Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Anisotropic Paramagnetic Peak Effect in Reversible Magnetization of Crystalline Miassite Superconductor Rh17S15subscriptRh17subscriptS15\text{Rh}_{17}\text{S}_{15}Rh start_POSTSUBSCRIPT 17 end_POSTSUBSCRIPT S start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT

Ruslan Prozorov prozorov@ameslab.gov 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    Alexei E. Koshelev Department of Physics and Astronomy, University of Notre Dame, Notre Dame, Indiana 46656, USA    Linlin Wang Ames National Laboratory, Ames, Iowa 50011, USA Department of Physics & Astronomy, Iowa State University, Ames, Iowa 50011, USA    Sergey L. Bud’ko Ames National Laboratory, Ames, Iowa 50011, USA Department of Physics & Astronomy, Iowa State University, Ames, Iowa 50011, USA    Paul C. Canfield Ames National Laboratory, Ames, Iowa 50011, USA Department of Physics & Astronomy, Iowa State University, Ames, Iowa 50011, USA
(1 June 2024)
Abstract

We report an unusual anisotropic paramagnetic peak effect observed in reversible magnetization of a single crystalline nodal superconductor Rh17S15subscriptRh17subscriptS15\text{Rh}_{17}\text{S}_{15}Rh start_POSTSUBSCRIPT 17 end_POSTSUBSCRIPT S start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT. Both temperature- and field-dependent magnetization measurements reveal a distinct novel vortex state above approximately 1 T. This peak effect is most pronounced when the magnetic field, H𝐻Hitalic_H, is applied parallel to the [111]delimited-[]111\left[111\right][ 111 ] direction, whereas it diminishes for H[110]conditional𝐻delimited-[]110H\parallel\left[110\right]italic_H ∥ [ 110 ]. Intriguingly, for H[100]conditional𝐻delimited-[]100H\parallel\left[100\right]italic_H ∥ [ 100 ], instead of a peak, we observe a step-like decrease in M(T)𝑀𝑇M(T)italic_M ( italic_T ), with the step amplitude increasing in larger applied magnetic fields. This behavior is opposite to the expectations of conventional Meissner expulsion. The magnitude of the peak effect, expressed in terms of dimensionless volume susceptibility, is on the order of Δχ=105Δ𝜒superscript105\Delta\chi=10^{-5}roman_Δ italic_χ = 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT (with full diamagnetic screening corresponding to χ=1𝜒1\chi=-1italic_χ = - 1). The observed anisotropic paramagnetic vortex response is unusual considering the cubic symmetry of Rh17S15subscriptRh17subscriptS15\text{Rh}_{17}\text{S}_{15}Rh start_POSTSUBSCRIPT 17 end_POSTSUBSCRIPT S start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT. We propose that in this distinct vortex phase, a small but finite attractive interaction between vortices below Hc2subscript𝐻𝑐2H_{c2}italic_H start_POSTSUBSCRIPT italic_c 2 end_POSTSUBSCRIPT may be responsible for this unusual phenomenon. Furthermore, the vortices seem to prefer aligning along the [111]delimited-[]111\left[111\right][ 111 ] direction, rotating toward it when the magnetic field is applied in other directions. Our findings add another item to the list of unusual properties of Rh17S15subscriptRh17subscriptS15\text{Rh}_{17}\text{S}_{15}Rh start_POSTSUBSCRIPT 17 end_POSTSUBSCRIPT S start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT that attracted recent attention as the first unconventional superconductor that has a mineral analog, miassite, found in nature.

I Introduction

Since the division of superconductors into two types by Abrikosov in 1957 [1], vortex physics has become one of the most active research areas in superconductivity due to its rich fundamental physics aspects and direct relevance for technological applications [2, 3, 4, 5, 6, 7, 8, 9, 10]. Various vortex lattice phases and behaviors, usually mapped on the vortex HT𝐻𝑇H-Titalic_H - italic_T phase diagram, are believed to be linked to the unconventional nature of superconductivity in different classes of superconductors, such as highTcsubscript𝑇𝑐-T_{c}- italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT cuprates [6, 8, 9, 11], borocarbides [12, 13] and iron-pnictides [14, 15, 16, 17, 11]. One of the most interesting mixed-state features is the non-monotonic dependence of the irreversible component of magnetization on a magnetic field or temperature. Depending on the context and author’s preferences, this feature can be called the “peak effect”, the “second magnetization peak” or the “fishtail” [18, 5, 19, 20, 21, 6, 7, 22, 23, 24, 25, 26, 27, 28]. Since any measurement has a certain experimental time window, the measured magnetic moment or current density is affected by magnetic relaxation, which is exponentially fast at current densities close to the critical current [6, 7, 8]. Therefore, there is always a question as to whether the peak effect is due to the actual non-monotonic behavior of jc(H,T)subscript𝑗𝑐𝐻𝑇j_{c}\left(H,T\right)italic_j start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_H , italic_T ), which would imply an unusual pinning mechanism, or whether it is the result of a non-monotonic magnetic relaxation. This latter “dynamic” mechanism is predicted in the weak collective pinning and creep model [6] and the former “static” mechanism can be, for example, due to the softening of the vortex lattice at low fields and close to Hc2subscript𝐻𝑐2H_{c2}italic_H start_POSTSUBSCRIPT italic_c 2 end_POSTSUBSCRIPT [2, 5], two different vortex phases [29] and a crossover from the collective to the plastic creep mechanism [30, 23]. Importantly, in either case, the reversible magnetization is always considered to be a monotonic function of a magnetic field and temperature. In fact, we are not aware of any reports of the non-monotonic behavior of reversible magnetization, except the so-called paramagnetic Meissner effect (PME), which seems to be related to extrinsic factors such as sample inhomogeneities, granularity, demagnetizing factors, form factor, or inhomogeneous cooling conditions [31, 32, 33].

Superconducting Rh17S15subscriptRh17subscriptS15\text{Rh}_{17}\text{S}_{15}Rh start_POSTSUBSCRIPT 17 end_POSTSUBSCRIPT S start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT with Tc=5.4Ksubscript𝑇𝑐5.4KT_{c}=5.4\>\text{K}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = 5.4 K has recently attracted attention as the rare case of a cubic compound with line nodes in its superconducting gap inferred from the Tlimit-from𝑇T-italic_T -linear variation of the London penetration depth measured down to 50 mK, and a strong suppression of Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT by non-magnetic disorder [34]. Thermal conductivity measurements down to 100 mK on the lower Tc=5.0Ksubscript𝑇𝑐5.0KT_{c}=5.0\>\text{K}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = 5.0 K sample show nodal-like concave field dependence but no residual term [35].

Despite its modest Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, Rh17S15subscriptRh17subscriptS15\text{Rh}_{17}\text{S}_{15}Rh start_POSTSUBSCRIPT 17 end_POSTSUBSCRIPT S start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT exhibits quite unusual superconducting properties. It has a very large upper critical field, Hc2(0)20.5Tsubscript𝐻𝑐2020.5TH_{c2}\left(0\right)\approx 20.5\>\text{T}italic_H start_POSTSUBSCRIPT italic_c 2 end_POSTSUBSCRIPT ( 0 ) ≈ 20.5 T, determined from the fit to Helfand and Werthamer theory [36] of the experimental zero resistivity data measured along the [111]delimited-[]111\left[111\right][ 111 ] direction where the last data point from the R(H)𝑅𝐻R(H)italic_R ( italic_H ) scan at 0.5 K shows Hc2subscript𝐻𝑐2H_{c2}italic_H start_POSTSUBSCRIPT italic_c 2 end_POSTSUBSCRIPT just below 20 T [37]. Data from polycrystalline samples yielded a similar value of Hc2(0)20Tsubscript𝐻𝑐2020TH_{c2}\left(0\right)\approx 20\>\text{T}italic_H start_POSTSUBSCRIPT italic_c 2 end_POSTSUBSCRIPT ( 0 ) ≈ 20 T [38, 39]. This Hc2(0)subscript𝐻𝑐20H_{c2}\left(0\right)italic_H start_POSTSUBSCRIPT italic_c 2 end_POSTSUBSCRIPT ( 0 ) exceeds the Pauli paramagnetic limiting field by a factor of two, Hp(0)=Δ(0)/(2μB)10Tsubscript𝐻𝑝0Δ02subscript𝜇𝐵10TH_{p}(0)=\Delta(0)/(\sqrt{2}\mu_{B})\approx 10\>\text{T}italic_H start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ( 0 ) = roman_Δ ( 0 ) / ( square-root start_ARG 2 end_ARG italic_μ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT ) ≈ 10 T, where the superconducting gap is estimated from the weak-coupling isotopic Bardeen-Cooper-Schrieffer (BCS) theory [40], Δ(0)=1.76kBTc0.82meVΔ01.76subscript𝑘𝐵subscript𝑇𝑐0.82meV\Delta(0)=1.76k_{B}T_{c}\approx 0.82\>\text{meV}roman_Δ ( 0 ) = 1.76 italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ≈ 0.82 meV. There is a large difference between the coherence length of about ξ(0)4.0𝜉04.0\xi\left(0\right)\approx 4.0italic_ξ ( 0 ) ≈ 4.0 nm, derived from Hc2(0)=ϕ0/(2πξ2)subscript𝐻𝑐20subscriptitalic-ϕ02𝜋superscript𝜉2H_{c2}(0)=\phi_{0}/(2\pi\xi^{2})italic_H start_POSTSUBSCRIPT italic_c 2 end_POSTSUBSCRIPT ( 0 ) = italic_ϕ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / ( 2 italic_π italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ), and the BCS length scale ξ0=vF/(πΔ(0))subscript𝜉0Planck-constant-over-2-pisubscript𝑣𝐹𝜋Δ0\xi_{0}=\hbar v_{F}/(\pi\Delta(0))italic_ξ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = roman_ℏ italic_v start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT / ( italic_π roman_Δ ( 0 ) )21.8nmabsent21.8nm\approx 21.8\>\text{nm}≈ 21.8 nm, which is more than five times greater. Here, vF=0.85×105m/ssubscript𝑣𝐹0.85superscript105msv_{F}=0.85\times 10^{5}\>\text{m}/\text{s}italic_v start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT = 0.85 × 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT m / s is the band-averaged Fermi velocity evaluated by our DFT calculations. In the weak-coupling BCS theory, in the clean limit (which is the case here), these two lengths are of the same order, ξ0/ξ=1.63subscript𝜉0𝜉1.63\xi_{0}/\xi=1.63italic_ξ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_ξ = 1.63 for isotropic slimit-from𝑠s-italic_s -wave [41, 42] and, similarly, with a slightly different numerical prefactor, for arbitrary klimit-from𝑘k-italic_k - dependent order parameter, including line nodal dlimit-from𝑑d-italic_d -wave [43]. Therefore, an extremely high Hc2(0)subscript𝐻𝑐20H_{c2}(0)italic_H start_POSTSUBSCRIPT italic_c 2 end_POSTSUBSCRIPT ( 0 ) alone represents a significant departure from the BCS theory for this relatively lowTcsubscript𝑇𝑐-T_{c}- italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT superconductor. All these properties imply that Rh17S15subscriptRh17subscriptS15\text{Rh}_{17}\text{S}_{15}Rh start_POSTSUBSCRIPT 17 end_POSTSUBSCRIPT S start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT is the first unconventional superconductor whose formula can be found in nature in the form of mineral miassite [34].

