HTML conversions sometimes display errors due to content that did not convert correctly from the source. This paper uses the following packages that are not yet supported by the HTML conversion tool. Feedback on these issues are not necessary; they are known and are being worked on.

  • failed: changes

Authors: achieve the best HTML results from your LaTeX submissions by following these best practices.

License: arXiv.org perpetual non-exclusive license
arXiv:2401.03013v1 [cond-mat.str-el] 05 Jan 2024
\definechangesauthor

[name=Dibya, color=cyan]d \definechangesauthor[name=Adam, color=purple]a \definechangesauthor[name=Herb, color=orange]h

Impact of a Lifshitz Transition on the onset of spontaneous coherence

Adam Eaton Department of Physics, Indiana University, Bloomington, Indiana 47405, USA    Dibya Mukherjee Department of Physics, Indiana University, Bloomington, Indiana 47405, USA    H. A. Fertig Department of Physics, Indiana University, Bloomington, Indiana 47405, USA Quantum Science and Engineering Center, Indiana University, Bloomington, Indiana 47405, USA
(January 5, 2024)
Abstract

Lifshitz transitions are topological transitions of a Fermi surface, whose signatures typically appear in the conduction properties of a host metal. Here, we demonstrate, using an extended Falicov-Kimball model of a two-flavor fermion system, that a Lifshitz transition which occurs in the non-interacting limit impacts interaction-induced insulating phases, even though they do not host Fermi surfaces. For strong interactions we find a first order transition between states of different polarization This transition line ends in a very unusual quantum critical endpoint, whose presence is stabilized by the onset of inter-flavor coherence. We demonstrate that the surfaces of maximum coherence in these states reflect the distinct Fermi surface topologies of the states separated by the non-interacting Lifshitz transition. The phenomenon is shown to be independent of the band topologies involved. Experimental realizations of our results are discussed for both electronic and optical lattice systems.

first keyword, second keyword, third keyword

Introduction.

—In recent years, topology has become increasingly appreciated in condensed matter physics as a framework for understading diverse physical phenomena. These include the Thouless pump [1, 2], topological defects [3, 4, 5, 6, 7, 8, 9], quantized Hall effects [10, 11, 12, 13, 14, 15, 16, 17, 18], magnetic breakdown [19, 20, 21, 22], and topological insulators [23, 24, 25, 26, 27, 28, 29, 30, 31, 32, 33]. In addition to its utility for theoretical understanding, topology is physically significant because it leads to phenomena that are robust with respect to various perturbations [34, 35, 36, 37, 38].

Lifshitz transitions [39, 40, 41, 42, 43, 44] are an example of this. They occur when the topology of a Fermi surface changes with system parameters such as pressure, do** or external magnetic field [40, 41, 42, 44], and typically are observable as anomalies in magneto-oscillation periods as the system passes through such transitions. Because a Lifshitz transition is a Fermi surface phenomenon, its impact is normally only expected in metallic systems. In this work, we demonstrate that such transitions can also impact systems outside their metallic regime. In particular, we report on a system where a Lifshitz transition in the non-interacting limit leaves a clear signature when gap-opening interactions eliminate the Fermi surface.

Refer to caption      (a)Refer to caption

Refer to caption      (b)Refer to caption

Refer to caption       (c)Refer to caption

Refer to caption         (d)Refer to caption

Refer to caption           (e)Refer to caption

Figure 1: (1), (1), (1): the coherence as a function of crystal momentum. Regions of largest coherence form topologically distinct curves. (1) and (1): Fermi surfaces on either side of the Lifshitz transition in absence of spontaneous coherence.

Specifically, we examine a two-component system with interactions such that inter-component coherent phases can be supported. We demonstrate that the system supports different phases where the loops of maximum coherence in the Brillouin Zone are topologically distinct. An example of this is presented in Fig. 1. The topologies of the Fermi surfaces on either side of the Lifshitz transition mirror the topologies of these maximum coherence loops, demonstrating that the Lifshitz transition “seeds” the quantum phase transition of the interacting system. Importantly, the two different coherent states are separated by a first order transition for relatively strong interactions. We find that, within the models we examine, the transition line hosts a very unusual quantum critical endpoint (QCEP) [45], reminiscent of a thermodynamic 𝒵2subscript𝒵2\mathcal{Z}_{2}caligraphic_Z start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT critical point [9, 46]. As in the the latter, beyond the endpoint the evolution between coherent phases becomes continuous, with no sharp distinction between the states. Examples of this are illustrated in Fig. 2.

Refer to caption                            (a)Refer to caption

Refer to caption                            (b)Refer to caption

Refer to caption                                                                     (c)Refer to caption Refer to caption                                                                     (d)Refer to caption

Refer to caption              (e)Refer to caption

Figure 2: (2) and (2): Energy states solving Hartree-Fock equations, demonstrating hysteresis for large U𝑈Uitalic_U which is absent for small U𝑈Uitalic_U. (c) and (d): Phase diagrams for four-band models for crossing bands of opposite and the same Chern numbers, respectively. (e): Phase diagram of a two-band model for crossing bands of opposite Chern numbers.

The phenomenon we describe in this work is robust with respect to band topology: we find it occurs whether the non-interacting bands that correlate through coherence have the same or different Chern numbers. In the latter case, the first order transition can also entail a transition in the occupied band topology [47, 48], but in such situations one does not find distinct coherent phases on either side of the transition. Our study demonstrates that for systems where a non-interacting Lifshitz transition resides in a setting where spontaneous breaking of a continuous symmetry can occur, a rich set of transitions results, separating phases that offer insight into the different Fermi surface topologies that the non-interacting system supports.

Model Hamiltonian.

—We adopt a minimal model capturing the physics in which we are interested, with Hamiltonian of the form

H^=𝐤=t,bp=±1[pE𝒌+Δ]c^𝐤,,pc^𝐤,,p+U𝐤ρ^𝒌,tρ^𝒌,b,^𝐻subscript𝐤subscript𝑡𝑏subscript𝑝plus-or-minus1delimited-[]𝑝subscript𝐸𝒌subscriptΔsuperscriptsubscript^𝑐𝐤𝑝subscript^𝑐𝐤𝑝𝑈subscript𝐤subscript^𝜌𝒌𝑡subscript^𝜌𝒌𝑏\hat{H}=\sum_{\bf k}\sum_{\ell=t,b}\sum_{p=\pm 1}\left[pE_{\bm{k}}+\Delta_{% \ell}\right]\hat{c}_{{\bf k},\ell,p}^{{\dagger}}\hat{c}_{{\bf k},\ell,p}+U\sum% _{\bf k}\hat{\rho}_{-\bm{k},t}\hat{\rho}_{\bm{k},b},over^ start_ARG italic_H end_ARG = ∑ start_POSTSUBSCRIPT bold_k end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT roman_ℓ = italic_t , italic_b end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_p = ± 1 end_POSTSUBSCRIPT [ italic_p italic_E start_POSTSUBSCRIPT bold_italic_k end_POSTSUBSCRIPT + roman_Δ start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT ] over^ start_ARG italic_c end_ARG start_POSTSUBSCRIPT bold_k , roman_ℓ , italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT over^ start_ARG italic_c end_ARG start_POSTSUBSCRIPT bold_k , roman_ℓ , italic_p end_POSTSUBSCRIPT + italic_U ∑ start_POSTSUBSCRIPT bold_k end_POSTSUBSCRIPT over^ start_ARG italic_ρ end_ARG start_POSTSUBSCRIPT - bold_italic_k , italic_t end_POSTSUBSCRIPT over^ start_ARG italic_ρ end_ARG start_POSTSUBSCRIPT bold_italic_k , italic_b end_POSTSUBSCRIPT , (1)

