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

Turaev–Viro invariants and cabling operations

Sanjay Kumar and Joseph M. Melby
Abstract

In this paper, we study the variation of the Turaev–Viro invariants for 33-manifolds with toroidal boundary under the operation of attaching a (p,q)(p,q)-cable space. We apply our results to a conjecture of Chen and Yang which relates the asymptotics of the Turaev–Viro invariants to the simplicial volume of a compact oriented 33-manifold. For pp and qq coprime, we show that the Chen–Yang volume conjecture is stable under (p,q)\left(p,q\right)-cabling. We achieve our results by studying the linear operator RTrRT_{r} associated to the torus knot cable spaces by the Reshetikhin–Turaev SO3SO_{3}-Topological Quantum Field Theory (TQFT), where the TQFT is well-known to be closely related to the desired Turaev–Viro invariants. In particular, our utilized method relies on the invertibility of the linear operator for which we provide necessary and sufficient conditions.

1 Introduction

For a compact 3-manifold MM, its Turaev–Viro invariants are a family of \mathbb{R}-valued homeomorphism invariants parameterized by an integer r3r\geq 3 depending on a 2r2r-th root of unity qq. We are primarily interested in the invariants when rr is odd and q=e2π1rq=e^{\frac{2\pi\sqrt{-1}}{r}}.

In this paper, we study the variation of the Turaev–Viro invariants of a 33-manifold with toroidal boundary when we attach a (p,q)(p,q)-cable space.

Definition 1.1.

Let VV be the standardly embedded solid torus in S3S^{3}, and let VV^{\prime} be a closed neighborhood of VV. For pp and qq coprime integers with q>0q>0, let Tp,qVT_{p,q}\subset\partial V be the torus knot of slope p/qp/q. The (p,q)(p,q)-cable space, denoted Cp,qC_{p,q}, is the complement of the torus knot Tp,qT_{p,q} in VV^{\prime}. Let MM be a 33-manifold with toroidal boundary. A manifold MM^{\prime} obtained from gluing a (p,q)(p,q)-cable space Cp,qC_{p,q} to a boundary component of MM along the exterior toroidal boundary component of Cp,qC_{p,q} is called a (p,q)(p,q)-cable of MM.

Our main theorem is the following.

Theorem 1.2.

Let MM be a manifold with toroidal boundary, let p,qp,q be coprime integers with q>0q>0, and let r3r\geq 3 be an odd integer coprime to qq. Suppose MM^{\prime} is a (p,q)(p,q)-cable of MM. Then there exists a constant C>0C>0 and natural number NN such that

1CrNTVr(M)TVr(M)CrNTVr(M).\frac{1}{Cr^{N}}TV_{r}(M)\leq TV_{r}(M^{\prime})\leq Cr^{N}TV_{r}(M).

Theorem 1.2 has notable applications to existing conjectures. In general, the Turaev–Viro invariants are difficult to compute; however, there is interest in the relationship between their rr-asymptotic behavior and classical invariants of 3-manifolds. Chen and Yang [5] conjectured that the growth rate of the Turaev–Viro invariants for hyperbolic manifolds is related to the manifold’s hyperbolic volume. They also provided computational evidence for the conjecture in [5].

This conjecture should be compared to the well-known conjectures of Kashaev [12] and Murakami-Murakami [16] relating the Kashaev and colored Jones invariants of hyperbolic link complements to their hyperbolic volumes. Detcherry and Kalfagianni [7] restated the Turaev–Viro invariant volume conjecture more generally in terms of the simplicial volume for manifolds which are not necessarily hyperbolic. In order to state the conjecture, we will first introduce a slightly weaker condition for the growth rate of the Turaev–Viro invariants.

Definition 1.3.

Define the following two asymptotics of the Turaev–Viro invariants for compact 3-manifolds as

lTV(M):=lim infr, r odd2πr\displaystyle\textit{lTV}(M):=\liminf_{r\rightarrow\infty,\text{ }r\text{ odd}}\frac{2\pi}{r} log|TVr(M;q=e2πir)|,\displaystyle\log\left|\text{TV}_{r}\left(M;q=e^{\frac{2\pi i}{r}}\right)\right|,

and

LTV(M):=lim supr, r odd2πr\displaystyle\textit{LTV}(M):=\limsup_{r\rightarrow\infty,\text{ }r\text{ odd}}\frac{2\pi}{r} log|TVr(M;q=e2πir)|.\displaystyle\log\left|\text{TV}_{r}\left(M;q=e^{\frac{2\pi i}{r}}\right)\right|.

Additionally, we will introduce the simplicial volume for compact orientable 33-manifolds with empty or toroidal boundary, originally defined by Gromov [9]. For MM a compact, orientable, irreducible 33-manifold, there is a unique collection of incompressible tori, up to isotopy, along which MM can be decomposed into atoroidal manifolds. This is known as the JSJ decomposition [10, 11]. By Thurston’s Geometrization Conjecture [23], famously completed in the work of Perelman [17, 18, 19], each of these atoroidal manifolds are either hyperbolic or Seifert-fibered, and by Thurston [22], the simplicial volume of MM coincides with the sum of the simplicial volumes of the resulting pieces. In the case where MM is hyperbolic, the simplicial volume is positive and is related to the hyperbolic volume of MM by

vol(M)=v3Mvol(M)=v_{3}\|M\|

where vol(M)vol(M) is the hyperbolic volume of MM, v31.0149v_{3}\approx 1.0149 is the volume of the regular ideal tetrahedron, and M\|M\| is the simplicial volume of MM.

This leads to a natural extension of Chen and Yang’s Turaev–Viro invariant volume conjecture [5].

Conjecture 1.4 ([5], [7]).

Let MM be a compact oriented 3-manifold. Then

LTV(M)=v3M,\text{LTV}(M)=v_{3}||M||,

where v3v_{3} is the volume of the regular ideal tetrahedron and \|\cdot\| is the simplicial volume.

If Conjecture 1.4 is true, since the simplicial volume is additive under the JSJ decomposition, it would imply that the asymptotics of the Turaev–Viro invariants are also additive under the decomposition. As an application of Theorem 1.2, we provide evidence of the additivity for the asymptotics of the Turaev–Viro invariants.

The asymptotic additivity property has been explored in several works. For a manifold MM which satisfies Conjecture 1.4, the property was proven for invertible cablings of MM by Detcherry and Kalfagianni [7], the figure-eight knot cabled with Whitehead chains by Wong [26], and an infinite family of manifolds with arbitrarily large simplicial volume by the authors of this paper [14]. Additionally, the property was proven to hold under the operation of attaching a (p,2)(p,2)-cable space by Detcherry [6], which we extend in this work.

A key property of the (p,q)(p,q)-cable spaces is that they have simplicial volume zero. Theorem 1.2 provides a way to construct new manifolds without changing the simplicial volume while controlling the growth of the Turaev–Viro invariants. This leads to many examples of manifolds satisfying Conjecture 1.4.

To the authors’ knowledge, in all of the proven examples of Conjecture 1.4, the stronger condition that the limit approaches the simplicial volume is verified, as opposed to only the limit superior. Theorem 1.2 implies the following corollaries. See Section 3 for more details.

Corollary 1.5.

Suppose MM satisfies Conjecture 1.4 and lTV(M)=v3MlTV(M)=v_{3}||M||. Then for any pp and qq coprime, any (p,q)(p,q)-cable MM^{\prime} also satisfies Conjecture 1.4.

Some examples which satisfy the hypothesis of Corollary 1.5 include the figure-eight knot and the Borromean rings by Detcherry-Kalfagianni-Yang [8], the Whitehead chains by Wong [25], the fundamental shadow links by Belletti-Detcherry-Kalfagianni-Yang [2], a family of hyperbolic links in S2×S1S^{2}\times S^{1} by Belletti [1], a large family of octahedral links in S3S^{3} by the first author of this paper [13], and a family of link complements in trivial S1S^{1}-bundles over oriented connected closed surfaces by the authors of this paper [14].

For general pp and qq coprime, Corollary 1.5 demonstrates the stability of Conjecture 1.4 under the (p,q)(p,q)-cabling operation. However, in the case when q=2nq=2^{n} for nn\in\mathbb{N}, we recover the full limit as shown in the following corollary.

Corollary 1.6.

Suppose MM satisfies Conjecture 1.4. Then for any odd pp and nn\in\mathbb{N}, any (p,2n)(p,2^{n})-cable MM^{\prime} also satisfies Conjecture 1.4. Moreover, if lTV(M)=v3MlTV(M)=v_{3}||M||, then lTV(M)=LTV(M)=v3MlTV(M^{\prime})=LTV(M^{\prime})=v_{3}||M^{\prime}||.

As a direct result of Corollaries 1.5 and 1.6, we extend the work of Detcherry [6], where the author considers the operation of attaching a (p,2)(p,2)-cable space. This allows us to construct manifolds satisfying Conjecture 1.4 from manifolds with toroidal boundary where the conjecture is already known. This includes all previously mentioned examples.

The general method of proof for Theorem 1.2 follows from the work of Detcherry [6]. Considering the cable space Cp,qC_{p,q} as a cobordism between tori, the Reshetikhin–Turaev SO3SO_{3}-TQFT at level rr, denoted by RTrRT_{r}, associates to it a linear operator

RTr(Cp,q):RTr(T2)RTr(T2).RT_{r}(C_{p,q}):RT_{r}(T^{2})\rightarrow RT_{r}(T^{2}).

For pp odd and q=2q=2, Detcherry presents RTr(Cp,2)RT_{r}(C_{p,2}) using the basis {e1,e3,,e2m1}\{e_{1},e_{3},\dots,e_{2m-1}\}, which is equivalent to the orthonormal basis {e1,e2,,em}\{e_{1},e_{2},\dots,e_{m}\} for RTr(T2)RT_{r}(T^{2}) given in [4] under the symmetry emi=em+i+1e_{m-i}=e_{m+i+1} for 0im10\leq i\leq m-1. More details of the construction are given in Section 2. With this basis, RTr(Cp,2)RT_{r}(C_{p,2}) can be presented as a product of two diagonal matrices and one triangular matrix. This allows the author to directly write the inverse of RTr(Cp,2)RT_{r}(C_{p,2}). From the inverse of this linear operator, Detcherry establishes a lower bound of the Turaev–Viro invariants under attaching a (p,2)(p,2)-cable space.

