diff --git "a/-tE3T4oBgHgl3EQfrgrp/content/tmp_files/2301.04661v1.pdf.txt" "b/-tE3T4oBgHgl3EQfrgrp/content/tmp_files/2301.04661v1.pdf.txt" new file mode 100644--- /dev/null +++ "b/-tE3T4oBgHgl3EQfrgrp/content/tmp_files/2301.04661v1.pdf.txt" @@ -0,0 +1,4184 @@ +Kondo Resonance, Pomeranchuk Effect, and Heavy Fermi Liquid in Twisted Bilayer +Graphene - A Numerical Renormalization Group Study +Geng-Dong Zhou1 and Zhi-Da Song1, ∗ +1International Center for Quantum Materials, School of Physics, Peking University, Beijing 100871, China +(Dated: January 13, 2023) +Low energy electron Hamiltonian in the magic-angle twisted bilayer graphene can be equivalently +reformulated as a topological heavy fermion model [Phys. Rev. Lett. 129, 047601 (2022)]. It consists +of effective localized f-electrons at AA-stacking regions and itinerant Dirac c-electrons. In this work, +we applied systematic analytical and numerical renormalization group analyses to a single-impurity +version of this model. +We obtained a phase diagram consisting of a Fermi liquid phase in the +Kondo regime, a Fermi liquid phase in the frozen impurity regime, and various local moment phases +with different spin momenta. Remarkably, this single-impurity phase diagram explains a series of +experimental discoveries reported recently: (i) the zero-energy peak at fillings 1 ≲ |ν| < 2 observed +in STM at low temperatures (T < 1K) [Nature 588, 610 (2020), Nature Physics 17, 1375 (2021), +Nature 600, 240 (2021), Nature 589, 536 (2021)], (ii) the cascade of transitions observed in STM at +higher temperatures [Nature 582, 198 (2020), Nature Physics 17, 1375 (2021)], (iii) the Pomeranchuk +effect at ν ≈ ±1 observed in transport and compressibility measurements [Nature 592, 214 (2021), +Nature 592, 220 (2021)], which show that the Fermi liquid ground state develops local moments +upon heating, and (iv) various transport experiments showing resistance peaks but no gaps around +ν ≈ ±1. For the first time, we point out that all these phenomena result from a simple unified +mechanism - the Kondo effect. The Fermi liquid state at ν ≈ ±1 exhibiting the zero-energy peak +is stabilized by the Kondo screening with a Kondo temperature TK ≈ 1.5K. A higher temperature +will suppress the Kondo screening and favor a local moment phase that obeys Curie’s law and +contributes to an entropy of the order of Boltzmann’s constant (per moir´e cell). +We computed +the spectral densities, entropies, and spin susceptibilities at various fillings and temperatures, and +obtained results quantitatively comparable to experiments. We also predict the heavy Fermi liquid +as the ground state in a wide range of fractional fillings and conjecture that it is the parent state +for the observed unconventional superconductivity. +I. +INTRODUCTION +Since the first discovery of the superconductivity [1] +and correlated insulators [2] in magic-angle twisted bi- +layer graphene (MATBG) [3], MATBG has become a new +platform to study novel correlation effects in flat-band +systems and has attracted extensive attentions. Remark- +ably rich physics, including interplay between supercon- +ductivity [4–11] and strong correlation [4, 6–8, 12–19], in- +teraction driven Chern insulators [20–26], strange metal +behaviors [27–29], and the Pomeranchuk effect [30, 31], +etc., have been observed in MATBG. Several theoretical +understandings of the correlated states have also been +achieved recently. The strong correlation arises from the +two topological flat bands [32–37], each of which is four- +fold degenerate due to the spin and valley d.o.f. A large +U(4) symmetry group [38–42] emerges in the flat-band +limit, where the actual bandwidth is counted as negligi- +ble. Then the observed correlated states at integer fillings +ν = 0, ±1, ±2, ±3 can be understood as flavor polarized +states [38–40, 42–58] that spontaneously break the U(4) +symmetry. Here |ν| is the number of electrons (ν > 0) +or holes (ν < 0) per moir´e cell counted from the charge +neutrality point (CNP). The continuous U(4) degeneracy +leads to Goldstone mode fluctuations [59, 60] that may +∗ songzd@pku.edu.cn +destroy the long-range order due to the Mermin-Wagner +theorem. +Less theoretical understandings have been achieved for +the gapless states, which are observed at both fractional +and integer fillings. For example, at ν ≈ ±1, depend- +ing on the experimental setup, both gapped correlated +insulators [4, 6, 7] and gapless Fermi liquid states [6, 25– +27, 29–31] have been observed in transport experiments, +suggesting that they are competing ground states with +close energies. Interestingly, the observed gapless Fermi +liquid states around ν = ±1 usually exhibit resistivity +peaks [25–27, 29, 31] above a few kelvins, and the peaks +could become increasingly pronounced as temperature +rises. +Scanning tunneling microscope (STM) measure- +ments [11, 19, 21, 22] have constantly seen that, at low +temperatures of about a few hundred millikelvins, the +conduction (valence) band will be pinned at the Fermi +level and form a sharp zero-energy peak for the fillings +1 ≲ ν < 2 (−2 < ν ≲ −1). The sharp peak does not fit +the intuition of Stoner instability given that the interac- +tion is indeed strong. When the temperature increases to +a few or ten kelvins, these peaks develop into a cascade +of transitions like a quantum dot model [17, 19]. +In this work, based on the recently developed topolog- +ical heavy fermion (THF) model [61, 62], we find that +the zero-energy peak, as well as the gapless Fermi liq- +uid states at ν ≈ ±1, are results of the Kondo resonance +[63–80]. At a higher temperature exceeding the Kondo +arXiv:2301.04661v1 [cond-mat.str-el] 11 Jan 2023 + +2 +energy scale, the local moments (LMs) formed by local- +ized electrons give rise to the transition cascades, and +the resistance peaks around ±1. We have numerically +reproduced the temperature-dependent features of the +spectral density. Our theory is also fully consistent with +the Pomeranchuk effect observed around fillings ν = ±1 +[30, 31], which show that local moments appear upon +heating. We have calculated LM entropies as functions +of the temperature, filling, and an external field, and +obtained curves comparable to the experimentally mea- +sured data in Ref. [30, 31]. +Our theory shows that the observed gapless states at +the fillings 1 ≲ |ν| < 2 are the strongly correlated heavy +Fermi liquid state. Since this filling range overlaps with +the superconductivity [1, 4–11] and the strange metal +[27–29] around ν = −2+δ (for small δ), the heavy Fermi +liquid state could be the parent state for the unconven- +tional superconductivity, as it is in the heavy fermion +materials with 4f or 5f electrons. +This opens a new +perspective - with a solid theoretical and experimental +basis at the same time - to study the superconductivity +in MATBG. +This work is organized as the followings. For this work +to be self-contained, in Sec. II we will review the THF +model and its symmetry shortly. In Sec. III, based on a +poor man’s scaling analysis and experimental facts, we +argue that the Kondo screening effect is irrelevant at +CNP, and hence the ground state at CNP is the pre- +viously identified symmetry-broken correlated insulator +[38–40, 42–44]. Then we derive a simpler effective peri- +odic Anderson model describing active excitations upon +the correlated ground state. +We further simplify the +model to a single-impurity version. In Sec. IV, by ap- +plying poor man’s scaling and Wilson’s numerical renor- +malization group (NRG) method [81–83] to the single- +impurity problem, we obtain a phase diagram charac- +terized by strong coupling fixed points and various LM +fixed points. The strong coupling phase is divided into +a Kondo regime and a frozen impurity (FI) regime. We +also present a detailed analysis of the RG flows at these +fixed points. The gapless 1 ≲ |ν| < 2 states are found +to be in the Kondo regime. In Secs. V and VI we cal- +culate the spectral densities, spin susceptibilities, and +entropies as functions of the filling ν and the temper- +ature T. The spectral densities feature sharp Kondo res- +onances at low temperatures smaller than the Kondo en- +ergy scale. Whereas at higher temperatures, the Kondo +resonances are suppressed and the Hubbard bands be- +come clearer, which periodically cross the Fermi level as +ν changes from 0 to 4 as that of a quantum dot model, +matching the STM experiments. The spin susceptibili- +ties obey Curie’s law at high temperatures, suggesting +the existence of LM, and approach constants at lower +temperatures, suggesting the Fermi liquid behavior. The +Kondo temperature TK can be estimated as the turning +temperature of the two behaviors of the spin susceptibil- +ity. For 1 ≲ |ν| < 2, we find kBTK ranges from 0.129meV +to 0.675meV. (Here kB is Boltzmann’s constant.) +At +c +c +Energy (meV) +0 +40 +-20 +(b) +ΓM +KM +MM +2|M| +(a) +-40 +-60 +20 +60 +KM +ΓM +MM +c +-80 +80 +f +f +(c) +G +ΓM +KM +MM +c +L=0,0 +L=1,-1 +(d) +G +15 +10 +0 +∆(ω) (meV) +ω (meV) +5 +FIG. 1. +The THF model. +(a) Top: red spheres represent +the effective f-electrons located at AA-stacking regions of +MATBG, and blue spheres represent the itinerant c-electrons. +Bottom: the moir´e Brillouin zone. (b) Black bands are given +by the THF model. The red and blue bands are the decou- +pled f- and c-bands, respectively. +M is a parameter that +determines the bandwidth of the flat bands. We focus on the +M → 0 limit in this work. (c) Excitation spectrum of the ac- +tive c-electrons upon the symmetry-breaking parent state for +ν > 0, where M and µc are set to zero. (d) The hybridization +function ∆(ω) contributed by the c-bands in (c). A nonzero +M will not change the asymptotic behavior of ∆(ω) around +the gap. +|ν| = 1 and kBTK ≈ 1meV, the entropy contributed by +the LM is around 2 ln 2 · kB ≈ 1.39kB. In the presence of +a strong in-plane magnetic field, the entropy is quenched +to about ln 2·kB ≈ 0.69kB. These values will be appreci- +ated if one notices that the measured entropies at ν ≈ 1 +and T ≈ 10K in the absence and presence of a strong in- +plane field are about 1.2kB and 0.6kB, respectively [30]. +In the Sec. VII, we discuss the heavy Fermi liquid states +at 1 ≲ |ν| < 2 and propose experiments to confirm them. +We estimate the heavy fermion bands and their quasi- +particle weights using spectral information provided by +the NRG calculation. The possible effects of the RKKY +interactions are also discussed. +II. +THE THF MODEL +One theoretical challenge in studying correlation +physics in MATBG is the lack of a fully symmetric lat- +tice model for low energy physics, which is forbidden +by the band topology protected by a C2zT symmetry +[32–34] and an emergent particle-hole symmetry P [37] +- even though extended Hubbard models [84–88] can be +constructed at the sacrifice of either symmetry or local- +ity. The band topology was thought as fragile [32–34] +but was later shown to be a stable symmetry anomaly +jointly protected by C2zT and P [37]. The THF model +[61, 62] resolved this problem by ascribing the strong +correlation to effective f-orbitals at the AA-stacking re- +gions, which form a triangular lattice, and leaving the +remaining low energy states to continuous c-bands de- +scribed by a topological Dirac Hamiltonian (Fig. 1). The +THF model faithfully reproduces the symmetry, topol- +ogy, dispersion, and Coulomb interaction of the continu- +ous Bistritzer-MacDonald model [3]. Its free part is given + +3 +by +ˆH0 = −µ ˆ +N + +� +ηs +� +aa′ +� +|k|<Λc +H(c,η) +aa′ (k)c† +kaηsckaηs ++ +� +ηsαa +� +|k|<Λc +� +e− |k|2λ2 +2 +H(cf,η) +aα +(k)c† +kaηsfkαηs + h.c. +� +. +(1) +Here µ is the chemical potential, ˆN is the particle-number +operator, ckaηs is the fermion operator for the c-electron +of the momentum k, orbital a (= 1, 2, 3, 4), valley η (= +±), and spin s (=↑, ↓), fkαηs is the corresponding fermion +operator for the f-electron of the orbital α (=1,2). The +summation over k for the c-bands is in principle limited +within the cutoff Λc. But the theory is well-defined and +yields the same low energy physics if we take the Λc → ∞ +limit. H(c,η)(k) = v⋆(ησx⊗σ0kx−σy⊗σzky)+02×2⊕Mσx +is the Dirac Hamiltonian of the c-bands. When M ̸= 0, +c-bands have a quadratic band touching at the zero en- +ergy, whereas when M = 0, c-bands become linear. The +two-by-two block of H(cf,η) +aα +(k) for a = 1, 2 is given by +γσ0 + v′ +⋆(ησxkx + σyky), and the two-by-two block of +H(cf,η) +aα +(k) for a = 3, 4 vanishes. +The parameter λ in +the second line of Eq. (1) is the spread of the Wan- +nier functions of f-electrons, and it truncates the hy- +bridization at |k| ≫ λ−1. +In this work we adopt the +w0/w1 = 0.8 parameters of Ref. [61]: γ = −24.75meV, +v⋆ = −4.303eV · ˚A, v′ +⋆ = 1.623eV · ˚A, λ = 1.4131/kθ, +kθ = 1.703˚A−1 · 2 sin θm +2 with θm = 1.05◦ being the first +magic angle. (w0 and w1 are the interlayer couplings of +MATBG at the AA-stacking and AB-stacking regions, +respectively. Due to the corrugation effect [89–92], w0 +is usually smaller than w1. Ref. [85] estimates w0/w1 as +0.817.) The resulting band structure with a nonzero M +(3.697meV) is shown in Fig. 1(b). One can see that the +topological flat bands result from the hybridization be- +tween c- and f-bands. As explained in detail in Ref. [61] +and shown in Fig. 1(b), the parameter M determines the +bandwidth of the flat-bands. +In each valley, the Hamiltonian ˆH0 respects a magnetic +space group P6′2′2 [32] (#177.151 in the BNS setting +[93]), generated by C3z, C2x, C2zT, and translation sym- +metries. The two f-orbitals and the four c-bands have +effective angular momenta L = −η, η and L = −η, η, +0, 0, respectively. +As shown in detail in [61], the six- +by-six representations of C3z, C2x, C2zT symmetries on +these orbitals are eiη 2π +3 σz ⊕ eiη 2π +3 σz ⊕ σ0, I3×3 ⊗ σx, and +I3×3 ⊗ σxK, respectively, where σ0,x,y,z are Pauli matri- +ces and K the complex conjugation. One can verify that +the Hamiltonian matrices H(c,η) and H(cf,η) given in the +last paragraph respect these crystalline symmetries. The +interaction Hamiltonian given in the following paragraph +also respects these crystalline symmetries. +The interaction Hamiltonian is given by +ˆHI =U1 +2 +� +R +δnf +Rδnf +R + U2 +2 +� +⟨RR′⟩ +δnf +Rδnf +R′ + +1 +2NM +� +qaa′ +V (q)δnc +−qa′δnc +qa + +1 +NM +� +Rqa +Wae−iq·Rδnf +Rδnc +qa − +J +2NM +� +ηη′αα′ +ss′ +� +|k|,|k′|<Λc +R +� +(ηη′ + (−1)α+α′)e−i(k−k′)·R− λ2(k2+k′2) +2 +(f † +Rα′η′s′fRαηs − 1 +2δηη′δαα′δss′)(c† +k,α+2ηsck′,α′+2η′s′ − 1 +2δkk′δηη′δαα′δss′) +� +, +(2) +where NM is the number of moir´e cells, fRαηs is the real +space fermion operator for the f-electrons, R are the tri- +angular lattice shown in Fig. 1(a), ⟨RR′⟩ represents near- +est neighbor pairs (ordered), δnf +R = � +αηs(f † +RαηsfRαηs − +1 +2) is the density operator of f-electrons, +δnc +qa += +� +ηsk(c† +k+qaηsckaηs − 1 +2δq0) is the density operator for c- +electrons of the orbital a with k and k + q being limited +within the cutoff Λc. U1,2, V (q), Wa are the density- +density interaction between ff, cc, cf electrons, respec- +tively, and J is an exchange interaction between cf elec- +trons. We adopt the w0/w1 = 0.8 parameters of Ref. [61]: +U1 = 57.95meV, U2 = 2.329meV, W1 = W2 = 44.03meV, +W3 = W4 = 50.20meV, J = 16.38meV. We choose +V (q) as the double-gate-screened Coulomb interaction, +πξ2Uξ +Ω0 +tanh(ξ|q|/2) +ξ|q|/2 +, with ξ = 10nm being distance between +MATBG and the gates, Uξ = 24meV the Coulomb inter- +action at the distance ξ, and Ω0 ≈ 156nm2 the area of +the moir´e cell. +Hereafter, we mainly focus on the flat-band limit, i.e., +M = 0. In this limit, an exact U(4) symmetry of ˆH0+ ˆHI +between the spin, valley, and orbital flavors emerge, as +previously recognized in the projected Coulomb Hamil- +tonian of the continuous model [38–42]. We emphasize +that this U(4) symmetry is not related to the so-called +chiral limit [35, 94], i.e., w0 = 0, which leads a distinct +U(4) symmetry [41, 39]. The U(4) symmetry in the flat- +band limit has been shown as a good approximation for +realistic parameters such as w0/w1 = 0.8 [41, 43]. The +sixteen U(4) generators acting on fRαηs, ckaηs (a = 1, 2), +and ckaηs (a = 3, 4) are +Σf +µν = {σ0τ0ςν, σyτxςν, σyτyςν, σ0τzςν} , +(3) +Σc12 +µν = {σ0τ0ςν, σyτxςν, σyτyςν, σ0τzςν} , +(4) +and +Σc34 +µν = {σ0τ0ςν, −σyτxςν, −σyτyςν, σ0τzςν} , +(5) + +4 +respectively, where ςν (ν = 0, x, y, z) are Pauli matrices +acting in the spin subspace, τµ (µ = 0, x, y, z) are Pauli +matrices acting in the valley subspace, and σ0,x,y,z are +Pauli matrices acting in the orbital subspace. With the +help of U(4) symmetry, the J term in Eq. (2) can be writ- +ten as a ferromagnetic coupling between the U(4) LM of +f-electrons and the U(4) LM of c-electrons [61]. When +M ̸= 0, only the µ = 0, z U(4) generators commute with +the Hamiltonian, leading to a lower U(2)×U(2) symme- +try group. The rotation generated by µ = z, ν = 0 is +referred to as the valley-U(1) symmetry. +Consistent with previous results [38–40, 42–44], a +Hartree-Fock treatment of the THF model has predicted +the ground state at CNP as a U(4) LM state that re- +spects a U(2)×U(2) subgroup [61]. The LM forms a 20- +fold multiplet belonging to the [2, 2]4 representation [43] +of the U(4) group. These states can be approximately +written as +|Ψ0⟩ = e−iθµν ˆΣµν � +R +f † +R1+↑f † +R1+↓f † +R2+↑f † +R2+↓|FS⟩ , +(6) +where the |FS⟩ is the Fermi sea state with the half-filled +c-bands, ˆΣµν are the U(4) generator operators defined +by the matrices in Eqs. (3) to (5), and θµν are the ro- +tation parameters, and an implicit summation over re- +peated µ, ν indices is assumed. When θµν’s are zero, |Ψ0⟩ +is the valley-polarized state because all the occupied f- +electrons are in the η = + valley, and the U(2)×U(2) +subgroup is generated by Σ0,ν and Σz,ν (ν = 0, x, y, z). +For nonzero θµν’s, |Ψ0⟩ respects an equivalent U(2)×U(2) +subgroup. The Kramers inter-valley coherent states can +be obtained by choosing nonzero θx0 and θy0 satisfying +θ2 +x0 + θ2 +y0 = (π/4)2. When M ̸= 0, the Kramers inter- +valley coherent states are the ground states, while the +valley polarized states have higher energy (∼ 0.1meV) +[43, 61]. +III. +EFFECTIVE ANDERSON MODEL FOR +ν > 0 STATES +A. +Irrelevance of Kondo screening at CNP +Here we argue that the Kondo screening effect is ir- +relevant at CNP; hence, the U(4) LM state in Eq. (6) +is valid as an approximate ground state. We first exam- +ine the energy scale of a fully symmetric Kondo state +at CNP. Since the f-sites are almost decoupled from +each other, a reasonable approximation is to view each +f-site as a single Anderson impurity coupled to a bath +of c-electrons. If we only consider the on-site U1 inter- +action and the single-particle hybridization between f- +and c-electrons (H(cf,η)(k) in Eq. (1)), then it is almost +a standard Anderson model with eight flavors. The ef- +fect of c-bath is described by the hybridization function +∆(ω), defined as the imaginary part of the self-energy of +a free f-electron (in the absence of U1) coupled to the c- +bath, i.e., Im[Σ(f) +αηs,α′η′s′(ω)] = δα,α′δηη′δss′sgn(ω)∆(ω). +The identity matrix form of Im[Σαηs,α′η′s′(ω)] is guar- +anteed by the spin-SU(2) (δss′), the valley-U(1) and the +time-reversal (δηη′), and crystalline (δαα′) symmetries. +Low energy c-bands (k → 0) are coupled to the impu- +rity through the constant coupling γ in H(cf,η)(k). Then +∆(ω) would be proportional to the density of states ρ(ω) +of c-bands. In the flat-band limit (M = 0), c-bands have +a linear dispersion (Fig. 1(b)) and hence the density of +states, as well as the hybridization function, linear in en- +ergy, i.e., ρ(ω) ∼ |ω|, ∆(ω) ∼ |ω|. As a consequence, +low-lying states of the impurity will see vanishing bath +electrons when the energy scale is small enough. Both nu- +merical [95–97] and analytical [98] RG studies have shown +that Anderson impurity models with such a ∆(ω) ∼ |ω| +hybridization function do not have the strong coupling +fixed point that exhibits Kondo screening. Instead, the +only stable fixed point is the LM phase. +With a finite M, the c-bands given by H(c,η)(k) in +Eq. (1) have a quadratic touching at the zero energy, +i.e., ±(−M/2 + +� +M 2/4 + v2⋆k2), leading to a finite den- +sity of states at the zero energy and hence a finite ∆(0). +Nevertheless, as explained in the following, the Kondo +energy scale resulting from realistic parameters is neg- +ligible. +In Appendix B 2 we derived an analytical ex- +pression of ∆(ω) for the symmetric state at CNP. For +|ω| > M, ∆(ω) is almost linear in |ω|, i.e., ∆(ω) ≈ b·|ω|. +For |ω| < M, ∆(ω) is a constant plus a linear term: +∆(ω) ≈ ∆(0)(1 + |ω|/M). Using the parameters given +in Sec. II and M = 3.697meV, we have b ≈ 0.129, +∆0 ≈ 0.239meV. A rough estimation of the Kondo en- +ergy scale can be made by replacing the ω-dependent +∆(ω) with the constant ∆(0) at ω = 0. Then naively +applying the large-N formula at second order [99], i.e., +kBTK ≈ De− +πU1 +4N ∆(0) with N = 8 being the number of +flavors and D = U1/2 the energy scale up to which the +perturbation theory applies, leads to an extremely small +kBTK ≈ U1 · 2 × 10−11. A better estimation is given by +a poor man’s scaling that considers the ω-dependence of +∆(ω). Readers may refer to Appendix B 2 for more de- +tails. Here we only present the main results. There are +two stages in the RG process: (i) a stage with energy +scale from U1/2 - scale up to which the perturbation the- +ory applies - to M. +(ii) a stage with an energy scale +below M. RG in the first stage effectively enhances ∆(0) +to g1∆(0) with g1 ≈ 2.34. Then, RG in the second stage +gives +kBTK ≈Meye− +πU1 +4N g1∆(0) +≈3.8 × 10−4meV +(2M = 7.39meV). +(7) +where y ≈ 1 is a factor contributed by the ω-dependence +of ∆(ω) in the second stage. This value is still much lower +than the energy gain of the symmetry-broken correlated +state [43, 59]. The bandwidth of the Goldstone modes +at CNP from ΓM to MM is about 8meV. (See Fig. 2 of +Ref. [59]). +If we understand this spectrum as a tight- +binding band of the Holstein–Primakoff bosons on the +f-sites, which form a triangular lattice, then the nearest + +5 +neighbor hopping is about 8meV/8=1meV. This hopping +indicates an RKKY interaction much larger than kBTK +evaluated in Eq. (7). +The actual TK can even be much smaller than the value +in Eq. (7). First, as ∆(0) → 0 when M → 0, TK decays +exponentially when M decreases. For example, a band- +width 2M = 5meV corresponds to +kBTK ≈ 5.1 × 10−6meV +(2M = 5meV) . +(8) +Second, because we only considered the U1 interaction +and the cf hybridization that gives all flavors of f- +electrons the same ∆(ω) (guaranteed by symmetries), the +single-impurity model has a U(8) symmetry. This U(8) +symmetry must be broken when other interaction terms, +e.g., J in Eq. (2), are taken into account, leading to a +multiplet splitting. +When the energy scale in the RG +is smaller than the multiplet splitting, the degeneracy +factor N should be reduced accordingly, and TK will be +further suppressed. (One can see section 17.2 of Ref. [99] +and Appendix B 3 for examples of how multiplet splitting +suppresses TK.) +The U(4) LM state at CNP is also supported by var- +ious experiments. In contrast to the Kondo resonance, +STM measurements have shown strong suppression of the +density of states at the zero energy at CNP [11, 12, 14– +17, 19, 21, 22]. Some transport experiments [4, 6, 7, 27] +also exhibits a gap behavior at CNP. Although there are +also transport experiments showing semimetal behavior, +the gaplessness can be explained if there are fluctuations +of the local moments from site to site, which is possi- +ble due to the Goldstone mode fluctuations [59, 60] and +possible inhomogeneity of the sample. +B. +Periodic Anderson model for ν > 0 states +We aim for an effective model describing the active ex- +citations upon the ground state |Ψ0⟩ (Eq. (6)) at CNP. +Let us first assume the valley-polarized state, where θµν’s +in Eq. (6) are all zero such that all the occupied f- +electrons are in the η = + valley. +As detailed in the +supplementary material of Ref. [61] and in Ref. [59], the +lowest particle and hole excitations are in the η = − +and η = + valleys, respectively. Thus, for ν > 0 states, +only particle excitations in the η = − valley will be in- +volved, and the electrons in the η = + valley can be +viewed as a static background. +The effective Hamil- +tonian can be obtained by replacing operators in the +η = + valley by their expectation values on |Ψ0⟩, which +are ⟨f † +Rα+sfR′α′+s′⟩ = δRR′δαα′δss′, ⟨c† +ka+sck′a′+s′⟩ ≈ +1 +2δkk′δaa′δss′, ⟨c† +ka+sfRα+s′⟩ = 0. Substituting these ex- +pectation values into ˆH0 + ˆHI, we obtain the effective +free Hamiltonian +ˆHeff +0 += +� +|k|<Λc +aa′s +(H(c) +aa′(k) − µδaa′)c† +kascka′s − µ +� +Rαs +nf +Rαs ++ +� +|k|<Λc +aα +� +e− 1 +2 λ2k2H(cf) +aα (k)c† +kasfkαs + h.c. +� +, +(9) +where nRαs = f † +RαsfRαs is the density operator of f- +electrons. Here we have dropped the valley index η as +they are limited to η = −. The H(c)(k) and H(cf)(k) +matrices are given by the H(c,−)(k) and H(cf,−)(k) ma- +trices defined after Eq. (1). We also obtain the effective +interaction Hamiltonian +ˆHeff +I += U1 +2 +� +R +nf +Rnf +R + U2 +2 +� +⟨RR′⟩ +nf +Rnf +R′ ++ +1 +2NM +� +qaa′ +V (q)δnc +−q,a′δnc +q,a + +1 +NM +� +Rqa +Wae−iq·Rnf +Rδnc +qa +− +J +NM +� +Rss′ +� +α +� +|k|,|k′|<Λc +e−i(k−k′)·R− λ2(k2+k′2) +2 +× (f † +Rαs′fRαs − 1 +2δss′)(c† +k,α+2,sck′,α+2,s′ − 1 +2δss′) , +(10) +where nf +R = � +αs nf +Rαs, δnc +qa = � +sk(c† +k+qasckas − 1 +2δq0) +with |k| and |k + q| being limited within the cutoff Λc. +The δnf +R operator in Eq. (2), which represents the density +deviation from the charge background at CNP, is now +replaced by nf +R, the total density in the η = − valley, +because the charge background is compensated by the +occupied η = + electrons. The J term in Eq. (2) is also +simplified: As active excitations are limited to η = −, +the factor ηη′ + (−1)α+α′ becomes 2δα,α′. +In the flat-band limit (M += 0), +ˆHeff +0 ++ ˆHeff +I +ap- +plies to arbitrary U(4) partners of the valley-polarized +state, including the so-called Krammers intervalley co- +herent state. To be specific, for a generic |Ψ0⟩ given in +Eq. (6), we can always define rotated operators fRαs = +UfRα−sU †, ckas = Ucka−sU †, where U = e−iθµν ˆΣµν is +the U(4) rotation defining |Ψ0⟩, such that the effective +Hamiltonian on the rotated basis is same as Eqs. (9) +and (10). +The effective Hamiltonian ˆHeff +0 ++ ˆHeff +I +respects all the +crystalline symmetries discussed in Sec. II. The effective +angular momenta of the active two f-orbitals and four +c-bands are L = 1, −1 and L = 1, −1, 0, 0, respectively. +And, the six-by-six representations of C3z, C2x, C2zT +on these orbitals are e−i 2π +3 σz ⊕ e−i 2π +3 σz ⊕ σ0, 13×3 ⊗ σx, +and 13×3 ⊗ σxK, respectively, with K being the complex +conjugation. +In the flat-band limit (M = 0), as |Ψ0⟩ +respects a U(2)×U(2) subgroup of the U(4) group, e.g., +independent spin-charge rotations in the two valleys for +the valley polarized |Ψ0⟩, one may expect a U(2)×U(2) +symmetry of ˆHeff +0 ++ ˆHeff +I . +However, since the effective +Hamiltonian only involves half of the d.o.f., e.g., the ac- +tive η = − valley for the valley polarized |Ψ0⟩, only a +single U(2) factor is meaningful for ˆHeff +0 ++ ˆHeff +I . There- +fore, hereafter we will say that ˆHeff +0 + ˆHeff +I +respects a U(2) +symmetry group. +When M ̸= 0, the U(4) symmetry is broken, and the +effective Hamiltonian will have an additional term. In +Appendix A we show that to the order of M 2, the cor- + +6 +ˆH0 + ˆHI +ˆHeff +0 ++ ˆHeff +I +ˆHSI +M = 0 +U(4) +U(2) +U(2)×U(2) +M ̸= 0 U(2)×U(2) +U(2) +U(2)×U(2) +TABLE I. Continuous symmetries of the Hamiltonians. ˆH0 + +ˆHI is the original THF model. For ν > 0 (ν < 0), ˆHeff +0 + ˆHeff +I +is the effective periodic Anderson model for the active particle +(hole) excitations upon the symmetry broken state at CNP. +ˆHSI is a single-impurity version of ˆHeff +0 ++ ˆHeff +I . +rection is simply an energy shift +M 2 +J +� +|k|<Λc +� +a=3,4 +� +s +c† +kasckas + O(M 4) . +(11) +Thus, the M-term breaks neither the crystalline symme- +try nor the U(2) symmetry, and it will play a minor role +in the effective theory. To avoid confusion, in Table I we +summarize the continuous symmetries of different Hamil- +tonians with M = 0 or M ̸= 0 discussed in this work. +The effective model for ν < 0 states, which only in- +volve hole excitations, can be obtained by applying the +particle-hole operation Pc [41, 61] to ˆHeff +0 ++ ˆHeff +I . +C. +Single impurity model for ν > 0 states +Hereafter we mainly focus on a single-impurity version +of ˆHeff +0 ++ ˆHeff +I , where only the correlation at the R = 0 +f-site is considered. The interactions involving other f- +sites will be treated at the mean-field level. The STM +spectra [11, 19, 21, 22] that show the zero-energy peaks +also clearly show a continuity between the gapped CNP +state and the gapless states at 1 ≲ |ν| < 2, implying +that they have the same symmetries. Therefore, in this +work, we assume no additional symmetry breaking. For +the completeness of the discussion, we also extend our +symmetric assumption to |ν| ≥ 2 states. One should be +aware that additional symmetry breaking may happen +at low temperatures in |ν| ≥ 2 states due to the effec- +tive RKKY interactions neglected in this work. Thus the +symmetric assumption is invalid for |ν| ≥ 2 states at low +temperatures. Nevertheless, the |ν| ≥ 2 states may re- +cover the symmetries at higher temperatures, where