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

The landscape law for tight binding Hamiltonians

Douglas Arnold, Marcel Filoche, Svitlana Mayboroda, Wei Wang, and Shiwen Zhang
Abstract.

The present paper extends the landscape theory pioneered in [FM, ADFJM2, DFM] to the tight-binding Schrödinger operator on d{\mathbb{Z}}^{d}. In particular, we establish upper and lower bounds for the integrated density of states in terms of the counting function based upon the localization landscape.

1. Introduction and main results

In this paper, we consider the discrete tight-binding Schrödinger operator H=Δ+VH=-\Delta+V on d{\mathbb{Z}}^{d}. The traditional approaches to the estimates for the integrated density of states of its continuous analogue in d\mathbb{R}^{d} can roughly be split into two groups, both pertaining to the asymptotic regimes. The first one are the results akin to the Weyl law, estimating the asymptotics of the spectrum as the eigenvalue μ+\mu\to+\infty in terms of the volume in the phase space of the set {(ξ,x):|ξ|2+V(x)μ}\bigl{\{}(\xi,x):\,|\xi|^{2}+V(x)\leq\mu\bigr{\}}, or in terms of the associated counting function according to the Fefferman-Phong uncertainty principle [Fe]. Informally speaking, letting μ+\mu\to+\infty corresponds to considering length scales which tend to zero, and hence such estimates, by design, are not relevant for a tight-binding model on d{\mathbb{Z}}^{d}. And indeed, deterministic potentials on d{\mathbb{Z}}^{d} were typically treated by more ad hoc approaches specific to their structure: periodicity, symmetries, etc, see, e.g., [DLY, HJ]. The second type of results pertains to the case when the potential is random. Then the integrated density of states exhibits the so-called Lifschitz tails, an exponential asymptotic behavior at the edge of spectrum. This regime is rather well-understood in both d{\mathbb{Z}}^{d} and d\mathbb{R}^{d}, but is in essence probabilistic, restricted to disordered potentials and insensitive to their individual features exhibited, for instance, on finite sets.

The present paper introduces another approach. It takes advantage of the landscape function uu from [FM, DFM] to build a box-counting somewhat analogous to the Weyl law and the Fefferman-Phong uncertainty principle, but associated to a different potential, the reciprocal of the landscape 1/u1/u. The use of the landscape in place of the original potential VV allows one to work in a non-asymptotic regime, contrary to aforementioned results, and in some sense to bring the ideas behind the original Weyl law to the lattice without restrictions on the potential or the pertinent eigenvalues, for both deterministic and random scenarios. In practice this approach provides a “black box”, in which the landscape, evaluated directly from the original Hamiltonian, yields an accurate approximation for the integrated density of states without any adjustable parameters, for deterministic and random potentials alike – see the numerical experiments in [FM, ADFJM1, ADFJM3]. The present paper addresses the estimates from above and below and makes the first step towards the mathematically rigorous understanding of the precision of the landscape predictions in the aforementioned works.

In order to describe our main results, we introduce some notations. Let Λ=(/K)d{1¯,,K¯}d\Lambda=({\mathbb{Z}}/K{\mathbb{Z}})^{d}\cong\{\bar{1},\cdots,\bar{K}\}^{d} be an integer torus, where KK\in\mathbb{N}, K3K\geq 3, and k¯\bar{k}, 1kK,1\leq k\leq K, is the congruence class, modulo KK. For simplicity, we will omit the bar from k¯\bar{k} when it is clear. Let V={vn}nΛ(Λ)V=\{v_{n}\}_{n\in{\Lambda}}\in\ell^{\infty}(\Lambda) be a real-valued, non-constant, non-negative potential. We denote by Vmax=maxnΛvnV_{\max}=\max_{n\in{\Lambda}}v_{n} the amplitude of the potential. The tight binding Hamiltonian HH is the linear operator on :=2(Λ)Kd\mathcal{H}:=\ell^{2}({\Lambda})\cong\mathbb{R}^{K^{d}} defined by

(1.1) (Hφ)n=|mn|1=1(φmφn)+vnφn,nΛ,(H\varphi)_{n}=-\sum_{|m-n|_{1}=1}\left(\varphi_{m}-\varphi_{n}\right)\,+v_{n}\varphi_{n},\ \ n\in{\Lambda},

where |n|1:=i=1d|ni||n|_{1}:=\sum_{i=1}^{d}|n_{i}| is the 11-norm on Λ\Lambda. We may think of φ\varphi either as a periodic sequence φn\varphi_{n} indexed by ndn\in{\mathbb{Z}}^{d} or as a periodic function φ(n)\varphi(n) on d{\mathbb{Z}}^{d}. We are interested in the normalized integrated density of states of HH, i.e., the eigenvalue counting function per unit volume:

(1.2) N(μ):=Kd×{the number of eigenvalues λofHsuchthatλμ}.N(\mu):=K^{-d}\,\times\left\{\,\textrm{the number of eigenvalues\ }\lambda\ {\rm of}\ H\ {\rm such\ that}\ \lambda\leq\mu\right\}.

In 2012, a new concept called the localization landscape was introduced in [FM]. Given an operator HH as above, the discrete localization landscape function is the unique solution u={un}nΛu=\{u_{n}\}_{n\in\Lambda}\in\mathcal{H} to the equation (Hu)n=1(Hu)_{n}=1. When V=0V=0 and uu vanishes on the boundary, the landscape is simply the torsion function of the Dirichlet Laplacian.

Next, let us define the landscape box counting function NuN_{u}. We start by defining, for any positive integer ss, a partition 𝒫(s){\mathcal{P}}(s) of the set {1,,K}d\{1,\cdots,K\}^{d} into subsets which are mostly boxes of side length ss, as follows. Writing K=qs+rK=qs+r (where the quotient qq and the remainder rr are non-negative integers and r<sr<s), we define a partition 𝒫1(s){\mathcal{P}}_{1}(s) of the set {1,,K}\{1,\cdots,K\} into qq subsets of ss consecutive elements, and, if r>0r>0, one additional subset of cardinality rr. The partition 𝒫(s){\mathcal{P}}(s) then consists of the boxes defined by the Cartesian products of dd subsets from 𝒫1(s){\mathcal{P}}_{1}(s), see Figure 1.

Refer to caption
Figure 1. The partition 𝒫(2){\mathcal{P}}(2) for Λ={1,,7}2\Lambda=\{1,\cdots,7\}^{2}.

For given μ>0\mu>0, we then set s(μ)=μ1/2s(\mu)=\left\lceil\mu^{-1/2}\right\rceil, and define Nu(μ)N_{u}(\mu) as the number of boxes on which the minimum of 1/un1/u_{n} does not exceed μ\mu, normalized by the size of the set Λ\Lambda:

(1.3) Nu(μ)=Kd×{the number of Q𝒫(s(μ))suchthatminnQ1unμ}.N_{u}(\mu)=K^{-d}\times\,\left\{\textrm{the number of \ }Q\in{\mathcal{P}}\big{(}s(\mu)\big{)}\ {\rm such\ that}\ \min_{n\in Q}\frac{1}{u_{n}}\leq\mu\right\}.

Our goal is to estimate the integrated density of states NN in terms of the landscape box counting function NuN_{u}. To this end, we will establish several estimates, stated below as Theorems 1, 2, and 3, which we collectively refer to as the Landscape Law.

The first result, which will be proven in Section 3.1, shows that, after a proper scaling, NuN_{u} provides an upper bound for N(μ)N(\mu) over the whole range of μ\mu.

Theorem 1.

Let V(Λ)V\in\ell^{\infty}(\Lambda) be a non-constant, non-negative potential. Then there is a dimensional constant C1>0C_{1}>0, such that

(1.4) N(μ)Nu(C1μ)for all μ>0.N\left(\mu\right)\leq N_{u}(C_{1}\mu)\quad\mbox{for all $\mu>0$.}

In saying that C1C_{1} is a dimensional constant, we mean that it depends only on dd, and, in particular, is independent of KK and VV. In fact, we can take C1=4dC_{1}=4d in (1.4).

The next theorem, proved in Section 3.2, contains the key estimate for obtaining a lower bound for N(μ)N(\mu).

Theorem 2.

Retain the hypotheses in Theorem 1. Then there are dimensional constants c,c0,c1,C0,α0,c0,c1,C0c_{\ast},c_{0},c_{1},C_{0},\alpha_{0},c^{\prime}_{0},c^{\prime}_{1},C^{\prime}_{0} (in particular, independent of KK and VV), such that

(1.5) N(μ)c0αdNu(c1αd+2μ)C0Nu(c1αd+4μ)for all 0<μcα4 and 0<α<α0,N(\mu)\geq c_{0}{\alpha^{d}}N_{u}(c_{1}\alpha^{d+2}\mu)-C_{0}N_{u}(c_{1}\alpha^{d+4}\mu)\quad\mbox{for all $0<\mu\leq c_{\ast}\alpha^{-4}$ and $0<\alpha<\alpha_{0}$,}

and

(1.6) N(μ)c0Nu(c1α2μ)C0Nu(c1α4μ)for all μ>cα4 and 0<α<α0.N(\mu)\geq c_{0}^{\prime}N_{u}(c^{\prime}_{1}\alpha^{2}\mu)-C_{0}^{\prime}N_{u}(c^{\prime}_{1}\alpha^{4}\mu)\quad\mbox{for all $\mu>c_{\ast}\alpha^{-4}$ and $0<\alpha<\alpha_{0}$.}

In order to get a positive lower bound on N(μ)N(\mu), we will remove the negative correction terms on the right-hand side of (1.5) and (1.6) through several complementary mechanisms. We roughly divide potentials into two regimes, corresponding to potentials satisfying a certain scaling condition and to certain disordered potentials. Note that these regimes could overlap and do not between them cover all potentials.

1.1. Deterministic potentials subject to the doubling scaling estimates

Theorem 3.

Retain the hypotheses of Theorem 1 and assume that uu satisfies the scaling condition

(1.7) n3Qun2CS(nQun2+d+4)\sum_{n\in 3Q}u_{n}^{2}\leq C_{S}\left(\sum_{n\in Q}u_{n}^{2}+\ell^{d+4}\right)

for every cube QΛQ\subset\Lambda of side length \ell. Here, 3Q3Q is the tripled cube concentric with QQ (see the definition in (2.10)). Then there exist positive constants C1,c3C_{1},\,c_{3} depending on dimension only and c2c_{2} depending on dd, VmaxV_{\max} and CSC_{S} such that

(1.8) c3Nu(c2μ)N(μ)Nu(C1μ)for all μ>0.c_{3}N_{u}(c_{2}\,\mu)\leq N(\mu)\leq N_{u}(C_{1}\mu)\quad\mbox{for all $\mu>0$}.

Assumption (1.7) is analogous to doubling hypotheses which are commonly used in the continuous case for elliptic PDEs. Such estimates are standard consequences of the Harnack and De Giorgi–Nash–Moser arguments which hold for homogeneous equations and for the Schrödinger equation with relatively slowly varying potentials, for instance, within the Kato class. See the discussion in [DFM, Ku, HL].

The scaling condition (1.7) also holds whenever VV is periodic. Indeed, suppose that {vn}\{v_{n}\} is periodic in each of the dd coordinate directions with period vector p=(p1,,pd)d\vec{p}=(p_{1},\cdots,p_{d})\in\mathbb{N}^{d} (see, e.g., [DLY, Ea, HJ, RS]). Assume that KK is divisible by each pip_{i}. Then, as we show in Section 3.4, the scaling condition (1.7) is satisfied with CSC_{S} depending on d,Vmaxd,V_{\max}, and p\vec{p}, but not on KK, which yields

Corollary 1.

Let H=Δ+VH=-\Delta+V be as in (1.1), with a non-trivial periodic potential V={vn}nΛV=\{v_{n}\}_{n\in\Lambda} as above. Then

(1.9) c3Nu(c2μ)N(μ)Nu(C1μ)for all μ>0,c_{3}\,N_{u}(c_{2}\,\mu)\,\leq\,N(\mu)\,\leq\,N_{u}(C_{1}\,\mu)\quad\mbox{for all $\mu>0$},

where C1,c3C_{1},c_{3} are dimensional constants and c2c_{2} depends on d,Vmaxd,V_{\max}, and p\vec{p} only.

Perhaps the major example when (1.7) might fail is that of disordered systems. Indeed, if on a 1-dimensional lattice we could have an arbitrarily long region of V=0V=0 followed by an arbitrarily long region of V=1V=1, that would correspond to a region of Δu=1-\Delta u=1 followed by a region of Δu+u=1-\Delta u+u=1. In the first case uu is quadratic and in the second exponential, which clearly destroys the “doubling” required by (1.7). Fortunately, there is a complementary mechanism to obtain an improved estimate akin to (1.8) from (1.5).

To illustrate it, suppose that for μ\mu belonging to some interval on the positive half-line, we have the bounds

aμβNu(μ)bμβ,a\,\mu^{\beta}\,\leq\,N_{u}(\mu)\,\leq\,b\,\mu^{\beta},

where the power β>d/2\beta>d/2. Substituting these bounds into (1.5) and choosing α\alpha sufficiently small, it is easy to deduce the lower bound (1.8) on N(μ)N(\mu) for μ\mu in the same interval. A similar argument can be used to obtain a lower bound for N(μ)N(\mu), or, more precisely, for the expectation of N(μ)N(\mu), in the case of Anderson potentials or any disordered potential near a fluctuation boundary. This is basically due to the fact that the aforementioned exponential nature of the Lifshitz tails “beats” the negative polynomial correction in (1.5).

1.2. Disordered potentials

Assume that the values {vn}nΛ\{v_{n}\}_{n\in\Lambda} are given by independent, identically distributed (i.i.d.) random variables, with common probability measure P0P_{0} on \mathbb{R}. Denote by F(δ)=P0(vnδ)F(\delta)=P_{0}(v_{n}\leq\delta) the common cumulative distribution function of vnv_{n} and by

suppP0={μ:P0(vn(με,μ+ε))>0,ε>0}{\rm supp}\,P_{0}=\Big{\{}\,\mu\in\mathbb{R}:\,P_{0}\bigl{(}v_{n}\in(\mu-\varepsilon,\mu+\varepsilon)\bigr{)}>0,\,\forall\,\varepsilon>0\,\Big{\}}

the support of the measure P0P_{0}. We assume that infsuppP0=0\inf{\rm supp}\,P_{0}=0 and supsuppP0=Vmax>0\sup{\rm supp}\,P_{0}=V_{\max}>0. We denote by 𝔼()\mathbb{E}(\cdot) the expectation with respect to the product measure on |Λ|\mathbb{R}^{|\Lambda|} generated by P0P_{0}.

Theorem 4.

Let V={vn}nΛV=\{v_{n}\}_{n\in\Lambda} be an Anderson-type potential as above. Let C1C_{1} be as in Theorem 1. Then there are constants c5,c6>0c_{5},c_{6}>0 depending on dd, the expectation of the random variable, and VmaxV_{\max}, such that

(1.10) c5𝔼Nu(c6μ)𝔼N(μ)𝔼Nu(C1μ)for allμ>0.c_{5}\,\mathbb{E}N_{u}(c_{6}\,\mu)\leq\mathbb{E}N(\mu)\leq\mathbb{E}N_{u}(C_{1}\,\mu)\ \ \ {\textrm{for all}}\ \ \mu>0.

Furthermore, there is a constant μ>0\mu_{\ast}>0 depending on dd and expectation of the random variable, such that if, in addition, μμ\mu\leq\mu_{\ast}, then (1.10) holds with the constants c5,c6c_{5},c_{6} independent of VmaxV_{\max}.

We note that in the course of the proof of Theorem 4 we prove the following universal bound on Lifschitz tails in terms of the cumulative distribution function FF.

Proposition 1.

Retain the setting of Theorem 4. Then there are constants μ0,K,ci\mu_{0},K_{\ast},c_{i}, depending only on the dimension and the expectation of the random variable (but independent of KK), such that

(1.11) c1μd/2F(c2μ)c3μd/2𝔼N(μ)c4μd/2F(c5μ)c6μd/2forallμ(K/K2,μ0).c_{1}\mu^{d/2}F(c_{2}\mu)^{c_{3}\mu^{-d/2}}\leq\mathbb{E}N(\mu)\leq c_{4}\,\mu^{d/2}\,F(c_{5}\,\mu)^{\,c_{6}\,\mu^{-d/2}}\ \ {\rm for\ all\ }\mu\in(K_{\ast}/K^{2},\mu_{0}).

To the best of our knowledge, this statement has never been formulated in this generality, even though perhaps it would not surprise the experts. The more traditional, weaker double log asymptotics are now considered classical (see [Li, KM2, Si, Ki]) and for certain classes of random potentials they have been improved in [BiKo, Ko, KM1] and other works. Here, Proposition 1 does not carry any a priori assumptions on the underlying probability distribution, and is a by-product of the landscape method.

1.3. The dual landscape and computational examples

Contrary to the continuous case, the spectrum of the discrete Schrödinger operator H=Δ+VH=-\Delta+V is a compact subset in [0,4d+Vmax][0,4d+V_{\max}]. The eigenvalue counting near the top of the spectrum for μ~\widetilde{\mu} close to 4d+Vmax4d+V_{\max} can be converted into the counting near the bottom of the spectrum for μ=4d+Vmaxμ~\mu=4d+V_{\max}-\widetilde{\mu} close to 0. Such a conversion is obtained via a dual model H~=Δ+VmaxV\widetilde{H}=-\Delta+V_{\max}-V, see [LMF, WZ]. One defines the dual landscape function u~\widetilde{u} as the solution to (H~u~)n=1(\widetilde{H}\,\widetilde{u})_{n}=1 and the box-counting function Nu~N_{\widetilde{u}} using (1.3), leading to

Corollary 2.

Retain the definitions in Theorem 4. Let C1C_{1} be as in Theorem 1. Suppose K3K\geq 3 is even. There are positive constants c~5,c~6\widetilde{c}_{5},\widetilde{c}_{6} depending on dd, the expectation of the random variable, and VmaxV_{\max}, such that

(1.12) 1𝔼Nu~(C1μ~)𝔼N(μ)1c~5𝔼Nu~(c~6μ~)for allμ<4d+Vmax,1-\mathbb{E}N_{\widetilde{u}}(C_{1}\,\widetilde{\mu})\leq\mathbb{E}N(\mu)\leq 1-\widetilde{c}_{5}\mathbb{E}N_{\widetilde{u}}(\widetilde{c}_{6}\,\widetilde{\mu})\ \ {\textrm{for all}}\ \mu<4d+V_{\max},

where μ~=4d+Vmaxμ\widetilde{\mu}=4d+V_{\max}-\mu and u~\widetilde{u} is the landscape function for Δ+VmaxV-\Delta+V_{\max}-V.

If one carefully tracks the values of the constants in (1.10) and (1.12) obtained in the proofs, they are of course far from optimal. However, the formulas emphasize the correct features of the spectrum and, as we have mentioned above, the numerical experiments actually yield even more satisfactory results than the formal estimates seem to warrant. In [DM+] and accompanying numerical work still in preparation, we show that there are very stable constants c1,c2c_{1},c_{2} such that a practical landscape law holds: N(μ)c1Nu(c2μ)N(\mu)\approx c_{1}N_{u}(c_{2}\mu), see Figure 2.

Refer to caption
Refer to caption
Figure 2. One-dimensional discrete Schrödinger operator on a periodic lattice /300{\mathbb{Z}}/300{\mathbb{Z}}, with a random potential V={vn}V=\{v_{n}\} uniformly distributed in [0,10][0,10]. Comparison between the true eigenvalue counting function NN, the scaled landscape box-counting function NuN_{u}, and the dual landscape box-counting function Nu~N_{\widetilde{u}}. The first plot shows the whole spectrum, while the second zooms in on the bottom spectrum.

We would like to point out that the landscape counting function, being a deterministic rather than a probabilistic tool, even picks up a spectral gap around the energy μ2.5\mu\approx 2.5 – a feature which would not be feasible, for instance, via the Lifschitz tail estimates. These details would of course disappear in the limit of an infinite domain but they demonstrate a surprising precision of the Landscape Law compared to any other currently available method.

The rest of the paper is organized as follows. We state preliminaries for tight-binding Hamiltonians and the discrete landscape theory in Section 2. In Section 3, we study deterministic potentials and prove Theorem 1, 2, 3, and Corollary 1. In Section 4, we concentrate on the Anderson model. We first prove the Lifshitz tail estimates for NuN_{u} and finally conclude Theorem 4 in Section 4.2. Section 4.3 is a discussion of the dual landscape theory. In the Appendix, we include some technical estimates for discrete harmonic functions and a well known probability result called Chernoff-Hoeffding bound. The key properties of the landscape function strongly rely on the foundations of the theory of elliptic PDEs. Many of these results require different techniques on d{\mathbb{Z}}^{d} compared to their continuous analogues. For instance, because of the lack of rotational symmetry and dilational invariance, many estimates for the Poisson kernel and the Green’s function are not known on a lattice, and are technically difficult to prove. A substantial portion of the paper is devoted to the discrete analogues of these elliptic estimates, and we hope they will be of independent interest.

Acknowledgments. Arnold is supported by the NSF grant DMS-1719694 and Simons Foundation grant 601937, DNA. Filoche is supported by Simons Foundation grant 601944, MF. Mayboroda is supported by NSF DMS 1839077 and the Simons Collaborations in MPS 563916, SM. Wang is supported by Simons Foundation grant 601937, DNA. Zhang is supported in part by the NSF grants DMS1344235, DMS-1839077, and Simons Foundation grant 563916, SM.

2. Preliminaries

In the tight-binding model, the Hilbert space is taken as the space of sequences 2(d)={{ϕi}id|id|ϕi|2<}\ell^{2}({\mathbb{Z}}^{d})=\left\{\{\phi_{i}\}_{i\in{\mathbb{Z}}^{d}}\,|\,\sum_{i\in{\mathbb{Z}}^{d}}|\phi_{i}|^{2}<\infty\right\} where we may think of ϕ={ϕi}id\phi=\{\phi_{i}\}_{i\in{\mathbb{Z}}^{d}} either as a function ϕ=ϕ(i)\phi=\phi(i) on d{\mathbb{Z}}^{d} or as a sequence {ϕi}\{\phi_{i}\} indexed by idi\in{\mathbb{Z}}^{d}. The d{\mathbb{Z}}^{d} lattice is equipped with the 11-norm:

(2.1) |n|1:=i=1d|ni|,|n|_{1}:=\sum_{i=1}^{d}|n_{i}|,

which reflects the graph structure of d{\mathbb{Z}}^{d}. We will also frequently need the infinity (maximum) norm

(2.2) |n|:=max1id|ni|.|n|_{\infty}:=\max_{1\leq i\leq d}|n_{i}|.

Two vertices m=(m1,,md),n=(n1,,nd)dm=(m_{1},\cdots,m_{d}),n=(n_{1},\cdots,n_{d})\in{\mathbb{Z}}^{d} are called the nearest neighbors if |mn|1=1|m-n|_{1}=1. We also say that nearest neighbors m,nm,n are connected by an edge of the discrete graph d{\mathbb{Z}}^{d}. We denote by ei=(0,,0,1,0,,0),i=1,,d,e_{i}=(0,\cdots,0,1,0,\cdots,0),\,i=1,\cdots,d, the elements of the canonical basis of d{\mathbb{Z}}^{d}. For ϕ={ϕn}ndd\phi=\{\phi_{n}\}_{n\in{\mathbb{Z}}^{d}}\in\mathbb{R}^{{\mathbb{Z}}^{d}}, its ii-th directional (forward) difference iϕ:d\nabla_{i}\phi:\,{\mathbb{Z}}^{d}\to\mathbb{R} is defined as

(2.3) iϕn=ϕn+eiϕn,i=1,,d,\nabla_{i}\phi_{n}=\phi_{n+e_{i}}-\phi_{n},\ \ i=1,\cdots,d,

and its gradient ϕ:dd\nabla\phi:{\mathbb{Z}}^{d}\to\mathbb{R}^{d} is

ϕn=(1ϕn,2ϕn,,dϕn).\nabla\phi_{n}=\left(\nabla_{1}\phi_{n},\nabla_{2}\phi_{n},\cdots,\nabla_{d}\phi_{n}\right).

We also denote the dot product and the induced norm of the resulting vectors by (gf)(n)=i=1dignifn(\nabla g\cdot\nabla f)(n)=\sum_{i=1}^{d}\nabla_{i}g_{n}\cdot\nabla_{i}f_{n}, and |f|(n):=(ff)(n).|\nabla f|(n):=\sqrt{(\nabla f\cdot\nabla f)(n)}. The discrete (graph) Laplacian Δ\Delta on d{\mathbb{Z}}^{d} is defined as usual, acting on ϕ={ϕn}nd\phi=\{\phi_{n}\}_{n\in{\mathbb{Z}}^{d}}, via

(2.4) (Δϕ)n=|mn|1=1(ϕmϕn)=1id(ϕn+ei+ϕnei 2ϕn).(\Delta\phi)_{n}=\sum_{|m-n|_{1}=1}\left(\phi_{m}-\phi_{n}\right)=\sum_{1\leq i\leq d}\left(\phi_{n+e_{i}}+\phi_{n-e_{i}}\,-\,2\phi_{n}\right).

For a real sequence {vn}nd\{v_{n}\}_{n\in{\mathbb{Z}}^{d}} on d{\mathbb{Z}}^{d}, the potential VV is a multiplication operator acting on ϕ2(d)\phi\in\ell^{2}({\mathbb{Z}}^{d}) as (Vϕ)n=vnϕn(V\phi)_{n}=v_{n}\phi_{n}. The operator Δ+V-\Delta+V is called the discrete Schrödinger operator on d{\mathbb{Z}}^{d}. If one takes vn=vn(ω)v_{n}=v_{n}(\omega) as independent, identically distributed random variables (in some probability space), the random operator Δ+V(ω)-\Delta+V(\omega) is usually referred to as the Anderson model. We refer readers to [AW, Ki] for more details and a complete introduction to tight-binding Hamiltonians and the Anderson model.

Throughout the rest of the paper, we consider the discrete Schrödinger operator Δ+V-\Delta+V restricted to a finite domain in d{\mathbb{Z}}^{d}. Let Λ=(/K)d{1¯,2¯,,K¯}d\Lambda=({\mathbb{Z}}/K{\mathbb{Z}})^{d}\cong\{\bar{1},\bar{2},\cdots,\bar{K}\}^{d}, where KK\in\mathbb{N} and k¯\bar{k}, 1kK,1\leq k\leq K, is the congruence class, modulo KK. For simplicity, we often treat Λ\Lambda as a subset of d{\mathbb{Z}}^{d}. Slightly abusing the notation, we denote by ||1|\cdot|_{1} the induced 1-norm of d{\mathbb{Z}}^{d} on the congruence class Λ\Lambda, where, for example, we consider two points (1,n2,,nd)(1,n_{2},\cdots,n_{d}) and (K,n2,,nd)(K,n_{2},\cdots,n_{d}) to be nearest neighbors and to have distance one from each other in Λ\Lambda. From now on, we will concentrate on the finite dimensional subspace 2(Λ)\ell^{2}({\Lambda}) of 2(d)\ell^{2}({\mathbb{Z}}^{d}). We frequently write :=2(Λ)Kd\mathcal{H}:=\ell^{2}({\Lambda})\cong\mathbb{R}^{K^{d}} for simplicity. The linear space \mathcal{H} is equipped with the usual inner product on Kd\mathbb{R}^{K^{d}}, which is denoted by ,=,\left\langle\cdot,\,\cdot\right\rangle=\left\langle\cdot,\,\cdot\right\rangle_{\mathcal{H}}. It is easy to check that for ϕ\phi\in\mathcal{H},

(2.5) ϕn=ϕn+Kei,nΛ,i=1,d,\phi_{n}=\phi_{n+Ke_{i}},\ n\in\Lambda,\ i=1\cdots,d,

which specifies the periodicity of ϕ\phi.

We assume that vn0v_{n}\geq 0 is real valued and non-constant, and that minΛvn0\min_{\Lambda}v_{n}\geq 0. We set Vmax:=maxΛvn>0V_{\max}:=\max_{\Lambda}v_{n}>0 and let H=HΛH=H_{\Lambda} be the restriction of Δ+V-\Delta+V to \mathcal{H}:

(2.6) (HΛϕ)n=(Δϕ)n+(Vϕ)n=|mn|1=1(ϕmϕn)+vnϕn,nΛ.(H_{\Lambda}\phi)_{n}=-(\Delta\phi)_{n}+(V\phi)_{n}=-\,\sum_{|m-n|_{1}=1}\,\left(\phi_{m}-\phi_{n}\right)\,+\,v_{n}\phi_{n},\ \ n\in{\Lambda}.

Similar to the continuous case, the discrete Hamiltonian can be written in its Dirichlet form on the periodic lattice Λ\Lambda:

ϕ,Hϕ=nΛϕn2+nΛvnϕn2,\left\langle\phi,\,H\phi\right\rangle_{\mathcal{H}}=\sum_{n\in\Lambda}\|\nabla\phi_{n}\|^{2}+\sum_{n\in\Lambda}v_{n}\phi_{n}^{2},

where ϕn2:=i=1d|iϕn|2\|\nabla\phi_{n}\|^{2}:=\sum_{i=1}^{d}|\nabla_{i}\phi_{n}|^{2}.

It is easy to check that all eigenvalues of HH in (2.6) are contained in [0,4d+Vmax][0,4d+V_{\max}] for any finite KK. For the Anderson model H=Δ+V(ω)H_{\infty}=-\Delta+V(\omega) acting on the entire space 2(d)\ell^{2}({\mathbb{Z}}^{d}), it is well known that the spectrum σ(H)\sigma(H_{\infty}) is (almost surely) the non-random set [0,4d]+suppV[0,4d+Vmax][0,4d]+{\rm supp}\,V\subset[0,4d+V_{\max}].

A linear operator acting a finite dimensional space may be viewed as a matrix. For example, HΛ=Δ+VH_{\Lambda}=-\Delta+V acting on Λ=/K\Lambda={\mathbb{Z}}/K{\mathbb{Z}} in (2.6) may be identified with the sum of the two K×KK\times K matrices,

(2.7) Δ=(21011200211012),V=(v10000v2000vK10000vK).-\Delta=\begin{pmatrix}2&-1&0&\cdots&-1\\ -1&2&\ddots&\vdots\\ 0&\ddots&\ddots&\ddots&0\\ \vdots&\ddots&\ddots&2&-1\\ -1&\cdots&0&-1&2\end{pmatrix},\quad\ \ V=\begin{pmatrix}v_{1}&0&0&\cdots&0\\ 0&v_{2}&0&\ddots&\vdots\\ 0&\ddots&\ddots&\ddots&0\\ \vdots&\ddots&\ddots&v_{K-1}&0\\ 0&\cdots&0&0&v_{K}\end{pmatrix}.

It is easy to verify that HH is invertible. Moreover, by the maximum principle (see Lemma A.2), all the matrix elements of its inverse are positive, H1(i,j)>0H^{-1}(i,j)>0 for all i,jΛi,j\in\Lambda. Therefore, there is a unique positive vector u2(Λ)u\in\ell^{2}(\Lambda) solving the equation (Hu)n=1,nΛ(Hu)_{n}=1,n\in\Lambda. The equation will be referred to as the landscape equation and the solution u={un}nΛu=\{u_{n}\}_{n\in{\Lambda}}, will be called the landscape function. The function uu thus defined is the discrete analogue of the landscape function in [FM] in the continuum setting. The discrete landscape function was first introduced in [LMF], for a one dimensional lattice with zero boundary conditions. It was studied on a higher dimensional lattice with periodic boundary conditions in [WZ]. The following result can be found in [WZ].

Theorem 5 (Theorem 2.10, Lemma 2.12 in [WZ]).

Assume that vn0v_{n}\geq 0 and is not identically zero. Let u={un}2(Λ)u=\{u_{n}\}\in\ell^{2}({\Lambda}) be the unique solution of the landscape equation (Hu)n=1(Hu)_{n}=1. Then

(2.8) minnΛun1maxnΛvn>0.\min_{{n\in{\Lambda}}}u_{n}\geq\frac{1}{\max_{n\in\Lambda}v_{n}}>0.

As shown in [ADFJM2] for the continuous case and in [WZ] for the discrete case, 1/u:={1/un}nΛ1/u:=\{1/u_{n}\}_{n\in\Lambda} serves as an effective potential via the following landscape uncertainty principle:

Theorem 6 (Lemma 2.14 in [WZ]).

For any f2(Λ)f\in\ell^{2}(\Lambda),

(2.9) f,Hf=nΛ1idun+eiun(ifnun)2+nΛ1unfn2nΛ1unfn2,\left\langle f,\,Hf\right\rangle_{\mathcal{H}}=\,\sum_{n\in{\Lambda}}\,\sum_{1\leq i\leq d}\,u_{n+e_{i}}u_{n}\,\left(\nabla_{i}\frac{f_{n}}{u_{n}}\right)^{2}\,\,+\,\sum_{n\in{\Lambda}}\,\frac{1}{u_{n}}f_{n}^{2}\,\geq\,\sum_{n\in{\Lambda}}\frac{1}{u_{n}}f_{n}^{2},

where ifnun=fn+ei/un+eifn/un\nabla_{i}\frac{f_{n}}{u_{n}}=f_{n+e_{i}}/u_{n+e_{i}}-f_{n}/u_{n}.

Let us introduce a few more notations. For a,b,aba,b\in{\mathbb{Z}},a\leq b, we denote by a,b={a,a+1,,b}\llbracket a,b\rrbracket=\{a,a+1,\cdots,b\} consecutive integers from aa to bb. We will frequently work with cubes in d{\mathbb{Z}}^{d}, and their images in Λ=(/K)d\Lambda=({\mathbb{Z}}/K{\mathbb{Z}})^{d}. For r,Q=1,rdr\in\mathbb{N},Q=\llbracket 1,r\rrbracket^{d}, we say that QQ is a cube in d{\mathbb{Z}}^{d} of side length (Q)=r\ell(Q)=r, which is the cardinality of QQ projected in each direction. We denote by |Q|=Card(Q)|Q|={\rm Card}\left(Q\right) the total cardinality of QQ and call it the volume of QQ when it is clear. For any ada\in{\mathbb{Z}}^{d}, a+Qa+Q is the translation of QQ in d{\mathbb{Z}}^{d}, respectively having the same side length and volume. For a cube QQ, we denote by 3Q3Q the cube concentric with QQ of side length 3(Q)3\ell(Q)

(2.10) 3Q:=1idki=0,±(Q)(Q+k1e1+k2e2++kded),3Q:=\bigcup_{\begin{subarray}{c}1\leq i\leq d\\ k_{i}=0,\pm\ell(Q)\end{subarray}}\,\left(Q+k_{1}e_{1}+k_{2}e_{2}+\cdots+k_{d}e_{d}\right),

see Figure 3.

Refer to caption
Figure 3. The center cube (in red) Q={4,5,6}2Q=\{4,5,6\}^{2} has side length 33. QQ is surrounded by its translations (in blue) Q+ae1+be2,a,b=0,±3Q+ae_{1}+be_{2},a,b=0,\pm 3. The union of all the small cubes is 3Q={1,,9}23Q=\{1,\cdots,9\}^{2} of side length 99.

Let Q\partial Q be the (inner) boundary of QQ:

(2.11) Q={nQ:n+eiQorneiQforsome 1id},\partial Q=\left\{\,n\in Q:\,n+e_{i}\not\in Q\ {\rm or}\ n-e_{i}\not\in Q\ {\rm for\ some}\,1\leq i\leq d\,\right\},

and let QQ\partial^{\circ}Q\subset\partial Q be the flat part of the boundary, removing all the corners:

(2.12) Q={nQ:n+eiQorneiQfor only one 1id}.\partial^{\circ}Q=\left\{n\in\partial Q:\,n+e_{i}\not\in Q\ {\rm or}\ n-e_{i}\not\in Q\ {\textrm{for only one}}\ 1\leq i\leq d\right\}.

