Turaev–Viro invariants and cabling operations
Abstract
In this paper, we study the variation of the Turaev–Viro invariants for -manifolds with toroidal boundary under the operation of attaching a -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 -manifold. For and coprime, we show that the Chen–Yang volume conjecture is stable under -cabling. We achieve our results by studying the linear operator associated to the torus knot cable spaces by the Reshetikhin–Turaev -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 , its Turaev–Viro invariants are a family of -valued homeomorphism invariants parameterized by an integer depending on a -th root of unity . We are primarily interested in the invariants when is odd and .
In this paper, we study the variation of the Turaev–Viro invariants of a -manifold with toroidal boundary when we attach a -cable space.
Definition 1.1.
Let be the standardly embedded solid torus in , and let be a closed neighborhood of . For and coprime integers with , let be the torus knot of slope . The -cable space, denoted , is the complement of the torus knot in . Let be a -manifold with toroidal boundary. A manifold obtained from gluing a -cable space to a boundary component of along the exterior toroidal boundary component of is called a -cable of .
Our main theorem is the following.
Theorem 1.2.
Let be a manifold with toroidal boundary, let be coprime integers with , and let be an odd integer coprime to . Suppose is a -cable of . Then there exists a constant and natural number such that
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 -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
and
Additionally, we will introduce the simplicial volume for compact orientable -manifolds with empty or toroidal boundary, originally defined by Gromov [9]. For a compact, orientable, irreducible -manifold, there is a unique collection of incompressible tori, up to isotopy, along which 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 coincides with the sum of the simplicial volumes of the resulting pieces. In the case where is hyperbolic, the simplicial volume is positive and is related to the hyperbolic volume of by
where is the hyperbolic volume of , is the volume of the regular ideal tetrahedron, and is the simplicial volume of .
This leads to a natural extension of Chen and Yang’s Turaev–Viro invariant volume conjecture [5].
Conjecture 1.4 ([5], [7]).
Let be a compact oriented 3-manifold. Then
where is the volume of the regular ideal tetrahedron and 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 which satisfies Conjecture 1.4, the property was proven for invertible cablings of 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 -cable space by Detcherry [6], which we extend in this work.
A key property of the -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.
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 by Belletti [1], a large family of octahedral links in by the first author of this paper [13], and a family of link complements in trivial -bundles over oriented connected closed surfaces by the authors of this paper [14].
For general and coprime, Corollary 1.5 demonstrates the stability of Conjecture 1.4 under the -cabling operation. However, in the case when for , we recover the full limit as shown in the following corollary.
Corollary 1.6.
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 -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 as a cobordism between tori, the Reshetikhin–Turaev -TQFT at level , denoted by , associates to it a linear operator
For odd and , Detcherry presents using the basis , which is equivalent to the orthonormal basis for given in [4] under the symmetry for . More details of the construction are given in Section 2. With this basis, can be presented as a product of two diagonal matrices and one triangular matrix. This allows the author to directly write the inverse of . From the inverse of this linear operator, Detcherry establishes a lower bound of the Turaev–Viro invariants under attaching a -cable space.
For general , does not have as simple a presentation under the same basis, making it more difficult to conclude that is invertible. In order to resolve this, we present using a different basis for , defined in Section 4, that allows us to also show directly that is invertible provided and are coprime. Following Detcherry’s argument, the invertibility of is integral in finding the lower bound from Theorem 1.2; however, the invertibility of is constrained by the condition that and are coprime, as outlined by Theorem 1.7.
Theorem 1.7.
Let be coprime to some positive integer . Then is invertible if and only if and are coprime. Moreover, the operator norm grows at most polynomially.
As we will show in Section 3, the coprime condition between and 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 -TQFTs, the 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 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 -TQFTs, which were defined by Reshetikhin and Turaev in [20]. Let be the category of -dimensional cobordisms and be the category of -vector spaces. For an odd integer and primitive -th root of unity , one associates a -dimensional TQFT . Blanchet, Habegger, Masbaum, and Vogel [4] gave a skein-theoretic framework for this -TQFT, and its main properties are the following:
-
1)
For a closed oriented surface , is a finite dimensional vector space over with the natural Hermitian form. For a disjoint union , we have .
