<<

HYPERBOLIC MANIFOLDS THAT FIBER ALGEBRAICALLY UP TO DIMENSION 8

GIOVANNI ITALIANO, BRUNO MARTELLI, AND MATTEO MIGLIORINI

Abstract. We construct some cusped finite-volume hyperbolic n-manifolds

Mn that fiber algebraically in all the dimensions 5 ≤ n ≤ 8. That is, there is a surjective homomorphism π1(Mn) → Z with finitely generated kernel. The kernel is also finitely presented in the dimensions n = 7, 8, and this leads to the first examples of hyperbolic n-manifolds Mfn whose fundamental group is finitely presented but not of finite type. These n-manifolds Mfn have infinitely many cusps of maximal rank and hence infinite Betti number bn−1. They cover the finite-volume manifold Mn. We obtain these examples by assigning some appropriate colours and states

to a family of right-angled hyperbolic polytopes P5,...,P8, and then applying some arguments of Jankiewicz – Norin – Wise [15] and Bestvina – Brady [6]. We exploit in an essential way the remarkable properties of the Gosset

polytopes dual to Pn, and the algebra of integral for the crucial dimensions n = 7, 8.

Introduction We prove here the following theorem. Every hyperbolic manifold in this paper is tacitly assumed to be connected, complete, and orientable.

Theorem 1. In every dimension 5 ≤ n ≤ 8 there are a finite volume hyperbolic 1 n-manifold Mn and a map f : Mn → S such that f∗ : π1(Mn) → Z is surjective n with finitely generated kernel. The cover Mfn = H /ker f∗ has infinitely many cusps of maximal rank. When n = 7, 8 the kernel is also finitely presented.

arXiv:2010.10200v3 [math.GT] 1 Mar 2021 We deduce in particular the following.

Corollary 2. In dimension n = 7, 8 there is a hyperbolic n-manifold with finitely presented fundamental group and infinitely many cusps of maximal rank. The man- ifold covers a finite-volume hyperbolic manifold.

The same assertion holds in the dimensions n = 5, 6 with “finitely generated” replacing “finitely presented”.

For every 5 ≤ n ≤ 8 the group π1(Mn) is a finite-index subgroup of the arithmetic O(n, 1, Z). Recall that a group Γ is of finite type, denoted F, if it has a finite classifying space, and of type Fm if it has a classifying space with finite m-skeleton.

1 2 GIOVANNI ITALIANO, BRUNO MARTELLI, AND MATTEO MIGLIORINI

Euler b1 b2 b3 b4 b5 b6 b7 Cusps

M5 0 24 120 136 39 0 0 0 40

M6 −64 18 183 411 207 26 0 0 27

M7 0 182 6321 41300 55139 24010 4031 0 4032

M8 278528 365 33670 583290 1783226 1346030 456595 65279 65280

Table 1. The Euler characteristic, Betti numbers, and the num-

ber of cusps of each hyperbolic n-manifold Mn. Each cusp of Mn is toric, that is diffeomorphic to T n−1 × [0, +∞).

When m = 1 or 2, being of type Fm is equivalent to Γ being finitely generated or finitely presented, respectively.

Corollary 3. In dimension n = 7, 8 the lattice O(n, 1, Z) contains a finitely pre- sented subgroup Γ without torsion and with infinite Betti number bn−1(Γ). In par- ticular Γ is F2 but not Fn−1.

Proof. Pick Γ = π1(Mfn) < π1(Mn) < O(n, 1, Z). Since Mfn has infinitely many cusps of maximal rank, it is homeomorphic to the interior of a manifold with in- finitely many compact boundary components and hence has infinite Betti number bn−1(Mfn) = bn−1(Γ). 

For every 5 ≤ n ≤ 8, all the finitely many cusps of Mn are toric, that is diffeo- morphic to T n−1 × [0, +∞), where we use T m to denote the m-torus. The cover

Mfn has infinitely many toric cusps, and finitely many cusps of rank n − 2, each diffeomorphic to T n−2 × R × [0, +∞). The manifolds Mn and the maps f are constructed explicitly and combinatorially, so some topological invariants may be calculated. The Euler characteristic, Betti numbers, and number of cusps of Mn are listed in Table 1. Outline of the construction. We use as building blocks a remarkable sequence n of finite-volume right-angled polytopes Pn ⊂ H , defined for 3 ≤ n ≤ 8. The reflection group of Pn is a finite-index subgroup of the integral lattice O(n, 1, Z). The polytope Pn has both ideal and real vertices. These polytopes made their first appearance in a paper of Agol, Long, and Reid [1]. Their combinatorics was then described by Potyagailo and Vinberg [26], and more information was later collected by Everitt, Ratcliffe, and Tschantz [12], who noticed in particular that P3,...,P8 are combinatorially dual to the Euclidean Gosset polytopes [14] discovered by Gosset in 1900 and usually indicated with the symbols −121, 021,..., 421. The Gosset polytopes form indeed a remarkable family of semi-regular polytopes.

The 1-skeleton of 221 is the configuration graph of the 27 lines in a generic cubic HYPERBOLIC MANIFOLDS THAT FIBER ALGEBRAICALLY 3

[10], while 321, 421 are intimately connected with the integral octonions and the

E8 lattice. It has been a great pleasure to study these beautiful objects for this project.

The hyperbolic manifold Mn is constructed by assembling some copies of Pn by means of a suitable colouring of its facets. This is a standard procedure that works with any right-angled polytope. It is important here that the colouring has few colours and many symmetries. Given the remarkable properties of the dual Gosset polytopes, it is natural to guess that some nice natural colourings for Pn should exist, and we show that this is indeed the case. In dimension n = 7, 8 we derive a natural colouring from the algebraic properties of the integral octonions. 1 Having constructed Mn, we build a map f : Mn → S by choosing an appro- priate state for Pn, that is a partition of facets into two sets In and Out. A state 1 defines a diagonal map f : Mn → S , as explained by Jankiewicz, Norin, and Wise [15]. The homomorphism f∗ : π1(Mn) → Z is often non-trivial, and its kernel may be studied through Bestvina – Brady theory of combinatorial Morse functions [6]. This fundamental paper furnishes in particular some conditions that, when satis-

fied, guarantee that ker f∗ is finitely generated or, even better, finitely presented. The conditions are the following: if some simplicial complexes (the ascending and descending links [6]) are all connected (respectively, simply connected), then the kernel is finitely generated (respectively, finitely presented).

The choice of an appropriate state for Pn is crucial here, and we have used again the exceptional properties of the dual Gosset polytope, and of the integral octonions for n = 7, 8, to select one state for which the above-mentioned conditions are satisfied. We took inspiration from a quaternions-generated state for the 24-cell that worked very well in [5] to design a similar octonions-generated state for P7 and

P8 here. After choosing the states, the conditions on the ascending and descending links have been verified with a computer code written in Sage, that may be downloaded from [34]. The symmetries of the polytopes, of the colourings, and of the states have reduced considerably the computations involved, to keep them within few hours of CPU time. Without all these exceptional symmetries, not only the proof of Theorem 1, but even the more straightforward calculation of the Betti numbers of Mn would have been problematic, especially in the higher dimensions n = 7, 8 where the combinatorics is highly not trivial, as one can guess by looking at the size of the numbers in Table 1. To the best of our knowledge the manifolds M7 and M8 are the first finite-volume hyperbolic manifolds in dimension n ≥ 7 for which the Betti numbers have been computed. Some notable examples exist in the literature in dimension 5 and 6, see [29, 12]. n The cover Mfn = H /ker f∗ has a finitely generated fundamental group, and also a finitely presented one for n = 7, 8. It has infinitely many cusps for all 5 ≤ n ≤ 8 4 GIOVANNI ITALIANO, BRUNO MARTELLI, AND MATTEO MIGLIORINI because f is homotopically trivial on some cusp of Mn, which therefore lifts to infinitely many copies of itself in Mfn.

Related work. We briefly discuss some works related to the present paper.

Coherence. The fundamental group of a hyperbolic 3-manifold M satisfies a number of nice properties, see [3] for a widely comprehensive discussion. In particular, Scott proved [31] that π1(M) is coherent: every finitely generated subgroup is also finitely presented. This is not the case in higher dimensions, as first experienced by Kapovich and Potyagailo who constructed [19] in 1991 a geometrically finite hyperbolic 4-manifold with non-coherent fundamental group, see also [25, 20]. A compact example was then built by Bowditch and Mess [7] in 1994. Later on, Kapovich, Potyagailo, and Vinberg [21] proved non-coherence for every non-uniform arithmetic lattice in Isom(Hn) with n ≥ 6, and then Kapovich [18] for every arithmetic hyperbolic lattice in dimension n ≥ 5, n 6= 7. He also conjectured in [18] that every hyperbolic lattice in dimension n ≥ 4 is non-coherent. Corollaries 2 and 3 describe an even wilder situation: in dimension n = 7, 8 there are finite-volume hyperbolic n-manifolds whose fundamental group contains subgroups that are F2 but not Fn−1. The first example of a group that is F2 but not Fm for some m ≥ 3 was provided by Stallings [32]. It would be interesting to know if such subgroups may also occur in the intermediate dimensions n = 4, 5, 6.

