1 Introduction

Let \((\Omega ,g)\) be an n-dimensional, compact \(C^{\infty }\) Riemannian manifold with boundary M and corresponding unit exterior normal \(\nu \). By some abuse of notation, we also let \(\nu \) denote a smooth vector field extension and \(\gamma _{M}: C^0(\Omega ) \rightarrow C^0(M)\) be the boundary restriction map. Let \({\mathcal {D}}: C^{\infty }(M) \rightarrow C^{\infty }(M)\) be the associated Dirichlet-to-Neumann (DtN) operator defined by

$$\begin{aligned} {\mathcal {D}} f := \gamma _{M} \partial _{\nu } u, \end{aligned}$$
(1.1)

where u solves the Dirichlet problem

$$\begin{aligned} \Delta _{g} u(x)&= 0, \,\,\, x \in \Omega , \nonumber \\ u(q)&= f(q), \,\,\, q \in M. \end{aligned}$$
(1.2)

The operator \({\mathcal {D}}\) is an elliptic, first-order, self-adjoint pseudodifferential operator (see for example [21, Section 7.11]) with an \(L^2\)-normalized basis of eigenfunctions \(\varphi _{j}; j=1,2,....\) It is useful here to work in the semiclassical setting from the outset. Choosing \(h^{-1} \in \text {Spec} \, {\mathcal {D}},\) the corresponding eigenfunction \(\varphi _h\) then satisfies the semiclassical eigenfunction equation

$$\begin{aligned} h {\mathcal {D}}\varphi _h = \varphi _h. \end{aligned}$$

The harmonic extension, \(u_{h} \in C^{\infty }(\Omega ),\) of a DtN eigenfunction \(\varphi _h\) is called a Steklov eigenfunction.

There has been a substantial amount of recent work devoted to the study of the asymptotic behavior of the DtN eigenvalues and both DtN and Steklov eigenfunctions, including the asymptotics of eigenfunction nodal sets (see for example [1, 3, 4, 6, 15, 16, 20, 26,27,28] and references therein).

For large eigenvalues, Steklov eigenfunctions possess both high oscillation inherited from the boundary DtN eigenfunctions and very sharp decay into the interior of \(\Omega .\) As a consequence, even though Steklov eigenfunctions decay rapidly, the oscillation implies, in particular, that the nodal sets have intricate structure. It has been conjectured [3] that the analogue of Yau’s conjecture [23, 24] for nodal volumes holds in the Steklov case. This was recently proved for real-analytic Riemann surfaces in [15].

The question of decay of Steklov eigenfunctions into the interior of M when (Mg) is real analytic was first raised by Hislop–Lutzer [6] where they conjecture that the Steklov eigenfunctions decay into the interior as \(e^{-d(x,\partial \Omega )/h}.\) In the special case where dim \(\Omega =2\) exponential decay with respect to \(d(x,\partial \Omega )\) was indeed proved in [15] and the eigenfunction decay is a key feature in their main results on nodal length. However, the analysis in [15] relies heavily on the assumption that the dimension equals two.

In Theorem 1, we prove an exponential decay result for Steklov eigenfunctions for general real-analytic metrics in arbitrary dimension that is sharp to first order at the boundary, \(\partial \Omega = M\), and we bound the quadratic error in the decay rate in terms of boundary curvature. In particular, we prove the conjecture due to Hislop–Lutzer [6].

One can heuristically view such exponential decay estimates for the Steklov eigenfunctions as describing the ‘tunnelling’ of the boundary DtN eigenfunctions on M into the interior of the manifold \(\Omega .\) In the Schrödinger case \(P(h)=-h^2\Delta _g+V-E\), one thinks of eigenfunctions as tunnelling from \(V\le E\) into the forbidden region \(V>E\). Since the interior eigenfunctions are harmonic, this decay is somewhat more subtle in the case of Steklov eigenfunctions. The boundary data \(\varphi _h\) concentrate microlocally on the cosphere bundle of the boundary, \(S^*\partial \Omega \), while the fact that \(-h^2\Delta _gu_h=0\) implies that \(u_h\) must concentrate at the zero section \(\{ (x,\xi ) \in T^*\Omega ; \xi = 0 \}.\) Thus, we think of \(u_h\) as tunnelling from boundary values with \(|\xi |_g=1\) to the interior where \(\xi =0\). For this reason, it is reasonable to view the decay estimates in Theorem 1 as a natural analogue of the well-known Agmon–Lithner estimates (see for example [7]) for Schrödinger eigenfunctions in classically forbidden regions. We note that the assumption that \((\Omega ,g)\) is \(C^{\omega }\) is necessary in Theorem 1 below because real-analyticity allows for accuracy up to exponential errors in h in the pseudodifferential calculus, whereas in the \(C^\infty \) case, one can only work to \(O(h^{\infty })\)-error. In particular, one can only microlocalize modulo such errors. Since Steklov eigenfunctions decay exponentially in h in the interior of \(\Omega \), the usual \(C^{\infty }\) semiclassical calculus of operators is not accurate enough to deal with these functions in a rigorous fashion. Similarly, our subsequent more general results in Theorems 2 and 3 hinge on the microlocal exponential weighted estimate in Proposition 2.5 which also requires real-analyticity to effectively control error terms.

Theorem 1

Let \((\Omega ^{n+1},g)\) be a compact, real-analytic (\(C^{\omega }\)) Riemannian manifold with \(C^{\omega }\) boundary \(\partial \Omega \) and \({\mathcal {D}}: C^{\infty }(\partial \Omega ) \rightarrow C^{\infty }(\partial \Omega )\) be the associated DtN operator. Then for all \(\delta >0\) there exist \(0<\varepsilon _0=\varepsilon _0(\Omega ^n,g,\delta )\) such that for \(\varphi _h\in C^\omega (\partial \Omega )\) with

$$\begin{aligned} (h\mathcal {D}-1)\varphi _h=0,\quad \Vert \varphi _h\Vert _{L^2(\partial \Omega )}=1, \end{aligned}$$

the harmonic extension \(u_{h}(x)\) satisfies the exponential decay estimate

$$\begin{aligned} | \partial _{x}^{\alpha } u_{h}(x) | \le C_{\alpha ,\delta } h^{ -\frac{{n}}{2} + \frac{1}{4} - |\alpha |} \, \exp \Big (- d(x) / h\Big ),\quad d_{\partial \Omega }(x)<\varepsilon _0, \end{aligned}$$
(1.3)

where \(C_{\alpha ,\delta }>0\) is a constant independent of h. In (1.3),

$$\begin{aligned} d(x) = d_{\partial \Omega }(x) +(C_{\Omega ,g}-\delta )d_{\partial \Omega }^2(x), \end{aligned}$$

where \(d_{\partial \Omega }(x)\) is the Riemannian distance to the boundary and

$$\begin{aligned} C_{\Omega ,g}=-\frac{3}{2}+\frac{1}{2}\inf _{(x',\xi ')\in S^*\partial \Omega }Q(x',\xi '). \end{aligned}$$
(1.4)

Here \(Q(x',\xi ')\) is the symbol of the second fundamental form of the boundary \(\partial \Omega .\)

Remark 1.1

The estimate (1.3) is only valid in a small collar neighborhood of \(\partial \Omega \) and indeed, since \(C_{\Omega ,g}\) may be negative, the rate function d(x) may cease to give exponential decay outside a collar neighborhood of \(\partial \Omega \). However, the maximum principle for the Laplace equation on \(\Omega _\varepsilon =\{x\in \Omega \mid d_{\partial \Omega }(x)>\varepsilon \}\) together with Theorem 1 also implies that for any \(\varepsilon >0\), there exists \(C,c>0\) so that

$$\begin{aligned} |u_h(x)|\le Ce^{-c/h},\quad d_{\partial \Omega }(x)>\varepsilon . \end{aligned}$$

Unfortunately, we lose control of the rate function outside a small collar neighborhood of the boundary.

We will see by examining the case of \(\Omega =B(0,R)\subset \mathbb {R}^2\) that the rate of decay in (1.3) is optimal to first order at the boundary. We do not expect to be able to obtain the optimal second order estimate since we are forced to throw away some of the oscillations in u when we apply the Cauchy–Schwarz inequality in (1.22), however, the behavior with respect to Q is optimal as we will see from several examples.

We point out that the estimate in Theorem 1 holds without change if one takes the \(\varphi _h\) to be \(L^2\)-normalized Laplace eigenfunctions on the boundary. Indeed, the bound in Theorem 1 can be adapted to the case of harmonic extensions of eigenfunctions of general elliptic, analytic, self-adjoint h-pseudodifferential operators on M. However, the bounds are somewhat cumbersome to state and we do not pursue this here.

It is worth noting that the proof of Theorem 1 is microlocal and thus the constant \(C_{\Omega ,g}\) can be made to depend on the nearest point in \(\partial \Omega \). In particular, let \((x',x_{{n+1}})\) be Fermi normal coordinates in a collar neighborhood of \(\partial \Omega \) so that \(x_{{n+1}}=d_{\partial \Omega }(x)\). Then the estimate (1.3) holds with d(x) replaced by

$$\begin{aligned} \tilde{d}(x)=x_{{n+1}}+(a_{\Omega ,g}(x')-\delta )x_{{n+1}}^2, \end{aligned}$$

where

$$\begin{aligned} a_{\Omega ,g}(x')=-\frac{3}{2}+\frac{1}{2}\inf _{p\in S^*_{x'}\partial \Omega }Q(p). \end{aligned}$$

1.1 Examples of Steklov Eigenfunctions: Sharpness of d(x) to First Order

We now examine a few examples to illustrate the results of Theorem 1.

1.1.1 The Disk

Let \(\Omega =B(0,R)\subset \mathbb {R}^2\). Then the Steklov eigenvalues are precisely \(\sigma =0,\frac{1}{R},\frac{2}{R}\dots \) with corresponding Steklov eigenfunctions given by

$$\begin{aligned} u^{\pm }_{{k}}=\frac{1}{\sqrt{2\pi R}R^n}r^{{k}}e^{\pm i{{k}}\theta },\quad \sigma =\frac{{{k}}}{R}. \end{aligned}$$
(1.5)

In particular, letting \(h=\sigma ^{-1}={{k}}^{-1}R\),

$$\begin{aligned} u^{\pm }_{{k}}=\frac{1}{\sqrt{2\pi R}}e^{[R\log (1-(R-r)/R))]/h}e^{i\theta /h}. \end{aligned}$$

Therefore, in this case

$$\begin{aligned} d(x)=-R\log (1-d_{\partial \Omega }(x)/R)=\sum _{{{j}}=1}^\infty \frac{[d_{\partial \Omega }(x)]^{{j}}}{{{j}}R^{{{j}}-1}}. \end{aligned}$$

This shows that to first order, the results of Theorem 1 are sharp. Moreover, notice that the second fundamental form of \(\partial B(0,R)\) is given by \(R^{-1}\) and thus, for the disk, the optimal quadratic term is

$$\begin{aligned} \frac{1}{2}\inf _{p\in S^*\partial \Omega }Q(p)[d_{\partial \Omega }(x)]^2 \end{aligned}$$
(1.6)

hence, modulo the \(\frac{3}{2}\) in (1.4), the constant \(C_{\Omega ,g}\) sharp. In particular, as the curvature of the boundary increases, the decay into the interior becomes more rapid.

The case of spheres in higher dimensions is nearly identical if we replace \(e^{\pm in\theta }\) by a spherical harmonic.

1.1.2 Cylinders

Let (Mg) be a real-analytic manifold of dimension n without boundary and \(\Omega =(-1,1)_t\times M_x\) with metric \(dt^2+g(x).\) Then

$$\begin{aligned} \Delta _\Omega =\partial _t^2+\Delta _M. \end{aligned}$$

Let \(\varphi _{{k}}\) be an orthonormal basis for \(L^2(M)\) with

$$\begin{aligned} (-\Delta _M-\lambda _{{k}}^2)\varphi _{{k}}=0. \end{aligned}$$

Then the Steklov eigenfunctions are given by

$$\begin{aligned} u_h(t,x)=\frac{\cosh (\lambda _{{k}}t)}{\cosh (\lambda _{{k}})}\varphi _{{k}}(x),\qquad v_h(x,t)=\frac{\sinh (\lambda _{{k}}t)}{\sinh (\lambda _{{k}})}\varphi _{{k}}(x) \end{aligned}$$

with Steklov eigenvalues \(\sigma _{{k}}=\lambda _{{k}}\tanh (\lambda _{{k}})\) and \(\sigma '_{{k}}=\lambda _{{k}}\coth (\lambda _{{k}})\) respectively. Notice that for \(\lambda _{{k}}\gg 1\),

$$\begin{aligned} \cosh (x)=\frac{1}{2}e^{|x|}+O(e^{-|x|}),\qquad \sinh (x)=\frac{{\text {sgn}}(x)}{2}e^{|x|}+O(e^{-|x|}). \end{aligned}$$

In particular, near \(|t|=1\),

$$\begin{aligned} |u_h(t,x)|=(e^{-\lambda _{{k}}(1-|t|)}+O(e^{-\lambda _{{k}}}))|\varphi _{{k}}(x)|. \end{aligned}$$

Then, notice that

$$\begin{aligned} \sigma _{{k}}=\lambda _{{k}}(1+O(e^{-\lambda _{{k}}})). \end{aligned}$$

So, using Hörmander’s \(L^\infty \) bounds (see e.g., [29, Chapter 7]) we have the estimate

$$\begin{aligned} |u_h(t,x)|=(e^{-\sigma _{{k}}(1-|t|)}+O(e^{-\sigma _{{k}}}))|\varphi _{{k}}(x)|\le C\sigma _{{k}}^{\frac{n-{{1}}}{2}}e^{-\sigma _{{k}}(1-|t|)}. \end{aligned}$$

Indeed, it is not hard to construct examples (e.g., where M is the sphere) so that this estimate is sharp. The analysis is similar for \(v_h\).

In particular, the best possible decay rate is given by

$$\begin{aligned} d(x)=d_{\partial \Omega }(x) \end{aligned}$$

and we again see that the first-order term in Theorem 1 is sharp and that the quadratic term is given by (1.6) since in this case the second fundamental form is 0.

1.1.3 The Annulus

Now, consider \(B(0,1){\setminus } B(0,r_0)\subset \mathbb {R}^2\). Then a simple computation shows that the Steklov eigenvalues are the roots, of

$$\begin{aligned} p_{{k}}(\sigma )=\sigma ^2-\sigma {{k}}\left( \frac{1+r_0}{r_0}\right) \left( \frac{1+r_0^{2{{k}}}}{1-r_0^{2{{k}}}}\right) +\frac{{{k}}^2}{r_0},\quad {{k}}=0,1,\dots \end{aligned}$$

with corresponding eigenfunctions

$$\begin{aligned} u^{\pm }_\sigma (r,\theta )=C_{{{k}},\sigma }e^{\pm in\theta }\left( r^{{k}}+\frac{{{k}}-\sigma }{{{k}}+\sigma }r^{-{{k}}}\right) . \end{aligned}$$
(1.7)

It is easy to show that the roots of \(p_n(\sigma )\) have

$$\begin{aligned} \sigma _{{{k}},1}={{k}}+O({{k}}r_0^{2{{k}}}),\quad \sigma _{{{k}},2}=\frac{{{k}}}{r_0}+O({{k}}r_0^{2{{k}}}). \end{aligned}$$

Then,

$$\begin{aligned} u^{\pm }_{\sigma _{{{k}},1}}=\frac{1}{\sqrt{2\pi }}e^{\pm i{{k}}\theta }r^{{k}}+O(e^{-cv}),\quad u^{\pm }_{\sigma _{{{k}},2}}=\frac{1}{\sqrt{2\pi r_0}}r_0^{{k}}r^{-{{k}}}e^{\pm i{{k}}\theta }+O(e^{-c{{k}}}). \end{aligned}$$

The case of \(u_{\sigma _{{{k}},1}}\) is identical to that for the disk, so we focus on \(u_{\sigma _{{{k}},2}}\). Let \(h=\sigma _{{{k}},2}^{-1}=r_0{{k}}^{-1}+O(e^{-c{{k}}}).\) Then,

$$\begin{aligned} |u^{\pm }_{\sigma _{{{k}},2}}(r,\theta )|=\frac{1}{\sqrt{2\pi r_0}}e^{-r_0\log [1+ (r-r_0)/r_0]/h}(1+O(e^{-c/h})). \end{aligned}$$

Therefore, in this case

$$\begin{aligned} d(x)=-r_0\log (1+d_{\partial \Omega }(x)/r_0)=-\sum _{{{j}}=1}^\infty \frac{[-d_{\partial \Omega }(x)]^{{j}}}{{{j}}r_0^{{{j}}-1}}. \end{aligned}$$

This shows again that to first order, the results of Theorem 1 are sharp. Moreover, notice that the second fundamental form of \(\partial \Omega \) near \(\partial B(0,r_0)\) is given by \(-r_0^{-1}\) and thus, near this boundary component, the optimal quadratic term is again given by (1.6).

1.2 Microlocal Estimates

It is clear from the approximate Poisson formula for \(u_h(x)\) (see (3.1) below) that the exponential eigenfunction decay in the interior \(\Omega \) is closely related to the precise rate of h-microlocal exponential decay of the boundary DtN eigenfunctions \(\varphi _h\) off the cosphere bundle \(S^*\partial \Omega = \{(y,\eta ) \in T^*\partial \Omega ; |\eta |_g =1 \}.\) To derive the requisite bounds, we prove weighted exponential h-microlocal estimates for the associated wave packets \(T(h) \varphi _h\) where \(T(h): C^{\infty }(\partial \Omega ) \rightarrow C^{\infty }(T^*\partial \Omega )\) is a globally defined FBI transform in the sense of [18]. Since these estimates seem of independent interest, we prove them for a rather general class of analytic h-pseudodifferential operators. An important consequence is the following exponential decay estimate for \(T(h) \varphi _h\) off the characteristic variety when \(\varphi _h\) solves \(P(h)\varphi _h=O(e^{-c/h}).\)

We recall that an operator \(P(h) \in Op_{h}(S^{0,k})\) with principal symbol \(p(y,\eta )\) is said to have simple characteristics provided \(dp \ne 0\) on the set \(\{p=0 \}.\) Moreover, \(p(y,\eta )\) is classically elliptic if \(|p(y,\eta )| \ge C' \langle \eta \rangle ^{k}\) for \(|\eta | \ge C\) with constants \(C,C' >0.\) Let \(S^{m,k}_{cla}\) denote the class of classical analytic symbols (see Sect. 2).

Theorem 2

Let \((M^{{n}},g)\) be a compact, closed, real-analytic manifold and \(P(h) \in Op_h (S^{0,k}_{cla})\) be an analytic, h-pseudodifferential operator with real, classically elliptic principal symbol \(p(x,\xi )\) having simple characteristics. Suppose that

$$\begin{aligned} P(h)\varphi _h=O_{L^2}(e^{-c/h}) ,\quad \Vert \varphi _h\Vert _{L^2}=1. \end{aligned}$$

Let \(T(h): C^{\infty }(M) \rightarrow C^{\infty }(T^*M)\) be a globally defined FBI transform as in (2.8) associated with an h-ellptic symbol \(a \in S^{3{{n}}/4,{{n}}/4}_{cla}\) and consider the weight function

$$\begin{aligned} \psi (x,\xi ) = \frac{ \delta \cdot p^2(x,\xi )}{ \langle \xi \rangle ^{2k}}. \end{aligned}$$

Then, provided \(\delta >0\) is a sufficiently small positive constant depending on (Mg),  it follows that for \(h \in (0,h_0(\delta )]\) with \(h_0(\delta )>0\) sufficiently small,

$$\begin{aligned} | e^{\psi /h} T(h) \varphi _h \Vert _{L^2(T^*M)} = O(1). \end{aligned}$$
(1.8)

The main technical ingredient needed for the proof of Theorem 2 is given in Proposition 2.5; it is essentially the manifold analogue of the microlocal exponential weighted estimates in \({\mathbb R}^n\) proved by Martinez [12, 13] and Nakamura [14] (see also [22]). As a direct application of Theorem 2, we prove the first-order exponential decay estimate in Theorem 1 (1.3).

Remark 1.2

Although weaker than Theorem 2 (since no rate of decay is specified), we point out that the following exponential decay estimate is an immediate consequence of Theorem 2.

Corollary 1.3

For fixed \(\varepsilon _0 >0\) consider the cutoff \(\chi _{\varepsilon _0}(x,\xi ) := \chi \Big ( \frac{ p(x,\xi )}{\varepsilon _0} \Big ).\) Then, for any \(\varphi _h\) satisfying the assumptions in Theorem 2, there exists constant \(C(\varepsilon _0)>0\) such that for \(h \le h(\varepsilon _0)\),

$$\begin{aligned} \Vert (1-\chi _{\varepsilon _0}) T(h) \varphi _h \Vert _{L^2(\partial \Omega )} = O(e^{-C(\varepsilon _0)/h}). \end{aligned}$$

We also note that Corollary 1.3 also follows from a rather standard parametrix construction for analytic h-pseudodifferential operators (see Sect. 2.2).

To prove Theorem 1 (1.4), one must estimate the constant \(C_{\Omega ,g}\) in the rate function \(d(x) = d_{\partial \Omega }(x) - C_{\Omega ,g} d^2_{\partial \Omega }(x).\) Unfortunately, the weighted \(L^2\) bound in Theorem 2 for \(T \varphi _h\) does not quite suffice for this since one needs an additional geometric bound for the constant \(\delta >0.\) To achieve this, we must h-microlocally refine the bound in Theorem 2. Global refinement of the decay estimate for \(T(h)\varphi _h\) by specifying sharp constant \(\delta >0\) in the Gaussian seems a rather intractable problem. However, with Corollary 1.3 in hand, we see that away from the characteristic variety, \(T(h)\varphi _h\) is exponentially small. Consequently, instead of attempting to refine the global estimate in Theorem 2, we h-microlocalize to a small neighborhood of the characteristic variety \(p^{-1}(0).\) In order to exploit the real-analyticity of (Mg) and P(h), we choose the neighborhood of M in \(T^*M\) to be the Grauert tube complexification \(M_{\tau }^{{\mathbb C}} \supset M,\) which we assume throughout contains the characteristic variety, \(p^{-1}(0).\) In addition, for the subsequent estimates, it will be important to choose a specific h-microlocal FBI transform \(T_{hol}(h): C^{\infty }(M) \rightarrow C^{\infty }(M_{\tau }^{{\mathbb C}})\) that is compatible with the complex structure on \(M_{\tau }^{{\mathbb C}}\). In view of [10, Theorem 0.1], it is natural to choose \(T_{hol}(h)\) to be the holomorphic continuation of the heat kernel on (Mg) at time \(t = \frac{h}{2}.\)

1.2.1 Motivating Example of FBI Transforms

Recall that the standard FBI transform on \({\mathbb R}^n\) is given by

$$\begin{aligned} T_{{\mathbb R}^n}u(\alpha _x,\alpha _\xi )=2^{-\frac{n}{2}}(\pi h)^{-\frac{3n}{4}}\int e^{-\frac{i}{h}\langle \alpha _x-y,\alpha _\xi \rangle -\frac{1}{2h}|\alpha _x-y|^2}u(y)\mathrm{d}y. \end{aligned}$$
(1.9)

We will introduce two globally defined FBI transforms below, \(T_{hol}\) and \(T_{geo}\). In the case of \({\mathbb R}^n\), these two FBI transforms agree and are given by (1.9). That is, on \({\mathbb R}^n\),

$$\begin{aligned} T_{hol}(\alpha _x-i\alpha _\xi )=T_{geo}(\alpha _x,\alpha _\xi )=T_{{\mathbb R}^n}(\alpha _x,\alpha _\xi ). \end{aligned}$$

Before formally stating our next result, we give some motivation. Consider the simple example of the circle \({\mathbb R}/ 2\pi {\mathbb Z}\) with a flat metric \(g = dx^2.\) That is, consider the boundary for the examples in Sects. 1.1.1 and 1.1.3. The functions \(\varphi _h(x) = e^{ix/h}\) appearing in (1.5) and (1.7) satisfy \((hD_{x}-1)\varphi _h=0.\) In this case, the complexification is

$$\begin{aligned} {\mathbb C}/ 2\pi {\mathbb Z}= \{ \alpha _x + i \alpha _\xi ; \alpha _x + 2\pi \equiv \alpha _x \} \cong T^* ({\mathbb R}/2\pi {\mathbb Z}). \end{aligned}$$

We compute \(T_{hol}\varphi _h\) by extending \(\varphi _h\) smoothly to \({\mathbb R}\) as a solution of \((hD_{x}-1)\varphi _h=0\) or using the definition of \(T_{hol}\) in (1.12). Then,

$$\begin{aligned} T_{hol}\varphi _h(\alpha _x-i\alpha _\xi )=h^{-1/4}e^{-\frac{1}{2h}\alpha _\xi ^2-\frac{1}{2h}}e^{i(\alpha _x-i\alpha _\xi )/h}=h^{-1/4}e^{i\alpha _x/h}e^{-\frac{1}{2h}(\alpha _\xi -1)^2}. \end{aligned}$$

In particular, with \(\psi _{hol}(\alpha ) = \frac{1}{2} (\alpha _\xi - 1)^2,\)

$$\begin{aligned} e^{\psi _{hol}/h}T_{hol}\varphi _h=h^{-1/4}e^{i\alpha _x/h}. \end{aligned}$$

Thus, for any \(\gamma <1\),

$$\begin{aligned} c\le \Vert e^{\frac{\gamma \psi _{hol}}{h}}T_{hol}\varphi _h\Vert _{L^2(T^*S^1)}\le C. \end{aligned}$$
(1.10)

We note that (1.10) is consistent with Theorem 2. However, it is useful to observe that with the precise weight function \(\psi _{hol} = \frac{1}{2}(\alpha _{\xi }-1)^2\) (with \(\gamma =1\)),

$$\begin{aligned} \Vert e^{\frac{\psi _{hol}}{h}}T_{hol}\varphi _h \Vert _{L^\infty (T^*S^1)}= h^{-1/4}. \end{aligned}$$
(1.11)

Thus, with the optimal weight function \(\psi _{hol}\), one has a polynomial gain of \(h^{-1/4}\) in the weighted \(L^2\) mass. The computations for eigenfunctions on higher-dimensional flat tori \({\mathbb R}^n/{\mathbb Z}^n\) are very similar.

