This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Single-gap Isotropic ss-wave Superconductivity
in Single Crystals AuSn4\text{AuSn}_{4}

Sunil Ghimire Ames National Laboratory, Ames, Iowa 50011, USA Department of Physics & Astronomy, Iowa State University, Ames, Iowa 50011, USA    Kamal R. Joshi Ames National Laboratory, Ames, Iowa 50011, USA Department of Physics & Astronomy, Iowa State University, Ames, Iowa 50011, USA    Elizabeth H. Krenkel Ames National Laboratory, Ames, Iowa 50011, USA Department of Physics & Astronomy, Iowa State University, Ames, Iowa 50011, USA    Makariy A. Tanatar Ames National Laboratory, Ames, Iowa 50011, USA Department of Physics & Astronomy, Iowa State University, Ames, Iowa 50011, USA    Marcin Kończykowski Laboratoire des Solides Irradiés, CEA/DRF/lRAMIS, École Polytechnique, CNRS, Institut Polytechnique de Paris, F-91128 Palaiseau, France    Romain Grasset Laboratoire des Solides Irradiés, CEA/DRF/lRAMIS, École Polytechnique, CNRS, Institut Polytechnique de Paris, F-91128 Palaiseau, France    Paul C. Canfield Ames National Laboratory, Ames, Iowa 50011, USA Department of Physics & Astronomy, Iowa State University, Ames, Iowa 50011, USA    Ruslan Prozorov [email protected] Ames National Laboratory, Ames, Iowa 50011, USA Department of Physics & Astronomy, Iowa State University, Ames, Iowa 50011, USA
(3 July 2024)
Abstract

London, λL(T)\lambda_{L}(T), and Campbell, λC(T)\lambda_{C}(T), penetration depths were measured in single crystals of a topological superconductor candidate AuSn4\text{AuSn}_{4}. At low temperatures, λL(T)\lambda_{L}(T) is exponentially attenuated and, if fitted with the power law, λ(T)Tn\lambda(T)\sim T^{n}, gives exponents n>4n>4, indistinguishable from the isotropic single ss-wave gap Bardeen-Cooper-Schrieffer (BCS) asymptotic. The superfluid density fits perfectly in the entire temperature range to the BCS theory. The superconducting transition temperature, Tc=2.40±0.05KT_{c}=2.40\pm 0.05\>\text{K}, does not change after 2.5 MeV electron irradiation, indicating the validity of the Anderson theorem for isotropic ss-wave superconductors. Campbell penetration depth before and after electron irradiation shows no hysteresis between the zero-field cooling (ZFC) and field cooling (FC) protocols, consistent with the parabolic pinning potential. Interestingly, the critical current density estimated from the original Campbell theory decreases after irradiation, implying that a more sophisticated theory involving collective effects is needed to describe vortex pinning in this system. In general, our thermodynamic measurements strongly suggest that the bulk response of the AuSn4\text{AuSn}_{4} crystals is fully consistent with the isotropic ss-wave weak-coupling BCS superconductivity.

I Introduction

In recent years, superconductors with topological features in their electronic bandstructure have attracted significant interest for various novel features predicted by a well-developed theory. For example, emerging zero-energy excitations called Majorana fermions  [1]. On the material side, the search for topological superconductors (TSCs) is very active but so far has yielded only a few “candidates” whose topological properties have not yet been fully confirmed experimentally, including UTe2 [2], Sr2RuO4 [3, 4, 5], UPt3 [6], 2M-WS2 [7], and MxBi2Se3 with M=Cu [8, 9]. The subject of this study, AuSn4\text{AuSn}_{4}, is another promising TSC candidate with theoretically predicted non-trivial topological characteristics [10, 11, 12, 13].

The superconductivity in orthorhombic AuSn4\text{AuSn}_{4} with a transition temperature to the superconducting state, Tc=2.4KT_{c}=2.4\>\text{K}, was discovered in 1962 [14]. This compound is isostructural to PtSn4\text{PtSn}_{4} [15] and PdSn4\text{PdSn}_{4} [16], which are not superconductors. The first principal study suggests semimetallic behavior with type I nodes [12]. The magneto-trasnport measurements show two-dimensional (2D) superconductivity in AuSn4\text{AuSn}_{4} [11, 17]. Recently, ARPES measurements supported by DFT calculations [13] revealed nearly degenerate polytypes in AuSn4\text{AuSn}_{4} crystals, making it a unique case of a three-dimensional (3D) electronic band structure with properties of a low-dimensional layered material. Thermodynamic magnetization and specific heat measurement in AuSn4\text{AuSn}_{4} single crystals are consistent with conventional nodeless ss-wave Bardeen-Cooper-Schrieffer (BCS) [18, 19] superconductivity [11]. Scanning tunneling microscopy (STM) measurements determined the superconducting gap to TcT_{c} ratio close to the ss-wave BCS value of Δ/Tc=1.76\Delta/T_{c}=1.76 [13]. However, other STM measurements suggest unconventional 2D superconductivity with a mixture of pp-wave surface states and ss-wave bulk [10]. Clearly, more measurements are required for an objective and conclusive determination of the nature of superconductivity in AuSn4\text{AuSn}_{4}.

