• 検索結果がありません。

The Symmetry of Intersection Numbers in Group Theory

N/A
N/A
Protected

Academic year: 2022

シェア "The Symmetry of Intersection Numbers in Group Theory"

Copied!
335
0
0

読み込み中.... (全文を見る)

全文

(1)

Geometry & Topology GGGG GG

GG G GGGGGG T T TTTTTTT TT

TT TT Volume 2 (1998) 1–10

Published: 16 January 1998

Einstein metrics and smooth structures

D Kotschick

Mathematisches Institut Universit¨at Basel

Rheinsprung 21 4051 Basel, Switzerland Email: dieter@math-lab.unibas.ch

Abstract

We prove that there are infinitely many pairs of homeomorphic non-diffeo- morphic smooth 4–manifolds, such that in each pair one manifold admits an Einstein metric and the other does not. We also show that there are closed 4–manifolds with two smooth structures which admit Einstein metrics with opposite signs of the scalar curvature.

AMS Classification numbers Primary: 57R55, 57R57, 53C25 Secondary: 14J29

Keywords: Einstein metric, smooth structure, four–manifold

Proposed: Peter Kronheimer Received: 8 September 1997

Seconded: Ronald Stern, Gang Tian Revised: 14 January 1998

(2)

In dimensions strictly smaller than four Einstein metrics have constant curva- ture and are therefore rare. In dimension four Einstein metrics of non-constant curvature exist, but it is still the case that existence of such a metric imposes non-trivial restrictions on the underlying manifold1. For closed orientable Ein- stein 4–manifolds X the Euler characteristic has to be non-negative, and, fur- thermore, the Hitchin–Thorpe inequality

e(X) 3

2|σ(X)| (1)

must hold, where e denotes the Euler characteristic and σ the signature. This condition is very crude, and is certainly homotopy invariant, as are the re- strictions coming from Gromov’s notion of simplicial volume [11], and from the existence of maps of non-zero degree to hyperbolic manifolds [21].

Our aim in this note is to discuss existence and non-existence of Einstein metrics as a property of the smooth structure. We shall exhibit infinitely many pairs of homeomorphic non-diffeomorphic smooth 4–manifolds, such that in each pair one manifold admits an Einstein metric and the other does not. This shows for the first time that the smooth structures of 4–manifolds form definite obstructions to the existence of an Einstein metric.

An isolated example of such a pair can be obtained as follows. Hitchin [12]

showed that Einstein manifolds for which (1) is an equality are either flat or quo- tients of aK3 surface with a Calabi–Yau metric. Thus, the existence of smooth manifolds homeomorphic but not diffeomorphic to the K3 surface, which is known from Donaldson theory [10] and also follows easily from Seiberg–Witten theory, see for example [7, 15], shows that by changing only the differentiable structure one can pass from a manifold with an Einstein metric to one with- out. The point of our examples is that there are lots of them, and they do not arise from the borderline case of a non-existence result. They are in some sense generic.

We shall also discuss a conjecture concerning uniqueness of Einstein metrics on 4–manifolds which complements the discussion of existence. This too depends on a consideration of different smooth structures on a fixed topological manifold.

1 Smooth structures as obstructions

We shall use Seiberg–Witten invariants to show that certain smooth structures obstruct the existence of Einstein metrics, and refer the reader to [24, 7, 15]

1No such restrictions are known in higher dimensions.

(3)

for the definitions and basic properties of the invariants. All manifolds in this section are closed, smooth, oriented 4–manifolds. For the sake of simplicity, we assume b+2 >1 throughout, though this is not essential.

We shall need the following result concerning the behaviour of the invariants under connected summing with CP2. This is usually referred to as a blowup formula.

Proposition 1.1 ([8, 16]) Let Pe(Y) be a Spinc–structure on Y, and X = Y#CP2, with E a generator of H2(CP2,Z). Then X has a Spinc–structure P(X)e withc1(Pe(X)) =c1(Pe(Y)) +E, such that the Seiberg–Witten invariants of Pe(Y) and of Pe(X) are equal (up to sign).

As the reflection in E in the cohomology of X is realised by a self-diffeo- morphism, the naturality of the invariants shows that there is another Spinc– structure with the same Seiberg–Witten invariant, up to sign, and with c1(Pe(X)) =c1(Pe(Y))−E.

Using this, we can prove the following version of a theorem of LeBrun [18]:

Theorem 1.2 Let Y be a manifold with a non-zero Seiberg–Witten invariant (of any degree), and X =Y#kCP2. If k > 23(2e(Y) + 3σ(Y)), then X does not admit an Einstein metric.

Proof If P(X) has a non-zero Seiberg–Witten invariant, then for every Rie-e mannian metric g there must be a solution (A, φ) of the monopole equations.

Denoting by ˆA the connection induced by the Spinc–connection A on the de- terminant bundle of the spinor bundle, we have

c21(Pe(X)) = 1 4π2

Z

X

(|F+ˆ

A|2− |Fˆ

A|2)dvolg 1 4π2

Z

X

|F+ˆ

A|2dvolg

= 1

32π2 Z

X

|φ|4dvolg 1 32π2

Z

X

s2gdvolg , where sg denotes the scalar curvature of g.

Given any class c H2(X,R), denote by c+ the projection of c into the subspace H2+ ⊂ H2 of g–self-dual harmonic forms along the subspace H2 of g–anti-self-dual harmonic forms. The argument above really proves

c1(Pe(X))+ 2

1 32π2

Z

X

s2gdvolg .

(4)

If Pe(Y) is a Spinc–structure on Y with non-zero Seiberg–Witten invariant, then, by Proposition 1.1 and the subsequent remark, there are Spinc–structures P(X) one X =Y#kCP2 with non-zero Seiberg–Witten invariants and with

c1(Pe(X)) =c1(Pe(Y)) + Xk

i=1

(1)iEi

for any choice of the signs (1)i. Choose the signs so that (1)iEi+·c1(P(Ye ))+0 . Then

1 32π2

Z

X

s2gdvolg

c1(Pe(X))+ 2

=

c1(Pe(Y))+ 2

+ 2 Xk i=1

(1)iEi+·c1(P(Ye ))++ Xk

i=1

(1)iEi+

!2

c1(Pe(Y))+ 2

≥c1(Pe(Y))2 2e(Y) + 3σ(Y)

= 2(e(X)−k) + 3(σ(X) +k) = 2e(X) + 3σ(X) +k , where we have used the inequality c1(Pe(Y))2 2e(Y) + 3σ(Y) which is equiv- alent to the assertion that the moduli space associated with P(Ye ) has non- negative dimension.

