We consider the chiral model of twisted bilayer graphene introduced by Tarnopolsky, Kruchkov, and Vishwanath (TKV). TKV proved that for inverse twist angles α such that the effective Fermi velocity at the moiré K point vanishes, the chiral model has a perfectly flat band at zero energy over the whole Brillouin zone. By a formal expansion, TKV found that the Fermi velocity vanishes at α ≈ 0.586. In this work, we give a proof that the Fermi velocity vanishes for at least one α between 0.57 and 0.61 by rigorously justifying TKV’s formal expansion of the Fermi velocity over a sufficiently large interval of α values. The idea of the proof is to project the TKV Hamiltonian onto a finite-dimensional subspace and then expand the Fermi velocity in terms of explicitly computable linear combinations of modes in the subspace while controlling the error. The proof relies on two propositions whose proofs are computer-assisted, i.e., numerical computation together with worst-case estimates on the accumulation of round-off error, which show that round-off error cannot possibly change the conclusion of computation. The propositions give a bound below on the spectral gap of the projected Hamiltonian, an Hermitian 80 × 80 matrix whose spectrum is symmetric about 0, and verify that two real 18-th order polynomials, which approximate the numerator of the Fermi velocity, take values with a definite sign when evaluated at specific values of α. Together with TKV’s work, our result proves the existence of at least one perfectly flat band of the chiral model.

Twisted bilayer graphene (TBG) is formed by stacking one layer of graphene on top of another in such a way that the Bravais lattices of the layers are twisted relative to each other. For generic twist angles, the lattices will be incommensurate, so the resulting structure will not be periodic. Bistritzer and MacDonald (BM)1 introduced an approximate model (BM model) for the electronic states of TBG, which is periodic over the scale of the bilayer moiré pattern, where the twist angle enters as a parameter. Using this model, BM showed that the Fermi velocity, the velocity of electrons at the Fermi level, vanishes at particular twist angles known as “magic angles.” The largest of these angles, known as the first magic angle, is at θ ≈ 1.1°. Numerical computations on the BM model show the stronger result that at magic angles, the Bloch band of the BM model at zero energy is approximately flat over the whole Brillouin zone.1,2 The flatness of the zero energy Bloch band is thought to be a critical ingredient for recently observed superconductivity of TBG,3 although the precise mechanism for superconductivity in TBG is not yet settled.

Aiming at a simplified model that explains the nearly flat band of TBG, Tarnopolsky, Kruchkov, and Vishwanath (TKV)4 introduced a simplification of the BM model, which has an additional “chiral” symmetry, known as the chiral model. TKV showed analytically that at magic angles (of the chiral model, still defined by vanishing of the Fermi velocity), the chiral model has exactly flat bands over the whole Brillouin zone. Using a formal perturbation theory (for the chiral model, the natural parameter is the reciprocal of the twist angle up to a constant), TKV derived approximate values for the magic angles of the chiral model. It is worth noting that the first magic angles of the chiral model and the BM model are nearby, but the higher magic angles are not very close.

Becker et al. introduced a spectral characterization of magic angles of the TKV model where the role of a non-normal operator is emphasized [the operator Dα appearing in (1)]. Using this characterization, they numerically computed precise values for the magic angles of the TKV model [see the discussion below (14)].5 In the same work, they also proved that the lowest band of the TKV model becomes exponentially close to flat even away from magic angles as the natural small parameter tends to zero. The same authors also investigated flat bands of the TKV model with more general interlayer coupling potentials and the spectrum of other special cases of the BM model.6 

In this work, we study the chiral model introduced by TKV and consider the problem of rigorously proving the existence of the first magic angle. We do this by justifying the formal perturbation theory of TKV to make a rigorous expansion of the Fermi velocity to high enough order, and over a large enough parameter range, so that we can prove the existence of a zero. By numerically verifying that the resulting expansion attains a negative value and proving that the result continues to hold when the effect of round-off error is included (Proposition II.2), we obtain existence of the magic angle (Theorem II.2).

The proof of validity of the expansion is challenging because the reciprocal of the twist angle at the zero of the Fermi velocity is large relative to the spectral gap of the unperturbed Hamiltonian, which means that the magic angle falls outside of the interval of twist angles where the perturbation series for the Fermi velocity is obviously convergent. To overcome this difficulty, we start by representing the chiral model Hamiltonian in a basis that takes full advantage of model symmetries. Then, using a rigorous bound on the high frequency components of the error, we reduce the error analysis to analysis of the eigenvalues of the chiral model projected onto finitely many low frequencies. The final stage of the error analysis (Theorem II.1) is to prove a proposition about the eigenvalues of the projected chiral model by a numerical computation that we prove continues to hold when the accumulation of round-off error is considered (Proposition IV.5). We discuss the limitations of our methods and, in particular, whether our methods might be generalized to the more general settings considered by Becker et al.5,6 in Remarks II.1–II.3.

We have made code for the numerical computations used in our proofs available at github.com/abwats/magic_angle. We give references to specific scripts in the text.

The chiral model, such as the Bistritzer–MacDonald model (B–M model) from which it is derived, is a formal continuum approximation to the atomistic tight-binding model of twisted bilayer graphene. The BM and chiral models aim to capture physics over the length-scale of the bilayer moiré pattern, which is, for small twist angles, much longer than the length-scale of the individual graphene layer lattices. Crucially, even when the graphene layers are incommensurate so that the bilayer is aperiodic on the atomistic scale, the chiral model and BM model are periodic (up to phases) with respect to the moiré lattice so that they can be analyzed via Bloch theory.

We define the moiré lattice to be the Bravais lattice

Λ=m1a1+m2a2:(m1,m2)Z2

generated by the moiré lattice vectors

a1=2π33,1,a2=2π33,1,

and we denote a fundamental cell of the moiré lattice by Ω. The moiré reciprocal lattice is the Bravais lattice

Λ*=n1b1+n2b2:(n1,n2)Z2

generated by the moiré reciprocal lattice vectors defined by ai · bj = 2πδij, given explicitly by

b1=123,3,b2=123,3.

We define q1=0,1, which is the (re-scaled) difference of the K points (Dirac points) of each layer, and

q1=0,1,q2=q1+b1=123,1,q3=q1+b2=123,1.

We write Ω* for a fundamental cell of the moiré reciprocal lattice and refer to such a cell as the Brillouin zone.

Let ϕ2π3. Tarnopolsky–Kruchkov–Vishwanath’s chiral Hamiltonian is defined as

Hα=0DαDα0,Dα=2īαU(r)αU(r)2ī,
(1)

where ̄=12(x+iy), U(r)=eiq1r+eiϕeiq2r+eiϕeiq3r, † denotes the adjoint (Hermitian transpose), and α is a real parameter, which we will take to be positive α ≥ 0 throughout [see (3)]. The chiral Hamiltonian Hα is an unbounded operator on H=L2(R2;C4) with domain H1(R2;C4). We will write functions in H as

ψ(r)=ψ1A(r),ψ2A(r),ψ1B(r),ψ2B(r),
(2)

where |ψτσ(r)|2 represents the electron density near to the K point (in momentum space) on sublattice σ and on layer τ. The diagonal terms of Dα arise from Taylor expanding the single layer graphene dispersion relation about the K point of each layer, while the off-diagonal terms of Dα couple the A and B sublattices of layers 1 and 2. The chiral model is identical to the BM model except that inter-layer coupling between sublattices of the same type is turned off in the chiral model. The precise form of the interlayer coupling potential U can be derived under quite general assumptions on the real space interlayer hopping.1,7 The parameter α is, up to unimportant constants, the ratio

αinterlayer hopping strength between A and B sublatticestwist angle.
(3)

Although the limit α → 0 can be thought of as the limit of vanishing interlayer hopping strength at fixed twist, it is physically more interesting to view the limit as modeling increasing twist angle at a fixed interlayer hopping strength.

Bistritzer and MacDonald studied the effective Fermi velocity of electrons in twisted bilayer graphene modeled by the BM model and computed values of the twist angle such that the Fermi velocity vanishes, which they called “magic angles.” One can similarly define an effective Fermi velocity for the chiral model and refer to values of α such that the Fermi velocity vanishes as “magic angles” [although technically α is related to the reciprocal of the twist angle (3)].

TKV proved the remarkable result that, at magic angles, the chiral model has a perfectly flat Bloch band at zero energy. Let LK2 denote the L2 space on a single moiré cell Ω with moiré K point Bloch boundary conditions. The starting point of TKV’s proof is an expression for the Fermi velocity as a function of α, v(α), as a functional of one of the Bloch eigenfunctions, ψαLK2, of Hα,

v(α)|ψα*(r)ψα(r)||ψαψα|,
(4)

where .. denotes the LK2 inner product. We give precise definitions of LK2, ψα, and v(α) in Definition III.2, Proposition III.6, and Definition III.3, respectively. We give a systematic formal derivation of why (4) is the effective Fermi velocity at the moiré K point in  Appendix A. To complete the proof, TKV showed that zeros of v(α) imply zeros of ψα at special “stacking points” of Ω and that such zeros of ψα allow for Bloch eigenfunctions with zero energy to be constructed for all k in the moiré Brillouin zone.

To derive approximate values for magic angles, TKV computed a formal perturbation series approximation of ψα,

ψα=Ψ0+αΨ1+,
(5)

and then substituted this expression into the functional for v(α) to obtain an expansion of v(α) in powers of α,

v(α)=13α2+α411149α6+143294α8+1+3α2+2α4+67α6+10798α8+.
(6)

By setting v(α) = 0, one obtains an approximation for the smallest magic angle: α ≈ 0.586.

Although TKV proved that flat bands occur at magic angles, they did not prove the existence of magic angles, and hence, they did not prove the existence of flat bands. The contribution of the present work is to prove rigorous estimates on the error in approximation (5), which are sufficiently high order and precise that, once substituted into (4), they suffice to rigorously prove the existence of a zero of v(α) and, hence, via TKV’s proof, the existence of at least one perfectly flat band.

It turns out to be relatively straightforward to prove that series (5) and (6) are uniformly convergent and to derive precise bounds on the error in truncating the series for |α|<13; see Proposition IV.3. The basic challenge, then, is to derive similar error bounds for α over an interval, which includes the expected location of the first magic angle, at 13. The first main theorem we will prove, roughly stated, is the following. See Theorem IV.1 for the more precise statement. The theorem relies on the existence of a spectral gap for an 80 × 80 Hermitian matrix, which requires numerical computation for its proof; see Proposition IV.5.

Theorem II.1.
TheKpoint Bloch functionψαsatisfies
ψα=n=08αnΨn+ηα,
(7)
whereηαn=08αnΨnwith respect to theLK2inner product and
ηαLK23α91520αforall0α710.
(8)

The functions Ψn for 0 ≤ n ≤ 8 are derived recursively: see  Appendix C. We stop at eighth order in the expansion because this is the minimal order such that we can guarantee the existence of a zero of v(α), but the functions Ψn are well-defined by a recursive procedure for arbitrary positive integers n; see Proposition IV.1.

Substituting (7) into the functional for Fermi velocity (4) and using ηαn=18αnΨn, we find that

v(α)=vN(α)vD(α),

where

vN(α)n=08αnΨn*(r)n=08αnΨn(r)+ηα*(r)n=08αnΨn(r)+n=08Ψn*(r)ηα(r)+ηα*(r)ηα(r)
(9)

and

vD(α)n=08αnΨnn=08αnΨn+ηαηα,

where .. denotes the LK2 inner product and ηα satisfies (8). The following is a straightforward calculation.

Proposition II.1.
The following identities hold:
n=08αnΨn*(r)n=08αnΨn(r)=13α2+α411149α6+143294α8753693311957764α10+459817233147460365316α1230028809212865451520327364608478700α14+4975014185899222712487856750603488800α16,
(10)
n=08αnΨnn=08αnΨn=1+3α2+2α4+67α6+10798α8+511948412α10+62026511356844852α12+355691470247113410497953025α14+2481663780475871337509641908202400α16.
(11)

We prove Proposition II.1 in  Appendix F. Naïvely, expansions (10) and (11) approximate the formal infinite series expansions of n=0αnΨn*(r)n=0αnΨn(r) and n=0αnΨnn=0αnΨn up to terms of order α9. We prove in Proposition F.2 that because of some simplifications, expansions (10) and (11) agree with the infinite series up to terms of order α10.

