Abstract
Altermagnetism, a recently discovered class of magnetic order characterized by vanishing net magnetization and spin-splitting band structures, has garnered significant research attention. In this work, we introduce a novel two-dimensional system that exhibits g-wave altermagnetism and undergoes a strain-induced transition from g-wave to d-wave altermagnetism. This system can be realized in an unconventional monolayer Cairo pentagonal lattice, for which we present a realistic tight-binding model that incorporates both magnetic and non-magnetic sites. Furthermore, we demonstrate that non-trivial band topology can emerge in this system by breaking the symmetry that protects the spin-polarized nodal points. Finally, ab initio calculations on several candidate materials, such as FeS2 and Nb2FeB2, which exhibit symmetry consistent with the proposed tight-binding Hamiltonian, are also presented. These findings open new avenues for exploring spintronic devices based on altermagnetic systems.
Similar content being viewed by others
Introduction
Recently, altermagnetism has attracted growing research interest due to its unconventional behavior, which is distinct from traditional collinear ferromagnetism and antiferromagnetism1,2,3,4,5,6,7,8,9. In altermagnet, magnetic moments form an antiferromagnetic-like order with zero net magnetization, while it exhibits energy splitting between states with opposite spins, similar to ferromagnets. A variety of materials have been proposed or confirmed to exhibit altermagnetism through first-principles calculations and experiments, including RuO210,11,12, MnTe13,14,15,16, FeSb217, and CrSb18,19. The unique electronic structure in altermagnets leads to numerous interesting effects and potential applications, such as spin-splitting torque phenomena20,21, unconventional superconductivity22,23,24,25,26,27,28,29, and distinct variants of Hall effect30,31,32,33,34,35,36,37,38. In addition, altermagnets hold great promise for spintronic applications due to their large spin-splitting and robustness against magnetic field perturbations2,39,40.
One of the most fundamental questions in the study of altermagnetic materials is how to understand the origin of this emergent behavior, which is essential for predicting new physical properties and exploring potential applications. At the microscopic level, several mechanisms for the emergence of altermagnetism have been proposed, including the interplay between magnetic and nonmagnetic atoms and the anisotropic ordering of local orbitals24,41. Thus, identifying realistic tight-binding models from a microscopic perspective is crucial for fully understanding this unconventional phenomenon. While several studies have explored altermagnetism using effective models, only a few have investigated the microscopic origin of altermagnetic properties24,31,41,42,43.
In our work, we investigate the microscopic theory for the origin of altermagnetism on a novel two-dimensional pentagonal structure44, known as the Cairo Pentagon. We uncover the mystery of altermagnetism on the Cairo pentagonal lattice at a microscopic level by constructing a simple but realistic tight-binding model containing both magnetic and non-magnetic atoms, and demonstrate that the interplay between them plays a key role in the origin of altermagnetism. Another remarkable feature of the pentagonal lattice is its sensitivity to strain, which induces strong in-plane anisotropy45,46,47. This anisotropy alters the spin-lattice symmetry, thereby impacting the structure of spin-splitting. Calculated band structure using the tight-binding model shows that this strong dependence on strain will lead to a transition between g-wave and d-wave altermagnetism. In addition, we find that this pentagonal altermagnet hosts symmetry protected polarized nodal points37,38,48,49, and we give examples demonstrating how breaking this symmetry could gap out these nodal points and lead to non-trivial topological bands. Finally, we examine two candidate materials FeS2 and Nb2FeB2 through ab initio calculations, the results successfully reproduce the altermagnetism and its transition under strain as we expect.
Our results represent the first demonstration of strain tuning to achieve different altermagnetic orders in realistic systems, along with a microscopic understanding of the underlying mechanisms. This paves a new avenue for the design and application of strain-tuned spintronic devices based on altermagnetism.
Results
Cairo pentagonal lattice structure and model Hamiltonian
We consider a Cairo pentagonal lattice with space group P4/mbm, shown in Fig. 1a. Due to the presence of non-magnetic sites, the opposite spin sublattices can not be mapped to each other by the combination of time-reversal with translation or inversion, which makes it an ideal platform to investigate the relationship between atomic interplay and altermagnetism.
a A schematics of the pentagonal lattice with space group P4/mbm and the tight-binding model parameters. The magnetic sites with spin-up, spin-down, and non-magnetic (nm) sites are represented by red, blue, and gray colors. t, tm, tnm,1 and tnm,2 represent the hopping strengths between the nearest-neighbor magnetic and non-magnetic sites, the nearest-neighbor magnetic sites, the first nearest-neighbor non-magnetic sites, and the second nearest-neighbor non-magnetic sites, respectively. The reflection planes corresponding to the mirror operators \({{\mathcal{M}}}_{x}\), \({{\mathcal{M}}}_{y}\), \({{\mathcal{M}}}_{xy}\) and \({{\mathcal{M}}}_{\bar{x}y}\) are represented by dashed lines. b A schematics of the same pentagonal lattice under the diagonal strain along xy or \(\bar{x}y\). c, d Visualization of the higher energy bands from the pair of spin-split bands on this lattice, without and with strain, in the first Brillouin zone. The bands with spin up and down are represented by red and blue, respectively. The yellow lines represent the spin-degenerate nodal lines that cross the Γ point. A number of high-symmetry momentum points are highlighted.
