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

thanks: correspondence should be addressed to javier.landaeta@cpfs.mpg.de or elena.hassinger@cpfs.mpg.dethanks: correspondence should be addressed to javier.landaeta@cpfs.mpg.de or elena.hassinger@cpfs.mpg.de

Field-angle dependence reveals odd-parity superconductivity in CeRh2As2

J. F. Landaeta Max Planck Institute for Chemical Physics of Solids, 01187 Dresden, Germany    P. Khanenko Max Planck Institute for Chemical Physics of Solids, 01187 Dresden, Germany    D.C. Cavanagh Department of Physics, University of Otago, P.O. Box 56, Dunedin 9054, New Zealand    C. Geibel Max Planck Institute for Chemical Physics of Solids, 01187 Dresden, Germany    S. Khim Max Planck Institute for Chemical Physics of Solids, 01187 Dresden, Germany    S. Mishra Laboratoire National des Champs Magnétiques Intenses (LNCMI-EMFL), CNRS, UGA, 38042 Grenoble, France    I. Sheikin Laboratoire National des Champs Magnétiques Intenses (LNCMI-EMFL), CNRS, UGA, 38042 Grenoble, France    P.M.R. Brydon Department of Physics and MacDiarmid Institute for Advanced Materials and Nanotechnology, University of Otago, P.O. Box 56, Dunedin 9054, New Zealand    D.F. Agterberg Department of Physics, University of Wisconsin-Milwaukee, Milwaukee, Wisconsin 53201, USA    M. Brando Max Planck Institute for Chemical Physics of Solids, 01187 Dresden, Germany    E. Hassinger Max Planck Institute for Chemical Physics of Solids, 01187 Dresden, Germany Technical University Munich, Physics department, 85748 Garching, Germany
(March 15, 2025)
Abstract

CeRh2As2 is an unconventional superconductor with multiple superconducting phases and Tc=0.26T_{\mathrm{c}}=0.26 K. When HcH\parallel c, it shows a field-induced transition at μ0H=4\mu_{0}H^{*}=4 T from a low-field superconducting state SC1 to a high-field state SC2 with a large critical field of μ0Hc2=14\mu_{0}H_{\mathrm{c2}}=14 T. In contrast, for HcH\perp c, only the SC1 with μ0Hc2=2\mu_{0}H_{\mathrm{c2}}=2 T is observed. A simple model based on the crystal symmetry was able to reproduce the phase-diagrams and their anisotropy, identifying SC1 and SC2 with even and odd parity superconducting states, respectively. However, additional orders were observed in the normal state which might have an influence on the change of the superconducting state at HH^{*}. Here, we present a comprehensive study of the angle dependence of the upper critical fields using magnetic ac-susceptibility, specific heat and torque on single crystals of CeRh2As2. The experiments show that the state SC2 is strongly suppressed when rotating the magnetic field away from the cc axis and it disappears for an angle of 35. This behavior agrees perfectly with our extended model of a pseudospin triplet state with d\vec{d} vector in the plane and hence allows to nail down that SC2 is indeed the suggested odd-parity state.

I Introduction

Odd-parity, or (pseudo-) spin triplet superconductivity is a rare phenomenon in nature. Only a few materials are promising candidates, including UPt3 Fisher et al. (1989), the ferromagnetic superconductors UCoGe, URhGe, UGe2Aoki et al. (2019), and UTe2 Ran et al. (2019). In these systems, important information on the superconducting (SC) state has come from an investigation of the angle dependence of the critical fields. Theoretically, when spin-orbit coupling is strong, the angle dependence of the Pauli-limiting field HpH_{p} offers a way to identify possible triplet states since HpH_{p} should depend on the orientation of the applied magnetic field with respect to the direction of the spins of the Cooper pairs Sigrist (2005); Machida et al. (1985); Mineev and Samokhin (1999). In experiment however, the observed angle dependencies of the upper critical fields in the above mentioned systems seem to be dominated by the orbital limit or by the interplay of superconductivity and field-enhanced spin fluctuations associated with an Ising quantum critical point Aoki et al. (2019); Tada et al. (2011).

Recently, CeRh2As2 was discovered to have unique and highly anisotropic superconducting critical field phase diagrams (Fig. 4a,f), with a suggested field-induced odd-parity state Khim et al. (2021). The aim of this study is to investigate its angle dependence in detail in order to find out more about the superconductivity itself and its relation with the normal state.

CeRh2As2 is a heavy-fermion compound with an electronic specific heat coefficient close to 1 J/molK2 at 0.5 K. At T0 0.4T_{0}\approx\leavevmode\nobreak\ 0.4 K, a phase transition to a suggested quadrupole-density-wave (QDW) state occurs Hafner et al. (2022). At Tc=0.26T_{\mathrm{c}}=0.26 K, CeRh2As2 enters a low-field superconducting state (SC1). A magnetic field applied along the cc axis induces a transition at μ0H4\mu_{0}H^{*}\approx 4 T into a second high-field superconducting state SC2, with an upper critical field Hc2=14H_{c2}=14 T. For in-plane fields, only the low-field state SC1 appears, with Hc2=2H_{c2}=2 T, whereas T0T_{0} increases with field so that the SC1 phase is found to be always included in the QDW phase for this field direction. Furthermore, in zero field within the SC state a broadening of the As(2) nuclear quadrupole resonance (NQR) line was interpreted as the onset of antiferromagnetic (AFM) order Kibune et al. (2022). However, no additional anomaly was detected in the bulk measurements such as specific heat and thermal expansion Khim et al. (2021); Hafner et al. (2022).

The crystal structure of CeRh2As2 is centrosymmetric, but the inversion symmetry is broken locally at the Ce sites enabling a Rashba spin orbit coupling with alternating sign on neighboring Ce layers Khim et al. (2021). Assuming dominant intralayer singlet SC pairing, a cc-axis field transforms an even-parity superconducting state with equal gap sign on both Ce layers into an odd-parity one with alternating gap sign Yoshida et al. (2012); Khim et al. (2021). The latter may be topological Skurativska et al. (2021); Nogaki et al. (2021) and can be described as a pseudospin triplet state with d\vec{d} vector in the plane leading to the absence of a Pauli limit for HcH\parallel c Khim et al. (2021); Yoshida et al. (2014).

The unusual results for the field along the cc axis may also have another origin. One possibility emerged from the observation that the QDW state below T00.4T_{0}\approx 0.4 K is suppressed by a cc-axis field of similar strength as HH^{*} Khanenko et al. ; Khim et al. (2021); Hafner et al. (2022). So, the SC1 phase is found to be inside the QDW phase as for in-plane fields, but the SC2 phase would be outside the QDW phase. The presence of the QDW state might lead to a drastic change of the superconducting properties and in particular of Hc2H_{c2} between the SC1 and SC2 states Khim et al. (2021). Ignoring the QDW and AFM phases, other possibilities have been suggested to explain the origin of the SC1 and SC2 phases as a change between different superconducting order parameters  Skurativska et al. (2021); Möckli and Ramires (2021a, b). There are scenarios which involve a field driven Fermi-surface Lifshitz transition Ptok et al. (2021) or the presence of a field-induced quantum critical point (QCP) Aoki et al. (2019); Tada et al. (2011). Eventually, because the normal state is also anisotropic, a study of the angle dependence of the superconducting states is a promising way to distinguish between those scenarios.

Hence, we study the superconducting phase diagram as a function of magnetic field direction by ac susceptibility, magnetic torque and specific heat. We find that the angle dependence supports the picture that SC2 is an odd parity state with pseudo-spin d\vec{d} vector in the plane. CeRh2As2 seems to be the first compound where the anisotropy of the Pauli limiting can be used to reveal triplet superconductivity.

