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

thanks: cThese two authors contributed equally

Superconductivity and Quantum Oscillations
in Single Crystals of the Compensated Semimetal CaSb2 

M. Oudah1,2,c [email protected]    J. Bannies1,3,c    D. A. Bonn1,2    M. C. Aronson1,2 1Stewart Blusson Quantum Matter Institute, University of British Columbia, Vancouver, British Columbia V6T 1Z4, Canada
2Department of Physics, University of British Columbia, Vancouver, Canada V6T 1Z1, Canada
3Department of Chemistry, University of British Columbia, Vancouver, Canada V6T 1Z1, Canada
Abstract

Bulk superconductivity in a topological semimetal is a first step towards realizing topological superconductors, which can host Majorana fermions allowing us to achieve quantum computing. Here, we report superconductivity and compensation of electrons and holes in single crystals of the nodal-line semimetal CaSb2. We characterize the superconducting state and find that Cooper pairs have moderate-weak coupling, and the superconducting transition in specific heat down to 0.22 K deviates from that of a BCS superconductor. The non-saturating magnetoresistance and electron-hole compensation at low temperature are consistent with density functional theory (DFT) calculations showing nodal-line features. Furthermore, we observe de Haas–van Alphen (dHvA) oscillations consistent with a small Fermi surface in the semimetallic state of CaSb2. Our DFT calculations show that the two electron bands crossing the Fermi level are associated with Sb1 zig-zag chains, while the hole band is associated with Sb2 zig-zag chains. The Sb1 zig-zag chains form a distorted square net, which may relate the MMSb2 family to the well known MMSbTe square net semimetals. Realization of superconductivity and a compensated semimetal state in single crystals of CaSb2 establishes the diantimonide family as a candidate class of materials for achieving topological superconductivity.

pacs:
Valid PACS appear here
preprint: APS/123-QED

I Introduction

A central theme in the field of topological materials has been the search for topological superconductors, which are expected to host Majorana fermions and have potential applications in achieving fault-tolerant quantum computing Beenakker (2013); Alicea (2012). One avenue for finding topological superconductivity relies on making a heterostructure of a superconductor and a topologically non-trivial material Fu and Kane (2008), in the hopes of inducing superconductivity via the proximity effect. In parallel, the quest to find bulk topological superconductors continues Fu and Berg (2010), and candidate materials include CuxBi2Se3 Sasaki et al. (2011), PdBi2 Matthias (1963); Imai et al. (2012), LaPt3Takayama et al. (2012), and UTe2 Ran et al. (2019). The search for superconductivity in topologically non-trivial materials includes work on topological insulators Hasan and Kane (2010) and topological semimetals Yan and Felser (2017). In topological insulators, superconductivity can be achieved by doping the material, but this approach has the disadvantage that doping can push the topological states away from the Fermi level Kondo et al. (2013).

In topological semimetals on the other hand, superconductivity may be realized without doping due to the finite density of states at the Fermi level EFE_{\rm{F}}. In addition, bulk bands form nodal points or nodal lines near EFE_{\rm{F}} in momentum space (k space) Burkov et al. (2011); Potter et al. (2014); Fang et al. (2016); Armitage et al. (2018). The nodal lines in topological semimetals can be gapped with inclusion of spin-orbit-coupling (SOC) Fang et al. (2015). However, non-symmorphic symmetry can protect the nodal lines, keeping them gapless even in the presence of SOC Zhao and Schnyder (2016). With this in mind, finding bulk superconductivity in topological semimetals with non-symmorphic space group symmetry is a promising path towards realizing topological superconductivity.

Refer to caption
Figure 1: (a) The X-ray diffraction pattern measured at room temperature of a CaSb2 single crystal (top, red), powder (middle, blue), and the simulated powder pattern (bottom, black). (b) Laue pattern of a CaSb2 single crystal. (c) Picture of a CaSb2 single crystal on mm-size grid. (d) Crystal structure of CaSb2 in the bcbc-plane showing Ca (blue) sites and the Sb zig-zag chains along the bb-direction associated with the Sb1 (red) and Sb2 (orange) sites. (e) Crystal structure of CaSb2 in the abab-plane showing zig-zag chains associated with the Sb1 site form a distorted square-net, where each of the distorted squares has two short (solid bonds) and two long (dashed bonds).

A number of materials with a square net of atoms in their crystal structure and non-symmorphic symmetry have been identified as topological semimetals in recent years, including SrMnSb2 Liu et al. (2017), ZrSiS Schoop et al. (2016) LaSbTe Singha et al. (2017); Wang et al. (2021), and GdSbTe Hosen et al. (2018); Lei et al. (2019, 2021). In these materials, a square net is formed such that the periodicity of the crystal structure is doubled compared to this square net itself. The doubling of the unit cell compared to the squares leads to folding of bands derived from pxp_{x} and pyp_{y} states of the square net atoms, and leads to a semimetallic state Klemenz et al. (2019), where the square net atom can be Si or Sb. The square net RERESbTe (RERE= rare earth) and the MMSb2 (MM= rare earth, alkaline earth) groups of materials are closely related, where the Sb square-net is distorted in the latter. We find the MMSb2 interesting, in particular CaSb2, where the topologically non-trivial character of the bands persists even when the Sb square-nets are distorted. Recently, it was theoretically predicted that CaSb2 is a nodal-line semimetal, and that this state is protected by the non-symmorphic symmetry of the crystal structure Funada et al. (2019). Experimental reports of a metal-to-insulator-like transition Funada et al. (2019) and superconductivity Ikeda et al. (2020) in CaSb2 have drawn interest to this material. Those experimental reports have led us to pursue the growth of single crystals, a critical step to making many of the measurements needed to explore the intrinsic properties of CaSb2.

In this paper, we confirm the existence of bulk superconductivity in high quality single crystals of CaSb2 and take several steps in identifying the nature of this superconducting state. Measurements of the phase diagram in magnetic field give us the fundamental superconducting length scales in the material: the London penetration depth and the superconducting coherence length. Low temperature specific heat measurements suggest unconventional aspects of the superconducting pairing state. Furthermore, we present band structure calculations demonstrating that states associated with the distorted Sb square-net are dominant near the Fermi level and thus play a central role in the superconducting state and electrical properties. We observe de Haas–van Alphen (dHvA) oscillations consistent with our calculated Fermi surfaces. In electrical resistivity measurements under magnetic fields, we observe a rise in resistivity at low temperature, consistent with previous reports. The magnetoresistance at low temperature reaches high values without saturation and the concentration of hole and electron carriers is almost equivalent below 50 K. All of these results show that superconductivity in CaSb2 occurs in the context of a well compensated semimetal.

II Methods

Single crystals of CaSb2 were grown from Sb self flux. Pieces of Ca (Alfa Aesar, 99.98 %) and Sb (Chempur, 99.999 %) with a molar ratio of 1:3.75 were loaded into an Al2O3 crucible and sealed in a quartz tube with a pressure of 0.3 atm argon. The mixture was slowly heated to 700 C and held at this temperature for 12 h. Subsequently, the mixture was cooled to 670 C at a rate of 10 K/h, followed by slow cooling to 592 C with 1 K/h. At this temperature, the crystals were separated from the flux by centrifuging at 2000-3000 rpm for about 15 s. Shiny, plate-like crystals with dimensions up to 3 x 3 x 0.5 mm3 were obtained by this method. Crystals are stable in air for several weeks. Phase purity and orientation of the crystals were checked by X-ray diffraction (XRD) using a Bruker D8 with Cu Kα1\alpha_{1} radiation (1.54056 Å). The composition and homogeneity were confirmed by means of energy dispersive X-ray spectroscopy (EDX) using a Philips XL30 scanning electron microscope equipped with a Bruker Quantax 200 energy-dispersion X-ray microanalysis system and an XFlash 6010 SDD detector.