The lattice is centrosymmetric due to the four-fold rotation symmetry C4z. In one primitive unit cell, there are two magnetic atoms with opposite collinear spins and four non-magnetic atoms. The presence of non-magnetic sites breaks symmetry \(\{{C}_{2\perp }| | {\mathcal{P}}{\boldsymbol{\tau }}\}\) while preserving symmetries \(\{{C}_{2\perp }| | {{\mathcal{M}}}_{x}(\frac{1}{2},\frac{1}{2})\}\), \(\{{C}_{2\perp }| | {{\mathcal{M}}}_{y}(\frac{1}{2},\frac{1}{2})\}\), \(\{{C}_{2\perp }| | {{\mathcal{M}}}_{xy}(\frac{1}{2},\frac{1}{2})\}\) and \(\{{C}_{2\perp }| | {{\mathcal{M}}}_{\bar{x}y}(\frac{1}{2},\frac{1}{2})\}\) relating two magnetic sublattices, where C2⊥ is the 180° rotation operator around an axis perpendicular to the spins, \({\mathcal{P}}{\boldsymbol{\tau }}\) is an operator combined by the inversion \({\mathcal{P}}\) with any translation τ, \({{\mathcal{M}}}_{x}\), \({{\mathcal{M}}}_{y}\), \({{\mathcal{M}}}_{xy}\) and \({{\mathcal{M}}}_{\bar{x}y}\) are the mirror operators about the x axis, y axis, and the diagonals xy and \(\bar{x}y\), and \(\left(\frac{1}{2},\frac{1}{2}\right)\) is a translation of \(\frac{1}{2}{{\bf{a}}}_{1}+\frac{1}{2}{{\bf{a}}}_{2}\), where a1 and a2 are primitive vectors. Due to symmetry considerations, the pair of electronic states with opposite spins split at general k points but remain degenerate along ky = 0, kx = 0, kx = − ky and kx = ky in the first Brillouin zone, as shown in Fig. 1c. This unconventional spin-splitting ensures that the magnetic pentagonal crystal is classified as an altermagnet.
To model the electronic structure in this pentagonal lattice, we consider a tight-binding Hamiltonian including both magnetic and non-magnetic sites in Eq. (1):
where \({c}_{i\sigma }^{(\dagger )}\) is the annihilation (creation) operator of an electron at site i with spin σ = ↑, ↓, and the hopping strength between electrons at sites i and j is described by tij. In our model, we take into account the hopping t between the nearest-neighbor magnetic site and non-magnetic site, the hopping tm between the nearest-neighbor magnetic sites, and the hopping tnm,1 (tnm,2) between the first (second) nearest-neighbor non-magnetic sites. J is the coupling between the electronic spins and localized magnetic moments Si, and σ is the Pauli matrix. In this work, we set \({{\bf{S}}}_{i}=(0,0,S{e}^{i{\bf{q}}{{\bf{r}}}_{i}})\) with magnetic wave vector q = (2π/a, 0) and S = 1. The on-site energy of magnetic and non-magnetic sites and their chemical potential are denoted by ϵm, ϵnm and μ.
We obtain the electronic band structure by diagonalizing the Hamiltonian matrix at each momentum point (see details in Supplementary Note 1). Without spin-orbit coupling, there are no interactions between electrons with opposite spins, and thus spin σ is a good quantum number. Finally, the Hamiltonian in Eq. (1) becomes
where \({f}_{n,{\bf{k}},\sigma }\,(\,{f}_{n,{\bf{k}},\sigma }^{\dagger })\) is a fermionic annihilation (creation) operator and En,σ(k) represents the corresponding energy dispersion of the nth band with spin σ (n = 1, …, 6, with energy increasing from low to high). For simplicity, we set t as the energy unit in the following discussion.
Altermagnetic band splittings of lattice with C 4z symmetry
The band dispersion of the above pentagonal lattice Hamiltonian exhibits g-wave altermagnetism, as shown in Fig. 2. Figure 2a shows six pairs of bands with inverse spin-splitting along Γ-C and Γ-D. Figure 2b, c plot the energy spin-splitting ΔEn(k) = En,↑(k) − En,↓(k) for the 4th and 5th pair of bands. In the first Brillouin zone, there are four spin-degenerate nodal lines, ky; = 0, kx = 0, kx = ky and kx = − ky, crossing the Γ point, as expected from the symmetry. Additionally, the spin-splitting ΔEn(k) for isolated pairs of bands near the Γ point is directly proportional to \({k}_{x}{k}_{y}({k}_{x}^{2}-{k}_{y}^{2})\) (see details in Supplementary Note 2). All these results illustrate that the tight-binding model successfully realizes g-wave altermagnetism.
a The electronic band structure along the chosen k path in the first Brillouin zone. The parameters used are tm = 0.2, tnm,1 = 0.9, tnm,2 = 0.6, J = 1, ϵm = ϵnm = 0 and μ = 0. The spin up and spin down bands are shown in red and blue, respectively. b, c False color plots of the spin-splitting energy ΔEn(k) = En,↑(k) − En,↓(k) for the 4th and 5th pair of bands in the first Brillouin zone. Four dashed lines represent the spin-degenerate nodal lines that cross the Γ point.
