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

Negative regularity mixing for random volume preserving diffeomorphisms

Jacob Bedrossian Department of Mathematics, University of California, Los Angeles, CA 90095, USA [email protected]. JB was supported by NSF Award DMS-2108633.    Patrick Flynn Department of Mathematics, University of California, Los Angeles, CA 90095, USA [email protected].    Sam Punshon-Smith Department of Mathematics, Tulane University, New Orleans, LA 70118, USA [email protected]. This material was based upon work supported by the NSF Award DMS-1803481.
Abstract

We consider the negative regularity mixing properties of random volume preserving diffeomorphisms on a compact manifold without boundary. We give general criteria so that the associated random transfer operator mixes HδH^{-\delta} observables exponentially fast in HδH^{-\delta} (with a deterministic rate), a property that is false in the deterministic setting. The criteria apply to a wide variety of random diffeomorphisms, such as discrete-time iid random diffeomorphisms, the solution maps of suitable classes of stochastic differential equations, and to the case of advection-diffusion by solutions of the stochastic incompressible Navier-Stokes equations on 𝕋2\mathbb{T}^{2}. In the latter case, we show that the zero diffusivity passive scalar with a stochastic source possesses a unique stationary measure describing “ideal” scalar turbulence. The proof is based on techniques inspired by the use of pseudodifferential operators and anisotropic Sobolev spaces in the deterministic setting.

1 Introduction

In this paper we study negative regularity mixing by volume-preserving random diffeomorphisms {ϕt:t𝕀}\{\phi^{t}\,:\,t\in\mathbb{I}\} on a smooth compact Riemannian manifold \mathcal{M} without boundary. Here, the index set 𝕀\mathbb{I} is either +\mathbb{R}_{+} or +\mathbb{Z}_{+}. One of our main motivations comes from fluid mechanics, namely the advection equation

tft+utft=0,f|t=0=f0\displaystyle\partial_{t}f_{t}+u_{t}\cdot\nabla f_{t}=0,\quad f|_{t=0}=f_{0} (1.1)

where ftf_{t} is a scalar field and (ut)t+(u_{t})_{t\in\mathbb{R}_{+}} is a Lipschitz regular, divergence-free, (time-dependent) velocity field with =𝕋d\mathcal{M}=\mathbb{T}^{d}. Naturally, the solution is given by ft=f0(ϕt)1f_{t}=f_{0}\circ(\phi^{t})^{-1} where ϕt=ϕu0,ωt\phi^{t}=\phi^{t}_{u_{0},\omega} is the flow map of associated to the velocity field utu_{t}. The solution operator

Tt:f0ft=f0(ϕt)1T^{t}:f_{0}\mapsto f_{t}=f_{0}\circ(\phi^{t})^{-1}

is often called the transfer operator in the dynamical systems literature.

In the context of Anosov maps, the study of mixing properties of the map ϕt\phi^{t} is closely related to the spectrum of the transfer operator, known as Ruelle Resonances. Introduced by Ruelle within the framework of thermodynamic formalism [43, 41, 42], these resonances offer refined insights into the decay of correlations. Subsequent work has significantly advanced our understanding of these resonances through the use of anisotropic Sobolev spaces [11, 28, 35, 1, 3, 16, 2] and microlocal analysis [20, 24, 21, 25, 26, 22, 23]. See section 1.2 for more context on some of this work in the context of our main result.

In the case when the velocity field is random (see [7, 9, 14, 17]), it is possible to prove a quenched mixing result (or equivalently almost-sure exponential decay of correlations), namely there exists a deterministic γ>0\gamma>0 and a random constant D=D(ω)D=D(\omega) satisfying 𝐄D2<\mathbf{E}D^{2}<\infty, such that for all mean zero f0H1f_{0}\in H^{1} there holds

ftH1D(ω)eγtf0H1.\displaystyle\left|\left|f_{t}\right|\right|_{H^{-1}}\leq D(\omega)e^{-\gamma t}\left|\left|f_{0}\right|\right|_{H^{1}}. (1.2)

In the case of (1.1) or similar examples, the random constant DD will also depend on the initial velocity field u0u_{0} (but not the initial data f0f_{0}).

Due to the time-reversibility of transport equations, it is relatively easy to conclude that one cannot obtain an almost-sure exponential decay result like (1.2) H1H^{-1} decay without assuming f0f_{0} having some small amount of regularity say HϵH^{\epsilon}. In particular one cannot expect decay in H1H^{-1} if f0f_{0} is also taken in H1H^{-1}.

The goal of this work is to prove that for many random maps, negative regularity mixing does hold in an averaged sense, namely that for all mean-zero f0Hδf_{0}\in H^{-\delta}, there exists a μ>0\mu>0 such that

𝐄ftHδ2eμtf0Hδ2.\displaystyle\mathbf{E}\left|\left|f_{t}\right|\right|_{H^{-\delta}}^{2}\lesssim e^{-\mu t}\left|\left|f_{0}\right|\right|_{H^{-\delta}}^{2}.

for some fixed μ,δ>0\mu,\delta>0 (independent of f0f_{0}), and also that it holds in a quenched sense but with a constant that depends on f0f_{0} (Corollary 1.13).

After we state the main results, we include a more extended discussion of both the results and the motivations, however, one can liken the result to something like the non-random multiplicative ergodic theorem [32]. Indeed, for each individual scalar field f0Hδf_{0}\in H^{-\delta} (and initial velocity field u0u_{0} if applicable) there may exist very specific realizations of (ut)(u_{t}) which obtain arbitrarily bad decay rates in HδH^{-\delta}, but these be so rare that if one averages over the ensemble of velocity fields, they are essentially negligible. The motivation and proof draws significant inspiration from the work on anisotropic Sobolev spaces, especially [20].

1.1 Abstract framework and statement of negative regularity mixing

Here we outline a general abstract setting for our theorem that applies broadly to a whole class of random maps, including iid random diffeomorphisms, stochastic transport on compact manifolds (i.e. particle trajectories solve SDEs) [27, 15], advection by velocity fields generated by the stochastic Navier-Stokes equations, alternating random shear flows [12], and other similar settings (see Section 2). Our abstract framework is similar to that of [12], but more general in order to treat the cases studied in [8, 9, 7, 10]. See Section 2 for explanation on how to connect our abstract result to some of the concrete examples.

Let MM be a dd-dimensional, smooth, compact Riemannian manifold without boundary and (Ω,,𝐏)(\Omega,\mathscr{F},\mathbf{P}) be a probability space, θ:ΩΩ\theta:\Omega\to\Omega be a 𝐏\mathbf{P} preserving transformation and (un)(u_{n}) be a Feller Markov chain on a Polish space UU with a unique stationary (probability) measure μ\mu and an associated continuous random dynamical system Φωn:UU\Phi^{n}_{\omega}:U\to U. We denote the corresponding skew product flow on Z=Ω×UZ=\Omega\times U by τ(ω,u):=(θω,Φω(u))\tau(\omega,u):=(\theta\omega,\Phi_{\omega}(u)) with invariant ergodic product measure 𝔪=𝐏×μ\mathfrak{m}=\mathbf{P}\times\mu.

We will consider ϕω,u:MM\phi_{\omega,u}:M\to M a Ck0C^{k_{0}}-regular (with k02k_{0}\geq 2), random volume preserving diffeomorphism and denote

ϕω,un=ϕτn(ω,u)ϕτ(ω,u)ϕω,u\phi^{n}_{\omega,u}=\phi_{\tau^{n}(\omega,u)}\circ\ldots\circ\phi_{\tau(\omega,u)}\circ\phi_{\omega,u}

the nn-fold composition along the skew product flow τ\tau. We emphasize that we do not assume that each map ϕω,u\phi_{\omega,u} in this composition is iid, so the associated process xn=ϕω,un(x)x_{n}=\phi^{n}_{\omega,u}(x), xMx\in M is not assumed Markovian in general. However, the joint process (xn,un)=(ϕω,un(x),Φωn(u))(x_{n},u_{n})=(\phi^{n}_{\omega,u}(x),\Phi^{n}_{\omega}(u)) on M×ZM\times Z is Markovian.

Remark 1.1.

To include the case of iid random diffeomorphisms in our framework, one simply dispenses with (un)(u_{n}) and the relevant Markov process becomes xn=ϕωn(x)x_{n}=\phi^{n}_{\omega}(x) where

ϕωn=ϕθnωϕθωϕω.\displaystyle\phi^{n}_{\omega}=\phi_{\theta^{n}\omega}\circ\ldots\circ\phi_{\theta\omega}\circ\phi_{\omega}.

This includes the time-1 maps of well-posed, volume-preserving SDEs; see [34].

1.1.1 Lyapunov structure

First, we make some assumptions which provide control on how large the derivatives of ϕω,un\phi^{n}_{\omega,u} can be. In most settings, (un)(u_{n}) is the velocity field, and the deviations on ϕn\phi^{n} are stated in terms of the deviations of (un)(u_{n}). Recall that a function V:U[1,)V:U\to[1,\infty) is called a Lyapunov function for (un)(u_{n}) if VV has bounded sublevel sets {Vr}\{V\leq r\} and the associated Markov kernel for (un)(u_{n}), P(u,A)=𝐏u(u1A)P(u,A)=\mathbf{P}_{u}(u_{1}\in A) satisfies a Lyapunov-Foster Drift condition: there exists δ(0,1)\delta\in(0,1) and K0K\geq 0 such that

PVδV+K,PV\leq\delta V+K, (1.3)

where PV(u):=UV(u)P(u,du)PV(u):=\int_{U}V(u^{\prime})P(u,\mathrm{d}u^{\prime}).

In general, we will require a stronger Lyapunov structure on (un)(u_{n}) and the derivatives of ϕω,u\phi_{\omega,u}. Specifically, we will assume the existence of a two parameter family of Lyapunov functions Vβ,η:U[1,)V_{\beta,\eta}:U\to[1,\infty) defined for β0\beta\geq 0 and η(0,1)\eta\in(0,1), satisfying the following conditions:

Assumption 1 (Lyapunov Structure).

There exists a two parameter family of Lyapunov functions

{Vβ,η:β1,η(0,1)},\{V_{\beta,\eta}\,:\,\beta\geq 1,\eta\in(0,1)\},

satisfying for each a(0,1/η)a\in(0,1/\eta), Vβ,ηa(u)=Vβa,ηa(u)V_{\beta,\eta}^{a}(u)=V_{\beta a,\eta a}(u), such that the following condition holds: a>1\exists a_{*}>1 such that β1\forall\beta\geq 1, k1\forall k\in\mathbb{N}_{\geq 1} with kk0k\leq k_{0} and b0\forall b\geq 0 there exists a β(k,b)β\beta(k,b)\geq\beta satisfying β(1,b)=β\beta(1,b)=\beta such that η(0,1/a)\forall\eta\in(0,1/a_{*}),

𝐄u(𝒬k(ϕ)bVβ,η(u1))ab,k,a,β,ηVβ(k,b),η(u),\mathbf{E}_{u}\left(\mathcal{Q}_{k}(\phi)^{b}V_{\beta,\eta}(u_{1})\right)^{a_{*}}\lesssim_{b,k,a_{*},\beta,\eta}V_{\beta(k,b),\eta}(u), (1.4)

where 𝒬k(ϕ)\mathcal{Q}_{k}(\phi) is given in Definition A.5.

In what follows, we will omit the parameters β,η\beta,\eta in contexts where they are not important.

Remark 1.2.

Since we are working on a Riemannian manifold, the definition of 𝒬k(ϕ)\mathcal{Q}_{k}(\phi) is somewhat technical and is defined over a suitable atlas. However, intuitively 𝒬k(ϕ)\mathcal{Q}_{k}(\phi) is a measure of the size of the kkth derivative of ϕ\phi and ϕ1\phi^{-1}, heuristically

𝒬k(ϕ)``="sup|α|kDαϕL+Dαϕ1L.\mathcal{Q}_{k}(\phi)\,``="\,\sup_{|\alpha|\leq k}\|D^{\alpha}\phi\|_{L^{\infty}}+\|D^{\alpha}\phi^{-1}\|_{L^{\infty}}.

The assumption on 𝒬k(ϕ)\mathcal{Q}_{k}(\phi) in (1.4) is necessary for our pseudodifferential operator method, and is particularly important in bounding the errors produced in Egorov’s theorem (see Appendix A below). See Section 2 for discussions on the possible forms of Vβ,ηV_{\beta,\eta} and proofs of (1.4).

Remark 1.3.

Note, that upon setting b=0b=0 in (1.4), an application of Jensen’s inequality implies that V=Vβ,ηV=V_{\beta,\eta} satisfies the super-Lyapunov property, namely that for every ϵ(0,1)\epsilon\in(0,1) there is a Kϵ>0K_{\epsilon}>0 such that

PV(CV)1/aϵV+Kϵ.PV\leq(CV)^{1/a_{*}}\leq\epsilon V+K_{\epsilon}.

This is a much stronger condition than a standard Lyapunov-Foster condition (1.3). Nevertheless, it is satisfied for many semilinear, parabolic stochastic PDEs where the nonlinearity is sub-critical with respect to a conserved quantity (e.g. the 2d Stochastic Navier-Stokes equations).

1.1.2 The linearization and projective process

Next, we will need some assumptions that encode dynamical information. The first is the assumption on the spectral properties of the ‘projective’ Markov semigroup, described below. A prominent role here is played by the inverse transpose of the derivative of the flow map, which we denote by Aˇω,u,x\check{A}_{\omega,u,x}, defined by

Aˇω,u,x:=(Dxϕω,u),\check{A}_{\omega,u,x}:=(D_{x}\phi_{\omega,u})^{-\top},

where the transpose is taken with respect to the Riemannian metric on MM. At each xMx\in M, Aˇω,u,x\check{A}_{\omega,u,x} naturally acts on covectors ξTxM\xi\in T^{*}_{x}M with Aˇω,u,xξTϕω,u(x)M\check{A}_{\omega,u,x}\xi\in T^{*}_{\phi_{\omega,u}(x)}M and Aˇω,u,xn=(Dxϕω,un)\check{A}_{\omega,u,x}^{n}=(D_{x}\phi^{n}_{\omega,u})^{-\top} is a linear cocycle on TMT^{*}M since

Aˇω,u,xn=(Dxϕω,un)=Aˇτn(ω,u),ϕω,un(x)Aˇω,u,x.\check{A}^{n}_{\omega,u,x}=(D_{x}\phi^{n}_{\omega,u})^{-\top}=\check{A}_{\tau^{n}(\omega,u),\phi^{n}_{\omega,u}(x)}\circ\ldots\circ\check{A}_{\omega,u,x}.
Remark 1.4.

Under Assumption 1, by the multiplicative ergodic theorem [32, 39, 37, 45], the linear cocycle Aˇω,u,xn\check{A}^{n}_{\omega,u,x} has a Lyapunov spectrum λ1ˇλ2ˇλdˇ\check{\lambda_{1}}\geq\check{\lambda_{2}}\geq\ldots\check{\lambda_{d}} which have the property that jλˇj=0\sum_{j}\check{\lambda}_{j}=0 since det(Aˇω,u,x)=1\mathrm{det}(\check{A}_{\omega,u,x})=1. This can be related to the Lyapunov spectrum of the linear cocycle Aω,u,xn=Dxϕω,unA^{n}_{\omega,u,x}=D_{x}\phi^{n}_{\omega,u} on TMTM λ1λ2λd\lambda_{1}\geq\lambda_{2}\geq\ldots\geq\lambda_{d} via

λˇj=λdj+1,\check{\lambda}_{j}=-\lambda_{d-j+1},

(see e.g. [12] Lemma B1). Specifically, if λ1>0\lambda_{1}>0, by volume preservation, it must be that λˇ1>0\check{\lambda}_{1}>0.

In what follows, we consider (xn,ξn)(x_{n},\xi_{n}), with ξn=Aˇω,u,xnξ\xi_{n}=\check{A}^{n}_{\omega,u,x}\xi, the induced dynamics on the cotangent bundle TMT^{*}M. Let TPT^{*}\!P denote the Markov semigroup associated to the Markov chain (un,xn,ξn)(u_{n},x_{n},\xi_{n}) on U×TMU\times T^{*}M defined for each bounded measurable φ\varphi by

TPφ(u0,x0,ξ0):=𝐄φ(u1,x1,ξ1).T^{*}\!P\varphi(u_{0},x_{0},\xi_{0}):=\mathbf{E}\varphi(u_{1},x_{1},\xi_{1}).

Additionally denote (xn,vn)(x_{n},v_{n}), with

vn=Aˇω,u,xnv/|Aˇω,u,xnv|,v_{n}=\check{A}^{n}_{\omega,u,x}v/|\check{A}^{n}_{\omega,u,x}v|,

the “projectivized” dynamics on the unit cosphere bundle 𝕊M={vTM:|v|=1}\mathbb{S}^{*}M=\{v\in T^{*}M\,:\,|v|_{*}=1\} and denote P^\hat{P} the Markov semigroup associated to the Markov chain (un,xn,vn)(u_{n},x_{n},v_{n}) on U×𝕊MU\times\mathbb{S}^{*}M defined analogously.

Given a Lyapunov function V(u)V(u) for (un)(u_{n}), we will denote the space LipV(U×𝕊M)\mathrm{Lip}_{V}(U\times\mathbb{S}^{*}M) to be the space of continuous functions ψ(u,x,v)\psi(u,x,v) on U×𝕊MU\times\mathbb{S}^{*}M with finite weighted Lipschitz norm

ψLipV:=supzU×SM|ψ(z)|V(u)+supz,zU×SMψ(z)ψ(z)(V(u)+V(u))d(z,z)<,\|\psi\|_{\mathrm{Lip}_{V}}:=\sup_{z\in U\times S^{*}M}\frac{|\psi(z)|}{V(u)}+\sup_{z,z^{\prime}\in U\times S^{*}M}\frac{\|\psi(z)-\psi(z)\|}{(V(u)+V(u^{\prime}))d(z,z^{\prime})}<\infty, (1.5)

where for z=(u,x,v)z=(u,x,v) and z=(u,x,v)z^{\prime}=(u^{\prime},x^{\prime},v^{\prime}) d(z,z)=d(u,u)+dSM((x,v),(x,v))d(z,z^{\prime})=d(u,u^{\prime})+d_{S^{*}M}\left((x,v),(x^{\prime},v^{\prime})\right), with dd being the metric on UU and dSMd_{S^{*}M} the metric on SMS^{*}M. It follows by the Feller property of (un)(u_{n}) and assumption (1.4) that P^\hat{P} restricts to a bounded linear operator on the space LipV(U×SM)\mathrm{Lip}_{V}(U\times S^{*}M), at least if one replaces the time-step 11 with N0N_{0} and considers the process (u~n,v~n)=(unN0,vnN0)(\tilde{u}_{n},\tilde{v}_{n})=(u_{nN_{0}},v_{nN_{0}}) (see [[9] Lemma 5.2] for more information). By relabeling, we can assume without loss of generality that N0=1N_{0}=1.

