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

A Microlocal Analysis of the Lévy Generator with Conjugate Points

Kevin Tully
Abstract.

We analyze the microlocal structure of the infinitesimal generator of a Lévy process on a closed Riemannian manifold when conjugate points are allowed. We show that if there are no singular conjugate pairs, then the infinitesimal generator can be written as a sum of pseudodifferential operators and Fourier integral operators. This extends and unifies known results for the flat torus, the sphere, and Anosov manifolds.

1. Introduction

In this paper, we will study the infinitesimal generator of a Lévy process on a Riemannian manifold (M,g)(M,g). Loosely speaking, a Lévy process on a manifold looks like Brownian motion interlaced with jumps along geodesics at random times. The infinitesimal generator encodes certain information about the process and reflects the geometry of (M,g)(M,g). For example, the generator of the most famous Lévy process, Brownian motion, is the Laplace-Beltrami operator. Though Brownian motion on manifolds is well studied [Elw88, Hsu02], the theory for more general Lévy processes has received less attention. Hunt introduced Lévy processes on Lie groups [Hun56], while Gangolli initiated the study of symmetric spaces [Gan64, Gan65]. Applebaum and Estrade were the first to construct Lévy processes on arbitrary Riemannian manifolds, under a natural assumption on the Lévy measure [AE00]. Our goal is to analyze the microlocal structure of the infinitesimal generator of this Lévy process.

One motivation for studying Lévy processes on manifolds is to shed light on the Lévy flight foraging hypothesis, which is the foundation of several biological models [Vis+96, BN13] and search algorithms [YD09, YD10, HP17, KKM22]. This controversial hypothesis claims that Lévy processes are a better model of animal foraging behavior than Brownian motion, in the sense that they can optimize search efficiencies [SK86, Vis+99]. Even though the underlying geometry is often curved, until recently this topic had only been studied in Euclidean space. The expected time for a pure jump Lévy process to find a small target on a manifold was first studied in [Cha+22], while [NTT21, NTT22, Nur+23] considered Brownian motion. A numerical comparison of [Cha+22] and [Nur+23] was performed in [TT23], confirming that Brownian motion might be the faster search strategy for small targets on the 22-torus. This suggests that the underlying geometry could be an important factor in determining whether the Lévy flight foraging hypothesis is valid in a given context.

In [Cha+22], conjugate points are one of the main geometric influences on the expected stopping time. Specifically, the authors show that on the sphere the expected stopping time exhibits singular behavior at antipodal points, but no such anomaly occurs on the flat torus or Anosov manifolds. (We will call (M,g)(M,g) an Anosov manifold if its geodesic flow is an Anosov flow on its unit sphere bundle [Ano69, Kni02], or, equivalently, if gg lies in the C2C^{2} interior of the set of metrics on MM without conjugate points [Rug91].) Conjugate points exert influence on the expected stopping time through the microlocal structure of the Lévy generator (i.e., the infinitesimal generator of the pure jump Lévy process). On the sphere the Lévy generator decomposes into a pseudodifferential operator and a Fourier integral operator, while on the flat torus and Anosov manifolds it is simply a pseudodifferential operator. Inspired by these results, as well as the microlocal analysis of the geodesic X-ray transform done in [SU12, MSU15, HU18], we prove a similar theorem in a more general geometric context.

Our main result, Theorem 3.8, shows that if there are no singular conjugate pairs, then the Lévy generator equals a sum of pseudodifferential operators and Fourier integral operators. Each Fourier integral operator is associated with conjugate pairs of a given order, and the order of the operator depends on the order of those conjugate pairs and the dimension of MM. The canonical relation of each Fourier integral operator is related to the geometry of the set of conjugate pairs. When (M,g)(M,g) is the sphere or an Anosov manifold we recover the results of [Cha+22]. Our theorem also covers all possibilities in two dimensions.

The structure of the paper is as follows. In Section 2 we discuss the necessary background, including an initial decomposition of the Lévy generator into pseudodifferential operators and a remainder term. In Section 3 we prove our main result by showing that the remainder term equals a smoothing operator plus a sum of Fourier integral operators.

Acknowledgements

The author would like to thank Gunther Uhlmann for proposing this problem and for his consistent support, guidance, and patience. This material is based upon work supported by the National Science Foundation Graduate Research Fellowship under Grant No. DGE-2140004.

2. Preliminaries

This section introduces our notation, provides the main definitions, and recalls a result from [Cha+22] that will guide our analysis of the Lévy generator. Throughout the paper, we will assume (M,g)(M,g) satisfies the following condition.

Assumption 2.1.

(M,g)(M,g) is a smooth, closed (that is, compact without boundary), and connected Riemannian manifold of dimension n2n\geq 2.

2.1. The Lévy Generator

Before defining the Lévy generator, let us specify our notation for vector bundles and introduce the exponential map.

Unless stated otherwise, we will denote the projection map associated with a vector bundle by π\pi with a subscript indicating the total space of the bundle (e.g., πTM\pi_{TM} is the projection map from the tangent bundle TMTM to MM). When referring to a point in a vector bundle, one must decide whether to specify the base point. For example, should a point in TMTM be denoted by (x,v)TM(x,v)\in TM or simply vTMv\in TM? We will often omit the base point, but we will include it when there is potential for confusion. We do follow the convention that (x,v)TM(x,v)\in TM means vTMv\in TM and x=πTM(v)x=\pi_{TM}(v), and similarly for points in other vector bundles.

For each vTMv\in TM, let γv\gamma_{v} be the maximal geodesic with initial data γ˙v(0)=v\dot{\gamma}_{v}(0)=v. Since we view γ˙v(0)\dot{\gamma}_{v}(0) as a point in TMTM, this initial data includes (loosely speaking) both the initial position and velocity of the geodesic. Each geodesic γv\gamma_{v} is defined on all of \mathbb{R} because MM is compact. Thus we can define the exponential map exp:TMM\exp:TM\to M by

exp(v)=γv(1).\exp(v)=\gamma_{v}(1).

For each xMx\in M, the restriction of the exponential map to TxMT_{x}M will be denoted by expx\exp_{x}.

Fix α(0,1)\alpha\in(0,1). It is well known that the fractional Laplacian of order α\alpha is the infinitesimal generator of a 2α2\alpha-stable jump process on Euclidean space. To ensure that the Lévy generator is consistent with the fractional Laplacian, we will include the constant

Cn,α=4αΓ(n/2+α)πn/2|Γ(α)|.C_{n,\alpha}=\frac{4^{\alpha}\Gamma(n/2+\alpha)}{\pi^{n/2}|\Gamma(-\alpha)|}.
Definition 2.2.

The Lévy generator 𝒜\boldsymbol{\mathscr{A}} is defined for uC(M)u\in C^{\infty}(M) and xMx\in M by

(𝒜u)(x)=Cn,αp.v.TxM{0}u(expx(v))u(x)|v|gn+2α𝑑Tx(v),(\mathscr{A}u)(x)=C_{n,\alpha}\,\operatorname{p{.}v.}\int_{T_{x}M\setminus\{0\}}\frac{u(\exp_{x}(v))-u(x)}{|v|_{g}^{n+2\alpha}}\,dT_{x}(v),

where dTxdT_{x} is the Riemannian density of TxMT_{x}M induced by the metric g|TxMg|_{T_{x}M}.

2.2. Averaging Along the Geodesic Flow

This subsection introduces one of the operators which will appear in our initial decomposition of the Lévy generator.

Let SMSM be the unit sphere bundle over MM. Its fiber over a point xMx\in M is

SxM:={vTxM:|v|g=1}.S_{x}M:=\{v\in T_{x}M:|v|_{g}=1\}.

Since MM is compact and nn-dimensional, SMSM is a compact manifold of dimension 2n12n-1, and it is an embedded submanifold of TMTM. We will write π\pi rather than πSM\pi_{SM} for the projection map SMMSM\to M, and we will use ιSM\iota_{SM} for the inclusion map ιSM:SMTM\iota_{SM}:SM\hookrightarrow TM.

The geodesic flow on SMSM is the smooth map Φ:SM×SM\Phi:SM\times\mathbb{R}\to SM defined by

Φ(v,s)=γ˙v(s).\Phi(v,s)=\dot{\gamma}_{v}(s).

It is a smooth submersion because Φ(,s):SMSM\Phi(\cdot,s):SM\to SM is a diffeomorphism for each ss\in\mathbb{R}. We will denote the differential of the map Φ(,s)\Phi(\cdot,s) at a vector vSMv\in SM by dvΦ|(v,s)d_{v}\Phi|_{(v,s)}. Also, the geodesic flow on TMTM will be denoted by Φ~\tilde{\Phi}.

Since (M,g)(M,g) is a closed Riemannian manifold, its injectivity radius rinjr_{\operatorname{inj}} is positive and finite. Choose a bump function χCc()\chi\in C_{c}^{\infty}(\mathbb{R}) satisfying χ(s)=1\chi(s)=1 for |s|<rinj2/4|s|<r_{\operatorname{inj}}^{2}/4 and χ(s)=0\chi(s)=0 for |s|>rinj2/2|s|>r_{\operatorname{inj}}^{2}/2. Let

a(s)={(1χ(s2))s12αifs0,0ifs0.a(s)=\begin{cases}(1-\chi(s^{2}))s^{-1-2\alpha}&\text{if}\ s\geq 0,\\ 0&\text{if}\ s\leq 0.\end{cases}

Then we can define an operator Ra:C(SM)C(SM)R_{a}:C^{\infty}(SM)\to C(SM) by

(1) (Raf)(v)=Cn,αa(s)f(Φ(v,s))𝑑s.(R_{a}f)(v)=C_{n,\alpha}\int_{\mathbb{R}}a(s)f(\Phi(v,s))\,ds.

We may think of RafR_{a}f as a certain average of ff along the geodesic flow. This operator will play a crucial role in our analysis of the Lévy generator.

2.3. Pushforward and Pullback by a Smooth Submersion

Next we will define the pushforward and pullback and set our notation for Fourier integral operators.

Definition 2.3.

Let XX and YY be smooth manifolds of dimension nXn_{X} and nYn_{Y}, respectively, with smooth positive densities dxdx and dydy. Let F:XYF:X\to Y be a smooth submersion. Then the pushforward F:Cc(X)Cc(Y)F_{*}:C_{c}^{\infty}(X)\to C_{c}^{\infty}(Y) and the pullback F:Cc(Y)C(X)F^{*}:C_{c}^{\infty}(Y)\to C^{\infty}(X) are defined by the requirement that

Y(Fφ)(y)ψ(y)𝑑y=Xφ(x)ψ(F(x))𝑑x=Xφ(x)(Fψ)(x)𝑑x\int_{Y}(F_{*}\varphi)(y)\psi(y)\,dy=\int_{X}\varphi(x)\psi(F(x))\,dx=\int_{X}\varphi(x)(F^{*}\psi)(x)\,dx

for all φCc(X)\varphi\in C_{c}^{\infty}(X) and ψCc(Y)\psi\in C_{c}^{\infty}(Y).

Explicitly, the pushforward by FF integrates over the level sets of FF, while the pullback precomposes with FF.

We would like to use Definition 2.3 to express RaR_{a} as a composition of simpler operators. Let a~m\tilde{a}^{m} be the operator which multiplies by the smooth function

a~(v,s):=Cn,αa(s).\tilde{a}(v,s):=C_{n,\alpha}\,a(s).

Then a~m\tilde{a}^{m} is a properly supported pseudodifferential operator of order 0 on SM×SM\times\mathbb{R}. Let

p:SM×SMp:SM\times\mathbb{R}\to SM

be the projection map. Since pp_{*} integrates a function over \mathbb{R}, (1) suggests that RaR_{a} equals the composition pa~mΦp_{*}\circ\tilde{a}^{m}\circ\Phi^{*}. This is not quite correct since a~mΦf\tilde{a}^{m}\circ\Phi^{*}f may not have compact support, even if fC(SM)f\in C^{\infty}(SM). But we can use a partition of unity to define the composition of pp_{*} and a~mΦ\tilde{a}^{m}\circ\Phi^{*}, and this is how we will view the operator RaR_{a}.

