Skip to main content
Nature Communications logoLink to Nature Communications
. 2025 Apr 1;16:3114. doi: 10.1038/s41467-025-58446-6

Altermagnetic ground state in distorted Kagome metal CsCr3Sb5

Chenchao Xu 1,2, Siqi Wu 3, Guo-Xiang Zhi 4, Guanghan Cao 3,5, Jianhui Dai 1,5, Chao Cao 2,3,5,, Xiaoqun Wang 3,5,, Hai-Qing Lin 3,5,
PMCID: PMC11961742  PMID: 40169584

Abstract

The CsCr3Sb5 exhibits superconductivity in close proximity to a density-wave (DW) like ground state at ambient pressure1, however details of the DW is still elusive. Using first-principles density-functional calculations, we found its ground state to be a 4 × 2 altermagnetic spin-density-wave (SDW) at ambient pressure, with an averaged effective moment of  ~ 1.7μB/Cr. The magnetic long range order is coupled to the lattice, generating 4a0 structural modulation. Multiple competing SDW phases are present and energetically close, suggesting strong magnetic fluctuation at finite temperature. The electronic states near Fermi level are dominated by Cr-3d orbitals, and the kagome flat bands are closer to the Fermi level than those in the AV3Sb5 family in paramagnetic state. When external pressure is applied, the energy differences between competing orders and structural modulations are suppressed. Yet, the magnetic fluctuation remains present and important even at high pressure because the high-symmetry kagome lattice is unstable in nonmagnetic phase up to 30 GPa. Our results suggest the crucial role of magnetism to stabilize the crystal structure, under both ambient and high pressure.

Subject terms: Superconducting properties and materials, Magnetic properties and materials


Recent work reported a kagome metal CsCr3Sb5 with frustrated magnetism, density-wave-like order at ambient pressure, and superconductivity under pressure. Using first principles calculations, Xu et al. reveal an altermagnetic spin-density-wave ground state at ambient pressure and phase competition with pressure.

Introduction

Kagome lattice has been an intriguing system2 since its original proposal in 19513. As one of the most frustrating geometry, the ground state of spin-1/2 kagome lattice is widely believed to be quantum spin liquid (QSL), but its nature is still under hot debate410. Once slightly doped, a superconducting phase may emerge from the proposed U(1) Dirac QSL11. For spin-1 kagome lattices, model studies point to a trimerized or simplex valence bond solid as its ground state1214. Beyond the spin models, the electronic and phonon band structure of the kagome lattice exhibits topological flat bands due to the destructive interference between Bloch waves, in addition to the Dirac cones and van Hove singularities (vHS)15,16. Notably, nontrivial physical properties, including kinetic ferromagnetism17, fractional Quantum Hall effect15, and large band-specific diamagnetism18, may emerge from the flat bands. Moreover, near the van Hove fillings, different orders including charge density wave (CDW), chiral spin density wave (cSDW), chiral d-wave superconductivity or even f-wave triplet superconductivity are proposed1922.

Of the many kagome materials, the ternary AV3Sb5 family has drawn much attention since its discovery23. Their much debated CDW order2431 is widely believed to be associated with the nesting between vHSs32,33, and may be chiral and break time-reversal symmetry (TRS)26,3438. The superconductivity was proposed to be unconventional31,39,40, but evidence for conventional BCS superconductivity is also presented4145. Under pressure, the CDW is suppressed, and the superconductivity exhibits a two-dome structure46,47. Nevertheless, despite its rich physics and phenomena, it is widely accepted that AV3Sb5 family is weakly correlated without intrinsic magnetism41,48.

Recently, a chromium-based kagome compound CsCr3Sb5 was reported to be superconducting at pressure p > pc = 4 GPa1. Under ambient pressure, the CsCr3Sb5 compound crystallizes in a hexagonal structure with space group P6/mmm (No. 191) at room temperature, similar to the AV3Sb5 family. Below 55 K, a stripe-like 4a0 structural modulation occurs, possibly due to CDW, with a concurrent SDW phase transition. External pressure continuously suppresses the long-range magnetic order as well as the CDW modulation, which eventually disappears at around 4 GPa where superconductivity appears with the highest Tcmax=6.4 K at pm = 4.2 GPa. Beyond pm, the Tc decreases, forming a dome-like structure. The upper critical field Hc2 is well beyond the Pauli limit around the Tcmax, suggesting an unconventional pairing mechanism. Several imminent questions, therefore, need to be addressed: (1) What is the long-range AFM pattern in this highly frustrated lattice? (2) What is the nature of the ordered magnetism? In particular, is the long-range order local moment or itinerant? (3) How does the electronic structure evolve under pressure?

In this paper, we perform a systematic study on CsCr3Sb5 using first-principles simulations, focusing on its ground state at ambient pressure and the effect of external pressure on the electronic structure. The ground state at ambient pressure is a collinear 4 × 2 altermagnetic spin-density-wave (SDW) order. In addition, multiple SDW orders are energetically close to each other, suggesting strong fluctuation at finite temperatures. Under pressure, the competition among different magnetic phases is enhanced, and a Lifshitz transition due to the metallization of a bonding state is observed at the critical pressure pc. Distinguished from AV3Sb5, the magnetism is crucial to stabilize the lattice structure even at high pressures.

