Abstract
Flat bands amplify correlation effects and are of extensive current interest. They provide a platform to explore both topology in correlated settings and correlation physics enriched by topology. Recent experiments in correlated kagome metals have found evidence for strange-metal behavior. A major theoretical challenge is to study the effect of local Coulomb repulsion when the band topology obstructs a real-space description. In a variant to the kagome lattice, we identify an orbital-selective Mott transition in any system of coupled topological flat and wide bands. This was made possible by the construction of exponentially localized and Kramers-doublet Wannier functions, which, in turn, leads to an effective Kondo-lattice description. Our findings show how quasiparticles are formed in such coupled topological flat-wide band systems and, equally important, how they are destroyed. Our work provides a conceptual framework for the understanding of the existing and emerging strange-metal properties in kagome metals and beyond.
Coupled topological flat and wide bands realize a metallic quantum critical point with a loss of quasiparticles.
INTRODUCTION
In flat electronic bands, Coulomb interactions are proportionally enhanced because of their reduced kinetic energy. As such, flat-band systems serve as a platform for strong correlation physics (1). Heavy fermion metals represent a canonical case of flat bands formed from highly localized atomic orbitals, and, indeed, they display rich correlation physics such as quantum criticality, strange metallicity, and unconventional superconductivity (2, 3). Studies here have led to the notion that quasiparticles are lost in strange metals (4–8). Another case, emerging in a growing list of materials, corresponds to flat bands formed by geometrical interference (9–13); such bands are often topological. These materials represent a playground to study both the strong correlations and topology (14–18) and have been found to show unusual properties such as exotic forms of charge density wave order (15, 19–23). Recently, experimental evidence for strange-metal behavior has emerged from correlated kagome metals (24, 25), which contains both flat and wide bands that intersect with each other. The observed behavior takes the form of a T-linear resistivity or a single-particle damping rate that is linear in frequency. This raises the question of how quasiparticles can be destroyed in such systems. The question is important not only for correlated kagome metals but also for related flat band settings such as moiré systems (26, 27).
Thus motivated, here we study the topological flat bands coupled to wide bands. Our work provides the first theoretical demonstration of an orbital-selective Mott transition in any system that involves coupled topological flat-wide bands. The orbital-selective Mott transition provides a framework to understand the quasiparticles’ formation and, equally important, their destruction. The latter allows for the understanding of the existing and emerging strange-metal properties of flat-band systems. Our work also connects the topological flat band systems to the orbital-selective physics of bulk correlated materials (28–30), in which the atomic orbitals exhibit different dispersions and yet [in contrast to models with fully decoupled orbitals (31)] are kinetically coupled with each other.
More specifically, the topological nature of the flat bands makes it difficult to describe them in terms of any localized orbitals. The band topology obstructs the formulation of exponentially localized and Kramers-doublet Wannier orbitals (32). This poses a challenge for treating the effect of sizable spatially local Coulomb repulsion and for connecting the behavior of the coupled topological flat-wide bands to the orbital-selective Mott physics.
Here, we make progress by considering a lattice with a lower symmetry, which retains the central issue of topological obstruction while allowing—in a particularly transparent way—for the construction of exponentially localized and Kramers-doublet Wannier orbitals. The lattice, as illustrated in Fig. 1A, is a variant of the kagome lattice. We are able to construct the Wannier centers, which turn out to form a triangular lattice (Fig. 1B). When the effective local Coulomb repulsion is larger than the width of the flat band (Dflat) but smaller than that of the wide band (Dwide), we identify a continuous orbital-selective Mott transition [a quantum critical point (QCP)] (Fig. 1C). The two involved ground states respectively feature a “large” and a “small” Fermi surface (expanded form of FS); in an effective Kondo-lattice formulation that we construct, the large Fermi surface is Kondo driven; the small Fermi surface, then, develops from a Kondo destruction (6–8). By analogy with the phase diagram of heavy fermion metals (1–3), the existence of the phases with large and small Fermi surfaces (33–35) opens up a regime of amplified quantum fluctuations (6) for beyond-Landau quantum criticality and the accompanying strange-metal behavior (36). Thus, our work provides a conceptual framework to address the aforementioned strange-metal behavior in the flat band–based metals. The connection with a Kondo-lattice description is also being pursued in moiré systems (37–40). The approach taken here is expected to shed light on the quantum phases and their transitions in those systems.
