28
G eometry & T opology G G G G G G G G GG G G G G G T T T T T T T T T T T T T T T Volume 3 (1999) 369ā€“396 Published: 26 October 1999 Circle-valued Morse theory and Reidemeister torsion Michael Hutchings Yi-Jen Lee Dept of Math, Stanford University, Stanford, CA 94305, USA Dept of Math, Princeton University, Princeton, NJ 08544, USA Email: [email protected] and [email protected] Abstract Let X be a closed manifold with Ļ‡(X) = 0, and let f : X ā†’ S 1 be a circle- valued Morse function. We deļ¬ne an invariant I which counts closed orbits of the gradient of f , together with ļ¬‚ow lines between the critical points. We show that our invariant equals a form of topological Reidemeister torsion deļ¬ned by Turaev [28]. We proved a similar result in our previous paper [7], but the present paper re- ļ¬nes this by separating closed orbits and ļ¬‚ow lines according to their homology classes. (Previously we only considered their intersection numbers with a ļ¬xed level set.) The proof here is independent of the proof in [7], and also simpler. Aside from its Morse-theoretic interest, this work is motivated by the fact that when X is three-dimensional and b 1 (X) > 0, the invariant I equals a count- ing invariant I 3 (X) which was conjectured in [7] to equal the Seibergā€“Witten invariant of X . Our result, together with this conjecture, implies that the Seibergā€“Witten invariant equals the Turaev torsion. This was conjectured by Turaev [28] and reļ¬nes the theorem of Meng and Taubes [14]. AMS Classiļ¬cation numbers Primary: 57R70 Secondary: 53C07, 57R19, 58F09 Keywords: Morseā€“Novikov complex, Reidemeister torsion, Seibergā€“Witten invariants Proposed: Ralph Cohen Received: 28 June 1999 Seconded: Robion Kirby, Steve Ferry Accepted: 21 October 1999 ISSN 1364-0380 Copyright Geometry and Topology 369

Circle-valued Morse theory and Reidemeister torsion

  • Upload
    others

  • View
    7

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Circle-valued Morse theory and Reidemeister torsion

Geometry & Topology GGGGGGGG G GG

GGGG

T TTTTTT

TTTTTTTT

Volume 3 (1999) 369ā€“396Published: 26 October 1999

Circle-valued Morse theory and Reidemeister torsion

Michael Hutchings

Yi-Jen Lee

Dept of Math, Stanford University, Stanford, CA 94305, USA

Dept of Math, Princeton University, Princeton, NJ 08544, USA

Email: [email protected] and [email protected]

Abstract

Let X be a closed manifold with Ļ‡(X) = 0, and let f : X ā†’ S1 be a circle-valued Morse function. We define an invariant I which counts closed orbits ofthe gradient of f , together with flow lines between the critical points. We showthat our invariant equals a form of topological Reidemeister torsion defined byTuraev [28].

We proved a similar result in our previous paper [7], but the present paper re-fines this by separating closed orbits and flow lines according to their homologyclasses. (Previously we only considered their intersection numbers with a fixedlevel set.) The proof here is independent of the proof in [7], and also simpler.

Aside from its Morse-theoretic interest, this work is motivated by the fact thatwhen X is three-dimensional and b1(X) > 0, the invariant I equals a count-ing invariant I3(X) which was conjectured in [7] to equal the Seibergā€“Witteninvariant of X . Our result, together with this conjecture, implies that theSeibergā€“Witten invariant equals the Turaev torsion. This was conjectured byTuraev [28] and refines the theorem of Meng and Taubes [14].

AMS Classification numbers Primary: 57R70

Secondary: 53C07, 57R19, 58F09

Keywords: Morseā€“Novikov complex, Reidemeister torsion, Seibergā€“Witteninvariants

Proposed: Ralph Cohen Received: 28 June 1999Seconded: Robion Kirby, Steve Ferry Accepted: 21 October 1999

ISSN 1364-0380

Copyright Geometry and Topology

369

Page 2: Circle-valued Morse theory and Reidemeister torsion

1 Introduction

Given a flow on a manifold, it is natural to ask how many closed orbits thereare. It turns out that for some well-behaved flows, the numbers of closed orbitsin different homology classes are related to the Reidemeister torsion of theunderlying manifold. For example, Fried [2] defined a ā€œtwisted Lefschetz zetafunctionā€ counting closed orbits of certain nonsingular hyperbolic flows andshowed that it equals a version of topological Reidemeister torsion, which isindependent of the flow.

In this paper, we are interested in the gradient flow of a circle-valued Morsefunction. For singular flows such as this one, the zeta function is no longerinvariant under deformation of the flow. It turns out that this lack of invariancecan be fixed by considering the Novikov complex, which counts gradient flowlines between critical points. We will show that one can obtain a topologicalinvariant by multiplying the zeta function by the Reidemeister torsion of theNovikov complex. We call the resulting invariant I .

In our previous work [7], we defined a weaker version of I and showed that itequals a form of topological Reidemeister torsion. Later we received a preprintfrom Turaev [28] defining a refined version of the latter invariant, which we callā€œTuraev torsionā€ here. Along similar lines we can refine the Morse theoreticinvariant in [7] to obtain the invariant I . The main result of this paper assertsthat I equals Turaev torsion.

Our previous methods are not quite sufficient to prove this refinement, so herewe introduce a different and simpler approach. This paper is independent of [7],except that the latter paper defines certain compactifications in Morse theorywhich we use here, and also provides more background and context.

We now proceed to define our invariant I more precisely and state our maintheorem. We then describe the application to three-dimensional Seibergā€“Wittentheory. In section 2 we give some background definitions, and in section 3 weprove the main theorem. In section 4 we give more details on the relation toSeibergā€“Witten theory.

1.1 Statement of results

The basic setup for this paper is as follows. Let X be a closed connectedoriented nā€“dimensional manifold. We assume throughout that Ļ‡(X) = 0, sothat we can define Reidemeister torsion. Also, our result is most interestingwhen b1(X) > 0.

Michael Hutchings and Yi-Jen Lee

Geometry and Topology, Volume 3 (1999)

370

Page 3: Circle-valued Morse theory and Reidemeister torsion

Let f : X ā†’ S1 . In order to consider the gradient flow of f , we endow X witha Riemannian metric. We make the following assumptions:

Assumption 1.1 (a) f is a Morse function.

(b) The ascending and descending manifolds of the critical points of f inter-sect transversely (see section 2.1).

(c) The closed orbits of the gradient āˆ‡f are nondegenerate (see below).

A standard transversality argument shows that these assumptions hold if f andthe metric are generic.

A closed orbit is a nonconstant map Ī³ : S1 ā†’ X with Ī³ā€²(t) = āˆ’Ī»āˆ‡f forsome Ī» > 0. We declare two closed orbits to be equivalent if they differ byreparametrization. The period p(Ī³) is the largest integer p such that Ī³ factorsthrough a pā€“fold covering S1 ā†’ S1 . A closed orbit is nondegenerate if det(1āˆ’dĻ†(x)) 6= 0, where Ļ† is the pth return map at a point x āˆˆ Ī³(S1) āŠ‚ X . If so,the Lefschetz sign Īµ(Ī³) is the sign of this determinant.

Notation 1.2 Let H1 := H1(X). Let Īø āˆˆ H1(X;Z) denote the pullback byf of the ā€œupwardā€ generator of H1(S1;Z).

Let Ī› = Nov(H1,āˆ’Īø) denote the Novikov ring [17, 5], consisting of functionsH1 ā†’ Z that are finitely supported on the set h āˆˆ H1 | āˆ’Īø(h) ā‰¤ C foreach C āˆˆ R. This ring has the obvious addition, and the convolution prod-uct. We denote a function a : H1 ā†’ Z by the (possibly infinite) formal sumāˆ‘

hāˆˆH1a(h) Ā· h .

Definition 1.3 [2, 20] We count closed orbits with the zeta function, whichis a function H1 ā†’ Z defined, in the above notation, by

Ī¶ := exp

āˆ‘Ī³āˆˆO

Īµ(Ī³)p(Ī³)

[Ī³]

. (1)

Here O denotes the set of closed orbits, and [Ī³] := Ī³āˆ—[S1] is the homology classof Ī³ in H1(X).

A compactness argument using Assumption 1.1, together with the observationthat āˆ’Īø([Ī³]) > 0 for all Ī³ , shows that Ī¶ āˆˆ Ī›āŠ—Q.

We remark that there is also a product formula [3, 7, 8]

Ī¶ =āˆĪ³āˆˆI

(1āˆ’ (āˆ’1)iāˆ’ [Ī³])āˆ’(āˆ’1)i0 . (2)

Circle-valued Morse theory and Reidemeister torsion

Geometry and Topology, Volume 3 (1999)

371

Page 4: Circle-valued Morse theory and Reidemeister torsion

Here I denotes the set of irreducible (period 1) closed orbits, and iāˆ’(Ī³) andi0(Ī³) denote the numbers of real eigenvalues of the return map in the intervals(āˆ’āˆž,āˆ’1) and (āˆ’1, 1), respectively. Equation (2) shows that in fact Ī¶ āˆˆ Ī›,ie, Ī¶ has integer coefficients. A third formula for the zeta function, in terms offixed points of return maps, is given in equation (13).

