Witten deformation for noncompact manifolds with bounded geometry
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 near infinity. We prove that the cohomology of the Witten deformation acting on the complex of smooth forms is isomorphic to the cohomology of Thom-Smale complex of as well as the relative cohomology of a certain pair for sufficiently large . 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 where is the usual exterior derivative on forms, and is a Morse function. Setting
Witten observed that when 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 . As a result, Witten deformation builds a direct bridge between the Betti numbers and the Morse indices of the critical points of .
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 , 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 (the superpotential) on a noncompact Kahler manifold (), then one considers the Witten-deformation of operator:
as its cohomology describes the quantum ground states of the Landau-Ginzburg model . If is also a Morse function with critical points, then complex Morse theoretic consideration leads to the expectation that
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 on a noncompact manifold and the deformed de Rham complex with respect to . 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 .
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 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 -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 be a noncompact connected complete Riemannian manifold with metric . is said to have bounded geometry, if the following conditions hold:
-
1.
the injectivity radius of is positive.
-
2.
, where is the -th covariant derivative of the curvature tensor and is a constant only depending on .
On such a manifold, the Sobolev constant is uniformly bounded, see e.g. [13]. Now let be a smooth function. In [5], the notion of strong tameness for the triple is introduced.
Definition 1.1.
The triple is said to be strongly tame, if has bounded geometry and
and
where are the gradient and Hessian of respectively.
Remark 1.2.
Fix , and let be the distance function induced by . Here simply means that
In this paper we only need the following weaker condition.
Definition 1.3.
The triple is said to be well tame, if has bounded geometry and
and
As usual, the metric induced a canonical metric (still denote it by ) on , which then defines an inner product on :
Let be the completion of with respect to , and for simplicity, we denote
For any , let be the so-called Witten deformation of de Rham operator . It is an unbounded operator on with domain . Also, has a formal adjoint operator , with such that
Set and we denote the Friedrichs extension of by . As we will see (Theorem 2.1), if is well tame, then is essentially self-adjoint (and hence is the unique self-adjoint extension). In Section 8, we will prove the Hodge-Kodaira decomposition when is well tame and large enough,
(1.1) |
where and are the graph extensions of and respectively.
Setting we have a chain complex (of unbounded operators)
Let denote the cohomology of this complex. In Section 8, we will show that , provided is well tame and is large enough.
Proposition 1.4.
The Hodge Laplacian has the following local expression:
(1.2) |
Here is a local frame on and is the dual frame on
1.3 Main results
In this subsection, we assume that has bounded geometry, is a Morse function with finite many critical points. Clearly this will be the case if is well tame and 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 be well tame, and be an eigenform of whose eigenvalue is uniformly bounded in . Then
for any (provided is sufficiently large and is a constant depending on the dimension , the function , the curvature bound, and ; for the precise choice of see the end of Section 3). Here the definition of the Agmon distance 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 . If is a critical point of , denote the Morse index of at Let be the number of critical points of with Morse index Then the strong Morse inequalities hold.
Theorem 1.2.
If is well tame, then we have the following strong Morse inequality
provided is large enough. And the equality holds for .
In general, may be very sensitive to . However we have the following result regarding the indepedence of in . Assume that the Morse function satisfies the Smale transversality condition. Let be the Thom-Smale complex given by . It is important to note that in general, since is noncompact, it could happen that . Also let be big enough, and be the relative de Rham complex.
Theorem 1.3.
If is well tame, then , and therefore the cohomology is well defined. Moreover, there exists (When is strongly tame, ), such that is isomorphic to for all . In addition, , hence is isomorphic to the relative de Rham cohomology .
Corollary 1.4.
If is well tame, then is independent of when is big enough. When is strongly tame, is independent of
As an another application of Theorem 1.3, we study the Morse cohomology for compact manifolds with boundary.
Let be a compact, oriented manifold of dimension with boundary . Let be the connected components of . We fix a collar neighborhood , and let be the standard coordinate on the factor.
Definition 1.5.
A smooth function on is called a transversal Morse function if it satisfies the following conditions:
-
1.
is a Morse function on the manifold ;
-
2.
is a Morse function on the manifold .
