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

Genus bounds for twisted quantum invariants

Daniel López Neumann and Roland van der Veen [email protected], [email protected]
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 HH. These are obtained by the Reshetikhin-Turaev construction extended to the Aut(H)\mathrm{Aut}(H)-twisted Drinfeld double of HH, provided HH is finite dimensional and m\mathbb{N}^{m}-graded.

We show that the degree of these polynomials is bounded above by 2g(K)d(H)2g(K)\cdot d(H) where g(K)g(K) is the Seifert genus of a knot KK and d(H)d(H) is the top degree of the Hopf algebra. When HH is an exterior algebra, our theorem recovers Friedl and Kim’s genus bounds for twisted Alexander polynomials. When HH is the Borel part of restricted quantum 𝔰𝔩2\mathfrak{sl}_{2} 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 𝔰𝔩(2,)\mathfrak{sl}(2,\mathbb{C}). 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 qq of the quantum group Uq(𝔤)U_{q}(\mathfrak{g}) 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 𝔤=𝔰𝔩2\mathfrak{g}=\mathfrak{sl}_{2} 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 SL(2,)SL(2,\mathbb{C})-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 XK=S3KX_{K}=S^{3}\setminus K, or in other words, invariants of a pair (K,ρ)(K,\rho) where ρ:π1(XK)G\rho:\pi_{1}(X_{K})\to G is a homomorphism into a group GG. For instance, the Alexander polynomial comes from the covering space corresponding to the projection π1(XK)H1(XK)=\pi_{1}(X_{K})\to H_{1}(X_{K})=\mathbb{Z}. Most applications depend on these invariants being polynomials, for which one needs G\mathbb{Z}\subset G, in particular, GG 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 (K,ρ)(K,\rho). This extension starts with a GG-crossed braided monoidal category, essentially a GG-graded category with a GG-action and a compatible braiding, and produces invariants of tangles endowed with a representation ρ:π1(XT)G\rho:\pi_{1}(X_{T})\to G, where XT=2×[0,1]TX_{T}=\mathbb{R}^{2}\times[0,1]\setminus T. The GG-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 GG-braided monoidal categories where GG is some infinite group.

In previous work, the first author [23] considered a special class of such categories: relative Drinfeld centers of crossed products Rep(H)Aut(H)\text{Rep}(H)\ltimes\mathrm{Aut}(H) where HH is a finite dimensional Hopf algebra and Aut(H)\mathrm{Aut}(H) its group of Hopf algebra automorphisms, or equivalently, Aut(H)\mathrm{Aut}(H)-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 D(H)D(H) so, in some sense, they do retain some covering space theory. Moreover, it was shown in [23, 22] that the SL(n,)SL(n,\mathbb{C})-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 HH be a finite dimensional Hopf algebra over a field 𝕂\mathbb{K} such that the Drinfeld double D(H)D(H) is ribbon. The twisted Drinfeld double D(H)¯\underline{D(H)} of HH is a family, indexed by αAut(H)\alpha\in\mathrm{Aut}(H), of deformations D(H)αD(H)_{\alpha} of the usual Drinfeld double of HH [34]. If TT is a framed, oriented, nn-component open tangle, the twisted Drinfeld double leads to a deformation

ZD(H)¯ρ(T)(HH)nZ^{\rho}_{\underline{D(H)}}(T)\in(H^{*}\otimes H)^{\otimes n}

of the usual universal invariant of tangles ZD(H)(T)Z_{D(H)}(T) (see e.g. [16]) that depends on a homomorphism ρ:π1(XT)Aut(H)\rho:\pi_{1}(X_{T})\to\mathrm{Aut}(H). This invariant is essentially a special case of Turaev’s construction [31]. Now, if in addition HH is m\mathbb{N}^{m}-graded, there is a canonical automorphism θAut(H)\theta\in\mathrm{Aut}(H^{\prime}), where H=H𝕂𝔽H^{\prime}=H\otimes_{\mathbb{K}}\mathbb{F} and 𝔽=𝕂[t1±1/2,,tm±1/2]\mathbb{F}=\mathbb{K}[t_{1}^{\pm 1/2},\dots,t_{m}^{\pm 1/2}], defined by θ(x)=t1k1tmkmx\theta(x)=t_{1}^{k_{1}}\dots t_{m}^{k_{m}}x for xHx\in H homogeneous of degree (k1,,km)(k_{1},\dots,k_{m}). Then any ρ\rho as above extends to ρθ:π1(XT)Aut(H)\rho\otimes\theta:\pi_{1}(X_{T})\to\mathrm{Aut}(H^{\prime}). This “degree twisted” extension generalizes a similar procedure in twisted Reidemeister torsion theory (e.g. [11]). We thus get an invariant

ZD(H)¯ρθ(T)(H𝔽H)nZ_{\underline{D(H^{\prime})}}^{\rho\otimes\theta}(T)\in(H^{\prime*}\otimes_{\mathbb{F}}H^{\prime})^{\otimes n}

defined for any ρ:π1(XT)Autm0(H)\rho:\pi_{1}(X_{T})\to\mathrm{Aut}^{0}_{\mathbb{N}^{m}}(H) (automorphisms preserving the degree and fixing the left cointegral of HH). If KoK_{o} is a framed, oriented, (1,1)(1,1)-tangle whose closure is a knot KK, then we define

PHρθ(K)=λϵD(H)(ZD(H)¯ρθ(Ko))𝕂[t1±1,,tm±1]P_{H}^{\rho\otimes\theta}(K)=\lambda\cdot\epsilon_{D(H^{\prime})}(Z_{\underline{D(H^{\prime})}}^{\rho\otimes\theta}(K_{o}))\in\mathbb{K}[t_{1}^{\pm 1},\dots,t_{m}^{\pm 1}]

where λ𝔽×\lambda\in\mathbb{F}^{\times} is a normalization factor making this an invariant of the unframed (K,ρ)(K,\rho). If ρ1\rho\equiv 1, PHθ(K)P_{H}^{\theta}(K) is a polynomial invariant of the oriented knot KK with no additional structure, but one still has to think that there is a canonical abelian representation H1(XK)Aut(H)H_{1}(X_{K})\to\mathrm{Aut}(H^{\prime}) sending the oriented meridian to θ\theta.

If PHρθ(K)=k¯ak¯t1k1tmkmP_{H}^{\rho\otimes\theta}(K)=\sum_{\overline{k}}a_{\overline{k}}t_{1}^{k_{1}}\dots t_{m}^{k_{m}} where ak¯𝕂a_{\overline{k}}\in\mathbb{K} and k¯=(k1,,km)m\overline{k}=(k_{1},\dots,k_{m})\in\mathbb{Z}^{m}, we define degPHρθ(K)\mathrm{deg}\ P_{H}^{\rho\otimes\theta}(K) as max{k1++km|ak¯0}min{k1++km|ak¯0}\max\{k_{1}+\dots+k_{m}\ |\ a_{\overline{k}}\neq 0\}-\min\{k_{1}+\dots+k_{m}\ |\ a_{\overline{k}}\neq 0\}. We denote d(H)=max{i1++im|H(i1,,im)0}d(H)=\max\{i_{1}+\dots+i_{m}\ |\ H_{(i_{1},\dots,i_{m})}\neq 0\} where H(i1,,im)H_{(i_{1},\dots,i_{m})} is the component of degree (i1,,im)m(i_{1},\dots,i_{m})\in\mathbb{N}^{m} of HH.

Theorem 1.

(Genus bound) Let HH be a m\mathbb{N}^{m}-graded Hopf algebra of finite dimension. Let KS3K\subset S^{3} be a knot and ρ:π1(S3K)Autm0(H)\rho:\pi_{1}(S^{3}\setminus K)\to\mathrm{Aut}^{0}_{\mathbb{N}^{m}}(H) a homomorphism. Then

degPHρθ(K)2g(K)d(H)\displaystyle\mathrm{deg}\ P_{H}^{\rho\otimes\theta}(K)\leq 2g(K)\cdot d(H)

where g(K)g(K) is the Seifert genus of KK.

In fact the proof shows that a similar bound holds for ZD(H)¯ρθ(Ko)Z_{\underline{D(H^{\prime})}}^{\rho\otimes\theta}(K_{o}). 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 ρθ\rho\otimes\theta, 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 H=Λ(n)H=\Lambda(\mathbb{C}^{n}) is an exterior algebra (with n\mathbb{C}^{n} concentrated in degree one), one has Aut0(H)SL(n,)\mathrm{Aut}^{0}_{\mathbb{N}}(H)\cong SL(n,\mathbb{C}), d(H)=nd(H)=n and PΛ(n)ρθ(K)P_{\Lambda(\mathbb{C}^{n})}^{\rho\otimes\theta}(K) coincides with the SL(n,)SL(n,\mathbb{C})-twisted relative Reidemeister torsion τρθ(XK,μ)\tau^{\rho\otimes\theta}(X_{K},\mu) where μXK\mu\subset\partial X_{K} is a meridian [23]. This torsion is essentially equivalent to the twisted Alexander polynomial ΔKρ(t)\Delta_{K}^{\rho}(t) of Lin and Wada [19, 35]. The above theorem implies

degτρθ(XK,μ)2g(K)n.\mathrm{deg}\ \tau^{\rho\otimes\theta}(X_{K},\mu)\leq 2g(K)n.

Since τρθ(XK,μ)=det(tρ(μ)In)τρθ(XK)\tau^{\rho\otimes\theta}(X_{K},\mu)=\det(t\rho(\mu)-I_{n})\tau^{\rho\otimes\theta}(X_{K}) this is equivalent to degτρθ(XK)n(2g(K)1)\mathrm{deg}\ \tau^{\rho\otimes\theta}(X_{K})\leq n(2g(K)-1) which is a special case of a result of Friedl and Kim [10].

1.4. ADO invariants

Now let HpH_{p} be the Borel part of the restricted quantum group U¯ζp(𝔰𝔩2)\overline{U}_{\zeta_{p}}(\mathfrak{sl}_{2}) at a 2p2p-th root of unity ζp\zeta_{p}. Then HpH_{p} has finite dimension, it is \mathbb{N}-graded (with d(Hp)=p1d(H_{p})=p-1) and D(Hp)D(H_{p}) is ribbon, so we have a polynomial invariant PHpθ(K)[ζp2][t±1].P_{H_{p}}^{\theta}(K)\in\mathbb{Z}[\zeta_{p}^{2}][t^{\pm 1}]. On the other hand, the unrolled restricted quantum group U¯ζpH(𝔰𝔩2)\overline{U}^{H}_{\zeta_{p}}(\mathfrak{sl}_{2}) leads to the ADO invariant ADOp(K,t)[ζp2][t±1]ADO_{p}(K,t)\in\mathbb{Z}[\zeta_{p}^{2}][t^{\pm 1}] [1]. We use the normalization of the ADO invariant in which ADO2(K,t)=ΔK(t)ADO_{2}(K,t)=\Delta_{K}(t).

Theorem 2.

We have

PHpθ(K)=ADOp(K,t)P_{H_{p}}^{\theta}(K)=ADO_{p}(K,t)

hence

degADOp(K,t)2g(K)(p1).\mathrm{deg}\ ADO_{p}(K,t)\leq 2g(K)(p-1).

As an example, the ADO4ADO_{4} invariants of the first 66 twist knots (denoted by their usual names in knot tables) are given as follows:

KK ADO𝔰𝔩2,4(K,t)ADO_{\mathfrak{sl}_{2},4}(K,t)
313_{1} (12i)+t3+it2(1+i)t1+(1+i)t+it2t3(1-2i)+t^{-3}+it^{-2}-(1+i)t^{-1}+(1+i)t+it^{2}-t^{3}
414_{1} 7+it33t26it1+6it3t2it37+it^{-3}-3t^{-2}-6it^{-1}+6it-3t^{2}-it^{3}
525_{2} (5+10i)(2+2i)t3+(35i)t2+(8+4i)t1(8+4i)t+(35i)t2+(2+2i)t3(-5+10i)-(2+2i)t^{-3}+(3-5i)t^{-2}+(8+4i)t^{-1}-(8+4i)t+(3-5i)t^{2}+(2+2i)t^{3}
616_{1} (7+10i)(22i)t3(3+5i)t2+(84i)t1(84i)t(3+5i)t2+(22i)t3(7+10i)-(2-2i)t^{-3}-(3+5i)t^{-2}+(8-4i)t^{-1}-(8-4i)t-(3+5i)t^{2}+(2-2i)t^{3}
727_{2} (5+4i)(2+i)t3+(32i)t2+(2+4i)t1(2+4i)t+(32i)t2+(2+i)t3(-5+4i)-(2+i)t^{-3}+(3-2i)t^{-2}+(2+4i)t^{-1}-(2+4i)t+(3-2i)t^{2}+(2+i)t^{3}
818_{1} (13+2i)(12i)t3(6+i)t2+(111i)t1(111i)t(6+i)t2+(12i)t3(13+2i)-(1-2i)t^{-3}-(6+i)t^{-2}+(1-11i)t^{-1}-(1-11i)t-(6+i)t^{2}+(1-2i)t^{3}

These all have degree 6\leq 6, since g(K)=1g(K)=1 for these knots, this matches with our theorem. Formulas for ADO invariants of double twist knots (genus 1) and torus knots T(2,2k+1)T_{(2,2k+1)} (genus kk) are given in [3]. All these satisfy the genus bound above.

1.5. The higher rank case

More generally, for a simple Lie algebra 𝔤\mathfrak{g} of rank mm and a primitive root of unity ζp2\zeta_{p}^{2} of order pp (say, pp coprime to the determinant of the Cartan matrix), our construction provides a polynomial invariant P𝔤,p(K)[ζp][t1±1,,tm±1]P_{\mathfrak{g},p}(K)\in\mathbb{Z}[\zeta_{p}][t_{1}^{\pm 1},\dots,t_{m}^{\pm 1}] that gives a lower bound to the genus. For instance

degP𝔰𝔩N+1,p(K)2g(K)(p1)16N(N+1)(N+2),\mathrm{deg}\ P_{\mathfrak{sl}_{N+1},p}(K)\leq 2g(K)(p-1)\frac{1}{6}N(N+1)(N+2),

which generalizes the above bound for ADO to all NN. The invariant P𝔤,pP_{\mathfrak{g},p} is obtained by letting HH be the Borel part of the corresponding restricted quantum group (which is m\mathbb{N}^{m}-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 𝔰𝔩3\mathfrak{sl}_{3}-ADO invariant at p=2p=2 (so ζp=i\zeta_{p}=i) defined in [17] is given as follows for the above twist knots:

KK ADO𝔰𝔩3,2(K,x,y)ADO_{\mathfrak{sl}_{3},2}(K,x,y)
313_{1} 1+x22x12x+x2+y2+x2y2x1y22y1x2y11+x^{-2}-2x^{-1}-2x+x^{2}+y^{-2}+x^{-2}y^{-2}-x^{-1}y^{-2}-2y^{-1}-x^{-2}y^{-1}
+2x1y1+xy12y+yx1+2xyx2y+y2xy2+x2y2+2x^{-1}y^{-1}+xy^{-1}-2y+yx^{-1}+2xy-x^{2}y+y^{2}-xy^{2}+x^{2}y^{2}
414_{1} 25+x212x112x+x2+y2+x2y23x1y212y13x2y125+x^{-2}-12x^{-1}-12x+x^{2}+y^{-2}+x^{-2}y^{-2}-3x^{-1}y^{-2}-12y^{-1}-3x^{-2}y^{-1}
+12x1y1+3xy112y+3yx1+12xy3x2y+y23xy2+x2y2+12x^{-1}y^{-1}+3xy^{-1}-12y+3yx^{-1}+12xy-3x^{2}y+y^{2}-3xy^{2}+x^{2}y^{2}
525_{2} 37+6x226x126x+6x2+6y2+6x2y210x1y226y110x2y137+6x^{-2}-26x^{-1}-26x+6x^{2}+6y^{-2}+6x^{-2}y^{-2}-10x^{-1}y^{-2}-26y^{-1}-10x^{-2}y^{-1}
+26x1y1+10xy126y+10yx1+26xy10x2y+6y210xy2+6x2y2+26x^{-1}y^{-1}+10xy^{-1}-26y+10yx^{-1}+26xy-10x^{2}y+6y^{2}-10xy^{2}+6x^{2}y^{2}
616_{1} 85+6x246x146x+6x2+6y2+6x2y214x1y246y114x2y185+6x^{-2}-46x^{-1}-46x+6x^{2}+6y^{-2}+6x^{-2}y^{-2}-14x^{-1}y^{-2}-46y^{-1}-14x^{-2}y^{-1}
+46x1y1+14xy146y+14yx1+46xy14x2y+6y214xy2+6x2y2+46x^{-1}y^{-1}+14xy^{-1}-46y+14yx^{-1}+46xy-14x^{2}y+6y^{2}-14xy^{2}+6x^{2}y^{2}
727_{2} 97+13x264x164x+13x2+13y2+13x2y223x1y264y123x2y197+13x^{-2}-64x^{-1}-64x+13x^{2}+13y^{-2}+13x^{-2}y^{-2}-23x^{-1}y^{-2}-64y^{-1}-23x^{-2}y^{-1}
+64x1y1+23xy164y+23yx1+64xy23x2y+13y223xy2+13x2y2+64x^{-1}y^{-1}+23xy^{-1}-64y+23yx^{-1}+64xy-23x^{2}y+13y^{2}-23xy^{2}+13x^{2}y^{2}

Our genus bound implies degP𝔰𝔩3,2(K)8g(K)\mathrm{deg}\ P_{\mathfrak{sl}_{3},2}(K)\leq 8g(K), 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 ΘK(x,y)\Theta_{K}(x,y). 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 ΘK¯(x,y)=ΘK(x,y)\Theta_{\overline{K}}(x,y)=-\Theta_{K}(x,y) 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 ADOpADO_{p} polynomials (for varying pp) 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 𝔰𝔩2\mathfrak{sl}_{2}, Bar-Natan and the second author studied the nn-loop polynomial using hh-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 nn and a representation ρ:π1(XK)U(n)\rho:\pi_{1}(X_{K})\to U(n) for which degτρθ(XK,μ)=2g(K)n\mathrm{deg}\ \tau^{\rho\otimes\theta}(X_{K},\mu)=2g(K)n. The authors do not know whether there is an algorithm to find such a ρ\rho 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 ρ1\rho\equiv 1 and varying HH 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 GG-coalgebras, we define twisted Drinfeld doubles and study ribbon elements in these. In Section 3 we define the invariant ZD(H)¯ρ(T)Z^{\rho}_{\underline{D(H)}}(T), the polynomial invariant PHρθ(K)P_{H}^{\rho\otimes\theta}(K), 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 HpH_{p} above and prove Theorem 2.

2. Hopf algebra preliminaries

For simplicity, we work over an algebraically closed field 𝕂\mathbb{K} of characteristic 0. In all that follows, GG 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 HH over 𝕂\mathbb{K} by m,Δ,1,ϵ,Sm,\Delta,1,\epsilon,S respectively. We will employ Sweedler’s notation for the coproduct of a Hopf algebra, that is,

Δ(x)=x(1)x(2),(Δid)Δ(x)=x(1)x(3)x(3),\Delta(x)=x_{(1)}\otimes x_{(2)},(\Delta\otimes\text{id})\Delta(x)=x_{(1)}\otimes x_{(3)}\otimes x_{(3)},

etc. The dual HH^{*} is also a Hopf algebra with

pq,x=p,x(1)q,x(2)\langle p\cdot q,x\rangle=\langle p,x_{(1)}\rangle\langle q,x_{(2)}\rangle

for each p,qHp,q\in H^{*} and xHx\in H. Here ,\langle\ ,\ \rangle denotes the usual pairing and p,x=x,p\langle p,x\rangle=\langle x,p\rangle for any pH,xHp\in H^{*},x\in H. We denote by Aut(H)\mathrm{Aut}(H) the group of Hopf algebra automorphisms of HH and Aut0(H)\mathrm{Aut}^{0}(H) the subgroup of Aut(H)\mathrm{Aut}(H) fixing a non-zero cointegral (either left or right, see Subsection 2.4 below for definitions).

