Abstract
We report on a class of gapped projected entangled pair states (PEPS) with non-trivial Euler topology motivated by recent progress in band geometry. In the non-interacting limit, these systems have optimal conditions relating to saturation of quantum geometrical bounds, allowing for parent Hamiltonians whose lowest bands are completely flat and which have the PEPS as unique ground states. Protected by crystalline symmetries, these states evade restrictions on capturing tenfold-way topological features with gapped PEPS. These PEPS thus form the first tensor network representative of a non-interacting, gapped two-dimensional topological phase, similar to the Kitaev chain in one dimension. Using unitary circuits, we then formulate interacting variants of these PEPS and corresponding gapped parent Hamiltonians. We reveal characteristic entanglement features shared between the free-fermionic and interacting states with Euler topology. Our results hence provide a rich platform of PEPS models that have, unexpectedly, a finite topological invariant, forming the basis for new spin liquids, quantum Hall physics, and quantum information pursuits.
Similar content being viewed by others
Introduction
Tensor network states (TNS) form a generally applicable tool for the description of quantum matter. A numerically efficient representation of the ground states of local gapped Hamiltonians1,2,3,4,5, TNS play a pivotal role both in the simulation of correlated systems6,7,8,9,10,11 and the analytical classification of topological phases12,13,14,15,16,17. Yet, due to the increased complexity in higher dimensions, TNS have not yet matched the success of non-interacting band theory18,19, both in simulations and the classification of topological phases20,21,22,23,24,25,26. Of particular interest are therefore systems which can be well captured with band theory but are marked by difficulties when it comes to TNS approaches. The most well-known such example is chiral topological systems: In topological band theory, they are characterized by occupied bands whose overall Chern number is non-vanishing, separated by a gap from the conduction bands. While TNS approaches are able to capture the chiral topological features of such systems, this comes at the cost of producing algebraically decaying correlations characteristic of critical systems27,28,29. Generally, it has been shown that TNS with exponentially decaying correlations cannot capture any higher-dimensional topological invariant30 of the ten Altland-Zirnbauer (AZ) classes31.
The severe restrictions of TNS to represent gapped non-interacting topological phases suggest that TNS might equally struggle to capture topological phases protected by crystalline symmetries. However, reinvigorated interests32,33 in relation to quantum geometry34,35 could provide a useful tool in that they outline flatband conditions. Under such conditions, it is possible to define topological flatband Hamiltonians that can be formulated as exemplary parent Hamiltonians. These are sums of projectors with local support that each annihilate the ground state(s). From these local projectors, a TNS ground state can in principle be constructed. However, whether the resulting state is non-vanishing and can be made the unique ground state of such a crystalline symmetry-protected topological Hamiltonian might be hampered by the previously mentioned hurdles. Intuitively, TNS with exponentially decaying correlations are incompatible with topological invariants, as they come with delocalized edge modes around a physical boundary (in more than one dimension); the local structure of tensor networks is incapable of separating such edge modes from the bulk modes, delocalizing them as well. Because of that, in the following, we consider crystalline symmetry-protected topological phases which do not have helical or chiral edge modes.
We show that a family of topological projected entangled pair states (PEPS)36 and generating Hamiltonians can be formulated in the context of the Euler class25,37,38,39. The Euler class is a multi-gap invariant25, pertaining to topological structures that emerge when groups of partitioned bands (band subspaces) carry non-trivial topological indices25. These topological charges of groups of bands can be altered by braiding nodes in momentum space, as band nodes residing between neighboring bands can carry non-Abelian charges37,38,40. The braiding of non-Abelian frame charges and multi-gap topologies have been increasingly related, both theoretically and experimentally, to physical systems that range from out-of-equilibrium settings41,42,43 and phonon spectra44,45, to electronic systems (twisted, magnetic and conventional)38,46,47 as well as metamaterials48,49,50.
The here introduced family of PEPS has non-trivial Euler class, evading no-go conditions as for tenfold-way topologies31, constituting a 2D analog of the Kitaev chain51. These PEPS differ from the higher-order topological insulators represented by PEPS in Ref. 52, which have a bond dimension in one spatial direction that grows with the system size. Within the non-interacting limit, from a band theory perspective, our PEPS enjoy ideal quantum geometrical properties, which we elucidate in the Supplementary Information (SI) in detail. More importantly, by applying shallow quantum circuits of diagonal unitaries, we transform these PEPS and their gapped parent Hamiltonians to interacting variants. We can signify the Euler phase both in the non-interacting and the interacting limit upon appealing to the entanglement spectrum. As such, our results set a benchmark for an exact class of PEPS parent Hamiltonians with finite topological invariant.
Results
Euler class
A pair of isolated (gapped from the rest of the spectrum) bands \(\left\vert {u}_{n}({{\boldsymbol{k}}})\right\rangle,\left\vert {u}_{n+1}({{\boldsymbol{k}}})\right\rangle\) can acquire a non-trivial Euler class χ when it is part of at least a three-band system that enjoys a reality condition assured by the presence of \({{{\mathcal{C}}}}_{2}{{\mathcal{T}}}\) [twofold rotations combined with time-reversal symmetry (TRS)], or \({{\mathcal{P}}}{{\mathcal{T}}}\) symmetry, involving parity and TRS. The Euler class is then concretely obtained as37,38
where one integrates the Euler curvature \({{\rm{Eu}}}=\langle {\partial }_{{k}_{1}}{u}_{n}({{\boldsymbol{k}}})| {\partial }_{{k}_{2}}{u}_{n+1}({{\boldsymbol{k}}})\rangle -\langle {\partial }_{{k}_{2}}{u}_{n}({{\boldsymbol{k}}})| {\partial }_{{k}_{1}}{u}_{n+1}({{\boldsymbol{k}}})\rangle\) over the Brillouin zone (BZ). The pair of bands can either be degenerate (and flat) or feature a number of 2χ nodal points that cannot be annihilated due to the topological nature25,37,38. Eq. (1) shows that the Euler class is the real analog of the Chern number. Similarly, the isolated two-band subspace does not admit exponentially-localized Wannier functions in a \({{{\mathcal{C}}}}_{2}{{\mathcal{T}}}\)-symmetric gauge, but unlike the Chern case, the system does not feature protected chiral or helical edge states, allowing for a PEPS representation. Our PEPS construction will break the Wannier restriction to \({{{\mathcal{C}}}}_{2}{{\mathcal{T}}}\) symmetric gauge and three bands; more precisely, the latter, as it starts with virtual particles occupying six bands. The projection from virtual to physical fermions reduces the number of bands back to three, constituting a (non-invertible) map between the gauge-symmetry breaking input state and the gapped Euler state. Hence, a representation by a PEPS with a finite bond dimension is made possible by the PEPS construction breaking the gauge symmetry.
