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 𝒦ki3\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 4.2 and 4.3. 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} is an arbitrary choice of a local point on JϵJ_{\epsilon} avoiding the zeros and poles of fP,fQ,fR,fSf_{P},f_{Q},f_{R},f_{S}.

Proof.

We know ηH1(GK,J[2])\eta\in H^{1}(G_{K},J[2]) corresponds to (a,b,c,d)(K/(K)2)4(a,b,c,d)\in(K^{*}/(K^{*})^{2})^{4} via taking the Weil pairing with P,Q,R,SP,Q,R,S. Hence, η\eta is represented by the cocycle

σb~σP+a~σQ+d~σR+c~σS,\sigma\mapsto\tilde{b}_{\sigma}P+\tilde{a}_{\sigma}Q+\tilde{d}_{\sigma}R+\tilde{c}_{\sigma}S,

where σGK\sigma\in G_{K} and for each element xK/(K)2x\in K^{*}/(K^{*})^{2}, we define x~σ{0,1}\tilde{x}_{\sigma}\in\{0,1\} such that (1)x~σ=σ(x)/x(-1)^{\tilde{x}_{\sigma}}=\sigma(\sqrt{x})/\sqrt{x}.

Then the image of η\eta in H1(GK,Pic0(Jϵ))H^{1}(G_{K},\text{Pic}^{0}(J_{\epsilon})) is represented by the cocycle that sends σGK\sigma\in G_{K} to

b~σϕϵ[τPΘΘ]+a~σϕϵ[τQΘΘ]+d~σϕϵ[τRΘΘ]+c~σϕϵ[τSΘΘ].\tilde{b}_{\sigma}\phi_{\epsilon}^{*}[\tau_{P}^{*}\Theta-\Theta]+\tilde{a}_{\sigma}\phi_{\epsilon}^{*}[\tau_{Q}^{*}\Theta-\Theta]+\tilde{d}_{\sigma}\phi_{\epsilon}^{*}[\tau_{R}^{*}\Theta-\Theta]+\tilde{c}_{\sigma}\phi_{\epsilon}^{*}[\tau_{S}^{*}\Theta-\Theta].

By Remark 3.2, there exist KK-rational divisors DP,DQ,DR,DSD_{P},D_{Q},D_{R},D_{S} on JϵJ_{\epsilon} such that the above cocycle sends σGK\sigma\in G_{K} to

b~σ[DPD]+a~σ[DQD]+d~σ[DRD]+c~σ[DSD].\tilde{b}_{\sigma}[D_{P}-D]+\tilde{a}_{\sigma}[D_{Q}-D]+\tilde{d}_{\sigma}[D_{R}-D]+\tilde{c}_{\sigma}[D_{S}-D].

We need to map this element in H1(GK,Pic0(Jϵ))H^{1}(G_{K},\text{Pic}^{0}(J_{\epsilon})) to an element in H2(GK,K¯(Jϵ)/K¯)H^{2}(G_{K},\bar{K}(J_{\epsilon})^{*}/\bar{K}^{*}) via the connecting map induced by the short exact sequence

0K¯(Jϵ)/K¯Div0(Jϵ)Pic0(Jϵ)0.0\rightarrow\bar{K}(J_{\epsilon})^{*}/\bar{K}^{*}\rightarrow\text{Div}^{0}(J_{\epsilon})\rightarrow\text{Pic}^{0}(J_{\epsilon})\rightarrow 0.

Hence, by the formula for the connecting map and the fact that the divisors D,DP,DQ,DR,D,D_{P},D_{Q},D_{R}, DSD_{S} are all KK-rational, we get that the corresponding element in H2(GK,K¯(Jϵ)/K¯)H^{2}(G_{K},\bar{K}(J_{\epsilon})^{*}/\bar{K}^{*}) has image in H2(GK,Div0(Jϵ))H^{2}(G_{K},\text{Div}^{0}(J_{\epsilon})) represented by the following cocycle:

(σ,τ)\displaystyle(\sigma,\tau)\mapsto (b~τb~στ+b~σ)(DPD)+(a~τa~στ+a~σ)(DQD)\displaystyle(\tilde{b}_{\tau}-\tilde{b}_{\sigma\tau}+\tilde{b}_{\sigma})(D_{P}-D)+(\tilde{a}_{\tau}-\tilde{a}_{\sigma\tau}+\tilde{a}_{\sigma})(D_{Q}-D)
+(d~τd~στ+d~σ)(DRD)+(c~τc~στ+c~σ)(DSD),\displaystyle+(\tilde{d}_{\tau}-\tilde{d}_{\sigma\tau}+\tilde{d}_{\sigma})(D_{R}-D)+(\tilde{c}_{\tau}-\tilde{c}_{\sigma\tau}+\tilde{c}_{\sigma})(D_{S}-D),

for σ,τGK\sigma,\tau\in G_{K}.

It can be checked that, for xK/(K)2x\in K^{*}/(K^{*})^{2} and σ,τGK\sigma,\tau\in G_{K}, we get x~τx~στ+x~σ=2\tilde{x}_{\tau}-\tilde{x}_{\sigma\tau}+\tilde{x}_{\sigma}=2 if both σ\sigma and τ\tau flip x\sqrt{x} and otherwise it is equal to zero. Define ισ,τ,x=1\iota_{\sigma,\tau,x}=1 if both σ\sigma and τ\tau flip x\sqrt{x} and otherwise ισ,τ,x=0\iota_{\sigma,\tau,x}=0. Note that the map that sends xK/(K)2x\in K^{*}/(K^{*})^{2} to the class of (σ,τ)ισ,τ,x(\sigma,\tau)\mapsto\iota_{\sigma,\tau,x} explicitly realizes the map K/(K)2H1(GK,12/)H1(GK,/)H2(GK,)K^{*}/(K^{*})^{2}\cong H^{1}(G_{K},\frac{1}{2}\mathbb{Z}/\mathbb{Z})\subset H^{1}(G_{K},\mathbb{Q}/\mathbb{Z})\rightarrow H^{2}(G_{K},\mathbb{Z}). Then, for σ,τGK\sigma,\tau\in G_{K}, the cocycle in the last paragraph sends (σ,τ)(\sigma,\tau) to

ισ,τ,b2(DPD)+ισ,τ,a2(DQD)+ισ,τ,d2(DRD)+ισ,τ,c2(DSD).\iota_{\sigma,\tau,b}\cdot 2(D_{P}-D)+\iota_{\sigma,\tau,a}\cdot 2(D_{Q}-D)+\iota_{\sigma,\tau,d}\cdot 2(D_{R}-D)+\iota_{\sigma,\tau,c}\cdot 2(D_{S}-D).

Hence, by Remark 3.2, the corresponding element in H2(GK,K¯(Jϵ)/K¯)H^{2}(G_{K},\bar{K}(J_{\epsilon})^{*}/\bar{K}^{*}) is represented by

(σ,τ)[fPισ,τ,bfQισ,τ,afRισ,τ,dfSισ,τ,c],(\sigma,\tau)\mapsto[f_{P}^{\iota_{\sigma,\tau,b}}\cdot f_{Q}^{\iota_{\sigma,\tau,a}}\cdot f_{R}^{\iota_{\sigma,\tau,d}}\cdot f_{S}^{\iota_{\sigma,\tau,c}}],

for all σ,τGK\sigma,\tau\in G_{K}.

For each place vv of KK, following the homogeneous space definition of ϵ,ηCT\langle\epsilon,\eta\rangle_{CT} as given in Section 2.5, we obtain an element in H2(GKv,Kv¯)H^{2}(G_{K_{v}},\bar{K_{v}^{*}}) from the long exact sequence induced by the short exact sequence 0Kv¯Kv¯(Jϵ)Kv¯(Jϵ)/Kv¯00\rightarrow\bar{K_{v}^{*}}\rightarrow\bar{K_{v}}(J_{\epsilon})^{*}\rightarrow\bar{K_{v}}(J_{\epsilon})^{*}/\bar{K_{v}}^{*}\rightarrow 0. The long exact sequence is the local version of (2.4) with XX replaced by JϵJ_{\epsilon}. By Remark 2.8(ii), this element in H2(GKv,Kv¯)H^{2}(G_{K_{v}},\bar{K_{v}}^{*}) can be represented by

(σ,τ)fP(Pv)ισ,τ,bfQ(Pv)ισ,τ,afR(Pv)ισ,τ,dfS(Pv)ισ,τ,c,(\sigma,\tau)\mapsto f_{P}(P_{v})^{\iota_{\sigma,\tau,b}}\cdot f_{Q}(P_{v})^{\iota_{\sigma,\tau,a}}\cdot f_{R}(P_{v})^{\iota_{\sigma,\tau,d}}\cdot f_{S}(P_{v})^{\iota_{\sigma,\tau,c}},

for all σ,τGK\sigma,\tau\in G_{K} and some local point PvJϵ(Kv)P_{v}\in J_{\epsilon}(K_{v}) avoiding the zeros and poles of fP,fQ,fR,fSf_{P},f_{Q},f_{R},f_{S}.

Hence, the above element in Br(Kv)H2(GKv,Kv¯)\text{Br}(K_{v})\cong H^{2}(G_{K_{v}},\bar{K_{v}}^{*}) is the class of the tensor product of quaternion algebras

(fP(Pv),b)+(fQ(Pv),a)+(fR(Pv),d)+(fS(Pv),c).(f_{P}(P_{v}),b)+(f_{Q}(P_{v}),a)+(f_{R}(P_{v}),d)+(f_{S}(P_{v}),c).

Then, we have that

inv ((fP(Pv),b)+(fQ(Pv),a)+(fR(Pv),d)+(fS(Pv),c))\displaystyle\big{(}(f_{P}(P_{v}),b)+(f_{Q}(P_{v}),a)+(f_{R}(P_{v}),d)+(f_{S}(P_{v}),c)\big{)}
=(fP(Pv,b)v(fQ(Pv),a)v(fR(Pv),d)v(fS(Pv),c)v,\displaystyle=(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: Kv×Kv{1,1}K_{v}^{*}\times K_{v}^{*}\rightarrow\{1,-1\}, as required.

Remark 3.4.

In Section 6, we will directly show that the formula for the Cassels-Tate pairing on Sel2(J)×Sel2(J)\text{Sel}^{2}(J)\times\text{Sel}^{2}(J) given in Theorem 3.3 is a finite product.