RESULTS
The lattice contains five sublattices marked as A, B, C, D, E (Fig. 1A). The model is written as , which contains the on-site Hubbard interactions
(1) |
as well as the noninteracting Hamiltonian . Included in are the nearest-neighbor tight binding hopping parameter t between two sites that are connected by a solid line in Fig. 1A, from an η electron (dz2 orbital) located at site r, sublattice i ∈ (A, B, C, D, and E) with spin σ, to its counterpart at r, j ∈ (A, B, C, D, and E) and spin σ; a chemical potential μ; a potential difference m between the sublattices C, D, and E and the sublattices A and B; and an additional potential difference γ between the sublattices D and E and the sublattice C (see Materials and Methods). To illustrate our case, we consider t = 1 = m without loss of generality.
To analyze a more tractable model with a lower symmetry, we focus on the case of a nonzero γ. Here, a C3z of the γ = 0 model (see Materials and Methods) is broken, and the middle band is not entirely flat, as seen in Fig. 2 (A and B). However, for small γ (which we illustrate with the case of γ = −0.1 t), the middle band remains relatively flat (the bandwidth is about 0.06 t) and, thus, will still be referred to as a flat band. Near Γ, there is a linear crossing between the flat and dispersive bands along the Γ − K line. The node is located near the Γ point and is protected by the Mx symmetry. When a spin-orbit coupling is further included (see the Supplementary Materials, fig. S1E), such a node would be gapped out, and the flat band acquires ±1 spin Chern number.
In this original lattice basis, the flat band comes from a linear superposition of the atomic states located in the five different sublattices. A single atomic–like representation of the flat band is needed to make progress, which we now turn to.
We will consider the Hubbard interaction U that is large compared to the width of the flat band but small compared to the width of the wide bands (Fig. 1C). Such a range is of interest to the experimentally studied systems mentioned earlier. Since there is an energy gap between the top three bands and the bottom two, we focus on the top three bands that include the flat band.
To express a band in terms of exponentially localized and Kramers-doublet Wannier orbitals, the Bloch states must have the same Mx-symmetry eigenvalue at all high-symmetry points (41, 42). In our case, the topological obstruction can be recognized by noticing that the different high-symmetry points of the flat band have different eigenvalues of the Mx symmetry. As shown in the Supplementary Materials (band 3 in table S1) and illustrated in Fig. 2B, the eigenvalue changes from +1 at K and M′ to −1 at γ. Similar changes of Mx-symmetry eigenvalues are also observed for the two wide bands (bands 1 and 2 in table S1). We therefore consider a combination of Bloch states so that the +1 eigenvalue portion of the flat band combines with the +1 eigenvalue portion of the wide band (band 2 in table S1 and illustrated in Fig. 2).
Through explicit construction, we indeed show the validity of the procedure. We find the exponentially localized Wannier functions, which are located on a triangular lattice (Fig. 1B), with the Mx-symmetry eigenvalue being +1. We call such a Wannier function a d orbital. The hopping integral of the two neighboring d orbitals is indeed weak in our construction: It turns out to be on the order of 0.03 t.
The remaining degrees of freedom can be represented by two Wannier orbitals with Mx-symmetry eigenvalue +1 and −1, respectively. These two orbitals mainly come from the top two dispersive bands, and we label them as c orbitals.
We have thus succeeded in constructing a single atomic–like state per unit cell to represent the flat band. The model takes the form of an effective multiorbital Hubbard model on a triangular lattice, which can also be viewed as an effective Anderson-lattice model. It is expressed in terms of the d and c orbitals, containing the kinetic term H0 and the interaction term HI.
The kinetic term, H0, is given in detail in the Supplementary Materials. It involves two types of electrons, with creating a d electron (electron in the orbital d), which represents the flat band’s degrees of freedom, at site R with spin σ. Similarly, creates a conduction c electron (electron in orbital c), which represents the itinerant degrees of freedom, at site R with spin σ and orbital a = 1,2. H0 contains the part Hd, describing the energy level (Ed) and their negligibly small hopping parameters; the part Hc, describing the inter-c-electron hopping parameters, and the energy levels Ea (a = 1,2); and the part Hdc, the d-c hybridization term with matrix element VR−R′,a,σ. These parameters are specified in the Supplementary Materials. As is seen there, the dominant hybridization is with one of the two conduction electron bands, a = 1. The corresponding hybridization is off-site because cR,1,σ is mirror odd.