Here, we probe the bulk nature of superconductivity in AuSn4\text{AuSn}_{4} single crystals by measuring London and Campbell penetration depths using a highly sensitive tunnel-diode resonator (TDR). Furthermore, we examine the response to a controlled non-magnetic point-like disorder induced by 2.5 MeV electron irradiation. We conclude that AuSn4\text{AuSn}_{4} is a robust isotropic ss-wave superconductor in the bulk. However, we cannot exclude the possibility that it could have a different type of superconductivity in the surface atomic layers, where the STM is most sensitive.

II Samples and Methods

Single crystals of AuSn4\text{AuSn}_{4} were grown with excess Sn flux [20, 21, 13]. High-purity Au and Sn were mixed in a 12:88 ratio in a fritted crucible and sealed in a quartz ampoule under an Ar gas atmosphere. The ampoule was heated to 11001100\>^{\circ}C over 12 hours, then cooled to 250250\>^{\circ}C in 12 hours, and significantly slower to 230230\>^{\circ}C over 90 hours. The ampoule was held at this temperature for 48 hours prior to removal from the furnace.

The London penetration depth, λ(T)\lambda(T), was measured using a sensitive frequency-domain self-oscillating tunnel-diode resonator (TDR) operating at a frequency of around 14 MHz. The measurements were performed in a cryostat 3 He with a base temperature of \approx 400 mK, which is 0.17Tc0.17T_{c}, allowing us to examine the low-temperature limit, which starts below approximately Tc/3T_{c}/3, where the superconducting gap is approximately constant. The experimental setup, measurement protocols, and calibration are described in detail elsewhere [22, 23, 24, 25, 26]. In the experiment, the variation, Δλ(T)=λ(T)λ(0.4K)\Delta\lambda(T)=\lambda(T)-\lambda(0.4\>\text{K}), is extracted from the resonant frequency shift. The small excitation magnetic field of 20Oe20\>\text{Oe} ensures a true Campbell regime when vortices are gently perturbed. For TDR measurements, the samples were cut into cuboids that are typically of size 0.6×0.4×0.1mm30.6\times 0.4\times 0.1\>\textrm{mm}^{3}.

Point-like disorder was introduced at the SIRIUS facility in the Laboratoire des Solides Irradiés at École Polytechnique in Palaiseau, France. Electrons, accelerated in a pelletron-type linear accelerator to 2.5 MeV, knock out ions, creating vacancy-interstitial Frenkel pairs [27, 28]. During irradiation, the sample is immersed in liquid hydrogen at around 20 K. This ensures efficient heat removal upon impact and prevents immediate recombination and migration of the produced atomic defects. The acquired irradiation dose is determined by measuring the total charge collected by a Faraday cage located behind the sample. As such, the acquired dose is measured in the “natural” units of C/cm2\text{C/cm}^{2}, which is equal to 1C/cm21/e6.24×10181\>\text{C/cm}^{2}\equiv 1/e\approx 6.24\times 10^{18} electrons per cm2. Upon warming to room temperature, some defects recombine, and some migrate to various sinks (dislocations, surfaces, etc.). This leaves a metastable population, about 70%, of point-like defects [29, 30]. Importantly, the same sample has been measured before and after electron irradiation.

III Results

III.1 London penetration depth

Refer to caption
Figure 1: Main Panel: Low-temperature temperature variation of the London penetration depth Δλ(T)=λ(T)λ(0.4K)\Delta\lambda(T)=\lambda(T)-\lambda(0.4\>\text{K}) as a function of normalized temperature, t=T/Tct=T/T_{c}, for pristine (blue circles) and irradiated at 2.5 C/cm2 (red circles) single crystal of AuSn4\text{AuSn}_{4}. Lines show fits to the power law, Δλ(T)Atn\Delta\lambda(T)\sim At^{n}, with the upper range of tmax=0.4t_{max}=0.4. The top right inset shows the Δλ(T)\Delta\lambda(T) in the whole temperature range, showing sharp superconducting transition with onset Tc=2.4KT_{c}=2.4\>\text{K} for both pristine and electron irradiated state. The top left inset shows the exponent nn versus the upper limit of the power-law fitting, tmax=Tmax/Tct_{max}=T_{max}/T_{c}, indicating robustness of the power law, experimentally indistinguishable from exponential.