where c^𝐤,,±1superscriptsubscript^𝑐𝐤plus-or-minus1\hat{c}_{{\bf k},\ell,\pm 1}^{{\dagger}}over^ start_ARG italic_c end_ARG start_POSTSUBSCRIPT bold_k , roman_ℓ , ± 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT creates a particle in an eigenstate of a Bernevig-Hughes-Zhang Hamiltonian [17] h,𝒌subscript𝒌h_{\ell,\bm{k}}italic_h start_POSTSUBSCRIPT roman_ℓ , bold_italic_k end_POSTSUBSCRIPT, with eigenvalues ±E𝐤plus-or-minussubscript𝐸𝐤\pm E_{\bf k}± italic_E start_POSTSUBSCRIPT bold_k end_POSTSUBSCRIPT and eigenvectors χ𝐪,±1subscript𝜒𝐪plus-or-minus1\chi_{{\bf q},\pm 1}italic_χ start_POSTSUBSCRIPT bold_q , ± 1 end_POSTSUBSCRIPT. (Details may be found in the Supplementary Material (SM) [49].) The second term is an interlayer contact interaction [30, 16] that involves density operators ρ^𝒌,=𝐪,p=±1χ𝐤+𝐪,pχ𝐪,pc^𝐪+𝐤,,pc^𝐪,,psubscript^𝜌𝒌subscript𝐪𝑝plus-or-minus1subscriptsuperscript𝜒𝐤𝐪𝑝subscript𝜒𝐪𝑝superscriptsubscript^𝑐𝐪𝐤𝑝subscript^𝑐𝐪𝑝\hat{\rho}_{\bm{k},\ell}=\sum_{{\bf q},p=\pm 1}\chi^{{\dagger}}_{{\bf k}+{\bf q% },p}\cdot\chi_{{\bf q},p}\hat{c}_{{\bf q}+{\bf k},\ell,p}^{{\dagger}}\hat{c}_{% {\bf q},\ell,p}over^ start_ARG italic_ρ end_ARG start_POSTSUBSCRIPT bold_italic_k , roman_ℓ end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT bold_q , italic_p = ± 1 end_POSTSUBSCRIPT italic_χ start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_k + bold_q , italic_p end_POSTSUBSCRIPT ⋅ italic_χ start_POSTSUBSCRIPT bold_q , italic_p end_POSTSUBSCRIPT over^ start_ARG italic_c end_ARG start_POSTSUBSCRIPT bold_q + bold_k , roman_ℓ , italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT over^ start_ARG italic_c end_ARG start_POSTSUBSCRIPT bold_q , roman_ℓ , italic_p end_POSTSUBSCRIPT. Eq. 1 preserves the total number of fermions of t𝑡titalic_t- and b𝑏bitalic_b-flavor separately, and supports a U(1)𝑈1U(1)italic_U ( 1 ) flavor symmetry. Such models belong to the general class of extended Falicov-Kimball models (EFKM), which have been studied as candidates for describing electronic excitonic physics, ferroelectricity and bilayers phases [50, 51, 52, 53, 54, 55, 56, 57, 58, 59]. The impacts of Lifshitz transitions and band topology on such systems, the focus of this study, are to our knowledge unknown.

The model can also be interpreted as a bilayer system without tunneling between layers [60, 61, 62, 63]), in which intralayer interactions have been neglected. In what follows we adopt this realization as a paradigm for such systems, and refer to the flavors as layers. We expect the behavior of this model at half-filling to apply when bands of each flavor are relatively far apart energetically, while there is a crossing of bands of two different flavors: with short-range interactions, Fermi statistics suppresses short-range intra-flavor interactions, allowing inter-flavor interactions to dominate. Beyond this, the model can also be mapped onto a two-species fermionic atomic gas system in an optical lattice, as we discuss below.

Hartree-Fock Analysis.

—We consider ground states of H^^𝐻\hat{H}over^ start_ARG italic_H end_ARG within the Hartree-Fock (HF) approximation, in situations where the system is half-filled. Details of the analysis may be found in the SM [49]. For Δ=0Δ0\Delta=0roman_Δ = 0, the spectrum consists of one filled and one empty band for each layer, with an intervening gap that is present even in the absence of interactions. With increasing |Δ|Δ|\Delta|| roman_Δ |, bands of opposite flavor approach one another, eventually crossing when spontaneous symmetry breaking is not considered. The Fermi surfaces consist of matching loops for each flavor surrounding the ΓΓ\Gammaroman_Γ point of the square Brillouin zone (BZ). With growing |Δ|Δ|\Delta|| roman_Δ | these loops eventually touch the M𝑀Mitalic_M points at the BZ edge, signaling a Lifshitz transition. For still larger |Δ|Δ|\Delta|| roman_Δ | the loop topology changes, now surrounding the X𝑋Xitalic_X points (corners) of the BZ.

Multi-flavor systems with matching Fermi surfaces are known to be unstable to spontaneous inter-flavor coherence in the presence of interactions [64, 65]. Figs. 2 (c) and (d) illustrates what happens, within our HF analysis, when two bands pass fully through one another as a function of ΔΔ\Deltaroman_Δ, with a Lifshitz transition separating different Fermi surface topologies (see SM [49] for an illustration.) We consider both crossing bands of the same Chern number 𝒞=1𝒞1\mathcal{C}=1caligraphic_C = 1 [panel (c)], and crossing bands of opposite Chern number 𝒞=±1𝒞plus-or-minus1\mathcal{C}=\pm 1caligraphic_C = ± 1 [panel (d)]. Continuous transitions are indicated as dashed lines while solid lines indicate first order transitions. The resulting phases include a non-interacting gap (NIG) phase, which is continuously connected to the Δ=0Δ0\Delta=0roman_Δ = 0 non-interacting state; inter-layer coherent phases (Coh I and II); and a layer-polarized (LP) phase.