To understand the origin of g-wave altermagnetism in this pentagonal lattice, we investigate the dependence of spin-splitting on parameters of the tight-binding Hamiltonian in Eq. (1). To quantify the spin-splitting, we introduce two measures: the maximal spin-splitting of the n-th pair of bands ∣ΔEn∣max, and the mean spin-splitting over the entire Brillouin zone ∣ΔE∣mean, which are expressed as
Figure 3 summarizes how these two quantities vary as functions of J, tm, tnm,1 and tnm,2. The spin-splitting exhibits a complex dependence on the parameters of the tight-binding model, and the different pairs of electronic bands show significant distinctions. It is evident that the hopping t between magnetic and non-magnetic sites, and the coupling J between the electronic spins and localized magnetic moments are necessary for the system to exhibit altermagnetism. If t = 0, the magnetic sites form a normal Néel antiferromagnet; while if J = 0, there is no distinction between the sub-Hamiltonians for spin up and down.
Plots of the maximal spin-splitting for the n-th pair of bands, ∣ΔEn∣max, and the mean spin-splitting over the entire Brillouin zone, ∣ΔE∣mean, as functions of a J, b tm, c tnm,1, and d tnm,2, while other parameters are fixed as tm = 0.2, tnm,1 = 0.9, tnm,2 = 0.6, J = 1, ϵm = ϵnm = 0 and μ = 0. Different colors represent the band pairs, ordered from lower to higher energy, while gray represents the mean spin-splitting ∣ΔE∣mean.
In addition to these factors, all spin-splittings ∣ΔEn∣max and ∣ΔE∣mean vanish when either tnm,1 or tnm,2 becomes zero. In these cases, the Hamiltonian in Eq. (1) always exhibits band degeneracy En,σ(kx, ky) = En,σ(kx, − ky) = En,σ( − kx, ky), as well as symmetries \(\{{C}_{2\perp }| | {{\mathcal{M}}}_{x}(\frac{1}{2},\frac{1}{2})\}\) and \(\{{C}_{2\perp }| | {{\mathcal{M}}}_{y}(\frac{1}{2},\frac{1}{2})\}\) (see details in Supplementary Note 3), which lead to band degeneracy En,↑(k) = En,↓(k) at all k points. The above reasoning indicates that the hopping terms between non-magnetic sites are also crucial for the emergence of altermagnetic spin-splitting in the pentagonal lattice.
Altermagnetic band splittings of lattice under x y and \({\bar{x}}y\) strains
Next, we study the impact of strain along the diagonal directions xy and \(\bar{x}y\) on the altermagnetism of the Cairo pentagonal lattice while keeping the Néel antiferromagnetic spin arrangement, as shown in Fig. 1b. After applying the strain, a significant in-plane anisotropy emerges due to the lattice’s sensitivity to it. This anisotropy breaks the lattice symmetry C4z, and the symmetries \(\{{C}_{2\perp }| | {{\mathcal{M}}}_{x}(\frac{1}{2},\frac{1}{2})\}\) and \(\{{C}_{2\perp }| | {{\mathcal{M}}}_{y}(\frac{1}{2},\frac{1}{2})\}\) relating two magnetic sublattices, while preserving two other symmetries \(\{{C}_{2\perp }| | {{\mathcal{M}}}_{xy}(\frac{1}{2},\frac{1}{2})\}\) and \(\{{C}_{2\perp }| | {{\mathcal{M}}}_{\bar{x}y}(\frac{1}{2},\frac{1}{2})\}\). The change in symmetry results in a transformation in the structure of unconventional spin-splitting, as shown in Fig. 1d.
To incorporate the effect of strain into the tight-binding Hamiltonian, we introduce anisotropic hopping between sites along the xy and \(\bar{x}y\) directions. For simplicity, we keep the positions of atoms unchanged. Consequently, the hopping term in Eq. (1) is modified as
where the anisotropy δ reflects the strength of strain.