The most important interactions here include the Hubbard interactions of the d electrons (Hu), and the Hund’s coupling between the d and c electrons (HHund). The interactions are labeled u and J1,2, respectively. The specific forms of the interactions are given in the Supplementary Materials.
In the limit of , the Hubbard interactions suppress the charge fluctuations of the d electrons and turn them into quantum spins. Correspondingly, the effective model acquires a representation in terms of a Kondo-lattice Hamiltonian on the same triangular lattice (see Materials and Methods). It contains two Kondo couplings JK,a (a = 1,2) to two conduction c electron bands and an inter-moment exchange coupling JH between the d spins. Both couplings are controlled by the parameter
(2) |
with ϕ0 being of order unity (see Materials and Methods and section SD).
For convenience in notation, we now treat of Eq. 2 as our tuning parameter (we note, however, that tuning is equivalent to varying the original Hubbard interaction U). At smaller , the Kondo effect dominates. Increasing enhances the relative effect of the antiferromagnetic exchange interactions between the local moments, which favors the correlations between the local moments and is detrimental to the development of the Kondo effect.
We have analyzed this competition through a set of saddle-point equations that are realized in a large-N limit (see Materials and Methods). We use a pseudo-fermion representation of the spin and solve the saddle-point equations in terms of the field ζR,R′a, which represents the hybridization of the Kondo-driven composite fermions and conduction-c fermions, and χR,R′, which characterizes the inter-moment spin singlets (43).
Figure 3 shows the resulting phase diagram. At small , we realize a topological heavy Fermi-liquid phase via the condensation of the hybridization field ζ. The latter converts the local moments into composite fermions that are represented by the f fields, which hybridize with the conduction electrons. Furthermore, the nonzero hybridization ζR1a2 attaches a +1 Mx-symmetry eigenvalue to the composite fermion f field, which is the opposite to the eigenvalue of the dominantly hybridizing conducting c1 band (see section SG). The correlated heavy bands are topological containing a Dirac node, as shown in Fig. 4A. The node is protected by Mx and the SU(2) symmetry: Because of the formation of the hybridized bands, the Fermi surface encloses the electrons residing on both the conduction and flat bands (44, 45). This is the large Fermi surface, which is shown in Fig. 4B.
Increasing leads to a Kondo destruction (6–8): The local moments couple to each other, and they no longer form a Kondo singlet with the spins of the conduction electrons. Consequently, the Fermi surface is derived from the dispersion of the conduction electrons, plotted in Fig. 4C. It is shown in Fig. 4D and is small in the sense that the Fermi surface encloses only the conduction electrons. The frustrated nature of the effective lattice (Fig. 1B) has led to the Kondo-destroyed phase via the bond variables χR,R′. However, the physical mechanism leading to the Kondo destruction is more universal, as characterized by a global phase diagram for the competition between the Kondo and RKKY/Heisenberg interactions (1).
DISCUSSION
We have demonstrated the first realization of orbital-selective Mott transition in models of topological flat bands in the presence of coupling with wide bands. The large-Fermi surface state provides a proper description for the development of the strongly correlated d-electron quasiparticles. In turn, this sets the stage for the transition into the small-Fermi surface phase, which represents the destruction of the correlated quasiparticles. This orbital-selective Mott transition is in the same universality class as the Kondo-destruction quantum criticality of heavy fermion metals (4, 6–8, 46), associated with which strange metallicity develops. Our theoretical description, thus, provides the conceptual framework to understand the recently emerged experimental evidence for strange metal behavior in correlated kagome metals (24, 25).