Thus, we have proved 32π12

R

Xs2gdvolg 2e(X) + 3σ(X) +k for every metric g on X.

Suppose now that g is Einstein. Then the Chern–Weil integrals for the Euler characteristic and the signature of X give

2e(X) + 3σ(X) = 1 4π2

Z

X

( 1

24s2g+ 2|W+|2)dvolg

1 96π2

Z

X

s2gdvolg

1

3(2e(X) + 3σ(X) +k) ,

where W+ denotes the self-dual part of the Weyl tensor of g. Therefore k 2(2e(X) + 3σ(X)) = 2(2e(Y) + 3σ(Y) −k), which implies k 23(2e(Y) + 3σ(Y)).

Theorem 1.2 was proved by LeBrun [18], who also discussed the borderline case k= 23(2e(Y) + 3σ(Y)), in the case where Y is complex or symplectic. In that

(5)

case the blown up manifoldX is also complex, respectively symplectic, so that Proposition 1.1 is not needed.

The following is the main result of this section, giving the examples mentioned in the introduction.

Theorem 1.3 There are infinitely many pairs (Xi, Zi) of simply connected closed oriented smooth 4–manifolds such that:

1) Xi is homeomorphic to Zi,

2) if i6=j, then Xi and Xj are not homotopy equivalent, 3) Zi admits an Einstein metric but Xi does not,

4) e(Xi)> 32|σ(Xi)|.

Note that 3) implies in particular that Xi and Zi are not diffeomorphic.

Proof We claim that there are simply connected minimal complex surfaces Yi, Zi of general type such that if we take Xi = Yi#kCP2, for a suitable k with k > 23(2e(Yi) + 3σ(Yi)), then the pairs (Xi, Zi) have all the desired properties. The last property, the strict Hitchin–Thorpe inequality, follows from the Noether and Miyaoka–Yau inequalities for Zi, which, by the first property, has the same Euler characteristic and signature as Xi.

If we take Zi to have ample canonical bundle, then the results of Aubin and Yau on the Calabi conjecture show that Zi admits a K¨ahler–Einstein metric, compare [3]. On the other hand, Xi does not admit any Einstein metric by Theorem 1.2.

The crucial issue then is to arrange that Zi, with ample canonical bundle, is homeomorphic to thek–fold blowup of Yi, with k > 23(2e(Yi) + 3σ(Yi)). As Xi

will be automatically non-spin,Xi andZi will be homeomorphic by Freedman’s classification [9] as soon as Zi is non-spin and has the same Euler characteristic and the same signature as Xi. One can find suitable surfaces using the known results on the geography of surfaces of general type, see [14] for a summary of the results.

To exhibit concrete examples, instead of working with the topological Euler characteristic and the signature, we shall use the first Chern number c21 = 2e+ 3σ and the Euler characteristic of the structure sheaf χ= 14(e+σ).

Under blowing up, c21 drops by one and χ is constant. Thus, the Miyaoka–Yau inequality for Yi implies c21(Xi)<3χ(Xi). In fact, c21(Xi) will be smaller still,

(6)

because simply connected surfaces Yi are not known to exist if we get too close to the Miyaoka–Yau line c21= 9χ.

The minimal surfaceZi satisfies the same inequalities on its characteristic num- bers as Xi. If the canonical bundle of Zi is very ample, then Castelnuovo’s theorem, see [2] page 228, gives c21(Zi)3χ(Zi)10, which will contradict the above upper bound for c21(Xi). Thus, Zi must be chosen to have ample but not very ample canonical bundle, and will be in the sector where

2χ(Zi)6≤c21(Zi)<3χ(Zi) ,

the first being the Noether inequality. The results of Xiao Gang and Z Chen, cf [14], show that all non-spin simply connected surfaces Zi with ample canon- ical bundle which are in this sector, and not too close to the line c21 = 3χ, will have companions Xi as required, obtained by blowing up minimal surfaces Yi. Note that by Beauville’s theorem on the canonical map, cf [2] page 228, all the Zi will be double covers of ruled surfaces.

We can avoid using the results of Xiao and Chen by taking for Zi the following family of Horikawa surfaces, cf [2]. Let Σi be the Hirzebruch surface whose section at infinityS has self-intersection −i, and let Zi be a double cover of Σi branched in a smooth curve homologous to B = 6S+ 2(2i+ 3)F, where F is the class of the fiber. The double cover is simply connected as B is ample, and KZi =π(KΣi+12B) =π(S+ (i+ 1)F) is not 2–divisible and soZi is not spin.

Moreover, KZi is the pullback of an ample line bundle and therefore ample, so that Zi admits a K¨ahler–Einstein metric. The characteristic numbers of Zi are c21(Zi) = 2i+ 4 and χ(Zi) =i+ 5.

Now, by the classical geography results of Persson [19], for all i large enough there are simply connected surfaces Yi of general type with c21(Yi) = 6i+ 13 and χ(Yi) =i+ 5, so that the (4i+ 9)–fold blowup Xi of Yi is homeomorphic to Zi.

The pairs (Xi, Zi) have all the desired properties.

Remark 1.4 The examples of manifolds without Einstein metrics given by LeBrun [18], namely blowups of hypersurfaces inCP3, cannot be used to prove Theorem 1.3 because they violate the Noether inequality. They are therefore not homeomorphic to minimal surfaces for which the resolution of the Calabi conjecture gives existence of an Einstein metric.

In view of Theorem 1.3, one can ask how many smooth structures with Einstein metrics and how many without, a given topological manifold has. On the

(7)

one hand, using for example the work of Fintushel–Stern, one can show that one has infinitely many choices for the smooth structures of the manifolds Yi in the proof of Theorem 1.3, which remain distinct under blowing up points.

