Genus bounds for twisted quantum invariants
Abstract.
By twisted quantum invariants we mean polynomial invariants of knots in the three-sphere endowed with a representation of the fundamental group into the automorphism group of a Hopf algebra . These are obtained by the Reshetikhin-Turaev construction extended to the -twisted Drinfeld double of , provided is finite dimensional and -graded.
We show that the degree of these polynomials is bounded above by where is the Seifert genus of a knot and is the top degree of the Hopf algebra. When is an exterior algebra, our theorem recovers Friedl and Kim’s genus bounds for twisted Alexander polynomials. When is the Borel part of restricted quantum at an even root of unity, we show that our invariant is the ADO invariant, therefore giving new genus bounds for these invariants.
1. Introduction
1.1. Background
Quantum invariants, as developed by Reshetikhin and Turaev [29, 30], are topological invariants of knots and 3-manifolds built from the representation theory of quantum groups, or more generally, the theory of braided monoidal categories. The Jones polynomial is the special case when the quantum group is that associated to the Lie algebra . When the braided categories involved are semisimple, or more precisely modular, these invariants can be nicely packaged into what is called a 3-dimensional topological quantum field theory (TQFT), a mathematical notion that formalizes the various cut-and-paste properties of quantum invariants [32]. As the name suggests, these objects originate from physics, namely from Witten’s interpretation of the Jones polynomial in terms of quantum field theory [37].
When the parameter of the quantum group is a root of unity, the corresponding representation category becomes non-semisimple and one can find continuous families of simple modules, such as Verma modules with complex highest weights. When this gives rise to a polynomial invariant, called the ADO invariant, after Akutsu-Deguchi-Ohtsuki [1]. These “non-semisimple quantum invariants” have been much less studied than their semisimple counterparts though, partially because extending these to 3-manifolds and eventually TQFTs is much more complicated and for more than 20 years they lacked a clear physical interpretation. Some of these hurdles have been overcome in the work of Geer-Patureau-Mirand and collaborators [13, 6], leading to TQFTs for ADO invariants with interesting topological features [5]. Moreover, connections between non-semisimple invariants and the physics of vertex operator algebras and logarithmic conformal field theory have recently been found [15, 7].
However, even though quantum invariants have nice cut-and-paste (TQFT) properties and physically interesting interpretations, their topological content remains mysterious. For instance, the Jones polynomial (as well as HOMFLY and Kauffman polynomials) is not clearly related to the Seifert genus of a knot, fibredness or sliceness. This is in sharp contrast to twisted Alexander polynomials (or equivalently, twisted Reidemeister torsion), which are related to all these topological properties [11] and even to hyperbolic geometry [27]. Relations between colored Jones polynomials and hyperbolic geometry or -character varieties are still major open conjectures in the field.
In order to find which aspects of the theory of braided monoidal categories capture interesting topology of knot complements, a first natural question is: which “structure” behind the Alexander polynomial and Reidemeister torsion is making it capture so much topological information? The answer is very simple: these are invariants of a covering space of the knot complement , or in other words, invariants of a pair where is a homomorphism into a group . For instance, the Alexander polynomial comes from the covering space corresponding to the projection . Most applications depend on these invariants being polynomials, for which one needs , in particular, is an infinite group.
It turns out that there is an extension of the theory of quantum invariants, due to Turaev [31], that produces invariants of such pairs . This extension starts with a -crossed braided monoidal category, essentially a -graded category with a -action and a compatible braiding, and produces invariants of tangles endowed with a representation , where . The -action condition turns out to pose a severe restriction: a semisimple monoidal category has only finitely many tensor autoequivalences up to isomorphism [8]. In light of this, we believe it is natural to consider non-semisimple -braided monoidal categories where is some infinite group.
In previous work, the first author [23] considered a special class of such categories: relative Drinfeld centers of crossed products where is a finite dimensional Hopf algebra and its group of Hopf algebra automorphisms, or equivalently, -twisted Drinfeld doubles 111From now on we refer to this simply as a twisted Drinfeld double as in [34]. Note that the term “twisted Drinfeld double” has another common meaning in the literature. as defined by Virelizier [34]. The reason to use such categories is that the twisted quantum invariants they define can be seen as “Fox calculus deformations” of the invariants coming from the usual Drinfeld double so, in some sense, they do retain some covering space theory. Moreover, it was shown in [23, 22] that the -twisted Reidemeister torsion of knot complements is obtained from the twisted Drinfeld double of an exterior algebra. Thus, one may expect that these quantum invariants contain topological information generalizing that of torsion.
In this work, we show that the twisted quantum invariants of [23] do indeed contain easily readable topological information about knot complements, namely, lower bounds to the Seifert genus. Moreover, we show that these invariants contain the ADO invariants, therefore giving genus bounds for these.
1.2. Main theorem
To state our main theorem, we briefly recall the construction of [23]. Let be a finite dimensional Hopf algebra over a field such that the Drinfeld double is ribbon. The twisted Drinfeld double of is a family, indexed by , of deformations of the usual Drinfeld double of [34]. If is a framed, oriented, -component open tangle, the twisted Drinfeld double leads to a deformation
of the usual universal invariant of tangles (see e.g. [16]) that depends on a homomorphism . This invariant is essentially a special case of Turaev’s construction [31]. Now, if in addition is -graded, there is a canonical automorphism , where and , defined by for homogeneous of degree . Then any as above extends to . This “degree twisted” extension generalizes a similar procedure in twisted Reidemeister torsion theory (e.g. [11]). We thus get an invariant
defined for any (automorphisms preserving the degree and fixing the left cointegral of ). If is a framed, oriented, -tangle whose closure is a knot , then we define
where is a normalization factor making this an invariant of the unframed . If , is a polynomial invariant of the oriented knot with no additional structure, but one still has to think that there is a canonical abelian representation sending the oriented meridian to .
If where and , we define as . We denote where is the component of degree of .
Theorem 1.
(Genus bound) Let be a -graded Hopf algebra of finite dimension. Let be a knot and a homomorphism. Then
where is the Seifert genus of .