Figure 1 shows the low-temperature dependence of the change in the London penetration depth, Δλ(T)=λ(T)λ(Tmin=0.4K)\Delta\lambda(T)=\lambda(T)-\lambda(T_{min}=0.4\>\text{K}) before (blue circles) and after 2.5 C/cm2 electron irradiation (red circles). The upper left inset shows the exponent nn determined from the power-law fitting, Δλ(T)Atn\Delta\lambda(T)\sim At^{n}, as a function of the upper fitting limit, tmax=Tmax/Tct_{max}=T_{max}/T_{c}. The solid lines in the main frame show an example of such a fitting with tmax=0.4t_{max}=0.4. The results show a robust and consistent behavior with n4n\geq 4, indicating experimentally indistinguishable from the exponential temperature dependence. The exponent, nn, decreased after irradiation as it should be in an ss-wave superconductor [31, 32].

The upper right inset of Fig.1 shows Δλ(T)\Delta\lambda(T) of the same sample in its pristine state and after 2.15 C/cm2 electron irradiation as a function of absolute temperature TT. One might think that for some reason (e.g., defect annealing and recombination), there was no increase in disorder after irradiation. This is not the case, as the saturation value above TcT_{c} increased substantially. The saturation is determined by the skin depth in the normal state, δskin=ρ/μ0πf\delta_{\text{skin}}=\sqrt{\rho/\mu_{0}\pi f}, where μ0=4π×107,H/m\mu_{0}=4\pi\times 10^{-7},\text{H/m} is the vacuum permeability, and ρ\rho is the resistance. We did not measure resistivity in this AuSn4\text{AuSn}_{4} sample, but we directly compared resistivity from transport measurements and extracted from the skin depth on the same samples in other compounds and always found good quantitative agreement [33, 34]. The upper critical fields are small, Hc2ab=130OeH^{\parallel ab}_{c2}=130\>\text{Oe} and Hc2c=90OeH^{\parallel c}_{c2}=90\>\text{Oe} [11]. Combined with the trend of measured magnetoresistance [13], the expected variation above TcT_{c} is negligible. An increase in δskin\delta_{\text{skin}} at a fixed frequency, ff, is due to an increase in resistivity, which is indicative of the increased scattering. Therefore, the fact that the superconducting transition TcT_{c} remains unchanged is consistent with the Anderson theorem for isotropic ss-wave superconductors [35, 36]. We observe similar robust superconductivity in another lowTc-T_{c} superconductor with non-trivial topology, LaNiGa2 [37].

Refer to caption
Figure 2: Top panel. Fit to the BCS low-temperature asymptotic, Δλ(T)=λ(0)πδ2teδt\Delta\lambda(T)=\lambda(0)\sqrt{\frac{\pi\delta}{2t}}e^{-\frac{\delta}{t}} with a fixed ratio δ=Δ(0)/Tc1.76\delta=\Delta(0)/T_{c}\approx 1.76 leaving only one free parameter, λ(0)=150nm\lambda(0)=150\>\text{nm} in the pristine sample (blue fitting curve and blue data symbols) and λ(0)=258nm\lambda(0)=258\>\text{nm} after 2.15 C/cm2 electron irradiation (red curve and symbols). Bottom panel: Superfluid density calculated from the data, ρs(T)=(1+Δλ(T)/λ(0))2\rho_{s}(T)=(1+\Delta\lambda(T)/\lambda(0))^{-2}. Solid lines show self-consistent full temperature range calculations using Eilenberger formalism for pristine (blue line) and irradiated (red line) states. The known analytical expression for the ss-wave dirty limit is shown in [38].

The exponential temperature dependence of λ(T)\lambda(T) can be fitted with the well-known low-temperature asymptotic BCS, Δλ(T)=λ(0)πδ2teδt\Delta\lambda(T)=\lambda(0)\sqrt{\frac{\pi\delta}{2t}}e^{-\frac{\delta}{t}} [19], where the ratio δ=Δ(0)/Tc\delta=\Delta(0)/T_{c} was fixed at δ1.76\delta\approx 1.76, leaving only one free parameter λ(0)\lambda(0). The fitting is shown in the top panel of Fig.2. It produces λ(0)=150nm\lambda(0)=150\>\text{nm} in the pristine state (blue fitting curve and blue data symbols) and λ(0)=258nm\lambda(0)=258\>\text{nm} after 2.15 C/cm2 electron irradiation (red curve and symbols). With these numbers, we can calculate the superfluid density in the full temperature range using ρs(T)(λ(0)/λ(T))2=(1+Δλ(T)/λ(0))2\rho_{s}(T)\equiv\left(\lambda(0)/\lambda(T)\right)^{2}=\left(1+\Delta\lambda(T)/\lambda(0)\right)^{-2}. The bottom panel of Fig.2 shows ρs(T)\rho_{s}(T) by blue and red circles for the pristine and irradiated states of the same sample, respectively. The theoretical lines of the clean (blue) and dirty (red) limits were calculated self-consistently using the Eilenberger formalism [39]. We note that the analytical dirty limit formula, ρs=(Δ(T)/Δ(0))tanh(Δ(T)/2T)\rho_{s}=\left(\Delta(T)/\Delta(0)\right)\tanh{\left(\Delta(T)/2T\right)} reproduces the numerical calculation precisely [38]. Examining Fig.2 we conclude that the classical BCS theory describes the experimental data well.