Electrical resistivity measurements were performed with conventional four- and five-probe geometries using a Quantum Design Physical Property Measurement System (PPMS) equipped with a 3He/4He dilution refrigerator insert. For these measurements Pt wires (25 μ\mum) were attached to the sample with silver epoxy. Measurements of the magnetic susceptibility were done using a Magnetic Property Measurements System 3 (MPMS3) also from Quantum Design, equipped with a 3He insert. Specific heat measurements were performed using a PPMS (Quantum Design).

Electronic structure calculations were performed within the framework of density functional theory (DFT) as implemented in the package Wien2k Blaha et al. (2001). The generalized gradient approximation with the PBE parametrization Perdew et al. (1996) was used. The basis set size was set to RmtKmax=8.5 and the irreducible Brillouin zone (BZ) was sampled with 2445 k points. The previously reported monoclinic crystal structure (P21/mP2_{1}/m, S.G.:11, a = 4.746 Å, b = 4.177 Å, c = 9.084 Å, β=106.3\beta=106.3^{\circ}) was used Deller and Eisenmann (1976).

III Results and Discussion

The CaSb2 single crystals used in this study were characterized using X-ray diffraction (XRD) and energy dispersive X-ray spectroscopy (EDX) to confirm the quality of the samples. The powder XRD pattern measured on our sample, shown in Fig. 1a in blue, is consistent with the previously reported monoclinic crystal structure (P21/mP2_{1}/m, S.G.:11) Deller and Eisenmann (1976), but we refined the lattice parameters of our samples to a = 4.7404 Å, b = 4.1796 Å, c = 9.0712 Å, and β\beta = 106.3 °. For comparison, we show in black in Fig. 1a the calculated diffraction pattern based on our refined structure. The Laue measurement in Fig. 1(b) shows sharp peaks confirming the high quality of the CaSb2 samples. EDX mapping shows a homogeneous distribution of Ca and Sb on the sample surface (Sup. Fig. 1 in in Supplemental Materials Sup ). The cc-axis diffraction pattern is shown in red in Fig. 1a and confirms that the direction perpendicular to the plate-like crystals (Fig. 1c) corresponds to the cc axis of the crystal structure (shown in Fig. 1d and e). The structure contains two zig-zag chains of antimony corresponding to two distinct crystallographic sites, Sb1 and Sb2 shown in red and orange, respectively, in Fig. 1d. The Ca atoms occupy one site, which sits between these two chains of Sb, as shown in blue in Fig. 1d. The zig-zag chains associated with the Sb1 site form a distorted square net (Fig. 1e), where squares become quadrilaterals with two short and two long distances of 2.915 Å (solid bonds) and 3.459 Å (dashed bonds), respectively. The zig-zag chains associated with Sb2 site have Sb-Sb distances of 2.939 Å and the chains are separated by 4.740 Å. The non-symmorphic space group contains a 21 screw axis along the bb direction, which is related to the symmetry of the Sb chains in the distorted square net.

Refer to caption
Figure 2: (a) Temperature dependence of the dc magnetic susceptibility measured in a 1 mT field applied parallel to the abab-plane (along plane of crystal) with zero-field-cooling (ZFC) and field-cooling (FC) procedures, presented in terms of the shielding fraction. (b) Temperature dependence of the dc susceptibility measured with different applied fields using a zero-field-cooling procedure. (c) The magnetization (MM) as a function of applied field measured at different temperatures below TcT_{\text{c}}. A degaussing procedure was carried out between measurements, and a linear fit was applied to the low field region of the 0.4 K data, as discussed in the text. (d) Temperature dependence of longitudinal resistivity ρxx\rho_{\mathrm{xx}} below 2 K measured in different magnetic fields, where the data are normalized by the 2 K data (ρxx,2K1μΩcm\rho_{\mathrm{xx,2~{}\rm{K}}}\sim 1~{}\mu\Omega\rm{cm}) measured in 0 mT. The superconducting transition is suppressed with increasing magnetic fields. (e) The TcT_{\text{c}} values extracted from transitions in χV\chi_{V} measured in different fields. Solid and dashed lines are WHH fits to the data used to estimate μ0Hc2(0)\mu_{\rm{0}}H_{\rm{c2}}(0). (f) Lower critical field μ0Hc1(0)\mu_{\rm{0}}H_{\rm{c1}}(0) estimated using the magnetization data in (c) by subtracting the linear fit to 0.4 K data from all the curves, shown in inset. The cutoff is chosen as estimate of μ0Hc1(0)\mu_{\rm{0}}H_{\rm{c1}}(0) at each temperature.

III.1 Superconductivity

We characterize the superconducting state of CaSb2 and show bulk superconductivity with sharp superconducting transitions, using magnetic susceptibility, specific heat, and resistivity measurements. The magnetic susceptibility at low temperature is shown in Fig. 2(a). The DC magnetization was measured between 0.4 K-1.9 K using zero-field-cooling (ZFC) processes and field-cooling processes (FC). The volume susceptibility (χV\chi_{V}) vs. temperature measured in a field of 1 mT is shown in Fig. 2(a), where χV\chi_{V} was corrected for demagnetization effects due to sample shape using the expression for a slab-like geometry in Ref. Brandt (1999). Both ZFC and FC measurements show evidence of bulk superconductivity with a strong diamagnetic signal that has an onset of 1.6 K in a field of 1 mT. The superconducting transition width ΔT\Delta T, defined as the temperature range over which the diamagnetic signal reaches 90% of its maximum value below TcT_{\text{c}}, is \sim0.25 K. The estimated superconducting volume fraction after demagnetization correction in the ZFC data is 100%. We note that demagnetization calculated using a rectangular cuboid in Ref. Prozorov and Kogan (2018) resulted in a volume fraction exceeding 100%. Thus, our data suggests near-perfect shielding in the superconducting state. The difference between the ZFC and FC values of χV\chi_{V} in our single crystal samples is suggestive of a type-II superconductor with flux pinning. However, the significant flux expulsion in the FC data suggests that pinning of flux vortices is weak.

In order to map out the field-temperature phase diagram, we track the shift in TcT_{\text{c}} for different applied magnetic fields in DC magnetization and specific heat measurements in Figs. 2(b) and  3(a), respectively. We measured a drop in electrical resistivity at the superconducting transition that is suppressed with applied field, as shown in Fig. 2(d), but due to noise in the data estimating TcT_{\text{c}} at each applied field is difficult. Also, we measured the AC susceptibility of the CaSb2 and the real (χ\chiAC) and imaginary (χ\chiAC) parts are shown in Sup. Fig. 2 in Supplemental Materials Sup . Although we measured a sharp superconducting transition with a peak in χ\chiAC at 1.67 K, we find this measurement is sensitive to the onset of superconductivity rather than the bulk TcT_{\text{c}}. From magnetization, TcT_{\text{c}} is defined as the temperature where the diamagnetic signal reaches 5% volume fraction at different magnetic fields. From specific heat, TcT_{\text{c}} is defined as the mid point of the transition seen in CelC_{\rm{el}}, inset of Fig. 3(a). The magnetic fields plotted against TcT_{\text{c}} as extracted from both measurements are shown in Fig. 2(e). To estimate μ0Hc2(0)\mu_{\rm{0}}H_{\rm{c2}}(0), we use the Werthamer-Helfand-Hohenberg (WHH) equation Werthamer et al. (1966). We get μ0Hc2(0)\mu_{\rm{0}}H_{\rm{c2}}(0) of 21.5 mT from specific heat and 24.4 mT from magnetization. Using the μ0Hc2(0)\mu_{\rm{0}}H_{\rm{c2}}(0) of 24.4 mT, we estimate the coherence length ξGL\xi_{\rm{GL}} using the equation

μ0Hc2(0)=Φ02πξGL2\mu_{0}H_{\text{c2}}(0)=\frac{\Phi_{0}}{2\pi\xi^{2}_{\text{GL}}}\ (1)

where Φ0\Phi_{0} is the quantum flux h/2eh/2e. We get ξGL\xi_{\rm{GL}} of 116 nm.