The model
To concretize the discussion, we consider spinless fermions hopping on the kagome lattice with nearest-neighbor hopping t = − 1, next-nearest-neighbor hopping \({t}^{{\prime} }=-1\) and third-nearest-neighbor hopping t″ = − 1 inside the hexagons (see Fig. 1). For chemical potential μ, the Hamiltonian thus reads
where 〈i, j〉, 〈〈i, j〉〉 correspond to nearest- and next-nearest neighbor pairs of sites i, j and 〈〈〈i, j〉〉〉⬡ to third-nearest neighbor pairs of the same hexagons. \({a}_{i}^{{\dagger} }\) (ai) are the fermionic creation (annihilation) operators and N the number of sites. The Hamiltonian has two degenerate flat bands at E = − 2 − μ and a dispersive band on top, separated by an energy gap Δ = 3; for more details on the model, see Methods. The flat bands have Euler number χ = 1, protected by \({{{\mathcal{C}}}}_{2}{{\mathcal{T}}}\) symmetry. At μ = − 2, both flat bands are at E = 0, and the ground states \(\left\vert {\psi }_{n}\right\rangle,\,n=0,1,\ldots,2N/3\) are macroscopically degenerate, characterized by fillings [0, 1, 2, …, 2N/3]. We now construct the ground state with the highest filling, which will become the unique ground state for − 2 < μ < 1. To that end, we note that the Hamiltonian for μ = − 2 can be rewritten as
where ⬡ denotes the hexagons of the kagome lattice, and we defined \({a}_{\hexagon}=\frac{1}{\sqrt{6}}{\sum }_{i\in \hexagon}{a}_{i}\) and \({h}_{\hexagon}={a}_{\hexagon}^{{\dagger} }{a}_{\hexagon}\). Hence, the ground states fulfill \({a}_{\hexagon}\left\vert {\psi }_{n}\right\rangle=0\) for all hexagons ⬡. The ground state with the highest occupation number is
where \(\left\vert 1\ldots 1\right\rangle\) is the fully occupied state. Due to \(\{{a}_{\hexagon},{a}_{\hexagon^{{\prime} }}\}=0\), the ordering in Eq. (4) is irrelevant. However, notably, for the other commutation relations, we have \(\{{a}_{\hexagon},{a}_{\hexagon^{{\prime} }}^{{\dagger} }\}={\delta }_{\hexagon,\hexagon^{{\prime} }}+\frac{1}{6}{\delta }_{\langle \hexagon,\hexagon^{{\prime} }\rangle }\), where \(\langle \hexagon,\hexagon^{{\prime} }\rangle\) denotes corner-sharing, neighboring hexagonal plaquettes. This anticommutation algebra shows that, while the operators \({a}_{\hexagon}^{{\dagger} }\) effectively create fermions in a superposition of six atomic orbitals, the Hamiltonian Eq. (3) is not adiabatically connected to a Hamiltonian of an atomic insulator, despite the functional similarity to such Hamiltonians: In particular, even if one deforms our quasiparticle operators a⬡ (see following section), their non-trivial anticommutation properties are protected by \({{{\mathcal{C}}}}_{2}{{\mathcal{T}}}\) symmetry. That is, the a⬡ cannot be deformed into single-site operators (and similarly for the Hamiltonian) without breaking \({{{\mathcal{C}}}}_{2}{{\mathcal{T}}}\) symmetry.
\(\vert {\psi }_{2N/3}\rangle\) can be written as a projected entangled simplex state (PESS)53 as follows: We start out with a virtual state of 2N spinless fermions – two assigned to each physical fermion. The virtual fermions are in the state \(\vert {\omega }_{v}\rangle=\frac{1}{\sqrt{6}}{\prod }_{\hexagon}{\sum }_{i=1}^{6}{c}_{\hexagon,i}\vert {1}_{v}\rangle\), where \(\vert {1}_{v}\rangle\) denotes the fully occupied virtual state. c⬡,i corresponds to virtual fermion i = 1, …, 6 within hexagon ⬡ (as opposed to the physical particles aj, there are now unique assignments within hexagons), cf. Fig. 2. The next step is to map each pair of virtual fermions around a site to one physical fermion. To that end, we use the operator \({\hat{M}}_{j}={a}_{j}^{{\dagger} }{c}_{j}^{{\prime} }{c}_{j}+{c}_{j}^{{\prime} }-{c}_{j}\). Here, \({c}_{j}^{{\prime} }\) corresponds to the virtual fermion located on the left of the site j and cj to the virtual fermion located on its right. We finally project on the vacuum of virtual particles, obtaining the overall state
where \(\vert {1}_{v}{0}_{p}\rangle\) corresponds to the vacuum of physical fermions and fully occupied virtual fermionic state. We already labeled the overall state as a “PEPS”, since it can also be written as the more familiar projected entangled pair state, as we show further below. We note that it is a fermionic PEPS where tensors are replaced by fermionic operators54. The map \({\hat{M}}_{j}={a}_{j}^{{\dagger} }{c}_{j}^{{\prime} }{c}_{j}+{c}_{j}^{{\prime} }-{c}_{j}\) has the following effect on the superposition of basis states in \({\prod }_{\hexagon}\frac{1}{\sqrt{6}}{\sum }_{i=1}^{6}{c}_{\hexagon,i}\vert {1}_{v}{0}_{p}\rangle\): If site \(j=\hexagon\cap \hexagon^{{\prime} }\) is neither affected by ci with i ∈ ⬡ nor ck with \(k\in \hexagon^{{\prime} }\) (i.e., there are two virtual fermions at site j), then a physical fermion is created via \({a}_{j}^{{\dagger} }\). If site j is affected by either ci∈ ⬡ or \({c}_{k\in \hexagon^{{\prime} }}\) (i.e., there is one virtual fermion at site j), then no physical fermion is created at site j, leaving it in the physical vacuum state. If site j is affected by both ci∈⬡ and \({c}_{k\in \hexagon^{{\prime} }}\) (i.e., there is no virtual fermion at site j), the basis state is annihilated. The negative sign in \({\hat{M}}_{j}\) is necessitated by the fermionic anticommutation relations in the PEPS construction.