Of particular interest are functions which are p-p homogeneous, i.e.

φ(u,x,ξ)=|ξ|pψ(u,x,ξ/|ξ|),\varphi(u,x,\xi)=|\xi|^{-p}\psi(u,x,\xi/|\xi|),

where ψ(u,)\psi(u,\cdot) is a smooth function on SMS^{*}M which is homogeneous of degree 0 in ξ\xi. Evaluating TPT^{*}P on such a symbol, we obtain

TPφ(u,x,ξ)=P^pψ(u,x,ξ/|ξ|),T^{*}P\varphi(u,x,\xi)=\hat{P}^{p}\psi(u,x,\xi/|\xi|),

where P^p\hat{P}^{p} is the ‘twisted’ semigroup defined by

P^pψ(u,x,v)=𝐄u,x,v|Aˇω,u,xv|pψ(u1,ϕω,u1(x),v1).\hat{P}^{p}\psi(u,x,v)=\mathbf{E}_{u,x,v}|\check{A}_{\omega,u,x}v|^{-p}\psi(u_{1},\phi^{1}_{\omega,u}(x),v_{1}).

Similar to P^\hat{P}, P^p\hat{P}^{p} forms a semigroup of bounded operators on Lip(U×SM)\mathrm{Lip}(U\times S^{*}M) (again, possibly by increasing the time-step and relabeling) and that when p=0p=0 one recovers the projective semigroup P^\hat{P}. We make the following spectral assumption on P^p\hat{P}^{p}.

Assumption 2 (Spectral Gap).
  1. (i)

    There exists a p0>0p_{0}>0 such that for all p[p0,p0]p\in[-p_{0},p_{0}], P^p\hat{P}^{p} admits a simple dominant eigenvalue eΛ(p)e^{-\Lambda(p)} on LipV(U×SM)\mathrm{Lip}_{V}(U\times S^{*}M) with a spectral gap, namely there exists an r(0,eΛ(p))r\in(0,e^{-\Lambda(p)}) such that the spectrum of P^p\hat{P}^{p} satisfies

    σ(P^p)\{eΛ(p)}B(0,r),\sigma(\hat{P}^{p})\backslash\{e^{-\Lambda(p)}\}\subseteq B(0,r),

    where B(0,r)B(0,r)\subset\mathbb{C} denotes the open ball of radius rr centered at 0.

  2. (ii)

    The eigenvalue satisfies Λ(p)>0\Lambda(p)>0 for p(0,p0)p\in(0,p_{0}).

  3. (iii)

    Let πp\pi_{p} be the rank-one spectral projector associated with eΛ(p)e^{-\Lambda(p)}. Then ψpπp𝟏\psi_{p}\equiv\pi_{p}\bm{1} is a dominant eigenfunction satisfying

    P^pψp=eΛ(p)ψp,\hat{P}^{p}\psi_{p}=e^{-\Lambda(p)}\psi_{p},

    and is bounded below by a positive constant on bounded sets. That is for each bounded set BUB\subset U there is a cB>0c_{B}>0 such that infuBψp>cB\inf_{u\in B}\psi_{p}>c_{B}.

Remark 1.5.

Assumption (iii) is essentially the assumption that (un,xn,vn)(u_{n},x_{n},v_{n}) is irreducible; see [12, 9] for discussion.

Remark 1.6.

Irreducibility and geometric ergodicity of the projective process (un,xn,vn)(u_{n},x_{n},v_{n}) in a suitable Wasserstein metric, usually proved via a weak Harris’ theorem (see [29]), implies Part (i) and (iii) with a straightforward spectral perturbation argument (see e.g. [12, 7] and the references therein).

Note that also, in particular, by setting p=0p=0 Assumption 2 also implies geometric ergodicity (again in a suitable metric) of P^\hat{P} and hence a unique stationary probability measure ν𝒫(U×SM)\nu\in\mathcal{P}(U\times S^{*}M).

Remark 1.7.

The quantity Λ(p)\Lambda(p) is called the moment Lyapunov function and, at least for |p|\left|p\right| close to 0, Part (iii) implies that for 𝐏×ν\mathbf{P}\times\nu-almost every (ω,u,x,v)(\omega,u,x,v),

Λ(p)=limn1nlog𝐄|Aˇω,u,xnv|p.\displaystyle\Lambda(p)=-\lim_{n\to\infty}\frac{1}{n}\log\mathbf{E}\left|\check{A}^{n}_{\omega,u,x}v\right|^{-p}.

The function Λ(p)\Lambda(p) is also closely related to the probability of finding |Aˇω,u,xnv|\left|\check{A}^{n}_{\omega,u,x}v\right| far from the expected eλˇ1ne^{\check{\lambda}_{1}n} for large nn. Particularly, it is related to a large deviation principle for the projective chain (un,xn,vn)(u_{n},x_{n},v_{n}). See e.g. [6, 5] for more discussions.

One can readily prove that Λ(0)=λˇ1\Lambda^{\prime}(0)=\check{\lambda}_{1} (see [Lemma 5.10 [9]] for the proof when the cocyle is Aω,u,xnA^{n}_{\omega,u,x} instead of Aˇω,u,xn\check{A}^{n}_{\omega,u,x} ). Hence, positivity of the Lyapunov exponent λˇ1>0\check{\lambda}_{1}>0 implies that for p1p\ll 1

Λ(p)=pλˇ1+o(p)>0,\Lambda(p)=p\check{\lambda}_{1}+o(p)>0,

which would then imply Part (ii) of Assumption 2.

Given the connection between TPT^{*}P and P^p\hat{P}^{p}, Assumption 2 implies that the function

ap(u,x,ξ):=|ξ|pψp(u,x,ξ/|ξ|)a_{p}(u,x,\xi):=|\xi|^{-p}\psi_{p}(u,x,\xi/|\xi|)

is an eigenfunction of TPT^{*}\!P with eigenvalue eΛ(p)e^{-\Lambda(p)}

TPap=eΛ(p)ap.T^{*}\!Pa_{p}=e^{-\Lambda(p)}a_{p}.

The function apa_{p} is the starting point of the symbol of a pseudo-differential operator that plays a key role in our proof below.

1.1.3 Two point geometric ergodicity

Closely related to the behavior of the twisted Markov semi-group P^p\hat{P}^{p} and the associated eigenfunctions ψp\psi_{p}, is the behavior of the two point motion. Namely, let xnx_{n} and yny_{n} be two different trajectories on M×MM\times M starting from distinct points x0x_{0} and y0y_{0}, that is xn=ϕω,un(x0)x_{n}=\phi^{n}_{\omega,u}(x_{0}) and yn=ϕω,un(y0)y_{n}=\phi^{n}_{\omega,u}(y_{0}). In order to avoid reducing to the one point motion, we assume that the starting points (x0,y0)(x_{0},y_{0}) do not belong to the diagonal set

Δ={(x,x):xM}M×M.\Delta=\{(x,x)\,:\,x\in M\}\subset M\times M.

We denote the associated Markov chain (un,xn,yn)(u_{n},x_{n},y_{n}) on U×ΔcU\times\Delta^{c} by P(2)P^{(2)}. We remark that U×ΔcU\times\Delta^{c} is an almost-surely invariant set for the two-point semigroup defined over U×M×MU\times M\times M. As the space Δc\Delta^{c} is not compact, one is likely to use a Lyapunov function 𝒱\mathcal{V} on U×ΔcU\times\Delta^{c} to control the behavior of the two point motion. As in [9, 7, 12] we consider the Lyapunov function 𝒱p\mathcal{V}_{p} with the following properties:

𝒱p(u,x,y)(d(x,y)p1)Vβ,η(u)\displaystyle\mathcal{V}_{p}(u,x,y)\approx(d(x,y)^{-p}\vee 1)V_{\beta,\eta}(u) (1.6)

for all (u,x,y)U×Δc(u,x,y)\in U\times\Delta^{c}. Depending on the setting, this kind of Lyapunov function can be constructed using ψp\psi_{p} [9, 7, 12] or using large deviation estimates on the exit times [17, 6]. We will need the following assumption on the exponential decay of the two point motion.

Assumption 3 (Exponential Decay of Two Point Motion).

We assume that for all 0<p10<p\ll 1 sufficiently small, for all β1\beta\geq 1, and all η(0,1)\eta\in(0,1), 𝒱p\exists\mathcal{V}_{p} satisfying (1.6) which is a Lyapunov function for the two point motion (un,xn,yn)(u_{n},x_{n},y_{n}) and for which the Markov process is 𝒱p\mathcal{V}_{p}-geometrically ergodic, namely there exists a γ>0\gamma>0 such that for all x0,y0Mx_{0},y_{0}\in M and all φC𝒱p(U×Δc)\varphi\in C_{\mathcal{V}_{p}}(U\times\Delta^{c}), there holds

|P(2)φ(u,x,y)UMMφdμdxdy|𝒱p(u,x,y)φC𝒱peγt.\displaystyle\left|P^{(2)}\varphi(u,x,y)-\int_{U}\int_{M}\int_{M}\varphi\,\mathrm{d}\mu\mathrm{d}x\mathrm{d}y\right|\lesssim\mathcal{V}_{p}(u,x,y)\left|\left|\varphi\right|\right|_{C_{\mathcal{V}_{p}}}e^{-\gamma t}.

Heuristically, we can consider Assumption 3 as a strict improvement over Assumption 2 above, extending the linearized information to the nonlinear two-point dynamics. Indeed, the proofs of [7, 9, 12] used Assumption 2 together with a Harris’ theorem argument using the Lyapunov function 𝒱p\mathcal{V}_{p} to prove Assumption 3. However, it is easier to state these as separate assumptions, rather than list all of the individual assumptions required for the Harris’ theorem argument to apply.

By a now-standard Borel-Cantelli argument together with a little additional Fourier analysis (see [17] and [9, 14, 12]), Assumption 3 implies quenched mixing estimates such as (1.2). We technically do not use such estimates directly, however, the manner in which we employ Assumption 3 is nevertheless quite similar to a quenched mixing estimate.

1.1.4 Main result

In what follows, denote fn=f0(ϕω,un)1f_{n}=f_{0}\circ(\phi^{n}_{\omega,u})^{-1}. Our main result is the following theorem.

Theorem 1.8.

Under Assumptions 13, there exists constants μ>0\mu>0, δ(0,1)\delta\in(0,1) and β1\beta_{*}\geq 1, such for each Lyapunov function V=Vβ,ηV=V_{\beta_{*},\eta}, with η>0\eta>0 sufficiently small there holds the following: for all distributions f0Hδf_{0}\in H^{-\delta} with f0dx=0\int f_{0}\mathrm{d}x=0, for all n0n\geq 0, u0Uu_{0}\in U

𝐄[V(un)1fnHδ2]V(u0)e2μnf0Hδ2,\mathbf{E}\left[V(u_{n})^{-1}\|f_{n}\|_{H^{-\delta}}^{2}\right]\lesssim V(u_{0})e^{-2\mu n}\|f_{0}\|_{H^{-\delta}}^{2}, (1.7)

As a corollary, there holds q(0,2)\forall q\in(0,2)

𝐄fnHδqqVq(u0)eqμnf0Hδq.\displaystyle\mathbf{E}\|f_{n}\|_{H^{-\delta}}^{q}\lesssim_{q}V^{q}(u_{0})e^{-q\mu n}\|f_{0}\|_{H^{-\delta}}^{q}. (1.8)
Remark 1.9.

We do not currently know if the result holds for q>2q>2, nor do we know how large one can make δ\delta; see Section 1.2 for more discussion.

Remark 1.10.

In the case that ϕω,u1=ϕω1\phi^{1}_{\omega,u}=\phi^{1}_{\omega} are iid diffeomorphisms (as would happen with the stochastic flow of diffeomorphisms generated by SDEs with smooth vector fields [34]) we can dispense with VV and hence (1.8) holds for ϵ=0\epsilon=0 and simply becomes

𝐄fnHδ2e2μnf0Hδ2.\mathbf{E}\|f_{n}\|_{H^{-\delta}}^{2}\lesssim e^{-2\mu n}\|f_{0}\|_{H^{-\delta}}^{2}. (1.9)
Remark 1.11.

For distributions that cannot be identified with an L1L^{1} function, we can define f0dx\int f_{0}\mathrm{d}x as the distribution acting on the smooth function ϕ1\phi\equiv 1.

Remark 1.12.

A related result was proven in [15], in the setting of the Kraichnan model, where the authors prove an exact identity for the averaged exponential decay of negative Sobolev norms. However, the proof in [15] relies on the specific structure of the Kraichnan model and does not extend to the general setting considered here. Additionally, as the self similarity required in [15] makes it challenging to define an associated flow map, their results do no immediately follow from Theorem 1.8.

As a corollary of Theorem 1.8, we obtain the following quenched mixing estimate with a random constant depending on the initial data.

Corollary 1.13 (Quenched Result).

Let μ\mu and δ\delta and β\beta_{*} be as in Theorem 1.8. Then, for all f0Hδf_{0}\in H^{-\delta} with f0dx=0\int f_{0}\mathrm{d}x=0, and V=Vβ,ηV=V_{\beta_{*},\eta} with η\eta suitably small, there exists a random constant K(f0,u0)K(f_{0},u_{0}) (depending on u0u_{0} and the initial f0f_{0}) such that there holds for all n0n\geq 0 and u0Uu_{0}\in U,

fnHδK(u0,f0)eμn/2V(u0)f0Hδ.\|f_{n}\|_{H^{-\delta}}\lesssim K(u_{0},f_{0})e^{-\mu n/2}V(u_{0})\|f_{0}\|_{H^{-\delta}}.

Moreover, KK has uniform (in f0f_{0} and u0u_{0}) moments 𝐄Kqq1\mathbf{E}K^{q}\lesssim_{q}1, q(0,2)\forall q\in(0,2).

Proof.

Clearly we can write

fnHδK(u0,f0)eμn/2V(u0)f0Hδ\|f_{n}\|_{H^{-\delta}}\lesssim K(u_{0},f_{0})e^{-\mu n/2}V(u_{0})\|f_{0}\|_{H^{-\delta}}

where KK is defined by

K(u0,f0)=maxn0eμn/2fnHδV(u0)f0Hδ.K(u_{0},f_{0})=\max_{n\geq 0}\frac{e^{\mu n/2}\|f_{n}\|_{H^{-\delta}}}{V(u_{0})\|f_{0}\|_{H^{-\delta}}}.

To see that KK has finite moments we estimate

𝐄Kqn0eμnq/2𝐄fnHδqV(u0)qf0Hδqqn0eqμn/2q1.\mathbf{E}K^{q}\leq\sum_{n\geq 0}\frac{e^{\mu nq/2}\mathbf{E}\|f_{n}\|_{H^{-\delta}}^{q}}{V(u_{0})^{q}\|f_{0}\|_{H^{-\delta}}^{q}}\lesssim_{q}\sum_{n\geq 0}e^{-q\mu n/2}\lesssim_{q}1.

1.2 Comments on the proof of Theorem 1.8

The proof of Theorem 1.8 is inspired primarily by the pioneering work of Faure, Roy and Sjostrand [20] and the subsequent works [24, 21, 25, 26, 22, 23]. In these works the authors use microlocal analysis to construct special anisotropic Sobolev spaces (i.e. spaces with different regularity in different directions in frequency space) on which one can prove spectral a gap, or at least quasi-compactness, for the transfer operator. This microlocal approach as also been used in the context of Anosov flows generated by a vectorfield [19, 18]. See also [11, 28, 35, 1, 3, 16, 2] for other works using anisotropic Banach spaces of functions to study spectra of the transfer operator.

The motivation of these anisotropic spaces is to find a space XX with

HαXHα,\displaystyle H^{-\alpha}\subset X\subset H^{\alpha},

such that for average-zero observables, the scalars decay exponentially in XX due to a spectral gap of the transfer operator 𝒯f:=fϕ\mathcal{T}f:=f\circ\phi (or to at least prove a quasi-compactness estimate that implies localization of the essential spectrum)

fϕnXeμnfX.\displaystyle\left|\left|f\circ\phi^{n}\right|\right|_{X}\lesssim e^{-\mu n}\left|\left|f\right|\right|_{X}. (1.10)

In settings where this is possible, it yields a significantly more precise result than quenched mixing estimates. In the microlocal approach of [20] (also in [18, 19]), these spaces are found by building a Lyapunov function for the linearized process on TMT^{\ast}M, a(x,ξ)a(x,\xi), and using it to define a pseudo-differential operator with a suitable quantization procedure. The space one obtains the essential spectrum estimates in is then defined via the norm (at least heuristically)

fX=Op(a)fL2,\displaystyle\left|\left|f\right|\right|_{X}=\|\mathrm{Op}(a)f\|_{L^{2}},

where Op(a)\mathrm{Op}(a) denotes the pseudo-differential operator associated to the symbol aa. In the case of Anosov diffeomorphisms, one chooses a(x,ξ)ξma(x,\xi)\approx\left\langle\xi\right\rangle^{m} for ξ\xi on the tangent to the unstable manifold through xx and ξm\approx\left\langle\xi\right\rangle^{-m} for ξ\xi on the tangent to the stable manifold through xx (the angle between the tangents is uniformly bounded away from zero by assumption of uniform hyperbolicity). After a suitable regularization procedure and a variable-order Egorov’s theorem, one can use this symbol to construct a norm XX and prove a Lasota-Yorke-type estimate which implies quasi-compactness of the transfer operator (see [20] for more details).

Again considering the case of Anosov diffeomorphisms, one can observe that a negative regularity mixing estimate such as Theorem 1.8 cannot possibly hold for deterministic maps. Indeed, by concentrating the initial distribution in HδH^{-\delta} close to the stable manifolds, one can obtain arbitrarily slow mixing rates for certain sequence of pathological initial data. For the random map case, the difference here is that for each fixed initial distribution, the assumptions in Section 1.1.2 basically rule out the possibility that a positive probability set of ωΩ\omega\in\Omega results in stable manifolds that line up badly to produce poor decay rates. One can wonder exactly how large of a moment one can take in Theorem 1.8, i.e. under what conditions the set of bad ω\omega remains negligible; our proof currently yields at most second moments. Similarly, one can wonder how large δ\delta can be taken, i.e. whether one can take δ\delta all the way down to δ>(d1)/2\delta>(d-1)/2 or even δ>d/2\delta>d/2 (going further is clearly impossible as point masses cannot be mixed).

In order to prove Theorem 1.8, we use a Lyapunov function (now again in the stochastic process sense) for the linearized process on U×TMU\times T^{\ast}M defined in Section 1.1.2 and use this as a symbol aa to obtain exponential decay estimates on the quantity

Op(a)f,f.\displaystyle\left\langle\mathrm{Op}(a)f,f\right\rangle.

