1 Introduction

Let (Mg) be a smooth compact connected Riemannian manifold of dimension n, \(\Delta _g\) be the negative definite Laplace-Beltrami operator acting on \(L^2(M)\), and \(\{\lambda _j^2\}_{j=0}^\infty \) be the eigenvalues of \(-\Delta _g\), repeated with multiplicity, \(0=\lambda _0^2 < \lambda _1^2 \le \lambda _2^2 \le \dots \). In this article we obtain improved asymptotics for both pointwise and integrated Weyl Laws. That is, we study asymptotics for the Schwartz kernel of the spectral projector

$$\begin{aligned} \Pi _\lambda :L^2(M,g)\rightarrow \bigoplus _{\lambda _j\le \lambda }\ker (-\Delta _g-\lambda _j^2), \end{aligned}$$

i.e. \(\Pi _\lambda \) is the orthogonal projection operator onto functions with frequency at most \(\lambda \). If \(\{\phi _{\lambda _j}\}_{j=1}^\infty \) is an orthonormal basis of eigenfunctions, \(-\Delta _g \phi _{\lambda _j}=\lambda _j^2 \phi _{\lambda _j}\), the Schwartz kernel of \(\Pi _\lambda \) is

$$\begin{aligned} \Pi _{\lambda }(x,y)=\sum _{\lambda _j\le \lambda } \phi _{\lambda _j}(x)\overline{\phi _{\lambda _j}}(y),\qquad (x,y)\in M\times M. \end{aligned}$$

Asymptotics for the spectral projector play a crucial role in the study of eigenvalues and eigenfunctions for the Laplacian, with applications to the study of physical phenomena such as wave propagation and quantum evolution. One of the oldest problems in spectral theory is to understand how eigenvalues distribute on the real line. Let \(N(\lambda ):=\#\{{j:\;}\lambda _j\le \lambda \}\) be the eigenvalue counting function. Motivated by black body radiation, Hilbert conjectured that, as \(\lambda \rightarrow \infty \),

$$\begin{aligned} N(\lambda )={(2\pi )^{-n}}{{\,\textrm{vol}\,}}_{{\mathbb {R}}^n}(B){{\,\textrm{vol}\,}}_g(M)\lambda ^n+E(\lambda ),\qquad \qquad {E(\lambda )=o(\lambda ^{n}).} \end{aligned}$$

Here, \({{\,\textrm{vol}\,}}_{{\mathbb {R}}^n}(B)\) is the volume of the unit ball \(B \subset {\mathbb {R}}^n\), \({{\,\textrm{vol}\,}}_g(M)\) is the Riemannian volume of M, and \({{\,\mathrm{{\text {dv}}}\,}}_g\) is the volume measure induced by the Riemannian metric. The conjecture was proved by Weyl  [46] and is known as the Weyl Law. We refer to \(E(\lambda )\) as a Weyl remainder. In 1968, Hörmander [25], provided a framework for the study of \(E(\lambda )\) and generalized the works of Avakumović [1] and Levitan [35], who proved \(E(\lambda )=O(\lambda ^{n-1})\); a result that is sharp on the round sphere and is thought of as the standard remainder.

The article [25] provided a framework for the study of Weyl remainders which led to many advances, including the work of Duistermaat–Guillemin [17] who showed \(E(\lambda )=o(\lambda ^{n-1})\) when the set of periodic geodesics has measure 0. Recently, [27] verified this dynamical condition on all product manifolds. A striking application of our main theorem on Weyl remainders is:

Theorem 1

Let \((M_i,g_i)\) be smooth compact connected Riemannian manifolds of dimension \(n_i\ge 1\) for \(i=1,2\). Then, with \(M=M_1\times M_2\), \(g=g_1\oplus g_2\), and \(n:=n_1+n_2\),

$$\begin{aligned} N(\lambda )={(2\pi )^{-n}}{{\,\textrm{vol}\,}}_{{\mathbb {R}}^n}(B){{\,\textrm{vol}\,}}_g(M)\lambda ^{n}+O\big (\lambda ^{n-1}/\,{\log \lambda }\big ),\qquad {\lambda \rightarrow \infty .} \end{aligned}$$

For future reference, we note that \( {N(\lambda )=\int _{M}\Pi _{\lambda }(x,x){{\,\mathrm{{\text {dv}}}\,}}_g(x)} \) and thus \(N(\lambda )\) can be studied by understanding the kernel of \(\Pi _{\lambda }\) restricted to the diagonal. We study both on and off diagonal Weyl remainders in this article. The main idea is to adapt the geodesic beam techniques developed by authors [9, 11, 22] to study Weyl remainders. These techniques were originally used to study averages of quasimodes over submanifolds by decomposing the quasimodes into geodesic beams and controlling the averages in terms of the \(L^2\) norms of these beams. In this work the key point is to study the eigenvalue counting function by viewing it as a sum of quasimodes averaged over the diagonal in \(M\times M\). We start our exposition in the setting of the on diagonal estimates.

1.1 On diagonal Weyl remainders

The connection between the spectrum of the Laplacian and the properties of periodic geodesics on M has been known since at least the works [15, 16, 45], with their relation to Weyl remainders first explored in the seminal work [17]. To control \(E(\lambda )\) we impose dynamical conditions on the periodicity properties of the geodesic flow \(\varphi _t:T^*M \setminus \{0\} \rightarrow T^*M{\setminus \{0\}}\), i.e., the Hamiltonian flow of \((x,\xi )\mapsto |\xi |_{g(x)}\). For \(t_0>0\), \(T>0\), and \(R>0\), define the set of near periodic directions in \(U \subset S^*M\) by

$$\begin{aligned} \mathcal {P}^R_{_{U}}(t_0,T):=\bigg \{ \rho \in U\, :\, \bigcup _{t_0\le |t|\le T}\varphi _t(B_{_{S^*\!M}}(\rho ,R))\cap B_{_{S^*\!M}}\!(\rho ,R)\ne \emptyset \bigg \}. \end{aligned}$$
(1.1)

Given two sets \(U\subset V\subset T^*M\), and \(R>0\), we write \(B_{_{V}}(U,R):=\{\rho \in V:\; d(U, \rho )<R\}\), where d is the distance induced by some fixed metric on \(T^*M\), \(B(U,R)=B_{_{T^*\!M}}(U,R)\), and \(B_{_{V}}(\rho ,R)=B_{_{V}}(\{\rho \},R)\). The set \(\mathcal {P}^R_{_{U}}(t_0,T)\) represents those points which come R close to being periodic with period between \(t_0\) and T and will be used to give a quantitative measure of how many near periodic geodesics there are.

We phrase our dynamical conditions in terms of a resolution function  \(\textbf{T}=\textbf{T}(R)\). This is a function of the scale, R, at which the manifold is resolved, which increases as \(R\rightarrow 0^+\). We use \(\textbf{T}\) to measure the time for which balls of radius R can be propagated under the geodesic flow while satisfying a given dynamical assumption, e.g. being non periodic.

Definition 1.1

We say a decreasing, continuous function \({\textbf{T}}:(0,\infty )\rightarrow (0,\infty )\) is a resolution function. In addition, we say a resolution function \(\textbf{T}\) is sub-logarithmic, if it is differentiable and

$$\begin{aligned} {(\log \log R^{-1})'}=-1\big /R\log R^{-1}\le [\log \textbf{T}(R)]'\le 0,\qquad 0<R<1.\, \end{aligned}$$

We measure how close \(\textbf{T}\) is to being logarithmic through

$$\begin{aligned} \Omega (\textbf{T}):=\limsup _{R\rightarrow 0^+}\textbf{T}(R)\big /\log R^{-1}. \end{aligned}$$
(1.2)

Simple examples of sub-logarithmic resolution functions are \(\textbf{T}(R)=\alpha (\log R^{-1})^\beta \) for any \(\alpha >0\) and \(0<\beta \le 1\). For an explanation for our use of resolution functions, see Remark 1.6.

For improved integrated Weyl remainders, we need a condition on the geodesic flow. We will use the notation that for \({U}\subset T^*\!M\) we write \(\mu _{_{\!{U}}}\) for the Liouville measure induced on U.

Definition 1.2

Let \(\textbf{T}\) be a resolution function. Then \(U\subset S^*M\) is said to be \(\textbf{T}\) non-periodic with constant \(C_{_{\!{\text {np}}}}\) provided there exists \(t_0>0\) such that

(1.3)

We say U is \(\textbf{T}\) non-periodic if there is such \(C_{_{\!{\text {np}}}}\), and \({W}\subset M\) is \(\textbf{T}\) non-periodic if \(S^*_{{W}}M\) is.

Below, for \(U\subset T^*\!M\), we write \(\dim _{box }{U}\) for the Minkowski box dimension of U (see e.g. [42, Page 333]). Note that if \({W}\subset M\) is open with smooth boundary then \(\dim _{box }\partial {W}=n-1\).

Theorem 2

Let (Mg) be a smooth compact connected Riemannian manifold of dimension n, \({W}\subset M\) be an open subset with \(\dim _{box }\partial {W}<n\), and \({\Omega _0}>0\). There exists \(C_{_{0}}>0\) such that if \(\textbf{T}\) is a sub-logarithmic rate function with \(\Omega (\textbf{T})<{\Omega _0}\) and \({W}\) is \(\textbf{T}\) non-periodic, then there is \(\lambda _0\) such that for all \(\lambda >\lambda _0\)

$$\begin{aligned} \Big |\int _{{W}} \Pi _\lambda (x,x) {{\,\mathrm{{\text {dv}}}\,}}_g(x)-{(2\pi )^{-n}}{{\,\textrm{vol}\,}}_{{\mathbb {R}}^n}(B){{\,\textrm{vol}\,}}_g({W})\lambda ^n\Big | \le C_{_{0}}\,\lambda ^{n-1}\big /\,\textbf{T}\big (\lambda ^{-1}\big ). \end{aligned}$$

In particular, if M is \(\textbf{T}\) non-periodic, then there is \(\lambda _0\) such that for all \(\lambda >\lambda _0\)

$$\begin{aligned} \Big |N(\lambda )-{(2\pi )^{-n}}{{\,\textrm{vol}\,}}_{{\mathbb {R}}^n}(B){{\,\textrm{vol}\,}}_g(M)\lambda ^n\Big | \le C_{_{0}}\,\lambda ^{n-1}\big /\,\textbf{T}\big (\lambda ^{-1}\big ). \end{aligned}$$

We illustrate an application of Theorem 2 in Fig. 1. In this example we construct a surface of revolution with both a periodic and a non-periodic set (see Definition 1.2). In particular, Theorem 2 applies with \({W}\) contained in the non-periodic (green) set. One can obtain little oh improvements for the statement in Theorem 2, but this requires the more general version given in Theorem 6 instead (see Remark 1.8). See Table 1 in Sect. 1.3 for some additional examples.

Fig. 1
figure 1

An example of a perturbation of the sphere with both a non-periodic (green) and a periodic (orange) physical space set. The perturbed metric coincides with the round metric outside the strip (ab). Trajectories which remain in the spherical strip are \(2\pi \) periodic, while those which enter the non-periodic set are mostly non-periodic. See Sect. B.2.1 for a precise description of this example (color figure online)

The assumptions of Theorem 2 apply to a wide variety of Riemannian manifolds. Indeed, in addition to the concrete examples in Sect. 1.3, the authors [12] use Theorem 2 to give a logarithmic improvement in the remainder for the Weyl law that works for ‘typical’ metrics on any smooth manifold. This result is the first quantitative estimate for the remainder in Weyl laws that holds for most metrics.

We next discuss \(E_\lambda (x)\), the remainder in the on diagonal pointwise Weyl law

$$\begin{aligned} \Pi _{\lambda }(x,x)={(2\pi )^{-n}}{{\,\textrm{vol}\,}}_{{\mathbb {R}}^n} (B)\lambda ^n +E_\lambda (x), \qquad x \in M. \end{aligned}$$
(1.4)

The Weyl remainder in [25] comes from the estimate \(E_{\lambda }(x)=O(\lambda ^{n-1})\) for \(x\in M\) (again, sharp on the round sphere). The connection between \(E_\lambda (x)\) and geodesic loops through x is studied in the works of Safarov, Sogge–Zelditch [38, 41] and often appears in estimates for sup-norms of eigenfunctions. To control the pointwise remainder \( E_\lambda (x)\) we impose dynamical conditions on the looping properties of geodesics joining x with itself. For \(t_0>0\), \(T>0\), \(R>0\), and \(x,y\in M\), define

$$\begin{aligned} {\mathcal {L}}_{x,y}^{{R}}(t_0, T)\!:=\!\bigg \{ {{\rho } \in \! S_x^*M}:\; \!\bigcup _{t_0\le |t|\le T}\varphi _t(B(\rho ,R)) \,\cap \, {B({S_y^*M},R)} \!\ne \!\emptyset \bigg \}.\qquad \end{aligned}$$
(1.5)

Similar to \(\mathcal {P}^R_{_{U}}(t_0,T)\), the set \({\mathcal {L}}_{x,y}^{{R}}(t_0, T)\) represents those points, \(\rho \), that are R close to x and such that the geodesic through \(\rho \) comes R close to passing through to y in some time between \(t_0\) and T. The set will be used to give a quantitative measure of how many near looping geodesics there are.

Definition 1.3

Let \(\textbf{T}\) be a resolution function, \(t_0> 0\), \(C_{_{\!{\text {nl}}}}>0\), and \(x, y \in M\). Then, (xy) is said to be a \((t_0, \textbf{T})\) non-looping pair with constant \(C_{_{\!{\text {nl}}}}\) when

We say x is \((t_0,\textbf{T})\) non-looping with constant \(C_{_{\!{\text {nl}}}}\) if (xx) is a \((t_0,\textbf{T})\) non-looping pair with constant \(C_{_{\!{\text {nl}}}}\).

Note that if \(t_0<{{\,\textrm{inj}\,}}(M)\), where \({{\,\textrm{inj}\,}}(M)\) is the injectivity radius of M, then for x to be \((t_0, \textbf{T})\) non-looping is the same as being \((\varepsilon , \textbf{T})\) non-looping for any \(0<\varepsilon \le t_0\). In this case, we write x is \((0, \textbf{T})\) non-looping.

To state our estimates on the pointwise Weyl remainder, we let \(\lambda >0\), and, for points \(x,y \in M\) with \(d(x,y)<{{\,\textrm{inj}\,}}M\), define

$$\begin{aligned} E_\lambda ^0(x,y):=\Pi _\lambda (x,y)-\frac{1}{(2\pi )^n}\int _{|\xi |_{g_y}<\lambda }e^{i\langle \exp _{y}^{-1}(x),\xi \rangle }\frac{d\xi }{\sqrt{|g_y|}} . \end{aligned}$$
(1.6)

Here, the integral is over \(T_y^*M\), \(\exp _x:T_x^*M\rightarrow M\) is the the exponential map, and \(|g_y|\) denotes the determinant of the metric g at y, when g is thought of as matrix in local coordinates.

Theorem 3

Let \(\alpha ,\beta \in {\mathbb {N}}^n\), \(0< \delta <\frac{1}{2}\), \(C_{_{\!{\text {nl}}}}>0\), and \({\Omega _0}>0\). There exists \(C_{_{0}}>0\) such that the following holds. If \(\textbf{T}\) is a sub-logarithmic resolution function  with \(\Omega (\textbf{T})<{\Omega _0}\), there is \(\lambda _0>0\) such that if \(x_0\in M\) is \((0,\textbf{T})\) non-looping with constant \(C_{_{\!{\text {nl}}}}\), then for all \(\lambda >\lambda _0\)

$$\begin{aligned} \sup _{x, y \in B(x_0, \lambda ^{-\delta })}\big |\partial _x^\alpha \partial _y^\beta E_\lambda ^0(x,y)\big |\le C_{_{0}}\,\lambda ^{n-1+|\alpha |+|\beta |}\big /\,\textbf{T}\big (\lambda ^{-1}\big ). \end{aligned}$$

See Table 2 in Sect. 1.3 for some examples to which Theorem 3 applies.

Remark 1.4

At first it may not be obvious that (1.6) is the correct remainder to estimate for off-diagonal Weyl asymptotics. However, one can check that the term we subtract comes from the singularities corresponding to the shortest geodesic from x to y and, when there are few additional loops from x to y, one expects these to give the main contribution. See also the discussion after Theorem 4.

Theorems 2 and 3 fit in a long history of work on asymptotics of the kernel of the spectral projector and the eigenvalue counting function. Many authors considered pointwise Weyl sums [1, 21, 25, 35, 36, 39], eventually proving the sharp remainder estimates. The article [25] provided a method which was used in many later works: [17] showed \(E(\lambda )=o(\lambda ^{n-1})\) under the assumption that the set of periodic trajectories has measure 0, [38, 41] improved estimates on \(E_\lambda (x)\) to \(o(\lambda ^{n-1})\) under the assumption that the set of looping directions through x has measure 0 (see also the book of Safarov–Vassiliev [37]). See [13, 14] for corresponding estimates that are uniform in a small neighborhood of the diagonal and Ivrii [28] for the case of manifolds with boundaries.

While o(1) improvements were available under dynamical assumptions, until now, quantitative improvements in remainders were available in geometries where one has an effective parametrix to \(\log \lambda \) times e.g. manifolds without conjugate points [2, 4, 31] or non-Zoll convex analytic rotation surfaces [43, 44]. We point out that the closest results to ours are those of Volovoy [43]. There, quantitative estimates on \(E(\lambda )\) are obtained under stronger assumptions than those of Theorem 2. In particular, \({W}\) is required to be equal to M and the volume in (1.3) is required to be bounded by a positive power of R, rather than \(\textbf{T}(R)^{-1}\).

The estimates in this article are available without additional geometric assumptions. This comes from our use of the ’geodesic beam techniques’ developed in the authors’ work [9, 11, 22] and which in turn draw upon the semiclassical approach of Koch–Tataru–Zworski [33]. Theorems 2 and 3 can be thought of as the quantitative analogs of the main results in [17] and of [38], [41] respectively. In fact, these results can be recovered from Theorems 2 and 3 by allowing \(\textbf{T}(R)\) to grow arbitrarily slowly as \(R\rightarrow 0^+\) (see [11, Appendix B]). We also note that our estimates include both \(C^\infty \) asymptotics for \(\Pi _\lambda (x,y)\) and uniformity in certain shrinking neighborhoods of the diagonal without any additional effort and hence include the results from [13, 14].

Remark 1.5

To recover the results of [13, 14, 38, 41] one needs uniformity in o(1) neighborhoods of points of interest. As stated, Theorem 3 does not quite include this since it works in a \(\lambda ^{-\delta }\) neighborhood of x. However, the full version of our estimates, Theorem 9, allows for the neighborhood of x to shrink arbitrarily slowly and thus recovers these earlier results.

Remark 1.6

(Resolution functions) There are several reasons why we state our theorems in terms of a general resolution function. First, it is necessary to allow \(\textbf{T}(R)\) to grow arbitrarily slowly as \(R\rightarrow 0\) to recover the o(1) results of [17, 38, 41] (see Remark 1.8). Second, while it may appear from Tables 1 and 2, that \(\textbf{T}(R)\) is always either \(c\log R^{-1}\) or the trivial case of \({{\,\textrm{inj}\,}}(M)\), this is not always true. In fact, one can check that many integrable examples are non-looping or non-periodic for \(\textbf{T}(R)\gg \log R^{-1}\). At the moment, the authors are not aware of concrete examples with \(\textbf{T}(R)\ll \log R\). However, it is likely that for any sub-logarithmic resolution function \(\textbf{T}\), with \(\textbf{T}(R)\rightarrow \infty \) as \(R\rightarrow 0^+\), a modification of the construction from [6] yields a metric on the sphere for which there is a point x such that x is not \((t_0,\textbf{T})\) non-looping for any \(t_0>0\), but there is a resolution function \(\textbf{T}_1\) with \(\textbf{T}_1(R){\longrightarrow } \infty \) as \(R\rightarrow 0^+\) and \(t_0>0\) such that x is \((t_0,\textbf{T}_1)\) non-looping. Also, note that our non-periodic, non-looping, and non-recurrent conditions are all monotonic in \(\textbf{T}\) in the sense that if \(\textbf{T}_1(R)\le \textbf{T}_2(R)\), and one of these conditions hold with the resolution function \(\textbf{T}_2\), then it also holds with \(\textbf{T}_1\).

1.2 Off diagonal Weyl remainders

The off diagonal behavior of \(\Pi _\lambda (x,y)\) plays a crucial role in understanding monochromatic random waves (see e.g. [7]) as well as in estimates for \(L^p\) norms of Laplace eigenfunctions (see e.g. [40, Section 5.1]). This problem is more complicated than the on diagonal situation since understanding the far off diagonal (i.e., \(d(x,y)>{{\,\textrm{inj}\,}}(M)\)) regime typically involves parametrices for \(e^{it\sqrt{-\Delta _g}}\) for \(t>{{\,\textrm{inj}\,}}(M)\), which are difficult to control. Notably, our geodesic beam techniques allow us to overcome this difficulty when estimating errors.

To control \(\Pi _\lambda (x,y)\) off-diagonal, we introduce a dynamical condition on the non-recurrence properties of the geodesics joining a point x with itself. To our knowledge, this is the first time non-recurrence is used in understanding off-diagonal Weyl remainders. For \(x \in M\), \({U} \subset S_x^*M\), \(t_0>0\), \(T>0\), and \(R>0\), let

$$\begin{aligned} \mathcal {R}^{R}_{_{{U},{\pm }}}(t_0,T):= \bigcup _{{t_0\le \pm t\le T}}\varphi _t\big (B({U},R) \big ) \cap B_{_{{S^*_xM}}}\!({U},R). \end{aligned}$$

Definition 1.7

Let \({\mathfrak {t}}\) and \(\textbf{T}\) be resolution functions  and \(R_0>0\). We say \(x \in M\) is \({({\mathfrak {t}},}\textbf{T})\) non-recurrent at scale \(R_0\) if for all \(\rho \in S^*_xM\) there exists a choice of ± such that for all \(A\subset B_{_{S^*_xM}}(\rho ,R_0)\), \(\varepsilon >0\), \(r>0\) with \(\textbf{T}(r)>{{\mathfrak {t}}(\varepsilon )}\), and \(0<R<R_0\),

Heuristically, the way to think about Definition 1.7 is as follows. Recall that the standard definition of recurrence of a set \(A\subset S^*_xM\) is that that for all \(B\subset A\) and \(\mu _{_{S^*_xM}}\)-almost every \(\rho \in B\), the geodesic through \(\rho \) returns to B infinitely often. Definition 1.7 is a strengthening of the statement that no recurrent set exists. Indeed, the set \(\mathcal {R}^{R}_{_{{U},{\pm }}}(t_0,T)\) consists of those points in U which return R close to U in times between \(t_0\) and T. Thus, a set is non-recurrent according to Definition 1.7 if every subset A of \(S^*_xM\) has the property that the collection of points which are close to A and almost return to A in time \({\mathfrak {t}}(\varepsilon )\) has volume smaller than \(\varepsilon \) times that of the ball of radius R around A. Thus, in particular, most points eventually do not come close to A and hence A is also non-recurrent in the traditional sense.

If (xy) is a \((t_0, \textbf{T})\) non looping pair for some \(t_0>0\) we measure the difference between \(\Pi _\lambda (x,y)\) and its smoothed version which takes into account propagation up to time \(t_0\). Let \(\rho \in \mathcal {S}({\mathbb {R}})\) with \({\hat{\rho }}(0)\equiv 1\) on \([-1,1]\) and \({{\,\textrm{supp}\,}}{\hat{\rho }}\subset [-2,2]\). For \(\sigma >0\) we define

$$\begin{aligned} \rho _{\sigma }(s):= \sigma \, \rho \big (\sigma \, s \big ). \end{aligned}$$
(1.7)

For \(x,y \in M\), \(t_0>0\), and \(\lambda >0\), let

$$\begin{aligned} E_\lambda ^{t_0}:=\Pi _\lambda -\rho _{_{t_0}}*\Pi _{\lambda }, \end{aligned}$$
(1.8)

where the convolution is taken in the \(\lambda \) variable. The quantity \(E_{\lambda }^{t_0}\) is the appropriate one to estimate since, under non-looping type assumptions, one expects the main contribution to the kernel of the spectral projector to come from short (fixed) time wave propagation.

Below is our first off diagonal result.

Theorem 4

Let \(\alpha ,\beta \in {\mathbb {N}}^n\), \(0<\delta <\frac{1}{2}\), \(C_{_{\!{\text {nl}}}}>0\), \({R_0>0}\), \({\Omega _0}>0\), \({\varepsilon }>0\), and \({\mathfrak {t}}\) be a resolution function, there is \(C_{_{0}}>0\) such that if \(\textbf{T}_j\) is a sub-logarithmic resolution function  with \(\Omega (\textbf{T}_j)<{\Omega _0}\) for \(j=1,2\) and \(\textbf{T}_{\max }=\max (\textbf{T}_1,\textbf{T}_2)\), then there is \(\lambda _0>0\) such the following holds. If \(x_0,y_0 \in M\) and \(t_0>0\) are such that \(x_0\) and \(y_0\) are respectively \(({\mathfrak {t}},\textbf{T}_1)\) and \(({\mathfrak {t}},\textbf{T}_2)\) non-recurrent at scale \(R_0\), and \((x_0,y_0)\) is a \((t_0,{\textbf{T}_{\max }})\) non-looping pair with constant \(C_{_{\!{\text {nl}}}}\), then for \(\lambda >\lambda _0\)

$$\begin{aligned}{} & {} \sup _{x \in B(x_0, \lambda ^{-\delta })}\sup _{y \in B(y_0, \lambda ^{-\delta })}\big |\partial _x^\alpha \partial _y^\beta E^{t_0{+\varepsilon }}_\lambda (x,y)\big | \le C_{_{0}}\,\lambda ^{n-1+|\alpha |+|\beta |}\bigg /\!\sqrt{\textbf{T}_1 \big (\lambda ^{-1}\big )\textbf{T}_2(\lambda ^{-1})}. \end{aligned}$$

See Table 2 in Sect. 1.3 for some examples to which Theorem 4 applies.

To compare Theorems 3 and 4, note that for \(x,y\in M\) with \(d(x,y)<\varepsilon <{{\,\textrm{inj}\,}}(M)\),

$$\begin{aligned}{} & {} \bigg |\partial ^\alpha _x\partial ^\beta _y \bigg (\rho _{_{\varepsilon \lambda }}*{\Pi _\lambda }(x,y)-\frac{1}{(2\pi )^n}\int _{|\xi |_{g_y}<\lambda }e^{i\langle \exp _{y}^{-1}(x),\xi \rangle }q_{{\lambda }}(x,y,\xi )\frac{d\xi }{\sqrt{|g_y|}}\bigg )\bigg |\\{} & {} \quad \le C_{_{0}}\lambda ^{n-2+|\alpha |+|\beta |} \end{aligned}$$

where \(q_{{\lambda }}(x,y, \xi ){=} 1+\lambda ^{-1}q_{-1}(x,y,\xi )\) and \(q_{-1}(x,y,\xi )=O(d(x,y))\) (see e.g. [13, Proof of Proposition 10]). Then, for points xy with \(d(x,y)<\lambda ^{-\delta }\), modulo terms smaller than our remainder, \(E^0_\lambda (x,y)\) as defined in (1.6) is the same as \(E^\varepsilon _\lambda (x,y)\).

For any \(t_0<\infty \), it is possible to write an oscillatory integral expression for \(\rho _{t_0}*\Pi _{\lambda }(x,y)\). However, its precise behavior in \(\lambda \) depends heavily on the geometry of (Mg); in particular, on the structure of the set of geodesics from x to y. This explains why we state our estimates in terms of \(E_{\lambda }^{t_0}\).

More generally, our results apply to averages of \(\Pi _\lambda (x,y)\) with \(x \in H_1\) and \(y \in H_2\), where \(H_1,H_2\) are any two smooth submanifolds of M. This type of integral is known as a Kuznecov sum [47] and appears in the analytic theory of automorphic forms [5, 23, 24, 29, 34]. All our dynamical assumptions for points \(x,y \in M\) above may be defined for the submanifolds \(H_1,H_2 \subset M\) instead. In doing so, the only change needed is to use the sets of unit co-normal directions \(S\!N^*\!H_1\) and \(S\!N^*\!H_2\), instead of \(S_x^*M\) and \(S_y^*M\). See Definitions 1.12 and 1.13 for a detailed explanation. In what follows \(d\sigma _{_{\!H_1}}\) and \(d\sigma _{_{\!H_2}}\) denote the volume measures induced by the Riemannian metric on \(H_1\) and \(H_2\) respectively.

Theorem 5

Let \(\alpha ,\beta \in {\mathbb {N}}^n\), \({1}\le k_1\le n\), \({1}\le k_2\le n\), \(C_{_{\!{\text {nl}}}}>0\), \({\Omega _0}>0\), \({\varepsilon }>0\), \({R_0}>0\), and \({\mathfrak {t}}\) be a resolution function. There is \(C_{_{0}}{=C_{_{0}}(\alpha ,\beta ,k_1,k_2,n,C_{_{\!{\text {nl}}}},\Omega _0,\varepsilon ,R_0,{\mathfrak {t}})}>0\) such that if \(\textbf{T}_j\) is a sub-logarithmic resolution function  with \(\Omega (\textbf{T}_j)<{\Omega _0}\) for \(j=1,2\) and \(\textbf{T}_{\max }=\max (\textbf{T}_1,\textbf{T}_2)\) the following holds. If \(t_0>0\), and \(H_j\subset M\) are submanifolds of codimension \(k_j\) such that \((H_1,H_2)\) is a \((t_0,{\textbf{T}_{\max }})\) non-looping pair with constant \(C_{_{\!{\text {nl}}}}\), and \(H_j\) is \(({\mathfrak {t}},\textbf{T}_j)\) non-recurrent at scale \(R_0\) for \(j=1,2\), then there is \(\lambda _0>0\) such that for \(\lambda >\lambda _0\)

$$\begin{aligned}{} & {} \left| \int _{H_1}\int _{H_2}\partial _x^\alpha \partial _y^\beta E^{t_0{+\varepsilon }}_\lambda (x,y)\;d\sigma _{_{\!H_1}}(x)d\sigma _{_{\!H_2}}(y)\right| \\{} & {} \quad \le C_{_{0}}\,\lambda ^{\frac{k_1+k_2}{2}-1+|\alpha |+|\beta |}\bigg /\!\sqrt{\textbf{T}_1 \big (\lambda ^{-1}\big )\textbf{T}_2(\lambda ^{-1})}. \end{aligned}$$

See Table 2 in Sect. 1.3 for some examples to which Theorem 5 applies.

To our knowledge, Theorem 5 is the first theorem to give improved remainders for Kuznecov sum remainders under dynamical assumptions. Theorems 3, 4, and 5 are consequences of our results for general semiclassical pseudodifferential operators (see Theorems 8 and 9).

1.3 Applications

In this section we present some examples to which our theorems apply. For each of them we give a reference for the detailed proofs that the relevant assumptions are satisfied. Note that Appendix B contains many examples not listed in Tables 1 and 2, and that the results from [8] can be used to find additional examples. With the exception of the final three rows of Table 1 with \(W=M\), all the estimates in Tables 1 and 2 are new.

In Table 1, we list examples where the assumptions of Theorem 2 hold. The final two examples are due to Volovoy [44].

Table 1 This table lists examples with \(\textbf{T}\) non-periodic subsets with \(\textbf{T}(R)= c \log R^{-1}\)

In Table 2 we list some examples for which Theorems 4 and 5 hold. In each case there exists \(t_0>0\) such that \((H_1, H_2)\) is a \((t_0, {\max (\textbf{T}_1,\textbf{T}_2)})\) non-looping pair. Note that we omit labeling points for which \(\textbf{T}_2={{\,\textrm{inj}\,}}(M)\) since being \({{\,\textrm{inj}\,}}(M)\) non-recurrent is an empty statement. In these cases the gain in the pointwise Weyl law is \(\sqrt{\log \lambda }\) instead of \(\log \lambda \).

Table 2 The table lists examples where Theorems 4 and 5 hold

1.4 Further improvements

Many experts believe that, for a Baire generic Riemannian metric on a smooth compact manifold, there is \(\delta >0\) such that \(E(\lambda )=O(\lambda ^{n-1-\delta })\). Presently, this type of improved remainder is only available when the geodesic flow has special structure e.g. the flat torus, non-Zoll convex analytic surfaces of revolution, or compact Lie groups of rank \(>1\) with bi-invariant metric [44]. Specifically, the geodesic flow must expand only polynomially in time, \(\Vert d\varphi _t\Vert _{L^\infty (TS^*\!M)}\le C\langle t\rangle ^N\) for some \(N>0\). Typically, geodesics will instead expand exponentially in some places and, because of this, Egorov’s theorem generally only holds to logarithmic times. In fact, the only ingredient in our proof which restricts us to logarithmic improvements is Egorov’s theorem. Under the assumption of polynomial expansion one can prove an Egorov theorem to polynomial times and hence obtain polynomially improved remainders using our methods. We do not pursue this here since the present article is intended to apply on a general manifold and the polynomial times involved in such an Egorov theorem are not explicit. We instead plan to address the integrable case specifically in a future article.