In fact the proof shows that a similar bound holds for . As mentioned previously, the motivation for this theorem comes from the Fox calculus formulas of [22, 23]. The actual proof given here makes no mention to Fox calculus, it is instead based on an argument of Habiro [16] but generalized to include the representation , which is essential for our theorem. We illustrate the theorem directly for (even) twist knots, which have genus one, in Subsection 3.7. Note that this theorem is in contrast with the Jones polynomial whose degree coincides with the crossing number for twist knots (since they are alternating), hence the degrees are not bounded.
We now list various special cases of our theorem.
1.3. Twisted Reidemeister torsion
When is an exterior algebra (with concentrated in degree one), one has , and coincides with the -twisted relative Reidemeister torsion where is a meridian [23]. This torsion is essentially equivalent to the twisted Alexander polynomial of Lin and Wada [19, 35]. The above theorem implies
Since this is equivalent to which is a special case of a result of Friedl and Kim [10].
1.4. ADO invariants
Now let be the Borel part of the restricted quantum group at a -th root of unity . Then has finite dimension, it is -graded (with ) and is ribbon, so we have a polynomial invariant On the other hand, the unrolled restricted quantum group leads to the ADO invariant [1]. We use the normalization of the ADO invariant in which .
Theorem 2.
We have
hence
As an example, the invariants of the first twist knots (denoted by their usual names in knot tables) are given as follows:
These all have degree , since for these knots, this matches with our theorem. Formulas for ADO invariants of double twist knots (genus 1) and torus knots (genus ) are given in [3]. All these satisfy the genus bound above.
1.5. The higher rank case
More generally, for a simple Lie algebra of rank and a primitive root of unity of order (say, coprime to the determinant of the Cartan matrix), our construction provides a polynomial invariant that gives a lower bound to the genus. For instance
which generalizes the above bound for ADO to all . The invariant is obtained by letting be the Borel part of the corresponding restricted quantum group (which is -graded) in our construction. We expect that this invariant recovers other “ADO-like” invariants defined in the literature, e.g. [17], [4], thus implying genus bounds for all of them. For instance, the -ADO invariant at (so ) defined in [17] is given as follows for the above twist knots:
Our genus bound implies , which is satisfied by the above polynomials. This will be studied in a separate paper.
1.6. Related results
In [25], Ohtsuki proved a genus bound for the 2-loop expansion of the Kontsevich integral, which is a 2-variable knot polynomial . No genus bound seems to be known for higher-loop expansions. The colored Jones polynomial is a specialization of the Kontsevich integral, so Ohtsuki’s theorem implies a genus bound for the 2-loop part of this expansion as well. However, this is a limited bound and no further results seem to be known for colored Jones polynomials. Note that Ohtsuki’s invariant satisfies so it vanishes on amphichiral knots, on the other hand, ADO invariants do not vanish on amphichiral knots. Now, the family of colored Jones polynomials turns out to be equivalent to the family of polynomials (for varying ) by a theorem of Willetts [36]. Thus, our theorem shows that non-semisimple quantum knot invariants contain more transparent topological information than their semisimple counterparts.
In the case of , Bar-Natan and the second author studied the -loop polynomial using -adic Hopf algebra techniques [2]. This gives another proof of Ohtsuki’s genus bound. This approach has the advantage of being polynomial-time computable but a genus bound was only given in the 2-loop case. Note that Seifert genus is in NP and co-NP [18].
Finally, we note that there already exist topological invariants detecting the Seifert genus. On the one hand, a theorem of Friedl and Vidussi [12] states that there is always some and a representation for which . The authors do not know whether there is an algorithm to find such a though. On the other hand, knot Floer homology, which categorifies the Alexander polynomial, detects the Seifert genus [26]. It is an interesting question whether one can actually detect the Seifert genus of a knot using Hopf algebra/representation theoretic techniques, e.g. by keeping and varying in our theorem.
1.7. Plan of the paper
We begin in Section 2 with some Hopf algebra preliminaries. Here we recall some definitions from Hopf -coalgebras, we define twisted Drinfeld doubles and study ribbon elements in these. In Section 3 we define the invariant , the polynomial invariant , and we prove some properties of these invariants. We illustrate our main theorem with the case of twist knots in Subsection 3.7. In Section 4 we prove our genus bound (Theorem 1). Finally, in Section 5, we study the case where the Hopf algebra is the above and prove Theorem 2.
2. Hopf algebra preliminaries
For simplicity, we work over an algebraically closed field of characteristic . In all that follows, denotes a group. Hopf algebras will be assumed finite dimensional unless otherwise stated.
2.1. Hopf algebras
For basic definitions on Hopf algebras, see e.g. [28]. We denote multiplication, coproduct, unit, counit and antipode of a Hopf algebra over by respectively. We will employ Sweedler’s notation for the coproduct of a Hopf algebra, that is,
etc. The dual is also a Hopf algebra with
for each and . Here denotes the usual pairing and for any . We denote by the group of Hopf algebra automorphisms of and the subgroup of fixing a non-zero cointegral (either left or right, see Subsection 2.4 below for definitions).
A Hopf algebra is -graded if with , and for each . We denote by the subgroup of automorphisms of preserving the -degree.
2.2. Hopf -coalgebras
A Hopf -coalgebra is a family where each is an algebra with unit together with a family of algebra morphisms for each , an algebra morphism and algebra antiautomorphisms for each satisfying graded versions of the Hopf algebra axioms, see [31, 33] for more details. Note that is a Hopf algebra in the usual sense (with the coproduct and antipode ).
A Hopf -coalgebra is said to be crossed if it comes with a family of algebra isomorphisms , simply denoted , preserving the Hopf structure and satisfying .
A crossed Hopf -coalgebra is quasi-triangular if it comes with a family of invertible elements in , satisfying
-
(1)
is -invariant, that is, ,
-
(2)
where denotes permutation of two factors (with a sign in the super-case),
-
(3)
,
-
(4)
for each , where we used the notation for any and similarly for (). If is quasi-triangular, the graded Drinfeld element is defined by
A quasi-triangular Hopf -coalgebra is -ribbon if it comes with a family of invertible elements satisfying, for each :
-
(1)
,
-
(2)
,
-
(3)
for all ,
-
(4)
.
