ABELIAN SUBGROUPS OF THE FUNDAMENTAL GROUP OF A SPACE WITH NO CONJUGATE POINTS
Abstract.
Each Abelian subgroup of the fundamental group of a compact and locally simply connected -dimensional length space with no conjugate points is isomorphic to for some . It follows from this and previously known results that each solvable subgroup of the fundamental group is a Bieberbach group. In the Riemannian setting, this may be proved using a novel property of the asymptotic norm of each Abelian subgroup.
2010 Mathematics Subject Classification:
Primary 20F65 and 53C20; Secondary 53C221. Introduction
A locally simply connected length space with universal cover has no conjugate points if any two points in can be joined by a unique geodesic. Let be a compact and locally simply connected length space with no conjugate points and finite Hausdorff dimension . In the Riemannian case, it has been believed for some time that Abelian subgroups of must be finitely generated; for example, this is stated in [2], although the argument there contains a gap. It will be shown here that each Abelian subgroup is isomorphic to for some .
Theorem 1.
Each Abelian subgroup of is isomorphic to for some .
For nonpositively curved manifolds, this is a consequence of the flat torus theorem of Gromoll–Wolf [3] and Lawson–Yau [6], which was generalized to manifolds with no focal points by O’Sullivan [10].
It was proved by Yau [11] in the case of nonpositive curvature, and O’Sullivan [10] for no focal points, that every solvable subgroup of the fundamental group is a Bieberbach group. Croke–Schroeder [2] mapped out a way to generalize this to spaces with no conjugate points: If a torsion-free solvable group has the property that its Abelian subgroups are all finitely generated and straight, then it must be a Bieberbach group. Lebedeva [7] showed that finitely generated Abelian subgroups of the fundamental group of a compact and locally simply connected length space with no conjugate points must be straight. Combining this with Theorem 1 completes the argument set out by Croke–Schroeder.
Theorem 2.
Each solvable subgroup of is a Bieberbach group.
This continues the theme, developed in [2], [7], [4], and unpublished work of Kleiner that, at the level of fundamental group, spaces with no conjugate points resemble those with nonpositive curvature.
Since the exponential map at each point of its universal cover is a diffeomorphism, a Riemannian manifold with no conjugate points must be aspherical. It’s worth pointing out that this condition isn’t enough to guarantee the conclusion of Theorem 1, as Mess [9, 8] showed that for each there exists a compact manifold with universal cover whose fundamental group contains a divisible Abelian subgroup, which cannot be finitely generated.
The second section contains a short proof of Theorem 1. The third section gives a different proof in the Riemannian setting, based on a property of Riemannian norms satisfied by the asymptotic norm of each Abelian subgroup of the fundamental group.
2. Proof of Theorem 1
Fix and a basepoint for . Overloading notation, each will be identified with the corresponding deck transformation of . Let be an Abelian subgroup of , in which the group operation is written additively, and suppose are linearly independent. Denote by the subgroup generated by the . The following are proved in [7]: On , the function
is positively homogeneous over . It is bounded below on by , the length of the shortest nontrivial geodesic loop in , so is torsion free. Its restriction to satisfies the triangle inequality, and, with respect to the isomorphism that takes each to the -th standard basis vector, extends to a norm on .
Denote by the Euclidean norm on . From the identifications , inherits the coordinate functions on . Since is a norm on , there exists such that
for all . The number is a Lipschitz constant for the on , and, as in the proof of Kirszbraun’s theorem [5], the functions
are Lipschitz extensions of the to . Each is -equivariant, in the sense that for all and all .
The map is Lipschitz, and for all and all . By construction, . Since is Abelian, there exists a map such that . Lift to a map . The composition descends to a map with surjective induced homomorphism, so by degree theory it must be surjective. Thus is surjective. Since a Lipschitz map cannot increase Hausdorff dimension, .
It follows that has rank at most . If it has rank zero, then the result is trivial. Without loss of generality, suppose it has rank . For any , there exist such that . It is well known that the function defined by and
for is a well-defined and injective homomorphism, so is an isomorphism onto its image . This map satisfies
for any . For any distinct , there exist distinct such that for each . For , one has that
Thus is a discrete subgroup of , and, consequently, .
3. Busemann functions in the Riemannian setting
For simplicity, it will be assumed in this section that is a smooth -dimensional Riemannian manifold, although what follows holds when is for some depending on . As before, let be an Abelian subgroup of generated by linearly independent . The key step in the proof of Theorem 1 is the construction of a -equivariant map such that . When is Riemannian, another such map may be constructed using a nondegenerate collection of Busemann functions.
