1 Introduction

Let (Xg) be a d-dimensional asymptotically Euclidean manifold. More explicitly, X belongs to a class of compact manifolds with boundary, whose interior is equipped with a Riemannian metric g which assumes a specific form close to the boundary \(\partial X\) (see Definition 29 in Section A.1 of “Appendix”). The elements of such class are also known as scattering manifolds, asymptotically conic manifolds, or manifolds with large conic ends. A typical example is the unit ball \({\mathbb {B}}^d\), equipped with a scattering metric.

On X, we consider a self-adjoint positive operator P, elliptic in the \({{\,\mathrm{SG}\,}}\)-calculus of order (mn) with \(m, n\in (0,\infty )\).Footnote 1 By the compact embedding of weighted Sobolev spaces, the resolvent is compact, and hence, the spectrum of P consists of a sequence of eigenvalues

$$\begin{aligned} 0 < \lambda _1 \le \lambda _2 \le \cdots \rightarrow +\infty . \end{aligned}$$

The goal of this article is to study the Weyl law of P, that is, the asymptotics of its counting function,

$$\begin{aligned} N(\lambda )= \#\{ j :\lambda _j < \lambda \}. \end{aligned}$$
(1)

Hörmander [18] proved, for a positive elliptic self-adjoint classical pseudodifferential operator of order \(m > 0\) on a compact manifold, the Weyl law

$$\begin{aligned} N(\lambda )=\gamma \cdot \lambda ^\frac{d}{m}+O\left( \lambda ^\frac{d-1}{m}\right) , \quad \lambda \rightarrow +\infty . \end{aligned}$$

It was pointed out that, in general, this is the sharp remainder estimate, since the exponent of \(\lambda \) in the remainder term cannot be improved for the Laplacian on the sphere. It was subsequently shown by Duistermaat and Guillemin [15] that, under a geometric assumption, there appears an additional term \(\gamma ' \lambda ^{(d-1)/m}\) and the remainder term becomes \(o(\lambda ^{(d-1)/m})\).

In the case of \({{\,\mathrm{SG}\,}}\)-operators on manifolds with cylindrical ends (see Definition 41 and the relationship with asymptotically Euclidean manifolds at the end of Section A.4 of “Appendix”), the leading order of the Weyl asymptotics was found by Maniccia and Panarese [22]. Battisti and Coriasco [3] improved the remainder estimate to \(O(\lambda ^{d/\max \{m,n\} - {\epsilon }})\) for some \({\epsilon }> 0\). For \(m \not = n\), Coriasco and Maniccia [10] proved the general sharp remainder estimate. We recall that \(\mathbb {R}^d\) is the simplest example of manifold with one cylindrical end.

In Theorem 1, we prove the analogue of Hörmander’s result for \(m = n\). This provides a more precise remainder term compared to the earlier result given in [3]. If the geodesic flow at infinity generated by the corner component \(p_ {\psi e} \) of the principal symbol of P is sufficiently generic, we have an even more refined estimate, parallel to the Duistermaat–Guillemin theorem, described in Theorem 2.

Theorem 1

Let \(P \in {{\,\mathrm{Op}\,}}{{\,\mathrm{SG}\,}}^{m,m}_\mathrm {cl}(X)\), \(m>0\), be a self-adjoint, positive, elliptic \({{\,\mathrm{SG}\,}}\)-classical pseudodifferential operator on an asymptotically Euclidean manifold X, and \(N(\lambda )\) its associated counting function. Then, the corresponding Weyl asymptotics reads as

$$\begin{aligned} N(\lambda ) = \gamma _2 \lambda ^{\frac{d}{m}}\log \lambda + \gamma _1 \lambda ^{\frac{d}{m}}+ O\left( \lambda ^{\frac{d-1}{m}}\log \lambda \right) . \end{aligned}$$

If \(X^o\) is a manifold with cylindrical ends, then the coefficients \(\gamma _j\), \(j=1,2\), are given by

$$\begin{aligned} \gamma _2&= \frac{{{\,\mathrm{TR}\,}}\left( P^{-\frac{d}{m}}\right) }{m\cdot d}, \\ \gamma _1&= \widehat{{{\,\mathrm{TR}\,}}}_{x,\xi }\left( P^{-\frac{d}{m}}\right) - \frac{{{\,\mathrm{TR}\,}}\left( P^{-\frac{d}{m}}\right) }{d^2}, \end{aligned}$$

where \({{\,\mathrm{TR}\,}}\) and \(\widehat{{{\,\mathrm{TR}\,}}}_{x,\xi }\) are suitable trace operators on the algebra of \({{\,\mathrm{SG}\,}}\)-operators on X.

Theorem 2

Let \(P \in {{\,\mathrm{Op}\,}}{{\,\mathrm{SG}\,}}^{m,m}_\mathrm {cl}(X)\) and \(N(\lambda )\) be as in Theorem 1. Denote by \(p_ {\psi e} \) the corner component of the principal symbol of P. If the set of periodic orbits of the Hamiltonian flow of \(\mathsf {X}_{f}\), \(f=(p_ {\psi e} )^\frac{1}{m}\), has measure zero on \({\mathcal {W}}^{\psi e}\), then we have the estimate

$$\begin{aligned} N(\lambda ) = \gamma _2 \lambda ^{\frac{d}{m}}\log \lambda + \gamma _1 \lambda ^\frac{d}{m} + \gamma _0 \lambda ^{\frac{d-1}{m}}\log \lambda + o\left( \lambda ^{\frac{d-1}{m}}\log \lambda \right) , \end{aligned}$$
(2)

with the coefficients \(\gamma _2\) and \(\gamma _1\) given in Theorem 1, and

$$\begin{aligned} \gamma _0 = \dfrac{{{\,\mathrm{TR}\,}}\left( P^{-\frac{d-1}{m}}\right) }{m\cdot (d-1)}, \end{aligned}$$

if \(X^o\) is a manifold with cylindrical ends.

Remark 3

The trace operators \({{\,\mathrm{TR}\,}}\) and \(\widehat{{{\,\mathrm{TR}\,}}}_{x,\xi }\) appearing in Theorems 1 and 2 were introduced in [3], see also Section A.3. in “Appendix”. The coefficient \(\gamma _0\) can be calculated as the Laurent coefficient of order \(-2\) at \(s = d-1\) of \(\zeta (s)\), the spectral \(\zeta \)-function associated with P.

Remark 4

To our best knowledge, this is the first result of a logarithmic Weyl law with the remainder being one order lower than the leading term. (We refer, for example, to [2] for a discussion of other settings with logarithmic Weyl laws.)

Remark 5

In view of the analysis at the end of Section A.4 in “Appendix”, we can apply our results to \({{\,\mathrm{SG}\,}}\)-operators on the manifold \(X={\mathbb {B}}^d\), equipped with an arbitrary scattering metric g, that is, to \({{\,\mathrm{SG}\,}}\)-operators on the manifold with one cylindrical end \(\mathbb {R}^d\), identified with \(({\mathbb {B}}^d)^o\) by radial compactification.

Next, we apply our results to the model operator P associated with the symbol \(p(x,\xi )=\left\langle x\right\rangle \!\cdot \!\left\langle \xi \right\rangle \), \(\left\langle z\right\rangle =\sqrt{1+|z|^2}\), \(z\in \mathbb {R}^d\), that is, \(P=\left\langle \cdot \right\rangle \sqrt{1-\varDelta }\). In particular, we observe that the condition on the underlying Hamiltonian flow in Theorem 2 is not satisfied and compute explicitly the coefficients \(\gamma _1\) and \(\gamma _2\).

Theorem 6

Let \(P = \left\langle \cdot \right\rangle \left\langle D\right\rangle \in {{\,\mathrm{Op}\,}}{{\,\mathrm{SG}\,}}^{1,1}_\mathrm {cl}(\mathbb {R}^d)\). Then,

$$\begin{aligned} N(\lambda ) = \gamma _2 \lambda ^d\log \lambda + \gamma _1 \lambda ^d + O(\lambda ^{d-1}\log \lambda ). \end{aligned}$$

Here, the coefficients are

$$\begin{aligned} \gamma _2&= \frac{[{{\,\mathrm{vol}\,}}(\mathbb {S}^{d-1})]^2}{(2\pi )^{d}} \cdot \frac{1}{d}, \\ \gamma _1&=- \frac{[{{\,\mathrm{vol}\,}}(\mathbb {S}^{d-1})]^2}{(2\pi )^{d}} \cdot \frac{1}{d}\cdot \!\left[ \varPsi \!\left( \frac{d}{2}\right) +\gamma +\frac{1}{d}\right] , \end{aligned}$$

where \(\displaystyle \gamma =\lim _{n\rightarrow +\infty }\left( \sum _{k=1}^n\frac{1}{k}-\log n\right) \) is the Euler–Mascheroni constant and

$$\begin{aligned} \varPsi (x)=\dfrac{d}{\mathrm{d}x}\log \varGamma (x) \end{aligned}$$
(3)

is the digamma function.

This implies that the Weyl asymptotics of the operator

$$\begin{aligned} Q = (1 + |x|^2) (1 - \varDelta ) \end{aligned}$$

is given by

$$\begin{aligned} N(\lambda ) = \frac{\gamma _2}{2} \lambda ^{\frac{d}{2}}\log \lambda + \gamma _1 \lambda ^{\frac{d}{2}} + O\left( \lambda ^{\frac{d-1}{2}}\log \lambda \right) , \end{aligned}$$

with the same coefficients given in Theorem 6.

Remark 7

It could be conjectured that many operators satisfying the assumptions of Theorem 1 also satisfy the additional geometric requirement which allows to obtain the refined Weyl formula (2) (cf. [14] for a proof of such fact in a different setting). However, we remark that it is still an open problem to construct explicitly an operator fulfilling the hypotheses of Theorem 2.

Remark 8

Operators like Q arise, for instance, as local representations of Schrödinger-type operators of the form \(H=-\varDelta _{\mathfrak {h}}+V\) on manifolds with ends, for appropriate choices of the metric \({\mathfrak {h}}\) and potential V (see Section A.4 in “Appendix” for a description of this class of manifolds adopted, e.g. in [3, 22]). We just sketch an example of construction of such an operator (see [8, Example 5.21] for the details). Consider the cylinder \(C=\{(u,v,z)\in \mathbb {R}^3:u^2+v^2=1, \, z>1\}={\mathbb {S}}^1\times (1,+\infty )\subset \mathbb {R}^3\) as the model of an end. Pulling back to the metric \({\mathfrak {h}}\) on C the metric \({\mathfrak {h}}^\prime \) on \(\mathbb {R}^3\) given by \(\mathfrak {h^\prime }=4^{-1}\mathrm {diag}(z^2\left\langle z\right\rangle ^{-n},z^2\left\langle z\right\rangle ^{-n}, 4\left\langle z\right\rangle ^{-n})\), \(n>0\), it turns out that, in suitable local coordinates \(x=(x_1,x_2)\in \mathbb {R}^2\) on C, the Laplace–Beltrami operator has the form

$$\begin{aligned} \varDelta _{\mathfrak {h}}=(1+x_1^2+x_2^2)^\frac{n}{2}(\partial _{x_1}^2 +\partial _{x_2}^2)=\left\langle x\right\rangle ^n\varDelta , \end{aligned}$$

with \(\varDelta \) the standard Laplacian. Choosing then, in local coordinates, \(V(x)=\left\langle x\right\rangle ^n\), we find

$$\begin{aligned} H=-\varDelta _{\mathfrak {h}}+V=\left\langle x\right\rangle ^n(1-\varDelta ). \end{aligned}$$

It is straightforward to see that \(H\in {{\,\mathrm{Op}\,}}{{\,\mathrm{SG}\,}}_\mathrm {cl}^{2,n}\) and it is elliptic (see Sect. 2), as claimed.

The proofs of Theorems 1 and 2 are broken into two parts. First, we will establish a connection between the wave trace near \(t=0\) and the zeta function, to calculate the coefficients of the wave trace. Then, we use a parametrix construction to relate the wave trace to the counting function.

The paper is organized as follows. In Sect. 2, we fix most of the notation used throughout the paper and recall the basic elements of the calculus of \({{\,\mathrm{SG}\,}}\)-classical pseudodifferential operators, the associated wavefront set, and the computation of the parametrix of Cauchy problems for \({{\,\mathrm{SG}\,}}\)-hyperbolic operators of order (1, 1). In particular, we quickly recall the invariance properties of the \({{\,\mathrm{SG}\,}}\)-calculus. In Sect. 3, we consider the wave trace of a \({{\,\mathrm{SG}\,}}\)-classical operator P of order (1, 1). Section 4 is devoted to study the relation between the wave trace and the spectral \(\zeta \)-function of P. In Sect. 5, we prove our main Theorems 1 and 2, while in Sect. 6, we examine the example given by the model operator \(P=\left\langle \cdot \right\rangle \left\langle D\right\rangle \) and prove Theorem 6. For the convenience of the reader, we conclude with an “Appendix”, including a few facts concerning asymptotically Euclidean manifolds and manifolds with cylindrical ends, including a comparison of the two notions at the end of Section A.4. We also give a short summary of the various trace operators and \({{\,\mathrm{SG}\,}}\)-Fourier integral operators.

2 SG-Calculus

The Fourier transform \({\mathcal {F}} : \mathcal {S}(\mathbb {R}^d) \rightarrow \mathcal {S}(\mathbb {R}^d)\) is defined by

$$\begin{aligned} ({\mathcal {F}} u)(\xi ) = {\widehat{u}}(\xi )=\int \hbox {e}^{-ix\xi }\,u(x)\,\mathrm{d}x, u \in \mathcal {S}(\mathbb {R}^d), \end{aligned}$$

and extends by duality to a bounded linear operator \({\mathcal {F}} : \mathcal {S}'(\mathbb {R}^d) \rightarrow \mathcal {S}'(\mathbb {R}^d)\).