To characterize the polarization of a phase, we define a polarization function P(𝒌)i=14ψi(𝒌)|τz𝟙|ψi(𝒌)f(Ei(𝒌))i=14Pi(𝒌)f(Ei(𝒌)),𝑃𝒌superscriptsubscript𝑖14quantum-operator-productsubscript𝜓𝑖𝒌tensor-productsubscript𝜏𝑧1subscript𝜓𝑖𝒌𝑓subscript𝐸𝑖𝒌superscriptsubscript𝑖14subscript𝑃𝑖𝒌𝑓subscript𝐸𝑖𝒌P(\bm{k})\equiv\sum_{i=1}^{4}\langle\psi_{i}(\bm{k})|\tau_{z}\otimes\mathbbm{1% }|\psi_{i}(\bm{k})\rangle f(E_{i}(\bm{k}))\equiv\sum_{i=1}^{4}P_{i}(\bm{k})f(E% _{i}(\bm{k})),italic_P ( bold_italic_k ) ≡ ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT ⟨ italic_ψ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( bold_italic_k ) | italic_τ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ⊗ blackboard_1 | italic_ψ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( bold_italic_k ) ⟩ italic_f ( italic_E start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( bold_italic_k ) ) ≡ ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT italic_P start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( bold_italic_k ) italic_f ( italic_E start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( bold_italic_k ) ) , where |ψi(𝐤)ketsubscript𝜓𝑖𝐤|\psi_{i}({\bf k})\rangle| italic_ψ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( bold_k ) ⟩ are the four HF wavefunctions at wavevector 𝐤𝐤{\bf k}bold_k with energy Ei(𝐤)subscript𝐸𝑖𝐤E_{i}({\bf k})italic_E start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( bold_k ), f𝑓fitalic_f is the Fermi function, and the τzsubscript𝜏𝑧\tau_{z}italic_τ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT is a Pauli matrix acting in layer-space. Spontaneous coherence in the system arises when order parameters of the form c𝐤,,pc𝐤,,p0delimited-⟨⟩subscriptsuperscript𝑐𝐤𝑝subscript𝑐𝐤superscriptsuperscript𝑝0\langle c^{{\dagger}}_{{\bf k},\ell,p}c_{{\bf k},\ell^{\prime},p^{\prime}}% \rangle\neq 0⟨ italic_c start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_k , roman_ℓ , italic_p end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT bold_k , roman_ℓ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_p start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ⟩ ≠ 0 for superscript\ell\neq\ell^{\prime}roman_ℓ ≠ roman_ℓ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT. This always entails values of Pi(𝐤)subscript𝑃𝑖𝐤P_{i}({\bf k})italic_P start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( bold_k ) which are not equal to either 11-1- 1 or 1111, so that non-vanishing values of a coherence function, C(𝒌)𝒌i=14(1|Pi(𝒌)|)f(Ei(𝒌)),𝐶𝒌subscript𝒌superscriptsubscript𝑖141subscript𝑃𝑖𝒌𝑓subscript𝐸𝑖𝒌C(\bm{k})\equiv\sum_{\bm{k}}\sum_{i=1}^{4}(1-|P_{i}(\bm{k})|)f(E_{i}(\bm{k})),italic_C ( bold_italic_k ) ≡ ∑ start_POSTSUBSCRIPT bold_italic_k end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT ( 1 - | italic_P start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( bold_italic_k ) | ) italic_f ( italic_E start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( bold_italic_k ) ) , signal the presence of interlayer coherence.

Fig. 1 illustrates the behavior of C(𝐤)𝐶𝐤C({\bf k})italic_C ( bold_k ) as a function of 𝐤𝐤{\bf k}bold_k for two coherent states, and their correlation with the non-interacting Fermi surface. The loci of maximum C(𝐤)𝐶𝐤C({\bf k})italic_C ( bold_k ) have two different topologies reflecting the behavior of the Fermi surfaces on either side of the non-interacting Lifshitz transition. For small U𝑈Uitalic_U the evolution from one behavior to the other as a function of ΔΔ\Deltaroman_Δ is continuous, while for larger values there is a first order transition between them (see Figs. 2 (a) and (b)). The transition line is quite interesting. It hosts a very unusual QCEP, which at the mean-field level is highly analogous to the critical point of a thermal 𝒵2subscript𝒵2\mathcal{Z}_{2}caligraphic_Z start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT transition. Indeed, the behavior of the full polarization, P=𝐤P(𝐤)𝑃subscript𝐤𝑃𝐤P=\sum_{{\bf k}}P({\bf k})italic_P = ∑ start_POSTSUBSCRIPT bold_k end_POSTSUBSCRIPT italic_P ( bold_k ), when the bias ΔΔ\Deltaroman_Δ is varied, displays a jump that continuously vanishes at the QCEP, as illustrated in Fig. 3.

Refer to caption
Figure 3: Plot of P𝑃Pitalic_P versus 1/Δ1Δ1/\Delta1 / roman_Δ for fixed U𝑈Uitalic_U, illustrating a closing discontinuous polarization jump.

Beyond this, as U𝑈Uitalic_U drops from large values, the transition line gives birth to two continuous transitions, in which coherence sets in from either the NIG or the LP phase. Moreover, although the situations illustrated in Figs. 2 (c) and (d) look very similar, they have an important difference. In (2), the transition out of the NIG phase involves a change in Chern number. The transition occurs without a gap closing at large U𝑈Uitalic_U [47, 66, 67, 28], while at smaller U𝑈Uitalic_U the onset of coherence occurs continuously and simultaneously with the topological transition. The topological transition is in this way tied up with the symmetry-breaking transition. By contrast, in (2) there is no topological transition of the filled bands, and there is no closing of the band gap at any of the quantum phase transitions. Fig. 4 illustrates this difference in behavior.

Two Band Model.

—Several of the results described above can be qualitatively understood within a two band model, in which one retains only the two bands that cross one another in the model. These can have the same or opposite topologies; Fig. 2(e) illustrates the resulting phase diagram for the case of opposite Chern numbers. Although there is a change in the locations of the phase boundaries, particularly at large U𝑈Uitalic_U, the system retains the same basic phases and features of the four band model. Most prominent is the first order transition line drop** from large U𝑈Uitalic_U, which ends at a QCEP.

The phases of this model are characterized by two order parameters: (i) BpU2V𝐤c^𝐤,tc^𝐤,tc^𝐤,bc^𝐤,bsubscript𝐵𝑝𝑈2𝑉subscript𝐤delimited-⟨⟩subscriptsuperscript^𝑐𝐤𝑡subscript^𝑐𝐤𝑡subscriptsuperscript^𝑐𝐤𝑏subscript^𝑐𝐤𝑏B_{p}\equiv{U\over 2V}\sum_{\bf k}\langle\hat{c}^{{\dagger}}_{{\bf k},t}\hat{c% }_{{\bf k},t}-\hat{c}^{{\dagger}}_{{\bf k},b}\hat{c}_{{\bf k},b}\rangleitalic_B start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ≡ divide start_ARG italic_U end_ARG start_ARG 2 italic_V end_ARG ∑ start_POSTSUBSCRIPT bold_k end_POSTSUBSCRIPT ⟨ over^ start_ARG italic_c end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_k , italic_t end_POSTSUBSCRIPT over^ start_ARG italic_c end_ARG start_POSTSUBSCRIPT bold_k , italic_t end_POSTSUBSCRIPT - over^ start_ARG italic_c end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_k , italic_b end_POSTSUBSCRIPT over^ start_ARG italic_c end_ARG start_POSTSUBSCRIPT bold_k , italic_b end_POSTSUBSCRIPT ⟩, where c^𝐤,subscript^𝑐𝐤\hat{c}_{{\bf k},\ell}over^ start_ARG italic_c end_ARG start_POSTSUBSCRIPT bold_k , roman_ℓ end_POSTSUBSCRIPT annihilates a particle in the retained band of layer \ellroman_ℓ, which is a measure of the polarization of the system; and (ii) an interlayer coherence Btb(𝐤)UV𝐤1|χ𝐤,tχ𝐤1,b|2c𝐤1,tc𝐤1,bsuperscript𝐵𝑡𝑏𝐤𝑈𝑉subscriptsubscript𝐤1superscriptsuperscriptsubscript𝜒𝐤𝑡subscript𝜒subscript𝐤1𝑏2delimited-⟨⟩subscriptsuperscript𝑐subscript𝐤1𝑡subscript𝑐subscript𝐤1𝑏B^{tb}({\bf k})\equiv{U\over V}\sum_{{\bf k}_{1}}|\chi_{{\bf k},t}^{{\dagger}}% \cdot\chi_{{\bf k}_{1},b}|^{2}\langle c^{{\dagger}}_{{\bf k}_{1},t}c_{{\bf k}_% {1},b}\rangleitalic_B start_POSTSUPERSCRIPT italic_t italic_b end_POSTSUPERSCRIPT ( bold_k ) ≡ divide start_ARG italic_U end_ARG start_ARG italic_V end_ARG ∑ start_POSTSUBSCRIPT bold_k start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUBSCRIPT | italic_χ start_POSTSUBSCRIPT bold_k , italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ⋅ italic_χ start_POSTSUBSCRIPT bold_k start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_b end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟨ italic_c start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_k start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_t end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT bold_k start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_b end_POSTSUBSCRIPT ⟩. (V𝑉Vitalic_V is the system area). Within mean-field theory the self-consistent equations for these have the form [49]