In order to demonstrate that \(\vert {\psi }_{2N/3}\rangle \propto \vert {\psi }_{{{\rm{PEPS}}}}\rangle\), we first verify that \({a}_{\hexagon}\vert {\psi }_{{{\rm{PEPS}}}}\rangle=0\) and later that the PEPS has filling 2N/3. For the first claim, we notice that
where […] denotes a sum of products of operators that do not act on the physical fermion at site j. We thus have
Second, we see that the initial state \(\vert {1}_{v}{0}_{p}\rangle\) contains 2N virtual and no physical fermions. The operator \({\prod }_{\hexagon}{\sum }_{k=1}^{6}{c}_{\hexagon,k}\) reduces that to 2N(1 − 1/6) = 5N/3 fermions. Finally, each operator \({\hat{M}}_{j}\) creates one physical fermion less than it annihilates virtual ones, i.e., we are left with 2N/3 physical fermions in \(\left\vert {\psi }_{{{\rm{PEPS}}}}\right\rangle\). Hence, \(\vert {\psi }_{{{\rm{PEPS}}}}\rangle \propto \vert {\psi }_{2N/3}\rangle\), as claimed.
The PESS we have considered so far can be converted into a PEPS by realizing that the simplex states are of the form \(\left\vert 011111\right\rangle+\left\vert 101111\right\rangle+\ldots+\left\vert 111110\right\rangle\), also known as a W-state55, which can be written as a non-translationally invariant matrix product state of bond dimension 2 or a translationally invariant one of bond dimension 6. \(\hat{M}\) can be represented as a rank-3 tensor \({M}_{ab}^{i}\) with \({M}_{11}^{1}={M}_{10}^{0}=-{M}_{01}^{0}=1\) and all other elements equal zero. The resulting PEPS tensor has rank 5 and bond dimension 2 or 6, respectively, see Fig. 3. We note that a similar construction in terms of PESS defined on triangles can be used to describe the ground state of the Hamiltonian (2) for nearest-neighbor hopping only, an Euler system with one flat bottom band touched by two dispersive bands from above48,49.
a Matrix product state representation of the simplex states residing on the hexagons. A can either be chosen to be of bond dimension D = 2, with \({A}_{12}^{0}=1/\sqrt{6},\,{A}_{11}^{1}=-{A}_{22}^{1}=1\), Q21 = 1 and all other elements of A and Q equal to zero, or D = 6 with \({A}_{61}^{0}=1/\sqrt{6},\,{A}_{l,l+1}^{1}=1\) (l = 1, …, 5), all other elements of A equal to zero and \(Q={\mathbb{1}}\) (translationally invariant representation). Incoming arrows denote left and outgoing arrows right lower indices. b By combining two A tensors with the tensor M, we obtain the tensor T constituting the PEPS. c PEPS with one rank-5 tensor located on each site of the kagome lattice (gray dashed lines).
We now show that the PEPS is the unique ground state for − 2 < μ < 1: For chemical potential μ = − 2, the ground state subspace S is the intersection of all null spaces of h⬡, i.e., \(S={{\rm{span}}}{\{{\prod }_{\hexagon}{a}_{\hexagon}{a}_{\hexagon}^{{\dagger} }\left\vert n\right\rangle \}}_{n}={{\rm{span}}}{\{{\prod }_{\hexagon}{a}_{\hexagon}\left\vert n\right\rangle \}}_{n}\), where \({\{\left\vert n\right\rangle \}}_{n}\) is a complete basis and we used that \({h}_{\hexagon}^{\perp }={a}_{\hexagon}{a}_{\hexagon}^{{\dagger} }\) is a projector onto the orthogonal complement of h⬡. Each a⬡ eliminates one fermion from the basis state \(\left\vert n\right\rangle\). Hence, the state with the highest occupation in S is \({\prod }_{\hexagon}{a}_{\hexagon}\left\vert 1\ldots 1\right\rangle\), and it has filling N − N/3 = 2N/3. Any other state contained in S has lower expectation value of the overall occupation. For − 2 < μ < 1, this highest occupation ground state becomes the unique ground state, as all other states contained in S have lower overall occupation expectation value and thus get penalized. This example shows that PEPS can be the unique ground states of local gapped Hamiltonians with non-trivial two-dimensional crystalline topological features, even in the non-interacting limit. This contrasts with the inability of free-fermionic PEPS to capture any higher-dimensional topological labels of the ten-fold classification unless the PEPS have algebraically decaying correlations27,28,31,56.
Free-fermion generalizations
A straightforward generalization is obtained by modifying the simplex states to a linear combination, \(\left\vert {\tilde{\omega }}_{v}\right\rangle={\prod }_{\hexagon}\mathop{\sum }_{i=1}^{6}{\beta }_{i}{c}_{\hexagon,i}\left\vert {1}_{v}\right\rangle\) with \({\beta }_{i}\in {\mathbb{C}}\), and keeping \({\hat{M}}_{j}\) the same. The new PEPS is annihilated by \({\tilde{a}}_{\hexagon}={\sum }_{i\in \hexagon}{\beta }_{i}{a}_{i}\), where we set \(\mathop{\sum }_{i=1}^{6}| {\beta }_{i}{| }^{2}=1\). This corresponds to the new Hamiltonian
One can check that \({{{\mathcal{C}}}}_{2}{{\mathcal{T}}}\) symmetry implies \({\beta }_{i+3}^{*}={\beta }_{i}\) (i = 1, 2, 3) up to an irrelevant overall phase. For any choice of \({\{{\beta }_{i}\}}_{i=1}^{6}\), the corresponding PEPS is an Euler insulator, as the corresponding parent Hamiltonians are always gapped and continuously connected to the original one. Correspondingly, we have \(\{\tilde{{a}}_{\hexagon},{\tilde{a}}_{\hexagon^{{\prime} }}^{{\dagger} }\}={\delta }_{\hexagon,\hexagon^{{\prime} }}+| {\beta }_{\hexagon\cap \hexagon^{{\prime} }}{| }^{2}{\delta }_{\langle \hexagon,\hexagon^{{\prime} }\rangle }\), where there are always \(i=\hexagon\cap \hexagon^{{\prime} }\) for which the final term is non-vanishing.