-
3.
For any point on the collar neighborhood, .
For a transversal Morse function on , since is continuous on any connected components of , so we can call to be positive (with respect to ) if , and negative if
Let be the union of all positive boundaries, be the union of all negative boundaries. Suppose we have a partition of positive boundaries , and a partition of negative boundaries Now We denote
-
•
by the set of internal critical points with Morse index , ;
-
•
by the set of critical points on positive boundary with Morse index , ;
-
•
by the set of critical points on negative boundary with Morse index , .
Let
Theorem 1.5.
There is a differential making a chain complex. Moreover, is isomorphic to the relative de Rham cohomology
In particular, for is isomorphic to the relative de Rham cohomology
Corollary 1.6.
Set then we have the following Morse inequalities:
1.4 Notations and Organization
In this article, we will generally use to denote differential forms, Morse function, functions, eigenforms, critical points of , general points, and 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 . Section 5 concerns the Thom-Smale theory in our setting. More specifically, we define the Thom-Smale complex . Then we define a morphism between the Witten instanton complex and the Thom-Smale complex, We prove that is well defined by using the Agamon estimate, deferring the proof that 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., . 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
Theorem 2.1.
On a complete Riemannian manifold, if
then is essentially self-adjoint.
2.2 On the spectrum of
From now on we will assume that is well tame. Then it follows that there exists a compact subset , which can be taken to be a compact submanifold with boundaries that contains the closure of a ball of sufficiently large radius of (we will make a more specific choice of later in section 5), , such that
(2.1) |
Let . First, we establish the following basic lemma.
Lemma 2.1.
Fix any , there exists so that whenever ,
(2.2) | |||||
Here is a constant depending only on the sectional curvature bounds of .
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
where . Thus, for any , there exists
(2.3) |
such that whenever we have
∎
Remark 2.2.
When is strongly tame, we can take but may depend on
Let be a new metric on (with discrete conical singularities).Fix let be the distance between and induced by Then we have a.e., where the gradient is induced by .
Theorem 2.2.
Let be the set of spectrum of Then for any positive number there is
(2.4) |
such that when , consists of a finite number of eigenvalues of finite multiplicity.
Proof.
Let be the integral of the spectral measure of on . It suffices to prove that is finite dimensional. For any we have
(2.5) |
Since on , , there is such that when ,
(2.6) |
Now define , by By (2.6), it’s easy to see that is injective, and Since is compact, must be finite. ∎
Remark 2.3.
Once again, if is strongly tame, we can take
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 is well tame. Then we have the Kodaira decomposition
Furthermore, the Hodge Theorem holds:
3 Exponential decay of eigenfunction
In this section, we assume that is well tame, and , where is described in Lemma 2.1. If is strongly tame, then we can just take
Recall that , the Agmon metric on Let be the compact set as in last section. In this and later sections we define the Agmon distance be the distance between and induced by Then we have a.e. , where the gradient is induced by .
For simplicity, denote by . We need the following two technical lemmas, whose proofs are postponed to Section 7.
Lemma 3.1.
Assume , and outside the compact subset in the weak sense. That is
Then there exists another compact subset of such that
(3.1) |
for .
Corollary 3.1.
If for some and , then
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 , and for some and , then
where the constant .
Proof.
Following the proof of Lemma 3.1 given in Section 7.1, put Then we deduce
for as above, and . We split the second integral on the right hand side into two; the one over will be absorbed into the first term. The second term is (we omit the volume form here)
Combining the above one arrives at
Thus, for ,
where . If is also bigger than , then outside . Hence
and consequently
for . ∎
Remark 3.2.
It may seem that and depend on as . However, notice that when becomes bigger, gets smaller. Hence we can choose , , s.t. they are independent of
Lemma 3.3 (De Giorgi-Nash-Moser Estimates).
For , let be the geodesic ball around with radius (in the metric ). Let , and on in the weak sense for some constant Then there exists constant depending only on the dimension , the Sobolev constant, and , such that
With these preparation we are now ready to prove our first main estimate for the eigenforms of .
Proof of Theorem 1.1.
