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

Sesquilinear pairings on elliptic curves

Katherine E. Stange
Abstract.

Let EE be an elliptic curve with complex multiplication by a ring RR, where RR is an order in an imaginary quadratic field or quaternion algebra. We define sesquilinear pairings (RR-linear in one variable and RR-conjugate linear in the other), taking values in an RR-module, generalizing the Weil and Tate-Lichtenbaum pairings.

Key words and phrases:
Elliptic curves, Weil pairing, Tate-Lichtenbaum pairing, complex multiplication
2020 Mathematics Subject Classification:
Primary: 11G05, 14H52
This work has been supported by NSF-CAREER CNS-1652238 and NSF DMS-2401580.

1. Introduction

The Weil and Tate-Lichtenbaum pairings are bilinear pairings on an elliptic curve EE with values in the multiplicative group 𝔾m\mathbb{G}_{m}. In the situation of complex multiplication, the points of the elliptic curve form more than just a \mathbb{Z}-module, but also an RR-module, for some ring RR which is an order in either an imaginary quadratic field or a quaternion algebra, both of which come equipped with an involution which we call conjugation. It is natural then to hope for a pairing with some type of RR-linearity. In this paper, we generalize these classical pairings to take values in an RR-module, so that the pairings can become sesquilinear, or conjugate linear in the following sense. If RR is commutative, an RR-sesquilinear pairing is a bilinear pairing ,\langle\cdot,\cdot\rangle on a pair of RR-modules, taking values in another RR-module, that satisfies

αx,βy=αβ¯x,y, for all α,βR.\langle\alpha x,\beta y\rangle={\alpha}{\overline{\beta}}\langle x,y\rangle,\text{ for all }\alpha,\beta\in R.

In the case that RR is non-commutative, we also consider a twisted version; see Section 4. For the remainder of the introduction, we assume RR is commutative; small adjustments are needed in the non-commutative case.

The Weil and Tate-Lichtenbaum pairings can be taken to act on divisor classes in Pic0(E)\operatorname{Pic}^{0}(E). By considering instead PicR0(E):=RPic0(E)\operatorname{Pic}_{R}^{0}(E):=R\otimes_{\mathbb{Z}}\operatorname{Pic}^{0}(E), we have an RR-module structure on divisor classes. To accommodate the values of the pairing, considering 𝔾m\mathbb{G}_{m} as a \mathbb{Z}-module in multiplicative notation, we can extend scalars to RR, writing 𝔾mR\mathbb{G}_{m}^{\otimes_{\mathbb{Z}}R}. (This multiplicative tensor notation is not without its pitfalls; see the end of the introduction for further discussion.) Write M[α]M[\alpha] for the α\alpha-torsion in an RR-module MM. For each αR\alpha\in R, we obtain Galois invariant sesquilinear pairings

Wα\displaystyle{W}_{\alpha} :PicR0(E)[α¯]×PicR0(E)[α]𝔾mR[α¯],\displaystyle:\operatorname{Pic}_{R}^{0}(E)[\overline{\alpha}]\times\operatorname{Pic}_{R}^{0}(E)[\alpha]\rightarrow\mathbb{G}_{m}^{\otimes_{\mathbb{Z}}R}[\overline{\alpha}],
Tα\displaystyle{T}_{\alpha} :PicR0(E)[α¯]×PicR0(E)/[α]PicR0(E)𝔾mR/(𝔾mR)α¯,\displaystyle:\operatorname{Pic}_{R}^{0}(E)[\overline{\alpha}]\times\operatorname{Pic}_{R}^{0}(E)/[\alpha]\operatorname{Pic}_{R}^{0}(E)\rightarrow\mathbb{G}_{m}^{\otimes_{\mathbb{Z}}R}/(\mathbb{G}_{m}^{\otimes_{\mathbb{Z}}R})^{\overline{\alpha}},

generalizing the classical Weil and Tate-Lichtenbaum pairings (these do not restrict to the classical pairings, but restrict to a sesquilinearization of such; see Proposition 4.4 and the discussion afterward). The pairing WαW_{\alpha} is also conjugate skew-Hermitian in the sense that

Wα(DP,DQ)=Wα¯(DQ,DP)¯1.W_{\alpha}(D_{P},D_{Q})=\overline{W_{\overline{\alpha}}(D_{Q},D_{P})}^{-1}.

These are defined by essentially imitating the definition of the classical pairings, including extending Weil reciprocity to RR-divisors.

However, this formal exercise is most interesting when applied to a curve with endomorphism ring containing a copy of RR. Consider an exact sequence

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}E\textstyle{E\ignorespaces\ignorespaces\ignorespaces\ignorespaces}η\scriptstyle{\eta}PicR0(E)\textstyle{\operatorname{Pic}^{0}_{R}(E)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ϵ\scriptstyle{\epsilon}E\textstyle{E\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}

given by

ϵ:iαi(Pi)i[αi]Pi.\epsilon:\sum_{i}\alpha_{i}(P_{i})\mapsto\sum_{i}[\alpha_{i}]P_{i}.

By restricting the pairing to the left-hand EE in the exact sequence, we obtain Galois invariant pairings

W^α\displaystyle\widehat{W}_{\alpha} :E[α¯]×E[α]𝔾mR[α],\displaystyle:E[\overline{\alpha}]\times E[\alpha]\rightarrow\mathbb{G}_{m}^{\otimes_{\mathbb{Z}}R}[{\alpha}],
T^α\displaystyle\widehat{T}_{\alpha} :E[α¯]×E/[α]E𝔾mR/(𝔾mR)α,\displaystyle:E[\overline{\alpha}]\times E/[\alpha]E\rightarrow\mathbb{G}_{m}^{\otimes_{\mathbb{Z}}R}/(\mathbb{G}_{m}^{\otimes_{\mathbb{Z}}R})^{{\alpha}},

which are sesquilinear, if RR is commutative, in the sense that for all γ,δR\gamma,\delta\in R and PE[α¯]P\in E[\overline{\alpha}], QEQ\in E,

T^α([γ]P,[δ]Q)=T^α(P,Q)γ¯δ,\widehat{T}_{\alpha}([\gamma]P,[\delta]Q)=\widehat{T}_{\alpha}(P,Q)^{\overline{\gamma}\delta},

and similarly for W^α\widehat{W}_{\alpha}. When RR is non-commutative, a similar construction is possible, but sesquilinearity in one entry is twisted by an action of α¯\overline{\alpha} (Section 4).

In the case that α=n\alpha=n\in\mathbb{Z}, these pairings can be interpreted as a ‘sesquilinearization’ of the usual Weil and Tate-Lichtenbaum pairings. For example if

tn:E[n]×E/[n]E𝔾m/𝔾mnt_{n}:E[n]\times E/[n]E\rightarrow\mathbb{G}_{m}/\mathbb{G}_{m}^{n}

represents the usual Tate-Lichtenbaum pairing, and R=[τ]R=\mathbb{Z}[\tau], then

T^n(P,Q)=(tn(P,Q)2N(τ)tn([τ¯]P,Q)Tr(τ))(tn([τ¯τ]P,Q))τ.\widehat{T}_{n}(P,Q)=\left(t_{n}(P,Q)^{2N(\tau)}t_{n}([-\overline{\tau}]P,Q)^{Tr(\tau)}\right)\left(t_{n}([\overline{\tau}-{\tau}]P,Q)\right)^{\tau}.

In the general case, one can only express T^α\widehat{T}_{\alpha} in terms of tnt_{n} if one computes certain preimages (See Remark 4.5).

We show that these new pairings are non-degenerate in most cases. The pairings are amenable to computation, for example for cryptographic purposes (see Algorithm 5.7).

Both the Tate-Lichtenbaum pairing and Weil pairing have a wide variety of interpretations in terms of cohomology, intersection pairings, Cartier duality, etc. In this paper we take an elementary approach in terms of divisors. However, the new pairings were discovered while revisiting an interpretation of these pairings in terms of the monodromy of the Poincaré biextension studied in the author’s PhD thesis [19]. A companion paper will explain these new pairings in that context, and their relationship with elliptic nets and height pairings.

Notations. Greek letters (α,β,\alpha,\beta,\ldots) generally refer to elements of the ring RR, with the exception of σ\sigma, which is an element of a Galois group, and η\eta and ϵ\epsilon, which are maps in Section 5. Roman letters in lower case (f,g,f,g,\ldots) will generally refer to elements of 𝔾m\mathbb{G}_{m} and capital roman letters (besides RR and EE) typically refer to points of an elliptic curve EE. We use the exponent R\otimes_{\mathbb{Z}}R for the extension of scalars from \mathbb{Z} to RR when viewing an abelian group in multiplicative notation as a \mathbb{Z}-module, as in 𝔾mR\mathbb{G}_{m}^{\otimes_{\mathbb{Z}}R}. Simple tensors are written gαg^{\otimes\alpha}, but we will suppress the \otimes, writing gαg^{\alpha}. Note, however, that we will continue to view this as a left RR-module. Regular exponents will be reserved for the module action of RR and \mathbb{Z} when in a multiplicative notational mode. In particular, we have the slightly counter-intuitive111We opted for this slight dissonance over the available alternatives, which were a switch to additive notation in the multiplicative group, or the use of notation (αx)β=βαx{}^{\beta}(^{\alpha}x)=\,^{\beta\alpha}x.

(xα)β=xβα.(x^{\alpha})^{\beta}=x^{\beta\alpha}.

For this reason we write (𝔾mR)Rα(\mathbb{G}_{m}^{\otimes_{\mathbb{Z}}R})^{R\alpha} for the image of the multiplicative left RR-module 𝔾mR\mathbb{G}_{m}^{\otimes_{\mathbb{Z}}R} under the action of the RR-submodule RαR\alpha, or equivalently, under RαRR\alpha R. We refer to this as the set of α\alpha-powers of 𝔾mR\mathbb{G}_{m}^{\otimes_{\mathbb{Z}}R}. (If α\alpha\in\mathbb{Z}, or more generally the centre of RR, we can simplify the notation from (𝔾mR)Rα(\mathbb{G}_{m}^{\otimes_{\mathbb{Z}}R})^{R\alpha} to (𝔾mR)α(\mathbb{G}_{m}^{\otimes_{\mathbb{Z}}R})^{\alpha}.)

We denote the algebraic closure of a field KK by K¯\overline{K}. We denote the action of an endomorphism αR\alpha\in R on PEP\in E by [α]P[\alpha]P. For an RR-module MM, write M[α]:={mM:αm=0}M[\alpha]:=\{m\in M:\alpha m=0\}. When RR is commutative, this is again an RR-module.

Acknowledgements. The author is grateful to Damien Robert for rekindling her interest through his recent work [15],[16], his interest in the author’s thesis, and several generous discussions, which inspired this work. The author also thanks Joseph Macula, Joseph H. Silverman, and Drew Sutherland for feedback on an earlier draft.

2. Classical pairings

2.1. The Weil pairing

This section follows Miller [12] and Silverman [17, Chap III, §8]. For the more general Weil pairing, see [8], [17, Exercise III.3.15].

Definition 2.1 (Weil pairing: first definition).

Let m>1m>1 be an integer. Let EE be an elliptic curve defined over a field KK which contains the field of definition of E[m]E[m], and with characteristic coprime to mm in the case of positive characteristic. Suppose that P,QE[m]P,Q\in E[m]. Choose divisors DPD_{P} and DQD_{Q} of disjoint support such that

DP(P)(𝒪),DQ(Q)(𝒪).D_{P}\sim(P)-(\mathcal{O}),\qquad D_{Q}\sim(Q)-(\mathcal{O}).

Then mDPmDQ0mD_{P}\sim mD_{Q}\sim 0, hence there are functions fPf_{P} and fQf_{Q} such that

div(fP)=mDP,div(fQ)=mDQ.\operatorname{div}(f_{P})=mD_{P},\qquad\operatorname{div}(f_{Q})=mD_{Q}.

The Weil pairing

em:E[m]×E[m]μme_{m}:E[m]\times E[m]\rightarrow\mu_{m}

is defined by

em(P,Q)=fP(DQ)fQ(DP).e_{m}(P,Q)=\frac{f_{P}(D_{Q})}{f_{Q}(D_{P})}.

For example, we can choose DPD_{P} and DQD_{Q} disjoint as follows: first choose some TT such that T{𝒪,P,Q,QP}T\not\in\{\mathcal{O},-P,Q,Q-P\}. Then set DP=(P+T)(T)D_{P}=(P+T)-(T) and DQ=(Q)(𝒪)D_{Q}=(Q)-(\mathcal{O}). Set the notation fm,Xf_{m,X} for the rational function with divisor m(X)m(𝒪)m(X)-m(\mathcal{O}). Then,

em(P,Q)=fP(DQ)fQ(DP)=fP(Q)fQ(T)fP(𝒪)fQ(P+T)=fm,P(QT)fm,Q(T)fm,P(T)fm,Q(P+T).e_{m}(P,Q)=\frac{f_{P}(D_{Q})}{f_{Q}(D_{P})}=\frac{f_{P}(Q)f_{Q}(T)}{f_{P}(\mathcal{O})f_{Q}(P+T)}=\frac{f_{m,P}(Q-T)f_{m,Q}(T)}{f_{m,P}(-T)f_{m,Q}(P+T)}.
Definition 2.2 (Weil pairing: second definition).