We are now in a position to state and prove our second result. This result also relies on a proposition, which requires numerical computation for its proof: that one real 18-th order polynomial in α attains a negative value and another attains a positive value, when evaluated at specific values of α; see Proposition II.2.

Theorem II.2.

There exist positive numbersαminandαmaxwith 0.57 < αmin < αmax < 0.61 such that the Fermi velocityv(α) defined by(4)has a zeroα*satisfyingαminα*αmax.

Proof.
Equation (9) and Cauchy and Schwarz imply that
vN(α)n=08αnΨn*(r)n=08αnΨn(r)2ηαn=08αnΨn+ηα2.
Using Theorem II.1 and Proposition C.3, we see that vN(α) is bounded above by the polynomial
13α2+α411149α6+143294α8753693311957764α10+459817233147460365316α1230028809212865451520327364608478700α14+4975014185899222712487856750603488800α16+E(α),
(12)
where
E(α)6α91520α1+3α+2α2+147α3+25842α4+19688373458α5+10652579931122α6+22129312323981473624696345α7+1836431197552144544997570760α8+9α18(1520α)2,
where we use Proposition C.3 to calculate the term in brackets for all 0α710. On the other hand, v(α) is bounded below for all 0α710 by the polynomial
13α2+α411149α6+143294α8753693311957764α10+459817233147460365316α1230028809212865451520327364608478700α14+4975014185899222712487856750603488800α16E(α).
(13)
We now claim the following.

Proposition II.2.

Expression(12), or equivalently the 18-th order polynomial obtained by multiplying(12)by (15 − 20α)2, is negative atα = 0.61. Similarly, expression(13)is positive atα = 0.57.

Proposition II.2 obviously implies by continuity (12), (13), and vN(α), each has at least one zero in the interval 0.57 < α < 0.61. We denote the largest zero of (12) in the interval by αmax and the smallest zero of (13) in the interval by αmin. Since the zeroes of vN(α) must lie between those of (12) and (13), we are done.□

Proof of Proposition II.2

(computer-assisted). We will first prove that (12) attains a negative value at 0.61 and then explain the modifications necessary to prove that (13) is positive at 0.57. Evaluating using the double-precision floating-point arithmetic, we find that at α = 0.61, (12) attains the negative value −0.020 263 (five significant figures; this value was computed by running the script compute_expansion_symbolically.py in the Github repo). It is straightforward to bound the numerical error, which accumulates when evaluating an 18-th order polynomial using the floating-point arithmetic. Even the simplest exact bound, which does not account for error cancellation (see, e.g., Eq. (8) of Oliver),8 yields an upper bound on the possible accumulated round-off error in the evaluation of an nth order polynomial j=0npjαj, for α ∈ [−1, 1], as (n+1)e(2n+1)ϵ1sup0jn|pj|, where ϵ is “machine epsilon”: roughly speaking, the maximum possible round-off error generated in a single arithmetic operation. Bounding the maximum coefficient in (12) by 1000, taking n = 18, and bounding ϵ by 3 × 10−16 (which was easily attained working in Python on our machine), the maximum possible numerical error in the evaluation of (12) is ≈10−11, which is much smaller than 0.020 263. We conclude that the first claim of Proposition II.2 must hold. Regarding the second, evaluating at α = 0.57, we find that (13) equals 0.029 138 (5sf). The same argument as before now shows that accumulated round-off error in the evaluation cannot possibly change the sign of (13) at α = 0.57.□

We do not attempt to rigorously estimate αmin and αmax precisely in this work, but numerically computing roots of polynomials (12) and (13) suggests αmin ≈ 0.576 83 (5sf) and αmax ≈ 0.601 77 (5sf), respectively, where 5sf is an abbreviation for five significant figure. Numerical computation of the first zero of n=08αnΨn*(r)n=08αnΨn(r) gives 0.585 97 (5sf); see Fig. 1 (the zero values were computed by running the script compute_expansion_symbolically.py in the Github repo).

FIG. 1.

Left: plot of the numerator vN(α) of the Fermi velocity approximated by eighth order TKV expansion (6) (orange) and of eighth order expansions with worst-case (12) (blue) and best-case (13) (green) errors. Right: detail showing computed roots of these functions near to α=13. Numerically computing the zeroes of each curve yields α = 0.585 97 (5sf), α = 0.601 77 (5sf), and α = 0.576 83 (5sf), respectively. The values of α (0.57 and 0.61) where we evaluated expressions (13) (green line) and (12) (blue line) to prove that vN(α) has a zero between 0.57 and 0.61 are shown with the black crosses.

FIG. 1.

Left: plot of the numerator vN(α) of the Fermi velocity approximated by eighth order TKV expansion (6) (orange) and of eighth order expansions with worst-case (12) (blue) and best-case (13) (green) errors. Right: detail showing computed roots of these functions near to α=13. Numerically computing the zeroes of each curve yields α = 0.585 97 (5sf), α = 0.601 77 (5sf), and α = 0.576 83 (5sf), respectively. The values of α (0.57 and 0.61) where we evaluated expressions (13) (green line) and (12) (blue line) to prove that vN(α) has a zero between 0.57 and 0.61 are shown with the black crosses.

Close modal

Using Proposition C.1 and the package Sympy9 for symbolic computation, we can compute the formal expansion of v(α) up to arbitrarily high order in α. In particular, we find the higher-order terms in expansion (6) to be

v(α)=13α2+α411149α6+143294α81022725711957764α10+688113701547460365316α12130055941435858531520327364608478700α14+1+3α2+2α4+67α6+10798α8+1601148412α10+134058653356844852α12+26407145691649226820995906050α14+.
(14)

Truncating the numerator after order α40 and setting the numerator equal to zero yield α = 0.585 663 558 389 56 (14sf) for the first zero of the Fermi velocity (to compute this value, run compute_expansion_symbolically.py in the Github repo with N = 40). This is consistent with the numerical computation of Becker et al.,5 who found α = 0.585 663 558 389 55 truncated (not rounded) to 14 digits, by diagonalizing a non-normal but compact operator whose reciprocal eigenvalues correspond to magic angles. Note that we do not attempt to rigorously justify series (14) to such large values of α and to such high order in this work; see Remark II.4.

Remark II.1

(Higher magic angles). The chiral model has been conjectured to have infinitely many magic angles,5 but it is not straightforward to extend our methods to prove the existence of such higher magic angles. The problem is that calculating the perturbation series centered atα = 0 requires diagonalizing the unperturbed operatorH0. In principle, it might be possible to calculate the perturbation series to higher order in order to get an accurate approximation of the Fermi velocity near to the higher magic angles. However, this would require significantly more calculation compared with the present work, and we have no guarantee that the error can be made small enough to prove the existence of another zero in that case.

Remark II.2

(More general interlayer hopping potentials). The chiral model(1)is an approximation to the full Hamiltonian of the twisted bilayer, even in the chiral limit where coupling between sublattices of the same type is turned off because the interlayer hopping potentialUonly allows for hopping between nearest neighbors in the momentum lattice (seeFig. 6 ). More general interlayer hopping potentials have been studied by Becker et al.6 In principle, such models should be amenable to the analysis of this work, but longer-range hopping would lead to much more involved calculations, and the construction of the finite-dimensional subspace Ξ of Proposition IV.4 would require more care: the fact that we can choose Ξ so thatPΞH1PΞ=1depends onH1only coupling nearest neighbors in the momentum lattice. Locality of hopping in the momentum space lattice has been exploited for efficient computation of density of states10 of twisted bilayers.

Remark II.3

(Generalization to BM model). Parts of our analysis should also apply to the full Bistritzer–MacDonald model. Specifically, one could study perturbation series for Bloch functions near to zero energy in powers of the inter-layer hopping strength, derive an equivalent expression for the Fermi velocity in terms of that series, and then study the zeroes of that series. However, there are various complications because of the lack of “chiral” symmetry. First, there is no reason for the continuation of the zero eigenvalue of the unperturbed operator to remain at zero. Second, the expression for the Fermi velocity in terms of the associated eigenfunction could be more complicated. Since zeros of the BM model Fermi velocity do not imply the existence of flat bands for that model, we do not consider these complications in thiswork.

Remark II.4

(Expanding to higher order). Our methods could, in principle, be continued to justify the expansion of the Fermi velocity to arbitrarily high order and potentially over larger intervals ofαvalues. However, these extensions are not immediate: pushing the expansion to higher order or to a larger interval ofαvalues would require a larger set Ξ in Lemma IV.1, and Proposition IV.5 would have to be re-proved for the new set Ξ. Note that the essential difficulty is justifying the perturbation series for largeα: the series are easily justified to all orders for|α|<13; see Proposition IV.3.

We review the symmetries, Bloch theory, and symmetry-protected zero modes of TKV’s chiral model in Sec. III. We prove Theorem II.1 in Sec. IV, postponing most details of the proofs to  Appendixes A– F. In  Appendix A, we show why (4) corresponds to the effective Fermi velocity at the moiré K point. In  Appendix B, we construct an orthonormal basis, which we refer to as the chiral basis, which allows for efficient computation and analysis of TKV’s formal expansion. We re-derive TKV’s formal expansions in  Appendix C. We give details of the Proof of Theorem II.1 in  Appendixes D and  E. We prove Proposition II.1 in  Appendix F. In the supplementary material, we list the basis functions of the subspace onto which we project the TKV Hamiltonian, give the explicit forms of the higher-order corrections in expansion (7), and present a derivation of the TKV Hamiltonian from the Bistritzer–MacDonald model.

In this section, we review the symmetries of the TKV model for the reader’s convenience and to fix notation. Becker et al.5 gave a group theoretical account of these symmetries, and further reviews can be found in the physics literature.11–13 Recall that ϕ=2π3, and let Rϕ denote the matrix, which rotates vectors counter-clockwise by ϕ, i.e.,

Rϕ=121331.

We define the following.

Definition III.1.
For anyv ∈ Λ, we define a phase-shifted translation operator acting on functionsfHby
τvfdiag1,eiq1v,1,eiq1vτ̃vf,τ̃vf(r)=f(r+v).
(15)
We define a phase-shifted version of the operator, which rotates functionsfHclockwise byϕby
Rfdiag1,1,eiϕ,eiϕR̃f,R̃f(r)=f(Rϕr).
(16)
For anyfH, we finally define the “chiral” symmetry operator
Sfdiag1,1,1,1f.
(17)

We then have the following.

Proposition III.1.
Operators(15)and(16)are symmetries in the sense that
Hα,τv=HατvτvHα=0
(18)
for all moiré lattice vectorsv ∈ Λ,
Hα,R=HαRRHα=0,
and operator(17)is a “chiral” symmetry in the sense that
Hα,S=HαS+SHα=0.
(19)

Proof.
The first claim is a direct calculation using the facts that for any v ∈ Λ,
τ̃vU(r)τ̃v=eiq1vU(r),τ̃v̄τ̃v=̄.
The second claim is a direct calculation using the facts that
R̃1U(r)R̃=eiϕU(r),R̃1̄R̃=eiϕ̄.
The final claim is trivial to check.□

The “chiral” symmetry (19) implies that the spectrum of Hα is symmetric about zero because

Hαψ=EψHαSψ=ESψ.

The same calculation implies that zero modes of Hα can always be chosen without loss of generality to be eigenfunctions of S.

We now want to reduce the eigenvalue problem for Hα using the symmetries just introduced. The symmetry (18) means that eigenfunctions of Hα can be chosen without loss of generality to be simultaneous eigenfunctions of τv for all v ∈ Λ. It therefore suffices to seek solutions of

Hαψ=Eψ

for r in a fundamental cell ΩR2/Λ of the moiré lattice in the symmetry-restricted spaces

Lk2fL2(Ω;C4):f(r+v)=eikvdiag(1,eiq1v,1,eiq1v)f(r)vΛ,
(20)

where k is known as the quasimomentum. Since Lk+w2=Lk2 for any w ∈ Λ*, it suffices to restrict attention to k in a fundamental cell of Λ*, which we denote Ω*R2/Λ* and refer to as the Brillouin zone. We also define symmetry-restricted Sobolev spaces Hks for each k ∈ Ω* and positive integer s by

HksfHs(Ω;C4):f(r+v)=eikvdiag(1,eiq1v,1,eiq1v)f(r)vΛ.