4 Explicit Computation

In this section, we explain how we explicitly compute the Cassels-Tate pairing on Sel2(J)×Sel2(J)\text{Sel}^{2}(J)\times\text{Sel}^{2}(J) using the formula given in Theorem 3.3, under the assumption that all points in J[2]J[2] are defined over KK. We fix ϵSel2(J)\epsilon\in\text{Sel}^{2}(J) and (Jϵ,[2]ϕϵ)(J_{\epsilon},[2]\circ\phi_{\epsilon}), the 2-covering of JJ corresponding to ϵ\epsilon with ϕϵ:Jϵ15J15\phi_{\epsilon}:J_{\epsilon}\subset\mathbb{P}^{15}\rightarrow J\subset\mathbb{P}^{15} given in Theorem 2.9. The statement of Theorem 3.3 suggests that we need to compute the KK-rational divisors D,DP,DQ,DR,DSD,D_{P},D_{Q},D_{R},D_{S} on JϵJ_{\epsilon} and the KK-rational function fP,fQ,fR,fSf_{P},f_{Q},f_{R},f_{S} on JϵJ_{\epsilon}, as in Remark 3.2.

4.1 Computing the twist of the Kummer surface

We describe a practical method for computing a linear isomorphism ψϵ:𝒦ϵ3𝒦3\psi_{\epsilon}:\mathcal{K}_{\epsilon}\subset\mathbb{P}^{3}\rightarrow\mathcal{K}\subset\mathbb{P}^{3} corresponding to ϵ\epsilon. More explicitly, we need to compute ψϵ\psi_{\epsilon} such that ψϵ(ψϵ1)σ\psi_{\epsilon}(\psi_{\epsilon}^{-1})^{\sigma} is the action of translation by ϵσJ[2]\epsilon_{\sigma}\in J[2] on 𝒦\mathcal{K} and (σϵσ)(\sigma\mapsto\epsilon_{\sigma}) is a cocycle representing ϵ\epsilon. Since all points in J[2]J[2] are defined over KK, the coboundaries in B1(GK,J[2])B^{1}(G_{K},J[2]) are trivial. Therefore, these conditions determine ψϵ\psi_{\epsilon} uniquely up to a change of linear automorphism of 𝒦ϵ3\mathcal{K}_{\epsilon}\subset\mathbb{P}^{3} over KK.

For each TJ[2]T\in J[2], we have an explicit formula for MTGL4(K)M_{T}\in\text{GL}_{4}(K), given in [thebook, Chapter 3 Section 2], representing the action of translation by TJ[2]T\in J[2] on the Kummer surface 𝒦3\mathcal{K}\subset\mathbb{P}^{3}. It can be checked that they form a basis of Mat4(K)\text{Mat}_{4}(K) and we suppose MP2=cPI,MQ2=cQI,MR2=cRI,andMS2=cSI.M_{P}^{2}=c_{P}I,M_{Q}^{2}=c_{Q}I,M_{R}^{2}=c_{R}I,\;\text{and}\;M_{S}^{2}=c_{S}I. The explicit formulae for cP,cQ,cR,cSc_{P},c_{Q},c_{R},c_{S} can also be found in [thebook, Chapter 3 Section 2]. Moreover, by [thebook, Chapter 3 Section 3] and the Weil pairing relationship among the generators P,Q,R,SP,Q,R,S of J[2]J[2] specified by (3.1), we know that [MP,MQ]=[MR,MS]=I[M_{P},M_{Q}]=[M_{R},M_{S}]=-I and the commutators of the other pairs are trivial.

Suppose (a,b,c,d)(K/(K)2)4(a,b,c,d)\in(K^{*}/(K^{*})^{2})^{4} represents ϵ\epsilon. Let AGL4(K¯)A\in\text{GL}_{4}(\bar{K}) represent the linear isomorphism ψϵ\psi_{\epsilon} and let MT=A1MTAGL4(K¯)M_{T}^{\prime}=A^{-1}M_{T}A\in\text{GL}_{4}(\bar{K}) represent the action of TT on the twisted Kummer 𝒦ϵ\mathcal{K}_{\epsilon}. It can be checked, see [thesis, Lemma 3.2.1] for details, that the set of matrices in PGL4(K¯)\text{PGL}_{4}(\bar{K}) that commute with MTM_{T} in PGL4(K¯)\text{PGL}_{4}(\bar{K}) for all TJ[2]T\in J[2] is {[MT],TJ[2]}\{[M_{T}],T\in J[2]\}. This implies that any BGL4(K¯)B\in\text{GL}_{4}(\bar{K}) such that [MT]=[B1MTB]PGL4(K¯)[M_{T}]^{\prime}=[B^{-1}M_{T}B]\in\text{PGL}_{4}(\bar{K}) for any TJ[2]T\in J[2] is equal to a multiple of MTM_{T} composed with AA and so can also represents ψϵ\psi_{\epsilon}. Hence, it will suffice to compute the matrices MTM_{T}^{\prime}.

Consider [MT]PGL4(K¯)[M_{T}^{\prime}]\in\text{PGL}_{4}(\bar{K}) and σGK\sigma\in G_{K}. We have

[MT]([MT]1)σ=[A1MTA(A1)σMT1Aσ]PGL4(K¯).[M_{T}^{\prime}]([M_{T}^{\prime}]^{-1})^{\sigma}=[A^{-1}M_{T}A(A^{-1})^{\sigma}M_{T}^{-1}A^{\sigma}]\in\text{PGL}_{4}(\bar{K}).

Recall that for each element xK/(K)2x\in K^{*}/(K^{*})^{2}, we define x~σ{0,1}\tilde{x}_{\sigma}\in\{0,1\} such that (1)x~σ=σ(x)/x(-1)^{\tilde{x}_{\sigma}}=\sigma(\sqrt{x})/\sqrt{x}. Since [A(A1)σ]=[MPb~σMQa~σMRd~σMSc~σ][A(A^{-1})^{\sigma}]=[M_{P}^{\tilde{b}_{\sigma}}M_{Q}^{\tilde{a}_{\sigma}}M_{R}^{\tilde{d}_{\sigma}}M_{S}^{\tilde{c}_{\sigma}}], we have [MT][M_{T}^{\prime}] is in PGL4(K)\text{PGL}_{4}(K). This means that we can redefine MT=λTA1MTAM_{T}^{\prime}=\lambda_{T}A^{-1}M_{T}A for some λTK¯\lambda_{T}\in\bar{K} such that MTGL4(K)M_{T}^{\prime}\in\text{GL}_{4}(K) by Hilbert Theorem 90.

Let NP=1/cPMP,NQ=1/cQMQ,NR=1/cRMR,NS=1/cSMSN_{P}=1/\sqrt{c_{P}}M_{P},N_{Q}=1/\sqrt{c_{Q}}M_{Q},N_{R}=1/\sqrt{c_{R}}M_{R},N_{S}=1/\sqrt{c_{S}}M_{S}. Then NT2=IN_{T}^{2}=I for T=P,Q,R,ST=P,Q,R,S. Define NT=A1NTAGL4(K¯)N_{T}^{\prime}=A^{-1}N_{T}A\in\text{GL}_{4}(\bar{K}) for T=P,Q,R,ST=P,Q,R,S. We note that NT,MTN_{T}^{\prime},M_{T}^{\prime} represent the same element in PGL4(K)\text{PGL}_{4}(K) and NT2=IN_{T}^{\prime 2}=I for each T=P,Q,R,ST=P,Q,R,S. Suppose MP2=αPI,MQ2=αQI,MR2=αRI,MS2=αSI.M_{P}^{\prime 2}=\alpha_{P}I,M_{Q}^{\prime 2}=\alpha_{Q}I,M_{R}^{\prime 2}=\alpha_{R}I,M_{S}^{\prime 2}=\alpha_{S}I. Then NT=1/αTMTN_{T}^{\prime}=1/\sqrt{\alpha_{T}}M_{T}^{\prime} for each T=P,Q,R,ST=P,Q,R,S. Note that there might be some sign issues here but they will not affect the later computation. Since

NP(NP1)σ=A1NPA(A1)σ(NP1)σAσ,N_{P}^{\prime}(N_{P}^{\prime-1})^{\sigma}=A^{-1}N_{P}A(A^{-1})^{\sigma}(N_{P}^{-1})^{\sigma}A^{\sigma},

using NP=1/cPMPN_{P}=1/\sqrt{c_{P}}M_{P} and NP=1/αPMPN_{P}^{\prime}=1/\sqrt{\alpha_{P}}M_{P}^{\prime} with MP,MPGL4(K)M_{P},M_{P}^{\prime}\in\text{GL}_{4}(K), we compute that

σ(αP)αP=σ(a)aσ(cP)cP,\frac{\sigma(\sqrt{\alpha_{P}})}{\sqrt{\alpha_{P}}}=\frac{\sigma(\sqrt{a})}{\sqrt{a}}\frac{\sigma(\sqrt{c_{P}})}{\sqrt{c_{P}}},

and similar equations for Q,R,S.Q,R,S.

This implies that αP=cPa\alpha_{P}=c_{P}a up to squares in KK and so via rescaling MPM_{P}^{\prime} by elements in KK, we have MP2=cPaIM_{P}^{\prime 2}=c_{P}aI. Similarly, MQ2=cQbI,MR2=cRcI,MS2=cSdIM_{Q}^{\prime 2}=c_{Q}bI,M_{R}^{\prime 2}=c_{R}cI,M_{S}^{\prime 2}=c_{S}dI. We note that we also have [MP,MQ]=[MR,MS]=I[M_{P}^{\prime},M_{Q}^{\prime}]=[M_{R}^{\prime},M_{S}^{\prime}]=-I and the commutators of the other pairs are trivial. This implies that

Mat4(K)(cPa,cQb)(cRc,cSd)\text{Mat}_{4}(K)\cong(c_{P}a,c_{Q}b)\otimes(c_{R}c,c_{S}d)
MPi11,MQj11,MR1i2,MS1j2,M_{P}^{\prime}\mapsto i_{1}\otimes 1,M_{Q}^{\prime}\mapsto j_{1}\otimes 1,M_{R}^{\prime}\mapsto 1\otimes i_{2},M_{S}^{\prime}\mapsto 1\otimes j_{2},

where (cPa,cQb)(c_{P}a,c_{Q}b) and (cRc,cSd)(c_{R}c,c_{S}d) are quaternion algebras with generators i1,j1i_{1},j_{1} and i2,j2i_{2},j_{2} respectively. In Section5, we will interpret this isomorphism as saying that the image of ϵ\epsilon via the obstruction map is trivial.

Let A=(cPa,cQb),B=(cRc,cSd)A=(c_{P}a,c_{Q}b),B=(c_{R}c,c_{S}d). By the argument above, we know ABA\otimes B represents the trivial element in Br(K)\text{Br}(K) and an explicit isomorphism ABMat4(K)A\otimes B\cong\text{Mat}_{4}(K) will give us the explicit matrices MP,MQ,MR,MSM_{P}^{\prime},M_{Q}^{\prime},M_{R}^{\prime},M_{S}^{\prime} we seek. Since the classes of A,BA,B are in Br[2]\text{Br}[2], we have A,BA,B representing the same element in Br(K)\text{Br}(K). This implies that ABA\cong B over KK, by Wedderburn’s Theorem. We have the following lemma.

