Totally geodesic surfaces in twist knot complements
Abstract
In this article, we give explicit examples of infinitely many non-commensurable (non-arithmetic) hyperbolic -manifolds admitting exactly totally geodesic surfaces for any positive integer , 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 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 -manifolds that contain them. In this paper, we will mainly be concerned with complete properly immersed totally geodesic surfaces in hyperbolic -manifolds. There has been considerable work in understanding the existence of totally geodesic surfaces in hyperbolic -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 -manifolds that do contain totally geodesic surfaces. Adams [1, Theorem 3.1] proved that any incompressible thrice-punctured sphere in a hyperbolic -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 -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 -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 -manifolds [9, Theorem 1.2]. In a subsequent work, Bader, Fisher, Miller and Stover showed that if a complete finite-volume hyperbolic -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 -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 , is there a hyperbolic -manifold containing exactly totally geodesic surfaces?
We answer yes to the above question by exhibiting explicit examples of hyperbolic -manifolds with exactly totally geodesic surfaces:
Theorem 1.2.
Let be any positive integer. There exist infinitely many non-commensurable hyperbolic -manifolds with exactly totally geodesic surfaces.
These are the first examples of hyperbolic -manifolds for which the number of totally geodesic surfaces is positive, finite, and precisely known. The idea behind 1.2 is to find hyperbolic -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 , let be the twist knot with exactly half-twists as shown in Figure 1. Let . These twist knots admit a complete finite-volume hyperbolic metric if and only if . In other words for , the knot group admits a discrete faithful representation into . By Adams [1, Theorem 3.1], the thrice-punctured sphere , seen as the dotted surface in Figure 1, is totally geodesic for . For simplification of notation, we will use to denote this surface, independent of .

By Reid [16, Theorem 3], among the twist knots with (all but the figure-8 knot, in our notation), there are infinitely many so that 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 for . The following theorem of our paper confirms this conjecture for infinitely many values of .
Theorem 1.3.
Let be the complement of a twist knot whose trace field has odd degree over and contains no proper real subfield besides . Then the thrice-punctured sphere is the unique totally geodesic surface in .
Since can be obtained by doing Dehn filling on the Whitehead link, the volume of 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 , the trace field of 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 for all prime . Using Magma [5] and SageMath [20], we also verify that the condition in 1.3 holds for the trace field of for odd and , see [19] for the code. Therefore, we obtain the following:
Corollary 1.4.
Suppose is an odd prime or is an odd number less than or equal to . The thrice-punctured sphere is the unique totally geodesic surface in .
The analysis in the proof of 1.3 has several interesting consequences. For all odd primes , we give explicit examples of non-elementary maximal Fuchsian subgroup of infinite covolume in . 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 -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 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 , our analysis of the geometry of in 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 is never a right angle.
Corollary 1.5.
The twist knot is not right-angled for 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 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 be the twist knot with half-twists, , and be its fundamental group. The group has a presentation where is given by a word of length in and , namely:
(1) |
The two generators , correspond to the two meridians of the twist knot chosen to have the same orientation. The longitude of the twist knots that commute with can be computed diagrammatically and is given by:
(2) |
where is spelled backwards. Finally, we note that the above presentation is different from that in [10, Equation 2.1] by an isomorphism interchanging and .
For , admits a complete hyperbolic structure of finite volume. The discrete faithful representation sends all conjugates of the meridians of the knot to parabolic isometries; i.e. conjugate in to:
Theorem 2.1.
Suppose we normalize so that:
Then defines the discrete faithful representation if and only if satisfies a polynomial . The polynomial is defined recursively by
(3) |
where , , and .
The fact that 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 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 under is contained in . In particular, the entries of the matrices in can be written as -linear combinations of elements in the -power basis.
Since the presentation of 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 on two generators and along with the surjective homomorphism sending , to , respectively. Let be the lift of given by Equation 1. Consider the homomorphism defined by
We claim that
(4) for some .
