Classification of non-free Kleinian groups generated by two parabolic transformations
Abstract.
We give a full proof to Agol’s announcement on the classification of non-free Kleinian groups generated by two parabolic transformations.
2010 Mathematics Subject Classification:
Primary 57M50, Secondary 57M251. Introduction
Motivated by knot theory, Riley studied Kleinian groups generated by two parabolic transformations (see [44, 45, 46, 47, 48]). In particular, the construction of the complete hyperbolic structure on the figure-eight knot complement [45] inspired Thurston to establish the uniformisation theorem of Haken manifolds. The space of marked subgroups of generated by two non-commuting parabolic transformations is parametrised by a non-zero complex number. There is an open set, , called the Riley slice of Schottky space, of Kleinian groups of this type that are free and discrete, and for which the quotient of the domain of discontinuity is a four times punctured sphere. For every group in , the Klein manifold (the quotient of union of the hyperbolic space and the domain of discontinuity) is homeomorphic to the complement of the 2-strand trivial tangle. Keen and Series [29] studied the Riley slice by applying their theory of pleating rays, and it was supplemented by Komori and Series [31]. Motivated by knot theory, Akiyoshi, Sakuma, Wada and Yamashita [6] studied the combinatorial structures of the Ford domains, by extending Jorgensen’s work [28] on punctured torus groups, which leads to a natural tessellation of (see Figure 0.2b in [6]). Ohshika and Miyachi [42] proved that the closure of is equal to the space of marked Kleinian groups with two parabolic generators which are free and discrete. Building on his joint work [24], [26] and [36] with Gehring, Hinkkanen and Marshall, respectively, Martin [35] identified the exterior of as the Julia set of a certain semigroup of polynomials and proved a “supergroup density theorem” for groups in the exterior of . The problem to detect freeness and non-freeness of (not necessarily discrete) groups generated by two non-commuting parabolic transformations has attracted attention of various researchers (see [33, 23, 56, 30] and references therein).
In this paper, we are interested in Kleinian groups that are in the complement of the closure of , namely the groups that are discrete but not free. The essential simple loops on the boundary of the complement of the 2-strand trivial tangle, which are not null homotopic in the ambient space, are parametrised by a slope in . The Heckoid groups, introduced by Riley [47] and formulated by Lee and Sakuma [32] following Agol [2], are Kleinian groups with two parabolic generators in which the element corresponding to the curve of slope has finite order. The most extreme case is the group where this element is the identity, in which case, the quotient of hyperbolic space by this group is the complement of a 2-bridge knot or link.
In [1, Theorem 4.3], Adams proved that a non-free and torsion-free Kleinian group is generated by two parabolic transformations if and only if the quotient hyperbolic manifold is homeomorphic to the complement of a -bridge link which is not a torus link. (We regard a knot as a one-component link.) This refines the result of Boileau and Zimmermann [11, Corollary 3.3] that a link in is a -bridge link if and only if its link group is generated by two meridians.
In 2002, Agol [2] announced the following classification theorem of non-free Kleinian groups generated by two parabolic transformations, which generalises Adams’ result. The main purpose of this paper is to give a full proof to this theorem.
Theorem 1.1.
A non-free Kleinian group is generated by two non-commuting parabolic elements if and only if one of the following holds.
-
(1)
is conjugate to the hyperbolic -bridge link group, , for some rational number , where and are coprime integers such that .
-
(2)
is conjugate to the Heckoid group, , for some and some .

In the remainder of the introduction, we explain the meaning of the theorem more precisely.
Recall that a -bridge link is a knot or a two-component link which is represented by a diagram in the - plane that has two maximal points and two minimal points with respect to the height function determined by the -coordinate. We may assume that the two maximal points and the two minimal points, respectively, have the same -coordinates. Such a diagram gives a plait (or plat) representation of the -bridge link consisting of two upper bridges, two lower bridges, and a -strand braid connecting the upper and lower bridges (see Figure 1(1)). The -bridge links are parametrized by the set , and the -bridge link corresponding to is denoted by and is called the -bridge link of slope (see Section 2 for the precise definition). If then is the -component trivial link, and if then is the trivial knot. If , where and are coprime integers, then is hyperbolic, i.e., admits a complete hyperbolic structure of finite volume, if and only if . In this case, there is a torsion-free Kleinian group , unique up to conjugation, such that is homeomorphic to the link complement as oriented manifold. We denote the Kleinian group , by , and call it the hyperbolic -bridge link group of slope .
The Heckoid groups were first introduced by Riley [47] as an analogy of the classical Hecke groups considered by Hecke [25]. The topological structure of their quotient orbifolds was worked out by Lee and Sakuma [32], following the description by Agol [2]. Specifically, they showed that the Heckoid groups are the orbifold fundamental groups of the Heckoid orbifolds illustrated in Figure 1(2)-(4). (See [7, 10, 19] for basic terminologies and facts concerning orbifolds.) These figures illustrate weighted graphs whose explicit descriptions are given by Definition 3.4. For each weighted graph in the figure, let be the pair of a compact -orbifold and a compact -suborbifold of determined by the rules described below. Let be the subgraph of consisting of the edges with weight , and let be the subgraph of consisting of the edges with integral weight.
-
(1)
The underlying space of the orbifold is the complement of an open regular neighbourhood of the subgraph .
-
(2)
The singular set of is , where the index of each edge of the singular set is given by the weight of the corresponding edge of .
-
(3)
For an edge of , let be the -suborbifold of defined as follows.
-
(a)
In Figure 1(2), consists of two annuli in whose cores, respectively, are meridians of the two edges of .
-
(b)
In Figure 1(3), consists of an annulus in whose core is a meridian of the single edge of .
-
(c)
In Figure 1(4), consists of two copies of the annular orbifold (the -orbifold with underlying space the disc and with two cone points of index ) in each of which is bounded by a meridian of an edge of .
-
(a)
By [32, Lemmas 6.3 and 6.6], the orbifold pair is a Haken pared orbifold (see Definition 3.1 or [10, Definition 8.3.7]) and admits a unique complete hyperbolic structure, which is geometrically finite (see Section 3 or [32, Proposition 6.7]). Namely there is a geometrically finite Kleinian group , unique up to conjugation, such that is isomorphic to the interior of the compact orbifold , such that represents the parabolic locus. The pair is also regarded as a relative compactification of the pair consisting of a non-cuspidal part of and its boundary (see Section 3).
We denote the pared orbifold by , , or according as it is described by the weighted graph in Figure 1(2), (3), or (4). We also denote the Kleinian group by .
Then the assertion (2) of the main Theorem 1.1 is equivalent to the following assertion (2’)
-
(2’)
is conjugate to the Kleinian group for some pared orbifold , , or in Definition 3.4.
Agol [2] also announced the following classification of parabolic generating pairs of the groups in Theorem 1.1, which refines and extends Adams’ results that every hyperbolic -bridge link group has only finitely many parabolic generating pairs [1, Corollary 4.1] and that the figure-eight knot group has precisely two parabolic generating pairs up to equivalence [1, Corollary 4.6].
Theorem 1.2.
(1) If is a hyperbolic -bridge link group, then it has precisely two parabolic generating pairs, up to equivalence.
(2) If is a Heckoid group, then it has a unique parabolic generating pair, up to equivalence.
Here, by a parabolic generating pair of a Kleinian group , we mean an unordered pair of parabolic transformations and that generate . Two parabolic generating pairs and are said to be equivalent if is equal to for some up to simultaneous conjugacy. In the companion [4] of this paper by Shunsuke Aimi, Donghi Lee, Shunsuke Sakai and the fourth author, an alternative proof of the theorem is given.
Theorems 1.1 and 1.2 are beautifully illustrated by a figure produced by Yasushi Yamashita upon request of Caroline Series, which is to be included in her article [52] in preparation. The figure is produced by using the results announced in [6, Section 3 of Preface]. (See also Figure 0.2b in [6], which was also produced by Yamashita.) For further properties of Heckoid groups, please see the article [5] in preparation.
This paper is organised as follows. In Section 2, we recall basic facts concerning -bridge links. In Section 3, we give the precise definitions of the Heckoid orbifolds and Heckoid groups. In Section 4, we give the classification of dihedral orbifolds, i.e., good orbifolds with dihedral orbifold fundamental groups (Theorem 4.1), which holds a key to the proof of the main theorem. In Section 5, we prove the relative tameness theorem for hyperbolic orbifolds (Theorem 5.1), following Bowditch’s proof of the tameness theorem for hyperbolic orbifolds ([15]). This theorem is used in the treatment of geometrically infinite two parabolic generator non-free Kleinian groups. In fact, it turns out there is no such groups. In Section 6, we introduce a convenient method for describing pared orbifolds (Convention 6.1) and the concept of an orbifold surgery (Definition 6.3), and then prove a simple but useful lemma for orbifold surgeries (Lemma 6.4). In Section 7, we follow Adams [1], and recall basic facts concerning two parabolic generator Kleinian groups, in particular an estimate of the length of parabolic generators with respect to the maximal cusp (Lemma 7.1). In Section 8, we give an outline of the proof of the main theorem. Sections 9, 10, and 11 are devoted to the proof of the main theorem. In the appendix, which consists of Sections 12 and 13, we give the classification of geometric dihedral orbifolds that is necessary for the proof Theorem 4.1.
Throughout this paper, we use the following notation.
Notation 1.3.
(1) For an orbifold , the symbol denotes the orbifold fundamental group of , denotes the abelianisation of , and denotes .
(2) For a natural number , denotes the cyclic group (or the ring) of order , and denotes the unit group of the ring .
(3) By a dihedral group, we mean a group generated by two elements of order . Thus it is isomorphic to the group for some . Note that has order or according to whether or . Note also that the order cyclic group is also regarded as a dihedral group.
Acknowledgement. M.S. would like to thank Ian Agol for sending the slide of his talk [2], encouraging him (and any of his collaborators) to write up the proof, and describing key ideas of the proof. He would also like to thank Michel Boileau for enlightening conversation in an early time. His sincere thanks also go to all the other authors for joining the project to give a proof to Agol’s announcement. J.P. would like to thank Sadayoshi Kojima for supporting his trip to Japan. H.A. was supported by JSPS Grants-in-Aid 19K03497. K.O. was supported by JSPS Grants-in-Aid 17H02843 and 18KK0071. M.S. was supported by JSPS Grants-in-Aid 15H03620.
2. Basic facts concerning -bridge links
In this section, we recall basic facts concerning -bridge links, which we use in the definitions of the Heckoid orbifolds and the Heckoid groups. The description of -bridge links given in this section is a mixture of those in [14, 51].
Let be the group of isometries of the Euclidean plane generated by the -rotations around the points in . Set and call it the Conway sphere. Then consists of four points in the -sphere . Let be the complementary -times punctured sphere. For each , let as be the simple loop in obtained as the projection of a line in of slope . Then is essential in , i.e., it does not bound a disc nor a once-punctured disc in . Conversely, any essential simple loop in is isotopic to for a unique : we call the slope of the essential loop. For each , let be the pair of mutually disjoint arcs in with , obtained as the image of the union of the lines in which intersect . Note that the union is a circle in containing , which divides into two discs and , where is the projection.

Let be the round -ball in , whose boundary contains the set consisting of the four marked points
Fix a homeomorphism satisfying the following conditions (see Figure 2).
-
(1)
maps the quadruple to the quadruple .
-
(2)
maps the circle to the equatorial circle , and maps the hemispheres and onto the hemispheres to , respectively.