Interacting generalizations
The modified simplex states give rise to non-interacting PEPS. We present two ways of generalizing these PEPS to the interacting regime. We will only briefly consider the first one, which employs a Gutzwiller projection and is likely to give rise to a new type of symmetry-enriched topological order57. The second example we consider applies symmetry-preserving shallow quantum circuits on the non-interacting PEPS. Therefore, it is guaranteed that the resulting interacting PEPSs are in the same topological phase as the original non-interacting one58.
To obtain the first type of interacting Euler state, we apply a Gutzwiller projection \({P}_{G}={\prod }_{i}({a}_{i\uparrow }^{{\dagger} }{a}_{i\uparrow }{a}_{i\downarrow }{a}_{i\downarrow }^{{\dagger} }+{a}_{i\uparrow }{a}_{i\uparrow }^{{\dagger} }{a}_{i\downarrow }^{{\dagger} }{a}_{i\downarrow })\) on two copies (denoted by ↑ and ↓) of the non-interacting PEPS, \(\vert {\psi }_{{{\rm{PEPS}}}}^{\uparrow }\rangle\) and \({{\mathcal{PH}}}\vert {\psi }_{{{\rm{PEPS}}}}^{\downarrow }\rangle\), where \({{\mathcal{PH}}}\) induces a particle-hole transformation. This is necessary, as the Gutzwiller projector PG enforces one fermion per site, i.e., in order for it not to annihilate the state, we need overall filling fraction 1 in the two PEPS copies. The resulting state
is expected to be an Euler quantum spin liquid and to have fractional statistics, similarly to the Gutzwiller projection of two p + ip superconducting states59. As the symmetries remain preserved under the Gutzwiller projection, \(\vert {\psi }_{{{\rm{PEPS}}}}^{S=1/2}\rangle\) is expected to be a symmetry-enriched topologically ordered state. This implies topological ground state degeneracy on the torus and anyonic excitations that survive breaking of the symmetry. If the protecting symmetry is not broken, additional features of non-interacting Euler insulators are expected, such as the cusp in the entanglement spectrum at K = 0 described below. However, studying the precise physical properties of this state goes beyond the scope of our work.
Second, we can also generalize the construction to interacting states by applying a shallow quantum circuit U of diagonal unitaries on the non-interacting PEPS, which makes it easy to ensure that \({{{\mathcal{C}}}}_{2}{{\mathcal{T}}}\) symmetry is preserved. Hence, by definition, we remain in the same topological phase. Furthermore, the new state will also be a PEPS of low bond dimension. We consider the simplest case of nearest-neighbor gates. We view these as being applied on all hexagons in a translationally invariant fashion. Within each hexagon, we label uj,j+1 as the unitary acting on sites j and j + 1 (j = 7 ≡ 1) inside a given hexagon, with sites enumerated as in Fig. 2. \({{{\mathcal{C}}}}_{2}{{\mathcal{T}}}\) symmetry is achieved if \({u}_{j,j+1}={u}_{j+3,j+4}^{*}\) for all j = 1, 2, 3. The simplest continuously tuneable case is \({u}_{j,j+1}={\mathbb{1}}-(1-{e}^{\pm i\alpha }){n}_{j}{n}_{j+1}\) with particle number operators nj, α ∈ [0, 2π), and positive (negative) sign for j = 1, 2, 3 (j = 4, 5, 6). The new PEPS is given by \(\left\vert {\psi }_{{{\rm{PEPS}}}}^{{\prime} }\right\rangle=U\left\vert {\psi }_{{{\rm{PEPS}}}}\right\rangle\) and the Hamiltonian gets transformed as
where we defined \({a}_{i}^{{\prime} }=U{a}_{i}{U}^{{\dagger} }\) and used that U commutes with \({n}_{i}={a}_{i}^{{\dagger} }{a}_{i}\). One can easily verify \({a}_{i}^{{\prime} {\dagger} }={a}_{i}^{{\dagger} }{\prod }_{j}^{\langle i,j\rangle }[{\mathbb{1}}-(1-{e}^{i{\sigma }_{\langle i,j\rangle }\alpha }){n}_{j}]\), where the product runs over all nearest neighbors of site i. σ〈i, j〉 = + 1 if 〈i, j〉 corresponds to one of the first three bonds in the hexagon that it lies in and − 1 if it corresponds to one of the last three bonds. This gives rise to the overall Hamiltonian
\({H}^{{\prime} }\) has the same spectrum as H for fixed μ, and \(\left\vert {\psi }_{{{\rm{PEPS}}}}^{{\prime} }\right\rangle\) is therefore its unique ground state for − 2 < μ < 1. The Hamiltonian is strictly local, acting on hexagons ⬡ and adjacent triangles. Its four-body interactions have amplitude \({{\mathcal{O}}}(\alpha )\) and higher-body interactions are of higher order. This is the first example of an interacting Euler insulator with a local gapped Hamiltonian. \(\vert {\psi }_{{{\rm{PEPS}}}}^{{\prime} }\rangle\) can be constructed by writing the phase matrix of \({u}_{j,j+1}={\mathbb{1}}-(1-{e}^{\pm i\alpha }){n}_{j}{n}_{j+1}\) as \({\sum }_{q=1}^{2}{R}_{q}^{ab}{R}_{q}^{cd}\) with \({R}_{1}^{ab}={\delta }_{ab}\) and \({R}_{2}^{ab}=\sqrt{-1+{e}^{\pm i\alpha }}{\delta }_{1a}{\delta }_{1b}\), a, b ∈ {0, 1}. (As the underlying operators are even, they can be decomposed into tensor products.) Four nearest-neighbor unitaries act on each site, such that the tensors of \(\vert {\psi }_{{{\rm{PEPS}}}}^{{\prime} }\rangle\) can be constructed by contracting the physical leg of T with four R-tensors, see Fig. 4. If the bond dimension of T was chosen to be 2, the interacting PEPS has bond dimension D = 4. We note that even though the protection of Euler topology under \({{{\mathcal{C}}}}_{2}{{\mathcal{T}}}\) symmetry in the presence of interactions is an open problem, in our case, we can guarantee that Euler topology is preserved, as our interacting PEPS \(\vert {\psi }_{{{\rm{PEPS}}}}^{{\prime} }\rangle\) and interacting model Hamiltonian are both related via a symmetry-preserving shallow quantum circuit to their non-interacting counterparts. A shallow quantum circuit does not change topological features, as these are global58. Hence, we are in the same symmetry-protected topological phase for all α ∈ [0, π]. This can also be seen from the fact that the Hamiltonians \({H}^{{\prime} }\) have the same energy spectrum (and gap) for all α.
