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

Semiclassical measures for higher dimensional quantum cat maps

Semyon Dyatlov dyatlov@math.mit.edu Department of Mathematics, Massachusetts Institute of Technology, Cambridge, MA 02139  and  Malo Jézéquel mpjez@mit.edu Department of Mathematics, Massachusetts Institute of Technology, Cambridge, MA 02139
Abstract.

Consider a quantum cat map MM associated to a matrix ASp(2n,)A\in\operatorname{Sp}(2n,\mathbb{Z}), which is a common toy model in quantum chaos. We show that the mass of eigenfunctions of MM on any nonempty open set in the position-frequency space satisfies a lower bound which is uniform in the semiclassical limit, under two assumptions: (1) there is a unique simple eigenvalue of AA of largest absolute value and (2) the characteristic polynomial of AA is irreducible over the rationals. This is similar to previous work [DJ18, DJN22] on negatively curved surfaces and [Sch21] on quantum cat maps with n=1n=1, but this paper gives the first results of this type which apply in any dimension. When condition (2) fails we provide a weaker version of the result and discuss relations to existing counterexamples. We also obtain corresponding statements regarding semiclassical measures and damped quantum cat maps.

In [DJN22], Dyatlov–Jin–Nonnenmacher proved the following lower bound on L2L^{2} mass of eigenfunctions: if (,g)(\mathcal{M},g) is a compact connected Riemannian surface with Anosov geodesic flow (e.g. a negatively curved surface) and uu is an eigenfunction of the Laplacian on \mathcal{M} with eigenvalue λ2-\lambda^{2}, then

uL2()CΩuL2(Ω)for any nonempty openΩ\|u\|_{L^{2}(\mathcal{M})}\leq C_{\Omega}\|u\|_{L^{2}(\Omega)}\quad\text{for any nonempty open}\quad\Omega\subset\mathcal{M} (1.1)

where the constant CΩ>0C_{\Omega}>0 depends on \mathcal{M} and Ω\Omega, but does not depend on λ\lambda. This result has applications to control for the Schrödinger equation, exponential energy decay for the damped wave equation, and semiclassical measures, and belongs to the domain of quantum chaos – see [DJN22] and §1.3 for a historical overview.

The paper [DJN22] only deals with the case of surfaces because the key ingredient, the fractal uncertainty principle of Bourgain–Dyatlov [BD18], is only known for subsets of \mathbb{R}. To prove an analogous result for manifolds of dimension n+1>2n+1>2 would require a fractal uncertainty principle for subsets of n\mathbb{R}^{n}. A naive extension of the fractal uncertainty principle to higher dimensions is false and no generalization suitable for applications to (1.1) is currently known – see the review of Dyatlov [Dya19, §6] and the paper of Han–Schlag [HS20].

In this paper we give a class of higher dimensional examples where a bound of type (1.1) can still be shown using the one-dimensional fractal uncertainty principle of [BD18]. We work in the setting of quantum cat maps, which are toy models commonly used in quantum chaos. In this setting, the geodesic flow on a Riemannian manifold is replaced by a classical cat map, which is the automorphism of the torus 𝕋2n:=2n/2n\mathbb{T}^{2n}:=\mathbb{R}^{2n}/\mathbb{Z}^{2n} induced by an integer symplectic matrix ASp(2n,)A\in\operatorname{Sp}(2n,\mathbb{Z}). The eigenfunctions of the Laplacian are replaced by those of a quantum cat map, an operator M𝐍,θM_{\mathbf{N},\theta} on an 𝐍n\mathbf{N}^{n}-dimensional space 𝐍(θ)\mathcal{H}_{\mathbf{N}}(\theta) which quantizes AA in the sense of (1.2) below. The high eigenvalue limit λ\lambda\to\infty is replaced by the limit 𝐍\mathbf{N}\to\infty. More general quantum maps have also been used in the study of continuous systems (such as Laplacians on Riemannian manifolds), where the quantum map corresponds to Poincaré map(s) of the original dynamical system, see in particular Bogomolny [Bog92] for a physics perspective and Sjöstrand–Zworski [SZ02] for an approach relevant to trace formulas.

For two-dimensional quantum cat maps (analogous to the case of Laplacian eigenfunctions on surfaces) an estimate similar to (1.1) was recently proved by Schwartz [Sch21]. The novelty of the present paper is that it also applies in higher dimensions.

1.1. Setting and lower bounds on eigenfunctions

To explain quantum cat maps in more detail, we use a semiclassical quantization procedure, mapping each classical observable (a function on a symplectic manifold called the phase space) to a quantum observable (an operator on some Hilbert space). In our setting the phase space is the torus 𝕋2n\mathbb{T}^{2n} and each classical observable is quantized to a family of operators (see §2.2.3):

aC(𝕋2n)Op𝐍,θ(a):𝐍(θ)𝐍(θ).a\in C^{\infty}(\mathbb{T}^{2n})\quad\mapsto\quad\operatorname{Op}_{\mathbf{N},\theta}(a):\mathcal{H}_{\mathbf{N}}(\theta)\to\mathcal{H}_{\mathbf{N}}(\theta).

Here θ𝕋2n\theta\in\mathbb{T}^{2n} is a parameter, 𝐍1\mathbf{N}\geq 1 is an integer, and the Hilbert spaces of quantum states 𝐍(θ)\mathcal{H}_{\mathbf{N}}(\theta), defined in §2.2.1, have dimension 𝐍n\mathbf{N}^{n}. We denote the inner product on these spaces by ,\langle\bullet,\bullet\rangle_{\mathcal{H}}. The semiclassical parameter is h:=(2π𝐍)1h:=(2\pi\mathbf{N})^{-1}.

Every matrix ASp(2n,)A\in\operatorname{Sp}(2n,\mathbb{Z}) defines a symplectic automorphism of the torus 𝕋2n\mathbb{T}^{2n}. This automorphism is quantized by a family of unitary maps

M𝐍,θ:𝐍(θ)𝐍(θ)M_{\mathbf{N},\theta}:\mathcal{H}_{\mathbf{N}}(\theta)\to\mathcal{H}_{\mathbf{N}}(\theta)

which satisfy the following exact Egorov’s theorem (2.58) intertwining the action of AA on 𝕋2n\mathbb{T}^{2n} with conjugation by M𝐍,θM_{\mathbf{N},\theta}:

M𝐍,θ1Op𝐍,θ(a)M𝐍,θ=Op𝐍,θ(aA)for allaC(𝕋2n).M_{\mathbf{N},\theta}^{-1}\operatorname{Op}_{\mathbf{N},\theta}(a)M_{\mathbf{N},\theta}=\operatorname{Op}_{\mathbf{N},\theta}(a\circ A)\quad\text{for all}\quad a\in C^{\infty}(\mathbb{T}^{2n}). (1.2)

To construct such M𝐍,θM_{\mathbf{N},\theta} we need to impose a quantization condition (2.57) on θ\theta. The constructed operators are unique up to multiplication by a unit complex constant. See §2.2.4 for more details and §2.2.5 for explicit formulas for Op𝐍,θ(a)\operatorname{Op}_{\mathbf{N},\theta}(a) and M𝐍,θM_{\mathbf{N},\theta}.

Throughout the paper we fix ASp(2n,)A\in\operatorname{Sp}(2n,\mathbb{Z}) and assume the following spectral gap condition on the spectrum Spec(A)\operatorname{Spec}(A):

A has a simple eigenvalue λ+ such thatmaxλSpec(A){λ+}|λ|<|λ+|.\text{$A$ has a simple eigenvalue $\lambda_{+}$ such that}\quad\max_{\lambda\in\operatorname{Spec}(A)\setminus\{\lambda_{+}\}}|\lambda|<|\lambda_{+}|. (1.3)

This condition is crucial in the proof because it means there is a one-dimensional ‘fast’ direction in which the powers of AA grow faster than in other directions, which makes it possible to apply the one-dimensional fractal uncertainty principle – see §1.4.

Our first result is the following analog of (1.1):

Theorem 1.

Assume that AA satisfies (1.3) and the characteristic polynomial of AA is irreducible over the rationals. Then for each aC(𝕋2n)a\in C^{\infty}(\mathbb{T}^{2n}), a0a\not\equiv 0, there exists Ca>0C_{a}>0 such that for all large enough 𝐍\mathbf{N} and every eigenfunction u𝐍(θ)u\in\mathcal{H}_{\mathbf{N}}(\theta) of M𝐍,θM_{\mathbf{N},\theta} we have

uCaOp𝐍,θ(a)u.\|u\|_{\mathcal{H}}\leq C_{a}\|\operatorname{Op}_{\mathbf{N},\theta}(a)u\|_{\mathcal{H}}. (1.4)

Here the norm Op𝐍,θ(a)u\|\operatorname{Op}_{\mathbf{N},\theta}(a)u\|_{\mathcal{H}} on the right-hand side of (1.4) plays a similar role to the norm uL2(Ω)\|u\|_{L^{2}(\Omega)} on the right-hand side of (1.1): if aa is supported in some 𝐍\mathbf{N}-independent subset of 𝕋2n\mathbb{T}^{2n}, then Op𝐍,θ(a)u\|\operatorname{Op}_{\mathbf{N},\theta}(a)u\|_{\mathcal{H}} can be thought of as the norm of uu localized in the position-frequency space to this set.

We remark that the conditions of Theorem 1 are always satisfied if n=1n=1 (i.e. AA is a 2×22\times 2 matrix) and AA is hyperbolic, that is it has no eigenvalues on the unit circle. Thus Theorem 1 (or more precisely, Theorem 2 below) implies the result of [Sch21]. See Figure 1 for a numerical illustration in the case n=1n=1. For n2n\geq 2, our assumption (1.3) does not require AA to be hyperbolic. We also remark that having characteristic polynomial irreducible over \mathbb{Q} is a generic property for integer symplectic matrices, see Rivin [Riv08] and the book of Kowalski [Kow08, Theorem 7.12].

Refer to caption
Refer to caption
Figure 1. Plots of concentration in the position-frequency space 𝕋2\mathbb{T}^{2} (more precisely, plots of the corresponding Wigner matrices convolved with appropriate Gaussians, on the logarithmic scale, see (2.60)) of two eigenfunctions of a quantum cat map corresponding to A=(2312)A=\begin{pmatrix}2&3\\ 1&2\end{pmatrix}, 𝐍=780\mathbf{N}=780. On the left is a typical eigenfunction, showing equidistribution consistent with the Quantum Ergodicity result of [BDB96]. On the right is a particular eigenfunction which exhibits scarring discovered in [FNDB03]. This eigenfunction does not violate Theorem 1 because the mass on the scar in the center is approximately equal to the mass elsewhere.

1.2. Further results

Theorem 1 is a consequence of a more general result, Theorem 2 below, that applies to quasimodes of M𝐍,θM_{\mathbf{N},\theta} and does not require the irreducibility over \mathbb{Q} of the characteristic polynomial of AA. Before stating it, we need to introduce more notation. In order to measure the strength of a quasimode, we introduce for u𝐍(θ)u\in\mathcal{H}_{\mathbf{N}}(\theta) the quantity

𝐫M(u):=minz𝕊1M𝐍,θuzu,\mathbf{r}_{M}(u):=\min_{z\in\mathbb{S}^{1}}\left\|M_{\mathbf{N},\theta}u-zu\right\|_{\mathcal{H}}, (1.5)

where 𝕊1\mathbb{S}^{1} denotes the unit circle in \mathbb{C}. Note that 𝐫M(u)=0\mathbf{r}_{M}(u)=0 if and only if uu is an eigenfunction of M𝐍,θM_{\mathbf{N},\theta}.

To relax the assumption on the characteristic polynomial of AA, let us notice that by (1.3) the leading eigenvalue λ+\lambda_{+} is real. Since the matrix AA is symplectic, its transpose is conjugate to its inverse, thus λ:=λ+1\lambda_{-}:=\lambda_{+}^{-1} is also a simple eigenvalue for AA and |λ|<1<|λ+||\lambda_{-}|<1<|\lambda_{+}|. Moreover, all other eigenvalues λ\lambda of AA satisfy |λ|<|λ|<|λ+||\lambda_{-}|<|\lambda|<|\lambda_{+}|.

Denote by E±2nE_{\pm}\subset\mathbb{R}^{2n} the (real) eigenspaces of AA associated to λ±\lambda_{\pm}. Let V±V_{\pm} be the smallest subspace of 2n\mathbb{Q}^{2n} such that E±E_{\pm} is contained in V±V_{\pm}\otimes\mathbb{R}. Note that V±V_{\pm} are invariant under AA. Denote by

𝕋±𝕋2n\mathbb{T}_{\pm}\ \subset\ \mathbb{T}^{2n} (1.6)

the subtori given by the projections of V±V_{\pm}\otimes\mathbb{R} to 𝕋2n\mathbb{T}^{2n}.

The tori 𝕋±\mathbb{T}_{\pm} are relevant here because of their alternative dynamical definition given in Lemma 4.3: if e±e_{\pm} is any eigenvector of AA associated to the eigenvalue λ±\lambda_{\pm}, then the closure of the orbit of a point x𝕋2nx\in\mathbb{T}^{2n} by the translation flow generated by e±e_{\pm} is x+𝕋±x+\mathbb{T}_{\pm}. In our setting, these translation flows will play the role that was played by the horocyclic flows on the unit tangent bundle of hyperbolic surfaces in [DJ18]. Let us also give

Definition 1.1.

Let UU and 𝕋\mathbb{T}^{\prime} be respectively an open subset and a subtorus of 𝕋2n\mathbb{T}^{2n} and assume that A(𝕋)=𝕋A(\mathbb{T}^{\prime})=\mathbb{T}^{\prime}. We say that UU satisfies the geometric control condition transversally to 𝕋\mathbb{T}^{\prime} if, for every x𝕋2nx\in\mathbb{T}^{2n}, there exists mm\in\mathbb{Z} such that Amx+𝕋A^{m}x+\mathbb{T}^{\prime} intersects UU.

We now state a more general version of Theorem 1:

Theorem 2.

Assume that AA satisfies (1.3). Let aC(𝕋2n)a\in C^{\infty}(\mathbb{T}^{2n}) be such that {a0}\left\{a\neq 0\right\} satisfies the geometric control condition transversally to 𝕋+\mathbb{T}_{+} and 𝕋\mathbb{T}_{-}. Then there exists Ca>0C_{a}>0 such that for all large enough 𝐍\mathbf{N} and every u𝐍(θ)u\in\mathcal{H}_{\mathbf{N}}(\theta), we have

uCaOp𝐍,θ(a)u+Ca𝐫M(u)log𝐍.\left\|u\right\|_{\mathcal{H}}\leq C_{a}\left\|\operatorname{Op}_{\mathbf{N},\theta}(a)u\right\|_{\mathcal{H}}+C_{a}\mathbf{r}_{M}(u)\log\mathbf{N}. (1.7)

Theorem 2 implies Theorem 1. Indeed, if the characteristic polynomial of AA is irreducible over the rationals, then 𝕋±=𝕋2n\mathbb{T}_{\pm}=\mathbb{T}^{2n} (see Lemma A.3), thus every nonempty open set satisfies the geometric control condition transversally to 𝕋+\mathbb{T}_{+} and 𝕋\mathbb{T}_{-}. However, one can find a matrix AA that satisfies (1.3) but for which (1.4) fails for certain choices of aa (in particular the characteristic polynomial of AA is reducible over \mathbb{Q}) – see the examples in §A.2.

We will also prove the following theorem about damped quantum cat maps. It is analogous to exponential energy decay for negatively curved surfaces proved in [DJN22] (following earlier work of Jin [Jin20] in the constant curvature case):

Theorem 3.

Assume that AA satisfies (1.3). Let bC(𝕋2n)b\in C^{\infty}(\mathbb{T}^{2n}) be such that |b|1\left|b\right|\leq 1. Assume that the set {|b|<1}\{|b|<1\} satisfies the geometric control condition transversally to 𝕋+\mathbb{T}_{+} and 𝕋\mathbb{T}_{-}. Then there exists 0<η<10<\eta<1 such that for all 𝐍\mathbf{N} large enough the spectral radius of the operator Op𝐍,θ(b)M𝐍,θ\operatorname{Op}_{\mathbf{N},\theta}(b)M_{\mathbf{N},\theta} is less than η\eta.

Note that by Lemma A.3, if the characteristic polynomial of AA is irreducible over \mathbb{Q} then the condition on bb simplifies to {|b|<1}\{|b|<1\}\neq\emptyset. See Figure 2 for a numerical illustration.

Refer to caption
Refer to caption
Figure 2. Numerically computed eigenvalues for two damped quantum cat maps Op𝐍,θ(b)M𝐍,θ\operatorname{Op}_{\mathbf{N},\theta}(b)M_{\mathbf{N},\theta}, with the unit circle in the background. In both cases n=2n=2 (i.e. AA is a 4×44\times 4 matrix), 𝐍=60\mathbf{N}=60, and the set {|b|<1}\{|b|<1\} is inside the 14{1\over 4}-neighborhood of the origin (0,0,0,0)𝕋4(0,0,0,0)\in\mathbb{T}^{4} in the \ell^{\infty} norm. On the left we use the matrix (A.4) with irreducible characteristic polynomial and the spectral gap of Theorem 3 is visible. On the right we use the matrix (A.6) and the condition of Theorem 3 is not satisfied; there does not appear to be a spectral gap.

1.3. Semiclassical measures and overview of history

We now give an application of Theorem 2 to semiclassical measures. These measures describe the possible ways in which the mass of eigenfunctions of M𝐍,θM_{\mathbf{N},\theta} can distribute in the position-frequency space in the limit 𝐍\mathbf{N}\to\infty, and they are defined as follows:

Definition 1.2.

Let 𝐍j\mathbf{N}_{j}\to\infty, θj𝕋2n\theta_{j}\in\mathbb{T}^{2n} be sequences such that the quantization condition (2.57) holds for all 𝐍j,θj\mathbf{N}_{j},\theta_{j}. Let uj𝐍j(θj)u_{j}\in\mathcal{H}_{\mathbf{N}_{j}}(\theta_{j}) be eigenfunctions of M𝐍j,θjM_{\mathbf{N}_{j},\theta_{j}} of norm 1. We say that the sequence uju_{j} converges weakly to a Borel measure μ\mu on 𝕋2n\mathbb{T}^{2n} if

Op𝐍j,θj(a)uj,uj𝕋2na𝑑μfor allaC(𝕋2n).\langle\operatorname{Op}_{\mathbf{N}_{j},\theta_{j}}(a)u_{j},u_{j}\rangle_{\mathcal{H}}\to\int_{\mathbb{T}^{2n}}a\,d\mu\quad\text{for all}\quad a\in C^{\infty}(\mathbb{T}^{2n}). (1.8)

A measure μ\mu on 𝕋2n\mathbb{T}^{2n} is called a semiclassical measure (associated to the toric automorphism AA) if there exist sequences 𝐍j,θj,uj\mathbf{N}_{j},\theta_{j},u_{j} such that (1.8) holds.

One can show (similarly to [Zwo12, Theorem 5.2] using a diagonal argument, the norm bound (2.51), Riesz representation theorem for the dual to C0(𝕋2n)C^{0}(\mathbb{T}^{2n}), and the sharp Gårding inequality (2.48)) that each norm 1 sequence uj𝐍j(θj)u_{j}\in\mathcal{H}_{\mathbf{N}_{j}}(\theta_{j}) has a subsequence which has a weak limit in the sense of (1.8). Every semiclassical measure is a probability measure on 𝕋2n\mathbb{T}^{2n} (as follows from taking a=1a=1 in (1.8) and using that Op𝐍,θ(1)=I\operatorname{Op}_{\mathbf{N},\theta}(1)=I) which is invariant under the map AA (as follows from (1.2)). Semiclassical measures for quantum cat maps are analogous to those for Laplacian eigenfunctions on a Riemannian manifold (,g)(\mathcal{M},g), which are probability measures on the cosphere bundle SS^{*}\mathcal{M} invariant under the geodesic flow – see [Zwo12, Chapter 5] for more information.

As explained in §3.3, Theorem 2 implies the following property of the support of semiclassical measures:

Theorem 4.

Assume that AA satisfies (1.3) and μ\mu is a semiclassical measure associated to AA. Then suppμ\operatorname{supp}\mu contains a set of the form x+𝕋+x+\mathbb{T}_{+} or x+𝕋x+\mathbb{T}_{-} for some x𝕋2nx\in\mathbb{T}^{2n}. In particular, if the characteristic polynomial of AA is irreducible over the rationals, then suppμ=𝕋2n\operatorname{supp}\mu=\mathbb{T}^{2n}.

In Appendix A, we discuss the algebraic properties of the spaces V+V_{+} and VV_{-}, and their implications in terms of the supports of semiclassical measures for AA. In particular, we give examples of the following situations:

  • the characteristic polynomial of AA is irreducible, and hence the semiclassical measures are fully supported;

  • 2n=V+V\mathbb{Q}^{2n}=V_{+}\oplus V_{-}, in which case there are semiclassical measures supported on translates of 𝕋+\mathbb{T}_{+} and 𝕋\mathbb{T}_{-} due to [Kel10];

  • V++V2nV_{+}+V_{-}\neq\mathbb{Q}^{2n} and there are semiclassical measures supported in 𝕋+\mathbb{T}_{+} and 𝕋\mathbb{T}_{-};

  • V++V2nV_{+}+V_{-}\neq\mathbb{Q}^{2n} but the supports of semiclassical measures (associated to a certain basis of eigenfunctions) are strictly larger than 𝕋+\mathbb{T}_{+} or 𝕋\mathbb{T}_{-}.

The last two cases emphasize the fact that when V++V2nV_{+}+V_{-}\neq\mathbb{Q}^{2n}, one has to study the action of AA on 2n/(V++V)\mathbb{Q}^{2n}/(V_{+}+V_{-}) in order to refine the information on the support of semiclassical measures given by Theorem 4.

We now give a brief overview of history of semiclassical measures in quantum chaos, referring the reader to review articles by Marklof [Mar06], Sarnak [Sar11], Nonnenmacher [Non13], and Dyatlov [Dya21] for more information. In the setting of eigenfunctions of the Laplacian on compact negatively curved Riemannian manifolds, the Quantum Ergodicity theorem of Shnirelman [Shn74], Zelditch [Zel87], and Colin de Verdière [CdV85] states that a density 1 sequence of eigenfunctions equidistributes, i.e. converges to the Liouville measure in a way analogous to (1.8). The Quantum Unique Ergodicity conjecture of Rudnick–Sarnak [RS94] claims that the whole sequence of eigenfunctions equidistributes, i.e. the Liouville measure is the only semiclassical measure. It is open in general but known in the particular cases of arithmetic hyperbolic surfaces for joint eigenfunctions of the Laplacian and the Hecke operators, which are additional symmetries present in the arithmetic case – see Lindenstrauss [Lin06] and Brooks–Lindenstrauss [BL14].

The works of Anantharaman [Ana08], Anantharaman–Nonnenmacher [AN07b], Anantharaman–Koch–Nonnenmacher [AKN09], Rivière [Riv10b, Riv10a], and Anantharaman–Silberman [AS13] establish entropy bounds, which are positive lower bounds on the Kolmogorov–Sinaĭ entropy of semiclassical measures. As mentioned in the beginning of the introduction, in the setting of negatively curved surfaces the paper [DJN22] gives another restriction: all semiclassical measures have full support. In the case of hyperbolic surfaces this was previously proved by Dyatlov–Jin [DJ18].

The study of quantum cat maps, which is the setting of the present paper, goes back to the work of Hannay–Berry [HB80]. Bouzouina–De Bièvre [BDB96] proved the analogue of quantum ergodicity in this setting. Faure–Nonnenmacher–De Bièvre [FNDB03] constructed examples of semiclassical measures for 2-dimensional quantum cat maps which are not the Liouville measure; this contradicts the Quantum Unique Ergodicity conjecture for quantum cat maps but it does not contradict Theorem 1 since these measures were supported on the entire 𝕋2\mathbb{T}^{2}. Faure–Nonnenmacher [FN04], Brooks [Bro10], and Rivière [Riv11] established ‘entropy-like’ bounds on semiclassical measures; see also Anantharaman–Nonnenmacher [AN07a] and Gutkin [Gut10] for entropy bounds for other models of quantum maps. As mentioned above, Schwartz [Sch21] obtained Theorem 1 for 2-dimensional quantum cat maps.

Kurlberg–Rudnick [KR00] introduced the analogue of Hecke operators in the setting of 2-dimensional quantum cat maps. For the joint eigenfunctions of a quantum cat map and the Hecke operators (known as arithmetic eigenfunctions) they showed that Quantum Unique Ergodicity holds; see also Gurevich–Hadani [GH11]. This does not contradict the counterexample in [FNDB03] since the quantum cat maps used there have eigenvalues of high multiplicity.

In the setting of higher dimensional quantum cat maps which have an isotropic invariant rational subspace, Kelmer [Kel10] constructed examples of semiclassical measures (for arithmetic eigenfunctions) which are supported on proper submanifolds of 𝕋2n\mathbb{T}^{2n}. On the other hand, if there are no isotropic invariant rational subspaces, then [Kel10] gives Quantum Unique Ergodicity for arithmetic eigenfunctions. Compared to Theorem 4, the conclusion of [Kel10] is stronger than suppμ=𝕋2n\operatorname{supp}\mu=\mathbb{T}^{2n} and the assumption is weaker than the characteristic polynomial of AA being irreducible over rationals (since [Kel10] only assumes that there are no isotropic invariant rational subspaces), however the result of [Kel10] only applies to arithmetic eigenfunctions. For a further discussion of the relation of our results with those of [Kel10], see Appendix A.

1.4. Outline of the proof

We now give an outline of the proof of Theorem 2. For simplicity we assume that the symbol aa satisfies 0a10\leq a\leq 1, as well as the following stronger version of the geometric control condition: there exists an open set 𝒰𝕋2n\mathcal{U}\subset\mathbb{T}^{2n} such that a=1a=1 on 𝒰\mathcal{U} and each shifted torus x+𝕋+x+\mathbb{T}_{+}, x+𝕋x+\mathbb{T}_{-} intersects 𝒰\mathcal{U}. We also assume that 𝐍\mathbf{N} is large and u𝐍(θ)u\in\mathcal{H}_{\mathbf{N}}(\theta) is an eigenfunction of M𝐍,θM_{\mathbf{N},\theta}.

1.4.1. Reduction to the key estimate

Take the partition of unity on 𝕋2n\mathbb{T}^{2n}

1=b1+b2,b1:=a,b2:=1a,suppb2𝒰=.1=b_{1}+b_{2},\quad b_{1}:=a,\quad b_{2}:=1-a,\quad\operatorname{supp}b_{2}\cap\mathcal{U}=\emptyset. (1.9)

Quantizing b1,b2b_{1},b_{2} to pseudodifferential operators, we get the quantum partition of unity

I=B1+B2,B1:=Op𝐍,θ(b1),B2:=Op𝐍,θ(b2).I=B_{1}+B_{2},\quad B_{1}:=\operatorname{Op}_{\mathbf{N},\theta}(b_{1}),\quad B_{2}:=\operatorname{Op}_{\mathbf{N},\theta}(b_{2}).

For an operator LL on 𝐍,θ\mathcal{H}_{\mathbf{N},\theta} and an integer TT, define the conjugated operator

L(T):=M𝐍,θTLM𝐍,θT:𝐍(θ)𝐍(θ).L(T):=M_{\mathbf{N},\theta}^{-T}LM_{\mathbf{N},\theta}^{T}:\mathcal{H}_{\mathbf{N}}(\theta)\to\mathcal{H}_{\mathbf{N}}(\theta).

Since uu is an eigenfunction of the unitary operator M𝐍,θM_{\mathbf{N},\theta}, we have for all TT and LL

L(T)u=LM𝐍,θTu=Lu.\|L(T)u\|_{\mathcal{H}}=\|LM_{\mathbf{N},\theta}^{T}u\|_{\mathcal{H}}=\|Lu\|_{\mathcal{H}}. (1.10)

We will propagate up to time 2T12T_{1}, where T1T_{1}\in\mathbb{N} is chosen so that

T1ρlog|λ+|log𝐍T_{1}\approx{\rho\over\log|\lambda_{+}|}\log\mathbf{N} (1.11)

and the constant ρ>0\rho>0 is chosen later. We write a refined quantum partition of unity

I=B𝒳+B𝒴,B𝒳:=B2(2T11)B2(1)B2(0),B𝒴:=j=02T11B2(2T11)B2(j+1)B1(j).\begin{gathered}I=B_{\mathcal{X}}+B_{\mathcal{Y}},\quad B_{\mathcal{X}}:=B_{2}(2T_{1}-1)\cdots B_{2}(1)B_{2}(0),\\ B_{\mathcal{Y}}:=\sum_{j=0}^{2T_{1}-1}B_{2}(2T_{1}-1)\cdots B_{2}(j+1)B_{1}(j).\end{gathered} (1.12)

Since |b2|1|b_{2}|\leq 1, we have B21+𝒪(𝐍12)\|B_{2}\|_{\mathcal{H}\to\mathcal{H}}\leq 1+\mathcal{O}(\mathbf{N}^{-{1\over 2}}) (see (2.51)), thus we can bound the norm of the product B2(2T11)B2(j+1)B_{2}(2T_{1}-1)\cdots B_{2}(j+1) by 2. Applying the partition (1.12) to uu and using (1.10) with L:=B1=Op𝐍,θ(a)L:=B_{1}=\operatorname{Op}_{\mathbf{N},\theta}(a), we then get

uB𝒳u+4T1Op𝐍,θ(a)u.\|u\|_{\mathcal{H}}\leq\|B_{\mathcal{X}}u\|_{\mathcal{H}}+4T_{1}\|\operatorname{Op}_{\mathbf{N},\theta}(a)u\|_{\mathcal{H}}. (1.13)

The key component of the proof is the following estimate valid for the right choice of T1T_{1}:

B𝒳=𝒪(𝐍β)for someβ>0as𝐍.\|B_{\mathcal{X}}\|_{\mathcal{H}\to\mathcal{H}}=\mathcal{O}({\mathbf{N}}^{-\beta})\quad\text{for some}\quad\beta>0\quad\text{as}\quad\mathbf{N}\to\infty. (1.14)

Here the exponent β\beta only depends on the matrix AA and the choice of the partition (1.9).

Together with (1.13), the key estimate (1.14) implies the following weaker version of Theorem 2:

uClog𝐍Op𝐍,θ(a)u.\|u\|_{\mathcal{H}}\leq C\log\mathbf{N}\|\operatorname{Op}_{\mathbf{N},\theta}(a)u\|_{\mathcal{H}}. (1.15)

To remove the log𝐍\log\mathbf{N} prefactor, we revise the decomposition I=B𝒳+B𝒴I=B_{\mathcal{X}}+B_{\mathcal{Y}} in the same way as in [DJ18, DJN22] (which in turn was inspired by [Ana08]), including more terms into B𝒳B_{\mathcal{X}} and using that the norm bound in (1.14) (or rather, its slight generalization) is a negative power of 𝐍\mathbf{N}. See §3.1.2 for details.

1.4.2. Microlocal structure of the propagated operators

We now give an outline of the proof of the estimate (1.14). To simplify the presentation, we assume the following stronger version of the gap condition (1.3):

A has a simple eigenvalue λ+ such thatmaxλSpec(A){λ+}|λ|<|λ+|.\text{$A$ has a simple eigenvalue $\lambda_{+}$ such that}\quad\max_{\lambda\in\operatorname{Spec}(A)\setminus\{\lambda_{+}\}}|\lambda|<\sqrt{|\lambda_{+}|}. (1.16)

Let E±2nE_{\pm}\subset\mathbb{R}^{2n} be the one-dimensional eigenspaces of AA corresponding to λ+\lambda_{+} and λ:=λ+1\lambda_{-}:=\lambda_{+}^{-1}. We also define LL_{\mp} to be the sum of the generalized eigenspaces corresponding to all eigenvalues of AA other than λ±\lambda_{\pm}. Then

2n=E+L=EL+.\mathbb{R}^{2n}=E_{+}\oplus L_{-}=E_{-}\oplus L_{+}.

By (1.16) we have the following norm bounds as TT\to\infty:

A±T=𝒪(|λ+|T),A±T|L=𝒪(|λ+|T2).\|A^{\pm T}\|=\mathcal{O}(|\lambda_{+}|^{T}),\quad\|A^{\pm T}|_{L_{\mp}}\|=\mathcal{O}(|\lambda_{+}|^{T\over 2}). (1.17)

Returning to the proof of (1.14), conjugating by M𝐍,θT1M_{\mathbf{N},\theta}^{T_{1}} we reduce it to the bound

BB+=𝒪(𝐍β)for someβ>0whereB:=B2(T11)B2(0),B+:=B2(1)B2(T1).\begin{gathered}\|B_{-}B_{+}\|_{\mathcal{H}\to\mathcal{H}}=\mathcal{O}(\mathbf{N}^{-\beta})\quad\text{for some}\quad\beta>0\\ \text{where}\quad B_{-}:=B_{2}(T_{1}-1)\cdots B_{2}(0),\quad B_{+}:=B_{2}(-1)\cdots B_{2}(-T_{1}).\end{gathered}

We would like to write the operators B±B_{\pm} as quantizations of some symbols. By the exact Egorov’s Theorem (1.2), we have B2(j)=Op𝐍,θ(b2Aj)B_{2}(j)=\operatorname{Op}_{\mathbf{N},\theta}(b_{2}\circ A^{j}) for all jj. However, when jj is too large the derivatives of the symbols b2Ajb_{2}\circ A^{j} grow too fast with 𝐍\mathbf{N} and the methods of semiclassical analysis no longer apply.

If 0j<T10\leq j<T_{1}, then by (1.17) the derivatives of b2Ajb_{2}\circ A^{j} in the eigendirection E+E_{+} are bounded by |λ+|T1|\lambda_{+}|^{T_{1}} but the derivatives along the complementary hyperplane LL_{-} are bounded by |λ+|T1/2|\lambda_{+}|^{T_{1}/2}. Thus b2Ajb_{2}\circ A^{j} lies in the anisotropic symbol class SL,ρ,ρ2(𝕋2n)S_{L_{-},\rho,{\rho\over 2}}(\mathbb{T}^{2n}), consisting of 𝐍\mathbf{N}-dependent functions in C(𝕋2n)C^{\infty}(\mathbb{T}^{2n}) such that each derivative along LL_{-} gives an 𝐍ρ2\mathbf{N}^{\rho\over 2} growth and derivatives in other directions give an 𝐍ρ\mathbf{N}^{\rho} growth – see §§2.1.4,2.2.3 for details. In order for the standard properties of semiclassical quantization to hold, we need ρ+ρ2<1\rho+{\rho\over 2}<1, that is

0ρ<23.0\leq\rho<\textstyle{2\over 3}. (1.18)

For such ρ\rho we can then show (see Lemma 3.11) that for any δ(0,13ρ2)\delta\in(0,1-{3\rho\over 2})

B=Op𝐍,θ(b)+𝒪(𝐍δ)whereb:=j=0T11b2Aj.B_{-}=\operatorname{Op}_{\mathbf{N},\theta}(b_{-})+\mathcal{O}(\mathbf{N}^{-\delta})_{\mathcal{H}\to\mathcal{H}}\quad\text{where}\quad b_{-}:=\prod_{j=0}^{T_{1}-1}b_{2}\circ A^{j}.

Reversing the direction of propagation and replacing LL_{-} with L+L_{+}, we similarly have

B+=Op𝐍,θ(b+)+𝒪(𝐍δ)whereb+:=j=1T1b2Aj.B_{+}=\operatorname{Op}_{\mathbf{N},\theta}(b_{+})+\mathcal{O}(\mathbf{N}^{-\delta})_{\mathcal{H}\to\mathcal{H}}\quad\text{where}\quad b_{+}:=\prod_{j=1}^{T_{1}}b_{2}\circ A^{-j}.

Then the key estimate (1.14) reduces to

Op𝐍,θ(b)Op𝐍,θ(b+)=𝒪(𝐍β)for someβ>0.\|\operatorname{Op}_{\mathbf{N},\theta}(b_{-})\operatorname{Op}_{\mathbf{N},\theta}(b_{+})\|_{\mathcal{H}\to\mathcal{H}}=\mathcal{O}(\mathbf{N}^{-\beta})\quad\text{for some}\quad\beta>0. (1.19)

The symbols b±b_{\pm} lie in the classes SL±,ρ+ε,ρ2+ε(𝕋2n)S_{L_{\pm},\rho+\varepsilon,{\rho\over 2}+\varepsilon}(\mathbb{T}^{2n}) for each ε>0\varepsilon>0 but when ρ>12\rho>{1\over 2} they cannot be put in the same symbol calculus. This is important because otherwise the norm of Op𝐍,θ(b)Op𝐍,θ(b+)\operatorname{Op}_{\mathbf{N},\theta}(b_{-})\operatorname{Op}_{\mathbf{N},\theta}(b_{+}) would converge as 𝐍\mathbf{N}\to\infty to sup|bb+|\sup|b_{-}b_{+}|, which would be equal to 1 as long as suppb2\operatorname{supp}b_{2} contains at least one periodic trajectory of AA.

1.4.3. Applying the fractal uncertainty principle

Of course, just not being in the same symbol calculus does not give the bound (1.19). This is where the fractal uncertainty principle of [BD18] enters.

We first discuss the case n=1n=1, that is, the matrix AA has size 2×22\times 2. Then the fractal uncertainty principle shows the decay bound (1.19) if the sets π±(suppb±)\pi_{\pm}(\operatorname{supp}b_{\pm}) are porous on scales C𝐍ρC\mathbf{N}^{-\rho} to 1 in the sense of Definition 4.1 below, and ρ>12\rho>{1\over 2}. Here we think of suppb±\operatorname{supp}b_{\pm} as subsets of [0,1]22[0,1]^{2}\subset\mathbb{R}^{2} and π,π+:2\pi_{-},\pi_{+}:\mathbb{R}^{2}\to\mathbb{R} are two linearly independent linear functionals. (The original result of [BD18] is stated in the particular case π(x,ξ)=x\pi_{-}(x,\xi)=x, π+(x,ξ)=ξ\pi_{+}(x,\xi)=\xi and the general case follows via conjugation by a metaplectic transformation.)

Refer to caption
Figure 3. A numerical illustration of the sets suppb\operatorname{supp}b_{-} (red) and suppb+\operatorname{supp}b_{+} (blue) for a 2-dimensional cat map. Each of these sets is ‘smooth’ in one of the eigendirections of AA and porous in the other eigendirection.

