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

\nouppercaseheads\setsecheadstyle\setsubsecheadstyle\externaldocument

[monsterbook-]monsterbook[http://www.math.toronto.edu/ivrii/monsterbook.pdf] \externaldocument[OOD2-]OODiagonal-2[https://arxiv.org/pdf/????]

Pointwise Spectral Asymptotics out of the Diagonal near Boundarythanks: 2010 Mathematics Subject Classification: 35P20.thanks: Key words and phrases: Microlocal Analysis, sharp spectral asymptotics.

Victor Ivrii This research was supported in part by National Science and Engineering Research Council (Canada) Discovery Grant RGPIN 13827
Abstract

We establish uniform (with respect to xx, yy) semiclassical asymptotics and estimates for the Schwartz kernel eh(x,y,τ)e_{h}(x,y,\tau) of spectral projector for a second order elliptic operator on the manifold with a boundary. While such asymptotics for its restriction to the diagonal eh(x,x,τ)e_{h}(x,x,\tau) and, especially, for its trace 𝖭h(τ)=eh(x,x,τ)𝑑x\mathsf{N}_{h}(\tau)=\int e_{h}(x,x,\tau)\,dx are well-known, the out-of-diagonal asymptotics are much less explored, especially uniform ones.

Our main tools: microlocal methods, improved successive approximations and geometric optics methods.

Our results would also lead to classical asymptotics of eh(x,y,τ)e_{h}(x,y,\tau) for fixed hh (say, h=1h=1) and τ\tau\to\infty.

Chapter 1 Introduction

Asymptotics away from the boundary.

Let us start from the case of asymptotics when points xx and yy are disjoint from the boundary. While sharp classical asymptotics of e(x,x,τ)e(x,x,\tau) for elliptic operators are known from L. Hörmander [], and could be traced to B. M. Levitan [Lev1, Lev2] and V. G. Avakumovič [Av1], I could not find asymptotics of e(x,y,τ)e(x,y,\tau), despite it could be easily derived by the method of Fourier integral operators combined with Tauberian theorem.

Theorem 1.1.

Let A=A𝗐(x,hD,h)A=A^{\mathsf{w}}(x,hD,h) be a self-adjoint scalar pseudo-differential operator with 𝒞K\mathscr{C}^{K}-symbol (K=K(d)K=K(d)) which is ξ\xi-microhyperbolic in B(0,1)XB(0,1)\subset X on the energy level τ\tau1)1)1) Which means that (1.1) |a(x,ξ)τ|+|ξa(x,ξ)|ϵ0\displaystyle|a(x,\xi)-\tau|+|\nabla_{\xi}a(x,\xi)|\geq\epsilon_{0} where a(x,ξ)A(x,ξ,0)a(x,\xi)\coloneqq A(x,\xi,0).. Further, let {ξ:a(x,ξ)=τ}\{\xi\colon a(x,\xi)=\tau\} be strongly convex. Then

(1.2) eh(x,y,τ)=eh𝖶(x,y,τ)+O(h1d)x,yB(0,ϵ)\displaystyle e_{h}(x,y,\tau)=e_{h}^{\mathsf{W}}(x,y,\tau)+O(h^{1-d})\qquad\forall x,y\in B(0,\epsilon)
where
(1.3) eh𝖶(x,y,τ)=(2πh)d{a(12(x+y),ξ)<τ}eih1xy,ξ𝑑ξ\displaystyle e_{h}^{\mathsf{W}}(x,y,\tau)=(2\pi h)^{-d}\int_{\{a(\frac{1}{2}(x+y),\xi)<\tau\}}e^{ih^{-1}\langle x-y,\xi\rangle}\,d\xi

is corresponding Weyl expression.

We will prove this theorem and discuss its generalizations and also more precise results in Section 2.

Asymptotics near the boundary.

Classical asymptotics of e(x,x,τ)e(x,x,\tau) for second order elliptic operators are known from R. Seeley [Se1, Se2], and sharp ones from V. Ivrii. The semiclassical version would carry remainder estimate O(h1dν(x)12)O(h^{1-d}{\nu}(x)^{-\frac{1}{2}}) and O(h1d)O(h^{1-d}) respectively while the trivial remainder estimate would be O(h1dν(x)1)O(h^{1-d}{\nu}(x)^{-1}) where ν(x)=dist(x,X){\nu}(x)=\operatorname{dist}(x,\partial X).

Let

0(x,y)\displaystyle\ell^{0}(x,y) |xy|,\displaystyle\coloneqq|x-y|,
(x,y)\displaystyle\ell\ (x,y) |xy|+ν(x)+ν(y).\displaystyle\coloneqq|x-y|+{\nu}(x)+{\nu}(y).
Theorem 1.2.

Let

(1.4) A=j,k(hDjVj(x))gjk(x)(hDkVk(x))+V(x),gjk=gkj,\displaystyle A=\sum_{j,k}\bigl{(}hD_{j}-V_{j}(x)\bigr{)}g^{jk}(x)\bigl{(}hD_{k}-V_{k}(x)\bigr{)}+V(x),\qquad g^{jk}=g^{kj},

be a self-adjoint scalar operator which is elliptic second order differential operator with 𝒞K\mathscr{C}^{K}-coefficients (K=K(d,δ)K=K(d,\delta)), and it is ξ\xi-microhyperbolic in B(0,1)B(0,1) on the energy level τ\tau, with either Dirichlet or Neumann boundary condition on XB(0,1)\partial X\cap B(0,1), and X𝒞K\partial X\in\mathscr{C}^{K}2)2)2) More general boundary conditions could be also considered.. Then

  1. (i)

    For d3d\geq 3 asymptotics (1.2) holds where now for Dirichlet or Neumann boundary condition

    (1.5) eh𝖶(x,y,τ)=eh0,𝖶(x,y,τ)[]eh0,𝖶(x,y~,τ)\displaystyle e_{h}^{\mathsf{W}}(x,y,\tau)=e_{h}^{0,\mathsf{W}}(x,y,\tau)\ [\mp]\ e_{h}^{0,\mathsf{W}}(x,\tilde{y},\tau)

    respectively, eh0,𝖶(x,y,τ)e_{h}^{0,\mathsf{W}}(x,y,\tau) defined by (1.3) and y~\tilde{y} is a reflected point.

  2. (ii)

    For d=2d=2 under one of the following assumptions

    1. (a)

      (x,y)h13+δ\ell(x,y)\leq h^{\frac{1}{3}+\delta},

    2. (b)

      (x,y)h13δ\ell(x,y)\geq h^{\frac{1}{3}-\delta},

    3. (c)

      ν(x)+ν(y)C0(2+h232δ){\nu}(x)+{\nu}(y)\geq C_{0}(\ell^{2}+h^{\frac{2}{3}-2\delta})

    where here and below δ>0\delta>0 is an arbitrarily small exponent, asymptotics

    (1.6) eh(x,y,τ)=eh𝖶(x,y,τ)+eh,𝖼𝗈𝗋𝗋(x,y,τ)+O(h1d)x,yB(0,ϵ)\displaystyle e_{h}(x,y,\tau)=e_{h}^{\mathsf{W}}(x,y,\tau)+e_{h,\mathsf{corr}}(x,y,\tau)+O(h^{1-d})\quad\forall x,y\in B(0,\epsilon)

    holds with correction term eh,𝖼𝗈𝗋𝗋(x,y,τ)e_{h,\mathsf{corr}}(x,y,\tau) tol be defined by (5.53) later.

  3. (iii)

    Finally, for d=2d=2 if neither of condition (a)–(c) is fulfilled

    (1.7) eh(x,y,τ)=O(h1δ).\displaystyle e_{h}(x,y,\tau)=O(h^{-1-\delta}).

Remark 1.3.
  1. (i)

    According to [Ivr2], Theorem LABEL:monsterbook-thm-8-1-6 asymptotics (1.2) holds for x=yx=y in the coordinate system such that x1x_{1} is the distance from xx to X\partial X in the metrics gjk(τV)1g^{jk}(\tau-V)^{-1}. Such precision is needed for d=2d=2 only.

  2. (ii)

    The leading term eh𝖶(x,y,τ)e_{h}^{\mathsf{W}}(x,y,\tau) is hd12d+12\asymp h^{-\frac{d-1}{2}}\ell^{-\frac{d+1}{2}}. In particular it is O(h1d)O(h^{1-d}) as hd1d+1\ell\gtrsim h^{\frac{d-1}{d+1}}.

  3. (iii)

    The correction term (as d=2d=2) is O(h3212)O(h^{-\frac{3}{2}}\ell^{\frac{1}{2}}) if h12\ell\leq h^{\frac{1}{2}} and O(h1232)O(h^{-\frac{1}{2}}\ell^{-\frac{3}{2}}) if h12ϵh^{\frac{1}{2}}\leq\ell\leq\epsilon.

  4. (iv)

    The trivial estimate (in much more general settings) is

    (1.8) |eh(x,y,τ)eh𝖶(x,y,τ)|Ch1d(1+1(x,y)).\displaystyle|e_{h}(x,y,\tau)-e^{\mathsf{W}}_{h}(x,y,\tau)|\leq Ch^{1-d}(1+\ell^{-1}(x,y)).

Ideas of proofs. I.

Consider x,yB(0,12)x,y\in B(0,\frac{1}{2}). We already know that under pretty general assumptions (which will be discussed later), in particular, in the frameworks of Theorems 1.1 and 1.2, the following estimate holds:

(1.9) e(x,x,τ+h)e(x,x,τ)Ch1dfor |ττ|ϵ0.\displaystyle e(x,x,\tau^{\prime}+h)-e(x,x,\tau^{\prime})\leq Ch^{1-d}\qquad\text{for\ \ }|\tau^{\prime}-\tau|\leq\epsilon_{0}.
Then
(1.10) |e(x,y,τ)e(x,y,τ)|Chd|ττ|+h1d\displaystyle|e(x,y,\tau)-e(x,y,\tau^{\prime})|\leq Ch^{-d}|\tau-\tau^{\prime}|+h^{1-d}
and therefore due to Tauberian methods
(1.11) |eh(x,y,τ)eT,h𝖳(x,y,τ)|CT1h1d\displaystyle|e_{h}(x,y,\tau)-e_{T,h}^{\mathsf{T}}(x,y,\tau)|\leq CT^{-1}h^{1-d}
where
(1.12) eT,h𝖳(x,y,τ)=h1τFth1τ(χ¯T(t)uh(x,y,t))𝑑τ\displaystyle e_{T,h}^{\mathsf{T}}(x,y,\tau)=h^{-1}\int_{-\infty}^{\tau}F_{t\to h^{-1}\tau^{\prime}}\bigl{(}\bar{\chi}_{T}(t)u_{h}(x,y,t)\bigr{)}\,d\tau^{\prime}

is the Tauberian expression, uh(x,y,t)u_{h}(x,y,t) is the Schwartz kernel of the propagator eih1Ate^{ih^{-1}At}, hTϵ1{h\leq T\leq\epsilon_{1}}, ϵ1\epsilon_{1} is a small constant and χ¯T(t)=χ¯(t/T)\bar{\chi}_{T}(t)=\bar{\chi}(t/T) and χ¯𝒞0([1,1])\bar{\chi}\in\mathscr{C}_{0}^{\infty}([-1,1]), χ¯(t)=1\bar{\chi}(t)=1 on (12,12)(-\frac{1}{2},\frac{1}{2}).

Assume first that B(0,1)XB(0,1)\subset X. Then for a small constant TT we can construct uh(x,y,t)u_{h}(x,y,t) as an oscillatory integral and we can construct eh𝖳(x,y,τ)e_{h}^{\mathsf{T}}(x,y,\tau) even as d=2d=2 or without strong convexity assumption. However the result could be simplified to eh𝖶(x,y,τ)e^{\mathsf{W}}_{h}(x,y,\tau) in the framework of Theorem 1.1 (for scalar operators under strong convexity assumption). This is done in Section 2.

Ideas of proofs. II.

Construction of the propagator is much more complicated close to the boundary, because there could be many generalized billiard rays with at least one reflection from X\partial X, from xx to yy on the energy level τ\tau. If

(1.13) (x,y)|xy|+ν(x)+ν(y)ϵ1\displaystyle\ell(x,y)\coloneqq|x-y|+{\nu}(x)+{\nu}(y)\leq\epsilon_{1}
the length of all such rays is (x,y)\asymp\ell(x,y). Further, if
(1.14) ν(x)+ν(y)σ(x,y)\displaystyle{\nu}(x)+{\nu}(y)\geq\sigma\ell(x,y)

with σC0(x,y)\sigma\geq C_{0}\ell(x,y) there is exactly one such ray and it has exactly one reflection, and the reflection angle ϵ1(ν(x)+ν(y))/(x,y)\geq\epsilon_{1}({\nu}(x)+{\nu}(y))/\ell(x,y). It allows us in the framework of Theorem 1.2 under assumption (1.14) with σ=h12δ(x,y)12+C0(x,y)\sigma=h^{\frac{1}{2}-\delta}\ell(x,y)^{-\frac{1}{2}}+C_{0}\ell(x,y) to construct a reflected wave as an oscillatory integral and simplify it. It will be done in Section 5.

So, we should consider a case when (1.14) is violated and we will do it in Section 4. As |xy|Ch12δ|x-y|\leq Ch^{\frac{1}{2}-\delta} the propagator could be constructed by the standard method of successive approximations with unperturbed constant coefficients operator.

Furthermore, in the framework of Theorem 1.2 under certain assumptions, which are complementary to (1.14) as h13+δ\ell\leq h^{\frac{1}{3}+\delta}, the reflected wave could be constructed also by the method of the successive approximations, but with unperturbed the operator

(1.15) A¯(z,hDz)=h2D12+A(0,w,hDz),\displaystyle\bar{A}(z,hD_{z})=h^{2}D_{1}^{2}+A(0,w^{\prime},hD_{z}^{\prime}),

where x=(x1;x)=(x1;x2,,xd)x=(x_{1};x^{\prime})=(x_{1};x_{2},\ldots,x_{d}) and XB(0,1)={x:x1>0}B(0,1)X\cap B(0,1)=\{x\colon x_{1}>0\}\cap B(0,1) and we make a change of variables w=12(x+y)w^{\prime}=\frac{1}{2}(x^{\prime}+y^{\prime}) and z=12(xy)z^{\prime}=\frac{1}{2}(x^{\prime}-y^{\prime}), leaving x1x_{1} and y1y_{1} unchanged.

Ideas of proofs. III.

Finally, to estimate the Tauberian expression rather than to find its asymptotics we apply in Section 3 only methods of propagations. Let us illustrate this away from the boundary. Let Q1,2Q_{1,2} be pseudodifferential operators with the symbols equal 11 in ρ\rho-vicinity of {ξΣ(w,τ):ξa(w,ξ)x¯y¯}\{\xi\in\Sigma(w,\tau)\colon\nabla_{\xi}a(w,\xi)\parallel\bar{x}-\bar{y}\}; this set consists of just two points. Then

(1.16) eh𝖳(x¯,y¯,τ)=Q2xeh𝖳(x,y,τ)tQ1y|x=x¯,y=y¯+O(hs)\displaystyle e_{h}^{\mathsf{T}}(\bar{x},\bar{y},\tau)=Q_{2x}e_{h}^{\mathsf{T}}(x,y,\tau)\,^{t}\!Q_{1y}|_{x=\bar{x},y=\bar{y}}+O(h^{s})

provided microlocal uncertainty principle ρ×ρh1δ\rho\times\rho\ell\geq h^{1-\delta} is fulfilled. It allows us to upgrade |eh𝖳(x,y,τ)|Ch1d1(x,y)|e^{\mathsf{T}}_{h}(x,y,\tau)|\leq Ch^{1-d}\ell^{-1}(x,y) to

(1.17) |eh𝖳(x,y,τ)|Cρd1h1d1(x,y).\displaystyle|e_{h}^{\mathsf{T}}(x,y,\tau)|\leq C\rho^{d-1}h^{1-d}\ell^{-1}(x,y).

We need also ρC0\rho\geq C_{0}\ell.

Chapter 2 Asymptotics inside domain

1 Tauberian asymptotics

As we already mentioned, asymptotics

(2.1) eh(x,x,τ)=κ0(x,τ)hd+O(h1d)\displaystyle e_{h}(x,x,\tau)=\kappa_{0}(x,\tau)h^{-d}+O(h^{1-d})

is known under much more general assumptions. It holds for matrix operators under ξ\xi-microhyperbolicity condition at point xx on the energy level τ\tau and for Schrödinger operator without any condition in dimension d3d\geq 3 (see [Ivr2], Sections LABEL:monsterbook-sect-5-4 and LABEL:monsterbook-sect-5-3 correspondingly).

Asymptotics (2.1) implies (1.9) and then (1.10). Then the standard Tauberian arguments (see f.e. [Ivr2], Section LABEL:monsterbook-sect-5-2) imply (1.11)–(1.12) with small constant TT.

2 Weyl asymptotics

While asymptotics (2.1) holds in much more general assumptions, to construct uh(x,y,t)u_{h}(x,y,t) we need to impose some restrictions. Let AA be a scalar operator3)3)3) Construction also works for matrix operators with characteristic roots of constant multiplicity.. While construction of propagator as an oscillatory integral is possible even if ξ\xi-microhyperbolicity condition is violated, we assume that it is fulfilled from the beginning, as we need it anyway.

Without any loss of the generality one can assume that

(2.2) a(y,0)<0.\displaystyle a(y,0)<0.
Indeed, otherwise we can achieve it by a corresponding gauge transformation. Then the strong convexity of Σ(y,τ)={θ:a(y,θ)=τ}\Sigma(y,\tau)=\{\theta\colon a(y,\theta)=\tau\} implies that
(2.3) θa(y,θ),θϵ0θΣ(y,τ).\displaystyle\langle\nabla_{\theta}a(y,\theta),\theta\rangle\geq\epsilon_{0}\qquad\forall\theta\in\Sigma(y,\tau).
Proposition 2.1.

Let AA be a scalar operator, ξ\xi-microhyperbolic in B(x¯,2ϵ)XB(\bar{x},2\epsilon)\subset X on energy level τ\tau and let x,yB(x¯,ϵ)x,y\in B(\bar{x},\epsilon). Let (2.3) be fulfilled. Then

(2.4) uh(x,y,t)eih1Φ(x,y,t,θ)B(x,y,t,θ,h)𝑑θ\displaystyle u_{h}(x,y,t)\equiv\int e^{ih^{-1}\Phi(x,y,t,\theta)}B(x,y,t,\theta,h)\,d\theta
modulo functions such that
(2.5) Fth1τχ¯T(t)vh(x,y,t)=O(hs)x,yB(0,ϵ)\displaystyle F_{t\to h^{-1}\tau^{\prime}}\bar{\chi}_{T}(t)v_{h}(x,y,t)=O(h^{s})\qquad\forall x,y\in B(0,\epsilon)
as Tϵ1T\leq\epsilon_{1}, |ττ|ϵ1|\tau^{\prime}-\tau|\leq\epsilon_{1} where
(2.6) Φ(x,y,t,θ)=φ(x,y,θ)+ta(y,θ),\displaystyle\Phi(x,y,t,\theta)=\varphi(x,y,\theta)+ta(y,\theta),
φ(x,y,θ)\varphi(x,y,\theta) satisfies stationary eikonal equation
(2.7) a(x,xφ)=a(y,θ)\displaystyle a(x,\nabla_{x}\varphi)=a(y,\theta)
and
(2.8) φ(x,y,θ)|xy,θ=0=0,\displaystyle\varphi(x,y,\theta)|_{\langle x-y,\theta\rangle=0}=0,
(2.9) xφ(x,y,θ)|x=y=θ\displaystyle\nabla_{x}\varphi(x,y,\theta)|_{x=y}=\theta
and
(2.10) B(x,y,t,θ,h)n0Bn(x,y,t,θ)hn,\displaystyle B(x,y,t,\theta,h)\sim\sum_{n\geq 0}B_{n}(x,y,t,\theta)h^{n},

and amplitudes BjnB_{jn} are defined from the Cauchy problems for transport equations; all functions are uniformly smooth; here and below ss is an arbitrarily large exponent.

Proof.

The proof is original L. Hörmander’s construction (see M. Shubin [Shu], Theorem 20.1).

While in [, Shu] this decomposition was used to calculate eh𝖳(x,y,τ)e^{\mathsf{T}}_{h}(x,y,\tau) as x=yx=y (actually, it’s classical variant), we will use it without this restriction. Due to ξ\xi-microhyperbolicity

(2.11) Fth1τ(χ¯T(t)uh(x,y,t))(2πh)dhΣ(y,τ)eih1φ(x,y,θ)B(x,y,θ)𝑑θ:da(y,θ)F_{t\to h^{-1}\tau^{\prime}}\bigl{(}\bar{\chi}_{T}(t)u_{h}(x,y,t)\bigr{)}\\ \sim(2\pi h)^{-d}h\int_{\Sigma(y,\tau^{\prime})}e^{ih^{-1}\varphi(x,y,\theta)}B^{\prime}(x,y,\theta)\,d\theta:da(y,\theta)

with Σ(y,τ)={θ:a(y,θ)=0}\Sigma(y,\tau)=\{\theta\colon a(y,\theta)=0\}, dθ:da(y,θ)d\theta:da(y,\theta) a natural density on Σ(y,τ)\Sigma(y,\tau^{\prime}) and B(x,y,θ)B^{\prime}(x,y,\theta) also allowing (2.10) decomposition with uniformly smooth BnB^{\prime}_{n}.

Proposition 2.2.

Let in B(x¯,2ϵ1)B(\bar{x},2\epsilon_{1}) both ξ\xi-microhyperbolicity and strong convexity conditions are fulfilled on energy level τ\tau. Then for |ττ|ϵ1|\tau^{\prime}-\tau|\leq\epsilon_{1} and xyx\neq y φ(x,y,θ)\varphi(x,y,\theta) has exactly two stationary points θ±(x,y,τ)\theta^{*}_{\pm}(x,y,\tau^{\prime}) on Σ(y,τ)\Sigma(y,\tau^{\prime}), defined from

(2.12) θφ(x,y,θ)=t±θa(y,θ),\displaystyle\nabla_{\theta}\varphi(x,y,\theta)=-t_{\pm}\nabla_{\theta}a(y,\theta), a(y,θ)=τ,\displaystyle a(y,\theta)=\tau^{\prime}, ±t±>0.\displaystyle\pm t_{\pm}>0.

These points are non-degenerate and t±|xy|t_{\pm}\asymp|x-y| in (2.12).

Furthermore,

(2.13) θ±(x,y,τ)=θ¯±(x,y,τ)+O(|xy|),\displaystyle\theta^{*}_{\pm}(x,y,\tau^{\prime})=\bar{\theta}_{\pm}(x,y,\tau^{\prime})+O(|x-y|),
where θ¯±(x,y,τ)\bar{\theta}_{\pm}(x,y,\tau^{\prime}) are defined from
(2.14) xy=t±θa(y,θ),a(y,θ)=τ,±t±>0\displaystyle x-y=-t^{\prime}_{\pm}\nabla_{\theta}a(y,\theta),\qquad a(y,\theta)=\tau^{\prime},\quad\pm t^{\prime}_{\pm}>0

and also t±=t±+O(|xy|2)t^{\prime}_{\pm}=t_{\pm}+O(|x-y|^{2}).

