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

Witten deformation for noncompact manifolds with bounded geometry

Xianzhe Dai Department of Mathematics, UCSB, Santa Barbara CA 93106, [email protected].          Partially supported by the Simons Foundation    Junrong Yan111Department of Mathematics, UCSB, Santa Barbara CA 93106, [email protected]
Abstract

Motivated by the Landau-Ginzburg model, we study the Witten deformation on a noncompact manifold with bounded geometry, together with some tameness condition on the growth of the Morse function ff near infinity. We prove that the cohomology of the Witten deformation dTfd_{Tf} acting on the complex of smooth L2L^{2} forms is isomorphic to the cohomology of Thom-Smale complex of ff as well as the relative cohomology of a certain pair (M,U)(M,U) for sufficiently large TT. We establish an Agmon estimate for eigenforms of the Witten Laplacian which plays an essential role in identifying these cohomologies via Witten’s instanton complex, defined in terms of eigenspaces of the Witten Laplacian for small eigenvalues. As an application we obtain the strong Morse inequalities in this setting.

1 Introduction

1.1 Overview

In the extremely influential paper [19], Witten introduced a deformation of the de Rham complex by considering the new differential df=d+df,d_{f}=d+df, where dd is the usual exterior derivative on forms, and ff is a Morse function. Setting

dTf:=d+Tdf,d_{Tf}:=d+Tdf,

Witten observed that when T>0T>0 is large enough, the eigenfunctions of the small eigenvalues for the corresponding deformed Hodge-Laplacian, the so called Witten Laplacian, concentrate at the critical points of ff. As a result, Witten deformation builds a direct bridge between the Betti numbers and the Morse indices of the critical points of ff.

Witten deformation on closed manifolds has produced a whole range of beautiful applications, from Demailly’s holomorphic Morse inequalities[14], to the proof of Ray-Singer conjecture and its generalization by Bismut-Zhang [2], to the instigation of the development of Floer homology theory.

Although the Witten deformation on noncompact manifolds are much less studied and understood, there are previous interesting work in the direction. In [4] the cohomology of an affine algebraic variety is related to that of the Witten complex of m\mathbb{C}^{m}, see also [7] for further development.

This paper is motivated by the study of Landau-Ginzburg models (c.f.[12]), which, according to Witten [20], are simply different phase of Calabi-Yau manifolds, and hence equivalent to Calabi-Yau manifolds. Suppose there is a non-trivial holomorphic function WW (the superpotential) on a noncompact Kahler manifold MnM^{n} (n=dimMn=\dim_{\mathbb{C}}M), then one considers the Witten-deformation of \partial operator:

¯W=¯i2W,\bar{\partial}_{W}=\bar{\partial}-\frac{i}{2}\partial W,

as its cohomology describes the quantum ground states of the Landau-Ginzburg model (M,W)(M,W). If WW is also a Morse function with kk critical points, then complex Morse theoretic consideration leads to the expectation that

H¯Wl(M)={k, if l=n0, otherwise.H^{l}_{\bar{\partial}_{W}}(M)=\begin{cases}\mathbb{C}^{k},\mbox{ if }l=n\\ 0,\mbox{ otherwise}.\end{cases}

For the mathematical study of LG models and their significant applications we point out the important work [9].

In this paper, we consider the more general case for Riemannian manifolds: we explore the relations between the Thom-Smale complex for a Morse function ff on a noncompact manifold MM and the deformed de Rham complex with respect to ff. The first difficulty one encounters here is the presence of continuous spectrum on a noncompact manifolds and for that one has to impose certain tameness conditions. This consists of the bounded geometry requirement for the manifold as well as growth conditions for the function. The notion of strong tameness is introduced in [5] in the Kähler setting which guarantees the discreteness of the spectrum for the Witten Laplacian. Here we introduce a slightly weaker notion which allows continuous spectrum but only outside a large interval starting from 0.

It is important to note that, and this is another new phenomenon in the noncompact case, the Thom-Smale complex may not be a complex in general. Namely, the square of its boundary operator need not be zero, since MM is noncompact. However we prove that with the tameness condition, it is.

The crucial technical part of our work is the Agmon estimate for eigenforms of the Witten Laplacian which is essential in extending the usual analysis from compact setting to the noncompact case. The Agmon estimate was discovered by S. Agmon in his study of NN-body Schrödinger operators in the Euclidean setting and has found many important applications. The exponential decay of the eigenfunction is expressed in terms of the so-called Agmon distance, Cf. [1]. We make essential use of this Agmon estimate to carry out the isomorphism between the Witten instanton complex defined in terms of eigenspaces corresponding to the small eigenvalues with the Thom-Smale complex defined in terms of the critical point data of the function. We remark that the Agmon estimate near the critical points also plays important role in the compact case , see, e.g. [2]. The novelty here is that we make essential use of the exponential decay at spatial infinity provided by the Agmon estimate.

As an application of our results on noncompact manifolds, we deduce corresponding results for manifolds with boundaries which generalize recent work of [16], [lu2017thomsmale].

Finally we would also like to point out the preprint [6] which has provided further motivation and inspiration for us.

In the rest of the introduction we give precise statements of our main results after setting up our notations. In subsequent work we will develop the local index theory and the Ray-Singer torsion for the Witten deformation in the noncompact setting.

Acknowledgment: We thank Shu Shen for interesting discussions and for providing us with an example of Thom-Smale complex not being a complex.

1.2 Notations and basic setup

Let (M,g)(M,g) be a noncompact connected complete Riemannian manifold with metric gg. (M,g)(M,g) is said to have bounded geometry, if the following conditions hold:

  1. 1.

    the injectivity radius r0r_{0} of MM is positive.

  2. 2.

    |mR|Cm|\nabla^{m}R|\leq C_{m}, where mR\nabla^{m}R is the mm-th covariant derivative of the curvature tensor and CmC_{m} is a constant only depending on mm.

On such a manifold, the Sobolev constant is uniformly bounded, see e.g. [13]. Now let f:Mf:M\mapsto\mathbb{R} be a smooth function. In [5], the notion of strong tameness for the triple (M,g,f)(M,g,f) is introduced.

Definition 1.1.

The triple (M,g,f)(M,g,f) is said to be strongly tame, if (M,g)(M,g) has bounded geometry and

limsupp|2f|(p)|f|2(p)=0,\lim\sup_{p\to\infty}\frac{|\nabla^{2}f|(p)}{|\nabla f|^{2}(p)}=0,

and

limp|f|,\lim_{p\to\infty}|\nabla f|\to\infty,

where f,2f\nabla f,\nabla^{2}f are the gradient and Hessian of ff respectively.

Remark 1.2.

Fix p0Mp_{0}\in M, and let dd be the distance function induced by gg. Here pp\to\infty simply means that d(p,p0).d(p,p_{0})\to\infty.

In this paper we only need the following weaker condition.

Definition 1.3.

The triple (M,g,f)(M,g,f) is said to be well tame, if (M,g)(M,g) has bounded geometry and

cf:=limsupp|2f|(p)|f|2(p)<,c_{f}:=\lim\sup_{p\to\infty}\frac{|\nabla^{2}f|(p)}{|\nabla f|^{2}(p)}<\infty,

and

ϵf:=liminfp|f|>0.\epsilon_{f}:=\lim\inf_{p\to\infty}|\nabla f|>0.

As usual, the metric gg induced a canonical metric (still denote it by gg) on Λ(M)\Lambda^{*}(M), which then defines an inner product (,)L2(\cdot,\cdot)_{L^{2}} on Ωc(M)\Omega^{*}_{c}(M):

(ϕ,ψ)L2=M(ϕ,ψ)g𝑑vol,ϕ,ψΩc(M).(\phi,\psi)_{L^{2}}=\int_{M}(\phi,\psi)_{g}dvol,\phi,\psi\in\Omega^{*}_{c}(M).

Let L2Λ(M)L^{2}\Lambda^{*}(M) be the completion of Ωc(M)\Omega_{c}^{*}(M) with respect to L2\|\cdot\|_{L^{2}}, and for simplicity, we denote L2(M):=L2Λ0(M).L^{2}(M):=L^{2}\Lambda^{0}(M).

For any T0T\geq 0, let dTf:=d+Tdf:Ω(M)Ω+1(M)d_{Tf}:=d+Tdf\wedge:\Omega^{*}(M)\mapsto\Omega^{*+1}(M) be the so-called Witten deformation of de Rham operator dd. It is an unbounded operator on L2Λ(M)L^{2}\Lambda^{*}(M) with domain Ωc(M)\Omega^{*}_{c}(M). Also, dTfd_{Tf} has a formal adjoint operator δTf\delta_{Tf}, with Dom(δTf)=Ωc(M),\mathrm{Dom}(\delta_{Tf})=\Omega^{*}_{c}(M), such that

(dTfϕ,ψ)L2=(ϕ,δTfψ)L2,ϕ,ψΩc(M).(d_{Tf}\phi,\psi)_{L^{2}}=(\phi,\delta_{Tf}\psi)_{L^{2}},\phi,\psi\in\Omega^{*}_{c}(M).

Set ΔH,Tf=(dTf+δTf)2,\Delta_{H,Tf}=(d_{Tf}+\delta_{Tf})^{2}, and we denote the Friedrichs extension of ΔH,Tf\Delta_{H,Tf} by Tf\Box_{Tf}. As we will see (Theorem 2.1), if (M,g,f)(M,g,f) is well tame, then ΔH,Tf\Delta_{H,Tf} is essentially self-adjoint (and hence Tf\Box_{Tf} is the unique self-adjoint extension). In Section 8, we will prove the Hodge-Kodaira decomposition when (M,g,f)(M,g,f) is well tame and TT large enough,

L2Λ(M)=kerTfImd¯TfImδ¯Tf,L^{2}\Lambda^{*}(M)=\ker\Box_{Tf}\oplus\mathrm{Im}\bar{d}_{Tf}\oplus\mathrm{Im}\bar{\delta}_{Tf}, (1.1)

where d¯Tf\bar{d}_{Tf} and δ¯Tf\bar{\delta}_{Tf} are the graph extensions of dTfd_{Tf} and δTf\delta_{Tf} respectively.

Setting Ω(2)(M):=L2Λ(M)Ω(M),\Omega_{(2)}^{*}(M):=L^{2}\Lambda^{*}(M)\cap\Omega^{*}(M), we have a chain complex (of unbounded operators)

dTfΩ(2)(M)dTfΩ(2)+1(M)dTf.\cdots\xrightarrow{d_{Tf}}\Omega_{(2)}^{*}(M)\xrightarrow{d_{Tf}}\Omega_{(2)}^{*+1}(M)\xrightarrow{d_{Tf}}\cdots.

Let H(2)(M,dTf)H^{*}_{(2)}(M,d_{Tf}) denote the cohomology of this complex. In Section 8, we will show that H(2)(M,dTf)kerTfH^{*}_{(2)}(M,d_{Tf})\cong\ker\Box_{Tf}, provided (M,g,f)(M,g,f) is well tame and TT is large enough.

Finally we note the following well known (Cf. [19, 21])

Proposition 1.4.

The Hodge Laplacian ΔH,Tf\Delta_{H,Tf} has the following local expression:

ΔH,Tf=ΔTei,ej2f[ei,ιej]+T2|f|2.\Delta_{H,Tf}=\Delta-T\nabla^{2}_{e_{i},e_{j}}f[e^{i}\wedge,\iota_{e_{j}}]+T^{2}|\nabla f|^{2}. (1.2)

Here {ei}\{e_{i}\} is a local frame on TMTM and {ei}\{e^{i}\} is the dual frame on TM.T^{*}M.

1.3 Main results

In this subsection, we assume that (M,g)(M,g) has bounded geometry, ff is a Morse function with finite many critical points. Clearly this will be the case if (M,g,f)(M,g,f) is well tame and ff is Morse.

As we mentioned the main technical result here is the Agmon estimate for the eigenforms of the Witten Laplacian.

Theorem 1.1.

Let (M,g,f)(M,g,f) be well tame, and ωDom(Tf)\omega\in\mathrm{Dom}(\Box_{Tf}) be an eigenform of Tf\Box_{Tf} whose eigenvalue is uniformly bounded in TT. Then

|ω(p)|CT(n+2)/2exp(aρT(p))ωL2,|\omega(p)|\leq CT^{(n+2)/2}\exp(-a\rho_{T}(p))\|\omega\|_{L^{2}},

for any a(0,1)a\in(0,1) (provided TT is sufficiently large and CC is a constant depending on the dimension nn, the function ff, the curvature bound, and aa; for the precise choice of T,CT,C see the end of Section 3). Here the definition of the Agmon distance ρT(p)\rho_{T}(p) will be given in Section 3.

The proof of the Agmon estimate, given in Section 7, is to carry out the idea of [1] in this more general setting.

Set bi(T)=dimH(2)i(M,dTf)b_{i}(T)=\dim H_{(2)}^{i}(M,d_{Tf}). If xx is a critical point of ff, denote nf(x)n_{f}(x) the Morse index of ff at x.x. Let mim_{i} be the number of critical points of ff with Morse index i.i. Then the strong Morse inequalities hold.

Theorem 1.2.

If (M,g,f)(M,g,f) is well tame, then we have the following strong Morse inequality

(1)ki=0k(1)ibi(T)(1)ki=0k(1)imi,kn,(-1)^{k}\sum_{i=0}^{k}(-1)^{i}b_{i}(T)\leq(-1)^{k}\sum_{i=0}^{k}(-1)^{i}m_{i},\ \ \forall k\leq n,

provided TT is large enough. And the equality holds for k=nk=n.

In general, bi(T)b_{i}(T) may be very sensitive to TT. However we have the following result regarding the indepedence of bi(T)b_{i}(T) in TT. Assume that the Morse function ff satisfies the Smale transversality condition. Let (C(Wu),~)(C^{*}(W^{u}),\tilde{\partial}^{\prime}) be the Thom-Smale complex given by ff. It is important to note that in general, since MM is noncompact, it could happen that (~)20(\tilde{\partial}^{\prime})^{2}\not=0. Also let c>0c>0 be big enough, Uc={pM:f(p)<c}U_{c}=\{p\in M:f(p)<-c\} and (Ω(M,Uc),d)(\Omega^{*}(M,U_{c}),d) be the relative de Rham complex.

Theorem 1.3.

If (M,g,f)(M,g,f) is well tame, then (~)2=0(\tilde{\partial}^{\prime})^{2}=0, and therefore the cohomology H(C(Wu),~)H^{*}(C^{\bullet}(W^{u}),\tilde{\partial}^{\prime}) is well defined. Moreover, there exists T00T_{0}\geq 0 (When (M,g,f)(M,g,f) is strongly tame, T0=0T_{0}=0), such that H(2)(M,dTf)H^{*}_{(2)}(M,d_{Tf}) is isomorphic to H(C(Wu),~)H^{*}(C^{\bullet}(W^{u}),\tilde{\partial}^{\prime}) for all T>T0T>T_{0}. In addition, H(C(Wu),~)H^{*}(C^{\bullet}(W^{u}),\tilde{\partial}^{\prime}), hence H(2)(M,dTf)H^{*}_{(2)}(M,d_{Tf}) is isomorphic to the relative de Rham cohomology HdR(M,Uc)H_{dR}^{*}(M,U_{c}).

By Theorem 1.3, we can refine our result of Theorem 1.2

Corollary 1.4.

If (M,g,f)(M,g,f) is well tame, then bi(T)b_{i}(T) is independent of TT when TT is big enough. When (M,g,f)(M,g,f) is strongly tame, bi(T)b_{i}(T) is independent of T>0.T>0.

As an another application of Theorem 1.3, we study the Morse cohomology for compact manifolds with boundary.

Let MM be a compact, oriented manifold of dimension nn with boundary M\partial M. Let Ni(iΛ)N_{i}(i\in\Lambda) be the connected components of M\partial M. We fix a collar neighborhood (0,1]×NiM(0,1]\times N_{i}\subset M, and let rr be the standard coordinate on the (0,1](0,1] factor.

Definition 1.5.

A smooth function ff on MM is called a transversal Morse function if it satisfies the following conditions:

  1. 1.

    f|MMf|_{M\setminus\partial M} is a Morse function on the manifold MMM\setminus\partial M;

  2. 2.

    f|Mf|_{\partial M} is a Morse function on the manifold M\partial M.

  3. 3.

    For any point xx on the collar neighborhood, fr|x0\left.-\frac{\partial f}{\partial r}\right|_{x}\neq 0.

For a transversal Morse function ff on MM, since fr-\frac{\partial f}{\partial r} is continuous on any connected components of M\partial M, so we can call NiN_{i} to be positive (with respect to ff) if fr|Ni>0-\frac{\partial f}{\partial r}|_{N_{i}}>0, and negative if fr|Ni<0.-\frac{\partial f}{\partial r}|_{N_{i}}<0.

Let N+N^{+} be the union of all positive boundaries, NN^{-} be the union of all negative boundaries. Suppose we have a partition of positive boundaries N+=N1+N2+N^{+}=N^{+}_{1}\sqcup N^{+}_{2}, and a partition of negative boundaries N=N1N2.N^{-}=N^{-}_{1}\sqcup N^{-}_{2}. Now We denote

  • by Crit,k(f)Crit^{\circ,k}(f) the set of internal critical points with Morse index kk, mk=|Crit,k(f)|m_{k}=|Crit^{\circ,k}(f)|;

  • by CritNj++,k(f)(j=1,2)Crit^{+,k}_{N^{+}_{j}}(f)(j=1,2) the set of critical points on positive boundary Nj+N^{+}_{j} with Morse index kk, nk,Nj+=|CritNj++,k(f)|n_{k,N^{+}_{j}}=|Crit^{+,k}_{N^{+}_{j}}(f)|;

  • by CritNj,k(f)(j=1,2)Crit^{-,k}_{N^{-}_{j}}(f)(j=1,2) the set of critical points on negative boundary NjN^{-}_{j} with Morse index k1k-1, lk,Nj=|CritNj,k(f)|l_{k,N^{-}_{j}}=|Crit^{-,k}_{N^{-}_{j}}(f)|.

Let Crit(f)=Cirt,(f)CirtN1++,(f)CirtN1,(f)\mathrm{Crit}^{*}(f)=Cirt^{\circ,*}(f)\cup Cirt^{+,*}_{N^{+}_{1}}(f)\cup Cirt^{-,*}_{N^{-}_{1}}(f)

Theorem 1.5.

There is a differential ~:Critk(f)Critk+1(f)\tilde{\partial}^{\prime}:\mathrm{Crit}^{k}(f)\mapsto\mathrm{Crit}^{k+1}(f) making (Crit(f),~)(\mathrm{Crit}^{*}(f)\otimes\mathbb{R},\tilde{\partial}^{\prime}) a chain complex. Moreover, H(Crit(f),~)H^{*}(\mathrm{Crit}^{\bullet}(f)\otimes\mathbb{R},\tilde{\partial}^{\prime}) is isomorphic to the relative de Rham cohomology H(M,N2N1+).H^{*}(M,N_{2}^{-}\cup N_{1}^{+}).

In particular, for N1+=N+,N1=,N_{1}^{+}=N^{+},N_{1}^{-}=\emptyset, H(Crit(f),~)H^{*}(\mathrm{Crit}^{\bullet}(f)\otimes\mathbb{R},\tilde{\partial}^{\prime}) is isomorphic to the relative de Rham cohomology H(M,M).H^{*}(M,\partial M).

