Rev. Mater.4, 044801 (2020). [Google Scholar]
- 13.Qi, Y. et al. Superconductivity from buckled-honeycomb-vacancy ordering. Sci. Bull.66, 327–331 (2021). [DOI] [PubMed] [Google Scholar]
- 14.Ying, T. et al. Fermi surface nesting, vacancy ordering and the emergence of superconductivity in IrSb compounds. arXiv, https://arxiv.org/abs/2108.13704 (2021).
- 15.Guguchia, Z. et al. Signatures of the topological s+− superconducting order parameter in the type-II Weyl semimetal Td-MoTe2. Nat. Commun.8, 1082 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Sonier, J. E., Brewer, J. H. & Kiefl, R. F. µSR studies of the vortex state in type-II superconductors. Rev. Modern Phys.72, 769 (2000).
- 17.Das, D. et al. Unconventional Pressure Dependence of the Superfluid Density in the Nodeless Topological Superconductor α-PdBi2. Phys. Rev. Lett.127, 217002 (2021). [DOI] [PubMed] [Google Scholar]
- 18.Luke, G. M. et al. Time-reversal symmetry-breaking superconductivity in Sr2RuO4. Nature394, 559 (1998). [Google Scholar]
- 19.Hillier, A. D., Jorge, Q. & Cywinski, R. Evidence for Time-Reversal Symmetry Breaking in the Noncentrosymmetric Superconductor LaNiC2. Phys. Rev. Lett.102, 117007 (2009). [DOI] [PubMed] [Google Scholar]
- 20.Guguchia, Z., Khasanov, R. & Luetkens, H. Unconventional charge order and superconductivity in kagome-lattice systems as seen by muon-spin rotation. npj Quantum Mater.8, 41 (2023). [Google Scholar]
- 21.Mielke III, C. et al. Time-reversal symmetry-breaking charge order in a kagome superconductor. Nature602, 245–250 (2022). [DOI] [PubMed] [Google Scholar]
- 22.Kubo, R. & Toyabe, T. In Magnetic Resonance and Relaxation, 810; T. Toyabe, M.S. thesis (University of Tokyo, 1966).
- 23.The fitting of Eq. 4 was performed using the additional library BMW developped by B. Wojek. Phys. Proc.30, 69 (2012).
- 24.Khasanov, R. et al. Experimental Evidence for Two Gaps in the High Temperature La1.83Sr0.17CuO4 Superconductor. Phys. Rev. Lett.98, 057007 (2007). [DOI] [PubMed] [Google Scholar]
- 25.Brandt, E. H. Flux distribution and penetration depth measured by muon spin rotation in high-Tc superconductors. Phys. Rev. B37, 2349 (1988). [DOI] [PubMed]
- 26.Tinkham, M. Introduction to Superconductivity (Krieger Publishing Company, 1975).
- 27.Carrington, A. & Manzano, F. Magnetic penetration depth of MgB2. Physica C385, 205 (2003).
- 28.Fang, M. H. et al. Superconductivity close to magnetic instability in Fe(Se1-xTex)0.82. Phys. Rev. B78, 224503 (2008).
- 29.Khasanov, R. et al. Correlation between oxygen isotope effects on transition temperature and magnetic penetration depth in high-temperature superconductors close to optimal doping. Phys. Rev. B74, 064504 (2006).
- 30.Sonier, J.E. μSR studies of cuprate superconductors. J. Phys. Soc. Japan85, 091005 (2016).
- 31.Luetkens, H. et al. Field and temperature dependence of the superfluid density in LaFeAsO1−xFx superconductors: a muon spin relaxation study. Phys. Rev. Lett. 101, 097009 (2008). [DOI] [PubMed]
- 32.Harshman, D. R. & Fiory, A.T. Concerning the superconducting gap symmetry in YBa2Cu3O7−δ, YBa2Cu4O8, and La2−xSrxCuO4 determined from muon spin rotation in mixed states of crystals and powders. J. Phys. Condens. Matter23, 315702 (2011). [DOI] [PubMed]
- 33.Leng, H. et al. Muon spin rotation study of the topological superconductor SrxBi2Se3. Phys. Rev. B97, 054503 (2018).
- 34.Neha, P., Biswas, P. K., Das, T. & Patnaik, S. Time-reversal symmetry breaking in topological superconductor Sr0.1Bi2Se3. Phys. Rev. Materials3, 074201 (2019).
- 35.Hirschfeld, P. J., Putikka, W. O. & Scalapino, D. J. d-wave Model for Microwave Response of High-Tc Superconductors. Phys. Rev. B50, 10250 (1994). [DOI] [PubMed] [Google Scholar]
- 36.Uemura, Y. J. et al. Universal Correlations between Tc and ns/m* (Carrier Density over Effective Mass) in High-Tc Cuprate Superconductors. Phys. Rev. Lett.62, 2317 (1989). [DOI] [PubMed] [Google Scholar]
- 37.Uemura, Y. J. et al. Basic Similarities among Cuprate, Bismuthate, Organic, Chevrel Phase, and Heavy-Fermion Superconductors Shown by Penetration Depth Measurements. Phys. Rev. Lett.66, 2665 (1991). [DOI] [PubMed] [Google Scholar]
- 38.Shengelaya, A. et al. Muon-Spin-Rotation Measurements of the Penetration Depth of the Infinite-Layer Electron-Doped Sr0.9La0.1CuO2 Cuprate Superconductor. Phys. Rev. Lett.94, 127001 (2005). [DOI] [PubMed] [Google Scholar]
- 39.von Rohr, F. O. et al. Unconventional Scaling of the Superfluid Density with the Critical Temperature in Transition Metal Dichalcogenides. Sci. Adv.5, eaav8465 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Guguchia, Z. et al. Direct evidence for the emergence of a pressure induced nodal superconducting gap in the iron-based superconductor Ba0.65Rb0.35Fe2As2. Nat. Commun.6, 8863 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Guguchia, Z. et al. Tunable unconventional kagome superconductivity in charge ordered RbV3Sb5 and KV3Sb5. Nat. Commun.14, 153 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Amato, A. et al. The new versatile general purpose surface-muon instrument (GPS) based on silicon photomultipliers for μSR measurements on a continuous-wave beam. Rev. Sci. Instrum.88, 093301 (2017). [DOI] [PubMed] [Google Scholar]
- 43.Sedlak, K., Scheuermann, R., Stoykov, A. & Amato, A. GEANT4 simulation and optimisation of the high-field μSR spectrometer. Phys. B404, 970–973 (2009). [Google Scholar]
- 44.Khasanov, R. et al. High pressure research using muons at the Paul Scherrer Institute. High. Press. Res.36, 140–166 (2016). [Google Scholar]
- 45.Khasanov, R. Perspective on the muon-spin rotation/relaxation under hydrostatic pressure. J. Appl. Phys.132, 190903 (2022). [Google Scholar]
- 46.Mielke III, C. et al. Nodeless kagome superconductivity in LaRu3Si2. Phys. Rev. Mat.5, 034803 (2021). [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Data Availability Statement
All relevant data are available from the authors. Alternatively, the data can be accessed through the data base at the following link http://musruser.psi.ch/cgi-bin/SearchDB.cgiusing the following details: 1. Area: HAL. Year: 2020. Run from 0107 to 0198 and from from 0294 to 0356. 2. Area: GPS. Year: 2020. Run from 0014 to 0037. Year: 2024. Run from 2004 to 2008.