-
(3)
is equivariant with respect to the natural -actions on and . Here the natural -action on is that which lifts to the group of isometries of the Euclidean plane generated by the -rotations around the points in , and the natural -action on is that generated by the -rotations about the coordinate axes of .
We identify with through the homeomorphism . Thus for , is regarded as an essential simple loop in , and is regarded as a union of two disjoint arcs in such that . Moreover, we can choose and so that they are -invariant.
For a rational number , let be a pair of arcs properly embedded in such that , which is obtained from by pushing its interior into . The pair is called the rational tangle of slope . We may assume is invariant by the natural -action on . In particular, the -axis intersects transversely in two points: Let be the subarc of the -axis they bound, and call it the core tunnel of (see Figure 4). Two meridional circles of near together with a subarc of forms a graph in homeomorphic to a pair of eyeglasses. This determines a canonical generating meridian pair of the rank free group .
By gluing the boundaries of the rational tangles and by the identity map, we obtain a link in the -sphere: we denote it by , and call it the -bridge link of slope . The number of components, , of is one or two (i.e., is a knot or a two-component link) according to whether the denominator is odd or even. The images of the core tunnels and in are called the upper tunnel and the lower tunnel of , respectively. We denote them by and , respectively. The canonical generating meridian pairs of and descend to generating meridian pairs of the link group : we call them the upper meridian pair and the lower meridian pair, respectively.
When we need to care about the orientation of the ambient -sphere , we regard as being obtained from and , where inherits the standard orientation of . In other words, we identify the ambient -sphere with the one-point compactification of , in such a way that the containing is identified with the original round ball via the identity map, whereas the containing is identified with via the inversion in . Thus . Under this orientation convention, a regular projection is read from the continued fraction expansion
in such a way that corresponds to the right-hand or left-hand half-twists according to whether is odd or even (see Figure 3).

The natural -actions on and can be glued to produce a -action on . Let and be the generators of the action whose restrictions to are the -rotations about the -axis and -axis, respectively (see Figure 4). We call , , and , respectively, the vertical involution, the horizontal involution, and the planar involution of . They are characterized by the following properties.
-
(1)
contains , whereas each of and intersects transversely in a single point.
-
(2)
The horizontal simple loop in is mapped by to itself preserving orientation, and it is mapped by to itself reversing orientation.

If the rational number satisfies the congruence , then admits an additional orientation-preserving symmetry which interchanges and . For a description of such symmetries, see e.g. [4, Sections 4 and 6], [49, Section 3].
We finally recall the classification theorem for -bridge links due to Schubert [53] (cf. [17, Chapter 12]).
Proposition 2.1.
For two rational numbers and , with and positive, the following holds.
(1) There is an orientation-preserving auto-homeomorphism of which maps to if and only if and either or . Moreover the following hold.
-
(a)
If and , then there there is an orientation-preserving auto-homeomorphism of which maps to and respects the -action. Moreover, the conjugate of the vertical involution of by is either the vertical or planar involution of , according to whether or .
-
(b)
If and , then there there is an orientation-preserving auto-homeomorphism of which maps to which respects the -action.
(2) There is an orientation-reversing auto-homeomorphism of which maps to if and only if and either or .
3. Heckoid orbifolds and Heckoid groups
In this section, we recall the definition of Heckoid orbifolds and Heckoid groups given by [32, Section 3].
Consider the quotient orbifold , where is the natural action illustrated in Figure 4. Note that its boundary is identified with , which is the quotient of by the group generated by the -rotations around the points in . Note that is identified with a normal subgroup of of index . For each , let be the simple loop in obtained as the projection of a line in of slope . The simple loop in doubly covers , and so we have as conjugacy classes in .
For and , consider the -orbifold , attach a 2-handle orbifold to it along the simple loop . Since divides into and , the resulting -orbifold has a spherical boundary , where is the dihedral group of order (cf. Notation 1.3(3)). Cap this spherical boundary with the -handle orbifold , and denote the resulting -orbifold by . (Though this orbifold was denoted by in [32], we employ this symbol, because we use the symbol to mainly denote spherical dihedral orbifolds.) Then we have
Let be the annular orbifold on , and continue to denote the orbifold pair by the symbol .
In [32, Section 6], it is proved that the orbifold pair is a pared -orbifold (see [10, Definition 8.3.7]).
Definition 3.1.
An orbifold pair is a pared -orbifold if it satisfies the following conditions
-
(1)
is a compact, orientable, irreducible -orbifold which is very good (i.e., has a finite manifold cover).
-
(2)
is a disjoint union of incompressible toric and annular -suborbifolds.
-
(3)
Every rank free abelian subgroup of is conjugate to a subgroup of some , where is a connected component.
-
(4)
Any properly embedded annular -suborbifold of whose boundary rests on essential loops in is parallel to .
It is also observed in [32, Section 6] that is a Haken pared orbifold (see [10, Definitions 8.0.1 and 8.3.7]). Hence, by the hyperbolization theorem of Haken pared orbifolds [10, Theorem 8.3.9], the pared orbifold admits a geometrically finite complete hyperbolic structure, namely, the interior of the orbifold admits a geometrically finite complete hyperbolic structure such that represents the parabolic locus (see Section 5 for definitions).
Moreover, such a hyperbolic structure is unique, because the ends of the non-cuspidal part of are isomorphic to , which are quasi-isometrically rigid, and every orbifold homeomorphism between two geometrically finite structures preserving the parabolicity in both directions is isotopic to a quasi-isometry, as can be seen by the same argument as Marden’s theorem [34]. We denote the unique (up to conjugation) Kleinian group that uniformises the pared orbifold by the symbol .
Now the Heckoid groups and the Heckoid orbifolds are defined as follows [32, p.242 and Definition 3.2].
Definition 3.2.
For and , the Heckoid group of slope and index is the Kleinian group that is obtained as the image of the natural homomorphism
The Heckoid orbifold of slope and index is the pared orbifold, that is obtained as the covering of the pared orbifold associated with the subgroup . We also denote the Kleinian group by .
Then we have the following proposition. (The main Theorem 1.1 implies that the converse to the first assertion of the proposition holds.)
Proposition 3.3.
For any and , the Heckoid group is a (non-free) Kleinian group with nontrivial torsion which is generated by two non-commuting parabolic transformations. Moreover, the image of the conjugacy class of the simple loop in is an elliptic transformation of rotation angle .
Proof.
Let be the canonical generating meridian pair of the rank free group (see Section 2). Then is generated by the image . Since is the Kleinian group which uniformises the pared orbifold , the generating pair of consists of non-commuting parabolic transformations. Since and since is a meridian of the singular set of of index , it follows that is an elliptic transformation of rotation angle . ∎
Next, we recall the topological description of the Heckoid orbifolds. In Definition 3.2, the Heckoid orbifold is defined as a covering of the pared orbifold . Their explicit topological description is given by [32, Propositions 5.2 and 5.3], which says that the Heckoid orbifold is isomorphic to one of the orbifold pairs depicted in Figure 1, that is specified by the following formula.
where is defined from by the following rule.
Thus the following precise definition of the orbifold pairs in Figure 1 gives an explicit topological picture of the Heckoid orbifold .
Definition 3.4.
(1) For and for a positive integer , denotes the orbifold pair determined by the weighted graph , where is given by
(2) For with odd and an odd integer , denotes the orbifold pair determined by the weighted graph , where is given by the following rule. Let and be the edges of the graph distinct from . Then
(3) For and an odd integer , denotes the orbifold pair determined by the weighted graph , where is given by the following rule. Let and be unions of two mutually disjoint edges of the graph distinct from . Moreover, if is even, then both and are preserved by the vertical involution of . (Thus interchanges the two components of each of and .) Then
In Defunition 3.4(3), the ‘identity’ means that for each edge contained in . Similarly, means that for each edge contained in . We employ this kind of convention throughout the paper.
Remark 3.5.
(1) Because of the -symmetry of -bridge links, the choice of the edges and in (2) and (3) does not affect the isomorphism class of the resulting orbifolds (see [32, Remark 5.4]).
(2) Suppose is odd. Then, in the definition of , the disjointness condition of and determines the pair up to the horizontal involution of . Moreover, according to whether is odd or even, both and are preserved by or , respectively (see Figure 5(1),(2)).
(3) Suppose is even. Then, in the definition of , the condition that both and are preserved by is not essential in the following sense. Let and be the components of for , such that and . Set and . Then and are unions of two mutually disjoint edges of the graph distinct from , such that both and are preserved by the planar involution , instead of the vertical involution (see Figure 5(3),(4)). Let be the weight function on the graph defined by
Then represents the orbifold , where for . This follows from the fact that there is a homeomorphism from to sending to (see Proposition 2.1(1a)).

4. Classification of dihedral orbifolds
In this section, we give a classification of the dihedral orbifolds, which plays a key role in the proof of the main theorem. We refer to [8, 9, 19] for standard terminologies for orbifolds.
By using the the orbifold theorem, the geometrisation theorem of compact orientable -manifolds, and the classification of geometric dihedral orbifolds (see Appendix), we obtain the following classification of good orbifolds with dihedral orbifold fundamental groups.
Theorem 4.1.
Let be a compact orientable -orbifold with nonempty singular set satisfying the following conditions.
-
(i)
does not contain a bad -suborbifold.
-
(ii)
Any component of is not spherical.
-
(iii)
is a dihedral group.
Then is isomorphic to one of the following orbifolds.
-
(1)
The spherical dihedral orbifold represented by the weighted graph for some and coprime positive integers and , where is given by the following rule (see Figure 6).
-
(2)
The orbifold represented by the weighted graph , where takes the value at each component of the -bridge link of slope , i.e. the -component trivial link.
-
(3)
The orbifold represented by the weighted graph , where is the trivial knot in the projective -space with .
-
(4)
The orbifold .
Remark 4.2.
For the orbifold , if (resp. ), then (resp. ) does not belong to the singular set (cf. Convention 6.2(1)). In particular, is the -orbifold associated with the -bridge link in the sense of [11], i.e. the orbifold with underlying space and with singular set , whose index is . In Adam’s classification of torsion-free Kleinian groups generated by two parabolic transformations [1, Theorem 4.3], the -orbifolds played a key role, whereas the orbifolds play the corresponding key role in this paper.

Proof.
Let be a -orbifold satisfying the three conditions. We first treat the case where is irreducible, i.e., any spherical -suborbifold of bounds a discal -suborbifold (a quotient of a -ball by a finite orthogonal group). We can observe that is topologically atoroidal as follows. Suppose on the contrary that contains an essential toric suborbifold . Then the inclusion map induces an injective homomorphism from into , as explained below. Since does not contain a bad -suborbifold by the condition (i), is very good, by [8, Corollary 1.3]. Thus by applying the equivariant loop theorem to the group action, embeds into (see [9, Corollary 3.20]). This contradicts the fact that the dihedral group does not contain .
Hence, by the orbifold theorem [8, Corollary 1.2], is geometric, i.e., either admits one of Thurston’s geometry or is a discal -orbifold. The latter possibility does not happen by the assumption (ii), and so admits one of Thurston’s geometry. If the geometry is , then by Proposition 12.2, is isomorphic to the orbifold in (1). If the geometry is , then by Proposition 13.1, is isomorphic to the orbifold in (2) or the orbifold in (3). (But this does not happen, because these orbifolds are reducible whereas we currently assume that is irreducible.) If the geometry is one of the remaining 6 geometries, then by Proposition 13.2, is isomorphic to the orbifold in (4).
