This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Totally geodesic surfaces in twist knot complements

Khanh Le and Rebekah Palmer
Abstract

In this article, we give explicit examples of infinitely many non-commensurable (non-arithmetic) hyperbolic 33-manifolds admitting exactly kk totally geodesic surfaces for any positive integer kk, answering a question of Bader, Fisher, Miller and Stover. The construction comes from a family of twist knot complements and their dihedral covers. The case k=1k=1 arises from the uniqueness of an immersed totally geodesic thrice-punctured sphere, answering a question of Reid. Applying the proof techniques of the main result, we explicitly construct non-elementary maximal Fuchsian subgroups of infinite covolume within twist knot groups, and we also show that no twist knot complement with odd prime half twists is right-angled in the sense of Champanerkar, Kofman, and Purcell.

1 Introduction

The study of surfaces has been essential in studying the geometry and topology of the 33-manifolds that contain them. In this paper, we will mainly be concerned with complete properly immersed totally geodesic surfaces in hyperbolic 33-manifolds. There has been considerable work in understanding the existence of totally geodesic surfaces in hyperbolic 33-manifolds. Menasco and Reid [15, Corollary 4] proved that the complement of a hyperbolic tunnel number one knot cannot contain a closed embedded totally geodesic surface. Calegari [6, Corollary 4.6] showed that a fibered knot complement in a rational homology sphere whose trace field has odd prime degree cannot contain an immersed totally geodesic surface. By Adams and Schoenfeld [2, Theorem 4.1], two-bridge knot and link complements do not contain any embedded orientable totally geodesic surfaces. On the other hand, there are examples of hyperbolic 33-manifolds that do contain totally geodesic surfaces. Adams [1, Theorem 3.1] proved that any incompressible thrice-punctured sphere in a hyperbolic 33-manifold is totally geodesic. Adams and Schoenfeld [2, Example 3.1] exhibited examples of balanced pretzel knot containing a totally geodesic Seifert surface. For arithmetic hyperbolic 33-manifolds, it is known that if there exists at least one totally geodesic surface, then there are in fact infinitely many totally geodesic surfaces.

Most recently, Fisher, Lafont, Miller and Stover provided the first examples of hyperbolic 33-manifolds whose set of totally geodesic surfaces is nonempty and finite [9]. They proved finiteness of maximal totally geodesic submanifolds of dimension at least 2 for a large class of non-arithmetic hyperbolic nn-manifolds [9, Theorem 1.2]. In a subsequent work, Bader, Fisher, Miller and Stover showed that if a complete finite-volume hyperbolic nn-manifold of dimension at least 3 contains infinitely many maximal totally geodesic submanifolds then it must be arithmetic [3, Theorem 1.1]. A similar result was also obtained for the case of closed hyperbolic 33-manifolds by Margulis and Mohammadi [12, Theorem 1.1]. Bader, Fisher, Miller and Stover [3, Question 5.4] asked the following natural question, which is the motivation of our paper:

Question 1.1.

For each k1k\geq 1, is there a hyperbolic 33-manifold containing exactly kk totally geodesic surfaces?

We answer yes to the above question by exhibiting explicit examples of hyperbolic 33-manifolds with exactly kk totally geodesic surfaces:

Theorem 1.2.

Let kk be any positive integer. There exist infinitely many non-commensurable hyperbolic 33-manifolds with exactly kk totally geodesic surfaces.

These are the first examples of hyperbolic 33-manifolds for which the number of totally geodesic surfaces is positive, finite, and precisely known. The idea behind 1.2 is to find hyperbolic 33-manifolds with exactly one totally geodesic surface and then consider particular covers of these manifolds.

We specifically study twist knot complements. For each positive integer jj, let KjK_{j} be the twist knot with exactly jj half-twists as shown in Figure 1. Let Mj:=S3KjM_{j}:=S^{3}\smallsetminus K_{j}. These twist knots admit a complete finite-volume hyperbolic metric if and only if j2j\geq 2. In other words for j2j\geq 2, the knot group Γj:=π1(S3Kj)\Gamma_{j}:=\pi_{1}(S^{3}\smallsetminus K_{j}) admits a discrete faithful representation ρj\rho_{j} into PSL2()\operatorname{PSL}_{2}({\mathbb{C}}). By Adams [1, Theorem 3.1], the thrice-punctured sphere NN, seen as the dotted surface in Figure 1, is totally geodesic for j2j\geq 2. For simplification of notation, we will use NN to denote this surface, independent of jj.

Refer to caption
Figure 1: Twist knot KjK_{j} and the thrice-punctured sphere NN

By Reid [16, Theorem 3], among the twist knots with j3j\geq 3 (all but the figure-8 knot, j=2j=2 in our notation), there are infinitely many jj so that MjM_{j} contains no closed totally geodesic surfaces and has exactly one commensurability class of cusped totally geodesic surfaces. In fact, Reid conjectured that the thrice-punctured sphere is the unique totally geodesic surface inside MjM_{j} for j3j\geq 3. The following theorem of our paper confirms this conjecture for infinitely many values of jj.

Theorem 1.3.

Let MjM_{j} be the complement of a twist knot whose trace field has odd degree over {\mathbb{Q}} and contains no proper real subfield besides {\mathbb{Q}}. Then the thrice-punctured sphere NN is the unique totally geodesic surface in MjM_{j}.

Since MjM_{j} can be obtained by doing Dehn filling on the Whitehead link, the volume of MjM_{j} is uniformly bounded above by that of the Whitehead link. We note that the condition on the trace field in 1.3 is the same as the one considered by Reid in [16, Proposition 2]. As noted in [16, Lemma 5], it was originally due to Riley [18] that for infinitely many primes jj, the trace field of MjM_{j} satisfies the condition in 1.3. Later work of Hoste and Shanahan [10] implies the condition in 1.3 holds for the trace field of MjM_{j} for all prime jj. Using Magma [5] and SageMath [20], we also verify that the condition in 1.3 holds for the trace field of MjM_{j} for jj odd and j99j\leq 99, see [19] for the code. Therefore, we obtain the following:

Corollary 1.4.

Suppose jj is an odd prime or j>1j>1 is an odd number less than or equal to 9999. The thrice-punctured sphere NN is the unique totally geodesic surface in MjM_{j}.

The analysis in the proof of 1.3 has several interesting consequences. For all odd primes jj, we give explicit examples of non-elementary maximal Fuchsian subgroup of infinite covolume in Γj\Gamma_{j}. By Maclachlan and Reid [14, Theorem 4], every non-elementary maximal Fuchsian subgroup of an arithmetic Kleinian group has finite covolume. Therefore, 1.4 illustrates a stark difference between arithmetic and non-arithmetic 33-manifolds. Infinite covolume Fuchsian groups of this kind are known to exist, for example using results in [9], but we do not know of any explicit examples in the literature.

Another consequence of our analysis relates to right-angled knots. A knot in S3S^{3} whose complement admits a decomposition into ideal hyperbolic right-angled polyhedra is called a right-angled knot. Such a decomposition gives immersed totally geodesic surfaces coming from the faces of the polyhedra which meet at right angles. For j3j\geq 3, our analysis of the geometry of NN in MjM_{j} gives a strong restriction on the boundary slopes of cusped totally geodesic surfaces. This allows us to show that the angle at an intersection of cusped totally geodesic surfaces in MjM_{j} is never a right angle.

Corollary 1.5.

The twist knot KjK_{j} is not right-angled for jj odd prime.

1.5 confirms for infinitely many knots the following conjecture by Champanerkar, Kofman and Purcell [7, Conjecture 5.12]:

Conjecture 1.6 (Champanerkar, Kofman, and Purcell).

There does not exist a right-angled knot.

Previously known evidence for this conjecture comes from knot complements with no totally geodesic surface, such as knot complements satisfying a condition described by Calegari [6, Corollary 4.6]. Since the knot 8208_{20} satisfies [6, Corollary 4.6], it does not contain totally geodesic surface and is not right-angled. Examples of knot complements with totally geodesic surfaces supporting 1.6 have been found among knots with small crossing numbers. In particular, the conjecture was verified for knots with up to 11 crossings by Champanerkar, Kofman, and Purcell [7]. As far as we know, 1.5 provides the first infinitely family of knots that contain at least one totally geodesic surface and are not right-angled.

Finally, our techniques allow us to investigate boundary slopes of totally geodesic surfaces in the figure-8 knot. We prove that any rational number occurs as a boundary slope of a totally geodesic surface in the figure-8 knot in 4.1.

Our article is organized as follows. In Section 2, we collect some facts about twist knot complements. In Section 3, we prove 1.3 and 1.2. At the end of Section 3, we establish 1.5. In Section 4, we study boundary slopes of totally geodesic surfaces in the figure-8 knot. Finally in Section 5, we discuss some computer experiments and some open questions.

Acknowledgements

We would like to thank our advisor Matthew Stover for asking us 1.1, for multiple helpful suggestions in proving the main theorems, and for the careful reading of the first draft of this paper. We also thank David Futer and Rose Kaplan-Kelly for pointing out the application of 3.5 towards right-angled knots. Finally, we would like to thank the referee for their comments and their suggestions to use Sage, an open source, and the initial Sage codes.

2 Preliminaries

2.1 Twist knots complements

Let KjK_{j} be the twist knot with jj half-twists, Mj:=S3KjM_{j}:=S^{3}\smallsetminus K_{j}, and Γj\Gamma_{j} be its fundamental group. The group Γj\Gamma_{j} has a presentation Γj=a,bb=wjawj1\Gamma_{j}=\langle a,b\mid b=w_{j}aw_{j}^{-1}\rangle where wjw_{j} is given by a word of length 2j2j in a±1a^{\pm 1} and b±1b^{\pm 1}, namely:

wj={(a1b1ab)(j1)/2a1b1j1mod2(a1bab1)j/2j0mod2w_{j}=\begin{cases}(a^{-1}b^{-1}ab)^{(j-1)/2}a^{-1}b^{-1}&j\equiv 1\mod 2\\ (a^{-1}bab^{-1})^{j/2}&j\equiv 0\mod 2\end{cases} (1)

The two generators aa, bb correspond to the two meridians of the twist knot chosen to have the same orientation. The longitude of the twist knots that commute with aa can be computed diagrammatically and is given by:

j={aw¯jb2wjaj1mod2w¯jwjj0mod2\ell_{j}=\begin{cases}a\overline{w}_{j}b^{2}w_{j}a&j\equiv 1\mod 2\\ \overline{w}_{j}w_{j}&j\equiv 0\mod 2\\ \end{cases} (2)

where w¯j\overline{w}_{j} is wjw_{j} spelled backwards. Finally, we note that the above presentation is different from that in [10, Equation 2.1] by an isomorphism interchanging aa and bb.

For j2j\geq 2, MjM_{j} admits a complete hyperbolic structure of finite volume. The discrete faithful representation ρj:ΓjSL2()\rho_{j}:\Gamma_{j}\to\operatorname{SL}_{2}({\mathbb{C}}) sends all conjugates of the meridians of the knot to parabolic isometries; i.e. conjugate in SL2()\operatorname{SL}_{2}({\mathbb{C}}) to:

(1101)\begin{pmatrix}1&1\\ 0&1\end{pmatrix}
Theorem 2.1.

Suppose we normalize so that:

ρj(a)=(1101)andρj(b)=(10zj1)\rho_{j}(a)=\begin{pmatrix}1&1\\ 0&1\end{pmatrix}\quad\text{and}\quad\rho_{j}(b)=\begin{pmatrix}1&0\\ z_{j}&1\end{pmatrix}

Then ρj\rho_{j} defines the discrete faithful representation if and only if zjz_{j} satisfies a polynomial Λj(z)[z]\Lambda_{j}(z)\in{\mathbb{Z}}[z]. The polynomial Λj(z)\Lambda_{j}(z) is defined recursively by

Λj+2(z)=(z2+2)Λj(z)Λj2(z)\Lambda_{j+2}(z)=(z^{2}+2)\Lambda_{j}(z)-\Lambda_{j-2}(z) (3)

where Λ1(z)=Λ0(z)=1\Lambda_{-1}(z)=\Lambda_{0}(z)=1, Λ1(z)=z+1\Lambda_{1}(z)=z+1, and Λ2(z)=z2z+1\Lambda_{2}(z)=z^{2}-z+1.

The fact that zjz_{j} must be a root of an integral polynomial was first observed by Riley for the class of two-bridge knots [18, Theorem 2]. For the twist knots, a recursion for the polynomials Λj(z)\Lambda_{j}(z) was given by Hoste and Shanahan [10, Theorem 2]. We make the following observation which will be used in the proof of 1.3:

Remark 2.2.

