This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Origin of band flatness and constraints of higher Chern numbers

Alexander Kruchkov Department of Physics, Harvard University, Cambridge, MA 02138, USA Department of Theoretical Physics, University of Geneva, Geneva CH 1211, Switzerland Branco Weiss Society in Science, ETH Zurich, Zurich, CH 8092, Switzerland
Abstract

Flat bands provide a natural platform for emergent electronic states beyond Landau paradigm. Among those of particular importance are flat Chern bands, including bands of higher Chern numbers (CC>>11). We introduce a new framework for band flatness through wave functions, and classify the existing isolated flat bands in a ”periodic table” according to tight binding features and wave function properties. Our flat band categorization encompasses seemingly different classes of flat bands ranging from atomic insulators to perfectly flat Chern bands and Landau Levels. The perfectly flat Chern bands satisfy Berry curvature condition Fxy=Tr𝒢ijF_{xy}=\text{Tr}\,\mathcal{G}_{ij} which on the tight-binding level is fulfilled only for infinite-range models. Most of the natural Chern bands fall into category of C=1C=1; the complexity of creating higher-CC flat bands is beyond the current technology. This is due to the breakdown of the microscopic stability for higher-CC flatness, seen atomistically e.g. in the increase of the hopping range bound as \proptoCa\sqrt{C}a. Within our new formalism, we indicate strategies for bypassing higher-CC constraints and thus dramatically decreasing the implementation complexity.

Electronic band flatness is a new condensed matter paradigm. The flat bands had been studied in the context of Landau levels Landau (1930); *Landau1977; Haldane (2018); Kapit and Mueller (2010); however a conceptually new approach in material science is engineering flat bands from the first principles Miyahara et al. (2005); Maimaiti et al. (2017); Rhim and Yang (2019). To date, we have accumulated sufficient numerical evidence of flat bands formation ranging from artificial crystallographic lattices (Kagome, Lieb) to controllable flat band engineering in twisted van der Waals heterostructures Trambly de Laissardière et al. (2010); Naik and Jain (2018); Carr et al. (2020). The controllable and predictable engineering of band flatness is of strategical importance, especially for strongly correlated phases such as unconventional superconductivity Cao et al. (2018); Yankowitz et al. (2019); Park et al. (2021); Hao et al. (2021), Fractional Chern Insulators Neupert et al. (2011); Tang et al. (2011); Sun et al. (2011); Haldane (2011), and the SYK phase Kitaev (2015); Sachdev and Ye (1993); Chen et al. (2018), to name a few. Of particular importance are flat Chern bands, however the higher-order Chern numbers are a rare case in reality, with the majority of natural flat bands restricted to |C|=1|C|=1.

Table 1: ”Periodic table” for perfectly flat gapped bands categorized through criteria (1), (7), (13).
atomic insulator fine-tuned flat generic trivial nonlocal generic topological nonlocal Landau Level TBG chiral
0 𝒪(1)\mathcal{O}(1) \infty \infty \infty \infty \infty (\infty); 𝒪(1)\mathcal{O}(1) Hopping range Λ\Lambda
- none any π/a\sim\pi/a π/a\sim\pi/a π/a\sim\pi/a - - (cancelled; π/λM\pi/\lambda_{M}) Singularity position hh
0 0 0 |C|=12+12|C|=\frac{1}{2}+\frac{1}{2} |C|=m|C|=m |C|>1|C|>1 C=±1C=\pm 1 CC=±1\pm 1 Chern number CC
not defined double-periodic, nonholomorphic double-periodic, nonholomorphic double-periodic nonholomorphic meromorphic non-double-periodic double-periodic meromorphic holomorphic quasiperiodic holomorphic quasiperiodic Periodicity in BZ and analyticity

It is certainly surprising that in many cases we can point out on the common origin for perfectly flat bands, which is (self)-trapping in a limited coordinate region. In case of Landau levels with ψLL\psi_{\text{LL}}\simer2/2lB2e^{-r^{2}/2l_{B}^{2}}, the self-trapping occurs in the area \simlB2l_{B}^{2}; for atomic insulator the self-trapping occurs on the lattice sites; for fine-tuned flat-bands (such as in Kagome, Liebe lattices) the self-trapping happens within the plaquette \sima2a^{2}; in twisted van der Waals heterostructures the self-trapping happens at small part of the moiré cell around high-symmetry points (AA stacking in twisted bilayer graphene). This motivates us to define the band flatness parameter through wave functions

f=R>Λ|Ψ(R)|2R>0|Ψ(R)|2.\displaystyle\vspace{- 1 mm}f=\frac{\sum_{R>\Lambda}|\Psi(R)|^{2}}{\sum_{R>0}|\Psi(R)|^{2}}. (1)

Here Λ\Lambda is a system-dependent real space hopping range (dimensionless). The parameter ff\ll11 is identifying band flatness through wave function localization. Clearly δ(R)\delta(R) gives a perfectly flat band with ff==0, and the flatness construction at small Λ\Lambda\sim11 is intuitively clear. Let’s illustrate that definition (1) works also for Λ\Lambda\to\infty, both for the gapped trivial and topological bands.