Recall that if F:XYF:X\to Y is a smooth map between smooth manifolds, then for each xXx\in X the differential dF|x:TxXTF(x)YdF|_{x}:T_{x}X\to T_{F(x)}Y yields a dual linear map

dF|xt:TF(x)YTxX.dF|_{x}^{t}:T_{F(x)}^{*}Y\to T_{x}^{*}X.

Since dF|xtdF|_{x}^{t} is given in coordinates by the transpose of the matrix of dF|xdF|_{x}, if FF is a smooth submersion then dF|xtdF|_{x}^{t} is injective for all xXx\in X.

It is known, going back at least to [GS75], that the pushforward and pullback by a smooth submersion are both Fourier integral operators. We will state the version from Lemma 11 of [HU18] to have this result in the precise format we need.

Lemma 2.4.

Suppose we are in the setting of Definition 2.3. Then the pushforward FF_{*} and the pullback FF^{*} are both Fourier integral operators of order (nYnX)/4(n_{Y}-n_{X})/4. The canonical relation of FF_{*} is

CF={(η,dF|xtη):xX,ηTF(x)Y{0}},C_{F_{*}}=\left\{\left(\eta,dF|_{x}^{t}\,\eta\right):x\in X,\ \eta\in T_{F(x)}^{*}Y\setminus\{0\}\right\},

while the canonical relation of FF^{*} is

CF={(dF|xtη,η):xX,ηTF(x)Y{0}}.C_{F^{*}}=\left\{\left(dF|_{x}^{t}\,\eta,\eta\right):x\in X,\ \eta\in T_{F(x)}^{*}Y\setminus\{0\}\right\}.

We will use the notation of [Hör85] for Fourier integral operators. Given smooth manifolds XX and YY, the set m(X×Y,C)\mathcal{I}^{m}(X\times Y,C^{\prime}) consists of all operators which map (Y)\mathcal{E}^{\prime}(Y) to 𝒟(X)\mathcal{D}^{\prime}(X), and whose Schwartz kernel is a Fourier integral of order mm with Lagrangian given by the twisted canonical relation CC^{\prime}. Unlike [Hör85], we will allow CC to be merely a local canonical relation, meaning it can be immersed rather than embedded.

In this setup, Lemma 2.4 implies that

Φ\displaystyle\Phi^{*} 1/4((SM×)×SM,CΦ),\displaystyle\in\mathcal{I}^{-1/4}\big{(}(SM\times\mathbb{R})\times SM,C_{\Phi^{*}}^{\prime}\big{)},
p\displaystyle p_{*} 1/4(SM×(SM×),Cp),\displaystyle\in\mathcal{I}^{-1/4}\big{(}SM\times(SM\times\mathbb{R}),C_{p_{*}}^{\prime}\big{)},
π\displaystyle\pi^{*} (1n)/4(SM×M,Cπ),\displaystyle\in\mathcal{I}^{(1-n)/4}(SM\times M,C_{\pi^{*}}^{\prime}),
π\displaystyle\pi_{*} (1n)/4(M×SM,Cπ).\displaystyle\in\mathcal{I}^{(1-n)/4}(M\times SM,C_{\pi_{*}}^{\prime}).

In the next subsection, we will see that the key to determining the microlocal structure of 𝒜\mathscr{A} is to show that a certain composition involving these operators is a Fourier integral operator. Our main tool will be the clean intersection calculus for Fourier integral operators, whose statement may be found in [DG75], [Wei75], and [Hör85].

2.4. Initial Decomposition of the Lévy Generator

The next result, a consequence of Theorem 1.61.6 in [Cha+22], is the starting point for our microlocal analysis of 𝒜\mathscr{A}. The idea of the proof is to split 𝒜\mathscr{A} into two parts: one where expx\exp_{x} is injective, and a remainder term involving π\pi_{*}, RaR_{a}, and π\pi^{*}. This decomposition is unnecessary if (M,g)(M,g) is the flat torus, since 𝒜-\mathscr{A} is the fractional Laplace-Beltrami operator in that case (Theorem 1.41.4 in [Cha+22]). Thus we may safely rule out that case throughout the paper.

Theorem 2.5.

If RaR_{a} preserves C(SM)C^{\infty}(SM), then

𝒜=𝒜2α+𝒜0+πRaπ,\mathscr{A}=\mathscr{A}_{2\alpha}+\mathscr{A}_{0}+\pi_{*}\circ R_{a}\circ\pi^{*},

where 𝒜2α\mathscr{A}_{2\alpha} (resp. 𝒜0\mathscr{A}_{0}) is a pseudodifferential operator of order 2α2\alpha (resp. 0) on MM.

Proof.

Fix uC(M)u\in C^{\infty}(M) and xMx\in M. Write 𝒜=A1+A2\mathscr{A}=A_{1}+A_{2}, where

(A1u)(x)\displaystyle(A_{1}u)(x) =Cn,αp.v.TxM{0}χ(|v|g2)u(expx(v))u(x)|v|gn+2α𝑑Tx(v),\displaystyle=C_{n,\alpha}\operatorname{p{.}v.}\int_{T_{x}M\setminus\{0\}}\chi\left(|v|_{g}^{2}\right)\frac{u(\exp_{x}(v))-u(x)}{|v|_{g}^{n+2\alpha}}\,dT_{x}(v),
(A2u)(x)\displaystyle(A_{2}u)(x) =Cn,αTxM{0}(1χ(|v|g2))u(expx(v))u(x)|v|gn+2α𝑑Tx(v).\displaystyle=C_{n,\alpha}\int_{T_{x}M\setminus\{0\}}\left(1-\chi\left(|v|_{g}^{2}\right)\right)\frac{u(\exp_{x}(v))-u(x)}{|v|_{g}^{n+2\alpha}}\,dT_{x}(v).

By our choice of χ\chi, the integrand of A1A_{1} vanishes whenever |v|g2>rinj2/2|v|_{g}^{2}>r_{\operatorname{inj}}^{2}/2, so we can make the change of variables y=expx(v)y=\exp_{x}(v). Then

(A1u)(x)=Cn,αp.v.Mχ(dg(x,y)2)u(y)u(x)dg(x,y)n+2αJ(x,y)𝑑Vg(y),(A_{1}u)(x)=C_{n,\alpha}\operatorname{p{.}v.}\int_{M}\chi(d_{g}(x,y)^{2})\frac{u(y)-u(x)}{d_{g}(x,y)^{n+2\alpha}}J(x,y)\,dV_{g}(y),

where dg(x,y)d_{g}(x,y) is the Riemannian distance from xx to yy, J(x,y)J(x,y) is the Jacobian determinant of the map yexpx1(y)y\mapsto\exp_{x}^{-1}(y), and dVgdV_{g} is the Riemannian density of (M,g)(M,g). Hence 𝒜2α:=A1\mathscr{A}_{2\alpha}:=A_{1} is a pseudodifferential operator of order 2α2\alpha on MM.

For A2A_{2}, use polar coordinates to write

(A2u)(x)=Cn,αSxM0(1χ(s2))u(expx(sv))u(x)s1+2α𝑑s𝑑Sx(v),(A_{2}u)(x)=C_{n,\alpha}\int_{S_{x}M}\int_{0}^{\infty}\left(1-\chi\left(s^{2}\right)\right)\frac{u(\exp_{x}(sv))-u(x)}{s^{1+2\alpha}}\,ds\,dS_{x}(v),

where dSxdS_{x} is the Riemannian density of SxMS_{x}M induced by the metric g|SxMg|_{S_{x}M}. Since RaR_{a} preserves C(SM)C^{\infty}(SM) and π\pi_{*} integrates over the fibers of π\pi, we can write

(A2u)(x)=(πRaπu)(x)(πRaπ1)(x)u(x).(A_{2}u)(x)=(\pi_{*}\circ R_{a}\circ\pi^{*}u)(x)-(\pi_{*}\circ R_{a}\circ\pi^{*}1)(x)u(x).

Let 𝒜0\mathscr{A}_{0} be the operator which multiplies by the constant function (πRaπ1)-(\pi_{*}\circ R_{a}\circ\pi^{*}1). Since 𝒜0\mathscr{A}_{0} is a pseudodifferential operator of order 0 on MM, this completes the proof. ∎

We will see in Theorem 3.1 that RaR_{a} is a Fourier integral operator, so it preserves C(SM)C^{\infty}(SM). Therefore the microlocal analysis of 𝒜\mathscr{A} boils down to showing that πRaπ\pi_{*}\circ R_{a}\circ\pi^{*} is a Fourier integral operator, a task which we will take up in the next section.

3. Microlocal Structure of the Lévy Generator

This section contains our main result: a decomposition of 𝒜\mathscr{A} into pseudodifferential operators and Fourier integral operators. In light of Theorem 2.5, a natural first step is to show that RaR_{a} is a Fourier integral operator. Since we will eventually need to split πRaπ\pi_{*}\circ R_{a}\circ\pi^{*} into several pieces, we will actually introduce an arbitrary smooth function into RaR_{a} and show this more general operator is a Fourier integral operator.

3.1. RaR_{a} is a Fourier Integral Operator

Givan any ψC(SM×)\psi\in C^{\infty}(SM\times\mathbb{R}), let

Ra,ψ=p(ψa~)mΦ,R_{a,\psi}=p_{*}\circ(\psi\tilde{a})^{m}\circ\Phi^{*},

where (ψa~)m(\psi\tilde{a})^{m} is the operator which multiplies by ψa~\psi\tilde{a}. Since (ψa~)m(\psi\tilde{a})^{m} is a properly supported pseudodifferential operator of order 0 on SM×SM\times\mathbb{R} and Φ\Phi^{*} is in 1/4((SM×)×SM,CΦ)\mathcal{I}^{-1/4}\big{(}(SM\times\mathbb{R})\times SM,C_{\Phi^{*}}^{\prime}\big{)}, the operator (ψa~)mΦ(\psi\tilde{a})^{m}\circ\Phi^{*} is also in 1/4((SM×)×SM,CΦ)\mathcal{I}^{-1/4}\big{(}(SM\times\mathbb{R})\times SM,C_{\Phi^{*}}^{\prime}\big{)}. Therefore, if we can show that the clean intersection calculus applies to the composition of pp_{*} and (ψa~)mΦ(\psi\tilde{a})^{m}\circ\Phi^{*}, then Ra,ψR_{a,\psi} is a Fourier integral operator. This is the content of the following theorem.

Theorem 3.1.

Let

CRa,ψ={(ξ,ξ~)TSM×TSM:\displaystyle C_{R_{a,\psi}}=\Big{\{}(\xi,\tilde{\xi})\in T^{*}SM\times T^{*}SM: ssuch that\displaystyle\ \exists\,s\in\mathbb{R}\ \text{such that}
dp|(πTSM(ξ),s)tξ=dΦ|(πTSM(ξ),s)tξ~}.\displaystyle\ dp|_{(\pi_{T^{*}SM}(\xi),s)}^{t}\,\xi=d\Phi|_{(\pi_{T^{*}SM}(\xi),s)}^{t}\,\tilde{\xi}\Big{\}}.

Then Ra,ψ1/2(SM×SM,CRa,ψ)R_{a,\psi}\in\mathcal{I}^{-1/2}(SM\times SM,C_{R_{a,\psi}}^{\prime}).

Proof.

As noted above, we can use a partition of unity to define the composition of pp_{*} and (ψa~)mΦ(\psi\tilde{a})^{m}\circ\Phi^{*}. Then by localizing and reducing to the case of two properly supported Fourier integral operators, the proof boils down to an analysis of the canonical relations.

Since CΦC_{\Phi^{*}} is the canonical relation of (ψa~)mΦ(\psi\tilde{a})^{m}\circ\Phi^{*}, Lemma 2.4 implies that

CRa,ψ=CpCΦ.C_{R_{a,\psi}}=C_{p_{*}}\circ C_{\Phi^{*}}.