Consider an eigenform of . That is , where the eigenvalue satisfies for some constant . Then letting , by a straightforward computation using the Bochner’s formula (for forms) and the Kato’s inequality, we have
where is the upper bound of curvature tensor. Hence by Corollary 3.2, we have, for ,
where the constant .
Recall that the compact set is chosen so that (2.1) is satisfied. Hence by Proposition 1.4, the conditions of Lemma 3.3 are satisfied for on . Also, the Agmon distance is the distance between and induced by and Suppose . Denote by the -geodesic ball around with radius . Set , and Then one can easily verify that
Remark 3.4.
The proof above gives the inequality for for some constant independent of , which is what we needed for later applications. For , using the same reasoning as in Remark 3.2, there exist constant , which is independent of , such that
(3.3) |
for all . Therefore via Moser iteration as in Lemma 3.3 and similar arguments as above, one can show that
4 Morse inequalities
In this and the next section, we assume that is a Morse function on , and . In fact, we assume that in a neighborhood of critical points of , we have coordinate system such that
(4.1) |
This is a generic condition. Without loss of generality we assume that is an Euclidean open ball around with radius Also, these open sets are disjoint.
Let be the space spanned by the eigenforms of with eigenvalue lying in By Theorem 2.2, is finite dimensional. Recall that denotes the number of critical points of with Morse index We have the following Proposition:
Proposition 4.1.
There exists big enough, so that whenever the number of eigenvalues (counted with multiplicity) in of equals . I.e. .
Remark 4.2.
See the definition of in (5.2). Also recall that, if is strongly tame,
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 , such that for any smooth form with and , one has
Here denotes the support of
Proof.
Since is well tame, there exist s.t. also on . Then our proposition follows from the same argument in Proposition 4.7 of [21]. ∎
5 Thom-Smale theory
In this section, we assume that is a compact subset of , is small enough (to be determined later), is big enough, such that outside , we have
(5.1) |
provided .
Moreover, we make a more judicious choice of Fix any . Set
where is the distance function induced by (note that the Agmon distance is the same distance function but between and a compact subset ). The second equality follows from the claim in the proof of Lemma 5.5. We choose so that that
where denotes the interior of
Remark 5.1.
We can take if is strongly tame.
Now we set
(5.2) |
(Note that the definition of does not involve ; Cf. (2.3) and (2.4) for the description of and .) Before defining the Thom-Smale complex, there is still a subtle issue for noncompact cases. That is, the gradient vector field 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 such that
Then we have
Lemma 5.2.
is a complete vector field.
Proof.
Let be the flow generated by . We show that for any , there exists a universal , s.t. is well defined on . Hence, is complete.
Let . It suffices to show that for any is well defined on , since is compact.
But on , , and is complete, hence , is a geodesic (See Lemma 5.5) inside for .
∎
Let be a critical point of the Morse function , and be the stable and unstable manifold of with respect to flow defined in Lemma 5.2 (See Chapter 6 in [21]). We will further assume that satisfies the Smale transversality condition, namely and intersect transversally. Then the Thom-Smale complex is defined by
and
To define the boundary operator, let and be critical points of , with .
For set
Here the integer is the signed counts of the flow lines in .
Remark 5.3.
With our nice choice of and , it is easy to see that for any , . Moreover, for any , the curve and Moreover, since , we can actually choose , such that it is independent of Thus, just like the compact case, by the transversality, is well defined.
We will prove in Section 7.3 that under our tameness condition, Thus, is a complex.
Recall that the Witten instanton complex is the finite dimensional space generated by the eigenforms of with eigenvalue lying in . By the discussion in the previous subsection, the cohomology of the Witten instanton comple is .
To prove Theorem 1.3, we now consider the following chain map Here denote the dual chain complex. Let be the dual basis of Then
However there is a technical issue here we need to address. When is compact, the integral is clearly well defined, but here may be noncompact. We will be content here only with the wel-definedness of the map and leave the proof that is indeed a chain map to Section 7.3.
Let small enough, be the -dimensional ball in with center and radius with respect to metric As before, let be the flow generated by . Then
Therefore, for any
The well definedness of is now reduced to the following two technical lemmas, as well as Theorem 1.1 and the well tameness of .