We use a regularization of the symbol

a(u,x,ξ)=1|ξ|pψp(u,x,ξ),\displaystyle a(u,x,\xi)=\frac{1}{\left|\xi\right|^{p}}\psi_{p}(u,x,\xi),

which by Assumption 2, satisfies a spectral-gap type estimate for the linearized process Markov semigroup. Formally, we could then expect by Egorov’s theorem

𝐄Op(a)fϕn,(fϕn)eΛ(p)nf,Op(a)f.\displaystyle\mathbf{E}\left\langle\mathrm{Op}(a)f\circ\phi^{n},(f\circ\phi^{n})\right\rangle\approx e^{-\Lambda(p)n}\left\langle f,\mathrm{Op}(a)f\right\rangle.

Given the lower bounds on ψp\psi_{p} in Part (iii) of Assumption 2, by Gårding’s inequality, we can heuristically expect that the quantization of this symbol would satisfy something like (ignoring VV for a moment)

Op(a)f,ffHp/22,\displaystyle\left\langle\mathrm{Op}(a)f,f\right\rangle\lesssim\left|\left|f\right|\right|_{H^{-p/2}}^{2},

which would suggest Theorem 1.8. The most obvious way this intuition fails is that both Egorov’s theorem and Gårding’s inequality only control high frequencies and both leave an error in lower regularity, for example the most one can hope for from Gårding’s inequality is something like the following for some constants ζ,C>0\zeta,C>0,

ζfHp/22CfHd+322f,Op(a)f.\displaystyle\zeta\left|\left|f\right|\right|_{H^{-p/2}}^{2}-C\left|\left|f\right|\right|_{H^{-\frac{d+3}{2}}}^{2}\leq\left\langle f,\mathrm{Op}(a)f\right\rangle. (1.11)

This is analogous to why one only directly obtains Lasota-Yorke type estimates in the deterministic setting, rather than direct exponential decay estimates. These errors will need to be dealt with carefully and in particular, require us to first prove a Hp/2Hd+32H^{-p/2}\to H^{-\frac{d+3}{2}}-type mixing estimate (this is done using Assumption 3; see Section 4.2). Note that even this kind of estimate is false for deterministic maps. However, there are two more reasons the above intuition is naïve: (A) the regularity of ψp\psi_{p} is limited to C1C^{1} and so the symbol must be suitably regularized and, perhaps most importantly, (B) the positivity and regularity of the symbol depend badly on the derivatives of ϕu0,ωt\phi^{t}_{u_{0},\omega} which are unbounded (measured by the surrogate uu), which means that the ζ,C\zeta,C in the Gårding’s inequality (1.11) and the errors in Egorov’s theorem would become time-dependent (or the regularization would need to be time-dependent). These issues present the main difficulties in proving Theorem 1.8. In Section 3, the regularization procedure of the symbol is presented and the basic properties are verified. In Section 4 the main arguments in Theorem 1.8 are given, namely an Hp/2Hd+32H^{-p/2}\to H^{-\frac{d+3}{2}}-type mixing estimate and its use, together with Assumption 1, to absorb the low frequency error terms coming from Egorov’s theorem and Gårding’s inequality to yield Theorem 1.8.

2 Applications

Here we outline several relevant applications of our general framework to examples of interest, most notably stochastic flow for SDEs and flows generated by the stochastic Navier-Stokes equations. Additionally, we state some important applications to the advection diffusion equation with a stochastic source term and the existence of a unique stationary measure in the zero-diffusivity limit.

2.1 Examples

2.1.1 IID diffeomorphisms and stochastic flows

A simple, but wide, class of examples our theorem applies to are iid random diffeomorphisms, such as the case of the stochastic flow of diffeomorphisms associated to an SDE on a compact, Riemannian manifold (,g)(\mathcal{M},g). Consider smooth divergence-free vector fields X0,X1,,XrX_{0},X_{1},...,X_{r} on (,g)(\mathcal{M},g), then the SDE

dxt=X0(xt)dt+j=1rXj(xt)dWt,\displaystyle\mathrm{d}x_{t}=X_{0}(x_{t})\mathrm{d}t+\sum_{j=1}^{r}X_{j}(x_{t})\circ\mathrm{d}W_{t},

defines a stochastic flow of diffeomorphisms xt=ϕωt(x0)x_{t}=\phi^{t}_{\omega}(x_{0}) (see [34] for details). As the vector fields are divergence free, ϕωt\phi^{t}_{\omega} is volume-preserving almost-surely, i.e. the Riemannian volume measure is almost-surely invariant. The projective process solves a similar SDE on the sphere bundle 𝕊\mathbb{S}\mathcal{M} [4] (i.e. the unit tangent bundle), denoted

dzt=X~0(zt)dt+j=1rX~j(zt)dWt,\displaystyle\mathrm{d}z_{t}=\tilde{X}_{0}(z_{t})\mathrm{d}t+\sum_{j=1}^{r}\tilde{X}_{j}(z_{t})\circ\mathrm{d}W_{t},

with z=(x,v)z=(x,v) and the ‘lifted’ vector fields satisfy

X~=(V(x)(Ivv)X(x)v).\displaystyle\tilde{X}=\begin{pmatrix}V(x)\\ (I-v\otimes v)\nabla X(x)v\end{pmatrix}.

As we explain in more detail below in Section 2.2, [17] provides checkable sufficient conditions for these flows to satisfy Assumption 23, at least when combined with the construction of ψp\psi_{p} found in [9]; see also the earlier work of [6, 13, 4]. In particular, a quenched mixing estimate (1.2) was proved in [17], specifically for any α(0,1)\alpha\in(0,1), γ>0\exists\gamma>0 (deterministic)

fϕωnC0,αD(ω)eγnfC0,α,\displaystyle\left|\left|f\circ\phi^{n}_{\omega}\right|\right|_{C^{0,-\alpha}}\lesssim D(\omega)e^{-\gamma n}\left|\left|f\right|\right|_{C^{0,\alpha}}, (2.1)

with 𝐄D2<\mathbf{E}D^{2}<\infty. Our results now additionally prove that for 0<p10<p\ll 1 (the lack of VV implies we can take ϵ=0\epsilon=0),

𝐄fϕnHp2eμnfHp2.\displaystyle\mathbf{E}\left|\left|f\circ\phi^{n}\right|\right|_{H^{-p}}^{2}\lesssim e^{-\mu n}\left|\left|f\right|\right|_{H^{-p}}^{2}.

Another concrete case of iid random diffeomorphisms are the Pierrehumbert flows [12], namely the case of transport by alternating shear flows on 𝕋d\mathbb{T}^{d}. For example, in 2D, the time-one map random map ϕω:𝕋d𝕋d\phi_{\omega}:\mathbb{T}^{d}\to\mathbb{T}^{d} could be given by ϕω((x,y))=(x,y)\phi_{\omega}((x,y))=(x_{\ast},y_{\ast}), where

x\displaystyle x_{\ast} =x+A(ω)sin(y+γ(ω))\displaystyle=x+A(\omega)\sin(y+\gamma(\omega))
y\displaystyle y_{\ast} =y+A(ω)sin(x+γ(ω)).\displaystyle=y+A^{\prime}(\omega)\sin(x_{\ast}+\gamma^{\prime}(\omega)).

where A,A,γ,γA,A^{\prime},\gamma,\gamma^{\prime} are suitable independent random variables, for example each drawn uniformly from (π,π)(-\pi,\pi) suffices. Assumptions 23 were proved in [12], wherein a quenched exponential mixing estimate such as (2.1) was proved and a suitable ψp\psi_{p} was constructed. Theorem 1.8 now provides also the estimate (1.9).

2.1.2 Stochastic Navier-Stokes equations

Let us briefly explain how to apply Theorem 1.8 to the setting of stochastic Navier-Stokes in 𝕋2\mathbb{T}^{2}, as in works of [9, 7] and [14]. The PDE in question is given by the following in 𝕋2\mathbb{T}^{2}

{dut+(utut+ptνΔut)dt+QdWtdivut=0;\begin{dcases}\,\mathrm{d}u_{t}+\left(u_{t}\cdot\nabla u_{t}+\nabla p_{t}-\nu\Delta u_{t}\right)\mathrm{d}t+Q\mathrm{d}W_{t}\\ \,\operatorname{\mathrm{div}}u_{t}=0\,;\end{dcases} (2.2)

in 3D the viscosity must be replaced by a suitable hyperviscosity νΔν(Δ)2-\nu\Delta\mapsto\nu(-\Delta)^{2} but this case otherwise also works. Let us now explain the operator QWtQW_{t}. Following the convention used in [46], we define the following real Fourier basis for functions on 𝕋d\mathbb{T}^{d} by

ek(x)={sin(kx),k+dcos(kx),kd,e_{k}(x)=\begin{cases}\sin(k\cdot x),\quad&k\in\mathbb{Z}^{d}_{+}\\ \cos(k\cdot x),\quad&k\in\mathbb{Z}^{d}_{-},\end{cases}

where +d={(k1,k2,kd)d:kd>0}{(k1,k2,kd)d:k1>0,kd=0}\mathbb{Z}_{+}^{d}=\{(k_{1},k_{2},\ldots k_{d})\in\mathbb{Z}^{d}:k_{d}>0\}\cup\{(k_{1},k_{2},\ldots k_{d})\in\mathbb{Z}^{d}\,:\,k_{1}>0,k_{d}=0\} and d=+d\mathbb{Z}_{-}^{d}=-\mathbb{Z}_{+}^{d}. We set 0d:=d{0,,0}\mathbb{Z}^{d}_{0}:=\mathbb{Z}^{d}\setminus\left\{0,\ldots,0\right\} and define {γk}k0d\{\gamma_{k}\}_{k\in\mathbb{Z}^{d}_{0}} a collection of full rank d×(d1)d\times(d-1) matrices satisfying γkk=0\gamma^{\top}_{k}k=0, γkγk=Id\gamma_{k}^{\top}\gamma_{k}=\operatorname{\mathrm{Id}}, and γk=γk\gamma_{-k}=-\gamma_{k}. Note that in dimension d=2d=2, γk\gamma_{k} is just a vector in 2\mathbb{R}^{2} and is therefore given by γk=±k/|k|\gamma_{k}=\pm k^{\perp}/|k|. In dimension three, the matrix γk\gamma_{k} defines a pair of orthogonal vectors γk1,γk2\gamma_{k}^{1},\gamma_{k}^{2} that span the space perpendicular to kk. Next, we define the natural Hilbert space on velocity fields u:𝕋ddu:\mathbb{T}^{d}\to\mathbb{R}^{d} by

𝐋2:={uL2(𝕋d;d):udx=0,divu=0},\displaystyle{\bf L}^{2}:=\left\{u\in L^{2}(\mathbb{T}^{d};\mathbb{R}^{d})\,:\,\int u\,\mathrm{d}x=0,\quad\operatorname{\mathrm{div}}u=0\right\}, (2.3)

with the natural L2L^{2} inner product. Let WtW_{t} be a cylindrical Wiener process on 𝐋2{\bf L}^{2} with respect to an associated canonical stochastic basis (Ω,,(t),𝐏)(\Omega,\mathscr{F},(\mathscr{F}_{t}),\mathbf{P}) and QQ a Hilbert-Schmidt operator on 𝐋2{\bf L}^{2}, diagonalizable with respect the Fourier basis on 𝐋2{\bf L}^{2}. In the works [7, 9], the operator QQ was assumed to satisfy the following regularity and non-degeneracy assumption

Assumption 4 (Assumptions in [7, 9]).

There exists α\alpha satisfying α>5d2\alpha>\frac{5d}{2} and a constant CC such that

1C(Δ)α/2u𝐋2Qu𝐋2C(Δ)α/2u𝐋2.\frac{1}{C}\|(-\Delta)^{-\alpha/2}u\|_{{\bf L}^{2}}\leq\|Qu\|_{{\bf L}^{2}}\leq C\|(-\Delta)^{-\alpha/2}u\|_{{\bf L}^{2}}.

In the work [14] the lower bound assumption was dropped and replaced with the following much weaker assumption.

Assumption 5 ((Essentially) Assumptions in [14]).

The QQ is compactly supported in frequency and satisfies the Hörmander hypoellipticity condition given in [29, 30] in 2D and [40] in 3D. We will additionally assume that all of the modes with |k|1\left|k\right|_{\ell^{\infty}}\leq 1 are forced.

Theorem 1.8 can be applied to both of these cases. We define our primary phase space of interest to be velocity fields with sufficient Sobolev regularity (under Assumption 5 we can choose σ>d2+3\sigma>\frac{d}{2}+3 arbitrarily):

𝐇:={uHσ(𝕋d,d):udx=0,divu=0},whereσ(α2(d1),αd2).{\bf H}:=\left\{u\in H^{\sigma}(\mathbb{T}^{d},\mathbb{R}^{d})\,:\,\int u\,\mathrm{d}x=0,\quad\operatorname{\mathrm{div}}u=0\right\},\quad\text{where}\quad\,\sigma\in(\alpha-2(d-1),\alpha-\tfrac{d}{2}).

Note we have chosen α\alpha sufficiently large to ensure that σ>d2+3\sigma>\frac{d}{2}+3 so that 𝐇C3{\bf H}\hookrightarrow C^{3}. Since we will need to take advantage of the “energy estimates” produced by the vorticity structure of the Navier-Stokes equations in 2D2D, we find it notationally convenient to define the following dimension dependent norm

u𝐖:={curlu𝐋2d=2u𝐋2d=3.\|u\|_{{\bf W}}:=\begin{cases}\|\operatorname{\mathrm{curl}}u\|_{{\bf L}^{2}}&d=2\\ \|u\|_{{\bf L}^{2}}&d=3.\end{cases} (2.4)

The following well-posedness theorem is classical (see e.g. [33]).

Proposition 2.1.

For all initial data u𝐇u\in{\bf H}, there exists a 𝐏\mathbf{P}-a.s. unique, global-in-time, t\mathscr{F}_{t}-adapted mild solution (ut)(u_{t}) to (2.2) satisfying u0=uu_{0}=u. Moreover, (ut)(u_{t}) defines a Feller Markov process on 𝐇{\bf H} and the corresponding Markov semigroup has a unique stationary probability measure μ\mu on 𝐇{\bf H}.

One then defines the Lagrangian flow map ϕu,ωt:𝕋d𝕋d\phi^{t}_{u,\omega}:\mathbb{T}^{d}\to\mathbb{T}^{d} as the solution map to the ODE

tϕu,ωt=utϕu,ωt,ϕu,ω0=Id.\partial_{t}\phi^{t}_{u,\omega}=u_{t}\circ\phi^{t}_{u,\omega},\quad\phi^{0}_{u,\omega}=\operatorname{\mathrm{Id}}. (2.5)

2.2 Checking the conditions for stochastic Navier-Stokes

In this section we sketch some ideas required to verify the conditions of Theorem 1.8 in the setting of stochastic Navier-Stokes. In the setting of stochastic flows, checking the conditions follow along similar lines and typically much easier to verify (see e.g [27]).

Assumption 1 will be show below in Section 2.2.1. Assumption 3 was proved for Assumption 4 in [9] and for Assumption 5 in [14]. Assumption 2 (i) – (iii) regarding the construction of ψp\psi_{p} was proved directly in [9] for the case of Assumption 4. In 2.2.2 below, we briefly explain how to obtain Assumption 2 (i) – (iii) for the degenerate noise Assumption 5.

2.2.1 Checking Assumption 1

We must check Assumption 1 for the Lyapunov function Vβ,ηV_{\beta,\eta} given by

Vβ,η(u)=u𝐇2βeηηu𝐖2,\displaystyle V_{\beta,\eta}(u)=\langle\left|\left|u\right|\right|_{{\bf H}}\rangle^{2\beta}e^{\eta\eta_{*}\left|\left|u\right|\right|_{{\bf W}}^{2}},

where η\eta_{*} is a This will be proved by making use of the following super Lypaunov bound proved in [9] (the proof works for both Assumption 4 and 5).

Lemma 2.2 ( [9] Lemma 3.7).

Let (ut)(u_{t}) be a solution to 2.2. There exists a γ>0\gamma_{*}>0, such that for all 0γ<γ0\leq\gamma<\gamma_{*}, T>0T>0, r(0,3)r\in(0,3), κ0\kappa\geq 0, and V(u)=Vβ,ηV(u)=V_{\beta,\eta} where β1\beta\geq 1 and 0<eγTη<10<e^{\gamma T}\eta<1, there exists a constant C=C(γ,T,r,κ,β,η)>0C=C(\gamma,T,r,\kappa,\beta,\eta)>0 such that the following estimate holds

𝐄uexp(κ0Tus𝐇rds)sup0tTVeγt(ut)CV(u).\mathbf{E}_{u}\exp\left(\kappa\int_{0}^{T}\left|\left|u_{s}\right|\right|_{{\bf H}^{r}}\mathrm{d}s\right)\sup_{0\leq t\leq T}V^{e^{\gamma t}}(u_{t})\leq CV(u). (2.6)

Assumption 1 reduces to proving the following:

Lemma 2.3.

Let (ut)(u_{t}) be a solution to (2.2). For all b0b\geq 0, η(0,1/a)\eta\in(0,1/a_{*}), β2\beta\geq 2 and kk0k\leq k_{0}, the following estimate holds,

𝐄u,x[(|Dkϕω,u1|+|Dk(ϕω,u1)1)|)bVβ,η(u1)]ab,k,a,β,ηVβ+b(k1),η(u).\mathbf{E}_{u,x}\left[\left(|D^{k}\phi^{1}_{\omega,u}|+|D^{k}(\phi^{1}_{\omega,u})^{-1})|\right)^{b}V_{\beta,\eta}(u_{1})\right]^{a_{*}}\lesssim_{b,k,a,\beta,\eta}V_{\beta+b(k-1),\eta}(u).
Proof.

To apply Lemma 2.2 we must first bound derivatives of ϕ\phi in terms of something that can be controlled

Claim 2.4.

For each kk\in\mathbb{N}, there exists a ckc_{k} satisfying ck=1c_{k}=1 such that the following estimate holds

sups[0,t]|Dkϕω,us|ksups[0,1]usCkk1exp(ck01DusLds)\sup_{s\in[0,t]}|D^{k}\phi^{s}_{\omega,u}|\lesssim_{k}\sup_{s\in[0,1]}\langle\|u_{s}\|_{C^{k}}\rangle^{k-1}\exp\left(c_{k}\int_{0}^{1}\|Du_{s}\|_{L^{\infty}}\mathrm{d}s\right)
Proof.