Therefore, by the clean intersection calculus, it is enough to prove the following:

  1. (i)

    The intersection

    (2) C:=(Cp×CΦ)(TSM×Δ(T(SM×))×TSM)C:=(C_{p_{*}}\times C_{\Phi^{*}})\cap\big{(}T^{*}SM\times\Delta\big{(}T^{*}(SM\times\mathbb{R})\big{)}\times T^{*}SM\big{)}

    is clean in the sense that CC is an embedded submanifold, and at every point cCc\in C the tangent space TcCT_{c}C equals the intersection of the tangent spaces of the two manifolds being intersected.

  2. (ii)

    The projection map πC:CTSM×TSM\pi_{C}:C\to T^{*}SM\times T^{*}SM is proper.

  3. (iii)

    For every (ξ,ξ~)TSM×TSM(\xi,\tilde{\xi})\in T^{*}SM\times T^{*}SM, the fiber πC1(ξ,ξ~)\pi_{C}^{-1}(\xi,\tilde{\xi}) is connected.

If we are willing to work with local canonical relations, then point (iii) can be omitted. Point (iii) will hold in this case, but in later results it may not.

Consider the map

G:TSM×T(SM×)G:T^{*}SM\times\mathbb{R}\to T^{*}(SM\times\mathbb{R})

defined by

G(ξ,s)=dp|(πTSM(ξ),s)tξ.G(\xi,s)=dp|_{(\pi_{T^{*}SM}(\xi),s)}^{t}\,\xi.

If we include the base points, then

(3) G(πTSM(ξ),ξ,s)=((πTSM(ξ),s),(ξ,0)),G(\pi_{T^{*}SM}(\xi),\xi,s)=\big{(}(\pi_{T^{*}SM}(\xi),s),(\xi,0)\big{)},

so GG is a smooth embedding. The injectivity of GG implies that πC\pi_{C} is injective. Hence point (iii) holds, and assuming points (i) and (ii) are true the excess of the intersection (2) is zero.

To begin proving point (i), let ZZ be the smooth rank-(2n1)(2n-1) subbundle of T(SM×)T^{*}(SM\times\mathbb{R}) whose fiber over each point (v,s)SM×(v,s)\in SM\times\mathbb{R} is

Z(v,s):=Range(dΦ|(v,s)t).Z_{(v,s)}:=\operatorname{Range}\left(d\Phi|_{(v,s)}^{t}\right).

In other words, Z(v,s)Z_{(v,s)} is the subspace of T(v,s)(SM×)T_{(v,s)}^{*}(SM\times\mathbb{R}) which is conormal to the kernel of dΦ|(v,s)d\Phi|_{(v,s)}. To see that ZZ is a smooth subbundle, just use the rank theorem to write

Φ(y1,,y2n)=(y1,,y2n1)\Phi\left(y^{1},\dots,y^{2n}\right)=\left(y^{1},\dots,y^{2n-1}\right)

locally, and note that (dy1,,dy2n1)\left(dy^{1},\dots,dy^{2n-1}\right) is a smooth local frame for ZZ. Because dΦ|(v,s)td\Phi|_{(v,s)}^{t} is a linear isomorphism from TΦ(v,s)SMT_{\Phi(v,s)}^{*}SM to Z(v,s)Z_{(v,s)}, we can define a smooth map

(4) dΦt:ZTSMd\Phi^{-t}:Z\to T^{*}SM

whose restriction to each fiber Z(v,s)Z_{(v,s)} is the inverse (dΦ|(v,s)t)1:Z(v,s)TΦ(v,s)SM(d\Phi|_{(v,s)}^{t})^{-1}:Z_{(v,s)}\to T_{\Phi(v,s)}^{*}SM.

The domain of our smooth parametrization of CC will be the set

𝒪:=Range(G)Z.\mathcal{O}:=\operatorname{Range}(G)\cap Z.

We claim that 𝒪\mathcal{O} is an embedded submanifold of dimension 4n24n-2. Let

(5) q:Zq:Z\to\mathbb{R}

be the restriction to ZZ of the projection map

T(SM×)((v,s),(ξ,σ))σ.T^{*}(SM\times\mathbb{R})\ni\big{(}(v,s),(\xi,\sigma)\big{)}\mapsto\sigma.

Since 𝒪\mathcal{O} equals the level set q1(0)q^{-1}(0), the claim is true if dq|ζdq|_{\zeta} is nonzero for all ζZ\zeta\in Z. Let πV(ζ)=(v,s)\pi_{V}(\zeta)=(v,s). Then ζ=dΦ|(v,s)tθ\zeta=d\Phi|_{(v,s)}^{t}\,\theta for some θTΦ(v,s)SM\theta\in T_{\Phi(v,s)}^{*}SM. Choose slice coordinates for SMSM near vv and near Φ(v,s)\Phi(v,s), and fix natural coordinates for TSMT^{*}SM associated with the latter. Then locally we can write Φ=(Φ1,,Φ2n1)\Phi=\left(\Phi^{1},\dots,\Phi^{2n-1}\right) and θ=(θ1,,θ2n1)\theta=(\theta_{1},\dots,\theta_{2n-1}). Since Φ(v,)\Phi(v,\cdot) is a unit-speed geodesic, we may suppose without loss of generality that

dds~|s~=sΦ1(v,s~)0.\left.\frac{d}{d\tilde{s}}\right|_{\tilde{s}=s}\Phi^{1}(v,\tilde{s})\neq 0.

Define a curve β=(β1,,β2n)\beta=\left(\beta^{1},\dots,\beta^{2n}\right) in TSMT^{*}SM as the composition of dΦ|(v,s)td\Phi|_{(v,s)}^{t} and the curve

τ(θ1+τ,θ2,,θ2n1)TΦ(v,s)SM.\mathbb{R}\ni\tau\mapsto(\theta_{1}+\tau,\theta_{2},\dots,\theta_{2n-1})\in T_{\Phi(v,s)}^{*}SM.

Then β\beta is a smooth curve in Z(v,s)Z_{(v,s)} such that β(0)=ζ\beta(0)=\zeta. Moreover,

(6) ddτ|τ=0(qβ)(τ)=ddτ|τ=0β2n(τ)=dds~|s~=sΦ1(v,s~)0.\left.\frac{d}{d\tau}\right|_{\tau=0}(q\circ\beta)(\tau)=\left.\frac{d}{d\tau}\right|_{\tau=0}\beta^{2n}(\tau)=\left.\frac{d}{d\tilde{s}}\right|_{\tilde{s}=s}\Phi^{1}(v,\tilde{s})\neq 0.

Therefore, as claimed, 𝒪\mathcal{O} is an embedded submanifold of dimension 4n24n-2.

Now we can use 𝒪\mathcal{O} to parametrize CC. Indeed, if Gξ1G_{\xi}^{-1} is the inverse of (3) composed with the projection onto the ξ\xi component, then the map

PC:𝒪TSM×𝒪×𝒪×TSMP_{C}:\mathcal{O}\to T^{*}SM\times\mathcal{O}\times\mathcal{O}\times T^{*}SM

defined by

PC(ζ)=(Gξ1(ζ),ζ,ζ,dΦtζ)P_{C}(\zeta)=\left(G_{\xi}^{-1}(\zeta),\zeta,\zeta,d\Phi^{-t}\zeta\right)

is a smooth embedding, so CC is an embedded submanifold of dimension 4n24n-2.

To finish proving point (i), fix cCc\in C. Since TcCT_{c}C is necessarily contained in the intersection of the tangent spaces of the manifolds on the right-hand side of (2), it is enough to show the reverse containment. Observe that Cp×CΦC_{p_{*}}\times C_{\Phi^{*}} can be parametrized by the map

PCp×CΦ:TSM××ZTSM×Range(G)×Z×TSMP_{C_{p_{*}}\times C_{\Phi^{*}}}:T^{*}SM\times\mathbb{R}\times Z\to T^{*}SM\times\operatorname{Range}(G)\times Z\times T^{*}SM

defined by

PCp×CΦ(ξ,s,ζ)=(ξ,G(ξ,s),ζ,dΦtζ).P_{C_{p_{*}}\times C_{\Phi^{*}}}(\xi,s,\zeta)=\left(\xi,G(\xi,s),\zeta,d\Phi^{-t}\zeta\right).

Hence any vector XTc(Cp×CΦ)X\in T_{c}(C_{p_{*}}\times C_{\Phi^{*}}) is the velocity of some smooth curve

τ(ξ(τ),s(τ),ζ(τ))TSM××Z,\mathbb{R}\ni\tau\mapsto(\xi(\tau),s(\tau),\zeta(\tau))\in T^{*}SM\times\mathbb{R}\times Z,

meaning PCp×CΦ(ξ(0),s(0),ζ(0))=cP_{C_{p_{*}}\times C_{\Phi^{*}}}(\xi(0),s(0),\zeta(0))=c and

X=ddτ|τ=0PCp×CΦ(ξ(τ),s(τ),ζ(τ)).X=\left.\frac{d}{d\tau}\right|_{\tau=0}P_{C_{p_{*}}\times C_{\Phi^{*}}}(\xi(\tau),s(\tau),\zeta(\tau)).

The vector XX is also in Tc(TSM×Δ(T(SM×))×TSM)T_{c}\big{(}T^{*}SM\times\Delta\big{(}T^{*}(SM\times\mathbb{R})\big{)}\times T^{*}SM\big{)} if and only if

ddτ|τ=0G(ξ(τ),s(τ))=ddτ|τ=0ζ(τ).\left.\frac{d}{d\tau}\right|_{\tau=0}G(\xi(\tau),s(\tau))=\left.\frac{d}{d\tau}\right|_{\tau=0}\zeta(\tau).

Thus, in any local coordinates, G(ξ,s)G(\xi,s) and ζ\zeta agree to first order at τ=0\tau=0. Then (3) implies that the same is true of ξ\xi and Gξ1(ζ)G_{\xi}^{-1}(\zeta). Hence

X\displaystyle X =ddτ|τ=0PCp×CΦ(ξ(τ),s(τ),ζ(τ))\displaystyle=\left.\frac{d}{d\tau}\right|_{\tau=0}P_{C_{p_{*}}\times C_{\Phi^{*}}}(\xi(\tau),s(\tau),\zeta(\tau))
=ddτ|τ=0PC(ζ(τ)),\displaystyle=\left.\frac{d}{d\tau}\right|_{\tau=0}P_{C}(\zeta(\tau)),

which means XTcCX\in T_{c}C. This completes the proof of point (i).

For point (ii), suppose we have a compact set KTSM×TSMK\subset T^{*}SM\times T^{*}SM. Then there exists a constant ρ>0\rho>0 such that

KS:={(ξ,ξ~)TSM×TSM:|ξ|g+|ξ~|gρ}.K\subset S:=\left\{(\xi,\tilde{\xi})\in T^{*}SM\times T^{*}SM:|\xi|_{g}+|\tilde{\xi}|_{g}\leq\rho\right\}.

Hence PC1(πC1(S))P_{C}^{-1}(\pi_{C}^{-1}(S)) is precisely the set

{ζ𝒪:|Gξ1(ζ)|g+|dΦtζ|gρ},\left\{\zeta\in\mathcal{O}:|G_{\xi}^{-1}(\zeta)|_{g}+|d\Phi^{-t}\zeta|_{g}\leq\rho\right\},

which is compact by continuity. Then PC1(πC1(K))P_{C}^{-1}(\pi_{C}^{-1}(K)) is compact, because it is a closed subset of the compact set PC1(πC1(S))P_{C}^{-1}(\pi_{C}^{-1}(S)). Since PCP_{C} is a diffeomorphism onto CC, this implies that πC\pi_{C} is a proper map. Thus point (ii) holds.

Since points (i)-(iii) hold and the excess is zero, the map πCPC\pi_{C}\circ P_{C} is a smooth embedding, so its image CRa,ψC_{R_{a,\psi}} is an embedded submanifold. Since pp_{*} and (ψa~)mΦ(\psi\tilde{a})^{m}\circ\Phi^{*} are Fourier integral operators of order 1/4-1/4, we conclude that Ra,ψ1/2(SM×SM,CRa,ψ)R_{a,\psi}\in\mathcal{I}^{-1/2}(SM\times SM,C_{R_{a,\psi}}^{\prime}). ∎

Because Ra,ψR_{a,\psi} is a Fourier integral operator with canonical relation CRa,ψC_{R_{a,\psi}}, its twisted wave front set must be contained in CRa,ψC_{R_{a,\psi}}. In the next lemma, we will leverage the fact that Ra,ψR_{a,\psi} is a composition of Fourier integral operators to say a bit more.

