2000 Mathematics Subject Classification. 58H15 (primary), 57M50, 20G10 (secondary).Supported by the INTAS project “CalcoMet-GT” 03-51-3663.
On deformations of hyperbolic
-manifolds with geodesic boundary
Roberto Frigerio
Dipartimento di Matematica, Universita di Pisa, Largo B. Pontecorvo 5, 56127 Pisa, Italy E-mail address : frigerio@mail.dm.unipi.it
-
Abstract.
Let
be a complete finite-volume hyperbolic
-manifold with compact non-empty geodesic boundary and let
be a geometric partially truncated triangulation of
. Also suppose that
has
toric cusps. We show that the set of solutions of consistency equations for
is a smooth manifold or real dimension
near the point representing the unique complete structure on
. As a consequence, small deformations of the solution representing the complete structure affect the compact tetrahedra and the hyperbolic structure on the geodesic boundary only at the second order.
The idea of constructing hyperbolic structures on manifolds by suitably gluing to each other geodesic polyhedra dates back to Thurston [?] . In the setting of cusped manifolds one employs ideal tetrahedra, which are parameterized by complex numbers, and tries to solve hyperbolicity equations. In [?] we explained how this approach can be adapted to the case of non-empty geodesic boundary: in the bounded case one has to consider truncated tetrahedra, whose parameterization is more complicated, but basically the whole scheme extends.
The conditions under which a gluing of truncated tetrahedra defines a non-singular hyperbolic metric are encoded by consistency equations, while completeness equations translate the conditions ensuring that such a metric is complete. For our purposes it is crucial to control the number of consistency equations, and this is the reason why the equations described here are quite different from those introduced in [?] . The set of solutions of consistency equations naturally provides a deformation space for finite-volume hyperbolic structures with geodesic boundary on a fixed
-manifold. Building on classical results in cohomology theory of representations, we prove that the complete structure is a smooth point of this deformation space and we explicitly describe local coordinates around it. This allows us to give a proof of Thurston's hyperbolic Dehn filling Theorem which applies to all the hyperbolic manifolds with geodesic boundary which admit a genuine geometric triangulation. It is actually conjectured that any complete finite-volume hyperbolic
-manifold with geodesic boundary admits a genuine triangulation; moreover, any such manifold admits a partially flat triangulation, and following [?] we could probably adapt our proof of Thurston's hyperbolic Dehn filling Theorem to deal also with this kind of triangulations. This would give a complete and self-contained proof of the filling Theorem via deformation theory of geometric triangulations.
In the last section we show that small deformations of the solution representing the complete structure affect the compact tetrahedra and the hyperbolic structure on the geodesic boundary only at the second order. These results are used in [?] in order to simplify computations in some concrete examples.
It is maybe worth mentioning that the deformation variety defined by the consistency equations for a cusped
-manifold without boundary has already been studied by several authors (see e. g. [?, ?] ). In particular, Choi has recently proved in [?] that in the cusped empty-boundary case the deformation variety is a smooth complex manifold at any point representing a non-degenerate (i. e. neither partially flat nor partially negatively oriented) geodesic ideal triangulation.
1 Triangulations and hyperbolicity equations
Let
be a complete finite-volume orientable hyperbolic 3-manifold with compact non-empty geodesic boundary (from now on we will usually summarize all this information saying just that
is hyperbolic). It is well-known that
consists of a compact portion containing
together with several cusps of the form
, where
is the torus, so
admits a natural compactification
obtained by adding some boundary tori. Since the components of
are totally geodesic, they inherit a hyperbolic metric, and have therefore negative Euler characteristic.
1.1 Partially truncated tetrahedra
A partially truncated tetrahedron is a pair
, where
is a tetrahedron and
is a set of vertices of
, that will be called ideal vertices. In the sequel we will always refer to
itself as a partially truncated tetrahedron, tacitly implying that
is also fixed. The topological realization
of
is obtained by removing from
the ideal vertices and small open stars of the non-ideal vertices. We call lateral hexagon and truncation triangle the intersection of
respectively with a face of
and with the link in
of a non-ideal vertex. The edges of the truncation triangles, which also belong to the lateral hexagons, are called boundary edges, and the other edges of
are called internal edges. If
has ideal vertices, a lateral hexagon of
may not be a hexagon, because some of its closed edges may be missing.