Corollary 1.6.

Set bi(M,N2N1+)=dim(Hi(M,N2N1+),b_{i}(M,N_{2}^{-}\cup N_{1}^{+})=\dim(H^{i}(M,N_{2}^{-}\cup N_{1}^{+}), then we have the following Morse inequalities:

(1)ki=0k(1)ibi(M,N2N1+)(1)ki=0k(1)i(mi+ni1,N1++li,N2).(-1)^{k}\sum_{i=0}^{k}(-1)^{i}b_{i}(M,N_{2}^{-}\cup N_{1}^{+})\leq(-1)^{k}\sum_{i=0}^{k}(-1)^{i}(m_{i}+n_{i-1,N^{+}_{1}}+l_{i,N_{2}^{-}}).
Remark 1.6.
  1. 1.

    Theorem 1.5 is a generalization of a result in [16].

  2. 2.

    Corollary 1.6 is a generalization of a result in [lu2017thomsmale].

1.4 Notations and Organization

In this article, we will generally use ϕ,ψ\phi,\psi to denote differential forms, ff Morse function, u,vu,v functions, ν,ω\nu,\omega eigenforms, x,y,zx,y,z critical points of ff, p,qp,q general points, and p0p_{0} a fixed point.

This paper is organized as follows. In Section 2, we discuss the spectral theory for the Witten Laplacian in our setting. We then proceed to establish the exponential decay estimate for eigenforms of the Witten Laplacian in Section 3. Assuming two technical results whose proofs are deferred to Section 7 (7.3) and using a lemma proved in Section 5 about the Agmon distance we prove Theorem 1.1, the Agmon estimate.

In Section 4 we present the proof of the Strong Morse Inequalities, Theorem 1.2, after introducing Witten instanton complex (FTf[0,1],,dTf)(F_{Tf}^{[0,1],*},d_{Tf}). Section 5 concerns the Thom-Smale theory in our setting. More specifically, we define the Thom-Smale complex C((Wu),~)C^{*}((W^{u})^{\prime},\tilde{\partial}^{\prime}). Then we define a morphism between the Witten instanton complex and the Thom-Smale complex, 𝒥:(FTf[0,1],,dTf)C((Wu),~).\mathcal{J}:(F_{Tf}^{[0,1],*},d_{Tf})\mapsto C^{*}((W^{u})^{\prime},\tilde{\partial}^{\prime}). We prove that 𝒥\mathcal{J} is well defined by using the Agamon estimate, deferring the proof that 𝒥\mathcal{J} is a chain map to Subsection 7.3.

In Section 6 we give an application of our results, namely Theorem 1.5. Section 7 collects the proofs of several technical results. In the first two subsections we prove the lemmas used in the proof of Agmon estimate. In the Subsection 7.3, we prove that our Thom-Smale complex is indeed a complex, i.e., ~2=0\tilde{\partial}^{\prime 2}=0. The rest of the proof for Theorem 1.3 is in Subsection 7.4 and Subsection 7.5. Finally in Section 8, which is an appendix, we discuss the Kodaira decomposition in a more general setting.

2 The Spectrum Of Witten Laplacian

In this section we study the spectral theory of the Witten Laplacian on noncompact manifolds. In particular we establish the Kodaira decomposition and the Hodge theorem for the Witten Laplacian under our tameness condition.

2.1 Essential self-adjointness of df+δfd_{f}+\delta_{f}

Theorem 2.1.

On a complete Riemannian manifold, if

limsupp|2f|(p)|f|2(p)<,\lim\sup_{p\to\infty}\frac{|\nabla^{2}f|(p)}{|\nabla f|^{2}(p)}<\infty,

then df+δfd_{f}+\delta_{f} is essentially self-adjoint.

Proof.

Since limsupp|2f|(p)|f|2(p)<\lim\sup_{p\to\infty}\frac{|\nabla^{2}f|(p)}{|\nabla f|^{2}(p)}<\infty, f\Box_{f} is bounded from below by Proposition 1.4. The rest of the proof is essentially the same as in Section 4 of [3]; see also the proof of Theorem 1.17 in [8]. ∎

2.2 On the spectrum of Tf\Box_{Tf}

From now on we will assume that (M,g,f)(M,g,f) is well tame. Then it follows that there exists a compact subset KK, which can be taken to be a compact submanifold with boundaries that contains the closure of a ball of sufficiently large radius of MM (we will make a more specific choice of KK later in section 5), δ1=12ϵf,δ2=2cf>0\delta_{1}=\frac{1}{2}\epsilon_{f},\delta_{2}=2c_{f}>0, such that

|f|>δ1,|2f|<δ2|f|2onMK.|\nabla f|>\delta_{1},\ \ \ |\nabla^{2}f|<\delta_{2}|\nabla f|^{2}\ \ \mbox{on}\ M-K. (2.1)

Let CK=maxK|2f|C_{K}=\max_{K}|\nabla^{2}f|. First, we establish the following basic lemma.

Lemma 2.1.

Fix any b(0,1)b\in(0,1), there exists T1=T1(cf,b)0T_{1}=T_{1}(c_{f},b)\geq 0 so that whenever TT1T\geq T_{1}, ϕDom(Tf)\phi\in\mathrm{Dom}(\Box_{Tf})

M(Tfϕ,ϕ)dvol\displaystyle\int_{M}(\Box_{Tf}\phi,\phi)\mathrm{dvol} \displaystyle\geq M(ϕ,ϕ)dvol+MKb2T2|f|2(ϕ,ϕ)dvol\displaystyle\int_{M}(\nabla\phi,\nabla\phi)\mathrm{dvol}+\int_{M-K}b^{2}T^{2}|\nabla f|^{2}(\phi,\phi)\mathrm{dvol} (2.2)
(CR+TCK)K(ϕ,ϕ)dvol.\displaystyle-(C_{R}+TC_{K})\int_{K}(\phi,\phi)\mathrm{dvol}.

Here CRC_{R} is a constant depending only on the sectional curvature bounds of gg.

Proof.

It suffices to show the inequality for a compactly supported smooth form. By Proposition 1.4, together with the Bochner-Weitzenböck formula, we have

M(Tfϕ,ϕ)dvol\displaystyle\int_{M}(\Box_{Tf}\phi,\phi)\mathrm{dvol} \displaystyle\geq M(ϕ,ϕ)dvol(CR+TCK)K(ϕ,ϕ)dvol\displaystyle\int_{M}(\nabla\phi,\nabla\phi)\mathrm{dvol}-(C_{R}+TC_{K})\int_{K}(\phi,\phi)\mathrm{dvol}
+MKeT(p)(ϕ,ϕ)dvol,\displaystyle+\int_{M-K}e_{T}(p)(\phi,\phi)\mathrm{dvol},

where eT=T2|f|2(14cfT)e_{T}=T^{2}|\nabla f|^{2}(1-\frac{4c_{f}}{T}). Thus, for any b(0,1)b\in(0,1), there exists

T1=T1(cf,b)T_{1}=T_{1}(c_{f},b) (2.3)

such that whenever TT1,T\geq T_{1}, we have

M(Tfϕ,ϕ)dvol\displaystyle\int_{M}(\Box_{Tf}\phi,\phi)\mathrm{dvol} \displaystyle\geq M(ϕ,ϕ)dvol+MK|bTf|2(ϕ,ϕ)dvol\displaystyle\int_{M}(\nabla\phi,\nabla\phi)\mathrm{dvol}+\int_{M-K}|bT\nabla f|^{2}(\phi,\phi)\mathrm{dvol}
(C+TCK)M(ϕ,ϕ)dvol.\displaystyle-(C+TC_{K})\int_{M}(\phi,\phi)\mathrm{dvol}.

Remark 2.2.

When (M,g,f)(M,g,f) is strongly tame, we can take T1=0,T_{1}=0, but KK may depend on T.T.

Let g~T:=b2T2|f|2g\tilde{g}_{T}:=b^{2}T^{2}|\nabla f|^{2}g be a new metric on MM (with discrete conical singularities).Fix p0K,p_{0}\in K, let ρT(p)\rho_{T}(p) be the distance between pp and p0p_{0} induced by g~T.\tilde{g}_{T}. Then we have |ρT|2=b2T2|f|2|\nabla\rho_{T}|^{2}=b^{2}T^{2}|\nabla f|^{2} a.e., where the gradient \nabla is induced by gg.

Theorem 2.2.

Let σ\sigma be the set of spectrum of Tf.\Box_{Tf}. Then for any positive number δ<bϵf2\delta<\frac{b\epsilon_{f}}{2} there is

T2=T2(δ,cf,ϵf,CR,CK)>0T_{2}=T_{2}(\delta,c_{f},\epsilon_{f},C_{R},C_{K})>0 (2.4)

such that when TT2T\geq T_{2}, σ[0,δ2T2]\sigma\cap[0,\delta^{2}T^{2}] consists of a finite number of eigenvalues of finite multiplicity.

Proof.

Let P:L2Λ(M)L2Λ(M)P:L^{2}\Lambda^{*}(M)\mapsto L^{2}\Lambda^{*}(M) be the integral of the spectral measure of Tf\Box_{Tf} on [0,δ2T2][0,\delta^{2}T^{2}]. It suffices to prove that L:=Im(P)L:=Im(P) is finite dimensional. For any ϕL,\phi\in L, we have

M(Tfϕ,ϕ)𝑑volδ2T2M|ϕ|2𝑑vol.\int_{M}(\Box_{Tf}\phi,\phi)dvol\leq\delta^{2}T^{2}\int_{M}|\phi|^{2}dvol. (2.5)

Combining with (2.2), we have

δ2T2M|ϕ|2𝑑vol\displaystyle\delta^{2}T^{2}\int_{M}|\phi|^{2}dvol \displaystyle\geq M(ϕ,ϕ)dvol+MK|bTf|2(ϕ,ϕ)dvol\displaystyle\int_{M}(\nabla\phi,\nabla\phi)\mathrm{dvol}+\int_{M-K}|bT\nabla f|^{2}(\phi,\phi)\mathrm{dvol}
(CR+TCK)M(ϕ,ϕ)dvol\displaystyle-(C_{R}+TC_{K})\int_{M}(\phi,\phi)\mathrm{dvol}

provided TT1T\geq T_{1}. That is,

M(ϕ,ϕ)dvol+MK|bTf|2(ϕ,ϕ)dvol\displaystyle\int_{M}(\nabla\phi,\nabla\phi)\mathrm{dvol}+\int_{M-K}|bT\nabla f|^{2}(\phi,\phi)\mathrm{dvol} \displaystyle\leq
δ2T2(1+CRδ2T2+CKδ2T)M(ϕ,ϕ)dvol\displaystyle\delta^{2}T^{2}(1+\frac{C_{R}}{\delta^{2}T^{2}}+\frac{C_{K}}{\delta^{2}T})\int_{M}(\phi,\phi)\mathrm{dvol}

Since |bTf|2>(bϵf2)2T2|bT\nabla f|^{2}>(\frac{b\epsilon_{f}}{2})^{2}T^{2} on MKM-K, δ<bϵf2\delta<\frac{b\epsilon_{f}}{2}, there is T2=T2(δ,cf,ϵf,CR,CK)T10T_{2}=T_{2}(\delta,c_{f},\epsilon_{f},C_{R},C_{K})\geq T_{1}\geq 0 such that when TT2T\geq T_{2},

M(ϕ,ϕ)dvolδ2T2(1+CRδ2T2+CKδ2T)K(ϕ,ϕ)dvol.\int_{M}(\nabla\phi,\nabla\phi)\mathrm{dvol}\leq\delta^{2}T^{2}(1+\frac{C_{R}}{\delta^{2}T^{2}}+\frac{C_{K}}{\delta^{2}T})\int_{K}(\phi,\phi)\mathrm{dvol}. (2.6)

Now define Q:LL2Λ(K)Q:L\mapsto L^{2}\Lambda^{*}(K), by Qu=u|K.Qu=u|_{K}. By (2.6), it’s easy to see that QQ is injective, and Im(Q)W1,2(ΛK).Im(Q)\subset W^{1,2}(\Lambda^{*}K). Since W1,2(ΛK)L2Λ(K)W^{1,2}(\Lambda^{*}K)\hookrightarrow L^{2}\Lambda^{*}(K) is compact, dim(L)=dim(Im(Q))dim(L)=dim(Im(Q)) must be finite. ∎

Remark 2.3.

Once again, if (M,g,f)(M,g,f) is strongly tame, we can take T2=0.T_{2}=0.

We now state the important consequence of this section. By combining Theorem 2.1 and Theorem 2.2 with Proposition 8.3, decomposition (8.4), we have

Theorem 2.3.

Assume that (M,g,f)(M,g,f) is well tame. Then we have the Kodaira decomposition

L2Λ(M)=kerfIm(df¯)Im(δ¯f),L^{2}\Lambda^{*}(M)=\ker\Box_{f}\oplus\mathrm{Im}(\bar{d_{f}})\oplus\mathrm{Im}(\bar{\delta}_{f}),

Furthermore, the Hodge Theorem holds:

H(2)(M,df)kerf.H^{*}_{(2)}(M,d_{f})\cong\ker\Box_{f}.

3 Exponential decay of eigenfunction

In this section, we assume that (M,g,f)(M,g,f) is well tame, and TT2T\geq T_{2}, where T2T_{2} is described in Lemma 2.1. If (M,g,f)(M,g,f) is strongly tame, then we can just take T2=0.T_{2}=0.

Recall that g~T:=b2T2|f|2g\tilde{g}_{T}:=b^{2}T^{2}|\nabla f|^{2}g, the Agmon metric on M.M. Let KK be the compact set as in last section. In this and later sections we define the Agmon distance ρT(p)\rho_{T}(p) be the distance between pp and KK induced by g~T.\tilde{g}_{T}. Then we have |ρT|2=b2T2|f|2|\nabla\rho_{T}|^{2}=b^{2}T^{2}|\nabla f|^{2} a.e. pKp\notin K, where the gradient \nabla is induced by gg.

For simplicity, denote b2T2|f|2b^{2}T^{2}|\nabla f|^{2} by λT\lambda_{T}. We need the following two technical lemmas, whose proofs are postponed to Section 7.

Lemma 3.1.

Assume wL2(M),0uL2(M)w\in L^{2}(M),0\leq u\in L^{2}(M), and (Δ+λT)uw(\Delta+\lambda_{T})u\leq w outside the compact subset KMK\subset M in the weak sense. That is

MKuv+λTuvdvolMKwvdvol, 0vCc(MK).\int_{M-K}\nabla u\nabla v+\lambda_{T}uv\mathrm{dvol}\leq\int_{M-K}w\cdot v\mathrm{dvol},\ \ \forall\ 0\leq v\in C_{c}^{\infty}(M-K).

Then there exists another compact subset LKL\supset K of MM such that

ML|u|2λTexp(2bρT)dvolC1[MK|w|2λT1exp(2bρT)dvol+LK|u|2λTexp(2bρT)dvol]\displaystyle\begin{split}\int_{M-L}|u|^{2}\lambda_{T}\exp(2b\rho_{T})\mathrm{dvol}&\leq C_{1}\left[\int_{M-K}|w|^{2}\lambda_{T}^{-1}\exp(2b\rho_{T})\mathrm{dvol}\right.\\ &+\left.\int_{L-K}|u|^{2}\lambda_{T}\exp(2b\rho_{T})\mathrm{dvol}\right]\end{split} (3.1)

for C1=8(1+b2)(1b2)2C_{1}=\frac{8(1+b^{2})}{(1-b^{2})^{2}}.

Corollary 3.1.

If w=cuw=cu for some c>0c>0 and T21+cbϵfT\geq\frac{2\sqrt{1+c}}{b\epsilon_{f}}, then

I(u):=M|u|2exp(2bρT)dvol<.I(u):=\int_{M}|u|^{2}\exp(2b\rho_{T})\mathrm{dvol}<\infty.
Proof.

With this choice of TT, λT>1+c\lambda_{T}>{1+c} outside K.K. Now replacing λT\lambda_{T} with λTc\lambda_{T}-c and ww with 0 in Lemma 3.1, we get

ML|u|2exp(2bρT)dvolML|u|2(λTc)exp(2bρT)dvol\displaystyle\int_{M-L}|u|^{2}\exp(2b\rho_{T})\mathrm{dvol}\leq\int_{M-L}|u|^{2}(\lambda_{T}-c)\exp(2b\rho_{T})\mathrm{dvol}
C1LK|u|2λTexp(2bρT)dvol<.\displaystyle\leq C_{1}\int_{L-K}|u|^{2}\lambda_{T}\exp(2b\rho_{T})\mathrm{dvol}<\infty.

By refining the argument above, we have the following corollary which will be used in the proof of our Agmon estimate for eigenforms.

Corollary 3.2.

If 0uL2(M)0\leq u\in L^{2}(M), and Tfu(c+T|2f|)u\Box_{Tf}u\leq(c+T|\nabla^{2}f|)u for some c>0c>0 and Tmax{(c+2cf)C2,1b2+8c1b2/(bϵf)}T\geq\max\{(c+2c_{f})C_{2},\sqrt{\frac{1-b^{2}+8c}{1-b^{2}}}/(b\epsilon_{f})\}, then

I(u):=M|u|2exp(2bρT)dvolCT2u2I(u):=\int_{M}|u|^{2}\exp(2b\rho_{T})\mathrm{dvol}\leq CT^{2}\|u\|^{2}

where the constant C=C(CL,CL,r0,b,c),L={pM:ρT(p)2},CL>maxL|f|2,CL>maxL|2f|C=C(C^{L},C_{L},r_{0},b,c),L=\{p\in M:\rho_{T}(p)\leq 2\},C^{L}>\max_{L}|\nabla f|^{2},C_{L}>\max_{L}|\nabla^{2}f|.

Proof.