BpU=12V𝐤2subscript𝐵𝑝𝑈12𝑉subscriptsubscript𝐤2\displaystyle\frac{B_{p}}{U}=\dfrac{1}{2V}\sum_{{\bf k}_{2}}divide start_ARG italic_B start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG start_ARG italic_U end_ARG = divide start_ARG 1 end_ARG start_ARG 2 italic_V end_ARG ∑ start_POSTSUBSCRIPT bold_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT ξ~(𝐤2)|ξ~(𝐤2)|2+|Btb(𝐤2)|2F(Bp+Δ);~𝜉subscript𝐤2superscript~𝜉subscript𝐤22superscriptsuperscript𝐵𝑡𝑏subscript𝐤22𝐹subscript𝐵𝑝Δ\displaystyle\dfrac{\widetilde{\xi}({\bf k}_{2})}{\sqrt{|\widetilde{\xi}({\bf k% }_{2})|^{2}+|B^{tb}({\bf k}_{2})|^{2}}}\equiv F(B_{p}+\Delta);divide start_ARG over~ start_ARG italic_ξ end_ARG ( bold_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) end_ARG start_ARG square-root start_ARG | over~ start_ARG italic_ξ end_ARG ( bold_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + | italic_B start_POSTSUPERSCRIPT italic_t italic_b end_POSTSUPERSCRIPT ( bold_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_ARG ≡ italic_F ( italic_B start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT + roman_Δ ) ; (2)
Btb(𝐤1)=U2V𝐤2superscript𝐵𝑡𝑏subscript𝐤1𝑈2𝑉subscriptsubscript𝐤2\displaystyle B^{tb}({\bf k}_{1})={U\over{2V}}\sum_{{\bf k}_{2}}italic_B start_POSTSUPERSCRIPT italic_t italic_b end_POSTSUPERSCRIPT ( bold_k start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) = divide start_ARG italic_U end_ARG start_ARG 2 italic_V end_ARG ∑ start_POSTSUBSCRIPT bold_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT f(θ𝐤1,θ𝐤1;𝒞rel)Btb(𝐤2)|ξ~(𝐤2)|2+|Btb(𝐤2)|2,𝑓subscript𝜃subscript𝐤1subscript𝜃subscript𝐤1subscript𝒞𝑟𝑒𝑙superscript𝐵𝑡𝑏subscript𝐤2superscript~𝜉subscript𝐤22superscriptsuperscript𝐵𝑡𝑏subscript𝐤22\displaystyle\dfrac{f(\theta_{{\bf k}_{1}},\theta_{{\bf k}_{1}};\mathcal{C}_{% rel})\,B^{tb}({\bf k}_{2})}{\sqrt{|\widetilde{\xi}({\bf k}_{2})|^{2}+|B^{tb}({% \bf k}_{2})|^{2}}},divide start_ARG italic_f ( italic_θ start_POSTSUBSCRIPT bold_k start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUBSCRIPT , italic_θ start_POSTSUBSCRIPT bold_k start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUBSCRIPT ; caligraphic_C start_POSTSUBSCRIPT italic_r italic_e italic_l end_POSTSUBSCRIPT ) italic_B start_POSTSUPERSCRIPT italic_t italic_b end_POSTSUPERSCRIPT ( bold_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) end_ARG start_ARG square-root start_ARG | over~ start_ARG italic_ξ end_ARG ( bold_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + | italic_B start_POSTSUPERSCRIPT italic_t italic_b end_POSTSUPERSCRIPT ( bold_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_ARG , (3)

where ξ~(𝐤2)=E𝐤2+Δ+Bp~𝜉subscript𝐤2subscript𝐸subscript𝐤2Δsubscript𝐵𝑝\tilde{\xi}({\bf k}_{2})=E_{{\bf k}_{2}}+\Delta+B_{p}over~ start_ARG italic_ξ end_ARG ( bold_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) = italic_E start_POSTSUBSCRIPT bold_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT + roman_Δ + italic_B start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT, 𝒞rel=1(1)subscript𝒞𝑟𝑒𝑙11\mathcal{C}_{rel}=1(-1)caligraphic_C start_POSTSUBSCRIPT italic_r italic_e italic_l end_POSTSUBSCRIPT = 1 ( - 1 ) when the product of the Chern numbers for the crossing bands is 1(-1), f(θ𝐤1,θ𝐤1;1)=(1+cosθ𝐤1cosθ𝐤2)/2𝑓subscript𝜃subscript𝐤1subscript𝜃subscript𝐤111subscript𝜃subscript𝐤1subscript𝜃subscript𝐤22f(\theta_{{\bf k}_{1}},\theta_{{\bf k}_{1}};1)=(1+\cos\theta_{{\bf k}_{1}}\cos% \theta_{{\bf k}_{2}})/{2}italic_f ( italic_θ start_POSTSUBSCRIPT bold_k start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUBSCRIPT , italic_θ start_POSTSUBSCRIPT bold_k start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUBSCRIPT ; 1 ) = ( 1 + roman_cos italic_θ start_POSTSUBSCRIPT bold_k start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUBSCRIPT roman_cos italic_θ start_POSTSUBSCRIPT bold_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT ) / 2 and f(θ𝐤1,θ𝐤1;1)=(cos2θ𝐤1/2)(cos2θ𝐤2/2)𝑓subscript𝜃subscript𝐤1subscript𝜃subscript𝐤11superscript2subscript𝜃subscript𝐤12superscript2subscript𝜃subscript𝐤22f(\theta_{{\bf k}_{1}},\theta_{{\bf k}_{1}};-1)=\left(\cos^{2}\theta_{{\bf k}_% {1}}/2\right)\left(\cos^{2}\theta_{{\bf k}_{2}}/2\right)italic_f ( italic_θ start_POSTSUBSCRIPT bold_k start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUBSCRIPT , italic_θ start_POSTSUBSCRIPT bold_k start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUBSCRIPT ; - 1 ) = ( roman_cos start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_θ start_POSTSUBSCRIPT bold_k start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUBSCRIPT / 2 ) ( roman_cos start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_θ start_POSTSUBSCRIPT bold_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT / 2 ). In Eq. 3 we have assumed that c𝐤1,tc𝐤1,bdelimited-⟨⟩subscriptsuperscript𝑐subscript𝐤1𝑡subscript𝑐subscript𝐤1𝑏\langle c^{{\dagger}}_{{\bf k}_{1},t}c_{{\bf k}_{1},b}\rangle⟨ italic_c start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_k start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_t end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT bold_k start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_b end_POSTSUBSCRIPT ⟩ is real and has C4subscript𝐶4C_{4}italic_C start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT rotational symmetry in the HF ground state, which we indeed find in our more general numerical analysis.

Refer to caption       (a)Refer to caption

Refer to caption       (b)Refer to caption

Figure 4: Gap vs. bias ΔΔ\Deltaroman_Δ for U=1.8𝑈1.8U=1.8italic_U = 1.8. (4): Crossing bands have the same Chern number 𝒞=1𝒞1\mathcal{C}=1caligraphic_C = 1. (4): Crossing bands have opposite Chern number 𝒞=±1𝒞plus-or-minus1\mathcal{C}=\pm 1caligraphic_C = ± 1. Insets: Gap remains open in first case but closes in second.