A geometric realization of
is an identification of
with a convex polyhedron in
such that the truncation triangles are geodesic triangles, the lateral hexagons are geodesic polygons with ideal vertices corresponding to missing edges, and truncation triangles and lateral hexagons lie at right angles to each other. An example of a geometric realization is shown in Fig. 1 , where truncation triangles are shadowed.
1.2 Triangulations
Let
be a compact orientable manifold and let
be obtained from
by removing the toric components of
. We define a partially truncated triangulation of
to be a realization of
as a gluing of some
's along a pairing of the lateral hexagons induced by a simplicial pairing of the faces of the
's. When
is endowed with a hyperbolic structure, a partially truncated triangulation of
is called geometric if, for each tetrahedron
of the triangulation, the pull-back to
of the Riemannian metric of
defines a geometric realization of
. Equivalently, the hyperbolic structure of
should be obtained by gluing geometric realizations of the
's along isometries of their lateral hexagons.
Definition 1.1.
A partially truncated triangulation
of an orientable
-manifold
is good if any tetrahedron in
has at most one ideal vertex.
Kojima proved in [?] that every hyperbolic
has a canonical decomposition into partially truncated polyhedra, rather than tetrahedra. Any polyhedron in a canonical decomposition has at most one ideal vertex, so triangulations arising as subdivisions of Kojima decompositions are examples of good triangulations. In the vast majority of cases the Kojima decomposition actually consists of tetrahedra, or at least can be subdivided into a geometric partially truncated triangulation. For instance, it is proved in [?] that there exist exactly
hyperbolic manifolds with non-empty geodesic boundary which can be (topologically) triangulated by at most four partially truncated tetrahedra: their Kojima decomposition can always be subdivided into a triangulation, and is itself a triangulation in
cases. These facts support the following:
Conjecture 1.2.
Any hyperbolic
with non-empty godesic boundary admits a good geometric triangulation.
1.3 Moduli for partially truncated tetrahedra
The following result implies that the dihedral angles can be used as moduli for geometric tetrahedra.
Theorem 1.3.
Let
be a partially truncated tetrahedron and let
be the set of edges of
. The geometric realizations of
are parameterized up to isometry by the dihedral angle assignements
such that for each vertex
of
, if
are the edges that emanate from
, then
is equal to
for ideal
and less than
for non-ideal
.
Having introduced moduli for geometric tetrahedra, our next task is to determine, given a triangulated manifold, which values of moduli define a global hyperbolic structure on the manifold. The following well-known hyperbolic trigonometry formulae will prove useful later:
Lemma 1.4.
With notation as in Fig. 2 we have
| |
| |
Let now
be a partially truncated tetrahedron with edges
as in Fig. 3 . We fix a geometric realization
of
determined by the dihedral angles
for
and we denote by
the length with respect to this realization. The boundary edges of the lateral hexagons of
correspond to the pairs of distinct non-opposite edges
, and will be denoted by
. Note that
disappears towards infinity, so it has length
, when the common vertex of
and
is ideal. Lemma 1.4 readily implies
|
(1)
|
Note that this result is correct also when the common end of
and
is ideal.
Turning to the length of an internal edge, we note that the edge is an infinite half-line or an infinite line when one or both its ends are ideal, respectively. Otherwise the length is computed using Lemma 1.4 . With notation as in Fig. 3 , and defining
as the vertex from which the edges
emanate, we set:
| |
| |
Proposition 1.5.
if and only if the vertex
is ideal. If
and
are both non-ideal then
|
(2)
|
Remark 1.6.
Let
be a partially truncated tetrahedron without ideal vertices. Then the geometric realizations of
are parameterized by the lengths of the internal edges. In fact, the map that associates to the dihedral angles of a geometric realization of
the lengths of its internal edges is a diffeomorphism between open subsets of
.