The set of pseudodifferential operators \(A = a^w(x,D) = {{\,\mathrm{Op}\,}}^w(a) : \mathcal {S}(\mathbb {R}^d) \rightarrow \mathcal {S}'(\mathbb {R}^d)\) on \(\mathbb {R}^d\) with Weyl symbol \(a \in \mathcal {S}'(\mathbb {R}^{2d})\) can be defined through the Weyl quantizationFootnote 2

$$\begin{aligned} Au(x) = (2\pi )^{-d}\iint \hbox {e}^{i(x-y)\xi } a( (x+y)/2, \xi ) u(y) \mathrm{d}y\,\mathrm{d}\xi , \quad u \in \mathcal {S}(\mathbb {R}^d). \end{aligned}$$
(4)

A smooth function \(a \in \mathcal {C}^\infty (\mathbb {R}^d\times \mathbb {R}^d)\) is a SG-symbol of order \((m_\psi ,m_e) \in \mathbb {R}^2\), and we write \(a \in {{\,\mathrm{SG}\,}}^{m_\psi ,m_e}(\mathbb {R}^{2d})\), if for all multiindices \(\alpha ,\beta \in \mathbb {N}_0^d\), there exists \(C_{\alpha \beta }>0\) such that, for all \(x,\xi \in \mathbb {R}^d\),

$$\begin{aligned} \left|{\partial }_x^\alpha {\partial }_\xi ^\beta a(x,\xi )\right| \le C_{\alpha \beta } \left\langle \xi \right\rangle ^{m_\psi -|\beta |} \left\langle x\right\rangle ^{m_e - |\alpha |}. \end{aligned}$$
(5)

The space \({{\,\mathrm{SG}\,}}^{m_\psi ,m_e}(\mathbb {R}^{2d})\) becomes a Fréchet space with the seminorms being the best constants in (5). The space of all \({{\,\mathrm{SG}\,}}\)-pseudodifferential operators of order \((m_\psi ,m_e)\) is denoted by

$$\begin{aligned} {{\,\mathrm{Op}\,}}{{\,\mathrm{SG}\,}}^{m_\psi ,m_e}(\mathbb {R}^d) = \{ {{\,\mathrm{Op}\,}}^w(a) :a \in {{\,\mathrm{SG}\,}}^{m_\psi ,m_e}(\mathbb {R}^{2d})\}. \end{aligned}$$

The corresponding calculus was established in the 70s by Cordes and Parenti (see, for example, [4, 32]). The letter “G” in the notation, after the usual initial “S” for “symbol space”, stands for “global”. This calculus of symbols of product type, globally defined on \(\mathbb {R}^d\), was also considered by Shubin (see [38]). Actually, the \({{\,\mathrm{SG}\,}}\)-calculus on \(\mathbb {R}^d\) is a special case of the Weyl calculus (see [19, Sections 18.4-18.6]), associated with the slowly varying Riemannian metric on \(\mathbb {R}^{2d}\) given by

$$\begin{aligned} g_{(y,\eta )}(x,\xi )=\left\langle y\right\rangle ^{-2}|x|^2+\left\langle \eta \right\rangle ^{-2}|\xi |^2 \end{aligned}$$

(see, for example, [26, p. 71]; see also [8, Section 2.3] for more general \({{\,\mathrm{SG}\,}}\)-classes of symbols and operators on \(\mathbb {R}^d\) as elements of the Weyl calculus).

We list below some basic properties of SG-symbols and operators. (We refer to, for example, [4] and [30, Section 3.1] for an overview of the SG-calculus.) Some more information is provided in “Appendix”, for the convenience of the reader.

  1. 1.

    \(\displaystyle {{\,\mathrm{Op}\,}}{{\,\mathrm{SG}\,}}(\mathbb {R}^d)=\bigcup _{(m_\psi ,m_e)\in \mathbb {R}^2}{{\,\mathrm{Op}\,}}{{\,\mathrm{SG}\,}}^{m_\psi ,m_e}(\mathbb {R}^d)\) is a graded *-algebra; its elements are linear continuous operators from \(\mathcal {S}(\mathbb {R}^d)\) to itself, extendable to linear continuous operators from \(\mathcal {S}'(\mathbb {R}^d)\) to itself;

  2. 2.

    the differential operators of the form

    $$\begin{aligned} \sum _{|\alpha | \le m_e, |\beta | \le m_\psi } a_{\alpha ,\beta } x^\alpha D^\beta , \quad m_e,m_\psi \in \mathbb {N}_0, \end{aligned}$$
    (6)

    are \({{\,\mathrm{SG}\,}}\) operators of order \((m_\psi ,m_e)\);

  3. 3.

    if \(A \in {{\,\mathrm{Op}\,}}{{\,\mathrm{SG}\,}}^{0,0}(\mathbb {R}^d)\), then A extends to a bounded linear operator

    $$\begin{aligned} A : L^2(\mathbb {R}^d) \rightarrow L^2(\mathbb {R}^d); \end{aligned}$$
  4. 4.

    there is an associated scale of \({{\,\mathrm{SG}\,}}\)-Sobolev spaces (also known as Sobolev–Kato spaces), defined by

    $$\begin{aligned} H^{s_\psi ,s_e}(\mathbb {R}^d) = \{u \in \mathcal {S}'(\mathbb {R}^d) :\Vert \left\langle \cdot \right\rangle ^{s_e} \left\langle D\right\rangle ^{s_\psi } u\Vert _{L^2(\mathbb {R}^d)} < \infty \}, \end{aligned}$$

    and for all \(m_\psi ,m_e,s_\psi ,s_e\in \mathbb {R}\) the operator \(A \in {{\,\mathrm{Op}\,}}{{\,\mathrm{SG}\,}}^{m_\psi ,m_e}(\mathbb {R}^d)\) extends to a bounded linear operator

    $$\begin{aligned} A : H^{s_\psi ,s_e}(\mathbb {R}^d) \rightarrow H^{s_\psi -m_\psi ,s_e-m_e}(\mathbb {R}^d); \end{aligned}$$
  5. 5.

    the inclusions \(H^{s_\psi ,s_e}(\mathbb {R}^d)\subset H^{r_\psi ,r_e}(\mathbb {R}^d)\), \(s_\psi \ge r_\psi \), \(s_e\ge r_e\), are continuous, compact when the order components inequalities are both strict; moreover, the scale of the \({{\,\mathrm{SG}\,}}\)-Sobolev spaces is global, in the sense that

    $$\begin{aligned} \bigcup _{s_\psi ,s_e} H^{s_\psi ,s_e}(\mathbb {R}^d) = \mathcal {S}'(\mathbb {R}^d), \quad \bigcap _{s_\psi ,s_e} H^{s_\psi ,s_e}(\mathbb {R}^d) = \mathcal {S}(\mathbb {R}^d); \end{aligned}$$
  6. 6.

    an operator \(A = {{\,\mathrm{Op}\,}}^w(a) \in {{\,\mathrm{Op}\,}}{{\,\mathrm{SG}\,}}^{m_\psi ,m_e}(\mathbb {R}^d)\) is elliptic if its symbol a is invertible for \(|x|+|\xi | \ge R>0\), and \(\chi (|x|+|\xi |)[a(x,\xi )]^{-1}\) is a symbol in \({{\,\mathrm{SG}\,}}^{-m_\psi ,-m_e}(\mathbb {R}^{2d})\), where \(\chi \in \mathcal {C}^\infty (\mathbb {R})\) with \(\chi (t) = 1\) for \(t > 2R\) and \(\chi (t) = 0\) for \(t < R\);

  7. 7.

    if \(A\in {{\,\mathrm{Op}\,}}{{\,\mathrm{SG}\,}}^{m_\psi ,m_e}(\mathbb {R}^d)\) is an elliptic operator, then there is a parametrix \(B \in {{\,\mathrm{Op}\,}}{{\,\mathrm{SG}\,}}^{-m_\psi ,-m_e}(\mathbb {R}^d)\) such that

    $$\begin{aligned} AB - {{\,\mathrm{I}\,}}\in {{\,\mathrm{Op}\,}}{{\,\mathrm{SG}\,}}^{-\infty ,-\infty }(\mathbb {R}^d), \quad BA - {{\,\mathrm{I}\,}}\in {{\,\mathrm{Op}\,}}{{\,\mathrm{SG}\,}}^{-\infty ,-\infty }(\mathbb {R}^d). \end{aligned}$$

2.1 SG-Classical Symbols

We first introduce two classes of \({{\,\mathrm{SG}\,}}\)-symbols which are homogeneous in the large with respect either to the variable or the covariable. For any \(\rho >0\), \(x_0\in \mathbb {R}^d\), we let \(B_\rho (x_0)=\{x\in \mathbb {R}^d:|x-x_0|<\rho \}\) and we fix a cut-off function \(\omega \in \mathcal {C}_c^\infty (\mathbb {R}^d)\) with \(\omega \equiv 1\) on the ball \(B_\frac{1}{2}(0)\). For proofs, we refer to [30, Section 3.2].

  1. 1.

    A symbol \(a = a(x, \xi )\) belongs to the class \({{\,\mathrm{SG}\,}}^{m_\psi ,m_e}_{\mathrm {cl}(\xi )}(\mathbb {R}^{2d})\) if there exist functions \(a_{m_\psi -i, \cdot } (x, \xi )\), \(i=0,1,\ldots \), homogeneous of degree \(m_\psi -i\) with respect to the variable \(\xi \), smooth with respect to the variable x, such that,

    $$\begin{aligned} a(x, \xi ) - \sum _{i=0}^{M-1} (1-\omega (\xi )) \, a_{m_\psi -i, \cdot } (x, \xi )\in {{\,\mathrm{SG}\,}}^{m_\psi -M, m_e}(\mathbb {R}^{2d}), \quad M=1,2, \ldots \end{aligned}$$
  2. 2.

    A symbol a belongs to the class \({{\,\mathrm{SG}\,}}_{\mathrm {cl}(x)}^{m_\psi ,m_e}(\mathbb {R}^{2d})\) if \(a \circ R \in {{\,\mathrm{SG}\,}}^{m_e,m_\psi }_{\mathrm {cl}(\xi )}(\mathbb {R}^{2d})\), where \(R(x,\xi ) = (\xi , x)\). This means that \(a(x,\xi )\) has an asymptotic expansion into homogeneous terms in x.

Definition 9

A symbol a is called \({{\,\mathrm{SG}\,}}\)-classical, and we write \(a \in {{\,\mathrm{SG}\,}}_{\mathrm {cl}(x,\xi )}^{m_\psi ,m_e}(\mathbb {R}^{2d})={{\,\mathrm{SG}\,}}_{\mathrm {cl}}^{m_\psi ,m_e}(\mathbb {R}^{2d})\), if the following two conditions hold true:

  1. (i)

    there exist functions \(a_{m_\psi -j, \cdot } (x, \xi )\), homogeneous of degree \(m_\psi -j\) with respect to \(\xi \) and smooth in x, such that \((1-\omega (\xi )) a_{m_\psi -j, \cdot } (x, \xi )\in {{\,\mathrm{SG}\,}}_{\mathrm {cl}(x)}^{m_\psi -j, m_e}(\mathbb {R}^{2d})\) and

    $$\begin{aligned} a(x, \xi )- \sum _{j=0}^{M-1} (1-\omega (\xi )) \, a_{m_\psi -j, \cdot }(x, \xi ) \in {{\,\mathrm{SG}\,}}^{m_\psi -M, m_e}_{\mathrm {cl}(x)}(\mathbb {R}^{2d}), \quad M=1,2,\ldots ; \end{aligned}$$
  2. (ii)

    there exist functions \(a_{\cdot , m_e-k}(x, \xi )\), homogeneous of degree \(m_e-k\) with respect to the x and smooth in \(\xi \), such that \((1-\omega (x))a_{\cdot , m_e-k}(x, \xi )\in {{\,\mathrm{SG}\,}}_{\mathrm {cl}(\xi )}^{m_\psi , m_e-k}(\mathbb {R}^{2d})\) and

    $$\begin{aligned} a(x, \xi ) - \sum _{k=0}^{M-1} (1-\omega (x)) \, a_{\cdot , m_e-k} (x,\xi )\in {{\,\mathrm{SG}\,}}^{m_\psi , m_e-M}_{\mathrm {cl}(\xi )}(\mathbb {R}^{2d}), \quad M=1,2,\ldots \end{aligned}$$

Note that the definition of \({{\,\mathrm{SG}\,}}\)-classical symbol implies a condition of compatibility for the terms of the expansions with respect to x and \(\xi \). In fact, defining \(\sigma ^\psi _{m_\psi -j}\) and \(\sigma ^e_{m_e-i}\) on \({{\,\mathrm{SG}\,}}_{\mathrm {cl}(\xi )}^{m_\psi ,m_e}\) and \({{\,\mathrm{SG}\,}}_{\mathrm {cl}(x)}^{m_\psi ,m_e}\), respectively, as

$$\begin{aligned} \sigma ^\psi _{m_\psi -j}(a)(x, \xi )&= a_{m_\psi -j, \cdot }(x, \xi ),\quad j=0, 1, \ldots , \\ \sigma ^e_{m_e-k}(a)(x, \xi )&= a_{\cdot , m_e-k}(x, \xi ),\quad k=0, 1, \ldots , \end{aligned}$$

it possible to prove that (cf. [30, (3.2.7)])

$$\begin{aligned} \begin{aligned} a_{m_\psi -j,m_e-k}=\sigma ^{\psi e}_{m_\psi -j,m_e-k}(a)=\sigma ^\psi _{m_\psi -j}(\sigma ^e_{m_e-k}(a))= \sigma ^e_{m_e-k}(\sigma ^\psi _{m_\psi -j}(a)) \end{aligned} \end{aligned}$$

for all \(j,k \in \mathbb {N}_0\).