Thus, one has infinitely many smooth manifolds one can use for each Xi, not admitting any Einstein metrics. On the other hand, it is known that there are homeomorphic non-diffeomorphic minimal surfaces of general type, cf [10], page 410, and the references cited there. In fact, the number of distinct smooth structures among sets of homeomorphic minimal surfaces of general type can be arbitrarily large [20]. It is not hard to check that all the examples in [20] and [10]

have ample canonical bundle, and therefore have K¨ahler–Einstein metrics of negative scalar curvature. However, all those examples have c21 > 3χ, and can therefore not be used as the Zi in the proof of Theorem 1.3. Still, those examples show that a given simply connected topological manifold can have an arbitrarily large number of smooth structures admitting Einstein metrics.

Compare Theorem 2.2 below.

2 Uniqueness for a given smooth structure

We have seen that existence of Einstein metrics on closed 4–manifolds depends in an essential way on the smooth structure. I believe that the issue of unique- ness, up to the sign of the scalar curvature, is also tied to the smooth structure.

More specifically:

Conjecture 2.1 A closed smooth 4–manifold admits Einstein metrics for at most one sign of the scalar curvature.

Such questions were raised in [3], pages 18–19, and are also addressed in [4].

What is new here, and in [4], is that the answer depends on the smooth struc- ture, and also seems to depend on the dimension. The conjecture is interesting because it is sharp — it would be false if one did not fix the smooth structure, but only the underlying topological manifold:

Theorem 2.2 There are simply connected homeomorphic but non-diffeomor- phic smooth 4–manifolds X and Y, such that X admits an Einstein metric of positive scalar curvature, and Y admits an Einstein metric of negative scalar curvature.

Proof We can take for X the 8–fold blowup of CP2. By the work of Tian–

Yau [22] this admits a K¨ahler–Einstein metric of positive scalar curvature.

(8)

ForY we take the simply connected numerical Godeaux surface constructed by Craighero–Gattazzo [5] and studied recently by Dolgachev–Werner [6]2. This has ample canonical bundle, so that by the work of Aubin and Yau it admits a K¨ahler–Einstein metric of negative scalar curvature.

By Freedman’s classification [9], X and Y are homeomorphic. That they are not diffeomorphic is clear from [13]. The argument carried out there for the Barlow surface, cf [1], works even more easily for the Craighero–Gattazzo sur- face as there is no complication arising from (2)–curves. Alternatively, the fact that X and Y have K¨ahler–Einstein metrics of opposite signs implies via Seiberg–Witten theory that they are non-diffeomorphic.

In higher dimensions, homeomorphic non-diffeomorphic manifolds with Ein- stein metrics are known [17, 23], though in those examples all the metrics have positive scalar curvature. Although the examples of [17, 23] are consistent with a higher-dimensional analogue of the above conjecture, such a generalisation is false:

Corollary 2.3 For everyi≥2there is a simply connected closed 4i–manifold which admits Einstein metrics of both positive and negative scalar curvature.

Proof Let X and Y be as in Theorem 2.2. Then the i–fold products Xi = X×. . .×X and Yi =Y ×. . .×Y have K¨ahler–Einstein metrics of positive, respectively negative, scalar curvature. However, as X and Y are simply con- nected and homeomorphic, they are h–cobordant. Therefore, Xi and Yi are also h–cobordant for all i, and for i≥2 are diffeomorphic by the h–cobordism theorem.

Theorem 2.2 and Corollary 2.3 have also been proved independently by Catanese and LeBrun [4]. Instead of the Craighero–Gattazzo surface they use the Bar- low surface, showing that it has deformations with ample canonical bundle.

Conjecturally, the Barlow and Craighero–Gattazzo surfaces are deformation equivalent, and therefore diffeomorphic.

Acknowledgement: I am grateful to the Department of Mathematics at Brown University and to the Max–Planck–Institut f¨ur Mathematik in Bonn for hospi- tality and support.

2I am grateful to Igor Dolgachev for telling me about this, and for providing an advance copy of [6].

(9)

References

[1] R Barlow, A simply connected surface of general type with pg = 0, In- vent. Math. 79 (1985) 293–301

[2] W Barth, C Peters,A Van de Ven,Compact Complex Surfaces, Springer–

Verlag (1984)

[3] A L Besse,Einstein Manifolds, Springer–Verlag (1987)

[4] F Catanese, C LeBrun, On the scalar curvature of Einstein manifolds, preprint (May 1997)

[5] P Craighero, R Gattazzo, Quintic surfaces of CP3 having a nonsingular model with q = pg = 0, P2 6= 0, Rend. Sem. Mat. Univ. Padova 91 (1994) 187–198

[6] I Dolgachev,C Werner,A simply connected numerical Godeaux surface with ample canonical class, preprint (April 1997)

[7] S K Donaldson, The Seiberg–Witten equations and 4–manifold topology, Bull. Amer. Math. Soc. 33 (1996) 45–70

[8] R Fintushel,R J Stern,Immersed spheres in 4–manifolds and the immersed Thom conjecture, Proc. of G¨okova Geometry–Topology Conference 1994, Turk- ish J. of Math. 19 (2) (1995) 27–39

[9] M H Freedman, The topology of four–manifolds, J. Differential Geometry 17 (1982) 357–454

[10] R Friedman, J W Morgan,Smooth Four–Manifolds and Complex Surfaces, Springer–Verlag (1994)

[11] M Gromov,Volume and bounded cohomology, Publ. Math. I.H.E.S. 56 (1982) 5–99

[12] N J Hitchin,Compact four–dimensional Einstein manifolds, J. Differential Ge- ometry 9 (1974) 435–441

[13] D Kotschick, On manifolds homeomorphic to CP2#8CP2, Invent. math. 95 (1989) 591–600

[14] D Kotschick, Orientation-reversing homeomorphisms in surface geography, Math. Annalen 292 (1992) 375–381

[15] D Kotschick, P B Kronheimer,T S Mrowka, monograph in preparation [16] D Kotschick,J W Morgan,C H Taubes, Four–manifolds without symplec-

tic structures but with non-trivial Seiberg–Witten invariants, Math. Research Letters 2 (1995) 119–124

[17] M Kreck,S Stolz,A diffeomorphism classification of 7–dimensional homoge- neous Einstein manifolds with SU(3)×SU(2)×U(1) symmetry, Ann. of Math.