Next, we treat the case when is reducible. Note that does not contain a non-separating spherical -suborbifold, because is finite. Thus we do not need to worry about the paradoxical problems concerning spherical splitting of -orbifolds pointed out by Petronio [43]. By [43, Theorems 0.1], there is a finite system of spherical -suborbifolds such that (a) no component of is punctured discal (a discal -orbifold minus regular neighbourhoods of a finite set) and (b) all prime factors of (the orbifolds obtained from the components of by capping the boundary components with discal orbifolds) are irreducible. It should be noted that some prime component may be a manifold, i.e., its branching locus is empty. By the geometrisation theorem of compact orientable -manifolds (see e.g. [7]) and the the geometrisation theorem of compact orientable -orbifolds (see e.g. [9, Theorem 3.27]), each prime component of admits a canonical decomposition into geometric pieces by a family of essential toric -orbifolds. In particular, each prime factor has a nontrivial orbifold fundamental group. Since the only nontrivial free product decomposition of a dihedral group is the decomposition of the infinite dihedral group into the free product , is the connected sum (along a -sphere with empty branching set) of two irreducible -orbifolds and , such that . Since is geometric, is isomorphic to (a) the discal -orbifold , (b) the orbifold , where is a trivial knot and , or (c) . By condition (ii), cannot be a discal orbifold. Since has nonempty ramification locus, at least one of is not isomorphic to . Hence, is isomorphic to the orbifold in (2) or the orbifold in (3). ∎
5. Relative tameness theorem for hyperbolic orbifolds
We first recall basic terminology for hyperbolic orbifolds, following [9, Chapter 6]. Let be a finitely generated Kleinian group and the quotient hyperbolic orbifold. For a real number , the -thin part of is the set of all points such that for some lift of to and some of order (including ). By the Margulis Lemma, there is a constant , such that for any real number , each component of is either a Margulis tube or a cuspidal end. Here a Margulis tube is a compact quotient of the -neighbourhood of a geodesic in by an elementary subgroup of which preserves the geodesic, and a cuspidal end is the quotient of a horoball in by an elementary parabolic subgroup of which preserves the horoball.
Topologically, a cuspidal end is a product , where is a Euclidean -orbifold. Thus we have the following possibilities for .
-
(1)
is the open annulus or , the quotient of by an involution.
-
(2)
is the torus or , the quotient of by an involution.
-
(3)
is , or , the quotient of by a finite cyclic group action of order , or , respectively.
A cusp is said to be rigid if , or . Otherwise it is said to be flexible. It is well-known that a cusp is rigid if and only if the holonomy representation of the orbifold fundamental group admits no nontrivial deformation (see [37, Proposition 1]).
Let be the union of the cuspidal ends of , and let be the non-cuspidal part of . Then is a disjoint union of euclidean -orbifolds, and is called the parabolic locus of . Note that and that consists of (closed) toric orbifolds (closed -orbifolds obtained as quotients of the -dimensional torus) and open annular orbifolds (open -orbifolds obtained as quotients of the open annulus ).
The following theorem is an orbifold version of (the relative version of) the tameness theorem established by Agol [2] and Calegari-Gabai [18] (see also Soma [55] and Bowditch [15]).
Theorem 5.1.
Let be a hyperbolic -orbifold with finitely generated orbifold fundamental group . Then there is a compact -orbifold and a compact suborbifold of , such that (i) and (ii) the interior of in is equal to .
Proof.
We give a proof following the arguments of Bowditch [15, Section 6.6] (cf. [2, Lemma 14.3]). By Selberg’s lemma, admits a finite regular manifold cover, namely there is a complete hyperbolic manifold and a finite group of orientation-preserving isometries of such that . The inverse image, , of in forms a -invariant non-cuspidal part of , and we have . By the relative version of the tameness theorem [18, Theorem 7.3] (cf. [15, Section 6]), there is a compact -manifold and a compact submanifold of , such that (i) and (ii) the interior of in is equal to . Let and be the double of along . Then the action of on extends to an action on , and is isomorphic to the double, , of along . Consider the double, , of along . Then is a compact manifold with interior . By [40, Theorem 8.5], the action of on extends to an action on , and is identified with . Let be the closure in of one of the two copies of in , and let be the image of in . Then the pair satisfies the desired conditions. ∎
The above theorem together with the following theorem enables us to reduce the treatment of geometrically infinite case to that of geometrically finite case.
Theorem 5.2.
Under the setting of Theorem 5.1, is a pared orbifold. Moreover, the pared orbifold admits a geometrically finite complete hyperbolic structure. Namely, there is a geometrically finite Kleinian group such that (i) the orbifold is isomorphic to the orbifold and (ii) is the parabolic locus of .
Proof.
The first assertion that is a pared orbifold can be proved as in the proof of [41, Corollary 6.10 in Chapter V]. So we prove the second assertion that the pared orbifold admits a geometrically finite hyperbolic structure. If the orbifold is Haken in the sense of [10, Definition 8.0.1] then it follows from [10, Theorem 8.3.9] that the pared orbifold admits a geometrically finite hyperbolic structure, as desired. So we may assume the orbifold is non-Haken, i.e., either it contains no essential -suborbifold or it contains an essential turnover. In the first case, consists only of turnovers by [9, Proposition 9.4]. This implies that every end of has a neighbourhood isomorphic to the product of . Since a hyperbolic turnover is always realised by a totally geodesic surface, each end has a neighbourhood containing no closed geodesics. Thus every end of the hyperbolic orbifold is geometrically finite and rigid. Thus admits a unique complete hyperbolic structure, and it is geometrically finite. In the latter case, by the turnover splitting theorem [9, Theorem 4.8], admits a decomposition by a finite disjoint family of essential hyperbolic turnovers into Haken orbifolds and small orbifolds. By the orbifold theorem, each piece admits a geometrically finite hyperbolic structure, respecting the parabolic locus. By gluing these hyperbolic structures along the totally geodesic hyperbolic turnovers, we obtain a geometrically finite hyperbolic structure on . ∎
Remark 5.3.
In [2], Agol suggested to prove the last assertion of Theorem 5.2 by using a relative version of the work of Feighn and Mess [22, Theorem 2] which proves the existence of a compact core of an orbifold with a finitely generated orbifold fundamental group . Such a relative version is proved by Matsuzaki [37, Lemma 2] under the assumption that is indecomposable (over finite cyclic groups and with respect to the parabolic subgroups) in the sense of [37, Definition in p.26]. But we are not sure if non-free two-parabolic generator Kleinian groups satisfy this property. Though Theorem 5.1, which is proved by using the deep tameness theorem, of course, guarantees the existence of a relative core of complete hyperbolic orbifolds with finitely generated fundamental groups, we are not sure if more ‘elementary’ proof is possible.
6. Orbifold surgery
In this section, we introduce a convenient method for representing pared orbifolds by weighted graphs, generalising the convention in the introduction (Convention 6.1). Then we introduce the concept of an orbifold surgery (Definition 6.3), which is a key ingredient of the proof of the main theorem, and prove a basic Lemma 6.4 for the orbifold surgery. At the end of this section, we also state another basic Lemma 6.5 concerning the -homology of an orbifold, which is repeatedly used in the proof of the main theorem.
Convention 6.1.
Consider a triple , where is a compact oriented -manifold, is a finite trivalent graph properly embedded in , and is a function on the edge set of which takes value in . Here, a loop component of is regarded as a single edge, is the set of degree vertices of , and all other vertices have degree . For each edge of , its value by is called the weight of the edge. We call the triple a weighted graph and call the weight function of the weighted graph. Let be the subgraph of consisting of the edges with weight , and let be the subgraph of consisting of the edges with integral weight.
We regard each component, , of as a -orbifold as follows: the underling space is the complement of an open regular neighbourhood of in , and the singular set is , where the index of a singular point is given by the weight of the corresponding edge of . We assume that the following condition (SC) is satisfied.
-
(SC)
For any sphere component of , the corresponding -orbifold is not a bad orbifold, a spherical orbifold, a discal orbifold, nor an annulus. Namely, (i) and (ii) if then , where () are the (germs of) edges of which have an endpoint in .
A trivalent vertex of is said to be spherical, euclidean or hyperbolic according to whether is bigger than, equal to, or smaller than , where () are the (germs of) edges incident on . Let (resp. ) be the set of the euclidean (resp. hyperbolic) vertices.
Let be the complement of an open regular neighbourhood of in . Then has the structure of an orbifold, with singular set , where the indices of the edge of are given by .
For each edge of , let be a meridian loop of , let be the disjoint union of the regular neighbourhoods in of , where runs over the edges of . The condition (SC) implies that each component of is either a euclidean or hyperbolic -orbifold. Let be the union of and the euclidean components of . Then is a disjoint union of euclidean -orbifolds.
We call the orbifold pair determined by the weighted graph .
Convention 6.2.
It is sometimes convenient to employ the following slight extension of Convention 6.1.
(1) We allow to have an edge with . In this case, we consider the weighted graph , where is the subgraph of consisting of those edges with and is the restriction of to . If is also trivalent graph properly embedded in and the condition (SC) is satisfied, then we define the orbifold pair determined by to be that determined by .
(2) We allow a quadrivalent vertex, , such that the four edge germs incident on it have index . In this case, represents a parabolic locus, , isomorphic to .
A key ingredient of the proof of the main theorem is an orbifold surgery.
Definition 6.3.
Let be a pared orbifold, represented by a weighted graph satisfying the condition (SC). By replacing the weight function with another weight function (which also takes value in ), we obtain another weighted graph . This fails to satisfy the condition (SC) only when some sphere component of the topological boundary determines a spherical -orbifold with three singular points. In this case, we cap all such sphere boundaries of with a cone over to obtain a new weighted graph, which we call the augmentation of . It satisfies the condition (SC), and determines an orbifold pair . We call the -orbifold the orbifold obtained from by the orbifold surgery determined by the replacement of the weight function with the new weight function .
The following simple lemma is used repeatedly in the proof of the main theorem.
Lemma 6.4.
Let be a pared orbifold, and let be the orbifold obtained from by an orbifold surgery. Then does not contain a bad -suborbifold and does not contain a spherical component. In particular, is very good.
Proof.
Let be a weighted graph representing the pared orbifold , and let be the weight function on that gives the orbifold , where is the orbifold pair that is represented by the augmentation of . Assume to the contrary that contains a bad -suborbifold, , which is either a teardrop or a spindle for some integers . Since the underlying space is disjoint from the vertex set of the singular set, , of , we may assume is a submanifold of transversal to . Then it determines a suborbifold, , of , such that . The singular set of is equal to , where is the subgraph of consisting of the edges of integral -weight, and the index of each singular point is given by the -weight of the corresponding edge of .
First, suppose that for . Let be the edge of such that is the singular point of . If is an edge of , then is isomorphic to the teardrop , which contradicts the fact that is good. If is an edge of , then is a disc whose boundary is an essential simple loop on . This contradicts the fact that is incompressible in .
Next, suppose that for . Let and be the edges of corresponding to the singular point of of index and , respectively. Then , and so and are distinct. If both and are contained in , then for some . Since does not contain a bad -suborbifold, and must be equal, and hence is an spherical suborbifold of . Since is irreducible, bounds a discal -orbifold. This implies and determine the same edge of . By the condition (SC), this in turn implies , a contradiction. If exactly one of and is contained in , then is a discal orbifold whose boundary is an essential simple loop on . This contradicts the assumption that is incompressible in . If none of and is contained in , then is an annulus whose boundary consists of a pair of essential simple loops on . Thus is parallel to by Definition 3.1(4), and so , a contradiction.