To measure μ0Hc1(0)\mu_{\rm{0}}H_{\rm{c1}}(0) we measured the field-dependent magnetization (MM) at different temperatures below the critical temperature, as shown in Fig. 2(c). The low-field linear fit for the 0.4 K measurement is shown with a black line from the origin (MfitM_{\rm{fit}}). This fit was subtracted from all the measured curves to construct a MM-MfitM_{\rm{fit}} plot. The field where the magnetization deviated from linear response, taken as the black line in the inset of Fig. 2(f), is the uncorrected μ0Hc1(0)\mu_{\rm{0}}H^{\rm{*}}_{\rm{c1}}(0) for that temperature. All μ0Hc1(0)\mu_{\rm{0}}H^{\rm{*}}_{\rm{c1}}(0) are plotted against the corresponding temperature in Fig. 2(f) and are fitted to the equation

μ0Hcl(T)=μ0Hcl(0)[1(TTc)2]\mu_{0}H^{*}_{cl}(T)=\mu_{0}H^{*}_{cl}(0)\left[1-\left(\frac{T}{T_{c}}\right)^{2}\right]\ (2)

where μ0Hc1(0)\mu_{\rm{0}}H^{\rm{*}}_{\rm{c1}}(0) is the lower critical field at 0 K and TcT_{\text{c}} is the superconducting critical temperature. μ0Hc1(0)\mu_{\rm{0}}H^{\rm{*}}_{\rm{c1}}(0) was calculated to be 3.1 mT. After correcting for the demagnetization factor, μ0Hc1(0)\mu_{\rm{0}}H_{\rm{c1}}(0) was calculated to be 4.9 mT.

Refer to caption
Figure 3: (a) Temperature dependence of CPC_{P} measured in varying magnetic fields. The superconducting transition seen at 1.52 K in 0 mT is suppressed with increasing magnetic field, emphasized in inset showing CelC_{\rm{el}}/TT. (b) Superconducting transition in the electronic part of the specific heat CelC_{\rm{el}}/TT as a function of TT in 0 T. Two fits are presented for a TcT_{\text{c}} of 1.52 K: dashed black line is for weak-coupling BCS behavior, while solid red line is a theoretical curve for dd-wave pairing on a simple cylindrical Fermi surface with line nodes Won and Maki (1994).

We estimate the penetration depth λ\lambda using ξGL=116.1\xi_{\rm{GL}}=116.1 nm and μ0Hc1(0)\mu_{\rm{0}}H_{\rm{c1}}(0) =4.9=4.9 mT with the equation

μ0Hcl=Φ04πλ2(lnλξ+12)\mu_{0}H_{\text{cl}}=\frac{\Phi_{0}}{4\pi\lambda^{2}}\left(\text{ln}\frac{\lambda}{\xi}+\frac{1}{2}\right)\ (3)

we get λ=89.5\lambda=89.5 nm for CaSb2. We calculate the Ginzburg-Landau constant κGL=λ/ξGL=0.771±0.038\kappa_{\rm{GL}}=\lambda/\xi_{\rm{GL}}=0.771\pm 0.038. Based on our calculated values, we classify CaSb2 as a type-II superconductor since κGL>12\kappa_{\rm{GL}}>\frac{1}{\sqrt{2}}. However, it is just barely type-II, so this property will be sensitive to defects and carrier density in this semi-metallic material.

In Fig. 3(b), the transition seen in CelC_{\rm{el}} with a TcT_{\text{c}} of 1.52 K in zero field and a ΔT\Delta T of 0.15 K provides further evidence for the bulk nature of the superconductivity. Magnetic fields push superconducting transitions seen in specific heat CPC_{P} to lower temperatures (Fig. 3(a)). The phonon contribution to CPC_{P} was determined by fitting the data measured in a field of 50 mT, where no transition is observed, between 0.55 K-4.0 K to the expression

Cp=γnT+βT3C_{p}=\gamma_{n}T+\beta T^{3}\ (4)

The result of this fit is γn=1.778\gamma_{n}=1.778 mJ mol-1 K-2 and β=0.471\beta=0.471 mJ mol-1 K-4. The phonon contribution represented by the cubic term βT3\beta T^{3} was subtracted from the total CPC_{P} to evaluate the electronic part of the specific heat CelC_{\rm{el}}. The electronic contribution to specific heat CelC_{\rm{el}} divided by temperature is plotted as a function of temperature in Fig. 3(b). The flat part of the graph above 1.55 K corresponds to a constant γ\gamma above the transition temperature consistent with metallic behavior. At TcT_{\text{c}} a jump is observed corresponding to a second-order phase transition into the superconducting state, and the narrow transition further confirms the high-quality of the crystals. We observed the superconducting transition in χV\chi_{V} and CelC_{\rm{el}} measurements on multiple crystals, grown in different batches, where the TcT_{\text{c}} as defined above differs by less than 1%.

The Debye model is used to calculate the Deybe temperature ΘD\Theta_{D}

ΘD=(12π45βnR)1/3\Theta_{D}=\left(\frac{12\pi^{4}}{5\beta}nR\right)^{1/3}\ (5)

where n=3n=3 and RR is the gas constant 8.314 J mol-1 K-1. The Debye temperature was calculated to be ΘD=231K\Theta_{D}=231~{}K, and with ΘD\Theta_{D} and TcT_{\text{c}}, we calculate the superconducting parameter λep\lambda_{ep} using the inverted McMillan McMillan (1968) equation

λep=1.04+μln(ΘD1.45Tc)(10.62μ)ln(ΘD1.45Tc)1.04\lambda_{\text{ep}}=\frac{1.04+\mu^{*}ln\left(\frac{\Theta_{D}}{1.45T_{c}}\right)}{(1-0.62\mu^{*})\text{ln}\left(\frac{\Theta_{D}}{1.45T_{c}}\right)-1.04}\ (6)

where μ=0.10\mu\rm{*}=0.10 and TcT_{\text{c}} =1.60=1.60 K. We find λep=0.458\lambda_{ep}=0.458 for CaSb2, suggesting it is a weak-moderate coupling superconductor. Using λep\lambda_{ep}, γ\gamma, and the Boltzmann constant kBk_{\rm{B}}, we calculate the electronic density of states at the Fermi energy N(EF)N(E_{\rm{F}}) with the equation

N(EF)=3γπ2kB2(1+λep)N(E_{F})=\frac{3\gamma}{\pi^{2}k_{B}^{2}(1+\lambda_{\text{ep}})}\ (7)

N(EF)N(E_{\rm{F}}) was estimated to be  0.52 states eV-1 per formula unit (fu) of CaSb2.