Results and discussion

High symmetry phase under ambient pressure

The high-temperature structure of CsCr3Sb5 without lattice distortion is isostructural to AV3Sb5, where the Cr atoms are arranged into corner-sharing triangles, forming the kagome lattice (Fig. 1a, b). Under ambient pressure, the fully optimized structure using nonmagnetic (NM) DFT calculations yields lattice constant a and all bond lengths smaller than the experimental value (refer to SI for detail). This is in sharp contrast to the AV3Sb5, whose lattice constants and bond lengths are slightly overestimated in NM DFT calculations due to the underestimate of bonding in general gradient approximation (GGA). Such phenomenon, however, is frequently observed in compounds where magnetic fluctuations are important, including iron-pnictides/chalcogenides4951 and other chromium-based superconductors52,53.

Fig. 1. High-temperature crystal structure and low-temperature SDW pattern.

Fig. 1

a Crystal structure of CsCr3Sb5 with P6/mmm symmetry. b Definition of atomic orbitals of Cr atoms. c AF-SOD SDW pattern inspired by inverse star-of-david charge order. The black solid hexagon indicates its Wigner–Seitz (WS) unit cell. d Ground state SDW pattern (failed AF-SOD) of CsCr3Sb5 at ambient pressure. The black solid hexagons and pink arrows indicate the original WS unit cells and the swapped moments compared to the AF-SOD pattern. The thick black dashed lines indicate the gliding-mirror planes. The dot-dashed line shows the 4 × 2 supercell. The blue/red shaded area helps to identify the symmetry between two spin-sublattices. In c, d, the spin-up and spin-down sublattices are marked in red and blue, respectively.

The electronic structure of the high-symmetry phase CsCr3Sb5 at high temperatures is shown in Fig. 2a. For undistorted NM CsCr3Sb5, each Cr atom is surrounded by 4 SbII atoms located directly above/below the center of kagome triangles, as well as 2 SbI atoms located at the center of kagome hexagons. Most of the electronic bands are dominated by the Cr-3d orbitals, except for one dispersive band from pz orbitals of Sb, which hybridizes with dyz/dxz orbitals along Γ-M and Γ-K. Flat bands formed by dxz/dyz orbitals can be identified around 300 meV above the Fermi level, which is considerably closer to the Fermi level compared to those in the AV3Sb5 family54. Other kagome band characteristics, including the vHSs and Dirac cones, can still be identified from the large-scale plots, but they are mostly far from the Fermi level. Interestingly, additional vHS can also be identified around K, most prominently around 270 meV below EF. Along K–M, this band exhibits rather flat dispersion, giving rise to an unambiguous vHS in the DOS (Fig. 2a). Density of state and symmetry analysis suggests it is a bonding state between Cr and SbII atoms. Another nearly flat band due to the dz2-orbitals can also be identified around K, but no vHS can be identified from the DOS. The bands are also very flat along Γ-A, implying the quasi-two-dimensionality of its electronic structure. The NM phonon spectrum of CsCr3Sb5 under ambient pressure using high symmetry kagome structure exhibits enormous imaginary frequencies at almost every q-point along the high symmetry lines (Fig. 3a, b), suggesting the high-symmetry lattice is highly unstable if the system is purely nonmagnetic.

Fig. 2. Electronic structures and Fermi surface of NM state and failed AF-SOD ground state.

Fig. 2

a, b Band structure and density of states (DOS) of CsCr3Sb5 in the high-symmetry NM state at a ambient pressure and b 5 GPa. The size of circles is proportional to the weight of corresponding atomic orbitals. The red circle is used to emphasize the most significant pressure effect on the NM band structure. The Cr-3d atomic orbitals are defined in Fig. 1b. c Band structure of CsCr3Sb5 in the failed AF-SOD ground state at ambient pressure. d Intersection of failed AF-SOD ground state Fermi surfaces at kz = 0 and kz = π. The First Brillouin zone is illustrated with a solid green line. For c, d, the red/blue lines correspond to spin-up/down, respectively; and the high symmetry points are chosen according to the standardized unit cell. e The intersection of NM Fermi surfaces at kz = 0 plane at 5 GPa. The additional Fermi surface pocket around M due to the Lifshitz transition is indicated by green circle.

Fig. 3. Phonon spectrum of NM state and failed AF-SOD ground state.

Fig. 3

a, b The phonon spectrum of CsCr3Sb5 in high-symmetry NM state at a ambient pressure and b 30 GPa. c The phonon spectrum of CsCr3Sb5 in failed AF-SOD ground state at ambient pressure. The corresponding magnetic pattern is illustrated in Fig. 1d. The zero frequency is marked with a green solid line.

Lattice distortion and magnetism under ambient pressure

In order to figure out the magnetic ground state of CsCr3Sb5 under ambient pressure, we first sort out several candidates proposed in previous model studies of spin-1/2 and spin-1 kagome lattices. In particular, collinear patterns including FM, up–up–down (UUD), A-type AFM, as well as noncollinear patterns including 1 × 1 120°-AFM, 3×3 in-plane AFM I/II, 2 × 2 cuboctahedron (cuboc) phases are considered. Inspired by the star-of-David and inverse star-of-David (SOD/ISOD) charge orders54, we have also considered 2 2 × 2 antiferromagnetic SOD patterns, dubbed as in-out SOD and all-in (or all-out) SOD patterns (refer to SI). In addition, a collinear equivalent of such a pattern, namely AF-SOD (Fig. 1c), is also considered. In the AF-SOD pattern, all Cr atoms are coordinated with exactly two nearest antiparallel Cr atoms, such that one can arrange the nearest neighboring AF bonds to form a SOD pattern. It is important here to point out that the AF-SOD pattern is a ferrimagnetic state because the spin-up sublattice is inequivalent to the spin-down sublattice, and there is no symmetry to guarantee an overall 0 net moment. In fact, there is a residue overall total moment of 1.36 μB/cell for the AF-SOD phase in the calculation.