1.4 Conditions for geometric gluing
Let
be obtained from an orientable compact
by removing all the tori in
and let
be a partially truncated triangulation of
. Let also
be a geometric realization of the tetrahedra in
and denote by
the length with respect to this realization. We now describe the conditions under which the realization
defines a hyperbolic structure on the whole of
. For our purposes it will be sufficient to deal only with good triangulations, so we assume from now on that
is good. The general case is treated in [?] .
In order to define a global hyperbolic structure on
, the tetrahedra of
must satisfy two obvious necessary conditions, which in fact are also sufficient. Namely, we should be able to glue the lateral hexagons by isometries, and we should have a total dihedral angle of
around each edge of the manifold. The first condition ensures that the hyperbolic structure defined by
on the complement of the
-skeleton of
extends to the complement of the
-skeleton. Since
is good, the second one ensures that the structure glues up without singularities also along the edges. The second condition is directly expressed in terms of moduli, and we will explain in a moment how to translate the first one into an equation on dihedral angles.
Remark 1.7.
If
were not good, requiring the dihedral angles around each edge to sum up to
would not be sufficient to obtain a non-singular hyperbolic metric on the
-skeleton of
. The point is that when some geometric tetrahedra are arranged one after the other around an edge
with two ideal endpoints, the first face of the first tetrahedron and the second face of the last tetrahedron may overlap without coinciding. Namely, the isometry which pairs these two faces may be a translation along
instead of being the identity. Of course the isometry has to be the identity if at least one endpoint of
is not ideal.
1.5 Exceptional hexagons
It is easily seen that gluings match ideal vertices to each other, because these notions are part of the initial topological information about a triangulation. When a pairing glues two compact lateral hexagons, to be sure that the gluing is an isometric one we may equivalently require the lengths of the internal edges or those of the boundary edges to match under the gluing. On the other hand, since
is a good triangulation, a non-compact lateral hexagon
is actually a pentagon with four right angles and an ideal vertex: we shall say in this case that
is an exceptional lateral hexagon. By Lemma 1.4 , the isometry class of an exceptional lateral hexagon is determined by the lengths of its boundary edges. However, in order to end up with a non-redundant set of consistency equations, it is convenient to find an alternative approach to moduli for exceptional hexagons.
To this aim we need now to be slightly more careful about orientation than we have been so far. Namely, we choose on the tetrahedra an orientation compatible with a global orientation of the manifold. As a result also the lateral hexagons have a fixed orientation, and the gluing maps reverse the orientation of the hexagons.
So, let us consider an exceptional hexagon
as in Fig. 4 , and recall that the hexagon is oriented and embedded in
by
. We consider the horospheres
and
centred at
and passing through the non-ideal ends of
and
respectively. We define
to be
, the sign being positive if
are arranged positively on
and
is contained in the horoball bounded by
, or if
are arranged negatively on
and
is contained in the horoball bounded by
, and negative otherwise.
Together with equation ( 1 ), the following proposition allows to compute
in terms of the dihedral angles.
Proposition 1.8.
We have
|
(3)
|
Proof: Let
and
be the angles and lengths shown in Fig. 5 . An easy computation in the upper half-plane model of
shows that
for
. Moreover we have
, so
.
Now for
we have
. Combining these equations we finally get
.
We now define
to be the length of
. The next proposition shows that the functions
and
provide a parameterization of isometry classes of exceptional hexagons.
Proposition 1.9.
Let
and
be paired exceptional lateral hexagons. Their pairing can be realized by an isometry if and only if
and
.
Proof: We concentrate on the “if ” part of the statement, the “only if ” part being obvious.
Let
and
be the boundary edges of of
and
respectively, and assume that the orientation-reversing pairing between
and
glues
to
and
to
. Since
, Lemma 1.4 easily implies that
if and only if
.
Moreover the assumption that
and Proposition 1.8 give
if and only if
. This forces
and
, whence the conclusion.
Remark 1.10.
Let
be a partially truncated tetrahedron and suppose that
is the unique ideal vertex of
. Let
be the faces of
incident to
and for
let
be the edge of
not containing
. The isometry classes of the geometric realizations of
are parameterized by the lengths of the
's and the values taken by
on the
's. More precisely, the map that associates to any geometric realization
of
the point
defines a diffeomorphism between open subsets of two affine hyperplanes of
.