Lemma 4.1.

Consider a tensor product of two quaternion algebras ABA\otimes B, where A=(α,β),B=(γ,δ),A=(\alpha,\beta),\;B=(\gamma,\delta), with generators i1,j1i_{1},j_{1} and i2,j2i_{2},j_{2} respectively. Suppose there is an isomorphism ψ:BA\psi:B\xrightarrow{\sim}A given by

i2a11+b1i1+c1j1+d1i1j1,\displaystyle i_{2}\mapsto a_{1}\cdot 1+b_{1}\cdot i_{1}+c_{1}\cdot j_{1}+d_{1}\cdot i_{1}j_{1},
j2a21+b2i1+c2j1+d2i1j1.\displaystyle j_{2}\mapsto a_{2}\cdot 1+b_{2}\cdot i_{1}+c_{2}\cdot j_{1}+d_{2}\cdot i_{1}j_{1}.

Then there is an explicit isomorphism

ABMat4(K)A\otimes B\cong\text{Mat}_{4}(K)

given by

i11Mi1:=[0α001000000α0010]i_{1}\otimes 1\mapsto M_{i_{1}}:=\begin{bmatrix}0&\alpha&0&0\\ 1&0&0&0\\ 0&0&0&\alpha\\ 0&0&1&0\\ \end{bmatrix}
j11Mj1:=[00β0000β10000100]j_{1}\otimes 1\mapsto M_{j_{1}}:=\begin{bmatrix}0&0&\beta&0\\ 0&0&0&-\beta\\ 1&0&0&0\\ 0&-1&0&0\\ \end{bmatrix}
1i2Mi2:=[a1b1αc1βd1αβb1a1d1βc1βc1d1αa1b1αd1c1b1a1]1\otimes i_{2}\mapsto M_{i_{2}}:=\begin{bmatrix}a_{1}&b_{1}\cdot\alpha&c_{1}\cdot\beta&-d_{1}\cdot\alpha\beta\\ b_{1}&a_{1}&-d_{1}\cdot\beta&c_{1}\cdot\beta\\ c_{1}&d_{1}\cdot\alpha&a_{1}&-b_{1}\cdot\alpha\\ d_{1}&c_{1}&-b_{1}&a_{1}\\ \end{bmatrix}
1j2Mj2:=[a2b2αc2βd2αβb2a2d2βc2βc2d2αa2b2αd2c2b2a2]1\otimes j_{2}\mapsto M_{j_{2}}:=\begin{bmatrix}a_{2}&b_{2}\cdot\alpha&c_{2}\cdot\beta&-d_{2}\cdot\alpha\beta\\ b_{2}&a_{2}&-d_{2}\cdot\beta&c_{2}\cdot\beta\\ c_{2}&d_{2}\cdot\alpha&a_{2}&-b_{2}\cdot\alpha\\ d_{2}&c_{2}&-b_{2}&a_{2}\\ \end{bmatrix}

Proof.

We have that AAopA\otimes A^{op} is isomorphic to a matrix algebra. More specifically, AAopEndK(A)A\otimes A^{op}\cong\text{End}_{K}(A) via uv(xuxv)u\otimes v\mapsto(x\mapsto uxv), which makes AAopMat4(K)A\otimes A^{op}\cong\text{Mat}_{4}(K) after picking a basis for AA. Hence,

ABopMat4(K)uv(xuxψ(v)).\begin{array}[]{ccc}A\otimes B^{op}&\cong&\text{Mat}_{4}(K)\\ u\otimes v&\mapsto&(x\mapsto ux\psi(v)).\\ \end{array}\;

More explicitly, fixing the basis of AA to be {1,i1,j1,i1j1}\{1,i_{1},j_{1},i_{1}j_{1}\}, the isomorphism is as given in the statement of the lemma.

By taking A=(cPa,cQb),B=(cRc,cSd)A=(c_{P}a,c_{Q}b),B=(c_{R}c,c_{S}d) in Lemma 4.1, we know that the matrices MP,MQ,MS,MRM_{P}^{\prime},M_{Q}^{\prime},M_{S}^{\prime},M_{R}^{\prime} and Mi1,Mj1,Mi2,Mj2M_{i_{1}},M_{j_{1}},M_{i_{2}},M_{j_{2}} are equal up to conjugation by a matrix CGL4(K)C\in\text{GL}_{4}(K) via the Noether Skolem Theorem. After a change of coordinates for 𝒦ϵ3\mathcal{K}_{\epsilon}\subset\mathbb{P}^{3} according to CC, we have that MP,MQ,MS,MRM_{P}^{\prime},M_{Q}^{\prime},M_{S}^{\prime},M_{R}^{\prime} are equal to Mi1,Mj1,Mi2,Mj2M_{i_{1}},M_{j_{1}},M_{i_{2}},M_{j_{2}}. Lemma 4.1 therefore reduces the problem of computing the MTM_{T}^{\prime} to that of computing an isomorphism between two quaternion algebras. See [thesis, Corollary 4.2.3 ] for the description of an explicit algorithm. Finally we solve for a matrix AA such that MT=λTA1MTAM_{T}^{\prime}=\lambda_{T}A^{-1}M_{T}A for some λTK¯\lambda_{T}\in\bar{K} and T=P,Q,R,ST=P,Q,R,S by linear algebra.

4.2 Explicit computation of DD

In this section, we explain a method for computing the KK-rational divisor DD on JϵJ_{\epsilon} representing the divisor class ϕϵ(2Θ)\phi_{\epsilon}^{*}(2\Theta). The idea is to compute it via the commutative diagram (2.3) in Remark 2.6.

By Theorem 2.9, there is an explicit isomorphism Jϵ15ϕϵJ15J_{\epsilon}\subset\mathbb{P}^{15}\xrightarrow{\phi_{\epsilon}}J\subset\mathbb{P}^{15}. We write u0,,u9,v1,,v6u_{0},...,u_{9},v_{1},...,v_{6} for the coordinates on the ambient space of Jϵ15J_{\epsilon}\subset\mathbb{P}^{15} and write k11,k12,,k44,b1,,b6k_{11},k_{12},...,k_{44},b_{1},...,b_{6} for the coordinates on the ambient space of J15J\subset\mathbb{P}^{15}. By the same theorem, ϕϵ\phi_{\epsilon} is represented by a block diagonal matrix consisting of a block of size 10 corresponding to the even basis elements and a block of size 6 corresponding to the odd basis elements. Following Section 4.1, we can compute an explicit isomorphism ψϵ:𝒦ϵ3𝒦3\psi_{\epsilon}:\mathcal{K}_{\epsilon}\subset\mathbb{P}^{3}\rightarrow\mathcal{K}\subset\mathbb{P}^{3} corresponding to ϵ\epsilon. We write k1,,k4k_{1}^{\prime},...,k_{4}^{\prime} for the coordinates on the ambient space of 𝒦ϵ3\mathcal{K}_{\epsilon}\subset\mathbb{P}^{3}. Recall that since all points in J[2]J[2] are defined over KK, all coboundaries in B1(GK,J[2])B^{1}(G_{K},J[2]) are trivial. So, we have that ϕϵ(ϕϵ1)σ\phi_{\epsilon}(\phi_{\epsilon}^{-1})^{\sigma} and ψϵ(ψϵ1)σ\psi_{\epsilon}(\psi_{\epsilon}^{-1})^{\sigma} both give the action of translation by some ϵσJ[2]\epsilon_{\sigma}\in J[2] such that (σϵσ)(\sigma\mapsto\epsilon_{\sigma}) represents ϵSel2(J)\epsilon\in\text{Sel}^{2}(J).

Define kij=kikjk_{ij}^{\prime}=k_{i}^{\prime}k_{j}^{\prime}. The isomorphism ψϵ:𝒦ϵki3𝒦ki3\psi_{\epsilon}:\mathcal{K}_{\epsilon}\subset\mathbb{P}_{k_{i}^{\prime}}^{3}\rightarrow\mathcal{K}\subset\mathbb{P}_{k_{i}}^{3}, induces a natural isomorphism ψϵ~:kij9kij9\tilde{\psi_{\epsilon}}:\mathbb{P}^{9}_{k_{ij}^{\prime}}\rightarrow\mathbb{P}^{9}_{k_{ij}}. More explicitly, suppose ψϵ\psi_{\epsilon} is represented by the 4×44\times 4 matrix AA where (k1:,k4)(i=14A1iki::i=14A4iki)(k_{1}^{\prime}:...,k_{4}^{\prime})\mapsto(\sum_{i=1}^{4}A_{1i}k_{i}^{\prime}:...:\sum_{i=1}^{4}A_{4i}k_{i}^{\prime}). Then ψϵ~:kij9kij9\tilde{\psi_{\epsilon}}:\mathbb{P}^{9}_{k_{ij}^{\prime}}\rightarrow\mathbb{P}^{9}_{k_{ij}} is given by (k11:k12::k44)(i,j=14A1iA1jkij:i,j=14A1iA2jkij::i,j=14A4iA4jkij)(k_{11}^{\prime}:k_{12}^{\prime}:...:k_{44}^{\prime})\mapsto(\sum_{i,j=1}^{4}A_{1i}A_{1j}k_{ij}^{\prime}:\sum_{i,j=1}^{4}A_{1i}A_{2j}k_{ij}^{\prime}:...:\sum_{i,j=1}^{4}A_{4i}A_{4j}k_{ij}^{\prime}).

