2000 Mathematics Subject Classification. 58H15 (primary), 57M50, 20G10 (secondary).Supported by the INTAS project “CalcoMet-GT” 03-51-3663.
<ph f="cmbx">On deformations of hyperbolic </ph> <math xmlns="http://www.w3.org/1998/Math/MathML"> <mn>3</mn> </math> <ph f="cmbx">-manifolds with geodesic boundary</ph>

Roberto Frigerio

Dipartimento di Matematica, Universita di Pisa, Largo B. Pontecorvo 5, 56127 Pisa, Italy E-mail address : frigerio@mail.dm.unipi.it
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 3   -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 3   -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 3   -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 N   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 N   is hyperbolic). It is well-known that N   consists of a compact portion containing N   together with several cusps of the form T × [ 0 , )   , where T   is the torus, so N   admits a natural compactification N ¯   obtained by adding some boundary tori. Since the components of N   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 ( Δ , )   , where Δ   is a tetrahedron and   is a set of vertices of Δ   , that will be called ideal vertices. In the sequel we will always refer to Δ   itself as a partially truncated tetrahedron, tacitly implying that   is also fixed. The topological realization Δ *   of Δ   is obtained by removing from Δ   the ideal vertices and small open stars of the non-ideal vertices. We call lateral hexagon and truncation triangle the intersection of Δ *   respectively with a face of Δ   and with the link in Δ   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 Δ *   are called internal edges. If Δ   has ideal vertices, a lateral hexagon of Δ *   may not be a hexagon, because some of its closed edges may be missing.
A geometric realization of Δ   is an identification of Δ *   with a convex polyhedron in H 3   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 N ¯   be a compact orientable manifold and let N   be obtained from N ¯   by removing the toric components of N ¯   . We define a partially truncated triangulation of N   to be a realization of N   as a gluing of some Δ *   's along a pairing of the lateral hexagons induced by a simplicial pairing of the faces of the Δ   's. When N   is endowed with a hyperbolic structure, a partially truncated triangulation of N   is called geometric if, for each tetrahedron Δ   of the triangulation, the pull-back to Δ *   of the Riemannian metric of N   defines a geometric realization of Δ   . Equivalently, the hyperbolic structure of N   should be obtained by gluing geometric realizations of the Δ   's along isometries of their lateral hexagons.
Definition 1.1. A partially truncated triangulation T   of an orientable 3   -manifold N   is good if any tetrahedron in T   has at most one ideal vertex.
Kojima proved in [? that every hyperbolic N   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 5192   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 5108   cases. These facts support the following:
Conjecture 1.2. Any hyperbolic N   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 Δ   be a partially truncated tetrahedron and let Δ ( 1 )   be the set of edges of Δ   . The geometric realizations of Δ   are parameterized up to isometry by the dihedral angle assignements θ : Δ ( 1 ) ( 0 , π )   such that for each vertex v   of Δ   , if e 1 , e 2 , e 3   are the edges that emanate from v   , then θ ( e 1 ) + θ ( e 2 ) + θ ( e 3 )   is equal to π   for ideal v   and less than π   for non-ideal v   .
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
cosh a 1 = ( cos α 2 cos α 3 + cos α 1 ) / ( sin α 2 sin α 3 ) ,
cosh b 1 = ( cosh c 2 cosh c 3 + cosh c 1 ) / ( sinh c 2 sinh c 3 ) .
Let now Δ   be a partially truncated tetrahedron with edges e 1 , , e 6   as in Fig.  3 . We fix a geometric realization θ   of Δ   determined by the dihedral angles θ i = θ ( e i )   for i = 1 , , 6   and we denote by L θ   the length with respect to this realization. The boundary edges of the lateral hexagons of Δ   correspond to the pairs of distinct non-opposite edges { e i , e j }   , and will be denoted by e i j   . Note that e i j   disappears towards infinity, so it has length 0   , when the common vertex of e i   and e j   is ideal. Lemma  1.4 readily implies
cosh L θ ( e 12 ) = ( cos θ 1 cos θ 2 + cos θ 3 ) / ( sin θ 1 sin θ 2 ) . (1)
Note that this result is correct also when the common end of e 1   and e 2   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 v i j k   as the vertex from which the edges e i , e j , e k   emanate, we set:
c θ ( e 1 ) = cos θ 1 ( cos θ 3 cos θ 6 + cos θ 2 cos θ 5 ) + cos θ 2 cos θ 6 + cos θ 3 cos θ 5 + cos θ 4 sin 2 θ 1 ;
d θ ( v 123 ) = 2 cos θ 1 cos θ 2 cos θ 3 + cos 2 θ 1 + cos 2 θ 2 + cos 2 θ 3 1 .
Proposition 1.5. d θ ( v 123 ) = 0   if and only if the vertex v 123   is ideal. If v 123   and v 156   are both non-ideal then
cosh L θ ( e 1 ) = c θ ( e 1 ) / d θ ( v 123 ) d θ ( v 156 ) . (2)
Remark 1.6. Let Δ   be a partially truncated tetrahedron without ideal vertices. Then the geometric realizations of Δ   are parameterized by the lengths of the internal edges. In fact, the map that associates to the dihedral angles of a geometric realization of Δ   the lengths of its internal edges is a diffeomorphism between open subsets of R 6   .