Eq. 2 provides particular insight into the connection between the first-order transition lines in Fig. 2 and a thermal 𝒵2subscript𝒵2\mathcal{Z}_{2}caligraphic_Z start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT transition. Direct plots of the left- and right-hand sides of the equation show that at large positive (negative) values of ΔΔ\Deltaroman_Δ, one finds a single solution with maximal negative (positive) values of Bpsubscript𝐵𝑝B_{p}italic_B start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT, but in some transition region of ΔΔ\Deltaroman_Δ there are three solutions. (See SM for details [49].) The physical state of the system jumps between two of the three solutions when their energies cross, behavior which is highly reminiscent of what one finds in a mean-field treatment of, for example, the liquid-gas transition [68]. Such a jump always occurs provided maxxdF(x)dx>1Usubscript𝑥𝑑𝐹𝑥𝑑𝑥1𝑈\max_{x}\frac{dF(x)}{dx}>\frac{1}{U}roman_max start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT divide start_ARG italic_d italic_F ( italic_x ) end_ARG start_ARG italic_d italic_x end_ARG > divide start_ARG 1 end_ARG start_ARG italic_U end_ARG. Interestingly, because of the Lifshitz transition, one finds a point for which the left-hand side of the inequality diverges when Btb=0superscript𝐵𝑡𝑏0B^{tb}=0italic_B start_POSTSUPERSCRIPT italic_t italic_b end_POSTSUPERSCRIPT = 0, so that there is a first order jump for any positive value of U𝑈Uitalic_U, and no QCEP is manifested. However, for |Btb|>0superscript𝐵𝑡𝑏0|B^{tb}|>0| italic_B start_POSTSUPERSCRIPT italic_t italic_b end_POSTSUPERSCRIPT | > 0, this divergence is smoothed over, maxxdF(x)dx<subscript𝑥𝑑𝐹𝑥𝑑𝑥\max_{x}\frac{dF(x)}{dx}<\inftyroman_max start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT divide start_ARG italic_d italic_F ( italic_x ) end_ARG start_ARG italic_d italic_x end_ARG < ∞, and for sufficiently small (but non-vanishing) U𝑈Uitalic_U the first order jump gives way to a continuous cross-over: a QCEP is stabilized. This mean-field phenomenology is highly reminiscent of that of a classical liquid-gas critical point. Remarkably, in order to be realized in this setting, spontaneous coherence, a purely quantum phenomenon, must be manifested between layers.

Eq. 3 also allows an understanding of the behaviors of the coherence onset at the NIG-Coh I boundary for small U𝑈Uitalic_U, which are different depending on whether Crel=±1subscript𝐶𝑟𝑒𝑙plus-or-minus1C_{rel}=\pm 1italic_C start_POSTSUBSCRIPT italic_r italic_e italic_l end_POSTSUBSCRIPT = ± 1. In this regime, the bias is such that the single-particle energies of the two bands are very close near the ΓΓ\Gammaroman_Γ point. Writing Δ0subscriptΔ0\Delta_{0}roman_Δ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT for the bias at which the two bands touch at the ΓΓ\Gammaroman_Γ point, we define Δ~ΔΔ0~ΔΔsubscriptΔ0\tilde{\Delta}\equiv\Delta-\Delta_{0}over~ start_ARG roman_Δ end_ARG ≡ roman_Δ - roman_Δ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, and simplify our model by assuming the bands to be quadratic, in which case they have a constant density of states g0subscript𝑔0g_{0}italic_g start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. Within this model and for small k𝑘kitalic_k one may assume cosθ𝐤=cosθ(E𝐤)1+α(E𝐤E0)subscript𝜃𝐤𝜃subscript𝐸𝐤1𝛼subscript𝐸𝐤subscript𝐸0\cos\theta_{{\bf k}}=\cos\theta(E_{\bf k})\approx-1+\alpha(E_{\bf k}-E_{0})roman_cos italic_θ start_POSTSUBSCRIPT bold_k end_POSTSUBSCRIPT = roman_cos italic_θ ( italic_E start_POSTSUBSCRIPT bold_k end_POSTSUBSCRIPT ) ≈ - 1 + italic_α ( italic_E start_POSTSUBSCRIPT bold_k end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ), with α𝛼\alphaitalic_α a parameter of order the bandwidth W𝑊Witalic_W. As shown in the SM [49], one may then show Btbb0(1cosθ𝐤)superscript𝐵𝑡𝑏subscript𝑏01subscript𝜃𝐤B^{tb}\approx b_{0}(1-\cos\theta_{\bf k})italic_B start_POSTSUPERSCRIPT italic_t italic_b end_POSTSUPERSCRIPT ≈ italic_b start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( 1 - roman_cos italic_θ start_POSTSUBSCRIPT bold_k end_POSTSUBSCRIPT ) for 𝒞rel=1subscript𝒞𝑟𝑒𝑙1\mathcal{C}_{rel}=1caligraphic_C start_POSTSUBSCRIPT italic_r italic_e italic_l end_POSTSUBSCRIPT = 1, and Btbb0cos2θ𝐤/2superscript𝐵𝑡𝑏subscript𝑏0superscript2subscript𝜃𝐤2B^{tb}\approx b_{0}\cos^{2}\theta_{\bf k}/2italic_B start_POSTSUPERSCRIPT italic_t italic_b end_POSTSUPERSCRIPT ≈ italic_b start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT roman_cos start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_θ start_POSTSUBSCRIPT bold_k end_POSTSUBSCRIPT / 2 for 𝒞rel=1subscript𝒞𝑟𝑒𝑙1\mathcal{C}_{rel}=-1caligraphic_C start_POSTSUBSCRIPT italic_r italic_e italic_l end_POSTSUBSCRIPT = - 1, with

b0Wsubscript𝑏0𝑊\displaystyle b_{0}\approx Witalic_b start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≈ italic_W exp[2Ug0]𝒞rel=1;2𝑈subscript𝑔0subscript𝒞𝑟𝑒𝑙1\displaystyle\exp\left[-\frac{2}{Ug_{0}}\right]\quad\quad\quad\mathcal{C}_{rel% }=1;roman_exp [ - divide start_ARG 2 end_ARG start_ARG italic_U italic_g start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ] caligraphic_C start_POSTSUBSCRIPT italic_r italic_e italic_l end_POSTSUBSCRIPT = 1 ;
b04Wα2|Δ~|subscript𝑏04𝑊superscript𝛼2~Δ\displaystyle b_{0}\approx 4\sqrt{\frac{W}{\alpha^{2}|\tilde{\Delta}|}}italic_b start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≈ 4 square-root start_ARG divide start_ARG italic_W end_ARG start_ARG italic_α start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT | over~ start_ARG roman_Δ end_ARG | end_ARG end_ARG exp[4α2Δ~2Ug0]Θ(Δ~)𝒞rel=1;4superscript𝛼2superscript~Δ2𝑈subscript𝑔0Θ~Δsubscript𝒞𝑟𝑒𝑙1\displaystyle\exp\left[\frac{-4}{\alpha^{2}\tilde{\Delta}^{2}Ug_{0}}\right]% \Theta(-\tilde{\Delta})\quad\mathcal{C}_{rel}=-1;roman_exp [ divide start_ARG - 4 end_ARG start_ARG italic_α start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT over~ start_ARG roman_Δ end_ARG start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_U italic_g start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ] roman_Θ ( - over~ start_ARG roman_Δ end_ARG ) caligraphic_C start_POSTSUBSCRIPT italic_r italic_e italic_l end_POSTSUBSCRIPT = - 1 ;