We claim the following.

Proposition III.2.

For each fixedk ∈ Ω*andα ≥ 0,Hα, defined on the domainHk1, extends to an unbounded self-adjoint elliptic operatorLk2Lk2with compact resolvent. In a complex neighborhood of everyα ≥ 0, the familyHαis a holomorphic family of type (A) in the sense of Kato.14 

Proof.

Ellipticity is immediate since the principal symbol of Hα is invertible. Self-adjointness is clear using the Fourier transform when α = 0 and for α ≠ 0 because αH1 is a bounded symmetric perturbation of H0 (see, e.g., Theorem 1.4 of Cycon et al.15). Elliptic regularity implies that the resolvent maps Lk2Hk1, and compactness of the resolvent then follows by Rellich’s theorem (see, e.g., Proposition 3.4 of Taylor16). The family Hα is holomorphic of type (A) since the domain of Hα is independent of α, and Hαf is holomorphic for every fHk1 (see Chap. 7 of Kato14).□

We now claim the following.

Proposition III.3.

LetfLk2. Then,RfLRϕ*k2.

Proof.
By definition, for any v ∈ Λ,
Rf(r+v)=diag(1,1,eiϕ,eiϕ)f(Rϕr+Rϕv).
By the definition of Lk2, we have
Rf(r+v)=ei(Rϕ*k)vdiag(1,ei(Rϕ*q1)v,1,ei(Rϕ*q1)v)Rf(r).
The conclusion now follows from Rϕ*q1=q1+b2 and b2 · v = 0 mod 2π for all v ∈ Λ.□

In particular, whenever Rϕ*k=k mod Λ*, we have RLk2=Lk2. Regarding such k, the following is a simple calculation.

Proposition III.4.

The moiréKandKpointsk = 0 andk = −q1and the moiré Γ pointk = q1 + b1satisfyRϕ*k=k mod Λ*.

The moiré K, K′, and Γ points are shown in Fig. 2. Note that the moiré K, K′, and Γ points should not be confused with the single layer K, K′, and Γ points. The moiré K point corresponds to the K point of layer 1, while the moiré K′ point corresponds to the K point of layer 2. Interactions with the K′ points of layers 1 and 2 are formally small for small twist angles and are hence ignored.

FIG. 2.

Diagram showing locations of moiré K (blue), K′ (red), and Γ (black) points within the moiré Brillouin zone (orange).

FIG. 2.

Diagram showing locations of moiré K (blue), K′ (red), and Γ (black) points within the moiré Brillouin zone (orange).

Close modal

In this work, we will be particularly interested in Bloch functions at the moiré K and K′ points. We therefore define the following.

Definition III.2.
LK2L02,LK2Lq12.

Let ω = e. Since the spaces LK2 and LK2 are invariant under R, they can be divided up into invariant subspaces corresponding to the eigenvalues of R,

LK2=LK,12LK,ω2LK,ω*2,LK2=LK,12LK,ω2LK,ω*2,

where

LK,σ2fLK2:Rf=σfσ=1,ω,ω*

and LK,σ2,σ=1,ω,ω*, are defined similarly.

The following, which is trivial to prove, will be important for studying the zero modes of Hα.

Proposition III.5.

The operatorScommutes withτvandRand hence maps theLK,σ2andLK,σ2spaces to themselves forσ = 1, ω, ω*.

Since S has eigenvalues ±1, we can define the spaces

LK,σ,±12=fLK,σ2:Sf=±fσ=1,ω,ω*

and spaces LK,σ,±12,σ=1,ω,ω* similarly.

Remark III.1.

Note that +1 eigenspaces ofScorrespond to wave-functions, which vanish in their third and fourth entries, which correspond, through(2), to wave-functions supported only onAsites of the layers. Similarly, −1 eigenspaces ofScorrespond to wave-functions, which vanish in their first and second entries, which are supported only onBsites of the layers.

We now want to investigate zero modes of Hα in detail. When α = 0, there are exactly four zero modes given by ej, j = 1, 2, 3, 4, where ej equals 1 in its jth entry and 0 in its other entries. It is easy to check that

e1LK,12,e2LK,12,e3LK,ω*2,e4LK,ω*2,
(21)

and hence, 0 is a simple eigenvalue of Hα when restricted to each of these subspaces. Recall that zero modes can always be chosen as eigenfunctions of S, and indeed, we have

e1LK,1,12,e2LK,1,12,e3LK,ω*,12,e4LK,ω*,12.
(22)

We now claim that these zero modes persist for all α. This was already established by TKV,4 and the following proposition is also similar to Proposition 3.1 of Becker et al.5 We re-state it using our notation and give the proof for completeness.

Proposition III.6.

There exist smooth functionsψαwithψα‖ = 1 in each of the spacesLK,1,12,LK,1,12,LK,ω*,12, andLK,ω*,12such thatψ0is as in(21),αψαis real-analytic, andHαψα = 0 for allα. The dimension of ker Hαrestricted to each of the spacesLK,12,LK,12,LK,ω*2, andLK,ω*2is always odd-dimensional.

Proof.

Since S preserves each of the spaces LK,12, LK,12, LK,ω*2, and LK,ω*2 and anti-commutes with Hα, the spectrum of Hα restricted to each space must be symmetric about 0 for all α. Since Hα restricted to each space has compact resolvent and Hα is a holomorphic family of type (A), the spectrum of Hα consists of finitely degenerate isolated eigenvalues depending real-analytically on α, with associated eigenfunctions also real-analytically depending on α (although the real-analytic choice of eigenfunction at an eigenvalue crossing may not respect ordering); see Theorem 3.9 of Chap. 7 of Kato.14 The null space of Hα in each of the spaces is one-dimensional at α = 0 by explicit calculation, with the zero modes given by (up to non-zero constants) (21). For small α > 0, real-analyticity and the chiral symmetry force the null space to remain simple and it is clear how to define ψα. For large α > 0, the non-zero eigenvalues of Hα may cross 0 at isolated values of α, and in this case, we define ψα to be the real-analytic continuation of the zero mode through the crossings. Note that real-analyticity prevents non-zero eigenvalues from equaling zero except at isolated points so that the real-analytic continuation of the zero mode through the crossing must indeed be a zero mode. At crossings, the null space must be odd-dimensional in order to preserve symmetry of the spectrum of Hα about 0. It remains to check that if ψ0 is in, say, LK,1,12, then ψα must remain in LK,1,12 for all α > 0. However, this must hold because the S-eigenvalue of ψα cannot change abruptly while preserving real-analyticity. Smoothness of ψα follows from elliptic regularity.□

In this work, we will restrict attention to the moiré K point and, especially, the family ψαLK,1,12. We expect that our analysis would go through with only minor modifications if we considered instead the moiré K′ point. The zero modes in LK,1,12 and LK,ω*,12 are related by the following symmetry.

Proposition III.7.

Letψ1αandψ1αdenote the zero modes ofHαin the spacesLK,1,12andLK,ω*,12, respectively. Then,ψ1α=(Φα,0), whereΦαL2(Ω;C2),Φα(r+v)=diag(1,eiq1v)Φα(r)for allv ∈ Λ, Φα(Rϕr) = Φα(r). Up to gauge transformationsψ1αeiϕ(α)ψ1α, which preserve real-analyticity ofψ1α, we haveψ1α(r)=(0,Φα*(r)).

Proof.

Since Sψ1α=ψ1α, the last two entries of ψ1α must vanish, so we can write ψ1α=(Φα,0). That Φα satisfies the stated symmetries follows immediately from ψ1αLK,12. It is straightforward to check using the definitions of R and τv that (0,Φα*(r))LK,ω*,12. To see that (0,Φα*(r)) is a zero mode, note that Φα satisfies DαΦα = 0, which implies that DαΦα*(r)=0 by a simple manipulation. To see that ψ1α(r)=(0,Φα*(r)) (up to real-analytic gauge transformations) for all α, note first that this clearly holds for α = 0 [the zero modes are explicit (22)]. For α > 0, the identity must continue to hold by uniqueness (up to real-analytic gauge transformations) of the real-analytic continuation of ψ1α starting from α = 0 and continuing first along the non-zero interval where ψ1α is non-degenerate in LK,1,12 and then through eigenvalue crossings as in the Proof of Proposition III.6.□

In  Appendix A, we use Proposition III.7 to derive the effective Dirac operator with the α-dependent Fermi velocity, which controls the Bloch band structure in a neighborhood of the moiré K point. The Fermi velocity of the effective Dirac operator is given by the following. Note that we drop the subscript +1 when referring to the zero mode of Hα in LK,1,12 since the zero mode of Hα in LK,ω*,12 plays no further role.

Definition III.3.
LetψαLK,1,12be as in Proposition III.6. Then, we define
v(α)|ψα*(r)ψα(r)||ψαψα|,
(23)
where..denotes theLK2inner product.

We now turn to approximating the zero mode ψαLK,1,12 by a series expansion in powers of α. We write Hα = H0 + αH1 and formally expand ψα as a series

ψα=Ψ0+αΨ1+,
(24)

where H0Ψ0 = 0 and

H0Ψn=H1Ψn1
(25)

for all n ≥ 1. To solve H0Ψ0 = 0, we take Ψ0 = e1. We prove the following in  Appendix C.

Proposition IV.1.
LetPdenote the projection operator inLK,12ontoe1andP = IP. The sequence ofEq. (25)has a unique solution such thatΨnLK,1,12for alln ≥ 0 andPΨn = 0 for alln ≥ 1, given by Ψ0 = e1and
Ψn=P(H0)1PH1Ψn1
(26)
for eachn ≥ 1.

Expansion (24) appears different from the series studied by TKV since we work only with the self-adjoint operators H0, H1, and Hα rather than the non-self-adjoint operator Dα [defined in (1)]. Since functions in LK,1,12 vanish in their last two components, there is no practical difference. However, working with only self-adjoint operators allows us to use the spectral theorem, which greatly simplifies the error analysis. We compute the first eight terms in expansion (24) in Proposition C.2 after developing some necessary machinery in  Appendix B.

In this section, we explain the essential challenge in proving error estimates for series (24) and explain how we overcome this challenge. Our goal is to prove the following.

Theorem IV.1.
LetψαLK,1,12be as in Proposition III.6. Then,
ψα=n=18αnΨn+ηα,
whereηαn=18αnΨnwith respect to theLK2inner product and
ηαLK,123α91520αforall0α710.

Proposition IV.1 guarantees that series (24) is well-defined up to arbitrarily many terms. A straightforward bound on the growth of terms in the series comes from the following proposition.

Proposition IV.2.
The spectrum ofH0inLK,12is
σLK,12(H0)={±|G|,±|q1+G|:GΛ*},
and hence,
P(H0)1PLK,12LK,12=1.
(27)
We also have
H1LK,12LK,12=3.
(28)

Proof.

This proposition is a combination of Propositions B.2, B.4, and B.7, proved in  Appendix B.□

Proposition IV.2 implies that P(H0)1PH1LK,12LK,123, which implies the following.

Proposition IV.3.

The formal series(24)converges toψαinLK2, with an explicit error rate, for all|α|<13. The formal series for the Fermi velocityv(α) obtained by substituting the series expansion ofψαinto(23)converges for the same range ofα, also with an explicit error rate.

Proof.