We calculate the mean free path (ll) using the following equation

l=2.372×1014(mme)2VM2N(EF)2ρl=2.372\times 10^{-14}\frac{\left(\frac{m^{*}}{m_{e}}\right)^{2}V_{M}^{2}}{N(E_{F})^{2}\rho}\ (8)

where N(EF)=0.52N(E_{\rm{F}})=0.52, the molar volume VM=52.0cm3/molV_{\rm{M}}=52.0~{}\rm{cm}^{3}/\rm{mol}, ρ=1μΩcm\rho=1~{}\mu\Omega\rm{cm} (Fig. 4(a)), and using m/me=0.165m^{*}/m_{e}=0.165, based on dHvA oscillation analysis in Sec. III.2, we get l=65.3nml=65.3~{}\rm{nm}. The resulting value of ξGL/l\xi_{\rm{GL}}/l is 1.78, places CaSb2 at the border between the clean limit and dirty limit.

The magnitude of the specific heat jump ΔC/Cen0.95\Delta C/C_{en}\sim 0.95, CenC_{en} being the electronic specific heat in the normal state, is about 34% smaller than the value of 1.43 expected for a BCS model in the weak-coupling limit. In Fig. 3(b) we show the BCS expectation of a jump at TcT_{\text{c}} and the exponential decay of CelC_{\rm{el}} below TcT_{\text{c}} with a black dotted line, and it is clear that the data in CaSb2 are not well explained with the BCS model. As we have near-perfect shielding from the Meissner signal in the magnetization measurements for multiple samples, we exclude the possibility of sample inhomogeneity as the origin of this smaller jump in the specific heat. We attempt a fit to the data based on the alpha model Padamsee et al. (1973); Johnston (2013), and we find that experimental data deviates from this model for α=1.425&1.50\alpha=1.425~{}\&~{}1.50, as shown in Sup. Fig. 3. More importantly, the decay in CelC_{\rm{el}} /T/T with decreasing temperature is weaker than that expected for a BCS superconductor. Even at the lowest temperature of 0.22 K the value of CelC_{\rm{el}} is not negligible and continues to be temperature-dependent. This unusual behaviour in CelC_{\rm{el}} suggests the presence of a small gap component, due perhaps to extreme gap anisotropy, differing gaps on different bands, or gaps with nodes.

Refer to caption
Figure 4: Electrical transport data and analysis by two-band model. (a) Temperature dependence of the longitudinal resistivity ρxx\rho_{\mathrm{xx}} for different magnetic fields with Ic\mathrm{I}\perp c in sample 1. (b) Magnetic field dependence of the resistance (MR) with Hc\mathrm{H}\parallel c in sample 1. (c) MR and Hall resistivity with Hc\mathrm{H}\parallel c in sample 2, respectively. (d-f) Charge carrier densities, their ratio, and carrier mobilities extracted from fits of MR and ρxy\rho_{\mathrm{xy}} of sample 2 to the two-band model Pippard (1989), respectively.
Table 1: Superconducting parameters derived from our measurements of CaSb2.
Parameter Units Value
TcT_{\text{c}} K 1.60
μ0Hc1(0)\mu_{\rm{0}}H_{\rm{c1}}(0) mT 4.9±0.44.9\pm 0.4
μ0Hc2(0)\mu_{\rm{0}}H_{\rm{c2}}(0) mT 24.4±0.724.4\pm 0.7
ξGL\xi_{\rm{GL}} nm 116
λ\lambda nm 89.5
κGL\kappa_{\rm{GL}} - 0.771±0.0380.771\pm 0.038
γn\gamma_{n} mJ mol-1 K-2 1.778
β\beta mJ mol-1 K-4 0.471
θD\theta_{D} K 231.3
λep\lambda_{ep} - 0.458
N(EF)N(E_{\rm{F}}) states eV-1 per fu 0.52

We consider a model at the extreme end of possible gap anisotropy, one that includes line nodes in the superconducting gap Won and Maki (1994), shown in red in Fig. 3(b). In this model, the average gap magnitude is smaller than that in the fully gapped case, and this is reflected in a reduced specific heat jump. Also, at low temperature the TT dependence of CelC_{\rm{el}} /T/T changes from exponential to power law Maeno et al. (2011), because the existence of nodes allows for the thermal excitation of quasiparticles down to very low temperatures. This nodal-gap function was recently suggested for a Bi2Se3 related material, namely Cu-intercalated (PbSe)5(Bi2Se3)6 Sasaki et al. (2014), considered to potentially be topological superconductors. In CaSb2 this model fits better to the superconducting jump at TcT_{\text{c}}, but the low temperature experimental data has lower values than expected in this model. The data clearly do not fit the BCS values or the simplest line-node model, which suggests strong anisotropy in the gap function. This may be related to one of the Fermi surfaces in our calculation being dominated by Sb chains in the distorted square-net (Sb1 site), discussed in Sec. III.2. We note the recent report of 121/123Sb-nuclear quadrupole resonance (NQR) measurement in a polycrystalline sample of CaSb2, which suggests ss-wave superconductivity, in part based on observing a coherence peak below TcT_{\text{c}} Takahashi et al. (2021). The suggestion of exponentially-activated behaviour in those NQR data at low temperatures is not clear, as is also the case for our specific heat data. A coherence peak was observed in nuclear magnetic resonance (NMR) measurements of the unconventional superconductors CePt3Si\rm{CePt}_{3}\rm{Si} Yogi et al. (2006a, b) and Li2(Pd1xPtx)3B\rm{Li}_{2}(\rm{Pd}_{1-x}\rm{Pt}_{x})_{3}\rm{B} Harada et al. (2010); Shimamura et al. (2007), which is composition dependent in the latter. Ultimately the resolution of the gap structure in CaSb2 will require lower temperature measurements on single crystal samples, and a variety of spectroscopic probes.

CaSb2 seems to fall at the border of type-I and type-II superconductivity. YbSb2, which crystalizes in a higher symmetry crystal structure (CmcmCmcm, S.G.:63), has been reported to be a type-I superconductor Zhao et al. (2012). The different crystal structures of YbSb2 and CaSb2 may be important to the nature of the emergent superconducting state in these compounds. Studying the effect of pinning centers on flux-trapping in CaSb2 and further improving the quality of grown crystals will help us clarify whether it is intrinsically a type-I or type-II superconductor, as well as further probing the presence of gap anisotropy. This will be the subject of future studies. A summary of observed superconducting parameters of CaSb2 is presented in Table 1. We firmly established and characterised bulk superconductivity in CaSb2. In the next section we place this superconductivity in the context of an unusual semimetal by examining the consistency of the calculated band structure with our transport measurements and de Haas–van Alphen (dHvA) oscillations.

Refer to caption
Figure 5: Electronic structure of CaSb2 from density functional theory calculations. (a) Band structure with bands crossing EFE_{\mathrm{F}} colored, spin-orbit coupling (SOC) not included. (b) Brillouin zone of CaSb2. (c) Three-dimensional visualization of the Fermi surface with SOC formed by one hole-like and two electron-like pockets. The color coding corresponds to that in panel (a).