To state the h-microlocal refinement of Theorem 2, we let M be a compact, closed, real-analytic manifold of dimension m and \(\widetilde{M}\) denote a Grauert tube complex thickening of M with M a totally real submanifold. By Bruhat-Whitney, \(\widetilde{M}\) can be identified with \(M^{{\mathbb C}}_{\tau } := \{ (\alpha _x, \alpha _\xi ) \in T^*M; \sqrt{2\rho }(\alpha _x,\alpha _\xi ) \le \tau \}\) where \(\sqrt{2\rho } = |\alpha _{\xi }|_g\) is the exhaustion function using the complex geodesic exponential map \( \kappa : M_{\tau }^{{\mathbb C}} \rightarrow \tilde{M}\) with \(\kappa (\alpha ) = \exp _{\alpha _x}( i \alpha _{\xi }).\)

Remark 1.4

We use the notation \(\alpha \) rather than \((x,\xi )\) because it is useful to think of \(\alpha \in \widetilde{M}\) at some times and \(\alpha \in T^*M\) at other times where we identify \(\widetilde{M}\) as a subset of \(T^*M\).

By possibly rescaling the semiclassical parameter h we assume without loss of generality that the characteristic manifold

$$\begin{aligned} p^{-1}(0) \subset M^{{\mathbb C}}_{\tau }. \end{aligned}$$

Now, let \(e^{h\Delta _g/2}\) have Schwartz kernel E(xyh) and \(E^{\mathbb {C}}(\alpha ,y,h)\) denote the holomorphic continuation of E(xyh) to \(M^{\mathbb {C}}_\tau \) in the outgoing x-variables. It is proved in [10] Theorem 0.1 (see also Sect. 2) that the operator given by

$$\begin{aligned} T_{hol}(h)u(\alpha ):=h^{-{{n}}/4}e^{-\rho (\alpha )}\int _M E^{\mathbb {C}}(\alpha ,y,h)u(y)\mathrm{d}y \end{aligned}$$
(1.12)

is an \(L^2\)-normalized, h-microlocal FBI transform defined for \(\alpha \in {\widetilde{M}}.\)

Theorem 3

Under the same assumptions as in Theorem 2 and with \(T=T_{hol}\) is as in (1.12), there exists \(\varepsilon >0\) small enough so that with \(\gamma <1/2\),

$$\begin{aligned} \Vert e^{\gamma a_0p^2/h}T_{hol}(h)\varphi _h\Vert _{L^2(\{|p|\le \varepsilon \})}=O(1), \quad a_0=\frac{1}{2|dp|_{g_s}^2} \end{aligned}$$
(1.13)

where \(g_s\) is the Sasaki metric on TM (see for example [2, Chapter 9]).

Moreover, there exists \(\psi _{hol} \in C^{\infty }(M_{\tau }^{{\mathbb C}})\) with

$$\begin{aligned} \psi _{hol}(\alpha ) = p^2(\alpha )\big ( \, a_0(\alpha )+O(p(\alpha )) \,\big ), \end{aligned}$$

where \(a_0\) is given in (1.13) such that for \(\varepsilon >0\) sufficiently small,

$$\begin{aligned} \Vert e^{\psi _{hol}/h}T_{hol}(h)\varphi _h\Vert _{L^2(\{|p|\le \varepsilon \})}=O(h^{-1/4}). \end{aligned}$$
(1.14)

We note that the example of the Laplace eigenfunctions on the circle (see (1.10) and (1.11) above), shows that the upper bound in Theorem 3 (1.14) is sharp.

Remark 1.5

We use (1.13) instead of (1.14) in the proof of Theorem 1. Using (1.14) results in better estimates as soon as \(d(x)\gg ch\log h^{-1}\), but for simplicity and since the function d(x) that we obtain is still not sharp, we do not state these estimates here.

1.3 Sketch of the Proof of Theorem 1

Let \({\mathcal {P}}: C^{\infty }(\partial \Omega ) \rightarrow C^{\infty }(\Omega )\) be the Poisson operator for the boundary value problem in (1.2), so that \( u_h(x) = {\mathcal {P}} \varphi _h(x).\) The first step in the proof of Theorem 1 amounts to understanding the microlocal structure of the Poisson operator, \({\mathcal {P}}\) following the analysis in [19].

1.3.1 Microlocal Analysis of the Poisson Operator

In view of [19, Section 3] and the fact that \(\varphi _h\) is microlocally supported away from the zero section one can write

$$\begin{aligned} u_h(x)&= {\mathcal {P}} \varphi _h (x) = U(h) \varphi _h(x) + O(e^{-C/h}), \end{aligned}$$
(1.15)

where \(U(h): C^{\infty }(\partial \Omega ) \rightarrow C^{\infty }(\Omega )\) is a semiclassical, complex-phase h-Fourier integral operator supported near diagonal. In terms of Fermi coordinates \((x_{{n+1}},x')\) in a collar neighborhood \(U= \{ (x',x_{{n+1}}); x_{{n+1}} \ge 0 \}\) of \(\partial \Omega = \{ x_{{n+1}} =0 \}\), U(h) has Schwartz kernel

$$\begin{aligned} K(x,y',h)= & {} (2\pi h)^{-n} \int _{{\mathbb R}^{n}} e^{ i \langle x' -y', {\xi '} \rangle /h} \, e^{- {{\Psi }}(x_{{n+1}},x',{\xi '}) /h} \,\nonumber \\&a(x,y',{\xi '},h) \, \chi (x'-y') \, d{\xi '}. \end{aligned}$$
(1.16)

Here,

$$\begin{aligned} {{\Psi }}(x_{{n+1}},x',\xi ')=x_{{n+1}}|\xi '|_{x'}+\frac{x_{{n+1}}^2(Q(x',\xi ')+i\left\langle \partial _{x'}|\xi '|_{x'},\xi '\right\rangle _{x'})}{2|\xi '|_{x'}}+O(x_{{n+1}}^3|\xi '|_{x'}) \end{aligned}$$

satisfies a complex eikonal equation (3.3) and \(a(x,y',{\xi '},h)\) is a semiclassical analytic symbol as in (2.3). Note that the error is exponentially decreasing since we work in the analytic setting.

In the Euclidean case, we note that one can derive the semiclassical Poisson formula (1.15) in an elementary fashion directly from the potential layer formulas using residue computations. For the benefit of the reader, we outline the argument here. Let \(\Omega \subset {\mathbb R}^{n}\) be a bounded Euclidean domain with real-analytic boundary \(\partial \Omega .\) Let \(G(z,z') \in {\mathcal {D}}'({\mathbb R}^{n} \times {\mathbb R}^{n}) \) be the free Green’s functions with \( \Delta _{z'} G(z,z') = \delta (z-z').\) From Green’s formula and the DtN eigenfunction condition, one gets

$$\begin{aligned} u_{h}(z)&= h^{-1} \int _{\partial \Omega } G(z,z') \varphi _{h}(z') d\sigma (z') - \int _{\partial \Omega } N(z,z') \varphi _{h}(z') d\sigma (z'). \end{aligned}$$
(1.17)

with \(N(z,z') = \partial _{\nu (z')} G(z,z'),\,\,\, (z,z') \in {\mathbb R}^{{{n+1}}} \times \partial \Omega .\) Writing \(G(z,z')\) and \(N(z,z')\) as Fourier integrals and rescaling the frequency variables \(\xi \rightarrow h^{-1} \xi \) one rewrites (1.18) in the form

$$\begin{aligned} u_h(z)&= (2\pi h)^{-{({n+1})}}h \int _{{\mathbb R}^{{{n+1}}}} \int _{\partial \Omega } e^{i \langle z -z', \xi \rangle /h } ( |\xi |^2 + i 0)^{-1} \, \varphi _h(z') d\sigma (z') \mathrm{d}\xi \nonumber \\&\quad + i (2\pi h)^{-{({n+1})}}h \int _{{\mathbb R}^{{{n+1}}}} \int _{\partial \Omega } e^{i \langle z-z',\xi \rangle /h} \langle \nu (z'), \xi \rangle (|\xi |^2 + i0)^{-1} \varphi _h(z') d\sigma (z') \mathrm{d}\xi . \end{aligned}$$
(1.18)

Let \(\chi \in C^{\infty }_0({\mathbb R})\) with \(\chi =1\) near the origin and supp \(\chi \subset [-\varepsilon _0,\varepsilon _0].\) We note that by making a change a change of contour \(\xi \mapsto \xi + i \delta \xi \) in (1.18) with \(0< \delta <1\) one can insert a spatial cutoff \(\chi (|z-z'|)\) in both integrals modulo an \(O(e^{-C/h})\) error. Next, we introduce convenient coordinates in a tubular neighborhood \(U_{\partial \Omega }\) of the boundary. Given a local \(C^{\omega }\) parameterization of the boundary \(q: U \rightarrow \partial \Omega \) with \(U \subset {\mathbb R}^n\) open, we write locally

$$\begin{aligned} z = q(x') + x_{{{n+1}}} \nu (x')\,\,\, \text {and} \,\, z' = q(y') \end{aligned}$$

By choosing \(\varepsilon _0>0\) sufficiently small, we can assume that z and \(z'\) lie in the same local coordinate chart. In terms of these new coordinates, one can rewrite the phase function

$$\begin{aligned} \langle z - z',\xi \rangle= & {} \langle q(x') - q(y'), \xi \rangle + x_{{{n+1}}} \langle \nu (x'), \xi \rangle \\= & {} \langle x' - y', dq^t(x',y') \xi \rangle + x_{{{n+1}}} \langle \nu (x'), \xi \rangle . \end{aligned}$$

We make the affine change of variables in (1.18) given by \(\xi \mapsto (\eta ',\eta _{{{n+1}}})\) where

$$\begin{aligned} \eta _{{{n+1}}} = \langle \nu (x'), \xi \rangle , \,\,\, \eta ' = dq^t(x',y') \xi \end{aligned}$$

Then, for \(x= (x',x_{{n+1}}) \in U_{\partial \Omega },\) using the fact that the DtN eigenfunctions are h-microlocally \(O(e^{-C/h})\) near the zero section \({\eta }' = 0\) (see Proposition 4.6), one can write

$$\begin{aligned}&u_h(x) + O(e^{-C(\varepsilon )/h}) \nonumber \\&\quad =(2\pi h)^{-{({n+1})}}h \int _{|\eta '| > \varepsilon } \int _{\partial \Omega } e^{i \langle x'-y', \eta ' \rangle /h } e^{i x_{{n+1}} \eta _{{n+1}} /h} \,b(x,y',\eta ) \chi (x'-y') \varphi _h\nonumber \\&\qquad (y') \mathrm{d}y' \mathrm{d}\eta _{{n+1}} \mathrm{d}\eta ', \end{aligned}$$
(1.19)

where \( b(x,y',\eta ) = (1 + i \eta _{{n+1}}) ( \eta _{{n+1}}^2 + |\eta '|_x^2 + i 0)^{-1} a(x',y').\ \)

For \(\eta ' \ne 0,\) a residue computation gives \( \int _{{\mathbb R}} e^{i x_{{n+1}} \eta _{{n+1}}/h} ( \eta _{{n+1}}^2 + |\eta '|_{x} + i0)^{-1} \mathrm{d}\eta _{{n+1}} = \frac{ \pi e^{- x_{{n+1}} |\eta '|_x/h}}{ |\eta '|_x}\) and similarily, \( \int _{{\mathbb R}} e^{i x_{{n+1}} \eta _{{n+1}}/h} \eta _{{n+1}} ( \eta _{{n+1}}^2 + |\eta '|_{x} + i0)^{-1} \mathrm{d}\eta _{{n+1}} = \pi e^{- x_{{n+1}} |\eta '|_x/h}.\) Substitution of these integral formulas in (1.19) gives modulo \(O(e^{-C(\varepsilon )/h})\) error,

$$\begin{aligned} u_{h}(x)&= (2\pi h)^{-{{n}}} \int _{|\eta '| >\varepsilon } \int _{\partial \Omega } e^{i \langle x'-y',\eta ' \rangle /h} e^{- x_{{n+1}} |\eta '|_x /h} |\eta '|_x^{-1} a(x',y') \\&\quad \times ( 1 + i |\eta '|_{x}^{-1}) \varphi _{\lambda }(y') \mathrm{d}y' \mathrm{d}\eta '. \end{aligned}$$

This is consistent with the general formula in (1.16).

1.3.2 Microlocal Lift of the Poisson Representation (1.15)

Given the representation of \(u_h\) in (1.15) in terms of semiclassical, complex-phase h-Fourier integral operator \(U(h): C^{\infty }(\partial \Omega ) \rightarrow C^{\infty }(\Omega ),\) the key idea in the proof of Theorem 1 is to lift (1.15) to the cotangent bundle of the boundary \(T^*\partial \Omega \) and then apply the weighted estimate in Theorem 2 to give the first-order approximation for the Steklov decay rate function d(x) in Theorem 1 (1.3). The quadratic term in d(x) is then bounded from above to prove Theorem 1 (1.4) using the refined h-microlocal weighted estimates in Theorem 3.

Roughly speaking, we do this as follows: Viewing \(x \in \Omega \) as parameters, we consider the family of functions \(K_{x,h} \in C^{\infty }(\partial \Omega )\) with

$$\begin{aligned} K_{x,h}(y') := K(x,y',h). \end{aligned}$$

Then, (1.15) can be written in the form

$$\begin{aligned} u_{h}(x) = \langle K_{x,h}, \overline{\varphi _h} \rangle _{L^2(\partial \Omega )} + O(e^{-C/h}). \end{aligned}$$
(1.20)

To lift (1.20) we let \(T(h): C^{\infty }(\partial \Omega ) \rightarrow C^{\infty }(T^*\partial \Omega )\) be an FBI transform in the sense of Sjöstrand [18] and \(S(h): C^{\infty }(T^*\partial \Omega ) \rightarrow C^{\infty }(\partial \Omega )\) be a left-parametrix with

$$\begin{aligned} S(h) T(h) = I + R(h), \end{aligned}$$

and R(h) exponentially small in the sense that

$$\begin{aligned} | \partial _{x}^{\alpha } \partial _{y}^{\beta } R(x,y)| = O_{\alpha ,\beta } (e^{-C/h}). \end{aligned}$$

Given the weight function \(\psi \in C^{\infty }(T^*\partial \Omega )\) in Theorem 2, one can write

$$\begin{aligned} u_{h}(x) = \langle e^{-\psi /h} S(h)^t K_{x,h}, \overline{e^{\psi /h} T(h) \varphi _h} \rangle _{L^2(T^*\partial \Omega )} + O(e^{-C/h}). \end{aligned}$$
(1.21)

Using the bound

$$\begin{aligned} \Vert e^{\psi /h} T(h) \varphi _h \Vert _{L^2(T^*M)} = O(1) \end{aligned}$$

in Theorem 2 and applying Cauchy–Schwarz in (1.21) one gets

$$\begin{aligned} |u_{h}(x) | = O(1) \, \Vert e^{-\psi /h} S(h)^t K_{x,h} \Vert _{L^2(T^*\partial \Omega )} + O(e^{-C/h}). \end{aligned}$$
(1.22)

Finally, a direct analysis of the first term on the RHS of (1.22) using the method of analytic stationary phase yields the bounds (1.3) and (1.4) in Theorem 1. We refer to Sect. 3 for a detailed proof of Theorem 1 using the weighted bounds in Theorems 2 and 3.

Remark 1.6

Both Theorems 2 and 3 have other applications to eigenfunction bounds, including the problem of obtaining geometric rates of decay for eigenfunctions in subdomains of configuration space M that correspond to classically forbidden regions that are geometrically more refined than the classical Agmon–Lithner estimates. Specific examples include (but are not limited to) joint eigenfunctions for quantum completely integrable (QCI) eigenfunctions. We hope to return to this elsewhere.

1.4 Outline of the Paper

In Sect. 2 we discuss the eigenfunction mass microlocalization results for the eigenfunctions \(\varphi _h\). The long range exponential decay estimates are proved in Proposition 2.3 and the short-range exponential weighted estimates near the characteristic variety (and inside the Grauert tube \(M_{\tau }^{{\mathbb C}}\)) are proved in Proposition 2.5. These estimates are combined to prove Theorem 2 in Sect. 2.6. The h-microlocally refined weighted estimates for \(T_{hol}(h) \varphi _h\) along with the proof of Theorem 3 are taken up in Sect. 2.7. In Sect. 3, the exponential weighted estimates in Sects. 2.6 and 2.7 are used to prove the decay estimates in Theorem 1 for the Steklov eigenfunctions. In Sect. 4, we prove the necessary h-microlocal exponential decay estimates for the Steklov eigenfunctions near the zero section of \(T^*\partial \Omega .\) This is necessary since the semiclassical DtN operator \( h {\mathcal {D}}: C^{\infty }(\partial \Omega ) \rightarrow C^{\infty }(\partial \Omega )\) fails to be an h-analytic pseudodifferential operator microlocally near the zero section.

2 Eigenfunction Mass Microlocalization

Let M be a compact, closed, real-analytic manifold of dimension m and \(\widetilde{M}\) denote a Grauert tube complex thickening of M with M a totally real submanifold. By Bruhat-Whitney, \(\widetilde{M}\) can be identified with \(M^{{\mathbb C}}_{\tau } := \{ (\alpha _x, \alpha _\xi ) \in T^*M; \sqrt{\rho }(\alpha _x,\alpha _\xi ) \le \tau \}\) where \(\sqrt{2\rho } = |\alpha _{\xi }|_g\) is the exhaustion function \(M^{{\mathbb C}}_{\tau }\), where we identify \(\widetilde{M}\) with \(M_{\tau }^{{\mathbb C}}\) using the complexified geodesic exponential map \( \kappa : M_{\tau }^{{\mathbb C}} \rightarrow \tilde{M}\) with \(\kappa (\alpha ) = \exp _{\alpha _x,{\mathbb C}}( i \alpha _{\xi })\) Viewed on \(\widetilde{M}\), the function \(\sqrt{\rho }(\alpha ) = \frac{-i}{2\sqrt{2}} r_{{\mathbb C}}(\alpha ,\bar{\alpha }),\) which satisfies homogeneous Monge–Ampere and its level sets exhaust the complex thickening \(\widetilde{M}\) (see Remark 2.6 and [5] for further details).

The example of the round sphere To illustrate these basic complex analytic entities, we consider the case of the n-dimensional round sphere

$$\begin{aligned} M = \{(x_1,...,x_{n+1}) \in {\mathbb R}^{n+1}; x_1^2 + \cdots + x_{n+1}^2 = 1 \} \subset {\mathbb R}^{n+1}. \end{aligned}$$

The complexification of M is the quadric

$$\begin{aligned} \widetilde{M}= & {} \{(z_1,...,z_{n+1}) \in {\mathbb C}^{n+1}; z_1^2 + \cdots z_{n+1}^2 = 1 \} \nonumber \\= & {} \{ (x,\xi ) \in {\mathbb R}^{2(n+1)} ; |x|^2 - |\xi |^2 = 1, \,\, \langle x, \xi \rangle = 0 \}. \end{aligned}$$
(2.1)

The Riemannian exponential map written in terms of affine ambient coordinates on \({\mathbb R}^{n+1}\) is

$$\begin{aligned} \exp _{x}(\xi ) = \cos (|\xi |) x + (\sin |\xi |) \frac{\xi }{|\xi |}, \end{aligned}$$

where \(\xi \in T_xM\), so \(\xi \in {\mathbb R}^{n+1}\) is orthogonal to x. The complexification of \(\exp _x(\xi )\) is then given by

$$\begin{aligned} \exp _{x,{\mathbb C}}( i \xi ) = ( \cosh |\xi | ) x + i (\sinh |\xi |) \frac{\xi }{|\xi |}. \end{aligned}$$

The distance function \(r(x,y) = 2 \sin ^{-1} \Big ( \frac{|x-y|}{2} \Big )\) complexifies to

$$\begin{aligned} r_{{\mathbb C}}(z,w) = 2 \sin ^{-1} \Big ( \frac{ \sqrt{ (z-w)^2} }{2} \Big ) \end{aligned}$$

and the associated exhaustion function on the complexification is

$$\begin{aligned} \frac{-i}{2\sqrt{2}} r_{{\mathbb C}}(z,\bar{z}) = \frac{1}{\sqrt{2}} \, \sinh ^{-1}( |{{\text {Im}}\,}z|), \quad z \in \tilde{M}. \end{aligned}$$

Pulling \(r_{{\mathbb C}}(z,\bar{z})\) back to \(T^*{\mathbb {S}}^2\) via the complexified exponential map gives

$$\begin{aligned} \sqrt{\rho }(z) = \exp _{x,{\mathbb C}}^{-1} \Big ( - \frac{i}{2\sqrt{2}} r_{{\mathbb C}}(z,\bar{z}) \Big ) = \frac{ |\xi |}{\sqrt{2}}. \end{aligned}$$

In terms of local coordinates \(\alpha =(\alpha _x,\alpha _{\xi }) \in T^*{\mathbb {S}}^2,\) this just gives \(\sqrt{\rho }(\alpha ) = \frac{1}{\sqrt{2}} |\alpha _{\xi }|_g.\) Of course, the multiplicative factor of \(\frac{-i}{2\sqrt{2}}\) above is just a computationally convenient normalization that we choose to adopt here.