Let ϕ:EE\phi:E\rightarrow E^{\prime} be an isogeny between elliptic curves defined over a perfect field KK which contains the field of definition of ker(ϕ)\ker(\phi) and ker(ϕ^)\ker(\widehat{\phi}), and with characteristic coprime to degϕ\deg\phi in the case of positive characteristic. Suppose that Pkerϕ^P\in\ker\widehat{\phi}, and QkerϕQ\in\ker{\phi}. Let gPg_{P} be a rational function with principal divisor

div(gP)=ϕ((P)(𝒪)).\operatorname{div}(g_{P})=\phi^{*}((P)-(\mathcal{O})).

(In the case that ϕ=[m]\phi=[m], this implies gPm=fm,P[m]g_{P}^{m}=f_{m,P}\circ[m].) The Weil pairing

eϕ:kerϕ^×kerϕμme_{\phi}:\ker\widehat{\phi}\times\ker\phi\rightarrow\mu_{m}

where mm is any positive integer with kerϕE[m]\ker\phi\subseteq E[m], and μm\mu_{m} denotes the mm-th roots of unity, is defined by

eϕ(P,Q)=gP(X+Q)gP(X),e_{\phi}(P,Q)=\frac{g_{P}(X+Q)}{g_{P}(X)},

where XX is any auxiliary point chosen disjoint from the supports of gPg_{P} and gPtQg_{P}\circ t_{Q} (the function gPg_{P} precomposed with translation by QQ).

The above definition generalizes naturally to a pairing associated to an isogeny; taking the isogeny to be the multiplication-by-m map [m][m] recovers the mm-Weil pairing.

The standard properties are as follows.

Proposition 2.3.

Suppose mm is coprime to char(K)\operatorname{char}(K) in the case of positive characteristic. Definitions 2.1 and 2.2 are well-defined, equal, and have the following properties (restricting to the case ϕ=[m]\phi=[m] for the first definition):

  1. (1)

    Bilinearity: for P,P1,P2kerϕ^P,P_{1},P_{2}\in\ker\widehat{\phi} and Q,Q1,Q2kerϕQ,Q_{1},Q_{2}\in\ker\phi,

    eϕ(P1+P2,Q)\displaystyle e_{\phi}(P_{1}+P_{2},Q) =eϕ(P1,Q)eϕ(P2,Q),\displaystyle=e_{\phi}(P_{1},Q)e_{\phi}(P_{2},Q),
    eϕ(P,Q1+Q2)\displaystyle e_{\phi}(P,Q_{1}+Q_{2}) =eϕ(P,Q1)eϕ(P,Q2).\displaystyle=e_{\phi}(P,Q_{1})e_{\phi}(P,Q_{2}).
  2. (2)

    Alternating: for PE[m]P\in E[m],

    em(P,P)=1.e_{m}(P,P)=1.
  3. (3)

    Skew-symmetry: for Pkerϕ^P\in\ker\widehat{\phi} and QkerϕQ\in\ker\phi,

    eϕ(P,Q)=eϕ^(Q,P)1.e_{\phi}(P,Q)=e_{\widehat{\phi}}(Q,P)^{-1}.
  4. (4)

    Non-degeneracy: for nonzero PE[m](K¯)P\in E[m](\overline{K}), there exists QE[m](K¯)Q\in E[m](\overline{K}) such that

    em(P,Q)1.e_{m}(P,Q)\neq 1.
  5. (5)

    Coherence: for Pkerϕ^ψ^P\in\ker\widehat{\phi}\circ\widehat{\psi}, and QkerϕQ\in\ker\phi,

    eψϕ(P,Q)=eϕ(ψ^P,Q).e_{\psi\circ\phi}(P,Q)=e_{\phi}(\widehat{\psi}P,Q).

    and for Pkerψ^P\in\ker\widehat{\psi}, and QkerψϕQ\in\ker\psi\circ\phi,

    eψϕ(P,Q)=eψ(P,ϕQ).e_{\psi\circ\phi}(P,Q)=e_{\psi}(P,\phi Q).
  6. (6)

    Compatibility: For mm-torsion points PP and QQ,

    em(ϕ^P,Q)=em(P,ϕQ).e_{m}(\widehat{\phi}P,Q)=e_{m}(P,\phi Q).
  7. (7)

    Galois invariance: for P,QE[m]P,Q\in E[m], and σGal(K¯/K)\sigma\in\operatorname{Gal}(\overline{K}/K),

    em(P,Q)σ=em(Pσ,Qσ).e_{m}(P,Q)^{\sigma}=e_{m}(P^{\sigma},Q^{\sigma}).
Proof.

For example, see [19, Chapter 16], [15], [2, Sec 3.1]. ∎

For elliptic curves over \mathbb{C}, the Weil pairing can be interpreted as a determinant, or an intersection pairing; see [6]. The Weil pairing also arises from the Cartier duality of the kernels of an isogeny and its dual; see Mumford [14, IV.§20, p.183-5] and Milne [13, §11,16].

2.2. The Tate-Lichtenbaum pairing

Another pairing intimately related to the Weil pairing is the Tate-Lichtenbaum pairing. This pairing was first defined by Tate [20] for abelian varieties over pp-adic number fields in 1958. In 1959, Lichtenbaum defined a pairing on Jacobian varieties and showed that it coincided with the pairing of Tate [10]. The pairing was introduced to cryptography by Frey and Rück [4]. Descriptions can be found in Silverman [17, VIII.2, X.1] and Duquesne-Frey [3]. For our version here, see for example [5].

Definition 2.4.

Let m>1m>1 be an integer. Let EE be an elliptic curve defined over a field KK. Suppose that PE(K)[m]P\in E(K)[m]. Choose divisors DPD_{P} and DQD_{Q} of disjoint support such that

DP(P)(𝒪),DQ(Q)(𝒪).D_{P}\sim(P)-(\mathcal{O}),\qquad D_{Q}\sim(Q)-(\mathcal{O}).

Then mDP0mD_{P}\sim 0, hence there is a function fPf_{P} such that

div(fP)=mDP.\operatorname{div}(f_{P})=mD_{P}.

The Tate-Lichtenbaum pairing

tm:E(K)[m]×E(K)/mE(K)K/(K)mt_{m}:E(K)[m]\times E(K)/mE(K)\rightarrow K^{*}/(K^{*})^{m}

is defined by

tm(P,Q)=fP(DQ).t_{m}(P,Q)=f_{P}(D_{Q}).
Proposition 2.5.

Definition 2.4 is well-defined, and has the following properties:

  1. (1)

    Bilinearity: for P,PE(K)[m]P,P^{\prime}\in E(K)[m] and Q,QE(K)Q,Q^{\prime}\in E(K)

    tm(P+P,Q)\displaystyle t_{m}(P+P^{\prime},Q) =tm(P,Q)tm(P,Q),\displaystyle=t_{m}(P,Q)t_{m}(P^{\prime},Q),
    tm(P,Q+Q)\displaystyle t_{m}(P,Q+Q^{\prime}) =tm(P,Q)tm(P,Q).\displaystyle=t_{m}(P,Q)t_{m}(P,Q^{\prime}).
  2. (2)

    Non-degeneracy: Let KK be a finite field containing the mm-th roots of unity μm\mu_{m}. For nonzero PE(K)[m]P\in E(K)[m], there exists QE(K)Q\in E(K) such that

    tm(P,Q)1.t_{m}(P,Q)\neq 1.

    Furthermore, for QE(K)\mE(K)Q\in E(K)\backslash mE(K), there exists PE(K)[m]P\in E(K)[m] such that

    tm(P,Q)1.t_{m}(P,Q)\neq 1.
  3. (3)

    Compatibility: For an mm-torsion point PEP\in E, an isogeny ϕ:EE\phi:E\rightarrow E^{\prime}, and a point QEQ\in E^{\prime},

    tm(ϕ^P,Q)=tm(P,ϕQ).t_{m}(\widehat{\phi}P,Q)=t_{m}(P,\phi Q).
  4. (4)

    Galois invariance: for P,QE[m]P,Q\in E[m], and σGal(K¯/K)\sigma\in\operatorname{Gal}(\overline{K}/K),

    tm(P,Q)σ=tm(Pσ,Qσ).t_{m}(P,Q)^{\sigma}=t_{m}(P^{\sigma},Q^{\sigma}).
Proof.

See for example [19, Chapter 16], [15] and [2, Sec 3.2]. ∎

Remark 2.6.

For purposes such as cryptography, where we wish to compare values of the Tate-Lichtenbaum pairing, it is typical to apply a final exponentiation by (q1)/m(q-1)/m in order to obtain values in μm\mu_{m}.

Including this final exponentiation, there is a more general notion of Tate pairing associated to a KK-rational isogeny ϕ:EE\phi:E\rightarrow E^{\prime} when K=𝔽qK=\mathbb{F}_{q}, that is,

tϕ:kerϕ^(K)×E(K)/ϕE(K)μm,t_{\phi}:\ker\widehat{\phi}(K)\times E^{\prime}(K)/\phi E(K)\rightarrow\mu_{m},

where mm is any positive integer so that kerϕE[m]E[q1]\ker\phi\subseteq E[m]\subseteq E[q-1]. This generalizes the definition above when ϕ=[m]\phi=[m], and can be given by

tϕ(P,Q)=eϕ(πq(T)T,P),t_{\phi}(P,Q)=e_{\phi}(\pi_{q}(T)-T,P),

where TT is an arbitrarily chosen ϕ\phi-preimage of QQ, πq\pi_{q} is the qq-power Frobenius, and eϕe_{\phi} is the Weil pairing. It has the property that its values agree with those of tmq1mt_{m}^{\frac{q-1}{m}} on the common codomain; in other words, it is a restriction. See [1], [15] and [2, Sec 3.2]; see also [8].

3. The calculus of RR-divisors

Let RR be an order in an imaginary quadratic field or quaternion algebra. Such a ring RR comes equipped with an involution which we term conjugation, denoted αα¯\alpha\mapsto\overline{\alpha}. In the quaternion algebra case, this is order reversing: αβ¯=β¯α¯\overline{\alpha\beta}=\overline{\beta}\overline{\alpha}.

Let EE be an elliptic curve with divisor group Div(E)\operatorname{Div}(E). We extend common notions from Div(E)\operatorname{Div}(E) to RDiv(E)R\otimes_{\mathbb{Z}}\operatorname{Div}(E). We emphasize that in this section we make no assumption that EE has complex multiplication.

In what follows we choose an integral basis: write R=[τi]:=iτiR=\mathbb{Z}[\tau_{i}]:=\sum_{i}\tau_{i}\mathbb{Z}, where τ0=1\tau_{0}=1 and we let ii range in {0,1}\{0,1\} or {0,1,2,3}\{0,1,2,3\} according to the rank r{2,4}r\in\{2,4\} of RR. When we sum over ii the range will be understood in context.

3.1. RR-divisors

We define DivR(E):=RDiv(E)\operatorname{Div}_{R}(E):=R\otimes_{\mathbb{Z}}\operatorname{Div}(E) to be the RR-module generated by all symbols (P)(P), where PP is a point of EE, i.e. finite formal RR-linear combinations PαP(P)\sum_{P}\alpha_{P}(P) of such symbols, which we call RR-divisors. (We will frequently suppress the \otimes for notational simplicity.) Then DivR(E)\operatorname{Div}_{R}(E) is an RR-module under the action α(βD)=αβD\alpha\cdot(\beta\otimes D)=\alpha\beta\otimes D. A divisor PαP(P)\sum_{P}\alpha_{P}(P) is of degree 0 if PαP=0\sum_{P}\alpha_{P}=0 in RR; these form a sub-RR-module DivR0(E)RDiv0(E)\operatorname{Div}^{0}_{R}(E)\cong R\otimes_{\mathbb{Z}}\operatorname{Div}^{0}(E).

In the presence of a preferred integral basis τi\tau_{i} for RR, we can write a sum over ii:

P(imi,Pτi)(P)=iτi(Pmi,P(P)).\sum_{P}\left(\sum_{i}m_{i,P}\tau_{i}\right)(P)=\sum_{i}\tau_{i}\left(\sum_{P}m_{i,P}(P)\right).

We say that a divisor of degree zero

D=iτiDi,DiDiv(E),D=\sum_{i}\tau_{i}D_{i},\quad D_{i}\in\operatorname{Div}(E),

is principal if DiD_{i}, i=0,,r1i=0,\ldots,r-1, are all principal in Div(E)\operatorname{Div}(E). We see that the principal divisors form a sub-RR-module and we define PicR(E)\operatorname{Pic}_{R}(E) and PicR0(E)\operatorname{Pic}_{R}^{0}(E) to be the RR-module quotient of DivR(E)\operatorname{Div}_{R}(E) and DivR0(E)\operatorname{Div}_{R}^{0}(E) by the principal divisors. Observe that being principal in PicR0(E)\operatorname{Pic}_{R}^{0}(E) is independent of basis, and that PicR(E)RPic(E)\operatorname{Pic}_{R}(E)\cong R\otimes_{\mathbb{Z}}\operatorname{Pic}(E), PicR0(E)RPic0(E)\operatorname{Pic}_{R}^{0}(E)\cong R\otimes_{\mathbb{Z}}\operatorname{Pic}^{0}(E).