The image of Γj\Gamma_{j} under ρj\rho_{j} is contained in SL2([zj])\operatorname{SL}_{2}({\mathbb{Z}}[z_{j}]). In particular, the entries of the matrices in ρj(Γj)\rho_{j}(\Gamma_{j}) can be written as {\mathbb{Z}}-linear combinations of elements in the zjz_{j}-power basis.

Since the presentation of Γj\Gamma_{j} that we are using is slightly different from that in [10, Equation 2.1], we give a proof of the above theorem for completeness.

  • Proof of 2.1.

    Consider the free group F2\operatorname{F}_{2} on two generators AA and BB along with the surjective homomorphism F2Γj\operatorname{F}_{2}\to\Gamma_{j} sending AA, BB to aa, bb respectively. Let WjW_{j} be the lift of wjw_{j} given by Equation 1. Consider the homomorphism P:F2SL2([z])P:\operatorname{F}_{2}\to\operatorname{SL}_{2}({\mathbb{Z}}[z]) defined by

    P(A)=(1101)andP(B)=(10z1)P(A)=\begin{pmatrix}1&1\\ 0&1\\ \end{pmatrix}\quad\text{and}\quad P(B)=\begin{pmatrix}1&0\\ z&1\end{pmatrix}

    We claim that

    P(Wj)=(Λj(z)μj(z)zμj(z)Λj1(z))P(W_{j})=\begin{pmatrix}\Lambda_{j}(z)&\mu_{j}(z)\\ z\mu_{j}(z)&\Lambda_{j-1}(z)\end{pmatrix} (4)

    for some μj(z)[z]\mu_{j}(z)\in{\mathbb{Z}}[z].

    For convenience of notation, we identify elements of F2\operatorname{F}_{2} with their image under PP. We consider two cases according to whether jj is even or odd. When jj is even, we have Wj=([A1,B])j/2W_{j}=([A^{-1},B])^{j/2}. Applying Cayley–Hamilton to [A1,B][A^{-1},B], we get

    ([A1,B])2tr([A1,B])[A1,B]+I=0([A^{-1},B])^{2}-\operatorname{tr}([A^{-1},B])[A^{-1},B]+I=0

    Note that tr([A1,B])=z2+2\displaystyle\operatorname{tr}([A^{-1},B])=z^{2}+2. Therefore, we have

    Wj+2=(z2+2)WjWj2W_{j+2}=(z^{2}+2)W_{j}-W_{j-2} (5)

    This gives us the recursion in Equation 3. Now we observe that the upper diagonal (resp. lower diagonal) entries of W2=[A1,B]W_{2}=[A^{-1},B] and of W0=IW_{0}=I give us the initial conditions for Λj(z)\Lambda_{j}(z) (resp. Λj1(z)\Lambda_{j-1}(z)). For the off-diagonal entries, we observe that they satisfy the same recursion as in Equation 3 and their initial conditions differ by a factor of zz. When jj is odd, we proceed similarly to compute P(Wj)P(W_{j}) where the initial conditions are given by W1=B1A1W_{-1}=B^{-1}A^{-1} and W1=A1B1W_{1}=A^{-1}B^{-1}.

    To finish the proof, we note that ρj\rho_{j} is obtained by factoring evjP\operatorname{ev}_{j}\circ P through the canonical projection F2Γj\operatorname{F}_{2}\to\Gamma_{j} where evj:SL2([z])SL2([zj])\operatorname{ev}_{j}:\operatorname{SL}_{2}({\mathbb{Z}}[z])\to\operatorname{SL}_{2}({\mathbb{Z}}[z_{j}]) is the homomorphism induced by the evaluation map [z][zj]{\mathbb{Z}}[z]\to{\mathbb{Z}}[z_{j}]. Using Equation 4, we see that the relation wjabwj=0w_{j}a-bw_{j}=0 is satisfied if and only if Λj(zj)=0\Lambda_{j}(z_{j})=0. ∎

Given a finitely-generated non-elementary subgroup Γ\Gamma of PSL2()\operatorname{PSL}_{2}({\mathbb{C}}), we can define the trace field of Γ\Gamma to be (trΓ):=(trγγΓ){\mathbb{Q}}(\operatorname{tr}\Gamma):={\mathbb{Q}}(\operatorname{tr}\gamma\mid\gamma\in\Gamma) and the invariant trace field of Γ\Gamma to be kΓ:=(trγ2γΓ)k\Gamma:={\mathbb{Q}}(\operatorname{tr}\gamma^{2}\mid\gamma\in\Gamma). If Γ\Gamma additionally has finite covolume, then (trΓ){\mathbb{Q}}(\operatorname{tr}\Gamma) and kΓk\Gamma are number fields [13, Theorem 3.1.2]. In general, the trace field and invariant trace field are different. The trace field is not an invariant of the commensurability class of Γ\Gamma, but the invariant trace field is. These fields may coincide, such as link complements in a /2{\mathbb{Z}}/2-homology sphere. [13, Corollary 4.2.2]. Finally, suppose that Γ\Gamma is generated by γ1\gamma_{1} and γ2\gamma_{2}; then (trΓ)=(trγ1,trγ2,trγ1γ2){\mathbb{Q}}(\operatorname{tr}\Gamma)={\mathbb{Q}}(\operatorname{tr}\gamma_{1},\operatorname{tr}\gamma_{2},\operatorname{tr}\gamma_{1}\gamma_{2}) [13, Equation 3.25].

It follows from the discussion above and 2.1 that the trace field of twist knot complements is (zj){\mathbb{Q}}(z_{j}). Hoste and Shanahan proved that Λj(z)\Lambda_{j}(z) is the monic minimal polynomial of zjz_{j} over {\mathbb{Z}} of degree jj [10, Theorem 1]. Therefore, [(zj):]=j[{\mathbb{Q}}(z_{j}):{\mathbb{Q}}]=j and so twist knots are pairwise non-commensurable.

Corollary 2.3.

Under the discrete faithful representation ρj\rho_{j}, we have

ρj(wj)=(0μjzjμj[(1)jzj2]μj)andρj(w¯j)=([(1)jzj2]μjμjzjμj0)\rho_{j}(w_{j})=\begin{pmatrix}0&\mu_{j}\\ z_{j}\mu_{j}&[(-1)^{j}z_{j}-2]\mu_{j}\\ \end{pmatrix}\quad\text{and}\quad\rho_{j}(\overline{w}_{j})=\begin{pmatrix}[(-1)^{j}z_{j}-2]\mu_{j}&\mu_{j}\\ z_{j}\mu_{j}&0\\ \end{pmatrix}

where μj=μj(zj)\mu_{j}=\mu_{j}(z_{j}). Furthermore, the image of j\ell_{j} is given by:

ρj(j)=(1τj01)\rho_{j}(\ell_{j})=\begin{pmatrix}-1&-\tau_{j}\\ 0&-1\end{pmatrix}

where τj=4μj2+2\tau_{j}=4\mu_{j}^{2}+2.

  • Proof.

    Let us consider W¯j\overline{W}_{j} to be the lift of w¯j\overline{w}_{j} in F2\operatorname{F}_{2} given by spelling WjW_{j} backwards. We first claim that:

    P(W¯j)=(Λj1(z)μj(z)zμj(z)Λj(z))P(\overline{W}_{j})=\begin{pmatrix}\Lambda_{j-1}(z)&\mu_{j}(z)\\ z\mu_{j}(z)&\Lambda_{j}(z)\end{pmatrix}

    Observe that W¯j\overline{W}_{j} satisfy the same recurrence in Equation 5, where W¯0\overline{W}_{0}, W¯1\overline{W}_{1} provide the base case.

    We next claim that:

    (1)j+1Λj(z)+Λj1(z)=[(1)jz2]μj(z)(-1)^{j+1}\Lambda_{j}(z)+\Lambda_{j-1}(z)=[(-1)^{j}z-2]\mu_{j}(z)

    We proceed by induction on jj. The base case is provided by W0=IW_{0}=I and W1=A1B1W_{1}=A^{-1}B^{-1}. Assume the statement for WkW_{k} for all k<j+2k<j+2. We have

    (1)j+3Λj+2(z)+Λj+1(z)\displaystyle(-1)^{j+3}\Lambda_{j+2}(z)+\Lambda_{j+1}(z) =(z2+2)[(1)j+3Λj(z)+Λj1(z)])[(1)j+3Λj2(z)+Λj3(z)]\displaystyle=(z^{2}+2)[(-1)^{j+3}\Lambda_{j}(z)+\Lambda_{j-1}(z)])-[(-1)^{j+3}\Lambda_{j-2}(z)+\Lambda_{j-3}(z)]
    =(z2+2)[(1)jz2]μj(z)[(1)jz2]μj2(z)\displaystyle=(z^{2}+2)[(-1)^{j}z-2]\mu_{j}(z)-[(-1)^{j}z-2]\mu_{j-2}(z)
    =[(1)j+2z2]μj+2(z)\displaystyle=[(-1)^{j+2}z-2]\mu_{j+2}(z)

    where the last equality follows from the recurrence of μj(z)\mu_{j}(z) given by Equation 5. Therefore, we have

    Λj1(zj)=[(1)jzj2]μj\Lambda_{j-1}(z_{j})=[(-1)^{j}z_{j}-2]\mu_{j}

    This identity gives us the image of wjw_{j} and w¯j\overline{w}_{j} under ρj\rho_{j}. We note that by the determinant of P(W¯j)P(\overline{W}_{j}),

    zjμj2=1z_{j}\mu_{j}^{2}=-1 (6)

    Applying ρj\rho_{j} to Equation 2 and using Equation 6, we get the image of j\ell_{j} under ρj\rho_{j} where

    τj=(4μj2+2)=4zj+2\tau_{j}=(4\mu_{j}^{2}+2)=-\frac{4}{z_{j}}+2 (7)

Remark 2.4.

Observe that since μj[zj]\mu_{j}\in{\mathbb{Z}}[z_{j}], Equation 6 implies that zjz_{j} and μj\mu_{j} are units in [zj]{\mathbb{Z}}[z_{j}]. We also note that Equations (6) and (7) are used extensively in the proof of 1.3.

We end this subsection by defining boundary slopes for cusped totally geodesic surfaces in MjM_{j}. Let us identify the universal cover 3{\mathbb{H}}^{3} of MjM_{j} with the upper-half space model and the ideal boundary of 3{\mathbb{H}}^{3} with {}{\mathbb{C}}\cup\{\infty\}. The action of Γj\Gamma_{j} on 3{\mathbb{H}}^{3} is given by ρj\rho_{j}. We say that P{}P\in{\mathbb{C}}\cup\{\infty\} is a cusp point of MjM_{j} if its stabilizer in Γj\Gamma_{j} is conjugate to a,j\langle a,\ell_{j}\rangle. In particular, the point \infty is the cusp point of MjM_{j} whose stabilizer in Γj\Gamma_{j} is the 2{\mathbb{Z}}^{2}-subgroup generated by aa and j\ell_{j} acting as translations by 11 and τj\tau_{j} respectively. At every cusp point in 3{\mathbb{H}}^{3}, we choose a sufficiently small horoball neighborhood such that the stabilizer is a conjugate of a,j\langle a,\ell_{j}\rangle and such that these horoball neighborhoods are invariant under the action of Γj\Gamma_{j}. We shall refer to the image of these horoball neighborhood as the cusp of MjM_{j}. Similarly given any hyperplane HH in 3{\mathbb{H}}^{3}, a point PP in the circle/line at infinity of HH is called a cusp point of HH if it is a fixed point of a parabolic element in StabΓj(H)\operatorname{Stab}_{\Gamma_{j}}(H).

Let Σ\Sigma be a properly immersed cusped totally geodesic surface in MjM_{j}. Since MjM_{j} has one cusp and Σ\Sigma is complete, all cusps of Σ\Sigma must be contained in the cusp of MM. Given any cusp cc of Σ\Sigma, we consider a lift Σ~\widetilde{\Sigma} of Σ\Sigma to 3{\mathbb{H}}^{3} by putting the cusp point corresponding to cc at \infty. The cusp cross-section of cc in Σ\Sigma is obtained by intersecting Σ\Sigma with the cusp cross-section of MM. This intersection lifts to a horocycle at \infty in Σ~\widetilde{\Sigma}. The stabilizer of this horocycle is an element of the form apjqa^{p}\ell_{j}^{q} in a,j\langle a,\ell_{j}\rangle. In this case, we say that p/qp/q is a boundary slope of Σ\Sigma.