For an integer interval I=a,a+r1I=\llbracket a,a+r-1\rrbracket of side length r3r\geq 3, we denote by I/3:=a+r/3,a+r/3+r/31I/3:=\llbracket a+\left\lceil r/3\right\rceil,a+\left\lceil r/3\right\rceil+\left\lfloor r/3\right\rfloor-1\rrbracket the middle third interval of II. For a cube Q=I1×I2××Id,Ii=ai,ai+r1,i=1,dQ=I_{1}\times I_{2}\times\cdots\times I_{d},I_{i}=\llbracket a_{i},a_{i}+r-1\rrbracket,i=1\cdots,d, we denote by Q/3Q/3 the middle third cube of QQ, defined as:

(2.13) Q/3:=(I1/3)×(I2/3)××(Id/3).Q/3:=(I_{1}/3)\times(I_{2}/3)\times\cdots\times(I_{d}/3).

It is easy to verify that Q/3Q/3 is the “thin” middle third part of QQ in the sense that (Q/3)=(Q)/3(Q)/3\ell(Q/3)=\left\lfloor\ell(Q)/3\right\rfloor\leq\ell(Q)/3 and 3(Q/3)Q3(Q/3)\subseteq Q. The relation becomes 3(Q/3)=Q3(Q/3)=Q if 3(Q)3\mid\ell(Q).

3. Landscape law: the general case and the self-improvement under the scaling condition

In this section, we study Theorems 1, 2, and 3. Let us recall some of the notations first. Let Λ1,Kd\Lambda\cong\llbracket 1,K\rrbracket^{d} be the periodic domain of side length KK. Let N(μ)N(\mu) be the (finite volume) integrated density of states (IDS) of HH on Λ\Lambda, as defined in (1.2).

Let u={un}u=\{u_{n}\} be the landscape function of HH defined in the introduction. For μ>0\mu>0, let s(μ)=μ1/2s(\mu)=\left\lceil\mu^{-1/2}\right\rceil and let Nu(μ)N_{u}(\mu) be the landscape box counting function, defined with respect to the partition 𝒫=𝒫(s(μ);Λ){\mathcal{P}}={\mathcal{P}}(s(\mu);\Lambda) as in (1.3), see Figures 4, 5 for examples on 1,2{\mathbb{Z}}^{1},{\mathbb{Z}}^{2}.

Refer to caption
Figure 4. The effective potential {1/un}n=111\{1/u_{n}\}_{n=1}^{11} is plotted in blue. The horizontal reference line is μ=1/9\mu=1/9 (in black). The partition 𝒫(3)={Qj}j=14{\mathcal{P}}(3)=\{Q_{j}\}_{j=1}^{4} contains four disjoint cubes. On Q2,Q3,Q4Q_{2},Q_{3},Q_{4}, min(1/un)\min(1/u_{n}) falls below μ\mu.
Refer to caption
Figure 5. The effective potential (in blue) on a 2{\mathbb{Z}}^{2} lattice. The reference energy μ=1/36\mu=1/36 (in pink). The partition 𝒫(6){\mathcal{P}}(6) contains 99 disjoint boxes, four regular boxes have side length 66, and five irregular boxes along the boundary.
Remark 3.1.

The landscape box counting function Nu=Nu𝒫N_{u}=N_{u}^{{\mathcal{P}}} is defined with respect to to the partition 𝒫=𝒫(r;Λ),r=μ1/2{\mathcal{P}}={\mathcal{P}}(r;\Lambda),r=\left\lceil\mu^{-1/2}\right\rceil. Consider any translation of 𝒫{\mathcal{P}} by a0,r1da\in\llbracket 0,r-1\rrbracket^{d}, i.e., 𝒫a={a+Q,Q𝒫}{\mathcal{P}}^{a}=\{a+Q,Q\in{\mathcal{P}}\}. Then 𝒫a{\mathcal{P}}^{a} is also a partition of Λ\Lambda of size rr since Λ\Lambda is a periodic torus. Each a+Q𝒫aa+Q\in{\mathcal{P}}^{a} can be covered by finitely many cubes Q𝒫Q^{\prime}\in{\mathcal{P}}, and vice versa. The number of cubes from one partition needed to cover a cube from another partition is at most 3d13^{d}-1. Therefore, if we use 𝒫a{\mathcal{P}}^{a} to define a landscape box counting function Nu𝒫aN_{u}^{{\mathcal{P}}^{a}}, the counting function will differ by at most a factor of 3±d3^{\pm d}, i.e., for any a0,r1da\in\llbracket 0,r-1\rrbracket^{d}

3dNu𝒫aNu𝒫3dNu𝒫a.3^{-d}N_{u}^{{\mathcal{P}}^{a}}\leq N_{u}^{{\mathcal{P}}}\leq 3^{d}N_{u}^{{\mathcal{P}}^{a}}.

Furthermore, if r<rr<r^{\prime}, then 𝒫(r;Λ){\mathcal{P}}(r;\Lambda) is a finer partition than 𝒫(r;Λ){\mathcal{P}}(r^{\prime};\Lambda). Each Q𝒫(r)Q^{\prime}\in{\mathcal{P}}(r^{\prime}) can be covered by at most (r/r+2)d(r^{\prime}/r+2)^{d} cubes Q𝒫(r;Λ)Q\in{\mathcal{P}}(r;\Lambda). Therefore, the number of Q𝒫(r)Q^{\prime}\in{\mathcal{P}}(r^{\prime}) such that minnQ1unμ\min_{n\in Q^{\prime}}\frac{1}{u_{n}}\,\leq\mu will differ from the number of Q𝒫(r)Q\in{\mathcal{P}}(r) such that minnQ1unμ\min_{n\in Q}\frac{1}{u_{n}}\,\leq\mu at most by a factor of (r/r+2)d(r^{\prime}/r+2)^{d}. In other words, for r<rr<r^{\prime},

Nu𝒫(r)Nu𝒫(r)(r/r+2)dNu𝒫(r).N_{u}^{{\mathcal{P}}(r^{\prime})}\leq N_{u}^{{\mathcal{P}}(r)}\leq\left({r^{\prime}}/{r}+2\right)^{d}\,N_{u}^{{\mathcal{P}}(r^{\prime})}.

Based on the above discussion, we are allowed to estimate Nu(μ)N_{u}(\mu) by either shifting the original partition 𝒫(μ1/2){\mathcal{P}}\left(\left\lceil\mu^{-1/2}\right\rceil\right) or tweaking the side length of the partition slightly. The change of the partition will lead to a different box-counting function, but the new counting function will differ from NuN_{u} only by some multiplicative dimensional constants. This will be very useful in the proofs.

3.1. Upper bound. Proof of Theorem 1

Let H=Δ+VH=-\Delta+V be as in (2.6) acting on =2(Λ)Kd\mathcal{H}=\ell^{2}(\Lambda)\cong\mathbb{R}^{K^{d}}. We denote by ,\langle\,\cdot\,,\,\cdot\,\rangle the inner product on 2(Λ)\ell^{2}(\Lambda) induced by the usual one on Kd\mathbb{R}^{K^{d}}.

Case I: μ<14d\mu<\frac{1}{4d}. We will actually start with μ<1\mu<1, and then consider the rescaling μ4d\frac{\mu}{4d} in the end. To get an upper bound for N(μ)N(\mu), it is enough to bound f,Hf\left\langle f,\,Hf\right\rangle from below on some subspace of \mathcal{H}. For r=μ1/22μ1/2r=\left\lceil\mu^{-1/2}\right\rceil\leq 2\mu^{-1/2} , let 𝒫(r)=𝒫(r;Λ){\mathcal{P}}(r)={\mathcal{P}}(r;\Lambda) be the partition of side length rr. Let

:={Q𝒫(r):minnQ1unμ}.{\mathcal{F}}:=\left\{Q\in{\mathcal{P}}(r)\,:\ \ \min_{n\in Q}\frac{1}{u_{n}}\,\leq\mu\right\}.

Let SS be the linear subspace of the vectors in \mathcal{H} whose average on each QQ\in\mathcal{F} is zero, i.e.,

S={f:1|Q|nQfn=0,Q}.S=\left\{f\in\mathcal{H}\,:\,\frac{1}{|Q|}\sum_{n\in Q}f_{n}=0,\ \ Q\in{\mathcal{F}}\right\}.

The subspace SS has Card(){\rm Card}\left(\mathcal{F}\right) many linear independent constraints since all Q𝒫Q\in{\mathcal{P}} are disjoint. Therefore, SS has codimension Card(){\rm Card}\left(\mathcal{F}\right).

By the landscape uncertainty principle (2.9),

f,Hf=nΛ(fn2+vnfn2)nΛ1unfn2.\left\langle f,\,Hf\right\rangle=\sum_{n\in\Lambda}\,\Big{(}\|\nabla f_{n}\|^{2}+v_{n}f_{n}^{2}\Big{)}\,\geq\,\sum_{n\in\Lambda}\frac{1}{u_{n}}f_{n}^{2}.

We see that

f,HfnΛfn2andf,HfnΛ1unfn2,\left\langle f,\,Hf\right\rangle\geq\sum_{n\in\Lambda}\|\nabla f_{n}\|^{2}\ \ {\rm and}\ \ \left\langle f,\,Hf\right\rangle\geq\,\sum_{n\in\Lambda}\frac{1}{u_{n}}f_{n}^{2},

which implies

2f,HfnΛ(fn2+1unfn2).2\left\langle f,\,Hf\right\rangle\geq\sum_{n\in\Lambda}\,\Big{(}\|\nabla f_{n}\|^{2}+\frac{1}{u_{n}}f_{n}^{2}\Big{)}.

Therefore, for fSf\in S,

(3.1) 2f,HfQ𝒫(r)nQ(fn2+1unfn2)QnQfn2+QnQ1unfn2.2\left\langle f,\,Hf\right\rangle\geq\sum_{Q\in{\mathcal{P}}(r)}\sum_{n\in Q}\,\left(\|\nabla f_{n}\|^{2}+\frac{1}{u_{n}}f_{n}^{2}\right)\geq\ \sum_{Q\in{\mathcal{F}}}\sum_{n\in Q}\,\|\nabla f_{n}\|^{2}\ +\ \sum_{Q\not\in{\mathcal{F}}}\sum_{n\in Q}\,\frac{1}{u_{n}}f_{n}^{2}.

In the second sum, 1/unmin1/un>μ1/u_{n}\geq\min 1/u_{n}>\mu since QQ\notin{\mathcal{F}}. Therefore,

QnQ1unfn2μQnQfn2.\sum_{Q\not\in{\mathcal{F}}}\,\sum_{n\in Q}\,\frac{1}{u_{n}}\,f_{n}^{2}\,\geq\,\mu\,\sum_{Q\not\in{\mathcal{F}}}\,\sum_{n\in Q}\,f_{n}^{2}\ \ .

To bound the gradient term on the right-hand side of (3.1), we need the discrete version of the Poincaré inequality (Lemma A.4 in Appendix A): for any cube QQ\in{\mathcal{F}} of side length (Q)=r=μ1/22μ1/2\ell(Q)=r=\left\lceil\mu^{-1/2}\right\rceil\leq 2\mu^{-1/2},

nQfn22(Q)2dnQ(fnf¯Q)22(2μ1/2)2dnQfn2=μ2dnQfn2.\sum_{n\in Q}\|\nabla f_{n}\|^{2}\geq\frac{2}{\ell(Q)^{2}d}\,\sum_{n\in Q}\,(f_{n}-\bar{f}_{Q})^{2}\,\geq\,\frac{2}{(2{\mu^{-1/2})}^{2}\,d}\,\sum_{n\in Q}f_{n}^{2}=\frac{\mu}{2d}\,\sum_{n\in Q}f_{n}^{2}.

Notice that the last Q𝒫Q\in{\mathcal{P}} in each direction may not be a regular box of equal side length, the above estimate remains the same since the side length of the irregular box does not exceed r=μ1/2r=\left\lceil\mu^{-1/2}\right\rceil, see (A.3) in Lemma A.4.

Putting these two parts together, one has for fSf\in S

f,Hf12Q𝒫nQ(fn2+1unfn2)\displaystyle\left\langle f,\,Hf\right\rangle\geq\,\frac{1}{2}\,\sum_{Q\in{\mathcal{P}}}\sum_{n\in Q}\,\left(\|\nabla f_{n}\|^{2}+\frac{1}{u_{n}}f_{n}^{2}\right)\,\geq\, 12Qμ2dnQfn2+12μQnQfn2\displaystyle\frac{1}{2}\,\sum_{Q\in{\mathcal{F}}}\frac{\mu}{2d}\,\sum_{n\in Q}f_{n}^{2}\ +\,\frac{1}{2}\,\mu\sum_{Q\not\in{\mathcal{F}}}\sum_{n\in Q}\,f_{n}^{2}
>\displaystyle>\, μ4dnΛfn2=μ4df,f.\displaystyle\frac{\mu}{4d}\sum_{n\in\Lambda}f_{n}^{2}=\frac{\mu}{4d}\left\langle f,\,f\right\rangle.

Therefore, using the mini-max characterization of eigenvalues, the number of eigenvalues of HH below μ4d\dfrac{\mu}{4d} is bounded from above by the codimension of the subspace SS, which is equal to Card(){\rm Card}\left(\mathcal{F}\right). Hence,

N(μ4d)Card()|Λ|=Nu(μ),forμ<1.N\left(\frac{\mu}{4d}\right)\leq\frac{{\rm Card}\left(\mathcal{F}\right)}{|\Lambda|}=N_{u}(\mu),\ \ {\rm for}\ \ \mu<1.

Equivalently,

(3.2) N(μ)Nu(4dμ),forμ<14d.N\left(\mu\right)\leq N_{u}(4d\mu),\ \ {\rm for}\ \ \mu<\frac{1}{4d}.

Case II: μ14d\mu\geq\frac{1}{4d}. Similar to Case I, we will work with μ1\mu\geq 1, and then consider the rescaling μ4d\frac{\mu}{4d} in the end. The construction is similar to the previous case. Let

={nΛ:1unμ},S={f:fn=0ifn}.{\mathcal{F}}=\left\{\,n\in\Lambda:\ \ \frac{1}{u_{n}}\leq\mu\right\},\ \ S=\left\{\,f\in\mathcal{H}\,:\,f_{n}=0\ \ {\rm if}\ \ n\in{\mathcal{F}}\,\right\}.

From the definition of NuN_{u} in (1.3) and the fact that μ1/2=1\left\lceil\mu^{-1/2}\right\rceil=1, we see that Nu(μ)=Card()/|Λ|N_{u}(\mu)={\rm Card}\left(\mathcal{F}\right)/{|\Lambda|}.

Due to (2.9),

f,Hfn1unfn2nμfn2=μnΛfn2=μf,fforallfS,\left\langle f,\,Hf\right\rangle\geq\sum_{n\notin{\mathcal{F}}}\,\frac{1}{u_{n}}\,f_{n}^{2}\,\geq\sum_{n\notin{\mathcal{F}}}\,\mu\,f_{n}^{2}=\mu\,\sum_{n\in\Lambda}\,f_{n}^{2}=\mu\left\langle f,\,f\right\rangle\ \ {\rm for\ all}\ f\in S,

where we have used the fact that fn=0f_{n}=0 on {\mathcal{F}} for the first equality. Notice that in this case, we do not need the Poincaré inequality.

Therefore, for all μ1\mu\geq 1, N(μ)Nu(μ)N\left({\mu}\right)\leq N_{u}(\mu). Since N(μ)N(\mu) is non-decreasing, it implies that N(μ/(4d))N(μ)Nu(μ)N\left(\mu/(4d)\right)\leq N\left(\mu\right)\leq N_{u}(\mu) for μ1\mu\geq 1. Equivalently,

(3.3) N(μ)Nu(4dμ),forμ14d.N\left(\mu\right)\leq N_{u}(4d\,\mu),\ \ {\rm for}\ \ \mu\geq\frac{1}{4d}.

Combing (3.2) and (3.3), we finish the proof for (1.4). \Box

3.2. General lower bound in the non-scaling case. Proof of Theorem 2

Similarly to the upper bound, if one can bound f,Hf\left\langle f,\,Hf\right\rangle from above on a subspace of Kd\mathcal{H}\cong\mathbb{R}^{K^{d}}, then the eigenvalue counting function will be bounded from below by the dimension of this subspace.

Proof.

Let

r=(cH/32)14μ12,r=\left\lceil\left({c_{H}}/{32}\right)^{\frac{1}{4}}\mu^{-\frac{1}{2}}\right\rceil,

where 0<cH<10<c_{H}<1 is a dimensional constant given by the discrete Moser-Harnack inequality, see Lemma A.11.

Given 0<α<(cH/32)1/4/180<\alpha<({c_{H}/32})^{-1/4}/18, let R=α1μ1/2rR=\left\lceil\alpha^{-1}\mu^{-1/2}\right\rceil\geq r.

Case I: rRKr\mid R\mid K, i.e., K=K0R,R=R0rK=K_{0}R,\,R=R_{0}r for some K0,R0K_{0},R_{0}\in\mathbb{N}. We will deal with the following three sub-cases for small, mild, and large μ\mu.

Case I(a): We first consider μ<(cH/32)12\mu<\left({c_{H}}/{32}\right)^{\frac{1}{2}} and therefore KRr2K\geq R\geq r\geq 2. In this case, one has (cH/32)14μ12r<2(cH/32)14μ12,α1μ1/2R<2α1μ1/2\left({c_{H}/32}\right)^{\frac{1}{4}}\mu^{-\frac{1}{2}}\leq r<2\left({c_{H}/32}\right)^{\frac{1}{4}}\mu^{-\frac{1}{2}},\alpha^{-1}\mu^{-1/2}\leq R<2\alpha^{-1}\mu^{-1/2}. Therefore,

(3.4) 12α1(32/cH)1/4R0=Rr2α1(32/cH)1/4.\frac{1}{2}\,\alpha^{-1}\,({32/c_{H}})^{1/4}\leq R_{0}=\frac{R}{r}\leq 2\alpha^{-1}\,({32/c_{H}})^{1/4}.

Consider the partition 𝒫(R;Λ){\mathcal{P}}(R;\Lambda) of Λ\Lambda of side length RR, and, for each QQ of side length R=R0rR=R_{0}r, consider the finer partition 𝒫(r;Q){\mathcal{P}}(r;Q) of side length rr. Clearly, the collection of all q𝒫(r;Q)q\in{\mathcal{P}}(r;Q) for all Q𝒫(R;Λ)Q\in{\mathcal{P}}(R;\Lambda) also forms a partition for Λ\Lambda of size rr:

(3.5) 𝒫(r;Λ)=Q𝒫(r;Q)={q:q𝒫(r;Q),Q𝒫(R;Λ)}.{\mathcal{P}}(r;\Lambda)=\bigcup_{Q}{\mathcal{P}}(r;Q)=\bigl{\{}q:\ q\in{\mathcal{P}}(r;Q),\ Q\in{\mathcal{P}}(R;\Lambda)\bigr{\}}.

Since α<(32/cH)1/4/18\alpha<({32/c_{H}})^{1/4}/18, then R/r=R09R/r=R_{0}\geq 9 due to (3.4). For each QQ, let qˇ\check{q} be a cube in 𝒫(r;Q){\mathcal{P}}(r;Q) such that distd(qˇ,cQ)d/2{\rm dist}_{\mathbb{R}^{d}}(\check{q},c_{Q})\leq\sqrt{d}/2, where cQc_{Q} is the center of QQ in d\mathbb{R}^{d} and the distance is measured in d\mathbb{R}^{d}. We call such qˇ\check{q} a centric cube with QQ. We see that a centric cube qˇ\check{q} satisfies 3qˇQ/33\check{q}\subset Q/3, cf. (2.10), (2.13), since R09R_{0}\geq 9. The notion of centric cube means that we consider cubes in 𝒫(r;Q){\mathcal{P}}(r;Q) and 𝒫(R;Q){\mathcal{P}}(R;Q) as subsets of d\mathbb{R}^{d}, and then pick qˇ\check{q} to be a cube in 𝒫(r;Q){\mathcal{P}}(r;Q) closest to the center of QQ in d\mathbb{R}^{d}. Note that the choice of a centric cube may not be unique. That would not effect the estimates below and the translation arguments in (3.16) due to Remark 3.1.

For any 0<α<(32/cH)1/4/180<\alpha<({32/c_{H}})^{1/4}/18, let

(3.6) ={Q𝒫(R;Λ):minnqˇ1unμandminnQ1unα2μ}.{\mathcal{F}}^{\prime}=\left\{Q\in{\mathcal{P}}(R;\Lambda)\,:\ \ \min_{n\in\check{q}}\frac{1}{u_{n}}\,\leq\,\mu\quad{\rm and}\quad\min_{n\in Q}\frac{1}{u_{n}}\,\geq\,\alpha^{2}\,\mu\right\}.

Given Q𝒫(R;Λ)Q\in{\mathcal{P}}(R;\Lambda), let Q/3QQ/3\subset Q be the middle third of QQ as usual. Let χQ={χnQ}nΛ\chi^{Q}=\{\chi^{Q}_{n}\}_{n\in\Lambda}\in\mathcal{H} be a discrete cut-off function supported on QQ and such that

(3.7) χnQ=1ifnQ/3,χnQ=0ifnQ,χnQ(0,1)otherwise,\chi^{Q}_{n}=1\ {\rm if}\ n\in Q/3,\quad\chi^{Q}_{n}=0\ {\rm if}\ n\notin Q,\quad\chi^{Q}_{n}\in(0,1)\,{\rm otherwise},

and

|χn+eiQχnQ|3R,ifn,n+eiQ,|χn+eiQχnQ|=0,ifnorn+eiQ,|\chi^{Q}_{n+e_{i}}-\chi^{Q}_{n}|\leq\frac{3}{R},\ {\rm if}\ n,n+e_{i}\in Q,\quad|\chi^{Q}_{n+e_{i}}-\chi^{Q}_{n}|=0,\ {\rm if}\ n\ {\rm or}\ n+e_{i}\notin Q,

for i=1,,di=1,\cdots,d. Such a cut-off function is the discrete analogue of a smooth bump function in the continuous case. We include an explicit construction in Appendix A.3 for reader’s convenience.

Let SS^{\prime} be the linear subspace of Kd\mathcal{H}\cong\mathbb{R}^{K^{d}} which is spanned by the cut-offs of uu to each QQ in \mathcal{F}^{\prime}. More precisely, we define

S=span{uQ={unQ}nΛ:unQ=unχnQ,Q}.S^{\prime}={\rm span}\big{\{}u^{Q}=\{u^{Q}_{n}\}_{n\in\Lambda}\in\mathcal{H}:\ u^{Q}_{n}=u_{n}\,\chi^{Q}_{n},\ Q\in{\mathcal{F}}^{\prime}\big{\}}.

The subspace SS^{\prime} has dimension Card(){\rm Card}\left(\mathcal{F}^{\prime}\right) since all QQ are disjoint and un>0u_{n}>0.

We aim to estimate uQ,HuQ/uQ,uQ{\left\langle u^{Q},\,H\,u^{Q}\right\rangle}/{\left\langle u^{Q},\,u^{Q}\right\rangle} from above for each uQu^{Q} in SS^{\prime}. First, by the landscape uncertainty principle (2.9),

uQ,HuQ=\displaystyle\left\langle u^{Q},\,Hu^{Q}\right\rangle= nΛ1idun+eiun(χn+eiQχnQ)2+nQ1un(unχnQ)2\displaystyle\sum_{n\in\Lambda}\,\sum_{1\leq i\leq d}u_{n+e_{i}}u_{n}\left(\chi^{Q}_{n+e_{i}}-\chi^{Q}_{n}\right)^{2}\,+\,\sum_{n\in Q}\,\frac{1}{u_{n}}\left(u_{n}\chi^{Q}_{n}\right)^{2}
(3.8) \displaystyle\leq 1idn,n+eiQun+eiun(3R)2+nQun9dRd2supQun2+RdsupQun.\displaystyle\sum_{1\leq i\leq d}\sum_{n,n+e_{i}\in Q}\,u_{n+e_{i}}u_{n}\left(\frac{3}{R}\right)^{2}\,+\,\sum_{n\in Q}\,u_{n}\leq{9d}\,{R^{d-2}}\sup_{Q}u_{n}^{2}+R^{d}\sup_{Q}u_{n}.

On the other hand, recall that 3qˇQ/33\check{q}\subset Q/3, so that

uQ,uQ=nQ(unχnQ)2nQ/3un2n3qˇun2.\left\langle u^{Q},\,u^{Q}\right\rangle=\sum_{n\in Q}\,\left(u_{n}\chi^{Q}_{n}\right)^{2}\geq\sum_{n\in Q/3}\,u^{2}_{n}\geq\sum_{n\in 3\check{q}}\,u^{2}_{n}.

Since (Δu)n=1vnun1-(\Delta u)_{n}=1-v_{n}u_{n}\leq 1, applying the discrete Moser-Harnack inequality in Lemma A.11, Appendix A.6, to unu_{n} on the smaller cubes qˇ\check{q} with (qˇ)=r\ell(\check{q})=r, one has

(3.9) n3qˇun2rd(cHsupqˇun2r4).\sum_{n\in 3\check{q}}\,u^{2}_{n}\geq r^{d}\Bigl{(}c_{H}\sup_{\check{q}}u^{2}_{n}-r^{4}\Bigr{)}.

By the definition of {\mathcal{F}}^{\prime} in (3.6), one has

(3.10) supnQunα2μ1,supnqˇunμ1.\sup_{n\in Q}u_{n}\leq\alpha^{-2}\mu^{-1},\ \ \ \sup_{n\in\check{q}}u_{n}\geq\mu^{-1}.

Notice that r=(cH/32)14μ12r=\left\lceil\left({c_{H}/32}\right)^{\frac{1}{4}}\mu^{-\frac{1}{2}}\right\rceil implies (cH/32)14μ12r<2(cH/32)14μ12\left({c_{H}/32}\right)^{\frac{1}{4}}\mu^{-\frac{1}{2}}\leq r<2\left({c_{H}/32}\right)^{\frac{1}{4}}\mu^{-\frac{1}{2}}, i.e.,

(3.11) r412cHμ2,r232/cHμ.r^{4}\leq\frac{1}{2}{c_{H}}\mu^{-2},\ \ r^{-2}\leq\sqrt{{32/c_{H}}}\,\mu.

Therefore, putting (3.8) and (3.9) together and using R=R0rR=R_{0}r,

uQ,HuQuQ,uQ\displaystyle\frac{\left\langle u^{Q},\,Hu^{Q}\right\rangle}{\left\langle u^{Q},\,u^{Q}\right\rangle}\leq 9dRd2supQun2+RdsupQunrd(cHsupqˇun2r4)9dR0d2rd2α4μ2+R0drdα2μ1rd(cHμ212cHμ2)\displaystyle\,\frac{{9d}\,{R^{d-2}}\sup_{Q}u_{n}^{2}+R^{d}\sup_{Q}u_{n}}{r^{d}\left(c_{H}\sup_{\check{q}}u^{2}_{n}-r^{4}\right)}\leq\,\frac{9d\,{R^{d-2}_{0}r^{d-2}}\alpha^{-4}\mu^{-2}+R^{d}_{0}\,r^{d}\,\alpha^{-2}\mu^{-1}}{r^{d}\left(c_{H}\mu^{-2}-{\frac{1}{2}{c_{H}}\mu^{-2}}\right)}
(3.12) \displaystyle\leq C(R0d2α4+R0dα2)μC2αd2μ,\displaystyle\,C\left(R^{d-2}_{0}\,\alpha^{-4}+R^{d}_{0}\,\alpha^{-2}\right)\mu\leq\,C_{2}\,\alpha^{-d-2}\,\mu,

where C2=C2(d,cH)C_{2}=C_{2}(d,c_{H}). In the last line, we used (3.4) with R0α1R_{0}\lesssim\alpha^{-1}.

Then, by orthogonality of the uQu^{Q} for QQ\in{\mathcal{F}}^{\prime}, we get for

(3.13) 𝒩0:=Card{eigenvaluesλofHsuchthatλC2αd2μ}{\mathcal{N}}_{0}:={\rm Card}\left\{\,{\textrm{eigenvalues}}\ \lambda\ {\rm of}\ H\ {\rm such\ that}\ \lambda\leq C_{2}\alpha^{-d-2}\,\mu\,\right\}

the estimate

(3.14) 𝒩0Card()=Card{Q𝒫(R;Λ):minnqˇ1unμandminnQ1unα2μ}Card{Q𝒫(R;Λ):minnqˇ1unμ}Card{Q𝒫(R;Λ):minnQ1unα2μ}:=𝒩1𝒩2.{\mathcal{N}}_{0}\geq\,{\rm Card}({\mathcal{F}}^{\prime})\\ ={\rm Card}\left\{Q\in{\mathcal{P}}(R;\Lambda)\,:\ \ \min_{n\in\check{q}}\frac{1}{u_{n}}\,\leq\,\mu\quad{\rm and}\quad\min_{n\in Q}\frac{1}{u_{n}}\,\geq\,\alpha^{2}\,\mu\right\}\\ \geq\,{\rm Card}\left\{\ Q\in{\mathcal{P}}(R;\Lambda)\,:\ \ \min_{n\in\check{q}}\frac{1}{u_{n}}\,\leq\,\mu\right\}\\ -{\rm Card}\left\{\ Q\in{\mathcal{P}}(R;\Lambda)\,:\min_{n\in Q}\frac{1}{u_{n}}\,\leq\,\alpha^{2}\,\mu\right\}:=\,{\mathcal{N}}_{1}-{\mathcal{N}}_{2}.

Given an integer jj, |j|R0/2|j|\leq\left\lfloor R_{0}/2\right\rfloor, we consider a translation Ti,j:ΛΛT^{i,j}:\Lambda\to\Lambda, by the vector jreijre_{i}, i.e., Ti,j(n)=n+jreiT^{i,j}(n)=n+jre_{i} for any nΛn\in\Lambda. For the partition 𝒫(R;Λ){\mathcal{P}}(R;\Lambda), denote by 𝒫i,j(R;Λ){\mathcal{P}}^{i,j}(R;\Lambda) the partition translated by Ti,jT^{i,j}. Recall that 𝒫(r;Λ){\mathcal{P}}(r;\Lambda) is the finer partition of side length rr, see (3.5). Denote by 𝒫i,j(r;Λ){\mathcal{P}}^{i,j}(r;\Lambda) the translation of 𝒫(r;Λ){\mathcal{P}}(r;\Lambda), which again is a refinement for 𝒫i,j(R;Λ){\mathcal{P}}^{i,j}(R;\Lambda). For any qˇQ𝒫(R;Λ)\check{q}\subset Q\in{\mathcal{P}}(R;\Lambda), it is easy to check that Qi,jTi,j(qˇ).Q\subset\bigcup_{i,j}T^{i,j}(\check{q}). In other words, the collection of all Ti,j(qˇ)T^{i,j}(\check{q}) will cover the entire QQ, provided enough tanslations of qˇ\check{q} (at most R0/2\left\lfloor R_{0}/2\right\rfloor many).

For each translated centric cube Ti,j(qˇ)T^{i,j}(\check{q}) in the corresponding Ti,j(Q)𝒫i,j(R;Λ)T^{i,j}(Q)\in{\mathcal{P}}^{i,j}(R;\Lambda), we repeat the construction of {\mathcal{F}}^{\prime} and 𝒮{\mathcal{S}}^{\prime} starting from (3.6). By exactly the same argument as for (3.14),

(3.15) 𝒩0𝒩1i,j𝒩2i,j,{\mathcal{N}}_{0}\geq{{\mathcal{N}}}^{i,j}_{1}-{{\mathcal{N}}}^{i,j}_{2},

where 𝒩0{\mathcal{N}}_{0} is the same as in (3.13) since the eigenvalue counting will be the same for all the translations, and

𝒩1i,j=\displaystyle{{\mathcal{N}}}^{i,j}_{1}= Card{Ti,j(Q)𝒫i,j(R;Λ):minnTi,j(qˇ)1unμ},\displaystyle\,{\rm Card}\left\{\ T^{i,j}(Q)\in{\mathcal{P}}^{i,j}(R;\Lambda)\,:\ \ \min_{n\in T^{i,j}(\check{q})}\frac{1}{u_{n}}\,\leq\,\mu\right\},
𝒩2i,j=\displaystyle{{\mathcal{N}}}^{i,j}_{2}= Card{Ti,j(Q)𝒫i,j(R;Λ):minnTi,j(Q)1unα2μ}.\displaystyle\,{\rm Card}\left\{\ T^{i,j}(Q)\in{\mathcal{P}}^{i,j}(R;\Lambda)\,:\min_{n\in T^{i,j}(Q)}\frac{1}{u_{n}}\,\leq\,\alpha^{2}\,\mu\right\}.

Recall that qˇ\check{q} has side length rr and is located near the center of each QQ. We repeat the above process exactly σ\sigma times so that j,iTi,j(qˇ)=Q\bigcup_{j,i}\,T^{i,j}(\check{q})\,=Q, and therefore, Qj,iTi,j(qˇ)=Λ\bigcup_{Q}\bigcup_{j,i}\,T^{i,j}(\check{q})=\Lambda, which is exactly the fine partition 𝒫(r;Λ){\mathcal{P}}(r;\Lambda) of the entire domain. One translated Ti,j(qˇ)T^{i,j}(\check{q}) corresponds to exactly one small cube qq in the original partition 𝒫(r;Λ){\mathcal{P}}(r;\Lambda). The number σ\sigma of the translations we need can be bounded from above by σ(2R0/2)dR0dCαd\sigma\leq\left(2\left\lfloor R_{0}/2\right\rfloor\right)^{d}\leq R_{0}^{d}\leq\,C\alpha^{-d}. Notice that for all j,ij,i, we have 𝒩2i,j<C𝒩2{{\mathcal{N}}}^{i,j}_{2}<\,C{\mathcal{N}}_{2} for some dimensional constant CC because of Remark 3.1. Then, summing up (3.15) over all possible translations,

(3.16) σ𝒩0\displaystyle\sigma{\mathcal{N}}_{0}\geq i,j𝒩1i,jσ𝒩2i,j\displaystyle\,\sum_{i,j}{{\mathcal{N}}}^{i,j}_{1}\ -\ \sigma{{\mathcal{N}}}^{i,j}_{2}
=\displaystyle= i,jCard{Ti,j(Q)𝒫i,j(R;Λ):minnTi,j(qˇ)1unμ}σ𝒩2i,j\displaystyle\,\sum_{i,j}{\rm Card}\left\{\ T^{i,j}(Q)\in{\mathcal{P}}^{i,j}(R;\Lambda)\,:\ \ \min_{n\in T^{i,j}(\check{q})}\frac{1}{u_{n}}\,\leq\,\mu\right\}-\ \sigma{{\mathcal{N}}}^{i,j}_{2}
=\displaystyle= j,iCard{Ti,j(qˇ):minnTi,j(qˇ)1unμ}σ𝒩2i,j\displaystyle\,\sum_{j,i}{\rm Card}\left\{\ T^{i,j}(\check{q}):\ \ \min_{n\in T^{i,j}(\check{q})}\frac{1}{u_{n}}\,\leq\,\mu\right\}-\ \sigma{{\mathcal{N}}}^{i,j}_{2}
\displaystyle\geq Card{q𝒫(r;Λ):minnq1unμ}Cσ𝒩2.\displaystyle\,{\rm Card}\left\{q\in{\mathcal{P}}(r;\Lambda):\ \ \min_{n\in q}\frac{1}{u_{n}}\,\leq\,\mu\right\}-\ C\sigma{{\mathcal{N}}}_{2}.

