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

\NewEnviron

equations

\BODY\begin{gathered}\BODY\end{gathered} (0.1)

The bulk-edge correspondence for curved interfaces

Alexis Drouot University of Washington, Seattle, USA. [email protected]  and  Xiaowen Zhu University of Washington, Seattle, USA. [email protected]
Abstract.

The bulk-edge correspondence is a condensed matter theorem that relates the conductance of a Hall insulator in a half-plane to that of its (straight) boundary. In this work, we extend this result to domains with curved boundaries. Under mild geometric assumptions, we prove that the edge conductance of a topological insulator sample is an integer multiple of its Hall conductance. This integer counts the algebraic number of times that the interface (suitably oriented) enters the measurement set. This result provides a rigorous proof of a well-known experimental observation: arbitrarily truncated topological insulators support edge currents, regardless of the shape of their boundary.

1. Introduction and Main results

1.1. Introduction

The study of topological insulators is a central topic in condensed matter physics. These materials are insulating phases of matter, described by a Hamiltonian with a spectral gap, to which one can associate a quantized topological invariant: the Hall conductance. When two insulators with distinct topological invariants are glued together, protected gapless currents emerge along the interface: the material becomes a conductor. For straight interfaces, the interface conductance is also quantized and equals the difference of the bulk topological invariants. This fundamental result is called the bulk-edge correspondence.

The quantization of the bulk conductance (i.e. the fact that it takes values in a discrete set) was first observed in the quantum Hall effect [L81, TKNN, H82]. The characterization of Hall conductances as a Chern number marked the birth of topological phases of matter. Haldane [H88] demonstrated on a famous model of magnetic graphene that this phenomenon is not restricted to quantum Hall systems. Hatsugai [Hatsugai] then showed that topological materials support gapless states along their boundary; this is known today as the bulk-edge correspondence. Since then, the bulk-edge correspondence has been extended to various situations: see [KM05a, KM05b, FKM07] for 2\mathbb{Z}_{2}-topological insulators, [GT18, ST19] for Floquet topological systems, and [HR08, RH08, YGS15, DMV17, PDV19, F19] for physical fields beyond quantum science, such as photonics, accounstics and fluid mechanics. By now mathematical proofs of the bulk-edge correspondence have spanned a wide variety of situations: see for instance [Hatsugai, KS02, EG02, KS04, L15, Ku17, Br18, LT22, L23] for discrete models, [EGS05, LH11, GS18, ST19] for disordered systems, [GP13, ASV13, FSSWY20] for 2\mathbb{Z}_{2}-topological insulators, and [B19, D21b, D21, SW22a] for continuous Hamiltonians.

Various experimental results have suggested that the bulk-edge correspondence is an extremely stable principle: it does not depend on the fine details of the interface [WMTHXHW17, NHCR18, FSSWY20, TJCIBBP21]. However, with the exception of some KK-theoretic approaches [Ku17, LT22, L23], most proofs of the bulk-edge correspondence have focused on straight interfaces. In this work, we provide a spectral proof of the bulk-edge correspondence for curved interfaces. We show equality between the edge conductance and an integer multiple of the Hall conductance; this goes beyond the aforementioned KK-theoretic results where the bulk index takes the form of a geometric Hall conductance. We provide a physical interpretation of the emerging integer as an intersection number between the boundary and the measurement set. To the best of our knowledge, this is the first time that such a quantity emerges in the study of topological insulators.

1.2. Main result

We briefly review standard facts from condensed matter physics. Electronic propagation in a quantum material follows the Schrödinger equation with a selfadjoint operator HH on 2(2,m)\ell^{2}(\mathbb{Z}^{2},\mathbb{C}^{m}). The operator HH describes the hopping of the electron between atomic sites. Its spectrum Σ(H)\Sigma(H) characterizes the electronic nature of the material: HH is a conductor at energy λ\lambda if and only if λΣ(H)\lambda\in\Sigma(H); and an insulator otherwise. In this paper, we will work with Hamiltonians that model limited hopping:

Definition 1 (ESR).

We say that a selfadjoint operator HH on 2(2,m)\ell^{2}(\mathbb{Z}^{2},\mathbb{C}^{m}) is exponentially-short-range (ESR) if its kernel satisfies, for some ν>0\nu>0,

|H(𝒙,𝒚)|ν1e2νd1(𝒙,𝒚),𝒙,𝒚2.|H({\bm{x}},{\bm{y}})|\leq\nu^{-1}e^{-2\nu d_{1}({\bm{x}},{\bm{y}})},\qquad\forall{\bm{x}},{\bm{y}}\in\mathbb{Z}^{2}. (1.1)

In (1.1), d1(𝒙,𝒚)d_{1}({\bm{x}},{\bm{y}}) denotes the 1\ell^{1}-distance between 𝒙,𝒚2{\bm{x}},{\bm{y}}\in\mathbb{Z}^{2}. When HH is an insulator at a certain energy, we can compute an invariant that characterizes the topological phase described by HH: the Hall conductance. It is the intrinsic conductance of the material originating from the quantum Hall effect, see [TKNN].

Definition 2 (Hall conductance).

Let HH be an ESR Hamiltonian with a spectral gap 𝒢{\mathcal{G}}, λ𝒢\lambda\in{\mathcal{G}} and PP be the spectral projection below energy λ\lambda: P:=𝟙(,λ)(H)P:=\mathds{1}_{(-\infty,\lambda)}(H). The Hall conductance of HH in the gap 𝒢{\mathcal{G}} is:

σb(H):=iTr(P[[P,𝟙{x2>0}],[P,𝟙{x1>0}]]).\sigma_{b}(H):=-i{\operatorname{Tr}}\left(P\big{[}[P,\mathds{1}_{\{x_{2}>0\}}],[P,\mathds{1}_{\{x_{1}>0\}}]\big{]}\right). (1.2)

The gap condition 𝒢Σ(H)c{\mathcal{G}}\subset\Sigma(H)^{c} ensures that (1.2) is well-defined; see e.g. [EGS05] (or Proposition 1 below). This work aims to prove the bulk-edge correspondence: for edge systems interpolating between two insulators, the conductance of the edge is equal to the difference between the two Hall conductances. We turn to the definition of edge systems:

Definition 3 (Edge operator).

Let H±H_{\pm} be two ESR Hamiltonians and U2U\subset\mathbb{R}^{2}. An edge Hamiltonian HeH_{e} associated to H+H_{+} and HH_{-} in UU and UcU^{c} is a selfadjoint operator on 2(2,m)\ell^{2}(\mathbb{Z}^{2},\mathbb{C}^{m}) that satisfies the kernel estimate:

𝒙,𝒚2,|E(𝒙,𝒚)|ν1e2νd1(𝒙,Ω),E:=He𝟙UH+𝟙U𝟙UcH𝟙Uc.\forall{\bm{x}},{\bm{y}}\in\mathbb{Z}^{2},\qquad\big{|}E({\bm{x}},{\bm{y}})\big{|}\leq\nu^{-1}e^{-2\nu d_{1}({\bm{x}},\partial\Omega)},\qquad E:=H_{e}-\mathds{1}_{U}H_{+}\mathds{1}_{U}-\mathds{1}_{U^{c}}H_{-}\mathds{1}_{U^{c}}. (1.3)

The condition (1.3) means that HeH_{e} describes two insulators glued along the edge U{\partial}U. To measure the conductance of HeH_{e} along U{\partial}U, we now discuss sets transverse to UU.

Definition 4 (Transversality).

We say that two sets U,V2U,V\subset\mathbb{R}^{2} are transverse if

lim inf|𝒙|+lnΨU,V(𝒙)ln|𝒙|>0,ΨU,V(𝒙)=def 1+d1(𝒙,U)+d1(𝒙,V).\liminf_{|{\bm{x}}|\to+\infty}\frac{\ln{\Psi_{U,V}}({\bm{x}})}{\ln|{\bm{x}}|}>0,\qquad{\Psi_{U,V}}({\bm{x}})\ \mathrel{\stackrel{{\scriptstyle\makebox[0.0pt]{\mbox{\tiny def}}}}{{=}}}\ 1+d_{1}({\bm{x}},{\partial}U)+d_{1}({\bm{x}},{\partial}V). (1.4)

In (1.4), d1(𝒙,U)d_{1}({\bm{x}},{\partial}U) denotes the 1\ell^{1}-distance between 𝒙2{\bm{x}}\in\mathbb{R}^{2} and the set U{\partial}U. Transversality is a geometric condition on the relative position of the boundaries U{\partial}U and V{\partial}V. It demands that these two sets get away from each other at a relatively mild rate, typically ΨU,V(𝒙)|𝒙|α{\Psi_{U,V}}({\bm{x}})\gtrsim|{\bm{x}}|^{\alpha} for some α>0\alpha>0. Under the transversality condition (1.4), we can define the conductance of HeH_{e} along U{\partial}U into the measurement set VV:

Definition 5 (Edge conductance).

Let HeH_{e} be an edge operator associated to two Hamiltonians H+,HH_{+},H_{-} with a joint spectral gap 𝒢{\mathcal{G}}, in the sets U,UcU,U^{c}. Assume that U,V2U,V\subset\mathbb{R}^{2} are transverse. The edge conductance of HeH_{e} into VV for energies in the bulk spectral gap 𝒢{\mathcal{G}} is:

σeU,V(He)=iTr(ρ(He)[He,𝟙V]),\sigma_{e}^{U,V}(H_{e})=i{\operatorname{Tr}}\big{(}\rho^{\prime}(H_{e})[H_{e},\mathds{1}_{V}]\big{)}, (1.5)

where ρC(;[0,1])\rho\in C^{\infty}(\mathbb{R};[0,1]) is a function such that:

ρ(x)={1,xsup𝒢,0,xinf𝒢.\rho(x)=\begin{cases}1,&x\geq\sup{\mathcal{G}},\\ 0,&x\leq\inf{\mathcal{G}}.\end{cases} (1.6)

In (1.5), [He,𝟙V][H_{e},\mathds{1}_{V}] measures the number of particles moving into the set VV per unit time, while ρ(He)\rho^{\prime}(H_{e}) is an energy density in 𝒢{\mathcal{G}}. As a result, σe(He)\sigma_{e}(H_{e}) captures the expected charge moving along U{\partial}U into VV per unit time and per unit energy: it is the conductance of HeH_{e} along U{\partial}U. Transversality of (U,V)(U,V) guarantees that σe(He)\sigma_{e}(H_{e}) is well defined; see Proposition 3 below. We also mention that σe(He)\sigma_{e}(H_{e}) does not depend on ρ\rho satisfying (1.6) – this follows, for instance, from Theorem 1 below. Definitions 1 - 5 serve as the basis for our setup:

Assumption 1.

In this work:

  1. (𝒜\mathcal{A}1)

    H±H_{\pm} are two selfadjoint ESR operators on 2(2,m)\ell^{2}(\mathbb{Z}^{2},\mathbb{C}^{m}) with a joint spectral gap 𝒢{\mathcal{G}}; P±=𝟙(,λ)(H±)P_{\pm}=\mathds{1}_{(-\infty,\lambda)}(H_{\pm}) are the spectral projectors below energy λ𝒢\lambda\in{\mathcal{G}}.

  2. (𝒜\mathcal{A}2)

    U,VU,V are transverse subsets of 2\mathbb{R}^{2}.

  3. (𝒜\mathcal{A}3)

    HeH_{e} is an edge Hamiltonian associated to H+,HH_{+},H_{-} in U,UcU,U^{c}.

We are now ready to state our main result:

Theorem 1 (Bulk-edge correspondence).

Under Assumption 1,

σeU,V(He)=𝒳U,V(σb(P+)σb(P)),\sigma_{e}^{U,V}(H_{e})=\mathcal{X}_{U,V}\cdot\big{(}\sigma_{b}(P_{+})-\sigma_{b}(P_{-})\big{)}, (1.7)

where 𝒳U,V\mathcal{X}_{U,V} is an integer that depends exclusively on the sets U,VU,V.

In rough terms, the integer 𝒳U,V\mathcal{X}_{U,V} emerging in the formula (1.7), which we call the intersection number, counts algebraically how many times the boundary of UU (suitably oriented) enters VV. See Figures 1 and 2 for a brief description on how to compute 𝒳U,V\mathcal{X}_{U,V} and §5.1, §LABEL:sec-7.2 for detailed definitions.

Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Figure 1. We define the intersection number 𝒳U,V\mathcal{X}_{U,V} between transverse simple sets U,VU,V in two steps. We first orient U{\partial}U such that UU is to its left according to the outward-pointing normal, see (a) and (b); then we count how many times the oriented U{\partial}U enters VV, see (c). Here 𝒳U,V=+111=1\mathcal{X}_{U,V}=+1-1-1=-1.

Theorem 1 is a version of the bulk-edge correspondence that goes beyond half-planes: it applies to sets (U,V)(U,V) with arbitrary geometry. In particular, leveraging flexibility on VV improves our knowledge on edge currents. When U{\partial}U is disconnected, our result shows that each connected component of U{\partial}U supports a current with conductance |σb(P+)σb(P)||\sigma_{b}(P_{+})-\sigma_{b}(P_{-})|, propagating according to the orientation of U{\partial}U; see Figures (2(a)) and (2(b)). Another application is when UU is a strip and VV is perpendicular to UU: Theorem 1 implies that the edge conductance vanishes since the intersection number is 0; see Figure (2(c)). Both these facts were heuristically known but rigorous proofs were missing.

1.3. Sketch of Proof

Our proof of Theorem 1 consists of two independent components.

Part 1 (§24). This part reduces the edge conductance σe(He)\sigma_{e}(H_{e}) to a bulk quantity (i.e. that depends only on H+H_{+}, HH_{-}). The key observation is that one can formally think of σe(He)\sigma_{e}(H_{e}) as a singular trace:

σeU,V(He)=Tr(iρ(He)[He,𝟙V])Tr(i[ρ(He),𝟙V]).\sigma_{e}^{U,V}(H_{e})={\operatorname{Tr}}\big{(}i\rho^{\prime}(H_{e})[H_{e},\mathds{1}_{V}]\big{)}\simeq{\operatorname{Tr}}\big{(}i[\rho(H_{e}),\mathds{1}_{V}]\big{)}. (1.8)

We call it “singular” because the operator [ρ(He),𝟙V][\rho(H_{e}),\mathds{1}_{V}] is not technically trace-class. However, if WW is a RR-tubular neighborhood of U{\partial}U then [ρ(He)𝟙W,𝟙V][\rho(H_{e})\mathds{1}_{W},\mathds{1}_{V}] is trace-class, and algebraic manipulations show that this trace vanishes. Therefore, as singular traces,

σeU,V(He)Tr([ρ(He)𝟙Wc,𝟙V]),\sigma_{e}^{U,V}(H_{e})\simeq{\operatorname{Tr}}\big{(}[\rho(H_{e})\mathds{1}_{W^{c}},\mathds{1}_{V}]\big{)}, (1.9)

which is essentially a bulk term: for R1R\gg 1, ρ(He)𝟙Wc\rho(H_{e})\mathds{1}_{W^{c}} depends essentially on the bulk Hamiltonians H+H_{+} and HH_{-}.

At the heuristic level, our proof is based on the above informal observation but its mathematical execution is more sophisticated. We follow a strategy due to Elgart–Graf–Schenker [EGS05], tailored to the more complex geometry under focus here. To avoid trace-class issues, we inject the cutoff 𝟙W\mathds{1}_{W} in the very first step: in the formula Tr(iρ(He)[He,𝟙V]){\operatorname{Tr}}(i\rho^{\prime}(H_{e})[H_{e},\mathds{1}_{V}]) rather than in the ill-defined quantity Tr([ρ(He),𝟙V]){\operatorname{Tr}}([\rho(H_{e}),\mathds{1}_{V}]). Getting to the rigorous analogue of (1.9) produces commutators which we analyze using Helffer–Sjöstrand formulas and cyclicity. Edge terms vanish as predicted, and this eventually yields

σeU,V(He)=σbU,V(P+)σbU,V(P),σbU,V(P):=iTr(P[[P,𝟙U],[P,𝟙V]]),\sigma_{e}^{U,V}(H_{e})=\sigma_{b}^{U,V}(P_{+})-\sigma_{b}^{U,V}(P_{-}),\qquad\sigma_{b}^{U,V}(P):=-i{\operatorname{Tr}}\big{(}P\big{[}[P,\mathds{1}_{U}],[P,\mathds{1}_{V}]\big{]}\big{)}, (1.10)

see Theorem 2 below. We call the emerging qauntities σbU,V(P±)\sigma_{b}^{U,V}(P_{\pm}) geometric bulk conductances.

Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Figure 2. The intersection number in subfigure (2(a)), (2(b)), (2(c)) are +1+1, 1-1 and 0; thus the edge conductance σeU,V1(He)=σb(P+)σb(P)=σeU,V2(He)\sigma_{e}^{U,V_{1}}(H_{e})=\sigma_{b}(P_{+})-\sigma_{b}(P_{-})=-\sigma_{e}^{U,V_{2}}(H_{e}), σeU,V3(He)=0\sigma_{e}^{U,V_{3}}(H_{e})=0.

Part 2 (§5LABEL:sec-7). In this part we reduce the geometric bulk conductances to integer multiples of Hall conductances (see Theorems 3 and LABEL:thm:2 below); we give an interpretation of the emerging integer as an intersection number between U{\partial}U and V{\partial}V. The key observation is that σbU,V(P)\sigma_{b}^{U,V}(P) is the trace of a commutator [A,B][A,B] where ABAB and BABA are not separately trace-class:

σbU,V(P)=iTr(P[[P,𝟙U],[P,𝟙V]])=iTr([P𝟙UP,P𝟙VP]).\sigma_{b}^{U,V}(P)=-i{\operatorname{Tr}}\big{(}P\big{[}[P,\mathds{1}_{U}],[P,\mathds{1}_{V}]\big{]}\big{)}=-i{\operatorname{Tr}}\big{(}[P\mathds{1}_{U}P,P\mathds{1}_{V}P]\big{)}. (1.11)

In particular, a compact perturbation UU^{\prime} of UU does not affect σbU,V(P)\sigma_{b}^{U,V}(P), because

σbU,V(P)σbU,V(P)=σbUΔU,V(P)=Tr([P𝟙UΔUP,P𝟙VP])=0,\sigma_{b}^{U^{\prime},V}(P)-\sigma_{b}^{U,V}(P)=\sigma_{b}^{U^{\prime}\Delta U,V}(P)={\operatorname{Tr}}\big{(}[P\mathds{1}_{U^{\prime}\Delta U}P,P\mathds{1}_{V}P]\big{)}=0, (1.12)

where the last identity comes from cyclicity, see Proposition 2 below.