Proof.

Proof follows trivially from

(2.15) φ(x,y,θ)=xy,θ+O(|xy|2),\displaystyle\varphi(x,y,\theta)=\langle x-y,\theta\rangle+O(|x-y|^{2}),

which follows from (2.8), (2.9).

Proposition 2.3.

In the framework of Proposition 2.2 the following asymptotics hold

(2.16) Fth1τ(χ¯T(t)uh(x,y,t))(2π)dh1d2ς=±eih1Sς(x,y,τ)Bς(x,y,τ)\displaystyle F_{t\to h^{-1}\tau}\bigl{(}\bar{\chi}_{T}(t)u_{h}(x,y,t)\bigr{)}\sim(2\pi)^{-d}h^{\frac{1-d}{2}}\sum_{\varsigma=\pm}e^{ih^{-1}S_{\varsigma}(x,y,\tau)}B^{\prime}_{\varsigma}(x,y,\tau)
and with cTϵc\ell\leq T\leq\epsilon
(2.17) eT,h𝖳(x,y,τ)(2π)dh1d2ς=±eih1Sς(x,y,τ)Bς′′(x,y,τ)\displaystyle e^{\mathsf{T}}_{T,h}(x,y,\tau)\sim(2\pi)^{-d}h^{\frac{1-d}{2}}\sum_{\varsigma=\pm}e^{ih^{-1}S_{\varsigma}(x,y,\tau)}B^{\prime\prime}_{\varsigma}(x,y,\tau)
with
(2.18) Sς(x,y,τ)=φ(x,y,θς(x,y,τ))\displaystyle S_{\varsigma}(x,y,\tau)=\varphi(x,y,\theta^{*}_{\varsigma}(x,y,\tau))

and BςB^{\prime}_{\varsigma}, Bς′′B^{\prime\prime}_{\varsigma} also decomposed into asymptotic series albeit with Bςnn+d12B^{\prime}_{\varsigma n}\ell^{n+\frac{d-1}{2}} and Bςn′′n+d+12B^{\prime\prime}_{\varsigma n}\ell^{n+\frac{d+1}{2}}uniformly smooth.

Proof.

Asymptotics (2.17) follows from Proposition 2.2 and stationary phase principle with effective semiclassical parameter =h1\hbar=h\ell^{-1}, =|xy|+h\ell=|x-y|+h.

To prove (2.18) we first rewrite (1.12) as the sum of two integrals with cut-off functions ϕ1(τ)\phi_{1}(\tau^{\prime}) and ϕ2)(τ)\phi_{2})(\tau^{\prime}); ϕj𝒞0\phi_{j}\in\mathscr{C}_{0}^{\infty}, supp(ϕ1)(τϵ,τ+ϵ)\operatorname{supp}(\phi_{1})\subset(\tau-\epsilon,\tau+\epsilon), supp(ϕ2)(,τϵ/2)\operatorname{supp}(\phi_{2})\subset(-\infty,\tau-\epsilon/2). In the integral with ϕ1\phi_{1} we plug (2.4), observe that there are no stationary points in {θ:τϵ<a(y,θ)<τ}\{\theta\colon\tau-\epsilon<a(y,\theta)<\tau\} and use Proposition 2.2 and stationary phase principle with effective semiclassical parameter =h1\hbar=h\ell^{-1}. In the integral with ϕ2\phi_{2} we simply observe that it has a complete asymptotics because it contains a mollification by τ\tau^{\prime} (see, f.e. Proposition 5.7 below).

Proposition 2.4.

In the framework of Proposition 2.2

  1. (i)

    The following estimates hold:

    (2.19) |Fth1τ(χ¯T(t)uh(x,y,t))|Ch1d21d2\displaystyle|F_{t\to h^{-1}\tau}\Bigl{(}\bar{\chi}_{T}(t)u_{h}(x,y,t)\Bigr{)}|\leq Ch^{\frac{1-d}{2}}\ell^{\frac{1-d}{2}}
    and as cTϵc\ell\leq T\leq\epsilon
    (2.20) |eT,h𝖳(x,y,τ)|Ch1d21d2.\displaystyle|e^{\mathsf{T}}_{T,h}(x,y,\tau)|\leq Ch^{\frac{1-d}{2}}\ell^{\frac{-1-d}{2}}.
  2. (ii)

    The following estimate holds as max(c,h1δ)Tϵ\max(c\ell,\,h^{1-\delta})\leq T\leq\epsilon:

    (2.21) |eT,h𝖳(x,y,τ)eh𝖶(x,y,τ)|Ch1d.\displaystyle|e^{\mathsf{T}}_{T,h}(x,y,\tau)-e^{\mathsf{W}}_{h}(x,y,\tau)|\leq Ch^{1-d}.

Proof.
  1. (a)

    Estimates (2.19)–(2.20) follow from (2.16)–(2.17).

  2. (b)

    To prove Statement (ii) for d3d\geq 3 one can observe that both |eT,h𝖳(x,y,τ||e^{\mathsf{T}}_{T,h}(x,y,\tau| and |eW(x,y,τ)||e^{W}(x,y,\tau)| do not exceed Ch1dCh^{1-d} as h12ϵh^{\frac{1}{2}}\leq\ell\leq\epsilon, while for h12\ell\leq h^{\frac{1}{2}}, due to (2.15), replacing Φ(x,y,t,θ)\Phi(x,y,t,\theta) by

    (2.22) Φ¯(x,y,θ)=xy,θ+ta(12(x+y),θ)\displaystyle\bar{\Phi}(x,y,\theta)=\langle x-y,\theta\rangle+ta(\frac{1}{2}(x+y),\theta)

    brings an error not exceeding the right-hand expression of (2.20) multiplied by C2h1C\ell^{2}h^{-1}, that is Ch1+d23d2Ch1dCh^{-\frac{1+d}{2}}\ell^{\frac{3-d}{2}}\leq Ch^{1-d}.

    Moreover, if we also replace B(x,y,t,θ,h)B(x,y,t,\theta,h) by 11 then the error will not exceed the right-hand expression of (2.20) multiplied by CC\ell, that is Ch1d21d2Ch1dCh^{\frac{1-d}{2}}\ell^{\frac{1-d}{2}}\leq Ch^{1-d}.

    Finally, with this choice of the phase function Φ¯(x,y,t,θ)\bar{\Phi}(x,y,t,\theta) and amplitude B=1B=1, we get eT,h𝖳(x,y,τ)=e𝖶(x,y,τ)+O(h)e^{\mathsf{T}}_{T,h}(x,y,\tau)=e^{\mathsf{W}}(x,y,\tau)+O(h^{\infty}).

  3. (c)

    As d=2d=2 we need more subtle arguments (which would also work for d3d\geq 3). Indeed, estimate (2.20) implies that |eT,h𝖳(x,y,τ)|Ch1|e^{\mathsf{T}}_{T,h}(x,y,\tau)|\leq Ch^{-1} as h13ϵ1h^{\frac{1}{3}}\leq\ell\leq\epsilon_{1} and estimate (2.21) leads to O(h1)O(h^{-1}) error only as h\ell\asymp h.

    We refer to Proposition 9.20(i).

Finally, Theorem 1.1 follows from estimate (1.11) and Proposition 2.4.

3 Improvements and generalizations

Two-term asymptotics.

We know that if AA is a scalar operator on the closed manifold, ξ\xi-microhyperbolic at point xx and if non-looping condition

(2.23) μx,τ{ξΣ(x,τ):t0,πxetHa(x,ξ)=x}=0\displaystyle\upmu_{x,\tau}\{\xi\in\Sigma(x,\tau)\colon\exists t\neq 0,\uppi_{x}e^{tH_{a}(x,\xi)}=x\}=0
is fulfilled at point xx then two-term asymptotics
(2.24) eh(x,x,τ)=κ0(x,τ)hd+κ1(x,τ)h1d+o(h1d)\displaystyle e_{h}(x,x,\tau)=\kappa_{0}(x,\tau)h^{-d}+\kappa_{1}(x,\tau)h^{1-d}+o(h^{1-d})

holds; here Σ(x,τ)={ξ:a(x,ξ)=τ}\Sigma(x,\tau)=\{\xi\colon a(x,\xi)=\tau\}, μx,τ\upmu_{x,\tau} is a natural measure on Σ(x,τ)\Sigma(x,\tau), corresponding to a density dxdξ:dξa(x,ξ)dxd\xi:d_{\xi}a(x,\xi), HaH_{a} is a Hamiltonian field generated by aa and etHae^{tH_{a}} is a Hamiltonian flow (see, f.e. [Ivr2], Section LABEL:monsterbook-sect-5-3). Can we get a two-term asymptotics for eh(x,y,τ)e_{h}(x,y,\tau)?

First of all, we need to improve Tauberian estimate (1.11). Asymptotics (2.24) at point xx implies

(2.25) |e(x,x,τ+T1h)e(x,x,τ)|CT1h1d+oT(h1d)\displaystyle|e(x,x,\tau+T^{-1}h)-e(x,x,\tau^{\prime})|\leq CT^{-1}h^{1-d}+o_{T}(h^{1-d})

and then

|e(x,y,τ)e(x,y,τ)|C|ττ|hd+C|ττ|12h12d+CT12h1d+oT(h1d)|e(x,y,\tau)-e(x,y,\tau^{\prime})|\\ \leq C|\tau-\tau^{\prime}|h^{-d}+C|\tau-\tau^{\prime}|^{\frac{1}{2}}h^{\frac{1}{2}-d}+CT^{-\frac{1}{2}}h^{1-d}+o_{T}(h^{1-d})

with arbitrarily large TT. Here condition (2.23) is fulfilled at one of points xx, yy. Then the standard Tauberian arguments imply

|eh(x,y,τ)eT,h𝖳(x,y,τ)|CT12h1d+oT(h1d)\displaystyle|e_{h}(x,y,\tau)-e_{T,h}^{\mathsf{T}}(x,y,\tau)|\leq CT^{-\frac{1}{2}}h^{1-d}+o_{T}(h^{1-d})

and propagation results combined with non-looping condition at one of points xx, yy imply that in the left hand expression one can replace TT by T=T(x,y)T^{\prime}=T(x,y) while still having arbitrary TT in the right hand-expression, and then we arrive to

(2.26) eh(x,y,τ)=eT,h𝖳(x,y,τ)+o(h1d)with T=T(x,y).\displaystyle e_{h}(x,y,\tau)=e_{T,h}^{\mathsf{T}}(x,y,\tau)+o(h^{1-d})\qquad\text{with\ \ }T=T(x,y).

Further, if we add another condition

(2.27) μx,τ{ξΣ(x,τ):t0,πxetHa(x,ξ)=y}=0\displaystyle\upmu_{x,\tau}\{\xi\in\Sigma(x,\tau)\colon\exists t\neq 0,\uppi_{x}e^{tH_{a}(x,\xi)}=y\}=0

or a similar condition, obtained by permutation of xx and yy (let’s call it (2.26)(\ref{eqn-2.26})^{\prime}), then from (2.26) and propagation results we obtain that

  1. (a)

    If (x,y)ϵ\ell(x,y)\geq\epsilon then eh(x,y,τ)=o(h1d)e_{h}(x,y,\tau)=o(h^{1-d}),

  2. (b)

    If (x,y)ϵ\ell(x,y)\leq\epsilon, then we can take T(x,y)=c(x,y)T(x,y)=c\ell(x,y) in (2.26).

Weyl asymptotics.

In particular, in the framework of Theorem 1.1, under extra assumptions (2.23) and either (2.26) or (2.26)(\ref{eqn-2.26})^{\prime}, we can apply the machinery developed in Subsection 2 (with the only exception d=2d=2 and neither h13\ell\gg h^{\frac{1}{3}} nor h13\ell\ll h^{\frac{1}{3}}). One can prove easily that then

(2.28) (e(x,y,τ)=e𝖶(x,y,τ))+o(h1d)x,yB(0,ϵ)\displaystyle\bigl{(}e(x,y,\tau)=e^{\mathsf{W}}(x,y,\tau)\bigr{)}+o(h^{1-d})\qquad\forall x,y\in B(0,\epsilon)

where eh𝖶(x,y,τ)e_{h}^{\mathsf{W}}(x,y,\tau) is defined by (1.3) with aa replaced by a+ha1a+ha_{1} where a1a_{1} is the subprincipal symbol of AA, that is with the main term defined by (1.3) and with the second term

(2.29) (2πh)1dΣ(z,τ)a1(z,θ)eixy,θ𝑑θ:dθa(z,θ),z=12(x+y).\displaystyle-(2\pi h)^{1-d}\int_{\Sigma(z,\tau)}a_{1}(z,\theta)e^{i\langle x-y,\theta\rangle}\,d\theta:d_{\theta}a(z,\theta),\qquad z=\frac{1}{2}(x+y).
Generalizations.
  1. (i)

    We can consider the case when a(x,ξ)a(x,\xi) has characteristic roots λj(x,ξ)\lambda_{j}(x,\xi) of constant multiplicity; and we need only to assume that it happens as |λj(x,ξ)τ|ϵ|\lambda_{j}(x,\xi)-\tau|\leq\epsilon.

  2. (ii)

    If we consider only y:yxΓy\colon y-x\in\Gamma where Γ\Gamma is a cone in d\mathbb{R}^{d} with a vertex at 0, then in virtue of propagation results (see [Ivr2], Section LABEL:monsterbook-sect-2-2) we need the previous assumption only for ξ\xi such that

    (2.30) Γ(K(x,ξ)K(x,ξ))\displaystyle\Gamma\cap(K(x,\xi)\cup-K(x,\xi))\neq\emptyset

    where K(x,ξ)K(x,\xi) is a cone dual to K(x,ξ)K^{\prime}(x,\xi) which is a connected component of

    {η:((ηξa(x,ξ))v,v)0;v:(aτ)vϵv}.\displaystyle\{\eta\colon((\eta\cdot\partial_{\xi}a(x,\xi))v,v)\geq 0\ ;\forall v\colon\|(a-\tau)v\|\leq\epsilon\|v\|\}.
  3. (iii)

    Also, instead of strong convexity we need to assume only that at points ξ\xi satisfying (2.30), the matrix of the curvatures of the surface {ξ:λ(x,ξ)=τ}\{\xi\colon\lambda(x,\xi)=\tau\} has at least two eigenvalues, disjoint from 0. Then it would be similar to the case d3d\geq 3 in Proposition 2.4. Probably, it would suffice to have just one eigenvalue, to recover more delicate arguments of the case d=2d=2 in Proposition 2.4.

Chapter 3 Microlocal methods

4 Propagation

Recall that we consider Schrödinger operator (1.4) and for such operator under ξ\xi-microhyperbolicity near boundary condition on the energy level τ\tau instead of on-diagonal asymptotics (2.1) asymptotics with a boundary-layer type term and remainder estimate O(h1d)O(h^{1-d}) holds; see [Ivr2], Section LABEL:monsterbook-sect-8-1. In particular, under Dirichlet or Neumann boundary conditions (and we consider only those for simplicity) asymptotics (1.5) holds as x=yx=y and also (1.9) holds. Then (1.10)–(1.12) hold and all we need is to rewrite Tauberian expression (1.12) in more explicit terms.

Let us decompose

(3.1) uh(x,y,t)=uh0(x,y,t)+uh1(x,y,t)\displaystyle u_{h}(x,y,t)=u_{h}^{0}(x,y,t)+u_{h}^{1}(x,y,t)
where uh0(x,y,t)u_{h}^{0}(x,y,t) is the solution for the “free space” and uh1(x,y,t)u_{h}^{1}(x,y,t) is a reflected wave, satisfying
(3.2) hDtuh1=A(x,hD,h)uh1,\displaystyle hD_{t}u_{h}^{1}=A(x,hD,h)u_{h}^{1},
(3.3) B(x,hD,h)uh1|X=B(x,hD,h)uh0|X,\displaystyle B(x,hD,h)u_{h}^{1}|_{\partial X}=-B(x,hD,h)u_{h}^{0}|_{\partial X},

where B(x,hD,h)B(x,hD,h) is a boundary operator. Then

(3.4) eh𝖳(x,y,τ)=eh0,𝖳(x,y,τ)+eh1,𝖳(x,y,τ),\displaystyle e^{\mathsf{T}}_{h}(x,y,\tau)=e^{0,\mathsf{T}}_{h}(x,y,\tau)+e^{1,\mathsf{T}}_{h}(x,y,\tau),

where eh0,𝖳(x,y,τ)e^{0,\mathsf{T}}_{h}(x,y,\tau) and eh1,𝖳(x,y,τ)e^{1,\mathsf{T}}_{h}(x,y,\tau) are derived by Tauberian expression (1.12) from uh0(x,y,t)u_{h}^{0}(x,y,t) and u1(x,y,t)u^{1}(x,y,t) correspondingly with T=ϵ1T=\epsilon_{1}.

While uh0(x,y,t)u^{0}_{h}(x,y,t) can be constructed as an oscillatory integral, our purpose is to construct uh1(x,y,t)u^{1}_{h}(x,y,t) in some way. To do this, in this section we apply the improved method of successive approximations near boundary, in the same way as the standard method was applied in [Ivr2], Section LABEL:monsterbook-sect-7-2 while in the next Section 5 we use the geometric optics method.

Recall that now AA is a second order elliptic operator and XB(0,1)={x:ν(x)>0}B(0,1)X\cap B(0,1)=\{x\colon{\nu}(x)>0\}\cap B(0,1) with |ν|ϵ0|\nabla{\nu}|\geq\epsilon_{0}. Without any loss of the generality one can assume that

Claim 1.

XB(0,1)={x:x1>0}B(0,1)X\cap B(0,1)=\{x\colon x_{1}>0\}\cap B(0,1) with V1(x)=0V_{1}(x)=0, g1k=δ1kg^{1k}=\updelta_{1k}

and therefore

(3.5) ν(x)=x1,(x,y)=|xy|+x1+y1\displaystyle{\nu}(x)=x_{1},\qquad\ell(x,y)=|x^{\prime}-y^{\prime}|+x_{1}+y_{1}

and AA is ξ\xi-microhyperbolic on the energy level τ\tau at xx as long as

(3.6) |V(x)τ|ϵ0,\displaystyle|V(x)-\tau|\geq\epsilon_{0},

and in this case {ξ:a(x,ξ)=τ}\{\xi\colon a(x,\xi)=\tau\} is strongly convex.

Proposition 4.1.
  1. (i)

    As Tϵ(x,y)T\leq\epsilon\ell(x,y) the following estimate holds:

    (3.7) |Fth1τ(χ¯T(t)uh1(x,y,t))|Chd(hT)s.\displaystyle|F_{t\to h^{-1}\tau}\Bigl{(}\bar{\chi}_{T}(t)u^{1}_{h}(x,y,t)\Bigr{)}|\leq Ch^{-d}\Bigl{(}\frac{h}{T}\Bigr{)}^{s}.
  2. (ii)

    Let condition (3.1) be fulfilled. Then for max(C(x,y),h1δ)TTϵ1\max(C\ell(x,y),\,h^{1-\delta})\leq T\leq T^{\prime}\leq\epsilon_{1} the following estimate holds:

    (3.8) |Fth1τ((χ¯T(t)χ¯T(t))uh1(x,y,t))|Chd(hT)s.\displaystyle|F_{t\to h^{-1}\tau}\Bigl{(}\bigl{(}\bar{\chi}_{T^{\prime}}(t)-\bar{\chi}_{T}(t)\bigr{)}u^{1}_{h}(x,y,t)\Bigr{)}|\leq Ch^{-d}\Bigl{(}\frac{h}{T}\Bigr{)}^{s}.

Proof 4.2.

Statement (i) follows from the finite speed of propagation and Statement (ii) follows from the fact that the speed of propagation with respect to xx is disjoint from 0 (either along X\partial X, if |ξ||\xi^{\prime}| is disjoint from 0 or in direction of x1x_{1} if ξ1\xi_{1} is disjoint from 0)–see [Ivr2], Chapter LABEL:monsterbook-sect-3.

Corollary 4.3.

Let eh1,𝖳(x,y,τ)e^{1,\mathsf{T}}_{h}(x,y,\tau) be defined by (1.12) with uh1(x,y,t)u^{1}_{h}(x,y,t) instead of uh(x,y,t)u_{h}(x,y,t).

  1. (i)

    Then in eh1,𝖳(x,y,τ)e^{1,\mathsf{T}}_{h}(x,y,\tau) one can replace T=ϵ1T=\epsilon_{1} by T=Cmax((x,y),h1δ)T=C\max(\ell(x,y),\,h^{1-\delta}).

  2. (ii)

    Further, in the framework of Proposition 4.1(ii) and (x,y)h1δ\ell(x,y)\geq h^{1-\delta} one can replace χ¯T(t)\bar{\chi}_{T}(t) with T=ϵ1T=\epsilon_{1} by (χ¯T(t)χ¯T(t))\bigl{(}\bar{\chi}_{T}(t)-\bar{\chi}_{T^{\prime}}(t)\bigr{)} with T=C0(x,y)T=C_{0}\ell(x,y) and T=C01(x,y)T^{\prime}=C_{0}^{-1}\ell(x,y).

We need to be more precise about propagation especially with respect to x1x_{1} and ξ1\xi_{1} as 0\ell^{0}\asymp\ell. Let us x¯\bar{x}, y¯\bar{y} and w¯\bar{w} be fixed final values of xx, yy and ww respectively and we need to consider case (x¯,y¯)h1δ\ell\coloneqq\ell(\bar{x},\bar{y})\geq h^{1-\delta}.

Proposition 4.4.

Let h1δTϵ1h^{1-\delta}\leq T\leq\epsilon_{1}, T=C0T=C_{0}\ell and

(4.4)1,2\textup{(\ref*{eqn-3.10})}_{1,2} ρTh1δ,ρT2.\displaystyle\rho T\geq h^{1-\delta},\qquad\rho\geq T^{2}.

Let b(y¯,ξ¯)=τb(\bar{y}^{\prime},\bar{\xi}^{\prime})=\tau and

(3.10) |yb(y,ξ¯)y=y¯|ρT1\displaystyle|\nabla_{y^{\prime}}b(y^{\prime},\bar{\xi})_{y^{\prime}=\bar{y^{\prime}}}|\leq\rho T^{-1}
where here and below
(3.11) b(x,ξ)a(x,ξ)|x1=ξ1=0;\displaystyle b(x^{\prime},\xi^{\prime})\coloneqq a(x,\xi)|_{x_{1}=\xi_{1}=0};

and let Q1(y,η)Q_{1}(y,\eta^{\prime}) have a symbol supported in (T,ρ)(T,\rho)-vicinity of (y¯,ξ¯)(\bar{y},\bar{\xi}^{\prime}).