II Results

In the following, we show results of ac susceptibility χac\chi_{ac}, magnetic torque τ\tau and specific heat CC as a function of temperature and magnetic field in different directions with an angle θ\theta away from the cc axis. These probes were chosen because they are sensitive to the bulk properties of the material and they can detect the transition inside the superconducting state Khim et al. (2021). Previously it was found that the TcT_{\mathrm{c}} from resistivity is higher than the bulk TcT_{\mathrm{c}} from specific heat or low-frequency ac-susceptibility Khim et al. (2021). It was suggested that this is due to percolating strain-induced superconductivity around impurities. Experimentally, the resistive TcT_{\mathrm{c}} follows the bulk TcT_{\mathrm{c}} in a parallel fashion for both HcH\parallel c and HaH\parallel a Hafner et al. (2022); Onishi et al. (2022). While the angle dependence of the superconducting state could have been investigated by resistivity as well, the transition at HH^{*} is invisible to this probe Khim et al. (2021); Hafner et al. (2022); Onishi et al. (2022).

Figure 1(a) shows the magnetic field dependence of χac\chi_{ac} at the temperature of 45 mK for different angles. Details on the experimental methods can be found in Landaeta (2021). We measured from θ=0\theta=0^{\circ} (HcH\parallel c) to 9090^{\circ} (HcH\perp c here named HabH\parallel ab). The in-plane field orientation was [110] for torque but not defined for χac\chi_{ac} and C/TC/T. A small but clear signature of the transition between SC1 and SC2 named HH^{*} appears for angles below 35. At 30 a small anomaly is visible at HH^{*} in the derivative dχac/dHd\chi_{ac}/dH highlighted in the inset of Figure 1(a).

Refer to caption
Figure 1: Magnetic field dependence of magnetic susceptibility and torque for different field directions. (a) Magnitude of the ac susceptibility χac\chi_{ac} at 45 mK for different angles as indicated. θ=0\theta=0^{\circ} corresponds to μ0Hc\mu_{0}H\parallel c. The bump labelled as HH^{*}, is the transition between SC1 and SC2. The inset shows the derivative of the curve at 30, indicating that HH^{*} is still present at that angle. The dashed horizontal line indicates the value at which Hc2H_{\mathrm{c2}} is defined. (b) Magnetic torque divided by H2sin(2θ)H^{2}\sin(2\theta) at 40 mK. The inset shows the torque for 2.52.5^{\circ} and 4.64.6^{\circ} where the change in the hysteresis of τ\tau is visible. Additionally, the red line gives the expected B2B^{2}-dependence. (c) Torque divided by H2sin(2θ)H^{2}\sin(2\theta) for angles close to the plane. The inset in (c) shows the additional phase transition appearing at Hcr9H_{\mathrm{cr}}\approx 9 T. In (b) and (c) curves are shifted for clarity.

We define the upper critical field Hc2H_{\mathrm{c2}} at the onset of the diamagnetic transition in order to be consistent with the previous analysis Khim et al. (2021) (horizontal dashed line in Figure 1(a)) and HH^{*} at the maximum of the small bump in the field dependence of χac\chi_{ac}. The torque τ\tau is shown in Figure 1(b,c). As expected, its field dependence depicted in the inset of panel (b) is quadratic in field in the normal state. Since we are interested in deviations from this behavior in the superconducting state, we present data as τ/H2\tau/H^{2} and scaled by sin(2θ)1\sin{(2\theta)}^{-1} which is the standard angle dependence of torque. It displays a similar step-like feature at HH^{*}. The strong hysteresis loop in the superconducting state was already observed in the magnetization in CeRh2As2 Khim et al. (2021); Landaeta (2021). While the hysteresis is counter-clockwise at small angles (Inset of Fig. 1b), as in the magnetization, it changes to a clockwise one at higher angles (Fig. 1b). This is related to a change in the magnetization anisotropy (see Landaeta (2021)). Hc2H_{\mathrm{c2}} is defined in the torque at the field where up and down sweep curves start to separate. Since this happens rather smoothly for small angles, there is a larger uncertainty on these points. Strictly speaking, this is the so-called "irreversibility field". As usually observed, it is slightly lower than Hc2H_{\mathrm{c2}} from χac\chi_{ac}, but it follows its angle dependence.

For fields near the in-plane direction, an additional transition is observed at HcrH_{\mathrm{cr}}\approx 9 T (Fig. 1c and Landaeta (2021)), as observed before Hafner et al. (2022), where it was associated with a change of the order below T0T_{0}. HcrH_{\mathrm{cr}} increases when the field is turned away from the abab plane towards the cc axis until, for angles below 58, this anomaly cannot be resolved experimentally any more. As can be seen in the inset, it is strongly hysteretic indicating a first order transition, in agreement with results from resistivity and magnetostriction Hafner et al. (2022). No further transitions have been detected for in-plane fields up to 36 T and cc-axis fields up to 26 T Landaeta (2021).

Refer to caption
Figure 2: Angle dependence of the critical fields at 45 mK. (a) different symbols represent different experimental techniques as indicated. HcrH_{\mathrm{cr}} indicates a transition between two ordered states I and II Hafner et al. (2022). Since it shows a strong hysteresis, only the critical fields from up sweeps are shown here. The critical fields from the down sweeps are typically around 1 T lower. (b) Angle dependence of the upper critical fields of SC1 and SC2 as well as the orbital limit HorbH_{\mathrm{orb}} of SC1 obtained from the phase diagrams in Fig. 4 and a fit of it using equation 1 (orange dashed line). Red and black lines are fits to equation 2 at 45 mK and the violet line reflects the calculated HH^{*}, as detailed in the main text and Landaeta (2021).

From these data we find the angle dependence of Hc2H_{\mathrm{c2}}, HH^{*} and HcrH_{\mathrm{cr}} shown in Figure 2(a). Here, the upper critical field Hc2H_{\mathrm{c2}} decreases rapidly as the angle is increased with respect to the cc axis until it approaches the value of HH^{*} at about 35. From this point on, we only detect one superconducting state in which the Hc2H_{\mathrm{c2}} slowly decreases with angle until it reaches the expected value of 2\approx 2 T for HabH\parallel ab.

Refer to caption
Figure 3: Temperature dependence of the specific heat C/TC/T in magnetic field at different angles. Here, a nuclear contribution has been removed Landaeta (2021).

For different magnetic field angles, temperature dependent ac-susceptibility (data in Landaeta (2021)) and specific heat were measured in magnetic field. A selection of the specific heat data are depicted in Fig. 3, where a nuclear contribution has been removed. In this paper, these data are used to extract the superconducting transition temperatures and a qualitative behavior of T0T_{0} (see below). A more detailed analysis of the specific heat and the angle dependence of T0T_{0} is subject of future studies.

Refer to caption
Figure 4: Magnetic field-temperature phase diagrams of CeRh2As2 for different directions of the magnetic field. The experimental points are from ac susceptibility and specific heat, as indicated. The data show in (a) and (f) were taken from Khim et al. (2021).

The superconducting critical fields phase diagrams are shown in Figure 4 where data from the temperature sweeps in Fig. 3 and from Landaeta (2021) and from the field sweeps in Fig. 1 are included. Note that the discrepancy between the critical field from specific heat and ac-susceptibility at 2020^{\circ} is ascribed to a difference in angle within the uncertainty because in this angle region, Hc2H_{\mathrm{c2}} changes strongly even with only a few degrees. For all angles below 35, a kink appears in Hc2(T)H_{\mathrm{c2}}(T) at a field that is close to HH^{*} obtained from the field-sweeps revealing that the HH^{*} line is almost temperature independent for all angles. As already observed for the high-symmetry directions Khim et al. (2021), the phase diagrams determined by ac-susceptibility are fully consistent with those obtained from the specific heat and hence reveal the bulk properties.