The required porosity property holds if we choose the projections π±\pi_{\pm} such that their kernels are given by the eigenspaces E±=L±E_{\pm}=L_{\pm}. This is illustrated on Figure 3, and can be proved (taking the case of suppb\operatorname{supp}b_{-} to fix notation) roughly speaking by combining the following observations:

  1. (1)

    Whether or not some z𝕋2z\in\mathbb{T}^{2} belongs to suppb\operatorname{supp}b_{-} depends on the forward trajectory {Ajz0j<T1}\{A^{j}z\mid 0\leq j<T_{1}\}, more specifically on whether the points in this trajectory all lie in suppb2\operatorname{supp}b_{2}.

  2. (2)

    The intersection of suppb\operatorname{supp}b_{-} with any line in the direction E+E_{+} is porous on scales C𝐍ρC\mathbf{N}^{-\rho} to 1. Indeed, the projection of E+E_{+} to 𝕋2\mathbb{T}^{2} is dense, so any sufficiently large line segment in the direction of E+E_{+} intersects the complement of suppb2\operatorname{supp}b_{2} (recalling (1.9)). This creates the pores on scale 1. To get pores on smaller scales, we use that segments in the direction of E+E_{+} are expanded by the map AjA^{j} by the factor |λ+|j|\lambda_{+}|^{j}, so the condition that suppbAj(suppb2)\operatorname{supp}b_{-}\subset A^{-j}(\operatorname{supp}b_{2}) creates pores on scales |λ+|j[𝐍ρ,1]|\lambda_{+}|^{-j}\in[\mathbf{N}^{-\rho},1].

  3. (3)

    If z,w𝕋2z,w\in\mathbb{T}^{2} lie on the same line segment in the direction EE_{-} (of bounded length), then the forward trajectories AjzA^{j}z, AjwA^{j}w converge to each other as jj\to\infty. Thus the projection π(suppb)\pi_{-}(\operatorname{supp}b_{-}) looks similar to the intersection of suppb\operatorname{supp}b_{-} with any fixed line segment in the direction of E+E_{+}, which we already know is porous.

Now, if we take 12<ρ<23{1\over 2}<\rho<{2\over 3}, then the fractal uncertainty principle applies and gives the key estimate. This roughly corresponds to the proof in [Sch21].

We now move on to the case of higher dimensions n>1n>1 which is the main novelty of this paper. As mentioned above, the higher dimensional version of fractal uncertainty principle is not available. However, we can still derive the key estimate (1.19) from the one-dimensional fractal uncertainty principle as long as the projections π±(suppb±)\pi_{\pm}(\operatorname{supp}b_{\pm}) are porous on scales C𝐍ρC\mathbf{N}^{-\rho} to 1 for some ρ>12\rho>{1\over 2}, where π±:2n\pi_{\pm}:\mathbb{R}^{2n}\to\mathbb{R} are some fixed linear maps such that kerπ+kerπ\ker\pi_{+}\cap\ker\pi_{-} is a codimension 2 symplectic subspace of 2n\mathbb{R}^{2n}. Following the case n=1n=1, it is natural to take π±\pi_{\pm} such that their kernels are given by the spaces L±L_{\pm}. However, we cannot expect π±(suppb±)\pi_{\pm}(\operatorname{supp}b_{\pm}) to be porous because the observation (3) above is no longer valid: there exist z,wz,w such that zwLz-w\in L_{-} but AjzAjw↛0A^{j}z-A^{j}w\not\to 0 as jj\to\infty, for example one can take zwz-w to be an eigenvector of AA with any eigenvalue λλ+\lambda\neq\lambda_{+} such that |λ|1|\lambda|\geq 1.

To deal with this issue, we split the product of operators in (1.19) into a sum of many pieces:

Op𝐍,θ(b)Op𝐍,θ(b+)=kOp𝐍,θ(b)Op𝐍,θ(ψk)Op𝐍,θ(b+)\operatorname{Op}_{\mathbf{N},\theta}(b_{-})\operatorname{Op}_{\mathbf{N},\theta}(b_{+})=\sum_{k}\operatorname{Op}_{\mathbf{N},\theta}(b_{-})\operatorname{Op}_{\mathbf{N},\theta}(\psi_{k})\operatorname{Op}_{\mathbf{N},\theta}(b_{+}) (1.20)

where {ψk}\{\psi_{k}\} is a partition of unity on 𝕋2n\mathbb{T}^{2n} such that each suppψk\operatorname{supp}\psi_{k} looks like a ball of radius 𝐍ρ2\mathbf{N}^{-{\rho\over 2}}. We then observe that:

Refer to caption
Refer to caption
Figure 4. Left: A numerical illustration of the set suppb\operatorname{supp}b_{-} for a 4-dimensional cat map, intersected with some 2-dimensional plane. The little squares correspond to the supports of the functions ψk\psi_{k}. Right: The zoomed in image of the blue square on the left. The projection onto the horizontal direction is π\pi_{-} and the set π(suppbsuppψk)\pi_{-}(\operatorname{supp}b_{-}\cap\operatorname{supp}\psi_{k}) is porous.
  • the terms B(k)B_{(k)} in the sum in (1.20) form an almost orthogonal family, namely

    B(k)B()=𝒪(𝐍),B(k)B()=𝒪(𝐍)whensuppψksuppψ=.B_{(k)}^{*}B_{(\ell)}=\mathcal{O}(\mathbf{N}^{-\infty}),\quad B_{(k)}B_{(\ell)}^{*}=\mathcal{O}(\mathbf{N}^{-\infty})\quad\text{when}\quad\operatorname{supp}\psi_{k}\cap\operatorname{supp}\psi_{\ell}=\emptyset.

    This follows from the nonintersecting support property of semiclassical calculus, where we use that the symbols ψk\psi_{k}, ψ\psi_{\ell} grow by 𝐍ρ2\mathbf{N}^{{\rho\over 2}} with each differentiation and thus belong to both the calculi SL±,ρ,ρ2(𝕋2n)S_{L_{\pm},\rho,{\rho\over 2}}(\mathbb{T}^{2n}) where the symbols b±b_{\pm} lie. By the Cotlar–Stein Theorem, this reduces the estimate (1.19) to a bound on the norms of the individual summands

    Op𝐍,θ(b)Op𝐍,θ(ψk)Op𝐍,θ(b+)=𝒪(𝐍β)for someβ>0.\|\operatorname{Op}_{\mathbf{N},\theta}(b_{-})\operatorname{Op}_{\mathbf{N},\theta}(\psi_{k})\operatorname{Op}_{\mathbf{N},\theta}(b_{+})\|_{\mathcal{H}\to\mathcal{H}}=\mathcal{O}(\mathbf{N}^{-\beta})\quad\text{for some}\quad\beta>0. (1.21)
  • Each of the localized symbols b±ψkb_{\pm}\psi_{k} does satisfy the porosity condition, namely π±(supp(b±ψk))\pi_{\pm}(\operatorname{supp}(b_{\pm}\psi_{k})) is porous on scales C𝐍ρC\mathbf{N}^{-\rho} to 1. This is because a localized version of the observation (3) above holds: if z,w2nz,w\in\mathbb{R}^{2n} and zwLz-w\in L_{-} has length 𝐍ρ2\leq\mathbf{N}^{-{\rho\over 2}}, then for 0jT10\leq j\leq T_{1} the difference AjzAjwA^{j}z-A^{j}w is small. Indeed, by (1.16) the expansion rate of AjA^{j} in the direction of LL_{-} is slower than |λ+|j2|\lambda_{+}|^{j\over 2}, and 𝐍ρ2|λ+|T121\mathbf{N}^{-{\rho\over 2}}|\lambda_{+}|^{T_{1}\over 2}\leq 1. (See Figure 4.) On the other hand, an analogue of the observation (2) above holds as well: the closure of the projection of E+E_{+} onto 𝕋2n\mathbb{T}^{2n} is equal to the subtorus 𝕋+\mathbb{T}_{+}, all of whose shifts intersect the complement of suppb2\operatorname{supp}b_{2} (see the beginning of §1.4). Thus the one-dimensional fractal uncertainty principle can be applied to give (1.21) and thus finish the proof.

In the above argument we again fix ρ(12,23)\rho\in({1\over 2},{2\over 3}). More precisely we need ρ>12\rho>{1\over 2} so that the fractal uncertainty principle can be applied (i.e. we get porosity down to some scale below 𝐍12\mathbf{N}^{-{1\over 2}}), and we need ρ<23\rho<{2\over 3} so that semiclassical calculus could be used to get the almost orthogonality property.

There are several differences of the above outline from the actual proof of the key estimate. First of all, since we make the weaker spectral gap assumption (1.3) instead of (1.16), we need to revise the choice of T1T_{1} and the parameters ρ,ρ\rho,\rho^{\prime} in the calculus – see §3.1.1. Secondly, at a certain point in the argument (see §3.5) we pass from the toric quantization Op𝐍,θ\operatorname{Op}_{\mathbf{N},\theta} to the Weyl quantization Oph\operatorname{Op}_{h} on n\mathbb{R}^{n}, with h:=(2π𝐍)1h:=(2\pi\mathbf{N})^{-1}.

1.5. Structure of the paper

  • In §2 we review semiclassical quantization and metaplectic operators on 2n\mathbb{R}^{2n} and on the torus 𝕋2n\mathbb{T}^{2n}, as well as the anisotropic symbol classes SL,ρ,ρS_{L,\rho,\rho^{\prime}}.

  • In §3 we prove Theorems 24 modulo the key estimate given by Proposition 3.10.

  • In §4 we prove Proposition 3.10 using the fractal uncertainty principle.

  • Finally, in Appendix A, we discuss the algebraic properties of the spaces V+V_{+} and VV_{-} and investigate the sharpness of Theorem 4 in different situations.

Notation: if (F,F)(F,\|\bullet\|_{F}) is a normed vector space and fhFf_{h}\in F is a family depending on a parameter h>0h>0, then we say that fh=𝒪(hα)Ff_{h}=\mathcal{O}(h^{\alpha})_{F} if fhF=𝒪(hα)\|f_{h}\|_{F}=\mathcal{O}(h^{\alpha}). We write fh=𝒪g(hα)Ff_{h}=\mathcal{O}_{g}(h^{\alpha})_{F} if the constant in 𝒪()\mathcal{O}(\bullet) depends on some additional parameter(s) gg.

2. Preliminaries

2.1. Semiclassical quantization

Here we briefly review semiclassical quantization on n\mathbb{R}^{n}, referring the reader to the book of Zworski [Zwo12] for details.

2.1.1. Basic properties

Let h(0,1]h\in(0,1] be the semiclassical parameter. Throughout the paper we denote by Oph\operatorname{Op}_{h} the Weyl quantization on n\mathbb{R}^{n}. For a Schwartz class symbol a𝒮(2n)a\in\mathscr{S}(\mathbb{R}^{2n}) it can be defined as follows [Zwo12, §4.1.1]:

Oph(a)f(x)=(2πh)n2neihxx,ξa(x+x2,ξ)f(x)𝑑x𝑑ξ,f𝒮(n).\operatorname{Op}_{h}(a)f(x)=(2\pi h)^{-n}\int_{\mathbb{R}^{2n}}e^{{i\over h}\langle x-x^{\prime},\xi\rangle}a\big{(}\textstyle{x+x^{\prime}\over 2},\xi\big{)}f(x^{\prime})\,dx^{\prime}d\xi,\quad f\in\mathscr{S}(\mathbb{R}^{n}). (2.1)

One can extend this definition to aa belonging to the set of symbols (see [Zwo12, §4.4.1])

S(1):={aC(2n):sup|(x,ξ)αa|<for all multiindicesα}.S(1):=\{a\in C^{\infty}(\mathbb{R}^{2n})\colon\sup|\partial^{\alpha}_{(x,\xi)}a|<\infty\ \text{for all multiindices}\ \alpha\}. (2.2)

A natural set of seminorms on S(1)S(1) is given by

aCm:=max|α|msup2n|(x,ξ)αa|,m0.\|a\|_{C^{m}}:=\max_{|\alpha|\leq m}\sup_{\mathbb{R}^{2n}}|\partial^{\alpha}_{(x,\xi)}a|,\quad m\in\mathbb{N}_{0}. (2.3)

For any aS(1)a\in S(1), the operator Oph(a)\operatorname{Op}_{h}(a) acts on the space of Schwartz functions 𝒮(n)\mathscr{S}(\mathbb{R}^{n}) and on the dual space of tempered distributions 𝒮(n)\mathscr{S}^{\prime}(\mathbb{R}^{n}), see [Zwo12, Theorem 4.16].

We define the standard symplectic form σ\sigma on 2n\mathbb{R}^{2n} by

σ(z,w):=ξ,yx,ηfor allz=(x,ξ),w=(y,η)2n\sigma(z,w):=\langle\xi,y\rangle-\langle x,\eta\rangle\quad\text{for all}\quad z=(x,\xi),\ w=(y,\eta)\in\mathbb{R}^{2n} (2.4)

where ,\langle\bullet,\bullet\rangle denotes the Euclidean inner product on n\mathbb{R}^{n}.

We now list several standard properties of the Weyl quantization:

  • Composition formula [Zwo12, Theorems 4.11–4.12, 4.17–4.18]: for a,bS(1)a,b\in S(1)

    Oph(a)Oph(b)=Oph(a#b)for some h-dependenta#bS(1),\operatorname{Op}_{h}(a)\operatorname{Op}_{h}(b)=\operatorname{Op}_{h}(a\#b)\quad\text{for some $h$-dependent}\quad a\#b\in S(1), (2.5)

    the function a#ba\#b is bounded in S(1)S(1) uniformly in hh and satisfies the following asymptotic expansion as h0h\to 0 for all N0N\in\mathbb{N}_{0}:

    a#b(z)=k=0N1(ih)k2kk!(σ(z,w)k(a(z)b(w)))|w=z+𝒪(hN)S(1)a\#b(z)=\sum_{k=0}^{N-1}{(-ih)^{k}\over 2^{k}k!}\big{(}\sigma(\partial_{z},\partial_{w})^{k}(a(z)b(w))\big{)}\big{|}_{w=z}+\mathcal{O}(h^{N})_{S(1)} (2.6)

    where σ(z,w)=ξ,yx,η\sigma(\partial_{z},\partial_{w})=\langle\partial_{\xi},\partial_{y}\rangle-\langle\partial_{x},\partial_{\eta}\rangle is a second order differential operator in z=(x,ξ)z=(x,\xi), w=(y,η)w=(y,\eta). Each CmC^{m}-seminorm of the 𝒪(hN)\mathcal{O}(h^{N}) remainder can be bounded by CaCNbCNhNC\|a\|_{C^{N^{\prime}}}\|b\|_{C^{N^{\prime}}}h^{N} for some CC and NN^{\prime} depending only on n,N,mn,N,m.

  • Adjoint formula [Zwo12, Theorem 4.1]: if aS(1)a\in S(1), then

    Oph(a)=Oph(a¯).\operatorname{Op}_{h}(a)^{*}=\operatorname{Op}_{h}(\overline{a}). (2.7)
  • L2L^{2}-boundedness [Zwo12, Theorem 4.23]: For aS(1)a\in S(1), the operator Oph(a):L2(n)L2(n)\operatorname{Op}_{h}(a):L^{2}(\mathbb{R}^{n})\to L^{2}(\mathbb{R}^{n}) is bounded uniformly in hh, more precisely its norm is bounded by CaCNC\|a\|_{C^{N}} for some C,NC,N depending only on nn.

  • Sharp Gårding inequality [Zwo12, Theorem 4.32]: if aS(1)a\in S(1) and a(x,ξ)0a(x,\xi)\geq 0 for all (x,ξ)(x,\xi), then

    Oph(a)f,fL2CahfL22for allfL2(n)\langle\operatorname{Op}_{h}(a)f,f\rangle_{L^{2}}\geq-C_{a}h\|f\|_{L^{2}}^{2}\quad\text{for all}\quad f\in L^{2}(\mathbb{R}^{n})

    where the constant CaC_{a} has the form CaCNC\|a\|_{C^{N}} for some C,NC,N depending only on nn.

The above properties imply in particular the nonintersecting support property

Oph(a)Oph(b)=𝒪(h)L2(n)L2(n)ifa,bS(1),suppasuppb=.\operatorname{Op}_{h}(a)\operatorname{Op}_{h}(b)=\mathcal{O}(h^{\infty})_{L^{2}(\mathbb{R}^{n})\to L^{2}(\mathbb{R}^{n})}\quad\text{if}\quad a,b\in S(1),\quad\operatorname{supp}a\cap\operatorname{supp}b=\emptyset. (2.8)

The next lemma gives an improved version of (2.8) in a special case when suppa,suppb\operatorname{supp}a,\operatorname{supp}b are separated far from each other in some direction. Lemma 2.1 is needed for the proof of Lemma 2.4, which is itself used in the proof of Lemma 4.5.

Lemma 2.1.

Let q:2nq:\mathbb{R}^{2n}\to\mathbb{R} be a linear functional of norm 1 and assume that a,bS(1)a,b\in S(1) satisfy for some r>0r>0

suppa{z2nq(z)r},suppb{z2nq(z)r}.\operatorname{supp}a\subset\{z\in\mathbb{R}^{2n}\mid q(z)\leq-r\},\quad\operatorname{supp}b\subset\{z\in\mathbb{R}^{2n}\mid q(z)\geq r\}. (2.9)

Then for each N>0N>0 we have

Oph(a)Oph(b)L2(n)L2(n)CNhN(1+r)N\|\operatorname{Op}_{h}(a)\operatorname{Op}_{h}(b)\|_{L^{2}(\mathbb{R}^{n})\to L^{2}(\mathbb{R}^{n})}\leq C_{N}h^{N}(1+r)^{-N} (2.10)

where the constant CNC_{N} depends only on nn, NN, and some seminorms aCN\|a\|_{C^{N^{\prime}}}, bCN\|b\|_{C^{N^{\prime}}} where NN^{\prime} depends only on n,Nn,N.

Proof.

If 0<r<10<r<1, then (2.10) follows immediately from (2.8). The constant CNC_{N} does not depend on rr or qq: we only use that all the terms in the expansion (2.6) for a#ba\#b are equal to 0. Thus we henceforth assume that r1r\geq 1.

1. Fix a function

FC(;(0,)),F(ρ)={ρ,ifρ1;|ρ|1,ifρ1.F\in C^{\infty}(\mathbb{R};(0,\infty)),\quad F(\rho)=\begin{cases}\rho,&\text{if}\quad\rho\geq 1;\\ |\rho|^{-1},&\text{if}\quad\rho\leq-1.\end{cases}

Then there exists a constant C0C_{0} such that

F(ρ)C0F(ρ)(1+|ρρ|2)for allρ,ρ.F(\rho^{\prime})\leq C_{0}F(\rho)(1+|\rho-\rho^{\prime}|^{2})\quad\text{for all}\quad\rho,\rho^{\prime}\in\mathbb{R}. (2.11)

Now define the function

𝐦:=FqC(2n;(0,)).\mathbf{m}:=F\circ q\ \in\ C^{\infty}(\mathbb{R}^{2n};(0,\infty)).

It follows from (2.11) that the function 𝐦\mathbf{m}, as well as any its power 𝐦s\mathbf{m}^{s}, is an order function in the sense of [Zwo12, §4.4.1].

Following [Zwo12, (4.4.2)], we define for an order function 𝐦~\widetilde{\mathbf{m}} the class of symbols S(𝐦~)S(\widetilde{\mathbf{m}}) consisting of functions aC(2n)a\in C^{\infty}(\mathbb{R}^{2n}) satisfying the derivative bounds

|zαa(z)|Cα𝐦~(z)for allz2n.|\partial^{\alpha}_{z}a(z)|\leq C_{\alpha}\widetilde{\mathbf{m}}(z)\quad\text{for all}\quad z\in\mathbb{R}^{2n}.

The space S(𝐦~)S(\widetilde{\mathbf{m}}) has a natural family of seminorms defined similarly to (2.3).

2. Take arbitrary NN\in\mathbb{N}. From the support property (2.9) we see that

a=𝒪(rN)S(𝐦N),b=𝒪(rN)S(𝐦N)a=\mathcal{O}(r^{-N})_{S(\mathbf{m}^{-N})},\quad b=\mathcal{O}(r^{-N})_{S(\mathbf{m}^{N})}

where the constants in 𝒪()\mathcal{O}(\bullet) depend only on some S(1)S(1)-seminorms of a,ba,b. The composition formula (2.6) is true for aS(𝐦N)a\in S(\mathbf{m}^{-N}), bS(𝐦N)b\in S(\mathbf{m}^{N}) with the remainder in the expansion still in S(1)S(1), see [Zwo12, Theorems 4.17–4.18]. Since suppasuppb=\operatorname{supp}a\cap\operatorname{supp}b=\emptyset, all the terms in the asymptotic expansion (2.6) are zero, so in particular

Oph(a)Oph(b)=Oph(a#b)wherea#b=𝒪(hNr2N)S(1).\operatorname{Op}_{h}(a)\operatorname{Op}_{h}(b)=\operatorname{Op}_{h}(a\#b)\quad\text{where}\quad a\#b=\mathcal{O}(h^{N}r^{-2N})_{S(1)}.

Now (2.10) follows from the L2L^{2}-boundedness property of quantizations of symbols in S(1)S(1). ∎

2.1.2. Quantum translations

For w2nw\in\mathbb{R}^{2n}, consider the operator

Uw:=Oph(aw)whereaw(z)=exp(ihσ(w,z)).U_{w}:=\operatorname{Op}_{h}(a_{w})\quad\text{where}\quad a_{w}(z)=\exp\big{(}\textstyle{i\over h}\sigma(w,z)\big{)}. (2.12)

Here the symbol awa_{w} lies in S(1)S(1), though its seminorms are not bounded uniformly in hh. By [Zwo12, Theorem 4.7] we have

Uwf(x)=eihη,xi2hy,ηf(xy),f𝒮(n)wherew=(y,η).U_{w}f(x)=e^{{i\over h}\langle\eta,x\rangle-{i\over 2h}\langle y,\eta\rangle}f(x-y),\quad f\in\mathscr{S}^{\prime}(\mathbb{R}^{n})\quad\text{where}\quad w=(y,\eta). (2.13)

In particular, UwU_{w} is a unitary operator on L2(n)L^{2}(\mathbb{R}^{n}).

We call UwU_{w} a quantum translation because it satisfies the following exact Egorov’s Theorem (which is easy to check for a𝒮(2n)a\in\mathscr{S}(\mathbb{R}^{2n}) using (2.1) and extends to the general case by density):

Uw1Oph(a)Uw=Oph(a~)for allaS(1),a~(z):=a(z+w).U_{w}^{-1}\operatorname{Op}_{h}(a)U_{w}=\operatorname{Op}_{h}(\tilde{a})\quad\text{for all}\quad a\in S(1),\quad\tilde{a}(z):=a(z+w). (2.14)

The map wUww\mapsto U_{w} is not a group homomorphism, instead we have

UwUw=ei2hσ(w,w)Uw+w.U_{w}U_{w^{\prime}}=e^{{i\over 2h}\sigma(w,w^{\prime})}U_{w+w^{\prime}}. (2.15)

This in particular implies the commutator formula

UwUw=eihσ(w,w)UwUw.U_{w}U_{w^{\prime}}=e^{{i\over h}\sigma(w,w^{\prime})}U_{w^{\prime}}U_{w}. (2.16)

2.1.3. Metaplectic transformations

Denote by Sp(2n,)\operatorname{Sp}(2n,\mathbb{R}) the group of real symplectic 2n×2n2n\times 2n matrices, i.e. linear isomorphisms A:2n2nA:\mathbb{R}^{2n}\to\mathbb{R}^{2n} such that, with the symplectic form σ\sigma defined in (2.4),

σ(Az,Aw)=σ(z,w)for allz,w2n.\sigma(Az,Aw)=\sigma(z,w)\quad\text{for all}\quad z,w\in\mathbb{R}^{2n}.

For each ASp(2n,)A\in\operatorname{Sp}(2n,\mathbb{R}), denote by A\mathcal{M}_{A} the set of all unitary transformations M:L2(n)L2(n)M:L^{2}(\mathbb{R}^{n})\to L^{2}(\mathbb{R}^{n}) satisfying the following exact Egorov’s theorem:

M1Oph(a)M=Oph(aA)for allaS(1).M^{-1}\operatorname{Op}_{h}(a)M=\operatorname{Op}_{h}(a\circ A)\quad\text{for all}\quad a\in S(1). (2.17)

Such transformations exist and are unique up to multiplication by a unit complex number, see [Zwo12, Theorem 11.9]. Moreover, they act 𝒮(n)𝒮(n)\mathscr{S}(\mathbb{R}^{n})\to\mathscr{S}(\mathbb{R}^{n}) and 𝒮(n)𝒮(n)\mathscr{S}^{\prime}(\mathbb{R}^{n})\to\mathscr{S}^{\prime}(\mathbb{R}^{n}). The set

:=ASp(2n,)A\mathcal{M}:=\bigcup_{A\in\operatorname{Sp}(2n,\mathbb{R})}\mathcal{M}_{A}

is a subgroup of the group of unitary transformations of L2(n)L^{2}(\mathbb{R}^{n}) called the metaplectic group and the map MAM\mapsto A is a group homomorphism Sp(2n,)\mathcal{M}\to\operatorname{Sp}(2n,\mathbb{R}).

The metaplectic transformations and the quantum translations are intertwined by the following corollary of (2.17):

M1UwM=UA1wfor allMA,w2n.M^{-1}U_{w}M=U_{A^{-1}w}\quad\text{for all}\quad M\in\mathcal{M}_{A},\ w\in\mathbb{R}^{2n}. (2.18)

2.1.4. Symbol calculus associated to a coisotropic space

We now introduce an exotic calculus corresponding to a linear foliation by coisotropic spaces. For each subspace L2nL\subset\mathbb{R}^{2n}, define its symplectic complement LσL^{\perp\sigma} as

Lσ={w2nσ(w,z)=0 for all zL}.L^{\perp\sigma}=\{w\in\mathbb{R}^{2n}\mid\sigma(w,z)=0\text{ for all }z\in L\}. (2.19)

Note that LσL^{\perp\sigma} is a subspace of dimension 2ndimL2n-\dim L.

We call LL coisotropic if LσLL^{\perp\sigma}\subset L. This in particular implies that dimLn\dim L\geq n. The case dimL=n\dim L=n corresponds to LL being Lagrangian, that is Lσ=LL^{\perp\sigma}=L. On the opposite end, both the whole 2n\mathbb{R}^{2n} and any codimension 1 subspace of it are coisotropic.

Our exotic calculus will use symbols in the class SL,ρ,ρS_{L,\rho,\rho^{\prime}}, defined as follows:

Definition 2.2.

Let L2nL\subset\mathbb{R}^{2n} be a coisotropic subspace. Fix

0ρρsuch thatρ+ρ<1.0\leq\rho^{\prime}\leq\rho\quad\text{such that}\quad\rho+\rho^{\prime}<1.

We say that an hh-dependent symbol a(x,ξ;h)C(2n)a(x,\xi;h)\in C^{\infty}(\mathbb{R}^{2n}) lies in SL,ρ,ρ(2n)S_{L,\rho,\rho^{\prime}}(\mathbb{R}^{2n}) if for any choice of constant vector fields X1,,Xk,Y1,,YmX_{1},\dots,X_{k},Y_{1},\dots,Y_{m} on 2n\mathbb{R}^{2n} such that Y1,,YmY_{1},\dots,Y_{m} are tangent to LL there exists a constant CC such that for all h(0,1]h\in(0,1]

sup2n|X1XkY1Yma|Chρkρm.\sup_{\mathbb{R}^{2n}}|X_{1}\dots X_{k}Y_{1}\dots Y_{m}a|\leq Ch^{-\rho k-\rho^{\prime}m}. (2.20)

Remarks. 1. The derivative bounds (2.20) can be interpreted as follows: the symbol aa can grow by hρh^{-\rho^{\prime}} when differentiated along LL and by hρh^{-\rho} when differentiated in other directions.

2. If LL is Lagrangian, then a version of the class SL,ρ,ρS_{L,\rho,\rho^{\prime}} corresponding to compactly supported symbols but an arbitrary (not necessarily constant) Lagrangian foliation previously appeared in [DJ18] which inspired part of the argument in the present paper. In the important special case ρ=0\rho^{\prime}=0 this class was introduced in [DZ16].

3. In the case ρ=ρ<12\rho=\rho^{\prime}<{1\over 2} the class SL,ρ,ρS_{L,\rho,\rho^{\prime}} does not depend on LL and becomes the standard mildly exotic pseudodifferential class Sρ(1)S_{\rho}(1), see [Zwo12, §4.4.1]. In particular, if ρ=ρ=0\rho=\rho^{\prime}=0 then we recover the class S(1)S(1) defined in (2.2).

4. In [DZ16, DJ18] the value of ρ\rho was taken close to 1 and ρ\rho^{\prime} was either 0 or very small. In the present paper we choose ρ,ρ\rho,\rho^{\prime} in a more complicated way depending on the size of the spectral gap of the matrix AA, see §3.1.1.

Each aSL,ρ,ρ(2n)a\in S_{L,\rho,\rho^{\prime}}(\mathbb{R}^{2n}) lies in S(1)S(1) for any fixed value of h>0h>0. Therefore, the quantization Oph(a)\operatorname{Op}_{h}(a) defines an operator on L2(n)L^{2}(\mathbb{R}^{n}). The S(1)S(1)-seminorms of aa are not bounded uniformly as h0h\to 0, so the standard properties of quantization from §2.1.1 do not apply. However, using that ρ+ρ<1\rho+\rho^{\prime}<1 and the fact that LL is coisotropic, we can establish analogues of these properties with weaker remainders:

Lemma 2.3.

1. Assume that a,bSL,ρ,ρ(2n)a,b\in S_{L,\rho,\rho^{\prime}}(\mathbb{R}^{2n}). Then Oph(a)Oph(b)=Oph(a#b)\operatorname{Op}_{h}(a)\operatorname{Op}_{h}(b)=\operatorname{Op}_{h}(a\#b) where a#bSL,ρ,ρ(2n)a\#b\in S_{L,\rho,\rho^{\prime}}(\mathbb{R}^{2n}) satisfies the following asymptotic expansion as h0h\to 0 for all N0N\in\mathbb{N}_{0}:

a#b(z)=k=0N1(ih)k2kk!(σ(z,w)k(a(z)b(w)))|w=z+𝒪(h(1ρρ)N)SL,ρ,ρ.a\#b(z)=\sum_{k=0}^{N-1}{(-ih)^{k}\over 2^{k}k!}\big{(}\sigma(\partial_{z},\partial_{w})^{k}(a(z)b(w))\big{)}\big{|}_{w=z}+\mathcal{O}(h^{(1-\rho-\rho^{\prime})N})_{S_{L,\rho,\rho^{\prime}}}. (2.21)

2. Assume that aSL,ρ,ρ(2n)a\in S_{L,\rho,\rho^{\prime}}(\mathbb{R}^{2n}). Then Oph(a)L2L2\|\operatorname{Op}_{h}(a)\|_{L^{2}\to L^{2}} is bounded uniformly in h(0,1]h\in(0,1].

3. Assume that aSL,ρ,ρ(2n)a\in S_{L,\rho,\rho^{\prime}}(\mathbb{R}^{2n}) and a0a\geq 0 everywhere. Then there exists a constant CC such that

Oph(a)f,fL2Ch1ρρfL22for allfL2(n),0<h1.\langle\operatorname{Op}_{h}(a)f,f\rangle_{L^{2}}\geq-Ch^{1-\rho-\rho^{\prime}}\|f\|_{L^{2}}^{2}\quad\text{for all}\quad f\in L^{2}(\mathbb{R}^{n}),\quad 0<h\leq 1. (2.22)

The constants in the above estimates depend only on certain SL,ρ,ρS_{L,\rho,\rho^{\prime}}-seminorms of a,ba,b similarly to the properties of the S(1)S(1)-calculus in §2.1.1, where the choice of the seminorms additionally depends on ρ,ρ\rho,\rho^{\prime}.

Proof.

1. Let dimL=n+p\dim L=n+p. Since LL is coisotropic, there exists a linear symplectomorphism

ASp(2n,)such thatA1(L)=L0:=span(x1,,xn,ξ1,,ξp).A\in\operatorname{Sp}(2n,\mathbb{R})\quad\text{such that}\quad A^{-1}(L)=L_{0}:=\operatorname{span}(\partial_{x_{1}},\dots,\partial_{x_{n}},\partial_{\xi_{1}},\dots,\partial_{\xi_{p}}). (2.23)

Denote ξ=(ξ,ξ′′)\xi=(\xi^{\prime},\xi^{\prime\prime}) where ξp\xi^{\prime}\in\mathbb{R}^{p} and ξ′′np\xi^{\prime\prime}\in\mathbb{R}^{n-p}. For a(x,ξ;h)C(2n)a(x,\xi;h)\in C^{\infty}(\mathbb{R}^{2n}), we have

aSL,ρ,ρ(2n)aASL0,ρ,ρ(2n)a\in S_{L,\rho,\rho^{\prime}}(\mathbb{R}^{2n})\ \iff\ a\circ A\in S_{L_{0},\rho,\rho^{\prime}}(\mathbb{R}^{2n})

where the space SL0,ρ,ρ(2n)S_{L_{0},\rho,\rho^{\prime}}(\mathbb{R}^{2n}) can be characterized by the following inequalities for all multiindices α,β\alpha,\beta:

sup2n|(x,ξ)αξ′′βa|Cαβhρ|α|ρ|β|.\sup_{\mathbb{R}^{2n}}|\partial^{\alpha}_{(x,\xi^{\prime})}\partial^{\beta}_{\xi^{\prime\prime}}a|\leq C_{\alpha\beta}h^{-\rho^{\prime}|\alpha|-\rho|\beta|}. (2.24)

Fix a metaplectic operator MAM\in\mathcal{M}_{A}. Then by (2.17) (which applies since the symbols in question are in S(1)S(1) for any fixed hh)

M1Oph(a)M=Oph(aA)for allaSL,ρ,ρ(2n).M^{-1}\operatorname{Op}_{h}(a)M=\operatorname{Op}_{h}(a\circ A)\quad\text{for all}\quad a\in S_{L,\rho,\rho^{\prime}}(\mathbb{R}^{2n}).

Since MM is unitary on L2(n)L^{2}(\mathbb{R}^{n}) and the terms in the expansion (2.21) are equivariant under composing a,ba,b with AA, we reduce the proof of the lemma to the case L=L0L=L_{0}.

2. Henceforth we assume that L=L0L=L_{0}. We use the unitary rescaling operator TT on L2(n)L^{2}(\mathbb{R}^{n}) defined by

Tf(x)=hρn2f(hρx).Tf(x)=h^{-{\rho^{\prime}n\over 2}}f(h^{-\rho^{\prime}}x).

A direct calculation using (2.1) shows that for all aSL0,ρ,ρ(2n)a\in S_{L_{0},\rho,\rho^{\prime}}(\mathbb{R}^{2n}) we have

T1Oph(a)T=Oph~(a~)whereh~:=h1ρρ,a~(x,ξ):=a(hρx,hρξ).T^{-1}\operatorname{Op}_{h}(a)T=\operatorname{Op}_{\tilde{h}}(\tilde{a})\quad\text{where}\quad\tilde{h}:=h^{1-\rho-\rho^{\prime}},\quad\tilde{a}(x,\xi):=a(h^{\rho^{\prime}}x,h^{\rho}\xi). (2.25)

By (2.24) we see that a~S(1)\tilde{a}\in S(1) uniformly in hh. This gives parts 2–3 of the lemma.

3. To show part 1 of the lemma, define the symbol a#~ba\widetilde{\#}b by the formula Oph~(a)Oph~(b)=Oph~(a#~b)\operatorname{Op}_{\tilde{h}}(a)\operatorname{Op}_{\tilde{h}}(b)=\operatorname{Op}_{\tilde{h}}(a\widetilde{\#}b) and consider the rescaling map B(x,ξ)=(hρx,hρξ)B(x,\xi)=(h^{\rho^{\prime}}x,h^{\rho}\xi). Then by (2.25)

(a#b)B=a~#~b~wherea~:=aB,b~:=bB.(a\#b)\circ B=\tilde{a}\widetilde{\#}\tilde{b}\quad\text{where}\quad\tilde{a}:=a\circ B,\quad\tilde{b}:=b\circ B. (2.26)

Define the remainder

rN(z):=a#b(z)k=0N1(ih)k2kk!(σ(z,w)k(a(z)b(w)))|w=z,r_{N}(z):=a\#b(z)-\sum_{k=0}^{N-1}{(-ih)^{k}\over 2^{k}k!}\big{(}\sigma(\partial_{z},\partial_{w})^{k}(a(z)b(w))\big{)}\big{|}_{w=z},

then by (2.26)

rN(B(z))=a~#~b~(z)k=0N1(ih~)k2kk!(σ(z,w)k(a~(z)b~(w)))|w=z.r_{N}(B(z))=\tilde{a}\widetilde{\#}\tilde{b}(z)-\sum_{k=0}^{N-1}{(-i\tilde{h})^{k}\over 2^{k}k!}\big{(}\sigma(\partial_{z},\partial_{w})^{k}(\tilde{a}(z)\tilde{b}(w))\big{)}\big{|}_{w=z}.

Since a~,b~\tilde{a},\tilde{b} are bounded in S(1)S(1) uniformly in hh, we see from (2.6) with hh replaced by h~\tilde{h} that

rN(B(z))=𝒪(h~N)S(1)r_{N}(B(z))=\mathcal{O}(\tilde{h}^{N})_{S(1)}

which implies that for all multiindices α,β\alpha,\beta

sup2n|xαξβrN|=𝒪(h(1ρρ)Nρ|α|ρ|β|).\sup_{\mathbb{R}^{2n}}|\partial^{\alpha}_{x}\partial^{\beta}_{\xi}r_{N}|=\mathcal{O}(h^{(1-\rho-\rho^{\prime})N-\rho^{\prime}|\alpha|-\rho|\beta|}). (2.27)

If p=0p=0, i.e. LL is Lagrangian, then this immediately gives the expansion (2.21) as (2.27) shows that rN=𝒪(h(1ρρ)N)SL0,ρ,ρr_{N}=\mathcal{O}(h^{(1-\rho-\rho^{\prime})N})_{S_{L_{0},\rho,\rho^{\prime}}}. In the general case, since 1ρρ>01-\rho-\rho^{\prime}>0 we get that all the derivatives of the remainder become rapidly decaying in hh when NN\to\infty, that is for each N0,α,βN_{0},\alpha,\beta there exists NN such that sup2n|xαξβrN|=𝒪(hN0)\sup_{\mathbb{R}^{2n}}|\partial^{\alpha}_{x}\partial^{\beta}_{\xi}r_{N}|=\mathcal{O}(h^{N_{0}}). Combining this with the fact that the kk-th term in (2.21) is 𝒪(h(1ρρ)k)SL0,ρ,ρ\mathcal{O}(h^{(1-\rho-\rho^{\prime})k})_{S_{L_{0},\rho,\rho^{\prime}}} we see that the expansion (2.21) holds. ∎

Remark. Lemma 2.3 does not hold for the standard quantization

Oph0(a)f(x)=(2πh)n2neihxx,ξa(x,ξ)f(x)𝑑x𝑑ξ.\operatorname{Op}^{0}_{h}(a)f(x)=(2\pi h)^{-n}\int_{\mathbb{R}^{2n}}e^{{i\over h}\langle x-x^{\prime},\xi\rangle}a(x,\xi)f(x^{\prime})\,dx^{\prime}d\xi.