A Hopf algebra HH is m\mathbb{N}^{m}-graded if H=ImHIH=\bigoplus_{I\in\mathbb{N}^{m}}H_{I} with m(HIHJ)HI+Jm(H_{I}\otimes H_{J})\subset H_{I+J}, Δ(HN)I+J=NHIHJ\Delta(H_{N})\subset\sum_{I+J=N}H_{I}\otimes H_{J} and S(HI)HIS(H_{I})\subset H_{I} for each I,J,NmI,J,N\in\mathbb{N}^{m}. We denote by Autm0(H)\mathrm{Aut}^{0}_{\mathbb{N}^{m}}(H) the subgroup of automorphisms of Aut0(H)\mathrm{Aut}^{0}(H) preserving the m\mathbb{N}^{m}-degree.

2.2. Hopf GG-coalgebras

A Hopf GG-coalgebra is a family H¯={Hα}αG\underline{H}=\{H_{\alpha}\}_{\alpha\in G} where each HαH_{\alpha} is an algebra with unit 1α1_{\alpha} together with a family of algebra morphisms Δα1,α2:Hα1α2Hα1Hα2\Delta_{\alpha_{1},\alpha_{2}}:H_{\alpha_{1}\alpha_{2}}\to H_{\alpha_{1}}\otimes H_{\alpha_{2}} for each α1,α2G\alpha_{1},\alpha_{2}\in G, an algebra morphism ϵ:H1𝕂\epsilon:H_{1}\to\mathbb{K} and algebra antiautomorphisms Sα:HαHα1S_{\alpha}:H_{\alpha}\to H_{\alpha^{-1}} for each α\alpha satisfying graded versions of the Hopf algebra axioms, see [31, 33] for more details. Note that H=H1H=H_{1} is a Hopf algebra in the usual sense (with the coproduct Δ1,1\Delta_{1,1} and antipode S1S_{1}).

A Hopf GG-coalgebra is said to be crossed if it comes with a family φ={φα,β}α,βG\varphi=\{\varphi_{\alpha,\beta}\}_{\alpha,\beta\in G} of algebra isomorphisms φα,β:HαHβαβ1\varphi_{\alpha,\beta}:H_{\alpha}\to H_{\beta\alpha\beta^{-1}}, simply denoted φβ\varphi_{\beta}, preserving the Hopf structure and satisfying φβ2φβ1=φβ2β1\varphi_{\beta_{2}}\varphi_{\beta_{1}}=\varphi_{\beta_{2}\beta_{1}}.

A crossed Hopf GG-coalgebra (H¯,φ)(\underline{H},\varphi) is quasi-triangular if it comes with a family R={Rα,β}α,βGR=\{R_{\alpha,\beta}\}_{\alpha,\beta\in G} of invertible elements in HαHβH_{\alpha}\otimes H_{\beta}, satisfying

  1. (1)

    Rα,βR_{\alpha,\beta} is φ\varphi-invariant, that is, (φγφγ)(Rα,β)=Rγαγ1,γβγ1(\varphi_{\gamma}\otimes\varphi_{\gamma})(R_{\alpha,\beta})=R_{\gamma\alpha\gamma^{-1},\gamma\beta\gamma^{-1}},

  2. (2)

    Rα,βΔα,β(x)=(τ[(φα1idHα)Δαβα1,α(x)])Rα,βR_{\alpha,\beta}\cdot\Delta_{\alpha,\beta}(x)=(\tau[(\varphi_{\alpha^{-1}}\otimes\text{id}_{H_{\alpha}})\Delta_{\alpha\beta\alpha^{-1},\alpha}(x)])\cdot R_{\alpha,\beta} where τ\tau denotes permutation of two factors (with a sign in the super-case),

  3. (3)

    (idHαΔβ,γ)(Rα,βγ)=(Rα,γ)1β3(Rα,β)12γ(\text{id}_{H_{\alpha}}\otimes\Delta_{\beta,\gamma})(R_{\alpha,\beta\gamma})=(R_{\alpha,\gamma})_{1\beta 3}\cdot(R_{\alpha,\beta})_{12\gamma},

  4. (4)

    (Δα,βidHγ)(Rαβ,γ)=((idHαφβ1)(Rα,βγβ1))1β3(Rβ,γ)α23(\Delta_{\alpha,\beta}\otimes\text{id}_{H_{\gamma}})(R_{\alpha\beta,\gamma})=((\text{id}_{H_{\alpha}}\otimes\varphi_{\beta^{-1}})(R_{\alpha,\beta\gamma\beta^{-1}}))_{1\beta 3}\cdot(R_{\beta,\gamma})_{\alpha 23}

for each α,β,γG\alpha,\beta,\gamma\in G, where we used the notation X12γ=xy1γX_{12\gamma}=x\otimes y\otimes 1_{\gamma} for any X=xyHαHβX=\sum x\otimes y\in H_{\alpha}\otimes H_{\beta} and similarly for Yα23,Z1β3Y_{\alpha 23},Z_{1\beta 3} (YHβHγ,ZHαHγY\in H_{\beta}\otimes H_{\gamma},Z\in H_{\alpha}\otimes H_{\gamma}). If (H¯,φ,R)(\underline{H},\varphi,R) is quasi-triangular, the graded Drinfeld element is u={uα}αGu=\{u_{\alpha}\}_{\alpha\in G} defined by

uα=mα(Sα1φαidHα)τα,α1(Rα,α1)Hα.u_{\alpha}=m_{\alpha}(S_{\alpha^{-1}}\varphi_{\alpha}\otimes\text{id}_{H_{\alpha}})\tau_{\alpha,\alpha^{-1}}(R_{\alpha,\alpha^{-1}})\in H_{\alpha}.

A quasi-triangular Hopf GG-coalgebra (H¯,φ,R)(\underline{H},\varphi,R) is GG-ribbon if it comes with a family v={vα}αGv=\{v_{\alpha}\}_{\alpha\in G} of invertible elements vαHαv_{\alpha}\in H_{\alpha} satisfying, for each α,βG\alpha,\beta\in G:

  1. (1)

    Δα,β(vαβ)=(vαvβ)(τβ,α[φα1id(Rαβα1,α)])Rα,β\Delta_{\alpha,\beta}(v_{\alpha\beta})=(v_{\alpha}\otimes v_{\beta})(\tau_{\beta,\alpha}[\varphi_{\alpha^{-1}}\otimes\text{id}(R_{\alpha\beta\alpha^{-1},\alpha})])R_{\alpha,\beta},

  2. (2)

    Sα(vα)=vα1S_{\alpha}(v_{\alpha})=v_{\alpha^{-1}},

  3. (3)

    φα(x)=vα1xvα\varphi_{\alpha}(x)=v_{\alpha}^{-1}xv_{\alpha} for all xHαx\in H_{\alpha},

  4. (4)

    φβ(vα)=vβαβ1\varphi_{\beta}(v_{\alpha})=v_{\beta\alpha\beta^{-1}}.

If 𝒈α=vαuα\boldsymbol{g}_{\alpha}=v_{\alpha}u_{\alpha}, we call 𝒈={𝒈α}αG\boldsymbol{g}=\{\boldsymbol{g}_{\alpha}\}_{\alpha\in G} the GG-pivot of H¯\underline{H}, this satisfies Sα1Sα(x)=𝒈αx𝒈α1S_{\alpha^{-1}}S_{\alpha}(x)=\boldsymbol{g}_{\alpha}x\boldsymbol{g}_{\alpha}^{-1} for each xHα,αGx\in H_{\alpha},\alpha\in G and is group-like in the sense that Δα,β(𝒈αβ)=𝒈α𝒈β\Delta_{\alpha,\beta}(\boldsymbol{g}_{\alpha\beta})=\boldsymbol{g}_{\alpha}\otimes\boldsymbol{g}_{\beta}.

2.3. Twisted Drinfeld doubles

We now define the twisted Drinfeld double of a Hopf algebra HH, which is a quasi-triangular Hopf Aut(H)\mathrm{Aut}(H)-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 αAut(H)\alpha\in\mathrm{Aut}(H) we define an algebra D(H)αD(H)_{\alpha} as follows: D(H)α=HHD(H)_{\alpha}=H^{*}\otimes H as a vector space and we define the product by

(pa)α(qb)=a(1),q(3)S1α1(a(3)),q(1)pq(2)a(2)b(p\otimes a)\cdot_{\alpha}(q\otimes b)=\langle a_{(1)},q_{(3)}\rangle\langle S^{-1}\alpha^{-1}(a_{(3)}),q_{(1)}\rangle p\cdot q_{(2)}\otimes a_{(2)}\cdot b

where p,qH,a,bHp,q\in H^{*},a,b\in H. This is an associative algebra with unit 1α=ϵH1H1_{\alpha}=\epsilon_{H}\otimes 1_{H}. For each α,βAut(H)\alpha,\beta\in\mathrm{Aut}(H) we define a coproduct Δα,β:D(H)αβD(H)αD(H)β\Delta_{\alpha,\beta}:D(H)_{\alpha\beta}\to D(H)_{\alpha}\otimes D(H)_{\beta} by

Δα,β(pa)=(p(2)a(1))(p(1)α1(a(2))).\Delta_{\alpha,\beta}(p\otimes a)=(p_{(2)}\otimes a_{(1)})\otimes(p_{(1)}\otimes\alpha^{-1}(a_{(2)})).

For each αAut(H)\alpha\in\mathrm{Aut}(H) we define an antipode Sα:D(H)αD(H)α1S_{\alpha}:D(H)_{\alpha}\to D(H)_{\alpha^{-1}} by

Sα(pa)=(ϵα1(S(a)))α1(pS11).S_{\alpha}(p\otimes a)=(\epsilon\otimes\alpha^{-1}(S(a)))\cdot_{\alpha^{-1}}(p\circ S^{-1}\otimes 1).

For each α,β\alpha,\beta define an algebra isomorphism φα:D(H)βD(H)αβα1\varphi_{\alpha}:D(H)_{\beta}\to D(H)_{\alpha\beta\alpha^{-1}} by

φα(pa)=pα1α(a).\varphi_{\alpha}(p\otimes a)=p\circ\alpha^{-1}\otimes\alpha(a).

Finally, for each α,β\alpha,\beta let

Rα,β=(ϵα(hi))(hi1)D(H)αD(H)βR_{\alpha,\beta}=\sum(\epsilon\otimes\alpha(h_{i}))\otimes(h^{i}\otimes 1)\in D(H)_{\alpha}\otimes D(H)_{\beta}

where (hi)(h_{i}) is any basis of HH and (hi)(h^{i}) is the dual basis. With all these structure maps, {D(H)α}αAut(H)\{D(H)_{\alpha}\}_{\alpha\in\mathrm{Aut}(H)} has the structure of a quasi-triangular crossed Hopf Aut(H)\mathrm{Aut}(H)-coalgebra. We denote D(H)¯={D(H)α}αAut(H)\underline{D(H)}=\{D(H)_{\alpha}\}_{\alpha\in\mathrm{Aut}(H)} and call it the twisted Drinfeld double of HH. If ϕ:GAut(H)\phi:G\to\mathrm{Aut}(H) is a group homomorphism, then D(H)¯|G={D(H)ϕ(α)}αG\underline{D(H)}|_{G}=\{D(H)_{\phi({\alpha})}\}_{\alpha\in G} is a quasi-triangular Hopf GG-coalgebra in an obvious way, we call it a GG-twisted Drinfeld double.

Remark 2.1.

The category of modules of the twisted Drinfeld double is braided Aut(H)\mathrm{Aut}(H)-crossed equivalent to the relative Drinfeld center 𝒵rel(Rep(H)Aut(H))\mathcal{Z}_{rel}(\text{Rep}(H)\rtimes\mathrm{Aut}(H)) of [14], where Aut(H)\mathrm{Aut}(H) acts on Rep(H)\text{Rep}(H) by VTα(V)V\to T_{\alpha}(V) for any VRep(H)V\in\text{Rep}(H), where Tα(V)T_{\alpha}(V) is VV as a vector space with HH-action hv:=α1(h)vh\cdot v:=\alpha^{-1}(h)v, see [23].

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 HH is an element ΛlH\Lambda_{l}\in H such that xΛl=ϵ(x)Λlx\Lambda_{l}=\epsilon(x)\Lambda_{l} for each xHx\in H. Dually, a right integral is λrH\lambda_{r}\in H^{*} such that λr(x(1))x(2)=λr(x)1H\lambda_{r}(x_{(1)})x_{(2)}=\lambda_{r}(x)1_{H} for each xHx\in H. As a consequence of uniqueness of integrals/cointegrals, there are unique group-likes 𝜶G(H),𝒂G(H)\boldsymbol{\alpha}\in G(H^{*}),\boldsymbol{a}\in G(H) satisfying

x(1)λr(x(2))=λr(x)𝒂,Λlx=𝜶(x)Λlx_{(1)}\lambda_{r}(x_{(2)})=\lambda_{r}(x)\boldsymbol{a},\hskip 28.45274pt\Lambda_{l}x=\boldsymbol{\alpha}(x)\Lambda_{l}

for all xHx\in H. These are the distinguished group-likes of HH.

Now, a theorem of Kauffman and Radford states that the Drinfeld double D(H)D(H) is ribbon if and only if there are group-likes 𝒃G(H),𝜷G(H)\boldsymbol{b}\in G(H),\boldsymbol{\beta}\in G(H^{*}) such that 𝒃2=𝒂,𝜷2=𝜶\boldsymbol{b}^{2}=\boldsymbol{a},\boldsymbol{\beta}^{2}=\boldsymbol{\alpha} and

S2=ad𝜷1ad𝒃.S^{2}=\text{ad}_{\boldsymbol{\beta}^{-1}}\circ\text{ad}_{\boldsymbol{b}}.

In such a case, the ribbon element is v=u1(𝜷𝒃)v=u^{-1}(\boldsymbol{\beta}\otimes\boldsymbol{b}) where uu is the Drinfeld element of D(H)D(H).

We can now give sufficient conditions for D(H)¯\underline{D(H)} to be ribbon (or rather, a restriction to a certain subgroup GAut(H)G\subset\mathrm{Aut}(H)). Let’s suppose that HH satisfies the above condition making D(H)D(H) ribbon. Let GAut(H)G\subset\mathrm{Aut}(H) be a subgroup fixing such 𝒃,𝜷\boldsymbol{b},\boldsymbol{\beta}. Let rH:Aut(H)𝕂×r_{H}:\mathrm{Aut}(H)\to\mathbb{K}^{\times} be the homomorphism characterized by α(Λl)=rH(α)Λl\alpha(\Lambda_{l})=r_{H}(\alpha)\Lambda_{l} (this is defined by uniqueness of cointegrals). Then, as shown in [23, Prop. 2.4], D(H)¯|G\underline{D(H)}|_{G} is GG-ribbon if and only if the homomorphism rH|G:G𝕂×r_{H}|_{G}:G\to\mathbb{K}^{\times} has a square root. In such a case, the GG-ribbon element is given by

vα=rH(α)1(idHα)(v)v_{\alpha}=\sqrt{r_{H}(\alpha)}^{-1}(\text{id}_{H^{*}}\otimes\alpha)(v)

for any αG\alpha\in G, where vv is the ribbon element of D(H)D(H). The GG-pivot is given by

𝒈α=rH(α)1(𝜷𝒃).\boldsymbol{g}_{\alpha}=\sqrt{r_{H}(\alpha)}^{-1}(\boldsymbol{\beta}\otimes\boldsymbol{b}).

The square root condition is immediately satisfied if G=Ker(rH)=Aut0(H)G=\text{Ker}\,(r_{H})=\mathrm{Aut}^{0}(H).

2.5. Trivializing the crossing

Suppose that GG is an abelian group with an homomorphism ϕ:GAut(H)\phi:G\to\mathrm{Aut}(H) and consider D(H)¯|G\underline{D(H)}|_{G} as defined above. Then GG acts on each D(H)αD(H)_{\alpha} by βφϕ(β)|D(H)α\beta\mapsto\varphi_{\phi(\beta)}|_{D(H)_{\alpha}} and setting Aα=𝕂[G]D(H)αA_{\alpha}=\mathbb{K}[G]\ltimes D(H)_{\alpha} we get another Hopf GG-coalgebra A¯={Aα}αG\underline{A}=\{A_{\alpha}\}_{\alpha\in G}. The Hopf structure is extended to A¯\underline{A} by declaring the elements of GG to be group-likes. In what follows we denote βG\beta\in G by φβ\varphi_{\beta} when considered as an element of AαA_{\alpha} so that φβx=φϕ(β)(x)φβ\varphi_{\beta}x=\varphi_{\phi(\beta)}(x)\varphi_{\beta} in AαA_{\alpha} for each xD(H)αx\in D(H)_{\alpha}. It is easy to see that

Rα,βA=(1φα)Rα,βAαAβR_{\alpha,\beta}^{A}=(1\otimes\varphi_{\alpha})R_{\alpha,\beta}\in A_{\alpha}\otimes A_{\beta}

is an RR-matrix for A¯\underline{A} in (almost) the usual sense, that is, it satisfies

Rα,βAΔα,βA(x)=Δβ,αA,op(x)Rα,βA.R^{A}_{\alpha,\beta}\cdot\Delta^{A}_{\alpha,\beta}(x)=\Delta^{A,op}_{\beta,\alpha}(x)\cdot R^{A}_{\alpha,\beta}.

In other words, the category 𝒞=αGRep(Aα)\mathcal{C}=\coprod_{\alpha\in G}\text{Rep}(A_{\alpha}) is a braided category in the usual sense. It is not difficult to see that the Drinfeld element of A¯\underline{A} is determined by uαA=uαφα1u^{A}_{\alpha}=u_{\alpha}\varphi_{\alpha}^{-1} where u={uα}u=\{u_{\alpha}\} is the Drinfeld element of D(H)¯\underline{D(H)}. If v={vα}v=\{v_{\alpha}\} is a ribbon element in D(H)¯\underline{D(H)}, then A¯\underline{A} is ribbon with vαA=vαφαv^{A}_{\alpha}=v_{\alpha}\varphi_{\alpha}. It follows that the pivot of A¯\underline{A} is the same as in D(H)¯\underline{D(H)}.

2.6. The super algebra case

Recall that a super vector space is a vector space VV with a mod 2 grading, that is, a decomposition V=V0V1V=V_{0}\oplus V_{1}. The category of super vector spaces is symmetric with τV,W(vw)=(1)|v||w|wv\tau_{V,W}(v\otimes w)=(-1)^{|v||w|}w\otimes v where |v|,|w||v|,|w| denotes the mod 2 degree of homogeneous elements vV,wWv\in V,w\in W. A super Hopf algebra is a Hopf algebra (H,m,1,Δ,ϵ,S)(H,m,1,\Delta,\epsilon,S) in the category of super vector spaces. This amounts to the same axioms as for a Hopf algebra, except that the coproduct Δ\Delta satisfies

Δm=(mm)(idτH,Hid)(ΔΔ).\Delta\circ m=(m\otimes m)\circ(\text{id}\otimes\tau_{H,H}\otimes\text{id})\circ(\Delta\otimes\Delta).