Moreover, the composition of two \({{\,\mathrm{SG}\,}}\)-classical operators is still classical. For \(A={{\,\mathrm{Op}\,}}{a}\in {{\,\mathrm{Op}\,}}{{\,\mathrm{SG}\,}}_\mathrm {cl}^{m_\psi ,m_e}(\mathbb {R}^d)\), the triple

$$\begin{aligned} \sigma (A) = (\sigma ^\psi (A),\sigma ^e(A),\sigma ^{\psi e}(A))=(a_\psi ,a_e,a_ {\psi e} ). \end{aligned}$$

where

$$\begin{aligned} \sigma ^\psi (A)(x,\xi ) = a_\psi (x,\xi )&= a_{m_\psi ,\cdot }\left( x,\frac{\xi }{|\xi |}\right) ,\\ \sigma ^e(A)(x,\xi ) = a_e(x,\xi )&= a_{\cdot ,m_e}\left( \frac{x}{|x|},\xi \right) ,\\ \sigma ^{\psi ,e}(A)(x,\xi ) = a_ {\psi e} (x,\xi )&= a_{m_\psi ,m_e}\left( \frac{x}{|x|},\frac{\xi }{|\xi |}\right) , \end{aligned}$$

is called the principal symbol of A. This definition keeps the usual multiplicative behaviour; that is, for any \(A\in {{\,\mathrm{Op}\,}}{{\,\mathrm{SG}\,}}_\mathrm {cl}^{m_\psi ,m_e}(\mathbb {R}^d)\), \(B\in {{\,\mathrm{Op}\,}}{{\,\mathrm{SG}\,}}_\mathrm {cl}^{r_\psi ,r_e}(\mathbb {R}^d)\), \((m_\psi ,m_e),(r_\psi ,r_e)\in \mathbb {R}^2\), the principal symbol of AB is given by

$$\begin{aligned} \sigma (AB) = \sigma (A) \cdot \sigma (B), \end{aligned}$$

where the product is taken componentwise. Proposition 10 allows to express the ellipticity of \({{\,\mathrm{SG}\,}}\)-classical operators in terms of their principal symbol.

Fixing a cut-off function \(\omega \in \mathcal {C}_c^\infty (\mathbb {R}^d)\) as above, we define the principal part of a to be

$$\begin{aligned} a_p(x,\xi ) = (1 - \omega (\xi ))a_\psi (x,\xi ) + (1 - \omega (x)) (a_e(x,\xi ) - (1 - \omega (\xi ))a_ {\psi e} (x,\xi )).\nonumber \\ \end{aligned}$$
(7)

2.2 SG-Wavefront Sets

We denote by \({\mathcal {W}}\) the disjoint union

$$\begin{aligned} {\mathcal {W}}= {\mathcal {W}}^\psi \sqcup {\mathcal {W}}^e\sqcup {\mathcal {W}}^{\psi e}=(\mathbb {R}^d \times \mathbb {S}^{d-1}) \sqcup (\mathbb {S}^{d-1}\times \mathbb {R}^d) \sqcup (\mathbb {S}^{d-1} \times \mathbb {S}^{d-1}), \end{aligned}$$

which may be viewed as the boundary of the (double) radial compactification of the phase space \(T^*\mathbb {R}^d\simeq \mathbb {R}^{d}\times \mathbb {R}^d\) (see, for example, [4] and “Appendix”). Therefore, it is natural to define smooth functions on \({\mathcal {W}}\) as follows:

$$\begin{aligned}&\mathcal {C}^\infty ({\mathcal {W}}) = \{ (f_\psi ,f_e,f_ {\psi e} ) \in \mathcal {C}^\infty ({\mathcal {W}}^\psi )\times \mathcal {C}^\infty ({\mathcal {W}}^e) \times \mathcal {C}^\infty ({\mathcal {W}}^{\psi e}) :\\&\quad \lim _{\lambda \rightarrow \infty } f_\psi (\lambda x, \xi ) = \lim _{\lambda \rightarrow \infty } f_e(x,\lambda \xi ) = f_ {\psi e} (x,\xi ) \text { for all } (x,\xi ) \in \mathbb {S}^{d-1}\times \mathbb {S}^{d-1}\}. \end{aligned}$$

By restriction, the principal symbol can be defined as a map \(\sigma : {{\,\mathrm{SG}\,}}_\mathrm {cl}^{m_\psi ,m_e}(\mathbb {R}^{2d})\ni a\mapsto \sigma (a) \in \mathcal {C}^\infty ({\mathcal {W}})\).

Proposition 10

An operator \(A\in {{\,\mathrm{Op}\,}}{{\,\mathrm{SG}\,}}_\mathrm {cl}^{m_\psi ,m_e}(\mathbb {R}^d)\) is elliptic if and only if \(\sigma (A)(x,\xi ) \not = 0\) for all \((x,\xi ) \in {\mathcal {W}}\).

For \(A \in {{\,\mathrm{Op}\,}}{{\,\mathrm{SG}\,}}_\mathrm {cl}^{m_\psi ,m_e}(\mathbb {R}^d)\) we define the following sets (see [9, 27]):

  1. 1.

    the elliptic set

    $$\begin{aligned} {{\,\mathrm{ell_{{{\,\mathrm{SG}\,}}}}\,}}(A) = \{ (x,\xi ) \in {\mathcal {W}}:\sigma (A)(x,\xi ) \not = 0\}, \end{aligned}$$
  2. 2.

    the characteristic set

    $$\begin{aligned} \Sigma _{{{\,\mathrm{SG}\,}}}(A) = {\mathcal {W}}{\setminus } {{\,\mathrm{ell_{{{\,\mathrm{SG}\,}}}}\,}}(A), \end{aligned}$$
  3. 3.

    the operator \({{\,\mathrm{SG}\,}}\)-wavefront set \({{\,\mathrm{WF}\,}}_{{{\,\mathrm{SG}\,}}} '(A) \subset {\mathcal {W}}\), via its complement: \((x,\xi ) \notin {{\,\mathrm{WF}\,}}_{{{\,\mathrm{SG}\,}}} '(A)\) if there exists \(B \in {{\,\mathrm{Op}\,}}{{\,\mathrm{SG}\,}}^{0,0}_\mathrm {cl}(\mathbb {R}^d)\) such that \((x,\xi ) \in {{\,\mathrm{ell_{{{\,\mathrm{SG}\,}}}}\,}}(B)\) and \(AB \in {{\,\mathrm{Op}\,}}{{\,\mathrm{SG}\,}}^{-\infty ,-\infty }(\mathbb {R}^d)\). More concisely,

    $$\begin{aligned} {{\,\mathrm{WF}\,}}_{{{\,\mathrm{SG}\,}}} '(A) = \bigcap _{\begin{array}{c} B \in {{\,\mathrm{Op}\,}}{{\,\mathrm{SG}\,}}_\mathrm {cl}^{0,0}\\ AB \in {{\,\mathrm{Op}\,}}{{\,\mathrm{SG}\,}}^{-\infty ,-\infty } \end{array}} \Sigma _{{{\,\mathrm{SG}\,}}}(B). \end{aligned}$$

The \({{\,\mathrm{SG}\,}}\)-wavefront set of a distribution \(u \in \mathcal {S}'(\mathbb {R}^d)\) is defined as

$$\begin{aligned} {{\,\mathrm{WF}\,}}_{{{\,\mathrm{SG}\,}}} (u) = \bigcap _{\begin{array}{c} A \in {{\,\mathrm{Op}\,}}{{\,\mathrm{SG}\,}}_\mathrm {cl}^{0,0}\\ Au \in \mathcal {S}(\mathbb {R}^d) \end{array}} \Sigma _{{{\,\mathrm{SG}\,}}}(A), \end{aligned}$$

see [4, 9, 27]. Following the concept of wavefront space by Cordes (see [4, Sect. 2.3]) and the approach in [9], we will decompose the \({{\,\mathrm{SG}\,}}\)-wavefront set of \(u\in \mathcal {S}'(\mathbb {R}^d)\) into its components in \({\mathcal {W}}\), namely,

$$\begin{aligned} {{\,\mathrm{WF}\,}}_{{{\,\mathrm{SG}\,}}} (u) = ({{\,\mathrm{WF}\,}}_{{{\,\mathrm{SG}\,}}} ^\psi (u), {{\,\mathrm{WF}\,}}_{{{\,\mathrm{SG}\,}}} ^e(u), {{\,\mathrm{WF}\,}}_{{{\,\mathrm{SG}\,}}} ^ {\psi e} (u)), \quad {{\,\mathrm{WF}\,}}_{{{\,\mathrm{SG}\,}}} ^\bullet (u) \subset {\mathcal {W}}^\bullet , \bullet \in \{\psi ,e, {\psi e} \}. \end{aligned}$$

Then, we have that

$$\begin{aligned} {{\,\mathrm{WF}\,}}_{{{\,\mathrm{SG}\,}}} ^\psi (u) = {{\,\mathrm{WF}\,}}_\mathrm {cl}(u), \end{aligned}$$

where \({{\,\mathrm{WF}\,}}_\mathrm {cl}(u)\) is the classical Hörmander’s wavefront set.

The \({{\,\mathrm{SG}\,}}\)-wavefront set is well behaved with respect to the Fourier transform (see, for example, [9, Lemma 2.4]):

$$\begin{aligned} (x,\xi ) \in {{\,\mathrm{WF}\,}}_{{{\,\mathrm{SG}\,}}} (u)\Longleftrightarrow (\xi ,-x)\in {{\,\mathrm{WF}\,}}_{{{\,\mathrm{SG}\,}}} ({\widehat{u}}). \end{aligned}$$

2.3 Complex Powers

As in the case of closed manifolds, it is possible to define complex powers of \({{\,\mathrm{SG}\,}}\)-pseudodifferential operators. We will only review the crucial properties of complex powers for a positive elliptic self-adjoint operator \(A \in {{\,\mathrm{Op}\,}}{{\,\mathrm{SG}\,}}_\mathrm {cl}^{m_\psi ,m_e}(\mathbb {R}^d)\), \(m_\psi ,m_e>0\). For the definition and proofs of the following properties, we refer to [3, Proposition 2.8, Theorems 3.1 and 4.2] (cf. also [23, 35]).

  1. (i)

    \(A^z A^s= A^{z+s}\) for all \(z, s \in \mathbb {C}\).

  2. (ii)

    \(A^k= \underbrace{A \circ \ldots \circ A}_{k \text { times}}\) for \(k \in \mathbb {N}_0\).

  3. (iii)

    If \(A \in {{\,\mathrm{Op}\,}}{{\,\mathrm{SG}\,}}_\mathrm {cl}^{m_\psi , m_e}(\mathbb {R}^d)\), then \(A^z \in {{\,\mathrm{Op}\,}}{{\,\mathrm{SG}\,}}_\mathrm {cl}^{m_\psi {\text {Re}}z, m_e {\text {Re}}z}(\mathbb {R}^d)\).

  4. (iv)

    If A is a classical \({{\,\mathrm{SG}\,}}\)-operator, then \(A^z\) is classical and its principal symbol is given by

    $$\begin{aligned} \sigma (A^z) = \sigma (A)^z. \end{aligned}$$
  5. (v)

    For \({\text {Re}}z < -d \cdot \min \{1/m_e, 1/m_\psi \}\), \(A^z\) is trace class.

For any \(A={{\,\mathrm{Op}\,}}(a)\in {{\,\mathrm{Op}\,}}{{\,\mathrm{SG}\,}}_\mathrm {cl}^{m_\psi , m_e}(\mathbb {R}^d)\) as above, the full symbolFootnote 3 of \(A^z\) will be denoted by

$$\begin{aligned} a(z) \in {{\,\mathrm{Op}\,}}{{\,\mathrm{SG}\,}}_\mathrm {cl}^{m_\psi {\text {Re}}z, m_e {\text {Re}}z}(\mathbb {R}^d). \end{aligned}$$

Let \(s \in \mathbb {C}\) with \({\text {Re}}(s) > \max \{d/m_e, d/m_\psi \}\). Using the property (v), it is possible to define \(\zeta (s)\) by

$$\begin{aligned} \zeta (s) = {{\,\mathrm{Tr}\,}}A^{-s} = \int K_{A^{-s}}(x, x) \mathrm{d}x = (2\pi )^{-d} \iint a(x,\xi ; -s) \mathrm{d}x\,\mathrm{d}\xi , \end{aligned}$$
(8)

where \(K_{A^z}\) is the Schwartz kernel of \(A^z\). We note that the \(\zeta \)-function may be written as

$$\begin{aligned} \zeta (s) = \sum _{j=1}^\infty \lambda _j^{-s}. \end{aligned}$$

with \((\lambda _j)_{j\in \mathbb {N}}\) the sequence of eigenvalues of A.

Theorem 11

(Battisti–Coriasco [3]) The function \(\zeta (s)\) is holomorphic for \({\text {Re}}(s)>d\cdot \max \{1/m_\psi , 1/m_e\}\). Moreover, it can be extended as a meromorphic function with possible poles at the points

$$\begin{aligned} s_j^1=\frac{d-j}{m_\psi }, \, j=0, 1, \ldots , \quad s^2_k=\frac{d-k}{m_e}, \, k=0, 1, \ldots \end{aligned}$$

Such poles can be of order two if and only if there exist integers jk such that

$$\begin{aligned} s_j^1=\frac{d-j}{m_\psi }=\frac{d-k}{m_e}=s_k^2. \end{aligned}$$
(9)

2.4 Parametrix of SG-Hyperbolic Cauchy Problems

Let \(P \in {{\,\mathrm{Op}\,}}{{\,\mathrm{SG}\,}}_\mathrm {cl}^{1,1}(\mathbb {R}^d)\) be a self-adjoint positive elliptic operator. By the construction from [11, Theorem 1.2] (cf. also [5, 6, 9]), it is possible to calculate a suitable parametrix for the Cauchy problem associated with the wave equation, namely

$$\begin{aligned} \left\{ \begin{aligned} (i{\partial }_t - P)u(t,x)&= 0\\ u(0,x)&= u_0(x). \end{aligned}\right. \end{aligned}$$
(10)

The solution operator of (10) exists by the spectral theorem and is denoted by \(U(t) = \hbox {e}^{-itP} = [\mathcal {F}_{\lambda \rightarrow t}(dE)](t)\), where dE is the spectral measure of P. There exists a short-time parametrix \({\widetilde{U}}(t)\), which is given by a regular family of \({{\,\mathrm{SG}\,}}\)-Fourier integral operators of type I (cf. Section A.2 in “Appendix”), defined through the integral kernel

$$\begin{aligned} K_{{\widetilde{U}}(t)}(x,y) = (2\pi )^{-d} \int \hbox {e}^{i(\phi (t,x,\xi ) - y\xi )} {\tilde{a}}(t,x,\xi ) \mathrm{d}\xi , \end{aligned}$$
(11)

where \({\tilde{a}} \in \mathcal {C}^\infty ( (-{\epsilon },{\epsilon }), {{\,\mathrm{SG}\,}}_\mathrm {cl}^{0,0})\) with \({\tilde{a}}(0) - 1 \in {{\,\mathrm{SG}\,}}^{-\infty ,-\infty }\) and \(\phi \in \mathcal {C}^\infty ( (-{\epsilon },{\epsilon }), {{\,\mathrm{SG}\,}}_\mathrm {cl}^{1,1})\).