The present torque and specific heat data also provide some new information on the phase diagram of the phase connected with T0T_{0} which was identified with a QDW order Khim et al. (2021); Hafner et al. (2022). An order of magnetic dipoles was ruled out since magnetic probes like ac-susceptibility or magnetization do not display any signature at this transition Khim et al. (2021). On the other hand, the small hump in the specific heat and a clear jump in the thermal expansion reveal a bulk phase transition. The electrical resistivity shows an increase below T0T_{0} possibly due to nesting. This indicates that itinerant ff-electrons forming the bands at the Fermi energy are involved in the T0T_{0} order. The field dependence of T0T_{0} is extremely anisotropic. For in-plane fields, an increase of T0T_{0} is observed Khim et al. (2021); Hafner et al. (2022). Since such an increase is typical for local quadrupolar order in cubic systems, it was one of the signs that the T0T_{0} order involves quadrupolar degrees of freedom of the itinerant electrons that are induced by a Kondo mixing of the crystal electric field doublets. This idea was supported by calculations of the quadrupole moment for such a system Hafner et al. (2022). Furthermore, at μ0Hcr=9\mu_{0}H_{\mathrm{cr}}=9 T, a first-order transition to another state occurs. In contrast, for cc-axis fields, the T0T_{0} order is suppressed at roughly H0=(4±1)H_{0}=(4\pm 1) T, seemingly coincident with the transition inside the superconducting state Khim et al. (2021). Both H0H_{0} and HcrH_{\mathrm{cr}} present a clear angle dependence. For 1010^{\circ} and 2020^{\circ}, the hump at T0T_{0} is suppressed with field, but remains visible up to roughly 4 T (Figure 3a,b), similar to what was observed for HcH\parallel c Khim et al. (2021). For 4545^{\circ} T0T_{0} stays approximately constant in applied fields up to 10 T (Figure 3c), far above the superconducting critical field, similar to the behavior observed for HabH\parallel ab Khim et al. (2021); Hafner et al. (2022). Therefore, H0H_{0} stays roughly constant at least up to 2020^{\circ} and then starts increasing strongly for angles between 2020^{\circ} and 4545^{\circ}. Furthermore, because HcrH_{\mathrm{cr}} shoots up for angles from 9090^{\circ} to 5858^{\circ}, it might meet the H0H_{0} line in a tricritical point in the range 30<θ<6030^{\circ}<\theta<60^{\circ} and μ0H>15\mu_{0}H>15 T. Another possibility is that the HcrH_{\mathrm{cr}} transition ends in a critical end point, since the anomaly just disappears going from 5858^{\circ} to 5050^{\circ}.
The absence of any further anomaly in the torque up to μ0H=36\mu_{0}H=36 T for θ=89.4\theta=89.4^{\circ} (HaH\parallel a) Landaeta (2021) indicates that the phase forming at H>HcrH>H_{\mathrm{cr}} is stable until very high field as observed e.g. for the quadrupolar phase II in CeB6 in high magnetic fields Effantin et al. (1985). Further high-field studies are necessary to determine the H0H_{0} boundary.
Notably, the angle dependence of H0H_{0} and HcrH_{\mathrm{cr}} are quite different from those of Hc2H_{\mathrm{c2}}, suggesting the interaction between this phase and superconductivity to be weak. On the other hand, for low angles, H0H_{0} nearly coincides with HH^{*} up to 2020^{\circ}.

Let us now concentrate on the angle dependence of the critical field of the state SC1. Previous analysis along the two crystallographic directions showed that the orbital limit HorbH_{\mathrm{orb}} in CeRh2As2 is large and strongly exceeds the experimental critical fields of SC1. Therefore, this state is strongly Pauli limited with enhanced Pauli fields compared to the bare weak coupling value of μ0Hp=1.84Tc0.5\mu_{0}H_{p}=1.84\,T_{\mathrm{c}}\approx 0.5 T Clogston (1962); Chandrasekhar (1962). We now proceed with the determination of the angle dependence of both the orbital limit HorbH_{\mathrm{orb}} and the Pauli limit HpH_{p}. In the clean-limit HorbH_{\mathrm{orb}} is given by the slope of the critical field Hc2(T)H_{\mathrm{c2}}(T) near TcT_{\mathrm{c}} as μ0Horb(T=0)=0.73TcdHc2dT|T=Tc\mu_{0}H_{\mathrm{orb}}(T=0)=-0.73T_{\mathrm{c}}\left.\frac{dH_{\mathrm{c2}}}{dT}\right|_{T=T_{\mathrm{c}}}Helfand and Werthamer (1966).

From the data in Fig. 4 we extract the values of the Hc2H_{\mathrm{c2}} slopes near TcT_{\mathrm{c}} Landaeta (2021) and hence HorbH_{\mathrm{orb}} for all angles, which are given in Figure 2b. We observe that TcT_{\mathrm{c}} varies slightly even in zero field for different sample orientations, which is probably due to the remanent magnetic field in the superconducting magnet and slightly different shapes of the ac susceptibility curves at different angles near the onset of the transition where we define TcT_{\mathrm{c}}. The values of TcT_{\mathrm{c}} are listed in Landaeta (2021). Note that the uncertainty of the slope and hence HorbH_{\mathrm{orb}} at each angle is quite large (error bars in Figure 2b).

The angle dependence of HorbH_{\mathrm{orb}} behaves as expected for an anisotropic bulk superconductor Tinkham (1996) according to

Hc2(θ)=Hc2cΓ2sinθ2+cosθ2H_{\mathrm{c2}}(\theta)=\frac{H_{\mathrm{c2}}^{c}}{\sqrt{\Gamma^{2}\sin{\theta}^{2}+\cos{\theta}^{2}}} (1)

where the anisotropy parameter Γ=Hc2c/Hc2ab\Gamma=H_{\mathrm{c2}}^{c}/H_{\mathrm{c2}}^{ab}. We find Γ=Horbc/Horbab=2.6\Gamma=H_{\mathrm{orb}}^{c}/H_{\mathrm{orb}}^{ab}=2.6 and μ0Horbc=16.2\mu_{0}H_{\mathrm{orb}}^{c}=16.2 T as shown by the dashed orange line in Figure 2b. For a fully gapped 3D superconductor in the Bardeen Cooper Schrieffer theory, this angle dependence reflects the anisotropy of the normal state since μ0Horb=Φ0/2πξ2(T)\mu_{0}H_{\mathrm{orb}}=\Phi_{0}/2\pi\xi^{2}(T) and ξ(0)=0.18vF/kBTc\xi(0)=0.18\hbar v_{\mathrm{F}}/k_{\mathrm{B}}T_{\mathrm{c}}. Using vF=kF/mv_{\mathrm{F}}=\hbar k_{\mathrm{F}}/m^{\ast} we find Horbc/Horbab=ma2/mcmama/mc=2.6H_{\mathrm{orb}}^{c}/H_{\mathrm{orb}}^{ab}=m_{a}^{\ast 2}/m_{c}^{\ast}m_{a}^{\ast}\Rightarrow m_{a}^{\ast}/m_{c}^{\ast}=2.6. This is comparable to the weak anisotropy in the magnetization at 2 K, where the in-plane value is a factor of 2 larger than the cc-axis value Khim et al. (2021). It will be interesting to see if band-structure calculations confirm this anisotropy in the future Hafner et al. (2022).