Our initial calculations show that the collinear orders prevail over the noncollinear orders at the same magnetic cell size (Table 1). In addition, patterns with nonzero net moments are energetically less favorable, indicating overall AFM tendency. Therefore, we employed a high-throughput algorithm (please refer to “Methods” for details) to search for the collinear AFM configurations with the lowest energy. Considering the structural modulation with a single Q vector (1/4, 0, 0) observed in experiment1, all 2 × 1, 2 × 2, 4 × 1, 8 × 1 and 4 × 2 collinear AFM orders are searched without considering spin-orbit coupling (SOC). Overall, the lattice constants and bond lengths of fully relaxed structures in most magnetic phases are closer to experimental observations. The averaged magnetic moment is quite robust, around 1.7–1.8 μB/Cr in all calculations even without explicitly considering the on-site Coulomb interactions, except for the in-plane FM cases (1.5 μB/Cr).

Table 1.

Total energies (in meV/f.u.) with respect to NM state and magnetic moment mCr (in μB) of typical magnetic configurations

Size Pattern SG 0 GPa 5 GPa
Etot mCr Etot mCr
1 × 1 × 1 FM P6/mmm −204.2 1.5 −187.3 1.4
UUD P2/m −295.5 1.7 −215.4 1.3
120°-AFM P6/mmm −246.4 1.9 to NM
1 × 1 × 2 A-type AFM P6/mmm −200.8 1.5 −184.2 1.4
3×3 AFM I P3¯1m −297.6 1.8 −238.6 1.6
×1 AFM II P3¯ m1 −254.9 1.8 −179.5 1.6
2 × 2 × 1 all-out SOD P6/mmm −291.5 1.8 −223.2 1.6
in-out SOD P6/mmm −228.9 1.8 −175.8.5 1.6
cuboc converge to 821
AF-SOD P6/mmm −338.4 1.9 (1.7) −275.8 1.8 (1.5)
821 Pm −382.8 1.7 −322.4.5 1.6
4 × 1 × 1 915 Pmm2 −387.3 1.8 −291.9 1.6
4 × 2 × 1 6692437a Pbam −404.0 1.7 −329.7 1.6
6707797 Pmm2 −401.5 1.7 −329.4 1.6
1550757 P2/m −369.9 1.7 −323.5 1.6
5917545 Cmmm −352.8 1.8 −296.4 1.6
4 × 2 NCL converge to 5917545
8 × 1 × 1 2987565 Pmm2 −394.3 1.7 −314.8 1.6

For AF-SOD state, the numbers in (outside) the parenthesis are the moments of Cr on the hexagons (triangles). The magnetic patterns can be found in SI. The SOC was included when comparing the energy between the collinear and noncollinear magnetic configurations.

aThe line in bold font (6692437) is the failed AF-SOD ground state.

The lowest energy magnetic configuration turns out to be a complex pattern (dubbed as “failed AF-SOD”, shown in Fig. 1d). It can be regarded as a decendent of the AF-SOD phase by swapping two pairs of the magnetic moment of next nearest neighboring atoms inside two neighboring 2 × 2 AF-SOD unit cells. The swapped atoms are uniformly distributed across all SOD structures and the staggered magnetic moment ranges from 1.4 to 1.9 μB for each Cr atom, aligning in parallel to minimize the total energy. We have also performed additional calculations with spin–orbit coupling (SOC) included, and the “failed AF-SOD" pattern remains the lowest energy phase. The SOC-included calculations also indicate magnetic anisotropy is small, as the in-plane alignment of moments is only  ~0.6 meV/Cr lower than the out-of-plane alignment, and the in-plane anisotropy appears to be negligible(please refer to SI for details). In such situation, it is recently argued that the spin-space group is more appropriate to describe the symmetry of the phase5557. The corresponding spin-space group is P−1b−1a1mm1 (please refer to SI for details). It is evident that the two spin-sublattices are not connected spatially by either simple translation or inversion, but by {Myz∣(0, 1/2, 0)} or {Mxz∣(1/2, 0, 0)}, a gliding-mirror operation purely on the spatial part. Thus, the ground state at the ambient pressure is altermagnetic58. We note that such altermagnetic ground state is robust under DFT+U calculations with abinitio U and J values determined from constrained random phase approximation (cRPA) calculations (please refer to SI for details).

The formation of magnetic long range order has nontrivial effect on the lattice structure by introducing structural modulations. In our calculations, all lowest energy configurations breaks the threefold rotational symmetry, and both the Cr and the SbII positions are strongly modulated. In particular, the failed AF-SOD ground state is formed by 2 4 × 1 stripes connected by gliding-mirror operation. Similarly, the second lowest (indexed with 6707797) naturally consists of 2 4 × 1 stripe order connected by translation only, which is also consistent with experimentally observed 4 × 1 charge order at low temperature. The calculated phonon spectrum of the failed AF-SOD phase is free of imaginary phonon frequency (Fig. 3c), indicating the dynamic stability of this phase. However, the phonon spectrum of the corresponding fully-relaxed non-magnetic structure is not (Fig. S3). Thus, the magnetism stabilizes the lattice structure in CsCr3Sb5.

The spin-degeneracy of the failed AF-SOD ground state is generally lifted (Fig. 2c) due to the lack of inversion or simple translation symmetry between sublattices (Fig. 1d). Although the quasi-Kramer’s degeneracy protected by SSG symmetries can still be identified along certain high symmetry lines55, the spin splitting is evident along Γ–S, with the largest splitting  ~80 meV close to the Fermi level. The splitting suggests separated spin-up/spin-down Fermi surface sheets (Fig. 2d, refer to SI for more details). These features may be verified through spin-resolved ARPES measurements to confirm the altermagnetic ground state59,60, if the measurement can be constrained within a single domain. In addition, our calculations show that moderate anomalous Hall conductivity may be present if the direction of magnetic moments is in-plane and the field direction is properly aligned (refer to SI for detail)61,62. It is worth noting that the flat-bands and vHS features are less obvious in the magnetic ground state compared to the undistorted NM CsCr3Sb5, due to the strong structure distortion of the altermagnetic ground state (refer to SI for detail).