When UU and VV are simple sets (i.e. their boundaries are connected), our strategy consists of using the robustness of the geometric bulk conductance to locally deform UU and VV to half-planes and recover the Hall conductance. In details, we construct sets UnU_{n}, VnV_{n}, that (up to permutation of UU and VV) resemble {x2>0}\{x_{2}>0\}, {x1>0}\{x_{1}>0\} in 𝔻2n(0)\mathbb{D}_{2n}(0) (the disk of radius nn, centered at the origin) and UU,VV outside 𝔻4n(0)\mathbb{D}_{4n}(0). Since UnU_{n}, VnV_{n} resemble UU, VV outside 𝔻4n(0)\mathbb{D}_{4n}(0), the robustness of geometric bulk conductances (Proposition 2) guarantees that

σbU,V(P)=σbUn,Vn(P)=limnσbUn,Vn(P).\sigma_{b}^{U,V}(P)=\sigma_{b}^{U_{n},V_{n}}(P)=\lim_{n\to\infty}\sigma_{b}^{U_{n},V_{n}}(P). (1.13)

On the other hand, 𝟙Un(𝒙)\mathds{1}_{U_{n}}({\bm{x}}), 𝟙Vn(𝒙)\mathds{1}_{V_{n}}({\bm{x}}) approach 𝟙{x2>0}(𝒙)\mathds{1}_{\{x_{2}>0\}}({\bm{x}}), 𝟙{x1>0}(𝒙)\mathds{1}_{\{x_{1}>0\}}({\bm{x}}) when nn\to\infty for every 𝒙{\bm{x}}. So formally,

σbU,V(P)=limnσbUn,Vn(P)=σb(P).\sigma_{b}^{U,V}(P)=\lim_{n\to\infty}\sigma_{b}^{U_{n},V_{n}}(P)=\sigma_{b}(P). (1.14)

The main challenge is to make (1.14) rigorous. We will rely on an application of the dominated convergence theorem on specifically designed deformations of (U,V)(U,V). Because (1.11) is a trace, the involved kernel is negligible outside large enough balls. Our construction of Un,VnU_{n},V_{n} preserve this negligibility uniformly, and this yield a rigorous proof of (1.14) for simple sets U,VU,V; see Proposition 5.

In §LABEL:sec-7 we will use additivity properties of the bulk conductances to extend (1.14) to arbitrary transverse sets U,VU,V, driving the emergence of the intersection number between UU and VV:

σbU,V(P)=𝒳U,Vσb(P).\sigma_{b}^{U,V}(P)=\mathcal{X}_{U,V}\cdot\sigma_{b}(P). (1.15)

See Theorem LABEL:thm:2. Theorem 1 will follow from combining (1.10) and (1.15).

The paper is organized as follows:

  • In §2, we provide basic estimates on short-range operators and transverse sets.

  • In §3, we introduce of bulk and edge conductance and detail their basic properties.

  • In §4, we reduce the edge conductance to the difference between geometric bulk conductances (Theorem 2).

  • In §5, we introduce the intersection number for simple transverse sets (transverse sets whose boundaries are connected).

  • In §6, we prove Theorem 3: for simple transverse sets, the geometric bulk conductance is equal to the intersection number times the Hall conductance.

  • In §LABEL:sec-7, we extend Theorem 3 to any pair of transverse sets. This leads to the emergence of the intersection number 𝒳U,V\mathcal{X}_{U,V} and completes the proof of Theorem 1.

1.4. Relation to existing results

The bulk-edge correspondence has been extensively studied in the past thirty years. Various perspectives have been developed around it: KK-theoretic approaches [KS02, KS04, T14, PS16, BKR, Ku17, Br18]; spectral flow methods [Hatsugai, Br18]; and tracial techniques [EG02, EGS05, GP13, GS18, GT18, ST19, B19, D21].

While many papers have focused on extending the bulk-edge correspondence to increasing degree of generality on the model – from Landau Hamiltonian to 2\mathbb{Z}_{2}-insulators, from discrete to continuous, from periodic to disordered – very few go beyond straight edges. The work [BBDKLW23] gave an explicit formula for edge states of an adiabatically modulated Dirac operator with a weakly curved interface. The result relies on semiclassical methods and gives precise results but is restricted to the adiabatic regime; see also [D22, B22, PPY22]. Our previous work [DZ23] shows the emergence of edge spectrum for topological insulators as long as both the domain and its complement contain arbitrarily large balls, but did not provide a formula for the conductance. Thiang [T20] derived similar results using KK-theory and coarse geometry.

Later in [LT22, L23], Ludewig and Thiang proved a K-theoretic version of the bulk-edge correspondence on flasque spaces. These are spaces that satisfy coarse-geometric properties; in contrast with our condition (1.4), flasqueness [R96, Definition 9.3] can be challenging to check on concrete examples beyond perturbations of half-spaces. Their K-theoretic bulk-edge correspondence result [L23, Theorem V.4] and edge-traveling interpretation [L23, Theorem VII.7] together connect the edge conductance to the geometric bulk conductance (1.10). Our approach goes beyond [LT22, L23] by connecting the geometric conductances themselves to a topological marker independent of the geometry: the Hall conductance (1.2). This drove the emergence of the intersection number 𝒳U,V\mathcal{X}_{U,V}.

1.5. Open problems

One natural question is whether our curved bulk-edge correspondence (1.7) holds for 2\mathbb{Z}_{2}-topological insulators. This would require a space-based trace formula for 2\mathbb{Z}_{2}-bulk indices, such as (1.2) for Hall insulators. At this point it is unclear whether such a formula is available or even possible; see for instance discussions in [LM19, FSSWY20].

Another interesting problem concerns the type of the edge spectrum; it is widely expected to include an absolutely continuous part. For instance, [BW22] proved that the edge spectrum is filled with absolutely continuous spectrum when the edge is straight, combining a form of the bulk-edge correspondence derived in [FSSWY20] with a result on the structure of unitary operators [MM21]. The present work provides the first step in the context of curved edges. In a follow-up project we will show that the emerging edge spectrum is absolutely continuous, extending the result of [BW22].

In [CB23], Chen and Bal showed that for a straight edge, the bulk index is equal to the difference between the number of modes reflected and transmitted. It would be very interesting to extend this result to curved edges. Yet this promises to be challenging, because the scattering techniques developed there rely on translational invariance.

Several authors [HM15, GMP17, BBDF18] have shown that the Hall conductance of interacting systems of particles is quantized (i.e. takes values in a discrete set). In our proof a geometric bulk conductance naturally emerges, see (1.10); it would be interesting to investigate whether the analogue of this quantity in interacting systems is quantized as well, and given by the integer 𝒳U,V\mathcal{X}_{U,V} times the Hall conductance.

1.6. Funding and/or competing interests

We gratefully acknowledge support from the National Science Foundation DMS 2054589 (AD) and the Pacific Institute for the Mathematical Sciences (XZ). The contents of this work are solely the responsibility of the authors and do not necessarily represent the official views of PIMS.

1.7. Notations

We use the following notations:

  • Regular font xx, xjx_{j}, zz denotes scalars in \mathbb{R}, \mathbb{Z}, or \mathbb{C}.

  • Bold font 𝒙=(x1,x2),𝒚=(y1,y2){\bm{x}}=(x_{1},x_{2}),{\bm{y}}=(y_{1},y_{2}) denotes points in 2\mathbb{Z}^{2}.

  • d(,)d(\cdot,\cdot) denote the Euclidean metric.

  • d1(𝒙,𝒚):=|x1y1|+|x2y2|d_{1}({\bm{x}},{\bm{y}}):=|x_{1}-y_{1}|+|x_{2}-y_{2}| denote the 1\ell^{1}-metric on 2\mathbb{R}^{2}.

  • dl(𝒙,𝒚)=ln(1+d1(𝒙,𝒚))d_{l}({\bm{x}},{\bm{y}})=\ln(1+d_{1}({\bm{x}},{\bm{y}})) denote the logarithm metric on 2\mathbb{R}^{2}.

  • \|\cdot\| and 1\|\cdot\|_{1} denote the operator norm and the trace-class norm, respectively.

  • Throughout the paper, CC denotes a constant that is allowed to vary from line to line, that may depend on ν\nu and ρ\rho (though this dependence is not emphasized). We will occasionally use CNC_{N} to emphasize dependence on an integer NN.

  • For A2A\subset\mathbb{R}^{2}, A𝒞A^{\mathcal{C}} denotes the open set int(Ac)\operatorname{int}(A^{c}).

  • 𝔻R(𝒛):={𝒙2:d1(𝒙,𝒛)<R}\mathbb{D}_{R}({\bm{z}}):=\{{\bm{x}}\in\mathbb{R}^{2}:d_{1}({\bm{x}},{\bm{z}})<R\} denotes the 1\ell^{1}-disk centered at zz, of radius RR.

  • 1={x1>0}{\mathbb{H}}_{1}=\{x_{1}>0\} and 2={x2>0}{\mathbb{H}}_{2}=\{x_{2}>0\}.

  • If KK is an operator on 2(2)\ell^{2}(\mathbb{Z}^{2}), K(𝒙,𝒚)K({\bm{x}},{\bm{y}}) denotes the Schwartz kernel of KK.

2. Preparation

In this section, we give some basic operator estimates that will be needed but leave the proof to Appendix LABEL:app-ESR.

2.1. Operator estimates

Under short-range assumptions, we give here estimates on resolvents and product of operators. We start by introducing the notion of textitpolynomially-short-range (PSR) Hamiltonians.

Definition 6 (PSR).

An operator HH on 2(2,m)\ell^{2}(\mathbb{Z}^{2},\mathbb{C}^{m}) is polynomially-short-range (PSR) if for any N>0N>0, there is CN>0C_{N}>0 such that

|H(𝒙,𝒚)|CN(1+d1(𝒙,𝒚))N=CNeNdl(𝒙,𝒚),𝒙,𝒚2.|H({\bm{x}},{\bm{y}})|\leq C_{N}(1+d_{1}({\bm{x}},{\bm{y}}))^{-N}=C_{N}e^{-Nd_{l}({\bm{x}},{\bm{y}})},\qquad\forall{\bm{x}},{\bm{y}}\in\mathbb{Z}^{2}.

Note that any ESR operator is also PSR. The next result shows that (1) resolvents of ESR for insulating energies are ESR; (2) functionals of ESR operators are PSR.

Lemma 2.1.

Assume HH is self-adjoint and ESR.

  1. (a)

    For 0<|Imz|<10<|\operatorname{Im}z|<1, (Hz)1(H-z)^{-1} is ESR. More precisely,

    |(Hz)1(𝒙,𝒚)|2|Imz|ec|Imz|d1(𝒙,𝒚),c=ν432.\left|(H-z)^{-1}({\bm{x}},{\bm{y}})\right|\leq\frac{2}{|\operatorname{Im}z|}e^{-c|\operatorname{Im}z|d_{1}({\bm{x}},{\bm{y}})},\qquad c=\dfrac{\nu^{4}}{32}. (2.1)
  2. (b)

    For any gCc()g\in C_{c}^{\infty}(\mathbb{R}), g(H)g(H) is PSR.

We now give estimates on products of ESR / PSR operators. We state them in a unifying way using the two metrics d1,dd_{1},d_{\ell}.

Lemma 2.2.

Let U2U\subset\mathbb{R}^{2}. Let SS be a selfadjoint operator on 2(2,m)\ell^{2}(\mathbb{Z}^{2},\mathbb{C}^{m}) that satisfies the kernel estimate

|S(𝒙,𝒚)|Cecd(𝒙,𝒚).|S({\bm{x}},{\bm{y}})|\leq Ce^{-cd_{*}({\bm{x}},{\bm{y}})}.

Then

𝟙US𝟙U(𝒙,𝒚)|Cecd(𝒙,𝒚)cd(𝒙,U)cd(𝒚,U),\displaystyle\|\mathds{1}_{U}S\mathds{1}_{U}({\bm{x}},{\bm{y}})|\leq Ce^{-cd_{*}({\bm{x}},{\bm{y}})-cd_{*}({\bm{x}},U)-cd_{*}({\bm{y}},U)}, (2.2)
|𝟙US𝟙Uc(𝒙,𝒚)|Cec3d(𝒙,𝒚)c3d(𝒙,U)c3d(𝒚,U),\displaystyle|\mathds{1}_{U}S\mathds{1}_{U^{c}}({\bm{x}},{\bm{y}})|\leq Ce^{-\frac{c}{3}d_{*}({\bm{x}},{\bm{y}})-\frac{c}{3}d_{*}({\bm{x}},{\partial}U)-\frac{c}{3}d_{*}({\bm{y}},{\partial}U)}, (2.3)
|[𝟙U,S](𝒙,𝒚)|2Cec3d(𝒙,𝒚)c3d(𝒙,U)c3d(𝒚,U).\displaystyle|[\mathds{1}_{U},S]({\bm{x}},{\bm{y}})|\leq 2Ce^{-\frac{c}{3}d_{*}({\bm{x}},{\bm{y}})-\frac{c}{3}d_{*}({\bm{x}},{\partial}U)-\frac{c}{3}d_{*}({\bm{y}},{\partial}U)}. (2.4)
Lemma 2.3.

Let U,V2U,V\subset\mathbb{R}^{2}. Assume that S1,,SnS_{1},\dots,S_{n} are self-adjoint operators on 2(2,m)\ell^{2}(\mathbb{Z}^{2},\mathbb{C}^{m}) that satisfy the following estimates: {equations} j [1,n],   —S_j(x, y)—C_j e^-cd_*(x, y);         
p [1,n],   —S_p(x, y)—C_p e^-cd_*(x, y) - cd_*(x, U ) - cd_*(y, U );
q [1,n],   —S_q(x, y)—C_qe^-cd_*(x, y) - cd_*(x, V ) - cd_*(y, V). Then we have

|(j=1nSj)(𝒙,𝒚)|(j=1nCj)Md,c4n1ec4d(𝒙,𝒚)c4d(𝒙,U)c4d(𝒚,U)c4d(𝒙,V)c4d(𝒚,V),\left|\left(\prod_{j=1}^{n}S_{j}\right)({\bm{x}},{\bm{y}})\right|\leq\left(\prod_{j=1}^{n}C_{j}\right)M_{d_{*},\frac{c}{4}}^{n-1}e^{-\frac{c}{4}d_{*}({\bm{x}},{\bm{y}})-\frac{c}{4}d_{*}({\bm{x}},{\partial}U)-\frac{c}{4}d_{*}({\bm{y}},{\partial}U)-\frac{c}{4}d_{*}({\bm{x}},{\partial}V)-\frac{c}{4}d_{*}({\bm{y}},{\partial}V)}, (2.5)

where Md,a:=𝐱2ead(𝐱,𝐲)M_{d_{*},a}:=\sum\limits_{{\bm{x}}\in\mathbb{Z}^{2}}e^{-ad_{*}({\bm{x}},{\bm{y}})}, {1,l}*\in\{1,l\}.

Note that Md,aM_{d_{*},a} is independent of 𝒚2{\bm{y}}\in\mathbb{Z}^{2} and

Md1,a16a2,a>0, and Mdl,a=𝒙2(1+|𝒙|1)a<+,a>2.\begin{split}&M_{d_{1},a}\leq\frac{16}{a^{2}},\quad\forall a>0,\quad\text{~{}and~{}}\quad M_{d_{l},a}=\sum\limits_{{\bm{x}}\in\mathbb{Z}^{2}}\big{(}1+|{\bm{x}}|_{1}\big{)}^{-a}<+\infty,\quad\forall a>2.\end{split} (2.6)

2.2. Transversality and trace-class

Recall that we say U,V2U,V\subset\mathbb{R}^{2} if

lim inf|𝒙|+lnΨU,V(𝒙)ln|𝒙|>0,ΨU,V(𝒙)=def 1+d1(𝒙,U)+d1(𝒙,V).\liminf_{|{\bm{x}}|\to+\infty}\frac{\ln{\Psi_{U,V}}({\bm{x}})}{\ln|{\bm{x}}|}>0,\qquad{\Psi_{U,V}}({\bm{x}})\ \mathrel{\stackrel{{\scriptstyle\makebox[0.0pt]{\mbox{\tiny def}}}}{{=}}}\ 1+d_{1}({\bm{x}},{\partial}U)+d_{1}({\bm{x}},{\partial}V). (2.7)

Note that it is equivalent to

c(0,1) such that |𝒙|c1,ΨU,V(𝒙)|𝒙|c.\exists c\in(0,1)\text{ such that }\forall|{\bm{x}}|\geq c^{-1},{\Psi_{U,V}}({\bm{x}})\geq|{\bm{x}}|^{c}. (2.8)

Meanwhile, because of the inequality

2ln(1+a+b)ln(1+a)+ln(1+b)ln(1+a+b),a,b>0,2\ln(1+a+b)\geq\ln(1+a)+\ln(1+b)\geq\ln(1+a+b),\qquad a,b>0,

we have:

2lnΨU,V(𝒙)dl(𝒙,U)+dl(𝒙,V)lnΨU,V(𝒙).2\ln\Psi_{U,V}({\bm{x}})\geq d_{l}({\bm{x}},{\partial}U)+d_{l}({\bm{x}},{\partial}V)\geq\ln\Psi_{U,V}({\bm{x}}). (2.9)

Therefore for any N>2cN>\frac{2}{c}, by (2.8)

𝒙2eN(dl(𝒙,U)+dl(𝒙,V))𝒙2ΨU,V(𝒙)N<+.\sum_{{\bm{x}}\in\mathbb{Z}^{2}}e^{-N(d_{l}({\bm{x}},{\partial}U)+d_{l}({\bm{x}},{\partial}V))}\leq\sum_{{\bm{x}}\in\mathbb{Z}^{2}}\Psi_{U,V}({\bm{x}})^{-N}<+\infty. (2.10)

The next result states that product of PSR operators that decay away from the boundaries of transverse sets U,VU,V are trace-class.

Corollary 2.4 (trace-class).

Assume U,V2U,V\subset\mathbb{R}^{2} are transverse sets. Assume SjS_{j}, j=1,,nj=1,\cdots,n, are PSR and for some p,q[1,n]p,q\in[1,n] and for any NN, there is CNC_{N} such that

|Sp(𝒙,𝒚)|CNeNdl(𝒙,U)Ndl(𝒚,U),|Sq(𝒙,𝒚)|CNeNdl(𝒙,V)Ndl(𝒚,V).|S_{p}({\bm{x}},{\bm{y}})|\leq C_{N}e^{-Nd_{l}({\bm{x}},{\partial}U)-Nd_{l}({\bm{y}},{\partial}U)},\ \ |S_{q}({\bm{x}},{\bm{y}})|\leq C_{N}e^{-Nd_{l}({\bm{x}},{\partial}V)-Nd_{l}({\bm{y}},{\partial}V)}.

Then j=1nSj\prod\limits_{j=1}^{n}S_{j} is trace-class.

3. Bulk and edge conductance

From now on, we always assume that Assumption 1 holds.

In this section, we show that the Hall conductance (1.2), the geometric bulk conductance (1.10), and the edge conductance (1.5), are well-defined.

3.1. Geometric bulk conductances

Proposition 1.