If , we call the -pivot of , this satisfies for each and is group-like in the sense that .
2.3. Twisted Drinfeld doubles
We now define the twisted Drinfeld double of a Hopf algebra , which is a quasi-triangular Hopf -coalgebra extending the Drinfeld double construction. This was introduced by Virelizier in [34], here we follow the conventions of [23] (see remark 2.1 below).
For each we define an algebra as follows: as a vector space and we define the product by
where . This is an associative algebra with unit . For each we define a coproduct by
For each we define an antipode by
For each define an algebra isomorphism by
Finally, for each let
where is any basis of and is the dual basis. With all these structure maps, has the structure of a quasi-triangular crossed Hopf -coalgebra. We denote and call it the twisted Drinfeld double of . If is a group homomorphism, then is a quasi-triangular Hopf -coalgebra in an obvious way, we call it a -twisted Drinfeld double.
2.4. Ribbon elements in the double
We will now establish under which conditions the twisted Drinfeld double has a graded ribbon element. For this, we briefly recall what happens for the usual Drinfeld double, for more details see [28].
Recall first that a left cointegral of a Hopf algebra is an element such that for each . Dually, a right integral is such that for each . As a consequence of uniqueness of integrals/cointegrals, there are unique group-likes satisfying
for all . These are the distinguished group-likes of .
Now, a theorem of Kauffman and Radford states that the Drinfeld double is ribbon if and only if there are group-likes such that and
In such a case, the ribbon element is where is the Drinfeld element of .
We can now give sufficient conditions for to be ribbon (or rather, a restriction to a certain subgroup ). Let’s suppose that satisfies the above condition making ribbon. Let be a subgroup fixing such . Let be the homomorphism characterized by (this is defined by uniqueness of cointegrals). Then, as shown in [23, Prop. 2.4], is -ribbon if and only if the homomorphism has a square root. In such a case, the -ribbon element is given by
for any , where is the ribbon element of . The -pivot is given by
The square root condition is immediately satisfied if .
2.5. Trivializing the crossing
Suppose that is an abelian group with an homomorphism and consider as defined above. Then acts on each by and setting we get another Hopf -coalgebra . The Hopf structure is extended to by declaring the elements of to be group-likes. In what follows we denote by when considered as an element of so that in for each . It is easy to see that
is an -matrix for in (almost) the usual sense, that is, it satisfies
In other words, the category is a braided category in the usual sense. It is not difficult to see that the Drinfeld element of is determined by where is the Drinfeld element of . If is a ribbon element in , then is ribbon with . It follows that the pivot of is the same as in .
2.6. The super algebra case
Recall that a super vector space is a vector space with a mod 2 grading, that is, a decomposition . The category of super vector spaces is symmetric with where denotes the mod 2 degree of homogeneous elements . A super Hopf algebra is a Hopf algebra in the category of super vector spaces. This amounts to the same axioms as for a Hopf algebra, except that the coproduct satisfies
More generally, we can talk about super Hopf -coalgebras, now it is that satisfies the above property (with in place of ). In the super case, denotes the automorphisms preserving the mod 2 degree. The twisted Drinfeld double of a super Hopf algebra is defined as before, but with some additional signs:
In this formula, represents the usual vector space pairing, but note that the sign comes from the fact that the right evaluation/coevaluation of the category of super vector spaces is not the same as that of vector spaces.
3. Twisted quantum invariants of -tangles
During this section, we let be a finite dimensional Hopf algebra. We suppose is ribbon with corresponding pivotal element and we let be such that is -ribbon. We let be a basis of and be the dual basis of .

3.1. -tangles
Let and let be a basepoint in . We think of as being on the plane of the page, so the -axis is transversal to the plane and the negative -axis is towards the reader. We also denote for . By a -tangle we mean a framed, oriented tangle in endowed with a representation where . Two -tangles and are isotopic if there is a basepoint-preserving isotopy , that is, as framed oriented tangles, and . Here is the induced map. By Van Kampen’s theorem, if and are -tangles satisfying that (where is the homomorphism induced by the corresponding inclusion) then we can stack on top of and rescale to define a new -tangle, called the composition of and . Similarly, we can stack a -tangle to the right of another. With these operations, -tangles form (the morphisms of) a monoidal category, and in fact, a -crossed ribbon category [31].
3.2. Twisted universal invariants of -tangles
Let be a -tangle with open components (and no closed components). We suppose the components of are ordered, say . We define now the invariant where each and is a meridian at the endpoint of as defined below. As usual, this invariant is defined from a planar diagram and then shown to be independent of it.
Let be an oriented planar diagram of where at all crossings both strands are oriented upwards. We assume comes with the blackboard framing. As usual in drawing a knot diagram we cut the projection close to each crossing to indicate where the knot passes under another strand. The resulting connected components in the plane are called the edges (arcs) of the diagram.
We define first meridians and “partial longitudes” associated to the edges of a diagram. For each edge of , there is an element defined as the homotopy class of the loop that starts at the basepoint , goes to the given edge via a linear path, encircles that edge once with linking number -1 and finally goes back to the basepoint by the same linear path. Then , so that all edges of the diagram are labelled by elements of :

For each we let where is the edge containing the endpoint of and we let . Now let be the core of , that is, as a framed tangle and let be a close oriented normal to (this uses the framing of ). Then, for each edge of the diagram of , we let be the “partial longitude” defined as the homotopy class of a path that goes linearly from to the endpoint of , follows with the opposite orientation until it reaches , and finally goes back to the basepoint by a linear path:

Note that for each edge of we have that . For a component of we denote where is the edge of containing the starting point of . Note that both and are independent of the diagram of .