Lemma 5.4.
Fix any we have
Hence,
Here is a constant independent of
Proof.
Let be a unit tangent vector of at Extend to a local unit vector field near via parallel transport along radial geodesics. Denote
(5.3) |
Noting that from (5.1)
we have
By a classical result in ODE, we have
Now our lemma follows from teh following, Lemma 5.5. ∎
Lemma 5.5.
Suppose is big enough, . Then there exists a constant s.t.
Proof.
For any , we claim that is one of the shortest smooth curve on connecting to
Granted, since for all , the claim gives
where is the distance induced by , is the diameter of with respect to metric
We now prove the claim (See [11] Lemma A 2.2 for another proof):
First, Let’s show that is a geodesic for any :
Let be a local orthomormal frame on with . In order to prove it suffices to prove
Let be the Levi-Civita connection induced by then
We now prove that is the shortest geodesic connecting and in , for all :
Assume that is another normal geodesic connecting and induced by Then Set then we have and Hence by a comparison theorem in ODE, we must have Assume that then we can see also we have Since is decreasing, we must have
By now, we can see that is one of shortest geodesic connecting and ∎
We now note the following lemma which plays an important role in estimating the eigenforms previously.
Lemma 5.6.
Suppose Then for any , there exists , such that
Proof.
Let be a normal minimal -geodesic connecting and . Then we have outside As is compact and is bounded on , we will assume without loss of generality that lies outside of .
Let then
Hence ∎
Here we are gives a direct proof of the isomorphism of and under the assumption that is proper.
Theorem 5.1.
Assume that is proper. Set and fix . Then for and are quasi-isomorphic.
Proof.
Set , . Let be the flow in generated by on , and on .
Define a map
By Theorem 1.1, and similar argument in Lemma 5.5, we can see that Hence, is well defined.
is a chain map, since
Hence induces a homomorphism (still denote it by ) between and The proof of the fact that is a bijection is tedious, which will be given in Subsection 7.6. ∎
6 An application of Theorem 1.3
Let be an oriented, compact Riemannian manifold with boundary and near the boundary. Let where . Then we can extend the metric to , s.t. near the infinity, the metric on is of product type, i.e. . It’s easy to see that has bounded geometry. We have the following technical lemma
Lemma 6.1.
Given a transversal Morse function , a partition of boundaries We are able to extend to a function on s.t.
-
1.
as ;
-
2.
on
-
3.
has critical points ;
-
4.
is well tame.
Proof.
Use notation to denote Since is a transversal Morse function, there exists , s.t. on
Hence, by considering the Taylor expansion of with respect to there is a smooth function on s.t.
-
1.
;
-
2.
near
-
3.
Assume that is close enough to , s.t. on
Let be a smooth function on s.t.
-
1.
on , and on , on ;
-
2.
;
Then , let
, let
It’s easy to verify that satisfy our conditions. ∎
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 Let ( large enough) be a smooth bump function such that
and
Set , and
Clearly a.e. and
Now set . Then by assumption, we have
Noting that we have
(7.1) |
Since (we now omit the volume form in what follows)
and
by (7.1), we have
(7.2) |
Letting by the monotone convergence theorem and the fact that , we have
Now let 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 is big enough and is small enough, s.t. Then on , we have
(7.3) |
weakly.
Set
Let be bump functions s.t.
and
Set , and Notice that and in Hence, by (7.3), we have
Hence, we have
By Sobolev inequality,
That is
Let we have
Consider By the same arguments as above, we have
As a consequence,
∎
We state two Lemmas that would be needed shortly,
Lemma 7.1.
Suppose , s.t. in weak sense (Here we assume ). For small enough, , let be the geodesic ball around with radius induced by . Then there exist , s.t.
where is a constant that depends only on dimension
Proof.
By the same argument as the proof of Theorem 1.1, we have
Lemma 7.2.
