• Keine Ergebnisse gefunden

Equivariant Euler characteristics and K -homology Euler classes for proper cocompact G-manifolds

N/A
N/A
Protected

Academic year: 2021

Aktie "Equivariant Euler characteristics and K -homology Euler classes for proper cocompact G-manifolds"

Copied!
46
0
0

Wird geladen.... (Jetzt Volltext ansehen)

Volltext

(1)

Equivariant Euler characteristics and K -homology Euler classes for proper cocompact G-manifolds

Wolfgang L¨uck Jonathan Rosenberg

Institut f¨ur Mathematik und Informatik Westf¨alische Wilhelms-Universtit¨at

Einsteinstr. 62 48149 M¨unster, Germany

and

Department of Mathematics University of Maryland College Park, MD 20742, USA

Email: lueck@math.uni-muenster.de and jmr@math.umd.edu URL: wwwmath.uni-muenster.de/u/lueck, www.math.umd.edu/~jmr Abstract

Let G be a countable discrete group and let M be a smooth proper cocompact G-manifold without boundary. The Euler operator defines via Kasparov theory an element, called the equivariant Euler class, in the equivariantKO-homology ofM. The universal equivariant Euler characteristic ofM, which lives in a group UG(M), counts the equivariant cells of M, taking the component structure of the various fixed point sets into account. We construct a natural homomorphism from UG(M) to the equivariant KO-homology of M. The main result of this paper says that this map sends the universal equivariant Euler characteristic to the equivariant Euler class. In particular this shows that there are no “higher”

equivariant Euler characteristics. We show that, rationally, the equivariant Euler class carries the same information as the collection of the orbifold Euler characteristics of the components of the L-fixed point sets ML, where L runs through the finite cyclic subgroups of G. However, we give an example of an action of the symmetric group S3 on the 3-sphere for which the equivariant Euler class has order 2, so there is also some torsion information.

AMS Classification numbers Primary: 19K33

(2)

Secondary: 19K35, 19K56, 19L47, 58J22, 57R91, 57S30, 55P91

Keywords: equivariantK-homology, de Rham operator, signature operator, Kasparov theory, equivariant Euler characteristic, fixed sets, cyclic subgroups, Burnside ring, Euler operator, equivariant Euler class, universal equivariant Euler characteristic

(3)

0 Background and statements of results

Given a countable discrete groupGand a cocompact proper smoothG-manifold M without boundary and with G-invariant Riemannian metric, the Euler char- acteristic operator defines via Kasparov theory an element, theequivariant Eu- ler class, in the equivariant real K-homology group of M

EulG(M) ∈ KOG0(M). (0.1)

The Euler characteristic operator is the minimal closure, or equivalently, the maximal closure, of the densely defined operator

(d+d) : Ω(M)⊆L2(M)→L2(M),

with the Z/2-grading coming from the degree of a differential p-form. The equivariant signature operator is the same underlying operator, but with a dif- ferent grading coming from the Hodge star operator. The signature operator also defines an element

SignG(M) ∈ K0G(M),

which carries a lot of geometric information about the action of G on M. (Rationally, when G = {1}, Sign(M) is the Poincar´e dual of the total L- class, the Atiyah-SingerL-class, which differs from the Hirzebruch L-class only by certain well-understood powers of 2, but in addition, it also carries quite interesting integral information [11], [22], [27]. A partial analysis of the class SignG(M) for G finite may be found in [26] and [24].)

We want to study how much information EulG(M) carries. This has already been done by the second author [23] in the non-equivariant case. Namely, given a closed Riemannian manifold M, not necessarily connected, let

e: M

π0(M)

Z= M

π0(M)

KO0({∗}) → KO0(M)

be the map induced by the various inclusions {∗} → M. This map is split injective; a splitting is given by the various projectionsC→ {∗}forC∈π0(M), and sends{χ(C)|C∈π0(M)} to Eul(M). Hence Eul(M) carries precisely the same information as the Euler characteristics of the various components of M, and there are no “higher” Euler classes. Thus the situation is totally different from what happens with the signature operator.

We will see that in the equivariant case there are again no “higher” Euler characteristics and that EulG(M) is determined by the universal equivariant

(4)

Euler characteristic (see Definition 2.5) χG(M) ∈ UG(M) = M

(H)consub(G)

M

WH\π0(MH)

Z.

Here and elsewhere consub(G) is the set of conjugacy classes of subgroups of G and NH ={g∈G|g−1Hg=H} is the normalizer of the subgroup H⊆G and WH := NH/H is its Weyl group. The component of χG(M) associated to (H) ∈ consub(G) and WH ·C ∈ WH\π0(MH) is the (ordinary) Euler characteristic χ(WHC\(C, C ∩M>H)), where WHC is the isotropy group of C∈π0(MH) under the WH-action. There is a natural homomorphism

eG(M) : UG(M) → KOG0(M). (0.2) It sends the basis element associated to (H) ⊆ consub(G) and WH ·C ∈ WH\π0(MH) to the image of the class of the trivial H-representation Runder the composition

RR(H) =KO0H({∗})−−→(α) KO0G(G/H) KO

G 0(x)

−−−−−→KO0G(M),

where (α) is the isomorphism coming from induction via the inclusion α: H

→G and x: G/H → M is any G-map with x(1H) ∈C. The main result of this paper is

Theorem 0.3 (Equivariant Euler class and Euler characteristic) Let G be a countable discrete group and letM be a cocompact proper smooth G-manifold without boundary. Then

eG(M)(χG(M)) = EulG(M).