Generic construction of flat bands.—Consider a generic tight binding Hamiltonian =ijΛtijαβciαcjβ\mathcal{H}=\sum_{ij}^{\Lambda}t_{ij}^{\alpha\beta}c^{{\dagger}}_{i\,\alpha}c^{{\dagger}}_{j\,\beta}. Here Λ\Lambda is the hopping range. Setting here lattice constant aa\to\infty (or tijt_{ij}==0) immediately recovers the ”atomic insulator”, a system with a perfectly flat band at E0E_{0}==0. The atomic insulator is the simplest perfectly flat band opening the ”periodic table” of perfectly flat bands (Table I). We can systematically build other classes of flat bands from the tight binding. In the momentum representation the generic tight-binding Hamiltonian reads =𝐤c𝐤ααβ(𝐤)c𝐤β\mathcal{H}=\sum_{\mathbf{k}}c^{{\dagger}}_{\mathbf{k}\alpha}\mathcal{H}_{\alpha\beta}(\mathbf{k})c_{\mathbf{k}\beta}. The matrix Hamiltonian αβ(𝐤)\mathcal{H}_{\alpha\beta}(\mathbf{k}) has electronic bands ε1(k),ε2(k)εN(k)\varepsilon_{1}(k),\varepsilon_{2}(k)...\varepsilon_{N}(k). Among them, we pick up the nnth{}^{\text{th}} band which is dispersive, gapped out from the others, and not crossing zero. We further introduce a new Hamiltonian hosting the perfectly flat band through procedure

tijTijflat=E0N𝐤(𝐤)εn(k)ei𝐤(𝐑i𝐑j).\displaystyle t_{ij}\ \ \to\ \ T^{\textbf{flat}}_{ij}=\frac{E_{0}}{N}\sum_{\mathbf{k}}\frac{\mathcal{H}(\mathbf{k})}{\varepsilon_{n}(k)}e^{i\mathbf{k}(\mathbf{R}_{i}-\mathbf{R}_{j})}. (2)

The new tight binding Hamiltonian with hopping (2) flat=ijTijflatciαcjβ\mathcal{H}^{\textbf{flat}}=\sum_{ij}^{\infty}T^{\textbf{flat}}_{ij}\,c^{{\dagger}}_{i\,\alpha}c^{{\dagger}}_{j\,\beta} contains at least one perfectly flat band positioned at E=E0E=E_{0}. However, this unique flat band comes at the expense that the tight-binding model becomes nonlocal (Λ\Lambda^{\prime}\to\infty). Of course, one can now take only the first few terms of TijflatT^{\textbf{flat}}_{ij}, since they are decaying exponentially fast with distance. However any real space truncation of Λ\Lambda inevitably introduces a ”parasitic dispersion”, so the band is no longer perfectly flat. The analysis above is useful for understanding the ”fine tuning” flat bands in Kagome, Lieb and other lattices. In the rare cases, one can find a lucky choice of fine-tuning parameter set upon truncation to nearest neighbors NN and NNN (Λ\Lambda\approx2,32,3), which gives flat or perfectly flat band through destructive interference. However, this construction is a fine tuning and inclusion of further-order hopping generically breaks flatness.

Clearly, the same band flattening procedure (2) can be applied to the Chern bands, as far as they are gapped. In this case the hopping parameters are generically complex-valued. For storytelling, consider the Haldane model Haldane (1988) on NN and NNN hoppings (Λ=2\Lambda=2),

FDMH=it0cici+ijtijNNcicj+ijtijNNNcicj,\displaystyle\mathcal{H}_{\text{FDMH}}=\sum_{i}t_{0}c^{{\dagger}}_{i}c_{i}+\sum_{\langle ij\rangle}t^{\text{NN}}_{ij}c^{{\dagger}}_{i}c^{{\dagger}}_{j}+\sum_{\langle\langle ij\rangle\rangle}t^{\text{NNN}}_{ij}c^{{\dagger}}_{i}c^{{\dagger}}_{j}, (3)

where we fix tijNNt^{\text{NN}}_{ij} as real and tijNNNt^{\text{NNN}}_{ij}==teiΦijt^{\prime}e^{i\Phi_{ij}}. Clearly, for Φij\Phi_{ij}==±π/2\pm\pi/2, the spectrum becomes particle-hole symmetric Haldane (1988). This yields upon transformation (1) two pefectly flat bands positioned at EE==±E0\pm E_{0}. The resulting FDMHflat\mathcal{H}_{\text{FDMH}}^{\textbf{flat}} becomes nonlocal (Λ\Lambda^{\prime}==\infty), though it is possible to truncate it and make the band arbitrary flat by choosing a corresponding truncation Λ\Lambda^{\prime} and minimizing bandwidth (see e.g. Ref.Neupert et al. (2011)). Thus it is always possible to construct a perfectly flat band for Λ\Lambda==\infty, be it topological or not. Since the range Λ\Lambda of effective tight-binding is related to the wave function’s spatial tails, the above statement is consistent with the flatness parameter (1).

Band flatness for topologically-trivial bands. Addressing localization properties is more natural in terms of Wannier functions. Topologically-trivial flat bands are maximally Wannierizable, thus without loss of generality we consider localization along one of the two spatial dimensions and investigate its asymptotics Kohn (1959); He and Vanderbilt (2001)

𝒲(xR)𝑑kxeikx(xR)uk.\displaystyle\mathcal{W}(x-R)\propto\int dk_{x}\,e^{ik_{x}(x-R)}\,u_{k}. (4)

A topologically trivial band may have a singularity in the form of a branch vertex in complex momentum k=kx+ihxk=k_{x}+ih_{x}, of a generic form u(kk)u0[i(kk)]αu(k\approx k_{*})\simeq u_{0}[i(k-k_{*})]^{\alpha}, with α\alpha>>1-1. Denote k=k0+ihk_{*}=k_{0}+ih as the position of the singularity in the complex plain. Hence the asymptotics of (4) reduces to integral representation of the Gamma function along a Hankel contour encompassing k0+ihk_{0}+ih. The result for large xx is

