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

Computing the Cassels-Tate Pairing for Genus Two Jacobians with Rational Two Torsion Points

Jiali Yan
Abstract

In this paper, we give an explicit formula as well as a practical algorithm for computing the Cassels-Tate pairing on Sel2(J)×Sel2(J)\text{Sel}^{2}(J)\times\text{Sel}^{2}(J) where JJ is the Jacobian variety of a genus two curve under the assumption that all points in J[2]J[2] are KK-rational. We also give an explicit formula for the Obstruction map Ob:H1(GK,J[2])Br(K)\text{Ob}:H^{1}(G_{K},J[2])\rightarrow\text{Br}(K) under the same assumption. Finally, we include a worked example demonstrating we can indeed improve the rank bound given by a 2-descent via computing the Cassels-Tate pairing.

1 Introduction

For any principally polarized abelian variety AA defined over a number field KK, Cassels and Tate [cassels1] [cassels2] and [tate] constructed a pairing

(A)×(A)/,\Sh(A)\times\Sh(A)\rightarrow\mathbb{Q}/\mathbb{Z},

that is nondegenerate after quotienting out the maximal divisible subgroup of (A)\Sh(A). This pairing is called the Cassels-Tate pairing and it naturally lifts to a pairing on Selmer groups. One application of this pairing is in improving the bound on the Mordell-Weil rank r(A)r(A) obtained by performing a standard descent calculation. More specifically, if (A)\Sh(A) is finite or if all the nn-torsion points of AA are defined over KK, the kernel of the Cassels-Tate pairing on Seln(A)×Seln(A)\text{Sel}^{n}(A)\times\text{Sel}^{n}(A) is equal to the image of the natural map Seln2(A)Seln(A)\text{Sel}^{n^{2}}(A)\rightarrow\text{Sel}^{n}(A) induced from the map A[n2]nA[n]A[n^{2}]\xrightarrow{n}A[n], see [thesis, Proposition 1.9.3] for details. This shows that carrying out an nn-descent and computing the Cassels-Tate pairing on Seln(A)×Seln(A)\text{Sel}^{n}(A)\times\text{Sel}^{n}(A) gives the same rank bound as obtained from n2n^{2}-descent where Seln2(A)\text{Sel}^{n^{2}}(A) needs to be computed.

There have been many results on computing the Cassels-Tate pairing in the case of elliptic curves, such as [cassels98] [steve] [binaryquartic] [monique] [3isogeny] [platonic] [3selmer]. We are interested in the natural problem of generalizing the different algorithms for computing the Cassels-Tate pairing for elliptic curves to compute the pairing for abelian varieties of higher dimension.

In Section 2, we give the preliminary results needed for the later sections, including the homogeneous space definition of the Cassels-Tate pairing. In Section 3, we state and prove an explicit formula for the pairing ,CT\langle\;,\;\rangle_{CT} on Sel2(J)×Sel2(J)\text{Sel}^{2}(J)\times\text{Sel}^{2}(J) where JJ is the Jacobian variety of a genus two curve under the assumption that all points in J[2]J[2] are KK-rational. This formula is analogous to that in the elliptic curve case in [cassels98]. In Section 4, we describe a practical algorithm for computing the pairing ,CT\langle\;,\;\rangle_{CT} using the formula in Section 3. In section 5, we also give an explicit formula for the Obstruction map Ob:H1(GK,J[2])Br(K)\text{Ob}:H^{1}(G_{K},J[2])\rightarrow\text{Br}(K) under the assumption that all points in J[2]J[2] are defined over KK generalizing the result in the elliptic curve case [oneil, Proposition 3.4], [clark, Theorem 6]. Finally, in Section 7, we include a worked example demonstrating that computing the Cassels-Tate pairing can indeed turn a 2-descent to a 4-descent and improve the rank bound given by a 2-descent. The content of this paper is based on Chapter 4 of the thesis of the author [thesis].

Acknowledgements

I express my sincere and deepest gratitute to my PhD supervisor, Dr. Tom Fisher, for his patient guidance and insightful comments at every stage during my research.

2 Preliminary Results

2.1 The set-up

In this paper, we are working over a number field KK. For any field kk, we let k¯\bar{k} denote its algebraic closure and let μnk¯\mu_{n}\subset\bar{k} denote the nthn^{th} roots of unity in k¯\bar{k}. We let GkG_{k} denote the absolute Galois group Gal(k¯/k)\text{Gal}(\bar{k}/k).

Let 𝒞\mathcal{C} be a general genus two curve defined over KK, which is a smooth projective curve. It can be given in the following hyperelliptic form:

𝒞:y2=f(x)=f6x6+f5x5+f4x4+f3x3+f2x2+f1x+f0,\mathcal{C}:y^{2}=f(x)=f_{6}x^{6}+f_{5}x^{5}+f_{4}x^{4}+f_{3}x^{3}+f_{2}x^{2}+f_{1}x+f_{0},

where fiKf_{i}\in K, f60f_{6}\neq 0 and the discriminant (f)0\triangle(f)\neq 0, which implies that ff has distinct roots in K¯\bar{K}.