More generally, suppose that ζ\zeta is a cusp point of a hyperplane HH in 3{\mathbb{H}}^{3} such that the parabolic element in the stabilizer of HH in Γj\Gamma_{j} fixing ζ\zeta is conjugate to apjqa^{p}\ell_{j}^{q} in Γj\Gamma_{j}. Then we say that p/qp/q is the boundary slope of HH at ζ\zeta.

2.2 The thrice-punctured sphere NN

For all j2j\geq 2, MjM_{j} contains an immersed thrice-punctured sphere NN as shown in Figure 1. By Adams [1, Theorem 3.1], NN is a totally geodesic surface. Moreover:

Proposition 2.5.

For all j2j\geq 2, the immersed surface NN is isotopic to a totally geodesic thrice-punctured sphere in MjM_{j} whose fundamental group is generated by x:=ax:=a and

yj:={awjaba1wj1a1if j1mod2awja1bawj1a1if j0mod2y_{j}:=\begin{cases}aw_{j}aba^{-1}w_{j}^{-1}a^{-1}&\text{if }j\equiv 1\mod 2\\ aw_{j}a^{-1}baw_{j}^{-1}a^{-1}&\text{if }j\equiv 0\mod 2\\ \end{cases}

The set of boundary slopes of NN is {1/0,1/0,2/1}\{1/0,1/0,-2/1\}.

  • Proof.

    From the diagram, we see that the fundamental group of NN is generated by bb and:

    {a1(bab1a1)(j1)/2bab1(aba1b1)(j1)/2aj1mod2(b1aba1)jb1ab(ab1a1b)jj0mod2\begin{cases}a^{-1}(bab^{-1}a^{-1})^{(j-1)/2}bab^{-1}(aba^{-1}b^{-1})^{(j-1)/2}a&j\equiv 1\mod 2\\ (b^{-1}aba^{-1})^{j}b^{-1}ab(ab^{-1}a^{-1}b)^{j}&j\equiv 0\mod 2\end{cases}

    Thus, we see that π1(N)\pi_{1}(N) is conjugate to the subgroup Δj\Delta_{j} generated by x:=ax:=a and yjy_{j} as stated. Under the discrete faithful representation, we have

    ρj(Δj)=(1101),(1143)\rho_{j}(\Delta_{j})=\left\langle\begin{pmatrix}1&1\\ 0&1\end{pmatrix},\begin{pmatrix}-1&1\\ -4&3\\ \end{pmatrix}\right\rangle

    which is conjugate to the principal congruence subgroup of level 2 of PSL2()\operatorname{PSL}_{2}({\mathbb{Z}}).

    Note that since both xx and yjy_{j} are conjugate to aa, the surface NN has two slopes 1/0. The parabolic element corresponding to the remaining cusp is given by

    x1yj1=(1041)=b2wjjwj1x^{-1}y_{j}^{-1}=\begin{pmatrix}-1&0\\ 4&-1\end{pmatrix}=b^{-2}w_{j}\ell_{j}w_{j}^{-1}

    Since b2wjjb^{-2}w_{j}\ell_{j} and a2ja^{-2}\ell_{j} are conjugate to each other in Γj\Gamma_{j}, the slope of the remaining cusp is 2/1-2/1. ∎

Remark 2.6.

Let 2{\mathbb{H}}^{2}_{\mathbb{R}} be the hyperplane in 3{\mathbb{H}}^{3} such that the line at infinity is the real line {\mathbb{R}}. The group ρj(Δj)\rho_{j}(\Delta_{j}) stabilizes 2{\mathbb{H}}^{2}_{\mathbb{R}} in 3{\mathbb{H}}^{3}. In fact, it is conjugate to the principal congruence subgroup of level 2 of PSL2()\operatorname{PSL}_{2}({\mathbb{Z}}) by the matrix

(1/2002)\begin{pmatrix}1/\sqrt{2}&0\\ 0&\sqrt{2}\end{pmatrix}

Thus, elements of ρj(Δj)\rho_{j}(\Delta_{j}) are precisely matrices in PSL2()\operatorname{PSL}_{2}({\mathbb{Z}}) such that the lower diagonal entry is congruent to 0mod40\mod 4. This group is also known as the Hecke congruence subgroup of level 44.

We will also need the following fact about the action of ρj(Δj)\rho_{j}(\Delta_{j}) on 2{\mathbb{H}}_{\mathbb{R}}^{2}.

Lemma 2.7.

Consider the action of ρj(Δj)\rho_{j}(\Delta_{j}) on 2{\mathbb{H}}^{2}_{\mathbb{R}}. The cusp points of this action is {}{\mathbb{Q}}\cup\{\infty\}, which is partitioned into 3 distinct orbits, namely

[0]\displaystyle[0] ={u/vv1mod2},\displaystyle=\{u/v\in{\mathbb{Q}}\mid v\equiv 1\mod 2\},
[12]\displaystyle[\tfrac{1}{2}] ={u/vv2mod4},\displaystyle=\{u/v\in{\mathbb{Q}}\mid v\equiv 2\mod 4\},
[]\displaystyle[\infty] ={u/vv0mod4}{}\displaystyle=\{u/v\in{\mathbb{Q}}\mid v\equiv 0\mod 4\}\cup\{\infty\}

where u,vu,v\in{\mathbb{Z}} are relatively prime.

  • Proof.

    Let CC denote the set of cusp points in 2{\mathbb{H}}^{2}_{\mathbb{R}} with respect to the action of ρj(Δj)\rho_{j}(\Delta_{j}). Since aρj(Δj)a\in\rho_{j}(\Delta_{j}), we see that C\infty\in C. The fundamental domain for the action is an ideal quadrilateral with vertices at ,0,12\infty,0,\frac{1}{2} and 11. There are three distinct orbits corresponding to 0, 12\frac{1}{2} and \infty. To see the description of these orbits, we first note that the integer matrices

    (vu4uv),(u2vvv2uu), and(uvvu)\begin{pmatrix}v^{\prime}&u\\ -4u^{\prime}&v\end{pmatrix},\quad\begin{pmatrix}u-2v^{\prime}&v^{\prime}\\ v-2u^{\prime}&u^{\prime}\end{pmatrix},\text{ and}\begin{pmatrix}u&v^{\prime}\\ v&u^{\prime}\end{pmatrix}

    map 0, 12\frac{1}{2}, and \infty to u/vu/v when vv is odd, 2mod42\mod 4, or 0mod40\mod 4, respectively. The determinants of these matrices can be chosen to be 1 because 4u4u is relatively prime to vv when both vv is odd and uu is relatively prime to vv. By Remark 2.6, to be in ρj(Δj)\rho_{j}(\Delta_{j}) (up to sign), it is sufficient to show that the lower diagonal entry of each matrix is congruent to 0mod40\mod 4. This is immediate for the first and the third matrices. For the second matrix, note that uu^{\prime} has to be an odd integer since vv is an even integer relatively prime to uu^{\prime}. Since vv is congruent to 2mod42\mod 4, the difference v2uv-2u^{\prime} is congruent to 0mod40\mod 4. ∎

We conclude this section by describing certain lifts of NN in 3{\mathbb{H}}^{3} explicitly. The parametrization of these lifts play an important role in the proof of 1.3. We call a totally geodesic hyperplane in 3{\mathbb{H}}^{3} vertical if it contains \infty in its circle/line at infinity. For convenience, we let vj:=wj1a(1)jwj1a1v_{j}:=w_{j}^{-1}a^{(-1)^{j}}w_{j}^{-1}a^{-1} and note that vjyjvj1=av_{j}y_{j}v_{j}^{-1}=a. Observe that up to the action of StabΓj()\operatorname{Stab}_{\Gamma_{j}}(\infty), the surface NN has three distinct vertical lifts in 3{\mathbb{H}}^{3}.

A direct computation shows that

wj1(0)=andvj(12)=w_{j}^{-1}(0)=\infty\quad\text{and}\quad v_{j}(\tfrac{1}{2})=\infty

Therefore, all vertical lifts of NN are orbits of 2{\mathbb{H}}^{2}_{\mathbb{R}}, wj1(2)w^{-1}_{j}({\mathbb{H}}^{2}_{\mathbb{R}}), and vj(2)v_{j}({\mathbb{H}}^{2}_{\mathbb{R}}) under the action of StabΓj()=a,j\operatorname{Stab}_{\Gamma_{j}}(\infty)=\langle a,\ell_{j}\rangle. Since the slope at 0 (resp. 12\frac{1}{2}) on 2{\mathbb{H}}^{2}_{\mathbb{R}} is 2/1-2/1 (resp. 1/01/0), the vertical hyperplane wj1(2)w^{-1}_{j}({\mathbb{H}}^{2}_{\mathbb{R}}) (resp. vj(2)v_{j}({\mathbb{H}}^{2}_{\mathbb{R}})) has slope 2/1-2/1 (resp. 1/01/0) at infinity,. Furthermore, we have

wj1()=(1)j+1+2zj=12τj+1+(1)j+1w_{j}^{-1}(\infty)=(-1)^{j+1}+\frac{2}{z_{j}}=-\frac{1}{2}\tau_{j}+1+(-1)^{j+1}

and

vj()=2μj2+12+(1)j+1=12τj+32+(1)j+1v_{j}(\infty)=-2\mu_{j}^{2}+\frac{1}{2}+(-1)^{j+1}=-\frac{1}{2}\tau_{j}+\frac{3}{2}+(-1)^{j+1}

Thus when jj is odd, the line at infinity of wj1(2)w_{j}^{-1}({\mathbb{H}}^{2}_{\mathbb{R}}) goes through the point 11 and has slope 2/1-2/1. Similarly, when jj is odd, the line at infinity of vj(2)v_{j}({\mathbb{H}}^{2}_{\mathbb{R}}) goes through the point τj/2+5/2-\tau_{j}/2+5/2 and has slope 1/01/0. Figure 2 depicts the lines at infinity of the vertical lifts of NN.

Refer to caption
Figure 2: Vertical lifts of NN and a fundamental domain of the cusp when jj is odd

Note that in Figure 2, the element aa acts by translating upward and j\ell_{j} acts by translating to the right. In other words, the real axis in {\mathbb{C}} is put in the vertical direction. Under this convention, the boundary slopes at infinity of a vertical hyperplane is precisely the slope of its line at infinity.

3 Main Theorems

For this section, we assume j3j\geq 3 and drop jj as the subscript notation.

3.1 Geometric and arithmetic constraints on totally geodesic surfaces

The following proposition of Reid [16, Proposition 2] gives an arithmetic constraint on the existence of totally geodesic surfaces inside hyperbolic 33-manifolds.

Proposition 3.1.

Let Γ\Gamma be a non-cocompact Kleinian group of finite covolume and satisfying the following two conditions:

  • (trΓ){\mathbb{Q}}(\operatorname{tr}\Gamma) is of odd degree over {\mathbb{Q}} and contains no proper real subfield other than {\mathbb{Q}}.

  • Γ\Gamma has integral traces.

Then Γ\Gamma contains no cocompact Fuchsian groups and at most one commensurability class (up to conjugacy in PSL2()\operatorname{PSL}_{2}({\mathbb{C}})) of non-cocompact Fuchsian subgroup of finite covolume.

This proposition gives a convenient way to rule out closed totally geodesic surfaces, as remarked in [16, Lemma 5].

Corollary 3.2.

Let MM have an odd prime number of half-twists. Then MM does not contain closed immersed totally geodesic surfaces.

  • Proof.

    Firstly, since [(trΓ):]=j[{\mathbb{Q}}(\operatorname{tr}\Gamma):{\mathbb{Q}}]=j is an odd prime [10, Theorem 1], the field (trΓ){\mathbb{Q}}(\operatorname{tr}\Gamma) has odd degree over \mathbb{Q} and contains no proper subfield except {\mathbb{Q}}. Secondly, the fact that Im(ρ)SL2([z])\text{Im}(\rho)\leq\operatorname{SL}_{2}({\mathbb{Z}}[z]) implies that Γ\Gamma has integral traces, see Remark 2.2. ∎

To prove the main theorems, we need to rule out cusped totally geodesic surfaces that are not freely homotopic to NN. To this end, we examine general behavior of the intersection of two totally geodesic surfaces, and then we show that any cusped totally geodesic surface in MM not freely homotopic to NN must intersect NN because this thrice-punctured sphere has two distinct boundary slopes.

We have the following lemma of Fisher, Lafont, Miller and Stover [9, Lemma 3.1] which describes the intersection of totally geodesic hypersurface and immersed totally geodesic submanifolds in finite-volume hyperbolic nn-manifold. We restate their lemma for dimension 33. A geodesic in a cusped finite-volume hyperbolic 33-manifold 3/Γ{\mathbb{H}}^{3}/\Gamma is called a cusp-to-cusp geodesic if it is the image of a geodesic in 3{\mathbb{H}}^{3} connecting two cusp points under the action of Γ\Gamma on 3{\mathbb{H}}^{3}.