More generally, we can talk about super Hopf GG-coalgebras, now it is Δα,β\Delta_{\alpha,\beta} that satisfies the above property (with τHβ,Hα\tau_{H_{\beta},H_{\alpha}} in place of τH,H\tau_{H,H}). In the super case, Aut(H)\mathrm{Aut}(H) denotes the automorphisms preserving the mod 2 degree. The twisted Drinfeld double of a super Hopf algebra HH is defined as before, but with some additional signs:

(pa)α(qb)\displaystyle(p\otimes a)\cdot_{\alpha}(q\otimes b) =(1)|a(1)|+|q(1)|+|a(2)||q(2)|+|a(1)||q(2)|+|a(2)||q(3)|\displaystyle=(-1)^{|a_{(1)}|+|q_{(1)}|+|a_{(2)}||q_{(2)}|+|a_{(1)}||q_{(2)}|+|a_{(2)}||q_{(3)}|}
a(1),q(3)S1α1(a(3)),q(1)pq(2)a(2)b.\displaystyle\langle a_{(1)},q_{(3)}\rangle\langle S^{-1}\alpha^{-1}(a_{(3)}),q_{(1)}\rangle p\cdot q_{(2)}\otimes a_{(2)}\cdot b.

In this formula, ,\langle\ ,\ \rangle represents the usual vector space pairing, but note that the sign (1)|a(1)|+|q(1)|(-1)^{|a_{(1)}|+|q_{(1)}|} 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 GG-tangles

During this section, we let HH be a finite dimensional Hopf algebra. We suppose D(H)D(H) is ribbon with corresponding pivotal element 𝜷𝒃\boldsymbol{\beta}\otimes\boldsymbol{b} and we let GAut(H)G\subset\mathrm{Aut}(H) be such that D(H)¯|G\underline{D(H)}|_{G} is GG-ribbon. We let (hi)(h_{i}) be a basis of HH and (hi)(h^{i}) be the dual basis of HH^{*}.

Refer to caption
Figure 1. A three strand tangle TT in the block XX.

3.1. GG-tangles

Let X=×[0,1]×[1,)X=\mathbb{R}\times[0,1]\times[-1,\infty) and let pp be a basepoint in ×[0,1]×{1}\mathbb{R}\times[0,1]\times\{-1\}. We think of ×[0,1]×{0}\mathbb{R}\times[0,1]\times\{0\} as being on the plane of the page, so the zz-axis is transversal to the plane and the negative zz-axis is towards the reader. We also denote iX=×{i}×[1,)\partial_{i}X=\mathbb{R}\times\{i\}\times[-1,\infty) for i=0,1i=0,1. By a GG-tangle we mean a framed, oriented tangle TT in ×[0,1]×(0,)\mathbb{R}\times[0,1]\times(0,\infty) endowed with a representation ρ:π1(XT,z)G\rho:\pi_{1}(X_{T},z)\to G where XT=XTX_{T}=X\setminus T. Two GG-tangles (T,ρ)(T,\rho) and (T,ρ)(T^{\prime},\rho^{\prime}) are isotopic if there is a basepoint-preserving isotopy dt:XXd_{t}:X\to X, that is, d0=idX,dt|X=id,d1(T)=Td_{0}=\text{id}_{X},\ d_{t}|_{\partial X}=\text{id},\ d_{1}(T)=T^{\prime} as framed oriented tangles, and ρ(d1)=ρ\rho^{\prime}\circ(d_{1})_{*}=\rho. Here (d1):π1(XT,p)π1(XT,p)(d_{1})_{*}:\pi_{1}(X_{T},p)\to\pi_{1}(X_{T^{\prime}},p) is the induced map. By Van Kampen’s theorem, if (T,ρ)(T,\rho) and (T,ρ)(T^{\prime},\rho^{\prime}) are GG-tangles satisfying that ρj1|π1(1XXT)=ρj0|π0(0XXT)\rho\circ j_{1}|_{\pi_{1}(\partial_{1}X\cap X_{T})}=\rho^{\prime}\circ j_{0}|_{\pi_{0}(\partial_{0}X\cap X_{T^{\prime}})} (where jij_{i} is the homomorphism induced by the corresponding inclusion) then we can stack (T,ρ)(T^{\prime},\rho^{\prime}) on top of (T,ρ)(T,\rho) and rescale to define a new GG-tangle, called the composition of (T,ρ)(T^{\prime},\rho^{\prime}) and (T,ρ)(T,\rho). Similarly, we can stack a GG-tangle to the right of another. With these operations, GG-tangles form (the morphisms of) a monoidal category, and in fact, a GG-crossed ribbon category [31].

3.2. Twisted universal invariants of GG-tangles

Let (T,ρ)(T,\rho) be a GG-tangle with nn open components (and no closed components). We suppose the components of TT are ordered, say T1,,TnT_{1},\dots,T_{n}. We define now the invariant ZD(H)¯ρ(T)D(H)α1D(H)αnZ^{\rho}_{\underline{D(H)}}(T)\in D(H)_{\alpha_{1}}\otimes\dots\otimes D(H)_{\alpha_{n}} where each αi=ρ(μi)\alpha_{i}=\rho(\mu_{i}) and μi\mu_{i} is a meridian at the endpoint of TiT_{i} as defined below. As usual, this invariant is defined from a planar diagram and then shown to be independent of it.

Let DD be an oriented planar diagram of TT where at all crossings both strands are oriented upwards. We assume DD 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 ee of DD, there is an element μeπ1(XT,p)\mu_{e}\in\pi_{1}(X_{T},p) defined as the homotopy class of the loop that starts at the basepoint pp, 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 ρ(μe)G\rho(\mu_{e})\in G, so that all edges of the diagram are labelled by elements of GG:

Refer to caption
Figure 2. On the right α=ρ(μe)\alpha=\rho(\mu_{e}) and β=ρ(μe)\beta=\rho(\mu_{e^{\prime}}).

For each i=1,,ni=1,\dots,n we let μi=μei\mu_{i}=\mu_{e_{i}} where eie_{i} is the edge containing the endpoint of TiT_{i} and we let αi=ρ(μi)G\alpha_{i}=\rho(\mu_{i})\in G. Now let cic_{i} be the core of TiT_{i}, that is, Tici×[0,1]T_{i}\cong c_{i}\times[0,1] as a framed tangle and let ciXTc^{\prime}_{i}\subset X_{T} be a close oriented normal to cic_{i} (this uses the framing of TiT_{i}). Then, for each edge eTie\subset T_{i} of the diagram DD of TT, we let leπ1(XT)l_{e}\in\pi_{1}(X_{T}) be the “partial longitude” defined as the homotopy class of a path that goes linearly from pp to the endpoint of cic^{\prime}_{i}, follows cic^{\prime}_{i} with the opposite orientation until it reaches ee, and finally goes back to the basepoint pp by a linear path:

Refer to caption
Figure 3. The partial longitude lel_{e} of the edge ee.

Note that for each edge eTie\subset T_{i} of DD we have that leμele1=μil_{e}\mu_{e}l_{e}^{-1}=\mu_{i}. For a component TiT_{i} of TT we denote [Ti]=lei[T_{i}]=l_{e^{\prime}_{i}} where eie^{\prime}_{i} is the edge of TiT_{i} containing the starting point of TiT_{i}. Note that both (α1,,αn)(\alpha_{1},\dots,\alpha_{n}) and (ρ([T1]),,ρ([Tn]))(\rho([T_{1}]),\dots,\rho([T_{n}])) are independent of the diagram DD of TT.

We now bring in the twisted Drinfeld double. To each positive crossing of DD whose bottom edges are labelled by α,β\alpha,\beta (as in Figure 2 above), we associate the RR-matrix Rα,βD(H)αD(H)βR_{\alpha,\beta}\in D(H)_{\alpha}\otimes D(H)_{\beta} and the crossing isomorphism φα:D(H)βD(H)αβα1\varphi_{\alpha}:D(H)_{\beta}\to D(H)_{\alpha\beta\alpha^{-1}}. We represent the two factors of the RR-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 φα\varphi_{\alpha} is represented by a white bead:

[Uncaptioned image]

If the crossing is negative and α,β\alpha,\beta are the labels of the edges at the top, we assign the inverse RR-matrix Rα,β1R_{\alpha,\beta}^{-1} and the crossing isomorphism φα1\varphi_{\alpha^{-1}}. These are again represented by two black and one white beads as above. To caps and cups labelled by α\alpha we associate the unit or the GG-pivot as follows:

[Uncaptioned image]

where 𝒈α=rH(α)1𝜷𝒃\boldsymbol{g}_{\alpha}=\sqrt{r_{H}(\alpha)}^{-1}\boldsymbol{\beta}\otimes\boldsymbol{b}. With these conventions, the black beads lying over a given edge ee of the diagram DD belong to D(H)ρ(μe)D(H)_{\rho(\mu_{e})}. Hence, we can multiply all these black beads (from right to left) as we follow the orientation along ee. Thus, we get a single black bead xeD(H)ρ(μe)x_{e}\in D(H)_{\rho(\mu_{e})} for each edge of the diagram. We now follow the orientation of a component TiT_{i} of TT and we multiply all these black beads as follows: if an edge e′′e^{\prime\prime} follows an edge ee^{\prime} and the label of the overpass separating these is α\alpha, then there is a white bead φαϵ\varphi_{\alpha^{\epsilon}} in between the black beads of ee^{\prime} and e′′e^{\prime\prime} (where ϵ=±1\epsilon=\pm 1 is the sign of the crossing). We evaluate φαϵ\varphi_{\alpha^{\epsilon}} on the bead xeD(H)ρ(μe)x_{e^{\prime}}\in D(H)_{\rho(\mu_{e^{\prime}})} and then we multiply this with the bead xe′′D(H)ρ(μe′′)x_{e^{\prime\prime}}\in D(H)_{\rho(\mu_{e^{\prime\prime}})}, this results in xe′′φαϵ(xe)D(H)ρ(μe′′)x_{e^{\prime\prime}}\varphi_{\alpha^{\epsilon}}(x_{e^{\prime}})\in D(H)_{\rho(\mu_{e^{\prime\prime}})}. 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 TT lying after an edge ee is equal to φρ(le)\varphi_{\rho(l_{e})}. Thus, we could equally evaluate φρ(le)(xe)\varphi_{\rho(l_{e})}(x_{e}) for each edge ee, which belongs to D(H)αiD(H)_{\alpha_{i}} for each eTie\subset T_{i}, and then multiply all the resulting beads on a given component (from right to left as we follow the orientation of TiT_{i}). If we do this for all components of TT we get an element

ZD(H)¯ρ(D)D(H)α1D(H)αn.Z_{\underline{D(H)}}^{\rho}(D)\in D(H)_{\alpha_{1}}\otimes\dots\otimes D(H)_{\alpha_{n}}.

This turns out to be independent of the diagram DD chosen, so it is a topological invariant of (T,ρ)(T,\rho), denoted ZD(H)¯ρ(T)Z^{\rho}_{\underline{D(H)}}(T). This is the invariant of GG-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 ZD(H)¯ρ(T)Z_{\underline{D(H)}}^{\rho}(T) as follows: let SS be the set consisting of the overpasses and underpasses of the diagram together with the right caps and cups. Suppose DD has NN crossings and kk right caps or cups, so |S|=2N+k|S|=2N+k. Order the set of crossings of the diagram DD arbitrarily, say c1,,cNc_{1},\dots,c_{N}. Order the set of right caps/cups arbitrarily too. This determines a total order of SS where the overpass of each crossing comes before the underpass and all caps/cups come after the crossings. We denote S1S_{1} the set SS with this order. Let αi,βi\alpha_{i},\beta_{i} be the labels of the bottom edges (resp. top edges) at cic_{i} if the crossing is positive (resp. negative). Let γi\gamma_{i} be the label at the ii-th right cap/cup for i=1,,ki=1,\dots,k and let ϵi\epsilon_{i} be the corresponding power of gγig_{\gamma_{i}}. For each xSx\in S, let D(H)xD(H)_{x} be D(H)αiD(H)_{\alpha_{i}}, D(H)βiD(H)_{\beta_{i}} or D(H)γiD(H)_{\gamma_{i}} depending on whether xx 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 TT, defines another total order on SS, denote it S2S_{2}. In other words, if did_{i} is the number of elements of SS corresponding to TiT_{i}, then the first d1d_{1} elements of S1S_{1} are those of T1T_{1} (ordered from right to left as we follow the orientation of T1T_{1}), the next d2d_{2} elements are those of T2T_{2} ordered in the same way, and so forth. This gives a permutation isomorphism P:xS1D(H)xxS2D(H)xP:\otimes_{x\in S_{1}}D(H)_{x}\to\otimes_{x\in S_{2}}D(H)_{x}. If D(H)¯\underline{D(H)} is a Hopf group-coalgebra in super-vector spaces, this means that signs are being introduced. Now, if

RD=(j=1NRαj,βj)(i=1kgγiϵi)R_{D}=\left(\bigotimes_{j=1}^{N}R_{\alpha_{j},\beta_{j}}\right)\otimes\left(\bigotimes_{i=1}^{k}g_{\gamma_{i}}^{\epsilon_{i}}\right)

(note that this belongs to xS1Dx\otimes_{x\in S_{1}}D_{x}) then

ZD(H)¯ρ(T)=(i=1nmαi(di))P(xS1φρ(lx))(RD),Z_{\underline{D(H)}}^{\rho}(T)=\left(\bigotimes_{i=1}^{n}m_{\alpha_{i}}^{(d_{i})}\right)\circ P\circ\left(\bigotimes_{x\in S_{1}}\varphi_{\rho(l_{x})}\right)(R_{D}),

where mαi(k):D(H)αikD(H)αim_{\alpha_{i}}^{(k)}:D(H)_{\alpha_{i}}^{\otimes k}\to D(H)_{\alpha_{i}} denotes iterated multiplication. Here we denote lx=lexl_{x}=l_{e_{x}} where exe_{x} is the edge of DD containing xx and lexl_{e_{x}} 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 RR-matrices and the pivots have degree zero.

Remark 3.1.

Our motivation to define ZD(H)¯ρ(T)Z^{\rho}_{\underline{D(H)}}(T) for super Hopf algebras is that the SL(n,)SL(n,\mathbb{C})-twisted Reidemeister torsion is the special case when HH is an exterior algebra Λ(n)\Lambda(\mathbb{C}^{n}) [23], which is a super Hopf algebra.

3.3. The abelian case

Suppose GG is an abelian group with an homomorphism ϕ:GAut(H)\phi:G\to\mathrm{Aut}(H). Note that, given a diagram of a GG-tangle (T=T1Tn,ρ)(T=T_{1}\cup\dots\cup T_{n},\rho), all edges of TiT_{i} will be labelled by the same αiG\alpha_{i}\in G. Consider D(H)¯|G\underline{D(H)}|_{G} and the Hopf GG-coalgebra A¯\underline{A} defined in Subsection 2.5. Then we can define a universal invariant ZA¯ρ(T)Aα1AαnZ^{\rho}_{\underline{A}}(T)\in A_{\alpha_{1}}\otimes\dots\otimes A_{\alpha_{n}} almost in the “usual sense”: we use the RR-matrix Rα,βAR^{A}_{\alpha,\beta} on crossings (with labels α,β\alpha,\beta as before) and the pivot of A¯\underline{A} (which is the same as that of D(H)¯\underline{D(H)}) on caps/cups and we run the above procedure. Every bead over a component TiT_{i} of a tangle TT belongs to the same algebra Aρ([Ti])A_{\rho([T_{i}])}, so we can multiply all of these directly. Using the semidirect product relation αx=φα(x)α,αG,xD(H)\alpha\cdot x=\varphi_{\alpha}(x)\cdot\alpha,\ \alpha\in G,\ x\in D(H) and that Rα,βA=hi(hiφα)R^{A}_{\alpha,\beta}=\sum h_{i}\otimes(h^{i}\cdot\varphi_{\alpha}) we can write this invariant as

(1) ZA¯ρ(T)=ZD(H)¯ρ(T)(φρ([T1])φρ([Tn])).\displaystyle Z^{\rho}_{\underline{A}}(T)=Z^{\rho}_{\underline{D(H)}}(T)\cdot(\varphi_{\rho([T_{1}])}\otimes\dots\otimes\varphi_{\rho([T_{n}])}).

This is because the product of all white beads over TiT_{i} equals φρ([Ti])\varphi_{\rho([T_{i}])}.

3.4. The m\mathbb{N}^{m}-graded case

In addition to the previous conditions on HH, suppose that HH is m\mathbb{N}^{m}-graded and let G=Autm0(H)G=\mathrm{Aut}^{0}_{\mathbb{N}^{m}}(H).

Let 𝔽=𝕂[t1±1/2,,tm±1/2]\mathbb{F}=\mathbb{K}[t_{1}^{\pm 1/2},\dots,t_{m}^{\pm 1/2}] and H=H𝕂𝔽H^{\prime}=H\otimes_{\mathbb{K}}\mathbb{F}. For every αAutm(H)\alpha\in\mathrm{Aut}_{\mathbb{N}^{m}}(H) and nn\in\mathbb{Z} we define an 𝔽\mathbb{F}-linear Hopf automorphism of HH^{\prime}, denoted αθnAut(H)\alpha\otimes\theta^{n}\in\mathrm{Aut}(H^{\prime}), by

αθn(x)=t1n|x|1tmn|x|mα(x)\alpha\otimes\theta^{n}(x)=t_{1}^{n|x|_{1}}\dots t_{m}^{n|x|_{m}}\alpha(x)

where (|x|1,,|x|m)(|x|_{1},\dots,|x|_{m}) is the m\mathbb{N}^{m}-degree of xHx\in H. It is easy to see that the map Autm(H)×Aut(H),(α,n)αθn\mathrm{Aut}_{\mathbb{N}^{m}}(H)\times\mathbb{Z}\to\mathrm{Aut}(H^{\prime}),\ (\alpha,n)\mapsto\alpha\otimes\theta^{n} is a homomorphism. In what follows we use the shorthand θ=idθ1Aut(H)\theta=\text{id}\otimes\theta^{1}\in\mathrm{Aut}(H^{\prime}). Let GAutm(H)G^{\prime}\subset\mathrm{Aut}_{\mathbb{N}^{m}}(H^{\prime}) be the subgroup of automorphisms of the form αθn\alpha\otimes\theta^{n} where αG,n\alpha\in G,\ n\in\mathbb{Z}. Since rH(α)=1r_{H}(\alpha)=1, it is easy to see that rH(αθn)=t1n|Λl|1tmn|Λl|mr_{H^{\prime}}(\alpha\otimes\theta^{n})=t_{1}^{n|\Lambda_{l}|_{1}}\dots t_{m}^{n|\Lambda_{l}|_{m}} (where rHr_{H^{\prime}} is as in Subsection 2.4) which has a square root

rH(αθn)=t1n|Λl|1/2tmn|Λl|m/2,\sqrt{r_{H^{\prime}}(\alpha\otimes\theta^{n})}=t_{1}^{n|\Lambda_{l}|_{1}/2}\dots t_{m}^{n|\Lambda_{l}|_{m}/2},

hence D(H)|GD(H^{\prime})|_{G^{\prime}} is GG^{\prime}-ribbon by [23, Prop. 2.4]. The pivot element of D(H)αθD(H^{\prime})_{\alpha\otimes\theta} is

𝒈αθ=t1|Λl|1/2tm|Λl|m/2𝜷𝒃.\boldsymbol{g}_{\alpha\otimes\theta}=t_{1}^{-|\Lambda_{l}|_{1}/2}\dots t_{m}^{-|\Lambda_{l}|_{m}/2}\boldsymbol{\beta}\otimes\boldsymbol{b}.