The parametrix \({\widetilde{U}}(t)\) solves the wave equation (10) in the sense that \({\tilde{u}}(t,x) = [{\widetilde{U}}(t) u_0](x)\) satisfies

$$\begin{aligned} \left\{ \begin{aligned} (i{\partial }_t - P){\tilde{u}}(t)&\in \mathcal {C}^\infty ( (-{\epsilon },{\epsilon }), \mathcal {S}(\mathbb {R}^d)) \\ {\tilde{u}}(0) - u_0&\in \mathcal {S}(\mathbb {R}^d). \end{aligned}\right. \end{aligned}$$
(12)

By a Duhamel argument, \(U(t)-{\widetilde{U}}(t) \in \mathcal {C}^\infty ( (-{\epsilon }, {\epsilon }), {\mathcal {L}}(\mathcal {S}'(\mathbb {R}^d), \mathcal {S}(\mathbb {R}^d)))\), (cf. [10, Theorem 16], [14, p. 284]). Since the error term is regularizing, we obtain that

$$\begin{aligned} K_{U(t)}(x,y) = (2\pi )^{-d}\int \hbox {e}^{i(\phi (t,x,\xi ) - y\xi )} a(t,x,\xi ) \mathrm{d}\xi , \end{aligned}$$
(13)

for \(a \in \mathcal {C}^\infty ( (-{\epsilon },{\epsilon }), {{\,\mathrm{SG}\,}}_\mathrm {cl}^{0,0})\) with \(a(0) = 1\) (cf. [7, Lemma 4.14]).

Let p be the principal part of the full Weyl-quantized symbol of P. The phase function \(\phi \) satisfies the eikonal equation

$$\begin{aligned} \left\{ \begin{aligned} {\partial }_t \phi (t,x,\xi ) + p(x,\phi ^\prime _x(t,x,\xi ))&= 0\\ \phi (0,x,\xi )&= x\xi . \end{aligned}\right. \end{aligned}$$
(14)

This implies that we have a Taylor expansion in t of the form

$$\begin{aligned} \phi (t,x,\xi ) = x\xi - t p(x,\xi ) + t^2 \mathcal {C}^\infty (\mathbb {R}_t, {{\,\mathrm{SG}\,}}^{1,1}_\mathrm {cl}) \end{aligned}$$
(15)

for t small enough.

For any \(f \in \mathcal {C}^\infty (\mathbb {R}^{2d})\), we define the Hamiltonian vector field by

$$\begin{aligned} \mathsf {X}_f = \left\langle {\partial }_x f, {\partial }_\xi \right\rangle - \left\langle {\partial }_\xi f, {\partial }_x\right\rangle \end{aligned}$$

and we denote its flow by \(t \mapsto \exp (t\mathsf {X}_f)\). For \(P\in {{\,\mathrm{Op}\,}}{{\,\mathrm{SG}\,}}_\mathrm {cl}^{1,1}(\mathbb {R}^d)\), we will collectively denote by \(\mathsf {X}_{\sigma (P)}\) the Hamiltonian vector fields on \({\mathcal {W}}^\bullet \) generated by \(\sigma ^\bullet (P)\), \(\bullet \in \{\psi ,e, {\psi e} \}\), and by \(t \mapsto \exp (t\mathsf {X}_{\sigma (P)})\) the three corresponding flows.

By the group property, \(U(t+s) = U(t) U(s)\), we can extend propagation of singularities results for small times to \(t \in \mathbb {R}\). In [9], the propagation of the \({{\,\mathrm{SG}\,}}\)-wavefront set under the action of \({{\,\mathrm{SG}\,}}\)-classical operators and operator families like U(t) has been studied. In particular, the following theorem was proved there, by means of \({{\,\mathrm{SG}\,}}\)-Fourier integral operators (see also [8] and the principal-type propagation result for the scattering wavefront set [27, Proposition 7]).

Theorem 12

Let \(u_0\in \mathcal {S}'(\mathbb {R}^d)\) and \(U(t) = \hbox {e}^{-itP}\). Then,

$$\begin{aligned} {{\,\mathrm{WF}\,}}_{{{\,\mathrm{SG}\,}}} ^\bullet (U(t)u_0)\subseteq \varPhi ^\bullet (t)({{\,\mathrm{WF}\,}}_{{{\,\mathrm{SG}\,}}} ^\bullet (u_0)), \end{aligned}$$

where \(\varPhi ^\bullet \) is the smooth family of canonical transformations on \({\mathcal {W}}^\bullet \) generated by \(\sigma ^\bullet (\phi )\) with \(\bullet \in \{\psi ,e, {\psi e} \}\).

Remark 13

In view of (15), Theorem 12 can also be stated in the following way: For any \(u_0\in \mathcal {S}'(\mathbb {R}^d)\) and \(t\in (-{\epsilon }/2,{\epsilon }/2)\), \({{\,\mathrm{WF}\,}}_{{{\,\mathrm{SG}\,}}} ^\bullet (U(t)u_0)\subset \exp (t\mathsf {X}_{\sigma ^\bullet (p)})({{\,\mathrm{WF}\,}}_{{{\,\mathrm{SG}\,}}} ^\bullet (u_0))\), where \(\bullet \in \{\psi ,e, {\psi e} \}\), and \(\mathsf {X}_f\) is the Hamiltonian vector field generated by f. In the sequel, we will express this fact in the compact form

$$\begin{aligned} {{\,\mathrm{WF}\,}}_{{{\,\mathrm{SG}\,}}} (U(t)u_0)\subset \exp (t\mathsf {X}_{\sigma (p)})({{\,\mathrm{WF}\,}}_{{{\,\mathrm{SG}\,}}} (u_0)), \quad u_0\in \mathcal {S}'(\mathbb {R}^d), t\in \mathbb {R}. \end{aligned}$$

2.5 SG-Operators on Manifolds

In the 80s, Schrohe [34] introduced the class of \({{\,\mathrm{SG}\,}}\)-manifolds, whose elements admit \(\mathcal {C}^\infty \) structures associated with finite atlases, where the changes of coordinates satisfy suitable estimates of \({{\,\mathrm{SG}\,}}\)-type. This class includes non-compact manifolds, namely, for instance, the Euclidean space \(\mathbb {R}^d\), the infinite-holed torus (see, for example, [36, Page 25]), and the manifolds with ends (see, for example, [36, Page 27]), including those considered in [3, 22] (see Section A.4 in “Appendix”). In particular, Schrohe showed that the concepts of Schwartz functions and distributions, as well as of weighted Sobolev space, have an invariant meaning on \({{\,\mathrm{SG}\,}}\)-manifolds, and the same holds true for \({{\,\mathrm{SG}\,}}\)-operators. In fact, the corresponding symbol and operator classes are preserved by the admissible changes of coordinates (cf. also [5, Section 4.4], for an alternative proof of this property). Maniccia and Panarese [22] considered a class of manifolds with ends and showed that a type of \({{\,\mathrm{SG}\,}}\)-classical operators can be defined there, with the principal symbol triple having an invariant meaning. In [3], this was employed, with reference to \({{\,\mathrm{SG}\,}}\)-classical operators locally described by the symbols recalled in Sect. 2.1, to study the \(\zeta \)-function of \({{\,\mathrm{SG}\,}}\)-operators on manifolds with ends satisfying suitable ellipticity properties, as well as to extend to such environment the concepts of Wodzicki residue and of the trace operators introduced by Nicola [31]. The latter appear in the statements of Theorems 1 and 2. The classical \({{\,\mathrm{SG}\,}}\)-operators have also been employed by Schulze (see, for example, [37]), with the terminology symbols with exit behaviour, in some steps of the construction of pseudodifferential calculi on singular manifolds.

Melrose [26, 27] introduced the so-called scattering calculus on asymptotically Euclidean manifolds. As it can be seen in the quoted references (see, for example, [26, Sect. 6.3]), up to a different choice of compactification with respect to the one we employ (cf. Section A.1 in “Appendix”), and an opposite sign for the e-order of the symbols (that is, the order with respect to the x variable), the operators belonging to the scattering calculus are locally represented by \({{\,\mathrm{SG}\,}}\) pseudodifferential operators. In particular, the principal symbol of the classical operators has an invariant meaning, and the same holds true for the wavefront set (about the latter, see also the comparison in [9, Section 6] between the scattering wavefront set and the \({{\,\mathrm{SG}\,}}\)-wavefront set on \(\mathbb {R}^d\) recalled in Sect. 2.2).

In view of the above observations about the invariance property of the \({{\,\mathrm{SG}\,}}\)-calculus on asymptotically Euclidean manifolds, in the sequel we will mostly work and prove our results for the locally defined operators, that is, on \(\mathbb {R}^d\), with the global results following by a partition of unity and local coordinates argument.

3 Wave Trace

We fix a positive elliptic operator \(P \in {{\,\mathrm{Op}\,}}{{\,\mathrm{SG}\,}}_\mathrm {cl}^{1,1}(\mathbb {R}^{d})\) with \( {\psi e} \)-principal symbol \(p_ {\psi e} = \sigma ^ {\psi e} (P)\). By the compactness of the embedding of \({{\,\mathrm{SG}\,}}\)-Sobolev spaces, we have that the resolvent \((\lambda - P)^{-1}\) is compact for \(\lambda > 0\), and hence, there exists an orthonormal basis \(\{\psi _j\}\) of \(L^2\) consisting of eigenfunctions of P with eigenvalues \(\lambda _j\) with the property that

$$\begin{aligned} 0 < \lambda _1 \le \lambda _2 \le \dots \rightarrow + \infty . \end{aligned}$$

Therefore, the spectral measure is given by \(dE(\lambda ) = \sum _{j=1}^\infty \delta _{\lambda _j}(\lambda )\left\langle \cdot , \psi _j\right\rangle \psi _j\), where \(\delta _\mu \) is the delta distribution centred at \(\mu \), and we have that

$$\begin{aligned} N(\lambda ) = {{\,\mathrm{Tr}\,}}\int _0^\lambda dE(\lambda )={{\,\mathrm{Tr}\,}}E_\lambda . \end{aligned}$$

The wave trace w(t) is (formally) defined as

$$\begin{aligned} w(t) = {{\,\mathrm{Tr}\,}}U(t) = \sum _{j=1}^\infty \hbox {e}^{-it\lambda _j}. \end{aligned}$$

As usual, w(t) is well defined as a distribution by means of integration by parts and the fact that \(P^{-N}\) is trace class for \(N > d\) (cf. Schrohe [35, Theorem 2.4]).

Theorem 12 directly implies Lemma 14.

Lemma 14

Choose \(t_0 \in \mathbb {R}\). Let \(\varGamma \subset {\mathcal {W}}\) be an open subset and such that \(\left[ \exp (t\mathsf {X}_{\sigma (P)}) (\varGamma )\right] \cap \varGamma = \emptyset \), for all \(t\in (t_0-\delta ,t_0+\delta )\) and \(\delta > 0\) small. Then, for all \(B \in {{\,\mathrm{Op}\,}}{{\,\mathrm{SG}\,}}_\mathrm {cl}^{0,0}(\mathbb {R}^{d})\) with \({{\,\mathrm{WF}\,}}_{{{\,\mathrm{SG}\,}}} '(B) \subset \varGamma \), and all \(t\in (t_0-\delta ,t_0+\delta )\), we have that \(B U(t) B \in {\mathcal {L}}(\mathcal {S}'(\mathbb {R}^d), \mathcal {S}(\mathbb {R}^d))\).

We will show that the improvement of the Weyl law is only related to the corner component

$$\begin{aligned} \{ t \in \mathbb {R}:\exp (t\mathsf {X}_{\sigma ^ {\psi e} (P)})(x,\xi ) = (x,\xi ) \text { for some } (x,\xi ) \in {\mathcal {W}}^ {\psi e} \}. \end{aligned}$$

The structure of the singularities of w(t) is more involved. This comes from the fact that the boundary at infinity is not a manifold or equivalently the flow is not homogeneous. In contrast to the case of a closed manifold, the distribution w(t) will not be a conormal distribution near 0, but it turns out that it is a log-polyhomogeneous distribution.

Let \({\epsilon }> 0\) as in Sect. 2.4 and choose a function \(\chi \in \mathcal {S}(\mathbb {R})\) with \({{\,\mathrm{supp}\,}}{\hat{\chi }} \subset (-{\epsilon }, {\epsilon })\) and \({\hat{\chi }} = 1\) on \((-{\epsilon }/2,{\epsilon }/2)\).

Proposition 15

Let \(B \in {{\,\mathrm{Op}\,}}{{\,\mathrm{SG}\,}}_\mathrm {cl}^{0,0}\) and denote by \(N_B(\lambda ) = {{\,\mathrm{Tr}\,}}(E_\lambda B B^*)\) the microlocalized counting function. There exist coefficients \(w_{jk} \in \mathbb {R}\) with \(k \in \mathbb {N}_0\) and \(j\in \{0,1\}\) independent of \(\chi \) such that

$$\begin{aligned} (N_B * \chi )(\lambda ) \sim \sum _{k=0}^\infty \sum _{j=0,1} w_{jk} \lambda ^{d-k} (\log \lambda )^j \end{aligned}$$
(16)

as \(\lambda \rightarrow \infty \).

Remark 16

Note that \([\mathcal {F}(N_B')](t) = {{\,\mathrm{Tr}\,}}(U(t) B B^*)\).