Therefore, the upper bound on the number of the translations σCαd\sigma\leq\,C\alpha^{-d} implies

(3.17) Cαd𝒩0Card{q𝒫(r;Λ):minnq1unμ}C~αd𝒩2.C\alpha^{-d}{\mathcal{N}}_{0}\geq{\rm Card}\left\{q\in{\mathcal{P}}(r;\Lambda):\ \ \min_{n\in q}\frac{1}{u_{n}}\,\leq\,\mu\right\}-\widetilde{C}\alpha^{-d}{\mathcal{N}}_{2}.

Notice that the partition in the counting of 𝒩2{\mathcal{N}}_{2} is of side length R=α1μ1/2=(α2μ)1/2R=\left\lceil\alpha^{-1}\mu^{-1/2}\right\rceil=\left\lceil(\alpha^{2}\mu)^{-1/2}\right\rceil, which is exactly the side length needed in the definition of Nu(α2μ)N_{u}(\alpha^{2}\mu). On the other hand, the partition in the counting of 𝒩1{\mathcal{N}}_{1} is of side length r=(cH/32)14μ12r=\left\lceil\left({c_{H}/32}\right)^{\frac{1}{4}}\mu^{-\frac{1}{2}}\right\rceil. The side length needed in the definition of Nu(μ)N_{u}(\mu) should be r=μ1/2r^{\prime}=\left\lceil\mu^{-1/2}\right\rceil, which is larger than rr used in 𝒩1{\mathcal{N}}_{1}. But the two counting functions defined by 𝒫(r){\mathcal{P}}(r) or 𝒫(r){\mathcal{P}}(r^{\prime}) will only differ by a dimensional factor since 1r/r2(32/cH)1/41\leq r^{\prime}/r\leq 2({{32/c_{H}}})^{1/4}, see Remark 3.1. Therefore,

Card{q𝒫(r;Λ):minnq1unμ}C1Card{q𝒫(r;Λ):minnq1unμ}.{\rm Card}\left\{q\in{\mathcal{P}}(r;\Lambda)\,:\ \ \min_{n\in q}\frac{1}{u_{n}}\,\leq\,\mu\right\}\geq C^{-1}{\rm Card}\left\{q^{\prime}\in{\mathcal{P}}(r^{\prime};\Lambda)\,:\ \ \min_{n\in q^{\prime}}\frac{1}{u_{n}}\,\leq\,\mu\right\}.

Then by (3.17)

αd𝒩0cCard{q𝒫(r;Λ):minnq1unμ}\displaystyle\alpha^{-d}{\mathcal{N}}_{0}\geq c\,{\rm Card}\left\{q^{\prime}\in{\mathcal{P}}(r^{\prime};\Lambda)\,:\ \ \min_{n\in q^{\prime}}\frac{1}{u_{n}}\,\leq\,\mu\right\}
CαdCard{Q𝒫(R;Λ):minnQ1unα2μ},\displaystyle-C\alpha^{-d}{\rm Card}\left\{\ Q\in{\mathcal{P}}(R;\Lambda)\,:\min_{n\in Q}\frac{1}{u_{n}}\,\leq\,\alpha^{2}\,\mu\right\},

which implies that

(3.18) N(C2αd2μ)cαdNu(μ)CNu(α2μ),forall 0<μ<(cH/32)1/2,N(C_{2}\alpha^{-d-2}\mu)\geq c\alpha^{d}N_{u}(\mu)-CN_{u}(\alpha^{2}\mu),\ {\rm for\ all}\ 0<\mu<(c_{H}/32)^{1/2},

provided that 0<α<(32/cH)1/4/180<\alpha<({32/c_{H}})^{1/4}/18, Kα1μ1/2K\geq\left\lceil\alpha^{-1}\mu^{-1/2}\right\rceil.

We notice that the construction above also needs the entire domain to be large enough, i.e., KRα1μ1/2K\geq R\geq\alpha^{-1}\mu^{-1/2}. The restrictions on KK can be removed easily. If 3r<K<α1μ1/23r<K<\left\lceil\alpha^{-1}\mu^{-1/2}\right\rceil, then we can repeat the above construction by setting R=KR=K directly, the proof for (3.18) is exactly the same. If K3rμ1/2K\leq 3r\lesssim\mu^{-1/2} then there are at most 3d3^{d} boxes in the partition. Therefore, Nu(μ)3d/KdN_{u}(\mu)\leq 3^{d}/K^{d}. On the other hand, the argument for (3.8) can be used to show that the ground state eigenvalue E0E_{0} of HH is bounded from above by E0CμE_{0}\leq C\mu, with a dimensional constant CC. Therefore, N(Cμ)1/KdNu(μ)N(C\mu)\geq 1/K^{d}\gtrsim N_{u}(\mu). Then (3.18) holds trivially by picking α\alpha small, and the smallness only depends on the dimension.

Case I(b): Next we consider (cH/32)12μ(3α)2\left({c_{H}}/{32}\right)^{\frac{1}{2}}\leq\mu\leq(3\alpha)^{-2}. Note that the range of μ\mu requires α<(32/cH)1/4/3\alpha<(32/c_{H})^{1/4}/3, which is fulfilled by the assumption of α\alpha. In this case, the side length (q)=r=(cH/32)14μ12=1\ell(q)=r=\left\lceil\left({c_{H}}/{32}\right)^{\frac{1}{4}}\mu^{-\frac{1}{2}}\right\rceil=1 and (Q)=R=α1μ1/23\ell(Q)=R=\left\lceil\alpha^{-1}\mu^{-1/2}\right\rceil\geq 3. For each Q𝒫(R)Q\in{\mathcal{P}}(R), we pick qˇ\check{q} to be a centric cube with QQ, then construct ,S{\mathcal{F}}^{\prime},S^{\prime} in the same way as (3.6). The upper bound (3.8) for uQ,HQ\left\langle u^{Q},\,H^{Q}\right\rangle remains the same. To bound uQ,uQ\left\langle u^{Q},\,u^{Q}\right\rangle from below, one has trivially uQ,uQQ/3un2qˇun2μ2\left\langle u^{Q},\,u^{Q}\right\rangle\geq\sum_{Q/3}u_{n}^{2}\geq\sum_{\check{q}}u_{n}^{2}\geq\mu^{-2} since qˇQ/3\check{q}\subset Q/3. The bounds on μ\mu and the definition of RR imply

R2α1μ1/2 2(cH/32)14α1,andR2(α1μ1/2)2=α2μ.R\leq 2\alpha^{-1}\mu^{-1/2}\leq\,2\left({c_{H}}/{32}\right)^{-\frac{1}{4}}\,\alpha^{-1},\ {\rm and}\ R^{-2}\leq\,(\alpha^{-1}\mu^{-1/2})^{-2}=\alpha^{2}\mu.

Hence,

uQ,HQuQ,uQ9dRd2α4μ2+Rdα2μ1μ2\displaystyle\frac{\left\langle u^{Q},\,H^{Q}\right\rangle}{\left\langle u^{Q},\,u^{Q}\right\rangle}\leq\frac{{9d}{R^{d-2}}\alpha^{-4}\mu^{-2}+R^{d}\alpha^{-2}\mu^{-1}}{\mu^{-2}}\leq  9d(Rdα4)R2+(Rdα2)μ\displaystyle\,{9d}({R^{d}}\alpha^{-4})R^{-2}+(R^{d}\alpha^{-2})\mu
(3.19) \displaystyle\leq C2αd2μ,\displaystyle\,{C}_{2}\,\alpha^{-d-2}\,\mu,

where the constant C2{C}_{2} only depends on dd and cHc_{H}. We obtain the bound as in (3.12), and therefore the same bound (3.18) holds for all 0<μ(3α)20<\mu\leq(3\alpha)^{-2}.

Repeating the proof of (3.18) for μ~=c1αd+2μ(3α)2\widetilde{\mu}=c_{1}\alpha^{d+2}\mu\leq(3\alpha)^{-2} with c1=C21c_{1}=C^{-1}_{2}, we obtain

(3.20) N(μ)cαdNu(c1αd+2μ)CNu(c1αd+4μ),forallμ<19c1αd4,N(\mu)\geq c{\alpha^{d}}N_{u}(c_{1}\alpha^{d+2}\mu)-CN_{u}(c_{1}\alpha^{d+4}\mu),\ {\rm for\ all}\ \mu<\frac{1}{9}c_{1}\,\alpha^{-d-4},

provided 0<α<(32/cH)1/4/180<\alpha<({32/c_{H}})^{1/4}/18.

Case I(c): The remaining case is μ(3α)2\mu\geq(3\alpha)^{-2}. Since α<(32/cH)1/4/18\alpha<({32/c_{H}})^{1/4}/18, then μ(cH/32)12\mu\geq\left({c_{H}}/{32}\right)^{\frac{1}{2}}. In this case, the range of μ\mu implies r=(cH/32)14μ12=1r=\left\lceil\left({c_{H}}/{32}\right)^{\frac{1}{4}}\mu^{-\frac{1}{2}}\right\rceil=1 and R=α1μ1/2<3R=\left\lceil\alpha^{-1}\mu^{-1/2}\right\rceil<3. We construct ,S{\mathcal{F}}^{\prime},S^{\prime} with cube QQ of side length R~=9\widetilde{R}=9 instead. Note that this will not change the counting for 𝒩0{\mathcal{N}}_{0} and 𝒩1{\mathcal{N}}_{1}. The change will only result a different counting for 𝒩2{\mathcal{N}}_{2}, which we denote by 𝒩~2\widetilde{\mathcal{N}}_{2} the new counting using (Q)=R~=9\ell(Q)=\widetilde{R}=9. Since 1R~/R91\leq\widetilde{R}/R\leq 9, one has 𝒩~2𝒩29d𝒩~2\widetilde{\mathcal{N}}_{2}\leq{\mathcal{N}}_{2}\leq 9^{d}\widetilde{\mathcal{N}}_{2}, due to Remark 3.1. Similar to (3.19), one has

uQ,HQuQ,uQ 9d9d2α4+9dα2μ(9d9d1+9d)α2μ,\frac{\left\langle u^{Q},\,H^{Q}\right\rangle}{\left\langle u^{Q},\,u^{Q}\right\rangle}\leq\,{9d}{9^{d-2}}\,\alpha^{-4}+9^{d}\alpha^{-2}\mu\leq\,({9d}{9^{d-1}}+9^{d})\,\alpha^{-2}\,\mu,

since α29μ\alpha^{-2}\leq 9\,\mu. Then we repeat the arguments for (3.15)-(3.18) using translations. Notice that the number σ\sigma of translations needed in this case is bounded from above by a dimensional constant σC\sigma\leq C. We obtain instead

(3.21) N(C2α2μ)cNu(μ)CNu(α2μ)forallμ(3α)2.N(C^{\prime}_{2}\alpha^{-2}\mu)\geq\,cN_{u}(\mu)-CN_{u}(\alpha^{2}\mu)\ \ \ {\rm for\ all}\ \mu\geq(3\alpha)^{-2}.

Repeating the proof of (3.21) for μ~=c1αd+2μ(3α)2\widetilde{\mu}=c^{\prime}_{1}\alpha^{d+2}\mu\geq(3\alpha)^{-2} where c1=(C2)1c^{\prime}_{1}=(C^{\prime}_{2})^{-1}, we reach

(3.22) N(μ)cNu(c1α2μ)CNu(c1α4μ)forallμ>19c1α4,N(\mu)\geq\,cN_{u}(c^{\prime}_{1}\alpha^{2}\mu)-CN_{u}(c^{\prime}_{1}\alpha^{4}\mu)\ \ \ {\rm for\ all}\ \mu>\frac{1}{9}c^{\prime}_{1}\,\alpha^{-4},

provided 0<α<(32/cH)1/4/180<\alpha<({32/c_{H}})^{1/4}/18.

Finally, we require further that α<(c1/c1)1/d\alpha<(c_{1}/c_{1}^{\prime})^{1/d} so that cα4<19c1αd4c_{\ast}\alpha^{-4}<\frac{1}{9}c_{1}\,\alpha^{-d-4} where c=19c1c_{\ast}=\frac{1}{9}c^{\prime}_{1}. Therefore, the estimates (3.20) and (3.22) cover μcα4\mu\leq c_{\ast}\alpha^{-4} and μ>cα4\mu>c_{\ast}\alpha^{-4} respectively. This completes the proof of Theorem 2.

Case II: either r∤Rr\not\mid R or R∤KR\not\mid K, where r=(cH/32)1/4μ1/2r=\left\lceil\left({c_{H}}/{32}\right)^{1/4}\mu^{-1/2}\right\rceil and R=α1μ1/2R=\left\lceil\alpha^{-1}\mu^{-1/2}\right\rceil are the same in Case I. Without loss of generality we assume that

K=(K01)R+R~, 0<R~<R,andR=(R01)r+r~, 0<r~<r.K=(K_{0}-1)\,R+\widetilde{R},\,0<\widetilde{R}<R,\ \ {\rm and}\ \ R=(R_{0}-1)\,r+\widetilde{r},\,0<\widetilde{r}<r.

The other two cases, where either R~=R\widetilde{R}=R or r~=r\widetilde{r}=r, are similar. Recall the construction of the partition for 𝒫(R;Λ){\mathcal{P}}(R;\Lambda) and 𝒫(r;Q){\mathcal{P}}(r;Q). In the last row and column of each direction we need to use a rectangular box instead of a cube. We denote the regular cube of side length RR or rr still by QQ and qq, and denote the remaining special rectangular boxes by Q~\widetilde{Q} and q~\widetilde{q}, whose side lengths are R~\widetilde{R} and r~\widetilde{r}, respectively, in at least one direction, and write

(3.23) 𝒫(R;Λ)={Q}{Q~},𝒫(r;Q)={q}{q~}.{\mathcal{P}}(R;\Lambda)=\{Q\}\cup\{\widetilde{Q}\},\ \ {\mathcal{P}}(r;Q)=\{q\}\cup\{\widetilde{q}\}.

Notice in this case, the union 𝒫~=Q𝒫(R;Λ)𝒫(r;Q)\widetilde{{\mathcal{P}}}=\bigcup_{Q\in{\mathcal{P}}(R;\Lambda)}{\mathcal{P}}(r;Q) is not the original partition 𝒫(r;Λ){\mathcal{P}}(r;\Lambda), but it is a finer one. Therefore, the counting function defined through 𝒫~\widetilde{{\mathcal{P}}} will be bounded from below by the counting function defined through 𝒫(r;Λ){\mathcal{P}}(r;\Lambda).

In this case, we define {\mathcal{F}}^{\prime} only using the regular cubes QQ and ignoring all the Q~\widetilde{Q}. Then by exact the same construction, we obtain (3.14), i.e., 𝒩0𝒩1𝒩2{\mathcal{N}}_{0}\geq{\mathcal{N}}_{1}-{\mathcal{N}}_{2}. Next, we need to translate the partition 𝒫(R;Λ){\mathcal{P}}(R;\Lambda) and 𝒫(r;Q){\mathcal{P}}(r;Q) by vectors of length rr several steps in each direction, so that the centric small cubes qˇ\check{q} can cover the large cubes QQ. In the previous case, we needed at most jR0/2j\sim\left\lfloor R_{0}/2\right\rfloor steps in each direction. In Case II, we want to continue the translation up to j~2R0\widetilde{j}\sim\left\lfloor 2R_{0}\right\rfloor steps in each direction, where the total number of translations in all directions is at most σ~(22R0)dCαd\widetilde{\sigma}\leq(22R_{0})^{d}\leq C\alpha^{-d}. By doing this, the translated centric cubes Ti,j(qˇ)T^{i,j}(\check{q}) will cover 3Q3Q. In particular, they will cover all the irregular boxes Q~\widetilde{Q} near the boundary of the domain. Notice that for the regular cubes QQ, translations up to j~2R0\widetilde{j}\sim\left\lfloor 2R_{0}\right\rfloor steps will cause an overlap, which leads to an overestimate in the corresponding sum i,j𝒩1i,j\sum_{i,j}{\mathcal{N}}^{i,j}_{1} as in (3.16). But since j~2R0\widetilde{j}\leq 2R_{0} and the Ti,j(qˇ)T^{i,j}(\check{q}) are contained in 5Q5Q for each QQ, the over-counting will be at most 5d5^{d} times more. In conclusion, we can obtain

σ~Card{eigenvaluesλofHsuchthatλC2αd2μ}\displaystyle\widetilde{\sigma}\,{\rm Card}\left\{\,{\textrm{eigenvalues}}\ \lambda\ {\rm of}\ H\ {\rm such\ that}\ \lambda\leq C_{2}\alpha^{-d-2}\,\mu\,\right\}
j,iCard{Ti,j(qˇ):minnTi,j(qˇ)1unμ}σ~Card{Q𝒫(R;Λ):minnQ1unα2μ}\displaystyle\geq\sum_{j,i}{\rm Card}\left\{T^{i,j}(\check{q}):\min_{n\in T^{i,j}(\check{q})}\frac{1}{u_{n}}\,\leq\,\mu\right\}-\widetilde{\sigma}\,{\rm Card}\left\{\ Q\in{\mathcal{P}}(R;\Lambda)\,:\min_{n\in Q}\frac{1}{u_{n}}\,\leq\,\alpha^{2}\,\mu\right\}
15dCard{q𝒫~:minnq1unμ}σ~Card{Q𝒫(R;Λ):minnQ1unα2μ}.\displaystyle\geq\frac{1}{5^{d}}{\rm Card}\left\{q\in\widetilde{{\mathcal{P}}}:\min_{n\in q}\frac{1}{u_{n}}\,\leq\,\mu\right\}-\ \widetilde{\sigma}\,{\rm Card}\left\{\ Q\in{\mathcal{P}}(R;\Lambda)\,:\min_{n\in Q}\frac{1}{u_{n}}\,\leq\,\alpha^{2}\,\mu\right\}.

Note that in the last line the collection of pertinent cubes qq or QQ already includes the irregular boxes q~\widetilde{q} or Q~\widetilde{Q} respectively. Together with the fact that 𝒫~\widetilde{\mathcal{P}} is finer than 𝒫(r;Λ){\mathcal{P}}(r;\Lambda), we obtain that

αdCard{eigenvaluesλofHsuchthatλC2αd2μ}\displaystyle\alpha^{-d}{\rm Card}\left\{\,{\textrm{eigenvalues}}\ \lambda\ {\rm of}\ H\ {\rm such\ that}\ \lambda\leq C_{2}\alpha^{-d-2}\,\mu\,\right\}
5dCard{q𝒫(r;Λ):minnq1unμ}CαdCard{Q𝒫(R;Λ):minnQ1unα2μ},\displaystyle\quad\geq{5^{-d}}{\rm Card}\left\{q\in{\mathcal{P}}(r;\Lambda):\min_{n\in q}\frac{1}{u_{n}}\leq\mu\right\}-C\alpha^{-d}{\rm Card}\left\{Q\in{\mathcal{P}}(R;\Lambda):\min_{n\in Q}\frac{1}{u_{n}}\leq\alpha^{2}\,\mu\right\},

which implies N(C2αd2μ)cαdNu(μ)CNu(α2μ)N(C_{2}\alpha^{-d-2}\mu)\geq c\alpha^{d}N_{u}(\mu)-CN_{u}(\alpha^{2}\mu). This is the same estimate as we obtained in (3.18). Replacing μ\mu by C21αd+2μC_{2}^{-1}\alpha^{d+2}\mu as in the remaining arguments in Case I, we can obtain (3.20) and (3.22) in a similar manner. ∎

3.3. Lower bound in the scaling case. Proof of Theorem 3

Let

(3.24) R=(12C′′)14μ12,R=\left\lceil\left(\frac{1}{2C^{\prime\prime}}\right)^{\frac{1}{4}}\mu^{-\frac{1}{2}}\right\rceil,

where C′′1C^{\prime\prime}\geq 1 is a dimensional constant that will be specified later.

We need to consider two cases, corresponding to RKR\mid K and R∤KR\not\mid K. Similar to the arguments in the previous subsection, the latter can be be reduced to the former by a translation argument. For simplicity, we will only deal with the case K=K0RK=K_{0}R for some K0K_{0}\in\mathbb{N}. Also, we first assume that μ\mu is small and R3R\geq 3, so that the cube of side length RR is large enough to construct cut off functions as in (3.7). Otherwise, when μ\mu is large and RR is small, we start with R~=9\widetilde{R}=9 and tweak the dimensional constants in the end by Remark 3.1 as in the previous subsection.

For (12C′′)14μ123\left(\frac{1}{2C^{\prime\prime}}\right)^{\frac{1}{4}}\mu^{-\frac{1}{2}}\geq 3, i.e., μ<19(12C′′)12\mu<\frac{1}{9}\left(\frac{1}{2C^{\prime\prime}}\right)^{-\frac{1}{2}}, one has R3R\geq 3. We consider the partition 𝒫(R;Λ){\mathcal{P}}(R;\Lambda) consisting cubes of side length RR as usual. Let

′′={Q𝒫(R;Λ):minnQ1unμ},S′′=span{uQ:unQ=unχnQ,Q′′},{\mathcal{F}}^{\prime\prime}=\left\{Q\in{\mathcal{P}}(R;\Lambda)\,:\ \min_{n\in Q}\frac{1}{u_{n}}\,\leq\,\mu\right\},\ \ S^{\prime\prime}={\rm span}\left\{u^{Q}\in\mathcal{H}:\,u^{Q}_{n}=u_{n}\chi^{Q}_{n},\,Q\in{\mathcal{F}}^{\prime\prime}\right\},

where χQ={χnQ}\chi^{Q}=\{\chi^{Q}_{n}\} is the cut-off function as in (A.7) on each QQ. The dimension of S′′S^{\prime\prime} equals Card(′′){\rm Card}\left(\mathcal{F}^{\prime\prime}\right).

Our goal, once again, is to establish estimates similar to (3.12). The upper bound for uQ,HuQ\left\langle u^{Q},\,Hu^{Q}\right\rangle remains the same as we obtained in (3.8):

(3.25) uQ,HuQ9dRd2supQun2+RdsupQun.\left\langle u^{Q},\,Hu^{Q}\right\rangle\leq{9d}\,{R^{d-2}}\sup_{Q}u_{n}^{2}+R^{d}\sup_{Q}u_{n}.

It remains to obtain the lower bound for uQ,uQ\left\langle u^{Q},\,u^{Q}\right\rangle. First, the Moser-Harnack inequality (A.41) implies that

n3Qun2Rd(cHsupQun2R4).\sum_{n\in 3Q}\,u^{2}_{n}\geq R^{d}\Bigl{(}c_{H}\sup_{Q}u^{2}_{n}-R^{4}\Bigr{)}.

Then we apply the scaling condition (1.7) twice, to write

n3Qun2CS(nQun2+Rd+4)\displaystyle\sum_{n\in 3Q}\,u^{2}_{n}\leq C_{S}\Bigl{(}\sum_{n\in Q}u_{n}^{2}+R^{d+4}\Bigr{)}\leq CS(CS(nQ/3un2+(R/3)d+4)+Rd+4)\displaystyle\,C_{S}\left(C_{S}\Bigl{(}\sum_{n\in Q/3}u_{n}^{2}+(R/3)^{d+4}\Bigr{)}+R^{d+4}\right)
\displaystyle\leq CS2nQ/3un2+CSRd+4.\displaystyle\,C_{S}^{2}\sum_{n\in Q/3}u_{n}^{2}+C_{S}^{\prime}R^{d+4}.

Hence,

nQ/3un2CS2cHRd(supQun2C′′R4),\sum_{n\in Q/3}u_{n}^{2}\geq C_{S}^{-2}c_{H}R^{d}\Bigl{(}\sup_{Q}u^{2}_{n}-C^{\prime\prime}R^{4}\Bigr{)},

where C′′C^{\prime\prime} depends on dd, cHc_{H} and CSC_{S}. By the choice of RR in (3.24),

(12C′′)14μ12R2(12C′′)14μ12,\left(\frac{1}{2C^{\prime\prime}}\right)^{\frac{1}{4}}\mu^{-\frac{1}{2}}\leq R\leq 2\left(\frac{1}{2C^{\prime\prime}}\right)^{\frac{1}{4}}\mu^{-\frac{1}{2}},

and by definition for Q′′Q\in{\mathcal{F}}^{\prime\prime}, supQun2μ2\sup_{Q}u^{2}_{n}\geq\mu^{-2}. Then we have

12supQun212μ212(2C′′)R4=C′′R4.\frac{1}{2}\sup_{Q}u^{2}_{n}\geq\frac{1}{2}\mu^{-2}\geq\frac{1}{2}(2C^{\prime\prime})R^{4}=C^{\prime\prime}R^{4}.

Therefore,

uQ,uQ=nQ(unχnQ)2nQ/3un2\displaystyle\left\langle u^{Q},\,u^{Q}\right\rangle\,=\,\sum_{n\in Q}\,(u_{n}\,\chi_{n}^{Q})^{2}\,\geq\,\sum_{n\in Q/3}\,u_{n}^{2}\,\geq\, CS2cHRd(supQun2C′′R4)\displaystyle\,C_{S}^{-2}\,c_{H}\,\,R^{d}\,\,\Bigl{(}\sup_{Q}u^{2}_{n}-C^{\prime\prime}R^{4}\Bigr{)}
\displaystyle\,\geq\, CS2cHRd(12supQun2).\displaystyle\,C_{S}^{-2}\,c_{H}\,\,R^{d}\,\,\Bigl{(}\,\frac{1}{2}\,\sup_{Q}u^{2}_{n}\Bigr{)}.

Putting the upper and lower bounds together, we have that

uQ,HuQuQ,uQ\displaystyle\frac{\left\langle u^{Q},\,Hu^{Q}\right\rangle}{\left\langle u^{Q},\,u^{Q}\right\rangle}\,\leq\,  9dRd2supQun2+RdsupQun12RdCS2cHsupQun2\displaystyle\,\frac{\,{9d}\,{R^{d-2}}\,\sup_{Q}\,u_{n}^{2}\,+\,R^{d}\,\,\sup_{Q}u_{n}}{\frac{1}{2}\,R^{d}\,C_{S}^{-2}\,c_{H}\,\sup_{Q}u^{2}_{n}}
=\displaystyle\,=\, C1R2+C2minQ1unC1(42C′′μ)+C2μ:=C3μ,\displaystyle\,C^{\prime}_{1}\,R^{-2}\,+\,C^{\prime}_{2}\,\min_{Q}\frac{1}{u_{n}}\leq\,C^{\prime}_{1}\,(4\sqrt{2C^{\prime\prime}}\,\mu\,)\,+C^{\prime}_{2}\,\mu\,:=\,C_{3}\,\mu\,,

where C3= 4C1 2C′′+C2C_{3}=\,4\,C^{\prime}_{1}\,\sqrt{\,2C^{\prime\prime}}\,+C^{\prime}_{2} depends only on cH,CSc_{H},C_{S} and the dimension dd. Therefore,

Card{eigenvaluesλofHsuchthatλC3μ}Card(′′).{\rm Card}\left\{\,{\textrm{eigenvalues}}\ \lambda\ {\rm of}\ H\ {\rm such\ that}\ \lambda\leq C_{3}\mu\,\right\}\geq{\rm Card}\left({\mathcal{F}}^{\prime\prime}\right).

Notice that in the definition of ′′{\mathcal{F}}^{\prime\prime}, the side length of the cube RR is smaller than the side length μ1/2\left\lceil\mu^{-1/2}\right\rceil required in the definition of Nu(μ)N_{u}(\mu). The box counting using 𝒫(R;Λ){\mathcal{P}}(R;\Lambda) can be bounded from below by the box counting using 𝒫(μ1/2;Λ){\mathcal{P}}(\left\lceil\mu^{-1/2}\right\rceil;\Lambda), due to Remark 3.1. The above estimates also require 3QΛ3Q\subset\Lambda, i.e., Kμ1/2K\gtrsim\mu^{-1/2}. The restriction on KK can be removed exactly in the same way as for the non-scaling case, by multiplying counting functions by a dimensional constant c~\widetilde{c}. Therefore, we obtain N(C3μ)c~Nu(μ)N(C_{3}\mu)\geq\,\widetilde{c}N_{u}(\mu) for all 0<μ<19(12C′′)12:=b0<\mu<\frac{1}{9}\left(\frac{1}{2C^{\prime\prime}}\right)^{-\frac{1}{2}}:=b. Equivalently, one has that

(3.26) N(μ)c~Nu(C31μ),forall 0<μ<C3b.N(\mu)\geq\,\widetilde{c}N_{u}(C^{-1}_{3}\mu),\ {\rm for\ all}\ 0<\mu<\,C_{3}\,b.

Next, we consider μb\mu\geq b and R=12C′′)1/4μ1/23R=\left\lceil\frac{1}{2C^{\prime\prime}})^{1/4}\mu^{-1/2}\right\rceil\leq 3. In this case, we construct ′′{\mathcal{F}}^{\prime\prime} and S′′S^{\prime\prime} using cubes of side length (Q)=R~=9\ell(Q)=\widetilde{R}=9 and tweak the dimensional constants in the counting in the end by Remark 3.1. The upper bound for uQ,HuQ\left\langle u^{Q},\,Hu^{Q}\right\rangle remains the same as in (3.25)

(3.27) uQ,HuQa(supQun2+supQun)a(μ2+μ1),\left\langle u^{Q},\,Hu^{Q}\right\rangle\leq\,a\,(\sup_{Q}u_{n}^{2}+\sup_{Q}u_{n})\leq\,a(\mu^{-2}+\mu^{-1}),

for some dimensional constant a>0a>0. For the lower bound on uQ,uQ\left\langle u^{Q},\,u^{Q}\right\rangle, we apply the Harnack inequality (A.43) to unu_{n} on QQ. We obtain

(3.28) uQ,uQQ/3un2infQun2cVsupQun2cVμ2,\left\langle u^{Q},\,u^{Q}\right\rangle\geq\,\sum_{Q/3}u_{n}^{2}\geq\,\inf_{Q}u_{n}^{2}\geq\,c_{V}\,\sup_{Q}\,u_{n}^{2}\geq\,c_{V}\,\mu^{-2},

for some constant cVc_{V} depending on VmaxV_{\max}. Therefore,

(3.29) uQ,HuQuQ,uQacV(1+μ)acV(b1+1)μ:=cVμ,\frac{\left\langle u^{Q},\,Hu^{Q}\right\rangle}{\left\langle u^{Q},\,u^{Q}\right\rangle}\leq\frac{a}{c_{V}}(1+\mu)\leq\frac{a}{c_{V}}\,(b^{-1}+1)\mu:=c_{V}^{\prime}\,\mu,

for some constant cVc^{\prime}_{V} depending on dd, VmaxV_{\max} and C′′C^{\prime\prime}. Then

N(cVμ)c~Nu(μ),forallμb.N(c_{V}^{\prime}\,\mu)\geq\,\widetilde{c}N_{u}(\mu),\ {\rm for\ all}\ \mu\geq b.

Equivalently, one has

(3.30) N(μ)c~Nu(cV1μ),forallμcVb.N(\mu)\geq\,\widetilde{c}N_{u}(c^{\prime-1}_{V}\mu),\ {\rm for\ all}\ \mu\geq\,c_{V}^{\prime}\,b.

Clearly, in (3.26), we can make C3cVC_{3}\geq c_{V}^{\prime} so that the estimates (3.26) and (3.30) cover all μ>0\mu>0. This completes the proof of Theorem 3.

\Box

3.4. Lower bound for the periodic potential

In this part, we prove Corollary 1 for a d{\mathbb{Z}}^{d} periodic potential V={vn}V=\{v_{n}\}. It is enough to show that the landscape function uu associated with the periodic potential satisfies the scaling condition (1.7). Informally, this is quite obvious. Indeed, at small scales (below pmax=maxdpdp_{\max}=\max_{d}p_{d}) we simply use the Harnack inequality. The emerging constant is roughly of the order of VmaxCpmaxV_{\max}^{\,Cp_{\max}} then. At large scales we simply use periodicity to reduce to small scales. Here are the details.

Suppose V={vn}V=\{v_{n}\} is d{\mathbb{Z}}^{d} periodic with period p=(p1,,pd)\vec{p}=(p_{1},\cdots,p_{d}). Let Γ=1,p1××1,pdd\Gamma=\llbracket 1,p_{1}\rrbracket\times\cdots\times\llbracket 1,p_{d}\rrbracket\subset{\mathbb{Z}}^{d} be the fundamental cell of VV. Notice that the condition piK,i=1,dp_{i}\mid K,i=1\cdots,d, guarantees that Λ=(/K)d\Lambda=({\mathbb{Z}}/K{\mathbb{Z}})^{d} contains finitely many copies of Γ\Gamma. Let HΓH_{\Gamma} be the restriction of HH on Γ\Gamma with the periodic boundary conditions, and let uΓ={unΓ}nΓu^{\Gamma}=\{u^{\Gamma}_{n}\}_{n\in\Gamma} be the landscape function for HΓH_{\Gamma}, i.e., (HΓuΓ)n=1,nΓ(H_{\Gamma}u^{\Gamma})_{n}=1,n\in\Gamma. By the uniqueness of the landscape function (Theorem 5), u={un}nΛu=\{u_{n}\}_{n\in\Lambda} will be the periodic extension of uΓ={unΓ}nΓu^{\Gamma}=\big{\{}u^{\Gamma}_{n}\big{\}}_{n\in\Gamma} to the entire domain Λ\Lambda.

For ss\in\mathbb{N} and aΛa\in\Lambda, let Q(s)=a+1,sdQ(s)=a+\llbracket 1,s\rrbracket^{d} be cube in Λ\Lambda of side length ss. Consider Γ+pd\Gamma+\,\vec{p}\,{\mathbb{Z}}^{d}, a collection of disjoint translations (copies) of the fundamental cell by pd\vec{p}\,{\mathbb{Z}}^{d}. Suppose s>pmax:=max{p1,,pd}s>p^{\max}:=\max\{p_{1},\cdots,p_{d}\}. It is easy to verify that the maximal number of copies of Γ\Gamma in the collection Γ+pd\Gamma+\,\vec{p}\,{\mathbb{Z}}^{d} which lie inside Q(s)Q(s) (we will call their union T1T_{1}) and the minimal number of copies of Γ\Gamma in the collection Γ+pd\Gamma+\,\vec{p}\,{\mathbb{Z}}^{d} which cover 3Q(s)3Q(s) (we will call their union T2T_{2}) differs by a dimensional multiplicative constant. Therefore, denoting the number of copies of Γ\Gamma in T1T_{1} by tt and using the periodicity of u={un}u=\{u_{n}\} with respect to all translations of Γ\Gamma, one has,

Q(s)un2T1un2=tΓ(unΓ)2,and3Q(s)un2T2un2CtΓ(unΓ)2,\sum_{Q(s)}\,u_{n}^{2}\,\geq\,\sum_{T_{1}}\,u_{n}^{2}\,=\,t\,\sum_{\Gamma}\,\big{(}u^{\Gamma}_{n}\big{)}^{2},\quad\mbox{and}\quad\sum_{3Q(s)}\,u_{n}^{2}\,\leq\,\sum_{T_{2}}\,u_{n}^{2}\,\leq\,Ct\,\sum_{\Gamma}\,\big{(}u^{\Gamma}_{n}\big{)}^{2},

which shows the scaling condition (1.7) is true for relatively large cubes.

If s<pmaxs<p^{\max}, we simply observe that the landscape function uu satisfies (Δu)n+vnun0,-(\Delta u)_{n}+v_{n}u_{n}\geq 0, and (Δu)n1nΛ.-(\Delta u)_{n}\leq 1\,\ \ n\in\Lambda. A combination of the Moser-Harnack inequality (A.41) and the Harnack inequality (A.43) implies that for some constant CC depending on dd and VmaxV_{\max}, one has

sup3Q(s)un2CsinfQ(s)un2CsinfQ(s)/3un2CssupQ(s)/3un2Cs(cH1(s/3)dQ(s)un2+cH1(s/3)4).\sup_{3Q(s)}\,u_{n}^{2}\,\leq\,C^{s}\inf_{Q(s)}u_{n}^{2}\,\leq\,C^{s}\inf_{Q(s)/3}u_{n}^{2}\,\leq\,C^{s}\sup_{Q(s)/3}\,u_{n}^{2}\leq\,C^{s}\Bigl{(}c_{H}^{-1}(s/3)^{-d}\,\sum_{Q(s)}u_{n}^{2}\,+\,c_{H}^{-1}(s/3)^{4}\Bigr{)}.