𝒲(x)2u0sin(πα)Γ(1+α)exp(hx)x1+α,\displaystyle\mathcal{W}(x)\simeq 2u_{0}\sin(\pi\alpha)\Gamma(1+\alpha)\frac{\exp(-hx)}{x^{1+\alpha}}, (5)

In case of several singularities we take h=min[Imk]h=\text{min}[\,\text{Im}\,k_{*}], corresponding to the one closest to the real axis. We further use this asymptotic to derive the behavior of the flatness parameter (1) as

f1Λ2(α+1)e2hΛa,\displaystyle f\sim\frac{1}{\Lambda^{2(\alpha+1)}}e^{-2h\Lambda a}, (6)

For this, we need to use analytical estimates of the sums of form Σn(Λ)\Sigma_{n}(\Lambda)==x=Λxnex\sum_{x=\Lambda}^{\infty}x^{-n}e^{-x}. We rewrite this sum as Σn(Λ)\Sigma_{n}(\Lambda)==eΛζ(i2π,n,Λ)e^{-\Lambda}\,\zeta(\frac{i}{2\pi},n,\Lambda) where ζ(ϕ,n,Λ)=x=0(x+Λ)ne2πiϕx\zeta(\phi,n,\Lambda)=\sum_{x=0}^{\infty}(x+\Lambda)^{-n}e^{2\pi i\phi x} is Lerch zeta function (see e.g. Apostol (1951); *Johnson1974). Up to 𝒪(1)\mathcal{O}(1) prefactor Lerch zeta function ζ(i2π,n,Λ)\zeta(\frac{i}{2\pi},n,\Lambda) behaves as 1/Λn1/\Lambda^{n} for Λ\Lambda\gg11, thus we obtain analytical estimate Σn(Λ)\Sigma_{n}(\Lambda)\simΛneΛ\Lambda^{-n}e^{-\Lambda} and Σn(1)\Sigma_{n}(1)\sim𝒪(1/e)\mathcal{O}(1/e). The flatness parameter involves summation of form Σn(Λ+1)/Σn(1)\Sigma_{n}(\Lambda+1)/\Sigma_{n}(1)\sim(Λ+1)neΛ(\Lambda+1)^{-n}e^{-\Lambda}. We confirm numerically that this estimate works good even for Λ\Lambda\sim22. Taking now n=2(α+1)n=2(\alpha+1) and restoring dimensional parameters, we obtain (6). For our purposes, we are not interested here in the power-law prefactor, and factor of 2 in the exponent. It is safe to rewrite the flatness criterion as

f0=ehΛa,for trivial bands.\displaystyle f_{0}=e^{-h\Lambda a},\ \ \ \text{for trivial bands. } (7)

The flatness parameter f0f_{0} of (7) sets a fundamental scale for achievable band flatness, and covers three distinguished classes of perfectly flat nontopological bands with f0=0f_{0}=0:

  • aa\to\infty, atomic insulator.

  • Λ\Lambda\to\infty, generic flat band construction, Eq. (2).

  • hh\to\infty, singularity removed to infinity (nonsingular perfectly flat band).

The first two cases are discussed on page 1; the examples for nonsingular perfectly flat bands are listed in Ref. Rhim and Yang (2019). Cases of topologically trivial, gapped perfectly flat bands are covered by the three classes above, and constitute the topologically-trivial sector of the flat band classification (Table I). We do not have evidence of perfectly flat, gapped topologially-trivial bands which do not fit into this classification. We now proceed to the Chern bands.

Band flatness for Chern bands. We start from the construction of higher-CC Chern bands on the basis of double periodic meromorphic functions. The essential toolbox is built upon implementation of theta functions, Weierstrass and Jacobi functions, and their combinations Akhiezer (sian). Independently of the choice of the elliptic function, we can use connection between the wave function singularities (poles) and the band Chern number CC==BZd2𝐤2πFxy\int_{\text{BZ}}\frac{d^{2}\mathbf{k}}{2\pi}F_{xy}, with FxyF_{xy}==xAy\partial_{x}A_{y}-yAx\partial_{y}A_{x}, 𝐀𝐤\mathbf{A}_{\mathbf{k}}==iu𝐤|𝐤u𝐤-i\langle u_{\mathbf{k}}|\partial_{\mathbf{k}}u_{\mathbf{k}}\rangle), in the complex plane zz==(kxk_{x},kyk_{y}) Jian et al. (2013); Baum and Bott (1970)

C=[γBZ+ziγzi]Az¯dz+Azdz¯4π=zipi(zi).\displaystyle C=\left[\oint_{\gamma_{\text{BZ}}}+\sum_{z^{*}_{i}}\oint_{\gamma_{z_{i}^{*}}}\right]\frac{A_{\bar{z}}dz+A_{z}d\bar{z}}{4\pi}=\sum_{z^{*}_{i}}p_{i}(z_{i}^{*}). (8)

The Chern number is expressed through the sum of all poles ziz^{*}_{i} in Brilloin zone (BZ), counting their multiplicity pi(zi)p_{i}(z^{*}_{i}) Baum and Bott (1970). The definition (8) is important since it is not connected to the band dispersion as such.

A theorem, tracing back to Thouless Thouless (1984), prevents Wannierizing Chern bands in 2D; see recent discussion in Po et al. (2018). However, as was pointed out by Qi, it does not prevent Wannierizing a Chern band along one of the 1D directions of a 2D Chern insulator Qi (2011). (Moreover, in this way a duality between a C=1C=1 flat Chern band and the Lowest Landau level is established Qi (2011)).