1.4 Conditions for geometric gluing

Let N   be obtained from an orientable compact N ¯   by removing all the tori in N ¯   and let T   be a partially truncated triangulation of N   . Let also θ   be a geometric realization of the tetrahedra in T   and denote by L θ   the length with respect to this realization. We now describe the conditions under which the realization θ   defines a hyperbolic structure on the whole of N   . For our purposes it will be sufficient to deal only with good triangulations, so we assume from now on that T   is good. The general case is treated in [?.
In order to define a global hyperbolic structure on N   , the tetrahedra of T   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 2 π   around each edge of the manifold. The first condition ensures that the hyperbolic structure defined by θ   on the complement of the 2   -skeleton of T   extends to the complement of the 1   -skeleton. Since T   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 T   were not good, requiring the dihedral angles around each edge to sum up to 2 π   would not be sufficient to obtain a non-singular hyperbolic metric on the 1   -skeleton of T   . The point is that when some geometric tetrahedra are arranged one after the other around an edge e   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 e   instead of being the identity. Of course the isometry has to be the identity if at least one endpoint of e   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 T   is a good triangulation, a non-compact lateral hexagon F   is actually a pentagon with four right angles and an ideal vertex: we shall say in this case that F   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 F 126 *   as in Fig.  4 , and recall that the hexagon is oriented and embedded in H 3   by θ   . We consider the horospheres O 1   and O 2   centred at e 12   and passing through the non-ideal ends of e 1   and e 2   respectively. We define σ θ ( F 126 )   to be ± d i s t ( O 1 , O 2 )   , the sign being positive if e 2 , e 12 , e 1   are arranged positively on F 126 *   and O 1   is contained in the horoball bounded by O 2   , or if e 2 , e 12 , e 1   are arranged negatively on F 126 *   and O 2   is contained in the horoball bounded by O 1   , and negative otherwise.
Together with equation ( 1 ), the following proposition allows to compute σ   in terms of the dihedral angles.
Proposition 1.8. We have
σ θ ( F 126 ) = ln ( sinh L θ ( e 16 ) / sinh L θ ( e 26 ) ) . (3)
Proof: Let α 1 , α 2   and R , R 1 , R 2   be the angles and lengths shown in Fig.  5 . An easy computation in the upper half-plane model of H 2   shows that L θ ( e i 6 ) = ln cot ( α i / 2 )   for i = 1 , 2   . Moreover we have R = R 1 tan α 1 = R 2 tan α 2   , so exp σ θ ( F 126 ) = R 1 / R 2 = tan α 2 / tan α 1   .

Now for i = 1 , 2   we have tan α i = 2 cot ( α i / 2 ) / ( cot 2 ( α i / 2 ) 1 ) = 2 exp L θ ( e i 6 ) / ( exp ( 2 L θ ( e i 6 ) ) 1 )   . Combining these equations we finally get exp σ θ ( F 126 ) = sinh L θ ( e 16 ) / sinh L θ ( e 26 )   .   We now define θ ( F 126 )   to be the length of e 6   . The next proposition shows that the functions σ   and   provide a parameterization of isometry classes of exceptional hexagons.
Proposition 1.9. Let F   and F   be paired exceptional lateral hexagons. Their pairing can be realized by an isometry if and only if σ θ ( F ) + σ θ ( F ) = 0   and θ ( F ) = θ ( F )   .
Proof: We concentrate on the “if ” part of the statement, the “only if ” part being obvious.
Let f 1 , f 2   and f 1 , f 2   be the boundary edges of of F   and F   respectively, and assume that the orientation-reversing pairing between F   and F   glues f 1   to f 2   and f 1   to f 2   . Since θ ( F ) = θ ( F )   , Lemma  1.4 easily implies that L ( f 1 ) L ( f 2 )   if and only if L ( f 2 ) L ( f 1 )   .
Moreover the assumption that σ θ ( F ) = σ θ ( F )   and Proposition  1.8 give L ( f 1 ) L ( f 2 )   if and only if L ( f 2 ) L ( f 1 )   . This forces L ( f 1 ) = L ( f 2 )   and L ( f 2 ) = L ( f 1 )   , whence the conclusion.  
Remark 1.10. Let Δ   be a partially truncated tetrahedron and suppose that v   is the unique ideal vertex of Δ   . Let F 1 , F 2 , F 3   be the faces of Δ   incident to v   and for i = 1 , 2 , 3   let e i   be the edge of F i   not containing v   . The isometry classes of the geometric realizations of Δ   are parameterized by the lengths of the e i   's and the values taken by σ   on the F i   's. More precisely, the map that associates to any geometric realization θ   of Δ   the point ( L θ ( e 1 ) , L θ ( e 2 ) , L θ ( e 3 ) , σ θ ( F 1 ) , σ θ ( F 2 ) , σ θ ( F 3 ) )   defines a diffeomorphism between open subsets of two affine hyperplanes of R 6   .

1.6 Consistency equations

Recall now that we are considering a candidate hyperbolic 3   -manifold N   endowed with a good triangulation T   , and that we have fixed a geometric realization θ   of the tetrahedra in T   . The above discussion implies the following:
Theorem 1.11. The parameterization θ   defines on N   a hyperbolic structure with geodesic boundary if and only if the following conditions hold:
  • (1) the total dihedral angle along any edge of T   in N   is equal to 2 π   ;
  • (2) L θ ( e ) = L θ ( e )   for all pairs ( e , e )   of matching compact internal edges;
  • (3) σ θ ( F ) + σ θ ( F ) = 0   for all pairs ( F , F )   of matching exceptional hexagons.
By Propositions  1.5 ,  1.8 , conditions (1), (2), (3) of Theorem  1.11 translate into a set C ( T )   of smooth equations on θ   , which are called consistency equations. Our next task is to compare the number of equations in C ( T )   with the dimension of the moduli space of geometric realizations of the tetrahedra of T   .
Let c   (resp. p   ) be the number of compact (resp. non-compact) tetrahedra of T   . By Theorem  1.3 , the number of parameters for the geometric realizations of T   is equal to 6 c + 5 p = 6 t p   , where t   is the total number of tetrahedra of T   . Let e 1 , , e l   be the edges of T   without ideal endpoints, and for i = 1 , , l   let x i   be the valence of e i   , i. e. the number of tetrahedra of T   incident to e i   , with multiplicity. Of course we have i = 1 l x i = 6 c + 3 p   . If we denote by h   the number of edges of T   with exactly one ideal endpoint, then the union of the boundary tori of N ¯   admits a triangulation with h   vertices and p   triangles, so h = p / 2   . Let us now turn to the number of equations. Theorem  1.11 determines l + h   equations arising from conditions (1), i = 1 l ( x i 1 )   equations arising from conditions (2) and 3 p / 2   equations arising from conditions (3). Since
l + h + ( i = 1 l ( x i 1 ) ) + 3 p / 2 = h + ( i = 1 l x i ) + 3 p / 2
= p / 2 + 6 c + 3 p + 3 p / 2 = 6 c + 5 p ,
we can conclude that the number of equations in C ( T )   is equal to the dimension of the moduli space of geometric realizations of the tetrahedra of T   .