The band dispersion relations are shown in Fig. 4a. Unlike the previous case, six pairs of bands now exhibit inverse spin-splitting along X-Γ-C and Y-Γ-D. The energy spin-splittings ΔEn(k) shown in Fig. 4b, c demonstrate that there are only two spin-degenerate nodal lines, kx = ± ky, crossing the Γ point, while the two former nodal lines are shifted away. The change in nodal lines is consistent with the symmetry breaking induced by diagonal strain, while the two preserved symmetries \(\{{C}_{2\perp }| | {{\mathcal{M}}}_{xy}(\frac{1}{2},\frac{1}{2})\}\) and \(\{{C}_{2\perp }| | {{\mathcal{M}}}_{\bar{x}y}(\frac{1}{2},\frac{1}{2})\}\) enforce the existence of the remaining two nodal lines. In Fig. 4d and e, we zoom in on the above plots near the Γ point, where the spin-splitting ΔEn(k) is directly proportional to \({k}_{x}^{2}-{k}_{y}^{2}\) (see details in Supplementary Note 2) and thus exhibits d-wave symmetry. Consequently, as a result of the applied strain, the system transitions from a g-wave to a d-wave altermagnetism in the neighborhood of the Γ point.
a The electronic band structure along the chosen k path in the first Brillouin zone. The parameters used are δ = 0.1, tm = 0.2, tnm,1 = 0.9, tnm,2 = 0.6, J = 1, ϵm = ϵnm = 0 and μ = 0. The bands of spin up and down are labeled by red and blue. b, c False color plots of the spin-splitting energy ΔEn(k) = En,↑(k) − En,↓(k) for the 4th and 5th pair of bands in the first Brillouin zone. d, e Plots of ΔEn(k) near the Γ points, within ∣kx∣ ≤ 10−2π/a and ∣ky∣ ≤ 10−2π/a.
Additionally, we examine the dependence of spin-splitting on the strain strength. Figure 5a illustrates the maximal spin-splitting ∣ΔEn∣max and the mean spin-splitting ∣ΔE∣mean as functions of anisotropy δ. The sensitivity of the band splitting to strain varies among different pairs of bands, while the variation in the mean spin-splitting ∣ΔE∣mean is relatively small. Within the range of ∣δ∣≤0.5, the maximum splitting can reach slightly higher than the energy level of t. For comparison, Fig. 5b depicts the case where the next-nearest-neighbor hopping between the magnetic and nonmagnetic sites tnm,2 is zero. As discussed earlier, the unstrained system exhibits normal spin-degeneracy in this case. Upon applying diagonal strain, the system transitions from a normal antiferromagnetic phase to a d-wave altermagnetic phase, with the maximal spin-splitting showing an almost linear dependence on δ for small values of δ.
Plots of the maximal spin-splitting for the n-th pair of bands, ∣ΔEn∣max, and the mean spin-splitting over the entire Brillouin zone, ∣ΔE∣mean, as functions of anisotropy δ under a tnm,2 = 0.6 and b tnm,2 = 0, with other parameters being tm = 0.2, tnm,1 = 0.9, nm,2 J = 1, ϵm = ϵnm = 0 and μ = 0. Different colors represent the band pairs, ordered from lower to higher energy, while gray represents the mean spin-splitting ∣ΔE∣mean.
Polarized nodal points and band topology
In addition to the nodal lines formed by spin-degeneracy, several spin-polarized nodal points are present in this altermagnetic pentagonal lattice. As shown in Fig. 6a, eight spin-polarized nodal points emerge at the band crossings of the 4th and 5th bands with the same spin. These nodal points are protected by the symmetry \(\{{\mathcal{T}}{C}_{2\perp }| | {\mathcal{T}}{\mathcal{P}}\}\), with each nodal point carrying a π Berry phase, where \({\mathcal{T}}\) represents the time-reversal operator. Thus, nontrivial band topology can be achieved by breaking this symmetry. Here, we explore two mechanisms that gap out the spin-polarized nodal points using toy models: breaking the lattice inversion symmetry \(\{E| | {\mathcal{P}}\}\) through the anisotropic hopping term between magnetic atoms, represented by
and breaking the lattice time-reversal symmetry \(\{{\mathcal{T}}{C}_{2\perp }| | {\mathcal{T}}\}\) through the complex hopping term between magnetic atoms, represented by
where \({\boldsymbol{\xi }}=\frac{1}{2}a\hat{x}\pm \frac{1}{2}a\hat{y}\).
a Eight spin-polarized nodal points formed by band crossing between the 4th and 5th pairs of bands. b The electronic band structure for the 4th and 5th pairs of bands under inversion symmetry \(\{E| | {\mathcal{P}}\}\) breaking with t1 = 0.2, and c, d the corresponding Berry curvature Ω(k) of the 5th pair of spin-up and spin-down bands. e The electronic band structure for the 4th and 5th pairs of bands under time-reversal symmetry \(\{{\mathcal{T}}{C}_{2\perp }| | {\mathcal{T}}\}\) breaking with t2 = 0.2, and f, g the corresponding Berry curvature Ω(k) of the 5th pair of spin-up and spin-down bands. The other parameters in the calculations are tm = 0.2, tnm,1 = 0.9, tnm,2 = 0.6, J = 1, ϵm = ϵnm = 0 and μ = 0.