Indeed, consider the case n=1n=1 and L=span(x+ξ)L=\operatorname{span}(\partial_{x}+\partial_{\xi}). The analog of the expansion (2.21) for the standard quantization is [Zwo12, Theorem 4.14]

Oph0(a)Oph0(b)=Oph0(a#0b),a#0b(x,ξ)k=0(ih)kk!ξka(x,ξ)xkb(x,ξ)ash0.\operatorname{Op}^{0}_{h}(a)\operatorname{Op}^{0}_{h}(b)=\operatorname{Op}^{0}_{h}(a\#^{0}b),\quad a\#^{0}b(x,\xi)\sim\sum_{k=0}^{\infty}{(-ih)^{k}\over k!}\partial^{k}_{\xi}a(x,\xi)\partial^{k}_{x}b(x,\xi)\quad\text{as}\quad h\to 0.

Take ρ:=0\rho^{\prime}:=0 and put a(x,ξ)=χ(hρ(xξ))a(x,\xi)=\chi(h^{-\rho}(x-\xi)) for some nonzero hh-independent function χCc()\chi\in C^{\infty}_{\mathrm{c}}(\mathbb{R}). Then aSL,ρ,0(2)a\in S_{L,\rho,0}(\mathbb{R}^{2}) and the kk-th term in the asymptotic expansion for a#0aa\#^{0}a is

ikk!(χ(k)(hρ(xξ)))2h(12ρ)k.{i^{k}\over k!}\big{(}\chi^{(k)}(h^{-\rho}(x-\xi))\big{)}^{2}h^{(1-2\rho)k}.

For 12<ρ<1{1\over 2}<\rho<1, each successive term in the expansion grows faster in hh than the previous one, which makes it impossible for this expansion to hold. The difference between the standard and the Weyl quantization exploited in the proof of Lemma 2.3 is that the Weyl quantization obeys the exact Egorov Theorem (2.17) and the related fact that the terms in the asymptotic expansion (2.21) are equivariant under Sp(2n,)\operatorname{Sp}(2n,\mathbb{R}).

We also prove here a statement on composition of operators whose symbols have well-separated supports, used in the proof of Lemma 4.5:

Lemma 2.4.

Let q:2nq:\mathbb{R}^{2n}\to\mathbb{R} be a linear functional of norm 1 and assume that a,bSL,ρ,ρ(2n)a,b\in S_{L,\rho,\rho^{\prime}}(\mathbb{R}^{2n}) satisfy for some r0r_{0}\in\mathbb{R} and r>0r>0

suppa{z2nq(z)r0r},suppb{z2nq(z)r0+r}.\operatorname{supp}a\subset\{z\in\mathbb{R}^{2n}\mid q(z)\leq r_{0}-r\},\quad\operatorname{supp}b\subset\{z\in\mathbb{R}^{2n}\mid q(z)\geq r_{0}+r\}. (2.28)

Then for each cSL,ρ,ρ(2n)c\in S_{L,\rho,\rho^{\prime}}(\mathbb{R}^{2n}) and N>0N>0 we have

Oph(a)Oph(c)Oph(b)L2(n)L2(n)CNhN(1+r)N\|\operatorname{Op}_{h}(a)\operatorname{Op}_{h}(c)\operatorname{Op}_{h}(b)\|_{L^{2}(\mathbb{R}^{n})\to L^{2}(\mathbb{R}^{n})}\leq C_{N}h^{N}(1+r)^{-N} (2.29)

where the constant CNC_{N} depends only on nn, NN, and some n,Nn,N-dependent SL,ρ,ρ(2n)S_{L,\rho,\rho^{\prime}}(\mathbb{R}^{2n})-seminorms of a,b,ca,b,c.

Proof.

1. We first show that for all a,ba,b satisfying (2.28) and all NN

Oph(a)Oph(b)L2(n)L2(n)CNhN(1+r)N.\|\operatorname{Op}_{h}(a)\operatorname{Op}_{h}(b)\|_{L^{2}(\mathbb{R}^{n})\to L^{2}(\mathbb{R}^{n})}\leq C_{N}h^{N}(1+r)^{-N}. (2.30)

We may shift the supports of a,ba,b by conjugating by a quantum translation (see (2.14)), so we may assume that r0=0r_{0}=0. We may also conjugate by a metaplectic transformation similarly to Step 1 in the proof of Lemma 2.3 to reduce to the case L=L0L=L_{0} where L0L_{0} is defined in (2.23). Next, using the dilation formula (2.25) we see that

Oph(a)Oph(b)L2(n)L2(n)=Oph~(a~)Oph~(b~)L2(n)L2(n)\|\operatorname{Op}_{h}(a)\operatorname{Op}_{h}(b)\|_{L^{2}(\mathbb{R}^{n})\to L^{2}(\mathbb{R}^{n})}=\|\operatorname{Op}_{\tilde{h}}(\tilde{a})\operatorname{Op}_{\tilde{h}}(\tilde{b})\|_{L^{2}(\mathbb{R}^{n})\to L^{2}(\mathbb{R}^{n})}

where h~=h1ρρ\tilde{h}=h^{1-\rho-\rho^{\prime}} and the rescaled symbols a~(x,ξ)=a(hρx,hρξ)\tilde{a}(x,\xi)=a(h^{\rho^{\prime}}x,h^{\rho}\xi), b~(x,ξ)=b(hρx,hρξ)\tilde{b}(x,\xi)=b(h^{\rho^{\prime}}x,h^{\rho}\xi) lie in S(1)S(1) uniformly in hh. The support condition (2.28) implies that

suppa~{z2nq~(z)r~},suppb~{z2nq~(z)r~}\operatorname{supp}\tilde{a}\subset\{z\in\mathbb{R}^{2n}\mid\tilde{q}(z)\leq-\tilde{r}\},\quad\operatorname{supp}\tilde{b}\subset\{z\in\mathbb{R}^{2n}\mid\tilde{q}(z)\geq\tilde{r}\}

where we put

q(x,ξ):=q(hρx,hρξ),q~:=qq,r~:=rq.q^{\prime}(x,\xi):=q(h^{\rho^{\prime}}x,h^{\rho}\xi),\quad\tilde{q}:={q^{\prime}\over\|q^{\prime}\|},\quad\tilde{r}:={r\over\|q^{\prime}\|}.

Note that q1\|q^{\prime}\|\leq 1 and thus r~r\tilde{r}\geq r.

We now apply Lemma 2.1 to a~,b~,h~,r~\tilde{a},\tilde{b},\tilde{h},\tilde{r} to get for all NN

Oph~(a~)Oph~(b~)L2(n)L2(n)CNh~N(1+r~)NCNhN(1ρρ)(1+r)N\|\operatorname{Op}_{\tilde{h}}(\tilde{a})\operatorname{Op}_{\tilde{h}}(\tilde{b})\|_{L^{2}(\mathbb{R}^{n})\to L^{2}(\mathbb{R}^{n})}\leq C_{N}\tilde{h}^{N}(1+\tilde{r})^{-N}\leq C_{N}h^{N(1-\rho-\rho^{\prime})}(1+r)^{-N}

which implies (2.30) since 1ρρ>01-\rho-\rho^{\prime}>0.

2. We now prove (2.29). If 0<r<10<r<1 then (2.29) follows by applying twice the composition formula (2.21) and using the L2L^{2} boundedness property of the class SL,ρ,ρS_{L,\rho,\rho^{\prime}} and the fact that suppasuppb=\operatorname{supp}a\cap\operatorname{supp}b=\emptyset. Henceforth we assume that r1r\geq 1. Fix

χC(;[0,1]),χ=1 on [12,),χ=0 on (,12]\chi\in C^{\infty}(\mathbb{R};[0,1]),\quad\chi=1\text{ on }[\textstyle{1\over 2},\infty),\quad\chi=0\text{ on }(-\infty,-\textstyle{1\over 2}]

and decompose

c=c1+c2,c1(z):=c(z)χ(q(z)r0r),c2(z):=c(z)(1χ(q(z)r0r)).c=c_{1}+c_{2},\quad c_{1}(z):=c(z)\chi\Big{(}{q(z)-r_{0}\over r}\Big{)},\quad c_{2}(z):=c(z)\bigg{(}1-\chi\Big{(}{q(z)-r_{0}\over r}\Big{)}\bigg{)}.

Then the symbols c1,c2c_{1},c_{2} are bounded in SL,ρ,ρ(2n)S_{L,\rho,\rho^{\prime}}(\mathbb{R}^{2n}) uniformly in rr. Moreover

suppc1{z2nq(z)r0r2},suppc2{z2nq(z)r0+r2}.\operatorname{supp}c_{1}\subset\{z\in\mathbb{R}^{2n}\mid q(z)\geq r_{0}-\textstyle{r\over 2}\},\quad\operatorname{supp}c_{2}\subset\{z\in\mathbb{R}^{2n}\mid q(z)\leq r_{0}+\textstyle{r\over 2}\}.

We now write Oph(c)=Oph(c1)+Oph(c2)\operatorname{Op}_{h}(c)=\operatorname{Op}_{h}(c_{1})+\operatorname{Op}_{h}(c_{2}) and estimate (using L2L^{2} boundedness of Oph(a)\operatorname{Op}_{h}(a), Oph(b)\operatorname{Op}_{h}(b))

Oph(a)Oph(c)Oph(b)L2L2C(Oph(a)Oph(c1)L2L2+Oph(c2)Oph(b)L2L2).\|\operatorname{Op}_{h}(a)\operatorname{Op}_{h}(c)\operatorname{Op}_{h}(b)\|_{L^{2}\to L^{2}}\leq C\big{(}\|\operatorname{Op}_{h}(a)\operatorname{Op}_{h}(c_{1})\|_{L^{2}\to L^{2}}+\|\operatorname{Op}_{h}(c_{2})\operatorname{Op}_{h}(b)\|_{L^{2}\to L^{2}}\big{)}.

We finally use (2.30) with r0r_{0} replaced by r0±3r4r_{0}\pm{3r\over 4} and rr replaced by r4r\over 4 to get

Oph(a)Oph(c1)L2L2,Oph(c2)Oph(b)L2L2CNhN(1+r)N\|\operatorname{Op}_{h}(a)\operatorname{Op}_{h}(c_{1})\|_{L^{2}\to L^{2}},\|\operatorname{Op}_{h}(c_{2})\operatorname{Op}_{h}(b)\|_{L^{2}\to L^{2}}\leq C_{N}h^{N}(1+r)^{-N}

which finishes the proof. ∎

2.2. Quantization on the torus

In this section we study quantizations of functions on the torus

𝕋2n:=2n/2n.\mathbb{T}^{2n}:=\mathbb{R}^{2n}/\mathbb{Z}^{2n}.

Each aC(𝕋2n)a\in C^{\infty}(\mathbb{T}^{2n}) can be identified with a 2n\mathbb{Z}^{2n}-periodic function on 2n\mathbb{R}^{2n}. This function lies in the symbol class S(1)S(1) defined in (2.2) and thus its Weyl quantization Oph(a)\operatorname{Op}_{h}(a) is an operator on L2(n)L^{2}(\mathbb{R}^{n}). We will decompose L2(n)L^{2}(\mathbb{R}^{n}) into a direct integral of finite dimensional spaces 𝐍(θ)\mathcal{H}_{\mathbf{N}}(\theta), θ𝕋2n\theta\in\mathbb{T}^{2n}, which we call the spaces of quantum states. The operator Oph(a)\operatorname{Op}_{h}(a) descends to these spaces and gives a quantization of the observable aC(𝕋2n)a\in C^{\infty}(\mathbb{T}^{2n}). Our presentation partially follows [BDB96].

To make sure that the spaces of quantum states are nontrivial, we henceforth make the following assumption (see [BDB96, Proposition 2.1]):

h=12π𝐍for some𝐍.h={1\over 2\pi\mathbf{N}}\quad\text{for some}\quad\mathbf{N}\in\mathbb{N}. (2.31)

2.2.1. The spaces of quantum states

Recall the quantum translations UwU_{w}, w2nw\in\mathbb{R}^{2n}, defined in §2.1.2. By (2.14) we have the commutation relations

Oph(a)Uw=UwOph(a)for allaC(𝕋2n),w2n.\operatorname{Op}_{h}(a)U_{w}=U_{w}\operatorname{Op}_{h}(a)\quad\text{for all}\quad a\in C^{\infty}(\mathbb{T}^{2n}),\quad w\in\mathbb{Z}^{2n}. (2.32)

This motivates the following definition of the spaces of quantum states: for each θ𝕋2n\theta\in\mathbb{T}^{2n}, put

𝐍(θ):={f𝒮(n)Uwf=e2πiσ(θ,w)+𝐍πiQ(w)f for all w2n}\mathcal{H}_{\mathbf{N}}(\theta):=\{f\in\mathscr{S}^{\prime}(\mathbb{R}^{n})\mid U_{w}f=e^{2\pi i\sigma(\theta,w)+\mathbf{N}\pi iQ(w)}f\text{ for all }w\in\mathbb{Z}^{2n}\} (2.33)

where the quadratic form QQ on n\mathbb{R}^{n} is defined by

Q(w):=y,ηwherew=(y,η)2n.Q(w):=\langle y,\eta\rangle\quad\text{where}\quad w=(y,\eta)\in\mathbb{R}^{2n}. (2.34)

Denote

𝐍:={0,,𝐍1}.\mathbb{Z}_{\mathbf{N}}:=\{0,\dots,\mathbf{N}-1\}.

The following description of the spaces 𝐍(θ)\mathcal{H}_{\mathbf{N}}(\theta) is a higher dimensional version of [BDB96, Proposition 2.1]:

Lemma 2.5.

The space 𝐍(θ)\mathcal{H}_{\mathbf{N}}(\theta) is 𝐍n{\mathbf{N}}^{n}-dimensional with a basis given by 𝐞jθ\mathbf{e}^{\theta}_{j}, j𝐍nj\in\mathbb{Z}_{\mathbf{N}}^{n}, where for θ=(θx,θξ)2n\theta=(\theta_{x},\theta_{\xi})\in\mathbb{R}^{2n} we define

𝐞jθ(x):=𝐍n2kne2πiθξ,kδ(x𝐍k+jθx𝐍),jn.\mathbf{e}^{\theta}_{j}(x):=\mathbf{N}^{-{n\over 2}}\sum_{k\in\mathbb{Z}^{n}}e^{-2\pi i\langle\theta_{\xi},k\rangle}\delta\bigg{(}x-{\mathbf{N}k+j-\theta_{x}\over\mathbf{N}}\bigg{)},\quad j\in\mathbb{Z}^{n}. (2.35)

Remark. The distributions 𝐞jθ\mathbf{e}^{\theta}_{j} satisfy the identities

𝐞jθ+w\displaystyle\mathbf{e}^{\theta+w}_{j} =𝐞jyθfor allw=(y,η)2n,\displaystyle=\mathbf{e}^{\theta}_{j-y}\quad\text{for all}\quad w=(y,\eta)\in\mathbb{Z}^{2n}, (2.36)
𝐞j+𝐍θ\displaystyle\mathbf{e}^{\theta}_{j+\mathbf{N}\ell} =e2πiθξ,𝐞jθfor alln.\displaystyle=e^{2\pi i\langle\theta_{\xi},\ell\rangle}\mathbf{e}^{\theta}_{j}\quad\text{for all}\quad\ell\in\mathbb{Z}^{n}.

In particular, even though the space 𝐍(θ)\mathcal{H}_{\mathbf{N}}(\theta) is canonically defined for θ\theta in the torus 𝕋2n\mathbb{T}^{2n}, in order to define the basis {𝐞jθ}\{\mathbf{e}^{\theta}_{j}\} we need to fix a representative θxn\theta_{x}\in\mathbb{R}^{n}. Note also that 𝐞jθ\mathbf{e}^{\theta}_{j} is supported on the shifted lattice 𝐍1(jθx)+n\mathbf{N}^{-1}(j-\theta_{x})+\mathbb{Z}^{n}.

Proof.

By (2.16), for each v2nv\in\mathbb{R}^{2n} the quantum translation UvU_{v} is an isomorphism from 𝐍(θ)\mathcal{H}_{\mathbf{N}}(\theta) onto 𝐍(θ𝐍v)\mathcal{H}_{\mathbf{N}}(\theta-\mathbf{N}v). On the other hand, we compute for all jnj\in\mathbb{Z}^{n} and v2nv\in\mathbb{R}^{2n}

Uv𝐞jθ=e2πiη,jθx+iπ𝐍Q(v)𝐞jθ𝐍vwherev=(y,η).U_{v}\mathbf{e}^{\theta}_{j}=e^{2\pi i\langle\eta,j-\theta_{x}\rangle+i\pi{\mathbf{N}}Q(v)}\mathbf{e}_{j}^{\theta-{\mathbf{N}}v}\quad\text{where}\quad v=(y,\eta). (2.37)

Thus it suffices to consider the case θ=0\theta=0.

Using w=(0,)w=(0,\ell) and w=(,0)w=(\ell,0) in the definition (2.33), as well as (2.15), we can characterize 𝐍(0)\mathcal{H}_{\mathbf{N}}(0) as the set of all f𝒮(n)f\in\mathscr{S}^{\prime}(\mathbb{R}^{n}) such that

e2πi𝐍,xf(x)=f(x),f(x)=f(x)for alln.e^{2\pi i\mathbf{N}\langle\ell,x\rangle}f(x)=f(x),\quad f(x-\ell)=f(x)\quad\text{for all}\quad\ell\in\mathbb{Z}^{n}. (2.38)

The first condition in (2.38) is equivalent to ff being a linear combination of delta functions at the points in the lattice 𝐍1n\mathbf{N}^{-1}\mathbb{Z}^{n}, that is

f(x)=rnfrδ(xr𝐍)for some(fr)rn.f(x)=\sum_{r\in\mathbb{Z}^{n}}f_{r}\delta\Big{(}x-{r\over\mathbf{N}}\Big{)}\quad\text{for some}\quad(f_{r}\in\mathbb{C})_{r\in\mathbb{Z}^{n}}.

The second condition in (2.38) is then equivalent to the periodic property fr𝐍=frf_{r-\mathbf{N}\ell}=f_{r} for all n\ell\in\mathbb{Z}^{n}. It follows that 𝐍(0)\mathcal{H}_{\mathbf{N}}(0) is the span of {𝐞j0j𝐍n}\{\mathbf{e}^{0}_{j}\mid j\in\mathbb{Z}_{\mathbf{N}}^{n}\}, which finishes the proof. ∎

We fix the inner product ,\langle\bullet,\bullet\rangle_{\mathcal{H}} on each 𝐍(θ)\mathcal{H}_{\mathbf{N}}(\theta) by requiring that {𝐞jθ}j𝐍n\{\mathbf{e}^{\theta}_{j}\}_{j\in\mathbb{Z}^{n}_{\mathbf{N}}} be an orthonormal basis. Note that while the basis {𝐞jθ}\{\mathbf{e}^{\theta}_{j}\} depends on the choice of the representative θxn\theta_{x}\in\mathbb{R}^{n}, the inner product only depends on θ𝕋2n\theta\in\mathbb{T}^{2n} as follows from (2.36).

Using the bases {𝐞jθ}\{\mathbf{e}^{\theta}_{j}\}, we can consider the spaces 𝐍(θ)\mathcal{H}_{\mathbf{N}}(\theta) as the fibers of a smooth 𝐍n\mathbf{N}^{n}-dimensional complex vector bundle over 𝕋2n\mathbb{T}^{2n}, which we denote by 𝐍\mathcal{H}_{\mathbf{N}}.

2.2.2. Decomposing L2L^{2}

We now construct a unitary isomorphism Π𝐍\Pi_{\mathbf{N}} between L2(n)L^{2}(\mathbb{R}^{n}) and the space of L2L^{2} sections L2(𝕋2n;𝐍)L^{2}(\mathbb{T}^{2n};\mathcal{H}_{\mathbf{N}}). This gives a decomposition of L2(n)L^{2}(\mathbb{R}^{n}) into the direct integral of the spaces 𝐍(θ)\mathcal{H}_{\mathbf{N}}(\theta) over θ𝕋2n\theta\in\mathbb{T}^{2n}.

Define the operators Π𝐍(θ):𝒮(n)𝐍(θ)\Pi_{\mathbf{N}}(\theta):\mathscr{S}(\mathbb{R}^{n})\to\mathcal{H}_{\mathbf{N}}(\theta) by

Π𝐍(θ)f:=j𝐍nf,𝐞jθL2𝐞jθ,θ𝕋2n.\Pi_{\mathbf{N}}(\theta)f:=\sum_{j\in\mathbb{Z}_{\mathbf{N}}^{n}}\langle f,\mathbf{e}^{\theta}_{j}\rangle_{L^{2}}\,\mathbf{e}^{\theta}_{j},\quad\theta\in\mathbb{T}^{2n}. (2.39)

Even though the basis {𝐞jθ}\{\mathbf{e}_{j}^{\theta}\} depends on the choice of the preimage θxn\theta_{x}\in\mathbb{R}^{n}, the operator Π𝐍(θ)\Pi_{\mathbf{N}}(\theta) does not depend on this choice as follows from (2.36). We next define the operator

Π𝐍:𝒮(n)C(𝕋2n;𝐍),Π𝐍f:=(Π𝐍(θ)f)θ𝕋2n.\Pi_{\mathbf{N}}:\mathscr{S}(\mathbb{R}^{n})\to C^{\infty}(\mathbb{T}^{2n};\mathcal{H}_{\mathbf{N}}),\quad\Pi_{\mathbf{N}}f:=(\Pi_{\mathbf{N}}(\theta)f)_{\theta\in\mathbb{T}^{2n}}.

We also define the operator Π𝐍:C(𝕋2n;𝐍)𝒮(n)\Pi_{\mathbf{N}}^{*}:C^{\infty}(\mathbb{T}^{2n};\mathcal{H}_{\mathbf{N}})\to\mathscr{S}(\mathbb{R}^{n}) as follows: for gC(𝕋2n;𝐍)g\in C^{\infty}(\mathbb{T}^{2n};\mathcal{H}_{\mathbf{N}})

Π𝐍g(x):=𝐍n2𝕋ng(𝐍x,θξ),𝐞0(𝐍x,θξ)𝑑θξ,xn.\begin{gathered}\Pi_{\mathbf{N}}^{*}g(x):={\mathbf{N}}^{{n\over 2}}\int_{\mathbb{T}^{n}}\big{\langle}g(-\mathbf{N}x,\theta_{\xi}),\mathbf{e}^{(-\mathbf{N}x,\theta_{\xi})}_{0}\big{\rangle}_{\mathcal{H}}\,d\theta_{\xi},\quad x\in\mathbb{R}^{n}.\end{gathered}

Here one can check that Π𝐍g𝒮(n)\Pi_{\mathbf{N}}^{*}g\in\mathscr{S}(\mathbb{R}^{n}) using a non-stationary phase argument and the following corollary of (2.36):

Π𝐍g(x)=Π𝐍(e2πi,θξg)(x)for alln.\Pi_{\mathbf{N}}^{*}g(x-\ell)=\Pi_{\mathbf{N}}^{*}(e^{2\pi i\langle\ell,\theta_{\xi}\rangle}g)(x)\quad\text{for all}\quad\ell\in\mathbb{Z}^{n}.
Lemma 2.6.

The map Π𝐍\Pi_{\mathbf{N}} extends to a unitary isomorphism from L2(n)L^{2}(\mathbb{R}^{n}) to L2(𝕋2n;𝐍)L^{2}(\mathbb{T}^{2n};\mathcal{H}_{\mathbf{N}}) and Π𝐍\Pi_{\mathbf{N}}^{*} extends to its adjoint.

Proof.

We argue similarly to [BDB96, Proposition 2.3].

1. We first show that Π𝐍\Pi_{\mathbf{N}} extends to an isometry from L2(n)L^{2}(\mathbb{R}^{n}) to L2(𝕋2n;𝐍)L^{2}(\mathbb{T}^{2n};\mathcal{H}_{\mathbf{N}}). For that it suffices to show the identity

𝕋2nΠ𝐍(θ)f2𝑑θ=fL2(n)2for allf𝒮(n).\int_{\mathbb{T}^{2n}}\|\Pi_{\mathbf{N}}(\theta)f\|_{\mathcal{H}}^{2}\,d\theta=\|f\|_{L^{2}(\mathbb{R}^{n})}^{2}\quad\text{for all}\quad f\in\mathscr{S}(\mathbb{R}^{n}). (2.40)

For jnj\in\mathbb{Z}^{n} and θxn\theta_{x}\in\mathbb{R}^{n}, define the function Fj,θxC(𝕋n)F_{j,\theta_{x}}\in C^{\infty}(\mathbb{T}^{n}) by

Fj,θx(θξ)=f,𝐞jθL2whereθ=(θx,θξ).F_{j,\theta_{x}}(\theta_{\xi})=\langle f,\mathbf{e}_{j}^{\theta}\rangle_{L^{2}}\quad\text{where}\quad\theta=(\theta_{x},\theta_{\xi}).

Then Fj,θxF_{j,\theta_{x}} can be written as a Fourier series:

Fj,θx(θξ)=𝐍n2knf(𝐍k+jθx𝐍)e2πiθξ,k.F_{j,\theta_{x}}(\theta_{\xi})={\mathbf{N}}^{-{n\over 2}}\sum_{k\in\mathbb{Z}^{n}}f\Big{(}{\mathbf{N}k+j-\theta_{x}\over\mathbf{N}}\Big{)}e^{2\pi i\langle\theta_{\xi},k\rangle}.

Therefore by Parseval’s Theorem

𝕋n|Fj,θx(θξ)|2𝑑θξ=𝐍nkn|f(𝐍k+jθx𝐍)|2.\int_{\mathbb{T}^{n}}|F_{j,\theta_{x}}(\theta_{\xi})|^{2}\,d\theta_{\xi}=\mathbf{N}^{-n}\sum_{k\in\mathbb{Z}^{n}}\bigg{|}f\Big{(}{\mathbf{N}k+j-\theta_{x}\over\mathbf{N}}\Big{)}\bigg{|}^{2}.

We then have

𝕋2nΠ𝐍(θ)f2𝑑θ\displaystyle\int_{\mathbb{T}^{2n}}\|\Pi_{\mathbf{N}}(\theta)f\|_{\mathcal{H}}^{2}\,d\theta =[0,1]n𝕋nj𝐍n|Fj,θx(θξ)|2dθξdθx\displaystyle=\int_{[0,1]^{n}}\int_{\mathbb{T}^{n}}\sum_{j\in\mathbb{Z}_{\mathbf{N}}^{n}}|F_{j,\theta_{x}}(\theta_{\xi})|^{2}\,d\theta_{\xi}d\theta_{x}
=𝐍nj𝐍n,kn[0,1]n|f(𝐍k+jθx𝐍)|2𝑑θx\displaystyle=\mathbf{N}^{-n}\sum_{j\in\mathbb{Z}_{\mathbf{N}}^{n},k\in\mathbb{Z}^{n}}\int_{[0,1]^{n}}\bigg{|}f\Big{(}{\mathbf{N}k+j-\theta_{x}\over\mathbf{N}}\Big{)}\bigg{|}^{2}\,d\theta_{x}
=𝐍nrn[0,1]n|f(rθx𝐍)|2𝑑θx=fL2(n)2\displaystyle=\mathbf{N}^{-n}\sum_{r\in\mathbb{Z}^{n}}\int_{[0,1]^{n}}\bigg{|}f\Big{(}{r-\theta_{x}\over\mathbf{N}}\Big{)}\bigg{|}^{2}\,d\theta_{x}=\|f\|_{L^{2}(\mathbb{R}^{n})}^{2}

which gives (2.40).

2. It remains to show that Π𝐍\Pi_{\mathbf{N}} is onto and Π𝐍\Pi_{\mathbf{N}}^{*} is the adjoint of Π𝐍\Pi_{\mathbf{N}}. For that it suffices to prove that for each gC(𝕋2n;𝐍)g\in C^{\infty}(\mathbb{T}^{2n};\mathcal{H}_{\mathbf{N}}) we have Π𝐍Π𝐍g=g\Pi_{\mathbf{N}}\Pi_{\mathbf{N}}^{*}g=g. We compute for all θ=(θx,θξ)2n\theta=(\theta_{x},\theta_{\xi})\in\mathbb{R}^{2n} and jnj\in\mathbb{Z}^{n}

Π𝐍g,𝐞jθL2\displaystyle\langle\Pi_{\mathbf{N}}^{*}g,\mathbf{e}_{j}^{\theta}\rangle_{L^{2}} =𝐍n2kne2πiθξ,kΠ𝐍g(𝐍k+jθx𝐍)\displaystyle=\mathbf{N}^{-{n\over 2}}\sum_{k\in\mathbb{Z}^{n}}e^{2\pi i\langle\theta_{\xi},k\rangle}\Pi_{\mathbf{N}}^{*}g\Big{(}{\mathbf{N}k+j-\theta_{x}\over\mathbf{N}}\Big{)}
=kne2πiθξ,k𝕋ng(θx,θ~ξ),𝐞0(θx𝐍kj,θ~ξ)𝑑θ~ξ\displaystyle=\sum_{k\in\mathbb{Z}^{n}}e^{2\pi i\langle\theta_{\xi},k\rangle}\int_{\mathbb{T}^{n}}\langle g(\theta_{x},\tilde{\theta}_{\xi}),\mathbf{e}_{0}^{(\theta_{x}-\mathbf{N}k-j,\tilde{\theta}_{\xi})}\rangle_{\mathcal{H}}\,d\tilde{\theta}_{\xi}
=g(θ),𝐞jθ.\displaystyle=\langle g(\theta),\mathbf{e}^{\theta}_{j}\rangle_{\mathcal{H}}.

Here in the last line we use that 𝐞0(θx𝐍kj,θ~ξ)=e2πiθ~ξ,k𝐞j(θx,θ~ξ)\mathbf{e}_{0}^{(\theta_{x}-\mathbf{N}k-j,\tilde{\theta}_{\xi})}=e^{2\pi i\langle\tilde{\theta}_{\xi},k\rangle}\mathbf{e}_{j}^{(\theta_{x},\tilde{\theta}_{\xi})} by (2.36), as well as convergence of the Fourier series of the function θξg(θx,θξ),𝐞j(θx,θξ)\theta_{\xi}\mapsto\langle g(\theta_{x},\theta_{\xi}),\mathbf{e}^{(\theta_{x},\theta_{\xi})}_{j}\rangle_{\mathcal{H}}. We now compute Π𝐍(θ)Π𝐍g=g(θ)\Pi_{\mathbf{N}}(\theta)\Pi_{\mathbf{N}}^{*}g=g(\theta) for all θ𝕋2n\theta\in\mathbb{T}^{2n}, finishing the proof. ∎

By duality, we may extend Π𝐍,Π𝐍\Pi_{\mathbf{N}},\Pi^{*}_{\mathbf{N}} to operators

Π𝐍:𝒮(n)𝒟(𝕋2n;𝐍),Π𝐍:𝒟(𝕋2n;𝐍)𝒮(n).\Pi_{\mathbf{N}}:\mathscr{S}^{\prime}(\mathbb{R}^{n})\to\mathcal{D}^{\prime}(\mathbb{T}^{2n};\mathcal{H}_{\mathbf{N}}),\quad\Pi_{\mathbf{N}}^{*}:\mathcal{D}^{\prime}(\mathbb{T}^{2n};\mathcal{H}_{\mathbf{N}})\to\mathscr{S}^{\prime}(\mathbb{R}^{n}).

We then have the natural formula

Π𝐍(δ(θθ0)f)=ffor allθ0𝕋2n,f𝐍(θ0)𝒮(n),\Pi_{\mathbf{N}}^{*}(\delta(\theta-\theta_{0})f)=f\quad\text{for all}\quad\theta_{0}\in\mathbb{T}^{2n},\quad f\in\mathcal{H}_{\mathbf{N}}(\theta_{0})\subset\mathscr{S}^{\prime}(\mathbb{R}^{n}), (2.41)

which follows by duality from the identity

δ(θθ0)f,Π𝐍f~L2(𝕋2n;𝐍)=f,Π𝐍(θ0)f~=j𝐍nf,𝐞jθ0𝐞jθ0,f~L2=f,f~L2(n)for allf~𝒮(n).\begin{split}\langle\delta(\theta-\theta_{0})f,\Pi_{\mathbf{N}}\tilde{f}\rangle_{L^{2}(\mathbb{T}^{2n};\mathcal{H}_{\mathbf{N}})}&=\langle f,\Pi_{\mathbf{N}}(\theta_{0})\tilde{f}\rangle_{\mathcal{H}}=\sum_{j\in\mathbb{Z}_{\mathbf{N}}^{n}}\langle f,\mathbf{e}^{\theta_{0}}_{j}\rangle_{\mathcal{H}}\,\langle\mathbf{e}^{\theta_{0}}_{j},\tilde{f}\rangle_{L^{2}}\\ &=\langle f,\tilde{f}\rangle_{L^{2}(\mathbb{R}^{n})}\quad\text{for all}\quad\tilde{f}\in\mathscr{S}(\mathbb{R}^{n}).\end{split}

2.2.3. Semiclassical quantization

Fix 𝐍\mathbf{N}\in\mathbb{N} and put h:=(2π𝐍)1h:=(2\pi\mathbf{N})^{-1} as before. Let aC(𝕋2n)a\in C^{\infty}(\mathbb{T}^{2n}). By (2.32) and (2.33), the operator Oph(a)\operatorname{Op}_{h}(a) maps each of the spaces 𝐍(θ)\mathcal{H}_{\mathbf{N}}(\theta) to itself. This defines the quantizations

Op𝐍,θ(a):=Oph(a)|𝐍(θ):𝐍(θ)𝐍(θ),θ𝕋2n\operatorname{Op}_{\mathbf{N},\theta}(a):=\operatorname{Op}_{h}(a)|_{\mathcal{H}_{\mathbf{N}}(\theta)}:\mathcal{H}_{\mathbf{N}}(\theta)\to\mathcal{H}_{\mathbf{N}}(\theta),\quad\theta\in\mathbb{T}^{2n}

which depend smoothly on θ\theta.

A special case is given by a(x,ξ)=a(x)a(x,\xi)=a(x) which is independent of ξ\xi and n\mathbb{Z}^{n}-periodic in xx. In this case Oph(a)\operatorname{Op}_{h}(a) is the multiplication operator by aa (see [Zwo12, Theorem 4.3]), so by (2.35)

Op𝐍,θ(a)𝐞jθ=a(jθx𝐍)𝐞jθfor allθ=(θx,θξ)2n,jn.\operatorname{Op}_{\mathbf{N},\theta}(a)\mathbf{e}^{\theta}_{j}=a\Big{(}{j-\theta_{x}\over\mathbf{N}}\Big{)}\mathbf{e}^{\theta}_{j}\quad\text{for all}\quad\theta=(\theta_{x},\theta_{\xi})\in\mathbb{R}^{2n},\quad j\in\mathbb{Z}^{n}. (2.42)

In particular, Op𝐍,θ(1)\operatorname{Op}_{\mathbf{N},\theta}(1) is the identity operator on 𝐍(θ)\mathcal{H}_{\mathbf{N}}(\theta).

Let Π𝐍,Π𝐍\Pi_{\mathbf{N}},\Pi_{\mathbf{N}}^{*} be the unitary operators constructed in §2.2.2. By (2.41), they relate the operator Oph(a):𝒮(n)𝒮(n)\operatorname{Op}_{h}(a):\mathscr{S}^{\prime}(\mathbb{R}^{n})\to\mathscr{S}^{\prime}(\mathbb{R}^{n}) to its restrictions Op𝐍,θ(a)\operatorname{Op}_{\mathbf{N},\theta}(a) as follows:

Π𝐍Oph(a)Π𝐍g(θ)=Op𝐍,θ(a)g(θ)for allg𝒟(𝕋2n;𝐍).\Pi_{\mathbf{N}}\operatorname{Op}_{h}(a)\Pi_{\mathbf{N}}^{*}g(\theta)=\operatorname{Op}_{\mathbf{N},\theta}(a)g(\theta)\quad\text{for all}\quad g\in\mathcal{D}^{\prime}(\mathbb{T}^{2n};\mathcal{H}_{\mathbf{N}}). (2.43)

Notice that (2.43) may also be deduced from the explicit expression (verified by an explicit computation using (2.13), (2.35), (2.39), and the Poisson summation formula; the series below converges in 𝒮(n)\mathscr{S}^{\prime}(\mathbb{R}^{n}))

Π𝐍(θ)f=w2neiπ𝐍Q(w)2iπσ(θ,w)Uwf for all f𝒮(n)\Pi_{\mathbf{N}}(\theta)f=\sum_{w\in\mathbb{Z}^{2n}}e^{i\pi\mathbf{N}Q(w)-2i\pi\sigma(\theta,w)}U_{w}f\quad\textup{ for all }\quad f\in\mathscr{S}(\mathbb{R}^{n}) (2.44)

and the commutation identity (2.32).