Algebraic fibrations. Theorem 1 furnishes some explicit examples of algebraically fibering fundamental groups of hyperbolic manifolds. We recall that a group G fibers algebraically if there is a surjective homomorphism G → Z with finitely generated kernel.

When G = π1(M) is the fundamental group of a 3-manifold, by a well-known theorem of Stallings [32] this condition is equivalent to the existence of a fibration M → S1. In higher dimensions this is false in general, and the first examples of algebraic fibrations on hyperbolic n-manifolds have appeared recently in dimension n = 4 in [2, 15]. The paper [15] is the main inspiration for our work. In [2] the algebraic fibration is obtained by constructing a RFRS tower and then applying a recent general theorem of Kielak [22] that transforms the RFRS property into an algebraic fibration (under some hypothesis). In dimension 4 the algebraic fibration can sometimes be promoted to a perfect circle-valued Morse function [5]. We do not know if similar geometric promotions may occur in dimension n ≥ 5. The algebraic fibrations constructed here have a serious obstruction at the cusps, see Section 2.14.

Infinitely many cusps. Theorem 1 produces some hyperbolic manifolds with finitely presented fundamental group and infinitely many cusps. HYPERBOLIC MANIFOLDS THAT FIBER ALGEBRAICALLY 5

In dimension 3, every hyperbolic manifold with finitely generated fundamental group has only finitely many cusps. This is yet another nice property of 3-manifolds that fails in higher dimensions: we already know from [16, 19] that there are some hyperbolic 4-manifolds with finitely generated fundamental group and infinitely many rank-1 cusps. With Theorem 1 we upgrade these examples by substituting “rank-1” with “maximal rank”, and “finitely generated” with “finitely presented”. The reader may consult [17] for a comprehensive survey on 3-dimensional theorems that are not valid in higher dimension (the paper contains also a lot of interesting material). It is conjectured in [16] that there is no hyperbolic n-manifold with finitely generated fundamental group and infinitely many cusps, all of maximal rank. We note that Theorem 1 does not disprove this conjecture, since Mfn also contains finitely many cusps of rank n − 2.

Structure of the paper. We introduce the polytopes Pn and construct the man- ifolds Mn in Section 1 by means of some appropriate colourings. Then in Section 1 2 we introduce the techniques of [15] and build the diagonal maps f : Mn → S via some carefully chosen states. By analysing the behaviour of f we finally prove Theorem 1.

1. The manifolds Mn We recall a general procedure to construct a manifold from a right-angled poly- tope P by colouring its facets. Then we turn to the polytopes P3,...,P8 and choose some nice colouring to generate the manifolds M3,...,M8. We will use the algebraic properties of the octonions to build M7 and M8.

1.1. Colours. Let P ⊂ Xn be a right-angled finite polytope in some space Xn = Hn, Rn or Sn. We always suppose that P has finite volume. When Xn = Hn, the polytope P may have both finite and ideal vertices. We can interpret P as an n orbifold P = X /Γ, where Γ is the right-angled generated by the reflections rF along the facets F of P . A presentation for Γ is

2 h rF | rF , [rF , rF 0 ] i where F varies among the facets of P and F,F 0 among the pairs of adjacent facets. A simple procedure to construct some manifold cover M → P consists of colouring the facets of P . This was used for instance in [23, 5]. We explain it briefly. A colouring of P is the assignment of a colour (taken from some finite set) to each facet of P , such that incident facets have distinct colours. We use generally {1, . . . , c} as a palette of colours and suppose that every colour is painted on at least one facet. c Let e1, . . . , ec be the canonical basis of the Z2-vector space Z2. A colouring on c P induces a homomorphism Γ → Z2 that sends rF of to ej, where j is the colour of 6 GIOVANNI ITALIANO, BRUNO MARTELLI, AND MATTEO MIGLIORINI

F . One verifies that the kernel Γ0 / Γ acts freely on Xn and hence we get a manifold n n c M = X /Γ0 that orbifold-covers P = X /Γ with degree 2 . The manifold M is hyperbolic, flat, or elliptic, according to the model Xn, and is tessellated into 2c copies of P . Geometrically, we may see M as constructed by mirroring P iteratively along its colours, in any order. The manifold M is canonically oriented by extending the orientation of P along the mirrors. c v c We can parametrize the 2 copies of P inside M as P with varying v ∈ Z2. Each P v is an identical copy of P . A facet F of P v is identified (via the identity map) to the same facet of P v+ej , where j is the colour of F . This fully determines the tessellation of M. We say that two colourings on P are isomorphic if they induce the same partition of facets, possibly after acting by some isometry of P . Isomorphic colourings yield isometric manifolds M. As an example, we can always colour P by assigning distinct colours to distinct c facets. In this case c equals the number of facets of P and Γ → Z2 is just the abelianization homomorphism. With this choice, the resulting manifold M can be quite big, so it is often preferable to construct a smaller manifold by minimizing the number of colours. Another reason for rejecting this inefficient colouring here will be given below in Corollary 10.

1.2. Facets and ideal vertices. A colouring of P induces some colourings also on the facets and on the vertex links of P , which in turn produce codimension-1 manifolds nicely embedded in M. Here are the details. Each facet F of P is itself a right-angled finite polytope in Xn−1. A colouring on P induces a colouring on F in the obvious way: assign to any facet of F the colour of the adjacent facet in P . This really gives a colouring of F because the link of any real vertex of P is a simplex, so two facets of F are adjacent if and only if the corresponding facets of P are. The facet F inherits some c0 ≤ c−1 of the c colours of P . The induced colouring on F generates a hyperbolic, flat, or elliptic (according to the geometry of Xn−1) 0 (n−1)-manifold N tessellated into 2c copies of F , and an orbifold-covering N → F .

Proposition 4. The counterimage of F along the orbifold covering M → P consists 0 of 2c−c −1 geodesically embedded copies of N.

Proof. To simplify the notation we may suppose that F has colour 1 and the facets 0 c adjacent to F are coloured with {2, . . . , c + 1}. For every vector v ∈ Z2 with first v v+e1 coordinate v1 = 0 we have a copy of F in M, that separates P from P . There are 2c−1 such copies of F overall. If we consider all the vectors v with fixed last coordinates vc0+2, . . . , vc, we find a geodesically embedded copy of N in M. There c−c0−1 are 2 ways of fixing these last coordinates.  HYPERBOLIC MANIFOLDS THAT FIBER ALGEBRAICALLY 7

An analogous reasoning works for the other strata of P . We are particularly interested in the vertices of P , especially the ideal ones in the hyperbolic case. If v is a real vertex of P , its link is a right-angled spherical (n − 1)-simplex ∆, which inherits a colouring with n distinct colours. This colouring generates a round sphere Sn−1, tessellated into 2n copies of ∆, which lifts to 2c−n round spheres around the 2c−n counterimages of v in M. If P is hyperbolic and v is an ideal vertex of P , its link is a right-angled Euclidean (n − 1)-parallelepiped C, which also inherits a colouring from that of P , with some c0 ≤ c distinct colours. The induced colouring on C generates a compact flat (n−1)- manifold N. (We prove that N is a torus below.) The counterimage of a small star 0 of P at v consists of 2c−c cusps in M, each with a section homothetic to N.

1.3. The Euclidean n-cube. One basic example is the Euclidean n-cube C.

Proposition 5. Up to isomorphism, there is a unique colouring on C with c colours for any n ≤ c ≤ 2n. It produces a flat n-torus for every c.

Proof. There are 2n − c pairs of opposite faces with the same colour and the re- maining 2(c − n) faces with distinct colours. This configuration is unique up to isomorphism. The proof that the resulting flat manifold is a torus is easy and left as an exercise (it can be shown by induction on n).  A cusp in a hyperbolic n-manifold is toric if its section is a flat (n − 1)-torus. We summarize our discussion as follows.

Corollary 6. If P ⊂ Hn is right-angled with ideal vertices, every colouring on P produces some hyperbolic n-manifold M whose cusps are all toric. 0 If P has c colours and v is an ideal vertex, there are 2c−c toric cusps in M above v, where c0 is the number of distinctly coloured facets incident to v.

c Remark 7. A more general notion of colouring consists of assigning a vector λF ∈ Z2 to each facet F of P , that is not necessarily a member of the canonical basis. With this more general notion, non-toric cusps may also appear: see for instance [13, 24].