For our purposes, localization along 1D is a good indicator of band flatness through (1), (4). We thus proceed with Wannierizing a Chern band along 1D, and finding its asymptotic behavior. For this, it is sufficient to replace the elliptic functions with their principal behavior around poles

u(k)nu0[i(kkn)]pn+Regular part.\displaystyle u(k)\simeq\sum_{n}\frac{u_{0}}{[i(k-k^{*}_{n})]^{p_{n}}}+\text{Regular part}. (9)

(here k=kx+ikyk=k_{x}+ik_{y} is in the first BZ). The associated Chern number is C=npnC=\sum_{n}p_{n}. The main contribution to the integral (4) is given by the pole (9) of multiplicity pnC/2p_{n}\leq C/2 closest to the the real axis. The residue at the pole is Resu(k)=iu0xpn1eixk/(pn1)!\text{Res}\,u(k)=-iu_{0}x^{p_{n}-1}e^{ixk_{*}}/(p_{n}-1)!, with k=k0+ihk_{*}=k_{0}+ih. Using the residue theorem, and an appropriate contour, one obtains asymptote

𝒲(x)2πu0(pn1)!xpn1ehx.\displaystyle\mathcal{W}(x)\simeq\frac{2\pi u_{0}}{(p_{n}-1)!}x^{p_{n}-1}e^{-hx}. (10)

To derive the flatness parameter, we need to evaluate sums of form Σm(Λ)=x>Λxmex\Sigma_{m}(\Lambda)=\sum_{x>\Lambda}^{\infty}x^{m}e^{-x} with mm==2(pn1)2(p_{n}-1). We can rewrite this sum through Lerch zeta function as Σm(Λ)=eΛζ(i2π,m,Λ)\Sigma_{m}(\Lambda)=e^{-\Lambda}\zeta(\frac{i}{2\pi},-m,\Lambda). Up to 𝒪(1)\mathcal{O}(1) prefactor Lerch zeta function ζ(i2π,m,Λ)\zeta(\frac{i}{2\pi},-m,\Lambda) behaves as Λm\Lambda^{m} for Λ\Lambda\gg11, thus Σm(Λ)\Sigma_{m}(\Lambda)\simΛmeΛ\Lambda^{m}e^{-\Lambda} and Σm(1)\Sigma_{m}(1)\sim𝒪(1/e)\mathcal{O}(1/e). Restoring dimensional units, we obtain the flatness criterion as fΛme2haΛf\sim\Lambda^{m}e^{-2ha\Lambda}, where m=2(pn1)m=2(p_{n}-1), i.e. depends on the nature of the wave function singularities. We confirm numerically that this is a good fit 111Numerical analysis gives [(Λ+Λ1)/Λ2]me2haΛ[(\Lambda+\Lambda_{1})/\Lambda_{2}]^{m}e^{-2ha\Lambda} with Λ1Λ2\Lambda_{1}\approx\Lambda_{2}..

Refer to caption
Figure 1: Flatness of higher-Chern bands and topological obstructions. (a) Band flatness bound of (1) vs. electronic hopping range Λ\Lambda for Chern bands with C=2,4,8C=2,4,8. Solid line represents asymptote (11) [31]; 2ha2ha==11 is set for all data points. Chern bands can be perfectly flat only for non-local hopping (Λ\Lambda\to\infty). (b) Flatness bound at NN hoping represents obstruction to band flatness such that inclusion of further-order hoppings is required (Fig.2). In case of Chern bands, fNN(C)f_{\text{NN}}(C) quickly saturates to the maximally-obstructed at C(h)C_{*}(h) (Fig.1b); in case of trivial bands fNNf_{\text{NN}} lower bound can be set to zero (7). This explains why most of natural flat Chern bands are limited to unit |C||C|=11.

We show that the finite Chern number inevitably leads to constraints on the band flatness (Fig.1). We note that high CC==NN in (8) can be attained in multiple ways. The simplest way is by having two poles of multiplicity NN each 222Double periodic meromorphic function cannot have less than two zeros, two poles counting multiplicity. This results into a higher Chern number C2NC_{2N}==2N2N\gg11 restraining band flatness as ff\simΛC2N2e2haΛ\Lambda^{C_{2N}-2}e^{-2ha\Lambda}. To have a topological band, the wave function singularity must reside inside the BZ. This leads to the limitation hh\leqπ/a\pi/a (square lattice), or haha\sim11 independently of aa and lattice symmetries. The flatness parameter is

fΛC2N2eΛ.\displaystyle f\sim\Lambda^{C_{2N}-2}e^{-\Lambda}. (11)

The only way to have a perfectly flat Chern band (ff==0) is for Λ\Lambda\to\infty. We come to the same conclusion if we consider a higher CC==NN Chern band constructed through adding NN poles of multiplicity 11 (or any other combination). In this case we need to replace C2NC_{2N} with 22×\timesC1C_{1}, C1C_{1}==11 in (11), however the same argument holds: the Chern band is perfectly flat only for Λ\Lambda\to\infty. We cannot fine-tune local tight-binding with NN and NNN hoppings in order to have perfectly flat Chern band on the lattice, as we did in the topologically-trivial case. The band flatness constraints in this case results from the inevitable singularity of the wave functions of form 1/(kk)α1/(k-k_{*})^{\alpha}, which is irremovable by definition, and is the source for the Berry curvature.

Thus we come to the conclusion that the Chern bands cannot be perfectly flat unless in the case of infinite hopping range Λ\Lambda. This observation is consistent with Chen theorem (2014), stating that in the double-periodic system the perfectly flat Chern bands are accessible only for Λ=\Lambda=\infty Chen et al. (2014).