We now introduce a notion of topological Reidemeister torsion following Turaev[28], and an analogous notion of Morse-theoretic torsion. Detailed definitionsare given in section 2.3.

Let X denote the universal (connected) abelian cover of X , whose automor-phism group is H1(X). A smooth triangulation of X lifts to X and gives rise toa chain complex Cāˆ—(X) over Z[H1]. The Reidemeister torsion of this complexis an element of Q(Z[H1])/ Ā±H1 , where Q(R) denotes the total quotient ringof R. The Ā±H1 ambiguity arises because the Reidemeister torsion depends ona choice of ordered basis.

Turaev [27] showed that the H1 ambiguity can be resolved by the choice of anā€œEuler structureā€. The space Eul(X) of Euler structures is a natural affinespace over H1(X), reviewed in section 2.2. One can also resolve the signambiguity by choosing a homology orientation of X , ie, an orientation o ofāŠ•

iHi(X;Q) (see [26]). We can then define the Turaev torsion

Ļ„(X; o) : Eul(X)ā†’ Q(Z[H1]). (3)

This is an H1ā€“equivariant map which does not depend on the triangulation. Wewrite Ļ„(X):=Ā±Ļ„(X, o); this is an H1ā€“equivariant map Eul(X)ā†’ Q(Z[H1])/Ā±1 .

Example 1.4 If X is the 3ā€“manifold obtained by zero surgery on a knotK āŠ‚ S3 , then for a suitable Euler structure Ī¾ ,

Ļ„(X)(Ī¾) =Alex(K)(1āˆ’ t)2

where Alex(K) āˆˆ Z[t, tāˆ’1]/ Ā± 1 is the Alexander polynomial of K and t is agenerator of H1(X) ' Z.

On the Morse theory side, the Novikov complex CNāˆ— is a chain complex overthe Novikov ring Ī›, whose chains are generated by critical points of the pullbackof f to X , and whose boundary operator counts gradient flow lines betweencritical points (see section 2.1). We can similarly define the Morse-theoretictorsion

Ļ„(CNāˆ—) : Eul(X)ā†’ Q(Ī›)/ Ā± 1. (4)

Michael Hutchings and Yi-Jen Lee

Geometry and Topology, Volume 3 (1999)

372

Page 5: Circle-valued Morse theory and Reidemeister torsion

Definition 1.5 Define I : Eul(X)ā†’ Q(Ī›)/Ā± 1 to be the product of the zetafunction and the Morse-theoretic torsion:

I := Ī¶ Ā· Ļ„(CNāˆ—).

Theorem 1.6 The Morse theory invariant I is independent of the metric anddepends only on the homotopy class of f , ie the cohomology class Īø .

One can prove this a priori ; see [7] for the rough idea and [6] for the details.Although this may help define related invariants in other contexts, in the presentcontext it is easier to compute I directly, which will prove Theorem 1.6 aposteriori . That is what we will do in this paper.

Theorem 1.7 (Main theorem) Our Morse theory invariant I is equal to thetopological torsion:

I = i(Ļ„(X))

as maps Eul(X)ā†’ Q(Ī›)/Ā± 1.

Here i : Q(Z[H1])ā†’ Q(Ī›) is induced by the inclusion Z[H1]ā†’ Ī›.

Remarks 1.8 (1) In the extreme case when there are no critical points, Xis a mapping torus and this theorem reduces to an equivariant version of theLefschetz fixed point theorem, cf [15, 3].

(2) The extreme case when Īø = 0, so that f lifts to a real-valued Morsefunction, is also essentially classical (cf [16]), and we treat it in section 2.4. Inthis case the Morse-theoretic torsion is a topological invariant; lack of invarianceand existence of closed orbits arise simultaneously when we pass from real-valued to circle-valued Morse theory.

(3) The class Īø āˆˆ H1(X), regarded as a map H1(X) ā†’ Z, induces a mapĪ› ā†’ Z((t)) sending h 7ā†’ tĪø(h) . This in turn induces a partially defined mapQ(Ī›) ā†’ Q((t)). One can deduce the main result of our previous paper [7] byapplying this map to Theorem 1.7.

(4) The statement of Theorem 1.7 makes sense when df is replaced by a genericclosed 1ā€“form Āµ and the Novikov ring is graded by āˆ’[Āµ]. It seems possible toextend Theorem 1.7 to this case by approximating Āµ by closed 1ā€“forms inrational cohomology classes, to which Theorem 1.7 applies. Another proof forclosed 1ā€“forms is given in [6] by first proving Theorem 1.6 for closed 1ā€“forms,and then using this to reduce to the real-valued case.

Circle-valued Morse theory and Reidemeister torsion

Geometry and Topology, Volume 3 (1999)

373

Page 6: Circle-valued Morse theory and Reidemeister torsion

(5) Some previous papers, such as [19], studied the torsion of the Novikov com-plex (or the Whitehead torsion, which is sharper but only defined ā€œrelativelyā€unless the Novikov complex is acyclic), without considering the zeta function.In this case one can still obtain a topological invariant by modding out by unitsin the Novikov ring with leading coefficient 1. This is useful for understandingthe obstructions to the existence of nonvanishing closed 1ā€“forms [10, 4]. How-ever, the extra information in the zeta function is important for the connectionwith Seibergā€“Witten theory below.

(6) A homology orientation of X can apparently remove the sign ambiguityin Ļ„(CNāˆ—). However we have not checked if Theorem 1.7 holds with the signambiguity removed this way.

1.2 Application to Seibergā€“Witten theory

We now consider the special case when dim(X) = 3 and b1(X) > 0. LetSpinc(X) denote the set of spin-c structures on X . Given a homology orienta-tion o, the Seibergā€“Witten invariant of X is a function

SWX,o : Spinc(X)ā†’ Z

which counts Rā€“invariant solutions to the Seibergā€“Witten equations on XƗR,modulo gauge equivalence. (See eg [13, 14, 18].)

Taubes [24] has shown that the SW invariant of a symplectic fourā€“manifoldequals a ā€œGromov invariantā€ counting pseudoholomorphic curves. In [7] weproposed that using similar analysis, one might be able to show that the SWinvariant of a 3ā€“manifold is equal to a Morse theory invariant

I3 : Spinc(X)ā†’ Z.

The invariant I3 counts certain unions of closed orbits and flow lines of thegradient vector field of a Morse function f : X ā†’ S1 with no index 0 or 3critical points. We review the definition of I3 in section 4.

Conjecture 1.9 [7] The Seibergā€“Witten invariant agrees with our Morse the-ory invariant:

SWX,o = Ā±I3.

(When b1(X) = 1, the SW invariant also depends on a choice of ā€œchamberā€,and in this conjecture we use the chamber determined by r āˆ— df for r >> 0.)

Michael Hutchings and Yi-Jen Lee

Geometry and Topology, Volume 3 (1999)

374

Page 7: Circle-valued Morse theory and Reidemeister torsion

Remark 1.10 If f has no critical points, and if we arrange for df to beharmonic, then this conjecture is a corollary of Taubesā€™ theorem [24] appliedto the symplectic fourā€“manifold (X Ɨ S1, df āˆ§ ds + āˆ—Xdf). Here s denotesthe S1 coordinate. The idea is that for a suitable homology orientation, ifs āˆˆ Spinc(X), then

SWX,o(s) = SWXƗS1(Ļ€āˆ—s) = GrXƗS1(Ļ€āˆ—s) = I3(s).

The first equality expresses the fact that all solutions to the SW equationson X Ɨ S1 are S1ā€“invariant; see [18] for details of this equality. The secondequality is Taubesā€™ theorem; here GrXƗS1(Ļ€āˆ—s) counts, in the sense of [25],pseudoholomorphic curves in a certain S1ā€“invariant homology class in XƗS1 .An energy argument shows that for a suitable almost complex structure, everysuch curve is a union of closed orbits of āˆ‡f crossed with S1 . This leads to thethird equality (cf [8, Thm. 0.1]), using the fact that I3 is a reparametrizationof the zeta function in this case.

Salamon has proved a statement equivalent to Conjecture 1.9 in this case usinga different method [22].

In another direction, Turaev [28] conjectured a combinatorial formula for theSeibergā€“Witten invariant as follows. If dim(X) = 3 and b1(X) > 1, then for anEuler structure Ī¾ , the torsion Ļ„(X; o)(Ī¾) is actually in the group ring Z[H1].If b1 = 1, then i(Ļ„(X; o)(Ī¾)) āˆˆ Ī›, rather than in the quotient ring. Given ahomology orientation o, one can then define a map

T (X; o) : Eul(X) āˆ’ā†’ Z,Ī¾ 7āˆ’ā†’ i(Ļ„(X; o)(Ī¾))(0).

where (0) indicates evaluation on 0 āˆˆ H1 . (This depends on the sign of [df ]when b1 = 1.) There is also a natural isomorphism ı : Spinc(X) ā†’ Eul(X)([27], see section 4.3).

Conjecture 1.11 (Turaev [28]) The Seibergā€“Witten invariant agrees withthe Turaev torsion:

SWX,o = T (X; o) ı : Spinc(X)ā†’ Z.

This statement is a refinement of the theorem of Meng and Taubes [14], whichgives an ā€œaveragedā€ version of this equivalence, in which one sums over spin-cstructures that differ by torsion elements of H2(X;Z).