The proof of Theorem 0.3 involves two independent steps. Let Ξ be an equiv- ariant vector field on M which is transverse to the zero-section. Let Zero(Ξ) be the set of points x ∈ M with Ξ(x) = 0. Then G\Zero(Ξ) is finite. The zero-section i: M → T M and the inclusion jx: TxM → T M induce an iso- morphism of Gx-representations

Txi⊕T0jx: TxM⊕TxM −=→Ti(x)(T M)

if we identifyT0(TxM) =TxM in the obvious way. If pri denotes the projection onto the i-th factor fori= 1,2 we obtain a linear Gx-equivariant isomorphism

dxΞ : TxM −−→TxΞ Ti(x)(T M) (TxiTxjx)

1

−−−−−−−−−→ TxM ⊕TxM −−→pr2 TxM. (0.4) Notice that we obtain the identity if we replace pr2 by pr1 in the expression (0.4) above. One can even achieve that Ξ iscanonically transverse to the zero- section, i.e., it is transverse to the zero-section and dxΞ induces the identity

(5)

on TxM/(TxM)Gx for Gx the isotropy group of x under the G-action. This is proved in [29, Theorem 1A on page 133] in the case of a finite group and the argument directly carries over to the proper cocompact case. Define the index of Ξ at a zero x by

s(Ξ, x) = det (dxΞ)Gx: (TxM)Gx →(TxM)Gx

|det ((dxΞ)Gx: (TxM)Gx →(TxM)Gx)| ∈ {±1}.

For x∈M let αx: Gx→G be the inclusion, (αx): RR(Gx) =KOG0x({∗})→ KO0G(G/Gx) be the map induced by induction via αx and let x: G/Gx →M be the G-map sending g to g·x. By perturbing the equivariant Euler operator using the vector field Ξ we will show :

Theorem 0.5 (Equivariant Euler class and vector fields) LetGbe a countable discrete group and let M be a cocompact proper smooth G-manifold without boundary. Let Ξ be an equivariant vector field which is canonically transverse to the zero-section. Then

EulG(M) = X

Gx∈G\Zero(Ξ)

s(Ξ, x)·KOG0(x)◦(αx)([R]),

where [R]∈ RR(Gx) =K0Gx({∗}) is the class of the trivial Gx-representation R, we consider x as a G-map G/Gx→M and αx: Gx →G is the inclusion.

In the second step one has to prove eG(M)(χG(M)) = X

Gx∈G\Zero(Ξ)

s(Ξ, x)·KOG0(x)◦(αx)([R]). (0.6) This is a direct conclusion of the equivariant Poincar´e-Hopf theorem proved in [20, Theorem 6.6] (in turn a consequence of the equivariant Lefschetz fixed point theorem proved in [20, Theorem 0.2]), which says

χG(M) = iG(Ξ). (0.7)

where iG(Ξ) is the equivariant index of the vector field Ξ defined in [20, (6.5)].

Since we get directly from the definitions eG(M)(iG(Ξ)) = X

Gx∈G\Zero(Ξ)

s(Ξ, x)·KO0G(x)◦(αx)([R]), (0.8) equation (0.6) follows from (0.7) and (0.8). Hence Theorem 0.3 is true if we can prove Theorem 0.5, which will be done in Section 1.

(6)

We will factorize eG(M) as eG(M) : UG(M) e

G 1(M)

−−−−→H0Or(G)(M;RQ) e

G 2(M)

−−−−→H0Or(G)(M;RR)

eG3(M)

−−−−→KO0G(M), where H0Or(G)(M;RF) is the Bredon homology of M with coefficients in the coefficient system which sends G/H to the representation ring RF(H) for the field F =Q,R. We will show that eG2(M) and eG3(M) are rationally injective (see Theorem 3.6). We will analyze the map eG1(M), which is not rationally injective in general, in Theorem 3.21.

The rational information carried by EulG(M) can be expressed in terms of orbifold Euler characteristics of the various components of the L-fixed point sets for all finite cyclic subgroups L ⊆ G. For a component C ∈ π0(MH) denote by WHC its isotropy group under the WH-action on π0(MH). For H⊆G finite WHC acts properly and cocompactly on C and itsorbifold Euler characteristic (see Definition 2.5), which agrees with the more general notion of L2-Euler characteristic,

χQWHC(C)∈Q,

is defined. Notice that for finite WHC the orbifold Euler characteristic is given in terms of the ordinary Euler characteristic by

χQWHC(C) = χ(C)

|WHC|. There is a character map (see (2.6))

chG(M) : UG(M)→ M

(H)∈consub(G)

M

WH\π0(MH)

Q

which sends χG(M) to the various L2-Euler characteristics χQWHC(C) for (H)∈consub(G) andWH·C ∈WH\π0(ML). Recall that rationally EulG(M) carries the same information as eG1(M)(χG(M)) since the rationally injec- tive map eG3(M) ◦ eG2(M) sends eG1(M)(χG(M)) to EulG(M). Rationally eG1(M)(χG(M)) is the same as the collection of all these orbifold Euler charac- teristics χQWHC(C) if one restricts to finite cyclic subgroups H. Namely, we will prove (see Theorem 3.21):

Theorem 0.9 There is a bijective natural map γQG: M

(L)∈consub(G) Lfinite cyclic

M

WL\π0(ML)

Q

=

−→ Q⊗ZH0Or(G)(X;RQ)

(7)

which maps

QWLC)(C)|(L)∈consub(G), L finite cyclic, WL·C ∈WL\π0(ML)} to 1⊗ZeG1(M)(χG(M)).

However, we will show that EulG(M) does carry some torsion information.

Namely, we will prove:

Theorem 0.10 There exists an action of the symmetric group S3 of order 3!

on the 3-sphere S3 such that EulS3(S3)∈KOS03(S3) has order 2.

The relationship between EulG(M) and the various notions of equivariant Euler characteristic is clarified in sections 2 and 4.3.

The paper is organized as follows:

1. Perturbing the equivariant Euler operator by a vector field 2. Review of notions of equivariant Euler characteristic 3. The transformation eG(M)

4. Examples

4.1. Finite groups and connected non-empty fixed point sets 4.2. The equivariant Euler class carries torsion information 4.3. Independence of EulG(M) andχGs(M)

4.4. The image of the equivariant Euler class under assembly References

This paper subsumes and replaces the preprint [25], which gave a much weaker version of Theorem 0.3.

This research was supported by Sonderforschungsbereich 478 (“Geometrische Strukturen in der Mathematik”) of the University of M¨unster. Jonathan Rosen- berg was partially supported by NSF grants DMS-9625336 and DMS-0103647.