For general qq, RTr(Cp,q)RT_{r}(C_{p,q}) does not have as simple a presentation under the same basis, making it more difficult to conclude that RTr(Cp,q)RT_{r}(C_{p,q}) is invertible. In order to resolve this, we present RTr(Cp,q)RT_{r}(C_{p,q}) using a different basis for RTr(T2)RT_{r}(T^{2}), defined in Section 4, that allows us to also show directly that RTr(Cp,q)RT_{r}(C_{p,q}) is invertible provided rr and qq are coprime. Following Detcherry’s argument, the invertibility of RTr(Cp,q)RT_{r}(C_{p,q}) is integral in finding the lower bound from Theorem 1.2; however, the invertibility of RTr(Cp,q)RT_{r}(C_{p,q}) is constrained by the condition that rr and qq are coprime, as outlined by Theorem 1.7.

Theorem 1.7.

Let pp be coprime to some positive integer qq. Then RTr(Cp,q)RT_{r}(C_{p,q}) is invertible if and only if rr and qq are coprime. Moreover, the operator norm |RTr(Cp,q)1||||RT_{r}(C_{p,q})^{-1}||| grows at most polynomially.

As we will show in Section 3, the coprime condition between rr and qq leads to the discrepancy between recovering the limit superior in Corollary 1.5 versus the full limit in Corollary 1.6.

The paper is organized as follows: We recall properties of the Reshetikhin–Turaev SO3SO_{3}-TQFTs, the RTrRT_{r} torus knot cabling formula, and relevant properties of the Turaev–Viro invariants in Section 2. In Section 3, we prove Theorem 1.2 assuming Theorem 1.7. In Section 4, the construction of the relevant basis for RTr(T2)RT_{r}(T^{2}) and the proof of Theorem 1.7 is given. Lastly, we consider future directions in Section 5.

Acknowledgements.

The authors would like to thank their advisor Efstratia Kalfagianni for guidance and helpful discussions. Additionally, we would like to thank Renaud Detcherry for his comments and suggestions on an early version of this paper. Finally, we thank the reviewer for their time and careful reading as well as their valuable suggestions for improving the overall clarity of the arguments in Section 4. Part of this research was supported in the form of graduate Research Assistantships by NSF Grants DMS-1708249 and DMS-2004155. For the second author, this research was partially supported by a Herbert T. Graham Scholarship through the Department of Mathematics at Michigan State University.

2 Preliminaries

2.1 Reshetikhin–Turaev TQFT

In this subsection, we outline relevant properties of the Reshetikhin–Turaev SO3SO_{3}-TQFTs, which were defined by Reshetikhin and Turaev in [20]. Let 𝔬𝔟\mathfrak{Cob} be the category of (2+1)(2+1)-dimensional cobordisms and Vect()Vect(\mathbb{C}) be the category of \mathbb{C}-vector spaces. For an odd integer r3r\geq 3 and primitive 2r2r-th root of unity AA, one associates a (2+1)(2+1)-dimensional TQFT RTr:𝔬𝔟Vect()RT_{r}:\mathfrak{Cob}\rightarrow Vect(\mathbb{C}). Blanchet, Habegger, Masbaum, and Vogel [4] gave a skein-theoretic framework for this SO3SO_{3}-TQFT, and its main properties are the following:

  1. 1)

    For a closed oriented surface Σ\Sigma, RTr(Σ)RT_{r}(\Sigma) is a finite dimensional vector space over \mathbb{C} with the natural Hermitian form. For a disjoint union ΣΣ\Sigma\sqcup\Sigma^{\prime}, we have RTr(ΣΣ)=RTr(Σ)RTr(Σ)RT_{r}(\Sigma\sqcup\Sigma^{\prime})=RT_{r}(\Sigma)\otimes RT_{r}(\Sigma^{\prime}).

  2. 2)

    For an oriented closed 3-manifold MM, RTr(M)RT_{r}(M)\in\mathbb{C} is a topological invariant.

  3. 3)

    For an oriented compact 3-manifold MM with boundary M\partial M, RTr(M)RTr(M)RT_{r}(M)\in RT_{r}(\partial M) is a vector.

  4. 4)

    For a cobordism (M,Σ1,Σ2)(M,\Sigma_{1},\Sigma_{2}), RTr(M):RTr(Σ1)RTr(Σ2)RT_{r}(M):RT_{r}(\Sigma_{1})\rightarrow RT_{r}(\Sigma_{2}) is a linear map.

In [4], the authors also give explicit bases for any surface. However, we will focus on RTr(T2)RT_{r}(T^{2}), which can be considered as a quotient of the Kauffman skein module of the genus 11 handlebody D2×S1D^{2}\times S^{1}.

We begin by coloring the core {0}×S1\{0\}\times S^{1} by the (i1)(i-1)-th Jones-Wenzl idempotent. This gives a family of elements eie_{i} of the Kauffman skein module of the solid torus. However, there are only finitely many Jones-Wenzl idempotents for a given odd r=2m+1r=2m+1 and 2r2r-th root of unity AA, namely e1,,e2m1e_{1},\dots,e_{2m-1} [4]. We can consider these eie_{i}’s as elements of the quotient RTr(T2)RT_{r}(T^{2}), giving us a basis.

Theorem 2.1 ([4], Theorem 4.10).

Let r=2m+13r=2m+1\geq 3. Then the family e1,,eme_{1},\dots,e_{m} is an orthonormal basis for RTr(T2)RT_{r}(T^{2}). Moreover, the relation emi=em+1+ie_{m-i}=e_{m+1+i} holds for 0im10\leq i\leq m-1.

The second part of the theorem implies that {e1,e3,,e2m1}\{e_{1},e_{3},\dots,e_{2m-1}\} is just a reordering of the basis {e1,,em}\{e_{1},\dots,e_{m}\}.

2.2 The Cabling Formula

Here, we will give an explicit description for the Reshetikhin–Turaev invariants of the torus knot cable spaces.

Let pp and qq be coprime integers where q>0q>0, and let Cp,qC_{p,q} be the (p,q)(p,q)-cable space Cp,qC_{p,q}. These spaces are Seifert-fibered and therefore have simplicial volume zero. For r=2m+13r=2m+1\geq 3, we extend the vectors eiRTr(T2)e_{i}\in RT_{r}(T^{2}) to all ii\in\mathbb{Z} in the following way. Let ei=eie_{-i}=-e_{i} for any i0i\geq 0, and let ei+kr=(1)keie_{i+kr}=(-1)^{k}e_{i} for any kk\in\mathbb{Z}. Note this means that er=e0=0e_{r}=e_{0}=0.

Regarding the cable space Cp,qC_{p,q} as a cobordism between tori, the Reshetikhin–Turaev SO3SO_{3}-TQFT gives a linear map

RTr(Cp,q):RTr(T2)RTr(T2).RT_{r}(C_{p,q}):RT_{r}(T^{2})\rightarrow RT_{r}(T^{2}).

The map RTr(Cp,q)RT_{r}(C_{p,q}) sends the element eie_{i} to the element of RTr(T2)RT_{r}(T^{2}) corresponding to a (p,q)(p,q)-torus knot embedded in the solid torus and colored by the (i1)(i-1)-th Jones-Wenzl idempotent. Morton [15] gives the following formula for the image of the basis elements under RTr(Cp,q)RT_{r}(C_{p,q}).

Theorem 2.2 ([15], Section 3, Cabling Formula).
RTr(Cp,q)(ei)=Apq(i21)/2kSiA2pk(qk+1)e2qk+1,RT_{r}(C_{p,q})(e_{i})=A^{pq(i^{2}-1)/2}\sum_{k\in S_{i}}A^{-2pk(qk+1)}e_{2qk+1},

where SiS_{i} is the set

Si={i12,i32,,i32,i12}.S_{i}=\{-\frac{i-1}{2},-\frac{i-3}{2},...,\frac{i-3}{2},\frac{i-1}{2}\}.

As we will see in Subsection 2.3, the Reshetikhin–Turaev TQFT is closely related to the Turaev–Viro invariants for 33-manifolds. By using their relationship, the explicit formula given in Theorem 2.2 will allow us to obtain a lower bound on the Turaev–Viro invariants under the cabling operation.

2.3 Properties of the Turaev–Viro invariants

In this subsection, we discuss properties of the Turaev–Viro invariants [24] as well as an important characterization in terms of the Reshetikhin–Turaev invariants.

The Turaev–Viro invariants were defined by Turaev and Viro [24] in terms of state sums over triangulations of a 3-manifold MM, but they are also closely related to the Reshetikhin–Turaev invariants. The following identity was originally proven for closed 3-manifolds by Roberts [21] and then extended to compact manifolds with boundary by Benedetti and Petronio [3].

Theorem 2.3 ([3, 21]).

Let r3r\geq 3 be an odd integer, and let qq be a primitive 2r2r-th root of unity. Then for a compact oriented manifold MM with toroidal boundary,

TVr(M;q)=RTr(M;q12)2TV_{r}\left(M;q\right)=\left\|RT_{r}\left(M;q^{\frac{1}{2}}\right)\right\|^{2}

where \|\cdot\| is the natural Hermitian norm on RTr(M).RT_{r}\left(\partial M\right).

We note that this identity holds more generally, but we have restricted to manifolds with toroidal boundary for simplicity.

In [7], Detcherry and Kalfagianni proved that the growth rate of the Turaev–Viro invariants has properties reminiscent of simplicial volume. We summarize their results in the following theorem.

Theorem 2.4 ([7]).

Let MM be a compact oriented 33-manifold, with empty or toroidal boundary.

  1. 1)

    If MM is a Seifert manifold, then there exist constants B>0B>0 and NN such that for any odd r3r\geq 3, we have TVr(M)BrNTV_{r}(M)\leq Br^{N} and LTV(M)0LTV(M)\leq 0.

  2. 2)

    If MM is a Dehn-filling of MM^{\prime}, then TVr(M)TVr(M)TV_{r}(M)\leq TV_{r}(M^{\prime}) and LTV(M)LTV(M)LTV(M)\leq LTV(M^{\prime}).

  3. 3)

    If M=M1𝑇M2M=M_{1}\underset{T}{\bigcup}M_{2} is obtained by gluing two 33-manifolds along a torus boundary component, then TVr(M)TVr(M1)TVr(M2)TV_{r}(M)\leq TV_{r}(M_{1})TV_{r}(M_{2}) and LTV(M)LTV(M1)+LTV(M2)LTV(M)\leq LTV(M_{1})+LTV(M_{2}).