Since Op𝐍,θ(a)\operatorname{Op}_{\mathbf{N},\theta}(a) depends smoothly on θ\theta, it follows from (2.43) and Lemma 2.6 that

maxθ𝕋2nOp𝐍,θ(a)𝐍(θ)𝐍(θ)=Oph(a)L2(n)L2(n).\max_{\theta\in\mathbb{T}^{2n}}\|\operatorname{Op}_{\mathbf{N},\theta}(a)\|_{\mathcal{H}_{\mathbf{N}}(\theta)\to\mathcal{H}_{\mathbf{N}}(\theta)}=\|\operatorname{Op}_{h}(a)\|_{L^{2}(\mathbb{R}^{n})\to L^{2}(\mathbb{R}^{n})}. (2.45)

Recall from Definition 2.2 the symbol class SL,ρ,ρ(2n)S_{L,\rho,\rho^{\prime}}(\mathbb{R}^{2n}) where L2nL\subset\mathbb{R}^{2n} is a coisotropic subspace and 0ρρ0\leq\rho^{\prime}\leq\rho, ρ+ρ<1\rho+\rho^{\prime}<1. We similarly define the corresponding symbol class on the torus

SL,ρ,ρ(𝕋2n)S_{L,\rho,\rho^{\prime}}(\mathbb{T}^{2n})

whose elements are the 2n\mathbb{Z}^{2n}-periodic symbols in SL,ρ,ρ(2n)S_{L,\rho,\rho^{\prime}}(\mathbb{R}^{2n}). Note that putting ρ=ρ=0\rho=\rho^{\prime}=0 we obtain the standard symbol class S(𝕋2n)S(\mathbb{T}^{2n}) consisting of functions in C(𝕋2n)C^{\infty}(\mathbb{T}^{2n}) with all derivatives bounded uniformly in hh.

Using (2.43) and (2.45), we see that Lemma 2.3 applies to the quantization Op𝐍,θ(a)\operatorname{Op}_{\mathbf{N},\theta}(a). In particular, we have the product formula for all a,bSL,ρ,ρ(𝕋2n)a,b\in S_{L,\rho,\rho^{\prime}}(\mathbb{T}^{2n})

Op𝐍,θ(a)Op𝐍,θ(b)=Op𝐍,θ(a#b)\operatorname{Op}_{\mathbf{N},\theta}(a)\operatorname{Op}_{\mathbf{N},\theta}(b)=\operatorname{Op}_{\mathbf{N},\theta}(a\#b) (2.46)

where a#bSL,ρ,ρ(𝕋2n)a\#b\in S_{L,\rho,\rho^{\prime}}(\mathbb{T}^{2n}) satisfies the expansion (2.21), the adjoint formula (following from (2.7))

Op𝐍,θ(a)=Op𝐍,θ(a¯),\operatorname{Op}_{\mathbf{N},\theta}(a)^{*}=\operatorname{Op}_{\mathbf{N},\theta}(\bar{a}), (2.47)

the norm Op𝐍,θ(a)𝐍(θ)𝐍(θ)\|\operatorname{Op}_{\mathbf{N},\theta}(a)\|_{\mathcal{H}_{\mathbf{N}}(\theta)\to\mathcal{H}_{\mathbf{N}}(\theta)} is bounded by some SL,ρ,ρS_{L,\rho,\rho^{\prime}}-seminorm of aa, and we have the sharp Gårding inequality for all aSL,ρ,ρ(𝕋2n)a\in S_{L,\rho,\rho^{\prime}}(\mathbb{T}^{2n}) such that a0a\geq 0 everywhere

Op𝐍,θ(a)f,fC𝐍ρ+ρ1f2for allf𝐍(θ)\langle\operatorname{Op}_{\mathbf{N},\theta}(a)f,f\rangle_{\mathcal{H}}\geq-C\mathbf{N}^{\rho+\rho^{\prime}-1}\|f\|_{\mathcal{H}}^{2}\quad\text{for all}\quad f\in\mathcal{H}_{\mathbf{N}}(\theta) (2.48)

where CC is some SL,ρ,ρS_{L,\rho,\rho^{\prime}}-seminorm of aa. The choice of seminorms above depends on ρ,ρ\rho,\rho^{\prime} but not on 𝐍\mathbf{N} or θ\theta. The inequality (2.48) follows from the usual sharp Gårding inequality, and Lemma 2.6 that implies that Op𝐍,θ(a)\operatorname{Op}_{\mathbf{N},\theta}(a) is self-adjoint and that its spectrum is contained in the L2L^{2} spectrum of Oph(a)\operatorname{Op}_{h}(a).

We now give several corollaries of the basic calculus above. First of all, from (2.46), the expansion (2.21) with N=1N=1, and the boundeness of the operator norm of Op𝐍,θ()\operatorname{Op}_{\mathbf{N},\theta}(\bullet) we get for all a,bSL,ρ,ρ(𝕋2n)a,b\in S_{L,\rho,\rho^{\prime}}(\mathbb{T}^{2n})

Op𝐍,θ(a)Op𝐍,θ(b)=Op𝐍,θ(ab)+𝒪(𝐍ρ+ρ1)𝐍(θ)𝐍(θ)\operatorname{Op}_{\mathbf{N},\theta}(a)\operatorname{Op}_{\mathbf{N},\theta}(b)=\operatorname{Op}_{\mathbf{N},\theta}(ab)+\mathcal{O}(\mathbf{N}^{\rho+\rho^{\prime}-1})_{\mathcal{H}_{\mathbf{N}}(\theta)\to\mathcal{H}_{\mathbf{N}}(\theta)} (2.49)

where the constant in 𝒪()\mathcal{O}(\bullet) depends only on some SL,ρ,ρS_{L,\rho,\rho^{\prime}}-seminorms of a,ba,b.

Next, we have the following inequality of norms:

Lemma 2.7.

Assume that a,bSL,ρ,ρ(𝕋2n)a,b\in S_{L,\rho,\rho^{\prime}}(\mathbb{T}^{2n}) and |a||b||a|\leq|b| everywhere. Then we have for all f𝐍(θ)f\in\mathcal{H}_{\mathbf{N}}(\theta)

Op𝐍,θ(a)fOp𝐍,θ(b)f+C𝐍ρ+ρ12f\|\operatorname{Op}_{\mathbf{N},\theta}(a)f\|_{\mathcal{H}}\leq\|\operatorname{Op}_{\mathbf{N},\theta}(b)f\|_{\mathcal{H}}+C\mathbf{N}^{\rho+\rho^{\prime}-1\over 2}\|f\|_{\mathcal{H}} (2.50)

where the constant CC depends only on some SL,ρ,ρS_{L,\rho,\rho^{\prime}}-seminorms of a,ba,b.

Remark. Taking bb to be the constant symbol b:=sup|a|b:=\sup|a|, we get

Op𝐍,θ(a)𝐍(θ)𝐍(θ)supz𝕋2n|a(z)|+C𝐍ρ+ρ12\|\operatorname{Op}_{\mathbf{N},\theta}(a)\|_{\mathcal{H}_{\mathbf{N}}(\theta)\to\mathcal{H}_{\mathbf{N}}(\theta)}\leq\sup_{z\in\mathbb{T}^{2n}}|a(z)|+C\mathbf{N}^{\rho+\rho^{\prime}-1\over 2} (2.51)

where CC only depends on some SL,ρ,ρS_{L,\rho,\rho^{\prime}}-seminorm of aa.

Proof.

By (2.47) and (2.49) we have

Op𝐍,θ(b)Op𝐍,θ(b)Op𝐍,θ(a)Op𝐍,θ(a)=Op𝐍,θ(|b|2|a|2)+𝒪(𝐍ρ+ρ1)𝐍(θ)𝐍(θ).\begin{gathered}\operatorname{Op}_{\mathbf{N},\theta}(b)^{*}\operatorname{Op}_{\mathbf{N},\theta}(b)-\operatorname{Op}_{\mathbf{N},\theta}(a)^{*}\operatorname{Op}_{\mathbf{N},\theta}(a)\\ =\operatorname{Op}_{\mathbf{N},\theta}(|b|^{2}-|a|^{2})+\mathcal{O}(\mathbf{N}^{\rho+\rho^{\prime}-1})_{\mathcal{H}_{\mathbf{N}}(\theta)\to\mathcal{H}_{\mathbf{N}}(\theta)}.\end{gathered} (2.52)

Since |a||b||a|\leq|b| everywhere, we have |b|2|a|20|b|^{2}-|a|^{2}\geq 0 everywhere, so by (2.48)

Op𝐍,θ(|b|2|a|2)f,fC𝐍ρ+ρ1f2.\langle\operatorname{Op}_{\mathbf{N},\theta}(|b|^{2}-|a|^{2})f,f\rangle_{\mathcal{H}}\geq-C\mathbf{N}^{\rho+\rho^{\prime}-1}\|f\|_{\mathcal{H}}^{2}. (2.53)

Together (2.52) and (2.53) give

Op𝐍,θ(b)f2Op𝐍,θ(a)f2C𝐍ρ+ρ1f2\|\operatorname{Op}_{\mathbf{N},\theta}(b)f\|_{\mathcal{H}}^{2}-\|\operatorname{Op}_{\mathbf{N},\theta}(a)f\|_{\mathcal{H}}^{2}\geq-C\mathbf{N}^{\rho+\rho^{\prime}-1}\|f\|_{\mathcal{H}}^{2}

which implies (2.50). ∎

Finally, we record here the following lemma regarding products of many quantized observables, which is analogous to [DJ18, Lemmas A.1 and A.6]:

Lemma 2.8.

Assume that a1,,aRSL,ρ,ρ(𝕋2n)a_{1},\dots,a_{R}\in S_{L,\rho,\rho^{\prime}}(\mathbb{T}^{2n}), where RC0log𝐍R\leq C_{0}\log\mathbf{N}, satisfy sup𝕋2n|aj|1\sup_{\mathbb{T}^{2n}}|a_{j}|\leq 1 and each SL,ρ,ρS_{L,\rho,\rho^{\prime}}-seminorm of aja_{j} is bounded uniformly in jj. Then:

1. The product a1aRa_{1}\cdots a_{R} lies in SL,ρ+ε,ρ+ε(𝕋2n)S_{L,\rho+\varepsilon,\rho^{\prime}+\varepsilon}(\mathbb{T}^{2n}) for all small ε>0\varepsilon>0.

2. We have for all ε>0\varepsilon>0

Op𝐍,θ(a1)Op𝐍,θ(aR)=Op𝐍,θ(a1aR)+𝒪(𝐍ρ+ρ1+ε)𝐍(θ)𝐍(θ).\operatorname{Op}_{\mathbf{N},\theta}(a_{1})\cdots\operatorname{Op}_{\mathbf{N},\theta}(a_{R})=\operatorname{Op}_{\mathbf{N},\theta}(a_{1}\cdots a_{R})+\mathcal{O}(\mathbf{N}^{\rho+\rho^{\prime}-1+\varepsilon})_{\mathcal{H}_{\mathbf{N}}(\theta)\to\mathcal{H}_{\mathbf{N}}(\theta)}.

Here the constant in 𝒪()\mathcal{O}(\bullet) depends only on ρ,ρ,ε,C0\rho,\rho^{\prime},\varepsilon,C_{0}, and on the maximum over jj of a certain SL,ρ,ρS_{L,\rho,\rho^{\prime}}-seminorm of aja_{j}.

Proof.

1. We have sup𝕋2n|a1aR|1\sup_{\mathbb{T}^{2n}}|a_{1}\cdots a_{R}|\leq 1, so it suffices to show that for each constant vector fields X1,,Xk,Y1,,YmX_{1},\dots,X_{k},Y_{1},\dots,Y_{m} on 𝕋2n\mathbb{T}^{2n} such that Y1,,YmY_{1},\dots,Y_{m} are tangent to LL we have

sup𝕋2n|X1XkY1Ym(a1aR)|=𝒪(hρkρm).\sup_{\mathbb{T}^{2n}}|X_{1}\dots X_{k}Y_{1}\dots Y_{m}(a_{1}\cdots a_{R})|=\mathcal{O}(h^{-\rho k-\rho^{\prime}m-}). (2.54)

Using the Leibniz Rule, we write X1XkY1Ym(a1aR)X_{1}\dots X_{k}Y_{1}\dots Y_{m}(a_{1}\cdots a_{R}) as a sum of Rm+k=𝒪(h0)R^{m+k}=\mathcal{O}(h^{0-}) terms. Each of these is a product of the form (P1a1)(PRaR)(P_{1}a_{1})\cdots(P_{R}a_{R}) where each PjP_{j} is a product of some subset of X1,,Xk,Y1,,YmX_{1},\dots,X_{k},Y_{1},\dots,Y_{m} and P1PR=X1XkY1YmP_{1}\dots P_{R}=X_{1}\dots X_{k}Y_{1}\dots Y_{m}. Using that sup|aj|1\sup|a_{j}|\leq 1 and each aja_{j} is bounded in SL,ρ,ρS_{L,\rho,\rho^{\prime}} uniformly in jj, we get (2.54).

2. We write

Op𝐍,θ(a1)Op𝐍,θ(aR)Op𝐍,θ(a1aR)=j=1RBjOp𝐍,θ(aj+1)Op𝐍,θ(aR)\operatorname{Op}_{\mathbf{N},\theta}(a_{1})\cdots\operatorname{Op}_{\mathbf{N},\theta}(a_{R})-\operatorname{Op}_{\mathbf{N},\theta}(a_{1}\cdots a_{R})=\sum_{j=1}^{R}B_{j}\operatorname{Op}_{\mathbf{N},\theta}(a_{j+1})\cdots\operatorname{Op}_{\mathbf{N},\theta}(a_{R})

where Bj:=Op𝐍,θ(a1aj1)Op𝐍,θ(aj)Op𝐍,θ(a1aj)B_{j}:=\operatorname{Op}_{\mathbf{N},\theta}(a_{1}\cdots a_{j-1})\operatorname{Op}_{\mathbf{N},\theta}(a_{j})-\operatorname{Op}_{\mathbf{N},\theta}(a_{1}\cdots a_{j}). By (2.51) and since sup|aj|1\sup|a_{j}|\leq 1 we have Op𝐍,θ(aj)𝐍(θ)𝐍(θ)1+C𝐍ρ+ρ12\|\operatorname{Op}_{\mathbf{N},\theta}(a_{j})\|_{\mathcal{H}_{\mathbf{N}}(\theta)\to\mathcal{H}_{\mathbf{N}}(\theta)}\leq 1+C\mathbf{N}^{\rho+\rho^{\prime}-1\over 2} where the constant CC is uniform in jj. Therefore (assuming 𝐍\mathbf{N} is large enough)

Op𝐍,θ(a1)Op𝐍,θ(aR)Op𝐍,θ(a1aR)𝐍(θ)𝐍(θ)2j=1RBj𝐍(θ)𝐍(θ).\|\operatorname{Op}_{\mathbf{N},\theta}(a_{1})\cdots\operatorname{Op}_{\mathbf{N},\theta}(a_{R})-\operatorname{Op}_{\mathbf{N},\theta}(a_{1}\cdots a_{R})\|_{\mathcal{H}_{\mathbf{N}}(\theta)\to\mathcal{H}_{\mathbf{N}}(\theta)}\leq 2\sum_{j=1}^{R}\|B_{j}\|_{\mathcal{H}_{\mathbf{N}}(\theta)\to\mathcal{H}_{\mathbf{N}}(\theta)}.

It remains to use that a1aja_{1}\cdots a_{j} is bounded in SL,ρ+ε,ρ+ε(𝕋2n)S_{L,\rho+\varepsilon,\rho^{\prime}+\varepsilon}(\mathbb{T}^{2n}) uniformly in jj for all ε>0\varepsilon>0, so by (2.49) we have Bj𝐍(θ)𝐍(θ)=𝒪(𝐍ρ+ρ1+)\|B_{j}\|_{\mathcal{H}_{\mathbf{N}}(\theta)\to\mathcal{H}_{\mathbf{N}}(\theta)}=\mathcal{O}(\mathbf{N}^{\rho+\rho^{\prime}-1+}) uniformly in jj. ∎

2.2.4. Quantization of toric automorphisms

Let Sp(2n,)Sp(2n,)\operatorname{Sp}(2n,\mathbb{Z})\subset\operatorname{Sp}(2n,\mathbb{R}) be the subgroup of integer symplectic matrices, i.e. elements of Sp(2n,)\operatorname{Sp}(2n,\mathbb{R}) which preserve the lattice 2n\mathbb{Z}^{2n}. We will quantize elements of Sp(2n,)\operatorname{Sp}(2n,\mathbb{Z}) as unitary operators on the spaces 𝐍(θ)\mathcal{H}_{\mathbf{N}}(\theta) provided that 𝐍,θ\mathbf{N},\theta satisfy the condition (2.57) below. To do this we need the following

Lemma 2.9.

Denote by 2:=/2\mathbb{Z}_{2}:=\mathbb{Z}/2\mathbb{Z} the field of order 2. Then for each ASp(2n,)A\in\operatorname{Sp}(2n,\mathbb{Z}) there exists unique φA22n\varphi_{A}\in\mathbb{Z}_{2}^{2n} such that, with the symplectic form σ\sigma defined in (2.4) and the quadratic form QQ defined in (2.34)

Q(A1w)Q(w)=σ(φA,w)mod2for allw2n.Q(A^{-1}w)-Q(w)=\sigma(\varphi_{A},w)\mod 2\mathbb{Z}\quad\text{for all}\quad w\in\mathbb{Z}^{2n}. (2.55)

Remark. Note that the map AφAA\mapsto\varphi_{A} satisfies for all A,BSp(2n,)A,B\in\operatorname{Sp}(2n,\mathbb{Z})

φAB=φA+AφB,φA1=A1φA.\varphi_{AB}=\varphi_{A}+A\varphi_{B},\quad\varphi_{A^{-1}}=A^{-1}\varphi_{A}.
Proof.

For w2nw\in\mathbb{Z}^{2n}, denote

Z(w):=(Q(A1w)Q(w))mod22.Z(w):=\big{(}Q(A^{-1}w)-Q(w)\big{)}\bmod 2\mathbb{Z}\ \in\ \mathbb{Z}_{2}.

We have

Q(w+w)=Q(w)+Q(w)+σ(w,w)mod2for allw,w2nQ(w+w^{\prime})=Q(w)+Q(w^{\prime})+\sigma(w,w^{\prime})\mod 2\mathbb{Z}\quad\text{for all}\quad w,w^{\prime}\in\mathbb{Z}^{2n}

which together with the fact that ASp(2n,)A\in\operatorname{Sp}(2n,\mathbb{Z}) implies that

Z(w+w)=Z(w)+Z(w)for allw,w2n.Z(w+w^{\prime})=Z(w)+Z(w^{\prime})\quad\text{for all}\quad w,w^{\prime}\in\mathbb{Z}^{2n}.

Thus ZZ is a group homomorphism 2n2\mathbb{Z}^{2n}\to\mathbb{Z}_{2}, which gives the existence and uniqueness of φA\varphi_{A} such that (2.55) holds. ∎

Now, fix ASp(2n,)A\in\operatorname{Sp}(2n,\mathbb{Z}) and choose a metaplectic operator (see §2.1.3)

MA.M\in\mathcal{M}_{A}.

Here we put h:=(2π𝐍)1h:=(2\pi\mathbf{N})^{-1} as before. Using (2.18) and (2.33) we see that

M(𝐍(θ))𝐍(Aθ+𝐍φA2)for allθ𝕋2n.M(\mathcal{H}_{\mathbf{N}}(\theta))\subset\mathcal{H}_{\mathbf{N}}\big{(}A\theta+\textstyle{\mathbf{N}\varphi_{A}\over 2}\big{)}\quad\text{for all}\quad\theta\in\mathbb{T}^{2n}.

Denote

M𝐍,θ:=M|𝐍(θ):𝐍(θ)𝐍(Aθ+𝐍φA2)M_{\mathbf{N},\theta}:=M|_{\mathcal{H}_{\mathbf{N}}(\theta)}:\mathcal{H}_{\mathbf{N}}(\theta)\to\mathcal{H}_{\mathbf{N}}\big{(}A\theta+\textstyle{\mathbf{N}\varphi_{A}\over 2}\big{)}

which depends smoothly on θ𝕋2n\theta\in\mathbb{T}^{2n}. Using (2.41) or (2.44), we see that the action of MM on L2(n)L^{2}(\mathbb{R}^{n}) is intertwined with the operators M𝐍,θM_{\mathbf{N},\theta} as follows:

Π𝐍MΠ𝐍g(Aθ+𝐍φA2)=M𝐍,θg(θ)for allg𝒟(𝕋2n;𝐍).\Pi_{\mathbf{N}}M\Pi_{\mathbf{N}}^{*}g\big{(}A\theta+\textstyle{\mathbf{N}\varphi_{A}\over 2}\big{)}=M_{\mathbf{N},\theta}\,g(\theta)\quad\text{for all}\quad g\in\mathcal{D}^{\prime}(\mathbb{T}^{2n};\mathcal{H}_{\mathbf{N}}). (2.56)

Since MM is unitary on L2(n)L^{2}(\mathbb{R}^{n}), it follows that each M𝐍,θM_{\mathbf{N},\theta} is a unitary operator as well.

We will be interested in the spectrum of M𝐍,θM_{\mathbf{N},\theta}, so we need its domain and range to be the same space 𝐍(θ)\mathcal{H}_{\mathbf{N}}(\theta). This is true if we choose θ𝕋2n\theta\in\mathbb{T}^{2n} such that the following quantization condition holds:

(IA)θ=𝐍φA2mod2n.(I-A)\theta={\mathbf{N}\varphi_{A}\over 2}\mod\mathbb{Z}^{2n}. (2.57)

Note that when 𝐍\mathbf{N} is even or φA=0\varphi_{A}=0, the equation (2.57) is satisfied in particular when θ=0\theta=0.

Assuming (2.57), from (2.17) we get the following exact Egorov’s Theorem:

M𝐍,θ1Op𝐍,θ(a)M𝐍,θ=Op𝐍,θ(aA)for allaC(𝕋2n).M_{\mathbf{N},\theta}^{-1}\operatorname{Op}_{\mathbf{N},\theta}(a)M_{\mathbf{N},\theta}=\operatorname{Op}_{\mathbf{N},\theta}(a\circ A)\quad\text{for all}\quad a\in C^{\infty}(\mathbb{T}^{2n}). (2.58)

2.2.5. Explicit formulas

Here we give some explicit formulas for the operators Op𝐍,θ(a)\operatorname{Op}_{\mathbf{N},\theta}(a) and M𝐍,θM_{\mathbf{N},\theta}. These are not used in the proofs but are helpful for implementing numerics. For simplicity we assume in this subsection that θ=0\theta=0.

For f𝐍(0)f\in\mathcal{H}_{\mathbf{N}}(0) define the coordinates

fj:=f,𝐞j0,jnf_{j}:=\langle f,\mathbf{e}^{0}_{j}\rangle_{\mathcal{H}},\quad j\in\mathbb{Z}^{n}

where 𝐞j0\mathbf{e}^{0}_{j} is defined in (2.35). By (2.36) we have fj+𝐍=fjf_{j+\mathbf{N}\ell}=f_{j} for all n\ell\in\mathbb{Z}^{n}.

Our first statement computes the expression Op𝐍,0(a)f,f\langle\operatorname{Op}_{\mathbf{N},0}(a)f,f\rangle_{\mathcal{H}} in terms of the values of aa at the points in the lattice 12𝐍2n{1\over 2\mathbf{N}}\mathbb{Z}^{2n}. As before we define 𝐍:={0,,𝐍1}\mathbb{Z}_{\mathbf{N}}:=\{0,\dots,\mathbf{N}-1\} and similarly 2𝐍:={0,,2𝐍1}\mathbb{Z}_{2\mathbf{N}}:=\{0,\dots,2\mathbf{N}-1\}.

Proposition 2.10.

Let aC(𝕋2n)a\in C^{\infty}(\mathbb{T}^{2n}). Then for all f𝐍(0)f\in\mathcal{H}_{\mathbf{N}}(0) we have

Op𝐍,0(a)f,f=(2𝐍)nj𝐍n,k,2𝐍neπi𝐍k,a(2j+k2𝐍,2𝐍)fjfj+k¯.\langle\operatorname{Op}_{\mathbf{N},0}(a)f,f\rangle_{\mathcal{H}}=(2\mathbf{N})^{-n}\sum_{j\in\mathbb{Z}_{\mathbf{N}}^{n},\ k,\ell\in\mathbb{Z}_{2\mathbf{N}}^{n}}e^{{\pi i\over\mathbf{N}}\langle k,\ell\rangle}a\Big{(}{2j+k\over 2\mathbf{N}},{\ell\over 2\mathbf{N}}\Big{)}f_{j}\overline{f_{j+k}}. (2.59)
Proof.

Since trigonometric polynomials are dense in C(𝕋2n)C^{\infty}(\mathbb{T}^{2n}), it suffices to consider the case of

a(z)=e2πiσ(w,z)for somew=(y,η)2n.a(z)=e^{2\pi i\sigma(w,z)}\quad\text{for some}\quad w=(y,\eta)\in\mathbb{Z}^{2n}.

Using (2.12), (2.37), and (2.36), we compute for all jnj\in\mathbb{Z}^{n}

Op𝐍,0(a)𝐞j0=Uw/𝐍𝐞j0=e2πi𝐍η,j+πi𝐍y,η𝐞j+y0.\operatorname{Op}_{\mathbf{N},0}(a)\mathbf{e}^{0}_{j}=U_{w/\mathbf{N}}\,\mathbf{e}^{0}_{j}=e^{{2\pi i\over\mathbf{N}}\langle\eta,j\rangle+{\pi i\over\mathbf{N}}\langle y,\eta\rangle}\mathbf{e}^{0}_{j+y}.

Therefore the left-hand side of (2.59) is

Op𝐍,0(a)f,f=j𝐍nfjOp𝐍,0(a)𝐞j0,f=j𝐍ne2πi𝐍η,j+πi𝐍y,ηfjfj+y¯.\langle\operatorname{Op}_{\mathbf{N},0}(a)f,f\rangle_{\mathcal{H}}=\sum_{j\in\mathbb{Z}_{\mathbf{N}}^{n}}f_{j}\langle\operatorname{Op}_{\mathbf{N},0}(a)\mathbf{e}_{j}^{0},f\rangle_{\mathcal{H}}=\sum_{j\in\mathbb{Z}_{\mathbf{N}}^{n}}e^{{2\pi i\over\mathbf{N}}\langle\eta,j\rangle+{\pi i\over\mathbf{N}}\langle y,\eta\rangle}f_{j}\overline{f_{j+y}}.

On the other hand, the right-hand side of (2.59) is equal to

(2𝐍)nj𝐍n,k,2𝐍neπi𝐍(ky,+η,2j+k)fjfj+k¯.(2\mathbf{N})^{-n}\sum_{j\in\mathbb{Z}_{\mathbf{N}}^{n},\ k,\ell\in\mathbb{Z}_{2\mathbf{N}}^{n}}e^{{\pi i\over\mathbf{N}}(\langle k-y,\ell\rangle+\langle\eta,2j+k\rangle)}f_{j}\overline{f_{j+k}}.

The sum over \ell is equal to 0 unless ky2𝐍nk-y\in 2\mathbf{N}\mathbb{Z}^{n} which happens for exactly one value of k2𝐍nk\in\mathbb{Z}_{2\mathbf{N}}^{n}. Using that fj+k=fj+yf_{j+k}=f_{j+y} for this kk we write the right-hand side of (2.59) as

j𝐍neπi𝐍η,2j+yfjfj+y¯.\sum_{j\in\mathbb{Z}_{\mathbf{N}}^{n}}e^{{\pi i\over\mathbf{N}}\langle\eta,2j+y\rangle}f_{j}\overline{f_{j+y}}.

This equals the left-hand side of (2.59) which finishes the proof. ∎

Proposition 2.10 can be interpreted as follows:

Op𝐍,0(a)f,f=(2𝐍)2np,q2𝐍na(p2𝐍,q2𝐍)W(f)pq\langle\operatorname{Op}_{\mathbf{N},0}(a)f,f\rangle_{\mathcal{H}}=(2\mathbf{N})^{-2n}\sum_{p,q\in\mathbb{Z}_{2\mathbf{N}}^{n}}a\Big{(}{p\over 2\mathbf{N}},{q\over 2\mathbf{N}}\Big{)}W(f)_{pq}

where the Wigner matrix W(f)pqW(f)_{pq} of ff is given by

W(f)pq=(2𝐍)nj𝐍neπi𝐍p2j,qfjfpj¯,p,q2𝐍n.W(f)_{pq}=(2\mathbf{N})^{n}\sum_{j\in\mathbb{Z}_{\mathbf{N}}^{n}}e^{{\pi i\over\mathbf{N}}\langle p-2j,q\rangle}f_{j}\overline{f_{p-j}},\quad p,q\in\mathbb{Z}_{2\mathbf{N}}^{n}. (2.60)

We now compute the action of metaplectic transformations M𝐍,0:𝐍(0)𝐍(0)M_{\mathbf{N},0}:\mathcal{H}_{\mathbf{N}}(0)\to\mathcal{H}_{\mathbf{N}}(0) where MAM\in\mathcal{M}_{A} for some ASp(2n,)A\in\operatorname{Sp}(2n,\mathbb{Z}) and we assume that 𝐍\mathbf{N} is even so that (2.57) is satisfied for θ=0\theta=0 and all AA. The general formulas are complicated, so instead we follow the approach of [Kel10, §1.2.1]. As proved for instance in [Koh97] the group Sp(2n,)\operatorname{Sp}(2n,\mathbb{Z}) is generated by matrices of the following block form, where ETE^{-T} denotes the transpose of E1E^{-1}:

SB\displaystyle S_{B} :=(I0BI),B is a symmetric n×n integer matrix;\displaystyle:=\begin{pmatrix}I&0\\ B&I\end{pmatrix},\quad B\text{ is a symmetric $n\times n$ integer matrix}; (2.61)
LE\displaystyle L_{E} :=(E00ET),EGL(n,),|detE|=1;\displaystyle:=\begin{pmatrix}E&0\\ 0&E^{-T}\end{pmatrix},\quad E\in\operatorname{GL}(n,\mathbb{Z}),\quad|\det E|=1;
F\displaystyle F :=(0II0).\displaystyle:=\begin{pmatrix}0&I\\ -I&0\end{pmatrix}.

Since the map MASp(2n,)M\in\mathcal{M}\mapsto A\in\operatorname{Sp}(2n,\mathbb{R}) such that MAM\in\mathcal{M}_{A} is a group homomorphism, it suffices to compute the operators M𝐍,0M_{\mathbf{N},0} for MAM\in\mathcal{M}_{A} where AA is in one of the forms (2.61). This is done in

Proposition 2.11.

Assume that 𝐍\mathbf{N} is even. Then there exist

MSB\displaystyle M\in\mathcal{M}_{S_{B}} such that(M𝐍,0f)j=eπi𝐍Bj,jfj,\displaystyle\quad\text{such that}\quad(M_{\mathbf{N},0}f)_{j}=e^{{\pi i\over\mathbf{N}}\langle Bj,j\rangle}f_{j}, (2.62)
MLE\displaystyle M\in\mathcal{M}_{L_{E}} such that(M𝐍,0f)j=fE1j,\displaystyle\quad\text{such that}\quad(M_{\mathbf{N},0}f)_{j}=f_{E^{-1}j}, (2.63)
MF\displaystyle M\in\mathcal{M}_{F} such that(M𝐍,0f)j=𝐍n2k𝐍ne2πi𝐍j,kfk\displaystyle\quad\text{such that}\quad(M_{\mathbf{N},0}f)_{j}=\mathbf{N}^{-{n\over 2}}\sum_{k\in\mathbb{Z}_{\mathbf{N}}^{n}}e^{-{2\pi i\over\mathbf{N}}\langle j,k\rangle}f_{k} (2.64)

for all f𝐍(0)f\in\mathcal{H}_{\mathbf{N}}(0) and jnj\in\mathbb{Z}^{n} where fj:=f,𝐞j0f_{j}:=\langle f,\mathbf{e}^{0}_{j}\rangle_{\mathcal{H}}.

Remark. The evenness assumption on 𝐍\mathbf{N} is only required in (2.62). The formulas (2.63)–(2.64) are valid for all 𝐍\mathbf{N} and one can check that φLE=φF=0\varphi_{L_{E}}=\varphi_{F}=0 where φA\varphi_{A} is defined in (2.55).

Proof.

This follows from the definition (2.35) of 𝐞j0\mathbf{e}^{0}_{j}, the Poisson summation formula (in case of (2.64)), and the following formulas for metaplectic transformations for which (2.17) can be verified directly using (2.1):

  • we have MSBM\in\mathcal{M}_{S_{B}} where

    Mf(x)=ei2hBx,xf(x);Mf(x)=e^{{i\over 2h}\langle Bx,x\rangle}f(x);
  • we have MLEM\in\mathcal{M}_{L_{E}} where

    Mf(x)=f(E1x);Mf(x)=f(E^{-1}x);
  • we have hF\mathcal{F}_{h}\in\mathcal{M}_{F} where

    hf(x)=(2πh)n2neihx,ηf(η)𝑑η.\mathcal{F}_{h}f(x)=(2\pi h)^{-{n\over 2}}\int_{\mathbb{R}^{n}}e^{-{i\over h}\langle x,\eta\rangle}f(\eta)\,d\eta. (2.65)

3. Proofs of Theorems 24

In this section, we reduce the proofs of Theorems 2 and 3 to a decay statement for long words, Proposition 3.10, that will be proved in §4. In §3.1, we introduce notation that will be used in the proofs of Theorems 2 and 3 and state the main estimates that we will need to write these proofs: Lemmas 3.1, 3.2, and 3.4. In §3.2, we explain how these estimates allow us to prove Theorems 2 and 3. In §3.3, we derive Theorem 4 from Theorem 2. In §3.4, we prove Lemmas 3.1 and 3.2. Finally, in §3.5, we reduce Lemma 3.4 to Proposition 3.10.

Our strategy in this section generally follows [DJ18, Jin20, DJN22]. However, the proofs have to be modified to adapt to the setting of quantum maps used here and to the assumption of geometric control transversally to 𝕋±\mathbb{T}_{\pm}. Another important difference is the choice of propagation times, see §3.1.1.

Throughout this section we fix ASp(2n,)A\in\operatorname{Sp}(2n,\mathbb{Z}) that satisfies the spectral gap condition (1.3) and choose a metaplectic operator MAM\in\mathcal{M}_{A} quantizing AA. We take 𝐍\mathbf{N} large and θ𝕋2n\theta\in\mathbb{T}^{2n} satisfying the quantization condition (2.57) and study the restrictions M𝐍,θM_{\mathbf{N},\theta} of MM to the spaces of quantum states 𝐍(θ)\mathcal{H}_{\mathbf{N}}(\theta), see §2.2.4. Following (2.31) we put h:=(2π𝐍)1h:=(2\pi\mathbf{N})^{-1}.

3.1. Words decomposition

In the proofs of both Theorems 2 and 3, we will consider two 𝐍\mathbf{N}-independent functions b1,b2b_{1},b_{2} on the torus 𝕋2n\mathbb{T}^{2n}. The choice of these functions will differ in the proof of each theorem, but we will always assume that they satisfy

b1,b2C(𝕋2n),|b1|+|b2|1.b_{1},b_{2}\in C^{\infty}(\mathbb{T}^{2n}),\quad|b_{1}|+|b_{2}|\leq 1. (3.1)

The functions b1b_{1} and b2b_{2} are supposed given for now, we will explain in the proofs of Theorems 2 and 3 how they are constructed. We will always explicitly point out when specific properties of b1b_{1} and b2b_{2} are required.

Let us write b:=b1+b2b:=b_{1}+b_{2} and take the quantizations (see §2.2.3)

B:=Op𝐍,θ(b),B1:=Op𝐍,θ(b1),B2:=Op𝐍,θ(b2).B:=\operatorname{Op}_{\mathbf{N},\theta}(b),\quad B_{1}:=\operatorname{Op}_{\mathbf{N},\theta}(b_{1}),\quad B_{2}:=\operatorname{Op}_{\mathbf{N},\theta}(b_{2}).

For any operator LL on 𝐍(θ)\mathcal{H}_{\mathbf{N}}(\theta) and TT\in\mathbb{Z}, we define the conjugated operator

L(T):=M𝐍,θTLM𝐍,θT:𝐍(θ)𝐍(θ).L(T):=M_{\mathbf{N},\theta}^{-T}\,L\,M^{T}_{\mathbf{N},\theta}:\mathcal{H}_{\mathbf{N}}(\theta)\to\mathcal{H}_{\mathbf{N}}(\theta). (3.2)

For mm\in\mathbb{N}, we introduce the set of words

𝒲(m):={1,2}m.\begin{split}\mathcal{W}(m):=\left\{1,2\right\}^{m}.\end{split}

For 𝚠=w0wm1𝒲(m)\mathtt{w}=w_{0}\dots w_{m-1}\in\mathcal{W}(m), define the operator

B𝚠Bwm1(m1)Bw1(1)Bw0(0)\begin{split}B_{\mathtt{w}}\coloneqq B_{w_{m-1}}(m-1)\cdots B_{w_{1}}(1)B_{w_{0}}(0)\end{split} (3.3)

and the corresponding symbol

b𝚠j=0m1bwjAj.\begin{split}b_{\mathtt{w}}\coloneqq\prod_{j=0}^{m-1}b_{w_{j}}\circ A^{j}.\end{split} (3.4)

To a function c:𝒲(m)c:\mathcal{W}(m)\to\mathbb{C}, we associate the operator

Bc𝚠𝒲(m)c(𝚠)B𝚠B_{c}\coloneqq\sum_{\mathtt{w}\in\mathcal{W}(m)}c(\mathtt{w})B_{\mathtt{w}} (3.5)

and accordingly the symbol

bc𝚠𝒲(m)c(𝚠)b𝚠.\begin{split}b_{c}\coloneqq\sum_{\mathtt{w}\in\mathcal{W}(m)}c(\mathtt{w})b_{\mathtt{w}}.\end{split}

If c=1lEc=\operatorname{1\hskip-2.75ptl}_{E} is the characteristic function of a subset E𝒲(m)E\subset\mathcal{W}(m), then we simply write BEB_{E} instead of B1lEB_{\operatorname{1\hskip-2.75ptl}_{E}}. Notice then that we have

B𝒲(m)=B(m1)B(1)B(0)=M𝐍,θ(m1)(BM𝐍,θ)m1B.B_{\mathcal{W}(m)}=B(m-1)\cdots B(1)B(0)=M_{\mathbf{N},\theta}^{-(m-1)}\left(BM_{\mathbf{N},\theta}\right)^{m-1}B. (3.6)

In the proof of Theorem 2, we will have B=IB=I, and thus B𝒲(m)=IB_{\mathcal{W}(m)}=I as well. On the other hand, Theorem 3 will be deduced from estimates on large powers of BM𝐍,θBM_{\mathbf{N},\theta} that will follow from estimates on B𝒲(m)B_{\mathcal{W}(m)} for mm large.

