Abstract
The detailed anisotropic dispersion of the low-temperature, low-energy magnetic excitations of the candidate spin-triplet superconductor UTe2 is revealed using inelastic neutron scattering. The magnetic excitations emerge from the Brillouin zone boundary at the high symmetry Y and T points and disperse along the crystallographic -axis. In applied magnetic fields to at least μ0H = 11 T along the , the magnetism is found to be field-independent in the (hk0) plane. The scattering intensity is consistent with that expected from U3+/U4+ f-electron spins with preferential orientation along the crystallographic -axis, and a fluctuating magnetic moment of μeff=1.7(5) μB. We propose interband spin excitons arising from f-electron hybridization as a possible origin of the magnetic excitations in UTe2.
Subject terms: Superconducting properties and materials, Magnetic properties and materials
Introduction
The heavy-fermion paramagnet UTe2 has been a subject of focused investigation within the strongly correlated electron community since the recent discovery of its highly nontrivial superconducting ground state1. UTe2 strongly violates the Pauli limit for BCS superconductors, which relates the temperature of the onset of superconductivity Tc ≈ 2K2,3 to the critical magnetic field. The Pauli limit is exceeded by factors of approximately 2, 4, and 2.5 for fields along the crystalline , , and directions respectively, and magnetization scaling strongly suggests proximity to ferromagnetism1,4 despite no observation of long-ranged magnetic order, which was taken as initial supporting evidence of a spin-triplet superconducting ground state in UTe2.
Nuclear magnetic resonance (NMR) studies indicate that the spin susceptibility remains constant upon cooling below Tc5–7. This violates expectations for spin-singlet BCS superconductivity, where a Knight shift signifies is associated with the condensation of the superconducting quasiparticles. This, along with other experimental results such as polar Kerr effect measurements8–10, exotic reentrant superconductivity at fields above 35 T11,12, and fluctuations in μSR13–15, are all taken to be supporting evidence for a potential spin-triplet superconducting ground state in UTe2. Such a superconducting pairing is exceedingly rare in condensed matter systems, with the only conclusive example being that of the rigorously studied superfluid phase of 3He16. A spin-triplet superconductor is of particular interest in the context of applications to quantum devices.
Many open questions remain regarding the ground state superconductivity in UTe2. The superconducting nodal gap function17–19, the order parameters associated with each of the superconducting phases5,8, time reversal symmetry breaking9, the role of magnetic interactions, and the electronic band structure and Fermi surface are all unresolved. As spin-triplet superconductivity can be mediated by ferromagnetism (although an antiferromagnetic mechanism20,21 is also possible) unambiguous understanding of the spin interactions is required to describe the paramagnetic ground state of UTe2. Additionally, band hybridization in heavy fermion systems involves the Kondo effect and RKKY interactions, and the delicate interplay of these effects with f-electron band has been linked to unconventional superconductivity and conduction electron mediated magnetic exchange interactions22. Their role in UTe2 is currently not well understood.
Inelastic neutron scattering (INS) experiments are the most direct probe of spin fluctuations and magnetism in heavy fermion superconductors, many of which possess many similarities to UTe2. In the cases of UCoGe23,24 and UGe224,25, ferromagnetic fluctuations are observed, whereas antiferromagnetic correlations associated with superconductivity have been observed in materials like URu2Si226,27 and the cuprates28. However, the imaginary susceptibility measured by INS in correlated f-electron systems often arises not only from magnetic excitations like magnons, but from band excitons as in the cases of CePd329 and SmB630. Further, it has recently been demonstrated that in f-electron systems even the magnons in ordered phases sensitively depend upon the underlying correlated band structure31. An applied magnetic field can help to distinguish between magnons and band excitons. For example, in the case of CeRu2Si232, magnetic field strongly suppresses magnon scattering due to interionic correlations, as opposed to the gentle perturbations expected in the band exciton picture.
Earlier INS studies of UTe2 report a condensation of excitations emerging at the Brillouin zone boundary upon cooling below the Kondo hybridization temperature of TK ≈ 40K33. Additional INS results consistently report the presence of a broad continuum of excitations in the superconducting state peaked in intensity at energies of ℏω = 4 meV34,35, where ℏω denotes neutron energy transfer, with the scattering being confined to the (hk0) plane. Additionally, an excitation observed a ℏω=1 meV was suggested to be a resonant excitation which originates from a bound state within the particle-hole continuum gap35,36. This was taken as evidence that the spin fluctuations were primarily antiferromagnetic in nature, as one would expect ferromagnetic fluctuations at integer Q vectors, where Q denotes momentum transfer.
Here, we report INS results on UTe2 in the (hk0) scattering plane with an applied field along the axis. Our results resolve the low energy excitations across multiple Brillouin zones in the (hk0) scattering plane at energies below 6 meV, accounting for nearly all the expected spectral weight from the magnetic moment of U3+. The measured scattering intensity is independent to an applied magnetic field of μ0H ≤11 T, suggesting that the magnetic response is dominated by interband excitations excited across the df Kondo hybridization gap, rather than ferromagnetic or antiferromagnetic exchange between local spins.