-
2)
For an oriented closed 3-manifold , is a topological invariant.
-
3)
For an oriented compact 3-manifold with boundary , is a vector.
-
4)
For a cobordism , is a linear map.
In [4], the authors also give explicit bases for any surface. However, we will focus on , which can be considered as a quotient of the Kauffman skein module of the genus handlebody .
We begin by coloring the core by the -th Jones-Wenzl idempotent. This gives a family of elements of the Kauffman skein module of the solid torus. However, there are only finitely many Jones-Wenzl idempotents for a given odd and -th root of unity , namely [4]. We can consider these ’s as elements of the quotient , giving us a basis.
Theorem 2.1 ([4], Theorem 4.10).
Let . Then the family is an orthonormal basis for . Moreover, the relation holds for .
The second part of the theorem implies that is just a reordering of the basis .
2.2 The Cabling Formula
Here, we will give an explicit description for the Reshetikhin–Turaev invariants of the torus knot cable spaces.
Let and be coprime integers where , and let be the -cable space . These spaces are Seifert-fibered and therefore have simplicial volume zero. For , we extend the vectors to all in the following way. Let for any , and let for any . Note this means that .
Regarding the cable space as a cobordism between tori, the Reshetikhin–Turaev -TQFT gives a linear map
The map sends the element to the element of corresponding to a -torus knot embedded in the solid torus and colored by the -th Jones-Wenzl idempotent. Morton [15] gives the following formula for the image of the basis elements under .
Theorem 2.2 ([15], Section 3, Cabling Formula).
where is the set
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 , 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 be an odd integer, and let be a primitive -th root of unity. Then for a compact oriented manifold with toroidal boundary,
where is the natural Hermitian norm on
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 be a compact oriented -manifold, with empty or toroidal boundary.
-
1)
If is a Seifert manifold, then there exist constants and such that for any odd , we have and .
-
2)
If is a Dehn-filling of , then and .
-
3)
If is obtained by gluing two -manifolds along a torus boundary component, then and .
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 is odd and was proven. For convenience, we will restate the main theorem.
Theorem 1.2.
Let be a manifold with toroidal boundary, let be coprime integers with , and let be an odd integer coprime to . Suppose is a -cable of . Then there exists a constant and natural number such that
We will now assume Theorem 1.7, which we also restate for convenience.
Theorem 1.7.
Let be coprime to some positive integer . Then is invertible if and only if and are coprime. Moreover, the operator norm grows at most polynomially.
Proof of Theorem 1.2.
As mentioned previously, the case when is odd and was shown by Detcherry [6], and our approach follows closely in structure. We let be a manifold with toroidal boundary, an integer, an integer coprime to , odd and coprime to , and a -cable of . We will proceed to prove Theorem 1.2 by showing the upper inequality of
followed by the lower inequality, where and . To obtain the upper inequality, we first remark that . By Theorem 2.4, this implies that
Since is a Seifert manifold, we have that
for some and also by Theorem 2.4. This leads to the upper inequality
For the lower inequality, we will use Theorem 1.7. From the properties of the Reshetikhin–Turaev -TQFT, we consider the linear map
If only has one boundary component, then
by the properties of a TQFT. If has other boundary components, then the invariant associated to any coloring of the other boundary components may be computed as
By the invertibility of from Theorem 1.7, we have the inequality
where is the norm induced by the Hermitian form of the TQFT and is the operator norm. Since the operator norm grows at most polynomially by Theorem 1.7, we obtain the inequality
for some and . 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
for some and . This leads to
where and . ∎
Corollary 1.5.
Proof.
By Theorem 2.4 Part , , and thus by Theorem 2.4 Part , . Since , the limit exists, and any subsequence also converges to . By Theorem 1.2 along odd ,
where
is the limit superior of the subsequence along which and are coprime.
Since
we have
where the final equality follows from the fact that the simplicial volume does not change under attaching a -cable space. ∎
Corollary 1.6.