1.5 Weyl laws for general operators

Let \(\Psi ^m(M)\) denote the class of semiclassical pseudodifferential operators of order \(m>0\) and \(P(h) \in \Psi ^m(M)\) be self-adjoint, with principal symbol p, that is positive and classically elliptic in the sense that there is \(C>0\) such that

$$\begin{aligned} p(x,\xi )\ge \tfrac{1}{C}|\xi |^m,\qquad |\xi |\ge C. \end{aligned}$$
(1.9)

Let \(\{E_j(h)\}_j\) be the eigenvalues of P repeated with multiplicity. For \(s\in {\mathbb {R}}\) we work with \(\Pi _h(s):=\mathbb {1}_{(-\infty ,s]}(P(h)),\) which is the orthogonal projection operator

$$\begin{aligned} \Pi _h(s): L^2(M,{g}) \rightarrow \bigoplus _{ E_j(h)\le s} \ker (P(h)-E_j(h)). \end{aligned}$$

For \(x,y \in M\) we write \(\Pi _h(s; x,y)\) for its kernel

$$\begin{aligned} \Pi _h(s;x,y):=\sum _{ E_j(h)\le s}\phi _{_{E_j(h)}}(x)\overline{\phi _{_{E_j(h)}}}(y), \end{aligned}$$
(1.10)

where \(\{\phi _{_{E_j(h)}}\}_j\) is an orthonormal basis for \(L^2(M)\) with \( P(h)\phi _{_{E_j(h)}}=E_j(h)\phi _{_{E_j(h)}}. \) Note that one integrates (1.10) against the Riemannian volume density \({{\,\mathrm{{\text {dv}}}\,}}_g(y)\).

Let \( \varphi _t:T^*M \rightarrow T^*M \) denote the Hamiltonian flow for p at time t. We recall the maximal expansion rate for the flow and the Ehrenfest time at frequency \(h^{-1}\) respectively:

$$\begin{aligned} \Lambda _{\max }{} & {} :=\limsup _{|t|\rightarrow \infty }\frac{1}{|t|}{\log } \sup _{{\{p\in [a-\varepsilon ,b+\varepsilon ]\}}}\Vert d\varphi _t(x,\xi )\Vert ,\nonumber \\ T_e(h){} & {} :=\frac{\log h^{-1}}{2\Lambda _{\max }}. \end{aligned}$$
(1.11)

Note that \(\Lambda _{\max }\in [0,\infty )\) and if \(\Lambda _{\max }=0\), we may replace it by an arbitrarily small constant.

Remark 1.8

(Little oh improvements) When the expansion rate \(\Lambda _{\max }=0\) (see (1.11)) and our dynamical assumptions hold for \(\textbf{T}(R)\gg \log R^{-1}\), our theorems can be used to obtain \(o(1/\log \lambda )\) improvements over standard remainders. In special situations where the geodesic flow has sub-exponential expansion, we expect similar results with improvements beyond \(o(1/\log \lambda )\).

Definition 1.9

Let \(a, b \in {\mathbb {R}}\) with \(a\le b\). Let \(t_0>0\) and \(\textbf{T}\) be a resolution function. A set \(U\subset T^*M\) is said to be \( \textbf{T}\) non-periodic for p in the window [ab] provided that for all \(E\in [a,b]\) Definition 1.2 holds with \(\varphi _t\) being the Hamiltonian flow for p, and with \(S^*M\) replaced by \(p^{-1}(E)\).

The following is our most general version of the Weyl Law. We write \(\pi _{_{M}}:T^*M \rightarrow M\) for the natural projection and \({{\textsf{H}}_p}\) for the Hamiltonian vector field for p.

Theorem 6

Let \(0< \delta <\frac{1}{2}\), \(\ell \in {\mathbb {R}}\), and \(\mathcal {V}\subset \Psi ^\ell (M)\) a bounded subset, \(U\subset T^*M\) open, \(t_0>0\), \(C_{_{U}}>0\), and \(a,b\in {\mathbb {R}}\) with \(a\le b\). Suppose \(d\pi _{_{M}}{{\textsf{H}}_p}\ne 0\) on \(p^{-1}([a,b])\cap {\overline{U}}\). Then, there is \(C_{_{0}}>0\) such that the following holds. Let \(K>0\), \(A\in \mathcal {V}\) with \({\hbox {WF}}_{\textrm{h}}(A)\subset U\), \(\Lambda >\Lambda _{\max }\), \(\textbf{T}\) be a sub-logarithmic resolution function with \(\Lambda \Omega (\textbf{T})<1-2\delta \), and suppose U is \(\textbf{T}\) non-periodic in the window [ab] with

$$\begin{aligned} \limsup _{R\rightarrow 0}\sup _{t\in [a,b]} \textbf{T}(R)\mu _{_{p^{-1}(t)}}(B(\partial U, R))\le C_{_{U}}. \end{aligned}$$
(1.12)

Then, there is \(h_0>0\) such that for all \(0<h<h_0\), and \(E\in [a,b+Kh]\)

$$\begin{aligned} \begin{aligned}&\bigg |\sum _{-\infty < E_j(h)\le E}\langle A\phi _{_{E_j(h)}},\phi _{_{E_j(h)}}\rangle -{\text {tr}}\big (A\,\rho _{_{t_0/h}}*\Pi _h(E) \big )\bigg | \le C_{_{0}}\,h^{1-n}\big /\,\textbf{T}(h). \end{aligned} \end{aligned}$$
(1.13)

Since the second term in (1.13) involves only short time propagation for the Schrödinger group \(e^{itP/h}\), its asymptotic expansion in powers of h can in principle be obtained. This calculation is routine, but long, so we do not include it here. For the details when \(P=-h^2\Delta _g\), we refer the reader to [17, Proposition 2.1]. In addition, if \(U\subset T^*M\) has smooth boundary which intersects \(p^{-1}(E)\) transversally for \(E\in [a,b]\), then (1.12) holds. Although the statement of Theorem 6 is cumbersome when U with rough boundary is allowed, it is natural to consider dynamical assumptions on this type of set. Indeed, many dynamical systems exhibit the so-called ‘chaotic sea’ with ‘integrable islands’ behavior where the dynamics are aperiodic in the sea; a set which typically has very rough boundary.

Next, we consider generalized Kuznecov [34] type sums of the form

$$\begin{aligned} \begin{aligned} \Pi _{_{H_1,H_2}}^{^{A_1,A_2}}(s)&:=\int _{H_1}\int _{H_2} A_1\Pi _h(s) A_2^*\, (x,y)\,d\sigma _{_{\!H_1}}(x)d\sigma _{_{\!H_2}}(y), \end{aligned} \end{aligned}$$

where \(A_1,A_2\in \Psi ^\infty (M)\) and \(H_1, H_2 \subset M\) are two submanifolds of M.

Let \(H\subset M\) be a smooth submanifold. For \(a,b \in {\mathbb {R}}\), \(a \le b\), define

$$\begin{aligned} \Sigma _{_{[a,b]}}^H:=p^{-1}([a,b])\cap N^*\!H. \end{aligned}$$
(1.14)

Definition 1.10

We say a submanifold \(H\subset M\) of codimension k is conormally transverse for p in the window [ab] if given \(f_1,\dots , f_{k}\in C_c^\infty (M;{{\mathbb {R}}})\) locally defining H, i.e. with \(H= \bigcap _{i=1}^k\{f_i=0\}\) and \(\{df_i\}\text { linearly independent on }H,\) we have

$$\begin{aligned} \Sigma _{_{[a,b]}}^H \subset \bigcup _{i=1}^k\{ {{\textsf{H}}_p}f_i \ne 0\}, \end{aligned}$$
(1.15)

Here, we interpret \(f_i\) as a function on the cotangent bundle by pulling it back through the canonical projection map.

Remark 1.11

If \(P(h)=-h^2\Delta _g\), then \(p(x,\xi )=|\xi |^2_{g(x)}\). Working with \(a=b=1\), we have \( \Sigma _{_{[a,b]}}^H= S\!N^*H\). In this setup every submanifold \(H \subset M\) is conormally transverse for p.

Definition 1.12

Let \(H_1, H_2 \subset M\) be two smooth submanifolds. Let \(a, b \in {\mathbb {R}}\) with \(a\le b\). Let \(t_0>0\), \(\textbf{T}\) a resolution function, and \(C_{_{\!{\text {nl}}}}>0\). We say \((H_1,H_2)\) is a \((t_0, \textbf{T})\) non-looping pair in the window [ab] with constant \(C_{_{\!{\text {nl}}}}\) provided that Definition 1.3 holds for all \(E\in [a,b]\) with \(\varphi _t\) being the Hamiltonian flow for p and with \({\mathcal {L}}_{x,y}^{{R}}\) changed to

$$\begin{aligned} {\mathcal {L}}_{_{H_1,H_2}}^{{R},{E}}(t_0, T):=\bigg \{ \rho \in \Sigma _{_{E}}^{H_1}:\; \bigcup _{t_0\le |t|\le T}\varphi _t({B(\rho ,R)}) \,\cap {B\big ( \Sigma _{_{E}}^{H_2},R\big )}\ne \emptyset \bigg \}, \end{aligned}$$

and with \({S_x^*M}\) and \({S_y^*M}\) replaced with \( \Sigma _{_{E}}^{H_1} \) and \( \Sigma _{_{E}}^{H_2}\) respectively. We say H is \((t_0,\textbf{T})\) non-looping if (HH) is a \((t_0,\textbf{T})\) non-looping pair.

Definition 1.13

Let \(H \subset M\) be a smooth submanifold. Let \(a, b \in {\mathbb {R}}\) with \(a\le b\). Let \(t_0>0\), \(R_0>0\), \(0<C_{_{\!{\text {nr}}}}<1\), and let \(\textbf{T}\) be a resolution function. H is said to be \(\textbf{T}\) non-recurrent in the window [ab] with constants \({(R_0,C_{_{\!{\text {nr}}}})}\) provided Definition 1.7 holds for any \(E\in [a,b]\) with \( S^*_xM\) replaced by \( \Sigma _{_{E}}^H\) and where \(\varphi _t\) is the Hamiltonian flow for p.

To state our main estimate for Kuznecov sums, let \(\rho \in \mathcal {S}({\mathbb {R}})\) with \({\hat{\rho }}(0)\equiv 1\) on \([-1,1]\) and \({{\,\textrm{supp}\,}}{\hat{\rho }}\subset [-2,2]\). For \(T>0\) we define

$$\begin{aligned} \rho _{_{h,T}}(t):= \tfrac{T}{h} \, \rho \Big (\tfrac{T}{h} \, t \Big ). \end{aligned}$$
(1.16)

We then introduce the remainder

$$\begin{aligned} E_{_{H_1,H_2}}^{^{A_1,A_2}}(T,h; {s})=\Pi _{_{H_1,H_2}}^{^{A_1,A_2}}({s})-\rho _{_{h,T}}*\Pi _{_{H_1,H_2}}^{^{A_1,A_2}}({s}). \end{aligned}$$
(1.17)

Theorem 7

Let \(P(h) \in \Psi ^m(M)\) be a self-adjoint semiclassical pseudodifferential operator with classically elliptic symbol p. Let \({\mathfrak {t}}\) be a resolution function and \({\varepsilon >0}\). For \(j=1,2,\) let \(H_j\subset M\) be submanifolds with co-dimension \(k_j\). Let \(a, b\in {\mathbb {R}}\) such that \(H_j\) is conormally transverse for p in the window [ab] for \(j=1,2\). Let \({R_0}>0,\) \(t_0>0\), and for \(j=1,2\), let \(\textbf{T}_j\) be sub-logarithmic resolution functions and \(\textbf{T}_{\max }=\max (\textbf{T}_1,\textbf{T}_2)\). Suppose \(H_j\) is \({({\mathfrak {t}},\textbf{T}_j)}\) non-recurrent in the window [ab] with constant \({R_0}\) for each \(j=1,2\), and \((H_1,H_2)\) is a \((t_0, {\textbf{T}_{\max }})\) non-looping pair in the window [ab] with constant \(C_{_{\!{\text {nl}}}}\). Then, for all \(A_1, A_2\in \Psi ^\infty (M)\), there exist \(h_0>0\) and \(C_{_{0}}>0\) such that for all \(0<h\le h_0\), \(K>0\), and \({s}\in [a-Kh,b+Kh]\)

$$\begin{aligned} \Big |E_{_{H_1,H_2}}^{^{A_1,A_2}}(t_0{+\varepsilon },h; {s})\Big |\le C_{_{0}}\, h^{1-\frac{k_1+k_2}{2}}\Big /\!\sqrt{\textbf{T}_ 1(h)\textbf{T}_2(h)}. \end{aligned}$$

Remark 1.14

We omit the precise dependence of the constant \(C_{_{0}}\) on various parameters in Theorem 7. Instead, we refer the reader to our main theorem on averages, Theorem 8, where we have introduced notation to handle uniformity in families of submanifolds \(H_1\) and \(H_2\).

1.6 Outline of the paper and ideas from the proof

In Sect. 2 we introduce the notion of good coverings by tubes and various assumptions on these coverings which allow us to adapt the results of [11] to our setup. We also state our main averages theorem in its full generality (Theorem 8). Section 3 studies how the dynamical assumptions in the introduction relate to the assumptions on coverings by tubes from Sect. 2. In Sect. 4 we adapt the crucial estimates coming from the geodesic beam techniques [11] so that they can be applied to the study of Weyl remainders. Next, in Sect. 5, we estimate the scale (in the energy) at which averages of the spectral projector behave like Lipschitz functions in the spectral parameter. With this in hand, we are able to approximate \(\Pi _h\) using \(\rho _{_{h,T(h)}}*\Pi _h\) with \(T(h)=\sqrt{\textbf{T}_1(h)\textbf{T}_2(h)}\). Finally, Sect. 6 shows that the \(\rho _{_{h,T(h)}}*\Pi _h\) approximation is close to \(\rho _{h,t_0}*\Pi _h\), finishing the proof of our main theorem on averages. Section 7 contains the proof of our theorems on the Weyl remainder. This section follows the same strategy as that for averages: an estimate for the Lipschitz scale of the trace of the spectral projector, followed by relating \(\rho _{_{h,T(h)}}*\Pi _h\) to \(\rho _{_{h,t_0}}*\Pi _h\). In Appendix A we present an index of notation and in Appendix B we give examples including those from Table 2 to which our theorems can be applied.

The main idea of this article is to view the kernel of the spectral projector \(\mathbb {1}_{[t-s,t]}(P)\) as a quasimode for P. This allows us to use the geodesic beam techniques from [11] to control the energy scale at which the projector behaves like a Lipschitz function and hence to estimate the error when the projector is smoothed at very small scales. This idea is used a second time when controlling \((\rho _{h,T(h)}-\rho _{h,t_0}) *\Pi _h\) to estimate the contribution from small volumes of the possibly looping tubes. A simple argument using Egorov’s theorem controls the remaining non-looping tubes. The crucial insight used to handle the Weyl law is to view the kernel of the spectral projector as a distribution on \(M\times M\), where it is a quasimode for \({\textbf{P}}:=P\otimes 1\), and to study the Weyl Law via integration of the kernel over the diagonal. By doing this, we are able to reduce the problem to bounding an average of a quasimode over a submanifold, a setting in which geodesic beam techniques apply.

Note that Theorems 2 and 6 are proved in Sects. 7.1.4 and 7.1.3 respectively. Theorem 1 is a corollary of Theorem 2; the necessary dynamical properties are proved in Appendix B.1.1. Theorems 345, and 7 follow from an application of Theorem 9 (See Sect. 2.4 for Theorems 34, and 5. Theorem 7 is a direct corollary of Theorem 9.). The fact that Theorem 9 follows from Theorem 8 is proved in Sect. 9 and Theorem 8 is proved in Sect. 6.2.

Acknowledgements. The authors would like to thank Dmitry Jakobson, Iosif Polterovich, John Toth, Dmitri Vassiliev and Steve Zelditch for helpful comments on the existing literature and Maciej Zworski for suggestions on how to improve the exposition and presentation, and Leonid Parnovski for comments on a previous draft. Thanks also to the anonymous referee who’s comments improved the exposition. The authors are grateful to the National Science Foundation for partial support under grants DMS-1900434 and DMS-1502661 (JG) and DMS-1900519 (YC). Y.C. is grateful to the Alfred P. Sloan Foundation.

2 Results with dynamical assumptions via coverings by tubes

We divide this section in four parts. In Sect. 2.1 we introduce the analogues of Definitions 1.12 and 1.13 via the use of coverings by bicharacteristic tubes. Microlocalization to these tubes will eventually be used to generate bicharacteristic beams. In Sect. 2.2 we introduce the uniformity assumptions that allow us to obtain uniform control of the constants in all our results. In Sect. 2.3 we state the most general version of our results, using the definitions via coverings by tubes, and the uniformity assumptions.

2.1 Dynamical assumptions via coverings by tubes

Let \(H\subset M\) be a smooth submanifold that is conormally transverse for p in the window [ab]. Let \({\mathcal {Z}}\subset T^*M\) with

$$\begin{aligned} \Sigma _{_{[a,b]}}^H \subset {\mathcal {Z}}\end{aligned}$$
(2.1)

be a hypersurface that is transverse to the flow, and \( \varphi _t \) continue to denote the Hamiltonian flow for p at time t. Given \(A \subset \Sigma _{_{[a,b]}}^{H}\), \(\tau >0\), and \(r>0\), we define

$$\begin{aligned} \Lambda _{_{\!A}}^\tau (r):=\bigcup _{|t|\le \tau +r}\varphi _t\big ({B_{_{{\mathcal {Z}}}}}(A,r)\big ). \end{aligned}$$
(2.2)

Let \(\tau _{_{{{\,\textrm{inj}\,}}_H}}>0\) be small enough so that the map

$$\begin{aligned} (-\tau _{_{{{\,\textrm{inj}\,}}_H}},\tau _{_{{{\,\textrm{inj}\,}}_H}})\times {\mathcal {Z}}\rightarrow T^*M, \qquad (t,q) \mapsto \varphi _t(q), \end{aligned}$$
(2.3)

is injective. Given \(r>0\), \(0<\tau <\tau _{_{{{\,\textrm{inj}\,}}_H}}\), and a collection of points \(\{\rho _j\}_{j\in \mathcal {J}(r)}\), we will work with the tubes

$$\begin{aligned} \mathcal {T}_j=\mathcal {T}_j(r):=\Lambda _{\rho _j}^\tau (r). \end{aligned}$$

A \((\tau , r)\)-cover for \(A\subset T^*M\) is a collection of tubes \(\{\mathcal {T}_j(r)\}_{j\in \mathcal {J}(r)}\) where \(\mathcal {J}(r)\subset {\mathbb {N}}\) for which

$$\begin{aligned} \Lambda _A^\tau (\tfrac{1}{2}r) \subset \bigcup _{j\in \mathcal {J}(r)}\mathcal {T}_j(r),\qquad \text {and}\qquad \, {\mathcal {T}}_j(r)\cap \Lambda ^\tau _A(\tfrac{1}{2}r)\ne \emptyset ,\quad \text {for all }{j\in \mathcal {J}(r)}. \end{aligned}$$

Let \({\mathfrak {D}}>0\). We say a \((\tau ,r)\)-cover is a \(({\mathfrak {D}},\tau ,r)\)-good cover, if there is a splitting \(\mathcal {J}(r)=\sqcup _{i=1}^{{\mathfrak {D}}}\mathcal {J}_i(r)\) such that for all \(1\le i\le {\mathfrak {D}}\) and \(k{\ne }\ell \in \mathcal {J}_i(r)\),

$$\begin{aligned} \mathcal {T}_k(3r)\cap \mathcal {T}_\ell (3r)=\emptyset . \end{aligned}$$
(2.4)

For \(E\in {\mathbb {R}}\) and \(r>0\), we adopt the notation

$$\begin{aligned} \mathcal {J}_{_{E}}(r):=\Big \{ j\in \mathcal {J}(r):\; \mathcal {T}_j(r)\, \cap {\mathcal {Z}}\,\cap B( \Sigma _{_{E}}^{H},r)\ne \emptyset \Big \}. \end{aligned}$$
(2.5)

We are now ready to introduce the definitions via coverings of our dynamical assumptions. First, for \(0<t_0<T_0\), we say \(A\subset T^*\!M\) is \([t_0,T_0]\) non-self looping if

$$\begin{aligned} \bigcup _{t=t_0}^{T_0}\varphi _t(A)\cap A=\emptyset \qquad \text { or }\qquad \bigcup _{t=-T_0}^{-t_0}\varphi _t(A)\cap A=\emptyset . \end{aligned}$$
(2.6)

Definition 2.1

(non looping pairs via coverings) Let \(t_0>0\), \({\tau _0}>0\), \({{\mathfrak {D}}}>0\), and \(\textbf{T}\) be a resolution function. Let \(H_1, H_2\) be two submanifolds and \(U_1\subset {N^*H_1}\), \(U_2\subset N^*H_2\). We say \((U_1,U_2)\) is a \((t_0, \textbf{T})\) non-looping pair in the window [ab] via \({\tau _0}\)-coverings with constant \({C_{_{\!{\text {nl}}}}}\) provided for all \(0<\tau <\tau _0\) there exists \(r_0>0\) such that for \(0<r<{r_0}\), any two \(({{\mathfrak {D}}},\tau , r)\)-good covers of \(U_1\cap \Sigma _{_{[a,b]}}^{H_1}\) and \(U_2\cap \Sigma _{_{[a,b]}}^{H_2}\), \(\{\mathcal {T}_j^1(r)\}_{j\in \mathcal {J}^1(r)}\) and \(\{\mathcal {T}_j^2(r)\}_{j\in \mathcal {J}^2(r)}\) respectively, and every \(E\in [a,b]\), there is splittings of indices

$$\begin{aligned} \begin{aligned} \mathcal {J}^1_{_{E}}(r)={\mathcal {B}_{_{E}}^1(r) \cup \mathcal {G}_{_{E}}^1(r)},\qquad \mathcal {J}_{_{E}}^2(r)=\mathcal {B}_{_{E}}^2(r) \cup \mathcal {G}_{_{E}}^2(r), \end{aligned} \end{aligned}$$

satisfying

  1. (1)

    for each \({i,k\in \{1,2\}}\), \(i\ne k\) every \(\ell \in \mathcal {G}_{_{E}}^i(r)\),

    $$\begin{aligned} \Bigg (\bigcup _{{t_0+\tau \le |t|\le \textbf{T}(r)-\tau }} \varphi _t \Big (\mathcal {T}_\ell ^i(r)\Big ) \Bigg ) \bigcap \Bigg ( \bigcup _{j \in \mathcal {J}^k_{_{E}}(r)} \mathcal {T}_j^k(r) \Bigg )=\emptyset , \end{aligned}$$
  2. (2)

    \(r^{2({n-1})}|\mathcal {B}_{_{E}}^1(r)||\mathcal {B}_{_{E}}^2(r)|\textbf{T}(r)^2\le {{\mathfrak {D}}^2}C_{_{\!{\text {nl}}}}.\)

We will say \((H_1,H_2)\) is a \((t_0,\textbf{T})\) non-looping pair in the window [ab] via \(\tau \)-coverings if \((N^*H_1,N^*H_2)\) is. We will also say H is \((t_0, \textbf{T})\) non-looping in the window [ab] via \(\tau \) coverings whenever (HH) is a non-looping pair.

In Definition 2.1, the sets \(\mathcal {B}_{_{E}}\) and \(\mathcal {G}_{_{E}}\) should be thought of as respectively ‘bad’ and ‘good’ tubes. The tubes \(\mathcal {B}_{_{E}}\) are ‘bad’ in the sense that they may connect \( \Sigma _{_{[a,b]}}^{H_1}\) and \( \Sigma _{_{[a,b]}}^{H_2}\) under the Hamiltonian flow for p in a relatively short time, while the tubes \(\mathcal {G}_{_{E}}\) are ‘good’ in the sense that they do not connect these two sets for some controlled amount of time (see part (1) of the definition). Part (2) of the definition guarantees that there are not too many bad tubes connecting \( \Sigma _{_{[a,b]}}^{H_1}\) and \( \Sigma _{_{[a,b]}}^{H_2}\).

In Sect. 3, we prove that non looping in the sense of Definition 1.12 is equivalent to non looping by coverings in the sense of Definition 2.1.

Definition 2.2

(non-recurrence via coverings) Let \({\tau _0}>0\), \({{\mathfrak {D}}}>0\), and \(\textbf{T}\) be a resolution function. We say H is \(\textbf{T}\) non-recurrent in the window [ab] via \({\tau _0}\)-coverings with constant \(C_{_{\!{\text {nr}}}}\) provided for all \(0<\tau <\tau _0\) there exists \(r_0>0\) such that for \(0<r<{r_0}\), every \(({{\mathfrak {D}}},\tau , r)\)-good cover of \( \Sigma _{_{[a,b]}}^{H}\), \(\{\mathcal {T}_j(r)\}_{j\in \mathcal {J}(r)}\), and \(E\in [a,b]\), there exists a finite collection of sets of indices \(\{\mathcal {G}_{_{E,\ell }}(r)\}_{\ell \in {\mathcal {L}_{_{E}}(r)}}\) with \( \mathcal {J}_{_{E}}(r)=\bigcup _{\ell \in {\mathcal {L}_{_{E}}(r)}}\mathcal {G}_{_{E,\ell }}(r), \) and so that for every \(\ell \in {\mathcal {L}_{_{E}}(r)}\) there exist functions \(t_\ell (r)>0\) and \({T_\ell (r)}>0\), with \(0\le t_\ell (r)\le T_\ell (r)\le \textbf{T}(r)\), so that

  1. (1)

    \(\bigcup _{j\in \mathcal {G}_{_{E,\ell }}(r)}\mathcal {T}_j(r)\;\;\text { is }\;\;[t_\ell (r),T_{\ell }(r)]\text { non-self looping},\)

  2. (2)

    \(r^{\frac{n-1}{2}}\sum _{\ell \in {{\mathcal {L}}_{_{E}}(r)}} {\big (|\mathcal {G}_{_{E,\ell }} (r)|t_\ell (r){T_\ell (r)}^{-1}\big )^{\frac{1}{2}}} \le {{\mathfrak {D}}^{\frac{1}{2}}}C_{_{\!{\text {nr}}}}\,\textbf{T}(r)^{-\frac{1}{2}}.\)

In Definition 2.2, the sets \(\mathcal {B}_{_{E}}\) and \(\mathcal {G}_{_{E}}\) should again be thought of as respectively ‘bad’ and ‘good’ tubes. The tubes \(\mathcal {B}_{_{E}}\) are ‘bad’ in the sense that they may self intersect under the Hamiltonian flow for p in a relatively short time, while the tubes \(\mathcal {G}_{_{E}}\) are ‘good’ in the sense that they do not self intersect these two sets for some controlled amount of time (see part (1) of the definition). Part (2) of the definition again guarantees that there are not too many bad tubes.

In Lemma 3.5 below we prove that non recurrence in the sense of Definition 1.13 implies non recurrence by coverings in the sense of Definition 2.2. At the moment, we are unable to determine whether these two definitions are equivalent.

2.2 Uniformity assumptions

Let \(H\subset M\) be a smooth submanifold. In practice, we prove estimates on \(\{{\tilde{H}}_h\}_h\), where \(\{{\tilde{H}}_h\}_h\) is a family of submanifolds such that

$$\begin{aligned} {\sup \Big \{ d\big (\rho , \Sigma _{_{[a,b]}}^{{{\tilde{H}}}_h}}\big )\,\,\big |\,\,{\rho \in \Sigma _{_{[a,b]}}^H}\Big \}\le {R(h)} \qquad \qquad h>0, \end{aligned}$$
(2.7)

where \(R(h)>0\) and for every multi-index \(\alpha \) there is \({\mathcal {K}}_{_{\alpha }}>0\) such that for all \(h>0\)

$$\begin{aligned} |\partial _x^\alpha \textbf{R}_{_{{{\tilde{H}}}_h}}|+|\partial _x^\alpha \mathbf{\Pi }_{_{{{\tilde{H}}}_h}}|\le {\mathcal {K}}_{_{\alpha }}. \end{aligned}$$
(2.8)

Here \(\textbf{R}_{_{{{\tilde{H}}}_h}}\) and \(\mathbf{\Pi }_{_{{{\tilde{H}}}_h}}\) denote the sectional curvature and the second fundamental form of \({{\tilde{H}}}_h\). Without loss of generality, we will assume \({\mathcal {Z}}\) is chosen so that there exist \(N>0\), \(C=C(p,a,b,\{{\mathcal {K}}_\alpha \}_{|\alpha |\le N})>0,\) and \(r_0>0\) such that for all \(E\in [a,b]\), \(A\subset \Sigma _{_{E}}^H\) and \(0<r<r_0\),

$$\begin{aligned} {{\,\textrm{vol}\,}}\Big (B_{_{{\mathcal {Z}}}}(A,r)\Big )\le Cr^{n}\mu _{ \Sigma _{_{E}}^{H}}\Big (B_{_{ \Sigma _{_{E}}^{H}}}\big (A,r\big )\Big ). \end{aligned}$$

We may do this since \(\dim {\mathcal {Z}}=2n-1\), \(\dim \Sigma _{_{E}}^{H}=n-1\), and \( \Sigma _{_{E}}^{H} \subset {\mathcal {Z}}\).

Note that when \(H=\{x_0\}\) is a point, the curvature bounds become trivial, and so in place of  (2.7) we work with \(d(x_0,{\tilde{x}}_h)<{R(h)}\) and may take \({{\mathcal {K}}_{_{\alpha }}}\) to be arbitrarily close to 0. In what follows, let \(r_{_{H}}:T^*M\rightarrow {\mathbb {R}}\) be the geodesic distance to H, i.e., \(r_{_{H}}(x,\xi )=d(x,H)\) for \((x,\xi )\in T^*M\), and write \(\pi _{_{M}}:T^*M \rightarrow M\) for the natural projection.

Definition 2.3

(regular families) We will say a family of submanifolds \(\{H_h\}_{h}\) is regular in the window [a, b] if it satisfies  (2.8) and there is \(\varepsilon >0\) so that for all \(h>0\), the map \((-\varepsilon ,\varepsilon )\times { \Sigma _{_{[a,b]}}^H}\rightarrow M\),

$$\begin{aligned} (t,\rho )\mapsto \pi _{_{M}}({\varphi _t(\rho )})\;\;\;\text { is a diffeomorphism}. \end{aligned}$$
(2.9)

Then, define \(|{{\textsf{H}}_p}r_{_{\!H}}|: \Sigma _{_{[a,b]}}^H\rightarrow {\mathbb {R}}\) by

$$\begin{aligned} |{{\textsf{H}}_p}r_{_{\!H}}|(\rho ):=\lim _{t\rightarrow 0}|{{\textsf{H}}_p}r_{_{\!H}}(\varphi _t(\rho ))|. \end{aligned}$$
(2.10)

Definition 2.4

(uniformly conormally transverse submanifolds) A family of submanifolds \(\{{{\tilde{H}}}_h\}_{h}\) is said to be uniformly conormally transverse for p in the window [a, b] provided

  1. (1)

    \({{\tilde{H}}}_h\) is conormally transverse for p in the window [ab] for all \(h>0\),

  2. (2)

    there exists \({\mathfrak {I}}_{_{\!0}}>0\) so that for all \(h>0\)

    $$\begin{aligned} \inf \Big \{|{{\textsf{H}}_p}r_{_{\!{{\tilde{H}}}_{\!h}}}|(\rho )\,\,\big |\,\,{\rho \in \Sigma _{_{[a,b]}}^H}\Big \} \ge {\mathfrak {I}}_{_{\!0}}. \end{aligned}$$
    (2.11)

When the constants involved in our estimates depend on \(\{{{\tilde{H}}}_h\}_h\), they will do so only through finitely many of the \({\mathcal {K}}_{_{\alpha }}\) constants and the constant \({\mathfrak {I}}_{_{\!0}}\).

Remark 2.5

We note that for \(p(x, \xi )=|\xi |^2_{g(x)}\), \( {a=b=1,}\) and \( \Sigma _{_{[a,b]}}^H=S\!N^*\!H\), we have \(|{{\textsf{H}}_p}r_{_{\!H}}|(\rho )=2\) for all \(\rho \in S\!N^*\!H\). It follows that every family of submanifolds is uniformly conormally transverse and we may take \({\mathfrak {I}}_{_{\!0}}=2\).

2.3 Main results

We now state the main results from which all of our Kuznecov type asymptotics follow. Throughout the text, the notation \(C=C(a_1,\dots , a_k)\) means that the constant C depends only on \(a_1, \dots , a_k\).

Theorem 8

For \(j=1,2\), let \(k_j\in \{1,\dots ,n\}\), \({\mathfrak {I}}_{_{\!0}}^j>0\), \(A_j\in \Psi ^\infty (M)\). Let \(C_{_{\!{\text {nr}}}}^1>0\), \(C_{_{\!{\text {nr}}}}^2>0\) and \(C_{_{\!{\text {nl}}}}>0\). There is

$$\begin{aligned} C_{_{0}}={C_{_{0}}(n, k_1, k_2, A_1, A_2, {\mathfrak {I}}_{_{\!0}}^1, {\mathfrak {I}}_{_{\!0}}^2,{C_{_{\!{\text {nr}}}}^1, C_{_{\!{\text {nr}}}}^2}, C_{_{\!{\text {nl}}}})}>0 \end{aligned}$$

such that the following holds.