Following the proof of Lemma 3.1 given in Section 7.1, put L={pM:ρT(p)2}.L=\{p\in M:\rho_{T}(p)\leq 2\}. Then we deduce

ML|u|2λTexp(2bρT)dvolC1LK|u|2λTexp(2bρT)dvol+C2MK(c+T|2f|)|u|2exp(2bρT)dvol\displaystyle\begin{split}\int_{M-L}|u|^{2}\lambda_{T}\exp(2b\rho_{T})\mathrm{dvol}&\leq C_{1}\int_{L-K}|u|^{2}\lambda_{T}\exp(2b\rho_{T})\mathrm{dvol}\\ &+C_{2}\int_{M-K}(c+T|\nabla^{2}f|)|u|^{2}\exp(2b\rho_{T})\mathrm{dvol}\end{split}

for C1=8(1+b2)(1b2)2C_{1}=\frac{8(1+b^{2})}{(1-b^{2})^{2}} as above, and C2=81b2C_{2}=\frac{8}{1-b^{2}}. We split the second integral on the right hand side into two; the one over LKL-K will be absorbed into the first term. The second term is (we omit the volume form here)

C2MLc|u|2exp(2bρT)+C2MLT|2f||u|2exp(2bρT)C2MLc|u|2exp(2bρT)+C2cfTcC2/ϵfML|u|2(λTcC2)exp(2bρT).\displaystyle\begin{split}C_{2}\int_{M-L}c|u|^{2}\exp(2b\rho_{T})&+C_{2}\int_{M-L}T|\nabla^{2}f||u|^{2}\exp(2b\rho_{T})\\ \leq C_{2}\int_{M-L}c|u|^{2}\exp(2b\rho_{T})&+\frac{C_{2}c_{f}}{T-cC_{2}/\epsilon_{f}}\int_{M-L}|u|^{2}(\lambda_{T}-cC_{2})\exp(2b\rho_{T}).\end{split}

Combining the above one arrives at

ML|u|2(λTcC2)exp(2bρT)C1LK|u|2(λT+c+T|2f|)exp(2bρT)+cfC2TcC2/ϵfML|u|2(λTcC2)exp(2bρT).\displaystyle\begin{split}\int_{M-L}|u|^{2}(\lambda_{T}-cC_{2})\exp(2b\rho_{T})&\leq C_{1}\int_{L-K}|u|^{2}(\lambda_{T}+c+T|\nabla^{2}f|)\exp(2b\rho_{T})\\ &+\frac{c_{f}C_{2}}{T-cC_{2}/\epsilon_{f}}\int_{M-L}|u|^{2}(\lambda_{T}-cC_{2})\exp(2b\rho_{T}).\end{split}

Thus, for T(c+2cf)C2T\geq(c+2c_{f})C_{2},

ML|u|2(λTcC2)exp(2bρT)2C1(CLb2T2+cLT+c)e4bu2,\displaystyle\int_{M-L}|u|^{2}(\lambda_{T}-cC_{2})\exp(2b\rho_{T})\leq 2C_{1}(C^{L}b^{2}T^{2}+c_{L}T+c)e^{4b}\|u\|^{2},

where CL>maxL|f|2,CL>maxL|2f|C^{L}>\max_{L}|\nabla f|^{2},C_{L}>\max_{L}|\nabla^{2}f|. If TT is also bigger than 1b2+8c1b2/(bϵf)\sqrt{\frac{1-b^{2}+8c}{1-b^{2}}}/(b\epsilon_{f}), then λT>1+cC2\lambda_{T}>1+cC_{2} outside LL. Hence

ML|u|2exp(2bρT)dvol\displaystyle\int_{M-L}|u|^{2}\exp(2b\rho_{T})\mathrm{dvol} ML|u|2(λTcC2)exp(2bρT)dvol\displaystyle\leq\int_{M-L}|u|^{2}(\lambda_{T}-cC-2)\exp(2b\rho_{T})\mathrm{dvol}
2C1(CLb2T2+cLT+c)e4bu2,\displaystyle\leq 2C_{1}(C^{L}b^{2}T^{2}+c_{L}T+c)e^{4b}\|u\|^{2},

and consequently

M|u|2exp(2bρT)dvol[2C1(CLb2T2+cLT+c)+1]e4bu2,\displaystyle\int_{M}|u|^{2}\exp(2b\rho_{T})\mathrm{dvol}\leq[2C_{1}(C^{L}b^{2}T^{2}+c_{L}T+c)+1]e^{4b}\|u\|^{2},

for Tmax{(c+2cf)C2,1b2+8c1b2/(bϵf)}T\geq\max\{(c+2c_{f})C_{2},\sqrt{\frac{1-b^{2}+8c}{1-b^{2}}}/(b\epsilon_{f})\}. ∎

Remark 3.2.

It may seem that CLC^{L} and CLC_{L} depend on TT as L={pM:ρT(p)<r0}L=\{p\in M:\rho_{T}(p)<r_{0}\}. However, notice that when TT becomes bigger, LL gets smaller. Hence we can choose CL>maxpL|f|(p)C^{L}>\max_{p\in L}|\nabla f|(p), CL>maxpL|2f(p)|C_{L}>\max_{p\in L}|\nabla^{2}f(p)|, s.t. they are independent of T.T.

Lemma 3.3 (De Giorgi-Nash-Moser Estimates).

For r>0r>0, let Br(p)B_{r}(p) be the geodesic ball around pp with radius rr (in the metric gg). Let 0uL2(M)0\leq u\in L^{2}(M), and Δucu\Delta u\leq cu on B2r(p)B_{2r}(p) in the weak sense for some constant c0.c\geq 0. Then there exists constant C2>0C_{2}>0 depending only on the dimension nn, the Sobolev constant, and cc, such that

supyBr(p)u(y)C2rn/2uL2(B2r(p)).\sup_{y\in B_{r}(p)}u(y)\leq\frac{C_{2}}{r^{n/2}}\|u\|_{L^{2}(B_{2r}(p))}.

With these preparation we are now ready to prove our first main estimate for the eigenforms of Tf\Box_{Tf}.

Proof of Theorem 1.1.

Consider an eigenform ω\omega of Tf\Box_{Tf}. That is Tfω=μ(T)ω\Box_{Tf}\omega=\mu(T)\omega, where the eigenvalue μ(T)\mu(T) satisfies |μ(T)|c|\mu(T)|\leq c for some constant cc. Then letting u=g(ω,ω)1/2u=g(\omega,\omega)^{1/2}, by a straightforward computation using the Bochner’s formula (for forms) and the Kato’s inequality, we have

Tfu(c+|R|+T|2f|)u,\Box_{Tf}u\leq(c+|R|+T|\nabla^{2}f|)u,

where |R||R| is the upper bound of curvature tensor. Hence by Corollary 3.2, we have, for Tmax{(c+|R|+2cf)C2,1b2+8c+8|R|1b2/(bϵf)}T\geq\max\{(c+|R|+2c_{f})C_{2},\sqrt{\frac{1-b^{2}+8c+8|R|}{1-b^{2}}}/(b\epsilon_{f})\},

I(u)=M|u|2exp(2bρT)dvolCT2u2I(u)=\int_{M}|u|^{2}\exp(2b\rho_{T})\mathrm{dvol}\leq CT^{2}\|u\|^{2}

where the constant C=C(CL,CL,r0,b,c,|R|)C=C(C^{L},C_{L},r_{0},b,c,|R|).

Recall that the compact set KK is chosen so that (2.1) is satisfied. Hence by Proposition 1.4, the conditions of Lemma 3.3 are satisfied for uu on MKM-K. Also, the Agmon distance ρT(p)\rho_{T}(p) is the distance between pp and KK induced by g~T\tilde{g}_{T} and L={pM:ρT(p)2}.L=\{p\in M:\rho_{T}(p)\leq 2\}. Suppose pMLp\in M-L. Denote by B~r(p)\tilde{B}_{r}(p) the g~T\tilde{g}_{T}-geodesic ball around pp with radius rr. Set l=supqB~2(p)|Tf|(q)l=\sup_{q\in\tilde{B}_{2}(p)}|T\nabla f|(q), and r=1/(2l).r=1/(2l). Then one can easily verify that B2r(q)B~2(p),qB~1(p).B_{2r}(q)\subset\tilde{B}_{2}(p),\ \forall q\in\tilde{B}_{1}(p).

Choose q0B~2(p)¯q_{0}\in\overline{\tilde{B}_{2}(p)} so that |Tf|(q0)(l/2,l].|T\nabla f|(q_{0})\in(l/2,l]. By Lemma 3.1 and de Giorgi-Nash-Moser estimate Lemma 3.3, we have

|u(p)|2exp(2bρT(p))\displaystyle|u(p)|^{2}\exp(2b\rho_{T}(p)) C2rnuL2(B2r(p))2exp(2bρT(p))\displaystyle\leq\frac{C_{2}}{r^{n}}\|u\|^{2}_{L^{2}(B_{2r}(p))}\exp(2b\rho_{T}(p))
C3rnB~2(p)|u|2(q)exp(2bρT(q))dvol\displaystyle\leq\frac{C_{3}}{r^{n}}\int_{\tilde{B}_{2}(p)}|u|^{2}(q)\exp(2b\rho_{T}(q))\mathrm{dvol}
C4|Tf(q0)|nI(u).\displaystyle\leq{C_{4}}|T\nabla f(q_{0})|^{n}I(u).

We will prove that

|f(q0)|2C5exp(2cfbTρT(q0))|\nabla f(q_{0})|^{2}\leq C_{5}\exp(\frac{2c_{f}}{bT}\rho_{T}(q_{0})) (3.2)

in Lemma 5.6. Hence,

|f(q0)|2C6exp(2cfbTρT(p))C6exp(ϵρT(p))|\nabla f(q_{0})|^{2}\leq C_{6}\exp(\frac{2c_{f}}{bT}\rho_{T}(p))\leq C_{6}\exp(\epsilon\rho_{T}(p))

for any small ϵ\epsilon, provided T2cfbϵT\geq\frac{2c_{f}}{b\epsilon}. It follows then that,

|u(p)|2C6I(u)Tnexp(2aρT(p)),|u(p)|^{2}\leq C_{6}I(u)T^{n}\exp(-2a\rho_{T}(p)),

for any a<ba<b provided Tncfb(ba)T\geq\frac{nc_{f}}{b(b-a)}. Hence if Tmax{(c+|R|+2cf)C2,1b2+8c+8|R|1b2/(bϵf),ncfb(ba)}T\geq\max\{(c+|R|+2c_{f})C_{2},\sqrt{\frac{1-b^{2}+8c+8|R|}{1-b^{2}}}/(b\epsilon_{f}),\frac{nc_{f}}{b(b-a)}\},

|u(p)|2C6C(CL,CL,r0,b,c,|R|)Tn+2exp(2aρT(p))u2.|u(p)|^{2}\leq C_{6}C(C^{L},C_{L},r_{0},b,c,|R|)T^{n+2}\exp(-2a\rho_{T}(p))\|u\|^{2}.

Remark 3.4.

The proof above gives the inequality for pML={ρT(p)>r0}p\in M-L=\{\rho_{T}(p)>r_{0}\} for some constant r0r_{0} independent of TT, which is what we needed for later applications. For pLp\in L, using the same reasoning as in Remark 3.2, there exist constant C>0C>0, which is independent of TT, such that

ΔuCTu.\Delta u\leq CTu. (3.3)

for all pLp\in L. Therefore via Moser iteration as in Lemma 3.3 and similar arguments as above, one can show that

|u|2(p)CTnuL22Cexp(2a)Tnexp(aρT)uL22.|u|^{2}(p)\leq C^{\prime}T^{n}\|u\|_{L^{2}}^{2}\leq C^{\prime}\exp(2a)T^{n}\exp(-a\rho_{T})\|u\|_{L^{2}}^{2}.

4 Morse inequalities

In this and the next section, we assume that ff is a Morse function on MM, and TT0T\geq T_{0}. In fact, we assume that in a neighborhood UxU_{x} of critical points xx of ff, we have coordinate system z=(z1,,zn),z=(z_{1},...,z_{n}), such that

f=z12znf(x)2+znf(x)+12++zn2,g=dz12++dzn2.\displaystyle\begin{split}f=-z_{1}^{2}-...-z_{n_{f}(x)}^{2}+z_{n_{f}(x)+1}^{2}+...+z_{n}^{2},g=dz_{1}^{2}+...+dz_{n}^{2}.\end{split} (4.1)

This is a generic condition. Without loss of generality we assume that UxU_{x} is an Euclidean open ball around xx with radius 1.1. Also, these open sets are disjoint.

Let FTf[0,1],F_{Tf}^{[0,1],*} be the space spanned by the eigenforms of Tf\Box_{Tf} with eigenvalue lying in [0,1].[0,1]. By Theorem 2.2, FTf[0,1],F_{Tf}^{[0,1],*} is finite dimensional. Recall that mim_{i} denotes the number of critical points of ff with Morse index i.i. We have the following Proposition:

Proposition 4.1.

There exists T3>T0T_{3}>T_{0} big enough, so that whenever TT3,T\geq T_{3}, the number of eigenvalues (counted with multiplicity) in [0,1][0,1] of Tf|Ω(2)i(M)\Box_{Tf}|_{\Omega^{i}_{(2)}(M)} equals mim_{i}. I.e. dimFTf[0,1],=mi\dim F_{Tf}^{[0,1],*}=m_{i}.

Remark 4.2.

See the definition of T0T_{0} in (5.2). Also recall that, if (M,g,f)(M,g,f) is strongly tame, T0=0.T_{0}=0.

The proof of Proposition 4.1 follows from that of Proposition 5.5 in [21], except for the proof of the following proposition:

Proposition 4.3.

There exist constants C>0C>0, T4>0T_{4}>0 such that for any smooth form ϕΩ(2)(M)\phi\in\Omega^{*}_{(2)}(M) with supp(ϕ)MxCritfUx{\rm supp}(\phi)\subset M-\cup_{x\in\mathrm{Crit}{f}}U_{x} and TT4T\geq T_{4}, one has

TfϕL2CTϕL2.\|\Box_{Tf}\phi\|_{L^{2}}\geq CT\|\phi\|_{L^{2}}.

Here supp(ϕ){\rm supp}(\phi) denotes the support of ϕ.\phi.

Proof.

Since ff is well tame, there exist δ1,δ2>0,\delta_{1},\delta_{2}>0, s.t. |f|δ1,|\nabla f|\geq\delta_{1}, also |2f|δ2|f|2|\nabla^{2}f|\leq\delta_{2}|\nabla f|^{2} on MxCritfUxM-\cup_{x\in\mathrm{Crit}{f}}U_{x}. Then our proposition follows from the same argument in Proposition 4.7 of [21]. ∎

On the other hand, (FTf[0,1],,dTf)(F_{Tf}^{[0,1],*},d_{Tf}) form a complex, the so called Witten instanton complex, whose cohomology is H(2)(M,df)H^{*}_{(2)}(M,d_{f}) by Theorem 2.3. As a result, our Theorem 1.2 (the strong Morse inequalities) follows from Proposition 4.1 and our Hodge theorem when T>T3T>T_{3}. For the case of T(T0,T3]T\in(T_{0},T_{3}], see section 7.

5 Thom-Smale theory

In this section, we assume that KK is a compact subset of MM, ϵ>0\epsilon>0 is small enough (to be determined later), T5=T5(ϵ)T_{5}=T_{5}(\epsilon) is big enough, such that outside KK, we have

T|2f|ϵT2|f|2,T|\nabla^{2}f|\leq\epsilon\,T^{2}|\nabla f|^{2}, (5.1)

provided TT5T\geq T_{5}.

Moreover, we make a more judicious choice of K.K. Fix any p0Mp_{0}\in M. Set

D\displaystyle D =\displaystyle= supx,yCrit(f)d~T(x,p0)+d~T(x,y)\displaystyle\sup_{x,y\in\mathrm{Crit}(f)}{\tilde{d}_{T}(x,p_{0})+\tilde{d}_{T}(x,y)}
=\displaystyle= supx,yCrit(f)bT(|f(x)f(p0)|+|f(x)f(y)|),\displaystyle\sup_{x,y\in\mathrm{Crit}(f)}bT(|f(x)-f(p_{0})|+|f(x)-f(y)|),

where d~T\tilde{d}_{T} is the distance function induced by g~T\tilde{g}_{T} (note that the Agmon distance ρT(x)\rho_{T}(x) is the same distance function but between xx and a compact subset KK). The second equality follows from the claim in the proof of Lemma 5.5. We choose KK so that that

B~D+1(p0)={pM:d~T(p,p0)D+1}K\tilde{B}_{D+1}(p_{0})=\{p\in M:\tilde{d}_{T}(p,p_{0})\leq D+1\}\subset K^{\circ}

where KK^{\circ} denotes the interior of K.K.

Remark 5.1.

We can take T5=0T_{5}=0 if (M,g,f)(M,g,f) is strongly tame.

Now we set

T0=max{T1,T2,T5}.T_{0}=\max\{T_{1},T_{2},T_{5}\}. (5.2)

(Note that the definition of T0T_{0} does not involve T3T_{3}; Cf. (2.3) and (2.4) for the description of T1T_{1} and T2T_{2}.) Before defining the Thom-Smale complex, there is still a subtle issue for noncompact cases. That is, the gradient vector field f-\nabla f may not be complete, i.e., its flow curves may not exist for all time. But notice that if we rescale the vector field by some positive function, we actually get the reparameterization of flow curves.

For this purpose, we fix a positve smooth function FF such that

F|MK=1T2|f|2.F|_{M_{K}}=\frac{1}{T^{2}|\nabla f|^{2}}.

Then we have

Lemma 5.2.

Ff-F\nabla f is a complete vector field.

Proof.

Let Φ~t\tilde{\Phi}^{t} be the flow generated by Ff-F\nabla f. We show that for any pMp\in M, there exists a universal ϵ0>0\epsilon_{0}>0, s.t. Φ~t(p)\tilde{\Phi}^{t}(p) is well defined on (ϵ0,ϵ0)(-\epsilon_{0},\epsilon_{0}). Hence, Ff-F\nabla f is complete.

Let L:={pM:d~T(p,K)1}L:=\{p\in M:\tilde{d}_{T}(p,K)\leq 1\}. It suffices to show that for any pML,p\in M-L, Φ~t(p)\tilde{\Phi}^{t}(p) is well defined on (1,1)(-1,1), since LL is compact.

But on MKM-K, F1g(Ff,Ff)=1T2F^{-1}g(-F\nabla f,-F\nabla f)=\frac{1}{T^{2}}, and (M,F1g)(M,F^{-1}g) is complete, hence Φ~t(p)\tilde{\Phi}^{t}(p), t(1,1)t\in(-1,1) is a geodesic (See Lemma 5.5) inside MKM-K for pLp\notin L.

Let xx be a critical point of the Morse function ff, Ws(x)W^{s}(x) and Wu(x)W^{u}(x) be the stable and unstable manifold of xx with respect to flow Φ~t\tilde{\Phi}^{t} defined in Lemma 5.2 (See Chapter 6 in [21]). We will further assume that ff satisfies the Smale transversality condition, namely Ws(x)W^{s}(x) and Wu(y)W^{u}(y) intersect transversally. Then the Thom-Smale complex (C(Wu),)(C_{*}(W^{u}),\partial) is defined by

C(Wfu)=xCrit(f)Wu(x),C_{*}(W^{u}_{f})=\oplus_{x\in\mathrm{Crit}(f)}\mathbb{R}W^{u}(x),

and

Ci(Wu)=xCrit(f),nf(x)=iWu(x).C_{i}(W^{u})=\oplus_{x\in\mathrm{Crit}(f),n_{f}(x)=i}\mathbb{R}W^{u}(x).

To define the boundary operator, let xx and yy be critical points of ff, with nf(y)=nf(x)1n_{f}(y)=n_{f}(x)-1.

For xCrit(f),x\in\mathrm{Crit}(f), set

~Wu(x)=yCrit(f),nf(y)=nf(x)1m(x,y)Wu(y).\tilde{\partial}W^{u}(x)=\sum_{y\in\mathrm{Crit}(f),n_{f}(y)=n_{f}(x)-1}m(x,y)W^{u}(y).

Here the integer m(x,y)m(x,y) is the signed counts of the flow lines in Ws(y)Wu(x),W^{s}(y)\cap W^{u}(x),.

Remark 5.3.

With our nice choice of KK and FF, it is easy to see that for any x,yCrit(f)x,y\in\mathrm{Crit}(f), Ws(x)Wu(y)KW^{s}(x)\cap W^{u}(y)\subset K^{\circ}. Moreover, for any pWu(x)Kp\in W^{u}(x)-K, the curve {Φ~t(p):t0}K=\{\tilde{\Phi}^{t}(p):t\geq 0\}\cap K=\emptyset and limt|f|(Φ~t(p))limtρT(Φ~t(p))=.\lim_{t\to\infty}|f|(\tilde{\Phi}^{t}(p))\approx\lim_{t\to\infty}\rho_{T}(\tilde{\Phi}^{t}(p))=\infty. Moreover, since d~T(p,p0)=Td~1(p,p0)\tilde{d}_{T}(p,p_{0})=T\tilde{d}_{1}(p,p_{0}), we can actually choose KK, such that it is independent of T.T. Thus, just like the compact case, by the transversality, m(x,y)m(x,y) is well defined.