Proof

From Sect. 2.4, through the calculus of \({{\,\mathrm{SG}\,}}\) FIOs (see [5] and Section A.2 in “Appendix”), we obtain that there is a parametrix \({\widetilde{U}}(t)\) for U(t) and we have the local representation of the kernel

$$\begin{aligned} K_{U(t) B B^*}(x,y) = (2\pi )^{-d} \int \hbox {e}^{i(\phi (t,x,\xi ) - y\xi )} a(t,x,\xi ) \,\mathrm{d}\xi \end{aligned}$$

for \(t \in (-{\epsilon },{\epsilon })\). The amplitude satisfies \(\sigma (a(0)) = \sigma (BB^*)\).

Set

$$\begin{aligned} \mathcal {T}_B(t) = {\hat{\chi }}(t) {{\,\mathrm{Tr}\,}}(U(t) B B^*). \end{aligned}$$

By the previous remark, we have that \(\mathcal {T}_B(t)\) is the Fourier transform of \((N_B' * \chi )(\lambda )\). We will now calculate the inverse Fourier transform of \(\mathcal {T}_B\).

Using the Taylor expansion of the phase function, we have that

$$\begin{aligned} \phi (t,x,\xi ) = x\xi + t\psi (t,x,\xi ), \end{aligned}$$

where \(\psi \) is smooth in t. Formally, we can write the trace as

$$\begin{aligned} \mathcal {T}_B(t) = (2\pi )^{-d} {\hat{\chi }}(t)\int \hbox {e}^{it\psi (t,x,\xi )} a(t,x,\xi ) \,\mathrm{d}x\,\mathrm{d}\xi . \end{aligned}$$

We proceed as in Hörmander [20, pp. 254–256], and we set

$$\begin{aligned} {\tilde{A}}_B(t,\lambda ) = (2\pi )^{-d} {\hat{\chi }}(t) \int _{\{-\psi (t,x,\xi ) \le \lambda \}} a(t,x,\xi ) \mathrm{d}x\,\mathrm{d}\xi . \end{aligned}$$
(17)

Note that ellipticity implies that \({\tilde{A}}_B(t,\lambda ) < \infty \).

Set \(X = {{\mathbb {B}}^d}\) and \({}^{\mathrm {sc}}\,\overline{T}^* X = {{\mathbb {B}}^d}\times {{\mathbb {B}}^d}\) with boundary defining functions \(\rho _X\) and \(\rho _\Xi \) as explained in Section A.1 of “Appendix”, and let

$$\begin{aligned} u(t,x,\xi ) = (2\pi )^{-d} {\hat{\chi }}(t) a(t,x,\xi )\,\mathrm{d}x\,\mathrm{d}\xi . \end{aligned}$$

Under the compactification \(\iota _2 = \iota \times \iota : \mathbb {R}^d\times \mathbb {R}^d \rightarrow {{\mathbb {B}}^d}\times {{\mathbb {B}}^d}= {}^{\mathrm {sc}}\,\overline{T}^* X\), we have that

$$\begin{aligned} (\iota _2)_* u \in \rho _X^{-d} \rho _\Xi ^{-d} \mathcal {C}^\infty ({}^{\mathrm {sc}}\,\overline{T}^* X, {}^b\Omega ). \end{aligned}$$

In the language of Melrose [25], \((\iota _2)_* u \in {\mathcal {A}}^{{\mathcal {K}}}_{{\text {phg}}}({}^{\mathrm {sc}}\,\overline{T}^* X, {}^b\Omega )\) with index set \({\mathcal {K}}\) given by \(K(\{\rho _X = 0\}) = K(\{\rho _\Xi = 0\}) = (-d + \mathbb {N}_0) \times \{0\}\).

It follows from (17) that

$$\begin{aligned} \partial _\lambda {\tilde{A}}_B(t,\lambda )&= \int _{-\psi = \lambda } u\\&= \left\langle (-\psi )_* u, \delta _\lambda \right\rangle . \end{aligned}$$

The function \({\tilde{\psi }}(t) = (\iota \circ (-\psi ) \circ \iota _2^{-1})(t) : {}^{\mathrm {sc}}\,\overline{T}^* X \rightarrow {\mathbb {B}}^1 =:{\overline{\mathbb {R}}}\) is a b-fibration with exponent matrix (1, 1) since the symbol \(\psi \) is of order (1, 1). Hence, the Push–Forward Theorem (cf. Melrose [25] and Grieser and Gruber [17]) implies that

$$\begin{aligned} (\iota \circ (-\psi ))_* u \in {\mathcal {A}}^{\mathcal {\psi _\# K}}_{{\text {phg}}}({\overline{\mathbb {R}}}, {}^b\Omega ), \end{aligned}$$

where \(\psi _\# {\mathcal {K}} = -d + \mathbb {N}_0 \times \{0,1\}\). Recall that the bundle \({}^b\Omega \) is generated by \(\rho ^{-1}d\rho \) and \(dy_j\) near the boundary \(\{\rho = 0\}\).

Pulling this distribution back to \(\mathbb {R}\) and pairing with \(\delta _\lambda \), we obtain the asymptotics

$$\begin{aligned} \partial _\lambda {\tilde{A}}_B(t,\lambda ) \sim \sum _{k=0}^\infty \sum _{j=0,1} a_{jk}(t) \lambda ^{d-1-k} (\log \lambda )^j + O(\lambda ^{-\infty }), \end{aligned}$$

where the coefficients \(a_{jk}\) are smooth and compactly supported in t. Defining \(A_B(\lambda ) = \hbox {e}^{iD_tD_\lambda }{\tilde{A}}_B(t,\lambda )|_{t=0}\), we find

$$\begin{aligned} \mathcal {T}_B(t) = \int _{\mathbb {R}} \hbox {e}^{-it\lambda } \partial _\lambda A_B(\lambda ) \mathrm{d}\lambda . \end{aligned}$$

The above calculation implies that \({\tilde{A}}_B\) and \(A_B\) have the asymptotics

$$\begin{aligned} A_B(\lambda ) = \sum _{k=0}^\infty \sum _{j=0,1} w_{jk}\lambda ^{d-k} (\log \lambda )^j + O(\lambda ^{-\infty }). \end{aligned}$$

We conclude that

$$\begin{aligned} (N_B * \chi )(\lambda )&= \int _{-\infty }^\lambda \mathcal {F}^{-1}_{t\rightarrow \lambda }\{\mathcal {T}_B\}(\lambda ) \,\mathrm{d}\lambda \\&= A_B(\lambda )\\&= \sum _{k=0}^\infty \sum _{j=0,1} w_{jk}\lambda ^{d-k} (\log \lambda )^j + O(\lambda ^{-\infty }). \end{aligned}$$

We note that the coefficients are determined by derivatives of \({\tilde{A}}_B(t,\lambda )\) at \(t = 0\) and since \({\hat{\chi }} = 1\) near \(t=0\), the specific choice of \(\chi \) does not change the coefficients. \(\square \)

4 Relation with the Spectral \(\zeta \)-Function

As in the case of pseudodifferential operators on closed manifolds (cf. Duistermaat and Guillemin [15, Corollary 2.2]), the wave trace at \(t = 0\) is related to the spectral \(\zeta \)-function. This relation extends to the \({{\,\mathrm{SG}\,}}\) setting.

Recall that for a positive self-adjoint elliptic operator \(P \in {{\,\mathrm{Op}\,}}{{\,\mathrm{SG}\,}}^{1,1}_\mathrm {cl}(\mathbb {R}^d)\), the function \(\zeta (s)\) is defined for \({\text {Re}}s > d\) by

$$\begin{aligned} \zeta (s) = {{\,\mathrm{Tr}\,}}P^{-s}. \end{aligned}$$

In addition, we consider the microlocalized version of \(\zeta (s)\), defined by

$$\begin{aligned} \zeta _B(s) = {{\,\mathrm{Tr}\,}}( P^{-s} B B^*) = \sum _{j=1}^\infty \lambda _j^{-s} \Vert B^* \psi _j\Vert ^2,\quad {\text {Re}}s > d, \end{aligned}$$

for \(B \in {{\,\mathrm{Op}\,}}{{\,\mathrm{SG}\,}}_\mathrm {cl}^{0,0}\). Of course, \(\zeta _{{{\,\mathrm{I}\,}}}(s) = \zeta (s)\).

By Theorem 11, \(\zeta (s)\) admits a meromorphic continuation to \(\mathbb {C}\) with poles of maximal order two at \(d-k\), \(k\in \mathbb {N}_0\). This result extends to \(\zeta _B(s)\), and we characterize the Laurent coefficients in terms of the wave trace expansion at \(t = 0\).

Proposition 17

The function \(\zeta _B(s)\) extends meromorphically to \(\mathbb {C}\) and has at most poles of order two at the points \(d - k\), \(k\in \mathbb {N}_0\). We have the expansion

$$\begin{aligned} \zeta _B(s) = \frac{A_{2,k}}{[s-(d-k)]^2} + \frac{A_{1,k}}{s-(d-k)} + f(s), \end{aligned}$$

where f is holomorphic near \(s = d-k\) and

$$\begin{aligned} \begin{aligned} A_{2,k}&= (d-k) w_{1k},\\ A_{1,k}&= w_{1k} + (d-k) w_{0k}, \end{aligned} \end{aligned}$$
(18)

where the \(w_{jk}\), \(k\in \mathbb {N}_0\), \(j=0,1\), are the coefficients appearing in the asymptotic expansion (16) of \(N_B(\lambda )\).

Proof

The meromorphic continuation and the possible location of the poles follow from similar arguments as in [3, Theorem 3.2] (see also the proof of Proposition 19). Hence, we only have to show that the poles are related to \(N_B(\lambda )\).

Let \({\epsilon }\in (0, \lambda _1)\) be sufficiently small. Choose an excision function \(\chi \in \mathcal {C}^\infty (\mathbb {R})\) such that \(\chi (\lambda ) = 0\) for \(\lambda < {\epsilon }\) and \(\chi (\lambda ) = 1\) for \(\lambda \ge \lambda _1\). Set \(\chi _s(\lambda ) = \chi (\lambda ) \lambda ^{-s}\). Then, using Remark 16,

$$\begin{aligned} \zeta _B(s) = \left\langle N_B', \chi _s\right\rangle = \left\langle {{\,\mathrm{Tr}\,}}(U(t)BB^*), \mathcal {F}^{-1}(\chi _s)\right\rangle . \end{aligned}$$

Let \(\rho \in \mathcal {S}(\mathbb {R})\) such that \(\rho \) is positive, \({\hat{\rho }}(0) = 1\), \({\hat{\rho }} \in \mathcal {C}_c^\infty (\mathbb {R})\), and \(\rho \) is even. By an argument similar to the one in [15, Corollary 2.2], we have that

$$\begin{aligned} \zeta _B(s) - \left\langle N_B' * \rho , \chi _s\right\rangle = \left\langle (1 - {\hat{\rho }}) {{\,\mathrm{Tr}\,}}(U(t)BB^*), \mathcal {F}^{-1}(\chi _s)\right\rangle \end{aligned}$$

is entire in s and polynomially bounded for \({\text {Re}}s > C\).

Now, we can insert the asymptotic expansion of \(N_B' * \rho \) to calculate the residues of \(\zeta _B(s)\). Taking the derivative of (16), we see that the asymptotic expansion of \(N_B' * \rho \) is given by

$$\begin{aligned} (N_B' * \rho )(\lambda ) = \sum _{k=0}^N\sum _{j=0,1} A_{j+1,k} \lambda ^{d-k-1} (\log \lambda )^j + o(\lambda ^{d-1-N}) \end{aligned}$$
(19)

for any \(N \in \mathbb {N}_0\) and \(A_{j,k}\) are given by (18).

Let \(k\in \mathbb {N}_0\) be arbitrary. If \(f \in \mathcal {C}^\infty (\mathbb {R})\) with \(f(\lambda ) = O(\lambda ^{d-k-1}\log \lambda )\) as \(\lambda \rightarrow \infty \), then \(\int f(\lambda ) \chi (\lambda ) \lambda ^{-s} \mathrm{d}\lambda \) is bounded and holomorphic in s for \({\text {Re}}s > d - k\). Let

$$\begin{aligned} I(s) = \int \lambda ^{d-k-1} \chi (\lambda ) \lambda ^{-s} \mathrm{d}\lambda . \end{aligned}$$

By partial integration, we obtain

$$\begin{aligned} I(s)&= \frac{\psi (s)}{s-(d-k)}. \end{aligned}$$

where \(\psi (s) = \int \lambda ^{d-k-s} \chi '(\lambda ) \mathrm{d}\lambda \) is holomorphic and \(\psi (d-k) = 1\). Therefore, we have

$$\begin{aligned} \int \lambda ^{d-k-s-1} (A_{1,k}+A_{2,k}\log \lambda ) \chi (\lambda ) \mathrm{d}\lambda = -A_{2,k} I'(s) + A_{1,k} I(s). \end{aligned}$$

Hence, the integral near \(s = d-k\) is given by

$$\begin{aligned} \int \lambda ^{d-k-s-1} (A_{1,k}+A_{2,k}\log \lambda ) \chi (\lambda ) \mathrm{d}\lambda = \frac{A_{2,k}}{[s-(d-k)]^2} + \frac{A_{1,k}}{s-(d-k)} + f(s), \end{aligned}$$

where f is holomorphic in a neighbourhood of \(s = d-k\). This shows that \(A_{k,2}\) and \(A_{k,1}\) as defined by (18) are the Laurent coefficients of \(\zeta (s)\) at \(s = d-k\)\(\square \)

The main advantage in employing the \(\zeta \)-function is that the coefficients are easier to calculate than for the wave trace.