Results
Inelastic Neutron Scattering Results
The INS measurements were performed on a mosaic of coaligned single crystalline samples of UTe2 ranging in Tc ∈ {1.8, 2.0} K, with bulk properties summarized in Fig. 1a, b. Scattering integrated over the elastic window (ℏω ∈ {−0.15, 0.15} meV) shows no evidence of magnetic or field-dependent scattering. A mosaic spread of full width at half maximum (FWHM) 5.0(1)o in the (hk0) plane is determined from the nuclear Bragg peaks. A sample environment background is determined using a momentum averaged scattering from the regions in Q, ω space where magnetic scattering from the sample is taken to be absent, which is described in detail in the Supplementary Information (SI).
Two representative constant energy slices of the magnetic scattering from UTe2 are shown in Fig. 2. The scattering in all cases has been symmetrized by the two-fold C2m lattice symmetry to enhance statistics. The lowest energy contribution to the scattering is shown in Fig. 2a, b. As in previous studies, the scattering is constrained to the Brillouin Zone (BZ) edge, emanating from the corners (T=(,0)-like points) and edges (Y=(0, 0.6, 0)-type points). Scattering at higher Brillioun zones is also observed.
Constant energy slices in energy window of maximum scattering intensity ℏω ∈ {3.2, 3.8} meV are shown for the 0 T measurement in Fig. 2c and 11 T in Fig. 2d. Here the scattering remains on the BZ edges, but the excitations have dispersed such that they have smeared in the (h00) direction. A minimum in intensity has emerged along the (0k0) line and zero intensity is observed at all Γ points. As is made more quantitatively clear in Fig. 3, there is no qualitative difference in scattering intensity for the 0 T and 11 T settings apart from a small deviation shown in Fig. 3(a). This also true for the elastic scattering, as shown in the SI.
In Fig. 3, representative intensity cuts along Q and energy transfer are shown. All data presented in cuts is not symmetrized. Figure 3(a) and (b) plot constant Q cuts along energy at the high symmetry points Y1 = (0, 0.6, 0) in (a) and T2 = (0.4, 1.4, 0) in (b). Both cuts are integrated in the Q dimension in a circle of radius δQ = 0.3 Å−1 (h ± 0.2 r.l.u., k ± 0.29 r.l.u.). The energy dependence of the intensity in both plots shows a monotonic increase in intensity from ℏω = 0, with a peak of intensity at 3.2(1) meV at Y1 and 4.0(1) meV T2 meV respectively. There is no statistically significant field-dependence to the scattering at the T2 point. At energies below ℏω = 3meV, the zero field scattering is slightly more intense at the Y1 point, as shown in Fig. 3a. The integrated intensity within ℏω ∈ {0.3, 3.0} is 2.0(3) b/U for μ0H = 0 T, and 1.5(3) b/U for μ0H = 11 T. We are unable to resolve the proposed spin-resonance excitation reported in refs. 34,35.
Figure 3c, d show constant energy cuts integrated in ℏω ∈ {1, 2} meV. This integration window accounts for most of the scattering before dispersion becomes significant. Consistent with previous studies33–37, we find broad peaks in intensity centered at the Brillouin zone edges. Again, there is no difference between the 0 T and 11 T scattering intensities. In (c) the cut is along the (h, 1.5, 0) direction, i.e. along the BZ edge, and in (d) the cut is along the (0, k, 0) direction. In both cases the scattering was integrated in a perpendicular Q range of 0.3 Å−1. The cuts are simultaneously fit to a model of nearest neighbor correlated spins described later in the text.
Finally, Fig. 4a, b present the full dispersion of the magnetic excitations in UTe2 along the zone boundary, with Fig. 4c including the Γ0 = (000), Γ1 = (010), and Γ2 = (020) points. The Q path in Fig. 4a is primarily along the (h00) direction and is depicted by the blue path in the schematic of the Brillouin zone in (c), and (b) is along the (0k0) direction and depicted by the red path in (f). The third path along (0k0) is shown in (c), revealing no dispersion at the Y-type points and zero intensity at the zone center. The scattering is broader in energy than the expected instrumental resolution (δEFWHM = 0.11 meV for ℏω=0 meV), and at every point in Q the energy dependent intensity I(ℏω) may be fit to the form of a simple Lorentzian peak of FWHM width ΔEFWHM=3.0(2) meV. For all values of Q with significant enough intensity to perform this fit, the extracted dispersion of magnetic excitations along the representative paths in Q is presented in Fig. 4d, e. Along both directions, the peak in scattering intensity has dispersion with clear periodicity. Due to the magnetic form factor of U3+/4+ and a net spin polarization factor in the neutron scattering cross-section, these excitations become faint at the highest accessible values of Q. The dispersion is steeper along the (0k0) direction, as shown both in Fig. 4b, e, and in all cases peaks at the T points.