  1. (i)

    Let Q2(x,ξ)Q_{2}(x,\xi^{\prime}) have a symbol, equal 0 in {(x,ξ):|ξξ¯|C1ρ}\{(x,\xi^{\prime})\colon|\xi^{\prime}-\bar{\xi}^{\prime}|\leq C_{1}\rho\}. Then

    (3.12) Fth1τ(χ¯T(t)Q2xuh1(x,y,t)tQ1y)=O(hs).\displaystyle F_{t\to h^{-1}\tau}\Bigl{(}\bar{\chi}_{T}(t)Q_{2x}u^{1}_{h}(x,y,t)\,^{t}\!Q_{1y}\Bigr{)}=O(h^{s}).
  2. (ii)

    Assume instead that Q2(x,hD)Q_{2}(x,hD^{\prime}) has a symbol, equal 0 as x1C1σTx_{1}\leq C_{1}\sigma T where here and below

    (3.13) σ=(ρ12+T).\displaystyle\sigma=(\rho^{\frac{1}{2}}+T).

    Then (3.12) holds.

  3. (iii)

    Assume instead that Q2(x,hD)Q_{2}(x,hD^{\prime}) has a symbol, equal 0 in

    (3.14) {(x,ξ):|xy¯+tξb(y¯,ξ)|ξ=ξ¯|C1ρT}t:ϵT±tC0T.\{(x^{\prime},\xi^{\prime})\colon|x^{\prime}-\bar{y}^{\prime}+t\nabla_{\xi^{\prime}}b(\bar{y}^{\prime},\xi^{\prime})|_{\xi=\bar{\xi}^{\prime}}|\leq C_{1}\rho T\}\\ \forall t\colon\epsilon T\leq\pm t\leq C_{0}T.

    Assume also that

    (3.15) ρ2Th1δ,ρC0T.\displaystyle\rho^{2}T\geq h^{1-\delta},\qquad\rho\geq C_{0}T.
    Then
    (3.16) Fth1τ(χT±(t)Q2xuh1(x,y,t)tQ1y)=O(hs).\displaystyle F_{t\to h^{-1}\tau}\Bigl{(}\chi^{\pm}_{T}(t)Q_{2x}u^{1}_{h}(x,y,t)\,^{t}\!Q_{1y}\Bigr{)}=O(h^{s}).

    where χ±𝒞0()\chi^{\pm}\in\mathscr{C}^{\infty}_{0}(\mathbb{R}) is supported in {t:ϵ±t1}\{t\colon\epsilon\leq\pm t\leq 1\}.

Proof 4.5.
  1. (i)

    Observe that under assumptions (4.4)2\textup{(\ref{eqn-3.10})}_{2} and (3.10) the propagation speed with respect to ξ\xi^{\prime} does not exceed C1(ρ+T)C_{1}(\rho+T) in

    {(x,ξ):|xy¯|+|x1|+|y1|C1T,|ξξ¯|C1ρ}.\displaystyle\{(x,\xi^{\prime})\colon|x^{\prime}-\bar{y}^{\prime}|+|x_{1}|+|y_{1}|\leq C_{1}T,\ |\xi^{\prime}-\bar{\xi}^{\prime}|\leq C_{1}\rho\}.

    This implies Statement (i) by the standard methods of propagation; see Subsection LABEL:monsterbook-sect-5-1-2 of [Ivr2]. Here (4.4)1\textup{(\ref*{eqn-3.10})}_{1} is the microlocal uncertainty principle).

  2. (ii)

    Due to Statement (i) the propagation is confined to {(x,ξ):|b(x,ξ)τ|C0σ2}\{(x,\xi^{\prime})\colon|b(x^{\prime},\xi^{\prime})-\tau|\leq C_{0}\sigma^{2}\} and thus to |ξ1|Cσ|\xi_{1}|\leq C\sigma and then to x1C1σx_{1}\leq C_{1}\sigma\ell. The microlocal uncertainty principle with respect to (x1,ξ1)(x_{1},\xi_{1}), that is σ×σh1δ\sigma\times\sigma\ell\geq h^{1-\delta} is fulfilled due to (4.4)1\textup{(\ref{eqn-3.10})}_{1} and therefore we can consider propagation with respect to (x1,ξ1)(x_{1},\xi_{1}) as x1σTx_{1}\geq\sigma T.

  3. (iii)

    Due to Statement (i) the propagation in the time direction ±t>0\pm t>0 is confined to the domain (3.14) and here (3.15) is a microlocal uncertainty principle.

Remark 4.6.
  1. (i)

    We can achieve (3.10) by gauge transformation, not affecting x1,ξ1x_{1},\xi_{1}. Further, if

    (3.17) ρT\displaystyle\rho\geq T

    then (3.10) is fulfilled automatically.

  2. (ii)

    Obviously, condition (3.10) is important only as ρT\rho\leq T, which is not the case as Th12T\leq h^{\frac{1}{2}}. Therefore we can cover also h1δ\ell\leq h^{1-\delta} with T=h1δT=h^{1-\delta}.

  3. (iii)

    Both statements of Proposition 4.4 hold for τ:|ττ|σ2=(ρ+2)\tau^{\prime}\colon|\tau-\tau^{\prime}|\leq\sigma^{2}=(\rho+\ell^{2}).

5 Spectral estimates

The following proposition will be useful in the proof of Theorem 1.2:

Proposition 5.7.
  1. (i)

    Let L0(x,y)h1δL\ell^{0}(x,y)\geq h^{1-\delta}. Let ϕ𝒞0()\phi\in\mathscr{C}_{0}^{\infty}(\mathbb{R}). Then

    (3.18) ϕ((ττ)/L)Fth1τ((χ¯T(t)uh(x,y,t))dτ=O(hs).\displaystyle\int\phi((\tau-\tau^{\prime})/L)F_{t\to h^{-1}\tau^{\prime}}\Bigl{(}(\bar{\chi}_{T}(t)u_{h}(x,y,t)\Bigr{)}\,d\tau^{\prime}=O(h^{s}).
  2. (ii)

    Let L(x,y)h1δL\ell(x,y)\geq h^{1-\delta}. Let ϕ𝒞0()\phi\in\mathscr{C}_{0}^{\infty}(\mathbb{R}). Then

    (3.19) ϕ((ττ)/L)Fth1τ((χ¯T(t)uh1(x,y,t))dτ=O(hs).\displaystyle\int\phi((\tau-\tau^{\prime})/L)F_{t\to h^{-1}\tau^{\prime}}\Bigl{(}(\bar{\chi}_{T}(t)u^{1}_{h}(x,y,t)\Bigr{)}\,d\tau^{\prime}=O(h^{s}).

Proof 5.8.

The left-hand expressions are just ϕ((hDtτ)/L)uh(x,y,t)|t=0\phi((hD_{t}-\tau)/L)u_{h}(x,y,t)\bigr{|}_{t=0} and ϕ((hDtτ)/L)uh1(x,y,t)|t=0\phi((hD_{t}-\tau)/L)u^{1}_{h}(x,y,t)\bigr{|}_{t=0} correspondingly and we apply finite speed of propagation.

Proposition 5.9.

Under ξ\xi-microhyperbolicity condition on the energy level τ\tau estimates (1.8) and

(3.20) |eh(x,y,τ)|Ch1d(1+01(x,y))\displaystyle|e_{h}(x,y,\tau)|\leq Ch^{1-d}\bigl{(}1+\ell^{0\,-1}(x,y)\bigr{)}

hold.

Proof 5.10.

In virtue of Subsection 1 for h𝖳ϵh\leq\mathsf{T}\leq\epsilon

eh(x,y,τ)=eh𝖳(x,y,τ)+O(T1h1d)=eh0,𝖳(x,y,τ)+eh1,𝖳(x,y,τ)+O(T1h1d)e_{h}(x,y,\tau)=e^{\mathsf{T}}_{h}(x,y,\tau)+O(T^{-1}h^{1-d})\\ =e^{0,\mathsf{T}}_{h}(x,y,\tau)+e^{1,\mathsf{T}}_{h}(x,y,\tau)+O(T^{-1}h^{1-d})

and for T=ϵ1(x,y)T=\epsilon_{1}\ell(x,y)

eh0,𝖳(x,y,τ)=e0,𝖶(x,y,τ)+O(hd(h1)s)\displaystyle e^{0,\mathsf{T}}_{h}(x,y,\tau)=e^{0,\mathsf{W}}(x,y,\tau)+O(h^{-d}(h\ell^{-1})^{s})
and
eh1,𝖳(x,y,τ)=O(hs)\displaystyle e^{1,\mathsf{T}}_{h}(x,y,\tau)=O(h^{s})

due to the finite speed of propagation, which implies (1.8). Estimate (3.20) is proven in the same way.

As h1δ(x¯,(¯y))ϵ1h^{1-\delta}\leq\ell(\bar{x},\bar{(}y))\leq\epsilon_{1} let us introduce ξ¯±\bar{\xi}^{\prime\pm}:

(3.21) ξb(w¯,ξ)|ξ=ξ¯±=t(y¯x¯),b(w¯,ξ¯±)=τ,±t>0\displaystyle\nabla_{\xi^{\prime}}b(\bar{w}^{\prime},\xi^{\prime})|_{\xi^{\prime}=\bar{\xi}^{\prime\pm}}=t(\bar{y}-\bar{x}),\qquad b(\bar{w}^{\prime},\bar{\xi}^{\prime\pm})=\tau,\qquad\pm t>0

with w=12(x¯+y¯)w^{\prime}=\frac{1}{2}(\bar{x}^{\prime}+\bar{y}^{\prime}). Due to strong convexity ξ¯±\bar{\xi}^{\prime\pm} are defined uniquely and t(x¯,(¯y))t\asymp\ell(\bar{x},\bar{(}y)).

Proposition 5.11.

Let h1δTϵ1h^{1-\delta}\leq T\leq\epsilon_{1}, T=C0T=C_{0}\ell, (3.17) and (3.15) be fulfilled. Then

(3.22) eh1,𝖳(x¯,y¯,τ)=Q2xeh1,𝖳(x,y,τ)tQ1y|x=x¯,y=y¯+O(hs)\displaystyle e_{h}^{1,\mathsf{T}}(\bar{x},\bar{y},\tau)=Q_{2x}e_{h}^{1,\mathsf{T}}(x,y,\tau)\,^{t}\!Q_{1y}|_{x=\bar{x},y=\bar{y}}+O(h^{s})

where Q1Q_{1}, Q2Q_{2} are operators with symbols equal 11 in (C1,C1ρ)(C_{1}\ell,C_{1}\rho) vicinities of (w¯,ξ¯±)(\bar{w}^{\prime},\bar{\xi}^{\prime\pm}).

Proof 5.12.

Due to Proposition 5.7

(3.23) eh1,𝖳(x,y,τ)=h1τϕ¯((ττ)L1)Fth1τ(χ¯T(t)uh1(x,y,t))𝑑τ+O(hs)e_{h}^{1,\mathsf{T}}(x,y,\tau)\\ =h^{-1}\int_{-\infty}^{\tau}\bar{\phi}((\tau-\tau^{\prime})L^{-1})F_{t\to h^{-1}\tau^{\prime}}\bigl{(}\bar{\chi}_{T}(t)u^{1}_{h}(x,y,t)\bigr{)}\,d\tau^{\prime}+O(h^{s})

where ϕ¯𝒞()\bar{\phi}\in\mathscr{C}^{\infty}(\mathbb{R}), ϕ¯(τ)=1\bar{\phi}(\tau)=1 as τ1\tau\leq 1 and ϕ¯(τ)=0\bar{\phi}(\tau)=0 as τ2\tau\geq 2, Lh1δL\ell\geq h^{1-\delta}.

Due to Proposition 4.1(ii) we can replace χ¯T(t)\bar{\chi}_{T}(t) by χT(t)=χ¯T(t)χ¯ϵT(t)\chi_{T}(t)=\bar{\chi}_{T}(t)-\bar{\chi}_{\epsilon T}(t). Then Proposition 4.4(iii) implies that if ψ1,ψ2𝒞0\psi_{1},\psi_{2}\in\mathscr{C}_{0}^{\infty} are supported in ϵ\epsilon\ell-vicinities of y¯\bar{y} and x¯\bar{x} respectively and Q=Q(hD)Q^{\prime}=Q^{\prime}(hD^{\prime}) is 0 in CρC\rho-vicinities of ξ¯±\bar{\xi}^{\prime\pm} then as |ττ|L=h1δL1|\tau^{\prime}-\tau|\leq L=h^{1-\delta}L^{-1}

Fth1τ(χT(t)Qxψ2(x)uh1(x,y,t)ψ1(y))=O(hs),\displaystyle F_{t\to h^{-1}\tau^{\prime}}\Bigl{(}\chi_{T}(t)Q^{\prime}_{x}\psi_{2}(x)u^{1}_{h}(x,y,t)\psi_{1}(y)\Bigr{)}=O(h^{s}),
Fth1τ(χT(t)ψ2(x)uh1(x,y,t)ψ1(y)tQy)=O(hs)\displaystyle F_{t\to h^{-1}\tau^{\prime}}\Bigl{(}\chi_{T}(t)\psi_{2}(x)u^{1}_{h}(x,y,t)\psi_{1}(y)\,^{t}\!Q^{\prime}_{y}\Bigr{)}=O(h^{s})

Indeed condition now are stronger than those of Proposition 4.4(iii).

Then these equalities and (3.23) imply (3.22).

Proposition 5.13.

In the framework of Proposition 5.11

(3.24) |eT,h1,𝖳(x,y,τ)|Ch1d1σρd2\displaystyle|e^{1,\mathsf{T}}_{T,h}(x,y,\tau)|\leq Ch^{1-d}\ell^{-1}\sigma\rho^{d-2}

with σ=h12δ12+\sigma=h^{\frac{1}{2}-\delta^{\prime}}\ell^{-\frac{1}{2}}+\ell.

Proof 5.14.

Due to Propositions 4.1(ii) and 5.11 we need to estimate the right-hand expression of (3.23) with symbols of Q1Q_{1} and Q2Q_{2} supported in Ω\Omega, the union of (C1,C1ρ)(C_{1}\ell,C_{1}\rho)-vicinities of (w¯,ξ¯+)(\bar{w}^{\prime},\bar{\xi}^{\prime+}) and (w¯,ξ¯)(\bar{w}^{\prime},\bar{\xi}^{\prime-}) intersected with

{ξ:|b(x,ξ)τ|Cσ2},\displaystyle\{\xi\colon|b(x^{\prime},\xi^{\prime})-\tau|\leq C\sigma^{2}\},

and with χ¯T(t)\bar{\chi}_{T}(t) replaced by χT(t)=χ¯T(t)χ¯ϵT(t)\chi_{T}(t)=\bar{\chi}_{T}(t)-\bar{\chi}_{\epsilon T}(t). Observe that modulo O(hs)O(h^{s})

eh1,𝖳(x,y,τ)h1τFth1τ(χT(t)uh1(x,y,t))𝑑τ=T1Fth1τ(βT(t)uh1(x,y,t))e_{h}^{1,\mathsf{T}}(x,y,\tau)\equiv h^{-1}\int_{-\infty}^{\tau}F_{t\to h^{-1}\tau^{\prime}}\Bigl{(}\chi_{T}(t)u^{1}_{h}(x,y,t)\Bigr{)}\,d\tau^{\prime}\\ =T^{-1}F_{t\to h^{-1}\tau}\Bigl{(}\beta_{T}(t)u^{1}_{h}(x,y,t)\Bigr{)}

with β(t)=t1χ(t)\beta(t)=t^{-1}\chi(t).

Since 0\ell\asymp\ell^{0} it sufficient to prove it for eT,h𝖳(x,y,τ)e^{\mathsf{T}}_{T,h}(x,y,\tau). Expressing uh(x,y,t)=eih1tτdτeh(x,y,τ)u_{h}(x,y,t)=\int e^{ih^{-1}t\tau}\,d_{\tau}e_{h}(x,y,\tau) we see that expression in question equals

β^((ττ)Th)dτQ2xeh(x,y,τ)tQ1y\displaystyle\int\hat{\beta}\bigl{(}\frac{(\tau-\tau^{\prime})T}{h}\bigr{)}\,d_{\tau^{\prime}}Q_{2x}e_{h}(x,y,\tau^{\prime})\,^{t}\!Q_{1y}

which does not exceed

Csupτ:|ττ|L\displaystyle C\sup_{\tau^{\prime}\colon|\tau^{\prime}-\tau|\leq L} |Q2xeh(x,y,τ+hT1,τ)tQ1y|+Chs.\displaystyle|Q_{2x}e_{h}(x,y,\tau^{\prime}+hT^{-1},\tau^{\prime})\,^{t}\!Q_{1y}|+C^{\prime}h^{s}.
Since eh(x,y,τ)e_{h}(x,y,\tau) is the Schwartz kernel of the orthogonal projector this does not exceed
Csupτ:|ττ|L\displaystyle C\sup_{\tau^{\prime}\colon|\tau^{\prime}-\tau|\leq L} |Q2xeh(x,z,τ+hT1,τ)tQ2z|z=x|12\displaystyle|Q_{2x}e_{h}(x,z,\tau^{\prime}+hT^{-1},\tau^{\prime})\,^{t}\!Q_{2z}|_{z=x}|^{\frac{1}{2}}
×\displaystyle\times |Q1zeh(z,y,τ+hT1,τ)tQ1y|z=y|12+Chs.\displaystyle|Q_{1z}e_{h}(z,y,\tau^{\prime}+hT^{-1},\tau^{\prime})\,^{t}\!Q_{1y}|_{z=y}|^{\frac{1}{2}}+C^{\prime}h^{s}.

Applying Tauberian estimate for x=yx=y we see that the right-hand expression does not exceed

Ch1d1supτ:|ττ|LΣ(x,τ)(Ω×ξ1)𝑑ξ:dξa.\displaystyle Ch^{1-d}\ell^{-1}\sup_{\tau^{\prime}\colon|\tau^{\prime}-\tau|\leq L}\int_{\Sigma(x,\tau^{\prime})\cap(\Omega\times\mathbb{R}_{\xi_{1}})}\,d\xi:d_{\xi}a.

The integral in the right-hand expression does not exceed Cσρd2C\sigma\rho^{d-2}. Finally, recall that TT\asymp\ell.

Taking ρ=max((h1δ1)12,)\rho=\max((h^{1-\delta^{\prime}}\ell^{-1})^{\frac{1}{2}},\,\ell) to satisfy (3.15) and (3.17), and thus σ=ρ\sigma=\rho we get

Corollary 5.15.

The following estimates hold as |x1|+|y1|Cσ|x_{1}|+|y_{1}|\leq C\sigma\ell

(3.25) eh1,𝖳(x,y,τ)=O(h1d)for {h12+δd4,h12δd=3,h13δd=2.\displaystyle e^{1,\mathsf{T}}_{h}(x,y,\tau)=O(h^{1-d})\ \ \text{for\ \ }\left\{\begin{aligned} &\ell\geq h^{\frac{1}{2}+\delta}&&d\geq 4,\\ &\ell\geq h^{\frac{1}{2}-\delta}&&d=3,\\ &\ell\geq h^{\frac{1}{3}-\delta}&&d=2.\end{aligned}\right.

Remark 5.16.
  1. (i)

    Then σh13δ\sigma\geq h^{\frac{1}{3}-\delta^{\prime}}.

  2. (ii)

    The similar arguments work under assumption |xy|ϵ|x1|+|y1||x^{\prime}-y^{\prime}|\geq\epsilon|x_{1}|+|y_{1}| instead of |x1|+|y1|Cσ|x_{1}|+|y_{1}|\leq C\sigma\ell.

  3. (iii)

    The similar but simpler arguments work for eh0,𝖳(x,y,τ)e^{0,\mathsf{T}}_{h}(x,y,\tau) as ρ20h1δ\rho^{2}\ell^{0}\geq h^{1-\delta}.

Chapter 4 Successive approximations

6 Successive approximations inside domain

6.1 Standard successive approximations

In this Section we are going to apply method of successive approximations to derive asymptotics of eh(x,y,τ)e_{h}(x,y,\tau) near boundary. However we start from the partial proof of Theorem 1.1 by this method away from the boundary. According to [Ivr2], Section LABEL:monsterbook-sect-5-3 (which is our standard reference here), we consider problem for propagator uh(x,y,t)u_{h}(x,y,t)

(hDtA)uh=0,\displaystyle(hD_{t}-A)u_{h}=0, u|t=0=δ(xy)\displaystyle u|_{t=0}=\updelta(x-y)

with A=A(x,hDx,h)A=A(x,hD_{x},h) and rewrite for uh±(x,y,t)uh(x,y,t)θ(±t)u^{\pm}_{h}(x,y,t)\coloneqq u_{h}(x,y,t)\uptheta(\pm t). We have equation

(4.1) (hDtA)uh±=ihδ(xy)δ(t)\displaystyle(hD_{t}-A)u_{h}^{\pm}=\mp ih\updelta(x-y)\updelta(t)

which we are going to solve by the successive approximations with unperturbed operator A¯=a(y,hDx)\bar{A}=a(y,hD_{x}). Then (4.1) yields the equality

(4.2) (hDtA¯)uh±=ihδ(xy)δ(t)+Ruh±\displaystyle(hD_{t}-\bar{A})u_{h}^{\pm}=\mp ih\updelta(x-y)\updelta(t)+Ru_{h}^{\pm}
and hence
(4.3) uh±=ihG¯±δ(xy)δ(t)+G¯±Ruh±,\displaystyle u_{h}^{\pm}=\mp ih\bar{G}^{\pm}\updelta(x-y)\updelta(t)+\bar{G}^{\pm}Ru_{h}^{\pm},

where G¯±\bar{G}^{\pm} and G±G^{\pm} are parametrices of the problems

(4.4) (hDtA¯)v=f,supp(v){±(tt0)0}\displaystyle(hD_{t}-\bar{A})v=f,\qquad\operatorname{supp}(v)\subset\{\pm(t-t_{0})\geq 0\}
and
(4.5) (hDtA)v=f,supp(v){±(tt0)0}\displaystyle(hD_{t}-{A})v=f,\qquad\operatorname{supp}(v)\subset\{\pm(t-t_{0})\geq 0\}

respectively with supp(f){±(tt0)0}\operatorname{supp}(f)\subset\{\pm(t-t_{0})\geq 0\} for some t0t_{0}\in\mathbb{R}; R=AA¯R=A-\bar{A}.

Moreover, equation (4.1) yields that

(4.6) uh±=ihG±δ(xy)δ(t).u_{h}^{\pm}=\mp ihG^{\pm}\updelta(x-y)\updelta(t).

Iterating (4.3) NN times and then substituting (4.6) we obtain the equality

(4.7) uh±=ih0nN1(G¯±R)nG¯±δ(xy)δ(t)ih(G¯±R)nG±δ(xy)δ(t).u_{h}^{\pm}=\mp ih\sum_{0\leq n\leq N-1}(\bar{G}^{\pm}R)^{n}\bar{G}^{\pm}\updelta(x-y)\updelta(t)\mp\\ ih(\bar{G}^{\pm}R)^{n}G^{\pm}\updelta(x-y)\updelta(t).