Let be well tame, , s.t.
for some If is a weak solution of then
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 be a compact Riemannian manifold, be a Morse function. Assume that satisfies Thom-Smale transversality condition. Then, for any critical point with Morse index , any , one has the following so called Stokes Formula
For our noncompact case with tame conditions and Thom-Smale transversality, we have similarly
Proposition 7.4.
For any critical point with Morse index , any , one has the following so called Stokes Formula
Before giving the proof of this proposition, we first draw a couple of consequneces.
Corollary 7.1.
Let be the map constructed in Section 5, then .
Proof.
Corollary 7.2.
Let , one has
Proof.
Now our Corollary follows from Proposition 7.4 trivially. ∎
Hence, the map introduced in Section 5 is a chain map.
The proof of Proposition 7.4 follows from the following observations:
Observation 7.5.
Let be compact manifold with boundary. Moreover, assume that near the boundary , the manifold is of product type Suppose that is a Morse function on . Then there exist a transversal Morse function on , s.t. Here is the standard coordinate on factor.
The proof is essentially the same with Theorem 2.5 in [18].
Observation 7.6.
Let be the ball with radius introduced in Section 5. Then for any , Moreover, if lies in an unstable manifold, then the curve Here denotes the interior of is the flow generated by
Proof.
It follows easily from the fact that for any , , and is decreasing along the flow ∎
Now we are ready to prove Proposition 7.4
Proof.
For any , let
we can find a compact submanifold with boundary, s.t. Here denotes the support of denote the interior of
Now consider the double of , By Observation 7.5, we can find a Morse function on , s.t. We may as well assume that satisfy Thom-Smale transversality condition. Then for any with let be the signed count of the number of flow lines in
7.3.1 An counterexample
On closing this subsection, let’s give a counterexample that when we drop the condition that has a positive lower bound near infinity, may fail.
Consider the following heart shaped topological sphere with obvious height function . Then we have four critical points as indicated below. Let be a flow line connecting and , and remove a point on . Making a conformal change of metric near point , s.t. is complete under this new metric. Now we can see that , as Since the flow line is invariant under the conformal change of metric, is still a (broken) flow line. However, in this case, , which is nonzero.
In our previous arguments, the fact that has a positive lower bounded near the infinity play a crucial role. It forces the value of Morse function goes to infinity along a flow line if that flow line flows to infinity (See also Remark 5.3 and Observation 7.6).
Remark 7.7.
We would like to thank Shu Shen for providing this interesting example.
7.4 Isomorphism of and
For simplicity, we assume that is self-indexed Morse function, i.e., if is a critical point of with Morse index , we require . Moreover,
Let , .
Recall that we assume in a neighborhood of critical points of , we have coordinate system such that
Moreover is an Euclidean open ball around with radius Also, these open balls are disjoint.
We have the following observation:
Lemma 7.8.
can be written as disjoint union of and , where is some open subset diffeomorphic to , is an Euclidean ball around with radius .
is diffeomorphic to
Proof.
Let , be the flow generated by . Then we have
This is because:
-
•
If , then . Hence .
-
•
If , and if for some with Morse index . Then , which implies . But this is impossible since .
Similarly, we can prove that is diffeomorphic to ∎
Let be complex of relative singular chains. Then we have
7.5 Isomorphism of and
We will first show that the chain map defined in Section 5 is in fact an isomorphism when is sufficiently large. Hence induced an isomorphism between and in that case.
More precisely we will follow the arguments in Chapter 6 of [21], with necessary modification, to show that there exists , such that is an isomorphism whenever (We point out that the explicit description of is more involved than .) In fact, the only difference is that we need a more refined estimate in Theorem 6.17 of [21], that is:
(7.4) |
where is the orthogonal projection from to , and C, c are positive constants.
Here is defined as follows. Notice that in Section 4, we require that in a neighborhoof of , the metric and Morse function is of the form (4.1). Hence, let be a bump function whose support is contained in , and in a neighborhood of . Now let
To obtain the estimate (7.4), pick a bump function with compact support, such that on . Then by our Agmon estimate, we have
The estimate of
now follows from exactly the same argument in the proof of Theorem 6.17 of [21].
Now it remains to prove that when , and are still isomorphic.