As a next step, the Pauli paramagnetic limit of SC1 is investigated. Here we consider the following expression for the upper critical field of a spin-singlet superconductor with both orbital and Pauli limiting

ln(t)=0𝑑u[1Fθ+Fθcos(HgθuHPt)]exp(Hu22Horbt2)1sinhu\ln(t)=\int\limits_{0}^{\infty}du\left\langle\frac{[1-F_{\theta}+F_{\theta}\cos(\frac{Hg_{\theta}u}{H_{P}t})]\exp(\frac{-Hu^{2}}{\sqrt{2}H_{\mathrm{orb}}t^{2}})-1}{\sinh u}\right\rangle (2)

where gθg_{\theta} is a field-angle dependent gg-factor, t=T/Tct=T/T_{\mathrm{c}}, and Fθ=1F_{\theta}=1 quantifies the pair-breaking due to Pauli limiting (as discussed later, this takes a different form for a pseudospin-triplet order parameter). \langle...\rangle indicates an angular average around the Fermi surface. Using the HorbH_{\mathrm{orb}} values from the fit of its angle dependence (in order to reduce the uncertainty at each angle), we find that HP/gθH_{\mathrm{P}}/g_{\theta} exhibits an anisotropy similar to Hc2(θ)H_{\mathrm{c2}}(\theta), that has the angular-dependence given by Eq. 1 with anisotropy parameter Γg=2.8\Gamma_{g}=2.8. For more details about the fitting procedure please refer to Landaeta (2021). Within error bars, there is no temperature dependence of the Hc2H_{\mathrm{c2}} anisotropy for SC1.

Now let us turn to the critical field of SC2 with the aim to understand what is causing the strong suppression of Hc2H_{\mathrm{c2}} when fields are turned away from the cc axis and the disappearance of SC2 at an angle of 3535^{\circ}. For HcH\parallel c, the temperature dependence of Hc2H_{\mathrm{c2}} follows qualitatively - and even quantitatively, when a lower TcT_{\mathrm{c}} is assumed - very well the expectations for a pure orbital limit Khim et al. (2021) without Pauli limiting effect Landaeta (2021).

To first order, the orbital limit of SC2 should be similar to the one of SC1 given in Fig. 2bfoo . Therefore, we can make the first statement that the angle dependence of the orbital limit is not strong enough to describe the steep decrease of the experimental critical field Hc2H_{\mathrm{c2}} with angle. Hence, the decrease must be related to the Pauli limiting kicking in when the field is tilted towards the abab plane. As discussed in the context of SC1, one place where the Pauli field introduces anisotropy is through the spin-orbit coupling renormalized gg-factor gθg_{\theta}. However, this anisotropy is also too small to account for the steep decrease of the experimental critical field Hc2H_{\mathrm{c2}} with angle since any renormalized HpH_{p} for a spin-singlet state should have the same anisotropy as the one observed in SC1. Consequently, there must be another source of anisotropy due to Pauli limiting. Indeed, this is exactly the behavior expected for triplet superconductors with d\vec{d} vector in the plane (helical state): an absence of Pauli limiting for HcH\parallel c (infinite HpcH_{p}^{c}) and presence of Pauli limiting for HabH\parallel ab which is isotropic in the plane. The latter is enhanced compared to the bare value Frigeri et al. (2004); Khim et al. (2021), when Rashba spin-orbit coupling is included. Here we consider such a helical triplet state subject to an orbital critical field, for which the critical field is given by the expression of Eq. 2 but now with Fθ=|d^h^θ|2F_{\theta}=|\hat{d}\cdot\hat{h}_{\theta}|^{2}, where d^\hat{d} is d/|d|\vec{d}/|\vec{d}| and h^θ\hat{h}_{\theta} is a unit vector that gives the direction of the Zeeman field projected onto to the pseudospin basis Khim et al. (2021). As long as the orbital field energy (gθμBHorbg_{\theta}\mu_{B}H_{\mathrm{orb}}) is smaller than the spin-orbit coupling energy, this is the expression for the odd-parity spin-singlet state in which the order parameter has opposite sign on the two inequivalent Ce-layers. In Fig. 2b, we show the calculated critical field for the state SC2. In the SM, we provide plots of the calculated HTH-T phase diagrams for the different field angles. The agreement with experiment is excellent. Note that even the angle above which only one SC phase occurs is reproduced.

This leads us to the main conclusion of our paper: The angle dependence of the superconducting critical field is dominated by the huge anisotropy of the Pauli field of an odd parity state with d-vector in the plane, in agreement with the interpretation in Khim et al. (2021) and in previous models of locally non-centrosymmetric SC Yoshida et al. (2014), although here the SC state is a pseudospin-triplet state with staggered dominant singlet intra-layer pairing.

As a final point, we have also calculated HH^{*}, the first-order transition between the SC1 and SC2 states. In the previous study for HcH\parallel c, the HH^{*} transition was found to be a result of a competition between the Pauli-limited SC1 (with Pauli field HP,1H_{\mathrm{P,1}}) and the Pauli-limit free SC2, so that HHP,1H^{*}\lessapprox H_{\mathrm{P,1}} Khim et al. (2021). Here, we calculate the phase transition between SC1 and SC2 for all magnetic field directions from the free energies taking only the Pauli limiting effect into account (violet line in Fig. 2b), with no additional parameters beyond those calculated from the fitting procedure. Neglecting the orbital limit leads to a slight overestimate of the critical field Schertenleib et al. (2021), but the angular dependence is in perfect agreement with experiment Landaeta (2021).

III Discussion

We would like to emphasize that the simple model established in Khim et al. (2021) and extended here to intermediate angles can reproduce all experimental observations based on only 3 free parameters that cannot be measured experimentally: the critical temperature of the superconducting state SC2 Tc,2T_{c,2}, the in-plane Pauli limit HP,1H_{\mathrm{P,1}} and the strength of the Rashba spin-orbit coupling relative to the interlayer hopping α~\tilde{\alpha}. This model has the minimal number of bands in this crystal symmetry and naturally contains both even and odd parity superconducting states. Even though CeRh2As2 certainly contains multiple bands Hafner et al. (2022), renormalized density functional theory calculations reveal the bulk of the density of states to be on symmetry-related Fermi surfaces near the zone boundary - which justifies the single-band approach used here Hafner et al. (2022); Cavanagh et al. (2022).

The angle dependence of the orbital limit of SC1 corroborates a rather 3D Fermi surface in agreement with the strongly warped cylinders calculated with renormalised band-structure calculations Hafner et al. (2022). Both the anisotropy of the effective mass given by the anisotropy of HorbH_{\mathrm{orb}} as well as the anisotropy of the g factor (and the magnetic susceptibility) are rather small. Furthermore, in quasi 2D systems such as Sr2RuO4Kittaka et al. (2009), CeCoIn5 Ikeda et al. (2001) or FeSe Farrar et al. (2020), the critical field is usually larger for in-plane fields than for cc-axis fields, because orbital motion perpendicular to the layers is hindered and almost no orbital limiting occurs with large HorbH_{\mathrm{orb}} for in-plane fields. In the 2D limit a cusp is expected for HorbH_{\mathrm{orb}} for fields close to the aa axis as observed for example in FeSe, K2Cr3As3 or in superlattices of CeCoIn5/YbCoIn5 Tinkham (1996); Farrar et al. (2020); Zuo et al. (2017); Goh et al. (2012), but not observed here. A previous theoretical study found that going from a quasi two dimensional Fermi surface to a three dimensional Fermi surface reduces the anisotropy of SC1 due to Rashba spin-orbit coupling, but shouldn’t affect qualitatively the anisotropy of SC2 for the suggested scenario here Skurativska et al. (2021). However, they investigated a Fermi surface at the Brillouin zone center and the situation may change for a Fermi surface at the zone boundary where – due to the non-symmorphic structure and symmetry-imposed degeneracies – large values of the Rashba strength over interlayer coupling are expected Cavanagh et al. (2022).