The invariant I3 turns out to be a reparametrization of the more general in-variant I . Thus we can apply Theorem 1.7 to prove:

Circle-valued Morse theory and Reidemeister torsion

Geometry and Topology, Volume 3 (1999)

375

Page 8: Circle-valued Morse theory and Reidemeister torsion

Theorem 1.12 Conjecture 1.9 is equivalent to Conjecture 1.11 (modulo signs).

The detailed proof is given in section 4.

Update (1) Three days after the first version of this paper was posted onthe internet, a preprint by Pajitnov [20] appeared, giving a result similar toTheorem 1.7, using Whitehead torsion.

(2) Turaev [29] has shown how to refine the methods of Meng and Taubes toprove Conjecture 1.11, modulo signs. Together with our results, this indirectlyproves Conjecture 1.9. However, one might still desire a direct analytic proof.The following is a summary of the situation:

Thm. 1.7

S1 Morse theory = Turaev torsionanalytic proof? @@@@ Meng-Taubes/Turaev

Seibergā€“Witten

Acknowledgments This paper would not exist were it not for Taubesā€™ workon Seibergā€“Witten and Gromov invariants. We thank him for sharing his ideasgenerously. We also thank R Bott, R Forman, and D Salamon for helpfulconversations.

2 Background

We now give some necessary background. Section 2.1 reviews the definition ofthe Novikov complex, which counts gradient flow lines between critical points.Section 2.2 reviews Turaevā€™s Euler structures, which are needed for the mostrefined version of Reidemeister torsion. Section 2.3 gives the precise definitionsof the versions of Reidemeister torsion that we use. Finally, section 2.4 provesTheorem 1.7 for real-valued Morse functions, as a warmup for some of thearguments in section 3.

2.1 The Novikov complex

We begin with some standard definitions from Morse theory. If p is a criticalpoint in X of f or in X of the pullback of f , the descending manifold D(p) isthe set of all points x such that upward gradient flow starting at x converges

Michael Hutchings and Yi-Jen Lee

Geometry and Topology, Volume 3 (1999)

376

Page 9: Circle-valued Morse theory and Reidemeister torsion

to p. Similarly the ascending manifold A(p) is the set of all points from whichdownward gradient flow converges to p. If ind(p) = i, then D(p) and A(p) areembedded open balls of dimensions i and nāˆ’ i respectively.

The Novikov complex (CNāˆ—, āˆ‚f ) is defined as follows. On X , we can lift ourMorse function to a map f : X ā†’ R. Let CNi denote the set of (possiblyinfinite) linear combinations Ī± of critical points of index i in X , such that foreach C āˆˆ R, the sum Ī± contains only finitely many critical points p āˆˆ X withf(p) > C . The action of H1 on the critical points by covering transformationsinduces an action of the Novikov ring Ī› on CNi . In fact CNi is a free Ī›ā€“module; one can specify a basis for CNāˆ— by choosing a lift of each critical pointin X to X .

We define āˆ‚f : CNi ā†’ CNiāˆ’1 as follows. If p, q āˆˆ X are critical points of indexi and iāˆ’ 1 respectively, let 怈āˆ‚fp, q怉 denote the signed number of gradient flowlines from p to q . If p is a critical point of index i, define

āˆ‚fp :=āˆ‘q

怈āˆ‚fp, q怉q

where the sum is over all critical points q āˆˆ X of index i āˆ’ 1. We count flowlines using the sign conventions from [7]. These conventions are chosen so that(āˆ‚f )2 = 0 and so that equation (16) holds.

Theorem 2.1 (Novikov) The homology of the Novikov complex is naturallyisomorphic to the homology of the ā€œhalf-infiniteā€ chains in X :

Hāˆ—(CNāˆ—, āˆ‚f ) ' Hāˆ—(Cāˆ—(X)āŠ— Ī›).

See eg [17, 19, 21, 7]. For example, if X = S1 and f : S1 ā†’ S1 has nonzerodegree, then the homology of the Novikov complex vanishes.

2.2 Euler structures

We now discuss three different notions of ā€œEuler structureā€ and how they relate.One can ignore this material at the expense of allowing an H1 ambiguity inReidemeister torsion.

Definition 2.2 (Turaev [27]) If X is a closed smooth manifold with Ļ‡(X) =0 and n = dim(X) > 1, a smooth Euler structure on X is a nonsingularvector field on X , where two such vector fields are declared equivalent if theirrestrictions to the complement of a ball in X are homotopic through nonsingularvector fields. We let Eul(X) denote the space of smooth Euler structures.

Circle-valued Morse theory and Reidemeister torsion

Geometry and Topology, Volume 3 (1999)

377

Page 10: Circle-valued Morse theory and Reidemeister torsion

By obstruction theory, Eul(X) is an affine space over Hnāˆ’1(X;Ļ€nāˆ’1(Snāˆ’1)) =H1(X). (It is nonempty since Ļ‡(X) = 0.)

The following alternate definition of smooth Euler structures is useful for Morsetheory, and also works well when n = 1. If v is a vector field on X withnondegenerate zeroes, let H1(X, v) denote the set of homology classes of 1ā€“chains Ī³ āˆˆ X with āˆ‚Ī³ = vāˆ’1(0), where the points in vāˆ’1(0) are orientedin the standard way. The set H1(X, v) is a subset of the relative homologyH1(X, vāˆ’1(0)) and is an affine space over H1(X).

Definition 2.3 One can show by a cobordism argument that the spacesH1(X, v) for different v ā€™s are canonically isomorphic to each other, and henceto a single space. We call this space Eul(X). We let iv : H1(X, v) ā†’ Eul(X)denote the canonical isomorphism.

If n > 1, we can go from Definition 2.3 to Definition 2.2 as follows. GivenĪ³ āˆˆ H1(X, v), we can represent Ī³ by disjoint paths connecting the zeroes of vin pairs. We then construct a nonsingular vector field by cancelling the zeroes ofv in a neighborhood of Ī³ . (If v has no zeroes, we send 0 āˆˆ H1(X) = H1(X, v)to the Euler structure represented by v and extend equivariantly.)

Definition 2.4 [27] Let (X,T ) be a finite connected CWā€“complex with cellsĻƒi. (X denotes the underlying topological space; T denotes the cell struc-ture.) A combinatorial Euler structure on (X,T ) is a choice of a lift of each cellto the universal abelian cover X , where two such sets of lifts Ļƒi and hiĻƒi,with hi āˆˆ H1(X) = Aut(X), are considered equivalent if

āˆ‘i(āˆ’1)dim(Ļƒi)hi = 0.

We let Eul(X,T ) denote the space of combinatorial Euler structures of theCWā€“complex (X,T ). This is clearly an affine space over H1(X).

Note that if T is a refinement of the cell-structure T with cells Ļ„j, thenthere is a canonical isomorphism Eul(X,T )ā†’ Eul(X, T ) sending Ļƒi to Ļ„j,where Ļ„j āŠ‚ Ļƒi if Ļ„j āŠ‚ Ļƒi .

Lemma 2.5 [27] If X is a closed smooth manifold with a smooth triangula-tion T , then there is a natural isomorphism between the spaces of smooth andcombinatorial Euler structures:

Eul(X) ' Eul(X,T ).

Michael Hutchings and Yi-Jen Lee

Geometry and Topology, Volume 3 (1999)

378

Page 11: Circle-valued Morse theory and Reidemeister torsion

The idea is that there is a natural vector field on each simplex with a zero at thecenter of each face and which points into the simplex near the boundary. Thesepiece together to give a continuous vector field on X . We can perturb this toa smooth vector field vT with a nondegenerate zero of sign (āˆ’1)i in the centerof each iā€“simplex. Then a smooth Euler structure Ī¾ can be represented by achain Ī³ consisting of paths connecting the zeroes in pairs, with [Ī³] = iāˆ’1

vT(Ī¾) āˆˆ

H1(X, vT ). We can lift the chain Ī³ to X , and the induced lifts of its endpointsdetermine a combinatorial Euler structure.

2.3 Reidemeister torsion

We now review the definition of Reidemeister torsion of certain chain complexes.We then use this algebra to define Reidemeister torsion for the two geometriccomplexes we are interested in.

2.3.1 Algebra

Let (Cāˆ—, āˆ‚) be a finite complex of finite dimensional vector spaces over a field F .The standard short exact sequences 0ā†’ Zi ā†’ Ci ā†’ Biāˆ’1 ā†’ 0 and 0ā†’ Bi ā†’Zi ā†’ Hi ā†’ 0 induce canonical isomorphisms det(Ci) = det(Zi)āŠ—det(Biāˆ’1) anddet(Zi) = det(Bi)āŠ—det(Hi), where ā€˜detā€™ denotes top exterior power. Combiningthese isomorphisms gives an isomorphism

Ī¦:āŠ—i

det(Ci)(āˆ’1)i ā†’āŠ—i

det(Hi)(āˆ’1)i . (5)

Let e be an ordered basis for Cāˆ— , ie, an ordered basis ei for each Ci . Let h bean ordered basis for Hāˆ— . Let [e] āˆˆ

āŠ—i det(Ci)(āˆ’1)i and [h] āˆˆ

āŠ—i det(Hi)(āˆ’1)i

denote the resulting volume forms.