We let JJ denote the Jacobian variety of 𝒞\mathcal{C}, which is an abelian variety of dimension two defined over KK that can be identified with Pic0(𝒞)\text{Pic}^{0}(\mathcal{C}). We denote the identity element of JJ by 𝒪J\mathcal{O}_{J}. Via the natural isomorphism Pic2(𝒞)Pic0(𝒞)\text{Pic}^{2}(\mathcal{C})\rightarrow\text{Pic}^{0}(\mathcal{C}) sending [P1+P2][P1+P2+][P_{1}+P_{2}]\mapsto[P_{1}+P_{2}-\infty^{+}-\infty^{-}], a point PJP\in J can be identified with an unordered pair of points of 𝒞\mathcal{C}, {P1,P2}\{P_{1},P_{2}\}. This identification is unique unless P=𝒪JP=\mathcal{O}_{J}, in which case it can be represented by any pair of points on 𝒞\mathcal{C} in the form {(x,y),(x,y)}\{(x,y),(x,-y)\} or {+,}\{\infty^{+},\infty^{-}\}. Suppose the roots of ff are denoted by ω1,,ω6\omega_{1},...,\omega_{6}. Then J[2]={𝒪J,{(ωi,0),(ωj,0)} for ij}J[2]=\{\mathcal{O}_{J},\{(\omega_{i},0),(\omega_{j},0)\}\text{ for }i\neq j\}. Also, for a point PJP\in J, we let τP:JJ\tau_{P}:J\rightarrow J denote the translation by PP on JJ.

As described in [thebook, Chapter 3, Section 3], suppose {P1,P2}\{P_{1},P_{2}\}and {Q1,Q2}\{Q_{1},Q_{2}\} represent P,QJ[2]P,Q\in J[2] where P1,P2,Q1,Q2P_{1},P_{2},Q_{1},Q_{2} are Weierstrass points, then

e2(P,Q)=(1)|{P1,P2}{Q1,Q2}|.e_{2}(P,Q)=(-1)^{|\{P_{1},P_{2}\}\cap\{Q_{1},Q_{2}\}|}.

2.2 Theta divisor and Kummer surface

The theta divisor, denoted by Θ\Theta, is defined to be the divisor on JJ that corresponds to the divisor {P}×𝒞+𝒞×{P}\{P\}\times\mathcal{C}+\mathcal{C}\times\{P\} on 𝒞×𝒞\mathcal{C}\times\mathcal{C} under the birational morphism Sym2𝒞J\text{Sym}^{2}\mathcal{C}\rightarrow J, for some Weierstrass point P𝒞P\in\mathcal{C}. The Jacobian variety JJ is principally polarized abelian variety via λ:JJ\lambda:J\rightarrow J^{\vee} sending PP to [τPΘΘ].[\tau_{P}^{*}\Theta-\Theta].

The Kummer surface, denoted by 𝒦\mathcal{K}, is the quotient of JJ via the involution [1]:PP[-1]:P\mapsto-P. The fixed points under the involution are the 16 points of order 2 on JJ and these map to the 16 nodal singular points of 𝒦\mathcal{K} (the nodes). General theory, as in [abelianvarieties, Theorem 11.1], [theta, page 150], shows that the linear system of nΘn\Theta of JJ has dimension n2n^{2}. Moreover, |2Θ||2\Theta| is base point free and |4Θ||4\Theta| is very ample.

2.3 Explicit embeddings of JJ and 𝒦\mathcal{K}

Denote a generic point on the Jacobian JJ of 𝒞\mathcal{C} by {(x,y),(u,v)}\{(x,y),(u,v)\}. Then, following [thebook, Chapter 3, Section 1], the morphism from JJ to 3\mathbb{P}^{3} is given by

k1=1,k2=(x+u),k3=xu,k4=β0,k_{1}=1,k_{2}=(x+u),k_{3}=xu,k_{4}=\beta_{0},

where

β0=F0(x,u)2yv(xu)2\beta_{0}=\frac{F_{0}(x,u)-2yv}{(x-u)^{2}}

with F0(x,u)=2f0+f1(x+u)+2f2(xu)+f3(x+u)(xu)+2f4(xu)2+f5(x+u)(xu)2+2f6(xu)3.F_{0}(x,u)=2f_{0}+f_{1}(x+u)+2f_{2}(xu)+f_{3}(x+u)(xu)+2f_{4}(xu)^{2}+f_{5}(x+u)(xu)^{2}+2f_{6}(xu)^{3}.

We denote the above morphism by J|2Θ|𝒦3J\xrightarrow{|2\Theta|}\mathcal{K}\subset\mathbb{P}^{3} and it maps 𝒪J\mathcal{O}_{J} to (0:0:0:1)(0:0:0:1). It is known that its image in 3\mathbb{P}^{3} is precisely the Kummer surface 𝒦\mathcal{K} and is given by the vanishing of the quartic G(k1,k2,k3,k4)G(k_{1},k_{2},k_{3},k_{4}) with explicit formula given in [thebook, Chapter 3, Section 1]. Therefore, the Kummer surface 𝒦3ki\mathcal{K}\subset\mathbb{P}^{3}_{k_{i}} is defined by G(k1,k2,k3,k4)=0.G(k_{1},k_{2},k_{3},k_{4})=0.

Remark 2.1.