1.7 Reducing the number of equations

When N   has cusps, some equations in C ( T )   turn out to be redundant. Let T   be a fixed toric component of N ¯   and let j   be the number of tetrahedra asymptotic to T   . Since any such tetrahedron contributes to the triangulation of T   with a Euclidean triangle, the sum of all the dihedral angles along all the edges of T   incident to T   is equal to j π   . So if we require condition (1) of Theorem  1.11 to hold for all but one edge incident to T   , then the same condition is automatically satisfied also along the remaining edge. This allows us to discard from C ( T )   one equation for each cusp of N   .
Moreover, let Δ   be a partially truncated tetrahedron with an ideal vertex v   incident to T   , and let F 1 , F 2 , F 3   be the faces of Δ   incident to v   . By the very definition of σ   it follows that σ ( F 1 ) + σ ( F 2 ) + σ ( F 3 ) = 0   . This implies that if   is the set of all the faces incident to T   of tetrahedra of T   , then we have F σ ( F ) = 0   . So if we require condition (3) of Theorem  1.11 to hold for all but one pair of matching exceptional hexagons incident to T   , then the same condition also holds for the remaining pair. This means that another equation of C ( T )   for each cusp of N   can be discarded. Supppose that N   compactifies to an orientable N ¯   with k   boundary tori. The above discussion is summarized by the following:
Proposition 1.12. We can discard 2 k   equations from C ( T )   thus obtaining an equivalent set C * ( T )   of n 2 k   equations, where n   is the dimension of the moduli space of geometric realizations of the tetrahedra of T   .
We have seen in the preceding subsection that the moduli space of geometric realizations of the tetrahedra of T   is given by a subset W   of R 6 t   , where t   is the number of tetrahedra of T   . More precisely, W   is an open convex subset of an affine subspace of dimension 6 t p   , where p   is the number of non-compact tetrahedra in T   . Recall that l   is the number of compact edges (considered as subsets of N   ) of T   . The above computation implies that equations in C * ( T )   corresponding to conditions (1) of Theorem  1.11 take the form A ( x ) = 0   , where A : W R l + p / 2 k   is an affine map, while equations corresponding to conditions (2) and (3) take the form F ( x ) = 0   , where F : W R 6 t l ( 3 / 2 ) p k   is constructed from formulae ( 2 ), ( 3 ), and is therefore smooth. From now on we denote by Ω ( T ) = F 1 ( 0 ) A 1 ( 0 ) W R 6 t   the set of solutions of consistency equations C * ( T )   .

1.8 Completeness

Let T 1 , , T k   be the boundary tori of N ¯   . From now on we denote by μ i , λ i   a fixed basis of H 1 ( T i ; Z ) = π 1 ( T i )   , i = 1 , , k   . Any point in Ω ( T )   naturally defines an A f f ( C )   -structure on T i   (see e. g. [?, ?). For x Ω ( T )   , we denote by a i ( x ) C   (resp. by b i ( x ) C   ) the linear component of the holonomy of μ i   (resp. of λ i   ) corresponding to the A f f ( C )   -structure defined by x   on T i   . It is well-known that the hyperbolic structure defined by x   on N   induces a complete metric on the i   -th cusp of N   if and only if a i ( x ) = b i ( x ) = 1   .
Moreover, one can explicitly compute a i   and b i   in terms of the dihedral angles as follows.
Let Δ   be a tetrahedron in T   , let v   be an ideal vertex of Δ   and L x ( v )   be the (similarity class of the) Euclidean triangle obtained by intersecting the geometric realization of Δ   parameterized by x   with a small horosphere centred at v   . The tetrahedron being oriented, this triangle is also oriented, so, once a vertex p   of L x ( v )   is fixed, we can associate to the similarity structure of L x ( v )   the unique complex number z x ( L ( v ) , p )   such that L x ( v )   is carried to the Euclidean triangle with vertices 0 , 1 , z x ( L ( v ) , p )   by an orientation-preserving similarity sending p   to 0   . Suppose that e 1 , e 2 , e 3   are the internal edges emanating form v   , and that they are positively arranged around v   . If p = L ( v 123 ) e 1   , then z x ( L ( v 123 ) , p ) = ( sin θ 2 / sin θ 3 ) e i θ 1 .   If γ   is an oriented simplicial loop on T i   and q   is a vertex of γ   , then the set of all triangles touching γ   in q   and lying on the right of γ   is well-defined and will be denoted by R ( γ , q )   .
Moreover, we shall denote by V ( γ )   the set of vertices of γ   . Let μ ^ i , λ ^ i   be simplicial loops on T i   representing μ i , λ i   . The following result is proved in [?, ?.
Proposition 1.13. We have
a i ( x ) = ( 1 ) # V ( μ ^ i ) q V ( μ ^ i ) T R ( μ ^ i , q ) z x ( T , q ) ,
b i ( x ) = ( 1 ) # V ( λ ^ i ) q V ( λ ^ i ) T R ( λ ^ i , q ) z x ( T , q ) .
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 Ω ( T )   that defines on N   a complete hyperbolic structure with geodesic boundary.
Proof: See [?.  