To summarize our findings from measurements of the London penetration depth, λ(T)\lambda(T), several independent parameters: (1) low-temperature behavior of λ(T)\lambda(T); (2) full temperature range behavior of ρs\rho_{s}; (3) disorder-independent TcT_{c} before and after electron irradiation, fully agree with the BCS theory for the isotropic swaves-wave gap with the ratio δ=Δ(0)/Tc1.76\delta=\Delta(0)/T_{c}\approx 1.76. This is the nature of superconductivity in the bulk of AuSn4\text{AuSn}_{4} crystals. However, our measurements would not pick up a tiny signal coming from the surface atomic layers, so unconventional topological features are still possible.

III.2 Campbell penetration depth

Refer to caption
Figure 3: Temperature variation of the measured magnetic penetration depth, λm\lambda_{m}, before (top panel) and after (bottom panel) electron irradiation, measured with the various dc magnetic fields applied along the cc- axis. The field values are shown. Solid lines correspond to zero-field cooling (ZFC), and dotted lines correspond to field cooling (FC) protocols. For one curve, this is shown by arrows. The ZFC and FC curves are indistinguishable, implying that the process is completely reversible, indicating the pinning potential’s parabolic shape. Note that the axes scales are the same in the top and bottom panels, aiding in a visual comparison of the effect of irradiation.

The temperature variation of the magnetic penetration depth before (top panel) and after (bottom panel) electron irradiation, measured in various dc magnetic fields applied along the cc- axis, is shown in Fig.3. The field values are shown next to each curve. Solid lines correspond to zero-field cooling (ZFC) in all curves, and dotted lines correspond to field cooling (FC) protocols. For one curve, this is shown by arrows. The ZFC and FC curves are indistinguishable, implying that the process is totally reversible, which indicates a parabolic shape of the pinning potential.

In the presence of an external DC magnetic field, Abrikosov vortices penetrate the sample and form a vortex lattice. Then the measured penetration depth, λm\lambda_{m}, has two contributions, the usual London penetration depth that in this section we explicitly denote as λL\lambda_{L}, and the Campbell penetration depth λC\lambda_{C}, which is a characteristic length scale over which a small ac perturbation is transmitted elastically by a vortex lattice into the sample [40, 41, 42, 43]. More specifically, the amplitude of the ac perturbation must be small enough so that the vortices remain in their potential well, and their motion is described by the reversible linear elastic response. In this case, λm2=λL2+λC2\lambda_{m}^{2}={\lambda_{L}^{2}+\lambda_{C}^{2}} [44, 45]. This requirement of a very small amplitude makes most conventional ac susceptibility techniques inapplicable for the measurements of the Campbell length. Specialized frequency domain resonators with sufficient sensitivity to a small excitation ac magnetic field are needed [46, 47]. Until now, only a few experimental studies have been published [46, 48, 49, 47, 50].

Figure 3 shows the temperature-dependent variation of the magnetic penetration depth, λm(T)=λL(0)+Δλm(T)\lambda_{m}(T)=\lambda_{L}(0)+\Delta\lambda_{m}(T), for different values of the dc magnetic field applied parallel to the sample cc-axis. For λL(0)\lambda_{L}(0), we have used the values obtained from the BCS fit; see the upper panel of Fig. 2. Then, we assumed that, above TcT_{c}, the resistivity is field independent, so we adjusted other curves to match that value. The top panel shows a pristine state, and the bottom panel shows the same sample after electron irradiation.

Generally speaking, the Campbell penetration depth can exhibit a hysteresis upon warming and cooling, indicating an anharmonic (non-parabolic) pinning potential and/or strong pinning [46, 51, 43, 50]. Therefore, there are two types of measurement protocols: zero-field cooling (ZFC) and field cooling (FC). In the ZFC protocol, the Campbell length is measured on warming after the sample was cooled in a zero magnetic field and the target field was applied at the base temperature (solid lines in Fig. 3). In the FC protocol, measurements are performed on cooling in a target magnetic field applied above TcT_{c} (dotted lines in Fig. 3). For both pristine and irradiated states, λm(T)\lambda_{m}(T) shows a monotonic increase with temperature, and there is no hysteresis between the ZFC and FC protocols. To aid in visualizing the effect of irradiation, the scales of the axes in Fig. 3 are the same in the top and bottom panels. It is clear that the measured penetration depth has increased after electron irradiation.