We will prove in Section 7.3 that under our tameness condition, ~2=0.\tilde{\partial}^{2}=0. Thus, C(Wu,~)C_{*}(W^{u},\tilde{\partial}) is a complex.

Recall that the Witten instanton complex FTf[0,1],F_{Tf}^{[0,1],*} is the finite dimensional space generated by the eigenforms of Tf\Box_{Tf} with eigenvalue lying in [0,1][0,1]. By the discussion in the previous subsection, the cohomology of the Witten instanton comple is H(2)(M,df)H^{*}_{(2)}(M,d_{f}).

To prove Theorem 1.3, we now consider the following chain map 𝒥:(FTf[0,1],,dTf)C((Wu),~).\mathcal{J}:(F_{Tf}^{[0,1],*},d_{Tf})\mapsto C^{*}((W^{u})^{\prime},\tilde{\partial}^{\prime}). Here C((Wu),~)C^{*}((W^{u})^{\prime},\tilde{\partial}^{\prime}) denote the dual chain complex. Let Wu(x)W^{u}(x)^{\prime} be the dual basis of Wu(x).W^{u}(x). Then

𝒥ω=xCrit(f)Wu(x)Wu(x)exp(Tf)ω.\mathcal{J}\omega=\sum_{x\in\mathrm{Crit}(f)}W^{u}(x)^{\prime}\int_{W^{u}(x)}\exp(Tf)\omega.

However there is a technical issue here we need to address. When Wu(x)¯\overline{W^{u}(x)} is compact, the integral Wu(x)exp(Tf)ω\int_{W^{u}(x)}\exp(Tf)\omega is clearly well defined, but Wu(x)¯\overline{W^{u}(x)} here may be noncompact. We will be content here only with the wel-definedness of the map and leave the proof that 𝒥\mathcal{J} is indeed a chain map to Section 7.3.

Let r>0r>0 small enough, Brnf(x)(x)KB_{r}^{n_{f}(x)}(x)\subset K be the nf(x)n_{f}(x)-dimensional ball in Wu(x)W^{u}(x) with center xx and radius rr with respect to metric g.g. As before, let Φ~t\tilde{\Phi}^{t} be the flow generated by Ff-F\nabla f. Then Wu(x)=limtΦ~t(Brnf(x)(x)).W^{u}(x)=\lim_{t\to\infty}\tilde{\Phi}^{t}(B_{r}^{n_{f}(x)}(x)).

Therefore, for any ωFTf[0,1],\omega\in F^{[0,1],*}_{Tf}

|Wu(x)exp(Tf)ω|=|limt(Φ~t)(Brnf(x)(x))exp(Tf)ω|\displaystyle|\int_{W^{u}(x)}\exp(Tf)\omega|=|\lim_{t\to\infty}\int_{(\tilde{\Phi}^{t})(B_{r}^{n_{f}(x)}(x))}\exp(Tf)\omega|
Cexp(Tf(x))limtBrnf(x)(x)|(Φ~t)ω||det((Φ~t))|dvol\displaystyle\leq C\exp(Tf(x))\lim_{t\to\infty}\int_{B_{r}^{n_{f}(x)}(x)}|(\tilde{\Phi}^{t})^{*}\omega||\det((\tilde{\Phi}^{t})_{*})|\mathrm{dvol}

The well definedness of 𝒥\mathcal{J} is now reduced to the following two technical lemmas, as well as Theorem 1.1 and the well tameness of (M,g,f)(M,g,f).

Lemma 5.4.

Fix any yBrnf(x)(x)Φ~tK,y\in B_{r}^{n_{f}(x)}(x)-\tilde{\Phi}^{-t}K, we have

|Φt(y)|\displaystyle|\Phi^{t}_{*}(y)| C7(T)exp(ϵρT(p))\displaystyle\leq C_{7}(T)\exp(\epsilon\rho_{T}(p))

Hence,

|det(Φt)(y)|C7(T)exp(nf(x)ϵρT(p))\displaystyle|\det(\Phi^{t})_{*}(y)|\leq C_{7}(T)\exp(n_{f}(x)\epsilon\rho_{T}(p))

Here C7C_{7} is a constant independent of y.y.

Proof.

Let ee be a unit tangent vector of Wu(x)W^{u}(x) at y,y, Extend ee to a local unit vector field near yy via parallel transport along radial geodesics. Denote

Yf=FfY_{f}=-F\nabla f (5.3)

Noting that from (5.1)

|Φ~te(Φ~t)(Yf)|ϵT|(Φ~t)e|,|\nabla_{\tilde{\Phi}^{t}_{*}e}(\tilde{\Phi}^{t})_{*}(Y_{f})|\leq\frac{\epsilon}{T}|(\tilde{\Phi}^{t})_{*}e|,

we have

|tg((Φ~t)e(y),(Φ~t)e(y))|\displaystyle|\frac{\partial}{\partial t}g((\tilde{\Phi}^{t})_{*}e(y),(\tilde{\Phi}^{t})_{*}e(y))|
=2|g((Φ~t)e(y)(Φ~t)Yf,(Φ~t)e(y))|\displaystyle=2|g(\nabla_{(\tilde{\Phi}^{t})_{*}e(y)}(\tilde{\Phi}^{t})_{*}Y_{f},(\tilde{\Phi}^{t})_{*}e(y))|
2ϵT|g((Φ~t)e(y),(Φ~t)e(y))|.\displaystyle\leq\frac{2\epsilon}{T}|g((\tilde{\Phi}^{t})_{*}e(y),(\tilde{\Phi}^{t})_{*}e(y))|.

By a classical result in ODE, we have

g((Φ~t)e(y),(Φ~t)e(y))C8exp(2ϵtT).g((\tilde{\Phi}^{t})_{*}e(y),(\tilde{\Phi}^{t})_{*}e(y))\leq C_{8}\exp(\frac{2\epsilon t}{T}).

Now our lemma follows from teh following, Lemma 5.5. ∎

Lemma 5.5.

Suppose t>0t>0 is big enough, yBrnf(x)(x)Φ~tKy\in B_{r}^{n_{f}(x)}(x)-\tilde{\Phi}^{-t}K. Then there exists a constant C9=C9(K,|f|2|K)>0,C_{9}=C_{9}(K,|\nabla f|^{2}|_{K})>0, s.t.

|ρT(Φ~t(y))tT|<C9T.|\rho_{T}(\tilde{\Phi}^{t}(y))-\frac{t}{T}|<C_{9}T.
Proof.

For any yBrnf(x)(x)Φ~tKy\in B_{r}^{n_{f}(x)}(x)-\tilde{\Phi}^{t}K, we claim that Φ(y)~s,s[0,t]\tilde{\Phi(y)}^{s},s\in[0,t] is one of the shortest smooth curve on (M,g~T)(M,\tilde{g}_{T}) connecting yy to Φ~t(y).\tilde{\Phi}^{t}(y).

Granted, since g~T(Yf(p),Yf(p))=1T2\tilde{g}_{T}(Y_{f}(p),Y_{f}(p))=\frac{1}{T^{2}} for all pKp\notin K, the claim gives

|ρT(Φt(y))tT|=|d~T(Φt(y),K)d~T(Φt(y),y)|\displaystyle|\rho_{T}(\Phi^{t}(y))-\frac{t}{T}|=|\tilde{d}_{T}(\Phi^{t}(y),K)-\tilde{d}_{T}(\Phi^{t}(y),y)|
TsuppK|f|diam(K),\displaystyle\leq T\sup_{p\in K}|\nabla f|\mathrm{diam}(K),

where d~T\tilde{d}_{T} is the distance induced by g~T\tilde{g}_{T}, diam(K)\mathrm{diam}(K) is the diameter of KK with respect to metric g.g.

We now prove the claim (See [11] Lemma A 2.2 for another proof):

First, Let’s show that γ(s):=Φ~s(y),s[0,r]\gamma(s):=\tilde{\Phi}^{s}(y),s\in[0,r] is a geodesic for any r>0r>0:

Let e~1T(s),,e~nT(s)\tilde{e}^{T}_{1}(s),...,\tilde{e}^{T}_{n}(s) be a local orthomormal frame on γ\gamma with e~1T=~Tf=γ\tilde{e}^{T}_{1}=-\tilde{\nabla}^{T}f=\gamma^{\prime}. In order to prove γ′′=0,\gamma^{\prime\prime}=0, it suffices to prove g~T(γ′′,e~iT)=0,i2.\tilde{g}_{T}(\gamma^{\prime\prime},\tilde{e}^{T}_{i})=0,i\geq 2.

Let ~T\tilde{\nabla}^{T} be the Levi-Civita connection induced by g~T,\tilde{g}_{T}, then

g~T(γ′′,e~iT)\displaystyle\tilde{g}_{T}(\gamma^{\prime\prime},\tilde{e}^{T}_{i}) =g~T(~~TfT~Tf,e~iT)\displaystyle=\tilde{g}_{T}(\tilde{\nabla}^{T}_{\tilde{\nabla}^{T}f}\tilde{\nabla}^{T}f,\tilde{e}^{T}_{i})
=g~T(~Tf,[~Tf,e~iT])\displaystyle=-\tilde{g}_{T}(\tilde{\nabla}^{T}f,[\tilde{\nabla}^{T}f,\tilde{e}^{T}_{i}])
=[~Tf,e~iT]f\displaystyle=-[\tilde{\nabla}^{T}f,\tilde{e}^{T}_{i}]f
=~Tfg~T(e~iT,~Tf)+e~iTg~T(~Tf,~Tf)\displaystyle=-\tilde{\nabla}^{T}f\tilde{g}_{T}(\tilde{e}^{T}_{i},\tilde{\nabla}^{T}f)+\tilde{e}^{T}_{i}\tilde{g}_{T}(\tilde{\nabla}^{T}f,\tilde{\nabla}^{T}f)
=0.\displaystyle=0.

We now prove that γ\gamma is the shortest geodesic connecting pp and γ(r)\gamma(r) in (M,g~T)(M,\tilde{g}_{T}), for all r>0r>0:

Assume that σ\sigma is another normal geodesic connecting pp and γ(r)\gamma(r) induced by g~T.\tilde{g}_{T}. Then g~T(σ(0),~Tf(p))<1.\tilde{g}_{T}(\sigma^{\prime}(0),\tilde{\nabla}^{T}f(p))<1. Set a(s)=fγ(s),b(s)=fσ(s),a(s)=f\circ\gamma(s),b(s)=f\circ\sigma(s), then we have a(0)=b(0),a(0)=b(0), and a(s)=1,b(s)=g~T(σ(s),~Tfγ(s))1.a^{\prime}(s)=-1,b^{\prime}(s)=-\tilde{g}_{T}(\sigma^{\prime}(s),\tilde{\nabla}^{T}f\circ\gamma(s))\geq-1. Hence by a comparison theorem in ODE, we must have a(s)b(s).a(s)\leq b(s). Assume that σ(r)=γ(r),\sigma(r^{\prime})=\gamma(r), then we can see r=Length(σ),r^{\prime}=Length(\sigma), also we have a(r)b(r)=a(r).a(r^{\prime})\leq b(r^{\prime})=a(r). Since aa is decreasing, we must have rr.r^{\prime}\geq r.

By now, we can see that Φ~s(y),s[0,t]\tilde{\Phi}^{s}(y),s\in[0,t] is one of shortest geodesic connecting yy and Φ~t(y).\tilde{\Phi}^{t}(y).

We now note the following lemma which plays an important role in estimating the eigenforms previously.

Lemma 5.6.

Suppose TT0.T\geq T_{0}. Then for any qMq\in M, there exists C>0C>0, such that

|f|2(q)Cexp(cfbTρT(q))|\nabla f|^{2}(q)\leq C\exp(\frac{c_{f}}{bT}\rho_{T}(q))
Proof.

Let γ:[0,ρT(q)]M\gamma:[0,\rho_{T}(q)]\mapsto M be a normal minimal g~T\tilde{g}_{T}-geodesic connecting p0p_{0} and qq. Then we have g(γ,γ)=1b2T2|f|2g(\gamma^{\prime},\gamma^{\prime})=\frac{1}{b^{2}T^{2}|\nabla f|^{2}} outside K.K. As KK is compact and |f||\nabla f| is bounded on KK, we will assume without loss of generality that γ\gamma lies outside of KK.

Let h(t)=|f|2γ,h(t)=|\nabla f|^{2}\circ\gamma, then

h(t)=g(γf,f)1bT|2f|cfbT|f|2=cfbTh(t),h^{\prime}(t)=g(\nabla_{\gamma^{\prime}}\nabla f,\nabla f)\leq\frac{1}{bT}|\nabla^{2}f|\leq\frac{c_{f}}{bT}|\nabla f|^{2}=\frac{c_{f}}{bT}h(t),

Hence |f|2(q)Cexp(cfbTρT(q)).|\nabla f|^{2}(q)\leq C\exp(\frac{c_{f}}{bT}\rho_{T}(q)).

Here we are gives a direct proof of the isomorphism of H(2)(M,dTf)H^{*}_{(2)}(M,d_{Tf}) and HdR(M,Uc)H^{*}_{dR}(M,U_{c}) under the assumption that ff is proper.

Theorem 5.1.

Assume that ff is proper. Set I=infpKf(p),S=suppKf(p)I=\inf_{p\in K}f(p),S=\sup_{p\in K}f(p) and fix c>|I|+|S|+2c>|I|+|S|+2. Then for Uc={pM:f(p)<c},U_{c}=\{p\in M:f(p)<-c\}, (Ω(2)(M),dTf)(\Omega_{(2)}^{*}(M),d_{Tf}) and (Ω(M,Uc),d)(\Omega^{*}(M,U_{c}),d) are quasi-isomorphic.

Proof.

We may as well set K=f1[I,S].K=f^{-1}[I,S]. Motivated by [7], consider (Cone,dC)(Cone^{*},d_{C}), where Conej=Ωj(M)Ωj1(Uc),Cone^{j}=\Omega^{j}(M)\oplus\Omega^{j-1}(U_{c}),

dC(ϕ,ϕ)=(dϕ,(1)j(d|Ucϕ+ϕ|Uc)).d_{C}(\phi,\phi^{\prime})=(d\phi,(-1)^{j}(-d|_{U_{c}}\phi^{\prime}+\phi|_{U_{c}})).

Then (Cone,dc)(Cone^{*},d_{c}) and (Ω(M,Uc),d)(\Omega^{*}(M,U_{c}),d) are quasi-isomorphic.

Set Uc={pM:f(p)>c}U_{c}^{\prime}=\{p\in M:f(p)>c\}, U=UcUcU=U_{c}\cup U_{c}^{\prime}. Let Φ¯t\bar{\Phi}^{t} be the flow in UU generated by Xf=fb2T2|f|2X_{f}=-\frac{\nabla f}{b^{2}T^{2}|\nabla f|^{2}} on UcU_{c}, and Xf=fb2T2|f|2X_{f}=\frac{\nabla f}{b^{2}T^{2}|\nabla f|^{2}} on UcU_{c}^{\prime}.

Define a map :FTf[0,1],jConej,\mathcal{L}:F_{Tf}^{[0,1],j}\mapsto Cone^{j},

ω(exp(Tf)ω,(1)j0(Φ¯s)(exp(Tf)ιXfω)𝑑s)\omega\mapsto(\exp(Tf)\omega,(-1)^{j}\int_{0}^{\infty}({\bar{\Phi}}^{s})^{*}(\exp(Tf)\iota_{X_{f}}\omega)ds)

By Theorem 1.1, and similar argument in Lemma 5.5, we can see that |(Φ¯s)ιXfω|Cexp(0saT𝑑t)Cexp(aTs).|({\bar{\Phi}}^{s})^{*}\iota_{X_{f}}\omega|\leq C\exp(-\int_{0}^{s}aTdt)\leq C\exp(-aTs). Hence, \mathcal{L} is well defined.

\mathcal{L} is a chain map, since

(dTfω)\displaystyle\mathcal{L}(d_{Tf}\omega) =(exp(Tf)dTfω,(1)j+10(Φ¯s)(exp(Tf)ιXfdTfω)𝑑s)\displaystyle=(\exp(Tf)d_{Tf}\omega,(-1)^{j+1}\int_{0}^{\infty}({\bar{\Phi}}^{s})^{*}(\exp(Tf)\iota_{X_{f}}d_{Tf}\omega)ds)
=(exp(Tf)dTfω,(1)j+10(Φ¯s)(exp(Tf)ιXfdTfω)𝑑s)\displaystyle=(\exp(Tf)d_{Tf}\omega,(-1)^{j+1}\int_{0}^{\infty}({\bar{\Phi}}^{s})^{*}(\exp(Tf)\iota_{X_{f}}d_{Tf}\omega)ds)
=(d(exp(Tf)ω),(1)j+10(Φ¯s)(ιXfd(exp(Tf)ω)ds))\displaystyle=(d(\exp(Tf)\omega),(-1)^{j+1}\int_{0}^{\infty}({\bar{\Phi}}^{s})^{*}(\iota_{X_{f}}d(\exp(Tf)\omega)ds))
=(d(exp(Tf)ω),(1)j+10(Φ¯s)(LXf(exp(Tf)ω)+dιXf(exp(Tf)ωds)\displaystyle=(d(\exp(Tf)\omega),(-1)^{j+1}\int_{0}^{\infty}({\bar{\Phi}}^{s})^{*}(L_{X_{f}}(\exp(Tf)\omega)+d\iota_{X_{f}}(\exp(Tf)\omega ds)
=(d(exp(Tf)ω),(1)j+10dds(Φ¯s)(exp(Tf)ω)+d(exp(Tf)ιXfω)ds)\displaystyle=(d(\exp(Tf)\omega),(-1)^{j+1}\int_{0}^{\infty}\frac{d}{ds}({\bar{\Phi}}^{s})^{*}(\exp(Tf)\omega)+d(\exp(Tf)\iota_{X_{f}}\omega)ds)
=(d(exp(Tf)ω),(1)j+1(exp(Tf)ω+d0(Φ¯s)(exp(Tf)ιXfω)ds)\displaystyle=(d(\exp(Tf)\omega),(-1)^{j+1}(-\exp(Tf)\omega+d\int_{0}^{\infty}({\bar{\Phi}}^{s})^{*}(\exp(Tf)\iota_{X_{f}}\omega)ds)
=dC(ω).\displaystyle=d_{C}\mathcal{L}(\omega).

Hence \mathcal{L} induces a homomorphism (still denote it by \mathcal{L}) between H(Ω(2)(M),dTf)H^{*}(\Omega_{(2)}^{\bullet}(M),d_{Tf}) and H(Cone,dC).H^{*}(Cone^{\bullet},d_{C}). The proof of the fact that \mathcal{L} is a bijection is tedious, which will be given in Subsection 7.6. ∎

6 An application of Theorem 1.3

Let (M,g)(M,g) be an oriented, compact Riemannian manifold with boundary M,\partial M, and near the boundary. Let M~:=MT,\tilde{M}:=M\cup T, where TM×[1,)T\cong\partial M\times[1,\infty). Then we can extend the metric gg to M~\tilde{M}, s.t. near the infinity, the metric gg on M~\tilde{M} is of product type, i.e. gM+dr2g_{\partial M}+dr^{2}. It’s easy to see that (M~,g)(\tilde{M},g) has bounded geometry. We have the following technical lemma

Lemma 6.1.