To understand the origin of the magnetism, we have tried to fit the DFT total energies of the 20 lowest magnetic configurations to a Heisenberg spin model with the bilinear exchange interactions. However, such fitting is not successful up to 6th-nearest-neighboring exchange interactions (~9.3 Å apart, please refer to SI for details). Considering the moments on Cr atoms have a wide distribution in all collinear magnetic configurations, our results suggest that the Fermi-surface-related itinerant magnetism may play an important role in the formation of the SDW order.

Pressure effect

With the knowledge of ambient pressure, we now investigate the pressure effect. The total energy, geometry, and Cr magnetic moment of the leading magnetic instabilities under different pressures are also shown in Table 1. Under pressure, the magnitude of magnetic moment on each Cr atom mCr is only slightly affected, but the total energy difference between the competing magnetic phases ΔE is significantly reduced. In particular, ΔE between the ground state and lowest energy competing phases of 2 × 2, 4 × 2, 4 × 1, and 8 × 1 is reduced to less than 1 meV/f.u. We argue that the long-range magnetic order would be suppressed at high pressure, if the quantum fluctuations are fully considered. Nevertheless, the magnetism remains important even at high pressure, as the phonon calculations indicate the structure instability is robust under pressure up to 30 GPa in NM state (Fig. 3b). Experimentally, however, the structural transition occurs at relatively low temperatures (~55 K) at ambient pressure and is quickly suppressed by applying external pressure at 5 GPa. Therefore, we conclude that the electronic correlations and magnetic fluctuations are crucial in the current system of interest, even at high pressure/temperatures where long-range magnetic orderings are suppressed. As a result, if the static altermagnetic long-range order is continuously suppressed, superconductivity with exotic pairing symmetry is possible due to the dynamic spin fluctuations58,63,64. Furthermore, the external pressure has a nontrivial effect on the electronic structure as well. Most remarkably, the bonding state originally forming the vHS around K becomes dispersive along M–K under pressure(red circle in Fig. 2b), and the vHS gradually vanishes. Around 5 GPa, a Lifshitz transition occurs, and the aforementioned bonding state crosses the Fermi level (green circle in Fig. 2e). Such metallization of the bonding state may also affect the superconductivity65,66.

In conclusion, we have performed a systematic study of the electronic structure, magnetism, and lattice stability of CsCr3Sb5. The ground state of CsCr3Sb5 at the ambient pressure is found to be 4 × 2 collinear altermagnetic SDW-type order, with considerable itinerant magnetism component. Under high pressure, the energy differences between competing orders are significantly suppressed, suggesting enhanced magnetic fluctuations. In addition, the NM phonon calculations indicate the high-symmetry kagome structure is unstable at low temperatures up to 30 GPa. Similarly, the NM phonon spectrum is also highly unstable even for the distorted structure of the ground state. In comparison, the distorted structure becomes stable after the magnetism is considered. These results suggest that magnetic fluctuation is extremely important and may couple to the lattice dynamics in CsCr3Sb5. Therefore, a simple electron-phonon coupling-based BCS mechanism is highly unlikely in this compound.

Methods

Electronic and phonon band structure

The calculations were performed based on density functional theory (DFT) with VIENNA ABINITIO SIMULATION PACKAGE (VASP)67,68 and cross-checked with Quantum ESPRESSO (QE)69. The energy cutoff of the plane-wave basis was 450 eV, and a Γ-centered 12 × 12 × 6 k-point mesh was employed in the self-consistent calculations with VASP. QE calculations were performed with ultrasoft-pseudopotentials with energy cutoffs 64 Ry and 720 Ry for the wavefunction and augmentation charge, respectively. In all calculations, the PBEsol approximation70 was used, and spin-orbit coupling (SOC) was not included unless otherwise stated. The lattice constants and atomic coordinates were fully relaxed until the force on each atom was less than 1 meV/Å and internal stress was less than 0.1 kbar. After full structure optimization, the phonon calculations were also performed with 4 × 4 × 3 q-point mesh for high-symmetry NM state using density-functional perturbation theory (DFPT) as implemented in QE. For the phonon spectrum of the failed AF-SOD state, the finite displacement method using 1 × 2 × 2 supercells is employed, as implemented in PHONOPY71.

High-throughput Search for magnetic configuration

The high-throughput calculations were performed for all possible collinear magnetic patterns up to 4 × 2 supercell.

  1. First, all possible magnetic configurations are indexed by assigning a N-bit binary integer to each configuration, where N is the number of the Cr sites within the supercell. For example, 7=000111_2 in a 2 × 1 configuration, it means the first 3 sites are spin-down and last 3 sites spin-up. The configurations with non-zero net moments are then discarded. Then, we search for the AFM symmetry connecting up/down-sublattices within each configuration and eliminate those without such symmetry. In addition, only 1 of all symmetrically equivalent configurations are kept. After this procedure, there are a total of 1911 configurations. The 8 × 1 and 4 × 2 supercells generate 1888 configurations, whereas the rest 3 generates 23 configurations.

  2. Secondly, for 2 × 1, 2 × 2, and 4 × 1 configurations, both static total energy calculations on perfect kagome lattice and full geometry relaxations were performed. Therefore, for each configuration, we obtain both E0 and Er. The former represents magnetic configuration energy difference only, while the latter is actual total energy. We note that for all these configurations, the full relaxation yields an energy difference order of 200 meV.