We now bring in the twisted Drinfeld double. To each positive crossing of whose bottom edges are labelled by (as in Figure 2 above), we associate the -matrix and the crossing isomorphism . We represent the two factors of the -matrix in the diagram by two black beads placed at the crossing, the first tensor factor on the overpass and the second on the underpass, while is represented by a white bead:
![[Uncaptioned image]](https://cdn.awesomepapers.org/papers/567a3e72-7ea8-4978-971c-e49f3f4b921c/crossing.png)
If the crossing is negative and are the labels of the edges at the top, we assign the inverse -matrix and the crossing isomorphism . These are again represented by two black and one white beads as above. To caps and cups labelled by we associate the unit or the -pivot as follows:
![[Uncaptioned image]](https://cdn.awesomepapers.org/papers/567a3e72-7ea8-4978-971c-e49f3f4b921c/pivots.png)
where . With these conventions, the black beads lying over a given edge of the diagram belong to . Hence, we can multiply all these black beads (from right to left) as we follow the orientation along . Thus, we get a single black bead for each edge of the diagram. We now follow the orientation of a component of and we multiply all these black beads as follows: if an edge follows an edge and the label of the overpass separating these is , then there is a white bead in between the black beads of and (where is the sign of the crossing). We evaluate on the bead and then we multiply this with the bead , this results in . In other words, we slide the black beads following the orientation of the diagram, and whenever a black bead crosses a white bead, we evaluate the corresponding crossing isomorphism on the black bead. Note that the product of all the white beads on a given component of lying after an edge is equal to . Thus, we could equally evaluate for each edge , which belongs to for each , and then multiply all the resulting beads on a given component (from right to left as we follow the orientation of ). If we do this for all components of we get an element
This turns out to be independent of the diagram chosen, so it is a topological invariant of , denoted . This is the invariant of -tangles defined in [23], we call it a twisted universal quantum invariant.
More precisely, and to be careful with the signs in the super-case, we define the invariant as follows: let be the set consisting of the overpasses and underpasses of the diagram together with the right caps and cups. Suppose has crossings and right caps or cups, so . Order the set of crossings of the diagram arbitrarily, say . Order the set of right caps/cups arbitrarily too. This determines a total order of where the overpass of each crossing comes before the underpass and all caps/cups come after the crossings. We denote the set with this order. Let be the labels of the bottom edges (resp. top edges) at if the crossing is positive (resp. negative). Let be the label at the -th right cap/cup for and let be the corresponding power of . For each , let be , or depending on whether is an overpass, an underpass or a cap/cup. Now, following the orientation of each component of the diagram, along with the order of the components of , defines another total order on , denote it . In other words, if is the number of elements of corresponding to , then the first elements of are those of (ordered from right to left as we follow the orientation of ), the next elements are those of ordered in the same way, and so forth. This gives a permutation isomorphism . If is a Hopf group-coalgebra in super-vector spaces, this means that signs are being introduced. Now, if
(note that this belongs to ) then
where denotes iterated multiplication. Here we denote where is the edge of containing and is the partial longitude defined above. Note that the order chosen on the set of crossings and in the set of caps/cups is irrelevant in the super-case, since both the -matrices and the pivots have degree zero.
Remark 3.1.
Our motivation to define for super Hopf algebras is that the -twisted Reidemeister torsion is the special case when is an exterior algebra [23], which is a super Hopf algebra.
3.3. The abelian case
Suppose is an abelian group with an homomorphism . Note that, given a diagram of a -tangle , all edges of will be labelled by the same . Consider and the Hopf -coalgebra defined in Subsection 2.5. Then we can define a universal invariant almost in the “usual sense”: we use the -matrix on crossings (with labels as before) and the pivot of (which is the same as that of ) on caps/cups and we run the above procedure. Every bead over a component of a tangle belongs to the same algebra , so we can multiply all of these directly. Using the semidirect product relation and that we can write this invariant as
(1) |
This is because the product of all white beads over equals .
3.4. The -graded case
In addition to the previous conditions on , suppose that is -graded and let .
Let and . For every and we define an -linear Hopf automorphism of , denoted , by
where is the -degree of . It is easy to see that the map is a homomorphism. In what follows we use the shorthand . Let be the subgroup of automorphisms of the form where . Since , it is easy to see that (where is as in Subsection 2.4) which has a square root
hence is -ribbon by [23, Prop. 2.4]. The pivot element of is
Now let be a tangle as in the previous section and let be an homomorphism. Then extends to a homomorphism defined by
for , where is the homology representation. Note that this depends on the orientation of the components of . When this is simply the abelian representation sending the canonical generators of to . Then the above construction leads to an invariant
In what follows, we identify with in the obvious way.
3.5. Knot polynomials
Suppose is -graded as above. If is a framed, oriented, --tangle whose -closure is a knot , then
depends on only up to conjugation hence it is an invariant of , see [23]. Here is given by for each . The map is not an algebra morphism over if so there is no reason for the above evaluation to be trivial. Note that is itself an invariant of when is abelian [23], in particular is an invariant of . If is the writhe of then
is an -valued invariant of the unframed -knot . Here is shorthand for where is the -degree of the cointegral of . When this simply becomes
(2) |
and this is a polynomial invariant of the unframed, oriented knot (without any additional structure). It is easy to show that this polynomial actually belongs to .
3.6. Elementary properties
In the untwisted case, the universal invariant of the composition of two tangles is obtained by multiplying the beads of the components being composed, doubling a strand corresponds to applying the coproduct, and reversing the orientation corresponds to applying the antipode (see e.g. [16]). These properties extend to our invariant , except that we have to take care of the ’s.
To simplify the statement, we only state the composition rule in the following special case. Suppose is a -tangle having two components with two endpoints in that are adjacent and labelled by the same , see Figure 4. We suppose that, at such endpoints, is oriented downwards and upwards and the endpoint of is to the left of that of . Thus, we can compose with an -labelled left cap on top. Let be the resulting -tangle.

Lemma 3.2.
The invariant is computed by
where is applied on the factors corresponding to and and is the -label of the endpoint of .
Proof.
With the above conventions, the new component in consists of , with its orientation, followed by . The bead of this component is obtained by sliding the bead on along , evaluating on all ’s encountered along , and multiplying with the bead of . Since the product of the ’s along equals , this proves the lemma. ∎
Now let be a component of a -tangle . Let be the tangle obtained by doubling along the framing. Suppose is a -tangle with and let . Let where are the meridians of the two copies of in (say, at the endpoints of the copies of ). Then where is the meridian corresponding to .

Lemma 3.3.
If is constructed by duplicating as above, then
where is applied on the tensor factor corresponding to .