If f=fiτi(K(E))Rf=\prod f_{i}^{\tau_{i}}\in(K(E)^{*})^{\otimes_{\mathbb{Z}}R}, we define

div(f)=iτidiv(fi).\operatorname{div}(f)=\sum_{i}\tau_{i}\operatorname{div}(f_{i}).

Thus principal divisors are those which are divisors of f(K(E))Rf\in(K(E)^{*})^{\otimes_{\mathbb{Z}}R}.

We define the usual push-foward and pull-back operations on divisors by extending RR-linearly. Suppose ϕ:EE\phi:E\rightarrow E^{\prime}. Then

ϕ(iτiDi)=iτiϕDi,ϕ(iτiDi)=iτiϕDi.\phi^{*}\left(\sum_{i}\tau_{i}D_{i}\right)=\sum_{i}\tau_{i}\phi^{*}D_{i},\quad\phi_{*}\left(\sum_{i}\tau_{i}D_{i}\right)=\sum_{i}\tau_{i}\phi_{*}D_{i}.

These inherit the usual desired properties:

  1. (1)

    ϕϕD=(degϕ)D\phi_{*}\phi^{*}D=(\deg\phi)D

  2. (2)

    ϕdiv(f)=div(ϕf)\phi^{*}\operatorname{div}(f)=\operatorname{div}(\phi^{*}f), ϕdiv(f)=div(ϕf)\phi_{*}\operatorname{div}(f)=\operatorname{div}(\phi_{*}f)

  3. (3)

    (ϕψ)=ϕψ(\phi\circ\psi)_{*}=\phi_{*}\psi_{*}, (ϕψ)=ψϕ(\phi\circ\psi)^{*}=\psi^{*}\phi^{*}

where we define ϕfiτi=(ϕfi)τi\phi_{*}\prod f_{i}^{\tau_{i}}=\prod(\phi_{*}f_{i})^{\tau_{i}} and ϕfiτi=(ϕfi)τi\phi^{*}\prod f_{i}^{\tau_{i}}=\prod(\phi^{*}f_{i})^{\tau_{i}}.

We also have a Galois action: (iτiDi)σ=iτiDiσ\left(\sum_{i}\tau_{i}D_{i}\right)^{\sigma}=\sum_{i}\tau_{i}D_{i}^{\sigma} for σGal(K¯/K)\sigma\in\operatorname{Gal}(\overline{K}/K).

For a divisor D=nP(P)Div(E)D=\sum n_{P}(P)\in\operatorname{Div}(E), nPn_{P}\in\mathbb{Z}, we define

DΣ:=[nP]PE.D^{\Sigma}:=\sum[n_{P}]P\in E.

Viewing EE as a \mathbb{Z}-module, we obtain an RR-module RER\otimes_{\mathbb{Z}}E. Then we have an RR-module isomorphism

PicR0(E)RE,τiDiτiDiΣ.\operatorname{Pic}_{R}^{0}(E)\cong R\otimes_{\mathbb{Z}}E,\quad\sum\tau_{i}D_{i}\mapsto\sum\tau_{i}\otimes D_{i}^{\Sigma}.

To show this is an isomorphism, we need to check that it is injective (surjectivity is clear). If D=iτiDi𝒪D=\sum_{i}\tau_{i}D_{i}\mapsto\mathcal{O} then DiΣ=𝒪D_{i}^{\Sigma}=\mathcal{O} for all ii, so DD is principal. In fact, an inverse is given by

τiPiiτi((Pi)(𝒪)).\sum\tau_{i}\otimes P_{i}\mapsto\sum_{i}\tau_{i}((P_{i})-(\mathcal{O})).

3.2. Evaluation of functions at divisors

Let 𝔾m\mathbb{G}_{m} be the multiplicative group. Then 𝔾mR\mathbb{G}_{m}^{\otimes_{\mathbb{Z}}R} is an RR-module whose action is written multiplicatively as αx=xα=xα\alpha\cdot x=x^{\otimes\alpha}=x^{\alpha}. As a reminder, the action is still a left action, so

(giτi)α=giατi\left({\prod g_{i}^{\tau_{i}}}\right)^{\alpha}=\prod g_{i}^{{\alpha\tau_{i}}}

It also has a conjugation which will be useful:

giτi¯:=giτi¯.\overline{\prod g_{i}^{\tau_{i}}}:=\prod g_{i}^{\overline{\tau_{i}}}.

Similarly, (K(E))R(K(E)^{*})^{\otimes_{\mathbb{Z}}R} has a left RR-module structure and conjugation. Therefore, for f(K(E))Rf\in(K(E)^{*})^{\otimes_{\mathbb{Z}}R}, we let

(1) div(fα)=αdiv(f),\operatorname{div}(f^{\alpha})={\alpha}\cdot\operatorname{div}(f),

so that div\operatorname{div} becomes an RR-module homomorphism.

We define evaluation of f=fiτik(E)Rf=\prod f_{i}^{\tau_{i}}\in k(E)^{\otimes_{\mathbb{Z}}R} at DDiv(E)D\in\operatorname{Div}(E) as f(D)=fi(D)τif(D)=\prod f_{i}(D)^{\tau_{i}}, and extend to DDivR(E)D\in\operatorname{Div}_{R}(E) by defining

f(αD)=f(D)α¯.f(\alpha\cdot D)=f(D)^{\overline{\alpha}}.

This definition requires that the supports of DD and div(f)\operatorname{div}(f) are disjoint. Observe the vinculum222Thank you to my brother and Wikipedia for teaching me this term for an \overline\backslash\text{overline}., which reflects the duality between ff and DD. Among other things, it allows for the two left RR-actions to communicate in the non-commutative setting:

f(αβD)=f(βD)α¯=f(D)β¯α¯=f(D)αβ¯.f(\alpha\beta\cdot D)=f(\beta\cdot D)^{\overline{\alpha}}=f(D)^{\overline{\beta}\overline{\alpha}}=f(D)^{\overline{\alpha\beta}}.

3.3. Weil reciprocity

A variation of Weil reciprocity ([9, Chapter VI, Corollary to Theorem 10]) holds for us:

Theorem 3.1.

Let f,g(K(E))Rf,g\in(K(E)^{*})^{\otimes_{\mathbb{Z}}R}. Then

f(div(g))=g(div(f))¯.f(\operatorname{div}(g))=\overline{g(\operatorname{div}(f))}.
Proof.

The proof relies on Weil reciprocity for Div(E)\operatorname{Div}(E). Suppose f=ifiτif=\prod_{i}f_{i}^{\tau_{i}} and g=jgjτjg=\prod_{j}g_{j}^{\tau_{j}}. We have

f(div(g))\displaystyle f(\operatorname{div}(g)) =ifi(div(g))τi=ijfi(div(gj))τj¯τi=ijgj(div(fi))τj¯τi\displaystyle=\prod_{i}f_{i}(\operatorname{div}(g))^{\tau_{i}}=\prod_{ij}f_{i}(\operatorname{div}(g_{j}))^{\overline{\tau_{j}}\tau_{i}}=\prod_{ij}g_{j}(\operatorname{div}(f_{i}))^{\overline{\tau_{j}}\tau_{i}}
=ijgj(div(fi))τi¯τj¯=jgj(div(f))τj¯=g(div(f))¯.\displaystyle=\overline{\prod_{ij}g_{j}(\operatorname{div}(f_{i}))^{\overline{\tau_{i}}\tau_{j}}}=\overline{\prod_{j}g_{j}(\operatorname{div}(f))^{\tau_{j}}}=\overline{g(\operatorname{div}(f))}.

4. Sesquilinear pairings

If RR is commutative, an RR-sesquilinear pairing is a bilinear pairing ,\langle\cdot,\cdot\rangle on a pair of RR-modules, taking values in another RR-module, that satisfies

αx,βy=αβ¯x,y, for all α,βR.\langle\alpha x,\beta y\rangle={\alpha}{\overline{\beta}}\cdot\langle x,y\rangle,\text{ for all }\alpha,\beta\in R.

For the non-commutative case, we need to add a type of twisting. Recall that RR is a maximal order in a division algebra. Thus we can set the notation Rγ:=γ1RγRR_{\gamma}:=\gamma^{-1}R\gamma\cap R, a subring of RR. For γR\gamma\in R and δRγ\delta\in R_{\gamma}, let δ(γ)\delta^{(\gamma)} be defined as that element of RR which satisfies δ(γ)γ=γδ\delta^{(\gamma)}\gamma=\gamma\delta. For us, a γ\gamma-twisted RR-sesquilinear pairing is a bilinear pairing ,\langle\cdot,\cdot\rangle on a pair of modules, the first an RγR_{\gamma}-module and the second an RR-module, taking values in another RR-module, that satisfies

αx,βy=β¯α(γ¯)x,y, for all αRγ,βR.\langle\alpha x,\beta y\rangle={\overline{\beta}}\;{{\alpha}^{(\overline{\gamma})}}\cdot\langle x,y\rangle,\text{ for all }\alpha\in R_{\gamma},\beta\in R.

Observe that for rank 22, commutativity implies δ(γ)=δ\delta^{(\gamma)}=\delta and Rγ=RR_{\gamma}=R, so the γ\gamma-twisting is vacuous, and we recover sesquilinear pairings in the traditional sense.

4.1. Generalization of Tate-Lichtenbaum pairing

For each αR\alpha\in R, we define an α\alpha-twisted RR-sesquilinear pairing generalizing the Tate-Lichtenbaum pairing:

Tα:PicR0(E)[α¯]×PicR0(E)/RαPicR0(E)𝔾mR/(𝔾mR)Rα¯,T_{\alpha}:\operatorname{Pic}_{R}^{0}(E)[\overline{\alpha}]\times\operatorname{Pic}_{R}^{0}(E)/R\alpha\operatorname{Pic}_{R}^{0}(E)\rightarrow\mathbb{G}_{m}^{\otimes_{\mathbb{Z}}R}/(\mathbb{G}_{m}^{\otimes_{\mathbb{Z}}R})^{R\overline{\alpha}},

by

Tα(DP,DQ)=fP(DQ) where div(fP)=α¯DP,T_{\alpha}(D_{P},D_{Q})=f_{P}(D_{Q})\quad\text{ where }\quad\operatorname{div}(f_{P})=\overline{\alpha}\cdot D_{P},

where DPD_{P} and DQD_{Q} are chosen to have disjoint support. Observe that PicR0(E)[α¯]\operatorname{Pic}_{R}^{0}(E)[\overline{\alpha}] is an RR-module when RR is commutative, but in general we can only assume it is an Rα¯R_{\overline{\alpha}}-module. Also, we use RαPicR0(E)R\alpha\operatorname{Pic}_{R}^{0}(E) since αPicR0(E)\alpha\operatorname{Pic}_{R}^{0}(E) may not be an RR-module in the non-commutative case.

To satisfy the condition on supports, observe that for any divisor DPicR0(E)D\in\operatorname{Pic}_{R}^{0}(E), there exist points P0,,Pr1EP_{0},\ldots,P_{r-1}\in E so that

(2) Diτi((Pi+S)(S))D\sim\sum_{i}\tau_{i}((P_{i}+S)-(S))

for any auxiliary point SES\in E. In particular, if P0,,Pr1P_{0},\ldots,P_{r-1} are such that DPiτi((Pi)(𝒪))D_{P}\sim\sum_{i}\tau_{i}((P_{i})-(\mathcal{O})), and

α¯τi=jαjiτj,\overline{\alpha}\tau_{i}=\sum_{j}\alpha_{ji}\tau_{j},

then we can take fP=ifiτi(K(E))Rf_{P}=\prod_{i}f_{i}^{\tau_{i}}\in(K(E)^{*})^{\otimes_{\mathbb{Z}}R}, where

(3) div(fi)=j=0r1αij(Pj)(j=0r1αij)(𝒪),\operatorname{div}(f_{i})=\sum_{j=0}^{r-1}\alpha_{ij}(P_{j})-\left(\sum_{j=0}^{r-1}\alpha_{ij}\right)(\mathcal{O}),

and then by a judicious choice of DQD_{Q} (choosing SS in the linearly equivalent form (2)), we can satisfy the condition on disjoint supports.

Remark 4.1.

The equations (3) allow for a Miller-style algorithm to compute this pairing [11] [7, §26.3.1]. This is polynomial time in the coefficients of the minimal polynomial of α\alpha. For example, if RR has basis 11 and τ\tau, and DP=((P0)(𝒪))+τ((P1)(𝒪))D_{P}=\left((P_{0})-(\mathcal{O})\right)+\tau\cdot\left((P_{1})-(\mathcal{O})\right), and

α¯=a+cτ,α¯τ=b+dτ,a,b,c,d,\overline{\alpha}=a+c\tau,\quad\overline{\alpha}\tau=b+d\tau,\quad a,b,c,d\in\mathbb{Z},

then fP=f0f1τ(K(E))Rf_{P}=f_{0}f_{1}^{\tau}\in(K(E)^{*})^{\otimes_{\mathbb{Z}}R}, where

(4) div(f0)=a(P0)+b(P1)(a+b)(𝒪),div(f1)=c(P0)+d(P1)(c+d)(𝒪).\operatorname{div}(f_{0})=a(P_{0})+b(P_{1})-(a+b)(\mathcal{O}),\quad\operatorname{div}(f_{1})=c(P_{0})+d(P_{1})-(c+d)(\mathcal{O}).