Since CsCpmaxC^{s}\leq C^{p^{\max}}, one has 3Q(s)un2(3s)dsup3Q(s)un2C~(Q(s)un2+sd+4),\sum_{3Q(s)}u_{n}^{2}\leq(3s)^{d}\sup_{3Q(s)}u_{n}^{2}\leq\widetilde{C}\left(\sum_{Q(s)}u_{n}^{2}+s^{d+4}\right), where the constant C~\widetilde{C} only depends on the dimension, VmaxV_{\max} and pmaxp^{\max}.

Therefore, for all cubes Q(s)Q(s), {un2}nΛ\{u_{n}^{2}\}_{n\in\Lambda} satisfies the scaling condition (1.7). The estimates for NN in the periodic case then follow directly from Theorem 3. \Box

4. Landscape law for the Anderson model

There must be changes propagating from changes in the previous section, at treatment of small scales. I do not bother about it for now. In the present section we will concentrate on the Anderson model. To this end, we consider V={vn}nΛV=\{v_{n}\}_{n\in\Lambda} with the values given by independent, identically distributed (i.i.d.) random variables, with common probability measure P0P_{0} on \mathbb{R}, subject to the conditions stated in the beginning of Section 1.2. In particular, denoting by F(δ)=P0(vnδ)F(\delta)=P_{0}(v_{n}\leq\delta) the common cumulative distribution function of vnv_{n}, we have F(δ)=0F(\delta)=0 if δ<0\delta<0, F(δ)=1F(\delta)=1 if δVmax\delta\geq V_{\max}, and there is a δ>0\delta_{\ast}>0, such that

(4.1) 0<F(δ)F(δ):=F<1forall 0<δδ.0<F(\delta)\leq F(\delta_{\ast}):=F_{\ast}<1\ \ {\rm for\ all}\ \ 0<\delta\leq\delta_{\ast}.

We note that δ\delta_{\ast} can be picked to be less than min(1,Vmax/2)\min(1,V_{\max}/2) since infsuppP0=0\inf{\rm supp}P_{0}=0. Hence FVmax/2𝔼(vn)FVmax/2+VmaxF_{\ast}V_{\max}/2\leq\mathbb{E}(v_{n})\leq F_{\ast}V_{\max}/2+V_{\max}. Some constants used in the proof of this section receive their dependence on 𝔼(vn)\mathbb{E}(v_{n}) through δ\delta_{\ast} and FF_{\ast}.

4.1. Estimates for NuN_{u} in the Anderson model

We will study the following tail estimates for NuN_{u} first.

Theorem 7.

Let V={vn}nΛV=\{v_{n}\}_{n\in\Lambda} be an Anderson-type potential as above. Then there are dimensional constants c3,c4,γ1,K>0c_{3},c_{4},\gamma_{1},K_{\ast}>0 such that

(4.2) 𝔼(Nu(μ))c4μd/2F(c3μ)γ1μd/2forallK/K2μ1.\mathbb{E}\left(N_{u}(\mu)\right)\,\geq\,c_{4}\,\mu^{d/2}\,F(c_{3}\mu)^{\,\gamma_{1}\,\mu^{-d/2}}\ \ {\rm for\ all}\ K_{\ast}/K^{2}\leq\mu\leq 1.

Furthermore, there are constants C3,C4,γ2C_{3},C_{4},\gamma_{2}, and μ>0\mu_{\ast}>0 depending on d,δ,Fd,\delta_{\ast},F_{\ast} only, such that

(4.3) 𝔼(Nu(μ))C4μd/2F(C3μ)γ2μd/2forallμ<μ.\mathbb{E}\left(N_{u}(\mu)\right)\,\leq\,C_{4}\,\mu^{d/2}\,F(C_{3}\mu)^{\,\gamma_{2}\,\mu^{-d/2}}\ \ {\rm for\ all}\ \mu<\mu_{\ast}.
Remark 4.1.

All the constants are independent of VmaxV_{\max}.

After we establish these tail estimates for NuN_{u}, we will combine them with the deterministic result Theorem 1,2 to prove Theorem 4, and (1.11).

Proof.  The proof of (4.2). For 0<μ10<\mu\leq 1, let r=μ12.r=\left\lceil\mu^{-\frac{1}{2}}\right\rceil. Let 𝒫(r;Λ){\mathcal{P}}(r;\Lambda) be partition of size rr as usual. It is enough to assume that K=K0rK=K_{0}r for some K0K_{0}\in\mathbb{N}, otherwise the counting can always be bounded from below by ignoring the irregular boxes in the last rows/columns of 𝒫(r;Λ){\mathcal{P}}(r;\Lambda). For all cubes Q𝒫=𝒫(r;Λ)Q\in{\mathcal{P}}={\mathcal{P}}(r;\Lambda), let ζQ=1\zeta_{Q}=1 if minnQ1unμ\min_{n\in Q}\frac{1}{u_{n}}\,\leq\mu and ζQ=0\zeta_{Q}=0 otherwise. Direct computation shows that

𝔼(Nu(μ))=\displaystyle\mathbb{E}\left(N_{u}(\mu)\right)= 1Kd𝔼(Card{Q𝒫(r):minnQ1unμ})\displaystyle\frac{1}{K^{d}}\,\mathbb{E}\left(\,{\rm Card}\left\{\,Q\in{\mathcal{P}}(r)\,:\ \ \min_{n\in Q}\frac{1}{u_{n}}\,\leq\mu\,\right\}\,\right)
(4.4) =\displaystyle= 1Kd𝔼(Q𝒫(r)ζQ)=1KdQ𝒫(r)𝔼(ζQ)=1K0drdQ𝒫(r)(minnQ1unμ).\displaystyle\frac{1}{K^{d}}\,\mathbb{E}\left(\sum_{Q\in{\mathcal{P}}(r)}\zeta_{Q}\right)=\frac{1}{K^{d}}\,\sum_{Q\in{\mathcal{P}}(r)}\mathbb{E}\left(\zeta_{Q}\right)=\frac{1}{K^{d}_{0}r^{d}}\sum_{Q\in{\mathcal{P}}(r)}{\mathbb{P}}\left(\,\min_{n\in Q}\frac{1}{u_{n}}\,\leq\mu\,\right).

For each Q𝒫Q\in{\mathcal{P}}, consider translations of QQ by the vectors kreikre_{i} for all 1id1\leq i\leq d directions and |k|m|k|\leq m. Here mm is some large integer that will be specified later. Let m=m(Q){\mathcal{B}}_{m}={\mathcal{B}}_{m}(Q) be the union of these translated cubes,

m=|k|m,1id(Q+krei).{\mathcal{B}}_{m}=\bigcup_{|k|\leq m,1\leq i\leq d}\left(Q+kre_{i}\right).

Similarly to (A.7), one can construct a discrete cut-off function χ={χn}Kd\chi=\{\chi_{n}\}\in\mathcal{H}\cong\mathbb{R}^{K^{d}}, supported on 2m{\mathcal{B}}_{2m}, and satisfying

χn=1,nm,andχn=0,n2m,\displaystyle\chi_{n}=1,\ \ n\in{\mathcal{B}}_{m},\ \ \ \ {\rm and}\ \ \chi_{n}=0,\ \ n\notin{\mathcal{B}}_{2m},
0χn1,n2m\m,\displaystyle 0\leq\chi_{n}\leq 1,\,n\in{\mathcal{B}}_{2m}\backslash{\mathcal{B}}_{m},
|iχn|=|χn+eiχn|=0,n2m,\displaystyle|\nabla_{i}\chi_{n}|=|\chi_{n+e_{i}}-\chi_{n}|=0,\,n\notin{\mathcal{B}}_{2m},
|iχn|=|χn+eiχn|<1m(Q)<1mμ12,n2m, 1id.\displaystyle|\nabla_{i}\chi_{n}|=|\chi_{n+e_{i}}-\chi_{n}|<\frac{1}{m\,\ell(Q)}<\frac{1}{m}\,\mu^{\frac{1}{2}},\ \ n\in{\mathcal{B}}_{2m},\ 1\leq i\leq d.

By the landscape uncertainty principle (2.9), one has

minm1un1|m|2mχn2+1|m|2mvnχn22ddm2μ+ 2dmax2mvn12μ+ 2dmax2mvn,\min_{{\mathcal{B}}_{m}}\frac{1}{u_{n}}\leq\,\frac{1}{|{\mathcal{B}}_{m}|}\sum_{{\mathcal{B}}_{2m}}\|\nabla\chi_{n}\|^{2}+\frac{1}{|{\mathcal{B}}_{m}|}\sum_{{\mathcal{B}}_{2m}}v_{n}\chi_{n}^{2}\leq\,\frac{2^{d}d}{m^{2}}\,\mu\,+\,2^{d}\,\max_{{\mathcal{B}}_{2m}}\,v_{n}\leq\,\frac{1}{2}\,\mu\,+\,2^{d}\,\max_{{\mathcal{B}}_{2m}}v_{n},

provided m22d+1dm^{2}\geq 2^{d+1}d. Therefore, for all Q𝒫(μ1/2;Λ)Q\in{\mathcal{P}}\bigl{(}\left\lceil\mu^{-1/2}\right\rceil;\Lambda\bigr{)},

{minnm1unμ}{maxn2mvn12d+1μ}=(F(cμ))|2m|(F(cμ))Cμd/2,\mathbb{P}\left\{\min_{n\in{\mathcal{B}}_{m}}\frac{1}{u_{n}}\,\leq\mu\right\}\geq\mathbb{P}\left\{\max_{n\in{\mathcal{B}}_{2m}}v_{n}\leq\frac{1}{2^{d+1}}\,\mu\right\}=\big{(}F\left(c\mu\right)\big{)}^{|{\mathcal{B}}_{2m}|}\geq\big{(}F\left(c\mu\right)\big{)}^{\,C\,\mu^{-d/2}},

where c=2d1c={2^{-d-1}} and C=(4m+1)dC=(4m+1)^{d}. On the other hand, notice that all the translations Q+kreiQ+kre_{i} still belong to 𝒫(r){\mathcal{P}}(r). Then we can rewrite m=m(Q){\mathcal{B}}_{m}={\mathcal{B}}_{m}(Q) as m(Q)=Q𝒫(r)m(Q)Q,{\mathcal{B}}_{m}(Q)=\bigcup_{Q^{\prime}\in{\mathcal{P}}(r)\cap{\mathcal{B}}_{m}(Q)}Q^{\prime}\,, which implies that for all Q𝒫(r)Q\in{\mathcal{P}}(r)

Q𝒫(r)m(Q){minnQ1unμ}{minnm(Q)1unμ}(F(cμ))Cμd/2.\sum_{Q^{\prime}\in{\mathcal{P}}(r)\cap{\mathcal{B}}_{m}(Q)}\mathbb{P}\Bigl{\{}\min_{n\in Q^{\prime}}\,\frac{1}{u_{n}}\,\leq\mu\Bigr{\}}\geq\mathbb{P}\Bigl{\{}\min_{n\in{\mathcal{B}}_{m}(Q)}\,\frac{1}{u_{n}}\,\leq\mu\Bigr{\}}\geq\,\bigl{(}F\left(c\mu\right)\bigr{)}^{\,C\,\mu^{-d/2}}.

Summing the left-hand side of the above inequality over all Q𝒫(r)Q\in{\mathcal{P}}(r), one has

Q𝒫(r)Q𝒫(r)m(Q){minnQ1unμ}=(2m+1)dQ𝒫(r){minnQ1unμ}.\sum_{Q\in{\mathcal{P}}(r)}\sum_{Q^{\prime}\in{\mathcal{P}}(r)\cap{\mathcal{B}}_{m}(Q)}\mathbb{P}\left\{\min_{n\in Q^{\prime}}\frac{1}{u_{n}}\,\leq\mu\right\}=(2m+1)^{d}\sum_{Q\in{\mathcal{P}}(r)}\mathbb{P}\left\{\min_{n\in Q}\frac{1}{u_{n}}\,\leq\mu\right\}.

Combining this together with (4.4), one has

𝔼(Nu(μ))\displaystyle\mathbb{E}\left(N_{u}(\mu)\right)\geq (2m+1)dK0drdQ𝒫(r)(Q𝒫(r)m(Q){minnQ1unμ})\displaystyle\,\frac{(2m+1)^{-d}}{K^{d}_{0}\,r^{d}}\,\sum_{Q\in{\mathcal{P}}(r)}\,\left(\sum_{Q^{\prime}\in{\mathcal{P}}(r)\cap{\mathcal{B}}_{m}(Q)}\mathbb{P}\left\{\min_{n\in Q^{\prime}}\frac{1}{u_{n}}\,\leq\mu\right\}\right)
\displaystyle\geq (2m+1)dK0drdQ𝒫(r)(F(cμ))Cμd/2(2m+1)d 2dμd2(F(cμ))Cμd/2.\displaystyle\,\frac{(2m+1)^{-d}}{K^{d}_{0}\,r^{d}}\,\sum_{Q\in{\mathcal{P}}(r)}\,\big{(}F\left(c\mu\right)\big{)}^{\,C\,\mu^{-d/2}}\geq\,(2m+1)^{-d}\,2^{-d}\,\mu^{\frac{d}{2}}\,\big{(}F\left(c\mu\right)\big{)}^{\,C\,\mu^{-d/2}}.

Note that we also need to impose the condition on the size of domain so that 2mΛ{\mathcal{B}}_{2m}\subset\Lambda, i.e., K(2m+1)r=Cμ1/2K\geq(2m+1)r=C^{\prime}\mu^{-1/2}.

The proof of (4.3). Using (4.4), it is enough to bound {minnQ1unμ}\mathbb{P}\bigl{\{}\min_{n\in Q}\frac{1}{u_{n}}\,\leq\mu\bigr{\}} from above since

(4.5) 𝔼(Nu(μ))1μ1/2dmaxQ𝒫(μ1/2){minnQ1unμ}.\mathbb{E}\left(N_{u}(\mu)\right)\leq\,\frac{1}{\left\lceil\mu^{-1/2}\right\rceil^{d}}\,\max_{Q\in{\mathcal{P}}(\left\lceil\mu^{-1/2}\right\rceil)}\,\mathbb{P}\left\{\min_{n\in Q}\,\frac{1}{u_{n}}\,\leq\mu\ \right\}.

This will be the most delicate part. We need several technical lemmas concerning the growth of the landscape function. Some of these estimates may have independent interest in the landscape theory.

For any r3r\geq 3, let BΛB\subset\Lambda be cube of side length (B)=r\ell(B)=r and let Bˇ=B/3\check{B}=B/3 be the middle third cube as defined in (2.13). We are going to show that there is a suitable MM (large, and only depending on the expectation of the random variable), such that for any μ\mu (small enough, depending on 𝔼(vn)\mathbb{E}(v_{n})), and any cube BB of side length r=(4Mμ)1/2r=\left\lceil(4M\mu)^{-1/2}\right\rceil

(4.6) {minnBˇ1unμ}A1Md/2F(A2Mμ)(Mμ)d/2/2,\mathbb{P}\Bigl{\{}\min_{n\in\check{B}}\frac{1}{u_{n}}\,\leq\mu\Bigr{\}}\leq\,A_{1}M^{d/2}F(A_{2}M\mu)^{\,(M\mu)^{-d/2}/2},

for some suitable constants A1,A2A_{1},A_{2} (depending only on the dimension and 𝔼(vn)\mathbb{E}(v_{n}), and independent of μ\mu).

We start from the following deterministic statement. The lemma states that the landscape function uu is forced to grow at a certain rate if VV is reasonably non-degenerate.

Lemma 4.1.

Let u={un}u=\{u_{n}\} be the landscape function given by Theorem 5. Let BΛB\subset\Lambda be a cube of side length r:=(B)3r:=\ell(B)\geq 3, and let Bˇ=B/3\check{B}=B/3 be the middle third as usual. For any 0<λ<10<\lambda<1, there is ε0(d,λ)>0\varepsilon_{0}(d,\lambda)>0 such that for all 0<ε<ε00<\varepsilon<\varepsilon_{0}, there are constants CP=CP(ε,λ,d)>0C_{P}=C_{P}(\varepsilon,\lambda,d)>0, M=M(ε,λ,d)>0M=M(\varepsilon,\lambda,d)>0 and r=r(ε,λ,d)>0r_{\ast}=r_{\ast}(\varepsilon,\lambda,d)>0, such that the following statement holds. If BB satisfies conditions

(i):
(4.7) Card{jB:vjCPr2}λ|B|,{\rm Card}\left\{\,j\in B:v_{j}\geq C_{P}r^{-2}\,\right\}\geq\lambda|B|,
(ii):

there is a ξBˇ\xi\in\check{B} such that

(4.8) uξMr2,u_{\xi}\geq Mr^{2},

then for all rrr\geq r_{\ast}, there is a ξΛ\xi^{\prime}\in\Lambda such that |ξξ|1+εr|\xi^{\prime}-\xi|_{\infty}\leq\left\lfloor\sqrt{1+\varepsilon}r\right\rfloor and

(4.9) uξ(1+ε)uξM1+εr2.u_{\xi^{\prime}}\geq(1+\varepsilon)\,u_{\xi}\geq M\left\lfloor\sqrt{1+\varepsilon}r\right\rfloor^{2}.
Remark 4.2.

This lemma holds for any λ\lambda, and works for any cube BB of side length rr. The choice of CPC_{P}, ε\varepsilon and MM only depend on λ\lambda, and is independent of the choice of BB, neither on its size nor the position.

Remark 4.3.

Note that this Lemma is a completely deterministic result. It has nothing to do with the randomness (structure of vnv_{n}). It can be applied to any VV and uu as long as (Hu)n=1(Hu)_{n}=1 locally on the lattice (containing the cube BB and its neighborhood). In our proof, Lemma 4.1 will lead to some important probability estimates. The small parameter λ\lambda will be picked at the very end when we are about to prove (4.6).

We need some technical preparations for the average of unu_{n}. We will frequently write u(n)=unu(n)=u_{n} to make the notations of the sub-index easier to read. For ξ=(ξ1,,ξd)Λ\xi=(\xi_{1},\cdots,\xi_{d})\in\Lambda, and r0r\in{\mathbb{Z}}_{\geq 0}, we denote by Q(r;ξ)Q(r;\xi) the box centered at ξ\xi of side length 2r+12r+1 :

Q(r;ξ)={m=(m1,,md)d:|mξ|r}.Q(r;\xi)=\left\{m=(m_{1},\cdots,m_{d})\in{\mathbb{Z}}^{d}:\,|m-\xi|_{\infty}\leq r\right\}.

We will omit the center ξ\xi (fixed) and write Q(r)=Q(r;ξ)Q(r)=Q(r;\xi) when it is clear. We denote by Q(r)Q(r)\partial Q(r)\subset Q(r) the inner boundary of Q(r)Q(r) as defined in (2.11), and by Q(r)Q(r)\partial^{\circ}Q(r)\subset\partial Q(r) the boundary removing the “corners” as defined in (2.12). For r=0r=0, we have the “degenerate cube” Q(0;ξ)=Q(0;ξ)={ξ}Q(0;\xi)\,=\,\partial Q(0;\xi)\,=\,\{\xi\}. Notice that Q(r)=ρ=0rQ(ρ)Q(r)=\cup_{\rho=0}^{r}\partial Q(\rho). Let ara_{r} be the average of unu_{n} on Q(r)\partial Q(r) with respect to the (discrete) Poisson kernel (see the definition and properties of PrP_{r} in (A.12) in Appendix A.4):

(4.10) ar=nQ(r;ξ)Pr(ξ,n)un,r1,a0=uξ,a_{r}=\sum_{n\in\partial Q(r;\xi)}P_{r}(\xi,n)\,u_{n},\quad r\geq 1,\quad a_{0}=u_{\xi},\,

in other words, a harmonic function with data uu on the boundary. Let ArA_{r} be the corresponding weighted average of unu_{n} on Q(r)Q(r):

(4.11) Ar=1|Q(r)|ρ=0r|Q(ρ)|aρ=1|Q(r)|nQ(r)pnun,A_{r}\,=\,\frac{1}{|Q(r)|}\,\sum_{\rho=0}^{r}\,|\partial Q(\rho)|\,a_{\rho}\,=\,\frac{1}{|Q(r)|}\,\sum_{n\in Q(r)}\,p_{n}\,u_{n},

where for any nQ(r)n\in Q(r), and ρ=|nξ|\rho=|n-\xi|_{\infty}

(4.12) pn=|Q(ρ)|Pρ(ξ,n).p_{n}=|\partial Q(\rho)|\,P_{\rho}(\xi,n).

By the properties of the discrete Poisson kernel, one has

(4.13) nQ(r;ξ)Pr(ξ,n)=1nQ(r;ξ)pn=|Q(r)|.\sum_{n\in\partial Q(r;\xi)}P_{r}(\xi,n)=1\Longrightarrow\sum_{n\in Q(r;\xi)}p_{n}=|Q(r)|.

The first two estimates are lower bounds on ara_{r} and ArA_{r}.

Lemma 4.2.

There is dimensional constant CC, such that for any ξΛ\xi\in\Lambda and r1r\geq 1

(4.14) aruξr2,\displaystyle a_{r}\geq u_{\xi}-r^{2},
(4.15) AruξCr2.\displaystyle A_{r}\geq u_{\xi}-Cr^{2}.
Proof.

Let u~\tilde{u} be the landscape function for the free Laplacian on Q(r)Q(r) with zero Dirichlet boundary condition:

{(Δu)n=1,nQ(r1),(u)n=0,nQ(r).\displaystyle\begin{cases}-(\Delta u^{\prime})_{n}=1,\,n\in Q(r-1),\\ (u^{\prime})_{n}=0,\,n\in\partial Q(r).\end{cases}

Let un′′=12r212di=1d(niξi)2u^{\prime\prime}_{n}=\frac{1}{2}r^{2}-\frac{1}{2d}\sum_{i=1}^{d}(n_{i}-\xi_{i})^{2}. Direct computation shows that

{(Δu′′)n=1,nQ(r1),(u′′)n0,nQ(r).\displaystyle\begin{cases}-(\Delta u^{\prime\prime})_{n}=1,n\in Q(r-1),\\ (u^{\prime\prime})_{n}\geq 0,n\in\partial Q(r).\end{cases}

By the maximum principle (Lemma A.1), one has for all nQ(r)n\in Q(r), unun′′r2{u^{\prime}}_{n}\leq u^{\prime\prime}_{n}\leq r^{2}. Let wnw_{n} be the harmonic function on Q(r)Q(r) with the boundary data equal to unu_{n}, i.e.,

{(Δw)n=0,nQ(r1),(w)n=un,nQ(r).\displaystyle\begin{cases}-(\Delta w)_{n}=0,\,n\in Q(r-1),\\ (w)_{n}=u_{n},\,n\in\partial Q(r).\end{cases}

Then by the Poisson integral formula (A.15), wξ=nQ(r;ξ)Pr(ξ,n)un=arw_{\xi}=\sum_{n\in\partial Q(r;\xi)}P_{r}(\xi,n)u_{n}=a_{r}. On the other hand,

{(Δ(w+uu))n=vnun0,nQ(r1),(w+uu)n=0,nQ(r).\displaystyle\begin{cases}-(\Delta(w+u^{\prime}-u))_{n}=v_{n}u_{n}\geq 0,\,n\in Q(r-1),\\ (w+u^{\prime}-u)_{n}=0,\,n\in\partial Q(r).\end{cases}

Therefore, (w+uu)n0(w+u^{\prime}-u)_{n}\geq 0 for all nQ(r)n\in Q(r). In particular, uξwξ+uξar+r2,u_{\xi}\leq w_{\xi}+u^{\prime}_{\xi}\leq a_{r}+r^{2}, proving (4.14), and a similar statement is true for 1ρr11\leq\rho\leq r-1, so that

|Q(ρ)|uξ|Q(ρ)|nQ(ρ)Pρ(ξ,n)un+|Q(ρ)|ρ2nQ(ρ)pnun+Cρd+1.|\partial Q(\rho)|u_{\xi}\leq|\partial Q(\rho)|\sum_{n\in\partial Q(\rho)}P_{\rho}(\xi,n)u_{n}+|\partial Q(\rho)|\,\rho^{2}\leq\sum_{n\in\partial Q(\rho)}p_{n}u_{n}+C\rho^{d+1}.

Summing over 1ρr11\leq\rho\leq r-1, one has

|Q(r)|uξ=ρ=1r1|Q(ρ)|uξρ=1r1nQ(ρ)pnun+Cρ=1r1ρd+1nQ(r;ξ)pnun+Crd+2,|Q(r)|u_{\xi}=\sum_{\rho=1}^{r-1}|\partial Q(\rho)|u_{\xi}\leq\sum_{\rho=1}^{r-1}\sum_{n\in\partial Q(\rho)}p_{n}u_{n}+C\sum_{\rho=1}^{r-1}\rho^{d+1}\leq\sum_{n\in Q(r;\xi)}p_{n}u_{n}+Cr^{d+2},

which implies uξAr+Cr2,u_{\xi}\leq\,A_{r}+C\,r^{2}, as desired. ∎

Lemma 4.3.

For any d1d\geq 1 and 0<η<1/40<\eta<1/4 there is a constant C1=C1(d)>0C_{1}=C_{1}(d)>0 (independent of η\eta) and a constant C2=C2(η,d)>0C_{2}=C_{2}(\eta,d)>0 such that for any cube Q(r)=Q(r;ξ)Q(r)=Q(r;\xi) of side length rC1/ηr\geq\,C_{1}/\eta, there is a subset Qη(r)Q(r)Q^{\eta}(r)\subset Q(r) such that |Q(r)\Qη(r)|C1ηrd|Q(r)\backslash Q^{\eta}(r)|\leq\,C_{1}\,\eta\,r^{d} and

(4.16) pnC2fornQη(r).p_{n}\geq\,C_{2}\ {\rm for\ }n\in Q^{\eta}(r).
Remark 4.4.

The lemma is true in all dimensions. However, for d=1d=1, we actually do not need to remove any portion of the cube (an interval in {\mathbb{Z}}) to obtain (4.16) since the 1-d Poisson kernel PrP_{r} is rather trivial (constantly 1/21/2), and so is pnp_{n}.

Proof.

Write Q(r)=ρ=0rQ(ρ)Q(r)=\cup_{\rho=0}^{r}\partial Q(\rho). The estimate follows from the lower bound of Pρ(ξ,n)P_{\rho}(\xi,n) on each Q(ρ)\partial Q(\rho) as long as nn is away from the edges (and the corners). Given 0<η<10<\eta<1, according to Lemma A.9, there exist c(ρ,η)c(\rho,\eta) and ρ0(η,d)\rho_{0}(\eta,d) such that for all ρρ0\rho\geq\rho_{0}, Pρ(ξ,n)cρ1dP_{\rho}(\xi,n)\geq c\rho^{1-d} on Q(ρ)\partial Q(\rho) except for Cηρd1C\eta\rho^{d-1} many nQ(ρ)n\in\partial Q(\rho), where CC only depends on the dimension. Therefore, pn=|Q(ρ)|Pρ(ξ,n)c~>0p_{n}=|\partial Q(\rho)|P_{\rho}(\xi,n)\geq\widetilde{c}>0 on ρ=ρ0rQ(ρ)\bigcup_{\rho=\rho_{0}}^{r}\partial Q(\rho) except for ρ=ρ0rCηρd1Cηrd\sum_{\rho=\rho_{0}}^{r}C\eta\rho^{d-1}\leq C\eta r^{d} many nn. For 0ρ<ρ00\leq\rho<\rho_{0}, we then have Pρ(ξ,n)>c(ρ,d)P_{\rho}(\xi,n)>c(\rho,d) except for those nn on the edges and corners, whose total cardinality is at most Cρd2C\rho^{d-2}. This again implies pnmin0ρ<ρ0c(ρ,d)p_{n}\geq\min_{0\leq\rho<\rho_{0}}c(\rho,d) for all n0ρ<ρ0Q(ρ)n\in\bigcup_{0\leq\rho<\rho_{0}}\partial Q(\rho) except for ρ=0ρ0ρd2rd1ηrd\sum_{\rho=0}^{\rho_{0}}\rho^{d-2}\lesssim r^{d-1}\lesssim\eta r^{d} many nn. Therefore, pnC2p_{n}\geq C_{2} for some constant C2C_{2} only depending on η\eta and dd, and the cardinality of the exceptional set of nQ(r)n\in Q(r) violating this, is at most C1ηrdC_{1}\eta r^{d}. ∎

With these two technical lemmas, we are ready to the

Proof of Lemma 4.1.

Let BB and ξBˇ\xi\in\check{B} be given as in Lemma 4.1, where (B)=r\ell(B)=r. Clearly, BQ(r;ξ)B\subset Q(r;\xi). Denote by JJ the set in condition (4.7), that is, J={jB:vjCPr2},J=\{j\in B:v_{j}\geq C_{P}r^{-2}\}, where CPC_{P} will be be picked later. Fix 0<λ<10<\lambda<1, we assume that |J|λ|B|=λrd|J|\geq\lambda|B|=\lambda r^{d}. Let Ar,pnA_{r},p_{n} be defined in (4.11). Let J0={jJ:uj<12Ar}J_{0}=\{j\in J:u_{j}<\,\frac{1}{2}\,A_{r}\,\} and let

(4.17) S(J0)=nJ0pn,S(J0C)=nQ(r)\J0pn=|Q(r)|S(J0),S(J_{0})=\sum_{n\in J_{0}}p_{n},\ \ S(J^{C}_{0})=\sum_{n\in Q(r)\backslash J_{0}}p_{n}=|Q(r)|-S(J_{0})\,,

where the last equality follows from (4.13).

Now we are ready to look for ξ\xi^{\prime} satisfying (4.9) in the following two cases:

Case I: |J0|12λ|B|=12λrd.|J_{0}|\geq\frac{1}{2}\lambda|B|=\frac{1}{2}\lambda r^{d}.

Let C1C_{1} be the constant from Lemma 4.3, then pick η=min(λ4C1,1/4)\eta=\min(\frac{\lambda}{4C_{1}},1/4) and let C2=C2(d,η)C_{2}=C_{2}(d,\eta) be also the constant from Lemma 4.3. Then

|J0|=nJ0Qη(r)1+nJ0\Qη(r)1C21nJ0Qη(r)pn+C1ηrdC21nJ0pn+14λrd.|J_{0}|=\sum_{n\in J_{0}\cap Q^{\eta}(r)}1+\sum_{n\in J_{0}\backslash Q^{\eta}(r)}1\leq\,C_{2}^{-1}\sum_{n\in J_{0}\cap Q^{\eta}(r)}p_{n}+C_{1}\eta r^{d}\leq\,C_{2}^{-1}\,\sum_{n\in J_{0}}p_{n}+\frac{1}{4}\lambda r^{d}.

Therefore,

S(J0)=nJ0pnC214λrd:=c3(λ)rd.S(J_{0})=\sum_{n\in J_{0}}p_{n}\geq\,C_{2}\frac{1}{4}\lambda\,r^{d}:=c_{3}(\lambda)\,r^{d}.

Direct computation shows that

|Q(r)|Ar=nQ(r)pnun=J0pnun+Q(r)\J0pnun12ArS(J0)+Q(r)\Q0pnun.|Q(r)|\,A_{r}=\sum_{n\in Q(r)}p_{n}u_{n}=\sum_{J_{0}}p_{n}u_{n}+\sum_{Q(r)\backslash J_{0}}p_{n}u_{n}\leq\frac{1}{2}\,A_{r}\,S(J_{0})+\sum_{Q(r)\backslash Q_{0}}p_{n}u_{n}.

By the definition of S(J0C)S(J_{0}^{C}) and (4.17), this implies that

1S(J0C)Q(r)\J0pnun|Q(r)|12S(J0)|Q(r)|S(J0)Ar(1+S(J0)2|Q(r)|)Ar(1+c4(d,λ))Ar.\frac{1}{S(J_{0}^{C})}\sum_{Q(r)\backslash J_{0}}p_{n}u_{n}\geq\frac{|Q(r)|-\frac{1}{2}S(J_{0})}{|Q(r)|-S(J_{0})}A_{r}\geq\Bigl{(}1+\frac{S(J_{0})}{2|Q(r)|}\Bigr{)}A_{r}\geq\bigl{(}1+c_{4}(d,\lambda)\bigr{)}\,A_{r}.

Therefore, there is one point ξQ(r)\J0\xi^{\prime}\in Q(r)\backslash J_{0} such that uξ(1+c4)Ar.u_{\xi^{\prime}}\geq\left(1+c_{4}\right)\,A_{r}. By (4.15) and (4.8),

uξ(1+c4)(uξCdr2)(1+c4)(1CdM)uξ(1+c42)uξ,u_{\xi^{\prime}}\geq\left(1+c_{4}\right)\,(u_{\xi}-C_{d}r^{2})\geq\left(1+c_{4}\right)\,\left(1-\frac{C_{d}}{M}\right)u_{\xi}\geq\left(1+\frac{c_{4}}{2}\right)\,u_{\xi},

provided M>2c4(Cd+1):=M0(d,λ).M>\frac{2}{c_{4}}(C_{d}+1):=M_{0}(d,\lambda). Therefore, (4.9) holds for ε<c4/2:=ε0(d,λ)\varepsilon<c_{4}/2:=\varepsilon_{0}(d,\lambda).

Case II: |J0|12λrd.|J_{0}|\leq\frac{1}{2}\lambda r^{d}.

Take R=1+εrR=\left\lfloor\sqrt{1+\varepsilon}\,r\right\rfloor for some small ε\varepsilon, which gives R2r2εr2R^{2}-r^{2}\leq\varepsilon r^{2}. Let ar,aRa_{r},a_{R} and Gr,GRG_{r},G_{R} be the surface average and the Green’s function on Q(r)=Q(r;ξ),Q(R)=Q(R;ξ)Q(r)=Q(r;\xi),Q(R)=Q(R;\xi) respectively, as defined in (4.10),(A.13). Applying the discrete Green’s identity (integration by parts formula) (A.15) on Q(r)Q(r) and Q(R)Q(R), one has

u(ξ)=aRmQ(R1)GR(ξ,m)Δu(m)=armQ(r1)Gr(ξ,m)Δu(m).u(\xi)=a_{R}-\sum_{m\in Q(R-1)}G_{R}(\xi,m)\Delta u(m)=a_{r}-\sum_{m\in Q(r-1)}G_{r}(\xi,m)\Delta u(m).

Then

aRar=\displaystyle a_{R}-a_{r}= mQ(R1)\Q(r1)GR(ξ,m)Δu(m)+mQ(r1)(GR(ξ,m)Gr(ξ,m))Δu(m)\displaystyle\sum_{m\in Q(R-1)\backslash Q(r-1)}G_{R}(\xi,m)\Delta u(m)+\sum_{m\in Q(r-1)}\,\big{(}G_{R}(\xi,m)-G_{r}(\xi,m)\big{)}\,\Delta u(m)
\displaystyle\geq mQ(R1)\Q(r1)GR(ξ,m)mQ(r1)(GR(ξ,m)Gr(ξ,m))\displaystyle-\sum_{m\in Q(R-1)\backslash Q(r-1)}G_{R}(\xi,m)-\sum_{m\in Q(r-1)}\,\big{(}G_{R}(\xi,m)-G_{r}(\xi,m)\big{)}
+mQ(r1)(GR(ξ,m)Gr(ξ,m))vmum.\displaystyle\qquad\qquad\qquad\qquad\qquad+\sum_{m\in Q(r-1)}\,\big{(}G_{R}(\xi,m)-G_{r}(\xi,m)\big{)}v_{m}u_{m}.