Entanglement spectra
We numerically calculated the entanglement spectra of \(\vert {\psi }_{{{\rm{PEPS}}}}^{{\prime} }\rangle\) for an infinitely long torus with circumference Ly. That is, the torus is bipartitioned with Ly unit cells located around the perimeter of the resulting cylinder. We obtained its entanglement spectrum by calculating the non-interacting \(\vert {\psi }_{{{\rm{PEPS}}}}\rangle\) using TenPy60 and applying the quantum circuit U on it to obtain the interacting \(\vert {\psi }_{{{\rm{PEPS}}}}^{{\prime} }\rangle\). The entanglement spectra for various values of α and Ly = 6 are shown in Fig. 5. We observe that the low-lying part of the entanglement spectrum possesses a cusp at momentum K = 0, which remains intact as α is increased: Our non-interacting model corresponds to two degenerate Chern bands with opposite Chern number C = ± 1. Due to continuous connection to the non-interacting limit, we expect that our K = 0 mode in the interacting case is not isolated, but connected to two isolated branches symmetric around the K = 0 axis (gapless counter-propagating modes). Revealing these branches would require vastly larger system sizes, which are beyond the scope of our work. We further detail the entanglement features of the non-interacting case, including the stable cusp at K = 0, in the Methods.
For small values of α, the low-lying spectrum strongly resembles the non-interacting one (α = 0). In particular, a cusp at K = 0 (highlighted by a red circle) is preserved as α is increased. Parallel to this, new entanglement energies appear at the top of the spectrum and eventually merge with its low-lying part.
Discussion
We leverage quantum geometric conditions (see Methods and SI) to define a class of exact PEPS with finite topological Euler invariant. The enigmatic nature of the Euler class allows to circumvent no-go conditions. Importantly, these models can be generalized to interacting variants and have definite entanglement signatures. As such, these PEPS set a benchmark for new pursuits. These potential pursuits involve studying exotic excitations and spin liquids realized from Euler many-body PEPS ground states. In particular, on introducing interactions, novel kinds of fractionalizations should emerge from the interplay of the many-body entanglement as well as emergent quantum anomalous Hall states61,62. In addition, as all our states can be created by shallow quantum circuits from product states and have topological features, they are also particularly interesting for implementations on noisy intermediate-scale quantum devices and the development of new quantum error correction protocols. We will report on this in the near future.
Methods
Momentum-space characterization of the model in the non-interacting limit
We demonstrate how the model introduced in the main text, Eq. (2), can be decomposed in momentum space. Furthermore, in the SI, we showcase the ideal non-Abelian quantum geometry realized in the topological Euler bands in the non-interacting limit.
We first Fourier transform the real-space creation and annihilation operators to the basis of Bloch orbitals: \({\tilde{a}}_{\alpha,{{\bf{k}}}}^{{\dagger} }=\frac{1}{\sqrt{N}}{\sum }_{l}{a}_{\alpha,l}^{{\dagger} }{e}^{-i{{\bf{k}}}\cdot ({{{\bf{R}}}}_{l}+{{{\bf{r}}}}_{\alpha })}\), \({\tilde{a}}_{\alpha,{{\bf{k}}}}=\frac{1}{\sqrt{N}}{\sum }_{l}{a}_{\alpha,l}{e}^{i{{\bf{k}}}\cdot ({{{\bf{R}}}}_{l}+{{{\bf{r}}}}_{\alpha })}\). Here, the operator \({a}_{\alpha,l}^{({\dagger} )}\) annihilates (creates) a single particle in an atomic orbital α = A, B, C situated at position rα with respect to the position vector of a unit cell center Rl, where l = 1, 2, …, N/3 indexes unit cells. Due to the three sites per unit cell, we therefore have a three-band model. By inserting \({a}_{\alpha,l}=\frac{1}{2\pi \sqrt{N}}{\sum }_{{{\bf{k}}}}{\tilde{a}}_{\alpha,{{\bf{k}}}}{e}^{-i{{\bf{k}}}\cdot ({{{\bf{R}}}}_{l}+{{{\bf{r}}}}_{\alpha })}\) into Eq. (2) in the main text, one obtains
Here, the Bloch Hamiltonian for the considered system on the kagome lattice, manifestly expressed in a real gauge, reads48,49
with the corresponding (real) matrix elements, on setting \(t={t}^{{\prime} }={t}^{{\prime\prime} }=-1\) and k = (k1, k2),
We recognize that the Bloch Hamiltonian can be further rewritten as
or, more compactly,
with \({{\bf{n}}}({{\bf{k}}})={\left(\cos ({k}_{1}/2),\cos ({k}_{2}/2),\cos ({k}_{1}/2+{k}_{2}/2)\right)}^{{{\rm{T}}}}\). Importantly, under such decomposition, the topology of the Euler bands in any three-band Hamiltonian satisfying a reality condition [H(k) = H*(k)] can be captured by the normalized vector \(\hat{{{\bf{n}}}}({{\bf{k}}})={{\bf{n}}}({{\bf{k}}})/| | {{\bf{n}}}({{\bf{k}}})| |\). In particular, in the considered model, the vector \(\hat{{{\bf{n}}}}({{\bf{k}}})\) reads
and it fully determines the Euler curvature as
The Euler curvature can be viewed as a skyrmion density in the momentum-space texture, with the skyrmion being spanned by \(\hat{{{\bf{n}}}}\) over the Brillouin zone (BZ) square/torus. In particular, the Euler invariant is given by37,48
and obtains χ = 1 in the case of interest, which corresponds to the momentum-space meron (half-skyrmion) with the half-skyrmion number Q = 1/249. Additionally, the vector n(k) fully captures the band dispersion present in the model, as Eq. (21) can be written as
explicitly determining the band dispersion in the third band as E3(k) = ( − μ − 2) + 4∣∣n(k)∣∣2, contrary to the flat-band dispersion in the bottom Euler bands E1(k) = E2(k) = ( − μ − 2). The band energies given by such dispersions manifestly have a gap across the entire Brillouin zone, as the norm of the vector n(k) is non-vanishing ∣∣n(k)∣∣ > 0 at every k-point. This follows from the fact that the components of the vector n(k), \(\cos ({k}_{1}/2)\), \(\cos ({k}_{2}/2)\), and \(\cos ({k}_{1}/2+{k}_{2}/2)\), are not independent, with at least one of those terms being necessarily non-vanishing at any k-point.