We close with several remarks on the generality and implications of our results. First, our Wannier construction procedure works equally well for the case with a spin-orbit coupling. Accordingly, our Kondo-lattice construction and the results for the quasiparticle formation and destruction readily apply in this case. Second, by emphasizing the interplay and coupling between the flat and wide bands, our work brings out an analogy between the correlated quantum materials that host topological flat bands (10–12, 15, 22, 23) and moiré systems (37–40), in which the strange-metal behavior has also been indicated (26). Compared to the topology of the graphene-based moiré flat bands (38, 47), the bands in our case share the traits of being topological but have the distinction of being much simpler. This simplification represents a crucial advantage in allowing us to provide a proof-of-principle demonstration of both the emergence and destruction of the proper quasiparticles in systems with coupled topological flat-wide bands through the Kondo route. Our explicit construction of the Kondo lattice also sets the stage for the study of a global quantum-phase diagram (1) in such topological flat band–anchored systems. Third, the topological nature of the flat band makes our Kondo lattice topological. This topological feature distinguishes our current model from the conventional Kondo-lattice models where the bands are usually topologically trivial. This topological feature naturally makes the heavy Fermi-liquid phase topologically nontrivial, featuring a Dirac node. Beyond this, at the QCP and in the non–Fermi-liquid region, the incoherent electronic excitations could also show nontrivial topology. However, any such topological feature goes beyond the quasiparticle description and is left for future studies.
We now discuss the stability of the QCP and the corresponding non–Fermi-liquid behavior. First, at finite N, the Kondo destruction part of the phase diagram could contain a magnetically ordered phase. The stability of the Kondo destruction QCP and the corresponding non–Fermi-liquid behavior remain robust, as has been shown in the Kondo-lattice model (4, 46). Second, with a nonzero spin-orbit coupling, the SU(2) spin symmetry will be broken. However, the existence of the Kondo destruction critical behavior relies on the competition of local moment fluctuations and Kondo effect instead of the specific symmetry of the local moment (4, 46, 48). Therefore, the Kondo destruction QCP and the non–Fermi-liquid behavior that we found here are robust.
In conclusion, we have advanced a realistic model to study the effect of local Coulomb repulsion for a system of coupled topological flat and wide bands. By constructing exponentially localized and Kramers-doublet Wannier functions for these bands, we are able to formulate a Kondo-lattice description. This has allowed us to provide the first demonstration of an orbital-selective Mott transition in any system of coupled topological flat and wide bands. The orbital-selective Mott transition provides a characterization for both the development and destruction of quasiparticles, leading to quantum phases with large and small Fermi surfaces. Our work provides a conceptual framework to describe the amplified quantum fluctuations of multiband systems with topological flat bands, sets the stage to understand the strange-metal properties that are emerging in kagome metals and other flat-band systems, and uncovers a linkage between these systems and both the f and d electron–based correlated bulk materials.
MATERIALS AND METHODS
The Hubbard model, , is defined on a variant of the kagome lattice defined in Fig. 1A, with the onsite interaction term, , given in Eq. 1. The noninteracting Hamiltonian is written as follows
(3) |
Here, creates an electron at site r, sublattice i ∈ (A, B, C, D, E) with spin σ.
At γ = 0, the system has a C3z rotational symmetry, an Mx mirror symmetry, and also an SU(2) spin rotational symmetry. This leads to a purely flat band as shown in the Supplementary Materials (fig. S1B). There is a quadratic band touching between the flat and a dispersive band at the center of the Brillouin zone, Γ = (0,0). The crossing is protected by both the Mx and C3z symmetries. To analyze a more tractable model with a lower symmetry, we focus on the case of a nonzero γ, as described in the Results section.
In the limit of , the charge fluctuations of the d electrons are suppressed, and these electrons are turned into quantum spins. By integrating out the high-energy degrees of freedom [in the presence of the Hund’s coupling, (49, 50)], we reach a Kondo-Heisenberg model with the Hund’s coupling. The effective Kondo-lattice Hamiltonian on the triangular lattice is
(4) |
Here, is the emergent spin-1/2 local moment formed by the localized d electrons. In the large u limit, the hoppings of d electrons induce a Heisenberg interaction of strength , where is specified by Eq. 2. The nonlocal hybridization terms between the d and c electrons lead to nonlocal Kondo couplings of strength . Here, the local moment of the d electron is Kondo coupled to a spin operator that are formed by two electron operators from the sites R1, R2 and the orbitals a1, a2 respectively. Last, the Hamiltonians for the c electrons and for the Hund’s coupling remain unchanged from those given in the effective multiorbital Hubbard (i.e., the effective Anderson lattice) model. In particular, the Hund’s coupling acts between the d and c electrons of the same site (see section SC). To analyze the competition between the inter-moment and Kondo exchange couplings, we introduce the pseudo-fermion representation of the spin operators and solve the model in the large-N limit [with a generalization of SU(2) to SU(N) and a suitable rescaling of the coupling constants in terms of 1/N; see section SF].