By possibly rescaling the semiclassical parameter h we assume without loss of generality that the characteristic manifold

$$\begin{aligned} p^{-1}(0) \subset M^{{\mathbb C}}_{\tau }. \end{aligned}$$

We will also have to consider a complexification of \(T^*M\) of the form

$$\begin{aligned} \widetilde{T^*M}:= \left\{ \alpha ; |{{\text {Im}}\,}\alpha _x | < \tau , \,\, |{{\text {Im}}\,}\alpha _{\xi }| \le \frac{1}{C} \langle \alpha _\xi , \rangle \right\} \end{aligned}$$
(2.2)

where \(C \gg 1\) is a sufficiently large constant and \(T^*M \subset \widetilde{T^*M}\) is then a totally real submanifold.

We recall that a complex m-dimensional submanifold, \(\Lambda ,\) of \(\widetilde{T^*M}\) is said to be I-Lagrangian if it is Lagrangian with respect to

$$\begin{aligned} {{\text {Im}}\,}\omega = {{\text {Im}}\,}( d\alpha _x \wedge d \alpha _{\xi }), \end{aligned}$$

where \(\omega = d\alpha _{x} \wedge d \alpha _{\xi }\) is the complex symplectic form on \(\widetilde{T^*M}\).

Let \(U\subset T^*M\) be open. Following [18], we say that \(a \in S^{m,k}_{cla}(U)\) provided \(a \sim h^{-m} (a_0 + h a_1 + \dots )\) in the sense that

$$\begin{aligned} \begin{aligned} \partial _{x}^{{l_1}} \partial _{\xi }^{{l_2}} \overline{\partial }_{(x,\xi )} a&= O_{{l_1},{l_2}}(1) e^{- \langle \xi \rangle /Ch}, \quad (x,\xi )\in U, \\ \Big | a - h^{-m} \sum _{0 \le j \le \langle \xi \rangle /C_0 h} h^{j} a_j \Big |&= O(1) e^{- \langle \xi \rangle /C_1 h},\\ |a_j|&\le C_0 C^{j} \, j ! \, \langle \xi \rangle ^{k-j},\qquad (x,\xi )\in U. \end{aligned} \end{aligned}$$
(2.3)

We sometimes write \(S^{m,k}_{cla}=S^{m,k}_{cla}(T^*M)\).

Following [19], we also define the notion of a homogeneous analytic symbol of order k and write \(a\in S^k_{ha}\) provided that there exist holomorphic functions \(a_k\) on a fixed complex conic neighborhood of \(T^*M{\setminus }\{0\}\) homogeneous of degree k in \(\xi \) so that there exists \(C_0>0\) so that

$$\begin{aligned} \left| a_{k-j}\left( x,\frac{\xi }{|\xi |}\right) \right| \le {C_0}^{j+1}j^j,\qquad j\ge 0 \end{aligned}$$
(2.4)

and for every \(C_1>0\) large enough, there exists \(C_2>0\) so that

$$\begin{aligned} \left| a(x,\xi )-\sum _{0\le j\le |\xi |/C_1}a_{k-j}(x,\xi )\right| \le C_2e^{-|\xi |/C_2},\quad |\xi |\ge 1. \end{aligned}$$
(2.5)

We say that an operator A(h) is a semiclassical analytic pseudodifferential operator of order m, k if its kernel can be written as \({A}(x,y;h)=K_{{1}}(x,y;h)+R_{{1}}(x,y;h)\) where for all \(\alpha ,\beta \),

$$\begin{aligned} |\partial _x^\alpha \partial _y^\beta R_{{1}}(x,y{;h})|\le C_{\alpha \beta } e^{-c_{\alpha \beta }/h}, \,\,\, c_{\alpha \beta } >0, \end{aligned}$$

and

$$\begin{aligned} K_{{1}}(x,y;h)=\frac{1}{(2\pi h)^{{n}}}\int e^{\frac{i}{h}\left\langle x-y,\xi \right\rangle }a(x,\xi ,h)\chi (|x-y|)\mathrm{d}\xi , \end{aligned}$$

where \(\chi \in C_c^\infty (\mathbb {R})\) is 1 near 0 and \(a\in S^{m,k}_{cla}\). We say A is h-elliptic if \(|a_0(x,\xi )|>ch^{-m}\langle \xi \rangle ^k\) where \(a_0\) is from (2.3). Recall also that A is classically elliptic if there is \(C>0\) so that if \(|\xi |>C\), \(|a_0(x,\xi )|>C^{-1}h^{-m}|\xi |^k\).

We say that an operator, B is a homogeneous analytic pseudodifferential operator of orderk if its kernel can be written as \({B}(x,y)=K_{{2}}(x,y)+R_{{2}}(x,y)\) where \(R_{{2}}(x,y)\) is real analytic and

$$\begin{aligned} K_{{2}}(x,y)=\frac{1}{(2\pi )^{{n}}}\int e^{i\left\langle x-y,\xi \right\rangle }{b}(x,\xi )\chi (|x-y|)\mathrm{d}\xi \end{aligned}$$

for some \({b}\in S^k_{ha}\).and \(\chi \in C_c^\infty (\mathbb {R})\) is 1 near 0. We say B is elliptic if there exists \(c>0\) so that \(b_k>c|\xi |^k\) on \(|\xi |\ge 1\) where \(b_k\) is from (2.4) and (2.5). For more details on the calculus of analytic pseudodifferential operators, we refer the reader to [17].

As in [18], given an h-elliptic, semiclassical analytic symbol \(a \in S^{3{{n}}/4,{{n}}/4}_{cla}(M \times (0,h_0]),\) we consider an intrinsic FBI transform \(T(h):C^{\infty }(M) \rightarrow C^{\infty }(T^*M)\) of the form

$$\begin{aligned} T u(\alpha ;h) = \int _{M} e^{i\varphi (\alpha ,y)/h} a(\alpha ,y,h)\chi ( \alpha _x, y) u(y) \, \mathrm{d}y \end{aligned}$$
(2.6)

with \(\alpha = (\alpha _x,\alpha _{\xi }) \in T^*M\) in the notation of [18].

Remark 2.1

The normalization \(a\in S^{3n/4,n/4}_{cla}\) appears so that T is \(L^2\) bounded with uniform bounds as \(h\rightarrow 0\) [18].

The phase function is required to satisfy \(\varphi (\alpha ,\alpha _x) = 0, \, \partial _y \varphi (\alpha ,\alpha _x) = - \alpha _{\xi }\) and

$$\begin{aligned} {{\text {Im}}\,}(\partial _y^2 \varphi )(\alpha ,\alpha _x) \sim |\langle \alpha _{\xi } \rangle | \, {\text {Id}}. \end{aligned}$$

Given \(T(h) :C^{\infty }(M) \rightarrow C^{\infty }(T^*M)\) it follows by an analytic stationary phase argument [18] that one can construct an operator \(S(h): C^{\infty }(T^*M) \rightarrow C^{\infty }(M)\) of the form

$$\begin{aligned} S v(x;h) = \int _{T^*M} e^{-i \, \overline{\varphi (x,\alpha )} /h} b(x,\alpha ,h) v(\alpha ) \, d\alpha \end{aligned}$$
(2.7)

with \(b \in S^{3{{n}}/4,{{n}}/4}_{cla}\) such S(h) is a left-parametrix for T(h) in the sense that

$$\begin{aligned} S(h) T(h) = {\text {Id}}+ R(h),\qquad \partial _{x}^{\alpha } \partial _{y}^{\beta } R(x,y,h) = O_{\alpha , \beta }(e^{-C/h}). \end{aligned}$$

We use two invariantly defined FBI transforms. The first transform \(T_{geo}(h): C^{\infty }(M) \rightarrow C^{\infty }(T^*M)\) is defined using only the Riemannian structure of (Mg) and has phase function

$$\begin{aligned} \varphi (\alpha ,y) = i \exp _{y}^{-1}(\alpha _x) \cdot \alpha _{\xi } - \frac{1}{2} r^{2}(\alpha _{x},y) \langle \alpha _{\xi } \rangle . \end{aligned}$$
(2.8)

Here, \(r(\cdot ,\cdot )\) is geodesic distance and \(\chi (\alpha _x,y) = \chi _0(r(\alpha _x,y))\) where \(\chi _0: {\mathbb R}\rightarrow [0,1]\) is an even cutoff with supp \(\chi _0 \subset [-inj(M,g), inj(M,g)]\) and \(\chi _0(r) =1\) when \(|r| < \frac{1}{2} inj(M,g).\)

The transform \(T_{geo}(h)\) will be used to derive h-microlocal exponential decay outside the Grauert tube \(M_{\tau }^{{\mathbb C}}\) and far from the characteristic variety \(p^{-1}(0).\) We will refer to the corresponding estimates as long-range.

To estimate h-microlocal eigenfunction mass inside the Grauert tube \(M_{\tau }^{{\mathbb C}}\) containing \(p^{-1}(0)\) we use instead another FBI-transform \(T_{hol}(h)\) which is defined in terms of the holomorphic continuation of the heat operator \(e^{t \Delta _g}\) at time \(t = h/2.\) We refer to the corresponding estimates as short-range.

Before continuing, we briefly recall here some background on the operator \(T_{hol}(h): C^{\infty }(M) \rightarrow C^{\infty }(M_{\tau }^{{\mathbb C}})\) and refer the reader to [10] for further details.

2.1 Complexified Heat Operator on Closed, Compact Manifolds

Consider the heat operator of (Mg) defined at time h / 2 by

$$\begin{aligned} E_{h}=e^{\frac{h}{2}\Delta _g}:C^{\infty }(M) \rightarrow C^{\infty }(M). \end{aligned}$$

By a result of Zelditch [25, Section 11.1], the maximal geometric tube radius \(\tau _{\max }\) agrees with the maximal analytic tube radius in the sense that for all \( 0<\tau < \tau _{\max }\), all the eigenfunctions \(\varphi _j\) extend holomorphically to \(M_\tau ^{\mathbb C}\) (see also [10, Prop. 2.1]). In particular, the kernel \(E(\cdot ,\cdot ;h)\) admits a holomorphic extension to \(M_\tau ^{\mathbb C}\times M_\tau ^{\mathbb C}\) for all \(0<\tau < \tau _{\max }\) and \(h \in (0,1)\), [10, Prop. 2.4]. We denote the complexification by \(E_h^{\mathbb C}( \cdot , \cdot )\). To recall asymptotics for \(E^{{\mathbb C}}_h\) we note that the squared geodesic distance on M

$$\begin{aligned} r^2(\cdot , \cdot ): M \times M \rightarrow {\mathbb R}\end{aligned}$$

holomorphically continues in both variables to \(M_{\tau } \times M_{\tau }\) in a straightforward fashion. More precisely, \(0<\tau <\tau _{\max }\), there exists a connected open neighborhood \(\tilde{\Delta }\subset M_\tau ^{\mathbb C}\times M_\tau ^{\mathbb C}\) of the diagonal \(\Delta \subset M \times M\) to which \(r^2(\cdot , \cdot )\) can be holomorphically extended [10, Corollary 1.24]. We denote the holomorphic extension by \(r_{\mathbb C}^2(\cdot ,\cdot ) .\) Moreover, one can easily recover the exhaustion function \(\sqrt{\rho _g}(\alpha _z)\) from \(r_{{\mathbb C}}\); indeed, \(\rho _g(\alpha _z)=-r^2_{\mathbb C}(\alpha _z, \bar{\alpha _z})\) for all \(\alpha _z \in M_\tau ^{\mathbb C}\).

To analyze the asymptotic behavior of \(E_h^{{\mathbb C}}(\alpha _z,y)\) with \((\alpha _z,y) \in M^{\mathbb C}_\tau \times M\), we split the kernel into two pieces where

  1. (i)

    the point \((\pi _{_M} \alpha _z,y) \in M \times M\) is close to the diagonal in terms of inj and the Grauert tube radius \(\tau \),

  2. (ii)

    the point \((\pi _{_M}\alpha _z ,y) \in M \times M\) is relatively far from the diagonal in terms of inj and \(\tau .\)

To control the behavior of the complexified heat kernel for a pair of points \((\pi _{_M} \alpha _z,y) \in M \times M\) that are relatively close or far from the diagonal, we need the following result [10].

Proposition 2.2

There exist \(0<\tau _0 \le \tau _{\max }\) and positive constants \(\beta , \delta _0, h_0\) and C, depending only on \(\tau _0>0\), such that for \(0<\tau \le \tau _0, 0<\delta \le \delta _0\) and \((\alpha _z,y)\in M_{\tau }^{{\mathbb C}}\times M\), the following is true:

  1. (i)

      When \(r(\pi _{M} \alpha _z, y)<\delta \) and \(h \in (0,h_0],\)

    $$\begin{aligned} E_h^{\mathbb C}(\alpha _z,y)=e^{-\frac{r^2_{\mathbb C}(\alpha _z,y)}{2h}} a^{\mathbb C}(\alpha _z,y; h) + O (e^{-\beta /h}). \end{aligned}$$
    (2.9)

    Here, \(a^{\mathbb C}(\alpha _z,y;h)\) is the polyhomogeneous sum

    $$\begin{aligned} a^{{\mathbb C}}(\alpha _z,y; h):=(2\pi h)^{-{{n}}/2} \sum _{0\le k\le D/h} a^{\mathbb C}_k(\alpha _z,y)h^k, \end{aligned}$$
    (2.10)

    where the \(a_k^{\mathbb C}\)’s denote the analytic continuation of the coefficients appearing in the formal solution of the heat equation on (Mg)

  2. (ii)

      There exists \(C>0\) so that when \(r(\pi _{M} \alpha _z, y)>\frac{\delta }{2}\) and \(h \in (0,1),\)

    $$\begin{aligned} \left| E_h^{\mathbb C}(\alpha _z,y) \right| \le C \;e^{- \frac{\delta ^2}{{C} h}}, \end{aligned}$$
    (2.11)

    where C is a positive constant depending only on (Mg).

From now on, we always carry out our analysis in the complex Grauert tubes \(M_\tau ^{\mathbb C}\) with \(0<\tau \le \tau _{{0}},\) where in view of Proposition 2.2, we have good control of the complexified heat kernel, \(E^{\mathbb C}_h(\cdot ,y)\) for \(y \in M\).

For \((\alpha _z,y) \in M_{\tau }^{{\mathbb C}} \times M\) with \(r(\mathrm Rez,y) <\varepsilon \) with \(\varepsilon >0\) small, one can show that the function \(y \mapsto - \mathrm Rer^2_{{\mathbb C}}(\alpha _z,y)\) attains a non-degenerate maximum at \(y = \mathrm Rez\). The corresponding strictly plurisubharmonic weight is the square of the exhaustion function given by

$$\begin{aligned} 2\rho (\alpha ) = {-} \mathrm Rer^2_{{\mathbb C}}(\alpha _z, \mathrm Rez) = {-} \frac{1}{4} r_{{\mathbb C}}^2(\alpha _z, \bar{\alpha _z}) = |\alpha _{\xi }|_{\alpha _x}^2, \end{aligned}$$

where

$$\begin{aligned} \alpha _z = \exp _{\alpha _x} ( -i \alpha _{\xi }). \end{aligned}$$

Using this observation and the expansion in Proposition 2.2 it is proved in [10, Theorem 0.1] that the operator \(T_{hol}(h): C^{\infty }(M) \rightarrow C^{\infty }(M_{\tau }^{{\mathbb C}})\) given by

$$\begin{aligned} T_{hol} \varphi _h (\alpha )= & {} h^{-{{n}}/4} \int _{M} e^{ [ - r_{{\mathbb C}}^2(\alpha _z,y)/2 - \rho (\alpha _z)]/h} a^{{\mathbb C}}(\alpha _z,y,h) \chi \nonumber \\&(\alpha _x,y) \varphi _h(y) \mathrm{d}y, \quad \alpha \in M_{\tau }^{{\mathbb C}} \end{aligned}$$
(2.12)

is also an FBI transform in the sense of (2.6) with amplitude \(a \in S^{m/2,0}_{cla}\) and phase function

$$\begin{aligned} \varphi (\alpha ,y)&= i \Big ( \, \frac{ r_{{\mathbb C}}^2(\alpha _z,y)}{2} + \rho (\alpha ) \, \Big ). \end{aligned}$$
(2.13)

In (2.12) the multiplicative factor \(h^{-{{n}}/4}\) is added to ensure \(L^2\)-normalization so that \(\Vert T_{hol} \varphi _h \Vert _{L^2(M_{\tau }^{{\mathbb C}})} \approx 1.\) The fact that the transform \(T_{hol}\) is compatible with the complex structure of the Grauert tube \(M_{\tau }^{{\mathbb C}}\) will be used a crucial way in the proof of the h-microlocal, short-range weighted \(L^2\) bounds in Proposition 2.5.

Since P(h) has simple characteristics and is classically elliptic \( p^{-1}(0)\) is a compact, real-analytic hypersurface and by assumption, \(p^{-1}(0) \subset M_{\tau }^{{\mathbb C}}.\)

2.2 Long-Range Estimates

Let \(p(\alpha , h)\sim \sum _{j=0}^\infty p_j(\alpha )h^j\in S^{m}\) be the full symbol of P(h) and assume that it lies in \(S^{0,m}_{cla}(W)\) where W is a neighborhood of \((x_0,\xi _0)\). Here, \(\xi _0\) is allowed to be a point at infinity in which case a neighborhood means a conic neighborhood of \(\xi _0\) near infinity. We say p is elliptic at \((x_0,\xi _0)\) if \(|p_0|\ge c\left\langle \xi \right\rangle ^m>0\) in a neighborhood of \((x_0,\xi _0)\).

Proposition 2.3

Suppose that P(h) is a semiclassical pseudodifferential operator analytic in a neighborhood of \((x_0,\xi _0)\) and elliptic at \((x_0,\xi _0)\). Suppose that

$$\begin{aligned} P(h)\varphi _h=O{}(e^{-c/h}),\quad \Vert \varphi _h\Vert _{L^2}=1. \end{aligned}$$

Then, for any FBI transform T(h), there exists \(c>0\) and W a neighborhood of \((x_0,\xi _0)\) so that

$$\begin{aligned} \Vert T(h)\varphi _h\Vert _{L^2(W)}=O(e^{-c/h}). \end{aligned}$$

Proof

Let \(\chi _1\in C^{\infty }_0(T^*M)\) so that \(\chi _1\equiv 1\) near \((x_0,\xi _0)\) and p is elliptic and analytic on \({{\text {supp}}}\chi _1\). Let T(h) be an FBI transform with symbol \(r\in S_{cla}^{3n/4,n/4}\) and phase function \(\varphi \). An application of analytic stationary phase [18] gives

$$\begin{aligned} \chi _1(\alpha )T(h) P(h) \varphi _h = \chi _1(\alpha )T_{b}(h) \varphi _h + O(e^{-C \langle \alpha _{\xi } \rangle / h}), \end{aligned}$$
(2.14)

where

$$\begin{aligned} T_{b}(h) \varphi _h (\alpha ) = \int _M e^{i\varphi (\alpha ,y)/h} b(\alpha ,y,h) \chi (|\alpha _x-y|) \varphi _h(y) \mathrm{d}y, \end{aligned}$$

with

$$\begin{aligned} b(\alpha ,y,h)= & {} \sum _{j=0}^{{C}h^{-1}} \tilde{p}_{j}(y, {\alpha },-d_y \varphi (\alpha ,y) ) h^j, \tilde{p}_{j} \in S^{3{{n}}/4-j, {{n}}/4+{m}-j}_{cla}\nonumber \\ \tilde{p}_0{(y,\alpha ,\eta )}= & {} {r_0(\alpha ,y)p_0(y,\eta )}. \end{aligned}$$
(2.15)

Here, \(r_0\) is the principal symbol of r from (2.3). Since

$$\begin{aligned} - d_{y} \varphi (\alpha ,y) = \alpha _{\xi } \, ( 1 + O(|\alpha _x-y|) ) \end{aligned}$$

and \(|\alpha _x -y| < \delta \ll 1\) on supp \(\chi (\alpha _x-y)\) it follows that \(p_0(y,-d_y \varphi )\) is h-elliptic near \(\alpha =(x_0,\xi _0)\). In particular,

$$\begin{aligned} \frac{1}{C} \langle \alpha _{\xi } \rangle ^m \le | p_0(y,-d_y \varphi (\alpha ,y)) | \le C \langle \alpha _{\xi } \rangle ^m,\qquad \alpha \text { near }(x_0,\xi _0). \end{aligned}$$
(2.16)

Then, from (2.14) and the eigenfunction equation \(P(h) \varphi _h =O(e^{-c/h}),\) it follows that

$$\begin{aligned} \Vert \chi _1 T_{b}(h) \varphi _h (\alpha )\Vert _{L^2} = O(e^{-C \langle \alpha _{\xi } \rangle / h}). \end{aligned}$$
(2.17)

We claim that (2.17) is independent of FBI transform; in particular,

$$\begin{aligned} \Vert \chi _1 T_{geo}(h) \varphi _h \Vert _{L^2} = O(e^{-C/h}). \end{aligned}$$
(2.18)

Since \(\chi _1(\alpha )b(\alpha ,y,h)\) is h-elliptic near \((x_0,\xi _0,x_0)\), [18, Proposition 6.2] proves the estimate. We review the proof here for the reader’s convenience. The operator given by

$$\begin{aligned} Au(x)= & {} h^{-3n/2}\iint \exp \left[ \frac{i}{h}\left( \varphi (\alpha ,x)-\overline{\varphi ({\alpha },y)}\right) \right] b(\alpha ,y)\chi _1(\alpha )\\&\quad \chi (|\alpha _x-y|)\chi (|\alpha _x-x|)u(y)\mathrm{d}yd\alpha \end{aligned}$$

is an h-pseudodifferential operator with elliptic symbol near \((x_0,\xi _0)\). So, for any \(a(\alpha ,x)\), supported near \((x_0,\xi _0,x_0)\), we can find a classical analytic symbol, \(\tilde{b}\) defined near \((x_0,\xi _0,x_0)\) so that

$$\begin{aligned} A(\tilde{b}(\alpha ,\cdot )e^{i\varphi /h})=a(\alpha ,x)e^{i\varphi /h} \end{aligned}$$

modulo exponential errors. In particular, for W a small enough neighborhood of \((x_0,\xi _0)\),

$$\begin{aligned} a(\alpha ,x)e^{i\varphi /h}=\int _{\beta \in W} b(\beta , x)e^{i\varphi (\beta ,x)/h} K(\alpha ,\beta )d\beta , \end{aligned}$$

where

$$\begin{aligned} K(\alpha ,\beta )=h^{-3n/2}\iint \exp \left[ \frac{i}{h}\left( \varphi (\alpha ,y)-\overline{\varphi (\beta ,y)}\right) \right] \tilde{b}(\alpha ,y)\mathrm{d}y. \end{aligned}$$

By an application of analytic stationary phase,

$$\begin{aligned} K(\alpha ,\beta )= & {} e^{i \Phi (\alpha ,\beta )/h} c(\alpha ,\beta ;h) \chi (\alpha _x- \beta _x) \chi ( |\alpha _{\xi }|^{-1} |\alpha _{\xi } - \beta _\xi |) \\&\quad +\, O_N( e^{-C/h} ) \max \{ |\alpha _{\xi }|,|\beta _{\xi }| \}^{-N}, \end{aligned}$$

where \( {{\text {Im}}\,}\Phi (\alpha , \beta ) \ge |\alpha - \beta |^2,\) and \(c \in S^{{n},{\infty }}_{cla}.\)

In particular, we can write

$$\begin{aligned} \chi _2(\alpha )T_{geo}= K \chi _1(\alpha )T_b+R(h) \end{aligned}$$

where K is tempered in h,

$$\begin{aligned} |\partial _{x}^{\alpha } \partial _{y}^{\beta } R(x,y,h)| = O_{\alpha ,\beta }(e^{-C/h}), \,\, C>0, \end{aligned}$$

and \(\chi _1\equiv 1\) on \({{\text {supp}}}\chi _2\). Henceforward, we write \(\chi _2 \Subset \chi _1\) to denote this. Thus,

$$\begin{aligned} \chi _2T_{geo} \varphi _h = K\chi _1 T_b \varphi _h + O(e^{-C/h}) =O_{}(e^{-C/h}). \end{aligned}$$

\(\square \)

2.3 Short-Range Estimates

Let \(\chi _{in} \in C^{\infty }_{0}(M_{\tau }^{{\mathbb C}};[0,1])\) and \(\tilde{\chi }_{in} \in C^{\infty }_{0} (M_{\tau }^{{\mathbb C}}; [0,1])\) be a cutoff with \(\tilde{\chi }_{in} \Supset \chi _{in}\).