The second difference between the observed scattering and previous reports is that the resonant excitation reported in refs. 35,36 is not seen in our measurement. While this resonance was initially reported at the Y1 = (0, 0.6, 0) point, independent measurements also observed the excitation at the Y2 = (0, 1.4, 0) point36. Here, our results do not show evidence of a peak in scattering at ℏω ≈ 1 meV at either Y1 or Y2. A more detailed examination is in the SI, but this does not preclude the existance of the resonant excitation UTe2. There are many different details of the experiment that may be responsible for this, such as mosaic distribution, detector coverage, or sample-dependence, and the resonance is not the focus of this work.
Discussion
Overall, the scattering presented is compatible with previous INS studies in the superconducting phase. There are two significant differences. First, these data allow for the full extraction of the dispersion of magnetic excitations in the (hk0) scattering plane. An important result that we extract from this is the quantity of the fluctuating moment for both the 0 T and 11 T configurations. Because the scattering has been normalized, the well-known total moment sum rule38,39 of the form
1 |
The dynamical structure factor S(Q, ω) is related to the scattering intensity by
2 |
Here, the scattering has been normalized by the kinematic ratio of final and initial momenta , the Lande g-factor g is taken to be 2, ∣F(Q)∣2 is the modulus of the magnetic form factor associated with scattering from U3+ or U4+, which are indistinguishable in this measurement, r0 = 5.391 ⋅ 10−13 cm is the characteristic magnetic neutron scattering length, and the summation is over the Cartesian x, y, z spin directions.
As the experiment cannot access to the (00l) scattering direction, to perform the integration over the full Brillouin zone we refer to the results of ref. 37 which suggest no dispersion along the (00l) direction and intensity with a Gaussian distribution centered at l = 0, with a width of σFWHM ≈ 0.5 rlu. The scattering was integrated in energy over the full range of observed excitations, ℏω ∈ {0.3, 6} meV, and in two Brillouin zones h ∈ {0, 1} and k ∈ {0.5, 2.5}. From this, we find an overall fluctuating moment of μeff =1.7(5) μB, which may be compared to the itinerant moment of approximately 1.1 μB as measured by high field magnetization40. It may also be compared with the moment found from the paramagnetic susceptibility of 3.6μB41 or free-ion U3+ of 3.62 μB. One may also compare the moment inferred from the total sum rule to the itinerant moment, which high-field magnetization finds to be near 1.1 μB40. By comparing the zero field and 11 T integrated Bragg peak intensities at the (110), (10), and (020) Bragg peaks we find an upper limit of the field-induced ordered moment of μ = 0.0(2) μB. This is slightly at odds with what one would expect from magnetization, which for fields along the b axis finds an induced moment on the order of μ = 0.15 μB1 at 11 T. A more detailed diffraction study using a large single crystal rather than a mosaic of coaligned samples would be more appropriate to resolve this moment.
The lowest energy part of the scattering emerges from the Y and T points. The Q-dependent intensity of the scattering can be directly calculated by assuming a form of S(Q, ω), where all spins in the unit cell have the same preferential spin orientation. The scattering takes the form
3 |
The sum is over all of the Y and T points, where the intensity is assumed to have a Gaussian line shape of amplitude AT for T-points and AY for Y-points. The peaks are of width σh and σk in the h and k directions, respectively. The prefactor is defined by
4 |
Here, the preferential spin orientation is constrained to be in the (hk0) plane, with the tilt away from the a axis towards the b axis being denoted as θa. No significant difference could be found between constraining the moment to the (hk0) plane and an entirely free orientation, apart from an upper limit of 35o from the -axis towards the -axis. Fitting the constant energy slices integrated within ℏ ∈ {1.7, 2.3} meV finds best fit values of θa=θa=17(3) degrees as depicted in Fig. 3e. The minimum at θa=17(3) degrees is shallow, but the scattering is clearly incompatible with a spin orientation beyond 30 degrees. This result demonstrates that the spins contributing to the scattering prefer to orient along the crystalline -axis within the ab plane, which is consistent with previous neutron studies34 and bulk magnetometry indicating that is the magnetic easy axis1.
The character of the scattering in Fig. 2a gives important information about the electronic band structure near the Fermi level ϵF. As shown by the fit in 3c, d, the U3+/4+ form factor is sufficient to describe the intensity modulation of the scattering with increasing ∣Q∣, indicating that the spin density of electrons in the conducting band originates from the 5f orbital. This is consistent with recent RIXS reports42 suggesting a 5f2 configuration. Secondly, we note that the scattering is found at Y = (0, 0.6, 0) and type points, but is absent at the type points. This cannot be explained exclusively through the effect of spin polarization, as there are many symmetry-equivalent M-type points available in the (hk0) scattering plane where no scattering is observed.