3 Bounding the Invariant Under Cabling

In this section, we will prove Theorem 1.2 with the assumption of a key theorem, and we reserve the technical details for Section 4. We remark that the major components of our argument follow from the work of Detcherry [6] where the case when pp is odd and q=2q=2 was proven. For convenience, we will restate the main theorem.

Theorem 1.2.

Let MM be a manifold with toroidal boundary, let p,qp,q be coprime integers with q>0q>0, and let r3r\geq 3 be an odd integer coprime to qq. Suppose MM^{\prime} is a (p,q)(p,q)-cable of MM. Then there exists a constant C>0C>0 and natural number NN such that

1CrNTVr(M)TVr(M)CrNTVr(M).\frac{1}{Cr^{N}}TV_{r}(M)\leq TV_{r}(M^{\prime})\leq Cr^{N}TV_{r}(M).

We will now assume Theorem 1.7, which we also restate for convenience.

Theorem 1.7.

Let pp be coprime to some positive integer qq. Then RTr(Cp,q)RT_{r}(C_{p,q}) is invertible if and only if rr and qq are coprime. Moreover, the operator norm |RTr(Cp,q)1||||RT_{r}(C_{p,q})^{-1}||| grows at most polynomially.

Proof of Theorem 1.2.

As mentioned previously, the case when pp is odd and q=2q=2 was shown by Detcherry [6], and our approach follows closely in structure. We let MM be a manifold with toroidal boundary, pp an integer, q>0q>0 an integer coprime to pp, r3r\geq 3 odd and coprime to qq, and MM^{\prime} a (p,q)(p,q)-cable of MM. We will proceed to prove Theorem 1.2 by showing the upper inequality of

1CrNTVr(M)TVr(M)CrNTVr(M)\frac{1}{Cr^{N}}TV_{r}(M)\leq TV_{r}(M^{\prime})\leq Cr^{N}TV_{r}(M)

followed by the lower inequality, where C>0C>0 and NN\in\mathbb{N}. To obtain the upper inequality, we first remark that M=Cp,q𝑇MM^{\prime}=C_{p,q}\underset{T}{\bigcup}M. By Theorem 2.4, this implies that

TVr(M)TVr(Cp,q)TVr(M).TV_{r}(M^{\prime})\leq TV_{r}(C_{p,q})TV_{r}(M).

Since Cp,qC_{p,q} is a Seifert manifold, we have that

TVr(Cp,q)C1rN1TV_{r}(C_{p,q})\leq C_{1}r^{N_{1}}

for some C1>0C_{1}>0 and N1N_{1}\in\mathbb{N} also by Theorem 2.4. This leads to the upper inequality

TVr(M)C1rN1TVr(M).TV_{r}(M^{\prime})\leq C_{1}r^{N_{1}}TV_{r}(M).

For the lower inequality, we will use Theorem 1.7. From the properties of the Reshetikhin–Turaev SO3SO_{3}-TQFT, we consider the linear map

RTr(Cp,q):RTr(T2)RTr(T2).RT_{r}(C_{p,q}):RT_{r}(T^{2})\rightarrow RT_{r}(T^{2}).

If MM only has one boundary component, then

RTr(M)=RTr(Cp,q)RTr(M)RT_{r}(M^{\prime})=RT_{r}(C_{p,q})RT_{r}(M)

by the properties of a TQFT. If MM has other boundary components, then the invariant associated to any coloring ii of the other boundary components may be computed as

RTr(M,i)=RTr(Cp,q)RTr(M,i).RT_{r}(M^{\prime},i)=RT_{r}(C_{p,q})RT_{r}(M,i).

By the invertibility of RTr(Cp,q)RT_{r}(C_{p,q}) from Theorem 1.7, we have the inequality

RTr(M)|RTr(Cp,q)1|RTr(M)||RT_{r}(M)||\leq|||{RT_{r}(C_{p,q})^{-1}}|||\cdot||RT_{r}(M^{\prime})||

where ||||||\cdot|| is the norm induced by the Hermitian form of the TQFT and |||||||||\cdot||| is the operator norm. Since the operator norm grows at most polynomially by Theorem 1.7, we obtain the inequality

1C2rN2RTr(M)RTr(M)\frac{1}{C_{2}r^{N_{2}}}||RT_{r}(M)||\leq||RT_{r}(M^{\prime})||

for some C2>0C_{2}>0 and N2N_{2}\in\mathbb{N}. Lastly, by Theorem 2.3, the norm of the Reshetikhin–Turaev invariant is related to the Turaev–Viro invariant such that we arrive to the desired inequality

1C3rN3TVr(M)TVr(M)\frac{1}{C_{3}r^{N_{3}}}TV_{r}(M)\leq TV_{r}(M^{\prime})

for some C3>0C_{3}>0 and N3N_{3}\in\mathbb{N}. This leads to

1CrNTVr(M)TVr(M)CrNTVr(M)\frac{1}{Cr^{N}}TV_{r}(M)\leq TV_{r}(M^{\prime})\leq Cr^{N}TV_{r}(M)

where C>0C>0 and NN\in\mathbb{N}. ∎

As discussed in Section 1, the following corollaries follow from Theorem 1.2.

Corollary 1.5.

Suppose MM satisfies Conjecture 1.4 and lTV(M)=v3MlTV(M)=v_{3}||M||. Then for any pp and qq coprime, any (p,q)(p,q)-cable MM^{\prime} also satisfies Conjecture 1.4.

Proof.

By Theorem 2.4 Part (1)(1), LTV(Cp,q)0LTV(C_{p,q})\leq 0, and thus by Theorem 2.4 Part (3)(3), LTV(M)LTV(M)LTV(M^{\prime})\leq LTV(M). Since lTV(M)=LTV(M)=v3MlTV(M)=LTV(M)=v_{3}\|M\|, the limit exists, and any subsequence also converges to v3Mv_{3}\|M\|. By Theorem 1.2 along odd rr,

lim supr, (r,q)=12πrlog|TVr(M)|\displaystyle\limsup_{r\rightarrow\infty,\text{ }(r,q)=1}\frac{2\pi}{r}\log\left|\text{TV}_{r}\left(M^{\prime}\right)\right| =lim supr, (r,q)=12πrlog|TVr(M)|=LTV(M)=v3M,\displaystyle=\limsup_{r\rightarrow\infty,\text{ }(r,q)=1}\frac{2\pi}{r}\log\left|\text{TV}_{r}\left(M\right)\right|=LTV(M)=v_{3}||M||,

where

lim supr, (r,q)=12πrlog|TVr()|\limsup_{r\rightarrow\infty,\text{ }(r,q)=1}\frac{2\pi}{r}\log\left|\text{TV}_{r}\left(-\right)\right|

is the limit superior of the subsequence along which rr and qq are coprime.

Since

v3M=lim supr, (r,q)=12πrlog|TVr(M)|\displaystyle v_{3}||M||=\limsup_{r\rightarrow\infty,\text{ }(r,q)=1}\frac{2\pi}{r}\log\left|\text{TV}_{r}\left(M^{\prime}\right)\right| LTV(M)LTV(M)=v3M,\displaystyle\leq LTV(M^{\prime})\leq LTV(M)=v_{3}||M||,

we have

LTV(M)=v3M=v3M,LTV(M^{\prime})=v_{3}||M||=v_{3}||M^{\prime}||,

where the final equality follows from the fact that the simplicial volume does not change under attaching a (p,q)(p,q)-cable space. ∎

Corollary 1.6.

Suppose MM satisfies Conjecture 1.4. Then for any odd pp and nn\in\mathbb{N}, any (p,2n)(p,2^{n})-cable MM^{\prime} also satisfies Conjecture 1.4. Moreover, if lTV(M)=v3MlTV(M)=v_{3}||M||, then lTV(M)=LTV(M)=v3MlTV(M^{\prime})=LTV(M^{\prime})=v_{3}||M^{\prime}||.

Proof.

Since rr is odd, (r,2n)=1(r,2^{n})=1 for any n1n\geq 1, which means Theorem 1.2 holds for any (p,2n)(p,2^{n})-cable of MM provided pp is odd. Since M=M||M||=||M^{\prime}||, this implies that

LTV(M)=LTV(M)=v3M=v3M,LTV(M^{\prime})=LTV(M)=v_{3}||M||=v_{3}||M^{\prime}||,

so MM^{\prime} also satisfies Conjecture 1.4.

Theorem 1.2 also implies that lTV(M)=lTV(M)lTV(M^{\prime})=lTV(M). In the case where lTV(M)=v3MlTV(M)=v_{3}||M||, we recover the full limit

lTV(M)=lTV(M)=v3||M|=v3||M||=LTV(M).lTV(M^{\prime})=lTV(M)=v_{3}||M|=v_{3}||M^{\prime}||=LTV(M^{\prime}).

4 Proof of Supporting Theorem

In this section, we will provide a proof of Theorem 1.7, which we restate here for convenience.

Theorem 1.7.

Let pp be coprime to some positive integer qq. Then RTr(Cp,q)RT_{r}(C_{p,q}) is invertible if and only if rr and qq are coprime. Moreover, the operator norm |RTr(Cp,q)1||||RT_{r}(C_{p,q})^{-1}||| grows at most polynomially.

We will use the following supporting proposition for the proof of Theorem 1.7, which is given in Subsection 4.1. The proof of this proposition is given in Subsection 4.2. We also use a couple of technical lemmas which are subsequently proven in Subsection 4.3. We begin by constructing a basis over which RTr(Cp,q)RT_{r}(C_{p,q}) admits a simpler expression.

By the cabling formula given by Theorem 2.2,

RTr(Cp,q)(ei)Span{e1,eql+1,eql1}l=1m1\displaystyle RT_{r}(C_{p,q})(e_{i})\in Span\{e_{1},e_{ql+1},e_{ql-1}\}_{l=1}^{m-1}

where m=r12m=\frac{r-1}{2}. Let Fm:={fl}l=0m1F_{m}:=\{f_{l}\}_{l=0}^{m-1}, where

f0:=\displaystyle f_{0}:= e1\displaystyle e_{1}
fl:=\displaystyle f_{l}:= eql+1A2pleql1l=1,,m.\displaystyle e_{ql+1}-A^{2pl}e_{ql-1}\qquad l=1,\dots,m.