3.1.1. Propagation times

We need now to fix a few parameters that will be used to choose a relevant value of mm. Recall from §1.2 that by the spectral gap assumption (1.3), AA has two particular simple eigenvalues

λ±,λ=λ+1,|λ|<1<|λ+|\lambda_{\pm}\in\mathbb{R},\quad\lambda_{-}=\lambda_{+}^{-1},\quad|\lambda_{-}|<1<|\lambda_{+}| (3.7)

and all other eigenvalues λ\lambda of AA are contained in the open annulus |λ|<|λ|<|λ+||\lambda_{-}|<|\lambda|<|\lambda_{+}|. Fix a constant γ\gamma such that

1<γ<|λ+|,Spec(A){λ+,λ}{λ:γ1<|λ|<γ}.1<\gamma<|\lambda_{+}|,\quad\operatorname{Spec}(A)\setminus\{\lambda_{+},\lambda_{-}\}\subset\{\lambda\colon\gamma^{-1}<|\lambda|<\gamma\}.

Define the hyperplanes

L:=Range(Aλ±I)2n.L_{\mp}:=\operatorname{Range}(A-\lambda_{\pm}I)\ \subset\ \mathbb{R}^{2n}. (3.8)

Note that LL_{\mp} are coisotropic (since they have codimension 1), invariant under AA and, denoting by E±E_{\pm} the eigenspaces of AA corresponding to λ±\lambda_{\pm},

2n=E+L=EL+.\mathbb{R}^{2n}=E_{+}\oplus L_{-}=E_{-}\oplus L_{+}. (3.9)

Moreover, since ASp(2n,)A\in\operatorname{Sp}(2n,\mathbb{R}) we have (where σ\sigma denotes the symplectic form from (2.4))

σ(z,w)=0for allzE±,wL±.\sigma(z,w)=0\quad\text{for all}\quad z\in E_{\pm},\quad w\in L_{\pm}. (3.10)

Next, LL_{\mp}\otimes\mathbb{C} is the sum of all generalized eigenspaces of AA with eigenvalues not equal to λ±\lambda_{\pm}. Therefore we have the spectral radius bounds

A±m\displaystyle\big{\|}A^{\pm m}\big{\|} =𝒪(|λ+|m)asm,\displaystyle=\mathcal{O}(|\lambda_{+}|^{m})\quad\text{as}\quad m\to\infty, (3.11)
A±m|L\displaystyle\big{\|}A^{\pm m}|_{L_{\mp}}\big{\|} =𝒪(γm)asm.\displaystyle=\mathcal{O}(\gamma^{m})\quad\text{as}\quad m\to\infty.

Let us now fix two numbers ρ,ρ(0,1)\rho,\rho^{\prime}\in(0,1) such that

ρ+ρ<1,ρlogγlog|λ+|<ρ<12<ρ.\rho+\rho^{\prime}<1,\quad\rho\frac{\log\gamma}{\log\left|\lambda_{+}\right|}<\rho^{\prime}<{1\over 2}<\rho. (3.12)

We also fix an integer JJ that satisfies

J>1+2log2log|λ+|.J>1+\frac{2\log 2}{\log\left|\lambda_{+}\right|}. (3.13)

We now set

T0:=ρlog𝐍Jlog|λ+| and T1:=JT0ρlog𝐍log|λ+|.T_{0}:=\left\lfloor\frac{\rho\log\mathbf{N}}{J\log\left|\lambda_{+}\right|}\right\rfloor\quad\textup{ and }\quad T_{1}:=JT_{0}\approx{\rho\log\mathbf{N}\over\log|\lambda_{+}|}. (3.14)

We call T0T_{0} the short logarithmic propagation time and T1T_{1} the long logarithmic propagation time.

Remark. Let us give some explanations regarding the choice of ρ,ρ\rho,\rho^{\prime} and the propagation times T0,T1T_{0},T_{1}:

  • We will use the semiclassical symbol classes SL±,ρ,ρ(𝕋2n)S_{L_{\pm},\rho,\rho^{\prime}}(\mathbb{T}^{2n}) introduced in §2.2.3 and for that we need 0ρρ0\leq\rho^{\prime}\leq\rho and ρ+ρ<1\rho+\rho^{\prime}<1. In particular, this is used in Lemma 3.11 and, most crucially, in Lemma 4.5.

  • From (3.12) and (3.14) we see that when 𝐍\mathbf{N} is large,

    |λ+|T1𝐍ρ,γT1𝐍ρ.|\lambda_{+}|^{T_{1}}\sim\mathbf{N}^{\rho},\quad\gamma^{T_{1}}\ll\mathbf{N}^{\rho^{\prime}}. (3.15)

    These inequalities are used to show that the symbols b𝚠b_{\mathtt{w}}, 𝚠𝒲(T1)\mathtt{w}\in\mathcal{W}(T_{1}), lie in the class SL,ρ+ε,ρ+ε(𝕋2n)S_{L_{-},\rho+\varepsilon,\rho^{\prime}+\varepsilon}(\mathbb{T}^{2n}), see Lemma 3.11. They are also used in the proof of the porosity property, Lemma 4.4.

  • The requirement ρ>12\rho>{1\over 2} ensures that the support of b𝚠b_{\mathtt{w}}, 𝚠𝒲(T1)\mathtt{w}\in\mathcal{W}(T_{1}), is porous in the direction of the eigenvector of AA with eigenvalue λ+\lambda_{+} on scales (almost) up to hρh1/2h^{\rho}\ll h^{1/2}, so that the fractal uncertainty principle can be applied – see Proposition 4.2, Lemma 4.4, and the last step of the proof of Lemma 4.6 in §4.3.3.

  • The inequality (3.13) on JJ ensures that the errors coming from the exotic calculus SL,ρJ,ρJS_{L_{-},{\rho\over J},{\rho^{\prime}\over J}} decay faster than the growth of the number of elements in 𝒲(T0)\mathcal{W}(T_{0}). More precisely, it makes the remainders in (3.22) and (3.34) decay as a negative power of 𝐍\mathbf{N}.

  • It is also useful to consider what happens in degenerate cases. Assume first that all the eigenvalues in Spec(A){λ+,λ}\operatorname{Spec}(A)\setminus\{\lambda_{+},\lambda_{-}\} lie on the unit circle (this is true in particular if n=1n=1). Then we could take ρ\rho to be any fixed number in (12,1)({1\over 2},1) and ρ\rho^{\prime} to be any sufficiently small positive number. This is the choice made in [DJ18] (which additionally took ρ\rho close to 1).

  • On the other hand, if γ\gamma is close to |λ+||\lambda_{+}| (i.e. the spectral gap of AA is small) then our conditions force ρ<12<ρ\rho^{\prime}<{1\over 2}<\rho to both be close to 121\over 2.

3.1.2. Partition into controlled/uncontrolled words and main estimates

We now decompose the operator B𝒲(m)B_{\mathcal{W}(m)}, m:=2T1m:=2T_{1}, into the sum of two operators corresponding to the controlled and uncontrolled region (see (3.20) below), and state the main estimates used in the proofs of Theorems 2 and 3.

Let F:𝒲(T0)[0,1]F:\mathcal{W}(T_{0})\to\left[0,1\right] be the function that gives the proportion of the digit 11 in a word, that is for 𝚠=w0wT01\mathtt{w}=w_{0}\dots w_{T_{0}-1} we have

F(𝚠)=#{j{0,,T01}:wj=1}T0.F(\mathtt{w})=\frac{\#\left\{j\in\left\{0,\dots,T_{0}-1\right\}:w_{j}=1\right\}}{T_{0}}. (3.16)

Let then α>0\alpha>0 be very small (small enough so that Lemma 3.4 below holds) and define

𝒵:={𝚠𝒲(T0):F(𝚠)α}.\mathcal{Z}:=\left\{\mathtt{w}\in\mathcal{W}(T_{0}):F(\mathtt{w})\geq\alpha\right\}. (3.17)

We call elements of 𝒵\mathcal{Z} controlled short logarithmic words.

We next use the set 𝒵\mathcal{Z} to split 𝒲(2T1)\mathcal{W}\left(2T_{1}\right) into two subsets: 𝒲(2T1)=𝒳𝒴\mathcal{W}(2T_{1})=\mathcal{X}\sqcup\mathcal{Y}, writing each word in 𝒲(2T1)\mathcal{W}(2T_{1}) as a concatenation of 2J2J words in 𝒲(T0)\mathcal{W}(T_{0}):

𝒳\displaystyle\mathcal{X} :={𝚠=𝚠(1)𝚠(2J):𝚠()𝒲(T0)𝒵 for all =1,,2J};\displaystyle\,:=\left\{\mathtt{w}=\mathtt{w}^{(1)}\dots\mathtt{w}^{(2J)}:\mathtt{w}^{(\ell)}\in\mathcal{W}(T_{0})\setminus\mathcal{Z}\textup{ for all }\ell=1,\dots,2J\right\}; (3.18)
𝒴\displaystyle\mathcal{Y} :={𝚠=𝚠(1)𝚠(2J):𝚠()𝒵 for some {1,,2J}}.\displaystyle\,:=\left\{\mathtt{w}=\mathtt{w}^{(1)}\dots\mathtt{w}^{(2J)}:\mathtt{w}^{(\ell)}\in\mathcal{Z}\textup{ for some }\ell\in\left\{1,\dots,2J\right\}\right\}. (3.19)

We call elements of 𝒳\mathcal{X} uncontrolled long logarithmic words and elements of 𝒴\mathcal{Y} controlled long logarithmic words.

It follows from (3.6) that (with the operators B𝒳,B𝒴B_{\mathcal{X}},B_{\mathcal{Y}} defined using (3.5))

M𝐍,θ(2T11)(BM𝐍,θ)2T11B=B𝒲(2T1)=B𝒳+B𝒴.M_{\mathbf{N},\theta}^{-(2T_{1}-1)}(BM_{\mathbf{N},\theta})^{2T_{1}-1}B=B_{\mathcal{W}(2T_{1})}=B_{\mathcal{X}}+B_{\mathcal{Y}}. (3.20)

In order to get an estimate on M𝐍,θ(2T11)(BM𝐍,θ)2T11BM_{\mathbf{N},\theta}^{-(2T_{1}-1)}(BM_{\mathbf{N},\theta})^{2T_{1}-1}B, we will control B𝒳B_{\mathcal{X}} and B𝒴B_{\mathcal{Y}} separately. Let us start with B𝒴B_{\mathcal{Y}}. It will be dealt with differently in the proofs of Theorem 2 and 3, but the main idea is the same: we make an assumption on the symbol b1b_{1} that translates into control on the operator B1B_{1} that is inherited by B𝒴B_{\mathcal{Y}}. In more practical terms, we will use the following lemma in the proof of Theorem 2.

Lemma 3.1.

Let aC(𝕋2n)a\in C^{\infty}\left(\mathbb{T}^{2n}\right). Assume that b1,b2C(𝕋2n)b_{1},b_{2}\in C^{\infty}(\mathbb{T}^{2n}) satisfy

b1,b20,b1+b2=1,suppb1mAm({a0}).b_{1},b_{2}\geq 0,\quad b_{1}+b_{2}=1,\quad\operatorname{supp}b_{1}\subset\bigcup_{m\in\mathbb{Z}}A^{m}\big{(}\{a\neq 0\}\big{)}. (3.21)

Let ε>0\varepsilon>0. Then there is a constant C>0C>0 such that for all 𝐍\mathbf{N} and u𝐍(θ)u\in\mathcal{H}_{\mathbf{N}}(\theta)

B𝒴uCOp𝐍,θ(a)u+C𝐫M(u)log𝐍+C𝐍12+12J(1+2log2log|λ+|)+εu.\left\|B_{\mathcal{Y}}u\right\|_{\mathcal{H}}\leq C\left\|\operatorname{Op}_{\mathbf{N},\theta}(a)u\right\|_{\mathcal{H}}+C\mathbf{r}_{M}(u)\log\mathbf{N}+C\mathbf{N}^{-\frac{1}{2}+\frac{1}{2J}\big{(}1+\frac{2\log 2}{\log\left|\lambda_{+}\right|}\big{)}+\varepsilon}\left\|u\right\|_{\mathcal{H}}. (3.22)

Here we recall that the quantity 𝐫M(u)\mathbf{r}_{M}(u), defined in (1.5), measures how close uu is to an eigenfunction of M𝐍,θM_{\mathbf{N},\theta}. Note that the condition (3.13) on JJ ensures that the power of 𝐍\mathbf{N} in the last term on the right-hand side of (3.22) is negative for ε\varepsilon small enough.

The proof of Theorem 3 will rely on the following estimate instead of Lemma 3.1.

Lemma 3.2.

Assume that b1,b2b_{1},b_{2} satisfy (3.1) and

|b1|+|b2|<1onsuppb1.|b_{1}|+|b_{2}|<1\quad\text{on}\quad\operatorname{supp}b_{1}. (3.23)

Then there exist C,δ>0C,\delta>0 such that for every 𝐍\mathbf{N} we have

B𝒴C𝐍δ.\left\|B_{\mathcal{Y}}\right\|_{\mathcal{H}\to\mathcal{H}}\leq C\mathbf{N}^{-\delta}. (3.24)

The proofs of Lemmas 3.13.2 can be found in §3.4.

The control on B𝒳B_{\mathcal{X}} will be more subtle to obtain. We will use the following estimate, whose proof ultimately relies on the Fractal Uncertainty Principle. Recall the subtori 𝕋±𝕋2n\mathbb{T}_{\pm}\subset\mathbb{T}^{2n} defined in (1.6), and make the following

Definition 3.3.

Let U𝕋2nU\subset\mathbb{T}^{2n} be a set. We say that UU is safe if and only if for every x𝕋2nx\in\mathbb{T}^{2n}, each of the shifted tori x+𝕋+x+\mathbb{T}_{+}, x+𝕋x+\mathbb{T}_{-} intersects UU.

Notice that being safe is slightly more restrictive than satisfying the geometric control condition with respect to 𝕋+\mathbb{T}_{+} and 𝕋\mathbb{T}_{-} (Definition 1.1) as we do not have here the flexibility to replace a point xx by its image by an iterate of AA.

The control on B𝒳B_{\mathcal{X}} is achieved in

Lemma 3.4.

Assume that b1,b2b_{1},b_{2} satisfy (3.1) and the complements 𝕋2nsuppb1\mathbb{T}^{2n}\setminus\operatorname{supp}b_{1}, 𝕋2nsuppb2\mathbb{T}^{2n}\setminus\operatorname{supp}b_{2} are both safe. Assume also that the constant α\alpha in (3.17) is chosen small enough. Then there are constants C,δ>0C,\delta>0 such that for every 𝐍\mathbf{N} we have

B𝒳C𝐍δ.\left\|B_{\mathcal{X}}\right\|_{\mathcal{H}\to\mathcal{H}}\leq C\mathbf{N}^{-\delta}.

The proof of Lemma 3.4 is reduced to a decay result for long logarithmic words, Proposition 3.10, in §3.5. The proof of the latter result is the point of §4.

3.2. Proofs of Theorems 2 and 3

Let us explain now how Lemmas 3.13.4 allow us to prove Theorems 2 and 3. First of all, we need to construct the functions b1b_{1} and b2b_{2} that appear in the statement of these lemmas. To do so, we will use the following two lemmas:

Lemma 3.5.

Assume that V𝕋2nV\subset\mathbb{T}^{2n} is a safe open set. Then VV contains a safe compact subset KK.

Proof.

We argue by contradiction: we write V=mKmV=\bigcup_{m\in\mathbb{N}}K_{m} where KmK_{m} are compact sets with KmKm+1K_{m}\subset K_{m+1}^{\circ} and assume that none of the KmK_{m}’s are safe. Then for every mm\in\mathbb{N} there exist

xm𝕋2n,σm{+,}such that(xm+𝕋σm)Km=.x_{m}\in\mathbb{T}^{2n},\quad\sigma_{m}\in\left\{+,-\right\}\quad\text{such that}\quad(x_{m}+\mathbb{T}_{\sigma_{m}})\cap K_{m}=\emptyset.

Up to extracting a subsequence, we may assume that (σm)m(\sigma_{m})_{m\in\mathbb{N}} is constant equal to some σ{+,}\sigma\in\left\{+,-\right\} and that (xm)m(x_{m})_{m\in\mathbb{N}} converges to some point x𝕋2nx\in\mathbb{T}^{2n}. Since VV is safe, the set x+𝕋σx+\mathbb{T}_{\sigma} intersects VV. Take y(x+𝕋σ)Vy\in(x+\mathbb{T}_{\sigma})\cap V and put ym:=xmx+yxm+𝕋σy_{m}:=x_{m}-x+y\in x_{m}+\mathbb{T}_{\sigma}. Then ymyVy_{m}\to y\in V, so ymKmy_{m}\in K_{m} for mm large enough. This gives a contradiction with our assumption that xm+𝕋σx_{m}+\mathbb{T}_{\sigma} does not intersect KmK_{m}. ∎

Lemma 3.6.

Let UU be an open subset of 𝕋2n\mathbb{T}^{2n} which is safe in the sense of Definition 3.3. Then there exist a1,a2C(𝕋2n)a_{1},a_{2}\in C^{\infty}(\mathbb{T}^{2n}) such that

a1,a20,a1+a2=1,suppa1Ua_{1},a_{2}\geq 0,\quad a_{1}+a_{2}=1,\quad\operatorname{supp}a_{1}\subset U

and the complements 𝕋2nsuppa1\mathbb{T}^{2n}\setminus\operatorname{supp}a_{1}, 𝕋2nsuppa2\mathbb{T}^{2n}\setminus\operatorname{supp}a_{2} are both safe.

Proof.

1. We show that there exist two compact sets

K1,K2𝕋2nsuch thatK1K2=,K1U,K_{1},K_{2}\subset\mathbb{T}^{2n}\quad\text{such that}\quad K_{1}\cap K_{2}=\emptyset,\quad K_{1}\subset U, (3.25)

and K1,K2K_{1},K_{2} are both safe. To do this, let H2nH\subset\mathbb{R}^{2n} be a hyperplane transverse to each of the tangent spaces V±V_{\pm}\otimes\mathbb{R} of 𝕋±\mathbb{T}_{\pm} where V±2nV_{\pm}\subset\mathbb{Q}^{2n} are the subspaces defined in the paragraph preceding (1.6). Denote by π:2n𝕋2n\pi:\mathbb{R}^{2n}\to\mathbb{T}^{2n} the projection map. Take large R>0R>0, denote by BH(R)B_{H}(R) the closed ball of radius RR in HH, and define

DR:=π(BH(R))𝕋2n.D_{R}:=\pi(B_{H}(R))\ \subset\ \mathbb{T}^{2n}.

Then, we can fix RR large enough so that the set DRD_{R} is safe. Indeed, every element of 𝕋2n\mathbb{T}^{2n} can be written as π(x)\pi(x) for some x[0,1]2n2nx\in[0,1]^{2n}\subset\mathbb{R}^{2n}. Then we can decompose x=x±+x±′′x=x^{\prime}_{\pm}+x^{\prime\prime}_{\pm} where x±Hx^{\prime}_{\pm}\in H, x±′′V±x^{\prime\prime}_{\pm}\in V_{\pm}\otimes\mathbb{R}. Moreover, if RR is large enough then we can choose this decomposition so that x±BH(R)x^{\prime}_{\pm}\in B_{H}(R). Then π(x±)(π(x)+𝕋±)DR\pi(x^{\prime}_{\pm})\in(\pi(x)+\mathbb{T}_{\pm})\cap D_{R}.

Now, the open set UDRU\setminus D_{R} is safe since UU is open and safe and each intersection DR(x+𝕋±)D_{R}\cap(x+\mathbb{T}_{\pm}) has empty interior in x+𝕋±x+\mathbb{T}_{\pm}. Then by Lemma 3.5 there exists a safe compact set K1UDRK_{1}\subset U\setminus D_{R}. The complement 𝕋2nK1\mathbb{T}^{2n}\setminus K_{1} contains DRD_{R} and thus is an open safe set. Using Lemma 3.5 again, let K2K_{2} be a compact safe subset of this complement, then K1,K2K_{1},K_{2} satisfy (3.25).

2. Using a partition of unity subordinate to the cover of 𝕋2n\mathbb{T}^{2n} by the sets UK2U\setminus K_{2}, 𝕋2nK1\mathbb{T}^{2n}\setminus K_{1} we choose a1,a2C(𝕋2n)a_{1},a_{2}\in C^{\infty}(\mathbb{T}^{2n}) such that

a1,a20,a1+a2=1,suppa1UK2,suppa2𝕋2nK1.a_{1},a_{2}\geq 0,\quad a_{1}+a_{2}=1,\quad\operatorname{supp}a_{1}\subset U\setminus K_{2},\quad\operatorname{supp}a_{2}\subset\mathbb{T}^{2n}\setminus K_{1}.

The complements of suppa1\operatorname{supp}a_{1}, suppa2\operatorname{supp}a_{2} contain the sets K2K_{2}, K1K_{1} and are thus both safe, finishing the proof. ∎

We are now ready to prove Theorems 2 and 3.

Proof of Theorem 2.

Let aC(𝕋2n)a\in C^{\infty}\left(\mathbb{T}^{2n}\right) be as in the statement of Theorem 2. Since {a0}\left\{a\neq 0\right\} satisfies the geometric control condition transversally to 𝕋+\mathbb{T}_{+} and 𝕋\mathbb{T}_{-}, the open set

U:=mAm({a0})U:=\bigcup_{m\in\mathbb{Z}}A^{m}\big{(}\{a\neq 0\}\big{)}

is safe. We apply Lemma 3.6 to UU to construct two functions a1,a2a_{1},a_{2} and we set b1:=a1b_{1}:=a_{1} and b2:=a2b_{2}:=a_{2} in §3.1. Notice that we have consequently b=b1+b2=1b=b_{1}+b_{2}=1 and (3.20) becomes

I=B𝒳+B𝒴.I=B_{\mathcal{X}}+B_{\mathcal{Y}}.

Fix α>0\alpha>0 be small enough so that Lemma 3.4 applies, that is there are C,δ>0C,\delta>0 such that for every u𝐍(θ)u\in\mathcal{H}_{\mathbf{N}}(\theta) we have

B𝒳uC𝐍δu.\begin{split}\left\|B_{\mathcal{X}}u\right\|_{\mathcal{H}}\leq C\mathbf{N}^{-\delta}\left\|u\right\|_{\mathcal{H}}.\end{split} (3.26)

Next, applying Lemma 3.1 with sufficiently small ε>0\varepsilon>0, recalling (3.13), and making CC larger and δ>0\delta>0 smaller if needed, we have for every u𝐍(θ)u\in\mathcal{H}_{\mathbf{N}}(\theta)

B𝒴uCOp𝐍,θ(a)u+C𝐫M(u)log𝐍+C𝐍δu.\left\|B_{\mathcal{Y}}u\right\|_{\mathcal{H}}\leq C\left\|\operatorname{Op}_{\mathbf{N},\theta}(a)u\right\|_{\mathcal{H}}+C\mathbf{r}_{M}(u)\log\mathbf{N}+C\mathbf{N}^{-\delta}\left\|u\right\|_{\mathcal{H}}. (3.27)

Putting (3.26) and (3.27) together, we find that, for every u𝐍(θ)u\in\mathcal{H}_{\mathbf{N}}(\theta), we have

uCOp𝐍,θ(a)u+C𝐫M(u)log𝐍+C𝐍δu.\left\|u\right\|_{\mathcal{H}}\leq C\left\|\operatorname{Op}_{\mathbf{N},\theta}(a)u\right\|_{\mathcal{H}}+C\mathbf{r}_{M}(u)\log\mathbf{N}+C\mathbf{N}^{-\delta}\left\|u\right\|_{\mathcal{H}}. (3.28)

Now, for 𝐍\mathbf{N} large enough we can remove the last term on the right-hand side of (3.28), obtaining (1.7) and finishing the proof.∎

Proof of Theorem 3.

We will find 𝐍\mathbf{N}-independent constants C,δ,κ>0C,\delta,\kappa>0 and an integer m00m_{0}\geq 0 such that

(Op𝐍,θ(b)M𝐍,θ)(2m0+1)κlog𝐍C𝐍δ.\left\|\left(\operatorname{Op}_{\mathbf{N},\theta}(b)M_{\mathbf{N},\theta}\right)^{(2m_{0}+1)\lfloor\kappa\log\mathbf{N}\rfloor}\right\|_{\mathcal{H}\to\mathcal{H}}\leq C\mathbf{N}^{-\delta}. (3.29)

It will then follow from (3.29) that the spectral radius of Op𝐍,θ(b)M𝐍,θ\operatorname{Op}_{\mathbf{N},\theta}(b)M_{\mathbf{N},\theta} is bounded above by

C1(2m0+1)κlog𝐍exp(δlog𝐍(2m0+1)κlog𝐍)𝐍+eδ(2m0+1)κ<1C^{\frac{1}{(2m_{0}+1)\lfloor\kappa\log\mathbf{N}\rfloor}}\exp\left(-\frac{\delta\log\mathbf{N}}{(2m_{0}+1)\lfloor\kappa\log\mathbf{N}\rfloor}\right)\quad\xrightarrow{\mathbf{N}\to+\infty}\quad e^{-\frac{\delta}{(2m_{0}+1)\kappa}}<1

which will give the conclusion of the theorem.

1. We first reduce to the situation when the set {|b|<1}\{|b|<1\} is safe (which is a stronger assumption than made in Theorem 3). Consider the open set

U:=mAm({|b|<1})=m01Um0whereUm0:=|m|m0Am({|b|<1}).U:=\bigcup_{m\in\mathbb{Z}}A^{m}\big{(}\{|b|<1\}\big{)}=\bigcup_{m_{0}\geq 1}U_{m_{0}}\quad\text{where}\quad U_{m_{0}}:=\bigcup_{|m|\leq m_{0}}A^{m}\big{(}\{|b|<1\}\big{)}.

By the assumption of the theorem, the set UU is safe. By Lemma 3.5, UU contains a safe compact subset KK. Since each Um0U_{m_{0}} is open, we may fix m0m_{0} such that Um0KU_{m_{0}}\supset K, which implies that Um0U_{m_{0}} is safe.

Using (2.58) and (2.49) (with ρ=ρ=0\rho=\rho^{\prime}=0), we next see that

(Op𝐍,θ(b)M𝐍,θ)2m0+1=M𝐍,θm0Op𝐍,θ(b~)M𝐍,θm0+1+𝒪(𝐍1)𝐍(θ)𝐍(θ),\left(\operatorname{Op}_{\mathbf{N},\theta}(b)M_{\mathbf{N},\theta}\right)^{2m_{0}+1}=M^{m_{0}}_{\mathbf{N},\theta}\operatorname{Op}_{\mathbf{N},\theta}(\tilde{b})M^{m_{0}+1}_{\mathbf{N},\theta}+\mathcal{O}\left(\mathbf{N}^{-1}\right)_{\mathcal{H}_{\mathbf{N}}(\theta)\to\mathcal{H}_{\mathbf{N}}(\theta)},

where

b~:=m=m0m0bAm.\tilde{b}:=\prod_{m=-m_{0}}^{m_{0}}b\circ A^{m}.

Since |b|1|b|\leq 1 everywhere, the set {|b~|<1}=Um0\{|\tilde{b}|<1\}=U_{m_{0}} is safe.

Using (2.51) to bound the operator norm of Op𝐍,θ(b~)\operatorname{Op}_{\mathbf{N},\theta}(\tilde{b}) by 1+𝒪(𝐍12)1+\mathcal{O}(\mathbf{N}^{-{1\over 2}}), we have for any fixed κ\kappa

(Op𝐍,θ(b)M𝐍,θ)(2m0+1)κlog𝐍=M𝐍,θm0(Op𝐍,θ(b~)M𝐍,θ2m0+1)κlog𝐍M𝐍,θm0+𝒪(𝐍1+)𝐍(θ)𝐍(θ)).\begin{split}&\left(\operatorname{Op}_{\mathbf{N},\theta}(b)M_{\mathbf{N},\theta}\right)^{(2m_{0}+1)\lfloor\kappa\log\mathbf{N}\rfloor}\\ &\qquad\qquad=M_{\mathbf{N},\theta}^{m_{0}}\left(\operatorname{Op}_{\mathbf{N},\theta}(\tilde{b})M_{\mathbf{N},\theta}^{2m_{0}+1}\right)^{\lfloor\kappa\log\mathbf{N}\rfloor}M_{\mathbf{N},\theta}^{-m_{0}}+\mathcal{O}(\mathbf{N}^{-1+})_{\mathcal{H}_{\mathbf{N}}(\theta)\to\mathcal{H}_{\mathbf{N}}(\theta))}.\end{split}

Hence to show (3.29) it suffices to prove

(Op𝐍,θ(b~)M𝐍,θ2m0+1)κlog𝐍C𝐍δ.\Big{\|}\big{(}\operatorname{Op}_{\mathbf{N},\theta}(\tilde{b})M_{\mathbf{N},\theta}^{2m_{0}+1}\big{)}^{\lfloor\kappa\log\mathbf{N}\rfloor}\Big{\|}_{\mathcal{H}\to\mathcal{H}}\leq C\mathbf{N}^{-\delta}. (3.30)

The operator M2m0+1M^{2m_{0}+1} lies in A2m0+1\mathcal{M}_{A^{2m_{0}+1}}, with the matrix A2m0+1A^{2m_{0}+1} still satisfying the spectral gap assumption (1.3) and producing the same tori 𝕋±\mathbb{T}_{\pm}. Therefore, to show (3.30) it suffices to prove the bound

(Op𝐍,θ(b)M𝐍,θ)κlog𝐍C𝐍δ\Big{\|}\big{(}\operatorname{Op}_{\mathbf{N},\theta}(b)M_{\mathbf{N},\theta}\big{)}^{\lfloor\kappa\log\mathbf{N}\rfloor}\Big{\|}_{\mathcal{H}\to\mathcal{H}}\leq C\mathbf{N}^{-\delta} (3.31)

for any AA satisfying (1.3) and any MAM\in\mathcal{M}_{A}, where we assume that bC(𝕋2n)b\in C^{\infty}(\mathbb{T}^{2n}), |b|1|b|\leq 1 everywhere, and the set {|b|<1}\{|b|<1\} is safe.

2. We now show (3.31). Using Lemma 3.6 for the set {|b|<1}\{|b|<1\}, we construct two cutoff functions a1,a2C(𝕋2n)a_{1},a_{2}\in C^{\infty}(\mathbb{T}^{2n}). Put

b1=a1b,b2=a2b.b_{1}=a_{1}b,\quad b_{2}=a_{2}b.

Note that b=b1+b2b=b_{1}+b_{2} in agreement with the convention of §3.1 and B=Op𝐍,θ(b)B=\operatorname{Op}_{\mathbf{N},\theta}(b). Moreover, |b1|+|b2|=|b|1|b_{1}|+|b_{2}|=|b|\leq 1. By construction, b1b_{1} and b2b_{2} satisfy the hypotheses of Lemma 3.4. Consequently, we can choose α\alpha small enough so that this lemma applies: there are C,δ>0C,\delta>0 such that

B𝒳C𝐍δ.\left\|B_{\mathcal{X}}\right\|_{\mathcal{H}\to\mathcal{H}}\leq C\mathbf{N}^{-\delta}.

As a1a_{1} is supported in {|b|<1}\left\{\left|b\right|<1\right\}, we see that |b1|+|b2|<1\left|b_{1}\right|+\left|b_{2}\right|<1 on suppb1\operatorname{supp}b_{1}, and we can consequently apply Lemma 3.2 to see that, up to making CC larger and δ\delta smaller, we also have

B𝒴C𝐍δ.\left\|B_{\mathcal{Y}}\right\|_{\mathcal{H}\to\mathcal{H}}\leq C\mathbf{N}^{-\delta}.

Recalling (3.20), we see that

(Op𝐍,θ(b)M𝐍,θ)2T1=M𝐍,θ2T11(B𝒳+B𝒴)M𝐍,θB𝒳+B𝒴2C𝐍δ.\begin{split}\left\|\left(\operatorname{Op}_{\mathbf{N},\theta}(b)M_{\mathbf{N},\theta}\right)^{2T_{1}}\right\|_{\mathcal{H}\to\mathcal{H}}&=\left\|M_{\mathbf{N},\theta}^{2T_{1}-1}\left(B_{\mathcal{X}}+B_{\mathcal{Y}}\right)M_{\mathbf{N},\theta}\right\|_{\mathcal{H}\to\mathcal{H}}\\ &\leq\left\|B_{\mathcal{X}}\right\|_{\mathcal{H}\to\mathcal{H}}+\left\|B_{\mathcal{Y}}\right\|_{\mathcal{H}\to\mathcal{H}}\leq 2C\mathbf{N}^{-\delta}.\end{split}

Since T1T_{1} is defined by (3.14), we just established (3.31), which ends the proof of the theorem. ∎

3.3. Proof of Theorem 4

We argue by contradiction. Let μ\mu be a semiclassical measure associated to AA and assume that suppμ\operatorname{supp}\mu does not contain any sets of the form x+𝕋+x+\mathbb{T}_{+} or x+𝕋x+\mathbb{T}_{-}. Then the complement 𝕋2nsuppμ\mathbb{T}^{2n}\setminus\operatorname{supp}\mu is an open safe set in the sense of Definition 3.3. By Lemma 3.5, there exists a compact safe set KK such that Ksuppμ=K\cap\operatorname{supp}\mu=\emptyset. Take aC(𝕋2n)a\in C^{\infty}(\mathbb{T}^{2n}) such that suppasuppμ=\operatorname{supp}a\cap\operatorname{supp}\mu=\emptyset and a=1a=1 on KK. Then the set {a0}\{a\neq 0\} is safe, so aa satisfies the hypothesis of Theorem 2.

Let uj𝐍j(θj)u_{j}\in\mathcal{H}_{\mathbf{N}_{j}}(\theta_{j}) be a sequence converging to μ\mu in the sense of (1.8). Since uju_{j} is an eigenfunction of M𝐍j,θjM_{\mathbf{N}_{j},\theta_{j}} and 𝐍j\mathbf{N}_{j}\to\infty, Theorem 2 shows that there exists a constant CaC_{a} such that for all large enough jj

1=ujCaOp𝐍j,θj(a)uj.1=\|u_{j}\|_{\mathcal{H}}\leq C_{a}\|\operatorname{Op}_{\mathbf{N}_{j},\theta_{j}}(a)u_{j}\|_{\mathcal{H}}. (3.32)

Now, by (2.47) and (2.49) (with ρ=ρ=0\rho=\rho^{\prime}=0) we have

Op𝐍j,θj(a)uj2\displaystyle\|\operatorname{Op}_{\mathbf{N}_{j},\theta_{j}}(a)u_{j}\|_{\mathcal{H}}^{2} =Op𝐍j,θj(a)Op𝐍j,θj(a)uj,uj\displaystyle=\langle\operatorname{Op}_{\mathbf{N}_{j},\theta_{j}}(a)^{*}\operatorname{Op}_{\mathbf{N}_{j},\theta_{j}}(a)u_{j},u_{j}\rangle_{\mathcal{H}} (3.33)
=Op𝐍j,θj(|a|2)uj,ujH+𝒪(𝐍j1).\displaystyle=\langle\operatorname{Op}_{\mathbf{N}_{j},\theta_{j}}(|a|^{2})u_{j},u_{j}\rangle_{H}+\mathcal{O}(\mathbf{N}_{j}^{-1}).

By (3.32), the left-hand side of (3.33) is bounded away from 0 as jj\to\infty. By (1.8), the right-hand side of (3.33) converges to 𝕋2n|a|2𝑑μ\int_{\mathbb{T}^{2n}}|a|^{2}\,d\mu; thus this integral is positive. This contradicts the fact that suppasuppμ=\operatorname{supp}a\cap\operatorname{supp}\mu=\emptyset and finishes the proof.

3.4. Estimates in the controlled region

Here we prove Lemmas 3.1 and 3.2, using the notation that we introduced in §3.1. We start by relating the operator B𝚠B_{\mathtt{w}} from (3.3) and the symbol b𝚠b_{\mathtt{w}} from (3.4). Recall the symbol classes SL,ρ,ρ(𝕋2n)S_{L,\rho,\rho^{\prime}}(\mathbb{T}^{2n}) from §2.2.3, the constants ρ,ρ\rho,\rho^{\prime} defined in (3.12), the integers J,T0J,T_{0} defined in (3.13) and (3.14), and the hyperplane L2nL_{-}\subset\mathbb{R}^{2n} defined in (3.8).

Lemma 3.7.

Let ε>0\varepsilon>0. For every 𝚠𝒲(T0)\mathtt{w}\in\mathcal{W}(T_{0}), the symbol b𝚠b_{\mathtt{w}} belongs to the symbol class SL,ρJ+ε,ρJ+ε(𝕋2n)S_{L_{-},\frac{\rho}{J}+\varepsilon,\frac{\rho^{\prime}}{J}+\varepsilon}(\mathbb{T}^{2n}) and

B𝚠=Op𝐍,θ(b𝚠)+𝒪(𝐍ρ+ρJ+ε1)𝐍(θ)𝐍(θ).B_{\mathtt{w}}=\operatorname{Op}_{\mathbf{N},\theta}(b_{\mathtt{w}})+\mathcal{O}\left(\mathbf{N}^{\frac{\rho+\rho^{\prime}}{J}+\varepsilon-1}\right)_{\mathcal{H}_{\mathbf{N}}(\theta)\to\mathcal{H}_{\mathbf{N}}(\theta)}.

Here the semi-norms of b𝚠b_{\mathtt{w}} and the constant in the 𝒪()\mathcal{O}(\bullet) are bounded uniformly in 𝚠\mathtt{w}.

Proof.

1. We first show that each SL,ρJ,ρJS_{L_{-},{\rho\over J},{\rho^{\prime}\over J}}-seminorm of the symbols

b1Aj and b2Aj,0jT0b_{1}\circ A^{j}\quad\textup{ and }\quad b_{2}\circ A^{j},\quad 0\leq j\leq T_{0}

is bounded uniformly in jj and 𝐍\mathbf{N}. To simplify notation we give the proof for b1b_{1}, which applies to b2b_{2} as well (in fact, it applies with b1b_{1} replaced by any fixed function in C(𝕋2n)C^{\infty}(\mathbb{T}^{2n})).

Let X1,,Xk,Y1,,YmX_{1},\dots,X_{k},Y_{1},\dots,Y_{m} be constant vector fields on 2n\mathbb{R}^{2n} such that Y1,,YmY_{1},\dots,Y_{m} are tangent to LL_{-}. Since AA is a linear map, we compute for x𝕋2nx\in\mathbb{T}^{2n}

X1XkY1Ym(b1Aj)(x)=Dk+mb1(Ajx)(AjX1,,AjXk,AjY1,,AjYm).X_{1}\dots X_{k}Y_{1}\dots Y_{m}(b_{1}\circ A^{j})(x)=D^{k+m}b_{1}(A^{j}x)\cdot\left(A^{j}X_{1},\dots,A^{j}X_{k},A^{j}Y_{1},\dots,A^{j}Y_{m}\right).