1.4. The right-angled hyperbolic polytopes. We refer to the excellent papers [26] and [12] for an introduction to the sequence of right-angled hyperbolic polytopes

P3,...,P8. These have many beautiful properties that we now briefly summarize. n Each Pn ⊂ H is a finite volume right-angled polytope with both finite and ideal vertices. The link of a finite and ideal vertex is respectively a right-angled spherical (n−1)-simplex and a Euclidean (n−1)-cube. The numbers of facets, ideal vertices, and finite vertices of Pn are listed in Table 2, together with the isometry group of

Pn and its order. The isometry group acts transitively on the facets, so in particular these are all isometric: in fact, every facet of Pn is isometric to Pn−1 when n ≥ 4.

The quotient of Pn by its isometry group is a simplex. 8 GIOVANNI ITALIANO, BRUNO MARTELLI, AND MATTEO MIGLIORINI

Facets Ideal Finite Isom(Pn) Order Dual

P3 6 3 2 A1 × A2 12 Triangular prism

P4 10 5 5 A4 120 Gosset 021

P5 16 10 16 D5 1920 Gosset 121

P6 27 27 72 E6 51840 Gosset 221

P7 56 126 576 E7 2903040 Gosset 321

P8 240 2160 17280 E8 696729600 Gosset 421

Table 2. The number of facets, ideal vertices, and finite vertices

of Pn, the isometry group Isom(Pn) expressed as a and

its order |Isom(Pn)|, and the dual Euclidean polytope.

1.5. Euler characteristic. Recall that the orbifold Euler characteristic of a hyper- bolic right-angled polyhedron P is zero in odd dimension while in even dimension it can be calculated via the simple formula

n X fi χ(P ) = (−1)i 2n−i i=0 where fi is the number of i-dimensional faces of P , including P itself (so fn = 1).

Only real vertices (not the ideal ones) contribute to f0. From this formula we deduce the well-known [12] values

χ(P4) = 1/16, χ(P6) = −1/8, χ(P8) = 17/2.

In even dimension the Euler characteristic and the volume are roughly the same thing, up to a constant that will be recalled below.

1.6. The Gosset polytopes. Combinatorially, the polytopes Pn are dual to the Gosset polytopes listed in the last column of Table 2 and discovered by Gosset [14] in 1900. Every Gosset polytope is a Euclidean polytope with regular facets, whose isometry group (which is the same as Isom(Pn)) acts transitively on the vertices.

Its regular facets are of two types: some simplexes (dual to the real vertices of Pn) and some hyperoctahedra (dual to the ideal vertices of Pn).

We will describe a colouring of Pn as a colouring on the vertices of the dual Gosset polytope, where we require of course that two vertices adjacent to the same edge must have distinct colours (so only the 1-skeleton of the dual Gosset polytope is important at this stage). We would like to find some colouring with a small number of colours, and possibly a high degree of symmetry: we are confident that some natural choices should arise from the exceptional properties of Pn and their dual Gosset polytopes, and this is indeed the case as we now see. HYPERBOLIC MANIFOLDS THAT FIBER ALGEBRAICALLY 9

Figure 1. The 1-skeleton of the triangular prism has a unique 3-colouring up to isomorphism, shown here.

3 1.7. The manifold M3. The hyperbolic polyhedron P3 ⊂ H is a double pyramid with triangular base, see [26, Figure 2]. The three vertices of the triangular base are ideal, while the two remaining vertices are real. Each of the 6 faces F of P3 is a triangle with a right-angled real vertex and two ideal vertices. The dual Gosset polytope is just a triangular prism. Its 1-skeleton is shown schematically in Figure 1. It can be coloured with 3 colours in a unique way (shown in the figure) up to isomorphism. Dually, we get a partition of the 6 faces of P3 into 3 pairs. The polyhedron cannot be coloured with less than 3 colours.

We equip P3 with this 3-colouring. This produces a hyperbolic 3-manifold M3, 3 tessellated into 2 = 8 copies of P3. Every face F of P3 inherits a 2-colouring and lifts to a thrice-punctured sphere geodesically embedded in M3, tessellated into 4 copies of F .

The link of an ideal vertex v of P3 is a square coloured with 3 colours. By

Corollary 6 a star of the vertex v lifts to a single cusp in M3. The hyperbolic manifold M3 has 3 cusps, one above each vertex v of P .

Using Sage we have calculated the Betti numbers of M3:

b0 = 1, b1 = 3, b2 = 2.

We get of course χ(M3) = 0. Here and in the following more complicated manifolds, to compute the Betti number of Mn we use the formula stated in [8, Theorem 1.1], also explained in [13, Section 2.2].

4 1.8. The manifold M4. The hyperbolic polytope P4 ⊂ H is fully described in [28]. It has 10 facets, all isometric to P3.

The dual Gosset polytope 021 is the 4-dimensional rectified simplex. A plane projection of its 1-skeleton is shown in Figure 2. The figure shows that there are 5 lines that cross the center of the figure and contain each a pair of non-adjacent vertices (like those painted in blue and yellow in the figure). This yields a colouring of 021, where the 10 vertices are partitioned into 5 pairs. 10 GIOVANNI ITALIANO, BRUNO MARTELLI, AND MATTEO MIGLIORINI

Figure 2. A projection of the 1-skeleton of the rectified simplex

021 on the plane. Some caution is needed here and in the sub- sequent figures, since some edges are superposed, so two vertices that are connected by an edge on the plane projection may not be

so in 021. To clarify this ambiguity, we have chosen a blue vertex and painted in red the 6 vertices adjacent to it, in two cases (all the other cases are obtained by rotation). The blue and yellow vertices are non incident. In general, two vertices that belong to a line that crosses the center of the figure are non incident. There are 5 such lines, and they partition the 10 vertices into 5 pairs.

Dually, we equip P4 with this 5-colouring. This produces a hyperbolic 4-manifold 5 M4, tessellated into 2 = 32 copies of P4. We have χ(M4) = 32/16 = 2.

Each of the 5 ideal vertices of P4 can be seen dually in Figure 2 as a set of 6 vertices that span the 1-skeleton of an octahedron. One can identify these sets and discover that they always contain 5 distinctly coloured vertices. The chase of these octahedra can be quite enduring for the human eye and is absolutely trivial for a computer, so we have used Sage here and all the more in the subsequent bigger polytopes [34].

We deduce from Corollary 6 that every ideal vertex v of P4 lifts to a unique cusp in M4. Hence M4 has 5 cusps overall.

Using Sage we have calculated the Betti numbers of M4:

b0 = 1, b1 = 5, b2 = 10, b3 = 4.

We get of χ(M4) = 2 again.

5 1.9. The manifold M5. The hyperbolic polytope P5 ⊂ H is fully described in [29]. It has 16 facets, all isometric to P4.

The dual Gosset polytope 121 has 16 vertices. A convenient planar projection of its 1-skeleton is shown in Figure 3. From the figure we deduce that two vertices that are symmetric with respect to the origin (like those painted in blue and yellow HYPERBOLIC MANIFOLDS THAT FIBER ALGEBRAICALLY 11

Figure 3. A projection of the 1-skeleton of 121 on the plane. Some edges are superposed, so two vertices that are connected by an

edge in the planar projection may not be so in 121. To clarify this ambiguity, we have selected a blue vertex and painted in red the 10 vertices adjacent to it, in two cases (all the other cases are obtained by rotation). We deduce that the blue and yellow vertices are non incident. In general, two vertices that are symmetric with respect to the center of the figure are non incident. There are 8 such pairs, and this is the colouring that we choose.

in the figure) are not incident. This yields a colouring for 121 where the 16 vertices are partitioned into 8 pairs.

We equip P5 with this 8-colouring. This produces a hyperbolic 5-manifold M5, 8 tessellated into 2 = 256 copies of P5.

The polytope P5 has 10 ideal vertices; each corresponds dually in Figure 3 to 8 vertices that span a hyperoctahedron. Using Sage we discover that 2 ideal vertices of P5 are adjacent to only 4 distinctly coloured facets, while the other 8 are adjacent to 8. Note that 4 and 8 are precisely the minimum and maximum number of colours on a 4-cube. By Corollary 6 there are 24 = 16 cusps above each vertex of the first kind and 1 cusp above each vertex of the second kind. Hence M5 has 2·16+8·1 = 40 cusps overall.

Using Sage we have calculated the Betti numbers of M5:

b0 = 1, b1 = 24, b2 = 120, b3 = 136, b4 = 39.

We get of course χ(M5) = 0.