The impact of breaking the symmetry \(\{E| | {\mathcal{P}}\}\) using Eq. (5) is illustrated in Fig. 6b. All spin-polarized nodal points are gapped out, while the spin-degeneracy nodal lines are preserved. We plot the Berry curvature Ω(k) of the 5th pairs of bands with opposite spins after inversion symmetry breaking in Fig. 6c, d, where the peaks of Ω(k) appear at the former nodal points (see details in Supplementary Note 4). In this case, the Berry curvature satisfies Ω(k) = − Ω( − k) for all k points, resulting in Chern number \({{\mathcal{C}}}_{\sigma }=0\). In contrast, the band structure after \(\{{\mathcal{T}}{C}_{2\perp }| | {\mathcal{T}}\}\) symmetry breaking using Eq. (6) and the Berry curvature of the 5th pairs of bands are shown in Fig. 6e, f and g. In each band, this symmetry-breaking term in Eq. (6) leads to an equivalent contribution from four nodal points. This mechanism mirrors the behavior seen in Haldane’s model, making the 5th pair of bands topologically non-trivial with a Chern number \({{\mathcal{C}}}_{\sigma }=\mp 2\).
Candidate materials
A two-dimensional form of pyrite, FeS2, has been predicted to exhibit a Cairo pentagonal structure with antiferromagnetic order50. It belongs to space group P4/mbm (No.127) and point group D4h. Fig. 7a shows the crystal structure, where each Fe atom is surrounded by four S atoms. To examine the magnetic ground state, we perform ab initio calculations to calculate the energy of different collinear magnetic states on this planar structure under geometry optimizations (see details in Supplementary Note 5). The result indicates that the most stable configuration exhibits the Néel antiferromagnetic spin arrangement on Fe atoms while S atoms are non-magnetic, as shown in Fig. 7a.
a Crystal structure of FeS2. The red and blue colors of Fe atoms indicate the directions of their localized magnetic moments, while the orange arrows represent the direction of diagonal strain. b The electronic band structure and the spin-splitting of the 30th pair of bands of FeS2 without strain. c, d Spin density isosurfaces for two pairs of bands without strain, illustrating narrow and large spin-splitting, respectively (the inset zooms in on energy by a factor of 5). e The electronic band structure of FeS2 with diagonal strain, where the angle γ between lattice vectors a and b is tuned to γ = 88°. f, g False color plot of the spin-splitting of the 30th pair of bands of FeS2 without and with diagonal strain (indicated by the blue arrow in b and e, see details in Supplementary Note 6). The coordinate (k1, k2) corresponds to reciprocal lattice vector br = k1b1 + k2b2.
Thus, this planar FeS2 structure corresponds to the pentagonal lattice system discussed above. In particular, it exhibits the symmetries \(\{{C}_{2\perp }| | {{\mathcal{M}}}_{x}(\frac{1}{2},\frac{1}{2})\}\), \(\{{C}_{2\perp }| | {{\mathcal{M}}}_{y}(\frac{1}{2},\frac{1}{2})\}\), \(\{{C}_{2\perp }| | {{\mathcal{M}}}_{xy}(\frac{1}{2},\frac{1}{2})\}\), and \(\{{C}_{2\perp }| | {{\mathcal{M}}}_{\bar{x}y}(\frac{1}{2},\frac{1}{2})\}\), which relate two magnetic sublattices. The electronic band structure and the spin-splitting of one pair of the bands obtained from ab initio calculations are shown in Fig. 7b, f, displaying features consistent with the tight-binding model result in Fig. 2, that inverse spin-splitting occurs along the Γ-C and Γ-D paths, while spin-degeneracy is observed along the Γ-X, Γ-Y, and Γ-M paths. These behaviors characterize planar pentagonal FeS2 as a g-wave altermagnet as expected. Additionally, we present the spin density isosurfaces for bands with different levels of spin-splitting in Fig. 7c, d, where a larger spin-splitting coincides with a more anisotropic spin density.
Next, we study the effect of strains along the diagonal directions on planar FeS2, as indicated by the arrows in Fig. 7a. We introduce the strain in ab initio calculations by slightly reducing the angle γ between lattice vectors a and b to less than 90°. For a small change Δγ = 2°, our ab initio calculation shows that the Néel antiferromagnetic spin arrangement is still the most stable configuration (see details in Supplementary Note 5). Fig. 7e and g present the electronic band structure and the spin-splitting for γ = 88. 0°, where inverse spin-splitting appears along the Γ-X-C-Γ and Γ-Y-D-Γ paths, while the bands remain degenerate along the Γ-M path, displaying the same features as in the model result shown in Fig. 4. This behavior indicates a transition from g-wave to d-wave altermagnetism, as predicted by the tight-binding model analysis.