Under Assumption 1, the geometric bulk conductances

σbU,V(P±):=iTr(P±[[P±,𝟙U],[P±,𝟙V]])\sigma_{b}^{U,V}(P_{\pm}):=-i{\operatorname{Tr}}\big{(}P_{\pm}\big{[}[P_{\pm},\mathds{1}_{U}],[P_{\pm},\mathds{1}_{V}]\big{]}\big{)} (3.1)

are well-defined and independent of λ𝒢\lambda\in{\mathcal{G}}.

Because the right and upper half-planes are transverse sets, the Hall conductance σb(H):=iTr(P[[P,𝟙{x2>0}],[P,𝟙{x1>0}]])\sigma_{b}(H):=-i{\operatorname{Tr}}\big{(}P\big{[}[P,\mathds{1}_{\{x_{2}>0\}}],[P,\mathds{1}_{\{x_{1}>0\}}]\big{]}\big{)} is well-defined.

Proof.

Since 𝒢{\mathcal{G}} is a spectral gap, for any λ,λ𝒢\lambda,\lambda^{\prime}\in{\mathcal{G}}, the spectral projectors 𝟙(,λ)(H)\mathds{1}_{(-\infty,\lambda)}(H) and 𝟙(,λ)(H)\mathds{1}_{(-\infty,\lambda^{\prime})}(H) are equal. This proves independence.

We then show that σbU,V(P)\sigma_{b}^{U,V}(P) is well-defined. Without loss of generalities we may assume that 𝒢{\mathcal{G}} is an interval; pick λ\lambda as the midpoint and fCc()f\in C_{c}^{\infty}(\mathbb{R}) such that f(H)=Pf(H)=P. By Lemma 2.1(a), P=f(H)P=f(H) is PSR (one can actually prove that it is ESR). By (2.4), for any NN, there exists CN>0C_{N}>0 such that

|P(𝒙,𝒚)|CNe4Ndl(𝒙,𝒚),|[P,𝟙U](𝒙,𝒚)|CNe4N(dl(𝒙,𝒚)+dl(𝒙,U)+dl(𝒚,U)),|[P,𝟙V](𝒙,𝒚)|CNe4N(dl(𝒙,𝒚)+dl(𝒙,V)+dl(𝒚,V)).\begin{split}&|P({\bm{x}},{\bm{y}})|\leq C_{N}e^{-4Nd_{l}({\bm{x}},{\bm{y}})},\\ &\big{|}[P,\mathds{1}_{U}]({\bm{x}},{\bm{y}})\big{|}\leq C_{N}e^{-4N(d_{l}({\bm{x}},{\bm{y}})+d_{l}({\bm{x}},{\partial}U)+d_{l}({\bm{y}},{\partial}U))},\\ &\big{|}[P,\mathds{1}_{V}]({\bm{x}},{\bm{y}})\big{|}\leq C_{N}e^{-4N(d_{l}({\bm{x}},{\bm{y}})+d_{l}({\bm{x}},{\partial}V)+d_{l}({\bm{y}},{\partial}V))}.\end{split} (3.2)

Introduce

KU,V=defP[[P,𝟙U],[P,𝟙V]].{K_{U,V}}\ \mathrel{\stackrel{{\scriptstyle\makebox[0.0pt]{\mbox{\tiny def}}}}{{=}}}\ P[[P,\mathds{1}_{U}],[P,\mathds{1}_{V}]]. (3.3)

By (2.5) in Lemma 2.3,

|KU,V(𝒙,𝒚)|CNeN(dl(𝒙,𝒚)+dl(𝒙,U)+dl(𝒙,V)+dl(𝒚,U)+dl(𝒚,V)).|{K_{U,V}}({\bm{x}},{\bm{y}})|\leq C_{N}e^{-N(d_{l}({\bm{x}},{\bm{y}})+d_{l}({\bm{x}},{\partial}U)+d_{l}({\bm{x}},{\partial}V)+d_{l}({\bm{y}},{\partial}U)+d_{l}({\bm{y}},{\partial}V))}. (3.4)

Since U,VU,V are transverse subsets of 2\mathbb{R}^{2}, by Corollary 2.4, KU,V=P[[P,𝟙U][P,𝟙V]]K_{U,V}=P\big{[}[P,\mathds{1}_{U}][P,\mathds{1}_{V}]\big{]} is trace-class; thus σbU,V(P)\sigma_{b}^{U,V}(P) is well-defined. This completes the proof. ∎

In particular, we have by (2.10) and (3.4)

|σbU,V(P)|=𝒙2|KU,V(𝒙,𝒙)|CN(𝒙2ΨU,V(𝒙)N)2.|{\sigma_{b}^{U,V}}(P)|=\sum\limits_{{\bm{x}}\in\mathbb{Z}^{2}}|K_{U,V}({\bm{x}},{\bm{x}})|\leq C_{N}\left(\sum_{{\bm{x}}\in\mathbb{Z}^{2}}\Psi_{U,V}({\bm{x}})^{-N}\right)^{2}. (3.5)

Given two sets A1,A2A_{1},A_{2}, we introduce the symmetric difference

A1ΔA2=(A1A2)(A2A1).A_{1}\Delta A_{2}=(A_{1}\setminus A_{2})\cup(A_{2}\setminus A_{1}). (3.6)

One of the key properties of geometric bulk conductances is:

Proposition 2.

[Robustness of geometric bulk conductances] Under Assumption 1, assume UU^{\prime} is such that for some R>0R>0,

UΔU{𝒙2:dl(𝒙,U)R}.U\Delta U^{\prime}\subset\{{\bm{x}}\in\mathbb{R}^{2}:d_{l}({\bm{x}},{\partial}U)\leq R\}. (3.7)

Then σbU,V(P±)=σbU,V(P±)\sigma_{b}^{U,V}(P_{\pm})=\sigma_{b}^{U^{\prime},V}(P_{\pm}).

Proof of Proposition 2.

First, U,VU^{\prime},V are transverse and σbU,V(P±)\sigma_{b}^{U^{\prime},V}(P_{\pm}) are well-defined. Indeed, by (2.9) and (3.7),

2lnΨU,V(𝒙)dl(𝒙,U)+dl(𝒙,V)dl(𝒙,U)+dl(𝒙,V)dl(U,U)lnΨU,V(𝒙)Rcln|𝒙|Rcln|𝒙|\begin{split}2\ln\Psi_{U^{\prime},V}({\bm{x}})&\geq d_{l}({\bm{x}},{\partial}U^{\prime})+d_{l}({\bm{x}},{\partial}V)\geq d_{l}({\bm{x}},{\partial}U)+d_{l}({\bm{x}},{\partial}V)-d_{l}({\partial}U,{\partial}U^{\prime})\\ &\geq\ln\Psi_{U,V}({\bm{x}})-R\geq c\ln|{\bm{x}}|-R\geq c^{\prime}\ln|{\bm{x}}|\end{split}

for any c>cc^{\prime}>c and large enough |𝒙||{\bm{x}}|. Hence U,VU^{\prime},V are transverse sets and σbU,V(P±)\sigma_{b}^{U^{\prime},V}(P_{\pm}) is well-defined. It remains to show:

σbD,V(P)=0whenD=UΔU,P=P±.\sigma_{b}^{D,V}(P)=0\quad\text{when}\quad D=U\Delta U^{\prime},\qquad P=P_{\pm}.

By direct expansion of definition (3.1) and 1=P+P1=P+P^{\perp}, we get

σbD,V(P)=Tr(P𝟙DP𝟙VPP𝟙VP𝟙DP)=Tr(P𝟙VP𝟙DPP𝟙DP𝟙VP).\sigma_{b}^{D,V}(P)={\operatorname{Tr}}\big{(}P\mathds{1}_{D}P\mathds{1}_{V}P-P\mathds{1}_{V}P\mathds{1}_{D}P\big{)}={\operatorname{Tr}}\big{(}P\mathds{1}_{V}P^{\perp}\mathds{1}_{D}P-P\mathds{1}_{D}P^{\perp}\mathds{1}_{V}P\big{)}.

Recall the cyclicity of trace-class operators – see e.g. [K89, Proposition 7.3]:

Tr(AB)=Tr(BA), when AB and BA are trace class.{\operatorname{Tr}}(AB)={\operatorname{Tr}}(BA),\text{~{}when~{}}AB\text{~{}and~{}}BA\text{~{}are~{}trace~{}class.} (3.8)

Let us first assume that every operator below is trace-class so we can freely use cyclicity above. We get

σbD,V(P)=Tr(P𝟙VP𝟙DP𝟙VP𝟙D)=Tr([P,𝟙V]𝟙D)=Tr(𝟙VcP𝟙V𝟙D1VP𝟙Vc𝟙D)=Tr(P𝟙V𝟙D𝟙VcP𝟙Vc𝟙D1V)=0,\begin{split}\sigma_{b}^{D,V}(P)&={\operatorname{Tr}}\big{(}P\mathds{1}_{V}P^{\perp}\mathds{1}_{D}-P^{\perp}\mathds{1}_{V}P\mathds{1}_{D}\big{)}={\operatorname{Tr}}\big{(}[P,\mathds{1}_{V}]\mathds{1}_{D}\big{)}\\ &={\operatorname{Tr}}\big{(}\mathds{1}_{V^{c}}P\mathds{1}_{V}\mathds{1}_{D}-1_{V}P\mathds{1}_{V^{c}}\mathds{1}_{D}\big{)}\\ &={\operatorname{Tr}}\big{(}P\mathds{1}_{V}\mathds{1}_{D}\mathds{1}_{V^{c}}-P\mathds{1}_{V^{c}}\mathds{1}_{D}1_{V}\big{)}=0,\end{split}

where we used [P,𝟙V]=𝟙VcP𝟙V𝟙VP𝟙Vc[P,\mathds{1}_{V}]=\mathds{1}_{V^{c}}P\mathds{1}_{V}-\mathds{1}_{V}P\mathds{1}_{V^{c}}.

Now we prove every operator mentioned above is trace-class. Note that P𝟙VP=P[1V,P]P^{\perp}\mathds{1}_{V}P=P^{\perp}[1_{V},P], P𝟙VP=[P,𝟙V]PP\mathds{1}_{V}P^{\perp}=[P,\mathds{1}_{V}]P^{\perp} and recall that PP is PSR (see the proof of Proposition 1). By (2.4),

|P𝟙VP(𝒙,𝒚)|,|P𝟙VP(𝒙,𝒚)|CNe6Ndl(𝒙,𝒚)6Ndl(𝒙,V)6Ndl(𝒚,V).\big{|}P\mathds{1}_{V}P^{\perp}({\bm{x}},{\bm{y}})\big{|},\big{|}P^{\perp}\mathds{1}_{V}P({\bm{x}},{\bm{y}})\big{|}\leq C_{N}e^{-6Nd_{l}({\bm{x}},{\bm{y}})-6Nd_{l}({\bm{x}},{\partial}V)-6Nd_{l}({\bm{y}},{\partial}V)}.

Meanwhile, 𝟙D(𝒙)e6Ndl(𝒙,D)e6Nln(1+R)e6Ndl(𝒙,U)\mathds{1}_{D}({\bm{x}})\leq e^{-6Nd_{l}({\bm{x}},D)}\leq e^{6N\ln(1+R)}e^{-6Nd_{l}({\bm{x}},{\partial}U)}. Hence

|𝟙DP(𝒙,𝒚)|,|P𝟙D(𝒙,𝒚)|CN(1+R)6Ne3Ndl(𝒙,𝒚)3Ndl(𝒙,U)3Ndl(𝒚,U).\big{|}\mathds{1}_{D}P({\bm{x}},{\bm{y}})\big{|},\big{|}P\mathds{1}_{D}({\bm{x}},{\bm{y}})\big{|}\leq C_{N}(1+R)^{6N}e^{-3Nd_{l}({\bm{x}},{\bm{y}})-3Nd_{l}({\bm{x}},{\partial}U)-3Nd_{l}({\bm{y}},{\partial}U)}.

Therefore by Corollary 2.4, P𝟙VP𝟙DP\mathds{1}_{V}P^{\perp}\mathds{1}_{D} and P𝟙VP𝟙DP^{\perp}\mathds{1}_{V}P\mathds{1}_{D} are both trace-class. Meanwhile, by (2.3) of Lemma 2.2 and 𝟙D(𝒙)(1+R)6Ne6Ndl(𝒙,U)\mathds{1}_{D}({\bm{x}})\leq(1+R)^{6N}e^{-6Nd_{l}({\bm{x}},{\partial}U)}, hence

|𝟙VcP𝟙V𝟙D(𝒙,𝒚)|C6N(1+R)6Ne2Ndl(𝒙,𝒚)2Ndl(𝒙,V)2Ndl(𝒚,V)6Ndl(𝒙,U).\big{|}\mathds{1}_{V^{c}}P\mathds{1}_{V}\mathds{1}_{D}({\bm{x}},{\bm{y}})\big{|}\leq C_{6N}(1+R)^{6N}e^{-2Nd_{l}({\bm{x}},{\bm{y}})-2Nd_{l}({\bm{x}},{\partial}V)-2Nd_{l}({\bm{y}},{\partial}V)-6Nd_{l}({\bm{x}},{\partial}U)}.

By (2.10),

𝒙,𝒚|𝟙VcP𝟙V𝟙D(𝒙,𝒚)|CN(1+R)6N𝒙e2Ndl(𝒙,V)2Ndl(𝒙,U)(𝒚e2Ndl(𝒙,𝒚))<+.\sum\limits_{{\bm{x}},{\bm{y}}}\big{|}\mathds{1}_{V^{c}}P\mathds{1}_{V}\mathds{1}_{D}({\bm{x}},{\bm{y}})\big{|}\leq C_{N}(1+R)^{6N}\sum\limits_{{\bm{x}}}e^{-2Nd_{l}({\bm{x}},{\partial}V)-2Nd_{l}({\bm{x}},{\partial}U)}\left(\sum\limits_{{\bm{y}}}e^{-2Nd_{l}({\bm{x}},{\bm{y}})}\right)<+\infty.

Hence 𝟙VcP𝟙V𝟙D\mathds{1}_{V^{c}}P\mathds{1}_{V}\mathds{1}_{D}, and similarly 1VP𝟙Vc𝟙D1_{V}P\mathds{1}_{V^{c}}\mathds{1}_{D}, are also trace-class. This completes the proof. ∎

3.2. Edge conductance

Now we can introduce the edge conductance.

Proposition 3 (Edge conductance).

Under Assumption 1, the edge conductance into VV

σeU,V(He)=iTr(ρ(He)[He,𝟙V])\sigma_{e}^{U,V}(H_{e})=i{\operatorname{Tr}}\big{(}\rho^{\prime}(H_{e})[H_{e},\mathds{1}_{V}]\big{)} (3.9)

is well-defined.

For the rest of the paper, we consider an arbitrary but fixed pair of transverse sets (U,V)(U,V) in 2\mathbb{R}^{2} and omit the superscripts U,VU,V from σeU,V(He)\sigma_{e}^{U,V}(H_{e}) for convenience.

Proof of Proposition 3.

1. By definition, supp(ρ)𝒢Σ(H+)cΣ(H)c\mathrm{supp}(\rho^{\prime})\subset{\mathcal{G}}\subset\Sigma(H_{+})^{c}\cap\Sigma(H_{-})^{c}; hence ρ(H±)=0\rho^{\prime}(H_{\pm})=0. Therefore, we have:

ρ(He)\displaystyle\rho^{\prime}(H_{e}) =ρ(He)ρ(H+)𝟙Uρ(H)𝟙Uc\displaystyle=\rho^{\prime}(H_{e})-\rho^{\prime}(H_{+})\mathds{1}_{U}-\rho^{\prime}(H_{-})\mathds{1}_{U^{c}} (3.10)
=(ρ(He)ρ(H+))𝟙U+(ρ(He)ρ(H))𝟙Uc.\displaystyle=\big{(}\rho^{\prime}(H_{e})-\rho^{\prime}(H_{+})\big{)}\mathds{1}_{U}+\big{(}\rho^{\prime}(H_{e})-\rho^{\prime}(H_{-})\big{)}\mathds{1}_{U^{c}}. (3.11)

2. We now estimate ρ(He)ρ(H+)\rho^{\prime}(H_{e})-\rho^{\prime}(H_{+}). By Helffer-Sjöstrand formula (LABEL:eq:_HS_1), we have

ρ(He)ρ(H+)=12πiρ~z¯((Hez)1(H+z)1)𝑑zdz¯=12πiρ~z¯(Hez)1(HeH+)(H+z)1𝑑zdz¯.\begin{split}\rho^{\prime}(H_{e})-\rho^{\prime}(H_{+})&=\frac{1}{2\pi i}\int_{\mathbb{C}}\frac{{\partial}\tilde{\rho}^{\prime}}{{\partial}\bar{z}}\left((H_{e}-z)^{-1}-(H_{+}-z)^{-1}\right)dz\wedge d\bar{z}\\ &=-\frac{1}{2\pi i}\int_{\mathbb{C}}\frac{{\partial}\tilde{\rho}^{\prime}}{{\partial}\bar{z}}(H_{e}-z)^{-1}(H_{e}-H_{+})(H_{+}-z)^{-1}dz\wedge d\bar{z}.\end{split} (3.12)

Note that

HeH+=E+𝟙UH+𝟙U+𝟙UcH𝟙UcH+=E(𝟙UcH+𝟙U+𝟙UH+𝟙Uc)+𝟙Uc(HH+)𝟙Uc:=EF+G.\begin{split}H_{e}-H_{+}&=E+\mathds{1}_{U}H_{+}\mathds{1}_{U}+\mathds{1}_{U^{c}}H_{-}\mathds{1}_{U^{c}}-H_{+}\\ &=E-\big{(}\mathds{1}_{U^{c}}H_{+}\mathds{1}_{U}+\mathds{1}_{U}H_{+}\mathds{1}_{U^{c}}\big{)}+\mathds{1}_{U^{c}}(H_{-}-H_{+})\mathds{1}_{U^{c}}:=E-F+G.\end{split}

And we have the following estimates:

  1. (1)

    By (2.1) in Lemma 2.1, when |Imz|<1|\operatorname{Im}z|<1, we have:

    |(H+z)1(𝒙,𝒚)|,|(Hez)1(𝒙,𝒚)|2|Imz|1eν4|Imz|32d1(𝒙,𝒚).\begin{split}\left|(H_{+}-z)^{-1}({\bm{x}},{\bm{y}})\right|,\ \left|(H_{e}-z)^{-1}({\bm{x}},{\bm{y}})\right|\leq 2|\operatorname{Im}z|^{-1}e^{-\frac{\nu^{4}|\operatorname{Im}z|}{32}d_{1}({\bm{x}},{\bm{y}})}.\end{split} (3.13)
  2. (2)

    By Lemma 2.2, {equations} —G(x, y)— ≤2ν^-1e^-2ν(d_1(x, y) + d_1(x, U^c) + d_1(y, U^c)).
    —F(x, y)—≤2ν^-1e^-2ν3(d_1(x, y) + d_1(x, ∂U) + d_1(y, ∂U)).