6 1.10. The manifold M6. The hyperbolic polytope P6 ⊂ H is fully described in [12]. It has 27 facets, each isometric to P5. 12 GIOVANNI ITALIANO, BRUNO MARTELLI, AND MATTEO MIGLIORINI

Figure 4. A projection of the 1-skeleton of 221 on the plane. Some edges are superposed, so two vertices that are connected by an

edge in the planar projection may not be so in 221. To clarify this ambiguity a little, we have picked a blue vertex and painted in red the 16 vertices adjacent to it. The only important information for us, that is obvious from the figure, is that the blue, yellow, and yellow vertices are mutually non incident. In general, three vertices that belong to a line that crosses the center of the figure are mutually non incident. There are 9 such lines, that partition the 27 vertices into 9 triplets. This is the colouring that we choose.

The dual Gosset polytope 221 has 27 vertices. Both P6 and 221 have many remarkable properties [10, 12]. To mention one, the 1-skeleton of 221 is the config- uration graph of the 27 lines in a general cubic surface!

A convenient planar projection of the 1-skeleton of 221 is shown in Figure 4, reproduced from [10, 12]. The figure shows that there are 9 lines that cross the center and contain each 3 mutually non-adjacent vertices (like those painted in blue and yellow in the figure), that we call triplets. We get a colouring for 221, where the 27 vertices are partitioned into 9 triplets. HYPERBOLIC MANIFOLDS THAT FIBER ALGEBRAICALLY 13

We equip P6 with this 9-colouring. This produces a hyperbolic 6-manifold M6, 9 tessellated into 2 = 512 copies of P6. We have χ(M6) = −512/8 = −64.

The polytope P6 has 27 ideal vertices. These can be visualized in Figure 4 as follows: for each vertex of 221, the 8 vertices that are not adjacent to it span a hyperoctahedron that is dual to an ideal vertex of P6. The 8 black or yellow vertices in the figure are one example.

Using Sage we discover that every ideal vertex of P6 is adjacent to 9 distinctly coloured facets, so it lifts to a unique cusp by Corollary 6. Hence M6 has 27 cusps.

Using Sage we have calculated the Betti numbers of M6:

b0 = 1, b1 = 18, b2 = 183, b3 = 411, b4 = 207, b5 = 26.

We get χ(M6) = −64 again.

7 1.11. The manifold M7. The hyperbolic polytope P7 ⊂ H has 56 facets, each isometric to P6.

The dual Gosset polytope 321 has 56 vertices. We will discover below that these can be partitioned into 14 sets of 4 mutually disjoint vertices, called quartets. This partition will be induced from a colouring of P8, that is in turn easily described using octonions.

We equip P7 with this 14-colouring. This produces a hyperbolic 7-manifold M7, 14 tessellated into 2 = 16384 copies of P7.

The polytope P7 has 126 ideal vertices. Using Sage we discover that 14 of these are incident to 6 distinctly-coloured facets, while the remaining 112 are incident to 12. Note that 6 and 12 are the minimum and maximum number of colours in a 8 2 6-cube. From Corollary 6 we deduce that M7 has 14 · 2 + 112 · 2 = 4032 cusps.

Using Sage we have calculated the Betti numbers of M7:

b0 = 1, b1 = 182, b2 = 6321, b3 = 41300, b4 = 55139, b5 = 24010, b6 = 4031.

We get of course χ(M7) = 0.

8 1.12. The manifold M8. The hyperbolic polytope P8 ⊂ H has 240 facets, each isometric to P7.

The dual Gosset polytope 421 has 240 vertices and can be described elegantly using octonions, much in the same way as the 4-dimensional 24-cell may be de- fined using quaternions. This viewpoint is crucial in this paper, so we introduce it carefully.

A 3-colouring for the 24-cell. To warm up, we start by recalling that the 24 vertices of the 24-cell are the quaternions

1 ±1, ±i, ±j, ±k, 2 (±1 ± i ± j ± k). 14 GIOVANNI ITALIANO, BRUNO MARTELLI, AND MATTEO MIGLIORINI

e6

e4 e1

e7

e3 e2 e5

Figure 5. The Fano plane. The circle should be interpreted as a line.

Two such vertices are adjacent along an edge if and only if their Euclidean scalar 1 4 product is 2 (we identify the quaternions space with the Euclidean R as usual). Every vertex is adjacent to 8 other vertices. We can assign 3 colours to the 24 vertices, by subdividing them into 3 sets of 8 vertices each, that we call octets. These are: (1) ±1, ±i, ±j, ±k; 1 (2) the elements 2 (±1 ± i ± j ± k) with an even number of minus signs; 1 (3) the elements 2 (±1 ± i ± j ± k) with an odd number of minus signs. The scalar product of two vertices lying in the same octet is an , so it is 1 never 2 . Therefore this indeed defines a colouring of the vertices of the 24-cell. Here is an algebraic description of this 3-colouring. The 24 vertices of the 24-cell described above form a group called the binary tetrahedral group. The 8 elements ±1, ±i, ±j, ±k form a normal subgroup of index 3, called the quaternion group and indicated with the symbol Q8. The octets are just the three lateral classes of Q8.

Octonions. We now turn to 421 and the octonions. For a nice introduction to the subject we recommend [4]. We describe an as a linear combination of 2 1, e1, e2, . . . , e7. We have ei = −1, and the multiplication of two distinct elements ei and ej is beautifully described by the Fano plane shown in Figure 5. The Fano plane is the projective plane over Z2 and it contains 7 points and 7 oriented lines: every line is a cyclically ordered triple of points as in the figure. For every i 6= j we have eiej = ±ek, where ek is the third vertex in the unique line containing ei and ej, and the sign is positive if and only if the line is cyclically oriented like ei → ej → ek. So for instance e1e2 = e4 and e1e6 = −e5. In general, we get

enen+1 = en+3 HYPERBOLIC MANIFOLDS THAT FIBER ALGEBRAICALLY 15 where the subscripts run modulo 7. The product is neither commutative nor asso- ciative: for every i, j, k we have

(eiej)ek = ±ei(ejek) where the sign is +1 if and only if ei, ej, ek belong to the same line in the Fano plane (which is always the case if i, j, k are not distinct).

A 15-colouring for the Gosset polytope 421. The 240 vertices of the Gosset polytope

421 are the octonions

±1, ±e1, ±e2, ±e3, ±e4, ±e5, ±e6, ±e7, 1 1 2 (±1 ± en ± en+1 ± en+3), 2 (±en+2 ± en+4 ± en+5 ± en+6) where n runs modulo 7. Although we will not need this information, we mention that these are (up to rescaling) precisely the 240 non-trivial elements in the E8 lattice of smallest .

We have 16 elements of type ±1 or ±ei. Each line l in the Fano plane contains 1 three vertices en, en+1, en+3 and determines 16 elements of type 2 (±1±en ±en+1 ± 1 en+3) and 16 elements of type 2 (±en+2 ± en+4 ± en+5 ± en+6), so we indeed get 16 + 7 · 16 + 7 · 16 = 15 · 16 = 240 vertices overall.

Two vertices of 421 are connected by an edge if and only if their Euclidean scalar 1 product is 2 . One can check that every vertex is adjacent to 56 other vertices (coherently with the fact that P7 has 56 facets). Similarly to what we did with the 24-cell, we can assign 15 colours at the vertices of 421 by subdividing the 240 vertices into 15 sets of 16 elements each; we call each such set a hextet. The hextets are:

(1) ±1, ±e1, ±e2, ±e3, ±e4, ±e5, ±e6, ±e7; 1 1 (2) the elements 2 (±1 ± en ± en+1 ± en+3) and 2 (±en+2 ± en+4 ± en+5 ± en+6) with an even number of minus signs; 1 1 (3) the elements 2 (±1 ± en ± en+1 ± en+3) and 2 (±en+2 ± en+4 ± en+5 ± en+6) with an odd number of minus signs. The hextets of type (2) and (3) depend on the choice of n modulo 7. So we get 1 + 2 · 7 = 15 hextets overall. One can check that the scalar product of two vertices 1 lying in the same hextet is always an integer, so it is never 2 . Therefore this is really a colouring of the vertices of 421.

Algebraic description. There is an algebraic interpretation for the colouring of 421 analogous to that of the 24-cell. Some caution is needed when passing from quater- nions to octonions: first, the product of octonions is notoriously nonassociative; second, contrary to a common mistake (see [9, Chapter 9] for a discussion), and as proved by Coxeter [11], the 240 vertices just described are not closed under 16 GIOVANNI ITALIANO, BRUNO MARTELLI, AND MATTEO MIGLIORINI multiplication! Indeed the product of the two vertices 1 1 1 (1 + e + e + e ) · (1 + e + e + e ) = (e + e + e + e ) 2 1 3 7 2 1 2 4 2 1 3 4 6 is not a vertex! We could fix this via a single reflection that transforms the 240 vertices into a multiplicatively closed set (this is explained in [9, Section 9.2]), thus obtaining another isometric description of 421, but we do not really need this here, so we just keep them as they are. The only thing that we need is that the 240 octonions are closed under left multiplication by each of the 16 elements in the hextet S = {±1, ±ei}, a fact that can be verified easily. The set S is closed under multiplication, but it is not a group since it is not associative. One can also verify that the left multiplication by each element of S preserves each hextet, and that this “action” of S is free and transitive, in the sense that for very pair of distinct elements in a hextet there is a unique element of S that sends the first to the second by left-multiplication.