Suppose PJ[2]P\in J[2]. We know τP(2Θ)2Θ\tau_{P}^{*}(2\Theta)\sim 2\Theta via the polarization. This implies that translation by PP on JJ induces a linear isomorphism on 𝒦3\mathcal{K}\subset\mathbb{P}^{3}.

We now look at the embedding of JJ in 15\mathbb{P}^{15} induced by |4Θ||4\Theta|. Let kij=kikjk_{ij}=k_{i}k_{j}, for 1ij41\leq i\leq j\leq 4. Since 𝒦\mathcal{K} is irreducible and defined by a polynomial of degree 4, k11,k12,,k44k_{11},k_{12},...,k_{44} are 10 linearly independent even elements in (2Θ++2Θ)\mathcal{L}(2\Theta^{+}+2\Theta^{-}). The six odd basis elements in (2Θ++2Θ)\mathcal{L}(2\Theta^{+}+2\Theta^{-}) are given explicitly in [explicittwist, Section 3]. A function gg on JJ is even when it is invariant under the involution [1]:PP[-1]:P\mapsto-P and is odd when g[1]=gg\circ[-1]=-g.

Unless stated otherwise, we will use the basis k11,k12,,k44,b1,,b6k_{11},k_{12},...,k_{44},b_{1},...,b_{6} for (2Θ++2Θ)\mathcal{L}(2\Theta^{+}+2\Theta^{-}), to embed JJ in 15\mathbb{P}^{15}. The following theorem gives the defining equations of JJ.

Theorem 2.2.

([72theorem, Therorem 1.2], [thegplawpaper, Therorem 1.2]) Let JJ be the Jacobian variety of the genus two curve 𝒞\mathcal{C} defined by y2=f6x6++f1x+f0y^{2}=f_{6}x^{6}+...+f_{1}x+f_{0}. The 7272 quadratic forms over [f0,,f6]\mathbb{Z}[f_{0},...,f_{6}] given in [72theorem, Appendix A] are a set of defining equations for the projective variety given by the embedding of JJ in 15\mathbb{P}^{15} induced by the basis of (2Θ++2Θ)\mathcal{L}(2\Theta^{+}+2\Theta^{-}) with explicit formulae given in [72theorem, Definition 1.1] or [thegplawpaper, Definition 1.1]. The change of basis between this basis of (2Θ++2Θ)\mathcal{L}(2\Theta^{+}+2\Theta^{-}) and k11,k12,,k44,b1,,b6k_{11},k_{12},...,k_{44},b_{1},...,b_{6} is given in [explicittwist, Section 3].

2.4 Principal homogeneous space and 2-coverings

A principal homogeneous space or torsor for JJ defined over a field KK is a variety VV together with a morphism μ:J×VV\mu:J\times V\rightarrow V, both defined over KK, that induces a simply transitive action on the K¯\bar{K}-points.

We say (V1,μ1)(V_{1},\mu_{1}) and (V2,μ2)(V_{2},\mu_{2}) are isomorphic over a field extension K1K_{1} of KK if there is an isomorphism ϕ:V1V2\phi:V_{1}\rightarrow V_{2} defined over K1K_{1} that respects the action of JJ.

A 2-covering of JJ is a variety XX defined over KK together with a morphism π:XJ\pi:X\rightarrow J defined over KK, such that there exists an isomorphism ϕ:XJ\phi:X\rightarrow J defined over K¯\bar{K} with π=[2]ϕ\pi=[2]\circ\phi. An isomorphism (X1,π1)(X2,π2)(X_{1},\pi_{1})\rightarrow(X_{2},\pi_{2}) between two 2-coverings is an isomorphism h:X1X2h:X_{1}\rightarrow X_{2} defined over KK with π1=π2h\pi_{1}=\pi_{2}\circ h. We sometimes denote (X,π)(X,\pi) by XX when the context is clear.

It can be checked that a 2-covering is a principal homogeneous space. The short exact sequence 0J[2]J2J00\rightarrow J[2]\rightarrow J\xrightarrow{2}J\rightarrow 0 induces the connecting map in the long exact sequence

δ:J(K)H1(GK,J[2]).\delta:J(K)\rightarrow H^{1}(G_{K},J[2]). (2.1)

The following two propositions are proved in [explicittwist].

Proposition 2.3.

[explicittwist, Lemma 2.14] Let (X,π)(X,\pi) be a 22-covering of an abelian variety JJ defined over KK and choose an isomorphism ϕ:XJ\phi:X\rightarrow J such that π=[2]ϕ\pi=[2]\circ\phi. Then for each σGK\sigma\in G_{K}, there is a unique point PJ[2](K¯)P\in J[2](\bar{K}) satisfying ϕσ(ϕ1)=τP\phi\circ\sigma(\phi^{-1})=\tau_{P}. The map σP\sigma\mapsto P is a cocycle whose class in H1(GK,J[2])H^{1}(G_{K},J[2]) does not depend on the choice of ϕ\phi. This yields a bijection between the set of isomorphism classes of 22-coverings of JJ and the set H1(GK,J[2])H^{1}(G_{K},J[2]).

Proposition 2.4.

[explicittwist, Proposition 2.15] Let XX be a 22-covering of JJ corresponding to the cocycle class ϵH1(GK,J[2])\epsilon\in H^{1}(G_{K},J[2]). Then XX contains a KK-rational point (equivalently XX is a trivial principle homogeneous space) if and only if ϵ\epsilon is in the image of the connecting map δ\delta in (2.1).