Lemma 3.3.

Let MM be a complete finite volume hyperbolic 33-manifold with at least 11 cusp. Suppose that Σ1\Sigma_{1} and Σ2\Sigma_{2} are two distinct properly immersed totally geodesic surfaces in MM such that Σ1Σ2\Sigma_{1}\cap\Sigma_{2} is non-empty. Then Σ1Σ2\Sigma_{1}\cap\Sigma_{2} is the union of closed geodesics and cusp-to-cusp geodesics.

Remark 3.4.

We observe that the lemma also holds for self-intersection of a totally geodesic surface. The key point is the following. For every 2-plane in the tangent space of a point in 3{\mathbb{H}}^{3}, there exists a unique hyperplane tangent to the 2-plane. It follows that the self-intersection of totally geodesic surface is transverse. Therefore, the components of the self-intersection are unions of properly immersed complete one-dimensional submanifolds.

3.2 Uniqueness of the thrice-punctured sphere

Before we prove 1.3 in this section, we make the following observations as well as restate that we drop the index jj from all notations. Let ΣM\Sigma\subset M be a cusped totally geodesic surface that is not freely homotopic to NN. Let Σ~\widetilde{\Sigma} be a hyperplane lift of Σ\Sigma to 3{\mathbb{H}}^{3}. Since Σ\Sigma is a cusped surface, the line/circle at infinity of Σ~\widetilde{\Sigma} contains a cusp point of MM. If Σ~\widetilde{\Sigma} is not a vertical hyperplane, we replace Σ~\widetilde{\Sigma} by a Γ\Gamma-translate that moves a cusp point of Σ~\widetilde{\Sigma} to infinity. Without loss of generality, we assume that Σ~\widetilde{\Sigma} is a vertical hyperplane.

Since Σ\Sigma is distinct from NN, we can see from Figure 2 that the line at infinity of Σ~\widetilde{\Sigma} must intersect the line at infinity of either 2{\mathbb{H}}^{2}_{\mathbb{R}} or w1(2)w^{-1}({\mathbb{H}}^{2}_{\mathbb{R}}) transversely. Therefore, Σ~\widetilde{\Sigma} intersects some vertical lift of NN along a geodesic (P,)(P,\infty) for some PP\in{\mathbb{C}}. Since \infty is a cusp point of MM, 3.3 implies that (P,)(P,\infty) projects onto a cusp-to-cusp geodesic in MM. In particular, PP is a cusp point of MM.

Using the geodesic (P,)(P,\infty), we construct parabolic elements in StabΓ(Σ~)\operatorname{Stab}_{\Gamma}(\widetilde{\Sigma}) fixing each end point and analyze the trace of the product of these elements. Since PP is a cusp point of MM, there exists γΓ\gamma\in\Gamma such that γ()=P\gamma(\infty)=P. Since PP and \infty are cusp points of Σ~\widetilde{\Sigma}, StabΓ(Σ~)\operatorname{Stab}_{\Gamma}(\widetilde{\Sigma}) contains the nontrivial elements apqa^{p}\ell^{q} and γamnγ1\gamma a^{m}\ell^{n}\gamma^{-1} fixing \infty and PP, respectively, where gcd(p,q)=gcd(m,n)=1\gcd(p,q)=\gcd(m,n)=1. In other words, p/qp/q (resp. m/nm/n) is the boundary slope of Σ~\widetilde{\Sigma} at \infty (resp. PP).

The following is a critical preliminary calculation in our analysis. We note that tr(ρ(π1Σ))\operatorname{tr}(\rho(\pi_{1}\Sigma)) must contain only real algebraic integers in (z){\mathbb{Q}}(z). The condition that (z){\mathbb{Q}}(z) does not contain any proper real subfield other than {\mathbb{Q}} implies that tr(ρ(π1Σ))\operatorname{tr}(\rho(\pi_{1}\Sigma))\subseteq{\mathbb{Z}}. Suppose that

ρ(γ)=(αβδη)\rho(\gamma)=\begin{pmatrix}\alpha&\beta\\ \delta&\eta\end{pmatrix}

Then, dropping the ρ\rho for convenience of notation, we must have

tr(γamnγ1apq)=(1)n+q+1[2+(m+nτ)(p+qτ)δ2]\operatorname{tr}(\gamma a^{m}\ell^{n}\gamma^{-1}a^{p}\ell^{q})=(-1)^{n+q+1}[-2+(m+n\tau)(p+q\tau)\delta^{2}]\in{\mathbb{Z}}

which is true if and only if

δ2(nqτ2+(mq+np)τ+mp)\delta^{2}(nq\tau^{2}+(mq+np)\tau+mp)\in{\mathbb{Z}} (8)

We shall refer to this as the trace condition, which gives us the following:

Lemma 3.5.

Let MM be the complement of a twist knot whose trace field is of odd degree over \mathbb{Q} and contains no proper real subfield other than \mathbb{Q}. Let NN be the totally geodesic thrice-punctured sphere as in Figure 1. Suppose that Σ\Sigma is an immersed cusped totally geodesic surface that is not freely homotopic NN. Then the set of boundary slopes of Σ\Sigma is precisely {1/0,2}\{1/0,-2\}, up to multiplicity.

  • Proof.

    It follows from the previous discussion that Σ\Sigma has a vertical lift Σ~\widetilde{\Sigma} to 3{\mathbb{H}}^{3}. The lift Σ~\widetilde{\Sigma} must intersect either 2{\mathbb{H}}^{2}_{\mathbb{R}} or w1(2)w^{-1}({\mathbb{H}}^{2}_{\mathbb{R}}), whose respective slopes are 1/01/0 and 2/1-2/1, along some vertical geodesic θ\theta. One end point of θ\theta is \infty, so by 3.3, θ\theta covers a cusp-to-cusp geodesic. This implies that the other end point of θ\theta must also be a cusp point in 2{\mathbb{H}}^{2}_{\mathbb{R}} or wj1(2)w^{-1}_{j}({\mathbb{H}}^{2}_{\mathbb{R}}). Therefore, either θ=(s/t,)\theta=(s/t,\infty) for s/ts/t\in{\mathbb{Q}} or θ=(w1(s/t),)\theta=(w^{-1}(s/t),\infty) for s/t{}s/t\in{\mathbb{Q}}\cup\{\infty\}.

    Suppose that the boundary slopes on Σ~\widetilde{\Sigma} are p/qp/q at \infty and m/nm/n at the other end point of θ\theta. As boundary slopes, neither (m,n)(m,n) nor (p,q)(p,q) is (0,0)(0,0). We have nontrivial elements apqa^{p}\ell^{q} and γamnγ1\gamma a^{m}\ell^{n}\gamma^{-1} in StabΓ(Σ~)\operatorname{Stab}_{\Gamma}(\widetilde{\Sigma}) where γΓ\gamma\in\Gamma has the property that γ1(s/t)=\gamma^{-1}(s/t)=\infty or γ1w1(s/t)=\gamma^{-1}w^{-1}(s/t)=\infty. We consider these two cases separately, and use the general form

    γ=(αβδη)\gamma=\begin{pmatrix}\alpha&\beta\\ \delta&\eta\end{pmatrix}

    Note that the end points of θ\theta are distinct since θ\theta is a geodesic in 3{\mathbb{H}}^{3}. Consequently, γ\gamma cannot fix \infty, and so δ0\delta\neq 0.

    Case 1:

    Suppose that Σ~\widetilde{\Sigma} intersects 2{\mathbb{H}}_{\mathbb{R}}^{2} nontrivially along a geodesic θ=(s/t,)\theta=(s/t,\infty) for some s/ts/t\in{\mathbb{Q}}, as seen in Figure 3. In view of 2.7, we may choose γ\gamma to be either

    (ts4st)w,(s2ttt2ss)v1, or (stts)\begin{pmatrix}t^{\prime}&s\\ -4s^{\prime}&t\end{pmatrix}w,\quad\begin{pmatrix}s-2t^{\prime}&t^{\prime}\\ t-2s^{\prime}&s^{\prime}\end{pmatrix}v^{-1},\text{ or }\begin{pmatrix}s&t^{\prime}\\ t&s^{\prime}\end{pmatrix}

    according to whether s/ts/t belongs to [0][0], [12][\frac{1}{2}] or [][\infty], respectively. It suffices to compute δ\delta in these three circumstances by using the trace condition Equation 8.

    Since v1()=12v^{-1}(\infty)=\frac{1}{2}, we see that v1v^{-1} is of the form

    ±(12)\pm\begin{pmatrix}1&*\\ 2&*\end{pmatrix}

    This implies that δ=t{0}\delta=t\in{\mathbb{Z}}\smallsetminus\{0\} in the last two situations. We show that this contradicts the assumption that (p,q)(0,0)(m,n)(p,q)\neq(0,0)\neq(m,n). It follows from (7) that the trace field is the same as the cusp field. That is, (z)=(τ){\mathbb{Q}}(z)={\mathbb{Q}}(\tau). Then

    [(τ):]=[(z):]3,[{\mathbb{Q}}(\tau):{\mathbb{Q}}]=[{\mathbb{Q}}(z):{\mathbb{Q}}]\geq 3,

    so the minimal polynomial of τ\tau over \mathbb{Q} has degree 3\geq 3. In particular, τ\tau cannot satisfy the quadratic polynomial over {\mathbb{Z}} given by the trace condition. Therefore, we have nq=0nq=0 and mq+np=0mq+np=0 since δ20\delta^{2}\neq 0. Since Σ~\widetilde{\Sigma} intersects 2{\mathbb{H}}^{2}_{\mathbb{R}} nontrivially, the boundary slope at \infty on Σ~\widetilde{\Sigma} is not meridional; in other words, q0q\neq 0. We thus have n=0n=0, which implies that mq=0mq=0, and so m=0m=0. This contradicts the assumption that (m,n)(0,0)(m,n)\neq(0,0).

    In the first situation, we get δ=tμz\delta=t\mu z and so δ2=t2z\delta^{2}=-t^{2}z. The trace condition becomes

    δ2(nqτ2+(mq+np)τ+mp)\displaystyle\delta^{2}(nq\tau^{2}+(mq+np)\tau+mp) \displaystyle\in{\mathbb{Z}}
    z[nq(4z1+2)2+(mq+np)(4z1+2)+mp]\displaystyle z[nq(-4z^{-1}+2)^{2}+(mq+np)(-4z^{-1}+2)+mp] \displaystyle\in{\mathbb{Q}}
    z[16nqz2(16nq+4mq+4np)z1+4nq+2mq+2np+mp]\displaystyle z[16nqz^{-2}-(16nq+4mq+4np)z^{-1}+4nq+2mq+2np+mp] \displaystyle\in{\mathbb{Q}}
    16nqz1+(2n+m)(2q+p)z\displaystyle 16nqz^{-1}+(2n+m)(2q+p)z \displaystyle\in{\mathbb{Q}}
    16nq+(2n+m)(2q+p)z2\displaystyle 16nq+(2n+m)(2q+p)z^{2} z\displaystyle\in{\mathbb{Q}}z

    Since the minimal polynomial of zz has degree at least 33, satisfying the above is equivalent to satisfying

    16nq=(2n+m)(2q+p)=016nq=(2n+m)(2q+p)=0

    Again, since Σ~\widetilde{\Sigma} intersects 2{\mathbb{H}}^{2}_{\mathbb{R}} nontrivially, we have q0q\neq 0, so m/n=1/0m/n=1/0 and p/q=2/1p/q=-2/1.