For convenience of notation, we identify elements of with their image under . We consider two cases according to whether is even or odd. When is even, we have . Applying Cayley–Hamilton to , we get
Note that . Therefore, we have
(5) This gives us the recursion in Equation 3. Now we observe that the upper diagonal (resp. lower diagonal) entries of and of give us the initial conditions for (resp. ). 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 . When is odd, we proceed similarly to compute where the initial conditions are given by and .
To finish the proof, we note that is obtained by factoring through the canonical projection where is the homomorphism induced by the evaluation map . Using Equation 4, we see that the relation is satisfied if and only if . ∎
Given a finitely-generated non-elementary subgroup of , we can define the trace field of to be and the invariant trace field of to be . If additionally has finite covolume, then and 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 , but the invariant trace field is. These fields may coincide, such as link complements in a -homology sphere. [13, Corollary 4.2.2]. Finally, suppose that is generated by and ; then [13, Equation 3.25].
It follows from the discussion above and 2.1 that the trace field of twist knot complements is . Hoste and Shanahan proved that is the monic minimal polynomial of over of degree [10, Theorem 1]. Therefore, and so twist knots are pairwise non-commensurable.
Corollary 2.3.
Under the discrete faithful representation , we have
where . Furthermore, the image of is given by:
where .
-
Proof.
Let us consider to be the lift of in given by spelling backwards. We first claim that:
Observe that satisfy the same recurrence in Equation 5, where , provide the base case.
We next claim that:
We proceed by induction on . The base case is provided by and . Assume the statement for for all . We have
where the last equality follows from the recurrence of given by Equation 5. Therefore, we have
This identity gives us the image of and under . We note that by the determinant of ,
(6)
Remark 2.4.
Observe that since , Equation 6 implies that and are units in . 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 . Let us identify the universal cover of with the upper-half space model and the ideal boundary of with . The action of on is given by . We say that is a cusp point of if its stabilizer in is conjugate to . In particular, the point is the cusp point of whose stabilizer in is the -subgroup generated by and acting as translations by and respectively. At every cusp point in , we choose a sufficiently small horoball neighborhood such that the stabilizer is a conjugate of and such that these horoball neighborhoods are invariant under the action of . We shall refer to the image of these horoball neighborhood as the cusp of . Similarly given any hyperplane in , a point in the circle/line at infinity of is called a cusp point of if it is a fixed point of a parabolic element in .
Let be a properly immersed cusped totally geodesic surface in . Since has one cusp and is complete, all cusps of must be contained in the cusp of . Given any cusp of , we consider a lift of to by putting the cusp point corresponding to at . The cusp cross-section of in is obtained by intersecting with the cusp cross-section of . This intersection lifts to a horocycle at in . The stabilizer of this horocycle is an element of the form in . In this case, we say that is a boundary slope of .
More generally, suppose that is a cusp point of a hyperplane in such that the parabolic element in the stabilizer of in fixing is conjugate to in . Then we say that is the boundary slope of at .
2.2 The thrice-punctured sphere
For all , contains an immersed thrice-punctured sphere as shown in Figure 1. By Adams [1, Theorem 3.1], is a totally geodesic surface. Moreover:
Proposition 2.5.
For all , the immersed surface is isotopic to a totally geodesic thrice-punctured sphere in whose fundamental group is generated by and
The set of boundary slopes of is .
-
Proof.
From the diagram, we see that the fundamental group of is generated by and:
Thus, we see that is conjugate to the subgroup generated by and as stated. Under the discrete faithful representation, we have
which is conjugate to the principal congruence subgroup of level 2 of .
Note that since both and are conjugate to , the surface has two slopes 1/0. The parabolic element corresponding to the remaining cusp is given by
Since and are conjugate to each other in , the slope of the remaining cusp is . ∎
Remark 2.6.