In ferromagnetic superconductors, which are the most prominent candidates of spin-triplet superconductivity, the critical fields are highly anisotropic Aoki et al. (2009, 2019), and the strong enhancements observed along certain directions or in the field-reentrant phases are related with strong Ising-type spin fluctuations and quantum criticality influencing the orbital limit Aoki et al. (2019); Tada et al. (2011). Here we find that this is not the case for CeRh2As2, where the anisotropy is accounted for by a pseudospin d\vec{d} that is oriented in the basal plane. An interesting open issue is why the orbital critical field and the Sommerfeld coefficient are so large for CeRh2As2. In the non-centrosymmetric superconductors CePt3Si and CeRhSi3 the origin of this was proposed to be antiferromagnetic quantum critical fluctuations that enhance the electron effective mass Kimura et al. (2005, 2007); Sugitani et al. (2006); Bauer and Sigrist (2012); Tada et al. (2011). This needs to be investigated further in CeRh2As2, where quadrupolar as well as antiferromagnetic degrees of freedom are suggested to play a role Hafner et al. (2022); Kibune et al. (2022); Kitagawa et al. (2022). CeIrSi3, CeRhSi3 and CeCoGe3 show a similar anisotropy of the superconducting state to CeRh2As2 with absence of Pauli limit for HcH\parallel c and strong but enhanced Pauli limit for HabH\parallel ab Kimura et al. (2007); Settai et al. (2008); Méasson et al. (2009), but a full determination of the angle dependence hasn’t been possible yet because the large anisotropy appears only under pressure, preventing an easy angle-dependent measurement.

At this point, we would like to discuss the possibility that the transition between the two superconducting states originates from the suppression of the T0T_{0} order Hafner et al. (2022); Khanenko et al. or of the AFM order Kibune et al. (2022) by a field along the cc axis. In the former scenario, the suppression of the T0T_{0} order is considered to be responsible for the HH^{*} transition line and the superconducting order parameter remains the same in the SC1 and SC2 states, i.e. below and above HH^{*}. Our results indicate that whenever the superconducting state coexists with the T0T_{0} order, it is Pauli limited, but when T0T_{0} is suppressed for fields larger than μ0H4\mu_{0}H^{*}\approx 4 T, the superconducting state is not anymore, or much less, Pauli limited. Knowing that the T0T_{0} order probably causes a nesting, i.e. partial gapping of the Fermi surface, it seems natural that it might affect the superconducting state. Accordingly, in order to understand the observed anisotropy of the superconducting state in this paper, the T0T_{0} order would at least have to suppress the spin-orbit coupling, for example by gapping out parts of the Fermi surface with large spin-orbit coupling. However, this seems difficult to explain microscopically, as it would imply that the bare spin-orbit coupling is significantly larger than the already substantial spin-orbit coupling used to fit the critical field data in Fig. 2. In the latter scenario, the AFM order is considered to be responsible for the HH^{*} line. The behavior of the AFM state with magnetic field and angle is not known yet, besides a single nuclear magnetic resonance (NMR) measurement showing a line-broadening starting between 0.2 and 0.3 K at 1.4 T along the cc axis (Supplementary material of Kibune et al. (2022)). The AFM order is expected to be suppressed by the magnetic field and the angle dependence of this suppression might be similar to the HH^{*} line. However, with a transition temperature of TN0.25T_{\mathrm{N}}\approx 0.25 K in zero field, it seems very unlikely that the suppression of TNT_{\mathrm{N}} would be responsible for the transition line at μ0H=4\mu_{0}H^{*}=4 T. In fact, this would imply TNT_{\mathrm{N}} to be almost constant up to 4 T followed by a first order phase transition. While such behavior is unexpected, it has been observed e.g., in systems which are near a ferromagnetic QCP and show field induced tricritical points, like Yb(Rh1-xCox)2Si2 with x=0.18x=0.18 Hamann et al. (2019): In this material the zero-field AFM phase decreases only very little with field along the cc axis, and then a first-order transition to a polarized state occurs. However, the underlying physics of this system is different, since it is antiferromagnetic but very close to a ferromagnetic state with an extremely large magnetocrystalline anisotropy of 10\approx 10. In CeRh2As2 the susceptibility is too small to be close to ferromagnetism and the magnetocrystalline anisotropy is only 2\lessapprox 2. So it is unlikely that similar physics is at play in CeRh2As2.

Although the angle-dependence cannot completely rule out these scenarios, a transition between two superconducting states seems more natural, especially given how well it fits the data.

IV Conclusion

The excellent agreement of the model with the experimental results strengthens the interpretation that the superconducting state changes from even to odd parity at HH^{*} and that the strong anisotropy of the critical field is rooted in the Pauli limiting effect of a helical pseudospin dd-vector. Since other orders are not included in this model and are not needed to obtain the good agreement, their influence on the superconducting phase diagram and especially on HH^{*} appears to be small. However, this point can only be resolved and other scenarios ruled out when more microscopic information on the orders and on the pairing mechanism will be available in the future, so that their interplay can be investigated and understood.

V Acknowledgements

We thank Konstantin Semeniuk, Aline Ramirez, David Moeckli, and Mark Fischer for stimulating discussions. JL and EH acknowledge support from the Max-Planck society for funding of the Max Planck research group "Physics of unconventional metals and superconductors". Torque measurements were carried out at the European High Magnetic Field lab in Grenoble. CG and EH are also supported by the joint Agence National de Recherche (ANR) and DFG program Fermi-NESt through grant GE602/4-1. DCC and PMRB were supported by the Marsden Fund Council from Government funding, managed by Royal Society Te Apārangi. DFA was supported by the US Department of Energy, Office of Basic Energy Sciences, Division of Materials Sciences and Engineering under Award DE-SC0021971.

VI Supplementary material

VI.1 Methods

VI.1.1 Samples

Single crystals of CeRh2As2 were grown in Bi flux as described elsewhere Khim et al. (2021). For ac-susceptibility and torque, we used the same batch as in Khim et al. (2021). For the specific heat, a different batch with almost identical TcT_{c} and T0T_{0} was used.

VI.1.2 Ac-susceptibility

The magnetic ac susceptibility χ\chi was measured using a homemade set of compensated pick-up coils as described in Khim et al. (2021). A superconducting modulation coil built in the main magnet produced an excitation field of 175 μ\muT at 5 Hz. The output signal of the pick-up coils was amplified using a low temperature transformer (LTT-m from CMR) with a winding ratio 1:100, a low noise amplifier SR560 from Stanford Research Systems and finally measured in a digital lock-in setup using a 24 bits PXI2-4492 as data acquisition system Khim et al. (2021). The ac susceptibility measurements were performed in a MX400 Oxford dilution refrigerator down to 45 mK and up to 15 T. For the angle dependence of χ\chi we used a Swedish rotator installed in the dilution refrigerator and both the pick-up coil and the sample rotate together. The angle between the external magnetic field and the sample is defined with respect to the tetragonal cc axis of a small single crystal of CeRh2As2 of a volume of 500×500×500μ\sim 500\times 500\times 500\,\mum3. The data from temperature sweeps were normalized to their respective value in the normal state at 0.5 K for all the magnetic fields applied. For the field sweeps, the absolute value of the signal at different temperatures is given, normalized to the curve measured at 0.45 K.