Notice that ΔGR(ξ,)=0\Delta G_{R}(\xi,\cdot)=0 in Q(R1;ξ)\{ξ}Q(R-1;\xi)\backslash\{\xi\}. By the maximum principle, Lemma A.3, one has

maxmQ(R1)\Q(r1)GR(ξ,m)maxmQ(r)GR(ξ,m),\max_{m\in Q(R-1)\backslash Q(r-1)}G_{R}(\xi,m)\leq\max_{m\in\partial Q(r)}G_{R}(\xi,m),

where we used GR(ξ,m)=0G_{R}(\xi,m)=0 for mQ(R)m\in\partial Q(R) and the discussion for an annular region after Lemma A.3.

On the other hand, Δ(GR(ξ,)Gr(ξ,))=0\Delta\big{(}G_{R}(\xi,\cdot)-G_{r}(\xi,\cdot)\big{)}=0 in Q(r1)Q(r-1) and GR(ξ,n)Gr(ξ,n)=GR(ξ,n)G_{R}(\xi,n)-G_{r}(\xi,n)=G_{R}(\xi,n) for nQ(r)n\in\partial Q(r). Then using once again the maximum principle, Lemma A.3, for any mQ(r1)m\in Q(r-1),

minmQ(r)GR(ξ,m)GR(ξ,m)Gr(ξ,m)maxmQ(r)GR(ξ,m).\min_{m^{\prime}\in\partial Q(r)}G_{R}(\xi,m^{\prime})\leq\,G_{R}(\xi,m)-G_{r}(\xi,m)\leq\max_{m^{\prime}\in\partial Q(r)}G_{R}(\xi,m^{\prime}).

By the choice of rr and RR, Q(r)\partial Q(r) is away from both Q(R)\partial Q(R) and the pole ξ\xi. Then Lemma A.7 implies that for rr large enough (depending only on ε\varepsilon) and all mQ(r)m\in\partial Q(r),

C1r2dGR(ξ,m)C2r2d,C_{1}r^{2-d}\leq G_{R}(\xi,m^{\prime})\leq C_{2}r^{2-d},

where C1,C2C_{1},C_{2} only depend on dd and ε\varepsilon. Therefore,

aRar\displaystyle a_{R}-a_{r}\geq C3(Rdrd)r2dC4rdr2d+r2dmQ(r1)vmum\displaystyle-C_{3}(R^{d}-r^{d})r^{2-d}-C_{4}r^{d}r^{2-d}+r^{2-d}\sum_{m\in Q(r-1)}\,v_{m}u_{m}
\displaystyle\geq C5(d,ε)r2+r2dmJ\J0vmum\displaystyle-C_{5}(d,\varepsilon)r^{2}+r^{2-d}\sum_{m\in J\backslash J_{0}}\,v_{m}u_{m}
\displaystyle\geq C5(d,ε)r2+C1r2dCPr212Ar(|J||J0|)C5(d,ε)r2+2εAr,\displaystyle-C_{5}(d,\varepsilon)r^{2}+C_{1}r^{2-d}\,C_{P}r^{-2}\frac{1}{2}A_{r}\big{(}|J|-|J_{0}|\big{)}\geq-C_{5}(d,\varepsilon)r^{2}+2\varepsilon A_{r},

provided that CPε/(C1λ).C_{P}\gtrsim\varepsilon/(C_{1}\lambda).

Finally, by the lower bounds for ar,Ara_{r},A_{r} in Lemma 4.2 and the condition (4.8) on uξu_{\xi}, we have

aR\displaystyle a_{R}\geq uξr2C5(d,ε)r2+2ε(uξCr2)(1+2ε)uξ(1+C5+C6)r2\displaystyle\,u_{\xi}-r^{2}-C_{5}(d,\varepsilon)r^{2}+2\varepsilon(u_{\xi}-Cr^{2})\geq(1+2\varepsilon)u_{\xi}-(1+C_{5}+C_{6})r^{2}
\displaystyle\geq (1+2ε)uξ(1+C5+C6)1Muξ(1+ε)uξ,\displaystyle\,(1+2\varepsilon)u_{\xi}-(1+C_{5}+C_{6})\,\frac{1}{M}\,u_{\xi}\geq(1+\varepsilon)u_{\xi},

provided that M(1+C5+C6)/ε:=M0(d,ε,λ).M\geq({1+C_{5}+C_{6}})/{\varepsilon}:=M_{0}(d,\varepsilon,\lambda). Recall that aR=Q(R)PR(ξ,n)una_{R}=\sum_{\partial Q(R)}P_{R}(\xi,n)u_{n} and Q(R)PR(ξ,n)=1\sum_{\partial Q(R)}P_{R}(\xi,n)=1, therefore, there is ξQ(R;ξ)\xi^{\prime}\in\partial Q(R;\xi) such that uξ(1+ε)uξ,u_{\xi^{\prime}}\geq(1+\varepsilon)u_{\xi}, which completes the proof of Lemma 4.1 in the second case. ∎

Notice that Lemma 4.1 is deterministic and requires no randomness of vnv_{n}. A direct consequence is the following estimate on the probability that uξu_{\xi} grows.

Lemma 4.4.

Let V(ω)={vn}nΛV(\omega)=\{v_{n}\}_{n\in\Lambda} be the Anderson-type potential as in Theorem 7. Fix 0<λ<10<\lambda<1, and retain ε<ε0,CP,M,r\varepsilon<\varepsilon_{0},C_{P},M,r_{\ast} from Lemma 4.1. For any cube QΛQ\subseteq\Lambda of side length (Q)\ell(Q), define the event

(4.18) Ω(Q):={ω:Card(jQ:vjCP(Q)2)λ|Q|}.\Omega(Q):=\Big{\{}\omega:\,{\rm Card}\left(j\in Q:v_{j}\geq C_{P}\,\ell(Q)^{-2}\right)\leq\lambda|Q|\Big{\}}.

Assume that r15/εr_{\ast}\geq 15/\varepsilon, otherwise, just reset rr_{\ast} to be max{15/ε,r}\max\{15/\varepsilon,r_{\ast}\}. For any r0rr_{0}\geq r_{\ast}, set

(4.19) rk+1=1+εrk,k=0,1,,kmax,r_{k+1}=\left\lfloor\sqrt{1+\varepsilon}\,r_{k}\right\rfloor,\ \ k=0,1,\cdots,k_{max},

where kmaxk_{\rm max} is the largest integer kk such that rk<Kr_{k}<K. Let Ωk=Ω(1,rkd),k=0,1,,kmax\Omega_{k}=\Omega\left(\llbracket 1,r_{k}\rrbracket^{d}\right),k=0,1,\cdots,k_{max}, and Ω=Ω(1,Kd)\Omega_{\infty}=\Omega\left(\llbracket 1,K\rrbracket^{d}\right).

Then for any cube BΛB\subset\Lambda of side length (B)=r0\ell(B)=r_{0}, and Bˇ=B/3\check{B}=B/3 as in (2.13),

(4.20) {ω:maxξBˇuξMr02}(Ω)+Cεdk=0kmax(Ωk),\mathbb{P}\Bigl{\{}\,\omega:\,\max_{\xi\in\check{B}}u_{\xi}\geq Mr_{0}^{2}\,\Bigr{\}}\,\leq{\mathbb{P}}\left(\,\Omega_{\infty}\,\right)+C\varepsilon^{-d}\sum_{k=0}^{k_{\rm max}}{\mathbb{P}}\left(\,\Omega_{k}\,\right),

for some dimensional constant CC.

Proof.

The idea is to repeatedly use Lemma 4.1 to construct a sequence of growing cubes and stop when the final cube exceeds the size of the entire domain.

We start with B0=BB_{0}=B of side length (B0)=r0\ell(B_{0})=r_{0}. Assume that maxBˇ0u(ξ)Mr02\max_{\check{B}_{0}}u({\xi})\geq Mr_{0}^{2}. We pick some ξ0Bˇ0\xi_{0}\in\check{B}_{0} such that u(ξ0)Mr02u(\xi_{0})\geq Mr_{0}^{2}. Then (4.8) is satisfied. Suppose furthermore that 0:=Ω0{\mathcal{E}}_{0}:=\Omega_{0} fails. Then (4.7) holds for B0B_{0}. All in all, Lemma 4.1 gives a point ξ1\xi_{1}, such that ξ1𝒞r1:={n:dist(n,Bˇ0)r1},\xi_{1}\in{{\mathcal{C}}_{r_{1}}}:=\{n:\,\mathrm{dist}\,(n,\check{B}_{0})\leq r_{1}\}, where dist\,\mathrm{dist}\, is measured in |||\cdot|_{\infty} for d{\mathbb{Z}}^{d} lattice points and

u(ξ1)(1+ε)u(ξ0)Mr12.u(\xi_{1})\geq(1+\varepsilon)u(\xi_{0})\geq Mr_{1}^{2}.

Additionally, we can require r015/εr_{0}\geq 15/\varepsilon which implies that r1>(1+ε/3)r0r_{1}>(1+\varepsilon/3)r_{0}. Clearly,

Card(𝒞r1)(2r0+2r1)d(12/ε)dr1d.{\rm Card}\left({{\mathcal{C}}_{r_{1}}}\right)\leq(2r_{0}+2r_{1})^{d}\leq({12}/{\varepsilon})^{d}r_{1}^{d}.

Therefore, 𝒞r1{{\mathcal{C}}_{r_{1}}} can be covered by at most n=36dεd+1n_{\ast}=\left\lfloor 36^{d}\varepsilon^{-d}\right\rfloor+1 disjoint cubes in d{\mathbb{Z}}^{d} of side length r1/3\left\lfloor r_{1}/3\right\rfloor, namely, Bˇ1(1),Bˇ1(2),,Bˇ1(n)\check{B}_{1}^{(1)},\check{B}_{1}^{(2)},\cdots,\check{B}_{1}^{(n_{\ast})}. Recall the definition of the middle third set in (2.13). Now extend each Bˇ1(j)\check{B}_{1}^{(j)} to a cube B1(j)B_{1}^{(j)} such that B1(j)B_{1}^{(j)} has side length r1r_{1} and contains each Bˇ1(j)\check{B}_{1}^{(j)} as a middle third part for j=1,2,,nj=1,2,\cdots,n_{\ast}. Note that these B1(j)B_{1}^{(j)} are not disjoint. But the overlap does not effect our estimate on the probability of the events from above.

In order for the induction driven by Lemma 4.1 to continue, we need to exclude the event that (4.7) fails for all B1(j)B_{1}^{(j)}. To this end, we define 1=j=1nΩ(B1(j)).{\mathcal{E}}_{1}=\bigcup_{j=1}^{n_{\ast}}\Omega(B_{1}^{(j)}). Assume that 1{\mathcal{E}}_{1} fails, which implies that (4.7) holds for all B1(j)B_{1}^{(j)}. Let B1=B1(j)B_{1}=B^{(j)}_{1} be the one that contains ξ1\xi_{1}. Now for ξ1Bˇ1(j)B1(j)\xi_{1}\in\check{B}_{1}^{(j)}\subsetneqq B_{1}^{(j)}, Lemma 4.1 gives ξ2\xi_{2} such that

(4.21) |ξ2ξ1|1+εr1=r2,u(ξ2)(1+ε)u(ξ1)Mr22.|\xi_{2}-\xi_{1}|\leq\left\lfloor\sqrt{1+\varepsilon}\,r_{1}\right\rfloor=r_{2},\qquad u(\xi_{2})\geq(1+\varepsilon)u(\xi_{1})\geq Mr_{2}^{2}.

Repeat the construction for r2>(1+ε/3)r1r_{2}>(1+\varepsilon/3)r_{1} and ξ2𝒞r2:={n:dist(n,Bˇ0)r1+r2}\xi_{2}\in{{\mathcal{C}}_{r_{2}}}:=\{n:\,\mathrm{dist}\,(n,\check{B}_{0})\leq r_{1}+r_{2}\},

Card(𝒞r2)(2r0+2r1+2r2)d(12/ε)dr2d.{\rm Card}\left({{\mathcal{C}}_{r_{2}}}\right)\leq(2r_{0}+2r_{1}+2r_{2})^{d}\leq({12}/{\varepsilon})^{d}r_{2}^{d}.

Therefore, 𝒞r2{{\mathcal{C}}_{r_{2}}} can be covered by at most n=36dεd+1n_{\ast}=\left\lfloor 36^{d}\varepsilon^{-d}\right\rfloor+1 disjoint cubes of side length r2/3\left\lfloor r_{2}/3\right\rfloor, Bˇ2(1),Bˇ2(2),,Bˇ2(n)\check{B}_{2}^{(1)},\check{B}_{2}^{(2)},\cdots,\check{B}_{2}^{(n_{\ast})}. Extend Bˇ2(j)\check{B}_{2}^{(j)} to B2(j)B_{2}^{(j)} in the same way, and define 2=j=1nΩ(B2(j)).{\mathcal{E}}_{2}=\bigcup_{j=1}^{n_{\ast}}\Omega(B_{2}^{(j)}). We assume that 2{\mathcal{E}}_{2} fails, and find ξ3\xi_{3} by Lemma 4.1. Inductively, at step kk, we assume that all the previous events, 1,2,,k1{\mathcal{E}}_{1},{\mathcal{E}}_{2},\cdots,{\mathcal{E}}_{k-1} fail, and obtain ξk,rk\xi_{k},r_{k} satisfying (4.9) and then define Brk,𝒞rkB_{r_{k}},{\mathcal{C}}_{r_{k}}. The same estimates as for r2,𝒞r2r_{2},{{\mathcal{C}}_{r_{2}}} hold for all kk, and hence,

(4.22) (1+ε/3)kr0(1+ε/3)rk1rk1+εrk1(1+ε)k/2r0(1+\varepsilon/3)^{k}r_{0}\leq\cdots\leq(1+\varepsilon/3)r_{k-1}\leq r_{k}\leq\sqrt{1+\varepsilon}\,r_{k-1}\leq\cdots\leq({1+\varepsilon})^{k/2}\,r_{0}

and

(4.23) Card(𝒞rk)(2r0+2r1+2rk)d2rk11(1+ε/3)1<(12/ε)drkd.{\rm Card}\left({{\mathcal{C}}_{r_{k}}}\right)\leq(2r_{0}+2r_{1}+\cdots 2r_{k})^{d}\leq 2r_{k}\ \frac{1}{1-(1+\varepsilon/3)^{-1}}<(12/\varepsilon)^{d}r_{k}^{d}.

Then Bˇk(j)\check{B}^{(j)}_{k}, Bk(j)B^{(j)}_{k} are defined in the same way as we did in the previous steps. Because of (4.23), for all kk, we need n=36dεd+1n_{\ast}=\left\lfloor 36^{d}\varepsilon^{-d}\right\rfloor+1 many Bˇk(j)\check{B}^{(j)}_{k} to cover 𝒞rk{{\mathcal{C}}_{r_{k}}}. Then we define the event k=j=1nΩ(Bk(j)).{\mathcal{E}}_{k}=\bigcup_{j=1}^{n_{\ast}}\Omega(B_{k}^{(j)}). Since vnv_{n} are i.i.d., the probability of each Ω(Bk(j))\Omega(B_{k}^{(j)}) is translation invariant, and only depends on the size of the cube Bk(j)B_{k}^{(j)}. In particular, {Ω(Bk(j))}=(Ωk).\mathbb{P}\bigl{\{}\Omega(B_{k}^{(j)})\bigr{\}}={\mathbb{P}}\left(\,\Omega_{k}\,\right). Therefore, for some dimensional constant CC,

(4.24) (k)n(Ωk)Cεd(Ωk),{\mathbb{P}}\left(\,{\mathcal{E}}_{k}\,\right)\leq\,n_{\ast}{\mathbb{P}}\left(\,\Omega_{k}\,\right)\leq\,C\varepsilon^{-d}\,{\mathbb{P}}\left(\,\Omega_{k}\,\right),

provided that ε<1\varepsilon<1.

We will continue the construction until we reach the kmaxk_{\rm max}-th step and obtain ξkmax,rkmax\xi_{k_{\rm max}},r_{k_{\rm max}}, BrkmaxB_{r_{k_{\max}}}, Bˇkmax(j)\check{B}^{(j)}_{k_{\max}}, Bkmax(j)B^{(j)}_{k_{\max}} and kmax{\mathcal{E}}_{k_{max}}.

We need to apply Lemma 4.1 two more times for the final step. However, according to our choice of kmaxk_{\max}, assuming that kmax{\mathcal{E}}_{k_{\max}} fails will already result a cube of side length 1+εrkmaxK\left\lfloor\sqrt{1+\varepsilon}\,r_{k_{\max}}\right\rfloor\geq K which may exceed the maximal size of the entire domain Λ\Lambda. To alleviate this issue, we need to enlarge the domain at this point for the last two steps, by making several copies of Λ\Lambda111We can do this from the very beginning of the construction, but it will not make any difference until we reach the size of KK.. We need pdp^{d} many copies where p:=31+ε+1p:=\left\lfloor 3\sqrt{1+\varepsilon}\right\rfloor+1.

Let K~=pK\widetilde{K}=pK so that K~>1+ε(3K)\widetilde{K}>\left\lfloor\sqrt{1+\varepsilon}\,(3K)\right\rfloor and denote

pΛ:=1,K~d=j(modp)d(Λ+jK).p\Lambda:=\llbracket 1,\widetilde{K}\rrbracket^{d}=\bigcup_{j\in({\mathbb{Z}}\,{\rm mod}\,p{\mathbb{Z}})^{d}}\left(\Lambda+j\,K\right).

We extend the potential V={vn}nΛV=\{v_{n}\}_{n\in\Lambda} periodically to V={vn}npΛV^{\prime}=\{v^{\prime}_{n}\}_{n\in p\Lambda}, where vm=vn,nΛ,v^{\prime}_{m}=v_{n},\ \ n\in\Lambda, and m=nmod(p)d.m=n\ {\rm mod}(p{\mathbb{Z}})^{d}. Now we consider the landscape equation on pΛp\Lambda for Δ+V-\Delta+V^{\prime} with periodic boundary conditions. The enlarged system has a unique solution uu^{\prime} by Theorem 5, which is a periodic extension of the original uu to pΛp\Lambda. Now we can return to the construction at the kmaxk_{\max}-th step.

Assume that the event kmax{\mathcal{E}}_{k_{\max}} fails. Then (4.7) holds for all possible Bkmax1,K~dB_{k_{\max}}\subset\llbracket 1,\widetilde{K}\rrbracket^{d} that may contain ξkmax\xi_{k_{\max}}. We have uξkmax=uξkmaxMrξkmax2.u^{\prime}_{\xi_{k_{\max}}}=u_{\xi_{k_{\max}}}\geq Mr^{2}_{\xi_{k_{\max}}}. Applying Lemma 4.1 to uΓu^{\Gamma} on Bkmax1,KdB_{k_{\max}}\subset\llbracket 1,K\rrbracket^{d}, we obtain ξ~1\widetilde{\xi}\in\llbracket 1, KdK\rrbracket^{d} such that

(4.25) uξ~(1+ε)uξkmaxM1+εrξkmax2MK2,u^{\prime}_{\widetilde{\xi}}\geq(1+\varepsilon)u_{\xi_{k_{\max}}}\geq M\left\lfloor\sqrt{1+\varepsilon}r_{\xi_{k_{\max}}}\right\rfloor^{2}\geq MK^{2}\,,

where the last inequality follows from the definition of kmaxk_{\max}. Now let ξ1,Kd\xi_{\infty}\in\llbracket 1,K\rrbracket^{d} be a point where uku_{k} attains its maximum. Clearly, u~k\widetilde{u}_{k} also attains its maximum at ξ\xi_{\infty},

(4.26) uξ=maxkΛuk=maxkpΛuk=uξ.u_{\xi_{\infty}}=\max_{k\in\Lambda}u_{k}=\max_{k\in p\Lambda}u^{\prime}_{k}=u^{\prime}_{\xi_{\infty}}\,.

Together with (4.25), we have uξ=uξuξ~MK2.u_{\xi_{\infty}}=u^{\prime}_{\xi_{\infty}}\geq u^{\prime}_{\widetilde{\xi}}\geq MK^{2}. Now consider Bˇ=Λ\check{B}_{\infty}=\Lambda, which is the middle third of B=3ΛpΛB_{\infty}=3\Lambda\subset p\Lambda. Let

:=\displaystyle{\mathcal{E}}_{\infty}:= {Card(j3Λ:vjCP(3K)2)λ|3Λ|}\displaystyle\left\{\,{\rm Card}\left(j\in 3\Lambda:v^{\prime}_{j}\geq C_{P}(3K)^{-2}\right)\leq\lambda|3\Lambda|\,\right\}
=\displaystyle= {Card(jΛ:vjCP9K2)λKd}.\displaystyle\left\{\,{\rm Card}\left(j\in\Lambda:v_{j}\geq\frac{C_{P}}{9}\,K^{-2}\right)\leq\lambda K^{d}\,\right\}.

Since Card(jΛ:vjCP9K2)Card(jΛ:vjCPK2){\rm Card}\left(j\in\Lambda:v_{j}\geq\frac{C_{P}}{9}\,K^{-2}\right)\geq{\rm Card}\left(j\in\Lambda:v_{j}\geq C_{P}\,K^{-2}\right), one has

(4.27) {Card(jΛ:vjCPK2)λKd}=Ω.\displaystyle{\mathcal{E}}_{\infty}\subseteq\left\{\,{\rm Card}\left(j\in\Lambda:v_{j}\geq C_{P}\,K^{-2}\right)\leq\lambda K^{d}\,\right\}=\Omega_{\infty}\,.

Now if {\mathcal{E}}_{\infty} fails, apply Lemma 4.1 one last time to uu^{\prime} on BˇB\check{B}_{\infty}\subsetneq B_{\infty}222One also needs to take MM nine times larger, so that M/9M0M/9\geq M_{0} in (4.26) to meet the requirement in Lemma 4.1.. Then (4.9) implies that there is a ξ1,K~d\xi^{\prime}\in\llbracket 1,\widetilde{K}\rrbracket^{d} such that uξ(1+ε)uξ>uξu^{\prime}_{\xi^{\prime}}\geq(1+\varepsilon)u^{\prime}_{\xi_{\infty}}>u^{\prime}_{\xi_{\infty}}. This is a contradiction. Recall that it happens when we start with uξ0Mr02u_{\xi_{0}}\geq Mr_{0}^{2} and assume that all j{\mathcal{E}}_{j} fail. Therefore, at least one j{\mathcal{E}}_{j} must be true. In other words,

{maxBˇ0uξMr02}(j=1kmaxj){maxBˇ0uξMr02}()+j=1kmax(j).\Bigl{\{}\,\max_{\check{B}_{0}}u_{\xi}\geq Mr_{0}^{2}\,\Bigr{\}}\subset{\mathcal{E}}_{\infty}\bigcup\Bigl{(}\bigcup_{j=1}^{k_{\max}}{\mathcal{E}}_{j}\Bigr{)}\Longrightarrow\mathbb{P}\Bigl{\{}\max_{\check{B}_{0}}u_{\xi}\geq Mr_{0}^{2}\Bigr{\}}\leq{\mathbb{P}}\left(\,{\mathcal{E}}_{\infty}\,\right)+\sum_{j=1}^{k_{\max}}{\mathbb{P}}\left(\,{\mathcal{E}}_{j}\,\right).

Together with (4.24) and (4.27), this completes the proof. ∎

The next lemma allows us to estimate the probability of each term on the right hand side of (4.20).

Lemma 4.5.

Let V={vj}jΛV=\{v_{j}\}_{j\in\Lambda} be the Anderson potential as in Theorem 7. Let FF and δ\delta_{\ast} be as in (4.1). For any BΛB\subset\Lambda and 0<λ<10<\lambda<1, if δ>0\delta>0 is such that 1λF(δ)>01-\lambda-F(\delta)>0, then

(4.28) {Card(jB:vjδ)λ|B|}eD(1λF)|B|,{\mathbb{P}}\left\{\,{\rm Card}\left(j\in B:v_{j}\geq\delta\right)\leq\lambda|B|\,\right\}\leq e^{-D(1-\lambda\|F)\,\,|B|},

where

(4.29) D(xy)=xlogxy+(1x)log1x1yD(x\|y)=x\log\frac{x}{y}+(1-x)\log\frac{1-x}{1-y}

is the Kullback–Leibler divergence between Bernoulli distributed random variables with parameters xx and yy respectively.

As a consequence, for any rr\in\mathbb{N},

(4.30) {Card(j1,rd:vjδ)λrd}(C(λ)F(δ)1λ)rd,{\mathbb{P}}\left\{\,{\rm Card}\left(j\in\llbracket 1,r\rrbracket^{d}:v_{j}\geq\delta\right)\leq\lambda r^{d}\,\right\}\,\leq\,\left(C(\lambda)F(\delta)^{1-\lambda}\right)^{r^{d}},

where C(λ)=(1λ)1λλC(\lambda)=(1-\lambda)^{-1}\lambda^{-\lambda}.

Furthermore, there is a λ>0\lambda_{\ast}>0, which only depends on F(δ),F(\delta_{\ast}), such that for all 0<δδ,0<λλ0<\delta\leq\delta_{\ast},0<\lambda\leq\lambda_{\ast} and any rr\in\mathbb{N}, one has

(4.31) {Card(j1,rd:vjδ)λrd}(F(δ))rd/2.{\mathbb{P}}\left\{\,{\rm Card}\left(j\in\llbracket 1,r\rrbracket^{d}:v_{j}\geq\delta\right)\leq\lambda r^{d}\,\right\}\,\leq\,\left(F(\delta)\right)^{{r^{d}}/{2}}.
Remark 4.5.

The λ\lambda_{\ast} can be taken as of order (1F(δ))2(1-F(\delta_{\ast}))^{2}, see (4.37).

Proof.

Let ζj\zeta_{j} be the characteristic function for the event vjδv_{j}\leq\delta, i.e., ζj=1\zeta_{j}=1 for vjδv_{j}\leq\delta and ζj=0\zeta_{j}=0 otherwise. Since {vj}\{v_{j}\} are i.i.d. random variables, all ζj\zeta_{j} are i.i.d. Bernoulli random variables, taking values in {0,1}\{0,1\}, with common expectation 𝔼(ζj)=(vjδ)=F(δ)\mathbb{E}\left(\zeta_{j}\right)={\mathbb{P}}\left(\,v_{j}\leq\delta\,\right)=F(\delta). Let SB:=jBζj.S_{B}:=\sum_{j\in B}\zeta_{j}. By the Chernoff–Hoeffding Theorem, [Ho] (see Lemma B.1),

(4.32) {SB(1λ)|B|}eD(1λF)|B|,{\mathbb{P}}\left\{\,S_{B}\geq(1-\lambda)|B|\,\right\}\leq e^{-D(1-\lambda\|F)\,|B|},

where F=F(δ)F=F(\delta) and D(xy)D(x\|y) is as in (4.29). Then (4.28) follows directly from (4.32) since

|B|Card(jB:vj>δ)=Card(jB:vjδ)=jBζj=SB.|B|-{\rm Card}\left(j\in B:v_{j}>\delta\right)={\rm Card}\left(j\in B:v_{j}\leq\delta\right)=\sum_{j\in B}\zeta_{j}=S_{B}.

Examining the the Kullback–Leibler divergence with parameter 1λ1-\lambda and FF, one has

D(1λF)=(1λ)log1λF+λlogλ1F\displaystyle D(1-\lambda\|F)=(1-\lambda)\log\frac{1-\lambda}{F}+\lambda\log\frac{\lambda}{1-F}\geq log((1λ)1λλλ)logF1λ\displaystyle\log\left((1-\lambda)^{1-\lambda}\lambda^{\lambda}\right)\,-\,\log F^{1-\lambda}
\displaystyle\geq log((1λ)λλ)logF1λ,\displaystyle\log\left((1-\lambda)\lambda^{\lambda}\right)\,-\,\log F^{1-\lambda},

where we used 1F<11-F<1 and (1λ)1λ1λ(1-\lambda)^{1-\lambda}\geq 1-\lambda. Therefore,

(4.33) {Card(j[1,r]dd:vjδ)λrd}eD(1λF)rd((1λ)1λλF1λ)rd,{\mathbb{P}}\left\{\,{\rm Card}\left(j\in[1,r]^{d}\cap{\mathbb{Z}}^{d}:v_{j}\geq\delta\right)\leq\lambda r^{d}\,\right\}\leq e^{-D(1-\lambda\|F)\,r^{d}}\leq\left((1-\lambda)^{-1}\lambda^{-\lambda}F^{1-\lambda}\right)^{r^{d}},

which yields (4.30).

Let q=1F(δ)(0,1)q=1-F(\delta_{\ast})\in(0,1). For 0<δδ0<\delta\leq\delta_{\ast},

(4.34) F(δ)F(δ)=1q<1logF(δ)log(1q)<0.F(\delta)\leq F(\delta_{\ast})=1-q<1\ \Longrightarrow\ \log F(\delta)\leq\log(1-q)<0.

On the other hand, it is easy to check that

(4.35) limλ0+log((1λ)λλ)1/2λ=0.\lim_{\lambda\to 0^{+}}\frac{\log\left((1-\lambda)\lambda^{\lambda}\right)}{1/2-\lambda}=0.

Then there is a λ1/2\lambda_{\ast}\leq 1/2 such that for all 0<λλ0<\lambda\leq\lambda_{\ast},

(4.36) log((1λ)λλ)1/2λ>log(1q)logF(δ)(1λ)1λλ<(F(δ))λ1/2.\frac{\log\left((1-\lambda)\lambda^{\lambda}\right)}{1/2-\lambda}>\log(1-q)\geq\log F(\delta)\Longrightarrow(1-\lambda)^{-1}\lambda^{-\lambda}<\left(F(\delta)\right)^{\lambda-1/2}.

Combined with (4.33), this gives

{Card(j1,rd:vjδ)λrd}(F(δ))rd/2,{\mathbb{P}}\left\{\,{\rm Card}\left(j\in\llbracket 1,r\rrbracket^{d}:v_{j}\geq\delta\right)\leq\lambda r^{d}\,\right\}\,\leq\,\left(F(\delta)\right)^{r^{d}/2},

which completes the proof of Lemma 4.5.

One can be more specific regarding the exact value of λ\lambda_{\ast}. For 0<δδ0<\delta\leq\delta_{\ast}, logF(δ)log(1q)<min(q,12)\log F(\delta)\leq\log(1-q)<-\min(q,\frac{1}{2}). If λ<1/4\lambda<1/4, then log(1λ)>2λ>λ\log(1-\lambda)>-2\lambda>-\sqrt{\lambda}, and λlogλ>λ.\lambda\,\log\lambda>-\sqrt{\lambda}. Let

(4.37) λ=(18min(q,1/2))2.\lambda_{\ast}=\left(\frac{1}{8}\min(q,1/2)\right)^{2}.

Then for λ<λ1/4\lambda<\lambda_{\ast}\leq 1/4,

0>log((1λ)λλ)1/2λ4(log((1λ)+λlogλ)8λmin(q,12)logF(δ),0>\frac{\log\left((1-\lambda)\lambda^{\lambda}\right)}{1/2-\lambda}\geq 4\left(\log((1-\lambda)+\lambda\log\lambda\right)\geq-8\sqrt{\lambda}\geq-\min(q,\frac{1}{2})\geq\log F(\delta),

which gives (4.36) similarly to the argument above. ∎

Combining Lemma 4.4 and Lemma 4.5 leads to

Lemma 4.6.

Let δ\delta_{\ast} and λ\lambda_{\ast} be as in Lemma 4.5. Fix λλ\lambda\leq\lambda_{\ast}, and take ε<ε0(λ,d)\varepsilon<\varepsilon_{0}(\lambda,d) and CP,M,rC_{P},M,r_{\ast} as in Lemma 4.4. Then for any cube BΛB\subsetneq\Lambda of side length (B)=rr\ell(B)=r\geq r_{\ast}, and its middle third part Bˇ\check{B}, one has

(4.38) {maxξBˇuξMr2}Cεd1F(δ)(F(CPr2))rd/2\mathbb{P}\Bigl{\{}\,\max_{\xi\in\check{B}}u_{\xi}\geq Mr^{2}\,\Bigr{\}}\,\leq\frac{C\varepsilon^{-d}}{1-F(\delta_{\ast})}\left(F(C_{P}r^{-2})\right)^{r^{d}/2}

for some dimensional constant C>0C>0.

Remark 4.6.

The exponent rd/2r^{d}/2 can be made arbitrarily close to rdr^{d}, by taking λ\lambda smaller, however, it will also result a large factor 1/λ1/\lambda in front of FF.

Proof.

Let r0=rr_{0}=r and define the sequence rkr_{k} as in Lemma 4.4 and r:=Kr_{\infty}:=K. Let δk=CPrk2,k=0,,kmax\delta_{k}=C_{P}r_{k}^{-2},k=0,\cdots,k_{\max}, and δ=CPK2\delta_{\infty}=C_{P}K^{-2}. By the construction of rkr_{k} and (4.22), one has

rkd(1+ε/3)dkr0d(1+kdε/3)r0dr0d+2k,k=0,,kmax.r^{d}_{k}\geq(1+\varepsilon/3)^{dk}r^{d}_{0}\geq(1+kd\varepsilon/3)r^{d}_{0}\geq r^{d}_{0}+2k,\ \ k=0,\cdots,k_{\max}.

For k=0,,kmaxk=0,\cdots,k_{\max} and k=k=\infty, one has δkδ0=CPr02δ\delta_{k}\leq\delta_{0}=C_{P}r_{0}^{-2}\leq\delta_{\ast} provided that r0CP/δr_{0}\geq\sqrt{C_{P}/\delta_{\ast}}. Notice that in the proof of Lemma 4.5, by the choice of δ\delta_{\ast} in (4.1), one has F(δ0)<F(δ):=1qF(\delta_{0})<F(\delta_{\ast}):=1-q. Therefore,

F(δk)1q<1,k=0,,kmax,andk=,F(\delta_{k})\leq 1-q<1,\ \ k=0,\cdots,k_{\max},\ {\rm and}\ k=\infty,

since the distribution FF is non-decreasing. Now apply Lemma 4.5 to all rkr_{k}. Combining (4.31) with (4.20), one has

{maxξBˇuξMr02}F(δ)Kd/2+Cεdk=0kmaxF(δk)rkd/2F(δ0)r0d/2+Cεdk=0kmaxF(δ0)(r0d+2k)/2F(δ0)r0d/2+CεdF(δ0)r0d/211F(δ0)F(δ0)r0d/2(1+Cεdq):=C(d,ε,δ)F(δ0)r0d/2,\mathbb{P}\left\{\,\max_{\xi\in\check{B}}u_{\xi}\geq Mr_{0}^{2}\,\right\}\,\leq F(\delta_{\infty})^{K^{d}/2}+C\varepsilon^{-d}\sum_{k=0}^{k_{\max}}F(\delta_{k})^{r_{k}^{d}/2}\\ \leq F(\delta_{0})^{r^{d}_{0}/2}+C\varepsilon^{-d}\sum_{k=0}^{k_{\max}}F(\delta_{0})^{(r_{0}^{d}+2k)/2}\leq F(\delta_{0})^{r_{0}^{d}/2}+C\varepsilon^{-d}F(\delta_{0})^{r_{0}^{d}/2}\frac{1}{1-F(\delta_{0})}\\ \leq F(\delta_{0})^{r_{0}^{d}/2}\left(1+\frac{C\varepsilon^{-d}}{q}\right):=C(d,\varepsilon,\delta_{\ast})F(\delta_{0})^{r_{0}^{d}/2},

which is the desired bound. ∎

Now we are ready to complete:

Proof of (4.3) in Theorem 7.