Refer to caption
Figure 4: Campbell penetration depth, λC2=λm2λL2\lambda_{C}^{2}=\sqrt{\lambda_{m}^{2}-\lambda_{L}^{2}} as a function of an applied magnetic field, HH, evaluated from the data shown in Fig. 3 at a fixed temperature of T=0.5KT=0.5\>\text{K}. for a FC protocol comparing pristine (blue symbols) and irradiated (red symbols) states of the same sample.

Figure 4 shows the Campbell penetration depth as a function of an applied magnetic field, HH, evaluated from the data shown in Fig. 3 at a fixed temperature of T=0.5KT=0.5\>\text{K} for a FC protocol comparing pristine (blue symbols) and irradiated (red symbols) states of the same sample. The Campbell length λC\lambda_{C} increases after irradiation. In the simple Campbell model [40, 41], λC2=ϕ0H/α\lambda_{C}^{2}=\phi_{0}H/\alpha, where ϕ0\phi_{0} is the magnetic flux quantum and α\alpha is the curvature of the pinning potential, α=d2U/dr2\alpha=d^{2}U/dr^{2}. The critical current density jc=αrp/ϕ0=Hrp/λC2j_{c}=\alpha r_{p}/\phi_{0}=Hr_{p}/\lambda_{C}^{2}, where rpr_{p} is the radius of the pinning potential, usually assumed to be the coherence length, ξ\xi. We note that this critical current is not the same as the persistent current in other magnetization measurements because of magnetic relaxation. With an operational frequency of 14 MHz, unattainable in conventional ac susceptometry, our estimate of the critical current density is much closer to the true critical value. The dc magnetization gives an even lower current density as a result of a long time window of data acquisition.

In a more general picture, α\alpha is determined by the elementary pinning forces [52, 42, 43]. In the original model with a fixed rpr_{p}, the Campbell length is expected to scale as λCH\lambda_{C}\sim\sqrt{H}, but Fig.4 shows a practically linear temperature dependence, especially after irradiation. This indicates that vortex pinning in AuSn4\text{AuSn}_{4} is more complicated with a field-dependent radius of the pinning potential, which is possible, for example, in a collective pinning theory when the vortex lattice evolves from the single-vortex pining regime to the vortex bundle regime [53]. In addition, it is known that the coherence length increases with the magnetic field [54]. Therefore, if ξH\xi\sim H, then λC\lambda_{C} will be a linear function of the applied field. As for the difference between pristine and irradiated states, it is possible that the collective pinning in the pristine state is replaced by the disordered vortex phase after electron irradiation, and one cannot directly compare the critical current densities using the same formula. In any case, the nature of pinning in AuSn4\text{AuSn}_{4} requires further investigation.

IV Conclusions

We report measurements of London, λL(T)\lambda_{L}(T), and Campbell, λC(T)\lambda_{C}(T), penetration depths in single crystals of the topological superconductor candidate AuSn4\text{AuSn}_{4} to elucidate the nature of superconductivity in the bulk. Several independent parameters studied before and after 2.5 MeV electron irradiation unambiguously point to isotropic single ss-wave gap weak coupling BCS superconductor. Specifically, the superfluid density before and after electron irradiation overlaps almost perfectly with the parameter-free theoretical BCS curves in the full temperature range for clean and dirty limits, respectively. The Campbell penetration depth before and after electron irradiation does not show hysteresis between the ZFC and FC data, indicating a parabolic shape of the pinning potential. However, the HH-linear behavior of λC\lambda_{C} implies either the field-dependent Labusch parameter, α\alpha, or the radius of the pinning potential, rpr_{p}, or both. Considering the low pinning in AuSn4 single crystals and the point-like nature of the induced defects, such a field dependence may be expected in the vortex bundle regimes within the collective pinning theory [53].

Acknowledgements.
We thank Hermann Suderow for the fruitful discussion. This work was supported by the US DOE, Office of Science, BES Materials Science and Engineering Division under the contract #\# DE-AC02-07CH11358. The authors acknowledge support from the EMIRA French network (FR CNRS 3618) on the SIRIUS platform.