The Figure 5 shows the temperature dependence of χ\chi with an external magnetic field μ0H\mu_{0}H applied at different angles. The dashed lines show the value of χ\chi where TcT_{c} is defined as the onset of the diamagnetic drop.

Refer to caption
Figure 5: Temperature dependence of ac magnetic susceptibility at differents magnetics field.(a)(b)(c)(d) are the data for 1010^{\circ}, 2020^{\circ}, 2525^{\circ} and 32.532.5^{\circ} respectively. The dashed line shows where TcT_{c} is defined.

VI.1.3 Torque

The magnetic torque τ=M×B\vec{\tau}=\vec{M}\times\vec{B} was measured using a 50 μ\mathrm{\mu}m thin CuBe cantilever via capacitive readout in a top-loading dilution refrigerator at base temperature in the M9 magnet at the Laboratoire National des Champs Magnétiques Intenses. A tiny piece of the resistivity sample from Khim et al. (2021); Hafner et al. (2022) was used. The cantilever plane was parallel to the field for H[110]H\parallel[110] and then rotated to the perpendicular orientation where HcH\parallel c axis. The field was swept always in the same direction out of the superconducting state and back to zero. For data analysis, we subtracted the zero-field capacitance at each angle and then plot the capacitance change as a function of field, which is proportional to the torque. We observe that in the normal state, the angle and field dependence follows rather well the expected sin(2θ)\sin(2\theta) and H2H^{2} dependencies, respectively, where θ\theta is again the angle measured from the c-axis. A scaling of the data with sin(2θ)\sin(2\theta) makes the torque at different angles comparable and dividing by H2H^{2} makes deviations from the normal state behavior more visible, especially when superconductivity sets in. The true torque can in principle be obtained by multiplying with a factor including the spring constant of the cantilever and the position of the sample. Here however, we leave the arbitrary units since we focus on the anomalies at phase transitions.

VI.1.4 Specific heat

The measurements were carried out with a quasi adiabatic heat pulse technique as described in Wilhelm et al. (2004). Compared to the data presented in Khim et al. (2021) where specific heat was measured on a sample with 5 mg mass, the sample here weighed 19 mg. Although samples were from different growth batches, the data in zero field lie exactly on top of each other. The sample was fixed to the platform with wedges of the desired angles. The uncertainty of angle in this method is estimated to ±2.5\pm 2.5^{\circ}.

Refer to caption
Figure 6: Magnetic field dependence of τ/sin(2θ)\tau/\sin(2\theta) for angles as indicated.(a) a strong increase of the hysteresis loop is observed for 2.52.5^{\circ}. (b) no other phase transitions than HqH_{q} and the superconducting state are observed up to 36 T for HabH\parallel ab and up to 26 T for HcH\parallel c. The inset shows the HqH_{q} transition at 9.2 T for an angle of 0.60.6^{\circ} away from HabH\parallel ab.

VI.2 Change of hysteresis loop in the magnetic torque for angles close to the cc axis

Refer to caption
Figure 7: (a) DC-magnetization along c and ab at 100 mK. The red line is the extrapolation of the magnetization to 12 T. (b) is the torque divided by the field, note that τ/HM\tau/H\propto M.

The raw data of magnetic torque for selected magnetic field angles is shown in Fig. 6. In the superconducting state, a strong hysteresis loop is observed for all angles as already observed in the magnetization (shown in Fig. 7a). We observe a change in the hysteresis loop which surprisingly changes sign as a function of field angle. For 2.52.5^{\circ} and 4.6 the up-sweep torque in the superconducting state joins the normal state behavior from the bottom. In contrast, for all larger angles, the up-sweeps approach the normal state behavior from the top (Fig. 6a). For fields very close to the c-axis, the ratio of torque and magnetic field, τ/H\tau/H, is expected to follow the magnetization. Our experimental result nicely follows this expectation as shown in Fig. 7 where we compare M(H)M(H) for field along cc (Fig. 7a) with τ/H\tau/H at 2.5 (Fig. 7b) Khim et al. (2021). The amplitude of the torque is proportional to the difference of the susceptibilities χaχc\chi_{a}-\chi_{c}. Since the magnetization is linear in field for both directions in the normal state with Mab2McM_{ab}\approx 2M_{c} as depicted in Fig. 7a, the normal state susceptibility has the same anisotropy as the magnetization. In the superconducting state at low angles, this difference is larger than in the normal state, implying that the hysteresis loop is larger in McM_{c} than in MaM_{a}. At larger angles, it is the inverse, meaning that the hysteresis loop is larger in MaM_{a} than in McM_{c}. This change causes the difference in the hysteresis loop. Since the hysteresis loop is caused by vortex physics, this is a sign of a change of the anisotropy of the vortex state with magnetic field direction with a strong enhancement of the hysteresis loop in McM_{c} when the magnetic field is close to the c-axis.

In Fig. 6a, the transition at HcrH_{cr} is also visible.

Fig. 6b shows field sweeps up to the highest measured fields at some angles. At 46 and 50 no other transition than Hc2H_{c2} is observed.

VI.3 Fits

Refer to caption
Figure 8: The blue circles are the TcT_{c} obtained form the temperature dependence of the ac-susceptibility (Figure 5) at different magnetic fields for angles of 10, 20, 25 and 32.5. For 0 and 90 the data were taken from Khim et al. (2021). The orange full lines are the fits for SC1, whereas the black doted lines are the expected behavior of SC2.
Refer to caption
Figure 9: Angle dependence of the upper critical field at 45 mK for SC1 (dashed lines) and SC2 (solid line) using an angle-independent HorbH_{orb} of 20 T, 15 T and 10 T.

To perform the theoretical fit to the experimental data of both the angle dependence at low temperature and the temperature dependence at fixed magnetic-field angle, we solve the linearized gap equation in the presence of both Pauli and orbital limiting effects,

ln(t)=0𝑑u[1Fθ+Fθcos(HgθuHPt)]exp(Hu22Horbt2)1sinhu\ln(t)=\int\limits_{0}^{\infty}du\langle\frac{[1-F_{\theta}+F_{\theta}\cos(\frac{Hg_{\theta}u}{H_{P}t})]\exp(\frac{-Hu^{2}}{\sqrt{2}H_{orb}t^{2}})-1}{\sinh u}\rangle (3)

as discussed in the main text, where \langle\ldots\rangle denotes the average over the Fermi surface, HPH_{P} is the Pauli limiting field for in-plane fields, HorbH_{orb} the orbital limiting field, t=T/Tct=T/T_{c}, gθg_{\theta} is the (angle-dependent) effective gg-factor, and FθF_{\theta} quantifies the pair-breaking effects of the Pauli field. For an even-parity superconductor, Fθ=1F_{\theta}=1 as the field will always be completely pair-breaking. On the other hand, for odd-parity states, Fθ=|d^h^θ|2F_{\theta}=|\hat{d}\cdot\hat{h}_{\theta}|^{2}, where d^\hat{d} is the normalized vector describing the pseudospin structure of the superconducting gap, and h^θ\hat{h}_{\theta} is a unit vector that gives the direction of the Zeeman field in the pseudospin basis. Therefore 0Fθ10\leq F_{\theta}\leq 1. The anisotropy of the Pauli limiting comes from both the intrinsic anisotropy of the gg-factor as measured by the susceptibility Khim et al. (2021), and from the spin-orbit coupling which introduces an additional anisotropy in the intraband component of the effective gg-factor, relevant for Pauli limiting, as well as in the pair-breaking parameter FθF_{\theta}. The forms of both gθg_{\theta} and FθF_{\theta} are determined by the spin-orbit coupling,