The scattering observed in this experiment confirms results from previous works33–37, and has direct physical consequences. The first is that at temperatures below TK ≈ 40 K, a dispersive magnetic excitation mode emerges which is peaked in intensity at 3.5(2) meV and is broad in energy (ΔEFWHM ≈ 1.5 meV). While the scattering intensity may persist to zero energy transfer, at the lowest accessible energies the intensity becomes vanishingly small and the maxima of all excitations bands are fully gapped. There is no evidence for long-range magnetic order, but the fluctuating moment has preferential spin orientation along the axis. The excitations themselves are constrained to the BZ edge, with no significant scattering at the Γ point at all measured energy transfers. Finally, the momentum dependence of the scattering is described well by ionic U3+/4+ and a total moment sum rule suggests that the scattering captures the majority of the relevant magnetic excitations. Though these results are informative, the interpretation of the origin of the magnetic scattering is less straightforward.
Previous INS studies of UTe2 have characterized these excitations at non-integer Q as antiferromagnetic spin fluctuations, suggesting a picture with dominant antiferromagnetic interactions between nearest neighbor U ions35,37. Subsequent analysis of neutron scattering spectra suggested the presence of both ferro and antiferromagnetic interactions in a spin-ladder model, where spin exchange interactions are ferromagnetic between the U3+ chains along the -axis, ferromagnetic within the U-dimers along the -axis, and antiferromagnetic between ladders along the -axis34. Additionally, a pressure-induced antiferromagnetic ordered phase in UTe2 was found at an experimentally modest hydrostatic pressure of about Pc=1.5 GPa43,44. These studies report an incommensurate AFM ordering wavevector of k = (0.07, 0.33, 1)44, and the close proximity of the ground state to an AFM ordered phase was taken as evidence for strong AFM magnetic interactions. This contradicts initial suggestions that spin-triplet superconductivity in UTe2 is promoted by proximity to a ferromagnetic critical point. While spin-triplet superconductivity has been suggested to be possible in the case of AFM interactions45, the precise nature of the spin interactions is still an important open question.
A magnon-like picture where the dispersion of the magnetic excitations originates from AFM interactions between spins raises a number of questions. The first is how magnons can emerge from a ground state that is paramagnetic. In this scenario, the excitations would be classified as paramagnons, where spins are proximate to local magnetic order due to strong spin correlations. Such a scenario is compatible with the broad signal observed in UTe2. However, our data show that the magnetic response has no dependence on the applied magnetic field up to a relatively large field scale of μ0H=11 T, contrary to what one would expect in the case of spins coupled with a Heisenberg interaction on the scale of 4 meV. Additionally, in the generic Heisenberg picture using linear spin wave theory the magnon dispersion minima occur at the magnetic ordering wave vector km. Here, and in previous works33,35,37, these minima are observed at the in equivalent Y and T points, requiring the further complexity of two different species of magnetic order or a multi-k structure. Because such a scenario quickly becomes quite complicated, we consider an alternative picture.
Neutron scattering from localized U-f spins in heavy fermion compounds has previously been treated in the context of an extended Anderson lattice model46,47. In the case of UTe2, this would mean that the Kondo effect hybridizes the U-6d, U-5f, and Te-5p electronic bands near the chemical potential. In the extended Anderson lattice picture, the newly formed hybridized bands are constrained by symmetry to have saddle points with a significant density of states at high symmetry points at the Brillouin zone edge. This treatment provides an alternative description of the scattering, which is that of interband electronic transitions between regions with high density of state between the hybridized bands, which are at the zone edge and zone boundary48 and naturally account for the field-independence of the excitations.
To explore this scenario, we compare to high resolution DFT+DMFT calculations of the band structure of UTe2. These calculations were performed using the Rutgers code implemented in the WIEN2k package using the experimentally determined crystal structure49, at temperatures of T=116 K, T=232 K, and T=580 K. A computational limitation of DFT+DMFT is the lower temperature limit, and thus the energy resolution of the band structure on the ≈ 10 meV scale. Thus, the detailed band structure near the chemical potential ϵf is not resolved in these calculations. Figure 5b, d plot the calculated quasiparticle band structure, which closely resembles our previously reported results in ref. 50. We also note that these calculations are similar to those presented in ref. 37, with the primary difference being that there is no infinite Hubbard U approximation. These results give insight into the hybridization associated with the f − band correlations, and while the precise energies within 10 meV of ϵf are unreliable due to the temperature constraint, the development of anticrossings of bands is evident. The anticrossings are denoted using arrows in Fig. 5c, and result in regions of high f-density of states at high symmetry points, between which interband excitations may occurr.
The calculated high temperature Fermi surface is shown in Fig. 5a, which finds two elliptical pockets similar to previous experimental and theoretical reports51–53. At low temperature, Fig. 5(b), the effect of hybridization may be observed with the shape of the pockets changing dramatically and a new pocket emerging around the Z-point. This can be compared directly to photoemission experiments performed at T = 20 K, which report a heavy electron pocket at the Z point53, and quantum oscillation measurements which suggest both 2D and 3D Fermi surface character, with conflicting conclusions as to the presence of such a pocket54,55. In both plots, the blue/red surfaces originate from Te-5p band, and the blue/yellow surfaces originate from the U-6d band.