Given a transversal Morse function ff, a partition of boundaries N+=N1+N2+,N^{+}=N^{+}_{1}\sqcup N^{+}_{2}, N=N1N2.N^{-}=N^{-}_{1}\sqcup N^{-}_{2}. We are able to extend ff to a function f~\tilde{f} on M~,\tilde{M}, s.t.

  1. 1.

    |f~|(x)|\tilde{f}|(x)\to\infty as xx\to\infty;

  2. 2.

    f~<0\tilde{f}<0 on (N1+N2)×(a,];(N_{1}^{+}\sqcup N_{2}^{-})\times(a,\infty];

  3. 3.

    f~\tilde{f} has critical points Crit(f~)=Cirt,(f)CirtN1++,(f)CirtN1,(f)\mathrm{Crit}^{*}(\tilde{f})=Cirt^{\circ,*}(f)\cup Cirt^{+,*}_{N^{+}_{1}}(f)\cup Cirt^{-,*}_{N^{-}_{1}}(f);

  4. 4.

    (M~,g,f~)(\tilde{M},g,\tilde{f}) is well tame.

Proof.

Use notation x=(x,r),xM,r(0,1]x=(x^{\prime},r),x^{\prime}\in\partial M,r\in(0,1] to denote xM×(0,1].x\in\partial M\times(0,1]. Since ff is a transversal Morse function, there exists s0<1s_{0}<1, s.t. |fr(x,r)|0|\frac{\partial f}{\partial r}(x^{\prime},r)|\neq 0 on M×(s0,1].\partial M\times(s_{0},1].

Hence, by considering the Taylor expansion of f(x,r)f(x^{\prime},r) with respect to r,r, there is a smooth function θ\theta on M×(s0,1],\partial M\times(s_{0},1], s.t.

  1. 1.

    f(x,r)=f(x,1)+fr(x,1)θ(x,r)f(x^{\prime},r)=f(x^{\prime},1)+\frac{\partial f}{\partial r}(x^{\prime},1)\theta(x^{\prime},r);

  2. 2.

    θ(x,r)=(r1)+o((r1))\theta(x^{\prime},r)=(r-1)+o((r-1)) near M×{1};\partial M\times\{1\};

  3. 3.

    θr(x,r)=1+o(1).\frac{\partial\theta}{\partial r}(x^{\prime},r^{\prime})=1+o(1).

Assume that s0s_{0} is close enough to 11, s.t. θ(x,r)<min{(r1)2,1/2(r1)},θr(x,r)>0\theta(x^{\prime},r)<\min\{-(r-1)^{2},1/2(r-1)\},\frac{\partial\theta}{\partial r}(x^{\prime},r^{\prime})>0 on M×(s0,1].\partial M\times(s_{0},1].

Let η1\eta_{1} be a smooth function on (,),(-\infty,\infty), s.t.

  1. 1.

    0<η<10<\eta<1 on (s0,1)(s_{0},1), and η0\eta\equiv 0 on (,s0)(-\infty,s_{0}), η1\eta\equiv 1 on (1,)(1,\infty);

  2. 2.

    η(r)>0,r(s0,1)\eta^{\prime}(r)>0,\forall r\in(s_{0},1);

Then xN1+N1\forall x^{\prime}\in N^{+}_{1}\cup N^{-}_{1}, let

f~(x,r)=f(x,r)=f(x,1)+fr(x,1)((1η(r))θ(x,r)+η(r)(r1)2);\tilde{f}(x^{\prime},r)=f(x^{\prime},r)=f(x^{\prime},1)+\frac{\partial f}{\partial r}(x^{\prime},1)((1-\eta(r))\theta(x^{\prime},r)+\eta(r)(r-1)^{2});

xN2+N2\forall x^{\prime}\in N^{+}_{2}\cup N^{-}_{2}, let

f~(x,r)=f(x,r)=f(x,1)+fr(x,1)((1η(r))θ(x,r)+η(r)1/2(r1)).\tilde{f}(x^{\prime},r)=f(x^{\prime},r)=f(x^{\prime},1)+\frac{\partial f}{\partial r}(x^{\prime},1)((1-\eta(r))\theta(x^{\prime},r)+\eta(r)1/2(r-1)).

It’s easy to verify that f~\tilde{f} satisfy our conditions. ∎

Since the Thom-Smale complex (C(Wu),~)(C^{*}(W^{u}),\tilde{\partial}^{\prime}) of f~\tilde{f} induces a differential operator ~\tilde{\partial}^{\prime} on Crit(f)\mathrm{Crit}^{*}(f), Theorem 1.5 follows easily from Theorem 1.3.

7 The Agmon Estimate

In this section we carry out the main technical estimates of the paper.

7.1 Proof of Lemma 3.1

Proof.

Our proof is adapted from Theorem 1.5 in [1].

Let L={pM:ρT(p)2}.L=\{p\in M:\rho_{T}(p)\leq 2\}. Let ηkCc()\eta_{k}\in C^{\infty}_{c}(\mathbb{R}) (kk large enough) be a smooth bump function such that

ηk(t)={0, If |t|<1 or |t|>k+1;1, If |t|(2,k),\eta_{k}(t)=\begin{cases}0,\mbox{ If $|t|<1$ or $|t|>k+1$;}\\ 1,\mbox{ If $|t|\in(2,k)$},\end{cases}

and |ηk(t)|2,|\eta_{k}^{\prime}(t)|\leq 2, ηk(t)[0,1],t.\eta_{k}(t)\in[0,1],\forall t\in\mathbb{R}.

Set ρT,j=min{ρT,j}\rho_{T,j}=\min\{\rho_{T},j\}, and

λT,j={λT, if ρT<j,0, otherwise.\lambda_{T,j}=\begin{cases}\lambda_{T},\mbox{ if }\rho_{T}<j,\\ 0,\mbox{ otherwise}\end{cases}.

Clearly |ρT,j|2=λT,j|\nabla\rho_{T,j}|^{2}=\lambda_{T,j} a.e. and λTλT,j.\lambda_{T}\geq\lambda_{T,j}.

Now set φk,j=(ηkρT)exp(bρT,j)\varphi_{k,j}=(\eta_{k}\circ\rho_{T})\exp(b\rho_{T,j}). Then by assumption, we have

Mu(φk,j2u)+λT(uφk,j)2dvolMwφk,j2udvol.\int_{M}\nabla u\nabla(\varphi_{k,j}^{2}u)+\lambda_{T}(u\varphi_{k,j})^{2}\mathrm{dvol}\leq\int_{M}w\varphi_{k,j}^{2}u\mathrm{dvol}.

Noting that u(φk,j2u)=|(φk,ju)|2|φk,j|2u2|φk,j|2u2,\nabla u\nabla(\varphi_{k,j}^{2}u)=|\nabla(\varphi_{k,j}u)|^{2}-|\nabla\varphi_{k,j}|^{2}u^{2}\geq-|\nabla\varphi_{k,j}|^{2}u^{2}, we have

MK(λT|uφk,j|2|u|2|φk,j|2)dvolMKwuφk,j2dvol.\int_{M-K}(\lambda_{T}|u\varphi_{k,j}|^{2}-|u|^{2}|\nabla\varphi_{k,j}|^{2})\mathrm{dvol}\leq\int_{M-K}wu\varphi_{k,j}^{2}\mathrm{dvol}. (7.1)

Since (we now omit the volume form dvol\mathrm{dvol} in what follows)

MKwuφk,j211b2MK(λT)1w2φk,j2+1b24MKλTu2φk,j2,\int_{M-K}wu\varphi_{k,j}^{2}\leq\frac{1}{1-b^{2}}\int_{M-K}(\lambda_{T})^{-1}w^{2}\varphi_{k,j}^{2}+\frac{1-b^{2}}{4}\int_{M-K}\lambda_{T}u^{2}\varphi_{k,j}^{2},

and

|φk,j|2\displaystyle|\nabla\varphi_{k,j}|^{2} 1+b22(ηkρT)2|ρT,j|2exp(2bρT,j)+1+b21b2(ηkρT)2|ρT|2exp(2bρT,j)\displaystyle\leq\frac{1+b^{2}}{2}(\eta_{k}\circ\rho_{T})^{2}|\nabla\rho_{T,j}|^{2}\exp(2b\rho_{T,j})+\frac{1+b^{2}}{1-b^{2}}(\eta_{k}^{\prime}\circ\rho_{T})^{2}|\nabla\rho_{T}|^{2}\exp(2b\rho_{T,j})
=1+b22(ηkρT)2λT,jexp(2bρT,j)+1+b21b2(ηkρT)2λTexp(2bρT,j),\displaystyle=\frac{1+b^{2}}{2}(\eta_{k}\circ\rho_{T})^{2}\lambda_{T,j}\exp(2b\rho_{T,j})+\frac{1+b^{2}}{1-b^{2}}(\eta_{k}^{\prime}\circ\rho_{T})^{2}\lambda_{T}\exp(2b\rho_{T,j}),

by (7.1), we have

3+b24MKλT(ηkρT)2u2exp(2bρT,j)1+b22MKλT,j(ηkρT)2u2exp(2bρT,j)11b2MKw2(ηkρT)2λT1exp(2bρT,j)+1+b21b2MKu2(ηkρT)2λTexp(2bρT,j)11b2MKw2λT1exp(2aρT)+21+b21b2LKu2λTexp(2bρT)+21+b21b2B~k+1B~ku2λTexp(2bj)\displaystyle\begin{split}&\frac{3+b^{2}}{4}\int_{M-K}\lambda_{T}(\eta_{k}\circ\rho_{T})^{2}u^{2}\exp(2b\rho_{T,j})-\frac{1+b^{2}}{2}\int_{M-K}\lambda_{T,j}(\eta_{k}\circ\rho_{T})^{2}u^{2}\exp(2b\rho_{T,j})\\ &\leq\frac{1}{1-b^{2}}\int_{M-K}w^{2}(\eta_{k}\circ\rho_{T})^{2}\lambda_{T}^{-1}\exp(2b\rho_{T,j})\\ &+\frac{1+b^{2}}{1-b^{2}}\int_{M-K}u^{2}(\eta_{k}^{\prime}\circ\rho_{T})^{2}\lambda_{T}\exp(2b\rho_{T,j})\\ &\leq\frac{1}{1-b^{2}}\int_{M-K}w^{2}\lambda_{T}^{-1}\exp(2a\rho_{T})\\ &+2\frac{1+b^{2}}{1-b^{2}}\int_{L-K}u^{2}\lambda_{T}\exp(2b\rho_{T})+2\frac{1+b^{2}}{1-b^{2}}\int_{\tilde{B}_{k+1}-\tilde{B}_{k}}u^{2}\lambda_{T}\exp(2bj)\end{split} (7.2)

Letting k,k\to\infty, by the monotone convergence theorem and the fact that MλTu2<\int_{M}\lambda_{T}u^{2}<\infty, we have

3+b24MLλTu2exp(2bρT,j)1+b22MLλT,ju2exp(2bρT,j)\displaystyle\frac{3+b^{2}}{4}\int_{M-L}\lambda_{T}u^{2}\exp(2b\rho_{T,j})-\frac{1+b^{2}}{2}\int_{M-L}\lambda_{T,j}u^{2}\exp(2b\rho_{T,j})
11b2MKw2λT1exp(2bρT)+21+b21b2LKu2λTexp(2bρT).\displaystyle\leq\frac{1}{1-b^{2}}\int_{M-K}w^{2}\lambda_{T}^{-1}\exp(2b\rho_{T})+2\frac{1+b^{2}}{1-b^{2}}\int_{L-K}u^{2}\lambda_{T}\exp(2b\rho_{T}).

Now let j.j\to\infty. By the monotone convergence theorem again, we arrive at (3.1), as desired. ∎

7.2 Proof of Lemma 3.3

Proof.

Our proof is a standard argument of Moser iteration.

Assume that ρT(p)\rho_{T}(p) is big enough and rr is small enough, s.t. B2r(p)K=.B_{2r}(p)\cap K=\emptyset. Then on B2r(p)B_{2r}(p), we have

cuTfuΔucu\geq\Box_{Tf}u\geq\Delta u (7.3)

weakly.

Set r1=2r,rk+1=rk(1/2)kr,nk=(n/(n2))k1r_{1}=2r,r_{k+1}=r_{k}-(1/2)^{k}r,n_{k}=(n/(n-2))^{k-1}

Let ηkCc(B2r)\eta_{k}\in C^{\infty}_{c}(B_{2r}) be bump functions s.t.

ηk={1 on Brk+1,0 on B2rBrk,\eta_{k}=\begin{cases}1\mbox{ on $B_{r_{k+1}}$,}\\ 0\mbox{ on $B_{2r}-B_{r_{k}}$,}\end{cases}

and |ηk(q)|<2rk+1rk,|\nabla\eta_{k}(q)|<\frac{2}{r_{k+1}-r_{k}}, ηk(q)[0,1],qB2r.\eta_{k}(q)\in[0,1],\forall q\in B_{2r}.

Set um=min{u,m}u_{m}=\min\{u,m\}, and ϕ1=η12umH01(B2r).\phi_{1}=\eta_{1}^{2}u_{m}\in H_{0}^{1}(B_{2r}). Notice that ϕ1=0\phi_{1}=0 and ϕ1=0\nabla\phi_{1}=0 in {um}.\{u\geq m\}. Hence, by (7.3), we have

Br1c(um)2𝑑volB2rcuϕ1𝑑volB2ruϕ1dvol\displaystyle\int_{B_{r_{1}}}c(u_{m})^{2}dvol\geq\int_{B_{2r}}cu\phi_{1}dvol\geq\int_{B_{2r}}\nabla u\nabla\phi_{1}dvol
=B2rη12|um|2+2η1ηumumdvol\displaystyle=\int_{B_{2r}}\eta_{1}^{2}|\nabla u_{m}|^{2}+2\eta_{1}\nabla\eta\nabla u_{m}u_{m}dvol
B2rη12|um|21/2η12|um|22|ηum|2dvol\displaystyle\geq\int_{B_{2r}}\eta_{1}^{2}|\nabla u_{m}|^{2}-1/2\eta_{1}^{2}|\nabla u_{m}|^{2}-2|\nabla\eta u_{m}|^{2}dvol
B2rη12|um|21/2η12|um|22|ηum|2dvol\displaystyle\geq\int_{B_{2r}}\eta_{1}^{2}|\nabla u_{m}|^{2}-1/2\eta_{1}^{2}|\nabla u_{m}|^{2}-2|\nabla\eta u_{m}|^{2}dvol
1/2Br2|um|2𝑑vol4/(r2r1)2Br1|um|2𝑑vol\displaystyle\geq 1/2\int_{B_{r_{2}}}|\nabla u_{m}|^{2}dvol-4/(r_{2}-r_{1})^{2}\int_{B_{r_{1}}}|u_{m}|^{2}dvol

Hence, we have

Br2|um|2𝑑vol(2c+8/(r1r2)2)Br2c(um)2𝑑volC(n)/(r1r2)2Br2c(um)2𝑑vol.\int_{B_{r_{2}}}|\nabla u_{m}|^{2}dvol\leq(2c+8/(r_{1}-r_{2})^{2})\int_{B_{r_{2}}}c(u_{m})^{2}dvol\leq C(n)/(r_{1}-r_{2})^{2}\int_{B_{r_{2}}}c(u_{m})^{2}dvol.

By Sobolev inequality,

(Br1|um|2n2𝑑vol)1/n2C(n)/(r1r2)2Br2c(um)2𝑑vol.(\int_{B_{r_{1}}}|u_{m}|^{2n_{2}}dvol)^{1/n_{2}}\leq C(n)/(r_{1}-r_{2})^{2}\int_{B_{r_{2}}}c(u_{m})^{2}dvol.

That is

umL2n2(Br2)(C(n)/(r1r2))umL2n1(Br1)\|u_{m}\|_{L^{2n_{2}}(B_{r_{2}})}\leq(C(n)/(r_{1}-r_{2}))\|u_{m}\|_{L^{2n_{1}}(B_{r_{1}})}

Let m,m\to\infty, we have

uL2n2(Br2)(C(n)/(r1r2))uL2n1(Br1)\|u\|_{L^{2n_{2}}(B_{r_{2}})}\leq(C(n)/(r_{1}-r_{2}))\|u\|_{L^{2n_{1}}(B_{r_{1}})}

Consider ϕk=ηk2(um2nk1)H01(B2r).\phi_{k}=\eta_{k}^{2}(u_{m}^{2n_{k}-1})\in H_{0}^{1}(B_{2r}). By the same arguments as above, we have

uL2nk+1(Brk+1)(C(n)/(rkrk+1))1/(nk)uL2nk(Brk).\|u\|_{L^{2n_{k+1}}(B_{r_{k+1}})}\leq(C(n)/(r_{k}-r_{k+1}))^{1/(n_{k})}\|u\|_{L^{2n_{k}}(B_{r_{k}})}.

As a consequence,

uL(Br)=limkuL2nk(Brk)\displaystyle\|u\|_{L^{\infty}(B_{r})}=\lim_{k\to\infty}\|u\|_{L^{2n_{k}}(B_{r_{k}})}
CΠk=1(C(n)/(rkrk+1))1/(nk)uL2(B2r)\displaystyle\leq C\Pi_{k=1}^{\infty}(C(n)/(r_{k}-r_{k+1}))^{1/(n_{k})}\|u\|_{L^{2}(B_{2r})}
=C(C(n)/r)(k=11/(nk))2k=1k/nkuL2(B2r)\displaystyle=C(C(n)/r)^{(\sum_{k=1}^{\infty}{1/(n_{k})})}2^{\sum_{k=1}^{\infty}k/n_{k}}\|u\|_{L^{2}(B_{2r})}
C/rn/2uL2(B2r).\displaystyle\leq C/r^{n/2}\|u\|_{L^{2}(B_{2r})}.

We state two Lemmas that would be needed shortly,

Lemma 7.1.

Suppose u,wL2(M)u,w\in L^{2}(M), s.t. Tfuw\Box_{Tf}u\leq w in weak sense (Here we assume u0.u\geq 0.). For r>0r>0 small enough, pLp\notin L, let Br(p)B_{r}(p) be the geodesic ball around pp with radius rr induced by gg. Then there exist C2>0C_{2}>0, s.t.

supyBr(p)u(y)C2rn/2(uL2(B2r(p))+wL2(B2r(p))),\sup_{y\in B_{r}(p)}u(y)\leq\frac{C_{2}}{r^{n/2}}(\|u\|_{L^{2}(B_{2r}(p))}+\|w\|_{L^{2}(B_{2r}(p))}),

where C2C_{2} is a constant that depends only on dimension n.n.

Proof.

The proof is actually similar to the proof Lemma 3.3, but a little more complicated. See Theorem 4.1 in [10] for a reference. ∎

By the same argument as the proof of Theorem 1.1, we have

Lemma 7.2.

Let (M,g,f)(M,g,f) be well tame, wL2(M)w\in L^{2}(M), s.t.