Now let TT be a tangle as in the previous section and let ρ:π1(XT)G\rho:\pi_{1}(X_{T})\to G be an homomorphism. Then ρ\rho extends to a homomorphism ρθ:π1(XT)GAut(H)\rho\otimes\theta:\pi_{1}(X_{T})\to G^{\prime}\subset\mathrm{Aut}(H^{\prime}) defined by

δρ(δ)θh(δ)\displaystyle\delta\mapsto\rho(\delta)\otimes\theta^{h(\delta)}

for δπ1(XT)\delta\in\pi_{1}(X_{T}), where h:π1(XT)H1(XT)h:\pi_{1}(X_{T})\to H_{1}(X_{T})\to\mathbb{Z} is the homology representation. Note that this depends on the orientation of the components of TT. When ρ1\rho\equiv 1 this is simply the abelian representation sending the canonical generators of H1(XT)H_{1}(X_{T}) to θ\theta. Then the above construction leads to an invariant

ZD(H)¯ρθ(T)D(H)α1θD(H)αnθ.Z_{\underline{D(H^{\prime})}}^{\rho\otimes\theta}(T)\in D(H^{\prime})_{\alpha_{1}\otimes\theta}\otimes\dots\otimes D(H^{\prime})_{\alpha_{n}\otimes\theta}.

In what follows, we identify D(H)αθD(H^{\prime})_{\alpha\otimes\theta} with (HH)[t1±1/2,,tm±1/2](H^{*}\otimes H)[t_{1}^{\pm 1/2},\dots,t_{m}^{\pm 1/2}] in the obvious way.

3.5. Knot polynomials

Suppose HH is m\mathbb{N}^{m}-graded as above. If (Ko,ρ)(K_{o},\rho) is a framed, oriented, (1,1)(1,1)-GG-tangle whose GG-closure is a knot KK, then

ϵD(H)(ZD(H)¯ρθ(Ko))𝔽\epsilon_{D(H^{\prime})}(Z_{\underline{D(H^{\prime})}}^{\rho\otimes\theta}(K_{o}))\in\mathbb{F}

depends on ρ\rho only up to conjugation hence it is an invariant of (K,ρ)(K,\rho), see [23]. Here ϵD(H):D(H)αθ𝔽\epsilon_{D(H^{\prime})}:D(H^{\prime})_{\alpha\otimes\theta}\to\mathbb{F} is given by ϵ(px)=p(1H)ϵH(x)\epsilon(p\otimes x)=p(1_{H})\epsilon_{H}(x) for each pH,xHp\in H^{*},\ x\in H. The map ϵD(H)\epsilon_{D(H^{\prime})} is not an algebra morphism over D(H)αθD(H^{\prime})_{\alpha\otimes\theta} if αid\alpha\neq\text{id} so there is no reason for the above evaluation to be trivial. Note that ZD(H)¯ρθ(Ko)Z_{\underline{D(H^{\prime})}}^{\rho\otimes\theta}(K_{o}) is itself an invariant of (K,ρ)(K,\rho) when ρ\rho is abelian [23], in particular ZD(H)¯θ(Ko)Z_{\underline{D(H^{\prime})}}^{\theta}(K_{o}) is an invariant of KK. If w(Ko)w(K_{o}) is the writhe of KoK_{o} then

PHρθ(K)=tw(Ko)|Λl|/2ϵD(H)(ZD(H)¯ρθ(Ko))P_{H}^{\rho\otimes\theta}(K)=t^{w(K_{o})|\Lambda_{l}|/2}\epsilon_{D(H^{\prime})}(Z_{\underline{D(H^{\prime})}}^{\rho\otimes\theta}(K_{o}))

is an 𝔽\mathbb{F}-valued invariant of the unframed GG-knot (K,ρ)(K,\rho). Here t|Λl|t^{|\Lambda_{l}|} is shorthand for t1|Λl|1tm|Λl|mt_{1}^{|\Lambda_{l}|_{1}}\dots t_{m}^{|\Lambda_{l}|_{m}} where (|Λl|1,,|Λl|m)(|\Lambda_{l}|_{1},\dots,|\Lambda_{l}|_{m}) is the m\mathbb{N}^{m}-degree of the cointegral Λl\Lambda_{l} of HH. When ρ1\rho\equiv 1 this simply becomes

(2) PHθ(K)=tw(Ko)|Λl|/2ϵD(H)(ZD(H)¯θ(Ko))\displaystyle P_{H}^{\theta}(K)=t^{w(K_{o})|\Lambda_{l}|/2}\epsilon_{D(H^{\prime})}(Z_{\underline{D(H^{\prime})}}^{\theta}(K_{o}))

and this is a polynomial invariant of the unframed, oriented knot KK (without any additional structure). It is easy to show that this polynomial actually belongs to 𝕂[t1±1,,tm±1]𝔽\mathbb{K}[t_{1}^{\pm 1},\dots,t_{m}^{\pm 1}]\subset\mathbb{F}.

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 ZD(H)¯ρZ_{\underline{D(H)}}^{\rho}, except that we have to take care of the φα\varphi_{\alpha}’s.

To simplify the statement, we only state the composition rule in the following special case. Suppose TT is a GG-tangle having two components Ti,Ti+1T_{i},T_{i+1} with two endpoints in 1X\partial_{1}X that are adjacent and labelled by the same αG\alpha\in G, see Figure 4. We suppose that, at such endpoints, TiT_{i} is oriented downwards and Ti+1T_{i+1} upwards and the endpoint of TiT_{i} is to the left of that of Ti+1T_{i+1}. Thus, we can compose TT with an α\alpha-labelled left cap on top. Let (T,ρ)(T^{\prime},\rho^{\prime}) be the resulting GG-tangle.

Refer to caption
Figure 4. TT^{\prime} is the composition of strands TiT_{i} and Ti+1T_{i+1}.
Lemma 3.2.

The invariant ZD(H)¯ρ(T)Z_{\underline{D(H)}}^{\rho^{\prime}}(T^{\prime}) is computed by

ZD(H)¯ρ(T)=(id(mγ(idφρ([Ti])))id)(ZD(H)¯ρ(T))Z_{\underline{D(H)}}^{\rho^{\prime}}(T^{\prime})=(\text{id}\otimes(m_{\gamma}\circ(\text{id}\otimes\varphi_{\rho([T_{i}])}))\otimes\text{id})(Z_{\underline{D(H)}}^{\rho}(T))

where mγm_{\gamma} is applied on the factors corresponding to TiT_{i} and Ti+1T_{i+1} and γ\gamma is the GG-label of the endpoint of TiT_{i}.

Proof.

With the above conventions, the new component in TT^{\prime} consists of Ti+1T_{i+1}, with its orientation, followed by TiT_{i}. The bead of this component is obtained by sliding the bead on Ti+1T_{i+1} along TiT_{i}, evaluating on all φβ\varphi_{\beta}’s encountered along TiT_{i}, and multiplying with the bead of TiT_{i}. Since the product of the φβ\varphi_{\beta}’s along TiT_{i} equals φρ([Ti])\varphi_{\rho([T_{i}])}, this proves the lemma. ∎

Now let T0T_{0} be a component of a GG-tangle (T,ρ)(T,\rho). Let TT^{\prime} be the tangle obtained by doubling T0T_{0} along the framing. Suppose TT^{\prime} is a GG-tangle with ρ:π1(XT)G\rho^{\prime}:\pi_{1}(X_{T^{\prime}})\to G and let ρ=ρ|π1(XT)\rho=\rho^{\prime}|_{\pi_{1}(X_{T})}. Let ρ(μ1)=α,ρ(μ2)=β\rho^{\prime}(\mu_{1})=\alpha,\rho^{\prime}(\mu_{2})=\beta where μ1,μ2\mu_{1},\mu_{2} are the meridians of the two copies of T0T_{0} in TT^{\prime} (say, at the endpoints of the copies of T0T_{0}). Then ρ(μ)=αβ\rho(\mu)=\alpha\beta where μ\mu is the meridian corresponding to T0T_{0}.

Refer to caption
Figure 5. On the left is the GG-tangle (T,ρ)(T,\rho). For simplicity, we just depict the component T0TT_{0}\subset T, but TT may have any number of components. On the right is (T,ρ)(T^{\prime},\rho^{\prime}) and the meridians μ1,μ2\mu_{1},\mu_{2}, which have linking number -1 with the corresponding strands of TT^{\prime}.
Lemma 3.3.

If TT^{\prime} is constructed by duplicating T0TT_{0}\subset T as above, then

(idΔα,βid)(ZD(H)¯ρ(T))=ZD(H)¯ρ(T)(\text{id}\otimes\dots\otimes\Delta_{\alpha,\beta}\otimes\dots\otimes\text{id})(Z_{\underline{D(H)}}^{\rho}(T))=Z_{\underline{D(H)}}^{\rho^{\prime}}(T^{\prime})

where Δα,β\Delta_{\alpha,\beta} is applied on the tensor factor corresponding to T0T_{0}.

Proof.

We need to check this on the elementary tangles. For the crossing tangles, this follows by the defining property of the graded RR-matrix. For right caps and cups, it follows from Δα,β(𝒈αβ)=𝒈α𝒈β\Delta_{\alpha,\beta}(\boldsymbol{g}_{\alpha\beta})=\boldsymbol{g}_{\alpha}\otimes\boldsymbol{g}_{\beta} and for left caps/cups it is obvious. The fact that the φα\varphi_{\alpha}’s are Hopf isomorphisms implies that the property holds for compositions of elementary tangles, thus for all tangles. ∎

Lemma 3.4.

Let T0TT_{0}\subset T be a component oriented upwards, that is, it begins on 0X\partial_{0}X and ends on 1X\partial_{1}X. Let δ\delta be the label at the endpoint of T0T_{0}. Let TT^{\prime} be the tangle obtained by reversing the orientation of T0T_{0}. If we use the same ρ\rho for both tangles, then

(idφρ([T0])1Sδid)(ZD(H)¯ρ(T))=ZD(H)¯ρ(T)(\text{id}\otimes\dots\otimes\varphi_{\rho([T_{0}])}^{-1}\circ S_{\delta}\otimes\dots\otimes\text{id})(Z_{\underline{D(H)}}^{\rho}(T))=Z_{\underline{D(H)}}^{\rho}(T^{\prime})

where φρ([T0])1Sδ\varphi^{-1}_{\rho([T_{0}])}\circ S_{\delta} is applied on the tensor factor corresponding to T0T_{0}.

Remark 3.5.

Recall that in our definition of ZD(H)¯ρZ^{\rho}_{\underline{D(H)}} we “gather” the beads at the endpoint of each component and then multiply. If we reverse the orientation of T0T_{0} and we gather the beads at the same point as before (which is the starting point of T0T^{\prime}_{0}), then we just need to apply the antipode SδS_{\delta}, but if we want to gather at the endpoint of T0T^{\prime}_{0} we need to slide our bead down through all of T0T_{0}, this gives the additional φρ([T0])1\varphi_{\rho([T_{0}])}^{-1} above.

Proof.

Suppose first we reverse the understrand of a positive crossing (oriented upwards) whose bottom labels are α,β\alpha,\beta. Thus, δ=αβα1\delta=\alpha\beta\alpha^{-1} and [T0]=α[T_{0}]=\alpha. Then the invariant is computed as on the left of the following figure:

[Uncaptioned image]

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:

α(S(hi))𝒈β1hiφα1(𝒈αβ1α11)\displaystyle\alpha(S(h_{i}))\otimes\boldsymbol{g}_{\beta^{-1}}h^{i}\varphi_{\alpha^{-1}}(\boldsymbol{g}^{-1}_{\alpha\beta^{-1}\alpha^{-1}}) =α(S(hi))𝒈β1hi𝒈β11\displaystyle=\alpha(S(h_{i}))\otimes\boldsymbol{g}_{\beta^{-1}}h^{i}\boldsymbol{g}^{-1}_{\beta^{-1}}
=α(S(hi))SβSβ1(hi)\displaystyle=\alpha(S(h_{i}))\otimes S_{\beta}S_{\beta^{-1}}(h^{i})
=α(S(hi))S2(hi)\displaystyle=\alpha(S(h_{i}))\otimes S^{2}(h^{i})
=α(hi)S(hi).\displaystyle=\alpha(h_{i})\otimes S(h^{i}).

But the picture on the right is exactly obtained by applying φα1Sδ\varphi_{\alpha}^{-1}S_{\delta} to the strand that is being reversed (note that SδS_{\delta} is simply SHS_{H^{*}} on HH^{*}). Similarly, if we reverse the overstrand of a positive crossing, the invariant is computed as on the left of the following figure:

[Uncaptioned image]

Here we have δ=α\delta=\alpha and [T0]=1[T_{0}]=1. It is easy to see that the right hand side is obtained by applying φ11Sα=Sα\varphi_{1}^{-1}S_{\alpha}=S_{\alpha} 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 T0T_{0} that has no caps or cups.

Now, suppose T0T_{0} has caps and cups. We will subdivide (a diagram of) T0T_{0} in subarcs a1,,aka_{1},\dots,a_{k} without caps/cups. Thus, as we follow the orientation of T0T_{0}, we follow first a1a_{1}, then a cap, then a2a_{2}, then a cup, and so on until aka_{k}. Each aia_{i} with odd ii is oriented upwards and for even ii it is oriented downwards. We let αi\alpha_{i} be the label at the starting point of aia_{i}, thus, αi+1\alpha_{i+1} is the label of the endpoint of each aia_{i} and we let α=α1\alpha=\alpha_{1} and αk+1=δ\alpha_{k+1}=\delta. We also let γi\gamma_{i} be the image of the partial longitude from the endpoint of T0T_{0} to the top of aia_{i}, and γi\gamma^{\prime}_{i} be the partial longitude for T0T^{\prime}_{0}, that is, γi\gamma^{\prime}_{i} goes from the starting point of T0T_{0} to the top of aia_{i}. We also denote γ=[T0]\gamma=[T_{0}], so γi(γi)1=γ\gamma_{i}(\gamma^{\prime}_{i})^{-1}=\gamma for each ii. For each ii (even or odd) we let xix_{i} be the total bead for that arc as if the arc was oriented upwards (and the bead is gathered at the top so xiD(H)αi+1x_{i}\in D(H)_{\alpha_{i+1}} for odd ii, xiD(H)αi1x_{i}\in D(H)_{\alpha_{i}^{-1}} for even ii). Thus, by what was shown above, for even ii the actual bead along that arc is Sαi1(xi)S_{\alpha_{i}^{-1}}(x_{i}) (if the bead is gathered at the top, see Remark 3.5 above). Then, the T0T_{0} tensor factor of ZD(H)¯ρ(T)Z_{\underline{D(H)}}^{\rho}(T) can be written as

xk𝒈βϵk𝒈βϵ3φγ2(Sα21(x2))𝒈βϵ2φγ1(x1)x_{k}\boldsymbol{g}_{\beta}^{\epsilon_{k}}\dots\boldsymbol{g}_{\beta}^{\epsilon_{3}}\varphi_{\gamma_{2}}(S_{\alpha_{2}^{-1}}(x_{2}))\boldsymbol{g}_{\beta}^{\epsilon_{2}}\varphi_{\gamma_{1}}(x_{1})

where ϵi\epsilon_{i} is the power of the pivot at the cap/cup right before aia_{i} (so ϵi{0,±1}\epsilon_{i}\in\{0,\pm 1\}). Now we apply φγ1Sδ=Sαφγ1\varphi_{\gamma^{-1}}S_{\delta}=S_{\alpha}\varphi_{\gamma^{-1}}. First we apply φγ1\varphi_{\gamma^{-1}} this results in

φγk(xk)𝒈αϵk𝒈αϵ3φγ2(Sα21(x2))𝒈αϵ2φγ1(x1).\varphi_{\gamma^{\prime}_{k}}(x_{k})\boldsymbol{g}_{\alpha}^{\epsilon_{k}}\dots\boldsymbol{g}_{\alpha}^{\epsilon_{3}}\varphi_{\gamma^{\prime}_{2}}(S_{\alpha_{2}^{-1}}(x_{2}))\boldsymbol{g}_{\alpha}^{\epsilon_{2}}\varphi_{\gamma^{\prime}_{1}}(x_{1}).

Applying SαS_{\alpha} gives

Sα(φγk(xk)𝒈αϵk𝒈αϵ3φγ2(Sα21(x2))𝒈αϵ2φγ1(x1))\displaystyle S_{\alpha}(\varphi_{\gamma^{\prime}_{k}}(x_{k})\boldsymbol{g}_{\alpha}^{\epsilon_{k}}\dots\boldsymbol{g}_{\alpha}^{\epsilon_{3}}\varphi_{\gamma^{\prime}_{2}}(S_{\alpha_{2}^{-1}}(x_{2}))\boldsymbol{g}_{\alpha}^{\epsilon_{2}}\varphi_{\gamma^{\prime}_{1}}(x_{1}))
=Sαφγ1(x1)𝒈α1ϵ2Sαφγ2Sα21(x2)𝒈α1ϵ3𝒈α1ϵkφγkSαφγk(xk)\displaystyle=S_{\alpha}\varphi_{\gamma^{\prime}_{1}}(x_{1})\boldsymbol{g}_{\alpha^{-1}}^{-\epsilon_{2}}S_{\alpha}\varphi_{\gamma^{\prime}_{2}}S_{\alpha_{2}^{-1}}(x_{2})\boldsymbol{g}_{\alpha^{-1}}^{-\epsilon_{3}}\dots\boldsymbol{g}_{\alpha^{-1}}^{-\epsilon_{k}}\varphi_{\gamma^{\prime}_{k}}S_{\alpha}\varphi_{\gamma^{\prime}_{k}}(x_{k})
=φγ1(Sα2(x1))𝒈α1ϵ2φγ2(Sα2Sα21(x2))𝒈α1ϵ3𝒈α1ϵkφγkSδ(xk)\displaystyle=\varphi_{\gamma^{\prime}_{1}}(S_{\alpha_{2}}(x_{1}))\boldsymbol{g}_{\alpha^{-1}}^{-\epsilon_{2}}\varphi_{\gamma^{\prime}_{2}}(S_{\alpha_{2}}S_{\alpha_{2}^{-1}}(x_{2}))\boldsymbol{g}_{\alpha^{-1}}^{-\epsilon_{3}}\dots\boldsymbol{g}_{\alpha^{-1}}^{-\epsilon_{k}}\varphi_{\gamma^{\prime}_{k}}S_{\delta}(x_{k})
=φγ1(Sα2(x1))𝒈α1ϵ2φγ2(𝒈α21x2𝒈α211)𝒈α1ϵ3𝒈α1ϵkφγkSδ(xk)\displaystyle=\varphi_{\gamma^{\prime}_{1}}(S_{\alpha_{2}}(x_{1}))\boldsymbol{g}_{\alpha^{-1}}^{-\epsilon_{2}}\varphi_{\gamma^{\prime}_{2}}(\boldsymbol{g}_{\alpha_{2}^{-1}}x_{2}\boldsymbol{g}_{\alpha_{2}^{-1}}^{-1})\boldsymbol{g}_{\alpha^{-1}}^{-\epsilon_{3}}\dots\boldsymbol{g}_{\alpha^{-1}}^{-\epsilon_{k}}\varphi_{\gamma^{\prime}_{k}}S_{\delta}(x_{k})
=φγ1(Sα2(x1))𝒈α1ϵ2+1φγ2(x2)𝒈α11ϵ3𝒈α1ϵkφγkSδ(xk)\displaystyle=\varphi_{\gamma^{\prime}_{1}}(S_{\alpha_{2}}(x_{1}))\boldsymbol{g}_{\alpha^{-1}}^{-\epsilon_{2}+1}\varphi_{\gamma^{\prime}_{2}}(x_{2})\boldsymbol{g}_{\alpha^{-1}}^{-1-\epsilon_{3}}\dots\boldsymbol{g}_{\alpha^{-1}}^{-\epsilon_{k}}\varphi_{\gamma^{\prime}_{k}}S_{\delta}(x_{k})

We used that SβSβ1(x)=𝒈β1x𝒈β11S_{\beta}S_{\beta^{-1}}(x)=\boldsymbol{g}_{\beta^{-1}}x\boldsymbol{g}_{\beta^{-1}}^{-1} for each xD(H)β1x\in D(H)_{\beta^{-1}} and β\beta and that Sαφγi=φγiS(γi)1αγiS_{\alpha}\varphi_{\gamma^{\prime}_{i}}=\varphi_{\gamma^{\prime}_{i}}S_{(\gamma^{\prime}_{i})^{-1}\alpha\gamma^{\prime}_{i}} and (γi)1αγi=αi+1(\gamma^{\prime}_{i})^{-1}\alpha\gamma^{\prime}_{i}=\alpha_{i+1} for odd ii or =αi=\alpha_{i} for even ii. Note that ϵi+ϵi=1\epsilon^{\prime}_{i}+\epsilon_{i}=1 for each even ii (which correspond to caps) and ϵi+ϵi=1\epsilon^{\prime}_{i}+\epsilon_{i}=-1 for each odd ii (corresponding to cups), where ϵi\epsilon^{\prime}_{i} is the power of the pivot at the given cap/cup when the orientation is reversed. Therefore, the above equals

φγ1(Sα2(x1))𝒈α1ϵ2φγ2(x2)𝒈α1ϵ3𝒈α1ϵkφγkSδ(xk)\varphi_{\gamma^{\prime}_{1}}(S_{\alpha_{2}}(x_{1}))\boldsymbol{g}_{\alpha^{-1}}^{\epsilon_{2}^{\prime}}\varphi_{\gamma^{\prime}_{2}}(x_{2})\boldsymbol{g}_{\alpha^{-1}}^{\epsilon_{3}^{\prime}}\dots\boldsymbol{g}_{\alpha^{-1}}^{\epsilon_{k}^{\prime}}\varphi_{\gamma^{\prime}_{k}}S_{\delta}(x_{k})

which is exactly the bead of T0T_{0} with its inverse orientation. ∎

Now let T0T_{0} be a component of a tangle TT as above, but suppose now that both endpoints of T0T_{0} lie on 1X\partial_{1}X and that the endpoint of T0T_{0} is to the right of its starting point. Let S~α:D(H)αD(H)α1\widetilde{S}_{\alpha}:D(H)_{\alpha}\to D(H)_{\alpha^{-1}} be defined by

(3) S~α(x)=Sα(x)𝒈α11.\displaystyle\widetilde{S}_{\alpha}(x)=S_{\alpha}(x)\boldsymbol{g}^{-1}_{\alpha^{-1}}.