References

  • Nayak et al. [2008] C. Nayak, S. H. Simon, A. Stern, M. Freedman, and S. D. Sarma, Non-abelian anyons and topological quantum computation, Reviews of Modern Physics 80, 1083 (2008).
  • Ran et al. [2019] S. Ran, C. Eckberg, Q.-P. Ding, Y. Furukawa, T. Metz, S. R. Saha, I.-L. Liu, M. Zic, H. Kim, J. Paglione, et al., Nearly ferromagnetic spin-triplet superconductivity, Science 365, 684 (2019).
  • Maeno et al. [2011] Y. Maeno, S. Kittaka, T. Nomura, S. Yonezawa, and K. Ishida, Evaluation of spin-triplet superconductivity in Sr2RuO4Journal of the Physical Society of Japan 81, 011009 (2011).
  • Kallin and Berlinsky [2009] C. Kallin and A. J. Berlinsky, Is Sr2RuO4 a chiral p-wave superconductor?, J. Physics: Cond. Mat. 21, 164210 (2009).
  • Kallin [2012] C. Kallin, Chiral p-wave order in Sr2RuO4Rep. Progr. Phys. 75, 042501 (2012).
  • Joynt and Taillefer [2002] R. Joynt and L. Taillefer, The superconducting phases of UPt3Reviews of Modern Physics 74, 235 (2002).
  • Yuqiang et al. [2019] F. Yuqiang, J. Pan, D. Zhang, D. Wang, H. Hirose, T. Terashima, S. Uji, Y. Yuan, W. Li, Z. Tian, J. Xue, Y. Ma, W. Zhao, Q. Xue, G. Mu, H. Zhang, and F. Huang, Discovery of Superconductivity in 2M WS2 with Possible Topological Surface States, Adv. Mater. , 1901942 (2019).
  • Fu and Berg [2010] L. Fu and E. Berg, Odd-Parity Topological Superconductors: Theory and Application to CuxBi2Se3Phys. Rev. Lett. 105, 097001 (2010).
  • Hor et al. [2010] Y. S. Hor, A. J. Williams, J. G. Checkelsky, P. Roushan, J. Seo, Q. Xu, H. W. Zandbergen, A. Yazdani, N. P. Ong, and R. J. Cava, Superconductivity in CuxBi2Se3 and its Implications for Pairing in the Undoped Topological Insulator, Phys. Rev. Lett. 104, 057001 (2010).
  • Zhu et al. [2023] W. Zhu, R. Song, J. Huang, Q.-W. Wang, Y. Cao, R. Zhai, Q. Bian, Z. Shao, H. Jing, L. Zhu, et al., Intrinsic surface p-wave superconductivity in layered AuSn4Nature Communications 14, 7012 (2023).
  • Shen et al. [2020] D. Shen, C. N. Kuo, T. W. Yang, I. N. Chen, C. S. Lue, and L. M. Wang, Two-dimensional superconductivity and magnetotransport from topological surface states in AuSn4 semimetal, Communications Materials 1, 56 (2020).
  • Karn et al. [2022] N. Karn, M. Sharma, and V. Awana, Non-trivial band topology in the superconductor AuSn4: a first principle study, Superconductor Science and Technology 35, 114002 (2022).
  • Herrera et al. [2023] E. Herrera, B. Wu, E. O’Leary, A. M. Ruiz, M. Águeda, P. G. Talavera, V. Barrena, J. Azpeitia, C. Munuera, M. García-Hernández, et al., Band structure, superconductivity, and polytypism in AuSn4Physical Review Materials 7, 024804 (2023).
  • Gendron and Jones [1962] M. Gendron and R. Jones, Superconductivity in the CuAl2 (C16) crystal class, J. Phys. Chem. Sol. 23, 405 (1962).
  • Wu et al. [2016] Y. Wu, L.-L. Wang, E. Mun, D. D. Johnson, D. Mou, L. Huang, Y. Lee, S. L. Bud’ko, P. C. Canfield, and A. Kaminski, Dirac node arcs in PtSn4Nature Physics 12, 667 (2016).
  • Jo et al. [2017] N. H. Jo, Y. Wu, L.-L. Wang, P. P. Orth, S. S. Downing, S. Manni, D. Mou, D. D. Johnson, A. Kaminski, S. L. Bud’ko, et al., Extremely large magnetoresistance and Kohler’s rule in PdSn4: A complete study of thermodynamic, transport, and band-structure properties, Physical Review B 96, 165145 (2017).
  • Sharma et al. [2023] M. Sharma, G. Gurjar, S. Patnaik, and V. Awana, Two-fold anisotropic superconducting state in topological superconductor Sn4Au, Europhysics Letters 142, 26004 (2023).
  • Bardeen et al. [1957a] J. Bardeen, L. N. Cooper, and J. R. Schrieffer, Microscopic theory of superconductivity, Phys. Rev. 106, 162 (1957a).