Proposition 18

Let \(B \in {{\,\mathrm{Op}\,}}{{\,\mathrm{SG}\,}}_\mathrm {cl}^{0,0}\) with principal \( {\psi e} \)-symbol \(b_ {\psi e} \). The function \(\zeta _B(s)\) has a pole of order two at \(s = d\) with leading Laurent coefficient

$$\begin{aligned} (2\pi )^{-d} \int _{\mathbb {S}^{d-1}}\int _{\mathbb {S}^{d-1}} [p_ {\psi e} (\theta ,\omega )]^{-d} \cdot b_ {\psi e} (\theta ,\omega ) \mathrm{d}\theta \,\mathrm{d}\omega . \end{aligned}$$

Proof

This follows from the same arguments as in [3, Theorems 3.2 and 4.2] (cf. the proof of Proposition 19), with the modification that the full symbol is \(a(z) = p(z) \# b\), where p(z) denotes the full symbol of \(P^z\). The principal \( {\psi e} \)-symbol of \(A(z) = P^z B\) is given by \(a_{z,z}(x,\xi ;z) = [p_ {\psi e} (x,\xi )]^z \cdot b_ {\psi e} (x,\xi )\)\(\square \)

For the three-term asymptotics, we compute the third coefficient more explicitly.

Proposition 19

Let \(p(s) = p(x,\xi ; s)\) be the full symbol of \(P^s\). The leading Laurent coefficient of \(\zeta (s)\) at \(s = d - 1\) is given by

$$\begin{aligned} A_{2,1} = (2\pi )^{-d} \int _{\mathbb {S}^{d-1}}\int _{\mathbb {S}^{d-1}} p_{-d,-d}(\theta ,\omega ;-d+1) \mathrm{d}\theta \,\mathrm{d}\omega = {{\,\mathrm{TR}\,}}(P^{-(d-1)}). \end{aligned}$$

Remark 20

The equality \(A_{2,1} = {{\,\mathrm{TR}\,}}(P^{-(d-1)})\) also holds on any manifold with cylindrical ends.

Proof

The second equality follows directly from Proposition 40 in “Appendix”.

As in [3], we split the zeta function into four parts

$$\begin{aligned} \zeta (s)=\sum _{j=1}^4\zeta _j(s), \end{aligned}$$

where, for \({\text {Re}}s>d\),

$$\begin{aligned} \zeta _j(s)=(2\pi )^{-d}\int _{\Omega _j}p(x,\xi ;-s)dxd\xi \end{aligned}$$

and

$$\begin{aligned} \Omega _1&=\{(x, \xi ):|x|\le 1, |\xi |\le 1\},\quad \Omega _2=\{(x, \xi ):|x|\le 1, |\xi |> 1\},\\ \Omega _3&=\{(x, \xi ):|x|>1, |\xi |\le 1\},\quad \Omega _4=\{(x, \xi ):|x|>1, |\xi |>1\}. \end{aligned}$$

Let us recall the main aspects of the proof of the properties of the four terms \(\zeta _j(s)\), \(j=1,\ldots ,4\), shown in [3].

  1. 1.

    \(\zeta _1(s)\) is holomorphic, since we integrate \(p(-s)\), a holomorphic function in s and smooth with respect to \((x,\xi )\), on a bounded set with respect to \((x,\xi )\).

  2. 2.

    Let us first assume \({\text {Re}}s > d\). Using the expansion of \(p(-s)\) with \(M\ge 1\) terms homogeneous with respect to \(\xi \), switching to polar coordinates in \(\xi \) and integrating the radial part, one can write

    $$\begin{aligned} \zeta _2(s)&=(2\pi )^{-d} \sum _{j=0}^{M-1}\frac{1}{s-(d-j)} \int _{|x|\le 1}\int _{{\mathbb {S}}^{d-1}} p_{-s-j, \cdot } (x, \omega ;-s) \mathrm{d}\omega \mathrm{d}x \\&\quad +(2\pi )^{-d}\iint _{\Omega _2} r_{-s-M, \cdot }(x, \xi ; -s) \mathrm{d}\xi \mathrm{d}x. \end{aligned}$$

    Notice that the last integral is convergent and provides a holomorphic function in s. Arguing similarly to the case of operators on smooth, compact manifolds, \(\zeta _2(s)\) turns out to be holomorphic for \({\text {Re}}(s)>d\), extendable as a meromorphic function to the whole complex plane with, at most, simple poles at the points \(s^1_j=d-j\), \(j=0,1,2,\ldots \)

  3. 3.

    Using now the expansion of \(p(-s)\) with respect to x, exchanging the role of variable and covariable with respect to the previous point, again first assuming \({\text {Re}}s > d\) and choosing \(M\ge 1\), one can write

    $$\begin{aligned} \zeta _3(s)&= (2\pi )^{-d} \sum _{k=0}^{M-1}\frac{1}{s-(d-k)}\int _{{\mathbb {S}}^{d-1}} \int _{|\xi |\le 1} p_{\cdot , -s-k}(\theta , \xi ;-s)\mathrm{d}\xi \mathrm{d}\theta \\&\quad +(2\pi )^{-d}\iint _{\Omega _3} t_{\cdot , -s-M}(x,\xi ;-s) \mathrm{d}\xi \mathrm{d}x. \end{aligned}$$

    Arguing as in point 2, \(\zeta _3(s)\) turns out to be holomorphic for \({\text {Re}}s>d\), extendable as a meromorphic function to the whole complex plane with, at most, simple poles at the points \(s^2_k= d-k\), \(k=0,1,2,\ldots \)

  4. 4.

    To treat the last term, both the expansions with respect to x and with respect to \(\xi \) are needed. We assume that \({\text {Re}}s > d\) and choose \(M \ge 1\). We argue as in point 2 to obtain

    $$\begin{aligned} \zeta _4(s)&= (2\pi )^{-d}\sum _{j=0}^{M-1} \frac{1}{s-(d-j)}\int _{|x|\ge 1} \int _{{\mathbb {S}}^{d-1}}p_{-s-j, \cdot }(x, \omega ;-s)\mathrm{d}\omega \mathrm{d}x\\&\quad + (2\pi )^{-d}\iint _{\Omega _4} r_{-s-M, \cdot }(x, \xi ;-s)\mathrm{d}\xi \mathrm{d}x. \end{aligned}$$

    Now, we introduce the expansion with respect to x, switching to polar coordinates and integrating the x-radial variable in the homogeneous terms, for both integrals

    $$\begin{aligned}&\int _{|x|\ge 1}\int _{{\mathbb {S}}^{d-1}}p_{-s-j, \cdot }(x, \omega ;-s)\mathrm{d}\omega \mathrm{d}x\\&\quad = \sum _{k=0}^{M-1} \frac{1}{s-(d-k)} \int _{\mathbb {S}^{d-1}}\int _{{\mathbb {S}}^{d-1}} p_{-s-j, -s-k}(\theta , \omega ;-s) \mathrm{d}\theta \mathrm{d}\omega \\&\qquad + \int _{|x|\ge 1} \int _{{\mathbb {S}}^{d-1}} t_{-s-j, -s-M}(x, \omega ;-s) \mathrm{d}x \mathrm{d}\omega \end{aligned}$$

    and

    $$\begin{aligned}&\iint _{\Omega _4} r_{-s-M, \cdot }(x, \xi ;-s)\mathrm{d}\xi \mathrm{d}x\\&\quad = \sum _{k=0}^{M-1} \frac{1}{s-(d-k)}\int _{{\mathbb {S}}^{d-1}}\int _{|\xi |\ge 1} r_{-s-M, -s-k}(\theta ,\xi ; -s) \mathrm{d}\xi \mathrm{d}\theta \\&\qquad + \iint _{\Omega _4}r_{-s-M, -s-M}(x, \xi ; -s)\mathrm{d}x \mathrm{d}\xi . \end{aligned}$$

    We end up with

    $$\begin{aligned} \zeta _4(s)&= \sum _{k=0}^{M-1} \sum _{j=0}^{M-1} \frac{1}{s-(d-j)} \frac{1}{s-(d-k)} I_{j}^{k}(s)\\&\quad + \sum _{j=0}^{M-1} \frac{1}{s-(d-j)} R_j^M(s) +\sum _{k=0}^{M-1} \frac{1}{s-(d-k)} R_M^k(s) + R^M_M(s), \end{aligned}$$

where

$$\begin{aligned} I_{j}^{k}(s)=(2\pi )^{-d} \int _{{\mathbb {S}}^{d-1}}\int _{{\mathbb {S}}^{d-1}} p_{-s-j, -s-k}(\theta ', \theta ; -s)\mathrm{d}\theta \mathrm{d}\theta ', \end{aligned}$$

and \(R^j_M\), \(R^M_k\), \(R_M^M\) are holomorphic in s for \({\text {Re}}s > M + d\), \(j,k=0,\ldots , M-1\). It follows that \(\zeta _4(s)\) is holomorphic for \({\text {Re}}(s)>d\) and can be extended as a meromorphic function to the whole complex plane with, at most, poles at the points \(s^1_j=d-j\), \(s^2_k= d-k\) with \(j,k \in \mathbb {N}_0\). Clearly, such poles can be of order two if and only if \(j=k\) (cf. Theorem 11).

In view of the properties of \(\zeta (s)\) recalled above, the limit

$$\begin{aligned} \lim _{s\rightarrow d-1}[s-(d-1)]^2\zeta (s)=\lim _{s\rightarrow d-1}[s-(d-1)]^2\zeta _4(s)=I^1_1(d-1) \end{aligned}$$

proves the desired claim. \(\square \)

5 Proof of the Main Theorems

We choose a positive function \(\rho \in \mathcal {S}(\mathbb {R})\) such that \({\hat{\rho }}(0) = 1\), \({{\,\mathrm{supp}\,}}{\hat{\rho }} \subset [-1,1]\), and \(\rho \) is even. For \(T > 0\), we set \(\rho _T(\lambda ) :=T \rho (T\lambda )\), which implies that \({\hat{\rho _T}}(t) = {\hat{\rho }}(t/T)\). Let \(\nu > 0\) be arbitrary. Then, it is possible to prove the next Tauberian theorem by following the proof in [33, Appendix B].

Theorem 21

(Tauberian theorem) Let \(N : \mathbb {R}\rightarrow \mathbb {R}\) such that N is monotonically non-decreasing, \(N(\lambda ) = 0\) for \(\lambda \le 0\), and is polynomially bounded as \(\lambda \rightarrow +\infty \). If

$$\begin{aligned} (\partial _\lambda N * \rho _T)(\lambda ) \le C_1 \lambda ^\nu \log \lambda , \quad \lambda \ge T^{-1} \end{aligned}$$

for \(C_1 > 0\), then

$$\begin{aligned} \left|N(\lambda ) - (N * \rho _T)(\lambda )\right| \le C\,C_1 T^{-1} \lambda ^\nu \log \lambda , \quad \lambda \ge T^{-1}. \end{aligned}$$

Proof of Theorem 1

The first part of Theorem 1 follows directly from the Tauberian theorem and Proposition 15, due to the identity

$$\begin{aligned}{}[\mathcal {F}(N^\prime )](t) = {{\,\mathrm{Tr}\,}}\hbox {e}^{-itP} \end{aligned}$$

and using a partition of unity to locally represent \(U(t) = \hbox {e}^{-itP}\) as a smooth family of \({{\,\mathrm{SG}\,}}\)-Fourier integral operators, as discussed in Sect. 2.4. To calculate the coefficients \(w_{j,0}\), we use Proposition 17, to see that

$$\begin{aligned} w_{1,0}&= \frac{A_{2,0}}{d},\\ w_{0,0}&= \frac{A_{1,0}}{d} - \frac{A_{2,0}}{d^2}. \end{aligned}$$

From the definition of the traces, recalled in Definition 39 in “Appendix” (see also [3, p. 247]), we have that

$$\begin{aligned} A_{2,0}&= {{\,\mathrm{TR}\,}}(P^{-d}),\\ A_{1,0}&= d \cdot \widehat{{{\,\mathrm{TR}\,}}}_{x,\xi }(P^{-d}), \end{aligned}$$

which gives the claimed coefficients. \(\square \)

To prove Theorem 2 it suffices to prove that

$$\begin{aligned} N(\lambda ) = (N*\rho )(\lambda ) + o(\lambda ^{d-1}\log \lambda ), \end{aligned}$$

where \((N*\rho )(\lambda )\) is obtained through Propositions 15 and 17.

We define the microlocal return time function \(\varPi : {\mathcal {W}}\rightarrow \mathbb {R}_+ \cup \{\infty \}\) by

$$\begin{aligned} \varPi (x,\xi ) = \inf \{t > 0 :\exp (t\mathsf {X}_{\sigma (P)})(x,\xi ) = (x,\xi )\}, \end{aligned}$$

and \(\varPi (x,\xi ) = \infty \) if no such t exists. For a set \(\varGamma \in {\mathcal {W}}\), we set \(\varPi _\varGamma = \inf _{z \in \varGamma } \varPi (z)\).

We will need a microlocalized version of the Poisson relation.

Proposition 22

Let \(\varGamma \subset {\mathcal {W}}\) and \({\hat{\chi }} \in \mathcal {C}_c^\infty (\mathbb {R})\) with \({{\,\mathrm{supp}\,}}{\hat{\chi }} \subset (0,\varPi _\varGamma )\). For all \(B \in {{\,\mathrm{Op}\,}}{{\,\mathrm{SG}\,}}_\mathrm {cl}^{0,0}\) with \({{\,\mathrm{WF}\,}}_{{{\,\mathrm{SG}\,}}} '(B) \subset \varGamma \), we have that

$$\begin{aligned} {\hat{\chi }}(t) {{\,\mathrm{Tr}\,}}(U(t) B B^*) \in \mathcal {C}_c^\infty (\mathbb {R}). \end{aligned}$$

In particular, \((\chi * N_B')(\lambda ) \in O(\lambda ^{-\infty }).\)

The proof is a standard argument (cf. Wunsch [39]) and is only sketched here.