1.6 Consistency equations
Recall now that we are considering a candidate hyperbolic
-manifold
endowed with a good triangulation
, and that we have fixed a geometric realization
of the tetrahedra in
. The above discussion implies the following:
Theorem 1.11.
The parameterization
defines on
a hyperbolic structure with geodesic boundary if and only if the following conditions hold:
-
(1)
the total dihedral angle along any edge of
in
is equal to
;
-
(2)
for all pairs
of matching compact internal edges;
-
(3)
for all pairs
of matching exceptional hexagons.
By Propositions 1.5 , 1.8 , conditions (1), (2), (3) of Theorem 1.11 translate into a set
of smooth equations on
, which are called consistency equations. Our next task is to compare the number of equations in
with the dimension of the moduli space of geometric realizations of the tetrahedra of
.
Let
(resp.
) be the number of compact (resp. non-compact) tetrahedra of
. By Theorem 1.3 , the number of parameters for the geometric realizations of
is equal to
, where
is the total number of tetrahedra of
. Let
be the edges of
without ideal endpoints, and for
let
be the valence of
, i. e. the number of tetrahedra of
incident to
, with multiplicity. Of course we have
. If we denote by
the number of edges of
with exactly one ideal endpoint, then the union of the boundary tori of
admits a triangulation with
vertices and
triangles, so
. Let us now turn to the number of equations. Theorem 1.11 determines
equations arising from conditions (1),
equations arising from conditions (2) and
equations arising from conditions (3). Since
| |
| |
we can conclude that the number of equations in
is equal to the dimension of the moduli space of geometric realizations of the tetrahedra of
.
1.7 Reducing the number of equations
When
has cusps, some equations in
turn out to be redundant. Let
be a fixed toric component of
and let
be the number of tetrahedra asymptotic to
. Since any such tetrahedron contributes to the triangulation of
with a Euclidean triangle, the sum of all the dihedral angles along all the edges of
incident to
is equal to
. So if we require condition (1) of Theorem 1.11 to hold for all but one edge incident to
, then the same condition is automatically satisfied also along the remaining edge. This allows us to discard from
one equation for each cusp of
.
Moreover, let
be a partially truncated tetrahedron with an ideal vertex
incident to
, and let
be the faces of
incident to
. By the very definition of
it follows that
. This implies that if
is the set of all the faces incident to
of tetrahedra of
, then we have
. So if we require condition (3) of Theorem 1.11 to hold for all but one pair of matching exceptional hexagons incident to
, then the same condition also holds for the remaining pair. This means that another equation of
for each cusp of
can be discarded. Supppose that
compactifies to an orientable
with
boundary tori. The above discussion is summarized by the following:
Proposition 1.12.
We can discard
equations from
thus obtaining an equivalent set
of
equations, where
is the dimension of the moduli space of geometric realizations of the tetrahedra of
.
We have seen in the preceding subsection that the moduli space of geometric realizations of the tetrahedra of
is given by a subset
of
, where
is the number of tetrahedra of
. More precisely,
is an open convex subset of an affine subspace of dimension
, where
is the number of non-compact tetrahedra in
. Recall that
is the number of compact edges (considered as subsets of
) of
. The above computation implies that equations in
corresponding to conditions (1) of Theorem 1.11 take the form
, where
is an affine map, while equations corresponding to conditions (2) and (3) take the form
, where
is constructed from formulae ( 2 ), ( 3 ), and is therefore smooth. From now on we denote by
the set of solutions of consistency equations
.
1.8 Completeness
Let
be the boundary tori of
. From now on we denote by
a fixed basis of
,
. Any point in
naturally defines an
-structure on
(see e. g. [?, ?] ). For
, we denote by
(resp. by
) the linear component of the holonomy of
(resp. of
) corresponding to the
-structure defined by
on
. It is well-known that the hyperbolic structure defined by
on
induces a complete metric on the
-th cusp of
if and only if
.
Moreover, one can explicitly compute
and
in terms of the dihedral angles as follows.