More details are given for the CM case in Algorithm 5.7.

Theorem 4.2.

The pairing defined above is well-defined, bilinear, and satisfies

  1. (1)

    Twisted sesquilinearity: For γRα¯\gamma\in R_{\overline{\alpha}} and δR\delta\in R,

    Tα(γDP,δDQ)=Tα(DP,DQ)δ¯γ(α¯).{T}_{\alpha}(\gamma\cdot D_{P},\delta\cdot D_{Q})={T}_{\alpha}(D_{P},D_{Q})^{{\overline{\delta}}\;{\gamma}^{(\overline{\alpha})}}.
  2. (2)

    Compatibility: Let ϕ:EE\phi:E\rightarrow E^{\prime}. Then

    Tα(ϕDP,ϕDQ)=Tα(DP,DQ)degϕ.T_{\alpha}(\phi_{*}D_{P},\phi_{*}D_{Q})=T_{\alpha}(D_{P},D_{Q})^{\deg\phi}.
  3. (3)

    Coherence: Suppose DPPicR0(E)[βα¯]D_{P}\in\operatorname{Pic}_{R}^{0}(E)[\overline{\beta\alpha}], and DQPicR0(E)/RβαPicR0(E)D_{Q}\in\operatorname{Pic}_{R}^{0}(E)/R\beta\alpha\operatorname{Pic}_{R}^{0}(E). Then

    Tβα(DP,DQ)mod(𝔾mR)Rα¯=Tα(β¯DP,DQmodRαPicR0(E)).T_{\beta\alpha}(D_{P},D_{Q})\bmod{(\mathbb{G}_{m}^{\otimes_{\mathbb{Z}}R})^{R\overline{\alpha}}}=T_{\alpha}(\overline{\beta}\cdot D_{P},D_{Q}\bmod R\alpha\operatorname{Pic}_{R}^{0}(E)).

    Suppose DPPicR0(E)[β¯]D_{P}\in\operatorname{Pic}_{R}^{0}(E)[\overline{\beta}], and DQPicR0(E)/RβαPicR0(E)D_{Q}\in\operatorname{Pic}_{R}^{0}(E)/R\beta\alpha\operatorname{Pic}_{R}^{0}(E). Then

    Tβα(DP,DQ)mod(𝔾mR)Rβ¯=Tβ(DP,αDQmodRβPicR0(E)).T_{\beta\alpha}(D_{P},D_{Q})\bmod{(\mathbb{G}_{m}^{\otimes_{\mathbb{Z}}R})^{R\overline{\beta}}}=T_{\beta}(D_{P},\alpha\cdot D_{Q}\bmod R\beta\operatorname{Pic}_{R}^{0}(E)).
  4. (4)

    Galois invariance: Suppose EE is defined over a field KK. Let σGal(K¯/K)\sigma\in\operatorname{Gal}(\overline{K}/K). Then

    Tα(DP,DQ)σ=Tα(DPσ,DQσ).T_{\alpha}(D_{P},D_{Q})^{\sigma}=T_{\alpha}(D_{P}^{\sigma},D_{Q}^{\sigma}).
Proof.

Choice of representative DQD_{Q} in the divisor class: Suppose DQDQD_{Q}\sim D_{Q}^{\prime}. Then for some g(K(E))Rg\in(K(E)^{*})^{\otimes_{\mathbb{Z}}R}, having divisor div(g)=DQDQ\operatorname{div}(g)=D_{Q}-D_{Q}^{\prime}, and using Weil reciprocity333There’s a subtlety here. Observe that (gβ)α¯=gαβ¯=gαβ¯=gβ¯α¯=gα¯(β¯)β¯=(gβ¯)α¯(β¯)\overline{(g^{\beta})^{\alpha}}=\overline{g^{\alpha\beta}}=g^{\overline{\alpha\beta}}=g^{\overline{\beta}\;\overline{\alpha}}=g^{\overline{\alpha}^{(\overline{\beta})}\overline{\beta}}=(g^{\overline{\beta}})^{\overline{\alpha}^{(\overline{\beta})}}, so that it is only in the case that RR is commutative that gα¯=g¯α¯\overline{g^{\alpha}}=\overline{g}^{\overline{\alpha}}. However, it is still true that gα¯(𝔾mR)Rα¯\overline{g^{\alpha}}\in(\mathbb{G}_{m}^{\otimes_{\mathbb{Z}}R})^{R\overline{\alpha}}. (Theorem 3.1),

fP(DQ)fP(DQ)1=fP(div(g))=g(div(fP))¯=g(α¯DP)¯=g(DP)α¯(𝔾mR)Rα¯.f_{P}(D_{Q})f_{P}(D_{Q}^{\prime})^{-1}=f_{P}(\operatorname{div}(g))=\overline{g(\operatorname{div}(f_{P}))}=\overline{g(\overline{\alpha}\cdot D_{P})}=\overline{g(D_{P})^{{\alpha}}}\in(\mathbb{G}_{m}^{\otimes_{\mathbb{Z}}R})^{R\overline{\alpha}}.

Choice of DQD_{Q} modulo RαPicR0(E)R\alpha\operatorname{Pic}_{R}^{0}(E):

fP(DQ+γαD)=fP(DQ)fP(D)α¯γ¯.f_{P}(D_{Q}+\gamma\alpha\cdot D^{\prime})=f_{P}(D_{Q})f_{P}(D^{\prime})^{\overline{\alpha}\;\overline{\gamma}}.

Choice of representative DPD_{P} in the divisor class: Suppose DPDPD_{P}\sim D_{P}^{\prime}. Notice that if we let div(fP)=α¯DP\operatorname{div}(f_{P})=\overline{\alpha}\cdot D_{P} and div(fP)=α¯DP\operatorname{div}(f_{P}^{\prime})=\overline{\alpha}\cdot D_{P}^{\prime}, then

div(fP)=div(fP)+α¯(DPDP).\operatorname{div}(f_{P}^{\prime})=\operatorname{div}(f_{P})+\overline{\alpha}\cdot(D_{P}-D_{P}^{\prime}).

Hence fP=fPgα¯f_{P}^{\prime}=f_{P}g^{\overline{\alpha}} where div(g)=DPDP\operatorname{div}(g)=D_{P}-D_{P}^{\prime}, which is principal by assumption. Then

fP(DQ)=fP(DQ)g(DQ)α¯.f_{P}^{\prime}(D_{Q})=f_{P}(D_{Q})g(D_{Q})^{\overline{\alpha}}.

Choice of fPf_{P}: Any two choices of fPf_{P} differ by a constant scalar, but DQD_{Q} has degree 0 by assumption, so the constant cancels in the formula fP(DQ)f_{P}(D_{Q}).

Bilinearity: Let DPD_{P}, DPDivR0(E)[α¯]D_{P}^{\prime}\in\operatorname{Div}^{0}_{R}(E)[\overline{\alpha}] and div(fP)=α¯DP\operatorname{div}(f_{P})=\overline{\alpha}\cdot D_{P}, div(fP)=α¯DP\operatorname{div}(f_{P}^{\prime})=\overline{\alpha}\cdot D_{P}^{\prime}. Then

Tα(DP+DP,DQ)=fP(DQ)fP(DQ)=Tα(DP,DQ)Tα(DP,DQ).{T}_{\alpha}(D_{P}+D_{P}^{\prime},D_{Q})=f_{P}(D_{Q})f_{P}^{\prime}(D_{Q})={T}_{\alpha}(D_{P},D_{Q}){T}_{\alpha}(D_{P}^{\prime},D_{Q}).

In the other factor,

Tα(DP,DQ+DQ)=fP(DQ+DQ)=fP(DQ)fP(DQ)=Tα(DP,DQ)Tα(DP,DQ).{T}_{\alpha}(D_{P},D_{Q}+D_{Q}^{\prime})=f_{P}(D_{Q}+D_{Q}^{\prime})=f_{P}(D_{Q})f_{P}(D_{Q}^{\prime})={T}_{\alpha}(D_{P},D_{Q}){T}_{\alpha}(D_{P},D_{Q}^{\prime}).

Twisted sesquilinearity: Suppose fPf_{P} has divisor α¯DP\overline{\alpha}\cdot D_{P}. In evaluating Tα(γDP,δDQ)T_{\alpha}(\gamma\cdot D_{P},\delta\cdot D_{Q}), we evaluate the function with divisor α¯γDP=γ(α¯)α¯DP\overline{\alpha}\cdot\gamma\cdot D_{P}=\gamma^{(\overline{\alpha})}\cdot\overline{\alpha}\cdot D_{P} at the divisor δDQ\delta\cdot D_{Q}. Since div(fPμ)=μdiv(fP)\operatorname{div}(f_{P}^{{\mu}})=\mu\cdot\operatorname{div}(f_{P}) by (1), this becomes

fP(δDQ)γ(α¯)=fP(DQ)δ¯γ(α¯).f_{P}(\delta\cdot D_{Q})^{{\gamma}^{(\overline{\alpha})}}=f_{P}(D_{Q})^{\overline{\delta}\;{\gamma}^{(\overline{\alpha})}}.

Compatibility: Observe that α¯ϕDP=ϕ(α¯DP)\overline{\alpha}\cdot\phi_{*}D_{P}=\phi_{*}(\overline{\alpha}\cdot D_{P}). Therefore, in the computation of Tα(ϕDP,ϕDQ)T_{\alpha}(\phi_{*}D_{P},\phi_{*}D_{Q}), we evaluate ϕfP\phi_{*}f_{P} at ϕDQ\phi_{*}D_{Q}. We have

ϕfP(ϕDQ)=fP(ϕϕDQ)=fP(DQ)degϕ,\phi_{*}f_{P}(\phi_{*}D_{Q})=f_{P}(\phi^{*}\phi_{*}D_{Q})=f_{P}(D_{Q})^{\deg\phi},

where the last equality depends upon the fact that ϕϕD(degϕ)D\phi^{*}\phi_{*}D\sim(\deg\phi)D for DPicR0(E)D\in\operatorname{Pic}^{0}_{R}(E).

Coherence: Both statements follow immediately from the definitions.

Galois invariance: This is immediate, since by our definition of the actions of RR on the various entities involved, we have (γD)σ=γDσ(\gamma\cdot D)^{\sigma}=\gamma\cdot D^{\sigma} for any γR\gamma\in R. ∎

Remark 4.3.

In cryptographic applications, we typically restrict to inputs defined over a field 𝔽q\mathbb{F}_{q}. If RR is commutative, to obtain canonical representatives of the codomain, it may be useful to post-compose with a map

(𝔽q)R/((𝔽q)R)α¯μα¯:={uμN(α)R(𝔽q)R:uα¯=1},(\mathbb{F}_{q}^{*})^{\otimes_{\mathbb{Z}}R}/((\mathbb{F}_{q}^{*})^{\otimes_{\mathbb{Z}}R})^{\overline{\alpha}}\rightarrow\mu_{\overline{\alpha}}:=\{u\in\mu_{N(\alpha)}^{\otimes_{\mathbb{Z}}R}\subseteq(\mathbb{F}_{q}^{*})^{\otimes_{\mathbb{Z}}R}:u^{\overline{\alpha}}=1\},

given by

xx(q1)α¯1.x\mapsto x^{(q-1)\overline{\alpha}^{-1}}.
Proposition 4.4.

Let nn\in\mathbb{Z}. For positive integers nn, let

tn:E[n]×E/[n]E𝔾m/𝔾mnt_{n}:E[n]\times E/[n]E\rightarrow\mathbb{G}_{m}/\mathbb{G}_{m}^{n}

denote the usual Tate-Lichtenbaum pairing as in Section 2.2. Let DPPicR0(E)[n]D_{P}\in\operatorname{Pic}_{R}^{0}(E)[n] and DQPicR0(E)D_{Q}\in\operatorname{Pic}_{R}^{0}(E). Suppose

DPiτi((Pi)(𝒪)),DQiτi((Qi)(𝒪)).D_{P}\sim\sum_{i}\tau_{i}\cdot\left((P_{i})-(\mathcal{O})\right),\quad D_{Q}\sim\sum_{i}\tau_{i}\cdot\left((Q_{i})-(\mathcal{O})\right).

Then

Tn(DP,DQ)=i,j=0r1tn(Pi,Qj)τj¯τi.T_{n}(D_{P},D_{Q})=\prod_{i,j=0}^{r-1}t_{n}(P_{i},Q_{j})^{\overline{\tau_{j}}\tau_{i}}.

Furthermore, when both of the following quantities are defined, we have

TN(α)(DP,DQ)Tα(DP,DQ)α(mod(𝔾mR)Rα¯){T}_{N(\alpha)}(D_{P},D_{Q})\equiv{T}_{\alpha}(D_{P},D_{Q})^{\alpha}\pmod{(\mathbb{G}_{m}^{\otimes_{\mathbb{Z}}R})^{R\overline{\alpha}}}
Proof.