Let \(P(h) \in \Psi ^m(M)\) be a self-adjoint semiclassical pseudodifferential operator, with classically elliptic symbol p. Let \(0<\delta <\frac{1}{2}\), \(K>0\), \(a,b \in {\mathbb {R}}\) with \(a\le b\), and for \(j=1,2\) let \(H_j{\subset M}\) be a submanifold with co-dimension \(k_j\) that is regular and uniformly conormally transverse for p in the window [ab] (with constant \({\mathfrak {I}}_{_{\!0}}^j\) as in (2.11)). Then, there exists \(\tau _0>0\) with the following property. Let \(\Lambda >\Lambda _{\max }\), and \(t_0>0\). For \(j=1,2\) let \(\textbf{T}_j\) be a sub-logarithmic resolution function  with \(\Lambda \Omega (\textbf{T}_j)<1-2\delta \) and such that the submanifold \(H_j\) is \(\textbf{T}_j\) non-recurrent in the window [ab] via \({\tau _0}\)-coverings with constant \(C_{_{\!{\text {nr}}}}^j\). Suppose \((H_1,H_2)\) is a \((t_0,\textbf{T}_{\max })\) non-looping pair in the window [ab] via \({\tau _0}\)-coverings with constant \(C_{_{\!{\text {nl}}}}\) where \(\textbf{T}_{\max }=\max (\textbf{T}_1,\textbf{T}_2)\). Let \(h^\delta \le R(h)=o(1)\) and for \(j=1,2\) let \(\{{\tilde{H}}_{j,h}\}_h\) be a family of submanifolds of codimension \(k_j\) that is regular, uniformly conormally transverse for p in the window [ab], and satisfies

$$\begin{aligned} \sup \Big \{d\big (\rho , \Sigma _{_{[a,b]}}^{{\tilde{H}}_{j,h}}\big )\,\big |\,\rho \in \Sigma _{_{[a,b]}}^{H_j}\Big \}\le R(h). \end{aligned}$$

Then, there is \(h_0>0\) such that for all \(0<h\le h_0\) and \({s}\in [a-Kh,b+Kh]\),

$$\begin{aligned} \Big |E_{_{{{\tilde{H}}}_{1,h}, {{\tilde{H}}}_{2,h}}}^{^{A_1,A_2}}(t_0,h; {s})\Big |\le C_{_{0}}\, h^{1-\frac{k_1+k_2}{2}}\Big /\!\sqrt{ \textbf{T}_1(R(h))\textbf{T}_2(R(h))}. \end{aligned}$$

We also have the following corollary involving the definitions of non-looping (Definition  1.12) and non-recurrence (Definition 1.13).

Theorem 9

Let \({\mathfrak {t}}\) be a resolution function, \(\Lambda >\Lambda _{\max }\), \(K>0\), \({\varepsilon }>0\), \(R_0>0\), \(0< \delta <\frac{1}{2}\), and for \(j=1,2\) let \(\textbf{T}_j\) be a sub-logarithmic resolution function with \(\Lambda \Omega (\textbf{T}_j)<1-2\delta \) and let \(\textbf{T}_{\max }=\max (\textbf{T}_1,\textbf{T}_2)\). Suppose the same assumptions as Theorem 8, but assume instead that for \(j=1,2\) the submanifold \(H_j\) is \({({\mathfrak {t}},\textbf{T}_j)}\) non-recurrent in the window [ab] at scale \(R_0\), and \((H_1,H_2)\) is a \((t_0, {\textbf{T}_{\max }})\) non-looping pair in the window [ab] with constant \(C_{_{\!{\text {nl}}}}\). Then, there exist \(C_{_{0}}={C_{_{0}}(n, k_1, k_2, A_1, A_2, {\mathfrak {I}}_{_{\!0}}^1, {\mathfrak {I}}_{_{\!0}}^2, {{\mathfrak {t}}}, C_{_{\!{\text {nl}}}})}\) and \(h_0>0\) such that for all \(0<h\le h_0\) and \({s}\in [a-Kh,b+Kh]\)

$$\begin{aligned} \Big |E_{_{{{\tilde{H}}}_{1,h}, {{\tilde{H}}}_{2,h}}}^{^{A_1,A_2}}(t_0+{\varepsilon },h; {s})\Big |\le C_{_{0}}\, h^{1-\frac{k_1+k_2}{2}}\Big /\!\sqrt{\textbf{T}_1(R(h))\textbf{T}_2(R(h))}. \end{aligned}$$

For the proof of Theorem 8, see Sect. 6.2 and for the proof of Theorem 9 see Sect. 9.

2.4 Application to the Laplacian

In this section we show how to obtain Theorems 34, and 5 from Theorem 9. It will be convenient here and below to use semiclassical Sobolev spaces defined for \(s\in {\mathbb {R}}\) by the norms

$$\begin{aligned} \Vert u\Vert _{H_{\text {scl}}^s(M)}^2:=\langle (-h^2\Delta _g+1)^su,u\rangle _{L^2(M)}. \end{aligned}$$
(2.12)

To pass from Theorem 9 to theorems about the Laplacian, we work with an operator Q such that \(\sigma (Q)(x,\xi )=|\xi |_{g(x)}\) near \(\{(x,\xi ):\,|\xi |_{g(x)}=1\}\), Theorem 9 applies with \(P=Q\), and for \(\lambda =h^{-1}\) and all \(N>0\)

$$\begin{aligned} \mathbb {1}_{(-\infty ,1]}(Q)= & {} \Pi _{\lambda },\quad (\rho _{_{h,t_0}}*\mathbb {1}_{(-\infty ,s]}(Q))(1)\nonumber \\= & {} \rho _{_{t_0}}*\Pi _{\lambda }+O(h^\infty )_{H_{scl }^{-N}\rightarrow H_{scl }^N}. \end{aligned}$$
(2.13)

Recall that \(\rho _{_{h,t_0}}\) is defined as in (1.16). To build Q, let \(\psi _1,\psi _2\in C_c^\infty ({\mathbb {R}};[0,1])\) with \({{\,\textrm{supp}\,}}\psi _1\subset (-1/4,1/4)\), \({{\,\textrm{supp}\,}}\psi _2\subset [-16,16]\), \(\psi _1 \equiv 1\) on \([-1/16,1/16]\) and \(\psi _2 \equiv 1\) on \([-4,4]\). We claim

$$\begin{aligned} Q= & {} (1-\psi _1(-h^2\Delta _g))\psi _2(-h^2\Delta _g)\sqrt{-h^2\Delta _g}\nonumber \\{} & {} -h^2\Delta _g(1-\psi _2(-h^2\Delta _g)) \end{aligned}$$
(2.14)

satisfies the desired properties. Observe that the second term in (2.14) is added to make Q classically elliptic, and that we use \(-h^2\Delta _g\) rather than \(\sqrt{-h^2\Delta _g}\) in order to apply [48, Theorem 14.9] to obtain \(Q\in \Psi ^2(M)\). Note also that Q is self-adjoint and \(\sigma (Q)=|\xi |_g\) on \(\{\frac{1}{2}\le |\xi |_g\le 2\}\),

$$\begin{aligned} \rho _{_{t_0}}*\Pi _{\lambda }{} & {} =\Big (\rho _{_{t_0,h}}*\mathbb {1}_{(-\infty ,s]}\Big (\sqrt{-h^2\Delta _g}\Big )\Big )(1),\nonumber \\ \Pi _\lambda{} & {} =\mathbb {1}_{(-\infty ,1]}\Big (\sqrt{-h^2\Delta _g}\Big ) \end{aligned}$$
(2.15)
$$\begin{aligned} \mathbb {1}_{(-\infty ,s]}(Q){} & {} =\mathbb {1}_{(-\infty ,s]}\Big (\sqrt{-h^2\Delta _g}\Big ),\qquad s\in [\tfrac{1}{2},2] \end{aligned}$$
(2.16)

and \(\mathbb {1}_{(-\infty ,s]}(Q)=\mathbb {1}_{(-\infty ,s]}(\sqrt{-h^2\Delta _g})=0\) for \(s<0\). Finally, we use the ellipticity of both Q and \(-h^2\Delta _g\) to obtain that for \(N\ge 0\)

$$\begin{aligned} \mathbb {1}_{(-\infty ,s]}(Q)&=O_{_{N}}(\langle s\rangle ^{N})_{H_{scl }^{-N}\rightarrow H_{scl }^N},\qquad \mathbb {1}_{(-\infty ,s]}\Big (\sqrt{-h^2\Delta _g}\Big )\\&=O_{_{N}}(\langle s\rangle ^{2N})_{H_{scl }^{-N}\rightarrow H_{scl }^N}. \end{aligned}$$

Now, for all \(N>0\) and \(L>1\) there is \(C_{_{N,L}}>0\) so that \(|\rho \Big (\frac{t_0}{h}(1-s)\Big )|\le C_{_{N,L}} h^{2N+L}\langle s\rangle ^{-2N-L}\) on \(|s-1|>\frac{1}{2}\). Therefore

$$\begin{aligned}{} & {} \Big [\rho _{t_0,h}*\Big (\mathbb {1}_{(-\infty ,s]}(Q)-\mathbb {1}_{(-\infty ,s]}(\sqrt{-h^2\Delta _g})\Big )\Big ](1)\nonumber \\{} & {} \quad =\int _{\begin{array}{c} s\notin [1/2,2]\\ s\ge 0 \end{array}}\frac{t_0}{h}\rho \Big (\frac{t_0}{h}(1-s)\Big )\Big (\mathbb {1}_{(-\infty ,s]}(Q)-\mathbb {1}_{(-\infty ,s]}\Big (\sqrt{-h^2\Delta _g}\Big )\Big )ds\nonumber \\{} & {} \quad =O_{_{N}}(h^{2N+L-1})_{H_{scl }^{-N}\rightarrow H_{scl }^N}. \end{aligned}$$
(2.17)

Combining (2.15) with (2.16) and (2.17), we obtain (2.13).

Now, every submanifold is conormally transverse for \(p(x,\xi )=|\xi |_{g(x)}\) at \(p^{-1}(1)\) with constant \({\mathfrak {I}}_{_{\!0}}=1\). Therefore, Theorems 34, and 5 follow from Theorem 9. To see this, we set \(P=Q\), \(a=b=1\), and observe that the Hamiltonian flow for \(\sigma (Q)\) near \(S^*_xM\) is equal to the geodesic flow. In particular, the dynamical definitions 1.12 and 1.13 applied to Q at \(E=1\) are exactly Definitions 1.3 and 1.7 with \(S^*_xM\) replaced by \(SN^*H\). This is true because Definitions 1.3 and 1.7 are stated with \(\varphi _t\) being the homogeneous geodesic flow, i.e., the flow generated by \(|\xi |_{g(x)}\). Next, we apply Theorem 5 with \(\Lambda =2\Lambda _{\max }{+1}\), \(h=\lambda ^{-1}\), and work with the resolution functions  \({\widetilde{\textbf{T}}}_j=(\Lambda {\Omega _0})^{-1}(1-2\delta )\textbf{T}_j\) for \(j=1,2\).

3 Dynamical assumptions and coverings

In this section we relate the non-looping and non-recurrence concepts introduced in Definitions 1.12 , 1.13, to their analogues via coverings given in Definitions 2.1, 2.2.

Proposition 3.1

Let \(H_1, H_2 \subset M\) be smooth submanifolds. Let \(a,b \in {\mathbb {R}}\) be such that \(H_1, H_2\) are conormally transverse for p in the window [ab], and \(\tau _0>0\). Let \(t_0>0\), \(\textbf{T}\) a resolution function, and suppose \((H_1,H_2)\) is a \((t_0,\textbf{T})\) non-looping pair in the window [ab] with constant \(C_{_{\!{\text {nl}}}}\). Then, there is \({\widetilde{C_{_{\!{\text {nl}}}}}}={\widetilde{C_{_{\!{\text {nl}}}}}}(p,a,b,n, C_{_{\!{\text {nl}}}},{H_1, H_2})>0\) such that \((H_1,H_2)\) is a \((t_0{+{3}\tau _0}, {{\widetilde{\textbf{T}}}})\) non-looping pair in the window [ab] via \(\tau _0\)-coverings with constant \({\widetilde{C_{_{\!{\text {nl}}}}}}\) and with \({\widetilde{\textbf{T}}}(R)=\textbf{T}(4R){-{3}\tau _0}\).

Before proving the proposition, we record some facts about sub-logarithmic resolution functions.

Lemma 3.2

Suppose \(\textbf{T}\) is a sub-logarithmic resolution function.

  1. (1)

    For \(0<a<b<1\),

    $$\begin{aligned} {\textbf{T}(b)\le \textbf{T}(a)}\le \frac{\log a}{\log b}\,{\textbf{T}(b)}. \end{aligned}$$

    In particular, \(\textbf{T}(R)\le \frac{\log R}{\log \mu +\log R} \textbf{T}(\mu R)\) for \(0<\mu <R^{-1}\).

  2. (2)

    Let \(f(s):=-\log (\textbf{T}^{-1}(s))\). Then, f extends to a differentiable function on \([0,\infty )\), \(f(0)=0\), and \(f(a)\le \frac{a}{b}f(b)\) for \(0<a<b\).

  3. (3)

    Let \(0<\delta <\frac{1}{2}\), and \(R(h)\ge h^\delta \) with \(R(h)=o(1)\). Then for all \(\Lambda >\Lambda _{\max }\), \(\varepsilon >0\), there is \(h_0>0\) such that for \(0<h<h_0\)

    $$\begin{aligned} \textbf{T}(R(h))\le (\Omega (\textbf{T})\Lambda +\varepsilon )T_e(h). \end{aligned}$$

Proof

Note that

$$\begin{aligned} 0\le \log {\frac{\textbf{T}(a)}{\textbf{T}(b)}}={-}\int ^{b}_{a}\frac{\textbf{T}'(s)}{\textbf{T}(s)}ds\le \int _{a}^{b}\frac{1}{s\log s^{-1}}ds= \log \Big (\frac{\log a^{-1}}{\log b^{-1}}\Big ), \end{aligned}$$

and hence the first claim holds. For the second claim, observe that since \(\textbf{T}\) is sub-logarithmic, \( f'(s)\ge -\frac{\log (\textbf{T}^{-1}(s))}{s}=\frac{f(s)}{s}. \)

To prove the last claim, observe that since \(R(h)=o(1)\), for all \(\Lambda >\Lambda _{\max }\) and \(\varepsilon >0\), there is \(h_0>0\) such that for \(0<h<h_0\),

$$\begin{aligned} \textbf{T}(R(h))\le (\Omega (\textbf{T})+\varepsilon {\Lambda ^{-1}})\log R(h)^{-1}\le (\Omega (\textbf{T})\Lambda +\varepsilon ) T_e(h). \end{aligned}$$

The second inequality follows from definitions (1.2), (1.11), and \(R(h)\ge h^\delta \) with \(0<\delta <\frac{1}{2}\). \(\square \)

In the following lemma we explain how to partition a \(({\mathfrak {D}},\tau ,r)\)-good cover for \( \Sigma _{_{E}}^{H_1}\) into tubes that do not loop through \( \Sigma _{_{E}}^{H_2}\) for times in \((t_0, T)\), and tubes that are ‘bad’ in the sense that they do loop through \( \Sigma _{_{E}}^{H_2}\). We do this while controlling the number of ‘bad‘ tubes in terms of the size of the set \( {{\mathcal {L}}_{_{{H_1,H_2}}}^{{S},E}(t_0, T)}\) for \(S>4r\).

Lemma 3.3

Let \(a,b\in {\mathbb {R}}\), \(H_1, H_2 \subset M\) be smooth submanifolds such that \(H_1, H_2\) are conormally transverse for p in the window [ab]. Then there is \(C_{_{0}}=C_{_{0}}(p,a,b,n,{H_1,H_2})\) such that the following holds. Let \(\tau _0>0\), \(r>0\), and \(0<\tau <\tau _0\). For \(i=1,2\) let \(\{\mathcal {T}_{j}^i(r)\}_{j\in {\mathcal {J}}^i(r)}\) be a \(({\mathfrak {D}},\tau ,r)\)-good cover of \(\Sigma ^{H_i}_{[a,b]}\). Let \(t_0>0\), \(T>0\). Then, for all \(E\in [a,b]\) and \( S\ge 4r \) there is a splitting \(\mathcal {J}_{_{E}}^1(r)=\mathcal {B}_{_{E}}^1(r)\cup \mathcal {G}_{_{E}}^1(r)\) such that

  1. (1)

    for \(j\in \mathcal {G}_{_{E}}^1(r)\) and \(k\in \mathcal {J}_{_{E}}^2(r)\)

    $$\begin{aligned} \bigcup _{t_0+{{2}(\tau +r)}\le |t|\le T-{{2}(\tau +r)}}\varphi _t(\mathcal {T}_{j}^1(r))\cap \mathcal {T}_{k}^2(r)=\emptyset , \end{aligned}$$
  2. (2)

    .

Proof

For \(j=1,2\) let \({\mathcal {Z}}_j\subset T^*M\) be the hypersurface transverse to the flow, with \( \Sigma _{_{[a,b]}}^{H_j} \subset {\mathcal {Z}}_j\), used to build the tubes of the cover, as explained in (2.1). Let \(E\in [a,b]\) and for \(S>0\) set

$$\begin{aligned} \mathcal {B}_{_{E}}^1(r):=\big \{ j\in \mathcal {J}_{_{E}}^1(r)\,:\, {\mathcal {T}}_j^1(r)\cap B_{_{{\mathcal {Z}}_{1}}}(\mathcal {L}_{_{H_1,H_2}}^{{S,E}}(t_0,T),2r)\ne \emptyset \big \}. \end{aligned}$$

Then, for \(j\in \mathcal {B}_{_{E}}^{1}(r)\),

$$\begin{aligned} {{\mathcal {T}}_j^1(r)}\cap {\mathcal {Z}}_{1}\;\subset B_{_{{\mathcal {Z}}_{1}}}(\mathcal {L}_{_{H_1,H_2}}^{{S,E}}(t_0,T),4r). \end{aligned}$$

In particular, there exists \(C_{_{0}}={C_{_{0}}(p,a,b,n,{H_1,H_2})}>0\) such that for all \(S\ge 4r\)

This proves the claim in (2).

To see the claim in (1), let \(j\in \mathcal {G}_{_{E}}^1(r):=\mathcal {J}_{_{E}}^1(r)\setminus \mathcal {B}_{_{E}}^1(r)\). Then, \({{\mathcal {T}}_j^{{1}}}(r)={\Lambda _{\rho _j}^\tau }(r)\) for some \(\rho _j\in {\mathcal {Z}}_{1}\) with \(d(\rho _j, \Sigma _{_{E}}^{H_1})<2r\) and . This yields that there is \(\rho _0\in \Sigma _{_{E}}^{H_1}\setminus \mathcal {L}_{_{H_1,H_2}}^{{S,E}}(t_0,T)\) such that \(d(\rho _0,\rho _j)<2r\). In particular, since \( \underset{t_0\le |t|\le T}{\bigcup }\varphi _t(B(\rho _0,S))\cap B( \Sigma _{_{E}}^{H_2},S)=\emptyset \) and \( {\mathcal {T}}_j^1(r)\subset \underset{|t|\le {\tau +r}}{\bigcup }\varphi _t(B(\rho _0,3r)), \) this yields

$$\begin{aligned} \bigcup _{t_0+{\tau +r}\le |t|\le T-{(\tau +r)}}\varphi _t({\mathcal {T}}_j^1(r))\cap B( \Sigma _{_{E}}^{H_2},S)=\emptyset \end{aligned}$$
(3.1)

for \(S\ge 4r\). On the other hand, since for all \(k \in \mathcal {J}_{_{E}}^2(r)\), we have \( {\mathcal {T}}_k^2(r)\cap {\mathcal {Z}}_{2}\subset B( \Sigma _{_{E}}^{H_2},3r), \)

$$\begin{aligned} {\mathcal {T}}_k^2(r)\subset \bigcup _{|t|\le {\tau +r}}\varphi _t(B( \Sigma _{_{E}}^{H_2},3r)) \end{aligned}$$
(3.2)

In particular, combining (3.1) and (3.2) we have

$$\begin{aligned} \bigcup _{t_0+{2}{(\tau +r)}\le |t|\le T-{{2}(\tau +r)}}\varphi _t({\mathcal {T}}_j^1(r))\cap B( \Sigma _{_{E}}^{H_2},S)=\emptyset . \end{aligned}$$

Thus, the claim (1) holds, provided \(S\ge 4r\). \(\square \)

With Lemmas 3.2 and 3.3 in place, we are now ready to prove Proposition 3.1.

Proof of Proposition 3.1

Let \(C_{_{0}}=C_{_{0}}(p, a, b,n,{H_1,H_2})\) be as in Lemma 3.3. We apply Lemma 3.3 with \(r=R\), \(T=\textbf{T}(S)\), \(S=4R\), \(0<R<\frac{1}{2}\tau _0\). This shows that \((H_1,H_2)\) is a \([t_0{+{3}\tau _0}, {\widetilde{\textbf{T}}}]\) non-looping pair in the window [ab] via \(\tau \)-coverings with constant \({\widetilde{C_{_{\!{\text {nl}}}}}}=C_0^2 C_{_{\!{\text {nl}}}}\). \(\square \)

Lemma 3.4

There is a constant \(C_n>0\), depending only on n, such that the following holds. Let \(\tau _0>0\), \(t_0>0\), \(H_1,H_2\subset M\) be smooth submanifolds such that \(H_1\) and \(H_2\) are conormally transverse for p in the window [ab]. Let \(\textbf{T}\) be a resolution function. If \((H_1,H_2)\) is a \((t_0,\textbf{T})\) non-looping pair in the window [ab] via \(\tau _0\)-coverings with constant \(C_{_{\!{\text {nl}}}}\), then \((H_1,H_2)\) is a \((t_0,{{\widetilde{\textbf{T}}}})\) non-looping pair in the window [ab] with constant \(C_{_{\!{\text {nl}}}}{C_n}\) and \({\widetilde{\textbf{T}}}(R)=\textbf{T}(2R)\).

Proof

Let \(E\in [a,b]\) and fix \(i,j \in \{1,2\}\), \(i\ne j\). For each \(R>0\) consider the non-looping partition \(\mathcal {J}_{_{E}}^i({R})=\mathcal {G}_{_{E}}^i(R)\sqcup \mathcal {B}_{_{E}}^i({R})\) given by Definition (2.1). Let \(\rho \in \mathcal {L}_{_{H_i,H_j}}^{{{R}/2,E}}(t_0,{\textbf{T}(R)})\). Then, there are \(\rho _1\in B(\rho ,{R}/2)\) and \(t_0\le |t|\le {\textbf{T}(R)}\) such that \(\varphi _t(\rho _1)\in B( \Sigma _{_{E}}^{H_j},{R}/2)\). Hence, there is \(\ell \in {\mathcal {B}_{_{E}}^i}({R})\) such that \(\rho _1\in {\mathcal {T}}^i_\ell ({R})\) and hence \(\rho \in {\mathcal {T}}^i_\ell (2{R})\). This implies \( B_{_{{ \Sigma _{_{E}}^{H_i}}}}(\rho ,{R}/2)\subset {{\mathcal {T}}^i_\ell (3{R})}. \) Thus,

In particular, there exists \(C_{{n}}>0\) such that

Therefore,

The lemma follows from Definition 1.12 after taking the limit \(R \rightarrow 0^+\) and redefining \(C_n\). \(\square \)

Proposition 3.5

Let \({\mathfrak {t}}\), \(\textbf{T}\) be resolution functions and \(H \subset M\) be a smooth submanifold. Let \(a,b \in {\mathbb {R}}\) be such that H is conormally transverse for p in the window [ab]. Suppose H is \({({\mathfrak {t}},\textbf{T})}\) non-recurrent in the window [ab] at scale \(R_0\).

Then, there exists \({C_{_{\!{\text {nr}}}}=C_{_{\!{\text {nr}}}}}(M,p,{{\mathfrak {t}}},{R_0})>0\) such that for all \(\tau _0>0\), there is a resolution function \({\widetilde{\textbf{T}}}\) such that the submanifold H is \({\widetilde{\textbf{T}}}\) non-recurrent in the window [ab] via \(\tau _0\)-coverings with constant \({C_{_{\!{\text {nr}}}}}\). Moreover, there is \(c>0\) such that if \(\textbf{T}\) is sub-logarithmic, then \({\widetilde{\textbf{T}}}(R)\ge c\textbf{T}(R)\) for all R.

The proof of this result hinges on two lemmas. To state the first one, we introduce a slight adaptation of  [8, Definition 3]. Let \(\varepsilon _0>0\), \(\digamma >0\), \({\mathfrak {t}}_0:[\varepsilon _0, +\infty ) \rightarrow [1, +\infty )\), and \(f:[0,\infty )\rightarrow [0,\infty )\). We say a set \(A_0\) is \((\varepsilon _0,{\mathfrak {t}}_0,\digamma , f)\) controlled up to time T provided it is \((\varepsilon _0, {\mathfrak {t}}_0, \digamma )\) controlled up to time T in the sense of  [8, Definition 3] except that we replace the condition on r by

$$\begin{aligned} 0<r<\tfrac{1}{\digamma }e^{-\digamma \Lambda T-f(T)}r_0 \end{aligned}$$
(3.3)

and replace point (3) by

$$\begin{aligned} \inf _{k}R_{1,k} \ge {\tfrac{1}{4}e^{-f(T)}}\inf _i R_{0,i}. \end{aligned}$$
(3.4)

Next, fix \(E \in [a,b]\). Since H is \( {({\mathfrak {t}},}\textbf{T})\) non-recurrent in the window [ab] at scale \({R_0}\), for all \(\rho \in \Sigma _{_{E}}^H\) there exists a choice of ± such that for all \(A \subset {B_{_{ \Sigma _{_{E}}^H}}(\rho ,R_0)}\), \(0<R<R_0\), \(\varepsilon >0\), and \(T>{{\mathfrak {t}}(\varepsilon )}\)

(3.5)

with f as in Lemma 3.2. Then, extract a finite cover of \( \Sigma _{_{E}}^H\) by balls \({\tilde{B}}_\rho =B(\rho ,R_{{0}}/2)\) and set

$$\begin{aligned} {\tilde{\mathcal {A}}}_{_{E}}:=\{{{\tilde{B}}}_{\rho _i}\}_{i=1}^K, \qquad \text {and}\qquad \mathcal {A}_{_{E}}:=\{{B}_{\rho _i}\}_{i=1}^K, \end{aligned}$$
(3.6)

where \({B}_\rho =B(\rho ,R_{{0}})\). Note that, again using that H is non-recurrent with at scale \(R_0\), we may assume \(K\le C_n R_0^{1-n}\) where \(C_n\) is a constant depending only on n.

Lemma 3.6

Let H, \({\mathfrak {t}}\) and \(\textbf{T}\) be as in Proposition 3.5 and \(f(T):= -\log (\textbf{T}^{-1}(T))\). Then, there exist \(c_n>0\) depending only on n and \(\digamma >0\) such that for all \(E\in [a,b]\) and \({T>1}\) every ball in \(\mathcal {A}_{_{E}}\) is \((0,{{\mathfrak {t}}_0},\digamma , f)\) controlled up to time T with \({\mathfrak {t}}_0(\varepsilon )={\mathfrak {t}}(c_n\varepsilon )\).

Proof

Let \(E\in [a,b]\). Let \(A_0:=B_{\rho _0}\) for some \(B_{\rho _0}\in \mathcal {A}_{_{E}}\), \(\varepsilon _1>0\), \(\Lambda {>} \Lambda _{\max }\), and \(0<\tau <\tfrac{1}{2}\tau _{_{{{\,\textrm{inj}\,}}_H}}\). Let \(T> {1}\) and \(0\le {{\tilde{R}}}_0 \le \frac{1}{\digamma }e^{-\digamma \Lambda T }\) for \(\digamma >{2R_0^{-1}}\) to be determined later. Let \(0<r_0<{{\tilde{R}}}_0\). Suppose \(A_1\subset A_0\) and \(\{B_{0,i}\}_{i=1}^N\) are balls centered in \(A_0\) with radii \(R_{0,i}\in [r_0,{{\tilde{R}}}_0]\) such that \( A_1\subset \cup _{i=1}^N B_{0,i}\subset A_0. \)

Let \(R:=\tfrac{1}{2}\inf _i R_{0,i}\). There exist \(C_n>0\), depending only on n, and a collection of balls \(\{{\tilde{B}}_{0,i}\}_{i=1}^{N_0}\) of radius R, such that

$$\begin{aligned} A_1\subset \bigcup _{i=1}^{N_0}{\tilde{B}}_{0,i},\qquad N_0R^{n-1}\le C_n\sum _{i=1}^NR_{0,i}^{n-1}. \end{aligned}$$
(3.7)

Fix \(0\le r \le \frac{1}{\digamma }e^{-\digamma \Lambda T{-f(T)} }r_0\). Next, let \(\{B(q_j, r)\}_{j\in \mathcal {J}} \subset \Sigma _{_{E}}^H\) be a cover of \( \Sigma _{_{E}}^H\) by balls of radius r such that there are at most \({\mathfrak {D}}_n\) balls over each point in \( \Sigma _{_{E}}^H\), where \({\mathfrak {D}}_n>0\) depends only on n. Assume, without loss of generality, that (3.5) holds for \(\rho _0\) with the choice \(\pm =+\). Next, set \(\mathcal {J}_{_{\!A_1}}:=\{j \in \mathcal {J}:\; B(q_j, {\frac{1}{2}}e^{-f(T)}R)\cap \mathcal {R}^{{e^{-f(T)}}R}_{_{{A_1, +}}}({\mathfrak {t}}{(\varepsilon _1)},T)\ne \emptyset \}\). Defining the collection

$$\begin{aligned} \{B_{1,i}\}_{i=1}^{N_1}:=\Big \{ B_{_{{ \Sigma _{_{E}}^H}}}\big (q_j, \tfrac{1}{2}e^{-f(T)}R\big ):\; j \in \mathcal {J}_{_{\!A_1}} \Big \}, \end{aligned}$$

we have . Then, letting \(R_{1,i}:=\tfrac{1}{2}e^{-f(T)}R\), we have \(R_{1,i}\in [0,\tfrac{1}{4}{{\tilde{R}}}_0]\), and using that \(R<R_0/2\) the bound in (3.5) applied to \(A_1\) yields

$$\begin{aligned} \sum _{i=1}^{N_1}R_{1,i}^{n-1} \le {\varepsilon _1}\,{\mathfrak {D}}_n\, \mu _{_{ \Sigma _{_{E}}^H}}(B_{_{ \Sigma _{_{E}}^H}}(A_1,R) \Big ). \end{aligned}$$
(3.8)

Next, by (3.7) note that \( B_{_{ \Sigma _{_{E}}^H}}(A_1,R)\subset \bigcup _{i=1}^{N_0}2{\tilde{B}}_{0,i}, \) where \(2{\tilde{B}}_{0,i}\) denotes the ball with the same center as \({\tilde{B}}_{0,i}\) but with radius 2R. Using (3.7) again there is \(C_n>0\) such that

$$\begin{aligned} \mu _{_{ \Sigma _{_{E}}^H}}(B_{_{ \Sigma _{_{E}}^H}}(A_1,R))\le {\mu _{_{ \Sigma _{_{E}}^H}} \Big (\bigcup _{i=1}^{N_0}2{\tilde{B}}_{0,i}\Big )} \le C_n\sum _{i=1}^N R_{0,i}^{n-1}. \end{aligned}$$
(3.9)

Let \({\varepsilon :=\varepsilon _1} C_n {\mathfrak {D}}_n\). Combining (3.8) and (3.9) yields point (2) of [8, Definition 3] with \({\mathfrak {t}}_0(\varepsilon )={\mathfrak {t}}(\varepsilon /(C_n{\mathfrak {D}}_n))\). By the definition of R, we also note that point (3), which was replaced by (3.4), also holds.