Here, Dk+mb1D^{k+m}b_{1} denotes the k+mk+m-th derivative of b1b_{1}, which is a k+mk+m-linear form, uniformly bounded since it does not depend on 𝐍\mathbf{N}. Consequently, for some 𝐍\mathbf{N}-independent constant C>0C>0, we have

sup|X1XkY1Ym(b1Aj)|C|AjX1||AjXk||AjY1||AjYm|.\sup\left|X_{1}\dots X_{k}Y_{1}\dots Y_{m}(b_{1}\circ A^{j})\right|\leq C\left|A^{j}X_{1}\right|\cdots\left|A^{j}X_{k}\right|\left|A^{j}Y_{1}\right|\cdots\left|A^{j}Y_{m}\right|.

Using the norm bounds (3.11) on Aj\|A^{j}\| and Aj|L\|A^{j}|_{L_{-}}\| we get (for a different choice of CC)

sup|X1XkY1Ym(b1Aj)|C|λ+|jkγjmC|λ+|kT0γmT0C𝐍ρJk+ρJm,\begin{split}\sup\left|X_{1}\dots X_{k}Y_{1}\dots Y_{m}(b_{1}\circ A^{j})\right|&\leq C\left|\lambda_{+}\right|^{jk}\gamma^{jm}\leq C\left|\lambda_{+}\right|^{kT_{0}}\gamma^{mT_{0}}\\ &\leq C\mathbf{N}^{\frac{\rho}{J}k+\frac{\rho^{\prime}}{J}m},\end{split}

which is precisely the required estimate. Here we used the definition (3.14) of T0T_{0} and the condition (3.12) that we imposed on ρ\rho^{\prime}.

2. Let 𝚠𝒲(T0)\mathtt{w}\in\mathcal{W}(T_{0}). By the exact Egorov theorem (2.58), and recalling (3.2), we have

Bwj(j)=Op𝐍,θ(bwjAj)for allj=0,,T01.B_{w_{j}}(j)=\operatorname{Op}_{\mathbf{N},\theta}(b_{w_{j}}\circ A^{j})\quad\text{for all}\quad j=0,\dots,T_{0}-1.

Now it remains to use Lemma 2.8, the definitions (3.3) and (3.4) of B𝚠B_{\mathtt{w}} and b𝚠b_{\mathtt{w}}, and the uniform bound etablished in Step 1 of this proof. ∎

With Lemma 3.7 at our disposal, we can produce the proof of Lemma 3.2.

Proof of Lemma 3.2.

1. Let us prove first the required estimate for the operator B𝒵B_{\mathcal{Z}} (instead of B𝒴B_{\mathcal{Y}}) associated to the set 𝒵\mathcal{Z} defined by (3.17). Put

η:=maxsuppb1(|b1|+|b2|)<1\eta:=\max_{\operatorname{supp}b_{1}}\big{(}|b_{1}|+|b_{2}|\big{)}<1

then |b1|η(1|b2|)\left|b_{1}\right|\leq\eta\left(1-\left|b_{2}\right|\right) everywhere. For 𝚠𝒲(T0)\mathtt{w}\in\mathcal{W}(T_{0}), denote

b~𝚠:=j=0T01b~wjAjwhereb~1:=1|b2|,b~2:=|b2|.\tilde{b}_{\mathtt{w}}:=\prod_{j=0}^{T_{0}-1}\tilde{b}_{w_{j}}\circ A^{j}\quad\text{where}\quad\tilde{b}_{1}:=1-|b_{2}|,\quad\tilde{b}_{2}:=|b_{2}|.

Recalling the function FF from (3.16) and the definition (3.4) of b𝚠b_{\mathtt{w}}, we see that

|b𝚠|ηαT0b~𝚠for all𝚠𝒵.\left|b_{\mathtt{w}}\right|\leq\eta^{\alpha T_{0}}\tilde{b}_{\mathtt{w}}\quad\text{for all}\quad\mathtt{w}\in\mathcal{Z}.

Summing over 𝚠𝒵\mathtt{w}\in\mathcal{Z}, we find that

|b𝒵|ηαT0𝚠𝒵b~𝚠ηαT0𝚠𝒲(T0)b~𝚠=ηαT0.\begin{split}\left|b_{\mathcal{Z}}\right|&\leq\eta^{\alpha T_{0}}\sum_{\mathtt{w}\in\mathcal{Z}}\tilde{b}_{\mathtt{w}}\leq\eta^{\alpha T_{0}}\sum_{\mathtt{w}\in\mathcal{W}(T_{0})}\tilde{b}_{\mathtt{w}}=\eta^{\alpha T_{0}}.\end{split}

Let ε>0\varepsilon>0 be very small. By (3.14), there are at most 2T0𝐍ρlog2Jlog|λ+|2^{T_{0}}\leq\mathbf{N}^{\frac{\rho\log 2}{J\log\left|\lambda_{+}\right|}} elements in 𝒵\mathcal{Z}. It follows from Lemma 3.7 that 𝐍ρlog2Jlog|λ+|b𝒵\mathbf{N}^{-\frac{\rho\log 2}{J\log\left|\lambda_{+}\right|}}b_{\mathcal{Z}} is bounded in SL,ρJ+ε,ρJ+εS_{L_{-},\frac{\rho}{J}+\varepsilon,\frac{\rho^{\prime}}{J}+\varepsilon} uniformly in 𝐍\mathbf{N} and that

𝐍ρlog2Jlog|λ+|B𝒵=Op𝐍,θ(𝐍ρlog2Jlog|λ+|b𝒵)+𝒪(𝐍ρ+ρJ+ε1)𝐍(θ)𝐍(θ).\mathbf{N}^{-\frac{\rho\log 2}{J\log\left|\lambda_{+}\right|}}B_{\mathcal{Z}}=\operatorname{Op}_{\mathbf{N},\theta}\left(\mathbf{N}^{-\frac{\rho\log 2}{J\log\left|\lambda_{+}\right|}}b_{\mathcal{Z}}\right)+\mathcal{O}\left(\mathbf{N}^{\frac{\rho+\rho^{\prime}}{J}+\varepsilon-1}\right)_{\mathcal{H}_{\mathbf{N}}(\theta)\to\mathcal{H}_{\mathbf{N}}(\theta)}.

We can consequently apply (2.51) to the symbol 𝐍ρlog2Jlog|λ+|b𝒵\mathbf{N}^{-\frac{\rho\log 2}{J\log\left|\lambda_{+}\right|}}b_{\mathcal{Z}} in order to find that, for some C>0C>0 that may vary from one line to another,

B𝒵𝐍ρlog2Jlog|λ+|Op𝐍,θ(𝐍ρlog2Jlog|λ+|b𝒵)+C𝐍ρlog2Jlog|λ+|+ρ+ρJ+ε1ηαT0+C𝐍ρlog2Jlog|λ+|+ρ+ρ2J+ε12C𝐍δ,\begin{split}\left\|B_{\mathcal{Z}}\right\|_{\mathcal{H}\to\mathcal{H}}&\leq\mathbf{N}^{\frac{\rho\log 2}{J\log\left|\lambda_{+}\right|}}\left\|\operatorname{Op}_{\mathbf{N},\theta}\left(\mathbf{N}^{-\frac{\rho\log 2}{J\log\left|\lambda_{+}\right|}}b_{\mathcal{Z}}\right)\right\|_{\mathcal{H}\to\mathcal{H}}+C\mathbf{N}^{\frac{\rho\log 2}{J\log\left|\lambda_{+}\right|}+\frac{\rho+\rho^{\prime}}{J}+\varepsilon-1}\\ &\leq\eta^{\alpha T_{0}}+C\mathbf{N}^{\frac{\rho\log 2}{J\log\left|\lambda_{+}\right|}+\frac{\rho+\rho^{\prime}}{2J}+\varepsilon-{1\over 2}}\\ &\leq C\mathbf{N}^{-\delta},\end{split} (3.34)

with

δ:=min(αρlogηJlog|λ+|,12ρlog2Jlog|λ+|ρ+ρ2Jε).\delta:=\min\left(-\frac{\alpha\rho\log\eta}{J\log\left|\lambda_{+}\right|},{1\over 2}-\frac{\rho\log 2}{J\log\left|\lambda_{+}\right|}-\frac{\rho+\rho^{\prime}}{2J}-\varepsilon\right).

Notice that the condition (3.13) that we imposed on JJ ensures that δ\delta is positive (provided ε\varepsilon is small enough).

2. The same proof with η\eta replaced by 11 gives

B𝒲(T0)1+C𝐍δ,B𝒲(T0)𝒵1+C𝐍δ.\left\|B_{\mathcal{W}(T_{0})}\right\|_{\mathcal{H}\to\mathcal{H}}\leq 1+C\mathbf{N}^{-\delta},\quad\left\|B_{\mathcal{W}(T_{0})\setminus\mathcal{Z}}\right\|_{\mathcal{H}\to\mathcal{H}}\leq 1+C\mathbf{N}^{-\delta}. (3.35)

3. Let us now use the estimates (3.34) and (3.35) to prove the lemma. If 𝚠=𝚠(1)𝚠(2J)𝒲(2T1)\mathtt{w}=\mathtt{w}^{(1)}\dots\mathtt{w}^{(2J)}\in\mathcal{W}(2T_{1}) is the concatenation of the words 𝚠(1),,𝚠(2J)𝒲(T0)\mathtt{w}^{(1)},\dots,\mathtt{w}^{(2J)}\in\mathcal{W}(T_{0}), then by (3.3) we have

B𝚠=B𝚠(2J)((2J1)T0)B𝚠(2)(T0)B𝚠(1)(0).B_{\mathtt{w}}=B_{\mathtt{w}^{(2J)}}((2J-1)T_{0})\cdots B_{\mathtt{w}^{(2)}}(T_{0})B_{\mathtt{w}^{(1)}}(0).

Using the definition (3.19) of 𝒴\mathcal{Y}, and splitting this set into the disjoint union of 2J2J subsets corresponding to the largest \ell such that 𝚠()𝒵\mathtt{w}^{(\ell)}\in\mathcal{Z}, we write

B𝒴==12JM𝐍,θ(2J1)T0(B𝒲(T0)𝒵M𝐍,θT0)2JB𝒵(M𝐍,θT0B𝒲(T0))1.\begin{split}B_{\mathcal{Y}}=\sum_{\ell=1}^{2J}M_{\mathbf{N},\theta}^{-(2J-1)T_{0}}\left(B_{\mathcal{W}(T_{0})\setminus\mathcal{Z}}M_{\mathbf{N},\theta}^{T_{0}}\right)^{2J-\ell}B_{\mathcal{Z}}\left(M_{\mathbf{N},\theta}^{T_{0}}B_{\mathcal{W}(T_{0})}\right)^{\ell-1}.\end{split} (3.36)

According to (3.34) and (3.35), the operator norm of each term in this sum is less than

C(1+C𝐍δ)2J1𝐍δ.C\left(1+C\mathbf{N}^{-\delta}\right)^{2J-1}\mathbf{N}^{-\delta}.

Since the number of terms in this sum is 2J2J, that does not depend on 𝐍\mathbf{N}, the estimate (3.24) follows. ∎

In order to prove Lemma 3.1, we need a few more preliminary results. We start with a norm estimate on the operators BcB_{c} defined in (3.5).

Lemma 3.8.

Assume that 0b1,b210\leq b_{1},b_{2}\leq 1. Let ε>0\varepsilon>0. Let

c,d:𝒲(T0),|c(𝚠)|d(𝚠)1for all𝚠𝒲(T0).c,d:\mathcal{W}(T_{0})\to\mathbb{C},\quad|c(\mathtt{w})|\leq d(\mathtt{w})\leq 1\quad\text{for all}\quad\mathtt{w}\in\mathcal{W}(T_{0}).

Then there is a constant C>0C>0, that does not depend on cc nor dd such that, for every u𝐍(θ)u\in\mathcal{H}_{\mathbf{N}}(\theta), we have

BcuBdu+C𝐍12+12J(1+2log2log|λ+|)+εu\left\|B_{c}u\right\|_{\mathcal{H}}\leq\left\|B_{d}u\right\|_{\mathcal{H}}+C\mathbf{N}^{-\frac{1}{2}+\frac{1}{2J}\left(1+\frac{2\log 2}{\log\left|\lambda_{+}\right|}\right)+\varepsilon}\left\|u\right\|_{\mathcal{H}}
Proof.

Since the number of elements in 𝒲(T0)\mathcal{W}(T_{0}) is 2T0𝐍ρlog2Jlog|λ+|2^{T_{0}}\leq\mathbf{N}^{\frac{\rho\log 2}{J\log\left|\lambda_{+}\right|}}, it follows from Lemma 3.7 that 𝐍ρlog2Jlog|λ+|bc\mathbf{N}^{-\frac{\rho\log 2}{J\log\left|\lambda_{+}\right|}}b_{c} and 𝐍ρlog2Jlog|λ+|bd\mathbf{N}^{-\frac{\rho\log 2}{J\log\left|\lambda_{+}\right|}}b_{d} are bounded in the symbol class SL,ρJ+ε,ρJ+εS_{L_{-},\frac{\rho}{J}+\varepsilon,\frac{\rho^{\prime}}{J}+\varepsilon} uniformly in 𝐍,c,d\mathbf{N},c,d and that

𝐍ρlog2Jlog|λ+|Bc=Op𝐍,θ(𝐍ρlog2Jlog|λ+|bc)+𝒪(𝐍ρ+ρJ+ε1)𝐍(θ)𝐍(θ).\mathbf{N}^{-\frac{\rho\log 2}{J\log\left|\lambda_{+}\right|}}B_{c}=\operatorname{Op}_{\mathbf{N},\theta}\left(\mathbf{N}^{-\frac{\rho\log 2}{J\log\left|\lambda_{+}\right|}}b_{c}\right)+\mathcal{O}\left(\mathbf{N}^{\frac{\rho+\rho^{\prime}}{J}+\varepsilon-1}\right)_{\mathcal{H}_{\mathbf{N}}(\theta)\to\mathcal{H}_{\mathbf{N}}(\theta)}.

The same estimate is satisfied by BdB_{d} and bdb_{d}. By our assumption, |bc|bd|b_{c}|\leq b_{d} everywhere. Thus by Lemma 2.7

BcuBdu+C𝐍12+ρ+ρ2J+ρlog2Jlog|λ+|+εu.\left\|B_{c}u\right\|_{\mathcal{H}}\leq\left\|B_{d}u\right\|_{\mathcal{H}}+C\mathbf{N}^{-{1\over 2}+\frac{\rho+\rho^{\prime}}{2J}+\frac{\rho\log 2}{J\log\left|\lambda_{+}\right|}+\varepsilon}\left\|u\right\|_{\mathcal{H}}.

The result then follows by using that ρρ+ρ1\rho\leq\rho+\rho^{\prime}\leq 1. ∎

We will also use a standard elliptic estimate.

Lemma 3.9.

Let aC(𝕋2n)a\in C^{\infty}(\mathbb{T}^{2n}) and assume that

suppb1A({a0}).\operatorname{supp}b_{1}\ \subset\ \bigcup_{\ell\in\mathbb{Z}}A^{\ell}(\left\{a\neq 0\right\}).

Then there is a constant C>0C>0 such that, for every mm\in\mathbb{Z} and every u𝐍(θ)u\in\mathcal{H}_{\mathbf{N}}(\theta), we have (with 𝐫M(u)\mathbf{r}_{M}(u) defined in (1.5))

B1(m)uCOp𝐍,θ(a)u+C(1+|m|)𝐫M(u)+C𝐍1u.\left\|B_{1}(m)u\right\|_{\mathcal{H}}\leq C\left\|\operatorname{Op}_{\mathbf{N},\theta}(a)u\right\|_{\mathcal{H}}+C(1+|m|)\mathbf{r}_{M}(u)+C\mathbf{N}^{-1}\left\|u\right\|_{\mathcal{H}}. (3.37)
Proof.

1. First of all, we may assume that a0a\geq 0 everywhere, since we may replace aa with |a|2|a|^{2} and use the following corollary of (2.49) (with ρ=ρ=0\rho=\rho^{\prime}=0):

Op𝐍,θ(|a|2)u\displaystyle\|\operatorname{Op}_{\mathbf{N},\theta}(|a|^{2})u\|_{\mathcal{H}} Op𝐍,θ(a¯)Op𝐍,θ(a)u+C𝐍1u\displaystyle\leq\|\operatorname{Op}_{\mathbf{N},\theta}(\bar{a})\operatorname{Op}_{\mathbf{N},\theta}(a)u\|_{\mathcal{H}}+C\mathbf{N}^{-1}\|u\|_{\mathcal{H}}
COp𝐍,θ(a)u+C𝐍1u.\displaystyle\leq C\|\operatorname{Op}_{\mathbf{N},\theta}(a)u\|_{\mathcal{H}}+C\mathbf{N}^{-1}\|u\|_{\mathcal{H}}.

Here CC denotes an 𝐍\mathbf{N}-independent constant whose precise value may change from line to line.

2. By a compactness argument, we see that there exists 0\ell_{0}\in\mathbb{N} such that

suppb1=00A({a>0})={a~>0}wherea~:==00aA.\operatorname{supp}b_{1}\ \subset\ \bigcup_{\ell=-\ell_{0}}^{\ell_{0}}A^{\ell}(\left\{a>0\right\})=\{\tilde{a}>0\}\quad\text{where}\quad\tilde{a}:=\sum_{\ell=-\ell_{0}}^{\ell_{0}}a\circ A^{\ell}.

We write b1=qa~b_{1}=q\tilde{a} for some qC(𝕋2n)q\in C^{\infty}(\mathbb{T}^{2n}). Then by (2.49) and (2.58)

Op𝐍,θ(b1)u\displaystyle\|\operatorname{Op}_{\mathbf{N},\theta}(b_{1})u\|_{\mathcal{H}} Op𝐍,θ(q)Op𝐍,θ(a~)u+C𝐍1u\displaystyle\leq\|\operatorname{Op}_{\mathbf{N},\theta}(q)\operatorname{Op}_{\mathbf{N},\theta}(\tilde{a})u\|_{\mathcal{H}}+C\mathbf{N}^{-1}\|u\|_{\mathcal{H}}
COp𝐍,θ(a~)u+C𝐍1u\displaystyle\leq C\|\operatorname{Op}_{\mathbf{N},\theta}(\tilde{a})u\|_{\mathcal{H}}+C\mathbf{N}^{-1}\|u\|_{\mathcal{H}}
=C=00M𝐍,θOp𝐍,θ(a)M𝐍,θu+C𝐍1u.\displaystyle=C\sum_{\ell=-\ell_{0}}^{\ell_{0}}\|M_{\mathbf{N},\theta}^{-\ell}\operatorname{Op}_{\mathbf{N},\theta}(a)M_{\mathbf{N},\theta}^{\ell}u\|_{\mathcal{H}}+C\mathbf{N}^{-1}\|u\|_{\mathcal{H}}.

Applying this with uu replaced by M𝐍,θmuM_{\mathbf{N},\theta}^{m}u, we see that for all mm, with the constant CC independent of mm,

B1(m)uC=00Op𝐍,θ(a)M𝐍,θm+u+C𝐍1u.\|B_{1}(m)u\|_{\mathcal{H}}\leq C\sum_{\ell=-\ell_{0}}^{\ell_{0}}\|\operatorname{Op}_{\mathbf{N},\theta}(a)M_{\mathbf{N},\theta}^{m+\ell}u\|_{\mathcal{H}}+C\mathbf{N}^{-1}\|u\|_{\mathcal{H}}. (3.38)

3. We have for every operator A:𝐍(θ)𝐍(θ)A:\mathcal{H}_{\mathbf{N}}(\theta)\to\mathcal{H}_{\mathbf{N}}(\theta), mm\in\mathbb{Z}, and u𝐍(θ)u\in\mathcal{H}_{\mathbf{N}}(\theta)

AM𝐍,θmuAu+A|m|𝐫M(u).\|AM_{\mathbf{N},\theta}^{m}u\|_{\mathcal{H}}\leq\|Au\|_{\mathcal{H}}+\|A\|_{\mathcal{H}\to\mathcal{H}}|m|\mathbf{r}_{M}(u). (3.39)

Indeed, take z𝕊1z\in\mathbb{S}^{1} such that 𝐫M(u)=M𝐍,θuzu\mathbf{r}_{M}(u)=\|M_{\mathbf{N},\theta}u-zu\|_{\mathcal{H}}. Then we have when m0m\geq 0,

M𝐍,θmuzmu=1mzmM𝐍,θ1(M𝐍,θz)u|m|𝐫M(u).\begin{split}\left\|M_{\mathbf{N},\theta}^{m}u-z^{m}u\right\|_{\mathcal{H}}&\leq\sum_{\ell=1}^{m}\left\|z^{m-\ell}M_{\mathbf{N},\theta}^{\ell-1}\left(M_{\mathbf{N},\theta}-z\right)u\right\|_{\mathcal{H}}\\ &\leq|m|\mathbf{r}_{M}(u).\end{split}

By a similar computation, we see that this estimate still holds when m<0m<0. These estimates imply (3.39).

4. Finally, putting together (3.38) and (3.39) (with A:=Op𝐍,θ(a)A:=\operatorname{Op}_{\mathbf{N},\theta}(a)) we get (3.37). ∎

We are now ready to prove Lemma 3.1.

Proof of Lemma 3.1.

We will prove the estimate (3.22) first for BFB_{F}, then for B𝒵B_{\mathcal{Z}}, and finally for B𝒴B_{\mathcal{Y}}.

1. We start by considering the operator BFB_{F} associated to the function FF defined by (3.16). Notice that the assumption that b1+b2=1b_{1}+b_{2}=1 implies that B1(j)+B2(j)=1B_{1}(j)+B_{2}(j)=1 for j=0,,T01j=0,\dots,T_{0}-1. It follows from the definition of FF that

BF=1T0𝚠𝒲(T0)(j=0T011l{wj=1})B𝚠=1T0j=0T01𝚠𝒲(T0)wj=1B𝚠=1T0j=0T01B1(j).B_{F}=\frac{1}{T_{0}}\sum_{\mathtt{w}\in\mathcal{W}(T_{0})}\left(\sum_{j=0}^{T_{0}-1}\operatorname{1\hskip-2.75ptl}_{\left\{w_{j}=1\right\}}\right)B_{\mathtt{w}}=\frac{1}{T_{0}}\sum_{j=0}^{T_{0}-1}\sum_{\begin{subarray}{c}\mathtt{w}\in\mathcal{W}(T_{0})\\ w_{j}=1\end{subarray}}B_{\mathtt{w}}=\frac{1}{T_{0}}\sum_{j=0}^{T_{0}-1}B_{1}(j).

Consequently, using Lemma 3.9, we find that for u𝐍(θ)u\in\mathcal{H}_{\mathbf{N}}(\theta) we have

BFuCOp𝐍,θ(a)u+CT0𝐫M(u)+C𝐍1u.\left\|B_{F}u\right\|_{\mathcal{H}}\leq C\left\|\operatorname{Op}_{\mathbf{N},\theta}(a)u\right\|_{\mathcal{H}}+CT_{0}\mathbf{r}_{M}(u)+C\mathbf{N}^{-1}\left\|u\right\|_{\mathcal{H}}.

Recalling the definition (3.14) of T0T_{0}, we find that, up to making CC larger, we have

BFuCOp𝐍,θ(a)u+C𝐫M(u)log𝐍+C𝐍1u.\left\|B_{F}u\right\|_{\mathcal{H}}\leq C\left\|\operatorname{Op}_{\mathbf{N},\theta}(a)u\right\|_{\mathcal{H}}+C\mathbf{r}_{M}(u)\log\mathbf{N}+C\mathbf{N}^{-1}\left\|u\right\|_{\mathcal{H}}. (3.40)

2. By the definition (3.17) of 𝒵\mathcal{Z}, we have Fα1l𝒵F\geq\alpha\operatorname{1\hskip-2.75ptl}_{\mathcal{Z}}. By Lemma 3.8, we deduce from (3.40) that, for some new C>0C>0 depending on α\alpha and every u𝐍(θ)u\in\mathcal{H}_{\mathbf{N}}(\theta), we have

B𝒵uCOp𝐍,θ(a)u+C𝐫M(u)log𝐍+C𝐍12+12J(1+2log2log|λ+|)+εu.\left\|B_{\mathcal{Z}}u\right\|_{\mathcal{H}}\leq C\left\|\operatorname{Op}_{\mathbf{N},\theta}(a)u\right\|_{\mathcal{H}}+C\mathbf{r}_{M}(u)\log\mathbf{N}+C\mathbf{N}^{-\frac{1}{2}+\frac{1}{2J}\left(1+\frac{2\log 2}{\log\left|\lambda_{+}\right|}\right)+\varepsilon}\left\|u\right\|_{\mathcal{H}}. (3.41)

3. From our assumption that b1+b2=1b_{1}+b_{2}=1, we deduce that B𝒲(T0)=IB_{\mathcal{W}(T_{0})}=I. Hence we have similarly to (3.36)

B𝒴==12JM𝐍,θ(2J1)T0(B𝒲(T0)𝒵M𝐍,θT0)2JB𝒵M𝐍,θ(1)T0.B_{\mathcal{Y}}=\sum_{\ell=1}^{2J}M_{\mathbf{N},\theta}^{-(2J-1)T_{0}}(B_{\mathcal{W}(T_{0})\setminus\mathcal{Z}}M_{\mathbf{N},\theta}^{T_{0}})^{2J-\ell}B_{\mathcal{Z}}M_{\mathbf{N},\theta}^{(\ell-1)T_{0}}.

Similarly to (3.35) we have B𝒵,B𝒲(T0)𝒵1+C𝐍δ\|B_{\mathcal{Z}}\|_{\mathcal{H}\to\mathcal{H}},\|B_{\mathcal{W}(T_{0})\setminus\mathcal{Z}}\|_{\mathcal{H}\to\mathcal{H}}\leq 1+C\mathbf{N}^{-\delta} for some δ>0\delta>0, so by (3.39)

B𝒴u\displaystyle\|B_{\mathcal{Y}}u\|_{\mathcal{H}} 2=12JB𝒵M𝐍,θ(1)T0u\displaystyle\leq 2\sum_{\ell=1}^{2J}\|B_{\mathcal{Z}}M_{\mathbf{N},\theta}^{(\ell-1)T_{0}}u\|_{\mathcal{H}} (3.42)
4JB𝒵u+C𝐫M(u)log𝐍.\displaystyle\leq 4J\|B_{\mathcal{Z}}u\|_{\mathcal{H}}+C\mathbf{r}_{M}(u)\log\mathbf{N}.

Now (3.22) follows from (3.41) and (3.42). ∎

3.5. Reduction to a fractal uncertainty principle

We now explain how Lemma 3.4 may be deduced from a Fractal Uncertainty Principle type statement, Proposition 3.10 below. The proof of Proposition 3.10 is given in §4.3.

Let 𝚠𝒲(2T1)\mathtt{w}\in\mathcal{W}(2T_{1}). Decompose the word 𝚠\mathtt{w} into two words of length T1T_{1}:

𝚠=𝚠+𝚠,𝚠±𝒲(T1).\mathtt{w}=\mathtt{w}_{+}\mathtt{w}_{-},\quad\mathtt{w}_{\pm}\in\mathcal{W}(T_{1}).

Then, we relabel 𝚠+\mathtt{w}_{+} and 𝚠\mathtt{w}_{-} as

𝚠+=wT1+w1+,𝚠=w0wT11,\mathtt{w}_{+}=w_{T_{1}}^{+}\dots w_{1}^{+},\quad\mathtt{w}_{-}=w_{0}^{-}\dots w_{T_{1}-1}^{-},

and define the symbols

b+=j=1T1bwj+Aj,b=j=0T11bwjAj.b_{+}=\prod_{j=1}^{T_{1}}b_{w_{j}^{+}}\circ A^{-j},\quad b_{-}=\prod_{j=0}^{T_{1}-1}b_{w_{j}^{-}}\circ A^{j}. (3.43)

In §4, we will prove the following estimate, where h=(2π𝐍)1h=(2\pi\mathbf{N})^{-1}, Oph\operatorname{Op}_{h} is the Weyl quantization on n\mathbb{R}^{n} defined in (2.1), and we treat b±C(𝕋2n)b_{\pm}\in C^{\infty}(\mathbb{T}^{2n}) as 2n\mathbb{Z}^{2n}-periodic functions in C(2n)C^{\infty}(\mathbb{R}^{2n}).

Proposition 3.10.

Assume that the complements 𝕋2nsuppb1\mathbb{T}^{2n}\setminus\operatorname{supp}b_{1}, 𝕋2nsuppb2\mathbb{T}^{2n}\setminus\operatorname{supp}b_{2} are both safe in the sense of Definition 3.3. Then there are constants C,β>0C,\beta>0 that do not depend on 𝚠\mathtt{w} nor 𝐍\mathbf{N} such that

Oph(b)Oph(b+)L2(n)L2(n)Chβ.\left\|\operatorname{Op}_{h}(b_{-})\operatorname{Op}_{h}(b_{+})\right\|_{L^{2}(\mathbb{R}^{n})\to L^{2}(\mathbb{R}^{n})}\leq Ch^{\beta}.

In order to take advantage of Proposition 3.10, let us first notice that the proof of Lemma 3.7 also gives without major modification:

Lemma 3.11.

Let ε>0\varepsilon>0. The symbols bb_{-} and b+b_{+} belong respectively to the symbol classes SL,ρ+ε,ρ+ε(𝕋2n)S_{L_{-},\rho+\varepsilon,\rho^{\prime}+\varepsilon}(\mathbb{T}^{2n}) and SL+,ρ+ε,ρ+ε(𝕋2n)S_{L_{+},\rho+\varepsilon,\rho^{\prime}+\varepsilon}(\mathbb{T}^{2n}), with bounds on the semi-norms that do not depend on 𝐍,𝚠\mathbf{N},\mathtt{w}. (Here L±L_{\pm} are defined in (3.8).) Moreover, we have

B𝚠+(T1)\displaystyle B_{\mathtt{w}_{+}}(-T_{1}) =Op𝐍,θ(b+)+𝒪(𝐍ρ+ρ+ε1)𝐍(θ)𝐍(θ),\displaystyle=\operatorname{Op}_{\mathbf{N},\theta}(b_{+})+\mathcal{O}\left(\mathbf{N}^{\rho+\rho^{\prime}+\varepsilon-1}\right)_{\mathcal{H}_{\mathbf{N}}(\theta)\to\mathcal{H}_{\mathbf{N}}(\theta)},
B𝚠\displaystyle B_{\mathtt{w}_{-}} =Op𝐍,θ(b)+𝒪(𝐍ρ+ρ+ε1)𝐍(θ)𝐍(θ),\displaystyle=\operatorname{Op}_{\mathbf{N},\theta}(b_{-})+\mathcal{O}\left(\mathbf{N}^{\rho+\rho^{\prime}+\varepsilon-1}\right)_{\mathcal{H}_{\mathbf{N}}(\theta)\to\mathcal{H}_{\mathbf{N}}(\theta)},

where the constants in 𝒪()\mathcal{O}(\bullet) are uniform in 𝐍,𝚠\mathbf{N},\mathtt{w}.

Using Proposition 3.10 and Lemma 3.11, we get a uniform bound on the operator norm of B𝚠B_{\mathtt{w}}.

Lemma 3.12.

Assume that the complements 𝕋2nsuppb1\mathbb{T}^{2n}\setminus\operatorname{supp}b_{1}, 𝕋2nsuppb2\mathbb{T}^{2n}\setminus\operatorname{supp}b_{2} are both safe. Then there are constants C,β>0C,\beta>0 that do not depend on 𝚠\mathtt{w} nor 𝐍\mathbf{N} such that

B𝚠C𝐍βfor all𝚠𝒲(2T1).\left\|B_{\mathtt{w}}\right\|_{\mathcal{H}\to\mathcal{H}}\leq C\mathbf{N}^{-\beta}\quad\text{for all}\quad\mathtt{w}\in\mathcal{W}(2T_{1}).
Proof.

We start by noticing that

B𝚠=M𝐍,θT1B𝚠B𝚠+(T1)M𝐍,θT1.B_{\mathtt{w}}=M_{\mathbf{N},\theta}^{-T_{1}}B_{\mathtt{w}_{-}}B_{\mathtt{w}_{+}}(-T_{1})M_{\mathbf{N},\theta}^{T_{1}}. (3.44)

By Lemma 3.11, the operators Op𝐍,θ(b±)\operatorname{Op}_{\mathbf{N},\theta}(b_{\pm}) are bounded in norm uniformly in 𝚠,𝐍\mathtt{w},\mathbf{N}. Hence, we deduce from Lemma 3.11 and (3.44) that, for some C>0C>0, we have

B𝚠COp𝐍,θ(b)Op𝐍,θ(b+)+C𝐍ρ+ρ+ε1.\left\|B_{\mathtt{w}}\right\|_{\mathcal{H}\to\mathcal{H}}\leq C\left\|\operatorname{Op}_{\mathbf{N},\theta}(b_{-})\operatorname{Op}_{\mathbf{N},\theta}(b_{+})\right\|_{\mathcal{H}\to\mathcal{H}}+C\mathbf{N}^{\rho+\rho^{\prime}+\varepsilon-1}.

Now, we deduce from (2.43) that for every gC(𝕋2n;N)g\in C^{\infty}(\mathbb{T}^{2n};\mathcal{H}_{N}) we have

ΠNOph(b)Oph(b+)ΠNg(θ)=Op𝐍,θ(b)Op𝐍,θ(b+)g(θ).\Pi_{N}\operatorname{Op}_{h}(b_{-})\operatorname{Op}_{h}(b_{+})\Pi_{N}^{*}g(\theta)=\operatorname{Op}_{\mathbf{N},\theta}(b_{-})\operatorname{Op}_{\mathbf{N},\theta}(b_{+})g(\theta).

Hence, using Lemma 2.6, we find as we did for (2.45) that

Op𝐍,θ(b)Op𝐍,θ(b+)Oph(b)Oph(b+)L2L2,\left\|\operatorname{Op}_{\mathbf{N},\theta}(b_{-})\operatorname{Op}_{\mathbf{N},\theta}(b_{+})\right\|_{\mathcal{H}\to\mathcal{H}}\leq\left\|\operatorname{Op}_{h}(b_{-})\operatorname{Op}_{h}(b_{+})\right\|_{L^{2}\to L^{2}},

and the result follows then immediately from Proposition 3.10. ∎

In order to get an estimate on the operator norm of B𝒳B_{\mathcal{X}} from Lemma 3.12, we will need the following bound on the cardinal of 𝒳\mathcal{X}.

Lemma 3.13.

There are a constant c>0c>0 (that does not depend on α\alpha) and a constant C>0C>0 (that may depend on α\alpha) such that for 𝐍\mathbf{N} large enough we have

#𝒳C(log𝐍)2J𝐍2cραlog|λ+|.\#\mathcal{X}\leq C(\log\mathbf{N})^{2J}\mathbf{N}^{\frac{2c\rho\sqrt{\alpha}}{\log\left|\lambda_{+}\right|}}.
Proof.

By (3.17), a word 𝚟=v0vT01\mathtt{v}=v_{0}\dots v_{T_{0}-1} belongs to 𝒲(T0)𝒵\mathcal{W}(T_{0})\setminus\mathcal{Z} if and only if the set {j{0,,T01}:vj=1}\left\{j\in\left\{0,\dots,T_{0}-1\right\}:v_{j}=1\right\} has fewer than αT0\alpha T_{0} elements. Hence, assuming α<1/2\alpha<1/2 and recalling the definition (3.14) of T0T_{0}, we have

#(𝒲(T0)𝒵)0αT0(T0)(αT0+1)(T0αT0)Clog𝐍exp((αlogα+(1α)log(1α))T0)C𝐍cραJlog|λ+|log𝐍.\begin{split}\#\big{(}\mathcal{W}(T_{0})\setminus\mathcal{Z}\big{)}&\leq\sum_{0\leq\ell\leq\alpha T_{0}}\begin{pmatrix}T_{0}\\ \ell\end{pmatrix}\leq\left(\alpha T_{0}+1\right)\begin{pmatrix}T_{0}\\ \lceil\alpha T_{0}\rceil\end{pmatrix}\\ &\leq C\log\mathbf{N}\exp\big{(}-(\alpha\log\alpha+(1-\alpha)\log(1-\alpha))T_{0}\big{)}\\ &\leq C\mathbf{N}^{\frac{c\rho\sqrt{\alpha}}{J\log\left|\lambda_{+}\right|}}\log\mathbf{N}.\end{split}

Here, we applied Stirling’s formula and the constant c>0c>0 is such that

(αlogα+(1α)log(1α))cα for all α(0,1).-(\alpha\log\alpha+(1-\alpha)\log(1-\alpha))\leq c\sqrt{\alpha}\quad\textup{ for all }\quad\alpha\in(0,1).

The result then follows from the fact that #𝒳=#(𝒲(T0)𝒵)2J\#\mathcal{X}=\#(\mathcal{W}(T_{0})\setminus\mathcal{Z})^{2J}. ∎

We have now all the tools required to prove Lemma 3.4.

Proof of Lemma 3.4.

We choose α>0\alpha>0 small enough so that

2cραlog|λ+|<β,\frac{2c\rho\sqrt{\alpha}}{\log\left|\lambda_{+}\right|}<\beta, (3.45)

where β\beta is from Lemma 3.12 and cc is from Lemma 3.13. Then we see that B𝒳B_{\mathcal{X}} is the sum of at most #𝒳\#\mathcal{X} terms each of whose operator norms is 𝒪(𝐍β)\mathcal{O}(\mathbf{N}^{-\beta}). Hence, we have for some C>0C>0,

B𝒳C(log𝐍)2J𝐍β+2cραlog|λ+|,\left\|B_{\mathcal{X}}\right\|_{\mathcal{H}\to\mathcal{H}}\leq C(\log\mathbf{N})^{2J}\mathbf{N}^{-\beta+\frac{2c\rho\sqrt{\alpha}}{\log\left|\lambda_{+}\right|}},

and the lemma follows due to (3.45). ∎

4. Decay for long words

In this section, we use the fractal uncertainty principle, Proposition 4.2 below, to prove Proposition 3.10 and end the proof of Theorems 2 and 3. In §4.1, we recall the definition of porosity and the statement of the fractal uncertainty principle. In §4.2, we establish porosity estimates for the supports of bb_{-} and b+b_{+} from Proposition 3.10, which allows us to use the fractal uncertainty principle in §4.3 to prove Proposition 3.10.