To deal with the short-range case, using analytic stationary phase one constructs an h-pseudodifferential intertwining operator \(Q(h) \in Op_h(S^{0,\infty }(T^*M))\) that is h-microlocally analytic on the Grauert tube \(M_{\tau }^{{\mathbb C}} \subset T^*M\) and satisfies

$$\begin{aligned} \chi _{in} T_{hol}(h) P(h) \varphi _h = \chi _{in} Q(h) T_{hol}(h) \varphi _h + O(h^{\infty }) \Vert \tilde{\chi }_{in} T_{hol}(h) \varphi _h \Vert _{L^2} + O(e^{-C/h}).\nonumber \\ \end{aligned}$$
(2.19)

To construct Q(h) in (2.19), using (2.14) we write

$$\begin{aligned} T_{hol}(h) P(h) \varphi _h = T_{b}(h) \varphi _h + O(e^{-C/h}), \end{aligned}$$

where \(b \in S^{3{{n}}/4,{{n}}/4+{m}}_{cla}(T^*M)\) is given by (2.15). Then, the symbol \(q(\alpha ,\alpha ^*;h) \sim _{h \rightarrow 0^+} \sum _{j=0}^{\infty } q_j(\alpha , \alpha ^*) h^j\) of Q(h) is determined by solving the equations

$$\begin{aligned} q_j(\alpha , d_{\alpha } \varphi (\alpha ,y))&= \tilde{p}_j (y,{\alpha }, -d_y \varphi (\alpha ,y)); \,\, j=0,1,2,3,... \nonumber \\ \tilde{p}_0(y,{\alpha },\eta )&= {r_0(y,\alpha )}p_0(y,\eta ), \end{aligned}$$
(2.20)

where \(\alpha ^*\) is the dual coordinate to \(\alpha \) and \(\tilde{p}_j\) are determined as in (2.15). To solve for the \(q_j\) in (2.20), we first consider the complexified equations

$$\begin{aligned} q_j^{{\mathbb C}}(\alpha , d_{\alpha } \varphi ^{{\mathbb C}}(\alpha ,y)) = \tilde{p}_j^{{\mathbb C}}(y, -d_y \varphi ^{{\mathbb C}}(\alpha ,y)); \,\, j=0,1,2,3,... \end{aligned}$$
(2.21)

Let \(\widetilde{M_{\tau }^{{\mathbb C}}}\) be a complex extension of \(M_{\tau }^{{\mathbb C}}\). Here, \(^{\mathbb C}\) denotes holomorphic continuation to \((\alpha ,y) \in \widetilde{ M_{\tau }^{{\mathbb C}} } \times M_{\tau }^{{\mathbb C}}\) with \(|\alpha _x - y| < \varepsilon _0\) and \(M_{\tau }^{{\mathbb C}} = \{ \alpha ; \sqrt{\rho }(\alpha ) \le \tau \}\) is identified with the complex thickening \(\tilde{M}.\) Since

$$\begin{aligned} \partial _{\alpha _x} \varphi ^{{\mathbb C}}&= - \partial _y \varphi ^{{\mathbb C}} + O(|\alpha _x -y|) = \alpha _{\xi } + O(|\alpha _x-y|), \nonumber \\ \partial _{\alpha _{\xi }} \varphi ^{{\mathbb C}}&= \alpha _{x}-y + O(|\alpha _x-y|^2), \end{aligned}$$
(2.22)

it follows that near \(\alpha _x=y\), \(\det \partial ^2_{y \, \alpha _\xi }\varphi \ne 0\) and so by the holomorphic implicit function theorem, \(d_{\alpha _{\xi }} \varphi (\alpha ,y)=w\) defines \(y=\beta _x^{{\mathbb C}}(\alpha ,w)\) with \(\beta _x^{{\mathbb C}}\) holomorphic in a neighborhood of \(\alpha _x=y.\) Hence, restricting to real points \((\alpha ,y) \in M_{\tau }^{{\mathbb C}} \times M\) with \(|\alpha _x -y| < \varepsilon _0,\) we can write

$$\begin{aligned} q_j(\alpha ,\alpha ^*)= & {} \tilde{p}_j(\beta _x(\alpha ,\alpha _{\xi }^*),-d_y\varphi (\alpha ,\beta _x(\alpha ,\alpha _\xi ^*)))\nonumber \\= & {} \tilde{p}_j(\beta _x(\alpha ,\alpha _\xi ^*),\beta _\xi (\alpha ,\alpha _{\xi }^*)); j=0,1,2,..., \end{aligned}$$
(2.23)

where \(\beta _x\) and \(\beta _{\xi }\) are locally \(C^{\omega }.\)

Since \(\tilde{p}_0 = p_0,\) for the principal symbols one gets

$$\begin{aligned} q_0(\alpha ,\alpha ^*) = p_0(\beta _x(\alpha ,\alpha _\xi ^*),\beta _\xi (\alpha ,\alpha _{\xi }^*)). \end{aligned}$$
(2.24)

It will be useful to introduce the principal symbol of conjugated operator \(e^{\psi /h} Q(h) e^{-\psi /h}\) given by

$$\begin{aligned} q_0^{\psi }(\alpha ,\alpha ^*) := q_0(\alpha , \alpha ^* + i d_{\alpha } \psi (\alpha )). \end{aligned}$$
(2.25)

Now, \(\partial ^2_{y, \alpha _\xi } \varphi |_{y=\alpha _x}=-{\text {Id}}\), and since \(\partial _y^2r(\alpha ,y)|_{y=\alpha _x}=2 {\text {Id}}\), \(\partial ^2_y\varphi |_{y=\alpha _x}=i{\text {Id}}\). Therefore, using also that

$$\begin{aligned} d_{\alpha _\xi }\varphi (\alpha ,y)=\alpha _x-y+O_{}(|\alpha _x-y|^2), \end{aligned}$$

It follows from (2.22) that,

$$\begin{aligned} \beta _{x}(\alpha ,\alpha _{\xi }^*) = \alpha _x-\alpha _\xi ^*+O(|\alpha _\xi ^*|^2), \quad \beta _{\xi }(\alpha , \alpha _{\xi }^*) = \alpha _{\xi }+i\alpha _{\xi }^*+O(|\alpha _\xi ^*|^2).\qquad \quad \end{aligned}$$
(2.26)

for \((\alpha ,y) \in M_{\tau }^{{\mathbb C}} \times M\) with \(r(\alpha _x,y) < \varepsilon _0.\) As an example, we note that in the \({\mathbb R}^n\) case with standard phase function \(\varphi (\alpha ,y) = (\alpha _x-y) \alpha _{\xi } + \frac{i}{2} |\alpha _x-y|^2,\) one has \(\beta _x = \alpha _x - \alpha _{\xi }^*\) and \(\beta _{\xi } = \alpha _{\xi }+i\alpha _{\xi }^*.\)

Given the intrinsic complex structure on the Grauert tube \(M_{\tau }^{{\mathbb C}},\) we denote the associated Cauchy–Riemann operators by \(\partial : C^{\infty }(M_{\tau }^{{\mathbb C}}) \rightarrow \Omega ^{1,0}(M_{\tau }^{{\mathbb C}})\) and \(\overline{\partial }: C^{\infty }(M_{\tau }^{{\mathbb C}}) \rightarrow \Omega ^{0,1}(M_{\tau }^{{\mathbb C}}).\) Moreover, given local coordinates \(\alpha _x\) in a chart \(U \subset M,\) the corresponding complex coordinates in \(U^{{\mathbb C}} \subset M_{\tau }^{{\mathbb C}}\) will be denoted by \(\alpha _{z}:= \exp _{\alpha _x}(-i \alpha _{\xi }), \,\, \bar{\alpha _z} = \exp _{\alpha _x}( i \alpha _{\xi }).\)

Given a smooth one-form \(\theta \in \Omega ^{1}(M_{\tau }^{{\mathbb C}})\) one can write it in local \((\alpha _x,\alpha _{\xi })\)-coordinates in the form

$$\begin{aligned} \theta = \alpha _x^* d\alpha _x + \alpha _{\xi }^* d \alpha _{\xi } \end{aligned}$$

and in terms of complex coordinates \((\alpha _z, \bar{\alpha _z})\) as

$$\begin{aligned} \theta = \alpha _z^* d\alpha _z + \bar{\alpha _z}^* d \bar{\alpha _z}. \end{aligned}$$

Consequently, in terms of the Cauchy–Riemann operators, \(\alpha _z^* = \sigma (\partial )(\alpha )\) and \( \bar{\alpha _{z}}^* = \sigma (\bar{\partial })(\alpha ).\) Here, \(\sigma \) denotes the principal symbol of a pseudodifferential operator.

Given a weight function \(\psi \in C^{\infty }(M_{\tau }^{{\mathbb C}})\) and the strictly plurisubharmonic weight \(\rho (\alpha _{z},\bar{\alpha _z}) =\frac{ |\alpha _{\xi }|_g^2}{2},\) we consider the associated submanifold \(\Lambda \subset \widetilde{M_{\tau }^{{\mathbb C}}} \,\) given by

$$\begin{aligned} \Lambda = \{ \,(\alpha , 2i \partial \psi (\alpha )+i(\bar{\partial }-\partial )\rho (\alpha ) ), \,\, \alpha \in M_{\tau }^{{\mathbb C}} \}. \end{aligned}$$
(2.27)

As we shall see below, the manifold \(\Lambda \) will play an important role in our main exponential weighted estimate in Proposition 2.5.

For future reference, we note that in terms of the local complex coordinates \((\alpha _z, \bar{\alpha }_z)\) in a geodesic normal coordinate chart U

$$\begin{aligned} \Lambda |_{U} := \{ ( \, \alpha _z, \bar{\alpha _z}; \alpha _z^*=(2i\partial _{\alpha _z}\psi - i \partial _{\alpha _z}\rho )(\alpha ), \, \bar{\alpha _z}^* = i \overline{\partial }_{\alpha _z}\rho (\alpha ) ), \,\, \alpha \in M_{\tau }^{{\mathbb C}} \}\nonumber \\ \end{aligned}$$
(2.28)

2.4 Complex Geometry of \(\Lambda \)

We first recall some basic complex symplectic geometry: Let X be a complex n-dimensional manifold with complex cotangent bundle \(T^*X.\) Viewing X as a real-analytic manifold, we let \(T^*X_{{{\mathbb R}}}\) denote the real 4n-dimensional cotangent bundle. There is a natural identification [9] of \( T^*X_{{\mathbb R}}\) with \(T^*X\) given as follows. Let \(v \in TX\) (a complex tangent vector) and \((z, \zeta ) \in T^*X\) (a complex covector). Then, the identification \(\iota : T^*X \rightarrow T^*X_{{\mathbb R}}\) is given by

$$\begin{aligned} \iota (z, \zeta (v)) = (z, \xi (v)); \quad \xi (v) = \mathrm Re\zeta (v). \end{aligned}$$

In terms of local coordinates \((\mathrm Rez, {{\text {Im}}\,}z): X \rightarrow {\mathbb R}^{2n}\) and the corresponding dual coordinates \((\xi ,\eta ) \in T_{(\mathrm Rez, {{\text {Im}}\,}z)}^*X_{{\mathbb R}},\)

$$\begin{aligned} \iota (\mathrm Rez, {{\text {Im}}\,}z;\xi ,\eta ) = (z, \zeta ), \quad \zeta = \xi + i \eta . \end{aligned}$$

Let \(\Gamma \subset T^*X\) be I-Lagrangian with respect to the complex symplectic form \(\Omega = dz \wedge d \zeta .\) Then, for any contractible coordinate chart U,  we recall that [9, Lemma 3.1] using the identification \(\iota ,\) one can locally characterize \(\Gamma \) as the graph of complex differential; that is,

$$\begin{aligned} \Gamma |_{U} = \{ (z, 2i \partial _{z} f), \,\, z \in U \}, \end{aligned}$$
(2.29)

with \(f \in C^{\infty }(U;{\mathbb R})\) and \( 2 \partial _z f = ( \partial _\mathrm{Rez} + i \partial _{{{\text {Im}}\,}z} ) f.\)

We claim that \(\Lambda \subset T^* ( M_{\tau }^{{\mathbb C}})\) in (2.27) can be naturally identified with an I-Lagrangian with respect to the canonical complex symplectic form. More precisely, consider

$$\begin{aligned} \tilde{\Lambda }: = \Lambda - i \, \text {graph}( d\rho ) \subset \widetilde{M_{\tau }^{{\mathbb C}}}. \end{aligned}$$

To see that \(\tilde{\Lambda }\) is indeed I-Lagrangian, we note that since \(d = \partial + \bar{\partial },\) one can write

$$\begin{aligned} 2 i \partial \psi + i(\bar{\partial }- \partial ) \rho = 2i \partial (\psi - \rho ) + i d \rho \end{aligned}$$

and so,

$$\begin{aligned} \tilde{\Lambda } = \{ (\alpha ,2i \partial (\psi - \rho )(\alpha )), \,\, \alpha \in M_{\tau }^{{\mathbb C}} \}. \end{aligned}$$

Consequently, in view of (2.29), \(\tilde{\Lambda }\) is indeed I-Lagrangian. Moreover, since \(\rho \) is strictly plurisubharmonic with \( \partial \bar{\partial }\rho >0\), it follows that with \(\Vert \psi \Vert _{C^2}\) sufficiently small, \(\tilde{\Lambda }\) is also \({\mathbb R}\)-symplectic.

We note that clearly one can write the Toeplitz multiplier \(q_0 |_{\Lambda }\) on the RHS of Proposition 2.5 as \(\tilde{q_0} |_{\ \tilde{\Lambda }}\) where \(\tilde{q_0}(\alpha ,\alpha ^*) = q_0(\alpha , \alpha ^* + i d_{\alpha } \rho (\alpha ))\) and \(\tilde{\Lambda }\) is the I-Lagrangian above. However, we find working with \(q_0\) (from (2.24)) and \(\Lambda \) (from (2.27)) computationally simpler and so we continue to work throughout with these instead.

Remark 2.4

Here we call \(q_0\) a Toeplitz multiplier in reference to the corresponding Toeplitz operator \(S_{hol}q_0T_{hol}\) (see e.g., [29, Chapter 13.4]).

The following h-microlocal manifold version of the microlocal Agmon estimates in \({\mathbb R}^n\) [13, 14] is central to the proofs of Theorems 1, 2.

Proposition 2.5

Let \(\Lambda \subset \widetilde{M_{\tau }^{{\mathbb C}}}\) be as in (2.27). Then, for any \(P(h) \in Op_h (S^{0,\infty }_{cla})\) there exists \(\delta >0\) so that for \(\psi \in S^0(1)\) with \(\Vert \psi \Vert _{C^1}{<\delta }\),

$$\begin{aligned}&\langle \chi _{in} e^{\psi /h} T_{hol}(h) P(h) \varphi _h, e^{\psi /h} T_{hol}(h) \varphi _h \rangle _{L^2} \\&\quad = \langle \chi _{in} e^{\psi /h} \, q_0 |_{\Lambda } \, T_{hol}(h) \varphi _h, e^{\psi /h} T_{hol}(h) \varphi _h \rangle _{L^2}\\&\qquad +\, O(h) \Vert \tilde{\chi }_{in} e^{\psi /h} T_{hol}(h) \varphi _h \Vert _{L^2}^2 + O(e^{-C/h}). \end{aligned}$$

Proof

The operator \(Q_{\psi }(h):= \chi _{in} e^{\psi /h} Q(h) e^{-\psi /h}\) has Schwartz kernel

$$\begin{aligned} Q_{\psi }(\alpha ,\beta ,h)&= (2\pi h)^{-2n} \lim _{\varepsilon \rightarrow 0^+} \int e^{i \langle \alpha - \beta , \alpha ^* \rangle /h} e^{- \varepsilon \langle \alpha ^* \rangle /h} e^{i [ \psi (\alpha ) - \psi (\beta )]/h} \chi _{in}(\alpha ) \,\nonumber \\&\quad q(\alpha ,\alpha ^*;h) \, d\alpha ^*. \end{aligned}$$
(2.30)

By Taylor expansion,

$$\begin{aligned} \psi (\alpha ) - \psi (\beta ) = \langle \Psi (\alpha ,\beta ), \alpha - \beta \rangle \end{aligned}$$

with \(| \Psi |< \Vert \psi \Vert _{C^1} < \delta .\) Since \(q(\alpha ,\alpha ^*,h)\) is analytic, for \(\delta >0\) small it follows by Stokes formula one can make the contour deformation

$$\begin{aligned} \alpha ^* \mapsto \alpha ^* + i \Psi (\alpha ,\beta ) \end{aligned}$$

in (2.30). Boundary terms as \(|\alpha ^*| \rightarrow \infty \) vanish and one gets that \(Q^{\psi }(h) \in Op_{h}(S^{0}(1))\) with symbol

$$\begin{aligned} q^{\psi }(\alpha ,\alpha ^*,h) \sim \sum _{j=0}^{\infty } \chi _{in}(\alpha ) q_j( \alpha , \alpha ^* + i d_{\alpha } \psi ) h^j. \end{aligned}$$

and principal symbol

$$\begin{aligned} q^{\psi }_0(\alpha ,\alpha ^*) = q_{0}(\alpha , \alpha ^* + i d_{\alpha } \psi (\alpha )), \end{aligned}$$

where \(q_0\) is defined in (2.24).

In view of (2.19) it follows that

$$\begin{aligned}&\langle \chi _{in} e^{\psi /h} T_{hol}(h) P(h) \varphi _h, e^{\psi /h} T_{hol}(h) \varphi _h \rangle _{L^2} \nonumber \\&\quad = \langle \chi _{in} Q^{\psi }(h) \, [e^{\psi /h} T_{hol}(h) \varphi _h ], \, e^{\psi /h} T_{hol}(h) \varphi _h \rangle _{L^2} \nonumber \\&\qquad + O(h^{\infty }) \Vert \tilde{\chi }_{in} e^{\psi /h} T_{hol}(h) \varphi _h \Vert _{L^2}^2 + O(e^{-C/h}). \end{aligned}$$
(2.31)

Next, in analogy with [13], we observe that with \(\bar{D}_{\alpha _z} = \frac{1}{i} \bar{\partial }_{\alpha _z},\)

$$\begin{aligned}&h \overline{D}_{\alpha _z} \Big ( e^{\psi /h} T_{hol}(h) \varphi _h \Big )\nonumber \\&\quad = \int _{M} h \overline{D}_{\alpha _z} \Big ( e^{ [ \, - r^{2}_{{\mathbb C}}(\alpha _z,y)/2 + \psi (\alpha _z,\bar{\alpha _z}) - \rho (\alpha ) \, ] /h } a(\alpha _z,y,h) \chi (\alpha _x,y) \Big ) \varphi _h(y) \mathrm{d}y \nonumber \\&\quad = \Big (-i\overline{\partial }_{\alpha _z} \psi (\alpha ) + i \overline{\partial }_{\alpha _z} \rho (\alpha ) \Big ) \, e^{\psi /h} T_{hol}(h) \varphi _h + O(e^{-C(\varepsilon _0)/h}) \nonumber \\&\quad = - i \overline{\partial }_{\alpha _z} \tilde{\psi }(\alpha ) \, e^{\psi /h} T_{hol}(h) \varphi _h + O(e^{-C(\varepsilon _0)/h}), \end{aligned}$$
(2.32)

where we have written

$$\begin{aligned} \tilde{\psi }=\psi -\rho . \end{aligned}$$

The last line in (2.32) follows since \(r^{2}_{{\mathbb C}}(\cdot ,y)\) and \(a(\cdot ,y,h)\) are holomorphic and the exponential error arises from differentiation of the cutoff \(\chi (\alpha _x,y).\) In particular, when \(r(\alpha _x,y)\ge \varepsilon _0>0\), there exists \(C(\varepsilon _0)>0\) so that

$$\begin{aligned} \mathrm{Re} [ \, - r^{2}_{{\mathbb C}}(\alpha _z,y)/2 + \psi (\alpha _z,\bar{\alpha _z}) - \rho (\alpha ) ] {\le -} C(\varepsilon _0){<}0. \end{aligned}$$

Similarly, taking complex conjugates, one gets that

$$\begin{aligned} hD_{\alpha _z} \Big ( \overline{ e^{\psi /h} T_{hol}(h) \varphi _h } \Big )&= i \partial _{\alpha _z} \tilde{\psi }(\alpha ) \, \overline{ e^{\psi /h} T_{hol}(h) \varphi _h} + O(e^{-C(\varepsilon _0)/h}). \end{aligned}$$
(2.33)

Taylor expansion of the principal symbol \(q^{\psi }_0\) of \(Q_{\psi }(h)\) around \(\bar{\alpha _z}^* = - i \overline{\partial }_{\alpha _z} \tilde{\psi }\) and \(\alpha _z^* = i \partial _{\alpha _z} \tilde{\psi }\) gives

$$\begin{aligned} q_0^{\psi }(\alpha ,\alpha ^*)&= q_0 (\alpha , \alpha ^* + i d_{\alpha } \psi (\alpha )) \nonumber \\&= q_0(\alpha _z,\bar{\alpha _z}; \alpha _z^* = i \partial _{\alpha _z} (\psi -\rho ) + i \partial _{\alpha _z} \psi , \, \bar{\alpha _z}^* = - i \overline{\partial }_{\alpha _z}(\psi -\rho ) + i \overline{\partial }_{\alpha _z}\psi ) \nonumber \\&\qquad + r_1(\alpha ,\alpha ^*) ( \bar{\alpha }_{z}^* + i \overline{\partial }_{\alpha _z} \tilde{\psi } ) + r_2(\alpha ,\alpha ^*) ( \alpha _{z}^* - i \partial _{\alpha _z} \tilde{\psi } ) \nonumber \\&= q_0 (\alpha _z, \bar{\alpha _z}; \alpha _z^* = 2 i \partial _{\alpha _z} \psi - i \partial _{\alpha _z} \rho , \, \bar{\alpha _z}^* = i \overline{\partial }_{\alpha _z}\rho ) \nonumber \\&\qquad + r_1(\alpha ,\alpha ^*) ( \bar{\alpha }_{z}^* + i \overline{\partial }_{\alpha _z} \tilde{\psi } ) + r_2(\alpha ,\alpha ^*) ( \alpha _{z}^* - i \partial _{\alpha _z} \tilde{\psi } ) \nonumber \\&= q_{0}(\alpha ) |_{\alpha \in \Lambda } + r_1(\alpha ,\alpha ^*) ( \bar{\alpha }_{z}^* + i \overline{\partial }_{\alpha _z} \tilde{\psi } ) + r_2(\alpha ,\alpha ^*) ( \alpha _{z}^* - i \partial _{\alpha _z} \tilde{\psi } )\nonumber \\ \end{aligned}$$
(2.34)