Let be the hyperplane in such that the line at infinity is the real line . The group stabilizes in . In fact, it is conjugate to the principal congruence subgroup of level 2 of by the matrix
Thus, elements of are precisely matrices in such that the lower diagonal entry is congruent to . This group is also known as the Hecke congruence subgroup of level .
We will also need the following fact about the action of on .
Lemma 2.7.
Consider the action of on . The cusp points of this action is , which is partitioned into 3 distinct orbits, namely
where are relatively prime.
-
Proof.
Let denote the set of cusp points in with respect to the action of . Since , we see that . The fundamental domain for the action is an ideal quadrilateral with vertices at and . There are three distinct orbits corresponding to , and . To see the description of these orbits, we first note that the integer matrices
map , , and to when is odd, , or , respectively. The determinants of these matrices can be chosen to be 1 because is relatively prime to when both is odd and is relatively prime to . By Remark 2.6, to be in (up to sign), it is sufficient to show that the lower diagonal entry of each matrix is congruent to . This is immediate for the first and the third matrices. For the second matrix, note that has to be an odd integer since is an even integer relatively prime to . Since is congruent to , the difference is congruent to . ∎
We conclude this section by describing certain lifts of in explicitly. The parametrization of these lifts play an important role in the proof of 1.3. We call a totally geodesic hyperplane in vertical if it contains in its circle/line at infinity. For convenience, we let and note that . Observe that up to the action of , the surface has three distinct vertical lifts in .
A direct computation shows that
Therefore, all vertical lifts of are orbits of , , and under the action of . Since the slope at (resp. ) on is (resp. ), the vertical hyperplane (resp. ) has slope (resp. ) at infinity,. Furthermore, we have
and
Thus when is odd, the line at infinity of goes through the point and has slope . Similarly, when is odd, the line at infinity of goes through the point and has slope . Figure 2 depicts the lines at infinity of the vertical lifts of .

Note that in Figure 2, the element acts by translating upward and acts by translating to the right. In other words, the real axis in 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 and drop 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 -manifolds.
Proposition 3.1.
Let be a non-cocompact Kleinian group of finite covolume and satisfying the following two conditions:
-
•
is of odd degree over and contains no proper real subfield other than .
-
•
has integral traces.
Then contains no cocompact Fuchsian groups and at most one commensurability class (up to conjugacy in ) 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 have an odd prime number of half-twists. Then does not contain closed immersed totally geodesic surfaces.
-
Proof.
Firstly, since is an odd prime [10, Theorem 1], the field has odd degree over and contains no proper subfield except . Secondly, the fact that implies that 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 . 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 not freely homotopic to must intersect 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 -manifold. We restate their lemma for dimension . A geodesic in a cusped finite-volume hyperbolic -manifold is called a cusp-to-cusp geodesic if it is the image of a geodesic in connecting two cusp points under the action of on .
Lemma 3.3.
Let be a complete finite volume hyperbolic -manifold with at least cusp. Suppose that and are two distinct properly immersed totally geodesic surfaces in such that is non-empty. Then 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 , 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 from all notations. Let be a cusped totally geodesic surface that is not freely homotopic to . Let be a hyperplane lift of to . Since is a cusped surface, the line/circle at infinity of contains a cusp point of . If is not a vertical hyperplane, we replace by a -translate that moves a cusp point of to infinity. Without loss of generality, we assume that is a vertical hyperplane.
Since is distinct from , we can see from Figure 2 that the line at infinity of must intersect the line at infinity of either or transversely. Therefore, intersects some vertical lift of along a geodesic for some . Since is a cusp point of , 3.3 implies that projects onto a cusp-to-cusp geodesic in . In particular, is a cusp point of .
Using the geodesic , we construct parabolic elements in fixing each end point and analyze the trace of the product of these elements. Since is a cusp point of , there exists such that . Since and are cusp points of , contains the nontrivial elements and fixing and , respectively, where . In other words, (resp. ) is the boundary slope of at (resp. ).