In addition to the two-dimensional cases, we highlight Nb2FeB2, a three-dimensional material predicted to exhibit g-wave altermagnetism51,52. Figure 8a shows the crystal structure of Nb2FeB2, which consists of alternate layers of Fe-B and Nb atoms along c axis. Each Fe-B layer exhibits the same Cairo pentagonal structure discussed earlier, and the magnetic Fe atoms carry an in-plane Néel antiferromagnetic spin arrangement, which is also examined through ab initio calculation (see details in Supplementary Note 5). Thus, the previous analysis of symmetry-induced altermagnetism also applies to Nb2FeB2. Its electronic band structures and spin-splitting without and with the diagonal strain are shown in Fig. 8c–f, where the angle γ between lattice vectors a and b is 90° and 88° respectively. These results demonstrate that Nb2FeB2 is a g-wave altermagnet and it undergoes a transition from g-wave to d-wave altermagnetism under diagonal strain, as expected. Furthermore, in both altermagnetic materials, the spin-splitting increases significantly after applying strain, consistent with the model analysis.
a Crystal structure of Nb2FeB2 shown from two perspectives. The purple arrows indicate the directions of localized magnetic moments on Fe atoms. b The first Brillouin zone of Nb2FeB2 and high symmetry moment points of our interest. c, d The electronic band structures of Nb2FeB2 under two conditions: without diagonal strain (where the angle γ between lattice vectors a and b is 90°) and with diagonal strain (where γ = 88°). e, f False color plot of the spin-splitting of the 45th pair of bands of Nb2FeB2 without and with diagonal strain (indicated by the blue arrow in (c, d), see details in Supplementary Note 6). The coordinate (k1, k2) corresponds to reciprocal lattice vector br = k1b1 + k2b2.
Discussion
Our study uncovers the first microscopic mechanism underlying g-wave altermagnetism in two dimension, realized in Cairo pentagonal lattice. Based on the symmetry analysis, we propose a simple but realistic tight-binding model containing both magnetic and non-magnetic sites that successfully realizes the altermagnetic band spin-splitting with g-wave symmetry. The lattice-spin symmetry ensured by the presence of non-magnetic atoms guarantees the existence of altermagnetism in this antiferromagnetically ordered pentagonal system. Our analysis of the dependence of spin-splitting on parameters in the tight-binding model provides one of the mechanisms, that both the hopping between magnetic site and non-magnetic site, and the hopping between non-magnetic sites are indispensable for the altermagnetic behavior. We note that there are other possible origins of altermagnetism, such as the anisotropic ordering of local orbitals, which are beyond our discussion in this work.
Another important finding from our model is that applying strain to the crystal can induce a transformation in the type of altermagnetism. In general, strain reduces the original lattice-spin symmetry and introduces anisotropic interactions, which makes the spin-splitting behavior more complex, even changes magnetism of the ground state. In this work, we focus on the cases where the strain is relatively small so that it does not affect the Néel antiferromagnetic spin arrangement in the ground state, which is confirmed by our ab initio calculations. Under this premise, near the Γ point we observe that the strain applied along the diagonal direction of this pentagonal structure causes a transition from g-wave to d-wave altermagnetism. Additionally, we find that this pentagonal altermagnet contains spin-polarized nodal points protected by \(\{{\mathcal{T}}{C}_{2\perp }| | {\mathcal{T}}{\mathcal{P}}\}\) symmetry. Breaking this symmetry could lead to non-trivial topological bands and the quantum Hall effect for electrons.
Finally, we identify several candidate materials through ab initio calculations, including monolayer FeS2 and bulk Nb2FeB2, which show promise for exploring altermagnetism in Cairo pentagonal crystal structures. In particular, we examine the effect of diagonal strain on these crystals, observing that altermagnetism transforms from g-wave to d-wave as expected. These findings suggest new paradigm of tuning the type of altermagnetism by strain, opening new possibilities for enabling strain-tuned spintronic devices and the interplay with other degrees of freedom.
Methods
We use Vienna Ab initio Simulation Package (VASP)53 for all density functional theory calculations. Projector-augmented-wave (PAW) potential was utilized for ion-electron interaction, and the generalized gradient approximation (GGA) with the Perdew-Burke-Ernzerhof (PBE) functional54 was applied to describe electron exchange-correlation interaction. The energy cutoff was set to be 600 eV. More details are in Supplementary Note 5.
Data availability
The data analyzed in the present study is available from the corresponding author upon reasonable request.
Code availability
The codes and scripts used to generate the data used in this study are available from the corresponding author upon reasonable request.
References
Šmejkal, L., Sinova, J. & Jungwirth, T. Beyond conventional ferromagnetism and antiferromagnetism: A phase with nonrelativistic spin and crystal rotation symmetry. Phys. Rev. X. 12, 031042 (2022).
Šmejkal, L., Sinova, J. & Jungwirth, T. Emerging research landscape of altermagnetism. Phys. Rev. X. 12, 040501 (2022).
Šmejkal, L., MacDonald, A. H., Sinova, J., Nakatsuji, S. & Jungwirth, T. Anomalous hall antiferromagnets. Nat. Rev. Mater. 7, 482–496 (2022).
Ma, H.-Y. et al. Multifunctional antiferromagnetic materials with giant piezomagnetism and noncollinear spin current. Nat. Commun. 12, 2846 (2021).
Hayami, S., Yanagi, Y. & Kusunose, H. Bottom-up design of spin-split and reshaped electronic band structures in antiferromagnets without spin-orbit coupling: Procedure on the basis of augmented multipoles. Phys. Rev. B. 102, 144441 (2020).