Let δ\delta_{\ast} and λ\lambda_{\ast} be given by Lemma 4.6. Fix λλ\lambda\leq\lambda_{\ast}, take ε<ε0(λ,d)\varepsilon<\varepsilon_{0}(\lambda,d) and CP,M,rC_{P},M,r_{\ast} as in Lemma 4.5.

For any μ1/(4M)\mu\leq 1/(4M), let r=(4Mμ)1/2r=\left\lceil(4M\mu)^{-1/2}\right\rceil so that μ1/4<Mr2μ1\mu^{-1}/4<Mr^{2}\leq\mu^{-1}. To apply Lemma 4.6, one also needs to ensure that rrr\geq r_{\ast}, which requires μ\mu to be taken in the range μμ=1/(Mr2)\mu\leq\mu_{\ast}=1/(Mr_{\ast}^{2}).

Now for any cube BB of side length r=(4Mμ)1/2r=\left\lceil(4M\mu)^{-1/2}\right\rceil and its middle third part Bˇ\check{B}, Lemma 4.6 implies that

{maxξBˇuξμ1}{maxξBˇuξMr2}C(F(CPr2))rd/2,\mathbb{P}\Bigl{\{}\,\max_{\xi\in\check{B}}u_{\xi}\geq\mu^{-1}\,\Bigr{\}}\leq\mathbb{P}\Bigl{\{}\,\max_{\xi\in\check{B}}u_{\xi}\geq Mr^{2}\,\Bigr{\}}\,\leq\,C\cdot\left(F(C_{P}r^{-2})\right)^{r^{d}/2},

where C>0C>0 in given by (4.38) depending on d,εd,\varepsilon and F(δ)F(\delta_{\ast}). Then

(4.39) {minnBˇ1unμ}C(F(CPr2))rd/2C(F(4CPMμ))(Mμ)d/2/2.\mathbb{P}\Bigl{\{}\,\min_{n\in\check{B}}\frac{1}{u_{n}}\leq\mu\,\Bigr{\}}\leq\,C\left(F(C_{P}r^{-2})\right)^{r^{d}/2}\leq\,C\big{(}F(4C_{P}M\,\mu)\big{)}^{(M\mu)^{-d/2}/2}.

Notice that (Bˇ)r/6(4Mμ)1/2/6\ell(\check{B})\geq r/6\geq(4M\mu)^{-1/2}/6. Recall that the cubes used in the definition of NuN_{u} have side length μ1/2\left\lceil\mu^{-1/2}\right\rceil. Any Q𝒫(μ1/2;Λ)Q\in{\mathcal{P}}\bigl{(}\left\lceil\mu^{-1/2}\right\rceil;\Lambda\bigr{)} can be covered by at most (μ1/2(4Mμ)1/2/6)d+1CMd/2\left(\frac{\left\lceil\mu^{-1/2}\right\rceil}{(4M\mu)^{-1/2}/6}\right)^{d}+1\leq\,C^{\prime}M^{d/2} disjoint cubes of side length (Bˇ)=(4Mμ)1/2/3\ell(\check{B})=\left\lceil\left\lceil(4M\mu)^{-1/2}\right\rceil/3\right\rceil for some CC^{\prime} which only depends on M,dM,d. Notice also that the estimate (4.39) is independent of the position of Bˇ\check{B}, and can be applied to all cubes Bˇ\check{B} of the same size. Therefore,

{minnQ1unμ}CMd/2{minnBˇ1unμ}C′′(F(C3μ))γ2μd/2\mathbb{P}\left\{\min_{n\in Q}\frac{1}{u_{n}}\,\leq\mu\right\}\leq\,C^{\prime}M^{d/2}\,\mathbb{P}\left\{\,\min_{n\in\check{B}}\frac{1}{u_{n}}\leq\mu\,\right\}\leq\,C^{\prime\prime}\big{(}F(C_{3}\,\mu)\big{)}^{\gamma_{2}\mu^{-d/2}}

for any QQ. Together with (4.5), we obtain the desired upper bound

𝔼(Nu(μ))1μ1/2dmaxQ𝒫(μ1/2){minnQ1unμ}C4μd/2(F(C3μ))γ2μd/2\mathbb{E}\left(N_{u}(\mu)\right)\leq\frac{1}{\left\lceil\mu^{-1/2}\right\rceil^{d}}\max_{Q\in{\mathcal{P}}(\left\lceil\mu^{-1/2}\right\rceil)}\mathbb{P}\left\{\min_{n\in Q}\frac{1}{u_{n}}\,\leq\mu\right\}\leq\,C_{4}\mu^{d/2}\big{(}F(C_{3}\,\mu)\big{)}^{\gamma_{2}\mu^{-d/2}}

for all μμ\mu\leq\mu_{\ast}. The constants C3,C4,γ2>0C_{3},C_{4},\gamma_{2}>0 only depend on d,Md,M and CPC_{P}, which eventually only depend on dd and F=F(δ)F_{\ast}=F(\delta_{\ast}). ∎

4.2. Lifschitz tails for the integrated density of states

Putting together the general upper/lower bounds in Theorem 1,2 for the deterministic case, and the Lifshitz tails in Theorem 7 for the Anderson model, we have

Theorem 8.

Let C1C_{1} be as in Theorem 1 and δ\delta_{\ast} be as in Theorem 7. Then there are constants c5,c6>0c_{5},c_{6}>0 depending on dd, δ\delta_{\ast} and VmaxV_{\max} such that

(4.40) c5𝔼Nu(c6μ)𝔼N(μ)𝔼Nu(C1μ),forallμ>0.c_{5}\,\mathbb{E}N_{u}(\,c_{6}\,\mu)\,\leq\,\mathbb{E}N(\mu)\,\leq\,\mathbb{E}N_{u}(C_{1}\,\mu),\ {\rm for\ all}\ \mu>0.

If furthermore μ,K\mu_{\ast},K_{\ast} are as in Theorem 7, depending only on dd and δ\delta_{\ast}, and μ<μ\mu<\mu_{\ast}, then the estimate (4.40) holds with constants c5,c6c_{5},c_{6} which are independent of VmaxV_{\max}.

If, in addition, K/K2<μ<μ/C1K_{\ast}/K^{2}<\mu<\mu_{\ast}/C_{1}, then there are constants C¯1,c¯1,C¯2,c¯1,γ¯1,γ¯2\bar{C}_{1},\bar{c}_{1},\bar{C}_{2},\bar{c}_{1},\bar{\gamma}_{1},\bar{\gamma}_{2} depending only on δ\delta_{\ast} such that

(4.41) c¯2μd/2F(c¯1μ)γ¯1μ2d/2𝔼N(μ)C¯2μd/2F(C¯1μ)γ¯2μd/2.\bar{c}_{2}\mu^{d/2}F(\bar{c}_{1}\mu)^{\bar{\gamma}_{1}\mu_{2}^{-d/2}}\,\leq\,\mathbb{E}N(\mu)\,\leq\,\bar{C}_{2}\,\mu^{d/2}F(\bar{C}_{1}\,\mu)^{\bar{\gamma}_{2}\mu^{-d/2}}.
Proof.

The upper bound in (4.40) is the average of the upper bound in Theorem 1. We only need to study the lower bound with the help of Theorem 2 and Theorem 7. Let c,c0,c1,C1c_{\ast},c_{0},c_{1},C_{1} and α<α0<1\alpha<\alpha_{0}<1, be as in Theorem 2. If μ4d+Vmax\mu\geq 4d+V_{\max} then N(μ)=1N(\mu)=1 and the left-hand side of (4.40) holds trivially. Fix μ<4d+Vmax\mu<4d+V_{\max}, let us denote μ2=c1αd+2μ\mu_{2}=c_{1}\alpha^{d+2}\mu and μ4=α2μ2=c1αd+4μ\mu_{4}=\alpha^{2}\mu_{2}=c_{1}\alpha^{d+4}\mu. Assume further that αα00:=(c/(4d+Vmax))1/4\alpha\leq\alpha_{00}:=\big{(}c_{\ast}/(4d+V_{\max})\big{)}^{-1/4}. Then (1.5) in Theorem 2 implies that

(4.42) 𝔼N(μ)c0αd𝔼Nu(μ2)C0𝔼Nu(μ4).\mathbb{E}N(\mu)\geq c_{0}\alpha^{d}\mathbb{E}N_{u}(\mu_{2})-C_{0}\mathbb{E}N_{u}(\mu_{4}).

Next, let c3,c4,C3,C4,γ1,γ2>0c_{3},c_{4},C_{3},C_{4},\gamma_{1},\gamma_{2}>0 and μ\mu_{\ast} be given by Theorem 7. Then by (4.2) and (4.3), if μ21\mu_{2}\leq 1 and μ4μ\mu_{4}\leq\mu_{\ast}, one has

(4.43) 𝔼Nu(μ2)c4μ2d/2F(c3μ2)γ1μ2d/2,and𝔼(Nu(μ4))C4μ4d/2F(C3μ4)γ2μ4d/2.\mathbb{E}{N_{u}(\mu_{2})}\geq c_{4}\mu_{2}^{d/2}F(c_{3}\mu_{2})^{\gamma_{1}\mu_{2}^{-d/2}},\ \ {\rm and}\ \ \mathbb{E}\left(N_{u}(\mu_{4})\right)\leq\,C_{4}\mu_{4}^{d/2}F(C_{3}\mu_{4})^{\gamma_{2}\mu_{4}^{-d/2}}.

Therefore,

𝔼N(μ)\displaystyle\mathbb{E}N(\mu)\geq c4αdμ2d/2F(c3μ2)γ1μ2d/2C4μ4d/2F(C3μ4)γ2μ4d/2\displaystyle\,c_{4}\alpha^{d}\mu_{2}^{d/2}F(c_{3}\mu_{2})^{\gamma_{1}\mu_{2}^{-d/2}}-C_{4}\mu_{4}^{d/2}F(C_{3}\mu_{4})^{\gamma_{2}\mu_{4}^{-d/2}}
(4.44) =\displaystyle= c4μ4d/2(F(c3μ2)γ1μ2d/2C4F(C3α2μ2)γ2μ4d/2)\displaystyle\,c_{4}{\mu_{4}^{d/2}}\left(\,F(c_{3}\mu_{2})^{\gamma_{1}\mu_{2}^{-d/2}}-C_{4}F(C_{3}\alpha^{2}\mu_{2})^{\gamma_{2}\mu_{4}^{-d/2}}\,\right)

for μ4μ,μ21\mu_{4}\leq\mu_{\ast},\mu_{2}\leq 1 and μ<4d+Vmax\mu<4d+V_{\max}. This requires αmin{α1,α2}\alpha\leq\min\{\alpha_{1},\alpha_{2}\}, where

α1:=(c1(4d+Vmax))1/(d+2),andα2:=μ1/(d+4)(c1(4d+Vmax))1/(d+4).\alpha_{1}:=\big{(}c_{1}(4d+V_{\max})\big{)}^{-1/(d+2)},\ \ {\rm and}\ \alpha_{2}:=\mu_{\ast}^{1/(d+4)}\big{(}c_{1}(4d+V_{\max})\big{)}^{-1/(d+4)}.

Let δ\delta_{\ast} be as in (4.1). If we assume, in addition, that α\alpha is smaller than both α3\alpha_{3} and α4\alpha_{4},

α3:=c3/C3,andα4:=δ1/(d+2)(c3c1(4d+Vmax))1/(d+2),\alpha_{3}:=\sqrt{c_{3}/C_{3}},\ \ {\rm and}\ \alpha_{4}:=\delta_{\ast}^{1/(d+2)}\big{(}c_{3}c_{1}(4d+V_{\max})\big{)}^{-1/(d+2)},

then for all μ<4d+Vmax\mu<4d+V_{\max}, one has C3μ4<c2μ2δC_{3}\mu_{4}<c_{2}\mu_{2}\leq\delta_{\ast}. Therefore, 0<F4F2F<10<F_{4}\leq F_{2}\leq F_{\ast}<1, where F4=F(C3μ4)=F(C3α2μ2)F_{4}=F(C_{3}\mu_{4})=F(C_{3}\alpha^{2}\mu_{2}), F2=F(c3μ2)F_{2}=F(c_{3}\mu_{2}), and F=F(δ).F_{\ast}=F(\delta_{\ast}). The difference term in (4.44) is then bounded from below by

F2γ1μ2d/2C4F4γ2μ4d/2F4γ1μ2d/2C4F4γ2μ4d/2.F_{2}^{\gamma_{1}\mu_{2}^{-d/2}}-C_{4}F_{4}^{\gamma_{2}\,\mu_{4}^{-d/2}}\geq F_{4}^{\gamma_{1}\mu_{2}^{-d/2}}-C_{4}F_{4}^{\gamma_{2}\,\mu_{4}^{-d/2}}.

We want to pick α\alpha small enough (independent of μ\mu) so that,

(4.45) F4γ1μ2d/2C4F4γ2μ4d/212F4γ1μ2d/2,F_{4}^{\gamma_{1}\mu_{2}^{-d/2}}-C_{4}F_{4}^{\gamma_{2}\,\mu_{4}^{-d/2}}\ \geq\ \frac{1}{2}F_{4}^{\gamma_{1}\mu_{2}^{-d/2}},

that is,

(4.46) (2C4)1F4γ2μ4d/2γ1μ2d/2=F4μ2d/2(γ2αdγ1).(2C_{4})^{-1}\geq\ F_{4}^{\gamma_{2}\,\mu_{4}^{-d/2}-\gamma_{1}\mu_{2}^{-d/2}}=F_{4}^{\mu_{2}^{-d/2}(\gamma_{2}\,\alpha^{-d}-\gamma_{1}\,)}.

Notice that μ2<1\mu_{2}<1, hence, μ2d/2>1\mu_{2}^{-d/2}>1 and μ2d/2(γ2αdγ1)>γ2αdγ1>0\mu_{2}^{-d/2}(\gamma_{2}\alpha^{-d}-\gamma_{1})>\gamma_{2}\alpha^{-d}-\gamma_{1}>0 provided that α<(γ2/γ1)1/d:=α5\alpha<(\gamma_{2}/\gamma_{1})^{1/d}:=\alpha_{5}. Then the fact that F4F<1F_{4}\leq F_{\ast}<1 implies that

F4μ2d/2(γ2αdγ1)Fμ2d/2(γ2αdγ1)Fγ2αdγ1.F_{4}^{\mu_{2}^{-d/2}(\gamma_{2}\alpha^{-d}-\gamma_{1})}\leq F_{\ast}^{\mu_{2}^{-d/2}(\gamma_{2}\alpha^{-d}-\gamma_{1})}\leq F_{\ast}^{\gamma_{2}\alpha^{-d}-\gamma_{1}}.

Solving 1(2C4)1Fγ2αdγ11\geq(2C_{4})^{-1}\geq F_{\ast}^{\gamma_{2}\alpha^{-d}-\gamma_{1}} for α\alpha, we observe that

γ2αdγ1log(2C4)logFα(γ21log(2C4)logF+γ21γ1)1/d:=α6\gamma_{2}\alpha^{-d}-\gamma_{1}\geq\frac{\log(2C_{4})}{\log F_{\ast}}\Longleftrightarrow\alpha\leq\left(\gamma_{2}^{-1}\frac{\log(2C_{4})}{\log F_{\ast}}+\gamma_{2}^{-1}\gamma_{1}\right)^{-1/d}:=\alpha_{6}

would yield (4.46).

Putting everything together, set

α=α:=min{α0,α00,α1,α2,,α6}.\alpha=\alpha_{\ast}:=\min\{\alpha_{0},\alpha_{00},\alpha_{1},\alpha_{2},\cdots,\alpha_{6}\}.

Then, for all μ<4d+Vmax\mu<4d+V_{\max}, (4.44), (4.45) and (4.43) imply that

𝔼N(μ)12c4μ4d/2F4γ1μ2d/212c4C41𝔼Nu(μ4)=12c4C41𝔼Nu(c1αd+4μ)=:c5𝔼Nu(c6μ),\mathbb{E}N(\mu)\geq\frac{1}{2}\,c_{4}{\mu_{4}^{d/2}}F_{4}^{\gamma_{1}\mu_{2}^{-d/2}}\geq\frac{1}{2}\,c_{4}C_{4}^{-1}\mathbb{E}N_{u}(\mu_{4})=\frac{1}{2}c_{4}C_{4}^{-1}\,\mathbb{E}N_{u}(c_{1}\alpha_{\ast}^{d+4}\mu)=:c_{5}\,\mathbb{E}N_{u}(c_{6}\ \mu)\,,

which completes the proof for the first inequality in (4.40).

It is also easy to verify that if we are only interested in small μ\mu, then all the αi\alpha_{i} can be picked independently of VmaxV_{\max}. Therefore, the final constants c5,c6c_{5},c_{6} are also independent of VmaxV_{\max}. In particular, let μ\mu_{\ast} be as in Theorem 7. Then for all cμ<C1μ<μc_{\ast}^{\prime}\mu<C_{1}\mu<\mu_{\ast}

𝔼N(μ)𝔼Nu(C1μ)C4(C1μ)d/2F(C3C1μ)γ2(C1μ)d/2=:C¯2μd/2F(C¯1μ)γ¯2μd/2\mathbb{E}N(\mu)\leq\mathbb{E}N_{u}(C_{1}\mu)\leq\,C_{4}(C_{1}\mu)^{d/2}F(C_{3}C_{1}\mu)^{\gamma_{2}(C_{1}\mu)^{-d/2}}=:\bar{C}_{2}\mu^{d/2}F(\bar{C}_{1}\mu)^{\bar{\gamma}_{2}\mu^{-d/2}}

and

𝔼N(μ)c5𝔼Nu(C6μ)c5C4(c6μ)d/2F(c3c6μ)γ1(C6μ)d/2=:c¯2μd/2F(c¯1μ)γ¯2μd/2\mathbb{E}N(\mu)\geq c_{5}\mathbb{E}N_{u}(C_{6}\mu)\geq c_{5}C_{4}(c_{6}\mu)^{d/2}F(c_{3}c_{6}\mu)^{\gamma_{1}(C_{6}\mu)^{-d/2}}=:\bar{c}_{2}\mu^{d/2}F(\bar{c}_{1}\mu)^{\bar{\gamma}_{2}\mu^{-d/2}}

where the constants c¯1,c¯2,C¯1,C¯2,γ¯1,γ¯2\bar{c}_{1},\bar{c}_{2},\bar{C}_{1},\bar{C}_{2},\bar{\gamma}_{1},\bar{\gamma}_{2} only depend on dd and μ\mu_{\ast}, and are independent of VmaxV_{\max}. ∎

4.3. Dual landscape and the top edge of the spectrum.

Let H=Δ+VH=-\Delta+V be as in (1.1) acting on =2(Λ),Λ=(/K)d\mathcal{H}=\ell^{2}(\Lambda),\Lambda=({\mathbb{Z}}/K{\mathbb{Z}})^{d}. In this part, we will briefly discuss the so-called dual landscape and see how it is applied to the eigenvalue-counting for high energy modes. We refer readers to Section 2.4 in [WZ] for more details. For φ=2(Λ)\varphi\in\mathcal{H}=\ell^{2}({\Lambda}), we define a dual vector φ~\widetilde{\varphi}

(4.47) φ~n=(1)s(n)φn,nΛ,\widetilde{\varphi}_{n}=(-1)^{s(n)}\varphi_{n},\ n\in{\Lambda},

where s(n)=j=1dnjs(n)=\sum_{j=1}^{d}n_{j} for n=(n1,n2,,nd)dn=(n_{1},n_{2},\cdots,n_{d})\in{\mathbb{Z}}^{d}. We assume, in addition, that KK is an even number so that φ~n=φ~n+Kei,nΛ,i=1,d\widetilde{\varphi}_{n}=\widetilde{\varphi}_{n+Ke_{i}},n\in\Lambda,i=1\cdots,d. Now suppose (μ,φ)(\mu,\varphi) is an eigenpair of H=Δ+VH=-\Delta+V in =2(Λ)\mathcal{H}=\ell^{2}({\Lambda}). A direct computation shows that

(4.48) (Δ+VmaxV)φ~=μ~φ~,(-\Delta+V_{\max}-V)\,\widetilde{\varphi}=\widetilde{\mu}\,\widetilde{\varphi},

where VmaxV={Vmaxvn}nΛV_{\max}-V=\{V_{\max}-v_{n}\}_{n\in\Lambda} is a non-negative potential and

(4.49) μ~=4d+Vmaxμ.\widetilde{\mu}=4d+V_{\max}-\mu\,.

In other words, (μ,φ)(\mu,\varphi) is an eigenpair of HH if and only if (μ~,φ~)(\widetilde{\mu},\widetilde{\varphi}) is an eigenpair of a dual operator H~:=Δ+VmaxV\widetilde{H}:=-\Delta+V_{\max}-V. This dual operator H~\widetilde{H} is the same type of discrete Schrödinger operator as HH, only with a different potential (and also taking values in [0,Vmax][0,V_{\max}]). We can define a dual landscape function u~\widetilde{u} satisfying (H~u~)n=1(\widetilde{H}\widetilde{u})_{n}=1, and a dual box-counting function Nu~(μ;H~)N_{\widetilde{u}}(\mu;\widetilde{H}) as in (1.3) for H~\widetilde{H}.

It is easy to check that the cardinality of the eigenvalues of HH which are smaller than or equal to μ\mu is the difference of the volume of Λ\Lambda and the cardinality of the eigenvalues of H~\widetilde{H} which are smaller than μ~\widetilde{\mu}. Therefore,

(4.50) N(μ;H)=1N(μ~;H~),N(\mu;H)=1-N^{-}\bigl{(}\widetilde{\mu};\widetilde{H}\bigr{)},

where N(;H)N(\cdot;H) and N(;H~)N^{-}(\cdot;{\widetilde{H}}) are the finite volume integrated density of states for HH and H~\widetilde{H}, respectively. Here, the counting N(;H~)N^{-}(\cdot;{\widetilde{H}}) is defined for eigenvalues strictly less than μ\mu, which is different from the definition of N(;H)N(\cdot;H) in (1.2). If VV is the Anderson-type potential with common distribution F(δ)=(vnδ)F(\delta)=\mathbb{P}(v_{n}\leq\delta), thenVmaxVV_{\max}-V is also an Anderson-type potential, with common distribution (Vmaxvnδ)\mathbb{P}(V_{\max}-v_{n}\leq\delta). We denote by F~(δ)=(Vmaxvn<δ)=1F(Vmaxδ)\widetilde{F}(\delta)=\mathbb{P}(V_{\max}-v_{n}<\delta)=1-F(V_{\max}-\delta). We now apply Theorems 1, 2, 4 and 7 to the dual operator H~\widetilde{H} and the dual counting function N(μ~;H~)N(\widetilde{\mu};\widetilde{H}) for μ~\widetilde{\mu} near 0. All the estimates still hold if we replace N,FN,F by NN^{-} and F~\widetilde{F}. In particular, the first part of Theorem 4 implies that there are constants c~5,c~6\widetilde{c}_{5},\widetilde{c}_{6} depending on dd, the expectation of the random variables, and VmaxV_{\max} such that for all μ~>0\widetilde{\mu}>0,

(4.51) c~5𝔼Nu~(c~6μ~;H~)𝔼N(μ~;H~)𝔼Nu~(C1μ~;H~).\widetilde{c}_{5}\,\mathbb{E}N_{\widetilde{u}}\,(\widetilde{c}_{6}\,\widetilde{\mu};\widetilde{H})\leq\,\mathbb{E}N^{-}\,(\widetilde{\mu};\widetilde{H})\leq\,\mathbb{E}N_{\widetilde{u}}(\,C_{1}\widetilde{\mu};\widetilde{H}).

Therefore, by (4.49) and (4.50), one has for all μ~=4d+Vmaxμ\widetilde{\mu}=4d+V_{\max}-\mu,

1𝔼Nu~(C1μ~;H~)𝔼N(μ;H)1c~5𝔼Nu~(c~6μ~;H~),1-\,\mathbb{E}N_{\widetilde{u}}\,(C_{1}\widetilde{\mu};\widetilde{H})\leq\mathbb{E}N(\mu;H)\leq 1-\widetilde{c}_{5}\,\mathbb{E}N_{\widetilde{u}}\,(\widetilde{c}_{6}\,\widetilde{\mu};\widetilde{H}),

which yields (1.12) in Corollary 2.

Appendix A Discrete Laplacian and harmonic functions

A.1. Maximum principle for sub-solutions

Lemma A.1 (The maximum principle for subharmonic functions).

Let Q=a1,b1××ad,bddQ=\llbracket a_{1},b_{1}\rrbracket\times\cdots\times\llbracket a_{d},b_{d}\rrbracket\subset{\mathbb{Z}}^{d} be a box in d{\mathbb{Z}}^{d} and let the inner boundary Q\partial Q be defined as in (2.11), and let QQ\partial^{\circ}Q\subset\partial Q be the flat part of the boundary as defined in (2.12). Let V={vn}nQV=\{v_{n}\}_{n\in Q} be a non-negative potential on QQ. A vector f={fn}nQf=\{f_{n}\}_{n\in Q} is called a sub-solution, on (the interior of) QQ if

(Δf)n+vnfn0,nQ\Q.-(\Delta f)_{n}+v_{n}f_{n}\geq 0,\ \ n\in Q\backslash\partial Q.

If ff is a sub-solution, then the minimum of fnf_{n} in Q\E(Q)Q\backslash E(Q) must be attained on Q\partial^{\circ}Q, i.e.,

(A.1) minnQ\QfnminnQfn.\min_{n\in Q\backslash\partial Q}\,f_{n}\,\geq\min_{n\in\partial^{\circ}Q}\,f_{n}.
Proof.

Let m=minnQfnm=\min_{n\in\partial^{\circ}Q}\,f_{n}. It is enough to prove that whenever fn0f_{n}\geq 0 for all nQn\in\partial^{\circ}Q, we have fn0f_{n}\geq 0 for all nQ\Qn\in Q\backslash\partial Q. Suppose not, then a:=minnQ\Qfn<0.-a:=\min_{n\in Q\backslash\partial Q}\,f_{n}<0. Let jQ\Qj\in Q\backslash\partial Q be such that the minimum is attained, i.e., fj=a<0f_{j}=-a<0 and fj±eifj,1idf_{j\pm e_{i}}\geq f_{j},1\leq i\leq d. Then (Δf)j+vjfj0-(\Delta f)_{j}+v_{j}f_{j}\geq 0 implies that 2dfj+vjfj1id(fj+ei+fjei)2d(a)2df_{j}+v_{j}f_{j}\geq\sum_{1\leq i\leq d}(f_{j+e_{i}}+f_{j-e_{i}})\geq 2d\cdot(-a). Therefore, fj±ei=fj=a,1idf_{j\pm e_{i}}=f_{j}=-a,1\leq i\leq d, and vj=0v_{j}=0. If any of j±eij\pm e_{i} belongs to the flat boundary Q\partial^{\circ}Q, then it is a contradiction with the assumption that fn0f_{n}\geq 0 for all nQn\in\partial^{\circ}Q. If not, then we pick any of them and repeat the procedure until eventually, after a finite number of steps, we reach the boundary and arrive at the contradiction again. ∎

There will be several direct corollaries of the above maximum principle. We will simply list them as independent lemmas and omit the details for the proof.

Lemma A.2 (Positivity of solutions for periodic boundary conditions).

Let Λ=(/K)d\Lambda=({\mathbb{Z}}/K{\mathbb{Z}})^{d}. If V={vn}nΛV=\{v_{n}\}_{n\in\Lambda} is a non-negative potential which is not constantly zero and (Δf+Vf)n0(-\Delta f+Vf)_{n}\geq 0 for all nΛn\in\Lambda, then fn0f_{n}\geq 0 for all nΛn\in\Lambda.

See [WZ], Lemma 2.12, for the proof.

Lemma A.3 (Maximum principle for discrete harmonic functions).

Let Q=a1,b1××ad,bddQ=\llbracket a_{1},b_{1}\rrbracket\times\cdots\times\llbracket a_{d},b_{d}\rrbracket\subset{\mathbb{Z}}^{d} be a box in d{\mathbb{Z}}^{d} and let Q,Q\partial Q,\partial^{\circ}Q be defined as in Lemma A.1. Suppose f={fn}nQf=\{f_{n}\}_{n\in Q} is a discrete harmonic function on (the interior of) QQ, i.e.,

(Δf)n=0,nQ\Q.(\Delta f)_{n}=0,\ \ n\in Q\backslash\partial Q.

Then for all nQ\Qn\in Q\backslash\partial Q

(A.2) minmQfmminmQfmfnmaxmQfmmaxmQfm.\min_{m\in\partial Q}\,f_{m}\leq\min_{m\in\partial^{\circ}Q}\,f_{m}\leq f_{n}\,\leq\max_{m\in\partial^{\circ}Q}\,f_{m}\leq\max_{m\in\partial Q}\,f_{m}.

This is a direct application of Lemma A.1, to ff and f-f. We only state maximum principles as above for the boxes in d{\mathbb{Z}}^{d} for simplicity, it is not hard to check that the same conclusion would hold for more general domains in d{\mathbb{Z}}^{d}, as long as they are “connected” with respect to the discrete Laplacian operator in a suitable sense. In particular, it works for the “annular” domain given by the difference of two cubes, A=Q2Q1A=Q_{2}\setminus Q_{1}. To be precise, if the boundary of AA is defined in the same as in (2.11), i.e., A={nA:n+eiAorneiAforsomeei}\partial A=\{n\in A:\,n+e_{i}\not\in A\ {\rm or}\ n-e_{i}\not\in A\ {\rm for\ some\ }e_{i}\} and (Δf)n=0(\Delta f)_{n}=0 for nA\An\in A\backslash\partial A, then

minmAfmfnmaxmAfm, for all nA\A.\min_{m\in\partial A}\,f_{m}\leq f_{n}\,\leq\max_{m\in\partial A}\,f_{m},\quad\mbox{ for all }n\in A\backslash\partial A.

A.2. The discrete Poincaré inequality

The result essentially can be generalized to any “connected” region in d{\mathbb{Z}}^{d}. We only need the version on a rectangular domain.

Lemma A.4.

Let Q=I1××IdQ=I_{1}\times\cdots\times I_{d} be a a rectangular domain in d{\mathbb{Z}}^{d}, where Ii=ai,biI_{i}=\llbracket a_{i},b_{i}\rrbracket for some ai<bia_{i}<b_{i}\in{\mathbb{Z}} and i=Card(Ii)\ell_{i}={\rm Card}\left(I_{i}\right)\in\mathbb{N}, i=1,di=1\cdots,d. For any (real-valued) sequence {fn}nQ\{f_{n}\}_{n\in Q}, let |Q|=Card(Q)|Q|={\rm Card}\left(Q\right) and f¯Q=1|Q|nQfn\bar{f}_{Q}=\frac{1}{|Q|}\sum_{n\in Q}f_{n}. Then

(A.3) nQ(fnf¯Q)2d2max2nQfn2.\displaystyle\sum_{n\in Q}(f_{n}-\bar{f}_{Q})^{2}\leq\frac{d}{2}\,\ell^{2}_{\max}\sum_{n\in Q}\|\nabla f_{n}\|^{2}.
Proof.

Without loss of generality, we assume that Q=1,1×1,2×1,dQ=\llbracket 1,\ell_{1}\rrbracket\times\llbracket 1,\ell_{2}\rrbracket\cdots\times\llbracket 1,\ell_{d}\rrbracket. It is enough to prove (A.3) for f¯Q=0\bar{f}_{Q}=0. It is easy to check that

(A.4) mQnQ(fmfn)2=2|Q|nQfn2.\displaystyle\sum_{m\in Q}\sum_{n\in Q}(f_{m}-f_{n})^{2}=2|Q|\sum_{n\in Q}f_{n}^{2}.

For m=(m1,,md)Q,n=(n1,,nd)Qm=(m_{1},\cdots,m_{d})\in Q,n=(n_{1},\cdots,n_{d})\in Q, let Γ(n,m)={γ1γ2γd+1}\Gamma(n,m)=\{\gamma^{1}\to\gamma^{2}\to\cdots\to\gamma^{d+1}\} be a discrete path in d{\mathbb{Z}}^{d} connecting nn and mm, defined taking the maximal steps along every coordinate. That is, all vertices γiQ\gamma^{i}\in Q, are given by γ1=n\gamma^{1}=n, γi+1=γi+tiei,i=1,d\gamma^{i+1}=\gamma^{i}+t_{i}e_{i},i=1\cdots,d, γd+1=m\gamma^{d+1}=m, and the all edges EiE^{i} connecting the consecutive vertices are parallel to eie_{i}. Then

(A.5) (fmfn)2=(1id(fγi+1fγi))2d1id(fγi+1fγi)2.\displaystyle(f_{m}-f_{n})^{2}=\Bigl{(}\sum_{1\leq i\leq d}(f_{\gamma^{i+1}}-f_{\gamma^{i}})\Bigr{)}^{2}\leq d\sum_{1\leq i\leq d}(f_{\gamma^{i+1}}-f_{\gamma^{i}})^{2}.

We claim that for each i=1,,di=1,\cdots,d,

(A.6) m,nQ(fγi+1fγi)2max2|Q|nQ|ifn|2,\displaystyle\sum_{m,n\in Q}(f_{\gamma^{i+1}}-f_{\gamma^{i}})^{2}\leq\ell_{\max}^{2}\,|Q|\,\sum_{n\in Q}|\nabla_{i}f_{n}|^{2},

where max=maxjj\ell_{\max}=\max_{j}\ell_{j}. Then (A.3) follows from (A.4)-(A.6).

We now prove (A.6) for i=1i=1. Fix γd+1=m=(m1,,md)\gamma^{d+1}=m=(m_{1},\cdots,m_{d}). Write γ1=n=(n1,nˇ)\gamma^{1}=n=(n_{1},\check{n}), where nˇ=(n2,,nd)1,2××1,d:=Qˇ\check{n}=(n_{2},\cdots,n_{d})\in\llbracket 1,\ell_{2}\rrbracket\times\cdots\times\llbracket 1,\ell_{d}\rrbracket:=\check{Q}. Assume that t1=m1n10t_{1}=m_{1}-n_{1}\geq 0. Then γ2=γ1+t1e1=(m1,nˇ)\gamma^{2}=\gamma^{1}+t_{1}e_{1}=(m_{1},\check{n}). Write fk=f(k1,,kd)f_{k}=f(k_{1},\cdots,k_{d}) for k=(k1,,kd)dk=(k_{1},\cdots,k_{d})\in{\mathbb{Z}}^{d}. Direct computation shows that

(fγ2fγ1)2=(k1=n1m11f(k1+1,nˇ)f(k1,nˇ))2\displaystyle(f_{\gamma^{2}}-f_{\gamma^{1}})^{2}=\left(\sum_{k_{1}=n_{1}}^{m_{1}-1}f(k_{1}+1,\check{n})-f(k_{1},\check{n})\right)^{2}\leq |t1|k1=n1m11(f(k1+1,nˇ)f(k1,nˇ))2\displaystyle|t_{1}|\sum_{k_{1}=n_{1}}^{m_{1}-1}\big{(}f(k_{1}+1,\check{n})-f(k_{1},\check{n})\big{)}^{2}
\displaystyle\leq maxk1=11(1f(k1,nˇ))2.\displaystyle\ell_{\max}\sum_{k_{1}=1}^{\ell_{1}}\big{(}\nabla_{1}f(k_{1},\check{n})\big{)}^{2}.