  3. (3)

    By short-range of H±H_{\pm}, HeH_{e} and definition of EE in (1.3), EE is ESR and decays away from U{\partial}U. By interpolating the two associated bounds, we get:

    |E(𝒙,𝒚)|2ν1eν2(d1(𝒙,𝒚)+d1(𝒙,U)+d1(𝒚,U)).\begin{split}|E({\bm{x}},{\bm{y}})|\leq 2\nu^{-1}e^{-\frac{\nu}{2}\big{(}d_{1}({\bm{x}},{\bm{y}})+d_{1}({\bm{x}},{\partial}U)+d_{1}({\bm{y}},{\partial}U)\big{)}}.\end{split}

Note that all the decay rate in the above bounds can be relaxed to ν4|Imz|32\frac{\nu^{4}|\operatorname{Im}z|}{32} because ν<1\nu<1 and |Imz|<1|\operatorname{Im}z|<1. Denote r:=ν432×4r:=\frac{\nu^{4}}{32\times 4}. Since d1(𝒙,U)d1(𝒙,Uc)d_{1}({\bm{x}},{\partial}U)\geq d_{1}({\bm{x}},U^{c}), we have

|(HeH+)(𝒙,𝒚)|\displaystyle\big{|}(H_{e}-H_{+})({\bm{x}},{\bm{y}})\big{|} =|(E+F+G)(𝒙,𝒚)|6ν1eν2(d1(𝒙,𝒚)+d1(𝒙,Uc)+d1(𝒚,Uc))\displaystyle=\big{|}(E+F+G)({\bm{x}},{\bm{y}})\big{|}\leq 6\nu^{-1}e^{-\frac{\nu}{2}\big{(}d_{1}({\bm{x}},{\bm{y}})+d_{1}({\bm{x}},U^{c})+d_{1}({\bm{y}},U^{c})\big{)}} (3.14)
6ν1er|Imz|(d1(𝒙,𝒚)+d1(𝒙,Uc)+d1(𝒚,Uc))\displaystyle\leq 6\nu^{-1}e^{-r|\operatorname{Im}z|\big{(}d_{1}({\bm{x}},{\bm{y}})+d_{1}({\bm{x}},U^{c})+d_{1}({\bm{y}},U^{c})\big{)}} (3.15)

By (2.5) in Lemma 2.3 and (2.6),

|(Hez)1(HeH+)(H+z)1(𝒙,𝒚)|CMd1,r|Imz|2|Imz|2er|Imz|(d1(𝒙,𝒚)+d1(𝒙,Uc)+d1(𝒚,Uc))C|Imz|6er|Imz|(d1(𝒙,𝒚)+d1(𝒙,Uc)+d1(𝒚,Uc)).\begin{split}\big{|}(H_{e}-z)^{-1}(H_{e}-H_{+})(H_{+}-z)^{-1}({\bm{x}},{\bm{y}})\big{|}&\leq\frac{CM_{d_{1},r|\operatorname{Im}z|}^{2}}{|\operatorname{Im}z|^{2}}e^{-r|\operatorname{Im}z|\big{(}d_{1}({\bm{x}},{\bm{y}})+d_{1}({\bm{x}},U^{c})+d_{1}({\bm{y}},U^{c})\big{)}}\\ &\leq\frac{C}{|\operatorname{Im}z|^{6}}e^{-r|\operatorname{Im}z|\big{(}d_{1}({\bm{x}},{\bm{y}})+d_{1}({\bm{x}},U^{c})+d_{1}({\bm{y}},U^{c})\big{)}}.\end{split}

As a result, using almost analyticity (LABEL:eq:_HS_1), when 𝒙𝒚{\bm{x}}\neq{\bm{y}}, let w=|Imz|w=|\operatorname{Im}z|,

|(ρ(He)ρ(H+))(𝒙,𝒚)|\displaystyle\big{|}\big{(}\rho^{\prime}(H_{e})-\rho^{\prime}(H_{+})\big{)}({\bm{x}},{\bm{y}})| supp(g)×[1,1]C|z¯g~||Imz|6er|Imz|(d1(𝒙,𝒚)+d1(𝒙,Uc)+d1(𝒚,Uc))|dzdz¯|\displaystyle\leq\int_{\mathrm{supp}(g)\times[-1,1]}\frac{C|{\partial}_{\bar{z}}\tilde{g}|}{|\operatorname{Im}z|^{6}}e^{-r|\operatorname{Im}z|\big{(}d_{1}({\bm{x}},{\bm{y}})+d_{1}({\bm{x}},U^{c})+d_{1}({\bm{y}},U^{c})\big{)}}|dz\wedge d\bar{z}| (3.16)
CN0wN6ewr(d1(𝒙,𝒚)+d1(𝒙,Uc)+d1(𝒚,Uc))𝑑w\displaystyle\leq C_{N}\int_{0}^{\infty}w^{N-6}e^{-wr\big{(}d_{1}({\bm{x}},{\bm{y}})+d_{1}({\bm{x}},U^{c})+d_{1}({\bm{y}},U^{c})\big{)}}dw (3.17)
CN(d1(𝒙,𝒚)+d1(𝒙,Uc)+d1(𝒚,Uc))N5.\displaystyle\leq C_{N}\big{(}d_{1}({\bm{x}},{\bm{y}})+d_{1}({\bm{x}},U^{c})+d_{1}({\bm{y}},U^{c})\big{)}^{N-5}. (3.18)

When 𝒙=𝒚{\bm{x}}={\bm{y}}, |(ρ(He)ρ(H+))(𝒙,𝒚)|ρ(He)ρ(H+)2g<+\big{|}(\rho^{\prime}(H_{e})-\rho^{\prime}(H_{+}))({\bm{x}},{\bm{y}})\big{|}\leq\|\rho^{\prime}(H_{e})-\rho^{\prime}(H_{+})\|\leq 2\|g\|_{\infty}<+\infty. Therefore, at the cost of potentially increasing CNC_{N}, we have for any 𝒙,𝒚{\bm{x}},{\bm{y}}:

|(ρ(He)ρ(H+))(𝒙,𝒚)|\displaystyle\big{|}\big{(}\rho^{\prime}(H_{e})-\rho^{\prime}(H_{+})\big{)}({\bm{x}},{\bm{y}})\big{|} CN(1+d1(𝒙,𝒚)+d1(𝒙,Uc)+d1(𝒚,Uc))3N\displaystyle\leq C_{N}\big{(}1+d_{1}({\bm{x}},{\bm{y}})+d_{1}({\bm{x}},U^{c})+d_{1}({\bm{y}},U^{c})\big{)}^{-3N} (3.19)
CNeN(dl(𝒙,𝒚)+dl(𝒙,Uc)+dl(𝒚,Uc)).\displaystyle\leq C_{N}e^{-N(d_{l}({\bm{x}},{\bm{y}})+d_{l}({\bm{x}},U^{c})+d_{l}({\bm{y}},U^{c}))}. (3.20)

3. Therefore, from (3.20):

|((ρ(He)ρ(H+))𝟙U)(𝒙,𝒚)|\displaystyle\big{|}\big{(}(\rho^{\prime}(H_{e})-\rho^{\prime}(H_{+}))\mathds{1}_{U}\big{)}({\bm{x}},{\bm{y}})\big{|} CNeNdl(𝒙,𝒚)Ndl(𝒙,Uc)Ndl(𝒚,Uc)𝟙U(𝒚)\displaystyle\leq C_{N}e^{-Nd_{l}({\bm{x}},{\bm{y}})-Nd_{l}({\bm{x}},U^{c})-Nd_{l}({\bm{y}},U^{c})}\mathds{1}_{U}({\bm{y}}) (3.21)
CNeNdl(𝒙,𝒚)Ndl(𝒙,Uc)Ndl(𝒚,Uc)eNdl(𝒚,U)\displaystyle\leq C_{N}e^{-Nd_{l}({\bm{x}},{\bm{y}})-Nd_{l}({\bm{x}},U^{c})-Nd_{l}({\bm{y}},U^{c})}\cdot e^{-Nd_{l}({\bm{y}},U)} (3.22)
CNeN2dl(𝒙,𝒚)N2dl(𝒙,Uc)N2dl(𝒚,Uc)N2dl(𝒚,U)N2dl(𝒙,U)\displaystyle\leq C_{N}e^{-\frac{N}{2}d_{l}({\bm{x}},{\bm{y}})-\frac{N}{2}d_{l}({\bm{x}},U^{c})-\frac{N}{2}d_{l}({\bm{y}},U^{c})-\frac{N}{2}d_{l}({\bm{y}},U)-\frac{N}{2}d_{l}({\bm{x}},U)} (3.23)
CNeN2dl(𝒙,𝒚)N2dl(𝒙,U)N2dl(𝒚,U).\displaystyle\leq C_{N}e^{-\frac{N}{2}d_{l}({\bm{x}},{\bm{y}})-\frac{N}{2}d_{l}({\bm{x}},{\partial}U)-\frac{N}{2}d_{l}({\bm{y}},{\partial}U)}. (3.24)

The same upper bound holds for (ρ(He)ρ(H))𝟙Uc\big{(}\rho^{\prime}(H_{e})-\rho^{\prime}(H_{-})\big{)}\mathds{1}_{U^{c}}. Going back to (3.11), we obtain

|ρ(He)(𝒙,𝒚)|CNeN2dl(𝒙,𝒚)N2dl(𝒙,U)N2dl(𝒚,U).\big{|}\rho^{\prime}(H_{e})({\bm{x}},{\bm{y}})\big{|}\leq C_{N}e^{-\frac{N}{2}d_{l}({\bm{x}},{\bm{y}})-\frac{N}{2}d_{l}({\bm{x}},{\partial}U)-\frac{N}{2}d_{l}({\bm{y}},{\partial}U)}. (3.25)

By (2.4) in Lemma 2.2, for any NN, there is CNC_{N} such that

|[He,𝟙V](𝒙,𝒚)|CNeNdl(𝒙,𝒚)Ndl(𝒙,V)Ndl(𝒚,V).\big{|}[H_{e},\mathds{1}_{V}]({\bm{x}},{\bm{y}})\big{|}\leq C_{N}e^{-Nd_{l}({\bm{x}},{\bm{y}})-Nd_{l}({\bm{x}},{\partial}V)-Nd_{l}({\bm{y}},{\partial}V)}. (3.26)

Since U,VU,V are transverse sets, by Corollary 2.4, ρ(He)[He,𝟙U]\rho^{\prime}(H_{e})[H_{e},\mathds{1}_{U}] is trace-class; hence σe(He)\sigma_{e}(H_{e}) is well-defined. ∎

4. Equality

Theorem 2.

Under Assumption 1,

σeU,V(He)=σbU,V(P+)σbU,V(P).\sigma_{e}^{U,V}(H_{e})=\sigma_{b}^{U,V}(P_{+})-\sigma_{b}^{U,V}(P_{-}). (4.1)
Proof.
Refer to caption
(a)
Refer to caption
(b)
Figure 3. The white, light blue, and dark blue region in (b)(b) represents URU_{R}^{-}, UR+U_{R}^{+}, and URU_{R}^{\partial} respectively.

We follow an approach due to Elgart–Graf–Schenker [EGS05], tailored here to our geometic setup. Define

UR:={𝒙:d(𝒙,U)<R},UR+:=(UR)cU,UR:=(UR)cUc,U_{R}^{{\partial}}:=\{{\bm{x}}:d({\bm{x}},{\partial}U)<R\},\qquad U_{R}^{+}:=\big{(}U_{R}^{{\partial}}\big{)}^{c}\cap U,\qquad U_{R}^{-}:=\big{(}U_{R}^{{\partial}}\big{)}^{c}\cap U^{c},

and correspondingly,

ζR(H):=Tr(ρ~z¯(Hz)1[H,𝟙V][(Hz)1,𝟙UR]dzdz¯),{,+,}.\zeta_{R}^{*}(H):={\operatorname{Tr}}\left(\int\frac{{\partial}{\tilde{\rho}}}{{\partial}\bar{z}}(H-z)^{-1}[H,\mathds{1}_{V}]\big{[}(H-z)^{-1},\mathds{1}_{U_{R}^{*}}\big{]}dz\wedge d\bar{z}\right),\qquad*\in\{{\partial},+,-\}. (4.2)

This quantity ζR(H)\zeta_{R}^{*}(H) plays an intermediate role in proving equality (4.1). In fact, we prove (4.1) by proving the following three equalities:

σe(He)\xlongequalClaim 1limRζR(He)\xlongequalClaim 2limRζR+(H+)ζR(H)\xlongequalClaim 3σbU,V(P+)σbU,V(P).\begin{split}\sigma_{e}(H_{e})\xlongequal{\text{Claim~{}}\ref{claim-4a}}\lim_{R\to\infty}\zeta_{R}^{\partial}(H_{e})\xlongequal{\text{Claim~{}}\ref{claim-4b}}\lim_{R\to\infty}-\zeta_{R}^{+}(H_{+})-\zeta_{R}^{-}(H_{-})\xlongequal{\text{Claim~{}}\ref{claim-4c}}\sigma_{b}^{U,V}(P_{+})-\sigma_{b}^{U,V}(P_{-}).\end{split} (4.3)

As explained in §1.4, to prove Claim 1, we inject a cutoff near U{\partial}U and justify that terms away from U{\partial}U are negligible. Claim 2 reduces this near-U{\partial}U edge quantity, ζR(He)\zeta_{R}^{\partial}(H_{e}), to the difference of two bulk quantities, ζR±(H±)\zeta_{R}^{\pm}(H_{\pm}). Claim 3 deform the bulk quantities, ζR±(H±)\zeta_{R}^{\pm}(H_{\pm}) (in the limit), to the geometric bulk conductances, σbU,V(P±)\sigma_{b}^{U,V}(P_{\pm}).

Now we state and prove these three claims.

Claim 1.
σe(He)=limRζR(He).\sigma_{e}(H_{e})=\lim\limits_{R\to\infty}\zeta_{R}^{\partial}(H_{e}).
Proof of Claim 1.

Since 2=URUR+UR\mathbb{R}^{2}=U_{R}^{\partial}\sqcup U_{R}^{+}\sqcup U_{R}^{-}, we split σe(He)\sigma_{e}(H_{e}) into two parts:

2πi[He,𝟙V]ρ(He)=2πi[He,𝟙V](𝟙UR+(𝟙UR++𝟙UR))ρ(He):=I+II.2\pi i[H_{e},\mathds{1}_{V}]\rho^{\prime}(H_{e})=2\pi i[H_{e},\mathds{1}_{V}]\big{(}\mathds{1}_{U_{R}^{\partial}}+(\mathds{1}_{U_{R}^{+}}+\mathds{1}_{U_{R}^{-}})\big{)}\rho^{\prime}(H_{e}):={\operatorname{I}}+{\operatorname{II}}. (4.4)

We show that II{\operatorname{II}} is trace-class and that Tr(II)0{\operatorname{Tr}}({\operatorname{II}})\to 0 when R+R\to+\infty. By (3.25), we have

|𝟙UR±ρ(He)(𝒙,𝒚)|\displaystyle\big{|}\mathds{1}_{U_{R}^{\pm}}\rho^{\prime}(H_{e})({\bm{x}},{\bm{y}})\big{|} CNe2Ndl(𝒙,(UR)c)2Ndl(𝒙,𝒚)2Ndl(𝒙,U)2Ndl(𝒚,U)\displaystyle\leq C_{N}e^{-2Nd_{l}\left({\bm{x}},(U_{R}^{\partial})^{c}\right)-2Nd_{l}({\bm{x}},{\bm{y}})-2Nd_{l}({\bm{x}},{\partial}U)-2Nd_{l}({\bm{y}},{\partial}U)} (4.5)
CN(1+R)NeNdl(𝒙,𝒚)Ndl(𝒙,U)Ndl(𝒚,U),\displaystyle\leq C_{N}(1+R)^{-N}e^{-Nd_{l}({\bm{x}},{\bm{y}})-Nd_{l}({\bm{x}},{\partial}U)-Nd_{l}({\bm{y}},{\partial}U)}, (4.6)

where we used |𝟙UR±(𝒙)|e2Ndl(𝒙,(UR)c)\big{|}\mathds{1}_{U_{R}^{\pm}}({\bm{x}})\big{|}\leq e^{-2Nd_{l}\left({\bm{x}},(U_{R}^{\partial})^{c}\right)} and dl(𝒙,(UR)c)+dl(𝒙,U)ln(1+R)d_{l}\left({\bm{x}},(U_{R}^{\partial})^{c}\right)+d_{l}({\bm{x}},{\partial}U)\geq\ln(1+R). Because of (3.26) and (4.6), Corollary 2.4 implies that II{\operatorname{II}} is trace-class. We now estimate its trace. By (3.26) and (2.5) in Lemma 2.3, we get

|II(𝒙,𝒚)|\displaystyle|\text{II}({\bm{x}},{\bm{y}})| CN(1+R)NeN4dl(𝒙,𝒚)N4dl(𝒙,U)N4dl(𝒚,U)N4dl(𝒙,V)N4dl(𝒚,V)\displaystyle\leq C_{N}(1+R)^{-N}e^{-\frac{N}{4}d_{l}({\bm{x}},{\bm{y}})-\frac{N}{4}d_{l}({\bm{x}},{\partial}U)-\frac{N}{4}d_{l}({\bm{y}},{\partial}U)-\frac{N}{4}d_{l}({\bm{x}},{\partial}V)-\frac{N}{4}d_{l}({\bm{y}},{\partial}V)} (4.7)

As a result,

|Tr(II)|𝒙|II(𝒙,𝒙)|CN(1+R)N𝒙eN2lnΨU,V(𝒙).\big{|}{\operatorname{Tr}}({\operatorname{II}})\big{|}\leq\sum\limits_{{\bm{x}}}|{\operatorname{II}}({\bm{x}},{\bm{x}})|\leq C_{N}(1+R)^{-N}\cdot\sum_{{\bm{x}}}e^{-\frac{N}{2}\ln\Psi_{U,V}({\bm{x}})}.

By (2.10), the sum is finite for N>4/cN>4/c. It follows that Tr(II)0{\operatorname{Tr}}({\operatorname{II}})\to 0 as RR\rightarrow\infty.