Entanglement spectra of the non-interacting PEPS
Here we present the entanglement spectrum of the non-interacting kagome Euler model of the main text. For this purpose, we start with the momentum space Hamiltonian defined on a thin torus, i.e., Lx ≫ Ly. In the insulating state, the bottom two flat bands are occupied and the dispersive conduction band is empty. We first write down the projector on the occupied state
The projector by definition has its eigenvalues restricted to 0 and 1. For the calculations performed on a lattice, we define the real space positions as r = n1a1 + n2a2, where \({n}_{1(2)}\in {\mathbb{Z}}\) and a1(2) are the lattice vectors. The corresponding reciprocal space momenta take the values as \({{\bf{k}}}=\frac{{k}_{1}}{2\pi }{{{\bf{b}}}}_{1}+\frac{{k}_{2}}{2\pi }{{{\bf{b}}}}_{2}\), where k1, k2 ∈ ( − π, π] and b1(2) are the reciprocal lattice vectors. For the kagome model here, we have chosen, \({{{\bf{a}}}}_{1}=(\sqrt{3}/2,\,1/2)\) and a2 = (0, 1) as the lattice vectors. The corresponding reciprocal lattice vectors are \({{{\bf{b}}}}_{1}=(4\pi /\sqrt{3},\,0)\) and \({{{\bf{b}}}}_{2}=(-2\pi /\sqrt{3},\,2\pi )\). From the projector, we obtain a one-body correlation operator
Since G is also a projector, its eigenvalues are also restricted to 0 and 1. We partition the system into subsystems A and B, such that the entanglement spectrum between the two subsystems is given by the eigenvalues of the reduced density matrix ρA. The spectrum of the reduced density matrix ρA can then be obtained from the spectrum of the reduced correlation matrix GA defined as63
In Fig. 6a and c, we show the spectrum of the reduced one-body correlation matrix GA. The plots are obtained for system sizes Lx = 120 and Ly = 6 (12) for (a) and (c) respectively. The eigenvalues Λi(k2) of GA are bounded to lie in [0, 1], although, unlike the projector eigenvalues, they are not restricted to be 0 and 1. Indeed the in-gap eigenvalues are related to the topological Euler class of the model64. However, unlike the well-known case of Chern insulators, these in-gap modes in the one-body correlation spectrum of the Euler topology are not related to the physical edge states due to non-trivial topology, which typically has a spectral flow between the bulk conduction and valence bands. In Fig. 6(e) we explicitly show the absence of such topological edge states with a spectral flow between flat valence bands at − 1 and dispersive conduction band. The physical energy spectrum is calculated for system size Lx = 120 and Ly = 12 with open boundary conditions along Lx and periodic boundary conditions along Ly.
a One-body correlation spectrum Λi on a thin torus with Ly = 6, b Many-body entanglement spectrum ϵi on an Ly = 6 torus. The red marker at k = 0 and ϵi = 0 is the ground state of the partition `A'. c and d are the one-body correlation spectrum Λi and many-body entanglement spectrum ϵi for Ly = 12, respectively, to clarify the variation along k. e The physical spectrum E on a cylinder of size Lx = 120 and Ly = 12 does not show an edge state with a spectral flow between the flat valence bands and dispersive conduction band.
From the spectrum of GA, we obtain the entanglement spectrum of the non-interacting model using the relation
To calculate the full many-body entanglement spectrum as shown in the figure, we first obtain the ground state of subsection A by occupying 2/3 of the highest eigenvalues Λi, which is commensurate with the 2/3 filling of the whole system. Here, one should keep in mind that since the eigenvalues of the projector \(\hat{P}\) of the occupied states lie at 1 and unoccupied states at 0, while constructing the ground state of GA, one should start counting from Λi → 1 as occupied, and going down in the eigenvalues Λi corresponds to going up in the excitation spectrum. Once the ground state is identified, we obtain a many-body entanglement spectrum by creating excitations on this ground state. Since for the entanglement spectrum, we partition the system, while the filling fraction 2/3 is a constraint for the whole system, the subsystem A has eigenstates that have particle numbers different to 2/3 filling of the subsystem itself. Therefore to calculate the full many-body ground state, we first partition the subsystem A into different total particle number channels and from there create all possible particle-hole excitations. Then for each excited state configuration described by a fixed fermionic occupation number, we can obtain the entanglement spectrum using Eq. (29).
The many-body entanglement spectrum of the non-interacting model is shown in Fig. 6b and d for the system size Lx = 120 and Ly = 6(12) respectively. We have taken the ground state entanglement energy to zero as a reference, which is shown by the red marker at k = 0 in Fig. 6b and d. Notice the presence of the zero entanglement energy state at k = π. This is obtained by considering a channel with one less (or more) particle in subsystem A than the exact 2/3 filling.
Comparing the many-body entanglement spectrum to the interacting case with α = 0 in the main text (left panel in Fig. 5), we see a good agreement in their low energy features. In particular, in both cases, the ground state corresponds to k = 0 with the lowest entanglement energy. As we move to a finite k, the entanglement energy increases and eventually comes back down to the ground state value at k = π creating a cusp-like feature in the low-energy entanglement spectrum. This low energy behavior can be traced back to the in-gap modes (around Λ = 0.5) in the one-body correlation spectrum shown in Fig. 6a and c. In the one-body correlation spectrum, for each mode near 0, there are two modes near 1, and therefore 2/3 filling corresponds to occupying all modes in the upper half of the correlation spectrum. The low energy excitations are then created near Λi = 0.5 with a very low energy cost, which leads to many low energy modes in the entanglement spectrum.