MλT1|w|2exp(a′′ρT)𝑑vol<\int_{M}\lambda_{T}^{-1}|w|^{2}\exp(a^{\prime\prime}\rho_{T})dvol<\infty

for some a′′(0,b).a^{\prime\prime}\in(0,b). If ϕL2(M)\phi\in L^{2}(M) is a weak solution of Tfϕw,\Box_{Tf}\phi\leq w, then

|ϕ(p)|Cexp(a′′ρT(p)).|\phi(p)|\leq C\exp(-a^{\prime\prime}\rho_{T}(p)).

7.3 On the Thom-Smale complex

First, let’s recall the situation of the compact case. The following is a restatement of Proposition 6 in [15].

Proposition 7.3.

Let (N,g)(N,g) be a compact Riemannian manifold, ff be a Morse function. Assume that (N,g,f)(N,g,f) satisfies Thom-Smale transversality condition. Then, for any critical point xCrit(f)x\in\mathrm{Crit}(f) with Morse index nf(x)n_{f}(x), any ϕΩnf(x)1(M)\phi\in\Omega^{n_{f}(x)-1}(M), one has the following so called Stokes Formula

Wu(x)𝑑ϕ=yCrit(f),nf(y)=nf(x)1m(x,y)Wu(y)ϕ.\int_{W^{u}(x)}d\phi=\sum_{y\in\mathrm{Crit}(f),n_{f}(y)=n_{f}(x)-1}m(x,y)\int_{W^{u}(y)}\phi.

For our noncompact case with tame conditions and Thom-Smale transversality, we have similarly

Proposition 7.4.

For any critical point xCrit(f)x\in\mathrm{Crit}(f) with Morse index nf(x)n_{f}(x), any ϕΩcnf(x)1(M)\phi\in\Omega^{n_{f}(x)-1}_{c}(M), one has the following so called Stokes Formula

Wu(x)𝑑ϕ=yCrit(f),nf(y)=nf(x)1m(x,y)Wu(y)ϕ.\int_{W^{u}(x)}d\phi=\sum_{y\in\mathrm{Crit}(f),n_{f}(y)=n_{f}(x)-1}m(x,y)\int_{W^{u}(y)}\phi.

Before giving the proof of this proposition, we first draw a couple of consequneces.

Corollary 7.1.

Let ~:C(Wu)C1(Wu)\tilde{\partial}:C_{*}(W^{u})\mapsto C_{*-1}(W^{u}) be the map constructed in Section 5, then ~2=0\tilde{\partial}^{2}=0.

Proof.

Otherwise, ~2Wu(x)0\tilde{\partial}^{2}W^{u}(x)\neq 0. Then there exists ϕΩcnf(x)2(M),\phi\in\Omega_{c}^{n_{f}(x)-2}(M), s.t.

~2Wu(x)ϕ0.\int_{\tilde{\partial}^{2}W^{u}(x)}\phi\neq 0.

But by Proposition 7.4,

~2Wu(x)ϕ=Wu(x)d2ϕ=0\int_{\tilde{\partial}^{2}W^{u}(x)}\phi=\int_{W^{u}(x)}d^{2}\phi=0

Corollary 7.2.

Let ωFTf[0,1],nf(x)1\omega\in F^{[0,1],n_{f}(x)-1}_{Tf}, one has

Wu(x)exp(Tf)dTfω=yCrit(f),nf(y)=nf(x)1m(x,y)Wu(y)exp(Tf)ω.\int_{W^{u}(x)}\exp(Tf)d_{Tf}\omega=\sum_{y\in\mathrm{Crit}(f),n_{f}(y)=n_{f}(x)-1}m(x,y)\int_{W^{u}(y)}\exp(Tf)\omega.
Proof.

By Theorem 1.1 and Lemma 5.4, for any ϵ>0\epsilon>0, there exist ϕΩcnf1(M)\phi\in\Omega^{n_{f}-1}_{c}(M), s.t. for any yCrit(f)y\in\mathrm{Crit}(f) with nf(y)=nf(x)1n_{f}(y)=n_{f}(x)-1,

Wu(x)|exp(Tf)dTfωdϕ|<ϵ,Wu(y)|exp(Tf)ωϕ|<ϵ.\int_{W^{u}(x)}|\exp(Tf)d_{Tf}\omega-d\phi|<\epsilon,\int_{W^{u}(y)}|\exp(Tf)\omega-\phi|<\epsilon.

Now our Corollary follows from Proposition 7.4 trivially. ∎

Hence, the map 𝒥\mathcal{J} introduced in Section 5 is a chain map.

The proof of Proposition 7.4 follows from the following observations:

Observation 7.5.

Let (N,N)(N,\partial N) be compact manifold with boundary. Moreover, assume that near the boundary N\partial N, the manifold is of product type (0,1]×N.(0,1]\times\partial N. Suppose that ff is a Morse function on N[1/2,1]×NN-[1/2,1]\times\partial N. Then there exist a transversal Morse function f¯\bar{f} on NN, s.t. f¯|N[1/4,1]×N=f,f~[3/4,1]×N=r.\bar{f}|_{N-[1/4,1]\times\partial N}=f,\tilde{f}_{[3/4,1]\times\partial N}=r. Here rr is the standard coordinate on (0,1](0,1] factor.

The proof is essentially the same with Theorem 2.5 in [18].

Observation 7.6.

Let B~D(p0)\tilde{B}_{D}(p_{0}) be the ball with radius DD introduced in Section 5. Then for any y,zCrit(f)y,z\in\mathrm{Crit}(f), Wu(y)Ws(z)B~D.W^{u}(y)\cap W^{s}(z)\subset\tilde{B}_{D}^{\circ}. Moreover, if pB~D(p0)p\notin\tilde{B}_{D}(p_{0}) lies in an unstable manifold, then the curve {Φt(p):t0}B~D=.\{\Phi^{t}(p):t\geq 0\}\cap\tilde{B}_{D}=\emptyset. Here B~D\tilde{B}_{D}^{\circ} denotes the interior of B~D=B~D(p0),\tilde{B}_{D}=\tilde{B}_{D}(p_{0}), Φt\Phi^{t} is the flow generated by f.-\nabla f.

Proof.

It follows easily from the fact that for any pMp\in M, d~T(p,Φt(p))=|f(p)f(Φt(p))|/b\tilde{d}_{T}(p,\Phi^{t}(p))=|f(p)-f(\Phi^{t}(p))|/b, and ff is decreasing along the flow Φt.\Phi^{t}.

Now we are ready to prove Proposition 7.4

Proof.

For any ϕΩcnf(x)(M)\phi\in\Omega^{n_{f}(x)}_{c}(M), let

D=bT(suppCritfsupp(ϕ)|f(p)f(p0)|+supp,qCritfsupp(ϕ)|f(p)f(q)|),D=bT(\sup_{p\in\mathrm{Crit}{f}\cup{\mathrm{s}upp}(\phi)}|f(p)-f(p_{0})|+\sup_{p,q\in\mathrm{Crit}{f}\cup{\mathrm{s}upp}(\phi)}|f(p)-f(q)|),

we can find a compact submanifold (N,N)(N,\partial N) with boundary, s.t. supp(ϕ)B~DN.{\mathrm{s}upp}(\phi)\cup\tilde{B}_{D}\subset N^{\circ}. Here supp(ϕ){\mathrm{s}upp}(\phi) denotes the support of ϕ,\phi, NN^{\circ} denote the interior of N.N.

Now consider the double (DN=N+N,gDN)(DN=N^{+}\cup N^{-},g_{DN}) of NN, gDN|supp(ϕ)B~D=g.g_{DN}|_{{\mathrm{s}upp}(\phi)\cup\tilde{B}_{D}}=g. By Observation 7.5, we can find a Morse function f¯\bar{f} on DNDN, s.t. f¯|supp(ϕ)B~D=f.\bar{f}|_{{\mathrm{s}upp}(\phi)\cup\tilde{B}_{D}}=f. We may as well assume that (DN,gDN,f~)(DN,g_{DN},\tilde{f}) satisfy Thom-Smale transversality condition. Then for any y,zCrit(f¯)y,z\in\mathrm{Crit}(\bar{f}) with nf¯(y)=nf¯(z)+1,n_{\bar{f}}(y)=n_{\bar{f}}(z)+1, let mDN(y,z)m_{DN}(y,z) be the signed count of the number of flow lines in Wf¯uu(y)Wf¯s(z).W^{u}_{\bar{f}}u(y)\cap W^{s}_{\bar{f}}(z).

Then we claim:

  1. 1.

    By Observation 7.6, if y,zsupp(ω)B~Dy,z\in{\mathrm{s}upp}(\omega)\cup\tilde{B}_{D} are critical points of f¯\bar{f} with nf¯(y)=nf¯(z)+1n_{\bar{f}}(y)=n_{\bar{f}}(z)+1, we have mDN(y,z)=m(y,z).m_{DN}(y,z)=m(y,z).

  2. 2.

    If zsupp(ϕ)B~D(p0),ysupp(ϕ)B~Dz\notin{\mathrm{s}upp}(\phi)\cup\tilde{B}_{D}(p_{0}),y\in{\mathrm{s}upp}(\phi)\cap\tilde{B}_{D} are critical points of f¯\bar{f}, and Wf¯s(z)Wf¯u(y)W^{s}_{\bar{f}}(z)\cap W^{u}_{\bar{f}}(y)\neq\emptyset, then Wf¯u(z)supp(ϕ)=.W^{u}_{\bar{f}}(z)\cap{\mathrm{s}upp}(\phi)=\emptyset. This is because, by definition of DD, Claim in Lemma 5.5 and properties of unstable manifolds, f¯(z)<f¯(y)+|f(y)f(p0)|D/(bT)infpsupp(ϕ)f¯(p).\bar{f}(z)<\bar{f}(y)+|f(y)-f(p_{0})|-D/(bT)\leq\inf_{p\in{\mathrm{s}upp}(\phi)}\bar{f}(p).

As a result, by Proposition 7.3

Wfu(x)𝑑ϕ\displaystyle\int_{W^{u}_{f}(x)}d\phi =Wf¯u(x)𝑑ϕ=zCrit(f¯),nf¯(z)=nf¯(x)1mDN(x,z)Wf¯u(z)ϕ\displaystyle=\int_{W^{u}_{\bar{f}}(x)}d\phi=\sum_{z\in\mathrm{Crit}(\bar{f}),n_{\bar{f}}(z)=n_{\bar{f}}(x)-1}m_{DN}(x,z)\int_{W^{u}_{\bar{f}}(z)}\phi
=yCrit(f),nf(y)=nf(x)1mDN(x,y)Wfu(y)ϕ (By Claim 2)\displaystyle=\sum_{y\in\mathrm{Crit}({f}),n_{f}(y)=n_{f}(x)-1}m_{DN}(x,y)\int_{W^{u}_{f}(y)}\phi\mbox{ (By Claim \ref{c2})}
=yCrit(f),nf(y)=nf(x)1m(x,y)Wfu(y)ϕ (By Claim 1).\displaystyle=\sum_{y\in\mathrm{Crit}({f}),n_{f}(y)=n_{f}(x)-1}m(x,y)\int_{W^{u}_{f}(y)}\phi\mbox{ (By Claim \ref{c1})}.

7.3.1 An counterexample

On closing this subsection, let’s give a counterexample that when we drop the condition that f\nabla f has a positive lower bound near infinity, ~2=0\tilde{\partial}^{2}=0 may fail.

Consider the following heart shaped topological sphere SS with obvious height function ff. Then we have four critical points x,y,z,wx,y,z,w as indicated below. Let γ\gamma be a flow line connecting yy and ww, and remove a point pp on γ\gamma. Making a conformal change of metric near point pp, s.t. SpS-p is complete under this new metric. Now we can see that |f(q)|0|\nabla f(q)|\to 0, as qp.q\to p. Since the flow line is invariant under the conformal change of metric, γp\gamma-p is still a (broken) flow line. However, in this case, ~2x=w\tilde{\partial}^{2}x=w, which is nonzero.

In our previous arguments, the fact that |f||\nabla f| has a positive lower bounded near the infinity play a crucial role. It forces the value of Morse function ff goes to infinity along a flow line if that flow line flows to infinity (See also Remark 5.3 and Observation 7.6).

pγ\gammayxzw
Remark 7.7.

We would like to thank Shu Shen for providing this interesting example.

7.4 Isomorphism of H(C(Wu),)H^{*}(C^{\bullet}(W^{u}),\partial) and HdR(M,Uc)H^{*}_{dR}(M,U_{c})

For simplicity, we assume that ff is self-indexed Morse function, i.e., if xx is a critical point of ff with Morse index ii, we require f(x)=if(x)=i. Moreover, B1(x)B_{1}(x)

Let Vi=f1(,i+12]V_{i}=f^{-1}(-\infty,i+\frac{1}{2}], 0in0\leq i\leq n.

Recall that we assume in a neighborhood UxU_{x} of critical points xx of ff, we have coordinate system z=(z1,,zn),z=(z_{1},...,z_{n}), such that

f=z12znf(x)2+znf(x)+12++zn2,f=-z_{1}^{2}-...-z_{n_{f}(x)}^{2}+z_{n_{f}(x)+1}^{2}+...+z_{n}^{2},
g=dz12++dzn2,g=dz_{1}^{2}+...+dz_{n}^{2},

Moreover UxU_{x} is an Euclidean open ball around xx with radius 1.1. Also, these open balls are disjoint.

We have the following observation:

Lemma 7.8.

V0V_{0} can be written as disjoint union of xCrit(f),nf(x)=0U~x\cup_{x\in Crit(f),n_{f}(x)=0}\tilde{U}_{x} and VV, where VV is some open subset diffeomorphic to UcU_{c}, U~x\tilde{U}_{x} is an Euclidean ball around xx with radius 12\frac{1}{2}.

VnV_{n} is diffeomorphic to M.M.

Proof.

Let Xf:=f|f|2X_{f}:=\frac{\nabla f}{|\nabla f|^{2}}, Φt\Phi^{t} be the flow generated by XfX_{f}. Then we have

(Φc+12(Uc))(xCrit(f),nf(x)=0U~x)=.\left(\Phi^{c+\frac{1}{2}}(U_{c})\right)\cap\left(\cup_{x\in Crit(f),n_{f}(x)=0}\tilde{U}_{x}\right)=\emptyset.

This is because:

  • If f(p)c12f(p)\leq c-\frac{1}{2}, then f(Φc+12(p))<0f(\Phi^{c+\frac{1}{2}}(p))<0. Hence Φc+12(p)xCrit(f),nf(x)=0U~x\Phi^{c+\frac{1}{2}}(p)\notin\cup_{x\in Crit(f),n_{f}(x)=0}\tilde{U}_{x}.

  • If c12f(p)<cc-\frac{1}{2}\leq f(p)<c, and if Φc+12(p)U~x\Phi^{c+\frac{1}{2}}(p)\in\tilde{U}_{x} for some xCrit(f)x\in Crit(f) with Morse index nf(x)=0n_{f}(x)=0. Then Φc+12(p)Ws(x)\Phi^{c+\frac{1}{2}}(p)\in W^{s}(x), which implies pWs(x)p\in W^{s}(x). But this is impossible since f(p)<c<0=f(x)f(p)<-c<0=f(x).

Similarly, we can prove that VnV_{n} is diffeomorphic to M.M.

Let C(Vi,Uc)C_{*}(V_{i},U_{c}) be complex of relative singular chains. Then we have

C(Vn,Uc)C(Vn1,Uc)C(V0,Uc).C_{*}(V_{n},U_{c})\supset C_{*}(V_{n-1},U_{c})\supset\cdots C_{*}(V_{0},U_{c}).

By a similar spectral sequence argument as in the proof of Theorem 1.6 in [2] and Lemma 7.8, one can show that

H(C(Wu),)H(M,Uc).H_{*}(C^{\bullet}(W^{u}),\partial)\simeq H_{*}(M,U_{c}).

Thus, it follows from the universal coefficient theorem that

H(C(Wu),)H(M,Uc).H^{*}(C^{\bullet}(W^{u}),\partial)\simeq H^{*}(M,U_{c}).

7.5 Isomorphism of H(2)(M,dTf)H^{*}_{(2)}(M,d_{Tf}) and H(C(Wu),)H^{*}(C^{\bullet}(W^{u}),\partial)

We will first show that the chain map 𝒥:(FTf[0,1],,dTf)C((Wu),~)\mathcal{J}:(F_{Tf}^{[0,1],*},d_{Tf})\mapsto C^{*}((W^{u})^{\prime},\tilde{\partial}^{\prime}) defined in Section 5 is in fact an isomorphism when TT is sufficiently large. Hence 𝒥\mathcal{J} induced an isomorphism between H(2)(M,dTf)H^{*}_{(2)}(M,d_{Tf}) and H(C(Wu),)H^{*}(C^{\bullet}(W^{u}),\partial) in that case.

More precisely we will follow the arguments in Chapter 6 of [21], with necessary modification, to show that there exists T6>T0T_{6}>T_{0}, such that 𝒥\mathcal{J} is an isomorphism whenever T>T6.T>T_{6}. (We point out that the explicit description of T6T_{6} is more involved than T0T_{0}.) In fact, the only difference is that we need a more refined estimate in Theorem 6.17 of [21], that is:

|𝒫τx,Tτx,T|Cexp(aT(ρ+c))τx,TL2,|\mathcal{P}\tau_{x,T}-\tau_{x,T}|\leq C\exp(-aT(\rho+c))\|\tau_{x,T}\|_{L^{2}}, (7.4)

where 𝒫\mathcal{P} is the orthogonal projection from L2Λ(M)L^{2}\Lambda(M) to F[0,1],F^{[0,1],*}, and C, c are positive constants.

Here τx,T\tau_{x,T} is defined as follows. Notice that in Section 4, we require that in a neighborhoof UU of xx, the metric and Morse function is of the form (4.1). Hence, let αx\alpha_{x} be a bump function whose support is contained in UU, and αx1\alpha_{x}\equiv 1 in a neighborhood VV of xx. Now let

τx,T=αxexp(T2|z|2)dz1dznf(x).\tau_{x,T}=\alpha_{x}\exp(-T^{2}|z|^{2})dz_{1}\wedge\cdots\wedge dz_{n_{f}(x)}.

To obtain the estimate (7.4), pick a bump function η\eta with compact support, such that η1\eta\equiv 1 on KK. Then by our Agmon estimate, we have

|(1η)(𝒫τx,Tτx,T)|Cexp(aT(ρ+c))τx,TL2.|(1-\eta)(\mathcal{P}\tau_{x,T}-\tau_{x,T})|\leq C\exp(-aT(\rho+c))\|\tau_{x,T}\|_{L^{2}}.

The estimate of

|η(𝒫τx,Tτx,T)|Cexp(cT)τx,TL2|\eta(\mathcal{P}\tau_{x,T}-\tau_{x,T})|\leq C\exp(-cT)\|\tau_{x,T}\|_{L^{2}}

now follows from exactly the same argument in the proof of Theorem 6.17 of [21].

Now it remains to prove that when T(T0,T6]T\in(T_{0},T_{6}], H(2)(M,dTf)H^{*}_{(2)}(M,d_{Tf}) and H(C(Wu),)H^{*}(C^{\bullet}(W^{u}),\partial) are still isomorphic.