The ground state is characterized by the bond fields χR,R′ (43) and the hybridization fields ζR,R′a
(5) |
and their maximum amplitudes ζ = maxR,R′,a{∣ζR,R′a∣} and χ = maxR,R′{∣χR,R′∣}. The results from solving the saddle-point equations have been given in the Results section. The Kondo destruction is captured by the suppression of the hybridization fields ζ, which appears not only in the SU(2)-symmetric setting but also in the cases with spin anisotropy (4, 46, 48).
Acknowledgments
We thank L. Chen, S. Paschen, C. Setty, M. Yi, and especially G. Aeppli for useful discussions.
Funding: The work was primarily supported by the U.S. DOE, BES, under award no. DE-SC0018197 (conceptualization for and construction of Wannier orbitals), by the Air Force Office of Scientific Research under grant no. FA9550-21-1-0356 (conceptualization for and determination of phase diagram), and additionally supported by the Robert A. Welch Foundation grant no. C-1411. Most of the computational calculations have been performed on the Shared University Grid at Rice funded by NSF under grant EIA-0216467, a partnership between Rice University, Sun Microsystems, and Sigma Solutions Inc., the Big-Data Private-Cloud Research Cyberinfrastructure MRI-award funded by NSF under grant no. CNS-1338099, and by Rice University and the Extreme Science and Engineering Discovery Environment (XSEDE) by NSF under grant no. DMR160057. The work of Q.S. was performed in part at the Aspen Center for Physics, which is supported by the NSF grant no. PHY-1607611.
Author contributions: Conceptualization: H.H. and Q.S. Methodology: H.H. and Q.S. Investigation: H.H. and Q.S. Supervision: Q.S. Writing—original draft: H.H. and Q.S. Writing—review and editing: H.H. and Q.S.
Competing interests: The authors declare that they have no competing interests.
Data and materials availability: All data needed to evaluate the conclusions in the paper are present in the paper and/or the Supplementary Materials. Additional data that have been used can be found at https://doi.org/10.5281/zenodo.8036791.
Supplementary Materials
This PDF file includes:
REFERENCES AND NOTES
- 1.S. Paschen, Q. Si, Quantum phases driven by strong correlations. Nat. Rev. Phys. 3, 9–26 (2021). [Google Scholar]
- 2.P. Coleman, A. J. Schofield, Quantum criticality. Nature 433, 226–229 (2005). [DOI] [PubMed] [Google Scholar]
- 3.S. Kirchner, S. Paschen, Q. Chen, S. Wirth, D. Feng, J. D. Thompson, Q. Si, Colloquium: Heavy-electron quantum criticality and single-particle spectroscopy. Rev. Mod. Phys. 92, 011002 (2020). [Google Scholar]
- 4.H. Hu, L. Chen, Q. Si, Quantum critical metals: Dynamical planckian scaling and loss of quasiparticles. arXiv:2210.14183 [cond-mat.str-el] (2022).
- 5.P. W. Phillips, N. E. Hussey, P. Abbamonte, Stranger than metals. Science 377, eabh4273 (2022). [DOI] [PubMed] [Google Scholar]
- 6.Q. Si, S. Rabello, K. Ingersent, J. L. Smith, Locally critical quantum phase transitions in strongly correlated metals. Nature 413, 804–808 (2001). [DOI] [PubMed] [Google Scholar]
- 7.P. Coleman, C. Pépin, Q. Si, R. Ramazashvili, How do Fermi liquids get heavy and die? J. Phys. Condens. Matter 13, R723–R738 (2001). [Google Scholar]
- 8.T. Senthil, M. Vojta, S. Sachdev, Weak magnetism and non-fermi liquids near heavy-fermion critical points. Phys. Rev. B 69, 035111 (2004). [Google Scholar]
- 9.A. Mielke, Ferromagnetic ground states for the hubbard model on line graphs. J. Phys. A: Math. Gen. 24, L73–L77 (1991). [Google Scholar]
- 10.L. Ye, M. Kang, J. Liu, F. von Cube, C. R. Wicker, T. Suzuki, C. Jozwiak, A. Bostwick, E. Rotenberg, D. C. Bell, L. Fu, R. Comin, J. G. Checkelsky, Massive dirac fermions in a ferromagnetic kagome metal. Nature 555, 638–642 (2018). [DOI] [PubMed] [Google Scholar]
- 11.M. Yao, H. Lee, N. Xu, Y. Wang, J. Ma, O. V. Yazyev, Y. Xiong, M. Shi, G. Aeppli, Y. Soh, Switchable Weyl nodes in topological kagome ferromagnet Fe3Sn2. arXiv:1810.01514 [cond-mat.str-el] (2018).