By a linear equivalence, assume that

DP=iτi((Pi)(𝒪)),DQ=jτj((Qj+S)(S)).D_{P}=\sum_{i}\tau_{i}\cdot\left((P_{i})-(\mathcal{O})\right),\quad D_{Q}=\sum_{j}\tau_{j}\cdot\left((Q_{j}+S)-(S)\right).

where SS is chosen to avoid intersections of supports. We have from (3), we have fP=ifiτif_{P}=\prod_{i}f_{i}^{\tau_{i}} where

div(fi)=n(Pi)n(𝒪).\operatorname{div}(f_{i})=n(P_{i})-n(\mathcal{O}).

We obtain

Tn(DP,DQ)=j(ifi((Qj+S)(S))τi)τj¯.T_{n}(D_{P},D_{Q})=\prod_{j}\left(\prod_{i}f_{i}((Q_{j}+S)-(S))^{\tau_{i}}\right)^{\overline{\tau_{j}}}.

That shows the first statement. For the second, suppose div(fP)=α¯DP\operatorname{div}(f_{P})=\overline{\alpha}\cdot D_{P}. Then for any divisor DQD_{Q} with sufficiently disjoint support,

(fPα)(DQ)=fP(DQ)α.(f_{P}^{{\alpha}})(D_{Q})=f_{P}(D_{Q})^{{\alpha}}.

On the left, we see this is by definition a representative of Tn(DP,DQ)T_{n}(D_{P},D_{Q}) in 𝔾mR/(𝔾mR)n\mathbb{G}_{m}^{\otimes_{\mathbb{Z}}R}/(\mathbb{G}_{m}^{\otimes_{\mathbb{Z}}R})^{n}, since div(fPα)=αdiv(fP)=nDP\operatorname{div}(f_{P}^{\alpha})=\alpha\cdot\operatorname{div}(f_{P})=nD_{P}. However, looking at the right, this is also a representative of Tα(DP,DQ)αT_{\alpha}(D_{P},D_{Q})^{\alpha} in 𝔾mR/(𝔾mR)Rα¯\mathbb{G}_{m}^{\otimes_{\mathbb{Z}}R}/(\mathbb{G}_{m}^{\otimes_{\mathbb{Z}}R})^{R\overline{\alpha}}. ∎

In particular, in the rank 22 case,

τ¯=Tr(τ)τ,τ¯τ=N(τ),\overline{\tau}=Tr(\tau)-\tau,\quad\overline{\tau}\tau=N(\tau),

which gives

(5) Tn(DP,DQ)=(tn(P0,Q0)tn(P1,Q1)N(τ)tn(P0,Q1)Tr(τ))(tn(P1,Q0)tn(P0,Q1)1)τ.{T}_{n}(D_{P},D_{Q})=\left(t_{n}(P_{0},Q_{0})t_{n}(P_{1},Q_{1})^{N(\tau)}t_{n}(P_{0},Q_{1})^{Tr(\tau)}\right)\left(t_{n}(P_{1},Q_{0})t_{n}(P_{0},Q_{1})^{-1}\right)^{\tau}.

Let x,y\langle x,y\rangle be a bilinear pairing on [τ]\mathbb{Z}[\tau]. Then

x1+τx2,y1+τy2\displaystyle\langle x_{1}+\tau x_{2},y_{1}+\tau y_{2}\rangle :=x1,y1+N(τ)x2,y2+Tr(τ)x1,y2+τ(x2,y1x1,y2)\displaystyle:=\langle x_{1},y_{1}\rangle+N(\tau)\langle x_{2},y_{2}\rangle+Tr(\tau)\langle x_{1},y_{2}\rangle+\tau\left(\langle x_{2},y_{1}\rangle-\langle x_{1},y_{2}\rangle\right)

defines a sesquilinear pairing (conjugate linear in second entry). This explains the formula (5), and in fact we could define the pairing Tn(DP,DQ)T_{n}(D_{P},D_{Q}) from tn(Pi,Qi)t_{n}(P_{i},Q_{i}) directly by using Proposition 4.4 as a definition.

Remark 4.5.

There does not seem to be an analogous construction for Tα(DP,DQ)T_{\alpha}(D_{P},D_{Q}) in terms of tn(Pi,Qi)t_{n}(P_{i},Q_{i}). The best we can do requires computing some preimages under multiplication maps. Specifically, by coherence,

Tα(DP,α¯DS)=Tn(DP,DS).T_{\alpha}(D_{P},\overline{\alpha}\cdot D_{S})=T_{n}(D_{P},D_{S}).

To use this for calculation, letting r=2r=2 for simplicity, suppose DS=(S0)(𝒪)+τ((S1)(𝒪))D_{S}=(S_{0})-(\mathcal{O})+\tau\cdot\left((S_{1})-(\mathcal{O})\right). Then suppose α¯=a+cτ,α¯τ=b+dτ\overline{\alpha}=a+c\tau,\overline{\alpha}\tau=b+d\tau, a,b,c,da,b,c,d\in\mathbb{Z}. Then

α¯DS\displaystyle\overline{\alpha}\cdot D_{S} =a(S0)+b(S1)(a+b)(𝒪)+τ(c(S0)+d(S1)(c+d)(𝒪))\displaystyle=a(S_{0})+b(S_{1})-(a+b)(\mathcal{O})+\tau\cdot\left(c(S_{0})+d(S_{1})-(c+d)(\mathcal{O})\right)
([a]S0+[b]S1)(𝒪)+τ(([c]S0+[d]S1)(𝒪)).\displaystyle\sim([a]S_{0}+[b]S_{1})-(\mathcal{O})+\tau\cdot\left(([c]S_{0}+[d]S_{1})-(\mathcal{O})\right).

Thus, we can give an expression for Tα(DP,DQ)T_{\alpha}(D_{P},D_{Q}) in terms of the classical Tate-Lichtenbaum pairing applied to combinations of P0,P1,S0,S1P_{0},P_{1},S_{0},S_{1} provided the SiS_{i} solve

[a]S0+[b]S1=Q0,[c]S0+[b]S1=Q1.[a]S_{0}+[b]S_{1}=Q_{0},\quad[c]S_{0}+[b]S_{1}=Q_{1}.

A principal ideal ring is one in which all right and left ideals are principal.

Lemma 4.6.

Let RR be a ring with an involution called conjugation, II be a principal two-sided ideal of RR, and suppose that R/IR/I is a finite principal ideal ring. Let t:A×BR/It:A\times B\rightarrow R/I be a sesquilinear form on RR-modules (conjugate linear in one variable). Suppose that tt is non-degenerate. Then if aAa\in A has annihilator II, then t(a,)t(a,\cdot) is surjective. Furthermore, if bBb\in B has annihilator II, then t(,b)t(\cdot,b) is surjective.

Proof.

Since R:=R/IR^{\prime}:=R/I is a principal ideal ring, we claim that there is no proper RR-submodule of RR^{\prime} with annihilator II. Indeed, every submodule R′′R^{\prime\prime} of RR^{\prime} is cyclic as an RR^{\prime} module, hence of the form R′′R/JR^{\prime\prime}\cong R^{\prime}/J for some ideal JJ which is the annihilator of R′′R^{\prime\prime}. By a cardinality argument, if R′′R^{\prime\prime} is a proper submodule of RR^{\prime}, then JJ is non-trivial and the annihilator of R′′R^{\prime\prime} as an RR-module is strictly larger than II.

Now let aAa\in A have annihilator II. Then t(a,B)t(a,B) is an RR-module with annihilator equal to the intersection of the annihilators of all elements t(a,b)R/It(a,b)\in R/I, bBb\in B. If this intersection is equal to II, then we have surjectivity, by the preceding argument. If not, then there exists some element rRr\in R which does not annihilate aa, but does annihilate t(a,B)t(a,B). These two properties, respectively, have the consequences that there exists bBb\in B such that t(ra,b)0t(ra,b)\neq 0 by non-degeneracy, but simultaneously that t(a,r¯b)=0t(a,\overline{r}b)=0. This contradiction completes the argument that t(a,)t(a,\cdot) is surjective. The argument that t(,b)t(\cdot,b) is surjective is similar. ∎

Theorem 4.7.

Let KK be a finite field over which the endomorphisms of RR are defined. Let αR\alpha\in R be coprime to char(K)char(K) and the discriminant of RR. Let n=N(α)n=N(\alpha). Suppose KK contains the nn-th roots of unity. Then

Tα:PicR0(E)[α¯](K)×PicR0(E)(K)/RαPicR0(E)(K)(K)R/((K)R)Rα¯T_{\alpha}:\operatorname{Pic}_{R}^{0}(E)[\overline{\alpha}](K)\times\operatorname{Pic}_{R}^{0}(E)(K)/R\alpha\operatorname{Pic}_{R}^{0}(E)(K)\rightarrow(K^{*})^{\otimes_{\mathbb{Z}}R}/((K^{*})^{\otimes_{\mathbb{Z}}R})^{R\overline{\alpha}}

is non-degenerate. Furthermore, if DPD_{P} has annihilator Rα¯RR\overline{\alpha}R, then Tα(DP,)T_{\alpha}(D_{P},\cdot) is surjective; and if DQD_{Q} has annihilator RαRR\alpha R, then Tα(,DQ)T_{\alpha}(\cdot,D_{Q}) is surjective.

Proof.

First, a few preliminaries. Using the fact that KK^{*} is cyclic of order divisible by α¯\overline{\alpha}, the target (K)R/((K)R)Rα¯R/Rα¯R(K^{*})^{\otimes_{\mathbb{Z}}R}/((K^{*})^{\otimes_{\mathbb{Z}}R})^{R\overline{\alpha}}\cong R/R\overline{\alpha}R as RR-modules, and this is finite. We wish to apply Lemma 4.6.

If RR is an imaginary quadratic order, then its quotient R/α¯RR/\overline{\alpha}R is a principal ideal ring (since α¯\overline{\alpha} is coprime to the discriminant).

If RR is an order in a quaternion algebra, then RpM2(p)R\otimes\mathbb{Q}_{p}\cong M_{2}(\mathbb{Q}_{p}) for pp not dividing the discriminant of RR. This implies, in particular, that R/pkRM2(/pk)R/p^{k}R\cong M_{2}(\mathbb{Z}/p^{k}\mathbb{Z}), which is a principal ideal ring. By assumption, α¯\overline{\alpha} is coprime to the discriminant. For any prime α¯\overline{\alpha}, the ring R/Rα¯RR/R\overline{\alpha}R is a quotient of such a ring, hence a principal ideal ring. In general, R/Rα¯RR/R\overline{\alpha}R is a product of principal ideal rings, hence a principal ideal ring.

So by Lemma 4.6, it suffices to check non-degeneracy. Consider first the non-degeneracy of TnT_{n}, nn\in\mathbb{Z}. Let DPD_{P} be given. We show non-degeneracy on the left by finding DQD_{Q} so that Tn(DP,DQ)T_{n}(D_{P},D_{Q}) is non-trivial. By Proposition 4.4, and the non-degeneracy of the traditional Tate pairing tnt_{n}, we can choose DQD_{Q} so that Tn(DP,DQ)T_{n}(D_{P},D_{Q}) is non-trivial (e.g., provided P0𝒪P_{0}\neq\mathcal{O}, choose QiQ_{i}, i>0i>0 to be 𝒪\mathcal{O} to simplify the condition). This depends upon the following fact: the image of TnT_{n} is taken modulo nn-th powers, hence a non-nn-th power entry in one position of 𝔾mR\mathbb{G}_{m}^{\otimes_{\mathbb{Z}}R} implies the element represents a non-trivial coset. Hence TnT_{n} is left-non-degenerate. An exactly similar argument shows TnT_{n} is right-non-degenerate.

Now we consider general α\alpha, with n=N(α)n=N(\alpha). Suppose div(fP)=α¯DP\operatorname{div}(f_{P})=\overline{\alpha}\cdot D_{P}. Then for any divisor DQD_{Q} with sufficiently disjoint support, as observed in the proof of Proposition 4.4,

(6) (fPα)(DQ)=fP(DQ)α.(f_{P}^{{\alpha}})(D_{Q})=f_{P}(D_{Q})^{{\alpha}}.

By non-degeneracy of TnT_{n}, fixing non-trivial DPPicR0(E)[α¯](K)PicR0(E)[n](K)D_{P}\in\operatorname{Pic}_{R}^{0}(E)[\overline{\alpha}](K)\subseteq\operatorname{Pic}_{R}^{0}(E)[n](K), one may choose DQPicR0(E)(K)D_{Q}\in\operatorname{Pic}_{R}^{0}(E)(K) so that Tn(DP,DQ)T_{n}(D_{P},D_{Q}) is not an nn-th power. The expression (6) is a representative of Tn(DP,DQ)T_{n}(D_{P},D_{Q}), so is not an nn-th power. Therefore fP(DQ)f_{P}(D_{Q}) cannot be an α¯\overline{\alpha}-power in 𝔾mR\mathbb{G}_{m}^{\otimes_{\mathbb{Z}}R}. However, this is a representative of Tα(DP,DQ)T_{\alpha}(D_{P},D_{Q}). Therefore we have shown left non-degeneracy.