The proof will be by induction on kk. Clearly it is true for k=1k=1. We assume it is true for all j(k1)j\leq(k-1). Using the equation (2.5) we can estimate |Dkϕω,ut||D^{k}\phi^{t}_{\omega,u}| using the Faà di Bruno formula

t|Dkϕω,ut|kn=1k|Dnusϕω,ut|Cn,kj=1k|Djϕω,ut|mj,\partial_{t}|D^{k}\phi^{t}_{\omega,u}|\lesssim_{k}\sum_{n=1}^{k}|D^{n}u_{s}\circ\phi^{t}_{\omega,u}|\sum_{C_{n,k}}\prod_{j=1}^{k}|D^{j}\phi^{t}_{\omega,u}|^{m_{j}},

where the sum is over the set

𝒞n,k={(m1,,mk)0k: 1m1+2m2++kmk=k,m1++mk=n}.\mathcal{C}_{n,k}=\{(m_{1},\ldots,m_{k})\in\mathbb{Z}^{k}_{\geq 0}\,:\,1\cdot m_{1}+2\cdot m_{2}+\ldots+k\cdot m_{k}=k,\quad m_{1}+\ldots+m_{k}=n\}.

Separating out the leading order term in the outer sum n=1n=1, |Duϕω,ut||Dkϕω,ut||Du\circ\phi^{t}_{\omega,u}||D^{k}\phi^{t}_{\omega,u}|, and applying Grönwall to the remaining terms, we obtain the following estimate

sups[0,1]|Dkϕω,us|ksups[0,1](n=2kDnusLCn,kj=1k1|Djϕω,us|mj)exp(01DusLds),\sup_{s\in[0,1]}|D^{k}\phi^{s}_{\omega,u}|\lesssim_{k}\sup_{s\in[0,1]}\left(\sum_{n=2}^{k}\|D^{n}u_{s}\|_{L^{\infty}}\sum_{C_{n,k}}\prod_{j=1}^{k-1}|D^{j}\phi^{s}_{\omega,u}|^{m_{j}}\right)\exp\left(\int_{0}^{1}\|Du_{s}\|_{L^{\infty}}\mathrm{d}s\right),

where we used that in mk=0m_{k}=0 if n2n\geq 2 and hence the product is only up to k1k-1. Using the induction hypothesis we have for n2n\geq 2 and (m1,,mk)Cn,k(m_{1},\ldots,m_{k})\in C_{n,k} that

j=1k1|Djϕω,ut|mj\displaystyle\prod_{j=1}^{k-1}|D^{j}\phi^{t}_{\omega,u}|^{m_{j}} ksups[0,1]usCk1j=2k1(j1)mjexp(j=2k1cjmj01DusLds),\displaystyle\lesssim_{k}\sup_{s\in[0,1]}\langle\|u_{s}\|_{C^{k-1}}\rangle^{\sum_{j=2}^{k-1}(j-1)m_{j}}\exp\left(\sum_{j=2}^{k-1}c_{j}m_{j}\int_{0}^{1}\|Du_{s}\|_{L^{\infty}}\mathrm{d}s\right),
ksups[0,1]usCk1k2exp((ck1)01DusLds),\displaystyle\lesssim_{k}\sup_{s\in[0,1]}\langle\|u_{s}\|_{C^{k-1}}\rangle^{k-2}\exp\left((c_{k}-1)\int_{0}^{1}\|Du_{s}\|_{L^{\infty}}\mathrm{d}s\right),

where in the second inequality we used that j=2k1(j1)mj=knk2\sum_{j=2}^{k-1}(j-1)m_{j}=k-n\leq k-2 and define ck:=1+supnsup(m1,,mk)Cn,kj=2k1cjmjc_{k}:=1+\sup_{n}\sup_{(m_{1},\ldots,m_{k})\in C_{n,k}}\sum_{j=2}^{k-1}c_{j}m_{j}. Substituting this back into the expression for |Dkϕω,ut||D^{k}\phi^{t}_{\omega,u}| we obtain the desired estimate. ∎

Likewise we obtain a similar estimate for the inverse:

Claim 2.5.

For each kk\in\mathbb{N}, there exists a c~k>0\tilde{c}_{k}>0 satisfying c~k=1\tilde{c}_{k}=1 such that the following estimate holds

sups[0,t]|Dk(ϕω,us)1|ksups[0,1]usCkk1exp(c~k01DusLds)\sup_{s\in[0,t]}|D^{k}(\phi^{s}_{\omega,u})^{-1}|\lesssim_{k}\sup_{s\in[0,1]}\langle\|u_{s}\|_{C^{k}}\rangle^{k-1}\exp\left(\tilde{c}_{k}\int_{0}^{1}\|Du_{s}\|_{L^{\infty}}\mathrm{d}s\right)
Proof.

To prove this, we will make use of the following estimate on derivatives of the inverse of a diffeomorphism which can be deduced from the Faà di Bruno formula,

|Dk(ϕω,u1)1|kkj=1kDjϕω,u1Lmj,|D^{k}(\phi^{1}_{\omega,u})^{-1}|\lesssim_{k}\sum_{\mathcal{B}_{k}}\prod_{j=1}^{k}\|D^{j}\phi^{1}_{\omega,u}\|^{m_{j}}_{L^{\infty}},

where the sum is over the set

k={(m1,,mk)0k: 1m1+2m2++kmk=2k2,m1++mk=k1}.\mathcal{B}_{k}=\{(m_{1},\ldots,m_{k})\in\mathbb{Z}^{k}_{\geq 0}\,:\,1\cdot m_{1}+2\cdot m_{2}+\ldots+k\cdot m_{k}=2k-2,\quad m_{1}+\ldots+m_{k}=k-1\}.

Substituting in the estimate for |Dkϕω,u1||D^{k}\phi^{1}_{\omega,u}| we obtain

|Dk(ϕω,u1)1|k(ksups[0,1]usCkj=1k(j1)mj)exp(c~k01DusLds),|D^{k}(\phi^{1}_{\omega,u})^{-1}|\lesssim_{k}\left(\sum_{\mathcal{B}_{k}}\sup_{s\in[0,1]}\langle\|u_{s}\|_{C^{k}}\rangle^{\sum_{j=1}^{k}(j-1)m_{j}}\right)\exp\left(\tilde{c}_{k}\int_{0}^{1}\|Du_{s}\|_{L^{\infty}}\mathrm{d}s\right),

for some suitable c~k>0\tilde{c}_{k}>0. Using that j=1k(j1)mj=2k2(k1)=k1\sum_{j=1}^{k}(j-1)m_{j}=2k-2-(k-1)=k-1, we obtain the desired estimate.

Using these estimates we can now prove the desired result by applying the super-Lyapunov bound (2.6) to the above estimates assuming that the regularity of uu is sufficiently high σk0\sigma\geq k_{0}.

2.2.2 Construction of ψp\psi_{p} with degenerate noise

Likely the simplest way to construct ψp\psi_{p} is through a spectral perturbation method applied to P^p\hat{P}^{p}, which is the method employed in e.g. [7, 9, 12]. The first step is to prove that 𝟏\mathbf{1} is the unique, dominant eigenvector for P^\hat{P} in CV1C^{1}_{V} (the corresponding eigenvalue is of course 11), which amounts to verifying a spectral gap for P^\hat{P}^{\ast} in the dual Lipschitz metric (i.e. the Wasserstein-1 norm). This is typically done with a weak Harris’ theorem; the proof of this geometric ergodicity found in [9], which closely follows [29], can be used for both Assumption 4 and Assumption 5. Then the observation that limp0P^p=P^\lim_{p\to 0}\hat{P}^{p}=\hat{P} in the strong operator topology implies the existence of a ψp\psi_{p} satisfying Assumption 2 (i) and (iii) through classical spectral perturbation theory. Verifying (ii) is a standard convexity ‘trick’ which proves that Λ(p)=λp+o(p)\Lambda(p)=\lambda p+o(p) as p0p\to 0, where λ\lambda is the Lyapunov exponent. See [9, 12] for expositions of this argument.

2.3 Passive scalars

As a final application of Theorem 1.8 we address specifically the advection-diffusion equation on 𝕋d\mathbb{T}^{d}

tft+utft=κΔft,f|t=0=f0.\displaystyle\partial_{t}f_{t}+u_{t}\cdot\nabla f_{t}=\kappa\Delta f_{t},\quad f|_{t=0}=f_{0}. (2.7)

The transfer operator for this equation can be written as follows

ft=𝐄W~f0(ϕω,u0t)1,\displaystyle f_{t}=\mathbf{E}_{\tilde{W}}f_{0}\circ(\phi^{t}_{\omega,u_{0}})^{-1}, (2.8)

where ϕω,u0t\phi^{t}_{\omega,u_{0}} solves the SDE

dϕω,u0t=ut(ϕω,u0t)+2κdW~t,ϕω,u00(x)=x.\displaystyle\mathrm{d}\phi^{t}_{\omega,u_{0}}=u_{t}(\phi^{t}_{\omega,u_{0}})+\sqrt{2\kappa}\mathrm{d}\tilde{W}_{t},\quad\phi^{0}_{\omega,u_{0}}(x)=x.

In (2.8), the notation 𝐄W~\mathbf{E}_{\tilde{W}} refers to the expectation with respect only to the Brownian motions W~t\tilde{W}_{t} (which are of course independent from (ut)(u_{t})). For the case of Pierrehumbert [12] and stochastic Navier-Stokes under Assumption 4 [7]. it was proved in those references that Assumptions 13 all hold uniformly in κ[0,κ0]\kappa\in[0,\kappa_{0}] for some small κ0\kappa_{0} (note that this is far from obvious). As the proof of Theorem 1.8 is quantitative as well, in these cases, Theorem 1.8 also holds uniformly in κ[0,κ0]\kappa\in[0,\kappa_{0}] for solutions to (2.7). It seems plausible that one could carry this out also for Assumption 5 however as of writing, this result has not yet appeared in the literature.

There is one last notable consequence of Theorem 1.8 which pertains to the limiting system arising in Batchelor-regime passive scalar turbulence [10], specifically the system

dut+(utut+ptνΔut)dt=QdWt\displaystyle\mathrm{d}u_{t}+(u_{t}\cdot\nabla u_{t}+\nabla p_{t}-\nu\Delta u_{t})\mathrm{d}t=Q\mathrm{d}W_{t} (2.9a)
divut=0\displaystyle\operatorname{\mathrm{div}}u_{t}=0 (2.9b)
dft+(utftκΔft)dt=bdξt.\displaystyle\mathrm{d}f_{t}+(u_{t}\cdot\nabla f_{t}-\kappa\Delta f_{t})\mathrm{d}t=b\mathrm{d}\xi_{t}. (2.9c)

For all κ>0\kappa>0, !\exists! stationary measure for the joint (ut,ft)(u_{t},f_{t}) process, which we denote μκ\mu^{\kappa}. In [10] it was proved that any sequence {μκn}n=0\left\{\mu^{\kappa_{n}}\right\}_{n=0}^{\infty} such that κn0\kappa_{n}\to 0 has a subsequence which converges weak-\ast to a measure supported on 𝐇×Hδ{\bf H}\times H^{-\delta} for all δ>0\delta>0 which is a stationary measure for the κ=0\kappa=0 system. One can use Theorem 1.8 to prove that in fact, there is a unique stationary measure for the κ=0\kappa=0 system supported on 𝐇×Hδ{\bf H}\times H^{-\delta} (and hence a unique limit for all such convergent subsequences as κ0\kappa\to 0).

Theorem 2.6.

Under Assumption 4 or Assumption 5, there exists a unique stationary measure to the κ=0\kappa=0 system (2.9a)–(2.9c) supported on 𝐇×Hδ{\bf H}\times H^{-\delta}.

Proof.

The proof is based on Birkhoff’s ergodic theorem and the contraction in HδH^{-\delta}. Suppose that there exist two measure μ1\mu^{1} and μ2\mu^{2} for the process (ut,ft)(u_{t},f_{t}) on 𝐇×Hδ{\bf H}\times H^{-\delta} which are stationary for the κ=0\kappa=0 system. Using ergodic decomposition we may assume that they are ergodic. Hence, by Birkhoff’s ergodic theorem, there exist two sets A^1,A^2\hat{A}_{1},\hat{A}_{2} such that for i,j{1,2}i,j\in\{1,2\}, μi(Aj)=δij\mu^{i}(A_{j})=\delta_{ij} and for every bounded Lipschitz φ\varphi on 𝐇×Hδ{\bf H}\times H^{-\delta}, and every initial data (u0,f0)A^i(u_{0},f_{0})\in\hat{A}_{i} we have

limT1Tt=0T1𝐄φ(ut,ft)φdμi\lim_{T\to\infty}\frac{1}{T}\sum_{t=0}^{T-1}\mathbf{E}\varphi(u_{t},f_{t})\to\int\varphi\mathrm{d}\mu^{i}

Moreover since μ1,μ2\mu^{1},\mu^{2} both project down to a unique stationary measure μ\mu for the Navier-Stokes system on 𝐇{\bf H}, it must be that the projections A1,A2𝐇A_{1},A_{2}\subset{\bf H} of A^1,A^2\hat{A}_{1},\hat{A}_{2} are both full μ\mu measure and hence have a non-empty intersection A1A2A_{1}\cap A_{2} which is also full μ\mu measure. It follows that we may choose initial data (u0,f01)A^1(u_{0},f^{1}_{0})\in\hat{A}_{1} and (u0,f02)A^2(u_{0},f^{2}_{0})\in\hat{A}_{2}, such that u0A1A2u_{0}\in A_{1}\cap A_{2} and we can write

|φdμ1φdμ2|limT1Tt=0T1𝐄|φ(ut,ft1)φ(ut,ft2)|limT1Tt=0T1𝐄ft1ft2Hδ.\left|\int\varphi\mathrm{d}\mu^{1}-\int\varphi\mathrm{d}\mu^{2}\right|\leq\lim_{T\to\infty}\frac{1}{T}\sum_{t=0}^{T-1}\mathbf{E}|\varphi(u_{t},f^{1}_{t})-\varphi(u_{t},f^{2}_{t})|\lesssim\lim_{T\to\infty}\frac{1}{T}\sum_{t=0}^{T-1}\mathbf{E}\|f^{1}_{t}-f^{2}_{t}\|_{H^{-\delta}}. (2.10)

Note that since the velocity utu_{t} is the same for ft1f^{1}_{t} ft2f^{2}_{t}, the difference f~t=ft1ft2\tilde{f}_{t}=f^{1}_{t}-f^{2}_{t} solves the advection-diffusion equation

tf~t+utf~t=0,\partial_{t}\tilde{f}_{t}+u_{t}\cdot\nabla\tilde{f}_{t}=0,

and hence by Theorem 1.8 the right-hand side of (2.10) goes to zero as TT\to\infty, implying that μ1=μ2\mu^{1}=\mu^{2}. ∎

3 Constructing the Symbol

In this section, we use the function ψp\psi_{p} from Assumption 2 to construct a pseudo-differential operator ap,ϵS1ϵpa_{p,\epsilon}\in S^{-p}_{1-\epsilon} which exhibits exponential decay under the map TPT^{*}P, up to a lower order remainder. This involves a certain regularization scheme of the finite regularity symbol ap(u,x,ξ)=ψp(u,x,ξ/|ξ|)|ξ|pa_{p}(u,x,\xi)=\psi_{p}(u,x,\xi/|\xi|)|\xi|^{-p}, where we mollify ψp\psi_{p} at the scale h>0h>0, where h|ξ|ϵh\sim|\xi|^{-\epsilon} as |ξ||\xi|\to\infty. This scheme is reminiscent of para-differential calculus [36], although it differs in the details.

3.1 Lower bound on ψp\psi_{p}

First we show that assumption that ψp\psi_{p} is lower bounded on bounded sets actually implies a 1/V(u)1/V(u) lowerbound on ψp\psi_{p} for all u𝐇u\in\mathbf{H}.

Lemma 3.1.

For all Lyapunov functions V(u)=Vβ,η(u)V(u)=V_{\beta,\eta}(u) with β1\beta\geq 1 and η(0,1/a)\eta\in(0,1/a_{*}) and |p|1\forall\left|p\right|\ll 1, ψp\psi_{p} satisfies the following lower bound

Vβ,η(u)1η|ψp(u,x,v)|.V_{\beta,\eta}(u)^{-1}\lesssim_{\eta}|\psi_{p}(u,x,v)|.
Proof.

Let r>0r>0 and define Br={V(u)r}B_{r}=\{V(u)\leq r\} for V(u)=Vβ,η(u)V(u)=V_{\beta,\eta}(u) with β,η\beta,\eta as above. Now let τ=inf{n>0unBR}\tau=\inf\{n>0\>\,u_{n}\in B_{R}\}. Our first step is to show that for each γ>0\gamma>0, there exists an r>0r>0 big enough so that

𝐏(τ>n)γV(u)eγn.\mathbf{P}(\tau>n)\lesssim_{\gamma}V(u)e^{-\gamma n}. (3.1)

Such a bound follows relatively easily by standard techniques in Markov chains, which we outline now for completeness. Indeed, the discrete Dynkin formula implies that

𝐄eγτV(uτ)=V(u0)+𝐄k=1τeγk(PV(uk1)eγV(uk1)).\mathbf{E}e^{\gamma\tau}V(u_{\tau})=V(u_{0})+\mathbf{E}\sum_{k=1}^{\tau}e^{\gamma k}\left(PV(u_{k-1})-e^{-\gamma}V(u_{k-1})\right).

By the super Lyapunov property PVeγV+KγPV\leq e^{-\gamma}V+K_{\gamma} mentioned in Remark 1.3 and the definition of τ\tau this implies the following exponential estimate on τ\tau

(rKγeγ1)𝐄eγτV(u0).\left(r-\frac{K_{\gamma}}{e^{\gamma}-1}\right)\mathbf{E}e^{\gamma\tau}\leq V(u_{0}).

Hence, taking r>Kγeγ1r>\frac{K_{\gamma}}{e^{\gamma}-1} and using Markov’s inequality gives (3.1).

Next, we note that by the spectral gap condition (Assumption 2) on P^p\hat{P}^{p} we have that ψp\psi_{p} is given by

ψp=limneΛ(p)P^p𝟏inLipV.\psi_{p}=\lim_{n\to\infty}e^{\Lambda(p)}\hat{P}^{p}\bm{1}\quad\text{in}\quad\mathrm{Lip}_{V}. (3.2)

Indeed, eΛ(p)P^pe^{\Lambda(p)}\hat{P}^{p} has a dominant simple eigenvalue 11 with a spectral gap. Let πp\pi_{p} be the rank one spectral projector associated with eΛ(p)e^{-\Lambda(p)} and recall that ψp=πp𝟏\psi_{p}=\pi_{p}\bm{1}. Then by the spectral gap assumption and Gelfand’s formula limneΛ(p)n(P^p)n(Iπp)𝟏=0\lim_{n\to\infty}e^{\Lambda(p)n}(\hat{P}^{p})^{n}(I-\pi_{p})\bm{1}=0 in LipV\mathrm{Lip}_{V}. It follows that

limneΛ(p)n(P^p)n𝟏=limneΛ(p)n(P^p)nπp𝟏=ψp,in LipV\lim_{n\to\infty}e^{\Lambda(p)n}(\hat{P}^{p})^{n}\bm{1}=\lim_{n\to\infty}e^{\Lambda(p)n}(\hat{P}^{p})^{n}\pi_{p}\bm{1}=\psi_{p},\quad\text{in }\mathrm{Lip}_{V}

The limit formula (3.2) implies that for each uBru\in B_{r}, we have

ψp(u,x,v)=limneΛ(p)n𝐄|Dxϕω,unv|p,\psi_{p}(u,x,v)=\lim_{n\to\infty}e^{\Lambda(p)n}\mathbf{E}|D_{x}\phi^{n}_{\omega,u}v|^{-p},

which is uniform over uu in BrB_{r}. By Assumption 2 there exists a cr>0c_{r}>0 such that

infuBr|ψp(u,x,v)|cr.\inf_{u\in B_{r}}|\psi_{p}(u,x,v)|\geq c_{r}.