We also state and prove the following proposition which is useful for the computation of the Cassels-Tate pairing later in Sections LABEL:sec:explicit-computation-of-d and LABEL:sec:explicit-computation-of-dp-dq-dr-ds. A Brauer-Severi variety is a variety that is isomorphic to a projective space over K¯\bar{K}.

Proposition 2.5.

Let (X,π)(X,\pi) be a 22-covering of J, with ϕ[2]=π\phi\circ[2]=\pi. Then the linear system |ϕ(2Θ)||\phi^{*}(2\Theta)| determines a map XSX\rightarrow S defined over KK, where SS is a Brauer-Severi variety. Also, there exists an isomorphism ψ\psi defined over K¯\bar{K} making the following diagram commute:

X{X}S{S}J{J}3.{\mathbb{P}^{3}.}|ϕ(2Θ)|\scriptstyle{|\phi^{*}(2\Theta)|}ϕ\scriptstyle{\phi}ψ\scriptstyle{\psi}|2Θ|\scriptstyle{|2\Theta|} (2.2)

In particular, if (X,π)(X,\pi) corresponds to a Selmer element via the correspondence in Proposition  2.3, then the Brauer-Severi variety SS is isomorphic to 3\mathbb{P}^{3}.

Proof.

Since (X,π)(X,\pi) is a 22-covering of JJ, by Proposition 2.3, we have that for each σGK\sigma\in G_{K}, ϕ(ϕ1)σ=τP\phi\circ(\phi^{-1})^{\sigma}=\tau_{P} for some PJ[2]P\in J[2]. The principal polarization gives τP(2Θ)2Θ\tau_{P}^{*}(2\Theta)\sim 2\Theta which implies that ϕ(2Θ)(ϕσ)(2Θ)\phi^{*}(2\Theta)\sim(\phi^{\sigma})^{*}(2\Theta), hence the morphism induced by |ϕ(2Θ)||\phi^{*}(2\Theta)| is defined over KK.

Now if (X,π)(X,\pi) corresponds to a Selmer element, then XX everywhere locally has a point by Proposition 2.4, and hence SS everywhere locally has a point. Since the Hasse principle holds for Brauer-Severi varieties by [bshasse, Corollary 2.6], we know that SS has a point over KK and hence it is isomorphic to 3\mathbb{P}^{3} by [QA, Therem 5.1.3].

We now make some observations and give some notation.

Remark 2.6.

Let ϵSel2(J)\epsilon\in\text{Sel}^{2}(J), and let (Jϵ,πϵ)(J_{\epsilon},\pi_{\epsilon}) denote the 2-covering corresponding to ϵ\epsilon. There exists an isomorphism ϕϵ\phi_{\epsilon} defined over K¯\bar{K} such that [2]ϕϵ=πϵ[2]\circ\phi_{\epsilon}=\pi_{\epsilon}. Then, by Proposition 2.5, we have the following commutative diagram:

Jϵ{J_{\epsilon}}𝒦ϵ3{\mathcal{K}_{\epsilon}\subset\mathbb{P}^{3}}J{J}𝒦3.{\mathcal{K}\subset\mathbb{P}^{3}.}|ϕϵ(2Θ)|\scriptstyle{|\phi_{\epsilon}^{*}(2\Theta)|}ϕϵ\scriptstyle{\phi_{\epsilon}}ψϵ\scriptstyle{\psi_{\epsilon}}|2Θ|\scriptstyle{|2\Theta|} (2.3)

The image of JϵJ_{\epsilon} under the morphism induced by |ϕϵ(2Θ)||\phi_{\epsilon}^{*}(2\Theta)| is a surface, denoted by 𝒦ϵ\mathcal{K}_{\epsilon}, which we call the twisted Kummer surface corresponding to ϵ\epsilon. Also ψϵ\psi_{\epsilon} is a linear isomorphism 33\mathbb{P}^{3}\rightarrow\mathbb{P}^{3} defined over K¯\bar{K}.

Notation 2.7.

Suppose (Jϵ,πϵ)(J_{\epsilon},\pi_{\epsilon}) is the 2-covering of JJ corresponding to ϵH1(GK,J[2])\epsilon\in H^{1}(G_{K},J[2]). The involution [1]:PP[-1]:P\mapsto-P on JJ induces an involution ιϵ\iota_{\epsilon} on JϵJ_{\epsilon} such that ϕϵιϵ=[1]ϕϵ\phi_{\epsilon}\circ\iota_{\epsilon}=[-1]\circ\phi_{\epsilon}, where [2]ϕϵ=πϵ[2]\circ\phi_{\epsilon}=\pi_{\epsilon}. Moreover, the degree 2 morphism Jϵ|ϕϵ(2Θ)|𝒦ϵ3J_{\epsilon}\xrightarrow{|\phi_{\epsilon}^{*}(2\Theta)|}\mathcal{K}_{\epsilon}\subset\mathbb{P}^{3} in (2.3) is precisely the quotient by ιϵ\iota_{\epsilon} and so an alternative definition of 𝒦ϵ\mathcal{K}_{\epsilon} is as the quotient of JϵJ_{\epsilon} by ιϵ\iota_{\epsilon}. We call a function gg on JϵJ_{\epsilon} even if it is invariant under ιϵ\iota_{\epsilon} and odd if gιϵ=gg\circ\iota_{\epsilon}=-g.