Since \(r_1,r_2 \in S^{0}(1),\) it then follows from (2.31)–(2.34), \(L^2\)-boundedness of pseudodifferential operators, and that \([Op_h(S^{m_1}),Op_h(S^{m_2})]\in hOp_h(S^{m_1+m_2})\) (see for example [29, Chapters 4,9]) that

$$\begin{aligned}&\langle \chi _{in} Q_{\psi }(h) \, [e^{\psi /h} T_{hol}(h) \varphi _h ] \,, e^{\psi /h} T_{hol}(h) \varphi _h \rangle _{L^2}\nonumber \\&\quad = \langle \chi _{in} e^{\psi /h} q_{0}(\alpha ) |_{\alpha \in \Lambda } T_{hol}(h) \varphi _h, \, e^{\psi /h} T_{hol}(h) \varphi _h \rangle _{L^2}\nonumber \\&\qquad + \langle \chi _{in} r_1(\alpha ,hD_\alpha ) [ h\overline{D}_{\alpha _z} + i \overline{\partial }_{\alpha _z} \tilde{\psi } ] e^{\psi /h} T_{hol}(h) \varphi _h, \, e^{\psi /h} T_{hol}(h) \varphi _h \rangle _{L^2}\nonumber \\&\qquad + \langle \chi _{in} e^{\psi /h} T_{hol}(h) \varphi _h, \, r_2( \alpha ,h D_{\alpha }) [ h\overline{D}_{\alpha _z} + i \overline{\partial }_{\alpha _z} \tilde{\psi } ] e^{\psi /h} T_{hol}(h) \varphi _h \rangle _{L^2}\nonumber \\&\qquad + O(h) \Vert \tilde{\chi }_{in} e^{\psi /h} T_{hol}(h) \varphi _h \Vert _{L^2}^2. \nonumber \\&\quad =\langle \chi _{in} e^{\psi /h} q_{0}(\alpha ) |_{\alpha \in \Lambda } T_{hol}(h) \varphi _h, \, e^{\psi /h} T_{hol}(h) \varphi _h \rangle _{L^2} \nonumber \\&\qquad + O(h) \Vert \tilde{\chi }_{in} e^{\psi /h} T_{hol}(h) \varphi _h \Vert _{L^2}^2.&\end{aligned}$$
(2.35)

This finishes the proof of the Proposition. \(\square \)

2.5 Microlocal Eigenfunction Decay Estimates

2.5.1 Estimation of the Multiplier \(q_0 |_{\Lambda }\)

In order to prove Theorem 2 we give an invariant characterization of the Toeplitz multiplier \(q_0 |_{\Lambda }.\)

In terms of the local coordinates \((\alpha _z, \bar{\alpha }_z)\) the Toeplitz multiplier in Proposition 2.5 is

$$\begin{aligned} q(\alpha _z,\bar{\alpha _z},i\partial _{\alpha _z}(2\psi -\rho ),i\partial _{\bar{\alpha }_z}\rho ). \end{aligned}$$

To give this in invariant meaning, we note that function \(q_{0} \in C^{\omega }(T^*M_{\tau }^{{\mathbb C}})\) and

$$\begin{aligned} (\alpha _z,\bar{\alpha _z},i\partial _{\alpha _z}(2\psi -\rho ),i\partial _{\bar{\alpha }_z}\rho ) \end{aligned}$$

are local coordinates for the point

$$\begin{aligned} \zeta&= i\partial _{\alpha _z}(2\psi -\rho )(\alpha ) \, d\alpha _z + i\partial _{\bar{\alpha }_z}\rho (\alpha ) \, d\bar{\alpha }_z \in T^*M_{\tau }^{{\mathbb C}}.\\&=2i\partial \psi (\alpha ) +i(\bar{\partial }-\partial )\rho (\alpha ) \in T^*M_{\tau }^{{\mathbb C}}. \end{aligned}$$

Consequently, the Toeplitz multiplier equals

$$\begin{aligned}&q_0^{\psi }( \alpha , i \partial \tilde{\psi }(\alpha ) - i \bar{\partial }\tilde{\psi }(\alpha )) = q_0 (\alpha , i (\partial - \bar{\partial }) \tilde{\psi }(\alpha ) + i d \psi (\alpha ) ) \nonumber \\&= q_0 (\alpha , 2 i \partial \psi (\alpha ) + i(\bar{\partial }- \partial ) \rho (\alpha ) ) = q_0 |_{\Lambda }(\alpha ), \quad \alpha \in M_{\tau }^{{\mathbb C}}, \end{aligned}$$
(2.36)

in view of the definition of \(\Lambda \) in (2.27), where

$$\begin{aligned} q_0 |_{\Lambda }(\alpha ) = q_0( \alpha ,2i\partial \psi (\alpha ) - i (\partial -\bar{\partial })\rho (\alpha ) ). \end{aligned}$$
(2.37)

Here, \(\partial , \bar{\partial }: C^{\infty }(M_{\tau }^{{\mathbb C}}) \rightarrow \Omega ^{1}(M_{\tau }^{{\mathbb C}})\) are the intrinsic Cauchy–Riemann operators.

The \(\rho \) portion, \(i\bar{\partial }\rho -i\partial \rho \), of the argument on the RHS of (2.37) can be readily computed. Given the strictly plurisubharmonic weight function \(\rho (\alpha ) = \frac{1}{2} |\alpha _{\xi }|^2_{\alpha _x}\), we recall that [5, p. 568],

$$\begin{aligned} i (\bar{\partial } - \partial ) \rho = \omega , \end{aligned}$$
(2.38)

where \(\omega = \sum _j \alpha _{\xi _j} d\alpha _{x_j}\) is the canonical one-form. Since both sides are invariant, one can easily verify the identity (2.38) by computing in geodesic normal coordinates at the center of the coordinate chart.

It follows from (2.38) that

$$\begin{aligned} \begin{aligned} \alpha _x^*(2i\partial \psi -i(\partial -\bar{\partial })\rho )&=\alpha _x^*(2i\partial \psi )+\alpha _\xi \\ \alpha _\xi ^*(2i\partial \psi -i(\partial -\bar{\partial })\rho )&=\alpha _\xi ^*(2i\partial \psi ). \end{aligned} \end{aligned}$$
(2.39)

From (2.39) one gets

$$\begin{aligned} q |_{\Lambda }(\alpha )&= p_0 \Big ( \, \beta _x( \alpha , \, \alpha _{\xi }^* ( 2i\partial \psi ) ) \, , \beta _{\xi }( \alpha ,\alpha _{\xi }^* ( 2i\partial \psi ) \, \Big ) \nonumber \\&= p_0 ( \, \alpha _x + O(|\partial \psi |), \, \alpha _{\xi } + O(|\partial \psi |) \, ), \end{aligned}$$
(2.40)

where in the last equality we have used (2.26).

Remark 2.6

We recall that \(\rho \) is a Kahler potential for the intrinsic Kahler form \(\Omega _g\) on the Grauert tube corresponding to the complex structure \(J_g\) induced by complexified Riemannian exponential map of the metric g; that is,

$$\begin{aligned} \partial \bar{\partial } \rho = \Omega _g. \end{aligned}$$

The corresponding exhaustion function \(\sqrt{2\rho }(\alpha ) = |\alpha _{\xi }|_g\) satisfies homogeneous Monge–Ampere,

$$\begin{aligned} \det \Big ( \frac{\partial ^2 \sqrt{\rho } }{\partial \alpha _{z_i} \partial \bar{\alpha _{z_j}} } \Big ) = 0. \end{aligned}$$

We refer to [5, 11] for further details.

From Proposition 2.5 and (2.40) we get the following useful estimate.

Proposition 2.7

Under the same assumptions as in Proposition 2.5,

$$\begin{aligned}&\langle \chi _{in} e^{\psi /h} T_{hol}(h) P(h) \varphi _h, e^{\psi /h} T_{hol}(h) \varphi _h \rangle _{L^2} \\&\qquad =\langle \chi _{in} e^{\psi /h} \, p(\alpha + O(|\partial \psi |) ) \, T_{hol}(h) \varphi _h, e^{\psi /h} T_{hol}(h) \varphi _h \rangle _{L^2} \\&\qquad \quad + O(h) \Vert \tilde{\chi }_{in} e^{\psi /h} T_{hol}(h) \varphi _h \Vert _{L^2}^2 + O(e^{-C/h}). \end{aligned}$$

2.6 Microlocal Concentration of the Eigenfunctions: Proof of Theorem 2

In this section, we prove the global weighted decay estimate in Theorem 2.

Proof

We first prove the weighted estimate in (1.8) h-microlocally on support of \(\chi _{in} \in C^{\infty }_{0}(M_{\tau }^{{\mathbb C}})\) and for the h-microlocal FBI transform \(T_{hol}(h).\)

$$\begin{aligned} \Vert \chi _{in} e^{\psi /h} T_{hol}(h) \varphi _h \Vert _{L^2(T^*M)} = O(1).\ \end{aligned}$$
(2.41)

The transform \(T_{hol}(h)\) is only defined on the Grauert tube \(M_{\tau }^{{\mathbb C}},\) which is generally a proper, bounded subset of \(T^*M.\) We then need to prove that the weighted bound in (2.41) still holds for \(T(h) = T_{geo}(h)\), after possibly shrinking \(\delta >0\) somewhat (independently of h).

In the following, we let \(P(h) \in Op_h(S^{0,k}_{cla})\) be as in the statement of Theorem 2 and \(\varphi _h\) be an exponential quasimode with

$$\begin{aligned} P(h) \varphi _h = O(e^{-c/h}). \end{aligned}$$

2.6.1 h-Microlocal Bounds for \(T_{hol}(h) \varphi _h\)

Recall \(\psi =\frac{\delta p^2}{2\langle \xi \rangle ^{2k}}.\) For \(\delta >0\) small,

$$\begin{aligned} \frac{1}{C(\delta )} p^2(\alpha _x,\alpha _{\xi }) \le \mathrm Rep^2(\alpha + O(|\partial _{\alpha } \psi |) ) \le C(\delta ) p^2(\alpha _x,\alpha _{\xi }). \end{aligned}$$
(2.42)

To see that (2.42) holds, we split into two cases:

  1. Case (i)

    \(\{ \alpha ; |p(\alpha )| \ll 1\}\) (near the characteristic variety). Here, we make a Taylor expansion to get

    $$\begin{aligned} p^2(\alpha + O(|\partial _{\alpha }\psi |) = p^2(\alpha ) + O(\delta ) |2p\partial _{\alpha }p|^2 = p^2(\alpha ) + O(\delta \, p^2(\alpha )).\nonumber \\ \end{aligned}$$
    (2.43)

    Consequently, near \(p=0\) it follows from (2.43) that for \(\delta >0\) small,

    $$\begin{aligned} \mathrm Rep^2(\alpha + O(|\partial _{\alpha }\psi |) ) \ge c p^2(\alpha ). \end{aligned}$$
  2. Case (ii)

    \(\{ \alpha ; |p(\alpha )| \gtrapprox 1 \}\) (far field). Here we use the fact that \(\psi \in S^{0}(1)\) and \( |\partial p^2| \lessapprox \langle \alpha _{\xi } \rangle ^{2k-1}\) and just make the first-order Taylor expansion

    $$\begin{aligned} p^2(\alpha + O(|\partial _{\alpha }\psi |) ) = p^2(\alpha ) + O( \delta \langle \alpha _{\xi } \rangle ^{2k-1} ). \end{aligned}$$

    Since \( p^2(\alpha ) \gtrapprox \langle \alpha _{\xi } \rangle ^{2k}\) in this range, (2.42) is also satisfied in this case, provided one chooses \(\delta >0\) small.

Since \(P(h) \in Op_h(S^{0,\infty }_{cla})\) implies that also \(P^2(h) \in Op_h(S^{0,\infty }_{cla}),\) and so Proposition 2.7 applies just as well with the latter. We note that by (2.42), there is a constant \(C>0\) such that

$$\begin{aligned} \frac{1}{C} p^2(\alpha ) \le \mathrm Rep^2|_{\Lambda }(\alpha ) \le C p^2(\alpha ), \quad \alpha \in \text {supp}\, \chi _{in} \end{aligned}$$

and so, by an application of Proposition 2.7 with the globally defined weight function \(\psi \) in the statement of Theorem 2, it follows that

$$\begin{aligned}&\langle \chi _{in} e^{\psi /h} T_{hol}(h) P^2(h) \varphi _h, e^{\psi /h} T_{hol}(h) \varphi _h \rangle _{L^2} \nonumber \\&\quad = \langle \chi _{in} e^{\psi /h} p^2|_{\Lambda } T_{hol}(h) \varphi _h, e^{\psi /h} T_{hol}(h) \varphi _h \rangle _{L^2} \nonumber \\&\qquad + O(h) \Vert \tilde{\chi }_{in} e^{\psi /h} T_{hol}(h) \varphi _h \Vert _{L^2}^2 + O(e^{-C/h}) \nonumber \\&\quad \ge \langle \chi _{in} e^{\psi /h} p^2 T_{hol}(h) \varphi _h, e^{\psi /h} T_{hol}(h) \varphi _h \rangle _{L^2} \nonumber \\&\qquad + O(h) \Vert \chi _{in} e^{\psi /h} T_{hol}(h) \varphi _h \Vert _{L^2}^2 + O(e^{-C/h}). \end{aligned}$$
(2.44)

In the last line of (2.44) we have used the long-range estimate (2.3) yet again to write

$$\begin{aligned} \Vert \tilde{\chi }_{in} e^{\psi /h} T_{hol}(h) \varphi _h \Vert _{L^2}^2 = \Vert \chi _{in} e^{\psi /h} T_{hol}(h) \varphi _h \Vert _{L^2}^2 + O(e^{-C/h}). \end{aligned}$$

To bound the LHS in (2.44) note that \(A_k(h):=S_{hol}(h) \left\langle \alpha _\xi \right\rangle ^{k}T_{hol}(h)\in Op_h(S^{0,k}_{cla})\). Hence, \(A_{-k}(h)P(h)\in {\text {Op}}_h(S^{0,0}_{cla})\) and since P(h) is classically elliptic with \(\Vert P(h) {\varphi _h}\Vert _{L^2}+\Vert {\varphi _h}\Vert _{L^2}=O(1)\), \(A_{k}(h){\varphi _h}\in L^2\). In particular,

$$\begin{aligned}&\left\langle \chi _{in} e^{\psi /h}T_{hol}(h)P^2(h) {\varphi _h}, e^{\psi /h}T_{hol}(h){\varphi _h}\right\rangle \\&=\left\langle \chi _{in} e^{\psi /h}\left\langle \alpha _\xi \right\rangle ^{-k}T_{hol}(h)P^2(h) {\varphi _h}, e^{\psi /h}\left\langle {\alpha _\xi }\right\rangle ^kT_{hol}(h){\varphi _h}\right\rangle \\&=\left\langle \chi _{in} e^{\psi /h}T_{hol}(h)A_{-k}(h)P^2(h) {\varphi _h},e^{\psi /h}T_{hol}(h)A_k(h) {\varphi _h}\right\rangle +O(e^{-C/h})\\&=O(e^{-C/h}), \end{aligned}$$

since by assumption \(P(h) \varphi _h = O_{L^2}(e^{-C/h}).\)

Thus, from (2.44) one gets that for \(\delta \) small enough and \(C=C(\delta )>0\),

$$\begin{aligned}&\langle \chi _{in} e^{\psi /h} p^2 T_{hol}(h) \varphi _h, e^{\psi /h} T_{hol}(h) \varphi _h \rangle _{L^2} \nonumber \\&\quad = O_{\delta }(h) \Vert \chi _{in} e^{\psi /h} T_{hol}(h) \varphi _h \Vert _{L^2}^2+O(e^{-C/h}). \end{aligned}$$
(2.45)

We note that for any \({{N}}>0,\)

$$\begin{aligned} \sup _{ p^2(\alpha ) \le {{N}} h } \psi (\alpha ) = O(h), \end{aligned}$$

so that \(e^{\psi (\alpha )/h} = O(1)\) when \(p^2(\alpha ) \le {{N}} h.\) It then follows from (2.45) that

$$\begin{aligned}&\langle \chi _{in} e^{\psi /h} [ p^2 + O(h)] T_{hol}(h) \varphi _h, e^{\psi /h} T_{hol}(h) \varphi _h \rangle _{L^2 (\{ p^2 \ge {{N}}h\}) } \nonumber \\&\quad = O_{\delta }(h) \Vert \chi _{in} T_{hol}(h)\varphi _h \Vert _{L^2 (\{ p^2 \le {{N}} h \}) }^2 + O(e^{-C/h}) = O(h). \end{aligned}$$
(2.46)

Choosing \({{N}={N(\delta )}}>0\) large enough to absorb the O(h) term on the LHS of (2.46) (which is independent of \({{N}}>0\)), it follows that

$$\begin{aligned} \Vert \chi _{in} e^{\psi /h} T_{hol}(h) \varphi _h \Vert _{L^2 (\{ p^2 \ge {{N}}h\}) } = O(1). \end{aligned}$$
(2.47)

Clearly, since \(\psi \approx p^2\) near \(p=0\) it also follows that

$$\begin{aligned} \Vert \chi _{in} e^{\psi /h} T_{hol}(h) \varphi _h \Vert _{L^2 (\{ p^2 \le {{N}}h\})} = O(1) \end{aligned}$$

which finishes the proof of (2.41).

2.6.2 Weighted Bounds in Terms of \(T_{geo}(h)\)

Since \(T_{hol}(h)\) is only h-microlocally defined, we need to show that essentially the same weighted bound holds for the globally defined FBI transform \(T_{geo}(h).\) More precisely, in this section we show that with \(\varepsilon _0 <1\) sufficiently small (but independent of h) the analogue of (2.41) holds for the globally defined FBI transform. That is

$$\begin{aligned} \Vert \chi _{in} e^{ \varepsilon _0 \psi / h} T_{geo}(h) \varphi _h \Vert _{L^2} = O(1). \end{aligned}$$
(2.48)

Given (2.48), the global result in Theorem 2 follows (after possibly shrinking \(\delta >0\) further) since we have already established the long range bound in (2.18).

To prove (2.48), we write

$$\begin{aligned}&\chi _{in} e^{ \varepsilon _0 \psi /h} T_{geo}(h) \varphi _h \nonumber \\&\quad = {\chi _{in}e^{\varepsilon _0\psi /h}T_{geo}(h)S_{hol}(h)T_{hol}(h) \varphi _h+O(e^{-C/h}) }\nonumber \\&\quad =\chi _{in} e^{\varepsilon _0 \psi /h} T_{geo}(h) S_{hol}(h) \tilde{\chi }_{in} T_{hol}(h) \varphi _h + O(e^{-C/h}) \nonumber \\&\quad =\chi _{in} e^{\varepsilon _0 \psi /h} T_{geo}(h) S_{hol}(h) e^{-\psi /h} \tilde{\chi }_{in} e^{\psi /h} T_{hol}(h) \varphi _h + O(e^{-C/h}). \end{aligned}$$
(2.49)

Note in the second line we us that \(T_{geo}S_{hol}\) is pseudolocal modulo exponential errors. Then, by an application of analytic stationary phase, after possibly shrinking \(\delta >0\), the Schwartz kernel of the operator \(\chi _{in} e^{\varepsilon _0 \psi /h} T_{geo}(h) S_{hol}(h) e^{-\psi /h}\) can be written in the form

$$\begin{aligned} e^{i \Phi (\alpha ,\beta )/h} c(\alpha ,\beta ;h) \chi (\alpha _x- \beta _x) \chi ( |\alpha _{\xi }|^{-1} |\alpha _{\xi } - \beta _\xi |) + O_N( e^{-C/h} ) \max \{ |\alpha _{\xi }|,|\beta _{\xi }| \}^{-N}, \end{aligned}$$

where,

$$\begin{aligned} {{\text {Im}}\,}\Phi (\alpha , \beta ) \ge - \varepsilon _0 p^2(\alpha ) + p^2(\beta ) + |\alpha - \beta |^2, \end{aligned}$$

\(c \in S^{{{n}},0}_{cla}\). To estimate \( \Phi (\alpha , \beta )\) when \(\alpha \) and \(\beta \) are near \(p^{-1}(0),\) we note that by Taylor expansion of \(p^2(\alpha )\) around \(\alpha = \beta \) one gets that

$$\begin{aligned} {{\text {Im}}\,}\Phi (\alpha , \beta ) \ge - \varepsilon _0 p^2(\alpha ) + p^2(\beta ) + C_0 ( p(\alpha ) - p(\beta ) )^2, \end{aligned}$$

with some \(C_0 >0.\) Writing \(x = p(\alpha )\) and \(y = p(\beta )\) it therefore suffices to consider the function

$$\begin{aligned} f(x,y)= - \varepsilon _0 x^2 + y^2 + C_0 (x-y)^2; \,\,\,\, (x,y) \in {\mathbb R}^2. \end{aligned}$$

An application of max/min shows that \(f(x,y) \ge 0\) provided \(\varepsilon _0 (C_0^2)>0\) is chosen sufficiently small and consequently, it follows that for small \(\varepsilon _0>0,\)

$$\begin{aligned} {{\text {Im}}\,}\Phi (\alpha ,\beta ) \ge 0. \end{aligned}$$

Then, from (2.49) and an application of the Schur‘s lemma, it follows that for \(\varepsilon _0>0\) sufficiently small,

$$\begin{aligned} \Vert e^{\varepsilon _0 \psi /h} T_{geo}(h) S_{hol}(h) e^{-\psi /h} \Vert _{L^2 \rightarrow L^2} = O(1) \end{aligned}$$

and so,

$$\begin{aligned} \Vert \chi _{in} e^{ \varepsilon _0 \psi /h} T_{geo}(h) \varphi _h \Vert _{L^2} = O(1) \Vert \tilde{\chi }_{in} e^{ \psi / h} T_{hol}(h) \varphi _h \Vert _{L^2} + O(e^{-C/h}) = O(1). \end{aligned}$$

Consequently, the h-microlocal bound (2.48) follows.

After possibly shrinking \(\varepsilon _0>0\) further, we know that by the long range bound in (2.18),

$$\begin{aligned} \Vert e^{\varepsilon _0 \psi /h} (1-\chi _{in}) T_{geo}(h) \varphi _h \Vert _{L^2 \rightarrow L^2} = O(1), \end{aligned}$$

and so, Theorem 2 follows. \(\square \)

2.7 Refinement of the Weight Function: Proof of Theorem 3

Proof

We first turn to the proof of (1.13). Since the results of Theorem 3 are h-microlocal and, moreover, away from a neighborhood of \(\{p=0\}\) we know that any FBI transform applied to \(\varphi _h\) is exponentially small we work in a small neighborhood of \(\{p=0\}\). In particular, this implies that for \(\chi \in C_c^\infty (\mathbb {R})\) with \(\chi \equiv 1\) near 0, and for fixed arbitrarily small \(\varepsilon _0>0\), there exists \(c = c(\varepsilon _0)>0\) so that

$$\begin{aligned} S_{hol}\chi (\varepsilon _0^{-1}p^2/2)T_{hol}(h)\varphi _h=\varphi _h +O_{L^2}(e^{-c/h}). \end{aligned}$$

Thus, we work exclusively with \(T_{hol}(h)\) here and start by reexamining the proof of (2.47). The key estimate is (2.44) where we use that \((P(h))^2\varphi _h=O_{H_h^{-k}}(e^{-c/h})\) (here \(H_h^{-k}\) is the semiclassical Sobolev space or order \(-k\); see for example [29, Chapter 14]). Notice that in order to conclude that \(e^{\psi /h}T_{hol}(h)\varphi _h\) is well controlled, we must have good control of the O(h) error term appearing in the right hand side of (2.44). In particular, we must have control of \(e^{\psi /h}T_{hol}(h)\varphi _h\) strictly away from \(\{p=0\}\). The long range estimates tell us that we have some exponential decay, however, we do not have any useful control over the constant. Therefore, in order to complete the arguments leading to (2.47), we must choose \(\psi \) so that

$$\begin{aligned} \psi = 0\text { on }p^2> 2\varepsilon ,\qquad p^2(\beta (\alpha ,\alpha _\xi ^*(2i\partial \psi )))\ge cp^2 \text { with } c>0 \text { when } p^2 < \varepsilon .\nonumber \\ \end{aligned}$$
(2.50)

\(\square \)

Lemma 2.8

For all \(\delta _0>0\), \(0<\gamma <1/2\), there exists \(\psi _0\in C^\infty (M_\tau )\) so that \(\Vert \psi _0\Vert _{C^1}\le \delta _0\), \({{\text {supp}}}\psi _0\subset \{|p|\le \delta _0\}\),

$$\begin{aligned} \mathrm Rep^2(\beta (\alpha ,\alpha _\xi ^*(2i\partial \psi _0)))\ge \frac{1-2\gamma }{2}p^2, \end{aligned}$$

and in a neighborhood of \(p=0\),

$$\begin{aligned} psi_0=\frac{\gamma }{2|dp|_{g_s}^2}, \end{aligned}$$

where \(g_s\) is the Sasaki metric on TM (see e.g., [2, Chapter 9]).