our +symmetric theory applies. +At a given filling ν, the symmetric mean field is char- +acterized by only a few parameters: νf = ⟨nf +R⟩, νc,a = +1 +NM ⟨δnc +q=0,a⟩, where νc,1 = νc,2, νc,3 = νc,4 due to the +crystalline symmetries. The actual values of νf and νc,a +can be determined self-consistently. The considered cor- +related site at R = 0 is described by the Hamiltonian +ˆHf = −µf +� +αs +nf +αs + U1 +� +(αs)<(α′s′) +nf +αsnf +α′s′ , +(12) +where the lattice index R is omitted for simplicity, µf = +−6νfU2−� +a νc,aWa− 1 +2U1+Jνc,3+µ is an effective chem- +ical potential for the f-site, and µ is the global chemical +potential determined by the total filling ν. The U2, Wa, J +terms in µf are contributed by the Hartree mean fields of +the U2, Wa, and J interactions in Eq. (10), respectively. +The U1 term in µf is from the diagonal U1 interactions +in Eq. (10), i.e., +1 +2U1 +� +αs nf +αsnf +αs. The U1 interaction +in ˆHf only contains the off-diagonal U1 interactions of +Eq. (10). +The effective Hamiltonian of c-electrons is given by +ˆHc = +� +|k|<Λc +� +aa′s +(H(c) +aa′(k) + ∆H(c) +aa′ − µcδaa′)c† +kascka′s , (13) +where H(c)(k) = −v⋆(σx ⊗ σ0kx + σy ⊗ σzky) is the free +Dirac Hamiltonian, ∆H(c) +aa′ = G·δaa′(δa3+δa4) is a mean- +field term that split the a = 1, 2 and a = 3, 4 c-electrons, +µc = −νfW1 −νc V +Ω0 +µ is an effective chemical potential +of c-electrons. The W1 and V terms in µc are contributed +by the W and V interactions in Eq. (10), respectively. +The mean field term G arises from two interaction terms: +(i) A mean field treatment of the J interaction in Eq. (10) +yields an energy shift J( 1 +2 − 1 +4νf) for a = 3, 4 c-electrons. +(ii) The Hartree mean fields of the Wa interactions in +Eq. (10) are νfW1 and νfW3 for a = 1, 2 and a = 3, 4 +c-electrons, respectively. As we have absorbed νfW1 to +µc, a = 3, 4 c-electrons have an extra energy shift (W3 − +W1)νf. Combining the two effects, the parameter G is +given by G = J +2 +(W3−W1− J +4 )νf. Since the interaction +V (q) of c-electrons is completely treated at the mean- +field level, ˆHc is an effective free-fermion system. +As +detailed in Appendix A, the band structure of Eq. (13) is +given by G/2± +� +G2/4 + v2⋆k2. In Fig. 1(c) we plot the c- +bands with µc = 0 and G = 10meV. It has a gap opened +by G, where, according to the symmetry representations +given in Sec. III B, the lowest conduction and highest +valence band states have angular momenta 0, 0 and 1, −1, +respectively. +The f-site is coupled to c-electrons via two terms. One +is the hybridization +ˆHhyb = +1 +√NM +� +|k|<Λc +� +aαs +� +e− λ2k2 +2 +H(cf) +aα (k)c† +kasfαs + h.c. +� +. +(14) +The other coupling term is the remaining ferromagnetic +exchange interaction +ˆHJ = − +J +NM +� +|k|,|k′|<Λc +� +αss′ +e− 1 +2 λ2(k2+k′2)(f † +αsfαs′ − 1 +4δss′νf) +× (c† +k′α+2s′ckα+2s − 1 +2δk,k′δss′νc,α+2) . +(15) +By “remaining” we mean that the mean field back- +grounds +1 +4νf and +1 +2νc,a are deducted in +ˆHJ. +As ex- +plained below, ˆHJ leads to an effective Hund’s coupling +that changes the symmetry of the single-impurity model. +There are also remaining density-density interactions be- +tween c- and f-electrons, i.e., Wa(nf − νf)(nc +q,a − νc,a). +However, these remaining density-density interactions do +not change the essence of the single-impurity problem as +ˆHJ does. + +7 +Thanks to the C3z symmetry, +ˆHhyb and ˆHJ couple +the f-electrons to two independent baths belonging to +different angular momenta. This allows us to treat the +two terms separately. In a polar coordinate, ˆHhyb only +couples f-electrons to +�ckαs = 1 +A +� +a +ˆ +dφ · H(cf) +αa (k)ckas +(α = 1, 2) , +(16) +where k = k(cos φ, sin φ), and A is a normalization fac- +tor. Explicitly, there are �ck1s ∼ +´ +dφ·(γck1s−v′ +⋆keiφck2s) +and �ck2s ∼ +´ +dφ · (γck2s − v′ +⋆ke−iφck1s). Under the C3z +operation (defined in Sec. III B), �ck1s and �ck2s have effec- +tive angular momenta 1, -1, respectively. On the other +hand, ˆHJ only couples f-electrons to +�ckas = 1 +A′ +ˆ +dφ · ckas +(a = 3, 4) . +(17) +Both ckas (a = 3, 4) have the effective angular momen- +tum 0 under C3z. +Because �ckas (a = 1, 2) and �ckas +(a = 3, 4) form different representations of C3z, they do +not couple to each other, hence the ˆHhyb-bath and the +ˆHJ-bath are indeed independent. +As a ferromagnetic coupling always flows to zero and +becomes irrelevant in low energy physics, we can inte- +grate out the ˆHJ-bath in a single attempt. This leads to +an effective Hund’s coupling (Appendix A) +ˆHH = JH +� +α +nα↑nα↓ , +(18) +where JH, estimated as 0.3meV, is the additional energy +that two electrons will acquire if they occupy the same +orbital. A nonzero M does not change the form of ˆHH. +Integrating out the ˆHhyb-bath leads to a self-energy +correction Σ(f) +αs,α′s′(ω) to the f-electrons, the imaginary +part of which defines the hybridization function ∆(ω), +i.e., Im[Σ(f) +αs,α′s′(ω)] = δαα′δss′sgn(ω)∆(ω). The identity +matrix structure of the self-energy is guaranteed by SU(2) +spin rotation symmetry and crystalline symmetries. In +Appendix A we derived an analytical expression of ∆(ω). +As shown in Fig. 1(d), where ∆(ω) for µc = 0 and G = +10meV is plotted, ∆(ω) has an abnormal ω-dependence +compared to those in usual metals. First, ∆(ω) = 0 for +ω in the gap of c-bands (Fig. 1(c)). Second, because f- +electrons (L = 1, −1) have different angular momenta +as the lowest conduction c-bands (L = 0), hybridization +between them vanishes as k → 0. +As a result, ∆(ω) +linearly approaches zero at the conduction band edge. +Third, as f-electrons have the same angular momenta as +the highest valence c-bands, ∆(ω) approaches a constant +at the valence band edge. In summary, around the gap +∆(ω) has the asymptotic behaviors +∆(ω) ∼ +� +� +� +� +� +|ω + µc − G|, +ω + µc → G + 0+ +0, +0 ≤ ω + µc ≤ G +const., +ω + µc → −0+ +. +(19) +A nonzero M does not change the asymptotic behav- +iors as these behaviors are guaranteed by the C3z sym- +metry that is also respected by M. Due to the damp- +ing factor e− 1 +2 λ2k2 in ˆHhyb, c-electrons with momenta +|k| ≫ 1/λ will not contribute to ∆(ω). Thus, ∆(ω) de- +cays exponentially when ω exceeds v⋆/λ ∼ 95meV. In +the rest of this work, we will restrict the hybridization to +|ω| < D = 100meV. +Baths giving rise to the same ∆(ω) are physically +equivalent. Therefore, we can choose a bath that is as +simple as possible. We introduce the following effective +single-impurity Hamiltonian +ˆHSI = ˆHf + ˆHH + +� +αs +ˆ D +−D +dϵ · ϵ · d† +αs(ϵ)dαs(ϵ) ++ +� +αs +ˆ D +−D +dϵ · +� +∆(ϵ) +π +(f † +αsdαs(ϵ) + h.c.) , +(20) +where ˆHf and ˆHH are given by Eqs. (12) and (18), re- +spectively, and dαs(ϵ) are the auxiliary bath fermions in- +troduced to reproduce the hybridization function. dαs(ϵ) +satisfy {dα′s′(ϵ′), d† +αs(ϵ)} = δα′αδs′sδ(ϵ′ − ϵ). ˆHSI is com- +pletely defined by the following parameters: µf the chem- +ical potential of f-electrons, U1 the Coulomb repulsion, +JH the Hund’s coupling, and ∆(ω) the hybridization +function, which further depends on µc the chemical po- +tential of c-electrons and G the gap of c-bands (Fig. 1(c)). +As explained at the beginning of this subsection, the ac- +tual values of µf, µc, and G depend on the occupations +νf, νc,a, which are further determined by self-consistent +calculations at given total fillings ν. We plot µf, µc, and +G as functions of ν in Fig. 2(a). For ν changing from 0 +to 4, µc changes from 0 to 64.70meV, µf changes from +-28.98meV to 124.13meV, and G changes from 8.19meV +to 14.21meV. We can now regard µc, µf, G as given pa- +rameters that define the single-impurity problem. +It is worth mentioning that Eq. (20) has an emergent +U(2)×U(2) symmetry - independent spin-charge rota- +tions in the α = 1, 2 orbitals - that is higher than the +U(2) symmetry of Eqs. (9) and (10). It is not surpris- +ing that a single-impurity model has a higher symmetry +than its lattice version. +For example, if J = 0, there +would be no Hund’s coupling JH, and ˆHSI would have +a U(4) symmetry, as expected in a four-flavor Anderson +impurity model without multiplet splitting. +IV. +PHASE DIAGRAM OF THE +SINGLE-IMPURITY MODEL +A. +Poor man’s scaling +Before going to numerical calculations, we first apply a +poor man’s scaling to the single impurity model Eq. (20). +The scaling theory helps us understand several features +of the data obtained by NRG, as will be discussed in +the following subsections. +And, it predicts the Kondo + +8 +temperatures in the same order as those predicted by +NRG. +We assume that the ground state of the isolated impu- +rity has nf (integer) occupied f-electrons. One should +not confuse nf with νf - the expectation value of f- +occupation after the impurity is coupled to the bath. The +chemical potential µf must be in the range (nf − 1)U1 < +µf < nfU1. +We apply a Schrieffer-Wolff transforma- +tion to Eq. (20) to obtain an effective Coqblin–Schrieff +model where the local Hilbert space of f-electrons is re- +stricted to nf particles. The transformation involves vir- +tual particle and hole excitations, the energies of which +are ∆E+ = nfU1 − µf and ∆E− = µf − (nf − 1)U1, re- +spectively. (We have ignored the JH term in ∆E± as it +is small compared to U1.) Adding the two contributions, +we have +ˆH = ˆHH + +� +αs +ˆ D +−D +dϵϵd† +αs(ϵ)dαs(ϵ) + 4g +πU1 +� +αα′ss′ +ˆ D +−D +dϵdϵ′ +� +× +� +∆(ϵ)∆(ϵ′)(f † +αsfα′s′ − xδαα′δss′)d† +α′s′(ϵ′)dαs(ϵ) +� +. +(21) +The parameters g, x are given by +g = U1 +4 +� +1 +∆E+ ++ +1 +∆E− +� +, +x = ∆E− +U1 +. +(22) +g is a dimensionless parameter characterizing the anti- +ferromagnetic coupling strength between the impurity +and the bath. x appears as a “charge background” of +the f-electrons. For µf = (νf − 1 +2)U1, there is g = 1, +x = 1 +2. For a generic (nf − 1)U1 < µf < nfU1, g ≥ 1 +and 0 < x < 1. Flow equations of g, x are derived in Ap- +pendix B 3, where the divergence of g indicates the strong +coupling fixed point that exhibits the Kondo screening. +We notice that x always flows to nf/4, i.e., the occupa- +tion fraction of f-electrons. +One should be careful about the cutoff D in Eq. (21) +First, it must be smaller than ∆E+ and ∆E− for the +Schrieffer-Wollf transformation to be valid. Second, for +analytical conveniences, we only keep the positive branch +of ∆(ω) (Eq. (19)) at ω > G − µc because when ν > +0 the negative branch is far away from the Fermi level +(Fig. 1(d)). Hence, we also require D < µc − G. We can +choose D = min(µc − G, ∆E+, ∆E−). +We first consider the case nf = 1. The flow equation +of g(t) as the cutoff is reduced to De−t is given by +dg +dt = 4∆(0) +πU +Ng2 + O(e−t) , +(23) +and the initial condition g(0) is given by Eq. (22). Here +N = 4 is the number of flavors. The local Hilbert space +for nf = 1 is four-fold. The Hund’s coupling JH does not +split the four-fold degeneracy and hence does not appear +in the flow equation. The O(e−t) terms are irrelevant +at small energy scales but they may affect the coupling +constant at an early stage of the RG process. Using a +linear approximation of ∆(ω), i.e., ∆(ω) ≈ ∆(0)(1 + +ω/(µc − G)), we obtain (Appendix B 3) +kBT (1) +K += Dey1 · e− +πU1 +4N ∆(0)g(0) +(24) +where y1 ≈ −0.75 +D +µc−G < 0 is factor contributed by the +irrelevant O(e−t) terms and will slightly suppress the +Kondo energy scale. The suppression factor ey1 appears +because, for the nf = 1 states, the virtual processes that +contribute to the RG equation involve more hole exci- +tations than particle excitations in the bath, such that +the smaller ∆(ω < 0) contributes more than the larger +∆(ω > 0). As a result, the resulting kBTK is smaller +than the standard case (y1 = 0) where the coupling is a +constant, i.e., ∆(ω) = ∆(0). +We then study the RG equations at nf = 2. Unlike +the nf = 1 case, Hund’s coupling JH splits the six- +dimensional local Hilbert space. According to Eq. (18), +the four states with (nf +1↑, nf +1↓; nf +2↑, nf +2↓) =(10;10), (10;01), +(01;10), (01;01) do not feel JH and have the energy +−2µf + U1, whereas the two states (11;00), (00;11) have +the energy −2µf + U1 + JH. We divide the RG process +into two stages: (i) a stage with an energy scale from D +to JH, (ii) a stage with an energy scale below JH. In the +first stage, JH plays a minor role; hence, a flow equation +similar to Eq. (23) applies. If g diverges before the energy +scale reaches JH, then the Kondo scale is given by +kBT (2)′ +K += D · e− +πU1 +4N ∆(0)g(0) +(25) +there is no y-factor as in Eq. (24) because, for the nf = 2 +states, the virtual processes that contribute to the RG +equations evolve equal holes and particles in the bath. +Otherwise, g will be renormalized to +g1 = +g(0) +1 − g(0) 4∆(0) +πU1 N ln D +JH +(26) +at the energy scale of JH. As detailed in Appendix B, RG +in the second stage is similar to Eq. (23) except that, due +to the multiplet splitting, the factor N = 4 is replaced +by 2. This leads to the Kondo energy scale +kBT (2)′′ +K += JH · e +− +πU1 +8∆(0)g1 = D +� D +JH +� N +2 −1 +e +− +πU1 +8∆(0)g(0) . (27) +Considering the g may diverge in either the first or the +second stage, the physical Kondo energy scale can be +written as +kBT (2) +K += +� +kBT (2)′ +K , +kBT (2)′ +K +> JH +kBT (2)′′ +K +, +otherwise +. +(28) +The theory at nf = 3 is almost the same as the theory +at nf = 1 except that the four single-particle states are +now replaced by the four single-hole states. Thus, the +local Hilbert space is also four-dimensional and will not +be split by Hund’s coupling JH. The Kondo energy scale +is given by +kBT (3) +K += Dey3 · e− +πU1 +4N ∆(0)g(0) +(29) +where y3 ≈ 0.75 +D +µc−G > 0 is a factor contributed by +the irrelevant O(e−t) terms and will slightly enhance the + +9 +ν +kBT (P ) +K +(meV) δK(meV) kBT (χ) +K +(meV) +0.75 +0.280 +0.165 +0.151 +1.00 +0.250 +0.158 +0.129 +1.25 +0.366 +0.305 +0.190 +1.50 +0.780 +0.558 +0.344 +1.75 +1.620 +1.674 +0.482 +2.00 +- +1.926 +0.675 +2.25 +2.73 +1.463 +0.670 +2.50 +3.22 +1.934 +0.736 +2.75 +4.18 +2.284 +0.872 +3.00 +4.12 +2.305 +1.024 +3.25 +4.05 +3.400 +1.271 +3.50 +2.86 +2.663 +1.391 +TABLE II. Kondo energy scales at various fillings ν. T (P ) +K , +δK, and T (χ) +K +are Kondo energy scales estimated by the poor +man’s scaling, the NRG spectral density, and the NRG spin +susceptibility, respectively. δK is defined as the half width at +half maximum of the spectral peak. T (χ) +K +is defined as the +turning temperature where χ(T) transitions from the Curie- +Weiss behavior to the Fermi liquid behavior. There is no data +of T (P ) +K +at ν = 2 because ν = 2 is close a mixed valence case +where µf ≈ U1 and the Schrieffer-Wolff transformation does +not apply. +Kondo temperature. The enhancement arises from the +fact that, in contrast to the case of nf = 1, for nf = 3 +states the virtual processes contributing to the RG equa- +tion evolve