The same estimate holds for t1<0t_{1}<0. Therefore, fix mm, summing over nQn\in Q gives

nQ(fγ2fγ1)2=nˇQˇn1=11(fγ2fγ1)2\displaystyle\sum_{n\in Q}(f_{\gamma^{2}}-f_{\gamma^{1}})^{2}=\sum_{\check{n}\in\check{Q}}\sum_{n_{1}=1}^{\ell_{1}}(f_{\gamma^{2}}-f_{\gamma^{1}})^{2}\leq (n1=11max)(nˇQˇk1=11(1f(k1,nˇ))2)\displaystyle\Big{(}\sum_{n_{1}=1}^{\ell_{1}}\ell_{\max}\Big{)}\Big{(}\sum_{\check{n}\in\check{Q}}\sum_{k_{1}=1}^{\ell_{1}}\big{(}\nabla_{1}f(k_{1},\check{n})\big{)}^{2}\Big{)}
\displaystyle\leq max2nQ|1fn|2.\displaystyle\ell_{\max}^{2}\sum_{n\in Q}|\nabla_{1}f_{n}|^{2}.

Then summing over mQm\in Q gives

m,nQ(fγ2fγ1)2|Q|max2nQ|1fn|2,\sum_{m,n\in Q}(f_{\gamma^{2}}-f_{\gamma^{1}})^{2}\leq|Q|\ell_{\max}^{2}\sum_{n\in Q}|\nabla_{1}f_{n}|^{2},

which proves (A.6) for i=1i=1. The cases i=2,,di=2,\cdots,d can be proved in a similar manner. This completes the proof of (A.6) and Lemma A.3.

A.3. Discrete cut-off functions

Let QQ be a cube of side length  R3R\geq 3 on d{\mathbb{Z}}^{d} and let jmax=R/3j_{\max}=\left\lfloor R/3\right\rfloor. Let Q\partial Q and Q/3Q/3 be given by (2.11) and (2.13). Let the distance dist(n,m)=|nm|{\rm dist}(n,m)=|n-m|_{\infty} be measured by the infinity norm on d{\mathbb{Z}}^{d}. Let (j),1jjmax{\mathcal{I}}(j),1\leq j\leq j_{\max}, be a d1d-1 dimensional subset of QQ which is distance jj away from Q/3Q/3:

(j):={nQ|dist(n,Q/3)=j}.{\mathcal{I}}(j):=\left\{n\in Q|\,\mathrm{dist}\,(n,Q/3)=j\ \right\}.

By the definition of Q/3Q/3 in (2.13), the side length of Q/3Q/3 satisfies (Q/3)=R/3R/3.\ell(Q/3)=\left\lfloor R/3\right\rfloor\leq R/3. It is easy to check that Q/3Q/3 and all (j){\mathcal{I}}(j) are pairwise disjoint for j=1,,jmaxj=1,\cdots,j_{\max}. And 3(Q/3)=Q/3(j=1jmax(j))Q3(Q/3)=Q/3\bigcup\Bigl{(}\bigcup_{j=1}^{j_{\max}}{\mathcal{I}}(j)\Bigr{)}\subset Q.

Now we can define the cut-off function χ={χn}\chi=\{\chi_{n}\} as

(A.7) χn={1,nQ/3,13Rj,n(j),j=1,,jmax,0,n3(Q/3).\displaystyle\chi_{n}=\begin{cases}1,&n\in Q/3,\\ 1-\frac{3}{R}j,&n\in{\mathcal{I}}(j),\ j=1,\cdots,j_{\max},\\ 0,&n\notin 3(Q/3).\end{cases}

It is easy to see that |χn+eiχn|3/R|\chi_{n+e_{i}}-\chi_{n}|\leq{3}/{R} if n,n+eiQn,n+e_{i}\in Q, and χn+eiχn=0\chi_{n+e_{i}}-\chi_{n}=0 otherwise, for all 1id1\leq i\leq d.

A.4. Dirichlet problem on a cube

We study the Dirichlet problem for the discrete Laplacian on a cube in d{\mathbb{Z}}^{d}. Recall the definitions of Q(r;ξ)Q(r;\xi), Q(r)\partial Q(r), Q(r)\partial^{\circ}Q(r), for ξ=(ξ1,,ξd)d\xi=(\xi_{1},\cdots,\xi_{d})\in{\mathbb{Z}}^{d}, and r0r\in{\mathbb{Z}}_{\geq 0}, given right before Lemma 4.2.

Lemma A.5 (Green’s formula).

For any f,g=2(Λ)f,g\in\mathcal{H}=\ell^{2}({\Lambda}),

(A.8) nQ(r1)gn(Δf)n=\displaystyle\sum_{n\in Q(r-1)}g_{n}(\Delta f)_{n}= n,n+eiQ(r)(g)n(f)n+nQ(r)gnf𝐍(n)\displaystyle-\sum_{n,n+e_{i}\in Q(r)}(\nabla g)_{n}(\nabla f)_{n}+\sum_{{n\in\partial^{\circ}Q(r)}}g_{n}\frac{\partial f}{\partial{\bf N}}(n)
(A.9) =\displaystyle= 12n,mQ(r)|mn|=1(gmgn)(fmfn)+nQ(r),mQ(r1):|mn|=1gn(fnfm).\displaystyle-\frac{1}{2}\sum_{\begin{subarray}{c}n,m\in Q(r)\\ |m-n|=1\end{subarray}}(g_{m}-g_{n})(f_{m}-f_{n})+\sum_{\begin{subarray}{c}n\in\partial^{\circ}Q(r),\,m\in\partial Q(r-1):\\ |m-n|=1\end{subarray}}g_{n}(f_{n}-f_{m}).

As a consequence,

nQ(r1)gn(Δf)n\displaystyle\sum_{n\in Q(r-1)}g_{n}(\Delta f)_{n}- nQ(r1)fn(Δg)n\displaystyle\sum_{n\in Q(r-1)}f_{n}(\Delta g)_{n}
(A.10) =\displaystyle= nQ(r),mQ(r1)|mn|=1fn(gngm)+nQ(r),mQ(r1)|mn|=1gn(fnfm).\displaystyle-\sum_{\begin{subarray}{c}n\in\partial^{\circ}Q(r),m\in\partial Q(r-1)\\ |m-n|=1\end{subarray}}f_{n}(g_{n}-g_{m})+\sum_{\begin{subarray}{c}n\in\partial^{\circ}Q(r),m\in\partial Q(r-1)\\ |m-n|=1\end{subarray}}g_{n}(f_{n}-f_{m}).

The Green’s formula for discrete graphs is rather standard, existing in various lecture notes, e.g. Theorem 1.37 in [Ba], see also in [Ch, Gu]. We omit the proof here.

Given {fn}nQ(r1)\{f_{n}\}_{n\in Q(r-1)} and {hn}nQ(r)\{h_{n}\}_{n\in\partial Q(r)}, we proceed to solve the linear system on Q(r)Q(r)

(A.11) {(Δu)n=fn,nQ(r1),un=hn,nQ(r).\displaystyle\begin{cases}-(\Delta u)_{n}=f_{n},\ \ n\in Q(r-1),\\ u_{n}=h_{n},\ n\in\partial Q(r).\end{cases}

The problem can be decomposed into the following two systems, which give us the discrete Poisson Kernel and the discrete Green’s function for the Dirichlet Laplacian.

The discrete Poisson kernel Pr(n;m):Q(r)×Q(r)[0,1]P_{r}(n;m):Q(r)\times\partial Q(r)\to[0,1] is the unique solution to the system

(A.12) {ΔPr(n,m)=0,nQ(r1),Pr(n,m)=δm(n),nQ(r),\displaystyle\begin{cases}&\Delta P_{r}(n,m)=0,\,n\in Q(r-1),\\ &P_{r}(n,m)=\delta_{m}(n),\ n\in\partial Q(r),\end{cases}

for a fixed mQ(r)m\in\partial Q(r). Similarly, the discrete Green’s function with pole at mm, Gr(n,m):Q(r)[0,1]G_{r}(n,m):Q(r)\to[0,1] is the unique solution to the system

(A.13) {ΔGr(n,m)=δm,nQ(r1),Gr(n,m)=0,nQ(r),\displaystyle\begin{cases}-\Delta G_{r}(n,m)=\delta_{m},\ &n\in Q(r-1),\\ G_{r}(n,m)=0,\ &n\in\partial Q(r),\end{cases}

for a fixed mQ(r1)m\in Q(r-1), Consider Δ-\Delta with zero boundary condition as an invertible matrix of the size |Q(r1)|×|Q(r1)||Q(r-1)|\times|Q(r-1)|. Clearly, for n,mQ(r1)n,m\in Q(r-1), G(n,m)=(Δ)1(n,m)=G(m,n)G(n,m)=(\Delta)^{-1}(n,m)=G(m,n) since Δ\Delta is self-adjoint.

Moreover, for fixed mQ(r)m\in\partial Q(r) and mQ(r)m^{\prime}\in Q(r), if we apply the Green’s formula (A.10) to gn=Pr(n,m)g_{n}=P_{r}(n,m) and fn=Gr(n,m)f_{n}=G_{r}(n,m^{\prime}), then

nQ(r1)Pr(n,m)(δm(n))=nQ(r),nQ(r1):|nn|=1δm(n)(Gr(n,m)Gr(n,m)),\sum_{n\in Q(r-1)}P_{r}(n,m)\big{(}-\delta_{m^{\prime}}(n)\big{)}=\sum_{\begin{subarray}{c}n\in\partial^{\circ}Q(r),\,n^{\prime}\in\partial Q(r-1):\\ |n^{\prime}-n|=1\end{subarray}}\delta_{m}(n)\,\big{(}G_{r}(n,m^{\prime})-G_{r}(n^{\prime},m^{\prime})\big{)},

which implies for any mQ(r)m\in\partial Q(r) and mQ(r)m^{\prime}\in Q(r),

(A.14) Pr(m,m)=Gr(n,m)=Gr(m,n),nQ(r1),|nm|=1.P_{r}(m^{\prime},m)=G_{r}(n^{\prime},m^{\prime})=G_{r}(m^{\prime},n^{\prime}),n^{\prime}\in\partial Q(r-1),|n^{\prime}-m|=1.

Notice that this can be considered as the (negative) normal derivative of Gr(m,)G_{r}(m^{\prime},\cdot) in the direction of the outward pointing to the surface of Q(r)Q(r).

Back to the system (A.11), using PrP_{r} and GrG_{r}, we can solve the system

un=mQ(r)Pr(n,m)hm+mQ(r1)Gr(n,m)fm.u_{n}=\sum_{m\in\partial Q(r)}P_{r}(n,m)h_{m}+\sum_{m^{\prime}\in Q(r-1)}G_{r}(n,m^{\prime})f_{m^{\prime}}.

In particular, we have the following integration by parts formula (Green’s identity) for any {un}nQ(r)\{u_{n}\}_{n\in\partial^{\circ}Q(r)} at the center of the box Q(r;ξ)Q(r;\xi):

(A.15) uξ=mQ(r)Pr(ξ,m)ummQ(r1)Gr(ξ,m)(Δu)m.u_{\xi}=\sum_{m\in\partial Q(r)}P_{r}(\xi,m)u_{m}-\sum_{m^{\prime}\in Q(r-1)}G_{r}(\xi,m^{\prime})(\Delta u)_{m^{\prime}}.

In particular, for any ξ\xi and rr,

(A.16) mQ(r)Pr(ξ,m)=1.\sum_{m\in\partial Q(r)}P_{r}(\xi,m)=1.

A.5. Estimates on the Green’s function and the Poisson kernel

Retain the definitions in the previous section, for RR\in\mathbb{N}, let Q(R)=Q(R;ξ)dQ(R)=Q(R;\xi)\subset{\mathbb{Z}}^{d} be the discrete cube centered at ξd\xi\in{\mathbb{Z}}^{d} of side length 2R+12R+1, and let GR(ξ,n)G_{R}(\xi,n) be the discrete Green’s function as defined in (A.13). In this part, we study the behavior of the discrete Green’s function away both from the pole ξ\xi and the boundary Q(R)\partial Q(R). We will approximate the discrete Green’s function by a continuous one to obtain the desired estimates. Let us also recall some of the definitions for the continuous case. Fix ξd\xi\in{\mathbb{Z}}^{d}, let 𝒬1=ξ+[1,1]d{\mathcal{Q}}_{1}=\xi+[-1,1]^{d} be a cube in d\mathbb{R}^{d} centered at ξ\xi of side length 22. Let 𝒢(ξ,){\mathcal{G}}(\xi,\cdot) be the continuous Green’s function on the cube 𝒬1{\mathcal{Q}}_{1} with zero boundary conditions:

(A.17) {Δc𝒢(ξ,x)=δξc(x),𝒢(ξ,x)=0,x𝒬1,\displaystyle\begin{cases}-\Delta_{\rm c}{\mathcal{G}}(\xi,x)=\delta^{c}_{\xi}(x),\\ {\mathcal{G}}(\xi,x)=0,x\in\partial{\mathcal{Q}}_{1},\end{cases}

where Δc=i=1d2xi2\Delta_{c}=\sum_{i=1}^{d}\frac{\partial^{2}}{\partial x_{i}^{2}} is the standard Laplacian on d\mathbb{R}^{d}, and δξc(x)\delta^{c}_{\xi}(x) is the Dirac delta function at ξ\xi in the distribution sense. For any RR\in\mathbb{N}, consider a square mesh of size h=1Rh=\frac{1}{R} on 𝒬1{\mathcal{Q}}_{1}. Denote the collection of all the mesh points by

(A.18) Ωh={τn=ξ+nh,nR,Rd}.\Omega_{h}=\big{\{}\tau_{n}=\xi+nh,n\in\llbracket-R,R\rrbracket^{d}\,\big{\}}.

We see that Ωh\Omega_{h} is indexed (one-to-one) by nQ(R)n\in Q(R), and hence 2(Ωh)(2R+1)2\ell^{2}(\Omega_{h})\cong\mathbb{R}^{(2R+1)^{2}}. For τnΩh\tau_{n}\in\Omega_{h}, let

(A.19) 𝒢h(ξ,τn):=1hd2GR(ξ,n)=1hd2GR(n,ξ).{\mathcal{G}}^{h}(\xi,\tau_{n}):=\frac{1}{h^{d-2}}G_{R}(\xi,n)=\frac{1}{h^{d-2}}G_{R}(n,\xi).

It is easy to verify that the equation for GR(n,ξ)G_{R}(n,\xi) in (A.13) implies that

(A.20) {(Δ𝒢h)(ξ,τn):=|nm|=1(𝒢h(ξ,τm)𝒢h(ξ,τn))=h2dδξ(n),nQ(R;ξ),𝒢h(ξ,τn)=0,nQ(R;ξ).\begin{cases}-(\Delta{\mathcal{G}}^{h})(\xi,\tau_{n}):=-\sum_{|n-m|=1}\big{(}{\mathcal{G}}^{h}(\xi,\tau_{m})-{\mathcal{G}}^{h}(\xi,\tau_{n})\big{)}={h^{2-d}}\delta_{\xi}(n),\quad\ n\in Q(R;\xi),\\ {\mathcal{G}}^{h}(\xi,\tau_{n})=0,\quad n\in\partial Q(R;\xi).\end{cases}

We see that 𝒢h(ξ,)2(Ωh){\mathcal{G}}^{h}(\xi,\cdot)\in\ell^{2}(\Omega_{h}) is the finite difference approximation to the solution of the continuous problem (A.17). The approximation can be quantified as follows.

Lemma A.6.

There are positive dimensional constants C,h0C,h_{0} such that if hh0h\leq h_{0}, then

(A.21) |𝒢h(ξ,τn)𝒢(ξ,τn)|C|logh|d+3h2for allτnΩh(𝒬1\𝒬1/2).\displaystyle|{\mathcal{G}}^{h}(\xi,\tau_{n})-{\mathcal{G}}(\xi,\tau_{n})|\leq\,C|\log h|^{d+3}h^{2}\ \ \ {\textrm{for all}}\ \tau_{n}\in\Omega_{h}\cap\left({\mathcal{Q}}_{1}\backslash{\mathcal{Q}}_{1/2}\right).
Remark A.1.

Such an approximation is proved in a rectangular domain in 2\mathbb{R}^{2} in [La]. It was later generalized to the interior of a domain of any dimension with smooth boundary by Schatz and Wahlbin, see Theorem 6.1 [SW], using the finite elements approach. The method in [SW] potentially can be generalized to any convex polyhedral domains, up to the boundary. Here we present a direct proof using the series expansion of 𝒢h{\mathcal{G}}^{h} and 𝒢{\mathcal{G}}. Similar estimate also holds for 𝒢h(y,x)𝒢(y,x){\mathcal{G}}^{h}(y,x)-{\mathcal{G}}(y,x) where the pole yy is not far away from the center ξ\xi. We will only deal with the case y=ξy=\xi which will be enough for our use.

Proof.

Without loss of generality we assume that ξ=0\xi=0. In this case, the mesh points τn=nh\tau_{n}=nh and Ωh=1RQ(R)𝒬1\Omega_{h}=\frac{1}{R}Q(R)\subset{\mathcal{Q}}_{1}, where h=1/Rh=1/R. Due to the symmetry of the problem, it is also enough to prove (A.21) on the upper half cube 𝒬1+={x𝒬1:xd0}{\mathcal{Q}}^{+}_{1}=\{x\in{\mathcal{Q}}_{1}:x_{d}\geq 0\}. Below, we construct the analytic series representations of 𝒢(0,x){\mathcal{G}}(0,x) and 𝒢h(0,x){\mathcal{G}}^{h}(0,x) on 𝒬1+{\mathcal{Q}}^{+}_{1}.

For the continuous case, the analytic expression of 𝒢(0,x){\mathcal{G}}(0,x) is well known by the method of the partial eigenfunction representation. Throughout the rest of the proof, we denote k=(k1,,kd1)+d1k=(k_{1},\cdots,k_{d-1})\in{\mathbb{Z}}_{+}^{d-1}, and x=(x~,xd)x=(\widetilde{x},x_{d}) where x~=(x1,,xd1)d1\widetilde{x}=(x_{1},\cdots,x_{d-1})\in\mathbb{R}^{d-1}. Let

(A.22) fk(x~)=i=1d1sin(kiπ2(xi+1)),andαk=π2k:=π2(i=1d1ki2)1/2.f_{k}(\widetilde{x})=\prod_{i=1}^{d-1}\sin\Big{(}\frac{k_{i}\pi}{2}(x_{i}+1)\Big{)},\ {\rm and}\ \alpha_{k}=\frac{\pi}{2}\|k\|:=\frac{\pi}{2}\Big{(}\sum_{i=1}^{d-1}k_{i}^{2}\Big{)}^{1/2}.

By separation of variables, for x=(x~,xd)𝒬1+,0<xd<1x=(\widetilde{x},x_{d})\in{\mathcal{Q}}^{+}_{1},0<x_{d}<1,

𝒢(0,x)=\displaystyle{\mathcal{G}}(0,x)= k+d11αksinh(2αk)sinh(αk)sinh(αk(1xd))fk(0)fk(x~)\displaystyle\sum_{k\in{\mathbb{Z}}_{+}^{d-1}}\frac{1}{\alpha_{k}\sinh(2\alpha_{k})}\sinh(\alpha_{k})\sinh\big{(}\alpha_{k}(1-x_{d})\big{)}f_{k}(0)f_{k}(\widetilde{x})
(A.23) =\displaystyle= k+d1sinh(αk(1xd))2αkcosh(αk)i=1d1sin(kiπ2)sin(kiπ2(xi+1)).\displaystyle\sum_{k\in{\mathbb{Z}}_{+}^{d-1}}\frac{\sinh(\alpha_{k}(1-x_{d}))}{2\alpha_{k}\cosh(\alpha_{k})}\,\prod_{i=1}^{d-1}\sin\bigl{(}\frac{k_{i}\pi}{2}\bigr{)}\sin\bigl{(}\frac{k_{i}\pi}{2}(x_{i}+1)\bigr{)}.

The partial eigenfunction representation can be used to derive a similar formula for 𝒢h(0,τn){\mathcal{G}}^{h}(0,\tau_{n}) solving (A.20) on the finite dimensional space. We may abuse the notation and write 𝒢h(0,n)=𝒢h(0,τn)=𝒢h(0,nh){\mathcal{G}}^{h}(0,n)={\mathcal{G}}^{h}(0,\tau_{n})={\mathcal{G}}^{h}(0,nh) when it is clear. Notice that 𝒢h(0,n){\mathcal{G}}^{h}(0,n) satisfies zero boundary condition on Q(R)\partial Q(R), it is enough to consider 𝒢h(0,n){\mathcal{G}}^{h}(0,n) as a discrete function only for nQ(R1)=R+1,R1dn\in Q(R-1)=\llbracket-R+1,R-1\rrbracket^{d}. Similar to the notations for the continuous case, we write n=(n~,nd)n=(\widetilde{n},n_{d}), where n~=(n1,,nd1)R+1,R1d1\widetilde{n}=(n_{1},\cdots,n_{d-1})\in\llbracket-R+1,R-1\rrbracket^{d-1}, and ndR+1,R1n_{d}\in\llbracket-R+1,R-1\rrbracket. Denote by TR:=1,2R1d1T_{R}:=\llbracket 1,2R-1\rrbracket^{d-1}. We first construct a basis for the subspace 2(R+1,R1d1)\ell^{2}(\llbracket-R+1,R-1\rrbracket^{d-1}). For k=(k1,,kd)TRk=(k_{1},\cdots,k_{d})\in T_{R} and n~=(n1,,nd1)R+1,R1d1\widetilde{n}=(n_{1},\cdots,n_{d-1})\in\llbracket-R+1,R-1\rrbracket^{d-1}, let

(A.24) fkh(n~)=hd1i=1d1sin(kiπ2(nih+1)).f^{h}_{k}(\widetilde{n})={\sqrt{h}}^{\,d-1}\prod_{i=1}^{d-1}\sin\bigl{(}\frac{k_{i}\pi}{2}(n_{i}h+1)\bigr{)}.

We claim that {fkh}kTR\{f^{h}_{k}\}_{k\in T_{R}} form a normalized basis for (2R1)d1\mathbb{R}^{(2R-1)^{d-1}}. This can be verified by direct computations of the finite dimensional inner product. For ni,min_{i},m_{i}\in{\mathbb{Z}}, let θ=πh(nimi)/2,φ=π(hni+hmi+2)/2\theta={\pi h(n_{i}-m_{i})}/{2},\varphi={\pi(hn_{i}+hm_{i}+2)}/{2}.

If ni=min_{i}=m_{i}, then

(A.25) 2ki=12R1sinkiπ(nih+1)2sinkiπ(mih+1)2=ki=12R1cos(kiθ)ki=12R1cos(kiφ)=(2R1)+12sin(2R1/2)φ2sinφ2=2R.2\sum_{k_{i}=1}^{2R-1}\sin\frac{k_{i}\pi(n_{i}h+1)}{2}\sin\frac{k_{i}\pi(m_{i}h+1)}{2}=\sum_{k_{i}=1}^{2R-1}\cos(k_{i}\theta)-\sum_{k_{i}=1}^{2R-1}\cos(k_{i}\varphi)\\ =(2R-1)+\frac{1}{2}-\frac{\sin(2R-1/2)\varphi}{2\sin\frac{\varphi}{2}}=2R.

If nimin_{i}\neq m_{i}, then 2Rθ=π(nimi),2Rφ=π(ni+mi)+2πRπ2R\theta=\pi(n_{i}-m_{i}),2R\varphi=\pi(n_{i}+m_{i})+2\pi R\in\pi{\mathbb{Z}}. Hence,

2ki=12R1sinkiπ(nih+1)2sinkiπ(mih+1)2=sin(2R1/2)θ2sinθ2sin(2R1/2)φ2sinφ2=sin(πni)sin(πmi)=0.2\sum_{k_{i}=1}^{2R-1}\sin\frac{k_{i}\pi(n_{i}h+1)}{2}\sin\frac{k_{i}\pi(m_{i}h+1)}{2}\\ =\frac{\sin(2R-1/2)\theta}{2\sin\frac{\theta}{2}}-\frac{\sin(2R-1/2)\varphi}{2\sin\frac{\varphi}{2}}=-\sin(\pi n_{i})\sin(\pi m_{i})=0.

Therefore,

ki=12R1hsinkiπ(nih+1)2hsinkiπ(mih+1)2=δ0(nimi),\sum_{k_{i}=1}^{2R-1}\sqrt{h}\sin\frac{k_{i}\pi(n_{i}h+1)}{2}\,\sqrt{h}\sin\frac{k_{i}\pi(m_{i}h+1)}{2}=\delta_{0}(n_{i}-m_{i}),

which implies for n~,m~R+1,R1d1\widetilde{n},\widetilde{m}\in\llbracket-R+1,R-1\rrbracket^{d-1},

kTRfkh(n~)fkh(m~)=i=1d1δ0(nimi)=δ0(n~m~).\sum_{k\in T_{R}}f^{h}_{k}(\widetilde{n})f^{h}_{k}(\widetilde{m})=\prod_{i=1}^{d-1}\delta_{0}(n_{i}-m_{i})=\delta_{0}(\widetilde{n}-\widetilde{m}).

This shows that the (2R1)d1×(2R1)d1(2R-1)^{d-1}\times(2R-1)^{d-1} dimensional matrix U(k,n~):=fkh(n~)U(k,\widetilde{n}):=f^{h}_{k}(\widetilde{n}) is unitary. Hence, its column vectors {fkh}kTR\{f^{h}_{k}\}_{k\in T_{R}} form a normalized basis.

Now we fix ndn_{d} and expand 𝒢h(0,nh){\mathcal{G}}^{h}(0,nh) with respect to the normalized basis {fkh}kTR\{f^{h}_{k}\}_{k\in T_{R}}:

(A.26) 𝒢h(0,n)=𝒢h(0,nh)=kTRFk(nd)fkh(n~),{\mathcal{G}}^{h}(0,n)={\mathcal{G}}^{h}(0,nh)=\sum_{k\in T_{R}}F_{k}(n_{d})f^{h}_{k}(\widetilde{n}),

where Fk(nd)F_{k}(n_{d}) is the coefficient function to be solved. We also expand the dd dimensional discrete delta function as δ0(n~,nd)=kδ0(nd)fkh(0)fkh(n~)\delta_{0}(\widetilde{n},n_{d})=\sum_{k}\delta_{0}(n_{d})f_{k}^{h}(0)f_{k}^{h}(\widetilde{n}).

Write the discrete Laplacian as Δ=Δ~+Δd\Delta=\widetilde{\Delta}+\Delta^{d}, where Δ~\widetilde{\Delta} is the second order difference with respect to the first d1d-1 variables n~=(n1,,nd1)\widetilde{n}=(n_{1},\cdots,n_{d-1}) and Δd\Delta^{d} is the second order difference with respect to the last variable ndn_{d}. Applying Δ~\widetilde{\Delta} to fkh(n~)f^{h}_{k}(\widetilde{n}) gives

Δ~fkh(n~)=fkh(n~)i=1d1(2coskiπh22).\widetilde{\Delta}f^{h}_{k}(\widetilde{n})=f^{h}_{k}(\widetilde{n})\,\sum_{i=1}^{d-1}\Bigl{(}2\cos\frac{k_{i}\pi h}{2}-2\Bigr{)}.

Combing with the expansion (A.26), we obtain

Δ𝒢h(0,n)=kTRFk(nd)(Δ~fkh(n~))+(ΔdFk(nd))fkh(n~)=kTR(Fk(nd)i=1d1(2coskiπh22)+(ΔdFk)(nd))fkh(n~).\Delta{\mathcal{G}}^{h}(0,n)=\sum_{k\in T_{R}}F_{k}(n_{d})\Big{(}\widetilde{\Delta}f^{h}_{k}(\widetilde{n})\Big{)}+\Big{(}\Delta^{d}F_{k}(n_{d})\Big{)}f^{h}_{k}(\widetilde{n})\\ =\sum_{k\in T_{R}}\Big{(}F_{k}(n_{d})\sum_{i=1}^{d-1}\bigl{(}2\cos\frac{k_{i}\pi h}{2}-2\bigr{)}+(\Delta^{d}F_{k})(n_{d})\Big{)}f^{h}_{k}(\widetilde{n}).

Hence, the dd dimensional equation Δ𝒢h(0,)=h2dδ0()\Delta{\mathcal{G}}^{h}(0,\cdot)=-h^{2-d}\delta_{0}(\cdot) can be reduced to a one dimensional difference equation of Fk(nd)F_{k}(n_{d}):

(A.27) Fk(nd)i=1d1(2coskiπh22)+(ΔdFk)(nd)=h2dfkh(0)δ0(nd).F_{k}(n_{d})\sum_{i=1}^{d-1}\Bigl{(}2\cos\frac{k_{i}\pi h}{2}-2\Bigr{)}+(\Delta^{d}F_{k})(n_{d})=-h^{2-d}f_{k}^{h}(0)\delta_{0}(n_{d}).

Define βk=βk(h)\beta_{k}=\beta_{k}(h) to be the positive solution solving

(A.28) 2coshβk2+i=1d1(2coskiπh22)=0.2\cosh\beta_{k}-2+\sum_{i=1}^{d-1}\left(2\cos\frac{k_{i}\pi h}{2}-2\right)=0.

For RndR-R\leq n_{d}\leq R, let v1(nd):=sinh(βk(Rnd))v_{1}(n_{d}):=\sinh(\beta_{k}(R-n_{d})) and v2(nd):=sinh(βk(R+nd))v_{2}(n_{d}):=\sinh(\beta_{k}(R+n_{d})). Then (Δdvj)(nd)=vj(nd)(2coshβk2),j=1,2(\Delta^{d}v_{j})(n_{d})=v_{j}(n_{d})(2\cosh\beta_{k}-2),j=1,2. Hence, vj,j=1,2v_{j},j=1,2 solves the homogeneous part of (A.27):

(A.29) vj(nd)i=1d1(2coskiπh22)+(Δdvj)(nd)=0,forallR+1ndR1,v_{j}(n_{d})\sum_{i=1}^{d-1}\Bigl{(}2\cos\frac{k_{i}\pi h}{2}-2\Bigr{)}+(\Delta^{d}v_{j})(n_{d})=0,\ \ {\rm for\ all}\ -R+1\leq n_{d}\leq R-1,

with boundary conditions v1(R)=0=v2(R)v_{1}(R)=0=v_{2}(-R). Define H(nd)=v1(nd)H(n_{d})=v_{1}(n_{d}) for 0ndR0\leq n_{d}\leq R, and H(nd)=v2(nd)H(n_{d})=v_{2}(n_{d}) for Rnd<0-R\leq n_{d}<0. Then H(nd)H(n_{d}) satisfies (A.29) for all R+1ndR1-R+1\leq n_{d}\leq R-1 except nd=0n_{d}=0. At nd=0n_{d}=0, the definition of H,v1H,v_{1} and v2v_{2} implies

H(0)(22coshβk)+H(1)+H(1)2H(0)=2cosh(βkR)sinh(βk).H(0)\left(2-2\cosh\beta_{k}\right)+H(1)+H(-1)-2H(0)=-2\cosh(\beta_{k}R)\sinh(\beta_{k}).

Finally, let Fk(nd)=h2dfkh(0)2cosh(βkR)sinh(βk)H(nd)F_{k}(n_{d})=\frac{h^{2-d}f_{k}^{h}(0)}{2\cosh(\beta_{k}R)\sinh(\beta_{k})}H(n_{d}). Then Fk(nd)F_{k}(n_{d}) solves the inhomogeneous equation (A.27) for all R+1ndR1-R+1\leq n_{d}\leq R-1, with zero boundary condition Fk(±R)=0F_{k}(\pm R)=0.

Together with (A.24) and (A.26), we obtain the analytic series expansion of 𝒢h(0,nh){\mathcal{G}}^{h}(0,nh), on the upper half cube 0ndR0\leq n_{d}\leq R:

𝒢h(0,nh)=\displaystyle{\mathcal{G}}^{h}(0,nh)= kTRFk(nd)fkh(n~)\displaystyle\sum_{k\in T_{R}}F_{k}(n_{d})f^{h}_{k}(\widetilde{n})
(A.30) =\displaystyle= kTRsinh(βk(Rnd))2Rsinh(βk)cosh(βkR)(i=1d1sin(kiπ2)sin(kiπ2(nih+1))).\displaystyle\sum_{k\in T_{R}}\frac{\sinh\big{(}\beta_{k}(R-n_{d})\big{)}}{2R\sinh(\beta_{k})\cosh(\beta_{k}R)}\left(\prod_{i=1}^{d-1}\sin\big{(}\frac{k_{i}\pi}{2}\big{)}\sin\big{(}\frac{k_{i}\pi}{2}(n_{i}h+1)\big{)}\right).

It was proved in [La] that for d=2d=2, and |n|>0|n|>0, one has |𝒢(0,nh)𝒢h(0,nh)|Ch2|n|2|{\mathcal{G}}(0,nh)-{\mathcal{G}}^{h}(0,nh)|\leq Ch^{2}|n|^{-2}. The method can be extended to higher dimensions. We will not bother to give the full generalization of the exact singularity of order 1/|n|21/|n|^{2}. We only need the version for RndR/2R\geq n_{d}\geq R/2 with some logarithm corrections as in (A.21). To do that, it is enough to study the asymptotic behavior of βk(h)\beta_{k}(h) in (A.28) as h0h\to 0.

Let βk\beta_{k} be given by (A.28). First, it was shown in [GuMa] (see also in [Gu]) that either βk1\beta_{k}\geq 1 or 2Rβkkd2R\beta_{k}\geq\|k\|_{d} for all R=1hR=\frac{1}{h} and 1ki2R11\leq k_{i}\leq 2R-1. We sketch the proof for reader’s convenience. If βk1\beta_{k}\leq 1, then

(A.31) 1+βk2coshβk=di=1d1cosπkih2di=1d1(1(kih2)2)=1+k2h24,1+\beta_{k}^{2}\geq\cosh\beta_{k}=d-\sum_{i=1}^{d-1}\cos\frac{\pi k_{i}h}{2}\geq d-\sum_{i=1}^{d-1}\left(1-\bigl{(}\frac{k_{i}h}{2}\bigr{)}^{2}\right)=1+\frac{\|k\|^{2}h^{2}}{4},

which gives βkkh/2\beta_{k}\geq\|k\|h/2. In (A.31), we used the elementary inequalities 1+x2coshx1+x^{2}\geq\cosh x, for x[0,1]x\in[0,1] and 1cosπxx21-\cos\pi x\geq x^{2} for x[0,1)x\in[0,1). Therefore, for any xd=nd/R[1/2,1]x_{d}=n_{d}/R\in[1/2,1], the coefficients of 𝒢h{\mathcal{G}}^{h} decays exponentially

|sinh(βk(Rnd))2Rsinh(βk)cosh(βkR)|1kek/4,or|sinh(βk(Rnd))2Rsinh(βk)cosh(βkR)|1keR/2.\left|\frac{\sinh(\beta_{k}(R-n_{d}))}{2R\sinh(\beta_{k})\cosh(\beta_{k}R)}\right|\leq\frac{1}{\|k\|}e^{-\|k\|/4},\ \ {\rm or}\ \left|\frac{\sinh(\beta_{k}(R-n_{d}))}{2R\sinh(\beta_{k})\cdot\cosh(\beta_{k}R)}\right|\leq\frac{1}{\|k\|}e^{-R/2}.

On the other hand, for any fixed kTRk\in T_{R}, we want to expand βk(h)\beta_{k}(h) in hh explicitly. Let αk=πk/2\alpha_{k}={\pi\|k\|}/{2} be as in (A.22), one has

βk=cosh1(di=1d1cosπkih2)=\displaystyle\beta_{k}=\cosh^{-1}\left(d-\sum_{i=1}^{d-1}\cos\frac{\pi k_{i}h}{2}\right)= cosh1(1+π2k2h28+O(k4h4))\displaystyle\cosh^{-1}\left(1+\frac{\pi^{2}\|k\|^{2}h^{2}}{8}+O(\|k\|^{4}h^{4})\right)
=\displaystyle= log(1+π2k2h28+πkh2+O(k3h3))\displaystyle\log\left(1+\frac{\pi^{2}\|k\|^{2}h^{2}}{8}+{\frac{\pi\|k\|h}{2}+O(\|k\|^{3}h^{3})}\right)
=\displaystyle= πkh2+O(k3h3)=αkh+O(k3h3).\displaystyle\frac{\pi\|k\|h}{2}+O(\|k\|^{3}h^{3})=\alpha_{k}h+O(\|k\|^{3}h^{3}).