Let δ\delta be the label of the endpoint of T0T_{0}. If TT^{\prime} is obtained by reversing the orientation of T0T_{0} then the above lemma implies that

(4) (idφρ([T0])1S~δid)(ZD(H)¯ρ(T))=ZD(H)¯ρ(T).\displaystyle(\text{id}\otimes\dots\otimes\varphi^{-1}_{\rho([T_{0}])}\widetilde{S}_{\delta}\otimes\dots\otimes\text{id})(Z_{\underline{D(H)}}^{\rho}(T))=Z_{\underline{D(H)}}^{\rho}(T^{\prime}).

When we use the abelian representation θT=θ\theta_{T}=\theta one has to be a bit more careful. This representation depends on the orientation of each component, in particular θT\theta_{T^{\prime}} is not equal to θT\theta_{T}. The beads at a negative crossing (of the first type) are given as follows:

[Uncaptioned image]

3.7. Twist knots

To illustrate our constructions we will compute the twisted universal invariant explicitly on the family of twist knots. Let KnK_{n} be the 2n2n-twist knot. This is an alternating knot with 2n+22n+2 crossings and genus 1. The knot K1K_{1} is the figure-eight knot, K2K_{2} (Stevedore’s knot) is drawn on the left of Figure 7. For simplicity, we suppose m=1m=1, that is, HH is an \mathbb{N}-graded Hopf algebra. Recall that the pivot of D(H)θD(H^{\prime})_{\theta} is td(H)/2𝒈t^{-d(H)/2}\boldsymbol{g} where 𝒈=𝜷𝒃\boldsymbol{g}=\boldsymbol{\beta}\otimes\boldsymbol{b} is the pivot of D(H)D(H) (note that d(H)=|Λl|d(H)=|\Lambda_{l}| for m=1m=1). We also take ρ1\rho\equiv 1 (but we keep θ:H1(XK)Aut(H)\theta:H_{1}(X_{K})\to\mathrm{Aut}(H)). Thus all white beads in our diagrams will be φθ±1\varphi_{\theta^{\pm 1}}, we will denote these simply by θ±1\theta^{\pm 1}. We will show directly that degPHθ(Kn)2d(H)\mathrm{deg}\ P_{H}^{\theta}(K_{n})\leq 2d(H) for any nn. This is in agreement with the genus bound, Theorem 1, since as mentioned the genus of every twist knot is 11. In what follows we denote Zθ=ZD(H)¯θ(Kn)Z^{\theta}=Z_{\underline{D(H^{\prime})}}^{\theta}(K_{n}).

Refer to caption
Figure 6. The figure eight knot and a 2-bridge presentation (opened to be a (1,1)(1,1)-tangle). The black dots represent elements of H,HH,H^{*}, the white dots represent θ±1\theta^{\pm 1}’s and the orange dots on right caps (resp. right cups) represent 𝒈=td(H)/2𝜷𝒃\boldsymbol{g}=t^{-d(H)/2}\boldsymbol{\beta}\otimes\boldsymbol{b} (resp. 𝒈1\boldsymbol{g}^{-1}). Here θ¯\overline{\theta} denotes θ1\theta^{-1}.

Let’s begin with the figure eight knot. We compute ZθZ^{\theta} 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 ZθZ^{\theta}. Note that this diagram has rotation number zero so all the powers of tt from pivots cancel out, in other words, we can simply put the untwisted pivot 𝒈=𝜷𝒃\boldsymbol{g}=\boldsymbol{\beta}\otimes\boldsymbol{b} on right caps and 𝒈1\boldsymbol{g}^{-1} on right cups, which is what we do. The diagram also has writhe zero so that ϵ(Zθ)=PHθ(K1)\epsilon(Z^{\theta})=P_{H}^{\theta}(K_{1}). In a bridge diagram, other from the 𝒈±1\boldsymbol{g}^{\pm 1}’s on caps/cups, all the beads over an overarc are in HH and all beads on an underarc are elements of HH^{*}. All white beads, which are θ±1\theta^{\pm 1}, lie on underarcs. We number the crossings from left to right as shown in Figure 6. If hjhj\sum h_{j}\otimes h^{j} represents the RR-matrix (of D(H)D(H)) on the jj-th crossing, then the invariant is given by (the using Einstein summation convention to hide the eightfold sum):

ZD(H)¯θ(K1)\displaystyle Z_{\underline{D(H^{\prime})}}^{\theta}(K_{1}) =h1h2θ1(h3h4)θS(h6)𝒈h1𝒈h8𝒈1S(h3)θh5h6θ1(h7h8)θ𝒈1S(h2)h5h4S(h7)\displaystyle=h^{1}h^{2}\theta^{-1}(h^{3}h^{4})\cdot_{\theta}S(h_{6})\boldsymbol{g}h_{1}\boldsymbol{g}h_{8}\boldsymbol{g}^{-1}S(h_{3})\cdot_{\theta}h^{5}h^{6}\theta^{-1}(h^{7}h^{8})\cdot_{\theta}\boldsymbol{g}^{-1}S(h_{2})h_{5}h_{4}S(h_{7})

Here θ\cdot_{\theta} denotes the multiplication of the twisted double D(H)θD(H^{\prime})_{\theta}. Note that

ϵD(H)(pθxθa)=ϵH(p)ϵH(a)ϵD(H)(x)\epsilon_{D(H^{\prime})}(p\cdot_{\theta}x\cdot_{\theta}a)=\epsilon_{H^{*}}(p)\epsilon_{H}(a)\epsilon_{D(H^{\prime})}(x)

for any pH,aH,xD(H)θp\in H^{*},a\in H,x\in D(H^{\prime})_{\theta}. Thus ϵ(Zθ)\epsilon(Z^{\theta}) equals

ϵ(h1h2θ1(h3h4))ϵ(𝒃1S(h2)h5h4S(h7))ϵ[(S(h6)𝒈h1𝒈h8𝒈1S(h3))θ(h5h6θ1(h7h8)𝜷1)]\displaystyle\epsilon(h^{1}h^{2}\theta^{-1}(h^{3}h^{4}))\epsilon(\boldsymbol{b}^{-1}S(h_{2})h_{5}h_{4}S(h_{7}))\epsilon[(S(h_{6})\boldsymbol{g}h_{1}\boldsymbol{g}h_{8}\boldsymbol{g}^{-1}S(h_{3}))\cdot_{\theta}(h^{5}h^{6}\theta^{-1}(h^{7}h^{8})\boldsymbol{\beta}^{-1})]
=ϵ[S(h6)𝒈2h8𝒈1θh6θ1(h8)𝜷1].\displaystyle=\epsilon[S(h_{6})\boldsymbol{g}^{2}h_{8}\boldsymbol{g}^{-1}\cdot_{\theta}h^{6}\theta^{-1}(h^{8})\boldsymbol{\beta}^{-1}].

Here we used that ϵ\epsilon is an algebra morphism over HH and over HH^{*} (though not over all of D(H)θD(H^{\prime})_{\theta}) and that ϵ(hj)hj=hjϵ(hj)=1\epsilon(h_{j})h^{j}=h_{j}\epsilon(h^{j})=1 for each jj. Note that

ϵ(xθy)=S1θ1(x(2)),y(1)x(1),y(2)\epsilon(x^{\prime}\cdot_{\theta}y^{\prime})=\langle S^{-1}\theta^{-1}(x^{\prime}_{(2)}),y^{\prime}_{(1)}\rangle\langle x^{\prime}_{(1)},y^{\prime}_{(2)}\rangle

for each xH,yHx^{\prime}\in H,y^{\prime}\in H^{*}. Thus, writing S(h6)𝒈2h8𝒈1=𝜷xS(h_{6})\boldsymbol{g}^{2}h_{8}\boldsymbol{g}^{-1}=\boldsymbol{\beta}x^{\prime} with xHx^{\prime}\in H and y=h6h8𝜷1Hy^{\prime}=h^{6}h^{8}\boldsymbol{\beta}^{-1}\in H^{*} (x,yx^{\prime},y^{\prime} have no meaning independently, but xyx^{\prime}\otimes y^{\prime} does) and using θ1(hi)=t|hi|hi=t|hi|hi\theta^{-1}(h^{i})=t^{-|h^{i}|}h^{i}=t^{|h_{i}|}h^{i} we obtain

ϵ(Zθ)\displaystyle\epsilon(Z^{\theta}) =ϵ(𝜷xθt|h8|y)\displaystyle=\epsilon(\boldsymbol{\beta}x^{\prime}\cdot_{\theta}t^{|h_{8}|}y^{\prime})
=t|h8|ϵ(xθy)\displaystyle=t^{|h_{8}|}\epsilon(x^{\prime}\cdot_{\theta}y^{\prime})
=t|h8|S1θ1(x(2)),y(1)x(1),y(2)\displaystyle=t^{|h_{8}|}\langle S^{-1}\theta^{-1}(x^{\prime}_{(2)}),y^{\prime}_{(1)}\rangle\langle x^{\prime}_{(1)},y^{\prime}_{(2)}\rangle
=t|h8||x(2)|S1(x(2)),y(1)x(1),y(2).\displaystyle=t^{|h_{8}|-|x^{\prime}_{(2)}|}\langle S^{-1}(x^{\prime}_{(2)}),y^{\prime}_{(1)}\rangle\langle x^{\prime}_{(1)},y^{\prime}_{(2)}\rangle.

Note that S1(x(2)),y(1)x(1),y(2)𝕂\langle S^{-1}(x^{\prime}_{(2)}),y^{\prime}_{(1)}\rangle\langle x^{\prime}_{(1)},y^{\prime}_{(2)}\rangle\in\mathbb{K}, so the maximal/minimal power of tt in ϵ(Zθ)\epsilon(Z^{\theta}) is the maximum/minimum of |h8||x(2)||h_{8}|-|x^{\prime}_{(2)}|. But since 0|h8|d(H),0|x(2)|d(H)0\leq|h_{8}|\leq d(H),0\leq|x^{\prime}_{(2)}|\leq d(H) (this uses that HH is \mathbb{N}-graded, not simply \mathbb{Z}-graded) it follows that d(H)|h8||x(2)|d(H)-d(H)\leq|h_{8}|-|x^{\prime}_{(2)}|\leq d(H), hence ϵ(Zθ)\epsilon(Z^{\theta}) is a polynomial of degree 2d(H)\leq 2d(H).

We now consider the case of general KnK_{n}. The knot KnK_{n} has a 2-bridge presentation as shown on the right of Figure 7, but with n1n-1 full rotations of the “middle band”. This diagram has rotation number and writhe zero as before so we can simply use 𝒈±1\boldsymbol{g}^{\pm 1} on right caps/cups. Let b1,a1,b2,a2b_{1},a_{1},b_{2},a_{2} be the arcs of the bridge presentation as one follows the orientation of KK. In Figure 7, b1b_{1} is blue and b2b_{2} is green. As in the example of the figure-eight knot, the product of all elements along a1a_{1} (resp. a2a_{2}) is an element qHq\in H^{*} (resp. pHp\in H^{*}). The beads over b1,b2b_{1},b_{2} are all in HH, except for the pivots 𝒈=𝜷𝒃\boldsymbol{g}=\boldsymbol{\beta}\otimes\boldsymbol{b} (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 b2,b1b_{2},b_{1} have the form 𝜷c,𝜷1d\boldsymbol{\beta}c,\boldsymbol{\beta}^{-1}d for some c,dHc,d\in H. Then the invariant is Zθ=pθ𝜷cθqθ𝜷1dZ^{\theta}=p\cdot_{\theta}\boldsymbol{\beta}c\cdot_{\theta}q\cdot_{\theta}\boldsymbol{\beta}^{-1}d and

ϵD(H)(Zθ)=ϵ(p𝜷)ϵ(d)ϵ(cθq).\epsilon_{D(H^{\prime})}(Z^{\theta})=\epsilon(p\boldsymbol{\beta})\epsilon(d)\epsilon(c\cdot_{\theta}q).

As before, hjϵ(hj)=ϵ(hj)hj=1D(H)h_{j}\epsilon(h^{j})=\epsilon(h_{j})h^{j}=1_{D(H)} so that evaluating p,dp,d on ϵD(H)\epsilon_{D(H^{\prime})} has the effect of killing all the D(H)D(H)-beads coming from a2a_{2} or b1b_{1}. The only beads that remain are those of the crossings between a1a_{1} and b2b_{2}, the θ±1\theta^{\pm 1}’s on a1a_{1} and the pivots 𝒈±1\boldsymbol{g}^{\pm 1}’s on b2b_{2}. In other words, except for the pivots, all the remaining beads are shown below (we denote θ¯=θ1\overline{\theta}=\theta^{-1}):

[Uncaptioned image]

Order the crossings between a1a_{1} and b2b_{2} by 1,2,,2n1,2,\dots,2n from left to right. Let hjhjh_{j}\otimes h^{j} be the D(H)D(H)-component of the RR-matrix on the jj-crossing. Then, the product of all beads along the red arc is

y\displaystyle y =h1θ1(h2)h3θ1(h4)θ1(h2n)\displaystyle=h^{1}\theta^{-1}(h^{2})h^{3}\theta^{-1}(h^{4})\dots\theta^{-1}(h^{2n})
=t|h2|++|h2n|y\displaystyle=t^{|h_{2}|+\dots+|h_{2n}|}y^{\prime}

where y=h1h2h2nHy^{\prime}=h^{1}h^{2}\dots h^{2n}\in H^{*}. The product along b2b_{2} is

x\displaystyle x =S(h1)𝒈S(h2n1)𝒈𝒈h2n𝒈1h4𝒈1h2𝒈1.\displaystyle=S(h_{1})\boldsymbol{g}\dots S(h_{2n-1})\boldsymbol{g}\cdot\boldsymbol{g}\cdot h_{2n}\boldsymbol{g}^{-1}\dots h_{4}\boldsymbol{g}^{-1}h_{2}\boldsymbol{g}^{-1}.

Since S2(a)=𝒈a𝒈1S^{2}(a)=\boldsymbol{g}a\boldsymbol{g}^{-1} for any aHa\in H, this has the form x=𝜷xx=\boldsymbol{\beta}x^{\prime} with xHx^{\prime}\in H. Thus

ϵ(Zθ)\displaystyle\epsilon(Z^{\theta}) =ϵ(xθy)=ϵ(xθy)\displaystyle=\epsilon(x\cdot_{\theta}y)=\epsilon(x^{\prime}\cdot_{\theta}y)
=S1θ1(x(2)),y(1)x(1),y(2)\displaystyle=\langle S^{-1}\theta^{-1}(x^{\prime}_{(2)}),y_{(1)}\rangle\langle x^{\prime}_{(1)},y_{(2)}\rangle
=t|h2|+|h4|++|h2n||x(2)|S1(x(2)),y(1)x(1),y(2)\displaystyle=t^{|h_{2}|+|h_{4}|+\dots+|h_{2n}|-|x^{\prime}_{(2)}|}\langle S^{-1}(x^{\prime}_{(2)}),y^{\prime}_{(1)}\rangle\langle x^{\prime}_{(1)},y^{\prime}_{(2)}\rangle

The term S1(x(2)),y(1)x(1),y(2)𝕂\langle S^{-1}(x^{\prime}_{(2)}),y^{\prime}_{(1)}\rangle\langle x^{\prime}_{(1)},y^{\prime}_{(2)}\rangle\in\mathbb{K}. Now, since HH is \mathbb{N}-graded the product h1h2nh^{1}\dots h^{2n} is zero if |h1|++|h2n|>d(H)|h_{1}|+\dots+|h_{2n}|>d(H), so the only terms that contribute to the above sum are those for which 0|h2|++|h2n|d(H)0\leq|h_{2}|+\dots+|h_{2n}|\leq d(H). Since also 0|x(2)|d(H)0\leq|x^{\prime}_{(2)}|\leq d(H), it follows that the degree of ϵ(Zθ)\epsilon(Z^{\theta}) is 2d(H)\leq 2d(H) as we wanted.

Refer to caption
Figure 7. On the left is the twist knot K2K_{2} and on the right, a bridge presentation of K2K_{2}, opened to be a (1,1)(1,1)-tangle. The horizontal red arcs are supposed to be the underarcs.

4. Proof of Theorem 1

Let HH be a finite dimensional m\mathbb{N}^{m}-graded Hopf algebra over 𝕂\mathbb{K} and let H=H𝕂𝔽H^{\prime}=H\otimes_{\mathbb{K}}\mathbb{F} where 𝔽=𝕂[t1±1/2,,tm±1/2]\mathbb{F}=\mathbb{K}[t_{1}^{\pm 1/2},\dots,t_{m}^{\pm 1/2}]. In what follows we denote G=Autm0(H)Aut(H)G=\mathrm{Aut}^{0}_{\mathbb{N}^{m}}(H)\subset\mathrm{Aut}(H) which we see as a subgroup of GAut(H)G^{\prime}\subset\mathrm{Aut}(H^{\prime}) via ααθ0\alpha\mapsto\alpha\otimes\theta^{0} and we denote D(H)αθ0D(H^{\prime})_{\alpha\otimes\theta^{0}} simply by D(H)αD(H^{\prime})_{\alpha}, that is, D(H)α=D(H)α𝔽D(H^{\prime})_{\alpha}=D(H)_{\alpha}\otimes\mathbb{F} for αG\alpha\in G. The reader interested on the case of ADO may simply suppose m=1m=1 and G=1G=1.