Proof.
We need to check this on the elementary tangles. For the crossing tangles, this follows by the defining property of the graded -matrix. For right caps and cups, it follows from and for left caps/cups it is obvious. The fact that the ’s are Hopf isomorphisms implies that the property holds for compositions of elementary tangles, thus for all tangles. ∎
Lemma 3.4.
Let be a component oriented upwards, that is, it begins on and ends on . Let be the label at the endpoint of . Let be the tangle obtained by reversing the orientation of . If we use the same for both tangles, then
where is applied on the tensor factor corresponding to .
Remark 3.5.
Recall that in our definition of we “gather” the beads at the endpoint of each component and then multiply. If we reverse the orientation of and we gather the beads at the same point as before (which is the starting point of ), then we just need to apply the antipode , but if we want to gather at the endpoint of we need to slide our bead down through all of , this gives the additional above.
Proof.
Suppose first we reverse the understrand of a positive crossing (oriented upwards) whose bottom labels are . Thus, and . Then the invariant is computed as on the left of the following figure:
![[Uncaptioned image]](https://cdn.awesomepapers.org/papers/567a3e72-7ea8-4978-971c-e49f3f4b921c/inverselemma1.png)
For convenience we keep track of the white beads in our pictures. That this equals the picture on the right follows from the following computation:
But the picture on the right is exactly obtained by applying to the strand that is being reversed (note that is simply on ). Similarly, if we reverse the overstrand of a positive crossing, the invariant is computed as on the left of the following figure:
![[Uncaptioned image]](https://cdn.awesomepapers.org/papers/567a3e72-7ea8-4978-971c-e49f3f4b921c/inverselemma2.png)
Here we have and . It is easy to see that the right hand side is obtained by applying to the tangle with the strand oriented upwards. For negative crossings, a similar computation shows that the lemma holds as well. Using the above composition property (and that the antipode is an algebra antiautomorphism), we deduce the lemma for in the case that has no caps or cups.
Now, suppose has caps and cups. We will subdivide (a diagram of) in subarcs without caps/cups. Thus, as we follow the orientation of , we follow first , then a cap, then , then a cup, and so on until . Each with odd is oriented upwards and for even it is oriented downwards. We let be the label at the starting point of , thus, is the label of the endpoint of each and we let and . We also let be the image of the partial longitude from the endpoint of to the top of , and be the partial longitude for , that is, goes from the starting point of to the top of . We also denote , so for each . For each (even or odd) we let be the total bead for that arc as if the arc was oriented upwards (and the bead is gathered at the top so for odd , for even ). Thus, by what was shown above, for even the actual bead along that arc is (if the bead is gathered at the top, see Remark 3.5 above). Then, the tensor factor of can be written as
where is the power of the pivot at the cap/cup right before (so ). Now we apply . First we apply this results in
Applying gives
We used that for each and and that and for odd or for even . Note that for each even (which correspond to caps) and for each odd (corresponding to cups), where is the power of the pivot at the given cap/cup when the orientation is reversed. Therefore, the above equals
which is exactly the bead of with its inverse orientation. ∎
Now let be a component of a tangle as above, but suppose now that both endpoints of lie on and that the endpoint of is to the right of its starting point. Let be defined by
(3) |
Let be the label of the endpoint of . If is obtained by reversing the orientation of then the above lemma implies that
(4) |
When we use the abelian representation one has to be a bit more careful. This representation depends on the orientation of each component, in particular is not equal to . The beads at a negative crossing (of the first type) are given as follows:
![[Uncaptioned image]](https://cdn.awesomepapers.org/papers/567a3e72-7ea8-4978-971c-e49f3f4b921c/thetas.png)
3.7. Twist knots
To illustrate our constructions we will compute the twisted universal invariant explicitly on the family of twist knots. Let be the -twist knot. This is an alternating knot with crossings and genus 1. The knot is the figure-eight knot, (Stevedore’s knot) is drawn on the left of Figure 7. For simplicity, we suppose , that is, is an -graded Hopf algebra. Recall that the pivot of is where is the pivot of (note that for ). We also take (but we keep ). Thus all white beads in our diagrams will be , we will denote these simply by . We will show directly that for any . This is in agreement with the genus bound, Theorem 1, since as mentioned the genus of every twist knot is . In what follows we denote .

Let’s begin with the figure eight knot. We compute from the 2-bridge diagram of the right of Figure 6. We use bridge presentations with minimal number of bridges to reduce the number of Drinfeld products to use, this simplifies considerably the expression for . Note that this diagram has rotation number zero so all the powers of from pivots cancel out, in other words, we can simply put the untwisted pivot on right caps and on right cups, which is what we do. The diagram also has writhe zero so that . In a bridge diagram, other from the ’s on caps/cups, all the beads over an overarc are in and all beads on an underarc are elements of . All white beads, which are , lie on underarcs. We number the crossings from left to right as shown in Figure 6. If represents the -matrix (of ) on the -th crossing, then the invariant is given by (the using Einstein summation convention to hide the eightfold sum):
Here denotes the multiplication of the twisted double . Note that
for any . Thus equals
Here we used that is an algebra morphism over and over (though not over all of ) and that for each . Note that
for each . Thus, writing with and ( have no meaning independently, but does) and using we obtain
Note that , so the maximal/minimal power of in is the maximum/minimum of . But since (this uses that is -graded, not simply -graded) it follows that , hence is a polynomial of degree .
We now consider the case of general . The knot has a 2-bridge presentation as shown on the right of Figure 7, but with full rotations of the “middle band”. This diagram has rotation number and writhe zero as before so we can simply use on right caps/cups. Let be the arcs of the bridge presentation as one follows the orientation of . In Figure 7, is blue and is green. As in the example of the figure-eight knot, the product of all elements along (resp. ) is an element (resp. ). The beads over are all in , except for the pivots (or inverses). As before, for each arc, we multiply the beads encountered along that arc and gather the product at the very end of that arc. The products along have the form for some . Then the invariant is and
As before, so that evaluating on has the effect of killing all the -beads coming from or . The only beads that remain are those of the crossings between and , the ’s on and the pivots ’s on . In other words, except for the pivots, all the remaining beads are shown below (we denote ):
![[Uncaptioned image]](https://cdn.awesomepapers.org/papers/567a3e72-7ea8-4978-971c-e49f3f4b921c/underarc.png)
Order the crossings between and by from left to right. Let be the -component of the -matrix on the -crossing. Then, the product of all beads along the red arc is
where . The product along is
Since for any , this has the form with . Thus
The term . Now, since is -graded the product is zero if , so the only terms that contribute to the above sum are those for which . Since also , it follows that the degree of is as we wanted.