Definition 2.6 We define the Reidemeister torsion

Ļ„(Cāˆ—, e, h) := Ī¦([e])/[h] āˆˆ FƗ.

We also define

Ļ„(Cāˆ—, e) :=Ļ„(Cāˆ—, e, 1) if Hāˆ— = 0,

0 otherwise.

Usually we will be interested in Ļ„ rather than Ļ„ . In practice, we can computethe torsion Ļ„ in terms of an alternating product of determinants as follows.

Circle-valued Morse theory and Reidemeister torsion

Geometry and Topology, Volume 3 (1999)

379

Page 12: Circle-valued Morse theory and Reidemeister torsion

Lemma 2.7 If Hāˆ— = 0, we can find a decomposition Cāˆ— = Aāˆ— āŠ•Bāˆ— such that(i) Ai and Bi are spanned by subbases of the basis ei , and (ii) the map

Ļ€Biāˆ’1 āˆ‚|Ai : Ai ā†’ Biāˆ’1

(which we abbreviate by āˆ‚ : Ai ā†’ Biāˆ’1 ) is an isomorphism. Then

Ļ„(Cāˆ—, e) := Ā±āˆi

det(āˆ‚ : Ai ā†’ Biāˆ’1)(āˆ’1)i .

Here the determinants are computed using the subbases of e.

We now extend the definition of torsion to complexes over certain rings whichmight not be fields.

Definition 2.8 [28] Let R be a ring, and assume that its total quotient ring(denoted by Q(R)) is a finite sum of fields, Q(R) = āŠ•jFj . Let (Cāˆ—, āˆ‚) be afinite complex of finitely generated free Rā€“modules with an ordered basis e.We define

Ļ„(Cāˆ—, e) :=āˆ‘j

Ļ„(Cāˆ— āŠ—R Fj , eāŠ— 1) āˆˆāŠ•j

Fj = Q(R).

In this case Hāˆ—(C) might not be free, in which case it does not have a basis inthe usual sense. However in this paper we call a set h := hj a ā€œbasisā€ forHāˆ—(C) when hj is a basis for Hāˆ—(Cāˆ— āŠ—R Fj) for each j . Given h = hj, wedefine

Ļ„(Cāˆ—, e, h) :=āˆ‘j

Ļ„(Cāˆ— āŠ—R Fj , eāŠ— 1, hj) āˆˆ Q(R)Ɨ.

Example 2.9 If 0 ā†’ C2āˆ‚ā†’ C1 ā†’ 0 is a 2ā€“term complex with C1 = C2 , and

if e is a basis which is identical on C1 and C2 , then

Ļ„(Cāˆ—, e) = det(āˆ‚).

We are interested in the rings Z[H1] and Ī›. Their quotients are finite sums offields (see eg [28]), and these decompositions are compatible with the inclusionZ[H1]ā†’ Ī›.

In section 3.2, we will need the following product formula for torsion. Let R be aring such that Q(R) is a finite sum of fields Fj . Let 0ā†’ Cāˆ— ā†’ C ā€²āˆ— ā†’ C ā€²ā€²āˆ— ā†’ 0 bea short exact sequence of finite complexes of finitely generated free Rā€“modules.Let e, eā€², eā€²ā€² be bases for Cāˆ—, C ā€²āˆ—, C ā€²ā€²āˆ— compatible with the exact sequence. Leth, hā€², hā€²ā€² be bases for the homology as in Definition 2.8. Let Lāˆ— denote the longexact sequence in homology, regarded as an acyclic chain complex, and let bdenote the basis for Lāˆ— obtained by combining h, hā€², hā€²ā€² .

Michael Hutchings and Yi-Jen Lee

Geometry and Topology, Volume 3 (1999)

380

Page 13: Circle-valued Morse theory and Reidemeister torsion

Lemma 2.10 We have the following product formula for torsion:

Ļ„(C ā€², eā€², hā€²) = Ļ„(C, e, h)Ļ„ (C ā€²ā€², eā€²ā€², hā€²ā€²)Ļ„(Lāˆ—, b).

Proof This follows from [16].

2.3.2 Geometry

We now define the Turaev torsion (3). In the notation of Definition 2.4, let(X,T ) be a finite connected CWā€“complex with universal abelian cover X . Lif-ing the cells gives a chain complex Cāˆ—(X, T ) over Z[H1(X)]. A combinatorialEuler structure Ī¾ determines a set of lifts of each cell to X , up to equiva-lence. Choose one set of lifts; this gives a basis for Cāˆ—(X, T ). A homologyorientation o determines an orientation of this basis, via the isomorphism (5)applied to Cāˆ—(X,T ). Let e(Ī¾, o) denote the resulting ordered basis. We de-fine the combinatorial Turaev torsion Ļ„(X,T ; o) to be the H1ā€“equivariant mapEul(X,T )ā†’ Q(Z[H1]) given by

Ļ„(X,T ; o)(Ī¾) := Ļ„(Cāˆ—(X, T ), e(Ī¾, o)) āˆˆ Q(Z[H1]).

Note that the right hand side of this equation does not depend on the choiceof a set of lifts. Furthermore, Ļ„(X,T ; o) = Ļ„(X, T ; o) under the canonicalisomorphism Eul(X,T )ā†’ Eul(X, T ), if T is a refinement of T .

Definition 2.11 Let X be a closed connected smooth manifold with Ļ‡(X) =0, with a smooth Euler structure Ī¾ āˆˆ Eul(X) and a homology orientationo. Choose a smooth triangulation T of X . Let Ī¾T āˆˆ Eul(X,T ) denote thecombinatorial Euler structure equivalent to Ī¾ via Lemma 2.5. We define theTuraev torsion

Ļ„(X; o)(Ī¾) := Ļ„(X,T ; o)(Ī¾T ).

The results of [27] show that the Turaev torsion does not depend on the choiceof smooth triangulation T .

We now define the Morse theoretic torsion (4). A smooth Euler structure Ī¾ canbe represented by a chain Ī³ connecting the critical points of f in pairs, with[Ī³] = iāˆ’1

āˆ‡f (Ī¾) āˆˆ H1(X,āˆ‡f). We can lift Ī³ to X , and the induced lifts of theendpoints determine a basis e(Ī¾) for CNāˆ— .

Definition 2.12 We define the Morse theoretic torsion Ļ„(CNāˆ—) : Eul(X) ā†’Q(Ī›)/ Ā± 1 by

Ļ„(CNāˆ—)(Ī¾) := Ļ„(CNāˆ—, e(Ī¾)).

Circle-valued Morse theory and Reidemeister torsion

Geometry and Topology, Volume 3 (1999)

381

Page 14: Circle-valued Morse theory and Reidemeister torsion

The map Ļ„(CNāˆ—) is H1ā€“equivariant, and again does not depend on the choiceof lifting. There is a sign ambiguity because the basis e(Ī¾) is unordered. Inthe special case when f has no critical points, we define Ļ„(CNāˆ—) to be the H1

equivariant map such that Ļ„(CNāˆ—)(Ī¾) = 1 for the smooth Euler structure Ī¾represented by āˆ‡f .

In the future, we call two bases for Cāˆ—(X, T ) or CNāˆ— equivalent if they corre-spond to the same Euler structure Ī¾ .

2.4 The real-valued case

Before proceeding more deeply into circle-valued Morse theory, it will be usefulto prove the main theorem for real-valued Morse functions.

Lemma 2.13 Theorem 1.7 holds when f : X ā†’ S1 lifts to a real-valued Morsefunction X ā†’ R.

Proof In this case Ī¶ = 1, so we just need to check that the Morse theoreticand topological torsions agree. This is essentially classical (cf [16]), except forthe identification of the bases determined by an Euler structure.

If Ī¾ is an Euler structure, then the bifurcation analysis in [11] shows thatĻ„(CNāˆ—)(Ī¾) is independent of the real-valued Morse function and the metric. Itis not hard to check that the Euler structures work out at each stage. In [11] itis assumed that the metric has a standard form near the critical points, but thiscan be arranged by a perturbation which does not affect the Novikov complex.

Now let T be a smooth triangulation of X . We can apparently perturb thevector field VT of section 2.2 so that it is the gradient of a Morse functionF : X ā†’ R with respect to some metric. In this case, the Novikov complex(CNāˆ—, āˆ‚F ) is identical to the chain complex (Cāˆ—(X;T ), āˆ‚). Moreover the basesdetermined by Ī¾ agree. Thus the Morse-theoretic and topological torsion areequal.

3 Proof of the main theorem

We will now prove the main theorem as follows. In section 3.1 we prepare for thecomputation of torsion by constructing a cell complex X ā€² which ā€œapproximatesā€X and is adapted to the vector field āˆ‡f . In section 3.2 we prove a technical

Michael Hutchings and Yi-Jen Lee

Geometry and Topology, Volume 3 (1999)

382

Page 15: Circle-valued Morse theory and Reidemeister torsion

lemma (Lemma 3.6) asserting that X and X ā€² have the same Reidemeistertorsion. The heart of the proof is in section 3.3 and section 3.4, where wedetermine the torsion of X ā€² by a short computation, and then interpret theanswer geometrically to recover the invariant I .