for small |Δ~|~Δ|\tilde{\Delta}|| over~ start_ARG roman_Δ end_ARG |. In these expressions, ΘΘ\Thetaroman_Θ is a Heaviside step function. These results show that for bands of opposite Chern number (𝒞rel=1subscript𝒞𝑟𝑒𝑙1\mathcal{C}_{rel}=-1caligraphic_C start_POSTSUBSCRIPT italic_r italic_e italic_l end_POSTSUBSCRIPT = - 1) the band closing (Δ~0~Δ0\tilde{\Delta}\rightarrow 0over~ start_ARG roman_Δ end_ARG → 0) associated with a change of topology of the occupied band occurs simultaneously with the onset of coherence (b00subscript𝑏00b_{0}\neq 0italic_b start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≠ 0). This contrasts with the situation for 𝒞rel=1subscript𝒞𝑟𝑒𝑙1\mathcal{C}_{rel}=1caligraphic_C start_POSTSUBSCRIPT italic_r italic_e italic_l end_POSTSUBSCRIPT = 1, in which b0subscript𝑏0b_{0}italic_b start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is non-zero for small |Δ~|~Δ|\tilde{\Delta}|| over~ start_ARG roman_Δ end_ARG | even if the corresponding non-interacting bands do not cross. Thus the single particle energy gap never closes in this case. The differing Berry’s curvatures of the two bands for 𝒞rel=1subscript𝒞𝑟𝑒𝑙1\mathcal{C}_{rel}=-1caligraphic_C start_POSTSUBSCRIPT italic_r italic_e italic_l end_POSTSUBSCRIPT = - 1 case frustrates the formation of coherence in the system. This is apparent in the insets of Fig. 4.

Discussion.

—These results suggest a number of interesting questions. One set of these addresses the universality classes of the critical point as well as those of the coherence onset regions when the bands first cross. In the former case, the presence of a broken U(1)𝑈1U(1)italic_U ( 1 ) symmetry and its accompanying Goldstone mode suggests that its critical behavior will be different than that of a classical thermal 𝒵2subscript𝒵2\mathcal{Z}_{2}caligraphic_Z start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT transition. In the latter case, the presence of gapless fermions for 𝒞rel=1subscript𝒞𝑟𝑒𝑙1\mathcal{C}_{rel}=-1caligraphic_C start_POSTSUBSCRIPT italic_r italic_e italic_l end_POSTSUBSCRIPT = - 1 suggests the transition will be in a different universality class than for 𝒞rel=1subscript𝒞𝑟𝑒𝑙1\mathcal{C}_{rel}=1caligraphic_C start_POSTSUBSCRIPT italic_r italic_e italic_l end_POSTSUBSCRIPT = 1. Effects of real thermal fluctuations on the system, and the form a quantum critical region [69, 70] takes, is import to understand in settings where temperature effects cannot be ignored. Another set of questions involve how this phase diagram might be manifested in different physical realizations. One involves an optical lattice [71] hosting two species of atoms with an inter-species Feshbach resonance [72, 73]. A particle-hole transformation maps this onto an EFKM; in this case coherence is realized in superconducting states of the system. For appropriate parameters, we expect two such states, separated by a first order transition. Finally, van der Waals materials [74] offer platforms for electron bilayer realizations of this system, which support layer polarized states [75, 76] and/or interlayer coherence [77, 78, 79, 80, 81], whose interaction and competition could lead to novel quantum phase boundaries and transitions such as those we have described in this study.

Acknowledgements

—The authors acknowledge useful discussions with Chunli Huang, Ganpathy Murthy, and Phil Richerme. This research was supported in part by Lilly Endowment, Inc., through its support for the Indiana University Pervasive Technology Institute. This work is supported in part by NSF Grant Nos. DMR-1914451 and ECCS-1936406. The authors thank the Aspen Center for Physics (NSF Grant No. PHY-1607611) where part of this work was done.