With this in mind, let N0>0N_{0}>0 be large enough so that

Gn(u,x,v):=eΛ(p)n𝐄|Dxϕω,unv|pcrG_{n}(u,x,v):=e^{\Lambda(p)n}\mathbf{E}|D_{x}\phi^{n}_{\omega,u}v|^{-p}\geq c_{r}

for all nN0n\geq N_{0} and uBru\in B_{r}, (x,v)P𝕋d(x,v)\in P\mathbb{T}^{d}. Now we can write Gn(u,x,v)G_{n}(u,x,v) as

Gn(u,x,v)\displaystyle G_{n}(u,x,v) =𝐄eΛ(p)τ|Dxϕω,uτv|peΛ(p)(nτ)|Dxτϕθτω,uτnτvτ|p\displaystyle=\mathbf{E}e^{\Lambda(p)\tau}|D_{x}\phi^{\tau}_{\omega,u}v|^{-p}\cdot e^{\Lambda(p)(n-\tau)}|D_{x_{\tau}}\phi^{n-\tau}_{\theta^{\tau}\omega,u_{\tau}}v_{\tau}|^{-p}
=𝐄[eΛ(p)τ|Dxϕω,uτv|p𝐄(eΛ(p)(nτ)|Dxτϕθτω,uτnτvτ|p|τ)]\displaystyle=\mathbf{E}\left[e^{\Lambda(p)\tau}|D_{x}\phi^{\tau}_{\omega,u}v|^{-p}\cdot\mathbf{E}\left(e^{\Lambda(p)(n-{\tau})}|D_{x_{\tau}}\phi^{n-\tau}_{\theta^{\tau}\omega,u_{\tau}}v_{{\tau}}|^{-p}\bigg{|}\mathcal{F}_{\tau}\right)\right]\,
=𝐄[eΛ(p)τ|Dxϕω,uτv|pGnτ(uτ,xτ,vτ)]\displaystyle=\mathbf{E}\left[e^{\Lambda(p)\tau}|D_{x}\phi^{\tau}_{\omega,u}v|^{-p}\cdot G_{n-\tau}(u_{\tau},x_{\tau},v_{\tau})\right]

Here, (n)(\mathcal{F}_{n}) denotes the standard filtration and τ\mathcal{F}_{\tau} is the corresponding stopped sigma-algebra, consisting of measurable sets KK for which K{τn}nK\cap\{\tau\leq n\}\in\mathcal{F}_{n} for all n0n\geq 0. In the last line of the above inequality, we used the strong Markov property to conclude

Gnτ(uτ,xτ,vτ)=𝐄(eΛ(p)(nτ)|Dxτϕθτω,uτnτvτ|p|τ)G_{n-\tau}(u_{\tau},x_{\tau},v_{\tau})=\mathbf{E}\left(e^{\Lambda(p)(n-{\tau})}|D_{x_{\tau}}\phi^{n-\tau}_{\theta^{\tau}\omega,u_{\tau}}v_{{\tau}}|^{-p}\bigg{|}\mathcal{F}_{\tau}\right)

We now bound ψp\psi_{p} for uBrcu\in B^{c}_{r} as follows. Since for nN0n\geq N_{0} and uBru\in B_{r}, Gn(u,x,v)crG_{n}(u,x,v)\geq c_{r}. Therefore, by (3.1), for any uBru\in B_{r}

Gn(u,x,v)\displaystyle G_{n}(u,x,v) 𝐄[𝟏N0nτeΛ(p)τ|Dxϕω,uτv|pGnτ(uτ,xτ,vτ)]\displaystyle\geq\mathbf{E}\left[{\bf 1}_{N_{0}\leq n-\tau}e^{\Lambda(p)\tau}|D_{x}\phi^{\tau}_{\omega,u}v|^{p}\cdot G_{n-\tau}(u_{\tau},x_{\tau},v_{\tau})\right] (3.3)
cr𝐄(𝟏N0nτeΛ(p)τ|Dxϕω,uτv|p).\displaystyle\geq c_{r}\mathbf{E}\left({\bf 1}_{N_{0}\leq n-\tau}e^{-\Lambda(p)\tau}|D_{x}\phi^{\tau}_{\omega,u}v|^{-p}\right).

Taking nn\to\infty and using that τ<\tau<\infty almost-surely, we conclude by monotone convergence that

ψp=limnGn(u,x,v)cr𝐄(eΛ(p)τ|Dxϕω,uτv|p)cr𝐄(eΛ(p)τ|Dxϕω,uτv|p)1.\psi_{p}=\lim_{n\to\infty}G_{n}(u,x,v)\geq c_{r}\mathbf{E}(e^{-\Lambda(p)\tau}|D_{x}\phi^{\tau}_{\omega,u}v|^{-p})\geq c_{r}\mathbf{E}(e^{\Lambda(p)\tau}|D_{x}\phi^{\tau}_{\omega,u}v|^{p})^{-1}\,.

Next, we note that upon choosing γ\gamma large enough (and consequently rr large enough), we have that

𝐄(eΛ(p)τ|Dxϕω,uτv|p)k>0𝐄(e2Λ(p)k|Dxϕω,ukv|2p)1/2P(τk)1/2γV(u)k>0Cke(2Λ(p)γ)k/2γV(u),\begin{aligned} \mathbf{E}(e^{\Lambda(p)\tau}|D_{x}\phi^{\tau}_{\omega,u}v|^{p})&\leq\sum_{k>0}\mathbf{E}(e^{2\Lambda(p)k}|D_{x}\phi^{k}_{\omega,u}v|^{2p})^{1/2}P(\tau\geq k)^{1/2}\\ &\lesssim_{\gamma}V(u)\sum_{k>0}C^{k}e^{(2\Lambda(p)-\gamma)k/2}\lesssim_{\gamma}V(u)\end{aligned},

where we used the Assumption 1 with k=1k=1 in the last line to bound using the Markov property

𝐄|Dxϕω,ukv|2pC𝐄[|Dxϕω,uk1v|2pV(uk1)]CkV(u),\mathbf{E}|D_{x}\phi^{k}_{\omega,u}v|^{2p}\leq C\mathbf{E}\left[|D_{x}\phi^{k-1}_{\omega,u}v|^{2p}V(u_{k-1})\right]\lesssim C^{k}V(u),

for some C1C\geq 1. ∎

3.2 Regularization scheme

Now we describe a regularization scheme that gives rise to a symbol ap,ϵa_{p,\epsilon} using properties of ψp\psi_{p}. The main result of this section, Lemma 3.2, shows that this symbol belongs to the symbol class S1ϵpS^{-p}_{1-\epsilon}, and is an approximate eigenfunction of TPT^{*}P (see appendix A for a definition of the symbol class).

For each h(0,1)h\in(0,1), we take {ψph}h(0,1)\{\psi_{p}^{h}\}_{h\in(0,1)} to be a family of a CC^{\infty} mollifications of ψp\psi_{p} in the variables (x,v)SM(x,v)\in S^{*}M up to scale hh. We impose the following requirements on such a mollification. First, we require that for all u𝐇u\in\mathbf{H}, we have

min(x,v)SMψp(u,x,v)\displaystyle\min_{(x,v)\in S^{*}M}\psi_{p}(u,x,v) min(x,v)SMψph(u,x,v)\displaystyle\leq\min_{(x,v)\in S^{*}M}\psi_{p}^{h}(u,x,v) (3.4)
max(x,v)SMψph(u,x,v)\displaystyle\max_{(x,v)\in S^{*}M}\psi_{p}^{h}(u,x,v) max(x,v)SMψp(u,x,v).\displaystyle\leq\max_{(x,v)\in S^{*}M}\psi_{p}(u,x,v). (3.5)

We emphasize that CC is independent of uu. Letting dSM:SM×SM[0,)d_{S^{*}M}:S^{*}M\times S^{*}M\to[0,\infty) be the geodesic distance between any two points on SMS^{*}M, and let Lip(SM)\mathrm{Lip}(S^{*}M) be the corresponding Lipschitz norm:

ψph(u)Lip(SM)\displaystyle\|\psi_{p}^{h}(u)\|_{\mathrm{Lip}(S^{*}M)} Cψp(u)Lip(SM),\displaystyle\leq C\|\psi_{p}(u)\|_{\mathrm{Lip}(S^{*}M)}, (3.6)
ψph(u)ψp(u)L(SM)\displaystyle\|\psi_{p}^{h}(u)-\psi_{p}(u)\|_{L^{\infty}(S^{*}M)} Chψp(u)Lip(SM).\displaystyle\leq Ch\|\psi_{p}(u)\|_{\mathrm{Lip}(S^{*}M)}. (3.7)

Regarding higher order derivatives, given any parametrization ν:WSM\nu:W\to S^{*}M where W2n1W\subset\mathbb{R}^{2n-1} is an open set, we require that

ψph(u,ν(z))Ck(W)\displaystyle\|\psi_{p}^{h}(u,\nu(z))\|_{C^{k}(W)} Ck,νh1kψp(u)Lip(SM), for all k>1.\displaystyle\leq C_{k,\nu}h^{1-k}\|\psi_{p}(u)\|_{\mathrm{Lip}(S^{*}M)},\text{ for all }k>1. (3.8)

The existence of the family {ψph}\{\psi_{p}^{h}\} is standard. For instance, when M=𝕋dM=\mathbb{T}^{d}, we can define ψph\psi_{p}^{h} using convolution by a standard mollifier. One can generalize such a mollification scheme to arbitrary MM by using the exponential map or a partition of unity.

Using the boundedness of ψp\psi_{p} w.r.t. the LipV\mathrm{Lip}_{V} norm in (1.5), the above conditions imply that for each η(0,1/a)\eta\in(0,1/a^{*}) and β1\beta\geq 1,

ψph(u)L(SM)+ψph(u)Lip(SM)+1hψph(u)ψp(u)L(SM)β,ηVβ,η(u)\displaystyle\|\psi_{p}^{h}(u)\|_{L^{\infty}(S^{*}M)}+\|\psi_{p}^{h}(u)\|_{\mathrm{Lip}(S^{*}M)}+\frac{1}{h}\|\psi_{p}^{h}(u)-\psi_{p}(u)\|_{L^{\infty}(S^{*}M)}\lesssim_{\beta,\eta}V_{\beta,\eta}(u) (3.9)

and

ψph(u,ν(z))Ck(W)\displaystyle\|\psi_{p}^{h}(u,\nu(z))\|_{C^{k}(W)} Ck,ν,β,ηh1kVβ,η(u)\displaystyle\leq C_{k,\nu,\beta,\eta}h^{1-k}V_{\beta,\eta}(u) (3.10)

for any k>1k>1 and ν:WSM\nu:W\to S^{*}M described above. Moreover,

1Cβ,ηVβ,η(u)ψph(u,x,ξ)\displaystyle\frac{1}{C_{\beta,\eta}V_{\beta,\eta}(u)}\leq\psi_{p}^{h}(u,x,\xi) (3.11)

independently of h>0h>0.

Fix a smooth, non-negative dyadic partition of unity {χN}N=2k,k0\{\chi_{N}\}_{N=2^{k},\ k\in\mathbb{N}_{0}} of \mathbb{R}, such that for each N2N\geq 2, we have sptχNB2NBN/2\mathrm{spt}\ \chi_{N}\subseteq B_{2N}\setminus B_{N/2}, χ1B1\chi_{1}\subseteq B_{1}, and

NχN(z)1\sum_{N}\chi_{N}(z)\equiv 1

for all zz\in\mathbb{R}. For convenience, given any (x,ξ)TM(x,\xi)\in T^{*}M, we define

ξ^:=ξ|ξ|SxM\hat{\xi}:=\frac{\xi}{|\xi|}\in S_{x}^{*}M

For longer expressions, we shall use the notation (ξ)=ξ^(\xi)^{\wedge}=\hat{\xi}. Let ϵ(0,12)\epsilon\in(0,\frac{1}{2}). We define

ap,ϵ(u,x,ξ):=N2χN(|ξ|)ψpNϵ(u,x,ξ^)1|ξ|p.\displaystyle a_{p,\epsilon}(u,x,\xi):=\sum_{N\geq 2}\chi_{N}(|\xi|)\psi^{N^{-\epsilon}}_{p}\left(u,x,\hat{\xi}\right)\frac{1}{|\xi|^{p}}. (3.12)

For each ϵ>0\epsilon>0, this defines an approximation of ap(u,x,ξ)a_{p}(u,x,\xi), which improves as |ξ||\xi|\to\infty. The parameter ϵ\epsilon determines the rate of convergence. Since the role of apa_{p} is that of an eigenfunction to the operator TPT^{*}P, we think of ap,ϵa_{p,\epsilon} as an approximate eigenfunction with error

rp,ϵ=TPap,ϵeΛ(p)ap,ϵ.r_{p,\epsilon}=T^{*}Pa_{p,\epsilon}-e^{-\Lambda(p)}a_{p,\epsilon}.

The following Lemma shows that ap,ϵa_{p,\epsilon} and rp,ϵr_{p,\epsilon} belong to appropriate symbol classes from which we can construct pseudo-differential operators, with rp,ϵr_{p,\epsilon} having strictly lower order than ap,ϵa_{p,\epsilon} (see Appendix A):

Lemma 3.2.

Let η(0,1)\eta\in(0,1) and ϵ(0,14)\epsilon\in(0,\frac{1}{4}). For each u𝐇u\in\mathbf{H}, ap,ϵ(u)S1ϵp(TM)a_{p,\epsilon}(u)\in S^{-p}_{1-\epsilon}(T^{*}M), and rp,ϵ(u)S12ϵpϵ(TM)r_{p,\epsilon}(u)\in S^{-p-\epsilon}_{1-2\epsilon}(T^{*}M). For each k0k\in\mathbb{N}_{0}, the seminorms of ap,ϵa_{p,\epsilon} (defined in Appendix A.1) are bounded as follows: for any β1\beta\geq 1,

[ap,ϵ(u)]kp,1ϵk,p,ϵ,β,ηVβ,η(u).\displaystyle[a_{p,\epsilon}(u)]^{-p,1-\epsilon}_{k}\lesssim_{k,p,\epsilon,\beta,\eta}V_{\beta,\eta}(u). (3.13)

On the other hand, there exists β(k,p,ϵ)1\beta(k,p,\epsilon)\geq 1 (depending only on kk, pp and ϵ\epsilon) such that

[rp,ϵ(u)]kpϵ,12ϵk,p,ϵ,ηVβ(k,p,ϵ),η(u).\displaystyle[r_{p,\epsilon}(u)]_{k}^{-p-\epsilon,1-2\epsilon}\lesssim_{k,p,\epsilon,\eta}V_{\beta({k,p,\epsilon}),\eta}(u). (3.14)
Proof.

We prove the bounds above in multiple steps:

Step 1: We prove (3.13). Fix a coordinate chart ϰι:UιUι\varkappa_{\iota}:U^{\prime}_{\iota}\to U_{\iota} from the atlas {ψι}ιJ\{\psi_{\iota}\}_{\iota\in J} as defined in Appendix A and set ζι=ϰι1\zeta_{\iota}=\varkappa_{\iota}^{-1}. It suffices to show that for any such ιJ\iota\in J, and any two multi-indices α,α\alpha,\alpha^{\prime}, we have

|xαξα{ap,ϵ(u,ζι(x),Dxζιξ)}|ι,α,α,β,η,p,ϵVβ,η(u)ξp+ϵ|α|(1ϵ)|α|.\displaystyle|\partial_{x}^{\alpha}\partial_{\xi}^{\alpha^{\prime}}\{a_{p,\epsilon}(u,\zeta_{\iota}(x),D_{x}\zeta_{\iota}^{-\top}\xi)\}|\lesssim_{\iota,\alpha,\alpha^{\prime},\beta,\eta,p,\epsilon}V_{\beta,\eta}(u)\langle\xi\rangle^{-p+\epsilon|\alpha|-(1-\epsilon)|\alpha^{\prime}|}. (3.15)

In proving this, we will omit the dependence of implicit constants on the parameters listed above. We will also drop the subscript on the coordinate chart and write ζ=ζι\zeta=\zeta_{\iota}. To show this, first observe that

ψpNϵ(u,ζ(x),(Dxζξ))1|ξ|p\psi^{N^{-\epsilon}}_{p}\left(u,\zeta(x),(D_{x}\zeta^{-\top}\xi)^{\wedge}\right)\frac{1}{|\xi|^{p}}

is a p-p-homogeneous function in ξ\xi. Moreover, given any cone 𝒞n\mathcal{C}\subsetneq\mathbb{R}^{n}, we can parameterize ξ0\xi\neq 0 in spherical coordinates using

ξ(λ,y)=λσ(y)\xi(\lambda,y)=\lambda\sigma(y)

where λ>0\lambda>0 and σ(z)\sigma(z) maps an open subset of n1\mathbb{R}^{n-1} to 𝕊d1\mathbb{S}^{d-1}. Then, with this parametrization, we have

ψpNϵ(u,ζ(x),(Dxζξ))1|ξ|p=ψpNϵ(u,ζ(x),(Dxζσ(y)))1λp|Dxζσ(z)|p\psi^{N^{-\epsilon}}_{p}\left(u,\zeta(x),(D_{x}\zeta^{-\top}\xi)^{\wedge}\right)\frac{1}{|\xi|^{p}}=\psi^{N^{-\epsilon}}_{p}\left(u,\zeta(x),(D_{x}\zeta^{-\top}\sigma(y))^{\wedge}\right)\frac{1}{\lambda^{p}|D_{x}\zeta^{-\top}\sigma(z)|^{p}}

We call z=(x,y)z=(x,y) and

ν(z)=(ζ(x),(Dxζσ(y)))SM,\nu(z)=(\zeta(x),(D_{x}\zeta^{-\top}\sigma(y))^{\wedge})\in S^{*}M,

so that ν\nu defines a smooth parametrization of SMS^{*}M in some open subset. Furthermore, the partial derivatives ξ\partial_{\xi} transform as follows under the change of coordinates:

ξi=σi(y)λ+1λj=1n1Sij(y)yj\frac{\partial}{\partial\xi_{i}}=\sigma_{i}(y)\frac{\partial}{\partial\lambda}+\frac{1}{\lambda}\sum_{j=1}^{n-1}S_{ij}(y)\frac{\partial}{\partial y_{j}}

where S(y)S(y) is a n×(n1)n\times(n-1) matrix given by the Moore-Penrose pseudo-inverse of DyσD_{y}\sigma^{\top}, i.e.

k=1n1Sik(y)σjyk=δijσiσj,\sum_{k=1}^{n-1}S_{ik}(y)\frac{\partial\sigma_{j}}{\partial y_{k}}=\delta_{ij}-\sigma_{i}\sigma_{j},

and for any yy, the range of S(y)S(y) (as a matrix) is {σ(y)}\{\sigma(y)\}^{\perp}.