Proof

Motivated by the construction of the weight in Theorem 2, we let \(\chi \in C_c^\infty (\mathbb {R})\) with \(\chi \equiv 1\) on \([-1,1]\), \({{\text {supp}}}\chi \subset [-2,2]\), \(0\le \chi \le 1\), \(\chi '(x)\le 0\) on \(x\ge 0\), and make the ansatz

$$\begin{aligned} \psi _0=\chi (\delta ^{-1}p^2)a p^2. \end{aligned}$$
(2.51)

Note that throughout this proof, all \(O(\cdot )\) statements are uniform in \(\delta \).

By (2.26),

$$\begin{aligned} \beta _x(\alpha ,\alpha _\xi ^*)=\alpha _x-\alpha _\xi ^*+O_{}(|\alpha _\xi ^*|^2),\quad \beta _\xi (\alpha ,\alpha _\xi ^*)=\alpha _\xi +i\alpha _\xi ^*+O_{}(|\alpha _\xi ^*|^2). \end{aligned}$$

We work in geodesic normal coordinates centered at \(\alpha _x\) (i.e., \(\alpha _z=\exp _{\alpha _x}(-i\alpha _\xi )\)) so that \(\partial \psi _0=\frac{1}{2}(\partial _{\alpha _x}\psi _0+i\partial _{\alpha _\xi }\psi _0)(d\alpha _x-id\alpha _\xi )\). Therefore,

$$\begin{aligned} \alpha _\xi ^*(2i\partial \psi _0)=\partial _{\alpha _x}\psi _0+i\partial _{\alpha _\xi }\psi _0. \end{aligned}$$

Now,

$$\begin{aligned} \partial \psi _0 =2 p \, \partial p \, a(\chi (\delta ^{-1}p^2)+\delta ^{-1}p^2 \chi '(\delta ^{-1}p^2)) + p^2 \chi (\delta ^{-1}p^2) \, \partial a. \end{aligned}$$

In particular, then

$$\begin{aligned} \begin{aligned} p^2(\beta (\alpha ,\alpha _\xi ^*(2i\partial \psi _0)))&=p^2 - 2p(\partial _{\alpha _x} p-i\partial _{\alpha _\xi } p)(\partial _{\alpha _x}\psi _0 +i\partial _{\alpha _\xi } \psi _0)+O(|\partial \psi _0|^2)\\&=p^2 \cdot \Big ( 1-4|dp|_{g_s}^2a(\chi (\delta ^{-1}p^2)+\delta ^{-1}p^2\chi '(\delta ^{-1}p^2))\\&\quad +O(p^{-1}|\partial \psi _0|^2) \Big ) \end{aligned} \end{aligned}$$

Therefore, choosing

$$\begin{aligned} a=\frac{\gamma }{2|dp|_{g_s}^2}, \end{aligned}$$

proves the lemma since \(\chi '(x)\le 0\) on \(x\ge 0\) implies

$$\begin{aligned} 1-4|dp|_{g_s}^2a(\chi (\delta ^{-1}p^2)+\delta ^{-1}p^2\chi '(\delta ^{-1}p^2))\ge & {} 1-2\gamma (\chi (\delta ^{-1}p^2)\\&\quad +\,\delta ^{-1}p^2\chi ' (\delta ^{-1}p^2))\ge 1-2\gamma \end{aligned}$$

and

$$\begin{aligned} p^{-1}|\partial \psi _0|^2=p^{-1}(O(\delta ^{-2}p^6+p^2))=O(p). \end{aligned}$$

\(\square \)

Using \(\psi _0\) from Lemma 2.8 in the analysis leading to (2.47) proves (1.13).

Next, we prove the bound in Theorem 3 (1.14). The fact that \((P(h))^2\varphi _h=O_{H_h^{-k}}(e^{-c/h})\) is a weaker condition than \(P(h)\varphi _h=O_{L^2}(e^{-c/h})\) and indeed, the example of \(e^{ix/h}\) on \(S^1\) shows that the weight \(\psi _0\) is not quite optimal. To remedy this, and obtain (2.41) we need to work directly with \(P(h)\varphi _h=O_{L^2}(e^{-c/h})\). Unlike \(p^2\), p does not have a fixed sign, so we need to work separately on \(p>0\) and \(p<0\). Therefore, we construct slightly different weights on \(p>0\) and \(p<0\). Since the weighted estimate naturally localizes to each region, we consider first the case \(p>0\) and then easily adapt the argument to the case \(p<0.\)

Lemma 2.9

For all \(\delta _0>0\), there exists \(\psi _+\in C^\infty (M_\tau )\) with \(\Vert \psi _+\Vert _{C^1}<\delta _0\), \({{\text {supp}}}\psi _+\subset \{|p|\le \delta _0\}\) so that

$$\begin{aligned} \mathrm Rep(\beta (\alpha ,\alpha _{\xi }^*(2i\partial \psi _+)))\ge p^2\quad \text { on }p\ge 0 \end{aligned}$$

and in a neighborhood of \(p=0\),

$$\begin{aligned} \psi _+=\frac{p^2}{2|dp|_{g_s}^2}+O(p^3), \end{aligned}$$

where \(g_s\) is the Sasaki metric on TM.

Proof

First, let \(\chi \in C_c^\infty (\mathbb {R})\) with \(\chi \equiv 1\) on \([-1,1]\), \({{\text {supp}}}\chi \subset [-2,2]\), and \(\chi '(x)\le 0\) on \(x\ge 0\). Then fix \(\delta >0\) to be chosen small enough later and let

$$\begin{aligned} \psi _+=\chi (\delta ^{-1}p)(a_0p^2+a_1p^3),\quad a_i\in C^\infty (M_\tau ). \end{aligned}$$

Note that throughout this proof, all \(O(\cdot )\) statements are uniform in \(\delta \).

Computing as in the proof of Lemma 2.8,

$$\begin{aligned} \beta _x(\alpha ,\alpha _\xi ^*(2i\partial \psi ))=\alpha _x-(\partial _{\alpha _x}\psi _++i\partial _{\alpha _\xi }\psi _0)+O(|d\psi |^2),\\ \beta _\xi (\alpha ,\alpha _\xi ^*(2i\partial \psi ))=\alpha _\xi +i(\partial _{\alpha _x}\psi _++i\partial _{\alpha _\xi }\psi _0)+O(|d\psi |^2). \end{aligned}$$

For convenience, let

$$\begin{aligned} \partial _{\bar{\alpha }_z}=\partial _{\alpha _x}+i\partial _{\alpha _\xi },\quad \partial _{\alpha _z}=\partial _{\alpha _x}-i\partial _{\alpha _\xi }. \end{aligned}$$

Then,

$$\begin{aligned} \partial _{\bar{\alpha }_z}\psi _+= & {} \chi (\delta ^{-1}p)(2a_0p\partial _{\bar{\alpha }_z}p+a_13p^2\partial _{\bar{\alpha }_z}p+p^2\partial _{\bar{\alpha }_z}a_0\\&\quad +\, O(p^3))+\delta ^{-1}\chi '(\delta ^{-1}p)(\partial _{\bar{\alpha }_z}p)(a_0p^2+O(p^3)) \end{aligned}$$

Then, Taylor expansion of \(p\circ \beta (\alpha ,\alpha _\xi ^*(2i\partial \psi _+))\) around \(\alpha \) gives

$$\begin{aligned}&p(\beta (\alpha ,\alpha _\xi ^*(2i\partial \psi _+)))\\&\quad =p\left[ \begin{aligned}&1-2|dp|_g^2\big (\delta ^{-1}\chi '(\delta ^{-1}p)a_0p+\chi (\delta ^{-1}p)(a_0+3pa_1)\big )\\&\qquad \qquad \qquad -\left\langle \partial _{\alpha _z}p,\chi (\delta ^{-1}p)p\partial _{\bar{\alpha }_z}a_0\right\rangle +\tilde{E}\end{aligned}\right] , \end{aligned}$$

where

$$\begin{aligned} |\tilde{E}|\le & {} \delta ^{-2}p^3|\chi '(\delta ^{-1}p)|^2+|p|\chi ^2(\delta ^{-1}p)\\&+\,p^2|\chi (\delta ^{-1}p)|+\delta ^{-1}p^2|\chi '(\delta ^{-1}p)| =O(p).\ \end{aligned}$$

Grouping terms by homogeneity in p, we have then

$$\begin{aligned}&p(\beta (\alpha ,\alpha _\xi ^*(2i\partial \psi _+)))\\&\quad =p\left[ \begin{aligned}&1-2|dp|_{g_s}^2\chi (\delta ^{-1}p)a_0 \\&\qquad \qquad \qquad -p\big (2|dp|_{g_s}^2[\chi (\delta ^{-1}p)a_1+\delta ^{-1}\chi '(\delta ^{-1}p)a_0]-\left\langle \partial _{\alpha _z}p,\chi (\delta ^{-1}p)\partial _{\bar{\alpha }_z}a_0\right\rangle \big )+\tilde{E}\end{aligned}\right] \end{aligned}$$

Now, let \({{N}}>0\) to be chosen large enough independently of \(\delta \) later

$$\begin{aligned} a_0:=\frac{1}{2|dp|_{g_s}^2},\qquad a_1=-{{N}}. \end{aligned}$$

Then, notice that on \(p\ge 0\), \(\chi '(\delta ^{-1}p)\le 0\), so, since \(\chi \ge 0\),

$$\begin{aligned}&\mathrm Rep(\beta (\alpha ,\alpha _\xi ^*(2i\partial \psi _+)))\\&\quad \ge p\left[ 1-\chi (\delta ^{-1}p)+p\left( 2|dp|_g^2\chi (\delta ^{-1}p){N}-\left\langle \partial _{\alpha _z}p,\chi (\delta ^{-1}p)\partial _{\bar{\alpha }_z}a_0\right\rangle \right) +O(p)\right] \end{aligned}$$

Therefore, on \(\{\chi (\delta ^{-1}p)\ge \frac{1}{2}\}\cap \{p\ge 0\}\),

$$\begin{aligned} \mathrm Rep(\beta (\alpha ,\alpha _\xi ^*(2i\partial \psi _+)))&\ge p\left[ p\big (|dp|_g^2{N}\right. \\&\quad \left. -\left\langle \partial _{\alpha _z}p,\chi (\delta ^{-1}p)\partial _{\bar{\alpha }_z}a_0\right\rangle \big )+O(p)\right] \ge \frac{{{N}}}{2}p^2 \end{aligned}$$

for N large enough. On the other hand, on \({{\text {supp}}}\chi (\delta ^{-1}p)\cap \{\chi (\delta ^{-1}p)\le \frac{1}{2}\}\cap \{p\ge 0\}\),

$$\begin{aligned} \mathrm Rep(\beta (\alpha ,\alpha _\xi ^*(2i\partial \psi _+))) \ge p\left[ \frac{1}{2}+O(p)\right] \ge \frac{1}{4}p \end{aligned}$$

if \(\delta >0\) is chosen small enough.

This completes the proof of the lemma since

$$\begin{aligned} |\psi _+|\le C\delta ^2,\qquad |\partial \psi _+|\le C\delta . \end{aligned}$$

\(\square \)

Remark 2.10

It is possible to construct \(\psi _+\) satisfying the following: For every \({{N}}>0\), \({{M_0}}>0\), and \(\delta _0>0\), there exists \(\psi _+\in C^\infty (M_\tau )\) with \(\Vert \psi _+\Vert _{C^1}\le \delta _0\), and \({{\text {supp}}}\psi _+\subset \{|p|\le \delta _0\}\) so that

$$\begin{aligned} \mathrm Rep(\beta (\alpha ,\alpha _\xi ^*(2i\partial \psi _+)))\ge \frac{1}{2}p^{{N}} +h{{M_0}},\quad \text { on }p\ge 0, \end{aligned}$$

and in a neighborhood of \(p=0\),

$$\begin{aligned} \psi _+=\sum _{j=0}^{{{N}}-1}a_jp^2-Ch{{M_0}}p,\quad a_0=\frac{1}{2|dp|_{g_s}^2}. \end{aligned}$$

Moreover, for any \(j<{{N}}-1\) and \(\delta >0\), there exist \(b\in C^\infty (M_\tau )\) with \(\Vert b\Vert _{C^1}<\delta \) so that if \(a_j\) is replaced by \(a_j+b,\) then

$$\begin{aligned} \inf _{0\le p<\delta _0}\mathrm Rep(\beta (\alpha ,\alpha _\xi ^*(2i\partial \psi _+)))<0. \end{aligned}$$

This allows one to improve the higher order terms in \(\psi _{hol}\) in Theorem 3 so that they are sharp for \(e^{ix/h}\) modulo \(p^{{N}}\) for any N.

Since we want to localize to \(p>0\), we insert a smooth cutoff \(\chi _+\) that approximates the indicator function \(\mathbf{1}_{[0, \varepsilon _0]}(p)\). However, in order to estimate error terms in the weighted \(L^2\) bounds corresponding to Proposition 2.5, we must ensure that \(\psi _+ = O(h)\) on supp \(\partial \chi _+\) so that, in particular,

$$\begin{aligned} \Vert e^{\psi _+/h} T_{hol}(h) \varphi _h \Vert _{L^2( \text {supp } \partial \chi _+ )} = O(1), \end{aligned}$$

where \({{\text {supp}}}\partial \chi _{+} \subset \{0 \le p \le h^{1/2} \}\cup \{p \ge \varepsilon _0 \}.\)

To construct \(\chi _+\), we let \(\chi _1\in C^\infty (\mathbb {R})\) with \({{\text {supp}}}\chi _1\subset (1,\infty )\) with \( \chi _1\equiv 1\) on \((2,\infty )\) and let

$$\begin{aligned} \chi _{+}:=\chi (\varepsilon _0^{-1}p)\chi _1( h^{-1/2}p), \end{aligned}$$

and \(\tilde{\chi }_{+}=\tilde{\chi }(\varepsilon _0^{-1}p^2/2)\tilde{\chi }_1(h^{-1/2}p)\) where \(\tilde{\chi }\in C_c^\infty (\mathbb {R})\) has \(\tilde{\chi }\equiv 1\) on \({{\text {supp}}}\chi \) and \(\tilde{\chi }_1\in C^\infty (\mathbb {R})\) with \(\tilde{\chi }_1\equiv 1\) on \({{\text {supp}}}\chi _1\) and \({{\text {supp}}}\tilde{\chi }_1\subset (1,\infty )\). Then \({{\text {supp}}}\chi _+\subset \{0\le p\le 2\varepsilon _0\}\) so that \(\psi _+=O(h)\) on \({{\text {supp}}}\partial \chi _+\).

We will need the following

Lemma 2.11

Under the same assumptions as in Proposition 2.5,

$$\begin{aligned}&\langle \chi _{+} e^{\psi _{+}/h} T_{hol}(h) P(h) \varphi _h, e^{\psi _{\pm }/h} T_{hol}(h) \varphi _h \rangle _{L^2} \nonumber \\&\qquad = \langle \chi _{+}e^{\psi _{\pm }/h} (p\circ \beta )|_{\Lambda } T_{hol}(h) \varphi _h, e^{\psi _{+}/h} T_{hol}(h) \varphi _h \rangle _{L^2} \nonumber \\&\quad \qquad +\, O(h) \Vert \tilde{\chi }_{+} e^{\psi _{+}/h} T_{hol}(h) \varphi _h \Vert _{L^2}^2\nonumber \\&\quad \qquad +\,O(h^{1/2})\Vert \tilde{\chi }_{+} e^{\psi _{+}/h} T_{hol}(h) \varphi _h \Vert _{L^2(\{|p|\le Ch^{1/2}\})}^2 + O(e^{-C/h}) \end{aligned}$$
(2.52)

Proof

We follow very closely the argument in the proof of Proposition 2.5. The only difference here occurs in the integration by parts arguments with respect to \(\partial _{\alpha _z}\) and \(\bar{\partial }_{\alpha _z}\) in (2.32) and (2.33), precisely when the cutoff \(\chi _+\) is differentiated in the amplitude since it is a singular semiclassical symbol with \(\chi _+ \in S^{0}_{1/2}(1)\) [29, Chapter 4] i.e.,

$$\begin{aligned} |\partial _{\alpha _z}^\alpha \chi _+|\le C_{\alpha \beta }h^{-|\alpha |/2}. \end{aligned}$$

Specifically, one needs to estimate a term of the form

$$\begin{aligned}&\left\langle h \partial _{\alpha _z} ( \chi _{\pm } ) \, r(\alpha ,hD_{\alpha } ) [ e^{\psi /h}T_{hol}(h)\varphi _h], e^{\psi /h}T_{hol}(h) \varphi \right\rangle \nonumber \\&\quad = \left\langle h \chi \partial _{\alpha _z} ( \chi _{1} ) \, r(\alpha ,hD_{\alpha } ) [ e^{\psi /h}T_{hol}(h)\varphi _h], e^{\psi /h}T_{hol}(h) \varphi \right\rangle \nonumber \\&\qquad + \, \left\langle h \chi _1 \partial _{\alpha _z} ( \chi ) \, r(\alpha ,hD_{\alpha } ) [ e^{\psi /h}T_{hol}(h)\varphi _h], e^{\psi /h}T_{hol}(h) \varphi \right\rangle \quad r \in S^{0}(1).\; \end{aligned}$$
(2.53)

Since \(\chi _+ \in S^{0}_{1/2}(1),\) it is in a singular symbol class. However, because it is a multiplier depending only the spatial coordinates \(\alpha \in M_{\tau }^{{\mathbb C}}\), it can be effectively composed with the standard h-pseudodifferential operators \(r(\alpha , hD_{\alpha }) \in Op_h (S^{0}(1)).\) In particular, symbols compose with \(h^{-j/2}\)-loss in the \(j^{\text {th}}\) term of the asymptotic expansion, \(L^2\)-boundedness and sharp Gårding still hold, as does h-pseudolocality. More precisely, for any \(R(h) \in Op_h (S^{0}(1))\) and spatial cutoff \(\chi _+ = \chi _+(\alpha ,h) \in S^{0}_{1/2}(1),\)

$$\begin{aligned} \chi _+ R(h) = \chi _+ R(h) \tilde{\chi }_+ + O(h^{\infty })_{L^2 \rightarrow L^2}. \end{aligned}$$

Since the cutoff \( \chi _1 \partial _{\alpha _z} \chi \in S^{0}_{1/2}(1),\) and depends only on the spatial \(\alpha \)-variables, it follows that for second term on the RHS of (2.53),

$$\begin{aligned}&\left\langle h \chi _1 \partial _{\alpha _z} ( \chi ) \, r(\alpha ,hD_{\alpha } ) [ e^{\psi /h}T_{hol}(h)\varphi _h], e^{\psi /h}T_{hol}(h) \varphi \right\rangle \\&\quad = O(h) \Vert \tilde{\chi }_{+} e^{\psi _{+}/h} T_{hol}(h) \varphi _h \Vert _{L^2}. \end{aligned}$$

As for the first term on the RHS of (2.53), \(h \chi \partial _{\alpha _z} (\chi _1) \in h^{-\frac{1}{2}} S_{1/2}^{0}(1)\) since there is a loss of \(h^{1/2}\) coming from differentiation of the \(\chi _1\)-term. Thus,

$$\begin{aligned} \left\langle h \chi \partial _{\alpha _z} ( \chi _{1} ) \, r(\alpha ,hD_{\alpha } ) [ e^{\psi /h}T_{hol}(h)\varphi _h], e^{\psi /h}T_{hol}(h) \varphi \right\rangle \\ =O(h^{1/2})\Vert \tilde{\chi }_{+} e^{\psi _{+}/h} T_{hol}(h) \varphi _h \Vert _{L^2(\{|p|\le Ch^{1/2}\})}^2. \end{aligned}$$

\(\square \)

Since \( P(h) \varphi _h = O_{L^2}(e^{-c/h}),\) it follows from Lemma 2.3 that for \(\varepsilon \) small enough and \({{N}}>0,\)

$$\begin{aligned}&\langle \chi _{+} e^{\psi _{+}/h} (p\circ \beta )|_{\Lambda } T_{hol}(h) \varphi _h, e^{\psi _{+}/h} T_{hol}(h) \varphi _h \rangle _{L^2} \nonumber \\&\quad = O(h) \Vert \tilde{\chi }_{+} e^{\psi _{+}/h} T_{hol}(h) \varphi _h \Vert _{L^2}^2\nonumber \\&\qquad +\,O(h^{1/2})\Vert \tilde{\chi }_{+} e^{\psi _{+}/h} T_{hol}(h) \varphi _h \Vert _{L^2(0 \le p \le {{N}}h^{1/2})}^2+O(e^{-c/h}). \end{aligned}$$
(2.54)

In analogy with the arguments used to prove (2.47), we substitute the lower bound \((p\circ \beta )|_{\Lambda } \ge p^2\) on \(p \ge 0\) from Lemma 2.9 in (2.54). Then, since \(\psi _+ = O(h)\) when \(p^2 = O(h)\) (so that \(e^{\psi _+/h} = O(1)\) on the latter set), it follows that

$$\begin{aligned}&\langle \chi _{+} e^{\psi _{+}/h} (c{{N}}h+O_{}(h))T_{hol}(h) \varphi _h, e^{\psi _{+}/h} T_{hol}(h) \varphi _h \rangle _{L^2(p \ge {{N}}h^{1/2})} \nonumber \\&\quad =O(h^{1/2}) \Vert \tilde{\chi }_{+} T_{hol}(h)\varphi _h \Vert _{L^2 (0 \le p\le {{N}}h^{1/2})}^2 \nonumber \\&\qquad +\, O(h) \Vert (\tilde{\chi }_{+} - \chi _+) T_{hol}(h)\varphi _h \Vert _{L^2 (p \ge {{N}}h^{1/2})}^2 \nonumber \\&\qquad +\, O(e^{-c/h}) = O(h^{1/2}). \end{aligned}$$
(2.55)

In (2.55) we have also used that \(e^{\psi _+(\alpha )/h} = O(1)\) for all \(\alpha \in \text {supp} (\tilde{\chi }_+ - \chi _{+}) \cap \{ p > C h^{1/2} \}\) since by construction \(\psi _+ = O(h)\) on the latter set (see also Figure 1).