3.1 The cell complex X ā€²

Assume 0 āˆˆ S1 = R/Z is a regular value of f (by composing f with a rotationif necessary). Let Ī£ := fāˆ’1(0). Let Y be the compact manifold with boundaryobtained by cutting X along Ī£. We can write āˆ‚Y = Ī£1 t Ī£0 , where Ī£i iscanonically isomorphic to Ī£, and āˆ’āˆ‡f points inward along Ī£1 .

We give Y a cell decomposition as follows. Let T1 be a smooth triangulationof Ī£1 such that each simplex is transverse to the ascending manifolds of thecritical points in Y . If p āˆˆ Y is a critical point, let D0(p) denote the descendingmanifold of p in Y . If Ļƒ āˆˆ T1 is a simplex, let F(Ļƒ) denote the set of ally āˆˆ Y such that upward gradient flow starting at y hits Ļƒ . Choose a celldecomposition T0 of Ī£0 , such that the intersections with Ī£0 of D0(p) andF(Ļƒ) are subcomplexes, for each critical point p and each simplex Ļƒ āˆˆ T1 .

Lemma 3.1 The cells in T1 and T0 , together with all the cells D0(p) andF(Ļƒ), give a legitimate cell decomposition, T ā€²Y , of Y .

Proof Recall that D0(p) and F(Ļƒ) have natural compactifications using bro-ken flow lines (cf [7]). It may be shown by ā€œinduction on heightā€ that thesecompactifications are homeomorphic to closed balls. There are moreover natu-ral continuous maps of the compactifications to Y which send the interiors ofthe balls homeomorphically to D0(p) and F(Ļƒ). The transversality conditionon T1 and Assumption 1.1(b) ensure that the boundary of a cell consists oflower dimensional cells in T ā€²Y .

We would like to glue the boundary components of Y back together to obtain anice cell decomposition of X , but usually T0 will not agree with T1 . To correctfor this, let Ļ : (Ī£0, T0)ā†’ (Ī£1, T1) be a cellular approximation to the canonicalidentification Ī£0 ā†’ Ī£1 . Consider the mapping cylinder of Ļ:

MĻ =(Ī£0 Ɨ [0, 1]) t Ī£1

(x, 1) āˆ¼ Ļ(x).

This has a cell decomposition consisting of T0 and T1 , together with the cellsāˆ† Ɨ (0, 1) for each āˆ† āˆˆ T0 . There is a canonical inclusion Ī£0 ā†’ MĻ sendingx 7ā†’ (x, 0), and there is also a canonical inclusion Ī£1 ā†’MĻ .

Circle-valued Morse theory and Reidemeister torsion

Geometry and Topology, Volume 3 (1999)

383

Page 16: Circle-valued Morse theory and Reidemeister torsion

Definition 3.2 Let X ā€² be the space obtained by gluing Y and MĻ alongĪ£0 tĪ£1 .

The space X ā€² inherits a cell decomposition, but for our computations we prefera simpler cell decomposition, obtained by fusing some cells together as follows.If āˆ† is a cell in Y of the form D0(p) or F(Ļƒ), we define a corresponding cellin X ā€² by

āˆ† := āˆ† āˆŖ ((āˆ‚āˆ† āˆ© Ī£0)Ɨ [0, 1)).

Here (āˆ‚āˆ† āˆ© Ī£0)Ɨ [0, 1) indicates a subset of MĻ .

Definition 3.3 Let T ā€² be the cell decomposition of X ā€² consisting of cells ofthe following types:

(a) D0(p) for p āˆˆ Y a critical point;

(b) simplices in T1 ;

(c) F(Ļƒ) for Ļƒ āˆˆ T1 .

3.2 X and X ā€² have the same Reidemeister torsion

We now show that X and X ā€² have the same Reidemeister torsion, if the Eulerstructures are compatible in an appropriate sense.

We begin by noting that Hāˆ—(X ā€²) = Hāˆ—(X), and Hāˆ—(X ā€²) = Hāˆ—(X) as Z[H1]ā€“modules, as one can see from the exact sequences (8), (9) below. Note that theuniversal abelian cover X ā€² of X ā€² is obtained from X by modifying a neighbor-hood of the inverse image of Ī£.

Notation 3.4 (1) If Z is a subset of X or X ā€² , then Z will denote the inverseimage of Z in X or X ā€² . So Z is usually not the universal abelian cover of Z .

(2) We omit the cell structures from the notation when they are clear fromcontext.

A smooth Euler structure Ī¾ on X corresponds to an equivalence class of liftsof the critical points of f to X , as in section 2.3.2. A combinatorial Eulerstructure Ī¾ā€² on X ā€² consists of an equivalence class of lifts of the cells T ā€² to X ā€² .

Definition 3.5 We say that Ī¾ and Ī¾ā€² are compatible if, within these equiva-lence classes, the lifts can be chosen so that:

Michael Hutchings and Yi-Jen Lee

Geometry and Topology, Volume 3 (1999)

384

Page 17: Circle-valued Morse theory and Reidemeister torsion

(a) The lift of each critical point p in Ī¾ is contained in the lift of the cellD0(p) in Ī¾ā€² .

(b) For each simplex Ļƒ āˆˆ T1 , the lift of Ļƒ in Ī¾ā€² agrees with the ā€œtopā€ of thelift of F(Ļƒ) in Ī¾ā€² .

The compatibility conditions in Definition 3.5 induce an isomorphism fromEul(X) to Eul(X ā€², T ā€²) as affine spaces over H1 .

Recall that Ļ„(X) := Ā±Ļ„(X, o); similarly write Ļ„(X ā€², T ā€²) := Ā±Ļ„(X ā€², T ā€², o).

Lemma 3.6 If the Euler structures Ī¾ āˆˆ Eul(X) and Ī¾ā€² āˆˆ Eul(X ā€², T ā€²) arecompatible as above, then

Ļ„(X)(Ī¾) = Ā±Ļ„(X ā€², T ā€²)(Ī¾ā€²).

Proof The strategy is to compute the torsion of X and X ā€² by cutting theminto pieces and using the product formula (Lemma 2.10) applied to variousexact sequences, and see that we obtain the same answer. We proceed in threesteps.

Step 1 Consider the cell decomposition on Ī£Ć—[0, 1] consisting of cells āˆ†Ć—0,āˆ†Ć— (0, 1), and āˆ†Ć—1, where āˆ† is a lift of a simplex āˆ† āˆˆ T1 . Also recall thatMĻ has a natural cell decomposition. We claim that with respect to these cellstructures,

Ļ„ (Cāˆ—(Ī£Ć— [0, 1]), eĪ£, h) = Ā±Ļ„(Cāˆ—(MĻ), eā€²Ī£, hā€²), (6)

provided that the bases eĪ£, h, eā€²Ī£, h

ā€² satisfy the following conditions:

(a) The bases h, hā€² for homology agree under the isomorphism induced bythe canonical map Ī£Ć— [0, 1]ā†’MĻ .

(b) The bases eĪ£, eā€²Ī£ are given by lifts of cells such that:

(i) The lifts of the cells in Ī£Ć— 1 āŠ‚ Ī£Ć— [0, 1] and Ī£1 āŠ‚MĻ agree.(ii) The lift of each cell āˆ† in Ī£Ć— 0 āŠ‚ Ī£Ć— [0, 1] or in Ī£0 Ɨ 0 āŠ‚MĻ

is adjacent to the lift of the cell āˆ† Ɨ (0, 1) in Ī£ Ɨ [0, 1] or MĻ

respectively.

To prove (6), we compute both sides by applying the product formula for torsionto the relative exact sequences

0ā†’ Cāˆ—(Ī£Ć— 1)ā†’ Cāˆ—(Ī£Ć— [0, 1])ā†’ Cāˆ—(Ī£Ć— [0, 1], Ī£ Ɨ 1)ā†’ 0,

0ā†’ Cāˆ—(Ī£1)ā†’ Cāˆ—(MĻ)ā†’ Cāˆ—(MĻ, Ī£1)ā†’ 0.

Circle-valued Morse theory and Reidemeister torsion

Geometry and Topology, Volume 3 (1999)

385

Page 18: Circle-valued Morse theory and Reidemeister torsion

The answers agree, because condition (a) implies that the Ļ„(Lāˆ—) factors agree,condition (b(i)) implies that the Ļ„(Cāˆ—(Ī£)) factors agree, and condition (b(ii))implies that

Ļ„(Cāˆ—(Ī£Ć— [0, 1], Ī£ Ɨ 1), eĪ£, 1) = Ā±Ļ„(Cāˆ—(MĻ, Ī£1), eā€²Ī£, 1) = Ā±1.

Step 2 Let TY be a smooth triangulation on Y whose restriction to eachcomponent of āˆ‚Y = Ī£0 tĪ£1 agrees with T1 . The smooth Euler structure Ī¾ onX determines an equivalence class of bases, eY , for Cāˆ—(Y , Ī£0;TY ), because TYglues to a smooth triangulation of X .

Let T ā€²Y denote the cell decomposition of Y given by Lemma 3.1. The com-binatorial Euler structure Ī¾ā€² on X ā€² determines an equivalence class of baseseā€²Y for Cāˆ—(Y , Ī£0;T ā€²Y ), because the cells of T ā€²Y in Y \ Ī£0 are in one to onecorrespondence with the cells of X ā€² .