  3. Thirdly, for 4 × 2 and 8 × 1 configurations, a static total energy calculation on perfect kagome lattice is performed first. Then, the configurations with E0 over 200 meV higher than the lowest E0 are screened out. The remaining configurations are fully relaxed to obtain Er.

  4. Finally, the configurations with the lowest Er are verified with dense K-mesh and SOC calculations.

Anomalous Hall conductivity

For the anomalous Hall conductivity calculations, we include SOC in the calculations. The DFT results were then fitted to a tight-binding (TB) Hamiltonian with a maximally projected Wannier function method72. Thirty atomic orbitals, including Cr-3d and Sb-5p, were chosen to construct the TB model Hamiltonian. The resulting Hamiltonian was symmetrized using WannSymm code73. With the symmetrized TB Hamiltonian, the intrinsic anomalous Hall conductivity (AHC) is calculated following the Kubo formula by integration of Berry curvature over the Brillouin zone74,75:

Ωnαβ(k)=2e2mnImnvαmmvβnEnEm2 1
σαβ=e2knd3k(2π)3fknΩnαβ(k) 2

The implementation in WannierTools76 was employed to calculated the AHC. To fix collinear moments in [100], [010], or [001] directions, SOC was applied. The Berry curvature was calculated with a 100 × 100 × 100 Γ-centered kmesh, with which the convergence of σαβ is achieved.

Constrained random-phase calculation

We performed the constrained random-phase approximation (cRPA) method using VASP to calculate the effective interaction parameter U and J. With the 30 atomic-like Wannier orbitals, we constructed the target space with Cr-3d orbitals:

wiσ=1NknkTinσ(k)ψnkσ 3

The Coulomb interaction matrix elements are written as:

Uijklσσ=1Nk3kkqn1n2n3n4Tin1*(k)Tjn2(kq)un1kei(q+G)run2kqUGG(q)×un3kqeiqGrun4kTkn3*kqTln4k 4

where the effective Coulomb kernel UGG(q) can be calculated as

UGGσσ(q,iω)=δGGχGGσσ(q,iω)χ~GGσσ(q,iω)VGG(q)1VGG(q) 5

χ~GGσσ(q,iω) denotes the polarizability within the target space. The screening effect due to the target polarizability is excluded from the total screening effects in the RPA method. Using projector method77, the target polarizability is obtained with:

χ~G,Gσ(q,iω)=1Nknnkfnkfnkqϵnkϵnkqiω×m1m2Pm1n*σ(k)um1kσei(G+q)rum2kqσPm2nσ(kq)×m1m2Pm2n*σ(kq)um2kqσeiGqrum1kσPm1nσ(k) 6

where the target projectors Pmnσ(k) for Cr-3d orbitals are defined as

Pmnσ(k)=iTTim*σ(k)Tinσ(k) 7

And the static limit (ω = 0) values are taken as the effective interaction parameters. The large bare interactions (Uiiiibare~18 eV, Uiijjbare~17 eV, Uijijbare~0.6 eV) are strongly screened to yield U = 1.0 eV and J = 0.4 eV.

Supplementary information

Source data

Source Data (28KB, zip)

Acknowledgements

The authors are grateful to Xiaofeng Xu, Jinke Bao, Yi Liu, Jiangfan Wang and Dexi Shao for stimulating discussions. C. C. acknowledges support from the National Key R&D Program of China (Nos. 2024YFA1408303 and 2022YFA1402202) and the National Natural Science Foundation of China (Nos. 12350710785 and 12274364). X. W. and H.-Q. L. acknowledge support from the National Key R&D Program of China (No. 2022YFA140271). G.-H. C. acknowledges the support from the Key R&D Program of Zhejiang Province (2021C01002). C. X. acknowledges the support from the National Natural Science Foundation of China (No. 12304175). G.-X. Z. acknowledge the support from the Zhejiang Provincial Natural Science Foundation of China (LQ23A040014). The calculations were performed on clusters at the High Performance Computing Cluster at Center of Correlated Matters Zhejiang University and High Performance Computing Center at Hangzhou Normal University.

Author contributions

C. C., G.-H. C., and C. X. initiated this work; C. C., X. W., and H.-Q. L. supervised the research. C. X., S. W., and C. C. performed the calculations; C. X. and C. C. were responsible for the data analysis; C. X. and C. C. drafted the paper with input from all authors. C. X., S. W., G.-X. Z., G.-H. C., J. D., C. C., X. W., and H.-Q. L. participated in the discussion and revised the paper.

Peer review

Peer review information

Nature Communications thanks the anonymous reviewers for their contribution to the peer review of this work. A peer review file is available.

Data availability

The data used to generate Figs. 2 and 3 are available via Figshare. The mcif files of the configurations in Table 1 are also available as source data. Any additional data relevant to the findings of this study are available from the corresponding author upon request. Source data are provided with this paper.

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.

Contributor Information

Chao Cao, Email: ccao@zju.edu.cn.

Xiaoqun Wang, Email: xiaoqunwang@zju.edu.cn.

Hai-Qing Lin, Email: hqlin@zju.edu.cn.

Supplementary information

The online version contains supplementary material available at 10.1038/s41467-025-58446-6.