The manifold M8. We equip P8 with this 15-colouring. This produces a hyperbolic 15 8-manifold M8, tessellated into 2 = 32768 copies of P8. We have χ(M8) = 215 · 17/2 = 278528.

The polytope P8 has 2160 ideal vertices. Using Sage we discover that 240 of them are incident to only 7 distinctly coloured facets, while the remaining 1920 are incident to 14. Note that 7 and 14 are the minimum and maximum number of 8 colours in a 7-cube. From Corollary 6 we deduce that M8 has 240·2 +190·2 = 65280 cusps.

Using Sage we have calculated the Betti numbers of M8:

b0 = 1, b1 = 365, b2 = 33670, b3 = 583290,

b4 = 1783226, b5 = 1346030, b6 = 456595, b7 = 65279.

We get χ(M8) = 278528 again.

1.13. Back to the polytope P7. The polytope P7 is a facet of P8. We think of

P7 as the facet dual to the vertex 1 of 421. As we already said, we equip P7 with the colouring induced by the 15-colouring of P8 just introduced.

We study this inherited colouring of P7. We think of 321 as the link figure of the vertex 1 of 421. The vertices of 421 adjacent to 1 are precisely those of the form

1 2 (1 ± en ± en+1 ± en+3) where n runs modulo 7. So we get 7 · 8 = 56 vertices, as required. These vertices 1 are contained in the hyperplane x0 = 2 and their is 321. Two such 1 vertices are connected by an edge in 321 if and only if their scalar product is 2 . The 15-colouring of 421 induces a 14-colouring of 321, that partitions the 56 vertices into 14 sets of four vertices each, that we call quartets. Each quartet HYPERBOLIC MANIFOLDS THAT FIBER ALGEBRAICALLY 17

P3 P4 P5 P6 P7 P8 3 pairs 5 pairs 8 pairs 9 triplets 14 quartets 15 hextets

Table 3. The colouring type of each P3,...,P8.

Volume χ Cusps

M3 8L(2) ∼ 7.28 0 3 2 M4 8π /3 ∼ 26.3 2 5

M5 224ζ(3) ∼ 269 0 40 3 3 M6 512π /15 ∼ 1.06 · 10 −64 27 5 M7 131072L(4) ∼ 1.30 · 10 0 4032 4 6 M8 4456448π /105 ∼ 4.13 · 10 278528 65280

Table 4. The volume, the Euler characteristic, and the number

of cusps of each hyperbolic n-manifold Mn.

1 consists of the vertices 2 (1 ± en ± en+1 ± en+3) that share the same n and the same parity of the minus signs.

Table 3 summarizes the chosen colouring type for each P3,...,P8.

1.14. Volumes. We show the volumes of the hyperbolic manifolds M3,...,M8 in Table 4. In even dimension n = 2m we have used the Gauss-Bonnet formula

Vol(P ) = (−2π)m/(n − 1)!! · χ(P ).

In odd dimension, we have

Vol(P3) = L(2) ∼ 0.91, Vol(P5) = 7ζ(3)/8 ∼ 1.05, Vol(P7) = 8L(4) ∼ 7.92. The symbols ζ and L indicate the Riemann and Dirichlet functions, see [27, 12].

2. The algebraic fibrations

We have constructed some hyperbolic manifolds M3,...,M8, and our aim is now 1 to build some nice map f : Mn → S for all n = 3,..., 8. We produce these maps by assigning to each Pn an appropriate state, as prescribed by [15]. We then study them by applying some fundamental results of [6].

2.1. States. Let P ⊂ Xn be a right-angled polytope in some space Xn = Hn, Rn, or Sn. Following [15], a state is a partition of the facets of P into two subsets, that we denote as I (in) and O (out). Every facet thus inherits a status I or O. Let P be equipped with a colouring with c colours. This induces a free action of c Z2 on the set of all the states of P , in the following way. For every j ∈ {1, . . . , c}, 18 GIOVANNI ITALIANO, BRUNO MARTELLI, AND MATTEO MIGLIORINI

001 101

011 111

000 100

010 110

Figure 6. The 1-skeleton of the dual cubulation C for M3, tessel- v lated into 8 polyhedra P3 . The polyhedron P3 has three colours, and each colour is painted in two faces. The vertices of C are 3 identified with Z2. There are two edges connecting v and v + ej corresponding to the two faces in P v with the same colour j, for each j = 1, 2, 3.

the basis element ej acts by reversing the IO status of every facet of P coloured by j, while leaving the status of the other facets unaffected. The action is free, hence each orbit consists of 2c distinct states.

2.2. Diagonal maps. As discovered in [15], the choice of a colouring and a state for a right-angled polytope P induce both a manifold M and a diagonal map M → S1. Shortly:

colouring + state on P =⇒ manifold M + diagonal map f : M → S1.

We explain how this works. We already know how to construct M from the colouring of P , so it remains to show how a state induces a map f : M → S1. c v c The manifold M is tessellated into the 2 polytopes P with varying v ∈ Z2. Since these are right-angled, the tessellation is dual to a cube complex C with 2c vertices. We work in the piecewise-linear category (see [30] for an introduction) and think of C as piecewise-linearly embedded inside M. If P has some ideal vertices n (as it will be the case with all the polytopes Pn ⊂ H that we consider here), the complement M \ C consists of open cusps, so there is a deformation retraction r : M → C. The cube complex C is a spine of M. We indicate the vertex of C dual to P v simply as v, so the vertices of C are c c identified with Z2. Here v stands both for a vector of Z2 and a vertex of C. The edges of C are dual to the facets of the tessellation: an edge of C connects v and v + ej if the dual facet F is coloured as j. So in particular there are k distinct HYPERBOLIC MANIFOLDS THAT FIBER ALGEBRAICALLY 19

Figure 7. Every square (and hence every k-cube) of the cubula- tion has its opposite edges oriented coherently as shown here.

edges connecting v to v + ej, where k is the number of facets in P coloured with j.

In all the colourings that we have chosen for the polytopes Pn the number k does not depend on the colour j. The 1-skeleton of C for P = P3 is shown in Figure 6.

15 Example 8. If we consider P = P8 with its 15-colouring, there are 2 vertices in

C, and 16 edges connecting v to v + ej for every v and every j.

Let now s be a fixed state for P . The state s induces an orientation on all the edges of C, in the simplest possible way: consider an edge connecting v and v + ej, where the j-th component of v is zero, that is vj = 0. The edge is dual to some facet of the tessellation that is a precise identical copy of a facet F of P . If the status of F is O, we orient the edge outward, that is from v to v + ej, while if it is

I we orient it inward, from v + ej to v. By construction, this orientation is coherent, that is on every square of C (and hence on any k-cube) the orientations of two opposite edges match as in Figure 7. This crucial fact allows us to apply Bestvina – Brady theory [6]. We identify every k-cube of C with the standard k-cube [0, 1]k ⊂ Rk, so that the orientations on the edges of C match with the orientations of the axis in Rk. The diagonal map on the standard k-cube is

k 1 [0, 1] −→ S = R/Z, x 7−→ x1 + ··· + xk. The diagonal maps on the k-cubes of C match to give a well-defined continuous piecewise-linear map C → S1. By pre-composing it with the deformation retraction r : M → C, we finally get a diagonal map

f : M → S1.

This is the main protagonist of our construction. The diagonal map induces a 1 homomorphism f∗ : π1(M) → π1(S ) = Z. A dichotomy arises here:

Proposition 9. Precisely one of the following holds: (1) The facets of P with the same colour also have the same status. In this case f is homotopic to a constant. 20 GIOVANNI ITALIANO, BRUNO MARTELLI, AND MATTEO MIGLIORINI

001 101

011 111

000 100

010 110

Figure 8. We assign a state to P3 where faces with the same colours have opposite status. We get the orientation of the 1- skeleton of C shown here.