4. Proof of Theorem 1
Let be a finite dimensional -graded Hopf algebra over and let where . In what follows we denote which we see as a subgroup of via and we denote simply by , that is, for . The reader interested on the case of ADO may simply suppose and .
4.1. Formulas from Seifert surfaces
Let be a Seifert surface for and let . We suppose is 0-framed, this is enough for our theorem. After an isotopy, we can suppose is obtained by thickening a -component (framed) tangle all of whose endpoints lie on and attaching a disk to the top as in Figure 8. We suppose this disk comes with a marked point on its boundary and let be the -tangle obtained by opening at that point. We denote the components of by from left to right and we suppose that each is oriented “to the right”, that is, the orientation at the rightmost endpoint is upward pointing. Then, is obtained by first doubling all the components of (with their orientations), reversing the orientation of the right parallel of each component and then composing with various left caps on the top. In the untwisted case, this implies that the universal invariant is obtained from by applying the coproduct to each tensor factor, applying antipodes on the even tensor factors (with a pivot) and finally multiplying:
(5) |
where is the permutation map (with signs in the super case) and . The additional pivot in the antipode comes from the fact that we reversed the orientation of a component of a tangle all of whose endpoints lie on . This idea comes back to Habiro [16]. However, what really makes our theorem work is the appearance of the representation , or rather, the degree twist . The essential observation is that, when is restricted to , cancels out, that is
This is because the orientations of are opposite on each pair of doubled components of . More precisely, let be the loops corresponding to the rightmost endpoints of the double of a component of and be the loop associated to the rightmost endpoint of itself so that . Since in we have and , where , so that . The effect of cancelling out is that, in the twisted version of (5) above, the invariant of has no powers of involved. Indeed, since takes values in and the -matrix of is induced from that of it follows that

Let be the partial longitudes associated to above. Let and so that for each (note that is the label of the endpoint of ). Note that are zero in so that and .
Proposition 4.1.
The twisted universal invariant of is computed as
where
Here denotes iterated multiplication of and is as in (3).
Proof.
This follows immediately from the lemmas of Subsection 3.6 and the above remarks. ∎
Since has no powers of , the above formula implies that all powers of in come from the tensor . Note that this tensor only depends on the genus (and the if is non-trivial). Thus, this already shows that the powers of are controlled. The next section shows how to keep track of the powers of coming from .
4.2. Filtration of
In what follows, if , we introduce the following notation:
-
•
,
-
•
,
-
•
-
•
and ,
-
•
if for each ,
-
•
If is an homogeneous element, , we denote and . Note that and for homogeneous . With this notation we have .
-
•
For every we denote so that .
-
•
Elements of are denoted by
We define a -linear -filtration on each , where , by
We will denote this simply by . In particular, an element of filtration-degree is a polynomial in of degree . We consider higher tensor products with the tensor product filtration.
Lemma 4.2.
The multiplication of is filtration-preserving.
Proof.
Let , say and , where and . Then
Since is -graded, we have . Thus , showing that the above direct sum is contained in as desired. ∎
In contrast to the above lemma, the maps are not filtration-preserving. However, we have the following:
Lemma 4.3.
The map is filtration-preserving.
Proof.
Let , say, with . We have
Note that in the third equality we used that preserves the -degree. By the previous lemma, the rightmost tensor factor of the last equality belongs to . The left tensor factor clearly belongs to so their tensor product belongs to which is contained in the -th term of the same filtration since
∎
4.3. Proof of Theorem 1
Let . We will prove something more general, namely, that the entire universal invariant is a polynomial of degree , after identifying with . Thus, any evaluation, in particular , is a polynomial of degree . To begin with, note that since there are no powers of involved in and is the top -degree of , we have
Recall that and that . Clearly, multiplication by preserves the filtration. Moreover, is filtration-preserving for any . Together with the previous lemmas, it follows that the map
is filtration-preserving (the factor is simply to eliminate the powers of from the -pivots in ). By Proposition 4.1, it follows that and by definition of our filtration, this implies that is a polynomial in of degree . This proves our theorem. ∎
Remark 4.4.
The construction of only requires to be -graded (instead of -graded). Proposition 4.1 is still valid in this case. Thus, we expect a genus bound to hold in this setting too. We prefer to work in the -graded case since the most relevant examples, Borel parts of quantum groups, are -graded.
5. Relation to ADO invariants
Let be the algebra generated by such that (Hopf algebra structure defined below). If is a primitive -th root of unity (with ), which we simply take as , we let be the quotient of by . This is -graded with During this whole section we denote and for any .
5.1. Quantum
Consider the algebra with generators satisfying
and Hopf algebra structure determined by
Since is a central group-like in , the quotient , where is the ideal generated by , is a Hopf algebra. The usual quantum group (at a root of unity) is the subalgebra generated by in this quotient.
5.2. The Drinfeld double of
The Drinfeld double is the quotient of by the relations
see [9]. For convenience, we will denote by the images of in . Let be the quotient of by the ideal generated by . We will also denote by the image in of the generators of .
The Hopf algebra is quasi-triangular with -matrix given by
where and for each . Note that this -matrix is simply the projection in of the usual -matrix of the double . Moreover, has a ribbon element determined by where and is the Drinfeld element (for an explicit expression of see [9]). Note that if and is defined by then satisfies the hypothesis of the Kauffman-Radford theorem (see 2.4). The pivot is the projection in of the pivot .