  • Bardeen et al. [1957b] J. Bardeen, L. N. Cooper, and J. R. Schrieffer, Theory of Superconductivity, Phys. Rev. 108, 1175 (1957b).
  • Okamoto [1993] H. Okamoto, Au-Sn (gold-tin), Journal of phase equilibria 14, 765 (1993).
  • Canfield [2019] P. C. Canfield, New materials physics, Reports on Progress in Physics 83, 016501 (2019).
  • Van Degrift [1975] C. T. Van Degrift, Tunnel diode oscillator for 0.001 ppm measurements at low temperatures, Review of Scientific Instruments 46, 599 (1975).
  • Prozorov et al. [2000a] R. Prozorov, R. W. Giannetta, A. Carrington, and F. M. Araujo-Moreira, Meissner-london state in superconductors of rectangular cross section in a perpendicular magnetic field, Phys. Rev. B 62, 115 (2000a).
  • Prozorov et al. [2000b] R. Prozorov, R. W. Giannetta, A. Carrington, P. Fournier, R. L. Greene, P. Guptasarma, D. G. Hinks, and A. R. Banks, Measurements of the absolute value of the penetration depth in high-T-c superconductors using a low-T-c superconductive coating, Appl. Phys. Lett. 77, 4202 (2000b).
  • Prozorov [2021] R. Prozorov, Meissner-london susceptibility of superconducting right circular cylinders in an axial magnetic field, Phys. Rev. App. 16, 024014 (2021).
  • Giannetta et al. [2022] R. Giannetta, A. Carrington, and R. Prozorov, London Penetration Depth Measurements Using Tunnel Diode Resonators, J. Low Temp. Phys. 208, 119 (2022).
  • Damask and Dienes [1963] A. C. Damask and G. J. Dienes, Point Defects in Metals (Gordon & Breach Science Publishers Ltd, 1963).
  • Thompson [1969] M. W. Thompson, Defects and Radiation Damage in Metals, revised september 27, 1974 ed., Cambridge Monographs on Physics (Cambridge University Press, 1969).
  • Prozorov et al. [2014] R. Prozorov, M. Kończykowski, M. A. Tanatar, A. Thaler, S. L. Bud’ko, P. C. Canfield, V. Mishra, and P. J. Hirschfeld, Effect of Electron Irradiation on Superconductivity in Single Crystals of Ba(Fe1xRux)2As2\mathrm{Ba}({\mathrm{Fe}}_{1-x}{\mathrm{Ru}}_{x}{)}_{2}{\mathrm{As}}_{2} (x = 0.24), Phys. Rev. X 4, 041032 (2014).
  • Prozorov et al. [2019] R. Prozorov, M. Kończykowski, M. A. Tanatar, H.-H. Wen, R. M. Fernandes, and P. C. Canfield, Interplay between superconductivity and itinerant magnetism in underdoped Ba1-xKxFe2As2 (x= 0.2) probed by the response to controlled point-like disorder, npj Quantum Materials 4, 34 (2019).
  • Tinkham [2004] M. Tinkham, Introduction to Superconductivity, 2nd ed. (Dover Publications, 2004).
  • Kogan et al. [2013] V. Kogan, R. Prozorov, and V. Mishra, London penetration depth and pair breaking, Phys. Rev. B 88, 224508 (2013).
  • Gordon et al. [2009] R. T. Gordon, N. Ni, C. Martin, M. A. Tanatar, M. D. Vannette, H. Kim, G. D. Samolyuk, J. Schmalian, S. Nandi, A. Kreyssig, A. I. Goldman, J. Q. Yan, S. L. Bud’ko, P. C. Canfield, and R. Prozorov, Unconventional London Penetration Depth in Single-Crystal Ba(Fe0.93Co0.07)2As2\mathrm{Ba}({\mathrm{Fe}}_{0.93}{\mathrm{Co}}_{0.07}{)}_{2}{\mathrm{As}}_{2} Superconductors, Phys. Rev. Lett. 102, 127004 (2009).
  • Kim et al. [2011] H. Kim, M. A. Tanatar, Y. J. Song, Y. S. Kwon, and R. Prozorov, Nodeless two-gap superconducting state in single crystals of the stoichiometric iron pnictide LiFeAs, Phys. Rev. B 83, 100502 (2011).
  • Timmons et al. [2020] E. I. Timmons, S. Teknowijoyo, M. Kończykowski, O. Cavani, M. A. Tanatar, S. Ghimire, K. Cho, Y. Lee, L. Ke, N. H. Jo, S. L. Bud’ko, P. C. Canfield, P. P. Orth, M. S. Scheurer, and R. Prozorov, Electron irradiation effects on superconductivity in PdTe2: An application of a generalized Anderson theorem, Phys. Rev. Res. 2, 023140 (2020).
  • Anderson [1959] P. W. Anderson, Theory of dirty superconductors, Journal of Physics and Chemistry of Solids 11, 26 (1959).