We only present the proof for the case when is strongly tame, the case of well tame being exactly the same except notationally. In this case, . The idea is to show that if then for any , and are isomorphic. Hence is independent of , which finishes the proof of isomorphism of and .
For simplicity, we prove that and are isomorphic, the general case being similar.
Define ; . Similarly ;
Clearly these are chain maps once we check that and are well defined. To this end, let’s verify that where is the fix point in defining Indeed, let be a normal minimal geodesic connecting and , in the metric . Then
Now the bound of (resp. ) follows by Theorem 1.1 and the standard volume comparison, Hence induces a homomorphism (still denote it by ) from to .
Our next step is to show that is injective. Suppose we have , s.t. is exact, which means that we can find s.t
Thus
By Lemma 7.2, Consequently, , and is exact.
As a result, is injective. Similarly, is also injective. Therefore, and are isomorphic.
7.6 is bijective
-
•
is injective:
Let . Assume that is exact, then there exists s.t.
where
Let , then choose a smooth function such that and Then denoting , we have
Also, on
(7.5) Thus we can see that for ,
Similarly, we have in
As a consequence,
on
Therefore, on , we have
(7.6) We claim that Since implies that , is trivial in
Now we prove the claim:
It suffices to prove that Let and endow they with induced metrics. Define a diffeomorphism as follow
Similarly, we can define a diffeomorphism
On , , hence for
(7.7) Then,
where is some positive number which is smaller than
For we have
For the same reason, we have .
-
•
is surjective: We claim that any cohomology class can be represented by a smooth closed -form so that Also, it behaves on as follows: and does not depend on for large . Then it follows that belongs to (follows from the similar argument as above). Let s.t. is exact. Then we can see that hence is surjective. This is because, we can find , s.t. As a result, on . Hence is of exponential decay in , which implies that is well define. Now implies that
It suffices to prove the claim:
It is clear that we may realize by a closed form on with and . Let denote a smooth function which is identically on and identically on . The form is cohomologous to and clearly satisfies what we claimed.
8 Appendix: Decomposition of space
In this section, we investigate the decomposition (1.1). For this purpose we first have to understand the Friedrichs extension of . Moreover, all operators considered in this section are closurable.
8.1 Review on Friedrichs extension
Let be a nonnegative, symmetric (unbounded) operator on Hilbert space , with i.e.
Define a norm on by
Let to be the completion of under Then for any , one can construct a bounded linear functional on as follows
(8.1) |
Since is indeed bounded functional on By Riesz representation, there exist , s.t.
Let then is bounded and injective. Take where is the identity map, then is the Friedrichs extension of with
Remark 8.1.
From the construction of Friedrichs extension of , we can see that
Let be two unbounded operators on Hilbert space , s.t.
-
(i)
-
(ii)
is a formal adjoint of
Let be the norm on given by
and be the completion of under the norm Then we can extend to with
Let be the closure of with . Namely, for any , since is dense in , by Riesz representation, there exists unique , such that Now define
Since , is symmetric and nonnegative with
Proposition 8.2.
The Friedrichs extension of is just
Proof.
Since we see that constructed in (8.1) is the same as Indeed, for any we have
Hence, we have
We now divide our discussion in two cases.
(a) We first prove that and
For any let
(8.2) |
Then for any we have
(8.3) |
where , and w.r.t. By the construction of Friedrichs extension and (LABEL:eq1), we deduce that and Comparing with (8.2), we obtain
(b) We then show that
Take any we can find , s.t. We now just need to show that For this, it suffices to prove that for some
In fact, by standard functional calculus,
∎
8.2 The Friedrichs extension of
By Proposition 8.2, we can see that the Friedichs extension of is .
If is an eigenvalue of with finite multiplicity, we have the following decomposition
(8.4) |
Could we say more about decomposition (8.4)?
Proposition 8.3.
Let be two unbounded operators on Hilbert space , such that
-
1.
-
2.
is orthogonal to , and
-
3.
is essential self-adjoint, i.e.
Then
Proof.
Since is the closure of under metric
Hence, Also, for any
where in the metric .
For each ,
Therefore and which means that ∎
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. Schrdinger equations, deformation theory and -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 -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 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.