It remains to prove Tr(I)=ζR(He){\operatorname{Tr}}({\operatorname{I}})=\zeta_{R}^{\partial}(H_{e}). Without loss of generalities, zρ~=zρ~\widetilde{{\partial}_{z}\rho}={\partial}_{z}\tilde{\rho} – see e.g. [D21, (2.3.5)]. Combining the Helffer–Sjöstrand formula (LABEL:eq:_HS_1) with integration by parts, we obtain

I=2πi[He,𝟙V]𝟙URρ(He)\displaystyle{\operatorname{I}}=2\pi i[H_{e},\mathds{1}_{V}]\mathds{1}_{U_{R}^{\partial}}\cdot\rho^{\prime}(H_{e}) =[He,𝟙V]𝟙URρ~z¯(Hez)1𝑑zdz¯\displaystyle=[H_{e},\mathds{1}_{V}]\mathds{1}_{U_{R}^{\partial}}\cdot\int\frac{{\partial}{\tilde{\rho}}^{\prime}}{{\partial}\bar{z}}(H_{e}-z)^{-1}dz\wedge d\bar{z} (4.8)
=[He,𝟙V]𝟙UR2ρ~zz¯(Hez)1𝑑zdz¯\displaystyle=[H_{e},\mathds{1}_{V}]\mathds{1}_{U_{R}^{\partial}}\cdot\int\frac{{\partial}^{2}{\tilde{\rho}}}{{\partial}z{\partial}\bar{z}}(H_{e}-z)^{-1}dz\wedge d\bar{z} (4.9)
=ρ~z¯[He,𝟙V]𝟙UR(Hez)2𝑑zdz¯.\displaystyle=-\int\frac{{\partial}{\tilde{\rho}}}{{\partial}\bar{z}}[H_{e},\mathds{1}_{V}]\mathds{1}_{U_{R}^{\partial}}\cdot(H_{e}-z)^{-2}dz\wedge d\bar{z}. (4.10)

If 𝒙UR{\bm{x}}\in U_{R}^{\partial}, then dl(𝒙,U)ln(1+R)d_{l}({\bm{x}},{\partial}U)\leq\ln(1+R). Thus for any NN,

|𝟙UR(𝒙)|eN(ln(1+R)dl(𝒙,U))(1+R)NeNdl(𝒙,U).|\mathds{1}_{U_{R}^{\partial}}({\bm{x}})|\leq e^{N\big{(}\ln(1+R)-d_{l}({\bm{x}},{\partial}U)\big{)}}\leq(1+R)^{N}e^{-Nd_{l}({\bm{x}},{\partial}U)}. (4.11)

Combining with (3.26) and using Corollary 2.4, we see that [He,𝟙V]𝟙UR[H_{e},\mathds{1}_{V}]\mathds{1}_{U_{R}^{\partial}} is trace-class. Using (Hz)1|Imz|1\|(H-z)^{-1}\|\leq|\operatorname{Im}z|^{-1}, we have:

[He,𝟙V]𝟙UR(Hez)21C|Imz|2.\left\|[H_{e},\mathds{1}_{V}]\mathds{1}_{U_{R}^{\partial}}(H_{e}-z)^{-2}\right\|_{1}\leq\dfrac{C}{|\operatorname{Im}z|^{2}}. (4.12)

Therefore, we can take the trace on both sides of (4.10) and switch trace and integral (using almost analyticity of ρ~\tilde{\rho}). This yields:

Tr(I)=ρ~z¯Tr([He,𝟙V]𝟙UR(Hez)2)𝑑zdz¯.{\operatorname{Tr}}({\operatorname{I}})=-\int\frac{{\partial}{\tilde{\rho}}}{{\partial}\bar{z}}{\operatorname{Tr}}\big{(}[H_{e},\mathds{1}_{V}]\mathds{1}_{U_{R}^{\partial}}(H_{e}-z)^{-2}\big{)}dz\wedge d\bar{z}. (4.13)

By (3.26), (4.11), and Corollary 2.4, the operators below are all trace-class. Thus we can apply cyclicity (3.8) to get

Tr(I)=ρ~z¯Tr((Hez)1[He,𝟙V]𝟙UR(Hez)1)𝑑zdz¯=ρ~z¯Tr((Hez)1[He,𝟙V](Hez)1𝟙UR)𝑑zdz¯+ρ~z¯Tr((Hez)1[He,𝟙V][(Hez)1,𝟙UR])𝑑zdz¯=ρ~z¯Tr([(Hez)1,𝟙V]𝟙UR)𝑑zdz¯+ζR(He)=Tr([ρ(He),𝟙V]𝟙UR)+ζR(He),\begin{split}{\operatorname{Tr}}({\operatorname{I}})&=-\int\frac{{\partial}{\tilde{\rho}}}{{\partial}\bar{z}}{\operatorname{Tr}}\left((H_{e}-z)^{-1}[H_{e},\mathds{1}_{V}]\mathds{1}_{U_{R}^{\partial}}(H_{e}-z)^{-1}\right)dz\wedge d\bar{z}\\ &=-\int\frac{{\partial}{\tilde{\rho}}}{{\partial}\bar{z}}{\operatorname{Tr}}\left((H_{e}-z)^{-1}[H_{e},\mathds{1}_{V}](H_{e}-z)^{-1}\mathds{1}_{U_{R}^{\partial}}\right)dz\wedge d\bar{z}\\ &\ \ \ +\int\frac{{\partial}{\tilde{\rho}}}{{\partial}\bar{z}}{\operatorname{Tr}}\left((H_{e}-z)^{-1}[H_{e},\mathds{1}_{V}]\big{[}(H_{e}-z)^{-1},\mathds{1}_{U_{R}^{\partial}}\big{]}\right)dz\wedge d\bar{z}\\ &=\int\frac{{\partial}{\tilde{\rho}}}{{\partial}\bar{z}}{\operatorname{Tr}}\left(\big{[}(H_{e}-z)^{-1},\mathds{1}_{V}\big{]}\mathds{1}_{U_{R}^{\partial}}\right)dz\wedge d\bar{z}+\zeta_{R}^{\partial}(H_{e})\\ &={\operatorname{Tr}}\big{(}\big{[}\rho(H_{e}),\mathds{1}_{V}\big{]}\mathds{1}_{U_{R}^{\partial}}\big{)}+\zeta_{R}^{\partial}(H_{e}),\end{split}

where we used Helffer-Sjöstrand formula (LABEL:eq:_HS_1) for the last equality. Finally,

Tr([ρ(He),𝟙V]𝟙UR)=Tr(𝟙Vcρ(He)𝟙V𝟙UR)Tr(𝟙Vρ(He)𝟙Vc𝟙UR)=Tr(𝟙V𝟙Vcρ(He)𝟙UR)Tr(𝟙V𝟙Vcρ(He)𝟙UR)=0\begin{split}{\operatorname{Tr}}\big{(}\big{[}\rho(H_{e}),\mathds{1}_{V}\big{]}\mathds{1}_{U_{R}^{\partial}}\big{)}&={\operatorname{Tr}}\big{(}\mathds{1}_{V^{c}}\rho(H_{e})\mathds{1}_{V}\mathds{1}_{U_{R}^{\partial}}\big{)}-{\operatorname{Tr}}\big{(}\mathds{1}_{V}\rho(H_{e})\mathds{1}_{V^{c}}\mathds{1}_{U_{R}^{\partial}}\big{)}\\ &={\operatorname{Tr}}\big{(}\mathds{1}_{V}\mathds{1}_{V^{c}}\rho(H_{e})\mathds{1}_{U_{R}^{\partial}}\big{)}-{\operatorname{Tr}}\big{(}\mathds{1}_{V}\mathds{1}_{V^{c}}\rho(H_{e})\mathds{1}_{U_{R}^{\partial}}\big{)}=0\end{split}

as 𝟙Vcρ(He)𝟙V𝟙UR\mathds{1}_{V^{c}}\rho(H_{e})\mathds{1}_{V}\mathds{1}_{U_{R}^{\partial}}, 𝟙Vρ(He)𝟙Vc𝟙UR\mathds{1}_{V}\rho(H_{e})\mathds{1}_{V^{c}}\mathds{1}_{U_{R}^{\partial}}, and 𝟙V𝟙Vcρ(He)𝟙UR\mathds{1}_{V}\mathds{1}_{V^{c}}\rho(H_{e})\mathds{1}_{U_{R}^{\partial}} are all trace-class by Lemma 2.1(b) and Corollary 2.4. Thus we get Tr(I)=ζR(He){\operatorname{Tr}}({\operatorname{I}})=\zeta_{R}^{\partial}(H_{e}); this completes the proof of Claim 1. ∎

Claim 2.

We have:

limR(ζR(He)+ζR+(H+)+ζR(H))=0.\lim\limits_{R\to\infty}\big{(}\zeta_{R}^{\partial}(H_{e})+\zeta_{R}^{+}(H_{+})+\zeta_{R}^{-}(H_{-})\big{)}=0. (4.14)
Proof of Claim 2.

Recall that 𝟙UR+𝟙UR++𝟙UR=1\mathds{1}_{U_{R}^{\partial}}+\mathds{1}_{U_{R}^{+}}+\mathds{1}_{U_{R}^{-}}=1. Since [1,(Hez)1]=0\big{[}1,(H_{e}-z)^{-1}\big{]}=0, from the definition (4.2), we get ζR(He)+ζR+(He)+ζR(He)=0\zeta_{R}^{\partial}(H_{e})+\zeta_{R}^{+}(H_{e})+\zeta_{R}^{-}(H_{e})=0. To show (4.14), it is enough to show

limR|ζR±(He)ζR±(H±)|=0.\lim_{R\to\infty}|\zeta_{R}^{\pm}(H_{e})-\zeta_{R}^{\pm}(H_{\pm})|=0. (4.15)

Denote AR+(H)=(Hz)1[H,𝟙V][(Hz)1,𝟙UR+]A_{R}^{+}(H)=(H-z)^{-1}[H,\mathds{1}_{V}]\big{[}(H-z)^{-1},\mathds{1}_{U_{R}^{+}}\big{]}. We have:

|ζR±(He)ζR±(H±)|\displaystyle\big{|}\zeta_{R}^{\pm}(H_{e})-\zeta_{R}^{\pm}(H_{\pm})\big{|} =|Trρ~z¯(AR+(He)AR+(H+))𝑑zdz¯|\displaystyle=\left|{\operatorname{Tr}}\int\frac{{\partial}{\tilde{\rho}}}{{\partial}\bar{z}}\left(A_{R}^{+}(H_{e})-A_{R}^{+}(H_{+})\right)dz\wedge d\bar{z}\right| (4.16)
|ρ~z¯|(AR+(He)AR+(H+))1|dzdz¯|.\displaystyle\leq\int\left|\frac{{\partial}{\tilde{\rho}}}{{\partial}\bar{z}}\right|\left\|\left(A_{R}^{+}(H_{e})-A_{R}^{+}(H_{+})\right)\right\|_{1}|dz\wedge d\bar{z}|. (4.17)

We rewrite

AR+(He)AR+(H+)=((Hez)1(H+z)1)[He,𝟙V][(Hez)1,𝟙UR+]+(H+z)1[HeH+,𝟙V][(Hez)1,𝟙UR+]+(H+z)1[H+,𝟙V][(Hez)1(H+z)1,𝟙UR+]=(Hez)1(H+He)(H+z)1[He,𝟙V][(Hez)1,𝟙UR+]+(H+z)1[HeH+,𝟙V][(Hez)1,𝟙UR+]+(H+z)1[H+,𝟙V][(Hez)1(H+He)(H+z)1,𝟙UR+]=:A1+A2+A3.\begin{split}A_{R}^{+}(H_{e})-A_{R}^{+}(H_{+})=&\left((H_{e}-z)^{-1}-(H_{+}-z)^{-1}\right)[H_{e},\mathds{1}_{V}]\big{[}(H_{e}-z)^{-1},\mathds{1}_{U_{R}^{+}}\big{]}\\ +&(H_{+}-z)^{-1}[H_{e}-H_{+},\mathds{1}_{V}]\big{[}(H_{e}-z)^{-1},\mathds{1}_{U_{R}^{+}}\big{]}\\ +&(H_{+}-z)^{-1}[H_{+},\mathds{1}_{V}]\big{[}(H_{e}-z)^{-1}-(H_{+}-z)^{-1},\mathds{1}_{U_{R}^{+}}\big{]}\\ =&(H_{e}-z)^{-1}(H_{+}-H_{e})(H_{+}-z)^{-1}[H_{e},\mathds{1}_{V}]\big{[}(H_{e}-z)^{-1},\mathds{1}_{U_{R}^{+}}\big{]}\\ +&(H_{+}-z)^{-1}[H_{e}-H_{+},\mathds{1}_{V}]\big{[}(H_{e}-z)^{-1},\mathds{1}_{U_{R}^{+}}\big{]}\\ +&(H_{+}-z)^{-1}[H_{+},\mathds{1}_{V}]\big{[}(H_{e}-z)^{-1}(H_{+}-H_{e})(H_{+}-z)^{-1},\mathds{1}_{U_{R}^{+}}\big{]}\\ =:&A_{1}+A_{2}+A_{3}.\end{split} (4.18)

Note that:

  1. (1)

    By (2.4) and using that H±H_{\pm}, HeH_{e} are PSR, for any NN, there exists CNC_{N} with

    |[H,𝟙V](𝒙,𝒚)|CNe8Ndl(𝒙,𝒚)8Ndl(𝒙,V)8Ndl(𝒙,V),{±,e}.\big{|}[H_{*},\mathds{1}_{V}]({\bm{x}},{\bm{y}})\big{|}\leq C_{N}e^{-8Nd_{l}({\bm{x}},{\bm{y}})-8Nd_{l}({\bm{x}},\partial V)-8Nd_{l}({\bm{x}},\partial V)},\qquad*\in\{\pm,\text{e}\}.
  2. (2)

    By an argument similar to that leading to (3.15), for any NN, there is CNC_{N} such that

    |(HeH+)(𝒙,𝒚)|CNe8Ndl(𝒙,𝒚)8Ndl(𝒙,Uc)8dl(𝒚,Uc).\big{|}(H_{e}-H_{+})({\bm{x}},{\bm{y}})\big{|}\leq C_{N}e^{-8Nd_{l}({\bm{x}},{\bm{y}})-8Nd_{l}({\bm{x}},U^{c})-8d_{l}({\bm{y}},U^{c})}.
  3. (3)

    Recall that 𝟙UR+(𝒙)e8Ndl(𝒙,UR+)\mathds{1}_{U_{R}^{+}}({\bm{x}})\leq e^{-8Nd_{l}({\bm{x}},U_{R}^{+})}.

Note that each AjA_{j}, j{1,2,3}j\in\{1,2,3\} contains HeH+H_{e}-H_{+}, [H,𝟙V][H_{*},\mathds{1}_{V}] and 𝟙UR+\mathds{1}_{U_{R}^{+}}. By (1)(1), (2)(2), (3)(3) above, the control of the resolvent norm provided by Lemma 2.1(a) and (2.5) in Lemma 2.3, we have

|Aj(𝒙,𝒚)|CN|Imz|3e2N(dl(𝒙,𝒚)+dl(𝒙,Uc)+dl(𝒚,Uc)+dl(𝒙,V)+dl(𝒚,V)+dl(𝒙,UR+)+dl(𝒚,UR+))CN|Imz|3(1+R)2NeN(dl(𝒙,𝒚)+dl(𝒙,U)+dl(𝒙,V)+dl(𝒚,U)+dl(𝒚,V))\begin{split}|A_{j}({\bm{x}},{\bm{y}})|&\leq\frac{C_{N}}{|\operatorname{Im}z|^{3}}e^{-2N\big{(}d_{l}({\bm{x}},{\bm{y}})+d_{l}({\bm{x}},U^{c})+d_{l}({\bm{y}},U^{c})+d_{l}({\bm{x}},{\partial}V)+d_{l}({\bm{y}},{\partial}V)+d_{l}({\bm{x}},U_{R}^{+})+d_{l}({\bm{y}},U_{R}^{+})\big{)}}\\ &\leq\frac{C_{N}}{|\operatorname{Im}z|^{3}}(1+R)^{-2N}e^{-N(d_{l}({\bm{x}},{\bm{y}})+d_{l}({\bm{x}},\partial U)+d_{l}({\bm{x}},{\partial}V)+d_{l}({\bm{y}},{\partial}U)+d_{l}({\bm{y}},{\partial}V))}\end{split}

where we used dl(𝒙,Uc)+dl(𝒙,UR+)dl(Uc,UR+)ln(1+R)d_{l}({\bm{x}},U^{c})+d_{l}({\bm{x}},U_{R}^{+})\geq d_{l}(U^{c},U_{R}^{+})\geq\ln(1+R) and dl(𝒙,Uc)+dl(𝒙,UR+)dl(𝒙,U)d_{l}({\bm{x}},U^{c})+d_{l}({\bm{x}},U_{R}^{+})\geq d_{l}({\bm{x}},\partial U). Therefore,

|(AR+(He)AR+(H+))(𝒙,𝒚)|CN|Imz|3(1+R)2NeN(dl(𝒙,𝒚)+dl(𝒙,U)+dl(𝒙,V)+dl(𝒚,U)+dl(𝒚,V)).\big{|}\big{(}A_{R}^{+}(H_{e})-A_{R}^{+}(H_{+})\big{)}({\bm{x}},{\bm{y}})\big{|}\leq\frac{C_{N}}{|\operatorname{Im}z|^{3}}(1+R)^{-2N}e^{-N(d_{l}({\bm{x}},{\bm{y}})+d_{l}({\bm{x}},\partial U)+d_{l}({\bm{x}},\partial V)+d_{l}({\bm{y}},{\partial}U)+d_{l}({\bm{y}},{\partial}V))}.

Since U,VU,V are transverse sets, when N2/cN\geq 2/c, by (2.10), we obtain:

AR+(He)AR+(H+)1𝒙,𝒚|(AR+(He)AR+(H+))(𝒙,𝒚)|CN|Imz|3(1+R)2N.\left\|A_{R}^{+}(H_{e})-A_{R}^{+}(H_{+})\right\|_{1}\leq\sum\limits_{{\bm{x}},{\bm{y}}}\big{|}\big{(}A_{R}^{+}(H_{e})-A_{R}^{+}(H_{+})\big{)}({\bm{x}},{\bm{y}})\big{|}\leq\frac{C_{N}}{|\operatorname{Im}z|^{3}}(1+R)^{-2N}.

By almost analyticity,

|ρ~z¯|(AR+(He)AR+(H+))1|dzdz¯|CN(1+R)2N.\int\left|\frac{{\partial}{\tilde{\rho}}}{{\partial}\bar{z}}\right|\left\|\left(A_{R}^{+}(H_{e})-A_{R}^{+}(H_{+})\right)\right\|_{1}|dz\wedge d\bar{z}|\leq C_{N}(1+R)^{-2N}.

It suffices to go back to (4.17) and take RR\rightarrow\infty to obtain (4.15). This completes the proof. ∎

Claim 3.

For any R>0R>0,

ζR±(H±)=σbUR±,V(P±)=σbU,V(P±).\zeta_{R}^{\pm}(H_{\pm})=-\sigma_{b}^{U_{R}^{\pm},V}(P_{\pm})=\mp\sigma_{b}^{U,V}(P_{\pm}). (4.19)
Proof of Claim 3.