127 (1988) 373–388

[18] C LeBrun, Four–manifolds without Einstein metrics, Math. Research Letters 3 (1996) 133–147

(10)

[19] U Persson, Chern invariants of surfaces of general type, Comp. Math. 43 (1981) 3–58

[20] M Salvetti, On the number of non-equivalent differentiable structures on 4 manifolds, Manuscr. Math. 63 (1989) 157–171

[21] A Sambusetti, An obstruction to the existence of Einstein metrics on 4 manifolds, C. R. Acad. Sci. Paris 322 (1996) 1213–1218

[22] G Tian, S-T Yau, ahler–Einstein metrics on complex surfaces with c1>0, Comm. Math. Phys. 112 (1987) 175–203

[23] McK Y Wang,W Ziller,Einstein metrics on principal torus bundles, J. Dif- ferential Geometry 31 (1990) 215–248

[24] E Witten, Monopoles and four–manifolds, Math. Research Letters 1 (1994) 769–796

(11)

Geometry & Topology GGGG GG

GG G GGGGGG T T TTTTTTT TT

TT TT Volume 2 (1998) 11–29

Published: 19 March 1998

The Symmetry of Intersection Numbers in Group Theory

Peter Scott

Mathematics Department University of Michigan Ann Arbor, MI 48109, USA Email: pscott@math.lsa.umich.edu

Abstract

For suitable subgroups of a finitely generated group, we define the intersection number of one subgroup with another subgroup and show that this number is symmetric. We also give an interpretation of this number.

AMS Classification numbers Primary: 20F32 Secondary: 20E06, 20E07, 20E08, 57M07

Keywords: Ends, amalgamated free products, trees

Proposed: Jean-Pierre Otal Received: 21 February 1997

Seconded: Cameron Gordon, Walter Neumann Revised: 13 March 1998

(12)

If one considers two simple closed curvesL and S on a closed orientable surface F, one can define their intersection number to be the least number of intersec- tion points obtainable by isotoping L and S transverse to each other. (Note that the count is to be made without any signs attached to the intersection points.) By definition, this number is symmetric, ie the roles of L and S are interchangeable. This can be regarded as a definition of the intersection num- ber of the two infinite cyclic subgroups Λ and Σ of the fundamental group of F which are carried by L and S. In this paper, we show that an analogous definition of intersection number of subgroups of a group can be given in much greater generality and proved to be symmetric. We also give an interpretation of these intersection numbers.

In [7], Rips and Sela considered a torsion free finitely presented group G and infinite cyclic subgroups Λ and Σ such that G splits over each. (A group G splits over a subgroup C if either G has a HNN decomposition G=A∗C, orG has an amalgamated free product structure G=A∗C B, where A 6=C 6=B.) They effectively considered the intersection number i(Λ,Σ) of Λ with Σ, and they proved that i(Λ,Σ) = 0 if and only if i(Σ,Λ) = 0. Using this, they proved that G has what they call a JSJ decomposition. If i(Λ,Σ) was not zero, it follows from their work that G can be expressed as the fundamental group of a graph of groups with some vertex group being a surface group H which contains Λ and Σ. Now it is intuitively clear (and we discuss it further at the end of section 2 of this paper) that the intersection number of Λ with Σ is the same whether it is measured in G or in H. Also the intersection numbers of Λ and Σ in H are symmetric because of their topological interpretation. So it follows at the end of all their work that the intersection numbers of Λ and Σ in G are also symmetric. In 1994, Rips asked if there was a simpler proof of this symmetry which does not depend on their proof of the JSJ splitting.

The answer is positive, and the ideas needed for the proof are all essentially contained in earlier papers of the author. This paper is a belated response to Rips’ question. The main idea is to reduce the natural, but not clearly symmetric, definition of intersection number to counting the intersections of suitably chosen sets. The most general possible algebraic situation in which to define intersection numbers seems to be that of a finitely generated groupGand two finitely generated subgroups Λ and Σ, not necessarily cyclic, such that the number of ends of each of the pairs (G,Λ) and (G,Σ) is more than one. Note that any infinite cyclic subgroup Λ of π1(F) satisfies e(π1(F),Λ) = 2. This is because F is closed and orientable so that the cover of F with fundamental group Λ is an open annulus which has two ends. In order to handle the general situation, we will need the concept of an almost invariant set, which is closely

(13)

related to the theory of ends. We should note that Kropholler and Roller [6]

introduced an intersection cohomology class in the special case of P D(n−1)–

subgroups of P Dn–groups. Their ideas are closely related to ours, and we will discuss the connections at the start of the last section of this paper. Finally, we should point out that since Rips asked the above question about symmetry of intersection numbers, Dunwoody and Sageev [2] have given a proof of the existence of a JSJ decomposition for any finitely presented group which is very much simpler and more elementary than that of Rips and Sela.

The preceding discussion is a little misleading, as the intersection numbers which we define are not determined simply by a choice of subgroups. In fact, we define intersection numbers for almost invariant sets. A special case occurs when one has a groupG and subgroups Λ and Σ such that G splits over each, as a splitting of G has a well defined almost invariant set associated. This is discussed in section 2. Thus we can define the intersection number of two splittings of G. In the case of cyclic subgroups of surface groups corresponding to simple closed curves, these curves determine splittings of the surface group over each cyclic subgroup, and the intersection number we define for these splittings is the same as the topological intersection number of the curves.

In the first section of this paper, we discuss in more detail intersection numbers of closed curves on surfaces. In the second section we introduce the concept of an almost invariant set and prove the symmetry results advertised in the title.

In the third section, we discuss the interpretation of intersection numbers when they are defined, and how our ideas are connected with those of Kropholler and Roller.

Acknowledgments This paper was written while the author was visiting the Mathematical Sciences Research Institute in Berkeley in 1996/7. Research at MSRI is supported in part by NSF grant DMS-9022140. He is also grateful for the partial support provided by NSF grants DMS-9306240 and DMS-9626537.

1 The symmetry for surface groups

In this section, we will discuss further the special case of two essential closed curves L and S on a compact surface F. This will serve to motivate the defi- nitions in the following section, and also show that the results of that section do indeed answer the question of Rips. It is not necessary to assume that F is closed or orientable, but we do need to assume thatLand S are two-sided on F.