Define f~lSpan{e1,,em}\tilde{f}_{l}\in Span\{e_{1},\dots,e_{m}\} to be the reduction of flf_{l} under the quotient induced by the symmetries ei=eie_{-i}=-e_{i} for any i0i\geq 0 and ei+nr=(1)neie_{i+nr}=(-1)^{n}e_{i} for any nn\in\mathbb{Z}. Note that for each ll, ql±1=kr+jql\pm 1=kr+j for some non-negative integers k,jk,j where 0j<r0\leq j<r. This means that up to sign, these symmetries imply

eql±1\displaystyle e_{ql\pm 1} =eqlkr±1=ej\displaystyle=e_{ql-kr\pm 1}=e_{j} for 0jm\displaystyle\text{for }0\leq j\leq m (1)
eql±1\displaystyle e_{ql\pm 1} =e(k+1)rql1=erj\displaystyle=e_{(k+1)r-ql\mp 1}=e_{r-j} for m+1j<r.\displaystyle\text{for }m+1\leq j<r. (2)

Finally, define F~m:={f~l}l=0m1\tilde{F}_{m}:=\{\tilde{f}_{l}\}_{l=0}^{m-1}, and let RmR_{m} be the (m×m)(m\times m)-matrix with columns corresponding to the reduced vectors f~l\tilde{f}_{l}, for l=0,,m1l=0,\dots,m-1. In particular, f~l\tilde{f}_{l} corresponds to col(l+1)col(l+1) of RmR_{m}, and the rows of RmR_{m} correspond to the original orthonormal basis {e1,,em}\{e_{1},\dots,e_{m}\} spanning RTr(T2)RT_{r}(T^{2}).

Remark 4.1.

We note that FmF_{m}, F~m\tilde{F}_{m}, and RmR_{m} are also dependent on pp and qq, but these dependencies are suppressed to avoid unwieldy notation.

The following proposition will be used to prove Theorem 1.7.

Proposition 4.2.

Let r=2m+13r=2m+1\geq 3 be coprime to qq. Then RmR_{m} is a change of basis from F~m{e1,,em}\tilde{F}_{m}\rightarrow\{e_{1},\dots,e_{m}\} and the operator norm |Rm1||||R_{m}^{-1}||| grows at most polynomially in mm. Moreover, for i{1,,m}i\in\{1,\dots,m\},

RTr(Cp,q)(ei)=\displaystyle RT_{r}(C_{p,q})(e_{i})= Aqp2(i21)lTiAp(q2l2+l)f~l,\displaystyle A^{\frac{qp}{2}(i^{2}-1)}\sum_{l\in T_{i}}A^{-p\left(\frac{q}{2}l^{2}+l\right)}\tilde{f}_{l}, (3)

where Ti={0,2,,i1}T_{i}=\{0,2,\dots,i-1\} for odd ii and Ti={1,3,,i1}T_{i}=\{1,3,\dots,i-1\} for even ii.

The idea of the proof is to leverage symmetric properties of the f~l\tilde{f}_{l} to give a presentation of Rm1R_{m}^{-1} and bound its operator norm. The assumption that (r,q)=1(r,q)=1 is necessary for invertibility, as indicated by the following proposition.

Proposition 4.3.

Suppose r=2m+13r=2m+1\geq 3 is odd and not coprime to qq. Then RmR_{m} is singular.

The proofs of Propositions 4.2 and 4.3 will be given in Subsection 4.2.

4.1 Proof of Theorem 1.7

We now can proceed with the proof of Theorem 1.7 assuming Proposition 4.2.

Proof of Theorem 1.7.

We begin with the necessary condition. Suppose (r,q)=d>1(r,q)=d>1. Then there are coprime q,rq^{\prime},r^{\prime} such that q=dqq=dq^{\prime} and r=drr=dr^{\prime}. We claim that row(nd)row(nd) of RTr(Cp,q)RT_{r}(C_{p,q}) consists of only zeros for each nn. Suppose some eql±1=ekr+je_{ql\pm 1}=e_{kr+j}, where 0jm0\leq j\leq m, reduces to ej=ende_{j}=e_{nd}. Then by Equation (1), qlkr±1=ndql-kr\pm 1=nd, which means d(qlrkn)=1d(q^{\prime}l-r^{\prime}k-n)=\mp 1, which is a contradiction. Similarly, if eql±1=ekr+je_{ql\pm 1}=e_{kr+j}, where m+1j<rm+1\leq j<r, reduces to erj=ende_{r-j}=e_{nd}. Then by Equation (2), d((1+k)rqln)=±1d((1+k)r^{\prime}-q^{\prime}l-n)=\pm 1, which is also a contradiction. This means that row(nd)=[0,,0]row(nd)=[0,\dots,0], thus RTr(Cp,q)RT_{r}(C_{p,q}) is singular.

For sufficiency, suppose (r,q)=1(r,q)=1. By Proposition 4.2, we can write RTr(Cp,q)RT_{r}(C_{p,q}) as a product of two diagonal matrices with an upper-triangular matrix and the change of basis RmR_{m}:

RTr(Cp,q)=\displaystyle RT_{r}(C_{p,q})= Rm(1000Ap((21)+q2(21)2)000Ap((m1)+q2(m1)2))×\displaystyle R_{m}\left(\begin{array}[]{cccc}1&0&\dots&0\\ 0&A^{-p\left((2-1)+\frac{q}{2}(2-1)^{2}\right)}&\ddots&\vdots\\ \vdots&\ddots&\ddots&0\\ 0&\dots&0&A^{-p\left((m-1)+\frac{q}{2}(m-1)^{2}\right)}\end{array}\right)\times
(101010010101101010110001)(1000Aqp2(221)000Aqp2(m21)).\displaystyle\left(\begin{array}[]{ccccccc}1&0&1&0&1&0&\dots\\ 0&1&0&1&0&1&\dots\\ &&1&0&1&0&\dots\\ \vdots&&&1&0&1&\dots\\ &&&&1&0&\\ &&&&&\ddots&\\ 0&&\dots&&&0&1\end{array}\right)\left(\begin{array}[]{cccc}1&0&\dots&0\\ 0&A^{\frac{qp}{2}(2^{2}-1)}&\ddots&\vdots\\ \vdots&\ddots&\ddots&0\\ 0&\dots&0&A^{\frac{qp}{2}(m^{2}-1)}\end{array}\right).

Note that the columns of the middle upper triangular matrix correspond to the index sets TiT_{i} of the sum in Equation (3). Inverting this product, we have

RTr(Cp,q)1=\displaystyle RT_{r}(C_{p,q})^{-1}= (1000Aqp2(122)000Aqp2(1m2))(10100001010010101010110001)\displaystyle\left(\begin{array}[]{cccc}1&0&\dots&0\\ 0&A^{\frac{qp}{2}(1-2^{2})}&\ddots&\vdots\\ \vdots&\ddots&\ddots&0\\ 0&\dots&0&A^{\frac{qp}{2}(1-m^{2})}\end{array}\right)\left(\begin{array}[]{ccccccc}1&0&-1&0&0&\dots&0\\ 0&1&0&-1&0&\dots&0\\ &&1&0&-1&0&\vdots\\ \vdots&&&1&0&-1\dots&0\\ &&&&1&\ddots&-1\\ &&&&&\ddots&0\\ 0&&&\dots&&0&1\end{array}\right)
×(1000Ap((21)+q2(21)2)000Ap((m1)+q2(m1)2))Rm1.\displaystyle\times\left(\begin{array}[]{cccc}1&0&\dots&0\\ 0&A^{p\left((2-1)+\frac{q}{2}(2-1)^{2}\right)}&\ddots&\vdots\\ \vdots&\ddots&\ddots&0\\ 0&\dots&0&A^{p\left((m-1)+\frac{q}{2}(m-1)^{2}\right)}\end{array}\right)R_{m}^{-1}.

By Proposition 4.2, |Rm1||||R_{m}^{-1}||| grows at most polynomially in mm, so it is bounded polynomially in rr. For the total bound, observe that both of the diagonal matrices are isometries, and the upper triangular matrix has operator norm bounded above by a polynomial in rr by the Cauchy-Schwartz inequality ∎

4.2 Proof of Propositions 4.2 and 4.3

We give proofs of Propositions 4.2 and 4.3 in this subsection. The following definitions and lemmas will be useful in the proofs.

By the symmetries ei=eie_{-i}=-e_{i} for any i0i\geq 0 and ei+kr=(1)keie_{i+kr}=(-1)^{k}e_{i} for any kk\in\mathbb{Z}, we may extend the definition of flf_{l} to all ll\in\mathbb{Z} using the following symmetries:

  • fl=e1+ql+A2ple1qlf_{l}=e_{1+ql}+A^{2pl}e_{1-ql} for any ll\in\mathbb{Z},

  • fl+r=(1)qflf_{l+r}=(-1)^{q}f_{l} for any ll\in\mathbb{Z}, and

  • fl=A2plflf_{l}=A^{2pl}f_{-l}.

The following Lemma will be used to present Rm1R_{m}^{-1}.

Lemma 4.4.

Let r=2m+13r=2m+1\geq 3 be coprime to qq, and let qq^{*} be the inverse of qq modulo rr. Then for l{0,,m1}l\in\{0,\dots,m-1\},

el+1={f0 if l=0fq if l=1fql+k=1l/2A2pq(kli=0k12i)fq(l2k) if l>1.e_{l+1}=\begin{cases}f_{0}&\text{ if }l=0\\ f_{q^{*}}&\text{ if }l=1\\ f_{q^{*}l}+\sum_{k=1}^{\lfloor l/2\rfloor}A^{2pq^{*}\left(kl-\sum_{i=0}^{k-1}2i\right)}f_{q^{*}(l-2k)}&\text{ if }l>1.\end{cases} (4)

Moreover, for i,j{0,,r1}i,j\in\{0,\dots,r-1\}, qiqjmodrq^{*}i\equiv q^{*}j\mod r if and only if i=ji=j.

Proof.