It remains to check point (1) i.e. there is \(\digamma >0\) such that \(\Lambda ^\tau _{A_1\setminus \cup _{k}B_{1,k}}(r)\) is \([{\mathfrak {t}}_0{(\varepsilon )},T]\) non-self looping for \( 0<r<\frac{1}{\digamma }e^{-\digamma \Lambda T{-f(T)}}R. \) For this, suppose \(\rho _1,\rho _2\in \Lambda ^{\tau }_{A_1\setminus \cup _{k}B_{1,k}}(r)\) and \(t\in [{{\mathfrak {t}}_0(\varepsilon )},T]\) such that \( \varphi _t(\rho _1)=\rho _2. \) Then, there are \(s_1,s_2\in {[-\tau -r,\tau +r]}\), \(q_1,q_2\in A_1\setminus \cup _k B_{1,k}\) such that \(d(\rho _i,\varphi _{s_i}(q_i))<{r}\). In particular, there is \(C_{_{0}}>0\) depending only on \((M,p,a,b,\Lambda )\) such that

$$\begin{aligned} d(\varphi _{s_2-t-s_1}(q_2), A_1)<(1+C_{_{0}}e^{\Lambda (|t|+{2\tau +2r})})r. \end{aligned}$$

Finally, let \(\digamma >0\) be large enough so that \(\frac{1}{\digamma }e^{-\digamma \Lambda T}< \min ((1+C_{_{0}} e^{\Lambda (|T|+{2\tau +2r})})^{-1},R_0/2)\). Note that the choice of \(\digamma \) does not need to depend on T. Then, since \( r<(1+C_{_{0}}e^{\Lambda (|{T}|+{2\tau +2r})})^{-1}{e^{-f(T)}}R, \) we have \(q_2\in \mathcal {R}^{{e^{-f(T)}}R}_{_{A_1{,+}}}({{\mathfrak {t}}_0(\varepsilon )},T)\), which is a contradiction since \( \mathcal {R}^{{e^{-f(T)}}R}_{_{A_1{,+}}}({{\mathfrak {t}}_0(\varepsilon )},T)\subset \cup _i B_{1,i}\). \(\square \)

In what follows we fix \(1<\beta _0 < \varepsilon _0^{-1}\) and define

$$\begin{aligned} {\textbf{F}}(T):=\sum _{k=0}^{\log _{\beta _0}T}f\big (\beta _0^{-k}\,T\big ). \end{aligned}$$

Lemma 3.7

Let \(B\subset \Sigma _{_{E}}^H\) be a ball of radius \(\delta >0\). Let \({0<\varepsilon _0<1}\), \({\mathfrak {t}}_0:[\varepsilon _0, +\infty ) \rightarrow [1, +\infty )\), \(f:[0,\infty )\rightarrow [0,\infty )\) increasing with \(f(e^{-x})\in L^1([0,\infty ))\), \(T_0>0\), and \(\digamma >0\), such that B can be \((\varepsilon _0, {{\mathfrak {t}}_0}, \digamma ,f)\)-controlled up to time \(T_0\). Let \(0<m<\frac{\log T_0-\log {{\mathfrak {t}}_0({\varepsilon _0})}}{\log \beta _0}\) be a positive integer, \(\Lambda > \Lambda _{\max }\),

$$\begin{aligned} 0< {{\tilde{R}}}_0\le {\min }\Big \{{\tfrac{1}{\digamma }}e^{-{\digamma }\Lambda T_0}, {\tfrac{\delta }{10}}\Big \}, \qquad 0<r_1<{\tfrac{1}{5\digamma }}e^{-({\digamma }\Lambda T_0+{\textbf{F}}(T_0)+{f(T_0)})}{{\tilde{R}}}_0, \end{aligned}$$

and \(B_0\subset B\) with \(d(B_0, B^c)>{{\tilde{R}}}_0\). Let \(0<\tau <\tau _0\) and suppose \(\{\Lambda _{_{\rho _j}}^\tau (r_1)\}_{j=1}^{N_{r_1}}\) is a \(({{\mathfrak {D}}},\tau , r_1)\) good cover of \(\Sigma _{_{\!H\!,p}}\) and set \( \mathcal {E}:=\{j \in \{1, \dots , {N_{r_1}}\}: \Lambda _{\rho _j}^\tau (r_1)\cap \Lambda ^\tau _{B_0}(\tfrac{r_1}{5})\ne \emptyset \}. \)

Then, there exist \(C_{_{\!M,p}}>0\) depending only on (Mp) and sets \(\{\mathcal {G}_{_{E,\ell }}\}_{\ell =0}^m\subset \{1,\dots N_{r_1}\}\), \(\mathcal {B}_{_{E}}\subset \{1,\dots N_{r_1}\}\) so that \(\mathcal {E}\;\subset \; \mathcal {B}_{_{E}}\cup \displaystyle \cup _{\ell =0}^m \mathcal {G}_{_{E,\ell }}\) and

$$\begin{aligned}&\bullet \;\bigcup _{i\in \mathcal {G}_{_{E,\ell }}}\Lambda _{\rho _i}^\tau (r_1)\;\text { is }\; \big [{\mathfrak {t}}_0(\varepsilon _0),\beta _0^{-\ell }T_0\big ]\; \text {non-self looping for}\, \ell \in \{0, \dots , m\}, \end{aligned}$$
(3.10)
$$\begin{aligned}&\bullet \; |\mathcal {G}_{_{E,\ell }}|\le C_{_{\!M,p}}{{\mathfrak {D}}}\varepsilon _0^\ell {\delta ^{n-1}} r_1^{1-n} \;\;\; {\text {for every}\;\; \ell \in \{0, \dots , m\}}, \end{aligned}$$
(3.11)
$$\begin{aligned}&\bullet \; |\mathcal {B}_{_{E}}|\le C_{_{\!M,p}}{{\mathfrak {D}}} \varepsilon _0^{m+1}{\delta ^{n-1}} r_1^{1-n}\Big .. \end{aligned}$$
(3.12)

Proof

The proof is the same as that of [8, Lemma 3.2], with a very minor modification. Namely, we replace \(R_0\) by \({\tilde{R}}_0\) and put \(r_0=e^{-{\textbf{F}}(T_0)}{{\tilde{R}}}_0\) instead of \(r_0=e^{2\textbf{D}\Lambda T_0}{{\tilde{R}}}_0\). We then obtain the following instead of the leftmost equation in [8, (3.21)]

$$\begin{aligned} \inf _{k}R_{2,k}\ge \tfrac{1}{4}e^{-f(T_0)}\inf _iR_{1,i}. \end{aligned}$$

Which in turn changes the leftmost equation in [8, (3.22)] to

$$\begin{aligned} \inf _{k}R_{\ell ,k}\ge e^{-{\textbf{F}}(T_0)}{{\tilde{R}}}_0=r_0. \end{aligned}$$

This follows from the argument below [8, Remark 8], that yields, since \(\ell \le m\),

$$\begin{aligned} \inf _{k}R_{\ell ,k}\ge \frac{1}{4^\ell }\prod _{j=0}^\ell e^{-f(\beta _0^{-j}T_0)}R_0=\frac{1}{4^\ell } e^{-\sum _{j=0}^\ell f(\beta _0^{-j}T_0)}{{\tilde{R}}}_0\ge e^{-{\textbf{F}}(T_0)}{{\tilde{R}}}_0. \end{aligned}$$

\(\square \)

With Lemmas 3.6 and 3.7 in place, we are now ready to prove Proposition 3.5.

Proof of Proposition 3.5

Let \(\{{\mathcal {T}}_j(R)\}_{j \in \mathcal {J}(h)}=\{\Lambda _{_{\rho _j}}^\tau (R)\}_{j\in \mathcal {J}(h)}\) be a \(({\mathfrak {D}}, \tau , R)\) good covering of \( \Sigma _{_{[a,b]}}^H\). Let \(E \in [a,b]\) and \(\mathcal {A}_{_{E}}:=\{B_{\rho _i}\}_{i=1}^K\) be the covering of \( \Sigma _{_{E}}^H\) as described in (3.6). Let \({\mathfrak {t}}_0\) be as in Lemma 3.6 and fix \(0<\varepsilon _0<\frac{1}{2}\). There exists \(\digamma >0\) such that each ball in \(\mathcal {A}_{_{E}}\) can be \((\varepsilon _0,{\mathfrak {t}}_0, \digamma , f)\) controlled for time \(T>1\).

We then apply Lemma 3.7 to each ball in \(\mathcal {A}_{_{E}}\). Let \(\delta _0:={R_0/2}\) be the radius of the balls in \(\mathcal {A}_{_{E}}\), and \({\textbf{T}_0=\textbf{T}_0(R)}\) such that \(\textbf{T}_0>{\mathfrak {t}}_0(\varepsilon _0)\) and

$$\begin{aligned} { R\le \frac{1}{10\digamma ^2}e^{-\big (2\digamma \Lambda \textbf{T}_0(R)+{\textbf{F}}(\textbf{T}_0(R))+{f(\textbf{T}_0(R))}\big )}.} \end{aligned}$$
(3.13)

Without loss of generality, we may assume \(\digamma \) is large enough so that \({\tfrac{1}{\digamma }}e^{-{\digamma }\Lambda {{\mathfrak {t}}_0(\varepsilon _0)}}\le {\tfrac{\delta _0}{10}}\). Then, putting \( {{\tilde{R}}}_0= \tfrac{1}{\digamma }e^{-{\digamma }\Lambda T_0}\) in Lemma 3.7, and using condition (3.13) allows us to set \(r_1={R}\) in Lemma 3.7 and apply it to each ball \(B_{\rho _0}\) in \(\mathcal {A}_{_{E}}\). Let \({{\tilde{B}}}_{\rho _0}\) be the ball with the same center as \(B_{\rho _0}\) but with a radius \({R_{{0}}/2}\) so that \(d({{\tilde{B}}}_{\rho _0}, B_{\rho _0}^c){=}R_0/2>{{\tilde{R}}_0}\). Let \(\tau _0>0\), \(0<\tau <\tau _0\), and set \(\mathcal {J}_{_{E}}^{\rho _0}({R})=\{j \in \mathcal {J}_{_{E}}({R}):\; \Lambda _{_{\rho _j}}^\tau (R) \cap \Lambda _{_{{{\tilde{B}}}_{\rho _0}}}^\tau (\tfrac{1}{5}R)\ne \emptyset \}\), there is \(C_{_{\!M,p}}>0\) and sets \(\{{\mathcal {G}_{_{E,\ell }}}\}_{\ell =0}^m\subset \mathcal {J}_{_{E}}({R})\), \(\mathcal {B}_{_{E}}\subset \mathcal {J}_{_{E}}({R})\) so that \(\mathcal {J}_{_{E}}^{\rho _0}({R})\;\subset \; \mathcal {B}_{_{E}}\cup \displaystyle \cup _{\ell =0}^m{\mathcal {G}_{_{E,\ell }}}\), and (3.10), (3.11), (3.12) hold.

Therefore, letting \(T_\ell = \beta _0^{-\ell } {\textbf{T}_0}\) and \(t_\ell ={\mathfrak {t}}_0(\varepsilon _0)\) for \(1 \le \ell \le m\), and setting \(\mathcal {G}_{m+1}:=\mathcal {B}_{_{E}}\), \(T_{m+1}=t_{m+1}=1\), yields that there exists \({C_{_{\!{\text {nr}}}}=C_{_{\!{\text {nr}}}}(M,p{,{\mathfrak {t}}})>0}\) such that

$$\begin{aligned} R^{\frac{n-1}{2}}\sum _{\ell =0}^{m+1}\bigg (\frac{|{{\mathcal {G}}}_\ell | t_\ell }{T_\ell }\bigg )^{\!\!1/2} \le \bigg (\frac{C_{_{M,p}}{\mathfrak {D}} \delta _0^{{n-1}}}{{\textbf{T}_0(R)}} \sum _{\ell =0}^{m+1}{(\beta _0\varepsilon _0)^\ell }\bigg )^{\tfrac{1}{2}}\le \frac{C_{_{\!{\text {nr}}}}{\mathfrak {D}}^{\frac{1}{2}}}{\sqrt{\textbf{T}_0(R)}}. \end{aligned}$$

The existence of \(C_{_{\!{\text {nr}}}}>0\) is justified since \(\beta _0\varepsilon _0<1\). Repeating for each ball \(B_{\rho _i} \in \mathcal {A}_{_{E}}\) and using \(K\le C_nR_0^{1-n}\), proves that H is \(\textbf{T}_0\) non-recurrent in the window [ab] via \(\tau _0\)-coverings with constant \(C_{_{\!{\text {nr}}}}C_nR_0^{1-n}\).

By Lemma 3.2, when \(\textbf{T}\) is sub-logarithmic and \(0<a<b\) we have \( f(b)\ge \frac{b}{a}f(a). \) In particular,

$$\begin{aligned} {\textbf{F}}(T_0)=\sum _j f(2^{-j}T_0)\le \sum _j 2^{-j}f(T_0)\le 2f(T_0). \end{aligned}$$

Therefore, using \(f(T)=-\log (\textbf{T}^{-1}(T))\), there exists \(c>0\) such that we may define

$$\begin{aligned} \textbf{T}_0(R)=c f^{-1}(\log R)\ge c \textbf{T}(R). \end{aligned}$$

\(\square \)

Remark 3.8

We note that our definition of recurrence (Definition 1.13) is equivalent to the following. There is \(\digamma >0\) such that for all \(\rho \in \Sigma _{_{E}}^{H}\) there is \(R_0>0\) such that \(B(\rho ,R_0)\) is \((\varepsilon _0,{\mathfrak {t}}_0,\digamma ,f)\) controlled with an additional small modification of the definition of \((\varepsilon _0,{\mathfrak {t}}_0,\digamma ,f)\) controlled (see (3.3) and (3.4)): One needs to replace (1) by

$$\begin{aligned} \bigcup _{t_0\le \pm t\le T}\Lambda _{A_1\setminus \cup {\tilde{B}}_{1,k}}^\tau (r)\cap \Lambda _{A_1}^\tau (r)=\emptyset . \end{aligned}$$

To see these are equivalent, we identify \({B(\rho ,R_0)}\) with \(A_0\) and A with \(A_1\).

One can check that all of the proofs of being \((\varepsilon _0,{\mathfrak {t}}_0,\digamma , f)\) controlled in [8] actually prove this slightly stronger condition with \(f(T)=CT\) for some \(C>0\).

4 Basic estimates for averages over submanifolds

Let \(P(h) \in \Psi ^m(M)\) be a self-adjoint semiclassical pseudodifferential operator, with classically elliptic symbol p. Throughout this section we assume \(H \subset M\) is a smooth submanifold of co-dimension k, and \(a, b\in {\mathbb {R}}\) are such that H is conormally transverse for p in the window [ab].

As explained in Sect. 1.6, we crucially view the kernel of the spectral projector \(\mathbb {1}_{[t-s,t]}(P)\) as a quasimode for P. We are then able to use estimates from [11] to estimate the error when the projector is smoothed at very small scales. This section is dedicated to adapting the estimates from [11] to the current setup.

All our estimates are made in terms of \(({\mathfrak {D}},\tau , R(h))\)-good covers and \(\delta \)-partitions associated to them. For the definition of a good cover see (2.4). Note, in addition, that there is a constant \({\mathfrak {D}}_n\) depending only on n such that we may work with a \(({\mathfrak {D}}_n,\tau , R(h))\) good cover [10, Lemma 2.2] [11, Proposition 3.3].

We now define the concept of \(\delta \)-partitions. For \(0\le \delta <\frac{1}{2}\), we write

$$\begin{aligned} S_\delta ^m(T^*M):=\left\{ \begin{array}{ccc} a\in C^\infty (T^*M): \\ |\partial _x^\alpha \partial _\xi ^\beta a(x,\xi )|\le C_{\alpha \beta }h^{-\delta (|\alpha |+|\beta |)} \langle \xi \rangle ^{m-|\beta |}\end{array}\right\} , \end{aligned}$$
(4.1)

and write \(\Psi _\delta ^m(M)\) for the corresponding semiclassical pseudodifferential operators. We refer the reader to [11, Appendix A.2], [48, Chapters 4,9], [19, Appendix E] for more detailed accounts of these operators.

Let \(\tau >0\), \(0< \delta <\tfrac{1}{2}\), and \(R(h)\ge {h^\delta }\). Let \(\{{\mathcal {T}}_j\}_{_{j\in \mathcal {J}(h)}}\) be a \((\tau ,R(h))\)-cover for \( \Sigma _{_{[a,b]}}^H\) with \({\mathcal {T}}_j=\Lambda _{\rho _j}^\tau (R(h))\), and for \(E\in [a,b]\) let \(\mathcal {J}_{_{E}}(h):=\mathcal {J}_{_{E}}(R(h))\) as defined in (2.5). We say

$$\begin{aligned} \{\chi _{_{{\mathcal {T}}_j}}\}_{j\in \mathcal {J}_{_{E}}(h)} \subset S_\delta (T^*M;[0,1]) \end{aligned}$$
(4.2)

is a \(\delta \)-partition for \( \Sigma _{_{E}}^H\) associated to \(\{{\mathcal {T}}_j\}_{j\in \mathcal {J}(h)}\) provided the families \(\{\chi _j\}_{j\in \mathcal {J}_{_{E}}(h)}\) and \( \{h^{-1}[P,\chi _j]\}_{j\in \mathcal {J}_{_{E}}(h)}\) are bounded in \(S_\delta (T^*M;[0,1])\) and

$$\begin{aligned}{} & {} \text {(1)} {{\,\textrm{supp}\,}}\chi _j \subset \Lambda _{\rho _j}^\tau (R(h)),\text { for all }j \in \mathcal {J}_{_{E}}(h),\\{} & {} \text {(2)} \underset{j\in \mathcal {J}_{_{E}}(h)}{\sum } \chi _j \ge 1\text { on }\Lambda _{ \Sigma _{_{E}}^H}^{\tau /2}(\tfrac{1}{2}R(h)). \end{aligned}$$

For the construction of such a partition we refer the reader to [11, Proposition 3.4].

The next lemma controls the average of Au over a submanifold H in terms of the \(L^2\) masses of the bicharacteristic beams intersecting the microsupport of A. Here, u is a quasimode for P and A is a pseudodifferential operator. When we apply this lemma, u will be the kernel of the spectral projector onto a small window, and A will either represent a localizer to a family of tubes or differentiation in one of the coordinates.

To ease notation, for \({E} \in {\mathbb {R}}\) we write \(P_{_{E}}=P_{_{E}}(h)\)

$$\begin{aligned} P_{_{E}}:=P-E. \end{aligned}$$
(4.3)

In addition, given \(A \in \Psi _\delta ^\infty (M)\), \(\psi \in C^\infty _0({\mathbb {R}};[0,1])\), \(E\in {\mathbb {R}}\), \(h>0\), \(C>0\), \(C_{_{N}}>0\), and \(u \in \mathcal {D}'(M)\) we set \(\alpha :=\frac{k-2m+1}{2}\) and

$$\begin{aligned}&Q^{A,\psi }_{E,h}(C, C_{_{N}}, u) :=\nonumber \\&\quad Ch^{-\frac{1}{2}-\delta }\big \Vert \big (1-\psi \big (\tfrac{P_{_{E}}}{h^\delta }\big )\big )P_{_{E}}Au\big \Vert _{_{H_{{\text {scl}}}^\alpha }}+C_{_{N}}h^N\Big (\Vert u\Vert _{_{\!L^2(M)}}+ \Vert P_{_{E}}u\Vert _{_{H_{{\text {scl}}}^\alpha }}\Big ). \end{aligned}$$
(4.4)

We fix \(\varepsilon _0>0\) and a continuous family \([a-\varepsilon _0,b+\varepsilon _0]\ni E\mapsto B_{_{E}}\in \Psi _\delta ^0(M)\) such that

$$\begin{aligned}{} & {} {\hbox {MS}}_{\textrm{h}}(B_{_{E}})\subset \Lambda ^{\tau _0+{\varepsilon _0}}_{ \Sigma _{_{E}}^H}(3R(h)) \qquad \text {and}\nonumber \\ {}{} & {} {\hbox {MS}}_{\textrm{h}}(I-B_{_{E}})\cap \Lambda ^{\tau _0+{\varepsilon _0}}_{ \Sigma _{_{E}}^H}(2R(h)))=\emptyset . \end{aligned}$$
(4.5)

This will serve as a microlocalizer to the region of interest. We recall the constants \(\mathcal {K}_0\), \(\tau _{{{\,\textrm{inj}\,}}}\), \({\mathfrak {I}}_{_{\!0}}\) defined in (2.8), (2.3), and (2.11) respectively.

Lemma 4.1

There exist \( \tau _0=\tau _0(M,p,\tau _{{{\,\textrm{inj}\,}}},{\mathfrak {I}}_{_{\!0}})>0\) and \(R_0=R_0(M,p,k, \mathcal {K}_0, \tau _{{{\,\textrm{inj}\,}}}, {\mathfrak {I}}_{_{\!0}})>0, \) such that the following holds.

Let \(0<\tau <\tau _0\), \(0<\delta <\tfrac{1}{2}\) and \({h^\delta } \le R(h) \le R_0\). For \(h>0\) let \(\{{\mathcal {T}}_j\}_{j\in \mathcal {J}(h)}\) be a \(({\mathfrak {D}}_n, \tau , R(h))\) good cover of \( \Sigma _{_{[a,b]}}^H\). Let \(\mathcal {V}\subset S_\delta (T^*M;[0,1])\) be bounded. Let \(\psi \in C^\infty _0({\mathbb {R}};[0,1])\) with \(\psi (t)=1\) for \(|t| \le \tfrac{1}{4}\) and \(\psi (t)=0\) for \(|t| \ge 1\). Let \(\ell \in {\mathbb {R}}\), \(\mathcal {W}\) and \(\widetilde{\mathcal {W}}\) be bounded subsets of \(\Psi _\delta (M)\) and \(\Psi ^{\ell }_\delta (M)\) respectively, and \(B_{_{E}}\) be as in (4.5).

Then, there exist \(C_{_{0}}=C_{_{0}}(n,k,{\mathfrak {I}}_{_{\!0}},\mathcal {V},\mathcal {W},\widetilde{\mathcal {W}})\), \(C>0\), and for all \(K>0\) there is \(h_0>0\), such that for all \(N>0\) there exists \(C_{_{N}}>0\), with the following properties. For all \(u\in \mathcal {D}'(M)\), \(0<h<h_0\), \(E\in [a-Kh,b+Kh]\), every \(\delta \)-partition \(\{\chi _{_{{\mathcal {T}}_j}}\}_{j\in \mathcal {J}_{_{E}}(h)}\subset {\mathcal {V}}\) associated to \(\{{\mathcal {T}}_j\}_{j\in \mathcal {J}_{_{E}}(h)}\), and every \(A\in \widetilde{\mathcal {W}}\) such that \(B_{_{E}}\frac{1}{h}[P,A]\in \mathcal {W}\),

$$\begin{aligned}&h^{\frac{k-1}{2}}\Big |\int _{H}Au\,d\sigma _{_{\!H}}\Big | \nonumber \\&\quad \le C_{_{0}}{R(h)^{\frac{n-1}{2}}} \sum _{j \in \mathcal {I}_{_{E}}(h)} \bigg (\frac{\Vert Op_h({{\tilde{\chi }}}_{_{{\mathcal {T}}_j}})u\Vert _{_{\!L^2(M)}}}{\tau ^{\frac{1}{2}}}+\frac{C}{h}\Vert Op_h({{\tilde{\chi }}}_{_{{\mathcal {T}}_j}})P_{_{E}}u\Vert _{_{\!L^2(M)}}\bigg ) \nonumber \\&+ Q^{A,\psi }_{E,h}(C, C_{_{N}}, u). \end{aligned}$$
(4.6)

Here, \(\mathcal {I}_{_{E}}(h):=\{j \in \mathcal {J}_{_{E}}(h): {\mathcal {T}}_j \cap {\hbox {MS}}_{\textrm{h}}(A) \cap \Lambda _{ \Sigma _{_{E}}^H}^\tau (R(h)/2) \ne \emptyset \}\), \(\psi \in S_\delta \cap C^\infty _c(T^*M;[0,1])\) is any symbol with \({{\,\textrm{supp}\,}}\psi \subset \big (\Lambda ^\tau _{\Sigma ^H_{_{E}}}(2h^\delta )\big )^c\), and for each \(j \in \mathcal {J}_{_{E}}(h)\) we let \({\tilde{\chi }}_{_{{\mathcal {T}}_j}}\) be any symbol in \(S_\delta (T^*M;[0,1])\cap C^\infty _c(T^*M;[0,1])\) such that \({\tilde{\chi }}_{_{{\mathcal {T}}_j}}\equiv 1 \) on \({{\,\textrm{supp}\,}}\chi _{_{{\mathcal {T}}_j}}\) and \({{\,\textrm{supp}\,}}{\tilde{\chi }}_{_{{\mathcal {T}}_j}}\subset {\mathcal {T}}_j.\) In addition, if \({\widetilde{\mathcal {W}}}\subset \Psi _0^{\ell }(M)\), then \(C_{_{0}}=C_{_{0}}(n,k,{\mathfrak {I}}_{_{\!0}},\mathcal {V},\widetilde{\mathcal {W}})\).

Proof

First, we prove the statement for the case \(A=I\). Note that in this case the sets \(\mathcal {W}\) and \(\widetilde{\mathcal {W}}\) play no role. The result for \(A=I\) is a direct combination of the estimate in [11, (3.16)] and [11, Proposition 3.2]. We recall the estimate [11, (3.16)] with \(w\equiv 1\) here:

(4.7)

In (4.7), \(C_{n,k}>0\) is a constant depending only on n and k, and \(\beta _\delta :T^*H \rightarrow {{\mathbb {R}}}\) is a localizer to near conormal directions defined by \(\beta _\delta (x', \xi ')=\chi \big (h^{-\delta }|\xi '|_{_{H}} \big )\) where \(\chi \in C^\infty _0({\mathbb {R}};[0,1])\) is a smooth cut-off with \(\chi (t)=1\) for \(t \le \tfrac{1}{2}\) and \(\chi (t)=0\) for \(t \ge 1\).

Indeed, [11, Proposition 3.2] yields the existence of \(\tau _0, R_0, h_0>0\) as claimed, and the estimate [11, (3.16)] yields the same bound as above, but with three modifications.

To obtain the desired estimate, observe that the constant \(C_{_{0}}=C_{_{0}}(n,k, {\mathfrak {I}}_{_{\!0}})>0\) is the constant \(C_{n,k}\) divided by \({\mathfrak {I}}_{_{\!0}}\), because we absorb the \(|{{\textsf{H}}_p}r_H(\rho _j)|\) factors in (4.7). Second, in (4.7) the estimate is given for for \(\Big |\int _{H} Op_h(\beta _{\delta }) u\,d\sigma _{_{\!H}}\Big |\). It turns out that this estimate is all we need since [11, Proposition 3.2] yields that for every \(N>0\) there exists \(c_{_{N}}>0\) such that for all \(u \in \mathcal {D}'(M)\)

$$\begin{aligned} \Big |\int _{H} (1-Op_h(\beta _{\delta })) u\,d\sigma _{_{\!H}}\Big | \le c_{_{N}}h^N \Big (\Vert u\Vert _{L^2(H)}+ \Vert P_{_{E}}u\Vert _{_{\!H_{{\text {scl}}}^{\!\!\frac{k-2m+1}{2}}\!\!\!(M)}}\Big ).\qquad \end{aligned}$$
(4.8)

The third modification is that in (4.7) the first error term is \(Ch^{-\frac{1}{2}-\delta } \big \Vert P_{_{E}}u\big \Vert {_{\!H_{{\text {scl}}}^{\!\!\frac{k-2m+1}{2}}\!\!\!(M)}}\) instead of \(Ch^{-\frac{1}{2}-\delta }\big \Vert \big (1-\psi \big (\tfrac{P_{_{E}}}{h^\delta }\big )\big )P_{_{E}}u\big \Vert {_{\!H_{{\text {scl}}}^{\!\!\frac{k-2m+1}{2}}\!\!\!(M)}}\). The operator \(\big (1-\psi \big (\tfrac{P_{_{E}}}{h^\delta }\big )\big )\) can be added since the error term is a consequence of the application of an elliptic parametrix applied to an operator supported away from \(P_{_{E}}=0\), in particular of the bound in [11, (3.10)], which is for \(Op_h(\chi ) u\) where \(\chi \) is supported in \(\{(x, \xi ):\; |p_{_{E}}(x, \xi )|\ge \tfrac{1}{3}h^\delta \}\). One then uses \({{\,\textrm{supp}\,}}\chi \subset {{\,\textrm{supp}\,}}\big (1-\psi \big (\tfrac{p_{_{E}}}{h^\delta }\big ) \big )\).

We note that the desired bound holds for every \(\delta \)-partition \(\{\chi _{_{{\mathcal {T}}_j}}\}_{j\in \mathcal {J}_{_{E}}(h)}\subset {\mathcal {V}}\) associated to \(\{{\mathcal {T}}_j\}_{j\in \mathcal {J}_{_{E}}(h)}\), since the constants \(C, C_{_{N}}, h_0\) provided by [11, Proposition 3.5] are uniform for \(\chi _{_{{\mathcal {T}}_j}}\) in bounded subsets of \(S_\delta \).

Given \(\varepsilon _0>0\) we note that the statement holds for every \(E \in [a-\varepsilon _0, b+\varepsilon _0]\) since the constants \(C, C_{_{N}}, h_0\) provided by [11, Proposition 3.5] depend on \(P_{_{E}}\) only through P. Therefore, given \(K>0\), the statement for \(A=I\) holds for \(E \in [a-Kh, b+Kh]\) provided \(h_0\) depends on K.

We now treat the case \(A \ne I\). Let \(\mathcal {V},\mathcal {W}, \widetilde{\mathcal {W}}\), and \(\{B_{_{E}}\}_{E\in [a-\varepsilon _0, b+\varepsilon _0]}\) be as in the assumptions. Let \(E\in [a-\varepsilon _0, b+\varepsilon _0]\). Let \(X\in \Psi _\delta (M)\) with \({\hbox {MS}}_{\textrm{h}}(I-X)\cap \Lambda _{ \Sigma _{_{E}}^H}^\tau (\tfrac{1}{3}R(h))=\emptyset \), \({\hbox {MS}}_{\textrm{h}}(X)\subset \Lambda _{ \Sigma _{_{E}}^H}^{\tau _0+{\varepsilon _0}}(\tfrac{1}{2}R(h))\) and \(B_{_{E}}[P,X]\in \Psi _\delta (M)\). Then, for all \(N>0\) there is \(C_{_{N}}>0\) depending on \(\mathcal {V}\)

$$\begin{aligned} \Big |\int _H (I-X)Aud\sigma _{_{\!H}}\Big |\le C_{_{N}}h^N, \end{aligned}$$

so we may replace A by XA and assume \({\hbox {MS}}_{\textrm{h}}(A)\subset \Lambda _{ \Sigma _{_{E}}}^{{\tau _0+{\varepsilon _0}}}(R(h)/2)\) from now on. Since the estimate holds when \(A=I\), there exist \(C_{_{0}}=C_{_{0}}(n,k,{\mathfrak {I}}_{_{\!0}})\), \(C>0\), and for all \(K>0\) there is \(h_0>0\) such that for all \(N>0\) there exists \(C_{_{N}}>0\) with the following properties. For all \(u\in \mathcal {D}'(M)\), \(0<h<h_0\), \(E\in [a-Kh,b+Kh]\), and every \(\delta \)-partition \(\{\chi _{_{{\mathcal {T}}_j}}\}_{j\in \mathcal {J}_{_{E}}(h)}\subset {\mathcal {V}}\) associated to \(\{{\mathcal {T}}_j\}_{j\in \mathcal {J}_{_{E}}(h)}\), the bound in (4.6) holds with I in place of A, and with Au in place of u:

$$\begin{aligned}&h^{\frac{k-1}{2}}\Big |\int _{H}Au\,d\sigma _{_{\!H}}\Big |\\&\quad \le C_{_{0}}{R(h)^{\frac{n-1}{2}}} \sum _{j \in \mathcal {I}_{_{E}}(h)} \bigg (\frac{\Vert Op_h({{\tilde{\chi }}}_{_{{\mathcal {T}}_j}})Au\Vert }{\tau ^{\frac{1}{2}}}+Ch^{-1}\Vert Op_h({{\tilde{\chi }}}_{_{{\mathcal {T}}_j}})P_{_{E}}Au\Vert \bigg ) \nonumber \\&+ Q^{I,\psi }_{E,h}(C, C_{_{N}}, Au). \end{aligned}$$

We may sum over \(j \in \mathcal {I}_{_{E}}(h)\) instead of \(j \in \mathcal {J}_{_{E}}(h)\) since \({\hbox {MS}}_{\textrm{h}}(A){\cap \Lambda _{ \Sigma _{_{E}}^H}^\tau (\tfrac{1}{2}R(h))}\subset \cup _{j \in \mathcal {I}_{_{E}}(h)} {\mathcal {T}}_j\).

Next, we explain how to write u in place of Au in each of the terms of the sum over \(j \in \mathcal {I}_{_{E}}(h)\) in (4.6). To replace the term \(\Vert Op_h({ \chi }_{_{{\mathcal {T}}_j}})Au\Vert _{_{\!L^2(M)}}\) with \(\Vert Op_h({{\tilde{\chi }}}_{_{{\mathcal {T}}_j}})u\Vert _{_{\!L^2(M)}}\), we use \({\hbox {MS}}_{\textrm{h}}(Op_h(\chi _{_{{\mathcal {T}}_j}})A) \subset {\text {Ell}}(Op_h({{\tilde{\chi }}}_{_{{\mathcal {T}}_j}}))\) and apply the elliptic parametrix construction to find \(F_1\in \Psi _\delta (M)\) with

$$\begin{aligned} Op_h(\chi _{_{{\mathcal {T}}_j}})A= F_1Op_h({\tilde{\chi }}_{_{{\mathcal {T}}_j}}). \end{aligned}$$
(4.9)

Next, to replace the term \(\Vert Op_h({ \chi }_{_{{\mathcal {T}}_j}})P_{_{E}}Au\Vert _{_{\!L^2(M)}}\) with \(\Vert Op_h({{\tilde{\chi }}}_{_{{\mathcal {T}}_j}})P_{_{E}}u\Vert _{_{\!L^2(M)}}\), we decompose

$$\begin{aligned} Op_h(\chi _{_{{\mathcal {T}}_j}})P_{_{E}}A=Op_h(\chi _{_{{\mathcal {T}}_j}})[P_{_{E}},A]+Op_h(\chi _{_{{\mathcal {T}}_j}})AP_{_{E}} \end{aligned}$$

for each \(j \in \mathcal {I}_{_{E}}(h)\), and apply the elliptic parametrix construction and find \( F_2\in \Psi _\delta (M)\) with

$$\begin{aligned} h^{-1}Op_h(\chi _{_{{\mathcal {T}}_j}})[P_{_{E}},A]=F_2Op_h({\tilde{\chi }}_{_{{\mathcal {T}}_j}}). \end{aligned}$$
(4.10)

To do this we used the assumptions: \(B_{_{E}}\) is microlocally the identity on \(\Lambda _{{ \Sigma _{_{E}}^H}}^{\tau _0+{\varepsilon _0}}({2R(h)})\), \({\hbox {MS}}_{\textrm{h}}({A}) \subset \Lambda _{{ \Sigma _{_{E}}^H}}^{\tau _0+{\varepsilon _0}}(\tfrac{1}{2}R(h))\), and A is such that \(B_{_{{E}}}\tfrac{1}{h}[P,A]\in \mathcal {W}\subset \Psi _\delta (M)\). This allows us to apply the parametrix construction to \(Op_h(\chi _{_{{\mathcal {T}}_j}})B_{_{E}}\tfrac{1}{h}[P_{_{E}},A]\).

Using (4.9) and (4.10), we may modify \(C_{_{0}}\), and having it now also depend on A, \(\mathcal {V}\) and \(\mathcal {W}\), to obtain the claim. Note that if \(A \in \Psi _0^\infty (M)\), then \(\tfrac{1}{h}[P_{_{E}},A] \in \Psi _\delta ^\infty (M)\) and so we may apply the elliptic parametrix construction to obtain (4.10) without the need of introducing the operator \(B_{_{{E}}}\) or the set \(\mathcal {W}\). In this case, we have \(C_{_{0}}=C_{_{0}}(n,k,{\mathfrak {I}}_{_{\!0}},\mathcal {V}, {\widetilde{\mathcal {W}}})\) as claimed. \(\square \)

Definition 4.2

Low density tubes Let \(\{{\mathcal {T}}_j\}_{j\in \mathcal {J}(h)}\) be a cover by tubes of \( \Sigma _{_{[a,b]}}^H\) and \(0<\delta <\tfrac{1}{2}\). Let \({{\mathcal {G}}}(h)\subset \mathcal {J}(h)\) and for each \(j \in {{{\mathcal {G}}}(h)}\) let \(1<t_j(E,h)\le T_j(E,h)\), where \(h>0\) and \(E\in {\mathbb {R}}\).