As described in the introduction in the case of simple curves, one defines their

(14)

intersection number to be the least number of intersection points obtainable by homotoping L and S transverse to each other, where the count is to be made without any signs attached to the intersection points. (One should also insist that L and S be in general position, in order to make the count correctly.) Of course, this number is symmetric, ie the roles of L and S are interchangeable.

We will show in section 2 that one can define these intersection numbers in an algebraically natural way. There is also an idea of self-intersection number for a curve on a surface and we will discuss a corresponding algebraic idea.

For the next discussion, we will restrict our attention to the case whenL and S are simple and introduce the algebraic approach to defining intersection num- bers taken by Rips and Sela in [7]. Let G denote π1(F). Suppose that L and S cannot be made disjoint and choose a basepoint on L∩S. Suppose that L represents the elementλ of G. This element λ cannot be trivial, nor can L be parallel to a boundary component of F, because of our assumption that L and S cannot be made disjoint. Thus L induces a splitting of G over the infinite cyclic subgroup Λ of Gwhich is generated by λ.Let σ denote the element of G represented by S.Defined(σ, λ) to be the length of σ when written as a word in cyclically reduced form in the splitting of G determined by L. Similarly, define d(λ, σ) to be the length of λ when written as a word in cyclically reduced form in the splitting of G determined by S. For convenience, suppose also that L and S are separating. Then each of these numbers is equal to the intersection number of L and S described above and therefore d(λ, σ) = d(σ, λ). What is interesting is that this symmetry is not obvious from the purely algebraic point of view, but it is obvious topologically because the intersection of two sets is symmetric.

In the above discussion, we restricted attention to simple closed curves on a surface F, because the algebraic analogue is clear. IfF is closed, then not only does a simple closed curve on F determine a splitting of π1(F) over the infinite cyclic subgroup carried by the curve, but any splitting ofπ1(F) over an infinite cyclic subgroup is induced in this way by some simple closed curve on F.Hence the algebraic situation described above exactly corresponds to the topological situation when F is closed.

Now we continue with further discussion of the intersection number of two closed curves L and S which need not be simple. As in [3], it will be convenient to assume that L and S are shortest closed geodesics in some Riemannian metric on F so that they automatically intersect minimally. Instead of defining the intersection number of L and S in the “obvious” way, we will interpret our intersection numbers in suitable covers of F, exactly as in [3] and [4]. Let FΛ denote the cover of F with fundamental group equal to Λ. Then L lifts to FΛ

(15)

and we denote its lift by L again. Let l denote the pre-image of this lift in the universal cover Fe of F. The full pre-image of L in Fe consists of disjoint lines which we callL–lines, which are all translates ofl by the action ofG. Similarly, we define FΣ, the line s and S–lines in F .e Now we consider the images of the L–lines in FΣ. Each L–line has image in FΣ which is a possibly singular line or circle. Then we define d(L, S) to be the number of images of L–lines in FΣ which meet S. Similarly, we define d(S, L) to be the number of images of S–lines in FΛ which meet L. It is shown in [3], using the assumption that L and S are shortest closed geodesics, that each L–line in FΣ crosses S at most once, and similarly for S–lines in FΛ. It follows that d(L, S) and d(S, L) are each equal to the number of points of L∩S, and so they are equal to each other. (This assumes that L and S are in general position.)

Here is an argument which shows that d(L, S) and d(S, L) are equal without reference to the situation in the surfaceF.Recall that theL–lines are translates of l by elements ofG. Of course, there is not a unique element of Gwhich sends l to a given L–line. In fact, the L–lines are in natural bijective correspondence with the cosets gΛ of Λ in G. (Our groups act on the left on covering spaces.) The images of the L–lines in FΣ are in natural bijective correspondence with the double cosets ΣgΛ, and d(L, S) counts the number of these double cosets such that the line gl crosses s. Similarly, d(S, L) counts the number of the double cosets ΛhΣ such that the line hs crosses l. Note that it is trivial that gl crosses s if and only if l crosses g 1s. Now we use the bijection from G to itself given by sending each element to its inverse. This induces a bijection between the set of all double cosets ΣgΛ and the set of all double cosets ΛhΣ by sending ΣgΛ to Λg1Σ. It follows that it also induces a bijection between those double cosets ΣgΛ such that gl crosses s and those double cosets ΛhΣ such that hs crosses l, which shows that d(L, S) equals d(S, L) as required.

This argument has more point when one applies it to a more complicated situ- ation than that of curves on surfaces. In [4], we considered least area maps of surfaces into a 3–manifold. The intersection number which we used there was defined in essentially the same way but it had no obvious topological interpre- tation such as the number of double curves of intersection. We proved that our intersection numbers were symmetric by the above double coset argument, in [4] just before Theorem 6.3.

(16)

2 Intersection Numbers in General

In order to handle the general case, we will need the idea of an almost invari- ant set. This idea was introduced by Cohen in [1] and was first used in the relative context by Houghton in [5]. We will introduce this idea and explain its connection with the foregoing.

LetE and F be sets. We say that E and F are almost equal, and write E=a F, if the symmetric difference (E−F)(F −E) is finite. If E is contained in some set W on which a group G acts on the right, we say that E is almost invariant ifEg=a E, for all g in G. An almost invariant subset E of W will be called non-trivial if it is infinite and has infinite complement. The connection of this idea with the theory of ends of groups is via the Cayley graph Γ of Gwith respect to some finite generating set of G. (Note that in this paper groups act on the left on covering spaces and, in particular, G acts on its Cayley graph on the left.) Using Z2 as coefficients, we can identify 0–cochains and 1–cochains on Γ with sets of vertices or edges. A subsetE of G represents a set of vertices of Γ which we also denote by E, and it is a beautiful fact, due to Cohen [1], that E is an almost invariant subset of G if and only if δE is finite, where δ is the coboundary operator. If H is a subgroup of G, we let H\G denote the set of cosets Hg of H in G, ie the quotient of G by the left action of H. Of course, G will no longer act on the left on this quotient, but it will still act on the right. Thus we have the idea of an almost invariant subset of H\G.