Note that is the usual braided extension of the restricted quantum group at ( is the Hopf subalgebra of generated by and is not braided)
5.3. The twisted Drinfeld double of
Consider the action of on given by defined by , that is, take in the degree twist action (here for any ). For each , let be the quotient of by the ideal generated by , in particular, is the Hopf algebra defined previously. Then has generators subject to the same relations as in except for the relation which now becomes
(6) |
The coproduct is given by
(7) |
These computations are carried out in [34] (in a slightly different convention, obtained by applying on each to our formulas). The -matrix is
where is as before. This is simply the projection in of the -matrix of defined in Subsection 2.3. The (left) cointegral of is and so that . This has an obvious square root , hence the Hopf group-coalgebra is -ribbon with ribbon element which is the image in of
where is the ribbon element of and the -pivot is
5.4. Relation to unrolled quantum
To relate the above twisted Drinfeld double to the usual quantum group we multiply each by and by . Let (note that there is one on each but we do not include in the notation). In these new generators, relation (6) becomes
and while the other relations among the generators remain unchanged. The coproduct (7) becomes
and the -matrix
Now let be the Hopf group-coalgebra defined in Subsection 2.5. Clearly is central in , hence we mod out by and we keep denoting the quotient by . For each , we denote by the element of corresponding to . Thus, is obtained by adding generators to satisfying
Here denote the above (rescaled) generators of . Then is a quasi-triangular Hopf group-coalgebra with trivial crossing and has an -matrix given by
where is the above -matrix of . The ribbon element becomes and the pivot element is the same as before, see Subsection 2.5.
As an algebra is isomorphic to (a ribbon extension of) , where is the quotient of by . However, the coproduct and the -matrix are different. It turns out that the difference is measured by a Drinfeld twist. In our setting (grading by an abelian group, no group action) a Drinfeld twist in is a collection where each is an invertible element of satisfying
-
(1)
,
-
(2)
.
for each . Drinfeld twisting a quasi-triangular Hopf group-coalgebra produces a new quasi-triangular Hopf group-coalgebra where as an algebra for each , the coproduct is and the -matrix is
where is the usual symmetry (with signs in the super case).
Lemma 5.1.
Let for each . Then is a (graded) Drinfeld twist on .
Proof.
It is clear that is a Drinfeld twist because the ’s are group-like. ∎
Thus, we obtain a new ribbon Hopf group-coalgebra whose coproduct is determined by
and
The antipode of is . The -matrix of is
where
and
Lemma 5.2.
The Drinfeld element, ribbon element and pivot of are the same of .
Proof.
We only check this for the Drinfeld element. Write so that . Then
where we used the above expression for in the third equality. ∎
The Hopf group-coalgebra is isomorphic, as a Hopf group-coalgebra, to (a ribbon extension of) . The latter appears in [24] as an example of “colored” Hopf algebra and is behind the definition of the non-semisimple quantum invariants of knots as we will see below. For the moment, we show the following.
Lemma 5.3.
The -valued and the -valued universal invariants of a knot are equal.
Proof.
It is a general fact that Drinfeld twisting does not changes the resulting quantum invariants. This is obvious if one understands a Drinfeld twist as an equivalence of monoidal categories. In the present case, we can see this directly as follows: let be the braid morphism corresponding to the -th generator of the braid group coming from the -matrix . Let be the one for . Then if it is easy to see that for all , hence for any braid where are the braid operators associated to respectively. Closing the -th strand in has the effect of multiplying with , so closing the last strands is invariant under conjugation by . Since and have the same pivot, it follows that the invariants are equal.
∎
5.5. Verma modules
For each , the Verma module is the -module with basis and action given by
see [21]. This becomes a -module if we set . Since acts by this is a module over . It is a module over if we set . The operator above acts over a tensor product as follows
where is the operator defined by
Thus, the action of over a tensor product is given by
where acts by .
5.6. ADO invariants
We define the ADO invariants following Murakami [21]. Let be a framed oriented knot which is the closure of a framed oriented -tangle . We suppose is given by a diagram with only upward crossings and left/right caps and cups and with the blackboard framing. Consider the above Verma module with the Yang-Baxter operator induced from above (that is where is the swap ), this is the same Yang-Baxter operator used in [21]. Then, to each positive crossing of we associate or its inverse if the crossing is negative. On left caps and cups we simply associate the evaluation/coevaluation of vector spaces. On right caps and cups we use
Pasting all these maps together as determined by the diagram defines a map . For generic , it can be shown that this map is multiplication by a scalar in where is the writhe of the diagram. This is called the ADO invariant of the framed knot and denoted (usually, the ADO invariant is a renormalized version of this that is no longer a polynomial, but since we only consider knots we prefer the unnormalized version given here). It is easily seen that a positive twist acts by multiplication by on . It follows that
is an invariant of the underlying unframed knot , where is the writhe of the diagram. Moreover, this invariant belongs to so setting , this determines a polynomial in which we denote simply by . Note that some authors use instead. This polynomial is not symmetric in , but it is after substituting [20]. Recall from (2) that is defined as
Proposition 5.4.
We have where is the degree twist of .
Proof.
In the above definition of ADO invariants, we used the Yang-Baxter operator . Since the factor also appears on the ribbon element (equivalently, on the action given by a positive twist) it does not affects the invariant of an unframed knot, so we could equally use to define ADO. Using , a positive twist acts by multiplication by . The maps associated to right caps and cups above are exactly given by multiplication by , which is the pivot of . Thus, the universal invariant defined from determines the ADO invariant by
where the second equality is Lemma 5.3. Now, by (1), is of the form
where is the projection (recall that we mod out by ). Note that is evaluated at since is defined through the action , that is, on the degree twist action. Write (rescaled generators of ), where . Using () we get
But and , so . Since also , it follows that . Setting this implies the theorem.
∎
References
- [1] Yasuhiro Akutsu, Tetsuo Deguchi, and Tomotada Ohtsuki, Invariants of colored links, J. Knot Theory Ramifications 1 (1992), no. 2, 161–184.
- [2] Dror Bar-Natan and Roland van der Veen, Perturbed gaussian generating functions for universal knot invariants, arXiv:2109.02057 (2021).