Finally, we apply Qyt=tQ(y,hDy){}^{t}\!Q_{y}=\,^{t}\!Q(y,hD_{y}) to the right of (4.7), where QQ is an operator with compactly supported symbol, equal 11 as a(y,ξ)τ+ϵa(y,\xi)\leq\tau+\epsilon. After this cut-off, according to [Ivr2], Section LABEL:monsterbook-sect-5-3, norms of the terms in (4.7) in the strip {(x,t):|t|T}\{(x,t)\colon|t|\leq T\} do not exceed ChMnT2nCh^{-M-n}T^{2n} and as Th12+δT\leq h^{\frac{1}{2}+\delta} we can ignore the remainder term. Since we need to consider Tc(x,y)T\leq c\ell(x,y) in the end we arrive to

(4.8) max(c,h1δ)Th12+δ.\displaystyle\max(c\ell,\,h^{1-\delta})\leq T\leq h^{\frac{1}{2}+\delta}.

Then as it was shown (see [Ivr2], simplified (LABEL:monsterbook-4-3-33))

(4.9) Fth1τ(χ¯Tuh±(x,y,t)tQy)i0nN(2π)d1hd+neih1xy,ξFn(y,ξ,τ)[q2(y,ξ,h)]𝑑ξ,±Imτ<0,F_{t\to h^{-1}\tau}\Bigl{(}\bar{\chi}_{T}u_{h}^{\pm}(x,y,t)\,^{t}\!Q_{y}\Bigr{)}\\ \equiv\mp i\sum_{0\leq n\leq N}(2\pi)^{-d-1}h^{-d+n}\int e^{ih^{-1}\langle x-y,\xi\rangle}F_{n}(y,\xi,\tau)[q_{2}(y,\xi,h)]\,d\xi,\quad\pm\operatorname{Im}\tau<0,

where Fn=β:|β|nFn,βξβF_{n}=\sum_{\beta\colon|\beta|\leq n}F_{n,\beta}\partial_{\xi}^{\beta} are differential operators applied to q2q_{2}, with coefficients FnβF_{n\beta} holomorphic with respect to ξ,τ\xi,\tau as Imτ0\operatorname{Im}\tau\neq 0, and have poles as τ\tau\in\mathbb{R}; denominators are (τa(y,ξ))2n+1+|β|(\tau-a(y,\xi))^{2n+1+|\beta|}. Then

(4.10) Fth1τ(χ¯Tuh(x,y,t))0nN(2π)d1hd+neih1xy,ξn(y,ξ,τ)𝑑ξ,τF_{t\to h^{-1}\tau}\Bigl{(}\bar{\chi}_{T}u_{h}(x,y,t)\Bigr{)}\\ \equiv\sum_{0\leq n\leq N}(2\pi)^{-d-1}h^{-d+n}\int e^{ih^{-1}\langle x-y,\xi\rangle}\mathcal{F}_{n}(y,\xi,\tau)\,d\xi,\qquad\tau\in\mathbb{R}

where n(y,ξ,τ)=i(Fn0(y,ξ,τi0)Fn0(y,ξ,τ+i0))\mathcal{F}_{n}(y,\xi,\tau)=i\bigl{(}F_{n0}(y,\xi,\tau-i0)-F_{n0}(y,\xi,\tau+i0)\bigr{)}. Then the principal term in eT,h𝖳(x,y,τ)e^{\mathsf{T}}_{T,h}(x,y,\tau) can be rewritten as eh𝖶(x,y,τ)e^{\mathsf{W}}_{h}(x,y,\tau) while all other terms as

(4.11) hd+n|α|α:|α|2n1(xy)αΣ(y,τ)Wn,α(y,ξ)eih1xy,ξ𝑑ξ\displaystyle h^{-d+n-|\alpha|}\sum_{\alpha\colon|\alpha|\leq 2n-1}(x-y)^{\alpha}\int_{\Sigma(y,\tau)}W_{n,\alpha}(y,\xi)e^{ih^{-1}\langle x-y,\xi\rangle}\,d\xi

with n1n\geq 1 and smooth Wn,αW_{n,\alpha}.

Indeed, taking ξ\xi-partition and using ξ\xi-microhyperbolicity, we can rewrite on each element n\mathcal{F}_{n} as a sum of Wn,α(y,ξ)ξαδ(τa(y,ξ)){W^{\prime}_{n,\alpha}(y,\xi)\partial_{\xi}^{\alpha}\updelta^{\prime}(\tau-a(y,\xi))}. Then, integrating by parts we will get (4.11) for terms in eT,h𝖳(x,y,τ)e^{\mathsf{T}}_{T,h}(x,y,\tau).

Now due to strong convexity and stationary phase principle these terms do not exceed

(4.12) Chd+n|α|(h/)d12|α|Chd+12+n|α|d12+|α|,\displaystyle Ch^{-d+n-|\alpha|}({h}/{\ell})^{\frac{d-1}{2}}\ell^{|\alpha|}\leq Ch^{-\frac{d+1}{2}+n-|\alpha|}\ell^{-\frac{d-1}{2}+|\alpha|},

which is O(h1d)O(h^{1-d}) as d3d\geq 3 and O(h3212)O(h^{-\frac{3}{2}}\ell^{\frac{1}{2}}) as d=2d=2. Recall that h12+δ\ell\leq h^{\frac{1}{2}+\delta}.

6.2 Improved successive approximations

To improve this approach we need to freeze symbol at point w=12(x+y)w=\frac{1}{2}(x+y) and to do this let us observe that

(hDtA(x,hDx,h))uh(x,y,t)=0,\displaystyle\bigl{(}hD_{t}-A(x,hD_{x},h)\bigr{)}u_{h}(x,y,t)=0,
(hDtA(y,hDy,h))uh(x,y,t)=0,\displaystyle\bigl{(}hD_{t}-A(y,-hD_{y},h)\bigr{)}u_{h}(x,y,t)=0, uh|t=0=δ(xy)\displaystyle u_{h}|_{t=0}=\updelta(x-y)

and therefore denoting vh(w,z,t)=uh(w+12z,w12z,t)v_{h}(w,z,t)=u_{h}(w+\frac{1}{2}z,w-\frac{1}{2}z,t) we have

(4.13) (hDt𝔄(w,z,hDz,hDw,h))vh(w,z,t)=0,vh|t=0=δ(z)\displaystyle\bigl{(}hD_{t}-\mathfrak{A}(w,z,hD_{z},hD_{w},h)\bigr{)}v_{h}(w,z,t)=0,\qquad v_{h}|_{t=0}=\updelta(z)
with
(4.14) 𝔄=12(A(w+12z,hDz+12hDw,h)+A(w12z,hDz12hDw,h)).\displaystyle\mathfrak{A}=\frac{1}{2}\bigl{(}A(w+\frac{1}{2}z,hD_{z}+\frac{1}{2}hD_{w},h)+A(w-\frac{1}{2}z,hD_{z}-\frac{1}{2}hD_{w},h)\bigr{)}.

Taking as unperturbed operator

(4.15) 𝔄¯=A(w,hDz,h)\displaystyle\bar{\mathfrak{A}}=A(w,hD_{z},h)

we get R=𝔄𝔄¯R=\mathfrak{A}-\bar{\mathfrak{A}} a a sum of zα(hw)βz^{\alpha}(h\partial_{w})^{\beta} with |α|+|β|2|\alpha|+|\beta|\geq 2.

With this modification we can write formula, similar to (4.7). Due to the propagation results we can add to RR factor χ¯cT(hw)\bar{\chi}_{cT}(h\partial_{w}) and then each new term in this formula adds an extra factor C(T2+h)Th1C(T^{2}+h)Th^{-1} instead of CT2h1CT^{2}h^{-1} and we can replace (4.8) by a weaker condition

(4.16) max(c,h1δ)Th13+δ.\displaystyle\max(c\ell,\,h^{1-\delta})\leq T\leq h^{\frac{1}{3}+\delta}.

Calculations, similar to those in the standard method lead us to the main term e𝖶(x,y,τ)e^{\mathsf{W}}(x,y,\tau) and to (4.11) replaced by

(4.17) hd+n|α|α:2|α|3n1(xy)αΣ(w,τ)Wn,α(w,ξ)eih1z,ξ𝑑ξ.\displaystyle h^{-d+n-|\alpha|}\sum_{\alpha\colon 2|\alpha|\leq 3n-1}(x-y)^{\alpha}\int_{\Sigma(w,\tau)}W_{n,\alpha}(w,\xi)e^{ih^{-1}\langle z,\xi\rangle}\,d\xi.

and to the same estimate (4.12) albeit with a restrictions |α|(3n1)/2|\alpha|\leq(3n-1)/2, h13\ell\leq h^{\frac{1}{3}}, and we get O(h1d)O(h^{1-d}) as d1d\geq 1.

7 Successive approximations near boundary

7.1 Standard successive approximations

We know (see [Ivr2], Section LABEL:monsterbook-sect-7-2) that uh0(x,y,t)u_{h}^{0}(x,y,t) and then uh1(x,y,t)u_{h}^{1}(x,y,t) for |t|T=h12+δ|t|\leq T=h^{\frac{1}{2}+\delta}, x1+y1C0Tx_{1}+y_{1}\leq C_{0}T could be constructed by the method of successive approximations with unperturbed operator A¯=a(0,y;hDx)\bar{A}=a(0,y^{\prime};hD_{x}) (so the principal part of AA is frozen at point (0,y)(0,y^{\prime})): it leads us to the expression for eh𝖳(x,y,τ)e^{\mathsf{T}}_{h}(x,y,\tau).

In this case unperturbed term would be as for A¯\bar{A} in the half-space and it will not exceed Chd(h/)d+1/2Ch^{-d}(h/\ell)^{{d+1}/2} and and one can prove that the perturbed term would acquire factor T2/hT^{2}/h with TT\asymp\ell and it does not exceed (4.12) (with redefined (x,y)\ell(x,y) now) and it is O(h1d)O(h^{1-d}) as d3d\geq 3 and Ch3212Ch^{-\frac{3}{2}}\ell^{\frac{1}{2}} as d=2d=2.

Since we assumed that the boundary conditions are either Dirichlet or Neumann, we arrive to the expression (1.3) for eh1,𝖳(x,y,τ)e^{1,\mathsf{T}}_{h}(x,y,\tau) as described in Theorem 1.2.

7.2 Improved successive approximations

Let us improve this construction in the same manner as we did inside domain. Consider problem for uh1u^{1}_{h}:

(4.18) (hDtA)uh1=0,uh1|t=0=0,\displaystyle(hD_{t}-A)u_{h}^{1}=0,\qquad u^{1}_{h}|_{t=0}=0,
(4.19) ðBuh1=ðBuh0|xX,\displaystyle\eth Bu_{h}^{1}=-\eth Bu_{h}^{0}|_{x\in\partial X},

where uh0(x,y,t)u^{0}_{h}(x,y,t) satisfies the same equation in the “whole space” and ð\eth is an operator restriction to Xx\partial X\ni x. Then

(4.20) (hDtA)uh1,±=0,\displaystyle(hD_{t}-A)u_{h}^{1,\pm}=0,
(4.21) ðBuh1±=ðBuh0±\displaystyle\eth Bu_{h}^{1\,\pm}=-\eth Bu_{h}^{0\,\pm}

where again uj±=θ(±t)uhj(x,y,t)u^{j\,\pm}=\uptheta(\pm t)u^{j}_{h}(x,y,t) and

(4.22) (hDtA¯)uh1±=Ru1±,\displaystyle(hD_{t}-\bar{A})u_{h}^{1\,\pm}=Ru^{1\,\pm},
(4.23) ðBuh1±=Buh0±,\displaystyle\eth Bu_{h}^{1\,\pm}=-Bu_{h}^{0\,\pm},

where uh0,±u_{h}^{0,\pm} satisfies (4.2); recall that B=IB=I or B=D1B=D_{1} for Dirichlet and Neumann boundary conditions correspondingly4)4)4) For more general boundary conditions we would need to replace BB by B¯+R1\bar{B}+R_{1}. Then like in [Ivr2], Section LABEL:monsterbook-sect-7-2

(4.24) uh1±=G¯±ðBuh0±+G¯±Ruh1±,\displaystyle u_{h}^{1\,\pm}=-\bar{G}^{\prime\,\pm}\eth Bu_{h}^{0\,\pm}+\bar{G}^{\pm}Ru_{h}^{1\,\pm},

where G¯±\bar{G}^{\pm} are parametrices for the problems

(4.25) (hDtA¯)v=f,\displaystyle(hD_{t}-\bar{A})v=f, ðBv=0,\displaystyle\eth Bv=0, supp(v){±(tt0)0}\displaystyle\operatorname{supp}(v)\subset\{\pm(t-t_{0})\geq 0\}
and G¯±\bar{G}^{\prime\,\pm} are parametrices for the problems
(4.26) (hDtA¯)v=0,\displaystyle(hD_{t}-\bar{A})v=0, ðBv=f,\displaystyle\eth Bv=f, supp(v){±(tt0)0}\displaystyle\operatorname{supp}(v)\subset\{\pm(t-t_{0})\geq 0\}

respectively with supp(f){±(tt0)0}\operatorname{supp}(f)\subset\{\pm(t-t_{0})\geq 0\} for some t0t_{0}\in\mathbb{R}; R=AA¯R=A-\bar{A}. Moreover,

(4.27) uh1±=G±ðBuh0±\displaystyle u_{h}^{1\,\pm}=-G^{\prime\,\pm}\eth Bu_{h}^{0\,\pm}

where G±{G}^{\pm} and G±{G}^{\prime\,\pm} are parametrices for the problems (4.25) and (4.26) but for operator AA. Iterating (4.24) NN times and then substituting (4.27) we arrive to formula similar to (4.7)

(4.28) uh1±=0nN1(G¯±R)nG¯±ðBuh0±(G¯±R)NG±ðBuh0±.\displaystyle u_{h}^{1\,\pm}=-\sum_{0\leq n\leq N-1}(\bar{G}^{\pm}R)^{n}\bar{G}^{\prime\,\pm}\eth Bu_{h}^{0\,\pm}-(\bar{G}^{\pm}R)^{N}G^{\prime\,\pm}\eth Bu_{h}^{0\,\pm}.

What is more, we plug uh0±u_{h}^{0\,\pm} given by (4.7) with G¯±\bar{G}^{\pm} and G±G^{\pm} replaced by G¯0±\bar{G}^{0\,\pm} and G0±G^{0\,\pm} which are parametrices for (4.4) and (4.5) in the “whole space”.

Finally, we apply Q1t=tQ(y1,z,hDz){}^{t}\!Q_{1}=\,^{t}\!Q(y_{1},z^{\prime},hD^{\prime}_{z}) to the right of (4.27) and (4.7), where Q1Q_{1} is an operator with the symbol, supported in

(4.29) ΩT,ρ,σ{(x1,z,ζ):|z|T,|ζξ¯|ρ,|x1|σT}\displaystyle\Omega_{T,\rho,\sigma}\coloneqq\{(x_{1},z^{\prime},\zeta)\colon|z^{\prime}|\leq T,\ |\zeta-\bar{\xi}^{\prime}|\leq\rho,\ |x_{1}|\leq\sigma T\}

where |a(0,x¯,0,ξ¯)τ)|cρ2|a(0,\bar{x}^{\prime},0,\bar{\xi}^{\prime})-\tau)|\leq c\rho^{2}. Then due to Proposition 4.4, under assumptions (3.10) and (4.4)1,2\textup{(\ref*{eqn-3.10})}_{1,2} we can insert Q2(x1,hDz)Q_{2}(x_{1},hD^{\prime}_{z}) to the right of each copy of RR with symbol supported in Ω3T,3ρ,3σ\Omega_{3T,3\rho,3\sigma} and equal 11 in Ω2T,2ρ,2σ\Omega_{2T,2\rho,2\sigma}.

So far we we followed exactly [Ivr2], Section LABEL:monsterbook-sect-7-2, except there we had ρ=c\rho=c. However now instead of A¯=a(0,y;hDx)\bar{A}=a(0,y^{\prime};hD_{x}) we take

(4.30) A¯=a(w,hDx),w=(0,w),w=12(x+y).\displaystyle\bar{A}=a(w,hD_{x}),\qquad w=(0,w^{\prime}),\qquad w^{\prime}=\frac{1}{2}(x^{\prime}+y^{\prime}).

Then, the norm of G¯±RQ2\bar{G}^{\pm}RQ_{2} does not exceed C(T2+σT+h)Th1C(T^{2}+\sigma T+h)Th^{-1} and we can replace (4.8) by a weaker condition (T2+σT+h)Th1+δ(T^{2}+\sigma T+h)T\leq h^{1+\delta}, which is equivalent to (4.16) plus

(4.31) σ2h1+δ.\displaystyle\sigma\ell^{2}\leq h^{1+\delta}.
Remark 7.1.

Recall that σ\sigma is defined by (3.13) and with ρ=h1δ1\rho=h^{1-\delta}\ell^{-1} condition (4.31) is also equivalent to (4.16). However, for us it is more important which value of the reflection angle is allowed in this method.

Thus we arrive to the following statement:

Proposition 7.2.

Let Q=Q(x1,z,hDz)Q=Q(x_{1},z^{\prime},hD^{\prime}_{z}) be an operator with the symbol q(x1,z,ζ)q(x_{1},z^{\prime},\zeta^{\prime}) supported in ΩT,ρ,σ\Omega_{T,\rho,\sigma} where condition (3.10) is fulfilled and let conditions (4.16) and (4.31) be also fulfilled.

Then we can skip remainder terms in both (4.27) and modified (4.7), which are O(hs)O(h^{s}), leaving us only with G¯±\bar{G}^{\pm}, G¯±\bar{G}^{\prime\,\pm} and G¯0±\bar{G}^{0\,\pm}.

Thus (4.28) without the last term becomes an asymptotic series. Now let us calculate eh1,𝖳(x,y,τ)e_{h}^{1,\mathsf{T}}(x,y,\tau) under these assumptions.

Proposition 7.3.

In the framework of Proposition 7.2

(4.32) Fth1τ(χ¯T(t)uh1±Q1t)n0hd+nTχ¯^((ττ)Th)𝑑τ𝑑ξγ+𝑑ξ1γ𝑑η1×eih1(x1ξ1y1η1+xy,ξ)Fn(w,ξ,ξ1,η1,τ)[q(w,ξ)]F_{t\to h^{-1}\tau}\Bigl{(}\bar{\chi}_{T}(t)u_{h}^{1\,\pm}\,{}^{t}\!Q_{1}\Bigr{)}\\ \sim\sum_{n\geq 0}h^{-d+n}\int T\hat{\bar{\chi}}\bigl{(}\frac{(\tau-\tau^{\prime})T}{h}\bigr{)}\,d\tau^{\prime}\int d\xi^{\prime}\int_{\gamma_{+}}d\xi_{1}\int_{\gamma_{-}}d\eta_{1}\\ \times e^{ih^{-1}(x_{1}\xi_{1}-y_{1}\eta_{1}+\langle x^{\prime}-y^{\prime},\xi^{\prime}\rangle)}F_{n}(w,\xi^{\prime},\xi_{1},\eta_{1},\tau^{\prime})[q(w,\xi^{\prime})]

for τ\tau\in\mathbb{C}_{\mp}, where γ±\gamma_{\pm} are closed contours in ±\mathbb{C}_{\pm} passing once around the poles of (τa(z,ξ1))1\bigl{(}\tau-a(z,\xi_{1})\bigr{)}^{-1} lying in ±\mathbb{C}_{\pm} and not passing around the poles lying in \mathbb{C}_{\mp}; for real zz and Imτ0\operatorname{Im}\tau\neq 0 there is no real pole, and FnF_{n} are sums of the terms

(4.33) ξ1jη1kWn,l,m,p,j,k(w,ξ,τ)×(τa(w,ξ1,ξ))l(τa(w,η1,ξ))m(τb(w,ξ))p\xi_{1}^{j}\eta_{1}^{k}W_{n,l,m,p,j,k}(w,\xi^{\prime},\tau)\\ \times\bigl{(}\tau-a(w,\xi_{1},\xi^{\prime})\bigr{)}^{-l}\bigl{(}\tau-a(w,\eta_{1},\xi^{\prime})\bigr{)}^{-m}\bigl{(}\tau-b(w,\xi^{\prime})\bigr{)}^{-p}

with

(4.34) b(w,ξ)=a(w,0,ξ),\displaystyle b(w,\xi^{\prime})=a(w,0,\xi^{\prime}),
(4.35) l1,m1, 2(l+m+p)3n+j+k+3,jl,km\displaystyle l\geq 1,\;m\geq 1,\;2(l+m+p)\leq 3n+j+k+3,\;j\leq l,\ k\leq m

and uniformly smooth Wn,l,m,p,j,kW_{n,l,m,p,j,k}.

Proof 7.4.

Proof follows the proof of Theorem LABEL:monsterbook-thm-7-2-13 of [Ivr2] with the some modifications, mainly introduced in Subsection 6.2. We rewrote AA as 𝔄\mathfrak{A} by (4.14) and defined 𝔄¯\bar{\mathfrak{A}} by (4.15) with the exception that ww^{\prime} and zz^{\prime} are now (d1)(d-1)-dimensional variables (see (4.30)) and we preserve x1x_{1} and y1y_{1}.

The main part of the perturbation RR is quadratic with respect to z,hDwz^{\prime},hD^{\prime}_{w} and linear with respect to x1x_{1}. In comparison with (LABEL:monsterbook-7-2-60) of [Ivr2] formula (4.33) is simpler, because we have scalar factors here and products collapse into powers. On the other hand, instead of claiming that WW of that proof are holomorphic satisfying (LABEL:monsterbook-7-2-38) of [Ivr2] here we do it in much explicit way since now we need to follow very carefully what are relations between j,l,l,m,pj,l,l,m,p and nn.

We have operators a(w,hD1,hDz,hDw)a(w,hD_{1},hD^{\prime}_{z},hD^{\prime}_{w}) with symbols which do not depend on xx. Let us make hh-Fourier transform by xx^{\prime} and tt.

Consider first uh0(x,y,t)u^{0}_{h}(x,y,t). We claim that

(4.36) Fth1τ(χ¯T(t)ð(hDx1)ruh0±Q1t)n0h1d+nTχ¯^((ττ)Th)𝑑τ𝑑ξγ𝑑η1×eih1(y1η1+xy,ξ)Fn0(w,ξ,η1,τ)[q(w,ξ)]F_{t\to h^{-1}\tau}\Bigl{(}\bar{\chi}_{T}(t)\eth(hD_{x_{1}})^{r}u_{h}^{0\,\pm}\,{}^{t}\!Q_{1}\Bigr{)}\\ \sim\sum_{n\geq 0}h^{1-d+n}\int T\hat{\bar{\chi}}\bigl{(}\frac{(\tau-\tau^{\prime})T}{h}\bigr{)}\,d\tau^{\prime}\int d\xi^{\prime}\int_{\gamma_{-}}d\eta_{1}\\ \times e^{ih^{-1}(-y_{1}\eta_{1}+\langle x^{\prime}-y^{\prime},\xi^{\prime}\rangle)}F^{0}_{n}(w,\xi^{\prime},\eta_{1},\tau^{\prime})[q(w,\xi^{\prime})]

for τ\tau\in\mathbb{C}_{\mp}, r=1r=1 and Fn0F^{0}_{n} are sums of the terms

(4.37) η1kWn,m,j,k0(w,ξ,τ)(τa(w,ξ1,ξ))l(τa(w,η1,ξ))m\displaystyle\eta_{1}^{k}W^{0}_{n,m,j,k}(w,\xi^{\prime},\tau)\bigl{(}\tau-a(w,\xi_{1},\xi^{\prime})\bigr{)}^{-l}\bigl{(}\tau-a(w,\eta_{1},\xi^{\prime})\bigr{)}^{-m}

with 2m3n+k+2r2m\leq 3n+k+2-r and k<mk<m with the only exception n=0n=0, m=1m=1, k=r=1k=r=1.