Lemma 3.2.

The twisted wave front set of Ra,ψR_{a,\psi} is contained in

{(ξ,ξ~)TSM×TSM:\displaystyle\Big{\{}(\xi,\tilde{\xi})\in T^{*}SM\times T^{*}SM: s0such that(πTSM(ξ),s)supp(ψ)\displaystyle\ \exists\,s\neq 0\ \text{such that}\ (\pi_{T^{*}SM}(\xi),s)\in\operatorname{supp}(\psi)\
anddp|(πTSM(ξ),s)tξ=dΦ|(πTSM(ξ),s)tξ~}.\displaystyle\ \text{and}\ dp|_{(\pi_{T^{*}SM}(\xi),s)}^{t}\,\xi=d\Phi|_{(\pi_{T^{*}SM}(\xi),s)}^{t}\,\tilde{\xi}\Big{\}}.
Proof.

Since Ra,ψR_{a,\psi} equals p(ψa~)mΦp_{*}\circ(\psi\tilde{a})^{m}\circ\Phi^{*}, we know

WF(Ra,ψ)CpWF((ψa~)m)CΦ.\operatorname{WF}^{\prime}(R_{a,\psi})\subset C_{p_{*}}\circ\operatorname{WF}^{\prime}((\psi\tilde{a})^{m})\circ C_{\Phi^{*}}.

Because the Schwartz kernel of (ψa~)m(\psi\tilde{a})^{m} is smooth away from supp(ψa~ψa~)\operatorname{supp}(\psi\tilde{a}\otimes\psi\tilde{a}), and aC()a\in C^{\infty}(\mathbb{R}) is supported away from 0, we also know

WF((ψa~)m){(ξ^,σ,ξ^,σ)Δ(T(SM×)):\displaystyle\operatorname{WF}^{\prime}((\psi\tilde{a})^{m})\subset\Big{\{}\big{(}\hat{\xi},\sigma,\hat{\xi},\sigma\big{)}\in\Delta\big{(}T^{*}(SM\times\mathbb{R})\big{)}: πT(σ)0and\displaystyle\ \pi_{T^{*}\mathbb{R}}(\sigma)\neq 0\ \text{and}
(πTSM(ξ^),πT(σ))supp(ψ)}.\displaystyle\ \left(\pi_{T^{*}SM}(\hat{\xi}),\pi_{T^{*}\mathbb{R}}(\sigma)\right)\in\operatorname{supp}(\psi)\Big{\}}.

Putting these two containments together yields the result. ∎

This simple observation will be useful when we split πRaπ\pi_{*}\circ R_{a}\circ\pi^{*} into several pieces.

3.2. Composition with π\pi^{*}

Recall that our goal is to understand the microlocal structure of πRaπ\pi_{*}\circ R_{a}\circ\pi^{*}. Theorem 3.1 demonstrated that Ra,ψR_{a,\psi} is a Fourier integral operator for any ψC(SM×)\psi\in C^{\infty}(SM\times\mathbb{R}). The next theorem shows the same is true of the operator

Lψ:=Ra,ψπ.L_{\psi}:=R_{a,\psi}\circ\pi^{*}.
Theorem 3.3.

Let

CLψ={(ξ,η~)TSM\displaystyle C_{L_{\psi}}=\big{\{}(\xi,\tilde{\eta})\in T^{*}SM ×TM:ssuch that\displaystyle\times T^{*}M:\exists\,s\in\mathbb{R}\ \text{such that}
dp|(πTSM(ξ),s)tξ=dΦ|(πTSM(ξ),s)tdπ|Φ(πTSM(ξ),s)tη~}.\displaystyle\ dp|_{(\pi_{T^{*}SM}(\xi),s)}^{t}\,\xi=d\Phi|_{(\pi_{T^{*}SM}(\xi),s)}^{t}\circ d\pi|_{\Phi(\pi_{T^{*}SM}(\xi),s)}^{t}\,\tilde{\eta}\big{\}}.

Then Lψ(n+1)/4(SM×M,CLψ)L_{\psi}\in\mathcal{I}^{-(n+1)/4}(SM\times M,C_{L_{\psi}}^{\prime}).

Proof.

Using Theorem 3.1 and Lemma 2.4, one can check that CLψC_{L_{\psi}} equals CRa,ψCπC_{R_{a,\psi}}\circ C_{\pi^{*}}, so it suffices to show points (i)-(ii) in the proof of Theorem 3.1 hold with (2) replaced by

(7) C:=(CRa,ψ×Cπ)(TSM×Δ(TSM)×TM).C:=(C_{R_{a,\psi}}\times C_{\pi^{*}})\cap(T^{*}SM\times\Delta(T^{*}SM)\times T^{*}M).

(Since we omit point (iii), in general CLψC_{L_{\psi}} will only be a local canonical relation.)

To begin proving point (i), let VV be the smooth rank-nn subbundle of TSMT^{*}SM whose fiber over each vector vSMv\in SM is

Vv:=Range(dπ|vt).V_{v}:=\operatorname{Range}\left(d\pi|_{v}^{t}\right).

Similar to (4), we can define a smooth map

dπt:VTMd\pi^{-t}:V\to T^{*}M

whose restriction to each fiber VvV_{v} is the inverse (dπ|vt)1:VvTπ(v)M(d\pi|_{v}^{t})^{-1}:V_{v}\to T_{\pi(v)}^{*}M. Let

Z~=(dΦt)1(V).\widetilde{Z}=\left(d\Phi^{-t}\right)^{-1}(V).

Since dΦt:ZTSMd\Phi^{-t}:Z\to T^{*}SM is a smooth submersion and VV is an embedded codimension-(n1)(n-1) submanifold of TSMT^{*}SM, we know Z~\widetilde{Z} is an embedded codimension-(n1)(n-1) submanifold of ZZ.

Using a similar argument as the proof of Theorem 3.1, we will show that

𝒪~:=𝒪Z~\widetilde{\mathcal{O}}:=\mathcal{O}\cap\widetilde{Z}

is an embedded submanifold of dimension 3n13n-1. Let q~:Z~\tilde{q}:\widetilde{Z}\to\mathbb{R} be the restriction to Z~\widetilde{Z} of the function qq defined in (5). Then 𝒪~\widetilde{\mathcal{O}} equals the level set q~1(0)\tilde{q}^{-1}(0), so it suffices to show that dq~|ζ~d\tilde{q}|_{\tilde{\zeta}} is nonzero for all ζ~Z~\tilde{\zeta}\in\widetilde{Z}. Let πV(ζ~)=(v,s)\pi_{V}(\tilde{\zeta})=(v,s). Then for some η~Tγv(s)M\tilde{\eta}\in T_{\gamma_{v}(s)}^{*}M, we have

ζ~=dΦ|(v,s)tdπ|Φ(v,s)tη~.\tilde{\zeta}=d\Phi|_{(v,s)}^{t}\circ d\pi|_{\Phi(v,s)}^{t}\,\tilde{\eta}.

Choose slice coordinates for SMSM near Φ(v,s)\Phi(v,s), which yield corresponding coordinates for MM near γv(s)\gamma_{v}(s). Fix natural coordinates on TMT^{*}M. Then locally we can write η~=(η~1,,η~n)\tilde{\eta}=(\tilde{\eta}_{1},\dots,\tilde{\eta}_{n}). Define a curve β\beta in TSMT^{*}SM as the composition of dΦ|(v,s)tdπ|Φ(v,s)td\Phi|_{(v,s)}^{t}\circ d\pi|_{\Phi(v,s)}^{t} and the curve

τ(η~1+τ,η~2,,η~n)Tγv(s)M.\mathbb{R}\ni\tau\mapsto(\tilde{\eta}_{1}+\tau,\tilde{\eta}_{2},\dots,\tilde{\eta}_{n})\in T_{\gamma_{v}(s)}^{*}M.

Then β\beta is a smooth curve in Z~\widetilde{Z} such that β(0)=ζ~\beta(0)=\tilde{\zeta}. Moreover, similar to (6),

ddτ|τ=0(q~β)(τ)0.\left.\frac{d}{d\tau}\right|_{\tau=0}(\tilde{q}\circ\beta)(\tau)\neq 0.

This proves that 𝒪~\widetilde{\mathcal{O}} is an embedded submanifold of dimension 3n13n-1.

Now we can use 𝒪~\widetilde{\mathcal{O}} to parametrize CC via the map

PC:𝒪~TSM×TSM×TSM×TMP_{C}:\widetilde{\mathcal{O}}\to T^{*}SM\times T^{*}SM\times T^{*}SM\times T^{*}M

defined by

PC(ζ~)=(Gξ1(ζ~),dΦtζ~,dΦtζ~,dπtdΦtζ~).P_{C}(\tilde{\zeta})=\left(G_{\xi}^{-1}(\tilde{\zeta}),d\Phi^{-t}\tilde{\zeta},d\Phi^{-t}\tilde{\zeta},d\pi^{-t}\circ d\Phi^{-t}\tilde{\zeta}\right).

By the last paragraph of the proof of Theorem 3.1, the map

𝒪ζ(Gξ1(ζ),dΦtζ)TSM×TSM\mathcal{O}\ni\zeta\mapsto\left(G_{\xi}^{-1}(\zeta),d\Phi^{-t}\zeta\right)\in T^{*}SM\times T^{*}SM

is a smooth embedding, and hence so is its restriction to 𝒪~\widetilde{\mathcal{O}}. Thus PCP_{C} is a smooth embedding, so CC is an embedded submanifold of dimension 3n13n-1.

To complete the proof of (i), note that CRa,ψ×CπC_{R_{a,\psi}}\times C_{\pi^{*}} is parametrized by the map

PCRa,ψ×Cπ:𝒪×VTSM×TSM×V×TMP_{C_{R_{a,\psi}}\times C_{\pi^{*}}}:\mathcal{O}\times V\to T^{*}SM\times T^{*}SM\times V\times T^{*}M

given by

PCRa,ψ×Cπ(ζ,θ)=(Gξ1(ζ),dΦtζ,θ,dπtθ).P_{C_{R_{a,\psi}}\times C_{\pi^{*}}}(\zeta,\theta)=\left(G_{\xi}^{-1}(\zeta),d\Phi^{-t}\zeta,\theta,d\pi^{-t}\theta\right).

Fix cCc\in C and suppose XTc(CRa,ψ×Cπ)X\in T_{c}(C_{R_{a,\psi}}\times C_{\pi^{*}}). Then there is a smooth curve

τ(ζ(τ),θ(τ))𝒪×V\mathbb{R}\ni\tau\mapsto(\zeta(\tau),\theta(\tau))\in\mathcal{O}\times V

such that PCRa,ψ×Cπ(ζ(0),θ(0))=cP_{C_{R_{a,\psi}}\times C_{\pi^{*}}}(\zeta(0),\theta(0))=c and the velocity of this curve at zero is XX. As in the proof of Theorem 3.1, if XTc(TSM×Δ(TSM)×TM)X\in T_{c}(T^{*}SM\times\Delta(T^{*}SM)\times T^{*}M) as well, then the derivatives of dΦtζd\Phi^{-t}\zeta and θ\theta agree at τ=0\tau=0 in any local coordinates. It follows that

X\displaystyle X =ddτ|τ=0PCRa,ψ×Cπ(ζ(τ),θ(τ))\displaystyle=\left.\frac{d}{d\tau}\right|_{\tau=0}P_{C_{R_{a,\psi}}\times C_{\pi^{*}}}(\zeta(\tau),\theta(\tau))
=ddτ|τ=0PC(ζ(τ)),\displaystyle=\left.\frac{d}{d\tau}\right|_{\tau=0}P_{C}(\zeta(\tau)),

which means XTcCX\in T_{c}C. Therefore the intersection (7) is clean and the excess is zero.