On the other hand, the isomorphism ϕϵ:Jϵ{ui,vi}15J{kij,bi}15\phi_{\epsilon}:J_{\epsilon}\subset\mathbb{P}^{15}_{\{u_{i},v_{i}\}}\rightarrow J\subset\mathbb{P}^{15}_{\{k_{ij},b_{i}\}} induces a natural isomorphism ϕϵ~:ui9kij9\tilde{\phi_{\epsilon}}:\mathbb{P}^{9}_{u_{i}}\rightarrow\mathbb{P}^{9}_{k_{ij}} represented by the 10×1010\times 10 block of the matrix representing ϕϵ\phi_{\epsilon}. Since ϕϵ(ϕϵ1)σ\phi_{\epsilon}(\phi_{\epsilon}^{-1})^{\sigma} and ψϵ(ψϵ1)σ\psi_{\epsilon}(\psi_{\epsilon}^{-1})^{\sigma} both give the action of translation by some ϵσJ[2]\epsilon_{\sigma}\in J[2], we get ϕϵ~(ϕϵ~1)σ=ψϵ~(ϕϵ~1)σ\tilde{\phi_{\epsilon}}(\tilde{\phi_{\epsilon}}^{-1})^{\sigma}=\tilde{\psi_{\epsilon}}(\tilde{\phi_{\epsilon}}^{-1})^{\sigma}. Therefore, ψϵ~1ϕϵ~\tilde{\psi_{\epsilon}}^{-1}\tilde{\phi_{\epsilon}} is defined over KK and we obtain the following commutative diagram that decomposes the standard commutative diagram (2.3):

Jϵ{ui,vi}15{J_{\epsilon}\subset\mathbb{P}^{15}_{\{u_{i},v_{i}\}}}ui9{\mathbb{P}^{9}_{u_{i}}}kij9{\mathbb{P}^{9}_{k_{ij}^{\prime}}}𝒦ϵki3{\mathcal{K}_{\epsilon}\subset\mathbb{P}^{3}_{k_{i}^{\prime}}}J{kij,bi}15{J\subset\mathbb{P}^{15}_{\{k_{ij},b_{i}\}}}kij9{\mathbb{P}^{9}_{k_{ij}}}𝒦ki3,{\mathcal{K}\subset\mathbb{P}^{3}_{k_{i}},}proj\scriptstyle{proj}ϕϵ\scriptstyle{\phi_{\epsilon}}(ψϵ~)1ϕϵ~\scriptstyle{(\tilde{\psi_{\epsilon}})^{-1}\tilde{\phi_{\epsilon}}}ϕϵ~\scriptstyle{\tilde{\phi_{\epsilon}}}ψϵ~\scriptstyle{\tilde{\psi_{\epsilon}}}g2\scriptstyle{g_{2}}ψϵ\scriptstyle{\psi_{\epsilon}}proj\scriptstyle{proj}g1\scriptstyle{g_{1}} (4.1)

where g1:(k11::k44)(k1::k4)g_{1}:(k_{11}:...:k_{44})\rightarrow(k_{1}:...:k_{4}) and g2:(k11::k44)(k1::k4)g_{2}:(k^{\prime}_{11}:...:k^{\prime}_{44})\rightarrow(k^{\prime}_{1}:...:k^{\prime}_{4}) are the projection maps. The composition of the morphisms on the bottom row gives the standard morphism J|2Θ|𝒦3J\xrightarrow{|2\Theta|}\mathcal{K}\subset\mathbb{P}^{3} and the composition of the morphisms on the top row gives Jϵ|ϕϵ(2Θ)|𝒦ϵ3J_{\epsilon}\xrightarrow{|\phi_{\epsilon}^{*}(2\Theta)|}\mathcal{K}_{\epsilon}\subset\mathbb{P}^{3}.

Let DD be the pull back on JϵJ_{\epsilon} via Jϵ{ui,vi}15projui9(ψϵ~)1ϕϵ~kij9projki3J_{\epsilon}\subset\mathbb{P}^{15}_{\{u_{i},v_{i}\}}\xrightarrow{proj}\mathbb{P}^{9}_{u_{i}}\xrightarrow{(\tilde{\psi_{\epsilon}})^{-1}\tilde{\phi_{\epsilon}}}\mathbb{P}^{9}_{k_{ij}^{\prime}}\xrightarrow{proj}\mathbb{P}^{3}_{k_{i}^{\prime}} of the hyperplane section given by k1=0k_{1}^{\prime}=0. This implies that DD is a KK-rational divisor on JϵJ_{\epsilon} representing the class of ϕϵ(2Θ)\phi_{\epsilon}^{*}(2\Theta). Moreover, the pull back on JϵJ_{\epsilon} via Jϵ{ui,vi}15projui9(ψϵ~)1ϕϵ~kij9J_{\epsilon}\subset\mathbb{P}^{15}_{\{u_{i},v_{i}\}}\xrightarrow{proj}\mathbb{P}^{9}_{u_{i}}\xrightarrow{(\tilde{\psi_{\epsilon}})^{-1}\tilde{\phi_{\epsilon}}}\mathbb{P}^{9}_{k_{ij}^{\prime}} of the hyperplane section given by k11=0k_{11}^{\prime}=0 is 2D2D.

4.3 Explicit computation of DP,DQ,DR,DSD_{P},D_{Q},D_{R},D_{S}

In this section, we explain how to compute the KK-rational divisors DP,DQ,DR,DSD_{P},D_{Q},D_{R},D_{S} defined in Remark 3.2. More explicitly, for TJ[2]T\in J[2], we give a method for computing a KK-rational divisor DTD_{T} on JϵJ_{\epsilon} representing the divisor class of ϕϵ(τT1(2Θ))\phi_{\epsilon}^{*}(\tau_{T_{1}}^{*}(2\Theta)) for some T1T_{1} on JJ such that 2T1=T2T_{1}=T. Recall that we assume all points in J[2]J[2] are defined over KK and we have an explicit isomorphism ϕϵ:JϵJ\phi_{\epsilon}:J_{\epsilon}\rightarrow J such that (Jϵ,[2]ϕϵ)(J_{\epsilon},[2]\circ\phi_{\epsilon}) is the 2-covering of JJ corresponding to ϵSel2(J)\epsilon\in\text{Sel}^{2}(J). Recall δ:J(K)H1(GK,J[2])\delta:J(K)\rightarrow H^{1}(G_{K},J[2]) in (2.1). We first prove the following lemma.

Lemma 4.2.

Let TJ(K)T\in J(K). Suppose ϕϵ+δ(T):Jϵ+δ(T)J\phi_{\epsilon+\delta(T)}:J_{\epsilon+\delta(T)}\rightarrow J is an isomorphism and (Jϵ+δ(T),[2]ϕϵ+δ(T))(J_{\epsilon+\delta(T)},[2]\circ\phi_{\epsilon+\delta(T)}) is the 2-covering of JJ corresponding to ϵ+δ(T)H1(GK,J[2])\epsilon+\delta(T)\in H^{1}(G_{K},J[2]). Let T1JT_{1}\in J such that 2T1=T2T_{1}=T. Then, ϕϵ+δ(T)1τT1ϕϵ:JϵJϵ+δ(T)\phi_{\epsilon+\delta(T)}^{-1}\circ\tau_{T_{1}}\circ\phi_{\epsilon}:J_{\epsilon}\rightarrow J_{\epsilon+\delta(T)} is defined over KK.

Proof.

Using the same argument as in the proof of Lemma 3.1(i), we know that (Jϵ,[2]τT1ϕϵ)(J_{\epsilon},[2]\circ\tau_{T_{1}}\circ\phi_{\epsilon}) is the 2-covering of JJ corresponding to ϵ+δ(T)H1(GK,J[2])\epsilon+\delta(T)\in H^{1}(G_{K},J[2]). Since all points in J[2]J[2] are defined over KK, we have τT1ϕϵ((τT1ϕϵ)1)σ=ϕϵ+δ(T)(ϕϵ+δ(T)1)σ\tau_{T_{1}}\circ\phi_{\epsilon}\circ((\tau_{T_{1}}\circ\phi_{\epsilon})^{-1})^{\sigma}=\phi_{\epsilon+\delta(T)}\circ(\phi_{\epsilon+\delta(T)}^{-1})^{\sigma}, as required.

Let TJ(K)T\in J(K) with 2T1=T2T_{1}=T. Consider the commutative diagram below which is formed by two copies of the standard diagram (2.3). Note that all the horizontal maps are defined over KK as is the composition of the vertical map on the left by Lemma 4.2. Hence, the composition of the thick arrows is defined over KK. Then the pull back on JϵJ_{\epsilon} via the thick arrows of a hyperplane section on 𝒦ϵ+δ(T)3\mathcal{K}_{\epsilon+\delta(T)}\subset\mathbb{P}^{3} is a KK-rational divisor DTD_{T} on JϵJ_{\epsilon} representing the divisor class ϕϵ(τT1(2Θ))\phi_{\epsilon}^{*}(\tau_{T_{1}}^{*}(2\Theta)). We note that in the case where TJ[2]T\in J[2], the composition of the vertical maps on the left hand side of the the diagram below is in fact given by a 16×1616\times 16 matrix defined over KK even though the individual maps are not defined over KK.

Jϵ15{J_{\epsilon}\subset\mathbb{P}^{15}}𝒦ϵ3{\mathcal{K}_{\epsilon}\subset\mathbb{P}^{3}}J15{J\subset\mathbb{P}^{15}}𝒦3{\mathcal{K}\subset\mathbb{P}^{3}}J15{J\subset\mathbb{P}^{15}}𝒦3{\mathcal{K}\subset\mathbb{P}^{3}}Jϵ+δ(T)15{J_{\epsilon+\delta(T)}\subset\mathbb{P}^{15}}𝒦ϵ+δ(T)3.{\mathcal{K}_{\epsilon+\delta(T)}\subset\mathbb{P}^{3}.}|ϕϵ(2Θ)|\scriptstyle{|\phi_{\epsilon}^{*}(2\Theta)|}ϕϵ\scriptstyle{\phi_{\epsilon}}ψϵ\scriptstyle{\psi_{\epsilon}}|2Θ|\scriptstyle{|2\Theta|}τT1\scriptstyle{\tau_{T_{1}}}|2Θ|\scriptstyle{|2\Theta|}ϕϵ+δ(T)1\scriptstyle{\phi_{\epsilon+\delta(T)}^{-1}}|ϕϵ+δ(T)(2Θ)|\scriptstyle{|\phi^{*}_{\epsilon+\delta(T)}(2\Theta)|}ψϵ+δ(T)\scriptstyle{\psi_{\epsilon+\delta(T)}} (4.2)

The bottom horizontal morphism Jϵ+δ(T)|ϕϵ+δ(T)(2Θ)|𝒦ϵ+δ(T)3J_{\epsilon+\delta(T)}\xrightarrow{|\phi^{*}_{\epsilon+\delta(T)}(2\Theta)|}\mathcal{K}_{\epsilon+\delta(T)}\subset\mathbb{P}^{3} can be explicitly computed using the algorithm in Section 4.2 with the Selmer element ϵ\epsilon replaced by ϵ+δ(T)\epsilon+\delta(T). Also, by Theorem 2.9, we have explicit formulae for ϕϵ\phi_{\epsilon} and ϕϵ+δ(T).\phi_{\epsilon+\delta(T)}. Hence, to explicitly compute DTD_{T}, we need to find a way to deal with τT1\tau_{T_{1}}, for some T1T_{1} such that 2T1=T2T_{1}=T.

Since we need to apply the above argument to T=P,Q,R,ST=P,Q,R,S, the basis for J[2]J[2], it would suffice to compute the translation maps τT1:J15J15\tau_{T_{1}}:J\subset\mathbb{P}^{15}\rightarrow J\subset\mathbb{P}^{15} when T1J[4]T_{1}\in J[4]. This map is given by a 16×1616\times 16 matrix. We show how to compute a 10×1610\times 16 matrix in the following proposition. We then explain below why this is sufficient for our purposes.

Proposition 4.3.