more particle excitations than hole excita- +tions in the bath, and particles have larger couplings than +holes. +In Table II we tabulate the Kondo energy scales ob- +tained by the poor man’s scaling at various fillings us- +ing the filling-dependent µc, µf, G parameters given in +Fig. 2(a), and compare them to the values obtained by +the NRG method. +B. +The NRG method +In the NRG method [81–83], the bath is alternatively +realized by a Wilson chain constructed in the way de- +scribed below. +First, the energy window [−D, D] is +discretized on a logarithmic scale, i.e., ωn = D/Λn−1 +(n = 1, 2 · · · ), where Λ > 1 is a scaling factor (cho- +sen as 3 in this work). +Then for each energy shell +ωn ≤ |ω| < ωn+1 two auxiliary bath electrons corre- +sponding to the positive and negative part of it are in- +troduced to reproduce the corresponding ∆(ω). These +auxiliary electrons are further recombined into a Wilson +chain, dnαs, such that (i) only the first site d1αs couples +to the impurity, (ii) the chain is a tight-binding model +with only on-site and nearest neighbor hopping terms. +The Hamiltonian Eq. (20) is now mapped to an impurity +plus a Wilson chain +ˆHN = ˆHf + ˆHH + +� +αs +t0(f † +αsd1αs + h.c.) ++ +N +� +n=1 +� +αs +ϵnd† +nαsdnαs + +N−1 +� +n=1 +� +αs +(tnd† +n+1αsdnαs + h.c.) , +(30) +where N is a large number. The parameters ϵn and tn +can be computed from ∆(ω) using a standard iterative +algorithm [83]. For n → ∞, ϵn ∼ Λ−n and tn ∼ Λ− 1 +2 n. +Thus, the right-most sites represent the lowest-lying bath +states. +One can define the Nth scaled Hamiltonian as �HN = +(Λ) +1 +2 N−1 ˆHN. They can be constructed iteratively +� +HN+1 =Λ +1 +2 � +HN + Λ +1 +2 (N−1) � +αs +� +ϵN+1d† +N+1,αsdN+1,αs ++ tNd† +N+1,αsdN,αs + tNd† +N,αsdN+1,αs +� +. +(31) +The Hilbert space dimension increases exponentially in +this iterative process. The NRG algorithm truncates the +Hilbert space by keeping a fixed number (chosen to be +∼1200 in this work) of the lowest-lying states at each +step. In order to keep the symmetry in the truncated +Hilbert space, in practice we keep all the states up to a +gap above the 1200th state. +C. +Phase diagram and fixed points +Two successive transformations (Eq. (31)) that take +�HN to �HN+2 can be thought as a renormalization group +operation [81, 82]. The system is said to achieve a fixed +point when �HN and �HN+2 have the same low-lying many- +body spectrum. We can obtain a zero temperature phase +diagram in the parameter space of µc, µf, G by analyz- +ing the fixed points. For the completeness of discussions, +here we let µf take value in [−0.5U1, 3.5U1] such that +the corresponding impurity occupation (in the decoupled +limit) nf = µf/U1 +1/2 takes value in [0, 4]. In Fig. 2(b) +and (c) we show the obtained phase diagrams in the pa- +rameter space of µc, µf for G = 8meV and G = 14meV, +respectively. 8meV and 14meV are chosen to be close to +the minimal (8.19meV) and maximal (14.21meV) values +of G (Fig. 2(a)), respectively. +Due to the U(2)×U(2) symmetry, all the many-body +levels can be classified into symmetry sectors labeled by +the good quantum numbers (Q1, Q2; S1, S2), where Qi +and Si are the charge and spin of the ith U(2) sym- +metry, respectively. +Here we take the convention that +Q1 + Q2 = 0 corresponds to a total occupation 2N + 2 +(2N) for odd (even) N. A fixed point is characterized by +low energy many-body levels and the associated quantum +numbers. In the whole phase diagram, we find two dis- +tinct types of stable fixed points: (i) the strong coupling +fixed point exhibiting a Fermi liquid behavior and (ii) the +LM fixed points exhibiting nonzero spin momenta. At a +strong coupling fixed point, as exampled in Fig. 2(d), +(e), for either even or odd N, the ground state is a sin- +glet and has (Q1, Q2; S1, S2) = (2k, 2k; 0, 0) for some or- +der one integer k, which in most cases equals to 0. The + +10 +(f) +(g) +(d) +(h) +EN (meV) +EN (meV) +(c) +(a) +(e) +200 +μf/U +μc (meV) +0 +1 +1 +2 +3 +60 +20 +0 +40 +LM1 +LM2 +LM3 +Kondo +FI +FI +δK (meV) +G=8meV +(b) +μf/U +60 +20 +0 +40 +LM1 +LM2 +LM3 +FI +Kondo +FI +μc (meV) +G=14meV +N +11 +21 +31 +41 +1 +(1, 1;1/2, 1/2) +(1, 1;1/2, 1/2) +(0, 0;0, 0) +(1, 0;1/2, 0) +(1, 0;1/2, 0) +(2, 0;0, 0) +-40 +40 +ω (meV) +N +11 +21 +31 +41 +1 +(2, 1;0, 1/2) +(2, 1;0, 1/2) +(1, 1;1/2, 1/2) +(1, 1;1/2, 1/2) +(2, 2;0, 0) +(1, 0;1/2, 0) +-40 +40 +ω (meV) +0 +100 +200 +150 +50 +11 +21 +31 +41 +N +(0, 0;0, 0) +(1, 0;1/2, 0) +(-1, 0;1/2, 0) +(-2, -1;0, 1/2) +(1, 0;1/2, 0) +(-1, 0;1/2, 0) +(0, 0;0, 0) +1 +-40 +40 +ω (meV) +E (meV) +0 +100 +150 +50 +11 +21 +31 +41 +1 +(-1, 0;1/2, 0) +(-1, 0;1/2, 0) +(0, 0;0, 0) +(0, 0;0, 0) +(1, 0;1/2, 0) +(1, 0;1/2, 0) +-40 +40 +ω (meV) +N +11 +21 +31 +41 +N +(0, 0;0, 0) +(1, 0;1/2, 0) +(1, 0;1/2, 0) +(-1, 0;1/2, 0) +1 +(1, 1;1/2, 1/2) +(1, 1;1/2, 1/2) +-40 +40 +ω (meV) +FIG. 2. Phase diagram and fixed points. (a) The self-consistent mean-field values of µc, µf, G as functions of the total filling ν +from ν = 0 to 4, where we have enforced the symmetries of the correlated insulator state at CNP. The left y-axis represents µc +and µf while the right y-axis represents G with a different range. (b) The phase diagram in the parameter space of µc, µf for +G = 8meV. The white lines are phase boundaries between the local moment (LM) phases and the strong coupling phase. The +dashed black lines are crossover boundaries between the frozen impurity (FI) and Kondo regimes of the strong coupling phase. +The color maps the half-width of the spectral density peak, reflecting the Kondo energy scale if in the Kondo regime. The +solid black line indicates the trajectory of µc and µf determined from a self-consistent calculation as ν changes from 0 to 4, +where the five arrows from left to right represent ν = 0, 1, 2, 3, 4, respectively. (c) is the same as (b) but a different parameter +G = 14meV is used. (d)(e) The RG flow of the many-body spectrum of the scaled Hamiltonian � +HN (N ∈ odd) in the Kondo +regime, where µc = 30.7meV, µf/U1 = 0.367, G = 9.49meV is the mean field value at ν = 1.25 for (d) and µc = 49.9meV, +µf/U1 = 1.286, G = 11.83meV is the mean field value at ν = 2.5 for (e). The spectral lines’ colors represent the symmetry +sectors labeled by good quantum numbers (Q1, Q2; S1, S2). Since the levels in sectors (Q1, Q2; S1, S2) and (Q2, Q1; S2, S1) are +identical, only |Q1| ≥ |Q2| sectors are shown for simplicity. The insets are the resulting single-particle spectral densities that +exhibit Kondo resonances. (f)(g)(h) The RG flow of many-body spectrum of the scaled Hamiltonian � +HN (N ∈ even) in the +LM1,2,3 phase, where µc = 5meV, µf/U1 = 0.5, 1.5, 2.5, G = 8meV respectively. The insets are the resulting single-particle +spectral densities that exhibit Hubbard bands. +low-lying many-body spectrum is identical to the one of +a free-fermion chain defined by ϵn and tn with an ad- +ditional chemical potential term. +In other words, the +impurity acts as if it was nonexistent. The underlying +mechanism is either the Kondo screening, where the im- +purity is an effective LM screened by the bath, or the +impurity freezing, where the impurity occupation νf is +effectively empty or full. We refer to the two cases as the +Kondo regime and the frozen impurity (FI) regime, re- +spectively, which are adiabatically connected. The fixed +points shown in Fig. 2(d), (e) are in the Kondo regime +because, if we continuously change µc to 0, they evolve +to the LM1 and LM2 states (discussed in the next para- +graph), respectively. There is a crossover between Kondo +and FI regimes as one changes µf, as indicated by the +dashed lines in Fig. 2(b), (c). Later we will determine +the crossover boundary using the spectral density. +At an LM fixed point, as exampled in Fig. 2(f), the low- +lying many-body spectrum is identical to a free-fermion +chain plus a detached LM. Depending on the represen- +tation of the ground state, the LM fixed points can be +further classified into LMn, where n = 1, 2, 3 is the +effective impurity occupation. +The flows of the spec- +tra towards these fixed points are shown in Fig. 2(f), +(g), (h), respectively. +LMn ground states have the +same SU(2)×SU(2) representations as ground states of +ˆHf + ˆHH (Eqs. (12) and (18)) with n impurity elec- +trons, where the Hubbard interaction freezes charge exci- +tations and the Hund’s coupling prefers states with elec- +trons lying in different orbitals. For n = 1, the ground +states are four-fold degenerate and belong to the sym- +metry sectors (Q1, Q2; S1, S2) = (2k + 1, 2k; 1 +2, 0) and +(2k, 2k + 1; 0, 1 +2), corresponding to the spin- 1 +2 states of +the two U(2)’s, respectively. The SU(2) representations +are the same as those of the four single-particle states of +ˆHf + ˆHH: (nf +1↑, nf +1↓; nf +2↑, nf +2↓) = (10;00), (01;00), (00;10), +(00;01). +For n = 2, the ground states are also four- +fold degenerate but belong to a different symmetry sec- +tor (2k + 1, 2k + 1; 1 +2, 1 +2). +One can understand them +as the product states of two spin- 1 +2 states of the two +U(2)’s. They have the same SU(2) representations as the +four two-particle states of ˆHf + ˆHH: (10;10), (10;01), +(01;10), (01;01). Without Hund’s coupling JH, the LM2 +ground states would be +�4 +2 +� += 6-fold degenerate. A fi- + +11 +0 +0.1 +0.2 +0.3 +0.4 +0.5 +-40 +0 +40 +(a) kBT=0 +-40 +0 +40 +0 +1 +2 +3 +4 +5 +(b) kBT=0 +(c) kBT=2meV +(d) kBT=5meV +0 +0.1 +0.2 +0.3 +0.4 +0.5 +-40 +0 +40 +0 +0.1 +0.2 +0.3 +0.4 +0.5 +-40 +40 +-5 5 +A(𝜔,T) (meV-1) +𝜔 (meV) +𝜔 (meV) +𝜔 (meV) +𝜔 (meV) +𝜈=0 +𝜈=1 +𝜈=2 +𝜈=3 +𝜈=4 +FIG. 3. Spectral densities A(ω, T) at various fillings ν and +temperatures. (a) Spectral densities at the zero temperature +for ν = 0, 0.1, 0.2 · · · 4. The curves are offset by 0.1ν (meV−1) +for clarity. (b) is the same as (a) but is shown with a smaller +vertical scale for clarity of Hubbard bands, marked by in- +verted triangles. The curves are offset by 0.01ν (meV−1). (c) +and (d) are spectral densities at finite temperatures, where +the curves are offset by 0.01ν (meV−1). +nite JH raises the energy of the two many-body states +with both electrons occupying the same orbital, i.e., +(11;00), (00;11). For n = 3, the ground states are still +four-fold degenerate but belong to the symmetry sec- +tors (2k − 1, 2k; 1 +2, 0) and (2k, 2k − 1; 0, 1 +2). Their SU(2) +representations are same as the four single-hole states +ˆHf + ˆHH: (01;11), (10;11), (11;01), (11;10). +It is also helpful to look at the global U(2) symme- +try representations, where the two U(2) rotations are the +same. The total charge and spin of LM1,2,3 are 1, 2, 3 +(mod 4) and 1 +2, 1 +2 ± 1 +2, 1 +2, respectively. +Phase boundaries between different LM phases and the +strong coupling phase are described by unstable fixed +points where different ground states cross with each +other. +The phase boundaries are shown by the white +lines in Fig. 2(b), (c). Starting from an LMn phase, in- +creasing µc will eventually drive it into a strong coupling +phase due to the enhancement of hybridization. The crit- +ical µc, as expected, is close to G, the conduction band +edge (Fig. 1(c), (d)). +V. +SPECTRAL DENSITY +We calculate the spectral density of the f-electrons, +A(ω, T) = − 1 +π +� +αs Gαs(ω, T), with Gαs(ω, T) being the +retarded Green’s function of fαs at the temperature T. +We use the method described in Ref. [100] to collect +the many-body levels at different RG steps to compute +A(ω, T). +The fixed points in Fig. 2(d), (e) are in the +Kondo regime and hence have sharp zero-energy peaks +due to the Kondo resonance, as shown in the insets of +Fig. 2(d), (e). +The fixed points in Fig. 2(f), (g), (h) +are in the LM1,2,3 phases, respectively, thus, their spec- +tral density are dominated by the upper and lower Hub- +bard bands. We compute the spectral densities for all +the points in the phase diagrams in Fig. 2(b), (c). We +identify a central peak for every calculation and measure +its half-width δK at half maximum. (If there is no cen- +tral peak, e.g., Fig. 2(f), δK = 0.) δK is indicated by the +color in Fig. 2(b), (c). We can distinguish the Kondo and +FI regimes in the strong coupling phase through spectral +density. Intuitively, a state in the Kondo regime should +have a Kondo resonance. By contrast, a state in the FI +regime should have its main spectral weight away from +zero energy because the impurity occupation is either +empty or full. +Thus, we identify a phase point in the +Kondo regime if δK covers the zero energy and otherwise +in the FI regime. The crossover between the two regimes +is indicated by dashed lines in Fig. 2(b), (c). +Several features of δK in Fig. 2(b), (c) can be un- +derstood using the poor man’s scaling developed in +Sec. IV A. First, there are three domes around µf = +1 +2U1, 3 +2U1, 5 +2U1 where δK is relatively small. They cor- +respond to the nf = 1, 2, 3 cases discussed in Sec. IV A. +From the poor man’s scaling perspective, these three µf’s +correspond to the minimal initial value of the coupling +constant g (Eq. (22)), which then leads to smaller TK’s. +Second, when µc is small (≲ 30meV), δK in the middle +dome is significantly smaller than those of the other two +domes. The reason is that the Kondo energy scale TK +for nf = 2 will be strongly suppressed due to the multi- +plet splitting if TK is smaller than JH. One can see that +the N factor in the exponential function of Eq. (27) is +replaced by 2. Third, for the same µc, the first dome has +lower δK than the third dome. This difference is a result +of the suppression factor y1 for nf = 1 (Eq. (24)) and the +enhancement factor y3 for nf = 3 (Eq. (29)) due to the +particle-hole asymmetry of ∆(ω), as discussed in detail +in Sec. IV A. +In order to compare our results with STM measure- +ments, we need to adopt physical µc, µf, G parameters. +As discussed in Sec. III C, µc, µf, G can be determined as +functions of the filling ν via a symmetric self-consistent +calculation of Eqs. (10). µc, µf, G as functions of ν are +shown in Fig. 2(a). The obtained spectral densities at the +zero temperature are shown in Fig. 3(a), (b). For ν = 0, +the state is in the FI regime with an (almost) zero occu- +pation; hence the spectral weight is mainly distributed at +positive energy. As ν increases, the spectral peak moves +to the zero energy and is eventually pinned at the zero en- +ergy to form a Kondo resonance. This is precisely what is +seen in STM experiments at low temperatures (T < 1K) +[11, 19, 21, 22]. One can also observe the evolution of +Hubbard bands at T = 0 as ν changes (Fig. 3(b)), but +they are relatively weak compared to the Kondo reso- +nance peaks. +At finite temperatures (Fig. 3(c), (d)), +the Kondo resonance peaks are smeared by thermal fluc- +tuations and the evolution of Hubbard bands becomes +clearer. As ν increases from 0 to 4, the