On the right, fix a non-trivial DQPicR0(E)(K)/RαPicR0(E)(K)D_{Q}\in\operatorname{Pic}_{R}^{0}(E)(K)/R\alpha\operatorname{Pic}_{R}^{0}(E)(K). Choose β[α]\beta\in\mathbb{Z}[\alpha] coprime to α\alpha such that m:=αβm:=\alpha\beta\in\mathbb{Z} and mm divides nn. By coprimality, we may choose a lift βDQPicR0(E)(K)/RmPicR0(E)(K)\beta\cdot D_{Q}^{\prime}\in\operatorname{Pic}_{R}^{0}(E)(K)/Rm\operatorname{Pic}_{R}^{0}(E)(K) of DQD_{Q}. We know there exists some DPPicR0(E)[m](K)D_{P}\in\operatorname{Pic}_{R}^{0}(E)[m](K) so that Tm(DP,DQ)T_{m}(D_{P},{D_{Q}^{\prime}}) is non-trivial, using the earlier case (since mm divides nn). Consider the two quantities

Tα(DP,DQ),Tm(DP,DQ).T_{\alpha}(D_{P},D_{Q}),\quad T_{m}(D_{P},{D_{Q}^{\prime}}).

Suppose div(fP)=mDP=α¯β¯DP\operatorname{div}(f_{P})=mD_{P}=\overline{\alpha}\cdot\overline{\beta}\cdot D_{P}. Then the quantity fP(DQ)(K)Rf_{P}(D_{Q}^{\prime})\in(K^{*})^{\otimes_{\mathbb{Z}}R} is a representative of both of the two quantities just displayed, in their respective domains. Since Tm(DP,DQ)T_{m}(D_{P},D_{Q}^{\prime}) is not an mm-th power in (K)R(K^{*})^{\otimes_{\mathbb{Z}}R}, we observe that Tα(DP,DQ)=Tα(DP,DQ)β¯T_{\alpha}(D_{P},D_{Q})=T_{\alpha}(D_{P},D_{Q}^{\prime})^{\overline{\beta}} is not a mm-th power, so Tα(DP,DQ)T_{\alpha}(D_{P},D_{Q}^{\prime}) is not an α¯\overline{\alpha} power. By coprimality, Tα(DP,DQ)=Tα(DP,DQ)β¯T_{\alpha}(D_{P},D_{Q})=T_{\alpha}(D_{P},D_{Q}^{\prime})^{\overline{\beta}} is not an α¯\overline{\alpha} power. ∎

4.2. Generalization of Weil pairing

Let 𝔾mR[α¯]={x𝔾mR:xα¯=10}\mathbb{G}_{m}^{\otimes_{\mathbb{Z}}R}[\overline{\alpha}]=\{x\in\mathbb{G}_{m}^{\otimes_{\mathbb{Z}}R}:x^{\overline{\alpha}}=1^{\otimes 0}\}, which444Keep in mind the multiplicative nature of our notation: 1τ=11=10=x01^{\otimes\tau}=1^{\otimes 1}=1^{\otimes 0}=x^{\otimes 0}, all representing the identity element of the RR-module. we might call the α¯\overline{\alpha}-th roots of unity in 𝔾mR\mathbb{G}_{m}^{\otimes_{\mathbb{Z}}R}. We can define a generalization of the Weil pairing

Wα:PicR0(E)[α¯]×PicR0(E)[α]𝔾mR[α¯],Wα(DP,DQ)=fP(DQ)fQ(DP)¯1,W_{\alpha}:\operatorname{Pic}_{R}^{0}(E)[\overline{\alpha}]\times\operatorname{Pic}^{0}_{R}(E)[{\alpha}]\rightarrow\mathbb{G}_{m}^{\otimes_{\mathbb{Z}}R}[\overline{\alpha}],\quad W_{\alpha}(D_{P},D_{Q})=f_{P}(D_{Q})\overline{f_{Q}(D_{P})}^{-1},

where div(fP)=α¯DP\operatorname{div}(f_{P})=\overline{\alpha}\cdot D_{P} and div(fQ)=αDQ\operatorname{div}(f_{Q})={\alpha}\cdot D_{Q}, where the pairs (fPf_{P}, DQD_{Q}) and (fQf_{Q}, DPD_{P}) have disjoint support; we reuse the notation from the definition of TαT_{\alpha} (Section 4.1).

Remark 4.8.

Comparing to TαT_{\alpha}, we may wish to write

Wα(DP,DQ)=?Tα(DP,DQ)Tα¯(DQ,DP)¯1,W_{\alpha}(D_{P},D_{Q})\stackrel{{\scriptstyle?}}{{=}}T_{\alpha}(D_{P},D_{Q}){\overline{T_{\overline{\alpha}}(D_{Q},D_{P})}}^{-1},

but a priori, this is not well-defined, because the validity of the equality depends on the correct choice of representative for the coset of Tα(DP,DQ)T_{\alpha}(D_{P},D_{Q}) or Tα¯(DQ,DP)T_{\overline{\alpha}}(D_{Q},D_{P}).

Theorem 4.9.

The definition above is well-defined, bilinear, and satisfies:

  1. (1)

    Restricted Sesquilinearity: For γ,δ\gamma,\delta such that γ(α)=γ\gamma^{(\alpha)}=\gamma and δ(α¯)=δ\delta^{(\overline{\alpha})}=\delta, we have

    Wα(γDP,δDQ)=Wα(DP,DQ)δ¯γ.{W}_{\alpha}(\gamma\cdot D_{P},\delta\cdot D_{Q})={W}_{\alpha}(D_{P},D_{Q})^{\overline{\delta}\gamma}.
  2. (2)

    Conjugate skew-Hermitianity:

    Wα(DP,DQ)=Wα¯(DQ,DP)¯1.W_{\alpha}(D_{P},D_{Q})=\overline{W_{\overline{\alpha}}(D_{Q},D_{P})}^{-1}.
  3. (3)

    Compatibility: Let ϕ:EE\phi:E\rightarrow E^{\prime}. Then

    Wα(ϕDP,ϕDQ)=Wα(DP,DQ)degϕ.W_{\alpha}(\phi_{*}D_{P},\phi_{*}D_{Q})=W_{\alpha}(D_{P},D_{Q})^{\deg\phi}.
  4. (4)

    Coherence: For DPPicR0(E)[βα¯]D_{P}\in\operatorname{Pic}_{R}^{0}(E)[\overline{\beta\alpha}], DQPicR0(E)[βα]D_{Q}\in\operatorname{Pic}^{0}_{R}(E)[{\beta\alpha}],

    Wβα(DP,DQ)=Wα(β¯DP,DQ)𝔾mR[α¯],Wβα(DP,DQ)=Wβ(DP,αDQ)𝔾mR[β¯].W_{\beta\alpha}(D_{P},D_{Q})=W_{\alpha}(\overline{\beta}\cdot D_{P},D_{Q})\in\mathbb{G}_{m}^{\otimes_{\mathbb{Z}}R}[\overline{\alpha}],\quad W_{\beta\alpha}(D_{P},D_{Q})=W_{\beta}(D_{P},{\alpha}\cdot D_{Q})\in\mathbb{G}_{m}^{\otimes_{\mathbb{Z}}R}[\overline{\beta}].
  5. (5)

    Galois invariance: Suppose EE is defined over a field KK. Let σGal(K¯/K)\sigma\in\operatorname{Gal}(\overline{K}/K); then

    Wα(DP,DQ)σ=Wα(DPσ,DQσ).W_{\alpha}(D_{P},D_{Q})^{\sigma}=W_{\alpha}(D_{P}^{\sigma},D_{Q}^{\sigma}).
Proof.

We begin with well-definition. Suppose DQDQD_{Q}\sim D_{Q}^{\prime} and DPDPD_{P}\sim D_{P}^{\prime}, and let div(g1)=DQDQ\operatorname{div}(g_{1})=D_{Q}-D_{Q}^{\prime} and div(g2)=DPDP\operatorname{div}(g_{2})=D_{P}-D_{P}^{\prime}. From Weil reciprocity,

fQ(DP)¯fQ(DP)¯=(fQfQ)(DP)¯=g1(DP)α¯=g1(α¯DP)¯=fP(DQ)fP(DQ).\frac{\overline{f_{Q}(D_{P})}}{\overline{f_{Q}^{\prime}(D_{P})}}=\overline{\left(\frac{f_{Q}}{f_{Q}^{\prime}}\right)(D_{P})}=\overline{g_{1}(D_{P})^{\alpha}}=\overline{g_{1}(\overline{\alpha}\cdot D_{P})}=\frac{f_{P}(D_{Q})}{f_{P}(D_{Q}^{\prime})}.

Therefore, Wα(DP,DQ)=Wα(DP,DQ)W_{\alpha}(D_{P},D_{Q})=W_{\alpha}(D_{P},D_{Q}^{\prime}). By a symmetrical argument, Wα(DP,DQ)=Wα(DP,DQ)W_{\alpha}(D_{P},D_{Q})=W_{\alpha}(D_{P}^{\prime},D_{Q}^{\prime}). Note that a scalar change of fPf_{P} or fQf_{Q} will cancel. Thus WαW_{\alpha} is well-defined taking values in 𝔾mR\mathbb{G}_{m}^{\otimes_{\mathbb{Z}}R}. The proof of bilinearity is as for TαT_{\alpha} in Theorem 4.2. From the definition, observe that Wα(DP,0)=Wα(0,DQ)=1W_{\alpha}(D_{P},0)=W_{\alpha}(0,D_{Q})=1. In particular, bilinearity implies the image is in 𝔾mR[α¯]\mathbb{G}_{m}^{\otimes_{\mathbb{Z}}R}[\overline{\alpha}].

The argument for sesquilinearity of TαT_{\alpha} in the proof of Theorem 4.2 works equally well here, as does the argument for compatibility. Conjugate skew-Hermitianity is exactly from the definition of WαW_{\alpha}. For coherence, recall that αβ¯=β¯α¯\overline{\alpha\beta}=\overline{\beta}\overline{\alpha} and apply the definitions. Galois invariance follows as in Theorem 4.2. ∎

Theorem 4.10.

Suppose EE has CM by [α][\alpha]. Suppose RR is an imaginary quadratic order. Let DPPicR0(E)[α¯]D_{P}\in\operatorname{Pic}_{R}^{0}(E)[\overline{\alpha}]. Fixing DPD_{P} as a representative in its class, let gPg_{P} be a function with divisor div(gP)=[α]DP\operatorname{div}(g_{P})=[\alpha]^{*}D_{P}. Suppose DQPicR0(E)[α]D_{Q}\in\operatorname{Pic}_{R}^{0}(E)[\alpha]. It is possible to choose the representative DQD_{Q} so that [α]DQ=0[\alpha]_{*}D_{Q}=0; do so. Then

Wα(DP,DQ)=gP(DQ+X)gP(X),{W}_{\alpha}(D_{P},D_{Q})=\frac{g_{P}(D_{Q}+X)}{g_{P}(X)},

where XX is any element of PicR0(E)\operatorname{Pic}_{R}^{0}(E) such that XX and DQ+XD_{Q}+X are not in the support of gPg_{P}.

Proof.

In the case EE has CM by αR\alpha\in R, ERE^{\otimes_{\mathbb{Z}}R} is a [α]\mathbb{Z}[\alpha]-module in two ways. To distinguish them, write [α](Pβ):=([α]P)β[\alpha](P^{\otimes\beta}):=([\alpha]P)^{\otimes\beta} versus (Pβ)α=Pαβ(P^{\otimes\beta})^{\alpha}=P^{\otimes\alpha\beta}. Fix fPf_{P} to have divisor α¯DP\overline{\alpha}\cdot D_{P}. Then the condition on a function gPg_{P} that gPα¯=fP[α]g_{P}^{\overline{\alpha}}=f_{P}\circ[\alpha] up to scaling is equivalent to gPα¯=[α]fPg_{P}^{\overline{\alpha}}=[\alpha]^{*}f_{P} up to scaling, which is equivalent to div(gP)=[α]DP\operatorname{div}(g_{P})=[\alpha]^{*}D_{P} because

div(gPα¯)=α¯div(gP),div([α]fP)=[α]α¯DP=α¯[α]DP.\operatorname{div}(g_{P}^{\overline{\alpha}})=\overline{\alpha}\cdot\operatorname{div}(g_{P}),\quad\operatorname{div}([\alpha]^{*}f_{P})=[\alpha]^{*}{\overline{\alpha}}\cdot D_{P}={\overline{\alpha}}\cdot[\alpha]^{*}D_{P}.

We now give a formula for a function gPg_{P} and show it has the equivalent properties above, so it must be an elliptic function with divisor [α¯]DP[\overline{\alpha}]^{*}D_{P}. Choose an auxiliary point DTDivR0(E)D_{T}\in\operatorname{Div}_{R}^{0}(E) with support disjoint from that of fPf_{P} but such that [α+1]DT=0[\alpha+1]_{*}D_{T}=0. Define for all divisors DXD_{X},

HDX:=[α]DX[α1]DT+DTα(DXDT)PicR0(E).H_{D_{X}}:=[{\alpha}]_{*}D_{X}-[{\alpha}-1]_{*}D_{T}+D_{T}-{\alpha}\cdot(D_{X}-D_{T})\in\operatorname{Pic}_{R}^{0}(E).

This is principal by construction, so we write HDX=div(hDX)H_{D_{X}}=\operatorname{div}(h_{D_{X}}). In order to specify gPg_{P} up to scaling, it suffices to give its values on PicR0(E)\operatorname{Pic}_{R}^{0}(E). Let DXD_{X} be an arbitrary divisor. Set

gP(DX):=fP(DX)hDX(DP)¯.g_{P}(D_{X}):=f_{P}(D_{X})\overline{h_{D_{X}}(D_{P})}.