gθ\displaystyle g_{\theta} =(gc,02cos2θ+gab,02sin2θ)+gab,02α~2sin2θsin2ϕ1+α~2\displaystyle=\sqrt{\frac{\left(g_{c,0}^{2}\cos^{2}\theta+g_{ab,0}^{2}\sin^{2}\theta\right)+g_{ab,0}^{2}\tilde{\alpha}^{2}\sin^{2}\theta\sin^{2}\phi}{1+\tilde{\alpha}^{2}}}
Fθ\displaystyle F_{\theta} =gab,02(1+α~2)sin2θsin2ϕ(gc,02cos2θ+gab,02sin2θ)+gab,02α~2sin2θsin2ϕ,\displaystyle=\frac{g_{ab,0}^{2}\left(1+\tilde{\alpha}^{2}\right)\sin^{2}\theta\sin^{2}\phi}{\left(g_{c,0}^{2}\cos^{2}\theta+g_{ab,0}^{2}\sin^{2}\theta\right)+g_{ab,0}^{2}\tilde{\alpha}^{2}\sin^{2}\theta\sin^{2}\phi}, (4)

where α~\tilde{\alpha} is the strength of the Rashba spin-orbit coupling relative to the interlayer hopping Khim et al. (2021), θ\theta is the angle from the cc-axis at which the field is applied, ϕ\phi is the angle around the Fermi surface which for simplicity we assume consists of a single circular sheet located at the centre of the Brillouin zone. gab,0g_{ab,0} and gc,0g_{c,0} are the values of the gg-factor for in-plane and cc-axis fields, respectively. The anisotropy in FθF_{\theta} plays a much more important role than that of gθg_{\theta}. In particular, for fields in plane, FθF_{\theta} implies that there is Pauli limiting. However, for all other field orientations, FθF_{\theta} implies a divergence in the Pauli field in the zero temperature limit. As temperature approaches zero, this divergence is strongly angle dependent: it is weak for fields close to the basal plane and strong for fields close to the cc-axis. This divergence is cut-off by the orbital field in our theory, but the underlying anisotropy of this divergence still manifests itself, and is the main source of the anisotropy in the SC2 phase. This is reflected in Fig. 9 where HorbH_{orb} is taken as a constant value and the critical field of SC2 is nevertheless highly anisotropic.

To determine the best fit to the data, we first consider the even-parity SC1, for which we set Fθ=1F_{\theta}=1. Using the orbital fields determined from the experimental data at each angle (as described in the main text), we vary the in-plane Pauli limiting field HPH_{P} and spin-orbit coupling strength α~\tilde{\alpha} for all angles simultaneously. The best fit to the SC1 data was found with HP,1=2.3H_{P,1}=2.3 T and α~=2.7\tilde{\alpha}=2.7. These fits are given as orange lines in Fig. 5. Differences in values for fit parameters compared to Khim et al. (2021) are due to the fact that we fit data from ac-susceptibility here (and not specific heat).

Turning to the odd-parity SC2, we must additionally determine the critical temperature Tc,2T_{c,2}, relative to the critical temperature for SC1 Tc,1T_{c,1}. As such, we fit the SC2 data by varying the ratio Tc,2/Tc,1T_{c,2}/T_{c,1}, with the spin-orbit coupling unchanged and the Pauli limiting field reduced from the SC1 value by the scaling HP,2=HP,1(Tc,2/Tc,1)H_{P,2}=H_{P,1}(T_{c,2}/T_{c,1}), where HP,1=2.3H_{P,1}=2.3 T is the Pauli field for SC1, and HP,2H_{P,2} the Pauli field for SC2. In order to keep the amount of parameters as low as possible, we assume that the orbital fields for SC2 take the same value as those for SC1. After determining the optimal Tc,2=0.80Tc,1T_{c,2}=0.80T_{c,1}, we finally allow the orbital fields to vary within the experimental margin of error, to further improve the fits of the temperature dependence. The resulting fits are shown in Fig. 8 as black dashed lines, and are qualitatively robust against small variations in the fitting parameters.

To calculate HH^{*}, the first order transition between SC1 and SC2, we calculated the free energy in the absence of orbital limiting, via the method described in Ref. Cavanagh et al. (2022) using the model parameters HP1H_{P1}, Tc2/Tc1T_{c2}/T_{c1} and α~\tilde{\alpha} found from the procedure described above, with no additional fitting parameters. We include the contribution to the SC2 free energy due to the Pauli limiting effect when θ0\theta\neq 0, accounted for by a modification of the superconducting quasiparticle energy defined by FθF_{\theta},

Eσ=(gθHHP)2+ξ2+Δ¯2+2σ(gθHHPξ)2+FθΔ¯2,E_{\sigma}=\sqrt{\left(g_{\theta}\frac{H}{H_{P}}\right)^{2}+\xi^{2}+\bar{\Delta}^{2}+2\sigma\sqrt{\left(g_{\theta}\frac{H}{H_{P}}\xi\right)^{2}+F_{\theta}\bar{\Delta}^{2}}}, (5)

where ξ\xi is the normal state energy dispersion and Δ¯\bar{\Delta} is the magnitude of the gap projected onto the Fermi surface. The calculated transition line HH^{*} is shown in Fig.2b, where it clearly agrees excellently with the angle dependence of the transition.

Parameter 0 10 20 25 32.5 90
TcT_{c} (K) 0.286 0.280 0.275 0.284 0.282 0.293
dH/dT|Tc-\left.dH/dT\right|_{T_{c}} (T/K) 80 77 71 57 50 27
HorbH_{orb} (T) 16.7 15.7 14.3 11.8 10.3 5.8
HorbH_{orb} (T), fit 16.2 15.7 14.7 11.8 10.3 5.8
Table 1: Experimental values of TcT_{c},dH/dT|Tc-dH/dT|_{T_{c}} and HorbH_{orb}, and HorbH_{orb} fit parameters. From the fitting, we find α~=2.7\tilde{\alpha}=2.7, HP,1=2.3TH_{P,1}=2.3T for SC1, and HP,2=1.84TH_{P,2}=1.84T for SC2, with Tc,2/Tc,1=0.80T_{c,2}/T_{c,1}=0.80 the ratio between the critical temperatures of SC2 and SC1.