- [3] Anna Beliakova and Kazuhiro Hikami, Non-semisimple invariants and Habiro’s series, Topology and geometry—a collection of essays dedicated to Vladimir G. Turaev, IRMA Lect. Math. Theor. Phys., vol. 33, Eur. Math. Soc., Zürich, [2021] ©2021, pp. 161–174.
- [4] Liudmila Bishler, Overview of knot invariants at roots of unity, arXiv:2207.01882 (2022).
- [5] C. Blanchet, F. Costantino, N. Geer, and B. Patureau-Mirand, Non-semi-simple TQFTs, Reidemeister torsion and Kashaev’s invariants, Adv. Math. 301 (2016), 1–78.
- [6] Francesco Costantino, Nathan Geer, and Bertrand Patureau-Mirand, Quantum invariants of 3-manifolds via link surgery presentations and non-semi-simple categories, J. Topol. 7 (2014), no. 4, 1005–1053.
- [7] Thomas Creutzig, Tudor Dimofte, Niklas Garner, and Nathan Geer, A QFT for non-semisimple TQFT, arXiv:2112.01559 (2021).
- [8] Pavel Etingof, Dmitri Nikshych, and Viktor Ostrik, On fusion categories, Ann. of Math. (2) 162 (2005), no. 2, 581–642.
- [9] B. L. Feigin, A. M. Gainutdinov, A. M. Semikhatov, and I. Yu. Tipunin, Modular group representations and fusion in logarithmic conformal field theories and in the quantum group center, Comm. Math. Phys. 265 (2006), no. 1, 47–93.
- [10] S. Friedl and T. Kim, The Thurston norm, fibered manifolds and twisted Alexander polynomials, Topology 45 (2006), no. 6, 929–953.
- [11] S. Friedl and S. Vidussi, A survey of twisted Alexander polynomials, The mathematics of knots, Contrib. Math. Comput. Sci., vol. 1, Springer, Heidelberg, 2011, pp. 45–94.
- [12] by same author, The Thurston norm and twisted Alexander polynomials, J. Reine Angew. Math. 707 (2015), 87–102.
- [13] Nathan Geer, Bertrand Patureau-Mirand, and Vladimir Turaev, Modified quantum dimensions and re-normalized link invariants, Compositio Mathematica 145 (2009), no. 1, 196–212.
- [14] Shlomo Gelaki, Deepak Naidu, and Dmitri Nikshych, Centers of graded fusion categories, Algebra Number Theory 3 (2009), no. 8, 959-990 (2009).
- [15] Sergei Gukov, Po-Shen Hsin, Hiraku Nakajima, Sunghyuk Park, Du Pei, and Nikita Sopenko, Rozansky-Witten geometry of Coulomb branches and logarithmic knot invariants, J. Geom. Phys. 168 (2021), Paper No. 104311, 22.
- [16] Kazuo Habiro, Bottom tangles and universal invariants, Algebr. Geom. Topol. 6 (2006), 1113–1214.
- [17] Matthew Harper, A non-abelian generalization of the Alexander polynomial from quantum , arXiv:2008.06983 (2020).
- [18] Marc Lackenby, The efficient certification of knottedness and Thurston norm, Advances in Mathematics 387 (2021), 107796.
- [19] X. S. Lin, Representations of knot groups and twisted Alexander polynomials, Acta Math. Sin. (Engl. Ser.) 17 (2001), no. 3, 361–380.
- [20] Jules Martel and Sonny Willetts, Unified invariant of knots from homological braid action on verma modules, arXiv:2112.15204 (2021).
- [21] Jun Murakami, Colored Alexander invariants and cone-manifolds, Osaka J. Math. 45 (2008), no. 2, 541–564.
- [22] Daniel López Neumann, Twisting Kuperberg invariants via Fox calculus and Reidemeister torsion, arXiv.1911.02925 (2019).
- [23] by same author, Twisted Kuperberg invariants of knots and Reidemeister torsion via twisted Drinfeld doubles, arXiv:2201.13382 (2022).
- [24] Tomotada Ohtsuki, Colored ribbon Hopf algebras and universal invariants of framed links, J. Knot Theory Ramifications 2 (1993), no. 2, 211–232.
- [25] Tomotada Ohtsuki, On the 2–loop polynomial of knots, Geometry & Topology 11 (2007), no. 3, 1357–1475.
- [26] Peter Ozsvath and Zoltan Szabo, Holomorphic disks and genus bounds, Geometry & Topology 8 (2004), no. 1, 311–334.
- [27] J. Porti, Reidemeister torsion, hyperbolic three-manifolds, and character varieties, Handbook of group actions. Vol. IV, Adv. Lect. Math. (ALM), vol. 41, Int. Press, Somerville, MA, 2018, pp. 447–507.
- [28] D. E. Radford, Hopf algebras, Series on Knots and Everything, vol. 49, World Scientific Publishing Co. Pte. Ltd., Hackensack, NJ, 2012.
- [29] N. Reshetikhin and V. G. Turaev, Ribbon graphs and their invariants derived from quantum groups, Comm. Math. Phys. 127 (1990), no. 1, 1–26.
- [30] by same author, Invariants of -manifolds via link polynomials and quantum groups, Invent. Math. 103 (1991), no. 3, 547–597.
- [31] V. G. Turaev, Homotopy field theory in dimension 3 and crossed group-categories, arXiv:math/0005291 (2000).
- [32] by same author, Introduction to combinatorial torsions, Lectures in Mathematics ETH Zürich, Birkhäuser Verlag, Basel, 2001, Notes taken by Felix Schlenk.
- [33] A. Virelizier, Hopf group-coalgebras, J. Pure Appl. Algebra 171 (2002), no. 1, 75–122.
- [34] by same author, Graded quantum groups and quasitriangular Hopf group-coalgebras, Comm. Algebra 33 (2005), no. 9, 3029–3050.
- [35] M. Wada, Twisted Alexander polynomial for finitely presentable groups, Topology 33 (1994), no. 2, 241–256.
- [36] Sonny Willetts, A unification of the ADO and colored Jones polynomials of a knot, Quantum Topol. 13 (2022), no. 1, 137–181.
- [37] E. Witten, Quantum field theory and the Jones polynomial, Comm. Math. Phys. 121 (1989), no. 3, 351–399.