  1. (a)

    Like in that proof of [Ivr2], each factor zrz_{r} and hDwrhD_{w_{r}} with r=2,,dr=2,\ldots,d is moved to the right towards δ(z)\updelta(z^{\prime}), using commutator relations

    (4.38) zrG¯0±=G¯0±zrG¯0±[𝔄¯,zr]G¯0±\displaystyle z_{r}\bar{G}^{0\,\pm}=\bar{G}^{0\,\pm}z_{r}-\bar{G}^{0\,\pm}[\bar{\mathfrak{A}},z_{r}]\bar{G}^{0\,\pm}

    which also hold for hDwrhD_{w_{r}}. So, each such commuting adds factor hh and increases either mm or pp by 11. However, each pair of those factors are accompanied by G¯0±\bar{G}^{0\,\pm}, and therefore in this process increment of the power of hh by 11 is “paid” by increment of mm with the factor 3/23/25)5)5) Compare with the standard method, when the factor was 22.. On the other hand, commuting with Q1t{}^{t}\!Q_{1}, each increment power of hh by 11 is “paid” by increment of (m+p)(m+p) with the factor 11.

  2. (b)

    Also recall that each factor x1x_{1} is moved to the left, towards ðB\eth B, also using (4.38) but for x1x_{1}. In this commutator factor hη1h\eta_{1} is also gained and mm is increased by 11 and also G¯0±\bar{G}^{0\,\pm} is applied so in the end mm is increased by 22. When x1x_{1} reaches ðB\eth B the term disappears for Dirichlet boundary condition, or this factor cancels with B=hx1B=h\partial_{x_{1}} for Neumann boundary condition (and factor hh is gained). process each increment power of hη1h\eta_{1} by 11 is “paid” by increment of mm with the factor 22. Therefore (4.36)–(4.37) has been proven.

  3. (c)

    Then for

    (4.39) Fth1τ(χ¯T(t)G¯±ðBuh0±Q1t)\displaystyle F_{t\to h^{-1}\tau}\Bigl{(}\bar{\chi}_{T}(t)\bar{G}^{\prime\,\pm}\eth Bu_{h}^{0\,\pm}\,{}^{t}\!Q_{1}\Bigr{)}

    formula (4.33) holds with l=1l=1, p=0p=0 and j=1,0j=1,0 for Dirichlet and Neumann boundary conditions respectively.

  4. (d)

    Again each factor zrz_{r} and hDwrhD_{w_{r}} with r=2,,dr=2,\ldots,d is moved to the right towards δ(z)\updelta(z^{\prime}), using commutator relations (4.38) for G¯±\bar{G}^{\pm} rather than for G¯0±\bar{G}^{0\,\pm} and also

    (4.40) zrG¯±=G¯±zrG¯±[𝔄¯,zr]G¯±\displaystyle z_{r}\bar{G}^{\prime\,\pm}=\bar{G}^{\prime\,\pm}z_{r}-\bar{G}^{\pm}[\bar{\mathfrak{A}},z_{r}]\bar{G}^{\prime\,\pm}

    which holds also for hDwrhD_{w_{r}}. Observe that that pp can increase but again as in (a) increment of the power of hh by 11 is “paid” by increment of m+pm+p with the factor 3/23/26)6)6) Indeed, it is sufficient to consider boundary value problem for ODE (τD2)v=f(\tau-D^{2})v=f with f(x)=γ+ξj(τξ2)leixξ𝑑ξf(x)=\int_{\gamma_{+}}\xi^{j}(\tau-\xi^{2})^{-l}e^{ix\xi}\,d\xi. One can see easily that v(x)=γ+(ξj(τξ2)l1κξjτ(jj)/2l(τξ2)1)eixξdξ\displaystyle v(x)=\int_{\gamma_{+}}\Bigl{(}\xi^{j}(\tau-\xi^{2})^{-l-1}-\kappa\xi^{j^{\prime}}\tau^{(j-j^{\prime})/2-l}(\tau-\xi^{2})^{-1}\Bigl{)}e^{ix\xi}\,d\xi satisfies this equation and v(0)=0v(0)=0 for some constant κ=κ𝖣jl\kappa=\kappa_{\mathsf{D}jl} and jjmod2j^{\prime}\equiv j\mod 2 and j=0,1j^{\prime}=0,1. Also v(0)=0v^{\prime}(0)=0 with some other constant κ=κ𝖭jl\kappa=\kappa_{\mathsf{N}jl}..

  5. (e)

    Also recall that each factor x1x_{1} which is to the left from G¯±ðB\bar{G}^{\prime\,\pm}\eth B is moved to the right, towards G¯±ðB\bar{G}^{\prime\,\pm}\eth B but instead of (4.38) we use

    (4.41) x1G¯±=G¯𝖣±x1G¯𝖣±[𝔄¯,x1]G¯±,\displaystyle x_{1}\bar{G}_{*}^{\pm}\ =\bar{G}_{\mathsf{D}}^{\pm}x_{1}-\bar{G}_{\mathsf{D}}^{\pm}[\bar{\mathfrak{A}},x_{1}]\bar{G}_{*}^{\pm},
    (4.42) x1G¯±=G¯𝖣±[𝔄¯,x1]G¯±\displaystyle x_{1}\bar{G}_{*}^{\prime\,\pm}=-\bar{G}_{\mathsf{D}}^{\pm}[\bar{\mathfrak{A}},x_{1}]\bar{G}_{*}^{\prime\,\pm}

    with =𝖣,𝖭*=\mathsf{D},\mathsf{N} (in particular, for G¯𝖣\bar{G}_{\mathsf{D}} (4.38) holds).

    Again, similarly to (c) either l+pl+p is increased by 22 and jj does not change, or jj is decreased by 11 and l+pl+p is increased by 11.

  6. (f)

    Then we have n=n+j+kn=n^{\prime}+j+k with l+p=l+2jl+p=l^{\prime}+2j, m=m+2km=m^{\prime}+2k and l+m32n+2l^{\prime}+m^{\prime}\leq\frac{3}{2}n^{\prime}+2. Those are “the worst case scenarios” when we consider the main part of perturbation RR, quadratic by (xy)(x^{\prime}-y^{\prime}) and linear by x1x_{1}. For the rest of perturbation increments of l+p,ml+p,m are relatively smaller. In the end we arrive to (4.35).

  7. (g)

    One can see easily that for the main term we have l=m=1l=m=1, p=0p=0 and j+k=1j+k=1 with j=0,1j=0,1 for Dirichlet and Neumann problems respectively and W=𝖼𝗈𝗇𝗌𝗍W_{*}=\mathsf{const}.

Proposition 7.5.

In the framework of Proposition 7.2

(4.43) Fth1τ(χ¯T(t)uh1±Q1t)n0,j0,k0hd+nTχ¯^((ττ)Th)𝑑τ𝑑ξγ+𝑑ξ1×eih1((x1+y1)ξ1+xy,ξ)Fn,j,k(w,ξ,ξ1,τ)[q(w,ξ)]F_{t\to h^{-1}\tau}\Bigl{(}\bar{\chi}_{T}(t)u_{h}^{1\,\pm}\,{}^{t}\!Q_{1}\Bigr{)}\\ \sim\sum_{n\geq 0,j\geq 0,k\geq 0}h^{-d+n}\int T\hat{\bar{\chi}}\bigl{(}\frac{(\tau-\tau^{\prime})T}{h}\bigr{)}\,d\tau^{\prime}\int d\xi^{\prime}\int_{\gamma_{+}}d\xi_{1}\\ \times e^{ih^{-1}((x_{1}+y_{1})\xi_{1}+\langle x^{\prime}-y^{\prime},\xi^{\prime}\rangle)}F_{n,j,k}(w,\xi^{\prime},\xi_{1},\tau^{\prime})[q(w,\xi^{\prime})]

for τ\tau\in\mathbb{C}_{\mp} where Fn,j,kF_{n,j,k} are sums of the terms

(4.44) ξ1jη1kWn,l,j,k(w,ξ,τ)(τa(y,ξ1,ξ))l\displaystyle\xi_{1}^{j}\eta_{1}^{k}W_{n,l,j,k}(w,\xi^{\prime},\tau)\bigl{(}\tau-a(y^{\prime},\xi_{1},\xi^{\prime})\bigr{)}^{-l}

with 1l32n+j+k+11\leq l\leq\frac{3}{2}n+j+k+1 and uniformly smooth Wn,l,j,kW_{n,l,j,k} and also

(4.45) Fth1τ(χ¯T(t)uh1Q1t)n0,j0,k0hd+nTχ¯^((ττ)Th)𝑑τ𝑑ξ×eih1((x1+y1)ξ1+xy,ξ)n,j,k(w,ξ,ξ1,τ)[q(w,ξ)],F_{t\to h^{-1}\tau}\Bigl{(}\bar{\chi}_{T}(t)u_{h}^{1}\,{}^{t}\!Q_{1}\Bigr{)}\\ \sim\sum_{n\geq 0,j\geq 0,k\geq 0}h^{-d+n}\int T\hat{\bar{\chi}}\bigl{(}\frac{(\tau-\tau^{\prime})T}{h}\bigr{)}\,d\tau^{\prime}\int d\xi\\ \times e^{ih^{-1}((x_{1}+y_{1})\xi_{1}+\langle x^{\prime}-y^{\prime},\xi^{\prime}\rangle)}\mathcal{F}_{n,j,k}(w,\xi^{\prime},\xi_{1},\tau^{\prime})[q(w,\xi^{\prime})],

where n,j,k(w,ξ,ξ1,τ)=Fn,j,k(w,ξ,ξ1,τi0)Fn,j,k(w,ξ,ξ1,τ+i0)\mathcal{F}_{n,j,k}(w,\xi^{\prime},\xi_{1},\tau^{\prime})=F_{n,j,k}(w,\xi^{\prime},\xi_{1},\tau^{\prime}-i0)-F_{n,j,k}(w,\xi^{\prime},\xi_{1},\tau^{\prime}+i0).

Proof 7.6.

(4.43)–(4.44) follows from Proposition 7.3 by calculating residues at zeroes of (τa(x,ξ1,ξ))(\tau-a(x,\xi_{1},\xi^{\prime})) and (τa(x,η1,ξ))(\tau-a(x,\eta_{1},\xi^{\prime})) in (4.33) and (4.43).

Next we observe that in (4.43) we can replace integral along γ+\gamma_{+} by integral along \mathbb{R}. Then (4.45) follows trivially.

Proposition 7.7.

Let ξ\xi-microhyperbolicity condition be fulfilled at energy level τ\tau and let

(4.46) ν(x)+ν(y)σ\displaystyle{\nu}(x)+{\nu}(y)\leq\sigma\ell
with
(4.47) (x,y)h13+δ,σ=h12+δ12.\displaystyle\ell(x,y)\leq h^{\frac{1}{3}+\delta},\quad\sigma=h^{\frac{1}{2}+\delta}\ell^{-\frac{1}{2}}.

Let T=TC0(x,y)T=T^{*}\geq C_{0}\ell(x,y). Then for any T:TTϵT\colon T^{*}\leq T\leq\epsilon

  1. (i)

    As d3d\geq 3 estimate

    (4.48) |eT,h1,𝖳(x,y,τ)[]eh0,𝖶(x,y~,τ)|Ch1d\displaystyle|e^{1,\mathsf{T}}_{T,h}(x,y,\tau)-[\mp]e_{h}^{0,\mathsf{W}}(x,\tilde{y},\tau)|\leq Ch^{1-d}

    with sign [][\mp] for Dirichlet and Neumann boundary conditions respectively.

  2. (ii)

    As d=2d=2 estimate

    (4.49) |eT,h1,𝖳(x,y,τ)[]eh0,𝖶(x,y~,τ)eh,𝖼𝗈𝗋𝗋(x,y,τ)|Ch1\displaystyle|e^{1,\mathsf{T}}_{T,h}(x,y,\tau)-[\mp]e_{h}^{0,\mathsf{W}}(x,\tilde{y},\tau)-e_{h,\mathsf{corr}}(x,y,\tau)|\leq Ch^{-1}

    with eh,𝖼𝗈𝗋𝗋(x,y,τ)e_{h,\mathsf{corr}}(x,y,\tau) given by (4.53) below.

Proof 7.8.
  1. (a)

    Consider first h12δ\ell\leq h^{\frac{1}{2}-\delta}. Let us take ρ=min((h/)12hδ,C2)\rho=\min((h/\ell)^{\frac{1}{2}}h^{-\delta},\,C_{2}). Then ρC0\rho\geq C_{0}\ell, σ=ρ12\sigma=\rho^{\frac{1}{2}} and

    (4.50) σ2h1+δas h37+δ\displaystyle\sigma\ell^{2}\leq h^{1+\delta}\qquad\text{as\ \ }\ell\leq h^{\frac{3}{7}+\delta}

    and the method of successive approximations works. Let us take take ρ\rho-admissible partition of unity by ζ\zeta^{\prime} in ρ\rho-vicinity of Σ(w,τ)\Sigma(w,\tau). In this case ρ\rho\geq\ell and σ=ρ12\sigma=\rho^{\frac{1}{2}}. Moreover, as ρ=C2\rho=C_{2} it covers the whole domain {ζ:b(w,ζ)<2τ}\{\zeta^{\prime}\colon b(w,\zeta^{\prime})<2\tau\} and as ρ<C2\rho<C_{2} and therefore h12δ\ell\geq h^{1-2\delta} and also ρ2h12δ\rho^{2}\ell\geq h^{1-2\delta} we can use the fact that

    Claim 2.

    If ρh1δ\rho\ell\geq h^{1-\delta}, ρC02\rho\geq C_{0}\ell^{2}, (3.10) is fulfilled and Q1Q_{1} is an operator with the symbol equal 0 in ρ\rho-vicinity of Σ(w,τ)\Sigma(w,\tau) and h1δ\ell\geq h^{1-\delta} then

    eh1,𝖳(x,y,τ)tQ1,z=O(hs)\displaystyle e^{1,\mathsf{T}}_{h}(x,y,\tau)\,^{t}\!Q_{1,z}=O(h^{s})

    which follows from Section 3. Therefore we can take cut-off operator Q1=IQ_{1}=I.

    Then the main term of the final expression equals to eh0,𝖶(x,y~,τ)e^{0,\mathsf{W}}_{h}(x,\tilde{y},\tau) and is hd(h1)(d+1)/2\asymp h^{-d}(h\ell^{-1})^{(d+1)/2} and nn-th term does not exceed

    (4.51) Chdδ(h)(d+1)/2×(ρ122h1)n1\displaystyle Ch^{-d-\delta^{\prime}}\bigl{(}\frac{h}{\ell}\bigr{)}^{(d+1)/2}\times(\rho^{\frac{1}{2}}\ell^{2}h^{-1})^{n-1}

    due to the same microlocal arguments as above. One can see easily that expression (4.51) is O(h1d+δ′′)O(h^{1-d+\delta^{\prime\prime}}) as d2d\geq 2 and n3n\geq 3 and h13+δ\ell\leq h^{\frac{1}{3}+\delta}. Furthermore, if we consider terms O(2)O(\ell^{2}) in RR then the corresponding second term does not exceed

    (4.52) Chdδ(h)(d+1)/2×3h1)n1=O(h1d+δ′′).\displaystyle Ch^{-d-\delta^{\prime}}\bigl{(}\frac{h}{\ell}\bigr{)}^{(d+1)/2}\times\ell^{3}h^{-1})^{n-1}=O(h^{1-d+\delta^{\prime\prime}}).

    Therefore we need to consider only the second term in the successive approximations. It is equal to

    (4.53) []eh,𝖼𝗈𝗋𝗋(x,y,τ)12(2πh)dΣ(w,τ)λ(w,ξ)(x1+y1)eih1xy~,ξ𝑑ξ:dξa(w,ξ)[\mp]e_{h,\mathsf{corr}}(x,y,\tau)\\ \coloneqq-\frac{1}{2}(2\pi h)^{-d}\int_{\Sigma(w,\tau)}\lambda(w^{\prime},\xi^{\prime})(x_{1}+y_{1})e^{ih^{-1}\langle x-\tilde{y},\xi\rangle}\,d\xi:d_{\xi}a(w,\xi)

    with

    (4.54) λ(w,ξ)=ax1(x1,w,ξ)|x1=ξ1=0.\displaystyle\lambda(w^{\prime},\xi^{\prime})=a_{x_{1}}(x_{1},w^{\prime},\xi)|_{x_{1}=\xi_{1}=0}.

    Indeed, one can see easily that the second term in the successive approximation for []Fth1τ(χ¯T(t)uh1±(x,y,t))[\mp]F_{t\to h^{-1}\tau}\bigl{(}\bar{\chi}_{T}(t)u^{1\,\pm}_{h}(x,y,t)\bigr{)} is equal to

    ±2h3(2πh)1dγ+×λ(w,ξ)ξ1(τa(x,ξ))3eixy~,ξ𝑑ξ=i2h2(2πh)1dγ+×λ(w,ξ)(x1+y1)(τa(x,ξ))2eixy~,ξImτ<0;\pm 2h^{3}(2\pi h)^{-1-d}\int_{\gamma_{+}\times\mathbb{R}}\lambda(w^{\prime},\xi^{\prime})\xi_{1}(\tau-a(x,\xi))^{-3}e^{i\langle x-\tilde{y},\xi\rangle}\,d\xi\\ =\mp\frac{i}{2}h^{2}(2\pi h)^{-1-d}\int_{\gamma_{+}\times\mathbb{R}}\lambda(w^{\prime},\xi^{\prime})(x_{1}+y_{1})(\tau-a(x,\xi))^{-2}e^{i\langle x-\tilde{y},\xi\rangle}\qquad\mp\operatorname{Im}\tau<0;

    then the second term in decomposition for []Fth1τ(χ¯T(t)uh1(x,y,t))[\mp]F_{t\to h^{-1}\tau}\bigl{(}\bar{\chi}_{T}(t)u^{1}_{h}(x,y,t)\bigr{)} is equal to

    12h(2πh)dλ(w,ξ)(x1+y1)eih1xy~,ξδ(τa(w,ξ))𝑑ξ\displaystyle-\frac{1}{2}h(2\pi h)^{-d}\int\lambda(w^{\prime},\xi^{\prime})(x_{1}+y_{1})e^{ih^{-1}\langle x-\tilde{y},\xi\rangle}\updelta^{\prime}(\tau-a(w,\xi))\,d\xi

    which implies (4.54).

  2. (b)

    Due to stationary phase integral in 4.53) does not exceed Chd(h)(d1)/2Ch^{-d}\bigl{(}\frac{h}{\ell}\bigr{)}^{(d-1)/2}, and expression (4.53) is O(h1d)O(h^{1-d}) as d3d\geq 3, x1+y1σx_{1}+y_{1}\leq\sigma\ell.

    Due to Corollary 5.15 as d3d\geq 3 we need to consider only this case h12δ\ell\leq h^{\frac{1}{2}-\delta}. Therefore Statement (i) has been proven.

  3. (c)

    Let d=2d=2. Then we need to consider also h12δh13+δh^{\frac{1}{2}-\delta}\leq\ell\leq h^{\frac{1}{3}+\delta} and with selected ρ=(h/)12hδ\rho=(h/\ell)^{\frac{1}{2}}h^{-\delta} condition σ2h1+δ\sigma\ell^{2}\leq h^{1+\delta} may fail. We need to consider lesser ρ\rho. Let us pick up ρ=h1δ1\rho=h^{1-\delta}\ell^{-1}. Then ρC02\rho\geq C_{0}\ell^{2} and σ2h1+δ\sigma\ell^{2}\leq h^{1+\delta} as h13δ\ell\leq h^{\frac{1}{3}-\delta}.

    However condition ρC0\rho\geq C_{0}\ell may fail and then we need to take care of (3.10) at each point of Σ(w,τ){ξ1=0}\Sigma(w,\tau)\cap\{\xi_{1}=0\}.

    The good news is that this set consists of two points ξ¯±\bar{\xi}^{\prime\pm} and we do not need ρ2h1δ\rho^{2}\ell\geq h^{1-\delta} because we can deal with each of these two points separately. Therefore we can replace ρ\rho-admissible operator Q1Q_{1} with the symbol, equal 11 in ρ\rho-vicinity of ξ¯±\bar{\xi}^{\prime\pm} by operator with the symbol equal 11 in ϵ\epsilon-vicinity of this point. Observing that the required gauge transformations are ξ2ξ2h1α±z2\xi_{2}\mapsto\xi_{2}-h^{-1}\alpha_{\pm}z_{2}, that are multiplications by eih1α±z22/2e^{ih^{-1}\alpha_{\pm}z_{2}^{2}/2} which do not affect e(x,y,τ)e(x,y,\tau) as z2=12(x2y2)z_{2}=\frac{1}{2}(x_{2}-y_{2}). Therefore again we can take Q2=IQ_{2}=I.

    Then nn-th term does not exceed (4.51) which is O(h1+δ′′)O(h^{-1+\delta^{\prime\prime}}) as n3n\geq 3 in current settings. Again, if we consider O(2)O(\ell^{2}) terms in RR then the corresponding second term does not exceed (4.52). Thus again we are left with expression (4.53) but this time we cannot claim that it is O(h1)O(h^{-1}) even for h12+δ\ell\leq h^{\frac{1}{2}+\delta} and ρ=h1δ1\rho=h^{1-\delta^{\prime}}\ell^{-1}. Therefore Statement (ii) has been proven.

Chapter 5 Geometric optics method

8 Constructing solution

Proposition 8.1.

Let

(5.1) ν(x)+ν(y)C02(x,y)\displaystyle{\nu}(x)+{\nu}(y)\geq C_{0}\ell^{2}(x,y)

and (x,y)ϵ\ell(x,y)\leq\epsilon. Then on the energy level τ\tau there exists a single generalized Hamiltonian billiard of the length cϵ\leq c\epsilon with at least one reflection from X\partial X, from yy to xx and one from xx to yy, and each has exactly one reflection and the incidence angles are (ν(x)+ν(y))1(x,y)\asymp\bigl{(}\nu(x)+\nu(y)\bigr{)}\ell^{-1}(x,y) (so, they are standard Hamiltonian billiards).

Proof 8.2.

The proof is obvious.

Remark 8.3.

Assumption (5.1) is essential. Indeed, if X\partial X is strongly concave7)7)7) With respect to Hamiltonian trajectories. then for ν(x)+ν(y)C02(x,y){\nu}(x)+{\nu}(y)\leq C_{0}\ell^{2}(x,y) there could be multiple reflections and if X\partial X is strongly convex7) then the incidence angle may be 0. In the former case there could be also multiple reflections and multiple billiard rays from xx to yy on the same energy level τ\tau while in the latter case there could be no rays at all.