2 Smoothness at the complete structure

Suppose now that x 0   is the unique point in Ω ( T )   which defines on N   a complete hyperbolic structure. For i = 1 , , k   and x Ω ( T )   let us define u i ( x ) = ln a i ( x ) , v i ( x ) = ln b i ( x ) ,   where ln   is the branch of the complex logarithm defined on { z C : ( z ) > 0 }   such that ln 1 = 0   . 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 x 0   in Ω ( T )   we have u i ( x ) = 0     v i ( x ) = 0     the hyperbolic structure defined by x   on the i   -th cusp of N   is complete.
Let F : W R 6 t l ( 3 / 2 ) p k   , A : W R l + p / 2 k   be the smooth functions previously defined such that Ω ( T ) = F 1 ( 0 ) A 1 ( 0 )   . We now set G : W R 6 t p 2 k × C k , G ( x ) = ( F ( x ) , A ( x ) , u 1 ( x ) , , u k ( x ) ) .   By Theorem  1.14 and Proposition  2.1 we have G 1 ( 0 ) = { x 0 }   . This section is entirely devoted to the proof of our main result:
Theorem 2.2. We have K e r d G x 0 = { 0 }   . Thus:
  • (1) G   induces a diffeomorphism of an open neighbourhood of x 0   in W   onto an open neighbourhood of 0   in R 6 t p 2 k × C k   ;
  • (2) Ω ( T )   is a smooth manifold of real dimension 2 k   near x 0   ;
  • (3) the map u : Ω ( T ) C k , u ( x ) = ( u 1 ( x ) , , u k ( x ) )   induces a diffeomorphism of an open neighbourhood of x 0   in Ω ( T )   onto an open neighbourhood of 0   in C k   .
Any point in Ω c ( T ) : = F 1 ( 0 )   defines a cone structure on N   having cone singularities along the edges of T   . The following immediate consequence of Theorem  2.2 was also proved in [?, ?.
Corollary 2.3. If N   is compact then Ω c ( T )   is parameterized in a neighbourhood of x 0   by the cone angles along the edges of T   .

2.1 Deforming cone structures

We begin with the following:
Proposition 2.4. The tangent map d F x 0 : T x 0 ( W ) R 6 t l ( 3 / 2 ) p k   has maximal rank, so Ω c ( T )   is a manifold of real dimension l + p / 2 + k   near x 0   .
Proof: By Remarks  1.6 ,  1.10 , the lengths of the compact internal edges and the values taken by σ   on the exceptional lateral hexagons of the tetrahedra of T   provide smooth coordinates on W   . It is easily seen that with respect to these coordinates the map F   is affine and has maximal rank at x 0   (whence at any point of W   ).   Let now α : ( ɛ , ɛ ) Ω c ( T )   be a smooth arc with α ( 0 ) = x 0   . We will study the deformation associated to α   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 N   the non-compact manifold obtained by drilling from N   all the edges of T   . For any t ( ɛ , ɛ )   the point α ( t )   determines a smooth hyperbolic structure M t   on N   , whose completion gives a hyperbolic cone structure on N   . First of all we describe how to deduce the shape of the geometric tetrahedra corresponding to the point α ( t )   just from the geometric structure M t   on N   . To this end we fix for t ( ɛ , ɛ )   a developing map D t : N ~ H 3   with associated holonomy representation ρ t : π 1 ( N ) P S L ( 2 , C )   (note that we can choose D t   and ρ t   to vary smoothly with t   ). Recall that N   compactifies to a manifold N ¯   with k   boundary tori, and denote by N ¯   the non-compact manifold obtained by drilling from N ¯   the closed properly embedded arcs corresponding to the edges of T   .
For i = 1 , , k   let T i   be the punctured torus T i = T i N ¯   , and denote by S p a r   the family of all the boundary components of the universal covering of N ¯   projecting to some T i   . We say that P π 1 ( N ) = π 1 ( N ¯ )   is a parabolic peripheral subgroup of π 1 ( N )   if P   is the stabilizer of some boundary component S S p a r   of the universal covering of N ¯   (so ρ 0 ( P )   is a Z + Z   parabolic subgroup of P S L ( 2 , C )   ). Let g d N ¯   be the portion of N ¯   corresponding to the geodesic boundary of N   , i. e. let g d N ¯ = N ¯ N   , and denote by S g d   the family of all the boundary components of the universal covering of N ¯   projecting to some component of g d N ¯   . We say that P π 1 ( N )   is a Fuchsian peripheral subgroup of π 1 ( N )   if P   is the stabilizer of some component S S g d   (so ρ 0 ( P )   is a Fuchsian subgroup of P S L ( 2 , C )   ).

2.2 Dual vectors to planes and horospheres

If S S g d   , by construction the image of S   under the developing map D t   is contained in a totally geodesic immersed surface in H 3   . It is easily seen that such a surface must in turn be contained in a geodesic plane S ^ ( t )   of H 3   . We now need to associate to each S ^ ( t )   a suitable ultra-ideal point b S ( t )   , which will be called the dual point of S ^ ( t )   . Such point naturally lies in 4-dimensional Minkowsky space, so we fix some notation about this space.
We denote by M 3 + 1   the space R 4   with coordinates x 0 , x 1 , x 2 , x 3   endowed with the Lorentzian inner product x , y = x 0 y 0 + x 1 y 1 + x 2 y 2 + x 3 y 3   . We set
3 = { x M 3 + 1 : x , x = 1 , x 0 > 0 } ,
3 + = { x M 3 + 1 : x , x = 1 } ,
L 3 + = { x M 3 + 1 : x , x = 0 , x 0 > 0 } .
We recall that 3   is the upper sheet of the two-sheeted hyperboloid, and that ,   restricts to a Riemannian metric on 3   . With this metric, 3   is the so-called hyperboloid model H 3 h y p   of hyperbolic space. The one-sheeted hyperboloid 3 +   turns out to have a bijective correspondence with the set of hyperbolic half-spaces in H 3 h y p   . Given b 3 +   , the corresponding half-space, called the dual of b   , is given by { v 3 : v , b 0 }   .
Similarly, the cone L 3 +   of future-oriented light-like vectors of M 3 + 1   corresponds to the set of horospheres in H 3 h y p   . The horosphere dual to u L 3 +   is given by { v 3 : v , u = 1 }   .
Note now that for any S S g d   the set D t ( N ~ )   locally lies on a definite side of S ^ ( t )   , so we can define b S ( t )   to be the dual vector to the half-space that locally contains D t ( N ~ )   and is bounded by S ^ ( t )   .
When S   belongs to S p a r   a vector b S ( t ) L 3 +   can also be defined as follows: take an oriented edge f ~   of T ~   ending in S   and set b S ( t )   to be the unique point in L 3 +   with x 0 ( b S ( t ) ) = 1   which projects to the endpoint of D t ( f ~ )   in H 3   . Such an endpoint exists because D t ( f ~ )   is a geodesic, and is clearly independent of the choice of f ~   , so b S ( t )   is indeed well-defined. Since developing maps vary smoothly with respect to t   we have the following:
Lemma 2.5. For any S S g d S p a r   the map t b S ( t )   is smooth with respect to t   .