The proof of point (ii) is the same as Theorem 3.1, so we omit the details. Thus CLψC_{L_{\psi}} is a local canonical relation. Since Ra,ψR_{a,\psi} and π\pi^{*} are Fourier integral operators of order 1/2-1/2 and (1n)/4(1-n)/4, respectively, we conclude that Lψ(n+1)/4(SM×M,CLψ)L_{\psi}\in\mathcal{I}^{-(n+1)/4}(SM\times M,C_{L_{\psi}}^{\prime}). ∎

When ψ=1\psi=1, we will write LL and CLC_{L} instead of L1L_{1} and CL1C_{L_{1}}. Then by Theorem 2.5,

𝒜=𝒜2α+𝒜0+πL.\mathscr{A}=\mathscr{A}_{2\alpha}+\mathscr{A}_{0}+\pi_{*}\circ L.

Hence the microlocal analysis of 𝒜\mathscr{A} reduces to understanding the composition πL\pi_{*}\circ L.

The main difficulty in this case is that CπCLC_{\pi_{*}}\circ C_{L} may have multiple connected components. One component corresponds to a smoothing operator, and the others appear only when there are conjugate points. If we rule out certain types of conjugate points, then these additional components give rise to Fourier integral operators whose canonical relations and orders can be determined. This is the content of Theorem 3.8, our main result. In the next subsection, we will provide the additional definitions and lemmas needed to state and prove it.

3.3. Conjugate Pairs

Though conjugate points along a geodesic are often defined in terms of vanishing Jacobi fields, it will be more convenient to work with the corresponding velocity vectors of the geodesic instead. This leads us to the notion of a conjugate pair.

Definition 3.4.

We call (v,s)SM×({0})(v,s)\in SM\times(\mathbb{R}\setminus\{0\}) a conjugate pair if

K(v,s):=ker(dπ|Φ(v,s)dvΦ|(v,s))ker(dπ|v){0}.K_{(v,s)}:=\ker\left(d\pi|_{\Phi(v,s)}\circ d_{v}\Phi|_{(v,s)}\right)\cap\ker(d\pi|_{v})\neq\{0\}.

If the dimension of K(v,s)K_{(v,s)} is 1kn11\leq k\leq n-1, then (v,s)(v,s) is a conjugate pair of order 𝐤\boldsymbol{k}. The set of regular conjugate pairs of order 𝐤\boldsymbol{k}, denoted by 𝒞R,k\mathcal{C}_{R,k}, is the set of conjugate pairs which have a neighborhood UU in SM×SM\times\mathbb{R} such that all other conjugate pairs in UU have order kk. The set of singular conjugate pairs, denoted by 𝒞S\mathcal{C}_{S}, is the set of conjugate pairs which are not in 𝒞R,k\mathcal{C}_{R,k} for any kk.

By Lemma 33 in [HU18], Definition 3.4 is equivalent to the traditional definition of conjugate points in terms of vanishing Jacobi fields along a geodesic.

The crucial assumption in Theorem 3.8 is that there are no singular conjugate pairs. This matters because, as the next lemma shows, the set of regular conjugate pairs of order kk is a smooth manifold, which may not be true of the set of all conjugate pairs.

Proposition 3.5.

For each integer 1kn11\leq k\leq n-1, the set 𝒞R,k\mathcal{C}_{R,k} is an embedded (2n1)(2n-1)-dimensional submanifold of SM×SM\times\mathbb{R}, and the set

ER,k:={((v,s),X)𝒞R,k×TSM:XK(v,s)}E_{R,k}:=\left\{\big{(}(v,s),X\big{)}\in\mathcal{C}_{R,k}\times TSM:X\in K_{(v,s)}\right\}

is a smooth vector bundle of rank kk over 𝒞R,k\mathcal{C}_{R,k}.

Proof.

For the first point, it is enough to show that each point in 𝒞R,k\mathcal{C}_{R,k} has a neighborhood UU in SM×SM\times\mathbb{R} such that 𝒞R,kU\mathcal{C}_{R,k}\cap U is an embedded submanifold of dimension 2n12n-1. To prove this local statement, we will extend the methods of [War65].

Fix (v,s)𝒞R,k(v,s)\in\mathcal{C}_{R,k}. Let dFexpd_{F}\exp be the differential in the fiber variables of the exponential map exp:TMM\exp:TM\to M. By [War65], we can find coordinate neighborhoods W1W_{1} of svsv in TMTM and W2W_{2} of exp(sv)\exp(sv) in MM such that the (k1)(k-1)st elementary symmetric polynomial in the eigenvalues of dFexpd_{F}\exp (denoted by σk1\sigma_{k-1}) has nonzero derivative in the radial direction. Then σk11(0)\sigma_{k-1}^{-1}(0) is an embedded (2n1)(2n-1)-dimensional submanifold of TM{0}TM\setminus\{0\}, and it equals the set of vectors in W1W_{1} with conjugate points of order kk or higher in W2W_{2}.

Choose a neighborhood UU of (v,s)(v,s) in SM×SM\times\mathbb{R} such that all other conjugate pairs in UU have order kk. Supposing without loss of generality that s>0s>0, we may assume USM×(0,)U\subset SM\times(0,\infty). Consider the smooth map f:TM{0}SM×f:TM\setminus\{0\}\to SM\times\mathbb{R} defined by

f(w)=(w|w|g,|w|g).f(w)=\left(\frac{w}{|w|_{g}},|w|_{g}\right).

Then ff is a smooth immersion and satisfies

f(σk11(0)f1(U))=𝒞R,kU.f\left(\sigma_{k-1}^{-1}(0)\cap f^{-1}(U)\right)=\mathcal{C}_{R,k}\cap U.

Since f|σk11(0)f1(U)f|_{\sigma_{k-1}^{-1}(0)\cap f^{-1}(U)} has a continuous inverse defined on its image by

𝒞R,kU(v~,s~)s~v~TM{0},\mathcal{C}_{R,k}\cap U\ni(\tilde{v},\tilde{s})\mapsto\tilde{s}\tilde{v}\in TM\setminus\{0\},

it follows that 𝒞R,kU\mathcal{C}_{R,k}\cap U is an embedded submanifold of dimension 2n12n-1.

To prove the second point, let 𝒱R,k\mathcal{V}_{R,k} be the pullback of ker(dπ)\ker(d\pi) by the map

𝒞R,k(v,s)vSM.\mathcal{C}_{R,k}\ni(v,s)\mapsto v\in SM.

Then ER,kE_{R,k} is the kernel of the smooth bundle homomorphism

𝒱R,k((v,s),X)(Φ(v,s),dπ|Φ(v,s)dvΦ|(v,s)X)TSM.\mathcal{V}_{R,k}\ni\big{(}(v,s),X\big{)}\mapsto\left(\Phi(v,s),\,d\pi|_{\Phi(v,s)}\circ d_{v}\Phi|_{(v,s)}X\right)\in TSM.

This map has constant rank n1kn-1-k, and we can view it as a bundle homomorphism over 𝒞R,k\mathcal{C}_{R,k} by pulling back TSMTSM by Φ\Phi. Hence ER,kE_{R,k} is a smooth rank-kk subbundle of 𝒱R,k\mathcal{V}_{R,k}. ∎

Next we will turn TMTM into a symplectic manifold and make some remarks. Let ω\omega be the canonical symplectic form on TMT^{*}M, and let g:TMTM\flat_{g}:TM\to T^{*}M be the musical isomorphism induced by the metric gg. Then we can define a symplectic form ωg\omega_{g} on TMTM by

ωg(X,Y)=ω(dgX,dgY),X,YT(TM).\omega_{g}(X,Y)=\omega(d\flat_{g}X,d\flat_{g}Y),\quad X,Y\in T(TM).

Then Φ~(,s)\tilde{\Phi}(\cdot,s) is a symplectomorphism for each ss\in\mathbb{R}, the kernel of dπTM|vd\pi_{TM}|_{v} is a Lagrangian subspace of Tv(TM)T_{v}(TM) for each vTMv\in TM, and in natural coordinates (xi,vi)\left(x^{i},v^{i}\right) on TMTM we have

(8) ωg=ξgixjdxjdxi+gijdvjdxi.\omega_{g}=\xi^{\ell}\frac{\partial g_{i\ell}}{\partial x^{j}}\,dx^{j}\wedge dx^{i}+g_{ij}\,dv^{j}\wedge dx^{i}.

In turn, ωg\omega_{g} induces a smooth bundle isomorphism ωg:T(TM)T(TM)\flat_{\omega_{g}}:T(TM)\to T^{*}(TM) defined by

[ωg(X)](Y)=ωg(X,Y):=(Xωg)(Y),[\flat_{\omega_{g}}(X)](Y)=\omega_{g}(X,Y):=(X\lrcorner\,\omega_{g})(Y),

where XωgX\lrcorner\,\omega_{g} is interior multiplication by XX. We will denote the inverse of ωg\flat_{\omega_{g}} by ωg\sharp_{\omega_{g}}.

The next lemma defines a smooth bundle homomorphism that will help us describe the canonical relations of the various pieces of πL\pi_{*}\circ L.

Lemma 3.6.

For each integer 1kn11\leq k\leq n-1, there is a smooth bundle homomorphism

Fk:ER,kT(M×M)=TM×TMF_{k}:E_{R,k}\to T^{*}(M\times M)=T^{*}M\times T^{*}M

defined by the requirement that

(9) (dιSM|vXωg,(dιSM|Φ(v,s)dvΦ|(v,s)X)ωg)=dπT(M×M)|(v,Φ(v,s))tFk((v,s),X).\left(d\iota_{SM}|_{v}X\lrcorner\,\omega_{g},\left(d\iota_{SM}|_{\Phi(v,s)}\circ d_{v}\Phi|_{(v,s)}X\right)\lrcorner\,\omega_{g}\right)=d\pi_{T(M\times M)}|_{(v,\Phi(v,s))}^{t}\,F_{k}\big{(}(v,s),X\big{)}.

The proof of this result is essentially the same as that of Lemma 44 in [HU18], so we do not repeat the details here.

Our final lemma is the key geometric tool in the proof of Theorem 3.8. It will allow us to split CπCLC_{\pi_{*}}\circ C_{L} into different pieces corresponding to different orders of conjugate pairs, each of which is associated with a Fourier integral operator.

Lemma 3.7.

Let (v,s)SM×(v,s)\in SM\times\mathbb{R}, v~=Φ(v,s)\tilde{v}=\Phi(v,s), ηTπ(v)M\eta\in T_{\pi(v)}^{*}M, and η~Tπ(v~)M\tilde{\eta}\in T_{\pi(\tilde{v})}^{*}M. Then

(10) dp|(v,s)tdπ|vtη=dΦ|(v,s)tdπ|v~tη~dp|_{(v,s)}^{t}\circ d\pi|_{v}^{t}\,\eta=d\Phi|_{(v,s)}^{t}\circ d\pi|_{\tilde{v}}^{t}\,\tilde{\eta}

if and only if

(11) dπ|vtη=dvΦ|(v,s)tdπ|v~tη~andη(v)=η~(v~)=0.d\pi|_{v}^{t}\,\eta=d_{v}\Phi|_{(v,s)}^{t}\circ d\pi|_{\tilde{v}}^{t}\,\tilde{\eta}\quad\text{and}\quad\eta(v)=\tilde{\eta}(\tilde{v})=0.

If (11) holds and s0s\neq 0 then (v,s)(v,s) is a conjugate pair, and if (v,s)𝒞R,k(v,s)\in\mathcal{C}_{R,k} then (η,η~)Fk(ER,k)(\eta,\tilde{\eta})\in F_{k}(E_{R,k}). Conversely, if (η,η~)Fk(ER,k)(\eta,\tilde{\eta})\in F_{k}(E_{R,k}) then (11) holds for some (v,s)𝒞R,k(v,s)\in\mathcal{C}_{R,k}.

Proof.