- 12.M. Kang, S. Fang, L. Ye, H. C. Po, J. Denlinger, C. Jozwiak, A. Bostwick, E. Rotenberg, E. Kaxiras, J. G. Checkelsky, R. Comin, Topological flat bands in frustrated kagome lattice CoSn. Nat. Commun. 11, 4004 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.C. Barreteau, F. Ducastelle, T. Mallah, A bird’s eye view on the flat and conic band world of the honeycomb and kagome lattices: Towards an understanding of 2D metal-organic frameworks electronic structure. J. Phys. Condens. Matter 29, 465302 (2017). [DOI] [PubMed] [Google Scholar]
- 14.C. Setty, H. Hu, L. Chen, Q. Si, Electron correlations and T-breaking density wave order in a ℤ2 kagome metal. arXiv:2105.15204 [cond-mat.str-el] (2021).
- 15.C. Setty, C. A. Lane, L. Chen, H. Hu, J.-X. Zhu, Q. Si, Electron correlations and charge density wave in the topological kagome metal FeGe. arXiv:2203.01930 [cond-mat.str-el] (2022).
- 16.S. D. Huber, E. Altman, Bose condensation in flat bands. Phys. Rev. B 82, 184502 (2010). [Google Scholar]
- 17.T. Mizoguchi, Y. Kuno, Y. Hatsugai, Construction of interacting flat-band models by molecular-orbital representation: Correlation functions, energy gap, and entanglement. Prog. Theor. Exp. Phys. 2022, 023I02 (2022). [Google Scholar]
- 18.Z. Gulácsi, A. Kampf, D. Vollhardt, Exact many-electron ground states on diamond and triangle hubbard chains. Prog. Theor. Phys. Suppl. 176, 1–21 (2008). [DOI] [PubMed] [Google Scholar]
- 19.Y.-X. Jiang, J.-X. Yin, M. M. Denner, N. Shumiya, B. R. Ortiz, G. Xu, Z. Guguchia, J. He, M. S. Hossain, X. Liu, J. Ruff, L. Kautzsch, S. S. Zhang, G. Chang, I. Belopolski, Q. Zhang, T. A. Cochran, D. Multer, M. Litskevich, Z.-J. Cheng, X. P. Yang, Z. Wang, R. Thomale, T. Neupert, S. D. Wilson, M. Z. Hasan, Unconventional chiral charge order in kagome superconductor KV3Sb5. Nat. Mater. 20, 1353–1357 (2021). [DOI] [PubMed] [Google Scholar]
- 20.C. Mielke III, D. Das, J.-X. Yin, H. Liu, R. Gupta, Y.-X. Jiang, M. Medarde, X. Wu, H. C. Lei, J. Chang, P. Dai, Q. Si, H. Miao, R. Thomale, T. Neupert, Y. Shi, R. Khasanov, M. Z. Hasan, H. Luetkens, Z. Guguchia, Time-reversal symmetry-breaking charge order in a kagome superconductor. Nature 602, 245–250 (2022). [DOI] [PubMed] [Google Scholar]
- 21.S. Zhou, Z. Wang, Doped orbital Chern insulator, Chern Fermi pockets, and chiral topological pair density wave in kagome superconductors. arXiv:2110.06266 [cond-mat.supr-con] (2021).