Proposition 8.4.
  1. (i)

    Let AA be ξ\xi-microhyperbolic on the energy level τ=a(y,θ){\tau=a(y,\theta)} and

    a(x,0)<τϵ.\displaystyle a(x,0)<\tau-\epsilon.

    Let φ0(x,y,θ)\varphi^{0}(x,y,\theta) be a solution of the stationary eikonal equation (2.7) satisfying (2.8) and (2.9). Then

    (5.2) |xαyβθγφ0(x,y,θ)|Cαβγ{Tα=β=0,1|α|+|β|1x,y:|xy|T|\partial_{x}^{\alpha}\partial_{y}^{\beta}\partial_{\theta}^{\gamma}\varphi^{0}(x,y,\theta)|\leq C_{\alpha\beta\gamma}\left\{\begin{aligned} &T&&\alpha=\beta=0,\\ &1&&|\alpha|+|\beta|\geq 1\end{aligned}\right.\\ \forall x,y\colon|x-y|\leq T

    provided TϵT\leq\epsilon with sufficiently small constant ϵ>0\epsilon>0.

  2. (ii)

    Let φ(x,y)\varphi(x,y) be another solution of the same equation satisfying

    (5.3) φ|x1=0=φ0|x1=0,x1φ|x1=0=x1φ0|x1=0.\displaystyle\varphi|_{x_{1}=0}=\varphi^{0}|_{x_{1}=0},\qquad\partial_{x_{1}}\varphi|_{x_{1}=0}=-\partial_{x_{1}}\varphi^{0}|_{x_{1}=0}.
    Assume that
    (5.4) σ=|x1φ|x1=0|C0T.\displaystyle\sigma=|\partial_{x_{1}}\varphi|_{x_{1}=0}|\geq C_{0}T.

    Then as |xy|T|x-y|\leq T the following inequalities hold:

    (5.5) |xαyβθγφ|Cαβγ{Tα=β=0,1α11,σ3α1|α||β||γ|α12.\displaystyle|\partial_{x}^{\alpha}\partial_{y}^{\beta}\partial_{\theta}^{\gamma}\varphi|\leq C_{\alpha\beta\gamma}\left\{\begin{aligned} &T&&\alpha=\beta=0,\\ &1&&\alpha_{1}\leq 1,\\ &\sigma^{3-\alpha_{1}-|\alpha|-|\beta|-|\gamma|}\qquad&&\alpha_{1}\geq 2.\end{aligned}\right.
  3. (iii)

    Furthermore,

    (5.6) |xαyβθγφ|Cαβγσ4α1|α||β||γ|α12,|γ|1.\displaystyle|\partial_{x}^{\alpha}\partial_{y}^{\beta}\partial_{\theta}^{\gamma}\varphi|\leq C_{\alpha\beta\gamma}\sigma^{4-\alpha_{1}-|\alpha|-|\beta|-|\gamma|}\qquad\alpha_{1}\geq 2,\ |\gamma^{\prime}|\geq 1.

Proof 8.5.
  1. (a)

    Consider first φ=φ0\varphi=\varphi^{0}. Then the left-hand expression of (5.2) does not exceed CαβγC_{\alpha\beta\gamma} and we need to consider only α=β=0\alpha=\beta=0. Applying θγ\partial_{\theta}^{\gamma} to eikonal equation we see that ddtθγφ\frac{d}{dt}\partial^{\gamma}_{\theta}\varphi is bounded where here and below

    (5.7) ddt(tka(k)(x,xφ)).\displaystyle\dfrac{d\ }{dt}\coloneqq\bigl{(}\partial_{t}-\sum_{k}a^{(k)}(x,\nabla_{x}\varphi)\bigr{)}.

    Since |θγφ0(x,y,t,θ)|CγT|\partial^{\gamma}_{\theta}\varphi^{0}(x,y,t,\theta)|\leq C_{\gamma}T as t=0t=0 and |xy|T|x-y|\leq T we conclude that the same is true for |t|T|t|\leq T.

  2. (b)

    First, we provide this estimate as x1=0x_{1}=0; from (5.2) and (5.3) we conclude that (5.5) holds as α1=0,1\alpha_{1}=0,1. Consider

    (5.8) φx12=a(y,θ)+b(x,xφ),b(x,xφ)Vj2gjk(φxjVj)(φxkVk).\varphi_{x_{1}}^{2}=a(y,\theta)+b(x,\nabla_{x^{\prime}}\varphi),\\ b(x,\nabla_{x^{\prime}}\varphi)\coloneqq-V-\sum_{j\geq 2}g^{jk}(\varphi_{x_{j}}-V_{j})(\varphi_{x_{k}}-V_{k}).

    Differentiating once by x1x_{1} we get

    (5.9) 2φx1φx1x1=bx1(x,xφ)+k2b(k)(x,x)φxkx1\displaystyle 2\varphi_{x_{1}}\varphi_{x_{1}x_{1}}=b_{x_{1}}(x,\nabla_{x^{\prime}}\varphi)+\sum_{k\geq 2}b^{(k)}(x,\nabla_{x^{\prime}})\varphi_{x_{k}x_{1}}

    which in virtue of (5.5) with α11\alpha_{1}\leq 1 does not exceed CC and therefore

    (5.10) |φx1x1|Cσ1.\displaystyle|\varphi_{x_{1}x_{1}}|\leq C\sigma^{-1}.

    Further, applying to (5.8) xαεyβθγ\partial_{x}^{\alpha-\upvarepsilon}\partial_{y}^{\beta}\partial_{\theta}^{\gamma} with ε=(1,0,,0)\upvarepsilon=(1,0,\ldots,0) and α1=2\alpha_{1}=2 and plugging x1=0x_{1}=0 we see that 2φx1φαβγ2\varphi_{x_{1}}\varphi_{\alpha\beta\gamma}8)8)8) Here and below we use notation φαβγ=x,tαyβθγφ\varphi_{\alpha\beta\gamma}=\partial_{x,t}^{\alpha}\partial_{y}^{\beta}\partial_{\theta}^{\gamma}\varphi. is a linear combination with bounded coefficients of φα1β1γ1\varphi_{\alpha^{1}\beta^{1}\gamma^{1}} with |α1|+|β1|+|γ1|<|α|+|β|+|γ||\alpha^{1}|+|\beta^{1}|+|\gamma^{1}|<|\alpha|+|\beta|+|\gamma| plus bounded terms, and then by induction by |α|+|β|+|γ||\alpha|+|\beta|+|\gamma| we arrive to (5.5) for α1=2\alpha_{1}=2.

    Consider induction by α1\alpha_{1}. Assume that (5.5) is proven for lesser α1\alpha_{1} and for arbitrary α,β,γ\alpha^{\prime},\beta,\gamma.

    Applying to (5.8) x,tαεyβθγ\partial_{x,t}^{\alpha-\upvarepsilon}\partial_{y}^{\beta}\partial_{\theta}^{\gamma} and plugging x1=0x_{1}=0 we get linear combinations with bounded coefficients of products

    (5.11) 1jmφαjβjγj\displaystyle\prod_{1\leq j\leq m}\varphi_{\alpha^{j}\beta^{j}\gamma^{j}}

    where j|αj||α|+m1\sum_{j}|\alpha^{j}|\leq|\alpha|+m-1, jβjβ\sum_{j}\beta^{j}\leq\beta, jγjγ\sum_{j}\gamma^{j}\leq\gamma. These terms appear when

    1. (A)

      We differentiate b(x,xφ)b(x,\nabla_{x^{\prime}}\varphi), in this case jα1j<α1\sum_{j}\alpha^{j}_{1}<\alpha_{1}, j|αj||α|\sum_{j}|\alpha^{j}|\leq|\alpha| and these terms due to induction assumption are O(σ4α1|α||β||γ)O(\sigma^{4-\alpha_{1}-|\alpha|-|\beta|-|\gamma}).

    2. (B)

      We differentiate φx12\varphi_{x_{1}}^{2} and m=2m=2, |α1|+|α2|=|α|+1|\alpha^{1}|+|\alpha^{2}|=|\alpha|+1, α11+α12=α1+1\alpha^{1}_{1}+\alpha_{1}^{2}=\alpha_{1}+1, α1j<α\alpha_{1}^{j}<\alpha and these terms due to induction also are O(σ4α1|α||β||γ|)O(\sigma^{4-\alpha_{1}-|\alpha|-|\beta|-|\gamma|}).

    3. (C)

      We differentiate φx12\varphi_{x_{1}}^{2} and m=1m=1, α11=α1\alpha^{1}_{1}=\alpha_{1}, α11<α1\alpha_{1}^{1}<\alpha_{1}. Then we apply induction by |α|+|β|+|γ||\alpha|+|\beta|+|\gamma|, and these terms due to induction assumption are O(σ4α1|α||β||γ|)O(\sigma^{4-\alpha_{1}-|\alpha|-|\beta|-|\gamma|}) as well.

    Then dividing by φx1\varphi_{x_{1}} and we get O(σ4α1|α||β||γ|)O(\sigma^{4-\alpha_{1}-|\alpha|-|\beta|-|\gamma|}). We need base of the induction in (C). However as |α|=α1|\alpha|=\alpha_{1}, β=γ=0\beta=\gamma=0 terms of type (C) do not appear.

  3. (c)

    We need to expand these estimates to x1>0x_{1}>0. Applying to equation (5.8) x,tαyβθγ\partial_{x,t}^{\alpha}\partial_{y}^{\beta}\partial_{\theta}^{\gamma} with |α|+|β|+|γ|=1|\alpha|+|\beta|+|\gamma|=1 we see that ddtφαβγ\dfrac{d\ }{dt}\varphi_{\alpha\beta\gamma} is bounded and therefore in this case we extend (5.5) from x1=0x_{1}=0 to x1>0x_{1}>0.

    Next, for |α|+|β|+|γ|=2|\alpha|+|\beta|+|\gamma|=2 we get that ddtφαβγ\dfrac{d\ }{dt}\varphi_{\alpha\beta\gamma} does not exceed C(S+1)C(S+1) with

    (5.12) S=α,β,γ|α|+|β|+|γ|=2|φαβγ|2\displaystyle S=\sum_{\alpha,\beta,\gamma\coloneqq|\alpha|+|\beta|+|\gamma|=2}|\varphi_{\alpha\beta\gamma}|^{2}

    and therefore |dS12dt|C(S+1)|\dfrac{dS^{\frac{1}{2}}}{dt}|\leq C(S+1); since S|x1=0=O(σ1)S|_{x_{1}=0}=O(\sigma^{-1}) due to (5.10) we conclude that for TϵσT\leq\epsilon\sigma estimate S12Cσ1S^{\frac{1}{2}}\leq C\sigma^{-1} holds. Then

    (5.13) |φαβγ|Cσ1for |α|+|β|+|γ|=2.\displaystyle|\varphi_{\alpha\beta\gamma}|\leq C\sigma^{-1}\qquad\text{for\ \ }|\alpha|+|\beta|+|\gamma|=2.

    But then, taking α11\alpha_{1}\leq 1 we get that ddtφαβγ\dfrac{d\ }{dt}\varphi_{\alpha\beta\gamma} does not exceed Cσ1(S12+1)C\sigma^{-1}(S^{\frac{1}{2}}+1) with SS defined by (5.12) with summation restricted to α11\alpha_{1}\leq 1 and then due to (5.5) for x1=0x_{1}=0 we conclude that

    (5.14) |φαβγ|Cfor |α|+|β|+|γ|=2,α11.\displaystyle|\varphi_{\alpha\beta\gamma}|\leq C\qquad\text{for\ \ }|\alpha|+|\beta|+|\gamma|=2,\ \alpha_{1}\leq 1.

    And then, taking α1=0\alpha_{1}=0 we get that ddtφαβγ\dfrac{d\ }{dt}\varphi_{\alpha\beta\gamma} does not exceed CC and then due to (5.5) for x1=0x_{1}=0 we conclude that

    (5.15) |φαβγ|CTfor α=β=0,|γ|=2.\displaystyle|\varphi_{\alpha\beta\gamma}|\leq CT\qquad\text{for\ \ }\alpha=\beta=0,\ |\gamma|=2.

    So, (5.5) holds as |α|+|β|+|γ|2|\alpha|+|\beta|+|\gamma|\leq 2, x1>0x_{1}>0.

  4. (d)

    Assume that for |α|+|β|+|γ|<p|\alpha|+|\beta|+|\gamma|<p estimate (5.5) has been proven. Applying to equation (5.8) x,tαyβθγ\partial_{x,t}^{\alpha}\partial_{y}^{\beta}\partial_{\theta}^{\gamma} with |α|+|β|+|γ|=p3|\alpha|+|\beta|+|\gamma|=p\geq 3 we again get a linear combination with bounded coefficients of (5.11) products (plus bounded terms) where this time j|αj||α|+m\sum_{j}|\alpha^{j}|\leq|\alpha|+m. Then, as α1=q\alpha_{1}=q, |α|+|β|+|γ|=p|\alpha|+|\beta|+|\gamma|=p we see from the same analysis as in (ii) that

    |ddtφαβγ|Cσ1S12+Cσ3pq+C\displaystyle|\frac{d\ }{dt}\varphi_{\alpha\beta\gamma}|\leq C\sigma^{-1}S^{\frac{1}{2}}+C\sigma^{3-p-q}+C

    and then (5.5) is extended from x1=0x_{1}=0 to x1>0x_{1}>0, but for α=β=0\alpha=\beta=0, when we proved so far that θγ=O(1)\partial_{\theta}^{\gamma}=O(1). But then |ddtφγ|C|\frac{d\ }{dt}\varphi_{\gamma}|\leq C and (5.5) is again extended from x1=0x_{1}=0 to x1>0x_{1}>0.

  5. (e)

    To prove Statement (iii) we again start from estimates at x1=0x_{1}=0. Then, exactly like in (b) we prove see that

    φx1φx1x1θj+φx1θjφx1x1=O(1)\displaystyle\varphi_{x_{1}}\varphi_{x_{1}x_{1}\theta_{j}}+\varphi_{x_{1}\theta_{j}}\varphi_{x_{1}x_{1}}=O(1)

    and since φx1θj=φx1θj0=O(σ)\varphi_{x_{1}\theta_{j}}=-\varphi^{0}_{x_{1}\theta_{j}}=O(\sigma) as j2j\geq 2, and we know that φx1x1=O(σ1)\varphi_{x_{1}x_{1}}=O(\sigma^{-1}), we conclude that φx1x1θj=O(σ1)\varphi_{x_{1}x_{1}\theta_{j}}=O(\sigma^{-1}). Again, like in (b), using double induction by |α|+|β|+|γ||\alpha|+|\beta|+|\gamma| and α1\alpha_{1} we prove (5.6) as x1=0x_{1}=0.

    Finally, like in (c) and (d) we expand this estimate to x1>0x_{1}>0. We leave easy details to the reader.

Proposition 8.6.

Let φ0(x,y,θ)\varphi^{0}(x,y,\theta) and φ(x,y,θ)\varphi(x,y,\theta) be defined in Proposition 8.4(i) and (ii) respectively. Consider asymptotic solution

(5.16) uh0(x,y,t)(2πh)deiΦ0(x,y,t,θ)n0Bn0(x,y,t,θ)hndθ\displaystyle u^{0}_{h}(x,y,t)\sim(2\pi h)^{-d}\int e^{i\Phi^{0}(x,y,t,\theta)}\sum_{n\geq 0}B^{0}_{n}(x,y,t,\theta)h^{n}\,d\theta
where Φ0(x,y,t,θ)=φ0(x,y,θ)+ta(y,θ)\Phi^{0}(x,y,t,\theta)=\varphi^{0}(x,y,\theta)+ta(y,\theta) and
(5.17) Φ(x,y,t,θ)=φ(x,y,θ)+ta(y,θ),\displaystyle\Phi(x,y,t,\theta)=\varphi(x,y,\theta)+ta(y,\theta),

and Bn0B^{0}_{n} are uniformly smooth and satisfying corresponding transport equations with initial conditions such that uh0(x,y,0)=δ(xy)u^{0}_{h}(x,y,0)=\updelta(x-y):

(5.18) |x,tαyβθγBn0(x,y,t,θ)|Cαβγx,y,t:|xy|+|t|Tα,β,γ.\displaystyle|\partial_{x,t}^{\alpha}\partial_{y}^{\beta}\partial_{\theta}^{\gamma}B^{0}_{n}(x,y,t,\theta)|\leq C_{\alpha\beta\gamma}\quad\forall x,y,t\colon|x-y|+|t|\leq T\ \forall\alpha,\beta,\gamma.

Consider formal asymptotic solution

(5.19) Uh1(x,y,t)=(2πh)deiφ(x,y,t,θ)n0Bn(x,y,t,θ)hndθ\displaystyle U^{1}_{h}(x,y,t)=(2\pi h)^{-d}\int e^{i\varphi(x,y,t,\theta)}\sum_{n\geq 0}B_{n}(x,y,t,\theta)h^{n}\,d\theta

where BnB_{n} satisfy corresponding transport equations and one of the boundary conditions

(5.20) Bn=Bn0\displaystyle B_{n}=-B^{0}_{n} as x1=0,\displaystyle\text{as\ \ }x_{1}=0,
(5.21) φx1Bnix1Bn1=φx1Bn0+ix1Bn10=0\displaystyle\varphi_{x_{1}}B_{n}-i\partial_{x_{1}}B_{n-1}=\varphi_{x_{1}}B^{0}_{n}+i\partial_{x_{1}}B^{0}_{n-1}=0 as x1=0,\displaystyle\text{as\ \ }x_{1}=0,

corresponding to uh1=uh0u^{1}_{h}=-u^{0}_{h} as x1=0x_{1}=0 and x1uh1=x1uh0\partial_{x_{1}}u^{1}_{h}=-\partial_{x_{1}}u^{0}_{h} as x1=0x_{1}=0 respectively. Then the following inequalities hold:

(5.22) |x,tαyβθγBn|Cnαβγ{1+Tσ1|α||β||γ|α1=n=0,σ3nα1|α||β||γ|α1+n1,\displaystyle|\partial_{x,t}^{\alpha}\partial_{y}^{\beta}\partial_{\theta}^{\gamma}B_{n}|\leq C_{n\alpha\beta\gamma}\left\{\begin{aligned} &1+T\sigma^{-1-|\alpha|-|\beta|-|\gamma|}&&\alpha_{1}=n=0,\\ &\sigma^{-3n-\alpha_{1}-|\alpha|-|\beta|-|\gamma|}\qquad&&\alpha_{1}+n\geq 1,\end{aligned}\right.
and if |γ|1|\gamma|\geq 1
(5.23) |x,tαyβθγBn|Cnαβγ{1+Tσ|α||β||γ|α1=n=0,σ13nα1|α||β||γ|α1+n1.\displaystyle|\partial_{x,t}^{\alpha}\partial_{y}^{\beta}\partial_{\theta}^{\gamma}B_{n}|\leq C_{n\alpha\beta\gamma}\left\{\begin{aligned} &1+T\sigma^{-|\alpha|-|\beta|-|\gamma|}&&\alpha_{1}=n=0,\\ &\sigma^{1-3n-\alpha_{1}-|\alpha|-|\beta|-|\gamma|}\qquad&&\alpha_{1}+n\geq 1.\end{aligned}\right.

Proof 8.7.
  1. (a)

    Observe that the transport equation is

    (5.24) (ddt+f)Bn=Bn1,\displaystyle\Bigl{(}\frac{d\ }{dt}+f\Bigr{)}B_{n}=\mathcal{L}B_{n-1},

    where ff is a linear combination with the smooth coefficients of second derivatives of Φ\Phi, =(x,Dx)\mathcal{L}=\mathcal{L}(x,D_{x}) is a second order differential operator with the smooth coefficients, and with the coefficient 11 at Dx12D_{x_{1}}^{2}, and B1=0B_{-1}=0. In virtue of (5.5)

    (5.25) |Dx,tαDyβDθγf|σ1α1|α||β||γ|.\displaystyle|D^{\alpha}_{x,t}D^{\beta}_{y}D_{\theta}^{\gamma}f|\leq\sigma^{-1-\alpha_{1}-|\alpha|-|\beta|-|\gamma|}.

    Consider first n=0n=0. Then equation (5.24) has a right-hand expression 0 and boundary condition (5.21) becomes B0=B00B_{0}=\mp B_{0}^{0} as x1=0x_{1}=0. Let us establish first (5.22) as x1=0x_{1}=0. As α1=0\alpha_{1}=0 these estimates follow from the above boundary condition and (5.18). Consider transport equation (5.24):

    (5.26) φx1B0,x1+k2b(k)(x,xφ)B0,xk12B0,t+12fB0=0\displaystyle\varphi_{x_{1}}B_{0,x_{1}}+\sum_{k\geq 2}b^{(k)}(x,\nabla_{x^{\prime}}\varphi)B_{0,x_{k}}-\frac{1}{2}B_{0,t}+\frac{1}{2}fB_{0}=0

    and set x1=0x_{1}=0. Then all terms in (5.26) are smooth, except the first and the last one, and the latter satisfies

    |x,tαεyβθγ(fB0)|Cσ|α||β||γ|α1=1\displaystyle|\partial^{\alpha-\upvarepsilon}_{x,t}\partial^{\beta}_{y}\partial^{\gamma}_{\theta}(f\,B_{0})|\leq C\sigma^{-|\alpha|-|\beta|-|\gamma|}\qquad\alpha_{1}=1

    due to (5.21) and smoothness of B0B_{0} at x1=0x_{1}=0. Sinnce x,tαyβθγφ\partial^{\alpha}_{x,t}\partial^{\beta}_{y}\partial^{\gamma}_{\theta}\varphi is obtained by division by φx1\varphi_{x_{1}} we arrive to (5.22) as n=0n=0, α1=0\alpha_{1}=0.

    Next we apply a double induction by α1\alpha_{1} and |α|+|β|+|γ||\alpha|+|\beta|+|\gamma| exactly as in the proof of Proposition 8.4, Part (b) with the following modifications:

    1. (A)

      We use transport equation (5.24) rather than eikonal equation.

    2. (B)

      We observe that x,tαεyβθγ\partial^{\alpha-\upvarepsilon}_{x,t}\partial^{\beta}_{y}\partial^{\gamma}_{\theta}, applied to the first term in (5.26) equals to φx1φαβγ\varphi_{x_{1}}\varphi_{\alpha\beta\gamma} plus a linear combination of

      (5.27) 1jmφαjβjγjx,tα0yβ0θγ0B0\displaystyle\prod_{1\leq j\leq m}\varphi_{\alpha^{j}\beta^{j}\gamma^{j}}\partial^{\alpha^{0}}_{x,t}\partial^{\beta^{0}}_{y}\partial^{\gamma^{0}}_{\theta}B_{0}

      with m=1m=1 and jαj=α+ε\sum_{j}\alpha^{j}=\alpha+\upvarepsilon, jβj=β\sum_{j}\beta^{j}=\beta, jγj=γ\sum_{j}\gamma^{j}=\gamma9)9)9) With summation over 0jm0\leq j\leq m and
      |α0|+|β0|+|γ0|<|α|+|β|+|γ||\alpha^{0}|+|\beta^{0}|+|\gamma^{0}|<|\alpha|+|\beta|+|\gamma|.