References

  • 1.Liu, Y. et al. Superconductivity under pressure in a chromium-based kagome metal. Nature632, 1032–1037 (2024). [DOI] [PubMed] [Google Scholar]
  • 2.Yin, J.-X., Lian, B. & Hasan, M. Z. Topological kagome magnets and superconductors. Nature612, 647–657 (2022). [DOI] [PubMed] [Google Scholar]
  • 3.Syôzi, I. Statistics of kagomé Lattice. Prog. Theor. Phys.6, 306–308 (1951). [Google Scholar]
  • 4.Yan, S., Huse, D. A. & White, S. R. Spin-liquid ground state of the S = 1/2 kagome Heisenberg antiferromagnet. Science332, 1173–1176 (2011). [DOI] [PubMed] [Google Scholar]
  • 5.Messio, L., Bernu, B. & Lhuillier, C. Kagome antiferromagnet: a chiral topological spin liquid? Phys. Rev. Lett.108, 207204 (2012). [DOI] [PubMed] [Google Scholar]
  • 6.Depenbrock, S., McCulloch, I. P. & Schollwöck, U. Nature of the spin-liquid ground state of the S = 1/2 Heisenberg model on the kagome lattice. Phys. Rev. Lett.109, 067201 (2012). [DOI] [PubMed] [Google Scholar]
  • 7.Xie, Z. Y. et al. Tensor renormalization of quantum many-body systems using projected entangled simplex states. Phys. Rev. X4, 011025 (2014). [Google Scholar]
  • 8.Gong, S.-S., Zhu, W., Balents, L. & Sheng, D. N. Global phase diagram of competing ordered and quantum spin-liquid phases on the kagome lattice. Phys. Rev. B91, 075112 (2015). [Google Scholar]
  • 9.Bieri, S., Messio, L., Bernu, B. & Lhuillier, C. Gapless chiral spin liquid in a kagome Heisenberg model. Phys. Rev. B92, 060407 (2015). [Google Scholar]
  • 10.Liao, H. J. et al. Gapless spin-liquid ground state in the S = 1 / 2 kagome antiferromagnet. Phys. Rev. Lett.118, 137202 (2017). [DOI] [PubMed] [Google Scholar]
  • 11.Jiang, Y.-F., Yao, H. & Yang, F. Possible superconductivity with a Bogoliubov Fermi surface in a lightly doped kagome U(1) spin liquid. Phys. Rev. Lett.127, 187003 (2021). [DOI] [PubMed] [Google Scholar]
  • 12.Cai, Z., Chen, S. & Wang, Y. Spontaneous trimerization in two-dimensional antiferromagnets. J. Phys.21, 456009 (2009). [DOI] [PubMed] [Google Scholar]
  • 13.Changlani, H. J. & Läuchli, A. M. Trimerized ground state of the spin-1 Heisenberg antiferromagnet on the kagome lattice. Phys. Rev. B91, 100407 (2015). [Google Scholar]
  • 14.Liu, T., Li, W., Weichselbaum, A., von Delft, J. & Su, G. Simplex valence-bond crystal in the spin-1 kagome Heisenberg antiferromagnet. Phys. Rev. B91, 060403 (2015). [Google Scholar]
  • 15.Tang, E., Mei, J.-W. & Wen, X.-G. High-temperature fractional quantum hall states. Phys. Rev. Lett.106, 236802 (2011). [DOI] [PubMed] [Google Scholar]
  • 16.Lin, Z. et al. Flatbands and emergent ferromagnetic ordering in Fe3Sn2 kagome lattices. Phys. Rev. Lett.121, 096401 (2018). [DOI] [PubMed] [Google Scholar]
  • 17.Pollmann, F., Fulde, P. & Shtengel, K. Kinetic ferromagnetism on a kagome lattice. Phys. Rev. Lett.100, 136404 (2008). [DOI] [PubMed] [Google Scholar]
  • 18.Yin, J.-X. et al. Negative flat band magnetism in a spin–orbit-coupled correlated kagome magnet. Nat. Phys.15, 443–448 (2019). [Google Scholar]
  • 19.Ko, W.-H., Lee, P. A. & Wen, X.-G. Doped kagome system as exotic superconductor. Phys. Rev. B79, 214502 (2009). [Google Scholar]
  • 20.Yu, S.-L. & Li, J.-X. Chiral superconducting phase and chiral spin-density-wave phase in a Hubbard model on the kagome lattice. Phys. Rev. B85, 144402 (2012). [Google Scholar]
  • 21.Kiesel, M. L., Platt, C. & Thomale, R. Unconventional fermi surface instabilities in the kagome Hubbard model. Phys. Rev. Lett.110, 126405 (2013). [DOI] [PubMed] [Google Scholar]
  • 22.Wang, W.-S., Li, Z.-Z., Xiang, Y.-Y. & Wang, Q.-H. Competing electronic orders on kagome lattices at van Hove filling. Phys. Rev. B87, 115135 (2013). [Google Scholar]
  • 23.Ortiz, B. R. et al. New kagome prototype materials: discovery of KV3Sb5, RbV3Sb5, and CsV3Sb5. Phys. Rev. Mater.3, 094407 (2019). [Google Scholar]
  • 24.Ortiz, B. R. et al. CsV3Sb5: a Inline graphic topological kagome metal with a superconducting ground state. Phys. Rev. Lett.125, 247002 (2020). [DOI] [PubMed] [Google Scholar]
  • 25.Li, H. et al. Observation of unconventional charge density wave without acoustic phonon anomaly in kagome superconductors AV3Sb5 (A= Rb, Cs). Phys. Rev. X11, 031050 (2021). [Google Scholar]
  • 26.Wang, Z. et al. Electronic nature of chiral charge order in the kagome superconductor CsV3Sb5. Phys. Rev. B104, 075148 (2021). [Google Scholar]
  • 27.Denner, M. M., Thomale, R. & Neupert, T. Analysis of charge order in the kagome metal AV3Sb5 (A = K, Rb, Cs). Phys. Rev. Lett.127, 217601 (2021). [DOI] [PubMed] [Google Scholar]