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 be coprime to some positive integer . Then is invertible if and only if and are coprime. Moreover, the operator norm 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 admits a simpler expression.
Define to be the reduction of under the quotient induced by the symmetries for any and for any . Note that for each , for some non-negative integers where . This means that up to sign, these symmetries imply
(1) | |||||
(2) |
Finally, define , and let be the -matrix with columns corresponding to the reduced vectors , for . In particular, corresponds to of , and the rows of correspond to the original orthonormal basis spanning .
Remark 4.1.
We note that , , and are also dependent on and , but these dependencies are suppressed to avoid unwieldy notation.
The following proposition will be used to prove Theorem 1.7.
Proposition 4.2.
Let be coprime to . Then is a change of basis from and the operator norm grows at most polynomially in . Moreover, for ,
(3) |
where for odd and for even .
The idea of the proof is to leverage symmetric properties of the to give a presentation of and bound its operator norm. The assumption that is necessary for invertibility, as indicated by the following proposition.
Proposition 4.3.
Suppose is odd and not coprime to . Then is singular.
4.1 Proof of Theorem 1.7
Proof of Theorem 1.7.
We begin with the necessary condition. Suppose . Then there are coprime such that and . We claim that of consists of only zeros for each . Suppose some , where , reduces to . Then by Equation (1), , which means , which is a contradiction. Similarly, if , where , reduces to . Then by Equation (2), , which is also a contradiction. This means that , thus is singular.
For sufficiency, suppose . By Proposition 4.2, we can write as a product of two diagonal matrices with an upper-triangular matrix and the change of basis :
Note that the columns of the middle upper triangular matrix correspond to the index sets of the sum in Equation (3). Inverting this product, we have
By Proposition 4.2, grows at most polynomially in , so it is bounded polynomially in . 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 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 for any and for any , we may extend the definition of to all using the following symmetries:
-
•
for any ,
-
•
for any , and
-
•
.
The following Lemma will be used to present .
Lemma 4.4.
Let be coprime to , and let be the inverse of modulo . Then for ,
(4) |
Moreover, for , if and only if .
Proof.
Since , there is a unique such that . Using the symmetries of and substituting in , we have
(5) |
We can then apply Equation (5) iteratively to express the ’s in terms of the ’s.
for . When , by definition, When , Equation (5) yields that For any , the iterative use of Equation (5) to express terminates when the final term is a scalar multiple of either or , depending on the parity of .
For the final statement, note that . This means there is a group isomorphism between the cyclic groups and sending the indices in Equation (4) to distinct for . Since are distinct for , this shows that if and only if . ∎
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 is coprime to . Then
where such that for .
Proof of Proposition 4.2.
It suffices to show that is nonsingular, in which case corresponds to the basis transformation . To establish nonsingularity, we will give a presentation of by expressing , for , in terms of where .
By Lemma 4.4, each , for , can be written in terms of where . These ’s reduce to ’s, where , using the above symmetries. This means that of dimension is contained in , a vector space of dimension at most .
Lemma 4.5 implies that , which means that
Since is a basis for the -dimensional vector space , is a set of vectors, and , then
From this, we conclude that is also a basis for . Since is a basis, this implies that is a change-of-basis matrix and is nonsingular, therefore, is nonsingular.
In order to bound the operator norm , we study the presentation of more closely. By Lemma 4.4, we may express each as
where is either zero or a root of unity and the summands correspond to the reduction of each index modulo . We remark that since is either zero or a root of unity, . Now after applying the symmetry for any , we may express as
where and . Additionally, by Lemma 4.5, we know that the coefficient of any summand of in terms of the basis is with . This means we may write
where Note that since and , we have
Hence every entry of has modulus bounded above by . For any complex unit vector such that for , the Cauchy-Schwartz inequality implies that
This shows that
so the operator norm is bounded polynomially.
Lastly, by the Cabling Formula in Theorem 2.2 and the definition of , the coefficient of in is given by
where for odd and for even . ∎
In order to prove Proposition 4.3, we establish the following definitions.
For , define . Observe that for . In addition, define to be the quotient of under the symmetries for any and for any . We will use the convention that and .