Flat Chern bands with C=1.—Now we can relax some of the conditions imposed in Eqs. (9), (8) on the flat band wave function. This can be done in two ways: by relaxing double periodicity or by relaxing holomorphicity conditions. First, we can relax condition of double-periodicity, but still require the flat band state to be a function of zz==kxk_{x}++ikyik_{y}. In this way we unlock all the odd Chern numbers in (8), including unitary |C|=1|C|=1. In this case the contribution along the BZ boundary, which vanishes due to double-periodicity in (8), may itself contribute to the Chern number. Interestingly, we can go even further by canceling the remaining singularity(-ies) inside the BZ and producing an effective magnetic flux; in this case the Chern number is given purely by the circulation along the BZ boundary γ𝐀𝑑𝐤=2πC\oint_{\gamma}\mathbf{A}d\mathbf{k}=2\pi C Zak (1989). This case corresponds to the continuum model of twisted bilayer graphene (TBG), which hosts perfectly flat Chern bands at the magic angle San-Jose et al. (2012); Tarnopolsky et al. (2019). Several authors has pointed out on the duality between the perfectly flat Chern bands in TBG and the lowest Landau level; we refer to Refs.San-Jose et al. (2012); Tarnopolsky et al. (2019); Popov and Milekhin (2021) for further details. In both cases we are dealing with effective magnetic fields which produce Berry curvature FxylB2F_{xy}\propto l_{B}^{2}, flux 2π2\pi through effective Brillouin zone (MBZ) and a perfectly flat band in a certain limit. Following Qi (2011), without loss of generality we can consider asymptote 𝒲(x)xnex2/2lB2\mathcal{W}(x)\propto x^{n}e^{-x^{2}/2l_{B}^{2}}. Thus for Landau Levels the flatness criterion (1) asymptotically reads

fLLΛ2n1eΛ2a2/lB2,\displaystyle f_{\text{LL}}\sim\Lambda^{2n-1}e^{-\Lambda^{2}a^{2}/l_{B}^{2}}, (12)

where we assume aa\lllBl_{B}. We see that the Landau levels are perfectly flat only in the nonlocal limit Λ\Lambda\gglB/al_{B}/a\to\infty. Bringing this system on the tight-binding lattice (finite aa, finite Λ\Lambda) inevitably broadens the Landau levels for any finite Λ\Lambda, see Refs.Hofstadter (1976); Kapit and Mueller (2010); Dong and Mueller (2020).

We can further relax the condition of holomorphicity, but keep the condition of double periodicity. This correspond to the Chern bands constructed from the pairs of Dirac points. A simplest construction is C=1/2+1/2C=1/2+1/2 Chern band constructed from two Dirac cones carrying Chern number |C|=1/2|C|=1/2 each. The total Chern number is integer since the Dirac nodes always come in pairs on the lattice. Since the normalized wave function associated with a (gapped) Dirac cone contains singularity of form u(k)u(k)\sim(kk)1/4(k-k_{*})^{-1/4}, the Wannier asymptotics is of form 𝒲(x)\mathcal{W}(x)\simx3/4ehxx^{-3/4}e^{-hx}. Since we again have restriction haha\sim11, we arrive to the flatness criterion of form fΛ3/2eΛf\sim\Lambda^{-3/2}e^{-\Lambda}. Once again we cannot have a perfectly flat band at local tight binding. Note that we can also construct higher Chern number C=NC=N by taking at least 2N2N Dirac cones and melting them into a flattened band (a strategy proved to be fruitful in graphene multilayers). This case completes our discussion on flatness of Chern bands, since further dropping both the condition of periodicity and holomorphicity breaks down the concept of Chern number as an integer topological invariant on the tight-binding level.

Independently of the method the flat Chern band has been composed, it is safe to rewrite the flatness criterion as

f0=ΛS[C]eF(Λ),for topological bands,\displaystyle f_{0}=\Lambda^{S[C]}e^{-F(\Lambda)},\ \ \ \text{for topological bands}, (13)

where S[C]|C|S[C]\propto|C|, F(Λ)Λ,Λ2F(\Lambda)\propto\Lambda,\Lambda^{2} with system-dependent coefficients which reflect the way the Chern number CC is constructed. The higher Chern number is, the harder is to make a good flat band. The topological bands can be perfectly flat only at infinite hopping range.

Comparing criterion (13) with the f0f_{0} criterion for trivial bands (7), we notice that criterion (13) is not reducible to the atomic insulator. This is because the Chern insulator and atomic insulator belong to different topological classes, and cannot be adiabatically connected Chiu et al. (2016).

Microscopic analysis and hopping range bounds.—We now look deeper into the microscopic details of topological constraints (Fig. 2). A representative parameter is the hopping range scale, defined for an isolated band as

rhop=(i,j=0(xixj)2𝒲(xi)𝒲(xj)i,j=0𝒲(xi)𝒲(xj))1/2.\displaystyle r_{\text{hop}}=\left(\frac{\sum\limits_{i,j=0}^{\infty}(x_{i}-x_{j})^{2}\mathcal{W}(x_{i})\mathcal{W}(x_{j})}{\sum\limits\limits_{i,j=0}^{\infty}\mathcal{W}(x_{i})\mathcal{W}(x_{j})}\right)^{1/2}.