Then

gP(DX)α¯=\displaystyle g_{P}(D_{X})^{\overline{\alpha}}= fP(αDX)hDX(α¯DP)¯\displaystyle f_{P}({\alpha}\cdot D_{X}){\overline{h_{D_{X}}(\overline{\alpha}\cdot D_{P})}}
=\displaystyle= fP(αDX+div(hDX))\displaystyle f_{P}({\alpha}\cdot D_{X}+\operatorname{div}(h_{D_{X}}))
=\displaystyle= fP(αDT+[α]DX[α+1]DT+DT)\displaystyle f_{P}({\alpha}\cdot D_{T}+[{\alpha}]_{*}D_{X}-[{\alpha}+1]_{*}D_{T}+D_{T})
=\displaystyle= fP([α]DX[α+1]DT)fP(DT)α+1\displaystyle f_{P}([{\alpha}]_{*}D_{X}-[{\alpha}+1]_{*}D_{T})f_{P}(D_{T})^{{\alpha}+1}
=\displaystyle= fP([α]DX)fP(DT)α+1.\displaystyle f_{P}([{\alpha}]_{*}D_{X})f_{P}(D_{T})^{{\alpha}+1}.

Replace gPg_{P} with a scalar multiple so that we obtain gPα¯=fP[α]g_{P}^{\overline{\alpha}}=f_{P}\circ[\alpha]. This provides us with a formula for gPg_{P}.

Now, we observe that since αDQ\alpha\cdot D_{Q} is principal, [α]DQ[\alpha]_{*}D_{Q} is principal. But then we can translate DQD_{Q} by a principal divisor (every principal divisor in the image of a pushforward [α][\alpha]_{*} has a principal preimage, since such principal divisors are supported on points in the image of [α][\alpha]) so that [α]DQ=0[\alpha]_{*}D_{Q}=0. This allows us to make the choice stipulated by the theorem statement. Then, using [α]DQ=0[\alpha]_{*}D_{Q}=0, the divisor

div(hDX)div(hDX+DQ)=αDQ[α]DQ=αDQ\operatorname{div}(h_{D_{X}})-\operatorname{div}(h_{D_{X}+D_{Q}})=\alpha\cdot D_{Q}-[\alpha]_{*}D_{Q}=\alpha\cdot D_{Q}

is the divisor of a function fQf_{Q}. We may now compute

gP(DX+DQ)gP(DX)\displaystyle\frac{g_{P}(D_{X}+D_{Q})}{g_{P}(D_{X})} =fP(DX+DQ)hDX+DQ(DP)¯fP(DX)hDX(DP)¯\displaystyle=\frac{f_{P}(D_{X}+D_{Q})\overline{h_{D_{X}+D_{Q}}(D_{P})}}{f_{P}(D_{X})\overline{h_{D_{X}}(D_{P})}}
=fP(DQ)fQ(DP)¯1\displaystyle={f_{P}(D_{Q})}{\overline{f_{Q}(D_{P})}^{-1}}
=Wα(DP,DQ).\displaystyle={W}_{\alpha}(D_{P},D_{Q}).

Analogously to Proposition 4.4, for WnW_{n}, we can give an expression in terms of the classical Weil pairing.

Proposition 4.11.

The following hold.

  1. (1)

    Let nn\in\mathbb{Z}. Let

    en:E[n]×E[n]μne_{n}:E[n]\times E[n]\rightarrow\mu_{n}

    denote the usual Weil pairing as in Section 2.1. Let DP,DQPicR0(E)[n]D_{P},D_{Q}\in\operatorname{Pic}_{R}^{0}(E)[n]. Suppose

    DPτi((Pi)(𝒪)),DQτi((Qi)(𝒪)).D_{P}\sim\sum\tau_{i}\cdot\left((P_{i})-(\mathcal{O})\right),\quad D_{Q}\sim\sum\tau_{i}\cdot\left((Q_{i})-(\mathcal{O})\right).

    Then

    Wn(DP,DQ)=i,j=0r1en(Pi,Qj)τj¯τi.W_{n}(D_{P},D_{Q})=\prod_{i,j=0}^{r-1}e_{n}(P_{i},Q_{j})^{\overline{\tau_{j}}\tau_{i}}.
  2. (2)

    Now suppose RR is an imaginary quadratic order, and αR\alpha\in R. Suppose

    eα:E[α¯]×E[α]μN(α)e_{\alpha}:E[\overline{\alpha}]\times E[\alpha]\rightarrow\mu_{N(\alpha)}

    denote the usual Weil pairing as in Section 2.1. Let DPPicR0(E)[α¯]D_{P}\in\operatorname{Pic}_{R}^{0}(E)[\overline{\alpha}], DQPicR0(E)[α]D_{Q}\in\operatorname{Pic}_{R}^{0}(E)[\alpha]. Suppose

    DPτi((Pi)(𝒪)),DQτi((Qi)(𝒪)).D_{P}\sim\sum\tau_{i}\cdot\left((P_{i})-(\mathcal{O})\right),\quad D_{Q}\sim\sum\tau_{i}\cdot\left((Q_{i})-(\mathcal{O})\right).

    Then

    Wα(DP,DQ)=i,j=01eα(Pi,Qj)τj¯τi.W_{\alpha}(D_{P},D_{Q})=\prod_{i,j=0}^{1}e_{\alpha}(P_{i},Q_{j})^{\overline{\tau_{j}}\tau_{i}}.
  3. (3)

    Finally, when both of the following quantities are defined, and when RR is an imaginary quadratic order, with αR\alpha\in R, then

    WN(α)(DP,DQ)=Wα(DP,DQ)α.{W}_{N(\alpha)}(D_{P},D_{Q})={W}_{\alpha}(D_{P},D_{Q})^{\alpha}.
Proof.

By a linear equivalence, assume that

DP=iτi((Pi)(𝒪)),DQ=jτj((Qj+S)(S)).D_{P}=\sum_{i}\tau_{i}\cdot\left((P_{i})-(\mathcal{O})\right),\quad D_{Q}=\sum_{j}\tau_{j}\cdot\left((Q_{j}+S)-(S)\right).

where SS is chosen to avoid intersections of supports. We have from (3), we have fP=ifi,Pτif_{P}=\prod_{i}f_{i,P}^{\tau_{i}}, fQ=ifj,Qτjf_{Q}=\prod_{i}f_{j,Q}^{\tau j} where

div(fi,P)=n(Pi)n(𝒪),div(fj,Q)=n(Qj+S)n(S).\operatorname{div}(f_{i,P})=n(P_{i})-n(\mathcal{O}),\quad\operatorname{div}(f_{j,Q})=n(Q_{j}+S)-n(S).

We obtain555In counterpoint to the footnote in the proof of Theorem 4.2, we do have gα¯=gα¯\overline{g^{\alpha}}=g^{\overline{\alpha}} when g𝔾m1g\in\mathbb{G}_{m}^{\otimes 1}.

Wn(DP,DQ)\displaystyle W_{n}(D_{P},D_{Q}) =fP(jτj((Qj+S)(S)))fQ(iτi((Pi)(𝒪)))¯1\displaystyle=f_{P}\left(\sum_{j}\tau_{j}((Q_{j}+S)-(S))\right)\overline{f_{Q}\left(\sum_{i}\tau_{i}((P_{i})-(\mathcal{O}))\right)}^{-1}
=jfP((Qj+S)(S))τj¯ifQ((Pi)(𝒪))τi¯¯1\displaystyle=\prod_{j}f_{P}((Q_{j}+S)-(S))^{\overline{\tau_{j}}}\overline{\prod_{i}f_{Q}((P_{i})-(\mathcal{O}))^{\overline{\tau_{i}}}}^{-1}
=j(ifi,P((Qj+S)(S))τi)τj¯i(jfj,Q((Pi)(𝒪))τj)τi¯¯1\displaystyle=\prod_{j}\left(\prod_{i}f_{i,P}((Q_{j}+S)-(S))^{\tau_{i}}\right)^{\overline{\tau_{j}}}\overline{\prod_{i}\left(\prod_{j}f_{j,Q}((P_{i})-(\mathcal{O}))^{\tau_{j}}\right)^{\overline{\tau_{i}}}}^{-1}
=jifi,P((Qj+S)(S))τj¯τifj,Q((Pi)(𝒪))τi¯τj¯1\displaystyle=\prod_{j}\prod_{i}f_{i,P}((Q_{j}+S)-(S))^{\overline{\tau_{j}}\tau_{i}}\overline{f_{j,Q}((P_{i})-(\mathcal{O}))^{\overline{\tau_{i}}\tau_{j}}}^{-1}
=jifi,P((Qj+S)(S))τj¯τi(fj,Q((Pi)(𝒪))τj¯τi)1\displaystyle=\prod_{j}\prod_{i}f_{i,P}((Q_{j}+S)-(S))^{\overline{\tau_{j}}\tau_{i}}\left(f_{j,Q}((P_{i})-(\mathcal{O}))^{\overline{\tau_{j}}{\tau_{i}}}\right)^{-1}

That shows the first statement. For the second and third, suppose div(fP)=α¯DP\operatorname{div}(f_{P})=\overline{\alpha}\cdot D_{P} and div(fQ)=αDQ\operatorname{div}(f_{Q})=\alpha\cdot D_{Q}.

For the second statement, we use the alternate definitions of eαe_{\alpha} and WαW_{\alpha} in terms of gPg_{P} (Definition 2.2 and Theorem 4.10). Using the notation gPg_{P} and XX from Theorem 4.10, we write gP=igi,Pτig_{P}=\prod_{i}g_{i,P}^{\tau_{i}}, DP=iτiDi,PD_{P}=\sum_{i}\tau_{i}D_{i,P} (so that div(gi,P)=Di,P\operatorname{div}(g_{i,P})=D_{i,P}), DQ=jτjDj,QD_{Q}=\sum_{j}\tau_{j}D_{j,Q} and X=jτjXjX=\sum_{j}\tau_{j}X_{j}. Then, much as in the computation above,

Wα(DP,DQ)=gP(DQ+X)gP(X)=i,j(gi,P(Dj,Q+Xj)τj¯τigi,P(Xj)τj¯τi)=i,jeα(Di,P,Dj,Q)τj¯τi.W_{\alpha}(D_{P},D_{Q})=\frac{g_{P}(D_{Q}+X)}{g_{P}(X)}=\prod_{i,j}\left(\frac{g_{i,P}(D_{j,Q}+X_{j})^{\overline{\tau_{j}}\tau_{i}}}{g_{i,P}(X_{j})^{\overline{\tau_{j}}\tau_{i}}}\right)=\prod_{i,j}e_{\alpha}(D_{i,P},D_{j,Q})^{\overline{\tau_{j}}\tau_{i}}.

For the final (third) statement, observe that for any divisor DQD_{Q} with sufficiently disjoint support,

(fPα)(DQ)(fQα¯)(DP)¯=(fP(DQ)fQ(DP)¯)α.\frac{(f_{P}^{{\alpha}})(D_{Q})}{\overline{(f_{Q}^{\overline{\alpha}})(D_{P})}}=\left(\frac{f_{P}(D_{Q})}{\overline{f_{Q}(D_{P})}}\right)^{{\alpha}}.

On the left, this is a representative of Wn(DP,DQ)W_{n}(D_{P},D_{Q}) in 𝔾mR[n]\mathbb{G}_{m}^{\otimes_{\mathbb{Z}}R}[n], since div(fPα)=αdiv(fP)=nDP\operatorname{div}(f_{P}^{\alpha})=\alpha\cdot\operatorname{div}(f_{P})=nD_{P} and div(fQα¯)=α¯div(fQ)=nDQ\operatorname{div}(f_{Q}^{\overline{\alpha}})=\overline{\alpha}\cdot\operatorname{div}(f_{Q})=nD_{Q}. However, looking at the right, this is also a representative of Wα(DP,DQ)αW_{\alpha}(D_{P},D_{Q})^{\alpha} in 𝔾mR[α¯]\mathbb{G}_{m}^{\otimes_{\mathbb{Z}}R}[\overline{\alpha}]. ∎

Remark 4.12.

Because of the footnote in the proof of Theorem 4.2, the last displayed equation of the proof above does not necessarily hold when RR is a quaternion algebra. Furthermore, if one is interested in the second statement of the theorem, in the case of RR a quaternion algebra, one could use the definition in Theorem 4.10 as the primary definition of the Weil pairing, but then one may wish to reprove Theorem 4.9; we have not attempted this.

When EE has CM by αR\alpha\in R, and RR is an imaginary quadratic order, then there is an alternate definition along the lines of the second definition in Section 2.1. Observe that for any field KK containing the nn-th roots of unity, where n=N(α)n=N(\alpha), we have (K)R[α¯](R/nR)[α¯]R/Rα¯R(K^{*})^{\otimes_{\mathbb{Z}}R}[\overline{\alpha}]\cong(R/nR)[\overline{\alpha}]\cong R/R\overline{\alpha}R.

Theorem 4.13.