To examine the redistribution of f − weight at low temperature, we examine constant energy kz=0 slices in Fig. 6. Figure. 6a, b show the kz=0 Fermi surface for T=116 K and T=580 K respectively. We take the temperature difference of the spectral function A(Q, ω) to be a measure of the U-5f band weight, which accumulates in pockets centered about the X point at E=Ef, as shown as a function of energy transfer in Fig. 6c–f. The f-weight disperses towards the zone corner at higher energy transfers, which is suggestive of the origin of the dispersion of the excitations observed in INS. This dispersion, which originates from hybridization, is also anisotropic in the (hk0) plane, as most clearly demonstrated at the chemical potential in Fig. 6d.
Using the distribution of the f-electron character presented in Fig. 6d, it is tempting to assign potential nesting vectors between regions within high spectral weight. This would lend support to the idea of an emergent charge density wave which has been observed in scanning tunneling microscopy (STM) studies, but has not been observed in bulk scattering measurements56–60. As no such wavevector is observed in the elastic component of the magnetic scattering, such a phase would need to be a surface state rather than in the bulk.
Due to the energy resolution of the calculation, it is not possible to determine the precise scattering pathways that would lead to the measured S(Q, ℏω) in INS. However, the calculations do show that the hybridized band structure features parabolic-like bands with extrema at the zone boundary, as suggested in the Anderson picture. Two essential conditions are required for interband scattering, the first is the presence of a high density of states near the chemical potential at the high symmetry points at which scattering is observed, and the second being anisotropy in the band hybridization resulting in the lack of scattering at the X point, both of which are captured in the DFT+DMFT calculations.
Our work resolves the magnetic excitations in UTe2 that peak at ℏω=3.5(2) meV, which we propose originate from electronic excitations near or across the Kondo hybridization gap. This conclusion is also supported by our previous work33, in which the temperature dependence of the magnetic scatting intensity is linked to the Kondo coherence tmperature. Here, the excitations capture a total fluctuating magnetic moment of μeff =1.7(5) μB, which is approximately compatible with what one would expect for free-ion U3+. Our analysis shows that their origin is likely not from magnetic interactions, but no conclusion is made regarding the proximity of UTe2 to any particular ordered state, or the nature of its magnetic interactions which have been reported as both ferromagnetic1, and antiferromagnetic37.
This is consistent with previous DFT+U studies, which suggest that ferromagnetic inter dimer interactions between U-ions may stabilize spin-triplet superconductivity61–63. The scattering captures the full expected magnetic moment in the system, and we expect that this scenario could be supported by further DMFT and calculations of similarly to the studies of refs. 29,30. As we are unable to perform such calculations of in this work, we cannot directly link the observed scattering to the calculated electronic band structure, and leave open the possibility of alternative scattering mechanisms. The DMFT results can be verified directly through photoemission high resolution experiments at low temperatures, which would be able to resolve the elliptical pockets predicted in our calculations. Additionally, we hope that this work will motivate the further development of DFT+DMFT methods to enable calculations of in the meV scale energy regime, which is relevant not only in the case of UTe2 but many other heavy fermion systems with Kondo hybridization.
Methods
The INS experiment was performed on a coaligned mosaic of 82 crystals of UTe2 with a total mass of 1.1 g. Most crystals were grown by chemical vapor transport64, with eight of total mass ≈ 80mg being grown by a salt flux method2,3. All batches of crystals were screened by either transport measurements or susceptibility, with a range of critical temperatures of Tc ∈ {1.8, 2.0} K indicating high sample quality. The crystals were mounted on a oxygen-free copper sample holder using CYTOP (AGC Chemicals Company) and aligned using a x-ray Laue diffractometer.
The INS measurement was performed using the CAMEA spectrometer65 at the Paul Scherrer Institut to measure the inelastic scattering in an energy transfer range of ℏω ∈ { − 1, 6} meV. The sample environment was the MB11 μ0H=11 T vertical cryomagnet with a dilution refrigerator insert with a base temperature of T=45 mK throughout the experiment based on the sample thermometer. All measurements were performed in the (hk0) scattering plane, with the magnetic field oriented along the direction. For the zero-field configuration, the total counting time was 61 hrs, and for the 11 T configuration 64 hrs. The scattering was normalized to absolute units of (b/meV/sr/U) using a vanadium standard of known mass. All analysis was performed using the MJOLNIR software66,67.
Supplementary information
Acknowledgements
This work is based on experiments performed at the Swiss spallation neutron source SINQ (proposal number 20230495), Paul Scherrer Institut, Villigen, Switzerland. Work at UMD was supported by NIST and by the National Science Foundation under the Division of Materials Research Grant NSF-DMR 2105191. MJ acknowledges funding by the Swiss National Science Foundation through the project ”Berry-Phase Tuning in Heavy f-Electron Metals (#200650). Identification of commercial equipment does not imply recommendation or endorsement by NIST. GK was supported by DOE BES CMS program. C.-J.K. was supported by the National Research Foundation of Korea (NRF) (Grant No. NRF-2022R1C1C1008200).