Now we again consider the situation of simple closed curves L and S on a compact surface F and letFe denote the universal cover of F.Pick a generating set for G which can be represented by a bouquet of circles embedded in F.We will assume that the wedge point of the bouquet does not lie on L or S. The pre-image of this bouquet in Fe will be a copy of the Cayley graph Γ ofG with respect to the chosen generating set. The pre-image in FΛ of the bouquet will be a copy of the graph Λ\Γ, the quotient of Γ by the action of Λ on the left.

Consider the closed curve L on FΛ.Let D denote the set of all vertices of Λ\Γ which lie on one side ofL. Then D has finite coboundary, as δD equals exactly the edges of Λ\Γ which cross L.Hence D is an almost invariant subset of Λ\G.

Let X denote the pre-image of D in Γ, so that X equals the set of vertices of Γ which lie on one side of the line l. There is an algebraic description of X in terms of canonical forms for elements of G as follows. Suppose that L separates F,so thatG=A∗ΛB.Also suppose thatL and Dare chosen so that all the vertices of Γ labelled with an element of Λ do not lie in X. Pick right transversals T and T0 for Λ in A and B respectively, both of which contain the identity e of G. (A right transversal of Λ in A consists of a choice of coset

(17)

representative for each coset aΛ.) Each element of G can be expressed uniquely in the form a1b1. . . anbnλ, where n 1, λ lies in Λ, each ai lies in T − {e} except that a1 may be trivial, and each bi lies in T0− {e} except that bn may be trivial. Then X consists of those elements for which a1 is non-trivial. If Λ is non-separating in F,there is a similar description for X. See Theorem 1.7 of [11] for details. Similarly, we can define a set E in FΣ and its pre-image Y in Fe which equals the set of vertices of Γ which lie on one side of the line s. Now finally the connection between the earlier arguments and almost invariant sets can be given. For we can decide whether the lines l and s cross by considering instead the sets X and Y. The lines l and s together divide G into the four sets X∩Y, X∩Y, X ∩Y and X ∩Y, where X denotes G−X, and l crosses s if and only if each of these four sets projects to an infinite subset of Σ\G. Equally, s crosses l if and only if each of these four sets projects to an infinite subset of Λ\G. As we know that l crosses sif and only if s crosses l, it follows that these conditions are equivalent. We will show that this symmetry holds in a far more general context.

Note that in the preceding example the subset X of G is Λ–invariant under the left action of Λ on G, ie λX =X, for all λ in Λ.

For the most general version of this symmetry result, we can consider any finitely generated group G. Note that the subgroups of G which we consider need not be finitely generated.

Definition 2.1 If G is a finitely generated group and H is a subgroup, then a subset X of G is H–almost invariant if X is invariant under the left action of H, and simultaneously the quotient set H\X is almost invariant under the right action of G. In addition, X is a non-trivial H–almost invariant subset of G if H\X and H\X are both infinite.

Note that if X is a non-trivial H–almost invariant subset of G, then e(G, H) is at least 2, as H\X is a non-trivial almost invariant subset of H\G.

Definition 2.2 Let X be a Λ–almost invariant subset of G and let Y be a Σ–almost invariant subset of G. We will say that X crosses Y if each of the four sets X∩Y, X∩Y, X∩Y and X∩Y projects to an infinite subset of Σ\G.

Note that it is obvious that if Y is trivial, then X cannot cross Y. Our first and most basic symmetry result is the following. This is essentially proved in Lemma 2.3 of [9], but the context there is less general.

(18)

Lemma 2.3 If G is a finitely generated group with subgroups Λ and Σ, and X is a non-trivial Λ–almost invariant subset of G and Y is a non-trivial Σ–

almost invariant subset of G, then X crosses Y if and only if Y crosses X.

Remark 2.4 If X and Y are both trivial, then neither can cross the other, so the above symmetry result is clear. However, this symmetry result fails if only one of X or Y is trivial. Here is a simple example. Let Λ and Σ denote infinite cyclic groups with generators λ and σ respectively, and let G denote the group Λ×Σ. We identify G with the set of integer points in the plane.

Let X = {(m, n) G: n > 0}, and let Y ={(m, n) G: m = 0}. Then X is a non-trivial Λ–almost invariant subset of G and Y is a trivial Σ–almost invariant subset of G. One can easily check that Y crosses X, although X cannot cross Y as Y is trivial.

Proof Suppose that X does not cross Y. By replacing one or both of X and Y by its complement if needed, we can assume that X∩Y projects to a finite subset of Σ\G. The fact that Y is non-trivial implies that Σ\Y is an infinite subset of Σ\G, so there is a point z in Σ\Y which is not in the image of X∩Y. Now we need to use some choice of generators for G and consider the corresponding Cayley graph Γ of G. The vertices of Γ are identified with G and the action of Gon itself on the left extends to an action on Γ.We consider z and the image of X ∩Y in the quotient graph Σ\Γ. As X ∩Y has finite image, there is a number d such that each point of its image can be joined to z by a path of length at most d. As the projection of Γ to Σ\Γ is a covering map, it follows that each point of X∩Y can be joined to some point lying above z by a path of length at most d. As any point above z lies in X, it follows that each point of X∩Y can be joined to some point of X by a path of length at most d. Hence each point of X∩Y lies at most distance d from δX. Thus the image of X∩Y in Λ\Γ lies within the d–neighbourhood of the compact set δ(Λ\X), and so must itself be finite. It follows that Y does not crossX, which completes the proof of the symmetry result.

At the start of this section, we explained how to connect the topological in- tersection of simple closed curves on a surface with crossing of sets. One can construct many other interesting examples in much the same way.

Example 2.5 As before, letF denote a closed surface with fundamental group G, and let Fe denote the universal cover of F.Pick a generating set of G which can be represented by a bouquet of circles embedded in F, so that Fe contains a copy of the Cayley graph Γ of G with respect to the chosen generators. Let

(19)

F1 denote a cover of F which is homeomorphic to a four punctured torus and let Λ denote its fundamental group. For example, if F is the closed orientable surface of genus three, we can consider a compact subsurface F0 of F which is homeomorphic to a torus with four open discs removed, and take the coverF1of F such that π1(F1) =π1(F0). For notational convenience, we identify F1 with S1×S1 with the four points (1,1),(1, i),(1,−1) and (1,−i) removed. Now we choose 1–dimensional submanifolds of F1 each consisting of two circles and each separating F1 into two pieces. Let L denote S1× {eπi/4, e5πi/4} and let S denote S1× {e3πi/4, e7πi/4}. As before, we let D denote all the vertices of the graph Λ\Γ in F1 which lie on one side of L, and let E denote all the vertices of the graph Λ\Γ in F1 which lie on one side of S. Let X and Y denote the pre-images of D and E in G. Now D is an almost invariant subset of Λ\G, as δD equals exactly the edges of Λ\Γ which cross L, and E is almost invariant for similar reasons. Hence X and Y are each Λ–almost invariant subsets of G.