The following is a critical preliminary calculation in our analysis. We note that must contain only real algebraic integers in . The condition that does not contain any proper real subfield other than implies that . Suppose that
Then, dropping the for convenience of notation, we must have
which is true if and only if
(8) |
We shall refer to this as the trace condition, which gives us the following:
Lemma 3.5.
Let be the complement of a twist knot whose trace field is of odd degree over and contains no proper real subfield other than . Let be the totally geodesic thrice-punctured sphere as in Figure 1. Suppose that is an immersed cusped totally geodesic surface that is not freely homotopic . Then the set of boundary slopes of is precisely , up to multiplicity.
-
Proof.
It follows from the previous discussion that has a vertical lift to . The lift must intersect either or , whose respective slopes are and , along some vertical geodesic . One end point of is , so by 3.3, covers a cusp-to-cusp geodesic. This implies that the other end point of must also be a cusp point in or . Therefore, either for or for .
Suppose that the boundary slopes on are at and at the other end point of . As boundary slopes, neither nor is . We have nontrivial elements and in where has the property that or . We consider these two cases separately, and use the general form
Note that the end points of are distinct since is a geodesic in . Consequently, cannot fix , and so .
- Case 1:
-
Suppose that intersects nontrivially along a geodesic for some , as seen in Figure 3. In view of 2.7, we may choose to be either
according to whether belongs to , or , respectively. It suffices to compute in these three circumstances by using the trace condition Equation 8.
Since , we see that is of the form
This implies that in the last two situations. We show that this contradicts the assumption that . It follows from (7) that the trace field is the same as the cusp field. That is, . Then
so the minimal polynomial of over has degree . In particular, cannot satisfy the quadratic polynomial over given by the trace condition. Therefore, we have and since . Since intersects nontrivially, the boundary slope at on is not meridional; in other words, . We thus have , which implies that , and so . This contradicts the assumption that .
In the first situation, we get and so . The trace condition becomes
Since the minimal polynomial of has degree at least , satisfying the above is equivalent to satisfying
Again, since intersects nontrivially, we have , so and .
Figure 3: Vertical lifts of and when odd - Case 2:
-
Suppose that is parallel to , or equivalently for some . We may assume that . In view of 2.7, we may choose to be either
When , we get , so
This is satisfied if and only if because the minimal polynomial of is at least cubic over . If , then since endpoints of must be distinct, therefore . When , we must have , which is a contradiction.
When or , we see that and use an equivalent calculation to Case 1 to conclude that the boundary slopes are and .
∎
We now prove the uniqueness of .
-
Proof of 1.3.
Throughout the proof, we assume that the trace field of has odd degree and has no proper real subfield except for . In particular, for some positive odd integer . Since has integral traces by 3.2 and its trace field satisfies the condition in 3.1, any totally geodesic surface in must have at least one cusp.
3.5 has the following consequence. If is a cusped totally geodesic surface in that is not , then there exists a vertical lift in with slope at . This lift intersects along a cusp-to-cusp geodesic . A careful look at the first case in the proof of 3.5 shows that must belong to the orbit of under the action of on ; in particular, is an odd integer. Using the action of , we may assume that , since is distinct from . Let be the vertical lift of to of slope at that intersects the hyperplane along the geodesic .
In order to prove 1.3, it suffices to show that each vertical hyperplane for with odd covers an infinite area totally geodesic surface, which implies that must be dense in . With this density in mind, we construct a non-closed non-cusp-to-cusp geodesic in as a self-intersection of .
Fix a rational number where odd and , relatively prime. Suppose that covers a complete properly immersed finite area totally geodesic surface. We move by an element of that takes to . To write down such an element, we first map to by an element of , then map to by . Consider the following element
taking to . Note that the latter matrix in the product belongs to as discussed in Remark 2.6. Therefore, . Since , the hyperplane must be vertical.