2.5 Definition of the Cassels-Tate Pairing

There are four equivalent definitions of the Cassels-Tate pairing stated and proved in [poonenstoll]. In this paper we will only be using the homogeneous space definition of the Cassels-Tate pairing. Suppose a,a(J)a,a^{\prime}\in\Sh(J). Via the polarization λ\lambda, we get aba^{\prime}\mapsto b where b(J).b\in\Sh(J^{\vee}). Let XX be the (locally trivial) principal homogeneous space defined over KK representing aa. Then Pic0(XK¯)\text{Pic}^{0}(X_{\bar{K}}) is canonically isomorphic as a GKG_{K}-module to Pic0(JK¯)=J(K¯).\text{Pic}^{0}(J_{\bar{K}})=J^{\vee}(\bar{K}). Therefore, b(J)H1(GK,J)b\in\Sh(J^{\vee})\subset H^{1}(G_{K},J^{\vee}) represents an element in H1(GK,Pic0(XK¯))H^{1}(G_{K},\text{Pic}^{0}(X_{\bar{K}})).

Now consider the exact sequence:

0K¯(X)/K¯Div0(XK¯)Pic0(XK¯)0.0\rightarrow\bar{K}(X)^{*}/\bar{K}^{*}\rightarrow\text{Div}^{0}(X_{\bar{K}})\rightarrow\text{Pic}^{0}(X_{\bar{K}})\rightarrow 0.

We can then map bb to an element bH2(GK,K¯(X)/K¯)b^{\prime}\in H^{2}(G_{K},\bar{K}(X)^{*}/\bar{K}^{*}) using the long exact sequence associated to the short exact sequence above. Since H3(GK,K¯)=0H^{3}(G_{K},\bar{K}^{*})=0, bb^{\prime} has a lift fH2(GK,K¯(X))f^{\prime}\in H^{2}(G_{K},\bar{K}(X)^{*}) via the long exact sequence induced by the short exact sequence 0K¯K¯(X)K¯(X)/K¯0:0\rightarrow\bar{K}^{*}\rightarrow\bar{K}(X)^{*}\rightarrow\bar{K}(X)^{*}/\bar{K}^{*}\rightarrow 0:

H2(GK,K¯)H2(GK,K¯(X))H2(GK,K¯(X)/K¯)H3(GK,K¯)=0.H^{2}(G_{K},\bar{K}^{*})\rightarrow H^{2}(G_{K},\bar{K}(X)^{*})\rightarrow H^{2}(G_{K},\bar{K}(X)^{*}/\bar{K}^{*})\rightarrow H^{3}(G_{K},\bar{K}^{*})=0. (2.4)

Next we show that fvH2(GKv,Kv¯(X))f^{\prime}_{v}\in H^{2}(G_{K_{v}},\bar{K_{v}}(X)^{*}) is the image of an element cvH2(GKv,Kv¯).c_{v}\in H^{2}(G_{K_{v}},\bar{K_{v}}^{*}). This is because b(J)b\in\Sh(J^{\vee}) is locally trivial which implies its image bb^{\prime} is locally trivial. Then the statement is true by the exactness of local version of sequence (2.4).

We then can define

a,b=vinvv(cv)/.\langle a,b\rangle=\sum_{v}\text{inv}_{v}(c_{v})\in\mathbb{Q}/\mathbb{Z}.

The Cassels-Tate pairing (J)×(J)/\Sh(J)\times\Sh(J)\rightarrow\mathbb{Q}/\mathbb{Z} is defined by

a,aCT:=a,λ(a).\langle a,a^{\prime}\rangle_{CT}:=\langle a,\lambda(a^{\prime})\rangle.

We sometimes refer to invv(cv)\text{inv}_{v}(c_{v}) above as the local Cassels-Tate pairing between a,a(J)a,a^{\prime}\in\Sh(J) for a place vv of KK. Note that the local Cassels-Tate pairing depends on the choice of fH2(GK,K¯(X))f^{\prime}\in H^{2}(G_{K},\bar{K}(X)^{*}). We make the following remarks that are useful for the computation for the Cassels-Tate pairing.

Remark 2.8.


  1. (i)

    By [poonenstoll], we know the homogeneous space definition of the Cassels-Tate pairing is independent of all the choices we make.

  2. (ii)

    Via the map Sel2(J)(J)[2]\text{Sel}^{2}(J)\rightarrow\Sh(J)[2], the definition of the Cassels-Tate pairing on (J)[2]×(J)[2]\Sh(J)[2]\times\Sh(J)[2] naturally lifts to a pairing on Sel2(J)×Sel2(J)\text{Sel}^{2}(J)\times\text{Sel}^{2}(J). In fact, from now on, we will only be considering ϵ,ηCT\langle\epsilon,\eta\rangle_{CT} for ϵ,ηSel2(J)\epsilon,\eta\in\text{Sel}^{2}(J). The principal homogeneous space XX in the definition is always taken to be the 22-covering of JJ corresponding to ϵ\epsilon. One can compute cvc_{v} by evaluating fvf_{v}^{\prime} at a point in X(Kv)X(K_{v}) provided that one avoids the zeros and poles of fvf_{v}^{\prime}. Note that X(Kv)X(K_{v})\neq\emptyset by Proposition 2.4.