1 Perturbing the equivariant Euler operator by a vector field

Let Mn be a complete Riemannian manifold without boundary, equipped with an isometric action of a discrete group G. Recall that the de Rham operator D = d+d, acting on differential forms on M (of all possible degrees) is a formally self-adjoint elliptic operator, and that on the Hilbert space of L2

(8)

forms, it is essentially self-adjoint [8]. With a certain grading on the form bundle (coming from the Hodge ∗-operator), D becomes the signature operator; with the more obvious grading of forms by parity of the degree,D becomes theEuler characteristic operator or simply theEuler operator. When M is compact and G is finite, the kernel of D, the space of harmonic forms, is naturally identified with the real or complex1 cohomology of M by the Hodge Theorem, and in this way one observes that the (equivariant) index of D (with respect to the parity grading) in the real representation ring of G is simply the (equivariant homological) Euler characteristic of M, whereas the index with respect to the other grading is the G-signature [2].

Now by Kasparov theory (good general references are [4] and [9]; for the de- tailed original papers, see [12] and [13]), an elliptic operator such as D gives rise to an equivariant K-homology class. In the case of a compact manifold, the equivariant index of the operator is recovered by looking at the image of this class under the map collapsing M to a point. However, the K-homology class usually carries far more information than the index alone; for example, it determines the G-index of the operator with coefficients in any G-vector bundle, and even determines the families index in KG(Y) of a family of twists of the operator, as determined by a G-vector bundle on M ×Y. (Y here is an auxiliary parameter space.) When M is non-compact, things are similar, except that usually there is no index, and the class lives in an appropriate Kasparov group KG−∗(C0(M)), which is locally finite KG-homology, i.e., the relative group KG(M ,{∞}), where M is the one-point compactification of M.2 We will be restricting attention to the case where the action of G is proper and cocompact, in which case KG−∗(C0(M)) may be viewed as a kind of orbifold K-homology for the compact orbifold G\M (see [4, Theorem 20.2.7].) We will work throughout with real scalars and realK-theory, and use a variant of the strategy found in [23] to prove Theorem 0.5.

Proof of Theorem 0.5 Recall that since Ξ is transverse to the zero-section, its zero set Zero(Ξ) is discrete, and sinceM is assumed G-cocompact, Zero(Ξ) consists of only finitely manyG-orbits. Write Zero(Ξ) = Zero(Ξ)+qZero(Ξ),

1depending on what scalars one is using

2HereC0(M) denotes continuous real- or complex-valued functions onM vanishing at infinity, depending on whether one is using real or complex scalars. This algebra is contravariant inM, so a contravariant functor ofC0(M) iscovariant inM. Excision in Kasparov theory identifies KG−∗(C0(M)) with KG−∗(C(M), C(pt)), which is identified with relative KG-homology. When M does not have finite G-homotopy type, KG- homology here means SteenrodKG-homology, as explained in [10].

(9)

according to the signs of the indices s(Ξ, x) of the zeros x ∈ Zero(Ξ). We fix a G-invariant Riemannian metric on M and use it to identify the form bundle ofM with the Clifford algebra bundle Cliff(T M) of the tangent bundle, with its standard grading in which vector fields are sections of Cliff(T M), and D with the Dirac operator on Cliff(T M).3 (This is legitimate by [15, II, Theorem 5.12].) Let H = H+ ⊕ H be the Z/2-graded Hilbert space of L2 sections of Cliff(T M). Let A be the operator on H defined by right Clifford multiplication by Ξ on Cliff(T M)+ (the even part of Cliff(T M)) and by right Clifford multiplication by −Ξ on Cliff(T M) (the odd part). We use right Clifford multiplication since it commutes with the symbol ofD. Observe thatA is self-adjoint, with square equal to multiplication by the non-negative function

|Ξ(x)|2. Furthermore, A is odd with respect to the grading and commutes with multiplication by scalar-valued functions.

For λ≥ 0, let Dλ = D+λA. As in [23], each Dλ defines an unbounded G- equivariant Kasparov module in the same Kasparov class asD. In the “bounded picture” of Kasparov theory, the corresponding operator is

Bλ =Dλ 1 +Dλ212

= 1 λDλ

1 λ2 + 1

λ2D2λ 1

2

. (1.1)

The axioms satisfied by this operator that insure that it defines a Kasparov KG-homology class (in the “bounded picture”) are the following:

(B1) It is self-adjoint, of norm ≤1, and commutes with the action of G.

(B2) It is odd with respect to the grading of Cliff(T M).

(B3) For f ∈C0(M), f Bλ ∼ Bλf and f Bλ2 ∼f, where ∼ denotes equality modulo compact operators.

We should point out that (B1) is somewhat stronger than it needs to be when G is infinite. In that case, we can replace invariance of Bλ under G by “G- continuity,” the requirement (see [13] and [4,§20.2.1]) that

(B10) f(g·Bλ−Bλ)∼0 for f ∈C0(M), g∈G.

In order to simplify the calculations that are coming next, we may assume with- out loss of generality that we’ve chosen the G-invariant Riemannian metric on M so that for each z∈Zero(Ξ), in some small open Gz-invariant neighborhood Uz of z, M is Gz-equivariantly isometric to a ball, say of radius 1, about the origin in Euclidean spaceRn with an orthogonal Gz-action, with z correspond- ing to the origin. This can be arranged since the exponential map induces a

3Since sign conventions differ, we emphasize that for us, unit tangent vectors onM have square1 in the Clifford algebra.

(10)

Gz-diffeomorphism of a small Gz-invariant neighborhood of 0∈ TzM onto a Gz-invariant neighborhood of z such that 0 is mapped to z and its differential at 0 is the identity onTzM under the standard identificationT0(TzM) =TzM. Thus the usual coordinates x1, x2, . . . , xn in Euclidean space give local coor- dinates in M for |x| < 1, and ∂x

1, ∂x

2, . . . , ∂x

n define a local orthonormal frame in T M near z. We can arrange that (Rn)Gz contains the points with x2 =. . . =xn = 0 if (Rn)Gz is different from {0}. In these exponential local coordinates, the point x1 = x2 = · · · = xn = 0 corresponds to z. We may assume we have chosen the vector field Ξ so that in these local coordinates, Ξ is given by the radial vector field

x1

∂x1

+x2

∂x2

+· · ·+xn

∂xn

(1.2) if z∈Zero(Ξ)+, or by the vector field

−x1

∂x1 +x2

∂x2 +· · ·+xn

∂xn (1.3)

if z ∈ Zero(Ξ). Thus |Ξ(x)| = 1 on ∂Uz for each z, and we can assume (rescaling Ξ if necessary) that |Ξ| ≥ 1 on the complement of S

zZero(Ξ)Uz. Recall that Dλ =D+λA.