    3. (C)

      We observe that x,tαεyβθγ\partial^{\alpha-\upvarepsilon}_{x,t}\partial^{\beta}_{y}\partial^{\gamma}_{\theta}, applied to the second term in (5.26) also is a linear combination with bounded coefficients of (5.27) with jβj=β\sum_{j}\beta^{j}=\beta, jγj=γ\sum_{j}\gamma^{j}=\gamma, j|αj||α|+m\sum_{j}|\alpha^{j}|\leq|\alpha|+m, jα1j<α1\sum_{j}\alpha^{j}_{1}<\alpha_{1}9) and |αj|+|βj|+|γj|2|\alpha^{j}|+|\beta^{j}|+|\gamma^{j}|\geq 2 for j1j\geq 1; it is possible that m=0m=0 here.

    We leave simple but tedious arguments to the reader.

  2. (b)

    Then we extend these estimates to x1>0x_{1}>0, TσT\leq\sigma. We apply arguments of Parts (c), (d) of the proof of Proposition 8.4. First, transport equation (5.24) and boundary condition B0=B00B_{0}=B^{0}_{0} as x1=0x_{1}=0 imply that |B0|C|B_{0}|\leq C for TC01σT\leq C_{0}^{-1}\sigma.

    Then applying x,tαyβθγ\partial_{x,t}^{\alpha}\partial_{y}^{\beta}\partial_{\theta}^{\gamma} with |α|+|β|+|γ|=1|\alpha|+|\beta|+|\gamma|=1 to transport equation (5.24) we get

    (5.28) |ddtB0,αβγ|Cσ1S12+Cσ3,S(α,β,γ|B0,αβγ|2)12\displaystyle|\frac{d\ }{dt}B_{0,\alpha\beta\gamma}|\leq C\sigma^{-1}S^{\frac{1}{2}}+C\sigma^{-3},\qquad S\coloneqq\Bigl{(}\sum_{\alpha,\beta,\gamma}|B_{0,\alpha\beta\gamma}|^{2}\Bigr{)}^{\frac{1}{2}}

    due to estimates |x,y,θ2φ|Cσ1|\partial_{x,y,\theta}^{2}\varphi|\leq C\sigma^{-1} and |x,t,y,θf|Cσ3|\partial_{x,t,y,\theta}f|\leq C\sigma^{-3} (Bn,αβx,tαyβθγBnB_{n,\alpha\beta}\coloneqq\partial_{x,t}^{\alpha}\partial_{y}^{\beta}\partial_{\theta}^{\gamma}B_{n}) and then for TC01σT\leq C_{0}^{-1}\sigma we get from this estimate and estimate |B0,αβγ|Cσ2|B_{0,\alpha\beta\gamma}|\leq C\sigma^{-2} at x1=0x_{1}=0 to estimate |B0,αβγ|Cσ2|B_{0,\alpha\beta\gamma}|\leq C\sigma^{-2}. So far the only restriction is |α|+|β|+|γ|=1|\alpha|+|\beta|+|\gamma|=1–here and in summation in the definition of SS.

    However, if we add an extra restriction α1=0\alpha_{1}=0, then due to this estimate (without restriction) and estimates |x,y,θ2φ|C|\partial_{x,y,\theta}^{2}\varphi|\leq C and |x,y,t,θf|Cσ2|\partial_{x,y,t,\theta}f|\leq C\sigma^{-2} if there is only one derivative with respect to x1x_{1}, we arrive to

    |ddtB0,αβγ|Cσ1S12+Cσ2\displaystyle|\frac{d\ }{dt}B_{0,\alpha\beta\gamma}|\leq C\sigma^{-1}S^{\frac{1}{2}}+C\sigma^{-2}

    where SS is defined by (5.28) with summation under the same restriction. Then due to this estimate and estimate |B0,αβγ|C|B_{0,\alpha\beta\gamma}|\leq C as x1=0x_{1}=0 and α1=0\alpha_{1}=0 we arrive to the estimate |B0,αβγ|C+CTσ2|B_{0,\alpha\beta\gamma}|\leq C+CT\sigma^{-2}.

    We apply induction by q|α|+|β|+|γ|q\coloneqq|\alpha|+|\beta|+|\gamma|. Assuming that for lesser values it is proven, we due to this induction assumption and estimates (5.5) and (5.25) arrive to

    (5.29) |ddtB0,αβγ|Cσ1S12+Cσp12|α|+|β|+|γ|,S(α,β,γ|B0,αβγ|2)12\displaystyle|\frac{d\ }{dt}B_{0,\alpha\beta\gamma}|\leq C\sigma^{-1}S^{\frac{1}{2}}+C\sigma^{p-1-2|\alpha|+|\beta|+|\gamma|},\quad S\coloneqq\Bigl{(}\sum_{\alpha,\beta,\gamma}|B_{0,\alpha\beta\gamma}|^{2}\Bigr{)}^{\frac{1}{2}}
    under restrictions
    (5.30) |α|+|β|+|γ|=q,α1|α|p\displaystyle|\alpha|+|\beta|+|\gamma|=q,\quad\alpha_{1}\leq|\alpha|-p
    with p=0p=0 and then we recover estimate
    (5.31) |B0,αβγ|Cσp2|α||β||γ|\displaystyle|B_{0,\alpha\beta\gamma}|\leq C\sigma^{p-2|\alpha|-|\beta|-|\gamma|}

    under these restrictions. Based on it we arrive to (5.29) then to (5.31) now under restriction (5.30) with p=1p=1; again due to (5.31) at x1=0x_{1}=0, and so on for p=2,p=2,\ldots until we reach p=|α|p=|\alpha| but in the latter case we use |B0,αβγ|Cαβγ|B_{0,\alpha\beta\gamma}|\leq C_{\alpha\beta\gamma} at x1=0x_{1}=0 as α1=0\alpha_{1}=0 and thus we achieve some improvement over (5.31): namely we get (5.22) as n=α1=0n=\alpha_{1}=0. We leave easy but tedious details to the reader.

  3. (c)

    Next we apply induction by n1n\geq 1 to estimate Bn,αβγB_{n,\alpha\beta\gamma} at x1=0x_{1}=0. However, first we consider n=1n=1 to establish the pattern. Observe first that under condition (5.20) Bn=Bn0B_{n}=-B^{0}_{n} and therefore

    (5.32) |Bn,αβγ|Cαβγat x1=0 as α1=0,\displaystyle|B_{n,\alpha\beta\gamma}|\leq C_{\alpha\beta\gamma}\qquad\text{at\ \ }x_{1}=0\text{\ \ as\ \ }\alpha_{1}=0,
    but under condition (5.21)
    Bn=Bn0+iφx11(Bn1,x1+Bn1,x10)at x1=0\displaystyle B_{n}=-B_{n}^{0}+i\varphi_{x_{1}}^{-1}\bigr{(}B_{n-1,x_{1}}+B^{0}_{n-1,x_{1}}\bigr{)}\qquad\text{at\ \ }x_{1}=0

    and therefore estimate (5.22) holds as n=1n=1, α1=0\alpha_{1}=0, x1=0x_{1}=0. It follows from estimates for |x1B0,αβγ|Cαβγσ2|α||β||γ||\partial_{x_{1}}B_{0,\alpha\beta\gamma}|\leq C_{\alpha\beta\gamma}\sigma^{-2-|\alpha|-|\beta|-|\gamma|} as α1=0\alpha_{1}=0, x1=0x_{1}=0 (but we divide by φx1\varphi_{x_{1}} which brings an extra factor σ1\sigma^{-1}).

    Recall that the transport equation is

    (5.33) (ddt+f)Bn=Gn,GnBn1,\displaystyle(\frac{d\ }{dt}+f)B_{n}=G_{n},\qquad G_{n}\coloneqq\mathcal{L}B_{n-1},
    and due to results of Part (b)
    (5.34) |Gn,(αε)βγ|Cσ13nα1|α||β||γ|\displaystyle|G_{n,(\alpha-\upvarepsilon)\beta\gamma}|\leq C\sigma^{1-3n-\alpha_{1}-|\alpha|-|\beta|-|\gamma|}

    and therefore in both cases estimate (5.22) holds (as n=1n=1, α1=1\alpha_{1}=1 and x1=0x_{1}=0).

    Applying the same argument as in Part (b) we can extend (5.34) and (5.22) to arbitrary α1\alpha_{1} as n=1n=1, and using induction by nn to arbitrary nn as well. Again, we leave easy but tedious details to the reader.

  4. (d)

    Using the same arguments as in Part (e) of the proof of Proposition 8.4, we can improve these estimates to (5.23) as |γ|1|\gamma|\geq 1.

Remark 8.8.

Under assumption (5.20) we can further improve estimates (5.22) and (5.23) for n1n\geq 1. However it would not improve our final result.

Remark 8.9.

One can wonder how sharp are our estimates.

  1. (i)

    Due to (5.2) and (5.3) φx1x1σ1\varphi_{x_{1}x_{1}}\asymp\sigma^{-1} as bx1(x,xφ)1b_{x_{1}}(x,\nabla_{x^{\prime}}\varphi)\asymp 1. Indeed

    2φx10φx1x10+k2b(k)(x,xφ0)φxkx10φx1t0+bx1(x,xφ0)=0,\displaystyle 2\varphi^{0}_{x_{1}}\varphi^{0}_{x_{1}x_{1}}+\sum_{k\geq 2}b^{(k)}(x,\nabla_{x^{\prime}}\varphi^{0})\varphi^{0}_{x_{k}x_{1}}-\varphi^{0}_{x_{1}t}+b_{x_{1}}(x,\nabla_{x^{\prime}}\varphi^{0})=0,
    2φx1φx1x1+k2b(k)(x,xφ)φxkx1φx1t+bx1(x,xφ)=0,\displaystyle 2\varphi_{x_{1}}\varphi_{x_{1}x_{1}}+\sum_{k\geq 2}b^{(k)}(x,\nabla_{x^{\prime}}\varphi)\varphi_{x_{k}x_{1}}-\varphi_{x_{1}t}+b_{x_{1}}(x,\nabla_{x^{\prime}}\varphi)=0,

    and since φ0\varphi^{0} is smooth function, we conclude that

    k2b(k)(x,xφ0)φxkx10φx1t0+bx1(x,xφ0)=O(σ)\displaystyle\sum_{k\geq 2}b^{(k)}(x,\nabla_{x^{\prime}}\varphi^{0})\varphi^{0}_{x_{k}x_{1}}-\varphi^{0}_{x_{1}t}+b_{x_{1}}(x,\nabla_{x^{\prime}}\varphi^{0})=O(\sigma)
    and then due to (5.3) that
    φx1φx1x1+bx1(x,xφ)=O(σ)as x1=0.\displaystyle\varphi_{x_{1}}\varphi_{x_{1}x_{1}}+b_{x_{1}}(x,\nabla_{x^{\prime}}\varphi)=O(\sigma)\qquad\text{as\ \ }x_{1}=0.

    Therefore estimates (5.5) cannot be improved, at least for derivatives only with respect to x1x_{1}.

  2. (ii)

    Repeating construction of Proposition 8.4, we conclude that under the same assumption bx1(x,xφ)1b_{x_{1}}(x,\nabla_{x^{\prime}}\varphi)\asymp 1, x1kφσ32k\partial_{x_{1}}^{k}\varphi\asymp\sigma^{3-2k} for all k2k\geq 2. Then from transport equation (5.24) we we conclude that B0,x1σ2B_{0,x_{1}}\asymp\sigma^{-2} and repeating construction of Proposition 8.6 we conclude that x1kB0σ2k\partial_{x_{1}}^{k}B_{0}\asymp\sigma^{-2k} for all k1k\geq 1 and finally x1kBnσ3n2k\partial_{x_{1}}^{k}B_{n}\asymp\sigma^{-3n-2k} for all n0n\geq 0, k1k\geq 1.

Proposition 8.10.
  1. (i)

    In the framework of Proposition 8.4(i) as |α|1|\alpha|\leq 1

    (5.35) |x,tαθγ(Φ0Φ¯0)|CγT2|α|\displaystyle|\partial_{x,t}^{\alpha}\partial_{\theta}^{\gamma}\bigl{(}\Phi^{0}-\bar{\Phi}^{0}\bigr{)}|\leq C_{\gamma}T^{2-|\alpha|} with Φ¯0xy,θ+ta(y,θ).\displaystyle\text{with\ \ }\bar{\Phi}^{0}\coloneqq\langle x-y,\theta\rangle+ta(y,\theta).
  2. (ii)

    In the framework of Proposition 8.4(ii) as |α|1|\alpha|\leq 1, α1=0\alpha_{1}=0

    (5.36) |x,tαθγ(ΦΦ¯)|Cγ(σ1|α||γ|T2+T2|α|)with Φ¯xy~,θ+ta(0,y,θ),|\partial_{x,t}^{\alpha}\partial_{\theta}^{\gamma}\bigl{(}\Phi\ -\bar{\Phi}\ \bigr{)}|\leq C_{\gamma}\bigl{(}\sigma^{1-|\alpha|-|\gamma|}T^{2}+T^{2-|\alpha|}\bigr{)}\\ \text{with\ \ }\bar{\Phi}\ \coloneqq\langle x-\tilde{y},\theta\rangle+ta(0,y^{\prime},\theta),

    where y~=(y1,y)\tilde{y}=(-y_{1},y^{\prime}) as y=(y1,y)y=(y_{1},y^{\prime}).

Proof 8.11.

Proof of both statements follows from decomposition of x,tαθγΦ0\partial^{\alpha}_{x,t}\partial_{\theta}^{\gamma}\Phi^{0} into Taylor series with quadratic error. First we prove (i) using estimates (5.2) and then (ii) using estimates (5.5).

For d=2d=2 we will need a better approximation; it will be proven later.

Then we have immediately

Corollary 8.12.
  1. (i)

    In the framework of Proposition 8.4(i) for Ch0(x,y)+|t|ϵCh\leq\ell^{0}(x,y)+|t|\leq\epsilon

    (5.37) a(x,xφ0)=τ,θΦ0(x,y,t,θ)=0\displaystyle a(x,\nabla_{x}\varphi^{0})=\tau,\quad\nabla_{\theta}\Phi^{0}(x,y,t,\theta)=0

    has no more than a single solution θ\theta and if it has, then |t|0|t|\asymp\ell^{0} and θ2Φ0\nabla^{2}_{\theta}\Phi^{0} is a positive definite matrix. This solution θ=θ¯+O(T)\theta=\bar{\theta}+O(T) where θ¯\bar{\theta} is a solution to tθa(x,θ)=yxt\nabla_{\theta}a(x,\theta)=y-x.

  2. (ii)

    In the framework of Proposition 8.4(ii) for Ch(x,y)+|t|ϵCh\leq\ell(x,y)+|t|\leq\epsilon

    (5.38) a(x,xφ)=τ,θΦ(x,y,t,θ)=0\displaystyle a(x,\nabla_{x}\varphi)=\tau,\quad\nabla_{\theta}\Phi(x,y,t,\theta)=0

    has has no more than a single solution θ\theta and if it has then |t||t|\asymp\ell and θ2Φ\nabla^{2}_{\theta}\Phi is a positive definite matrix. This solution θ=θ¯+O(T)\theta=\bar{\theta}+O(T) where θ¯\bar{\theta} is a solution to tθa(x,θ)=y~xt\nabla_{\theta}a(x,\theta)=\tilde{y}-x.

Proposition 8.13.

In the framework of Proposition 8.4(ii) under extra assumption

(5.39) σ2h1δ\displaystyle\sigma^{2}\ell\geq h^{1-\delta}
the following estimate holds
(5.40) |Fth1τ(χ¯T(t)uh1(x,y,t)χ¯T(t)Uh1(x,y,t))|Chs\displaystyle|F_{t\to h^{-1}\tau}\Bigl{(}\bar{\chi}_{T^{\prime}}(t)u^{1}_{h}(x,y,t)-\bar{\chi}_{T}(t)U^{1}_{h}(x,y,t)\Bigr{)}|\leq Ch^{s}

with Uh1(x,y,t)U^{1}_{h}(x,y,t) defined by (5.19) with Bn(x,y,t,θ)B_{n}(x,y,t,\theta) described above and multiplied by ϕ(θ)\phi(\theta) supported in 2ϵ3ρ2\epsilon_{3}\rho-vicinity and equal 11 in ϵ3ρ\epsilon_{3}\rho-vicinity of θ\theta described in Corollary 8.12(ii) and with ChTTϵCh\leq T\leq T^{\prime}\leq\epsilon, T=cT=c\ell, ρ=σ2\rho=\sigma^{2}.

Proof 8.14.

Let us fix x=x¯x=\bar{x} and y=y¯y=\bar{y} in (5.40) and use xx as a variable. Due to Proposition 8.1 there is a single generalized billiard of the length ϵ\leq\epsilon with reflections at X\partial X from x¯\bar{x} to y¯\bar{y}. It has the length \asymp\ell, exactly one reflection and a reflection angle ρ=(ν(x¯)+ν(y¯))1h13δ\asymp\rho=({\nu}(\bar{x})+{\nu}(\bar{y}))\ell^{-1}\gtrsim h^{\frac{1}{3}-\delta}.

Further, Proposition 8.6(ii) implies that for σh13δ\sigma\geq h^{\frac{1}{3}-\delta} (which is due to (5.39) and σC0\sigma\geq C_{0}\ell) (5.19) is a proper asymptotic series as |t|T|t|\leq T and we define Uh1(x,y,t)U^{1}_{h}(x,y,t) through it (with cut-off).

Furthermore, Corollary 8.12 implies that then Uh1(x,y,t)U^{1}_{h}(x,y,t) is negligible outside Ω2T,3ρ,3σ\Omega_{2T,3\rho,3\sigma} (by (x1,x,ξ)(x_{1},x^{\prime},\xi^{\prime})) of this billiard while the standard propagation results imply that uh1(x,y,t)tQyu^{1}_{h}(x,y,t)\,^{t}\!Q_{y} is also negligible outside of this vicinity provided symbol of QQ is supported in the similar vicinity of (y¯,θ¯)(\bar{y},\bar{\theta}^{\prime}), θ¯\bar{\theta}^{\prime} corresponds to this billiard. Let symbol of QQ be 11 in the similar vicinity of (y¯,ξ¯)(\bar{y},\bar{\xi}^{\prime}). Then uh1(x,y,t)tQyUh1(x,y,t)u^{1}_{h}(x,y,t)\,^{t}\!Q_{y}\equiv U^{1}_{h}(x,y,t) for TtT-T\leq t\leq T with 0<T=C00<T=C_{0}\ell.

On the other hand, in this case uh1(x,y,t)(ItQy)0u^{1}_{h}(x,y,t)(I-^{t}\!Q_{y})\equiv 0 in (ϵρ,ϵ)(\epsilon\rho\ell,\epsilon\ell)-vicinity of x¯\bar{x} as TtϵT-T\leq t\leq\epsilon T. Case of ϵTtT-\epsilon T\leq t\leq T is considered in the same way, albeit with the billiard from x¯\bar{x} to y¯\bar{y}. Then we arrive to (5.40) with T=TT^{\prime}=T.

Finally, the standard propagation results imply that

(5.41) |Fth1τ((χ¯T(t)χ¯T(t))uh1(x,y,t))|Chs\displaystyle|F_{t\to h^{-1}\tau}\Bigl{(}\bigl{(}\bar{\chi}_{T^{\prime}}(t)-\bar{\chi}_{T}(t)\bigr{)}u^{1}_{h}(x,y,t)\Bigr{)}|\leq Ch^{s}

as TTϵT\leq T^{\prime}\leq\epsilon. Then (5.40) expands to T:TTϵT^{\prime}\colon T\leq T^{\prime}\leq\epsilon.

9 Spectral asymptotics

Proposition 9.15.

In the framework of Proposition 8.4(ii) under extra assumption

(5.42) σ2h1δ,h35\displaystyle\sigma^{2}\ell\geq h^{1-\delta},\qquad\ell\geq h^{\frac{3}{5}}
the following asymptotics holds
(5.43) eh1,𝖳(x,y,τ)=(2πh)da(y,θ)<τeih1φ(x,y,θ)𝑑θ+O(h1d).\displaystyle e^{1,\mathsf{T}}_{h}(x,y,\tau)=(2\pi h)^{-d}\int_{a(y,\theta)<\tau}e^{ih^{-1}\varphi(x,y,\theta)}\,d\theta+O(h^{1-d}).

Proof 9.16.

Let us plug Uh1U^{1}_{h} instead of uh1u^{1}_{h} into Tauberian expression eh1,𝖳(x,y,τ)e^{1,\mathsf{T}}_{h}(x,y,\tau) with T=ϵT=\epsilon. Let us decompose Bn(x,y,t,θ)B_{n}(x,y,t,\theta) into asymptotic series by tkt^{k}.

  1. (a)

    Then terms with k=0k=0 become

    (5.44) (2πh)da(y,θ)<τn0eih1φ(x,y,θ)Bn(x,y,0,θ)hndθ\displaystyle(2\pi h)^{-d}\int_{a(y,\theta)<\tau}\sum_{n\geq 0}e^{ih^{-1}\varphi(x,y,\theta)}B_{n}(x,y,0,\theta)h^{n}\,d\theta

    and we claim that these terms do not exceed Cσ3nhd+n(h1)(d+1)/2C\sigma^{-3n}h^{-d+n}(h\ell^{-1})^{(d+1)/2}. Indeed, we know from Proposition 8.4(ii) that φ\varphi is uniformly infinitely smooth by θ\theta and from Corollary 8.12 that there are no stationary points by θ\theta and restriction of φ(x,y,θ)\varphi(x,y,\theta) to Σ(y,τ)\Sigma(y,\tau) has two non-degenerate critical points. The leading terms would be of this magnitude, while all other terms would have extra factors not exceeding h/σ2h/\sigma^{2}\ell due to (5.22).

    We want to estimate all extra terms by Ch1dCh^{1-d}. One can see easily, that d=2d=2 is the worst case but even then therms with n1n\geq 1 are O(h1d)O(h^{1-d}) and terms with h/σ2h/\sigma^{2}\ell appear when we differentiate B0(x,y,0,θ)B_{0}(x,y,0,\theta) by θ\theta. Then due to (5.23) these terms do not exceed

    Chd(h1)d+32(1+σ2)\displaystyle Ch^{-d}(h\ell^{-1})^{\frac{d+3}{2}}(1+\ell\sigma^{-2})

    and this does not exceed Ch1dCh^{1-d} under assumption (5.32).

    Finally, since B0(x,y,0,θ)=1+O(σ1)B_{0}(x,y,0,\theta)=1+O(\ell\sigma^{-1}) we can replace B0(x,y,0,θ)B_{0}(x,y,0,\theta) by 11 resulting in the main part of asymptotics (5.43).