References

  • Kraus et al. [2012] Y. E. Kraus, Y. Lahini, Z. Ringel, M. Verbin, and O. Zilberberg, Topological states and adiabatic pum** in quasicrystals, Phys. Rev. Lett. 109, 106402 (2012).
  • Madsen et al. [2013] K. A. Madsen, E. J. Bergholtz, and P. W. Brouwer, Topological equivalence of crystal and quasicrystal band structures, Phys. Rev. B 88, 125118 (2013).
  • Ringel et al. [2012] Z. Ringel, Y. E. Kraus, and A. Stern, Strong side of weak topological insulators, Phys. Rev. B 86, 045102 (2012).
  • Teo and Hughes [2013] J. C. Y. Teo and T. L. Hughes, Existence of majorana-fermion bound states on disclinations and the classification of topological crystalline superconductors in two dimensions, Phys. Rev. Lett. 111, 047006 (2013).
  • Veldhorst et al. [2012] M. Veldhorst, M. Snelder, M. Hoek, T. Gang, V. K. Guduru, X. L. Wang, U. Zeitler, W. G. van der Wiel, A. A. Golubov, H. Hilgenkamp, and A. Brinkman, Josephson supercurrent through a topological insulator surface state, Nature Materials 11, 417 (2012).
  • Fu and Kane [2008] L. Fu and C. L. Kane, Superconducting proximity effect and majorana fermions at the surface of a topological insulator, Phys. Rev. Lett. 100, 096407 (2008).
  • Fu and Kane [2009] L. Fu and C. L. Kane, Josephson current and noise at a superconductor/quantum-spin-hall-insulator/superconductor junction, Phys. Rev. B 79, 161408 (2009).
  • Nelson [2002] D. R. Nelson, Defects and geometry in condensed matter physics (Cambridge University Press, 2002).
  • Chaikin et al. [1995] P. M. Chaikin, T. C. Lubensky, and T. A. Witten, Principles of condensed matter physics, Vol. 10 (Cambridge university press Cambridge, 1995).
  • Thouless et al. [1982] D. J. Thouless, M. Kohmoto, M. P. Nightingale, and M. den Nijs, Quantized hall conductance in a two-dimensional periodic potential, Phys. Rev. Lett. 49, 405 (1982).
  • Väyrynen et al. [2013] J. I. Väyrynen, M. Goldstein, and L. I. Glazman, Helical edge resistance introduced by charge puddles, Phys. Rev. Lett. 110, 216402 (2013).
  • Hart et al. [2014] S. Hart, H. Ren, T. Wagner, P. Leubner, M. Mühlbauer, C. Brüne, H. Buhmann, L. W. Molenkamp, and A. Yacoby, Induced superconductivity in the quantum spin hall edge, Nature Physics 10, 638 (2014).
  • Weeks et al. [2011] C. Weeks, J. Hu, J. Alicea, M. Franz, and R. Wu, Engineering a robust quantum spin hall state in graphene via adatom deposition, Phys. Rev. X 1, 021001 (2011).
  • Hou et al. [2009] C.-Y. Hou, E.-A. Kim, and C. Chamon, Corner junction as a probe of helical edge states, Phys. Rev. Lett. 102, 076602 (2009).
  • Du et al. [2015] L. Du, I. Knez, G. Sullivan, and R.-R. Du, Robust helical edge transport in gated InAs/GaSbInAsGaSb\mathrm{InAs}/\mathrm{GaSb}roman_InAs / roman_GaSb bilayers, Phys. Rev. Lett. 114, 096802 (2015).
  • Bozkurt et al. [2018] A. M. Bozkurt, B. Pekerten, and İ. Adagideli, Work extraction and landauer’s principle in a quantum spin hall device, Physical Review B 97, 245414 (2018).
  • Bernevig et al. [2006] B. A. Bernevig, T. L. Hughes, and S.-C. Zhang, Quantum spin hall effect and topological phase transition in HgTe quantum wells, Science 314, 1757 (2006).
  • Girvin and Yang [2019] S. Girvin and K. Yang, Modern Condensed Matter Physics (Cambridge University Press, 2019).
  • Lu and Fertig [2014] C.-K. Lu and H. A. Fertig, Magnetic breakdown in twisted bilayer graphene, Phys. Rev. B 89, 085408 (2014).
  • Chapai et al. [2023] R. Chapai, M. Leroux, V. Oliviero, D. Vignolles, N. Bruyant, M. P. Smylie, D. Y. Chung, M. G. Kanatzidis, W.-K. Kwok, J. F. Mitchell, and U. Welp, Magnetic breakdown and topology in the kagome superconductor csv3sb5subscriptcsv3subscriptsb5{\mathrm{csv}}_{3}{\mathrm{sb}}_{5}roman_csv start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT roman_sb start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT under high magnetic field, Phys. Rev. Lett. 130, 126401 (2023).
  • Alexandradinata and Glazman [2017] A. Alexandradinata and L. Glazman, Geometric phase and orbital moment in quantization rules for magnetic breakdown, Phys. Rev. Lett. 119, 256601 (2017).
  • Lemut et al. [2020] G. Lemut, A. D. Vela, M. J. Pacholski, J. Tworzydło, and C. W. J. Beenakker, Magnetic breakdown spectrum of a kramers–weyl semimetal, New Journal of Physics 22, 093022 (2020).
  • Hasan and Kane [2010] M. Z. Hasan and C. L. Kane, Colloquium: Topological insulators, Rev. Mod. Phys. 82, 3045 (2010).
  • Qi and Zhang [2011] X.-L. Qi and S.-C. Zhang, Topological insulators and superconductors, Rev. Mod. Phys. 83, 1057 (2011).
  • Cooper and Moessner [2012] N. R. Cooper and R. Moessner, Designing topological bands in reciprocal space, Phys. Rev. Lett. 109, 215302 (2012).
  • Wang et al. [2015] J. Wang, B. Lian, and S.-C. Zhang, Quantum anomalous hall effect in magnetic topological insulators, Physica Scripta 2015, 014003 (2015).
  • Neupert et al. [2011] T. Neupert, L. Santos, C. Chamon, and C. Mudry, Fractional quantum hall states at zero magnetic field, Phys. Rev. Lett. 106, 236804 (2011).
  • Rachel [2016] S. Rachel, Quantum phase transitions of topological insulators without gap closing, Journal of Physics: Condensed Matter 28, 405502 (2016).
  • Ren et al. [2020] Y. Ren, Z. Qiao, and Q. Niu, Engineering corner states from two-dimensional topological insulators, Phys. Rev. Lett. 124, 166804 (2020).
  • Lunde and Platero [2013] A. M. Lunde and G. Platero, Hyperfine interactions in two-dimensional hgte topological insulators, Phys. Rev. B 88, 115411 (2013).
  • Seshadri et al. [2019] R. Seshadri, A. Dutta, and D. Sen, Generating a second-order topological insulator with multiple corner states by periodic driving, Phys. Rev. B 100, 115403 (2019).
  • Liu et al. [2010] C.-X. Liu, X.-L. Qi, H. Zhang, X. Dai, Z. Fang, and S.-C. Zhang, Model hamiltonian for topological insulators, Phys. Rev. B 82, 045122 (2010).
  • Asbóth et al. [2016] J. K. Asbóth, L. Oroszlány, and A. Pályi, Time-reversal symmetric two-dimensional topological insulators: The bernevig–hughes–zhang model, in A Short Course on Topological Insulators: Band Structure and Edge States in One and Two Dimensions (Springer International Publishing, Cham, 2016) pp. 119–138.
  • Nayak et al. [2008] C. Nayak, S. H. Simon, A. Stern, M. Freedman, and S. Das Sarma, Non-abelian anyons and topological quantum computation, Rev. Mod. Phys. 80, 1083 (2008).
  • Lohse et al. [2018] M. Lohse, C. Schweizer, H. M. Price, O. Zilberberg, and I. Bloch, Exploring 4d quantum hall physics with a 2d topological charge pump, Nature 553, 55 (2018).
  • Hu et al. [2021] B. Hu, Z. Zhang, H. Zhang, L. Zheng, W. Xiong, Z. Yue, X. Wang, J. Xu, Y. Cheng, X. Liu, and J. Christensen, Non-hermitian topological whispering gallery, Nature 597, 655 (2021).
  • Abanov [2017] A. Abanov, Topology, geometry and quantum interference in condensed matter physics, in Topology and Condensed Matter Physics, edited by S. M. Bhattacharjee, M. Mj, and A. Bandyopadhyay (Springer Singapore, Singapore, 2017) pp. 281–331.
  • Hamma et al. [2013] A. Hamma, L. Cincio, S. Santra, P. Zanardi, and L. Amico, Local response of topological order to an external perturbation, Phys. Rev. Lett. 110, 210602 (2013).
  • Volovik [2018] G. E. Volovik, Exotic lifshitz transitions in topological materials, Physics-Uspekhi 61, 89 (2018).
  • Li et al. [2022] Y. Li, A. Eaton, H. A. Fertig, and B. Seradjeh, Dirac magic and lifshitz transitions in aa-stacked twisted multilayer graphene, Phys. Rev. Lett. 128, 026404 (2022).