Fig. 1
figure 1

We show the various cutoffs used in the proof of Theorem 3

Choosing \({{N}}>0\) large enough to absorb the O(h) term on the LHS of (2.55) it follows that

$$\begin{aligned} \Vert \chi _{+} e^{\psi _{+}/h} T_{hol}(h) \varphi _h \Vert _{L^2 } = O(h^{-1/4}). \end{aligned}$$
(2.56)

The analysis on the set \(p<0\) follows in the same way as above, except one uses the reflected cutoff function \(\chi _{-}(p) = \chi _+(-p), \) and the ansatz for the corresponding weight function is

$$\begin{aligned} \psi _{-}= \chi (-\delta ^{-1}p) p^2(a_0+a_1p), \quad p < 0, \, a_j \in C^{\infty }(M_\tau ); j=0,1. \end{aligned}$$
(2.57)

so that

$$\begin{aligned} -\mathrm Rep(\beta (\alpha ,\alpha _{\xi }^*(2i\partial \psi _-)))\ge p^2\qquad p\le 0. \end{aligned}$$

That finishes the proof of Theorem 3 (1.14). \(\square \)

3 Exponential Decay of the Harmonic Extensions of Boundary Eigenfunctions: Proof of Theorem 1

Proof

Recall that \(M:=\partial \Omega \) with \(\Omega \) the domain of the Steklov problem. Given the FBI transform \(T(h): C^{\infty }(M) \rightarrow C^{\infty }(T^*M)\) we construct a left-parametrix \(S(h): C^{\infty }(T^*M) \rightarrow C^{\infty }(M)\) and then given \(q\in S^{m}(1)\) we define the anti-Wick h-pseudodifferential quantization by

$$\begin{aligned} q^{aw}_{h}:= S(h) q T(h). \end{aligned}$$

Let \({\mathcal {P}}: C^{\infty }(M) \rightarrow C^{\infty }(\Omega )\) be the Poisson operator and \(\chi \in C^{\infty }_{0}(T^*M; [0,1])\) be a cutoff supported near the zero section, with \(\chi ({x',\xi '}) =1\) for \(\{ 0 \le |{\xi '}| \le \varepsilon \}\) and \(\chi ({x',\xi '}) =0\) for \(|{\xi '}| > 2 \varepsilon .\) Here, \(\varepsilon >0\) is some arbitrarily small but fixed constant. Then, one can write

$$\begin{aligned} u_h(x)&= {\mathcal {P}} \varphi _h (x_{{n+1}},{x'}) = {\mathcal {P}} (1 - \chi _h^{aw}) \varphi _h (x) + {\mathcal {P}} \chi _{h}^{aw} \varphi _h(x). \end{aligned}$$
(3.1)

Here, \( x=(x_{{n+1}},x')\) denote Fermi coordinates in a neighborhood of the boundary with \(\Omega = \{ x_{{n+1}} \ge 0 \}\) and \(M = \{ x_{{n+1}} = 0 \}.\) We show in Sect. 4.2 that for \(\chi _{h}^{aw} \varphi _h\) (the piece supported near the zero section), one has the apriori bound

$$\begin{aligned} \Vert \chi _{h}^{aw} \varphi _h \Vert _{L^2} = O(e^{-C/h}) \end{aligned}$$
(3.2)

for some \( C>0.\)

It follows from [19] that \({\mathcal {P}} (1 - \chi _h^{aw}) \varphi _h (x)\) (the piece h-microlocally supported away from the zero section) has Schwartz kernel of the form

$$\begin{aligned} K(x,y',h)= & {} (2\pi h)^{-{{n}}} \int _{{\mathbb R}^{{{n}}}} e^{ i({\Psi }(x_{{n+1}},x',{\xi '})- \langle {y', \xi '}\rangle ) /h} \, c(x,{y'},{\xi '},h) \, \chi (x'-y') \, d{\xi '}\\&\quad +\, O(e^{-C/h}), \end{aligned}$$

with \(c \in S^{0}(1)\) supported away from \(|{\xi '}|=0\) and in \(S^{0,0}_{cla}(|{\xi '}|>2\varepsilon )\), \({\Psi }\) solving

$$\begin{aligned} (\partial _{x_{{n+1}}}{\Psi })^2+r(x,\partial _{x'}{\Psi })=0,\quad {\Psi }(0,x',\xi ')=\left\langle x',\xi '\right\rangle ,\quad {{\text {Im}}\,}{\Psi }\ge 0 \end{aligned}$$
(3.3)

with \(r(a,x',\xi ')\) the symbol of the Laplacian induced on \(\{x_{{n+1}}=a\}.\) In particular,

$$\begin{aligned} r(x_{{n+1}},x',\xi ')=|\xi '|_{x'}^2+2x_{{n+1}}Q(x',\xi ')+O(x_{{n+1}}^2), \end{aligned}$$

where \(Q(x',\xi ')\) is the symbol of the second fundamental form. It follows by Taylor expansion in \(x_{{n+1}}\) that

$$\begin{aligned} {\Psi }=i{\Psi }_1(x_{{n+1}},x',\xi ')+\left\langle x',\xi '\right\rangle , \end{aligned}$$

where

$$\begin{aligned} {\Psi }_1(x_{{n+1}},x',\xi ')= & {} x_{{n+1}}|\xi '|_{x'}+\frac{x_{{n+1}}^2(Q(x',\xi ')+i\left\langle \partial _{x'}|\xi '|_{x'},\xi '\right\rangle _{x'})}{2|\xi '|_{x'}}\\&\quad +\, O(x_{{n+1}}^3|\xi '|_{x'}) \end{aligned}$$

is homogeneous of degree 1 in \(\xi '\). The near-diagonal cutoff \(\chi (x'-y')\) appears above since \(K(x,y',h)\) is the part of the Poisson operator arising h-microlocally from the complement of the zero section.

Consequently, from (3.1) and (3.2) we have

$$\begin{aligned} u_{h}(x)&= (2\pi h)^{-{{n}}} \int _{{\mathbb R}^{{n}}} \int _{M} e^{ i({\Psi }(x_{{n+1}},x',{\xi '})- \langle y', {\xi '}\rangle ) /h} \, c(x,y',{\xi '},h) \, \nonumber \\&\quad \chi (x'-y') \, \varphi _h(y') \, \mathrm{d}y' d{\xi '} +\, O(e^{-C/h}). \end{aligned}$$
(3.4)

Next we make an analytic resolution of the identity and write

$$\begin{aligned} Id = S(h) T(h) + O(e^{-C/h}), \end{aligned}$$

where \(T(h): C^{\infty }(M) \rightarrow C^{\infty }(T^*M)\) is an FBI transform with phase function \(\varphi \) (not to be confused with the Steklov eigenfunction \(\varphi _h\)) and \(S(h): C^{\infty }(T^*M \rightarrow C^{\infty }(M)\) is a left-parametrix. Notice that by Theorem 2 or Corollary 1.3 together with the analysis in Sect. 4.2 for \(\chi _1\in C_c(\mathbb {R})\) with \(\chi _1 \equiv 1\) on \([-1,1]\), and \({{\text {supp}}}\chi _1\subset [-2,2]\), for any \(\varepsilon >0\),

$$\begin{aligned} (1-\chi _1(\varepsilon ^{-1}(|\alpha _{\xi '}|_{\alpha _{x'}}-1))T(h)=O_{L^2}(e^{-c/h}). \end{aligned}$$

Let \(\chi _{1,\varepsilon }(\alpha )=\chi _1(\varepsilon ^{-1}(|\alpha _{\xi '}|_{\alpha _{x'}}-1)).\) Then,

$$\begin{aligned} \varphi _h=S(h)\chi _{1,\varepsilon }(\alpha )T(h)\varphi _h+O_{L^2}(e^{-c/h}) \end{aligned}$$

and substitution in the integral formula for \(u_h\) gives

$$\begin{aligned} u_h(x) = (2\pi h)^{-{{n}}} \int _{M} K_{h,x}(y') \, S(h)\chi _{1,\varepsilon }(\alpha ) T(h) \varphi _h(y') \mathrm{d}y' + O(e^{-C/h}) \end{aligned}$$
(3.5)

with

$$\begin{aligned} K_{h,x}(y') = \int _{{\mathbb R}^{{{n}}}} e^{ i({\Psi }(x_{{n+1}},x',{\xi '})- \langle y', {\xi '}\rangle ) /h} c(x',y',{\xi '},h) \chi (x'-y') \, d{\xi '}. \end{aligned}$$

One can then rewrite the formula (3.5) in the form

$$\begin{aligned} u_h(x)&= (2\pi h)^{-{{n}}} \langle \chi _{1,\varepsilon }(\alpha )S(h)^t K_{h,x}, \overline{T(h) \varphi _h} \rangle _{L^2(T^*M)} + O(e^{-C/h}). \end{aligned}$$
(3.6)

Thus, \(\chi _{1,\varepsilon }(\alpha )S(h)^t K_{h,x}(\alpha )\) equals

$$\begin{aligned}&(2\pi h)^{-{{n}}} \int _{M} \int _{{\mathbb R}^{{{n}}}} e^{i [ \, -\overline{\varphi (\alpha ,y')}+ {\Psi }(x_{{n+1}},x',{\xi '})- \langle y', {\xi '} \rangle \, ] /h} \\&\quad c(x',y',{\xi '},h) \chi (\alpha _x -y') a(y',\alpha _{x'},h) \chi (x'-y') \chi _{1,\varepsilon }(\alpha )\, \mathrm{d}y' d{\xi '} \end{aligned}$$

with \(\alpha = (\alpha _{x'}, \alpha _{\xi '}) \in T^*M\) and \(c \in S^{0,0}_{cla}(|{\xi '}|>2\varepsilon ), \, a \in S^{3{{n}}/4, {{n}}/4}_{cla}.\) Next, we apply analytic stationary phase in the \((y',{\xi '})\) variables. We can do for \(x_{{n+1}}\) small since \(\psi (x_{{n+1}},x',\eta ')=\left\langle x',\eta '\right\rangle +O_{C^\infty }(x_{{n+1}})\). Writing

$$\begin{aligned} \Phi (x,{\xi '};\alpha ,y'):= - \overline{\varphi (\alpha ,y')} + {\Psi }(x,{\xi '})- \langle y', {\xi '}\rangle \end{aligned}$$

for the total phase and computing in geodesic normal coordinates centered at \(\alpha _{x'},\) the critical point equations are

$$\begin{aligned} \partial _{{\xi '}} \Phi= & {} x'-y' +i\partial _{{\xi '}}{\Psi }_1(x',{\xi '})=0,\\ \partial _{y'}\Phi= & {} -\partial _{y'} \overline{\varphi (\alpha ,y')} - {\xi '} = \alpha _{\xi '} - {\xi '} + O(|\alpha _{x'}-y'|) = 0. \end{aligned}$$

It follows that, denoting the complex critical points by \(y'_c\), \(\eta '_c\), we have, using the fact that \(\psi _1\) is homogeneous of degree 1

$$\begin{aligned} \left\langle x'-y_c',{\xi }_c'\right\rangle =-i\left\langle \partial _{{\xi }'} {\Psi }_1 (x',{\xi }'_c),{\xi }'_c\right\rangle =-i{\Psi }_1(x,{\xi }'_c). \end{aligned}$$

Therefore,

$$\begin{aligned} \Phi (x,{\xi }'_c,\alpha ,y'_c)= & {} -\overline{\varphi (\alpha ,y'_c)},\\ y'_{c}= & {} x'+i\partial _{{\xi }'}\psi _1|_{{\xi }'={\xi }'_c},\quad {\xi }_c'=\alpha _{\xi '}+O(|\alpha _{x'}-y_c'|). \end{aligned}$$

Here, we implicitly analytically continue \(\overline{\varphi (\alpha ,y')}\). In particular, notice that since on the support of the integrand, \(\left| |\alpha _{\xi }|_g-1\right| \ll 1\), and \({\xi }_c'=\alpha _{\xi '}+O(|\alpha _{x'}-y_c'|)\), we have that \(||{\xi }_c'|_g-1|\ll 1\) and hence the fact that the amplitude is not analytic in \({\xi }'\) near \({\xi }'=0\), does not cause issues in the analytic stationary phase argument.

Consequently, the \((2 \pi h)^{-{{n}}}\) factor in front of (3.5) gets killed upon application of analytic stationary phase in \((y',\eta ')\) and we get the bound

$$\begin{aligned} |\chi _{1,\varepsilon }(\alpha )S(h)^t K_{x,h}(\alpha )|&\lessapprox h^{-3{{n}}/4}e^{{{\text {Im}}\,}\overline{\varphi (\alpha ,y_c'(x,\alpha ))}/h} \chi _{1,\varepsilon }(\alpha )\chi (\alpha _{x'}-y_c') . \end{aligned}$$
(3.7)

Substitution of (3.7) together with the weighted \(L^2\) estimate \(\Vert e^{\psi /h} T(h) \varphi _h \Vert _{L^2} = O(1)\) with \(\psi = \gamma p^2, \, \gamma < \frac{1}{2}\) (see Theorem 3 (1.13) ) in (3.6) gives, by Cauchy–Schwarz,

$$\begin{aligned} | u_h(x) |&\le | \, \langle e^{-\psi /h} \chi _{1,\varepsilon }(\alpha )S(h)^t K_{h,x}, \overline{e^{\psi /h} T(h) \varphi _h} \rangle _{L^2(T^*M)} \, | + O(e^{-C/h}) \nonumber \\&\le h^{-3{{n}}/4} \Vert e^{-\psi (\alpha )/h} e^{{{\text {Im}}\,}\overline{\varphi (\alpha ,y_c'(x,\alpha ))}/h} \Vert _{L^2(T^*M,d\alpha )} + O(e^{-C/h}). \end{aligned}$$
(3.8)

Now, recall that we may work with \(T_{hol}\) and \(S_{hol}\) since for some \(\varepsilon >0\),

$$\begin{aligned} S_{hol}\chi (\varepsilon ^{-1}p^2)T_{hol}\varphi _h=\varphi _h+O_{L^2}(e^{-c/h}). \end{aligned}$$

In this case the analytic continuation, of \(-\overline{\varphi }\) is given by

$$\begin{aligned} i(\rho (\bar{\alpha })+r^2_{\mathbb {C}}(\bar{\alpha },y))/2. \end{aligned}$$

Computing in normal geodesic coordinates centered at \(\alpha _x\), observe that

$$\begin{aligned} r^2(\alpha _x,y)=\left\langle \alpha _x-y,\alpha _x-y\right\rangle \end{aligned}$$

Therefore,

$$\begin{aligned} 2i\overline{\varphi (\alpha ,y_c'(x,\alpha ))}&=\left\langle \alpha _{x'}+i\alpha _{\xi '}-y_c',\alpha _{x'}+i\alpha _{\xi '}-y_c'\right\rangle +|\alpha _{\xi '}|^2\\&=| \alpha _{x'}-\mathrm Rey_c'|^2-|{{\text {Im}}\,}y_c'|^2+2i\left\langle \alpha _{x'}-\mathrm Rey_c',{{\text {Im}}\,}y_c'\right\rangle \\&\quad +2i\left\langle \alpha _{x'}-y_c',\alpha _{\xi '}\right\rangle \end{aligned}$$

Now, by Taylor expansion in \({\xi '}\),

$$\begin{aligned}\left\langle \partial _{{\xi '}}{\Psi }_1|_{{\xi }'={\xi }_c'},\alpha _{\xi '}\right\rangle&={\Psi }_1(x,\alpha _{\xi '})+x_{{n+1}}O(|{\xi }_c'-\alpha _{\xi '}|^2)\\&={\Psi }_1(x,\alpha _{\xi '})+x_{{n+1}}O(|\alpha _{x'}-y_c'|^2)\\&={\Psi }_1(x,\alpha _{\xi '})+x_{{n+1}}O((|\alpha _{x'}-\mathrm Rey_c'|^2+|{{\text {Im}}\,}y_c'|^2)) \end{aligned}$$

so,

$$\begin{aligned} {{\text {Im}}\,}\overline{\varphi (\alpha ,y_c'(x,\alpha ))}&=-\frac{1}{2}(| \alpha _{x'}-\mathrm Rey_c'|^2(1+O_{C^\infty }(x_{{n+1}}))\\&\quad -|{{\text {Im}}\,}y_c'|^2(1+O_{C^\infty }(x_{{n+1}}))) -{\Psi }_1(x,\alpha _{\xi '}) \end{aligned}$$

Written out explicitly, the last line in (3.8) says that for \(\gamma <1/2\),

$$\begin{aligned}&|u_h(x)| \\&\quad \lessapprox h^{-3{{n}}/4} \Big ( \int _{T^*M} e^{2 {{\text {Im}}\,}\overline{\varphi (\alpha ,y_c'(x,\alpha ))}/h} e^{-2 \psi (\alpha ) /h} \, d\alpha \Big )^{1/2} \\&\quad \lessapprox h^{-3{{n}}/4} \Big ( \int _{T^*M} e^{-[(|\alpha _{x'}-\mathrm Rey_c'|^2-|{{\text {Im}}\,}y_c'|^2)(1+O_{C^\infty }(x_{{n+1}}))+2\mathrm Re{\Psi }_1(x,\alpha _{\xi '})+2 \psi (\alpha )] /h} \, d\alpha \Big )^{1/2} \\&\quad \lessapprox h^{-3{{n}}/4} \Big ( \int _{T^*M} e^{-[(|\alpha _{x'}-\mathrm Rey_c'|^2-|{{\text {Im}}\,}y_c'|^2)(1+O_{C^\infty }(x_{{n+1}}))+2\mathrm Re{ \Psi }_1(x,\alpha _{\xi '})+2 \psi (\alpha )] /h} \, d\alpha \Big )^{1/2} \\&\quad \lessapprox h^{-3{{n}}/4} \Big ( \int _{S^{{{n-1}}}}\int _{0}^\infty \int _{M} e^{-[(|\alpha _{x'}-\mathrm Rey_c'|^2-|{{\text {Im}}\,}y_c'|^2)(1+O_{C^\infty }(x_{{n+1}}))+2\mathrm Re{\Psi }_1(x,r\omega )+ \gamma (r-1)^2] /h} \, \\&\qquad \mathrm{d}\alpha _{x'}\mathrm{d}r\mathrm{d}\omega \Big )^{1/2} \end{aligned}$$

Now, recall that \( {{\text {Im}}\,}y_c'=O(x_{{n+1}})\) and \(\partial _{\alpha }y_c'=O(x_{{n+1}})\) and hence applying the method of steepest descent, in the \(\alpha _{x'},r\) variables, the critical point of the phase occurs at

$$\begin{aligned}&2(\alpha _{x'}-\mathrm Rey_{c}')(I-\partial _{\alpha _{x'}}\mathrm Rey_c')+(\alpha _{x'}-\mathrm Rey_c')^2O(x_{{n+1}})\nonumber \\&\quad \quad -2({{\text {Im}}\,}y_c')\partial _{\alpha _{x'}}{{\text {Im}}\,}y_c'=O(x_{{n+1}}^3)\end{aligned}$$
(3.9)
$$\begin{aligned}&2(\alpha _{x'}-\mathrm Rey_{c'}')(-\partial _r\mathrm Rey_c')+(\alpha _{x'}-\mathrm Rey_c')^2O(x_{{n+1}})-2({{\text {Im}}\,}y_c')\partial _r {{\text {Im}}\,}y_c'\nonumber \\&\qquad +2\partial _r\mathrm Re{\Psi }_1(x,r\omega )+2\gamma (r-1)=O(x_{{n+1}}^3) \end{aligned}$$
(3.10)

Using (3.9), we have that at the stationary point

$$\begin{aligned} \alpha _{x'}-\mathrm Rey_c'=({{\text {Im}}\,}y_c')\partial _{\alpha _{x'}}{{\text {Im}}\,}y_c'+O(x_{{n+1}}^3). \end{aligned}$$

Putting this into (3.10) gives

$$\begin{aligned} 2\partial _r\mathrm Re{\Psi }_1(x,r\omega )+2\gamma (r-1)=2({{\text {Im}}\,}y_c')\partial _r{{\text {Im}}\,}y_c'(x,\alpha _x,r\omega )+O(x_{{n+1}}^3). \end{aligned}$$

Now,

$$\begin{aligned} 2\partial _r{{\text {Im}}\,}y_c'= & {} 0,\quad 2\partial _r\mathrm Re{\Psi }_1(x,r\omega )=a(x,\omega )\\ a(x,\omega )= & {} 2x_{{n+1}}+x_{{n+1}}^2\partial _{\xi '}Q(x',\omega )+x_{{n+1}}^2Q(x',\omega )+O(x_{{n+1}}^3) \end{aligned}$$

So, the critical point occurs at

$$\begin{aligned} r=1-\frac{a(x,\omega )}{2\gamma }+O(x_{{n+1}}^3). \end{aligned}$$

Noting that \({{\text {Im}}\,}y_c'=x_{{n+1}}+O(x_{{n+1}}^2)\) and evaluating the exponential at this point yields

$$\begin{aligned} |u_h(x)|&\lessapprox h^{-{{n}}/2+1/4}\left( \int _{S^{{{n-1}}}}e^{-2(x_{{n+1}}+x_{{n+1}}^2(Q(x',\omega )-\gamma ^{-1}-1)/2+O(x_{{n+1}}^3))/h}\mathrm{d}\omega \right) ^{1/2}\\&\lessapprox h^{-{{n}}/2+1/4}e^{-(x_{{n+1}}-x_{{n+1}}^2(-\inf _{\omega \in S^{{{n}}-1}}Q(x',\omega )+\gamma ^{-1}+1)/2+O(x_{{n+1}}^3))/h}. \end{aligned}$$

The same argument works for derivatives with each differentiation creating a power of \(h^{-1}.\)\(\square \)

Remark 3.1

We note that the first order bound with \(d(x) = d_{\partial \Omega }(x) + O( d_{\partial \Omega }^2(x))\) in Theorem 1 (1.3) follows from the weighted \(L^2\)-bound in Theorem 2 by essentially the same argument as above. The proof in that case is slightly simpler since one need only keep track of the analytic stationary phase computations to \(O(x_{{n+1}}^2)\) error (rather than \(O(x_{{n+1}}^3)\)).

4 Microlocal Estimates Near the Zero Section

While the Dirichlet-to-Neumann map is a homogeneous analytic pseudodifferential operator, it is not well behaved as a semiclassical analytic pseudodifferential operator near the zero section. In particular, when semiclassically rescaled, the full symbol of a homogeneous pseudodifferential operator, a, typically has singularities of the form

$$\begin{aligned} |\partial _\xi ^\alpha a(x,\xi /h)|\sim C_\alpha h^{-|\alpha |},\quad |\xi |\ll 1. \end{aligned}$$

Apriori, this singular h dependence may result in the transport of semiclassical analytic wavefront sets away from the zero section under the action of a (homogeneous) pseudodifferential operator. The purpose of this section is to show that no such transport occurs and then to use this information to estimate Steklov eigenfunctions near the zero section.

4.1 Cauchy Estimates and the Euclidean FBI Transform

For \(0\le \tilde{h}\le h\), let

$$\begin{aligned} T_{\tilde{h} }u(x,\xi ,h;\mu ):=\int e^{\frac{i}{\tilde{h}}\left\langle x-y,\xi \right\rangle -\frac{1}{2\tilde{h}}|x-y|^2}u(y)\mathrm{d}y. \end{aligned}$$

Let also

$$\begin{aligned} S_{\tilde{h}}v(x):=2^{-n}(\pi \tilde{h})^{-3n/2}\int e^{\frac{i}{\tilde{h}}\left\langle y-x,\eta \right\rangle -\frac{1}{2\tilde{h}}|x-y|^2}v(y,\eta )\mathrm{d}y\mathrm{d}\eta . \end{aligned}$$

Then for all \(u\in \mathcal {S}(\mathbb {R}^{{n}})\), \(u=S_{\tilde{h}}T_{\tilde{h}}u\).

Define

$$\begin{aligned} T'_{a,\mu ,\tilde{h}}u:=\int e^{\frac{i}{\tilde{h}}\left\langle x-y,\xi \right\rangle -\frac{\mu }{2\tilde{h}}|x-y|^2}a(x,y,\xi )u(y)\mathrm{d}y. \end{aligned}$$

The next proposition is very similar to [8, Proposition 2.1]. The difference is that we do not require \(\mu \ge 1\). Instead, we keep track of the dependence of various estimates on \(\mu \ge \mu _0\).