We say \(\{{\mathcal {T}}_j\}_{j \in {{\mathcal {G}}}(h)}\) has \(\{(t_j, T_j)\}_{j\in {{\mathcal {G}}}(h)}\) density on [ab] if the following holds. For all \(\mathcal {V}\subset S_\delta \) bounded, \(K>0\) there is \(h_0>0\) such that for all \(0<h<h_0\), \(E\in [a-Kh,b+Kh]\), every \(\delta \)-partition \(\{\chi _j\}_{j\in \mathcal {G}_{_{E}}(h)}\subset \mathcal {V}\) associated to \(\{{\mathcal {T}}_j\}_{j \in \mathcal {G}_{_{E}}(h)}\), and all \(u \in \mathcal {D}'(M)\),

$$\begin{aligned} \sum _{j\in {\mathcal {G}_{_{E}}(h)}}\Vert Op_h(\chi _j)u\Vert _{_{\!L^2(M)}}^2\frac{T_{j}(E,h)}{t_j(E,h)}\le 4\Vert u\Vert _{_{\!L^2(M)}}^2+ 4\max _{j \in {\mathcal {G}_{_{E}}(h)}}\frac{T_{j}(E,h)^2}{h^2} \Vert P_{_{E}}u\Vert ^2_{_{\!L^2(M)}}, \end{aligned}$$

where \(\mathcal {G}_{_{E}}(h)={{\mathcal {G}}}(h)\cap \mathcal {J}_{_{E}}(h)\).

The statement of [11, Lemma 4.1] can be reformulated as: if a collection of families of tubes is non self-looping for different times, then the tubes have a low density dictated by those times. More precisely, the following lemma is a restatement of [11, Lemma 4.1].

Lemma 4.3

Let \(R_0,\) \(\tau _0\), \(\delta \), R(h), \(\tau \), and \(\{{\mathcal {T}}_j\}_{j\in \mathcal {J}(h)}\) be as in Lemma 4.1. Let \(0< \alpha < 1-\limsup _{h\rightarrow 0^+} 2\tfrac{\log R(h)}{\log h}\) and \(K>0\). There exists \(h_0>0\) such that the following holds. Let \(0<h<h_0\), \(E\in [a-Kh, b+Kh]\), and \(\mathcal {G}_{_{E}}(h)\subset \mathcal {J}_{_{E}}(h)\) with \({\mathcal {G}_{_{E}}(h)=\sqcup _{\ell \in \mathcal {L}_{_{E}}(h)} \mathcal {G}_{_{E,\ell }}}(h)\). For every \(\ell \in {\mathcal {L}}_{_{E}}(h)\) suppose \(t_\ell (E,h)>0\), \(0<{T_\ell (E,h)}\le 2\alpha \, T_e(h),\) and

$$\begin{aligned} \bigcup _{j\in {{{\mathcal {G}}}_{_{E,\ell }}(h)}}{\mathcal {T}}_j \qquad \text {is} \;\; [t_\ell ,T_\ell ]\;\;\text { non-self looping for every}\;\; \ell \in \mathcal {L}_{_{E}}(h). \end{aligned}$$

Then, \(\{{\mathcal {T}}_j\}_{j \in {{\mathcal {G}}}(h)}\) has \(\{(t_j, T_j)\}_{j\in {{\mathcal {G}}}(h)}\) density on [ab], where for \(0<h<h_0\), \(j \in \mathcal {J}(h)\), and \(E\in [a-Kh, b+Kh]\), we set \((t_j(E,h), T_j(E,h)):=(t_\ell (E,h), T_\ell (E,h))\) whenever \(j \in \mathcal {G}_{_{E,\ell }}(h)\).

We note that the statement of [11, Lemma 4.1] does not provide the requisite uniformity for \(E\in [a-Kh,b+Kh]\); however, this follows from the same argument.

Our next estimate shows that if a family of tubes has low density, then averages of a quasimode over H can be controlled in terms of the density times.

Lemma 4.4

Let \(R_0,\) \(\tau _0\), \(\delta \), R(h), \(\tau \), \(\{{\mathcal {T}}_j\}_{j\in \mathcal {J}(h)}\), \(\mathcal {W}\), \(\widetilde{\mathcal {W}}\), and \(\psi \) be as in Lemma 4.1. Then, there exist \(C_{_{0}}=C_{_{0}}(n,k,p,{\mathfrak {I}}_{_{\!0}},\mathcal {W})\) and \(C>0\), and for all \(N>0\), \(K>0\) there are \(h_0>0\) and \(C_{_{N}}>0\), such that the following holds.

Suppose that for all \(0<h<h_0\) and \(E\in [a-Kh, b+Kh]\) there exists \(\mathcal {G}_{_{E}}(h)\subset \mathcal {J}_{_{E}}(h)\) with \({\mathcal {G}_{_{E}}(h)=\sqcup _{\ell \in {\mathcal {L}}_{E}(h)} \mathcal {G}_{_{E,\ell }}}(h)\), such that for every \(\ell \in {\mathcal {L}}_{_{E}}(h)\) there exist \(t_\ell =t_\ell (E,h)>0\) and \({T_\ell =T_\ell (E,h)}>0\) so that, with \((t_j, T_j):=(t_\ell , T_\ell )\) for every \(j \in \mathcal {G}_{_{E,\ell }}(h)\), then

$$\begin{aligned}{} & {} (1)\{{\mathcal {T}}_j\}_{j \in {{\mathcal {G}}}(h)}\text { has }\{(t_j, T_j)\}_{j\in {{\mathcal {G}}}(h)}\text { density on }\,[a,b],\\ {}{} & {} (2) {\hbox {MS}}_{\textrm{h}}(A)\cap \Lambda _{ \Sigma _{_{E}}^H}^\tau (\tfrac{1}{2}R(h))\subset \, {\bigcup _{j \in \mathcal {G}_{_{E}}(h)}{\mathcal {T}}_j}. \end{aligned}$$

Then, for all \(u\in \mathcal {D}'(M)\), \(0<h<h_0\), \(E\in [a-Kh,b+Kh]\), and every \(A\in {\widetilde{\mathcal {W}}}\) with \(B_{_{E}}\frac{1}{h}[P,A]\in \mathcal {W}\),

$$\begin{aligned}&h^{\frac{k-1}{2}}\Big |\int _{H}Au\,d\sigma _{_{\!H}}\Big | \\&\quad \le C_{_{0}}R(h)^{\frac{n-1}{2}} \sum _{\ell \in \mathcal {L}_{_{E}}(h)}\!\!\bigg (\frac{(|{{\mathcal {G}}}_{_{E,\ell }}|t_\ell )^{\frac{1}{2}}}{\tau ^{\frac{1}{2}} T_\ell ^{\frac{1}{2}}}\Vert u\Vert _{_{\!L^2(M)}}+\frac{(|{{\mathcal {G}}}_{_{E,\ell }}| t_\ell T_\ell )^{\frac{1}{2}}}{h}\Vert P_{_{E}}u\Vert _{_{\!L^2(M)}}\!\!\bigg )\\&+ Q^{A,\psi }_{E,h}(C, C_{_{N}}, u). \end{aligned}$$

In addition, if \({\widetilde{\mathcal {W}} \subset } \Psi _0^\infty (M)\), the estimate holds with \(C_{_{0}}=C_{_{0}}(n,k,p,{\mathfrak {I}}_{_{\!0}},{\widetilde{\mathcal {W}}})\).

Proof

Let \(\mathcal {V}\) a bounded subset of \(S_\delta (T^*M;[0,1])\). By Lemma 4.1 there exist \(C_{_{0}}=C_{_{0}}(n,k,{\mathfrak {I}}_{_{\!0}},\mathcal {V},\mathcal {W})\), \(C>0\), and \(h_0>0\), such that for all \(N>0\) there exist \(C_{_{N}}>0\), with the following properties. For all \(u\in \mathcal {D}'(M)\), \(K>0\), \(0<h<h_0\), \(E\in [a-Kh,b+Kh]\), and every \(\delta \)-partition \(\{\chi _{_{{\mathcal {T}}_j}}\}_{j\in \mathcal {J}_{_{E}}(h)}\subset {\mathcal {V}}\) associated to \(\{{\mathcal {T}}_j\}_{j\in \mathcal {J}_{_{E}}(h)}\),

$$\begin{aligned}&h^{\frac{k-1}{2}}\Big |\int _{H}Auds{H}\Big | \\&\quad \le C_{_{0}}{R(h)^{\frac{n-1}{2}}} \!\!\!\sum _{j \in \mathcal {I}_{_{E}}(h)} \!\!\!\bigg (\frac{\Vert Op_h({{\tilde{\chi }}}_{_{{\mathcal {T}}_j}})u\Vert _{L^2}}{\tau ^{\frac{1}{2}}}+\frac{C}{h}\Vert Op_h({{\tilde{\chi }}}_{_{{\mathcal {T}}_j}})P_{_{E}}u\Vert _{L^2}\!\!\bigg )\\&\qquad +Q^{A,\psi }_{E,h}(C,\! C_{_{N}}, \!u), \end{aligned}$$

where \(\mathcal {I}_{_{E}}(h):=\bigcup _{\ell \in {\mathcal {L}}_{h,E}}{\mathcal {G}_{_{E,\ell }}}\). Note that if \(A\in \Psi _0^\infty (M)\), then the estimate holds with \(C_{_{0}}=C_{_{0}}(n,k,p, {\mathfrak {I}}_{_{\!0}}, \mathcal {V}, {\widetilde{\mathcal {W}}})\). Next, note that

$$\begin{aligned} \sum _{j\in \mathcal {I}_{_{E}}(h)}\Vert Op_h({\tilde{\chi }}_{_{{\mathcal {T}}_j}})P_{_{E}}u\Vert \le |\mathcal {J}_{_{E}}(h)|^{\frac{1}{2}} \Big (\sum _{j\in \mathcal {J}_{_{E}}(h)}\Vert Op_h({\tilde{\chi }}_{_{{\mathcal {T}}_j}})P_{_{E}}u\Vert ^2\Big )^{\frac{1}{2}}, \end{aligned}$$

and so, since \(|\mathcal {J}_{_{E}}(h)| \le C_n{{\,\textrm{vol}\,}}({ \Sigma _{_{E}}^H})R(h)^{{1-n}}\) for some \(C_n>0\), we have, after adjusting \(C>0\), that for all \(0<h<h_0\)

$$\begin{aligned} h^{\frac{k-1}{2}}\Big |\int _{H}Au\,d\sigma _{_{\!H}}\Big |&\le C_{_{0}}\frac{R(h)^{\frac{n-1}{2}}}{\tau ^{\frac{1}{2}}}\!\!\sum _{j\in \mathcal {I}_{_{E}}(h)} \!\Vert Op_h({\tilde{\chi }}_{_{{\mathcal {T}}_j}}) u\Vert _{_{\!L^2(M)}}\!\!\nonumber \\ {}&\quad + \frac{C}{h}\Vert P_{_{E}}u\Vert _{_{\!L^2(M)}}\!+ Q^{A,\psi }_{E,h}(C,\! C_{_{N}},\! u). \end{aligned}$$
(4.11)

Since we are working with a \(({{\mathfrak {D}}_n},\tau ,R(h))\)-good cover, we split each \({\mathcal {G}_{_{E,\ell }}}\) into \({\mathfrak {D}}_n\) families \(\{\mathcal {G}_{_{E,\ell , i}}\}_{i=1}^{{\mathfrak {D}}_n}\) of disjoint tubes. Note that

$$\begin{aligned} \sum _{j\in \mathcal {I}_{_{E}}(h)}\Vert Op_h({{\tilde{\chi }}}_j)u\Vert _{_{\!L^2(M)}}\le \sum _{\ell \in {\mathcal {L}}}{\sum _{i=1}^{{\mathfrak {D}}_n} \sum _{j\in \mathcal {G}_{_{E,\ell ,i}}}}\Vert Op_h({{\tilde{\chi }}}_j)u\Vert _{_{\!L^2(M)}}. \end{aligned}$$
(4.12)

Next, since \(\{{\mathcal {T}}_j\}_{j \in {{\mathcal {G}}}(h)}\) has \(\{(t_j, T_j)\}_{j\in {{\mathcal {G}}}(h)}\) density on [ab], after possibly shrinking \(h_0\) (depending on the \(S_\delta \) bounds for \({\tilde{\chi }}_j\) and \(K>0\)), Cauchy-Schwarz yields that for all \(0<h<h_0\)

$$\begin{aligned} {\sum _{j \in \mathcal {G}_{_{E,\ell ,i}}}}\Vert Op_h({{\tilde{\chi }}}_j)u\Vert _{_{\!L^2(M)}}\le 2\Big (\frac{t_\ell |{\mathcal {G}_{_{E,\ell }}}|}{T_\ell }\Big )^{\frac{1}{2}}\Big (\Vert u\Vert _{_{\!L^2(M)}}^2+ \frac{T_{\ell }^2}{h^2}\, \Vert P_{_{E}}u\Vert ^2_{_{\!L^2(M)}}\Big )^{\frac{1}{2}}. \end{aligned}$$
(4.13)

The result follows from combining (4.13) and (4.12), and feeding this to (4.11). Note that \(C_{_{0}}\) needs to be modified, but only in a way that depends on n via \({\mathfrak {D}}_n\). \(\square \)

We also need the following basic estimate for averages over submanifolds to control averages of \(u={\textbf{1}}_{(-\infty , s]}(P)\) when s is large.

Lemma 4.5

Suppose \(H\subset M\) is a submanifold of codimension k and \(P\in \Psi ^m(M)\), with \(m>0\), is such that there exists \(C>0\) for which

$$\begin{aligned} |\sigma (P)(x,\xi )|\ge |\xi |^m/C,\qquad (x,\xi )\in N^*H,\qquad |\xi |\ge C. \end{aligned}$$

Let \(\psi \in S^0(T^*M; [0,1])\) with \(\psi \equiv 1\) on \(N^*H\), and let \(\ell \in {\mathbb {R}}\). Let \(A\in \Psi _\delta ^{\ell }(M)\) and \(r>\frac{k+2\ell }{2m}\). Then, there are \(C_{_{0}}>0\) and \(h_{0}>0\) such that for all \(N>0\) there is \(C_{_{N}}>0\) satisfying

$$\begin{aligned} h^{\frac{k}{2}}\Big |\int _H \!\!\!Aud\sigma _{_{H}}\Big |\le&C_{_{0}}\Big ( \Vert Op_h(\psi )u\Vert _{{_{\!L^2(M)}}}+\Vert Op_h(\psi )P_{_{E}}^ru\Vert _{{_{\!L^2(M)}}}\Big )\nonumber \\&+C_{_{N}}h^N\Vert u\Vert _{H_{scl }^{-N}(M)},\qquad 0<h<h_0. \end{aligned}$$

Proof

Let \({{\tilde{\psi }}} \in S^0(T^*M; [0,1])\) with \({\tilde{\psi }}\equiv 1\) on \(N^*H\), \({{\,\textrm{supp}\,}}{\tilde{\psi }}\subset \{\psi \equiv 1\}\), and such that

$$\begin{aligned} |\sigma (P_{_{E}})(x,\xi )|\ge \tfrac{1}{C}|\xi |^m,\qquad (x,\xi )\in {{\,\textrm{supp}\,}}{\tilde{\psi }},\qquad |\xi |\ge C. \end{aligned}$$

Then, since \({\hbox {WF}}_{\textrm{h}}(\delta _{H})=N^*\!H\), for any \(N>0\) there is \(C_{_{N}}>0\) such that

$$\begin{aligned} \Big |\int _HAOp_h(1-{\tilde{\psi }})ud\sigma _{_{H}}\Big |\le C_{_{N}}h^N\Vert u\Vert _{H_{scl }^{-N}(M)}. \end{aligned}$$
(4.14)

Next, by the Sobolev embedding theorem, for any \(\varepsilon >0\) there exists \(C_{_{0}}>0\) such that

$$\begin{aligned} \Big | \int _{H}AOp_h({\tilde{\psi }})u d\sigma _{_{H}}\Big |&\le C_{_{0}}h^{-\frac{k}{2}}\Vert Op_h({\tilde{\psi }})u\Vert _{H_{scl }^{\frac{k}{2}+\varepsilon +\ell }(M)}. \end{aligned}$$

Taking r with \(rm>\frac{k}{2}+\ell \) and using an elliptic parametrix, for any \(N>0\) there is \(C_{_{N}}>0\) with

$$\begin{aligned} h^{\frac{k}{2}}\Big | \int _{H}AOp_h(\psi )u d\sigma _{_{H}}\Big |&\le C_{_{0}}\Vert Op_h({\tilde{\psi }})u\Vert _{H_{scl }^{rm}(M)} \nonumber \\&\le C_{_{0}} \big (\Vert Op_h(\psi )u\Vert _{{_{\!L^2(M)}}}+\Vert Op_h(\psi ) P_{_{E}}^ru\Vert _{{_{\!L^2(M)}}}\big )\nonumber \\&\quad +C_{_{N}}h^N\Vert u\Vert _{H_{scl }^{-N}(M)}. \end{aligned}$$
(4.15)

Indeed, this follows from letting \(\chi \in S^0(T^*M;[0,1])\) so that \( |\sigma (P_{_{E}})(x,\xi )|\ge \tfrac{1}{C}|\xi |^m\) in the support of \({{\tilde{\psi }}} (1-\chi )\), and then using the elliptic parametrix construction to find \(F_1, F_2 \in \Psi ^0(M)\) such that

$$\begin{aligned} \langle hD\rangle ^{rm} Op_h({{\tilde{\psi }}})(1-Op_h(\chi ))=F_1 Op_h(\psi )P_{_{E}}^r + O(h^\infty )_{\Psi ^{-\infty }},\\ \langle hD\rangle ^{rm} Op_h({{\tilde{\psi }}})Op_h(\chi )=F_2 Op_h(\psi ) + O(h^\infty )_{\Psi ^{-\infty }}. \end{aligned}$$

Combining with (4.14) and (4.15) completes the proof. \(\square \)

5 Lipschitz scale for spectral projectors

In this section we estimate the scale at which averages of the spectral projector behave like Lipschitz functions of the spectral parameter, and use this to approximate \(\Pi _h\) using \(\rho _{h,T(h)}*\Pi _h\).

Throughout this section we assume \(H_1,H_2 \subset M\) are two smooth submanifolds of co-dimension \(k_1\) and \(k_2\) respectively. The goal for this section is to prove the following proposition.

Proposition 5.1

Suppose \(a, b\in {\mathbb {R}}\) such that \(H_1, H_2\) are uniformly conormally transverse for p in the window [ab]. Let \(\tau _0, R_0\) be as in Lemma 4.1. Let \(0<\tau <\tau _0\) and \(0<\delta <\tfrac{1}{2}\). For \(i=1,2\), let \(\textbf{T}_i\) be sub-logarithmic resolution functions  with \(\Omega (\textbf{T}_i)\Lambda <1-2\delta \) and suppose \(H_i\) is \({\textbf{T}_i}\) non-recurrent in the window [ab] via \(\tau \)-coverings with constant \(C_{_{\!{\text {nr}}}}^i\).

Let \(A_1,A_2\in \Psi ^\infty (M)\), \({K>0}\), \( R(h)\ge {h^\delta }\), and \(\textbf{T}:=\sqrt{\textbf{T}_1\textbf{T}_2}\). Then, there exist \(h_0>0\) and

$$\begin{aligned} C_{_{0}}=C_{_{0}}(n, k_1, k_2, {\mathfrak {I}}_{_{\!0}}^1,{\mathfrak {I}}_{_{\!0}}^2, A_1, A_2, C_{_{\!{\text {nr}}}}^1, C_{_{\!{\text {nr}}}}^2)>0, \end{aligned}$$

such that for all \(0<h\le h_0\) and \(E\in [a-Kh,b+Kh]\),

$$\begin{aligned} \Big |\Pi _{_{H_1,H_2}}^{^{A_1,A_2}}(E)-\rho _{_{h,T_{\max }(h)}}*\Pi _{_{H_1,H_2}}^{^{A_1,A_2}}(E)\Big |\le C_{_{0}}h^{\frac{2-k_1-k_2}{2}}\Big /\textbf{T}(R(h)). \end{aligned}$$

Remark 5.2

To ease notation, throughout this section we write \(T_i(h):=\textbf{T}_i(R(h))\), \(T(h):=\textbf{T}(R(h))\), and \({T_{\max }(h):=\max (\textbf{T}_1(R(h),\textbf{T}_2(R(h))))}.\)

Proof

We split the proof into Lemmas 5.35.4, and 5.5 below. Lemmas 5.4 and 5.5 show that there exist \(C_{_{0}}=C_{_{0}}(n, k_1, k_2, {\mathfrak {I}}_{_{\!0}}^1,{\mathfrak {I}}_{_{\!0}}^2, A_1, A_2, C_{_{\!{\text {nr}}}}^1, C_{_{\!{\text {nr}}}}^2)>0\), \(C_1>0\), and \(h_0>0\) such that \(w_h(E):=\Pi _{_{H_1,H_2}}^{^{A_1,A_2}}(E)\) satisfies the hypotheses of Lemma 5.3 with \(I_h:= [a-Kh,b+Kh]\), \(\rho _h:=\rho _{_{h,T_{\max }(h)}}\), \(\sigma _h:= T_{\max }(h)/h\),

$$\begin{aligned} L_h:=C_{_{0}}h^{\frac{2-k_1-k_2}{2}}\Big /T(h)\qquad \text {and}\qquad B_h:= C_{_{1}}{h^{-\frac{k_1+k_2}{2}}}, \end{aligned}$$

and \(0<h<h_0\). Next, let \(\{K_j\}_{j=1}^\infty \subset {\mathbb {R}}_+\) be given by the choice of \(\rho \) in (1.16). Since \(\Big \langle \frac{T_1(h)s}{h}\Big \rangle ^{\frac{1}{2}}\Big \langle \frac{T_2(h) s}{h}\Big \rangle ^{\frac{1}{2}}\le \langle \sigma _{_{h}} s\rangle \) for all \(s\in {\mathbb {R}}\), Lemma 5.3 yields that there exists \(C_{_{\rho }}>0\) and for all \(N>0\) there exists \(C_{_{N}}>0\) such that

$$\begin{aligned} \Big |\Pi _{_{H_1,H_2}}^{^{A_1,A_2}}(E)-\rho _{_{h,T(h)}}*\Pi _{_{H_1,H_2}}^{^{A_1,A_2}}(E)\Big |\le C_{_{\rho }} C_{_{0}} \frac{h^{\frac{2-k_1-k_2}{2}}}{T(h)}+ C_{_{N}} C_{_{1}}{h^{-\frac{k_1+k_2}{2}}} \Big (\frac{h}{T_{\max }(h)}\Big )^N, \end{aligned}$$

for all \(0<h<h_0\). This completes the proof after choosing \(h_0\) small enough. \(\square \)

We now present the lemmas used in the proof of Proposition 5.1. The first shows that if a family of functions \(\{w_h\}_h\) is Lipstchitz at scale \(\sigma _h^{-1}\) with (at most) polynomial growth at infinity, then the family can be well approximated by its convolution \(\rho _{h} *w_h\) where \(\{\rho _h\}_h\) is a family of Schwartz functions

Lemma 5.3

Let \(\{K_j\}_{j={0}}^\infty \subset {\mathbb {R}}_+\). Then, there exists \(C>0\) and for all \(N_0\in {\mathbb {R}}\), \(N>0\) there exists \(C_{_{N}}>0\), such that the following holds. Let \(\{\rho _{h}\}_{h>0}\subset {\mathcal {S}}({\mathbb {R}})\) be a family of functions and \(\{\sigma _h\}_{h>0}\subset {\mathbb {R}}_+\) such that for all \(j\ge 1\) and \(h>0\),

$$\begin{aligned} |\rho _h(s)|\le {\sigma _h K_j}\,{\langle \sigma _h s\rangle ^{-j}}\qquad \text {for all}\; s \in {\mathbb {R}}. \end{aligned}$$

Let \(\{L_h\}_{h>0}\subset {\mathbb {R}}_{+}\), \(\{B_h\}_{h>0}\subset {\mathbb {R}}_{+}\), \(\{w_h:{\mathbb {R}}\rightarrow {\mathbb {R}}\}_{h>0}\), \(I_h\subset [-K_0,K_0]\), \(h_0>0\) and \(\varepsilon _0>0\), be so that for all \(0<h<h_0\)

  • \(|w_h(t-s)-w_h(t)|\le L_h\langle \sigma _h \,s\rangle \) for all \(t\in I_h\) and \({|s|\le \varepsilon _0}\),

  • \(|w_h(s)|\le B_h\langle s\rangle ^{N_0}\) for all \(s\in {\mathbb {R}}\).

Then, for all \(0<h<h_0\) and \(t\in I_h\)

$$\begin{aligned} \Big |(\rho _{h} *w_h)(t)-w_h(t)\int _{{\mathbb {R}}} \rho _{h}(s) ds\Big |\le C L_h+ C_{_{N}} B_h \sigma _h^{-N}\varepsilon _0^{-N}. \end{aligned}$$

Proof

For all \(0<h<h_0\) and \(t\in I_h\)

$$\begin{aligned}&\Big |(\rho _h *w_h)(t)-w_h(t)\int _{{\mathbb {R}}} \rho _h(s)ds\Big | =\Big |\int _{{\mathbb {R}}} \rho _h(s)\big (w_h(t-s)-w_h(t)\big )ds\Big |\\&\le L_h\int _{|s|\le \varepsilon _0} |\rho _h(s)|\langle \sigma _h s\rangle ds +B_h\int _{|s|\ge \varepsilon _0}|\rho _h(s)|\Big ( \langle t-s\rangle ^{N_0}+\langle t\rangle ^{N_0}\Big )ds\\&\le L_h\int _{|s|\le \varepsilon _0}\!\!\! \sigma _h K_{3} \langle \sigma _h s\rangle ^{-2} ds +B_h\!\!\int _{|s|\ge \varepsilon _0} \!\!\!K_{_{N_0+2+N}}\sigma _h \langle \sigma _h s\rangle ^{-(N_0+2+N)}\\&\quad \Big ( \langle t-s\rangle ^{N_0}+\!\langle t\rangle ^{N_0}\Big )ds. \end{aligned}$$

The existence of C and \(C_{_{N}}\) follows from integrability of each term and the boundedness of \(I_h\). \(\square \)

The next lemma shows that the family of functions \(w_h(t)=\Pi _{_{H_1,H_2}}^{^{A_1,A_2}}(t)\) is Lipschitz at scales dictated by the non-recurrence times for \(H_1\) and \(H_2\).

Lemma 5.4

Suppose \(a,b\in {\mathbb {R}}\), \(\varepsilon _0>0\) are such that \(H_1\), \(H_2\) are conormally transverse for p in the window \([a-\varepsilon _0,b+\varepsilon _0]\). Let \(A_1\), \(A_2\), \(\tau _0\), \(R_0\), \(\tau \), \(\delta \), R(h), and \(\alpha \) be as in Proposition 5.1. Let \(C_{_{\!{\text {nr}}}}>\) and \(K>0\). Then, there exist \(h_0>0\) and

$$\begin{aligned} C_{_{0}}=C_{_{0}}(n, k_1, k_2, {\mathfrak {I}}_{_{\!0}}^1,{\mathfrak {I}}_{_{\!0}}^2, A_1, A_2, C_{_{\!{\text {nr}}}})>0 \end{aligned}$$

such that the following holds.

For \(i=1,2\), let \(\textbf{T}_i\) be a sub-logarithmic resolution function with \(\Omega (\textbf{T}_i)\Lambda <1-2\delta \). Suppose \(H_i\) is \(\textbf{T}_i\) non-recurrent in the window [ab] via \(\tau \)-coverings with constant \(C_{_{\!{\text {nr}}}}^i{\le C_{_{\!{\text {nr}}}}}\). Then for all \(0<h\le h_0\), \(|s|\le \varepsilon _0\), and \(t\in [a-Kh,b+Kh]\),

$$\begin{aligned} \Big |\Pi _{_{H_1,H_2}}^{^{A_1,A_2}}(t)-\Pi _{_{H_1,H_2}}^{^{A_1,A_2}}(t-s)\Big |\le C_{_{0}}\frac{h^{\frac{2-k_1-k_2}{2}}}{\sqrt{T_1(h)T_2(h)}}\Big \langle \frac{T_1(h)s}{h}\Big \rangle ^{\frac{1}{2}}\Big \langle \frac{T_2(h) s}{h}\Big \rangle ^{\frac{1}{2}}. \end{aligned}$$

Proof

We first assume the statement for \(|s| \le 2h\). Suppose \(s\ge 2h\). The case of \(s\le -2h\) being similar. Define \(k_0:=\lfloor \frac{s}{h}\rfloor \) and \(t_k:=t-s+kh\) for \(0\le k\le k_0-1,\) and \(t_k:=t\) for \(k=k_0\). Then

$$\begin{aligned} \Pi _{_{H_1,H_2}}^{^{A_1,A_2}}(t)-\Pi _{_{H_1,H_2}}^{^{A_1,A_2}}(t-s)=\sum _{k=0}^{k_0-1}\Pi _{_{H_1,H_2}}^{^{A_1,A_2}}(t_{k+1})-\Pi _{_{H_1,H_2}}^{^{A_1,A_2}}(t_{k}). \end{aligned}$$

Using \(|t_{k+1}-t_k|\le 2h\), and putting \(t=t_{k+1}\), \(s=t_{k+1}-t_k\), we apply the case \(|s|\le 2h\) with \(T_1=T_2=1\) for each term to obtain

$$\begin{aligned} \Big |\Pi _{_{H_1,H_2}}^{^{A_1,A_2}}(t)-\Pi _{_{H_1,H_2}}^{^{A_1,A_2}}(t-s)\Big |&\le C_{_{0}}k_0 h^{\frac{2-k_1-k_2}{2}}\le C_{_{0}}h^{\frac{2-k_1-k_2}{2}} |s/h|, \end{aligned}$$

and this proves the claim provided the statement holds for \(|s| \le 2h\).

We proceed to prove the statement for \(|s| \le 2h\). First, note that by (1.10) and Cauchy-Schwarz

$$\begin{aligned}&\Big |\Pi _{_{H_1,H_2}}^{^{A_1,A_2}}(t)-\Pi _{_{H_1,H_2}}^{^{A_1,A_2}}(t-s)\Big |^2\nonumber \\ {}&\le \sum _{t-s\le E_k\le t} \Big |\int _{H_1}A_1\phi _{_{E_k}}d\sigma _{_{\!H_1}}\Big |^2\,\,\cdot \,\, \sum _{t-s\le E_j\le t}\Big |\int _{H_2} A_2\phi _{_{E_j}} d\sigma _{_{\!H_2}}\Big |^2. \end{aligned}$$
(5.1)

Now, for each \(i=1,2,\)

$$\begin{aligned} \sum _{t-s\le E_j\le t}\Big |\int _{H_i}\!\!A_i\phi _{_{E_j}}d\sigma _{_{\!H_i}}\Big |^2{} & {} =\Vert \mathbb {1}_{[t-s,t]}(P)\, A_i^*\, \delta _{_{H_i}}\Vert ^2_{_{L^2(M)}}\nonumber \\ {}{} & {} =\!\!\!\!\!\sup _{\Vert w\Vert _{_{\!L^2(M)}}=1}\Big | \int _{H_i}\!\!A_i\mathbb {1}_{[t-s,t]}(P) w\,d\sigma _{_{\!H_i}}\Big |^2, \qquad \end{aligned}$$
(5.2)

where \(\delta _{_{H_i}}\) is the delta distribution at \(H_i\) and the last equality follows by duality.

We now use the non-recurrence assumption on \(H_1\) and \(H_2\). Since for each \(i=1,2\), the submanifold \(H_i\) is \(\textbf{T}_i\) non-recurrent in the window [ab] via \(\tau _0\)-coverings, there is \(h_0>0\) small enough depending on R(h), K so that for all \(0<h<h_0\) and \(t \in {[E-Kh,E+Kh]}\) there is a partition of indices \(\mathcal {J}_{_{t}}^i(h)=\cup _{\ell \in {\mathcal {L}}_{_{t}}^i(h)} {{\mathcal {G}}}_{_{t,{\ell }}}^i(h)\), and times \(\{T_\ell ^i(h)\}_{\ell \in {\mathcal {L}}_{_{t}}^i(h)}\), and \(\{t_\ell ^i(h)\}_{\ell \in {\mathcal {L}}_{_{t}}^i(h)}\) as in Definition 2.2.