(2) There are two facets in P with equal colour and opposite status. In this case 1 the homomorphism f∗ : π1(M) → π1(S ) = Z is non-trivial with image 2Z. Proof. If (1) holds, all the edges joining two given vertices of C are oriented in the same way, and we may lift the map f : M → S1 to a map f˜: M → R as follows: c send every vertex v ∈ Z2 of C to the maximum number of edges exiting from v and pointing outward to distinct vertices, then extend f˜ diagonally to cubes. If (2) holds, there are two edges joining the same pair of vertices with opposite orientation, that form a loop that is sent to ±2 along f∗. Moreover 1 6∈ Im(f∗) because the 1-skeleton of C is naturally bipartited into even and odd vertices, according to the parity of v1 + ··· + vc.  The case (1) is not so interesting: all the examples that we construct here on the hyperbolic manifolds Mn will be of type (2). In (2), since Imf∗ = 2Z, one may 1 decide to replace f with a lift along a degree-2 covering of S to get a surjective f∗.

Corollary 10. If all the facets of P have distinct colours, the diagonal map f is homotopically trivial, for every choice of a state.

This partially explains why we insisted in finding colourings for Pn with the smallest possible number of colours.

Example 11. In P3 we will choose the following state: for every pair of faces with the same colour, assign I to one face and O to the other (the choice will not affect much the result, as we will see). The resulting 1-skeleton of C is then oriented as in Figure 8. By Proposition 9 the homomorphism f∗ is not trivial.

2.3. States and orbits. Let a right-angled P ⊂ Xn be equipped with a colouring and a state s. These determine a diagonal map f : M → S1. We now would like HYPERBOLIC MANIFOLDS THAT FIBER ALGEBRAICALLY 21 to study f and how it depends on s. A powerful machinery is already available for this task and is described in [6]. Recall that s induces a coherent orientation of the edges of the dual cubulation C. It also induces a state on every polytope P v of the tessellation, as follows: every facet F of P v is dual to an edge e of C, and hence inherits a transverse orientation from that of e. We assign the status O or I to F according to whether the transverse orientation points outward or inward with respect to P v. It is easy to see that the state induced on P v is precisely v(s), the result of the action of v on the initial state s, as described in Section 2.1. Summing up, the polyhedron P 0 has the initial state s, while P v inherits the c state v(s) for any v ∈ Z2. The following proposition says that states that lie in the same orbits produce equivalent diagonal maps.

0 c Proposition 12. Two states s, s that lie in the same orbit with respect to the Z2 action produce two diagonal maps f, f 0 : M → S1 that are equivalent up to some isometry of M, that is there is an isometry ψ : M → M with f = f 0 ◦ ψ.

0 c Proof. If s = w(s) for some w ∈ Z2, we pick the isometry ψ : M → M that sends v v+w 0 P to P via the identity map. We get f = f ◦ ψ.  2.4. Ascending and descending links. Let a right-angled P ⊂ Xn be equipped with a state s. Let Q be a Euclidean polytope combinatorially dual to P . When

P = Pn, then Q is a Gosset polytope. The state s induces a dual state on Q, that is the assignment of a status I or O to each vertex of Q. We note that if we remove the hyperoctahedra from Q (that correspond to the ideal vertices of P ) we get a flag simplicial complex Q˙ (because P is right-angled, and hence simple). Following [6], we define the ascending link (respectively, descending link) as the subcomplex of Q˙ generated by the vertices with status O (respectively, I). Since Q˙ is a flag complex, these subcomplexes are determined by their 1-skeleton. Let now P be equipped with both a colouring and a state. We get a manifold M 1 c and a diagonal map f : M → S . For every vertex v ∈ Z2 of the dual cubulation C, the link of v in C is precisely link(v) = Q˙ , and it inherits the state v(s) of P v. The status of a vertex of link(v) is I or O according to whether the corresponding oriented edge of C points inward or outward with respect to v. The ascending and descending links at v are denoted respectively as link↑(v) and link↓(v), and they are disjoint subcomplexes of link(v). 1 The diagonal map f : M → S induces a homomorphism f∗ : π1(M) → Z. We are interested in its kernel H.

Theorem 13. [6, Theorem 4.1] The following holds:

• If link↑(v), link↓(v) are connected for every v, then H is finitely generated.

• If link↑(v), link↓(v) are simply connected for every v, H is finitely presented. 22 GIOVANNI ITALIANO, BRUNO MARTELLI, AND MATTEO MIGLIORINI

2.5. Legal states. Following [15], a state s on P is legal if the ascending and descending links that it determines on the dual flag simplicial complex are both c connected. The group Z2 acts on the set of all states of P , and an orbit is legal if it consists only of legal states. As noted in [15], Theorem 13 implies the following.

Corollary 14. A legal orbit defines a diagonal map f : M → S1 with finitely generated H = ker f∗.

The chase of a legal orbit is the combinatorial game introduced in [15]. After introducing the rules of the game, the authors exhibited some legal orbits on two remarkable right-angled polytopes in H4, namely the ideal 24-cell and the compact right-angled 120-cell [15], so providing the first algebraically fibering hyperbolic 4- manifolds. Here we play with the right-angled polytopes Pn and find some legal orbits on all of them. More than that, we find some even better kind of orbits in the cases n = 3, 7, 8, that we call 1-legal.

2.6. 1-legal states. We extend the nomenclature of [15] by saying that a state s is 1-legal if its ascending and descending links are both simply connected. An orbit is 1-legal if it consists only of 1-legal states. Here is a consequence of Theorem 13.

Corollary 15. A 1-legal orbit defines a diagonal map f : M → S1 with finitely presented H = ker f∗.

2.7. The Euler characteristic check. In the following pages we will double count the Euler characteristic of our manifolds as a safety check. If a colouring and a state on P produce a manifold M and a diagonal function f : M → S1, we always have X  (1) χ(M) = 1 − χ(link↑(v)) . c v∈Z2

The same formula holds with the descending link link↓(v). We say that the integer

1 − χ(link↑(v)) is the contribution of v to the Euler characteristic of M. Note that a contractible ascending link contributes with zero, while a k-sphere contributes with (−1)k+1.

We now construct a legal orbit on each individual polytope P3,...,P8. We have used a code written with Sage to analyse all these cases; both the code and the resulting data are available from [34].

2.8. A legal orbit for P3. In the chosen colouring for P3 the facets are partitioned into three pairs. For every pair, assign the status O to one facet and I to the other, arbitrarily. The orbit of this state s is independent of this choice and consists precisely of all the 23 = 8 states that can be constructed in this way. The 8 states reduce to 2 up to isomorphism, and they are shown in Figure 9. The ascending and descending links are both either a triangle or two segments connected along an endpoint. In both cases they are contractible. One can in fact HYPERBOLIC MANIFOLDS THAT FIBER ALGEBRAICALLY 23

Figure 9. We exhibit a state by colouring the vertices in black and white, with black (white) corresponding to the status I (O).

There are only two states in the orbit of P3 up to isomorphism, and they both yield contractible simplicial complexes as ascending and descending links: a triangle and two segments joined along an endpoint. verify that the conditions of [5, Theorem 9] are satisfied and hence the diagonal 1 map f : M3 → S can be smoothened to become a fibration (but we will not need this here). The ascending and descending links are simply connected and hence the orbit is 1-legal. The formula (1) holds since χ(M3) = 0 and each contractible link con- tributes with zero to the sum.

2.9. A legal orbit for P4. In the chosen colouring for P4 the facets are partitioned into five pairs. As in the previous case, we consider the orbit that consists of all the states that assign distinct statuses to each pair. We get 25 = 32 states. These states reduce quite easily to 4 up to isomorphism, depicted in Figure 10. As shown in the figure, the ascending and descending links are always connected, so the orbit is legal. However, we note that the orbit is not 1-legal, since in the first case both the ascending and descending link are circles. The first case occurs only in 2 of the 32 states. In fact, one can verify that the ascending and descending links in the first case form a Hopf link in S3, if considered in the boundary of the Gosset polytope, and that the conditions of [5, Theorem 9] are satisfied, so the diagonal function 1 f : M4 → S can be smoothened to a circle-valued Morse function with two index-2 critical points. This is the best that we can get in dimension 4 (but we will not need this fact here).

The formula (1) holds since χ(M4) = 2 and the 2 states of the first kind con- tribute each with 1, while all the others contribute with 0. 24 GIOVANNI ITALIANO, BRUNO MARTELLI, AND MATTEO MIGLIORINI

Figure 10. We exhibit a state by colouring the vertices in black and white, with black (white) corresponding to the status I (O).

There are only four states in the orbit of P4 up to isomorphism. We show here the descending link, generated by the black vertices. In the first case we get a circle, while in the other cases we always get a contractible complex made of 3 triangles, 2 triangles, and 1 tetrahedron and 1 triangle respectively.