Suppose T1J[4]T_{1}\in J[4]. Given the coordinates of T1J{kij,bi}15T_{1}\in J\subset\mathbb{P}^{15}_{\{k_{ij},b_{i}\}}, we can compute the following composition of morphisms:

Ψ:J{kij,bi}15τT1J{kij,bi}15projkij9.\Psi:J\subset\mathbb{P}^{15}_{\{k_{ij},b_{i}\}}\xrightarrow{\tau_{T_{1}}}J\subset\mathbb{P}^{15}_{\{k_{ij},b_{i}\}}\xrightarrow{proj}\mathbb{P}^{9}_{k_{ij}}.

Proof.

Let T=2T1J[2]T=2T_{1}\in J[2]. Recall that we let MTM_{T} denote the action of translation by TT on 𝒦3\mathcal{K}\subset\mathbb{P}^{3}. Then for any PJP\in J, we have ki(P+T)=j=14(MT)ijkj(P)k_{i}(P+T)=\sum_{j=1}^{4}(M_{T})_{ij}k_{j}(P) projectively as a vector of length 4, and projectively as a vector of length 10, kij(P+T1)k_{ij}(P+T_{1}) is equal to

ki(P+T1)kj(P+T1)=ki(P+T1)l=14(MT)jlkl(PT1)=l=14(MT)jlkl(PT1)ki(P+T1).k_{i}(P+T_{1})k_{j}(P+T_{1})=k_{i}(P+T_{1})\sum_{l=1}^{4}(M_{T})_{jl}k_{l}(P-T_{1})=\sum_{l=1}^{4}(M_{T})_{jl}k_{l}(P-T_{1})k_{i}(P+T_{1}).

By [thegplawpaper, Theorem 3.2], there exists a 4×44\times 4 matrix of bilinear forms ϕij(P,T1)\phi_{ij}(P,T_{1}), with explicit formula, that is projectively equal to the matrix ki(PT1)kj(P+T1)k_{i}(P-T_{1})k_{j}(P+T_{1}). Since we have an explicit formula for MTM_{T} in [thebook, Chapter 3, Section 2], we can partially compute the linear isomorphism τT1\tau_{T_{1}}:

Ψ:J{kij,bi}15τT1J{kij,bi}15projkij9,\Psi:J\subset\mathbb{P}^{15}_{\{k_{ij},b_{i}\}}\xrightarrow{\tau_{T_{1}}}J\subset\mathbb{P}^{15}_{\{k_{ij},b_{i}\}}\xrightarrow{proj}\mathbb{P}^{9}_{k_{ij}},

as required.

Remark 4.4.

Suppose 2T1=TJ[2]2T_{1}=T\in J[2]. From the doubling formula on 𝒦\mathcal{K} as in [thegplawpaper, Appendix C], we can compute the coordinates of the image of T1T_{1} on 𝒦3\mathcal{K}\subset\mathbb{P}^{3} from the coordinates of the image of TT on 𝒦3\mathcal{K}\subset\mathbb{P}^{3}. This gives the 10 even coordinates, kij(T1)k_{ij}(T_{1}) and we can solve for the odd coordinates by the 72 defining equations of JJ given in Theorem 2.2. Note that by Lemma 4.2, we know the field of definition of T1T_{1} is contained in the composition of the field of definition of ϕϵ\phi_{\epsilon} and ϕϵ+δ(T)\phi_{\epsilon+\delta(T)}. Hence, we can compute this field explicitly which helps solving for this point using MAGMA [magma].

Consider TJ[2]T\in J[2] with T1J[4]T_{1}\in J[4] such that 2T1=T2T_{1}=T. We follow the discussion in Section 4.2 with ϵ\epsilon replaced by ϵ+δ(T)\epsilon+\delta(T). This gives a diagram analogous to (4.1). Let k1,T,,k4,Tk_{1,T}^{\prime},...,k_{4,T}^{\prime} be the coordinates on the ambient space of 𝒦ϵ+δ(T)3\mathcal{K}_{\epsilon+\delta(T)}\subset\mathbb{P}^{3} and let u0,T,,u9,T,v1,T,,v6,Tu_{0,T},...,u_{9,T},v_{1,T},...,v_{6,T} be the coordinates on the ambient space of Jϵ+δ(T)15J_{\epsilon+\delta(T)}\subset\mathbb{P}^{15}. Let kij,T=ki,Tkj,Tk_{ij,T}^{\prime}=k_{i,T}^{\prime}k_{j,T}^{\prime}. Decomposing the lower half of the diagram (4.2) gives the commutative diagram below:

Jϵ{ui,vi}15{J_{\epsilon}\subset\mathbb{P}^{15}_{\{u_{i},v_{i}\}}}𝒦ϵki3{\mathcal{K}_{\epsilon}\subset\mathbb{P}^{3}_{k_{i}^{\prime}}}J{kij,bi}15{J\subset\mathbb{P}^{15}_{\{k_{ij},b_{i}\}}}𝒦ki3{\mathcal{K}\subset\mathbb{P}_{k_{i}}^{3}}J{kij,bi}15{J\subset\mathbb{P}^{15}_{\{k_{ij},b_{i}\}}}kij9{\mathbb{P}_{k_{ij}}^{9}}𝒦ki3{\mathcal{K}\subset\mathbb{P}_{k_{i}}^{3}}Jϵ+δ(T){ui,T,vi,T}15{J_{\epsilon+\delta(T)}\subset\mathbb{P}^{15}_{\{u_{i,T},v_{i,T}\}}}kij,T9{\mathbb{P}_{k_{ij,T}^{\prime}}^{9}}𝒦ϵ+δ(T)ki,T3.{\mathcal{K}_{\epsilon+\delta(T)}\subset\mathbb{P}_{k_{i,T}^{\prime}}^{3}.}|ϕϵ(2Θ)|\scriptstyle{|\phi_{\epsilon}^{*}(2\Theta)|}ϕϵ\scriptstyle{\phi_{\epsilon}}ψϵ\scriptstyle{\psi_{\epsilon}}|2Θ|\scriptstyle{|2\Theta|}τT1\scriptstyle{\tau_{T_{1}}}Ψ\scriptstyle{\Psi}proj\scriptstyle{proj}g1\scriptstyle{g_{1}}(ψ~ϵ+δ(T))1\scriptstyle{(\tilde{\psi}_{\epsilon+\delta(T)})^{-1}}ϕϵ+δ(T)\scriptstyle{{\phi}_{\epsilon+\delta(T)}}g2\scriptstyle{g_{2}}ψϵ+δ(T)\scriptstyle{\psi_{\epsilon+\delta(T)}} (4.3)

Recall Proposition 4.3 explains how Ψ\Psi can be explicitly computed and the composition of the thick arrows in (4.3) is defined over KK by Lemma 4.2. Let DTD_{T} be the pull back on JϵJ_{\epsilon} via the thick arrows in (4.3) of the hyperplane section given by k1,T=0k_{1,T}^{\prime}=0. This implies that DTD_{T} is a KK-rational divisor on JϵJ_{\epsilon} representing the class of ϕϵ(τT1(2Θ))\phi_{\epsilon}^{*}(\tau_{T_{1}}^{*}(2\Theta)). Moreover, the pull back on JϵJ_{\epsilon} via

Jϵ{ui,vi}15ϕϵJ{kij,bi}15Ψkij9(ψ~ϵ+δ(T))1kij,T9J_{\epsilon}\subset\mathbb{P}^{15}_{\{u_{i},v_{i}\}}\xrightarrow{\phi_{\epsilon}}J\subset\mathbb{P}^{15}_{\{k_{ij,b_{i}}\}}\xrightarrow{\Psi}\mathbb{P}^{9}_{k{ij}}\xrightarrow{(\tilde{\psi}_{\epsilon+\delta(T)})^{-1}}\mathbb{P}^{9}_{k_{ij,T}^{\prime}}

of the hyperplane section given by k11,T=0k_{11,T}^{\prime}=0 is 2DT2D_{T}.

We now apply the above discussion with T=P,Q,R,ST=P,Q,R,S and get that the divisors DP,DQ,DR,DSD_{P},D_{Q},D_{R},D_{S} on JϵJ_{\epsilon} described in Remark 3.2 as required.

Remark 4.5.

From the above discussion and the discussion in Section 4.2, the KK-rational functions fP,fQ,fR,fSf_{P},f_{Q},f_{R},f_{S} in the formula for the Cassels-Tate pairing in Theorem 3.3 are quotients of linear forms in the coordinates of the ambient space of Jϵ15J_{\epsilon}\subset\mathbb{P}^{15}. They all have the same denominator, this being the linear form that cuts out the divisor 2D2D.

5 The Obstruction Map

In this section, we will state and prove 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). See below for the definition of this map. This generalizes a formula in the elliptic curve case due to O’Neil [oneil, Proposition 3.4], and later refined by Clark [clark, Theorem 6]. Although this is not needed for the computation of the Cassels-Tate pairing, it explains why we needed to work with quaternion algebras in Section 4.1.

Definition 5.1.

The obstruction map

Ob:H1(GK,J[2])H2(GK,K¯)Br(K)\text{Ob}:H^{1}(G_{K},J[2])\rightarrow H^{2}(G_{K},\bar{K}^{*})\cong\text{Br}(K)

is the composition of the map H1(GK,J[2])H1(GK,PGL4(K¯))H^{1}(G_{K},J[2])\rightarrow H^{1}(G_{K},\text{PGL}_{4}(\bar{K})) induced by the action of translation of J[2]J[2] on 𝒦3\mathcal{K}\subset\mathbb{P}^{3}, and the injective map H1(GK,PGL4(K¯))H2(GK,K¯)H^{1}(G_{K},\text{PGL}_{4}(\bar{K}))\rightarrow H^{2}(G_{K},\bar{K}^{*}) induced from the short exact sequence 0K¯GL4(K¯)PGL4(K¯)00\rightarrow\bar{K}^{*}\rightarrow\text{GL}_{4}(\bar{K})\rightarrow\text{PGL}_{4}(\bar{K})\rightarrow 0.

Theorem 5.2.

Let JJ be the Jacobian variety of a genus two curve defined over a field KK with char(K)2char(K)\neq 2. Suppose all points in J[2]J[2] are defined over KK. For ϵH1(GK,J[2])\epsilon\in H^{1}(G_{K},J[2]), represented by (a,b,c,d)(K/(K)2)4(a,b,c,d)\in(K^{*}/(K^{*})^{2})^{4} as in Section 3, the obstruction map Ob:H1(GK,J[2])Br(K)\mathrm{Ob}:H^{1}(G_{K},J[2])\rightarrow\mathrm{Br}(K) sends ϵ\epsilon to the class of the tensor product of two quaternion algebras:

Ob(ϵ)=(cPa,cQb)+(cRc,cSd),\mathrm{Ob}(\epsilon)=(c_{P}a,c_{Q}b)+(c_{R}c,c_{S}d),

where cP,cQ,cR,cSKc_{P},c_{Q},c_{R},c_{S}\in K are such that MP2=cPI,MQ2=cQI,MR2=cRI,andMS2=cSIM_{P}^{2}=c_{P}I,M_{Q}^{2}=c_{Q}I,M_{R}^{2}=c_{R}I,\;\text{and}\;M_{S}^{2}=c_{S}I as defined in Section 4.1.