  2. (b)

    Consider now terms with k1k\geq 1 in the decomposition of Bn(x,y,t,θ)B_{n}(x,y,t,\theta). We can rewrite these terms as

    (2πh)1dτk1Σ(y,τ)Bn,k(x,y,θ)hn+k1𝑑θ:dθa(y,θ)\displaystyle(2\pi h)^{1-d}\partial_{\tau}^{k-1}\int_{\Sigma(y,\tau)}B^{\prime}_{n,k}(x,y,\theta)h^{n+k-1}\,d\theta:d_{\theta}a(y,\theta)
    and using ξ\xi-microhyperbolicity rewrite it as
    (5.45) (2πh)1dΣ(y,τ)Bn,k′′(x,y,θ)hn(h1)k𝑑θ:dθa(y,θ).\displaystyle(2\pi h)^{1-d}\int_{\Sigma(y,\tau)}B^{\prime\prime}_{n,k}(x,y,\theta)h^{n}(h\ell^{-1})^{k}\,d\theta:d_{\theta}a(y,\theta).

    with Bn,k′′B^{\prime\prime}_{n,k} coming from BnB_{n} with kk derivatives by tt and no more than (k1)(k-1) by θ\theta. Then using (5.22)–(5.23) we can estimate these terms by Chd12d12Ch1dCh^{-\frac{d-1}{2}}\ell^{-\frac{d-1}{2}}\leq Ch^{1-d}.

Remark 9.17.

The similar asymptotics

(5.46) eh0,𝖳(x,y,τ)=(2πh)da(y,θ)<τeih1φ0(x,y,θ)𝑑θ+O(h1d)\displaystyle e^{0,\mathsf{T}}_{h}(x,y,\tau)=(2\pi h)^{-d}\int_{a(y,\theta)<\tau}e^{ih^{-1}\varphi^{0}(x,y,\theta)}\,d\theta+O(h^{1-d})

is well-known.

Remark 9.18.

Consider in the framework of Proposition 8.4(i), (ii) integrals Ih(φ)I_{h}(\varphi) and Ih(φ0)I_{h}(\varphi^{0}) in the right-hand expressions of (5.46) and (5.43) correspondingly:

(5.47) Ih(φ)(2πh)da(y,θ)<τeih1φ(x,y,θ)𝑑θ.\displaystyle I_{h}(\varphi)\coloneqq(2\pi h)^{-d}\int_{a(y,\theta)<\tau}e^{ih^{-1}\varphi(x,y,\theta)}\,d\theta.
  1. (i)

    Let d3d\geq 3. Then Ih(φ0)=Ih(φ¯0)+O(h1d)I_{h}(\varphi^{0})=I_{h}(\bar{\varphi}^{0})+O(h^{1-d}) and Ih(φ)=Ih(φ¯)+O(h1d)I_{h}(\varphi)=I_{h}(\bar{\varphi})+O(h^{1-d}) with φ¯0(x,y,θ)=xy,θ\bar{\varphi}^{0}(x,y,\theta)=\langle x-y,\theta\rangle, φ¯0(x,y,θ)=xy~,θ\bar{\varphi}^{0}(x,y,\theta)=\langle x-\tilde{y},\theta\rangle.

  2. (ii)

    Let d=2d=2. Then Ih(φ0)Ih(φ¯0)I_{h}(\varphi^{0})\equiv I_{h}(\bar{\varphi}^{0}) with an error O(h1)O(h^{-1}) as 0h13\ell^{0}\geq h^{\frac{1}{3}}, O(h1)O(h^{-1}) as 0(x,y)h13\ell^{0}(x,y)\geq h^{\frac{1}{3}}, O(h1232)O(h^{-\frac{1}{2}}\ell^{-\frac{3}{2}}) as h120(x,y)h13h^{\frac{1}{2}}\leq\ell^{0}(x,y)\leq h^{\frac{1}{3}} and O(h3212)O(h^{-\frac{3}{2}}\ell^{\frac{1}{2}}) as h0(x,y)h12h\leq\ell^{0}(x,y)\leq h^{\frac{1}{2}}.

    Also Ih(φ)Ih(φ¯)I_{h}(\varphi)\equiv I_{h}(\bar{\varphi}^{)} with an error O(h1)O(h^{-1}) as 0h13\ell^{0}\geq h^{\frac{1}{3}}, O(h1)O(h^{-1}) as 0(x,y)h13\ell^{0}(x,y)\geq h^{\frac{1}{3}}, O(h1232)O(h^{-\frac{1}{2}}\ell^{-\frac{3}{2}}) as h12(x,y)h13h^{\frac{1}{2}}\leq\ell(x,y)\leq h^{\frac{1}{3}} and O(h3212)O(h^{-\frac{3}{2}}\ell^{\frac{1}{2}}) as h(x,y)h12h\leq\ell(x,y)\leq h^{\frac{1}{2}}.

Proof 9.19.

Observe first that due to the stationary phase principle Ih(φ)=Chd(h1)(d+1)/2I_{h}(\varphi)=Ch^{-d}(h\ell^{-1})^{(d+1)/2} and the same is true for Ih(φ¯)I_{h}(\bar{\varphi}). Therefore for h12\ell\geq h^{\frac{1}{2}} we have simply estimates rather than the error estimates.

On the other hand due to proposition 8.10 φ(x,y,θ)=xy~,θ+O(2)\varphi(x,y,\theta)=\langle x-\tilde{y},\theta\rangle+O(\ell^{2}), for h12\ell\leq h^{\frac{1}{2}} the error does not exceed Chd(h1)(d+1)/2×2h1Ch^{-d}(h\ell^{-1})^{(d+1)/2}\times\ell^{2}h^{-1} which is O(h1d)O(h^{1-d}).

The same is true for Ih(φ0)I_{h}(\varphi^{0}) and Ih(φ¯0)I_{h}(\bar{\varphi}^{0}) albeith with 0\ell^{0} rather than \ell.

Now we need to deal with d=2d=2 and 0h13\ell^{0}\leq h^{\frac{1}{3}} or h13\ell\leq h^{\frac{1}{3}}.

Proposition 9.20.

Let d=2d=2. Then

  1. (i)

    In the framework of Proposition 8.4(i) for h1δ0(x,y)h13h^{1-\delta}\leq\ell^{0}(x,y)\leq h^{\frac{1}{3}}

    (5.48) a(y,θ)<τeih1φ0(x,y,θ)𝑑θ=a(w,ξ)<τeih1xy,ξ𝑑ξ+O(h)\displaystyle\int_{a(y,\theta)<\tau}e^{ih^{-1}\varphi^{0}(x,y,\theta)}\,d\theta=\int_{a(w,\xi)<\tau}e^{ih^{-1}\langle x-y,\xi\rangle}\,d\xi+O(h)

    with w=12(x+y)w=\frac{1}{2}(x+y).

  2. (ii)

    In the framework of Proposition 8.4(ii) for h1δ(x,y)h13h^{1-\delta}\leq\ell(x,y)\leq h^{\frac{1}{3}}

    (5.49) a(y,θ)<τeih1φ(x,y,θ)𝑑θ=a(w,ξ)<τeih1(xy~,ξ+12ϰ(w,ξ,τ)(x12+y12))𝑑ξ+O(h)\int_{a(y,\theta)<\tau}e^{ih^{-1}\varphi(x,y,\theta)}\,d\theta\\ =\int_{a(w,\xi)<\tau}e^{ih^{-1}(\langle x-\tilde{y},\xi\rangle+\frac{1}{2}\varkappa(w^{\prime},\xi^{\prime},\tau)(x_{1}^{2}+y_{1}^{2}))}\,d\xi+O(h)

    with w=(0,12(x+y))w=(0,\frac{1}{2}(x^{\prime}+y^{\prime})) and with

    (5.50) ϰ(x,ξ,τ)=(τb(x2,ξ2))12ax1(x,ξ)|x1=ξ1=0;\displaystyle\varkappa(x^{\prime},\xi^{\prime},\tau)=(\tau-b(x_{2},\xi_{2}))^{-\frac{1}{2}}a_{x_{1}}(x,\xi)|_{x_{1}=\xi_{1}=0};

    recall that b(w2,ξ2)=a(x,ξ)|x1=ξ1=0b(w_{2},\xi_{2})=a(x,\xi)|_{x_{1}=\xi_{1}=0}.

Proof 9.21.
  1. (i)

    Consider (z,ξ¯):a(w¯,ξ¯)=τ(z,\bar{\xi})\colon a(\bar{w},\bar{\xi})=\tau and Hamiltonian trajectory Ψt(w¯,ξ¯)\Psi_{t}(\bar{w},\bar{\xi}) of a(x,ξ)a(x,\xi), Ψ0(w¯,ξ¯)=(w¯,ξ¯)\Psi_{0}(\bar{w},\bar{\xi})=(\bar{w},\bar{\xi}). Then as (y(t),θ(t))=Ψt(w¯,ξ¯)(y(t),\theta(t))=\Psi_{t}(\bar{w},\bar{\xi}) and (x(t),.)=Ψt(w¯,ξ¯)(x(t),.)=\Psi_{-t}(\bar{w},\bar{\xi}), |t|C0|t|\leq C_{0}\ell we have

    (5.51) φ0(x(t),y(t),θ(t))=x(t)y(t),ξ¯+O(3)\displaystyle\varphi^{0}(x(t),y(t),\theta(t))=\langle x(t)-y(t),\bar{\xi}\rangle+O(\ell^{3})

    where in this part of the proof for simplicity we write \ell rather than 0\ell^{0}.

    Observe that as h1δ\ell\geq h^{1-\delta} we can reduce integrals in (5.48) to Σ(y,τ)\Sigma(y,\tau) and Σ(w¯,τ)\Sigma(\bar{w},\tau) respectively, gaining factor h1h\ell^{-1}. Consider first those reduced integrals over ϵ\epsilon-vicinities of θ(t)\theta(t) and ξ¯\bar{\xi}. We see that there is a diffeomorphism of these vicinities and

    φ0(x(t),y(t),θ(w¯,ξ(s),t))=x(t)y(t),ξ+O(3+s2)\displaystyle\varphi^{0}(x(t),y(t),\theta(\bar{w},\xi(s),t))=\langle x(t)-y(t),\xi\rangle+O(\ell^{3}+s^{2}\ell)

    provided aθθa_{\theta}\parallel\theta at (w¯,ξ¯)(\bar{w},\bar{\xi}) (we can assume it without any loss of the generality), ss is an angle parameter on Σ(w¯,τ)\Sigma(\bar{w},\tau), ξ(0)=ξ¯\xi(0)=\bar{\xi}. Making change of variables we estimate an error by C(h1)32×(3h1+ε2h1+)C(h\ell^{-1})^{\frac{3}{2}}\times(\ell^{3}h^{-1}+\varepsilon^{2}\ell h^{-1}+\ell) when we integrate over |s|ε|s|\leq\varepsilon, and when we integrate over |s|ε|s|\geq\varepsilon, we can multiply it by an arbitrary large power of h/ε2h/\varepsilon^{2}\ell. Taking ε=(h1)12\varepsilon=(h\ell^{-1})^{\frac{1}{2}} we make both errors O(h)O(h).

    On the other hand, fix x¯\bar{x} and y¯\bar{y} and find corresponding w¯\bar{w}, ξ¯\bar{\xi} and tt. One can see easily that w¯=w+O(2)\bar{w}=w+O(\ell^{2}) and the error in the integral in the right hand expression of (5.48) when we redefine ww does not exceed C(h)122=O(h)C(h\ell)^{\frac{1}{2}}\ell^{2}=O(h).

  2. (ii)

    Making change of variables 𝗑1=x1+12ϰ(w2,ξ,τ)x12\mathsf{x}_{1}=x_{1}+\frac{1}{2}\varkappa(w_{2},\xi^{\prime},\tau)x_{1}^{2} and therefore 𝗒1=y1+12ϰ(w2,ξ,τ)y12\mathsf{y}_{1}=y_{1}+\frac{1}{2}\varkappa(w_{2},\xi^{\prime},\tau)y_{1}^{2} we would arrive to operator with the symbol which, after division by (aτ)(a-\tau), modulo O(2)O(\ell^{2}) does not depend on 𝗑1\mathsf{x}_{1} and therefore with the corresponding phase function

    φ=(𝗑1+𝗒1)ξ1+(x2y2)ξ2+O(3).\displaystyle\varphi=(\mathsf{x}_{1}+\mathsf{y}_{1})\xi_{1}+(x_{2}-y_{2})\xi_{2}+O(\ell^{3}).
    Then we can use the the method of reflection and after this, Statement (i). So we get
    (5.52) a(w,ξ)<τeih1((𝗑1+𝗒1)ξ1+(x2y2)ξ2𝑑ξ+O(h)\displaystyle\int_{a(w,\xi)<\tau}e^{ih^{-1}((\mathsf{x}_{1}+\mathsf{y}_{1})\xi_{1}+(x_{2}-y_{2})\xi_{2}}\,d\xi+O(h)

    with w=(0,12(x2+y2))w=(0,\frac{1}{2}(x_{2}+y_{2})); which is exactly the right-hand expression of (5.49).

Remark 9.22.
  1. (i)

    Therefore in Theorem 1.2

    (5.53) eh,𝖼𝗈𝗋𝗋(x,y,τ)=[](2πh)1{a(12(x+y),ξ)<τ}eih1xy~,ξ(eih112ϰ(12(x2+y2),ξ2,τ)1)𝑑ξe_{h,\mathsf{corr}}(x,y,\tau)\\ =[\mp](2\pi h)^{-1}\int_{\{a(\frac{1}{2}(x+y),\xi)<\tau\}}e^{ih^{-1}\langle x-\tilde{y},\xi\rangle}\Bigl{(}e^{ih^{-1}\frac{1}{2}\varkappa(\frac{1}{2}(x_{2}+y_{2}),\xi_{2},\tau)}-1\Bigr{)}\,d\xi

    with ϰ(x2,ξ2,τ)\varkappa(x_{2},\xi_{2},\tau) defined by (5.50).

  2. (ii)

    One can see easily that in the framework of Proposition 7.7 expressions (5.53) and (4.49) coincide modulo O(h1)O(h^{-1}).

Chapter 6 Synthesis and final remarks

Proof 9.1 (Proof of Theorems 1.1 and 1.2).

We know that for (x,y)ϵ\ell(x,y)\leq\epsilon in the frameworks of Theorems 1.1 and1.2 estimate (1.8) holds.

eh(x,y,τ)=eT,h𝖳(x,y,τ)+O(h1d).\displaystyle e_{h}(x,y,\tau)=e^{\mathsf{T}}_{T,h}(x,y,\tau)+O(h^{1-d}).
  1. (i)

    We also know from Section 2 that in the framework of Theorem 1.1 eT,h𝖳(x,y,τ)e^{\mathsf{T}}_{T,h}(x,y,\tau) can be defined as an oscillatory integral and Proposition 9.20(i) implies that modulo O(h1d)O(h^{1-d}) this oscillatory integral can be rewritten as eh𝖶(x,y,τ)e^{\mathsf{W}}_{h}(x,y,\tau) defined by (1.3). This proves Theorem 1.1.

  2. (ii)

    In the framework of Theorem 1.2 we established in Section 3

    eT,h1,𝖳(x,y,τ)=O(h1d)\displaystyle e^{1,\mathsf{T}}_{T,h}(x,y,\tau)=O(h^{1-d})

    as d3d\geq 3 and (x,y)h12δ\ell(x,y)\geq h^{\frac{1}{2}-\delta} and as d=2d=2 and (x,y)h13δ\ell(x,y)\geq h^{\frac{1}{3}-\delta}. Further, we proved there that eT,h1,𝖳(x,y,τ)=O(h1dδ)e^{1,\mathsf{T}}_{T,h}(x,y,\tau)=O(h^{1-d-\delta}) as d=2d=2 and (x,y)h13+δ\ell(x,y)\geq h^{\frac{1}{3}+\delta}. In both cases we also assume there that |xy|ϵ(x,y)|x^{\prime}-y^{\prime}|\geq\epsilon\ell(x,y). However it follows from Section 5 that this latter condition is not necessary. Combined with Remark 1.3(ii) it proves Theorem 1.2 in this case.

    Let d3d\geq 3. In Section 4 we proved asymptotics

    (6.1) eT,h1,𝖳(x,y)=[]eh0,𝖶(x,y~,τ)+O(h1d)\displaystyle e^{1,\mathsf{T}}_{T,h}(x,y)=[\mp]e^{0,\mathsf{W}}_{h}(x,\tilde{y},\tau)+O(h^{1-d})

    if either (x,y)h12+δ\ell(x,y)\leq h^{\frac{1}{2}+\delta} or h12+δ(x,y)h12δh^{\frac{1}{2}+\delta}\leq\ell(x,y)\leq h^{\frac{1}{2}-\delta} and

    (6.2) ν(x)+ν(y)σ(x,y)\displaystyle{\nu}(x)+{\nu}(y)\leq\sigma\ell(x,y)

    with σ=h2δ\sigma=h^{2\delta}. In Section 5 we proved the same asymptotics as

    (6.3) ν(x)+ν(y)σ(x,y)\displaystyle{\nu}(x)+{\nu}(y)\geq\sigma\ell(x,y)

    with σ=h12δ12+C0\sigma=h^{\frac{1}{2}-\delta}\ell^{-\frac{1}{2}}+C_{0}\ell. These domains overlap and cover all values of h12δ\ell\leq h^{\frac{1}{2}-\delta} and σ\sigma.

  3. (iii)

    Let d=2d=2. Then instead of asymptotics (6.1) we have

    (6.4) eT,h1,𝖳(x,y)=[]eh0,𝖶(x,y~,τ)+eh,𝖼𝗈𝗋𝗋(x,y,τ)+O(h1d)\displaystyle e^{1,\mathsf{T}}_{T,h}(x,y)=[\mp]e^{0,\mathsf{W}}_{h}(x,\tilde{y},\tau)+e_{h,\mathsf{corr}}(x,y,\tau)+O(h^{1-d})

    and Sections 4 and 5 cover cases (6.2) with σ=h1+δ2\sigma=h^{1+\delta}\ell^{-2} and (6.3) with σ=h13δ\sigma=h^{\frac{1}{3}-\delta}, (x,y)h13+δ\ell(x,y)\leq h^{\frac{1}{3}+\delta}. These domains overlap and cover all values h13+δ\ell\leq h^{\frac{1}{3}+\delta} and σ\sigma. Here we are left with domain h13+δh13δh^{\frac{1}{3}+\delta}\leq\ell\leq h^{\frac{1}{3}-\delta}, σ=h13δ\sigma=h^{\frac{1}{3}-\delta} where we have less precise estimate eT,h1,𝖳(x,y,τ)=O(h1dδ)e^{1,\mathsf{T}}_{T,h}(x,y,\tau)=O(h^{1-d-\delta}).

Remark 9.2.
  1. (i)

    We would like to derive remainder estimate O(h1)O(h^{-1}) as d=2d=2 and (x,y)ϵ\ell(x,y)\leq\epsilon, thus removing the gap h13+δh13δh^{\frac{1}{3}+\delta}\leq\ell\leq h^{\frac{1}{3}-\delta}, ν(x)+ν(y)h13δ{\nu}(x)+{\nu}(y)\leq h^{\frac{1}{3}-\delta}.

  2. (ii)

    For Schrödinger operator (1.4) can get rid of ξ\xi-microhyperbolicity assumption by rescaling method using scaling function

    (6.5) γx=(ϵ|V(x)τ|+h23).\displaystyle\gamma_{x}=(\epsilon|V(x)-\tau|+h^{\frac{2}{3}}).

    Then using arguments of Proposition LABEL:OOD2-prop-3.2 and Remark LABEL:OOD2-remark-4.1 of [Ivr3] we have a remainder estimate O(h1dγx(d3)/2γy(d3)/2)O(h^{1-d}\gamma_{x}^{(d-3)/2}\gamma_{y}^{(d-3)/2}) which as d3d\geq 3 is O(h1d)O(h^{1-d}); however as d=2d=2 it is as bad as O(h43)O(h^{-\frac{4}{3}}).

  3. (iii)

    As d=2d=2 away from X\partial X remainder estimate O(h1)O(h^{-1}) in the regular zone and O(h1615)O(h^{-\frac{16}{15}}) or better in the singular zone has been derived in [Ivr3]. However it required the analysis of the Hamiltonian trajectories and introduction of the correction term. This does not look feasible near the boundary especially because both the boundary and degeneration define correction term.

References

  • [Arn] V. I. Arnold. Mathematical Methods of Classical Mechanics. Springer-Verlag, 1990.
  • [Av1] V. G. Avakumovič. Über die eigenfunktionen auf geschlossen riemannschen mannigfaltigkeiten. Math. Z., 65:324–344 (1956).
  • [Car1] T. Carleman. Propriétes asymptotiques des fonctions fondamentales des membranes vibrantes. In C. R. 8-ème Congr. Math. Scand., Stockholm, 1934, pages 34–44, Lund (1935).
  • [Car2] T. Carleman. Über die asymptotische Verteilung der Eigenwerte partieller Differentialgleichungen. Ber. Sachs. Acad. Wiss. Leipzig, 88:119–132 (1936).
  • [Hö] L. Hörmander. The spectral function of an elliptic operator. Acta Math., 121:193–218 (1968).
  • [Ivr1] V. Ivrii. Precise Spectral Asymptotics for Elliptic Operators. Lect. Notes Math. Springer-Verlag 1100 (1984).
  • [Ivr2] V. Ivrii. Microlocal Analysis, Sharp Spectral, Asymptotics and Applications. Volume I. Semiclassical Microlocal Analysis and Local and Microlocal Semiclassical Asymptotics. Springer-Verlag (2019).
  • [Ivr3] V. Ivrii. Pointwise spectral asymptotics out of the diagonal near degeneration.
    ArXiv:108.13675, August 2021.
  • [Ivr4] Pointwise spectral asymptotics out of the diagonal. Presentation, July 26, 2021.
  • [Lev1] B. M. Levitan. On the asymptotic behaviour of the spectral function of the second order elliptic equation. Izv. AN SSSR, Ser. Mat., 16(1):325–352 (1952). In Russian.
  • [Lev2] B. M. Levitan. Asymptotic behaviour of the spectral function of elliptic operator. Russian Math. Surveys, 26(6):165–232 (1971).
  • [Sa] Yu. G. Safarov. The asymptotics of the spectral function of positive elliptic operator without non-trapping condition. Funct. Anal. Appl., 22(3):213–223 (1988).
  • [SaVa] Yu. G. Safarov, D. G. Vassiliev. The asymptotic distribution of eigenvalues of differential operators. AMS Transl., Ser. 2, 150 (1992).
  • [Se1] R. Seeley. A sharp asymptotic estimate for the eigenvalues of the Laplacian in a domain of 𝐑3{\boldsymbol{R}}^{3}. Advances in Math., 102(3):244–264 (1978).
  • [Se2] R. Seeley. An estimate near the boundary for the spectral function of the Laplace operator. Amer. J. Math., 102(3):869–902 (1980).
  • [Shu] M. A. Shubin. Pseudodifferential Operators and Spectral Theory. Springer-Verlag (2001).