    Refer to caption
    Figure 3: Vertical lifts of NN and Σ\Sigma when jj odd
    Case 2:

    Suppose that Σ~\widetilde{\Sigma} is parallel to 2{\mathbb{H}}^{2}_{\mathbb{R}}, or equivalently θ=(w1(s/t),)\theta=(w^{-1}(s/t),\infty) for some s/ts/t\in{\mathbb{Q}}. We may assume that p/q=1/0p/q=1/0. In view of 2.7, we may choose γ=(αβδη)\gamma=\begin{pmatrix}\alpha&\beta\\ \delta&\eta\end{pmatrix} to be either

    w1(ts4st)w, or w1(s2ttt2ss)v1, or w1(stts)w^{-1}\begin{pmatrix}-t^{\prime}&s\\ -4s^{\prime}&t\end{pmatrix}w,\text{ or }w^{-1}\begin{pmatrix}s-2t^{\prime}&t^{\prime}\\ t-2s^{\prime}&s^{\prime}\end{pmatrix}v^{-1},\text{ or }w^{-1}\begin{pmatrix}s&t^{\prime}\\ t&s^{\prime}\end{pmatrix}

    When s/t[0]s/t\in[0], we get δ=sz\delta=sz, so

    δ2(m+nτ)\displaystyle\delta^{2}(m+n\tau) \displaystyle\in{\mathbb{Z}}
    s2z2(m+n(4z1+2))\displaystyle s^{2}z^{2}(m+n(-4z^{-1}+2)) \displaystyle\in{\mathbb{Z}}

    This is satisfied if and only if s2n=s2(2n+m)=0s^{2}n=s^{2}(2n+m)=0 because the minimal polynomial of zz is at least cubic over {\mathbb{Q}}. If s=0s=0, then w1(0)=w^{-1}(0)=\infty since endpoints of θ\theta must be distinct, therefore s0s\neq 0. When s0s\neq 0, we must have m=n=0m=n=0, which is a contradiction.

    When s/t[12]s/t\in[\frac{1}{2}] or [][\infty], we see that δ=sμz\delta=-s\mu z and use an equivalent calculation to Case 1 to conclude that the boundary slopes are p/q=1/0p/q=1/0 and m/n=2/1m/n=-2/1.

We now prove the uniqueness of NN.

  • Proof of 1.3.

    Throughout the proof, we assume that the trace field of Γ\Gamma has odd degree and has no proper real subfield except for {\mathbb{Q}}. In particular, Γ=Γj\Gamma=\Gamma_{j} for some positive odd integer jj. Since Γ\Gamma has integral traces by 3.2 and its trace field satisfies the condition in 3.1, any totally geodesic surface in MM must have at least one cusp.

    3.5 has the following consequence. If Σ\Sigma is a cusped totally geodesic surface in MM that is not NN, then there exists a vertical lift Σ~\widetilde{\Sigma} in 3{\mathbb{H}}^{3} with slope 2/1-2/1 at \infty. This lift intersects 2{\mathbb{H}}^{2}_{\mathbb{R}} along a cusp-to-cusp geodesic (s/t,)(s/t,\infty). A careful look at the first case in the proof of 3.5 shows that s/ts/t must belong to the orbit of [0][0] under the action of Δ\Delta on 2{\mathbb{H}}^{2}_{\mathbb{R}}; in particular, tt is an odd integer. Using the action of aa, we may assume that 0<s/t<10<s/t<1, since Σ\Sigma is distinct from NN. Let Hs/tH_{s/t} be the vertical lift of Σ\Sigma to 3{\mathbb{H}}^{3} of slope 2/1-2/1 at \infty that intersects the hyperplane 2{\mathbb{H}}^{2}_{\mathbb{R}} along the geodesic (s/t,)(s/t,\infty).

    In order to prove 1.3, it suffices to show that each vertical hyperplane Hs/tH_{s/t} for 0<s/t<10<s/t<1 with tt odd covers an infinite area totally geodesic surface, which implies that Hs/t/StabΓ(Hs/t)H_{s/t}/\operatorname{Stab}_{\Gamma}(H_{s/t}) must be dense in MM. With this density in mind, we construct a non-closed non-cusp-to-cusp geodesic in MM as a self-intersection of Hs/t/StabΓ(Hs/t)H_{s/t}/\operatorname{Stab}_{\Gamma}(H_{s/t}).

    Fix a rational number 0<s/t<10<s/t<1 where tt odd and ss, tt relatively prime. Suppose that Hs/tH_{s/t} covers a complete properly immersed finite area totally geodesic surface. We move Hs/tH_{s/t} by an element of Γ\Gamma that takes s/ts/t to \infty. To write down such an element, we first map s/ts/t to 0 by an element of Δ\Delta, then map 0 to \infty by w1w^{-1}. Consider the following element

    ζ=w1(ts4st)\zeta=w^{-1}\begin{pmatrix}t&-s\\ 4s^{\prime}&-t^{\prime}\end{pmatrix}

    taking s/ts/t to \infty. Note that the latter matrix in the product belongs to Δ\Delta as discussed in Remark 2.6. Therefore, ζΓ\zeta\in\Gamma. Since ζ(s/t)=\zeta(s/t)=\infty, the hyperplane ζ(Hs/t)\zeta(H_{s/t}) must be vertical.

    The analysis in the first case of 3.5 shows that the slope of Hs/tH_{s/t} at s/ts/t is 1/01/0. Consequently, the slope of ζ(Hs/t)\zeta(H_{s/t}) is at \infty is 1/01/0. Since Hs/tH_{s/t} has slope 2/1-2/1, the vertical hyperplane ζ(Hs/t)\zeta(H_{s/t}) intersects Hs/tH_{s/t} along a geodesic θ=(θ,)\theta=(\theta_{-},\infty) for some θ\theta_{-}\in{\mathbb{C}}. We have

    ζ()=1+(12+st)4z\zeta(\infty)=1+\left(\frac{1}{2}+\frac{s^{\prime}}{t}\right)\frac{4}{z}

    Since the slope of Hs/tH_{s/t} at \infty is 2/1-2/1, the line at infinity of Hs/tH_{s/t} is parametrized by

    st+r(τ2)=str4z\frac{s}{t}+r(\tau-2)=\frac{s}{t}-r\frac{4}{z}

    for some rr\in{\mathbb{R}} while that of ζ(Hs/t)\zeta(H_{s/t}) is parametrized by

    ζ()+r\zeta(\infty)+r^{\prime}

    for some rr^{\prime}\in{\mathbb{R}}. It follows that

    θ=st+(12+st)4z=sz+4s+2ttz.\theta_{-}=\frac{s}{t}+\left(\frac{1}{2}+\frac{s^{\prime}}{t}\right)\frac{4}{z}=\frac{sz+4s^{\prime}+2t}{tz}. (9)

    Since one end point of θ\theta is at \infty, a cusp point of Γ\Gamma, Remark 3.4 implies that θ\theta covers a cusp-to-cusp geodesic. In particular, θ\theta_{-} is a cusp point of Γ\Gamma. Thus, there exists

    γ=(αβδη)ΓSL2([z])\gamma=\begin{pmatrix}\alpha&\beta\\ \delta&\eta\end{pmatrix}\in\Gamma\leq\operatorname{SL}_{2}({\mathbb{Z}}[z])

    such that γ(θ)=\gamma(\theta_{-})=\infty, see Remark 2.2. We must have

    η=δθ\eta=-\delta\theta_{-} (10)

    By 3.5, the slope at θ\theta_{-} of Hs/tH_{s/t} is either 1/01/0 or 2/1-2/1. We consider these two cases separately.

    Case 1:

    Suppose that the slope at θ\theta_{-} is 1/01/0. Consequently, StabΓ(Hs/t)\operatorname{Stab}_{\Gamma}(H_{s/t}) contains γ1aγ\gamma^{-1}a\gamma. The trace condition Equation 8 implies that

    tr(γ1aγa21)=δ2(2τ)=4μ2δ2.\operatorname{tr}(\gamma^{-1}a\gamma a^{2}\ell^{-1})=\delta^{2}(2-\tau)=-4\mu^{2}\delta^{2}\in{\mathbb{Z}}.

    Let us write δ2=zk/4\delta^{2}=z\,k/4 for some kk\in{\mathbb{Z}}. Note that δ0\delta\neq 0 as Γ\Gamma contains no trace 0 element. Since δ[z]\delta\in{\mathbb{Z}}[z], we have k/4k/4\in{\mathbb{Z}}. Using Equation 6, we have δ=±μ1k/4\delta=\pm\mu^{-1}\sqrt{-k/4} or μδ=±k/4\mu\delta=\pm\sqrt{-k/4}. Since [(z):][{\mathbb{Q}}(z):{\mathbb{Q}}] is odd by [10, Theorem 1], (z){\mathbb{Q}}(z) does not contain the square root of non-square integer. It follows that k/4\sqrt{-k/4}\in{\mathbb{Z}}. In other words, δ=±mμ1\delta=\pm m\mu^{-1} for some mm\in{\mathbb{Z}}. Substituting this and Equation 9 into Equation 10, we have

    ±ημ1=m(z)sz+4s+2ttz=mstz+m(4s+2t)t[z].\pm\eta\mu^{-1}=-m(-z)\frac{sz+4s^{\prime}+2t}{tz}=\frac{ms}{t}z+\frac{m(4s^{\prime}+2t)}{t}\in{\mathbb{Z}}[z].

    The integer tt must divide mm because ss and tt are relatively prime. Since μ\mu is a unit in [z]{\mathbb{Z}}[z], we have the following containment of ideals η[z]mt[z]\eta{\mathbb{Z}}[z]\subseteq\frac{m}{t}{\mathbb{Z}}[z] and δ[z]m[z]mt[z]\delta{\mathbb{Z}}[z]\subseteq m{\mathbb{Z}}[z]\subseteq\frac{m}{t}{\mathbb{Z}}[z]. In particular, the ideal (δ,η)[z](\delta,\eta){\mathbb{Z}}[z] generated by δ\delta and η\eta in [z]{\mathbb{Z}}[z] is contained in mt[z]\frac{m}{t}{\mathbb{Z}}[z]. Since αηβδ=1\alpha\eta-\beta\delta=1, (δ,η)[z]=[z](\delta,\eta){\mathbb{Z}}[z]={\mathbb{Z}}[z]. This fact implies that m=±tm=\pm t. We obtain the final expressions for η\eta and δ\delta.

    η=±(sz+4s+2t)μandδ=±tμ1\eta=\pm(sz+4s^{\prime}+2t)\mu\quad\text{and}\quad\delta=\pm t\mu^{-1} (11)

    The slope of Hs/tH_{s/t} at θ\theta_{-} is 1/01/0 and γ(θ)=\gamma(\theta_{-})=\infty, so γ(Hs/t)\gamma(H_{s/t}) is a vertical hyperplane with 1/01/0 slope at \infty. Thus, this hyperplane intersects w1(2)w^{-1}({\mathbb{H}}^{2}_{\mathbb{R}}) along a vertical geodesic σ\sigma. By 3.3, σ\sigma covers a cusp-to-cusp geodesic in MM, since one end point of σ\sigma is at \infty. Therefore, the other end point of σ\sigma must be a cusp point of Γ\Gamma.

    The boundary of w1(2)w^{-1}({\mathbb{H}}^{2}_{\mathbb{R}}) is parametrized by 1+r(2τ)1+r^{\prime}(2-\tau) for rr^{\prime}\in{\mathbb{R}}, and the boundary of γ(Hs/t)\gamma(H_{s/t}) is parametrized by γ()+r\gamma(\infty)+r for some rr\in{\mathbb{R}}. Since cusp points of Γ\Gamma are contained in (z){}{\mathbb{Q}}(z)\cup\{\infty\}, there exist r,rr,r^{\prime}\in{\mathbb{R}} such that

    ±αtμ+r=1+r(2τ)=1+4rz(z)\pm\frac{\alpha}{t}\mu+r=1+r^{\prime}(2-\tau)=1+\frac{4r^{\prime}}{z}\in{\mathbb{Q}}(z)

    Therefore, r,r(z)=r,r^{\prime}\in{\mathbb{Q}}(z)\cap{\mathbb{R}}={\mathbb{Q}} because (z){\mathbb{Q}}(z) has no proper real subfield except {\mathbb{Q}}. Using Equation 6, we see that

    ±α/μ=(4rt+(1r)tz)\pm\alpha/\mu=(4r^{\prime}t+(1-r)tz)

    Since α[z]\alpha\in{\mathbb{Z}}[z] and μ\mu is a unit of [z]{\mathbb{Z}}[z], the right-hand side of the equation must belong to [z]{\mathbb{Z}}[z]. In particular, 4rt4r^{\prime}t and (1r)t(1-r)t are integers. To simplify our notation, we write α=(α1z+α0)μ\alpha=(\alpha_{1}z+\alpha_{0})\mu where αi\alpha_{i}\in{\mathbb{Z}}. Now we substitute the expression for α\alpha, η\eta, and δ\delta (see Equation 11) into the determinant of γ\gamma to obtain

    (α1z+α0)(sz+4s+2t)μ2±tμ1β\displaystyle(\alpha_{1}z+\alpha_{0})(sz+4s^{\prime}+2t)\mu^{2}\pm t\mu^{-1}\beta =1\displaystyle=1
    (α1z+α0)(sz+4s+2t)±tμβ\displaystyle(\alpha_{1}z+\alpha_{0})(sz+4s^{\prime}+2t)\pm t\mu\beta =z\displaystyle=-z

    Since μ,β[z]\mu,\beta\in{\mathbb{Z}}[z], we can write the equation above as

    α0(4s+2t)+[α1(4s+2t)+α0s]z+α1sz2+ti=0j1kizi=z\alpha_{0}(4s^{\prime}+2t)+[\alpha_{1}(4s^{\prime}+2t)+\alpha_{0}s]z+\alpha_{1}sz^{2}+t\sum_{i=0}^{j-1}k_{i}z^{i}=-z

    for kik_{i}\in{\mathbb{Z}}. Noting that [(z):]3[{\mathbb{Q}}(z):{\mathbb{Q}}]\geq 3, the equality above implies that

    α0(4s+2t)+tk0\displaystyle\alpha_{0}(4s^{\prime}+2t)+tk_{0} =0,\displaystyle=0,
    α1(4s+2t)+α0s+tk1\displaystyle\alpha_{1}(4s^{\prime}+2t)+\alpha_{0}s+tk_{1} =1,\displaystyle=-1,
    (α1s+tk2)\displaystyle(\alpha_{1}s+tk_{2}) =0.\displaystyle=0.