Thus, for all ξd\xi\in\mathbb{R}^{d} such that ξ|ξ|\frac{\xi}{|\xi|} is in the range of σ\sigma, we use the p-p homogeneity in λ\lambda to bound

|xαξα(ψpNϵ(u,ζ(x),(Dxζξ))1|ξ|p)|\displaystyle\left|\partial_{x}^{\alpha}\partial_{\xi}^{\alpha^{\prime}}\left(\psi^{N^{-\epsilon}}_{p}\left(u,\zeta(x),(D_{x}\zeta^{-\top}\xi)^{\wedge}\right)\frac{1}{|\xi|^{p}}\right)\right| ψpNϵ(u,ν(z))Czα+αλp+|α|\displaystyle\lesssim\frac{\|\psi_{p}^{N^{-\epsilon}}(u,\nu(z))\|_{C^{\alpha+\alpha^{\prime}}_{z}}}{\lambda^{-p+|\alpha^{\prime}|}} (3.16)
Vβ,η(u)Nϵ(α+α)λp+|α|\displaystyle\lesssim V_{\beta,\eta}(u)\frac{N^{\epsilon(\alpha+\alpha^{\prime})}}{\lambda^{-p+|\alpha^{\prime}|}} (3.17)
Vβ,η(u)Nϵ(α+α)|Dxζξ|p+|α|.\displaystyle\lesssim V_{\beta,\eta}(u)\frac{N^{\epsilon(\alpha+\alpha^{\prime})}}{|D_{x}\zeta^{-\top}\xi|^{-p+|\alpha^{\prime}|}}. (3.18)

By covering ξn\xi\in\mathbb{R}^{n} by finitely many cones for we which can construct such a smooth map σ\sigma, we recover the bound for all possible ξn\xi\in\mathbb{R}^{n}. In particular, when |Dxζξ|N|D_{x}\zeta^{-\top}\xi|\sim N (as is the case when |Dxζξ||D_{x}\zeta^{-\top}\xi| is in the support of χN\chi_{N}) the above is bounded by

Vα,η(u)|Dxζξ|p(1ϵ)|α|+ϵ|α|.\lesssim V_{\alpha^{\prime},\eta}(u)|D_{x}\zeta^{-\top}\xi|^{-p-(1-\epsilon)|\alpha^{\prime}|+\epsilon|\alpha|}.

On the other hand, using the parametrization σ(y)=ξ|ξ|\sigma(y)=\frac{\xi}{|\xi|} once more, we have

χN(|Dxζξ|)=χ(λN|Dxζσ(y)|).\chi_{N}(|D_{x}\zeta^{-\top}\xi|)=\chi\left(\frac{\lambda}{N}|D_{x}\zeta^{-\top}\sigma(y)|\right).

Then, it is straightforward to show that for all λN\lambda\sim N for which the above expression is nonzero that

|xαξα(χN(|Dxζξ|))|1λ|α|1|Dxζξ||α|.|\partial_{x}^{\alpha}\partial_{\xi}^{\alpha^{\prime}}(\chi_{N}(|D_{x}\zeta^{-\top}\xi|))|\lesssim\frac{1}{\lambda^{|\alpha^{\prime}|}}\lesssim\frac{1}{|D_{x}\zeta^{-\top}\xi|^{|\alpha^{\prime}|}}.

We then apply the multivariable chain rule to each term in the sum which defines ap,ϵa_{p,\epsilon} as in (3.12) to recover (3.13).

Step 2a: We now show (3.14) for k=0k=0. It suffices to show that for all β1\beta\geq 1, η(0,1)\eta\in(0,1) and (x,ξ)TM(x,\xi)\in T^{*}M, we have

|rp,ϵ(u,x,ξ)|β,η,p,ϵVβ,η(u)ξpϵ,|r_{p,\epsilon}(u,x,\xi)|\lesssim_{\beta,\eta,p,\epsilon}V_{\beta,\eta}(u)\langle\xi\rangle^{-p-\epsilon},

and then take β=β(k,p,ϵ)\beta=\beta(k,p,\epsilon) as defined in Step 2b below. Once again, dependence of constants on the parameters enumerated above is implied. Using the identity TPap=eΛ(p)apT^{*}Pa_{p}=e^{-\Lambda(p)}a_{p}, we write

TPap,ϵ(u)eΛ(p)ap,ϵ(u)\displaystyle T^{*}Pa_{p,\epsilon}(u)-e^{-\Lambda(p)}a_{p,\epsilon}(u) =TP[ap,ϵ(u)ap(u)]eΛ(p)(ap,ϵ(u)ap(u)).\displaystyle=T^{*}P[a_{p,\epsilon}(u)-a_{p}(u)]-e^{-\Lambda(p)}(a_{p,\epsilon}(u)-a_{p}(u)). (3.19)

The difference appearing above can be expressed as follows:

ap,ϵ(u,x,ξ)ap(u,x,ξ)=χ1(|ξ|)ψp(u,x,ξ^)1|ξ|p+N2χN(|ξ|)[ψpNϵψp](u,x,ξ^)1|ξ|p.a_{p,\epsilon}(u,x,\xi)-a_{p}(u,x,\xi)=-\chi_{1}(|\xi|)\psi_{p}(u,x,\hat{\xi})\frac{1}{|\xi|^{p}}+\sum_{N\geq 2}\chi_{N}(|\xi|)[\psi^{N^{-\epsilon}}_{p}-\psi_{p}](u,x,\hat{\xi})\frac{1}{|\xi|^{p}}.

Since the first term is compactly supported, it is trivial to bound it by V(u)|ξ|p+ϵ\frac{V(u)}{|\xi|^{p+\epsilon}}. As for the second term, we have

|N2χN(|ξ|)[ψpNϵψp](u,x,ξ^)1|ξ|p|\displaystyle\left|\sum_{N\geq 2}\chi_{N}(|\xi|)[\psi^{N^{-\epsilon}}_{p}-\psi_{p}](u,x,\hat{\xi})\frac{1}{|\xi|^{p}}\right| N2χN(|ξ|)ψpNϵ(u)ψp(u)L(SM)1|ξ|p\displaystyle\lesssim\sum_{N\geq 2}\chi_{N}(|\xi|)\|\psi^{N^{-\epsilon}}_{p}(u)-\psi_{p}(u)\|_{L^{\infty}(S^{*}M)}\frac{1}{|\xi|^{p}} (3.20)
N2χN(|ξ|)Vβ,η(u)Nϵ|ξ|p\displaystyle\lesssim\sum_{N\geq 2}\chi_{N}(|\xi|)\frac{V_{\beta,\eta}(u)N^{-\epsilon}}{|\xi|^{p}} (3.21)
Vβ,η(u)|ξ|p+ϵ.\displaystyle\lesssim\frac{V_{\beta,\eta}(u)}{|\xi|^{p+\epsilon}}. (3.22)

Thus,

|ap,ϵ(u,x,ξ)ap(u,x,ξ)|Vβ,η(u)|ξ|p+ϵ.|a_{p,\epsilon}(u,x,\xi)-a_{p}(u,x,\xi)|\lesssim\frac{V_{\beta,\eta}(u)}{|\xi|^{p+\epsilon}}.

We can apply this bound to the same difference under TPT^{*}P, in conjunction with Assumption 1,

|TP[ap,ϵ(u,x,ξ)ap(u,x,ξ)]|\displaystyle|T^{*}P[a_{p,\epsilon}(u,x,\xi)-a_{p}(u,x,\xi)]| =𝐄|ap,ϵ(u1,ϕ1(x),(Dxϕ1)ξ)ap(u,ϕ1(x),(Dxϕ1)ξ))|\displaystyle=\mathbf{E}|a_{p,\epsilon}(u_{1},\phi^{1}(x),(D_{x}\phi^{1})^{-\top}\xi)-a_{p}(u,\phi^{1}(x),(D_{x}\phi^{1})^{-\top}\xi))| (3.23)
𝐄[Vβ,η(u1)|(Dxϕ1)ξ|p+ϵ]Vβ,η(u)|ξ|p+ϵ.\displaystyle\lesssim\mathbf{E}\left[\frac{V_{\beta,\eta}(u_{1})}{|(D_{x}\phi^{1})^{-\top}\xi|^{p+\epsilon}}\right]\lesssim\frac{V_{\beta,\eta}(u)}{|\xi|^{p+\epsilon}}. (3.24)

In summary,

|rp,ϵ(u,x,ξ)|Vβ,η(u)|ξ|p+ϵ.\displaystyle|r_{p,\epsilon}(u,x,\xi)|\lesssim\frac{V_{\beta,\eta}(u)}{|\xi|^{p+\epsilon}}. (3.25)

On the other hand, we have the uniform bound |rp,ϵ(u,x,ξ)|Vβ,η(u)|r_{p,\epsilon}(u,x,\xi)|\lesssim V_{\beta,\eta}(u), since ap,ϵa_{p,\epsilon} is zero near the pole of 1|ξ|p\frac{1}{|\xi|^{p}}. Taking the minimum of these two bounds, we conclude (3.2).

Step 2b: As in step 1, we fix a coordinate chart ζ1=ζι1\zeta^{-1}=\zeta_{\iota}^{-1}, and suppress the dependence of constants on the parameters. For higher order derivatives of rp,ϵr_{p,\epsilon}, we let |α|+|β|=k1|\alpha|+|\beta|=k\geq 1, and use the crude bound

|xαξβ{rp,ϵ(u,ζ(x),Dxζξ)}|\displaystyle|\partial_{x}^{\alpha}\partial_{\xi}^{\beta}\{r_{p,\epsilon}(u,\zeta(x),D_{x}\zeta^{-\top}\xi)\}| |xαξβ{ap,ϵ(u,ζ(x),Dxζξ)}|\displaystyle\leq|\partial_{x}^{\alpha}\partial_{\xi}^{\beta}\{a_{p,\epsilon}(u,\zeta(x),D_{x}\zeta^{-\top}\xi)\}| (3.26)
+𝐄[|xαξβ{ap,ϵ(u1,ϕ1ζ(x),(Dζ(x)ϕ1)Dxζξ)}|].\displaystyle\quad+\mathbf{E}[|\partial_{x}^{\alpha}\partial_{\xi}^{\beta}\{a_{p,\epsilon}(u_{1},\phi^{1}\circ\zeta(x),(D_{\zeta(x)}\phi^{1})^{-\top}D_{x}\zeta^{-\top}\xi)\}|]. (3.27)

Then, we recycle (3.13) to bound the above by

|xαξβ{rp,ϵ(u,ζ(x),Dxζξ)}|\displaystyle|\partial_{x}^{\alpha}\partial_{\xi}^{\beta}\{r_{p,\epsilon}(u,\zeta(x),D_{x}\zeta^{-\top}\xi)\}| Vβ,η(u)ξp+ϵ|α|(1ϵ)|β|\displaystyle\lesssim V_{\beta,\eta}(u)\langle\xi\rangle^{-p+\epsilon|\alpha|-(1-\epsilon)|\beta|} (3.28)
+𝐄[Qk(ϕ1)kVβ,η(u1)(Dζ(x)ϕ1)ξp+ϵ|α|(1ϵ)|β|].\displaystyle\quad+\mathbf{E}[Q_{k}(\phi^{1})^{k}V_{\beta,\eta}(u_{1})\langle(D_{\zeta(x)}\phi^{1})^{-\top}\xi\rangle^{-p+\epsilon|\alpha|-(1-\epsilon)|\beta|}]. (3.29)

Now, in the latter term, we have

(Dζ(x)ϕ1)ξp+ϵ|α|(1ϵ)|β|𝒬1(ϕ1)p+ϵ|α|+(1ϵ)|β|ξp+ϵ|α|(1ϵ)|β|\langle(D_{\zeta(x)}\phi^{1})^{-\top}\xi\rangle^{-p+\epsilon|\alpha|-(1-\epsilon)|\beta|}\leq\mathcal{Q}_{1}(\phi^{1})^{p+\epsilon|\alpha|+(1-\epsilon)|\beta|}\langle\xi\rangle^{-p+\epsilon|\alpha|-(1-\epsilon)|\beta|}

Hence, applying Assumption 1,

|xαξβ{rp,ϵ(u,ζ(x),Dxζξ)}|\displaystyle|\partial_{x}^{\alpha}\partial_{\xi}^{\beta}\{r_{p,\epsilon}(u,\zeta(x),D_{x}\zeta^{-\top}\xi)\}| (3.30)
(Vβ,η(u)+𝐄[𝒬k(ϕ1)k+p+ϵ|α|+(1ϵ)|β|Vβ,η(u1)])ξp+ϵ|α|(1ϵ)|β|\displaystyle\quad\lesssim\left(V_{\beta,\eta}(u)+\mathbf{E}[\mathcal{Q}_{k}(\phi^{1})^{k+p+\epsilon|\alpha|+(1-\epsilon)|\beta|}V_{\beta,\eta}(u_{1})]\right)\langle\xi\rangle^{-p+\epsilon|\alpha|-(1-\epsilon)|\beta|} (3.31)
Vβ,η(u)ξp+ϵ|α|(1ϵ)|β|.\displaystyle\quad\lesssim V_{\beta,\eta}(u)\langle\xi\rangle^{-p+\epsilon|\alpha|-(1-\epsilon)|\beta|}. (3.32)

Now, observe that since k1k\geq 1,

p+ϵ|α|(1ϵ)|β|pϵ+2ϵ|α|(12ϵ)|β|.-p+\epsilon|\alpha|-(1-\epsilon)|\beta|\leq-p-\epsilon+2\epsilon|\alpha|-(1-2\epsilon)|\beta|.

Combining this with Step 2a, we conclude (3.14) for all k0k\geq 0. ∎

As an application of Gårding’s inequality (Proposition A.7), the lemma above implies that the “norm” fOp(ap,ϵ)f,ff\mapsto\langle\mathrm{Op}(a_{p,\epsilon})f,f\rangle defines the same topology as Hp2(M)H^{-\frac{p}{2}}(M), provided one has control on lower frequencies.

Corollary 3.3 (Quantitative Gårding’s inequality).

Under the hypotheses of Lemma 3.2, we have

1Cp,ϵ,β,ηVβ,η(u)fHp22Cp,ϵ,β,ηVβ,η(u)fHd+322Op(ap,ϵ(u))f,fCp,ϵ,β,ηVβ,η(u)fHp22.\displaystyle\frac{1}{C_{p,\epsilon,\beta,\eta}V_{\beta,\eta}(u)}\|f\|_{H^{-\frac{p}{2}}}^{2}-C_{p,\epsilon,\beta,\eta}V_{\beta,\eta}(u)\|f\|_{H^{-\frac{d+3}{2}}}^{2}\leq\langle\mathrm{Op}(a_{p,\epsilon}(u))f,f\rangle\leq C_{p,\epsilon,\beta,\eta}V_{\beta,\eta}(u)\|f\|_{H^{-\frac{p}{2}}}^{2}. (3.33)
Proof.

Observe that by (3.11) and the definition of ap,ϵa_{p,\epsilon} in equation (3.12), we have

1Cp,ϵ,β,ηVβ,η(u)|ξ|pap,ϵ(u,x,ξ).\displaystyle\frac{1}{C_{p,\epsilon,\beta,\eta}V_{\beta,\eta}(u)|\xi|^{p}}\leq a_{p,\epsilon}(u,x,\xi). (3.34)

for all |ξ|1|\xi|\geq 1, and for any β1\beta\geq 1. Thus, we combine Gårding’s inequality in Proposition A.7 in the case of m=pm=-p and m2s=d+32\frac{m}{2}-s=-\frac{d+3}{2} with the seminorm bounds (3.13) and (3.34). ∎

4 Low regularity mixing

In this section, we use the results of Section 3 to show prove Theorem 1.8, i.e. that fnf_{n} decays exponentially fast in Hp2H^{-\frac{p}{2}}. We break this up into two steps: first, we exhibit time 1 exponential decay of high frequencies of f1f_{1} through the pseudo-differential operator Op(ap,ϵ)\mathrm{Op}(a_{p,\epsilon}) as in Lemma 4.1. Second, we show that low frequencies decay exponentially for long times, as in Lemma 4.2. The latter of these two results does not rely on ap,ϵa_{p,\epsilon}. Rather, it is a consequence of the two-point geometric ergodicity in Assumption 3. We combine these two lemmas to prove Theorem 1.8 in Section 4.3.

4.1 Short time, high frequency decay

Lemma 3.2 and Egorov’s Theorem (Theorem A.8) imply that under conjugation by the time-1 flow map ϕ1\phi^{1}, the operator Op(ap,ϵ)\mathrm{Op}(a_{p,\epsilon}) decays by a factor of eΛ(p)e^{-\Lambda(p)}, plus a lower order remainder:

Lemma 4.1.

Let ϵ(0,14)\epsilon\in(0,\frac{1}{4}), p(0,1)p\in(0,1), and η(0,1/a)\eta\in(0,1/a^{*}). Then there exists some β=β(p,ϵ)1\beta^{\prime}=\beta^{\prime}(p,\epsilon)\geq 1 (depending only on pp and ϵ\epsilon) such that

|𝐄[Op(ap,ϵ(u1))f1,f1]eΛ(p)Op(ap,ϵ(u))f,f|p,ϵ,ηVβ,η(u)fHp+ϵ22.|\mathbf{E}[\langle\mathrm{Op}(a_{p,\epsilon}(u_{1}))f_{1},f_{1}\rangle]-e^{-\Lambda(p)}\langle\mathrm{Op}(a_{p,\epsilon}(u))f,f\rangle|\lesssim_{p,\epsilon,\eta}V_{\beta^{\prime},\eta}(u)\|f\|_{H^{-\frac{p+\epsilon}{2}}}^{2}. (4.1)
Proof.