Proof.

Let NP=1/cPMP,NQ=1/cQMQ,NR=1/cRMR,NS=1/cSMSGL4(K¯)N_{P}=1/\sqrt{c_{P}}M_{P},N_{Q}=1/\sqrt{c_{Q}}M_{Q},N_{R}=1/\sqrt{c_{R}}M_{R},N_{S}=1/\sqrt{c_{S}}M_{S}\in\text{GL}_{4}(\bar{K}). Then NPN_{P} is a normalized representation in GL4(K¯)\text{GL}_{4}(\bar{K}) of [MP]PGL4(K)[M_{P}]\in\text{PGL}_{4}(K). Similar statements are true for Q,R,SQ,R,S. Notice that NP2=NQ2=NR2=NS2=IN_{P}^{2}=N_{Q}^{2}=N_{R}^{2}=N_{S}^{2}=I. So there is a uniform way of picking a representation in GL4(K¯)\text{GL}_{4}(\bar{K}) for the translation induced by α1P+α2Q+α3R+α4S\alpha_{1}P+\alpha_{2}Q+\alpha_{3}R+\alpha_{4}S for αi\alpha_{i}\in\mathbb{Z}, namely NPα1NQα2NRα3NSα4.N_{P}^{\alpha_{1}}N_{Q}^{\alpha_{2}}N_{R}^{\alpha_{3}}N_{S}^{\alpha_{4}}.

Since ϵH1(K,J[2])\epsilon\in H^{1}(K,J[2]) is represented by (a,b,c,d)(K/K2)4(a,b,c,d)\in(K^{*}/{K^{*}}^{2})^{4} and P,Q,R,SP,Q,R,S satisfy the Weil pairing matrix (3.1), a cocycle representation of ϵ\epsilon is:

σb~σP+a~σQ+d~σR+c~σS,\sigma\mapsto\tilde{b}_{\sigma}P+\tilde{a}_{\sigma}Q+\tilde{d}_{\sigma}R+\tilde{c}_{\sigma}S,

where for each element xK/(K)2x\in K^{*}/(K^{*})^{2}, we define x~σ{0,1}\tilde{x}_{\sigma}\in\{0,1\} such that (1)x~σ=σ(x)/x(-1)^{\tilde{x}_{\sigma}}=\sigma(\sqrt{x})/\sqrt{x}.

Now consider the following commutative diagram of cochains:

C1(GK,K¯){C^{1}(G_{K},\bar{K}^{*})}C1(GK,GL4){C^{1}(G_{K},\text{GL}_{4})}C1(GK,PGL4){C^{1}(G_{K},\text{PGL}_{4})}C2(GK,K¯){C^{2}(G_{K},\bar{K}^{*})}C2(GK,GL4){C^{2}(G_{K},\text{GL}_{4})}C2(GK,PGL4).{C^{2}(G_{K},\text{PGL}_{4}).}d\scriptstyle{d}d\scriptstyle{d}d\scriptstyle{d}

Defining Nσ=NPb~σNQa~σNRd~σNSc~σN_{\sigma}=N_{P}^{\tilde{b}_{\sigma}}N_{Q}^{\tilde{a}_{\sigma}}N_{R}^{\tilde{d}_{\sigma}}N_{S}^{\tilde{c}_{\sigma}}, we have

H1(K,J[2])H1(GK,PGL4)(a,b,c,d)(σ[Nσ]).\begin{array}[]{ccc}H^{1}(K,J[2])&\rightarrow&H^{1}(G_{K},\text{PGL}_{4})\\ (a,b,c,d)&\mapsto&(\sigma\mapsto[N_{\sigma}]).\end{array}

Then (σ[Nσ])C1(GK,PGL4)(\sigma\mapsto[N_{\sigma}])\in C^{1}(G_{K},\text{PGL}_{4}) lifts to (σNσ)C1(GK,GL4)(\sigma\mapsto N_{\sigma})\in C^{1}(G_{K},\text{GL}_{4}) which is then mapped to

((σ,τ)(Nτ)σNστ1Nσ)C2(GK,GL4).((\sigma,\tau)\mapsto(N_{\tau})^{\sigma}N_{\sigma\tau}^{-1}N_{\sigma})\in C^{2}(G_{K},\text{GL}_{4}).

Note that

NPσ=(1cPMP)σ=1σ(cP)MP=cPσ(cP)NP=(1)(cP~)σNP,N_{P}^{\sigma}=(\frac{1}{\sqrt{c_{P}}}M_{P})^{\sigma}=\frac{1}{\sigma(\sqrt{c_{P}})}M_{P}=\frac{\sqrt{c_{P}}}{\sigma(\sqrt{c_{P}})}N_{P}=(-1)^{(\widetilde{c_{P}})_{\sigma}}N_{P},

treating cPc_{P} in K/(K)2K^{*}/(K^{*})^{2}. Similar results also hold for Q,R,SQ,R,S. Observe that for any xK/(K)2x\in K^{*}/(K^{*})^{2} and σ,τGK\sigma,\tau\in G_{K}, we have x~σx~στ+x~σ\tilde{x}_{\sigma}-\tilde{x}_{\sigma\tau}+\tilde{x}_{\sigma} is equal to 0 or 2. Since NP2=NQ2=NR2=NS2=IN_{P}^{2}=N_{Q}^{2}=N_{R}^{2}=N_{S}^{2}=I, [NP,NQ]=[NR,NS]=I[N_{P},N_{Q}]=[N_{R},N_{S}]=-I and the commutators of the other pairs are trivial, we have

(Nτ)σNστ1Nσ=\displaystyle(N_{\tau})^{\sigma}N_{\sigma\tau}^{-1}N_{\sigma}= (NPb~τNQa~τNRd~τNSc~τ)σNSc~στNRd~στNQa~στNPb~στNPb~σNQa~σNRd~σNSc~σ\displaystyle(N_{P}^{\tilde{b}_{\tau}}N_{Q}^{\tilde{a}_{\tau}}N_{R}^{\tilde{d}_{\tau}}N_{S}^{\tilde{c}_{\tau}})^{\sigma}\cdot N_{S}^{-\tilde{c}_{\sigma\tau}}N_{R}^{-\tilde{d}_{\sigma\tau}}N_{Q}^{-\tilde{a}_{\sigma\tau}}N_{P}^{-\tilde{b}_{\sigma\tau}}\cdot N_{P}^{\tilde{b}_{\sigma}}N_{Q}^{\tilde{a}_{\sigma}}N_{R}^{\tilde{d}_{\sigma}}N_{S}^{\tilde{c}_{\sigma}}
=\displaystyle= (1)(cP~)σb~τ(1)(cQ~)σa~τ(1)(cR~)σd~τ(1)(cS~)σc~τ\displaystyle(-1)^{(\widetilde{c_{P}})_{\sigma}\cdot\tilde{b}_{\tau}}\cdot(-1)^{(\widetilde{c_{Q}})_{\sigma}\cdot\tilde{a}_{\tau}}\cdot(-1)^{(\widetilde{c_{R}})_{\sigma}\cdot\tilde{d}_{\tau}}\cdot(-1)^{(\widetilde{c_{S}})_{\sigma}\cdot\tilde{c}_{\tau}}
NPb~τNQa~τNRd~τNSc~τNSc~στNRd~στNQa~στNPb~στNPb~σNQa~σNRd~σNSc~σ\displaystyle\cdot N_{P}^{\tilde{b}_{\tau}}N_{Q}^{\tilde{a}_{\tau}}\cdot N_{R}^{\tilde{d}_{\tau}}N_{S}^{\tilde{c}_{\tau}}N_{S}^{-\tilde{c}_{\sigma\tau}}N_{R}^{-\tilde{d}_{\sigma\tau}}\cdot N_{Q}^{-\tilde{a}_{\sigma\tau}}N_{P}^{-\tilde{b}_{\sigma\tau}}N_{P}^{\tilde{b}_{\sigma}}N_{Q}^{\tilde{a}_{\sigma}}\cdot N_{R}^{\tilde{d}_{\sigma}}N_{S}^{\tilde{c}_{\sigma}}
=\displaystyle= (1)(cP~)σb~τ(1)(cQ~)σa~τ(1)(cR~)σd~τ(1)(cS~)σc~τ\displaystyle(-1)^{(\widetilde{c_{P}})_{\sigma}\cdot\tilde{b}_{\tau}}\cdot(-1)^{(\widetilde{c_{Q}})_{\sigma}\cdot\tilde{a}_{\tau}}\cdot(-1)^{(\widetilde{c_{R}})_{\sigma}\cdot\tilde{d}_{\tau}}\cdot(-1)^{(\widetilde{c_{S}})_{\sigma}\cdot\tilde{c}_{\tau}}
NPb~τNQa~τNQa~στNPb~στNPb~σNQa~σNRd~τNSc~τNSc~στNRd~στNRd~σNSc~σ\displaystyle\cdot N_{P}^{\tilde{b}_{\tau}}N_{Q}^{\tilde{a}_{\tau}}N_{Q}^{-\tilde{a}_{\sigma\tau}}N_{P}^{-\tilde{b}_{\sigma\tau}}N_{P}^{\tilde{b}_{\sigma}}N_{Q}^{\tilde{a}_{\sigma}}\cdot N_{R}^{\tilde{d}_{\tau}}N_{S}^{\tilde{c}_{\tau}}N_{S}^{-\tilde{c}_{\sigma\tau}}N_{R}^{-\tilde{d}_{\sigma\tau}}N_{R}^{\tilde{d}_{\sigma}}N_{S}^{\tilde{c}_{\sigma}}
=\displaystyle= (1)(cP~)σb~τ(1)(cQ~)σa~τ(1)(cR~)σd~τ(1)(cS~)σc~τ(1)a~σb~τ(1)c~σd~τI.\displaystyle(-1)^{(\widetilde{c_{P}})_{\sigma}\cdot\tilde{b}_{\tau}}\cdot(-1)^{(\widetilde{c_{Q}})_{\sigma}\cdot\tilde{a}_{\tau}}\cdot(-1)^{(\widetilde{c_{R}})_{\sigma}\cdot\tilde{d}_{\tau}}\cdot(-1)^{(\widetilde{c_{S}})_{\sigma}\cdot\tilde{c}_{\tau}}\cdot(-1)^{\tilde{a}_{\sigma}\cdot\tilde{b}_{\tau}}\cdot(-1)^{\tilde{c}_{\sigma}\cdot\tilde{d}_{\tau}}\cdot I.