An important theorem of Ivanov–Kapovitch [4] states that, whenever commute, the change in the Busemann functions of axes of under the action of is constant on . This was previously proved by Croke–Schroeder [2] for analytic . Thus one may define a function by setting equal to that change.
Because for all , one might hope to show that extends to an inner product and, consequently, that is Riemannian. In fact, satisfies a number of the properties of an inner product: It is linear over in the first slot (see Corollary 4.2 of [4]), for all , and it satisfies a version of the Cauchy–Schwarz inequality,
(1) |
with equality if and only if and are rationally related. It follows that extends to an inner product if and only if it is symmetric, but it’s far from clear that symmetry holds in general (cf. [1]). Regardless, also resembles an inner product in the following way.
Lemma 3.
For each , there exist such that the matrix is nonsingular.
If are as in Lemma 3 and are Busemann functions of respective axes, then up to composition with an affine isomorphism the map is -equivariant and satisfies . The Riemannian version of Theorem 1 follows.
The proof of Lemma 3 is by induction. When , the conclusion holds with . Suppose the conclusion holds for some . If the conclusion fails when , then there exists a nonzero in the null space of the matrix . The following lemma then completes the inductive step.
Lemma 4.
There exists a solid cone centered around the ray such that, if , for , and , then the matrix is nonsingular.
The proof of Lemma 4 uses the following elementary fact.
Lemma 5.
Let . Suppose is a sequence of matrices of the form
for a fixed matrix and a sequence such that . Suppose also that is a sequence of vectors in of the form
for satisfying and . If satisfy , then . Consequently, has nontrivial null space.
Proof of Lemma 4.
Without loss of generality, one may suppose that . Assume for the sake of contradiction that the result is false. Then, for each and any fixed sequence , there exists a sequence of rational numbers such that and, when for and , each matrix is singular.
Let for , and write
(2) | ||||
Let . Then
(3) | ||||
The inductive hypothesis and the linearity of in the first slot imply that are linearly independent. The word norm of with respect to the subgroup of generated by is
Because the corresponding norms on are equivalent, there exists , depending only on , such that
By the Cauchy–Schwarz inequality (1), for each ,
(4) |
Similarly,
(5) |
Let , , , and for . Write
and
By (2) and (3), ; by (4), ; and, by (5), for all sufficiently large . Since is nonsingular, it follows from Lemma 5 that the matrices
are nonsingular for all such . The corresponding must also be nonsingular, which is a contradiction. ∎
When in Lemma 3, inequality (1) implies that one may take and . When has no focal points, one may, by the flat torus theorem, take for all . However, in the general case for , there is no apparent local structure that forces the Busemann functions of the axes of the to have linearly independent gradients, and it is not clear that the conclusion of Lemma 3 holds with for all .
Question 6.
Must the matrix be nonsingular?
References
- [1] Dmitri Burago and Sergei Ivanov, Riemannian tori without conjugate points are flat, Geom. Funct. Anal. 4 (1994), 259–269.
- [2] Christopher B. Croke and Viktor Schroeder, The fundamental group of compact manifolds without conjugate points, Comment. Math. Helv. 61 (1986), 161–175.
- [3] Detlef Gromoll and Joseph A. Wolf, Some relations between the metric structure and the algebraic structure of the fundamental group in manifolds of nonpositive curvature, Bull. Amer. Math. Soc. 77 (1971), 545–552.
- [4] Sergei Ivanov and Vitali Kapovitch, Manifolds without conjugate points and their fundamental groups, J. Differential Geom. 96 (2014), 223–240.
- [5] Mojżesz D. Kirszbraun, Über die zusammenziehende und Lipschitzsche Transformationen, Fund. Math. 22 (1934), 77–108 (German).
- [6] Blaine H. Lawson and Shing Tung Yau, Compact manifolds of nonpositive curvature, J. Differential Geom. 7 (1972), 211–228.
- [7] Nina Lebedeva, On the fundamental group of a compact space without conjugate points, PDMI preprint, www.pdmi.ras.ru/preprint/2002/02-05.html (2002).
- [8] Wolfgang Lück, Aspherical manifolds, Bulletin of the Manifold Atlas (2012), 1–17.
- [9] Geoffrey Mess, Examples of Poincaré duality groups, Proc. Amer. Math. Soc. 110 (1990), 1145–1146.
- [10] J.J. O’Sullivan, Riemannian manifolds without focal points, J. Differential Geometry 11 (1976), 321–333.
- [11] Shing Tung Yau, On the fundamental group of compact manifolds of non-positive curvature, Ann. of Math. 93 (1971), 579–585.