There is only limited information on the vortex properties of Rh17S15subscriptRh17subscriptS15\text{Rh}_{17}\text{S}_{15}Rh start_POSTSUBSCRIPT 17 end_POSTSUBSCRIPT S start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT, because most studies have been performed on polycrystalline samples. For single crystals, only critical fields were reported, but without direction-resolved data. The broad peak effect in magnetization was observed as a function of the magnetic field and some non-monotonic signatures were observed in magnetic susceptibility as a function of temperature [27]. The authors found a strong dependence of the persistent current density on the time window of the experiment and suggested a mix of static and dynamic scenarios where the vortex lattice becomes progressively more disordered with an increasing magnetic field. Perhaps, because of its cubic crystal structure, it was believed that this does not make much difference. However, cubic materials can exhibit electronic anisotropy [44], including in the vortex state, for example of the upper critical field of niobium [45].

In this paper, we report the significant anisotropy of the vortex response in single crystals of Rh17S15subscriptRh17subscriptS15\text{Rh}_{17}\text{S}_{15}Rh start_POSTSUBSCRIPT 17 end_POSTSUBSCRIPT S start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT below the upper critical field, Hc2subscript𝐻𝑐2H_{c2}italic_H start_POSTSUBSCRIPT italic_c 2 end_POSTSUBSCRIPT. A highly unusual paramagnetic reversible peak effect was observed in magnetization as a function of temperature and magnetic field for H[111]conditional𝐻delimited-[]111H\parallel\left[111\right]italic_H ∥ [ 111 ] and it is reduced when H[110]conditional𝐻delimited-[]110H\parallel\left[110\right]italic_H ∥ [ 110 ]. For H[100]conditional𝐻delimited-[]100H\parallel\left[100\right]italic_H ∥ [ 100 ] instead of a peak, a step in M(T)𝑀𝑇M(T)italic_M ( italic_T ) develops. These obsevations indicate that vortices prefer to align along the [111]delimited-[]111\left[111\right][ 111 ] “easy direction”. The effect is suppressed by disorder. It appears that there is a novel vortex lattice phase that has never been reported before.

II Results

The basic properties and characterization of our single crystals with a focus on the superconducting gap structure have previously been reported [34]. In low fields, there is a sharp superconducting transition in magnetization and resistivity. In small magnetic fields, there is a significant difference between zero-field-cooled (ZFC) and field-cooled (FC) magnetization. For example, following the ZFC protocol by applying H=10Oe𝐻10OeH=10\,\mathrm{Oe}italic_H = 10 roman_Oe after cooling to a base temperature, T=2K𝑇2KT=2\,\mathrm{K}italic_T = 2 roman_K, results in volume magnetic susceptibility, χ1𝜒1\chi\approx-1italic_χ ≈ - 1, indicating a complete diamagnetic screening in all three orientations. Cooling in the same field from above Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT gives only χ0.018,0.023,0.027𝜒0.0180.0230.027\chi\approx-0.018,\,-0.023,\,-0.027italic_χ ≈ - 0.018 , - 0.023 , - 0.027 in the [100]delimited-[]100\left[100\right][ 100 ], [110]delimited-[]110\left[110\right][ 110 ] and [111]delimited-[]111\left[111\right][ 111 ] directions, respectively. This indicates that the pinning is weakest in the [111]delimited-[]111\left[111\right][ 111 ] direction. With an increasing magnetic field, the transition curves smear and at H=1.5T𝐻1.5TH=1.5\>\text{T}italic_H = 1.5 T become reversible, showing practically no hysteresis between ZFC and FC magnetization.

There is a small but notable paramagnetic signal in the normal state, also reported previously in polycrystalline samples [38, 39]. Just above Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, the volume magnetic susceptibility is on the order of χ=1×104𝜒1superscript104\chi=1\times 10^{-4}italic_χ = 1 × 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT. This is comparable to the magnetic susceptibility of tungsten at room temperature, which is regarded as “non magnetic”. Interestingly, this paramagnetic signal is temperature dependent. It was suggested that sharp peaks in the density of states at the Fermi level could result in temperature-dependent Pauli spin susceptibility. This mechanism has also been suggested for V3SisubscriptV3Si\text{V}_{3}\text{Si}V start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT Si where the paramagnetic susceptibility per vanadium ion is of the same order as for rhosium in Rh17S15subscriptRh17subscriptS15\text{Rh}_{17}\text{S}_{15}Rh start_POSTSUBSCRIPT 17 end_POSTSUBSCRIPT S start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT [46, 39]. Perhaps, this active magnetic channel is related to the peak effect reported here.

In discussing magnetic anisotropy, it is important to analyze the possible effect of the sample shape, which enters the magnetic response via demagnetizing factors N𝑁Nitalic_N [47]. For our fields of interest, above 1 T, demagnetizing correction can be safely neglected. Specifically, the effective magnetic field strength, seen by the sample, is Heff=HappNM/Vsubscript𝐻effsubscript𝐻app𝑁𝑀𝑉H_{\text{eff}}=H_{\text{app}}-NM/Vitalic_H start_POSTSUBSCRIPT eff end_POSTSUBSCRIPT = italic_H start_POSTSUBSCRIPT app end_POSTSUBSCRIPT - italic_N italic_M / italic_V, where Happsubscript𝐻appH_{\text{app}}italic_H start_POSTSUBSCRIPT app end_POSTSUBSCRIPT is the applied magnetic field and V=1.7×109m3𝑉1.7superscript109superscriptm3V=1.7\times 10^{-9}\>\text{m}^{3}italic_V = 1.7 × 10 start_POSTSUPERSCRIPT - 9 end_POSTSUPERSCRIPT m start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT is the volume of the sample. Even if we take an exaggerated value of the measured magnetic moment, M=103erg/G=106Am2𝑀superscript103erg/Gsuperscript106superscriptAm2M=10^{-3}\>\text{erg/G}=10^{-6}\>\text{Am}^{2}italic_M = 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT erg/G = 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT Am start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, the corresponding volume magnetization is M/V=580A/m𝑀𝑉580A/mM/V=580\>\text{A/m}italic_M / italic_V = 580 A/m. The applied magnetic field strength corresponding to a magnetic induction of 1 T is Happ=Bapp/μ08×105A/msubscript𝐻𝑎𝑝𝑝subscript𝐵𝑎𝑝𝑝subscript𝜇08superscript105A/mH_{app}=B_{app}/\mu_{0}\approx 8\times 10^{5}\>\text{A/m}italic_H start_POSTSUBSCRIPT italic_a italic_p italic_p end_POSTSUBSCRIPT = italic_B start_POSTSUBSCRIPT italic_a italic_p italic_p end_POSTSUBSCRIPT / italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≈ 8 × 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT A/m. Since the demagnetizing factor is bound by 0N<10𝑁10\leq N<10 ≤ italic_N < 1 (in our case N0.350.55𝑁0.350.55N\approx 0.35-0.55italic_N ≈ 0.35 - 0.55 depending on the orientation), the NM/V𝑁𝑀𝑉NM/Vitalic_N italic_M / italic_V correction term is of the order of 200400A/m200400A/m200-400\>\text{A/m}200 - 400 A/m (2.55.0G2.55.0G2.5-5.0\>\text{G}2.5 - 5.0 G), which is more than three orders of magnitude smaller than the applied field strength (10000G10000G10000\>\text{G}10000 G).