Hayami, S., Yanagi, Y. & Kusunose, H. Momentum-dependent spin splitting by collinear antiferromagnetic ordering. J. Phys. Soc. Jpn. 88, 123702 (2019).
González-Hernández, R. et al. Efficient electrical spin splitter based on nonrelativistic collinear antiferromagnetism. Phys. Rev. Lett. 126, 127701 (2021).
Yuan, L.-D., Wang, Z., Luo, J.-W., Rashba, E. I. & Zunger, A. Giant momentum-dependent spin splitting in centrosymmetric low-z antiferromagnets. Phys. Rev. B. 102, 014422 (2020).
Yuan, L.-D., Wang, Z., Luo, J.-W. & Zunger, A. Prediction of low-z collinear and noncollinear antiferromagnetic compounds having momentum-dependent spin splitting even without spin-orbit coupling. Phys. Rev. Mater. 5, 014409 (2021).
Feng, Z. et al. An anomalous hall effect in altermagnetic ruthenium dioxide. Nat. Electron. 5, 735–743 (2022).
Fedchenko, O. et al. Observation of time-reversal symmetry breaking in the band structure of altermagnetic ruo2. Sci. Adv. 10, eadj4883 (2024).
Zhou, X. et al. Crystal thermal transport in altermagnetic ruo2 (2023). Phys. Rev. Lett. 132, 056701 (2024).
Lee, S. et al. Broken kramers degeneracy in altermagnetic mnte. Phys. Rev. Lett. 132, 036702 (2024).
Osumi, T. et al. Observation of a giant band splitting in altermagnetic mnte. Phys. Rev. B. 109, 115102 (2024).
Krempasky`, J. et al. Altermagnetic lifting of kramers spin degeneracy. Nature. 626, 517–522 (2024).
Liu, Z., Ozeki, M., Asai, S., Itoh, S. & Masuda, T. Chiral split magnon in altermagnetic mnte. Phys. Rev. Lett. 133, 156702 (2024).
Mazin, I. I., Koepernik, K., Johannes, M. D., González-Hernández, R. & Šmejkal, L. Prediction of unconventional magnetism in doped fesb2. Proc. Natl Acad. Sci. 118, e2108924118 (2021).
Reimers, S. et al. Direct observation of altermagnetic band splitting in crsb thin films. Nat. Commun. 15, 2116 (2024).
Yang, G. et al. Three-dimensional mapping of the altermagnetic spin splitting in crsb. Nat. Commun. 16, 1442 (2025).
Bai, H. et al. Observation of spin splitting torque in a collinear antiferromagnet ruo 2. Phys. Rev. Lett. 128, 197202 (2022).
Karube, S. et al. Observation of spin-splitter torque in collinear antiferromagnetic ruo 2. Phys. Rev. Lett. 129, 137201 (2022).
Sumita, S., Naka, M. & Seo, H. Fulde-ferrell-larkin-ovchinnikov state induced by antiferromagnetic order in κ-type organic conductors. Phys. Rev. Res. 5, 043171 (2023).
Zhu, D., Zhuang, Z.-Y., Wu, Z. & Yan, Z. Topological superconductivity in two-dimensional altermagnetic metals. Phys. Rev. B. 108, 184505 (2023).
Brekke, B., Brataas, A. & Sudbø, A. Two-dimensional altermagnets: Superconductivity in a minimal microscopic model. Phys. Rev. B. 108, 224421 (2023).
Zhang, S.-B., Hu, L.-H. & Neupert, T. Finite-momentum cooper pairing in proximitized altermagnets. Nat. Commun. 15, 1801 (2024).
Chakraborty, D. & Black-Schaffer, A. M. Zero-field finite-momentum and field-induced superconductivity in altermagnets. Phys. Rev. B. 110, L060508 (2024).
Wei, M. et al. Gapless superconducting state and mirage gap in altermagnets. Phys. Rev. B. 109, L201404 (2024).
Mæland, K., Brekke, B. & Sudbø, A. Many-body effects on superconductivity mediated by double-magnon processes in altermagnets. Phys. Rev. B. 109, 134515 (2024).
Chourasia, S., Svetogorov, A., Kamra, A. & Belzig, W. Thermodynamic properties of a superconductor interfaced with an altermagnet. Phys. Rev. B. 111, 224503 (2025).
Leiviskä, M. et al. Anisotropy of the anomalous hall effect in thin films of the altermagnet candidate mn 5 si 3. Phys. Rev. B. 109, 224430 (2024).
Sato, T., Haddad, S., Fulga, I. C., Assaad, F. F. & van den Brink, J. Altermagnetic anomalous hall effect emerging from electronic correlations. Phys. Rev. Lett. 133, 086503 (2024).
Attias, L., Levchenko, A. & Khodas, M. Intrinsic anomalous hall effect in altermagnets. Phys. Rev. B. 110, 094425 (2024).
Liao, C.-T., Wang, Y.-C., Tien, Y.-C., Huang, S.-Y. & Qu, D. Separation of inverse altermagnetic spin-splitting effect from inverse spin hall effect in ruo 2. Phys. Rev. Lett. 133, 056701 (2024).