Data availability
The datasets for the plots are available upon request.
Code availability
The codes for our simulations are available upon request. They are based on the publicly available TeNPy library60.
References
Verstraete, F. & Cirac, J. I. Matrix product states represent ground states faithfully. Phys. Rev. B 73, 094423 (2006).
Huang, Y. Area law in one dimension: Degenerate ground states and Renyi entanglement entropy, Preprint at https://arxiv.org/abs/1403.0327 (2014).
Molnar, A., Schuch, N., Verstraete, F. & Cirac, J. I. Approximating Gibbs states of local Hamiltonians efficiently with projected entangled pair states. Phys. Rev. B 91, 045138 (2015).
Dalzell, A. M. & Brandão, F. G. S. L. Locally accurate MPS approximations for ground states of one-dimensional gapped local Hamiltonians. Quantum 3, 187 (2019).
Huang, Y. Approximating local properties by tensor network states with constant bond dimension, Preprint at https://arxiv.org/abs/1903.10048 (2019).
White, S. R. & Scalapino, D. J. Density matrix renormalization group study of the striped phase in the 2D t−J model. Phys. Rev. Lett. 80, 1272 (1998).
Yan, S., Huse, D. A. & White, S. R. Spin-liquid ground state of the S=1/2 Kagome Heisenberg antiferromagnet. Science 332, 1201080 (2011).
Corboz, P. & Mila, F. Tensor network study of the Shastry-Sutherland model in zero magnetic field. Phys. Rev. B 87, 115144 (2013).
Corboz, P., Rice, T. M. & Troyer, M. Competing states in the t−J-model: uniform d-wave state versus stripe state. Phys. Rev. Lett. 113, 046402 (2014).
He, Y.-C., Zaletel, M. P., Oshikawa, M. & Pollmann, F. Signatures of Dirac cones in a DMRG study of the kagome Heisenberg model. Phys. Rev. X 7, 031020 (2017).
Gohlke, M., Wachtel, G., Yamaji, Y., Pollmann, F. & Kim, Y. B. Quantum spin liquid signatures in Kitaev-like frustrated magnets. Phys. Rev. B 97, 075126 (2018).
Pollmann, F., Turner, A. M., Berg, E. & Oshikawa, M. Entanglement spectrum of a topological phase in one dimension. Phys. Rev. B 81, 064439 (2010).
Schuch, N., Pérez-García, D. & Cirac, I. Classifying quantum phases using matrix product states and projected entangled pair states. Phys. Rev. B 84, 165139 (2011).
Williamson, D. J. et al. Matrix product operators for symmetry-protected topological phases: gauging and edge theories. Phys. Rev. B 94, 205150 (2016).
Wahl, T. B. Tensor networks demonstrate the robustness of localization and symmetry-protected topological phases. Phys. Rev. B 98, 054204 (2018).
Chan, A. & Wahl, T. B. Classification of symmetry-protected topological many-body localized phases in one dimension. J. Phys. Cond. Mat. 32, 305601 (2020).
Li, J., Chan, A. & Wahl, T. B. Classification of symmetry-protected topological phases in two-dimensional many-body localized systems. Phys. Rev. B 102, 014205 (2020).
Qi, X.-L. & Zhang, S.-C. Topological insulators and superconductors. Rev. Mod. Phys. 83, 1057 (2011).
Hasan, M. Z. & Kane, C. L. Colloquium. Rev. Mod. Phys. 82, 3045 (2010).
Fu, L. Topological Crystalline Insulators. Phys. Rev. Lett. 106, 106802 (2011).
Slager, R.-J., Mesaros, A., Juričić, V. & Zaanen, J. The space group classification of topological band-insulators. Nat. Phys. 9, 98 (2013).
Po, H. C., Vishwanath, A. & Watanabe, H. Symmetry-based indicators of band topology in the 230 space groups. Nat. Commun. 8, 50 (2017).
Bradlyn, B. et al. Topological quantum chemistry. Nature 547, 298 (2017).
Slager, R.-J. The translational side of topological band insulators. J. Phys. Chem. Solids 128, 24 (2019).
Bouhon, A., Bzdušek, T. & Slager, R.-J. Geometric approach to fragile topology beyond symmetry indicators. Phys. Rev. B 102, 115135 (2020).
Shiozaki, K. & Sato, M. Topology of crystalline insulators and superconductors. Phys. Rev. B 90, 165114 (2014).
Dubail, J. & Read, N. Tensor network trial states for chiral topological phases in two dimensions and a no-go theorem in any dimension. Phys. Rev. B 92, 205307 (2015).
Wahl, T. B., Tu, H.-H., Schuch, N. & Cirac, J. I. Projected entangled-pair states can describe chiral topological states. Phys. Rev. Lett. 111, 236805 (2013).
Yang, S., Wahl, T. B., Tu, H.-H., Schuch, N. & Cirac, J. I. Chiral projected entangled-pair state with topological order. Phys. Rev. Lett. 114, 106803 (2015).
Kitaev, A. Periodic table for topological insulators and superconductors. AIP Conf. Proc. 1134, 22 (2009).
Read, N. Compactly supported Wannier functions and algebraic K-theory. Phys. Rev. B 95, 115309 (2017).
Törmä, P. Essay: where can quantum geometry lead us? Phys. Rev. Lett. 131, 240001 (2023).
Bouhon, A., Timmel, A. & Slager, R.-J., Quantum geometry beyond projective single bands https://arxiv.org/abs/2303.02180 (2023)
Provost, J. & Vallee, G. Riemannian structure on manifolds of quantum states. Commun. Math. Phys. 76, 289 (1980).
Resta, R. The insulating state of matter: a geometrical theory. Euro. Phys. Jour. B 79, 121 (2011).
Verstraete, F. & Cirac, J. I. Renormalization algorithms for quantum-many body systems in two and higher dimensions, Preprint at https://arxiv.org/abs/cond-mat/0407066 (2004).
Bouhon, A. et al. Non-Abelian reciprocal braiding of Weyl points and its manifestation in ZrTe. Nat. Phys. 16, 1137 (2020).
Ahn, J., Park, S. & Yang, B.-J. Failure of Nielsen-Ninomiya theorem and fragile topology in two-dimensional systems with space-time inversion symmetry: application to twisted bilayer graphene at magic angle. Phys. Rev. X 9, 021013 (2019).