References

  • Fisher et al. (1989) R. A. Fisher, S. Kim, B. F. Woodfield, N. E. Phillips, L. Taillefer, K. Hasselbach, J. Flouquet, A. L. Giorgi, and J. L. Smith, Phys. Rev. Lett. 62, 1411 (1989).
  • Aoki et al. (2019) D. Aoki, K. Ishida, and J. Flouquet, J. Phys. Soc. Jpn. 88 (2019).
  • 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, and N. P. Butch, Science 365, 684+ (2019).
  • Sigrist (2005) M. Sigrist, AIP Conference Proceedings 789, 165 (2005).
  • Machida et al. (1985) K. Machida, T. Ohmi, , and M.-A. Ozaki, J. Phys. Soc. Jpn. 54, 1552 (1985).
  • Mineev and Samokhin (1999) V. Mineev and K. Samokhin, Introduction to unconventional superconductivity (Taylor & Francis, 1999).
  • Tada et al. (2011) Y. Tada, N. Kawakami, and S. Fujimoto, J. Phys. Soc. Jpn. 80, SA006 (2011).
  • Khim et al. (2021) S. Khim, J. F. Landaeta, J. Banda, N. Bannor, M. Brando, P. M. R. Brydon, D. Hafner, R. Küchler, R. Cardoso-Gil, U. Stockert, A. P. Mackenzie, D. F. Agterberg, C. Geibel, and E. Hassinger, Science 373, 1012 (2021).
  • Hafner et al. (2022) D. Hafner, P. Khanenko, E.-O. Eljaouhari, R. Küchler, J. Banda, N. Bannor, T. Lühmann, J. F. Landaeta, S. Mishra, I. Sheikin, E. Hassinger, S. Khim, C. Geibel, G. Zwicknagl, and M. Brando, Phys. Rev. X 12, 011023 (2022).
  • Kibune et al. (2022) M. Kibune, S. Kitagawa, K. Kinjo, S. Ogata, M. Manago, T. Taniguchi, K. Ishida, M. Brando, E. Hassinger, H. Rosner, C. Geibel, and S. Khim, Phys. Rev. Lett. 128, 057002 (2022).
  • Yoshida et al. (2012) T. Yoshida, M. Sigrist, and Y. Yanase, Phys. Rev. B 86, 134514 (2012).
  • Skurativska et al. (2021) A. Skurativska, M. Sigrist, and M. H. Fischer, Phys. Rev. Res. 3, 033133 (2021).
  • Nogaki et al. (2021) K. Nogaki, A. Daido, J. Ishizuka, and Y. Yanase, Phys. Rev. Res. 3, L032071 (2021).
  • Yoshida et al. (2014) T. Yoshida, M. Sigrist, and Y. Yanase, J. Phys. Soc. Jpn. 83, 013703 (2014).
  • (15) P. Khanenko et al., to be published .
  • Möckli and Ramires (2021a) D. Möckli and A. Ramires, Phys. Rev. Res. 3, 023204 (2021a).
  • Möckli and Ramires (2021b) D. Möckli and A. Ramires, Phys. Rev. B 104, 134517 (2021b).
  • Ptok et al. (2021) A. Ptok, K. J. Kapcia, P. T. Jochym, J. Łażewski, A. M. Oleś, and P. Piekarz, Phys. Rev. B 104, L041109 (2021).
  • Onishi et al. (2022) S. Onishi, U. Stockert, S. Khim, J. Banda, M. Brando, and E. Hassinger, Frontiers in El. Mat.  (2022).
  • Landaeta (2021) J. F. Landaeta, Supplementary Material  (2021).
  • Effantin et al. (1985) J. M. Effantin, J. Rossat-Mignod, P. Burlet, H. Bartholin, S. Kunii, and T. Kasuya, J. Magn. Magn. Mat. 47,48, 145 (1985).
  • Clogston (1962) A. M. Clogston, Phys. Rev. Lett. 9, 266 (1962).
  • Chandrasekhar (1962) B. S. Chandrasekhar, Appl. Phys. Lett. 1, 7 (1962).
  • Helfand and Werthamer (1966) E. Helfand and N. R. Werthamer, Phys. Rev. 147, 288 (1966).
  • Tinkham (1996) M. Tinkham, Introduction to Superconductivity, Vol. 2nd ed. (McGraw- Hill, New York, 1996).
  • (26) Note that this is true only if both SC states live on the same Fermi surface since Horb{H}_{orb} depends mainly on the Fermi velocity. If different Fermi surfaces are responsible for SC1 and SC2 the anisotropy of Horb{H}_{orb} of SC2 and SC1 could be different. In addition, the orbital limit can also strongly depend on the nodal structure of the gap function Tada et al. (2011), this is currently unknown for CeRh2As2 .
  • Frigeri et al. (2004) P. A. Frigeri, D. F. Agterberg, A. Koga, and M. Sigrist, Phys. Rev. Lett. 92, 097001 (2004).
  • Schertenleib et al. (2021) E. G. Schertenleib, M. H. Fischer, and M. Sigrist, Phys. Rev. Research 3, 023179 (2021).
  • Cavanagh et al. (2022) D. C. Cavanagh, T. Shishidou, M. Weinert, P. M. R. Brydon, and D. F. Agterberg, Phys. Rev. B 105, L020505 (2022).
  • Kittaka et al. (2009) S. Kittaka, T. Nakamura, Y. Aono, S. Yonezawa, K. Ishida, and Y. Maeno, Phys. Rev. B 80, 174514 (2009).
  • Ikeda et al. (2001) S. Ikeda, H. Shishido, M. Nakashima, R. Settai, D. Aoki, Y. Haga, H. Harima, Y. Aoki, T. Namiki, H. Sato, and Y. Ōnuki, J. Phys. Soc. Jpn. 70, 2248 (2001).
  • Farrar et al. (2020) L. S. Farrar, M. Bristow, A. A. Haghighirad, A. McCollam, S. J. Bending, and A. I. Coldea, NPJ Quant. Mat. 5, 29 (2020).
  • Zuo et al. (2017) H. Zuo, J.-K. Bao, Y. Liu, J. Wang, Z. Jin, Z. Xia, L. Li, Z. Xu, J. Kang, Z. Zhu, and G.-H. Cao, Phys. Rev. B 95, 014502 (2017).
  • Goh et al. (2012) S. K. Goh, Y. Mizukami, H. Shishido, D. Watanabe, S. Yasumoto, M. Shimozawa, M. Yamashita, T. Terashima, Y. Yanase, T. Shibauchi, A. I. Buzdin, and Y. Matsuda, Phys. Rev. Lett. 109, 157006 (2012).
  • Aoki et al. (2009) D. Aoki, T. Matsuda, V. Taufour, E. Hassinger, G. Knebel, and J. Flouquet, J. Phys. Soc. Jpn. 78, 113709 (2009).
  • Kimura et al. (2005) N. Kimura, K. Ito, K. Saitoh, Y. Umeda, H. Aoki, and T. Terashima, Phys. Rev. Lett. 95, 247004 (2005).
  • Kimura et al. (2007) N. Kimura, K. Ito, H. Aoki, S. Uji, and T. Terashima, Phys. Rev. Lett. 98, 197001 (2007).
  • Sugitani et al. (2006) I. Sugitani, Y. Okuda, H. Shishido, T. Yamada, A. Thamizhavel, E. Yamamoto, T. D. Matsuda, Y. Haga, T. Takeuchi, R. Settai, and Y. Ōnuki, J. Phys. Soc. Jpn. 75, 043703 (2006).
  • Bauer and Sigrist (2012) E. Bauer and M. Sigrist, Non-Centrosymmetric Superconductors : Introduction and Overview, Lecture Notes in Physics, Vol. 847 (Springer-Verlag, Berlin Heidelberg, 2012).
  • Kitagawa et al. (2022) S. Kitagawa, M. Kibune, K. Kinjo, M. Manago, T. Taniguchi, K. Ishida, M. Brando, E. Hassinger, C. Geibel, and S. Khim, J. Phys. Soc. of Jpn. 91, 043702 (2022)https://doi.org/10.7566/JPSJ.91.043702 .
  • Settai et al. (2008) R. Settai, Y. Miyauchi, T. Takeuchi, F. Lévy, I. Sheikin, and Y. Ōnuki, J. Phys. Soc. Jpn. 77, 073705 (2008).
  • Méasson et al. (2009) M.-A. Méasson, H. Muranaka, T. Kawai, Y. Ota, K. Sugiyama, M. Hagiwara, K. Kindo, T. Takeuchi, K. Shimizu, F. Honda, R. Settai, and Y. Ōnuki, J. Phys. Soc. Jpn. 78, 124713 (2009).
  • Hamann et al. (2019) S. Hamann, J. Zhang, D. Jang, A. Hannaske, L. Steinke, S. Lausberg, L. Pedrero, C. Klingner, M. Baenitz, F. Steglich, C. Krellner, C. Geibel, and M. Brando, Phys. Rev. Lett. 122, 077202 (2019).
  • Wilhelm et al. (2004) H. Wilhelm, T. Lühmann, T. Rus, and F. Steglich, Rev. Sci. Instrum. 75, 2700 (2004).