Proof of Proposition 22

For \(t_0 \in {{\,\mathrm{supp}\,}}{\hat{\chi }}\) and \((x,\xi ) \in \varGamma \), we choose a conic neighbourhood U of \((x,\xi )\) such that

$$\begin{aligned} {[}\varPhi (t) U] \cap U = \emptyset \end{aligned}$$

for all \(t \in (t_0-{\epsilon },t_0+{\epsilon })\) with \({\epsilon }> 0\) sufficiently small. The existence of this neighbourhood is guaranteed by the conditions on \(\varGamma \) and \({{\,\mathrm{supp}\,}}{\hat{\chi }}\). Choose \({{\tilde{B}}} \in {{\,\mathrm{Op}\,}}{{\,\mathrm{SG}\,}}_\mathrm {cl}^{0,0}\) with \({{\,\mathrm{WF}\,}}_{{{\,\mathrm{SG}\,}}} '({{\tilde{B}}}) \subset U\). Lemma 14 implies that for any \(k \in \mathbb {N}\),

$$\begin{aligned} {\partial }_t^k \left( {\tilde{B}} U(t) {\tilde{B}}\right) = {\tilde{B}} P^k U(t) {\tilde{B}} \in {\mathcal {L}}(\mathcal {S}'(\mathbb {R}^d),\mathcal {S}(\mathbb {R}^d)), \end{aligned}$$

hence \({\tilde{B}} U(t) {\tilde{B}}\) and all its derivatives are trace class. We obtain the claim by using a partition of unity. \(\square \)

We also define the modified return time

$$\begin{aligned} {\tilde{\varPi }}(x,\xi ) = \max \{\varPi (x,\xi ), {\epsilon }\}, \end{aligned}$$

where \({\epsilon }\) is given as in (12), and set \({\tilde{\varPi }}_\varGamma =\inf _{z\in \varGamma }{\tilde{\varPi }}(z)\). The main tool to prove Theorem 2 is Proposition 23.

Proposition 23

It holds true that

$$\begin{aligned} \limsup _{\lambda \rightarrow \infty } \frac{|N(\lambda ) - (N * \rho )(\lambda )|}{\lambda ^{d-1}\log \lambda } \le C \int _{{\mathcal {W}}^{\psi e}} {\tilde{\varPi }}(x,\xi )^{-1} \frac{\mathrm{d}S}{p_{1,1}(x,\xi )}. \end{aligned}$$

Proof of Theorem 2

The claim follows immediately by Proposition 23, since the assumptions imply that \(\varPi (x,\xi )^{-1} = 0\) almost everywhere on \({\mathcal {W}}^{\psi e}\). From Proposition 19, we obtain the coefficient \(\gamma _0\). \(\square \)

Proof of Proposition 23

Consider an open covering \(\{\varGamma ^\bullet _j\}\) of \({\mathcal {W}}\) with \(\bullet \in \{\psi ,e, {\psi e} \}\) and \(j \in \{1,\cdots , n_\bullet \}\) such that \(\varGamma _j^\psi \subset {\mathcal {W}}^\psi \) and \(\varGamma _j^e \subset {\mathcal {W}}^e\) do not intersect \({\mathcal {W}}^{\psi e}\), and \(\varGamma _j^ {\psi e} \cap {\mathcal {W}}^ {\psi e} \not =\emptyset \).

We consider a partition of unity on the level of operators such that

$$\begin{aligned} I = \sum _{j=1}^{n_\psi } A^\psi _j (A^\psi _j)^* + \sum _{j=1}^{n_e} A^e_j(A^e_j)^* + \sum _{j=1}^{n_ {\psi e} } A^ {\psi e} _j(A^ {\psi e} _j)^* + R, \end{aligned}$$

where \(A^\psi _j \in {{\,\mathrm{Op}\,}}{{\,\mathrm{SG}\,}}_\mathrm {cl}^{0,-\infty }\), \(A^e_j \in {{\,\mathrm{Op}\,}}{{\,\mathrm{SG}\,}}_\mathrm {cl}^{-\infty ,0}\), \(A^ {\psi e} _j \in {{\,\mathrm{Op}\,}}{{\,\mathrm{SG}\,}}_\mathrm {cl}^{0,0}\), and \(R \in {\mathcal {L}}(\mathcal {S}',\mathcal {S})\). Furthermore, we assume that \({{\,\mathrm{WF}\,}}_{{{\,\mathrm{SG}\,}}} '(A^\bullet _j) \subset \varGamma _j^\bullet \).

Inserting the partition of unity into the counting function yields

$$\begin{aligned} N(\lambda ) = \sum _{j=1}^{n_\psi }N^\psi _j(\lambda ) + \sum _{j=1}^{n_e} N^e_j(\lambda ) + \sum _{j=1}^{n_ {\psi e} } N^ {\psi e} _j(\lambda ) + {{\,\mathrm{Tr}\,}}(E_\lambda R), \end{aligned}$$

where as before \(N^\bullet _j(\lambda ) = {{\,\mathrm{Tr}\,}}E_\lambda A^\bullet _j (A^\bullet _j)^* = \sum _{\lambda _k < \lambda } \Vert (A^\bullet _j)^* \psi _k\Vert ^2\). Here, \(\psi _k\) are the eigenfunctions of P with eigenvalue \(\lambda _k\).

Since \(A^e_j\) and \(A^ {\psi e} _j\) have wavefront set near the boundary of X, we can choose local coordinates such that \(A^e_j\) and \(A^ {\psi e} _j\) become SG-operators on \(\mathbb {R}^d\) of order \((-\infty ,0)\) and (0, 0), respectively. As in the proof of Theorem 1, the parametrix of the wave equation then is locally a smooth family of \({{\,\mathrm{SG}\,}}\)-Fourier integral operators.

The asymptotics for \(N^\psi _j\) and \(N^e_j\) are standard and follow from similar arguments as in Hörmander [18], exchanging the roles of variable and covariable for \(N^e_j\). We can also adapt the proof of Proposition 15. By observing that for \(N^e_j(\lambda )\) and \(N^\psi _j(\lambda )\), the amplitude is supported only near one boundary face, we obtain

$$\begin{aligned} (N^e_j * \rho )(\lambda )&\sim \sum _{k = 0}^\infty C_{e,j,k} \lambda ^{d-k},\\ (N^\psi _j * \rho )(\lambda )&\sim \sum _{k = 0}^\infty C_{\psi ,j,k} \lambda ^{d-k}. \end{aligned}$$

Using the standard Tauberian theorem [33, Theorem B.2.1] yields

$$\begin{aligned} (N^e_j * \rho )(\lambda )&= (N^e_j * \rho )(\lambda ) + O(\lambda ^{d-1}),\\ (N^\psi _j * \rho )(\lambda )&= (N^\psi _j * \rho )(\lambda ) + O(\lambda ^{d-1}). \end{aligned}$$

The operator \(E_\lambda R\) is regularising; thus, its trace is uniformly bounded. We arrive at

$$\begin{aligned} N(\lambda ) - (N * \rho )(\lambda ) = \sum _{j=1}^{n_ {\psi e} } \left[ N^ {\psi e} _j(\lambda ) - (N^ {\psi e} _j*\rho )(\lambda )\right] + O(\lambda ^{d-1}). \end{aligned}$$

It remains to estimate the terms \(N^ {\psi e} _j(\lambda ) - (N^ {\psi e} _j*\rho )(\lambda )\). For this, let

$$\begin{aligned} \varPi _j = \inf _{(x,\xi ) \in \varGamma _j^ {\psi e} }\varPi (x,\xi ),\quad {\tilde{\varPi }}_j = \max \{ \varPi _j, {\epsilon }\}. \end{aligned}$$

For \(1/T < {\epsilon }\), we have by Proposition 15 that

$$\begin{aligned} (N^ {\psi e} _j * \rho )(\lambda ) \sim \sum _{k=0}^\infty \sum _{j=0,1} w_{jk} \lambda ^{d-k} (\log \lambda )^j. \end{aligned}$$

This implies that the derivative is given by

$$\begin{aligned} (\partial _\lambda N^ {\psi e} _j * \rho )(\lambda ) = d \cdot w_{1,0} \lambda ^{d-1}\log \lambda + O(\lambda ^{d-1}), \end{aligned}$$

where \(w_{1,0}\) is given by Proposition 18. Namely,

$$\begin{aligned} w_{1,0} = \frac{1}{d} \int _{\mathbb {S}^{d-1}}\int _{\mathbb {S}^{d-1}} [p_{ {\psi e} }(\theta ,\omega )]^{-d} \cdot |\sigma ^ {\psi e} (A^ {\psi e} _j)(\theta ,\omega )|^2 \mathrm{d}\theta \,\mathrm{d}\omega . \end{aligned}$$

Together with Proposition 22, this implies that

$$\begin{aligned} (N^ {\psi e} _j * \rho _T)(\lambda )&= (N^ {\psi e} _j * \rho )(\lambda ) + O(\lambda ^{-\infty })\\&= \sum _{k=0}^\infty \sum _{j=0,1} w_{jk} \lambda ^{d-k} (\log \lambda )^j + O(\lambda ^{-\infty }) \end{aligned}$$

for \(1/T < {{\tilde{\varPi }}}_j\).

Applying the Tauberian theorem to \(N^ {\psi e} _j * \rho _T\) yields

$$\begin{aligned}&\frac{|N^ {\psi e} _j(\lambda ) - (N^ {\psi e} _j * \rho )(\lambda )|}{\lambda ^{d-1}\log \lambda }\\&\quad \lesssim {{\tilde{\varPi }}}_j^{-1} \int _{\mathbb {S}^{d-1}} \int _{\mathbb {S}^{d-1}} |\sigma ^ {\psi e} (A_j^ {\psi e} )(\omega ,\theta )|^2 p_{1,1}(\omega ,\theta )^{-d} \mathrm{d}\theta \,\mathrm{d}\omega \end{aligned}$$

for \(\lambda \ge {{\tilde{\varPi }}}_j\). Taking the \(\limsup \) and summing over all j gives

$$\begin{aligned}&\limsup _{\lambda \rightarrow \infty }\frac{|N(\lambda ) - (N * \rho )(\lambda )|}{\lambda ^{d-1}\log \lambda }\\&\lesssim \sum _{j=1}^{n_ {\psi e} } {{\tilde{\varPi }}}_j^{-1} \int _{\mathbb {S}^{d-1}} \int _{\mathbb {S}^{d-1}} |\sigma ^ {\psi e} (A_j^ {\psi e} )(\omega ,\theta )|^2 p_{1,1}(\omega ,\theta )^{-d} \mathrm{d}\theta \,\mathrm{d}\omega . \end{aligned}$$

The right-hand side is an upper Riemann sum; therefore, we obtain the claim by shrinking the partition of unity. \(\square \)

6 An Example: The Model Operator \(P=\left\langle \cdot \right\rangle \left\langle D\right\rangle \)

In this section, we will consider the case of the operator \(P = \left\langle \cdot \right\rangle \left\langle D\right\rangle \) on \(\mathbb {R}^d\). First, we compute the full symbol of P near the corner:

$$\begin{aligned} \left\langle x\right\rangle \left\langle \xi \right\rangle&= |x|\cdot |\xi |\cdot \left( 1+\frac{1}{|x|^2}\right) ^\frac{1}{2} \left( 1+\frac{1}{|\xi |^2}\right) ^\frac{1}{2} \\&=|x|\cdot |\xi |\cdot \sum _{j,k=0}^\infty \begin{pmatrix}\frac{1}{2}\\ j \end{pmatrix}\begin{pmatrix}\frac{1}{2}\\ k\end{pmatrix} (-1)^{j+k}\frac{1}{|x|^{2j} \cdot |\xi |^{2k}} \\&=\sum _{j,k=0}^\infty \begin{pmatrix}\frac{1}{2}\\ j\end{pmatrix}\begin{pmatrix} \frac{1}{2}\\ k\end{pmatrix} (-1)^{j+k}|x|^{1-2j} \cdot |\xi |^{1-2k}. \end{aligned}$$

It follows that \(p_ {\psi e} (x,\xi ) =\sigma ^ {\psi e} (P)(x,\xi )= |x||\xi |\), \(p_\psi (x,\xi ) = |\xi | \left\langle x\right\rangle \), and \(p_e(x,\xi ) = |x|\left\langle \xi \right\rangle \).

We have to investigate the flow of the principal symbol \(p_ {\psi e} \) in the corner. The Hamiltonian vector field on \(\mathbb {R}^{2d}\) is given by

$$\begin{aligned} \mathsf {X}_{p_ {\psi e} } = {\partial }_\xi p_ {\psi e} {\partial }_x - {\partial }_x p_ {\psi e} {\partial }_\xi . \end{aligned}$$

First, we show that the angle between x and \(\xi \) is invariant under the flow. This follows from

$$\begin{aligned} {\partial }_t \left\langle x,\xi \right\rangle&= \left\langle {\partial }_t x, \xi \right\rangle + \left\langle x, {\partial }_t \xi \right\rangle \\&= \frac{|x|}{|\xi |} \left\langle \xi ,\xi \right\rangle - \frac{|\xi |}{|x|} \left\langle x,x\right\rangle \\&= |x| |\xi | - |x| |\xi | = 0. \end{aligned}$$

Hence, the quantity

$$\begin{aligned} c = c(x_0,\xi _0) = \frac{\left\langle x_0,\xi _0\right\rangle }{|x_0| |\xi _0|} \end{aligned}$$

is preserved by the flow. The Hamiltonian flow \(\varPhi ^ {\psi e} (t) : {\mathcal {W}}^ {\psi e} \rightarrow {\mathcal {W}}^ {\psi e} \) is given by the angular part.