Note that UR+ΔUURU_{R}^{+}\Delta U\subset U_{R}^{\partial} and URΔUcURU_{R}^{-}\Delta U^{c}\subset U_{R}^{\partial}. By Proposition 2,

σbUR+,V(P+)=σbU,V(P+),σbUR,V(P)=σbUc,V(P)=σbU,V(P).\begin{split}&\sigma_{b}^{U_{R}^{+},V}(P_{+})=\sigma_{b}^{U,V}(P_{+}),\qquad\sigma_{b}^{U_{R}^{-},V}(P_{-})=\sigma_{b}^{U^{c},V}(P_{-})=-\sigma_{b}^{U,V}(P_{-}).\end{split}

Hence we get the second equality in (4.19). WLOG we prove ζR+(H+)=σbUR+,V(P+)\zeta_{R}^{+}(H_{+})=-\sigma_{b}^{U_{R}^{+},V}(P_{+}) below. Recall that

ζR+(H+):=Tr(ρ~z¯(H+z)1[H+,𝟙V][(H+z)1,𝟙UR+]dzdz¯)=:Tr(A).\begin{split}\zeta_{R}^{+}(H_{+}):={\operatorname{Tr}}\left(\int_{\mathbb{C}}\frac{{\partial}{\tilde{\rho}}}{{\partial}\bar{z}}(H_{+}-z)^{-1}[H_{+},\mathds{1}_{V}]\big{[}(H_{+}-z)^{-1},\mathds{1}_{U_{R}^{+}}\big{]}dz\wedge d\bar{z}\right)=:{\operatorname{Tr}}(A).\end{split} (4.20)

Expand [(H+z)1,𝟙UR+][(H_{+}-z)^{-1},\mathds{1}_{U_{R}^{+}}] and use (H+z)1[H+,𝟙V](H+z)1=[(H+z)1,𝟙V](H_{+}-z)^{-1}[H_{+},\mathds{1}_{V}](H_{+}-z)^{-1}=-\big{[}(H_{+}-z)^{-1},\mathds{1}_{V}\big{]}, we get

A=ρ~z¯([(H+z)1,𝟙V]𝟙UR+(H+z)1[H+,𝟙V]𝟙UR+(H+z)1)𝑑zdz¯=:A1A2.\begin{split}A&=\int_{\mathbb{C}}\frac{{\partial}{\tilde{\rho}}}{{\partial}\bar{z}}\left(-\big{[}(H_{+}-z)^{-1},\mathds{1}_{V}\big{]}\mathds{1}_{U_{R}^{+}}-(H_{+}-z)^{-1}[H_{+},\mathds{1}_{V}]\mathds{1}_{U_{R}^{+}}(H_{+}-z)^{-1}\right)dz\wedge d\bar{z}\\ &=:-A_{1}-A_{2}.\end{split}

By the Helffer-Sjöstrand formula (LABEL:eq:_HS_1), we have

A1=[ρ(H+),𝟙V]𝟙UR+.A_{1}=\big{[}\rho(H_{+}),\mathds{1}_{V}\big{]}\mathds{1}_{U_{R}^{+}}. (4.21)

To conlude, we adapt a trick from [EGS05].

Lemma 4.1.

We have P+A2P+=P+A2P+=0P_{+}A_{2}P_{+}=P_{+}^{\perp}A_{2}P_{+}^{\perp}=0.

Proof.

By functional calculus,

P+A2P+=ρ~z¯P+(H+z)1[H+,𝟙V]𝟙UR+(H+z)1P+𝑑zdz¯=ρ~z¯P+(P+H+P+z)1[H+,𝟙V]𝟙UR+(P+H+P+z)1P+𝑑zdz¯=:ρ~z¯K(z)dzdz¯.\begin{split}P_{+}A_{2}P_{+}&=\int_{\mathbb{C}}\frac{{\partial}{\tilde{\rho}}}{{\partial}\bar{z}}P_{+}(H_{+}-z)^{-1}[H_{+},\mathds{1}_{V}]\mathds{1}_{U_{R}^{+}}(H_{+}-z)^{-1}P_{+}dz\wedge d\bar{z}\\ &=\int_{\mathbb{C}}\frac{{\partial}{\tilde{\rho}}}{{\partial}\bar{z}}P_{+}(P_{+}H_{+}P_{+}-z)^{-1}[H_{+},\mathds{1}_{V}]\mathds{1}_{U_{R}^{+}}(P_{+}H_{+}P_{+}-z)^{-1}P_{+}dz\wedge d\bar{z}\\ &=:\int_{\mathbb{C}}\frac{{\partial}{\tilde{\rho}}}{{\partial}\bar{z}}K(z)dz\wedge d\bar{z}.\end{split}

Since λ𝒢Σ(H+)c\lambda\in{\mathcal{G}}\subset\Sigma(H_{+})^{c} while ρ𝒢\rho\in{\mathcal{E}}_{\mathcal{G}}, ρ~\tilde{\rho} is the almost analytic extension of ρ\rho, Σ(P+H+P+)supp(ρ~)=\Sigma(P_{+}H_{+}P_{+})\cap\mathrm{supp}(\tilde{\rho})=\emptyset. Hence (P+H+P+z)1(P_{+}H_{+}P_{+}-z)^{-1}, and thus the whole integrand, denoted as K(z)K(z), is analytic on supp(ρ~)\mathrm{supp}(\tilde{\rho}). Choose RR large enough such that supp(ρ~)𝔻R(0)\mathrm{supp}(\tilde{\rho})\subset\mathbb{D}_{R}(0). By Stokes theorem,

ρ~z¯K(z)𝑑zdz¯=𝔻R(0)ρ~(z,z¯)K(z)𝑑z=0.\int_{\mathbb{C}}\frac{{\partial}{\tilde{\rho}}}{{\partial}\bar{z}}K(z)dz\wedge d\bar{z}=\oint_{{\partial}\mathbb{D}_{R}(0)}\tilde{\rho}(z,\bar{z})K(z)dz=0.

A similar computation gives

P+A2P+=(1ρ~)z¯P+(P+H+P+z)1[H+,𝟙V]𝟙UR+(P+H+P+z)1P+𝑑zdz¯=:(1ρ~)z¯J(z)dzdz¯=𝔻R(0)(1ρ~)(z,z¯)J(z)𝑑z=𝔻R(0)J(z)𝑑z.\begin{split}P_{+}^{\perp}A_{2}P_{+}^{\perp}&=\int_{\mathbb{C}}\frac{{\partial}(1-\tilde{\rho})}{{\partial}\bar{z}}P_{+}^{\perp}(P_{+}^{\perp}H_{+}P_{+}^{\perp}-z)^{-1}[H_{+},\mathds{1}_{V}]\mathds{1}_{U_{R}^{+}}(P_{+}^{\perp}H_{+}P_{+}^{\perp}-z)^{-1}P_{+}dz\wedge d\bar{z}\\ &=:\int_{\mathbb{C}}\frac{{\partial}(1-\tilde{\rho})}{{\partial}\bar{z}}J(z)dz\wedge d\bar{z}\\ &=\oint_{{\partial}\mathbb{D}_{R}(0)}(1-\tilde{\rho})(z,\bar{z})J(z)dz=\oint_{{\partial}\mathbb{D}_{R}(0)}J(z)dz.\end{split}

However, when 𝒛𝔻R(0){\bm{z}}\in{\partial}\mathbb{D}_{R}(0),

J(z)Cd(z,Σ(H+))2C(RH)2.\|J(z)\|\leq Cd\big{(}z,\Sigma(H_{+})\big{)}^{-2}\leq C(R-\|H\|)^{-2}.

Hence

P+A2P+CR(RH+)0,R.\|P_{+}^{\perp}A_{2}P_{+}^{\perp}\|\leq\frac{CR}{(R-\|H_{+}\|)}\to 0,\qquad R\to\infty.

This completes the proof. ∎

Now write A=P+2A+(P+)2AA=P_{+}^{2}A+(P_{+}^{\perp})^{2}A. Since AA is trace-class, Lemma 4.1 and cyclicity (3.8) yields:

Tr(A)=Tr(P+2A+(P+)2A)=Tr(P+AP++P+AP+)=Tr(P+A1P++P+A1P+).\begin{split}{\operatorname{Tr}}(A)&={\operatorname{Tr}}\big{(}P_{+}^{2}A+(P_{+}^{\perp})^{2}A\big{)}={\operatorname{Tr}}\big{(}P_{+}AP_{+}+P_{+}^{\perp}AP_{+}^{\perp}\big{)}\\ &=-{\operatorname{Tr}}\big{(}P_{+}A_{1}P_{+}+P_{+}^{\perp}A_{1}P_{+}^{\perp}\big{)}.\end{split}

Using A1=[ρ(H+),𝟙V]𝟙UR+A_{1}=[\rho(H_{+}),\mathds{1}_{V}]\mathds{1}_{U_{R}^{+}} and ρ(H+)=1P+=P+\rho(H_{+})=1-P_{+}=P_{+}^{\perp}, we see

P+A1P++P+A1P+=P+[P+,𝟙V]𝟙UR+P++P+[P+,𝟙V]𝟙UR+P+=P+𝟙VP+𝟙UR+P++P+𝟙V𝟙UR+P+P+𝟙VP+𝟙UR+P+=P+𝟙VP+𝟙UR+P++P+𝟙VP+𝟙UR+P+\begin{split}P_{+}A_{1}P_{+}+P_{+}^{\perp}A_{1}P_{+}^{\perp}&=P_{+}[P_{+}^{\perp},\mathds{1}_{V}]\mathds{1}_{U_{R}^{+}}P_{+}+P_{+}^{\perp}[P_{+}^{\perp},\mathds{1}_{V}]\mathds{1}_{U_{R}^{+}}P_{+}^{\perp}\\ &=-P_{+}\mathds{1}_{V}P_{+}^{\perp}\mathds{1}_{U_{R}^{+}}P_{+}+P_{+}^{\perp}\mathds{1}_{V}\mathds{1}_{U_{R}^{+}}P_{+}^{\perp}-P_{+}^{\perp}\mathds{1}_{V}P_{+}^{\perp}\mathds{1}_{U_{R}^{+}}P_{+}^{\perp}\\ &=-P_{+}\mathds{1}_{V}P_{+}^{\perp}\mathds{1}_{U_{R}^{+}}P_{+}+P_{+}^{\perp}\mathds{1}_{V}P_{+}\mathds{1}_{U_{R}^{+}}P_{+}^{\perp}\end{split}

where

Tr(P+𝟙VP+𝟙UR+P+)=Tr(P+𝟙VP+P+𝟙UR+P+)=Tr([P+,𝟙V]P+P+[𝟙UR+,P+])=Tr(P+[𝟙UR+,P+][P+,𝟙V]P+)=Tr(P+𝟙UR+P+𝟙VP+).\begin{split}{\operatorname{Tr}}\big{(}P_{+}^{\perp}\mathds{1}_{V}P_{+}\mathds{1}_{U_{R}^{+}}P_{+}^{\perp}\big{)}&={\operatorname{Tr}}\big{(}P_{+}^{\perp}\mathds{1}_{V}P_{+}\cdot P_{+}\mathds{1}_{U_{R}^{+}}P_{+}^{\perp}\big{)}={\operatorname{Tr}}\big{(}[P_{+}^{\perp},\mathds{1}_{V}]P_{+}\cdot P_{+}[\mathds{1}_{U_{R}^{+}},P_{+}^{\perp}]\big{)}\\ &={\operatorname{Tr}}\big{(}P_{+}[\mathds{1}_{U_{R}^{+}},P_{+}^{\perp}][P_{+}^{\perp},\mathds{1}_{V}]P_{+}\big{)}={\operatorname{Tr}}\big{(}P_{+}\mathds{1}_{U_{R}^{+}}P_{+}^{\perp}\mathds{1}_{V}P_{+}\big{)}.\end{split}

Thus

ζR+(H+)=Tr(A)=Tr(P+A1P++P+A1P+)=Tr(P+𝟙VP+𝟙UR+P+P+𝟙UR+P+𝟙VP+)=Tr(P+𝟙UR+P+𝟙VP+P+𝟙VP+𝟙UR+P+).\begin{split}\zeta_{R}^{+}(H_{+})&={\operatorname{Tr}}(A)=-{\operatorname{Tr}}\big{(}P_{+}A_{1}P_{+}+P_{+}^{\perp}A_{1}P_{+}^{\perp}\big{)}={\operatorname{Tr}}\big{(}P_{+}\mathds{1}_{V}P_{+}^{\perp}\mathds{1}_{U_{R}^{+}}P_{+}-P_{+}\mathds{1}_{U_{R}^{+}}P_{+}^{\perp}\mathds{1}_{V}P_{+}\big{)}\\ &={\operatorname{Tr}}\big{(}P_{+}\mathds{1}_{U_{R}^{+}}P_{+}\mathds{1}_{V}P_{+}-P_{+}\mathds{1}_{V}P_{+}\mathds{1}_{U_{R}^{+}}P_{+}\big{)}.\end{split}

On the other hand, by direct computation,

σbUR+,V(P+)=Tr(P+[[P+,𝟙UR+],[P+,𝟙V]])=Tr(P+𝟙VP+𝟙UR+P+P+𝟙UR+P+𝟙VP+).\sigma_{b}^{U_{R}^{+},V}(P_{+})=-{\operatorname{Tr}}\big{(}P_{+}\big{[}[P_{+},\mathds{1}_{U_{R}^{+}}],[P_{+},\mathds{1}_{V}]\big{]}\big{)}={\operatorname{Tr}}\big{(}P_{+}\mathds{1}_{V}P_{+}\mathds{1}_{U_{R}^{+}}P_{+}-P_{+}\mathds{1}_{U_{R}^{+}}P_{+}\mathds{1}_{V}P_{+}\big{)}.

Thus ζR+(H+)=σbUR+,V(P+)\zeta_{R}^{+}(H_{+})=-\sigma_{b}^{U_{R}^{+},V}(P_{+}). ∎

5. Intersection number between simple transverse sets

In this section, we define an intersection number 𝒳U,V\mathcal{X}_{U,V} between simple transverse sets U,VU,V (see Definition 8). This integer will emerge when expressing geometric bulk conductances in terms of Hall conductances: see Theorem 3 below.

Definition 7 (Simple path).

A continuous map γ:2\gamma:\mathbb{R}\rightarrow\mathbb{R}^{2} is called a simple path if:

  • (i)

    γ\gamma is injective and proper (the preimage of a compact set is compact);

  • (ii)

    There exists a discrete closed set SS\subset\mathbb{R} such that γ\gamma is smooth on S\mathbb{R}\setminus S;

  • (iii)

    For all tt\in\mathbb{R}, the left and right derivatives of γ\gamma at tt exist and have norm 11.

If (ii) and (iii) hold but γ\gamma is periodic and injective over its period, we call γ\gamma a simple loop.

Proposition 4.

Let γ1,γ2\gamma_{1},\gamma_{2} be two simple paths with ranges Γ1,Γ2\Gamma_{1},\Gamma_{2}, such that Γ1ΔΓ2\Gamma_{1}\Delta\Gamma_{2} is compactly supported. Let A1A_{1} be a connected component of 2Γ1\mathbb{R}^{2}\setminus\Gamma_{1}. Then there exists a unique connected component A2A_{2} of 2Γ2\mathbb{R}^{2}\setminus\Gamma_{2} such that A1ΔA2A_{1}\Delta A_{2} is bounded.

\floatbox

[\capbeside\thisfloatsetupcapbesideposition=right,center,capbesidewidth=0.6]figure[\FBwidth] Refer to caption

Figure 4. Simple path splits 2\mathbb{R}^{2} into two halves. When two simple paths Γ1\Gamma_{1}, Γ2\Gamma_{2} differ by a compact set, given A1A_{1} connected component of 2Γ1\mathbb{R}^{2}\setminus\Gamma_{1}, there is a connected component A2A_{2} of 2Γ2\mathbb{R}^{2}\setminus\Gamma_{2} such that A1A_{1}, A2A_{2} differ by a compact set.

Proposition 4 has a flavor reminiscent of the Jordan curve Theorem. While visually obvious (see Figure 4), its proof requires some work. We give a sketch here and defer the proof to Appendix LABEL:app:A:

  • We show that if γ1\gamma_{1} is a proper injective curve, then 2γ1()\mathbb{R}^{2}\setminus\gamma_{1}(\mathbb{R}) has two connected components (Lemma LABEL:lem:4);

  • We then justify that a perturbation Γ2\Gamma_{2} of Γ1\Gamma_{1} on a compact set perturbs the two connected components by a bounded set only (Lemma LABEL:lem:5);

  • We unambiguously define the left side of a simple path (Proposition LABEL:prop:1) and justify that the left components of 2γ1()\mathbb{R}^{2}\setminus\gamma_{1}(\mathbb{R}), 2γ2()\mathbb{R}^{2}\setminus\gamma_{2}(\mathbb{R}) have bounded symmetric difference (see Figure 5 for a pictorial representation and Proposition LABEL:prop:1 for a proper definition of the left of a simple curve).

\floatbox

[\capbeside\thisfloatsetupcapbesideposition=right,center,capbesidewidth=0.6]figure[\FBwidth] Refer to caption

Figure 5. At each point on γ\gamma, we can find small enough disk split in two by γ\gamma and define the left side of γ\gamma as the “conventional” left side when travelling along the path γ\gamma.
Definition 8 (Simple set).

An open subset AA of 2\mathbb{R}^{2} is simple if it is connected and its boundary is the range of a simple path or of a simple loop.

Associated with the distinction between simple paths and simple loops, there are two types of simple sets: those with bounded boundaries (given by simple loops) and those with unbounded boundaries (given by simple paths). By Jordan’s theorem, those with bounded boundaries are bounded or have bounded complements.

5.1. Intersection number

Definition 9 (Intersection number for simple sets).

Let U,VU,V be transverse sets such that UU is simple. If U{\partial}U is unbounded, let γ\gamma be a simple path with range U{\partial}U, such that UU lies to the left of γ\gamma (see Figure 5 for a pictorial representation and Proposition LABEL:prop:1 for a proper definition). We define the intersection number between U,VU,V as:

𝒳U,V=def𝒳+(U,V)𝒳(U,V),𝒳±(U,V)=deflimt±𝟙Vγ(t).\mathcal{X}_{U,V}\ \mathrel{\stackrel{{\scriptstyle\makebox[0.0pt]{\mbox{\tiny def}}}}{{=}}}\ \mathcal{X}_{+}(U,V)-\mathcal{X}_{-}(U,V),\qquad\mathcal{X}_{\pm}(U,V)\ \mathrel{\stackrel{{\scriptstyle\makebox[0.0pt]{\mbox{\tiny def}}}}{{=}}}\ \lim_{t\rightarrow\pm\infty}\mathds{1}_{V}\circ\gamma(t). (5.1)

If U{\partial}U is bounded, we set 𝒳U,V=0\mathcal{X}_{U,V}=0.

We show in Appendix LABEL:app:B that 𝒳U,V\mathcal{X}_{U,V} is correctly defined and independent of the choice of curve γ\gamma describing its boundary. In §LABEL:sec-7.2 we extend Definition 9 to more general transverse sets U,VU,V, such that (loosely speaking) U{\partial}U is made of well-separated curves.

6. Bulk index computation

6.1. Theorem 3 for transverse simple sets

The main theorem here is:

Theorem 3.