By Egorov’s theorem (Theorem A.8), Lemma 3.2 and Assumption 1,

|𝐄[Op(ap,ϵ(u1))f1,f1]Op(TPap,ϵ(u))f,f|\displaystyle|\mathbf{E}[\langle\mathrm{Op}(a_{p,\epsilon}(u_{1}))f_{1},f_{1}\rangle]-\langle\mathrm{Op}(T^{*}Pa_{p,\epsilon}(u))f,f\rangle| (4.2)
𝐄[𝒬k0(ϕ)k0V0,η(u1)]fHp+12ϵ22Vβ1(k0),η(u),\displaystyle\quad\lesssim\mathbf{E}\Big{[}\mathcal{Q}_{k_{0}}(\phi)^{k_{0}}V_{0,\eta}(u_{1})\Big{]}\|f\|_{H^{-\frac{p+1-2\epsilon}{2}}}^{2}\lesssim V_{\beta_{1}(k_{0}),\eta}(u), (4.3)

for some β1(k0)1\beta_{1}(k_{0})\geq 1. On the other hand, by the boundedness of pseudo-differential operators (Proposition A.6) and Lemma 3.2 above,

|Op(TPap,ϵ(u))f,feΛ(p)Op(ap,ϵ(u))f,f|ϵVβ2(k0,p,ϵ),η(u)fHp+ϵ22,\left|\langle\mathrm{Op}(T^{*}Pa_{p,\epsilon}(u))f,f\rangle-e^{-\Lambda(p)}\langle\mathrm{Op}(a_{p,\epsilon}(u))f,f\rangle\right|\lesssim_{\epsilon}V_{\beta_{2}(k_{0},p,\epsilon),\eta}(u)\|f\|_{H^{-\frac{p+\epsilon}{2}}}^{2},

for some β2(k0,p,ϵ)\beta_{2}(k_{0},p,\epsilon). Combining these two bounds, and fact that ϵ<12ϵ\epsilon<1-2\epsilon, we finish the proof choosing β\beta^{\prime} to be the bigger of β1(k0)\beta_{1}(k_{0}) and β2(k0,p,ϵ)\beta_{2}(k_{0},p,\epsilon). ∎

4.2 Long time, low frequency decay

Next, we show how the assumption of 2-point mixing in Assumption 3 implies the exponential decay of fnf_{n} in a very low-regularity norm.

Lemma 4.2.

There are absolute constants α,p(0,1)\alpha,p\in(0,1) such that the following holds. For any β1\beta\geq 1, η(0,1/a)\eta\in(0,1/a^{*}) and any f0Hp/2f_{0}\in H^{-p/2} such that Mf0dx=0\int_{M}f_{0}\,\mathrm{d}x=0, we have

𝐄[Vβ,η(un)fnHd+322]eαnVβ,η(u0)f0Hp22.\displaystyle\mathbf{E}[V_{\beta,\eta}(u_{n})\|f_{n}\|_{H^{-\frac{d+3}{2}}}^{2}]\lesssim{e^{-\alpha n}}V_{\beta,\eta}(u_{0})\|f_{0}\|_{H^{-\frac{p}{2}}}^{2}. (4.4)

Here, we recall that fn=Tnf0=f0(ϕu0,ωn)1f_{n}=T^{n}f_{0}=f_{0}\circ(\phi^{n}_{u_{0},\omega})^{-1}.

Proof.

We break the proof into two steps. For convenience, we write with V(u)=Vβ,η(u)V(u)=V_{\beta,\eta}(u).

Step 1: First, we show that there exists α0>0\alpha_{0}>0 such that

𝐄[V(un)fnHd+322]eα0nV(u)f0L22,\displaystyle\mathbf{E}[V(u_{n})\|f_{n}\|_{H^{-\frac{d+3}{2}}}^{2}]\lesssim e^{-\alpha_{0}n}V(u)\|f_{0}\|_{L^{2}}^{2}, (4.5)

for all n0n\in\mathbb{N}_{0}. To prove the bound above, let K(x,y)K(x,y) be the kernel of the operator

Op(ξd+32)Op(ξd+32)S1d3(TM).\mathrm{Op}(\langle\xi\rangle^{-\frac{d+3}{2}})^{*}\mathrm{Op}(\langle\xi\rangle^{-\frac{d+3}{2}})\in S^{-d-3}_{1}(T^{*}M).

In particular, KC3ϵ(×)K\in C^{3-\epsilon}(\mathcal{M}\times\mathcal{M}) for all ϵ(0,1)\epsilon\in(0,1). We define

K×(x,y)=K(x,y)M×MK(x,y)dxdy,K_{\times}(x,y)=K(x,y)-\iint_{M\times M}K(x^{\prime},y^{\prime})\mathrm{d}x^{\prime}\mathrm{d}y^{\prime},

so that K×K_{\times} is mean zero. Then, since f0f_{0} is mean zero,

𝐄[V(un)fnHd+322]\displaystyle\mathbf{E}[V(u_{n})\|f_{n}\|_{H^{-\frac{d+3}{2}}}^{2}] =M×M𝐄[V(un)K×(ϕn(x),ϕn(y))]f0(x)f0(y)dxdy\displaystyle=\iint_{M\times M}\mathbf{E}[V(u_{n})K_{\times}(\phi^{n}(x),\phi^{n}(y))]f_{0}(x)f_{0}(y)\mathrm{d}x\mathrm{d}y (4.6)
𝐄[V(un)K×(ϕn(x),ϕn(y))]Lx,y2(M2)f0L2(M)2.\displaystyle\leq\|\mathbf{E}[V(u_{n})K_{\times}(\phi^{n}(x),\phi^{n}(y))]\|_{L^{2}_{x,y}(M^{2})}\|f_{0}\|_{L^{2}(M)}^{2}. (4.7)

Then, estimate (4.5) follows from the two-point mixing bound of Assumption 3: for all q>0q>0 sufficiently small, and all (x,y)M×M(x,y)\in M\times M,

|𝐄[V(un)K×(ϕn(x),ϕn(y))]|qV(u0)dM(x,y)qeα0nKL|\mathbf{E}[V(u_{n})K_{\times}(\phi^{n}(x),\phi^{n}(y))]|\lesssim_{q}V(u_{0})d_{M}(x,y)^{-q}e^{-\alpha_{0}n}\|K\|_{L^{\infty}}

for all n0n\in\mathbb{N}_{0}, and dM(x,y)d_{M}(x,y) denotes the distance between xx and yy, Thus, by taking q<d2q<\frac{d}{2}, we have

𝐄[V(un)K×(ϕn(x),ϕn(y))]L22eα0n.\|\mathbf{E}[V(u_{n})K_{\times}(\phi^{n}(x),\phi^{n}(y))]\|_{L^{2}}^{2}\lesssim e^{-\alpha_{0}n}.

Step 2: We show that there exists α1>0\alpha_{1}>0 such that

𝐄[V(un)fnHd+322]eα1nV(u0)f0H12.\displaystyle\mathbf{E}[V(u_{n})\|f_{n}\|_{H^{-\frac{d+3}{2}}}^{2}]\lesssim e^{\alpha_{1}n}V(u_{0})\|f_{0}\|_{H^{-1}}^{2}. (4.8)

To prove this, we use the cocycle property of DϕnD\phi^{n} and Assumption 1 in the case k=1k=1, and take α1>0\alpha_{1}>0 large enough so that

𝐄[V(un)|Dϕn|2]L\displaystyle\|\mathbf{E}[V(u_{n})|D\phi^{n}|^{2}]\|_{L^{\infty}} eα1𝐄[V(un1)|Dϕn1|2]\displaystyle\leq e^{\alpha_{1}}\mathbf{E}[V(u_{n-1})|D\phi^{n-1}|^{2}] (4.9)
e2α1𝐄[V(un2)|Dϕn2|2]\displaystyle\leq e^{2\alpha_{1}}\mathbf{E}[V(u_{n-2})|D\phi^{n-2}|^{2}] (4.10)
\displaystyle\leq\ldots (4.11)
eα1(n1)𝐄[V(u1)|Dϕ1|2]eα1nV(u0).\displaystyle\leq e^{\alpha_{1}(n-1)}\mathbf{E}[V(u_{1})|D\phi^{1}|^{2}]\leq e^{\alpha_{1}n}V(u_{0}). (4.12)

Then, by writing f0=div(Δ1f0)f_{0}=\mathrm{div}(\nabla\Delta^{-1}f_{0}), where Δ\Delta is the Laplace-Beltrami operator and div\mathrm{div} denotes the divergence on MM, we can integrate by parts to get

𝐄[V(un)fnHd+322]\displaystyle\mathbf{E}[V(u_{n})\|f_{n}\|_{H^{-\frac{d+3}{2}}}^{2}] 2KL𝐄[V(un)|Dxϕn|2]]LΔ1f0L12eα1nΔ1f0L22.\displaystyle\lesssim\|\nabla^{2}K\|_{L^{\infty}}\|\mathbf{E}[V(u_{n})|D_{x}\phi^{n}|^{2}]]\|_{L^{\infty}}\|\nabla\Delta^{-1}f_{0}\|_{L^{1}}^{2}\lesssim e^{\alpha_{1}n}\|\nabla\Delta^{-1}f_{0}\|_{L^{2}}^{2}. (4.13)

Finally, we note

Δ1f0L22=(Δ)1f0,f0f0H12,\|\nabla\Delta^{-1}f_{0}\|_{L^{2}}^{2}=\langle(-\Delta)^{-1}f_{0},f_{0}\rangle\lesssim\|f_{0}\|_{H^{-1}}^{2},

from which (4.8) follows.

Step 3: In the final step, we interpolate the two estimates to get (4.4). For each h>0h>0, let f0(h)f^{(h)}_{0} be a mollification of f0f_{0} that satisfies

f0(h)L2hp2f0Hp2,f0f0(h)H1h1p2f0Hp2.\|f^{(h)}_{0}\|_{L^{2}}\lesssim h^{-\frac{p}{2}}\|f_{0}\|_{H^{-\frac{p}{2}}},\quad\|f_{0}-f^{(h)}_{0}\|_{H^{-1}}\lesssim h^{1-\frac{p}{2}}\|f_{0}\|_{H^{-\frac{p}{2}}}.

Then, combining (4.5) and (4.8), we have

𝐄[V(un)fnHd+322]\displaystyle\mathbf{E}[V(u_{n})\|f_{n}\|_{H^{-\frac{d+3}{2}}}^{2}] 𝐄[V(un)Snf0(h)Hd+322]+𝐄[V(un)Sn(f0f0(h))Hd+322]\displaystyle\lesssim\mathbf{E}[V(u_{n})\|S_{n}f^{(h)}_{0}\|_{H^{-\frac{d+3}{2}}}^{2}]+\mathbf{E}[V(u_{n})\|S_{n}(f_{0}-f^{(h)}_{0})\|_{H^{-\frac{d+3}{2}}}^{2}] (4.14)
V(u0)(eα0nhp+eα1nh2p)f0Hp22.\displaystyle\lesssim V(u_{0})(e^{-\alpha_{0}n}h^{-p}+e^{\alpha_{1}n}h^{2-p})\|f_{0}\|_{H^{-\frac{p}{2}}}^{2}. (4.15)

We now choose h=eα0n2ph=e^{-\frac{\alpha_{0}n}{2p}}, and

0<pα02(α0+α1)0<p\leq\frac{\alpha_{0}}{2(\alpha_{0}+\alpha_{1})}

so that

eα0nhpeα0n/2, and eα1nh2p=e(α02+α1α02p)neα0n/2.e^{-\alpha_{0}n}h^{-p}\leq e^{-\alpha_{0}n/2},\quad\text{ and }\quad e^{\alpha_{1}n}h^{2-p}=e^{(\frac{\alpha_{0}}{2}+\alpha_{1}-\frac{\alpha_{0}}{2p})n}\leq e^{-\alpha_{0}n/2}.

Taking α=α02\alpha=\frac{\alpha_{0}}{2}, this gives us (4.4).

4.3 Proof of the main theorem

Now we are ready to prove the main Theorem 1.8

Proof of Theorem 1.8.

We fix the constants α,p(0,1)\alpha,p\in(0,1) as in Lemma 4.2 and set ϵ=15\epsilon=\frac{1}{5}. In the statement of the main theorem, we shall prove the estimate with δ=p2\delta=\frac{p}{2}.

By Lemma 4.1, there exists some β\beta^{\prime} such that for each η(0,1/a)\eta^{\prime}\in(0,1/a^{*})

𝐄[Op(ap,ϵ(u1))f1,f1]eΛ(p)Op(ap,ϵ(u))f0,f0+CηVβ,η(u)f0Hp+ϵ22.\displaystyle\mathbf{E}[\langle\mathrm{Op}(a_{p,\epsilon}(u_{1}))f_{1},f_{1}\rangle]\leq e^{-\Lambda(p)}\langle\mathrm{Op}(a_{p,\epsilon}(u))f_{0},f_{0}\rangle+C_{\eta^{\prime}}V_{\beta^{\prime},\eta^{\prime}}(u)\|f_{0}\|_{H^{-\frac{p+\epsilon}{2}}}^{2}. (4.16)

Using interpolation, there exists ϱ>0\varrho>0 such that for any h>0h>0, there is some constant ChC_{h} for which the following holds:

Vβ,η(u)f0Hp+ϵ22hf0Hp22+ChVβ,η(u)ϱf0Hd+322.V_{\beta^{\prime},\eta^{\prime}}(u)\|f_{0}\|_{H^{-\frac{p+\epsilon}{2}}}^{2}\leq h\|f_{0}\|_{H^{-\frac{p}{2}}}^{2}+C_{h}V_{\beta^{\prime},\eta^{\prime}}(u)^{\varrho}\|f_{0}\|_{H^{-\frac{d+3}{2}}}^{2}.

Applying Gårding’s inequality, Corollary 3.3, to the first term on the right-hand side, we have

Vβ,η(u)f0Hp+ϵ22hOp(ap,ϵ(u))f0,f0+ChVβ,η(u)ϱf0Hd+322.V_{\beta^{\prime},\eta^{\prime}}(u)\|f_{0}\|_{H^{-\frac{p+\epsilon}{2}}}^{2}\leq h\langle\mathrm{Op}(a_{p,\epsilon}(u))f_{0},f_{0}\rangle+C_{h}V_{\beta^{\prime},\eta^{\prime}}(u)^{\varrho}\|f_{0}\|_{H^{-\frac{d+3}{2}}}^{2}. (4.17)

We now set

β:=ϱβ,η:=ϱη,\beta_{*}:=\varrho\beta^{\prime},\quad\eta_{*}:=\varrho\eta^{\prime},

so that Vβ,η(u)ϱ=Vβ,η(u)V_{\beta^{\prime},\eta^{\prime}}(u)^{\varrho}=V_{\beta_{*},\eta_{*}}(u). Under this re-scaling, β1\beta_{*}\geq 1 is an absolute constant, while η(0,1/(aϱ))\eta_{*}\in(0,1/(a^{*}\varrho)). For brevity sake, since β\beta_{*} and η\eta_{*} are now fixed, we will drop the subscripts and write V=Vβ,ηV=V_{\beta_{*},\eta_{*}}.

Combined with (4.16), this yields

𝐄[Op(ap,ϵ(u1))f1,f1](eΛ(p)+h)Op(ap,ϵ(u))f0,f0+ChV(u)f0Hd+322.\displaystyle\mathbf{E}[\langle\mathrm{Op}(a_{p,\epsilon}(u_{1}))f_{1},f_{1}\rangle]\leq(e^{-\Lambda(p)}+h)\langle\mathrm{Op}(a_{p,\epsilon}(u))f_{0},f_{0}\rangle+C_{h}V(u)\|f_{0}\|_{H^{-\frac{d+3}{2}}}^{2}. (4.18)

We now take h=eΛ(p)/2eΛ(p)h=e^{-\Lambda(p)/2}-e^{-\Lambda(p)} and set μ=min{Λ(p)/2,α}\mu=\min\{\Lambda(p)/2,\alpha\}. Then, by the Markov property,

𝐄[Op(ap,ϵ(un+1))fn+1,fn+1|n]eμOp(ap,ϵ(un))fn,fn+CV(un)fnHd+322.\displaystyle\mathbf{E}[\langle\mathrm{Op}(a_{p,\epsilon}(u_{n+1}))f_{n+1},f_{n+1}\rangle|\mathcal{F}_{n}]\leq e^{-\mu}\langle\mathrm{Op}(a_{p,\epsilon}(u_{n}))f_{n},f_{n}\rangle+CV(u_{n})\|f_{n}\|_{H^{-\frac{d+3}{2}}}^{2}. (4.19)

By taking the full expectation of both sides, and iterating the above inequality, we have

𝐄[Op(ap,ϵ(un))fn,fn]eμnOp(ap,ϵ(u))f0,f0+Ck=0n1e(n1k)μ𝐄[V(uk)fkHd+322].\mathbf{E}[\langle\mathrm{Op}(a_{p,\epsilon}(u_{n}))f_{n},f_{n}\rangle]\leq e^{-\mu n}\langle\mathrm{Op}(a_{p,\epsilon}(u))f_{0},f_{0}\rangle+C\sum_{k=0}^{n-1}e^{-(n-1-k)\mu}\mathbf{E}[V(u_{k})\|f_{k}\|_{H^{-\frac{d+3}{2}}}^{2}].

Then, by applying Lemma 4.2 to the sum (recalling that μα\mu\leq\alpha) to get

𝐄[Op(ap,ϵ(un))fn,fn]\displaystyle\mathbf{E}[\langle\mathrm{Op}(a_{p,\epsilon}(u_{n}))f_{n},f_{n}\rangle] eμnOp(ap,ϵ(u))f0,f0+CηV(u)eμnf0Hp22\displaystyle\leq e^{-\mu n}\langle\mathrm{Op}(a_{p,\epsilon}(u))f_{0},f_{0}\rangle+C_{\eta}V(u)e^{-\mu n}\|f_{0}\|_{H^{-\frac{p}{2}}}^{2} (4.20)
ηV(u)eμnf0Hp22.\displaystyle\lesssim_{\eta}V(u)e^{-\mu n}\|f_{0}\|_{H^{-\frac{p}{2}}}^{2}. (4.21)

We apply Corollary 3.3 once more to deduce

𝐄[1V(un)fnHp22]ηV(u)eμnf0Hp22+𝐄[V(un)fnHd+322].\mathbf{E}\left[\frac{1}{V(u_{n})}\|f_{n}\|_{H^{-\frac{p}{2}}}^{2}\right]\lesssim_{\eta}V(u)e^{-\mu n}\|f_{0}\|_{H^{-\frac{p}{2}}}^{2}+\mathbf{E}[V(u_{n})\|f_{n}\|_{H^{-\frac{d+3}{2}}}^{2}].

for each nn. Applying Lemma 4.2 once more to the second term on the right-hand side completes the proof. ∎

Appendix A Pseudo-differential calculus

Here, we review the requisite results on pseudo-differential calculus on compact dd-dimensional Riemannian manifolds (M,g)(M,g).

In this section we fix a finite atlas {ϰι:UιUι}ιJ\{\varkappa_{\iota}:U_{\iota}\to U^{\prime}_{\iota}\}_{\iota\in J}, where each ιJ\iota\in J, we have that UιU_{\iota} is an open set in MM, and VιV_{\iota} is an open set in n\mathbb{R}^{n}. It will also be convenient to define ζι=ϰι1\zeta_{\iota}=\varkappa^{-1}_{\iota}.

A.1 Quantization on manifolds

Definition A.1.