Since (r,q)=1(r,q)=1, there is a unique qrq^{*}\in\mathbb{Z}_{r} such that qq1modrqq^{*}\equiv 1\mod r. Using the symmetries of flf_{l} and substituting in qlq^{*}l, we have

e1+l\displaystyle e_{1+l} =fqlA2pqle1l=fql+A2pqlel1=fql+A2pqle1+(l2).\displaystyle=f_{q^{*}l}-A^{2pq^{*}l}e_{1-l}=f_{q^{*}l}+A^{2pq^{*}l}e_{l-1}=f_{q^{*}l}+A^{2pq^{*}l}e_{1+(l-2)}. (5)

We can then apply Equation (5) iteratively to express the eie_{i}’s in terms of the fjf_{j}’s.

el+1=fql+A2pqlfq(l2)+A2pq(2l2)fq(l4)++A2pq(l/2li=0l/212i)fq(l2l/2)\displaystyle e_{l+1}=f_{q^{*}l}+A^{2pq^{*}l}f_{q^{*}(l-2)}+A^{2pq^{*}(2l-2)}f_{q^{*}(l-4)}+\cdots+A^{2pq^{*}\left(\lfloor l/2\rfloor l-\sum_{i=0}^{\lfloor l/2\rfloor-1}2i\right)}f_{q^{*}(l-2\lfloor l/2\rfloor)}

for l{2,,m1}l\in\{2,\dots,m-1\}. When l=0l=0, by definition, e1=f0.e_{1}=f_{0}. When l=1l=1, Equation (5) yields that e2=fq.e_{2}=f_{q^{*}}. For any l{0,,m1}l\in\{0,\dots,m-1\}, the iterative use of Equation (5) to express el+1e_{l+1} terminates when the final term is a scalar multiple of either e1e_{1} or e2e_{2}, depending on the parity of ll.

For the final statement, note that (q,r)=1(q^{*},r)=1. This means there is a group isomorphism between the cyclic groups {qkmodr|kr}\{q^{*}k\mod r|k\in\mathbb{Z}_{r}\} and r\mathbb{Z}_{r} sending the indices qkmodrq^{*}k\mod r in Equation (4) to distinct jj for j{0,1,,r1}j\in\{0,1,\dots,r-1\}. Since qkmodrq^{*}k\mod r are distinct for krk\in\mathbb{Z}_{r}, this shows that qiqjmodrq^{*}i\equiv q^{*}j\mod r if and only if i=ji=j. ∎

In order to prove Proposition 4.2, we will use the following lemma. The proof of this lemma is given in Subsection 4.3.

Lemma 4.5.

Suppose r=2m+13r=2m+1\geq 3 is coprime to qq. Then

f~m=j=0m1Cjfj,\tilde{f}_{m}=\sum_{j=0}^{m-1}C_{j}f_{j},

where CjC_{j}\in\mathbb{C} such that |Cj|=1|C_{j}|=1 for j{0,,m1}j\in\{0,\dots,m-1\}.

Proof of Proposition 4.2.

It suffices to show that RmR_{m} is nonsingular, in which case RmR_{m} corresponds to the basis transformation F~m{e1,,em}\tilde{F}_{m}\rightarrow\{e_{1},\dots,e_{m}\}. To establish nonsingularity, we will give a presentation of Rm1R_{m}^{-1} by expressing eie_{i}, for i{1,,m}i\in\{1,\dots,m\}, in terms of fjf_{j} where j{0,m1}j\in\{0,\dots m-1\}.

By Lemma 4.4, each eie_{i}, for i{1,,m}i\in\{1,\dots,m\}, can be written in terms of fjf_{j} where jj\in\mathbb{Z}. These fjf_{j}’s reduce to flf_{l}’s, where l{0,,m}l\in\{0,\dots,m\}, using the above symmetries. This means that Span{e1,,em}Span\{e_{1},\dots,e_{m}\} of dimension mm is contained in Span{f0,,fm}Span\{f_{0},\dots,f_{m}\}, a vector space of dimension at most m+1m+1.

Lemma 4.5 implies that f~mSpan{f0,,fm1}\tilde{f}_{m}\in Span\{f_{0},\dots,f_{m-1}\}, which means that

Span{f0,,fm1}=Span{f0,,fm}Span{e1,,em}.Span\{f_{0},\dots,f_{m-1}\}=Span\{f_{0},\dots,f_{m}\}\supseteq Span\{e_{1},\dots,e_{m}\}.

Since {e1,,em}\{e_{1},\dots,e_{m}\} is a basis for the mm-dimensional vector space RTr(T2)RT_{r}(T^{2}), {f0,,fm1}\{f_{0},\dots,f_{m-1}\} is a set of mm vectors, and Span{f0,,fm1}Span{e1,,em}Span\{f_{0},\dots,f_{m-1}\}\supseteq Span\{e_{1},\dots,e_{m}\}, then

Span{f0,,fm1}=Span{e1,,em}=RTr(T2).Span\{f_{0},\dots,f_{m-1}\}=Span\{e_{1},\dots,e_{m}\}=RT_{r}(T^{2}).

From this, we conclude that {f0,,fm1}\{f_{0},\dots,f_{m-1}\} is also a basis for RTr(T2)RT_{r}(T^{2}). Since {f0,,fm1}\{f_{0},\dots,f_{m-1}\} is a basis, this implies that Rm1R_{m}^{-1} is a change-of-basis matrix and is nonsingular, therefore, RmR_{m} is nonsingular.

In order to bound the operator norm |Rm1||||R_{m}^{-1}|||, we study the presentation of Rm1R_{m}^{-1} more closely. By Lemma 4.4, we may express each eie_{i} as

ei=j=0r1Bjifj,e_{i}=\sum_{j=0}^{r-1}B_{j}^{i}f_{j},

where BjiB_{j}^{i} is either zero or a root of unity and the summands correspond to the reduction of each index modulo rr. We remark that since BjiB_{j}^{i} is either zero or a root of unity, |Bji|1|B_{j}^{i}|\leq 1. Now after applying the symmetry fl=A2plfl=A2plfrlf_{l}=A^{2pl}f_{-l}=A^{2pl}f_{r-l} for any l>ml>m, we may express eie_{i} as

ei=j=0m(Bji+A2p(rj)Brji)fj=j=0mDjifj,e_{i}=\sum_{j=0}^{m}(B_{j}^{i}+A^{2p(r-j)}B_{r-j}^{i})f_{j}=\sum_{j=0}^{m}D_{j}^{i}f_{j},

where Dji=Bji+A2p(rj)BrjiD_{j}^{i}=B_{j}^{i}+A^{2p(r-j)}B_{r-j}^{i} and |Dji|2|D_{j}^{i}|\leq 2. Additionally, by Lemma 4.5, we know that the coefficient of any summand of fmf_{m} in terms of the basis {f0,,fm1}\{f_{0},\dots,f_{m-1}\} is CjiC_{j}^{i} with |Cji|=1|C_{j}^{i}|=1. This means we may write

ei\displaystyle e_{i} =(j=0m1Djifj)+Dmifm\displaystyle=\left(\sum_{j=0}^{m-1}D_{j}^{i}f_{j}\right)+D_{m}^{i}f_{m}
=(j=0m1Djifj)+Dmi(k=0m1Ckifk)\displaystyle=\left(\sum_{j=0}^{m-1}D_{j}^{i}f_{j}\right)+D_{m}^{i}\left(\sum_{k=0}^{m-1}C_{k}^{i}f_{k}\right)
=j=0m1[DmiCji+Dji]fj.\displaystyle=\sum_{j=0}^{m-1}\left[D_{m}^{i}C_{j}^{i}+D_{j}^{i}\right]f_{j}.
=j=0m1Ejifj\displaystyle=\sum_{j=0}^{m-1}E_{j}^{i}f_{j}

where Eji=DmiCji+Dji.E_{j}^{i}=D_{m}^{i}C_{j}^{i}+D_{j}^{i}. Note that since |Dji|2|D_{j}^{i}|\leq 2 and |Cji|=1|C_{j}^{i}|=1, we have

|Eji|=|DmiCji+Dji||DmiCji|+|Dji|=|Dmi||Cji|+|Dji|4.|E_{j}^{i}|=|D_{m}^{i}C_{j}^{i}+D_{j}^{i}|\leq|D_{m}^{i}C_{j}^{i}|+|D_{j}^{i}|=|D_{m}^{i}||C_{j}^{i}|+|D_{j}^{i}|\leq 4.

Hence every entry of Rm1R_{m}^{-1} has modulus bounded above by 44. For any complex unit vector v=[v0,,vm1]Tv=[v_{0},\dots,v_{m-1}]^{T} such that |vi|1|v_{i}|\leq 1 for i{0,m1}i\in\{0,\dots m-1\}, the Cauchy-Schwartz inequality implies that

Rm1v\displaystyle\|R_{m}^{-1}v\| =[i=0m1E0ivi,,i=0m1Em1ivi]T\displaystyle=\left\|\left[\sum_{i=0}^{m-1}E_{0}^{i}v_{i},\dots,\sum_{i=0}^{m-1}E_{m-1}^{i}v_{i}\right]^{T}\right\|
=(|i=0m1E0ivi|2++|i=0m1Em1ivi|2)12\displaystyle=\left(\left|\sum_{i=0}^{m-1}E_{0}^{i}v_{i}\right|^{2}+\dots+\left|\sum_{i=0}^{m-1}E_{m-1}^{i}v_{i}\right|^{2}\right)^{\frac{1}{2}}
(i,j=0m1|Eji|2|vi|2)12(i,j=0m1|Eji|2)12\displaystyle\leq\left(\sum_{i,j=0}^{m-1}\left|E_{j}^{i}\right|^{2}\left|v_{i}\right|^{2}\right)^{\frac{1}{2}}\leq\left(\sum_{i,j=0}^{m-1}\left|E_{j}^{i}\right|^{2}\right)^{\frac{1}{2}}
(i,j=0m142)12=(i,j=0m116)12=(16m2)12=4m.\displaystyle\leq\left(\sum_{i,j=0}^{m-1}4^{2}\right)^{\frac{1}{2}}=\left(\sum_{i,j=0}^{m-1}16\right)^{\frac{1}{2}}=\left(16m^{2}\right)^{\frac{1}{2}}=4m.

This shows that

Rm1vO(m),||R_{m}^{-1}v||\leq O(m),

so the operator norm |Rm1||||R_{m}^{-1}||| is bounded polynomially.

Lastly, by the Cabling Formula in Theorem 2.2 and the definition of flf_{l}, the coefficient of flf_{l} in RTr(Cp,q)(ei)RT_{r}(C_{p,q})(e_{i}) is given by

RTr(Cp,q)(ei)=\displaystyle RT_{r}(C_{p,q})(e_{i})= Aqp2(i21)lTiAp(q2l2+l)f~l,\displaystyle A^{\frac{qp}{2}(i^{2}-1)}\sum_{l\in T_{i}}A^{-p\left(\frac{q}{2}l^{2}+l\right)}\tilde{f}_{l},

where Ti={0,2,,i1}T_{i}=\{0,2,\dots,i-1\} for odd ii and Ti={1,3,,i1}T_{i}=\{1,3,\dots,i-1\} for even ii. ∎

In order to prove Proposition 4.3, we establish the following definitions.