Lemma 1.4 Fix a small number ε > 0, and let Pλ denote the spectral pro- jection of D2λ corresponding to [0, ε]. Then for λ sufficiently large, rangePλ is G-isomorphic to L2(Zero(Ξ)) (a Hilbert space with Zero(Ξ) as orthonormal basis, with the obvious unitary action of G coming from the action of G on Zero(Ξ)), and there is a constant C > 0 such that (1−Pλ)D2λ ≥ Cλ. (In other words, (ε, Cλ)∩(specD2λ) =∅.) Furthermore, the functions in rangePλ become increasingly concentrated near Zero(Ξ) as λ→ ∞.

Proof First observe that in Euclidean space Rn, if Ξ is defined by (1.2) or (1.3) and A and Dλ are defined from Ξ as on M, then Sλ =D2λ is basically a Schr¨odinger operator for a harmonic oscillator, so one can compute its spectral decomposition explicitly. (For example, if n= 1, then Sλ =−dxd222x2±λ, the sign depending on whether z∈Zero(Ξ)+ or z∈Zero(Ξ) and whether one considers the action on H+ or H.) When z∈Zero(Ξ)+, the kernel of Sλ in L2 sections of Cliff(TRn) is spanned by the Gaussian function

(x1, x2, . . . , xn)7→e−λ|x|2/2,

and ifz∈Zero(Ξ), theL2 kernel is spanned by a similar section of Cliff(T M), e−λ|x|2/2∂x

1. Also, in both cases, Sλ has discrete spectrum lying on an arith- metic progression, with one-dimensional kernel (in L2) and first non-zero eigen- value given by 2nλ.

(11)

Now let’s go back to the operator on M. Just as in [23, Lemma 2], we have the estimate

−Kλ≤Dλ2−(D22A2)≤Kλ, (1.5) where K >0 is some constant (depending on the size of the covariant deriva- tives of Ξ).4 But Dλ2 ≥0, and also, from (1.5),

1

λ2Dλ2 ≥A2+ 1

λ2D2−K

λ, (1.6)

which implies that 1

λ2Dλ2 ≥multiplication by |Ξ(x)|2−K

λ. (1.7)

So if ξλ is a unit vector in rangePλ, we have ε

λ2 ≥ 1

λ2Dλ2ξλ, ξλ

≥ Z

M

|Ξ(x)|2λ(x)|2dvol−K λ

ξλ

2. (1.8) Now kξλk = 1, and if we fix η > 0, we only make the integral smaller by replacing |Ξ(x)|2 by η on the set Eη ={x :|Ξ(x)|2 ≥η} and by 0 elsewhere.

So ε

λ2 ≥ −K λ +η

Z

Eη

λ(x)|2dvol or

ξλχEη

2≤ K ηλ + ε

ηλ2. (1.9)

This being true for any η, we have verified that as λ → ∞, ξλ becomes in- creasingly concentrated near the zeros of Ξ, in the sense that the L2 norm of its restriction to the complement of any neighborhood of Zero(Ξ) goes to 0.

It remains to compute rangePλ (as a unitary representation space of G) and to prove that D2λ has the desired spectral gap. Define a C2 cut-off function ϕ(t), 0≤t <∞, so that 0≤ϕ(t)≤1, ϕ(t) = 1 for 0≤t≤ 12, ϕ(t) = 0 for t≥1, and ϕ is decreasing on the interval 1

2,1

. In other words, ϕ is supposed to have a graph like this:

4We are using the cocompactness of the G-action to obtain a uniform estimate.

(12)

We can arrange that |ϕ0(t)| ≤ 3 and that |ϕ00(t)| ≤ 20. For each element z of Zero(Ξ), recall that we have a Gz-invariant neighborhood Uz that can be identified with the unit ball in Rn equipped with an orthogonal Gz-action. So the function ψz,λ(x) = ϕ(r)e−λr2/2, where r = |x| is the radial coordinate in Rn, makes sense as a function in C2(M), with support in Uz. For simplicity suppose z∈Zero(Ξ)+; the other case is exactly analogous except that we need a 1-form instead of a function. Then D2λ, acting on radial functions, becomes

−∆ +λ2|x|2−nλ=− ∂2

∂r2 −(n−1)1 r

∂r+λ2r2−nλ.

As we mentioned before, this operator on Rn annihilates x 7→ e−λr2/2, so we have

kDλψz,λk2z,λk2 =

Dλ2ψz,λ, ψz,λz,λ, ψz,λi

= R1

0 ϕ(r) −rϕ00(r) + (1−n+ 2r2λ)ϕ0(r)

eλr2rn2dr R1

0 ϕ(r)2e−λr2rn−1dr

≤ R1

1/2(20r+ 6λ)e−λr2rn−2dr R1/2

0 e−λr2rn−1dr . (1.10)

The expression (1.10) goes to 0 faster than λk for any k ≥ 1, since the numerator dies rapidly and the denominator behaves like a constant timesλ−n/2 for large λ, so Pλψz,λ is non-zero and very close to ψz,λ. Rescaling constructs a unit vector in rangePλ concentrated near z, regardless of the value of ε, provided λ is sufficiently large (depending on ε). And the action of g ∈ G sends this unit vector to the corresponding unit vector concentrated near g·z. In particular, rangePλ contains a Hilbert space G-isomorphic to L2(Zero(Ξ)).