Author contributions
Synthesis of samples used in this study was performed by C.F., P.C., and G.S. The INS experiment was conceived by N.B. and M.J., and performed by T.H., P.C., and J.A.R., with assistance from D.G.M. and J.L. DFT+DMFT calculations were performed by C.J.K. and G.K. The manuscript was written by T.H. and N.B. with input from all authors.
Data availability
All data are available upon reasonable request to Thomas Halloran.
Code availability
All analysis was performed using the MJOLNIR software66,67, and code used in the analysis is available upon reasonable request to Thomas Halloran.
Competing interests
The authors declare no competing interests.
Footnotes
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Supplementary information
The online version contains supplementary material available at 10.1038/s41535-024-00720-9.
References
- 1.Ran, S. et al. Nearly ferromagnetic spin-triplet superconductivity. Science365, 684–687 (2019). [DOI] [PubMed] [Google Scholar]
- 2.Sakai, H. et al. Single crystal growth of superconducting UTe2 by molten salt flux method. Phys. Rev. Mater.6, 073401 (2022). [Google Scholar]
- 3.Aoki, D. Molten salt flux liquid transport method for ultra clean single crystals UTe2. J. Phys. Soc. Jpn.93, 043703 (2024). [Google Scholar]
- 4.Tokiwa, Y. et al. Reinforcement of superconductivity by quantum critical fluctuations of metamagnetism in ute2. Phys. Rev. B109, L140502 (2024). [Google Scholar]
- 5.Fujibayashi, H. et al. Superconducting Order Parameter in UTe2 Determined by Knight Shift Measurement. J. Phys. Soc. Jpn.91, 043705 (2022). [Google Scholar]
- 6.Nakamine, G. et al. Anisotropic response of spin susceptibility in the superconducting state of UTe2 probed with 125Te − NMR measurement. Phys. Rev. B103, L100503 (2021). [Google Scholar]
- 7.Ambika, D. V. et al. Possible coexistence of antiferromagnetic and ferromagnetic spin fluctuations in the spin-triplet superconductor UTe2 revealed by 125Te NMR under pressure. Phys. Rev. B105, L220403 (2022). [Google Scholar]
- 8.Hayes, I. M. et al. Multicomponent superconducting order parameter in UTe2. Science373, 797–801 (2021). [DOI] [PubMed] [Google Scholar]
- 9.Wei, D. S. et al. Interplay between magnetism and superconductivity in UTe2. Phys. Rev. B105, 024521 (2022). [Google Scholar]
- 10.Ajeesh, M. et al. Fate of time-reversal symmetry breaking in UTe2. Phys. Rev. X13, 041019 (2023). [Google Scholar]
- 11.Ran, S. et al. Extreme magnetic field-boosted superconductivity. Nat. Phys.15, 1250–1254 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Lewin, S. K., Frank, C. E., Ran, S., Paglione, J. & Butch, N. P. A review of UTe2 at high magnetic fields. Rep. Prog. Phys.86, 114501 (2023). [DOI] [PubMed] [Google Scholar]
- 13.Sundar, S. et al. Coexistence of ferromagnetic fluctuations and superconductivity in the actinide superconductor UTe2. Phys. Rev. B100, 140502 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Azari, N. et al. Absence of Spontaneous Magnetic Fields due to Time-Reversal Symmetry Breaking in Bulk Superconducting UTe2. Phys. Rev. Lett.131, 226504 (2023). [DOI] [PubMed] [Google Scholar]
- 15.Sundar, S. et al. Ubiquitous spin freezing in the superconducting state of UTe2. Commun. Phys.6, 1–11 (2023). [Google Scholar]
- 16.Leggett, A. J. A theoretical description of the new phases of liquid 3He. Rev. Mod. Phys.47, 331–414 (1975). [Google Scholar]
- 17.Suetsugu, S. et al. Fully gapped pairing state in spin-triplet superconductor UTe2. Sci. Adv.10, eadk3772 (2024). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Ishizuka, J., Sumita, S., Daido, A. & Yanase, Y. Insulator-Metal Transition and Topological Superconductivity in UTe2 from a First-Principles Calculation. Phys. Rev. Lett.123, 217001 (2019). [DOI] [PubMed] [Google Scholar]
- 19.Ishihara, K. et al. Chiral superconductivity in UTe2 probed by anisotropic low-energy excitations. Nat. Commun.14, 2966 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Hu, H., Cai, A., Chen, L. & Si, Q. Spin-singlet and spin-triplet pairing correlations in antiferromagnetically coupled Kondo systems http://arxiv.org/abs/2109.12794 (2021).