Hubbard bands +periodically pass through the zero energy, matching the +cascade of transitions seen in STM experiments at higher + +12 +(a) +kBT (meV) +4 +102 +101 +100 +10-1 +10-2 +10-3 +𝜈 +0 +1 +2 +3 +101 +100 +10-1 +10-2 +102 +101 +100 +10-1 +10-2 +10-3 +kBT (meV) +102 +101 +100 +10-1 +10-2 +10-3 +kBT (meV) +100 +10-1 +10-2 +10-3 +0.0 +0.5 +1.0 +1.5 +2.0 +2.5 +3.0 +3.5 +4.0 +𝜈 +(b) +(d) +(c) +102 +101 +100 +10-1 +10-2 +10-3 +kBT (meV) +𝜈 +0 +1 +2 +3 +4 +0 +1 +2 +3 +(e) +𝜔 (meV) +0 +20 +40 +-20 +-40 +0 +0.1 +0.2 +0.3 +A(ω,T) (meV-1) +𝜈=1 +0 +0.1 +0.5 +1 +2 +5 +kBT (meV) +kBT =0.82meV +(f) +𝜈 +0 +1 +2 +3 +4 +0 +0.4 +0.8 +1.2 +1.6 +Bloc(T) +0 +5 +10 +15 +20 +0.0 +0.5 +1.0 +1.5 +2.0 +2.5 +3.0 +3.5 +4.0 +𝜈 +0 +1 +2 +3 +4 +Simp +×kBln2 +FIG. 4. Spin susceptibility and entropy contributed by the im- +purity. (a) χloc(T)/χloc(0) as a function of filling ν and tem- +perature T. (b) The local spin susceptibilities χloc(T) at fill- +ings ν = 0, 0.5, 1 · · · 4. (c) The entropy contributed by the im- +purity as a function of ν and T. (d) The entropy contributed +by the impurity Simp(T)/(kB ln 2) at fillings ν = 0, 0.5, 1 · · · 4. +(e) The spectral densities at ν = 1 at various temperatures. +(f) The entropy contributed by the impurity as a function of +ν at B = 0, 5, 10, 15, 20 T and temperature kBT = 0.82 meV. +temperatures [17, 19]. +One can use δK to estimate the Kondo energy scale. +Using the ν-dependent µc, µf, G parameters given in +Fig. 2(a), we tabulate the δK’s at different fillings in Ta- +ble II. Comparing it to TK estimated by the poor man’s +scaling, denoted as T (P ) +K , we find T (P ) +K +is about δK ∼ 2δK. +VI. +LOCAL MOMENTS AND THE +POMERANCHUK EFFECT +At a temperature exceeding the Kondo energy scale, +the LM will become effectively decoupled from the bath +and visible in experimental measurements. This is the +mechanism of the Pomeranchuk effect [30, 31]. Refs. [30] +observed a higher entropy (∼ 1kB per moir´e cell with kB +being the Boltzmann’s constant) state at ν ≈ 1 at the +temperature T ≈ 10K. As this entropy can be quenched +by an in-plane magnetic field, it is ascribed to a free local +moment. Ref. [31] observed a similar effect at ν ≈ −1 +and showed that an additional resistivity peak that is +absent at T = 0 develops in the higher entropy state at +T ≈ 10K. These observations can be naturally explained +by the transition from the Fermi liquid phase to the LM +phase as the temperature increases. +To demonstrate the LM phase at higher temperatures, +we calculate the local spin susceptibilities χloc(T) us- +ing the filling-dependent µc, µf, G parameters given in +Fig. 2(a). +χloc(T) is defined as +dMloc +dBloc [101, 102], with +Mloc being the spin momenta contributed by the impu- +rity and Bloc a local magnetic field that only acts on the +impurity. As shown in Fig. 4(a), (b), for ν ≥ 0.5, χloc(T) +approaches a constant as T → 0, and obeys the Curie’s +law, i.e., χloc(T) ∼ T −1, when T is larger than the Kondo +energy scale. Obeying Curie’s law is a clear indication of +a free LM. One may notice that the T-dependences of +χloc(T)’s for ν < 0.5 are non-monotonous. The ν < 0.5 +states lie in the FI regime, thus the spin susceptibili- +ties are extremely small when T → 0, and will start to +increase when T is able to excite the LM1 states. For +ν > 0.5, we can define the Kondo temperature TK as +the turning point between Curie’s behavior and Fermi +liquid behavior. Specifically, it can be obtained as the +crossing of the extended T −1 line from the LM side and +the extended horizontal line from the Fermi liquid side +(Fig. 4(b)). We tabulate the resulting TK in Table II. As +shown in the table, such defined TK is about 1 +3δK ∼ δK +with δK being the half-width of the spectral density dis- +cussed in Sec. V. +We also calculate the impurity entropy Simp(T) for +comparison with experiments. Simp(T) is defined as the +difference of the entropy of �HN and that of a reference +free-fermion chain (without the impurity site) defined +by the same ϵn, tn parameters as in �HN. As shown in +Fig. 4(c) and (d), Simp(T) is zero in the Fermi liquid +phase at sufficiently low T and starts to increase when +T reaches the Kondo energy scale. For ν = 1, Simp(T) +climbs to about 2 ln 2 · kB at about kBT ≈ 1meV and +(approximately) stays at this value until kBT reaches +10meV. The entropy 2 ln 2 · kB ≈ 1.39kB is close to the +measured value (∼ 1kB) in Refs. [30, 31] and can be un- +derstood as contributed by the four degenerate states in +the LM1 phase. Higher excited states will be involved +when kBT is larger than 10meV, and the entropy contin- +ues to increase for larger kBT. We also show the spectral +density at ν = 1 in Fig. 4(e), one can see that the Kondo +resonance peak becomes weak for kBT > 1meV, which +is consistent with the entropy and spin susceptibility re- +sults. +An in-plane magnetic field will polarize the spin and +hence suppress the entropy. +However, as discussed in +Sec. IV C, the four-fold degenerate LM1 states consist of +two spin- 1 +2 states due to the orbital degeneracy, hence +a strong field will not completely quench the entropy. +Instead, due to the orbital degeneracy, the remaining en- +tropy will be ln 2 · kB ≈ 0.69kB. This is also consistent +with observations in Ref. [30]. In Fig. 4(f), we plot the + +13 +(a) 𝜈=1.2 +(b) 𝜈=1.8 +(c) 𝜈=2.4 +Band Energy (meV) +0 +20 +40 +60 +-20 +-40 +-60 +ΓM +KM +MM +ΓM +KM +MM +ΓM +KM +MM +0 +0.2 +0.4 +0.6 +0.8 +1 +FIG. 5. Heavy Fermi liquid bands at ν = 1.2 (a), 1.8 (b), 2.4 +(c). As explained in the text, the effective hybridization be- +tween c- and f-bands is suppressed by a factor z +1 +2 with z being +the quasi-particle weight of f-electrons, which are estimated +as 0.038, 0.27, 0.16 for (a), (b), and (c), respectively. The +color of the bands represents the total quasi-particle weight, +which is always larger than z. +impurity entropy as a function of the filling at various +magnetic fields Bloc. At ν = 1, the entropy saturates +to a constant ∼ ln 2 · kB under a strong field. One can +see that the entropy values, the shapes of the curves, and +their field-dependences in Fig. 4(f) are comparable to the +experimental results in Fig. 2(e) of Ref. [30]. +VII. +DISCUSSIONS +Based on the NRG calculations, the poor man’s scal- +ing, and various experimental observations, we have +shown that the gapless states at 1 ≲ |ν| < 2 are in the +Kondo regime. Considering the translation invariance of +the actual system, these states must be the heavy Fermi +liquid states. Here we estimate the heavy Fermi liquid +bands from the information provided by the NRG cal- +culation. At the zero temperature, a spectral density in +the Kondo regime possesses a Lorentz peak around the +zero energy, i.e., A(ω) ≈ 4z +π +δK +ω2+δ2 +K , where z is the quasi- +particle weight, δK is the half-width given in Fig. 2(b), +(c), and the factor 4 is from orbital and spin degenera- +cies. Then the quasi-particle weight can be estimated as +z = πA(0)δK/4. Substituting the quasi-particle part of +Gαs(iω), i.e., +z +iω, into Dyson’s equation of c-electrons +G(c)(iω, k) = G(c,0)(iω, k) ++ G(c,0)(iω, k)H(cf)(k) z +iω H(cf)†(k)G(c)(iω, k) , +(32) +one can see that the cf hybridization is effectively sup- +pressed by a factor of z +1 +2 . Here ω is the Matsubara fre- +quency, G(c,0)(iω, k) = (iω − H(c)(k))−1 is the free prop- +agator of c-electrons, and H(c)(k), H(cf)(k) are Hamilto- +nian matrices in Eq. (9). If z = 0 the effective hybridiza- +tion will be zero, corresponding to the LM phase where +the Fermi surface is solely contributed by the c-electrons. +Using this method we obtain the bands at ν = 1.2 and +1.8, as shown in Fig. 5(a), (b), respectively. One should +be aware that the Hubbard band information is com- +pletely neglected in this method. For future reference, +we also estimate the heavy Fermi liquid bands at ν = 2.4 +(Fig. 5(c)) by assuming the ν = 2.4 state in the Kondo +regime. +The heavy Fermi liquid states at 1 ≲ |ν| < 2 can be +further confirmed by future experimental research. For +example, the Fermi surface can dramatically change as +one tunes the filling and temperature or applies an ex- +ternal field. The Fermi surface change will be reflected +in spectral measurements such as quasi-particle interfer- +ence. It is also in principle possible to directly measure +the scattering phase shift [103]. +c-bands will induce RKKY interactions between LMs +at different f-sites, which can lead to further symme- +try breaking and should be crucial to stabilize the ob- +served correlated insulator states at |ν| = 2. We have +ignored these RKKY interactions and further symmetry +breaking in the current work. A full self-consistent treat- +ment including both RKKY and Kondo screening effects +may result in more complicated ν-dependencies of µc, µf +than those shown Fig. 2(a), (b), (c). +For example, at +ν = 2, µc given by a self-consistent symmetry breaking +Hartree-Fock mean field [61] is about 26meV, which is +significantly lower than the one (∼40meV) in Fig. 2(a). +Thus, a possible mechanism for the correlated insulators +to win the Kondo screening is that µc drops to a small +value such that the Kondo energy scale becomes irrele- +vant. (See Fig. 2(b), (c).) Observations in Refs. [18, 26] +also suggest that the ν-dependencies of µc, µf are com- +plicated. The competition between RKKY and Kondo +screening is also a potential mechanism for the observed +strange metal behaviors [27–29] and could play an impor- +tant role in the unconventional superconductivity [1, 4– +11]. We leave this for future studies. +Note added. +During the preparation of the current +work, a related work [104] appeared. This work studied +the symmetric Kondo state using a slave-fermion mean +field in a Kondo lattice model derived from the THF +model. Our theory is based on the symmetry-broken cor- +related state at CNP. We are also aware of related works +on the Kondo problem in MATBG by A. M. Tsvelik’s and +B. A. Bernevig’s group [105, 106] and P. Coleman’s group +[107] that will appear soon, and a generalization of the +THF model to the magic-angle twisted trilayer graphene +[108]. Ref. [105] also obtains a Kondo temperature about +1 ∼ 2K around |ν| ≈ 1. +ACKNOWLEDGMENTS +We are grateful to B. Andrei Bernevig, Ning-Hua +Tong, Xi Dai, Jia-Bin Yu, Xiao-Bo Lu, Yong-Long Xie, +Yi-Lin Wang, and Chang-Ming Yue for helpful discus- +sions. Z.-D. S. and G.-D. Z. were supported by National +Natural Science Foundation of China (General Program +No. 12274005), National Key Research and Development +Program of China (No. 2021YFA1401900). + +60 +0.9 +40 +0.8 +0.7 +20 +0.6 +0 +0.5 +0.4 +-20 +0.3 +0.2 +-40 +0.1 +-60 +0 +K +G +M60 +0.9 +40 +0.8 +0.7 +20 +0.6 +0 +0.5 +0.4 +-20 +0.3 +0.2 +-40 +0.1 +-60 +0 +K +G +M60 +0.9 +40 +0.8 +0.7 +20 +0.6 +0 +0.5 +0.4 +-20 +0.3 +0.2 +-40 +0.1 +-60 +0 +K +G +M60 +0.9 +40 +0.8 +0.7 +20 +0.6 +0 +0.5 +0.4 +-20 +0.3 +0.2 +-40 +0.1 +-60 +0 +K +G +M14 +Appendix A: More details about the effective +Hamiltonian +1. +Nonzero M term +A generic trial ground state at CNP is given by +(Eq. (6)) +|Ψ0⟩ = U +� +R +f † +R1+↑f † +R1+↓f † +R2+↑f † +R2+↓|FS⟩ , +(A1) +where U = exp(−iθµν ˆΣµν) is a U(4) rotation operator +and an implicit summation over repeated µ, ν indices is +assumed. We can always define the rotated fermion op- +erators �ckaηs = UckaηsU †, �fRaηs = UfRaηsU † such that +�fRα+s’s are occupied in |Ψ0⟩ and �fRα−s’s are empty in +|Ψ0⟩. According to the discussions in the supplementary +material section S4B of Ref. [61], in the flat-band limit +(M = 0), the lowest particle (hole) excitations only in- +volve �cka−s and �fRa−s (�cka+s and �fRa+s). +Thus, the +effective periodic Anderson model for ν > 0 derived +in Sec. III B is written in terms of ckas = �cka−s and +fRαs = �fRα−s. Here we give the explicit forms of the +rotated operators +�fRαηs = +� +α′η′s′ +� +eiθµνΣf +µν +� +αηs,α′η′s′ fRα′η′s′ , +(A2) +and +�ckaηs = +� +a′=1,2 +η′s′ +� +eiθµνΣc12 +µν +� +aηs,a′η′s′ cka′η′s′ +(a = 1, 2), (A3) +�ckaηs = +� +a′=3,4 +η′s′ +� +eiθµνΣc34 +µν +� +aηs,a′η′s′ cka′η′s′ +(a = 3, 4), (A4) +where the eight-by-eight matrices Σf +µν, Σc12 +µν , Σc34 +µν +are +defined in Eqs. (3) to (5). +The M-term in the original basis of the THF model +(Eq. (1)) is +M +� +aa′=3,4 +� +|k|<Λc +� +ηs +[σx]aa′c† +kaηscka′ηs . +(A5) +It favors the Kramers inter-valley coherent state dis- +cussed at the end of Sec. II, where θx0 and θy0 are nonzero +and satisfy θ2 +x0 + θ2 +y0 = (π/4)2. Without loss of general- +ity, we assume U = exp(−i π +4 ˆΣx0) for the Kramers inter- +valley coherent state. Writing this M-term in terms of +the rotated operators, we obtain +M +� +|k|<Λc +� +a,a′=3,4 +� +ηη′ss′ +�c† +kaηsOaηs,a′η′s′�cka′η′s′ , +(A6) +where O = ei π +4 Σc34 +x0 σxτ0ς0e−i π +4 Σc34 +x0 += −σzτxς0. The τx +matrix in O represents couplings between the empty +and occupied single-particle states. If we simply project +this M-term onto the empty states, it vanishes, i.e., +[Oa−s,a′−s′] = 0. +A better approximation is applying +a Schrieffer-Wolff transformation to decouple the η = ± +states, leading to a second-order correction to the effec- +tive Hamiltonian. As ⟨Ψ0| �f † +αηs �fαηs|Ψ0⟩ = (1 + η)/2, the +J term in Eq. (2) yields the following mean field term (see +also the supplementary material section S4B of Ref. [61]) +− J +2 +� +a=3,4 +� +ηs +η · �c† +aηs�caηs +(A7) +Then, regarding the J +2 term as the zeroth order Hamilto- +nian and M as a perturbation, a Schrieffer-Wolff trans- +formation leads to the correction +− M 2 +J +� +|k|<Λc +� +a=3,4 +� +ηs +η · �c† +kaηs�ckaηs + O(M 4) . +(A8) +The resulting energy levels ±(J/2 + M/J2) at k = 0 is +fully consistent with a Taylor expansion of the one-shot +energy levels ± +� +J2/4 + M 2 derived in Ref. [61]. Pro- +jecting the correction to the active d.o.f., i.e., ckas = +�cka−s, we obtain the correction to the effective Hamilto- +nian +M 2 +J +� +|k|<Λc +� +a=3,4 +� +s +c† +kasckas + O(M 4) . +(A9) +2. +Hund’s coupling +The four-by-four Hamiltonian matrix H(c)(k)+∆H(c) +in Eq. (13), i.e., −v⋆(σx⊗σ0kx+σy ⊗σzky)+02×2⊕Gσ0, +can be diagonalized analytically. As discussed at the end +of the last subsection, to O(M 2), the M term simply +shifts the energy of a = 3, 4 electrons by M 2/J. Thus, +all the analysis below applies to the M ̸= 0 after G is +replaced by G + M 2/J. We find the energy eigenvalues +and wave-functions of the H(c)(k) + ∆H(c) as +ϵ1(k) =ϵ+(k) = G +2 + +� +G2 +4 + v2⋆k2 +u1(k) = +� +sin θk +2 e−iφk 0 − cos θk +2 +0 +�T +, +(A10) +ϵ2(k) =ϵ+(k) = G +2 + +� +G2 +4 + v2⋆k2 +u2(k) = +� +0 sin θk +2 eiφk 0 − cos θk +2 +�T +, +(A11) +ϵ3(k) =ϵ−(k) = G +2 − +� +G2 +4 + v2⋆k2 +u3(k) = +� +cos θk +2 e−iφk 0 sin θk +2 +0 +�T +, +(A12) +ϵ4(k) =ϵ−(k) = G +2 − +� +G2 +4 + v2⋆k2 +u4(k) = +� +0 cos θk +2 eiφk 0 sin θk +2 +�T +. +(A13) + +15 +where +θk = arccos +G/2 +� +G2/4 + v2⋆k2 +(A14) +and φk = arg(kx + iky). +We now derive the effective Hund’s coupling ˆHH. Ap- +plying a second-order perturbation in terms of ˆHJ, we +obtained the correction to the Hamiltonian +∆ ˆH = − J2 +N 2 +M +� +I +� +α1α2 +s1s′ +1s2s′ +2 +� +k1,k′ +1 +k2,k′ +2 +(f † +α1s1fα1s′ +1 − νf +4 δs1s′ +1) +× (f † +α2s′ +2fα2s2 − νf +4 δs2s′ +2) · e− λ2 +2 (k2 +1+k′2 +1 +k2 +2+k′2 +2 ) +× +⟨Ψ0|c† +k′ +1α1+2s′ +1ck1α1+2s1|ΨI⟩⟨ΨI|c† +k2α2+2s2ck′ +2α2+2s′ +2|Ψ0⟩ +EI − E0 +, +(A15) +where |ΨI⟩ are excited states with a single particle-hole +pair and EI are the energies of the excited states. k1,2, +k′ +1,2 are limited within the cutoff Λc. Due to the mo- +mentum and spin conservation, for the matrix element +to be nonzero, there must be k1 = k2, s1 = s2, k′ +1 = k′ +2, +s′ +1 = s′ +2. For simplicity, we rewrite k1, k′ +1, s1, and s′ +1 +as k, k′, s, and s′, respectively. (k, s) and (k′, s′) label +the particle and the hole excitations, respectively. Then +the matrix element in the third line of Eq. (A15) can be +written as +nF (ϵ+(k′) − µc)(1 − nF (ϵ−(k) − µc)) +× ⟨Ψ0|c† +k′α1+2s′ckα1+2sc† +kα2+2sck′α2+2s′|Ψ0⟩ +(A16) +According +to +the +wave +functions +given +in +Eqs. +(A10)-(A13), +there are ⟨Ψ0|c† +k′α1+2s′ck′α2+2s′|Ψ0⟩ += +δα1α2 sin2 θk′ +2 , ⟨Ψ0|ckα1+2sc† +kα2+2s|Ψ0⟩ = δα1α2 cos2 θk +2 . +The excitation energy EI −E0 is given by ϵ+(k)−ϵ−(k′). +Thus, ∆ ˆH is simplified to +∆ ˆH = − J2 +N 2 +M +� +αss′ +kk′ +(f † +αsfαs′ − νf +4 δss′)(f † +αs′fαs − νf +4 δss′) +× nF (ϵ−(k′) − µc)(1 − nF (ϵ+(k) − µc)) sin2 θk′ +2 cos2 θk +2 +ϵ+(k) − ϵ−(k′) +× e−λ2(k2+k′2) +(A17) +The s = s′ contribution is an effective chemical potential +shift, estimated as 0.17meV at CNP, of the f-electrons. +As it is much smaller than U1, we will omit the s = s′ +contribution. The s ̸= s′ contribution can be written as +ˆHH = JH +� +α +nf +α↑nf +α↓ +(A18) +with JH given by +JH =2J2 +�Ω0 +2π +�2 ˆ Λc +0 +dk′ · k′ +ˆ Λc +k0 +dk · k · e−λ2(k2+k′2) +× sin2 θk′ +2 cos2 θk +2 +ϵ+(k) − ϵ−(k′) , +(A19) +where k0 is determined by ϵ+(k0) = µc. Here we have +made use of the fact that ϵ±(k) and θk only depends +on |k| but not φk. At CNP, µc = 0 and G = J/2 = +8.19meV, taking the limit Λc → ∞, we obtain JH ≈ +0.34meV. Using the self-consistent values of µc and G +shown in Fig. 2(a), we find JH at ν = 1, 2, 3, 4 are given +by 0.29meV, 0.26meV, 0.21meV, 0.19meV, respectively. +As JH is small and does not change significantly with ν, +in this work, we simply set JH = 0.34meV for simplicity. +3. +Hybridization function +By definition, the hybridization function ∆(ω) is given +by +∆(ω) = π +N +� +k +� +n +|Vnα(k)|2δ(ω − ϵn(k)) +(A20) +where Vnα(k) = � +a u∗ +an(k)H(cf) +aα (k)e− λ2k2 +2 +is the hy- +bridization between fαs and the n-th energy band of c- +electrons. +∆(ω) does not depend on α because of the +C2zT or C2x symmetry that flips the α index. Substitut- +ing ϵn(k) and uan(k) in Eqs. (A10)-(A13) into the above +equation, we obtain Vnα(k) for α = 1 as +V11(k) =γ sin θk +2 eiφk +V21(k) =v′ +⋆(−kx + iky) sin θk +2 e−iφk +V31(k) =γ cos θk +2 eiφk +V41(k) =v′ +⋆(−kx + iky) cos θk +2 e−iφk . +(A21) +Using the energy eigenvalues in Eqs. (A10)-(A13) and +the Vnα(k) matrix elements given above, it is direct to +obtain +∆(ω) = Ω0 +2v2⋆ +����ω + µc − G +2 +���� +� +γ2 + v′2 +⋆ k2 +F +� +e−k2 +F λ2 +� +θ(ω + µc − G) sin2 θkF +2 ++ θ(−ω − µc) cos2 θkF +2 +� +(A22) +where kF is given by +kF = 1 +v⋆ +� +(ω + µc − G/2)2 − (G/2)2 . +(A23) +We now verify the asymptotic behaviors of ∆(ω). +When ω + µc → G+, kF → 0 and only the first term +in the second line of Eq. (A22) contributes to ∆(ω). Ac- +cording to Eq. (A14), there is cos θkF = +G/2 +ω+µc−G/2 and +hence sin2 θkF +2 += 1 +2− 1 +2 cos θkF ≈ (ω+µc−G)/G. Then we +obtain the asymptotic behavior of ∆(ω) as ω + µc → G+ +∆(ω) = Ω0 +4v2⋆ +γ2 · (ω + µc − G) + O((ω + µc − G)2) . +(A24) +When ω + µc → −0+, kF → 0 and only the second +term in the second line of Eq. (A22) contributes to ∆(ω). + +16 +According to Eq. (A14), there is cos θkF = +G/2 +G/2−ω−µc and +hence cos2 θkF +2 += 1 +2 + 1 +2 cos θkF ≈ 1. Then we obtain the +asymptotic behavior of ∆(ω) as ω + µc → −0+ +∆(ω) = Ω0 +4v2⋆ +Gγ2 + O((ω + µc)) . +(A25) +Appendix B: Poor man’s scaling of Anderson models +with energy-dependent couplings +1. +Generic theory for U(N) models +We consider the Anderson impurity model with N +symmetric flavors +ˆH = − µf ˆ +Nf + U +2 +ˆ +Nf( ˆ +Nf − 1) + +N +� +µ=1 +ˆ D +−D +dϵ · ϵ · d† +µ(ϵ)dµ(ϵ) ++ +N +� +µ=1 +ˆ D +−D +dϵ +� +∆(ϵ) +π +(f † +µdµ(ϵ) + h.c.) , +(B1) +where µ is the flavor index and Nf = �N +µ=1 f † +µfµ. We +assume the ground state of the isolated impurity has nf +f-electrons, which can take the values 1, 2 · · · (N − 1). +(We do not consider the empty case (nf = 0), the full +case (nf = N), and the mixed valence case where ground +states with different nf are exactly degenerate.) We fur- +ther assume the charge gaps to nf − 1 and nf + 1 elec- +trons are ∆E− and ∆E+ = U − ∆E−, respectively. We +then apply a Schrieffer-Wolff transformation to obtain an +effective Coqblin–Schrieffer model for the Hilbert space +restricted to ˆNf = nf +ˆH = +N +� +µ=1 +ˆ D +−D +dϵ · ϵ · d† +µ(ϵ)dµ(ϵ) + 4g +πU +N +� +µ,µ′=1 +ˆ D +−D +dϵdϵ′ +� +� +∆(ϵ)∆(ϵ′)(f † +µfµ′ − xδµµ′)d† +µ′(ϵ′)dµ(ϵ) +� +. +(B2) +Terms that only involve ˆNf are omitted because they +only contribute to an energy shift. +The bandwidth D +should be smaller than min(∆E+, ∆E−), otherwise, the +Schrieffer-Wolff transformation is invalid. The parame- +ters g and x are given by +g = U +4 +� +1 +∆E+ ++ +1 +∆E− +� +, +x = ∆E− +U +, +(B3) +respectively. If µf = (nf − 1 +2)U, there is ∆E+ = ∆E− = +1 +2U and g = 1, x = 1 +2. +We now truncate the bandwidth at D−dD = D(1−dt) +(dt ≪ 1) and consider second order (in g) corrections +form the virtual particle (D − dD < ϵ < D) and hole +(−D < ϵ < −D + dD) excitations. The particle excita- +tion contributes to the correction +− (4g)2 +(πU)2 +1 +D +� +µ1µ2µ′ +1µ′ +2 +ˆ D−dD +−D+dD +dϵ1dϵ2d +ˆ D +D−dD +dϵ′ +1dϵ′ +2 +× +� +∆(ϵ1)∆(ϵ2)∆(ϵ′ +1)∆(ϵ′ +2)d† +µ1(ϵ1)⟨dµ′ +1(ϵ′ +1)d† +µ′ +2(ϵ′ +2)⟩dµ2(ϵ2) +× (f † +µ′ +1fµ1 − xδµ1µ′ +1)P(f † +µ2fµ′ +2 − xδµ2µ′ +2) . +(B4) +The denominator D in the factor is the excitation en- +ergy of a virtual particle. +P is a projector to the re- +stricted Hilbert space, where ˆNf = nf. +The expecta- +tion ⟨dµ′ +1(ϵ′ +1)d† +µ′ +2(ϵ′ +2)⟩ evaluated on the ground state is +δ(ϵ′ +1 − ϵ′ +2)δµ′ +1µ′ +2. Then we have +− (4g)2 +(πU)2 +dD +D ∆(D) +� +µ1µ2µ′ +ˆ D−dD +−D+dD +dϵ1dϵ2 +� +∆(ϵ1)∆(ϵ2) +× d† +µ1(ϵ1)dµ2(ϵ2)(f † +µ′fµ1 − xδµ1µ′)(f † +µ2fµ′ − xδµ2µ′) , (B5) +where P is omitted as it commutes with f † +µ2fµ′ and +f † +µ′fµ1. +After a few steps of algebra, the four-fermion +operator � +µ′ f † +µ′fµ1f † +µ2fµ′ can be simplified to +f † +µ2fµ1 + +� +µ′ +f † +µ′fµ′fµ1f † +µ2 = f † +µ2fµ1(1−nf)+nfδµ1µ2 , (B6) +where we have made use of the fact that the Hilbert space +is restricted to ˆNf = nf. Substituting this into Eq. (B5), +we obtain the corrections to parameters g and xg as +dg +dt +���� +p += 4∆(D(t)) +πU +((nf − 1) + 2x) g2 , +(B7) +d(xg) +dt +���� +p += 4∆(D(t)) +πU +� +x2 + nf +� +g2 . +(B8) +Here t is the RG parameter and D(t) = De−t is the +reduced bandwidth after successive t/dt RG steps. +We then calculate the contribution from virtual hole +excitation. +Following the same process as in the last +paragraph, we obtain +− (4g)2 +(πU)2 +dD +D ∆(−D) +� +µ1µ2µ′ +ˆ D−dD +−D+dD +dϵ1dϵ2 +� +∆(ϵ1)∆(ϵ2) +× dµ1(ϵ1)d† +µ2(ϵ2)(f † +µ1fµ′ − xδµ1µ′)P(f † +µ′fµ2 − xδµ2µ′) += (4g)2 +(πU)2 dt∆(−D) +� +µ1µ2µ′ +ˆ D−dD +−D+dD +dϵ1dϵ2 +� +∆(ϵ1)∆(ϵ2) +× d† +µ2(ϵ2)dµ1(ϵ1)(f † +µ1fµ′ − xδµ1µ′)(f † +µ′fµ2 − xδµ2µ′) . +(B9) +In the second equation, we have omitted an energy con- +stant term from the anti-commutator {d† +µ2(ϵ2), dµ1(ϵ1)}. +P is the projector to the restricted Hilbert space, where +ˆNf = nf. +It is omitted in the second equation be- +cause it commutes with f † +µ′fµ2 and f † +µ1fµ′. +The four- +fermion operator � +µ′ f † +µ1fµ′f † +µ′fµ2 can be simplified to +(N − nf + 1)f † +µ1fµ2 as the Hilbert space is restricted to +ˆNf = nf. Then the corrections to g, xg from Eq. (B9) +can be read out as +dg +dt +���� +h += 4∆(−D(t)) +πU +(N − nf + 1 − 2x) g2 , +(B10) +d(xg) +dt +���� +h += 4∆(−D(t)) +πU +� +−x2� +g2 . +(B11) + +17 +Adding up the particle and the hole contributions we +can obtain the RG equations for g and (xg). The Kondo +energy scale TK can be estimated as the reduced band- +width D(t) where g diverges. For a constant ∆(ω) = ∆0, +we obtain +dg +dt = 4∆0 +πU Ng2, +d(xg) +dt += 4∆0 +πU nfg2 +(B12) +and the solution +g(t) = +g(0) +1 − g(0) 4∆0 +πU N · t , +(B13) +x(t) = x(0)g(0) +g(t) + nf +N · g(t) − g(0) +g(t) +. +(B14) +where g(0) is the initial condition given in Eq. (B3). g(t) +diverges at tK = +πU +4N g(0)∆0 , corresponding the Kondo en- +ergy scale De−tK = De− +πU +4N g(0)∆0 . +As g(t) diverges as +t → tK, the second term in x(t) dominates and there +must be x → nf +N . In other words, x flows to the occupa- +tion fraction. +2. +Application to the symmetric state at CNP +We assume a symmetric state of the THF model at +CNP and examine its Kondo energy scale. Following the +calculations in Appendix A 3, we obtain the hybridiza- +tion function contributed by the fully symmetric c-bands +(Fig. 1(b)) +∆(ω) = Ω0 +4v2⋆ +� ����|ω| − M +2 +���� θ(|ω| − M) +� +γ2 + v′2 +⋆ k2 +F 1 +� +sin2 θkF 1 +2 +× e−k2 +F 1λ2 + +����|ω| + M +2 +���� +� +γ2 + v′2 +⋆ k2 +F 2 +� +cos2 θkF 2 +2 +e−k2 +F 2λ2� +, +(B15) +where +kF 1 = 1 +v⋆ +� +(|ω| − M/2)2 − (M/2)2, +(B16) +kF 2 = 1 +v⋆ +� +(|ω| + M/2)2 − (M/2)2, +(B17) +θk = arccos +M/2 +� +M 2/4 + v2⋆k2 . +(B18) +We should choose the initial cutoff D += +1 +2U1 be- +yond which the Schrieffer-Wolff transformation is invalid. +For these states kF 1,2 ≲ +U1 +2v⋆ and hence v′2 +⋆ k2 +F 1,2 ≲ +119.4meV2, which is significantly smaller than γ2 ≈ +612.6meV2. The damping factors e−λ2kF 1,22 ≳ 0.74 are +also large. Thus, in the following we approximate ∆(ω) +(|ω| < U1/2) as +∆(ω) ≈ Ω0 +4v2⋆ +� ����|ω| − M +2 +���� θ(|ω| − M)γ2 sin2 θkF 1 +2 ++ +����|ω| + M +2 +���� γ2 cos2 θkF 2 +2 +� +. +(B19) +We first consider the flat-band limit (M = 0), where +the parameter θk is always π +2 . Thus, we have +∆(ω) = b|ω|, +b = Ω0 +4v2⋆ +γ2 ≈ 0.1290 . +(B20) +We also assume that there is no multiplet splitting in the +symmetric state such that the effective Anderson model +should be a U(8) theory with nf = 4. Naively applying +the RG equations derived in the last subsection gives +d�g +dt = −�g + 4bD +πU1 +N �g2, +(B21) +where N = 8, D = U1/2, �g = ge−t. Due to the particle- +hole symmetry at CNP, the initial condition (Eq. (B3)) +is �g(0) = 1. It seems that there would be an unstable +fixed point �g∗ = +2π +4N b, the initial �g below (above) which +flows to zero (infinity). Using the actual parameters we +find g∗ ≈ 1.52, hence the system would not be in the +Kondo phase. This result differs from the standard case +with a constant ∆(ω), where a positive g always flows +to infinity. Furthermore, a more careful RG analysis [98] +shows that the fixed point �g∗ does not really exist. It is +a false result of the weak coupling expansion, which fails +for ∆(ω) ∼ |ω|r with r > 1 +2. Thus, a ∆(ω) ∼ |ω| bath +does not have a strong coupling phase. This conclusion +is also consistent with numerical studies [95–97]. +We then consider the case with M ̸= 0. We use the +value M = 3.697meV. The RG process can be divided +into two stages: (i) When D(t) = 1 +2U1e−t > M, there is +approximately ∆(D(t)) ≈ bD(t). (ii) When D(t) < M, +the first line of Eq. (B19) vanishes, and cos2 θkF 2 +2 +in the +second line equals to +M +4ω+2M + 1 +2. Then there is ∆(D(t)) ≈ +∆0(1 + D(t)/M), with ∆0 = Ω0Mγ2 +8v2⋆ +≈ 0.239meV. The +boundary between the two stages is t1 = ln U1 +2M . The RG +equation in the first stage reads +dg +dt = 2Nb +π +g2e−t ⇒ g(t) = +1 +1 − 2N b +π (1 − e−t) . +(B22) +Due to the particle-hole symmetry, the initial condition +given by Eq. (B3) is g(0) = 1. We have g1 = g(t1) ≈ 2.34. +The RG equation in the second stage is given by +dg +dt =4N∆0 +πU1 +g2 + 4N∆0 +πU1 +g2 · e−(t−t1) +⇒ g(t) = +1 +g−1 +1 +− 4N ∆0 +πU1 (t − t1) − 4N ∆0 +πU1 (1 − e−(t−t1)) . +(B23) + +18 +g(t) diverges at tK − t1 ≈ +πU1 +4g1N ∆0 − y with y = 1, corre- +sponding to the Kondo energy scale +kBTK = Mey · e− +πU1 +4g1N ∆0 ≈ 3.8 × 10−4meV. +(B24) +3. +Application to the effective model for ν > 0 +states +In the absence of the Hund’s coupling JH, we can re- +gard (α, s) as a composite index so that ˆHSI (Eq. (20)) +is a U(N) theory with N = 4. Then the flow equations +in Appendix B 1 apply. For simplicity, we omit the neg- +ative branch of ∆(ω) (Eq. (19)) at ω < −µc because +it is far away from the Fermi level for ν > 0 (Fig. 1(d)). +The positive branch of ∆(ω) can be well approximated by +∆(ω) = ∆(0)(1+ω/(µc−G)) for |ω| < µc−G (Fig. 1(d)). +We choose the initial cutoff D to be the minimum value +of µc − G and ∆E±. +Substituting this ∆(ω) into the +general RG equations in Appendix B 1, we obtain +dg +dt = 4∆(0) +πU1 Ng2 + +4∆(0)D +πU1(µc − G)(4x + 2nf − 2 − N)g2e−t +(B25) +and +d(xg) +dt += 4∆(0) +πU1 nfg2 + +4∆(0)D +πU1(µc − G)(2x2 + nf)g2e−t (B26) +The O(e−t) terms will eventually become irrelevant when +t is sufficiently large. +After the O(e−t) terms become +irrelevant, we have +d(xg) +dg += nf/N, implying x → +nf +N +at the divergence of g. We then approximate the flow +equation of g by setting x to its fixed point value +nf +N , +i.e., +dg +dt ≈ 4∆(0) +πU1 Ng2 + +4∆(0)D +πU1(µc − G)(3nf − 6)g2e−t +(B27) +The solution of g is +g(t) ≈ +1 +g−1(0) − 4∆(0) +πU1 N +� +t + ynf (1 − e−t) +� , +(B28) +where ynf = +D +µc−G(3nf − 6). The Kondo energy scale +is determined t = tK at which g diverges. +Assuming +tK ≫ 1, we have +tK ≈ +πU1 +4Ng(0)∆(0) − ynf +(B29) +and hence +kBTK ≈ D · eynf · e− +πU1 +4N g(0)∆(0) . +(B30) +a. +The nf = 1, 3 cases +In the presence of the Hund’s coupling, we have to +examine the derivations in Appendix B 1 carefully. The +most important effect of ˆHH is to change the local Hilbert +space at small energy scales. In general, JH leads to a +multiplet splitting. When the RG energy scale is smaller +than the splitting, the higher energy multiplet will be- +come inaccessible, and the local Hilbert space is effec- +tively reduced. A minor effect is that the charge gaps +∆E± will depend on JH and the resulted coupling be- +tween f-spin and d-spin in the Coqblin–Schrieffer model +will break the U(N) symmetry. +In the following, we study how ˆHH changes the RG +equations. We first consider the nf = 1 case. In the vir- +tual particle excitation process (Eq. (B5)), the interme- +diate f-multiplet is given by |F ′⟩ = (f † +µ2fµ′ − δµ2µ′)|F⟩, +where F is the initial f-multiplet. (µ should be regarded +as the composite index (α, s).) As |F ′⟩ has the same par- +ticle number as |F⟩, it must be one of the four states with +(n1↑, n1↓; n2↑, n2↓) = (10;00), (01;00), (00;10), (00;01). +All of the possible intermediate states do not feel the +Hund’s coupling (JH +� +α nα↑nα↓) and hence they have +the same energy as |F⟩. +Hence, the excitation energy +of the intermediate state is purely contributed by d- +electrons. Then all the following derivations apply. The +same argument applies to the virtual hole excitation +(Eq. (B9)). Therefore, the RG equations for nf = 1 will +not be affected by JH. For the same reason, RG equa- +tions for nf = 3 will also not be affected by JH, where +the initial and intermediate states are single-hole states +that do not feel JH. The TK for nf = 1, 3 is given by +Eq. (B30). +b. +The nf = 2 case +The Hilbert space with two particles has six states: +(n1↑, n1↓; n2↑, n2↓) = (10;10), (10;01), (01;10), (01;01), +(11;00), (00;11). The former four states have the energy +−2µf + U1, and the latter two states have the energy +−2µf + U1 + JH. Thus JH leads to a multiplet splitting. +We divide the RG into two stages. In the first stage D(t) +is larger than JH, then the splitting JH only plays a +minor role and can be neglected. Thus the RG equations +in the first stage are given by Eq. (B27). The first stage +ends at t1 = ln(D/JH). If g diverges before t reaches +t1, the Kondo energy scale should be given by Eq. (B30) +with y2 = 0, i.e., +kBT ′ +K = D · e− +πU1 +4N g(0)∆(0) . +(B31) +If g is still finite at t1 +g1 = +g(0) +1 − g(0) 4∆(0) +πU1 N ln D +JH +, +(B32) +then the RG goes into the second stage. +The effective cutoff and the initial g of the second stage +are JH and g1, respectively. We first examine the virtual +particle excitation process (Eq. (B5)), where the interme- +diate f-multiplet is given by |F ′⟩ = (f † +µ2fµ′ − δµ2µ′)|F⟩. + +19 +Here F is the initial f-multiplet. +µ′, µ2 should be re- +garded as the composite indices (α′, s′), (α2, s2), respec- +tively. Suppose |F⟩ is one of the four low energy states, +where each orbital (α = 1, 2) has one electron; then, for +|F ′⟩ to be a low energy state, the index µ′ must have the +same orbital index with µ2, i.e., α′ = α2, such that each +orbital (α = 1, 2) in |F ′⟩ still has one electron. +With +this restriction, the four-fermion operator in Eq. (B6) +becomes +f † +α2s2fα1s1 + +� +s′ +f † +α2s′fα2s′fα1s1f † +α2s2 +(B33) +� +s′ f † +α2s′fα2s′ acting on the bra state (final state) gives +nf +α2, which must equal to 1 given that the bra state is one +of the four low energy states. Thus the four-fermion op- +erator equals to δα2α1δs2s1. The resulting contributions +to the RG equation are +dg +dt +���� +p += 4∆(D(t)) +πU +(2x) g2 , +(B34) +d(xg) +dt +���� +p += 4∆(D(t)) +πU +� +x2 + 1 +� +g2 . +(B35) +We second examine the virtual hole excitation process +(Eq. (B9)), where the intermediate f-multiplet is given +by |F ′⟩ = (f † +µ′fµ2 − δµ′µ2)|F⟩. Suppose |F⟩ is one of the +four low energy states; then, for |F ′⟩ to be in the low +energy state, the index µ′ must have the same orbital +index with µ2, i.e., α′ = α2. With this restriction, the +four-fermion operator in Eq. (B9) can be written as +� +s′ +f † +α1s1fα2s′f † +α2s′fα2s2 . +(B36) +If |F⟩ is one of the four low energy states, it at most +occupies one electron in the α2 orbital. The α2 orbital +of fα2s2|F⟩ must be empty, implying � +s′ fα2s′f † +α2s′ = 2. +Thus the four-fermion operator equals 2f † +α1s1fα2s2. The +resulting contributions to the RG equation are +dg +dt +���� +h += 4∆(D(t)) +πU +(2 − 2x) g2 , +(B37) +d(xg) +dt +���� +h += 4∆(D(t)) +πU +� +−x2� +g2 . +(B38) +Eqs. (B34), (B35), (B37) and (B38) are identical to equa- +tions of the U(2) case where N = 2, nf = 1. Following +the steps of deriving Eq. (B30), we find x still flows to 1 +2, +and +kBT ′′ +K ≈ JH · e− +πU1 +8g1∆(0) . +(B39) +The final expression for the Kondo energy scale at nf = +2 is +kBTK = +� +kBT ′ +K, +kBTK > JH +kBT ′′ +K, +otherwise +. +(B40) + +20 +[1] Yuan Cao, Valla Fatemi, Shiang Fang, Kenji Watan- +abe, Takashi Taniguchi, Efthimios Kaxiras, and Pablo +Jarillo-Herrero, “Unconventional superconductivity in +magic-angle graphene superlattices,” Nature 556, 43– +50 (2018). +[2] Yuan Cao, Valla Fatemi, Ahmet Demir, Shiang Fang, +Spencer L. Tomarken, Jason Y. Luo, Javier D. Sanchez- +Yamagishi, +Kenji +Watanabe, +Takashi +Taniguchi, +Efthimios Kaxiras, Ray C. Ashoori, and Pablo Jarillo- +Herrero, “Correlated insulator behaviour at half-filling +in magic-angle graphene superlattices,” Nature 556, 80– +84 (2018). +[3] Rafi Bistritzer and Allan H. MacDonald, “Moir´e bands +in twisted double-layer graphene,” Proceedings of the +National Academy of Sciences 108, 12233–12237 (2011). +[4] Xiaobo Lu, Petr Stepanov, Wei Yang, Ming Xie, Mo- +hammed Ali Aamir, Ipsita Das, Carles Urgell, Kenji +Watanabe, Takashi Taniguchi, Guangyu Zhang, Adrian +Bachtold, Allan H. MacDonald, +and Dmitri K. Efe- +tov, “Superconductors, orbital magnets and correlated +states in magic-angle bilayer graphene,” Nature 574, +653–657 (2019), see Fig. 1(c) and Extended Data Fig. +5 for gapped insulator at ν = 1, inferred by the low +temperature resistivity behaviour. +[5] Matthew Yankowitz, Shaowen Chen, Hryhoriy Polshyn, +Yuxuan Zhang, K. Watanabe, T. Taniguchi, David +Graf, Andrea F. Young, +and Cory R. Dean, “Tuning +superconductivity in twisted bilayer graphene,” Science +363, 1059–1064 (2019). +[6] Yu Saito, +Jingyuan Ge, +Kenji Watanabe, +Takashi +Taniguchi, and Andrea F. Young, “Independent super- +conductors and correlated insulators in twisted bilayer +graphene,” Nature Physics 16, 926–930 (2020), see Fig. +1 (a) and Extended Data Fig. 4 (a) for the low tem- +perature resistivity exhibiting insulator behavior near +ν = 1. +[7] Petr Stepanov, Ipsita Das, Xiaobo Lu, Ali Fahimniya, +Kenji Watanabe, +Takashi Taniguchi, +Frank H. L. +Koppens, Johannes Lischner, Leonid Levitov, +and +Dmitri K. Efetov, “Untying the insulating and super- +conducting orders in magic-angle graphene,” Nature +583, 375–378 (2020), see Fig. 2(a) D3 for insulator at +ν = 1. +[8] Xiaoxue Liu, Zhi Wang, K. Watanabe, T. Taniguchi, +Oskar Vafek, +and J. I. A. Li, “Tuning electron cor- +relation in magic-angle twisted bilayer graphene using +Coulomb screening,” Science 371, 1261–1265 (2021). +[9] Harpreet Singh Arora, Robert Polski, Yiran Zhang, +Alex Thomson, Youngjoon Choi, Hyunjin Kim, Zhong +Lin, +Ilham Zaky Wilson, +Xiaodong Xu, +Jiun-Haw +Chu, Kenji Watanabe, Takashi Taniguchi, Jason Alicea, +and Stevan Nadj-Perge, “Superconductivity in metallic +twisted bilayer graphene stabilized by WSe2,” Nature +583, 379–384 (2020), number: 7816 Publisher: Nature +Publishing Group. +[10] Yuan Cao, Daniel Rodan-Legrain, Jeong Min Park, +Noah F. Q. Yuan, Kenji Watanabe, Takashi Taniguchi, +Rafael M. Fernandes, Liang Fu, +and Pablo Jarillo- +Herrero, “Nematicity and competing orders in super- +conducting magic-angle graphene,” Science 372, 264– +271 (2021), publisher: American Association for the Ad- +vancement of Science. +[11] Myungchul Oh, +Kevin P. Nuckolls, +Dillon Wong, +Ryan L. Lee, Xiaomeng Liu, Kenji Watanabe, Takashi +Taniguchi, +and Ali Yazdani, “Evidence for unconven- +tional superconductivity in twisted bilayer graphene,” +Nature 600, 240–245 (2021), see Fig. 1(c) for the the +zero-energy peak, where the conduction (valence) bands +are found to be pinned to the Fermi energy at negative +(positive) fillings. +[12] Yonglong Xie, Biao Lian, Berthold J¨ack, Xiaomeng Liu, +Cheng-Li Chiu, Kenji Watanabe, Takashi Taniguchi, +B. Andrei Bernevig, +and Ali Yazdani, “Spectroscopic +signatures of many-body correlations in magic-angle +twisted bilayer graphene,” Nature 572, 101–105 (2019). +[13] Aaron L. Sharpe, Eli J. Fox, Arthur W. Barnard, Joe +Finney, Kenji Watanabe, Takashi Taniguchi, M. A. +Kastner, and David Goldhaber-Gordon, “Emergent fer- +romagnetism near three-quarters filling in twisted bi- +layer graphene,” Science 365, 605–608 (2019). +[14] Youngjoon Choi, Jeannette Kemmer, Yang Peng, Alex +Thomson, Harpreet Arora, Robert Polski, Yiran Zhang, +Hechen Ren, Jason Alicea, Gil Refael, Felix von Op- +pen, Kenji Watanabe, Takashi Taniguchi, +and Stevan +Nadj-Perge, “Electronic correlations in twisted bilayer +graphene near the magic angle,” Nature Physics 15, +1174–1180 (2019). +[15] Alexander Kerelsky, Leo J. McGilly, Dante M. Kennes, +Lede +Xian, +Matthew +Yankowitz, +Shaowen +Chen, +K. Watanabe, T. Taniguchi, James Hone, Cory Dean, +Angel Rubio, +and Abhay N. Pasupathy, “Maximized +electron interactions at the magic angle in twisted bi- +layer graphene,” Nature 572, 95–100 (2019). +[16] Yuhang Jiang, Xinyuan Lai, Kenji Watanabe, Takashi +Taniguchi, Kristjan Haule, Jinhai Mao, +and Eva Y. +Andrei, “Charge order and broken rotational symmetry +in magic-angle twisted bilayer graphene,” Nature 573, +91–95 (2019). +[17] Dillon Wong, Kevin P. Nuckolls, Myungchul Oh, Biao +Lian, Yonglong Xie, Sangjun Jeon, Kenji Watanabe, +Takashi Taniguchi, B. Andrei Bernevig, +and Ali Yaz- +dani, “Cascade of electronic transitions in magic-angle +twisted bilayer graphene,” Nature 582, 198–202 (2020). +[18] U. Zondiner, A. Rozen, D. Rodan-Legrain, Y. Cao, +R. Queiroz, T. Taniguchi, K. Watanabe, Y. Oreg, F. von +Oppen, Ady Stern, E. Berg, P. Jarillo-Herrero, +and +S. Ilani, “Cascade of phase transitions and Dirac re- +vivals in magic-angle graphene,” Nature 582, 203–208 +(2020). +[19] Youngjoon Choi, Hyunjin Kim, Cyprian Lewandowski, +Yang Peng, Alex Thomson, Robert Polski, Yiran Zhang, +Kenji Watanabe, Takashi Taniguchi, Jason Alicea, and +Stevan Nadj-Perge, “Interaction-driven band flattening +and correlated phases in twisted bilayer graphene,” Na- +ture Physics 17, 1375–1381 (2021), see Fig. 4(a,f-i) for +the evolutions of the zero-energy peak (around ν = −1) +and Hubbard bands as the temperature is increased +from 400mK to 20K. +[20] M. Serlin, C. L. Tschirhart, H. Polshyn, Y. Zhang, +J. Zhu, K. Watanabe, T. Taniguchi, L. Balents, +and +A. F. Young, “Intrinsic quantized anomalous Hall ef- +fect in a moir´e heterostructure,” Science 367, 900–903 + +21 +(2020). +[21] Kevin P. Nuckolls, Myungchul Oh, Dillon Wong, Biao +Lian, Kenji Watanabe, Takashi Taniguchi, B. Andrei +Bernevig, and Ali Yazdani, “Strongly correlated Chern +insulators in magic-angle twisted bilayer graphene,” Na- +ture 588, 610–615 (2020), see Fig 1.c for the zero- energy +peak from ν = ±1 to ±2 at T = 200mK and B⊥ = 1T. +[22] Youngjoon Choi, +Hyunjin Kim, +Yang Peng, +Alex +Thomson, Cyprian Lewandowski, Robert Polski, Yi- +ran Zhang, Harpreet Singh Arora, Kenji Watanabe, +Takashi Taniguchi, Jason Alicea, +and Stevan Nadj- +Perge, “Correlation-driven topological phases in magic- +angle twisted bilayer graphene,” Nature 589, 536–541 +(2021), see Fig. 1(a) for zero-energy peak around ν = −1 +and Hubbard band at T = 2K. +[23] Yu Saito, Jingyuan Ge, Louk Rademaker, Kenji Watan- +abe, Takashi Taniguchi, Dmitry A. Abanin, +and An- +drea F. Young, “Hofstadter subband ferromagnetism +and symmetry-broken Chern insulators in twisted bi- +layer graphene,” Nature Physics 17, 478–481 (2021). +[24] Ipsita Das, Xiaobo Lu, Jonah Herzog-Arbeitman, Zhi- +Da Song, Kenji Watanabe, Takashi Taniguchi, B. An- +drei Bernevig, +and Dmitri K. Efetov, “Symmetry- +broken Chern insulators and Rashba-like Landau-level +crossings in magic-angle bilayer graphene,” Nature +Physics 17, 710–714 (2021). +[25] Shuang +Wu, +Zhenyuan +Zhang, +K. +Watanabe, +T. Taniguchi, +and Eva Y. Andrei, “Chern insula- +tors, van Hove singularities and topological flat bands +in +magic-angle +twisted +bilayer +graphene,” +Nature +Materials 20, 488–494 (2021), fig. 1(b), near ν = ±1 a +resistivity peak emerges at around T = 11K. +[26] Jeong Min Park, Yuan Cao, Kenji Watanabe, Takashi +Taniguchi, and Pablo Jarillo-Herrero, “Flavour Hund’s +coupling, Chern gaps and charge diffusivity in moir´e +graphene,” Nature 592, 43–48 (2021), fig. 4(a), near +ν = ±1, resistivity peaks occur and above around 10K. +[27] Hryhoriy Polshyn, Matthew Yankowitz, Shaowen Chen, +Yuxuan Zhang, K. Watanabe, T. Taniguchi, Cory R. +Dean, +and +Andrea +F. +Young, +“Large +linear-in- +temperature resistivity in twisted bilayer graphene,” +Nature Physics , 1–6 (2019), in fig. 1 (a)(b), a peak +rises as T increases at ν = −1. +[28] Yuan +Cao, +Debanjan +Chowdhury, +Daniel +Rodan- +Legrain, +Oriol +Rubies-Bigorda, +Kenji +Watanabe, +Takashi Taniguchi, T. Senthil, +and Pablo Jarillo- +Herrero, “Strange Metal in Magic-Angle Graphene with +near Planckian Dissipation,” Physical Review Letters +124, 076801 (2020), publisher: American Physical So- +ciety. +[29] Alexandre Jaoui, Ipsita Das, Giorgio Di Battista, Jaime +D´ıez-M´erida, Xiaobo Lu, Kenji Watanabe, Takashi +Taniguchi, Hiroaki Ishizuka, Leonid Levitov, +and +Dmitri K. Efetov, “Quantum critical behaviour in +magic-angle twisted bilayer graphene,” Nature Physics +18, 633–638 (2022), in Fig. 1 (a), near ν = ±1 peaks +occurs when temperature rises. Notably, the peak near +ν = 1 is rather sharp in contrast to other experiments +mentioned, where usually the peak near ν = −1 is rec- +ognized more easily. +[30] Asaf Rozen, Jeong Min Park, Uri Zondiner, Yuan +Cao, Daniel Rodan-Legrain, Takashi Taniguchi, Kenji +Watanabe, Yuval Oreg, Ady Stern, Erez Berg, Pablo +Jarillo-Herrero, +and Shahal Ilani, “Entropic evidence +for a Pomeranchuk effect in magic-angle graphene,” Na- +ture 592, 214–219 (2021), see Fig. 2(e) for entropies at +different fillings and temperatures. +[31] Yu Saito, Fangyuan Yang, Jingyuan Ge, Xiaoxue Liu, +Takashi Taniguchi, Kenji Watanabe, J. I. A. Li, Erez +Berg, and Andrea F. Young, “Isospin Pomeranchuk ef- +fect in twisted bilayer graphene,” Nature 592, 220–224 +(2021), see Fig. 1(b) and Extended Data Fig. 2 (a-f) for +resistivity peaks occur at around 10K near ν=-1. +[32] Zhida Song, Zhijun Wang, Wujun Shi, Gang Li, Chen +Fang, +and B. Andrei Bernevig, “All Magic Angles in +Twisted Bilayer Graphene are Topological,” Physical +Review Letters 123, 036401 (2019). +[33] Hoi Chun Po, Liujun Zou, T. Senthil, +and Ashvin +Vishwanath, “Faithful tight-binding models and frag- +ile topology of magic-angle bilayer graphene,” Physical +Review B 99, 195455 (2019). +[34] Junyeong Ahn, Sungjoon Park, and Bohm-Jung Yang, +“Failure of Nielsen-Ninomiya Theorem and Fragile +Topology in Two-Dimensional Systems with Space- +Time Inversion Symmetry: Application to Twisted Bi- +layer Graphene at Magic Angle,” Physical Review X 9, +021013 (2019). +[35] Grigory Tarnopolsky, Alex Jura Kruchkov, and Ashvin +Vishwanath, “Origin of Magic Angles in Twisted Bi- +layer Graphene,” Physical Review Letters 122, 106405 +(2019). +[36] Jianpeng Liu, Junwei Liu, +and Xi Dai, “The pseudo- +Landau-level representation of twisted bilayer graphene: +band topology and the implications on the correlated +insulating phase,” Physical Review B 99, 155415 (2019), +arXiv: 1810.03103. +[37] Zhi-Da Song, Biao Lian, Nicolas Regnault, and