Thus we have proved that does not contain a bad -suborbifold. The assertion that does not contain a spherical orbifold follows from the fact that is represented by the augmentation of . The assertion that is very good follows from [8, Corollary 1.3], which is a consequence of the orbifold theorem. ∎
Another key tool for the proof of the main theorem is the homology with coefficient. Under Notation 1.3, we have the following lemma, which can be easily deduced from the definition of and the Alexander duality.
Lemma 6.5.
Suppose an orbifold is represented by a weighted graph in . Let be the subgraph of spanned by the edges of even weight. Then is determined by . To be precise, we have the following natural isomorphisms.
In particular, the following hold.
-
(1)
is generated by the meridians of edges of .
-
(2)
The meridian of an edge of of odd degree represents the trivial element of .
-
(3)
Let () be edges of incident on a vertex of , and suppose that is odd and and are even. Then the meridians of and represent the same element of .
7. Canonical horoball pairs for Kleinian groups generated by two parabolic transformations
Throughout Sections 7 11, denotes a non-elementary Kleinian group generated by two parabolic transformations and , and denotes the quotient hyperbolic -orbifold. Let be the geodesic joining the parabolic fixed points of and , and let be the -rotation around . Then we have
We call the inverting elliptic element for the parabolic generating pair of the Kleinian group . As shown in [57, Section 5.4], we can find a geodesic intersecting orthogonally, such that the -rotation, , around it satisfies the following identity.
We call the exchanging elliptic element for the parabolic generating pair of the Kleinian group . It should be noted that is the exchanging elliptic element for the parabolic generating pair of .
By abuse of notation, we denote the isometries of induced by and by the same symbols and , respectively. Each of them is either the identity map or a (nontrivial) involution of , i.e., its order is or . We call the isometries and , the exchanging involution and the inverting involution of associated with the parabolic generating pair . It should be noted that if is isomorphic to a hyperbolic -bridge link group and is the upper-meridian pair, then the involutions and on are the restrictions of the vertical and horizontal involutions of (see Figure 4). This is the reason why we use the symbols and with two different meanings.
Let be the group generated by and the exchanging elliptic element associated with the parabolic generating pair of . Then is a Kleinian group which is either equal to or a -extension of according to whether belongs to or not. Let be the quotient hyperbolic orbifold, and let be the maximal cusp of corresponding to the conjugacy class of containing both and . Then the inverse image of by the projection is a union of horoballs with disjoint interiors but whose boundaries have nonempty tangential intersections. We call it the canonical horoball system associated with the parabolic generating pair of . If a parabolic element of stabilises a member of the canonical horoball system, we denote the horoball by . We denote the translation length of on the horosphere by the symbol , and call it the length of in the canonical horosphere. We call the pair the canonical horoball pair for the parabolic generating pair of the Kleinian group .
Note that the definition of depends on the parabolic generating pair , because the exchanging elliptic element is involved in the definition. However, it actually depends only on the pair , because any orientation-preserving isometry, which exchanges and , also exchanges the members and of the canonical horoball pair associated with . (Otherwise, the product of and an unexpected involution, which exchanges and but does not exchange and , gives a loxodromic transformation which fixes the parabolic fixed points and . This contradicts the assumption that is discrete.)
The following lemmas are proved by Adams [1, Lemma 3.1, Theorem 3.2, and p.197] (see also Brenner [16]). Since they holds a key to the proof of the main theorem and since we described the setting in a slightly different way, we include the proof.
Lemma 7.1.
Under the above setting, the following hold.
-
(1)
For any parabolic element which stabilises a member of the canonical horoball system, we have .
-
(2)
.
-
(3)
If is non-free then .
Proof.
(1) We may assume is the horosphere in the upper half space model . Then some other member, , of the canonical horoball system touches and hence has Euclidean diameter . Since is also a member of the canonical horoball system, and have disjoint interiors. Hence we have .
(2) Since and are conjugate in , and are equal. Moreover, is by (1).
Lemma 7.2.
Both and are primitive in .
Proof.
If is a free, then the assertion follows from the fact that any member of a free-generating system of a free group is primitive. So, we may assume is non-free. Suppose on the contrary that one of the two elements, say , is imprimitive, namely there is an element and an integer such that . Then by Lemma 7.1(1). But, this contradicts Lemma 7.1(3). ∎
8. Outline of the proof of Theorem 1.1
We now state an outline of the proof of Theorem 1.1. Since the if part is clear (cf. Proposition 3.3), we prove the only if part. To this end, we summarise the setting of Theorem 1.1.
Assumption 8.1.
Let be a non-free Kleinian group generated by two non-commuting parabolic transformations and , and let be the quotient hyperbolic orbifold. Let be the non-cuspidal part of , and the parabolic locus. By Theorem 5.1, admits a relative compactification , which is a pared orbifold by Theorem 5.2. The pared orbifold can be represented by a weighted graph , where is a compact -manifold, is a trivalent graph properly embedded in , and is a weight function on the edge set of (see Convention 6.1). We abuse notation to denote the (compact) pared orbifold by . We denote the components of , which is now denoted by , corresponding to the cusps and by and , respectively.
Outline of the proof of Theorem 1.1.
Under Assumption 8.1, the proof is divided into the following two cases.
- Case 1.
-
Case 2.
is a rigid cusp (Section 10).
In both cases, the first task is to find an orbifold surgery that yields an orbifold with dihedral orbifold fundamental group.
In Case 1, this can be generically done by using Lemma 7.2. In fact, if is a flexible cusp, then Lemma 7.2 implies that each of the parabolic elements and can be represented by simple loops of and , respectively. Generically, these simple loops are disjoint, and such an surgery obviously exists. This generic case is treated in Section 9.
However, there is an exceptional case where and the simple loops representing and intersect nontrivially (Lemma 9.1). In this case, the exchanging elliptic element does not belong to , and we need to consider the -extension of and consider the corresponding pared orbifold , where is isomorphic to the rigid cusp . The treatment of this case is deferred to Section 11, after the treatment of the rigid cusp Case 2 in Section 10, described below.
In Case 2, if is isomorphic to either or , the dihedral surgery can be found by using an estimate of the shortest, second shortest, and third shortest lengths of parabolic elements on the maximal rigid cusp, which in turn is based on Lemma 7.1. If is isomorphic to , the inverting parabolic element does not belong to , and we consider the -extension and the corresponding pared orbifold . The images of and in this quotient is isomorphic to , and this case can be treated by using arguments in the case where .
After finding an orbifold surgery that yields an orbifold with dihedral orbifold fundamental group, we can appeal to the classification Theorem 4.1 of the dihedral orbifolds, because Lemma 6.4 guarantees that the orbifold satisfies the three conditions in Theorem 4.1. So, belongs to the list in the theorem. The original pared orbifold is obtained from the dihedral orbifold by inverse surgery operations. Through case-by-case arguments, by using the homology with -coefficients, a result concerning the symmetries of the spherical dihedral orbifold (Corollary 12.7), and a ‘surgery trick’ (the last paragraph in Case 1 in Section 10 and Case 1 in Section 11), we prove the following.
- (1)
-
(2)
If is a rigid cusp, then we encounter a contradiction (Section 10).
This ends an outline of the proof of the main Theorem 1.1. ∎
9. Proof of Theorem 1.1 - flexible cusp: generic case -
Under Assumption 8.1, suppose that is a flexible cusp. Then the -orbifold is isomorphic to the torus , the pillowcase , the annulus , or . The following fact is the starting point of this section.
Lemma 9.1.
Under the above setting, and are represented by simple loops on and , respectively. Moreover, if , then one of the following holds.
-
(1)
The parabolic elements and are represented by the same (possibly oppositely oriented) simple loop.
-
(2)
, , and , where the first is the exchanging elliptic element associated with and the last two ’s denote the involution on induced by the exchanging elliptic element (see Figure 11).
Proof.
The first assertion directly follows from Lemma 7.2, because any primitive parabolic element in the orbifold fundamental group of the -dimensional orbifold , , , or is represented by a simple loop on the -orbifold. For the proof of the second assertion, suppose that . If the exchanging elliptic element belongs to , then is conjugate to in , and so they are represented by the same simple loop. Thus we may suppose . Then descends to a nontrivial orientation-preserving involution on , which we continue to denote by , on the flexible cusp . By the classification of orientation-preserving involutions on flexible cusps, we can observe that either (a) the involution on preserves or reverses the homotopy class of each essential simple loop on , or (b) and . In the first case, and are represented by the same simple loop, and so we obtain the desired conclusion. ∎
In this section, we treat the case where either or and the conclusion (1) in Lemma 9.1 holds. Thus we assume the following condition in the remainder of this section. The other case is treated in Section 11.
Assumption 9.2.
Under this assumption, we can apply an orbifold surgery on to the pared orbifold to obtain a dihedral orbifold, , as follows. Note that Assumption 9.2 implies that the pared orbifold is represented by a weighted graph , such that there are (possibly identical) edges and of whose meridians represent and , respectively. Let be a weight function on which is identical with , except that . Then the orbifold represented by the augmentation of the weighted graph is a result of an “order ” orbifold surgery on , and is dihedral, as shown below.
Note that there is a natural epimorphism from to , and the images of and in have order . Moreover, the images of and have the same order, because (a) if then and are conjugate in and so in , and (b) if then descends to an involution on which interchanges the images of and . So is either the trivial group or a dihedral group. Since is very good by Lemma 6.4 and since has nonempty singular set, is nontrivial and so isomorphic to a dihedral group.
Thus satisfies the three conditions in Theorem 4.1 and so belongs to the list in the theorem. We have the following lemma.
Lemma 9.3.
The orbifold is isomorphic to the spherical dihedral orbifold for some and coprime positive integers and .
Proof.
We show that the possibilities (2), (3) and (4) in Theorem 4.1 cannot happen. Suppose (2) happens. Then we can see that one of the following holds, by recalling the fact that is obtained from the pared orbifold an order orbifold surgery.
-
(i)
is the exterior of the two-component trivial link, , and the singular set of is empty.
-
(ii)
The underlying space of is the solid torus (the exterior of a trivial knot), , and the singular set is a trivial knot in the solid torus with index .
In each case, is reducible, a contradiction.
By the same reasoning, we can see that (3) cannot happen.
If (4) happens, then as in the above, we can see that one of the following holds, where is a two-strand trivial tangle.
-
(i)
and the singular set of is empty.
-
(ii)
and the singular set of is with index .
In the first case, is a rank free group, which contradicts the assumption that is non-free. In the second case, note that , the abelianization of the orbifold fundamental group , is . On the other hand, both and are represented by the core loop of the annulus , and the pair cannot generate , a contradiction. ∎
By Lemma 9.3, the original orbifold is recovered from by applying the inverse orbifold surgery operation. This leads us to the following proposition.
Proposition 9.4.
Under the notation in Lemma 9.3, the following hold, if necessary by replacing with where or .
-
(1)
If , then one of the following holds.
-
(i)
and , where .
-
(ii)
, , and .
-
(iii)
, is an odd integer , and .
-
(iv)
, is an odd integer , and .
-
(i)
-
(2)
If , then one of the following holds.
-
(i)
and , where .
-
(ii)
, , and .
-
(iii)
, is an odd integer , and .