To see that (10) and (11) are equivalent, first note that

dp|(v,s)tdπ|vtη\displaystyle dp|_{(v,s)}^{t}\circ d\pi|_{v}^{t}\,\eta =(dπ|vtη,0),\displaystyle=\left(d\pi|_{v}^{t}\,\eta,0\right),
dΦ|(v,s)tdπ|v~tη~\displaystyle d\Phi|_{(v,s)}^{t}\circ d\pi|_{\tilde{v}}^{t}\,\tilde{\eta} =(dvΦ|(v,s)tdπ|v~tη~,η~(v~)),\displaystyle=\left(d_{v}\Phi|_{(v,s)}^{t}\circ d\pi|_{\tilde{v}}^{t}\,\tilde{\eta},\tilde{\eta}(\tilde{v})\right),

by (3) and the fact that dπ|v~tη~(Φ˙(v,s))=η~(v~)d\pi|_{\tilde{v}}^{t}\,\tilde{\eta}(\dot{\Phi}(v,s))=\tilde{\eta}(\tilde{v}). Hence (10) holds if and only if

dπ|vtη=dvΦ|(v,s)tdπ|v~tη~andη~(v~)=0.d\pi|_{v}^{t}\,\eta=d_{v}\Phi|_{(v,s)}^{t}\circ d\pi|_{\tilde{v}}^{t}\,\tilde{\eta}\quad\text{and}\quad\tilde{\eta}(\tilde{v})=0.

Thus (11) implies (10). For the converse, just apply both sides of (10) to the vector (Φ˙(v,0),0)T(v,s)(SM×)(\dot{\Phi}(v,0),0)\in T_{(v,s)}(SM\times\mathbb{R}) and deduce that η(v)=η~(v~)\eta(v)=\tilde{\eta}(\tilde{v}).

Before addressing the claims about conjugate pairs, let us make a few observations. It will be useful to work with TMTM rather than SMSM. To connect the two, observe that

(12) dπ|vt\displaystyle d\pi|_{v}^{t} =dιSM|vtdπTM|vt,\displaystyle=d\iota_{SM}|_{v}^{t}\circ d\pi_{TM}|_{v}^{t},
(13) dvΦ|(v,s)tdιSM|v~t\displaystyle d_{v}\Phi|_{(v,s)}^{t}\circ d\iota_{SM}|_{\tilde{v}}^{t} =dιSM|vtdvΦ~|(v,s)t,\displaystyle=d\iota_{SM}|_{v}^{t}\circ d_{v}\tilde{\Phi}|_{(v,s)}^{t},

due to the identities π=πTMιSM\pi=\pi_{TM}\circ\iota_{SM} and ιSMΦ=Φ~(ιSM(),)\iota_{SM}\circ\Phi=\tilde{\Phi}(\iota_{SM}(\cdot),\cdot). Let

(14) X=(dπTM|vtη)ωgTv(TM).X=\left(d\pi_{TM}|_{v}^{t}\,\eta\right)^{\sharp_{\omega_{g}}}\in T_{v}(TM).

Equivalently, applying ωg\flat_{\omega_{g}}, we have

(15) Xωg=dπTM|vtη.X\lrcorner\,\omega_{g}=d\pi_{TM}|_{v}^{t}\,\eta.

In the next paragraph, we will prove the following analogue of the first condition in (11):

(16) dπTM|vtη=dvΦ~|(v,s)tdπTM|v~tη~.d\pi_{TM}|_{v}^{t}\,\eta=d_{v}\tilde{\Phi}|_{(v,s)}^{t}\circ d\pi_{TM}|_{\tilde{v}}^{t}\,\tilde{\eta}.

Assuming (16) for the moment, the fact that Φ~(,s)\tilde{\Phi}(\cdot,s) is a symplectomorphism implies

(17) dπTM|v~tη~=dvΦ~|(v,s)Xωg.d\pi_{TM}|_{\tilde{v}}^{t}\,\tilde{\eta}=d_{v}\tilde{\Phi}|_{(v,s)}X\lrcorner\,\omega_{g}.

Now suppose (11) holds and s0s\neq 0. We will divide the proof that (v,s)(v,s) is a conjugate pair into three steps. The first one is to prove (16). By (12), (13), and (11),

dιSM|vtdπTM|vtη=dιSM|vtdvΦ~|(v,s)tdπTM|v~tη~.d\iota_{SM}|_{v}^{t}\circ d\pi_{TM}|_{v}^{t}\,\eta=d\iota_{SM}|_{v}^{t}\circ d_{v}\tilde{\Phi}|_{(v,s)}^{t}\circ d\pi_{TM}|_{\tilde{v}}^{t}\,\tilde{\eta}.

Since ker(dιSM|vt)\ker(d\iota_{SM}|_{v}^{t}) is the span of the differential of TMw|w|g2TM\ni w\mapsto|w|_{g}^{2}, this means

dπTM|vtη=dvΦ~|(v,s)tdπTM|v~tη~+τ2d(|w|g2)|vd\pi_{TM}|_{v}^{t}\,\eta=d_{v}\tilde{\Phi}|_{(v,s)}^{t}\circ d\pi_{TM}|_{\tilde{v}}^{t}\,\tilde{\eta}+\frac{\tau}{2}\left.d(|w|_{g}^{2})\right|_{v}

for some τ\tau\in\mathbb{R}. Applying both sides to a radial vector rTv(TM)r\in T_{v}(TM), we find

0=η~(dπTM|v~dvΦ~|(v,s)r)+τ.0=\tilde{\eta}\left(d\pi_{TM}|_{\tilde{v}}\circ d_{v}\tilde{\Phi}|_{(v,s)}\,r\right)+\tau.

Since η~(v~)=0\tilde{\eta}(\tilde{v})=0 by assumption and the vector dπTM|v~dvΦ~|(v,s)rd\pi_{TM}|_{\tilde{v}}\circ d_{v}\tilde{\Phi}|_{(v,s)}\,r is parallel to v~\tilde{v}, this implies that τ=0\tau=0 and hence completes the proof of (16).

The second step is to show that (14) is in Range(dιSM|v)\operatorname{Range}(d\iota_{SM}|_{v}). Since (15) implies that XωgX\lrcorner\,\omega_{g} vanishes on ker(dπTM|v)\ker(d\pi_{TM}|_{v}), which is a Lagrangian subspace, XX must be in ker(dπTM|v)\ker(d\pi_{TM}|_{v}). Choose normal coordinates (xi)\left(x^{i}\right) on MM centered at π(v)\pi(v) such that

(18) v=dπTM|vx1,v=d\pi_{TM}|_{v}\,\frac{\partial}{\partial x^{1}},

and let (xi,vi)\left(x^{i},v^{i}\right) be natural coordinates on TMTM. Then Range(dιSM|v)\operatorname{Range}(d\iota_{SM}|_{v}) is the span of the vectors /v2,,/vn\partial/\partial v^{2},\dots,\partial/\partial v^{n}. Since XX is in ker(dπTM|v)\ker(d\pi_{TM}|_{v}), we can write

X=ajvjX=a^{j}\frac{\partial}{\partial v^{j}}

for some aja^{j}\in\mathbb{R}. Using (15), (18), and the assumption η(v)=0\eta(v)=0, we find

ωg(X,x1)=0.\omega_{g}\left(X,\frac{\partial}{\partial x^{1}}\right)=0.

Using (8) and the fact that gij=δijg_{ij}=\delta_{ij} at π(v)\pi(v), this implies a1=0a^{1}=0. Hence XX is in Range(dιSM|v)\operatorname{Range}(d\iota_{SM}|_{v}), so we can define the vector dιSM|v1XTvSMd\iota_{SM}|_{v}^{-1}X\in T_{v}SM.

The third step is to show that dιSM|v1Xd\iota_{SM}|_{v}^{-1}X is in K(v,s)K_{(v,s)}, meaning

dιSM|v1Xker(dπ|v~dvΦ|(v,s))ker(dπ|v).d\iota_{SM}|_{v}^{-1}X\in\ker\left(d\pi|_{\tilde{v}}\circ d_{v}\Phi|_{(v,s)}\right)\cap\ker(d\pi|_{v}).

Since XX is in ker(dπTM|v)\ker(d\pi_{TM}|_{v}) and dπ|v=dπTM|vdιSM|vd\pi|_{v}=d\pi_{TM}|_{v}\circ d\iota_{SM}|_{v}, we know dιSM|v1Xd\iota_{SM}|_{v}^{-1}X is in ker(dπ|v)\ker(d\pi|_{v}). Hence it is enough to prove that

dπ|v~dvΦ|(v,s)dιSM|v1X=0.d\pi|_{\tilde{v}}\circ d_{v}\Phi|_{(v,s)}\circ d\iota_{SM}|_{v}^{-1}X=0.

By the transposes of (12) and (13), this is equivalent to showing

dπTM|v~dvΦ~|(v,s)X=0.d\pi_{TM}|_{\tilde{v}}\circ d_{v}\tilde{\Phi}|_{(v,s)}X=0.

But (17) implies that dvΦ~|(v,s)Xωgd_{v}\tilde{\Phi}|_{(v,s)}X\lrcorner\,\omega_{g} vanishes on the Lagrangian subspace ker(dπTM|v~)\ker(d\pi_{TM}|_{\tilde{v}}), so dvΦ~|(v,s)Xd_{v}\tilde{\Phi}|_{(v,s)}X is indeed in ker(dπTM|v~)\ker(d\pi_{TM}|_{\tilde{v}}). This proves that (v,s)(v,s) is a conjugate pair.

Now suppose (v,s)𝒞R,k(v,s)\in\mathcal{C}_{R,k}. Since dιSM|v1Xd\iota_{SM}|_{v}^{-1}X is in K(v,s)K_{(v,s)}, this means

((v,s),dιSM|v1X)ER,k.\left((v,s),d\iota_{SM}|_{v}^{-1}X\right)\in E_{R,k}.

Using (15) in the first line below and (13) (transposed) and (17) in the second, we find

(dιSM|vdιSM|v1X)ωg\displaystyle\left(d\iota_{SM}|_{v}\circ d\iota_{SM}|_{v}^{-1}X\right)\lrcorner\,\omega_{g} =Xωg=dπTM|vtη,\displaystyle=X\lrcorner\,\omega_{g}=d\pi_{TM}|_{v}^{t}\,\eta,
(dιSM|v~dvΦ|(v,s)dιSM|v1X)ωg\displaystyle\left(d\iota_{SM}|_{\tilde{v}}\circ d_{v}\Phi|_{(v,s)}\circ d\iota_{SM}|_{v}^{-1}X\right)\lrcorner\,\omega_{g} =dvΦ~|(v,s)Xωg=dπTM|v~tη~.\displaystyle=d_{v}\tilde{\Phi}|_{(v,s)}X\lrcorner\,\omega_{g}=d\pi_{TM}|_{\tilde{v}}^{t}\,\tilde{\eta}.

Hence (η,η~)=Fk((v,s),dιSM|v1X)(\eta,\tilde{\eta})=F_{k}\big{(}(v,s),d\iota_{SM}|_{v}^{-1}X\big{)}, which proves that (η,η~)Fk(ER,k)(\eta,\tilde{\eta})\in F_{k}(E_{R,k}).

Conversely, suppose (η,η~)=Fk((v,s),X)(\eta,\tilde{\eta})=F_{k}\big{(}(v,s),X\big{)}. Unpacking (9), this means

(19) dιSM|vXωg\displaystyle d\iota_{SM}|_{v}X\lrcorner\,\omega_{g} =dπTM|vtη,\displaystyle=d\pi_{TM}|_{v}^{t}\,\eta,
(20) (dιSM|v~dvΦ|(v,s)X)ωg\displaystyle\left(d\iota_{SM}|_{\tilde{v}}\circ d_{v}\Phi|_{(v,s)}X\right)\lrcorner\,\omega_{g} =dπTM|v~tη~.\displaystyle=d\pi_{TM}|_{\tilde{v}}^{t}\,\tilde{\eta}.