To complete the proof of the Lemma, it will suffice to show that if ξ is a unit vector in the domain ofD which is orthogonal to eachψz,λ, then kDλξk2 ≥Cλ for some constantC >0, provided λis sufficiently large LetE =S

z∈Zero(Ξ)Vz, where Vz corresponds to the ball about the origin of radius 12 when we identify Uz with the ball about the origin inRn of radius 1. LetχE be the characteristic function of E. Then

1 =kξk2 =kχEξk2+k(1−χE)ξk2. Hence we must be in one of the following two cases:

(a) k(1−χE)ξk212. (b) kχEξk212.

(13)

In case (1), we can argue just as in the inequalities (1.8) and (1.9) with η= 14, since E is precisely the set where |Ξ(x)|2 < 14. So we obtain

1

λ2kDλξk2 = 1

λ2Dλ2ξ, ξ

≥ −K λ +1

4k(1−χE)ξk2 ≥ 1 8 −K

λ,

which gives kDλξk2 const·λ2 once λ is sufficiently large. So now consider case (2). Then for some z, we must have kχG·Vzξk22|G\Zero(Ξ)|1 . But by assumption, ξ ⊥ψg·z,λ (for this same z and all g∈G). Assume for simplicity that ξ ∈ H+ and z ∈ Zero(Ξ)+. If ξ ∈ H+ and z ∈ Zero(Ξ), there is no essential difference, and if ξ ∈ H, the calculations are similar, but we need 1-forms in place of functions. Anyway, if we let ξg denote ξ|Ug·z transported to Rn, we have

0 = Z

Rn

ϕ(|x|)ξg(x)eλ|x|2/2dx (∀g), 1≥X

g

Z

|x|≤12

ϕ(|x|)2 ξg(x)

2dx≥ 1

2|G\Zero(Ξ)|.

Now we use the fact that the Schr¨odinger operator Sλ on Rn has one-dimen- sional kernel in L2 spanned by x7→eλ|x|2/2 (if z∈Zero(Ξ)+), and spectrum bounded below by 2nλ on the orthogonal complement of this kernel. (If z ∈ Zero(Ξ), the entire spectrum of Sλ on H+ is bounded below by 2nλ.) So compute as follows:

Dλ ϕ(|x|)ξg

2 =

Dλ2 ϕ(|x|)ξg

, ϕ(|x|)ξg

≥2nλhϕ(|x|)ξg, ϕ(|x|)ξgi. (1.11) Letω be the function on M which is 0 on the complement of S

gUg·z and given by ϕ(|x|) on Ug·z (when we use the local coordinate system there centered at g·z). Then:

kDλξk2=

Dλ ωξ

2+

Dλ 1−ω ξ

2

+ 2

D2λ 1−ω ξ

, ωξ

. (1.12)

Since Dλ is local and ω is supported on the Ug·z, g∈G, the first term on the right is simply

Dλ ωξ

2 =X

g

Dλ ϕ(|x|)ξg

2 ≥ 2nλ

2|G\Zero(Ξ)| (1.13) by (1.11). In the inner product term in (1.12), since ωξ is a sum of pieces with disjoint supports Ug·z, we can split this as a sum over terms we can transfer to

(14)

Rn, getting 2X

g

Sλ 1−ϕ(|x|) ξg

, ϕ(|x|)ξg

= 2X

g

D2 1−ϕ(|x|) ξg

, ϕ(|x|)ξg + 2X

g

Z

1 2≤|x|≤1

2|x|2+T λ)ϕ(|x|) 1−ϕ(|x|)

g(x)|2dx,

where−K ≤T ≤K. Sinceλ2|x|2+T λ >0 on 12 ≤ |x| ≤1 for large enough λ, the integral here is nonnegative, and the only possible negative contributions to kDλξk2 are the terms 2

D2 1−ϕ(|x|) ξg

, ϕ(|x|)ξg

, which do not grow with λ. So from (1.12), (1.13), and (1.11), kDλξk2≥const·λ for large enough λ, which completes the proof.

Proof of Theorem 0.5, continuedWe begin by defining a continuous fieldE of Z/2-graded Hilbert spaces over the closed interval [0,+∞]. Over the open interval [0,+∞), the field is just the trivial one, with fiber Eλ = H, the L2 sections of Cliff(T M). But the fiber E over +∞ will be the direct sum of H ⊕V, where V =L2(Zero(Ξ)) is a Hilbert space with orthonormal basis vz, z ∈ Zero(Ξ). We put a Z/2-grading on V by letting V+ = L2(Zero(Ξ)+), V =L2(Zero(Ξ)). To define the continuous field structure, it is enough by [7, Proposition 10.2.3] to define a suitable total set of continuous sections near the exceptional point λ = ∞. We will declare ordinary continuous functions [0,∞] → H to be continuous, but will also allow additional continuous fields that become increasingly concentrated near the points of Zero(Ξ). Namely, suppose z ∈ Zero(Ξ). By Lemma 1.4, for λ large, Dλ has an element ψz,λ in its “approximate kernel” increasingly supported close to z, and we have a formula for it. So we declare (ξ(λ))λ< to define a continuous field converging to cvz at λ=∞ if for any neighborhood U of z,

Z

MrU

|ξ(λ)(m)|2dvol(m)→0 asλ→ ∞, and if (assuming z∈Zero(Ξ)+) ξ(λ)∈ H+ and

ξ(λ)−c λ

π n4

ψz,λ

→0 asλ→ ∞. (1.14)

The constant reflects the fact that the L2-norm of e−λ|x|2/2 is πλn4

. If z ∈ Zero(Ξ), we use the same definition, but require ξ(λ)∈ H.

This concludes the definition of the continuous field of Hilbert spaces E, which we can think of as a HilbertC-module over C(I), I the interval [0,+∞]. We

(15)

will use this to define a Kasparov (C0(M), C(I))-bimodule, or in other words, a homotopy of Kasparov (C0(M),R)-modules. The action of C0(M) on E is the obvious one: C0(M) acts on H the usual way, and it acts on V (the other summand of E) by evaluation of functions at the points of Zero(Ξ):

f ·vz=f(z)vz, z∈Zero(Ξ), f ∈C0(M).