Refer to caption
Figure 1: Zero-field-cooled (ZFC) and field-cooled (FC) magnetization of the same single crystal Rh17S15subscriptRh17subscriptS15\text{Rh}_{17}\text{S}_{15}Rh start_POSTSUBSCRIPT 17 end_POSTSUBSCRIPT S start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT in three different orientations, (a) H[100]conditional𝐻delimited-[]100H\parallel\left[100\right]italic_H ∥ [ 100 ], (b) H[110]conditional𝐻delimited-[]110H\parallel\left[110\right]italic_H ∥ [ 110 ], (c) H[111]conditional𝐻delimited-[]111H\parallel\left[111\right]italic_H ∥ [ 111 ]. Each curve was measured in an indicated magnetic field and contains data on warming (ZFC) and cooling (FC), as shown by red arrows. The ZFC and FC curves coincide within the noise, indicating practically reversible magnetization. A pronounced reversible peak effect develops above H=1.5T𝐻1.5TH=1.5\>\text{T}italic_H = 1.5 T in [111]delimited-[]111\left[111\right][ 111 ] orientation. Red dashed lines show that without the peak, there would be a continuous transition starting at Hc2subscript𝐻𝑐2H_{c2}italic_H start_POSTSUBSCRIPT italic_c 2 end_POSTSUBSCRIPT. The peak is reduced in the [110]delimited-[]110\left[110\right][ 110 ] direction. Instead of a peak, [100]delimited-[]100\left[100\right][ 100 ] direction shows a step-like diamagnetic response with the step amplitude increasing with the increase of an applied magnetic field, unexpected for conventional Meissner expulsion.

Figure 1 shows the ZFC and FC magnetization of the same Rh17S15subscriptRh17subscriptS15\text{Rh}_{17}\text{S}_{15}Rh start_POSTSUBSCRIPT 17 end_POSTSUBSCRIPT S start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT single crystal in three different orientations, (a) H[100]conditional𝐻delimited-[]100H\parallel\left[100\right]italic_H ∥ [ 100 ], (b) H[110]conditional𝐻delimited-[]110H\parallel\left[110\right]italic_H ∥ [ 110 ], and (c) H[111]conditional𝐻delimited-[]111H\parallel\left[111\right]italic_H ∥ [ 111 ], measured in indicated magnetic fields. Each curve contains the ZFC and FC data, which are indistinguishable within the experimental noise, as shown by red arrows. The magnetization is practically reversible. Note that we chose not to convert magnetization (shown as raw data in CGS erg/G units) to avoid division by a magnetic field, which would make the graph less clear.

The central result of this paper is an unexpected pronounced, reversible paramagnetic peak developing roughly above H=1.5T𝐻1.5TH=1.5\>\text{T}italic_H = 1.5 T in a [111]delimited-[]111\left[111\right][ 111 ] orientation, but not present in the [100]delimited-[]100\left[100\right][ 100 ] direction where there is a pronounced, steplike diamagnetic decrease in M(T)𝑀𝑇M(T)italic_M ( italic_T ) on cooling, but no peak. In the intermediate orientation, [110]delimited-[]110\left[110\right][ 110 ], there is a reduced peak effect. The noise developing below the transition in Fig. 1 (a) and (b) is not due to instrumentation but is generated by the sample and most likely reflects the enhanced motion of vortices. The red dashed lines in Figure 1(c) show that without the peak, there would be a continuous transition starting at Hc2subscript𝐻𝑐2H_{c2}italic_H start_POSTSUBSCRIPT italic_c 2 end_POSTSUBSCRIPT. This indicates that a new distinct vortex phase is formed inside the superconducting phase, disconnected from the Hc2(T)subscript𝐻𝑐2𝑇H_{c2}(T)italic_H start_POSTSUBSCRIPT italic_c 2 end_POSTSUBSCRIPT ( italic_T ) line.

The peak in a [111]delimited-[]111\left[111\right][ 111 ] direction is truly paramagnetic, as it exceeds the value in the normal state above Tc(Hc2)subscript𝑇𝑐subscript𝐻𝑐2T_{c}(H_{c2})italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_H start_POSTSUBSCRIPT italic_c 2 end_POSTSUBSCRIPT ), where the dimensionless paramagnetic volume susceptibility is of the order of χ=1×104𝜒1superscript104\chi=1\times 10^{-4}italic_χ = 1 × 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT, and the peak height from the baseline is approximately ten times smaller; see the inset in Fig.6. We also stress that, while the [100]delimited-[]100\left[100\right][ 100 ] direction does not show a peak, its temperature dependence is also unusual. There is a step in M(T)𝑀𝑇M(T)italic_M ( italic_T ) below Hc2subscript𝐻𝑐2H_{c2}italic_H start_POSTSUBSCRIPT italic_c 2 end_POSTSUBSCRIPT and, importantly, its amplitude increases with the increase of the applied magnetic field, contrary to the expectations of the standard Meissner expulsion. As can be seen in Fig. 1(a), there is practically no step in M(T)𝑀𝑇M(T)italic_M ( italic_T ) at 1.5 T, but it grows with the applied field and its amplitude is comparable to the amplitude of the peak in the [100]delimited-[]100\left[100\right][ 100 ] direction.

With such unusual observations, the first step is to verify the effect on other samples from a different batch. This is shown in Fig. 2 where the temperature-dependent magnetic moment measured along the [111]delimited-[]111\left[111\right][ 111 ] orientation is plotted for two different samples, (a) sample A and (b) sample B from a different batch. We measured several other samples and confirmed that the effect is reproducible.

Refer to caption
Figure 2: Zero-field-cooled (ZFC) and field-cooled (FC) magnetization of two different crystals of Rh17S15subscriptRh17subscriptS15\text{Rh}_{17}\text{S}_{15}Rh start_POSTSUBSCRIPT 17 end_POSTSUBSCRIPT S start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT measured with H[111]conditional𝐻delimited-[]111H\parallel\left[111\right]italic_H ∥ [ 111 ] in indicated magnetic fields. The curves are reversible as indicated by red arrows.

The next step is to examine the magnetic field dependence of the magnetic moment measured at a fixed temperature. Here we observed another surprising result, the peak effect, but unlike the peak effects and fishtails reported so far, this peak effect has a significant, dominant, reversible component, which is consistent with the peak shown in Fig. 1. Figure 3 shows the magnetic moment measured along the [111]delimited-[]111\left[111\right][ 111 ] direction at T=3.6K𝑇3.6KT=3.6\>\text{K}italic_T = 3.6 K in sample A. The external magnetic field was swept as shown in the inset in Fig. 3, repeating down and up sweeps at the same rate, indicated by arrows, and then switching the ramp rate and repeating the measurement. Four different ramp rates, from 200 Oe/s, 100 Oe/s, 50 Oe/s and 12 Oe/s, were used, a seventeen-fold reduction. There is a significant magnetic relaxation with the blue curve obtained at the slowest 12 Oe/s showing a practically reversible nonmonotonic M(H)𝑀𝐻M(H)italic_M ( italic_H ) loop.

Refer to caption
Figure 3: Magnetic moment in sample A as a function of the applied magnetic field measured along the [111]delimited-[]111\left[111\right][ 111 ] direction at T=3.6K𝑇3.6KT=3.6\>\text{K}italic_T = 3.6 K. The up and down sweeps are performed at the same rate in the directions indicated by the arrows. Then the sweeps were repeated at different ramp rates indicated in the inset, which shows the time profile of the magnetic field sweeps. There is a significant magnetic relaxation with the blue curve obtained at 12Oe/s12Oe/s12\>\text{Oe/s}12 Oe/s showing practically reversible M(H)𝑀𝐻M(H)italic_M ( italic_H ) loop.