    Both ss and ss^{\prime} are units in /t{\mathbb{Z}}/t{\mathbb{Z}}, so the first and the third equations imply that

    αi0modt.\alpha_{i}\equiv 0\mod t.

    This contradicts the second equation. Therefore, there exists no such γΓ\gamma\in\Gamma in this case.

    Case 2:

    Suppose that the slope at θ\theta_{-} is 2/1-2/1. It follows that StabΓ(Hs/t)\operatorname{Stab}_{\Gamma}(H_{s/t}) contains γ1a2j1γ\gamma^{-1}a^{2}\ell_{j}^{-1}\gamma. The trace condition Equation 8, implies that

    tr(γ1a21γa21)=δ2(2τ)2=16δ2z2.\operatorname{tr}(\gamma^{-1}a^{2}\ell^{-1}\gamma a^{2}\ell^{-1})=\delta^{2}(2-\tau)^{2}=\frac{16\delta^{2}}{z^{2}}\in{\mathbb{Z}}.

    Let us write δ2=zk/16\delta^{2}=z\,k/16 for some kk\in{\mathbb{Z}}. Since δ[z]\delta\in{\mathbb{Z}}[z], we must have k/16k/16\in{\mathbb{Z}}. Now we write δ=±zk/16\delta=\pm z\sqrt{k/16}. Note that δ0\delta\neq 0 since Γ\Gamma does not contain element with trace 0. Since [(z):][{\mathbb{Q}}(z):{\mathbb{Q}}] is odd by [10, Theorem 1], the trace field (z){\mathbb{Q}}(z) does not contain the square root of any non-square integer. It follows that k/16\sqrt{k/16}\in{\mathbb{Z}}. In other words, δ=mz\delta=mz for some mm\in{\mathbb{Z}}. Substituting this into Equation 10, we have

    η=mzsz+4s+2ttz=mstzm(4s+2t)t[z].\eta=-mz\frac{sz+4s^{\prime}+2t}{tz}=-\frac{ms}{t}z-\frac{m(4s^{\prime}+2t)}{t}\in{\mathbb{Z}}[z].

    Since η\eta\in{\mathbb{Z}}, we see that ms/tms/t\in{\mathbb{Z}}. Since ss and tt are relatively prime, tt divides mm. We have the following containment of ideals η[z]mt[z]\eta{\mathbb{Z}}[z]\subseteq\frac{m}{t}{\mathbb{Z}}[z] and δ[z]=m[z]mt[z]\delta{\mathbb{Z}}[z]=m{\mathbb{Z}}[z]\subseteq\frac{m}{t}{\mathbb{Z}}[z]. In particular, (δ,η)[z]mt[z](\delta,\eta){\mathbb{Z}}[z]\subseteq\frac{m}{t}{\mathbb{Z}}[z]. Since αηβδ=1\alpha\eta-\beta\delta=1, (δ,η)[z]=[z](\delta,\eta){\mathbb{Z}}[z]={\mathbb{Z}}[z]. This implies that m=±tm=\pm t. We obtain the following expressions for η\eta and δ\delta

    η=±(sz+4s+2t)andδ=±tz.\eta=\pm(sz+4s^{\prime}+2t)\quad\text{and}\quad\delta=\pm tz. (12)

    Since the slope of Hs/tH_{s/t} at θ\theta_{-} is 2/1-2/1 and γ(θ)=\gamma(\theta_{-})=\infty, we observe that γ(Hs/t)\gamma(H_{s/t}) is a vertical plane with slope 2/1-2/1 at \infty. It must intersect the hyperplane 2{\mathbb{H}}^{2}_{\mathbb{R}} at a vertical geodesic σ\sigma. By 3.3, σ\sigma covers a cusp-to-cusp geodesic in MM, since one end point of σ\sigma is \infty. The other end point of σ\sigma is a cusp point on the line at infinity of 2{\mathbb{H}}^{2}_{\mathbb{R}} and so must be a rational number. In other words, there exists a real number rr such that

    αtz+r4z.\frac{\alpha}{tz}+r\frac{4}{z}\in{\mathbb{Q}}.

    This implies that r(z)=r\in{\mathbb{Q}}(z)\cap{\mathbb{R}}={\mathbb{Q}} since (z){\mathbb{Q}}(z) has no proper real subfield except {\mathbb{Q}}. Therefore α=α1z+α0[z]\alpha=\alpha_{1}z+\alpha_{0}\in{\mathbb{Z}}[z]. Now we substitute the expression for α\alpha, η\eta and δ\delta (see Equation 12) into the determinant of γ\gamma to obtain

    (α1z+α0)(sz+4s+2t)tzβ=±1modt(\alpha_{1}z+\alpha_{0})(sz+4s^{\prime}+2t)-tz\beta=\pm 1\mod t

    Since μ,β[z]\mu,\beta\in{\mathbb{Z}}[z], we can write the equation above as

    α0(4s+2t)+[α1(4s+2t)+α0s]z+α1sz2+ti=0j1kizi=±1\alpha_{0}(4s^{\prime}+2t)+[\alpha_{1}(4s^{\prime}+2t)+\alpha_{0}s]z+\alpha_{1}sz^{2}+t\sum_{i=0}^{j-1}k_{i}z^{i}=\pm 1

    Since [(z):]3[{\mathbb{Q}}(z):{\mathbb{Q}}]\geq 3, the equality above implies that

    α0(4s+2t)+tk0\displaystyle\alpha_{0}(4s^{\prime}+2t)+tk_{0} =±1\displaystyle=\pm 1
    α1(4s+2t)+α0s+tk1\displaystyle\alpha_{1}(4s^{\prime}+2t)+\alpha_{0}s+tk_{1} =0\displaystyle=0
    α1s+tk2\displaystyle\alpha_{1}s+tk_{2} =0\displaystyle=0

    Since ss is a unit in /t{\mathbb{Z}}/t{\mathbb{Z}}, the third equation implies that α10modt\alpha_{1}\equiv 0\mod t. Given that α10modt\alpha_{1}\equiv 0\mod t and s(/t)×s\in({\mathbb{Z}}/t{\mathbb{Z}})^{\times}, the second equation implies that α00modt\alpha_{0}\equiv 0\mod t. These conditions on αi\alpha_{i} contradict the first equation. Consequently, there exists no such γ\gamma in this case.

    The analysis of both cases shows that θ\theta_{-} is not a cusp point of Γ\Gamma because there is no γΓ\gamma\in\Gamma such that γ()=θ\gamma(\infty)=\theta_{-}. This contradiction implies that Hs/tH_{s/t} does not cover a finite area totally geodesic surface. This completes the proof of 1.3. ∎

The existence of infinitely many twist knot complements with a unique totally geodesic surface follows as a consequence.

  • Proof of 1.4.

    Let MM have an odd prime number of twists. Note that tr(ρ(π1Σ))\operatorname{tr}(\rho(\pi_{1}\Sigma))\subset{\mathbb{Z}} because Γ\Gamma has integral traces, (trΓ){\mathbb{Q}}(\operatorname{tr}\Gamma) has no proper real subfield (except {\mathbb{Q}}) because it has odd prime degree over \mathbb{Q} by [10, Theorem 1], and ρ(π1Σ)\rho(\pi_{1}\Sigma) is a Fuchsian subgroup. Thus the conditions of 1.3 are satisfied.

    Using Magma [5] and SageMath [20], we verify that Γj\Gamma_{j} also satisfies the conditions of 1.3 for odd integers with 3j993\leq j\leq 99, see [19] for the code.

    Twist knot complements with distinct numbers of half-twists are always non-commensurable because their invariant trace fields (zj){\mathbb{Q}}(z_{j}) are pairwise distinct. Thus we have a family of infinitely many non-commensurable 33-manifolds with exactly one totally geodesic surface. ∎

3.3 Lifting the thrice-punctured sphere

We at last deduce the proof of 1.2, reinstating the index j.

  • Proof of 1.2.

    In view of 1.4, to obtain a finite volume hyperbolic 33-manifold with exactly kk totally geodesic surfaces, it suffices to find covers of MjM_{j} such that NN has exactly kk lifts for some odd prime jj. We construct these manifolds from dihedral covers of the twist knot complements. We first count the number of lifts of NN to dihedral covers of twist knot with an odd number of half-twists. For this purpose, let jj be a positive odd integer.

    We write down an explicit surjective homomorphism from Γj\Gamma_{j} to the dihedral group DnD_{n} of order 2n2n. The group DnD_{n} has the following presentation

    Dn=α,βα2,β2,(αβ)n.D_{n}=\langle\alpha,\beta\mid\alpha^{2},\beta^{2},(\alpha\beta)^{n}\rangle.

    Let us fix an integer n>1n>1 dividing 2j+12j+1. We have a surjective homomorphism φ:ΓjDn\varphi:\Gamma_{j}\to D_{n} defined by φ(a)=α\varphi(a)=\alpha and φ(b)=β\varphi(b)=\beta since φ(wjawj1b1)=φ(ab)2j+1\varphi(w_{j}aw_{j}^{-1}b^{-1})=\varphi(ab)^{2j+1}. Let M^j\hat{M}_{j} be the cover corresponding to kerφ\ker\varphi.

    We observe that

    φ(yj)=φ((awj)2a(awj)2)=α\varphi(y_{j})=\varphi((aw_{j})^{2}a(aw_{j})^{-2})=\alpha

    since φ(awj)\varphi(aw_{j}) has order 2. Therefore, φ(Δj)=α\varphi(\Delta_{j})=\langle\alpha\rangle is a cyclic subgroup of order 2. Let N0N_{0} be a lift of NN to M^j\hat{M}_{j} corresponding to kerφ|Δj\ker\varphi|_{\Delta_{j}}. Since M^j\hat{M}_{j} is a regular cover, the deck group DnD_{n} acts transitively on the lifts of NN to M^j\hat{M}_{j}. We identify these lifts with the orbit of N0N_{0} under the action of DnD_{n}. Since the stabilizer of N0N_{0} under the action of DnD_{n} is α\langle\alpha\rangle, we identify the lifts of NN to M^j\hat{M}_{j} with the left cosets of α\langle\alpha\rangle in DnD_{n}. In particular, NN lifts to nn distinct totally geodesic surfaces {Nm}m=0n1\left\{N_{m}\right\}_{m=0}^{n-1} in M^j\hat{M}_{j}.

    Observe that the cyclic subgroup αβ\langle\alpha\beta\rangle of order nn forms a complete left coset representatives of α\langle\alpha\rangle in DnD_{n}. Up to reordering, we identify NmN_{m} with the coset (αβ)mα(\alpha\beta)^{m}\langle\alpha\rangle for 0mn10\leq m\leq n-1. The action of α\alpha on the lifts NN is given by

    αNm=Nnm,\alpha\cdot N_{m}=N_{n-m},

    where Nn:=N0N_{n}:=N_{0}, since α(αβ)mα=(αβ)nmα\alpha\cdot(\alpha\beta)^{m}\langle\alpha\rangle=(\alpha\beta)^{n-m}\langle\alpha\rangle. Since nn is an odd integer, α\alpha leaves invariant exactly one surface N0N_{0} and permutes n1n-1 surfaces pairwise. It follows that M^j/α\hat{M}_{j}/\langle\alpha\rangle contains exactly 1+n12=(n+1)/21+\frac{n-1}{2}=(n+1)/2 lifts of NN.