Recall that for each , for some non-negative integers where . This means that up to sign,
(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 , then there are coprime and such that and . We claim that of consists of only zeros for each . Suppose some , then . This implies that which is a contradiction. Similarly, if , then which is also a contradiction. This means that , thus 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.
Recall is the -matrix with columns corresponding to . We also define to be the -matrix obtained by appending the column corresponding to to . The following technical lemmas will be used in the proof of Lemma 4.5.
Lemma 4.7.
Suppose is coprime to , and let be the multiplicative inverse of in the ring . Then
-
(i)
Each column of has at most two nonzero entries. Moreover, for each column with two nonzero entries, their corresponding row indices differ by at most .
-
(ii)
Let
Then in , and where is a root of unity. Moreover, every other column of has exactly two nonzero entries which are roots of unity.
Proof.
Part : Each column of corresponds to the reduced vector , . Since is a linear combination of at most two vectors in , there are at most two nonzero entries in .
Now suppose the index of is , where . Then the index of is , where either (for ) or (for ). We split into cases:
-
•
If , then , , and .
-
•
If , then , , and . This implies that . Since and are coprime, , for some . However, if , we have , which is a contradiction. Thus , so , corresponding to .
-
•
If , then .
-
•
If , then and .
-
•
If , then and .
-
•
If , then .
This implies that the row indices of the nonzero entries in each column differ by at most two for every column except . In particular, the only case where the row indices differ by exactly occurs when .
Part : Since , we have for . Note that has exactly one nonzero entry if and only if one of the following occurs:
-
(1)
Either and are equal or opposite modulo .
-
(2)
Either or vanishes modulo .
Case occurs if and only if , corresponding to . In this case, .
Case occurs if and only if either or modulo . Define
Note that if vanishes, then . Define to be the coefficient of the vector obtained from Equation (3). This means that is the unique column with exactly one nonzero entry except for .
Finally, the conclusion follows from the uniqueness of and Part . ∎
The second technical lemma makes use of Lemma 4.7 in its proof.
Lemma 4.8.
Suppose is coprime to . Then
-
(i)
Each row of has exactly two nonzero entries.
-
(ii)
There is a unique , , such that , where are roots of unity.
The following lemma will be useful in the proof of Lemma 4.8.
Lemma 4.9.
Suppose is coprime to , and let and . Then for with ,
-
(i)
, , and do not have integer solutions,
-
(ii)
, , and may each have integer solutions.
Proof.
Note and encode the two families of indices of the reduced vectors given in Equation (6). There are six equations relating pairs of expressions in .
Part : This follows from the fact that and the bounds on and . We show the case and note that the other two cases follow analogously. Assume for distinct and that . This implies
Since and , must have a nontrivial factor of , which contradicts the bounds on and .
Part : We have the following:
-
•
if and only if ,
-
•
if and only if , and
-
•
if and only if .
All three of these equations may have integer solutions for . ∎
Proof of Lemma 4.8.
Part : It is a corollary of Lemma 4.9 that every row of has at most two nonzero entries. In particular, let be an integral solution to one of the equations of Lemma 4.9 Part . Suppose is such that and are both solutions to equations in Lemma 4.9 Part . Then by Lemma 4.9 Part , either or .
Note that by Lemma 4.7 Part , has exactly nonzero entries since there are 2 in each column other than and , which each have exactly 1. This means that every row of must have exactly 2 nonzero entries.
Part : In the proof of Lemma 4.7 Part , we saw that the only value of corresponding to a column with the nonzero row entry indices differing by 1 is . By Part , of has exactly 2 nonzero entries. This implies that there are some such that has nonzero entries in and and has nonzero entries in and . Take . Finally, define and to be the coefficients of the vectors and defined by Equation (3), respectively. Note that if , and has only 1 nonzero entry. ∎
Lastly, we are ready to prove Lemma 4.5.
Proof of Lemma 4.5.
The last column of the matrix represents the reduced vector written in terms of the basis . We will prove Lemma 4.5 by showing that can be written as a linear combination of the first columns. From this linear combination, we will see that the coefficients will have the required bounds from the statement.