We now compare the magnetic hysteresis M(H)𝑀𝐻M\left(H\right)italic_M ( italic_H ) loops measured in the same sample A in three principal orientations. Figure 4(a) shows the significant anisotropy of the response. There is a conventionally looking “fishtail” in the [100]delimited-[]100\left[100\right][ 100 ] orientation (blue curve), but an asymmetric, almost reversible, peak effect in the [111]delimited-[]111\left[111\right][ 111 ] orientation (red curve). Estimated from the width of the magnetic hysteresis, the maximum persistent current for the [100]delimited-[]100\left[100\right][ 100 ] orientation is about ten times greater than for the [111]delimited-[]111\left[111\right][ 111 ] orientation, and is intermediate for the [110]delimited-[]110\left[110\right][ 110 ] orientation. The location of the peak shifts to the lower fields from [111]delimited-[]111\left[111\right][ 111 ] to [110]delimited-[]110\left[110\right][ 110 ] to [100]delimited-[]100\left[100\right][ 100 ]. Such significant anisotropy in an electronically isotropic cubic system is remarkable but apparently not impossible in the vortex state. In order to better compare the loops, a small paramagnetic background, linear in H𝐻Hitalic_H, was subtracted. The original loops are shown in the insets with backgrounds shown by red lines. In the pristine state, a full M(H)𝑀𝐻M(H)italic_M ( italic_H ) loop is shown, including an often-observed sharp peak near zero field. In this scale, the background is practically zero. For the irradiated sample, discussed later, a truncated loop is shown to better visualize the background contribution. Importantly, the paramagnetic background is practically identical for all three orientations as well as before and after the electron irradiation.

Refer to caption
Figure 4: Magnetic hysteresis M(H)𝑀𝐻M(H)italic_M ( italic_H ) loops of the same single crystal of Rh17S15subscriptRh17subscriptS15\text{Rh}_{17}\text{S}_{15}Rh start_POSTSUBSCRIPT 17 end_POSTSUBSCRIPT S start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT measured in sample A along three principal directions at T=3.4K𝑇3.4KT=3.4\>\text{K}italic_T = 3.4 K with small paramagnetic background subtracted. The magnetic field ramp rate was 100 Oe/s. Panel (a) shows the results in the pristine crystal and panel (b) shows the same crystal but after 2.5MeV2.5MeV2.5\>\text{MeV}2.5 MeV electron irradiation with the dose of 6.2×1018electrons/cm26.2superscript1018electronssuperscriptcm26.2\times 10^{18}\>\text{electrons}/\text{cm}^{2}6.2 × 10 start_POSTSUPERSCRIPT 18 end_POSTSUPERSCRIPT electrons / cm start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. The xlimit-from𝑥x-italic_x - and ylimit-from𝑦y-italic_y - axes scales are shared in both panels to allow for direct comparison. Insets show M(H)𝑀𝐻M(H)italic_M ( italic_H ) loops before background subtraction: (a) a full loop including a pronounced zero-field peak; (b) a truncated M(H)𝑀𝐻M(H)italic_M ( italic_H ) loop shown to emphasize the background line.

Another important metric of a superconductor is the response to a controlled disorder. Previously, we used electron irradiation to probe the superconducting transition temperature of Rh17S15 and found results consistent with the line-nodal superconducting gap [34]. Here we examine the effect of non-magnetic pointlike disorder on the observed paramagnetic reversible peak effect. Figure 4(b) shows magnetization loops in the peak region for three orientations after 2.5MeV2.5MeV2.5\>\text{MeV}2.5 MeV electron irradiation with the dose of 6.2×1018electrons/cm26.2superscript1018electronssuperscriptcm26.2\times 10^{18}\>\text{electrons}/\text{cm}^{2}6.2 × 10 start_POSTSUPERSCRIPT 18 end_POSTSUPERSCRIPT electrons / cm start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. For direct comparison with the pristine case, the xlimit-from𝑥x-italic_x - and ylimit-from𝑦y-italic_y - axes scales are shared in both panels (a) and (b) of Fig. 4, pristine and irradiated, respectively. The initially almost reversible peak in the [111]delimited-[]111\left[111\right][ 111 ] orientation evolves with disorder into a “normal” fishtail dominated by an irreversible component. Considering the very low pinning in our Rh17S15subscriptRh17subscriptS15\text{Rh}_{17}\text{S}_{15}Rh start_POSTSUBSCRIPT 17 end_POSTSUBSCRIPT S start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT crystals, it is possible that the peak effect observed in other materials with larger pinning may hide the reversible phase in their hypothetical clean state.

The destructive effect of disorder on the observed reversible peak effect is further illustrated in Fig. 5, where M(T)𝑀𝑇M(T)italic_M ( italic_T ) ZFC-FC temperature scans are compared for the same sample A in a [111]delimited-[]111\left[111\right][ 111 ] orientation in (a) pristine and (b) electron irradiated states. The peak is smeared and, again, magnetic noise appears similar to that in Fig. 1(a),(b), probably indicative of significant vortex displacements and jumps between metastable configurations.

Refer to caption
Figure 5: ZFC and FC magnetization of the same Rh17S15subscriptRh17subscriptS15\text{Rh}_{17}\text{S}_{15}Rh start_POSTSUBSCRIPT 17 end_POSTSUBSCRIPT S start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT single crystal A measured along [111]delimited-[]111\left[111\right][ 111 ] direction, (a) before and (b) after 2.5MeV2.5MeV2.5\>\text{MeV}2.5 MeV electron irradiation with the dose of 6.2×1018electrons/cm26.2superscript1018electronssuperscriptcm26.2\times 10^{18}\>\text{electrons}/\text{cm}^{2}6.2 × 10 start_POSTSUPERSCRIPT 18 end_POSTSUPERSCRIPT electrons / cm start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT.

Finally, we construct an HT𝐻𝑇H-Titalic_H - italic_T phase diagram that includes the novel reversible vortex phase. Figure 6 shows different H(T)𝐻𝑇H(T)italic_H ( italic_T ) lines of characteristic temperatures defined as shown in the inset. In the inset, the paramagnetic background above Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT was subtracted to show the paramagnetic volume susceptibility of the peak in absolute units. The blue circles mark the temperature of the peak maximum, Tmaxsubscript𝑇𝑚𝑎𝑥T_{max}italic_T start_POSTSUBSCRIPT italic_m italic_a italic_x end_POSTSUBSCRIPT, and violet circles show the peak shoulder, Tmax2subscript𝑇𝑚𝑎𝑥2T_{max2}italic_T start_POSTSUBSCRIPT italic_m italic_a italic_x 2 end_POSTSUBSCRIPT. The orange area is bound by the onset and offset temperatures. It represents the domain inside which the novel vortex phase exists in our range of fields and temperatures. For comparison, additional lines are obtained tracing the peak position in the M(H)𝑀𝐻M\left(H\right)italic_M ( italic_H ) loops measured at different temperatures. Green- and violet-filled stars show the results for the pristine and irradiated samples, respectively. Remarkably, the peak positions determined from the two measurements coincide in pristine samples. There is no peak in the M(T)𝑀𝑇M(T)italic_M ( italic_T ) measurements after irradiation and the peak in the M(H)𝑀𝐻M(H)italic_M ( italic_H ) loops is shifted to lower temperatures. To complete the phase diagram, Fig. 6 shows the upper critical field in a [111]delimited-[]111\left[111\right][ 111 ] orientation before (yellow squares) and after (yellow circles) electron irradiation. The line Hc2(T)subscript𝐻𝑐2𝑇H_{c2}(T)italic_H start_POSTSUBSCRIPT italic_c 2 end_POSTSUBSCRIPT ( italic_T ) is shifted to lower temperatures after irradiation, consistent with our earlier conclusion that Rh17S15subscriptRh17subscriptS15\text{Rh}_{17}\text{S}_{15}Rh start_POSTSUBSCRIPT 17 end_POSTSUBSCRIPT S start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT is a line nodal superconductor.

Refer to caption
Figure 6: The HT𝐻𝑇H-Titalic_H - italic_T state phase diagram including the novel vortex phase. The definitions of characteristic temperatures of the reversible M(T)𝑀𝑇M\left(T\right)italic_M ( italic_T ) peak are shown in the inset. The blue circles mark the temperature of the peak maximum Tmaxsubscript𝑇𝑚𝑎𝑥T_{max}italic_T start_POSTSUBSCRIPT italic_m italic_a italic_x end_POSTSUBSCRIPT, and the violet circles show the temperature of the knee (can also be seen as the second maximum) of the peak, Tmax2subscript𝑇𝑚𝑎𝑥2T_{max2}italic_T start_POSTSUBSCRIPT italic_m italic_a italic_x 2 end_POSTSUBSCRIPT. Onset and offset temperatures are the boundaries of the orange area inside which the peak is located. Green stars and purple stars mark the location of the peak effect maximum in the M(H)𝑀𝐻M\left(H\right)italic_M ( italic_H ) measurements in pristine and electron-irradiated samples, respectively. The yellow squares and circles show the upper critical field in a [111]delimited-[]111\left[111\right][ 111 ] direction before and after electron irradiation, respectively. In the inset, the paramagnetic background above Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT was subtracted to show the volume susceptibility of the peak in absolute units.

III Discussion

Although we do not have a microscopic explanation for the unusual anisotropic reversible vortex behavior, we can state that we observed a distinct intermediate vortex state likely related to another order which may be either magnetic or from a competing superconducting channel. Due to the reversible nature of the ZFC-FC magnetization and the tendency to such a behavior in M(H)𝑀𝐻M(H)italic_M ( italic_H ) loops after relaxation of the irreversible component, we can exclude non-equilibrium flux trapping and vortex density gradients associated with pinning. This implies that a uniform vortex lattice experiences some compression, which means that the vortices attract each other. Despite a very low pinning the field-cooled magnetic susceptibility is surprisingly small, which may also indicate that the Meissner expulsion is hindered by an attractive force.