Clearly X and Y cross. An important feature of this example is that although X and Y cross, the boundaries L and S of the corresponding surfaces in F1 are disjoint. This is quite different from the example with which we introduced almost invariant sets, but this is a much more typical situation.

Definition 2.6 Let Λ and Σ be subgroups of a finitely generated group G.

Let D denote a non-trivial almost invariant subset of Λ\G, let E denote a non-trivial almost invariant subset of Σ\G and let X and Y denote the pre- images in G of D and E respectively. We define i(D, E) to equal the number of double cosets ΣgΛ such that gX crosses Y.

For this definition to be interesting, we need to show that i(D, E) is finite, which is not obvious from the definition in this general situation. In fact, it may well be false if one does not assume that the groups Λ and Σ are finitely generated, although we have no examples. From now on, we will assume that Λ and Σ are finitely generated.

Lemma 2.7 Let Λ and Σ be finitely generated subgroups of a finitely gen- erated group G. Let D denote a non-trivial almost invariant subset of Λ\G, and let E denote a non-trivial almost invariant subset of Σ\G. Then i(D, E) is finite.

Proof This is again proved by using the Cayley graph, so it appears to depend on the fact that G is finitely generated. However, we have no examples where i(D, E) is not finite when G is not finitely generated. The proof we give is essentially contained in that of Lemmas 4.3 and 4.4 of [8]. Start by considering

(20)

the finite graph δD in Λ\Γ. As Λ is finitely generated, we can add edges and vertices toδD to obtain a finite connected subgraphδ1Dof Λ\Γ which contains δD and has the property that its inclusion in Λ\Γ induces a surjection of its fundamental group to Λ. Thus the pre-image of δ1D in Γ is a connected graph which we denote by δ1X. Similarly, we obtain a finite connected graph δ1E of Σ\Γ which contains δE and has connected pre-image δ1Y in Γ. As usual, we will denote the pre-images of D and E in G by X and Y respectively.

Next we claim that if gX crosses Y then g(δ1X) intersects δ1Y. (The converse need not be true.) Suppose that g(δ1X) and δ1Y are disjoint. Then g(δ1X) cannot meet δY. As g(δ1X) is connected, it must lie in Y or Y. It follows that g(δX) lies in Y orY, so that one of the four sets X∩Y, X∩Y, X∩Y and X∩Y must be empty, which implies that gX does not cross Y.

Now we can show that i(D, E) must be finite. Recall that i(D, E) is defined to be the number of double cosets ΣgΛ such that gX crosses Y. The preceding paragraph implies that i(D, E) is bounded above by the number of double cosets ΣgΛ such that g(δ1X) meets δ1Y. Let P and Q be finite subgraphs of δ1X and δ1Y which project onto δ1D and δ1E respectively. If g(δ1X) meets δ1Y, then there exist elements λ of Λ and σ of Σ such that g(λP) meets σQ.

Thus σ1gλP meets Q.Now there are only finitely many elements of G which can translate P to meet Q, and it follows that i(D, E) is bounded above by this number.

We have just shown that, as in the preceding section, the intersection numbers we have defined are symmetric, but we will need a little more information.

Lemma 2.8 Let G be a finitely generated group with subgroups Λ and Σ, let D denote a non-trivial almost invariant subset of Λ\G, and let E denote a non-trivial almost invariant subset of Σ\G. Then the following statements hold:

1) i(D, E) =i(E, D),

2) i(D, E) =i(D, E) =i(D, E) =i(D, E),

3) if D0 is almost equal to D and E0 is almost equal to E, and X, X0 and Y, Y0 denote their pre-images in G, then X crosses Y if and only if X0 crosses Y0, so that i(D, E) =i(D0, E0).

Proof The first part is proved by using the bijection from G to itself given by sending each element to its inverse. This induces a bijection between all

(21)

double cosets ΣgΛ and ΛhΣ by sending ΣgΛ to Λg1Σ,and it further induces a bijection between those double cosets ΣgΛ such that gX crosses Y and those double cosets ΛhΣ such that hY crosses X.

The second part is clear from the definitions.

For the third part, we note that, as E and E0 are almost equal, so are their complements in Σ\G, and it follows that X crosses Y if and only if it crosses Y0. Hence the symmetry proved in Lemma 2.3, shows that Y crosses X if and only Y0 crosses X. Now the same argument reversing the roles of D and E yields the required result.

At this point, we have defined in a natural way a number which can reasonably be called the intersection number of D and E, but have not yet defined an intersection number for subgroups of G. First note that if e(G,Λ) is equal to 2, then all choices of non-trivial almost invariant sets in Λ\G are almost equal or almost complementary. Let D denote some choice here. Suppose that e(G,Σ) is also equal to 2, and let E denote a non-trivial almost invariant subset of Σ\G. The third part of the preceding lemma implies that i(D, E) is independent of the choices of D and E and so depends only on the subgroups Λ and Σ. This is then the definition of the intersection number i(Λ,Σ). In the special case when G is the fundamental group of a closed orientable surface and Λ and Σ are cyclic subgroups of G, it is automatic that e(G,Λ) and e(G,Σ) are each equal to 2. The discussion of the previous section clearly shows that this definition coincides with the topological definition of intersection number of loops representing generators of these subgroups, whether or not those loops are simple. Note that one can also define the self-intersection number of an almost invariant subset D of Λ\G to be i(D, D), and hence can define the self-intersection number of a subgroup Λ of G such that e(G,Λ) = 2. Again this idea generalises the topological idea of self-intersection number of a loop on a surface.