The analysis in the first case of 3.5 shows that the slope of at is . Consequently, the slope of is at is . Since has slope , the vertical hyperplane intersects along a geodesic for some . We have
Since the slope of at is , the line at infinity of is parametrized by
for some while that of is parametrized by
for some . It follows that
(9) Since one end point of is at , a cusp point of , Remark 3.4 implies that covers a cusp-to-cusp geodesic. In particular, is a cusp point of . Thus, there exists
such that , see Remark 2.2. We must have
(10) By 3.5, the slope at of is either or . We consider these two cases separately.
- Case 1:
-
Suppose that the slope at is . Consequently, contains . The trace condition Equation 8 implies that
Let us write for some . Note that as contains no trace element. Since , we have . Using Equation 6, we have or . Since is odd by [10, Theorem 1], does not contain the square root of non-square integer. It follows that . In other words, for some . Substituting this and Equation 9 into Equation 10, we have
The integer must divide because and are relatively prime. Since is a unit in , we have the following containment of ideals and . In particular, the ideal generated by and in is contained in . Since , . This fact implies that . We obtain the final expressions for and .
(11) The slope of at is and , so is a vertical hyperplane with slope at . Thus, this hyperplane intersects along a vertical geodesic . By 3.3, covers a cusp-to-cusp geodesic in , since one end point of is at . Therefore, the other end point of must be a cusp point of .
The boundary of is parametrized by for , and the boundary of is parametrized by for some . Since cusp points of are contained in , there exist such that
Therefore, because has no proper real subfield except . Using Equation 6, we see that
Since and is a unit of , the right-hand side of the equation must belong to . In particular, and are integers. To simplify our notation, we write where . Now we substitute the expression for , , and (see Equation 11) into the determinant of to obtain
Since , we can write the equation above as
for . Noting that , the equality above implies that
Both and are units in , so the first and the third equations imply that
This contradicts the second equation. Therefore, there exists no such in this case.
- Case 2:
-
Suppose that the slope at is . It follows that contains . The trace condition Equation 8, implies that
Let us write for some . Since , we must have . Now we write . Note that since does not contain element with trace . Since is odd by [10, Theorem 1], the trace field does not contain the square root of any non-square integer. It follows that . In other words, for some . Substituting this into Equation 10, we have
Since , we see that . Since and are relatively prime, divides . We have the following containment of ideals and . In particular, . Since , . This implies that . We obtain the following expressions for and
(12) Since the slope of at is and , we observe that is a vertical plane with slope at . It must intersect the hyperplane at a vertical geodesic . By 3.3, covers a cusp-to-cusp geodesic in , since one end point of is . The other end point of is a cusp point on the line at infinity of and so must be a rational number. In other words, there exists a real number such that
This implies that since has no proper real subfield except . Therefore . Now we substitute the expression for , and (see Equation 12) into the determinant of to obtain
Since , we can write the equation above as
Since , the equality above implies that
Since is a unit in , the third equation implies that . Given that and , the second equation implies that . These conditions on contradict the first equation. Consequently, there exists no such in this case.
The analysis of both cases shows that is not a cusp point of because there is no such that . This contradiction implies that 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 have an odd prime number of twists. Note that because has integral traces, has no proper real subfield (except ) because it has odd prime degree over by [10, Theorem 1], and is a Fuchsian subgroup. Thus the conditions of 1.3 are satisfied.
Using Magma [5] and SageMath [20], we verify that also satisfies the conditions of 1.3 for odd integers with , see [19] for the code.
Twist knot complements with distinct numbers of half-twists are always non-commensurable because their invariant trace fields are pairwise distinct. Thus we have a family of infinitely many non-commensurable -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 -manifold with exactly totally geodesic surfaces, it suffices to find covers of such that has exactly lifts for some odd prime . We construct these manifolds from dihedral covers of the twist knot complements. We first count the number of lifts of to dihedral covers of twist knot with an odd number of half-twists. For this purpose, let be a positive odd integer.