2.3 Lifting geometric tetrahedra

Let now Δ   be a tetrahedron in T   , and denote by Δ ~   a lift of Δ   to N ~   . Let S i   be the boundary component of the universal covering of N ¯   that corresponds to v i   , where v 1 , , v 4   are the vertices of Δ ~   . If Δ ¯ ( t )   is the convex hull of b S 1 ( t ) , , b S 4 ( t )   in M 3 + 1   , then projecting Δ ¯ ( t )   to H 3   and truncating its infinite-volume ends with the corresponding S ^ i ( t )   's gives back a truncated tetrahedron Δ * ( t )   isometric to the geometric realization of Δ   parameterized by α ( t )   . It is easily seen that the dihedral angles of Δ * ( t )   smoothly depend on b S 1 ( t ) , , b S 4 ( t )   , so Theorem  2.2 is now reduced to the following:
Proposition 2.6. If α ˙ ( 0 ) K e r d G x 0   , then we can choose the D t   's in such a way that b ˙ S ( 0 ) = 0   for all S S g d S p a r   .

2.4 The tangent vector to ρ t  

Let us consider the double D N   of N   obtained by mirroring N   along its boundary (which is now given by some punctured surfaces of negative Euler characteristic). Since N   is totally geodesic with respect to the hyperbolic structure M t   , this structure can be doubled to a smooth hyperbolic metric D M t   on D N   .
Let D ρ t : π 1 ( D N ) P S L ( 2 , C )   be a holonomy representation corresponding to a developing map for D M t   . It is easily seen that we can assume D ρ t   to vary smoothly with respect to t   . We set D ρ ˙ : π 1 ( D N ) s l ( 2 , C ) , D ρ ˙ ( γ ) = d d t | t = 0 ( D ρ t ( γ ) D ρ 0 ( γ ) 1 ) .   As explained in Appendix  A , we have D ρ ˙ Z 1 ( π 1 ( D N ) ; s l ( 2 , C ) ; D ρ 0 )   . This subsection is devoted to the proof of the following:
Proposition 2.7. If α ˙ ( 0 ) K e r d G x 0   , then D ρ ˙ B 1 ( π 1 ( D N ) ; s l ( 2 , C ) ; D ρ 0 )   .
From now on we suppose α ˙ ( 0 ) K e r d G x 0   . Let f 1 , , f m D N   be the doubles of the edges of T   and for all j = 1 , , m   let j   be a small loop in D N   encircling f j   . We denote by γ j   an element in π 1 ( D N )   representing j   (such a γ j   is well-defined only up to conjugation). Now our hypothesis implies that if θ j ( t )   is the cone angle of (the completion of ) D M t   along f j   , then θ ˙ j ( 0 ) = 0   . Also observe that we have D ρ t ( γ j ) = g t [ ( exp ( i θ j ( t ) / 2 ) 0 0 exp ( i θ j ( t ) / 2 ) ) ] g t 1   for some smooth path g : ( ɛ , ɛ ) P S L ( 2 , C )   . Differentiating this relation we easily get D ρ ˙ ( γ j ) = 0   . Let K   be the kernel of the map π 1 ( D N ) π 1 ( D N )   induced by the inclusion, and observe that K   is the smallest normal subgroup of π 1 ( D N )   generated by γ 1 , , γ m   . Let D ρ ¯ 0 : π 1 ( D N ) P S L ( 2 , C )   be the natural representation associated to D ρ 0   . Since D ρ 0 ( γ ) = 1   and D ρ ˙ ( γ ) = 0   for all γ K   , Lemma  A.1 implies the following:
Proposition 2.8. D ρ ˙   naturally defines a cocycle z ρ Z 1 ( π 1 ( D N ) ; s l ( 2 , C ) ; D ρ ¯ 0 )   .
Moreover, D ρ ˙   belongs to B 1 ( π 1 ( D N ) ; s l ( 2 , C ) ; D ρ 0 )   if and only if z ρ   belongs to B 1 ( π 1 ( D N ) ; s l ( 2 , C ) ; D ρ ¯ 0 )   .
Now if N   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 N   is non-compact, then [ z ρ ] H p a r 1 ( π 1 ( D N ) ; s l ( 2 , C ) ; D ρ ¯ 0 )   .
Proof: Let γ π 1 ( D N )   be such that D ρ ¯ 0 ( γ )   is non-trivial parabolic, and let γ   be the infinite cyclic group generated by γ   . We have to check that z ρ   restricts to a coboundary in B 1 ( γ ; s l ( 2 , C ) ; D ρ ¯ 0 i )   , where i : γ ¯ π 1 ( D N )   is the natural inclusion.
Without loss of generality we can suppose γ π 1 ( T i ) π 1 ( N ) π 1 ( D N )   for some i = 1 , , k   . Recall that a preferred element μ i π 1 ( T i )   was previously fixed, set u i ( t ) = u i ( α ( t ) )   and observe that since α ˙ ( 0 ) K e r d G x 0   we have u ˙ i ( 0 ) = 0   . Let μ i , γ   be elements in π 1 ( T i )   projecting respectively to μ i , γ   . By Lemma  2.5 a smooth path g : ( ɛ , ɛ ) P S L ( 2 , C )   exists such that both g t 1 ρ t ( μ i ) g t   and g t 1 ρ t ( γ ) g t   fixes H 3   for t ( ɛ , ɛ )   , so that
D ρ t ( μ i ) = g t [ ( exp ( u i ( t ) / 2 ) τ i ( t ) 0 exp ( u i ( t ) / 2 ) ) ] g t 1 ,
D ρ t ( γ ) = g t [ ( a ( t ) b ( t ) 0 a ( t ) 1 ) ] g t 1 ,
where τ i , a , b : ( ɛ , ɛ ) C   are smooth arcs with a ( 0 ) = 1 , b ( 0 ) 0 , τ i ( 0 ) 0   .
Since ρ 0 ( π 1 ( T i ) ) = Z + Z   is Abelian, an element k K   exists such that μ i γ = k γ μ i   in π 1 ( D N )   . By Proposition  2.8 , this readily implies d d t | t = 0 ( D ρ t ( μ i ) D ρ t ( γ ) ) = d d t | t = 0 ( D ρ t ( γ ) D ρ t ( μ i ) ) ,   which after some computations gives a ˙ ( 0 ) = 0   (here we use u ˙ i ( 0 ) = 0   ). Let us consider the deformation φ t : γ P S L ( 2 , C )   defined by φ t ( ( γ ) n ) = D ρ t ( ( γ ) n )   for any n Z   .
We claim that φ ˙ = 0   in H 1 ( γ ; s l ( 2 , C ) ; φ 0 )   . This will easily give that z ρ   restricts to a coboundary in B 1 ( γ ; s l ( 2 , C ) ; D ρ ¯ 0 i )   , whence the conclusion. Since derivatives of conjugated deformations differ by a coboundary, we can suppose φ t ( γ ) = [ ( a ( t ) b ( t ) 0 a ( t ) 1 ) ] .   Setting v = ( b ˙ ( 0 ) 2 b ( 0 ) 0 0 b ˙ ( 0 ) 2 b ( 0 ) ) s l ( 2 , C ) ,   an easy computation shows that φ ˙ ( γ ) = v A d ( φ 0 ( γ ) ) ( v )   . This readily implies that φ ˙   is a coboundary.  