The observed paramagnetic effect is small. In terms of dimensionless volume magnetic susceptibility, the peak height increases with increasing magnetic field and appears to saturate around Δχ+105Δ𝜒superscript105\Delta\chi\approx+10^{-5}roman_Δ italic_χ ≈ + 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT. This is on top of the paramagnetic background above Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, which is about χ+104𝜒superscript104\chi\approx+10^{-4}italic_χ ≈ + 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT. We can estimate how much the distance between the vortices should change to cause such a peak. For a triangular lattice, the intervortex distance is, a=2ϕ0/3B𝑎2subscriptitalic-ϕ03𝐵a=\sqrt{2\phi_{0}/\sqrt{3}B}italic_a = square-root start_ARG 2 italic_ϕ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / square-root start_ARG 3 end_ARG italic_B end_ARG. The magnetic susceptibility for uniform B𝐵Bitalic_B is, χ=(B/Happ1)𝜒𝐵subscript𝐻𝑎𝑝𝑝1\chi=\left(B/H_{app}-1\right)italic_χ = ( italic_B / italic_H start_POSTSUBSCRIPT italic_a italic_p italic_p end_POSTSUBSCRIPT - 1 ) (as discussed above, the demagnetizing correction is not important in large fields of the effect). Therefore, to the first order, the change in the intervortex spacing, Δa=aa0Δχa0/2Δ𝑎𝑎subscript𝑎0Δ𝜒subscript𝑎02\Delta a=a-a_{0}\approx-\Delta\chi a_{0}/2roman_Δ italic_a = italic_a - italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≈ - roman_Δ italic_χ italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / 2. Positive ΔχΔ𝜒\Delta\chiroman_Δ italic_χ means compression. Numerically, for μ0Happ=9Tsubscript𝜇0subscript𝐻𝑎𝑝𝑝9T\mu_{0}H_{app}=9\>\text{T}italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_H start_POSTSUBSCRIPT italic_a italic_p italic_p end_POSTSUBSCRIPT = 9 T, the spacing is a0=16.3nmsubscript𝑎016.3nma_{0}=16.3\>\text{nm}italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 16.3 nm and we find that the vortices must move by a small distance, Δa9.7×105nmΔ𝑎9.7superscript105nm\Delta a\approx 9.7\times 10^{-5}\>\text{nm}roman_Δ italic_a ≈ 9.7 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT nm. Therefore, if some attractive force causes the observed paramagnetic peak, it is extremely small yet finite. An interesting and important aspect is that the situation is highly anisotropic with a step instead of a peak observed in the [100]delimited-[]100\left[100\right][ 100 ] direction. To explain this, we suggest that the vortices tend to alight along the [111]delimited-[]111\left[111\right][ 111 ] crystallographic direction, so that when a magnetic field is aligned in other directions and in the absence of pinning, vortices rotate toward [111]delimited-[]111\left[111\right][ 111 ] causing a reduction of the [100]delimited-[]100\left[100\right][ 100 ] component of the magnetization vector. Since our magnetic measurements are always performed along the direction of an applied magnetic field, this leads to a step-like feature in [100]delimited-[]100\left[100\right][ 100 ] orientation.

Throughout the paper, we emphasized a small but finite paramagnetic background. Considering the paramagnetic nature of the described effect, it is possible that normal-state magnetism plays some role, although the mechanism is unclear.

In the discussion of the unusual effects observed, it is important to estimate some relevant parameters for Rh17S15subscriptRh17subscriptS15\text{Rh}_{17}\text{S}_{15}Rh start_POSTSUBSCRIPT 17 end_POSTSUBSCRIPT S start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT. First, we evaluate the fundamental length scales. Using the thermodynamic Rutgers relation, we previously obtained the value of London penetration depth of λ(0)=550nm𝜆0550nm\lambda\left(0\right)=550\>\text{nm}italic_λ ( 0 ) = 550 nm [34]. The only other work on single crystals used magnetization measurements to determine the lower critical field, which yielded a similar value of λ(0)=490nm𝜆0490nm\lambda\left(0\right)=490\>\text{nm}italic_λ ( 0 ) = 490 nm [37], although such estimates have a very large error bar. The coherence length is ξ(0)=4nm𝜉04nm\xi(0)=4\>\text{nm}italic_ξ ( 0 ) = 4 nm for Hc2(0)=20.5Tsubscript𝐻𝑐2020.5TH_{c2}(0)=20.5\>\text{T}italic_H start_POSTSUBSCRIPT italic_c 2 end_POSTSUBSCRIPT ( 0 ) = 20.5 T. Therefore, the Ginzburg-Landau parameter κ137.3𝜅137.3\kappa\approx 137.3italic_κ ≈ 137.3 places Rh17S15subscriptRh17subscriptS15\text{Rh}_{17}\text{S}_{15}Rh start_POSTSUBSCRIPT 17 end_POSTSUBSCRIPT S start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT in an extreme type-II limit. Overall, single crystals show a very small amount of hysteresis with reversible magnetization dominating over the irreversible part, even in the regions of the “fishtail” in M(H)𝑀𝐻M\left(H\right)italic_M ( italic_H ) loops. This is the case for all orientations of the applied magnetic field. However, in the peak effect region, the hysteresis in the [111]delimited-[]111\left[111\right][ 111 ] orientation is about ten time smaller than in the [100]delimited-[]100\left[100\right][ 100 ] orientation. This clean-limit behavior is also consistent with the nearly perfect Tlimit-from𝑇T-italic_T -linear behavior of λ(T)𝜆𝑇\lambda\left(T\right)italic_λ ( italic_T ) at low temperatures [34]. Furthermore, from the resistivity at the transition temperature, ρ(Tc)12μΩcm𝜌subscript𝑇𝑐12𝜇Ωcm\rho(T_{c})\approx 12\;\mu\Omega\cdot\text{cm}italic_ρ ( italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ) ≈ 12 italic_μ roman_Ω ⋅ cm, the mean free path can be estimated using the electronic band-structure parameters, =3/(vFN(0)e2ρ)91nm3subscript𝑣𝐹𝑁0superscript𝑒2𝜌91nm\ell=3/(v_{F}N\left(0\right)e^{2}\rho)\approx 91\>\text{nm}roman_ℓ = 3 / ( italic_v start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT italic_N ( 0 ) italic_e start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ρ ) ≈ 91 nm, where the band-averaged product vFN(0)=1.1×1052s1m2J1subscript𝑣𝐹𝑁01.1superscript1052superscripts1superscriptm2superscriptJ1v_{F}N\left(0\right)=1.1\times 10^{52}\>\text{s}^{-1}\text{m}^{-2}\text{J}^{-1}italic_v start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT italic_N ( 0 ) = 1.1 × 10 start_POSTSUPERSCRIPT 52 end_POSTSUPERSCRIPT s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT m start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT J start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT. This yields the dimensionless scattering rate, Γ=0.88ξ0/=0.2<1Γ0.88subscript𝜉00.21\Gamma=0.88\xi_{0}/\ell=0.2<1roman_Γ = 0.88 italic_ξ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / roman_ℓ = 0.2 < 1, placing the material in the clean limit.