For 1lm1\leq l\leq m, define fl±:=eql±1f_{l}^{\pm}:=e_{ql\pm 1}. Observe that fl=fl+A2plflf_{l}=f_{l}^{+}-A^{2pl}f_{l}^{-} for 1lm1\leq l\leq m. In addition, define f~l±\tilde{f}_{l}^{\pm} to be the quotient of fl±f_{l}^{\pm} under the symmetries ei=eie_{-i}=-e_{i} for any i0i\geq 0 and ei+kr=(1)keie_{i+kr}=(-1)^{k}e_{i} for any kk\in\mathbb{Z}. We will use the convention that f0+=f0=e1f_{0}^{+}=f_{0}=e_{1} and f0=0f_{0}^{-}=0.

Recall that for each ll, ql±1=kr+jql\pm 1=kr+j for some non-negative integers k,jk,j where 0j<r0\leq j<r. This means that up to sign,

f~l±={ej=eqlkr±10jmerj=e(k+1)rql1m+1j<r.\displaystyle\tilde{f}_{l}^{\pm}=\begin{cases}e_{j}=e_{ql-kr\pm 1}&0\leq j\leq m\\ e_{r-j}=e_{(k+1)r-ql\mp 1}&m+1\leq j<r.\end{cases} (6)

We can now prove Proposition 4.3.

Proof of Proposition 4.3.

We will repeat the argument given in the proof of Theorem 1.7. Suppose (r,q)=d>1(r,q)=d>1, then there are coprime qq^{\prime} and rr^{\prime} such that q=dqq=dq^{\prime} and r=drr=dr^{\prime}. We claim that row(nd)row(nd) of RmR_{m} consists of only zeros for each nn. Suppose some f~l±=ej=end\tilde{f}^{\pm}_{l}=e_{j}=e_{nd}, then qlkr±1=ndql-kr\pm 1=nd. This implies that d(qlrkn)=1d(q^{\prime}l-r^{\prime}k-n)=\mp 1 which is a contradiction. Similarly, if f~l±=erj=end\tilde{f}^{\pm}_{l}=e_{r-j}=e_{nd}, then d((1+k)rqln)=±1d((1+k)r^{\prime}-q^{\prime}l-n)=\pm 1 which is also a contradiction. This means that row(nd)=[0,,0]row(nd)=[0,\dots,0], thus RmR_{m} is singular. ∎

4.3 Proof of Lemma 4.5

In this subsection, we provide a proof for Lemma 4.5. We use the notation introduced in Subsection 4.2.

Remark 4.6.

For the following arguments, we use the convention that equalities between vectors eie_{i} are necessarily taken up to sign. This ultimately has no effect on the arguments for Proposition 4.2 and Theorem 1.7.

Recall RmR_{m} is the (m×m)(m\times m)-matrix with columns corresponding to F~m={f0,,fm1}\tilde{F}_{m}=\{f_{0},\dots,f_{m-1}\}. We also define SmS_{m} to be the (m×(m+1))(m\times(m+1))-matrix obtained by appending the column corresponding to f~m\tilde{f}_{m} to RmR_{m}. The following technical lemmas will be used in the proof of Lemma 4.5.

Lemma 4.7.

Suppose r=2m+13r=2m+1\geq 3 is coprime to qq, and let qq^{*} be the multiplicative inverse of qq in the ring r\mathbb{Z}_{r}. Then

  1. (i)

    Each column of SmS_{m} has at most two nonzero entries. Moreover, for each column with two nonzero entries, their corresponding row indices differ by at most 22.

  2. (ii)

    Let

    l:={qif qmrqif q>m.l^{*}:=\begin{cases}q^{*}&\text{if }q^{*}\leq m\\ r-q^{*}&\text{if }q^{*}>m.\end{cases}

    Then in SmS_{m}, col(1)=[1,0,,0]Tcol(1)=[1,0,\dots,0]^{T} and col(l+1)=[0,Dl,0,,0]Tcol(l^{*}+1)=[0,D_{l^{*}},0,\dots,0]^{T} where DlD_{l^{*}} is a root of unity. Moreover, every other column of SmS_{m} has exactly two nonzero entries which are roots of unity.

Proof.

Part (i)(\ref{matrixlem_i}): Each column of SmS_{m} corresponds to the reduced vector f~l\tilde{f}_{l}, 0lm0\leq l\leq m. Since flf_{l} is a linear combination of at most two vectors in Span{e1,eql+1,eql1}l=1m1Span\{e_{1},e_{ql+1},e_{ql-1}\}_{l=1}^{m-1}, there are at most two nonzero entries in col(l+1)col(l+1).

Now suppose the index of fl+f_{l}^{+} is ql+1=kr+jql+1=kr+j, where 0j<r0\leq j<r. Then the index of flf_{l}^{-} is ql1=kr+j2=kr+jql-1=kr+j-2=k^{\prime}r+j^{\prime}, where either (k,j)=(k1,r+j2)(k^{\prime},j^{\prime})=(k-1,r+j-2) (for j{0,1}j\in\{0,1\}) or (k,j)=(k,j2)(k^{\prime},j^{\prime})=(k,j-2) (for j2j\geq 2). We split into cases:

  • If j=0j=0, then j=r2j^{\prime}=r-2, f~l+=e0=0\tilde{f}_{l}^{+}=e_{0}=0, and f~l=e2\tilde{f}_{l}^{-}=e_{2}.

  • If j=1j=1, then j=r1j^{\prime}=r-1, f~l+=eqlkr+1=e1\tilde{f}_{l}^{+}=e_{ql-kr+1}=e_{1}, and f~l=e(k+1)rql+1=e1\tilde{f}_{l}^{-}=e_{(k+1)r-ql+1}=e_{1}. This implies that l=krql=\frac{kr}{q}. Since ll\in\mathbb{Z} and r,qr,q are coprime, k=qnk=qn, for some n0n\geq 0. However, if n1n\geq 1, we have l1+r>ml\geq 1+r>m, which is a contradiction. Thus n=0n=0, so l=0l=0, corresponding to col(1)col(1).

  • If 2jm2\leq j\leq m, then f~l+=ejej2=f~l\tilde{f}_{l}^{+}=e_{j}\neq e_{j-2}=\tilde{f}_{l}^{-}.

  • If j=m+1j=m+1, then j=m1j^{\prime}=m-1 and f~l+=emem1=f~l\tilde{f}_{l}^{+}=e_{m}\neq e_{m-1}=\tilde{f}_{l}^{-}.

  • If j=m+2j=m+2, then j=mj^{\prime}=m and f~l+=em2em=f~l\tilde{f}_{l}^{+}=e_{m-2}\neq e_{m}=\tilde{f}_{l}^{-}.

  • If m+3j<rm+3\leq j<r, then f~l+=erjerj+2=f~l\tilde{f}_{l}^{+}=e_{r-j}\neq e_{r-j+2}=\tilde{f}_{l}^{-}.

This implies that the row indices of the nonzero entries in each column differ by at most two for every column except col(1)col(1). In particular, the only case where the row indices differ by exactly 11 occurs when j=m+1j=m+1.

Part (ii)(\ref{matrixlem_ii}): Since ei=eie_{-i}=-e_{i}, we have fl=e1+ql+A2ple1qlf_{l}=e_{1+ql}+A^{2pl}e_{1-ql} for 1lm1\leq l\leq m. Note that col(l+1)col(l+1) has exactly one nonzero entry if and only if one of the following occurs:

  1. (1)

    Either 1+ql1+ql and 1ql1-ql are equal or opposite modulo rr.

  2. (2)

    Either 1+ql1+ql or 1ql1-ql vanishes modulo rr.

Case (1)(1) occurs if and only if l=0l=0, corresponding to f~0=f0=e0\tilde{f}_{0}=f_{0}=e_{0}. In this case, col(1)=[1,0,,0]Tcol(1)=[1,0,\dots,0]^{T}.

Case (2)(2) occurs if and only if either l=ql=q^{*} or l=ql=-q^{*} modulo rr. Define

l:={qif qmrqif q>m.l^{*}:=\begin{cases}q^{*}&\text{if }q^{*}\leq m\\ r-q^{*}&\text{if }q^{*}>m.\end{cases}

Note that if ql±1ql\pm 1 vanishes, then |ql1|=2|ql\mp 1|=2. Define DlD_{l^{*}} to be the coefficient of the vector e|ql1|=e2e_{|ql\mp 1|}=e_{2} obtained from Equation (3). This means that col(l+1)col(l^{*}+1) is the unique column with exactly one nonzero entry except for col(1)col(1).

Finally, the conclusion follows from the uniqueness of ll^{*} and Part (i)(\ref{matrixlem_i}). ∎

The second technical lemma makes use of Lemma 4.7 in its proof.

Lemma 4.8.

Suppose r=2m+13r=2m+1\geq 3 is coprime to qq. Then

  1. (i)

    Each row of SmS_{m} has exactly two nonzero entries.

  2. (ii)

    There is a unique ll^{\prime}, 1lm1\leq l^{\prime}\leq m, such that col(l+1)=[0,,0,Dl,El]Tcol(l^{\prime}+1)=[0,\dots,0,D_{l^{\prime}},E_{l^{\prime}}]^{T}, where Dl,ElD_{l^{\prime}},E_{l^{\prime}} are roots of unity.

The following lemma will be useful in the proof of Lemma 4.8.

Lemma 4.9.

Suppose r3r\geq 3 is coprime to qq, and let gl±:=qlkr±1g_{l}^{\pm}:=ql-kr\pm 1 and hl±=(1+k)rql1h_{l}^{\pm}=(1+k)r-ql\mp 1. Then for 0l1,l2m0\leq l_{1},l_{2}\leq m with l1l2l_{1}\neq l_{2},

  1. (i)

    gl1±=gl2±g_{l_{1}}^{\pm}=g_{l_{2}}^{\pm}, gl1±=hl2g_{l_{1}}^{\pm}=h_{l_{2}}^{\mp}, and hl1±=hl2±h_{l_{1}}^{\pm}=h_{l_{2}}^{\pm} do not have integer solutions,

  2. (ii)

    gl1±=gl2g_{l_{1}}^{\pm}=g_{l_{2}}^{\mp}, gl1±=hl2±g_{l_{1}}^{\pm}=h_{l_{2}}^{\pm}, and hl1±=hl2h_{l_{1}}^{\pm}=h_{l_{2}}^{\mp} may each have integer solutions.