  • 28.Xu, H.-S. et al. Multiband superconductivity with sign-preserving order parameter in kagome superconductor CsV3Sb5. Phys. Rev. Lett.127, 187004 (2021). [DOI] [PubMed] [Google Scholar]
  • 29.Zhao, H. et al. Cascade of correlated electron states in the kagome superconductor CsV3Sb5. Nature599, 216–221 (2021). [DOI] [PubMed] [Google Scholar]
  • 30.Li, H. et al. Rotation symmetry breaking in the normal state of a kagome superconductor KV3Sb5. Nat. Phys.18, 265–270 (2022). [Google Scholar]
  • 31.Chen, H. et al. Roton pair density wave in a strong-coupling kagome superconductor. Nature599, 222–228 (2021). [DOI] [PubMed] [Google Scholar]
  • 32.Liu, Z. et al. Charge-density-wave-induced bands renormalization and energy gaps in a kagome superconductor RbV3Sb5. Phys. Rev. X11, 041010 (2021). [Google Scholar]
  • 33.Kang, M. et al. Twofold van Hove singularity and origin of charge order in topological kagome superconductor CsV3Sb5. Nat. Phys.18, 301–308 (2022). [Google Scholar]
  • 34.Jiang, Y.-X. et al. Unconventional chiral charge order in kagome superconductor KV3Sb5. Nat. Mater.20, 1353–1357 (2021). [DOI] [PubMed] [Google Scholar]
  • 35.Feng, X., Jiang, K., Wang, Z. & Hu, J. Chiral flux phase in the kagome superconductor AV3Sb5. Sci. Bull.66, 1384–1388 (2021). [DOI] [PubMed] [Google Scholar]
  • 36.Yu, L. et al. Evidence of a hidden flux phase in the topological kagome metal CsV3Sb5. Preprint at arXiv10.48550/arXiv.2107.10714 (2021).
  • 37.Shumiya, N. et al. Intrinsic nature of chiral charge order in the kagome superconductor RbV3Sb5. Phys. Rev. B104, 035131 (2021). [Google Scholar]
  • 38.Mielke, C. et al. Time-reversal symmetry-breaking charge order in a kagome superconductor. Nature602, 245–250 (2022). [DOI] [PubMed] [Google Scholar]
  • 39.Wu, X. et al. Nature of unconventional pairing in the kagome superconductors AV3Sb5 (A = K, Rb, Cs). Phys. Rev. Lett.127, 177001 (2021). [DOI] [PubMed] [Google Scholar]
  • 40.Zhao, C. C. et al. Nodal superconductivity and superconducting domes in the topological kagome metal CsV3Sb5. Preprint at arXiv10.48550/arXiv.2102.08356 (2021).
  • 41.Zhao, J., Wu, W., Wang, Y. & Yang, S. A. Electronic correlations in the normal state of the kagome superconductor KV3Sb5. Phys. Rev. B103, L241117 (2021). [Google Scholar]
  • 42.Duan, W. et al. Nodeless superconductivity in the kagome metal CsV3Sb5. Sci. China Phys. Mech. Astron.64, 107462 (2021). [Google Scholar]
  • 43.Mu, C. et al. S-wave superconductivity in kagome metal CsV3Sb5 revealed by 121/123Sb NQR and 51V NMR measurements. Chin. Phys. Lett.38, 077402 (2021). [Google Scholar]
  • 44.Wang, C., Jia, Y., Zhang, Z. & Cho, J.-H. Phonon-mediated s-wave superconductivity in the kagome metal CsV3Sb5 under pressure. Phys. Rev. B108, L060503 (2023). [Google Scholar]
  • 45.Zhong, Y. et al. Testing electron–phonon coupling for the superconductivity in kagome metal CsV3Sb5. Nat. Commun.14, 1945 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Du, F. et al. Pressure-induced double superconducting domes and charge instability in the kagome metal KV3Sb5. Phys. Rev. B103, L220504 (2021). [Google Scholar]
  • 47.Zhu, C. C. et al. Double-dome superconductivity under pressure in the V-based kagome metals AV3Sb5 (A = Rb and K). Phys. Rev. B105, 094507 (2022). [Google Scholar]
  • 48.Kenney, E. M., Ortiz, B. R., Wang, C., Wilson, S. D. & Graf, M. J. Absence of local moments in the kagome metal KV3Sb5 as determined by muon spin spectroscopy. J. Phys.33, 235801 (2021). [DOI] [PubMed] [Google Scholar]
  • 49.Cao, C., Hirschfeld, P. J. & Cheng, H.-P. Proximity of antiferromagnetism and superconductivity in LaFeAsO1−xFx: effective Hamiltonian from ab initio studies. Phys. Rev. B77, 220506 (2008). [Google Scholar]
  • 50.Singh, D. J. & Du, M.-H. Density functional study of LaFeAsO1−xFx: a low carrier density superconductor near itinerant magnetism. Phys. Rev. Lett.100, 237003 (2008). [DOI] [PubMed] [Google Scholar]
  • 51.Mazin, I. I., Johannes, M. D., Boeri, L., Koepernik, K. & Singh, D. J. Problems with reconciling density functional theory calculations with experiment in ferropnictides. Phys. Rev. B78, 085104 (2008). [Google Scholar]
  • 52.Jiang, H., Cao, G. & Cao, C. Electronic structure of quasi-one-dimensional superconductor K2Cr3As3 from first-principles calculations. Sci. Rep.5, 16054 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Wu, X.-X., Le, C.-C., Yuan, J., Fan, H. & Hu, J.-P. Magnetism in quasi-one-dimensional A2Cr3As3 (A=K,Rb) superconductors. Chin. Phys. Lett.32, 057401 (2015). [Google Scholar]