We claim that

Ļ„(Cāˆ—(Y , Ī£0;T ā€²Y ), eā€²Y , hā€²) = Ā±Ļ„(Cāˆ—(Y , Ī£0;TY ), eY , h) (7)

provided that the bases hā€² and h on homology agree.

To prove (7), note that the pullback of the Morse function f to Y lifts to areal-valued function f : Y ā†’ R. Let CMāˆ—(f) denote the Morse complex of fon the covering Y . A direct computation, using the compatibility of Ī¾ and Ī¾ā€² ,shows that

Ļ„(Cāˆ—(Y , Ī£0;T ā€²Y ), eā€²Y , hā€²) = Ā±Ļ„(CMāˆ—(f), eā€²ā€², hā€²ā€²).

(For similar calculations see section 3.3 and section 3.4; the result here corre-sponds essentially to setting t = 0 in (11) and Lemma 3.7.) Here the basis eā€²ā€²

for CMāˆ—(f) is determined by the lifts of the critical points determined by Ī¾as before, and we assume that the bases hā€², hā€²ā€² on homology agree under thestandard isomorphism Hāˆ—(Y , Ī£0) ' Hāˆ—(CMāˆ—(f)).

We also have

Ļ„(CMāˆ—(f), eā€²ā€², hā€²ā€²) = Ā±Ļ„(Cāˆ—(Y , Ī£0;TY ), eY , h).

The idea of the proof is to vary f in the space of Morse functions on Y such that(i) the gradient points outward along Ī£0 and does not point outward along Ī£1 ,and (ii) wherever the gradient is tangent to Ī£1 , the inward covariant derivativeof the gradient points inward. As in section 2.4, one can show that the resultingtorsion is independent of the Morse function. Deforming f to a Morse functionFY adapted to the triangulation TY , such that CMāˆ—(FY ) = Cāˆ—(Y , Ī£0, TY ),we have Ļ„(CMāˆ—(f), eā€²ā€², hā€²ā€²) = Ā±Ļ„(CMāˆ—(FY ), ifF (eā€²ā€²), hā€²ā€²), where ifF (eā€²ā€²) is the

Michael Hutchings and Yi-Jen Lee

Geometry and Topology, Volume 3 (1999)

386

Page 19: Circle-valued Morse theory and Reidemeister torsion

equivalence class of lifts (which correspond to bases of CMāˆ— ) induced from eā€²ā€²

via the homotopy from f to FY . But ifF (eā€²ā€²) = eY because the homotopy fromāˆ‡f to āˆ‡FY extends to a homotopy of vector fields on X = Y āˆŖ (Ī£Ć— I), whichis homotopic to a homotopy from āˆ‡f to the standard vector field associated tothe triangulation of X (cf end of section 2.2).

The above two equations prove (7).

Step 3 We now use (6) and (7) to compute the torsion of X and X ā€² .

We can regard X as the union of Y and Ī£Ć— [0, 1] along Ī£0tĪ£1 . Let T denotethe cell decomposition of X obtained by gluing the triangulation TY of Y tothe product cell structure on Ī£Ć— [0, 1] obtained from T1 . We then have a shortexact sequence

0ā†’ Cāˆ—(

Ī£0 t Ī£1

)ā†’ Cāˆ—(Ī£ Ɨ [0, 1]) āŠ• Cāˆ—(Y ;TY )ā†’ Cāˆ—(X ;T )ā†’ 0. (8)

Let T ā€² denote the ā€œunfusedā€ cell decomposition of X ā€² from section 3.1. Wethen have a short exact sequence

0ā†’ Cāˆ—(

Ī£0 t Ī£1

)ā†’ Cāˆ—(MĻ)āŠ• Cāˆ—(Y ;T ā€²Y )ā†’ Cāˆ—(X ā€²;T ā€²)ā†’ 0. (9)

We can choose representatives eY , eā€²Y such that they agree on Ī£ Ɨ 1 and

Ī£1 with eĪ£, eā€²Ī£ respectively. Let e(Ī¾) denote the basis for Cāˆ—(X ;T ) obtained

by combining the bases eĪ£, eY of Step 1 and Step 2 respectively. Similarly leteā€²(Ī¾ā€²) denote the basis for Cāˆ—(X ā€²;T ā€²) obtained by combining the bases eā€²Ī£, e

ā€²Y

of Step 1 and Step 2. Then e(Ī¾) is a representative of the combinatorial Eu-ler structure on (X, T ) corresponding to the smooth Euler structure Ī¾ , andeā€²(Ī¾ā€²) is a representative of the image of Ī¾ā€² under the canonical isomorphismEul(X ā€², T ā€²)ā†’ Eul(X ā€², T ā€²).

Applying the product formula to the above exact sequences, and using equations(6) and (7), we obtain

Ļ„(Cāˆ—(X;T ), e(Ī¾), h) = Ā±Ļ„(Cāˆ—(X ā€²;T ā€²), eā€²(Ī¾ā€²), hā€²). (10)

Here we are assuming that the bases h, hā€² for homology agree under the naturalisomorphism Hāˆ—(X) ' Hāˆ—(X ā€²). Also, to apply (7) in the above computation,one relates Y to the pair (Y ,Ī£0) as in Step 1.

In particular, equation (10) implies that

Ļ„(Cāˆ—(X ;T ), e(Ī¾)) = Ā±Ļ„(Cāˆ—(X ā€²;T ā€²), eā€²(Ī¾ā€²)).

This implies the lemma because Ļ„(Cāˆ—(X ;T ), e(Ī¾)) = Ļ„(X)(Ī¾), since the in-sertion of Ī£ Ɨ [0, 1] changes nothing, and similarly Ļ„(Cāˆ—(X ā€²;T ā€²), eā€²(Ī¾ā€²)) =Ļ„(X ā€², T ā€²)(Ī¾ā€²).

Circle-valued Morse theory and Reidemeister torsion

Geometry and Topology, Volume 3 (1999)

387

Page 20: Circle-valued Morse theory and Reidemeister torsion

3.3 Computing the torsion

We now compute the torsion of the approximating cell complex (X ā€², T ā€²), for acombinatorial Euler structure Ī¾ā€² compatible with a smooth Euler structure Ī¾on X as in the previous section.

Since Lemma 2.13 proves the theorem for real-valued Morse functions, we as-sume from now on that

Īø 6= 0.

Without loss of generality, we may also assume that Īø is indivisible in H1(X;Z).(If Īø is divisible by k , we can lift f to a kā€“fold cover of S1 without changingthe invariant I .) Let V := Ker(Īø), and choose a splitting

H1(X) = V āŠ• Z.Let t denote the generator of the Z component with Īø(t) = āˆ’1. Then theNovikov ring can be identified with the ring of formal Laurent series in t withcoefficients in Z[V ]:

Ī› = Z[V ]((t)).

Recall that Q(Ī›) is a finite sum of fields. To prove Theorem 1.7, it suffices toshow that it holds after projecting to each such field. Let K be a field com-ponent of Q(Ī›). By the Novikov isomorphism (Theorem 2.1), the complexesCNāˆ— āŠ—K and Cāˆ—(X)āŠ—K have isomorphic homology. So we will assume thatthese complexes are both acyclic, since otherwise they both have zero torsionĻ„ , and there is nothing to prove. In all of the calculations below, we implicitlytensor everything with the field K .

We can decompose

Ci(X ā€²) = Di āŠ•Ei āŠ• Fiwhere the three summands are generated by the cells of types (a), (b), and (c)respectively from Definition 3.3. Let us choose a basis e(Ī¾ā€²) for Ci(X ā€²) as inDefinition 3.5. We can identify

Fi ' Eiāˆ’1.

The matrix for the boundary operator on Ci(X ā€²) can then be written as

āˆ‚i =

Di Ei Fi

Diāˆ’1 Ni 0 Wi

Eiāˆ’1 āˆ’tMi āˆ‚Ī£i 1āˆ’ tĻ†iāˆ’1

Fiāˆ’1 0 0 āˆ’āˆ‚Ī£iāˆ’1

. (11)

Michael Hutchings and Yi-Jen Lee

Geometry and Topology, Volume 3 (1999)

388

Page 21: Circle-valued Morse theory and Reidemeister torsion

We remark that āˆ‚Ī£i is the boundary operator on Cāˆ—(Ī£). Also Ļ†iāˆ’1 is a matrix

with entries in Z[V ], which can be interpreted as the return map of the gradientflow from Ī£ to Ī£, after perturbation by our cellular approximation Ļ. LikewiseMi sends D0(p), where p āˆˆ X is a critical point, to a perturbation of theā€œdescending sliceā€ D0(p) āˆ© Ī£.

Continuing the calculation, due to the acyclicity assumption we may choosedecompositions Di = DA

i āŠ• DBi such that DA

i and DBi are spanned by (cells

corresponding to) critical points, and the differential āˆ‚f induces an isomor-phism DA

i ā†’ DBiāˆ’1 . In the notation below, we denote matrices with domain or

range Dāˆ— by boldface letters, and we denote their restrictions to DAi and/or

projections to DBiāˆ’1 by plain letters.