Vortex-relevant parameters include: the lower critical field, Hc1(0)=ϕ0(lnκ+0.497)/(4πλ2)29Gsubscript𝐻𝑐10subscriptitalic-ϕ0𝜅0.4974𝜋superscript𝜆229GH_{c1}\left(0\right)=\phi_{0}\left(\ln\kappa+0.497\right)/(4\pi\lambda^{2})% \approx 29\>\text{G}italic_H start_POSTSUBSCRIPT italic_c 1 end_POSTSUBSCRIPT ( 0 ) = italic_ϕ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( roman_ln italic_κ + 0.497 ) / ( 4 italic_π italic_λ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) ≈ 29 G for λ(0)=550nm𝜆0550nm\lambda(0)=550\>\text{nm}italic_λ ( 0 ) = 550 nm and ξ(0)=4nm𝜉04nm\xi(0)=4\>\text{nm}italic_ξ ( 0 ) = 4 nm; the characteristic energy scale, ε0=ϕ02/(4πμ0λ2(0))=5.6meV/nmsubscript𝜀0superscriptsubscriptitalic-ϕ024𝜋subscript𝜇0superscript𝜆205.6meVnm\varepsilon_{0}=\phi_{0}^{2}/\left(4\pi\mu_{0}\lambda^{2}(0)\right)=5.6\>\text% {meV}/\text{nm}italic_ε start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = italic_ϕ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / ( 4 italic_π italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_λ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( 0 ) ) = 5.6 meV / nm; and vortex line energy, εl=ε0(lnκ+0.497)=30.2meV/nmsubscript𝜀𝑙subscript𝜀0𝜅0.49730.2meVnm\varepsilon_{l}=\varepsilon_{0}(\ln\kappa+0.497)=30.2\>\text{meV}/\text{nm}italic_ε start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT = italic_ε start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( roman_ln italic_κ + 0.497 ) = 30.2 meV / nm. The thermodynamic critical field, Hc(0)=ϕ0/(22πξ(0)λ(0))=0.11Tsubscript𝐻𝑐0subscriptitalic-ϕ022𝜋𝜉0𝜆00.11TH_{c}(0)=\phi_{0}/\left(2\sqrt{2}\pi\xi(0)\lambda(0)\right)=0.11\>\text{T}italic_H start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( 0 ) = italic_ϕ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / ( 2 square-root start_ARG 2 end_ARG italic_π italic_ξ ( 0 ) italic_λ ( 0 ) ) = 0.11 T and the Ginzburg number that characterizes the width of the critical fluctuations, Gi=(γkBTc/ε0ξ(0))2/8=5.3×105𝐺𝑖superscript𝛾subscript𝑘𝐵subscript𝑇𝑐subscript𝜀0𝜉0285.3superscript105Gi=\left(\gamma k_{B}T_{c}/\varepsilon_{0}\xi(0)\right)^{2}/8=5.3\times 10^{-5}italic_G italic_i = ( italic_γ italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / italic_ε start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_ξ ( 0 ) ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 8 = 5.3 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT; here we assumed anisotropy parameter γ=1𝛾1\gamma=1italic_γ = 1. A practical formula for this parameter is: Gi=3.25×1016[nm2K2](κγλ(0)[nm]Tc[K])2𝐺𝑖3.25superscript1016delimited-[]superscriptnm2superscriptK2superscript𝜅𝛾𝜆0delimited-[]nmsubscript𝑇𝑐delimited-[]K2Gi=3.25\times 10^{-16}[\text{nm}^{-2}\text{K}^{-2}]\left(\kappa\gamma\lambda(0% )[\text{nm}]T_{c}[\text{K}]\right)^{2}italic_G italic_i = 3.25 × 10 start_POSTSUPERSCRIPT - 16 end_POSTSUPERSCRIPT [ nm start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT K start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT ] ( italic_κ italic_γ italic_λ ( 0 ) [ nm ] italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT [ K ] ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. The largest Gi𝐺𝑖Giitalic_G italic_i value is observed in high-Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT cuprates, of the order of 102superscript10210^{-2}10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT or more. For low-Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT conventional superconductors it is in the range of <107absentsuperscript107<10^{-7}< 10 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT and much less. For example, in niobium, assuming the clean case values of λ(0)=33nm𝜆033nm\lambda(0)=33\>\text{nm}italic_λ ( 0 ) = 33 nm, ξ(0)=93nm𝜉093nm\xi(0)=93\>\text{nm}italic_ξ ( 0 ) = 93 nm and Tc=9.4Ksubscript𝑇𝑐9.4KT_{c}=9.4\>\text{K}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = 9.4 K [48], Gi=4×1012𝐺𝑖4superscript1012Gi=4\times 10^{-12}italic_G italic_i = 4 × 10 start_POSTSUPERSCRIPT - 12 end_POSTSUPERSCRIPT. In iron-based superconductors, with the typical values of λ(0)=200nm𝜆0200nm\lambda(0)=200\>\text{nm}italic_λ ( 0 ) = 200 nm, ξ(0)=3nm𝜉03nm\xi(0)=3\>\text{nm}italic_ξ ( 0 ) = 3 nm and Tc=33Ksubscript𝑇𝑐33KT_{c}=33\>\text{K}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = 33 K [49], we obtain Gi=8×105𝐺𝑖8superscript105Gi=8\times 10^{-5}italic_G italic_i = 8 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT, see also [50]. Therefore, the Ginzburg number of Rh17S15subscriptRh17subscriptS15\text{Rh}_{17}\text{S}_{15}Rh start_POSTSUBSCRIPT 17 end_POSTSUBSCRIPT S start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT is comparable with iron pnictides and it is unusually large for low-Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT superconductors.

IV Conclusions

We report an unusual highly anisotropic reversible vortex phase in single crystals Rh17S15subscriptRh17subscriptS15\text{Rh}_{17}\text{S}_{15}Rh start_POSTSUBSCRIPT 17 end_POSTSUBSCRIPT S start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT. For H[111]conditional𝐻delimited-[]111H\parallel\left[111\right]italic_H ∥ [ 111 ], there is a pronounced peak in M(T)𝑀𝑇M(T)italic_M ( italic_T ). This peak is reduced for H[110]conditional𝐻delimited-[]110H\parallel[110]italic_H ∥ [ 110 ]. For H[100]conditional𝐻delimited-[]100H\parallel\left[100\right]italic_H ∥ [ 100 ], instead of a peak, a steep step develops. Both the peak and the step amplitudes increase with the increase of an applied magnetic field. In all orientations, the warming and cooling scans are reversible. For H[111]conditional𝐻delimited-[]111H\parallel\left[111\right]italic_H ∥ [ 111 ], there is also a reversible peak effect (inside the irreversible “fishtail”) in M(H)𝑀𝐻M(H)italic_M ( italic_H ) loops with the same location as in M(T)𝑀𝑇M(T)italic_M ( italic_T ) scans when plotted on a TH𝑇𝐻T-Hitalic_T - italic_H phase diagram. Non-magnetic point-like disorder induced by electron irradiation suppresses the unusual features. We suggest that the observed peak effect may be caused by a weak attractive interaction between vortices. Furthermore, H[111]conditional𝐻delimited-[]111H\parallel\left[111\right]italic_H ∥ [ 111 ] appears to be the “easy axis” for vortices. When a magnetic field is applied in a different direction, the vortices rotate toward H[111]conditional𝐻delimited-[]111H\parallel\left[111\right]italic_H ∥ [ 111 ], leading to a reduction in the magnetization projection on the measurement axis, which explains the step observed in the H[100]conditional𝐻delimited-[]100H\parallel\left[100\right]italic_H ∥ [ 100 ] orientation.

V Methods

Single crystal growth: Single crystalline samples of Rh17S15subscriptRh17subscriptS15\text{Rh}_{17}\text{S}_{15}Rh start_POSTSUBSCRIPT 17 end_POSTSUBSCRIPT S start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT were synthesized out of the Rh-S eutectic by using a high-temperature solution growth technique. The details of growth and characterization are provided elsewhere [51, 34]. Briefly, elemental rhodium powder and sulfur were combined in a fritted Canfield Crucible set[52], sealed in a silica ampoule, slowly heated (over 12 hours) to 1150 \celsius  and then slowly cooled from 1150 \celsius  to 920 \celsius  over 50 hours and decanted[53]. Millimeter-sized single crystals of Rh17S15subscriptRh17subscriptS15\text{Rh}_{17}\text{S}_{15}Rh start_POSTSUBSCRIPT 17 end_POSTSUBSCRIPT S start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT with well-defined facets were obtained.

Samples: Cuboid-shaped samples with visible (100)100\left(100\right)( 100 ), (110)110\left(110\right)( 110 ) and (111)111\left(111\right)( 111 ) facets were selected for measurements. Sample A, used for most figures and quantitative analysis, had dimensions: 1.35×1.25×1.10mm31.351.251.10superscriptmm31.35\times 1.25\times 1.10\>\text{mm}^{3}1.35 × 1.25 × 1.10 mm start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT along three orthogonal 100delimited-⟨⟩100\left<100\right>⟨ 100 ⟩ directions. It had volume V=1.72mm3𝑉1.72superscriptmm3V=1.72\>\text{mm}^{3}italic_V = 1.72 mm start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT and weighed 12.68mg12.68mg12.68\>\text{mg}12.68 mg. Sample B was 1.35×1.0×0.95mm31.351.00.95superscriptmm31.35\times 1.0\times 0.95\>\text{mm}^{3}1.35 × 1.0 × 0.95 mm start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT, V=1.28mm3𝑉1.28superscriptmm3V=1.28\>\text{mm}^{3}italic_V = 1.28 mm start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT and weight 9.43mg9.43mg9.43\>\text{mg}9.43 mg. We measured several other samples for statistics and consistently observed reported here results.

Magnetization measurements: Magnetic moment was measured using a Quantum Design vibrating sample magnetometer (VSM) in a 9T9T9\>\text{T}9 T Physical Property Measurement System (PPMS). Note that the magnetic moment is measured in CGS units, 1emu=1erg/G=103Am21emu1erg/Gsuperscript103superscriptAm21~{}\text{emu}=1~{}\text{erg/G}=10^{-3}\text{Am}^{2}1 emu = 1 erg/G = 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT Am start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. The samples were glued in proper orientation on Pyrex cylinders fitted inside the brass half-cylinder, a standar sample holder suitable for our samples. The orientation was done by eye following well-defined crystallographic planes. The same sample was repeatedly oriented and measured, then irradiated, and then the same set of measurements was repeated. Other samples from different batches were measured.