Jin, H., Tan, Z., Gong, Z. & Wang, J. Anomalous hall effect in two-dimensional vanadium tetrahalogen with altermagnetic phase. Phys. Rev. B. 110, 155125 (2024).
Fang, Y., Cano, J. & Ghorashi, S. A. A. Quantum geometry induced nonlinear transport in altermagnets. Phys. Rev. Lett. 133, 106701 (2024).
Rao, P., Mook, A. & Knolle, J. Tunable band topology and optical conductivity in altermagnets. Phys. Rev. B. 110, 024425 (2024).
Parshukov, K., Wiedmann, R. & Schnyder, A. P. Topological crossings in two-dimensional altermagnets: Symmetry classification and topological responses. Phys. Rev. B. 111, 224406 (2025).
Tan, C.-Y. et al. Bipolarized weyl semimetals and quantum crystal valley hall effect in two-dimensional altermagnetic materials. arXiv preprint arXiv:2406.16603 (2024).
Bai, L. et al. Altermagnetism: Exploring new frontiers in magnetism and spintronics. Adv. Funct. Mater. 34, 2409327 (2024).
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. X. 12, 011028 (2022).
Leeb, V., Mook, A., Šmejkal, L. & Knolle, J. Spontaneous formation of altermagnetism from orbital ordering. Phys. Rev. Lett. 132, 236701 (2024).
Roig, M., Kreisel, A., Yu, Y., Andersen, B. M. & Agterberg, D. F. Minimal models for altermagnetism. Phys. Rev. B. 110, 144412 (2024).
Das, P., Leeb, V., Knolle, J. & Knap, M. Realizing altermagnetism in fermi-hubbard models with ultracold atoms. Phys. Rev. Lett. 132, 263402 (2024).
Shen, Y. & Wang, Q. Pentagon-based 2d materials: Classification, properties and applications. Phys. Rep. 964, 1–42 (2022).
Wang, Y., Li, Y. & Chen, Z. Not your familiar two dimensional transition metal disulfide: structural and electronic properties of the pds 2 monolayer. J. Mater. Chem. C. 3, 9603–9608 (2015).
Shao, X. et al. Electronic properties of a π-conjugated cairo pentagonal lattice: Direct band gap, ultrahigh carrier mobility, and slanted dirac cones. Phys. Rev. B. 98, 085437 (2018).
Li, P. et al. Penta-pdpse: a new 2d pentagonal material with highly in-plane optical, electronic, and optoelectronic anisotropy. Adv. Mater. 33, 2102541 (2021).
Antonenko, D. S., Fernandes, R. M. & Venderbos, J. W. Mirror chern bands and weyl nodal loops in altermagnets. Phys. Rev. Lett. 134, 096703 (2025).
Del Re, L. Dirac points and topological phases in correlated altermagnets. arXiv preprint arXiv:2408.14288 (2024).
Wei, S., Zheng, S., Wen, X., Xie, C. & Liang, J. A novel antiferromagnetic semiconductor hidden in pyrite. Comput. Mater. Sci. 183, 109852 (2020).
Hou, X.-Y., Yang, H.-C., Liu, Z.-X., Guo, P.-J. & Lu, Z.-Y. Large intrinsic anomalous hall effect in both nb2feb2 and ta2feb2 with collinear antiferromagnetism. Phys. Rev. B. 107, L161109 (2023).
Gao, Z. F. et al. Ai-accelerated discovery of altermagnetic materials. Natl. Sci. Rev. 12, nwaf066 (2025).
Kresse, G. & Hafner, J. Ab initio molecular-dynamics simulation of the liquid-metal–amorphous-semiconductor transition in germanium. Phys. Rev. B. 49, 14251–14269 (1994).
Perdew, J. P., Burke, K. & Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 77, 3865–3868 (1996).
Acknowledgements
This work is supported by the Center for Molecular Magnetic Quantum Materials, an Energy Frontier Research Center funded by the U.S. Department of Energy, Office of Science, Basic Energy Sciences under Award no. DE-SC0019330. Computations were done using the utilities of the University of Florida Research Computing.
Author information
Authors and Affiliations
Contributions
C.J. and S.L. conceived, and C.J. supervised the work. S.L. performed calculations based on the tight-binding model. Y.Z. performed DFT calculations. S.L. analyzed the data. S.L., Y.Z., X.Z., A.B. and C.J. wrote the manuscript, with input from all authors. All authors reviewed the manuscript.
Corresponding author
Ethics declarations
Competing interests
The authors declare no competing interests.
Additional information
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Supplementary information
Rights and permissions
Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.
About this article
Cite this article
Li, S., Zhang, Y., Bahri, A. et al. Altermagnetism and strain induced altermagnetic transition in Cairo pentagonal monolayer. npj Quantum Mater. 10, 83 (2025). https://doi.org/10.1038/s41535-025-00793-0
Received:
Accepted:
Published:
DOI: https://doi.org/10.1038/s41535-025-00793-0