  • Chen et al. [2012] K.-S. Chen, Z. Y. Meng, T. Pruschke, J. Moreno, and M. Jarrell, Lifshitz transition in the two-dimensional hubbard model, Phys. Rev. B 86, 165136 (2012).
  • Lemonik et al. [2010] Y. Lemonik, I. L. Aleiner, C. Toke, and V. I. Fal’ko, Spontaneous symmetry breaking and lifshitz transition in bilayer graphene, Phys. Rev. B 82, 201408 (2010).
  • Akzyanov [2021] R. S. Akzyanov, Lifshitz transition in dirty doped topological insulator with nematic superconductivity, Phys. Rev. B 104, 224502 (2021).
  • Balents and Starykh [2016] L. Balents and O. A. Starykh, Quantum lifshitz field theory of a frustrated ferromagnet, Phys. Rev. Lett. 116, 177201 (2016).
  • Brando et al. [2016] M. Brando, D. Belitz, F. M. Grosche, and T. R. Kirkpatrick, Metallic quantum ferromagnets, Rev. Mod. Phys. 88, 025006 (2016).
  • Pelissetto and Vicari [2002] A. Pelissetto and E. Vicari, Critical phenomena and renormalization-group theory, Physics Reports 368, 549 (2002).
  • Amaricci et al. [2015] A. Amaricci, J. C. Budich, M. Capone, B. Trauzettel, and G. Sangiovanni, First-order character and observable signatures of topological quantum phase transitions, Phys. Rev. Lett. 114, 185701 (2015).
  • Juričić et al. [2017] V. Juričić, D. S. L. Abergel, and A. V. Balatsky, First-order quantum phase transition in three-dimensional topological band insulators, Phys. Rev. B 95, 161403 (2017).
  • [49] More information is available in the online Supplementary Material (SM).
  • Batista [2002] C. D. Batista, Electronic ferroelectricity in the falicov-kimball model, Phys. Rev. Lett. 89, 166403 (2002).
  • Freericks and Zlatić [2003] J. K. Freericks and V. Zlatić, Exact dynamical mean-field theory of the falicov-kimball model, Rev. Mod. Phys. 75, 1333 (2003).
  • Batista et al. [2004] C. D. Batista, J. E. Gubernatis, J. Bonča, and H. Q. Lin, Intermediate coupling theory of electronic ferroelectricity, Phys. Rev. Lett. 92, 187601 (2004).
  • Ihle et al. [2008] D. Ihle, M. Pfafferott, E. Burovski, F. X. Bronold, and H. Fehske, Bound state formation and the nature of the excitonic insulator phase in the extended falicov-kimball model, Phys. Rev. B 78, 193103 (2008).
  • Phan et al. [2011] V.-N. Phan, H. Fehske, and K. W. Becker, Excitonic resonances in the 2d extended falicov-kimball model, Europhysics Letters 95, 17006 (2011).
  • Golosov [2012] D. Golosov, Collective excitations and stability of the exciton phase in the extended falicov-kimball model, Physical Review B 86, 155134 (2012).
  • Zenker et al. [2012] B. Zenker, D. Ihle, F. X. Bronold, and H. Fehske, Electron-hole pair condensation at the semimetal-semiconductor transition: A bcs-bec crossover scenario, Phys. Rev. B 85, 121102 (2012).
  • Kaneko et al. [2013] T. Kaneko, S. Ejima, H. Fehske, and Y. Ohta, Exact-diagonalization study of exciton condensation in electron bilayers, Phys. Rev. B 88, 035312 (2013).
  • Ejima et al. [2014] S. Ejima, T. Kaneko, Y. Ohta, and H. Fehske, Order, criticality, and excitations in the extended falicov-kimball model, Phys. Rev. Lett. 112, 026401 (2014).
  • Farkašovský [2023] P. Farkašovský, Hartree-fock exploration of electronic ferroelectricity, valence transitions, and metal-insulator transitions in the extended falicov-kimball model, Phys. Rev. B 108, 075161 (2023).
  • Zheng et al. [1997] L. Zheng, M. W. Ortalano, and S. Das Sarma, Exchange instabilities in semiconductor double-quantum-well systems, Phys. Rev. B 55, 4506 (1997).
  • Hanna et al. [2000] C. B. Hanna, D. Haas, and J. C. Díaz-Vélez, Double-layer systems at zero magnetic field, Phys. Rev. B 61, 13882 (2000).
  • Cookmeyer and Das Sarma [2023] T. Cookmeyer and S. Das Sarma, Symmetry breaking in zero field 2d electron bilayers (2023), arXiv:2312.10791 [cond-mat.mes-hall] .
  • Zhu and Das Sarma [2023] J. Zhu and S. Das Sarma, Interaction and coherence in 2d bilayers (2023), arXiv:2312.10838 [cond-mat.str-el] .
  • Min et al. [2008] H. Min, R. Bistritzer, J.-J. Su, and A. H. MacDonald, Room-temperature superfluidity in graphene bilayers, Phys. Rev. B 78, 121401 (2008).
  • Seradjeh et al. [2009] B. Seradjeh, J. E. Moore, and M. Franz, Exciton condensation and charge fractionalization in a topological insulator film, Phys. Rev. Lett. 103, 066402 (2009).
  • Ezawa et al. [2013] M. Ezawa, Y. Tanaka, and N. Nagaosa, Topological phase transition without gap closing, Scientific Reports 3, 2790 (2013).
  • Yang et al. [2013] Y. Yang, H. Li, L. Sheng, R. Shen, D. N. Sheng, and D. Y. Xing, Topological phase transitions with and without energy gap closing, New Journal of Physics 15, 083042 (2013).
  • Pathria and Beale [2022] R. Pathria and P. D. Beale, Statistical Mechanics (Fourth Edition) (Academic Press, 2022).
  • Sachdev [2011] S. Sachdev, Quantum Phase Transitions, 2nd ed. (Cambridge University Press, 2011).
  • Sondhi et al. [1997] S. L. Sondhi, S. M. Girvin, J. P. Carini, and D. Shahar, Continuous quantum phase transitions, Rev. Mod. Phys. 69, 315 (1997).
  • Jotzu et al. [2014] G. Jotzu, M. Messer, R. Desbuquois, M. Lebrat, T. Uehlinger, D. Greif, and T. Esslinger, Experimental realization of the topological haldane model with ultracold fermions, Nature 515, 237 (2014).
  • Dao et al. [2012] T.-L. Dao, M. Ferrero, P. S. Cornaglia, and M. Capone, Mott transition of fermionic mixtures with mass imbalance in optical lattices, Phys. Rev. A 85, 013606 (2012).
  • Gröbner et al. [2017] M. Gröbner, P. Weinmann, E. Kirilov, H.-C. Nägerl, P. S. Julienne, C. R. Le Sueur, and J. M. Hutson, Observation of interspecies feshbach resonances in an ultracold K39133Cssuperscript133superscriptK39Cs{}^{39}\mathrm{K}-^{133}\mathrm{Cs}start_FLOATSUPERSCRIPT 39 end_FLOATSUPERSCRIPT roman_K - start_POSTSUPERSCRIPT 133 end_POSTSUPERSCRIPT roman_Cs mixture and refinement of interaction potentials, Phys. Rev. A 95, 022715 (2017).
  • Ajayan et al. [2016] P. Ajayan, P. Kim, and K. Banerjee, Two-dimensional van der Waals materials, Physics Today 69, 38 (2016)https://pubs.aip.org/physicstoday/article-pdf/69/9/38/10117753/38_1_online.pdf .
  • Young and Levitov [2011] A. F. Young and L. S. Levitov, Capacitance of graphene bilayer as a probe of layer-specific properties, Phys. Rev. B 84, 085441 (2011).
  • Hunt et al. [2017] B. M. Hunt, J. I. A. Li, A. A. Zibrov, L. Wang, T. Taniguchi, K. Watanabe, J. Hone, C. R. Dean, M. Zaletel, R. C. Ashoori, and A. F. Young, Direct measurement of discrete valley and orbital quantum numbers in bilayer graphene, Nature Communications 8, 948 (2017).
  • Moon et al. [1995] K. Moon, H. Mori, K. Yang, S. M. Girvin, A. H. MacDonald, L. Zheng, D. Yoshioka, and S.-C. Zhang, Spontaneous interlayer coherence in double-layer quantum hall systems: Charged vortices and kosterlitz-thouless phase transitions, Phys. Rev. B 51, 5138 (1995).
  • Eisenstein [2003] J. Eisenstein, Evidence for spontaneous interlayer phase coherence in a bilayer quantum hall exciton condensate, Sol. St. Comm. 127, 123 (2003).
  • Lutchyn et al. [2010] R. M. Lutchyn, E. Rossi, and S. Das Sarma, Spontaneous interlayer superfluidity in bilayer systems of cold polar molecules, Phys. Rev. A 82, 061604 (2010).
  • Murthy et al. [2017] G. Murthy, E. Shimshoni, and H. A. Fertig, Spin-valley coherent phases of the ν=0𝜈0\nu=0italic_ν = 0 quantum hall state in bilayer graphene, Phys. Rev. B 96, 245125 (2017).
  • Li et al. [2019] J. Li, H. Fu, Z. Yin, K. Watanabe, T. Taniguchi, and J. Zhu, Metallic phase and temperature dependence of the ν=0𝜈0\nu=0italic_ν = 0 quantum hall state in bilayer graphene, Phys. Rev. Lett. 122, 097701 (2019).