Let
be a tetrahedron in
, let
be an ideal vertex of
and
be the (similarity class of the) Euclidean triangle obtained by intersecting the geometric realization of
parameterized by
with a small horosphere centred at
. The tetrahedron being oriented, this triangle is also oriented, so, once a vertex
of
is fixed, we can associate to the similarity structure of
the unique complex number
such that
is carried to the Euclidean triangle with vertices
by an orientation-preserving similarity sending
to
. Suppose that
are the internal edges emanating form
, and that they are positively arranged around
. If
, then
If
is an oriented simplicial loop on
and
is a vertex of
, then the set of all triangles touching
in
and lying on the right of
is well-defined and will be denoted by
.
Moreover, we shall denote by
the set of vertices of
. Let
be simplicial loops on
representing
. The following result is proved in [?, ?] .
As a consequence of Mostow-Prasad's rigidity Theorem for hyperbolic manifolds with geodesic boundary [?, ?] we get the following:
Theorem 1.14.
There exists at most one point in
that defines on
a complete hyperbolic structure with geodesic boundary.
Proof: See [?] .
2 Smoothness at the complete structure
Suppose now that
is the unique point in
which defines on
a complete hyperbolic structure. For
and
let us define
where
is the branch of the complex logarithm defined on
such that
. Since a non-trivial parabolic isometry does not commute with a non-trivial orientation-preserving hyperbolic isometry, we have the following:
Proposition 2.1.
In a neighbourhood of
in
we have
the hyperbolic structure defined by
on the
-th cusp of
is complete.
Let
,
be the smooth functions previously defined such that
. We now set
By Theorem 1.14 and Proposition 2.1 we have
. This section is entirely devoted to the proof of our main result:
Theorem 2.2.
We have
. Thus:
-
(1)
induces a diffeomorphism of an open neighbourhood of
in
onto an open neighbourhood of
in
;
-
(2)
is a smooth manifold of real dimension
near
;
-
(3)
the map
induces a diffeomorphism of an open neighbourhood of
in
onto an open neighbourhood of
in
.
Any point in
defines a cone structure on
having cone singularities along the edges of
. The following immediate consequence of Theorem 2.2 was also proved in [?, ?] .
Corollary 2.3.
If
is compact then
is parameterized in a neighbourhood of
by the cone angles along the edges of
.
2.1 Deforming cone structures
We begin with the following:
Proposition 2.4.
The tangent map
has maximal rank, so
is a manifold of real dimension
near
.
Proof: By Remarks 1.6 , 1.10 , the lengths of the compact internal edges and the values taken by
on the exceptional lateral hexagons of the tetrahedra of
provide smooth coordinates on
. It is easily seen that with respect to these coordinates the map
is affine and has maximal rank at
(whence at any point of
).
Let now
be a smooth arc with
. We will study the deformation associated to
using tools from the cohomology theory of representations:
notation is as in Appendix A , where we give some basic definitions and results.
From now on we denote by
the non-compact manifold obtained by drilling from
all the edges of
. For any
the point
determines a smooth hyperbolic structure
on
, whose completion gives a hyperbolic cone structure on
. First of all we describe how to deduce the shape of the geometric tetrahedra corresponding to the point
just from the geometric structure
on
. To this end we fix for
a developing map
with associated holonomy representation
(note that we can choose
and
to vary smoothly with
). Recall that
compactifies to a manifold
with
boundary tori, and denote by
the non-compact manifold obtained by drilling from
the closed properly embedded arcs corresponding to the edges of
.
For
let
be the punctured torus
, and denote by
the family of all the boundary components of the universal covering of
projecting to some
. We say that
is a parabolic peripheral subgroup of
if
is the stabilizer of some boundary component
of the universal covering of
(so
is a
parabolic subgroup of
). Let
be the portion of
corresponding to the geodesic boundary of
, i. e. let
, and denote by
the family of all the boundary components of the universal covering of
projecting to some component of
. We say that
is a Fuchsian peripheral subgroup of
if
is the stabilizer of some component
(so
is a Fuchsian subgroup of
).