-
(i)

Proof.
Recall that is represented by the weighted graph for some and for some coprime positive integers and , and is given by the following rule (see Figure 6):
Then is represented by the weighted graph , where is obtained from by replacing the label of the edges and , which correspond to and respectively, with the label . By Remark 4.3, we may assume , if necessary by replacing with where .
Case 1. . Then is the -bridge link . Thus , where either (a) or (b) is a -component link and . In the first case, , and so is the hyperbolic -bridge link complement, : in particular, . In the second case, both and are meridians of the component , which contradicts the fact that .
Case 2. .
Subcase 2.1. (see Figure 7(1)). Then the edge set of consists of and the two components , of . Let , and be the meridians of , and , respectively. By Lemma 6.5, is freely generated by , and moreover we have . Since is generated by (the images of) and , we may assume and . Thus is represented by , where and . Hence .
Subcase 2.2. (see Figure 7(2)). Then the edge set of consists of and the two subarcs and of bounded by . Let , and be the meridians of , and , respectively.
Suppose first that is odd. Then we see by Lemma 6.5 that in and that is generated by . Hence one of the following holds.
-
(1)
and so is represented by , where and . Hence .
-
(2)
for or . By the symmetry of , we may assume and so is represented by , where , and . Hence .
Suppose next that is even. Then in . Since is generated by and , we have . This implies that the exchanging elliptic element for does not belong to , and descends to an involution on interchanging with . We now use Corollary 12.7 on the symmetry of the orbifold . We first consider the generic case where (i.e., is a nontrivial knot) or . (Recall the current assumption .) Then, by Corollary 12.7(1), any orientation-preserving involution of preserves . So, and are different from , and therefore . Hence, as in the previous case, we can conclude . In the exceptional case where and , The orbifold has the -fold cyclic symmetry as illustrated in Figure 15. Thus, if necessary after applying this symmetry, we may assume . Hence we have .
Since we repeatedly use the above argument in the remainder of the proof of Proposition 9.4, we state an expanded version of the argument as a lemma.
Lemma 9.5.
Under the setting of Proposition 9.4, suppose and . Then , and the exchanging elliptic element does not belong to and it descends to an orientation-preserving involution of interchanging and . Moreover, the following hold.
-
(1)
Except when and , and are different from .
-
(2)
If , then the inverting elliptic element belongs to .
Proof.
We have only to prove (2). If does not belong to , then it descends to an orientation-preserving involution of which preserves both and . However, if , then by Corollary 12.7(2), no orientation-preserving involution of preserves an edge of the singular set different from . This contradicts the assertion (1). ∎
Case 3. . Since and are coprime, we see and one of and is odd.
Subcase 3.1. (see Figure 7(3)). Let and be the components of , and let () be the edges of such that for and that the vertical involution of interchanges and for . Let and be the meridians of and , respectively. Then by using Lemma 6.5 and the fact that one of and is odd, we see that is freely generated by and : moreover we have . Hence, we may assume and . Since the horizontal involution of interchanges and (), we may assume and for some or . By Lemma 9.5(2), we have , and so is homeomorphic to or . Since , we must have . If , i.e. , then is given by
Since the vertical involution of preserves , we see that is isomorphic to . If , i.e. , then the planar involution of preserves . Hence, we see by Remark 3.5 that is isomorphic to , where .
Subcase 3.2. (see Figure 7(4)). Suppose first that one of and is even. Then by Lemma 6.5. Hence, by Lemma 9.5(2), both and are contained in , and . In particular, or . Let () be the edges of the singular set of contained in the knot in this cyclic order. We also assume that and . Since the -symmetry of acts transitively on the edge set (see Figure 14), we may assume and so . Since joins with and since are coprime integers such that , the condition that or implies that and . This in turn implies that . Since is isotopic to the simple loop in , we must have . Thus is equal to or . However, if then , , and the odd index edge share a vertex, it follows from Lemma 6.5(3) that the meridian of and the meridian of represent the same element of , a contradiction. Hence . Set and . Then and are disjoint, and the following hold.
Hence we have (cf. Remark 3.5(2)).
Suppose finally that both and are odd. Then, by Lemma 6.5, the meridians of represent the trivial element of , and hence both and are contained in . On the other hand, since , we have is homeomorphic to an annulus, and hence the inverting elliptic element descends to an involution of which preserves each of the two mutually different edges and and restricts to an orientation-reversing involution on each of the edges. But, such an involution does not exist by Corollary 12.7(2), a contradiction.
This completes the proof of Proposition 9.4. ∎
10. Proof of Theorem 1.1 - rigid cusp case -
Under Assumption 8.1, suppose that is a rigid cusp. Thus the -orbifold is isomorphic to where , , or .
Let be the orbifold fundamental group , and let be the subgroup of consisting of parabolic transformations. We may assume that (a) stabilises the ideal point of the upper-half space model of , and (b) the boundary of the canonical horoball is identified with the horosphere . For each element , let be the length of in the canonical horosphere (see Section 7), namely , the translation length of in , and simply call it the length of . Let be the minimum of the lengths of nontrivial elements of . More generally, for each , let be the -th shortest length of nontrivial elements of .
Case 1. . Then , and is the rank free abelian group with free basis . We may assume the action of on the horosphere is given by the following rule. There is a positive real such that is the rotation about , and and are the rotations about and , respectively. We can easily observe the following.
-
(i)
The shortest length is equal to , and it is attained precisely by the conjugates of in . (Note that is conjugate to .)
-
(ii)
The second shortest length is equal to , and it is attained precisely by the conjugates of in .
-
(iii)
The third shortest length is equal to , and it is attained precisely by the conjugates of in .
By Lemma 7.1(1), , and so . Since is non-free, Lemma 7.1(3) implies that the length of the parabolic element is less than . Since , is equal to either or . By using this fact, we obtain the following lemma.
Lemma 10.1.
The parabolic element is conjugate to or in . Moreover the following hold.
-
(1)
If is conjugate to , then the images of by the natural epimorphisms from to , , and have order .
-
(2)
If is conjugate to , then the images of by the natural epimorphisms from to , , and have order , and , respectively. Moreover, the -homology class of vanishes.
Proof.
The assertion in the first line follows from the observations preceding the lemma. The assertions (1) and (2) can be checked easily, by using the fact that is conjugate to in . ∎

Now let be the orbifold obtained from the pared orbifold by the orbifold surgery as illustrated in Figure 8. Namely, for each index edge of the singular set which has an endpoint in or , we replace the index with the index , and then cap all resulting spherical boundary components with discal -orbifolds. Then each of and shrinks into a vertex of with link , which we denote by and , respectively. We denote by (resp. ) the edge of the singular set incident on (resp. ) whose index in the original orbifold is .
Lemma 10.2.
The orbifold is isomorphic to a spherical dihedral orbifold for some and coprime positive integers and . Moreover, is conjugate to in , and the homology class determined by is equal to the meridian of the edge . Similarly, the homology class is equal to the meridian of the edge .
Proof.
By Lemma 10.1, is conjugate to or in , its image in has order or accordingly. Hence the image of in has order . Moreover, the images of and have the same order, because (a) if the exchanging involution belongs to then and are conjugate in and so in , and (b) if then descends to an involution on which interchanges the images of and . Hence is either the trivial group or a dihedral group. Since is very good by Lemma 6.4 and since it has a singular point with link , is a noncyclic dihedral group. Hence, by Theorem 4.1, is isomorphic to a spherical dihedral orbifold .
We prove the remaining assertions. If is conjugate to , then it descends to the trivial element of , and so it represents the trivial element of . This contradicts the fact that is a dihedral group generated by the images of and . Hence is conjugate to . This implies that the -homology class is equal to that represented by the element , and so it is the meridian of the edge . The existence of the exchanging elliptic element implies the corresponding assertion for . ∎
Lemma 10.3.
The pared orbifold is represented by the weighted graph for some , where is determined by the following rule (see Figure 9):
for some odd integer , where and are unions of two mutually disjoint edges of the graph distinct from , such that . Moreover, and correspond to distinct endpoints of (in the sense of Convention 6.1(3)).
Proof.
By Lemma 10.2, is represented by the weighted graph for some , where is given by the rule
for some coprime positive integers and . Since is obtained from by an orbifold surgery, there is a weight function on the graph such that the pared orbifold is represented by the weighted graph . By Remark 4.3, we may assume is odd, if necessary by replacing with where . Hence, we see by Lemma 6.5. Since is generated by and , which are the meridians of the edges and , respectively (see Lemma 10.2), we have .
Since the links of and are isomorphic to , we see and . Since (resp. ) is the unique edge of the trivalent graph incident on the vertex (resp. ) with -weight , and since , we see that and are distinct endpoints of . (If , then its ‘link’ in is of the form .) Hence and correspond to distinct endpoints of .
We observe that and are not equal to . If, say was equal to , then it is incident on . Since , this implies we have , a contradiction. This observation implies that both and are contained in .
We next observe that . If , then the endpoints of and are all contained in . This together with the previous observation implies that the ‘links’ of and in are isomorphic to , a contradiction.
We now show that and are disjoint. If they are not disjoint, then they share an endpoint of , which has odd weight . This implies that the meridians of and represent an identical element of (see Lemma 6.5(3)), and so , a contradiction.
Set and let . Then and satisfy the desired conclusion with . ∎

We show that the situation described in Lemma 10.3 cannot happen. To this end, we perform another orbifold surgery on which replaces the weight of with . To be precise, we consider the orbifold represented by the weighted graph for some , where is given by the following rule.
Note that shrinks into a singular point of with link or . Since is conjugate to in , we see by Lemma 10.1(1) that the image of in has order . The same argument can be applied to and we see that the image of in also has order . Since is very good by Lemma 6.4 and since the singular set of contains a trivalent vertex, is a noncyclic dihedral group. Since the singular set of contains four trivalent vertices, Theorem 4.1 implies that must be isomorphic to a spherical dihedral orbifold with . In particular, the singular set of must contain precisely four or five edges with index . This contradicts the fact that contains precisely three edges of index (see Figure 9).
Case 2. . Then , and is the rank free abelian group with free basis . We may assume the action of on the horosphere is given by the following rule. There is a positive real such that is the rotation about , is the rotation about , and is the rotations about . The action of the generators of is given by
We can easily observe the following.
-
(i)
, and it is attained precisely by the conjugates of in .
-
(ii)
, and it is attained precisely by the conjugates of in .
-
(iii)
, and it is attained precisely by the conjugates of in .
By Lemma 7.1(1), , and so . Since is non-free, Lemma 7.1(3) implies that the length of the parabolic element is less than . Hence we obtain the following.
Lemma 10.4.
The parabolic element is conjugate to or in .

Now let be the orbifold obtained from the pared orbifold by the orbifold surgery as illustrated in Figure 10. Namely, for each edge of the singular set which has the index cone point of or as an endpoint, we replace the weight with the new weight , and then cap all resulting spherical boundary components with discal -orbifolds. Then and shrink into singular points, and , of with link .
Lemma 10.5.
The image of by the natural epimorphism from to has order .
Proof.
By Lemma 10.4, is conjugate to either or in . Moreover, the images of and in have order . This is obvious for , and the assertion for is verified as follows. In , we have and so . Hence the image of in is equal to . Thus it is conjugate to , and so has order , as desired. ∎
By the above lemma, the image of in has order . The existence of the exchanging elliptic element implies that the images of and in have the same order. Thus is either a dihedral group or the trivial group. Since is very good by Lemma 6.4 and since the singular set of contains a trivalent vertex, is a noncyclic dihedral group. Hence, Theorem 4.1 implies that must be isomorphic to a spherical dihedral orbifold with . However, the orbifold does not contain a singular point with link , a contradiction.