Fix mm\in\mathbb{R} and ρ(12,1]\rho\in(\frac{1}{2},1]. We define the Hörmander symbol classes Sρm(TM)S_{\rho}^{m}(T^{*}M) to be the set of aC(TM)a\in C^{\infty}(T^{*}M) satisfying the following. Given any trivialization Φ:TM|U2n\Phi:T^{*}M|_{U}\to\mathbb{R}^{2n}, we have

|xαξα(aΦ1)(x,ξ)|Cα,α,ΦΦ1(x,ξ)mρ|α|+(1ρ)|α|,\Big{|}\partial_{x}^{\alpha}\partial_{\xi}^{{\alpha^{\prime}}}(a\circ\Phi^{-1})(x,\xi)\Big{|}\leq C_{\alpha,{\alpha^{\prime}},\Phi}\langle\Phi^{-1}(x,\xi)\rangle^{m-\rho|{\alpha^{\prime}}|+(1-\rho)|\alpha|},

for all multi-indices α,α\alpha,{\alpha^{\prime}}, and all (x,ξ)(x,\xi) in the range of Φ\Phi. In the above context, the bracket z\langle z\rangle is given by 1+|z|2\sqrt{1+|z|^{2}} for any zTMz\in T^{*}M, where is defined in terms of the metric |z|2=gπ(z)(z,z)|z|^{2}=g_{\pi(z)}(z,z).

We define a family of seminorms that determine a Fréchet topology on SρmS_{\rho}^{m} as follows. For each k>0k>0, we define the seminorms

[a]km,ρ=supιsup(x,ξ)Uι×nsup|α|,|α|k|xαξα(a(ζι(x),Dxζιξ))Dxζιξm+ρ|α|(1ρ)|α||.\displaystyle[a]_{k}^{m,\rho}=\sup_{\iota}\sup_{(x,\xi)\in U^{\prime}_{\iota}\times\mathbb{R}^{n}}\sup_{|\alpha|,|{\alpha^{\prime}}|\leq k}\Big{|}\partial_{x}^{\alpha}\partial_{\xi}^{\alpha^{\prime}}(a(\zeta_{\iota}(x),D_{x}\zeta_{\iota}^{-\top}\xi))\langle D_{x}\zeta_{\iota}^{-\top}\xi\rangle^{-m+\rho|{\alpha^{\prime}}|-(1-\rho)|\alpha|}\Big{|}. (A.1)

Below, we give a generalization of the Kohn-Nirenberg quantization scheme for manifolds, following [38].

Definition A.2.

Fix χC(TM)\chi\in C^{\infty}(TM) to be a non-negative cutoff function satisfying the following: first, χ1\chi\equiv 1 in an open neighborhood of 0TM0_{TM}; second, the map (x,v)(x,expx(v))(x,v)\mapsto(x,\exp_{x}(v)) defines a diffeomorphism from the support of χ\chi to an open neighborhood of the diagonal {(x,x):xM}M×M\{(x,x)\ :\ x\in M\}\subset M\times M.

Then, given any symbol aSρma\in S^{m}_{\rho}, we set

Op(a)f(x)=1(2π)nTxM×TxMχ(x,v)eiv,ξa(x,ξ)f(expx(v))dvdξ\mathrm{Op}(a)f(x)=\frac{1}{(2\pi)^{n}}\iint_{T_{x}M\times T_{x}^{*}M}\chi(x,v)e^{i\langle v,\xi\rangle}a(x,\xi)f(\exp_{x}(v))\,\mathrm{d}v\mathrm{d}\xi

for any fC(M)f\in C^{\infty}(M).

Remark A.3.

We remark that while this quantization depends on the choice of χ\chi, this construction is unique modulo the addition of infinitely smoothing operators. In contrast to [20], we are only concerned with the principal symbol of our operator, although using this quantization scheme is convenient in Lemma 4.2. An alternative, but equally viable quantization scheme can be found in Theorem 14.1 in [47].

A.2 Function spaces on Riemannian manifolds

Throughout this paper, we shall consider a general compact, smooth, dd-dimensional Riemannian manifold MM. Here, we fix the function spaces throughout the paper. We define L2(M)L^{2}(M) to be the L2L^{2} space on MM with respect to the usual Riemannian volume form, with inner product

f,g=Mf(x)g(x)dx,\langle f,g\rangle=\int_{M}f(x)g(x)\,\mathrm{d}x,

for any two real valued f,gL2(M)f,g\in L^{2}(M). Here, dx\mathrm{d}x is shorthand for dVol(x)\mathrm{d}\mathrm{Vol}(x).

Definition A.4.

We say fHs(M)f\in H^{s}(M) to be the closure of C(M;)C^{\infty}(M;\mathbb{C}) with respect to the norm

fHs2:=fL22+Op(ξs)fL22.\|f\|_{H^{s}}^{2}:=\|f\|_{L^{2}}^{2}+\|\mathrm{Op}(\langle\xi\rangle^{s})f\|_{L^{2}}^{2}.

We now define a norm-like quantity on the group of diffeomorphisms in Ck(M,M)C^{k}(M,M), in order to track the regularity of such maps.

Definition A.5.

Let k1k\geq 1. Given any CkC^{k} diffeomorphism ϕ:MM\phi:M\to M, we define the functional

𝒬k(ϕ):=1+supι,ζsup1|α|kmax{xα(ϰζϕϰι1)L(Vιϕ1(Uζ);n),xα(ϰζϕ1ϰι1)L(Vιϕ(Uζ);n)}.\mathcal{Q}_{k}(\phi):=1+\sup_{\iota,\zeta}\sup_{1\leq|\alpha|\leq k}\max\{\|\partial_{x}^{\alpha}(\varkappa_{\zeta}\circ\phi\circ\varkappa_{\iota}^{-1})\|_{L^{\infty}(V_{\iota}\cap\phi^{-1}(U_{\zeta});\mathbb{R}^{n})},\|\partial_{x}^{\alpha}(\varkappa_{\zeta}\circ\phi^{-1}\circ\varkappa_{\iota}^{-1})\|_{L^{\infty}(V_{\iota}\cap\phi(U_{\zeta});\mathbb{R}^{n})}\}.

In the above, we set L()=0\|\cdot\|_{L^{\infty}(\emptyset)}=0.

A.3 Estimates on pseudo-differential operators

The first result we need is that the operators defined above are bounded between Sobolev spaces. These results are classical when MM is replaced with Euclidean space n\mathbb{R}^{n}.

Proposition A.6.

Let mm\in\mathbb{R} and ρ(12,1]\rho\in(\frac{1}{2},1]. Let aSρma\in S^{m}_{\rho}. Then for each ss\in\mathbb{R}, Op(a)\mathrm{Op}(a) is a bounded operator from HsH^{s} to HsmH^{s-m}. More precisely, for each ss\in\mathbb{R}, there exists some k0k\geq 0 such that

Op(a)HsHsmCm,ρ,s[a]km,ρ.\|\mathrm{Op}(a)\|_{H^{s}\to H^{s-m}}\leq C_{m,\rho,s}[a]_{k}^{m,\rho}.

In particular, for any fHm2f\in H^{\frac{m}{2}},

Op(a)f,fCm,ρ[a]km,ρfHm22.\langle\mathrm{Op}(a)f,f\rangle\leq C_{m,\rho}[a]_{k}^{m,\rho}\|f\|_{H^{\frac{m}{2}}}^{2}.

We now state the weak Gårding inequality on MM.

Proposition A.7.

Let mm\in\mathbb{R} and ρ(12,1]\rho\in(\frac{1}{2},1], and c0,c1(0,)c_{0},c_{1}\in(0,\infty). Let aSρm(TM)a\in S^{m}_{\rho}(T^{*}M), and suppose that it satisfies

c0|ξ|mRe(a(x,ξ))c_{0}|\xi|^{m}\leq\mathrm{Re}(a(x,\xi))

for all (x,ξ)TM(x,\xi)\in T^{*}M, with |ξ|>c1|\xi|>c_{1}. Then, for all fHm2(M)f\in H^{\frac{m}{2}}(M), and s>0s>0, there exists some kk such that

c02fHm22Cm,s,ρ,c0,c1[a]km,ρfHm2s2Op(a)f,fL2.\frac{c_{0}}{2}\|f\|_{H^{\frac{m}{2}}}^{2}-C_{m,s,\rho,c_{0},c_{1}}\left\langle[a]^{m,\rho}_{k}\right\rangle\|f\|_{H^{\frac{m}{2}-s}}^{2}\leq\langle\mathrm{Op}(a)f,f\rangle_{L^{2}}.

We now state a version of Egorov’s theorem on manifolds.

Theorem A.8.

Let mm\in\mathbb{R} and ρ(12,1]\rho\in(\frac{1}{2},1]. Let aSρm(TM)a\in S^{m}_{\rho}(T^{*}M), A=Op(a)A=\mathrm{Op}(a), and let ϕ:MM\phi:M\to M be a smooth diffeomorphism. Define the cotangent lift of ϕ\phi to be the map ϕ~:TMTM\widetilde{\phi}:T^{*}M\to T^{*}M given by

ϕ~:(x,ξ)(ϕ(x),Dxϕξ).\widetilde{\phi}:(x,\xi)\mapsto(\phi(x),D_{x}\phi^{-\top}\xi).

for any xMx\in M and ξTxM\xi\in T_{x}^{*}M. Moreover, given any map ψ\psi between manifolds, we define its pullback ψf=fψ\psi^{*}f=f\circ\psi for scalar valued ff where the composition is well-defined.

The operator Aϕ:=ϕA(ϕ1)A^{\phi}:=\phi^{*}A(\phi^{-1})^{*} can be decomposed

Aϕ=Op(ϕ~a)+R.A^{\phi}=\mathrm{Op}({\widetilde{\phi}}^{*}a)+R.

Here, ϕ~aSρm\widetilde{\phi}^{*}a\in S^{m}_{\rho}. Moreover, for all ss, there exists k00k_{0}\in\mathbb{N}_{0} such that

AϕHsHsm+RHsHsm+2ρ1Cm,s,ρ𝒬k0(ϕ)k0[a]k0m,ρ.\|A^{\phi}\|_{H^{s}\to H^{s-m}}+\|R\|_{H^{s}\to H^{s-m+2\rho-1}}\lesssim C_{m,s,\rho}\mathcal{Q}_{k_{0}}(\phi)^{k_{0}}[a]_{k_{0}}^{m,\rho}.
Remark A.9.

While these results do not appear to be available in the literature as stated, their analogues on n\mathbb{R}^{n} are well-known, with the caveat that such results usually do not quantify the implicit constant in terms of seminorms on aa. See [44, 31]. To recover the results on MM, one can use a partition of unity and an atlas via a standard construction, and Egorov’s theorem for properly supported pseudo-differential operators on n\mathbb{R}^{n} (see for instance [31], Theorem 18.1.17).

The quantitative bounds in terms of the seminorms on aa, as well as 𝒬k(ϕ)\mathcal{Q}_{k}(\phi) in the case of Theorem A.8, are direct consequences of the proofs of these results, as these seminorms appear in bounding error terms in the asymptotic expansions of pseudo-differential operators (see [31], Theorems 18.1.7 and 18.1.8).

References

  • [1] Viviane Baladi. Anisotropic sobolev spaces and dynamical transfer operators: cc^{\infty} foliations. pages 123–135.
  • [2] Viviane Baladi. The quest for the ultimate anisotropic banach space. Journal of Statistical Physics, 166:525–557, 2017.
  • [3] Viviane Baladi and Masato Tsujii. Anisotropic hölder and sobolev spaces for hyperbolic diffeomorphisms. 57:127–154.
  • [4] Peter H Baxendale. Lyapunov exponents and relative entropy for a stochastic flow of diffeomorphisms. Probability Theory and Related Fields, 81:521–554, 1989.
  • [5] PH Baxendale and BL Rozovskii. Kinematic dynamo and intermittence in a turbulent flow. Geophysical & Astrophysical Fluid Dynamics, 73(1-4):33–60, 1993.
  • [6] PH Baxendale and DW Stroock. Large deviations and stochastic flows of diffeomorphisms. Probability Theory and Related Fields, 80(2):169–215, 1988.
  • [7] Jacob Bedrossian, Alex Blumenthal, and Sam Punshon-Smith. Almost-sure enhanced dissipation and uniform-in-diffusivity exponential mixing for advection–diffusion by stochastic navier–stokes. Probability Theory and Related Fields, 179(3):777–834, 2021.
  • [8] Jacob Bedrossian, Alex Blumenthal, and Sam Punshon-Smith. Lagrangian chaos and scalar advection in stochastic fluid mechanics. Journal of the European Mathematical Society, 24(6):1893–1990, 2022.
  • [9] Jacob Bedrossian, Alex Blumenthal, and Samuel Punshon-Smith. Almost-sure exponential mixing of passive scalars by the stochastic navier–stokes equations. The Annals of Probability, 50(1):241–303, 2022.
  • [10] Jacob Bedrossian, Alex Blumenthal, and Samuel Punshon-Smith. The batchelor spectrum of passive scalar turbulence in stochastic fluid mechanics at fixed reynolds number. Communications on Pure and Applied Mathematics, 75(6):1237–1291, 2022.
  • [11] Michael Blank, Gerhard Keller, and Carlangelo Liverani. Ruelle perron frobenius spectrum for anosov maps. 15:1905–1973.
  • [12] Alex Blumenthal, Michele Coti Zelati, and Rishabh S Gvalani. Exponential mixing for random dynamical systems and an example of pierrehumbert. The Annals of Probability, 51(4):1559–1601, 2023.
  • [13] Andrew Carverhill. Furstenberg’s theorem for nonlinear stochastic systems. Probability theory and related fields, 74:529–534, 1987.
  • [14] William Cooperman and Keefer Rowan. Exponential scalar mixing for the 2d navier-stokes equations with degenerate stochastic forcing. arXiv preprint arXiv:2408.02459, 2024.
  • [15] Michele Coti Zelati, Theodore D Drivas, and Rishabh S Gvalani. Mixing by statistically self-similar gaussian random fields. 191:1–11.
  • [16] Mark F Demers. A gentle introduction to anisotropic banach spaces. Chaos, Solitons & Fractals, 116:29–42, 2018.
  • [17] Dmitry Dolgopyat, Vadim Kaloshin, and Leonid Koralov. Sample path properties of the stochastic flows. The Annals of Probability, 32(1A):1–27, 2004.
  • [18] Semyon Dyatlov and Maciej Zworski. Stochastic stability of Pollicott–Ruelle resonances. Nonlinearity, 28(10):3511, September 2015.
  • [19] Semyon Dyatlov and Maciej Zworski. Dynamical zeta functions for anosov flows via microlocal analysis. Ann. Sci. Ec. Norm. Super., 49(3):543–577, 2016.
  • [20] Frederic Faure, Nicolas Roy, and Johannes Sjostrand. Semi-classical approach for anosov diffeomorphisms and ruelle resonances. The Open Mathematics Journal, 1(1):35–81, 2008.
  • [21] Frédéric Faure and Johannes Sjöstrand. Upper bound on the density of ruelle resonances for anosov flows. 308:325–364.
  • [22] Frédéric Faure and Masato Tsujii. Fractal weyl law for the ruelle spectrum of anosov flows. 6:331–426.
  • [23] Frédéric Faure and Masato Tsujii. Micro-local analysis of contact anosov flows and band structure of the ruelle spectrum.
  • [24] Frédéric Faure and Masato Tsujii. Prequantum transfer operator for symplectic anosov diffeomorphism.
  • [25] Frédéric Faure and Masato Tsujii. Semiclassical approach for the Ruelle-pollicott spectrum of hyperbolic dynamics, volume 9 of Springer INdAM series, pages 65–135. Springer International Publishing.
  • [26] Frédéric Faure and Masato Tsujii. The semiclassical zeta function for geodesic flows on negatively curved manifolds. 208:851–998.
  • [27] Benjamin Gess and Ivan Yaroslavtsev. Stabilization by transport noise and enhanced dissipation in the kraichnan model.
  • [28] Sébastien Gouëzel and Carlangelo Liverani. Banach spaces adapted to anosov systems. 26:189.
  • [29] Martin Hairer and Jonathan C. Mattingly. Ergodicity of the 2D Navier-Stokes equations with degenerate stochastic forcing. Ann. of Math., 164(3):993–1032, 2006.
  • [30] Martin Hairer and Jonathan C. Mattingly. Spectral gaps in Wasserstein distances and the 2D stochastic Navier-Stokes equations. Ann. Probab., 36(6):2050–2091, 2008.
  • [31] Lars Hörmander. The analysis of linear partial differential operators III: Pseudo-differential operators. Springer Science & Business Media, 2007.
  • [32] Yuri Kifer. Ergodic theory of random transformations, volume 10. Springer Science & Business Media, 2012.
  • [33] Sergei Kuksin and Armen Shirikyan. Mathematics of two-dimensional turbulence, volume 194. Cambridge University Press, 2012.
  • [34] Hiroshi Kunita. Stochastic Flows and Stochastic Differential Equations, volume 24. Cambridge University Press.
  • [35] Carlangelo Liverani and ,Dipartimento di Matematica, Università di Roma “Tor Vergata”, Via della Ricerca Scientifica, I-00133 Roma. Fredholm determinants, anosov maps and ruelle resonances. 13:1203–1215.
  • [36] Guy Métivier. Para-differential calculus and applications to the cauchy problem for nonlinear systems. 2008.
  • [37] Valery Iustinovich Oseledets. A multiplicative ergodic theorem. characteristic Ljapunov exponents of dynamical systems. Trudy Moskovskogo Matematicheskogo Obshchestva, 19:179–210, 1968.
  • [38] Markus J Pflaum. The normal symbol on riemannian manifolds. New York J. Math, 4:97–125, 1998.
  • [39] Madabusi S Raghunathan. A proof of Oseledec’s multiplicative ergodic theorem. Israel Journal of Mathematics, 32(4):356–362, 1979.
  • [40] Marco Romito and Lihu Xu. Ergodicity of the 3d stochastic Navier–Stokes equations driven by mildly degenerate noise. Stochastic Processes and their Applications, 121(4):673–700, 2011.
  • [41] D Ruelle. Resonances of chaotic dynamical systems. 56:405–407.
  • [42] David Ruelle. Locating resonances for AxiomA dynamical systems. 44:281–292.
  • [43] David Ruelle. Thermodynamic formalism: The mathematical structure of equilibrium statistical mechanics. Cambridge University Press.
  • [44] Michael E Taylor. Pseudodifferential operators. princeton univ. Press, Zb1, 453:47026, 1981.
  • [45] Marcelo Viana. Lectures on Lyapunov exponents, volume 145. Cambridge University Press, 2014.
  • [46] E Weinan and Jonathan C Mattingly. Ergodicity for the navier-stokes equation with degenerate random forcing: finite-dimensional approximation. Comm. Pure Appl. Math, 54(11):1386–1402, 2001.
  • [47] Maciej Zworski. Semiclassical analysis, volume 138. American Mathematical Society, 2022.