Note that we have chosen \(h_0\) small enough so that \(\mathcal {J}_{_{E}}^i(h)\) is a \((\tau ,R(h))\) good covering of \(\Sigma ^{H_i}_t\) for \(t \in [E-Kh,E+Kh]\). In particular, for \(i=1, 2\) and \(t\in [E-Kh,\,E+Kh]\)

$$\begin{aligned}{} & {} R(h)^{\frac{n-1}{2}}\sum _{\ell \in \mathcal {L}_{_{E}}^i(h)} \frac{(|{{\mathcal {G}}}_{_{t,{\ell }}}^i|t_\ell ^i)^{\frac{1}{2}} }{{(T_\ell ^i)}^{\frac{1}{2}}} \le \frac{C_{_{\!{\text {nr}}}}^i}{T_i^{\frac{1}{2}}},\nonumber \\{} & {} R(h)^{\frac{n-1}{2}}\sum _{\ell \in \mathcal {L}_{_{E}}^i(h)} (|{{\mathcal {G}}}_{_{t,{\ell }}}^i|t_\ell ^i)^{\frac{1}{2}}(T_\ell ^i)^{\frac{1}{2}}\le C_{_{\!{\text {nr}}}}^i T_i^{\frac{1}{2}}. \end{aligned}$$
(5.3)

The first bound is condition (2) in Definition 2.2, and the second bound follows from the first one together with the \(T_\ell ^i \le T_i\) for all \(\ell \in \mathcal {L}^i_{h,E}\). Next, for \(\ell \in \mathcal {L}^i_{E}\) let

$$\begin{aligned} {{\tilde{T}}}_{\ell }^i(h)&:={\left\{ \begin{array}{ll}T_{\ell }^i(h) \big \langle \tfrac{T_i(h) s}{h}\big \rangle ^{-1}t_\ell ^i\le T_\ell ^i\big \langle \tfrac{T_i(h)s}{h}\big \rangle ^{-1}\\ 1&{}\text {else} \end{array}\right. },\nonumber \\ {\tilde{t}}_{\ell }^i(h)&:={\left\{ \begin{array}{ll}t_{\ell }^i(h) &{}t_\ell ^i\le T_\ell ^i\big \langle \tfrac{T_i(h)s}{h}\big \rangle ^{-1}\\ 1&{}\text {else} \end{array}\right. } \end{aligned}$$
(5.4)

and note that \(\sum _{{\tilde{t}}_\ell ^i={\tilde{T}}_\ell ^i=1}|{{\mathcal {G}}}_{_{t,{\ell }}}^i|^{\frac{1}{2}}\le C_{_{\!{\text {nr}}}}^i \sqrt{\frac{1}{T_i}\Big \langle \frac{T_i s}{h}\Big \rangle }. \) In particular,

$$\begin{aligned} \begin{aligned} \!\!\!\!\sum _{\ell \in \mathcal {L}_{_{E}}^i(h)}\!\!\frac{(|{{\mathcal {G}}}_{_{t,{\ell }}}^i|{\tilde{t}}_\ell ^i)^{\frac{1}{2}}}{({\tilde{T}}_\ell ^i)^{\frac{1}{2}}}\le 2C_{_{\!{\text {nr}}}}^i \sqrt{\frac{1}{T_i}\Big \langle \frac{T_i s}{h}\Big \rangle }, \qquad \sum _{\ell \in \mathcal {L}_{_{E}}^i(h)}\!\!\sqrt{|{{\mathcal {G}}}_{_{t,{\ell }}}^i|{\tilde{t}}_\ell ^i{\tilde{T}}_\ell ^i}\le 2C_{_{\!{\text {nr}}}}^i \Big (\frac{1}{T_i}\Big \langle \frac{T_i s}{h}\Big \rangle \Big )^{-\frac{1}{2}}. \end{aligned} \end{aligned}$$
(5.5)

Then, since for each \(\ell \in \mathcal {L}_{_{E}}^i(h)\) the union of tubes with indices in \({\mathcal {G}^i_{_{E,\ell }}}\) is also \([{\tilde{t}}_\ell ^i(h),{{\tilde{T}}}_{\ell }^i(h)]\) non-self looping, we may apply Lemma 4.3 with the sets \(\{{{\mathcal {G}}}_{_{t,{\ell }}}^i(h)\}_{\ell \in \mathcal {L}_{_{E}}^i(h)}\), \(\{{{\tilde{T}}}_\ell ^i(h)\}_{\ell \in \mathcal {L}_{_{E}}^i(h)}\), \(\{t_\ell ^i(h)\}_{\ell \in \mathcal {L}_{_{E}}^i(h)}\) to see that \(\{{\mathcal {T}}_j\}_{j\in {{\mathcal {G}}}_{_{t,{\ell }}}^i(h)}\) has \(\{(t_j,T_j)\}\) density on [ab] where \(t_j={\tilde{t}}_j^i(h)\), \(T_j={\tilde{T}}_j^i(h)\). Then, using Lemma 4.4 with operators \(A_i \in \Psi ^\infty (M)\), \(\psi \in C^\infty _0({\mathbb {R}};[0,1])\) with \(\psi (t)=1\) for \(|t| \le \tfrac{1}{4}\) and \(\psi (t)=0\) for \(|t| \ge 1\), and for \(s \in {\mathbb {R}}\) let \(u=\mathbb {1}_{[t-s,t]}(P)w\), where w is any function in \(L^2(M)\) with \(\Vert w\Vert _{_{\!L^2(M)}}=1\). Next, by Lemma 4.4, for \(i=1,2\), there exist \(C_{_{0}}^i=C_{_{0}}(n,k_i,{\mathfrak {I}}_{_{\!0}}^i, A_i)\), \(C>0\), and for all N there is \(C_{_{N}}>0\) such that for all \(0<h<h_0\), \(s\in {\mathbb {R}}\), and \(t \in [E-Kh, E+Kh]\)

$$\begin{aligned}&h^{\frac{k_i-1}{2}}\Big |\int _{H_i}A_i\mathbb {1}_{[t-s,t]}(P)w\,d\sigma _{_{\!H_i}}\Big |\nonumber \\&\le C_{_{0}}^iR(h)^{\frac{n-1}{2}}\!\!\!\! \sum _{\ell \in \mathcal {L}_{_{E}}^i(h)}\!\!\frac{(|{{\mathcal {G}}}_{_{t,{\ell }}}^i|{{\tilde{t}}}_\ell ^i)^{\frac{1}{2}}}{(\tau {{\tilde{T}}}_\ell ^i)^{\frac{1}{2}}}\Vert \mathbb {1}_{[t-s,t]}(P)w\Vert _{_{\!L^2(M)}}+C_{_{0}}^iR(h)^{\frac{n-1}{2}}\!\!\!\! \sum _{\ell \in \mathcal {L}_{_{E}}^i(h)}\frac{(|{{\mathcal {G}}}_{_{t,{\ell }}}^i| {\tilde{t}}_\ell ^i {{\tilde{T}}}_\ell ^i)^{\frac{1}{2}}}{h}\Vert P_t \mathbb {1}_{[t-s,t]}(P)w\Vert _{_{\!L^2(M)}}\nonumber \\&\quad + Q^{A,\psi }_{t,h}(C, C_{_{N}}, \mathbb {1}_{[t-s,t]}(P)w). \end{aligned}$$
(5.6)

Note that for all N there is \(C_{_{N}}>0\) such that for all \(t\in [a{-Kh},b{+Kh}]\), \(|s|\le 10\) and \(0<h<1\)

$$\begin{aligned} \Vert P_t\mathbb {1}_{[t-s,t]}(P)\Vert _{L^2\rightarrow H_{scl }^N}\le C_{_{N}}|s|,\qquad \Vert \mathbb {1}_{[t-s,t]}(P)\Vert _{L^2\rightarrow L^2}\le 1. \end{aligned}$$
(5.7)

In addition, we use the elliptic parametrix construction, together with \(|s|\le 2h\) to obtain

$$\begin{aligned} \Vert \big (1-\psi \big (\tfrac{P_t}{h^\delta }\big )\big )P_tA_i\mathbb {1}_{[t-s,t]}(P)\Vert _{L^2\rightarrow H_{scl }^N}\le C_{_{N}} h^N. \end{aligned}$$
(5.8)

We combine these estimates with (5.3) and the definition of \({{\tilde{T}}}_\ell ^i\) into (5.2) to obtain that for all \(0<h<h_0\), \(|s|\le 2h\), \(K>0\), and \(t \in [E-Kh, E+Kh]\),

$$\begin{aligned}&h^{\frac{k_i-1}{2}}\Vert \mathbb {1}_{[t-s,t]}(P)\, A_i^*\, \delta _{_{H_i}}\Vert {_{\!L^2(M)}}\\&\quad \le C_{_{0}}^iC_{_{\!{\text {nr}}}}^i \bigg ( \frac{1}{\tau ^{\frac{1}{2}}}\Big (\frac{1}{ T_i} \Big \langle \frac{T_i s}{h} \Big \rangle \Big )^\frac{1}{2} + \frac{|s|}{h} \Big (\frac{1}{ T_i} \Big \langle \frac{T_i s}{h} \Big \rangle \Big )^{\!-\frac{1}{2}}\bigg ) +C_{_{N}}h^N. \end{aligned}$$

In particular, since \(\tau <1\), using this estimate in (5.2) we conclude that for all \(0<h<h_0\), \(|s|\le 2h\), \(K>0\), and \(t \in [E-Kh, E+Kh]\)

$$\begin{aligned} h^{\frac{k_i-1}{2}}\Big (\sum _{t-s\le E_j\le t}\Big |\int _{H_i}A_i\phi _{_{E_j}}d\sigma _{_{\!H_i}}\Big |^2\Big )^{\tfrac{1}{2}}\le \frac{C_{_{0}}^iC_{_{\!{\text {nr}}}}^i}{\sqrt{\tau T_i(h)}}\Big \langle \frac{T_i(h)s}{h}\Big \rangle ^{\tfrac{1}{2}} +C_{_{N}}h^N. \end{aligned}$$

Combining estimates for \(H_1\) and \(H_2\) using (5.1), and \(C_{_{\!{\text {nr}}}}^i\le C_{_{\!{\text {nr}}}}\) completes the proof. \(\square \)

The last lemma shows that \(w_h(s)=\Pi _{_{H_1,H_2}}^{^{A_1,A_2}}(s)\) has at most polynomial growth at infinity.

Lemma 5.5

Let \(\ell _1, \ell _2\in {\mathbb {R}}\). Then, there is \(N_0>0\) such that for all \(A_1\in \Psi ^{\ell _1}_\delta (M)\), \(A_2\in \Psi ^{\ell _2}_\delta (M)\), there are \(C_{_{1}}>0\), \(h_0>0\), such that for all \(0<h<h_0\) and \(s\in {\mathbb {R}}\),

$$\begin{aligned} |\Pi _{_{H_1,H_2}}^{^{A_1,A_2}}(s)|\le C_{_{1}}h^{-\frac{k_1+k_2}{2}}\langle s\rangle ^{N_0}. \end{aligned}$$

Proof

Arguing as in (5.1), and (5.2), it is enough to prove that there is \(C_{_{1}}>0\) such that for each \(i=1,2\) there is \(N_i>0\) for which

$$\begin{aligned} \sup _{\Vert w\Vert _{_{\!L^2(M)}}=1}\Big | \int _{H_i}A_i\mathbb {1}_{(-\infty ,s]}(P) w\,d\sigma _{_{\!H_i}}\Big |\le C_{_{1}}h^{-\frac{k_i}{2}}\langle s\rangle ^{N_i}. \end{aligned}$$

Applying Lemma 4.5 with \(u=\mathbb {1}_{(-\infty ,s]}(P) w\) yields that for any \(\psi \in S^0(T^*M;[0,1])\) with \(\psi \equiv 1\) on \(N^*\!H\) and \(r_i> \tfrac{k_i+2\ell _i}{2m}\) there exist \(C_{_{1}}>0\) and \(h_0>0\) such that for all \(N>0\) there is \(C_{_{N}}>0\) satisfying for \(0<h<h_1\) and \(s\in {\mathbb {R}}\),

$$\begin{aligned} \begin{aligned}&h^{\frac{k_i}{2}}\Big | \int _{H_i}\!A_i\mathbb {1}_{(-\infty ,s]}(P) w\,d\sigma _{_{\!H_i}}\Big | \le C_{_{N}}h^N\Vert \mathbb {1}_{(-\infty ,s]}(P) w\Vert _{H^{-N}_{{\text {scl}}}(M)}\\&\qquad + C_{_{1}} \big (\Vert {Op_h(\psi )}\mathbb {1}_{(-\infty ,s]}(P) w\Vert _{{_{\!L^2(M)}}}\!\!+\Vert {Op_h(\psi )} P_s^{r_i}\mathbb {1}_{(-\infty ,s]}(P) w\Vert _{{_{\!L^2(M)}}}\big ). \end{aligned} \end{aligned}$$
(5.9)

Finally, the last term is bounded by \(C_{_{1}} (1+|s|^{r_i})\) since \(\Vert f(P)\Vert _{L^2\rightarrow L^2}\le \Vert f\Vert _{L^\infty }.\) \(\square \)

6 Smoothed projector with non-looping condition

This section is dedicated to the proof of Theorems 8 and 9. The crucial step, completed in Sect. 6.1, is to bound \( (\rho _{_{h,{\widetilde{T}}(h)}}-\rho _{_{h,t_0}})*\Pi _{_{H_1,H_2}}^{^{A_1,A_2}}\) when the pair \((H_1,H_2)\) is \((t_0,\textbf{T})\) non-looping and \({\widetilde{T}}(h)=\frac{1}{2}\textbf{T}(R(h))\). In Sect. 6.2 we prove Theorem 8 by combining the estimates from §6.1 with Proposition 5.1. In §6.3 we derive Theorem 9 from Theorem 8.

6.1 Comparing against a short fixed time

Throughout this section we continue to assume \(H_1 \subset M\) and \(H_2\subset M\) are two submanifolds of co-dimension \(k_1\) and \(k_2\) respectively. The goal is to show that, under the assumption \((H_1,H_2)\) is a \((t_0,\textbf{T})\) non-looping pair in the window [ab], we can control \(\rho _{\sigma _{h,{\widetilde{T}}(h)}}*\Pi _h\) by comparing it to \(\rho _{_{h,t_0}}*\Pi _h\). For the rest of the section we write

$$\begin{aligned} {\widetilde{T}}(h):=\tfrac{1}{2}{\textbf{T}(R(h))},\qquad T(h):=\textbf{T}(R(h)). \end{aligned}$$

Proposition 6.1

Suppose \(a, b\in {\mathbb {R}}\) are such that \(H_1, H_2\) are conormally transverse for p in the window [ab]. Let \(\tau _0, R_0\) be as in Lemma 4.1. Let \(0<\tau <\tau _0\), \(0<\delta <\tfrac{1}{2}\), and \(\textbf{T}\) a sub-logarithmic resolution function with \(\Omega (\textbf{T})\Lambda <1-2\delta \).

Suppose \((H_1,H_2)\) is a \((t_0, {\textbf{T}})\) non-looping pair in the window [ab] via \(\tau \)-coverings with constant \(C_{_{\!{\text {nl}}}}\). Let \(A_1,A_2\in \Psi ^\infty (M)\), \( {h^\delta } \le R(h) \le R_0\), and \(K>0\). There exist

$$\begin{aligned} C_{_{0}}=C_{_{0}} (n,k_1, k_2,{\mathfrak {I}}_{_{\!0}}^1, {\mathfrak {I}}_{_{\!0}}^2,A_1,A_2,C_{_{\!{\text {nl}}}})>0 \end{aligned}$$

and \(h_0>0\) such that for all \(0<h<h_0\) and all \(E\in [a-Kh,b+Kh]\),

$$\begin{aligned} \Big | (\rho _{_{h,{\widetilde{T}}(h)}}-\rho _{_{h,t_0}})*\Pi _{_{H_1,H_2}}^{^{A_1,A_2}}(E) \Big |\le C_{_{0}} h^{\frac{2-k_1-k_2}{2}}\Big /{\textbf{T}(R(h))}. \end{aligned}$$
(6.1)

We prove the proposition at the end of the section. The proof hinges on four lemmas. The first one, Lemma 6.3, rewrites the left hand side in (6.1) in terms of the function

$$\begin{aligned} f_{_{S,T,h}}(\lambda ):= & {} f_{_{S,T}}(h^{-1}\lambda ), \nonumber \\ f_{_{S,T}}(\lambda ):= & {} \frac{1}{i}\int _{\mathbb {R}}\frac{1}{\tau }\,{\hat{\rho }}\big (\tfrac{\tau }{T}\big )\big (1-{\hat{\rho }}\big (\tfrac{\tau }{S}\big )\big ) e^{-i{\tau }\lambda }d\tau , \end{aligned}$$
(6.2)

where ST are two positive constants with \(S<T\), and \(\rho \) is as in (1.16)

Remark 6.2

We note that for all \(N>0\)

$$\begin{aligned} |f_{_{S, T}}(\lambda )|\le C_{_{N}}\langle \lambda S\rangle ^{-N},\qquad {{\,\textrm{supp}\,}}{\hat{\rho }}\big (\tfrac{\tau }{T}\big )\big (1-{\hat{\rho }}\big (\tfrac{\tau }{S}\big )\big ) \subset \{ \tau \in {{\mathbb {R}}}:\; |\tau | \in [S, 2T]\}. \end{aligned}$$
(6.3)

Lemma 6.3

Suppose \(k>0\) and \(P\in \Psi ^k(M)\) is self-adjoint with symbol satisfying (1.9). Then, for all \(N>0\),

$$\begin{aligned} (\rho _{_{h,{\widetilde{T}}}}-\rho _{_{h,t_0}})*\Pi _h(E)={f_{_{t_0,{\widetilde{T}},h}}\big (P_{_{E}}\big )}+O(h^N)_{H_{scl }^{-N}\rightarrow H_{scl }^N}. \end{aligned}$$

Proof

First, we prove that if P is self-adjoint \(E_1, E_2 \in {\mathbb {R}}\), then

$$\begin{aligned} \int _{E_1}^{E_2}(\rho _{_{h,{\widetilde{T}}(h)}}-\rho _{_{h,t_0}})*\partial _s\Pi _h(s)ds= {f_{_{t_0,{\widetilde{T}}(h),h}}\big (P_{_{E_2}}\big )-f_{_{t_0,{\widetilde{T}}(h),h}}\big (P_{_{E_1}}\big ).}\nonumber \\ \end{aligned}$$
(6.4)

To ease notation write \({\widetilde{T}}\) for \({\widetilde{T}}(h)\). To prove (6.4) we write

$$\begin{aligned} \int _{E_1}^{E_2}(\rho _{_{h,{\widetilde{T}}}}-\rho _{_{h,t_0}})*\partial _s\Pi _h(s)ds&=\int _{E_1}^{E_2}\int _{\mathbb {R}}{\hat{\rho }}\big (\tfrac{w}{\sigma _{_{h,{\widetilde{T}}}}}\big )\big [1-{\hat{\rho }}\big (\tfrac{w}{\sigma _{_{h,t_0}}}\big )\big ]e^{-iw(P-s)}dw ds, \end{aligned}$$

where we use \({\hat{\rho }}\big (\tfrac{w}{\sigma _{_{h,t_0}}}\big )={\hat{\rho }}\big (\tfrac{w}{\sigma _{_{h,{\widetilde{T}}}}}\big ){\hat{\rho }}\big (\tfrac{w}{\sigma _{_{h,t_0}}}\big )\). Putting \(\tau :=hw\), (6.4) follows.

Next, let \(N>0\). By (6.4) it suffices to find \(E_1 \in {\mathbb {R}}\) such that for all \({t>c>0}\)

$$\begin{aligned}{} & {} \big \Vert {f_{t_0,{\widetilde{T}},h}\big (P_{_{E_1}}\big )}\big \Vert _{H_{scl }^{-N}\rightarrow H_{scl }^N}\le C_{_{N}} h^{2N},\nonumber \\ {}{} & {} \Vert \rho _{_{h,t}}*\Pi _h(E_1)\Vert _{H_{scl }^{-N}\rightarrow H_{scl }^N} = O(h^N). \end{aligned}$$
(6.5)

To prove the first claim in (6.5), note that by (6.3) for all \(N>0\) there is \(C_{_{N}}>0\) such that

$$\begin{aligned} \big \Vert P_{_{E_1}}^N{f_{t_0,{\widetilde{T}},h}\big (P_{_{E_1}}\big )}P_{_{E_1}}^N\big \Vert _{L^2\rightarrow L^2}\le C_{_{N}}h^{2N}. \end{aligned}$$

Next, since P satisfies (1.9), there is \(a>0\) such that \(p(x,\xi )>-a\) for all \((x, \xi ) \in T^*M\). In particular, for \(E_1\le -2a\), \(P_{_{E_1}}\) is elliptic and we have \( P_{_{E_1}}^{-1}:H_{scl }^s(M)\rightarrow H_{scl }^{s+k}(M)=O_s(1) \) for all \(s \in {\mathbb {R}}\). Then, for \(E_1\le {-2a}\) the first claim in (6.5) follows.

Next, by the sharp Gårding inequality, there is \(C>0\) such that \(\Pi _h(s)\equiv 0\) for \(s\le -a-Ch\). Thus, for \(E_1\le -3a\) and all \(N,M\ge 0\) there is \(C_{_{M,N}}>0\) such that

$$\begin{aligned} \Vert ({\rho _{_{h,t}}}*\Pi _h)(E_1)\Vert _{H_{scl }^{-N}\rightarrow H_{scl }^N}{} & {} \le \int _{{\mathbb {R}}}\tfrac{t}{h}\rho \big (\tfrac{t}{h}s\big )\Vert \Pi _h(E_1-s)\Vert _{H_{scl }^{-N}\rightarrow H_{scl }^N}ds\\{} & {} \le C_{_{M,N}} \int _{s\le -a}\tfrac{t}{h}\big \langle \tfrac{t}{h}s\big \rangle ^{-M}\langle s\rangle ^{2N/k}. \end{aligned}$$

The claim follows after choosing M large enough. \(\square \)

Let \(H_1, H_2\), \(t_0, T(h)\), \(\tau ,\) and R(h) be as in Proposition 6.1. Since \((H_1,H_2)\) is a \((t_0, \textbf{T})\) non-looping pair in the window [ab] via \(\tau _0\)-coverings, for \(i=1,2\) and \(h>0\) we let

$$\begin{aligned}{} & {} \{{\mathcal {T}}_j^i\}_{j \in \mathcal {J}^i(h)}\quad \text {a}\, ({\mathfrak {D}}_n,\tau , R(h))\text {-good cover of}\, \Sigma _{_{[a,b]}}^{H_i} \,\text {satisfying}\, (1) \,\text {and}\, (2) \nonumber \\{} & {} \quad \text {in Definition} 2.1. \end{aligned}$$
(6.6)

We study \(A_1{f_{_{t_0,{\widetilde{T}},h}}\big (P_{_{E}}\big )}\,A_2^*\) by understanding the behavior of

$$\begin{aligned} {F_{{j,\ell }}^{^{A_1,A_2}}(E,h)}:= Op_h(\chi _{_{{\mathcal {T}}_j^1}})A_1{f_{_{t_0,{\widetilde{T}},h}}\big (P_{_{E}}\big )}\,A_2^*Op_h(\chi _{_{{\mathcal {T}}_{\ell }^2}}) \end{aligned}$$
(6.7)

for \(j \in \mathcal {J}^1(h)\) and \(k \in \mathcal {J}^2(h)\). Next, we study the case when \({\mathcal {T}}_j^1\) does not loop through \({\mathcal {T}}_k^2\).

Lemma 6.4

Assume \(H_1\) and \(H_2\) are conormally transverse for p in the window [ab]. For \(i=1,2\) let \(\{{\mathcal {T}}_j^i\}_{j\in \mathcal {J}^i(h)}\) as in (6.6) and \(j \in \mathcal {J}^1(h)\), \(\ell \in \mathcal {J}^2(h)\) be such that

$$\begin{aligned} \varphi _t({\mathcal {T}}_j^1)\cap {\mathcal {T}}_{\ell }^2=\emptyset ,\qquad |t|\in [t_0{+\tau },T(h){-\tau }]. \end{aligned}$$

Let \(K>0\) and \(\mathcal {V}\) be a bounded subset of \(S_\delta (T^*M;[0,1])\). Then, there exists \(h_0>0\) and for all \(N>0\) there exists \(C_{_{N}}>0\) such that for all \(0<h<h_0\), \(E\in [a-Kh,b+Kh]\), and every \(\delta \)-partition \(\{\chi _{_{{\mathcal {T}}_j^i}}\}_{j\in \mathcal {J}_{_{E}}^i(h)}\subset {\mathcal {V}}\) associated to \(\{{\mathcal {T}}_j\}_{j\in \mathcal {J}_{_{E}}^i(h)}\), \(i=1,2\),

$$\begin{aligned} \Vert {F_{{j,\ell }}^{^{A_1,A_2}}(E,h)}\Vert _{H^{-N}_{{\text {scl}}}(M) \rightarrow H^N_{{\text {scl}}}(M)}\le C_{_{N}}h^N. \end{aligned}$$

Proof

By Egorov’s theorem, for all \(N>0\) there exist \(h_0>0\) and \(C_{_{N}}>0\) such that for all \(0<h<h_0\), \(E\in [a-Kh,b+Kh]\), and \(|t|\in [t_0+\tau , T(h)-\tau ]\)

$$\begin{aligned}{} & {} \big \Vert Op_h(\chi _{_{{\mathcal {T}}_j^1}})A_1\, e^{-i t \frac{P_{_{E}}}{h}}A_2^*\,Op_h(\chi _{_{{\mathcal {T}}_{\ell }^2}})\big \Vert _{H^{-N}_{{\text {scl}}}(M) \rightarrow H^N_{{\text {scl}}}(M)}\le C_{_{N}}h^N , \end{aligned}$$

(see e.g. [18, Proposition 3.9]). The claim follows from the definition (6.2) together with the facts that by (6.3) the support of its integrand has \(\tau \in [t_0, 2 {\widetilde{T}}(h)]\), and \({\widetilde{T}}(h)=\tfrac{1}{2}T(h)\). \(\square \)

The next lemma provides an estimate for \({F_{{j,\ell }}^{^{A_1,A_2}}(E,h)}\) based on volumes of tubes.

Lemma 6.5

Assume \(H_1\) and \(H_2\) are conormally transverse for p in the window [ab]. Let \(A_1\), \(A_2\), \(\tau _0\), \(R_0\), \(\tau \), \(\delta \), and R(h) be as in Proposition 6.1. For \(i=1,2\) let \(\{{\mathcal {T}}_j^i\}_{j\in \mathcal {J}^i(h)}\) be a \(({\mathfrak {D}}_n, \tau , R(h))\)-good covering of \( \Sigma _{_{[a,b]}}^{H_i}\). Let \(K>0\) and \(\mathcal {V}\) a bounded subset of \(S_\delta (T^*M;[0,1])\). Then, there are \(C_{_{0}}=C_{_{0}}(n,k_1, k_2,{\mathfrak {I}}_{_{\!0}}^1, {\mathfrak {I}}_{_{\!0}}^2, A_1,A_2,\mathcal {V})\) and \(h_0>0\), and for all \(N>0\) there exists \(C_{_{N}}>0\) such that the following holds. For all \(0<h<h_0\), \(E\in [a-Kh,b+Kh]\), all \(\delta \)-partitions \(\{\chi _{_{{\mathcal {T}}_j^i}}\}_{j\in \mathcal {J}_{_{E}}^i(h)}\subset {\mathcal {V}}\) and \(\mathcal {I}_i \subset \mathcal {J}_{_{E}}^i(h)\) for \(i=1,2\), and all \(t_0, {{\tilde{T}}}\) with \(0<t_0<{\widetilde{T}}\),

$$\begin{aligned}&\bigg |\int _{H_1}\int _{H_2} \sum _{\ell \in \mathcal {I}_1,j\in \mathcal {I}_2}\!\!\! {F_{{j,\ell }}^{^{A_1,A_2}}(E,h)}(x,y)d\sigma _{_{\!H_2}}(y)d\sigma _{_{\!H_1}}(x)\bigg |\\ {}&\le C_{_{0}} \tau ^{-1} h^{\frac{2-k_1-k_2}{2}}R(h)^{n-1}|\mathcal {I}_1|^{\frac{1}{2}}|\mathcal {I} _2|^{\frac{1}{2}} + C_{_{N}}h^N. \end{aligned}$$

Proof

The first step in our proof is to define for \(0<t_0<{\widetilde{T}}\) the functions

$$\begin{aligned} g^2_{_{t_0,{\widetilde{T}}}}(\lambda )g^1_{_{t_0,{\widetilde{T}}}}(\lambda ):=f_{_{t_0,{\widetilde{T}}}}(\lambda ), \qquad \qquad g^2_{_{t_0,{\widetilde{T}}}}(\lambda ):=\langle t_0\lambda \rangle ^{-N_0}, \end{aligned}$$

where \(N_0 \ge 1\) will be chosen later. Note that by (6.3) for all \(L>0\) there is \(C_{L}>0\) such that

$$\begin{aligned} |g^1_{_{t_0,{\widetilde{T}}}}(\lambda )|\le C_{L}\langle t_0 \lambda \rangle ^{-L+1}. \end{aligned}$$
(6.8)

Since \({f_{_{t_0,{\widetilde{T}},h}}\big (P_{_{E}}\big )= g^1_{_{t_0,{\widetilde{T}},h}}\big (P_{_{E}}\big )g^2_{_{t_0,{\widetilde{T}},h}}\big (P_{_{E}}\big )},\) we may use Cauchy-Schwarz to bound

$$\begin{aligned}&\Big |\int _{H_1}\int _{H_2} \sum _{\ell \in \mathcal {I}_1,j\in \mathcal {I}_2}\Big [ {F_{{j,\ell }}^{^{A_1,A_2}}(E,h)}\Big ](x,y)d\sigma _{_{\!H_2}}(y)d\sigma _{_{\!H_1}}(x)\Big |\\&\qquad \le \Big \Vert \sum _{\ell \in \mathcal {I}_1} {g^1_{_{t_0,{\widetilde{T}}}}\big (P_{_{E}}\big )}A_1^*\,Op_h(\chi _{_{{\mathcal {T}}_\ell ^1}})\delta _{_{H_1}}\Big \Vert _{_{\!L^2(M)}}\\&\Big \Vert \sum _{\ell \in \mathcal {I}_2}{g^2_{_{t_0,{\widetilde{T}},h}}\big (P_{_{E}}\big )}A_2^*\,Op_h(\chi _{_{{\mathcal {T}}_\ell ^2}})\delta _{_{H_2}}\Big \Vert _{_{\!L^2(M)}}. \end{aligned}$$

Next, we use that for \(i=1,2\),

$$\begin{aligned}&\Big \Vert \sum _{\ell \in \mathcal {I}_i} {g^i_{_{t_0,{\widetilde{T}},h}}\big (P_{_{E}}\big )}A_i^*\,Op_h(\chi _{_{{\mathcal {T}}_\ell ^i}})\delta _{_{H_i}}\Big \Vert _{_{\!L^2(M)}}\\&\quad \le \sup _{\Vert w\Vert =1} \Big |\int _{H_i}\sum _{\ell \in \mathcal {I}_i}Op_h(\chi _{_{{\mathcal {T}}_{\ell }^i}})A_i \, {g^i_{_{t_0,{\widetilde{T}},h}}\big (P_{_{E}}\big )}w \,d\sigma _{_{\!H_i}}\Big |. \end{aligned}$$

Thus, let \(w \in L^2(M)\) and fix \(i \in \{1,2\}\). We next apply Lemma 4.4 to the function \(u={g^i_{_{t_0,{\widetilde{T}},h}}\big (P_{_{E}}\big )}w\) and operator \(A=\sum _{j\in \mathcal {I}_i}Op_h(\chi _{_{{\mathcal {T}}_{j}^i}}) A_i\in \Psi _\delta ^\infty (M)\). Here, we use that \({\hbox {MS}}_{\textrm{h}}(A) \subset \cup _{j \in \mathcal {I}_i}{\mathcal {T}}_j^i\) and that \(\tfrac{1}{h}[P_{_{E}},A] \in \Psi ^\infty _\delta (M)\) (see the definition of a \(\delta \)-partition (4.2)). In particular, we may fix \(\mathcal {W}\subset \Psi ^\infty _\delta (M)\) such that \(\tfrac{1}{h}[P_{_{E}},A] \in \mathcal {W}\) regardless of the choice of cover and \(\delta \)-partition contained in \(\mathcal {V}\). Then, the constant \(C_{_{0}}^i\) provided by the Lemma depends on \(A_i\) instead of \(\mathcal {W}\).