2.2 Dual vectors to planes and horospheres
If
, by construction the image of
under the developing map
is contained in a totally geodesic immersed surface in
. It is easily seen that such a surface must in turn be contained in a geodesic plane
of
. We now need to associate to each
a suitable ultra-ideal point
, which will be called the dual point of
. Such point naturally lies in 4-dimensional Minkowsky space, so we fix some notation about this space.
We denote by
the space
with coordinates
endowed with the Lorentzian inner product
. We set
| |
| |
| |
We recall that
is the upper sheet of the two-sheeted hyperboloid, and that
restricts to a Riemannian metric on
. With this metric,
is the so-called hyperboloid model
of hyperbolic space. The one-sheeted hyperboloid
turns out to have a bijective correspondence with the set of hyperbolic half-spaces in
. Given
, the corresponding half-space, called the dual of
, is given by
.
Similarly, the cone
of future-oriented light-like vectors of
corresponds to the set of horospheres in
. The horosphere dual to
is given by
.
Note now that for any
the set
locally lies on a definite side of
, so we can define
to be the dual vector to the half-space that locally contains
and is bounded by
.
When
belongs to
a vector
can also be defined as follows: take an oriented edge
of
ending in
and set
to be the unique point in
with
which projects to the endpoint of
in
. Such an endpoint exists because
is a geodesic, and is clearly independent of the choice of
, so
is indeed well-defined. Since developing maps vary smoothly with respect to
we have the following:
Lemma 2.5.
For any
the map
is smooth with respect to
.
2.3 Lifting geometric tetrahedra
Let now
be a tetrahedron in
, and denote by
a lift of
to
. Let
be the boundary component of the universal covering of
that corresponds to
, where
are the vertices of
. If
is the convex hull of
in
, then projecting
to
and truncating its infinite-volume ends with the corresponding
's gives back a truncated tetrahedron
isometric to the geometric realization of
parameterized by
. It is easily seen that the dihedral angles of
smoothly depend on
, so Theorem 2.2 is now reduced to the following:
Proposition 2.6.
If
, then we can choose the
's in such a way that
for all
.
2.4 The tangent vector to
Let us consider the double
of
obtained by mirroring
along its boundary (which is now given by some punctured surfaces of negative Euler characteristic). Since
is totally geodesic with respect to the hyperbolic structure
, this structure can be doubled to a smooth hyperbolic metric
on
.
Let
be a holonomy representation corresponding to a developing map for
. It is easily seen that we can assume
to vary smoothly with respect to
. We set
As explained in Appendix A , we have
. This subsection is devoted to the proof of the following:
Proposition 2.7.
If
, then
.
From now on we suppose
. Let
be the doubles of the edges of
and for all
let
be a small loop in
encircling
. We denote by
an element in
representing
(such a
is well-defined only up to conjugation). Now our hypothesis implies that if
is the cone angle of (the completion of )
along
, then
. Also observe that we have
for some smooth path
. Differentiating this relation we easily get
. Let
be the kernel of the map
induced by the inclusion, and observe that
is the smallest normal subgroup of
generated by
. Let
be the natural representation associated to
. Since
and
for all
, Lemma A.1 implies the following:
Proposition 2.8.
naturally defines a cocycle
.
Moreover,
belongs to
if and only if
belongs to
.
Now if
is compact, i. e. if no cusps are involved, Theorem A.2 directly applies concluding the proof of Proposition 2.7 . When there are cusps, Proposition 2.7 can be deduced from Theorem A.3 and the following:
Proposition 2.9.
If
is non-compact, then
.
Proof: Let
be such that
is non-trivial parabolic, and let
be the infinite cyclic group generated by
. We have to check that
restricts to a coboundary in
, where
is the natural inclusion.
Without loss of generality we can suppose
for some
. Recall that a preferred element
was previously fixed, set
and observe that since
we have
. Let
be elements in
projecting respectively to
. By Lemma 2.5 a smooth path
exists such that both
and
fixes
for
, so that
| |
| |
where
are smooth arcs with
.
Since
is Abelian, an element
exists such that
in
. By Proposition 2.8 , this readily implies
which after some computations gives
(here we use
). Let us consider the deformation
defined by
for any
.