2.5 The final step

For γ π 1 ( D N )   let t r γ : ( ɛ , ɛ ) C   be the map defined as follows: t r γ ( t ) = t r a c e ( D ρ ^ t ( γ ) )   , where t D ρ ^ t ( γ ) S L ( 2 , C )   is a smooth lift of t D ρ t ( γ ) P S L ( 2 , C )   (so t r γ   is well-defined only up to the sign). Also recall that D t   and ρ t   are respectively a developing map and a holonomy representation for the hyperbolic structure M t   on N   .
The following result can be easily deduced from the proof of the previous proposition.
Lemma 2.10. Let γ π 1 ( D N )   be such that D ρ 0 ( γ )   is non-trivial parabolic.
Then t ¨ r γ ( 0 ) = 0   .
As a consequence of Proposition  2.7 and Lemma  2.10 we obtain the following:
Proposition 2.11. D t   and ρ t   can be chosen in such a way that d d t | t = 0 ρ t ( γ ) = 0   for all γ π 1 ( N )   . Moreover, if γ   is a non-trivial element of a peripheral parabolic subgroup of π 1 ( N )   , then t ¨ r γ ( 0 ) = 0   .
Let γ π 1 ( N )   be such that ρ 0 ( γ ) 1   and suppose r γ : ( ɛ , ɛ ) H 3   is a smooth path such that r γ ( t )   is a fixed point for ρ t ( γ )   for any t ( ɛ , ɛ )   . We want to study how the derivative of r ( t )   is related to the derivative of ρ t ( γ )   . We identify H 3   with C { }   and we set ρ t ( γ ) = [ ( a ( t ) b ( t ) c ( t ) d ( t ) ) ] .   Without loss of generality we can suppose that ρ 0 ( γ )   does not fix   and that r γ ( 0 ) = 0 C H 3   , so b ( 0 ) = 0   , c ( 0 ) 0   . An easy computation now shows that r γ ( t ) = ( a ( t ) d ( t ) ± ( t r γ ( t ) + 2 ) ( t r γ ( t ) 2 ) ) / ( 2 c ( t ) ) .   From this formula we can readily deduce the following lemmas.
Lemma 2.12. If ρ 0 ( γ )   is non-trivial loxodromic and ρ ˙ ( γ ) = 0   , then r ˙ γ ( 0 ) = 0   .
On the other hand, let ρ 0 ( γ )   be non-trivial parabolic. Also assume that ρ ˙ ( γ ) = 0   and t ¨ r γ ( 0 ) = 0   . Then r ˙ γ ( 0 ) = 0   .
We can now conclude the proof of Proposition  2.6 . Choose D t   and ρ t   as in the statement of Proposition  2.11 . Let S   be a boundary component of the universal covering of N ¯   which belongs to S p a r   , denote by P S   the stabilizer of S   in π 1 ( N )   and choose an element γ P S   with ρ 0 ( γ ) 1   . By construction the projection of b S ( t )   to H 3   is fixed by ρ t ( γ )   , so Lemma  2.12 applies ensuring b ˙ S ( 0 ) = 0   .
Suppose now that S   belongs to S g d   , and let P S   be the stabilizer of S   in π 1 ( N )   . For γ P S   with ρ 0 ( γ ) 1   let p γ ( t ) , q γ ( t )   be the fixed points of ρ t ( γ )   on H 3   . Note that since ρ 0 ( γ )   is loxodromic we can choose p γ , q γ   to vary smoothly with respect to t   , at least in a small neighbourhood of 0   . This gives p ˙ γ = q ˙ γ = 0   by Lemma  2.12 . Now a standard result in Kleinian group theory ensures that the set { p γ ( 0 ) , q γ ( 0 ) : γ P S , ρ 0 ( γ ) 1 }   is dense in the closure at infinity of D 0 ( S ) H 3   , and this easily implies b ˙ S ( 0 ) = 0   .