Fix \(\psi \in C^\infty _0({\mathbb {R}};[0,1])\) with \(\psi (t)=1\) for \(|t| \le \tfrac{1}{4}\) and \(\psi (t)=0\) for \(|t| \ge 1\). By Lemma 4.4 with \(t_1=t_0\), \(T_1=t_0\), and \({{\mathcal {G}}}_\ell =\emptyset \) for all \(\ell >1\), we obtain that there are \(C_{_{0}}^i=C_{_{0}}^i(n,k_i,{\mathfrak {I}}_{_{\!0}}^i,A_i)>0\), \(C>0\), there exist \(h_0>0\) and for all \(N>0\) there is \(C_{_{N}}>0\) such that for all \(0<h<h_0\)

$$\begin{aligned}&h^{\frac{k_i-1}{2}}\Big |\int _{H_i} \sum _{j\in \mathcal {I}_i} Op_h(\chi _{_{{\mathcal {T}}_j^i}})A_i \,{g^i_{_{t_0,{\widetilde{T}},h}}\big (P_{_{E}}\big )}w\, d\sigma _{_{\!H_i}}\Big |\le Q^{A,\psi }_{E,h}(C, C_{_{N}}, {g^i_{_{t_0,{\widetilde{T}},h}}\big (P_{_{E}}\big )}w)\\&\qquad \qquad \qquad + C_{_{0}}^i\,R(h)^{\frac{n-1}{2}}|\mathcal {I}_i|^{\frac{1}{2}} \Big (\frac{1}{\tau ^{\frac{1}{2}} }\big \Vert {g^i_{_{t_0,{\widetilde{T}},h}}\big (P_{_{E}}\big )}w\big \Vert _{_{\!L^2(M)}}+ \tfrac{t_0}{h}\big \Vert P_{_{E}}{g^i_{_{t_0,{\widetilde{T}},h}}\big (P_{_{E}}\big )}w\big \Vert _{_{\!L^2(M)}}\Big ). \end{aligned}$$

By the definitions \(g^i_{_{t_0,{\widetilde{T}}}}\), \(i=1,2\) and (6.8) there exists \(C>0\) such that for all \(t_0,{\widetilde{T}}\) with \(t_0<{\widetilde{T}}\),

$$\begin{aligned} \big \Vert {g^i_{_{t_0,{\widetilde{T}},h}}\big (P_{_{E}}\big )}\big \Vert _{L^2\rightarrow L^2}\le C,\qquad \big \Vert P_{_{E}}{g^i_{_{t_0,{\widetilde{T}},h}}\big (P_{_{E}}\big )}\big \Vert _{L^2\rightarrow L^2}\le C\frac{h}{t_0},\quad i=1,2. \end{aligned}$$

In addition, note that for \(i=1,2\) there exists \(C_{_{N_0}}>0\) such that

$$\begin{aligned} \big \Vert \big (1-\psi \big (\tfrac{P_{_{E}}}{h^\delta }\big )\big )P_{_{E}}A\,{g^i_{_{t_0,{\widetilde{T}},h}}\big (P_{_{E}}\big )}\big \Vert _{L^2 \rightarrow L^2}\le C_{_{N_0}}h^{N_0(1-\delta )+\delta }. \end{aligned}$$

The claim follows from choosing \(N_0\) large enough that \(N_0(1-\delta )+\delta \ge N\). \(\square \)

Lemma 6.6

Assume the same assumptions as in Proposition 6.1. For \(i=1,2\) let \(\{{\mathcal {T}}_j^i\}_{j\in \mathcal {J}^i(h)}\) be as in (6.6), \(\mathcal {V}\) be a bounded subset of \(S_\delta (T^*M;[0,1])\) and \(K>0\). There exists \(h_0>0\), and for all \(N>0\) there exists \(C_{_{N}}>0\) such that for all \(0<h<h_0\), \(E\in [a-Kh,b+Kh]\), and every \(\delta \)-partition \(\{\chi _{_{{\mathcal {T}}_j^i}}\}_{j\in \mathcal {J}_{_{E}}^i(h)}\subset {\mathcal {V}}\) associated to \(\{{\mathcal {T}}_j^i\}_{j\in \mathcal {J}_{_{E}}^i(h)}\),

$$\begin{aligned}{} & {} \Big \Vert \gamma _{_{H_1}}A_1{f_{_{t_0, {\widetilde{T}},h}}\big (P_{_{E}}\big )}A_2^*\delta _{_{H_2}} -\sum _{j\in \mathcal {J}^1_{_{{E}}}(h),\, \ell \in \mathcal {J}^2_{_{{E}}}(h)}\!\!\!\!\!\!\!\!\!\!\!\!\!\gamma _{_{H_1}} {F_{{j,\ell }}^{^{A_1,A_2}}(E,h)}\delta _{_{H_2}}\Big \Vert _{H_{_{{\text {scl}}}}^{-N}(H_2)\rightarrow H_{_{{\text {scl}}}}^N(H_1)}\\{} & {} \quad \le C_{_{N}}h^N. \end{aligned}$$

Proof

Let \(K>0\) and \(\psi \in C^\infty _c((-1,1);[0,1])\) with \(\psi (t)=1\) for \(|t| \le \tfrac{1}{4}\). We claim there exists \(h_0>0\) such that for all \(N>0\) there is \(C_{_{N}}>0\) so that for \(0<h<h_0\), \(E\in [a-Kh,b+Kh]\).

$$\begin{aligned}{} & {} \Vert \big (1-{\psi }\big (\tfrac{P_{_{E}}}{h^\delta }\big )\big )f_{_{t_0, {\widetilde{T}},h}}\big (P_{_{E}}\big )\Vert _{H_{scl }^{-N}(M)\rightarrow H_{scl }^N(M)}\le C_{_{N}} h^{N}. \end{aligned}$$
(6.9)

To see this, first note that for \({{{\tilde{\psi }}}}\in C_c^\infty \) with \({{\,\textrm{supp}\,}}{{\tilde{\psi }}} \subset \{{\psi }\equiv 1\}\) and \(L>0\),

$$\begin{aligned}{} & {} \big (1-{\psi }\big (\tfrac{P_{_{E}}}{h^\delta }\big )\big )f_{_{t_0, {\widetilde{T}},h}}\big (P_{_{E}}\big ) \\{} & {} \quad =P_{_{E}}^{-L}\big (1-{\psi }\big (\tfrac{P_{_{E}}}{h^\delta }\big )\big )P_{_{E}}^L f_{_{t_0, {\widetilde{T}},h}}\big (P_{_{E}}\big )P_{_{E}}^LP_{_{E}}^{-L}\big (1-{{\tilde{\psi }}}\big (\tfrac{P_{_{E}}}{h^\delta }\big )\big ). \end{aligned}$$

Now, since \(P_{_{E}}\) is classically elliptic in \(\Psi ^m(M)\), for all \(s\in {\mathbb {R}}\),

$$\begin{aligned} P_{_{E}}^{-L}\big (1-\psi \big (\tfrac{P_{_{E}}}{h^\delta }\big )\big )=O_{L,s}(h^{-\delta L})_{{H_{scl }^{s}(M)\rightarrow H_{scl }^{s+mL}(M)}}. \end{aligned}$$
(6.10)

Note that (6.10) also holds with \({{\tilde{\psi }}}\) in place of \(\psi \). In addition, by (6.3)

$$\begin{aligned} P_{_{E}}^L f_{_{t_0, {\widetilde{T}},h}}\big (P_{_{E}}\big )P_{_{E}}^L=O_{_{L}}(h^{2L})_{L^2(M)\rightarrow L^2(M)}. \end{aligned}$$
(6.11)

Taking \(L>\max (N/m,N/(2(1-\delta )))\) and combining (6.10) and (6.11) we obtain (6.9).

Next, for \(i=1,2\) we define \( {G_i}:={\text {Id}}-\sum _{j \in \mathcal {J}_{_{{E}}}^i(h)} \!\!\!Op_h(\chi _{_{\mathcal {T}^i_j}}), \) and note that \({\hbox {MS}}_{\textrm{h}}(G_i)\cap \Lambda ^\tau _{ \Sigma _{_{E}}^{H_i}}(R(h)/2)=\emptyset \). Therefore, combining Lemma 4.1 together with (6.9), there exists \(h_0>0\) such that for all \(N>0\) there is \(C_{_{N}}>0\) so that for all \(0<h<h_0\), \(E\in [a-Kh,b+Kh]\).

$$\begin{aligned}{} & {} \big \Vert \gamma _{_{H_1}}A_1G_1 {f_{_{t_0, {\widetilde{T}},h}}}\big (P_{_{E}}\big )A_2^*\delta _{_{H_2}} \big \Vert _{H_\text {scl}^{-N}(H_2)\rightarrow H_\text {scl}^N(H_1)}\le C_{_{N}}h^N, \end{aligned}$$
(6.12)

In particular, the lemma follows from applying (6.12) and its analogs since

$$\begin{aligned}&{\gamma _{_{H_1}}A_1{f_{_{t_0, {\widetilde{T}},h}}\big (P_{_{E}}\big )}A_2^*\delta _{_{H_2}} -\!\!\!\!\!\!\!\!\!\sum _{j\in \mathcal {J}_E^1(h),\ell \in \mathcal {J}_E^2(h)}\!\!\!\!\!\!\!\!\!\!\gamma _{_{H_1}} {F_{{j,\ell }}^{^{A_1,A_2}}(E,h)}\delta _{_{H_2}}}\\&\quad =\gamma _{_{H_1}}A_1G_1f_{_{t_0, {\widetilde{T}},h}}\big (P_{_{E}}\big )A_2^*\delta _{_{H_2}}+\gamma _{_{H_1}}A_1 f_{_{t_0, {\widetilde{T}},h}}\big (P_{_{E}}\big )G_2A_2^*\delta _{_{H_2}}\\&\qquad +\gamma _{_{H_1}}A_1 G_1 f_{_{t_0, {\widetilde{T}},h}}\big (P_{_{E}}\big )G_2A_2^*\delta _{_{H_2}}. \end{aligned}$$

\(\square \)

Proof of Proposition 6.1. Since \((H_1,H_2)\) is a \((t_0, \textbf{T})\) non-looping pair in the window [ab] via \(\tau _0\)-coverings, for \(i=1,2\) and \(h>0\) we may work with \(\{{\mathcal {T}}_j^i\}_{j\in \mathcal {J}^i(h)}\), as in (6.6) and \(\{\chi _{_{{\mathcal {T}}_j^i}}\}_{j\in \mathcal {J}^i(h)}\) a \(\delta \)-partition associated \(\{{\mathcal {T}}^i_j\}\) For each \(E\in [a,b]\) and \(i=1,2,\) let \(\mathcal {J}^i_{_{E,h}}=\mathcal {B}_{_{E}}^i(h)\cup \mathcal {G}^i_{_{E}}(h)\) be a partition of indices such that property (1) of Definition 2.1 with \(r=R(h)\). Then, by Lemma 6.4, for \(K>0\) there exists \(h_0>0\) such that the following holds: For all \(N>0\) there is \(C_{_{N}}>0\) so that for all \(0<h<h_0\), \(E\in [a-Kh,b+Kh]\), and \(i,k=1,2\) with \(i\ne k\),

$$\begin{aligned} \begin{aligned} \Big |\int _{H_1} \int _{H_2}\sum _{j \in {\mathcal {J}^k_{_{E}}(h)}}\sum _{\ell \in {\mathcal {G}^i_{_{E}}(h)}}[{F_{{j,\ell }}^{^{A_1,A_2}}(E,h)}](x,y) d\sigma _{_{\!H_2}}(y)d\sigma _{_{\!H_1}}(x)\Big | \le C_{_{N}}h^N. \end{aligned} \nonumber \\ \end{aligned}$$
(6.13)

Therefore, considering the remaining term, and applying Lemma 6.5 we obtain the following. There is \(C_{_{0}}=C_{_{0}} (n,k_1, k_2,{\mathfrak {I}}_{_{\!0}}^1, {\mathfrak {I}}_{_{\!0}}^2,A_1,A_2)>0\) and for \(K>0\) there exists \(h_0>0\) such that the following holds: For all \(N>0\) there is \(C_{_{N}}>0\) so that for all \(0<h<h_0\), \(E\in [a-Kh,b+Kh]\),

$$\begin{aligned}&\Big |\int _{H_1} \int _{H_2}\sum _{j \in \mathcal {B}^1_{_{E}}(h)}\sum _{\ell \in \mathcal {B}_{_{E}}^2(h)}[{F_{{j,\ell }}^{^{A_1,A_2}}(E,h)}](x,y)d\sigma _{_{\!H_2}}(y)d\sigma _{_{\!H_1}}(x)\Big | \nonumber \\&\qquad \quad \le C_{_{0}}h^{\frac{2-k_1-k_2}{2}}R(h)^{n-1}{|\mathcal {B}^1_{_{E}}(h)|^{\frac{1}{2}}}|\mathcal {B}^2_{_{E}}(h)|^{\frac{1}{2}}+C_{_{N}}h^N \le C_{_{0}}C_{_{\!{\text {nl}}}}h^{\frac{2-k_1-k_2}{2}}\big /T(h). \qquad \end{aligned}$$
(6.14)

To get the last line we used that our covering satisfies property (2) of Definition 2.1. Combining Lemma 6.6 with  (6.6),  (6.13), and (6.14), we obtain the claim.

6.2 Proof of Theorem 8

Since for \(i=1,2\) the submanifold \(H_i\) is \( T_i(h)\) non-recurrent in the window [ab] via \({\tau _0}\)-coverings with constant \(C_{_{\!{\text {nr}}}}^i\), we may apply Proposition 5.1 to obtain the existence of \(C_{_{0}}=C_{_{0}}(n,k_1, k_2,{\mathfrak {I}}_{_{\!0}}^1, {\mathfrak {I}}_{_{\!0}}^2,A_1,A_2,C_{_{\!{\text {nr}}}}^1, C_{_{\!{\text {nr}}}}^2)\) and for all \(K>0\) obtain \(h_0>0\) such that for all \(0<h\le h_0\) and \({s}\in [a-Kh,b+Kh]\),

$$\begin{aligned} \Big |\Pi _{_{H_1,H_2}}^{^{A_1,A_2}}(s)-\rho _{_{h,{{\widetilde{T}}_{\max }(h)}}}*\Pi _{_{H_1,H_2}}^{^{A_1,A_2}}(s)\Big |\le C_{_{0}}\,h^{\frac{2-k_1-k_2}{2}}\big /\,T(h), \end{aligned}$$
(6.15)

where \(T(h)=(T_1(h)T_2(h))^{\tfrac{1}{2}}\) and \(T_{\max }(h)=\max (T_1(h),T_2(h))\). Note that we are actually applying the proposition only using that \(H_i\) is \(\tfrac{1}{2} T_i(h)\) non-recurrent.

On the other hand, since \((H_1,H_2)\) is a \((t_0, {\textbf{T}_{\max }})\) non-looping pair in the window [ab] via \(\tau _0\) coverings, we may apply Proposition 6.1 to obtain that there exist \(C_{_{1}}=C_{_{1}}(n,k_1, k_2,{\mathfrak {I}}_{_{\!0}}^1, {\mathfrak {I}}_{_{\!0}}^2,A_1,A_2,C_{_{\!{\text {nl}}}})>0\) and for all \(K>0\) there is \(h_0>0\) such that for all \(0<h<h_0\) and all \(s\in [a-Kh,b+Kh]\)

$$\begin{aligned} \Big | (\rho _{_{h,{{\widetilde{T}}_{\max }}(h)}}-\rho _{_{h,t_0}})*\Pi _{_{H_1,H_2}}^{^{A_1,A_2}}(s) \Big |\le C_{_{1}} \,h^{\frac{2-k_1-k_2}{2}}\big /\,T(h). \end{aligned}$$
(6.16)

The result follows from combining (6.15) with (6.16). We note that \(H_1\) and \(H_2\) may be replaced by \({{\tilde{H}}}_{1,h}\) and \({{\tilde{H}}}_{2,h}\) since \(C_{_{\!{\text {nl}}}}\), \(C_{_{\!{\text {nr}}}}^1\), and \(C_{_{\!{\text {nr}}}}^2\) are uniform for \(\{{{\tilde{H}}}_{1,h}\}_h\) and \(\{{{\tilde{H}}}_{2,h}\}_h\).

6.3 Proof of Theorem 9

Let \(0<\tau {<\min (\tau _0,\varepsilon /3)}\). By Proposition 3.5 there exists \(c_0>0\), \(C_{_{\!{\text {nr}}}}=C_{_{\!{\text {nr}}}}(M,p,{\mathfrak {t}},R_0)>0\) such that for \(j=1,2\), the submanifold \(H_j\) is \(c \textbf{T}_i(R)\) non-recurrent in the window [ab] via \(\tau \) coverings with constant \(C_{_{\!{\text {nr}}}}\).

Now, since \((H_1,H_2)\) is a \((t_0,{\textbf{T}}_{\max })\) non-looping pair in the window [ab] with constant \(C_{_{\!{\text {nl}}}}\). Proposition 3.1 implies there is \({\widetilde{C_{_{\!{\text {nl}}}}}}={\widetilde{C_{_{\!{\text {nl}}}}}}(p,a,b,n,C_{_{\!{\text {nl}}}}, H_1, H_2)\) such that \((H_1,H_2)\) is a \((t_0+3\tau _0,{\tilde{\textbf{T}}})\) non-looping pair in the window [ab] via \(\tau _0\)-coverings with constant \({\widetilde{C_{_{\!{\text {nl}}}}}}\) where \({\tilde{\textbf{T}}}(R)=\textbf{T}_{\max }(4R)-3\tau _0\). Since \(\textbf{T}_j\) are sub-logarithmic, there is \(c_1>0\) such that \({\tilde{\textbf{T}}}(R)\ge c_1 \textbf{T}_{\max }(R)\). The proof now follows from a direct application of Theorem 8 with \(\textbf{T}_j\) replaced by \(\min (c_0,c_1)\textbf{T}_j\) and \(t_0\) by \(t_0+\varepsilon \).

7 The Weyl law

In order to improve remainders in the Weyl law itself, we let \({\mathbf {\Delta }} \subset M \times M\) be the diagonal, and for \(A_1, A_2\in \Psi ^\infty (M)\) consider the integral

$$\begin{aligned}&\int _{M}[A_1\mathbb {1}_{(-\infty ,s]}(P)A_2](x,x){{\,\mathrm{{\text {dv}}}\,}}_g(x) \\&\quad = \int _{{\mathbf {\Delta }}}\Big ((A_1\otimes A_2^*)\mathbb {1}_{(-\infty ,s]}(P)\Big )(x,y) d\sigma _{_{\!{\mathbf {\Delta }}}}(x,y), \end{aligned}$$

where \(d\sigma _{_{\!{\mathbf {\Delta }}}}\) is the Riemannian volume form induced on \(\Delta \) by the product metric on \(M \times M\). To ease notation, we write \({\textbf{P}}_t=(P-t)\otimes 1={P\otimes 1-t{\text {Id}}}.\) We will view \({\mathbf {\Delta }}\) as a hypersurface of codimension n in \(M\times M\), and the kernel of \(\mathbb {1}_{[t-s,t]}(P)\) as a quasimode for \({\textbf{P}}_t\). In particular, observe that for any operator \(B:L^2(M)\rightarrow L^2(M)\)

$$\begin{aligned} \Vert {\textbf{P}}_t\mathbb {1}_{[t-s,t]}(P){B}\Vert _{_{L^2(M\times M)}}\le |s|\Vert \mathbb {1}_{[t-s,t]}(P){B}\Vert _{_{L^2(M\times M)}}. \end{aligned}$$
(7.1)

In addition, note that for \((x,\xi ,y,\eta ) \in T^*M \times T^*M\)

$$\begin{aligned} \sigma ({\textbf{P}}_t)(x,\xi ,y,\eta )=p(x,\xi )-t=:{{\textbf{p}}}(x,\xi ,y,\eta )-t=:{\textbf{p}}_t(x,\xi ,y,\eta ). \end{aligned}$$

Therefore, for all \(c>0\), there is \(C>0\) such that if \(c|\eta |\le |\xi |\) and \(|\xi |\ge C\), then

$$\begin{aligned} |\sigma ({\textbf{P}}_t)(x,\xi ,y,\eta )|\ge \tfrac{1}{C}|(\xi ,\eta )|^m. \end{aligned}$$

In particular, since we work near the \({\textbf{p}}\) flow-out of \( N^*{\mathbf {\Delta }}\cap \{{\textbf{p}}=t\} \) where \(t\in [a,b]\), and

$$\begin{aligned} N^*{\mathbf {\Delta }}=\{ (x,\xi ,x,-\xi ):\; (x,\xi )\in T^*M\}, \end{aligned}$$

we may work as though \({\textbf{P}}_t\) were elliptic in \(\Psi ^m(M\times M)\), and apply the results of the previous sections by accepting \(O(h^\infty )\) errors. We will do this without further comment.

We next describe the tubes relevant in this section. We will work microlocally near a point \(\rho _0\in N^*{\mathbf {\Delta }}\cap {\textbf{p}}^{-1}([a,b])\). Let \(\pi _{_{R}},\pi _{_{L}}:T^*(M\times M)\rightarrow T^*M\) denote the projections to the right and left factor, and let \({\mathcal {Z}}_{\pi _{_{L}}(\rho _0)}\subset T^*M\) be a transversal to the flow for p containing \(\pi _{_{L}}(\rho _0)\). (Such a hypersurface exists since \(dp(\rho )\ne 0\) on \(p^{-1}([a,b])\).) Define a transversal to the flow for \({\textbf{p}}\) by

$$\begin{aligned} {\mathcal {Z}}_{\rho _0}:={\mathcal {Z}}_{\pi _{_{L}}(\rho _0)}\times T^*M, \end{aligned}$$

and let U be a neighborhood of \(\rho _0\) in \(N^*{\mathbf {\Delta }}\) such that \( U\cap {\textbf{p}}^{-1}([a,b])\subset {\mathcal {Z}}_{\rho _0}. \) We will use the metric \({{\tilde{d}}}\) on \(T^*M\times M\) defined by \( {{\tilde{d}}}\Big ((\rho _{_{L}},\rho _{_{R}}),(q_{_{L}},q_{_{R}})\Big ):=\max \Big ( d(\rho _{_{L}},q_{_{L}}), d(\rho _{_{R}},q_{_{R}})\Big ), \) for \((\rho _{_{L}},\rho _{_{R}}),(q_{_{L}},q_{_{R}})\in T^*M \times M.\) With this definition, for \(\rho =(\rho _{_{L}},\rho _{_{R}})\in N^*{\mathbf {\Delta }}\cap \{{\textbf{p}}_t=0\}\),

$$\begin{aligned} {\mathcal {T}}_\rho :=\Lambda _\rho ^\tau (r)= {\tilde{\Lambda }}_{\rho _{_{L}}}^\tau (r)\times B(\rho _{_{R}},r) \end{aligned}$$

where \(\Lambda _A^\tau (r)\) is defined by (2.2) with \(\varphi _t\) the Hamiltonian flow for \({\textbf{p}}\) and \({{\tilde{{\mathcal {T}}}}}={\tilde{\Lambda }}_{\rho _{_{L}}}^\tau (r)\) denotes a tube with respect to p and the hypersurface \({\mathcal {Z}}_{\pi _{_{L}}(\rho _0)}\). In particular, when we use cutoffs with respect to a tube \({\mathcal {T}}\), we will always work with cutoffs of the form

$$\begin{aligned} \chi _{_{{\mathcal {T}}}}(x,\xi , y, \eta )=\chi _{_{{\tilde{{\mathcal {T}}}}}}(x,\xi )\chi _{\rho _{_{R}}}(y,{\eta }),\qquad \qquad {{\,\textrm{supp}\,}}\chi _{\rho _{_{R}}}\subset B(\rho _{_{R}},{r}). \end{aligned}$$

We will refer only to this tube in \(T^*M\), leaving the other implicit and will think of the kernel of \(A_1\mathbb {1}_{[a,b]}(P)A_2\) as that of \(\mathbb {1}_{[a,b]}(P)\) acted on by \(A_1\otimes A_2^{{t}}.\) Before we start our proof of the improved Weyl remainder, we need a dynamical lemma.

Lemma 7.1

Let \(C_{_{\!{\text {np}}}}>0\), \(a\le b\), and \(U\subset T^*M\) satisfying \(d\pi _{_{M}}{{\textsf{H}}_p}\ne 0\) on \(p^{-1}([a,b])\cap {\overline{U}}\). Then there are \(\tau _0>0\) and \({\widetilde{C_{_{\!{\text {np}}}}}}={\widetilde{C_{_{\!{\text {np}}}}}}(p,U,C_{_{\!{\text {np}}}})\) such that the following holds. If U is \((t_0,\textbf{T})\) non-periodic for p in the window [ab] with constant \(C_{_{\!{\text {np}}}}\), then \(N^*\mathbf{\Delta }\cap (U\times T^*M)\) is \((t_0+{3}\tau _0,\textbf{T}({16}R)-{3}\tau _0)\) non-looping for \({\textbf{p}}\) via \(\tau _0\)-coverings in the window [ab] with constant \({\widetilde{C_{_{\!{\text {np}}}}}}\).

Proof

Let \(E\in [a,b]\). We work with \(\mathcal {L}_{_{\mathbf{\Delta },\mathbf{\Delta }}}^{{R,E}}(t_0,T)\) as defined in Definition 1.12 but with p replaced by \({\textbf{p}}\), \(\varphi _t^{{\textbf{p}}}:=\exp (t H_\textbf{p})\), and \(\Sigma _{_{E}}^{\mathbf{\Delta }}=N^*\!\mathbf{\Delta } \cap \{{\textbf{p}}=\!E\}\). First, we claim

$$\begin{aligned} \pi _{_{L}} \Big (B_{_{\Sigma _{_{E}}^{{\mathbf {\Delta }}}}}\!\big ({\scriptstyle \mathcal {L}_{_{{\mathbf {\Delta }_U},{\mathbf {\Delta }_U}}}^{{R,E}}\!(t_0,T)}, R\big ) \Big ) \subset B_{_{\Sigma _{_{E}}^{{\mathbf {\Delta }}}}}\!\big ({\scriptstyle \mathcal {P}_{{U}}^{R}(t_0,T)}, {2}R\big ). \end{aligned}$$
(7.2)

Here, through a slight abuse of notation, we write \(\mathcal {L}^{R,E}_{_{{\mathbf {\Delta }_U},{\mathbf {\Delta }_U}}}\) for (1.5) with \(S^*_xM\) and \(S^*_yM\) replaced by \({\mathbf {\Delta }_U}:=N^*\mathbf{\Delta } \cap (U\times T^*M)\) and \(\varphi _t=\exp (tH_{\textbf{p}})\). To prove (7.2) suppose \(\rho _0\in B_{_{\Sigma _{_{E}}^{{\mathbf {\Delta }}}}}\!\big ({\scriptstyle \mathcal {L}_{_{{\mathbf {\Delta }_U},{\mathbf {\Delta }_U}}}^{{R,E}}\!(t_0,T)}, R\big )\). Then, there are \(\rho _1 \in \Sigma _{_{E}}^{\mathbf{\Delta }}{\cap {\mathbf {\Delta }_U}}\) and \(\rho _1'\in T^*(M\times M)\) such that

$$\begin{aligned} {\tilde{d}}(\rho _0, \rho _1)<R,\quad {{\tilde{d}}(\rho _1,\rho _1')<R}, \quad \text {and} \quad \bigcup _{t_0\le |t|\le T}\varphi _t^{\textbf{p}}({\rho _1'})\cap B({\scriptstyle \Sigma _{_{E}}^{\mathbf{\Delta }}},R)\ne \emptyset . \end{aligned}$$

Therefore, there is \(\rho _2 \in \Sigma _{_{E}}^{\mathbf{\Delta }}\) such that \({{\tilde{d}}}\big (\varphi _t^{\textbf{p}}(\rho _1'), \rho _2\big )<R\) for some \(t_0\le |t|\le T\). Let \(\rho _1'=(x',\xi ',y',-\eta ')\) with \((x',\xi '),(y',\eta ')\in T^*M\). Then, since \(\rho _1=(x,\xi ,x,-\xi )\) and \(\rho _2=(y,\eta ,y,-\eta )\) for some \((x,\xi )\in T^*M\) and \((y,\eta )\in T^*M\), we have \(d(\varphi _t(x',\xi '),(x',\xi '))<4R\) and \(\pi _{_{L}}({\rho _1'})=({x',\xi '})\in \mathcal {P}^{{4}R}_{_{U}}(t_0,T).\) On the other hand, since \(d(\pi _{_{L}}(\rho _0),\pi _{_{L}}({\rho '_1}))<{2}R\) we obtain \(\pi _{_{L}}(\rho _0)\in B_{_{S^*\!M}}\!\big ({\scriptstyle \mathcal {P}_U^{{4}R}(t_0,T)}, {2}R\big )\). This proves claim (7.2).

Next, note that since \(\pi _{_{L}}:{{\mathbf {\Delta }_U}}\cap { \Sigma _{_{E}}^{{\mathbf {\Delta }}} \rightarrow \{p=E\}\cap U}\) is a diffeomorphism for \(E\in [a,b]\), it follows that there exists \(C={C(p)}>0\) such that for all \(E \in [a,b]\)

$$\begin{aligned} \mu _{_{E}}\Big ( B_{_{\Sigma _{_{E}}^{{\mathbf {\Delta }}}}}\!\big ({\scriptstyle \mathcal {L}_{_{{\mathbf {\Delta }_U},{\mathbf {\Delta }_U}}}^{{R,E}}\!(t_0,T)}, R\big )\Big )\le C \mu _{_{{S^*M}}}\Big (B_{_{S^*\!M}}\!\big ({\scriptstyle \mathcal {P}_U^{{4}R}(t_0,T)}, {2}R\big )\Big ). \end{aligned}$$

Hence, if U is \((t_0,{\textbf{T}})\) non-periodic for p at energy E, we have

$$\begin{aligned}{} & {} \mu _{_{E}}\Big ( B_{_{\Sigma _{_{E}}^{{\mathbf {\Delta }}}}}\!\big ({\scriptstyle \mathcal {L}_{_{{\mathbf {\Delta }_U},{\mathbf {\Delta }_U}}}^{{R,E}}\!(t_0,{\textbf{T}({4}R))}}, R\big )\Big )\,{\textbf{T}}({4}R)\\ {}{} & {} \le C\mu _{_{{S^*M}}}\Big (B_{_{S^*\!M}}\!\big ({\scriptstyle \mathcal {P}_U^{{4}R}(t_0,{\textbf{T}({4}R)})}, {4}R\big )\Big )\,{{\textbf{T}}({4}R)}\le CC_{_{\!{\text {np}}}}, \end{aligned}$$

and so \({\mathbf {\Delta }_U}\) is \((t_0,\textbf{T}({4}R))\) non-looping for \(\textbf{p}\) at energy E. The result follows from Corollary 3.1. \(\square \)

In what follows, we write \(\Vert \cdot \Vert _{_{HS}}\) for the Hilbert-Schmidt norm.

Lemma 7.2

Let \(\mathcal {V}\subset S_\delta (T^*M;[0,1])\) be a bounded subset. Then, there are \(C>0\) and \(h_0>0\), and for all \(N>0\) there exists \(C_{_{N}}>0\), such that for all \(t \in [a,b]\), \(\chi \in \mathcal {V}\), \(0<h<h_0\), and \(|s|\le 2h\),

$$\begin{aligned} \Vert \mathbb {1}_{[t-s,t]}(P)Op_h(\chi )\Vert ^2_{_{HS}}\le C h^{1-n} {\mu _{_{{p^{-1}(t)}}}}({{\,\textrm{supp}\,}}\chi \cap p^{-1}(t)) +C_{_{N}}h^N, \nonumber \\ \end{aligned}$$
(7.3)
$$\begin{aligned} h^{-2}\Vert P_t\mathbb {1}_{[t-s,t]}(P)Op_h(\chi )\Vert ^2_{_{HS}}\le C h^{1-n}{\mu _{_{{p^{-1}(t)}}}}({{\,\textrm{supp}\,}}\chi \cap p^{-1}(t)) +C_{_{N}}h^N. \nonumber \\ \end{aligned}$$
(7.4)

Proof

We follow the proof of [18, Lemma 3.11]. Let \(\psi \in \mathcal {S}({\mathbb {R}})\) with \(\psi (0)=1\) and \({{\,\textrm{supp}\,}}{\hat{\psi }}\subset [-1,1]\). Define \(\psi _\varepsilon (s):=\psi (\varepsilon s).\) Then, there is \(\varepsilon _0>0\) small enough so that \(\psi _{\varepsilon _0}(s)>\frac{1}{2}\) on \([-2,2]\). Abusing notation slightly, put \(\psi =\psi _{\varepsilon _0}\). Then, there exists an operator \(Z_s\) such that \( \mathbb {1}_{[t-s,t]}(P)={Z_s}\psi \big (\tfrac{P_t}{h}\big ), \) \([Z_s,P]=0\), and \(\Vert Z_s\Vert _{L^2\rightarrow L^2}\le 3\) for \(|s|\le 2h\). Therefore, \( \Vert \mathbb {1}_{[t-s,t]}(P)Op_h(\chi )\Vert _{_{HS}}\le {3}\big \Vert \psi \big (\tfrac{P_t}{h}\big )Op_h(\chi )\big \Vert _{_{HS}} \) and the Hilbert–Schmidt norm is the \(L^2\) norm of the kernel. Next, we recall that after application of a microlocal partition of unity, we may write

$$\begin{aligned} \psi \Big (\tfrac{P_t}{h}\Big )(x,y)=h^{-n}\int _{\mathbb {R}}\int _{{\mathbb {R}}^n} {\hat{\psi }}(\tau )e^{\frac{i}{h}(\varphi (\tau ,x,\eta )-\langle y,\eta \rangle -t\tau )}a(\tau ,x,y,\eta )d\eta d\tau +O(h^\infty )_{_{HS}} \end{aligned}$$

for a symbol \(a\sim \sum _j h^ja_j\) and phase \(\varphi \) solving \( \partial _t\varphi =p(x,\partial _x\varphi ) \) and \( \varphi (0,x,\eta )=\langle x,\eta \rangle . \) At this point the proof of (7.3) follows exactly as in [18, Lemma 3.11].

To obtain (7.4), we write \( P_t\mathbb {1}_{[t-s,t]}(P)=Z_sP_t\psi \big (\tfrac{P_t}{h}\big ) \) and note that \( \tfrac{P_t}{h}\psi \big (\tfrac{P_t}{h}\big )=(t\psi )\big (\tfrac{P_t}{h}\big ). \) Hence the same argument applies with \(\widehat{t\psi }(\tau )=-i\partial _\tau {\hat{\psi }}(\tau )\) replacing \({\hat{\psi }}(\tau )\). \(\square \)

We will also need the following trace bound for \(\mathbb {1}_{[t-s,t]}\).