For any non-negative integer N, let ψN,αn=0NαnΨn, where Ψn are as in (26). Since Ψ0 ⊥ Ψn for all n ≥ 1 and ‖Ψ0‖ = 1, we have that ‖ψN,α‖ ≥ 1 for all N. Let ϕN,αψN,αψN,α; then, we can decompose ϕN,α = α + ηα for some constant c and where ηαψα. Applying Hα to both sides, we have that HαϕN,α=αN+1H1ΨNψN,α=Hαηα. Now, fix α ≥ 0 such that |α|<13. Then, αH1‖ < 1, and hence, the first non-zero eigenvalue of Hα is bounded away from 0 by 1 − 3α (recall that the first non-zero eigenvalues of H0 are ±1). Since ηαψα, where ψα spans the eigenspace of the zero eigenvalue of Hα, we have that ηα|αN+1|H1ΨN|13α|ψN,α. Using the bound ‖ΨN‖ ≤ (3α)N and the bound below on ‖ψN,α‖, we have that ηα(3α)N+1|13α|, which clearly → 0 as N → ∞, so that limN→∞ϕN,α = ψα (up to a non-zero constant). Now, consider v(α) defined by (23). Assuming WLOG that ‖ψα‖ = 1 and substituting ψα = ϕN,α + ηα, we find immediately, using Cauchy–Schwarz, that v(α)ϕN,*(r)ϕN(r)2ηα+ηα2. In terms of ψN, we have v(α)ψN,*(r)ψN(r)ψNψN2ηα+ηα2.□

Proposition IV.3 shows that for |α|<13, series (24) converges to ψα and can be used to compute the Fermi velocity. However, this restriction is too strong to prove that the Fermi velocity has a zero, which occurs at the larger value α13. Of course, Proposition IV.2 establishes only the most pessimistic possible bound on the expansion functions Ψn, and this bound appears to be far from sharp from explicit calculation of each Ψn; see Proposition C.3. We briefly discuss a possible route to a tighter bound in Remark C.2 but do not otherwise pursue this approach in this work.

We now explain how to obtain error estimates over a large enough range of α values to prove v(α) has a zero. We seek a solution of Hαψα = 0 in LK,1,12 with the form

ψα=ψN,α+ηα,ψN,αn=0NαnΨn.
(29)

For arbitrary α, let Qα denote the projection in LK,12 onto ψN,α, and Qα,⊥IQα (note that Q0 = P). Note that Qα depends on N, but we suppress this to avoid clutter. We assume WLOG that Qαηα(r) = 0. It follows that ηα satisfies

Qα,HαQα,ηα=αN+1Qα,H1ΨN.

To obtain a bound on ηα in L2(Ω), we require a lower bound on the operator Qα,HαQα,:Qα,LK,12Qα,LK,12. The following lemma gives a lower bound on this operator in terms of a lower bound on the projection of this operator onto the finite-dimensional subspace of LK,12 corresponding to a finite subset of the eigenfunctions of H0. The importance of this result is that since H1 only couples finitely many modes of H0, for fixedN, by taking the subset sufficiently large, we can always arrange that ψN,α lies in this subspace.

Lemma IV.1.
LetPΞdenote the projection onto a subset Ξ of the eigenfunctions ofH0inLK,12, and letμ ≥ 0 be maximal such that
PΞH0PΞfμffHK,11,PΞIPΞ,
(30)
(with this notation, the operatorPintroduced in Proposition IV.1 corresponds toPΞwith Ξ being the set {e1} andμ = 1). Suppose thatQαPΞ = PΞQα = Qα, i.e., thatψN,αlies in ranPΞ. Definegαby
gαmin|E|:EisaneigenvalueofthematrixQα,PΞHαPΞQα,actingQα,PΞLK,12Qα,PΞLK,12.
We note thatPΞQα,⊥is the projection onto the subspace ofPΞLK,12orthogonal toψN,α. As long as
3αμandαQα,PΞH1PΞ<min(gα,μ3α),
then
Qα,HαQα,ηαmin(gα,μ3α)αQα,PΞH1PΞQα,ηα.
(31)

Note that gα would be identically zero if not for the restriction that the matrix acts on Qα,PΞLK,12 since otherwise ψN,α would be an eigenfunction with eigenvalue zero for all α. As it is, g0 = 1 and αgα is real-analytic so that gα must be positive for a non-zero interval of positive α values.

Proof.
Using QαPΞ = PΞQα, we have PΞQα,=Qα,PΞ=PΞ, and hence,
Qα,HαQα,ηα=Qα,(PΞ+PΞ)Hα(PΞ+PΞ)Qα,ηα=Qα,PΞHαPΞQα,ηα+αQα,PΞH1PΞηα+αPΞHαPΞQα,ηα+PΞHαPΞηα.
By the reverse triangle inequality,
Qα,HαQα,ηαQα,PΞHαPΞQα,ηα+PΞHαPΞηααQα,PΞH1PΞηα+PΞHαPΞQα,ηα.
(32)
We want to bound the second term above and the first term below. We start with the second term
Qα,PΞH1PΞηα+PΞHαPΞQα,ηα2=Qα,PΞH1PΞηα2+PΞH1PΞQα,ηα2Qα,PΞH1PΞ2PΞηα2+PΞQα,ηα2=Qα,PΞH1PΞ2Qα,ηα2,
where we use Pythagoras’ theorem, PΞH1PΞQα,ηα=PΞH1PΞQα,PΞQα,ηα since PΞQα,⊥ is a projection, and Qα,PΞH1PΞ=PΞH1PΞQα,. Hence, we can bound
Qα,PΞH1PΞηα+PΞHαPΞQα,ηαQα,PΞH1PΞQα,ηα.
(33)
For the first term, first note that using Proposition IV.2 and the spectral theorem,
Qα,PΞHαPΞQα,ηα|Qα,PΞH0PΞQα,ηααQα,PΞH1PΞQα,ηα|(μ3α)PΞQα,ηα
as long as μ ≥ 3α. We now estimate
Qα,PΞHαPΞQα,ηα+PΞHαPΞηα2=Qα,PΞHαPΞQα,ηα2+PΞHαPΞηα2(gα)2Qα,PΞηα2+(μ3α)2PΞηα2min(gα)2,(μ3α)2Qα,PΞηα2+PΞηα2=min(gα)2,(μ3α)2Qα,ηα2.
It follows that as long as 3α ≤ μ,
Qα,PΞHαPΞQα,ηα+PΞHαPΞηαmin(gα,μ3α)Qα,ηα.
(34)
The conclusion now holds as long as 3α ≤ μ and αQα,PΞH1PΞmin(gα,μ3α) upon substituting (33) and (34) into (32).□

For Lemma IV.1 to be useful, we must check that it is possible to choose Ξ so that the bound (31) is non-trivial, i.e., so that the constant is positive. We will prove the following in  Appendix D.

Proposition IV.4.

There exists a subset Ξ of the eigenfunctions ofH0such that the following holds:

  1. The maximalμsuch that(30)holds isμ = 7.

  2. ψ8,αdefined by(29)lies in ran PΞ.

  3. PΞH1PΞ=1, and hence,Qα,PΞH1PΞ1.

The set Ξ constructed in Proposition IV.4 is the set of LK,12-eigenfunctions of H0 with eigenvalues with magnitude 43, augmented with two extra basis functions to ensure that PΞH1PΞ=1. Including all LK,12-eigenfunctions of H0 with eigenvalue magnitudes up to and including 43 ensure that ψ8,α lies in ran PΞ.

We now require the following.

Proposition IV.5.

Let Ξ be as in Proposition IV.4. Then,gα34for all0α710.

Proof

(computer-assisted). Consider HΞαQα,PΞHαPΞQα, acting on PΞLK,12. Assuming that α is restricted to an interval such that the zero eigenspace of HΞα is simple, then, using orthogonality of eigenvectors corresponding to different eigenvalues and the fact that Qα is the spectral projection onto the unique zero mode of HΞα, HΞα has the same non-zero eigenvalues as the matrix Qα,⊥PΞHαPΞQα,⊥ acting on Qα,PΞLK,12. The matrix HΞα is an 81 × 81 matrix whose spectrum is symmetric about 0 because of the chiral symmetry. When α = 0, the spectrum is explicit: 0 is a simple eigenvalue, and the smallest non-zero eigenvalues are ±1, both also simple. Proposition IV.5 is proved if we can prove that the first positive eigenvalue of HΞα is bounded away from zero by 34 for all 0α710. Note that if this holds, the zero eigenspace of HΞα must be simple for all 0α710, and hence, our basic assumption is justified. The strategy of the proof is as follows.

  1. Define a grid G7n10N:n{0,1,,N}, where N is a positive integer taken sufficiently large that the grid spacing h710N<1388831 (the number 388 831 comes from Proposition IV.7).

  2. Numerically compute the eigenvalues of HΞα for αG. We find that the numerically computed first positive eigenvalues of these matrices are uniformly bounded below by 810>34.

  3. Perform a backward error analysis that fully accounts for round-off error in the numerical computation in order to prove that the exact first positive eigenvalues of the matrices HΞα must also be bounded below by 810 at each αG.

  4. Use perturbation theory to bound the exact first positive eigenvalue of HΞα below by 34 over the whole interval of α values between 0 and 710.

When discussing round-off error due to working in the floating-point arithmetic, we will denote “machine epsilon” by ϵ. The significance of this number is that we will assume that all complex numbers a can be represented by floating-point numbers ã such that |aã|ϵa. We will also make the standard assumption about creation of round-off error in the floating-point arithmetic operations: if ã and b̃ are floating-point complex numbers and if (ãOb̃)comp and ãOb̃ represent the numerically computed value and exact value of an arithmetic operation on the numbers ã and b̃, then (ãOb̃)comp=ãOb̃+e, where |e|(ãOb̃)ϵ. In Python, this is indeed the case, for all reasonably sized (such that stack overflow does not occur) complex numbers, with ϵ = 2.220 44 × 10−16 (5sf). We now present the main points of parts 2–4 of the strategy, postponing proofs of intermediate lemmas to  Appendix E.

For part 2 of the strategy, for each αG, we let H̃Ξα denote HΞα (which is known exactly) evaluated as floating-point numbers. We generate numerically computed eigenpairs λ̃j,ṽj for 1 ≤ j ≤ 81 for each H̃Ξα using numpy’s Hermitian eigensolver eigh. We find that the smallest first positive eigenvalue of H̃Ξα for αG is 0.814 719 126 144 543 6 (computed using compute_PHalphaP_enclosures.py in the Github repo). Note that the difference between this number and 810 is bounded below by 0.01.

The main tool for part 3 of the strategy is the following theorem.

Theorem IV.2.
Letmandndenote positive integers withmn. LetAbe a Hermitiann × nmatrix, and let{vj}1jmbe orthonormaln-vectors satisfying (AλjI)vj = rjfor scalarsλjandn-vectorsrjfor each 1 ≤ jm. Then, there arem eigenvalues {αj}1jmofA, which can be put into one-to-one correspondence withλj’s such that
|λjαj|2msup1imri2forall1jm.

Proof.

See Appendix E 1.□

Naïvely, one would hope to be able to calculate enclosure intervals for every eigenvalue of HΞα and, in particular, a lower bound on the first positive eigenvalue of HΞα by directly applying Theorem IV.2 with A=HΞα, m = 81, and λj and vj given for each 1 ≤ j ≤ 81 by the approximate eigenpairs λ̃j,ṽj computed in part 2. However, we cannot directly apply the theorem because {ṽj}1j81 are not exactly orthonormal because of round-off error. Hence, we will prove the existence of an exactly orthonormal set {v̂j}1j81 close to the set {ṽj}1j81 and apply Theorem IV.2 to the set {v̂j}1j81 (with the same λ̃j) instead. Note that to carry out this strategy, we must bound the residuals r̂j(HΞαλ̃j)v̂j. The result we need to implement this strategy is the following. Note that the result requires numerical computation of inner products and residuals, and we account for round-off error in these computations.

Theorem IV.3.
Letmandnbe positive integers withmn. LetAbe ann × nHermitian matrix, letÃdenoteAevaluated in floating-point numbers, and letṽj,λ̃jfor 1 ≤ jmbe a set ofn-dimensional vectors and real numbers, respectively. Letṽiṽjcompdenote their numerically computed inner products, and letr̃j,compÃλ̃jIṽjcompdenote their numerically computed residuals. Letϵdenote machine epsilon, and assume < 0.01. Letμbe
μ(1.01)n2ϵsup1imṽi2+sup1im|ṽiṽicomp1|+supij1i,jm|ṽiṽjcomp|.
Then, as long asmμ<12, there is an orthonormal set ofn-vectors{v̂j}1jmwhose residualsr̂j(Aλ̃jI)v̂jsatisfy the bound
sup1jmr̂j221/2nA2+sup1jm|λ̃j|μ+n1/2sup1jmr̃j,comp+(1.01)n5/2ϵÃmax+sup1jm|λ̃j|sup1jmṽj+nϵAmaxsup1jmṽj,
(35)
whereAmaxdenotes the largest of the absolute values of the elements of the matrixA.