For definiteness, we use the construction of higher-Chern flat band as in Eqs.(10-11). We observe numerically that rhop2r_{\text{hop}}^{2} behaves monotonically with h~\tilde{h}==haha; for haha\lesssim11 we have rhop2r_{\text{hop}}^{2}\sim1/h~21/\tilde{h}^{2}; rhop2r_{\text{hop}}^{2}\simC\,C. To understand this asymptotics analytically, we replace sum with integrals in definition above, rhop2r_{\text{hop}}^{2}\simeqh2I[pnh^{-2}I[p_{n}-1,2]/I[pn1,2]/I[p_{n}-1,0]1,0], with I[q,s]I[q,s]==0dx1dx2(x1x2)q(x1\iint\limits_{0}^{\infty}dx_{1}dx_{2}(x_{1}x_{2})^{q}(x_{1}-x2)se(x1+x2)x_{2})^{s}e^{-(x_{1}+x_{2})}. We arrive to the analytical expression rhop2r_{\text{hop}}^{2}==2Γ(pn+1)/h2Γ(pn)2\Gamma(p_{n}+1)/h^{2}\Gamma(p_{n}). Since in this case pnp_{n} are integers, we have rhop2r_{\text{hop}}^{2}==2pn/h22p_{n}/h^{2}. Restoring connection between the poles and the Chern number as in Eq.(11), we obtain

rhop1hC.\displaystyle r_{\text{hop}}\simeq\frac{1}{h}\sqrt{C}. (14)

Hence the higher-CC bands require longer-range hopping.

The wave function singularity position hh=hmaxh_{\text{max}} results into the lower bound for hopping range (14). For the Chern bands, it is impossible to remove this singularity to infinity, thus hmaxh_{\text{max}} is bounded from above. For a square lattice, the first estimate on hmaxh_{\text{max}} gives as π/a\pi/a. Independently of lattice symmetries, we can use hmaxah_{\text{max}}a\sim11, hence the lower bound for hopping range in case of topological bands scales as (Fig.2a)

rhopCa.\displaystyle r_{\text{hop}}\simeq\sqrt{C}a. (15)

This agrees with Ref.Jian et al. (2013). Note that the actual hopping range cutoff Λ\Lambda_{*} required to stabilize the band of desirable flatness ff_{*} with respect to next-order hoppings can be much higher than the lower bound (15) scaling as \proptoCaCa (Fig.2b).

Refer to caption
Figure 2: Microscopical analysis of topological constraints on the band flatness. (a) Characteristic hopping range scale rhopr_{\text{hop}} demonstrates the lower bound \simeqCa\sqrt{C}a, Eqs.(14-15). This lower bound is fundamental Jian et al. (2013). (b) Required tight-binding cut-off Λ\Lambda_{*} to stabilize band flatness ff_{*}\ll11 is higher than the hopping range lower bound r0r_{0}. For all data points, we take haha==11 and f(Λ)f_{*}(\Lambda_{*})==0.10.1. For higher-Chern flat bands, both (a,b) require precise experimental control of multiple hopping parameters beyond NN and NNN, thus restricting realizations of |C||C|\gg11 flat bands. The strategy to bypass this limitation is proposed in the main text.

Bypassing higher Chern constraints.—We notice that the microscopic constraints of form (14),(15), which were derived assuming higher-order poles of multiplicity NN==C/2C/2, could be bypassed in the experiment. Instead, we can take (a physically reasonable number) NN of simple poles to construct CC==NN×\timesC1C_{1} flat band using C1C_{1}==11. In this case 𝒲(x)\mathcal{W}(x)\simxC11ehaxx^{C_{1}-1}e^{-hax}, hence we need to replace CC with C1C_{1} in (15), and the lower bound C1\sqrt{C_{1}}aaa\sim a remains fixed on the lattice size for all the Chern numbers. Consider a thin film material (e.g. a monolayer or bilayer) with its band structure hosting a flat Chern band of unit |C1||C_{1}|==11. We can stack NN such layers into a multilayer heterostructure, represented by matrix Hamiltonian \mathcal{H} with monolayer Hamiltonians on the main diagonal and interlayer coupling terms placed on the diagonals just below and above. The topological invariant of this composition can be found through computing C[𝔾]C[\mathbb{G}]==13!ijkϵijkTr𝔾i𝔾1𝔾j𝔾1𝔾k𝔾1,\frac{1}{3!}\sum_{ijk}\epsilon_{ijk}\text{Tr}\,\mathbb{G}\partial_{i}\mathbb{G}^{-1}\mathbb{G}\partial_{j}\mathbb{G}^{-1}\mathbb{G}\partial_{k}\mathbb{G}^{-1}, where i,j,ki,j,k==ω\omega,kxk_{x},kyk_{y} and 𝔾=(ω)1\mathbb{G}=(\omega-\mathcal{H})^{-1} is matrix Green’s function Gurarie (2011). Consider a case of vanishingly weak interlayer interaction; in this case the Hamiltonian approximately factorizes into matrix tensor product. If the Chern number, associated with a single layer is C1C_{1}, the NN layers in this case give

C[𝔾N]=N×C1,rhopa.\displaystyle C[\mathbb{G}_{N}]=N\times C_{1},\ \ \ \to\ \ \ r_{\text{hop}}\sim a. (16)

This argument is expected to be valid when the weak interlayer interactions are switched on while preserving the finite gap; a similar algorithm was implemented by Trescher and Bergholtz Trescher and Bergholtz (2012). Importantly, the construction (16) does not obstruct the band flatness, i.e. the flat Chern bands can be constructed for any number of layers NN. Clearly, this argument is also valid for constructing higher Chern numbers by bringing together 2n2n (gapped) Dirac cones in van der Waals multilayers for reaching CC==1,2,31,2,3... nearly flat Chern bands Zhang et al. (2019).