We claim that
in
. This will easily give that
restricts to a coboundary in
, whence the conclusion. Since derivatives of conjugated deformations differ by a coboundary, we can suppose
Setting
an easy computation shows that
. This readily implies that
is a coboundary.
2.5 The final step
For
let
be the map defined as follows:
, where
is a smooth lift of
(so
is well-defined only up to the sign). Also recall that
and
are respectively a developing map and a holonomy representation for the hyperbolic structure
on
.
The following result can be easily deduced from the proof of the previous proposition.
Lemma 2.10.
Let
be such that
is non-trivial parabolic.
Then
.
As a consequence of Proposition 2.7 and Lemma 2.10 we obtain the following:
Proposition 2.11.
and
can be chosen in such a way that
for all
. Moreover, if
is a non-trivial element of a peripheral parabolic subgroup of
, then
.
Let
be such that
and suppose
is a smooth path such that
is a fixed point for
for any
. We want to study how the derivative of
is related to the derivative of
. We identify
with
and we set
Without loss of generality we can suppose that
does not fix
and that
, so
,
. An easy computation now shows that
From this formula we can readily deduce the following lemmas.
Lemma 2.12.
If
is non-trivial loxodromic and
, then
.
On the other hand, let
be non-trivial parabolic. Also assume that
and
. Then
.
We can now conclude the proof of Proposition 2.6 . Choose
and
as in the statement of Proposition 2.11 . Let
be a boundary component of the universal covering of
which belongs to
, denote by
the stabilizer of
in
and choose an element
with
. By construction the projection of
to
is fixed by
, so Lemma 2.12 applies ensuring
.
Suppose now that
belongs to
, and let
be the stabilizer of
in
. For
with
let
be the fixed points of
on
. Note that since
is loxodromic we can choose
to vary smoothly with respect to
, at least in a small neighbourhood of
. This gives
by Lemma 2.12 . Now a standard result in Kleinian group theory ensures that the set
is dense in the closure at infinity of
, and this easily implies
.
3 Dehn filling
Once the smoothness of
at
is established, one can prove Thurston's hyperbolic Dehn filling Theorem just by following the strategy described in [?] . At this stage, this argument applies only to those hyperbolic manifolds with geodesic boundary which admit a good geodesic triangulation (but see Conjecture 1.2 ). The following result is taken from [?] .
Lemma 3.1.
Let
. Then there exists a complex number
with non-zero imaginary part such that if
is a sequence with
and
for every
, then
.
3.1 Thurston's hyperbolic Dehn filling Theorem
Let
be a sufficiently small neighbourhood of
in
and let
. For
, we define the
-Dehn filling coefficient
as follows: if
, then
; otherwise,
are the unique real solutions of the equation
(Existence and uniqueness of such solutions near
can be easily deduced from Theorem 2.2 and Lemma 3.1 .) Let us set
As a consequence of Theorem 2.2 and Lemma 3.1 we have the following:
Theorem 3.2.
If
is small enough, the map
defines a diffeomorphism onto an open neighbourhood of
in
.
For
we denote by
the hyperbolic structure induced on
by
, and by
the metric completion of
. Recall that a preferred basis
of
is fixed for every
. We also set
| |
| |
Theorem 3.3.
If
is sufficiently small and
belongs to
, then
admits a complete finite-volume smooth hyperbolic structure which is obtained by adding to
a closed geodesic at any cusp with non-infinite Dehn filling coefficient. From a topological point of view,
is obtained by Dehn filling the
-th cusp of
along the slope
if
, and by leaving the
-th cusp of
unfilled if
,
.
Proof: See e. g. [?, ?, ?, ?] .
The following proposition will prove useful in the last subsection.
Proposition 3.4.
Let
be any smooth manifold and let
be a smooth map. Suppose that there exists a small neighbourhood
of
in
such that for all
with
we have
. Then
.
Proof: Since
defines a chart around
, it is sufficient to observe that
accumulates to
along any direction in
.
3.2 Infinitesimal deformations of compact tetrahedra
We know from Mostow's rigidity Theorem that compact hyperbolic
-manifolds do not admit deformations. The following results seem to suggest that in the non-compact case deformations take place mostly near the cusps: even if it has to be affected by any non-trivial deformation, the compact core offers resistance to changing its shape. More precisely, we now show that deformations of
near the complete structure affect compact tetrahedra only at the second order.