Proof.

Note gl±g_{l}^{\pm} and hl±h_{l}^{\pm} encode the two families of indices of the reduced vectors f~l±\tilde{f}_{l}^{\pm} given in Equation (6). There are six equations relating pairs of expressions in {gl+,gl,hl+,hl}\{g_{l}^{+},g_{l}^{-},h_{l}^{+},h_{l}^{-}\}.

Part (i)(\ref{indlem(a)}): This follows from the fact that (r,q)=1(r,q)=1 and the bounds on l1l_{1} and l2l_{2}. We show the case gl1±=gl2±g_{l_{1}}^{\pm}=g_{l_{2}}^{\pm} and note that the other two cases follow analogously. Assume for distinct l1,l2{0,,m}l_{1},l_{2}\in\{0,\dots,m\} and k1,k2k_{1},k_{2}\in\mathbb{Z} that ql1k1r±1=ql2k2r±1ql_{1}-k_{1}r\pm 1=ql_{2}-k_{2}r\pm 1. This implies

k1k2=q(l1l2)r.k_{1}-k_{2}=\frac{q(l_{1}-l_{2})}{r}\in\mathbb{Z}.

Since l1l2l_{1}\neq l_{2} and (r,q)=1(r,q)=1, l1l2l_{1}-l_{2} must have a nontrivial factor of rr, which contradicts the bounds on l1l_{1} and l2l_{2}.

Part (ii)(\ref{indlem(b)}): We have the following:

  • gl1±=gl2g_{l_{1}}^{\pm}=g_{l_{2}}^{\mp} if and only if q(l1l2)=(k1k2)r2q(l_{1}-l_{2})=(k_{1}-k_{2})r\mp 2,

  • gl1±=hl2±g_{l_{1}}^{\pm}=h_{l_{2}}^{\pm} if and only if q(l1+l2)=(1+k1+k2)r2q(l_{1}+l_{2})=(1+k_{1}+k_{2})r\mp 2, and

  • hl1±=hl2h_{l_{1}}^{\pm}=h_{l_{2}}^{\mp} if and only if q(l1l2)=(k1k2)r2q(l_{1}-l_{2})=(k_{1}-k_{2})r\mp 2.

All three of these equations may have integer solutions for l1,l2{0,,m}l_{1},l_{2}\in\{0,\dots,m\}. ∎

Proof of Lemma 4.8.

We will use the same notation as in the proof of Lemma 4.7 and in Lemma 4.9.

Part (i)(\ref{matrixlem2_i}): It is a corollary of Lemma 4.9 that every row of SmS_{m} has at most two nonzero entries. In particular, let (l1,l2)(l_{1},l_{2}) be an integral solution to one of the equations of Lemma 4.9 Part (ii)(\ref{indlem(b)}). Suppose l3{0,,m1}l_{3}\in\{0,\dots,m-1\} is such that (l1,l3)(l_{1},l_{3}) and (l2,l3)(l_{2},l_{3}) are both solutions to equations in Lemma 4.9 Part (ii)(\ref{indlem(b)}). Then by Lemma 4.9 Part (i)(\ref{indlem(a)}), either l3=l1l_{3}=l_{1} or l3=l2l_{3}=l_{2}.

Note that by Lemma 4.7 Part (ii)(\ref{matrixlem_ii}), SmS_{m} has exactly 2m2m nonzero entries since there are 2 in each column other than col(1)col(1) and col(l+1)col(l^{*}+1), which each have exactly 1. This means that every row of SmS_{m} must have exactly 2 nonzero entries.

Part (ii)(\ref{matrixlem2_ii}): In the proof of Lemma 4.7 Part (i)(\ref{matrixlem_i}), we saw that the only value of jj corresponding to a column with the nonzero row entry indices differing by 1 is j=m+1j=m+1. By Part (i)(\ref{matrixlem2_i}), row(m)row(m) of SmS_{m} has exactly 2 nonzero entries. This implies that there are some l1,l2l_{1},l_{2} such that col(l1+1)col(l_{1}+1) has nonzero entries in row(m)row(m) and row(m1)row(m-1) and col(l2+1)col(l_{2}+1) has nonzero entries in row(m)row(m) and row(m2)row(m-2). Take l=l1l^{\prime}=l_{1}. Finally, define DlD_{l^{\prime}} and ElE_{l^{\prime}} to be the coefficients of the vectors em1e_{m-1} and eme_{m} defined by Equation (3), respectively. Note that if m=2m=2, l2=ll_{2}=l^{*} and col(l2+1)col(l_{2}+1) has only 1 nonzero entry. ∎

Lastly, we are ready to prove Lemma 4.5.

Proof of Lemma 4.5.

The last column col(m+1)col(m+1) of the matrix SmS_{m} represents the reduced vector fmf_{m} written in terms of the basis {e1,,em}\{e_{1},\dots,e_{m}\}. We will prove Lemma 4.5 by showing that col(m+1)col(m+1) can be written as a linear combination of the first mm columns. From this linear combination, we will see that the coefficients will have the required bounds from the statement.

We claim that col(m+1)col(m+1) of SmS_{m} can be written as a linear combination of elements in {f0,,fm1}\{f_{0},\dots,f_{m-1}\}. From Lemma 4.7 Part (ii)(\ref{matrixlem_ii}), if l=ml^{*}=m, then col(m+1)col(m+1) has exactly one nonzero entry, and if l<ml^{*}<m, then col(m+1)col(m+1) has exactly two nonzero entries.

Case 1:

We first consider the case l=ml^{*}=m. Here, the nonzero entry of col(m+1)col(m+1) lies in row(2)row(2). This implies that f~mSpan{e1,,em}\tilde{f}_{m}\in Span\{e_{1},\dots,e_{m}\} has a scalar of e2e_{2} as a summand. By Lemma 4.8 Part (i)(\ref{matrixlem2_i}), we know that there is exactly one other nonzero entry in row(2)row(2) in some column j1j_{1}. From the argument of Lemma 4.7 Part (i)(\ref{matrixlem_i}), there exists a nonzero entry in row(4)row(4) of col(j1)col(j_{1}). Lemma 4.8 Part (i)(\ref{matrixlem2_i}) implies there exists a nonzero entry in some column j2j_{2} and row(4)row(4). From the argument of Lemma 4.7 Part (i)(\ref{matrixlem_i}), there exists a nonzero entry in row(6)row(6) of col(j2)col(j_{2}). Again, we pick the other nonzero entry of row(6)row(6) which lies in some column j3j_{3}. Note that col(j3)col(j_{3}) cannot be equal to any of the previous columns. If it were a previous column, it would contradict our bound on the number of nonzero entries in a column. We continue this iteration until we reach either row(m1)row(m-1) or row(m)row(m), depending on the parity of mm.

If m1m-1 is even, by Lemma 4.8 Part (ii)(\ref{matrixlem2_ii}), the next corresponding row with a nonzero entry will be row(m)row(m) where mm is odd. Similarly, if mm is even, by Lemma 4.8 Part (ii)(\ref{matrixlem2_ii}), the next corresponding row with a nonzero entry will be row(m1)row(m-1) where m1m-1 is odd. Now when we continue the algorithm, our subsequent row indices will be odd and decrease by 22 until we reach row(1)row(1). By Lemma 4.8 Part (i)(\ref{matrixlem2_i}) and Lemma 4.7 Part (ii)(\ref{matrixlem_ii}), there exists a nonzero entry in row(1)row(1) of col(1)col(1), and it is the only nonzero entry in col(1)col(1). Since every entry of our matrix is a root of unity by Lemma 4.7 Part (ii)(\ref{matrixlem_ii}) and terminates at row(1)row(1), scalars by roots of unity of the columns appearing in our sequence gives fmf_{m} as a linear combination of elements of {f0,,fm1}\{f_{0},\dots,f_{m-1}\} where all coefficients are roots of unity.

Case 2:

Now suppose col(m+1)col(m+1) has exactly two nonzero entries. We denote the row indices of these entries by i1i_{1}^{-} and i1+i_{1}^{+}, where i1<i1+i_{1}^{-}<i_{1}^{+}. By Lemma 4.8 Part (i)(\ref{matrixlem2_i}), row(i1)row(i_{1}^{-}) has another nonzero entry in some other column j1j_{1}^{-}. Similarly, row(i1+)row(i_{1}^{+}) has another nonzero entry in some column j1+j_{1}^{+}. We make the following claim, which we prove at the end.

Claim: j1j1+j_{1}^{-}\neq j_{1}^{+}.

We will proceed similarly to the first case. Consider the column col(j1+)col(j_{1}^{+}), which has exactly two nonzero entries and cannot correspond to either col(1)col(1) or col(l+1)col(l^{*}+1) since i1+>2i_{1}^{+}>2. By Lemma 4.8 Part (i)(\ref{matrixlem2_i}) and the claim, there exists another nonzero entry in some row(i2+)row(i_{2}^{+}) of col(j1+)col(j_{1}^{+}) such that (i2+i1+){1,1,2}(i_{2}^{+}-i_{1}^{+})\in\{-1,1,2\}. The case when (i2+i1+)=1(i_{2}^{+}-i_{1}^{+})=-1 corresponds to i1+=mi_{1}^{+}=m, and the case when (i2+i1+)=1(i_{2}^{+}-i_{1}^{+})=1 corresponds to i1+=m1i_{1}^{+}=m-1. We now implement the same argument as the case with one entry in col(m+1)col(m+1). Note that, in this procedure, we do not utilize any rows with index less than i1i_{1}^{-} with the same parity as i1i_{1}^{-}. If i1=m1i_{1}^{-}=m-1 and i1+=mi_{1}^{+}=m, they will have different parities. In the other case, i1i_{1}^{-} will have the parity of ik+i_{k}^{+} until we have a kk such that (ik+ik1+){1,1}(i_{k}^{+}-i_{k-1}^{+})\in\{-1,1\}. This implies that for all kkk^{\prime}\geq k, ik+i_{k^{\prime}}^{+} will have opposite parity to i1i_{1}^{-}.