- 21.Kreisel, A., Quan, Y. & Hirschfeld, P. J. Spin-triplet superconductivity driven by finite-momentum spin fluctuations. Phys. Rev. B105, 104507 (2022). [Google Scholar]
- 22.White, B. D., Thompson, J. D. & Maple, M. B. Unconventional superconductivity in heavy-fermion compounds. Phys. C: Superconductivity its Appl.514, 246–278 (2015). [Google Scholar]
- 23.Stock, C. et al. Anisotropic Critical Magnetic Fluctuations in the Ferromagnetic Superconductor UCoGe. Phys. Rev. Lett.107, 187202 (2011). [DOI] [PubMed] [Google Scholar]
- 24.Prokeš, K., de Visser, A., Huang, Y. K., Fåk, B. & Ressouche, E. Anomalous spin distribution in the superconducting ferromagnet UCoGe studied by polarized neutron diffraction. Phys. Rev. B81, 180407 (2010). [Google Scholar]
- 25.Haslbeck, F. et al. Ultrahigh-resolution neutron spectroscopy of low-energy spin dynamics in UGe2. Phys. Rev. B99, 014429 (2019). [Google Scholar]
- 26.Butch, N. P. et al. Symmetry and correlations underlying hidden order in URu2Si2. Phys. Rev. B91, 035128 (2015). [DOI] [PubMed] [Google Scholar]
- 27.Broholm, C. et al. Magnetic excitations in the heavy-fermion superconductorURu2Si2. Phys. Rev. B43, 12809–12822 (1991). [DOI] [PubMed] [Google Scholar]
- 28.Fujita, M. et al. Progress in Neutron Scattering Studies of Spin Excitations in High-Tc Cuprates. J. Phys. Soc. Jpn.81, 011007 (2012). [Google Scholar]
- 29.Goremychkin, E. A. et al. Coherent band excitations in CePd3: A comparison of neutron scattering and ab initio theory. Science359, 186–191 (2018). [DOI] [PubMed] [Google Scholar]
- 30.Fuhrman, W. et al. Interaction Driven Subgap Spin Exciton in the Kondo Insulator SmB6. Phys. Rev. Lett.114, 036401 (2015). [DOI] [PubMed] [Google Scholar]
- 31.Simeth, W. et al. A microscopic kondo lattice model for the heavy fermion antiferromagnet cein3. Nat. Commun.14, 8239 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Rossat-Mignod, J. et al. Inelastic neutron scattering study of cerium heavy fermion compounds. J. Magn. Magn. Mater.76-77, 376–384 (1988). [Google Scholar]
- 33.Butch, N. P. et al. Symmetry of magnetic correlations in spin-triplet superconductor UTe2. npj Quantum Mater.7, 1–6 (2022). [Google Scholar]
- 34.Knafo, W. et al. Low-dimensional antiferromagnetic fluctuations in the heavy-fermion paramagnetic ladder compound UTe2. Phys. Rev. B104, L100409 (2021). [Google Scholar]
- 35.Duan, C. et al. Resonance from antiferromagnetic spin fluctuations for superconductivity in UTe2. Nature600, 636–640 (2021). [DOI] [PubMed] [Google Scholar]
- 36.Raymond, S. et al. Feedback of Superconductivity on the Magnetic Excitation Spectrum of UTe2. J. Phys. Soc. Jpn.90, 113706 (2021). [Google Scholar]
- 37.Duan, C. et al. Incommensurate Spin Fluctuations in the Spin-Triplet Superconductor Candidate UTe2. Phys. Rev. Lett.125, 237003 (2020). [DOI] [PubMed] [Google Scholar]
- 38.Hohenberg, P. C. & Brinkman, W. F. Sum rules for the frequency spectrum of linear magnetic chains. Phys. Rev. B10, 128–131 (1974). [Google Scholar]
- 39.Zaliznyak, I. A. & Lee, S.-H. Magnetic neutron scattering. In Spin, vol. 5 (2005).
- 40.Miyake, A. et al. Metamagnetic transition in heavy fermion superconductor UTe2. J. Phys. Soc. Jpn.88, 063706 (2019). [Google Scholar]
- 41.Ikeda, S. et al. Single Crystal Growth and Magnetic Properties of UTe2. J. Phys. Soc. Jpn.75, 116–118 (2006). [Google Scholar]
- 42.Christovam, D. S. et al. Stabilization of U 5f2 configuration in UTe2 through U 6d dimers in the presence of Te2 chains http://arxiv.org/abs/2402.03852 (2024).
- 43.Thomas, S. M. et al. Evidence for a pressure-induced antiferromagnetic quantum critical point in intermediate-valence UTe2. Sci. Adv.6, eabc8709 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Knafo, W. et al. Incommensurate antiferromagnetism in UTe2 under pressure. arXiv preprint arXiv:2311.05455 (2023).