2.6 Explicit 2-coverings of JJ

Let Ω\Omega represent the set of 6 roots of ff, denoted by ω1,,ω6\omega_{1},...,\omega_{6}. Recall, as in Proposition 2.3, the isomorphism classes of 2-coverings of JJ are parameterized by H1(GK,J[2])H^{1}(G_{K},J[2]). For the explicit computation of the Cassels-Tate pairing, we need the following result on the explicit 2-coverings of JJ corresponding to elements in Sel2(J)\text{Sel}^{2}(J). We note that this theorem in fact works over any field of characteristic different from 2.

Theorem 2.9.

[explicittwist, Proposition 7.2, Theorem 7.4, Appendix B] Let JJ be the Jacobian variety of a genus two curve defined by y2=f(x)y^{2}=f(x) where ff is a degree 6 polynomial and ϵSel2(J)\epsilon\in\mathrm{Sel}^{2}(J). Embed JJ in 15\mathbb{P}^{15} via the coordinates k11,k12,,k44,b1,,b6k_{11},k_{12},...,k_{44},b_{1},...,b_{6}. There exists Jϵ15J_{\epsilon}\subset\mathbb{P}^{15} defined over KK with Galois invariant coordinates u0,,u9,v1,,v6u_{0},...,u_{9},v_{1},...,v_{6} and a linear isomorphism ϕϵ:JϵJ\phi_{\epsilon}:J_{\epsilon}\rightarrow J such that (Jϵ,[2]ϕϵ)(J_{\epsilon},[2]\circ\phi_{\epsilon}) is a 22-covering of JJ whose isomorphism class corresponds to the cocycle class ϵ\epsilon. Moreover, ϕϵ\phi_{\epsilon} can be explicitly represented by the 16×1616\times 16 matrix R=[R100R2]R=\begin{bmatrix}R_{1}&0\\ 0&R_{2}\\ \end{bmatrix} for some 10×1010\times 10 matrix R1R_{1} and some 6×66\times 6 matrix R2R_{2}.

Remark 2.10.

The explicit formula for ϕϵ\phi_{\epsilon} is given in the beginning of [explicittwist, Section 7] and depends only on ϵ\epsilon and the underlying genus two curve. Note that the coordinates u0,,u9,v1,,v6u_{0},...,u_{9},v_{1},...,v_{6} are derived from another set of coordinates c0,,c9,d1,,d6c_{0},...,c_{9},d_{1},...,d_{6} defined in [explicittwist, Definitions 6.9, 6.11] where c0,,c9c_{0},...,c_{9} are even and d1,,d6d_{1},...,d_{6} are odd. This set of coordinates are in general not Galois invariant, however, they are in the case where all points of J[2]J[2] are defined over the base field.

3 Formula for the Cassels-Tate Pairing

From now on, we always assume that the genus two curve 𝒞\mathcal{C} is defined by y2=f(x)y^{2}=f(x) such that all roots of ff are defined over KK. Note that this implies that all points in J[2]J[2] are defined over KK which is equivalent to all the Weierstrass points defined over KK. In this section, under the above assumption, we state and prove an explicit formula for the Cassels-Tate pairing on Sel2(J)×Sel2(J)\text{Sel}^{2}(J)\times\text{Sel}^{2}(J).

Let the genus two curve 𝒞\mathcal{C} be of the form

𝒞:y2=λ(xω1)(xω2)(xω3)(xω4)(xω5)(xω6),\mathcal{C}:y^{2}=\lambda(x-\omega_{1})(x-\omega_{2})(x-\omega_{3})(x-\omega_{4})(x-\omega_{5})(x-\omega_{6}),

where λ,ωiK\lambda,\omega_{i}\in K and λ0\lambda\neq 0. Its Jacobian variety is denoted by JJ.

The 2-torsion subgroup J[2]J[2] has basis

P={(ω1,0),(ω2,0)},\displaystyle P=\{(\omega_{1},0),(\omega_{2},0)\}, Q={(ω1,0),(ω3,0)},\displaystyle\;\;\;Q=\{(\omega_{1},0),(\omega_{3},0)\},
R={(ω4,0),(ω5,0)},\displaystyle R=\{(\omega_{4},0),(\omega_{5},0)\}, S={(ω4,0),(ω6,0)}.\displaystyle\;\;\;S=\{(\omega_{4},0),(\omega_{6},0)\}.

By the discussion at the end of Section 2.1, the Weil pairing is given relative to this basis by:

W=[1111111111111111].W=\begin{bmatrix}1&-1&1&1\\ -1&1&1&1\\ 1&1&1&-1\\ 1&1&-1&1\\ \end{bmatrix}. (3.1)

More explicitly, WijW_{ij} denotes the Weil pairing between the ithi^{th} and jthj^{th} generators.