  • Ghimire et al. [2024a] S. Ghimire, K. R. Joshi, E. H. Krenkel, M. A. Tanatar, Y. Shi, M. Kończykowski, R. Grasset, V. Taufour, P. P. Orth, M. S. Scheurer, and R. Prozorov, Electron irradiation reveals robust fully gapped superconductivity in LaNiGa2{\mathrm{LaNiGa}}_{2}Phys. Rev. B 109, 024515 (2024a).
  • Kogan [2013] V. G. Kogan, Homes scaling and BCS, Phys. Rev. B 87, 220507 (2013).
  • Eilenberger [1968] G. Eilenberger, Transformation of Gorkov’s equation for type II superconductors into transport-like equations, Zeitschrift für Physik A Hadrons and nuclei 214, 195 (1968).
  • Campbell [1969] A. M. Campbell, The response of pinned flux vortices to low-frequency fields, Journal of Physics C: Solid State Physics 2, 1492 (1969).
  • Campbell [1971] A. M. Campbell, The interaction distance between flux lines and pinning centres, Journal of Physics C: Solid State Physics 4, 3186 (1971).
  • Willa et al. [2015a] R. Willa, V. B. Geshkenbein, and G. Blatter, Campbell penetration in the critical state of type-II superconductors, Phys. Rev. B 92, 134501 (2015a).
  • Gaggioli et al. [2022] F. Gaggioli, G. Blatter, and V. B. Geshkenbein, Creep effects on the Campbell response in type-II superconductors, Phys. Rev. Res. 4, 013143 (2022).
  • Brandt [1991] E. H. Brandt, Penetration of magnetic ac fields into type-II superconductors, Phys. Rev. Lett. 67, 2219 (1991).
  • Koshelev and Vinokur [1991] A. E. Koshelev and V. M. Vinokur, Frequency response of pinned vortex lattice, Physica C 173, 465 (1991).
  • Prozorov et al. [2003] R. Prozorov, R. W. Giannetta, N. Kameda, T. Tamegai, J. A. Schlueter, and P. Fournier, Campbell penetration depth of a superconductor in the critical state, Phys. Rev. B 67, 184501 (2003).
  • Ghigo et al. [2023] G. Ghigo, D. Torsello, L. Gozzelino, M. Fracasso, M. Bartoli, C. Pira, D. Ford, G. Marconato, M. Fretto, I. De Carlo, N. Pompeo, and E. Silva, Vortex dynamics in NbTi films at high frequency and high DC magnetic fields, Sci. Rep. 13, 1 (2023).
  • Kim et al. [2021] H. Kim, M. A. Tanatar, H. Hodovanets, K. Wang, J. Paglione, and R. Prozorov, Campbell penetration depth in low carrier density superconductor YPtBi, Phys. Rev. B 104, 014510 (2021).
  • Srpcic et al. [2021] J. Srpcic, M. Ainslie, Y. Shi, and J. Durrell, The Campbell penetration depth in type-II superconductors, in 7th International Workshop on Numerical Modelling of High Temperature Superconductors (HTS 2020) (2021).
  • Ghimire et al. [2024b] S. Ghimire, F. Gaggioli, K. R. Joshi, M. Konczykowski, R. Grasset, E. H. Krenkel, A. Datta, M. A. Tanatar, S. Chen, C. Petrovic, V. B. Geshkenbein, and R. Prozorov, Nonmonotonic relaxation of Campbell penetration depth from creep-enhanced vortex pinning, arXiv:2403.14891 10.48550/arXiv.2403.14891 (2024b).
  • Gordon et al. [2013] R. Gordon, N. Zhigadlo, S. Weyeneth, S. Katrych, and R. Prozorov, Conventional superconductivity and hysteretic Campbell penetration depth in single crystals MgCNi3Phys. Rev. B 87, 94520 (2013).
  • Willa et al. [2015b] R. Willa, V. B. Geshkenbein, R. Prozorov, and G. Blatter, Campbell response in type-II superconductors under strong pinning conditions, Phys. Rev. Lett. 115, 207001 (2015b).
  • Blatter et al. [1994] G. Blatter, M. V. Feigel’man, V. B. Geshkenbein, A. I. Larkin, and V. M. Vinokur, Vortices in high-temperature superconductors, Rev. Mod. Phys. 66, 1125 (1994).
  • Prozorov et al. [2007] R. Prozorov, V. G. Kogan, M. D. Vannette, S. L. Bud’ko, and P. C. Canfield, Radio-frequency magnetic response of vortex lattices undergoing structural transformations in superconducting borocarbide crystals, Phys. Rev. B 76, 094520 (2007).