Case 3. . Then the inverting elliptic element does not belong to , and the group, , obtained from by adding is a -extension of . Consider the hyperbolic orbifold . Then is the quotient of by the isometric involution induced by , which we continue to denote by . Set , and . Then is isomorphic to . Thus . Since the subgroup of generated by and is non-free, we see by the arguments in Case 2 that is conjugate to or in .
Let be the orbifold obtained from the pared orbifold by the orbifold surgery as illustrated in Figure 10 at both and . Then and shrink into singular points, and , of with link . The images of and in have the same order , and so the subgroup of they generate is either a dihedral group or the trivial group. This subgroup has index in , because has index in . Hence the group is a trivial group, a dihedral group, (the -extension of the trivial group) or a -extension of a dihedral group.
Since is very good by Lemma 6.4 and contains a singular point with link , is either a noncyclic dihedral group or a -extension of a noncyclic dihedral group. Hence Theorem 4.1 implies that is isomorphic to (a) a spherical dihedral orbifold or (b) the quotient of by an isometric involution, where . Since does not have a singular point with link , (a) cannot happen, and so we may assume (b) holds. Since does not have an index subgroup, the link of an inverse image of the singular point in the double cover of is also isomorphic to . But, this is impossible. Hence cannot be isomorphic to .
Thus we have proved that cannot be a rigid cusp.
11. Proof of Theorem 1.1 - flexible cusp: exceptional case -
In this section, we treat the case where the following assumption is satisfied, and prove that this assumption is never satisfied.
Assumption 11.1.
Let be the group generated by and . Let be the quotient hyperbolic orbifold. Let be the non-cuspidal part of , and the parabolic locus. By abuse of notation, we denote the pared orbifold obtained as the relative compactification of by the same symbol . We denote the component of the compact euclidean -orbifold corresponding to the conjugacy class containing and by . Thus and , where denotes the involution on the pared orbifold induced by the exchanging involution . In particular, is the double orbifold covering of , associated with the homomorphism such that and . We denote the homomorphism induced by by the same symbol.
Note that . As in Case 1 in Section 10, we identify with the stabiliser . Then the proof of Lemma 10.1 also works in this setting, because generates the non-free subgroup of the Kleinian group , and we have the following lemma.
Lemma 11.2.
The parabolic element is conjugate to or in , and of course, the assertions (1) and (2) in Lemma 10.1 also hold.

Let and () be the edges of the singular sets and as illustrated in Figure 11. Thus and are contained in the fixed point set of the involution on , and is the image of by the covering projection . (Note that it can happen that some of them are identical, though their germs near the parabolic locus are different.) Then the following holds.
Lemma 11.3.
The homomorphism , that determines the double orbifold covering , satisfies
where denotes the meridian of the edge . Moreover, the homology class determined by is equal to either or .
Proof.
The formula for follows from the fact that the fixed point set of the involution on contains and , which project to and , respectively. It is also obvious that . So, if , we have . Suppose that . Then, since is generated by and , we see . So, is conjugate to by Lemma 11.2. (Otherwise is conjugate to and so .) Therefore is contained in . Thus is generated by . Since , it follows that is equal to either or . ∎
Let be the orbifold obtained from the pared orbifold by the orbifold surgery that replaces the index of the edges and with the index . Then shrinks into a singular point, , with link , and the image of in has order by Lemma 11.2. Since is generated by and , is either trivial, or a noncyclic dihedral group. By using Lemma 6.4, Theorem 4.1 and the fact that has a singular point with link , we see that is isomorphic to a spherical dihedral orbifold with noncyclic dihedral orbifold fundamental group. Moreover, we may assume that and that is an endpoint of . By Lemma 6.5, we have .
Lemma 11.4.
Under the above setting, and so the edges () are all distinct.
Proof.
We first observe that cannot be conjugate to . In fact, if was conjugate to , then its image in is trivial by Lemma 11.2 (cf. Lemma 10.1(2)), and so is generated by the image of . This contradicts the fact that is a noncyclic dihedral group. This observation together with Lemma 11.2 implies that is conjugate to and so . Moreover, or by Lemma 11.3. Hence is generated by the meridians of the three edges () incident on the vertex .
Now suppose on the contrary that . Then we see, by using Lemma 6.5, that the meridian of represents the trivial element of and the meridians of the remaining two edges incident on represent the identical element of . This contradicts the fact that . Hence . This implies that the the edges () incident on are all distinct, as desired. ∎
Recall that the weights of the edges , , of are . Since by Lemma 11.4, we can apply the orbifold surgery on of “type ”, namely we can replace the index of the edge of the singular set with the index , and leave the other indices, including the index of , unchanged. We denote the resulting orbifold by . By Lemma 11.2, has order at most in . Hence, by using Lemma 6.4, Theorem 4.1, and the fact that has a singular point with link , we see that is isomorphic to a spherical dihedral orbifold with noncyclic dihedral orbifold fundamental group. Moreover, we may assume and that the parabolic locus degenerates into a singular point, , which is an endpoint of . It should be noted that the edge of corresponds to . (Here, we reset the notation, and the symbols and now represent objects different from those they had represented in the paragraph preceding Lemma 11.4.)
Case 1. . We apply the orbifold surgery on of “type ”, namely we replace the index of the edge of the singular set with the index , and leave the other indices, including the index of , unchanged. (This is possible by Lemma 11.4.) We denote the resulting orbifold by . By Lemma 11.2, has order at most in . Hence, again by using Lemma 6.4, Theorem 4.1, and the fact that has a singular point with link , we see that is isomorphic to a spherical dihedral orbifold with noncyclic dihedral orbifold fundamental group. Note that the edges and of , which have indices and , respectively, share a common endpoint (see Figure 12). But this cannot happen in any spherical dihedral orbifold, a contradiction.

Case 2. . Then is represented by a weighted graph , such that is a knot, , and
Recall that the subset of are the images of the fixed point set of the involution on . This implies that the map induced by the orbifold covering is the double branched covering branched over . Hence is represented by the weighted graph illustrated in Figure 13. Here, we assume the extended Convention 6.2, and the two -valent vertices represent parabolic loci isomorphic to . Hence we see by Lemma 6.4 that , a contradiction.
Thus we have proved that the situation in Assumption 11.1 cannot occur. This completes the proof of the main Theorem 1.1.

12. Appendix 1: Spherical orbifolds with dihedral orbifold fundamental groups
In this appendix, we classify the orientable spherical -orbifolds with dihedral orbifold fundamental groups (Proposition 12.2), and determine the (orientation-preserving) isometry groups of these orbifolds (Propositions 12.5 and 12.6). Proposition 12.2 is used in the proof of Theorem 4.1, and Corollary 12.7 is used in Section 9. Propositions 12.5 and 12.6 are used in the companion [4] of this paper. The classification of the spherical dihedral orbifolds is implicitly contained in Dunber’s work [21], which classifies the Seifert fibered orbifolds. The isometry groups of the dihedral spherical orbifolds obtained as the -orbifolds associated with -bridge links are calculated by [50, 27]. Moreover, in the recent papers [38, 39], Mecchia and Seppi classified the Seifert fibered spherical -orbifolds and calculated the isometry groups of such orbifolds. Since every spherical dihedral orbifold is Seifert fibered, the results in this section are implicitly contained in [38, 39]. However, we give a self-contained proof, because it is not a simple task to translate their results into the form we need.
We first recall basic facts concerning the -dimensional spherical geometry following [54, 50]. Let be the quaternion skew field. We use the symbol to denote a generic quaternion
(We believe this does not cause any confusion, even though is also used to denote the numerator of a rational number .) For each , denotes its conjugate, denotes its real part, and denotes its norm . We identify () with the following subspaces of .
The norm induces a Euclidean metric on , and () are endowed with the induced metrics. The subspaces and form a Lie group with respect to the restriction of the product in . The group acts on itself by conjugation leaving invariant. This gives an epimorphism , with , defined by
If with , then is the rotation of , by angle , with fixed points .
For a positive integer , any cyclic subgroup of order (resp. any dihedral subgroup of order ) of is conjugate to the subgroup (resp. ), where and with . Note that these groups are contained in the subgroup of . Then the following hold (see, e.g. [50, Proposition 2.6]).
Lemma 12.1.
(1) If , then the normaliser of in is equal to .
(2) If , then the normaliser of in is equal to . If , then is equal to the binary octahedral group , where is the octahedral group obtained as the subgroup of preserving the regular octahedron in the -dimensional Euclidean subspace of spanned by the vertices .
Let be the homomorphism defined by
Then is an ephimorphism with .
We occasionally identify with the unit sphere
in by the correspondence . Let be the injective homomorphism defined by
When (), where is a rational number, we write
(1) |
because its restriction to the circles and are the ‘-rotation’ and ‘-rotation’, respectively. Though the symbol is used in two different ways, we believe this does not cause any confusion, because its meaning is clearly understood from the context according to whether is a unit complex or a rational number.
Observe that
In particular, we have
(2) |
Consider the isometries , and , which acts on as follows.
Observe and that
(3) |
In fact, is the split extension of by , where the action of on by conjugation is given by the following formula.
(4) |
The following proposition gives a classification of the orientable spherical -orbifolds with dihedral orbifold fundamental groups.
Proposition 12.2.
Let be an oriented spherical -orbifold. Then is isomorphic to a dihedral group, if and only if is isomorphic to the orbifold, , represented by the weighted graph in Figure 6 for some and coprime positive integers and , where is given by the following rule.
In fact, with is isomorphic to , where is the subgroup of given by
(5) |
for some integers and such that
(6) |
Moreover, the spherical structure of is unique, i.e., if is a subgroup of such that is isomorphic to as oriented orbifolds, then is conjugate to the subgroup defined by (5).
Proof.
We first prove the only if part of the first assertion. Let be a subgroup of isomorphic to the dihedral group , and let and be the elements of such that
(Though the symbols , and are used in different meanings in the previous sections, we believe this does not cause any confusion.) We show that is isomorphic to some , such that .
Claim 12.3.
After taking conjugation in , we may assume , where , , , , and are positive integers such that , , , and .
Proof of Claim 12.3.
Since any element of is conjugate to an element in , we may assume, by taking conjugation, that (see (2)). Since has order , we may assume for some integers and such that . For , set , and , so that , where . Note also that . Set and (). Then . Thus and so . Similarly we have . Hence we have and . ∎
Now consider the subgroup generated by . Since , we have
Note that
Hence we have
Thus is the orbifold with underlying space and with singular set the Hopf link, where one component has index and the other component has index . To give a precise description of this orbifold, identify with the join , by the correspondence . Thus the first and second factor circles of correspond to the circles and in , respectively. For , let be the isometry of defined by (). Then the isometry is identified with the self-homeomorphism of , defined by
Under the above convention, the orbifold is described as follows. The underlying space of the orbifold is given by
and the singular set is the union of the two circles which gives the join structure of , where the first factor circle (which corresponds to ) has index and the second factor circle (which corresponds to ) has index . Here, denotes as in (1).