    Any integer k2k\geq 2 can be written as (n+1)/2(n+1)/2 where n=2k13n=2k-1\geq 3. To prove the theorem, it suffices to show that for any odd integer nn there exist infinitely many primes pp such that nn divides 2p+12p+1. Let n=2q+1n=2q+1 and thus gcd(q,n)=1\gcd(q,n)=1. By Dirichlet’s theorem, the arithmetic progression {q+nk}k=0\{q+nk\}_{k=0}^{\infty} contains infinitely many primes pi=q+nkip_{i}=q+nk_{i}. Note that nn divides 2pi+12p_{i}+1 for all ii. This completes the proof of 1.2. ∎

3.4 Non-elementary Fuchsian subgroups

The proof of 3.5 gives infinitely many explicit examples of non-elementary maximal Fuchsian subgroups of infinite covolume in Γj\Gamma_{j}.

Proposition 3.6.

Let MM be a twist knot complement with at least 33 half-twists. Let Hs/tH_{s/t} be the vertical hyperplane intersecting 2{\mathbb{H}}^{2}_{\mathbb{R}} at s/ts/t\in{\mathbb{Q}} with slope 2/1-2/1 at \infty. For all odd tt, the stabilizer of Hs/tH_{s/t} in Γ\Gamma is non-elementary.

  • Proof.

    We observe that StabΓ(Hs/t)\operatorname{Stab}_{\Gamma}(H_{s/t}) contains a subgroup generated by

    a2and(ts4st)b(ts4st)a^{-2}\ell\quad\text{and}\quad\begin{pmatrix}-t^{\prime}&s\\ -4s^{\prime}&t\end{pmatrix}b\begin{pmatrix}t&-s\\ 4s^{\prime}&-t^{\prime}\end{pmatrix}

    where 4sstt=14ss^{\prime}-tt^{\prime}=1. The fact that the latter element belongs to Γj\Gamma_{j} follows from Remark 2.6. It is clear that a2a^{-2}\ell stabilizes Hs/tH_{s/t}, so the line at infinity of Hs/tH_{s/t} has slope 2/1-2/1 and thus is parametrized by

    st+r(2τ)\frac{s}{t}+r(2-\tau)

    for rr\in{\mathbb{R}}. We have

    (ts4st)b(ts4st)(st+4rz)=st+r4rt2+14z=st+r4rt2+1(2τ)\begin{pmatrix}-t^{\prime}&s\\ -4s^{\prime}&t\end{pmatrix}b\begin{pmatrix}t&-s\\ 4s^{\prime}&-t^{\prime}\end{pmatrix}\left(\frac{s}{t}+\frac{4r}{z}\right)=\frac{s}{t}+\frac{r}{4rt^{2}+1}\frac{4}{z}=\frac{s}{t}+\frac{r}{4rt^{2}+1}(2-\tau)

    for any rr\in{\mathbb{R}}. Since r/(4rt2+1)r/(4rt^{2}+1)\in{\mathbb{R}}, the image point also belongs to the line at infinity of Hs/tH_{s/t}. Therefore, the latter element stabilizes Hs/tH_{s/t}.

    These are independent parabolic element with respect to the action of the stabilizer of Hs/tH_{s/t} on Hs/tH_{s/t} because they have distinct fixed points. They thus generate a non-elementary subgroup of the stabilizer of Hs/tH_{s/t}. ∎

Since all vertical lifts of NN are described at the end of Section 2, for s/ts/t\in{\mathbb{Q}}\smallsetminus{\mathbb{Z}} and odd tt, the image of Hs/tH_{s/t} in MjM_{j} is not freely homotopic to NN. In particular, StabΓj(Hs/t)\operatorname{Stab}_{\Gamma_{j}}(H_{s/t}) and Δj\Delta_{j} are not conjugate in Γj\Gamma_{j}. By [3, Theorem 1.1], at most finitely many Γj\Gamma_{j}-orbits of hyperplanes Hs/tH_{s/t} cover a finite area totally geodesic surface. Other quotients of these hyperplanes must be of infinite area. Furthermore, if the trace field of MjM_{j} satisfies the condition in 1.3, then 1.3 says that Δj\Delta_{j} is the unique (up to conjugacy) finite covolume maximal Fuchsian subgroup of Γj\Gamma_{j}; that is,

Corollary 3.7.

Let jj be an odd integer such that (zj){\mathbb{Q}}(z_{j}) does not contain any proper real subfield. Then for all s/ts/t\in{\mathbb{Q}}\smallsetminus{\mathbb{Z}} and odd tt, StabΓj(Hs/t)\operatorname{Stab}_{\Gamma_{j}}(H_{s/t}) is a non-elementary Fuchsian group of infinite covolume.

In other words, 3.7 provides examples of infinitely many hyperbolic 33-manifolds that contain non-elementary Fuchsian subgroups of infinite covolume. This behavior demonstrates a drastic difference between arithmetic and non-arithmetic hyperbolic manifolds. In an arithmetic hyperbolic 33-manifold, every non-elementary Fuchsian subgroup has finite covolume [14, Theorem 4]. Infinite covolume Fuchsian groups of this kind are known to exist, such as using results in [9], but we do not know of any explicit examples in the literature.

3.5 Right-angled knot complements

We now discuss an application of our boundary slope analysis. A hyperbolic manifold is called right-angled if it can be constructed by gluing together a set of hyperbolic polyhedra whose dihedral angles are all π/2\pi/2. Being right-angled is ever-present in 33-dimensional topology, such as showing that a manifold is LERF [8, Corollary 1.4] and addressing the virtually fibered conjecture [8, Corollary 1.2].

It is known the the faces of these polyhedra give rise to immersed totally geodesic surfaces [7]. In the case of a cusped hyperbolic 33-manifold, at least one polyhedron must admit an ideal vertex. The intersection of this polyhedron with a sufficiently small horosphere based at the ideal vertex must be a Euclidean right-angled polygon. Sides of this polygon being perpendicular may be checked by looking at the angle between two intersecting hyperplanes.

There are examples in [6] of knot complements that are not right-angled because they contain no totally geodesic surface. It is also known that all knots up to 1111 crossings are not right-angled because of volume obstructions [7]. However, we authors do not know of any existing literature that provides an infinite list of knot complements that are not right-angled. We give the first such infinite family.

  • Proof of 1.5.

    For jj odd prime, we check that the right-angled property fails by finding the angle between the real hyperplane through 0 and the hyperplane of slope 2/1-2/1 through 0. This angle is π/2\pi/2 if and only if the argument of 1+τ2-1+\frac{\tau}{2} is π/2\pi/2.

    Refer to caption
    Figure 4: Vertical lifts of NN for jj odd along with the angle that must be π/2\pi/2

We know that τ=4z1+2\tau=-4z^{-1}+2, so 1+τ2=2z1-1+\frac{\tau}{2}=-2z^{-1}, whose argument is bounded away from π/2\pi/2 by [10, Theorem 1]. Therefore, no twist knot complement of this form is right-angled. ∎

4 Totally geodesic surfaces in the figure-8 knot complement

The figure-8 knot complement MM is exceptional among twist knot complements. It is the sole arithmetic knot complement [17] and therefore admits infinitely many totally geodesic surfaces [13]. Two such surfaces are shown in Figure 5. One is the previously defined NN. The other is a dual surface NN^{\prime} whose boundary slopes are 1/01/0 and 2/12/1; this information can be calculated using the techniques in Section 2 and visualized in Figure 6.

Refer to caption
(a)
Refer to caption
(b)
Figure 5: Two distinct totally geodesic thrice-punctured spheres NN and NN^{\prime} in the figure-8 knot complement
Refer to caption
Figure 6: Vertical lifts of NN (blue) and NN^{\prime} (red, dashed) for jj odd

The occurrence of the boundary slope 2/12/1 for NN^{\prime} is in stark contrast with the restricted set of boundary slopes available to other twist knot complements found in 3.5. The approach of the proof of 3.5 breaks down for the figure-8 knot, primarily because the minimal polynomial of zz is quadratic. Moreover,

Theorem 4.1.

Every number in {}\mathbb{Q}\cup\{\infty\} occurs as a boundary slope for some totally geodesic surface in the figure-8 knot complement.

  • Proof.

    Let apqa^{p}\ell^{q} be a parabolic element fixing \infty. We can find wamnw1wa^{m}\ell^{n}w^{-1} such that the trace condition Equation 8 is satisfied. As computed in the proof of 3.5, we must satisfy

    16nq+(2n+m)(2q+p)z2z16nq+(2n+m)(2q+p)z^{2}\in\mathbb{Q}z

    Since the minimal polynomial of zz is z2z+1=0z^{2}-z+1=0, satisfying the above is equivalent to satisfying

    16nq(2n+m)(2q+p)=016nq-(2n+m)(2q+p)=0

    For any pair (p,q)(p,q), there exists such a pair (m,n)(m,n). Thus the group Δ\Delta generated by apqa^{p}\ell^{q} and wamnw1wa^{m}\ell^{n}w^{-1} is a non-elementary Fuchsian subgroup. Because the figure-8 knot is arithmetic, every non-elementary Fuchsian subgroup in π1(M)\pi_{1}(M) sits inside a finite coarea arithmetic Fuchsian subgroup of π1(M)\pi_{1}(M) [14, Theorem 4]. Thus there exists ΔΔ\Delta^{\prime}\supset\Delta such that Δ\Delta^{\prime} is the stabilizer of a hyperplane which corresponds to a totally geodesic surface in the figure-8 knot complement admitting boundary slope p/qp/q. ∎

5 Computer experiments and questions

5.1 Trace fields of twist knots

The condition that the trace field (zj){\mathbb{Q}}(z_{j}) of KjK_{j} for an odd prime jj has no proper real subfield is used frequently in the proof of 1.3 and the lemmas leading up to it. Using Magma, we observe that for 1<j<1001<j<100, the trace field (zj){\mathbb{Q}}(z_{j}) has no proper subfield except for {\mathbb{Q}}, see [19] for the code. The following question arises naturally:

Question 5.1.

Is {\mathbb{Q}} the unique proper subfield of the trace field (zj){\mathbb{Q}}(z_{j}) of KjK_{j}?

We also checked that the Galois group of the Riley polynomial Gal(Λj)\operatorname{Gal}(\Lambda_{j}) is congruent to SjS_{j}, the full symmetric group on jj letters for 1<j991<j\leq 99, see [19]. This implies that there are no proper subfields between (zj){\mathbb{Q}}(z_{j}) and {\mathbb{Q}} since a subgroup of index jj in SjS_{j} is necessarily maximal. Therefore, it is natural to ask:

Question 5.2.

Is Gal(Λj)Sj\operatorname{Gal}(\Lambda_{j})\cong S_{j} for all jj?

We conjecture that the answer is yes to both questions. By 1.3, an affirmative answer to either question implies that NN is the unique totally geodesic surface in MjM_{j} all for odd jj.

Finally, using Magma we compute the class number of (zj){\mathbb{Q}}(z_{j}) and observe that it is 11 for all 2j192\leq j\leq 19, see [19]. As a consequence of the argument in the proof of 1.3, the cusp points of Γj\Gamma_{j} form a proper subset of the trace field (zj){\mathbb{Q}}(z_{j}). In other words, for odd integer 2j192\leq j\leq 19, the group Γj\Gamma_{j} gives an example of a Kleinian group whose trace field has class number 11 and contains cusp points as a proper subset. Examples of this phenomenon are only known to exist for Kleinian groups whose trace field has class number strictly bigger than 1 [11, Theorem 6.1]. One may ask if the class number of (zj){\mathbb{Q}}(z_{j}) is 11 for all jj. However, the class number of (z20)\mathbb{Q}(z_{20}) is computed to be 22, see [19].

5.2 Homological behavior of NN in congruence covers

We study the homological behavior of totally geodesic surfaces in congruence covers of Γj\Gamma_{j}. This study is motivated by the analogy between totally geodesic surfaces in (arithmetic) hyperbolic manifolds and generic hyperplane sections in the projective variety described in [4]. For a (complex) projective variety MM of dimension nn, the Lefschetz Hyperplane Theorem says that the map

Hi(MH)Hi(M)H_{i}(M\cap H)\to H_{i}(M)

is injective for all ini\geq n and for a generic hyperplane HH. Replacing ”hyperplane section” by ”totally geodesic surface in hyperbolic 33-manifold”, one can ask similar questions about the behavior of the induced map between the first homology of totally geodesic surfaces and the ambient hyperbolic 33-manifold.