We only present the proof for the case when (M,g,f)(M,g,f) is strongly tame, the case of well tame being exactly the same except notationally. In this case, T0=0T_{0}=0. The idea is to show that if S>0,S>0, then for any T[7/8S,S]T\in[7/8S,S], H(2)(M,dTf)H^{*}_{(2)}(M,d_{Tf}) and H(2)(M,dSf)H^{*}_{(2)}(M,d_{Sf}) are isomorphic. Hence H(2)(M,dTf)H^{*}_{(2)}(M,d_{Tf}) is independent of T(0,)T\in(0,\infty), which finishes the proof of isomorphism of H(2)(M,dTf)H^{*}_{(2)}(M,d_{Tf}) and H(C(Wu),)H^{*}(C^{\bullet}(W^{u}),\partial).

For simplicity, we prove that H(2)(M,d7f)H^{*}_{(2)}(M,d_{7f}) and H(2)(M,d8f)H^{*}_{(2)}(M,d_{8f}) are isomorphic, the general case being similar.

Thus fix coefficient b=6364b=\frac{63}{64} in Lemma 2.1 and Theorem 1.1.

Define Mf:(F8f,[0,1],d8f)(Ω(2)(M),d7f)M_{f}:(F^{*,[0,1]}_{8f},d_{8f})\mapsto(\Omega^{*}_{(2)}(M),d_{7f}); wF8f,[0,1],\forall w\in F^{*,[0,1]}_{8f}, Mf(w)=exp(f)wM_{f}(w)=\exp(f)w. Similarly Mf:(F7f,[0,1],d7f)(Ω(2)(M),d8f)M_{-f}:(F^{*,[0,1]}_{7f},d_{7f})\mapsto(\Omega^{*}_{(2)}(M),d_{8f}); wF7f,[0,1],\forall w\in F^{*,[0,1]}_{7f}, Mf(w)=exp(f)w.M_{-f}(w)=\exp(-f)w.

Clearly these are chain maps once we check that MfM_{f} and MfM_{-f} are well defined. To this end, let’s verify that |f(p)||f(p0)|+1bTρT(p),|f(p)|\leq|f(p_{0})|+\frac{1}{bT}\rho_{T}(p), where p0p_{0} is the fix point in defining ρT.\rho_{T}. Indeed, let γ:[0,ρT(p)]\gamma:[0,\rho_{T}(p)] be a normal minimal geodesic connecting p0p_{0} and pp, in the metric g~T\tilde{g}_{T}. Then

|ddtfγ(t)|=|<~f,γ>g~T|1bT.|\frac{d}{dt}f\circ\gamma(t)|=|<\tilde{\nabla}f,\gamma^{\prime}>_{\tilde{g}_{T}}|\leq\frac{1}{bT}.

Now the L2L^{2} bound of Mf(w)M_{f}(w) (resp. Mf(w)M_{-f}(w)) follows by Theorem 1.1 and the standard volume comparison, Hence MfM_{f} induces a homomorphism (still denote it by MfM_{f}) from H(2)(M,d8f)H^{*}_{(2)}(M,d_{8f}) to H(2)(M,d7f)H^{*}_{(2)}(M,d_{7f}).

Our next step is to show that MfM_{f} is injective. Suppose we have wker(8f)w\in\ker(\Box_{8f}), s.t. MfwM_{f}w is exact, which means that we can find αIm(δ7f),\alpha\in Im(\delta_{7f}), s.t exp(f)w=d7fα(=(d7f+δ7f)α).\exp(f)w=d_{7f}\alpha(=(d_{7f}+\delta_{7f})\alpha).

Thus

7fα=(d7f+δ7f)exp(f)w=exp(f)d8fw+exp(2f)δ6fw\Box_{7f}\alpha=(d_{7f}+\delta_{7f})\exp(f)w=\exp(f)d_{8f}w+\exp(2f)\delta_{6f}w
=0+exp(2f)(δ8fwι2fw)=exp(2f)ι2fw.=0+\exp(2f)(\delta_{8f}w-\iota_{2f}w)=-\exp(2f)\iota_{2f}w.

By Lemma 7.2, |α|Cexp(1/3ρ7).|\alpha|\leq C\exp(-1/3\rho_{7}). Consequently, exp(f)αL2Λ(M)\exp(-f)\alpha\in L^{2}\Lambda^{*}(M), and w=d8fexp(f)αw=d_{8f}\exp(-f)\alpha is exact.

As a result, MfM_{f} is injective. Similarly, MfM_{-f} is also injective. Therefore, H(2)(M,d8f)H^{*}_{(2)}(M,d_{8f}) and H(2)(M,d7f)H^{*}_{(2)}(M,d_{7f}) are isomorphic.

7.6 \mathcal{L} is bijective

  • \mathcal{L} is injective:

    Let ωFTf[0,1],j\omega\in F^{[0,1],j}_{Tf}. Assume that (ω)\mathcal{L}(\omega) is exact, then there exists ϕΩj1(M),ϕΩj2(Uc),\phi\in\Omega^{j-1}(M),\phi^{\prime}\in\Omega^{j-2}(U_{c}), s.t.

    exp(Tf)ω=dϕ,ω=(1)j(ϕdϕ),\exp(Tf)\omega=d\phi,\omega^{\prime}=(-1)^{j}(\phi-d\phi^{\prime}),

    where ω=(1)j0(Φ¯s)(exp(Tf)ιXfω)𝑑s.\omega^{\prime}=(-1)^{j}\int_{0}^{\infty}(\bar{\Phi}^{s})^{*}(\exp(Tf)\iota_{X_{f}}\omega)ds.

    Let c=c+2c^{\prime}=c+2, then choose a smooth function χ:M\chi:M\mapsto\mathbb{R} such that χ|Uc=1\chi|_{U_{c^{\prime}}}=1 and χ|MUc=0\chi|_{M-U_{c}}=0 Then denoting ψ=ϕd(χϕ)\psi=\phi-d(\chi\phi^{\prime}), we have

    dψ=exp(Tf)ω and ψ|Uc=(1)jω.d\psi=\exp(Tf)\omega\mbox{ and }\psi|_{U_{c^{\prime}}}=(-1)^{j}\omega^{\prime}.

    Also, on Uc,U_{c^{\prime}},

    ιXfψ=ιXf0(Φ¯s)(exp(Tf)ιXfω)𝑑s=0(Φ¯s)(exp(Tf)ιXfιXfω)𝑑s=0\iota_{X_{f}}\psi=\iota_{X_{f}}\int_{0}^{\infty}(\bar{\Phi}^{s})^{*}(\exp(Tf)\iota_{X_{f}}\omega)ds=\int_{0}^{\infty}(\bar{\Phi}^{s})^{*}(\exp(Tf)\iota_{X_{f}}\iota_{X_{f}}\omega)ds=0 (7.5)

    Next, choose a smooth function η,\eta, s.t. η=0\eta=0 on KK, and η=1\eta=1 in f1((S+1,)(,I1))f^{-1}((S+1,\infty)\cup(-\infty,-I-1)). Then we set

    ψ(p)=ψ(p)d(ηf(p)+I/2+S/20(Φ¯s)ιXfψ𝑑s),pUc,\psi^{\prime}(p)=\psi(p)-d(\eta\int_{-f(p)+I/2+S/2}^{0}(\bar{\Phi}^{s})^{*}\iota_{X_{f}}\psi ds),p\in U_{c}^{\prime},
    ψ(p)=ψ(p)d(η0f(p)+I/2+S/2(Φ¯s)ιXfψ𝑑s),pUc,\psi^{\prime}(p)=\psi(p)-d(\eta\int^{-f(p)+I/2+S/2}_{0}(\bar{\Phi}^{s})^{*}\iota_{X_{f}}\psi ds),p\in U_{c},

    By (7.5), we have ψ=ψ\psi^{\prime}=\psi on UcU_{c^{\prime}}.

    Thus we can see that for pUcp\in U_{c^{\prime}}^{\prime},

    ιXfψ\displaystyle\iota_{X_{f}}\psi^{\prime} =ιXfψιXfd(ηf(p)+I/2+S/20(Φ¯s)ιXfψ𝑑s)\displaystyle=\iota_{X_{f}}\psi-\iota_{X_{f}}d(\eta\int_{-f(p)+I/2+S/2}^{0}(\bar{\Phi}^{s})^{*}\iota_{X_{f}}\psi ds)
    =ιXfψ(f(p)+I/2+S/20ιXfd(Φ¯s)ιXfψ𝑑s)\displaystyle=\iota_{X_{f}}\psi-(\int_{-f(p)+I/2+S/2}^{0}\iota_{X_{f}}d(\bar{\Phi}^{s})^{*}\iota_{X_{f}}\psi ds)
    ιXfdf(Φ¯f(p)+I/2+S/2)ιXfψ\displaystyle-\iota_{X_{f}}df(\bar{\Phi}^{-f(p)+I/2+S/2})^{*}\iota_{X_{f}}\psi
    =ιXfψ(f(p)+I/2+S/20dds(Φ¯s)ιXfψds(Φ¯f(p)+I/2+S/2)ιXfψ\displaystyle=\iota_{X_{f}}\psi-(\int_{-f(p)+I/2+S/2}^{0}\frac{d}{ds}(\bar{\Phi}^{s})^{*}\iota_{X_{f}}\psi ds-(\bar{\Phi}^{-f(p)+I/2+S/2})^{*}\iota_{X_{f}}\psi
    =ιXfψιXfηψ+(Φ¯f(p)+I/2+S/2)ιXfψ(Φ¯f(p)+I/2+S/2)ιXfψ\displaystyle=\iota_{X_{f}}\psi-\iota_{X_{f}}\eta\psi+(\bar{\Phi}^{-f(p)+I/2+S/2})^{*}\iota_{X_{f}}\psi-(\bar{\Phi}^{-f(p)+I/2+S/2})^{*}\iota_{X_{f}}\psi
    =0.\displaystyle=0.

    Similarly, we have ιXfψ=0\iota_{X_{f}}\psi^{\prime}=0 in Uc.U_{c^{\prime}}.

    As a consequence,

    dds(Φ¯s)ψ|s=t=ιXf(Φ¯t)dψ=ιXf(Φ¯t)exp(Tf)ω\frac{d}{ds}(\bar{\Phi}^{s})^{*}\psi^{\prime}|_{s=t}=\iota_{X_{f}}(\bar{\Phi}^{t})^{*}d\psi^{\prime}=\iota_{X_{f}}(\bar{\Phi}^{t})^{*}\exp(Tf)\omega

    on U.U.

    Therefore, on UcU_{c}^{\prime}, we have

    |(Φ¯t)ψ|=|0tιXf(Φ¯s)exp(Tf)ω𝑑s|0t|ιXf(Φ¯s)exp(Tf)ω|𝑑s0t|(Φ¯s)exp(Tf)ω|𝑑sCexp((1a)Tt) (Since on Uc(Φ¯t)ωCexp(aTt)).\displaystyle\begin{split}|(\bar{\Phi}^{t})^{*}\psi^{\prime}|&=|\int_{0}^{t}\iota_{X_{f}}(\bar{\Phi}^{s})^{*}\exp(Tf)\omega ds|\\ &\leq\int_{0}^{t}|\iota_{X_{f}}(\bar{\Phi}^{s})^{*}\exp(Tf)\omega|ds\\ &\leq\int_{0}^{t}|(\bar{\Phi}^{s})^{*}\exp(Tf)\omega|ds\\ &\leq C\exp((1-a)Tt)\mbox{ (Since on $U_{c}^{\prime}$, $(\bar{\Phi}^{t})^{*}\omega\leq C\exp(-aTt))$.}\end{split} (7.6)

    We claim that exp(Tf)ψL2Λ(M).\exp(-Tf)\psi^{\prime}\in L^{2}\Lambda^{*}(M). Since dψ=exp(Tf)ωd\psi^{\prime}=\exp(Tf)\omega implies that dTfexp(Tf)ψ=ωd_{Tf}\exp(-Tf)\psi^{\prime}=\omega, ω\omega is trivial in H(Ω(2),dTf).H^{*}(\Omega_{(2)}^{\bullet},d_{Tf}).

    Now we prove the claim:

    It suffices to prove that UcUc|exp(Tf)ψ|2𝑑vol<.\int_{U_{c^{\prime}}\cup U_{c}^{\prime}}|\exp(-Tf)\psi^{\prime}|^{2}dvol<\infty. Let Kc=f1{c},Kc=f1{c},K_{c^{\prime}}=f^{-1}\{-c^{\prime}\},K_{c}^{\prime}=f^{-1}\{c\}, and endow they with induced metrics. Define a diffeomorphism Ψc:Kc×(0,)Uc\Psi_{c^{\prime}}:K_{c^{\prime}}\times(0,\infty)\mapsto U_{c^{\prime}} as follow

    Ψc(p,t)=Φ¯t(p).\Psi_{c^{\prime}}(p,t)=\bar{\Phi}^{t}(p).

    Similarly, we can define a diffeomorphism Ψc:Kc×(0,)Uc.\Psi_{c}^{\prime}:K_{c}^{\prime}\times(0,\infty)\mapsto U_{c}^{\prime}.

    On UcU_{c^{\prime}}, |ψ|=|ω||\psi^{\prime}|=|\omega^{\prime}|, hence for pKcp\in K_{c^{\prime}}

    |(Φ¯t)(exp(Tf)ψ)(p)|=(Φ¯t)(exp(Tf)(p)0(Φ¯s)(exp(Tf)ιXfo(p))ds)=exp(Tc+Tt)0((exp(TcT(s+t))(Φ¯s+t)ιXfω(p))ds)0((exp(Ts)|(Φ¯s+t)ιXfω(p))|dsCexp(aTt)0((exp((a+1)Ts)ds (Since (Φ¯s+t)ωCexp(T(s+t)))Cexp(aTt).\displaystyle\begin{split}&|(\bar{\Phi}^{t})^{*}(\exp(-Tf)\psi^{\prime})(p)|=(\bar{\Phi}^{t})^{*}(\exp(-Tf)(p)\int_{0}^{\infty}(\bar{\Phi}^{s})^{*}(\exp(Tf)\iota_{X_{f}}\\ o(p))ds)\\ &=\exp(Tc^{\prime}+Tt)\int_{0}^{\infty}((\exp(-Tc^{\prime}-T(s+t))(\bar{\Phi}^{s+t})^{*}\iota_{X_{f}}\omega(p))ds)\\ &\leq\int_{0}^{\infty}((\exp(-Ts)|(\bar{\Phi}^{s+t})^{*}\iota_{X_{f}}\omega(p))|ds\\ &\leq C\exp(-aTt)\int_{0}^{\infty}((\exp(-(a+1)Ts)ds\mbox{ (Since $(\bar{\Phi}^{s+t})^{*}\omega\leq C\exp(-T(s+t)))$}\\ &\leq C\exp(-aTt).\end{split} (7.7)

    Then,

    |Uc|exp(Tf)ψ|2𝑑vol|=|0Kc(Φ¯t)(|exp(Tf)ω|2dvolKcdt)|\displaystyle|\int_{U_{c^{\prime}}}|\exp(-Tf)\psi^{\prime}|^{2}dvol|=|\int_{0}^{\infty}\int_{K_{c^{\prime}}}(\bar{\Phi}^{t})^{*}(|\exp(-Tf)\omega^{\prime}|^{2}dvol_{K_{c^{\prime}}}dt)|
    0Kcexp(2aTt)(Φ¯t)(dvolKcdt)\displaystyle\leq\int_{0}^{\infty}\int_{K_{c^{\prime}}}\exp(-2aTt)(\bar{\Phi}^{t})^{*}(dvol_{K_{c^{\prime}}}dt)
    C0Kcexp(2aTt)𝑑volKc𝑑t<, (Similar with Lemma 5.4)\displaystyle\leq C\int_{0}^{\infty}\int_{K_{c^{\prime}}}\exp(-2a^{\prime}Tt)dvol_{K_{c^{\prime}}}dt<\infty,\mbox{ (Similar with Lemma \ref{vol})}

    where aa^{\prime} is some positive number which is smaller than a.a.

    For pUc,p\in U_{c}^{\prime}, we have

    |(Φ¯t)(exp(Tf)ψ)(p)|=|(exp(TcTt)(Φ¯t)ψ)(p)|\displaystyle|(\bar{\Phi}^{t})^{*}(\exp(-Tf)\psi^{\prime})(p)|=|(\exp(-Tc-Tt)(\bar{\Phi}^{t})^{*}\psi^{\prime})(p)|
    Cexp(aTt) (By (7.6)).\displaystyle\leq C\exp(-aTt)\mbox{ (By (\ref{ucc}))}.

    For the same reason, we have Uc|exp(Tf)ψ|2𝑑vol<\int_{U_{c}^{\prime}}|\exp(-Tf)\psi^{\prime}|^{2}dvol<\infty.

  • \mathcal{L} is surjective: We claim that any cohomology class ξHj(M,Uc)\xi\in H^{j}(M,U_{c}) can be represented by a smooth closed jj-form ϕ\phi so that ϕ|Uc=0.\phi|_{U_{c}}=0. Also, it behaves on UU as follows: ιXfϕ=0\iota_{X_{f}}\phi=0 and (Φ¯t)ϕ(\bar{\Phi}^{t})^{*}\phi does not depend on tt for large tt. Then it follows that exp(Tf)ϕ\exp(-Tf)\phi belongs to L2Λ(M)L^{2}\Lambda^{*}(M) (follows from the similar argument as above). Let νKerTf,\nu\in Ker\Box_{Tf}, s.t. νexp(Tf)ϕ\nu-\exp(-Tf)\phi is exact. Then we can see that (ν)ξ\mathcal{L}(\nu)\in\xi hence \mathcal{L} is surjective. This is because, we can find ψIm(δTf)\psi\in Im(\delta_{Tf}), s.t. νexp(Tf)ϕ=dTfψ=(dTf+δTf)ψ.\nu-\exp(-Tf)\phi=d_{Tf}\psi=(d_{Tf}+\delta_{Tf})\psi. As a result, Tfψ=0\Box_{Tf}\psi=0 on UcU_{c}. Hence ψ\psi is of exponential decay in UcU_{c}, which implies that (ψ)\mathcal{L}(\psi) is well define. Now (ν)(exp(Tf)ϕ)=dC(ψ)\mathcal{L}(\nu)-\mathcal{L}(\exp(-Tf)\phi)=d_{C}\mathcal{L}(\psi) implies that (ν)ξ.\mathcal{L}(\nu)\in\xi.

    It suffices to prove the claim:

    It is clear that we may realize ξ\xi by a closed form ϕ\phi on MM with dϕ=0d\phi=0 and ϕ|Uc=0\phi|_{U_{c}}=0. Let η:M\eta:M\mapsto\mathbb{R} denote a smooth function which is identically 0 on KK and identically 11 on f1((S+1,)(,I1))f^{-1}((S+1,\infty)\cup(-\infty,-I-1)). The form ϕ(p)=ϕ(p)d(ηf(p)+I/2+S/20(Φ¯s)ιXfϕ𝑑s)\phi^{\prime}(p)=\phi(p)-d(\eta\int_{-f(p)+I/2+S/2}^{0}(\bar{\Phi}^{s})^{*}\iota_{X_{f}}\phi ds) is cohomologous to ϕ\phi and clearly satisfies what we claimed.

8 Appendix: Decomposition of L2L^{2} space

In this section, we investigate the decomposition (1.1). For this purpose we first have to understand the Friedrichs extension of ΔH,f\Delta_{H,f}. Moreover, all operators considered in this section are closurable.

8.1 Review on Friedrichs extension

Let A\operatorname{A} be a nonnegative, symmetric (unbounded) operator on Hilbert space \mathcal{H}, with Dom(A)=V,\mathrm{Dom}(\operatorname{A})=V, i.e.