4.1. Formulas from Seifert surfaces

Let SS be a Seifert surface for KK and let g=g(S)g=g(S). We suppose KK is 0-framed, this is enough for our theorem. After an isotopy, we can suppose SS is obtained by thickening a 2g2g-component (framed) tangle TT all of whose endpoints lie on 1X\partial_{1}X and attaching a disk to the top as in Figure 8. We suppose this disk comes with a marked point pp on its boundary and let KoK_{o} be the (1,1)(1,1)-tangle obtained by opening K=SK=\partial S at that point. We denote the components of TT by T1,,T2gT_{1},\dots,T_{2g} from left to right and we suppose that each TiT_{i} is oriented “to the right”, that is, the orientation at the rightmost endpoint is upward pointing. Then, KoK_{o} is obtained by first doubling all the components of TT (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 ZD(H)(K)=ZD(H)(Ko)D(H)Z_{D(H)}(K)=Z_{D(H)}(K_{o})\in D(H) is obtained from ZD(H)(T)D(H)2gZ_{D(H)}(T)\in D(H)^{\otimes 2g} by applying the coproduct to each tensor factor, applying antipodes on the even tensor factors (with a pivot) and finally multiplying:

(5) ZD(H)(Ko)=mD(H)(4g1)Pg(idD(H)S~D(H))2gΔD(H)2g(ZD(H)(T))\displaystyle Z_{D(H)}(K_{o})=m_{D(H)}^{(4g-1)}\circ P^{\otimes g}\circ(\text{id}_{D(H)}\otimes\widetilde{S}_{D(H)})^{\otimes 2g}\circ\Delta_{D(H)}^{\otimes 2g}\left(Z_{D(H)}(T)\right)

where PP is the permutation map P(uvxy)=vxuyP(u\otimes v\otimes x\otimes y)=v\otimes x\otimes u\otimes y (with signs in the super case) and S~D(H)(x)=SD(H)(x)𝒈1\widetilde{S}_{D(H)}(x)=S_{D(H)}(x)\boldsymbol{g}^{-1}. 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 2×1\mathbb{R}^{2}\times 1. This idea comes back to Habiro [16]. However, what really makes our theorem work is the appearance of the representation ρθ\rho\otimes\theta, or rather, the degree twist θ\theta. The essential observation is that, when ρθ\rho\otimes\theta is restricted to π1(XT)\pi_{1}(X_{T}), θ\theta cancels out, that is

ρθ|π1(XT)=ρ|π1(XT).\rho\otimes\theta|_{\pi_{1}(X_{T})}=\rho|_{\pi_{1}(X_{T})}.

This is because the orientations of KK are opposite on each pair of doubled components of TT. More precisely, let μi,μi′′\mu^{\prime}_{i},\mu^{\prime\prime}_{i} be the loops corresponding to the rightmost endpoints of the double of a component TiT_{i} of TT and μi\mu_{i} be the loop associated to the rightmost endpoint of TiT_{i} itself so that μi=μiμi′′\mu_{i}=\mu^{\prime}_{i}\mu^{\prime\prime}_{i}. Since [μi]=+1=[μi′′][\mu^{\prime}_{i}]=+1=-[\mu^{\prime\prime}_{i}] in H1(XK)H_{1}(X_{K})\cong\mathbb{Z} we have ρθ(μi)=αiθ\rho\otimes\theta(\mu^{\prime}_{i})=\alpha_{i}\otimes\theta and ρθ(μi′′)=βiθ1\rho\otimes\theta(\mu^{\prime\prime}_{i})=\beta_{i}\otimes\theta^{-1}, where αi=ρ(μi),βi=ρ(μi′′)\alpha_{i}=\rho(\mu^{\prime}_{i}),\beta_{i}=\rho(\mu^{\prime\prime}_{i}), so that ρθ(μi)=ρ(μi)\rho\otimes\theta(\mu_{i})=\rho(\mu_{i}). The effect of θ\theta cancelling out is that, in the twisted version of (5) above, the invariant of TT has no powers of tt involved. Indeed, since ρ\rho takes values in GGG\subset G^{\prime} and the RR-matrix of D(H)¯|G\underline{D(H^{\prime})}|_{G} is induced from that of D(H)¯|G\underline{D(H)}|_{G} it follows that

ZD(H)¯ρ(T)=ZD(H)¯ρ(T)i=12gD(H)αiβii=12gD(H)αiβiθ0.Z_{\underline{D(H^{\prime})}}^{\rho}(T)=Z_{\underline{D(H)}}^{\rho}(T)\in\bigotimes_{i=1}^{2g}D(H)_{\alpha_{i}\beta_{i}}\subset\bigotimes_{i=1}^{2g}D(H^{\prime})_{\alpha_{i}\beta_{i}\otimes\theta^{0}}.
Refer to caption
Figure 8. TT is a tangle that is thickened to give the lower portion of the surface. TT is an arbitrary tangle with 2g2g components T1,,T2gT_{1},\dots,T_{2g} where the endpoints of T2i1,T2iT_{2i-1},T_{2i} are adjacent as shown above.

Let li,li′′π1(XKo)l^{\prime}_{i},l^{\prime\prime}_{i}\in\pi_{1}(X_{K_{o}}) be the partial longitudes associated to μi,μi′′\mu^{\prime}_{i},\mu^{\prime\prime}_{i} above. Let γi=ρ(li)\gamma_{i}=\rho(l^{\prime}_{i}) and δi=ρ(li′′)\delta_{i}=\rho(l^{\prime\prime}_{i}) so that γiαiγi1=δiβi1δi1=β11\gamma_{i}\alpha_{i}\gamma_{i}^{-1}=\delta_{i}\beta_{i}^{-1}\delta_{i}^{-1}=\beta_{1}^{-1} for each ii (note that β11\beta_{1}^{-1} is the label of the endpoint of KoK_{o}). Note that li,li′′l^{\prime}_{i},l^{\prime\prime}_{i} are zero in H1(XK)H_{1}(X_{K}) so that ρθ(li)=γi\rho\otimes\theta(l^{\prime}_{i})=\gamma_{i} and ρθ(li′′)=δi\rho\otimes\theta(l^{\prime\prime}_{i})=\delta_{i}.

Proposition 4.1.

The twisted universal invariant of (Ko,ρθ)(K_{o},\rho\otimes\theta) is computed as

ZD(H)¯ρθ(Ko)=F(ZD(H)¯ρ(T))Z_{\underline{D(H^{\prime})}}^{\rho\otimes\theta}(K_{o})=F\left(Z_{\underline{D(H)}}^{\rho}(T)\right)

where

F=m(4g1)Pg(i=12gφγiφδi)(i=12g(idD(H)αiθS~βiθ1)Δαiθ,βiθ1).F=m^{(4g-1)}\circ P^{\otimes g}\circ\left(\bigotimes_{i=1}^{2g}\varphi_{\gamma_{i}}\otimes\varphi_{\delta_{i}}\right)\left(\bigotimes_{i=1}^{2g}(\text{id}_{D(H^{\prime})_{\alpha_{i}\otimes\theta}}\otimes\widetilde{S}_{\beta_{i}\otimes\theta^{-1}})\circ\Delta_{\alpha_{i}\otimes\theta,\beta_{i}\otimes\theta^{-1}}\right).

Here m(4g1)m^{(4g-1)} denotes iterated multiplication of D(H)β11θD(H^{\prime})_{\beta_{1}^{-1}\otimes\theta} and S~\widetilde{S} is as in (3).

Proof.

This follows immediately from the lemmas of Subsection 3.6 and the above remarks. ∎

Since ZD(H)¯ρ(T)Z_{\underline{D(H)}}^{\rho}(T) has no powers of tt, the above formula implies that all powers of tt in ZD(H)¯ρθ(Ko)Z_{\underline{D(H^{\prime})}}^{\rho\otimes\theta}(K_{o}) come from the tensor FF. Note that this tensor only depends on the genus (and the αi,βi,γi,δiG\alpha_{i},\beta_{i},\gamma_{i},\delta_{i}\in G if ρ\rho is non-trivial). Thus, this already shows that the powers of tt are controlled. The next section shows how to keep track of the powers of tt coming from FF.

4.2. Filtration of D(H)¯\underline{D(H^{\prime})}

In what follows, if I=(i1,,im),J=(j1,,jm)mI=(i_{1},\dots,i_{m}),J=(j_{1},\dots,j_{m})\in\mathbb{N}^{m}, we introduce the following notation:

  • H=ImHIH=\bigoplus_{I\in\mathbb{N}^{m}}H_{I},

  • tI=t1i1tmimt^{I}=t_{1}^{i_{1}}\dots t_{m}^{i_{m}},

  • d(I)=i1++imd(I)=i_{1}+\dots+i_{m}

  • I+J=(i1+j1,,im+jm)I+J=(i_{1}+j_{1},\dots,i_{m}+j_{m}) and I=(i1,,im)-I=(-i_{1},\dots,-i_{m}),

  • IJI\leq J if ikjki_{k}\leq j_{k} for each k=1,,mk=1,\dots,m,

  • If aHa\in H is an homogeneous element, aHIa\in H_{I}, we denote |a|=I|a|=I and d(a)=d(I)d(a)=d(I). Note that |ab|=|a|+|b||ab|=|a|+|b| and d(ab)=d(a)+d(b)d(ab)=d(a)+d(b) for homogeneous a,bHa,b\in H. With this notation we have θ(a)=t|a|a\theta(a)=t^{|a|}a.

  • For every nn\in\mathbb{N} we denote Hn=d(I)=nHIH_{n}=\oplus_{d(I)=n}H_{I} so that H=nHnH=\oplus_{n\in\mathbb{N}}H_{n}.

  • Elements of Autm0(H)\mathrm{Aut}^{0}_{\mathbb{N}^{m}}(H) are denoted by α,β,γ,\alpha,\beta,\gamma,...

We define a 𝕂\mathbb{K}-linear \mathbb{N}-filtration on each D(H)αθn=(HH)[t1±1/2,,tm±1/2]D(H^{\prime})_{\alpha\otimes\theta^{n}}=(H^{*}\otimes H)[t_{1}^{\pm 1/2},\dots,t_{m}^{\pm 1/2}], where nn\in\mathbb{Z}, by

D(H)αθn[k]:=I,J0,d(I)+d(J)k(HHI)tJ.D(H^{\prime})_{\alpha\otimes\theta^{n}}[k]:=\bigoplus_{I,J\geq 0,d(I)+d(J)\leq k}(H^{*}\otimes H_{I})\cdot t^{-J}.

We will denote this simply by Dαθn[k]D^{\prime}_{\alpha\otimes\theta^{n}}[k]. In particular, an element of filtration-degree kk is a polynomial in HHH^{*}\otimes H of degree k\leq k. We consider higher tensor products i=1nDαiθ\otimes_{i=1}^{n}D^{\prime}_{\alpha_{i}\otimes\theta} with the tensor product filtration.

Lemma 4.2.

The multiplication of D(H)αθD(H^{\prime})_{\alpha\otimes\theta} is filtration-preserving.

Proof.

Let xDαθ[n],yDαθ[m]x\in D^{\prime}_{\alpha\otimes\theta}[n],y\in D^{\prime}_{\alpha\otimes\theta}[m], say x=patJx=p\otimes a\cdot t^{-J} and y=qbtLy=q\otimes b\cdot t^{-L}, where a,bH,p,qHa,b\in H,\ p,q\in H^{*} and d(a)+d(J)n,d(b)+d(L)md(a)+d(J)\leq n,d(b)+d(L)\leq m. Then

xαθy\displaystyle x\cdot_{\alpha\otimes\theta}y =tJLa(1),q(3)S1((αθ)1(a(3))),q(1)pq(2)a(2)b\displaystyle=t^{-J-L}\langle a_{(1)},q_{(3)}\rangle\langle S^{-1}((\alpha\otimes\theta)^{-1}(a_{(3)})),q_{(1)}\rangle p\cdot q_{(2)}\otimes a_{(2)}\cdot b
=tJL|a(3)|a(1),q(3)S1(α1(a(3))),q(1)pq(2)a(2)b\displaystyle=t^{-J-L-|a_{(3)}|}\langle a_{(1)},q_{(3)}\rangle\langle S^{-1}(\alpha^{-1}(a_{(3)})),q_{(1)}\rangle p\cdot q_{(2)}\otimes a_{(2)}\cdot b
(HH|a(2)b|)tJL|a(3)|.\displaystyle\in\bigoplus(H^{*}\otimes H_{|a_{(2)}b|})t^{-J-L-|a_{(3)}|}.

Since HH is m\mathbb{N}^{m}-graded, we have d(a(2))+d(a(3))d(a)d(a_{(2)})+d(a_{(3)})\leq d(a). Thus d(a(2)b)+d(J+L+a(3))d(a)+d(b)+d(J)+d(L)n+md(a_{(2)}b)+d(J+L+a_{(3)})\leq d(a)+d(b)+d(J)+d(L)\leq n+m, showing that the above direct sum is contained in Dαθ[n+m]D^{\prime}_{\alpha\otimes\theta}[n+m] as desired. ∎

In contrast to the above lemma, the maps Δαθ,βθ1,Sβθ1\Delta_{\alpha\otimes\theta,\beta\otimes\theta^{-1}},S_{\beta\otimes\theta^{-1}} are not filtration-preserving. However, we have the following:

Lemma 4.3.

The map (idD(H)αθSβθ1)Δαθ,βθ1:D(H)αβD(H)αθD(H)β1θ(\text{id}_{D(H^{\prime})_{\alpha\otimes\theta}}\otimes S_{\beta\otimes\theta^{-1}})\circ\Delta_{\alpha\otimes\theta,\beta\otimes\theta^{-1}}:D(H^{\prime})_{\alpha\beta}\to D(H^{\prime})_{\alpha\otimes\theta}\otimes D(H^{\prime})_{\beta^{-1}\otimes\theta} is filtration-preserving.

Proof.

Let xDαβ[n]x\in D^{\prime}_{\alpha\beta}[n], say, x=patJx=p\otimes a\cdot t^{-J} with d(a)+d(J)nd(a)+d(J)\leq n. We have

(idD(H)αθSβθ1)\displaystyle(\text{id}_{D(H^{\prime})_{\alpha\otimes\theta}}\otimes S_{\beta\otimes\theta^{-1}}) Δαθ,βθ1(x)\displaystyle\Delta_{\alpha\otimes\theta,\beta\otimes\theta^{-1}}(x)
=tJ(idD(H)αθSβθ1)(p(2)a(1)p(1)t|a(2)|α1(a(2)))\displaystyle=t^{-J}(\text{id}_{D(H)_{\alpha\otimes\theta}}\otimes S_{\beta\otimes\theta^{-1}})(p_{(2)}\otimes a_{(1)}\otimes p_{(1)}\otimes t^{-|a_{(2)}|}\alpha^{-1}(a_{(2)}))
=tJp(2)a(1)(t|a(2)|t|α1(a(2))|β1Sα1(a(2))β1θp(1)S1)\displaystyle=t^{-J}p_{(2)}\otimes a_{(1)}\otimes(t^{-|a_{(2)}|}t^{|\alpha^{-1}(a_{(2)})|}\beta^{-1}S\alpha^{-1}(a_{(2)})\cdot_{\beta^{-1}\otimes\theta}p_{(1)}\circ S^{-1})
=(tJp(2)a(1))(β1Sα1(a(2))β1θp(1)S1).\displaystyle=(t^{-J}p_{(2)}\otimes a_{(1)})\otimes(\beta^{-1}S\alpha^{-1}(a_{(2)})\cdot_{\beta^{-1}\otimes\theta}p_{(1)}\circ S^{-1}).

Note that in the third equality we used that αAut(H)\alpha\in\mathrm{Aut}(H) preserves the m\mathbb{N}^{m}-degree. By the previous lemma, the rightmost tensor factor of the last equality belongs to Dβ1θ[d(a(2))]D^{\prime}_{\beta^{-1}\otimes\theta}[d(a_{(2)})]. The left tensor factor clearly belongs to Dαθ[d(a(1))+d(J)]D^{\prime}_{\alpha\otimes\theta}[d(a_{(1)})+d(J)] so their tensor product belongs to (DαθDβ1θ)[d(a(1))+d(a(2))+d(J)](D^{\prime}_{\alpha\otimes\theta}\otimes D^{\prime}_{\beta^{-1}\otimes\theta})[d(a_{(1)})+d(a_{(2)})+d(J)] which is contained in the nn-th term of the same filtration since

d(a(1))+d(a(2))+d(J)d(a)+d(J)n.d(a_{(1)})+d(a_{(2)})+d(J)\leq d(a)+d(J)\leq n.

4.3. Proof of Theorem 1

Let N=d(H)N=d(H). We will prove something more general, namely, that the entire universal invariant ZD(H)¯ρθ(Ko)D(H)Z_{\underline{D(H^{\prime})}}^{\rho\otimes\theta}(K_{o})\in D(H^{\prime}) is a polynomial of degree 2gN\leq 2gN, after identifying D(H)D(H^{\prime}) with (HH)[t1±1/2,,tm±1/2](H^{*}\otimes H)[t_{1}^{\pm 1/2},\dots,t_{m}^{\pm 1/2}]. Thus, any evaluation, in particular PHρθP_{H}^{\rho\otimes\theta}, is a polynomial of degree 2gN\leq 2gN. To begin with, note that since there are no powers of tt involved in ZD(H)¯ρ(T)Z_{\underline{D(H)}}^{\rho}(T) and NN is the top \mathbb{N}-degree of HH, we have

ZD(H)¯ρ(T)i=12g(Dαiβi[N])(i=12gDαiβi)[2gN].Z_{\underline{D(H)}}^{\rho}(T)\in\bigotimes_{i=1}^{2g}(D^{\prime}_{\alpha_{i}\beta_{i}}[N])\subset\left(\bigotimes_{i=1}^{2g}D^{\prime}_{\alpha_{i}\beta_{i}}\right)[2gN].

Recall that S~βθ1(x)=Sβθ1(x)𝒈β1θ1\widetilde{S}_{\beta\otimes\theta^{-1}}(x)=S_{\beta\otimes\theta^{-1}}(x)\boldsymbol{g}_{\beta^{-1}\otimes\theta}^{-1} and that 𝒈β1θ=t|Λl|/2𝜷𝒃\boldsymbol{g}_{\beta^{-1}\otimes\theta}=t^{-|\Lambda_{l}|/2}\boldsymbol{\beta}\otimes\boldsymbol{b}. Clearly, multiplication by 𝜷𝒃\boldsymbol{\beta}\otimes\boldsymbol{b} preserves the filtration. Moreover, φα\varphi_{\alpha} is filtration-preserving for any αG\alpha\in G. Together with the previous lemmas, it follows that the map

tg|Λl|m(4g1)Pg(i=12gφγiφδi)(i=12g(idDαiθS~βiθ1)Δαiθ,βiθ1)t^{-g|\Lambda_{l}|}\cdot m^{(4g-1)}\circ P^{\otimes g}\circ\left(\bigotimes_{i=1}^{2g}\varphi_{\gamma_{i}}\otimes\varphi_{\delta_{i}}\right)\left(\bigotimes_{i=1}^{2g}(\text{id}_{D^{\prime}_{\alpha_{i}\otimes\theta}}\otimes\widetilde{S}_{\beta_{i}\otimes\theta^{-1}})\circ\Delta_{\alpha_{i}\otimes\theta,\beta_{i}\otimes\theta^{-1}}\right)

is filtration-preserving (the tg|Λl|t^{-g|\Lambda_{l}|} factor is simply to eliminate the powers of tt from the GG^{\prime}-pivots in S~\widetilde{S}). By Proposition 4.1, it follows that tg|Λl|ZD(H)¯ρθ(Ko)Dβ11θ[2gN]t^{-g|\Lambda_{l}|}Z_{\underline{D(H^{\prime})}}^{\rho\otimes\theta}(K_{o})\in D^{\prime}_{\beta_{1}^{-1}\otimes\theta}[2gN] and by definition of our filtration, this implies that ZD(H)¯ρθ(Ko)Z_{\underline{D(H^{\prime})}}^{\rho\otimes\theta}(K_{o}) is a polynomial in HHH^{*}\otimes H of degree 2gN\leq 2gN. This proves our theorem. ∎

Remark 4.4.