Electron irradiation: The low-temperature 2.5 MeV electron irradiation was performed at the “SIRIUS” facility of the Laboratoire des Solides Irradiés (LSI) at École Polytechnique in Palaiseau, France. The general description of such experiments and specific details are provided elsewhere [54, 55, 56]. Relativistic electrons are particularly suitable for the creation of point-like defects in solids, because the energy transfer upon their collisions with ions matches the knockout energy barriers, typically in the 10100eV10100eV10-100\>\text{eV}10 - 100 eV range in metallic compounds. A knocked out ion becomes an interstitial and, together with the created vacancy, forms the so-called Frenkel pair. The irradiation is carried out with the sample immersed in liquid hydrogen at about 22K22K22\>\text{K}22 K to avoid immediate recombination of Frenkel pairs and clustering of the produced defects. Upon warming to room temperature, a metastable population of defects remains after some defects recombined and migrated to surfaces and extended defects [54, 55]. The degree of annealing depends on the material and in Rh17S15subscriptRh17subscriptS15\text{Rh}_{17}\text{S}_{15}Rh start_POSTSUBSCRIPT 17 end_POSTSUBSCRIPT S start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT is about 30%. The amount of the produced disorder is gauged by the increase in resistivity after irradiation. Public domain software developed by NIST (https://physics.nist.gov) was used for electron propagation and energy loss calculations and custom SECTE software developed at the LSI at École Polytechnique was used for the scattering cross-section calculations. In this work, a single dose of 6.24×10186.24superscript10186.24\times 10^{18}6.24 × 10 start_POSTSUPERSCRIPT 18 end_POSTSUPERSCRIPT electrons/cm2 was used. The total cross-section to knock out either Rh or S at 2.5 MeV is 76.4 barn, assuming a typical displacement threshold of 30 eV. This gives about one defect per approximately 30 conventional formula units (Z=2𝑍2Z=2italic_Z = 2) with a mean distance between the defects of approximately 3 nm. The resistivity at Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT increased from 12μΩcm12𝜇Ωcm12\>\mu\Omega\cdot\text{cm}12 italic_μ roman_Ω ⋅ cm to approximately 30μΩcm30𝜇Ωcm30\>\mu\Omega\cdot\text{cm}30 italic_μ roman_Ω ⋅ cm, corresponding to a threefold increase in the scattering rate ΓΓ\Gammaroman_Γ, proportional to the concentration of defects. With the above estimate, Γ=0.2Γ0.2\Gamma=0.2roman_Γ = 0.2 in the clean case, and it increases to Γ=0.6Γ0.6\Gamma=0.6roman_Γ = 0.6 upon irradiation that moves closer to the dirty limit of Γ>1Γ1\Gamma>1roman_Γ > 1. The increase in pinning due to additional disorder is evident in the magnetization M(H)𝑀𝐻M(H)italic_M ( italic_H ) loops presented.

Acknowledgements.
This work was supported by the U.S. Department of Energy (DOE), Office of Science, Basic Energy Sciences, Materials Science and Engineering Division. Ames Laboratory is operated for the U.S. DOE by Iowa State University under contract DE-AC02-07CH11358. Electron irradiation was performed on the SIRIUS platform supported by the French National network of accelerators for irradiation and analysis of molecules and materials EMIR&A (FR CNRS 3618).