- 22.X. Teng, L. Chen, F. Ye, E. Rosenberg, Z. Liu, J. X. Yin, Y. X. Jiang, J. S. Oh, M. Z. Hasan, K. J. Neubauer, B. Gao, Y. Xie, M. Hashimoto, D. Lu, C. Jozwiak, A. Bostwick, E. Rotenberg, R. J. Birgeneau, J. H. Chu, M. Yi, P. Dai, Discovery of charge density wave in a kagome lattice antiferromagnet. Nature 609, 490–495 (2022). [DOI] [PubMed] [Google Scholar]
- 23.J.-X. Yin, Y. X. Jiang, X. Teng, M. S. Hossain, S. Mardanya, T. R. Chang, Z. Ye, G. Xu, M. M. Denner, T. Neupert, B. Lienhard, H. B. Deng, C. Setty, Q. Si, G. Chang, Z. Guguchia, B. Gao, N. Shumiya, Q. Zhang, T. A. Cochran, D. Multer, M. Yi, P. Dai, M. Z. Hasan, Discovery of charge order and corresponding edge state in kagome magnet FeGe. Phys. Rev. Lett. 129, 166401 (2022). [DOI] [PubMed] [Google Scholar]
- 24.L. Ye, S. Fang, M. G. Kang, J. Kaufmann, Y. Lee, J. Denlinger, C. Jozwiak, A. Bostwick, E. Rotenberg, E. Kaxiras, D. C. Bell, O. Janson, R. Comin, J. G. Checkelsky, A flat band-induced correlated kagome metal. arXiv:2106.10824 [cond-mat.mtrl-sci] (2021).
- 25.S. A. Ekahana, Y. Soh, A. Tamai, D. Gosálbez-Martínez, M. Yao, A. Hunter, W. Fan, Y. Wang, J. Li, A. Kleibert, C. A. F. Vaz, J. Ma, Y. Xiong, O. V. Yazyev, F. Baumberger, M. Shi, G. Aeppli, Anomalous quasiparticles in the zone center electron pocket of the kagomé ferromagnet Fe3Sn2. arXiv:2206.13750 (2022).
- 26.A. Jaoui, I. das, G. di Battista, J. Díez-Mérida, X. Lu, K. Watanabe, T. Taniguchi, H. Ishizuka, L. Levitov, D. K. Efetov, Quantum critical behaviour in magic-angle twisted bilayer graphene. Nat. Phys. 18, 633–638 (2022). [Google Scholar]
- 27.W. Zhao, B. Shen, Z. Tao, Z. Han, K. Kang, K. Watanabe, T. Taniguchi, K. F. Mak, J. Shan, Gate-tunable heavy fermions in a moiré Kondo lattice. arXiv:2211.00263 [cond-mat.str-el] (2022). [DOI] [PubMed]
- 28.J. Huang, R. Yu, Z. Xu, J. X. Zhu, J. S. Oh, Q. Jiang, M. Wang, H. Wu, T. Chen, J. D. Denlinger, S. K. Mo, M. Hashimoto, M. Michiardi, T. M. Pedersen, S. Gorovikov, S. Zhdanovich, A. Damascelli, G. Gu, P. Dai, J. H. Chu, D. Lu, Q. Si, R. J. Birgeneau, M. Yi, Correlation-driven electronic reconstruction in FeTe1-xSex. Commun. Phys. 5, 29 (2022). [Google Scholar]
- 29.R. Yu, H. Hu, E. M. Nica, J.-X. Zhu, Q. Si, Orbital selectivity in electron correlations and superconducting pairing of iron-based superconductors. Front. Phys. 9, 578347 (2021). [Google Scholar]
- 30.H. Hu, L. Chen, J.-X. Zhu, R. Yu, Q. Si, Orbital-selective mott phase as a dehybridization fixed point. arXiv:2203.06140 [cond-mat.str-el] (2022).