The isometry descends to the periodic isomorphism of the orbifold given by , because the periodic map (resp. ) on descends to the periodic map (resp. ) on the circle (resp. ). Note that , with , where is the inverse of in the multiplicative group (cf. Notation 1.3(2)). Hence we see that the orbifold is isomorphic to the orbifold, , with underlying space the lens space, , and with singular set the union of the core circles of the standard genus Heegaard splitting of with indices and , respectively. (Though the notation looks similar to the notation in (1), we believe there is no fear of confusion.)
Since and , we see by using Lemma 12.1(1) that for some . Since any element of is conjugate to by an element of , we may assume , and so . This implies that the involution of descends to the hyper-elliptic involution of . Recall that (i) the quotient map determined by the hyper-elliptic involution gives the double branched covering of branched over the -bridge link and that (ii) the core circles of the genus Heegaard splitting project to the upper and lower tunnels, respectively. Hence, the quotient , with , is isomorphic to the orbifold . This completes the proof of the only if part of the first assertion. The proof also shows that the group is given by the formula (5) for some integers and satisfying the condition (6).
The if part of the first assertion follows from the above argument and the following claim.
Claim 12.4.
For any rational number and a pair of coprime integers , there is a pair of integers which satisfies the condition (6).
Proof of Claim 12.4.
Consider the homomorphism
where the first homomorphism is the product of the natural projections and the second homomorphism maps to . Then both of the two homomorphisms are surjective and so is their composition . Regard the numerator of the rational number as an element of , and let be a pair of integers which projects to an element in the inverse image . Then satisfies the condition (6). ∎
Finally we prove the uniqueness of the spherical structure on the orbifold . The preceding arguments show that the triple uniquely determines a dihedral subgroup , up to conjugation, such that is isomorphic to as oriented orbifolds. Thus we have only to show that there are no unexpected orientation-preserving topological isomorphism between two orbifolds, and .
Assume that and are isomorphic as oriented orbifolds. Then and , because they have isomorphic orbifold fundamental groups and the same index sets of the singular sets. In particular we have .
Suppose first that . Then has the unique cyclic subgroup of index , and so each of and has the unique double orbifold covering with cyclic orbifold fundamental group. The underlying spaces of the covering orbifolds are the lens spaces and , respectively. Hence, by the classification of lens spaces, we have and either or . Moreover, by using the uniqueness of the genus one Heegaard splittings (see [12, 13]), we see that and are isomorphic as oriented orbifolds if and only if one of the following conditions hold.
-
(1)
, , and .
-
(2)
, , and .
In both cases, we can see that the subgroups and are conjugate in .
In the exceptional case when , we have either (i) and or (ii) and . We can easily see that the subgroups and are conjugate in .
This completes the uniqueness of the spherical structure. ∎
Next, we calculate the (orientation-preserving) isometry group of the dihedral spherical -orbifold . If , then is the -orbifold, , associated with the -bridge link (cf. [11]) and its isometry group is calculated by [50, Theorem 4.1] and [27, Corollary 3.2.11]. (There are errors in [50, Theorem 4.1] for the special case when , . There are also misprints for the generic case in the statement of Theorem 4.1, though the correct result can be found in the tables in [50, p.184].)
Proposition 12.5.
The orientation-preserving isometry group of the spherical orbifold is described as follows.
-
(1)
If , then the following holds.
-
(2)
If , then then the following holds.
In the remainder of this section, we treat the remaining case . In the very special case, when and , we call the trivial -orbifold, because its singular set is the trivial -curve in . Then we have the following proposition.
Proposition 12.6.
The orientation-preserving isometry group of the spherical dihedral orbifold with is described as follows.
-
(1)
, except when is isomorphic to the trivial -orbifold , i.e. except when and .
-
(2)
For the the trivial -orbifold , we have .


Before proving the proposition, we give the following consequence of the proposition (and the orbifold theorem), which is used in the proof of the main theorem.
Corollary 12.7.
Consider a spherical orbifold with , and let be an orientation-preserving involution of the orbifold. Then the following hold.
-
(1)
Except when and , (i.e. except when is isomorphic to the trivial -orbifold ), stabilises the edges and of the singular set (when it is contained in the singular set).
-
(2)
If , then does not stabilise any edge of the singular set different from .
Proof of Corollary 12.7.
By the orbifold theorem, we may assume is an isometry of the spherical orbifold. (This is proved by applying the orbifold theorem to the finite group action on the universal cover of generated by a lift of and the covering transformation group.) On the other hand, the action of in the generic case is as illustrated in Figure 14. (See also [4, Figure 6-8], and replace the weights with , then we obtain the desired visualisation, besides the exceptional case.) The exceptional case where the orbifold is the trivial -orbifold is illustrated in Figure 15. The assertion (1) is now obvious from Figure 14. The assertion (2) also follows from the figure by noting that consists of four edges if (otherwise, consists of two edges). ∎
The proof of Proposition 12.6 presented below is parallel to that of [50, Theorem 4.1]. Consider the dihedral spherical -orbifold with . By Proposition 12.2, the orbifold fundamental group is identified with the subgroup
where , for some integers and satisfying the condition (6). Pick such that . Then . Set
Then , where is the normaliser of in .
For , set , where is the projection to the -th factor. Then for some positive integer . Then the following lemma is obvious from the definition of , where denotes the order of a group element.
Lemma 12.8.
(1) .
(2) If is even, then . If is odd, then is either or .
Note that the orientation-reversing isometry , defined by , acts on by conjugation, as follows:
Hence, we assume without loss of generality.
Lemma 12.9.
(1) .
(2) If , then for some odd integer , and . Moreover, if and only if .
Proof.
(1) Suppose on the contrary that . Then , and so . This implies and therefore . Since , we have , a contradiction.
(2) Suppose . Then , and so . By using Lemma 12.8, we can verify the following, from which the assertion (2) follows.
-
(i)
If is odd, then and so as in (1), a contradiction.
-
(ii)
If for some odd integer , then and so . Hence we have and .
-
(iii)
If for some integer , then and so . Hence , a contradiction.
∎
Lemma 12.10.
Except for the special case where and , namely except when is the trivial -orbifold, we have
Proof.
If , then Lemma 12.1 implies for each (because by assumption), and hence we have .
Since by Lemma 12.9(1), we have only to treat the case where . Since we exclude the case where and , Lemma 12.9(2) implies , and so . On the other hand, since , we see by Lemma 12.1 that . Hence .
Now observe that the decomposition induces the decomposition of into the following two non-empty subsets.
Note that and . Pick an arbitrary element . Then , and so the inner-automorphism of determined by preserves the subgroup . Thus the inner-automorphism of determined by preserves the subset of . Hence the inner-automorphism of determined by preserves the subgroup , and so it preserves the subset , i.e., . By the description of in Lemma 12.1, this implies that . Hence , as desired. ∎
Lemma 12.11.
The normaliser of in is contained in .
Proof.
Lemma 12.12.
Except when is the trivial -orbifold, we have the following.
Proof.
Recall that where and . Set and . Suppose an element belongs to . Then , and hence belongs to the subgroup .
Proof of Proposition 12.6.
Consider the spherical dihedral orbifold with . We first treat the generic case where is not the trivial -orbifold . Then, by using Lemma 12.12 and the fact that , is isomorphic to the quotient of the group by its subgroup . Set , and . It should be noted that and that the subset is equal to the set of all order elements of .
Note that the order of is equal to or according to whether is odd or even. Similarly, the order of is equal to or according to whether is odd or even. Thus the order of is or , where the latter happens if and only if both and are even.
Case 1. . Then the element has order . Hence it is equal to one of the elements of . Thus . This implies that for some . Since corresponds to the subgroup , we have
Case 2. . Since , we have , and so is odd. Thus , and therefore . Hence, we have
This completes the proof of Proposition 12.6 in the generic case.
In the exceptional case, where is the trivial -orbifold , we may assume
Then, by using Lemma 12.1, we can see that . Hence we have
∎
13. Appendix 2: Non-spherical geometric orbifolds with dihedral orbifold fundamental groups
In this section, we classify the non-spherical geometric orbifolds with dihedral orbifold fundamental groups (Propositions 13.1 and 13.2). These results are used in the proof of Theorem 4.1.
We first deal with the dihedral orbifolds with geometry.
Proposition 13.1.
Let be a compact orientable orbifold with nonempty singular set which satisfies the following conditions.
-
(i)
No component of is spherical.
-
(ii)
is a dihedral group.
Then is isomorphic to one of the following orbifolds.
-
(1)
, the orbifold represented by the weighted graph , where takes the value at each component of the -bridge link of slope , i.e. the -component trivial link.
-
(2)
, the orbifold with underlying space whose singular set is the trivial knot (i.e., the boundary of an embedded disc in ) with index .
Proof.
By the assumption that has the geometry , we have and . By the condition (ii), for some .
If , then the action of the finite dihedral group on extends to an action on the compact -manifold , where is a compactification of , and is identified with . Thus has a spherical boundary component, which contradicts the condition (i). So and . Since the action of on is properly discontinuous, is the product of a (possibly trivial) rotation of and a nontrivial translation of . Thus the orbifold is homeomorphic to the manifold . The isometry descends to a fiber-preserving involution of which acts on the second factor as a reflection. Thus is the quotient of by an equivalence relation and where and are orientation-reversing involutions of . Thus is conjugate to either the reflection in a great circle or the antipodal map. According to the combination (reflection, reflection), (reflection, antipodal map), or (antipodal map, antipodal map), is isomorphic to , , or . The last case cannot happen because has the empty singular set. ∎
The following proposition deals with the dihedral orbifolds with the remaining geometries.
Proposition 13.2.
Let be a compact orientable -orbifold with nonempty singular set which has one of the geometries different from and and satisfies the following conditions.
-
(i)
is a dihedral group.
-
(ii)
No component of is spherical.
Then is isomorphic to .
Proof.
Let be the geometry which possesses. Then is , , , or , and is isomorphic to for some discrete subgroup of . Note that the underlying topological space of is homeomorphic to . The proof is divided into two cases according to whether is finite or infinite.
Case 1. Suppose that is a finite dihedral group . Then, as will be shown below, the action of on has a global fixed point . Then the exponential map from , the tangent space to at , to is a -equivariant homeomorphism. This implies that , contradicting the condition (ii).
The existence of a global fixed point can be proved as follows. For the constant curvature case or , this is well-known. We shall first deal with the case where is , , or . Recall that there is an exact sequence
where is the Euclidean plane when is and the hyperbolic plane when is or . We also note that the projection above is induced by a fibration . Let be the image of in and the kernel in of the projection to . Then the action of on has a global fixed point , and the action of on the fibre has a global fixed point since both of them are finite. Thus has a global fixed point on when is , , or .
We shall now show the same property when . In this case, there is an exact sequence
and the projection is induced by a fibration . Let be the projection of in . Then is either trivial or generated by a reflection on . In either case, it fixes a point on . In the former case, acts on the fibre by Euclidean isometries in such a way that and correspond to reflections, and hence fixes a point on the fibre. In the latter case, the kernel of the projection is isomorphic to , and fixes a point on the fibre in the same way. Thus we have shown that has a fixed point also in the case when .
Case 2. Suppose is the infinite dihedral group . First we shall consider the case when has constant curvature. Then and are order elliptic transformations, and hence fix pointwise axes and respectively. They do not meet each other since otherwise the action fixes their intersection and cannot be faithful and discrete. Let be the common perpendicular to and if it exists. (This does not exist when and touches at infinity. This exceptional case will be considered later.) Let be the totally geodesic plane containing and perpendicular to . We define in the same way. Then the region cobounded by and constitutes a fundamental domain of the action of . Suppose now that and touches at infinity. Then there is a totally geodesic plane containing both and and it is preserved by . We then let and be totally geodesic planes containing and respectively, which are perpendicular to . Then a fundamental domain is cobounded by and again. Therefore, in either case, we can see and so .