We define a field T of operators on E as follows. For λ < ∞, Tλ ∈ L(Eλ) = L(H) is simply Bλ as defined in (1.1), where recall that Dλ = D+λA. For λ=∞,E=H⊕V and T is 0 onV and is given on Hby the operatorB= A/|A| which is right Clifford multiplication by |Ξ(x)|Ξ(x) (an L, but possibly discontinuous, vector field) on H+ and by −Ξ(x)|Ξ(x)| on H. Note that T2 is 0 on V and the identity on H. While 1V is not compact onV if Zero(Ξ) is infinite, this is not a problem since for f ∈Cc(M), the action of f on V has finite rank (since f annihilates vz for z6∈suppf).

Now we check the axioms for (E, T) to define a homotopy of Kasparov modules from [D] to the class of

C0(M),E, T

= C0(M),H, B

⊕ C0(M), V, 0 . But C0(M),H, B

is a degenerate Kasparov module, since B commutes with multiplication by functions and has square 1. So the class of C0(M),E, T

is just the class of C0(M), V, 0

, which (essentially by definition) is the image under the inclusion Zero(Ξ) ,→ M of the sum (over G\Zero(Ξ)) of +1 times the canonical class KO0G(z)◦(αz)([R]) for G·z⊆Zero(Ξ)+ and of −1 times this class if G·z⊆Zero(Ξ). This will establish Theorem 0.5, assuming we can verify that we have a homotopy of Kasparov modules.

The first thing to check is that the action of C0(M) on E is continuous, i.e., given by a ∗-homomorphism C0(M) → L(E). The only issue is continuity at λ=∞. In other words, since the action onH is constant, we just need to know that if ξ is a continuous field converging as λ→ ∞ to a vector v in V, then for f ∈C0(M), f ·ξ(λ)→ f·v. But it’s enough to consider the special kinds of continuous fields discussed above, since they generate the structure, and if ξ(λ)→cvz, then ξ(λ) becomes increasingly concentrated at z (in the sense of L2 norm), and hence f·ξ(λ)→cf(z)vz, as required.

Next, we need to check thatT ∈ L(E). Again, the only issue is (strong operator) continuity at λ = ∞. Because of the way continuous fields are defined at λ=∞, there are basically two cases to check. First, ifξ ∈ H, we need to check that Bλξ →Bξ as λ→ ∞. Since the Bλ’s all have norm ≤1, we also only need to check this on a dense set of ξ’s. First, fix ε >0 small and suppose ξ

(16)

is smooth and supported on the open set where |Ξ(x)|2 > ε. Then for λ large, Lemma 1.4 implies that there is a constant C >0 (depending on ε but not on λ) such that hD2λξ, ξi> Cλkξk2. In fact, if Pλ is the spectral projection of Dλ2 for the interval [0, Cλ], Lemma 1.4 implies thatkPλξk ≤εkξk forλsufficiently large. (This is because the condition on the support of ξ forces ξ to be almost orthogonal to the spectral subspace where D2λ ≤ ε.) Now let Eλ+ and Eλ be the spectral projections for Dλ corresponding to the intervals (0,∞) and (−∞,0), and let F+ and F be the spectral projections for A corresponding to the same intervals. Since the vector field Ξ vanishes only on a discrete set, the operator A has no kernel, and hence F++F= 1. Now we appeal to two results in Chapter VIII of [14]: Corollary 1.6 in §1, and Theorem 1.15 in §2.

The former shows that the operatorsA+λ1D, all defined on domD, “converge strongly in the generalized sense” toA. Since the positive and negative spectral subspaces for A+λ1D are the same as for Dλ (since the operators only differ by a homothety), [14, Chapter VIII,§2, Theorem 1.15] then shows that Eλ+→F+ and Eλ→F+ in the strong operator topology. Note that the fact that A has no kernel is needed in these results.

Now since kPλξk ≤εkξk for λ sufficiently large, we also have Bξ=F+ξ−Fξ, and kBλξ−(Eλ+ξ−Eλξ)k ≤2ε for λ sufficiently large. Hence

kBλξ−Bξk ≤2ε+k(E+λξ−F+ξ)−(Eλξ−Fξ)k →2ε.

Now let ε → 0. Since, with ε tending to zero, ξ’s satisfying our support condition are dense, we have the required strong convergence.

There is one other case to check, that where ξ(λ) → cvz in the sense of the continuous field structure of E. In this case, we need to show that Bλξ(λ)→0.

This case is much easier: ξ(λ)→cvz means

ξ(λ)−c λ

π n

4

ψz,λ

→0 by (1.14), while kBλk ≤1 and

Dλ λ

π n

4

ψz,λ

!

→0 by (1.10), so Bλξ(λ)→0 in norm.

Thus T ∈ L(E). Obviously, T satisfies (B1) and (B2) of page 9, so we need to check the analogues of (B3), which are that f(1−T2) and [T, f] lie in K(E)

(17)

forf ∈C0(M). First consider 1−T2. 1−Tλ2 is locally compact (i.e., compact after multiplying by f ∈Cc(M)) for each λ, since

1−Tλ2 = 1−B2λ= (1 +D2λ)1 = 1 + (D+λA)2−1

is locally compact for λ < ∞, and 1−T2 is just projection onto V, where functions f of compact support act by finite-rank operators. So we just need to check that 1−Tλ2 is a norm-continuous field of operators on E. Continuity for λ <∞ is routine, and implicit in [3, Remarques 2.5]. To check continuity atλ=∞, we use Lemma 1.4, which shows that (1 +D2λ)1 =Pλ+O λ1

, and also that Pλ is increasingly concentrated near Zero(Ξ). So near λ = ∞, we can write the field of operators (1 +Dλ2)−1 as a sum of rank-one projections onto vector fields converging to the various vz’s (in the sense of our continuous field structure) and another locally compact operator converging in norm to 0.

This leaves just one more thing to check, that for f ∈ C0(M), [f, Tλ] lies in K(E). We already know that [f, Bλ]∈ K(H) for fixedλand is norm-continuous inλforλ <∞, so sinceT commutes with multiplication operators, it suffices to show that [f, Bλ] converges to 0 in norm as λ→0. We follow the method of proof in [23, p. 3473], pointing out the changes needed because of the zeros of the vector field Ξ.