III.2 Semimetallic Properties and Band Structure

The resistivity measurements reveal metallic behavior with a longitudinal resistivity ρxx\rho_{\mathrm{xx}} of ca. 1 μΩ\mu\Omega\cdotcm at 2 K (Fig. 4a). The high residual resistivity ratio ρ300K/ρ2K\rho_{\mathrm{300~{}\rm{K}}}/\rho_{\mathrm{2~{}\rm{K}}} of ca. 100 indicates the good quality of the crystals. At low temperature, the resistivity rises rapidly with increasing field. While the zero field resistivity has the temperature dependence of a good metal, in higher fields the resistivity begins to increase with decreasing temperature as in an insulator. The rise in resistivity with decreasing temperature in an applied field larger than 3T3~{}\rm{T} plateaus for temperatures below 10K\sim 10~{}\rm{K}. This behavior in our measurements on single crystals of CaSb2 is similar with those measurements reported on polycrystalline samples Fu and Kane (2008), as well as other topological semimetals such as WP2 Wang et al. (2017), NbAs2 Wang et al. (2016), and ZrP2 Bannies et al. (2021). As shown in Fig. 4b, the magnetoresistance MR, which is defined as MR=ρxx(B)/ρxx(B=0)1MR=\rho_{\mathrm{xx}}(B)/\rho_{\mathrm{xx}}(B=0)-1, remains unsaturated up to 9 T, where it reaches large values of 2500% at 2 K. The field dependence of the MR can be described by a power law MR=a+bBcMR=\mathrm{a}+\mathrm{b}\cdot B^{\mathrm{c}} with c=1.79\mathrm{c}=1.79. This behavior is reproduced in a second sample (see Fig. 4c inset). The Shubnikov-de Haas oscillations superimposed on the MR background (Fig. 4b) further demonstrate good sample quality but are not analyzed here due to the low number of oscillations observed. Instead we will present de Haas-van Alphen (dHvA) oscillations in detail below. Together, the field-driven metal-to-insulator-like transition and the large and unsaturated MR suggest a semimetallic state in CaSb2.

To verify this scenario, we measured the Hall resistivity ρxy\rho_{\mathrm{xy}} in the same field and temperature ranges as the MR (Fig. 4c). While ρxy(B)\rho_{\mathrm{xy}}(B) is roughly linear for all temperatures, it cannot be fit with a semi-classical one-band model because ρxx\rho_{\mathrm{xx}} is independent of the field in this model, which is contrast to our data (see Fig. 4b). Instead, the data can be fitted with the semi-classical two-band model Pippard (1989):

ρxx=1e(nhμh+neμe)+(nhμe+neμh)μhμeB2(nhμh+neμe)2+(nhne)2μh2μe2B2\rho_{\mathrm{xx}}=\frac{1}{e}\frac{(n_{\mathrm{h}}\mu_{\mathrm{h}}+n_{\mathrm{e}}\mu_{\mathrm{e}})+(n_{\mathrm{h}}\mu_{\mathrm{e}}+n_{\mathrm{e}}\mu_{\mathrm{h}})\mu_{\mathrm{h}}\mu_{\mathrm{e}}B^{2}}{(n_{\mathrm{h}}\mu_{\mathrm{h}}+n_{\mathrm{e}}\mu_{\mathrm{e}})^{2}+(n_{\mathrm{h}}-n_{\mathrm{e}})^{2}\mu_{\mathrm{h}}^{2}\mu_{\mathrm{e}}^{2}B^{2}} (9)

and

ρxy=Benhμh2neμe2+(nhne)μh2μe2B2(nhμh+neμe)2+(nhne)2μh2μe2B2\rho_{\mathrm{xy}}=\frac{B}{e}\frac{n_{\mathrm{h}}\mu_{\mathrm{h}}^{2}-n_{\mathrm{e}}\mu_{\mathrm{e}}^{2}+(n_{\mathrm{h}}-n_{\mathrm{e}})\mu_{\mathrm{h}}^{2}\mu_{\mathrm{e}}^{2}B^{2}}{(n_{\mathrm{h}}\mu_{\mathrm{h}}+n_{\mathrm{e}}\mu_{\mathrm{e}})^{2}+(n_{\mathrm{h}}-n_{\mathrm{e}})^{2}\mu_{\mathrm{h}}^{2}\mu_{\mathrm{e}}^{2}B^{2}} (10)

based on the carrier densities ne,hn_{\mathrm{e,h}} and mobilities μe,h\mu_{\mathrm{e,h}} for electrons and holes. By simultaneously fitting ρxx\rho_{\mathrm{xx}} and ρxy\rho_{\mathrm{xy}} to the two-band model (see Sup. Fig. 4 for an exemplary fit), we extracted the carrier densities and mobilities of the charge carriers, which are shown as functions of temperature in Fig. 4d-f (the full dataset is shown in Sup. Fig. 5 in Supplemental Materials Sup ). While electrons are the majority charge carriers at high temperatures, the carrier densities are essentially compensated below 50 K with ne/nh=0.99(1)n_{\mathrm{e}}/n_{\mathrm{h}}=0.99(1) at 2 K. The carrier mobilities increase monotonically with decreasing temperature, presumably due to reduced scattering of charge carriers by phonons. The high values of approximately 3000 cm2/Vs at 2 K further indicate good crystal quality. Based on the two-band model fits, we conclude that CaSb2 is a compensated semimetal in which large MR arises at low temperature from the combination of charge carrier compensation and high carrier mobility.

Refer to caption
Figure 6: Quantum oscillations and fat band representations based on density functional theory calculations. (a) Oscillatory component ΔM\Delta M corresponding to the de Haas-van Alphen (dHvA) oscillations at different temperatures. (b) Fast Fourier Transform spectrum of the dHvA oscillations at 1.8 K. (c) Fit of eq. 11 to the dHvA oscillations at a representative temperature of 1.8 K, using three fundamental frequencies. (d, e) Fat band representations of the band structure of CaSb2. The size of symbols represents the projected weight of the Sb1 and Sb2 atomic wave functions, respectively.

We confirmed our experimental findings by band structure calculations based on density functional theory (DFT). The resulting band structure of CaSb2 is shown in Fig. 5(a). The Fermi surface is composed of one hole-like pocket centered in the ky=0k_{\mathrm{y}}=0 plane and two electron-like pockets at the edge of the Brillouin zone (BZ), see Fig. 5c. While the hole pocket exhibits a complex three-dimensional shape, the two electron-like bands have weak dispersion along kzk_{\mathrm{z}}, giving rise to a quasi-two-dimensional character. The presence of two electron pockets in the calculations seemingly contradicts our analysis of the magnetotransport. However, the volume of the smaller electron pocket is only 1/8 of the volume of the bigger electron pocket. We argue that the two electrons pockets can be effectively described as one combined electron pocket if they have similar average carrier mobilities. This assumption is validated by the good quality of the two-band model fits to the experimental data. Attempts to fit the data with a three-band model containing two electron pockets and one hole pocket did not improve the quality of the fit and were therefore discarded.