We now apply Lemma 2.7 with Ai = DAi āŠ• Fi and Bi = DB

i āŠ• Ei . (We willexplain in a moment why this choice of Ai and Bi is legitimate.) We obtain

Ļ„(Cāˆ—(X ā€², e(Ī¾ā€²)) =nāˆi=1

det(Ī©i)(āˆ’1)i+1

where

Ī©i =( DA

i Fi

DBiāˆ’1 Ni Wi

Eiāˆ’1 āˆ’tMi 1āˆ’ tĻ†iāˆ’1

).

We note that 1āˆ’ tĻ†iāˆ’1 is invertible because Ļ†iāˆ’1 has entries in Z[V ]. It followsthat

det(Ī©i) = det(1āˆ’ tĻ†iāˆ’1) det(Ki)

where

Ki := Ni + tWi(1āˆ’ tĻ†iāˆ’1)āˆ’1Mi : Di ā†’ Diāˆ’1. (12)

It will follow from Lemma 3.9(b) and the choice of DAi ,D

Bi that Ki is non-

singular, provided that the triangulation T1 is sufficiently fine and the cellularapproximation Ļ is sufficiently close to the identity. In particular, the matricesĪ©i are then nonsingular, so that Lemma 2.7 legitimately applies to the Ai andBi chosen above.

In conclusion, the above calculations imply the following lemma.

Lemma 3.7 If T1 is sufficiently fine and Ļ is sufficiently close to the identity,then

Ļ„(Cāˆ—(X ā€²))(Ī¾ā€²) =nāˆi=1

(det(1āˆ’ tĻ†iāˆ’1) det(Ki)

)(āˆ’1)i+1

.

Circle-valued Morse theory and Reidemeister torsion

Geometry and Topology, Volume 3 (1999)

389

Page 22: Circle-valued Morse theory and Reidemeister torsion

3.4 Geometric interpretation

We will now interpret the factors on the right side of Lemma 3.7 in terms ofMorse theory.

Notation 3.8 Suppose x, y are elements of Ī› = Z[V ]((t)), or matrices withentries in Ī›, which might depend on the choice of triangulation T1 and cellularapproximation Ļ. We write

x ā‰ˆ yif x āˆ’ y = O(tk), where k can be made arbitrarily large by choosing T1 suffi-ciently fine and Ļ sufficiently close to the identity.

Lemma 3.9 (a)āˆnāˆ’1i=0 det(1āˆ’ tĻ†i)(āˆ’1)i ā‰ˆ Ī¶ .

(b) Under the natural identification Dāˆ— ' CNāˆ— , we have

Ki ā‰ˆ āˆ‚fi .

Proof (a) Let f : X ā†’ R be a lift of f , and let Ī£ := fāˆ’1(0). The downwardgradient flow of f induces partially defined return maps

Ļ•k : Ī£ā†’ tkĪ£.

The definition (1) of Ī¶ is equivalent to

Ī¶ = exp

āˆ‘k>0, gāˆˆV

Fix(Ļ•k tāˆ’kgāˆ’1)gtk

k

āˆˆ Z[V ]((t)) = Ī›. (13)

Here Fix(s) counts fixed points of the equivariant map s modulo coveringtransformations, with their Lefschetz signs.

Suppose to begin that Ļ = id. By the machinery used to prove the Lefschetzfixed point theorem in [1], for each k we haveāˆ‘

gāˆˆVFix(Ļ•k tāˆ’kgāˆ’1) Ā· g =

nāˆ’1āˆ‘i=0

(āˆ’1)i Tr(Ļ†ki ) āˆˆ Z[V ]. (14)

In this case we have

Ī¶ =nāˆ’1āˆi=0

det(1āˆ’ tĻ†i)(āˆ’1)i . (15)

To see this, it is enough to check that the logarithmic derivatives of both sidesare equal, which follows from equations (13) and (14).

Michael Hutchings and Yi-Jen Lee

Geometry and Topology, Volume 3 (1999)

390

Page 23: Circle-valued Morse theory and Reidemeister torsion

In general, let H : Ī£Ć—[0, 1]ā†’ Ī£ be the homotopy from id to Ļ. In [7] we defineda natural compactification Ī“ āŠ‚ Ī£ Ɨ Ī£ of the graph of Ļ•. Using this one candefine a compactified graph Ī“it of (H(Ā·, t)Ļ•)i) in a similar manner. Now thereexists a positive integer N such that if the cells in T1 are all contained in ballsof radius Īµ, then the homotopy H can be chosen so that dist(H(t, x), x) < NĪµfor all t āˆˆ [0, 1] and x āˆˆ Ī£. (Cf the construction of H in [23]; by carefullycontrolling each intermediate step in the homotopy, the above claim may beachieved.) Also, the set of fixed points of Ļ•i lies in the interior of Ī“i under thediagonal map Ī£ā†’ Ī£Ć—Ī£ by definition of Ī“i , and is compact as a consequenceof Assumption 1.1. It follows that for any positive integer k we can choose Īµ sothat for all i ā‰¤ k and all t āˆˆ [0, 1], the compactified graph Ī“it does not cross thediagonal in Ī£ Ɨ Ī£. Then equation (14) will hold up to order k , and thereforeso will equation (15).

(b) If p is a critical point of index i, then

āˆ‚[D(p)] = [D(āˆ‚fp)] (16)

where the brackets indicate the fundamental class of the compactification of thedescending manifold [7]. Now suppose again that Ļ = id. Recall from equation(12) that the matrix Ki sends D0(p) to a linear combination of cells of the formD0(q), where q is a critical point of index iāˆ’ 1. (We will henceforth omit thehatsĖ†on D0 or F when Ļ = id, since in this case hatted and unhatted versionscan be identified.) In fact,

Ki(D0(p)) = D0(āˆ‚f (p)).

To see this, note that from the definition of Mi in equation (11), we have

[D(p)] = D0(p) +āˆžāˆ‘k=0

tk+1F(Ļ†kMi(D0(p))).

(Here the initial descending manifolds D0 and initial downward flow F aredefined as in section 3.1, but using X and Ī£ instead of X and Ī£.) Applyingequation (11) to this gives

āˆ‚[D(p)] = Ki(D0(p)) + (terms without initial descending manifolds). (17)

Equations (16) and (17) imply that āˆ‚fi = Ki when Dāˆ— (the domain/range ofKāˆ— ) is identified with CNāˆ— .

The case Ļ 6= id can be handled similarly to part (a).

Circle-valued Morse theory and Reidemeister torsion

Geometry and Topology, Volume 3 (1999)

391

Page 24: Circle-valued Morse theory and Reidemeister torsion

We can now complete the proof of Theorem 1.7. Lemmas 3.9(b) and 2.7 implythat

nāˆi=1

det(Ki))(āˆ’1)i+1 ā‰ˆ Ļ„(CNāˆ—)(Ī¾).

Together with Lemmas 3.6, 3.7, and 3.9(a), this implies that Theorem 1.7 holdsup to order k for all k .

4 The 3ā€“dimensional case and Seibergā€“Witten the-

ory

We now review from [7] the definition of the Morse-theoretic invariant I3 , andthe background for Conjecture 1.9 relating this invariant to Seibergā€“Wittentheory. We will then prove Theorem 1.12, relating this invariant to Turaevtorsion.

4.1 Motivation from Seibergā€“Witten theory

Let X be a closed connected oriented smooth 3ā€“manifold with b1(X) > 0. Lets be a spin-c structure on X . This determines a U(2)ā€“bundle S ā†’ X with aClifford action of TX on S . A section Ļˆ of S and a connection A on det(S)satisfy the Seibergā€“Witten equations with perturbation Ļ‰ if, in the notation of[9],

6 āˆ‚AĻˆ = 0,Ļ(FA) = iĻƒ(Ļˆ,Ļˆ) + iĻ(Ļ‰).

The Seibergā€“Witten invariant SW(s) counts solutions to these equations mod-ulo gauge equivalence. (For more on 3ā€“dimensional Seibergā€“Witten invariantssee eg [13, 14, 18].)

Let us choose the perturbation to be Ļ‰ = rāˆ—df , where f : X ā†’ S1 is harmonic,āˆ— denotes the Hodge star, and r is a real number. By perturbing the metric,we may arrange that f is a Morse function. Away from the critical points, thespinor bundle S splits into eigenspaces of Clifford multiplication by df ,

S = E āŠ• (E āŠ—Kāˆ’1), (18)

where Kāˆ’1 := Ker(df : TX ā†’ R).

Michael Hutchings and Yi-Jen Lee

Geometry and Topology, Volume 3 (1999)

392

Page 25: Circle-valued Morse theory and Reidemeister torsion

Taking r ā†’āˆž, one expects that for a Seibergā€“Witten solution the zero set ofthe E component of Ļˆ to become parallel to āˆ‡f . (The energy of the Seibergā€“Witten solution will be concentrated along this zero set. For detailed analysissee [24] and its sequels.) This suggests that SW(s) counts unions of closedorbits and flow lines of āˆ‡f starting and ending at critical points, whose totalhomology class is Poincare dual to c1(E).