3 Dehn filling

Once the smoothness of Ω ( T )   at x 0   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 j { 1 , , k }   . Then there exists a complex number τ j   with non-zero imaginary part such that if { y n } n N Ω ( T )   is a sequence with lim n y n = x 0   and u j ( y n ) 0   for every n N   , then lim n v j ( y n ) / u j ( y n ) = τ j   .

3.1 Thurston's hyperbolic Dehn filling Theorem

Let U   be a sufficiently small neighbourhood of x 0   in Ω ( T )   and let x U   . For j = 1 , , k   , we define the j   -Dehn filling coefficient ( p j ( x ) , q j ( x ) ) R 2 { }   as follows: if u j ( x ) = 0   , then ( p j ( x ) , q j ( x ) ) =   ; otherwise, p j ( x ) , q j ( x )   are the unique real solutions of the equation p j ( x ) u j ( x ) + q j ( x ) v j ( x ) = 2 π i .   (Existence and uniqueness of such solutions near x 0   can be easily deduced from Theorem  2.2 and Lemma  3.1 .) Let us set d = ( d 1 , , d k ) : U k i = 1 S 2 , d j ( x ) = ( p j ( x ) , q j ( x ) ) S 2 = R 2 { } .   As a consequence of Theorem  2.2 and Lemma  3.1 we have the following:
Theorem 3.2. If U   is small enough, the map d   defines a diffeomorphism onto an open neighbourhood of ( , , )   in S 2 × × S 2   .
For x Ω ( T )   we denote by N ( x )   the hyperbolic structure induced on N   by x   , and by N ^ ( x )   the metric completion of N ( x )   . Recall that a preferred basis μ i , λ i   of H 1 ( T i ; Z )   is fixed for every i = 1 , , k   . We also set
I Ω ( T ) = { x U Ω ( T ) : f o r i = 1 , , k t h e i t h D e h n f i l l i n g c o e f f i c i e n t
a s s o c i a t e d t o x i s e q u a l e i t h e r t o o r t o a p a i r o f c o p r i m e i n t e g e r s } .
Theorem 3.3. If U   is sufficiently small and x   belongs to I Ω ( T ) U   , then N ^ ( x )   admits a complete finite-volume smooth hyperbolic structure which is obtained by adding to N ( x )   a closed geodesic at any cusp with non-infinite Dehn filling coefficient. From a topological point of view, N ^ ( x )   is obtained by Dehn filling the i   -th cusp of N   along the slope p i ( x ) μ i + q i ( x ) λ i   if ( p i ( x ) , q i ( x ) )   , and by leaving the i   -th cusp of N   unfilled if ( p i ( x ) , q i ( x ) ) =   , i = 1 , , k   .
Proof: See e. g. [?, ?, ?, ?.   The following proposition will prove useful in the last subsection.
Proposition 3.4. Let X   be any smooth manifold and let g : Ω ( T ) X   be a smooth map. Suppose that there exists a small neighbourhood U   of x 0   in Ω ( T )   such that for all x , x U I Ω ( T )   with d ( x ) = d ( x )   we have g ( x ) = g ( x )   . Then d g x 0 = 0   .
Proof: Since d : U i = 1 k S 2   defines a chart around x 0   , it is sufficient to observe that U I Ω ( T )   accumulates to x 0   along any direction in T x 0 Ω ( T )   .  

3.2 Infinitesimal deformations of compact tetrahedra

We know from Mostow's rigidity Theorem that compact hyperbolic 3   -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 T   near the complete structure affect compact tetrahedra only at the second order.
Let f   be any compact internal edge of T   . For x Ω ( T )   we denote by f ( x )   the length of f   with respect to the metric structure defined by x   . By Lemma  1.4 the map f : Ω ( T ) R   is smooth.
Proposition 3.5. We have d x 0 f = 0   .
Proof: Let x , x U I Ω ( T )   be such that d ( x ) = d ( x )   . Then the identity of N   extends to a homeomorphism between N ^ ( x )   and N ^ ( x )   . By Mostow-Prasad's rigidity Theorem, such a homeomorphism is homotopic to an isometry ψ : N ^ ( x ) N ^ ( x )   via a homotopy which preserves the geodesic boundary (see e. g. [?). For y U I Ω ( T )   let f ( y ) N ( y ) N ^ ( y )   be the geodesic segment corresponding to f   . From the above discussion it follows that ψ ( f ( x ) )   is homotopic to f ( x )   relatively to N ^ ( x )   . Since both ψ ( f ( x ) )   and f ( x )   intersect N ^ ( x )   perpendicularly, this easily implies that ψ ( f ( x ) ) = f ( x )   , whence f ( x ) = f ( x )   . 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 f   lying on a compact lateral hexagon of some tetrahedron of T   , and let f : Ω ( T ) R   be the function which associates to x Ω ( T )   the length of f   in the geometric realization parameterized by x   .
Then d x 0 f = 0   .
Corollary 3.7. Fix an internal edge f   in a compact tetrahedron of T   and let a f : Ω ( T ) R   be the function that associates to x Ω ( T )   the dihedral angle assigned to f   by x   . Then d a x 0 f = 0   .

3.3 Infinitesimal deformations of the geodesic boundary

Let T e i c h ( N )   be the Teichmüller space of hyperbolic structures on N   , i. e. the space of equivalence classes of hyperbolic metrics on N   , where two such metrics are considered equivalent if they are isometric through a diffeomorphism homotopic to the identity of N   . For x Ω ( T )   we denote by B ( x ) T e i c h ( N )   the equivalence class of the hyperbolic structure induced by N ( x )   on N   . It is well-known that T e i c h ( N )   admits a structure of differentiable manifold such that B : Ω ( T ) T e i c h ( N )   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 d B x 0 = 0   .