Table 2: Extracted parameters from the fit of the de Haas-van Alphen oscillations with eq. 11. Extremal areas AA were calculated based on frequencies FF by use of the Onsager relation Shoenberg (2009). Effective masses mm^{*} and Dingle temperatures TDT_{\mathrm{D}} were used to calculate quantum relaxation times τq=/(2πkBTD)\tau_{\mathrm{q}}=\hbar/(2\pi k_{\mathrm{B}}T_{\mathrm{D}}) and quantum mobilities μq=eτq/m\mu_{\mathrm{q}}=e\tau_{\mathrm{q}}/m^{*}.
FF [T] AA [10-3Å-2] mm^{*} [mem_{e}] TDT_{\mathrm{D}} [K] τq\tau_{\mathrm{q}} [10-13 s] μq\mu_{\mathrm{q}} [m2/Vs]
71.5(1) 6.829(9) 0.168(1) 6.5(2) 1.88(7) 0.197(9)
77.5(1) 7.398(7) 0.160(1) 6.3(2) 1.94(7) 0.213(9)
115.8(1) 11.050(3) 0.166(1) 5.2(1) 2.32(4) 0.247(6)

To complement our findings about the Fermi surface based on electrical transport experiments, we have studied the magnetization in the normal state at low temperatures. At 1.8 K, we observe de Haas-van Alphen (dHvA) oscillations in fields larger than 2.5 T for BcB\parallel c. In Fig. 6(a), we show the oscillatory component of the dHvA oscillations as a function of inverse magnetic field at selected temperatures between 1.8 K and 10 K. The fast Fourier transform (FFT) spectrum of the data at 1.8 K, shown in Fig. 6b, reveals three peaks at Fα=(70±4)F_{\alpha}=(70\pm 4) T, Fβ=(78±4)F_{\beta}=(78\pm 4) T and Fγ=(116±4)F_{\gamma}=(116\pm 4) T. Given the limited number of periods collected, and the presence of closely-overlapping frequencies, it is advantageous to fit the dHvA oscillations to a standard model for a Fermi liquid in order to avoid limitations of the FFT method Ramshaw et al. (2011). In this procedure, we fitted the oscillatory component ΔM\Delta M of the dHvA oscillations as a function of magnetic field and temperature to the Lifshitz-Kosevich formula Shoenberg (2009):

ΔMB1/2RTRDRSsin(2π(FB12+ϕ))\Delta M\sim-B^{1/2}R_{\mathrm{T}}R_{\mathrm{D}}R_{\mathrm{S}}\sin{\left(2\pi\left(\frac{F}{B}-\frac{1}{2}+\phi\right)\right)} (11)

with RT=XTsinh(XT)R_{\mathrm{T}}=\frac{XT}{\sinh(XT)}, RD=exp(XTD)R_{\mathrm{D}}=\exp{(-XT_{\mathrm{D}})}, and RS=cos(πgm2m0)R_{\mathrm{S}}=\cos{\left(\pi g\frac{m^{*}}{2m_{0}}\right)}, where X=2π2kBm/(eB)X=2\pi^{2}k_{\mathrm{B}}m^{*}/(\hbar eB). TDT_{\mathrm{D}} is the Dingle temperature and ϕ\phi is a phase factor that depends on the dimensionality of the Fermi surface and the Berry phase among other factors.

Accordingly, the temperature and magnetic field dependent damping of the oscillations are described by RTR_{\mathrm{T}} and RDR_{\mathrm{D}}. Fitting ΔM(B,T)\Delta M(B,T) to eq. 11 therefore gives access to the effective masses and Dingle temperatures. Furthermore, it allows us to refine the initially imprecise frequencies from the FFT. In Fig. 6c, we show a representative part of the fit at a constant temperature of 1.8 K, which demonstrates the good quality of the fit. The data extracted from the fit at all temperatures are summarized in Tab. 2.

The three fundamental frequencies are 71.5(1) T, 77.5(1) T, and 115.8(1) T. Using the Onsager relation AF=2πeFA_{\mathrm{F}}=\frac{2\pi e}{\hbar}F Shoenberg (2009), we find that the extremal orbits associated with these frequencies cover 0.357(5)%, 0.387(4)%, and 0.578(2)% of the BZ projected along [001], respectively. The small values suggest small volumes of these Fermi surface pockets, in line with the semimetallic character. All three orbits show similar low effective masses of ~ 0.16 mem_{e} and the quantum mobilities range between ~ 2000 and 2400 cm2/Vs, indicative of high sample quality. Altogether, the small extremal areas, low effective masses and high quantum mobilities are consistent with the semimetallic character implied by the transport measurements.

Next, we compare the extremal orbits from dHvA oscillations with those predicted by DFT calculations. We do not find a quantitative match between theory and experiment. The mismatch can be caused by two possibly concomitant reasons: (i) the as-grown single crystals are self-doped, resulting in an experimental Fermi energy different from the theoretical one or (ii) the functional used in DFT calculations does not describe the semimetallicity accurately, resulting in a wrong Fermi surface. Without studies of the angle-dependence of the dHvA oscillations, it is impossible to deduce the reason for the mismatch. Future studies should therefore aim at obtaining a 3D Fermi surface from quantum oscillations. Importantly, our experimental results confirm the semimetallicity independent of the DFT calculations and thus remain valid despite the mismatch with the DFT calculations.

The bands around the Fermi level are dominated by contributions from Sb states, as shown by the fat band representations of Sb1 and Sb2 atoms in Fig. 6(d) and (e), respectively. The two electron bands crossing the Fermi level along the Γ\Gamma-Z line close to the Z point have strong Sb1 character, whereas the hole band centered around the Γ\Gamma point is dominated by Sb2 character. The Ca contribution to these bands is very small as shown in Sup. Fig. 6 in Supplemental Materials Sup . The electron bands crossing the Fermi level along C-Z form nodal lines in the ky=πk_{y}=\pi plane in the presence of SOC (see Sup. Fig. 7 in Supplemental Materials Sup ), in agreement with previous calculations Funada et al. (2019). These electron bands have a strong Sb1 character near the Fermi level and significant Sb2 character at higher energy; the Sb1 site forms the distorted square-net in CaSb2. The similarities in the crystal structure between the MMSb2 distorted square net family and the MMSbTe square net family may result in similarities in the band structure of these materials. We highlight that the periodicity of atoms compared with the periodicity of the unit cell in a particular direction results in non-symmorphic symmetry, which is important for the nodal-line semimetal state in both cases, rather than the perfect square-net of Sb in MMSbTe. The connection between the nodal-line semimetal states in both classes of materials warrants further theoretical and experimental investigation. The combined evidence for semimetallic character from electrical transport and dHvA oscillations support our calculated band structure, which includes non-trivial band topology along the C-Z direction at the ky=πk_{y}=\pi plane.

Interestingly, CaSb2 is one of only a handful of materials that displays at ambient pressure both a large MR (i.e. semimetallic character) and superconductivity, see Tab. 3. These materials include MoTe2 Qi et al. (2016), TaSe3 Sambongi et al. (1977); Saleheen et al. (2020), and α\alpha-Ga Chen et al. (2018). Topological states were experimentally observed by angle-resolved photoemission spectroscopy (ARPES) in MoTe2 Deng et al. (2016) and TaSe3 Lin et al. (2021). Within superconducting semimetals, CaSb2 offers a good balance between a moderately high TcT_{\text{c}} of 1.6 K and large MR of 25 at 2 K and 9 T, similar to TaSe3. In contrast, the low TcT_{\text{c}} of MoTe2 is difficult to reach, posing a significant disadvantage for applications. Among the materials listed in Tab. 3, α\alpha-Ga has superior characteristics both in terms of TcT_{\text{c}} and MR, but Ga is difficult to handle due to its low melting point of 29.8 Sostman (1977). The combination of TcT_{\text{c}}, MR, and stability of CaSb2 make it an excellent candidate for exploring the interplay between superconductivity and semimetallicity.