To describe our question, we need to set up some notations. Let 𝒪j\mathcal{O}_{j} be the ring of integers of (zj){\mathbb{Q}}(z_{j}) and identify Γj\Gamma_{j} with its image in PSL2(𝒪j)\operatorname{PSL}_{2}(\mathcal{O}_{j}). Let Γj(I):=ker(ΓjPSL2(𝒪j/I))\Gamma_{j}(I):=\ker(\Gamma_{j}\to\operatorname{PSL}_{2}(\mathcal{O}_{j}/I)) be the principal congruence cover of level II for some ideal II and Mj(I)M_{j}(I) be the cover of MjM_{j} that corresponds to Γj(I)\Gamma_{j}(I). Let N1,,NmN_{1},\dots,N_{m} be the lifts of NN in Mj(I)M_{j}(I).

Question 5.3.

What are the possible behaviors of the map

H1(N1Nm;)H1(Mj(I);)H_{1}(N_{1}\cup\dots\cup N_{m};{\mathbb{Z}})\to H_{1}(M_{j}(I);{\mathbb{Z}})

as II varies over the prime ideals in 𝒪j\mathcal{O}_{j}?

We first investigate this question when II is a prime with residue degree 11, i.e. when 𝒪j/I𝔽p\mathcal{O}_{j}/I\cong\mathbb{F}_{p}, a finite field of prime order pp. Suppose that 𝔭\mathfrak{p} is a prime ideal in 𝒪j\mathcal{O}_{j} lying above an odd prime pp such that 𝒪j/𝔭𝔽p\mathcal{O}_{j}/\mathfrak{p}\cong\mathbb{F}_{p}. In [18, Proposition 4], Riley showed that the representation ΓjPSL2(𝔽p)\Gamma_{j}\to\operatorname{PSL}_{2}(\mathbb{F}_{p}) coming from the reduction map is surjective for all odd prime pp. For all odd prime pp, Δj\Delta_{j} maps onto the group

(1101),(1041)=PSL2(𝔽p).\left\langle\begin{pmatrix}1&1\\ 0&1\\ \end{pmatrix},\begin{pmatrix}-1&0\\ 4&-1\\ \end{pmatrix}\right\rangle=\operatorname{PSL}_{2}(\mathbb{F}_{p}).

under the reduction map Therefore for any prime ideal 𝔭\mathfrak{p} with residue degree 11 lying above an odd prime pp, Mj(𝔭)M_{j}(\mathfrak{p}) contains exactly one unique lift of the totally geodesic thrice-puncture sphere, and this cover has fundamental group Δj(𝔭)=ΔΓj(𝔭)\Delta_{j}(\mathfrak{p})=\Delta\cap\Gamma_{j}(\mathfrak{p}). For all examples that we computed with Magma, we observed that H1(Δj(𝔭))H_{1}(\Delta_{j}(\mathfrak{p})) maps onto H1(Γj(𝔭))H_{1}(\Gamma_{j}(\mathfrak{p})) under the induced map for prime ideals 𝔭\mathfrak{p} of residue degree 11 lying above an odd prime pp. In particular, we verified that H1(Δj(𝔭))H_{1}(\Delta_{j}(\mathfrak{p})) surjects H1(Γj(𝔭))H_{1}(\Gamma_{j}(\mathfrak{p})) for 3j203\leq j\leq 20 that has a prime ideal of residue degree 11 lying above 33 and 55, see [19].

Question 5.4.

Is the induced map H1(Δj(𝔭))H1(Γj(𝔭))H_{1}(\Delta_{j}(\mathfrak{p}))\to H_{1}(\Gamma_{j}(\mathfrak{p})) surjective for all prime ideals 𝔭\mathfrak{p} of residue degree 1 lying above odd prime pp?

Finally, we record some observations of patterns in the first homology of these principal congruence covers in Table 1. In the table, we denote the homology group

0r(/a1)r1(/as)rs{\mathbb{Z}}^{r}_{0}\oplus({\mathbb{Z}}/a_{1}{\mathbb{Z}})^{r_{1}}\oplus\dots\oplus({\mathbb{Z}}/a_{s}{\mathbb{Z}})^{r_{s}}

by [0r0,a1r1,,asrs][0^{r_{0}},a_{1}^{r_{1}},\dots,a_{s}^{r_{s}}] where a1,,asa_{1},\dots,a_{s} are the invariant factors. Furthermore, repetition of a prime pp indicates a prime with different prime ideals lying above of residue degree 1. See , see [19] for the code used to produce the table.

pp jj kk H1(Γj(𝔭))H_{1}(\Gamma_{j}(\mathfrak{p}))
3 4k+14k+1 1k131\leq k\leq 13 [04,(3k+1)2][0^{4},(3k+1)^{2}]
5 10k+110k+1 1k41\leq k\leq 4 [012,(6k+1)5][0^{12},(6k+1)^{5}]
5 10k+310k+3 1k41\leq k\leq 4 [017,(3k+1)4,6k+2][0^{17},(3k+1)^{4},6k+2]
Table 1: Homology of Principal Congruence Covers

5.3 Finding surfaces in the figure-8 knot complement

Given the linear relation between p/qp/q and m/nm/n that is described in both 3.5 and the proof of 4.1, we ask how restrictive such relations between boundary slopes must be:

Question 5.5.

Can any pair of numbers in {}\mathbb{Q}\cup\{\infty\} occur simultaneously as boundary slopes for some totally geodesic surface of MM?

We give some computed examples of totally geodesic surfaces Σ\Sigma in MM below. We identify these by specifying a vertical hyperplane in 3{\mathbb{H}}^{3} and its stabilizer. These hyperplanes have the form (s/t,p/q){\mathbb{H}}(s/t,p/q) where 0s/t<10\leq s/t<1 is the point of intersection with H2H^{2}_{\mathbb{R}} and p/qp/q is the slope of the vertical plane at infinity. We denote the stabilizers of these hyperplanes at infinity by Δ(s/t,p/q)\Delta(s/t,p/q).

Example 5.6.

The plane (0,1/0){\mathbb{H}}(0,1/0) is 2{\mathbb{H}}^{2}_{{\mathbb{R}}} and has stabilizer

Δ(0,1/0)=a,aawa1w=(1101),(1143)\Delta(0,1/0)=\langle a,{}^{awa^{-1}w}a\rangle=\left\langle\begin{pmatrix}1&1\\ 0&1\end{pmatrix},\begin{pmatrix}-1&1\\ -4&3\end{pmatrix}\right\rangle

This is the thrice-punctured sphere NN.

Example 5.7.

The plane (0,2/1){\mathbb{H}}(0,2/1) has stabilizer

Δ(0,2/1)a2,aab1=(14z01),(2zz+10)\displaystyle\Delta(0,2/1)\supseteq\langle a^{2}\ell,{}^{ab^{-1}}a\rangle=\left\langle\begin{pmatrix}-1&-4z\\ 0&-1\end{pmatrix},\begin{pmatrix}2&-z\\ -z+1&0\end{pmatrix}\right\rangle

Note that

(14z01)(2zz+10)=(2zz10)\begin{pmatrix}-1&-4z\\ 0&-1\end{pmatrix}\begin{pmatrix}2&-z\\ -z+1&0\end{pmatrix}=\begin{pmatrix}2&z\\ z-1&0\end{pmatrix}

This is the thrice-punctured sphere NN^{\prime}.

Both NN and NN^{\prime} admit precisely two boundary slopes, but that is not necessarily true for all totally geodesic surfaces:

Example 5.8.

The plane (0,0){\mathbb{H}}(0,0) has stabilizer

Δ(0,0)\displaystyle\Delta(0,0) ,a6w,a6w1,w1\displaystyle\supseteq\langle\ell,{}^{w}a^{6}\ell,{}^{\ell w^{-1}}a^{6}\ell,{}^{\ell w^{-1}}\ell\rangle
=(14z+201),(108z+41),(1124z+128z+413)\displaystyle\phantom{\supseteq}=\left\langle\begin{pmatrix}-1&-4z+2\\ 0&-1\end{pmatrix},\begin{pmatrix}-1&0\\ -8z+4&-1\end{pmatrix},\begin{pmatrix}11&-24z+12\\ -8z+4&-13\end{pmatrix}\right\rangle

The set of boundary slopes of Δ(0,0)\Delta(0,0) contains but may not be limited to {0,6/1}\{0,6/1\}.

Example 5.9.

The plane (0,1/1){\mathbb{H}}(0,1/1) has stabilizer

Δ(0,1/1)\displaystyle\Delta(0,1/1) a,a10w3,av1\displaystyle\supseteq\langle a\ell,{}^{w}a^{10}\ell^{3},{}^{\ell v}a\ell^{-1}\rangle
=(14z+101),(1016z+121)(2552z+1316z+1227)\displaystyle\phantom{\supseteq}=\left\langle\begin{pmatrix}-1&-4z+1\\ 0&-1\end{pmatrix},\begin{pmatrix}-1&0\\ -16z+12&-1\end{pmatrix}\begin{pmatrix}25&-52z+13\\ -16z+12&-27\end{pmatrix}\right\rangle

The set of boundary slopes of Δ(0,1/1)\Delta(0,1/1) contains but may not be limited to {1/1,10/3,1/1}\{1/1,10/3,-1/1\}.

Given the existence of sets of boundary slopes that are not precisely two numbers in {}\mathbb{Q}\cup\{\infty\}, we ask:

Question 5.10.

Does there exist a totally geodesic surface of MM admitting exactly 11 boundary slope? arbitrarily many boundary slopes?

References

  • [1] C. Adams, Thrice-punctured spheres in hyperbolic 33-manifolds, Trans. Am. Math. Soc. 287 (1985) 645–656.
  • [2] C. Adams, E. Schoenfeld, Totally geodesic Seifert surfaces in hyperbolic knot and link complement I, Geom. Dedicata 116 (2005) 237–247.
  • [3] U. Bader, D. Fisher, N. Miller, M. Stover, Arithmeticity, superrigidity, and totally geodesic submanifolds, Ann. of Math. 193 (2021) 837–861.
  • [4] N. Bergeron, Lefschetz properties for arithmetic real and complex hyperbolic manifolds, Int. Math. Res. Not. 20 (2003) 1089–1122.
  • [5] W. Bosma, J. Cannon, and C. Playoust, The Magma algebra system. I. The user language, J. Symbolic Comput., 24 (1997), 235–265.
  • [6] D. Calegari, Real places and torus bundles, Geom. Dedicata 118 (2006) 209–227.
  • [7] A. Champanerkar, I. Kofman, J. Purcell, Right-angled polyhedra and alternating links, arXiv:1910.13131 (2019).
  • [8] E. Chesebro, J. Deblois, H. Wilton, Some virtually special hyperbolic 33-manifold groups, European Mathematical Society, 87 (2012) 727–787.
  • [9] D. Fisher, J. Lafont, N. Miller, M. Stover, Finiteness of maximal geodesic submanifolds in hyperbolic hybrids, J. Eur. Math. Soc. (JEMS) 23 (2021), no. 11, 3591–3623.
  • [10] J. Hoste, P. Shananhan, Trace fields of twist knots, J. Knot Theory Ramifications 10(4) (2001) 625–639.
  • [11] D. Long, A. Reid, Fields of definition of canonical curves, Contemporary Math. 541, A.M.S. Publications, (2011) 247–257.
  • [12] G. Margulis, A. Mohammadi, Arithmeticity of hyperbolic 33-manifolds containing infinitely many totally geodesic surfaces, Ergodic Theory and Dynamical Systems, (2021) 1–32.
  • [13] C. Maclachlan, A. Reid, The Arithmetic of Hyperbolic 33-manifolds, GTM 219, Springer-Verlag New York Inc. (2003).
  • [14] C. Maclachlan, A. Reid, Commensurability classes of arithmetic Kleinian groups and their Fuchsian subgroups, Math. Proc. Camb. Phil. Soc. 102 (1987) 251–258.
  • [15] W. Menasco, A. Reid, Totally geodesic surfaces in hyperbolic link complements, Topology 90 (1992) 215–226.
  • [16] A. Reid, Totally geodesic surfaces in hyperbolic 33-manifolds, Proc. Edinburgh Math. Soc. (2) 34 (1991) no. 1, 77–88
  • [17] A. Reid, Arithmeticity of knot complements, J. London Math. Soc. 43 (1991), 171–184.
  • [18] R. Riley, Parabolic representations of knot groups I, Proc. L.M.S. 24 (1972) 217–242.
  • [19] aux/* in the source of the arXiv version of this preprint [2009.04637].
  • [20] SageMath, the Sage Mathematics Software System (Version 8.4), The Sage Developers, 2018, https://www.sagemath.org.

Department of Mathematics, Temple University,
Wachman Hall Rm. 638
1805 N. Broad St
Philadelphia PA, 19122 USA
E-mail addresses:
[email protected], [email protected],