  • 54.Tan, H., Liu, Y., Wang, Z. & Yan, B. Charge density waves and electronic properties of superconducting kagome metals. Phys. Rev. Lett.127, 046401 (2021). [DOI] [PubMed] [Google Scholar]
  • 55.Liu, P., Li, J., Han, J., Wan, X. & Liu, Q. Spin-group symmetry in magnetic materials with negligible spin-orbit coupling. Phys. Rev. X12, 021016 (2022). [Google Scholar]
  • 56.Šmejkal, L., Sinova, J. & Jungwirth, T. Beyond conventional ferromagnetism and antiferromagnetism: a phase with nonrelativistic spin and crystal rotation symmetry. Phys. Rev. X12, 031042 (2022). [Google Scholar]
  • 57.Chen, X. et al. Enumeration and representation theory of spin space groups. Phys. Rev. X14, 031038 (2024). [Google Scholar]
  • 58.Šmejkal, L., Sinova, J. & Jungwirth, T. Emerging research landscape of altermagnetism. Phys. Rev. X12, 040501 (2022). [Google Scholar]
  • 59.Krempaský, J. et al. Altermagnetic lifting of Kramers spin degeneracy. Nature626, 517–522 (2024). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Lee, S. et al. Broken kramers degeneracy in altermagnetic MnTe. Phys. Rev. Lett.132, 036702 (2024). [DOI] [PubMed] [Google Scholar]
  • 61.Šmejkal, L., Hellenes, A. B., González-Hernández, R., Sinova, J. & Jungwirth, T. Giant and tunneling magnetoresistance in unconventional collinear antiferromagnets with nonrelativistic spin-momentum coupling. Phys. Rev. X12, 011028 (2022). [Google Scholar]
  • 62.Šmejkal, L., MacDonald, A. H., Sinova, J., Nakatsuji, S. & Jungwirth, T. Anomalous hall antiferromagnets. Nat. Rev. Mater.7, 482–496 (2022). [Google Scholar]
  • 63.Mazin, I. I. Notes on altermagnetism and superconductivity. Preprint at arXiv10.48550/arXiv.2203.05000 (2022).
  • 64.Zhang, S.-B., Hu, L.-H. & Neupert, T. Finite-momentum Cooper pairing in proximitized altermagnets. Nat. Commun.15, 1801 (2024). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Gao, M., Yan, X.-W., Lu, Z.-Y. & Xiang, T. Phonon-mediated high-temperature superconductivity in the ternary borohydride KB2H8 under pressure near 12 GPa. Phys. Rev. B104, L100504 (2021). [Google Scholar]
  • 66.Sun, H. et al. Signatures of superconductivity near 80 k in a nickelate under high pressure. Nature621, 493–498 (2023). [DOI] [PubMed] [Google Scholar]
  • 67.Kresse, G. & Hafner, J. Ab initio molecular dynamics for liquid metals. Phys. Rev. B47, 558–561 (1993). [DOI] [PubMed] [Google Scholar]
  • 68.Kresse, G. & Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B59, 1758–1775 (1999). [Google Scholar]
  • 69.Giannozzi, P. et al. QUANTUM ESPRESSO: a modular and open-source software project for quantum simulations of materials. J. Phys.21, 395502 (2009). [DOI] [PubMed] [Google Scholar]
  • 70.Perdew, J. P. et al. Restoring the density-gradient expansion for exchange in solids and surfaces. Phys. Rev. Lett.100, 136406 (2008). [DOI] [PubMed] [Google Scholar]
  • 71.Togo, A., Chaput, L., Tadano, T. & Tanaka, I. Implementation strategies in phonopy and phono3py. J. Phys. Condens. Matter35, 353001 (2023). [DOI] [PubMed] [Google Scholar]
  • 72.Pizzi, G. et al. Wannier90 as a community code: new features and applications. J. Phys.32, 165902 (2020). [DOI] [PubMed] [Google Scholar]
  • 73.Zhi, G.-X., Xu, C., Wu, S.-Q., Ning, F. & Cao, C. Wannsymm: a symmetry analysis code for wannier orbitals. Comput. Phys. Commun.271, 108196 (2022). [Google Scholar]
  • 74.Wang, X., Yates, J. R., Souza, I. & Vanderbilt, D. Ab initio calculation of the anomalous hall conductivity by wannier interpolation. Phys. Rev. B74, 195118 (2006). [Google Scholar]
  • 75.Yao, Y. et al. First principles calculation of anomalous hall conductivity in ferromagnetic BCC Fe. Phys. Rev. Lett.92, 037204 (2004). [DOI] [PubMed] [Google Scholar]
  • 76.Wu, Q., Zhang, S., Song, H.-F., Troyer, M. & Soluyanov, A. A. Wanniertools: an open-source software package for novel topological materials. Comput. Phys. Commun.224, 405–416 (2018). [Google Scholar]
  • 77.Kaltak, M. Merging GW With DMFT. Ph.D. thesis. (Universität Wien, Wien, 2015) Dissertation. 10.25365/thesis.38099.

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Source Data (28KB, zip)

Data Availability Statement

The data used to generate Figs. 2 and 3 are available via Figshare. The mcif files of the configurations in Table 1 are also available as source data. Any additional data relevant to the findings of this study are available from the corresponding author upon request. Source data are provided with this paper.


Articles from Nature Communications are provided here courtesy of Nature Publishing Group

RESOURCES