Table 3: Materials that exhibit large magnetoresistance (MR) and superconductivity at ambient pressure, their superconducting transition temperatures TcT_{\text{c}}, and MR values at 9 T.
Material TcT_{\text{c}} [K] MR Ref.
MoTe2 0.1 40\sim 40 at 1.4 K, 9 T Qi et al. (2016)
α\alpha-Ga 0.9 1.6\sim 1.6\cdot104 at 2 K, 9 T Chen et al. (2018)
TaSe3 2.1\sim 2.1 30\sim 30 at 1.9 K, 9 T Sambongi et al. (1977); Saleheen et al. (2020)
CaSb2 1.6 25 at 1.8 K, 9 T This work

In theoretical work, it was suggested that the nodal-line structure of the normal state Fermi surface of CaSb2 could survive in the superconducting state when weak-coupling is present Ono et al. (2021). This allows for realizing topological superconductivity in CaSb2. We find experimentally that superconductivity in single crystals of CaSb2 is in the weak-moderate coupling limit with λep=0.456\lambda_{ep}=0.456 and the decay in the electronic part of the specific heat CelC_{\rm{el}} in the superconducting state down to 0.22 K is slower and does not follow the exponential decay expected for a conventional BCS superconductor. Future experiments will help us clarify the extent of influence of topology in the normal state of CaSb2 on the superconducting properties.

IV Conclusions

In this Article, which is the first report on single crystals of CaSb2, we report evidence for bulk superconductivity in CaSb2 with an onset of about 1.6 K. The CPC_{P} jump at TcT_{\text{c}} is inconsistent with that expected in a BCS superconductor, and the decay in the CelC_{\rm{el}} at low temperature hints at anisotropy of the superconducting gap function. Our transport measurements support a well-compensated semimetal picture for CaSb2, where electrons and holes are almost perfectly compensated below 20 K. The high quality of the single crystals allows us to measure de Haas-van Alphen (dHvA) oscillations and we identify three frequencies associated with a small Fermi surface. Measuring the intrinsic superconducting and compensated semimetal properties of CaSb2 was only possible due to availability of single crystals. Our calculations reveal that bands at the Fermi level are dominated by contributions from the distorted Sb square-net. The exact nature of the topologically non-trivial band crossing can be elucidated in future experiments, including ARPES. The family of square-net materials and diantimonides with distorted square-nets host a rich variety of properties, and the present work will serve as an important step towards connecting these two sub-fields. Our results promote CaSb2, and the family of MMSb2, for further investigation of topological properties and superconductivity.

Note added. After submitting this manuscript, we noticed a report on the properties of polycrystalline CaSb2 showing a peak in TcT_{\text{c}} under pressure, and suggesting an unconventional nature of the superconductivity Kitagawa et al. (2021).

V Acknowledgements

We thank Mengxing (Ketty) Na for help with the Laue measurement, Jacob Kabel for help with the EDX measurement, Silvia Luescher-Folk for help with low temperature electrical transport measurement, and Ilya Elfimov for discussion on DFT calculations. MO acknowledges the support by Stewart Blusson Quantum Matter Institute and the Max Planck-UBC-UTokyo Center for Quantum Materials. JB, DAB, and MCA acknowledge the support by the Natural Sciences and Engineering Research Council of Canada (NSERC).