2.10. A legal orbit for P5. In the chosen colouring for P5 the facets are parti- tioned into 8 pairs. As in the previous cases, we consider the orbit that consists of all the states that assign different statuses to each pair. We get 28 = 256 states. Each state produces a pair of ascending and descending links. Since these are flag simplicial complexes, only the graphs formed by their 1-skeleta are important. Using Sage, we have listed all the resulting 512 graphs and discovered that they reduce to only 7 up to isomorphism. These 7 graphs are those generated by the black vertices in Figure 11. We can check by hand (or with Sage) that the first 4 graphs in the figure gen- erate a contractible simplicial complex. The remaining three generate a simplicial 2 1 3 complex that collapses respectively to S , a wedge of three circles ∨3S , and S . 2 1 The complexes that collapse to S and ∨3S are shown in Figure 12. The complex that collapses to S3 is actually homeomorphic to S3, and it is a hyperoctahedron (with empty interior), decomposed into 16 tetrahedra. It corresponds dually to an ideal vertex of P5. In all cases the simplcial complex is connected, so the orbit is legal. It is not always simply connected, so the orbit is not 1-legal.

The formula (1) holds since χ(M5) = 0 and using Sage we find that we get 2 1 3 32 occurrences of S , 8 of ∨3S , and 8 of S . Their contributions to the Euler characteristic are 32 · (−1) + 8 · 3 + 8 · 1 = 0.

2.11. A legal orbit for P6. In the chosen colouring for P6 the 27 facets are par- titioned into 9 triplets. As opposite to the previous cases, there does not seem to be a natural choice of an orbit here. However, a brute computer search shows HYPERBOLIC MANIFOLDS THAT FIBER ALGEBRAICALLY 25

Figure 11. We exhibit a state by colouring the vertices in black and white, with black (white) corresponding to the status I (O).

Every ascending or descending link for P5 is isomorphic to one of the 7 descending links shown here. 26 GIOVANNI ITALIANO, BRUNO MARTELLI, AND MATTEO MIGLIORINI

Figure 12. Two simplicial complexes that collapse to S2 and 1 ∨3S . The first is an octahedron (with empty interior) with two tetrahedra attached to a pair of opposite faces. The second consists of two tetrahedra and four edges joining them.

that there are many legal orbits for P6. For instance, we may take s as the state where the 9 exterior vertices of 221 in Figure 4 have status O and the remaining 18 vertices have status I. With this choice the orbit of 29 states produces 210 = 1024 graphs to analyse (a descending and an ascending link for each vertex: recall that since these are flag only the 1-skeleton is important). Sage reduces them to 44 up to isomorphism, and shows that they are all connected. We could not find a single 1-legal orbit, but admittedly we have not checked all the possible orbits. For instance, with the state s just described we discover that 2 of the 44 simplicial complexes are not simply connected (their fundamental group is Z). Sage also checks that (1) holds: the non-contractible ascending links are homologically like 10 copies of S1, 81 copies of S2, 28 copies of S3, 9 copies of S4, 2 3 15 copies of ∨2S , and 9 copies of ∨2S . By summing their contributions we get

10 − 81 + 28 − 9 − 30 + 18 = −64 = χ(M6).

2.12. A legal orbit for P7. In the chosen colouring for P7 the 56 facets are par- titioned into 14 quartets. We see P7 as the facet of P8 dual to the vertex 1 in 421.

We will define below a state for P8, and this induces one s for P7 in the obvious way: every facet of P7 inherits the status of the adjacent facet in P8. This state s inherited in this way turns out to be balanced with respect to the colouring, in the following sense: there is an even number 2m of facets sharing the same colour, and precisely half of them are given the status I, and the other half the status O. If a state is balanced, then every other state in the orbit also is. Here m = 2 and in each quartet two facets receive the status I and two the status O. The orbit of s consists of 214 states, each contributing with an ascending and descending link. Using Sage we are pleased to discover that the resulting 215 = 32768 graphs reduce to only 106 up to isomorphism. (This is probably due to the many symmetries of s that are inherited from P8.) HYPERBOLIC MANIFOLDS THAT FIBER ALGEBRAICALLY 27

Using Sage we also see that all the simplicial complexes generated by the 106 graphs are simply connected. Therefore the orbit is 1-legal. With Sage we have also checked (1). All the data can be found in [34].

2.13. A legal orbit for P8. In the chosen colouring for P8 the 240 facets are partitioned into 15 hextets. How can we find a legal orbit for P8? The numbers 240 are overwhelming: the polytope P8 has 240 facets, so there are 2 possible states to choose from. Each orbit consists of 215 distinct states, and we would like to find one orbit where all these 215 states are legal, or even better 1-legal. A brute force computer search is out of reach. To define a legal state we take inspiration from the 24-cell sitting inside quater- nions space, since this has some strong analogies with the Gosset polytope 421 sitting in octonions space as we already noticed above. We have already exploited this analogy when we fixed a convenient colouring for P8, and we do it again now to design a convenient state.

A state for the 24-cell. A nice legal state for the 24-cell was constructed in [5] as follows. Recall that its 24 vertices are divided into three octets: these are 1 ±1, ±i, ±j, ±k, the elements 2 (±1 ± i ± j ± k) with an even number of minus signs, and those with an odd number of minus signs. Consider the group G = {±1, ±i} and its action on the 24 vertices by left- multiplication. We can verify easily that each octet is invariant by this action, and is subdivided into two orbits of four elements each. We assign arbitrarily the status I to one orbit, and O to the other. The resulting state s is balanced (see the definition above), and also legal, as it was in fact already observed in [15]. The ascending and descending links are each homeomorphic to a G-invariant annulus as in Figure 13, so they are connected but not simply connected (the state is not 1-legal). The two G-invariant annuli form altogether a banded Hopf link in S3. 3 3 The orbit of s along the action of Z2 consists of the 2 states obtained from s by interchanging the I/O status of some octet. The geometry of the 24-cell is so extraordinary that these 23 states are all isomorphic [5]. In particular, the orbit is legal (but it is not 1-legal). The choice of which orbit is I and which is O inside each octet is irrelevant, since different choices lead to the same orbit.

A state for 421. We now try to mimic the above construction for 421. The 240 ver- tices are partitioned into 15 hextets, that is ±1, ±e1, ±e2, ±e3, ±e4, ±e5, ±e6, ±e7, 1 1 the elements 2 (±1 ± en ± en+1 ± en+3) and 2 (±en+2 ± en+4 ± en+5 ± en+6) with an even number of minus signs, and those with an odd number of minus signs, with the integer n varying modulo 7.

It is now natural to consider the quaternion group Q = {±1, ±e1, ±e2, ±e4} and its “action” on the 240 vertices of 421 by left-multiplication. This is the analogue of the group G = {±1, ±i} defined above, roughly because taking quaternions inside 28 GIOVANNI ITALIANO, BRUNO MARTELLI, AND MATTEO MIGLIORINI

Figure 13. The ascending and descending link are both an an- nulus decomposed into 12 triangles, and altogether they form two annuli that collapse onto a Hopf link in S3. The figure (taken from [5]) shows the vertices of the 24-cell, with their 3-colouring (Blue / Red / Yellow) and their state (the vertices with a O status have an additional black circle). Only some edges of the 24-cell are shown for the sake of clarity: more edges should be added between each yellow vertex and its 8 neighbours. octonions looks like taking complex numbers inside quaternions. However, this is not really a because octonions are not associative and some caution is needed.

We already know that the set S = {±1, ±e1,..., ±e7} acts freely and transitively by left-multiplication on every hextet (that is, for every pair of elements in the hextet there is a unique element in S whose left-multiplication sends the first to the second). We pick the following 15 base elements, one inside each hextet:

1, 1 + en + en+1 + en+3, −1 + en + en+1 + en+3 where n runs modulo 7. We consider inside each hextet the 8 elements obtained by left-multiplying the base element by the elements of Q. We assign to these 8 elements the status O, and the status I to the remaining 8 of the hextet. We have defined a balanced state s. The orbit consists of 215 balanced states. By analogy with the 24-cell, it would be tempting to guess that the 215 states in the orbit are all isomorphic, and maybe that the ascending and descending links are homotopic to two copies of S3 linked in S7. This is however impossible by a Euler characteristic argument, due to the fact that the 24-cell has χ = 1 while HYPERBOLIC MANIFOLDS THAT FIBER ALGEBRAICALLY 29

χ(P8) = 17/2 is much bigger. In general, one should not push the analogies too far: the situation is intrinsically more complicated here. We will come back to this point below. Using Sage we have determined the ascending and descending links of each of the 215 = 32768 states. Note that each graph can have up to 240 vertices, and we have 216 = 65536 graphs to analyse. Luckily, these graphs reduce up to isomorphism to only 185. With Sage we discover that each of the 185 simplicial complexes generated by these graphs is connected and simply connected. Therefore the orbit is 1-legal. We also checked (1). Both sides give (quite reassuringly) the same number 278528. This equations also explains a fact we alluded to few lines above. Since

χ(P8) = 17/2, the average contribution to the Euler characteristic of an ascending link is 17/2, which is a relatively big number if compared to the Euler characteristic of the other polytopes considered above. It is certainly impossible that all the ascending links are S3, since their individual contribution is 1. The ascending links that we find with Sage are indeed quite complicated, much more than those discovered with P3,...,P7. They can be found in [34].