We claim that of can be written as a linear combination of elements in . From Lemma 4.7 Part , if , then has exactly one nonzero entry, and if , then has exactly two nonzero entries.
Case 1:
We first consider the case . Here, the nonzero entry of lies in . This implies that has a scalar of as a summand. By Lemma 4.8 Part , we know that there is exactly one other nonzero entry in in some column . From the argument of Lemma 4.7 Part , there exists a nonzero entry in of . Lemma 4.8 Part implies there exists a nonzero entry in some column and . From the argument of Lemma 4.7 Part , there exists a nonzero entry in of . Again, we pick the other nonzero entry of which lies in some column . Note that 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 or , depending on the parity of .
If is even, by Lemma 4.8 Part , the next corresponding row with a nonzero entry will be where is odd. Similarly, if is even, by Lemma 4.8 Part , the next corresponding row with a nonzero entry will be where is odd. Now when we continue the algorithm, our subsequent row indices will be odd and decrease by until we reach . By Lemma 4.8 Part and Lemma 4.7 Part , there exists a nonzero entry in of , and it is the only nonzero entry in . Since every entry of our matrix is a root of unity by Lemma 4.7 Part and terminates at , scalars by roots of unity of the columns appearing in our sequence gives as a linear combination of elements of where all coefficients are roots of unity.
Case 2:
Now suppose has exactly two nonzero entries. We denote the row indices of these entries by and , where . By Lemma 4.8 Part , has another nonzero entry in some other column . Similarly, has another nonzero entry in some column . We make the following claim, which we prove at the end.
Claim: .
We will proceed similarly to the first case. Consider the column , which has exactly two nonzero entries and cannot correspond to either or since . By Lemma 4.8 Part and the claim, there exists another nonzero entry in some of such that . The case when corresponds to , and the case when corresponds to . We now implement the same argument as the case with one entry in . Note that, in this procedure, we do not utilize any rows with index less than with the same parity as . If and , they will have different parities. In the other case, will have the parity of until we have a such that . This implies that for all , will have opposite parity to .
We now follow the same algorithm beginning with . By the claim, the indices of our subsequent rows must be decreasing. Otherwise, this would contradict Lemma 4.8 Part .
Since both cases in total utilize every row exactly once, is given by a linear combination of elements of where, by Lemma 4.7 Part , all coefficients are roots of unity.
Proof of Claim:
It now suffices to prove that . By contradiction, let us assume that , and we will denote .
If , then either or . If , then since , we will have two columns with nonzero entries in the last two rows. This contradicts Lemma 4.8 Part , which states that there is a unique such column. If , then there are two distinct columns with nonzero entries in and . By Lemma 4.8 Part , there must exist a different column with nonzero entries in and , which contradicts there being at most entries in .
If , then , and there are no other columns with nonzero entries in besides and . By Lemma 4.8 Part , there must exist a different column with nonzero entries in and , which contradicts there being at most entries in .
In the general case, we assume , and we will define . Since , and already have two nonzero entries. Since , these entries cannot be in either or since they only have entries in the first two rows. This implies that the columns which correspond to nonzero entries in must have exactly two nonzero entries in some columns and such that . Since has two nonzero entries in and , the other nonzero entries in and must be in some , where .
-
•
If , we have and . Here, we reach the same contradiction as when and .
-
•
If , then and . This gives the same contradiction as when and .
-
•
If with and , then our argument is the same as when and .
Finally, we consider when and . 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 on the cable space as well as a polynomial bound on its operator norm. The same methodology could apply in the context for the operator 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 where is a orientable compact genus surface with boundary components. Now let such that is a collection of vertical fibers in . We define the Seifert cable space where to be the manifold obtained by performing -Dehn surgery along the -th vertical fiber in .
If an analogous result to Theorem 1.7 holds for the Seifert cable space , 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 and 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 invertible when is sufficiently large and coprime to every ?
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 -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 -manifolds. Mem. Amer. Math. Soc., 21(220):viii+192, 1979.
- [11] K. Johannson. Homotopy equivalences of -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 . 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 -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 -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 -manifolds and quantum -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]