Proof.

See Appendix E 2.□

Numerical computation (using the script compute_PHalphaP_enclosures.py in the Github repo) shows that the maximum of sup1im|ṽiṽicomp1| and supij1i,jm|ṽiṽjcomp| over αG is bounded by 7 × 10−15. Hence, we can apply Theorem IV.3 with A=HΞα and λ̃j,ṽj given by the numerically computed eigenpairs of H̃Ξα to obtain orthonormal sets {v̂j}1j81 whose residuals with respect to HΞα satisfy (35). The following is straightforward.

Proposition IV.6.
sup0α710HΞα210,sup0α710HΞαmax7.

Proof.

The first estimate follows from ‖PΞH0PΞ‖ ≤ 7 and ‖H1‖ ≤ 3. The second estimate follows immediately from writing the matrix HΞα in the chiral basis.□

We can now apply Theorem IV.2 with A=HΞα and λj, vj given by the numerically computed λ̃j from part 2 and v̂j coming from Theorem IV.3, in order to derive rigorous enclosure intervals for every eigenvalue of HΞα. We find that (using the script compute_PHalphaP_enclosures.py in the Github repo) the suprema over αG of sup1jmṽj, sup1jmr̃j,comp,H̃Ξαmax, and sup1jm|λ̃j| are bounded by 1, 5 × 10−14, 7, and 8, respectively. It is then easy to see that 2 × 81 times the right-hand side of (35) is much smaller than 0.01 and is hence smaller than the distance between the minimum over αG of the numerically computed first positive eigenvalues of H̃Ξα and 810. We can therefore conclude that the first positive eigenvalues of HΞα are bounded below by 810 at every αG.

The main tool for part 4 of the strategy is the following.

Theorem IV.4.
LetAαbe ann × nHermitian matrix real-analytically depending on a real parameterα. Denote the ordered eigenvalues ofAαbyλjαfor 1 ≤ jn. Then, for anyαandα0,
|λjαλjα0||αα0|supβ[α0,α]βAβ2forall1jn.

Proof.

See Appendix E 3.□

We would like to apply Theorem IV.4 to bound the variation of eigenvalues of HΞα. To this end, we require the following proposition, which bounds the derivative of HΞα with respect to α over the interval 0α710.

Proposition IV.7.
sup0α710αHΞα238883.

Proof.

See Appendix E 4.□

Proposition IV.7 combined with Theorem IV.4 explains the choice of distance h=1388831 between grid points. Assuming that the first positive eigenvalue of HΞα is bounded below by 810 at grid points between 0 and 710 separated by h, we see that as long as

38883h2<81034h<1388830,

Proposition IV.7 and Theorem IV.4 guarantee that the first eigenvalue of HΞα must be greater than 34 over the whole interval 0α710.□

Remark IV.1.

In the Proof of Proposition IV.5, we bound the round-off error, which can occur in our numerical computations in order to draw rigorous conclusions. A common approach to this is the interval arithmetic; see Rump17 and references therein. Our approach applies directly to the present problem and is just as rigorous.

The results of a computation of the eigenvalues of HΞα are shown in Fig. 3.

FIG. 3.

Plot of numerically computed eigenvalues of the 81 × 81 matrix HΞα acting on PΞLK,12 (blue lines), showing that the first non-zero eigenvalues are bounded away from 0 by 34 (red lines) when α is less than 710 (black line). The zero eigenvalue corresponds to the subspace spanned by ψ8,α, and the non-zero eigenvalues equal those of the 80 × 80 matrix Qα,⊥PΞHαPΞQα,⊥ acting on Qα,PΞLK,12 since non-zero eigenvectors v of Qα,⊥PΞHαPΞQα,⊥ must be orthogonal to ψ8,α by orthogonality of eigenvectors corresponding to different eigenvalues.

FIG. 3.

Plot of numerically computed eigenvalues of the 81 × 81 matrix HΞα acting on PΞLK,12 (blue lines), showing that the first non-zero eigenvalues are bounded away from 0 by 34 (red lines) when α is less than 710 (black line). The zero eigenvalue corresponds to the subspace spanned by ψ8,α, and the non-zero eigenvalues equal those of the 80 × 80 matrix Qα,⊥PΞHαPΞQα,⊥ acting on Qα,PΞLK,12 since non-zero eigenvectors v of Qα,⊥PΞHαPΞQα,⊥ must be orthogonal to ψ8,α by orthogonality of eigenvectors corresponding to different eigenvalues.

Close modal

Assuming Proposition IV.4 and Proposition IV.5, the bound (31) becomes, for all 0α710,

Qα,HαQα,ηα34αQα,ηα.

We now assume the following, proved in  Appendix C.

Proposition IV.8.

H1Ψ8320.

We can now give the Proof of Theorem IV.1.

Proof of Theorem IV.1.

The proof follows immediately from Lemma IV.1, Proposition IV.4, Assumption IV.5, and Proposition IV.8.□

In the supplementary material, we list the chiral basis functions, which span the space Ξ, list terms Ψ5 − Ψ8 in the formal expansion of ψα, and derive the TKV Hamiltonian from the Bistritzer–MacDonald model.

M.L. and A.B.W. were supported, in part, by ARO MURI Award No. W911NF-14-0247 and NSF DMREF Award No. 1922165.

The data that support the findings of this study are available within the article and its supplementary material.

The Bloch eigenvalue problem for the TKV Hamiltonian at quasi-momentum k is

Hαψkα=Ekψkα,

where Hα is as in (1) and

ψkα(r+v)=eikvdiag(1,eiq1v,1,eiq1v)ψkα(r)vΛ.

By Propositions III.6 and III.7, 0 is a twofold (at least) degenerate eigenvalue at the moiré K point k = 0, with associated eigenfunctions ψ±1α as in Proposition III.7. In what follows, we assume that 0 is exactly twofold degenerate so that ψ±1α form a basis of the degenerate eigenspace. This assumption is clearly true for small α but could, in principle, be violated for α > 0.

Introducing χkαeikrψkα, we derive the equivalent Bloch eigenvalue problem with k-independent boundary conditions

Hkαχkα=Ekχkα,
(A1)

where

Hkα0DkαDkα0,Dkα=Dx+kx+i(Dy+ky)αU(r)αU(r)Dx+kx+i(Dy+ky),

where Dx,y ≔ −i∂x,y and

χkα(r+v)=diag(1,eiq1v,1,eiq1v)χkα(r)vΛ.

Clearly, ψ±1α remain a basis of the zero eigenspace for the problem (A1) at k = 0.

Differentiating the operator Dkα, we find kxDkα=I2 and kyDkα=iI2, where I2 denotes the 2 × 2 identity matrix, so that

kxHkα=0I2I20,kyHkα=0iI2iI20.
(A2)

By degenerate perturbation theory,18 for small k, we have that eigenfunctions χkα of (A1) are given by

χkασ=±1cσ,kψσα,

where the coefficients cσ,k and associated eigenvalues Ekϵk are found by solving the matrix eigenvalue problem

ψ1αkkH0αψ1αψ1αψ1αψ1αkkH0αψ1αψ1αψ1αψ1αkkH0αψ1αψ1αψ1αψ1αkkH0αψ1αψ1αψ1αc+1,kc1,k=ϵkc+1,kc1,k.
(A3)

Using (A2) and the explicit forms of ψ±1α given by Proposition III.7, we find that the matrix on the left-hand side of (A3) can be simplified to

0λ(α)(kxiky)λ*(α)(kx+iky)0,λ(α)ψ1α(r)ψ1α*(r)ψ1αψ1α.

It follows that, for small k, we have Ek ≈ ±v(α)|k|, where v(α) = |λ(α)| is as in (23).

1. The spectrum and eigenfunctions of H0 in LK2

The first task is to understand the spectrum and eigenfunctions of H0 in LK2. In the Appendix B 2, we will discuss the spectrum and eigenfunctions of H0 in LK,12. Recall that

H0=0D0D00,D0=2ī002ī,

where ̄=12(x+iy). To describe the eigenfunctions of H0 in LK2, we introduce some notation. Let v=v1,v2 be a vector in R2. Then, we will write

zv=v1+iv2,ẑv=v1+iv2|v|.

Finally, let V denote the area of the moiré cell Ω.

Proposition B.1.
The zero eigenspace ofH0inLK2is spanned by
χ±0=12V1,0,±1,0.
For allG ≠ 0 in the reciprocal lattice, then
χ±G(r)=12V1,0,±ẑG,0eiGr
are eigenfunctions with eigenvalues ±|G|. For allGin the reciprocal lattice,
χ±q1+G(r)=12V0,1,0,±ẑG+q1ei(q1+G)r
are eigenfunctions with eigenvalues ±|q1 + G|. The operatorH0has no other eigenfunctions inLK2other than linear combinations of these, and hence, the spectrum ofH0inLK2is
σLK2(H0)=±|G|,±|q1+G|:GΛ*.

Proof.

The proof is a straightforward calculation taking into account the LK2 boundary conditions given by (20) with k = 0. For example, e2 and e4 are zero eigenfunctions of H0 but in LK2, not LK2.□

Note that (as it must be because of the chiral symmetry) the spectrum is symmetric about 0 and all of the eigenfunctions with negative eigenvalues are given by applying S to the eigenfunctions with positive eigenvalues.

The union of the lattices Λ* and Λ* + q1 has the form of a honeycomb lattice in momentum space, where the lattice Λ* corresponds to “A” sites and Λ* + q1 corresponds to “B” sites (or vice versa); see Fig. 4.

FIG. 4.

Diagram showing A (blue) and B (red) sites of the momentum space lattice. Each site corresponds to two LK2-eigenvalues of H0, given by ± the distance between the site and the origin (black). The lattice vectors b1 and b2 are shown as well as the A site nearest neighbor vectors q1, q2, and q3.

FIG. 4.

Diagram showing A (blue) and B (red) sites of the momentum space lattice. Each site corresponds to two LK2-eigenvalues of H0, given by ± the distance between the site and the origin (black). The lattice vectors b1 and b2 are shown as well as the A site nearest neighbor vectors q1, q2, and q3.

Close modal

2. The spectrum and eigenfunctions of H0 in LK,12

We now discuss the spectrum of H0 in LK,12.

Proposition B.2.
The zero eigenspace ofH0inLK,12is spanned by
χ0̃1Ve1.
For allG ≠ 0 in the reciprocal lattice Λ*,
χ±G̃13k=02Rkχ±G=13k=02χ±(Rϕ*)kG
are eigenfunctions ofH0inLK,12with associated eigenvalues ±|G|. For allGin the reciprocal lattice Λ*,
χ±G+q1̃=13k=02Rkχ±G+q1=13k=02χ±(Rϕ*)k(G+q1)
are eigenfunctions ofH0inLK,12with associated eigenvalues ±|q1 + G|. The operatorH0has no other eigenfunctions inLK,12other than linear combinations of these, and hence, the spectrum ofH0inLK,12is
σLK,12(H0)=±|G|,±|q1+G|:GΛ*.

Proof.

The proof is another straightforward calculation starting from Proposition B.1.□

For an illustration of the support of the LK,12-eigenfunctions of H0 on the momentum space lattice, see Fig. 5. It is important to note that the notation introduced in Proposition B.2 is not one-to-one because, for example,

χ±G̃=χ±Rϕ*G̃=χ±(Rϕ*)2G̃

for any G ≠ 0 in Λ*.

FIG. 5.

Diagram showing support of LK,12-eigenfunctions of H0 superposed on the momentum space lattice. Each eigenfunction is given by superposing an LK2-eigenfunction of H0 with its rotations by 2π3 and 4π3. The support of the eigenfunctions χ±q1̃ with eigenvalues ±1 is shown with the black crosses, while the support of the eigenfunctions χ±b1̃ with eigenvalues ±3 is shown with the black circles.

FIG. 5.