The construction of ZD(H)¯ρθ(T)Z_{\underline{D(H^{\prime})}}^{\rho\otimes\theta}(T) only requires HH to be m\mathbb{Z}^{m}-graded (instead of m\mathbb{N}^{m}-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 m\mathbb{N}^{m}-graded case since the most relevant examples, Borel parts of quantum groups, are m\mathbb{N}^{m}-graded.

5. Relation to ADO invariants

Let HH be the algebra generated by k,Ek,E such that kE=ζpEkkE=\zeta_{p}Ek (Hopf algebra structure defined below). If ζp\zeta_{p} is a primitive 2p2p-th root of unity (with p2p\geq 2), which we simply take as ζp=eπi/p\zeta_{p}=e^{\pi i/p}, we let HpH_{p} be the quotient of HH by k4p=1,Ep=0k^{4p}=1,E^{p}=0. This is \mathbb{N}-graded with |E|=1,|k|=0.|E|=1,|k|=0. During this whole section we denote ζpα=eπiα/p\zeta_{p}^{\alpha}=e^{\pi i\alpha/p} and [α]=ζpαζpαζpζp1[\alpha]=\frac{\zeta_{p}^{\alpha}-\zeta_{p}^{-\alpha}}{\zeta_{p}-\zeta_{p}^{-1}} for any α\alpha\in\mathbb{C}.

5.1. Quantum 𝔰𝔩2\mathfrak{sl}_{2}

Consider the algebra D(H)D(H) with generators E,F,k±1,κ±1E,F,k^{\pm 1},\kappa^{\pm 1} satisfying

kE\displaystyle kE =ζpEk,\displaystyle=\zeta_{p}Ek, κE\displaystyle\kappa E =ζp1Eκ,\displaystyle=\zeta_{p}^{-1}E\kappa, kF\displaystyle kF =ζp1Fk,\displaystyle=\zeta_{p}^{-1}Fk, κF\displaystyle\kappa F =ζpFκ,\displaystyle=\zeta_{p}F\kappa,
[E,F]\displaystyle[E,F] =k2κ2ζpζp1,\displaystyle=\frac{k^{2}-\kappa^{2}}{\zeta_{p}-\zeta_{p}^{-1}}, kκ\displaystyle k\kappa =κk,\displaystyle=\kappa k, kk1\displaystyle kk^{-1} =1=κκ1,\displaystyle=1=\kappa\kappa^{-1},

and Hopf algebra structure determined by

Δ(E)\displaystyle\Delta(E) =Ek2+1E,\displaystyle=E\otimes k^{2}+1\otimes E, Δ(F)\displaystyle\Delta(F) =κ2F+F1,\displaystyle=\kappa^{2}\otimes F+F\otimes 1, Δ(k±1)\displaystyle\Delta(k^{\pm 1}) =k±1k±1,\displaystyle=k^{\pm 1}\otimes k^{\pm 1}, Δ(κ±1)\displaystyle\Delta(\kappa^{\pm 1}) =κ±1κ±1,\displaystyle=\kappa^{\pm 1}\otimes\kappa^{\pm 1},
ϵ(E)\displaystyle\epsilon(E) =ϵ(F)=0,\displaystyle=\epsilon(F)=0, ϵ(k)\displaystyle\epsilon(k) =ϵ(κ)=1,\displaystyle=\epsilon(\kappa)=1, S(E)\displaystyle S(E) =Ek2,\displaystyle=-Ek^{-2}, S(F)\displaystyle S(F) =κ2F,\displaystyle=-\kappa^{-2}F,
S(k)\displaystyle S(k) =k1,\displaystyle=k^{-1}, S(κ)\displaystyle S(\kappa) =κ1.\displaystyle=\kappa^{-1}.

Since kκk\kappa is a central group-like in D(H)D(H), the quotient D(H)/ID(H)/I, where II is the ideal generated by kκ1k\kappa-1, is a Hopf algebra. The usual quantum group Uζp(𝔰𝔩2)U_{\zeta_{p}}(\mathfrak{sl}_{2}) (at a root of unity) is the subalgebra generated by E,F,K±1:=k±2E,F,K^{\pm 1}:=k^{\pm 2} in this quotient.

5.2. The Drinfeld double of HpH_{p}

The Drinfeld double D(Hp)D(H_{p}) is the quotient of D(H)D(H) by the relations

Ep\displaystyle E^{p} =Fp=0,\displaystyle=F^{p}=0, k4p\displaystyle k^{4p} =κ4p=1,\displaystyle=\kappa^{4p}=1,

see [9]. For convenience, we will denote by e,f,k¯,κ¯e,f,\overline{k},\overline{\kappa} the images of E,F,k,κE,F,k,\kappa in D(Hp)D(H_{p}). Let D(p)D(p) be the quotient of D(Hp)D(H_{p}) by the ideal generated by k¯κ¯1\overline{k}\overline{\kappa}-1. We will also denote by e,f,k¯e,f,\overline{k} the image in D(p)D(p) of the generators of D(Hp)D(H_{p}).

The Hopf algebra D(p)D(p) is quasi-triangular with RR-matrix given by

R=14pi,j=04p1ζpij/2k¯ik¯jR=\frac{1}{4p}\sum_{i,j=0}^{4p-1}\zeta_{p}^{-ij/2}\overline{k}^{i}\otimes\overline{k}^{j}\cdot\mathcal{R}

where =m=0p1cmemfm\mathcal{R}=\sum_{m=0}^{p-1}c_{m}e^{m}\otimes f^{m} and cm=(ζpζp1)m[m]!ζpm(m1)/2c_{m}=\frac{(\zeta_{p}-\zeta_{p}^{-1})^{m}}{[m]!}\zeta_{p}^{m(m-1)/2} for each m0m\geq 0. Note that this RR-matrix is simply the projection in D(p)D(p) of the usual RR-matrix of the double D(Hp)D(H_{p}). Moreover, D(p)D(p) has a ribbon element determined by 𝒈=vu\boldsymbol{g}=vu where 𝒈=k2p+2\boldsymbol{g}=k^{2p+2} and uu is the Drinfeld element (for an explicit expression of v1v^{-1} see [9]). Note that if 𝒃:=kp+1\boldsymbol{b}:=k^{p+1} and 𝜷Hp\boldsymbol{\beta}\in H_{p}^{*} is defined by 𝜷(k)=ζpp+12,𝜷(E)=0\boldsymbol{\beta}(k)=\zeta_{p}^{\frac{p+1}{2}},\boldsymbol{\beta}(E)=0 then (𝜷,𝒃)(\boldsymbol{\beta},\boldsymbol{b}) satisfies the hypothesis of the Kauffman-Radford theorem (see 2.4). The pivot 𝒈=k2p+2\boldsymbol{g}=k^{2p+2} is the projection in D(p)D(p) of the pivot 𝜷𝒃D(Hp)\boldsymbol{\beta}\otimes\boldsymbol{b}\in D(H_{p}).

Note that D(p)D(p) is the usual braided extension of the restricted quantum group U¯q(𝔰𝔩2)\overline{U}_{q}(\mathfrak{sl}_{2}) at q=ζpq=\zeta_{p} (U¯q(𝔰𝔩2)\overline{U}_{q}(\mathfrak{sl}_{2}) is the Hopf subalgebra of D(p)D(p) generated by e,f,k¯2e,f,\overline{k}^{2} and is not braided)

5.3. The twisted Drinfeld double of HpH_{p}

Consider the action of (,+)(\mathbb{C},+) on HpH_{p} given by αϕαAut(Hp)\alpha\mapsto\phi_{\alpha}\in\mathrm{Aut}(H_{p}) defined by ϕα(k)=k,ϕα(e)=ζp2αe\phi_{\alpha}(k)=k,\phi_{\alpha}(e)=\zeta_{p}^{2\alpha}e, that is, take t=ζp2αt=\zeta_{p}^{2\alpha} in the degree twist action (here ζpα:=eiπα/p\zeta_{p}^{\alpha}:=e^{i\pi\alpha/p} for any α\alpha\in\mathbb{C}). For each α\alpha, let DαD_{\alpha} be the quotient of D(Hp)ϕαD(H_{p})_{\phi_{\alpha}} by the ideal generated by kκ1k\kappa-1, in particular, D0D_{0} is the Hopf algebra D(p)D(p) defined previously. Then DαD_{\alpha} has generators e,f,k¯e,f,\overline{k} subject to the same relations as in D0D_{0} except for the relation [e,f]=k¯2k¯2ζpζp1[e,f]=\frac{\overline{k}^{2}-\overline{k}^{-2}}{\zeta_{p}-\zeta_{p}^{-1}} which now becomes

(6) [e,f]=k¯2ζp2αk¯2ζpζp1.[e,f]=\frac{\overline{k}^{2}-\zeta_{p}^{-2\alpha}\overline{k}^{-2}}{\zeta_{p}-\zeta_{p}^{-1}}.

The coproduct Δα,β:Dα+βDαDβ\Delta_{\alpha,\beta}:D_{\alpha+\beta}\to D_{\alpha}\otimes D_{\beta} is given by

(7) Δα,β(e)=ζp2α1e+ek¯2,Δα,β(f)=k¯2f+f1.\Delta_{\alpha,\beta}(e)=\zeta_{p}^{-2\alpha}1\otimes e+e\otimes\overline{k}^{2},\hskip 28.45274pt\Delta_{\alpha,\beta}(f)=\overline{k}^{-2}\otimes f+f\otimes 1.

These computations are carried out in [34] (in a slightly different convention, obtained by applying idHα\text{id}_{H^{*}}\otimes\alpha on each DαD_{\alpha} to our formulas). The RR-matrix is

Rα,β=14pi,j=04p1ζpij/2k¯ik¯jm=0p1cmζp2αmemfmR_{\alpha,\beta}=\frac{1}{4p}\sum_{i,j=0}^{4p-1}\zeta_{p}^{-ij/2}\overline{k}^{i}\otimes\overline{k}^{j}\cdot\sum_{m=0}^{p-1}c_{m}\zeta_{p}^{2\alpha m}e^{m}\otimes f^{m}

where cmc_{m} is as before. This is simply the projection in DαDβD_{\alpha}\otimes D_{\beta} of the RR-matrix of D(Hp)¯\underline{D(H_{p})} defined in Subsection 2.3. The (left) cointegral of HpH_{p} is Λl=i=04p1kiep1\Lambda_{l}=\sum_{i=0}^{4p-1}k^{i}e^{p-1} and ϕα(Λl)=ζp2α(p1)Λl\phi_{\alpha}(\Lambda_{l})=\zeta_{p}^{2\alpha(p-1)}\Lambda_{l} so that rHp(ϕα)=ζp2α(p1)r_{H_{p}}(\phi_{\alpha})=\zeta_{p}^{2\alpha(p-1)}. This has an obvious square root rHp(ϕα)=ζpα(p1)\sqrt{r_{H_{p}}(\phi_{\alpha})}=\zeta_{p}^{\alpha(p-1)}, hence the Hopf group-coalgebra {Dα}α\{D_{\alpha}\}_{\alpha\in\mathbb{C}} is \mathbb{C}-ribbon with ribbon element vαv_{\alpha} which is the image in DαD_{\alpha} of

ζpα(p1)(idHα)(v)\displaystyle\zeta_{p}^{-\alpha(p-1)}(\text{id}_{H^{*}}\otimes\alpha)(v)

where vv is the ribbon element of D(Hp)D(H_{p}) and the GG-pivot is

𝒈α=ζpα(p1)k¯2p+2.\boldsymbol{g}_{\alpha}=\zeta_{p}^{-\alpha(p-1)}\overline{k}^{2p+2}.

5.4. Relation to unrolled quantum 𝔰𝔩2\mathfrak{sl}_{2}

To relate the above twisted Drinfeld double to the usual quantum group Uq(𝔰𝔩2)U_{q}(\mathfrak{sl}_{2}) we multiply each eDαe\in D_{\alpha} by ζpα\zeta_{p}^{\alpha} and k¯Dα\overline{k}\in D_{\alpha} by ζpα2\zeta_{p}^{\frac{\alpha}{2}}. Let E=ζpαe,F=f,k=ζpα2k¯E=\zeta_{p}^{\alpha}e,F=f,k=\zeta_{p}^{\frac{\alpha}{2}}\overline{k} (note that there is one E,kE,k on each DαD_{\alpha} but we do not include α\alpha in the notation). In these new generators, relation (6) becomes

[E,F]=k2k2ζpζp1[E,F]=\frac{k^{2}-k^{-2}}{\zeta_{p}-\zeta_{p}^{-1}}

and k4p=ζp2pαk^{4p}=\zeta_{p}^{2p\alpha} while the other relations among the generators remain unchanged. The coproduct (7) becomes

Δα,β(E)=Ek2+ζpα1E,Δα,β(F)=ζpαk2F+F1\Delta_{\alpha,\beta}(E)=E\otimes k^{2}+\zeta_{p}^{-\alpha}1\otimes E,\hskip 28.45274pt\Delta_{\alpha,\beta}(F)=\zeta_{p}^{\alpha}k^{-2}\otimes F+F\otimes 1

and the RR-matrix

Rα,β\displaystyle R_{\alpha,\beta} =14pi,j=04p1ζpij/2k¯ik¯jm=0p1cmζp2αmemfm\displaystyle=\frac{1}{4p}\sum_{i,j=0}^{4p-1}\zeta_{p}^{-ij/2}\overline{k}^{i}\otimes\overline{k}^{j}\sum_{m=0}^{p-1}c_{m}\zeta_{p}^{2\alpha m}e^{m}\otimes f^{m}
=14pi,j=04p1ζpij/2αi/2βj/2kikjm=0p1cmζpαmEmFm\displaystyle=\frac{1}{4p}\sum_{i,j=0}^{4p-1}\zeta_{p}^{-ij/2-\alpha i/2-\beta j/2}k^{i}\otimes k^{j}\sum_{m=0}^{p-1}c_{m}\zeta_{p}^{\alpha m}E^{m}\otimes F^{m}
=ζpαβ/24pi,j=04p1ζp12(α+j)(β+i)kikjm=0p1cmζpαmEmFm\displaystyle=\frac{\zeta_{p}^{\alpha\beta/2}}{4p}\sum_{i,j=0}^{4p-1}\zeta_{p}^{-\frac{1}{2}(\alpha+j)(\beta+i)}k^{i}\otimes k^{j}\sum_{m=0}^{p-1}c_{m}\zeta_{p}^{\alpha m}E^{m}\otimes F^{m}

Now let Aα=[(,+)]DαA_{\alpha}=\mathbb{C}[(\mathbb{C},+)]\ltimes D_{\alpha} be the Hopf group-coalgebra defined in Subsection 2.5. Clearly φ1/2k1\varphi_{1/2}k^{-1} is central in AαA_{\alpha}, hence we mod out by φ1/2=k\varphi_{1/2}=k and we keep denoting the quotient by AαA_{\alpha}. For each β\beta\in\mathbb{C}, we denote by kβk^{\beta} the element of AαA_{\alpha} corresponding to φβ/2\varphi_{\beta/2}. Thus, AαA_{\alpha} is obtained by adding generators kβ,βk^{\beta},\beta\in\mathbb{C} to DαD_{\alpha} satisfying

kβE\displaystyle k^{\beta}E =ζpβEkβ,\displaystyle=\zeta_{p}^{\beta}Ek^{\beta}, kβF\displaystyle k^{\beta}F =ζpβFkβ,\displaystyle=\zeta_{p}^{-\beta}Fk^{\beta}, kβ+γ\displaystyle k^{\beta+\gamma} =kβkγ,\displaystyle=k^{\beta}k^{\gamma},
Δα,βA(kγ)\displaystyle\Delta^{A}_{\alpha,\beta}(k^{\gamma}) =kγkγ,\displaystyle=k^{\gamma}\otimes k^{\gamma}, k0\displaystyle k^{0} =1α,\displaystyle=1_{\alpha}, k1\displaystyle k^{1} =k.\displaystyle=k.

Here E,F,kE,F,k denote the above (rescaled) generators of DαD_{\alpha}. Then A¯={Aα}α\underline{A}=\{A_{\alpha}\}_{\alpha\in\mathbb{C}} is a quasi-triangular Hopf group-coalgebra with trivial crossing and has an RR-matrix given by

Rα,βA=(1k2α)Rα,βR^{A}_{\alpha,\beta}=(1\otimes k^{2\alpha})R_{\alpha,\beta}

where Rα,βR_{\alpha,\beta} is the above RR-matrix of {Dα}\{D_{\alpha}\}. The ribbon element becomes vαA=vαk2αv^{A}_{\alpha}=v_{\alpha}\cdot k^{2\alpha} and the pivot element is the same 𝒈α\boldsymbol{g}_{\alpha} as before, see Subsection 2.5.

As an algebra AαA_{\alpha} is isomorphic to (a ribbon extension of) [(,+)]Uα\mathbb{C}[(\mathbb{C},+)]\ltimes U_{\alpha}, where UαU_{\alpha} is the quotient of Uζp(𝔰𝔩2)U_{\zeta_{p}}(\mathfrak{sl}_{2}) by Ep=Fp=0,K2p=ζp2pαE^{p}=F^{p}=0,K^{2p}=\zeta_{p}^{2p\alpha}. However, the coproduct and the RR-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 A¯={Aα}α\underline{A}=\{A_{\alpha}\}_{\alpha\in\mathbb{C}} is a collection J={Jα,β}α,βJ=\{J_{\alpha,\beta}\}_{\alpha,\beta\in\mathbb{C}} where each Jα,βJ_{\alpha,\beta} is an invertible element of AαAβA_{\alpha}\otimes A_{\beta} satisfying

  1. (1)

    1αJβ,γ(idAαΔβ,γA(Jα,β+γ))=Jα,β1γ(Δα,βAidAγ(Jα+β,γ))1_{\alpha}\otimes J_{\beta,\gamma}\cdot(\text{id}_{A_{\alpha}}\otimes\Delta^{A}_{\beta,\gamma}(J_{\alpha,\beta+\gamma}))=J_{\alpha,\beta}\otimes 1_{\gamma}\cdot(\Delta^{A}_{\alpha,\beta}\otimes\text{id}_{A_{\gamma}}(J_{\alpha+\beta,\gamma})),

  2. (2)

    ϵ0idAα(J0,α)=1α=idAαϵ0(Jα,0)\epsilon_{0}\otimes\text{id}_{A_{\alpha}}(J_{0,\alpha})=1_{\alpha}=\text{id}_{A_{\alpha}}\otimes\epsilon_{0}(J_{\alpha,0}).

for each α,β,γ\alpha,\beta,\gamma\in\mathbb{C}. Drinfeld twisting a quasi-triangular Hopf group-coalgebra produces a new quasi-triangular Hopf group-coalgebra A¯J={AαJ}\underline{A}^{J}=\{A_{\alpha}^{J}\} where AαJ=AαA_{\alpha}^{J}=A_{\alpha} as an algebra for each α\alpha, the coproduct is Δα,βJ=Jα,βΔα,βAJα,β1\Delta_{\alpha,\beta}^{J}=J_{\alpha,\beta}\Delta^{A}_{\alpha,\beta}J_{\alpha,\beta}^{-1} and the RR-matrix is

Rα,βJ=τβ,α(Jβ,α)Rα,βAJα,β1R_{\alpha,\beta}^{J}=\tau_{\beta,\alpha}(J_{\beta,\alpha})\cdot R^{A}_{\alpha,\beta}\cdot J_{\alpha,\beta}^{-1}

where τβ,α:AβAαAαAβ\tau_{\beta,\alpha}:A_{\beta}\otimes A_{\alpha}\to A_{\alpha}\otimes A_{\beta} is the usual symmetry (with signs in the super case).