Quantum geometry and flatness constraints.—Finally, we make connection between the band geometry and flatness criterion (1). The band topology and geometry is described by the ”quantum geometric” tensor Roy (2014); Jackson et al. (2015):

𝔊ij=iu𝐤|(1|u𝐤u𝐤|)|ju𝐤,\displaystyle\mathfrak{G}_{ij}=\langle\partial_{i}u_{\mathbf{k}}|\left(1-{|u_{\mathbf{k}}\rangle\langle u_{\mathbf{k}}|}\right)|\partial_{j}u_{\mathbf{k}}\rangle,

where i=ki\partial_{i}=\frac{\partial}{\partial k_{i}}. The imaginary part of 𝔊\mathfrak{G} is responsible for topology, and gives (off-diagonal) Berry curvature Fij=Im𝔊ijF_{ij}=\text{Im}\mathfrak{G}_{ij}; the real part 𝒢ij=Re𝔊ij\mathcal{G}_{ij}=\text{Re}\mathfrak{G}_{ij} is Fubini-Study metrics and is responsible for the band geometry and its flatness. The ideal flat Chern bands satisfy the Berry-geometric condition Fxy=Tr𝒢ijF_{xy}=\text{Tr}\,\mathcal{G}_{ij} (see Refs.Haldane (2011); Roy (2014); Claassen et al. (2015)). The holomorphic (and meromorphic) perfectly flat Chern bands automatically satisfy this criterion 333The lowest Landau level automatically satisfies FxyF_{xy}==Tr𝒢ij\text{Tr}\,\mathcal{G}_{ij}; a delicate moment here is discussion of the higher Landau levels which at first glance violate this condition. However, since the wave functions of the higher Landau levels can be chosen holomorphic Haldane (2018); Chen and Biswas (2020), surprisingly this argument still holds. . This identity requires Λ=\Lambda=\infty, thus consistent with our classification of Table I. We can further rewrite

Fxy=Tr𝒢ij=u𝐤||𝐫^|2|u𝐤,\displaystyle F_{xy}=\text{Tr}\,\mathcal{G}_{ij}=\langle u_{\mathbf{k}}|\,|\hat{\mathbf{r}}|^{2}\,|u_{\mathbf{k}}\rangle, (17)

where r^\hat{r} is the generalized position operator. Integrating (17) over Brillouin zone, one obtains r02r_{0}^{2}\proptoCC, hence the localization length is r0r_{0}\simCa\sqrt{C}a. For the Chern bands it is impossible to minimize localization length r0r_{0} independently from the hopping range bound (15); thus the flatness parameter (1) cannot be made arbitrary small for any finite Λ\Lambda, there are always finite tails. The Chern bands can be made perfectly flat only for Λ=\Lambda=\infty, as was demonstrated through Eq. (13). This is an intuitive interpretation of Chen theorem Chen et al. (2014). Clearly now, the higher Berry fluxes in (17), hence the higher Chern numbers |C||C|>>11, present stronger constraints on electronic band flatness.

Conclusions.—To conclude, the new criterion for band flatness (1) allows us to derive two concise cases f0f_{0}==ehaΛe^{-ha\Lambda} for trivial bands and f0f_{0}==ΛS[C]eF[Λ]\Lambda^{S[C]}e^{-F[\Lambda]} for topological bands of Chern number CC. The criteria allows us to systematize the known classes of perfectly flat bands (Table I) as the fundamental building blocks for realistic nearly flat bands. The large Chern number presents an obstruction to the band flatness, which can be seen in e.g. in the required range \propto Ca\sqrt{C}a. The most feasible route for overcoming this obstruction is using multilayers for gluing together higher Chern numbers. We can hope that the strategy for building higher-CC bands with proposed pathways (16) shall become a route towards realizing elusive Kitaev E8E_{8} state, a bosonic Quantum Hall phase which can be build on the CC==8n8n flat band Kitaev (2006).

\bigodot

Acknowledgements. The author thanks Bertrand Halperin, Subir Sachdev, Philip Kim, and Emil Bergholtz for enlightening discussions. This work was financed by the Branco Weiss Society in Science, ETH Zurich, through the research grant on flat bands, strong interactions and the SYK physics.