Diagram showing support of LK,12-eigenfunctions of H0 superposed on the momentum space lattice. Each eigenfunction is given by superposing an LK2-eigenfunction of H0 with its rotations by 2π3 and 4π3. The support of the eigenfunctions χ±q1̃ with eigenvalues ±1 is shown with the black crosses, while the support of the eigenfunctions χ±b1̃ with eigenvalues ±3 is shown with the black circles.

Close modal

3. The chiral basis of LK,12

Recall that zero modes of Hα can be assumed to be eigenfunctions of the chiral symmetry operator S. It follows that the most convenient basis for our purposes is not be the spectral basis just introduced but the basis of LK,12 consisting of eigenfunctions of S. We call this basis the chiral basis.

Definition B.1.
The chiral basis ofLK,12is defined as the union of the functions
χ0̃=1Ve1,
χG̃,±112χG̃±χG̃,GΛ*\{0},
and
χq1+G̃,±112χq1+G̃±χq1+G̃,GΛ*.

The following is straightforward.

Proposition B.3.

The chiral basis is an orthonormal basis ofLK,12. The modesχ0̃,χG̃,1, andχq1+G̃,1are +1 eigenfunctions ofS, while the modesχG̃,1andχq1+G̃,1are −1 eigenfunctions ofS.

Written out, chiral basis functions have a very simple form. We have

χ0̃=1Ve1,
(B1)

and for all G ∈ Λ*\{0},

χG̃,1(r)=13Ve1k=02ei((Rϕ*)kG)r,χG̃,1(r)=13VẑGe3k=02eikϕei((Rϕ*)kG)r,
(B2)

and for all G ∈ Λ*,

χG+q1̃,1(r)=13Ve2k=02ei((Rϕ*)k(q1+G))r,χG+q1̃,1(r)=13VẑG+q1e4k=02eikϕei((Rϕ*)k(q1+G))r.
(B3)

We use the chiral basis to divide up LK,12 as follows.

Definition B.2.

We define spacesLK,1,±12to be the spans of the ±1 eigenfunctions ofSinLK,12, respectively.

Clearly, we have

LK,12=LK,1,12LK,1,12.

We can divide up the chiral basis more finely as follows.

Definition B.3.
We define
LK,1,1,A2χ0̃χG̃,1:GΛ*\{0},LK,1,1,B2χG+q1̃,1:GΛ*,LK,1,1,A2χG̃,1:GΛ*\{0},LK,1,1,B2χG+q1̃,1:GΛ*.

Remark B.1.

Note that notationsAandBin Definition B.3 refer toAandBsites of the momentum space lattice, not to theAandBsites of the real space lattice. Recalling Remark III.1 and comparing(B2)and(B3)with(2), we see thatLK,1,1,A2corresponds to wave-functions supported onAsites of layer 1,LK,1,1,B2corresponds to wave-functions supported onAsites of layer 2,LK,1,1,A2corresponds to wave-functions supported onBsites of layer 1, andLK,1,1,B2corresponds to wave-functions supported onBsites of layer 2.

Clearly, we have

LK,12=LK,1,1,A2LK,1,1,B2LK,1,1,A2LK,1,1,B2.

The following propositions are straightforward to prove. For the first claim, note that {S,H0}=0.

Proposition B.4.
OperatorH0mapsLK,1,±1,σ2LK,1,1,σ2forσ = A, B. The action ofH0on chiral basis functions is as follows:
H0χ0̃=0
for allG ∈ Λ*withG ≠ 0,
H0χG̃,±1=|G|χG̃,1,
and for allG ∈ Λ*,
H0χq1+G̃,±1=|q1+G|χq1+G̃,1.

Proposition B.5.
LetPdenote the projection operator ontoχ0̃inLK,12, andP = 1 − P. Then, the operatorP(H0)1PmapsLK,1,±1,σ2LK,1,1,σ2forσ = A, B, and
P(H0)1PχG̃,±1=1|G|χG̃,1
for allG ∈ Λ*withG ≠ 0, and
P(H0)1Pχq1+G̃,±1=1|q1+G|χq1+G̃,1
for allG ∈ Λ*.

In the Appendixes B 4 and 5, we will study the action of the operator H1 on LK,12 with respect to the chiral basis.

4. The spectrum of H1 in LK2 and LK,12

Recall that

H1=0D1D10,D1=0U(r)U(r)0,

where U(r)=eiq1r+eiϕeiq2r+eiϕeiq3r. We claim the following.

Proposition B.6.
For eachr0 ∈ Ω, ±|U(r0)| and ±|U(−r0)| are eigenvalues ofH1:LK2LK2. Forr0such thatU(r0) ≠ 0, the ±|U(r0)| eigenvectors are
0,1,±U(r0)|U(r0)|,0δ(rr0).
Forr0such thatU(−r0) ≠ 0, the ±|U(−r0)| eigenvectors are
1,0,0,±U(r0)|U(r0)|δ(rr0).
WhenU(r0) = 0, zero is a degenerate eigenvalue with associated eigenfunctionse2δ(rr0) ande3δ(rr0). WhenU(−r0) = 0, zero is a degenerate eigenvalue with associated eigenfunctionse1δ(rr0) ande4δ(rr0). Finally,
σLK2(H1)=[3,3].
(B4)

Proof.
We prove only (B4) since the other assertions are clear. The triangle inequality yields the obvious bound
|U(r0)|3
so that the LK2 spectrum of H1 must be contained in the interval [−3, 3]. To see that the spectrum actually equals [−3, 3], note that if r0=4π33,0; then,
q1r0=0,(q1+b1)r0=123,1r0=2π3,(q1+b2)r0=123,1r0=2π3,
and hence, U(r0) = 3. On the other hand, when r0 = 0, we have U(r0) = 0 so that the spectrum of H1 in LK2 equals [−3, 3].□

By taking linear combinations of rotated copies of the H1 eigenfunctions, just as we did with the H0 eigenfunctions, it is straightforward to prove an analogous result to Proposition B.6 in LK,12. We record only the following.

Proposition B.7.
σLK,12(H1)=[3,3].

5. The action of H1 on LK,12 with respect to the chiral basis

We now want to study the action of H1 on LK,12 with respect to the chiral basis. We will prove two propositions, which parallel Proposition B.4.

Proposition B.8.
The operatorH1mapsLK,1,1,A2LK,1,1,B2andLK,1,1,B2LK,1,1,A2. The action ofH1on chiral basis functions is asfollows:
H1χ0̃=3ẑq1̄χq1̃,1
(B5)
and
H1χq1̃,1=eiϕẑq1q2̄χq1q2̃,1+eiϕẑq1q3̄χq1q3̃,1.
(B6)
For allG ∈ Λ*\{0},
H1χG̃,1=ẑG+q1̄χG+q1̃,1+eiϕẑG+q2̄χG+q2̃,1+eiϕẑG+q3̄χG+q3̃,1.
(B7)
For allG ∈ Λ*\{0},
H1χG+q1̃,1=ẑḠχG̃,1+eiϕẑG+q1q2̄χG+q1q2̃,1+eiϕẑG+q1q3̄χG+q1q3̃,1.
(B8)

Note that H1 exchanges chirality (S eigenvalue) and the A and B momentum space sublattices, while H0 only exchanges chirality. Proposition B.8 has a simple interpretation in terms of nearest neighbor hopping in the momentum space lattice; see Figs. 6 and 7.

FIG. 6.

Illustration of the action of H1 in LK,12 as hopping in the momentum space lattice described by Eq. (B7) (left, starting at b1) and (B8) (right, starting at q1 + b1b2). The origin is marked by a black dot.

FIG. 6.

Illustration of the action of H1 in LK,12 as hopping in the momentum space lattice described by Eq. (B7) (left, starting at b1) and (B8) (right, starting at q1 + b1b2). The origin is marked by a black dot.

Close modal
FIG. 7.

Illustration of the action of H1 as hopping in the momentum space lattice described by Eq. (B5) (left, starting at 0) and (B6) (right, starting at q1). Although it appears that the hopping in these cases does not respect 2π3 rotation symmetry, this is an artifact of working with chiral basis functions, which individually respect the rotation symmetry; see (B9).

FIG. 7.

Illustration of the action of H1 as hopping in the momentum space lattice described by Eq. (B5) (left, starting at 0) and (B6) (right, starting at q1). Although it appears that the hopping in these cases does not respect 2π3 rotation symmetry, this is an artifact of working with chiral basis functions, which individually respect the rotation symmetry; see (B9).

Close modal

Remark B.2.
At first glance,(B5) and (B6) appear different from(B7)and(B8)because they appear to violate2π3rotation symmetry. However, this is not the case since every chiral basis function individually respects this symmetry. For example, usingχq1̃,1=χq2̃,1=χq3̃,1andẑq1̄=eiϕẑq2̄=eiϕẑq3̄, we can re-write(B5)in a way that manifestly respects the2π3rotation symmetry as
H1χ0̃=13ẑq1̄χq1̃,1+eiϕẑq2̄χq2̃,1+eiϕẑq3̄χq3̃,1.
(B9)
Equation (B6)can also be written in a manifestly rotationally invariant way, but the expression is long, and hence, we omit it. Note that(B6)cannot have a term proportional toχ0̃sinceχ0̃LK,1,12andH1mapsLK,1,12LK,1,12.

Proof of Proposition B.8.
We will prove (B7); the proofs of the other identities are similar and hence omitted. We have
H1χG̃,1=13Veiq1r+eiϕei(q1+b1)r+eiϕei(q1+b2)reiGr+ei(Rϕ*G)r+ei((Rϕ*)2G)re4.
Multiplying out, we have
13Vei(q1+G)r+eiϕei(q1+G+b1)r+eiϕei(q1+G+b2)r+ei(q1+(Rϕ*G))r+eiϕei(q1+(Rϕ*G)+b1)r+eiϕei(q1+(Rϕ*G)+b2)r+ei(q1+((Rϕ*)2G))r+eiϕei(q1+((Rϕ*)2G)+b1)r+eiϕei(q1+((Rϕ*)2G)+b2)r=13Vei(q1+G)r+eiϕei(q1+G+b1)r+eiϕei(q1+G+b2)rei(Rϕ*(q1+G+b1))r+eiϕei(Rϕ*(q1+G+b2))r+eiϕei(Rϕ*(q1+G))r+ei((Rϕ*)2(q1+G+b2))r+eiϕei((Rϕ*)2(q1+G))r+eiϕei(Rϕ*)2(q1+G+b1)r=13Vei(q1+G)r+eiϕei(Rϕ*(q1+G))r+eiϕei((Rϕ*)2(q1+G))r+13Veiϕei(q1+G+b1)r+eiϕei(Rϕ*(q1+G+b1))r+eiϕei((Rϕ*)2(q1+G+b1))r+13Veiϕei(q1+G+b2)r+eiϕei(Rϕ*(q1+G+b2))r+eiϕei((Rϕ*)2(q1+G+b2))r,
from which (B7) follows.□

Proposition B.9.
The operatorH1mapsLK,1,1,A2LK,1,1,B2andLK,1,1,B2LK,1,1,A2. The action ofH1on chiral basis functions is as follows:
H1χq1̃,1=ẑq13χ0̃+eiϕχq1q2̃,1+eiϕχq1q3̃,1.
For allG ∈ Λ*\{0},
H1χG̃,1=ẑGχG+q1̃,1+eiϕχG+q2̃,1+eiϕχG+q3̃,1.
For allG ∈ Λ*,
H1χG+q1̃,1=ẑG+q1χG̃,1+eiϕχG+q1q2̃,1+eiϕχG+q1q3̃,1.

Proof.

The proof is similar to that of Proposition B.8 and is hence omitted.□

We now bring to bear the developments of the  Appendix B on the asymptotic expansion of the zero mode ψαLK,1,12 starting from Ψ0=e1=χ0̃. We first give the Proof of Proposition IV.1.