Let YTv~(TM)Y\in T_{\tilde{v}}(TM). Then (20) and the transpose of (13) imply that

ωg(dvΦ~|(v,s)dιSM|vX,Y)=dπTM|v~tη~(Y).\omega_{g}\left(d_{v}\tilde{\Phi}|_{(v,s)}\circ d\iota_{SM}|_{v}X,Y\right)=d\pi_{TM}|_{\tilde{v}}^{t}\,\tilde{\eta}(Y).

Using that Φ~(,s)\tilde{\Phi}(\cdot,s) is a symplectomorphism together with (19), we find

(dvΦ~|(v,s)t)1dπTM|vtη(Y)=dπTM|v~tη~(Y).\left(d_{v}\tilde{\Phi}|_{(v,s)}^{t}\right)^{-1}\circ d\pi_{TM}|_{v}^{t}\,\eta(Y)=d\pi_{TM}|_{\tilde{v}}^{t}\,\tilde{\eta}(Y).

Hence (16) holds in this direction of the proof as well. Applying dιSM|vtd\iota_{SM}|_{v}^{t} to both sides of that equation and rewriting with (12) and (13) establishes the first condition in (11).

Now take the same natural coordinates on TMTM described above (18). Then

dιSM|vX=j=2najvjd\iota_{SM}|_{v}X=\sum_{j=2}^{n}a^{j}\frac{\partial}{\partial v^{j}}

for some aja^{j}\in\mathbb{R}. By (18), (19), and (8), we get

η(v)=dπTM|vtη(x1)=(dιSM|vXωg)(x1)=ωg(dιSM|vX,x1)=0,\eta(v)=d\pi_{TM}|_{v}^{t}\,\eta\left(\frac{\partial}{\partial x^{1}}\right)=(d\iota_{SM}|_{v}X\lrcorner\,\omega_{g})\left(\frac{\partial}{\partial x^{1}}\right)=\omega_{g}\left(d\iota_{SM}|_{v}X,\frac{\partial}{\partial x^{1}}\right)=0,

and η~(v~)=0\tilde{\eta}(\tilde{v})=0 by a similar argument. This establishes the second condition in (11). ∎

3.4. Final Decomposition of the Lévy Generator

Now we can state our main result. It refines Theorem 2.5 by decomposing πRaπ\pi_{*}\circ R_{a}\circ\pi^{*} into a smoothing operator and a sum of Fourier integral operators, assuming there are no singular conjugate pairs.

Theorem 3.8.

Suppose 𝒞S=\mathcal{C}_{S}=\emptyset. Then for k=1k=1 to n1n-1, the sets

CAk=Fk(ER,k)T(M×M)C_{A_{k}}=F_{k}(E_{R,k})\subset T^{*}(M\times M)

are either local canonical relations or empty. Let CAk,1,,CAk,MkC_{A_{k,1}},\dots,C_{A_{k,M_{k}}} be the connected components of CAkC_{A_{k}}. Let 𝒜2α\mathscr{A}_{2\alpha} and 𝒜0\mathscr{A}_{0} be the pseudodifferential operators from Theorem 2.5. Then

𝒜=𝒜2α+𝒜0+𝒜+k=1n1(m=1MkAk,m),\mathscr{A}=\mathscr{A}_{2\alpha}+\mathscr{A}_{0}+\mathscr{A}_{-\infty}+\sum_{k=1}^{n-1}\left(\sum_{m=1}^{M_{k}}A_{k,m}\right),

where 𝒜\mathscr{A}_{-\infty} is a smoothing operator, and for each kk either

Ak,m(nk+1)/2(M×M,CAk,m),A_{k,m}\in\mathcal{I}^{-(n-k+1)/2}(M\times M,C_{A_{k,m}}^{\prime}),

or Mk=1M_{k}=1 and Ak,1=0A_{k,1}=0 if CAk=C_{A_{k}}=\emptyset.

Proof.

We must decompose πL\pi_{*}\circ L into a smoothing operator and a sum of Fourier integral operators. Though the clean intersection calculus does not directly apply to πL\pi_{*}\circ L, we will cut up this operator so that it applies to each separate piece.

First let us describe CπCLC_{\pi_{*}}\circ C_{L}. By Lemma 2.4 and Theorem 3.3,

CπCL={(η,η~)TM\displaystyle C_{\pi_{*}}\circ C_{L}=\Big{\{}(\eta,\tilde{\eta})\in T^{*}M ×TM:(v,s)SM×such that\displaystyle\times T^{*}M:\exists\,(v,s)\in SM\times\mathbb{R}\ \text{such that}
dp|(v,s)tdπ|vtη=dΦ|(v,s)tdπ|Φ(v,s)tη~}.\displaystyle\ dp|_{(v,s)}^{t}\circ d\pi|_{v}^{t}\,\eta=d\Phi|_{(v,s)}^{t}\circ d\pi|_{\Phi(v,s)}^{t}\,\tilde{\eta}\Big{\}}.

The requirement in CπCLC_{\pi_{*}}\circ C_{L} is precisely (10), which is equivalent to (11) by Lemma 3.7. We can also use this lemma to cut up CπCLC_{\pi_{*}}\circ C_{L} into several pieces according to different orders of conjugate pairs. Indeed, if we take s=0s=0 and any vSMv\in SM such that η(v)=0\eta(v)=0, then vv, ss, and η\eta satisfy (11), so one piece is the diagonal

Δ:={(η,η)TM×TM}.\Delta:=\{(\eta,\eta)\in T^{*}M\times T^{*}M\}.

If (η,η~)CπCL(\eta,\tilde{\eta})\in C_{\pi_{*}}\circ C_{L} and ηη~\eta\neq\tilde{\eta}, then (11) is satisfied for some (v,s)SM×(v,s)\in SM\times\mathbb{R}, and s0s\neq 0 because dπ|vtd\pi|_{v}^{t} is injective. Then Lemma 3.7 and the assumption 𝒞S=\mathcal{C}_{S}=\emptyset imply that

CπCL=Δ(k=1n1CAk).C_{\pi_{*}}\circ C_{L}=\Delta\cup\left(\bigcup_{k=1}^{n-1}C_{A_{k}}\right).

Our goal is to write πL\pi_{*}\circ L as a sum of Fourier integral operators, each having a canonical relation contained in a single set of this union. Since 𝒞S=\mathcal{C}_{S}=\emptyset, we can find open subsets UkU_{k} of SM×SM\times\mathbb{R} with disjoint closures such that 𝒞R,kUk\mathcal{C}_{R,k}\subset U_{k} for k=1k=1 to n1n-1. Write

Uk=m=1MkUk,m,U_{k}=\bigcup_{m=1}^{M_{k}}U_{k,m},

where each open set Uk,mU_{k,m} contains exactly one of the connected components of 𝒞R,k\mathcal{C}_{R,k}. Then we can construct a partition of unity {ψk,m}\{\psi_{k,m}\} on SM×SM\times\mathbb{R} such that

supp(ψ0,1)(SM×)(k=1n1𝒞R,k),\operatorname{supp}(\psi_{0,1})\subset(SM\times\mathbb{R})\setminus\left(\bigcup_{k=1}^{n-1}\mathcal{C}_{R,k}\right),

and supp(ψk,m)Uk,m\operatorname{supp}(\psi_{k,m})\subset U_{k,m} for k=1k=1 to n1n-1 and m=1m=1 to MkM_{k}. Setting M0=1M_{0}=1, we have

πL=k=0n1m=1MkπLψk,m.\pi_{*}\circ L=\sum_{k=0}^{n-1}\sum_{m=1}^{M_{k}}\pi_{*}\circ L_{\psi_{k,m}}.

By Lemma 3.2, the twisted wave front set of Lψk,mL_{\psi_{k,m}} is contained in

CL,k,m:={(ξ,η~)TSM×TM:s0such that(πTSM(ξ),s)supp(ψk,m)\displaystyle C_{L,k,m}:=\big{\{}(\xi,\tilde{\eta})\in T^{*}SM\times T^{*}M:\exists\,s\neq 0\ \text{such that}\ (\pi_{T^{*}SM}(\xi),s)\in\operatorname{supp}(\psi_{k,m})
anddp|(πTSM(ξ),s)tξ=dΦ|(πTSM(ξ),s)tdπ|Φ(πTSM(ξ),s)tη~}.\displaystyle\ \text{and}\ dp|_{(\pi_{T^{*}SM}(\xi),s)}^{t}\,\xi=d\Phi|_{(\pi_{T^{*}SM}(\xi),s)}^{t}\circ d\pi|_{\Phi(\pi_{T^{*}SM}(\xi),s)}^{t}\,\tilde{\eta}\big{\}}.

Therefore we obtain smoothing operators except near points in

(CπCL,k,m)={\displaystyle(C_{\pi_{*}}\circ C_{L,k,m})^{\prime}=\Big{\{} (η,η~)TM×TM:(v,s)SM×({0})such that\displaystyle(\eta,-\tilde{\eta})\in T^{*}M\times T^{*}M:\exists\,(v,s)\in SM\times(\mathbb{R}\setminus\{0\})\ \text{such that}
(v,s)supp(ψk,m)anddp|(v,s)tdπ|vtη=dΦ|(v,s)tdπ|Φ(v,s)tη~}.\displaystyle\ (v,s)\in\operatorname{supp}(\psi_{k,m})\ \text{and}\ dp|_{(v,s)}^{t}\circ d\pi|_{v}^{t}\,\eta=d\Phi|_{(v,s)}^{t}\circ d\pi|_{\Phi(v,s)}^{t}\,\tilde{\eta}\Big{\}}.

Since Lemma 3.7 implies that CπCL,0,1C_{\pi_{*}}\circ C_{L,0,1} is empty,

𝒜:=πLψ0,1\mathscr{A}_{-\infty}:=\pi_{*}\circ L_{\psi_{0,1}}

is a smoothing operator. Lemma 3.7 also implies that CπCL,k,m=CAk,mC_{\pi_{*}}\circ C_{L,k,m}=C_{A_{k,m}} for k=1k=1 to n1n-1 and m=1m=1 to MkM_{k}. If any of these compositions are empty, we can absorb the corresponding operator into 𝒜\mathscr{A}_{-\infty} and set Ak,m=0A_{k,m}=0.

It remains to show that if CπCL,k,mC_{\pi_{*}}\circ C_{L,k,m} is nonempty, then the clean intersection calculus applies to πLψk,m\pi_{*}\circ L_{\psi_{k,m}}. We again refer to points (i)-(ii) in the proof of Theorem 3.1. (Since we omit point (iii), we only obtain local canonical relations in general.) Define

Ck,m=(Cπ×CL,k,m)(TM×Δ(TSM)×TM).C_{k,m}=(C_{\pi_{*}}\times C_{L,k,m})\cap(T^{*}M\times\Delta(T^{*}SM)\times T^{*}M).

Let ER,k,mE_{R,k,m} be the restriction of ER,kE_{R,k} to the mmth connected component of 𝒞R,k\mathcal{C}_{R,k}, and let FkF_{k}^{\ell} be the \ellth component function of FkF_{k} for =1,2\ell=1,2. Then Lemma 3.7 implies that Ck,mC_{k,m} is a connected component of

(21) {(Fk1((v,s),X),dιSM|vt(dιSM|vXωg),dιSM|vt(dιSM|vXωg),Fk2((v,s),X)):\displaystyle\Big{\{}\left(F_{k}^{1}\big{(}(v,s),X\big{)},d\iota_{SM}|_{v}^{t}(d\iota_{SM}|_{v}X\lrcorner\,\omega_{g}),d\iota_{SM}|_{v}^{t}(d\iota_{SM}|_{v}X\lrcorner\,\omega_{g}),F_{k}^{2}\big{(}(v,s),X\big{)}\right):
((v,s),X)ER,k,m}.\displaystyle\big{(}(v,s),X\big{)}\in E_{R,k,m}\Big{\}}.