Next, we shall consider the case when is or or . As before, let be the projection of to , and the kernel of the projection. We first observe that the images and of and in are nontrivial. In fact, if say is trivial, then acts on as a nontrivial order isometry. Thus acts on by a reflection, and so the image must be an orientation-reversing isometry on , which contradicts our assumption that is trivial.
Since and have order , their images and in also have order . We first deal with the case where both of them are orientation-preserving, i.e. -rotations. Let and be the centres of the -rotations and , respectively. Then and are -rotations about the geodesics and , respectively. Since the action of is faithful, we have . Now consider the geodesic line in containing and , and let and be the lines which intersects perpendicularly at and , respectively. Then and are disjoint, and they cobound a region in . We see that is a fundamental region of the action of , and we have .
We next treat the case where both of and are orientation-reversing, i.e. reflections. Let and be the axes of the reflections and , respectively. Then and are the ‘symmetries’ with respect to the geodesics and , respectively, where and are lifts of and , respectively. If and are disjoint, they cobound a region in , and is a fundamental region of the action of , and we have . If and meet each other at a point . Then acts effectively and discretely on the fiber , and so the axes and intersect perpendicularly at distinct points, and , respectively. Let and be the ruled surfaces in obtained as the unions of the geodesics which intersect perpendicularly at and , respectively. Then and are disjoint planes in , and the domain they cobound is a fundamental domain of , and we can see .
We now treat the case where one of and is orientation-preserving and the other is orientation-reversing. We may assume is orientation-preserving and is orientation-reversing. Let be the center of the -rotation , and let be the axis of the reflection . If belongs to , then the axes of the -rotations of and intersect, and the action of cannot be discrete and faithful. So is not contained in . Let be a geodesic line in which passes through and is disjoint from . Let be the region in bounded by and . Then is a fundamental region of the action of , and we have .
Finally, suppose that . Then the projection of to is either trivial or or itself. In the case when is trivial, the generators and act on each fibre by -rotations, and their fixed points must differ. Thus is a bundle over with fiber . So we have . In the case when is , the action of is a reflection with respect to a point . We set . Then the and act on by reflections, and by the same argument as in the previous paragraph, we have a homeomorphism . Finally, suppose that . Then and fix points and on respectively, and they differ. We consider fibres and . The elements act on and as reflections with axes and . Then the region cobounded by and constitutes a fundamental region for the action of , and we see that . ∎
References
- [1] C. Adams, Hyperbolic -manifolds with two generators, Comm. Anal. Geom. 4 (1996), 181–206.
- [2] I. Agol, The classification of non-free -parabolic generator Kleinian groups, Slides of talks given at Austin AMS Meeting and Budapest Bolyai conference, July 2002, Budapest, Hungary.
- [3] I. Agol, Tameness of hyperbolic 3-manifolds, arXiv:math/0405568.
- [4] S. Aimi, D. Lee, S. Sakai, and M. Sakuma, Classification of parabolic generating pairs of Kleinian groups with two parabolic generators, preprint.
- [5] H. Akiyoshi, J. Parker, and M. Sakuma, On Hecke groups and Heckoid groups, in preparation.
- [6] H. Akiyoshi, M. Sakuma, M. Wada, and Y. Yamashita, Punctured torus groups and -bridge knot groups (I), Lecture Notes in Mathematics 1909, Springer, Berlin, 2007.
- [7] L. Bessiéres, G. Besson, S. Maillot, M. Boileau, J. Porti, Geometrisation of 3-manifolds, EMS Tracts in Mathematics, 13. European Mathematical Society (EMS), Zürich, 2010. x+237 pp.
- [8] M. Boileau, B. Leeb, and J. Porti, Geometrization of 3-dimensional orbifolds, Ann. of Math. 162 (2005), 195–290.
- [9] M. Boileau, S, Maillot, and J. Porti, Three-dimensional orbifolds and their geometric structures, Panoramas et Synthèses [Panoramas and Syntheses], 15, Société Mathématique de France, Paris, 2003. viii+167 pp.
- [10] M. Boileau and J. Porti, Geometrization of -orbifolds of cyclic type, Appendix A by Michael Heusener and Porti, Astérisque No. 272 (2001).
- [11] M. Boileau and B. Zimmermann, The -orbifold group of a link, Math. Z. 200 (1989), 187–208.
- [12] F. Bonahon, Difféotopies des espaces lenticulaires, Topology 22 (1983), 305–314.
- [13] F. Bonahon and J. P. Otal, Scindements de Heegaard des espaces lenticulaires, C. R. Acad. Sci. Paris Sér. I Math. 294 (1982), 585–587.
- [14] F. Bonahon and L. Siebenmann, New geometric splittings of classical knots, and the classification and symmetries of arborescent knots, http://www-bcf.usc.edu/~fbonahon/Research/Publications.html
- [15] B. H. Bowditch, Notes on tameness, Enseign. Math. 56 (2010), 229–285.
- [16] J. L. Brenner, Quelques groupes libres de matrices, C. R. Acad. Sci. Paris 241 (1955), 1689–1691.
- [17] G. Burde, H. Zieschang and M. Heusener, Knots. Third, fully revised and extended edition, De Gruyter Studies in Mathematics, 5. De Gruyter, Berlin, 2014. xiv+417
- [18] D. Calegari and D. Gabai, Shrinkwrapping and the taming of hyperbolic 3-manifolds, J. Amer. Math. Soc. 19 (2006), 385–446.
- [19] D. Cooper, C. Hodgson and S. Kerckhoff, Three-dimensional orbifolds and cone-manifolds MSJ Memoirs 5, Mathematical Society of Japan, Tokyo, 2000. x+170 pp.
- [20] W. D. Dunbar, Hierarchies for 3-orbifolds, Topology Appl. 29 (1988), 267–283.
- [21] W. D. Dunbar, Geometric orbifolds, Rev. Mat. Univ. Complut. Madrid 1 (1988), 67–99.
- [22] M. Feighn and G. Mess, Conjugacy classes of finite subgroups of Kleinian groups, Amer. J. Math. 113 (1991), 179–188.
- [23] J. Gilman, The structure of two-parabolic space: parabolic dust and iteration, Geom. Dedicata 131 (2008), 27–48.
- [24] F. W. Gehring and G. J. Martin, Commutators, collars and the geometry of Möbius groups, J. Anal. Math. 63 (1994), 175–219.
- [25] E. Hecke, Über die Bestimmung Dirichletscher Reihen durch ihre Funktionalgleichung, Math. Ann. 112 (1936), 664–699.
- [26] A. Hinkkanen and G. J. Martin, The dynamics of semigroups of rational functions. I, Proc. London Math. Soc. 73 (1996), 358–384.
- [27] T. L. Jeevanjee, Isometries of orbifolds double-covered by lens spaces, J. Knot Theory Ramifications 12 (2003), 819–832.
- [28] T. Jorgensen, On pairs of once-punctured tori, Kleinian groups and hyperbolic 3-manifolds (Warwick, 2001), 183–207, London Math. Soc. Lecture Note Ser., 299, Cambridge Univ. Press, Cambridge, 2003.
- [29] L. Keen and C. Series, The Riley slice of Schottky space, Proc. London Math. Soc. 69 (1994), 72–90.
- [30] S. Kim and T. Koberda, Non-freeness of groups generated by two parabolic elements, arXiv:1901.06375.
- [31] Y. Komori and C. Series, The Riley slice revisited, The Epstein birthday schrift, 303–316, Geom. Topol. Monogr., 1, Geom. Topol. Publ., Coventry, 1998.
- [32] D. Lee and M. Sakuma, Epimorphisms from -bridge link groups onto Heckoid groups (I), Hiroshima Math. J. 43 (2013), 239–264.
- [33] R. C. Lyndon and J. L. Ullman, Groups generated by two parabolic linear fractional transformations, Canadian J. Math. 21 (1969), 1388–1403.
- [34] A. Marden, The Geometry of Finitely Generated Kleinian Groups, Ann. Math. 99 (1974), 383–462.
- [35] G. J. Martin, Nondiscrete parabolic characters of the free group : supergroup density and Nielsen classes in the complement of the Riley slice, preprint.
- [36] G. J. Martin and T. H. Marshall, Polynomial Trace Identities in , Quaternion Algebras, and Two-generator Kleinian Groups, arXiv 1911.11643.
- [37] K. Matsuzaki, Structural stability of Kleinian groups, Michigan Math. J. 44 (1997), 21–36.
- [38] M. Mecchia and A. Seppi, Fibered spherical -orbifolds, Rev. Mat. Iberoam. 31 (2015), 811–840.
- [39] M. Mecchia and A. Seppi, Isometry groups and mapping class groups of spherical 3-orbifolds, arXiv:1607.06281.
- [40] W.H. Meeks and P. Scott, Finite group actions on 3-manifolds, Invent. Math. 86 (1986), 287–346.
- [41] J.W. Morgan and H. Bass, The Smith conjecture, Pure Appl. Math., 112, Academic Press, Orlando, FL, 1984.
- [42] K. Ohshika and H. Miyachi, Uniform models for the closure of the Riley slice, In the tradition of Ahlfors-Bers. V, 249–306, Contemp. Math., 510, Amer. Math. Soc., Providence, RI, 2010.
- [43] C. Petronio, Spherical splitting of 3-orbifolds, Math. Proc. Cambridge Philos. Soc. 142 (2007), 269–287.
- [44] R. Riley, Parabolic representations of knot groups. I, Proc. London Math. Soc. 24 (1972), 217–242.
- [45] R. Riley, A quadratic parabolic group, Math. Proc. Cambridge Philos. Soc. 77 (1975), 281–288.
- [46] R. Riley, Parabolic representations of knot groups. II, Proc. London Math. Soc. 31 (1975), 495–512.
- [47] R. Riley, Algebra for Heckoid groups, Trans. Amer. Math. Soc. 334 (1992), 389–409.
- [48] R. Riley, Groups generated by two parabolics and for in the first quadrant, Output of an computer experiment.
- [49] M. Sakuma, On strongly invertible knots, Algebraic and topological theories (Kinosaki, 1984), 176–196, Kinokuniya, Tokyo, 1986.
- [50] M. Sakuma, The geometries of spherical Montesinos links, Kobe J. Math. 7 (1990), 167–190.
- [51] M. Sakuma and J. Weeks, Examples of canonical decompositions of hyperbolic link complements, Japan. J. Math. (N.S.) 21 (1995), 393–439.
- [52] C. Series, All about the Riley slice, in preparation.
- [53] H. Schubert, Knoten mit zwei Brücken, Math. Z. 65 (1956), 133–170.
- [54] P. Scott, The geometries of -manifolds, Bull. London Math. Soc. 15 (1983), 401–487.
- [55] T. Soma, Existence of ruled wrappings in hyperbolic 3-manifolds, Geom. Topol. 10 (2006), 1173–1184.
- [56] E. C. Tan and S. P. Tan, Quadratic Diophantine equations and two generator Möbius groups, J. Austral. Math. Soc. Ser. A 61 (1996), 360–368.
- [57] W. Thurston, The geometry and topology of three-manifolds, available from http://library.msri.org/books/gt3m/