A Cohomology theory of representations

A.1 The tangent space to a representation

Let G   be a Lie group with associated Lie algebra g   and Γ   be any group, and denote by ( Γ , G )   the set of representations of Γ   in G   . We say that a path { ρ t ( Γ , G ) : t ( ɛ , ɛ ) }   is smooth if ρ t ( γ )   is a smooth function of t   for any γ Γ   . If { ρ t }   is a smooth path of representations, the tangent vector to the map t ρ t ( γ )   at 0   gives an element in T ρ 0 ( γ ) G   .
Identifying this tangent space with g = T 1 G   by right translation we get an element ρ ˙ ( γ )   of the Lie algebra g   : ρ ˙ ( γ ) = d d t | t = 0 ( ρ t ( γ ) ρ 0 ( γ ) 1 ) .   Differentiating the homomorphism relation ρ t ( γ 1 γ 2 ) = ρ t ( γ 1 ) ρ t ( γ 2 )   we see that ρ ˙ : Γ g   satisfies the so-called cocycle relation ρ ˙ ( γ 1 γ 2 ) = ρ ˙ ( γ 1 ) + A d ( ρ 0 ( γ 1 ) ) ( ρ ˙ ( γ 2 ) ) ,   where A d : G G L ( g )   is the usual adjoint representation.
Consider now a trivial deformation of ρ 0   , i. e. let t g t   be a smooth path in G   starting at the identity and set ρ t ( γ ) = g t ρ 0 ( γ ) g t 1   for all γ Γ   . Then differentiation shows that ρ ˙ ( γ ) = g ˙ A d ( ρ 0 ( γ ) ) ( g ˙ )   for any γ Γ   , where g ˙ g   is the tangent vector to t g t   at t = 0   . We now set
Z 1 ( Γ ; g ; ρ 0 ) = { c : Γ g : c ( γ 1 γ 2 ) = c ( γ 1 ) + A d ( ρ 0 ( γ 1 ) ) ( c ( γ 2 ) ) } ,
B 1 ( Γ ; g ; ρ 0 ) = { b : Γ g : b ( γ ) = m A d ( ρ 0 ( γ ) ) ( m ) f o r s o m e m g } ,
H 1 ( Γ ; g ; ρ 0 ) = Z 1 ( Γ ; g ; ρ 0 ) / B 1 ( Γ ; g ; ρ 0 ) .
The above discussion shows that Z 1 ( Γ ; g ; ρ 0 )   corresponds in some sense to the tangent space of ( Γ , G )   at ρ 0   . Under this identification the module B 1 ( Γ ; g ; ρ 0 )   should represent the tangent space to trivial deformations of ρ 0   , so H 1 ( Γ ; g ; ρ 0 )   should give the tangent space of ( Γ , G ) / G   at [ ρ 0 ]   (however, this holds true only in the setting of algebraic schemes).
Let c Z 1 ( Γ ; g ; ρ )   , and suppose that Γ 0   is a normal subgroup of Γ   such that ρ ( γ ) = 1 G   , c ( γ ) = 0   for all γ Γ 0   . Let also ρ ¯ : Γ / Γ 0 G   be the representation induced by ρ   .
Lemma A.1. The map c ¯ : Γ / Γ 0 M   defined by c ¯ ( [ γ ] ) = c ( γ )   is well-defined and gives a cocycle c ¯ Z 1 ( Γ / Γ 0 ; g ; ρ ¯ )   . Moreover, we have c ¯ B 1 ( Γ / Γ 0 ; g ; ρ ¯ )   if and only if c B 1 ( Γ ; g ; ρ )   .

A.2 Classical rigidity results

Let N   be a smooth 3   -manifold without boundary and suppose ρ 0 : π 1 ( N ) P S L ( 2 , C )   is the holonomy representation for a complete finite-volume hyperbolic structure on N   .
The following result is due to Weil [?, and can be considered as a local version of Mostow's rigidity Theorem for compact hyperbolic 3   -manifolds.
Theorem A.2. Suppose N   is compact. Then H 1 ( π 1 ( N ) ; s l ( 2 , C ) ; ρ 0 ) = 0   .
Suppose now that N   compactifies to a manifold N ¯   with non-empty boundary N ¯ = T 1 T k   . In this case ρ 0   admits non-trivial deformations, so we cannot expect H 1 ( π 1 ( N ) ; s l ( 2 , C ) ; ρ 0 )   to be trivial. If K   is a subgroup of π 1 ( N )   , the natural injection i K : K π 1 ( N )   induces a map on cohomology i K * : H 1 ( π 1 ( N ) ; s l ( 2 , C ) ; ρ 0 ) H 1 ( K ; s l ( 2 , C ) ; ρ 0 i K ) .   For γ π 1 ( B )   we denote by γ π 1 ( N )   the cyclic subgroup generated by γ   . If P = { γ π 1 ( N ) : ρ 0 ( γ ) i s p a r a b o l i c }   , we set H p a r 1 ( π 1 ( N ) ; s l ( 2 , C ) ; ρ 0 ) = γ P K e r i γ * H 1 ( π 1 ( N ) ; s l ( 2 , C ) ; ρ 0 ) .   The naïve correspondence between H 1 ( π 1 ( N ) ; s l ( 2 , C ) ; ρ 0 )   and the tangent space of ( π 1 ( N ) , P S L ( 2 , C ) ) / P S L ( 2 , C )   at [ ρ 0 ]   induces an identification of H p a r 1 ( π 1 ( N ) ; s l ( 2 , C ) ; ρ 0 )   with the space of infinitesimal deformations of ρ 0   through holonomies for complete structures on N   . Thus the following result [?, ? can be considered an infinitesimal version of Mostow-Prasad's rigidity Theorem for complete finite-volume hyperbolic 3   -manifolds:
Theorem A.3. We have H p a r 1 ( π 1 ( N ) ; s l ( 2 , C ) ; ρ 0 ) = 0   .
Dipartimento di Matematica, Universita di Pisa, Largo B. Pontecorvo 5, 56127 Pisa, Italy E-mail address : frigerio@mail.dm.unipi.it