Proof of Proposition IV.1.
We have seen that χ0̃LK,1,12. By the calculations of the Appendix B 5, H1χ0̃LK,1,12, which is orthogonal to the null space of H0. The general solution of H0Ψ1 = −H1Ψ0 is
Ψ1=P(H0)1PH1Ψ0+CΨ0,
where C is an arbitrary constant, which is in LK,1,12 by Proposition B.4. To ensure that Ψ1 is orthogonal to Ψ0, we take C = 0. It is clear that this procedure can be repeated to derive an expansion to all orders satisfying the conditions of Proposition IV.1.□

Our goal is to calculate ΨnLK,1,12 satisfying the conditions of Proposition B.4 up to n = 8. This amounts to calculating, for n = 1 to n = 8,

Ψn=P(H0)1PH1Ψn1.

We do this algorithmically by repeated application of the following proposition, which combines Proposition B.8 and Proposition B.5.

Proposition C.1.
The operatorP(H0)1PH1mapsLK,1,1,A2LK,1,1,B2andLK,1,1,B2LK,1,1,A2. Its action on chiral basis functions is as follows:
P(H0)1PH1χ0̃=3ẑq1̄χq1̃,1
(C1)
and
P(H0)1PH1χq1̃,1=eiϕẑq1q2̄|q1q2|χq1q2̃,1eiϕẑq1q3̄|q1q3|χq1q3̃,1.
(C2)
For allG ∈ Λ*\{0},
P(H0)1PH1χG̃,1=ẑG+q1̄|G+q1|χG+q1̃,1eiϕẑG+q2̄|G+q2|χG+q2̃,1eiϕẑG+q3̄|G+q3|χG+q3̃,1.
(C3)
For allG ∈ Λ*\{0},
P(H0)1PH1χG+q1̃,1=ẑḠ|G|χG̃,1eiϕẑG+q1q2̄|G+q1q2|χG+q1q2̃,1eiϕẑG+q1q3̄|G+q1q3|χG+q1q3̃,1.

We now claim the following.

Proposition C.2.
Let Ψnbe the sequence defined by Proposition IV.1. Then,
Ψ1=3iχq1̃,1,
(C4)
Ψ2=3i2χb1̃,1+3+i2χb2̃,1,
(C5)
Ψ3=177321i14χq1b2̃,1+177321i14χq1b1̃,1,
(C6)
Ψ4=12157+21i14χb2̃,1+122127+21i7χ2b2̃,1+1215721i14χb1̃,1+12212721i7χ2b1̃+2321χb1b2̃,1.
(C7)

Proof.
Equations (C4) and (C5) follow immediately from (C1) and (C2) and using q2 = q1 + b1 and q3 = q1 + b2. The derivation of Eq. (C6) is more involved, so we give details. Using linearity and applying (C3) twice, we find that
P(H0)1PH1Ψ2=3i2ẑq1b2̄|q1b2|χq1b2̃,1+eiϕẑq1+b1b2̄|q1+b1b2|χq1+b1b2̃,1+eiϕẑq1̄χq1̃,1+3+i2ẑq1b1̄|q1b1|χq1b1̃,1+eiϕẑq1̄χq1̃,1+eiϕẑq1+b2b1̄|q1+b2b1|χq1+b2b1̃,1.
First, the terms proportional to χq1̃,1 cancel. Next, since Rϕ(q1 + b1b2) = q1 + b2b1, we have χq1+b1b2̃,1=χq1+b2b1̃,1. These terms also cancel, leaving (C6). The derivation of (C7) (and the higher corrections) is involved but does not depend on any new ideas and is therefore omitted.□

We give the explicit forms of Ψ58 in the supplementary material.

Remark C.1.
Written out,(C4)and(C5)become
Ψ1=3i13Ve2eiq1r+eiq2r+eiq3r
and
Ψ2=ieiϕ13Ve1eib1r+ei(b2b1)r+eib2r+ieiϕ13Ve1eib2r+eib1r+ei(b1b2)r,
which agree withEq. (24)of Tarnopolsky et al.4 up to an overall factor ofV(this factor cancels in the Fermi velocity, so there is no discrepancy).

Using orthonormality of the chiral basis functions, it is straightforward to calculate the norms of each of the Ψn. We have the following.

Proposition C.3.
Ψ0=1,Ψ1=3,Ψ2=2,Ψ3=147,Ψ4=25842,Ψ5=19688373458,
Ψ6=10652579931122,Ψ7=22129312323981473624696345,Ψ8=1836431197552144544997570760.

Remark C.2.

Note that the sequence of norms of the expansion functions grows much slower than the pessimistic bound ‖ΨN+1‖ ≤ 3‖ΨN‖, N = 0, 1, 2, … guaranteed by Proposition IV.2. The reason is that bounds(27)and(28)are never attained. AsNbecomes larger, the bound(27)is very pessimistic because ΨNis mostly made up of eigenfunctions ofH0with eigenvalues strictly larger than 1. The bound(28)is also very pessimistic because it is attained only at delta functions, which can only be approximated with a superposition of a large number of eigenfunctions ofH0. It seems possible that a sharper bound could be proved starting from these observations, but we do not pursue this in this work.

We finally give the Proof of Proposition IV.8.

Proof of Proposition IV.8.
Explicit computation using Proposition B.8 and orthonormality of the chiral basis functions gives
H1Ψ8=4855076200233765642149927122800.147320.

We choose Ξ as

ΞLK,12-eigenfunctions of H0 witheigenvalues with magnitude 43χq14b1+b2̃,±1,χq1+b14b2̃,±1.

Part 1 of Proposition IV.4 follows immediately from observing that χq12b12b2̃,±1 is not in Ξ, but |q1 − 2b1 − 2b2| = 7. That μ = 7 is optimal can be seen from Fig. 8.

FIG. 8.

Illustration of Ξ in the momentum space lattice. The circle has radius 43 so that every dot within the circle corresponds to two chiral basis vectors included in Ξ. Chiral basis vectors exactly 43 away from the origin, marked with black dots, are also included in Ξ. We also include in Ξ the chiral basis vectors χq14b1+b2̃,±1,χq1+b14b2̃,±1, which correspond to the dots marked with circles, which are distance 7 (NB. 7>43) from the origin. We do not include the chiral basis vectors χq12b12b2̃,±1, marked with the black crosses, which are also a distance 7 from the origin. The reason for this is so that part 3 of Proposition IV.4 holds. With this choice, every dot in Ξ has at most one nearest neighbor lattice point outside of Ξ. It follows immediately from Propositions B.8 and B.9 (H1 acts by nearest neighbor hopping in the momentum space lattice) that PΞH1PΞ=1. Note that if we chose Ξ to include χq12b12b2̃,±1, this would no longer hold because these basis functions would have two nearest neighbors outside Ξ, resulting in the worse bound PΞH1PΞ2.

FIG. 8.

Illustration of Ξ in the momentum space lattice. The circle has radius 43 so that every dot within the circle corresponds to two chiral basis vectors included in Ξ. Chiral basis vectors exactly 43 away from the origin, marked with black dots, are also included in Ξ. We also include in Ξ the chiral basis vectors χq14b1+b2̃,±1,χq1+b14b2̃,±1, which correspond to the dots marked with circles, which are distance 7 (NB. 7>43) from the origin. We do not include the chiral basis vectors χq12b12b2̃,±1, marked with the black crosses, which are also a distance 7 from the origin. The reason for this is so that part 3 of Proposition IV.4 holds. With this choice, every dot in Ξ has at most one nearest neighbor lattice point outside of Ξ. It follows immediately from Propositions B.8 and B.9 (H1 acts by nearest neighbor hopping in the momentum space lattice) that PΞH1PΞ=1. Note that if we chose Ξ to include χq12b12b2̃,±1, this would no longer hold because these basis functions would have two nearest neighbors outside Ξ, resulting in the worse bound PΞH1PΞ2.

Close modal

Part 2 follows from the fact that ψ8,α depends only on eigenfunctions of H0 with eigenvalues with magnitude less than or equal to 43. The largest eigenvalue is 43, coming from dependence of Ψ8 on χ4b2̃,1, since |4b2|=43.

Part 3 can be seen from Fig. 8.

1. Proof of Theorem IV.2

We will prove Theorem IV.2 starting from Theorem 11.5.1 of Parlett,19 where the proof can be found.

Lemma E.1.
LetQbe ann × mmatrix, which satisfiesQQ = Im. DefineH = QAQandR = AQQH. Let{θj}1jmdenote the eigenvalues ofH(the Ritz values). Then,mofA’s eigenvalues{αj}1jmcan be put into one-to-one correspondence with the{θj}1jmin such a waythat
|θjαj|R21jm.

Proof of Theorem IV.2.
Let Q be the matrix whose columns are v1, …, vm. Using orthonormality of the vj, QQ = Im and
H=QAQ=diag(λ1,,λm)+v1r1v1rmvnr1vnrn.
We now prove that the eigenvalues of H, denoted by θj, are close to λj’s. By the Gershgorin circle theorem, we have
|θiλi+viri|jim|virj|,
which implies, using ‖vj2 = 1,
|θiλi|=|θiλiviri+viri|j=1m|virj|msup1imri2.
We can now use Lemma E.1 to bound the difference between λjs and exact eigenvalues αj,
|λjαj|=|λjθj+θjαj|msup1imri2+R2,
where RAQQH = (IQQ)AQ. Since QQ projects onto vj, R simplifies to
R=(IQQ)R,Rr1rm.
Since QQ is a projection, so is IQQ, and hence, ‖R2 ≤ ‖R′‖2. To bound ‖R′‖2, note that for any v with ‖v2 = 1, we have
Rv2=e1vr1++emvrmmsup1imri2,
where ej denote the standard orthonormal basis vectors. The result now follows.□

2. Proof of Theorem IV.3

Proof of Theorem IV.3.

We start with the following lemma, which guarantees that numerically computed approximately orthonormal sets can be approximated by exactly orthonormal sets.

Lemma E.2.
Letṽ1,,ṽmben-dimensional vectors, letṽiṽjcompfor 1 ≤ i, jmdenote their numerically computed inner products, letϵdenote machine epsilon, and assume that < 0.01. Define
μ(1.01)n2ϵsup1imṽi2+supi|ṽiṽicomp1|+supij|ṽiṽjcomp|.
(E1)
Then, as long asmμ<12, there is a set ofn-dimensional orthonormal vectorsv̂1,,v̂m, which satisfy
v̂jṽj221/2mμ,1jm.

Proof.
Bounding the round-off error in computing inner products in the usual way (see, for example, Chap. 2.7 of Golub and Van Loan20) and assuming that < 0.01, we have that for each 1 ≤ i, jm, ṽiṽj=ṽiṽjcomp+eij, where |eij|(1.01)nϵ|ṽi||ṽj|(1.01)nϵṽi2ṽj2. Letting Q̃ denote the matrix whose columns are ṽi’s, then Q̃Q̃Im=E, where, for all ij, |Eij||ṽiṽjcomp|+(1.01)nϵṽi2ṽj2, and for all i, |Eii||1ṽiṽicomp|+(1.01)nϵṽi22. Paying the price of factors of n to replace ‖·‖2 norms by ‖·‖ norms, we can obtain a trivial bound on the maximal element of E, denoted as ‖Emax, by
Emax(1.01)n2ϵsupiṽi2+supi|ṽiṽicomp1|+supi,j|ṽiṽjcomp|.
Note that this is nothing but μ in the statement of the theorem. Using the Gershgorin circle theorem, we then have that the eigenvalues λ of Q̃Q̃ satisfy |λ − 1| ≤ mEmax. We claim that there are exact orthonormal vectors v̂j near (in the ‖·‖2-norm) to the ṽj. To see this, note that Q̂Q̃(Q̃Q̃)1/2 satisfies Q̂Q̂=Im and
Q̂Q̃2Q̂2(Q̃Q̃)1/2Im2=(Q̃Q̃)1/2Im2.
Let λmax and λmin denote the maximum and minimum eigenvalues of Q̃Q̃, respectively. Then, (Q̃Q̃)1/2Im2max|λmin1/21|,|λmax1/21|. Since λmin is bounded below by 1 − mEmax and λmax is bounded above by 1 + mEmax, we have
(Q̃Q̃)1/2Im2max|(1+mEmax)1/21|,|(1mEmax)1/21|.
Using Taylor’s theorem, for |x|<12, we have that |(1 + x)1/2 − 1| ≤ 2−1/2|x| and |(1 − x)1/2 − 1| ≤ 2−1/2|x|. Since by assumption mEmax<12, we conclude
Q̃Q̂2(Q̃Q̃)1/2Im221/2mEmax.
Letting v̂j denote the columns of Q̂ and noting that v̂jṽj2Q̂Q̃2 for all 1 ≤ jm, the result is proved.□

Using Lemma E.2, we have that there exists an exactly orthonormal set {v̂j}1jm nearby to the set {ṽj}1jm. We now want to bound the residuals of the v̂j in terms of numerically computable quantities. We start with the following easy lemma whose proof is a straightforward manipulation.