Under Assumption 1, we have

σbU,V(P±)=𝒳U,Vσb(P±).\sigma_{b}^{U,V}(P_{\pm})=\mathcal{X}_{U,V}\cdot\sigma_{b}(P_{\pm}). (6.1)

Since UU is simple, by Definition 9, 𝒳U,V{0,±1}\mathcal{X}_{U,V}\in\{0,\pm 1\}. It suffices to prove the following:

  • 1.

    If 𝒳U,V=0\mathcal{X}_{U,V}=0, then σbU,V(P)=0\sigma_{b}^{U,V}(P)=0 (Lemma 6.1).

  • 2.

    If Theorem 3 holds when 𝒳U,V=1\mathcal{X}_{U,V}=1, then it also holds when 𝒳U,V=1\mathcal{X}_{U,V}=-1 (Lemma 6.2).

  • 3.

    Theorem 3 holds when 𝒳U,V=1\mathcal{X}_{U,V}=16.3 - §6.5).

The core of the proof is the third step.

6.2. Steps 1 and 2

Lemma 6.1.

Let U,VU,V be transverse simple sets such that 𝒳U,V=0\mathcal{X}_{U,V}=0, and PP be an ESR projector. Then σbU,V(P)=0\sigma_{b}^{U,V}(P)=0.

Proof.

We recall the notation A𝒞=intAcA^{\mathcal{C}}=\operatorname{int}A^{c}.

1. Assume first that U{\partial}U is bounded. Since UU is a simple set, by definition, U{\partial}U is a simple loop. Hence either UU or U𝒞U^{\mathcal{C}} is bounded by Jordan’s theorem. In the former case, Proposition 2 implies σbU,V(P)=σb,V(P)=0\sigma_{b}^{U,V}(P)=\sigma_{b}^{\emptyset,V}(P)=0. In the latter case, Proposition 2 implies σbU,V(P)=σb2,V(P)=0\sigma_{b}^{U,V}(P)=\sigma_{b}^{\mathbb{R}^{2},V}(P)=0. Therefore we assume U{\partial}U is unbounded in the rest of the proof.

2. Since 𝒳U,V=0\mathcal{X}_{U,V}=0, 𝒳+(U,V)=𝒳(U,V)\mathcal{X}_{+}(U,V)=\mathcal{X}_{-}(U,V); potentially replacing VV by V𝒞V^{\mathcal{C}} we may assume that 𝒳+(U,V)=𝒳(U,V)=1=limt±𝟙Vγ(t)\mathcal{X}_{+}(U,V)=\mathcal{X}_{-}(U,V)=1=\lim\limits_{t\to\pm\infty}\mathds{1}_{V}\circ\gamma(t). Hence when TT is large enough, γ({|t|>T})V\gamma(\{|t|>T\})\subset V. Given R>0R>0, define

VR=def𝔻R(V)={𝒙2:dl(𝒙,V)<R}.V_{R}\ \mathrel{\stackrel{{\scriptstyle\makebox[0.0pt]{\mbox{\tiny def}}}}{{=}}}\ \mathbb{D}_{R}(V)=\big{\{}{\bm{x}}\in\mathbb{R}^{2}:\ d_{l}({\bm{x}},V)<R\big{\}}.

When R>max{dl(γ(t),V):t[T,T]}R>\max\big{\{}d_{l}(\gamma(t),V):t\in[-T,T]\big{\}}, we have UVR{\partial}U\subset V_{R}.

3. By continuity of distance function, dl(VR,V)Rd_{l}({\partial}V_{R},V)\leq R and dl(VR,V)Rd_{l}({\partial}V_{R},{\partial}V)\leq R. As a result,

dl(𝒙,V2R)dl(𝒙,V)dl(V,V2R)dl(𝒙,V)2R\begin{split}d_{l}({\bm{x}},{\partial}V_{2R})&\geq d_{l}({\bm{x}},{\partial}V)-d_{l}({\partial}V,{\partial}V_{2R})\geq d_{l}({\bm{x}},{\partial}V)-2R\end{split}

By (2.9),

2lnΨU,V2R(𝒙)\displaystyle 2\ln\Psi_{U,V_{2R}}({\bm{x}}) dl(𝒙,V2R)+dl(𝒙,U)\displaystyle\geq d_{l}({\bm{x}},{\partial}V_{2R})+d_{l}({\bm{x}},{\partial}U) (6.2)
dl(𝒙,V)+dl(𝒙,U)2RlnΨU,V(𝒙)2R.\displaystyle\geq d_{l}({\bm{x}},{\partial}V)+d_{l}({\bm{x}},{\partial}U)-2R\geq\ln{\Psi_{U,V}}({\bm{x}})-2R. (6.3)

On the other hand, if dl(VR,V)<Rd_{l}({\partial}V_{R},V)<R, then there is 𝒙VR{\bm{x}}\in{\partial}V_{R} such that dl(𝒙,V)<Rd_{l}({\bm{x}},V)<R, by definition this implies 𝒙VR{\bm{x}}\in V_{R}. But VRV_{R} is an open set so 𝒙VRVR={\bm{x}}\in V_{R}\cap{\partial}V_{R}=\emptyset. We get a contradiction; thus d(VR,V)=Rd({\partial}V_{R},V)=R. In particular, for any 𝒙V2R{\bm{x}}\in{\partial}V_{2R},

dl(𝒙,VR)dl(𝒙,VR)dl(V,VR)2RR=Rdl(VR,V2R)R.d_{l}({\bm{x}},V_{R})\geq d_{l}({\bm{x}},V_{R})-d_{l}(V,V_{R})\geq 2R-R=R\quad\Rightarrow\quad d_{l}(V_{R},{\partial}V_{2R})\geq R.

As a result, by (2.9),

2lnΨU,V2R(𝒙)dl(U,V2R)dl(VR,V2R)R.2\ln\Psi_{U,V_{2R}}({\bm{x}})\geq d_{l}({\partial}U,{\partial}V_{2R})\geq d_{l}(V_{R},{\partial}V_{2R})\geq R. (6.4)

Interpolating between (6.3) and (6.4) gives:

2lnΨU,V2R(𝒙)14(lnΨU,V(𝒙)2R)+34R=lnΨU,V(𝒙)+R4.2\ln\Psi_{U,V_{2R}}({\bm{x}})\geq\dfrac{1}{4}(\ln\Psi_{U,V}({\bm{x}})-2R)+\dfrac{3}{4}R=\dfrac{\ln\Psi_{U,V}({\bm{x}})+R}{4}. (6.5)

4. Since V2RΔV{𝒙:dl(𝒙,V)2R}V_{2R}\Delta V\subset\{{\bm{x}}:d_{l}({\bm{x}},{\partial}V)\leq 2R\}, by Proposition 2:

|σbU,V(P)|=|σbU,V2R(P)|.\big{|}\sigma_{b}^{U,V}(P)\big{|}=\big{|}\sigma_{b}^{U,V_{2R}}(P)\big{|}. (6.6)

By (3.5) and (6.5), we have

|σbU,V2R(P)|CN(𝒙2eNlnΨU,V2R(𝒙))2CNeNR4(𝒙2eN8lnΨU,V(𝒙))2.\big{|}\sigma_{b}^{U,V_{2R}}(P)\big{|}\leq C_{N}\left(\sum_{{\bm{x}}\in\mathbb{Z}^{2}}e^{-N\ln\Psi_{U,V_{2R}}({\bm{x}})}\right)^{2}\leq C_{N}e^{-\frac{NR}{4}}\left(\sum_{{\bm{x}}\in\mathbb{Z}^{2}}e^{-\frac{N}{8}\ln\Psi_{U,V}({\bm{x}})}\right)^{2}. (6.7)

Since U,VU,V are transverse sets, by (2.10), the last sum is finite when NN is large enough and the sum does not depend on RR. Therefore, taking RR\rightarrow\infty yields σbU,V(P)=0=𝒳U,Vσb(P)\sigma_{b}^{U,V}(P)=0=\mathcal{X}_{U,V}\cdot\sigma_{b}(P). This proves Theorem 3 when 𝒳U,V=0\mathcal{X}_{U,V}=0. ∎

Lemma 6.2.

Let PP be an ESR projector. Assume that for all transverse simple sets U,VU,V such that 𝒳U,V=1\mathcal{X}_{U,V}=1, we have σbU,V(P)=σb(P)\sigma_{b}^{U,V}(P)=\sigma_{b}(P). Then for all transverse simple sets U,VU,V such that 𝒳U,V=1\mathcal{X}_{U,V}=-1, we have σbU,V(P)=σb(P)\sigma_{b}^{U,V}(P)=-\sigma_{b}(P).

Proof.

Assume that 𝒳U,V=1\mathcal{X}_{U,V}=-1. Then, 𝒳U,Vc=1\mathcal{X}_{U,V^{c}}=1, and so by assumption we have σbU,Vc(P)=σb(P)\sigma_{b}^{U,V^{c}}(P)=\sigma_{b}(P). But since σbU,Vc(P)=σbU,V(P)\sigma_{b}^{U,V^{c}}(P)=-\sigma_{b}^{U,V}(P), we deduce that σbU,V(P)=σb(P)\sigma_{b}^{U,V}(P)=-\sigma_{b}(P).∎

Therefore, it remains to prove that Theorem 3 holds for pairs (U,V)(U,V) such that 𝒳U,V=1\mathcal{X}_{U,V}=1.

6.3. Uniformly transverse families

Recall that transversality condition (1.4) is equivalent to (2.8):

c(0,1)such that, |𝒙|c1,ΨU,V(𝒙)|𝒙|c.\exists c\in(0,1)\ \text{such that, }\ \forall|{\bm{x}}|\geq c^{-1},\ {\Psi_{U,V}}({\bm{x}})\geq|{\bm{x}}|^{c}. (6.8)

When (6.8) holds, we refer to (U,V)(U,V) as cc-transverse.

Definition 10 (Uniformly transversality).

Let ={(Un,Vn):n}\mathcal{F}=\big{\{}(U_{n},V_{n}):n\in\mathbb{N}\big{\}} be a family of transverse simple sets. We say that \mathcal{F} is uniformly transverse if there exists c(0,1)c\in(0,1) such that for all nn, (Un,Vn)(U_{n},V_{n}) is cc-transverse. We say that \mathcal{F} is equivalent to (U,V)(U,V) if for all nn, UΔUnU\Delta U_{n} and VΔVnV\Delta V_{n} are bounded.

Uniform transversality boils down to (6.8) holding uniformly in nn:

c(0,1) such that, n,|𝒙|c1,ΨUn,Vn(𝒙)|𝒙|c.\exists c\in(0,1)\text{ such that, }\ \forall n,\ \ \forall|{\bm{x}}|\geq c^{-1},\ \Psi_{U_{n},V_{n}}({\bm{x}})\geq|{\bm{x}}|^{c}. (6.9)

Recall that i:={xi0}{\mathbb{H}}_{i}:=\{x_{i}\geq 0\}, i=1,2i=1,2 denote the right/upper half-plane.

Our strategy to prove Theorem 3 goes as follows. We will construct a family of uniformly transverse sets (Un,Vn)(U_{n},V_{n}) (see Proposition 5 and §6.5) such that (Un,Vn)(U_{n},V_{n}) are compact perturbations of (U,V)(U,V) and they look like (2,1)({\mathbb{H}}_{2},{\mathbb{H}}_{1}) inside the disk 𝔻n(0)\mathbb{D}_{n}(0). Because (Un,Vn)(U_{n},V_{n}) are compact perturbations of (U,V)(U,V), we will have σbUn,Vn(P)=σbU,V(P)\sigma_{b}^{U_{n},V_{n}}(P)=\sigma_{b}^{U,V}(P) by the robustness of geometric bulk conductance (Proposition 3). Meanwhile, since (Un,Vn)(U_{n},V_{n}) look like 2{\mathbb{H}}_{2} and 1{\mathbb{H}}_{1} in 𝔻n(0)\mathbb{D}_{n}(0), we will show σbUn,Vn(P)\sigma_{b}^{U_{n},V_{n}}(P) converges to σb2,1(P)\sigma_{b}^{{\mathbb{H}}_{2},{\mathbb{H}}_{1}}(P) as nn\to\infty by showing

  • KUn,Vn(𝒙,𝒙)K2,1(𝒙,𝒙)K_{U_{n},V_{n}}({\bm{x}},{\bm{x}})\to K_{{\mathbb{H}}_{2},{\mathbb{H}}_{1}}({\bm{x}},{\bm{x}}) as nn\to\infty for each 𝒙{\bm{x}} (Lemma 6.3);

  • the convergence is uniform because of uniform transversality.

This will prove that σbUn,Vn(P)\sigma_{b}^{U_{n},V_{n}}(P) converges as nn\rightarrow\infty to σb2,1(P)\sigma_{b}^{{\mathbb{H}}_{2},{\mathbb{H}}_{1}}(P), and in particular σbU,V(P)=limnσbUn,Vn(P)=σb2,1(P)=σb(P)\sigma_{b}^{U,V}(P)=\lim\limits_{n\to\infty}\sigma_{b}^{U_{n},V_{n}}(P)=\sigma_{b}^{{\mathbb{H}}_{2},{\mathbb{H}}_{1}}(P)=\sigma_{b}(P).

Lemma 6.3.

Assume that {(Un,Vn):n}\{(U_{n},V_{n}):\ n\in\mathbb{N}\} is a family of transverse sets in 2\mathbb{R}^{2} such that for all nn, Un𝔻n(0)=2𝔻n(0)U_{n}\cap\mathbb{D}_{n}(0)={\mathbb{H}}_{2}\cap\mathbb{D}_{n}(0) and Vn𝔻n(0)=1𝔻n(0)V_{n}\cap\mathbb{D}_{n}(0)={\mathbb{H}}_{1}\cap\mathbb{D}_{n}(0). Then for any 𝐱2{\bm{x}}\in\mathbb{Z}^{2}:

limnKUn,Vn(𝒙,𝒙)=K2,1(𝒙,𝒙).\lim_{n\rightarrow\infty}K_{U_{n},V_{n}}({\bm{x}},{\bm{x}})=K_{{\mathbb{H}}_{2},{\mathbb{H}}_{1}}({\bm{x}},{\bm{x}}). (6.10)

Lemma 6.3 means that the geometric bulk conductance can be computed locally. This observation was key to the work [DZ23] on emergence of edge spectrum for truncated topological insulators.

Proof.

1. Recall that KA,B=P[[P,𝟙A],[P,𝟙B]]K_{A,B}=P\big{[}[P,\mathds{1}_{A}],[P,\mathds{1}_{B}]\big{]} (3.3) satisfies the kernel estimate (3.4):

|KA,B(𝒙,𝒚)|CNeN(dl(𝒙,A)+dl(𝒙,A)+dl(𝒚,B)+dl(𝒚,B)+dl(𝒙,𝒚)).|K_{A,B}({\bm{x}},{\bm{y}})|\leq C_{N}e^{-N(d_{l}({\bm{x}},{\partial}A)+d_{l}({\bm{x}},{\partial}A)+d_{l}({\bm{y}},{\partial}B)+d_{l}({\bm{y}},{\partial}B)+d_{l}({\bm{x}},{\bm{y}}))}. (6.11)

Since KA,BK_{A,B} is linear in 𝟙A\mathds{1}_{A} and 𝟙B\mathds{1}_{B}, we have

KA,BKA𝔻n(0),B𝔻n(0)=KA𝔻n(0)c,B+KA𝔻n(0)c,B𝔻n(0)c.K_{A,B}-K_{A\cap\mathbb{D}_{n}(0),B\cap\mathbb{D}_{n}(0)}=K_{A\cap\mathbb{D}_{n}(0)^{c},B}+K_{A\cap\mathbb{D}_{n}(0)^{c},B\cap\mathbb{D}_{n}(0)^{c}}.

Note that

dl(𝒙,(A𝔻n(0)c))dl(0,(A𝔻n(0)c)))dl(𝒙,0)ln(1+n)dl(𝒙,0).d_{l}({\bm{x}},{\partial}(A\cap\mathbb{D}_{n}(0)^{c}))\geq d_{l}(0,{\partial}(A\cap\mathbb{D}_{n}(0)^{c})))-d_{l}({\bm{x}},0)\geq\ln(1+n)-d_{l}({\bm{x}},0).

Hence we have

|KA𝔻n(0)c,B(𝒙,𝒙)|CNeNdl(𝒙,(A𝔻n(0)c))CNeNln(1+n)+Ndl(𝒙,0).\big{|}K_{A\cap\mathbb{D}_{n}(0)^{c},B}({\bm{x}},{\bm{x}})\big{|}\leq C_{N}e^{-Nd_{l}({\bm{x}},{\partial}(A\cap\mathbb{D}_{n}(0)^{c}))}\leq C_{N}e^{-N\ln(1+n)+Nd_{l}({\bm{x}},0)}. (6.12)

The same estimate holds for |KA𝔻n(0)c,B𝔻n(0)c(𝒙,𝒙)|\big{|}K_{A\cap\mathbb{D}_{n}(0)^{c},B\cap\mathbb{D}_{n}(0)^{c}}({\bm{x}},{\bm{x}})\big{|}. Hence we obtain

|KA,BKA𝔻n(0),B𝔻n(0)(𝒙,𝒙)|CNeNln(1+n)+Ndl(𝒙,0)=CN(1+n)NeNdl(𝒙,0).\big{|}K_{A,B}-K_{A\cap\mathbb{D}_{n}(0),B\cap\mathbb{D}_{n}(0)}({\bm{x}},{\bm{x}})\big{|}\leq C_{N}e^{-N\ln(1+n)+Nd_{l}({\bm{x}},0)}=C_{N}(1+n)^{-N}e^{-Nd_{l}({\bm{x}},0)}. (6.13)

2. We now apply (6.13) to the pair (Un,Vn)(U_{n},V_{n}), then to the pair (2,1)({\mathbb{H}}_{2},{\mathbb{H}}_{1}): {equations} K_U_n,V_n(x,x) - K_U_n ∩D_n(0),V_n ∩D_n(0)(x,x)≤C_N(1 + n)^-N e^Nd_l(x, 0), {equations} K_H_2, H_1(x,x) - K_H_2 ∩D_n(0),H_1 ∩D_n(0)(x,x) ≤C_N(1 + n)^-N e^Nd_l(x, 0) . Because 2𝔻n(0)=Un𝔻n(0){\mathbb{H}}_{2}\cap\mathbb{D}_{n}(0)=U_{n}\cap\mathbb{D}_{n}(0) and 1𝔻n(0)=Vn𝔻n(0){\mathbb{H}}_{1}\cap\mathbb{D}_{n}(0)=V_{n}\cap\mathbb{D}_{n}(0), summing these two bounds gives {equations} K_U_n ,V_n(x,x) - K_H_2, H_1(x,x) ≤2C_N(1 + n)^-N e^Nd_l(x, 0) . It suffices to take the limit as nn goes to \infty to conclude. ∎

Proposition 5.