We write down an explicit surjective homomorphism from to the dihedral group of order . The group has the following presentation
Let us fix an integer dividing . We have a surjective homomorphism defined by and since . Let be the cover corresponding to .
We observe that
since has order 2. Therefore, is a cyclic subgroup of order 2. Let be a lift of to corresponding to . Since is a regular cover, the deck group acts transitively on the lifts of to . We identify these lifts with the orbit of under the action of . Since the stabilizer of under the action of is , we identify the lifts of to with the left cosets of in . In particular, lifts to distinct totally geodesic surfaces in .
Observe that the cyclic subgroup of order forms a complete left coset representatives of in . Up to reordering, we identify with the coset for . The action of on the lifts is given by
where , since . Since is an odd integer, leaves invariant exactly one surface and permutes surfaces pairwise. It follows that contains exactly lifts of .
Any integer can be written as where . To prove the theorem, it suffices to show that for any odd integer there exist infinitely many primes such that divides . Let and thus . By Dirichlet’s theorem, the arithmetic progression contains infinitely many primes . Note that divides for all . 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 .
Proposition 3.6.
Let be a twist knot complement with at least half-twists. Let be the vertical hyperplane intersecting at with slope at . For all odd , the stabilizer of in is non-elementary.
-
Proof.
We observe that contains a subgroup generated by
where . The fact that the latter element belongs to follows from Remark 2.6. It is clear that stabilizes , so the line at infinity of has slope and thus is parametrized by
for . We have
for any . Since , the image point also belongs to the line at infinity of . Therefore, the latter element stabilizes .
These are independent parabolic element with respect to the action of the stabilizer of on because they have distinct fixed points. They thus generate a non-elementary subgroup of the stabilizer of . ∎
Since all vertical lifts of are described at the end of Section 2, for and odd , the image of in is not freely homotopic to . In particular, and are not conjugate in . By [3, Theorem 1.1], at most finitely many -orbits of hyperplanes cover a finite area totally geodesic surface. Other quotients of these hyperplanes must be of infinite area. Furthermore, if the trace field of satisfies the condition in 1.3, then 1.3 says that is the unique (up to conjugacy) finite covolume maximal Fuchsian subgroup of ; that is,
Corollary 3.7.
Let be an odd integer such that does not contain any proper real subfield. Then for all and odd , is a non-elementary Fuchsian group of infinite covolume.
In other words, 3.7 provides examples of infinitely many hyperbolic -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 -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 . Being right-angled is ever-present in -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 -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 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 odd prime, we check that the right-angled property fails by finding the angle between the real hyperplane through and the hyperplane of slope through . This angle is if and only if the argument of is .
Figure 4: Vertical lifts of for odd along with the angle that must be
We know that , so , whose argument is bounded away from 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 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 . The other is a dual surface whose boundary slopes are and ; this information can be calculated using the techniques in Section 2 and visualized in Figure 6.



The occurrence of the boundary slope for 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 is quadratic. Moreover,
Theorem 4.1.
Every number in occurs as a boundary slope for some totally geodesic surface in the figure-8 knot complement.
-
Proof.
Let be a parabolic element fixing . We can find such that the trace condition Equation 8 is satisfied. As computed in the proof of 3.5, we must satisfy
Since the minimal polynomial of is , satisfying the above is equivalent to satisfying
For any pair , there exists such a pair . Thus the group generated by and is a non-elementary Fuchsian subgroup. Because the figure-8 knot is arithmetic, every non-elementary Fuchsian subgroup in sits inside a finite coarea arithmetic Fuchsian subgroup of [14, Theorem 4]. Thus there exists such that is the stabilizer of a hyperplane which corresponds to a totally geodesic surface in the figure-8 knot complement admitting boundary slope . ∎
5 Computer experiments and questions
5.1 Trace fields of twist knots
The condition that the trace field of for an odd prime 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 , the trace field has no proper subfield except for , see [19] for the code. The following question arises naturally:
Question 5.1.