References

  • Beenakker (2013) C. Beenakker, Annu. Rev. Condens. Matter Phys. 4, 113 (2013).
  • Alicea (2012) J. Alicea, Rep. Prog. Phys. 75, 076501 (2012).
  • Fu and Kane (2008) L. Fu and C. L. Kane, Phys. Rev. Lett. 100, 096407 (2008).
  • Fu and Berg (2010) L. Fu and E. Berg, Phys. Rev. Lett. 105, 097001 (2010).
  • Sasaki et al. (2011) S. Sasaki, M. Kriener, K. Segawa, K. Yada, Y. Tanaka, M. Sato, and Y. Ando, Phys. Rev. Lett. 107, 217001 (2011).
  • Matthias (1963) B. Matthias, Rev. Mod. Phys 35, 1 (1963).
  • Imai et al. (2012) Y. Imai, F. Nabeshima, T. Yoshinaka, K. Miyatani, R. Kondo, S. Komiya, I. Tsukada, and A. Maeda, JPSJ 81, 113708 (2012).
  • Takayama et al. (2012) T. Takayama, K. Kuwano, D. Hirai, Y. Katsura, A. Yamamoto, and H. Takagi, Phys. Rev. Lett. 108, 237001 (2012).
  • 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., Science 365, 684 (2019).
  • Hasan and Kane (2010) M. Z. Hasan and C. L. Kane, Rev. Mod. Phys. 82, 3045 (2010).
  • Yan and Felser (2017) B. Yan and C. Felser, Annu. Rev. Condens. Matter Phys. 8, 337 (2017).
  • Kondo et al. (2013) T. Kondo, Y. Nakashima, Y. Ota, Y. Ishida, W. Malaeb, K. Okazaki, S. Shin, M. Kriener, S. Sasaki, K. Segawa, et al., Phys. Rev. Lett. 110, 217601 (2013).
  • Burkov et al. (2011) A. Burkov, M. Hook, and L. Balents, Phys. Rev. B 84, 235126 (2011).
  • Potter et al. (2014) A. C. Potter, I. Kimchi, and A. Vishwanath, Nat. Commun. 5, 1 (2014).
  • Fang et al. (2016) C. Fang, H. Weng, X. Dai, and Z. Fang, Chin. Phys. B 25, 117106 (2016).
  • Armitage et al. (2018) N. Armitage, E. Mele, and A. Vishwanath, Rev. Mod. Phys. 90, 015001 (2018).
  • Fang et al. (2015) C. Fang, Y. Chen, H.-Y. Kee, and L. Fu, Phys. Rev. B 92, 081201 (2015).
  • Zhao and Schnyder (2016) Y. Zhao and A. P. Schnyder, Phys. Rev. B 94, 195109 (2016).
  • Liu et al. (2017) J. Liu, J. Hu, Q. Zhang, D. Graf, H. B. Cao, S. Radmanesh, D. Adams, Y. Zhu, G. Cheng, X. Liu, et al., Nat. Mater. 16, 905 (2017).
  • Schoop et al. (2016) L. M. Schoop, M. N. Ali, C. Straßer, A. Topp, A. Varykhalov, D. Marchenko, V. Duppel, S. S. Parkin, B. V. Lotsch, and C. R. Ast, Nat. Commun. 7, 1 (2016).
  • Singha et al. (2017) R. Singha, A. Pariari, B. Satpati, and P. Mandal, Phys. Rev. B 96, 245138 (2017).
  • Wang et al. (2021) Y. Wang, Y. Qian, M. Yang, H. Chen, C. Li, Z. Tan, Y. Cai, W. Zhao, S. Gao, Y. Feng, et al., Phys. Rev. B 103, 125131 (2021).
  • Hosen et al. (2018) M. M. Hosen, G. Dhakal, K. Dimitri, P. Maldonado, A. Aperis, F. Kabir, C. Sims, P. Riseborough, P. M. Oppeneer, D. Kaczorowski, et al., Sci. Rep. 8, 1 (2018).
  • Lei et al. (2019) S. Lei, V. Duppel, J. M. Lippmann, J. Nuss, B. V. Lotsch, and L. M. Schoop, Adv. Quantum Technol. 2, 1900045 (2019).
  • Lei et al. (2021) S. Lei, A. Saltzman, and L. M. Schoop, Phys. Rev. B 103, 134418 (2021).
  • Klemenz et al. (2019) S. Klemenz, S. Lei, and L. M. Schoop, Annu. Rev. Mater. Res. 49, 185 (2019).
  • Funada et al. (2019) K. Funada, A. Yamakage, N. Yamashina, and H. Kageyama, J. Phys. Soc. Japan 88, 044711 (2019).
  • Ikeda et al. (2020) A. Ikeda, M. Kawaguchi, S. Koibuchi, T. Hashimoto, T. Kawakami, S. Yonezawa, M. Sato, and Y. Maeno, Phys. Rev. Mater. 4, 041801 (2020).
  • Blaha et al. (2001) P. Blaha, K. Schwarz, G. K. Madsen, D. Kvasnicka, and J. Luitz, An augmented plane wave+ local orbitals program for calculating crystal properties (2001).
  • Perdew et al. (1996) J. P. Perdew, K. Burke, and Y. Wang, Phys. Rev. B 54, 16533 (1996).
  • Deller and Eisenmann (1976) K. Deller and B. Eisenmann, Zeitschrift für anorganische und allgemeine Chemie 425, 104 (1976).
  • (32) SeeSupplementalMaterialforSEM/EDXresults,ACsusceptibilitymeasurement,transportresultsforCaSb2S2andDFTbandstructurecalculations\rm{See~{}Supplemental~{}Material~{}for~{}SEM/EDX~{}results,~{}AC}\\ \rm{susceptibility~{}measurement,~{}transport~{}results~{}for~{}CaSb}_{2}~{}\rm{S2}\\ \rm{~{}and~{}DFT~{}band~{}structure~{}calculations}.
  • Brandt (1999) E. H. Brandt, Phys. Rev. B 60, 11939 (1999).
  • Prozorov and Kogan (2018) R. Prozorov and V. G. Kogan, Phys. Rev. Appl. 10, 014030 (2018).
  • Werthamer et al. (1966) N. Werthamer, E. Helfand, and P. Hohenberg, Phys. Rev. 147, 295 (1966).
  • Won and Maki (1994) H. Won and K. Maki, Phys. Rev. B 49, 1397 (1994).
  • McMillan (1968) W. McMillan, Phys. Rev. 167, 331 (1968).
  • Padamsee et al. (1973) H. Padamsee, J. Neighbor, and C. Shiffman, J. Low Temp. Phys. 12, 387 (1973).
  • Johnston (2013) D. C. Johnston, Supercond. Sci. Technol. 26, 115011 (2013).
  • Pippard (1989) A. B. Pippard, Magnetoresistance in metals, vol. 2 (Cambridge university press, 1989).
  • Maeno et al. (2011) Y. Maeno, S. Kittaka, T. Nomura, S. Yonezawa, and K. Ishida, J. Phys. Soc. Jpn. 81, 011009 (2011).
  • Sasaki et al. (2014) S. Sasaki, K. Segawa, and Y. Ando, Phys. Rev. B 90, 220504 (2014).
  • Takahashi et al. (2021) H. Takahashi, S. Kitagawa, K. Ishida, M. Kawaguchi, A. Ikeda, S. Yonezawa, and Y. Maeno, J. Phys. Soc. Jpn. 90, 073702 (2021).
  • Yogi et al. (2006a) M. Yogi, H. Mukuda, Y. Kitaoka, S. Hashimoto, T. Yasuda, R. Settai, T. D. Matsuda, Y. Haga, Y. Ōnuki, P. Rogl, et al., J. Phys. Soc. Jpn. 75, 013709 (2006a).
  • Yogi et al. (2006b) M. Yogi, Y. Kitaoka, S. Hashimoto, T. Yasuda, R. Settai, T. Matsuda, Y. Haga, Y. Ōnuki, P. Rogl, and E. Bauer, J. Phys. Chem. Solids 67, 522 (2006b).
  • Harada et al. (2010) S. Harada, Y. Inada, and G.-Q. Zheng, Physica C Supercond 470, 1089 (2010).
  • Shimamura et al. (2007) A. Shimamura, Y. Furukawa, K. Kumagai, H. Takeya, and K. Hirata, Physica C Supercond 460, 663 (2007).
  • Zhao et al. (2012) L. L. Zhao, S. Lausberg, H. Kim, M. A. Tanatar, M. Brando, R. Prozorov, and E. Morosan, Phys. Rev. B 85, 214526 (2012).
  • Wang et al. (2017) A. Wang, D. Graf, Y. Liu, Q. Du, J. Zheng, H. Lei, C. Petrovic, et al., Phys. Rev. B 96, 121107 (2017).
  • Wang et al. (2016) Y.-Y. Wang, Q.-H. Yu, P.-J. Guo, K. Liu, and T.-L. Xia, Phys. Rev. B 94, 041103 (2016).
  • Bannies et al. (2021) J. Bannies, E. Razzoli, M. Michiardi, H.-H. Kung, I. Elfimov, M. Yao, A. Fedorov, J. Fink, C. Jozwiak, A. Bostwick, et al., Phys. Rev. B 103, 155144 (2021).
  • Shoenberg (2009) D. Shoenberg, Magnetic oscillations in metals (Cambridge university press, 2009).
  • Ramshaw et al. (2011) B. Ramshaw, B. Vignolle, J. Day, R. Liang, W. Hardy, C. Proust, and D. Bonn, Nat. Phys. 7, 234 (2011).
  • Qi et al. (2016) Y. Qi, P. G. Naumov, M. N. Ali, C. R. Rajamathi, W. Schnelle, O. Barkalov, M. Hanfland, S.-C. Wu, C. Shekhar, Y. Sun, et al., Nat. Commun. 7, 1 (2016).
  • Sambongi et al. (1977) T. Sambongi, M. Yamamoto, K. Tsutsumi, Y. Shiozaki, K. Yamaya, and Y. Abe, J. Phys. Soc. Japan 42, 1421 (1977).
  • Saleheen et al. (2020) A. I. U. Saleheen, R. Chapai, L. Xing, R. Nepal, D. Gong, X. Gui, W. Xie, D. P. Young, E. Plummer, and R. Jin, npj Quantum Materials 5, 1 (2020).
  • Chen et al. (2018) B. Chen, X. Duan, H. Wang, J. Du, Y. Zhou, C. Xu, Y. Zhang, L. Zhang, M. Wei, Z. Xia, et al., npj Quantum Mater. 3, 1 (2018).
  • Deng et al. (2016) K. Deng, G. Wan, P. Deng, K. Zhang, S. Ding, E. Wang, M. Yan, H. Huang, H. Zhang, Z. Xu, et al., Nat. Phys. 12, 1105 (2016).
  • Lin et al. (2021) C. Lin, M. Ochi, R. Noguchi, K. Kuroda, M. Sakoda, A. Nomura, M. Tsubota, P. Zhang, C. Bareille, K. Kurokawa, et al., Nat. Mater. pp. 1–7 (2021).
  • Sostman (1977) H. E. Sostman, Rev. Sci. Instrum. 48, 127 (1977).
  • Ono et al. (2021) S. Ono, H. C. Po, and K. Shiozaki, Phys. Rev. Res. 3, 023086 (2021).
  • Kitagawa et al. (2021) S. Kitagawa, K. Ishida, A. Ikeda, M. Kawaguchi, S. Yonezawa, and Y. Maeno, Phys. Rev. B 104, L060504 (2021).