Lemma 7.3

Suppose \(a,b\in {\mathbb {R}}\), \(\varepsilon _0>0\), \(\ell _1,\ell _2\in {\mathbb {R}}\), \(\mathcal {V}_1\subset \Psi ^{\ell _1}(M)\), and \(\mathcal {V}_2\subset \Psi _\delta ^{\ell _2}(M)\) bounded subsets, \(U\subset T^*M\) open such that \(|d\pi _{_{M}}{{\textsf{H}}_p}|>c >0\) on \(p^{-1}([a-\varepsilon _0,b+\varepsilon _0])\cap U\). Let \(\tau _0\),\(R_0\), \(\delta \), R(h), and \(\tau \) be as in Lemma 4.1. Let \(\{{\mathcal {T}}_j\}_{j\in \mathcal {J}(h)}\) be a \(({\mathfrak {D}},\tau ,R(h))\) good covering of \({\textbf{p}}^{-1}([a,b])\cap N^*{\mathbf {\Delta }}{\cap (U\times T^*M)}\) and \(\mathcal {V}\subset S_\delta (T^*M\times T^*M; [0,1])\) bounded. Then, there is \(C_{_{0}}>0\) such that for all \(\{\chi _{_{{\mathcal {T}}_j}}\}_{j\in \mathcal {J}(h)}\subset \mathcal {V}\) partitions for \(\{{\mathcal {T}}_j\}_{j\in \mathcal {J}(h)}\), \(j \in \mathcal {J}(h)\), \(A_1\in \mathcal {V}_1\), \(A_2\in \mathcal {V}_2\), and \(|s|\le \varepsilon _0\)

$$\begin{aligned} \Big |\int _{{\mathbf {\Delta }}} Op_h(\chi _{_{{\mathcal {T}}_j}})A_1\mathbb {1}_{[t-s,t]}(P)A_2 d\sigma _{_{\!{\mathbf {\Delta }}}}\Big |\le C_{_{0}}h^{1-n}R(h)^{2n-1}\Big \langle \frac{s}{h}\Big \rangle . \end{aligned}$$

Proof

We first note that it suffices to prove the statement for \(|s|\le 2h\). Indeed, this is because we may apply the arguments from Lemma 5.4 and decompose \( \mathbb {1}_{[t-s,t]}(P)=\sum _{k=0}^{k_0-1}\mathbb {1}_{[t_k,t_{k+1}]}(P), \) with \(|t_{k+1}-t_k|\le 2h\). This allows us to obtain the result for \(|s|\le \varepsilon _0\).

Suppose \(|s|\le 2h\). Let \({{\tilde{U}}} \supset {B(U,2R(h))}\), \(j \in \mathcal {J}(h)\), and \(A:=Op_h(\chi _{_{{\mathcal {T}}_j}})(A_1 \otimes A_2)\). Note that

$$\begin{aligned}{}[{\mathbf{P_t}},A]= & {} [{\mathbf{P_t}},Op_h(\chi _{_{{\mathcal {T}}_j}})](A_1\otimes A_2)\nonumber \\{} & {} +Op_h(\chi _{_{{\mathcal {T}}_j}})[P-t,A_1]\otimes A_2\in \Psi _\delta (M) \end{aligned}$$
(7.5)

with seminorms bounded by those of \(\chi _{_{{\mathcal {T}}_j}}\), \(A_1\), and \(A_2\). We next apply Lemma 4.1 with \(A:=Op_h(\chi _{_{{\mathcal {T}}_j}})(A_1 \otimes A_2)\), \(\mathbf{P_t}\) in place of \(P_t\), \(k=n\), \(M\times M\) in place of M, and \( u:=\mathbb {1}_{[t-s,t]}(P){Op_h(\chi _{_{{\tilde{U}}}})}, \) where the latter is viewed as a kernel on \(M \times M\). Here, \(\chi _{_{{\tilde{U}}}}\in S_\delta (T^*M)\) with \(\chi _{_{{\tilde{U}}}}\equiv 1\) on B(UR(h)), \({{\,\textrm{supp}\,}}\chi _{_{{\tilde{U}}}}\subset {\tilde{U}}\). Let \({\tilde{\chi }}_{_{{\mathcal {T}}_j}}\in \mathcal {V}\) with \({{\,\textrm{supp}\,}}{\tilde{\chi }}_{_{{\mathcal {T}}_j}}\subset {\mathcal {T}}_j\) and \({\tilde{\chi }}_{_{{\mathcal {T}}_j}}\equiv 1\) on \({{\,\textrm{supp}\,}}\chi _{_{{\mathcal {T}}_j}}\). Then, since \({\hbox {MS}}_{\textrm{h}}(A) \subset {\mathcal {T}}_j\), by Lemma 4.1 there exist \(C_{_{0}}>0\) and \(C>0\), such that

$$\begin{aligned}{} & {} h^{\frac{n-1}{2}}\Big |\int _{{\mathbf {\Delta }}} Op_h(\chi _{_{{\mathcal {T}}_j}})A_1\mathbb {1}_{[t-s,t]}(P)A_2 d\sigma _{_{\!{\mathbf {\Delta }}}}\Big | \\ {}{} & {} \le C_{_{0}}{R(h)^{\frac{2n-1}{2}}} \Big ({\Vert Op_h({{\tilde{\chi }}}_{_{{\mathcal {T}}_j}})u\Vert _{_{\!L^2(M)}}}+\frac{C}{h}\Vert Op_h({{\tilde{\chi }}}_{_{{\mathcal {T}}_j}}){\textbf{P}}_{t}u\Vert _{_{\!L^2(M)}}\Big ). \end{aligned}$$

Note that we omit the analogous error terms appearing in the estimate of Lemma 4.1 since these error terms can be dealt with by applying the bounds in (5.7) and (5.8) in combination with (7.1).

Next, since \(Op_h({\tilde{\chi }}_{_{{\mathcal {T}}_j}})=Op_h({\tilde{\chi }}_{_{{\widetilde{{\mathcal {T}}}}_j}})\otimes Op_h({\tilde{\chi }}_{\rho _j})\), where \({\tilde{\chi }}_{\rho _j}\) and \({\tilde{\chi }}_{_{{\widetilde{{\mathcal {T}}}}_j}}\) are bounded in \(S_\delta (T^*M;[0,1])\) by the seminorms in the set \(\mathcal {V}\), we obtain

$$\begin{aligned}&h^{\frac{n-1}{2}}R(h)^{-\frac{2n-1}{2}}\Big |\int _{{\mathbf {\Delta }}} Op_h(\chi _{_{{\mathcal {T}}_j}})A_1\mathbb {1}_{[t-s,t]}(P)A_2d\sigma _{_{\!{\mathbf {\Delta }}}}\Big | \\&\quad \le C_{_{0}}\Vert Op_h({\tilde{\chi }}_{_{{\widetilde{{\mathcal {T}}}}_j}})u Op_h({\tilde{\chi }}_{\rho _j})\Vert _{_{HS}}+ C_{_{0}}Ch^{-1}\Vert Op_h({\tilde{\chi }}_{_{{\widetilde{{\mathcal {T}}}}_j}}) {P}_{t}uOp_h({\tilde{\chi }}_{\rho _j})\Vert _{_{HS}}\\&\quad \le C_{_{0}}h^{\frac{1-n}{2}}R(h)^{\frac{2n-1}{2}}, \end{aligned}$$

where u is now viewed as an operator. In the last line we used Lemma 7.2 and the existence of \(C>0\) such that \( \mu _t\Big (({{\,\textrm{supp}\,}}{\tilde{\chi }}_{\rho _j})\cap p^{-1}(t)\Big )\le CR(h)^{2n-1}. \) This finishes the proof when \(|s|\le 2h\). \(\square \)

Lemma 7.4

Let \(a,b,\varepsilon _0\), \(\tau _0\), \(\mathcal {V}_1,\mathcal {V}_2\) \(R_0\), \(\tau \), \(\delta \), R(h) and \(\alpha \) as in Lemma 4.4. Let \({N^*{\mathbf {\Delta }}\cap (U\times T^*M)}\) be \(\textbf{T}\) non-looping for \({\textbf{p}}\) in the window [ab] via \(\tau \)-coverings and let \(C_{_{\!{\text {np}}}}\) be the constant \(C_{_{\!{\text {nl}}}}\) in Definition 2.1. Then, there is \(C_{_{0}}=C_{_{0}}(n,P, {\mathcal {V}_1, \mathcal {V}_2}, C_{_{\!{\text {np}}}},\varepsilon _0)>0\) and for all \(K>0\) there is \(h_0>0\) such that for all \(0<h\le h_0\), \(A_1\in \mathcal {V}_1\), \(A_2 \in \mathcal {V}_2\) with \({{\hbox {MS}}_{\textrm{h}}(A_2)\subset U}\), \(|s|\le 2h\), and \(t\in [a-Kh,b+Kh]\),

$$\begin{aligned} h^{n-1}\Big |\int _{{\mathbf {\Delta }}}A_1\mathbb {1}_{[t-s,t]}(P)A_2d\sigma _{_{\!{\mathbf {\Delta }}}} \Big |^2\le C_{_{0}} \frac{1}{T(h)}\Big \langle \frac{T(h)s}{h}\Big \rangle \Vert \mathbb {1}_{[t-s,t]}(P){Op_h(\chi _{_{{\tilde{U}}}})}\Vert ^2_{L^2}, \end{aligned}$$

where \({\tilde{U}}(h)\supset B(U,2R(h))\), \(\chi _{_{{\tilde{U}}}}\in S_\delta \), \(\chi _{_{{\tilde{U}}}}\equiv 1\) on B(UR(h)), and \({{\,\textrm{supp}\,}}\chi _{_{{\tilde{U}}}}\subset {\tilde{U}}\).

Proof

Since \(N^*{\mathbf {\Delta }} \cap (U\times T^*M)\) is \(\textbf{T}\) non-looping in the window [ab] via \(\tau _0\)-coverings, for all \(t\in [a-Kh,b+Kh]\), there is a partition of indices \(\mathcal {J}_{_{t}}(h)={{\mathcal {G}}}_{_{t,0}}(h)\sqcup {{\mathcal {G}}}_{_{t,1}}(h)\) as described in Definition 2.1 (with \(H={\mathbf {\Delta }}\)). Let \(t_0=t_0\), \(t_1=1\), \(T_0(h)=T(h)\) and \(T_1(h)=1\). Then, there is \(C_{_{\!{\text {np}}}}>0\) such that for all \(t\in [a-Kh,b+Kh]\)

$$\begin{aligned}{} & {} \sum _{\ell =0}^1\sqrt{\frac{|{{\mathcal {G}}}_{_{t,\ell }}(h)|t_\ell }{T_\ell }}\le \frac{C_{_{\!{\text {np}}}}R(h)^{\frac{1-2n}{2}}}{\sqrt{T(h)}}, \nonumber \\{} & {} \sum _{\ell =0}^1\sqrt{|{{\mathcal {G}}}_{_{t,\ell }}(h)| t_\ell T_\ell }\le C_{_{\!{\text {np}}}}R(h)^{\frac{1-2n}{2}}\sqrt{T(h)}. \end{aligned}$$
(7.6)

Next, we argue as in (5.5), and then apply a combination of Lemma 4.3 and Lemma 4.4 with \(A:=A_1 \otimes A_2\), \(\mathbf{P_t}\) in place of \(P_{_{E}}\), 2n in place of n, \(M\times M\) in place of M, \(k=n\), and \(u:=\mathbb {1}_{[t-s,t]}(P){Op_h(\chi _{_{{\tilde{U}}}})}\), where u is viewed as a kernel on \(M \times M\). Then, there is \(C_{_{0}}>0\) so that

$$\begin{aligned}&h^{\frac{n-1}{2}}\Big |\int _{{\mathbf {\Delta }}} \!\!\!A_1\mathbb {1}_{[t-s,t]}(P)A_2 d\sigma _{_{\!{\mathbf {\Delta }}}}\Big |\\&\quad \le C_{_{0}}R(h)^{\frac{2n-1}{2}} \bigg (\sum _{\ell =0}^1\!\!\frac{(|{{\mathcal {G}}}_{_{t,\ell }}(h)|{{\tilde{t}}}_\ell )^{\frac{1}{2}}}{(\tau {{\tilde{T}}}_\ell )^{\frac{1}{2}}}\Vert u\Vert _{L^2} +\!\!\sum _{\ell }\!\!\frac{(|{{\mathcal {G}}}_{_{t,\ell }}(h)| {\tilde{t}}_\ell {\tilde{T}}_\ell )^{\frac{1}{2}}}{h}\Vert {\textbf{P}}_tu\Vert _{L^2}\!\bigg ), \end{aligned}$$

where \({\tilde{t}}_\ell \) and \({\tilde{T}}_\ell \) are as in (5.4). We have used that, since \({\hbox {MS}}_{\textrm{h}}(A)\subset U \times T^*M\) and the tubes are a covering for \({\textbf{p}}^{-1}([a,b])\cap N^*{\mathbf {\Delta }}{\cap (U\times T^*M)}\), then \({\hbox {MS}}_{\textrm{h}}(A)\cap \Lambda _{\Sigma ^{{\mathbf {\Delta }}}_t}^\tau (R(h)/2) \subset \bigcup _{j \in \mathcal {J}_{_{t}}(h)}{\mathcal {T}}_j\). Also, note that we omit the analogous error terms appearing in the estimate of Lemma 4.4 since these error terms can be dealt with by applying the bounds in (5.7) and (5.8) in combination with (7.1).

The proof follows from applying the bounds in (5.5) in combination with (7.1). \(\square \)

Lemma 7.5

Let \(\ell _i \in {\mathbb {R}}\), \(\mathcal {V}_i\subset \Psi _\delta ^{\ell _i}(M)\) bounded for \(i=1,2\). Then, there are \(N_0>0\), \(C>0\), \(h_0>0\) such that for all \(A_1\in \mathcal {V}_1\) and \(A_2\in \mathcal {V}_2\), \(s\in {\mathbb {R}}\) and \(0<h<h_0\)

$$\begin{aligned} \Big |\int A_1\mathbb {1}_{(-\infty ,s]}(P)A_2 d\sigma _{_{\!{\mathbf {\Delta }}}}\Big |\le Ch^{-\frac{n}{2}}\langle s\rangle ^{N_0}\Vert \mathbb {1}_{(-\infty ,s]}(P)\Vert _{L^2}. \end{aligned}$$

Proof

We apply Lemma 4.5 with \(H={\mathbf {\Delta }}\), \(A=A_1 \otimes A_2\), and \(u=\mathbb {1}_{(-\infty ,s]}(P)\). Then, for \(r>\frac{n+2(\ell _1+\ell _2)}{2m}\), there is \(C>0\) such that for all \(N>0\) there is \(C_N>0\) such that

$$\begin{aligned}&h^{\frac{n}{2}}\Big |\int _{{\mathbf {\Delta }}}A_1\mathbb {1}_{(-\infty ,s]}(P)A_2 d\sigma _{_{\!{\mathbf {\Delta }}}}\Big | \\&\quad \le C(\Vert \mathbb {1}_{(-\infty ,s]}(P)\Vert _{L^2}+\Vert {\textbf{P}}^r \mathbb {1}_{(-\infty ,s]}(P)\Vert _{L^2})+C_Nh^N\Vert \mathbb {1}_{(-\infty ,s]}(P)\Vert _{L^2}. \end{aligned}$$

It follows from (7.1) that the last term can be bounded by \(C(1+|s|^r) \Vert \mathbb {1}_{(-\infty ,s]}(P)\Vert _{L^2}\). \(\square \)

7.1 Proofs of Theorems 2 and 6

We claim that for \({E}\in [a-Kh,b+Kh]\) and \(A_1\in \mathcal {V}_1\), and \(A_2\in \mathcal {V}_2\) with \({\hbox {MS}}_{\textrm{h}}(A_2)\subset U\),

$$\begin{aligned} h^{n-1}\Big |\int _{{\mathbf {\Delta }}}A_1\Big (\mathbb {1}_{(-\infty ,E]}(P)-\big (\rho _{h,t_0}*\mathbb {1}_{(-\infty ,\cdot ]}(P)\big )(E)\Big )A_2d\sigma _{_{\!{\mathbf {\Delta }}}}\Big |\le C_0\big /T(h). \nonumber \\ \end{aligned}$$
(7.7)

We start by showing under the same assumptions that

$$\begin{aligned}{} & {} h^{n-1}\Big |\int _{{\mathbf {\Delta }}}A_1\Big (\big (\rho _{_{h,T(h)}}*\mathbb {1}_{(-\infty ,\,\cdot \,]}(P) \big )({E})-\mathbb {1}_{(-\infty ,{E}]}(P)\Big )A_2 d\sigma _{_{\!{\mathbf {\Delta }}}}\Big | \nonumber \\{} & {} \quad \le C_0\big /T(h), \end{aligned}$$
(7.8)
$$\begin{aligned}{} & {} h^{n-1}\Big |\int _{{\mathbf {\Delta }}} A_1\Big ( \big (\rho _{_{h,T(h)}}*\mathbb {1}_{(-\infty ,\,\cdot \,]}(P)\big ){(E)} - \big (\rho _{_{h,t_0}}*\mathbb {1}_{(-\infty ,\,\cdot \,]}(P)\big ){(E)}\Big )A_2d\sigma _{_{\!{\mathbf {\Delta }}}} \Big |\nonumber \\{} & {} \quad \le C_0\big /T(h). \end{aligned}$$
(7.9)

for some \(t_0\) independent of h. At the end of the section we will derive Theorems 2 and 6 from (7.7).

7.1.1 Proof of (7.8).

Let \({{\tilde{U}}},U_0 \subset T^*M\) with \(B(U_0,2R(h))\subset U\subset B(U_0,4R(h)) \subset {{\tilde{U}}}\). Then, let \(\chi _{_{{\tilde{U}}}},\chi _{_{U_0}},\chi _{_{{\tilde{U}}\setminus {U_0}}}\in S_\delta (T^*M;[0,1])\) with \(\chi _{_{{\tilde{U}}}}\equiv 1\) on U, \({{\,\textrm{supp}\,}}\chi _{_{{\tilde{U}}}}\subset B(U_0,3R(h))\), \(\chi _{_{U_0}}\equiv 1\) on \(B(U_0, R(h))\), \({{\,\textrm{supp}\,}}\chi _{_{U_0}}\subset U\), \(\chi _{_{{\tilde{U}}\setminus {U_0}}}\equiv 1\) on \({{\,\textrm{supp}\,}}\chi _{_{{\tilde{U}}}}(1-\chi _{_{U_0}})\), \({{\,\textrm{supp}\,}}\chi _{_{{\tilde{U}}\setminus {U_0}}}\subset {\tilde{U}}\setminus U_0\). By Lemma 7.2 and (1.12) there exists \(C_0>0\) such that for \(|s|\le 2h\),

$$\begin{aligned}{} & {} h^{n-1}\big \Vert \mathbb {1}_{[t-s,t]}(P)Op_h(\chi _{_{{\tilde{U}}\setminus U_0}})\big \Vert _{_{HS}}^2\nonumber \\{} & {} \quad \le C_0\mu _{_{{p^{-1}(t)}}}(p^{-1}(t)\cap ({\tilde{U}}\setminus U_0)) \le C_0 C_{_{U}}\big /T(h). \end{aligned}$$
(7.10)

Note that when \(U=T^*M\) this is an empty statement. Then, for \({|s|\le 2h}\), by Lemma 7.4

$$\begin{aligned}&h^{n-1}{\text {tr}}\Big (\mathbb {1}_{[t-s,t]}(P){Op_h(\chi _{_{U_0}})}\Big )^2 \Big (\frac{1}{T(h)}\Big \langle \frac{T(h)s}{h}\Big \rangle \Big )^{-1}\le C_{_{0}}\Vert \mathbb {1}_{[t-s,t]}(P)Op_h(\chi _{_{{\tilde{U}}}})\Vert ^2_{L^2}\\&\le C_{_{0}}{\text {tr}}\mathbb {1}_{[t-s,t]}(P)Op_h(\chi _{_{U_0}})+C_{_{0}}\Vert \mathbb {1}_{[t-s,t]}(P)Op_h(\chi _{_{{\tilde{U}}\setminus U_0}})\Vert _{_{HS}}^2 +C_{_{N}}h^N. \end{aligned}$$

Then, applying the quadratic formula with \(x={\text {tr}}\mathbb {1}_{[t-s,t]}(P){Op_h(\chi _{_{U_0}})}\), for \(|s|\le 2h\) we have

$$\begin{aligned} 0\le h^{n-1}{\text {tr}}\mathbb {1}_{[t-s,t]}(P){Op_h(\chi _{_{U_0}})}&\le \frac{C_0}{T(h)}\Big \langle \frac{T(h)s}{h}\Big \rangle + \frac{C_{_{U}}C_0}{T(h)}+c_Nh^N. \end{aligned}$$

Next, for \(|s|\le \varepsilon _0\), splitting \(\mathbb {1}_{[t-s,t]}(P)=\sum _{k=0}^{k_0-1}\mathbb {1}_{[t_k,t_{k+1}]}(P)\) as before, we have by Lemma 7.4 and Lemma 7.5 that there exists \(N_0>0\) such that

$$\begin{aligned}{} & {} h^{n-1}\Big |\int _{{\mathbf {\Delta }}}A_1\mathbb {1}_{[t-s,t]}(P)A_2d\sigma _{_{\!{\mathbf {\Delta }}}}\Big |\le C_{_{0}}\frac{1}{T(h)}\Big \langle \frac{T(h)s}{h}\Big \rangle , \end{aligned}$$
(7.11)
$$\begin{aligned}{} & {} h^{\frac{n}{2}}\Big |\int _{{\mathbf {\Delta }}}A_1\mathbb {1}_{(-\infty ,s]}(P)A_2d\sigma _{_{\!{\mathbf {\Delta }}}}\Big |\le C\langle s \rangle ^{N_0}\Vert \mathbb {1}_{(-\infty ,s]}(P)\Vert _{L^2}\le Ch^{-\frac{n}{2}}(1+|s|^{2N_0}), \end{aligned}$$
(7.12)

where to get the last inequality, we use Lemma 7.5 with \(U=M\), \(A_1=A_2={\text {Id}}\).

In particular, combining  (7.11) and (7.12) together with Lemma 5.3 implies (7.8) holds.

7.1.2 Proof of (7.9).

Using Lemma 6.3, the proof of (7.9) amounts to understanding

$$\begin{aligned}{} & {} A_1\big (({\rho _{_{_{h,{{\widetilde{T}}(h)}}}}}-\rho _{_{h,t_0}})*\mathbb {1}_{(-\infty ,\,\cdot \,]}(P)\big ){(E)}A_2 \\{} & {} \quad = A_1 {f_{t_0,{\widetilde{T}}(h),h}}\big (P_{_{E}}\big )A_2+O(h^\infty )_{H_{scl }^{-N}\rightarrow H_{scl }^N}, \end{aligned}$$

where \(f_{S,T,h}\) is given by (6.2), and \({\tilde{T}}(h)=\frac{T(h)}{2}\). In particular, for \(E\in [a-Kh,b+Kh]\), we consider \( {\text {tr}}A_1{f_{t_0,{\tilde{T}},h}\big (P_E\big )}A_2.\) For this, we let \(\{{\mathcal {T}}_j\}_{j\in \mathcal {J}(h)}\) be a \(({\mathfrak {D}},\tau ,R(h))\)-good covering of \({\textbf{p}}^{-1}([a,b])\cap N^*{\mathbf {\Delta }}\cap (U\times T^*M)\) and \(\mathcal {V}\subset S_\delta (T^*M\times M;[0,1])\) a bounded subset. Let \(\{\chi _{_{{\mathcal {T}}_j}}\}_{j\in \mathcal {J}(h)}\subset \mathcal {V}\) be a partition associated to \(\{{\mathcal {T}}_j\}_{j\in \mathcal {J}_{_{E}}(h)}\).

Lemma 7.6

Let \(\mathcal {I}\subset \mathcal {J}_{_{E}}(h)\), \(\mathcal {V}_1\subset \Psi ^{\ell _1}(M)\), \(\mathcal {V}_2\subset \Psi _\delta ^{\ell _2}(M)\) bounded subsets. Then, there exist \(C_{_{0}}>0\) and \(h_0>0\) such that for all \(A_1\in \mathcal {V}_1\), \(A_2\in \mathcal {V}_2\), \(0<h<h_0\)

$$\begin{aligned} \Big |\int _{{\mathbf {\Delta }}} \sum _{j\in \mathcal {I}} Op_h(\chi _{_{{\mathcal {T}}_j}})A_1 {f_{t_0,{\tilde{T}},h}\big (P_{_{E}}\big )}A_2d\sigma _{_{\!{\mathbf {\Delta }}}}\Big |\le C_{_{0}} h^{1-n}R(h)^{2n-1}|\mathcal {I}|. \end{aligned}$$

Proof

We first note that \( {f_{t_0,{\widetilde{T}}(h),h}(P_{_{E}})}= {\varrho _h}*\partial _s \mathbb {1}_{(-\infty ,\,\cdot \,]}({P})({E}), \) where \( {\varrho _h(s)}:={f_{t_0,{\widetilde{T}}(h),h}(-s)}. \) Then, since \(\widehat{f_{t_0,{\widetilde{T}}(h)}}(0)=0\), we have \(\int _{\mathbb {R}}\partial _s\varrho _h(s)ds=0\). In particular, by the estimates (6.3), Lemma 5.3 applies with \(\sigma _h=h^{-1}\). Note that by Lemma 7.3, for \(t\in [a-Kh,b+Kh]\), and \(|s|\le 1\),

$$\begin{aligned} \Big |\int _{{\mathbf {\Delta }}}Op_h(\chi _{_{{\mathcal {T}}_j}})A_1(\mathbb {1}_{(-\infty ,t]}-\mathbb {1}_{(-\infty ,t-s]})A_2 d\sigma _{_{\!{\mathbf {\Delta }}}}\Big |\le Ch^{1-n}R(h)^{2n-1}\Big \langle \frac{s}{h}\Big \rangle . \nonumber \\ \end{aligned}$$
(7.13)

Also, by Lemma 7.5, there exists \(N_0\) such that for \(s\in {\mathbb {R}}\),

$$\begin{aligned} \Big |\int _{{\mathbf {\Delta }}}Op_h(\chi _{_{{\mathcal {T}}_j}})A_1\mathbb {1}_{(-\infty ,s]}(P)A_2 d\sigma _{_{\!{\mathbf {\Delta }}}}\Big |\le Ch^{-n}\langle s\rangle ^{N_0}. \end{aligned}$$
(7.14)

The proof follows from Lemma 5.3 using (7.13) and (7.14), and by summing in \(j\in {\mathcal {I}}\). \(\square \)

Lemma 7.7

Let \(\mathcal {V}_1, \mathcal {V}_2\) as in Lemma 7.6 and suppose \({\mathcal {T}}_j\) is a tube such that \({\tilde{{\mathcal {T}}}}_j\), its corresponding tube in \(T^*M\), satisfies \( \varphi _t({\tilde{{\mathcal {T}}}}_j)\cap {\tilde{{\mathcal {T}}}}_j=\emptyset \) for \(|t|\in [t_0,T(h)].\) Then for all \(N>0\) there is \(C_{_{N}}>0\) such that for all \(A_1\in \mathcal {V}_1\), and \(A_2\in \mathcal {V}_2\),

$$\begin{aligned} \Big |\int _{{\mathbf {\Delta }}} Op_h(\chi _{_{{\mathcal {T}}_j}})A_1{f_{t_0,{\tilde{T}},h}\big (P_{_{E}}\big )}A_2 d\sigma _{_{\!{\mathbf {\Delta }}}}\Big |\le C_{_{N}}h^N. \end{aligned}$$

Proof

Note that the assumption on \({\tilde{{\mathcal {T}}}}_j\) implies \( \exp (tH_{\textbf{p}})({\mathcal {T}}_j)\cap N^*{\mathbf {\Delta }} =\emptyset \) for \( |t|\in [t_0,T(h)]. \) Therefore, the same application of Egorov’s theorem as in Lemma 6.4, completes the proof. \(\square \)

Since U is \(\textbf{T}\) non-periodic in the window [ab] via \(\tau \)-coverings, for all \(E\in [a-Kh,b+Kh]\), there is a splitting \( \mathcal {J}_{_{E}}(h)=\mathcal {B}_{_{E}}(h)\cup \mathcal {G}_{_{E}}(h) \) such that \(\varphi _t({\tilde{{\mathcal {T}}}}_j)\cap {\tilde{{\mathcal {T}}}}_j=\emptyset \) for \(|t|\in [t_0,T(h)]\) for \(j\in \mathcal {G}_{_{E}}(h)\), and \( |\mathcal {B}_{_{E}}(h)|R(h)^{2n-1}\le T^{-1}(h). \) We write, using \({\hbox {MS}}_{\textrm{h}}(A_1\otimes A_2)\cap \Lambda _{\Sigma ^{{\mathbf {\Delta }}}_t}^\tau (R(h)/2)\subset \bigcup _{j\in \mathcal {J}_{h,E}}{\mathcal {T}}_j,\)

$$\begin{aligned}&\int _{{\mathbf {\Delta }}} A_1{f_{t_0,{\tilde{T}},h}\big (P_{_{E}}\big )}A_2 d\sigma _{_{\!{\mathbf {\Delta }}}} \\&\quad =\sum _{{j\in \mathcal {G}_{_{E}}(h) \cup \mathcal {B}_{_{E}}(h)}}\int _{{\mathbf {\Delta }}} Op_h(\chi _{_{{\mathcal {T}}_j}})A_1{f_{t_0,{\tilde{T}},h}\big (P_{_{E}}\big )}A_2 d\sigma _{_{\!{\mathbf {\Delta }}}} +O(h^\infty ). \end{aligned}$$

Applying Lemma 7.7 to the sum over \(\mathcal {G}_{_{E}}(h)\) and Lemma 7.6 to the sum over \(\mathcal {B}_{_{E}}(h)\), we have

$$\begin{aligned} \Big |\int _{{\mathbf {\Delta }}} A_1{f_{t_0,{\tilde{T}},h}\big (P_{_{E}}\big )}A_2 d\sigma _{_{\!{\mathbf {\Delta }}}}\Big |\le C h^{1-n}|\mathcal {B}_{_{E}}(h)|R(h)^{2n-1}+O(h^\infty )\le C\big /T(h) \end{aligned}$$

for any \(E\in [a-Kh,b+Kh]\). In particular (7.9) holds.

7.1.3 Completion of the proof of Theorem 6

In order to complete the proof of Theorem 6, we take \(A_1={\text {Id}}\) and \(A_2=A^{{t}}\) and apply (7.7) to obtain the theorem. \(\square \)

7.1.4 Proof of Theorem 2

We assume \({W}\subset M\) is \(\textbf{T}\) non-periodic and let \(P=Q\) as in (2.14). Then \(|d\pi _{_{M}}{{\textsf{H}}_p}|>c>0\) on \(|\xi |_g>\frac{1}{2}>0\) so we may apply (7.7) for \(E>\frac{1}{2}\). Let \(0<\delta <\tfrac{1}{2}\). Let \(\chi _h\in C_c^\infty (M)\) as in [9, (19)] i.e. such that \(\chi _h\equiv 1\) in a neighborhood of \(\partial {W}\), \({{\,\textrm{supp}\,}}\chi _h\subset \{ d(x,\partial {W})<2h^\delta \}\), \(|\partial _x^\alpha \chi |\le C_\alpha h^{-|\alpha |\delta },\) \({{\,\textrm{vol}\,}}_{_{\!M}}({{\,\textrm{supp}\,}}\chi _h)\le C h^{\delta (n-\dim _{box }\partial {W})}.\)

Let \(R(h)\ge {h^\delta }\), and \(T(h)=\textbf{T}(R(h))\). Then, put \(A_1=1\) and \(A_2=(1-\chi _h)1_{{W}}\) in (7.7) to obtain

$$\begin{aligned} \Big |\int _{{\mathbf {\Delta }}}\Big (\mathbb {1}_{(-\infty ,1]}(P)-\rho _{h,t_0}*\mathbb {1}_{(-\infty ,\cdot ]}(P)({1})\Big )(1-\chi _h)1_{{W}}d\sigma _{_{\!{\mathbf {\Delta }}}}\Big |\le C_0h^{1-n}\big /T(h). \end{aligned}$$

Next, since \( \rho _{h,t_0}*\mathbb {1}_{(-\infty ,\cdot ]}(P)({1})(x,x)=\frac{{{\,\textrm{vol}\,}}_{{\mathbb {R}}^n}(B^n)}{(2\pi h)^{n}}+O(h^{-n+2}) \) (apply Theorem 3 with \(\textbf{T}=1\)),

$$\begin{aligned} \Big |\int _{{W}} (1-\chi _h(x))\Big ({\Pi _h(1,x,x) }-(2\pi h)^{-n}{{\,\textrm{vol}\,}}_{{\mathbb {R}}^n}(B^n)\Big ){{\,\mathrm{{\text {dv}}}\,}}_g(x)\Big |\le C_0h^{1-n}\big /T(h). \end{aligned}$$

Also, since \(\Pi _h(1,x,x)=(2\pi h)^{-n}{{\,\textrm{vol}\,}}_{{\mathbb {R}}^n}(B^n)|= O(h^{1-n})\) (apply Theorem 3 with \(\textbf{T}={{\,\textrm{inj}\,}}M\)),

$$\begin{aligned} \Big |\int _{{W}}\chi _h(x)\Big (\Pi _h(1,x,x)-(2\pi h)^{-n}{{\,\textrm{vol}\,}}_{{\mathbb {R}}^n}(B^n)\Big ){{\,\mathrm{{\text {dv}}}\,}}_g(x)\Big |\le Ch^{1-n+\delta (n-\dim _{box }(\partial {W}))}, \end{aligned}$$

where we used \( {{\,\textrm{vol}\,}}({{\,\textrm{supp}\,}}\chi _h)\le h^{\delta (n-\dim _{box }(\partial {W}))}\). In particular,

$$\begin{aligned}{} & {} \Big |\int _{{W}}\Pi _h(1,x,x){{\,\mathrm{{\text {dv}}}\,}}_g(x)-(2\pi h)^{-n}{{\,\textrm{vol}\,}}_{{\mathbb {R}}^n}(B^n){{\,\textrm{vol}\,}}_M({W})\Big |\\{} & {} \quad \le Ch^{1-n}\Big (T(h)^{-1}+Ch^{\delta (n-\dim _{box }\partial {W})}\Big ). \end{aligned}$$

\(\square \)