Is the unique proper subfield of the trace field of ?
We also checked that the Galois group of the Riley polynomial is congruent to , the full symmetric group on letters for , see [19]. This implies that there are no proper subfields between and since a subgroup of index in is necessarily maximal. Therefore, it is natural to ask:
Question 5.2.
Is for all ?
We conjecture that the answer is yes to both questions. By 1.3, an affirmative answer to either question implies that is the unique totally geodesic surface in all for odd .
Finally, using Magma we compute the class number of and observe that it is for all , see [19]. As a consequence of the argument in the proof of 1.3, the cusp points of form a proper subset of the trace field . In other words, for odd integer , the group gives an example of a Kleinian group whose trace field has class number 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 is for all . However, the class number of is computed to be , see [19].
5.2 Homological behavior of in congruence covers
We study the homological behavior of totally geodesic surfaces in congruence covers of . 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 of dimension , the Lefschetz Hyperplane Theorem says that the map
is injective for all and for a generic hyperplane . Replacing ”hyperplane section” by ”totally geodesic surface in hyperbolic -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 -manifold.
To describe our question, we need to set up some notations. Let be the ring of integers of and identify with its image in . Let be the principal congruence cover of level for some ideal and be the cover of that corresponds to . Let be the lifts of in .
Question 5.3.
What are the possible behaviors of the map
as varies over the prime ideals in ?
We first investigate this question when is a prime with residue degree , i.e. when , a finite field of prime order . Suppose that is a prime ideal in lying above an odd prime such that . In [18, Proposition 4], Riley showed that the representation coming from the reduction map is surjective for all odd prime . For all odd prime , maps onto the group
under the reduction map Therefore for any prime ideal with residue degree lying above an odd prime , contains exactly one unique lift of the totally geodesic thrice-puncture sphere, and this cover has fundamental group . For all examples that we computed with Magma, we observed that maps onto under the induced map for prime ideals of residue degree lying above an odd prime . In particular, we verified that surjects for that has a prime ideal of residue degree lying above and , see [19].
Question 5.4.
Is the induced map surjective for all prime ideals of residue degree 1 lying above odd prime ?
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
by where are the invariant factors. Furthermore, repetition of a prime indicates a prime with different prime ideals lying above of residue degree 1. See , see [19] for the code used to produce the table.
3 | |||
5 | |||
5 |
5.3 Finding surfaces in the figure-8 knot complement
Given the linear relation between and 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 occur simultaneously as boundary slopes for some totally geodesic surface of ?
We give some computed examples of totally geodesic surfaces in below. We identify these by specifying a vertical hyperplane in and its stabilizer. These hyperplanes have the form where is the point of intersection with and is the slope of the vertical plane at infinity. We denote the stabilizers of these hyperplanes at infinity by .
Example 5.6.
The plane is and has stabilizer
This is the thrice-punctured sphere .
Example 5.7.
The plane has stabilizer
Note that
This is the thrice-punctured sphere .
Both and admit precisely two boundary slopes, but that is not necessarily true for all totally geodesic surfaces:
Example 5.8.
The plane has stabilizer
The set of boundary slopes of contains but may not be limited to .
Example 5.9.
The plane has stabilizer
The set of boundary slopes of contains but may not be limited to .
Given the existence of sets of boundary slopes that are not precisely two numbers in , we ask:
Question 5.10.
Does there exist a totally geodesic surface of admitting exactly boundary slope? arbitrarily many boundary slopes?
References
- [1] C. Adams, Thrice-punctured spheres in hyperbolic -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 -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 -manifolds containing infinitely many totally geodesic surfaces, Ergodic Theory and Dynamical Systems, (2021) 1–32.
- [13] C. Maclachlan, A. Reid, The Arithmetic of Hyperbolic -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 -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],