If one considers subgroups Λ and Σ such that e(G,Λ) ore(G,Σ) is greater than 2, there are possibly different ideas for their intersection number depending on which almost invariant sets we pick. (It is tempting to simply define i(Λ,Σ) to be the minimum possible value for i(D, E), where D is a non-trivial Λ–almost invariant subset of G and E is a non-trivial Σ–almost invariant subset of G.

But this does not seem to be the “right” definition.) However, there is a natural way to choose these almost invariant sets if we are given splittings of G over Λ and Σ. As discussed in the previous section in the case of surface groups, the standard way to do this when G =A∗ΛB is in terms of canonical forms for

(22)

elements of G as follows. Pick right transversals T and T0 for Λ in A and B respectively, both of which contain the identity e of G. Then each element can be expressed uniquely in the forma1b1. . . anbnλ, wheren≥1, λlies in Λ,each ai lies in T − {e} except that a1 may be trivial, and each bi lies in T0− {e} except that bn may be trivial. Let X denote the subset of G consisting of elements for which a1 is non-trivial, and let D denote Λ\X. It is easy to check directly that X is Λ–almost invariant. One must check that λX =X, for all λ in Λ and that Dg=a D, for all g in G. The first equation is trivial, and the second is easily checked when g lies in A or B, which implies that it holds for all g in G. Note also that the definition of X is independent of the choices of transversals of Λ in A and B. Then D is the almost invariant set determined by the given splitting of G. This definition seems asymmetric, but if instead we consider the Λ–almost invariant subset of G consisting of elements whose canonical form begins with a non-trivial element ofB, we will obtain an almost invariant subset of Λ\G which is almost equal to the complement of D. There is a similar description of D when G=A∗Λ. For details see Theorem 1.7 of [11]. The connection between D and the given splitting of G can be seen in several ways. From the topologists’ point of view, one sees this as described earlier for surface groups. From the point of view of groups acting on trees, there is also a very natural description. One identifies a splitting of G with an action of G on a tree T without inversions, such that the quotient G\T has a single edge. Let e denote the edge of T with stabiliser Λ, let v denote the vertex ofewith stabiliser A,and let E denote the component ofT−{e} which contains v. Then we can define X = {g G : ge E}. It is easy to check directly that this set is the same as the set X defined above using canonical forms.

In the preceding paragraph, we showed how to obtain a well defined intersection number of given splittings over Λ and Σ. An important point to notice is that this intersection number is not determined by the subgroups Λ and Σ of G only. It depends on the given splittings. In the case when G is a surface group, this is irrelevant as there can be at most one splitting of a surface group over a given infinite cyclic subgroup. But in general, a group G with subgroup Λ can have many different splittings over Λ.

Example 2.9 Here is a simple example to show that intersection numbers depend on splittings, not just on subgroups. First we note that the self- intersection number of any splitting is zero. Now construct a group G by amalgamating four groups G1, G2, G3 and G4 along a common subgroup Λ.

Thus G can be expressed as G12ΛG34, where Gij is the subgroup of G gen- erated byGi and Gj,but it can also be expressed as G13ΛG24 or G14ΛG23.

(23)

The intersection number of any distinct pair of these splittings ofGis non-zero, but all the splittings being considered are splittings over the same group Λ.

A question which arose in our introduction in connection with the work of Rips and Sela was how the intersection number of two subgroups of a group G alters if one replaces G by a subgroup. In general, nothing can be said, but in interesting cases one can understand the answer to this question. The particular case considered by Rips and Sela was of a finitely presented group G which is expressed as the fundamental group of a graph of groups with some vertex group being a groupH which contains infinite cyclic subgroups Λ and Σ.

Further H is the fundamental group of a surface F and Λ and Σ are carried by simple closed curves L and S on F. A point deliberately left unclear in our earlier discussion of their work was that F is not a closed surface. It is a compact surface with non-empty boundary. The curves L and S are not homotopic to boundary components and so define splittings of H. The edges in the graph of groups which are attached toH all carry some subgroup of the fundamental group of a boundary component of F. This implies that L and S also define splittings of G. It is clear from this picture that the intersection number of Λ and Σ should be the same whether measured in G or in H, as it should equal the intersection number of the curves L and S, but this needs a little more thought to make precise. As usual, the first point to make is that we are really talking about the intersection numbers of the splittings defined by L and S, rather than intersection numbers of Λ and Σ. For the number of ends e(H,Λ) and e(H,Σ) are infinite when F is a surface with boundary. As G is finitely presented, we can attach cells to the boundary of F to construct a finite complex K with fundamental group G. Now the identification of the intersection number of the given splittings of G with the intersection number of L and S proceeds exactly as at the start of this section, where we showed how to identify the intersection number of the given splittings of H with the intersection number of L and S.

3 Interpreting intersection numbers

It is natural to ask what is the meaning of the intersection numbers defined in the previous section. The answer is already clear in the case of a surface group with cyclic subgroups. In this section, we will give an interpretation of the intersection number of two splittings of a finitely generated group G over finitely generated subgroups. We start by discussing the connection with the work of Kropholler and Roller.

参照

関連したドキュメント

In this work we study spacelike and timelike surfaces of revolution in Minkowski space E 3 1 that satisfy the linear Weingarten relation aH + bK = c, where H and K denote the

In Section 5, we study the contact of a 1-lightlike surface with an anti de Sitter 3-sphere as an application of the theory of Legendrian singularities and discuss the

For example, a maximal embedded collection of tori in an irreducible manifold is complete as each of the component manifolds is indecomposable (any additional surface would have to

Nicolaescu and the author formulated a conjecture which relates the geometric genus of a complex analytic normal surface singularity (X, 0) — whose link M is a rational homology

Kilbas; Conditions of the existence of a classical solution of a Cauchy type problem for the diffusion equation with the Riemann-Liouville partial derivative, Differential Equations,

We develop a theory of convex cocompact subgroups of the mapping class group M CG of a closed, oriented surface S of genus at least 2, in terms of the action on Teichm¨ uller

The main problems which are solved in this paper are: how to systematically enumerate combinatorial braid foliations of a disc; how to verify whether a com- binatorial foliation can

A Darboux type problem for a model hyperbolic equation of the third order with multiple characteristics is considered in the case of two independent variables.. In the class