The above homological condition implies that in our union of closed orbits andflow lines there is precisely one flow line starting at each index 2 critical pointand ending at each index 1 critical point. (See [7]. There are no index 0 or 3critical points because f is harmonic, and there are equally many index 1 andindex 2 points because Ļ‡(X) = 0.) In other words, in the notation of section2.2, our union of closed orbits and flow lines lives in H1(X, v), where v = āˆ’āˆ‡f .As in Remark 1.10, the counting of closed orbits is related to Taubesā€™ countingof pseudoholomorphic tori in symplectic 4ā€“manifolds [25], which indicates thatwe should allow closed orbits to be multiply covered when they are elliptic, butnot when they are hyperbolic.

4.2 The definition of I3

We now want to define I3(s) to be a signed count of such unions of closedorbits and flow lines. A convenient way to do so is to use generating functionsas follows. Choose orderings of the index 1 and index 2 critical points. Let Pijdenote the set of flow lines from the ith index 2 point to the jth index 1 point.Define the path matrix P by

P ij :=āˆ‘Ī³āˆˆPij

Īµ(Ī³)[Ī³].

Here [Ī³] āˆˆ H1(X, vāˆ’1(0)) is the homology class of Ī³ (oriented downward), andĪµ(Ī³) is the sign of Ī³ as in section 2.1. The entries of P live in the Novikovring of the relative homology group H1(X, vāˆ’1(0)) with grading given by minusintersection number with Ī£.

Note that det(P ), regarded as a Zā€“valued function on H1(X, vāˆ’1(0)), is sup-ported on H1(X, v). Also, the subset of Nov(H1(X, vāˆ’1(0))) consisting of func-tions supported on H1(X, v) is a Ī›ā€“submodule. A generating function countingunions of closed orbits and flow lines of the type we want is now given by

Iv3 := Ī¶ Ā· det(P ) āˆˆ Nov(H1(X, v)).

In the above equation, ā€˜Ā·ā€™ denotes the Ī› action. The closed orbits are countedcorrectly as a result of the product formula for the zeta function (2).

Circle-valued Morse theory and Reidemeister torsion

Geometry and Topology, Volume 3 (1999)

393

Page 26: Circle-valued Morse theory and Reidemeister torsion

Let jv : Spinc(X) ā†’ H1(X, v) denote the map that sends a spin-c structureto the Poincareā€“Lefschetz dual of c1(E), where E is the line bundle defined in(18). The map jv is an H1ā€“equivariant isomorphism.

Definition 4.1 [7] Regarding Iv3 as a function H1(X, v)ā†’ Z, we define

I3 := Iv3 jv : Spinc(X)ā†’ Z.

This definition makes sense for any Morse function f : X ā†’ S1 with no index 0or 3 critical points, even if f is not harmonic. The calculation below will showthat I3 does not depend on f , except that there is a global sign ambiguity inI3 due to the orientation choices we made. (Also I3 depends on the sign ofĪø āˆˆ H1(X;Z) when b1(X) = 1.)

4.3 Relation with Turaev torsion

To relate I3 to torsion, we note that the isomorphism iv jv : Spinc(X) ā†’Eul(X) does not depend on v . It follows that there is a canonical isomorphismĪ¹ : Spinc(X)ā†’ Eul(X). (This isomorphism was first defined by Turaev [27] ina different but equivalent way. The inverse map sends a smooth Euler structurerepresented by a nonsingular vector field u to the spin-c structure whose spinbundle is CuāŠ• uāŠ„ with a standard Clifford action.) In summary, we have thefollowing commutative triangle:

H1(X, v) ivāˆ’ā†’ Eul(X)jv Ī¹

Spinc(X)

Proof of Theorem 1.12 It is enough to show that

Iv3 iāˆ’1v = Ā±T (X; o) : Eul(X)ā†’ Z.

By Theorem 1.7 and the definition of T , this is equivalent to asserting that forsome orientation choice, we have

det(P )(iāˆ’1v (Ī¾)) = Ļ„(CNāˆ—)(Ī¾)(0)

for all Euler structures Ī¾ . If Ī¾0 is a reference Euler structure and Ī³ := iāˆ’1v (Ī¾0) āˆˆ

H1(X, v), then this equation is equivalent to

det(P )(Ā·+ Ī³) = Ļ„(CNāˆ—)(Ī¾0)(Ā·) : H1 ā†’ Z.

This last equation follows from the definition of Ļ„(CNāˆ—) and Example 2.9.

Michael Hutchings and Yi-Jen Lee

Geometry and Topology, Volume 3 (1999)

394

Page 27: Circle-valued Morse theory and Reidemeister torsion

References

[1] R F Brown, The Lefschetz fixed point theorem, Scott, Foresman and Co. (1971)

[2] D Fried, Homological identities for closed orbits, Invent. Math. 71 (1983) 419ā€“442

[3] D Fried, Lefschetz formulas for flows, from: ā€œThe Lefschetz centennial con-ference, part IIIā€, Contemp. Math. 58, III, Amer. Math. Soc. Providence, R.I.(1987) 19ā€“69

[4] K Fukaya, The symplectic sā€“cobordism conjecture: a summary, from: ā€œGeom-etry and Physics (Aarhus, 1995)ā€, Lecture Notes in Pure and Appl. Math. 184,Dekker (1997) 209ā€“219

[5] H Hofer, D Salamon, Floer homology and Novikov rings, from: ā€œThe Floermemorial volumeā€, Progr. Math. 133, Birkhauser, Basel (1995) 483ā€“524

[6] M Hutchings, Reidemeister torsion in generalized Morse theory, e-printmath.DG/9907066

[7] M Hutchings, Y-J Lee, Circle-valued Morse theory, Reidemeister torsion,and Seibergā€“Witten invariants of threeā€“manifolds, Topology, 38 (1999) 861ā€“888

[8] E Ionel, T Parker, Gromov invariants and symplectic maps, e-printdg-ga/9703013

[9] P B Kronheimer, T S Mrowka, The genus of embedded surfaces in the pro-jective plane, Math. Res. Lett. (1994) 797ā€“806

[10] F Latour, Existence de 1ā€“formes fermees non singulieres dans une classe decohomologie de de Rham, Inst. Hautes Etudes Sci. Publ. Math. No. 80 (1994)135ā€“194

[11] F Laudenbach, On the Thomā€“Smale complex, appendix to J-M Bismut &W Zhang, An extension of a theorem by Cheeger and Muller, Asterisque, 205(1992)

[12] Y-J Lee, Morse theory and Seibergā€“Witten monopoles on 3ā€“manifolds, HarvardUniversity thesis (1997)

[13] Y Lim, Seibergā€“Witten theory of 3ā€“manifolds, preprint (1996)

[14] G Meng, C H Taubes, SW = Milnor torsion, Math. Res. Lett. 3 (1996)661ā€“674

[15] J Milnor, Infinite cyclic coverings, Collected works vol. 2, Publish or Perish(1996)

[16] J Milnor, Whitehead torsion, Bull. Amer. Math. Soc. 72 (1966) 358ā€“426

[17] S P Novikov, Multivalued functions and functionals. An analogue of the Morsetheory, Soviet Math. Dokl. 24 (1981) 222ā€“226

[18] C Okonek, A Teleman, 3ā€“dimensional Seibergā€“Witten invariants and non-Kahlerian geometry, Math. Ann. 312 (1998) 261ā€“288

Circle-valued Morse theory and Reidemeister torsion

Geometry and Topology, Volume 3 (1999)

395

Page 28: Circle-valued Morse theory and Reidemeister torsion

[19] A Pazhitnov, On the Novikov complex for rational Morse forms, Ann. Fac.Sci. Toulouse Math. (6) 4 (1995) 297ā€“338

[20] A Pajitnov, Simple homotopy type of Novikov complex for closed 1ā€“forms andLefschetz zeta functions of the gradient flow, eprint dg-ga/9706014

[21] M Pozniak, Floer homology, Novikov rings, and clean intersections, Univ. ofWarwick thesis (1994)

[22] D Salamon, Seibergā€“Witten invariants of mapping tori, symplectic fixed points,and Lefschetz numbers, from: ā€œProceedings of Gokova Geometryā€“Topology Con-ference 1998ā€, (S Akbulut, T Onder and R Stern, editors), International Press(1999) 117ā€“143

[23] E Spanier, Algebraic topology, Springerā€“Verlag corrected reprint (1989)

[24] C H Taubes, The Seibergā€“Witten and Gromov invariants, Math. Res. Lett. 2(1995) 221ā€“238

[25] C H Taubes, Counting pseudo-holomorphic submanifolds in dimension 4, J.Differential Geom. 44 (1996) 818ā€“893

[26] V Turaev, Reidemeister torsion in knot theory, Russian Math. Surveys 41(1986) 119ā€“182

[27] V Turaev, Euler structures, nonsingular vector fields, and torsions of Reide-meister type, Math. USSR-Izv. 34 (1990) 627ā€“662

[28] V Turaev, Torsion invariants of Spinc structures on 3ā€“manifolds, Math. Res.Lett. 4 (1997) 679ā€“695

[29] V Turaev, A combinatorial formulation for the Seibergā€“Witten invariants of3ā€“manifolds, Math. Res. Lett. 5 (1998) 583ā€“598

[30] A Weil, Numbers of solutions of equations in finite fields, Bull. Amer. Math.Soc. 55 (1949) 497ā€“508

Michael Hutchings and Yi-Jen Lee

Geometry and Topology, Volume 3 (1999)

396