B. An- +drei Bernevig, “Twisted bilayer graphene. II. Stable +symmetry anomaly,” Physical Review B 103, 205412 +(2021), publisher: American Physical Society. +[38] Jian Kang and Oskar Vafek, “Strong Coupling Phases +of Partially Filled Twisted Bilayer Graphene Narrow +Bands,” Phys. Rev. Lett. 122, 246401 (2019), publisher: +American Physical Society. +[39] Nick Bultinck, Eslam Khalaf, Shang Liu, Shubhayu +Chatterjee, Ashvin Vishwanath, and Michael P. Zaletel, +“Ground State and Hidden Symmetry of Magic-Angle +Graphene at Even Integer Filling,” Physical Review X +10, 031034 (2020), publisher: American Physical Soci- +ety. +[40] Kangjun Seo, Valeri N. Kotov, and Bruno Uchoa, “Fer- +romagnetic Mott state in Twisted Graphene Bilayers at +the Magic Angle,” Phys. Rev. Lett. 122, 246402 (2019), +publisher: American Physical Society. +[41] B. Andrei Bernevig, Zhi-Da Song, Nicolas Regnault, +and Biao Lian, “Twisted bilayer graphene. III. Interact- +ing Hamiltonian and exact symmetries,” Physical Re- +view B 103, 205413 (2021), publisher: American Phys- +ical Society. +[42] Oskar Vafek and Jian Kang, “Renormalization Group +Study +of +Hidden +Symmetry +in +Twisted +Bilayer +Graphene with Coulomb Interactions,” Physical Review +Letters 125, 257602 (2020), publisher: American Phys- +ical Society. +[43] Biao Lian, Zhi-Da Song, Nicolas Regnault, Dmitri K. +Efetov, Ali Yazdani, and B. Andrei Bernevig, “Twisted +bilayer graphene. IV. Exact insulator ground states and +phase diagram,” Physical Review B 103, 205414 (2021), + +22 +publisher: American Physical Society. +[44] Fang Xie, Aditya Cowsik, Zhi-Da Song, Biao Lian, +B. Andrei Bernevig, +and Nicolas Regnault, “Twisted +bilayer graphene. VI. An exact diagonalization study at +nonzero integer filling,” Physical Review B 103, 205416 +(2021), publisher: American Physical Society. +[45] Ming Xie and A. H. MacDonald, “Nature of the Cor- +related Insulator States in Twisted Bilayer Graphene,” +Phys. Rev. Lett. 124, 097601 (2020), publisher: Amer- +ican Physical Society. +[46] Jianpeng Liu and Xi Dai, “Theories for the correlated +insulating states and quantum anomalous Hall effect +phenomena in twisted bilayer graphene,” Phys. Rev. B +103, 035427 (2021), publisher: American Physical So- +ciety. +[47] Tommaso Cea and Francisco Guinea, “Band structure +and insulating states driven by Coulomb interaction +in twisted bilayer graphene,” Physical Review B 102, +045107 (2020), publisher: American Physical Society. +[48] J¨orn W. F. Venderbos and Rafael M. Fernandes, “Corre- +lations and electronic order in a two-orbital honeycomb +lattice model for twisted bilayer graphene,” Physical Re- +view B 98, 245103 (2018), publisher: American Physical +Society. +[49] Masayuki Ochi, Mikito Koshino, and Kazuhiko Kuroki, +“Possible correlated insulating states in magic-angle +twisted bilayer graphene under strongly competing in- +teractions,” Phys. Rev. B 98, 081102 (2018), publisher: +American Physical Society. +[50] Yi Zhang, Kun Jiang, Ziqiang Wang, +and Fuchun +Zhang, “Correlated insulating phases of twisted bilayer +graphene at commensurate filling fractions: A Hartree- +Fock study,” Physical Review B 102, 035136 (2020), +publisher: American Physical Society. +[51] Shang Liu, Eslam Khalaf, Jong Yeon Lee, and Ashvin +Vishwanath, “Nematic topological semimetal and insu- +lator in magic-angle bilayer graphene at charge neutral- +ity,” Physical Review Research 3, 013033 (2021), pub- +lisher: American Physical Society. +[52] Yuan Da Liao, Jian Kang, Clara N. Breiø, Xiao Yan +Xu, Han-Qing Wu, Brian M. Andersen, Rafael M. Fer- +nandes, +and Zi Yang Meng, “Correlation-Induced In- +sulating Topological Phases at Charge Neutrality in +Twisted Bilayer Graphene,” Physical Review X 11, +011014 (2021), publisher: American Physical Society. +[53] Jian Kang and Oskar Vafek, “Non-Abelian Dirac node +braiding and near-degeneracy of correlated phases +at odd integer filling in magic-angle twisted bilayer +graphene,” Physical Review B 102, 035161 (2020), pub- +lisher: American Physical Society. +[54] Tomohiro Soejima, Daniel E. Parker, Nick Bultinck, Jo- +hannes Hauschild, +and Michael P. Zaletel, “Efficient +simulation of moir´e materials using the density matrix +renormalization group,” Physical Review B 102, 205111 +(2020), publisher: American Physical Society. +[55] Kasra Hejazi, Xiao Chen, +and Leon Balents, “Hybrid +Wannier Chern bands in magic angle twisted bilayer +graphene and the quantized anomalous Hall effect,” +Physical Review Research 3, 013242 (2021), publisher: +American Physical Society. +[56] Yuan Da Liao, Zi Yang Meng, and Xiao Yan Xu, “Va- +lence bond orders at charge neutrality in a possible +two-orbital extended hubbard model for twisted bilayer +graphene,” Phys. Rev. Lett. 123, 157601 (2019). +[57] Dante M. Kennes, Johannes Lischner, +and Christoph +Karrasch, “Strong correlations and d + id superconduc- +tivity in twisted bilayer graphene,” Physical Review B +98, 241407 (2018), publisher: American Physical Soci- +ety. +[58] Laura Classen, Carsten Honerkamp, +and Michael M. +Scherer, “Competing phases of interacting electrons on +triangular lattices in moir´e heterostructures,” Physical +Review B 99, 195120 (2019), publisher: American Phys- +ical Society. +[59] B. Andrei Bernevig, Biao Lian, Aditya Cowsik, Fang +Xie, Nicolas Regnault, and Zhi-Da Song, “Twisted bi- +layer graphene. V. Exact analytic many-body excita- +tions in Coulomb Hamiltonians: Charge gap, Goldstone +modes, and absence of Cooper pairing,” Physical Re- +view B 103, 205415 (2021), publisher: American Phys- +ical Society. +[60] Eslam Khalaf, Nick Bultinck, Ashvin Vishwanath, and +Michael P Zaletel, “Soft modes in magic angle twisted +bilayer graphene,” arXiv preprint arXiv:2009.14827 +(2020). +[61] Zhi-Da Song and B. Andrei Bernevig, “Magic-Angle +Twisted Bilayer Graphene as a Topological Heavy +Fermion +Problem,” +Physical +Review +Letters +129, +047601 (2022). +[62] Hao Shi and Xi Dai, “Heavy-fermion representation for +twisted bilayer graphene systems,” Physical Review B +106, 245129 (2022), publisher: American Physical So- +ciety. +[63] Ph Nozi`eres and A. Blandin, “Kondo effect in real met- +als,” Journal de Physique 41, 193–211 (1980), publisher: +Soci´et´e Fran¸caise de Physique. +[64] A.M. Tsvelick and P.B. Wiegmann, “Exact results in +the theory of magnetic alloys,” Advances in Physics 32, +453–713 (1983), publisher: +Taylor & Francis +eprint: +https://doi.org/10.1080/00018738300101581. +[65] Piers Coleman, “New approach to the mixed-valence +problem,” Physical Review B 29, 3035–3044 (1984), +publisher: American Physical Society. +[66] N. E. Bickers, “Review of techniques in the large-$N$ +expansion for dilute magnetic alloys,” Reviews of Mod- +ern Physics 59, 845–939 (1987). +[67] Ian Affleck and Andreas W. W. Ludwig, “Critical theory +of overscreened Kondo fixed points,” Nuclear Physics B +360, 641–696 (1991). +[68] Ian Affleck and Andreas W. W. Ludwig, “The Kondo +effect, conformal field theory and fusion rules,” Nuclear +Physics B 352, 849–862 (1991). +[69] V. J. Emery and S. Kivelson, “Mapping of the two- +channel Kondo problem to a resonant-level model,” +Physical Review B 46, 10812–10817 (1992), publisher: +American Physical Society. +[70] A. M. Tsvelik and M. Reizer, “Phenomenological the- +ory of non-Fermi-liquid heavy-fermion alloys,” Physical +Review B 48, 9887–9889 (1993), publisher: American +Physical Society. +[71] Akira Furusaki and Naoto Nagaosa, “Kondo effect in +a Tomonaga-Luttinger liquid,” Physical Review Letters +72, 892–895 (1994), publisher: American Physical Soci- +ety. +[72] D. L. Cox and A. Zawadowski, “Exotic Kondo effects +in metals: Magnetic ions in a crystalline electric field +and tunnelling centres,” Advances in Physics 47 (1998), +10.1080/000187398243500. + +23 +[73] Olivier Parcollet, Antoine Georges, Gabriel Kotliar, +and Anirvan Sengupta, “Overscreened multichannel +$\mathrm{SU}(N)$ Kondo model: Large-$N$ solution +and conformal field theory,” Physical Review B 58, +3794–3813 (1998). +[74] Piers Coleman and Andrew J Schofield, “Quantum crit- +icality,” Nature 433, 226–229 (2005). +[75] G. Kotliar, S. Y. Savrasov, K. Haule, V. S. Oudovenko, +O. Parcollet, and C. A. Marianetti, “Electronic struc- +ture calculations with dynamical mean-field theory,” +Reviews of Modern Physics 78, 865–951 (2006), pub- +lisher: American Physical Society. +[76] Philipp Werner, Armin Comanac, Luca de’ Medici, +Matthias Troyer, +and Andrew J. Millis, “Continuous- +Time Solver for Quantum Impurity Models,” Physical +Review Letters 97, 076405 (2006), publisher: American +Physical Society. +[77] Philipp Gegenwart, Qimiao Si, +and Frank Steglich, +“Quantum criticality in heavy-fermion metals,” Nature +Physics 4, 186–197 (2008). +[78] Qimiao +Si +and +Frank +Steglich, +“Heavy +fermions +and +quantum +phase +transi- +tions,” +Science +329, +1161–1166 +(2010), +https://www.science.org/doi/pdf/10.1126/science.1191195. +[79] Maxim Dzero, Kai Sun, Victor Galitski, and Piers Cole- +man, “Topological Kondo Insulators,” Physical Review +Letters 104, 106408 (2010), publisher: American Phys- +ical Society. +[80] Feng Lu, JianZhou Zhao, Hongming Weng, Zhong Fang, +and Xi Dai, “Correlated Topological Insulators with +Mixed Valence,” Physical Review Letters 110, 096401 +(2013), publisher: American Physical Society. +[81] H. R. Krishna-murthy, J. W. Wilkins, and K. G. Wil- +son, “Renormalization-group approach to the Anderson +model of dilute magnetic alloys. I. Static properties for +the symmetric case,” Physical Review B 21, 1003–1043 +(1980), publisher: American Physical Society. +[82] H. R. Krishna-murthy, J. W. Wilkins, and K. G. Wil- +son, “Renormalization-group approach to the Anderson +model of dilute magnetic alloys. II. Static properties for +the asymmetric case,” Physical Review B 21, 1044–1083 +(1980), publisher: American Physical Society. +[83] Ralf Bulla, Theo A. Costi, and Thomas Pruschke, “Nu- +merical renormalization group method for quantum im- +purity systems,” Reviews of Modern Physics 80, 395– +450 (2008), publisher: American Physical Society. +[84] Jian Kang and Oskar Vafek, “Symmetry, Maximally Lo- +calized Wannier States, and a Low-Energy Model for +Twisted Bilayer Graphene Narrow Bands,” Phys. Rev. +X 8, 031088 (2018). +[85] Mikito Koshino, Noah F. Q. Yuan, Takashi Koretsune, +Masayuki Ochi, Kazuhiko Kuroki, and Liang Fu, “Max- +imally Localized Wannier Orbitals and the Extended +Hubbard Model for Twisted Bilayer Graphene,” Physi- +cal Review X 8, 031087 (2018). +[86] Noah F. Q. Yuan and Liang Fu, “Model for the metal- +insulator transition in graphene superlattices and be- +yond,” Physical Review B 98, 045103 (2018). +[87] Liujun Zou, Hoi Chun Po, Ashvin Vishwanath, +and +T. Senthil, “Band structure of twisted bilayer graphene: +Emergent symmetries, +commensurate approximants, +and Wannier obstructions,” Physical Review B 98, +085435 (2018), publisher: American Physical Society. +[88] Jiawei Zang, Jie Wang, Antoine Georges, Jennifer Cano, +and Andrew J. Millis, “Real space representation of +topological system: twisted bilayer graphene as an ex- +ample,” (2022), arXiv:2210.11573 [cond-mat]. +[89] Kazuyuki Uchida, Shinnosuke Furuya, Jun-Ichi Iwata, +and Atsushi Oshiyama, “Atomic corrugation and elec- +tron localization due to moir´e patterns in twisted bilayer +graphenes,” Phys. Rev. B 90, 155451 (2014). +[90] MM Van Wijk, A Schuring, MI Katsnelson, and A Fa- +solino, “Relaxation of moir´e patterns for slightly mis- +aligned identical lattices: graphene on graphite,” 2D +Materials 2, 034010 (2015). +[91] Shuyang Dai, Yang Xiang, +and David J Srolovitz, +“Twisted bilayer graphene: Moir´e with a twist,” Nano +letters 16, 5923–5927 (2016). +[92] Sandeep K Jain, Vladimir Juriˇci´c, +and Gerard T +Barkema, “Structure of twisted and buckled bilayer +graphene,” 2D Materials 4, 015018 (2016). +[93] Samuel V. Gallego, Emre S. Tasci, Gemma de la Flor, +J. Manuel Perez-Mato, and Mois I. Aroyo, “Magnetic +symmetry in the Bilbao Crystallographic Server: a com- +puter program to provide systematic absences of mag- +netic neutron diffraction,” Journal of Applied Crystal- +lography 45, 1236–1247 (2012). +[94] Jie Wang, Yunqin Zheng, Andrew J. Millis, +and Jen- +nifer Cano, “Chiral approximation to twisted bilayer +graphene: Exact intravalley inversion symmetry, nodal +structure, and implications for higher magic angles,” +Phys. Rev. Res. 3, 023155 (2021). +[95] Kan Chen and C. Jayaprakash, “The Kondo effect in +pseudo-gap Fermi systems: +a renormalization group +study,” Journal of Physics: Condensed Matter 7, L491 +(1995). +[96] Carlos +Gonzalez-Buxton +and +Kevin +Ingersent, +“Renormalization-group +study +of +Anderson +and +Kondo impurities in gapless Fermi systems,” Physical +Review B 57, 14254–14293 (1998). +[97] Kevin Ingersent and Qimiao Si, “Critical Local-Moment +Fluctuations, Anomalous Exponents, and ω/T Scaling +in the Kondo Problem with a Pseudogap,” Physical Re- +view Letters 89, 076403 (2002). +[98] Lars Fritz and Matthias Vojta, “Phase transitions in +the pseudogap Anderson and Kondo models: Critical +dimensions, renormalization group, and local-moment +criticality,” Physical Review B 70, 214427 (2004). +[99] Piers +Coleman, +Introduction +to +many-body +physics +(Cambridge University Press, 2015) see section 17.3 for +the Kondo temperature in the large-N limit. +[100] R. Bulla, T. A. Costi, +and D. Vollhardt, “Finite- +temperature numerical renormalization group study of +the Mott transition,” Physical Review B 64, 045103 +(2001), publisher: American Physical Society. +[101] Ralf Bulla, Hyun-Jung Lee, Ning-Hua Tong, +and +Matthias Vojta, “Numerical renormalization group for +quantum impurities in a bosonic bath,” Physical Review +B 71, 045122 (2005), publisher: American Physical So- +ciety. +[102] Tie-Feng Fang, Ning-Hua Tong, Zhan Cao, Qing-Feng +Sun, and Hong-Gang Luo, “Spin susceptibility of An- +derson impurities in arbitrary conduction bands,” Phys. +Rev. B 92, 155129 (2015), publisher: American Physical +Society. +[103] Ulrich Gerland, Jan von Delft, T. A. Costi, +and Yu- +val Oreg, “Transmission phase shift of a quantum dot +with kondo correlations,” Phys. Rev. Lett. 84, 3710– + +24 +3713 (2000). +[104] Yang-Zhi Chou and Sankar Das Sarma, “Kondo lattice +model in magic-angle twisted bilayer graphene,” arXiv +preprint arXiv:2211.15682 (2022). +[105] Haoyu Hu, G. Rai, L. Crippa, T. Wehling, G. Sangio- +vanni, R. Valen´t, Alexei M. Tsvelik, +and B. Andrei +Bernevig, (2023), to appear. +[106] Haoyu Hu, B. Andrei Bernevig, and Alexei M. Tsvelik, +(2023), to appear. +[107] Piers Coleman, (2023), to appear. +[108] Jiabin Yu, +Ming Xie, +B. Andrei Bernevig, +and +Sankar Das Sarma, “Magic angle twisted symmetric tri- +layer graphene as a topological heavy fermion problem,” +(2023), to appear. +