4.1. Fractal uncertainty principle

The central tool of our proof is the fractal uncertainty principle, due originally to Bourgain–Dyatlov [BD18]. Roughly speaking, it states that a function in L2()L^{2}(\mathbb{R}) cannot be localized in both position and (semiclassically rescaled) frequency near a fractal set. To make the statement precise, we use the following

Definition 4.1.

Let ν(0,1)\nu\in(0,1) and τ0τ1\tau_{0}\leq\tau_{1} be nonnegative real numbers. Let XX be a subset of \mathbb{R}. We say that XX is ν\nu-porous on scales τ0\tau_{0} to τ1\tau_{1} if for every interval II\subset\mathbb{R} of length |I|[τ0,τ1]\left|I\right|\in\left[\tau_{0},\tau_{1}\right], there is a subinterval JIJ\subset I of length |J|=ν|I|\left|J\right|=\nu\left|I\right| such that JX=J\cap X=\emptyset.

We also recall the definition of the 11-dimensional semiclassical Fourier transform:

hf(x)=(2πh)12eihxηf(η)𝑑η,fL2().\mathcal{F}_{h}f(x)=(2\pi h)^{-\frac{1}{2}}\int_{\mathbb{R}}e^{-\frac{i}{h}x\eta}f(\eta)\,d\eta,\quad f\in L^{2}(\mathbb{R}). (4.1)

Denote by 1lX:L2()L2()\operatorname{1\hskip-2.75ptl}_{X}:L^{2}(\mathbb{R})\to L^{2}(\mathbb{R}) the multiplication operator by the indicator function of XX. We use the following extension of the fractal uncertainty principle of [BD18] proved by Dyatlov–Jin–Nonnenmacher [DJN22, Proposition 2.10] (where we put γ0±:=ϱ\gamma_{0}^{\pm}:=\varrho, γ1±:=0\gamma_{1}^{\pm}:=0 in the notation of [DJN22]):

Proposition 4.2.

Let ν(0,1)\nu\in(0,1) and ϱ(12,1]\varrho\in(\frac{1}{2},1]. Then there exist C,β>0C,\beta>0 such that for every h(0,1)h\in(0,1) and every X,YX,Y\subset\mathbb{R} which are ν\nu-porous on scales hϱh^{\varrho} to 11, we have

1lXh1lYL2()L2()Chβ.\left\|\operatorname{1\hskip-2.75ptl}_{X}\mathcal{F}_{h}\operatorname{1\hskip-2.75ptl}_{Y}\right\|_{L^{2}\left(\mathbb{R}\right)\to L^{2}\left(\mathbb{R}\right)}\leq Ch^{\beta}.

Remark. The condition that ϱ>12\varrho>{1\over 2} is essential. Indeed, the sets X=Y=[110h,110h]X=Y=[-{1\over 10}\sqrt{h},{1\over 10}\sqrt{h}] are 13{1\over 3}-porous on scales h\sqrt{h} to 1, but 1lXh1lYL2L2\left\|\operatorname{1\hskip-2.75ptl}_{X}\mathcal{F}_{h}\operatorname{1\hskip-2.75ptl}_{Y}\right\|_{L^{2}\to L^{2}} does not go to 0 as h0h\to 0, as can be checked by applying the operator in question to the function h14χ(h12x)h^{-{1\over 4}}\chi(h^{-{1\over 2}}x) where χCc((110,110))\chi\in C^{\infty}_{\mathrm{c}}((-{1\over 10},{1\over 10})) has L2L^{2} norm 1.

4.2. Porosity property

In order to use Proposition 4.2, we need to establish certain porosity properties for sets related to the support of bb_{-} and b+b_{+} from Proposition 3.10. Recall that the symbols b±b_{\pm} are defined in (3.43) using arbitrary words 𝚠±𝒲(T1)\mathtt{w}_{\pm}\in\mathcal{W}(T_{1}), where the long logarithmic propagation time T1T_{1} is defined in (3.14). The functions b1,b2b_{1},b_{2} used in the definitions of b±b_{\pm} satisfy (3.1) and the complements 𝕋2nsuppb1\mathbb{T}^{2n}\setminus\operatorname{supp}b_{1}, 𝕋2nsuppb2\mathbb{T}^{2n}\setminus\operatorname{supp}b_{2} are both safe in the sense of Definition 3.3.

Fix eigenvectors e±2n{0}e_{\pm}\in\mathbb{R}^{2n}\setminus\{0\} of AA associated to the eigenvalues λ±\lambda_{\pm} (see (3.7)). Note that by (3.9) and (3.10) we have σ(e+,e)0\sigma(e_{+},e_{-})\neq 0. We choose e±e_{\pm} so that we get the following identity used in §4.3.3 below:

σ(e+,e)=1.\sigma(e_{+},e_{-})=1. (4.2)

We let φ±t\varphi_{\pm}^{t} be the translation flows on the torus corresponding to e±e_{\pm}, that is

φ±t(z)=z+te±mod2nforz𝕋2n,t.\varphi_{\pm}^{t}(z)=z+te_{\pm}\bmod\mathbb{Z}^{2n}\quad\text{for}\quad z\in\mathbb{T}^{2n},\ t\in\mathbb{R}. (4.3)

Recall that the subtori 𝕋±\mathbb{T}_{\pm} are defined as the projections to 𝕋2n\mathbb{T}^{2n} of the spaces V±V_{\pm}\otimes\mathbb{R} where V±2nV_{\pm}\subset\mathbb{Q}^{2n} are minimal subspaces such that e±V±e_{\pm}\in V_{\pm}\otimes\mathbb{R}.

Lemma 4.3.

Let z𝕋2nz\in\mathbb{T}^{2n}. Then the closure in 𝕋2n\mathbb{T}^{2n} of the orbit of zz under φ±t\varphi_{\pm}^{t} is

{φ±t(z)t}¯=z+𝕋±.\overline{\{\varphi_{\pm}^{t}(z)\mid t\in\mathbb{R}\}}=z+\mathbb{T}_{\pm}. (4.4)
Proof.

Let us consider for instance the case of φ+t\varphi_{+}^{t}. By (4.3), it suffices to show that G=𝕋+G=\mathbb{T}_{+} where

G:=e+mod2n¯𝕋2n.G:=\overline{\mathbb{R}e_{+}\bmod\mathbb{Z}^{2n}}\ \subset\ \mathbb{T}^{2n}.

Since e+mod2n𝕋+\mathbb{R}e_{+}\bmod\mathbb{Z}^{2n}\subset\mathbb{T}_{+}, we have G𝕋+G\subset\mathbb{T}_{+}, and let us prove the reciprocal inclusion.

The set GG is a closed subgroup of 𝕋2n\mathbb{T}^{2n}, thus it is a Lie subgroup. Let 𝔤2n\mathfrak{g}\subset\mathbb{R}^{2n} be the Lie algebra of GG; since GG is connected, the exponential map 𝔤G\mathfrak{g}\to G is onto and thus G𝔤/ZG\simeq\mathfrak{g}/Z where Z:=𝔤2nZ:=\mathfrak{g}\cap\mathbb{Z}^{2n}. Since GG is compact, the rank of the lattice ZZ is equal to the dimension of 𝔤\mathfrak{g}, thus 𝔤=V\mathfrak{g}=V\otimes\mathbb{R} where V2nV\subset\mathbb{Q}^{2n} is the subspace generated by ZZ. Since e+𝔤e_{+}\in\mathfrak{g}, by the definition of V+V_{+} we have V+VV_{+}\subset V. This implies that V+𝔤V_{+}\otimes\mathbb{R}\subset\mathfrak{g} and thus 𝕋+G\mathbb{T}_{+}\subset G as needed. ∎

We now fix a constant C0>0C_{0}>0 (to be chosen in Step 2 of the proof of Lemma 4.6 in §4.3.3 below) and introduce for z𝕋2nz\in\mathbb{T}^{2n} the hh-dependent sets

Ω±(z):={t:vL such that |v|C0hρ and φ±t(z)+vsupp b}.\Omega_{\pm}(z):=\left\{t\in\mathbb{R}\colon\exists v\in L_{\mp}\text{ such that }\left|v\right|\leq C_{0}h^{\rho^{\prime}}\textup{ and }\varphi_{\pm}^{t}(z)+v\in\textup{supp }b_{\mp}\right\}. (4.5)

Here 𝐍=(2πh)1\mathbf{N}=(2\pi h)^{-1} as before; L±L_{\pm} and ρ\rho^{\prime} have been introduced in §3.1.1, see in particular (3.8) and (3.12). We can visualize the sets Ω±(z)\Omega_{\pm}(z) as follows: let us lift suppb\operatorname{supp}b_{\mp} to a subset of 2n\mathbb{R}^{2n} and zz to a point in 2n\mathbb{R}^{2n}. The set

{z+te±+v:t,vL,|v|C0hρ}\{z+te_{\pm}+v\colon t\in\mathbb{R},\ v\in L_{\mp},\ |v|\leq C_{0}h^{\rho^{\prime}}\}

is a cylinder in 2n=e±L\mathbb{R}^{2n}=\mathbb{R}e_{\pm}\oplus L_{\mp}, and Ω±(z)\Omega_{\pm}(z) is the projection onto the \mathbb{R} direction of the intersection of this cylinder with suppb\operatorname{supp}b_{\mp}.

The porosity statement needed in order to apply Proposition 4.2 is the following

Lemma 4.4.

Let ϱ(0,ρ)\varrho\in(0,\rho). Then there exist ν,h0(0,1)\nu,h_{0}\in(0,1), independent of 𝐍,𝚠\mathbf{N},\mathtt{w} such that, if 0<hh00<h\leq h_{0}, then for every z𝕋2nz\in\mathbb{T}^{2n}, the sets Ω+(z)\Omega_{+}(z) and Ω(z)\Omega_{-}(z) are ν\nu-porous on scales hϱh^{\varrho} to 11.

Remarks. 1. Lemma 4.4 is where we use the assumption that the complements 𝕋2nsuppb1\mathbb{T}^{2n}\setminus\operatorname{supp}b_{1}, 𝕋2nsuppb2\mathbb{T}^{2n}\setminus\operatorname{supp}b_{2} are both safe.

2. The choice of scales in Lemma 4.4 can be explained as follows using (3.15) (taking Ω+\Omega_{+} to simplify notation). On one hand, since Ae+=λ+e+Ae_{+}=\lambda_{+}e_{+}, the map AT1A^{T_{1}} expands the vector e+e_{+} by |λ+|T1hρhϱ|\lambda_{+}|^{T_{1}}\sim h^{-\rho}\gg h^{-\varrho}. Thus we expect porosity of suppb\operatorname{supp}b_{-} in the direction of e+e_{+} on scales from hϱh^{\varrho} to 1. On the other hand, by (3.11), the same map AT1A^{T_{1}} sends the ball {vL:|v|C0hρ}\{v\in L_{-}\colon|v|\leq C_{0}h^{\rho^{\prime}}\} to a set of diameter C0γT1hρ1\leq C_{0}\gamma^{T_{1}}h^{\rho^{\prime}}\ll 1, so changing φ+t(z)\varphi_{+}^{t}(z) by an element of this ball does not change much the forward trajectory under AA up to time T1T_{1}, which is used to define the symbol bb_{-} in (3.43).

Proof.

We show the porosity of Ω+(z)\Omega_{+}(z), with the case of Ω(z)\Omega_{-}(z) handled similarly (reversing the direction of time).

1. Since the complements 𝕋2nsuppb1\mathbb{T}^{2n}\setminus\operatorname{supp}b_{1}, 𝕋2nsuppb2\mathbb{T}^{2n}\setminus\operatorname{supp}b_{2} are safe, by Lemma 3.5 there exist compact subsets K1,K2𝕋2nK_{1},K_{2}\subset\mathbb{T}^{2n} such that the interiors K1,K2K_{1}^{\circ},K_{2}^{\circ} are safe and

K1suppb1=K2suppb2=.K_{1}\cap\operatorname{supp}b_{1}=K_{2}\cap\operatorname{supp}b_{2}=\emptyset.

We claim that there exist constants R>1R>1, r>0r>0 such that for any =1,2\ell=1,2, the intersection of every length RR flow line of φ+t\varphi_{+}^{t} with KK_{\ell} contains a segment of length rr. (Here the length of flow lines is defined using the parametrization by tt.)

We argue by contradiction: assume that such R,rR,r do not exist, then there is a sequence zm𝕋2nz_{m}\in\mathbb{T}^{2n} such that for each mm the intersection K{φ+t(zm):|t|m}K_{\ell}\cap\{\varphi_{+}^{t}(z_{m})\colon|t|\leq m\} does not contain any segment of length 1/m1/m. Passing to a subsequence, we may assume that the sequence zmz_{m} converges to some z𝕋2nz_{\infty}\in\mathbb{T}^{2n} as mm\to\infty. Since the interior KK_{\ell}^{\circ} is safe, it intersects z+𝕋+z_{\infty}+\mathbb{T}_{+}. Then by Lemma 4.3 there exists tt\in\mathbb{R} such that φ+t(z)K\varphi_{+}^{t}(z_{\infty})\in K_{\ell}^{\circ}. For mm large enough the segment {φ+s(zm):s[t,t+1/m]}\big{\{}\varphi_{+}^{s}(z_{m})\colon s\in[t,t+1/m]\big{\}} lies inside KK_{\ell}^{\circ}. This contradicts our assumption and proves the claim.

2. Let z𝕋2nz\in\mathbb{T}^{2n}. We will show that Ω+(z)\Omega_{+}(z) is ν\nu-porous on scales hϱh^{\varrho} to 1 for

ν:=rR|λ+|.\nu:={r\over R|\lambda_{+}|}.

Let II\subset\mathbb{R} be an interval of length between hϱh^{\varrho} and 11. Let jj denote the smallest integer such that |λ+|j|I|R\left|\lambda_{+}\right|^{j}\left|I\right|\geq R. By the definition (3.14) of T1T_{1}, and recalling that 𝐍=(2πh)1\mathbf{N}=(2\pi h)^{-1},

|λ+|T11|I|(2π)ρ|λ+|(1+J)hϱρh0+,\left|\lambda_{+}\right|^{T_{1}-1}\left|I\right|\geq(2\pi)^{-\rho}\left|\lambda_{+}\right|^{-(1+J)}h^{\varrho-\rho}\underset{h\to 0}{\to}+\infty,

thus 0<j<T10<j<T_{1}, provided hh is small enough.

Since Ae+=λ+e+Ae_{+}=\lambda_{+}e_{+}, the set Aj({φ+t(z):tI})A^{j}\left(\left\{\varphi_{+}^{t}(z):t\in I\right\}\right) is a flow line of φ+\varphi_{+} of length |λ+|j|I|R\left|\lambda_{+}\right|^{j}\left|I\right|\geq R. Consequently, the intersection of this set with KwjK_{w_{j}^{-}} contains a segment of length rr. It follows that there exists a segment JIJ\subset I of length |λ+|jr|\lambda_{+}|^{-j}r such that

Ajφ+t(z)Kwjfor alltJ.A^{j}\varphi_{+}^{t}(z)\in K_{w_{j}^{-}}\quad\text{for all}\quad t\in J. (4.6)

By our choice of jj, we have |J|ν|I||J|\geq\nu|I|. It remains to prove that JΩ+(z)=J\cap\Omega_{+}(z)=\emptyset. Recalling the definition (4.5) of Ω+(z)\Omega_{+}(z), we see that it suffices to show that for each tJt\in J we have

φ+t(z)+vsuppbfor allvLsuch that|v|C0hρ.\varphi_{+}^{t}(z)+v\not\in\operatorname{supp}b_{-}\quad\text{for all}\quad v\in L_{-}\quad\text{such that}\quad|v|\leq C_{0}h^{\rho^{\prime}}.

We have Aj(φ+t(z)+v)=Ajφ+t(z)+AjvA^{j}(\varphi_{+}^{t}(z)+v)=A^{j}\varphi_{+}^{t}(z)+A^{j}v. Recalling the bound (3.11) on the norm of AjA^{j} restricted to LL_{-}, as well as the condition (3.12) we imposed on ρ\rho^{\prime}, we see that (with the constant CC depending on C0C_{0})

|Ajv|CγjhρCγT1hρC(2π)ρlogγlog|λ+|hρρlogγlog|λ+|h00.\left|A^{j}v\right|\leq C\gamma^{j}h^{\rho^{\prime}}\leq C\gamma^{T_{1}}h^{\rho^{\prime}}\leq C(2\pi)^{-\rho\frac{\log\gamma}{\log\left|\lambda_{+}\right|}}h^{\rho^{\prime}-\rho\frac{\log\gamma}{\log\left|\lambda_{+}\right|}}\underset{h\to 0}{\to}0.

By (4.6) and since Kwjsuppbwj=K_{w_{j}^{-}}\cap\operatorname{supp}b_{w_{j}^{-}}=\emptyset, we see that Aj(φ+t(z)+v)suppbwjA^{j}(\varphi_{+}^{t}(z)+v)\not\in\operatorname{supp}b_{w_{j}^{-}}, provided hh is small enough so that |Ajv||A^{j}v| is less than the distance between KwjK_{w_{j}^{-}} and suppbwj\operatorname{supp}b_{w_{j}^{-}}. Recalling the definition (3.43) of bb_{-}, it follows that φ+t(z)+vsuppb\varphi_{+}^{t}(z)+v\not\in\operatorname{supp}b_{-}, which finishes the proof. ∎

4.3. Proof of Proposition 3.10

We now give the proof of Proposition 3.10, relying on the following three ingredients:

  • the fact that for any ε>0\varepsilon>0, we have b±SL±,ρ+ε,ρ+ε(𝕋2n)b_{\pm}\in S_{L_{\pm},\rho+\varepsilon,\rho^{\prime}+\varepsilon}(\mathbb{T}^{2n}) uniformly in hh and in the words 𝚠±\mathtt{w}_{\pm} (see Lemma 3.11);

  • the porosity property of the supports suppb±\operatorname{supp}b_{\pm} given by Lemma 4.4;

  • and the fractal uncertainty princple in the form of Proposition 4.2.

Henceforth we treat b±b_{\pm} as 2n\mathbb{Z}^{2n}-invariant symbols in SL±,ρ+ε,ρ+ε(2n)S_{L_{\pm},\rho+\varepsilon,\rho^{\prime}+\varepsilon}(\mathbb{R}^{2n}). All the constants in the estimates below are independent of h,𝚠±h,\mathtt{w}_{\pm}.

4.3.1. Decomposing the operator and the scheme of the proof

We start by decomposing the operator Oph(b)Oph(b+)\operatorname{Op}_{h}(b_{-})\operatorname{Op}_{h}(b_{+}) into a series, see (4.8) below. For that, fix a function

ψ~Cc((1,1)2n;),k2nψ~(zk)2=1for allz2n.\widetilde{\psi}\in C^{\infty}_{\mathrm{c}}((-1,1)^{2n};\mathbb{R}),\quad\sum_{k\in\mathbb{Z}^{2n}}\widetilde{\psi}(z-k)^{2}=1\quad\text{for all}\quad z\in\mathbb{R}^{2n}.

For instance, we can start with χCc((1,1)2n;[0,1])\chi\in C^{\infty}_{\mathrm{c}}((-1,1)^{2n};[0,1]) such that the 2n\mathbb{Z}^{2n}-periodic function F(x):=k2nχ(xk)2F(x):=\sum_{k\in\mathbb{Z}^{2n}}\chi(x-k)^{2} is everywhere positive, and put ψ~(x):=F(x)12χ(x)\widetilde{\psi}(x):=F(x)^{-{1\over 2}}\chi(x).

Now, consider the partition of unity 1=k2nψk21=\sum_{k\in\mathbb{Z}^{2n}}\psi_{k}^{2} where the hh-dependent symbol ψkCc(2n)\psi_{k}\in C^{\infty}_{\mathrm{c}}(\mathbb{R}^{2n}) is given by

ψk(z):=ψ~(zhρk),k2n.\psi_{k}(z):=\widetilde{\psi}\Big{(}{z\over h^{\rho^{\prime}}}-k\Big{)},\quad k\in\mathbb{Z}^{2n}. (4.7)

Recalling Definition 2.2, we see that ψk\psi_{k} lies in SL,ρ,ρ(2n)S_{L,\rho,\rho^{\prime}}(\mathbb{R}^{2n}) uniformly in h,kh,k for any coisotropic subspace L2nL\subset\mathbb{R}^{2n}.

We have the following decomposition, with the series below converging in the strong operator topology as an operator L2(n)L2(n)L^{2}(\mathbb{R}^{n})\to L^{2}(\mathbb{R}^{n}), which can be checked by applying it to a function in 𝒮(n)\mathscr{S}(\mathbb{R}^{n}):

Oph(b)Oph(b+)=k2nPkwherePk:=Oph(b)Oph(ψk2)Oph(b+).\operatorname{Op}_{h}(b_{-})\operatorname{Op}_{h}(b_{+})=\sum_{k\in\mathbb{Z}^{2n}}P_{k}\quad\text{where}\quad P_{k}:=\operatorname{Op}_{h}(b_{-})\operatorname{Op}_{h}(\psi_{k}^{2})\operatorname{Op}_{h}(b_{+}). (4.8)

We now state two estimates on the operators PkP_{k} which together will give Proposition 3.10. The first one is an almost orthogonality type statement when |k||k-\ell| is sufficiently large:

Lemma 4.5.

For every m>0m>0 there exists a constant Cm>0C_{m}>0 such that for every k,2nk,\ell\in\mathbb{Z}^{2n} such that |k|10n|k-\ell|\geq 10\sqrt{n} we have

PkPL2L2Cmhm|k|m and PkPL2L2Cmhm|k|m.\left\|P_{k}^{*}P_{\ell}\right\|_{L^{2}\to L^{2}}\leq C_{m}h^{m}\left|k-\ell\right|^{-m}\quad\textup{ and }\quad\left\|P_{k}P_{\ell}^{*}\right\|_{L^{2}\to L^{2}}\leq C_{m}h^{m}\left|k-\ell\right|^{-m}.

The second one is the norm bound on each individual PkP_{k}, which uses the fractal uncertainty principle:

Lemma 4.6.

There exist constants C,β>0C,\beta>0 such that for every k2nk\in\mathbb{Z}^{2n} we have

PkL2L2Chβ.\|P_{k}\|_{L^{2}\to L^{2}}\leq Ch^{\beta}.

Here the constant β\beta only depends on the porosity constant ν\nu in Lemma 4.4 and on ρ\rho.

Before proving Lemmas 4.5 and 4.6, let us explain how they imply Proposition 3.10:

Proof of Proposition 3.10.

It follows from Lemma 4.5 (with m:=4n+1m:=4n+1) and Lemma 4.6 that there are constants C,β>0C,\beta>0 such that

supk2n2nPkPL2L212Chβ and supk2n2nPkPL2L212Chβ.\sup_{k\in\mathbb{Z}^{2n}}\sum_{\ell\in\mathbb{Z}^{2n}}\left\|P_{k}^{*}P_{\ell}\right\|_{L^{2}\to L^{2}}^{\frac{1}{2}}\leq Ch^{\beta}\quad\textup{ and }\quad\sup_{k\in\mathbb{Z}^{2n}}\sum_{\ell\in\mathbb{Z}^{2n}}\left\|P_{k}P_{\ell}^{*}\right\|_{L^{2}\to L^{2}}^{\frac{1}{2}}\leq Ch^{\beta}.

Hence, it follows from the Cotlar–Stein Theorem [Zwo12, Theorem C.5] and the decomposition (4.8) that Oph(b)Oph(b+)L2L2Chβ\|\operatorname{Op}_{h}(b_{-})\operatorname{Op}_{h}(b_{+})\|_{L^{2}\to L^{2}}\leq Ch^{\beta} as needed. ∎

4.3.2. Almost orthogonality

We are left with the proofs of Lemmas 4.5 and 4.6. We start with Lemma 4.5:

Proof of Lemma 4.5.

1. Let k,2nk,\ell\in\mathbb{Z}^{2n} be such that |k|10n|k-\ell|\geq 10\sqrt{n}. Define the linear functional q:2nq:\mathbb{R}^{2n}\to\mathbb{R} by

q(z)=z,k|k|,z2n.q(z)={\langle z,\ell-k\rangle\over|k-\ell|},\quad z\in\mathbb{R}^{2n}.

Note that qq has norm 1. Putting r0:=hρq(k+2)r_{0}:=h^{\rho^{\prime}}q({k+\ell\over 2}), we have

suppψk\displaystyle\operatorname{supp}\psi_{k} {z2n|q(z)r0hρ|k|4},\displaystyle\subset\big{\{}z\in\mathbb{R}^{2n}\,\big{|}\,q(z)\leq r_{0}-h^{\rho^{\prime}}\textstyle{|k-\ell|\over 4}\}, (4.9)
suppψ\displaystyle\operatorname{supp}\psi_{\ell} {z2n|q(z)r0+hρ|k|4}.\displaystyle\subset\big{\{}z\in\mathbb{R}^{2n}\,\big{|}\,q(z)\geq r_{0}+h^{\rho^{\prime}}\textstyle{|k-\ell|\over 4}\}.

Indeed, assume that zsuppψkz\in\operatorname{supp}\psi_{k}. Then |q(hρzk)||hρzk|2n|q(h^{-\rho^{\prime}}z-k)|\leq|h^{-\rho^{\prime}}z-k|\leq\sqrt{2n}. Since hρq(k)=r0hρ|k|2h^{\rho^{\prime}}q(k)=r_{0}-h^{\rho^{\prime}}{|k-\ell|\over 2} and |k|10n|k-\ell|\geq 10\sqrt{n}, we have q(z)r0hρ|k|4q(z)\leq r_{0}-h^{\rho^{\prime}}{|k-\ell|\over 4}. This gives the first statement in (4.9), with the second one proved similarly. Note that (4.9) implies in particular that suppψksuppψ=\operatorname{supp}\psi_{k}\cap\operatorname{supp}\psi_{\ell}=\emptyset.

2. We estimate the norm PkPL2L2\|P_{k}^{*}P_{\ell}\|_{L^{2}\to L^{2}}, with PkPL2L2\|P_{k}P_{\ell}^{*}\|_{L^{2}\to L^{2}} estimated in a similar way. We write using (2.5) and (2.7)

PkP=Oph(b¯+)Oph(ψk2)Oph(b¯#b)Oph(ψ2)Oph(b+).P_{k}^{*}P_{\ell}=\operatorname{Op}_{h}(\bar{b}_{+})\operatorname{Op}_{h}(\psi_{k}^{2})\operatorname{Op}_{h}(\bar{b}_{-}\#b_{-})\operatorname{Op}_{h}(\psi_{\ell}^{2})\operatorname{Op}_{h}(b_{+}).

It follows from Lemmas 2.3 and 3.11 that Oph(b¯+)\operatorname{Op}_{h}(\bar{b}_{+}) is uniformly bounded on L2L^{2}. Thus it suffices to show that

Oph(ψk2)Oph(b¯#b)Oph(ψ2)L2L2Cmhm|k|m.\|\operatorname{Op}_{h}(\psi_{k}^{2})\operatorname{Op}_{h}(\bar{b}_{-}\#b_{-})\operatorname{Op}_{h}(\psi_{\ell}^{2})\|_{L^{2}\to L^{2}}\leq C_{m}h^{m}|k-\ell|^{-m}. (4.10)

To show (4.10), it suffices to apply Lemma 2.4 with

r:=hρ|k|4,a:=ψk2,b:=ψ2,c:=b¯#b.r:=h^{\rho^{\prime}}\textstyle\frac{|k-\ell|}{4},\quad a:=\psi_{k}^{2},\quad b:=\psi_{\ell}^{2},\quad c:=\bar{b}_{-}\#b_{-}.

Here for each ε>0\varepsilon>0 the symbols a,b,ca,b,c are bounded in the class SL,ρ+ε,ρ+ε(2n)S_{L_{-},\rho+\varepsilon,\rho^{\prime}+\varepsilon}(\mathbb{R}^{2n}) uniformly in hh by Lemmas 2.3 and 3.11, as well as (4.7). The support condition of Lemma 2.4 is satisfied by (4.9). ∎

4.3.3. Decay for an individual summand

Finally, we apply the fractal uncertainty principle to prove Lemma 4.6.

Proof of Lemma 4.6.

1. Recall that the ψk\psi_{k}’s belong uniformly to both the symbol classes SL+,ρ,ρS_{L_{+},\rho,\rho^{\prime}} and SL,ρ,ρS_{L_{-},\rho,\rho^{\prime}}. Recalling Lemma 3.11, we can apply the product formula from Lemma 2.3 to find that for every ε>0\varepsilon>0

Pk=Oph(bψk)Oph(b+ψk)+𝒪(h1ρρε)L2L2.P_{k}=\operatorname{Op}_{h}(b_{-}\psi_{k})\operatorname{Op}_{h}(b_{+}\psi_{k})+\mathcal{O}(h^{1-\rho-\rho^{\prime}-\varepsilon})_{L^{2}\to L^{2}}.

Therefore it suffices to show that

Oph(bψk)Oph(b+ψk)L2L2Chβ.\|\operatorname{Op}_{h}(b_{-}\psi_{k})\operatorname{Op}_{h}(b_{+}\psi_{k})\|_{L^{2}\to L^{2}}\leq Ch^{\beta}. (4.11)

2. We next study the supports of the symbols b±ψkb_{\pm}\psi_{k}. We have from (4.7)

suppψkhρk+(hρ,hρ)2n.\operatorname{supp}\psi_{k}\ \subset\ h^{\rho^{\prime}}k+(-h^{\rho^{\prime}},h^{\rho^{\prime}})^{2n}.

Thus by (3.9) any zsuppψkz\in\operatorname{supp}\psi_{k} can be written as z=hρk+t±e±+vz=h^{\rho^{\prime}}k+t_{\pm}e_{\pm}+v_{\mp} where t±t_{\pm}\in\mathbb{R}, vLv_{\mp}\in L_{\mp}, and |v|C0hρ|v_{\mp}|\leq C_{0}h^{\rho^{\prime}} for some constant C0C_{0} depending only on the matrix AA. Choose s±(k)s_{\pm}^{(k)}\in\mathbb{R} such that hρks±(k)e±+Lh^{\rho^{\prime}}k\in s_{\pm}^{(k)}e_{\pm}+L_{\mp}. Put z(k):=hρkmod2n𝕋2nz^{(k)}:=h^{\rho^{\prime}}k\bmod\mathbb{Z}^{2n}\in\mathbb{T}^{2n}. Recalling the definition (4.5) of the sets Ω±(z)\Omega_{\pm}(z), we get

supp(bψk)tΩ~±(te±+L)whereΩ~±:=s±(k)+Ω±(z(k)).\operatorname{supp}(b_{\mp}\psi_{k})\ \subset\ \bigcup_{t\in\widetilde{\Omega}_{\pm}}\big{(}te_{\pm}+L_{\mp}\big{)}\quad\text{where}\quad\widetilde{\Omega}_{\pm}:=s_{\pm}^{(k)}+\Omega_{\pm}(z^{(k)})\ \subset\ \mathbb{R}. (4.12)

3. We now conjugate by a metaplectic transformation which ‘straightens out’ the vectors e±e_{\pm} and the subspaces L±L_{\pm}. Using (3.10), (4.2), and the linear version of Darboux’s Theorem, we construct a symplectic matrix QSp(2n,)Q\in\operatorname{Sp}(2n,\mathbb{R}) such that

  • Qx1=eQ\partial_{x_{1}}=e_{-} and Qξ1=e+Q\partial_{\xi_{1}}=e_{+};

  • Qspan(x1,,xn,ξ2,,ξn)=LQ\operatorname{span}(\partial_{x_{1}},\dots,\partial_{x_{n}},\partial_{\xi_{2}},\dots,\partial_{\xi_{n}})=L_{-};

  • Qspan(x2,,xn,ξ1,,ξn)=L+Q\operatorname{span}(\partial_{x_{2}},\dots,\partial_{x_{n}},\partial_{\xi_{1}},\dots,\partial_{\xi_{n}})=L_{+}.

Let M~Q\widetilde{M}\in\mathcal{M}_{Q} be a metaplectic operator associated to QQ (see §2.1.3). Then by (2.17)

Oph(bψk)Oph(b+ψk)L2L2=M~1Oph(bψk)Oph(b+ψk)M~L2L2=Oph((bψk)Q)Oph((b+ψk)Q)L2L2.\begin{split}\left\|\operatorname{Op}_{h}(b_{-}\psi_{k})\operatorname{Op}_{h}(b_{+}\psi_{k})\right\|_{L^{2}\to L^{2}}&=\left\|\widetilde{M}^{-1}\operatorname{Op}_{h}(b_{-}\psi_{k})\operatorname{Op}_{h}(b_{+}\psi_{k})\widetilde{M}\right\|_{L^{2}\to L^{2}}\\ &=\left\|\operatorname{Op}_{h}\big{(}(b_{-}\psi_{k})\circ Q\big{)}\operatorname{Op}_{h}\big{(}(b_{+}\psi_{k})\circ Q\big{)}\right\|_{L^{2}\to L^{2}}.\end{split}

Thus (4.11) reduces to

Oph((bψk)Q)Oph((b+ψk)Q)L2L2Chβ\left\|\operatorname{Op}_{h}\big{(}(b_{-}\psi_{k})\circ Q\big{)}\operatorname{Op}_{h}\big{(}(b_{+}\psi_{k})\circ Q\big{)}\right\|_{L^{2}\to L^{2}}\leq Ch^{\beta} (4.13)

and the support condition (4.12) becomes

supp((bψk)Q)\displaystyle\operatorname{supp}\big{(}(b_{-}\psi_{k})\circ Q\big{)} {(x,ξ)ξ1Ω~+},\displaystyle\subset\ \{(x,\xi)\mid\xi_{1}\in\widetilde{\Omega}_{+}\}, (4.14)
supp((b+ψk)Q)\displaystyle\operatorname{supp}\big{(}(b_{+}\psi_{k})\circ Q\big{)} {(x,ξ)x1Ω~}.\displaystyle\subset\ \{(x,\xi)\mid x_{1}\in\widetilde{\Omega}_{-}\}.

4. For δ>0\delta>0, denote the δ\delta-neighborhood of Ω~±\widetilde{\Omega}_{\pm} by

Ω~±(δ):=Ω~±+[δ,δ].\widetilde{\Omega}_{\pm}(\delta):=\widetilde{\Omega}_{\pm}+[-\delta,\delta].

Let χ\chi_{\mp} be the convolutions of the indicator functions of Ω~±(12hρ)\widetilde{\Omega}_{\pm}({1\over 2}h^{\rho}) with the function hρχ(hρt)h^{-\rho}\chi(h^{-\rho}t) where χCc((12,12))\chi\in C^{\infty}_{\mathrm{c}}((-{1\over 2},{1\over 2})) is a nonnegative function integrating to 1. Then

χC(;[0,1]),suppχΩ~±(hρ),χ=1onΩ~±\chi_{\mp}\in C^{\infty}(\mathbb{R};[0,1]),\quad\operatorname{supp}\chi_{\mp}\subset\widetilde{\Omega}_{\pm}(h^{\rho}),\quad\chi_{\mp}=1\quad\text{on}\quad\widetilde{\Omega}_{\pm}

and for each \ell there exists a constant CC_{\ell} (depending only on \ell and the choice of χ\chi) such that

supt|tχ±(t)|Chρ.\sup_{t\in\mathbb{R}}|\partial_{t}^{\ell}\chi_{\pm}(t)|\leq C_{\ell}h^{-\rho\ell}.

Define the symbols χ~±C(2n)\widetilde{\chi}_{\pm}\in C^{\infty}(\mathbb{R}^{2n}) by

χ~(x,ξ)=χ(ξ1),χ~+(x,ξ)=χ+(x1).\widetilde{\chi}_{-}(x,\xi)=\chi_{-}(\xi_{1}),\quad\widetilde{\chi}_{+}(x,\xi)=\chi_{+}(x_{1}).

Then χ~±\widetilde{\chi}_{\pm} lie in the symbol class SQ1L±,ρ,0(2n)S_{Q^{-1}L_{\pm},\rho,0}(\mathbb{R}^{2n}) uniformly in hh. On the other hand, by Lemma 3.11 the symbols (b±ψk)Q(b_{\pm}\psi_{k})\circ Q lie in the larger class SQ1L±,ρ+ε,ρ+ε(2n)S_{Q^{-1}L_{\pm},\rho+\varepsilon,\rho^{\prime}+\varepsilon}(\mathbb{R}^{2n}) uniformly in hh for every fixed ε>0\varepsilon>0. By Lemma 2.3 and since (b±ψk)Q=χ~±((b±ψk)Q)(b_{\pm}\psi_{k})\circ Q=\widetilde{\chi}_{\pm}((b_{\pm}\psi_{k})\circ Q) by (4.14), we have

Oph((bψk)Q)\displaystyle\operatorname{Op}_{h}\big{(}(b_{-}\psi_{k})\circ Q\big{)} =Oph((bψk)Q)Oph(χ~)+𝒪(h1ρρε)L2L2,\displaystyle=\operatorname{Op}_{h}\big{(}(b_{-}\psi_{k})\circ Q\big{)}\operatorname{Op}_{h}(\widetilde{\chi}_{-})+\mathcal{O}(h^{1-\rho-\rho^{\prime}-\varepsilon})_{L^{2}\to L^{2}},
Oph((b+ψk)Q)\displaystyle\operatorname{Op}_{h}\big{(}(b_{+}\psi_{k})\circ Q\big{)} =Oph(χ~+)Oph((b+ψk)Q)+𝒪(h1ρρε)L2L2.\displaystyle=\operatorname{Op}_{h}(\widetilde{\chi}_{+})\operatorname{Op}_{h}\big{(}(b_{+}\psi_{k})\circ Q\big{)}+\mathcal{O}(h^{1-\rho-\rho^{\prime}-\varepsilon})_{L^{2}\to L^{2}}.