2.14. The restriction of f to the cusps. In our analysis we have briefly men- tioned the fact that when n = 3, 4 the chosen orbits satisfy the conditions of [5, 1 Theorem 9] and f : Mn → S can be smoothened to become a smooth circle-valued Morse function of the nicest possible type: a fibration for n = 3, and a perfect circle-valued Morse function for n = 4. One may wonder if this is also the case when n ≥ 5, and indeed this was our hope at the beginning of our study: it would be extremely interesting to find a fibration on an odd-dimensional hyperbolic manifold of dimension 5 or 7. We show that this is not the case, for any possible choice of initial state, a serious obstruction being the restriction of f to the cusps of Mn.

∼ n−1 Proposition 16. For n = 5,..., 8, there is some cusp C = T × [0, +∞) ⊂ Mn where the restriction f|C is homotopic to a constant. This holds for any possible choice of a state for Pn.

Proof. By our analysis above, it follows that when n = 5, 7, 8 there is an ideal vertex v of Pn whose 2(n−1) incident facets have all distinct colours. By Corollary 10, the restriction of f to any cusp C lying above v is homotopically trivial. When n = 6, each of the 27 vertices v has a 5-cube link with exactly one pair of opposite faces sharing the same colour. In each of the 9 triplets, every pair is an opposite pair of facet of this kind, for some ideal vertex v. For any choice of a state, there will be one such pair with the same status, because the three statuses on a triple cannot be all distinct. By Proposition 9 the restriction of f to this cusp is null-homotopic.  30 GIOVANNI ITALIANO, BRUNO MARTELLI, AND MATTEO MIGLIORINI

2.15. The geometrically infinite coverings. For every n = 3,..., 8, the kernel 1 of f∗ : π1(Mn) → π1(S ) = Z is a normal subgroup H / π1(Mn) of infinite index. Corollaries 14 and 15 imply the following.

Theorem 17. The normal subgroup H is finitely generated, and also finitely pre- sented when n = 3, 7, 8. The limit set of H is the whole sphere ∂Hn = Sn−1.

The limit set is the whole sphere because H is normal in π1(Mn) and Mn has finite volume. In particular, the hyperbolic n-manifold

n Mfn = H /H that covers Mn is geometrically infinite. We are particularly interested in the dimensions n = 5,..., 8.

Theorem 18. When 5 ≤ n ≤ 8, the hyperbolic manifold Mfn has infinitely many toric cusps. In particular the Betti number bn−1(Mfn) = ∞ is infinite and π1(Mfn) = H is not Fn−1.

Proof. The restriction of f to some cusp is null-homotopic by Proposition 16. Therefore the cusp lifts to infinitely many copies of itself in Mfn. 

Recall that a group H is of type Fm if there exists a K(H, 1) with finite m- skeleton [6]. If H is Fm, the Betti number bm(H) is obviously finite.

Corollary 19. When n = 7, 8, the fundamental group of the hyperbolic manifold

Mfn is finitely presented but not Fn−1.

References

[1] I. Agol – D. Long – A. Reid, The Bianchi groups are separable on geometrically finite subgroups, Ann. Math., 153 (2001), 599–621. [2] I. Agol – M. Stover, Congruence RFRS towers, arXiv:1912.10283 [3] M. Aschenbrenner – S. Friedl – H. Wilton, “3-manifold groups,” EMS Series of Lectures in Mathematics (2015), 215 pages. [4] J. Baez, The octonions, Bull. Amer. Math. Soc. 39 (2001), 145–205. [5] L. Battista – B. Martelli, Hyperbolic 4-manifolds with perfect circle-valued Morse func- tions, arXiv:2009.04997 [6] M. Bestvina – N. Brady, Morse theory and finiteness properties of groups, Invent. Math., 129 (1997), 445–470. [7] B. Bowditch – G. Mess, A 4-dimensional Kleinian group, Transactions of AMS 14 (1994), 391–405. [8] S. Choi – H. Park, Multiplicative structure of the cohomology ring of real toric spaces, Homology, Homotopy and Applications 22 (2020), 97–115. [9] J. Conway – D. Smith, “On Quaternions and Octonions: Their Geometry, Arithmetic, and Symmetry”, A K Peters, Ltd., Natick, MA, 2003, 159 pages. [10] H. Coxeter, The polytope 221 whose twenty-seven vertices correspond to the lines on the general cubic surface. Amer. J. Math, 62 (1940), 457–486. [11] , Integral Cayley numbers (1946), Duke Math. J., 13, 561–578. HYPERBOLIC MANIFOLDS THAT FIBER ALGEBRAICALLY 31

[12] B. Everitt – J. Ratcliffe – S. Tschantz, Right-angled Coxeter polytopes, hyperbolic six- manifolds, and a problem of Siegel, Math. Ann. 354 (2012), 871–905. [13] L Ferrari – A. Kolpakov – L. Slavich, Cusps of Hyperbolic 4-Manifolds and Rational Homology Spheres, arXiv:2009.09995 [14] T. Gosset, On the regular and semi-regular figures in space of n dimensions, Messenger Math. 29 (1900), 43-48. [15] K. Jankiewicz – S. Norin – D. T. Wise, Virtually fibering right-angled Coxeter groups, to appear in J. Inst. Math. Jussieu. [16] M. Kapovich, On the absence of Sullivan’s cusp finiteness theorem in higher dimensions, in “Algebra and analysis” (Irkutsk, 1989), Amer. Math. Soc., Providence, RI, 1995, 77–89. [17] , Kleinian groups in higher dimensions, Geometry and dynamics of groups and spaces, 487–564, Progr. Math., 265, Birkhuser, Basel, 2008. [18] , Non-coherence of arithmetic hyperbolic lattices, Geometry & Topology 17 (2013), 39–71. [19] M. Kapovich – L. Potyagailo, On the absence of Ahlfors’ finiteness theorem for Kleinian groups in dimension three, Topol. Appl. 40 (1991), 83–91. [20] , On absence of Ahlfors’ and Sullivan’s finiteness theorems for Kleinian groups in higher dimensions, Siberian Math. Journ., 32 (1991), 227–237. [21] M. Kapovich – L. Potyagailo – E. Vinberg, Non-coherence of some non-uniform lattices n in Isom(H ), Geometry and Topology Monographs 14 (2000), 335–351. [22] D. Kielak, Residually finite rationally-solvable groups and virtual fibring. arXiv:1809.09386, To appear in J. Amer. Math. Soc. [23] A. Kolpakov – B. Martelli, Hyperbolic four-manifolds with one cusp, Geom. & Funct. Anal. 23 (2013), 1903–1933. [24] A. Kolpakov – L. Slavich, Hyperbolic four-manifolds, colourings and mutations, Proc. London Math. Soc. 113 (2016), 163–184. [25] L. Potyagailo, The problem of finiteness for Kleinian groups in 3-space, “Knots 90”, de Gruyter (1992). [26] L. Potyagailo – E- V. Vinberg, On right-angled reflection groups in hyperbolic spaces, Comment. Math. Helv. 80 (2005), 63–73. [27] J. Ratcliffe – S. Tschantz, Volumes of integral congruence hyperbolic manifolds, J. Reine Angew. Math. 488 (1997), 55–78. [28] , The volume spectrum of hyperbolic 4-manifolds, Experiment. Math. 9 (2000), 101– 125. [29] , Integral congruence two hyperbolic 5-manifolds, Geom. Dedicata 107 (2004), 187– 209. [30] C. Rourke – B. Sanderson, “Introduction to piecewise-linear topology,” Springer–Verlag 1972. [31] G. Scott, Finitely generated 3-manifold groups are finitely presented, J. London Math. Soc. 6 (1973), 437–440. [32] J Stallings, On fibering certain 3-manifolds, in ‘Topology of 3-manifolds and related topics” (Proc. The Univ. of Georgia Institute, 1961), pages 95–100. Prentice-Hall, Englewood Cliffs, N.J., 1962. [33] , A finitely presented group whose 3-dimensional integral homology is not finitely generated, Amer. J. Math. 85 (1963), 541–543. [34] http://people.dm.unipi.it/martelli/research.html 32 GIOVANNI ITALIANO, BRUNO MARTELLI, AND MATTEO MIGLIORINI