Therefore, Rβk=αk+O(|logh|3h2)R\,\beta_{k}=\alpha_{k}+O(|\log h|^{3}h^{2}) for kCd|logh|\|k\|\leq C_{d}|\log h| (with any constant CdC_{d} only depending on the dimension). For any xd=nd/R[1/2,1]x_{d}=n_{d}/R\in[1/2,1] and t=1xd[0,1/2]t=1-x_{d}\in[0,1/2], we compare the coefficients of 𝒢{\mathcal{G}} and 𝒢h{\mathcal{G}}^{h} in (A.23) and (A.30) up to kCd|logh|\|k\|\leq C_{d}|\log h|. Let f(x)=sinh(xt)cosh(x)f(x)=\frac{\sinh(xt)}{\cosh(x)}. Then 0f(x)1,|f(x)|20\leq f(x)\leq 1,|f^{\prime}(x)|\leq 2 for any x0x\geq 0. Therefore, f(βkR)=f(αk)+O(logh3h2)f(\beta_{k}R)=f(\alpha_{k})+O(\|\log h\|^{3}h^{2}). Notice that 1/(Rsinh(xh))=1x(1+O(x2h2)1/({R\sinh(xh)})=\frac{1}{x}(1+O(x^{2}h^{2}) implies

1Rsinh(βk)=1βkR(1+O(βk2h2))=1αk(1+O(|logh|3h2)).\displaystyle\frac{1}{{R\sinh(\beta_{k})}}=\frac{1}{\beta_{k}R}\bigl{(}1+O(\beta_{k}^{2}h^{2})\bigr{)}=\frac{1}{\alpha_{k}}\bigl{(}1+O(|\log h|^{3}h^{2})\bigr{)}.

Putting all these together, we have

sinh(βk(Rnd))2Rsinh(βk)cosh(βkR)=sinh(αk(1xd))2αkcosh(αk)+O(|logh|3h2).\displaystyle\frac{\sinh(\beta_{k}(R-n_{d}))}{2R\sinh(\beta_{k})\cosh(\beta_{k}R)}=\frac{\sinh(\alpha_{k}(1-x_{d}))}{2\alpha_{k}\cosh(\alpha_{k})}+O(|\log h|^{3}h^{2}).

Then we split the series expression of 𝒢(0,nh)𝒢h(0,nh){\mathcal{G}}(0,nh)-{\mathcal{G}}^{h}(0,nh) into low frequency and high frequency part for a=8|logh|=8logRa=8|\log h|=8\log R,

|𝒢(0,nh)𝒢h(0,nh)|\displaystyle|{\mathcal{G}}(0,nh)-{\mathcal{G}}^{h}(0,nh)|\leq ka()+k>a()\displaystyle\,\sum_{\|k\|\leq a}\left(\cdot\right)+\sum_{\|k\|>a}\left(\cdot\right)
\displaystyle\leq kaO(|logh|3h2)+ak<+11e/4+(2R1)d1eR\displaystyle\,\sum_{\|k\|\leq a}O(|\log h|^{3}h^{2})+\sum_{\ell\geq a}\sum_{\ell\leq\|k\|<\ell+1}\frac{1}{\ell}e^{-\ell/4}\,+(2R-1)^{d-1}e^{-R}
\displaystyle\leq adO(|logh|3h2)+ad2ea/4+(2R1)d1eRCd|logh|d+3h2,\displaystyle\,a^{d}O(|\log h|^{3}h^{2})+a^{d-2}e^{-a/4}\,+(2R-1)^{d-1}e^{-R}\leq C_{d}|\log h|^{d+3}h^{2},

for sufficiently large RRdR\geq R_{d}. This completes the proof of (A.21). ∎

For any 0<ε<1/40<\varepsilon<1/4, by the positivity and smoothness (away from the pole) of the the continuous Green’s function 𝒢{\mathcal{G}}, there are c1(ε,d)>0,c2(ε,d)>0c_{1}(\varepsilon,d)>0,c_{2}(\varepsilon,d)>0 such that 2c1𝒢(ξ,x)c2/22c_{1}\leq{\mathcal{G}}(\xi,x)\leq c_{2}/2 for all 𝒬1ε/4\𝒬1/2{\mathcal{Q}}_{1-\varepsilon/4}\backslash{\mathcal{Q}}_{1/2}. Combining this with the approximation in (A.21), we have c1𝒢h(ξ,τn)c2c_{1}\leq{\mathcal{G}}^{h}(\xi,\tau_{n})\leq c_{2} for h<h~0(ε,d)h<\widetilde{h}_{0}(\varepsilon,d) and τn=ξ+nhΩh(𝒬1ε/4\𝒬1/2)\tau_{n}=\xi+nh\in\Omega_{h}\cap\left({\mathcal{Q}}_{1-\varepsilon/4}\backslash{\mathcal{Q}}_{1/2}\right). Then by (A.19), one has c1Rd2GR(ξ,n)c2c_{1}\leq R^{d-2}G_{R}(\xi,n)\leq c_{2} for R=1/h1/h~0R=1/{h}\geq 1/\widetilde{h}_{0}. Notice that τn=ξ+nhΩh(𝒬1ε/4\𝒬1/2)\tau_{n}=\xi+nh\in\Omega_{h}\cap\left({\mathcal{Q}}_{1-\varepsilon/4}\backslash{\mathcal{Q}}_{1/2}\right) is equivalent to

(A.32) nQ(R;ξ),andR/2|nξ|<(1ε/4)R.\displaystyle n\in Q(R;\xi),\ \ {\rm and}\ R/2\leq|n-\xi|_{\infty}<(1-\varepsilon/4)R.

For any 0<ε<1/40<\varepsilon<1/4 and r15/εr\geq 15/\varepsilon, if we set R=(1+ε)1/2rR=\left\lfloor(1+\varepsilon)^{1/2}r\right\rfloor, then it is easy to verify that (1+ε/3)rR(1+ε/2)r(1+\varepsilon/3)r\leq R\leq(1+\varepsilon/2)r, which implies that R/2r(1+ε/3)1R<(1ε/4)R.R/2\leq r\leq(1+\varepsilon/3)^{-1}R<(1-\varepsilon/4)R. In other words, if nQ(r;ξ)n\in\partial Q(r;\xi), then nn will satisfy (A.32). In conclusion, we have obtained

Lemma A.7.

For any 0<ε<1/40<\varepsilon<1/4, and rr\in\mathbb{N}, let R=(1+ε)1/2rR=\left\lfloor(1+\varepsilon)^{1/2}r\right\rfloor. There are constants c1,c2,r0c_{1},c_{2},r_{0} depending on dd and ε\varepsilon such that if rr0r\geq r_{0}, then c1r2dGR(ξ,m)c2r2dc_{1}r^{2-d}\leq G_{R}(\xi,m)\leq c_{2}r^{2-d} for all mQ(r;ξ)m\in Q(r;\xi)

We are also interested in the behavior of 𝒢{\mathcal{G}} and GrG_{r} near the boundary.

Lemma A.8.

Let xc=ξ+(0,,0,1)𝒬1x^{\,c}=\xi+(0,\cdots,0,1)\in\partial{\mathcal{Q}}_{1} be the center of the top surface of 𝒬1{\mathcal{Q}}_{1}. For any 0<η<1/40<\eta<1/4, let T1ηT_{1-\eta} be the semi cube contained in 𝒬1{\mathcal{Q}}_{1}, cetered at xcx^{c}, and away from the other surfaces of 𝒬1{\mathcal{Q}}_{1} by distance η\eta:

T1η={x𝒬1:max1id|xixic|1η}.T_{1-\eta}=\{x\in{\mathcal{Q}}_{1}:\max_{1\leq i\leq d}|x_{i}-x_{i}^{\,c}|\leq 1-\eta\}.

There is a constant c(η,d)c(\eta,d) such that for all x=(x1,,xd)T1ηx=(x_{1},\cdots,x_{d})\in T_{1-\eta} , one has

(A.33) 𝒢(ξ,x)c(η,d)|ξd+1xd|=c(η,d)dist(x,𝒬1).\displaystyle{\mathcal{G}}(\xi,x)\geq c(\eta,d)|\xi_{d}+1-x_{d}|=c(\eta,d)\,{\rm dist}(x,\partial{\mathcal{Q}}_{1}).
Proof.

Consider a larger semi-cube T1η/2T_{1-\eta/2} such that T1ηT1η/2T1.T_{1-\eta}\subsetneq T_{1-\eta/2}\subsetneq T_{1}. Let h(x)=ξd+1xd.h(x)=\xi_{d}+1-x_{d}. Clearly, 𝒢(ξ,x){\mathcal{G}}(\xi,x) and h(x)h(x) are two strictly positive harmonic functions on the interior of T1η/2T_{1-\eta/2}. By the comparison principle for harmonic functions, see, e.g. [Da, Ke], there is a constant cc only depends on dd and η\eta such that 𝒢(ξ,x)/h(x)c𝒢(ξ,y)/h(y){{\mathcal{G}}(\xi,x)}/{h(x)}\geq c{{\mathcal{G}}(\xi,y)}/{h(y)} for all x,yT1ηx,y\in T_{1-\eta}. Take y=(0,,0,ξd+2η)T1ηy=(0,\cdots,0,\xi_{d}+2\eta)\in T_{1-\eta} so that h(y)=12ηh(y)=1-2\eta. Then 𝒢(ξ,y)C(η,d){\mathcal{G}}(\xi,y)\geq C(\eta,d), and therefore, 𝒢(ξ,x)ch(x)=c(ξd+1xd),{{\mathcal{G}}(\xi,x)}\geq ch(x)=c\left(\xi_{d}+1-x_{d}\right), where cc only depends on dd and η\eta. ∎

Combing Lemma A.6 and Lemma A.8, we have

Lemma A.9.

Let Q(r)=Q(r;ξ)Q(r)=Q(r;\xi) be a cube in d{\mathbb{Z}}^{d} centered at ξd\xi\in{\mathbb{Z}}^{d}, with side length rr. Let Pr(ξ,m)P_{r}(\xi,m) be the discrete Poisson kernel on Q(r)Q(r) given by (A.12) with pole at the center ξ\xi. Let SQ(r)S\subset\partial Q(r) be the “top surface” of Q(r)Q(r):

S:={m=(m1,,md)Q(r):md=ξd+r}.S:=\big{\{}m=(m_{1},\cdots,m_{d})\in\partial Q(r):\ m_{d}=\xi_{d}+r\big{\}}.

Let nc=(n1c,n2c,,ξd+r)Sn^{c}=(n^{c}_{1},n_{2}^{c},\cdots,\xi_{d}+r)\in S be the center of SS. Suppose 0<η<1/40<\eta<1/4. There are constants c=c(η,d)c=c(\eta,d) and r0(η,d)r_{0}(\eta,d) depending only on the dimension and η\eta such that

(A.34) Pr(ξ,m)c1rd1.\displaystyle P_{r}(\xi,m)\geq c\frac{1}{r^{d-1}}.

for all rr0r\geq r_{0} and m=(m1,,md)Sm=(m_{1},\cdots,m_{d})\in S satisfying |mnc|(1η)r.|m-n^{c}|_{\infty}\leq(1-\eta)r.

The same estimate holds for Pr(ξ,m)P_{r}(\xi,m) on the other 2d12d-1 surfaces Si={mQ(r):mi=ξi±r},i=1,d1S^{i}=\big{\{}m\in\partial Q(r):\ m_{i}=\xi_{i}\pm r\big{\}},i=1\cdots,d-1 when mm is η\eta-away from the edges and the corners.

Proof.

It is enough to prove the result for Q(r)=Q(r;0)Q(r)=Q(r;0) centered at ξ=0\xi=0. We consider the approximation (A.21) on 𝒬1=[1,1]d{\mathcal{Q}}_{1}=[-1,1]^{d} with the mesh size h=1rh0h=\frac{1}{r}\leq h_{0}. Retain the definitions of 𝒢,𝒢h,Ωh{\mathcal{G}},{\mathcal{G}}^{h},\Omega_{h} and τn\tau_{n} in Lemma A.6. By the definition of the mesh Ωh\Omega_{h}, for n=(n1,,nd1,r1)Q(r1)n=(n_{1},\cdots,n_{d-1},r-1)\in\partial Q(r-1), the mesh points τn=n1rΩh(𝒬1\𝒬1/2)\tau_{n}=n\frac{1}{r}\in\Omega_{h}\cap\left({\mathcal{Q}}_{1}\backslash{\mathcal{Q}}_{1/2}\right). Then Lemma A.6 implies that

|𝒢(0,τn)𝒢1/r(0,τn)|C(logr)d+3r2,|{\mathcal{G}}(0,\tau_{n})-{\mathcal{G}}^{1/r}(0,\tau_{n})|\leq C\frac{(\log r)^{d+3}}{r^{2}},

for n=(n1,,nd1,r1)Q(r1)n=(n_{1},\cdots,n_{d-1},r-1)\in\partial Q(r-1).

For these n=(n1,,nd1,r1)Q(r1)n=(n_{1},\cdots,n_{d-1},r-1)\in\partial Q(r-1), assume further that |ni|<(1η)r,1id1|n_{i}|<(1-\eta)r,1\leq i\leq d-1. Then for τn=n/r\tau_{n}=n/r and xc=(0,,0,1)x^{c}=(0,\cdots,0,1), one has

τnxc=(n1r,,nd1r,r1r).\tau_{n}-x^{c}=\big{(}\frac{n_{1}}{r},\cdots,\frac{n_{d-1}}{r},\frac{r-1}{r}\big{)}.

Then τnT1η\tau_{n}\in T_{1-\eta} provided that r1/ηr\leq 1/\eta, and hence, (A.33) implies that 𝒢(0,τn)cdist(τn,𝒬1)=c1r.{\mathcal{G}}(0,\tau_{n})\geq c{\rm dist}(\tau_{n},\partial{\mathcal{Q}}_{1})=c\frac{1}{r}. Therefore, for rr(d,η)r\geq r_{\ast}(d,\eta)

𝒢1/r(0,τn)𝒢(0,τn)C(logr)d+3r2c1rC(logr)d+3r2c~1r.{\mathcal{G}}^{1/r}(0,\tau_{n})\geq{\mathcal{G}}(0,\tau_{n})-C\frac{(\log r)^{d+3}}{r^{2}}\geq c\frac{1}{r}-C\frac{(\log r)^{d+3}}{r^{2}}\geq\widetilde{c}\,\frac{1}{r}.

By (A.19) (for h=1/rh=1/r), we have

(A.35) Gr(0,n)=hd2𝒢1/r(0,τn)c~1rd1,G_{r}(0,n)=h^{d-2}{\mathcal{G}}^{1/r}(0,\tau_{n})\geq\widetilde{c}\frac{1}{r^{d-1}},

for n=(n1,,nd1,r1)Q(r1)n=(n_{1},\cdots,n_{d-1},r-1)\in\partial Q(r-1) and |ni|<(1η)r,i=1,,d1|n_{i}|<(1-\eta)r,i=1,\cdots,d-1.

Notice that if mSQ(r)m\in S\subset\partial Q(r) and nQ(r1)n\in\partial Q(r-1) with |nm|1=1|n-m|_{1}=1, then |ni|<(1η)r,i=1,,d1|n_{i}|<(1-\eta)r,i=1,\cdots,d-1, and nd=r1n_{d}=r-1. Combing (A.35) with (A.14), one has

Pr(0,m)=Gr(0,n)c~1rd1.P_{r}(0,m)=G_{r}(0,n)\geq\widetilde{c}\frac{1}{r^{d-1}}.

This completes the proof of the lemma. ∎

A.6. Harnack type inequalities

We prove the discrete sub-mean value property and the Moser-Harnack inequality first.

Lemma A.10.

Suppose f={fn}f=\{f_{n}\} is a discrete nonnegative subharmonic function on Q(r;ξ)Q(r;\xi) in the sense that

(Δf)n=1id(fn+ei+fnei)+2dfn0,nQ(r1;ξ).-(\Delta f)_{n}=-\sum_{1\leq i\leq d}(f_{n+e_{i}}+f_{n-e_{i}})+2df_{n}\leq 0,\ n\in Q(r-1;\xi).

There is a dimensional constant C=C(d)>0C=C(d)>0 such that

(A.36) fξCr1dnQ(r;ξ)fnandfξCrdnQ(r;ξ)fn.f_{\xi}\leq\,C{r^{1-d}}\sum_{n\in\partial Q(r;\xi)}f_{n}\quad\mbox{and}\quad f_{\xi}\leq\,Cr^{-d}\sum_{n\in Q(r;\xi)}f_{n}.

As a consequence, for any cube Ωd\Omega\subset{\mathbb{Z}}^{d}, if fnf_{n} is non-negative and subharmonic on a domain containing the tripled cube 3Ω3\Omega, then

(A.37) |(Ω)|dsupξΩfξ2Cn3Ωfn2.\displaystyle|\ell(\Omega)|^{d}\sup_{\xi\in\Omega}f^{2}_{\xi}\leq C\sum_{n\in 3\Omega}f^{2}_{n}.
Proof.

Fix ξd\xi\in{\mathbb{Z}}^{d}, let f~\widetilde{f} be the discrete harmonic function on Q(r;ξ)Q(r;\xi) which coincides ff on Q(r;ξ)\partial Q(r;\xi), i.e., (Δf~)=0,nQ(r1;ξ)(\Delta\widetilde{f})=0,n\in Q(r-1;\xi), and f~n=fn,nQ(r;ξ).\widetilde{f}_{n}=f_{n},n\in\partial Q(r;\xi). By the integration by parts formula (A.15), f~ξ=mQ(r)Pr(ξ,m)fm,\widetilde{f}_{\xi}=\sum_{m\in\partial Q(r)}P_{r}(\xi,m){f}_{m}, where Pr(n,m)P_{r}(n,m) is the discrete Poisson kernel on Q(r)Q(r) from (A.12). It was showed in [Gu, GuMa] that there is a dimensional constant C>0C>0 such that Pr(ξ,m)Cr1dP_{r}(\xi,m)\leq Cr^{1-d} for all mQ(r;ξ)m\in\partial Q(r;\xi).

Let wn=f~nfnw_{n}=\widetilde{f}_{n}-f_{n}. Clearly, (Δw)n=(Δf)n0-(\Delta w)_{n}=(\Delta f)_{n}\geq 0 for all nQ(r1;ξ)n\in Q(r-1;\xi) and wn=0w_{n}=0 for nQ(r;ξ)n\in\partial Q(r;\xi). By the maximum principle Lemma A.1, one has wn=f~nfn0w_{n}=\widetilde{f}_{n}-f_{n}\geq 0 for nQ(r1;ξ)n\in Q(r-1;\xi). In particular, for all 0rr0\leq r^{\prime}\leq r,

(A.38) fξf~ξ=mQ(r;ξ)Pr(ξ,m)fmCr1dmQ(r;ξ)fmf_{\xi}\leq\widetilde{f}_{\xi}=\sum_{m\in\partial Q(r^{\prime};\xi)}P_{r^{\prime}}(\xi,m){f}_{m}\leq Cr^{\prime 1-d}\sum_{m\in\partial Q(r^{\prime};\xi)}f_{m}

since fn0f_{n}\geq 0. We multiply (A.38) by (r)d1(r^{\prime})^{d-1}, and then sum for all 1rr1\leq r^{\prime}\leq r to obtain

fξr=1r(r)d1Cr=1rnQ(r;ξ)fnCnQ(r;ξ)fn.f_{\xi}\sum_{r^{\prime}=1}^{r}(r^{\prime})^{d-1}\leq C\sum_{r^{\prime}=1}^{r}\sum_{n\in\partial Q(r^{\prime};\xi)}f_{n}\leq C\sum_{n\in Q(r;\xi)}f_{n}.

Therefore,

(A.39) fξC~rdnQ(r;ξ)fn,f_{\xi}\leq\widetilde{C}r^{-d}\sum_{n\in Q(r;\xi)}f_{n},

which proves (A.36).

Furthermore, (A.39) and Hölder inequality imply that for any ξ\xi and rr,

(A.40) rdfξ2C~nQ(r;ξ)fn2,r^{d}f^{2}_{\xi}\leq\widetilde{C}\sum_{n\in Q(r;\xi)}f^{2}_{n},

which easily yields (A.37). ∎

As a direct consequence, we have

Lemma A.11 (Moser-Harnack inequality for sub-solutions).

Let Ωd\Omega\subset{\mathbb{Z}}^{d} be a cube of side length (Ω)\ell(\Omega), and let 3Ω3\Omega be the tripled cube (2.10). Suppose gng_{n} is a non-negative sub-solution to an inhomogeneous equation on a domain containing 3Ω3\Omega, so that (Δg)n1-(\Delta g)_{n}\leq 1 and gn0g_{n}\geq 0 for n3Ω.n\in 3\Omega. Then there is a dimensional constant cH>0c_{H}>0 such that

(A.41) n3Ωgn2|(Ω)|d(cHsupΩgn2|(Ω)|4).\sum_{n\in 3\Omega}g_{n}^{2}\geq|\ell(\Omega)|^{d}\,\Bigl{(}c_{H}\sup_{\Omega}g^{2}_{n}-|\ell(\Omega)|^{4}\Bigr{)}.
Proof.

Suppose 3Ω=a1,b1××ad,bd3\Omega=\llbracket a_{1},b_{1}\rrbracket\times\cdots\times\llbracket a_{d},b_{d}\rrbracket, biai+1=3,i=1,,db_{i}-a_{i}+1=3\ell,i=1,\cdots,d, where =(Ω)\ell=\ell(\Omega) is the side length of Ω\Omega. Denote by |3Ω|=3dd|3\Omega|=3^{d}\ell^{d} its cardinality as usual. For n=(n1,,nd)3Ωn=(n_{1},\cdots,n_{d})\in 3\Omega, let

hn=98212di=1d(niai)(bini).h_{n}=\frac{9}{8}\ell^{2}-\frac{1}{2d}\sum_{i=1}^{d}(n_{i}-a_{i})(b_{i}-n_{i}).

Direct computations show that

(Δh)n=1and 0hn982,n3Ω.(\Delta h)_{n}=1\quad{\mbox{and }}\quad 0\leq h_{n}\leq\frac{9}{8}\ell^{2},\quad n\in 3\Omega.

Let fn=hn+gnf_{n}=h_{n}+g_{n}. Then (Δf)n=(Δh)n(Δg)n0-(\Delta f)_{n}=-(\Delta h)_{n}-(\Delta g)_{n}\leq 0 and fngn0f_{n}\geq g_{n}\geq 0. We can apply Lemma A.10 to the non-negative subharmonic function fnf_{n}. The estimate (A.37) implies that

|(Ω)|dsupnΩgn2|(Ω)|dsupnΩfn2Cn3Ωfn2C(|(Ω)|4+d+n3Ωgn2),|\ell(\Omega)|^{d}\sup_{n\in\Omega}g^{2}_{n}\leq|\ell(\Omega)|^{d}\sup_{n\in\Omega}f^{2}_{n}\leq C\sum_{n\in 3\Omega}f^{2}_{n}\leq C\Bigl{(}|\ell(\Omega)|^{4+d}+\sum_{n\in 3\Omega}g_{n}^{2}\Bigr{)},

as desired.

Next, we study the discrete Harnack inequality for sup-solutions of a homogeneous Schrödinger equation with a bounded potential.

Lemma A.12.

Suppose v={vn}v=\{v_{n}\}, vnVmaxv_{n}\leq V_{\max}, is a bounded potential. Let ff be a non-negative super-solution for the Schrödinger equation on a cube Ωd\Omega\subset{\mathbb{Z}}^{d} so that

(A.42) (Δf)n+vnfn0,fn0,nΩ.-(\Delta f)_{n}+v_{n}f_{n}\geq 0,\ \ f_{n}\geq 0,\ \ n\in\Omega.

There is a constant C>0C>0 depending on dd and VmaxV_{\max} such that for any cube QΩQ\subset\Omega of side length (Q)\ell(Q),

(A.43) supnQfnC(Q)infnQfn.\sup_{n\in Q}f_{n}\leq C^{\ell(Q)}\inf_{n\in Q}f_{n}.
Proof.

Assume that the finite dimensional vector {fn}nQ\{f_{n}\}_{n\in Q} attains its minimum and maximum at m,nQm,n\in Q respectively. Connect m,nm,n by a discrete path {γj}j=0s\{\gamma^{j}\}_{j=0}^{s} in QQ, where γ0=m,γs=n\gamma_{0}=m,\gamma_{s}=n and γj+1=γj±ei\gamma^{j+1}=\gamma^{j}\pm e_{i} for some ii. It is easy to check that the minimum number of steps needed to reach nn from mm is s=|mn|1d(Q)s=|m-n|_{1}\leq d\,\ell(Q). The upper bound for vnv_{n} and (A.42) imply (2d+Vmax)fk|kk|1=1fk,(2d+V_{\max})f_{k}\geq\sum_{|k^{\prime}-k|_{1}=1}f_{k^{\prime}}, for all kΩk\in\Omega. Then

fm12d+Vmax|mm|1=1fm12d+Vmaxfγ11(2d+Vmax)2fγ21(2d+Vmax)sfγs.f_{m}\geq\frac{1}{2d+V_{\max}}\sum_{|m^{\prime}-m|_{1}=1}f_{m^{\prime}}\geq\frac{1}{2d+V_{\max}}f_{\gamma^{1}}\geq\frac{1}{(2d+V_{\max})^{2}}f_{\gamma^{2}}\geq\cdots\frac{1}{(2d+V_{\max})^{s}}f_{\gamma^{s}}.

Therefore,

infnQfn(2d+Vmax)d(Q)supnQfn,\inf_{n\in Q}f_{n}\geq(2d+V_{\max})^{-d\,\ell(Q)}\sup_{n\in Q}f_{n},

which gives (A.43). ∎

Appendix B Chernoff bound

Lemma B.1 (Chernoff–Hoeffding Theorem, [Ho]).

Suppose BdB\subset{\mathbb{Z}}^{d} and {ζj}jB\{\zeta_{j}\}_{j\in B} are i.i.d. Bernoulli random variables, taking values in {0,1}\{0,1\} with common expectation p=𝔼(ζj)(0,1)p=\mathbb{E}\left(\zeta_{j}\right)\in(0,1). Then for any 0<λ<1p0<\lambda<1-p,

(B.1) {jBζj(1λ)|B|}eD(1λp)|B|,\mathbb{P}\Bigl{\{}\sum_{j\in B}\zeta_{j}\geq(1-\lambda)|B|\Bigr{\}}\leq e^{-D(1-\lambda\|p)\,|B|},

where

(B.2) D(xy)=xlogxy+(1x)log1x1yD(x\|y)=x\log\frac{x}{y}+(1-x)\log\frac{1-x}{1-y}

is the Kullback–Leibler divergence between Bernoulli distributed random variables with parameters xx and yy respectively.

We sketch the proof, following the arguments used in [DFM] (which is also close to the original proof of Hoeffding), for readers’ convenience.

Proof.

Let S=jBζjS=\sum_{j\in B}\zeta_{j}. For any t>0t>0,

{S(1λ)|B|}={etSet(1λ)|B|}et(1λ)|B|𝔼(etS)=et(1λ)|B|(𝔼(etζ1))|B|.{\mathbb{P}}\left\{\,S\geq(1-\lambda)|B|\,\right\}={\mathbb{P}}\left\{\,e^{tS}\geq e^{t(1-\lambda)|B|}\,\right\}\leq e^{-t(1-\lambda)|B|}\,\mathbb{E}\left(e^{tS}\right)=e^{-t(1-\lambda)|B|}\,\left(\mathbb{E}\left(e^{t\zeta_{1}}\right)\right)^{|B|}.

For any jj, the expectation of etζje^{t\zeta_{j}} is 𝔼(etζ1)=etp+1p.\mathbb{E}\left(e^{t\zeta_{1}}\right)=e^{t}p+1-p. Therefore, for all t>0t>0,

(B.3) log{S(1λ)|B|}t(1λ)|B|+|B|log(etp+1p)=:|B|f(t).\log{\mathbb{P}}\left\{\,S\geq(1-\lambda)|B|\,\right\}\leq-t(1-\lambda)|B|+|B|\log\left(e^{t}p+1-p\right)=:-|B|\,f(t).

It is enough to optimize f(t)f(t) in tt. Under the condition of 1λp>01-\lambda-p>0, the function f(t)f(t) attains its only local maximum at t=log(1λλ1pp).{t^{\ast}}=\log\left(\frac{1-\lambda}{\lambda}\frac{1-p}{p}\right). Direct computations show that

f(t)=t(1λ)log(etp+1p)=(1λ)log1λp+λlogλ1p=D(1λp)f(t^{\ast})=t^{\ast}(1-\lambda)-\log\left(e^{t^{\ast}}p+1-p\right)=(1-\lambda)\log\frac{1-\lambda}{p}\,+\lambda\log\frac{\lambda}{1-p}=D(1-\lambda\|p)

where D(xy)D(x\|y) is defined in (B.2). Combing with (B.3), we have (B.1). ∎

References

  • [AW] Aizenman, M., and Warzel, S., Random Operators: Disorder Effects on Quantum Spectra and Dynamics. Vol. 168. American Mathematical Soc., 2015.
  • [ADFJM1] Arnold, D. N., David, G., Jerison, D., Mayboroda, S., and Filoche, M., Effective confining potential of quantum states in disordered media. Phys. Rev. Lett. 116.5 (2016): 056602.
  • [ADFJM2] Arnold, D. N., David, G., Filoche, M., Jerison, D., and Mayboroda, S., Localization of eigenfunctions via an effective potential. Comm. Partial Differential Equations 44.11 (2019): 1186-1216.
  • [ADFJM3] Arnold, D. N., David, G., Filoche, M., Jerison, D., and Mayboroda, S., Computing spectra without solving eigenvalue problems. SIAM J. Sci. Comput. 41.1 (2019): B69-B92.
  • [Ba] Barlow, M.T., Random walks and heat kernels on graphs. volume 438 of London Mathematical Society Lecture Note Series. Cambridge University Press, Cambridge, 2017.
  • [BiKo] Biskup, M., and König, W., Long-time tails in the parabolic Anderson model. Ann. Probab. (2001): 636-682.
  • [Ch] Chung, F. R., Spectral graph theory. CBMS Lectures, Fresno 6.92 (1996): 17-21.
  • [Da] Dahlberg, B., On estimates for harmonic measure. Arch . Rat. Mech . Anal. 65.3 (1977): 275-288.
  • [DFM] David, G., Filoche, M., and Mayboroda, S., The landscape law for the integrated density of states. arXiv preprint arXiv:1909.10558 (2019).
  • [DLY] Damanik, D., Lukic, M., and Yessen, W., Quantum dynamics of periodic and limit-periodic Jacobi and block Jacobi matrices with application to some quantum many body problems. Comm. Math. Phys. 337(3), 1535–1561 (2015).
  • [DM+] Desforges, P., Mayboroda, S., Zhang, S., David, G., Arnold, D. N., Wang, W., and Filoche, M., Sharp estimates for the integrated density of states in Anderson tight-binding models. arXiv preprint arXiv:2010.09287 (2020).
  • [Ea] Eastham, M. S. P., The spectral theory of periodic differential equations. Scottish Academic Press [distributed by Chatto & Windus, London, 1973.
  • [Fe] Fefferman, C., The uncertainty principle. Bull. Amer. Math. Soc. 9.2 (1983): 129-206.
  • [FM] Filoche, M., and Mayboroda, S., Universal mechanism for Anderson and weak localization. Proc. Natl. Acad. Sci. USA 109.37 (2012): 14761-14766.
  • [Gu] Guadie, M., Harmonic functions on square lattices: uniqueness sets and growth properties. PhD thesis. Norwegian University of Science and Technology, Trondheim (2013).
  • [GuMa] Guadie, M., and Malinnikova, E., On Three Balls Theorem for Discrete Harmonic Functions. Comput. Methods Funct. Theory 14.4 (2014): 721-734.
  • [Ho] Hoeffding, W., Probability Inequalities for Sums of Bounded Random Variables. J. Amer. Statist. Assoc. 58.301 (1963): 13-30.
  • [HJ] Han, R., and Jitomirskaya, S., Discrete Bethe–Sommerfeld conjecture. Comm. Math. Phys. 361(1) (2018): 205-216 .
  • [HL] Han, Q., and Lin, F., Elliptic partial differential equations. Vol. 1. American Mathematical Soc., 2011.
  • [Ke] Kenig, C. E., Harmonic analysis techniques for second order elliptic boundary value problems. Vol. 83. American Mathematical Soc., 1994.
  • [Ki] Kirsch, W., An invitation to random Schrödinger operators. arXiv preprint arXiv:0709.3707 (2007).
  • [KM1] Kirsch, W., and Metzger, B., The integrated density of states for random Schrödinger operators in spectral theory and mathematical physics: a Festschrift in honor of Barry Simon’s 60th birthday. Proc. Sympos. Pure Math. Vol. 76. 2007.
  • [KM2] Kirsch, W., and Martinelli, F., Large deviations and Lifschitz singularity of the integrated density of states of random Hamiltonians, Comm. Math. Phys. 89.1 (1983): 27-40.
  • [Ko] König, W., The parabolic Anderson model. Random walk in random potential. P Birkhäuser, 2016.
  • [Ku] Kurata, K., On doubling properties for non-negative weak solutions of elliptic and parabolic PDE. Israel J. Math. 115.1 (2000): 285-302.
  • [La] Laasonen, P., On the solution of Poisson’s difference equation. J. ACM 5.4 (1958): 370-382.
  • [Li] Lifshitz, I. Y. M., Energy spectrum structure and quantum states of disordered condensed systems. Sov. Phy. Usp. 7.4 (1965): 549.
  • [LMF] Lyra, M. L., Mayboroda, S., and Filoche, M., Dual hidden landscapes for Anderson localization in discrete lattices. Europhys. Lett. EPL 109.4 (2015): 47001.
  • [RS] Reed, M., and Simon, B., Methods of modern mathematical physics. IV: Analysis of operators. New York - San Francisco - London: Academic Press (1978).
  • [SW] Schatz, A. H., and Wahlbin, L. B., Interior maximum norm estimates for finite element methods. Math. Comp. 31.138 (1977): 414-442.
  • [Si] Simon, B., Lifshitz Tails for the Anderson Model. J. Stat. Phys. 38.1-2 (1985): 65-76.
  • [WZ] Wang, W., and Zhang, S., The exponential decay of eigenfunctions for tight binding Hamiltonians via landscape and dual landscape functions. Ann. Henri Poincaré. 22.5 (2021): 1429-1457.

————————————–

D.  Arnold, School of Mathematics, University of Minnesota, 206 Church St SE, Minneapolis, MN 55455 USA

E-mail address: arnold@umn.edu


M. Filoche, Physique de la Matière Condensée, Ecole Polytechnique, CNRS, Institut Polytechnique de Paris, Palaiseau, France

E-mail address: marcel.filoche@polytechnique.edu

S. Mayboroda, School of Mathematics, University of Minnesota, 206 Church St SE, Minneapolis, MN 55455 USA

E-mail address: svitlana@math.umn.edu


W.  Wang, School of Mathematics, University of Minnesota, 206 Church St SE, Minneapolis, MN 55455 USA

E-mail address: wang9585@umn.edu


S. Zhang, School of Mathematics, University of Minnesota, 206 Church St SE, Minneapolis, MN 55455 USA

E-mail address: zhan7294@umn.edu