We can take f ∈ Cc(M) with critical points at all of the points of the set Zero(Ξ), since such functions are dense in C0(M). Then estimate as follows:

[f, Bλ] = h

f, Dλ(1 +D2λ)−1/2 i

= [f, Dλ](1 +Dλ2)−1/2+Dλh

f, (1 +D2λ)−1/2i

. (1.15) We have [f, Dλ] = [f, D], which is a 0’th order operator determined by the derivatives of f, of compact support since f has compact support, and we’ve seen that (1 +D2λ)−1/2 converges as λ → ∞ (in the norm of our continuous field) to projection onto the space V = L2(Zero(Ξ)). Since the derivatives of f vanish on Zero(Ξ), the product [f, Dλ](1 +D2λ)−1/2, which is the first term in (1.15), goes to 0 in norm. As for the second term, we have (following [4, p.

199]) Dλ

h

f, (1 +D2λ)−1/2 i

= 1 π

Z

0

µ12Dλ

f, (1 +D2λ+µ)−1

dµ, (1.16) and

Dλ

f, (1 +D2λ+µ)−1

=Dλ(1 +D2λ+µ)−1

1 +Dλ2+µ, f

(1 +D2λ+µ)−1. Now use the fact that

1 +Dλ2+µ, f

= D2λ, f

=Dλ[Dλ, f] + [Dλ, f]Dλ =Dλ[D, f] + [D, f]Dλ.

(18)

We obtain that Dλ

f, (1 +D2λ+µ)−1

= Dλ2

1 +Dλ2+µ[D, f] 1

1 +D2λ+µ+ Dλ

1 +D2λ+µ[D, f] Dλ

1 +Dλ2+µ. (1.17) Again a slight modification of the argument in [23, p. 3473] is needed, since Dλ has an “approximate kernel” concentrated near the points of Zero(Ξ). So we estimate the norm of the right side of (1.17) as follows:

Dλ

f, (1 +Dλ2+µ)−1

D2λ

1 +D2λ+µ[D, f] 1 1 +D2λ

(1.18) +

Dλ

1 +D2λ+µ[D, f] Dλ 1 +Dλ2

. (1.19)

The first term, (1.18), is bounded by the second, (1.19), plus an additional commutator term:

Dλ

1 +D2λ+µ[D,[D, f]] 1 1 +D2λ

. (1.20)

Now the contribution of the term (1.19) is estimated by observing that the function

x

1 +x2+µ, −∞< x <∞ has maximum value 211+µ at x =√

1 +µ, is increasing for 0< x < √ 1 +µ, and is decreasing to 0 for x >√

1 +µ. Fix ε >0 small. Since, by Lemma 1.4,

|Dλ| has spectrum contained in [0,√ ε]∪[√

Cλ, ∞), we find that

Dλ 1 +Dλ2

















1 2

1+µ,

√Cλ≤√

1 +µ, orµ≥Cλ−1, max

ε 1+ε+µ,

1+µ+Cλ

√ ,

Cλ≥√

1 +µ, or 0≤µ≤Cλ−1.

(1.21)

(19)

Thus the contribution of the term (1.19) to the integral in (1.16) is bounded by k[D, f]k

π

Z

0

Dλ 1 +Dλ2

2 1

√µdµ

≤ k[D, f]k π

Z

Cλ−1

√1µ 1

4(1 +µ)dµ (1.22)

+

Z Cλ−1

0

√1µ max

(1 +µ+Cλ)2, ε (1 +ε+µ)2

≤ k[D, f]k π

1 4

Z

Cλ−1

µ32 dµ+ Z

0

√1µ Cλ (Cλ)2 dµ+

Z

0

√µ(1 +ε µ)2

= k[D, f]k π

1 2√

Cλ−1 + 2

√Cλ +πε 2

→ k[D, f]k

2 ε. (1.23)

We can make this as small as we like by taking ε small enough. Similarly, the contribution of term (1.20) to the integral in (1.16) is bounded by

k[D,[D, f]]k π

Z

0

Dλ 1 +Dλ2

1 1 +µ

√1µdµ

≤ k[D,[D, f]]k π

Z

1

1 2√

1 +µ 1 1 +µ

√1µdµ

+

Z Cλ−1

0

√Cλ (1 +µ+Cλ)

1 1 +µ

√1µdµ+ Z

0

√µ(1 +ε µ)2

≤ k[D,[D, f]]k π

Z

Cλ−1

1 2µ2 dµ+

Z

0

√1 Cλ

√µ(1 +1 µ)dµ

+ Z

0

√µ(1 +ε µ)2

≤ k[D,[D, f]]k π

1

2(Cλ−1)+ π

√Cλ+ πε 2

→ k[D,[D, f]]k

2 ε, (1.24)

which again can be taken as small as we like. This completes the proof.

2 Review of notions of equivariant Euler character- istic

Next we briefly review the universal equivariant Euler characteristic, as well as some other notions of equivariant Euler characteristic, so we can see exactly how they are related to the KOG-Euler class EulG(M). We will use the following notation in the sequel.

(20)

Notation 2.1 Let G be a discrete group and H ⊆ G be a subgroup. Let NH ={g ∈G |gHg1 =H} be its normalizer and let WH := NH/H be its Weyl group.

Denote by consub(G) the set of conjugacy classes (H) of subgroups H⊆G.

Let X be a G-CW-complex. Put

XH := {x∈X|H⊆Gx}; X>H := {x∈X|H(Gx}, where Gx is the isotropy group of x under the G-action.

Let x: G/H→X be a G-map. Let XH(x) be the component of XH contain- ing x(1H). Put

X>H(x) =XH(x)∩X>H.

Let WHx be the isotropy group of XH(x)∈π0(XH) under the WH-action.

Next we define the group UG(X), in which the universal equivariant Euler characteristic takes its values. Let Π0(G, X) be the component category of the G-space X in the sense of tom Dieck [6, I.10.3]. Objects are G-maps x: G/H → X. A morphism σ from x: G/H → X to y: G/K → X is a G-map σ: G/H → G/K such that y◦σ and x are G-homotopic. A G-map f: X→Y induces a functor Π0(G, f) : Π0(G, X)→Π0(G, Y) by composition with f. Denote by Is Π0(G, X) the set of isomorphism classes [x] of objects x: G/H→X in Π0(G, X). Define

UG(X) := Z[Is Π0(G, X)], (2.2) where for a set S we denote by Z[S] the free abelian group with basisS. Thus we obtain a covariant functor from the category of G-spaces to the category of abelian groups. Obviously UG(f) =UG(g) if f, g: X →Y are G-homotopic.