On the other hand, (cP,cQ)(cR,cS)(c_{P},c_{Q})\otimes(c_{R},c_{S}) is isomorphic to MP,MQ,MR,MS=Mat4(K)\langle M_{P},M_{Q},M_{R},M_{S}\rangle=\text{Mat}_{4}(K) which represents the identity element in the Brauer group. Hence, we have

(cPa,cQb)+(cRc,cSd)=(a,b)+(c,d)+(cP,b)+(cQ,a)+(cR,d)+(cS,c),(c_{P}a,c_{Q}b)+(c_{R}c,c_{S}d)=(a,b)+(c,d)+(c_{P},b)+(c_{Q},a)+(c_{R},d)+(c_{S},c),

which is precisely represented by a cocycle that sends (σ,τ)(\sigma,\tau) to

(1)(cP~)σb~τ(1)(cQ~)σa~τ(1)(cR~)σd~τ(1)(cS~)σc~τ(1)a~σb~τ(1)c~σd~τ,(-1)^{(\widetilde{c_{P}})_{\sigma}\cdot\tilde{b}_{\tau}}\cdot(-1)^{(\widetilde{c_{Q}})_{\sigma}\cdot\tilde{a}_{\tau}}\cdot(-1)^{(\widetilde{c_{R}})_{\sigma}\cdot\tilde{d}_{\tau}}\cdot(-1)^{(\widetilde{c_{S}})_{\sigma}\cdot\tilde{c}_{\tau}}\cdot(-1)^{\tilde{a}_{\sigma}\cdot\tilde{b}_{\tau}}\cdot(-1)^{\tilde{c}_{\sigma}\cdot\tilde{d}_{\tau}},

for all σ,τGK\sigma,\tau\in G_{K} as required.

6 Bounding the Set of Primes

In this section, we directly show that the formula for ϵ,ηCT\langle\epsilon,\eta\rangle_{CT} in Theorem 3.3 is actually always a finite product, as mentioned in Remark 3.4. Since for a local field with odd residue characteristic, the Hilbert symbol between xx and yy is trivial when the valuations of xx, yy are both 0, it suffices to find a finite set SS of places of KK, such that outside SS the first arguments of the Hilbert symbols in the formula for ϵ,ηCT\langle\epsilon,\eta\rangle_{CT} have valuation 0 for some choice of the local point PvP_{v}.

Let 𝒪K\mathcal{O}_{K} be the ring of integers for the number field KK. By rescaling the variables, we assume the genus two curve is defined by y2=f(x)=f6x6++f0y^{2}=f(x)=f_{6}x^{6}+...+f_{0} where the fif_{i} are in 𝒪K\mathcal{O}_{K}.

The first arguments of the Hilbert symbols in the formula for ϵ,ηCT\langle\epsilon,\eta\rangle_{CT} are fP(Pv)f_{P}(P_{v}), fQ(Pv)f_{Q}(P_{v}), fR(Pv)f_{R}(P_{v}) or fS(Pv),f_{S}(P_{v}), where fP,fQ,fR,fSf_{P},f_{Q},f_{R},f_{S} can be computed as the quotients of two linear forms in 15\mathbb{P}^{15} with the denominators being the same, as explained in Remark 4.5. Since we know that the Cassels-Tate pairing is independent of the choice of the local points PvP_{v} as long as these are chosen to avoid all the zeros and poles, it suffices to make sure that there exists at least one local point PvP_{v} on JϵJ_{\epsilon} for which the values of the quotients of the linear forms all have valuation 0 for all vv outside SS. The idea is to first reduce the problem to the residue field.

By Theorem 2.9 and Remark 2.10, we have an explicit formula for the linear isomorphism

Jϵ15ϕϵJ15,J_{\epsilon}\subset\mathbb{P}^{15}\xrightarrow{\phi_{\epsilon}}J\subset\mathbb{P}^{15},

which is defined over K=K(a,b,c,d)K^{\prime}=K(\sqrt{a},\sqrt{b},\sqrt{c},\sqrt{d}) where ϵ=(a,b,c,d)(K/(K)2)4\epsilon=(a,b,c,d)\in(K^{*}/(K^{*})^{2})^{4}. Suppose ϕϵ\phi_{\epsilon} is represented by MϵGL16(K)M_{\epsilon}\in\text{GL}_{16}(K^{\prime}). Note we can assume that all entries of MϵM_{\epsilon} are in 𝒪K\mathcal{O}_{K^{\prime}}, the ring of integers of KK^{\prime}.

Notation 6.1.

Let KK be a local field with valuation ring 𝒪K\mathcal{O}_{K}, uniformizer π\pi and residue field kk. Let XNX\subset\mathbb{P}^{N} be a variety defined over KK and I(X)K[x0,,xN]I(X)\subset K[x_{0},...,x_{N}] be the ideal of XX. Then the reduction of XX, denoted by X¯\bar{X}, is the variety defined by the polynomials {f¯:fI(X)𝒪K[x0,,xN]}\{\bar{f}:f\in I(X)\cap\mathcal{O}_{K}[x_{0},...,x_{N}]\}. Here f¯\bar{f} is the polynomial obtained by reducing all the coefficients of ff modulo π\pi. Note that this definition of the reduction of a variety XNX\subset\mathbb{P}^{N} defined over a local field KK is equivalent to taking the special fibre of the closure of XX in SN\mathbb{P}^{N}_{S}, where S=Spec𝒪KS=\operatorname{Spec}\mathcal{O}_{K}.

Let S0={places of bad reduction for 𝒞}{places dividing 2}{infinite places}S_{0}=\{\text{places of bad reduction for }\mathcal{C}\}\cup\{\text{places dividing 2}\}\cup\{\text{infinite places}\}. Fix a place vS0v\notin S_{0} and suppose it is above the prime pp. We now treat J,JϵJ,J_{\epsilon} and 𝒞\mathcal{C} as varieties defined over the local field KvK_{v}. Let 𝒪v\mathcal{O}_{v} denote the valuation ring of KvK_{v} and 𝔽q\mathbb{F}_{q} denote its residue field, where qq is some power of pp. It can be shown that J¯\bar{J} is also an abelian variety as the defining equations of JJ are defined over 𝒪v\mathcal{O}_{v} and are derived algebraically in terms of the coefficients of the defining equation of the genus two curve 𝒞\mathcal{C} by Theorem 2.2. In fact, J¯\bar{J} is the Jacobian variety of 𝒞¯\bar{\mathcal{C}}, the reduction of 𝒞\mathcal{C}.

Now fix a place vv^{\prime} of KK^{\prime} above the place vv of KK. Let 𝒪v\mathcal{O}_{v^{\prime}} and 𝔽qr\mathbb{F}_{q^{r}} denote the valuation ring and the residue field of KvK^{\prime}_{v^{\prime}}. It can be checked that as long as vv^{\prime} does not divide detMϵ𝒪K\det M_{\epsilon}\in\mathcal{O}_{K^{\prime}}, the following diagram commutes and Mϵ¯\bar{M_{\epsilon}} is a well defined linear isomorphism defined over the residue field 𝔽qr\mathbb{F}_{q^{r}} between two varieties defined over 𝔽q\mathbb{F}_{q}:

Jϵ15{J_{\epsilon}\subset\mathbb{P}^{15}}J15{J\subset\mathbb{P}^{15}}Jϵ¯15{\bar{J_{\epsilon}}\subset\mathbb{P}^{15}}J¯15,{\bar{J}\subset\mathbb{P}^{15},}Mϵ\scriptstyle{M_{\epsilon}}reductionreductionMϵ¯\scriptstyle{\bar{M_{\epsilon}}}

where Mϵ¯\bar{M_{\epsilon}} denotes the reduction of the matrix MϵM_{\epsilon} over the residue field 𝔽qr\mathbb{F}_{q^{r}}.

This linear isomorphism Mϵ¯\bar{M_{\epsilon}} implies that Jϵ¯\bar{J_{\epsilon}} is smooth whenever J¯\bar{J} is. In this case, Jϵ¯\bar{J_{\epsilon}} is a twist of J¯\bar{J} and it in fact a 2-covering of J¯\bar{J}. Indeed, the surjectivity of the natural map Gal(Kv/Kv)Gal(𝔽qr/𝔽q)\text{Gal}(K^{\prime}_{v^{\prime}}/K_{v})\rightarrow\text{Gal}(\mathbb{F}_{q^{r}}/\mathbb{F}_{q}) shows that Mϵ(Mϵ1)σ=τPσM_{\epsilon}(M_{\epsilon}^{-1})^{\sigma}=\tau_{P_{\sigma}} for all σGal(Kv/Kv)\sigma\in\text{Gal}(K^{\prime}_{v^{\prime}}/K_{v}) implies that Mϵ¯(Mϵ¯1)σ¯=τPσ¯¯\bar{M_{\epsilon}}(\bar{M_{\epsilon}}^{-1})^{\bar{\sigma}}=\tau_{\bar{P_{\bar{\sigma}}}} for all σ¯Gal(𝔽qr/𝔽q)\bar{\sigma}\in\text{Gal}(\mathbb{F}_{q^{r}}/\mathbb{F}_{q}). We know any principal homogeneous space of J¯\bar{J} over a finite field has a point by [lang, Theorem 2] and so is trivial by Proposition 2.4. Therefore, there exists an isomorphism Jϵ¯𝜓J¯\bar{J_{\epsilon}}\xrightarrow{\psi}\bar{J} defined over 𝔽q\mathbb{F}_{q}. Hence, as long as vS0v\notin S_{0} and vv does not divide NK/K(detMϵ)N_{K^{\prime}/K}(\det M_{\epsilon}), Jϵ¯\bar{J_{\epsilon}} has the same number of 𝔽q\mathbb{F}_{q}-points as J¯\bar{J}. By the Hasse-Weil bound, we know the number of 𝔽q\mathbb{F}_{q}-points on 𝒞\mathcal{C} is bounded below by q14qq-1-4\sqrt{q}. Since we can represent points on J¯\bar{J} by pairs of points on 𝒞¯\bar{\mathcal{C}} and this representation is unique other than the identity point on J¯\bar{J}. The number of 𝔽q\mathbb{F}_{q}-points on J¯\bar{J} is bounded below by (q14q)(q34q)/2(q-1-4\sqrt{q})(q-3-4\sqrt{q})/2.