Lemma 5.1.

Let Jα,β=1kαAαAβJ_{\alpha,\beta}=1\otimes k^{\alpha}\in A_{\alpha}\otimes A_{\beta} for each α,β\alpha,\beta\in\mathbb{C}. Then J={Jα,β}α,βJ=\{J_{\alpha,\beta}\}_{\alpha,\beta\in\mathbb{C}} is a (graded) Drinfeld twist on A¯\underline{A}.

Proof.

It is clear that JJ is a Drinfeld twist because the kαk^{\alpha}’s are group-like. ∎

Thus, we obtain a new ribbon Hopf group-coalgebra A¯J\underline{A}^{J} whose coproduct is determined by

Δα,βJ(E)=Jα,βΔα,βA(E)Jα,β1\displaystyle\Delta_{\alpha,\beta}^{J}(E)=J_{\alpha,\beta}\Delta^{A}_{\alpha,\beta}(E)J_{\alpha,\beta}^{-1} =Ek2+ζpα1kαEkα\displaystyle=E\otimes k^{2}+\zeta_{p}^{-\alpha}1\otimes k^{\alpha}Ek^{-\alpha}
=Ek2+1E\displaystyle=E\otimes k^{2}+1\otimes E

and

Δα,βJ(F)=Jα,βΔα,βA(F)Jα,β1\displaystyle\Delta_{\alpha,\beta}^{J}(F)=J_{\alpha,\beta}\Delta^{A}_{\alpha,\beta}(F)J_{\alpha,\beta}^{-1} =ζpαk2kαFkα+F1\displaystyle=\zeta_{p}^{\alpha}k^{-2}\otimes k^{\alpha}Fk^{-\alpha}+F\otimes 1
=k2F+F1.\displaystyle=k^{-2}\otimes F+F\otimes 1.

The antipode of A¯J\underline{A}^{J} is SαJ(x)=kαSαA(x)kαS^{J}_{\alpha}(x)=k^{\alpha}S^{A}_{\alpha}(x)k^{-\alpha}. The RR-matrix of A¯J\underline{A}^{J} is

Rα,βJ\displaystyle R_{\alpha,\beta}^{J} =(kβ1)Rα,βA(1kα)\displaystyle=(k^{\beta}\otimes 1)R^{A}_{\alpha,\beta}(1\otimes k^{-\alpha})
=(kβ1)ζpαβ/24pi,j=04p1ζp12(α+j)(β+i)kikj+2αm=0p1cmζpαmEmFm(1kα)\displaystyle=(k^{\beta}\otimes 1)\frac{\zeta_{p}^{\alpha\beta/2}}{4p}\sum_{i,j=0}^{4p-1}\zeta_{p}^{-\frac{1}{2}(\alpha+j)(\beta+i)}k^{i}\otimes k^{j+2\alpha}\sum_{m=0}^{p-1}c_{m}\zeta_{p}^{\alpha m}E^{m}\otimes F^{m}(1\otimes k^{-\alpha})
=ζpαβ/24pi,j=04p1ζp12(α+j)(β+i)ki+βkj+αm=0p1cmEmFm\displaystyle=\frac{\zeta_{p}^{\alpha\beta/2}}{4p}\sum_{i,j=0}^{4p-1}\zeta_{p}^{-\frac{1}{2}(\alpha+j)(\beta+i)}k^{i+\beta}\otimes k^{j+\alpha}\sum_{m=0}^{p-1}c_{m}E^{m}\otimes F^{m}
=α,βJ\displaystyle=\mathcal{H}_{\alpha,\beta}\cdot\mathcal{R}^{J}

where

α,β=ζpαβ/24pi,j=04p1ζp12(α+j)(β+i)ki+βkj+α\mathcal{H}_{\alpha,\beta}=\frac{\zeta_{p}^{\alpha\beta/2}}{4p}\sum_{i,j=0}^{4p-1}\zeta_{p}^{-\frac{1}{2}(\alpha+j)(\beta+i)}k^{i+\beta}\otimes k^{j+\alpha}

and J=m=0p1cmEmFm.\mathcal{R}^{J}=\sum_{m=0}^{p-1}c_{m}E^{m}\otimes F^{m}.

Lemma 5.2.

The Drinfeld element, ribbon element and pivot of A¯J\underline{A}^{J} are the same of A¯\underline{A}.

Proof.

We only check this for the Drinfeld element. Write Rα,βA=sαtβR^{A}_{\alpha,\beta}=\sum s^{\prime}_{\alpha}\otimes t^{\prime}_{\beta} so that Rα,βJ=kβsαtβkαR^{J}_{\alpha,\beta}=\sum k^{\beta}s^{\prime}_{\alpha}\otimes t^{\prime}_{\beta}k^{-\alpha}. Then

uαJ\displaystyle u_{\alpha}^{J} =SαJ(tαkα)kαsα\displaystyle=\sum S^{J}_{-\alpha}(t^{\prime}_{-\alpha}k^{-\alpha})k^{-\alpha}s^{\prime}_{\alpha}
=kαSαJ(tα)kαsα\displaystyle=\sum k^{\alpha}S^{J}_{-\alpha}(t^{\prime}_{\alpha})k^{-\alpha}s^{\prime}_{\alpha}
=Sα(tα)sα=uα\displaystyle=\sum S_{-\alpha}(t^{\prime}_{\alpha})s^{\prime}_{\alpha}=u_{\alpha}

where we used the above expression for SJS^{J} in the third equality. ∎

The Hopf group-coalgebra {AαJ}\{A_{\alpha}^{J}\} is isomorphic, as a Hopf group-coalgebra, to (a ribbon extension of) {[(,+)]Uα}α\{\mathbb{C}[(\mathbb{C},+)]\ltimes U_{\alpha}\}_{\alpha\in\mathbb{C}}. 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 A¯\underline{A}-valued and the A¯J\underline{A}^{J}-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 σiJ:AαnAαn\sigma^{J}_{i}:A_{\alpha}^{\otimes n}\to A_{\alpha}^{\otimes n} be the braid morphism corresponding to the ii-th generator of the braid group BnB_{n} coming from the RR-matrix Rα,αJR_{\alpha,\alpha}^{J}. Let σi\sigma_{i} be the one for Rα,αAR^{A}_{\alpha,\alpha}. Then if J=1kαk2αkα(n1)J=1\otimes k^{\alpha}\otimes k^{2\alpha}\otimes\dots\otimes k^{\alpha(n-1)} it is easy to see that J1σiJJ=σiJ^{-1}\sigma_{i}^{J}J=\sigma_{i} for all ii, hence J1σ(β)JJ=σ(β)J^{-1}\sigma(\beta)^{J}J=\sigma(\beta) for any braid βBn\beta\in B_{n} where σ(β)J,σ(β)\sigma(\beta)^{J},\sigma(\beta) are the braid operators associated to A¯J,A¯\underline{A}^{J},\underline{A} respectively. Closing the ii-th strand in J1σ(β)JJJ^{-1}\sigma(\beta)^{J}J has the effect of multiplying k(i1)αk^{(i-1)\alpha} with k(i1)αk^{-(i-1)\alpha}, so closing the last n1n-1 strands is invariant under conjugation by JJ. Since A¯J\underline{A}^{J} and A¯\underline{A} have the same pivot, it follows that the invariants are equal.

5.5. Verma modules

For each α\alpha\in\mathbb{C}, the Verma module VαV_{\alpha} is the Uζp(𝔰𝔩2)U_{\zeta_{p}}(\mathfrak{sl}_{2})-module with basis v0,,vp1v_{0},\dots,v_{p-1} and action given by

Evi\displaystyle Ev_{i} =[αi+1]vi1,\displaystyle=[\alpha-i+1]v_{i-1}, Fvi\displaystyle Fv_{i} =[i+1]vi+1,\displaystyle=[i+1]v_{i+1}, Kvi\displaystyle Kv_{i} =ζpα2ivi,\displaystyle=\zeta_{p}^{\alpha-2i}v_{i},

see [21]. This becomes a D(H)/(kκ1)D(H)/(k\kappa-1)-module if we set kvi=ζpα2ivikv_{i}=\zeta_{p}^{\frac{\alpha}{2}-i}v_{i}. Since k4pk^{4p} acts by ζp2pα\zeta_{p}^{2p\alpha} this is a module over DαD_{\alpha}. It is a module over Aα=[(,+)]DαA_{\alpha}=\mathbb{C}[(\mathbb{C},+)]\ltimes D_{\alpha} if we set kβvi=ζpβ(α/2i)vik^{\beta}v_{i}=\zeta_{p}^{\beta(\alpha/2-i)}v_{i}. The operator α,βAαAβ\mathcal{H}_{\alpha,\beta}\in A_{\alpha}\otimes A_{\beta} above acts over a tensor product VαVβV_{\alpha}\otimes V_{\beta} as follows

α,β(vavb)\displaystyle\mathcal{H}_{\alpha,\beta}(v_{a}\otimes v_{b}) =ζpαβ/24pi,j=04p1ζp12(α+j)(β+i)ζp(i+β)(α/2a)ζp(j+α)(β/2b)(vavb)\displaystyle=\frac{\zeta_{p}^{\alpha\beta/2}}{4p}\sum_{i,j=0}^{4p-1}\zeta_{p}^{-\frac{1}{2}(\alpha+j)(\beta+i)}\zeta_{p}^{(i+\beta)(\alpha/2-a)}\zeta_{p}^{(j+\alpha)(\beta/2-b)}(v_{a}\otimes v_{b})
=ζpαβ4pi,j=04p1ζpaβaibαbjij/2(vavb)\displaystyle=\frac{\zeta_{p}^{\alpha\beta}}{4p}\sum_{i,j=0}^{4p-1}\zeta_{p}^{-a\beta-ai-b\alpha-bj-ij/2}(v_{a}\otimes v_{b})
=ζpαβ4pi=04p1ζpaβaibαj=04p1(ζpbi/2)j(vavb)\displaystyle=\frac{\zeta_{p}^{\alpha\beta}}{4p}\sum_{i=0}^{4p-1}\zeta_{p}^{-a\beta-ai-b\alpha}\sum_{j=0}^{4p-1}(\zeta_{p}^{-b-i/2})^{j}(v_{a}\otimes v_{b})
=ζpαβ4pζpaβa(2b)bα(4p)(vavb)\displaystyle=\frac{\zeta_{p}^{\alpha\beta}}{4p}\zeta_{p}^{-a\beta-a(-2b)-b\alpha}(4p)(v_{a}\otimes v_{b})
=ζpαβaβ+2abbα(vavb)\displaystyle=\zeta_{p}^{\alpha\beta-a\beta+2ab-b\alpha}(v_{a}\otimes v_{b})
=ζpαβ/2ζpHH/2(vavb)\displaystyle=\zeta_{p}^{\alpha\beta/2}\zeta_{p}^{H\otimes H/2}(v_{a}\otimes v_{b})

where ζpHH/2\zeta_{p}^{H\otimes H/2} is the operator defined by

ζpHH/2(vavb)=ζp(α2a)(β2b)/2vavb.\zeta_{p}^{H\otimes H/2}(v_{a}\otimes v_{b})=\zeta_{p}^{(\alpha-2a)(\beta-2b)/2}v_{a}\otimes v_{b}.

Thus, the action of Rα,βJ=α,βR^{J}_{\alpha,\beta}=\mathcal{H}_{\alpha,\beta}\cdot\mathcal{R} over a tensor product VαVβV_{\alpha}\otimes V_{\beta} is given by

Rα,βJ=ζpαβ/2R¯R^{J}_{\alpha,\beta}=\zeta_{p}^{\alpha\beta/2}\overline{R}

where R¯:VαVβVαVβ\overline{R}:V_{\alpha}\otimes V_{\beta}\to V_{\alpha}\otimes V_{\beta} acts by ζpHH2\zeta_{p}^{\frac{H\otimes H}{2}}\mathcal{R}.

5.6. ADO invariants

We define the ADO invariants following Murakami [21]. Let KK be a framed oriented knot which is the closure of a framed oriented (1,1)(1,1)-tangle KoK_{o}. We suppose KoK_{o} is given by a diagram DD with only upward crossings and left/right caps and cups and with the blackboard framing. Consider the above Verma module V=VαV=V_{\alpha} with the Yang-Baxter operator c¯:VVVV\overline{c}:V\otimes V\to V\otimes V induced from R¯\overline{R} above (that is c¯=PR¯\overline{c}=P\circ\overline{R} where PP is the swap vwwvv\otimes w\mapsto w\otimes v), this is the same Yang-Baxter operator used in [21]. Then, to each positive crossing of DD we associate R¯\overline{R} 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

vi,vi=ζpα(1p)2i, 1i=0p1ζpα(1p)2ivivi.\langle v_{i},v^{*}_{i}\rangle=\zeta_{p}^{\alpha(1-p)-2i},\ 1\mapsto\sum_{i=0}^{p-1}\zeta_{p}^{-\alpha(1-p)-2i}v^{*}_{i}\otimes v_{i}.

Pasting all these maps together as determined by the diagram DD defines a map VαVαV_{\alpha}\to V_{\alpha}. For generic α\alpha, it can be shown that this map is multiplication by a scalar in ζpα2w(D)/2[ζp,ζp±α]\zeta_{p}^{\alpha^{2}w(D)/2}\cdot\mathbb{Z}[\zeta_{p},\zeta_{p}^{\pm\alpha}] where w(D)w(D) is the writhe of the diagram. This is called the ADO invariant of the framed knot KK and denoted ADOp(K)ADO^{\prime}_{p}(K) (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 ζpα2/2α(p1)\zeta_{p}^{\alpha^{2}/2-\alpha(p-1)} on VαV_{\alpha}. It follows that

ζpw(D)α2/2ζpw(D)α(p1))ADOp(K)\zeta_{p}^{-w(D)\alpha^{2}/2}\zeta_{p}^{w(D)\alpha(p-1))}ADO^{\prime}_{p}(K)

is an invariant of the underlying unframed knot KK, where w(D)w(D) is the writhe of the diagram. Moreover, this invariant belongs to [ζp,ζp±α]\mathbb{Z}[\zeta_{p},\zeta_{p}^{\pm\alpha}] so setting t=ζp2αt=\zeta_{p}^{2\alpha}, this determines a polynomial in [ζp][t±1/2]\mathbb{Z}[\zeta_{p}][t^{\pm 1/2}] which we denote simply by ADOp(K,t)ADO_{p}(K,t). Note that some authors use t=ζpαt=\zeta_{p}^{\alpha} instead. This polynomial is not symmetric in tt, but it is after substituting t=ζp2xt=\zeta_{p}^{-2}x [20]. Recall from (2) that PHpθP_{H_{p}}^{\theta} is defined as

PHpθ(K)=tw(D)(p1)/2ϵD(Hp)(ZD(Hp)¯θ(Ko)).P_{H_{p}}^{\theta}(K)=t^{w(D)(p-1)/2}\epsilon_{D(H_{p})}(Z^{\theta}_{\underline{D(H_{p})}}(K_{o})).
Proposition 5.4.

We have ADOp(K,t)=PHpθ(K)ADO_{p}(K,t)=P_{H_{p}}^{\theta}(K) where θ\theta is the degree twist of HpH_{p}.

Proof.

In the above definition of ADO invariants, we used the Yang-Baxter operator c¯=ζpα2/2cα,αJ\overline{c}=\zeta_{p}^{-\alpha^{2}/2}c^{J}_{\alpha,\alpha}. Since the ζpα2/2\zeta_{p}^{-\alpha^{2}/2} 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 cα,αJc^{J}_{\alpha,\alpha} to define ADO. Using cα,αJc^{J}_{\alpha,\alpha}, a positive twist acts by multiplication by ζpα2α(p1)\zeta_{p}^{\alpha^{2}-\alpha(p-1)}. The maps associated to right caps and cups above are exactly given by multiplication by 𝒈α±1\boldsymbol{g}_{\alpha}^{\pm 1}, which is the pivot of A¯J\underline{A}^{J}. Thus, the universal invariant ZαJAαZ^{J}_{\alpha}\in A_{\alpha} defined from A¯J\underline{A}^{J} determines the ADO invariant by

ADOp(K,ζpα)v0=ζpw(D)(α2+α(p1))ZαJv0=ζpw(D)(α2+α(p1))Zαv0ADO_{p}(K,\zeta_{p}^{\alpha})v_{0}=\zeta_{p}^{w(D)(-\alpha^{2}+\alpha(p-1))}Z_{\alpha}^{J}v_{0}=\zeta_{p}^{w(D)(-\alpha^{2}+\alpha(p-1))}Z_{\alpha}v_{0}

where the second equality is Lemma 5.3. Now, by (1), ZαZ_{\alpha} is of the form

Zα=π(ZD(Hp)¯θ|t=ζp2α)k2αw(D)Z_{\alpha}=\pi(Z^{\theta}_{\underline{D(H_{p})}}|_{t=\zeta_{p}^{2\alpha}})\cdot k^{2\alpha w(D)}

where π:D(Hp)αDα\pi:D(H_{p})_{\alpha}\to D_{\alpha} is the projection (recall that we mod out by kκ1k\kappa-1). Note that ZD(Hp)¯θZ_{\underline{D(H_{p})}}^{\theta} is evaluated at t=ζp2αt=\zeta_{p}^{2\alpha} since AαA_{\alpha} is defined through the action ϕα(e)=ζp2αe\phi_{\alpha}(e)=\zeta_{p}^{2\alpha}e, that is, t=ζp2αt=\zeta_{p}^{2\alpha} on the degree twist action. Write ZD(Hp)¯θ(K)=λijlkiκjFlElZ^{\theta}_{\underline{D(H_{p})}}(K)=\sum\lambda_{ijl}k^{i}\kappa^{j}F^{l}E^{l} (rescaled generators of D(Hp)αD(H_{p})_{\alpha}), where λijl[t±1]\lambda_{ijl}\in\mathbb{C}[t^{\pm 1}]. Using Eiv0=0E^{i}v_{0}=0 (i>0i>0) we get

Zαv0=i,j,lλijlkijFlElk2αw(D)v0=ζpw(D)α2i,jλij0ζpα(ij)/2v0.\displaystyle Z_{\alpha}v_{0}=\sum_{i,j,l}\lambda_{ijl}k^{i-j}F^{l}E^{l}k^{2\alpha w(D)}v_{0}=\zeta_{p}^{w(D)\alpha^{2}}\sum_{i,j}\lambda_{ij0}\zeta_{p}^{\alpha(i-j)/2}v_{0}.

But k=ζpα/2k¯k=\zeta_{p}^{\alpha/2}\overline{k} and ϵ(k¯)=1\epsilon(\overline{k})=1, so ϵ(kij)=ζpα(ij)/2\epsilon(k^{i-j})=\zeta_{p}^{\alpha(i-j)/2}. Since also ϵ(E)=ϵ(F)=0\epsilon(E)=\epsilon(F)=0, it follows that ϵ(ZD(Hp)¯θ(K))|t=ζp2αv0=Zαv0\epsilon(Z^{\theta}_{\underline{D(H_{p})}}(K))|_{t=\zeta_{p}^{2\alpha}}v_{0}=Z_{\alpha}v_{0}. Setting ζp2α=t\zeta_{p}^{2\alpha}=t 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 𝔰𝔩3\mathfrak{sl}_{3}, 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 33-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.