- 31.V. Anisimov, I. Nekrasov, D. Kondakov, T. Rice, M. Sigrist, Orbital-selective Mott-insulator transition in Ca2-xSrxRuO4. Eur. Phys. J. B. 25, 191–201 (2002). [Google Scholar]
- 32.N. Marzari, A. A. Mostofi, J. R. Yates, I. Souza, D. Vanderbilt, Maximally localized wannier functions: Theory and applications. Rev. Mod. Phys. 84, 1419–1475 (2012). [Google Scholar]
- 33.S. Paschen, T. Lühmann, S. Wirth, P. Gegenwart, O. Trovarelli, C. Geibel, F. Steglich, P. Coleman, Q. Si, Hall-effect evolution across a heavy-fermion quantum critical point. Nature 432, 881–885 (2004). [DOI] [PubMed] [Google Scholar]
- 34.H. Shishido, R. Settai, H. Harima, Y. Ōnuki, A drastic change of the Fermi surface at a critical pressure in CeRhIn5: dHvA study under pressure. J. Phys. Soc. Jpn. 74, 1103–1106 (2005). [Google Scholar]
- 35.T. Park, F. Ronning, H. Q. Yuan, M. B. Salamon, R. Movshovich, J. L. Sarrao, J. D. Thompson, Hidden magnetism and quantum criticality in the heavy fermion superconductor CeRhIn5. Nature 440, 65–68 (2006). [DOI] [PubMed] [Google Scholar]
- 36.L. Prochaska, X. Li, D. C. MacFarland, A. M. Andrews, M. Bonta, E. F. Bianco, S. Yazdi, W. Schrenk, H. Detz, A. Limbeck, Q. Si, E. Ringe, G. Strasser, J. Kono, S. Paschen, Singular charge fluctuations at a magnetic quantum critical point. Science 367, 285–288 (2020). [DOI] [PubMed] [Google Scholar]
- 37.A. Ramires, J. L. Lado, Emulating heavy fermions in twisted trilayer graphene. Phys. Rev. Lett. 127, 026401 (2021). [DOI] [PubMed] [Google Scholar]
- 38.Z.-D. Song, B. A. Bernevig, Magic-angle twisted bilayer graphene as a topological heavy fermion problem. Phys. Rev. Lett. 129, 047601 (2022). [DOI] [PubMed] [Google Scholar]
- 39.A. Kumar, N. C. Hu, A. H. MacDonald, A. C. Potter, Gate-tunable heavy fermion quantum criticality in a moiré Kondo lattice. Phys. Rev. B 106, L041116 (2022). [Google Scholar]
- 40.D. Guerci, J. Wang, J. Zang, J. Cano, J. H. Pixley, A. Millis, Chiral Kondo lattice in doped MoTe2/WSe2 bilayers. arXiv:2207.06476 [cond-mat.str-el] (2022). [DOI] [PMC free article] [PubMed]
- 41.J. Cano, B. Bradlyn, Band representations and topological quantum chemistry. Annu. Rev. Condens. Matter Phys. 12, 225–246 (2021). [Google Scholar]
- 42.J. Kruthoff, J. de Boer, J. van Wezel, C. L. Kane, R.-J. Slager, Topological classification of crystalline insulators through band structure combinatorics. Phys. Rev. X 7, 041069 (2017). [Google Scholar]
- 43.I. Affleck, J. B. Marston, Large-n limit of the Heisenberg-Hubbard model: Implications for high-Tc superconductors. Phys. Rev. B 37, 3774–3777 (1988). [DOI] [PubMed] [Google Scholar]
- 44.E. Pivovarov, Q. Si, Transitions from small to large Fermi momenta in a one-dimensional Kondo lattice model. Phys. Rev. B 69, 115104 (2004). [Google Scholar]
- 45.M. Oshikawa, Topological approach to Luttinger’s theorem and the Fermi surface of a Kondo lattice. Phys. Rev. Lett. 84, 3370–3373 (2000). [DOI] [PubMed] [Google Scholar]
- 46.H. Hu, L. Chen, Q. Si, Extended dynamical mean field theory for correlated electron models. arXiv:2210.14197 [cond-mat.str-el] (2022).
- 47.H. C. Po, L. Zou, T. Senthil, A. Vishwanath, Faithful tight-binding models and fragile topology of magic-angle bilayer graphene. Phys. Rev. B 99, 195455 (2019). [Google Scholar]
- 48.J.-X. Zhu, D. R. Grempel, Q. Si, Continuous quantum phase transition in a Kondo lattice model. Phys. Rev. Lett. 91, 156404 (2003). [DOI] [PubMed] [Google Scholar]
- 49.W. Ding, R. Yu, Q. Si, E. Abrahams, Effective exchange interactions for bad metals and implications for iron-based superconductors. Phys. Rev. B 100, 235113 (2019). [Google Scholar]
- 50.O. N. Meetei, O. Erten, M. Randeria, N. Trivedi, P. Woodward, Theory of high Tc ferrimagnetism in a multiorbital mott insulator. Phys. Rev. Lett. 110, 087203 (2013). [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.