References

  • Landau (1930) L. Landau, Zeitschrift für Physik 64, 629 (1930).
  • Landau and Lifshitz (1977) L. D. Landau and E. M. Lifshitz, Quantum mechanics: non-relativistic theory (Pergamon Press, 1977).
  • Haldane (2018) F. Haldane, Journal of Mathematical Physics 59, 081901 (2018).
  • Kapit and Mueller (2010) E. Kapit and E. Mueller, Physical Review Letters 105, 215303 (2010).
  • Miyahara et al. (2005) S. Miyahara, K. Kubo, H. Ono, Y. Shimomura,  and N. Furukawa, Journal of the Physical Society of Japan 74, 1918 (2005).
  • Maimaiti et al. (2017) W. Maimaiti, A. Andreanov, H. C. Park, O. Gendelman,  and S. Flach, Physical Review B 95, 115135 (2017).
  • Rhim and Yang (2019) J.-W. Rhim and B.-J. Yang, Phys. Rev. B 99, 045107 (2019).
  • Trambly de Laissardière et al. (2010) G. Trambly de Laissardière, D. Mayou,  and L. Magaud, Nano Letters 10, 804 (2010).
  • Naik and Jain (2018) M. H. Naik and M. Jain, Physical Review Letters 121, 266401 (2018).
  • Carr et al. (2020) S. Carr, C. Li, Z. Zhu, E. Kaxiras, S. Sachdev,  and A. Kruchkov, Nano Letters 20, 3030 (2020).
  • Cao et al. (2018) Y. Cao, V. Fatemi, S. Fang, K. Watanabe, T. Taniguchi, E. Kaxiras,  and P. Jarillo-Herrero, Nature 556, 43 (2018).
  • Yankowitz et al. (2019) M. Yankowitz, S. Chen, H. Polshyn, Y. Zhang, K. Watanabe, T. Taniguchi, D. Graf, A. F. Young,  and C. R. Dean, Science 363, 1059 (2019).
  • Park et al. (2021) J. M. Park, Y. Cao, K. Watanabe, T. Taniguchi,  and P. Jarillo-Herrero, Nature 590, 249 (2021).
  • Hao et al. (2021) Z. Hao, A. Zimmerman, P. Ledwith, E. Khalaf, D. H. Najafabadi, K. Watanabe, T. Taniguchi, A. Vishwanath,  and P. Kim, Science 371, 1133 (2021).
  • Neupert et al. (2011) T. Neupert, L. Santos, C. Chamon,  and C. Mudry, Physical Review Letters 106, 236804 (2011).
  • Tang et al. (2011) E. Tang, J.-W. Mei,  and X.-G. Wen, Physical Review Letters 106, 236802 (2011).
  • Sun et al. (2011) K. Sun, Z. Gu, H. Katsura,  and S. D. Sarma, Physical Review Letters 106, 236803 (2011).
  • Haldane (2011) F. Haldane, Physical Review Letters 107, 116801 (2011).
  • Kitaev (2015) A. Kitaev, Talk at KITP, University of California, Santa Barbara  (2015).
  • Sachdev and Ye (1993) S. Sachdev and J. Ye, Physical Review Letters 70, 3339 (1993).
  • Chen et al. (2018) A. Chen, R. Ilan, F. De Juan, D. Pikulin,  and M. Franz, Physical Review Letters 121, 036403 (2018).
  • Haldane (1988) F. D. M. Haldane, Physical Review Letters 61, 2015 (1988).
  • Kohn (1959) W. Kohn, Physical Review 115, 809 (1959).
  • He and Vanderbilt (2001) L. He and D. Vanderbilt, Physical Review Letters 86, 5341 (2001).
  • Apostol (1951) T. M. Apostol, Pacific Journal of Mathematics 1, 161 (1951).
  • Johnson (1974) B. Johnson, Pacific Journal of Mathematics 53, 189 (1974).
  • Akhiezer (sian) N. I. Akhiezer, Elements of the theory of elliptic functions (American Mathematical Society, 1990; Gos.Tech.Izdat. 1948 [in Russian]).
  • Jian et al. (2013) C.-M. Jian, Z.-C. Gu,  and X.-L. Qi, Physica Status Solidi RRL 7, 154 (2013).
  • Baum and Bott (1970) P. F. Baum and R. Bott, in Essays on Topology and related topics (Springer, 1970) pp. 29–47.
  • Thouless (1984) D. Thouless, Journal of Physics C: Solid State Physics 17, L325 (1984).
  • Po et al. (2018) H. C. Po, H. Watanabe,  and A. Vishwanath, Phys. Rev. Lett. 121, 126402 (2018).
  • Qi (2011) X.-L. Qi, Physical Review Letters 107, 126803 (2011).
  • Note (1) Numerical analysis gives [(Λ+Λ1)/Λ2]me2haΛ[(\Lambda+\Lambda_{1})/\Lambda_{2}]^{m}e^{-2ha\Lambda} with Λ1Λ2\Lambda_{1}\approx\Lambda_{2}.
  • Note (2) Double periodic meromorphic function cannot have less than two zeros, two poles counting multiplicity.
  • Chen et al. (2014) L. Chen, T. Mazaheri, A. Seidel,  and X. Tang, Journal of Physics A 47, 152001 (2014).
  • Zak (1989) J. Zak, Physical Review Letters 62, 2747 (1989).
  • San-Jose et al. (2012) P. San-Jose, J. Gonzalez,  and F. Guinea, Physical Review Letters 108, 216802 (2012).
  • Tarnopolsky et al. (2019) G. Tarnopolsky, A. J. Kruchkov,  and A. Vishwanath, Physical Review Letters 122, 106405 (2019).
  • Popov and Milekhin (2021) F. K. Popov and A. Milekhin, Physical Review B 103, 155150 (2021).
  • Hofstadter (1976) D. R. Hofstadter, Physical Review B 14, 2239 (1976).
  • Dong and Mueller (2020) J. Dong and E. J. Mueller, Physical Review A 101, 013629 (2020).
  • Chiu et al. (2016) C.-K. Chiu, J. C. Teo, A. P. Schnyder,  and S. Ryu, Reviews of Modern Physics 88, 035005 (2016).
  • Gurarie (2011) V. Gurarie, Physical Review B 83, 085426 (2011).
  • Trescher and Bergholtz (2012) M. Trescher and E. J. Bergholtz, Physical Review B 86, 241111 (2012).
  • Zhang et al. (2019) Y.-H. Zhang, D. Mao, Y. Cao, P. Jarillo-Herrero,  and T. Senthil, Physical Review B 99, 075127 (2019).
  • Roy (2014) R. Roy, Physical Review B 90, 165139 (2014).
  • Jackson et al. (2015) T. S. Jackson, G. Möller,  and R. Roy, Nature communications 6, 1 (2015).
  • Claassen et al. (2015) M. Claassen, C. H. Lee, R. Thomale, X.-L. Qi,  and T. P. Devereaux, Phys. Rev. Lett. 114, 236802 (2015).
  • Note (3) The lowest Landau level automatically satisfies FxyF_{xy}==Tr\tmspace+.1667em𝒢ij\text{Tr}\tmspace+{.1667em}\mathcal{G}_{ij}; a delicate moment here is discussion of the higher Landau levels which at first glance violate this condition. However, since the wave functions of the higher Landau levels can be chosen holomorphic Haldane (2018); Chen and Biswas (2020), surprisingly this argument still holds.
  • Chen and Biswas (2020) Y. Chen and R. R. Biswas, Physical Review B 102, 165313 (2020).
  • Kitaev (2006) A. Kitaev, Annals of Physics 321, 2 (2006).