Note that Oph(χ~+)=χ+(x1)\operatorname{Op}_{h}(\widetilde{\chi}_{+})=\chi_{+}(x_{1}) is a multiplication operator and Oph(χ~)=χ(ihx1)\operatorname{Op}_{h}(\widetilde{\chi}_{-})=\chi_{-}(-ih\partial_{x_{1}}) is a Fourier multiplier (see [Zwo12, Theorem 4.9]). Multiplying the above estimates and using that Oph((b±ψk)Q)\operatorname{Op}_{h}((b_{\pm}\psi_{k})\circ Q) are bounded uniformly as operators on L2(n)L^{2}(\mathbb{R}^{n}), we reduce (4.13) to the following estimate:

χ(ihx1)χ+(x1)L2(n)L2(n)Chβ.\|\chi_{-}(-ih\partial_{x_{1}})\chi_{+}(x_{1})\|_{L^{2}(\mathbb{R}^{n})\to L^{2}(\mathbb{R}^{n})}\leq Ch^{\beta}. (4.15)

5. If we consider L2(n)L^{2}(\mathbb{R}^{n}) as the Hilbert tensor product L2()L2(n1)L^{2}(\mathbb{R})\otimes L^{2}(\mathbb{R}^{n-1}) corresponding to the decomposition x=(x1,x)x=(x_{1},x^{\prime}), x:=(x2,,xn)x^{\prime}:=(x_{2},\dots,x_{n}), then the operators χ+(x1)\chi_{+}(x_{1}) and χ(ihx1)\chi_{-}(-ih\partial_{x_{1}}) are the tensor products of the same operators in one variable with the identity operator on L2(n1)L^{2}(\mathbb{R}^{n-1}). Thus (4.15) is equivalent to

χ(ihx1)χ+(x1)L2()L2()Chβ\|\chi_{-}(-ih\partial_{x_{1}})\chi_{+}(x_{1})\|_{L^{2}(\mathbb{R})\to L^{2}(\mathbb{R})}\leq Ch^{\beta} (4.16)

where we now treat the factors in the product as operators on L2()L^{2}(\mathbb{R}). Denote by h:L2()L2()\mathcal{F}_{h}:L^{2}(\mathbb{R})\to L^{2}(\mathbb{R}) the unitary semiclassical Fourier transform, see (4.1). Then χ(ihx1)=h1χ(x1)h\chi_{-}(-ih\partial_{x_{1}})=\mathcal{F}_{h}^{-1}\chi_{-}(x_{1})\mathcal{F}_{h}. Thus the left-hand side of (4.16) is equal to χ(x1)hχ+(x1)L2()L2()\|\chi_{-}(x_{1})\mathcal{F}_{h}\chi_{+}(x_{1})\|_{L^{2}(\mathbb{R})\to L^{2}(\mathbb{R})}. Since χ±=χ±1lΩ~(hρ)\chi_{\pm}=\chi_{\pm}\operatorname{1\hskip-2.75ptl}_{\widetilde{\Omega}_{\mp}(h^{\rho})} and |χ±|1|\chi_{\pm}|\leq 1, the bound (4.16) reduces to

1lΩ~+(hρ)h1lΩ~(hρ)L2()L2()Chβ.\|\operatorname{1\hskip-2.75ptl}_{\widetilde{\Omega}_{+}(h^{\rho})}\mathcal{F}_{h}\operatorname{1\hskip-2.75ptl}_{\widetilde{\Omega}_{-}(h^{\rho})}\|_{L^{2}(\mathbb{R})\to L^{2}(\mathbb{R})}\leq Ch^{\beta}. (4.17)

6. We finally apply the fractal uncertainty principle. Fix ϱ(12,ρ)\varrho\in({1\over 2},\rho), which is possible since ρ>12\rho>{1\over 2} by (3.12). By Lemma 4.4, there exists ν>0\nu>0 such that the sets Ω~±\widetilde{\Omega}_{\pm} are ν\nu-porous on scales hϱh^{\varrho} to 1. Since hρhϱh^{\rho}\ll h^{\varrho} for h1h\ll 1, the neighborhoods Ω~±(hρ)\widetilde{\Omega}_{\pm}(h^{\rho}) are ν3\nu\over 3-porous on scales hϱh^{\varrho} to 1 – see for example [DJN22, Lemma 2.11]. Now (4.17) follows from the fractal uncertainty principle of Proposition 4.2, and the proof is finished. ∎

Appendix A Properties of integer symplectic matrices

In this Appendix, we discuss the algebraic hypotheses made on the matrix AA in Theorems 2, 3, and 4. More precisely, we investigate the spaces V+V_{+} and VV_{-} (and hence the tori 𝕋+\mathbb{T}_{+} and 𝕋\mathbb{T}_{-}) defined in (1.6). In particular, we prove Lemma A.3 that allows us to deduce Theorem 1 from Theorem 2.

A.1. Algebraic considerations

We start by giving a new characterization of V+V_{+} and VV_{-}. Let ASp(2n,)A\in\operatorname{Sp}(2n,\mathbb{Z}) satisfy the spectral gap condition (1.3) and recall from the introduction that V±V_{\pm} were defined as the smallest subspaces of 2n\mathbb{Q}^{2n} such that E±V±E_{\pm}\subset V_{\pm}\otimes\mathbb{R} where E±2nE_{\pm}\subset\mathbb{R}^{2n} are the eigenspaces of AA corresponding to the eigenvalues λ+\lambda_{+} and λ:=λ+1\lambda_{-}:=\lambda_{+}^{-1}.

We will be using basic field theory, see e.g. [DF04, Chapter 13]. Recall that for an algebraic number λ\lambda\in\mathbb{C}, its minimal polynomial (over \mathbb{Q}) is the unique irreducible monic polynomial P[x]P\in\mathbb{Q}[x] such that P(λ)=0P(\lambda)=0. Two algebraic numbers are called Galois conjugates if they have the same minimal polynomial.

Lemma A.1.

Let P±P_{\pm} denote the minimal polynomials of λ±\lambda_{\pm}. Then V±=kerP±(A)V_{\pm}=\ker P_{\pm}(A). The dimensions of V±V_{\pm} are equal to each other and to the degrees of P±P_{\pm}. Moreover, we have the following two cases:

  1. (1)

    if λ+\lambda_{+} is a Galois conjugate of λ\lambda_{-}, then V+=VV_{+}=V_{-};

  2. (2)

    otherwise V+V={0}V_{+}\cap V_{-}=\{0\}.

Proof.

1. We first show that V+=kerP+(A)V_{+}=\ker P_{+}(A) and dimV+=degP+\dim V_{+}=\deg P_{+}. (The case of VV_{-} is treated similarly.) Note that kerP+(A)\ker P_{+}(A) is an AA-invariant subspace of 2n\mathbb{Q}^{2n}. Any (complex) eigenvalue of the endomorphism A|kerP+(A)A|_{\ker P_{+}(A)} has to be a root of P+P_{+}, thus the characteristic polynomial P~+[x]\widetilde{P}_{+}\in\mathbb{Q}[x] of A|kerP+(A)A|_{\ker P_{+}(A)} is a power of P+P_{+}. On the other hand, P~+\widetilde{P}_{+} divides the characteristic polynomial of AA. Since λ+\lambda_{+} is a simple eigenvalue of AA, we see that P~+=P+\widetilde{P}_{+}=P_{+}.

Since P+(λ+)=0P_{+}(\lambda_{+})=0, we see that E+kerP+(A)E_{+}\subset\ker P_{+}(A)\otimes\mathbb{R}, and consequently we have V+kerP+(A)V_{+}\subset\ker P_{+}(A). As V+V_{+} is AA-invariant, the characteristic polynomial of the endomorphism A|V+A|_{V_{+}} divides the characteristic polynomial P+P_{+} of the endomorphism A|kerP+(A)A|_{\ker P_{+}(A)}. Since P+P_{+} is irreducible over \mathbb{Q} and dimV+>0\dim V_{+}>0, we see that these two characteristic polynomials are equal. It follows that V+=kerP+(A)V_{+}=\ker P_{+}(A) and dimV+=degP+\dim V_{+}=\deg P_{+}.

2. Recall that the degree of P±P_{\pm} is the dimension of the field (λ±)\mathbb{Q}(\lambda_{\pm}) as a vector field over \mathbb{Q}. Since λ+=λ1\lambda_{+}=\lambda_{-}^{-1}, we have (λ+)=(λ)\mathbb{Q}(\lambda_{+})=\mathbb{Q}(\lambda_{-}), so that degP+=degP\deg P_{+}=\deg P_{-}. It follows that dimV+=dimV\dim V_{+}=\dim V_{-}.

3. Since P+P_{+} and PP_{-} are irreducible over the rationals, either they are coprime, in which case V+V={0}V_{+}\cap V_{-}=\{0\}, or they are equal, in which case V+=VV_{+}=V_{-}, due to the characterization we just proved. If P+=PP_{+}=P_{-}, then P+(λ)=0P_{+}(\lambda_{-})=0 and λ\lambda_{-} is a Galois conjugate of λ+\lambda_{+}. Reciprocally, if P+(λ)=0P_{+}(\lambda_{-})=0, then P+P_{+} and PP_{-} are not coprime, so that P+=PP_{+}=P_{-}. ∎

In order to discuss the sharpness of Theorem 4, we introduce a decomposition of 2n\mathbb{Q}^{2n}. For a subspace V2nV\subset\mathbb{Q}^{2n}, denote by Vσ2nV^{\perp\sigma}\subset\mathbb{Q}^{2n} its symplectic complement, see (2.19). Recall that VV is called symplectic if VVσ={0}V\cap V^{\perp\sigma}=\{0\}.

Lemma A.2.

We have the following two cases:

  1. (1)

    if λ+\lambda_{+} is a Galois conjugate of λ\lambda_{-}, then V+=VV_{+}=V_{-} is symplectic;

  2. (2)

    otherwise V±V_{\pm} are both isotropic and the symplectic form σ\sigma is nondegenerate on V++VV_{+}+V_{-}.

Consequently, we have a decomposition of 2n\mathbb{Q}^{2n} into

2n=V0V1,\mathbb{Q}^{2n}=V_{0}\oplus V_{1},

where V0=V++VV_{0}=V_{+}+V_{-} and V1=(V++V)σV_{1}=(V_{+}+V_{-})^{\perp\sigma} are symplectic.

Proof.

1. For each complex eigenvalue λSpec(A)\lambda\in\operatorname{Spec}(A), define the space of generalized eigenvectors

V(λ):={v2n0:(AλI)v=0}.V(\lambda):=\{v\in\mathbb{C}^{2n}\mid\exists\ell\geq 0:\ (A-\lambda I)^{\ell}v=0\}.

Then we have the decomposition

2n=λSpec(A)V(λ).\mathbb{C}^{2n}=\bigoplus_{\lambda\in\operatorname{Spec}(A)}V(\lambda). (A.1)

We claim that for all λ,λSpec(A)\lambda,\lambda^{\prime}\in\operatorname{Spec}(A) such that λλ1\lambda\lambda^{\prime}\neq 1,

σ(v,v)=0for allvV(λ),vV(λ).\sigma(v,v^{\prime})=0\quad\text{for all}\quad v\in V(\lambda),\ v^{\prime}\in V(\lambda^{\prime}). (A.2)

To prove (A.2), we argue by induction on +\ell+\ell^{\prime} where ,0\ell,\ell^{\prime}\geq 0 are the smallest numbers such that (AλI)v=(AλI)v=0(A-\lambda I)^{\ell}v=(A-\lambda^{\prime}I)^{\ell^{\prime}}v^{\prime}=0. If =0\ell=0 or =0\ell^{\prime}=0, then σ(v,v)=0\sigma(v,v^{\prime})=0 since v=0v=0 or v=0v^{\prime}=0. Otherwise we use that AA is symplectic to write

σ(v,v)=σ(Av,Av)=λλσ(v,v)+σ((AλI)v,Av)+σ(λv,(AλI)v).\sigma(v,v^{\prime})=\sigma(Av,Av^{\prime})=\lambda\lambda^{\prime}\sigma(v,v^{\prime})+\sigma((A-\lambda I)v,Av^{\prime})+\sigma(\lambda v,(A-\lambda^{\prime}I)v^{\prime}).

Using the inductive hypothesis we see that the last two terms on the right-hand side are 0, which gives σ(v,v)=0\sigma(v,v^{\prime})=0 as needed.

By (A.1) and (A.2), we see that for all λSpec(A)\lambda\in\operatorname{Spec}(A)

V(λ)σ=λSpec(A),λλ1V(λ).V(\lambda)^{\perp\sigma}=\bigoplus_{\lambda^{\prime}\in\operatorname{Spec}(A),\,\lambda^{\prime}\neq\lambda^{-1}}V(\lambda^{\prime}). (A.3)

2. Since λ±\lambda_{\pm} is a simple eigenvalue of AA and P±P_{\pm} is its minimal polynomial, each root of P±P_{\pm} is a simple eigenvalue of AA. By Lemma A.1 we have

V±=λ,P±(λ)=0V(λ).V_{\pm}\otimes\mathbb{C}=\bigoplus_{\lambda,\,P_{\pm}(\lambda)=0}V(\lambda).

Since P±P_{\pm} are the minimal polynomials of λ±\lambda_{\pm} and λ+=λ1\lambda_{+}=\lambda_{-}^{-1}, we have P(λ)=cλdegP+P+(λ1)P_{-}(\lambda)=c\lambda^{\deg P_{+}}P_{+}(\lambda^{-1}) for some c{0}c\in\mathbb{Q}\setminus\{0\}. It follows from (A.3) that

(V±)σ=λSpec(A),P(λ)0V(λ).(V_{\pm}\otimes\mathbb{C})^{\perp\sigma}=\bigoplus_{\lambda\in\operatorname{Spec}(A),\,P_{\mp}(\lambda)\neq 0}V(\lambda).

If λ+\lambda_{+} is a Galois conjugate of λ\lambda_{-}, then P+=PP_{+}=P_{-}, so V+=VV_{+}=V_{-} is symplectic. Otherwise P+P_{+} and PP_{-} are coprime, so V±V_{\pm} are both isotropic and σ\sigma is nondegenerate on V++VV_{+}+V_{-}. ∎

Remark. Using Lemma 4.3, the algebraic consideration from this section have dynamical implication. Lemma A.1 that if z𝕋2nz\in\mathbb{T}^{2n} then the closure of the orbits of zz for the flows (φ+t)t(\varphi^{t}_{+})_{t\in\mathbb{R}} and (φt)t(\varphi^{t}_{-})_{t\in\mathbb{R}} are either identical (if λ+\lambda_{+} is a Galois conjugate of λ\lambda_{-}) or have a finite number of points of intersection.

From Lemma A.2, we know that if z𝕋2nz\in\mathbb{T}^{2n} then the closure of the orbit of zz under the action by translation of the 22-dimensional vector space generated by e+e_{+} and ee_{-} is always a symplectic subtorus of 𝕋2n\mathbb{T}^{2n}.

A.2. Most favorable cases

Theorem 4 gives a condition on the support of semiclassical measures for AA in terms of the spaces V+V_{+} and VV_{-}. The larger these spaces are, the stronger the conclusion of Theorem 4 is. Considering the decomposition from Lemma A.2, the most favorable case is when V1V_{1} is trivial. In that situation, there are still two possibilities according to Lemma A.1:

  1. (1)

    2n=V+=V\mathbb{Q}^{2n}=V_{+}=V_{-}, or

  2. (2)

    2n=V+V\mathbb{Q}^{2n}=V_{+}\oplus V_{-} and V±V_{\pm} are Lagrangian.

In case (1), Theorem 4 says that all semiclassical measures for AA are fully supported. Actually, this is exactly the setting of Theorem 1, as we prove now.

Lemma A.3.

The characteristic polynomial of AA is irreducible over \mathbb{Q} if and only if 2n=V+=V\mathbb{Q}^{2n}=V_{+}=V_{-} (that is 𝕋2n=𝕋+=𝕋\mathbb{T}^{2n}=\mathbb{T}_{+}=\mathbb{T}_{-}, or equivalently the flows (φ+t)t(\varphi_{+}^{t})_{t\in\mathbb{R}} and (φt)t(\varphi_{-}^{t})_{t\in\mathbb{R}} are minimal).

Proof.

Notice that P+P_{+} divides the characteristic polynomial of AA and recall that the dimension of V+V_{+} is the degree of P+P_{+}. Hence, if V+V_{+} is equal to 2n\mathbb{Q}^{2n}, the degree of P+P_{+} is 2n2n and P+P_{+} must be the characteristic polynomial of AA, which is consequently irreducible. Reciprocally, if the characteristic polynomial of AA is irreducible, it must be equal to P+P_{+}, so that V+=2nV_{+}=\mathbb{Q}^{2n}. ∎

Of course, when V+=V=2nV_{+}=V_{-}=\mathbb{Q}^{2n}, the control of the support of semiclassical measures for AA given by Theorem 4 is sharp. When n=1n=1, AA satisfies the spectral gap condition (1.3) if and only if it is hyperbolic (i.e. it has no eigenvalues on the unit circle), and in this case we always have 2=V+=V\mathbb{Q}^{2}=V_{+}=V_{-}. When n>1n>1, one can easily construct examples of matrices satisfying (1.3) with irreducible characteristic polynomials. For example, when n=2n=2 one can take

A=(0010000110010112)A=\begin{pmatrix}0&0&1&0\\ 0&0&0&1\\ -1&0&0&1\\ 0&-1&1&2\end{pmatrix} (A.4)

with the characteristic polynomial

P(λ)=λ42λ3+λ22λ+1=(λ2(1+2)λ+1)(λ2(12)λ+1)P(\lambda)=\lambda^{4}-2\lambda^{3}+\lambda^{2}-2\lambda+1=(\lambda^{2}-(1+\sqrt{2})\lambda+1)(\lambda^{2}-(1-\sqrt{2})\lambda+1)

which has a root in (0,1)(0,1), a root in (1,)(1,\infty), and two complex roots on the unit circle.

Let us now consider the case (2), when 2n=V+V\mathbb{Q}^{2n}=V_{+}\oplus V_{-}. Our result is still sharp in this situation since, under some mild additional assumptions, Kelmer [Kel10, Theorem 1] constructed semiclassical measures supported in some translate of 𝕋+\mathbb{T}_{+} and semiclassical measures supported in some translate of 𝕋\mathbb{T}_{-}. A basic example (previously presented by Gurevich [Gur05] and Kelmer [Kel10]) is

A=(B00BT),A=\begin{pmatrix}B&0\\ 0&B^{-T}\end{pmatrix}, (A.5)

where BGL(n,)B\in\operatorname{GL}(n,\mathbb{Z}), |detB|=1|\det B|=1, has irreducible characteristic polynomial and a leading simple eigenvalue, that also dominates the inverses of the eigenvalues of BB (so that AA satisfies the spectral gap condition (1.3)). One can take for instance

B=(010001110).B=\begin{pmatrix}0&1&0\\ 0&0&1\\ 1&1&0\end{pmatrix}.

Using the coordinates (x,ξ)(x,\xi) on 2n\mathbb{R}^{2n}, we see that the spaces V±V_{\pm} are given by V+={ξ=0}V_{+}=\{\xi=0\}, V={x=0}V_{-}=\{x=0\}. Note that if we allow BB to be in GL(n,)\operatorname{GL}(n,\mathbb{Q}), then, when 2n=V+V\mathbb{Q}^{2n}=V_{+}\oplus V_{-}, the matrix AA is always of the form (A.5) after a symplectic (rational) change of coordinates.

Using Proposition 2.11, we see that for a matrix AA of the form (A.5) and θ=0\theta=0, the following elements of 𝐍(0)\mathcal{H}_{\mathbf{N}}(0) are eigenvectors for the quantizations M𝐍,0M_{\mathbf{N},0} of AA:

𝐞00 and 𝐍n2j𝐍n𝐞j0.\mathbf{e}_{0}^{0}\quad\textup{ and }\quad\mathbf{N}^{-\frac{n}{2}}\sum_{j\in\mathbb{Z}_{\mathbf{N}}^{n}}\mathbf{e}_{j}^{0}.

It follows from Proposition 2.10 that these eigenvectors converge respectively to the semiclassical measures

a𝕋na(0,ξ)𝑑ξ and a𝕋na(x,0)𝑑x.a\mapsto\int_{\mathbb{T}^{n}}a(0,\xi)\,d\xi\quad\textup{ and }\quad a\mapsto\int_{\mathbb{T}^{n}}a(x,0)\,dx.

These measures are supported respectively in 𝕋\mathbb{T}_{-} and 𝕋+\mathbb{T}_{+}.

A.3. General case

For now, we only considered the case in which the space V1V_{1} from Lemma A.2 is trivial. Let us now discuss what happens when V1V_{1} is non-trivial. Let 𝕋0\mathbb{T}_{0} and 𝕋1\mathbb{T}_{1} be the subtori of 𝕋2n\mathbb{T}^{2n} tangent respectively to V0V_{0} and V1V_{1}. As before, we consider two cases:

  1. (1)

    If λ+\lambda_{+} is a Galois conjugate of λ\lambda_{-}, then Theorem 4 shows that the support of every semiclassical measure contains a translate of 𝕋0=𝕋±\mathbb{T}_{0}=\mathbb{T}_{\pm}.

  2. (2)

    Otherwise Theorem 4 shows that the support of every semiclassical measure contains a translate of 𝕋+\mathbb{T}_{+} or 𝕋\mathbb{T}_{-}, which are different tori (their tangent spaces intersect trivially). On the other hand, from [Kel10, Theorem 1], we know that (under mild additional assumptions) there are semiclassical measures supported in some translate of 𝕋++𝕋1\mathbb{T}_{+}+\mathbb{T}_{1} and semiclassical measures supported in some translate of 𝕋+𝕋1\mathbb{T}_{-}+\mathbb{T}_{1}.

Note that in both cases the conclusion of Theorem 4 is not sharp. However, we cannot say more on the support of the semiclassical measures for AA without further information on the action of AA on 𝕋1\mathbb{T}_{1}.

To illustrate this fact, take a matrix BSp(2n,)B\in\operatorname{Sp}(2n,\mathbb{Z}) that satisfies (1.3) and a matrix CSp(2n,)C\in\operatorname{Sp}(2n^{\prime},\mathbb{Z}) whose eigenvalues are dominated by the leading eigenvalue of BB. Assume in addition that in the decomposition from Lemma A.2 for the matrix BB, the factor V1V_{1} is trivial. Then, we form the matrix

A:=BCSp(2(n+n),).A:=B\oplus C\ \in\ \operatorname{Sp}(2(n+n^{\prime}),\mathbb{Z}).

Notice that the matrix AA satisfies the condition (1.3) and that the action of AA on the spaces V0V_{0} and V1V_{1} is given respectively by the matrices BB and CC. The quantizations M𝐍,θM_{\mathbf{N},\theta} of AA are tensor products of quantizations of BB with quantizations of CC, with a basis of eigenfunctions consisting of tensor products of eigenfunctions corresponding to BB with those corresponding to CC. The torus decomposed as 𝕋2(n+n)=𝕋0×𝕋1\mathbb{T}^{2(n+n^{\prime})}=\mathbb{T}_{0}\times\mathbb{T}_{1} and the semiclassical measures for AA associated to eigenfunctions of product type are of the form μ=μ0×μ1\mu=\mu_{0}\times\mu_{1} where μ0\mu_{0} and μ1\mu_{1} are semiclassical measures respectively for BB and CC. We know that the support of μ0\mu_{0} must contain a translate of 𝕋+\mathbb{T}_{+} or 𝕋\mathbb{T}_{-} (and this estimate cannot be improved as discussed in §A.2).

If the semiclassical measures for CC have large supports, then Theorem 4 is not sharp. For instance, if CC satisfies (1.3) and has an irreducible characteristic polynomial over the rationals, then μ1\mu_{1} must be fully supported, so that the supports of the semiclassical measures for AA (associated to eigenfunctions of product type) contain a translate of 𝕋+×𝕋1\mathbb{T}_{+}\times\mathbb{T}_{1} or 𝕋×𝕋1\mathbb{T}_{-}\times\mathbb{T}_{1}.

However, it is not true in general that the semiclassical measures for CC have a large support. The most extreme case is when CC is given by the symplectic rotation matrix FF from (2.61). In that case, the Dirac mass at 0 is a semiclassical measure for CC (as proved in Lemma A.4 below). Hence, all the measures of the form μ0×δ0\mu_{0}\times\delta_{0} are semiclassical measures for AA, and we see that Theorem 4 is sharp in that case. A concrete example of a matrix AA for which this happens is

A=(2312)(0110)=(2030000110200100).A=\begin{pmatrix}2&3\\ 1&2\end{pmatrix}\oplus\begin{pmatrix}0&1\\ -1&0\end{pmatrix}=\begin{pmatrix}2&0&3&0\\ 0&0&0&1\\ 1&0&2&0\\ 0&-1&0&0\end{pmatrix}. (A.6)

We end this section with an example of matrix with the Dirac mass at 0 as a semiclassical measure, that was needed for our discussion above.

Lemma A.4.

Let FF be the symplectic matrix from (2.61). Then the Dirac mass at 0 is a semiclassical measure associated to FF.

Proof.

Note that φF=0\varphi_{F}=0, so the quantization condition (2.57) holds for θ=0\theta=0 and all NN. We will construct an eigenvector for the quantizations of FF using a Gaussian function localized near 0 in phase space. We start with the function

f(x)=e|x|22hL2(n),hf=ff(x)=e^{-\frac{|x|^{2}}{2h}}\in L^{2}(\mathbb{R}^{n}),\quad\mathcal{F}_{h}f=f

where hF\mathcal{F}_{h}\in\mathcal{M}_{F} is the semiclassical Fourier transform on L2(n)L^{2}(\mathbb{R}^{n}) defined in (2.65) and a quantization of FF on 𝐍(0)\mathcal{H}_{\mathbf{N}}(0) is given by h|𝐍(0)\mathcal{F}_{h}|_{\mathcal{H}_{\mathbf{N}}(0)}. Using the projector Π𝐍\Pi_{\mathbf{N}} from §2.2.2, define the state

f𝐍Π𝐍(0)f=j𝐍nf𝐍,j𝐞j0𝐍(0),f_{\mathbf{N}}\coloneqq\Pi_{\mathbf{N}}(0)f=\sum_{j\in\mathbb{Z}_{\mathbf{N}}^{n}}f_{\mathbf{N},j}\mathbf{e}_{j}^{0}\ \in\ \mathcal{H}_{\mathbf{N}}(0),

where f𝐍,j=f,𝐞j0L2f_{\mathbf{N},j}=\langle f,\mathbf{e}_{j}^{0}\rangle_{L^{2}} for j𝐍nj\in\mathbb{Z}_{\mathbf{N}}^{n}. More explicitly, recalling (2.35) we have

f𝐍,j=𝐍n2kneπ𝐍|k+j𝐍|2.f_{\mathbf{N},j}=\mathbf{N}^{-\frac{n}{2}}\sum_{k\in\mathbb{Z}^{n}}e^{-\pi\mathbf{N}\left|k+\frac{j}{\mathbf{N}}\right|^{2}}. (A.7)

Since hf=f\mathcal{F}_{h}f=f, it follows from Lemma 2.6 and the intertwining relation (2.56) that f𝐍f_{\mathbf{N}} is an eigenvector for the quantizations of FF on 𝐍(0)\mathcal{H}_{\mathbf{N}}(0). By a diagonal argument (similarly to [Zwo12, Theorem 5.2]) there is a sequence of even numbers 𝐍pp\mathbf{N}_{p}\underset{p\to\infty}{\to}\infty and a semiclassical measure μ\mu for FF such that, for every aC(𝕋2n)a\in C^{\infty}(\mathbb{T}^{2n}),

Op𝐍p,0(a)f𝐍p,f𝐍pf𝐍p2p𝕋2na𝑑μ.\frac{\langle\operatorname{Op}_{\mathbf{N}_{p},0}(a)f_{\mathbf{N}_{p}},f_{\mathbf{N}_{p}}\rangle_{\mathcal{H}}}{\left\|f_{\mathbf{N}_{p}}\right\|^{2}_{\mathcal{H}}}\underset{p\to\infty}{\to}\int_{\mathbb{T}^{2n}}a\,d\mu. (A.8)

We will show that μ\mu is the delta measure at (0,0)(0,0), which (by the diagonal argument again and since the limit of every convergent subsequence is the same) implies that the convergence statement (A.8) holds for the entire sequence f𝐍f_{\mathbf{N}}. Let us prove first that μ\mu is supported in {x=0}\{x=0\}. Let a(x)C(𝕋n)a(x)\in C^{\infty}(\mathbb{T}^{n}) be such that the ball centered at 0 of some small radius ε>0\varepsilon>0 does not intersect suppa\operatorname{supp}a. By (2.42), we have

Op𝐍,0(a)f𝐍,f𝐍=j𝐍na(j𝐍)|f𝐍,j|2.\langle\operatorname{Op}_{\mathbf{N},0}(a)f_{\mathbf{N}},f_{\mathbf{N}}\rangle_{\mathcal{H}}=\sum_{j\in\mathbb{Z}_{\mathbf{N}}^{n}}a\Big{(}{j\over\mathbf{N}}\Big{)}|f_{\mathbf{N},j}|^{2}. (A.9)

From (A.7) we get |f𝐍,j|2C𝐍ne2πε2𝐍|f_{\mathbf{N},j}|^{2}\leq C\mathbf{N}^{-n}e^{-2\pi\varepsilon^{2}\mathbf{N}} for all j𝐍nj\in\mathbb{Z}_{\mathbf{N}}^{n} such that j/𝐍suppaj/\mathbf{N}\in\operatorname{supp}a. On the other hand f𝐍|f𝐍,0|𝐍n2\|f_{\mathbf{N}}\|_{\mathcal{H}}\geq|f_{\mathbf{N},0}|\geq\mathbf{N}^{-{n\over 2}}. Thus

𝕋2na(x)𝑑μ=lim𝐍Op𝐍,0(a)f𝐍,f𝐍f𝐍2=0\int_{\mathbb{T}^{2n}}a(x)\,d\mu=\lim_{\mathbf{N}\to\infty}{\langle\operatorname{Op}_{\mathbf{N},0}(a)f_{\mathbf{N}},f_{\mathbf{N}}\rangle_{\mathcal{H}}\over\|f_{\mathbf{N}}\|_{\mathcal{H}}^{2}}=0

for all aCc(𝕋n{0})a\in C^{\infty}_{\mathrm{c}}(\mathbb{T}^{n}\setminus\{0\}), which gives that suppμ{x=0}\operatorname{supp}\mu\subset\{x=0\}. Since μ\mu is a semiclassical measure associated to FF, it is invariant under FF. Thus suppμF({x=0})={ξ=0}\operatorname{supp}\mu\subset F(\{x=0\})=\{\xi=0\}. It follows that suppμ={(0,0)}\operatorname{supp}\mu=\{(0,0)\}. Since μ\mu is a probability measure, it has to be the delta measure at (0,0)(0,0). ∎

Acknowledgements. We would like to thank the anonymous referee for a careful reading of the paper and many useful comments. SD was supported by NSF CAREER grant DMS-1749858 and a Sloan Research Fellowship. Most of this work was done while MJ was supported by the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation programme (grant agreement No 787304), and working at LPSM111Laboratoire de Probabilités, Statistique et Modélisation (LPSM), CNRS, Sorbonne Université, Université de Paris, 4, Place Jussieu, 75005 Paris, France.

References

  • [AKN09] Nalini Anantharaman, Herbert Koch, and Stéphane Nonnenmacher. Entropy of eigenfunctions. In Vladas Sidoravičius, editor, New Trends in Mathematical Physics, pages 1–22, Dordrecht, 2009. Springer Netherlands.
  • [AN07a] Nalini Anantharaman and Stéphane Nonnenmacher. Entropy of semiclassical measures of the Walsh-quantized baker’s map. Ann. Henri Poincaré, 8(1):37–74, 2007.
  • [AN07b] Nalini Anantharaman and Stéphane Nonnenmacher. Half-delocalization of eigenfunctions for the Laplacian on an Anosov manifold. Ann. Inst. Fourier (Grenoble), 57(7):2465–2523, 2007. Festival Yves Colin de Verdière.
  • [Ana08] Nalini Anantharaman. Entropy and the localization of eigenfunctions. Ann. of Math. (2), 168(2):435–475, 2008.
  • [AS13] Nalini Anantharaman and Lior Silberman. A Haar component for quantum limits on locally symmetric spaces. Israel J. Math., 195(1):393–447, 2013.
  • [BD18] Jean Bourgain and Semyon Dyatlov. Spectral gaps without the pressure condition. Ann. of Math. (2), 187(3):825–867, 2018.
  • [BDB96] Abdelkader Bouzouina and Stephan De Bièvre. Equipartition of the eigenfunctions of quantized ergodic maps on the torus. Comm. Math. Phys., 178(1):83–105, 1996.
  • [BL14] Shimon Brooks and Elon Lindenstrauss. Joint quasimodes, positive entropy, and quantum unique ergodicity. Invent. Math., 198(1):219–259, 2014.
  • [Bog92] Evgeny Bogomolny. Semiclassical quantization of multidimensional systems. Nonlinearity, 5(4):805–866, 1992.
  • [Bro10] Shimon Brooks. On the entropy of quantum limits for 2-dimensional cat maps. Comm. Math. Phys., 293(1):231–255, 2010.
  • [CdV85] Yves Colin de Verdière. Ergodicité et fonctions propres du laplacien. In Bony-Sjöstrand-Meyer seminar, 1984–1985, pages Exp. No. 13, 8. École Polytech., Palaiseau, 1985.
  • [DF04] David S. Dummit and Richard M. Foote. Abstract algebra. John Wiley & Sons, Inc., Hoboken, NJ, third edition, 2004.
  • [DJ18] Semyon Dyatlov and Long Jin. Semiclassical measures on hyperbolic surfaces have full support. Acta Math., 220(2):297–339, 2018.
  • [DJN22] Semyon Dyatlov, Long Jin, and Stéphane Nonnenmacher. Control of eigenfunctions on surfaces of variable curvature. J. Amer. Math. Soc., 35(2):361–465, 2022.
  • [Dya19] Semyon Dyatlov. An introduction to fractal uncertainty principle. J. Math. Phys., 60(8):081505, 31, 2019.
  • [Dya21] Semyon Dyatlov. Around quantum ergodicity. Annales mathématiques du Québec, 2021.
  • [DZ16] Semyon Dyatlov and Joshua Zahl. Spectral gaps, additive energy, and a fractal uncertainty principle. Geom. Funct. Anal., 26(4):1011–1094, 2016.
  • [FN04] Frédéric Faure and Stéphane Nonnenmacher. On the maximal scarring for quantum cat map eigenstates. Comm. Math. Phys., 245(1):201–214, 2004.
  • [FNDB03] Frédéric Faure, Stéphane Nonnenmacher, and Stephan De Bièvre. Scarred eigenstates for quantum cat maps of minimal periods. Comm. Math. Phys., 239(3):449–492, 2003.
  • [GH11] Shamgar Gurevich and Ronny Hadani. Proof of the Kurlberg-Rudnick rate conjecture. Ann. of Math. (2), 174(1):1–54, 2011.
  • [Gur05] Shamgar Gurevich. Weil representation, deligne sheaf and proof of the kurlberg-rudnick conjecture, 2005. Ph.D. thesis; arXiv:math-ph/0601031.
  • [Gut10] Boris Gutkin. Entropic bounds on semiclassical measures for quantized one-dimensional maps. Comm. Math. Phys., 294(2):303–342, 2010.
  • [HB80] John Hannay and Michael Berry. Quantization of linear maps on a torus-Fresnel diffraction by a periodic grating. Phys. D, 1(3):267–290, 1980.
  • [HS20] Rui Han and Wilhelm Schlag. A higher-dimensional Bourgain-Dyatlov fractal uncertainty principle. Anal. PDE, 13(3):813–863, 2020.
  • [Jin20] Long Jin. Damped wave equations on compact hyperbolic surfaces. Communications in Mathematical Physics, 373(3):771–794, 2020.
  • [Kel10] Dubi Kelmer. Arithmetic quantum unique ergodicity for symplectic linear maps of the multidimensional torus. Ann. of Math. (2), 171(2):815–879, 2010.
  • [Koh97] Winfried Kohnen. On the generators of Spn(𝐙){\rm Sp}_{n}({\bf Z}). Linear Algebra Appl., 253:363–367, 1997.
  • [Kow08] Emmanuel Kowalski. The large sieve and its applications, volume 175 of Cambridge Tracts in Mathematics. Cambridge University Press, Cambridge, 2008. Arithmetic geometry, random walks and discrete groups.
  • [KR00] Pär Kurlberg and Zeév Rudnick. Hecke theory and equidistribution for the quantization of linear maps of the torus. Duke Math. J., 103(1):47–77, 2000.
  • [Lin06] Elon Lindenstrauss. Invariant measures and arithmetic quantum unique ergodicity. Ann. of Math. (2), 163(1):165–219, 2006.
  • [Mar06] Jens Marklof. Arithmetic quantum chaos. In Jean-Pierre Françoise, Gregory L. Naber, and Tsou Sheung Tsun, editors, Encyclopedia of mathematical physics. Vol. 1, 2, 3, 4, 5, pages 212–220. Academic Press/Elsevier Science, Oxford, 2006.
  • [Non13] Stéphane Nonnenmacher. Anatomy of quantum chaotic eigenstates. In Chaos, volume 66 of Prog. Math. Phys., pages 193–238. Birkhäuser/Springer, Basel, 2013.
  • [Riv08] Igor Rivin. Walks on groups, counting reducible matrices, polynomials, and surface and free group automorphisms. Duke Math. J., 142(2):353–379, 2008.
  • [Riv10a] Gabriel Rivière. Entropy of semiclassical measures for nonpositively curved surfaces. Ann. Henri Poincaré, 11(6):1085–1116, 2010.
  • [Riv10b] Gabriel Rivière. Entropy of semiclassical measures in dimension 2. Duke Math. J., 155(2):271–336, 2010.
  • [Riv11] Gabriel Rivière. Entropy of semiclassical measures for symplectic linear maps of the multidimensional torus. Int. Math. Res. Not. IMRN, (11):2396–2443, 2011.
  • [RS94] Zeév Rudnick and Peter Sarnak. The behaviour of eigenstates of arithmetic hyperbolic manifolds. Comm. Math. Phys., 161(1):195–213, 1994.
  • [Sar11] Peter Sarnak. Recent progress on the quantum unique ergodicity conjecture. Bull. Amer. Math. Soc. (N.S.), 48(2):211–228, 2011.
  • [Sch21] Nir Schwartz. The full delocalization of eigenstates for the quantized cat map, 2021. arXiv:2103.06633.
  • [Shn74] Alexander Shnirelman. Ergodic properties of eigenfunctions. Uspehi Mat. Nauk, 29(6(180)):181–182, 1974.
  • [SZ02] Johannes Sjöstrand and Maciej Zworski. Quantum monodromy and semi-classical trace formulae. J. Math. Pures Appl. (9), 81(1):1–33, 2002.
  • [Zel87] Steven Zelditch. Uniform distribution of eigenfunctions on compact hyperbolic surfaces. Duke Math. J., 55(4):919–941, 1987.
  • [Zwo12] Maciej Zworski. Semiclassical analysis, volume 138 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, 2012.