Let (U,V)(U,V) be transverse simple sets in 2\mathbb{R}^{2} such that 𝒳U,V=1\mathcal{X}_{U,V}=1. There exists a family ={(Un,Vn):n}\mathcal{F}=\{(U_{n},V_{n}):\ n\in\mathbb{N}\} with the following properties:

  • (a)

    \mathcal{F} is uniformly transverse;

  • (b)

    \mathcal{F} is equivalent to (U,V)(U,V);

  • (c)

    In the disk 𝔻n(0)\mathbb{D}_{n}(0), UnU_{n} is the upper half-plane 2={x2>0}{\mathbb{H}}_{2}=\{x_{2}>0\} and VnV_{n} is the right half-plane 1={x1>0}{\mathbb{H}}_{1}=\{x_{1}>0\}.

Proposition 5 is the key construction in the proof of Theorem 3 and we defer its proof to §6.4 - §6.5.

Proof of Theorem 3 assuming Proposition 5.

By Proposition 5, there is a family ={(Un,Vn):n}\mathcal{F}=\{(U_{n},V_{n}):\ n\in\mathbb{N}\} satisfying (a), (b), (c) above. Since \mathcal{F} is equivalent to (U,V)(U,V) and UnΔUU_{n}\Delta U, VnΔVV_{n}\Delta V are bounded, by Proposition 2, we deduce that σbU,V(P)=σbUn,Vn(P)\sigma_{b}^{U,V}(P)=\sigma_{b}^{U_{n},V_{n}}(P) for all nn. In particular,

σbU,V(P)=limnσbUn,Vn(P)=limn𝒙2KUn,Vn(𝒙,𝒙).\sigma_{b}^{U,V}(P)=\lim_{n\rightarrow\infty}\sigma_{b}^{U_{n},V_{n}}(P)=\lim_{n\rightarrow\infty}\sum_{{\bm{x}}\in\mathbb{Z}^{2}}K_{U_{n},V_{n}}({\bm{x}},{\bm{x}}). (6.14)

Our plan is now to apply the dominated convergence theorem to the above series.

Define kn(𝒙)=KUn,Vn(𝒙)k_{n}({\bm{x}})=K_{U_{n},V_{n}}({\bm{x}}). By Lemma 6.3, for every 𝒙2{\bm{x}}\in\mathbb{Z}^{2}, kn(𝒙)k_{n}({\bm{x}}) converges to K2,1(𝒙,𝒙)K_{{\mathbb{H}}_{2},{\mathbb{H}}_{1}}({\bm{x}},{\bm{x}}) as nn\rightarrow\infty. Moreover, by (6.11) and uniformly-admissiblility, when |𝒙|c1|{\bm{x}}|\geq c^{-1},

|kn(𝒙)|=|KUn,Vn(𝒙,𝒙)|CNΨUn,Vn(𝒙)2NCN|𝒙|2Nc=:k(𝒙).\big{|}k_{n}({\bm{x}})\big{|}=\big{|}K_{U_{n},V_{n}}({\bm{x}},{\bm{x}})\big{|}\leq C_{N}\Psi_{U_{n},V_{n}}({\bm{x}})^{-2N}\leq C_{N}|{\bm{x}}|^{-2Nc}=:k({\bm{x}}). (6.15)

When NN is large enough, we have k(𝒙)1(2)k({\bm{x}})\in\ell^{1}(\mathbb{Z}^{2}). Thus we can apply dominated convergence theorem to get:

limn𝒙2KUn,Vn(𝒙,𝒙)\displaystyle\lim_{n\rightarrow\infty}\sum_{{\bm{x}}\in\mathbb{Z}^{2}}K_{U_{n},V_{n}}({\bm{x}},{\bm{x}}) =limn𝒙2kn(𝒙)\displaystyle=\lim_{n\rightarrow\infty}\sum_{{\bm{x}}\in\mathbb{Z}^{2}}k_{n}({\bm{x}}) (6.16)
=𝒙2limnkn(𝒙)=𝒙2K2,1(𝒙,𝒙)=σb2,1(P)=σb(P).\displaystyle=\sum_{{\bm{x}}\in\mathbb{Z}^{2}}\lim_{n\rightarrow\infty}k_{n}({\bm{x}})=\sum_{{\bm{x}}\in\mathbb{Z}^{2}}K_{{\mathbb{H}}_{2},{\mathbb{H}}_{1}}({\bm{x}},{\bm{x}})=\sigma_{b}^{{\mathbb{H}}_{2},{\mathbb{H}}_{1}}(P)=\sigma_{b}(P). (6.17)

Going back to (6.14) completes the proof of Theorem 3. ∎

6.4. Ordering entrance / exit points

In the following two subsections, we aim to prove Proposition 5. Let U,VU,V be transverse simple sets. For r>0r>0, J=U,VJ=U,V, define: {equations} t_J^+(r)  =def  sup{ t :  γ_J(t) ∈¯D_r(0) },  t_J^-(r)  =def  inf{ t :  γ_J(t) ∈¯D_r(0) },  z_J^±(r)  =def  γ_J ∘t_J^±(r).

Lemma 6.4.

Let U,VU,V be transverse simple sets such that 𝒳U,V=1\mathcal{X}_{U,V}=1. Let R>0R>0 such that UV𝔻R(0){\partial}U\cap{\partial}V\subset\mathbb{D}_{R}(0). For all r>Rr>R, there exists θ1<θ2<θ3<θ4<θ1+2π\theta_{1}<\theta_{2}<\theta_{3}<\theta_{4}<\theta_{1}+2\pi (see Figure 6), such that

𝒛V(r)=reiθ1,𝒛U(r)=reiθ2,𝒛V+(r)=reiθ3,𝒛U+(r)=reiθ4.{\bm{z}}_{V}^{-}(r)=re^{i\theta_{1}},\quad{\bm{z}}_{U}^{-}(r)=re^{i\theta_{2}},\quad{\bm{z}}_{V}^{+}(r)=re^{i\theta_{3}},\quad{\bm{z}}_{U}^{+}(r)=re^{i\theta_{4}}. (6.18)
\floatbox

[\capbeside\thisfloatsetupcapbesideposition=right,center,capbesidewidth=0.6]figure[\FBwidth] Refer to caption

Figure 6. The sets U,VU,V with the arguments θ1,θ2,θ3,θ4\theta_{1},\theta_{2},\theta_{3},\theta_{4}.
Proof.

1. Fix r>Rr>R. Define

ΓU+=γU((tU+(r),+)),ΓU=γU((,tU(r)))\Gamma_{U}^{+}=\gamma_{U}\big{(}(t_{U}^{+}(r),+\infty)\big{)},\qquad\Gamma_{U}^{-}=\gamma_{U}\big{(}(-\infty,t_{U}^{-}(r))\big{)} (6.19)

These are connected sets which do not intersect V{\partial}V. In particular they lies in VV or in V𝒞V^{\mathcal{C}}. But since 𝒳+(U,V)=1\mathcal{X}_{+}(U,V)=1, for tt sufficiently large γU(t)V\gamma_{U}(t)\in V. It follows that ΓU+V\Gamma_{U}^{+}\subset V. Likewise ΓUV𝒞\Gamma_{U}^{-}\subset V^{\mathcal{C}}.

2. Let θ1\theta_{1} such that 𝒛V(r)=reiθ1{\bm{z}}_{V}^{-}(r)=re^{i\theta_{1}}; let θ3(θ1,θ1+2π)\theta_{3}\in(\theta_{1},\theta_{1}+2\pi) such that 𝒛V+(r)=reiθ3{\bm{z}}_{V}^{+}(r)=re^{i\theta_{3}}. Let γW\gamma_{W} be the curve defined by:

γW=γV|ttV(r)reiθ|θ(θ1,2πθ3)γV|ttV+(r).\gamma_{W}=\gamma_{V}\big{|}_{t\leq t_{V}^{-}(r)}\oplus re^{-i\theta}\big{|}_{\theta\in(-\theta_{1},2\pi-\theta_{3})}\oplus\gamma_{V}\big{|}_{t\geq t_{V}^{+}(r)}. (6.20)

where we use the symbol \oplus to denote the concatenation of two curves. See Figure 7. Let WW the connected component of γW()\mathbb{C}\setminus\gamma_{W}(\mathbb{R}) to the left of γW\gamma_{W}. The function θeiθ\theta\mapsto e^{-i\theta} runs clockwise. Therefore (e.g. by considering a sufficiently small disk centered at reiθre^{i\theta}, θ(θ1,θ2)\theta\in(\theta_{1},\theta_{2}), simply split by γW\gamma_{W}) we see that 𝔻r(0)\mathbb{D}_{r}(0) lies in the component of ΓW\mathbb{C}\setminus\Gamma_{W} to the right of γW\gamma_{W}, in particular it does not intersect WW.

3. Note that ΓU+\Gamma_{U}^{+} is an unbounded connected subset of VV. Because VΔWV\Delta W is bounded (see Proposition 4) ΓU+\Gamma_{U}^{+} intersects WW. Moreover, it does not intersect W{\partial}W, so we have ΓU+W\Gamma_{U}^{+}\subset W. In particular,

𝒛U+(r)=limttU+(r)γU(t)W¯.{\bm{z}}_{U}^{+}(r)=\lim_{t\rightarrow t_{U}^{+}(r)}\gamma_{U}(t)\in\overline{W}. (6.21)

Because |𝒛U+(r)|=r|{\bm{z}}_{U}^{+}(r)|=r, we obtain

𝒛U+(r)W¯𝔻r(0)={reiθ:θ(θ1,2πθ3)}{\bm{z}}_{U}^{+}(r)\in\overline{W}\cap{\partial}\mathbb{D}_{r}(0)=\big{\{}re^{-i\theta}:\ \theta\in(-\theta_{1},2\pi-\theta_{3})\big{\}} (6.22)

Therefore, 𝒛U+(r)=reθ4{\bm{z}}_{U}^{+}(r)=re^{\theta_{4}} for some θ4(θ3,θ1+2π)\theta_{4}\in(\theta_{3},\theta_{1}+2\pi).

\floatbox

[\capbeside\thisfloatsetupcapbesideposition=right,center,capbesidewidth=0.6]figure[\FBwidth] Refer to caption

Figure 7. Construction of γW\gamma_{W} from γV\gamma_{V} and determination of θ4(θ3,θ1+2π)\theta_{4}\in(\theta_{3},\theta_{1}+2\pi).

4. By a similar argument, ΓU\Gamma_{U}^{-} lies in the component of ΓW\mathbb{C}\setminus\Gamma_{W} to the right of γW\gamma_{W}. Because ΓU\Gamma_{U}^{-} does not intersect 𝔻r(0)\mathbb{D}_{r}(0) and |𝒛U(r)|=r|{\bm{z}}_{U}^{-}(r)|=r, we deduce that as in (6.22) that

𝒛U(r){reiθ:θ(θ1,θ3)},{\bm{z}}_{U}^{-}(r)\in\big{\{}re^{i\theta}:\ \theta\in(\theta_{1},\theta_{3})\big{\}}, (6.23)

which implies 𝒛U(r)=reθ2{\bm{z}}_{U}^{-}(r)=re^{\theta_{2}} for some θ3(θ1,θ3)\theta_{3}\in(\theta_{1},\theta_{3}). This completes the proof. ∎

6.5. Construction of uniformly transverse family of sets

Let (U,V)(U,V) be a cc-transverse pair with 𝒳U,V=1\mathcal{X}_{U,V}=1. Now we can construct a family of uniformly transverse sets (Un,Vn)(U_{n},V_{n}) that is equivalent to (U,V)(U,V) such that Un𝔻n(0)=1𝔻n(0)U_{n}\cap\mathbb{D}_{n}(0)={\mathbb{H}}_{1}\cap\mathbb{D}_{n}(0), Vn𝔻n(0)=2𝔻n(0)V_{n}\cap\mathbb{D}_{n}(0)={\mathbb{H}}_{2}\cap\mathbb{D}_{n}(0).

By Lemma 6.4, when nn is large enough, for some θ1<θ2<θ3<θ4<θ1+2π\theta_{1}<\theta_{2}<\theta_{3}<\theta_{4}<\theta_{1}+2\pi, we have

𝒛V(4n)=4neiθ1,𝒛U(4n)=4neiθ2,𝒛V+(4n)=4neiθ3,𝒛U+(4n)=4neiθ4.{\bm{z}}_{V}^{-}(4n)=4ne^{i\theta_{1}},\quad{\bm{z}}_{U}^{-}(4n)=4ne^{i\theta_{2}},\quad{\bm{z}}_{V}^{+}(4n)=4ne^{i\theta_{3}},\quad{\bm{z}}_{U}^{+}(4n)=4ne^{i\theta_{4}}. (6.24)

For such an nn, we define four functions αk:[0,4]\alpha_{k}:[0,4]\rightarrow\mathbb{R}, k{1,2,3,4}k\in\{1,2,3,4\} by

αk(s)={kπ2,s[0,2](3s)kπ2+(s2)θk,s[2,3]θk,s[3,4]\alpha_{k}(s)=\left\{\begin{matrix}\frac{k\pi}{2},&s\in[0,2]\\ (3-s)\frac{k\pi}{2}+(s-2)\theta_{k},&s\in[2,3]\\ \theta_{k},&s\in[3,4]\end{matrix}\right. (6.25)

and four curves 𝒛k:[0,4]{\bm{z}}_{k}:[0,4]\rightarrow\mathbb{C} (see Figure 8) by

𝒛k(s)=nseiαk(s).{\bm{z}}_{k}(s)=ns\cdot e^{i\alpha_{k}(s)}. (6.26)
Refer to caption
Figure 8. The curves 𝒛1,𝒛2,𝒛3,𝒛4{\bm{z}}_{1},{\bm{z}}_{2},{\bm{z}}_{3},{\bm{z}}_{4}.

We will make use of the following result:

Lemma 6.5.

Let c[0,1],nc\in[0,1],n\in\mathbb{N}. Assume that t,s[0,4]t,s\in[0,4] are such that |𝐳1(t)||𝐳2(s)|<27nc|{\bm{z}}_{1}(t)|-|{\bm{z}}_{2}(s)|<2^{-7}n^{c}. Then

2n|sin(α1(t)α2(s)2)|24nc.2n\left|\sin\left(\frac{\alpha_{1}(t)-\alpha_{2}(s)}{2}\right)\right|\geq 2^{-4}n^{c}. (6.27)
Proof.

We first estimate α1(4)α2(4)=θ1θ2\alpha_{1}(4)-\alpha_{2}(4)=\theta_{1}-\theta_{2}. Without loss of generality, we can assume |θ1θ2|<π|\theta_{1}-\theta_{2}|<\pi by choosing θ2(θ1π,θ1+π)\theta_{2}\in(\theta_{1}-\pi,\theta_{1}+\pi). Recall that 𝒛1(4)=𝒛V(4n)V{\bm{z}}_{1}(4)={\bm{z}}_{V}^{-}(4n)\in{\partial}V by definition. By transversality, we see

(4n)cΨU,V(𝒛1(4))=d(𝒛1(4),U)|𝒛1(4)𝒛2(4)|=4n|eiθ1eiθ2|4n|θ1θ2|.\begin{split}(4n)^{c}&\leq{\Psi_{U,V}}\big{(}{\bm{z}}_{1}(4)\big{)}=d\big{(}{\bm{z}}_{1}(4),{\partial}U\big{)}\leq\big{|}{\bm{z}}_{1}(4)-{\bm{z}}_{2}(4)\big{|}\\ &=4n\big{|}e^{i\theta_{1}}-e^{i\theta_{2}}\big{|}\leq 4n|\theta_{1}-\theta_{2}|.\end{split}

Since c<1c<1, for any n1n\geq 1, (4n)c11π2(4n)^{c-1}\leq 1\leq\frac{\pi}{2}. Recall that by definition, α1(t)α2(t)\alpha_{1}(t)-\alpha_{2}(t) is monotone over [0,4][0,4]. Hence we obtain

|α1(t)α2(t)|min{|α1(4)α2(4)|,|α1(0)α2(0)|}min{|θ1θ2|,π2}min{(4n)c1,π2}(4n)c114nc1.\begin{split}|\alpha_{1}(t)-\alpha_{2}(t)|&\geq\min\big{\{}|\alpha_{1}(4)-\alpha_{2}(4)|,|\alpha_{1}(0)-\alpha_{2}(0)|\big{\}}\geq\min\big{\{}|\theta_{1}-\theta_{2}|,\frac{\pi}{2}\big{\}}\\ &\geq\min\left\{(4n)^{c-1},\frac{\pi}{2}\right\}\geq(4n)^{c-1}\geq\frac{1}{4}n^{c-1}.\end{split} (6.28)

Meanwhile, by the definition of αk(s)\alpha_{k}(s) and the assumption |𝒛1(t)||𝒛2(s)|=ntns27nc|{\bm{z}}_{1}(t)|-|{\bm{z}}_{2}(s)|=nt-ns\leq 2^{-7}n^{c}, we have

|α2(t)α2(s)||ts|maxr[0,4]|α2(r)||ts|4π23nc1.|\alpha_{2}(t)-\alpha_{2}(s)|\leq|t-s|\cdot\max\limits_{r\in[0,4]}|\alpha_{2}^{\prime}(r)|\leq|t-s|\cdot 4\pi\leq 2^{-3}n^{c-1}. (6.29)

Combining (6.28) and (6.29), we obtain

|α1(t)α2(s)||α1(t)α2(t)||α2(t)α2(s)|14nc118nc1=18nc1.\begin{split}|\alpha_{1}(t)-\alpha_{2}(s)|&\geq|\alpha_{1}(t)-\alpha_{2}(t)|-|\alpha_{2}(t)-\alpha_{2}(s)|\\ &\geq\frac{1}{4}n^{c-1}-\frac{1}{8}n^{c-1}=\frac{1}{8}n^{c-1}.\end{split}

Since |α1(t)α2(s)2|=|θ1θ22|π2\left|\frac{\alpha_{1}(t)-\alpha_{2}(s)}{2}\right|=\left|\frac{\theta_{1}-\theta_{2}}{2}\right|\leq\frac{\pi}{2} and when |α|π2|\alpha|\leq\frac{\pi}{2}, |sinα||r|/2|\sin\alpha|\geq|r|/2, we see that

2n|sin(α1(t)α2(s)2)|2n|α1(t)α2(s)|424nc2n\left|\sin\left(\frac{\alpha_{1}(t)-\alpha_{2}(s)}{2}\right)\right|\geq\frac{2n|\alpha_{1}(t)-\alpha_{2}(s)|}{4}\geq 2^{-4}n^{c}

for any n1n\geq 1. This proves Lemma 6.5. ∎

We now deform U,VU,V by using 𝒛k{\bm{z}}_{k} as boundary functions. Specifically, we define two sets Un,VnU_{n},V_{n} as the set lying to the left of the following boundaries (see Figure LABEL:fig:P12) – recall that αk\alpha_{k} and 𝒛k{\bm{z}}_{k} depends on nn, see (6.26): {equations} ∂U_n = γ_U((-∞,t_U^-(4n)))z_2([0,4])z_4([0,4])γ_U((t_U^+(4n), +∞)),
∂V_n = γ_V((-∞,t_V^-(4n)))z_1([0,4]) ∪z