Lemma 24

The differential equation for \(\omega = x/\left|x\right|\) and \(\theta = \xi /\left|\xi \right|\) describing the Hamiltonian flow \(\varPhi ^ {\psi e} (t) : {\mathcal {W}}^ {\psi e} \rightarrow {\mathcal {W}}^ {\psi e} \) is given by

$$\begin{aligned} \left\{ \begin{aligned} {\partial }_t \omega&= -c\omega + \theta \\ {\partial }_t \theta&= -\omega + c\theta .\\ \end{aligned}\right. \end{aligned}$$
(20)

Proof

We observe that

$$\begin{aligned} {\partial }_t \frac{x(t)}{\left|x(t)\right|}&= \frac{{\partial }_t x(t)}{\left|x(t)\right|} - \frac{x(t) {\partial }_t \left|x(t)\right|}{\left|x(t)\right|^2}. \end{aligned}$$

The calculation of \({\partial }_t |x|\) is straightforward:

$$\begin{aligned} {\partial }_t \left|x\right|&= \frac{\left\langle x,\xi \right\rangle }{\left|x\right| \left|\xi \right|} \cdot \left|x\right| = \frac{\left\langle x_0,\xi _0\right\rangle }{\left|x_0\right| \left|\xi _0\right|} \cdot \left|x\right|, \end{aligned}$$

This implies

$$\begin{aligned} {\partial }_t \frac{x(t)}{\left|x(t)\right|} = \frac{\xi (t)}{\left|\xi (t)\right|} - c \, \frac{x(t)}{\left|x(t)\right|}, \end{aligned}$$

as claimed. The second equation follows likewise. \(\square \)

Proposition 25

The return time function \(\varPi : {\mathcal {W}}^ {\psi e} \rightarrow \mathbb {R}\) is given by

$$\begin{aligned} \varPi (\omega ,\theta ) = {\left\{ \begin{array}{ll} \dfrac{2\pi }{\sqrt{1 - \left\langle \omega ,\theta \right\rangle ^2}}, &{} \quad \left\langle \omega ,\theta \right\rangle ^2 \not = 1\\ 0, &{} \quad \left\langle \omega ,\theta \right\rangle ^2 = 1. \end{array}\right. } \end{aligned}$$

Proof

The system of differential equations (20) decomposes into d decoupled systems of the form

$$\begin{aligned} {\partial }_t \upsilon (t) = A \upsilon (t), \end{aligned}$$

where

$$\begin{aligned} A = \begin{pmatrix} -c &{} \quad 1 \\ -1 &{} \quad c\end{pmatrix} \end{aligned}$$

We note that the eigenvalues of the matrix A are given by \(\lambda _\pm = \pm i\sqrt{1 - c^2}\). Thus, we have that the fundamental solution to the differential equation (20) for \((\omega , \theta )\) is given by

$$\begin{aligned} S \cdot \begin{pmatrix}\hbox {e}^{-it\sqrt{1-c^2}}{{\,\mathrm{I}\,}}_d &{} 0\\ 0 &{} \hbox {e}^{it\sqrt{1-c^2}}{{\,\mathrm{I}\,}}_d\end{pmatrix} \cdot S^{-1} \end{aligned}$$

for some unitary matrix \(S = S(c)\). The claim follows by choosing the minimal \(t > 0\) with \(t\sqrt{1-c^2} \in 2\pi \mathbb {Z}\) and noting that \(c = \left\langle \omega (0),\theta (0)\right\rangle =\left\langle \omega _0,\theta _0\right\rangle \) for \(\omega _0, \theta _0 \in \mathbb {S}^{d-1}\). \(\square \)

Remark 26

Proposition 25 shows that Theorem 2 cannot be applied to P. Nevertheless, we calculate \(\gamma _0\) in Proposition 28.

Proof of Theorem 6

By the Weyl law, Theorem 1, we have that

$$\begin{aligned} N(\lambda ) = \gamma _2 \lambda ^d \log \lambda + \gamma _1 \lambda ^d + O(\lambda ^{d-1} \log \lambda ). \end{aligned}$$

So it remains to calculate the corresponding Laurent coefficients of \(\zeta (s)\). By Proposition 40, we have

$$\begin{aligned} \gamma _2&= \frac{{{\,\mathrm{TR}\,}}(P^{-d})}{d} = \frac{{{\,\mathrm{Tr}\,}}_{\psi ,e}(P^{-d})}{d} = \frac{(2\pi )^{-d}}{d} \int _{\mathbb {S}^{d-1}}\int _{\mathbb {S}^{d-1}} p_ {\psi e} (\theta ,\omega )^{-d} \mathrm{d}\theta \, \mathrm{d}\omega \\&= \frac{(2\pi )^{-d}}{d} \int _{\mathbb {S}^{d-1}}\int _{\mathbb {S}^{d-1}} \mathrm{d}\theta \, \mathrm{d}\omega \\&= \frac{[{{\,\mathrm{vol}\,}}(\mathbb {S}^{d-1})]^2}{(2\pi )^{d}} \frac{1}{d}. \end{aligned}$$

Again by Proposition 40,

$$\begin{aligned} \gamma _1&= \widehat{{{\,\mathrm{TR}\,}}}_{x,\xi }(P^{-d}) - \frac{{{\,\mathrm{TR}\,}}(P^{-d})}{d^2}\\&= \frac{1}{d} \left( \frac{1}{d} \widehat{{{\,\mathrm{TR}\,}}}_\theta (P^{-d}) + \widehat{{{\,\mathrm{Tr}\,}}}_\psi (P^{-d}) + \widehat{{{\,\mathrm{Tr}\,}}}_e(P^{-d}) - \frac{1}{d} {{\,\mathrm{TR}\,}}(P^{-d})\right) \end{aligned}$$

First, we note that \(\widehat{{{\,\mathrm{TR}\,}}}_\psi (P^{-d}) = \widehat{{{\,\mathrm{TR}\,}}}_e(P^{-d})\) and the last term we already calculated for \(\gamma _2\). We recall that \(p_ {\psi e} = 1\) on \(\mathbb {S}^{d-1} \times \mathbb {S}^{d-1}\). Thus, we have for \(\widehat{{{\,\mathrm{TR}\,}}}_\theta (P^{-d})\) that

$$\begin{aligned} \widehat{{{\,\mathrm{TR}\,}}}_\theta (P^{-d})&= \frac{1}{(2\pi )^d} \int _{\mathbb {S}^{d-1}} \int _{\mathbb {S}^{d-1}} p_ {\psi e} (\theta ,\omega )^{-d} \log \left( p_ {\psi e} (\theta ,\omega )^{-d} \right) \mathrm{d}\theta \, \mathrm{d}\omega \\&= 0. \end{aligned}$$

This implies

$$\begin{aligned} \gamma _1 = \frac{2 \cdot \widehat{{{\,\mathrm{TR}\,}}}_e(P^{-d})}{d} - \frac{{{\,\mathrm{TR}\,}}(P^{-d})}{d^2}. \end{aligned}$$
(21)

Hence, we only have to calculate \(\widehat{{{\,\mathrm{TR}\,}}}_e(P^{-d})\):

$$\begin{aligned} \widehat{{{\,\mathrm{TR}\,}}}_e(P^{-d})&=\frac{1}{(2\pi )^d} \lim _{\tau \rightarrow +\infty }\Big \{ \int _{\mathbb {S}^{d-1}}\int _{|\xi |\le \tau }p_e(\theta ,\xi )^{-d} \mathrm{d}\theta \,\mathrm{d}\xi \\&\quad -(\log \tau )\int _{\mathbb {S}^{d-1}}\int _{\mathbb {S}^{d-1}} p_ {\psi e} (\theta ,\omega )^{-d} \mathrm{d}\theta \, \mathrm{d}\omega \Big \}\\&= \frac{{{\,\mathrm{vol}\,}}(\mathbb {S}^{d-1})^2}{(2\pi )^d} \lim _{\tau \rightarrow +\infty }\left[ {{\,\mathrm{vol}\,}}(\mathbb {S}^{d-1})^{-1}\int _{|x|\le \tau }\left\langle x\right\rangle ^{-d}\mathrm{d}x-\log \tau \right] . \end{aligned}$$

Using polar coordinates, we see that

$$\begin{aligned} {{\,\mathrm{vol}\,}}(\mathbb {S}^{d-1})^{-1}\int _{|x|\le \tau }\left\langle x\right\rangle ^{-d}\mathrm{d}x = \int _0^\tau (1 + r^2)^{-d/2} r^{d-1} \mathrm{d}r. \end{aligned}$$

Now, we perform a change of variables \(r=t^{-\frac{1}{2}}\Leftrightarrow t=r^{-2}>0\), so that

$$\begin{aligned} \int _0^\tau (1+r^2)^{-\frac{d}{2}}r^{d-1}\,\mathrm{d}r&= \frac{1}{2} \int _{\tau ^{-2}}^{+\infty } (t+1)^{-\frac{d}{2}}t^{-1}\,\mathrm{d}t\\&= \frac{1}{2} \int _{\tau ^{-2}}^{+\infty }\frac{\mathrm{d}t}{t(t+1)} - \frac{1}{2} \int _{\tau ^{-2}}^{+\infty } \left[ (1+t)^{-1}-(t+1)^{-\frac{d}{2}}\right] \frac{\mathrm{d}t}{t}. \end{aligned}$$

For \({\text {Re}}z > 0\), we have that (cf. [16, #8.36])

$$\begin{aligned} \varPsi (z)=\int _0^{+\infty } \left[ (1+t)^{-1}-(t+1)^{-z}\right] \frac{\mathrm{d}t}{t}-\gamma , \end{aligned}$$

where \(\varPsi (z)\) is the digamma function, defined by (3). By elementary computations, we obtain

$$\begin{aligned} \int _{\tau ^{-2}}^{+\infty }\frac{\mathrm{d}t}{t(t+1)}+\log \tau ^{-2}&= \lim _{\kappa \rightarrow +\infty }\left[ \log \frac{\kappa }{\kappa +1}-\log \tau ^{-2}+\log (1+\tau ^{-2}) \right] \\&\quad +\log \tau ^{-2} \\&=\log (1+\tau ^{-2})\longrightarrow 0 \text { for }\tau \rightarrow +\infty . \end{aligned}$$

Hence, we have that

$$\begin{aligned} \lim _{\tau \rightarrow +\infty }&\left[ {{\,\mathrm{vol}\,}}(\mathbb {S}^{d-1})^{-1}\int _{|x|\le \tau }\left\langle x\right\rangle ^{-d}\mathrm{d}x-\log \tau \right] \\&= \frac{1}{2} \lim _{\tau \rightarrow \infty }\left[ \int _{\tau ^{-2}}^{+\infty } \frac{\mathrm{d}t}{t(t+1)} + \log \tau ^{-2}\right] \\&\quad - \frac{1}{2} \lim _{\tau \rightarrow \infty } \int _{\tau ^{-2}}^{+\infty } \left[ (1+t)^{-1}-(t+1)^{-\frac{d}{2}}\right] \frac{\mathrm{d}t}{t}\\&= -\frac{1}{2} \left[ \varPsi (d/2) + \gamma \right] . \end{aligned}$$

Summing up, we have obtained

$$\begin{aligned} \gamma _1 = - \frac{[{{\,\mathrm{vol}\,}}(\mathbb {S}^{d-1})]^2}{(2\pi )^{d}} \frac{1}{d}\cdot \!\left[ \varPsi \!\left( \frac{d}{2}\right) +\gamma +\frac{1}{d}\right] . \end{aligned}$$
(22)

The proof is complete. \(\square \)

Remark 27

Using the properties of the function \(\varPsi \), we can make (22) more explicit. Indeed (see, for example, [16, #8.366, page 945]), we find:

$$\begin{aligned} \gamma _1= {\left\{ \begin{array}{ll} \displaystyle - \frac{[{{\,\mathrm{vol}\,}}(\mathbb {S}^{d-1})]^2}{(2\pi )^{d}} \frac{1}{d} \left( \frac{1}{d} + 2\sum _{k=1}^{\frac{d-1}{2}}\frac{1}{2k-1} - 2 \log 2\right) \!, &{} \text { if }d \text { is odd,} \\ \displaystyle -\frac{[{{\,\mathrm{vol}\,}}(\mathbb {S}^{d-1})]^2}{(2\pi )^{d}} \frac{1}{d} \left( \frac{1}{d}+\sum _{k=1}^{\frac{d}{2}-1}\frac{1}{k}\right) , &{} \text { if }d \text { is even.} \end{array}\right. } \end{aligned}$$

In particular, we have that

$$\begin{aligned} \gamma _1 = {\left\{ \begin{array}{ll} -\dfrac{2}{\pi }(1 - 2 \log 2), &{} d=1,\\ -\dfrac{1}{4}, &{} d=2. \end{array}\right. } \end{aligned}$$

Even though the coefficient \(\gamma _0\) does not appear in the Weyl law of P, for the sake of completeness we show that it is not hard to compute in this case.

Proposition 28

For the operator \(P = \left\langle \cdot \right\rangle \left\langle D\right\rangle \), we have that \(\gamma _0 = 0\).

Proof

To calculate \(\gamma _0\), we switch to Weyl quantization. Letting \(p^w(x,D) = P\), the symbol p has an asymptotic expansion

$$\begin{aligned} p = p_ {\psi e} + p_{0,0} \mod {{\,\mathrm{SG}\,}}^{-1,0} + {{\,\mathrm{SG}\,}}^{0,-1}, \end{aligned}$$

where \(p_{0,0}(x,\xi ) = \frac{\left\langle x,\xi \right\rangle }{|x||\xi |}\) is the subprincipal symbol of p. Hence, for the symbol of \(P^{d-1} = [a^w(x,D)]^{d-1}\) we find an asymptotic expansion

$$\begin{aligned} p^{d-1} = (p_ {\psi e} )^{d-1} + (d-1)(p_ {\psi e} )^{d-2} p_{0,0} \mod {{\,\mathrm{SG}\,}}^{d-3,d-2} + {{\,\mathrm{SG}\,}}^{d-2,d-3}. \end{aligned}$$

This implies for the symbol \(p^{-d+1}\) of the inverse of \([p^w(x,D)]^{d-1}\) that

$$\begin{aligned} p^{-d+1} = (p_ {\psi e} )^{-d+1} - (d-1)\frac{p_{0,0}}{(p_ {\psi e} )^d} \mod {{\,\mathrm{SG}\,}}^{-d-1,-d} + {{\,\mathrm{SG}\,}}^{-d,-d-1}. \end{aligned}$$

Therefore, by Propositions 17 and 19, we have that

$$\begin{aligned} \gamma _0&= \frac{1}{d-1} A_{2,1}\\&= -(2\pi )^{-d} \int _{\mathbb {S}^{d-1}}\int _{\mathbb {S}^{d-1}} \frac{p_{0,0}}{(p_ {\psi e} )^d} \mathrm{d}\theta \,\mathrm{d}\omega \\&= -(2\pi )^{-d} \int _{\mathbb {S}^{d-1}}\int _{\mathbb {S}^{d-1}} \left\langle \theta ,\omega \right\rangle \mathrm{d}\theta \,\mathrm{d}\omega = 0. \end{aligned}$$

\(\square \)