We now follow the same algorithm beginning with row(i1)row(i_{1}^{-}). By the claim, the indices of our subsequent rows iki_{k}^{-} must be decreasing. Otherwise, this would contradict Lemma 4.8 Part (i)(\ref{matrixlem2_i}).

Since both cases in total utilize every row exactly once, fmf_{m} is given by a linear combination of elements of {f0,,fm1}\{f_{0},\dots,f_{m-1}\} where, by Lemma 4.7 Part (ii)(\ref{matrixlem_ii}), all coefficients are roots of unity.

Proof of Claim:

It now suffices to prove that j1+j1j_{1}^{+}\neq j_{1}^{-}. By contradiction, let us assume that j1+=j1j_{1}^{+}=j_{1}^{-}, and we will denote i1=i1+i_{1}=i_{1}^{+}.

If i1=mi_{1}=m, then either i1=m1i_{1}^{-}=m-1 or i1=m2i_{1}^{-}=m-2. If i1=m1i_{1}^{-}=m-1, then since j1+=j1j_{1}^{+}=j_{1}^{-}, we will have two columns with nonzero entries in the last two rows. This contradicts Lemma 4.8 Part (ii)(\ref{matrixlem2_ii}), which states that there is a unique such column. If i1=m2i_{1}^{-}=m-2, then there are two distinct columns with nonzero entries in row(m2)row(m-2) and row(m)row(m). By Lemma 4.8 Part (ii)(\ref{matrixlem2_ii}), there must exist a different column with nonzero entries in row(m1)row(m-1) and row(m)row(m), which contradicts there being at most 22 entries in row(m)row(m).

If i1=m1i_{1}=m-1, then i1<m1i_{1}^{-}<m-1, and there are no other columns with nonzero entries in row(m1)row(m-1) besides col(j1+)col(j_{1}^{+}) and col(m+1)col(m+1). By Lemma 4.8 Part (ii)(\ref{matrixlem2_ii}), there must exist a different column with nonzero entries in row(m1)row(m-1) and row(m)row(m), which contradicts there being at most 22 entries in row(m1)row(m-1).

In the general case, we assume i1m2i_{1}\leq m-2, and we will define i2=i1+2i_{2}=i_{1}+2. Since j1+=j1j_{1}^{+}=j_{1}^{-}, col(j1+)col(j_{1}^{+}) and col(m+1)col(m+1) already have two nonzero entries. Since i2>2i_{2}>2, these entries cannot be in either col(1)col(1) or col(l+1)col(l^{*}+1) since they only have entries in the first two rows. This implies that the columns which correspond to nonzero entries in row(i2)row(i_{2}) must have exactly two nonzero entries in some columns col(j2+)col(j_{2}^{+}) and col(j2)col(j_{2}^{-}) such that j2+,j2{j1+,m+1}j_{2}^{+},j_{2}^{-}\notin\{j_{1}^{+},m+1\}. Since row(i1)row(i_{1}) has two nonzero entries in col(j1+)col(j_{1}^{+}) and col(m+1)col(m+1), the other nonzero entries in col(j2+)col(j_{2}^{+}) and col(j2)col(j_{2}^{-}) must be in some row(i3)row(i_{3}), where i3i2{1,1,2}i_{3}-i_{2}\in\{-1,1,2\}.

  • If i3i2=1i_{3}-i_{2}=-1, we have i2=mi_{2}=m and i3=m1i_{3}=m-1. Here, we reach the same contradiction as when i1=mi_{1}=m and i1=m1i_{1}^{-}=m-1.

  • If i3i2=1i_{3}-i_{2}=1, then i2=m1i_{2}=m-1 and i3=mi_{3}=m. This gives the same contradiction as when i1=mi_{1}=m and i1=m1i_{1}^{-}=m-1.

  • If i3i2=2i_{3}-i_{2}=2 with i2=m2i_{2}=m-2 and i3=mi_{3}=m, then our argument is the same as when i1=mi_{1}=m and i1=m2i_{1}^{-}=m-2.

Finally, we consider when i3i2=2i_{3}-i_{2}=2 and i3mi_{3}\neq m. In this case, we can continue to iterate the same algorithm until we reach the same contradictions.

5 Further Directions

The primary approach of this paper utilizes the invertibility of the operator RTrRT_{r} on the cable space Cp,qC_{p,q} as well as a polynomial bound on its operator norm. The same methodology could apply in the context for the operator RTrRT_{r} for other cable spaces.

Although the technique may apply in the case when the cable space has positive simplicial volume, a more natural approach would be to generalize our argument to other cable spaces with simplicial volume zero. For example, we may consider the manifold defined as follows. Let N=Σg,2×S1N=\Sigma_{g,2}\times S^{1} where Σg,2\Sigma_{g,2} is a orientable compact genus gg surface with 22 boundary components. Now let {xi}i=1mΣg,2\{x_{i}\}_{i=1}^{m}\subset\Sigma_{g,2} such that {xi}i=1m×S1\{x_{i}\}_{i=1}^{m}\times S^{1} is a collection of mm vertical fibers in NN. We define the Seifert cable space C(s1,sm)C\left(s_{1},\dots s_{m}\right) where si=piqis_{i}=\frac{p_{i}}{q_{i}}\in\mathbb{Q} to be the manifold obtained by performing sis_{i}-Dehn surgery along the ii-th vertical fiber in NN.

If an analogous result to Theorem 1.7 holds for the Seifert cable space C(s1,sm)C\left(s_{1},\dots s_{m}\right), the corresponding Theorem 1.2 will also follow as well as its applications to Conjecture 1.4. Similar to the constraint of Theorem 1.7 where rr and qq must be coprime, the analogous result for the Seifert cable space may require a related caveat. This leads to the following concluding question.

Question 5.1.

Is RTr(C(s1,sm))RT_{r}(C\left(s_{1},\dots s_{m}\right)) invertible when rr is sufficiently large and coprime to every qiq_{i}?

References

  • [1] G. Belletti. The maximum volume of hyperbolic polyhedra. Trans. Amer. Math. Soc., 374(2):1125–1153, 2021.
  • [2] G. Belletti, R. Detcherry, E. Kalfagianni, and T. Yang. Growth of quantum 6j6j-symbols and applications to the volume conjecture. J. Differential Geom., 120(2):199–229, 2022.
  • [3] R. Benedetti and C. Petronio. On Roberts’ proof of the Turaev-Walker theorem. J. Knot Theory Ramifications, 5(4):427–439, 1996.
  • [4] C. Blanchet, N. Habegger, G. Masbaum, and P. Vogel. Topological quantum field theories derived from the Kauffman bracket. Topology, 34(4):883–927, 1995.
  • [5] Q. Chen and T. Yang. Volume conjectures for the Reshetikhin-Turaev and the Turaev-Viro invariants. Quantum Topol., 9(3):419–460, 2018.
  • [6] R. Detcherry. Growth of Turaev-Viro invariants and cabling. J. Knot Theory Ramifications, 28(14):1950041, 8, 2019.
  • [7] R. Detcherry and E. Kalfagianni. Gromov norm and Turaev-Viro invariants of 3-manifolds. Ann. Sci. Éc. Norm. Supér. (4), 53(6):1363–1391, 2020.
  • [8] R. Detcherry, E. Kalfagianni, and T. Yang. Turaev-Viro invariants, colored Jones polynomials, and volume. Quantum Topol., 9(4):775–813, 2018.
  • [9] M. Gromov. Volume and bounded cohomology. Inst. Hautes Études Sci. Publ. Math., 56:5–99 (1983), 1982.
  • [10] W. H. Jaco and P. B. Shalen. Seifert fibered spaces in 33-manifolds. Mem. Amer. Math. Soc., 21(220):viii+192, 1979.
  • [11] K. Johannson. Homotopy equivalences of 33-manifolds with boundaries, volume 761 of Lecture Notes in Mathematics. Springer, Berlin, 1979.
  • [12] R. M. Kashaev. The hyperbolic volume of knots from the quantum dilogarithm. Lett. Math. Phys., 39(3):269–275, 1997.
  • [13] S. Kumar. Fundamental shadow links realized as links in S3{S}^{3}. Algebraic & Geometric Topology, 21(6):3153–3198, 2021.
  • [14] S. Kumar and J. M. Melby. Asymptotic additivity of the Turaev-Viro invariants for a family of 33-manifolds. Journal of the London Mathematical Society, to appear, 2022.
  • [15] H. R. Morton. The coloured Jones function and Alexander polynomial for torus knots. Math. Proc. Cambridge Philos. Soc., 117(1):129–135, 1995.
  • [16] H. Murakami and J. Murakami. The colored Jones polynomials and the simplicial volume of a knot. Acta Math., 186(1):85–104, 2001.
  • [17] G. Perelman. The entropy formula for the Ricci flow and its geometric applications. arXiv preprint at arXiv:math/0211159, 2002.
  • [18] G. Perelman. Finite extinction time for the solutions to the Ricci flow on certain three-manifolds. arXiv preprint at arXiv:math/0307245, 2003.
  • [19] G. Perelman. Ricci flow with surgery on three-manifolds. arXiv preprint at arXiv:math/0303109, 2003.
  • [20] N. Reshetikhin and V. G. Turaev. Invariants of 33-manifolds via link polynomials and quantum groups. Invent. Math., 103(3):547–597, 1991.
  • [21] J. Roberts. Skein theory and Turaev-Viro invariants. Topology, 34(4):771–787, 1995.
  • [22] W. P. Thurston. The geometry and topology of three-manifolds. Princeton University Math Department Notes, 1979.
  • [23] W. P. Thurston. Three-dimensional manifolds, Kleinian groups and hyperbolic geometry. Bull. Amer. Math. Soc. (N.S.), 6(3):357–381, 1982.
  • [24] V. G. Turaev and O. Y. Viro. State sum invariants of 33-manifolds and quantum 6j6j-symbols. Topology, 31(4):865–902, 1992.
  • [25] K. H. Wong. Asymptotics of some quantum invariants of the Whitehead chains. arXiv preprint at arXiv:1912.10638, 2019.
  • [26] K. H. Wong. Volume conjecture, geometric decomposition and deformation of hyperbolic structures. arXiv preprint at arXiv:1912.11779, 2020.

Department of Mathematics, University of California, Santa Barbara, Santa Barbara, CA, 93106-6105, USA

E-mail address: [email protected]

Department of Mathematics, Michigan State University, East Lansing, MI, 48824, USA

E-mail address: [email protected]