- 45.Tei, J., Mizushima, T. & Fujimoto, S. Pairing symmetries of multiple superconducting phases in UTe2: Competition between ferromagnetic and antiferromagnetic fluctuations. Phys. Rev. B109, 064516 (2024). [Google Scholar]
- 46.Brandow, B. H. Variational theory of valence fluctuations: Ground states and quasiparticle excitations of the anderson lattice model. Phys. Rev. B33, 215–238 (1986). [DOI] [PubMed] [Google Scholar]
- 47.Brandow, B. H. Finite-temperature behavior of the anderson lattice. Phys. Rev. B37, 250–260 (1988). [DOI] [PubMed] [Google Scholar]
- 48.Auerbach, A., Kim, J. H., Levin, K. & Norman, M. R. Theory of antiferromagnetic correlations and neutron-scattering cross section in heavy-fermion metals. Phys. Rev. Lett.60, 623–626 (1988). [DOI] [PubMed] [Google Scholar]
- 49.Hutanu, V. et al. Low-temperature crystal structure of the unconventional spin-triplet superconductor ute2 from single-crystal neutron diffraction. Acta Crystallogr. Sect. B: Struct. Sci., Cryst. Eng. Mater.76, 137–143 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Mekonen, S. M. et al. Optical investigation of the heavy-fermion normal state in superconducting UTe2. Phys. Rev. B106, 085125 (2022). [Google Scholar]
- 51.Eaton, A. et al. Quasi-2d fermi surface in the anomalous superconductor UTe2. Nat. Commun.15, 223 (2024). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Aoki, D. et al. First observation of the de haas–van alphen effect and fermi surfaces in the unconventional superconductor ute2. J. Phys. Soc. Jpn.91, 083704 (2022). [Google Scholar]
- 53.Miao, L. et al. Low energy band structure and symmetries of UTe2 from angle-resolved photoemission spectroscopy. Phys. Rev. Lett.124, 076401 (2020). [DOI] [PubMed] [Google Scholar]
- 54.Broyles, C. et al. Revealing a 3d fermi surface pocket and electron-hole tunneling in UTe2 with quantum oscillations. Phys. Rev. Lett.131, 036501 (2023). [DOI] [PubMed] [Google Scholar]
- 55.Weinberger, T. et al. Quantum interference between quasi-2d fermi surface sheets in UTe2. Phys. Rev. Lett.132, 266503 (2024). [DOI] [PubMed] [Google Scholar]
- 56.Aishwarya, A. et al. Magnetic-field-sensitive charge density waves in the superconductor UTe2. Nature618, 928–933 (2023). [DOI] [PubMed] [Google Scholar]
- 57.Aishwarya, A. et al. Melting of the charge density wave by generation of pairs of topological defects in ute2. Nat. Phys.20, 1–6 (2024). [Google Scholar]
- 58.Kengle, C. S. et al. Absence of bulk charge density wave order in the normal state of ute2. Nat. Commun.15, 9713 (2024). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Kengle, C. S. et al. Absence of a bulk signature of a charge density wave in hard x-ray measurements of ute2. Phys. Rev. B110, 145101 (2024). [Google Scholar]
- 60.Theuss, F. et al. Absence of a bulk thermodynamic phase transition to a density wave phase in ute2. Phys. Rev. B110, 144507 (2024). [Google Scholar]
- 61.Xu, Y., Sheng, Y. & Yang, Y.-f Quasi-two-dimensional fermi surfaces and unitary spin-triplet pairing in the heavy fermion superconductor UTe2. Phys. Rev. Lett.123, 217002 (2019). [DOI] [PubMed] [Google Scholar]
- 62.Shishidou, T., Suh, H. G., Brydon, P. M. R., Weinert, M. & Agterberg, D. F. Topological band and superconductivity in UTe2. Phys. Rev. B103, 104504 (2021). [Google Scholar]
- 63.Kang, B., Choi, S. & Kim, H. Orbital selective kondo effect in heavy fermion superconductor UTe2. npj Quantum Mater.7, 1–8 (2022). [Google Scholar]
- 64.Yao, S. et al. Controllable growth of centimetre-sized UTe2 single crystals by the chemical vapor transport method. CrystEngComm.24, 6262–6268 (2022). [Google Scholar]
- 65.Lass, J. et al. Commissioning of the novel Continuous Angle Multi-energy Analysis spectrometer at the Paul Scherrer Institut. Rev. Sci. Instrum.94, 023302 (2023). [DOI] [PubMed] [Google Scholar]
- 66.Lass, J., Jacobsen, H., Mazzone, D. G. & Lefmann, K. MJOLNIR: A software package for multiplexing neutron spectrometers. SoftwareX12, 100600 (2020). [Google Scholar]
- 67.Lass, J. MJOLNIRPackage & abehersan. Mjolnirpackage/mjolnir: Version 1.3.2. 10.5281/zenodo.10572477 (2024).
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Data Availability Statement
All data are available upon reasonable request to Thomas Halloran.
All analysis was performed using the MJOLNIR software66,67, and code used in the analysis is available upon reasonable request to Thomas Halloran.