There is a natural bijection

Is Π0(G, X) −=→ a

(H)∈consub(G)

WH\π0(XH), (2.3) which sends x: G/H → X to the orbit under the WH-action on π0(XH) of the component XH(x) of XH which contains the point x(1H). It induces a natural isomorphism

UG(X) −→= M

(H)consub(G)

M

WH\π0(XH)

Z. (2.4)

(21)

Definition 2.5 Let X be a finite G-CW-complex X. We define theuniversal equivariant Euler characteristic of X

χG(X) ∈ UG(X)

by assigning to [x: G/H → X]∈Is Π0(G, X) the (ordinary) Euler character- istic of the pair of finite CW-complexes (WHx\XH(x), WHx\X>H(x)).

If the action of G on X is proper (so that the isotropy group of any open cell in X is finite), we define the orbifold Euler characteristic of X by:

χQG(X) := X

p0

X

G·eG\Ip(X)

|Ge|−1 ∈Q,

where Ip(X) is the set of open cells of X (after forgetting the group action).

The orbifold Euler characteristic χQG(X) can be identified with the more gen- eral notion of the L2-Euler characteristic χ(2)(X;N(G)), where N(G) is the group von Neumann algebra of G. One can compute χ(2)(X;N(G)) in terms of L2-homology

χ(2)(X;N(G)) = X

p≥0

(−1)p·dimN(G)

Hp(2)(X;N(G) ,

where dimN(G) denotes the von Neumann dimension (see for instance [17, Sec- tion 6.6]).

Next we define for a proper G-CW-complex X thecharacter map chG(X) : UG(X) → M

Is Π0(G,X)

Q = M

(H)∈consub(G)

M

WH\π0(XH)

Q. (2.6) We have to define for an isomorphism class [x] of objects x: G/H → X in Π0(G, X) the component chG(X)([x])[y] of chG(X)([x]) which belongs to an isomorphism class [y] of objects y: G/K → X in Π0(G, X), and check that χG(X)([x])[y] is different from zero for at most finitely many [y]. Denote by mor(y, x) the set of morphisms from y to x in Π0(G, X). We have the left operation

aut(y, y)×mor(y, x)→mor(y, x), (σ, τ)7→τ ◦σ1. There is an isomorphism of groups

WKy=→aut(y, y)

which sends gK ∈WKy to the automorphism of y given by the G-map Rg−1: G/K →G/K, g0K 7→g0g−1K.

(22)

Thus mor(y, x) becomes a left WKy-set.

The WKy-set mor(y, x) can be rewritten as

mor(y, x) = {g∈G/HK |g·x(1H)∈XK(y)},

where the left operation of WKy on {g ∈G/HK |g·x(1H) ∈YK(y)} comes from the canonical left action of G on G/H. Since H is finite and hence contains only finitely many subgroups, the set WK\(G/HK) is finite for each K ⊆G and is non-empty for only finitely many conjugacy classes (K) of sub- groups K ⊆G. This shows that mor(y, x) 6= ∅ for at most finitely many iso- morphism classes [y] of objects y∈Π0(G, X) and that the WKy-set mor(y, x) decomposes into finitely many WKy orbits with finite isotropy groups for each object y ∈Π0(G, X). We define

chG(X)([x])[y] := X

WKy·σ∈

WKy\mor(y,x)

|(WKy)σ|−1, (2.7)

where (WKy)σ is the isotropy group of σ ∈mor(y, x) under the WKy-action.

Lemma 2.8 Let X be a finite proper G-CW-complex. Then the map chG(X) of (2.6) is injective and satisfies

chG(X)(χG(X))[y] = χQWKy(XK(y)).

The induced map

idQZchG(X) : Q⊗ZUG(X) −→= M

(H)consub(G)

M

WH\π0(XH)

Q

is bijective.

Proof Injectivity of χG(X) and chG(X)(χG(X))[y] = χQWKy(XK(y)). is proved in [20, Lemma 5.3]. The bijectivity of idQZchG(X) follows since its source and its target are Q-vector spaces of the same finite Q-dimension.

Now let us briefly summarize the various notions of equivariant Euler charac- teristic and the relations among them. Since some of these are only defined when M is compact and G is finite, we temporarily make these assumptions for the rest of this section only.

Definition 2.9 If G is a finite group, the Burnside ring A(G) of G is the Grothendieck group of the (additive) monoid of finiteG-sets, where the addition comes from disjoint union. This becomes a ring under the obvious multiplication

Referenzen

ÄHNLICHE DOKUMENTE

These notes are based on a series of lectures given at the meeting Journ´ ees EDP in Roscoff in June 2015 on recent developments con- cerning weak solutions of the Euler equations

We note in passing that in the time-dependent case [8, 11, 1] has lead to solutions with H¨older regularity, a question that has been the focus of interest in view of Onsager’s

Theorem 4 implies that there can not be wild solutions on an annulus with smooth rotational initial data that are H¨ older continuous?. Indeed, any admissible H¨ older

By 9J ( we denote the set of all i-open polyhedra.. On the other hand we notice that there are rings 3 which admit no Euler characteristic, and others which admit more than one.

In this paper we prove a local in time existence result for the solutions of (1.1), with space periodic boundary conditions, under natural minimal regularity assumptions on the

Hereby, the directed length of a segment P Q will be denoted by P Q (of course, this directed length is only de…ned if the line through the points P and Q is directed, but we can

Jede Menge 6= {} von reellen Zahlen hat eine kleinste obere Schranke in IR ∪ {∞}. ∞ ergibt sich genau dann als kleinste obere Schranke, wenn die Menge nach oben unbeschr¨

Key words: hyperbolic systems, wave equation, evolution Galerkin schemes, Maxwell equa- tions, linearized Euler equations, divergence-free, vorticity, dispersion..