Let
be any compact internal edge of
. For
we denote by
the length of
with respect to the metric structure defined by
. By Lemma 1.4 the map
is smooth.
Proof: Let
be such that
. Then the identity of
extends to a homeomorphism between
and
. By Mostow-Prasad's rigidity Theorem, such a homeomorphism is homotopic to an isometry
via a homotopy which preserves the geodesic boundary (see e. g. [?] ). For
let
be the geodesic segment corresponding to
. From the above discussion it follows that
is homotopic to
relatively to
. Since both
and
intersect
perpendicularly, this easily implies that
, whence
. Now the conlcusion follows from Proposition 3.4 .
The lengths of the boundary edges of a compact lateral hexagon smoothly depend on the lengths of its internal edges, and the dihedral angles of a compact truncated tetrahedron smoothly depend on the lengths of its internal edges. Thus Proposition 3.5 imply the following results.
Corollary 3.6.
Fix a boundary edge
lying on a compact lateral hexagon of some tetrahedron of
, and let
be the function which associates to
the length of
in the geometric realization parameterized by
.
Then
.
3.3 Infinitesimal deformations of the geodesic boundary
Let
be the Teichmüller space of hyperbolic structures on
, i. e. the space of equivalence classes of hyperbolic metrics on
, where two such metrics are considered equivalent if they are isometric through a diffeomorphism homotopic to the identity of
. For
we denote by
the equivalence class of the hyperbolic structure induced by
on
. It is well-known that
admits a structure of differentiable manifold such that
is smooth. As a consequence of Mostow-Prasad's rigidity Theorem and of Proposition 3.4 we get the following:
Proposition 3.8.
We have
.
A Cohomology theory of representations
A.1 The tangent space to a representation
Let
be a Lie group with associated Lie algebra
and
be any group, and denote by
the set of representations of
in
. We say that a path
is smooth if
is a smooth function of
for any
. If
is a smooth path of representations, the tangent vector to the map
at
gives an element in
.
Identifying this tangent space with
by right translation we get an element
of the Lie algebra
:
Differentiating the homomorphism relation
we see that
satisfies the so-called cocycle relation
where
is the usual adjoint representation.
Consider now a trivial deformation of
, i. e. let
be a smooth path in
starting at the identity and set
for all
. Then differentiation shows that
for any
, where
is the tangent vector to
at
. We now set
| |
| |
| |
The above discussion shows that
corresponds in some sense to the tangent space of
at
. Under this identification the module
should represent the tangent space to trivial deformations of
, so
should give the tangent space of
at
(however, this holds true only in the setting of algebraic schemes).
Let
, and suppose that
is a normal subgroup of
such that
,
for all
. Let also
be the representation induced by
.
Lemma A.1.
The map
defined by
is well-defined and gives a cocycle
. Moreover, we have
if and only if
.
A.2 Classical rigidity results
Let
be a smooth
-manifold without boundary and suppose
is the holonomy representation for a complete finite-volume hyperbolic structure on
.
The following result is due to Weil [?] , and can be considered as a local version of Mostow's rigidity Theorem for compact hyperbolic
-manifolds.
Theorem A.2.
Suppose
is compact. Then
.
Suppose now that
compactifies to a manifold
with non-empty boundary
. In this case
admits non-trivial deformations, so we cannot expect
to be trivial. If
is a subgroup of
, the natural injection
induces a map on cohomology
For
we denote by
the cyclic subgroup generated by
. If
, we set
The naïve correspondence between
and the tangent space of
at
induces an identification of
with the space of infinitesimal deformations of
through holonomies for complete structures on
. Thus the following result [?, ?] can be considered an infinitesimal version of Mostow-Prasad's rigidity Theorem for complete finite-volume hyperbolic
-manifolds:
Theorem A.3.
We have
.
Dipartimento di Matematica, Universita di Pisa, Largo B. Pontecorvo 5, 56127 Pisa, Italy E-mail address : frigerio@mail.dm.unipi.it