Thus, to show Ck,mC_{k,m} is an embedded submanifold, it is enough to prove that

(22) ER,k,m((v,s),X)(v,dιSM|vt(dιSM|vXωg))TSME_{R,k,m}\ni\big{(}(v,s),X\big{)}\mapsto\left(v,d\iota_{SM}|_{v}^{t}(d\iota_{SM}|_{v}X\lrcorner\,\omega_{g})\right)\in T^{*}SM

is a smooth embedding. Let pk,mp_{k,m} be the projection of the mmth connected component of 𝒞R,k\mathcal{C}_{R,k} onto SMSM. Let 𝒱R,k,m\mathcal{V}_{R,k,m} be the pullback of ker(dπ)\ker(d\pi) by pk,mp_{k,m}. Then ER,k,mE_{R,k,m} is a smooth subbundle of 𝒱R,k,m\mathcal{V}_{R,k,m} (as in the proof of Proposition 3.5), so it suffices to show that the extension of (22) to 𝒱R,k,m\mathcal{V}_{R,k,m} is a smooth embedding. Observe that the bundle homomorphism

ER,k,m((v,s),X)(v,X)TSME_{R,k,m}\ni\big{(}(v,s),X\big{)}\mapsto(v,X)\in TSM

covers pk,mp_{k,m} and is a smooth embedding (because it is a proper injective immersion). This implies that pk,mp_{k,m} is a smooth embedding, so 𝒱R,k,m\mathcal{V}_{R,k,m} is smoothly isomorphic to the restriction of ker(dπ)\ker(d\pi) to Range(pk,m)\operatorname{Range}(p_{k,m}). Hence it suffices to show that

ker(dπ)|Range(pk,m)(v,X)(v,dιSM|vt(dιSM|vXωg))TSM\ker(d\pi)|_{\operatorname{Range}(p_{k,m})}\ni(v,X)\mapsto\left(v,d\iota_{SM}|_{v}^{t}(d\iota_{SM}|_{v}X\lrcorner\,\omega_{g})\right)\in T^{*}SM

is a smooth embedding. But (8) implies that this map is injective in each fiber, and hence a smooth bundle isomorphism onto its image. Therefore (22) is a smooth embedding, which proves that Ck,mC_{k,m} is a connected embedded submanifold of dimension 2n+k12n+k-1.

Now let c=(c1,c2,c3,c4)Ck,mc=(c_{1},c_{2},c_{3},c_{4})\in C_{k,m} be arbitrary, and consider the set

Dk,m\displaystyle D_{k,m} :=Tc(Cπ×CL,k,m)Tc(TM×Δ(TSM)×TM)\displaystyle:=T_{c}(C_{\pi_{*}}\times C_{L,k,m})\cap T_{c}(T^{*}M\times\Delta(T^{*}SM)\times T^{*}M)
Tc1(TM)×Tc2(TSM)×Tc3(TSM)×Tc4(TM).\displaystyle\subset T_{c_{1}}(T^{*}M)\times T_{c_{2}}(T^{*}SM)\times T_{c_{3}}(T^{*}SM)\times T_{c_{4}}(T^{*}M).

Since TcCk,mT_{c}C_{k,m} has dimension 2n+k12n+k-1 and is contained in Dk,mD_{k,m}, the intersection is clean if the dimension of Dk,mD_{k,m} is at most 2n+k12n+k-1. Suppose (Y1,Y2,Y3,Y4)Dk,m(Y_{1},Y_{2},Y_{3},Y_{4})\in D_{k,m}. Then Y2=Y3Y_{2}=Y_{3}, and an examination of CL,k,mC_{L,k,m} shows that Y3Y_{3} determines Y4Y_{4}. Hence the dimension of Dk,mD_{k,m} is at most 3n13n-1 (the dimension of CπC_{\pi_{*}}). But for fixed (v,s)(v,s), the set

{dιSM|vt(dιSM|vXωg):((v,s),X)ER,k,m}\left\{d\iota_{SM}|_{v}^{t}(d\iota_{SM}|_{v}X\lrcorner\,\omega_{g}):\big{(}(v,s),X\big{)}\in E_{R,k,m}\right\}

is a kk-dimensional vector space contained in

{dπ|vtη:ηTπ(v)M}.\left\{d\pi|_{v}^{t}\,\eta:\eta\in T_{\pi(v)}^{*}M\right\}.

Therefore the dimension of Dk,mD_{k,m} is at most 3n1(nk)=2n+k13n-1-(n-k)=2n+k-1, so the intersection is clean with excess k1k-1. Because Ck,mC_{k,m} is a component of (21), the projection map

πk,m:Ck,mTM×TM\pi_{k,m}:C_{k,m}\to T^{*}M\times T^{*}M

is proper by the same argument as the proof of Theorem 3.1. Since π\pi_{*} and Lψk,mL_{\psi_{k,m}} are Fourier integral operators of order (1n)/4(1-n)/4 and (n+1)/4-(n+1)/4, respectively, we conclude that

Ak,m:=πLψk,mA_{k,m}:=\pi_{*}\circ L_{\psi_{k,m}}

is in (nk+1)/2(M×M,CAk,m)\mathcal{I}^{-(n-k+1)/2}(M\times M,C_{A_{k,m}}^{\prime}) whenever the set CAk,mC_{A_{k,m}} is nonempty. ∎

Two special cases of Theorem 3.8 are worth mentioning. First, if (M,g)(M,g) is Anosov then it has no conjugate points [Rug91], so each Fourier integral operator Ak,mA_{k,m} is zero and we recover Theorem 1.61.6 in [Cha+22]. Second, Theorem 3.8 covers all possibilities in two dimensions because singular conjugate pairs cannot exist (since conjugate pairs can only have order 11). In higher dimensions, the generic case includes singular conjugate pairs [Arn72, Klo83].

References

  • [Ano69] D.V.“@ Anosov “Geodesic flows on closed Riemann manifolds with negative curvature” American Mathematical Society, 1969
  • [AE00] D.“@ Applebaum and A.“@ Estrade “Isotropic Lévy processes on Riemannian manifolds” In Ann. Probab. JSTOR, 2000, pp. 166–184
  • [Arn72] V.I.“@ Arnol’d “Normal forms for functions near degenerate critical points, the Weyl groups of AkA_{k}, DkD_{k}, EkE_{k} and Lagrangian singularities” In Funct. Anal. App. 6 Springer, 1972, pp. 254–272
  • [BN13] P.C.“@ Bressloff and J.M.“@ Newby “Stochastic models of intracellular transport” In Rev. Mod. Phys. 85.1 APS, 2013, pp. 135
  • [Cha+22] Y.“@ Chaubet, Y.G.“@ Bonthonneau, T.“@ Lefeuvre and L.“@ Tzou “Geodesic Lévy Flights and Expected Stopping Time for Random Searches” In arXiv preprint arXiv:2211.13973, 2022
  • [DG75] J.J.“@ Duistermaat and V.“@ Guillemin “The spectrum of positive operators and periodic geodesies” In Invent. Math 29, 1975, pp. 39–79
  • [Elw88] K.D.“@ Elworthy “Geometric aspects of diffusions on manifolds”, 1988 Springer
  • [Gan64] R.“@ Gangolli “Isotropic infinitely divisible measures on symmetric spaces” In Acta Math. 111, 1964, pp. 213–46
  • [Gan65] R.“@ Gangolli “Sample functions of certain differential processes on symmetric spaces” In Pac. J. Math. 15.2 Mathematical Sciences Publishers, 1965, pp. 477–496
  • [GS75] V.“@ Guillemin and D.“@ Schaeffer “Fourier integral operators from the Radon transform point of view” In Proc. Symp. Pure Math. 27, 1975, pp. 297–300
  • [HP17] A.A.“@ Heidari and P.“@ Pahlavani “An efficient modified grey wolf optimizer with Lévy flight for optimization tasks” In Appl. Soft Comput. 60 Elsevier, 2017, pp. 115–134
  • [HU18] S.“@ Holman and G.“@ Uhlmann “On the microlocal analysis of the geodesic X-ray transform with conjugate points” In J. Differ. Geom. 108.3 Lehigh University, 2018, pp. 459–494
  • [Hör85] L.“@ Hörmander “The analysis of linear partial differential operators IV: Fourier integral operators” Springer, 1985
  • [Hsu02] E.P.“@ Hsu “Stochastic analysis on manifolds” American Mathematical Society, 2002
  • [Hun56] G.A.“@ Hunt “Semi-groups of measures on Lie groups” In Trans. Am. Math. Soc. 81.2, 1956, pp. 264–293
  • [KKM22] W.“@ Kaidi, M.“@ Khishe and M.“@ Mohammadi “Dynamic levy flight chimp optimization” In Knowl.-Based Syst. 235 Elsevier, 2022, pp. 107625
  • [Klo83] F.“@ Klok “Generic singularities of the exponential map on Riemannian manifolds” In Geom. Dedicata 14.4 Springer, 1983, pp. 317–342
  • [Kni02] G.“@ Knieper “Hyperbolic dynamics and Riemannian geometry” In Handbook of dynamical systems 1 Elsevier, 2002, pp. 239–319
  • [MSU15] F.“@ Monard, P.“@ Stefanov and G.“@ Uhlmann “The geodesic ray transform on Riemannian surfaces with conjugate points” In Commun. Math. Phys. 337.3 Springer, 2015, pp. 1491–1513
  • [Nur+23] M.“@ Nursultanov, W.“@ Trad, J.C.“@ Tzou and L.“@ Tzou “The narrow capture problem on general Riemannian surfaces” In Differ. Integral Equ. 36.11/12 Khayyam Publishing, Inc. West Palm Beach, FL, USA, 2023, pp. 877–906
  • [NTT22] M.“@ Nursultanov, W.“@ Trad and L.“@ Tzou “Narrow escape problem in the presence of the force field” In Math. Methods Appl. Sci. 45.16 Wiley Online Library, 2022, pp. 10027–10051
  • [NTT21] M.“@ Nursultanov, J.C.“@ Tzou and L.“@ Tzou “On the mean first arrival time of Brownian particles on Riemannian manifolds” In J. Math. Pures Appl. 150 Elsevier, 2021, pp. 202–240
  • [Rug91] R.O.“@ Ruggiero “On the creation of conjugate points” In Math. Z. 208 Springer, 1991, pp. 41–55
  • [SK86] M.F.“@ Shlesinger and J.“@ Klafter “Lévy walks versus Lévy flights” In On growth and form: Fractal and non-fractal patterns in physics Springer, 1986, pp. 279–283
  • [SU12] P.“@ Stefanov and G.“@ Uhlmann “The geodesic X-ray transform with fold caustics” In Anal. PDE 5.2 Mathematical Sciences Publishers, 2012, pp. 219–260
  • [TT23] J.C.“@ Tzou and L.“@ Tzou “Challenging the Lévy Flight Foraging Hypothesis-A Joint Monte Carlo and Numerical PDE Approach” In arXiv preprint arXiv:2302.13976, 2023
  • [Vis+96] G.M.“@ Viswanathan, V.“@ Afanasyev, S.V.“@ Buldyrev, E.J.“@ Murphy, P.A.“@ Prince and H.E.“@ Stanley “Lévy flight search patterns of wandering albatrosses” In Nature 381.6581 Nature Publishing Group UK London, 1996, pp. 413–415
  • [Vis+99] G.M.“@ Viswanathan, S.V.“@ Buldyrev, S.“@ Havlin, M.G.E.“@ Da Luz, E.P.“@ Raposo and H.E.“@ Stanley “Optimizing the success of random searches” In Nature 401.6756 Nature Publishing Group UK London, 1999, pp. 911–914
  • [War65] F.W.“@ Warner “The conjugate locus of a Riemannian manifold” In Am. J. Math 87.3 JSTOR, 1965, pp. 575–604
  • [Wei75] A.“@ Weinstein “On Maslov’s quantization condition, Fourier Integral Operators and Partial Differential Equations, J. Chazarain, ed.” Springer-Verlag, BerlinNew York, 1975
  • [YD09] X.S.“@ Yang and S.“@ Deb “Cuckoo search via Lévy flights” In 2009 World congress on nature & biologically inspired computing (NaBIC), 2009, pp. 210–214 Ieee
  • [YD10] X.S.“@ Yang and S.“@ Deb “Eagle strategy using Lévy walk and firefly algorithms for stochastic optimization” In Nature inspired cooperative strategies for optimization (NICSO 2010) Springer, 2010, pp. 101–111