Bouhon, A., Black-Schaffer, A. M. & Slager, R.-J. Wilson loop approach to fragile topology of split elementary band representations and topological crystalline insulators with time-reversal symmetry. Phys. Rev. B 100, 195135 (2019).
Wu, Q., Soluyanov, A. A. & Bzdušek, T. Non-Abelian band topology in noninteracting metals. Science 365, 1273 (2019).
Slager, R.-J., Bouhon, A. & Ünal, F. N. Non-Abelian Floquet braiding and anomalous Dirac string phase in periodically driven systems. Nat Commun 15, 1144 (2024).
Ünal, F. N., Bouhon, A. & Slager, R.-J. Topological Euler class as a dynamical observable in optical lattices. Phys. Rev. Lett. 125, 053601 (2020).
Zhao, W. et al. Quantum simulation for topological Euler insulators. Commun. Phys. 5, 223 (2022).
Peng, B., Bouhon, A., Monserrat, B. & Slager, R.-J. Phonons as a platform for non-Abelian braiding and its manifestation in layered silicates. Nat. Commun. 13, 423 (2022).
Peng, B., Bouhon, A., Slager, R.-J. & Monserrat, B. Multigap topology and non-Abelian braiding of phonons from first principles. Phys. Rev. B 105, 085115 (2022).
Bouhon, A., Lange, G. F. & Slager, R.-J. Topological correspondence between magnetic space group representations and subdimensions. Phys. Rev. B 103, 245127 (2021).
Lee, S. H., Qian, Y. & Yang, B.-J. Euler band topology in spin-orbit coupled magnetic systems, Preprint at https://arxiv.org/abs/2404.16383 (2024).
Jiang, B. et al. Experimental observation of non-Abelian topological acoustic semimetals and their phase transitions. Nat. Phys. 17, 1239 (2021).
Jiang, B. et al. Observation of an acoustic topological Euler insulator with meronic waves. Science Bulletin 69, 1653 (2024).
Guo, Q. et al. Experimental observation of non-Abelian topological charges and edge states. Nature 594, 195 (2021).
Kitaev, A. Y. Unpaired Majorana fermions in quantum wires. Physics-Uspekhi 44, 131 (2001).
Hackenbroich, A., Bernevig, B. A., Schuch, N. & Regnault, N. Fermionic tensor networks for higher-order topological insulators from charge pumping. Phys. Rev. B 101, 115134 (2020).
Xie, Z. et al. Tensor renormalization of quantum many-body systems using projected entangled simplex states. Phys. Rev. X 4, 011025 (2014).
Kraus, C. V., Schuch, N., Verstraete, F. & Cirac, J. I. Fermionic projected entangled pair states. Phys. Rev. A 81, 052338 (2010).
Dür, W., Vidal, G. & Cirac, J. I. Three qubits can be entangled in two inequivalent ways. Phys. Rev. A 62, 062314 (2000).
Wahl, T. B., Haßler, S. T., Tu, H.-H., Cirac, J. I. & Schuch, N. Symmetries and boundary theories for chiral projected entangled pair states. Phys. Rev. B 90, 115133 (2014).
Chen, X., Gu, Z.-C., Liu, Z.-X. & Wen, X.-G. Symmetry protected topological orders and the group cohomology of their symmetry group. Phys. Rev. B 87, 155114 (2013).
Chen, X., Gu, Z.-C. & Wen, X.-G. Local unitary transformation, long-range quantum entanglement, wave function renormalization, and topological order. Phys. Rev. B 82, 155138 (2010).
Tu, H.-H. Projected BCS states and spin Hamiltonians for the SO(n)1 Wess-Zumino-Witten model. Phys. Rev. B 87, 041103 (2013).
Hauschild, J. & Pollmann, F. Efficient numerical simulations with Tensor Networks: Tensor Network Python (TeNPy), SciPost Phys. Lect. Notes, 5 https://doi.org/10.21468/SciPostPhysLectNotes.5 (2018), code available from https://github.com/tenpy/tenpy
Bouhon, A. & Slager, R.-J. Multi-gap topological conversion of Euler class via band-node braiding: minimal models, PT-linked nodal rings, and chiral heirs, Preprint at https://arxiv.org/abs/2203.16741 (2022).
Jo, N. H. et al. Intrinsic axion insulating behavior in antiferromagnetic MnBi6Te10. Phys. Rev. B 102, 045130 (2020).
Peschel, I. Calculation of reduced density matrices from correlation functions. J. Phys. A Math. Gen. 36, L205 (2003).
Takahashi, R. & Ozawa, T. Bulk-edge correspondence of Stiefel-Whitney and Euler insulators through the entanglement spectrum and cutting procedure. Phys. Rev. B 108, 075129 (2023).
Acknowledgements
We acknowledge funding from a New Investigator Award, EPSRC grant EP/W00187X/1, an EPSRC ERC underwrite grant EP/X025829/1, and a Royal Society exchange grant IES/R1/221060 as well as Trinity College, Cambridge (R.-J.S., G.C., and T.B.W.). This project was also supported by funding from the Rod Smallwood Studentship at Trinity College, Cambridge (W.J.J.).
Author information
Authors and Affiliations
Contributions
R.-J.S. and T.B.W. initiated the project. T.B.W. constructed the PEPS and performed the numerical simulations with input on the geometry and specific models from R.-J.S, W.J.J., A.B and G.C. W.J.J. performed the momentum-space characterisation of the model and quantum geometry with inputs from R.J.S., T.B.W., and A.B. G.C. numerically benchmarked the entanglement of the free fermion spectra. All authors discussed the results. T.B.W. with input of R.-J.S. took the lead in the writing of the manuscript, but the final form of the manuscript benefitted from input from all authors.
Corresponding authors
Ethics declarations
Competing interests
The authors declare no competing interests.
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.
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 4.0 International License, which permits use, sharing, adaptation, 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 changes were made. 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/4.0/.
About this article
Cite this article
Wahl, T.B., Jankowski, W.J., Bouhon, A. et al. Exact projected entangled pair ground states with topological Euler invariant. Nat Commun 16, 284 (2025). https://doi.org/10.1038/s41467-024-55484-4
Received:
Accepted:
Published:
DOI: https://doi.org/10.1038/s41467-024-55484-4