We now show that this choice of basis determines an isomorphism H1(GK,J[2])(K/(K)2)4H^{1}(G_{K},J[2])\cong(K^{*}/(K^{*})^{2})^{4}. Consider the map J[2]w2(μ2(K¯))4,J[2]\xrightarrow{w_{2}}(\mu_{2}(\bar{K}))^{4}, where w2w_{2} denotes taking the Weil pairing with P,Q,R,SP,Q,R,S. Since P,Q,R,SP,Q,R,S form a basis for J[2]J[2] and the Weil pairing is a nondegerate bilinear pairing, we get that w2w_{2} is injective. This implies that w2w_{2} is an isomorphism as |J[2]|=|(μ2(K¯))4|=16|J[2]|=|(\mu_{2}(\bar{K}))^{4}|=16. We then get

H1(GK,J[2])w2,H1(GK,(μ2(K¯))4)(K/(K)2)4,H^{1}(G_{K},J[2])\xrightarrow{w_{2,*}}H^{1}(G_{K},(\mu_{2}(\bar{K}))^{4})\cong(K^{*}/(K^{*})^{2})^{4}, (3.2)

where w2,w_{2,*} is induced by w2w_{2} and \cong is the Kummer isomorphism derived from Hilbert’s Theorem 90. Since the map (3.2) is an isomorphism, we can represent elements in H1(GK,J[2])H^{1}(G_{K},J[2]) by elements in (K/(K)2)4(K^{*}/(K^{*})^{2})^{4}.

Before stating and proving the formula for the Cassels-Tate pairing, we first state and prove the following lemma.

Lemma 3.1.

For ϵSel2(J)\epsilon\in\mathrm{Sel}^{2}(J), let (Jϵ,πϵ)(J_{\epsilon},\pi_{\epsilon}) denote the corresponding 22-covering of JJ. Hence, there exists an isomorphism ϕϵ:JϵJ\phi_{\epsilon}:J_{\epsilon}\rightarrow J defined over K¯\bar{K} such that [2]ϕϵ=πϵ[2]\circ\phi_{\epsilon}=\pi_{\epsilon}. Suppose TJ(K)T\in J(K) and T1J(K¯)T_{1}\in J(\bar{K}) satisfy 2T1=T2T_{1}=T. Then

  1. (i)

    There exists a KK-rational divisor DTD_{T} on JϵJ_{\epsilon} which represents the divisor class of ϕϵ(τT1(2Θ))\phi_{\epsilon}^{*}(\tau_{T_{1}}^{*}(2\Theta)).

  2. (ii)

    Let DD and DTD_{T} be KK-rational divisors on JϵJ_{\epsilon} representing the divisor class of ϕϵ(2Θ)\phi_{\epsilon}^{*}(2\Theta) and ϕϵ(τT1(2Θ))\phi_{\epsilon}^{*}(\tau_{T_{1}}^{*}(2\Theta)) respectively. Then DTDϕϵ(τTΘΘ)D_{T}-D\sim\phi_{\epsilon}^{*}(\tau_{T}^{*}\Theta-\Theta). Suppose TT is a two torsion point. Then 2DT2D2D_{T}-2D is a KK-rational principal divisor. Hence, there exists a KK-rational function fTf_{T} on JϵJ_{\epsilon} such that div(fT)=2DT2D\text{div}(f_{T})=2D_{T}-2D.

Proof.

By definition of a 2-covering, [2]ϕϵ=πϵ[2]\circ\phi_{\epsilon}=\pi_{\epsilon} is a morphism defined over KK. Also, by Proposition 2.3, ϕϵ(ϕϵ1)σ=τϵσ\phi_{\epsilon}\circ(\phi_{\epsilon}^{-1})^{\sigma}=\tau_{\epsilon_{\sigma}} for all σGK\sigma\in G_{K}, where (σϵσ)(\sigma\mapsto\epsilon_{\sigma}) is a cocycle representing ϵ\epsilon. Since [2]τT1ϕϵ=τT[2]ϕϵ=τTπϵ[2]\circ\tau_{T_{1}}\circ\phi_{\epsilon}=\tau_{T}\circ[2]\circ\phi_{\epsilon}=\tau_{T}\circ\pi_{\epsilon} and τT\tau_{T} is defined over KK, (Jϵ,τTπϵ)(J_{\epsilon},\tau_{T}\circ\pi_{\epsilon}) is also a 2-covering of JJ. We compute τT1ϕϵ((τT1ϕϵ)1)σ=τT1ϕϵ(ϕϵ1)στσ(T1)=τϵστT1τσ(T1),\tau_{T_{1}}\circ\phi_{\epsilon}\circ((\tau_{T_{1}}\circ\phi_{\epsilon})^{-1})^{\sigma}=\tau_{T_{1}}\circ\phi_{\epsilon}\circ(\phi_{\epsilon}^{-1})^{\sigma}\circ\tau_{-\sigma(T_{1})}=\tau_{\epsilon_{\sigma}}\circ\tau_{T_{1}}\circ\tau_{-\sigma(T_{1})}, for all σGK\sigma\in G_{K}. This implies the 2-covering (Jϵ,τTπϵ)(J_{\epsilon},\tau_{T}\circ\pi_{\epsilon}) corresponds to the element in H1(GK,J[2])H^{1}(G_{K},J[2]) that is represented by the cocycle (σϵσ+T1σ(T1))(\sigma\mapsto\epsilon_{\sigma}+T_{1}-\sigma(T_{1})). Hence, (Jϵ,τTπϵ)(J_{\epsilon},\tau_{T}\circ\pi_{\epsilon}) is the 2-covering of JJ corresponding to ϵ+δ(T)\epsilon+\delta(T), where δ\delta is the connecting map as in (2.1). By Proposition 2.5, there exists a commutative diagram:

Jϵ{J_{\epsilon}}3{\mathbb{P}^{3}}J{J}3,{\mathbb{P}^{3},}|ϕϵ(τT1(2Θ))|\scriptstyle{|\phi_{\epsilon}^{*}(\tau_{T_{1}}^{*}(2\Theta))|}τT1ϕϵ\scriptstyle{\tau_{T_{1}}\circ\phi_{\epsilon}}ψϵ\scriptstyle{\psi_{\epsilon}}|2Θ|\scriptstyle{|2\Theta|}

where the morphism Jϵ|ϕϵ(τT1(2Θ))|3J_{\epsilon}\xrightarrow{|\phi_{\epsilon}^{*}(\tau_{T_{1}}^{*}(2\Theta))|}\mathbb{P}^{3} is defined over KK. So the pull back of a hyperplane section via this morphism gives us a rational divisor DTD_{T} representing the divisor class of ϕϵ(τT1(2Θ))\phi_{\epsilon}^{*}(\tau_{T_{1}}^{*}(2\Theta)) as required by (i).

Since the polarization λ:JJ\lambda:J\rightarrow J^{\vee} is an isomorphism and 2T1=T2T_{1}=T, we have ϕϵ(λ(T))=[ϕϵ(τTΘΘ)]=[ϕϵ(τT1(2Θ))][ϕϵ(2Θ)]=[DT][D]\phi_{\epsilon}^{*}(\lambda(T))=[\phi_{\epsilon}^{*}(\tau_{T}^{*}\Theta-\Theta)]=[\phi_{\epsilon}^{*}(\tau_{T_{1}}^{*}(2\Theta))]-[\phi_{\epsilon}^{*}(2\Theta)]=[D_{T}]-[D]. The fact that TT is a two torsion point implies that 2ϕϵ(λ(P))=02\phi_{\epsilon}^{*}(\lambda(P))=0. Hence, 2DT2D2D_{T}-2D is a KK-rational principal divisor which gives (ii).

The following remark explains how we will use Lemma 3.1 in the formula for the Cassels-Tate pairing on Sel2(J)×Sel2(J)\text{Sel}^{2}(J)\times\text{Sel}^{2}(J).

Remark 3.2.

Applying Lemma 3.1(i) with T=𝒪J,P,Q,R,SJ[2]T=\mathcal{O}_{J},P,Q,R,S\in J[2] gives divisors D=D𝒪JD=D_{\mathcal{O}_{J}} and DP,DQ,DRDSD_{P},D_{Q},D_{R}D_{S}. Then by Lemma 3.1(ii), there exist KK-rational functions fP,fQ,fR,fSf_{P},f_{Q},f_{R},f_{S} on JϵJ_{\epsilon} such that div(fT)=2DT2D\text{div}(f_{T})=2D_{T}-2D for T=P,Q,R,ST=P,Q,R,S.

Theorem 3.3.

Let JJ be the Jacobian variety of a genus two curve 𝒞\mathcal{C} defined over a number field KK where all points in J[2]J[2] are defined over KK. For any ϵ,ηSel2(J)\epsilon,\eta\in\mathrm{Sel}^{2}(J), let (Jϵ,[2]ϕϵ)(J_{\epsilon},[2]\circ\phi_{\epsilon}) be the 22-covering of JJ corresponding to ϵ\epsilon where ϕϵ:JϵJ\phi_{\epsilon}:J_{\epsilon}\rightarrow J is an isomorphism defined over K¯\bar{K}. Fix a choice of basis P,Q,R,SP,Q,R,S for J[2]J[2], with the Weil pairing given by matrix (3.1). Let (a,b,c,d)(a,b,c,d) denote the image of η\eta via H1(GK,J[2])(K/(K)2)4H^{1}(G_{K},J[2])\cong(K^{*}/(K^{*})^{2})^{4}, where this is the isomorphism induced by taking the Weil pairing with P,Q,R,SP,Q,R,S. Let fP,fQ,fR,fSf_{P},f_{Q},f_{R},f_{S} be the KK-rational functions on JϵJ_{\epsilon} defined in Remark 3.2. Then the Cassels-Tate pairing ,CT:Sel2(J)×Sel2(J){±1}\langle\;,\;\rangle_{CT}:\mathrm{Sel}^{2}(J)\times\mathrm{Sel}^{2}(J)\rightarrow\{\pm 1\} is given by

ϵ,ηCT=place v(fP(Pv),b)v(fQ(Pv),a)v(fR(Pv),d)v(fS(Pv),c)v,\langle\epsilon,\eta\rangle_{CT}=\prod_{\text{place }v}(f_{P}(P_{v}),b)_{v}(f_{Q}(P_{v}),a)_{v}(f_{R}(P_{v}),d)_{v}(f_{S}(P_{v}),c)_{v},

where (,)v(\;,\;)_{v} denotes the Hilbert symbol for a given place vv of KK and PvP_{v} i