Lemma E.3.
LetAbe ann × nHermitian matrix, and suppose thatr̂(Aλ̃I)v̂andr̃(Aλ̃I)ṽ. Then,
r̂2A2+|λ̃|v̂ṽ2+r̃2.

The following lemma quantifies the error in approximating exact residuals by numerically computed values.

Lemma E.4.
LetAbe a Hermitiann × nmatrix, and letÃdenote the matrix whose entries are those ofAevaluated as floating-point numbers. LetÃλ̃Iṽcompdenote the numerically computed value ofÃλ̃Iṽin the floating-point arithmetic. Then,r̃(Aλ̃I)ṽsatisfies
r̃2n1/2(Ãλ̃I)ṽcompmax+(1.01)n5/2ϵÃλ̃Imaxṽ+nϵAmaxṽ.

Proof.
For matrices A and B with entries Aij and Bij, we will write |A| to denote the matrix with entries |Aij| for all i, j, and |A| ≤ |B| if |Aij| ≤ |Bij| for all i, j. It is straightforward to see that (see Chap. 2.7 of Golub and Van Loan20) Ã=A+F, where |F| < ϵ|A|. In addition, (Ãλ̃I)ṽ=(Ãλ̃I)ṽcomp+g, where |g|(1.01)nϵ|(Ãλ̃I)||ṽ|. Now, note that (Aλ̃I)ṽ=(Ãλ̃I)ṽFṽ so that
r̃2(Ãλ̃I)ṽcomp2+g2+Fṽ2.
Noting that
g2=(1.01)nϵ|Ãλ̃I||ṽ|2(1.01)n5/2ϵÃλ̃Imaxṽ,
Fṽ2ϵ|A||ṽ|2nϵAmaxṽ,
and
(Ãλ̃I)ṽcomp2n1/2(Ãλ̃I)ṽcompmax,
the result is proved.□

We now prove estimate (35). Applying Lemma E.2 to the set {ṽj}1jm yields an orthonormal set {v̂j}1jm such that v̂jṽj221/2mμ, where μ is as in (E1). By Lemma E.3, we have that

r̂j221/2mA2+|λ̃j|μ+r̃j2,1jm.

The estimate now follows easily upon applying Lemma E.4 and taking the sup over j.□

3. Proof of Theorem IV.4

Proof of Theorem IV.4.
The proof is a simple consequence of the min–max characterization of eigenvalues of Hermitian matrices. By min–max (here, U denotes a subspace of Cn),
|λjαλjα0|=mindimU=jmaxvUv0v(AαAα0)vvv.
On the other hand, for any fixed v ≠ 0, we have by Taylor’s theorem
v(AαAα0)vvv|αα0|supβ[α0,α]βv(AβAα0)vvv|αα0|supβ[α0,α]βAβ2,
and the result follows immediately.□

4. Proof of Proposition IV.7

Proof of Proposition IV.7.
Differentiating HΞα yields
αHΞα=(αQα)PΞHαPΞQα,+Qα,PΞH1PΞQα,+Qα,PΞHαPΞ(αQα).
For α < 1, we have ‖PΞHαPΞ2 ≤ 10, and ‖H12 ≤ 3. It remains only to estimate ‖αQα2. Using the Dirac notation to represent LK,12-projections, we have
Qα=Ψ(8)Ψ(8)=m=08αmΨmn=08αnΨn=m=08n=08αm+nΨmΨn
so that
αQα=m=08n=08(m+n)αm+n1ΨmΨn.
Using ΨmΨn2Ψm2Ψn2 and max0j8Ψj23 by Proposition C.3, we have, for α ≤ 1,
αQα23m=08n=08(m+n)=1944.
Putting everything together, we conclude
sup0α710αHΞα22×10×1944+3=38883.

We can now prove Proposition II.1. We start by proving (11).

1. Proof of (11)

We now prove (11). It is straightforward to derive

n=08αnΨnn=08αnΨn=n=08j=0nΨjΨnjαn+n=07j=0nΨ8jΨ8(nj)α16n.
(F1)

We now make two observations which simplify the computation. First, recall that the operator P(H0)1PH1 maps LK,1,1,A2LK,1,1,B2 and LK,1,1,B2LK,1,1,A2. It follows that Ψ0LK,1,1,A2, Ψ1LK,1,1,B2, Ψ2LK,1,1,A2, and so on, and hence,

Ψ2iΨ2j+1=0i,j{0,1,2,}.

It follows that all terms in (F1) with odd powers of α vanish. Second, note that since Ψ0 ∈ ranP, while Ψn ∈ ranP for all n ≥ 1, we have that

ΨnΨ0=Ψ0Ψn=0n{1,2,}.

Deriving (11) is then just a matter of computation using the properties of the chiral basis. For the leading term, we have

Ψ0Ψ0=χ0̃χ0̃=1.

For the α2 term, the only non-zero term is

Ψ1Ψ1=3iχq1̃,13iχq1̃,1=3,

using (C4). For the α4 term, the possible non-zero terms are

Ψ3Ψ1+Ψ2Ψ2+Ψ1Ψ3,

but Ψ3 and Ψ1 depend on orthogonal chiral basis vectors [see (C4) and (C6)], so we are left with

Ψ2Ψ2=3i2χb1̃,1+3+i2χb2̃,13i2χb1̃,1+3+i2χb2̃,1=2,

using (C5) and orthgonality of χb1̃,1 and χb2̃,1. We omit the derivation of the higher terms since the derivations do not require any new ideas.

2. Proof of (10)

It is straightforward to derive

n=08αnΨn*(r)n=08αnΨn(r)=n=08j=0nΨj*(r)Ψnj(r)αn+n=07j=0nΨ8j*(r)Ψ8(nj)(r)α16n.
(F2)

We now note the following.

Proposition F.1.

Letχbe a chiral basis function inLK,1,12. Then,χ*(−r) = χ(r).

Proof.

The proof follows immediately from the explicit forms of the chiral basis functions in LK,1,12 given by (B1)–(B2)–(B3) and the observation that for any kR2, eik(r)*=eikr.□

Using Proposition F.1 and the same two observations as in the Appendix F 1, we have that the only non-zero terms in (F2) are those with even powers of α and that other than the leading term, terms involving Ψ0 do not contribute. The calculation is then similar to the previous case. For the leading order term, we have

Ψ0*(r)Ψ0(r)=χ0̃χ0̃=1.

The only non-zero α2 term is

Ψ1*(r)Ψ1(r)=3iχq1̃,13iχq1̃,1=3.

The only non-zero α4 term is

Ψ2*(r)Ψ2(r)=3+i2χb1̃,1+3i2χb2̃,13i2χb1̃,1+3+i2χb2̃,1=3i22+3+i22=1.

We omit the derivation of the higher terms since the derivations do not require any new ideas.

Proposition IV.1 implies that the series expansion of ψα exists up to any order. We can therefore define formal infinite series by

n=0αnΨn*(r)n=0αnΨn(r),
(F3)
n=0αnΨnn=0αnΨn.
(F4)

We then have the following.

Proposition F.2.

Expansions(10)and(11)approximate the formal series(F3)and(F4)up to terms of orderα10.

Proof.

The series agree exactly without any simplifications up to terms of α9. However, because the even and odd terms in the expansion of ψα are orthogonal (since they lie in LK,1,1,A2 and LK,1,1,B2, respectively), all terms with odd powers of α vanish in expansions (F3) and (F4). The series may disagree at order α10 because the infinite series includes terms arising from inner products of Ψ1 and Ψ9.□

1.
R.
Bistritzer
and
A. H.
MacDonald
,
Proc. Natl. Acad. Sci. U. S. A.
108
,
12233
(
2011
); arXiv:1009.4203.
2.
P.
San-Jose
,
J.
González
, and
F.
Guinea
,
Phys. Rev. Lett.
108
,
216802
(
2012
); arXiv:1110.2883.
3.
Y.
Cao
,
V.
Fatemi
,
S.
Fang
,
K.
Watanabe
,
T.
Taniguchi
,
E.
Kaxiras
, and
P.
Jarillo-Herrero
,
Nature
556
,
43
(
2018
).
4.
G.
Tarnopolsky
,
A. J.
Kruchkov
, and
A.
Vishwanath
,
Phys. Rev. Lett.
122
,
106405
(
2019
); arXiv:1808.05250.
5.
S.
Becker
,
M.
Embree
,
J.
Wittsten
, and
M.
Zworski
, “
Mathematics of magic angles in a model of twisted bilayer graphene
,” arXiv:2008.08489 [math-ph] (
2020
).
6.
S.
Becker
,
M.
Embree
,
J.
Wittsten
, and
M.
Zworski
,
Phys. Rev. B
103
,
165113
(
2020
); arXiv:2010.05279.
7.
G.
Catarina
,
B.
Amorim
,
E. V.
Castro
,
J. M. V. P.
Lopes
, and
N.
Peres
,
Handbook of Graphene
(
Wiley
,
2019
), pp.
177
231
.
8.
J.
Oliver
,
J. Comput. Appl. Math.
5
,
85
(
1979
).
9.
A.
Meurer
,
C. P.
Smith
,
M.
Paprocki
,
O.
Čertík
,
S. B.
Kirpichev
,
M.
Rocklin
,
A.
Kumar
,
S.
Ivanov
,
J. K.
Moore
,
S.
Singh
,
T.
Rathnayake
,
S.
Vig
,
B. E.
Granger
,
R. P.
Muller
,
F.
Bonazzi
,
H.
Gupta
,
S.
Vats
,
F.
Johansson
,
F.
Pedregosa
,
M. J.
Curry
,
A. R.
Terrel
,
Š.
Roučka
,
A.
Saboo
,
I.
Fernando
,
S.
Kulal
,
R.
Cimrman
, and
A.
Scopatz
,
PeerJ Comput. Sci.
3
,
e103
(
2017
).
10.
D.
Massatt
,
S.
Carr
,
M.
Luskin
, and
C.
Ortner
,
Multiscale Model. Simul.
16
,
429
(
2018
).
11.
L.
Zou
,
H. C.
Po
,
A.
Vishwanath
, and
T.
Senthil
,
Phys. Rev. B
98
,
085435
(
2018
); arXiv:1806.07873.
12.
H. C.
Po
,
L.
Zou
,
A.
Vishwanath
, and
T.
Senthil
,
Phys. Rev. X
8
,
031089
(
2018
); arXiv:1803.09742.
13.
H. C.
Po
,
L.
Zou
,
T.
Senthil
, and
A.
Vishwanath
,
Phys. Rev. B
99
,
195455
(
2019
); arXiv:1808.02482.
14.
T.
Kato
,
Perturbation Theory for Linear Operators
(
Springer-Verlag Berlin Heidelberg
,
1995
), Vol. 132.
15.
H. L.
Cycon
,
R. G.
Froese
,
W.
Kirsch
, and
B.
Simon
,
Schrödinger Operators: With Applications to Quantum Mechanics and Global Geometry
(
Springer-Verlag Berlin Heidelberg
,
1987
).
16.
M. E.
Taylor
,
Partial Differential Equations I
(
Springer
,
New York, NY
,
1996
).
17.
S. M.
Rump
,
Acta Numer.
19
,
287
449
(
2010
).
18.
A.
Messiah
,
Quantum Mechanics: Volume II
(
North-Holland Publishing Company
,
Amsterdam
,
1962
).
19.
B. N.
Parlett
,
The Symmetric Eigenvalue Problem
(
Society for Industrial and Applied Mathematics
,
1998
).
20.
G. H.
Golub
and
C. F.
Van Loan
,
Matrix Computations
, 4th ed. (
The Johns Hopkins University Press
,
2013
).

Supplementary Material