Proposition 4.1

Suppose that a has tempered growth in \((x,y,\xi )\) and \(0<\mu _0\le \mu <\mu _1\). Then suppose that for W a neighborhood of \((x_0,\xi _0)\),

$$\begin{aligned} \sup _{W}|T_{\tilde{h}} u |\le Ce^{-\delta /\tilde{h}}. \end{aligned}$$

Then there is a neighborhood V of \((x_0,\xi _0)\)

$$\begin{aligned} \sup _{{V}}|T'_{a,\mu ,\tilde{h}}u|\le Ce^{-c\min ( \mu _0,\mu _1^{-1},\delta )/\tilde{h}}. \end{aligned}$$

Proof Sketch

In order to prove the lemma, one writes \(T'_{a,\mu ,\tilde{h}}u=T'_{a,\mu ,\tilde{h}}S_{\tilde{h}}T_{\tilde{h}}u\). Then, one can easily estimate the kernel of \(T'_{a,\mu ,\tilde{h}}S_{\tilde{h}}\) using a simple change of variables. \(\square \)

The next proposition is similar to [8, Proposition 2.2], except that we obtain a quantitative estimate on distances to the zero section.

Proposition 4.2

Fix \(x_0\in M\), \(\varepsilon >0\). Let X be a neighborhood of \(x_0\) in M. Suppose that \(\Vert u\Vert _{L^\infty ({X})}\le Ch^{-N}\). Then the following are equivalent:

  • there exist \(C,c>0\), \(h_0>0\) and an \(\varepsilon \)-independent neighborhood, W, of \(x_0\) so that for every \(0\le \tilde{h}\le h\le h_0\),

    $$\begin{aligned} |T_{\tilde{h}}u(x,\xi ,h)|\le Ce^{-c /\tilde{h}}\Vert u\Vert _{L^\infty (X)},\quad x\in W ,\,|\xi |\ge \varepsilon , \end{aligned}$$
    (4.1)
  • there exists \(C_1>0\), \(h_0>0\) and an \(\varepsilon \)-independent neighborhood, W, of \(x_0\) and constant \(c_1>0\) so that for \(0<h<h_0\)

    $$\begin{aligned} \sup _\mathrm{Rex\in W,\,|{{\text {Im}}\,}x|\le c_1}|u|\le C_1e^{\frac{\varepsilon }{2h}}{\Vert u\Vert _{L^\infty (X)}} \end{aligned}$$
    (4.2)
  • there exists \(C>0\), \({h_0>0}\) and an \(\varepsilon \)-independent neighborhood, W, of \(x_0\) so that for \(0<h<h_0\)

    $$\begin{aligned} |(hD)^\alpha u{(x)}|\le C\Vert u\Vert _{L^\infty (X)}C^{|\alpha |}(h|\alpha |+\varepsilon )^{|\alpha |},\quad x\in W. \end{aligned}$$
    (4.3)

Proof Sketch

The fact that (4.2) implies (4.3) follows from basic Cauchy estimates, while that (4.3) implies (4.2) follows from writing out the Taylor formula with remainder.

The equivalence of (4.2) and (4.1) follows from writing down a resolution of the identity in terms of the FBI transform and deforming the contour into the complex domain. \(\square \)

The next estimate proves that the application of an analytic homogeneous pseudodifferential operator preserves estimates of the form (4.3). Roughly speaking, we show that the Sobolev mapping properties of such an operator behave like the Sobolev mapping properties of multiplication by an h-independent analytic function. Throughout the proof of the next proposition, we will use the following elementary estimates without comment

$$\begin{aligned} \left( 1+\frac{s}{t}\right) ^t\le & {} e^s\le \left( 1+\frac{s}{t}\right) ^{t+s},\quad t,s>0\end{aligned}$$
(4.4)
$$\begin{aligned} (ne)^{-|\alpha |}|\alpha |^{|\alpha |}\le & {} \alpha !\le |\alpha |^{|\alpha |},\quad \alpha \in \mathbb {N}^n \end{aligned}$$
(4.5)

Proposition 4.3

Let P be a homogeneous analytic pseudodifferential operator of order k. Suppose that u satisfies (4.3) in a neighborhood, U of \(x_0\) with some constant C. Then there exists a neighborhood, W of \(x_0\) so that Pu satisfies

$$\begin{aligned} |(hD)^\beta Pu(x)|\le h^{-k-{{n}}}C^{|\beta |+1}(\varepsilon +h|\beta |)^{|\beta |},\quad x\in W. \end{aligned}$$

Proof

Let \(\chi \in C_c(\mathbb {R})\) with \(\chi \equiv 1\) on \([-1,1]\) and \({{\text {supp}}}\chi \subset [-2,2]\). Then for any \(\delta >0\), the kernel of P is given by \(K_\delta (x,y)+R_\delta (x,y)\) where \(R_\delta \) is real analytic and

$$\begin{aligned} K_\delta (x,y)=(2\pi )^{-{{n}}}\int e^{i\left\langle x-y,\xi \right\rangle } {{p}}(x,y,\xi )\chi ((3\delta )^{-1}|x-y|)\mathrm{d}\xi \end{aligned}$$

with \({{p}}\in S^k_{ha}\).

The kernel of \(\partial _x^\beta P\) is given by \(\partial _x^\beta K+\partial _x^\beta R_\delta \). Since \(R_\delta \) is analytic,

$$\begin{aligned} \left| \int \partial _x^\beta R_\delta (x,y)u(y)\mathrm{d}y\right| \le \Vert \partial _x^\beta R(x,y)\Vert _{L^2_y}\Vert u\Vert _{L^2}\le C_\delta ^\beta \beta !\Vert u\Vert _{L^2}, \end{aligned}$$

so we need only consider \(\partial _x^\beta K_\delta \).

Observe that

$$\begin{aligned} \partial _x^\beta K_\delta (x,y)=(2\pi )^{-{{n}}}\int e^{i\left\langle x-y,\xi \right\rangle }\sum _{\beta '+\beta ''=\beta }\xi ^{\beta '}\partial _{x}^{\beta ''}({{p}}\chi ((3\delta )^{-1}|x-y|))\mathrm{d}\xi . \end{aligned}$$

Deforming the contour in \(\xi \) to

$$\begin{aligned} \xi \mapsto \xi +iR\frac{x-y}{|x-y|} \end{aligned}$$

for some \(R>0\), we can, modulo an analytic error, replace the kernel by

$$\begin{aligned} (2\pi )^{-{{n}}}\int e^{i\left\langle x-y,\xi \right\rangle }\sum _{\beta '+\beta ''=\beta }\xi ^{\beta '}\partial _{x}^{\beta ''}({{p}})\chi (\delta ^{-1}|x-y|)\mathrm{d}\xi . \end{aligned}$$
(4.6)

Let \(\partial _x^\beta \tilde{P}\) be the operator with kernel as in (4.6). Then, let \(\psi \in C_c^\infty (M)\) have support on a neighborhood, W of \(x_0\) so that \(W\subset U\) and \(d(W,\partial U)\ge 2\delta \) with U as in the statement of the proposition. By [29, Theorem 4.23] there exists \(N>0\) a constant (independent of \(h,\beta ,P\)) so that

$$\begin{aligned} \Vert \psi \partial _x^\beta \tilde{P}u\Vert _{L^2}&\le \sum _{\beta '+\beta ''=\beta }\Vert (\xi ^{\beta '}\left\langle \xi \right\rangle ^{-|\beta '|-{k}}\partial _x^{\beta ''}{{p}})(x,D)\Vert _{L^2\rightarrow L^2}\Vert u\Vert _{H^{k+|\beta '|}(U)}\\&\le \sum _{\beta '+\beta ''=\beta }\sum _{|\alpha |\le {{Nn}}}\Vert \partial ^\alpha (\xi ^{\beta '}\left\langle \xi \right\rangle ^{-|\beta '|-k}\partial _x^{\beta ''}{{p}})\Vert _{L^\infty }\Vert u\Vert _{H^{k+|\beta '|}(U)}\\&\le C\sum _{j=1}^{|\beta |}\sum _{|\alpha |\le {{Nn}}}\frac{|\beta |!}{(|\beta |-j)!j!}C^{|\beta |-j+|\alpha |}\\&\quad \times (|\beta |-j+|\alpha |)!h^{-k-j}C^{k+j}(\varepsilon +h(k+j))^{k+j}\\&\le C\sum _{j=1}^{|\beta |}\sum _{|\alpha |\le {{Nn}}}\frac{|\beta |!({n}e)^{|\beta |}}{(|\beta |-j)^{|\beta |-j}j^j}C^{|\beta |-j+|\alpha |}\\&\qquad (|\beta |-j+|\alpha |)^{|\beta |-j+|\alpha |}h^{-k-j}C^{k+j}(\varepsilon +h(k+j))^{k+j}\\ \Vert \psi \partial _x^\beta \tilde{P}u\Vert _{L^2}&\le C^{|\beta |+1}|\beta |!({n}e)^{|\beta |}\left( h+\frac{\varepsilon }{|\beta |}\right) ^{|\beta |}h^{-|\beta |-k}\\&\qquad \sum _{\begin{array}{c} 0\le j\le |\beta |\\ |\alpha |\le {Nn} \end{array}}j^{k}C^{|\beta |-j+|\alpha |}e^{|\alpha |} (|\beta |-j+|\alpha |)^{|\alpha |}C^{k+j}\\&\le C^{|\beta |+1}\left( h|\beta |+\varepsilon \right) ^{|\beta |}h^{-|\beta |-k} \end{aligned}$$

Relabeling \(|\beta |=|\beta |+{{n}}\) gives the desired estimates using Sobolev embeddings. The result follows from taking \(\psi \equiv 1\) on a slightly smaller neighborhood \(W'\subset W\). \(\square \)

Let \(\chi \in C_c^\infty (\mathbb {R})\) with \({{\text {supp}}}\chi \subset [-1,1]\), \(\varepsilon >0\), and define

$$\begin{aligned} S^{\varepsilon }_{\tilde{h}}v(x):=2^{-n}(\pi h)^{-3n/2}\int e^{\frac{i}{\tilde{h}}\left\langle y-x,\eta \right\rangle -\frac{1}{2\tilde{h}}|x-y|^2}\chi (\varepsilon ^{-1}|\eta |)v(y,\eta )\mathrm{d}y\mathrm{d}\eta . \end{aligned}$$

Proposition 4.4

Let \(\chi \in C_c^\infty (\mathbb {R})\) with \({{\text {supp}}}\chi \subset [-1,1]\). Then for for all \(\varepsilon >0\), \(|\xi |\ge 2\varepsilon \), and \(0<\tilde{h}\le h\),

$$\begin{aligned} |T_{\tilde{h}}S_h^{\varepsilon }u(x,\xi )|\le Ce^{-\frac{|\xi |^2}{2\tilde{h}}}\Vert u\Vert _{L^2}. \end{aligned}$$

Proof

The kernel of \(T_{\tilde{h}}S_h^{\varepsilon }\) is given by

$$\begin{aligned}&=2^{-n}(\pi h)^{-3n/2}\iint e^{\frac{i}{\tilde{h}}\left( \left\langle x-w,\xi \right\rangle +\frac{\tilde{h}}{h}\left\langle w-y,\eta \right\rangle \right) -\frac{1}{2\tilde{h}}\left( (x-w)^2+\frac{\tilde{h}}{h}(w-y)^2\right) }\chi (\varepsilon ^{-1}|\eta |)dw \end{aligned}$$

In particular, letting \(\mu =\frac{\tilde{h}}{h}\),

$$\begin{aligned} T_{\tilde{h}}S_h^{\varepsilon }v(x,\xi )=2^{-n}(\pi h)^{-3n/2}h^n\tilde{h}^{-n}\iint b(x,\xi ,y,\eta )v(y,\eta )\mathrm{d}y\mathrm{d}\eta , \end{aligned}$$

where

$$\begin{aligned} b(x,\xi ,y,\eta )= & {} e^{\frac{i}{\tilde{h}}\left\langle x-y,\xi +\mu ^2 \eta \right\rangle -\frac{\mu }{2(\mu +1)\tilde{h}}(x-y)^2-\frac{1}{2(\mu +1)\tilde{h}}(\mu \eta -\xi )^2}\\&\quad \quad \chi (\varepsilon ^{-1}|\eta |)\int e^{-\frac{\mu +1}{2\tilde{h}}\left( w-\left( \frac{\mu x+y+i(\mu \eta -\xi )}{\mu +1}\right) \right) ^2}dw \end{aligned}$$

Shifting contours shows that

$$\begin{aligned} |b(x,\xi ,y,\eta )|\le \chi (\varepsilon ^{-1}|\eta |)e^{-\frac{1}{2(\mu +1)\tilde{h}}(\mu \eta -\xi )^2}. \end{aligned}$$

On \({{\text {supp}}}\chi (\varepsilon ^{-1}|\eta |)\), \(|\eta |\le \varepsilon \) and \(|\mu \eta |\le \varepsilon \). Therefore, taking \(|\xi |\ge 2\varepsilon \) proves the lemma. \(\square \)

Our next proposition is the key proposition for this section and shows that homogeneous analytic pseudodifferential operators do not transport semiclassical analytic wavefront sets away from the zero section.

We say that u is compactly microlocalized if there exists \(\chi \in C_c^\infty (\mathbb {R})\) so that for some global FBI transform \(\tilde{T}\),

$$\begin{aligned} (1-\chi (|\alpha _{\xi }|_{\alpha _x}))\tilde{T}u=O(e^{-c/h}). \end{aligned}$$

Proposition 4.5

Let M be a compact, real-analytic manifold, and P be a homogeneous analytic pseudodifferential operator. Let \(\tilde{T}\) be a global FBI transform on M with left-parametrix \(\tilde{S}\). Let \(\chi _0,\chi _1\in C_c^\infty (\mathbb {R})\) have \(\chi _0\equiv 1\) on \([-1,1]\) with \({{\text {supp}}}\chi _0\subset [-2,2]\) and \(\chi _1\equiv 1\) on \([-{{N}},{{N}}]\) with \({{\text {supp}}}\chi _1\subset [-2{{N}},2{{N}}]\). Then, for \(\varepsilon >0\) small enough and \({{N}}>0\) large enough, and u compactly microlocalized, there exists \(c>0\) so that

$$\begin{aligned} (1-\chi _1(\varepsilon ^{-1}|\alpha _{\xi }|_{\alpha _x}))\tilde{T}Q\tilde{S}\chi _0(\varepsilon ^{-1}|\alpha _{\xi }|_{\alpha _x})\tilde{T}u=O_{ C^\infty }(e^{-c/h}). \end{aligned}$$

Proof

Notice that

$$\begin{aligned} \tilde{S}\chi _0(\varepsilon ^{-1}|\alpha _\xi |_{\alpha _x}))\tilde{T}u \end{aligned}$$

is microlocally vanishing for \(|\xi |\ge 2\varepsilon \). Therefore, for \(\psi \) with small enough support so that the Euclidean FBI transforms are well-defined,

$$\begin{aligned}&\psi \tilde{S}\chi _0(\varepsilon ^{-1}|\alpha _\xi |_{\alpha _x})\tilde{Tu}\\&\quad =\psi S_hT_h\tilde{S}\chi _0(\varepsilon ^{-1}|\alpha _\xi |_{\alpha _x})\tilde{T}u\\&\quad =\psi S_hT_h \tilde{S}\chi _0(\varepsilon ^{-1}|\alpha _\xi |_{\alpha _x})\chi (|\alpha _{\xi }|_{\alpha _x})\tilde{T}u+O{}(e^{-c/h})\\&\quad =\psi S_h\chi _0((2\varepsilon )^{-1}|\xi |)T_h \tilde{S}\chi _0(\varepsilon ^{-1}|\alpha _\xi |_{\alpha _x})\chi (|\alpha _{\xi }|_{\alpha _x})\tilde{T}u+O{}(e^{-c/h}) \end{aligned}$$

Now, by Proposition 4.4, (4.1) holds for the image of \(S_h\chi _0((2\varepsilon )^{-1}|\xi |)\). In particular, Lemma 4.3 applies and hence taking a partition of unity, \(Q\tilde{S}\chi _0(\varepsilon ^{-1}|\alpha _\xi |_{\alpha _x}))\tilde{T}u\) satisfies (4.3) and hence, for any \(\tilde{\psi }\) supported in a small enough region so that the Euclidean FBI transform is well-defined,

$$\begin{aligned} \tilde{\psi }Q\tilde{S}\chi _0(\varepsilon ^{-1}|\alpha _\xi |_{\alpha _x}))\tilde{T}u \end{aligned}$$

is microlocally vanishing on \(|\xi |\ge 2\varepsilon ^{-1}\) and since this is independent of the choice of FBI transform, taking M large enough gives

$$\begin{aligned} (1-\chi _1(\varepsilon ^{-1}|\alpha _{\xi }|_{\alpha _x}))\tilde{T}\tilde{\psi }Q\psi \tilde{S}\chi _0(\varepsilon ^{-1}|\alpha _{\xi }|_{\alpha _x})\tilde{T}u=O_{ C^\infty }(e^{-c/h}). \end{aligned}$$

Taking a partition of unity then proves the lemma. \(\square \)

4.2 Application to Eigenvalue Problems

Let Q be an (h-independent) homogeneous, elliptic, classical analytic pseudodifferential operator of order \(k>0\). Let \(\varphi _h\) denote a solution to

$$\begin{aligned} (h^kQ-1)\varphi _h=0. \end{aligned}$$

We are now in a position to analyze the behavior of \(\varphi _h\) near \(|\xi |=0\).

Notice that for any \(\tilde{\chi }\in C_c^\infty (\mathbb {R}^d)\) with \(\tilde{\chi }\equiv 1\) near 0, and \({{\text {supp}}}\tilde{\chi }\subset B(0,1)\), \((1-\tilde{\chi }(hD))h^kQ\) is a semiclassical pseudodifferential operator whose symbol is analytic in the interior of \(\{\tilde{\chi }= 0\}\). Thus, since Q is elliptic,

$$\begin{aligned} \sigma (h^k(1-\tilde{\chi }(hD))Q)|\ge c|\xi |^k,\quad |\xi |\ge 1 \end{aligned}$$

and hence by Proposition 2.3 for \(\chi \in C_c^\infty (\mathbb {R})\) with \(\chi \equiv 1\) on \([-2,2]\), for any \(\delta >0\), there exists \(c>0\) so that

$$\begin{aligned} (1-\chi (\delta ^{-1}q(\alpha _x,\alpha _\xi ))(1-\chi (\delta ^{-1}|\alpha _\xi |_{\alpha _x}))T_{geo}\varphi _h=O{}(e^{-c/h}). \end{aligned}$$
(4.7)

In the next proposition, we estimate the eigenfunctions \(\varphi _h\) near the zero section by their norm in a small annulus around the zero section. In particular, this shows that the cutoff \((1-\chi (\delta ^{-1}|\alpha _\xi |_{\alpha _x}))\) can be removed from (4.7).

Proposition 4.6

Let \(\varphi _h\) be a solution to

$$\begin{aligned} (h^kQ-1)\varphi _h=O_{L^2}(e^{-c/h}),\quad \Vert \varphi _h\Vert _{L^2}=1. \end{aligned}$$
(4.8)

For \(\varepsilon >0\), let \(\chi _\varepsilon =\chi (\varepsilon ^{-1}|\alpha _{\xi }|_{\alpha _x}).\) Then for \(\varepsilon >0\) small enough, there exists \(c>0\) so that

$$\begin{aligned} \Vert \chi _\varepsilon T_{geo}\varphi _h\Vert _{L^2}=O{}(e^{-c/h}). \end{aligned}$$

Proof

We first use the apriori estimate (4.7) together with Proposition 4.5 to decompose the eigenfunction equation (4.8) into two (essentially) orthogonal pieces—one supported near and one supported away from the zero section. We are then able to use a simple Neumann series argument to obtain estimates near the zero section.

Fix \(N>0\) to be chosen large enough later.

$$\begin{aligned} O(e^{-c/h})&=T_{geo}(h^kQ-1)S_{geo}T_{geo}\varphi _h=T_{geo}(h^kQ-1)S_{geo}(\chi _\varepsilon +(1-\chi _\varepsilon ))T_{geo}\varphi _h\nonumber \\&=T_{geo}(h^kQ-1)S_{geo}(\chi _\varepsilon +(1-\chi _{2^{3N}\varepsilon }))T_{geo}\varphi _h+O(e^{-c/h})\end{aligned}$$
(4.9)
$$\begin{aligned}&=(\chi _{2^{N}\varepsilon }T_{geo}h^kQS_{geo}\chi _{2\varepsilon }-1)\chi _\varepsilon T_{geo}\varphi _h\nonumber \\&\quad +\,(1-\chi _{2^{2N}\varepsilon })T_{geo}(h^kQ-1)S_{geo}(1-\chi _{2^{3N}\varepsilon })T_{geo}\varphi _h+O(e^{-c/h}) \end{aligned}$$
(4.10)

In (4.9), we used (4.7) to see that for any \(\psi \) with

$$\begin{aligned} {{\text {supp}}}\psi \subset \{|\sigma (q)(\alpha _x,\alpha _\xi )|<1\}\cap \{|\alpha _\xi |_{\alpha _x}>0\},\\ \psi T_{geo}\varphi _h=O{}(e^{-c/h}). \end{aligned}$$

In (4.10), we choose N large enough so that Proposition 4.5 implies

$$\begin{aligned} (1-\chi _{2^{N}\varepsilon })T_{geo}h^kQS_{geo}\chi _{2\varepsilon },\, \chi _{2^{2N}\varepsilon }T_{geo}h^kQS_{geo}(1-\chi _{2^{3N}\varepsilon })=O(e^{-c/h}). \end{aligned}$$

In order to invert \((\chi _{2^{N}\varepsilon }T_{geo}h^kQS_{geo}\chi _{2\varepsilon }-1)\) by Neumann series, we obtain estimates on \(\chi _{2^{N}\varepsilon }T_{geo}h^kQS_{geo}\chi _{2\varepsilon }\). Since \(S_{geo}\chi _{2\varepsilon }T_{geo}\) is the anti-Wick quantization of \(\chi _{2\varepsilon }\), it is a pseudodifferential operator with symbol \(\chi _{2\varepsilon }(\alpha ),\)

$$\begin{aligned} \Vert S_{geo}\chi _{2\varepsilon }T_{geo}\Vert _{L^2\rightarrow L^2}\le 1+O{}(h),\quad \Vert (h\nabla )^k S_{geo}\chi _{2\varepsilon }T_{geo}\Vert _{L^2\rightarrow L^2}\le C\varepsilon ^k +O{}(h). \end{aligned}$$

Therefore, \(\Vert Q\Vert _{H^k\rightarrow L^2}\le C\) implies

$$\begin{aligned} \Vert T_{geo}h^kQS_{geo}\chi _{2\varepsilon } T_{geo}\Vert _{L^2\rightarrow L^2}\le (C\varepsilon ^k+O{}(h)). \end{aligned}$$

In particular,

$$\begin{aligned} (\chi _{2^{N}\varepsilon }T_{geo}h^kQS_{geo}\chi _{2\varepsilon }-1)^{-1}=-\sum _{k=0}^\infty (\chi _{2^{N}\varepsilon }T_{geo}h^kQS_{geo}\chi _{2\varepsilon })^k. \end{aligned}$$
(4.11)

Hence, applying (4.11) on the left of (4.10)

$$\begin{aligned} \chi _\varepsilon T_{geo}\varphi _h +(1-\chi _{2^{2N}\varepsilon })T_{geo}(h^kQ-1)S_{geo}(1-\chi _{2^{3N}\varepsilon })T_{geo}\varphi _h=O{}(e^{-c/h}) \end{aligned}$$

and, multiplying by \(\chi _{\varepsilon /2}\) on the left, we have

$$\begin{aligned} \chi _{\varepsilon /2}T_{geo}\varphi _h=O{}(e^{-c/h}). \end{aligned}$$

\(\square \)