References

  • Abrikosov [1957] A. A. Abrikosov, On the Magnetic Properties of Superconductors of the Second Group, Sov. Phys. JETP [ZhETF 32, 1442-1450 (1957)] 5, 1174 (1957).
  • Pippard [1969] A. B. Pippard, A possible mechanism for the peak effect in type II superconductors, Philosophical Magazine 19, 217 (1969).
  • Campbell and Evetts [1972] A. M. Campbell and J. E. Evetts, Critical currents in superconductors, Monographs on physics (Taylor & Francis Ltd. (London), 1972).
  • Dew-Huges [1974] D. Dew-Huges, Flux pinning mechanisms in type-II superconductors, Phil. Mag. 30, 293 (1974).
  • Larkin and Ovchinnikov [1979] A. I. Larkin and Y. N. Ovchinnikov, Pinning in type II superconductors, Journal of Low Temperature Physics 34, 409 (1979).
  • 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).
  • Brandt [1995] E. H. Brandt, The flux-line lattice in superconductors, Rep. Prog. Phys. 58, 1465 (1995).
  • Yeshurun et al. [1996] Y. Yeshurun, A. P. Malozemoff, and A. Shaulov, Magnetic relaxation in high-temperature superconductors, Reviews of Modern Physics 68, 911 (1996).
  • Hull [2003] J. R. Hull, Applications of high-temperature superconductors in power technology, Rep. Prog. Phys. 66, 1865 (2003).
  • Tinkham [2004] M. Tinkham, Introduction to Superconductivity, 2nd ed., Dover Books on Physics (Dover Publications, 2004).
  • Kwok et al. [2016] W.-K. Kwok, U. Welp, A. Glatz, A. E. Koshelev, K. J. Kihlstrom, and G. W. Crabtree, Vortices in high-performance high-temperature superconductors, Rep. Prog. Phys. 79, 116501 (2016).
  • Bud’ko and Canfield [2006] S. L. Bud’ko and P. C. Canfield, Magnetism and superconductivity in rare earth-nickel-borocarbides, Comptes Rendus Phys. 7, 56 (2006).
  • Müller and Narozhnyi [2001] K.-H. Müller and V. N. Narozhnyi, Interaction of superconductivity and magnetism in borocarbide superconductors, Rep. Prog. Phys. 64, 943 (2001).
  • Canfield and Bud’ko [2010] P. C. Canfield and S. L. Bud’ko, FeAs-Based Superconductivity: A Case Study of the Effects of Transition Metal Doping on BaFe2As2Ann. Rev. Cond. Matt. Phys. 1, 27 (2010).
  • Wen and Li [2011] H.-H. Wen and S. Li, Materials and Novel Superconductivity in Iron Pnictide Superconductors, Ann. Rev. Cond. Matt. Phys. 2, 121 (2011).
  • Stewart [2011] G. R. Stewart, Superconductivity in iron compounds, Rev. Mod. Phys. 83, 1589 (2011).
  • Chubukov and Hirschfeld [2015] A. Chubukov and P. J. Hirschfeld, Iron-based superconductors, seven years later, Physics Today 68, 46 (2015).
  • DeSorbo [1964] W. DeSorbo, The Peak Effect in Substitutional and Interstitial Solid Solutions of High-Field Superconductors, Rev. Mod. Phys. 36, 90 (1964).
  • Daeumling et al. [1990] M. Daeumling, J. M. Seuntjens, and D. C. Larbalestier, Oxygen-defect flux pinning, anomalous magnetization and intra-grain granularity in YBa2Cu307-δNature 346, 332 (1990).
  • Krusin-Elbaum et al. [1992] L. Krusin-Elbaum, L. Civale, V. M. Vinokur, and F. Holtzberg, “‘Phase diagram”’ of the vortex-solid phase in Y-Ba-Cu-O crystals: A crossover from single-vortex (1D) to collective (3D) pinning regimes, Phys. Rev. Lett. 69, 2280 (1992).
  • Klein et al. [1994] L. Klein, E. R. Yacoby, Y. Yeshurun, A. Erb, G. Müller-Vogt, V. Breit, and H. Wühl, Peak effect and scaling of irreversible properties in untwinned Y-Ba-Cu-O crystals, Phys. Rev. B 49, 4403 (1994).
  • Tang et al. [1996] C. Tang, X. Ling, S. Bhattacharya, and P. M. Chaikin, Peak effect in superconductors: melting of Larkin domains, Europhysics Letters (EPL) 35, 597 (1996).
  • Giller et al. [1997] D. Giller, A. Shaulov, R. Prozorov, Y. Abulafia, Y. Wolfus, L. Burlachkov, Y. Yeshurun, E. Zeldov, V. M. Vinokur, J. L. Peng, and R. L. Greene, Disorder-induced transition to entangled vortex solid in Nd-Ce-Cu-O crystal, Phys. Rev. Lett. 79, 2542 (1997).
  • Banerjee et al. [2000] S. S. Banerjee, S. Ramakrishnan, A. K. Grover, G. Ravikumar, P. K. Mishra, V. C. Sahni, C. V. Tomy, G. Balakrishnan, D. M. Paul, P. L. Gammel, D. J. Bishop, E. Bucher, M. J. Higgins, and S. Bhattacharya, Peak effect, plateau effect, and fishtail anomaly: The reentrant amorphization of vortex matter in 2H-NbSe2, Phys. Rev. B 62, 11838 (2000).
  • Mikitik and Brandt [2001] G. P. Mikitik and E. H. Brandt, Peak effect, vortex-lattice melting line, and order-disorder transition in conventional and high-T-c superconductors, Phys. Rev. B 64, 184514 (2001).
  • Prozorov et al. [2008] R. Prozorov, N. Ni, M. A. Tanatar, V. G. Kogan, R. T. Gordon, C. Martin, E. C. Blomberg, P. Prommapan, J. Q. Yan, S. L. Bud’ko, and P. C. Canfield, Vortex phase diagram of Ba(Fe0.93Co0.07)2As2 single crystals, Phys. Rev. B 78, 224506 (2008).
  • Naren et al. [2009] H. R. Naren, A. Tamizhavel, S. Ramakrishnan, and A. K. Grover, Vortex phase diagram in weakly pinned Rh17S15subscriptRh17subscriptS15\text{Rh}_{17}\text{S}_{15}Rh start_POSTSUBSCRIPT 17 end_POSTSUBSCRIPT S start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPTJournal of Physics: Conference Series 150, 052183 (2009).
  • Gao et al. [2022] L. Gao, T. Ying, Y. Zhao, W. Cao, C. Li, L. Xiong, Q. Wang, C. Pei, J.-Y. Ge, H. Hosono, and Y. Qi, Fishtail effect and the vortex phase diagram of high-entropy alloy superconductor, Appl. Phys. Lett. 120, 092602 (2022).
  • Konczykowski et al. [2000] M. Konczykowski, S. Colson, C. J. van der Beek, M. V. Indenbom, P. H. Kes, and E. Zeldov, Magnetic relaxation in the vicinity of second magnetization peak in BSCCO crystals, Physica C 332, 219 (2000).
  • Abulafia et al. [1996] Y. Abulafia, A. Shaulov, Y. Wolfus, R. Prozorov, L. Burlachkov, Y. Yeshurun, D. Majer, E. Zeldov, H. Wuhl, V. B. Geshkenbein, and V. M. Vinokur, Plastic vortex creep in YBa2Cu3O7-x crystals, Phys. Rev. Lett. 77, 1596 (1996).
  • Koshelev and Larkin [1995] A. E. Koshelev and A. I. Larkin, Paramagnetic moment in field-cooled superconducting plates: Paramagnetic Meissner effect, Phys. Rev. B 52, 13559 (1995).
  • Nusran et al. [2018] N. M. Nusran, K. R. Joshi, K. Cho, M. A. Tanatar, W. R. Meier, S. L. Bud’ko, P. C. Canfield, Y. Liu, T. A. Lograsso, and R. Prozorov, Spatially-resolved study of the Meissner effect in superconductors using NV-centers-in-diamond optical magnetometry, New Journal of Physics 20, 043010 (2018).
  • Koblischka et al. [2023] M. R. Koblischka, L. Půst, C.-S. Chang, T. Hauet, and A. Koblischka-Veneva, The Paramagnetic Meissner Effect (PME) in Metallic Superconductors, Metals 13, 1140 (2023).
  • Kim et al. [2024] H. Kim, M. A. Tanatar, M. Kończykowski, R. Grasset, U. S. Kaluarachchi, S. Teknowijoyo, K. Cho, A. Sapkota, J. M. Wilde, M. J. Krogstad, S. L. Bud’ko, P. M. R. Brydon, P. C. Canfield, and R. Prozorov, Nodal superconductivity in miassite Rh71{}_{1}7start_FLOATSUBSCRIPT 1 end_FLOATSUBSCRIPT 7S51{}_{1}5start_FLOATSUBSCRIPT 1 end_FLOATSUBSCRIPT 5Communications Materials 5, 17 (2024).
  • Nie et al. [2024] J. Y. Nie, C. C. Zhao, C. Q. Xu, B. Li, C. P. Tu, X. Zhang, D. Z. Dai, H. R. Wang, S. Xu, W. Jiao, B. M. Wang, Z. Xu, X. Xu, and S. Y. Li, On the superconducting gap structure of the miassite Rh71{}_{1}7start_FLOATSUBSCRIPT 1 end_FLOATSUBSCRIPT 7S51{}_{1}5start_FLOATSUBSCRIPT 1 end_FLOATSUBSCRIPT 5: Nodal or nodeless?, arXiv:2405.08633  (2024).
  • Helfand and Werthamer [1966] E. Helfand and N. R. Werthamer, Temperature and Purity Dependence of the Superconducting Critical Field, Hc2. II, Phys. Rev. 147, 288 (1966).
  • Settai et al. [2010] R. Settai, K. Katayama, H. Muranaka, T. Takeuchi, A. Thamizhavel, I. Sheikin, and Y. Ōnuki, Superconducting properties in Rh71{}_{1}7start_FLOATSUBSCRIPT 1 end_FLOATSUBSCRIPT 7S51{}_{1}5start_FLOATSUBSCRIPT 1 end_FLOATSUBSCRIPT 5 under magnetic field and pressure, J. Phys. Chem. Sol. 71, 700 (2010).
  • Naren et al. [2011a] R. H. Naren, A. Tamizhavel, and R. S, Superconductivity in Rh17S15subscriptRh17subscriptS15\text{Rh}_{17}\text{S}_{15}Rh start_POSTSUBSCRIPT 17 end_POSTSUBSCRIPT S start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT and Pd17Se15subscriptPd17subscriptSe15\text{Pd}_{17}\text{Se}_{15}Pd start_POSTSUBSCRIPT 17 end_POSTSUBSCRIPT Se start_POSTSUBSCRIPT 15 end_POSTSUBSCRIPT: A comparative study, Journal of Physics: Conference Series 273, 012074 (2011a).
  • Naren et al. [2011b] R. H. Naren, A. Thamizhavel, K. A. Nigam, and S. Ramakrishan, Superconductivity in rh17s15and pd17se15: A comparative study, J. Phys. Soc. Jpn. 80, SA014 (2011b).
  • Bardeen et al. [1957] J. Bardeen, L. N. Cooper, and J. R. Schrieffer, Microscopic Theory of Superconductivity, Physical Review 106, 162 (1957).
  • Werthamer et al. [1966] N. R. Werthamer, E. Helfand, and P. C. Hohenberg, Temperature and Purity Dependence of the Superconducting Critical Field, Hc2subscript𝐻𝑐2H_{c2}italic_H start_POSTSUBSCRIPT italic_c 2 end_POSTSUBSCRIPT. III. Electron Spin and Spin-Orbit Effects, Phys. Rev. 147, 295 (1966).
  • Kogan and Prozorov [2012] V. G. Kogan and R. Prozorov, Orbital upper critical field and its anisotropy of clean one- and two-band superconductors, Rep. Prog. Phys. 75, 114502 (2012).
  • Benfatto et al. [2002] L. Benfatto, A. Toschi, S. Caprara, and C. Castellani, Coherence length in superconductors from weak to strong coupling, Phys. Rev. B 66, 054515 (2002).
  • Zarea et al. [2023] M. Zarea, H. Ueki, and J. A. Sauls, Effects of anisotropy and disorder on the superconducting properties of niobium, Frontiers in Physics 11, :1269872 (2023).
  • Butler [1980] W. H. Butler, Upper Critical Field of Nb: Calculated Temperature Dependence and Anisotropy, Phys. Rev. Lett. 44, 1516 (1980).
  • Labbe [1967] J. Labbe, Paramagnetic susceptibility in the V3SisubscriptV3Si\text{V}_{3}\text{Si}V start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT Si type of compounds in the normal state, Phys. Rev. 158, 647 (1967).
  • Prozorov and Kogan [2018] R. Prozorov and V. G. Kogan, Effective demagnetizing factors of diamagnetic samples of various shapes, Phys. Rev. Applied 10, 014030 (2018).
  • Prozorov et al. [2022] R. Prozorov, M. Zarea, and J. A. Sauls, Niobium in the clean limit: An intrinsic type-I superconductor, Phys. Rev. B 106, L180505 (2022).
  • Prozorov and Kogan [2011] R. Prozorov and V. G. Kogan, London penetration depth in iron-based superconductors, Rep. Prog. Phys. 74, 124505 (2011).
  • Koshelev et al. [2019] A. E. Koshelev, K. Willa, R. Willa, M. P. Smylie, J.-K. Bao, D. Y. Chung, M. G. Kanatzidis, W.-K. Kwok, and U. Welp, Melting of vortex lattice in the magnetic superconductor RbEuFe4As4subscriptRbEuFe4subscriptAs4\text{RbEuFe}_{4}\text{As}_{4}RbEuFe start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT As start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPTPhys. Rev. B 100, 094518 (2019).
  • Canfield and Slade [2023] P. C. Canfield and T. J. Slade, Single Crystal Growth of Synthetic Sulfide- and Phosphide-Based Minerals for Physical Measurements, Minerals 13, 429 (2023).
  • Canfield et al. [2016] P. C. Canfield, T. Kong, U. S. Kaluarachchi, and N. H. Jo, Use of frit-disc crucibles for routine and exploratory solution growth of single crystalline samples, Philos. Mag. 96, 84 (2016).
  • Canfield [2019] P. C. Canfield, New materials physics, Rep. Prog. Phys. 83, 016501 (2019).
  • Damask and Dienes [1963] A. Damask and G. Dienes, Point Defects in Metals (Gordon and Breach, 1963).
  • Thompson [1969] M. Thompson, Defects and Radiation Damage in Metals, Cambridge Monographs on Physics (Cambridge University Press, 1969).
  • Cho et al. [2018] K. Cho, M. Kończykowski, S. Teknowijoyo, M. A. Tanatar, and R. Prozorov, Using electron irradiation to probe iron-based superconductors, Supercond. Sci. Technol. 31, 064002 (2018).