On the other hand, let l1,,l5l_{1},...,l_{5} be the 5 linear forms that appear as numerator or denominator of fP,fQ,fR,fSf_{P},f_{Q},f_{R},f_{S}. We can assume that the coefficients of lil_{i} are in 𝒪K\mathcal{O}_{K} by scaling, for all i=1,,5i=1,...,5. Fix a place vv of KK that does not divide all the coefficients of lil_{i}, for any i=1,,5i=1,...,5. Let HiH_{i} be the hyperplane defined by the linear form lil_{i} and Hi¯\bar{H_{i}} be its reduction, which is a hyperplane defined over the residue field 𝔽q\mathbb{F}_{q}, We need to bound the number of 𝔽q\mathbb{F}_{q}-points of Jϵ¯\bar{J_{\epsilon}} that lie on one of the hyperplanes Hi¯\bar{H_{i}}. Let rir_{i} be the number of irreducible components of Jϵ¯Hi¯\bar{J_{\epsilon}}\cap\bar{H_{i}}. By [hartshorne, Chapter 1, Theorem 7.2 (Projective Dimension Theorem) and Theorem 7.7], we know that each irreducible component CjiC^{i}_{j} of Jϵ¯Hi¯\bar{J_{\epsilon}}\cap\bar{H_{i}}, where j=1,,rij=1,...,r_{i}, is a curve and the sum of degrees of all the irreducible components counting intersection multiplicity is degJϵ¯=32.\deg\bar{J_{\epsilon}}=32. Leting dji=degCjid^{i}_{j}=\deg C^{i}_{j}, we have j=1ridji32\sum_{j=1}^{r_{i}}d^{i}_{j}\leq 32 for all ii.

Lemma 6.2.

Let CNC\subset\mathbb{P}^{N} be a curve of degree dd. Then #C(𝔽q)d(q+1)\#C(\mathbb{F}_{q})\leq d(q+1).

Proof.

We may assume that CC is contained in no hyperplane. Then projection to the first two coordinates gives a nonconstant morphism C1C\rightarrow\mathbb{P}^{1} of degree d\leq d. Since #1(𝔽𝕢)=q+1\#\mathbb{P}^{1}(\mathbb{F_{q}})=q+1, this gives the required bound.

By applying the above lemma to each CjiC_{j}^{i}, we get the number of 𝔽q\mathbb{F}_{q}-points of Jϵ¯\bar{J_{\epsilon}} that lie on one of the hyperplanes Hi¯,i=1,,5,\bar{H_{i}},i=1,...,5, is no more than

i=15j=1ridji(q+1)160(q+1).\sum_{i=1}^{5}\sum_{j=1}^{r_{i}}d^{i}_{j}\cdot(q+1)\leq 160(q+1).

We compute that for any x>500x>500, we have (x14x)(x34x)/2>160(x+1)(x-1-4\sqrt{x})(x-3-4\sqrt{x})/2>160(x+1). Recall qq is a power of pp. Hence, if vv is a place of KK above the prime p>500p>500 such that vS0v\notin S_{0} and vv does not divide NK/K(detMϵ)N_{K^{\prime}/K}(\det M_{\epsilon}) or all the coefficients of lil_{i} for some ii, we have a smooth 𝔽q\mathbb{F}_{q}-point on Jϵ¯\bar{J_{\epsilon}} which by Hensel’s Lemma [hensel, Exercise C.9(c)] lifts to the point PvP_{v} as required. This implies that the first arguments of the Hilbert symbols in the formula for the local Cassels-Tate pairing of ϵ,ηCT\langle\epsilon,\eta\rangle_{CT} have valuation 0. It can be checked that since vS0v\notin S_{0}, the second arguments of these Hilbert symbols also have valuation 0. Hence, the formula for the Cassels-Tate is indeed always a finite product.

Note that in the case where K=K=\mathbb{Q} or more generally if KK has class number 1, we can always make the linear forms primitive by scaling. Therefore, in this case, the subset {\{places dividing all the coefficients of the denominator or the numerator of fP,fQ,fR or fS}f_{P},f_{Q},f_{R}\text{ or }f_{S}\} is empty.

7 Worked Example

Now we demonstrate the algorithm with a worked example computed using MAGMA [magma]. In particular, we will see with this example, that computing the Cassels-Tate pairing on Sel2(J)\text{Sel}^{2}(J) does improve the rank bound obtained via a 2-descent. This genus two curve was kindly provided by my PhD supervisor, Tom Fisher, along with a list of other genus two curves for me to test the algorithm.

Consider the following genus two curve

𝒞:y2=10x(x+10)(x+5)(x10)(x5)(x1).\mathcal{C}:y^{2}=-10x(x+10)(x+5)(x-10)(x-5)(x-1).

Its Jacobian variety JJ has all its two torsion points defined over \mathbb{Q}. A set of generators of J[2]J[2] compatible with the Weil pairing matrix (3.1) are P={(0,0),(10,0)},Q={(0,0),(5,0)},R={(10,0),(5,0)},S={(10,0),(1,0)}.P=\{(0,0),(-10,0)\},Q=\{(0,0),(-5,0)\},R=\{(10,0),(5,0)\},S=\{(10,0),(1,0)\}. We identify H1(GK,J[2])=(/()2)4H^{1}(G_{K},J[2])=(\mathbb{Q}^{*}/(\mathbb{Q}^{*})^{2})^{4} as in Section 3. Consider ϵ,ηSel2(J)\epsilon,\eta\in\text{Sel}^{2}(J) represented by (33,1,1,11)(-33,1,-1,-11) and (11,1,1,11)(11,1,-1,-11) respectively. The images of [P],[Q],[R],[S][P],[Q],[R],[S] via δ:J()/2J()H1(G,J[2])\delta:J(\mathbb{Q})/2J(\mathbb{Q})\rightarrow H^{1}(G_{\mathbb{Q}},J[2]), computed via the explicit formula as in [thebook, Chapter 6, Section 1], are δ([P])=(66,1,6,22),δ([Q])=(1,1,3,1),δ([R])=(6,3,1,3),δ([S])=(22,1,3,11).\delta([P])=(-66,1,6,22),\delta([Q])=(-1,1,3,1),\delta([R])=(6,3,1,3),\delta([S])=(22,1,-3,-11). Now following the discussions in Sections 4.2 and 4.3, we can compute, using the coordinates c0,,c9,d1,,d6c_{0},...,c_{9},d_{1},...,d_{6} for Jϵ15J_{\epsilon}\in\mathbb{P}^{15} as described in Remark 2.10. we have

k11\displaystyle k_{11}^{\prime} =618874080c0496218440c1390547052c3+205551080c4\displaystyle=618874080c_{0}-496218440c_{1}-390547052c_{3}+205551080c_{4}
+384569291c6+52868640c8;\displaystyle+384569291c_{6}+52868640c_{8};
k11,P\displaystyle k_{11,P}^{\prime} =36051078800000c2+8111492730000c3+265237150000c7\displaystyle=-36051078800000c_{2}+8111492730000c_{3}+265237150000c_{7}
196928587500c86786529337500c9+22531924250d2\displaystyle-196928587500c_{8}-6786529337500c_{9}+22531924250d_{2}
126449158891d4117221870375d5+937774963000d6;\displaystyle-126449158891d_{4}-117221870375d_{5}+937774963000d_{6};
k11,Q\displaystyle k_{11,Q}^{\prime} =134800c1+235600c3+62000c4+52235c6+60016d15456d5;\displaystyle=134800c_{1}+235600c_{3}+62000c_{4}+52235c_{6}+60016d_{1}-5456d_{5};
k11,R\displaystyle k_{11,R}^{\prime} =30223125c6+4050000c849750d3+709236d4\displaystyle=-30223125c_{6}+4050000c_{8}-49750d_{3}+709236d_{4}
k11,S\displaystyle k_{11,S}^{\prime} =4724524800c1+8557722360c3+13102732800c4+1258642935c6\displaystyle=4724524800c_{1}+8557722360c_{3}+13102732800c_{4}+1258642935c_{6}
+7291944000c92709362304d1+97246845d2+8475710d3\displaystyle+7291944000c_{9}-2709362304d_{1}+97246845d_{2}+8475710d_{3}
+30788208d5.\displaystyle+30788208d_{5}.

Hence, we have explicit formulae for

fP=k11,Pk11,fQ=k11,Qk11,fR=k11,Rk11,fS=k11,Sk11.f_{P}=\frac{k_{11,P}^{\prime}}{k_{11}^{\prime}},f_{Q}=\frac{k_{11,Q}^{\prime}}{k_{11}^{\prime}},f_{R}=\frac{k_{11,R}^{\prime}}{k_{11}^{\prime}},f_{S}=\frac{k_{11,S}^{\prime}}{k_{11}^{\prime}}.

In particular, they are defined over \mathbb{Q} as claimed. From Section 6, we compute that only primes below 500 can potentially contribute to ϵ,ηCT\langle\epsilon,\eta\rangle_{CT}. Then, it turns out that the only nontrivial local Cassels-Tate pairings between ϵ\epsilon and η\eta are at places 11,19,11,19,\infty and ϵ,ηCT=1\langle\epsilon,\eta\rangle_{CT}=-1.

Under the isomorphism H1(G,J[2])(/()2)4H^{1}(G_{\mathbb{Q}},J[2])\rightarrow(\mathbb{Q}^{*}/(\mathbb{Q}^{*})^{2})^{4}, Sel2(J)\text{Sel}^{2}(J) has size 262^{6} and is generated by (33,1,1,11),(11,1,1,11),(66,1,2,22),(11,1,2,22),(3,3,3,3),(3,1,3,1).(-33,1,-1,-11),(11,1,-1,-11),(66,1,2,22),(11,1,2,22),(3,3,3,3),(3,1,3,1). Since 𝒞\mathcal{C} has rational points, the Cassels-Tate pairing can be shown to be alternating using [poonenstoll, Corollary 7]. Since all the two torsion points on JJ are rational and ϵ,ηCT=1\langle\epsilon,\eta\rangle_{CT}=-1, we get |ker,CT|=24|\ker\langle\;,\;\rangle_{CT}|=2^{4}.

Indeed, we verified that the Cassels-Tate pairing matrix, with the generators of Sel2(J)\text{Sel}^{2}(J) listed above, is

[111111111111111111111111111111111111],\begin{bmatrix}1&-1&1&1&-1&-1\\ -1&1&1&-1&1&-1\\ 1&1&1&1&1&1\\ 1&-1&1&1&-1&-1\\ -1&1&1&-1&1&-1\\ -1&-1&1&-1&-1&1\\ \end{bmatrix},

which is a rank 2 matrix.

As shown in [thesis, Remark 1.9.4(ii)], in the case where all points in J[2]J[2] are defined over the base field, computing the Cassels-Tate pairing on Sel2(J)\text{Sel}^{2}(J) gives the same rank bound as obtained from carrying out a 44-descent, i.e. computing Sel4(J)\text{Sel}^{4}(J), which can potentially give a better rank bound than the one given by a 2-descent. In this example, the rank bound coming from 2-decent was rank(J())2\operatorname{rank}(J(\mathbb{Q}))\leq 2. Our calculations of the Cassels-Tate pairing on Sel2(J)\text{Sel}^{2}(J) improves this bound and in fact shows that rank(J())=0\operatorname{rank}(J(\mathbb{Q}))=0.

References