(Aα,β)=(α,Aβ),α,βV;(Aα,α)0.(\operatorname{A}\alpha,\beta)_{\mathcal{H}}=(\alpha,\operatorname{A}\beta)_{\mathcal{H}},\ \forall\alpha,\beta\in V;\ \ (\operatorname{A}\alpha,\alpha)_{\mathcal{H}}\geq 0.

Define a norm V1\|\cdot\|_{V_{1}} on VV by

αV12=(α,α)+(α,Aα).\|\alpha\|_{V_{1}}^{2}=(\alpha,\alpha)_{\mathcal{H}}+(\alpha,\operatorname{A}\alpha)_{\mathcal{H}}.

Let V1V_{1} to be the completion of VV under V1.\|\cdot\|_{V_{1}}. Then for any β\beta\in\mathcal{H}, one can construct a bounded linear functional LβL_{\beta} on V1V_{1} as follows

Lβ(ϕ)=(ϕ,β),ϕV1.L_{\beta}(\phi)=(\phi,\beta)_{\mathcal{H}},\phi\in V_{1}. (8.1)

Since |(ϕ,β)|ϕβϕβV1,|(\phi,\beta)_{\mathcal{H}}|\leq\|\phi\|_{\mathcal{H}}\|\beta\|_{\mathcal{H}}\leq\|\phi\|_{\mathcal{H}}\|\beta\|_{V_{1}}, LβL_{\beta} is indeed bounded functional on V1.V_{1}. By Riesz representation, there exist γV1\gamma\in V_{1}, s.t. (ϕ,γ)V1=(ϕ,β).(\phi,\gamma)_{V_{1}}=(\phi,\beta)_{\mathcal{H}}.

Let B:V1,βγ,\operatorname{B}:\mathcal{H}\mapsto V_{1},\beta\mapsto\gamma, then B\operatorname{B} is bounded and injective. Take =B1I,\Box=\operatorname{B}^{-1}-\operatorname{I}, where I\operatorname{I} is the identity map, then \Box is the Friedrichs extension of A,\operatorname{A}, with Dom()=Im(B).\mathrm{Dom}(\Box)=\mathrm{Im}(\operatorname{B}).

Remark 8.1.

From the construction of Friedrichs extension \Box of A\operatorname{A}, we can see that Dom()=Im((I+)1).Dom(\Box)=\mathrm{Im}((\operatorname{I}+\Box)^{-1}).

Let T,S\operatorname{T},\operatorname{S} be two unbounded operators on Hilbert space \mathcal{H}, s.t.

  1. (i)
    V=Dom(T)=Dom(S),TVV.V=\mathrm{Dom}(\operatorname{T})=\mathrm{Dom}(\operatorname{S}),\operatorname{T}V\subset V.
  2. (ii)

    SS is a formal adjoint of T:T:

    (Tα,β)=(α,Sβ),(\operatorname{T}\alpha,\beta)_{\mathcal{H}}=(\alpha,\operatorname{S}\beta)_{\mathcal{H}},

Let W\|\cdot\|_{W} be the norm on VV given by

αW2=(α,α)+(Tα,Tα),αV,\|\alpha\|^{2}_{W}=(\alpha,\alpha)_{\mathcal{H}}+(\operatorname{T}\alpha,\operatorname{T}\alpha)_{\mathcal{H}},\alpha\in V,

and WW be the completion of VV under the norm W.\|\cdot\|_{W}. Then we can extend T\operatorname{T} to T¯min\bar{\operatorname{T}}_{min} with Dom(T¯min)=W.\mathrm{Dom}(\bar{\operatorname{T}}_{min})=W.

Let S¯max\bar{\operatorname{S}}_{max} be the closure of S\operatorname{S} with Dom(S¯max)={α:|(α,Tϕ)|Mαϕ,ϕV}\mathrm{Dom}(\bar{\operatorname{S}}_{max})=\{\alpha\in\mathcal{H}:|(\alpha,\operatorname{T}\phi)_{\mathcal{H}}|\leq M_{\alpha}\|\phi\|_{\mathcal{H}},\forall\phi\in V\}. Namely, for any αDom(S¯max)\alpha\in\mathrm{Dom}(\bar{\operatorname{S}}_{max}), since VV is dense in \mathcal{H}, by Riesz representation, there exists unique ν\nu\in\mathcal{H}, such that (ν,ϕ)H=(α,Tϕ).(\nu,\phi)_{H}=(\alpha,\operatorname{T}\phi). Now define S¯max(α)=ν.\bar{\operatorname{S}}_{max}(\alpha)=\nu.

Since TVV\operatorname{T}V\subset V, ST\operatorname{S}\operatorname{T} is symmetric and nonnegative with Dom(ST)=V.\mathrm{Dom}(ST)=V.

Proposition 8.2.

The Friedrichs extension Δ\Delta of ST\operatorname{S}\operatorname{T} is just S¯maxT¯min.\bar{\operatorname{S}}_{max}\bar{\operatorname{T}}_{min}.

Proof.

Since TVV,\operatorname{T}V\subset V, we see that V1V_{1} constructed in (8.1) is the same as W.W. Indeed, for any ϕ,ψV,\phi,\psi\in V, we have

(ψ,ϕ)+(Tψ,Tϕ)=(ψ,ϕ)+(STψ,ϕ)\displaystyle(\psi,\phi)_{\mathcal{H}}+(\operatorname{T}\psi,\operatorname{T}\phi)_{\mathcal{H}}=(\psi,\phi)_{\mathcal{H}}+(\operatorname{S}\operatorname{T}\psi,\phi)_{\mathcal{H}}

Hence, we have

Dom(Δ)={αW:α=(I+Δ)1f,f},\mathrm{Dom}(\Delta)=\{\alpha\in W:\alpha=(I+\Delta)^{-1}f,f\in\mathcal{H}\},
Dom(S¯maxT¯min)={αW:T¯minαDom(S¯max)}.\mathrm{Dom}(\bar{S}_{max}\bar{T}_{min})=\{\alpha\in W:\bar{\operatorname{T}}_{min}\alpha\in\mathrm{Dom}(\bar{\operatorname{S}}_{max})\}.

We now divide our discussion in two cases.

(a) We first prove that DomS¯maxT¯minDom(Δ),\mathrm{Dom}{\bar{\operatorname{S}}_{max}\bar{\operatorname{T}}_{min}}\subset\mathrm{Dom}(\Delta), and αDom(S¯max),\forall\alpha\in\mathrm{Dom}(\bar{\operatorname{S}}_{max}), S¯maxT¯minα=Δα.\bar{\operatorname{S}}_{max}\bar{\operatorname{T}}_{min}\alpha=\Delta\alpha.

For any αDomS¯maxT¯min,\alpha\in\mathrm{Dom}{\bar{\operatorname{S}}_{max}\bar{\operatorname{T}}_{min}}, let

β=α+S¯maxT¯minα.\beta=\alpha+\bar{\operatorname{S}}_{max}\bar{\operatorname{T}}_{min}\alpha. (8.2)

Then for any ϕW,\phi\in W, we have

(α,ϕ)W=limn(α,ϕn)W=limn(α,ϕn)+(T¯minα,Tϕn)=limn(α,ϕn)+(S¯maxT¯minα,ϕn) (Since ϕnV,T¯minαDom(S¯max) )=limn(α+S¯maxT¯minα,ϕn)=(α+S¯maxT¯minα,ϕ)=(β,ϕ),\displaystyle\begin{split}&(\alpha,\phi)_{W}=\lim_{n\rightarrow\infty}(\alpha,\phi_{n})_{W}\\ &=\lim_{n\rightarrow\infty}(\alpha,\phi_{n})_{\mathcal{H}}+(\bar{\operatorname{T}}_{min}\alpha,\operatorname{T}\phi_{n})_{\mathcal{H}}\\ &=\lim_{n\rightarrow\infty}(\alpha,\phi_{n})_{\mathcal{H}}+(\bar{\operatorname{S}}_{max}\bar{\operatorname{T}}_{min}\alpha,\phi_{n})_{\mathcal{H}}\mbox{ (Since $\phi_{n}\in V,\bar{\operatorname{T}}_{min}\alpha\in\mathrm{Dom}({\bar{\operatorname{S}}_{max}})$ )}\\ &=\lim_{n\rightarrow\infty}(\alpha+\bar{\operatorname{S}}_{max}\bar{\operatorname{T}}_{min}\alpha,\phi_{n})_{\mathcal{H}}=(\alpha+\bar{\operatorname{S}}_{max}\bar{\operatorname{T}}_{min}\alpha,\phi)_{\mathcal{H}}=(\beta,\phi)_{\mathcal{H}},\end{split} (8.3)

where ϕnV\phi_{n}\in V, and ϕnϕ\phi_{n}\rightarrow\phi w.r.t. W.\|\cdot\|_{W}. By the construction of Friedrichs extension and (LABEL:eq1), we deduce that α(I+Δ)1\alpha\in(I+\Delta)^{-1}\mathcal{H} and (I+Δ)α=β.(I+\Delta)\alpha=\beta. Comparing with (8.2), we obtain S¯maxT¯minα=Δα.\bar{\operatorname{S}}_{max}\bar{\operatorname{T}}_{min}\alpha=\Delta\alpha.

(b) We then show that Dom(Δ)Dom(S¯maxT¯min).\mathrm{Dom}(\Delta)\subset\mathrm{Dom}(\bar{\operatorname{S}}_{max}\bar{\operatorname{T}}_{min}).

Take any αDom(Δ)W,\alpha\in\mathrm{Dom}(\Delta)\subset W, we can find ff\in\mathcal{H}, s.t. α=(I+Δ)1f.\alpha=(I+\Delta)^{-1}f. We now just need to show that T¯minαDom(S¯max).\bar{\operatorname{T}}_{min}\alpha\in\mathrm{Dom}(\bar{\operatorname{S}}_{max}). For this, it suffices to prove that gV,\forall g\in V, |(T¯minα,Tg)|Mg|(\bar{\operatorname{T}}_{min}\alpha,\operatorname{T}g)_{\mathcal{H}}|\leq M\|g\|_{\mathcal{H}} for some M>0.M>0.

In fact, by standard functional calculus,

|(T¯minα,Tg)|=|(α,STg)| (via αnV,αnα w.r.t W)\displaystyle|(\bar{\operatorname{T}}_{min}\alpha,\operatorname{T}g)_{\mathcal{H}}|=|(\alpha,STg)_{\mathcal{H}}|\mbox{ (via $\alpha_{n}\in V,\alpha_{n}\rightarrow\alpha$ w.r.t $\|\|_{W}$)}
=|((I+Δ)1f,Δg)|\displaystyle=|((I+\Delta)^{-1}f,\Delta g)_{\mathcal{H}}|
=|(f,(I+Δ)1Δg)|\displaystyle=|(f,(I+\Delta)^{-1}\Delta g)_{\mathcal{H}}|
Mg\displaystyle\leq M\|g\|_{\mathcal{H}}

8.2 The Friedrichs extension of ΔH,f\Delta_{H,f}

By Proposition 8.2, we can see that the Friedichs extension f\Box_{f} of ΔH,f\Delta_{H,f} is (df+δf¯)max(df+δf¯)min(\overline{d_{f}+\delta_{f}})_{max}(\overline{d_{f}+\delta_{f}})_{min}.

If 0 is an eigenvalue of f\Box_{f} with finite multiplicity, we have the following decomposition

L2Λ(M)=kerfIm(df+δf¯)max.L^{2}\Lambda^{*}(M)=\ker\Box_{f}\oplus\mathrm{Im}(\overline{d_{f}+\delta_{f}})_{max}. (8.4)

Could we say more about decomposition (8.4)?

Proposition 8.3.

Let T,S\operatorname{T},\operatorname{S} be two unbounded operators on Hilbert space \mathcal{H}, such that

  1. 1.
    V=Dom(T)=Dom(S),TVV.V=\mathrm{Dom}(\operatorname{T})=\mathrm{Dom}(\operatorname{S}),\operatorname{T}V\subset V.
  2. 2.

    Im(T)\mathrm{Im}(T) is orthogonal to Im(S)\mathrm{Im}(S), and

    (Tα,β)=(α,Sβ).(\operatorname{T}\alpha,\beta)_{\mathcal{H}}=(\alpha,\operatorname{S}\beta)_{\mathcal{H}}.
  3. 3.

    T+ST+S is essential self-adjoint, i.e. (T+S)¯min=(T+S)¯max.\overline{(T+S)}_{min}=\overline{(T+S)}_{max}.

Then

T+S¯=T¯min|DomS¯minDomT¯min+S¯min|DomS¯minDomT¯min\overline{T+S}=\bar{T}_{min}|_{\mathrm{Dom}{\bar{S}_{min}\cap\mathrm{Dom}\bar{T}_{min}}}+\bar{S}_{min}|_{\mathrm{Dom}{\bar{S}_{min}\cap\mathrm{Dom}\bar{T}_{min}}}
=T¯max|DomS¯maxDomT¯max+S¯max|DomS¯maxDomT¯max=\bar{T}_{max}|_{\mathrm{Dom}{\bar{S}_{max}\cap\mathrm{Dom}\bar{T}_{max}}}+\bar{S}_{max}|_{\mathrm{Dom}{\bar{S}_{max}\cap\mathrm{Dom}\bar{T}_{max}}}
Proof.

Since Dom(T+S)¯min\mathrm{Dom}{\overline{(T+S)}}_{min} is the closure of VV under metric

(ϕ,ϕ)+((T+S)ϕ,(T+S)ϕ)=(ϕ,ϕ)+(Tϕ,Tϕ)+(Sϕ,Sϕ),ϕV,()(\phi,\phi)_{\mathcal{H}}+((T+S)\phi,(T+S)\phi)_{\mathcal{H}}=(\phi,\phi)_{\mathcal{H}}+(T\phi,T\phi)_{\mathcal{H}}+(S\phi,S\phi)_{\mathcal{H}},\phi\in V,(**)

Hence, Dom(T+S)¯minDomS¯minDomT¯min.\mathrm{Dom}{\overline{(T+S)}}_{min}\subset\mathrm{Dom}{\bar{S}_{min}\cap\mathrm{Dom}\bar{T}_{min}}. Also, for any ϕDom(T+S)min\phi\in\mathrm{Dom}{(T+S)}_{min}

(T+S)¯minϕ=limn(T+S)ϕn=limnTϕn+Sϕn=Tminϕ+Sminϕ,\overline{(T+S)}_{min}\phi=\lim_{n\rightarrow\infty}(T+S)\phi_{n}=\lim_{n\rightarrow\infty}T\phi_{n}+S\phi_{n}=\operatorname{T}_{min}\phi+\operatorname{S}_{min}\phi,

where ϕnVϕ\phi_{n}\in V\rightarrow\phi in the metric ()(**).

For each ϕDomS¯maxDomT¯max,\phi\in\mathrm{Dom}{\bar{S}_{max}\cap\mathrm{Dom}\bar{T}_{max}}, ψV\psi\in V,

(ϕ,(T+S)ψ)=(ϕ,Tψ)+(ϕ,Sψ)\displaystyle(\phi,(T+S)\psi)_{\mathcal{H}}=(\phi,T\psi)_{\mathcal{H}}+(\phi,S\psi)_{\mathcal{H}}
=(T¯maxϕ,ψ)+(S¯maxϕ,ψ)\displaystyle=(\bar{\operatorname{T}}_{max}\phi,\psi)_{\mathcal{H}}+(\bar{\operatorname{S}}_{max}\phi,\psi)_{\mathcal{H}}
Cψ.\displaystyle\leq C\|\psi\|_{\mathcal{H}}.

Therefore ϕDom((T+S)¯max),\phi\in\mathrm{Dom}(\overline{(T+S)}_{max}), and (T+S)¯maxϕ=T¯maxϕ+S¯maxϕ,\overline{(T+S)}_{max}\phi=\bar{\operatorname{T}}_{max}\phi+\bar{\operatorname{S}}_{max}\phi, which means that DomS¯minDomT¯minDom((T+S)¯max).\mathrm{Dom}{\bar{S}_{min}\cap\mathrm{Dom}\bar{T}_{min}}\subset\mathrm{Dom}(\overline{(T+S)}_{max}).

References

  • [1] S. Agmon. Lectures on exponential decay of solutions of second-order elliptic equations: bounds on eigenfunctions of N-body Schrödinger operations.(MN-29). Princeton University Press, 2014.
  • [2] J.-M. Bismut and W. Zhang. An extension of a theorem by Cheeger and Müller. Astérisque, 205, 1992.
  • [3] P. Chernoff. Essential self-adjointness of powers of generators of hyperbolic equations. Journal of Functional Analysis, 12(1):401–414.
  • [4] A. Dimca and M. Saito. On the cohomology of a general fiber of a polynomial map. Compositio Math., 85:299–309, 1993.
  • [5] H. Fan. Schro¨\mathrm{\ddot{o}}dinger equations, deformation theory and tttt^{*}-geometry. arXiv preprint arXiv:1107.1290, 2011.
  • [6] H. Fan and H. Fang. Torsion type invariants of singularities. arXiv preprint arXiv:1603.0653, 2016.
  • [7] M. Farber and E. Shustin. Witten deformation and polynomial differential forms. Geom. Dedicata, 80:125–155, 2000.
  • [8] M. Gromov and H. B. Lawson. Positive scalar curvature and the Dirac operator on complete Riemannian manifolds. Publication IHES, 58(1):83–196.
  • [9] H. Fan, T. Jarvis and Y. Ruan. The Witten equation, mirror symmetry and quantum singularity theory. Ann. Math., 178:1–106, 2013.
  • [10] Q. Han and F. Lin. Elliptic partial differential equations, volume 1. American Mathematical Soc., 2011.
  • [11] B. Helffer and J. Sjöstrand. Puits multiples en mecanique semi-classique iv etude du complexe de Witten. Communications in Partial Differential Equations, 10(3):245–340, 1985.
  • [12] K. Hori, R. Thomas, S. Katz, C. Vafa, R. Pandharipande, A. Klemm, R. Vakil and E. Zaslow. Mirror symmetry, volume 1. American Mathematical Soc., 2003.
  • [13] J. Cheeger, M. Gromov and M. Taylor. Finite propogation speed, kernel estimate for functions of the Laplace operator, and the geometry of complete Riemannian manifolds. J. Diff. Geom., 17:15–53, 1982.
  • [14] J. P. Demailly. Champs magnétiques et inégalités de Morse pour la d-cohomologie,. C. R. Acad. Sci., and Ann. Inst. Fourier, 301, 35:119–122, 185–229, 1985.
  • [15] F. Laudenbach. On the Thom-Smale complex. Astérisque, 205:219–233, 1992.
  • [16] F. Laudenbach. A Morse complex on manifolds with boundary. Geometriae Dedicata, 153(1):47–57, 2011.
  • [17] W. Lu. A Thom-Smale-Witten theorem on manifolds with boundary. Mathematical Research Letters, 24(1):119–151, 2017.
  • [18] J. Milnor. Lectures on the hh-cobordism theorem, volume 2258. Princeton University Press, 2015.
  • [19] E. Witten. Supersymmetry and Morse theory. J. Diff. Geom, 17(4):661–692, 1982.
  • [20] E. Witten. Phases of N=2N=2 theories in two dimensions. Nuclear Physcis B, 403:159–222, 1993.
  • [21] W. Zhang. Lectures on Chern-Weil theory and Witten deformations, volume 4. World Scientific, 2001.