Let αR\alpha\in R have norm n=N(α)n=N(\alpha). Let K¯\overline{K} be an algebraically closed field with characteristic coprime to nn. Suppose nn is also coprime to the discriminant of RR. The pairing

Wα:PicR0(E)[α¯](K¯)×PicR0(E)[α](K¯)(R/nR)[α¯]{W}_{\alpha}:\operatorname{Pic}_{R}^{0}(E)[\overline{\alpha}](\overline{K})\times\operatorname{Pic}_{R}^{0}(E)[\alpha](\overline{K})\rightarrow(R/nR)[\overline{\alpha}]

is non-degenerate.

Proof.

As in the proof of Theorem 4.7, for WnW_{n} it suffices to use Proposition 4.11 and the non-degeneracy of ene_{n} (Proposition 2.3). Now consider the general case. Fix DPPicR0(E)[α¯](K¯)D_{P}\in\operatorname{Pic}_{R}^{0}(E)[\overline{\alpha}](\overline{K}). Suppose Wα(DP,DQ)=1W_{\alpha}(D_{P},D_{Q})=1 for all DQPicR0(E)[α](K¯)D_{Q}\in\operatorname{Pic}_{R}^{0}(E)[\alpha](\overline{K}). Then for all DQPicR0(E)[N(α)](K¯)D_{Q}\in\operatorname{Pic}_{R}^{0}(E)[N(\alpha)](\overline{K}), we have α¯DQPicR0(E)[α](K¯)\overline{\alpha}\cdot D_{Q}\in\operatorname{Pic}_{R}^{0}(E)[\alpha](\overline{K}), and therefore WN(α)(DP,DQ)=Wα(DP,α¯DQ)=1W_{N(\alpha)}(D_{P},D_{Q})=W_{\alpha}(D_{P},\overline{\alpha}\cdot D_{Q})=1. So we have DP0D_{P}\sim 0 by the first case. ∎

5. Curves with complex multiplication

Thus far the pairings we have constructed are somewhat abstract, being defined even for elliptic curves having no complex multiplication. In this section, we ‘transport’ these pairings to curves with complex multiplication by subrings of RR, and see that the pairings interact with the endomorphisms.

If we have an RR-module homomorphism into PicR0(E)\operatorname{Pic}_{R}^{0}(E), this transports a pairing and its properties from the target to the source.

5.1. Transport via CM subrings

Suppose SRS\subseteq R is a subring, and suppose that EE has CM by SS. Fix a map []:SEnd(E)[\cdot]:S\rightarrow\operatorname{End}(E), γ[γ]\gamma\mapsto[\gamma].

Then for γS\gamma\in S, [γ][\gamma]_{*} acts on Pic0(E)\operatorname{Pic}^{0}(E). Then there is a surjective RR-module homomorphism

PicR0(E)RPic0(E)RSPic0(E).\operatorname{Pic}_{R}^{0}(E)\cong R\otimes_{\mathbb{Z}}\operatorname{Pic}^{0}(E)\rightarrow R\otimes_{S}\operatorname{Pic}^{0}(E).

which in particular takes

γDγD[γ]D\gamma\cdot D\rightarrow\gamma\cdot D\sim[\gamma]_{*}D

for all γS\gamma\in S. This gives rise to an exact sequence of RR-modules defining PicR,S0(E)\operatorname{Pic}_{R,S}^{0}(E) as follows:

(7) 0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}PicR,S0(E)\textstyle{\operatorname{Pic}_{R,S}^{0}(E)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}η\scriptstyle{\eta}PicR0(E)\textstyle{\operatorname{Pic}^{0}_{R}(E)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ϵ\scriptstyle{\epsilon}RSPic0(E)\textstyle{R\otimes_{S}\operatorname{Pic}^{0}(E)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}

Thus we can transport pairings to PicR,S0(E)\operatorname{Pic}_{R,S}^{0}(E). When R=SR=S, we can identify PicR,S0(E)\operatorname{Pic}_{R,S}^{0}(E) with Er1E^{r-1} via

Er1PicR,S0(E),(P1,,Pr1)([τi]Pi)(𝒪)+τi((Pi)(𝒪)).E^{r-1}\rightarrow\operatorname{Pic}_{R,S}^{0}(E),\quad(P_{1},\ldots,P_{r-1})\mapsto\left(\sum[-\tau_{i}]P_{i}\right)-(\mathcal{O})+\sum\tau_{i}\left((P_{i})-(\mathcal{O})\right).

(This is not canonical; there’s a choice of automorphism of Er1E^{r-1}.) Thus we obtain pairings on Er1E^{r-1}.

5.2. Imaginary quadratic case

Suppose EE defined over KK has CM by RR, an order in an imaginary quadratic field. To fix a map REnd(E)R\rightarrow\operatorname{End}(E), denoted γ[γ]\gamma\rightarrow[\gamma], we first fix an injection ι:RK¯\iota:R\rightarrow\overline{K}, and then we can take that which is normalized as in [18, II.1.1], i.e. [γ]ω=ι(γ)ω[\gamma]^{*}\omega=\iota(\gamma)\omega for the invariant differential ω\omega of EE and γR\gamma\in R. The situation of the last subsection becomes

(8) 0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}E\textstyle{E\ignorespaces\ignorespaces\ignorespaces\ignorespaces}η\scriptstyle{\eta}PicR0(E)\textstyle{\operatorname{Pic}^{0}_{R}(E)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ϵ\scriptstyle{\epsilon}E\textstyle{E\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}

given by RR-module homomorphism

ϵ:PicR0(E)E,αi(Pi)[αi]Pi.\epsilon:\operatorname{Pic}^{0}_{R}(E)\rightarrow E,\quad\sum\alpha_{i}(P_{i})\mapsto\sum[\alpha_{i}]P_{i}.

The kernel is an RR-module, identified with EE via

η:EPicR0(E),P([τ]P)(𝒪)+τ((P)(𝒪)).\eta:E\rightarrow\operatorname{Pic}_{R}^{0}(E),\quad P\mapsto([-\tau]P)-(\mathcal{O})+\tau((P)-(\mathcal{O})).

but note that the RR-module action on this EE is twisted:

(9) η([α]P)=α¯η(P),\eta([\alpha]P)=\overline{\alpha}\cdot\eta(P),

because if α=a+cτ\alpha=a+c\tau and ατ=b+dτ\alpha\tau=b+d\tau, then α¯=dcτ\overline{\alpha}=d-c\tau and α¯τ=b+aτ\overline{\alpha}\tau=-b+a\tau, so

η([α]P)=([τα]P)(𝒪)+τ(([α]P)(𝒪))(d([τ]P)b(P)+τ(c([τ]P)+a(P)))=α¯η(P).\eta([\alpha]P)=([-\tau\alpha]P)-(\mathcal{O})+\tau(([\alpha]P)-(\mathcal{O}))\sim(d([-\tau]P)-b(P)+\tau(-c([-\tau]P)+a(P)))=\overline{\alpha}\cdot\eta(P).

Observe that η\eta is not actually dependent on the choice of τ\tau; a map fitting the exact sequence is unique up to automorphism of EE. Notice η\eta respects the action of any isogeny ϕ:EE\phi:E\rightarrow E^{\prime} which itself respects CM by RR, i.e., if ϕ[τ]=[τ]ϕ\phi\circ[\tau]=[\tau]\circ\phi, then

η(ϕP)=ϕη(P).\eta(\phi P)=\phi_{*}\eta(P).

Finally, we discuss the Galois action. Let σGal(K¯/K)\sigma\in\operatorname{Gal}(\overline{K}/K). Recall that the exact sequence (8) depends upon the normalized choice of map REnd(E)R\rightarrow\operatorname{End}(E) and the injection ι\iota. Write ηE\eta_{E} and ηEσ\eta_{E^{\sigma}} to distinguish. When we conjugate EE to EσE^{\sigma}, making these normalized choices, there is an isomorphism End(E)End(Eσ)\operatorname{End}(E)\cong\operatorname{End}(E^{\sigma}) given by ([α]E)σ=[ασ]Eσ([\alpha]_{E})^{\sigma}=[\alpha^{\sigma}]_{E^{\sigma}} (this follows as in [18, II.2.2(a)]). Then the following commutes:

(10) 0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}E\textstyle{E\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ηE\scriptstyle{\eta_{E}}σ\scriptstyle{\sigma}PicR0(E)\textstyle{\operatorname{Pic}^{0}_{R}(E)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}σ\scriptstyle{\sigma}E\textstyle{E\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}σ\scriptstyle{\sigma}0\textstyle{0}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Eσ\textstyle{E^{\sigma}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ηEσ\scriptstyle{\eta_{E^{\sigma}}}PicRσ0(Eσ)\textstyle{\operatorname{Pic}^{0}_{R^{\sigma}}(E^{\sigma})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Eσ\textstyle{E^{\sigma}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}

where the notation RσR^{\sigma} indicates that we use the injection ισ:RK\iota\circ\sigma:R\rightarrow K in defining ηEσ\eta_{E^{\sigma}}, i.e. we initially replace RR with RσR^{\sigma} so that

ηEσ:EσPicRσ0(E),P([τσ]P)(𝒪)+τσ((P)(𝒪)).\eta_{E^{\sigma}}:E^{\sigma}\rightarrow\operatorname{Pic}_{R^{\sigma}}^{0}(E),\quad P\mapsto([-\tau^{\sigma}]P)-(\mathcal{O})+\tau^{\sigma}((P)-(\mathcal{O})).

This preserves the Galois action on PicR0\operatorname{Pic}_{R}^{0} as given before:

(γηE(P))σ=ηE([γ¯]EP)σ=ηEσ([γ¯σ]EσPσ)=γηEσ(Pσ)=γ(ηE(P))σ.(\gamma\cdot\eta_{E}(P))^{\sigma}=\eta_{E}([\overline{\gamma}]_{E}P)^{\sigma}=\eta_{E^{\sigma}}([\overline{\gamma}^{\sigma}]_{E^{\sigma}}P^{\sigma})={\gamma}\cdot\eta_{E^{\sigma}}(P^{\sigma})=\gamma\cdot(\eta_{E}(P))^{\sigma}.

5.3. Imaginary quadratic pairings

Define

W^α:E[α¯]×E[α]𝔾mR[α],W^α(P,Q)=Wα¯(η(P),η(Q)),\widehat{W}_{\alpha}:E[\overline{\alpha}]\times E[\alpha]\rightarrow\mathbb{G}_{m}^{\otimes_{\mathbb{Z}}R}[\alpha],\quad\widehat{W}_{\alpha}(P,Q)=W_{\overline{\alpha}}({\eta}(P),\eta(Q)),

where η\eta is as in the previous section.

Theorem 5.1.

The pairing defined above is well-defined, bilinear, and satisfies

  1. (1)

    Restricted Sesquilinearity: For γ,δ\gamma,\delta such that γ(α)=γ\gamma^{(\alpha)}=\gamma and δ(α¯)=δ\delta^{(\overline{\alpha})}=\delta, we have

    W^α([γ]P,[δ]Q)=W^α(P,Q)δγ¯.\widehat{W}_{\alpha}([\gamma]P,[\delta]Q)=\widehat{W}_{\alpha}(P,Q)^{\delta\overline{\gamma}}.
  2. (2)

    Conjugate skew-Hermitianity:

    W^α(P,Q)=W^α¯(Q,P)¯1.\widehat{W}_{\alpha}(P,Q)=\overline{\widehat{W}_{\overline{\alpha}}(Q,P)}^{-1}.
  3. (3)

    Compatibility: Let ϕ:EE\phi:E\rightarrow E^{\prime} be an isogeny between curves with CM by RR and satisfy [α]ϕ=ϕ[α][\alpha]\circ\phi=\phi\circ[\alpha]. Then for PE[α¯]P\in E[\overline{\alpha}] and QE[α]Q\in E[\alpha],

    W^α(ϕP,ϕQ)=W^α(P,Q)degϕ.\widehat{W}_{\alpha}(\phi P,\phi Q)=\widehat{W}_{\alpha}(P,Q)^{\deg\phi}.
  4. (4)

    Coherence: For PE[αβ¯]P\in E[\overline{\alpha\beta}], QE[αβ]Q\in E[{\alpha\beta}],

    W^αβ(P,Q)=W^α([β¯]P,Q)𝔾mR[α],W^αβ(P,Q)=W^β(P,[α]Q)𝔾mR[β].\widehat{W}_{\alpha\beta}(P,Q)=\widehat{W}_{\alpha}([\overline{\beta}]P,Q)\in\mathbb{G}_{m}^{\otimes_{\mathbb{Z}}R}[\alpha],\quad\widehat{W}_{\alpha\beta}(P,Q)=\widehat{W}_{\beta}(P,[{\alpha}]Q)\in\mathbb{G}_{m}^{\otimes_{\mathbb{Z}}R}[\beta].
  5. (5)

    Galois invariance: Suppose EE is defined over a field KK, and suppose there is an injection ι:RK¯\iota:R\rightarrow\overline{K}; indicate this in the notation for the pairing as discussed above. For σGal(K¯/K)\sigma\in\operatorname{Gal}(\overline{K}/K),

    W^αι(P,Q)σ=W^αισ(Pσ,Qσ).\widehat{W}^{\iota}_{\alpha}(P,Q)^{\sigma}=\widehat{W}^{\iota\circ\sigma}_{\alpha}(P^{\sigma},Q^{\sigma}).
Proof.

We see immediately that this pairing is sesquilinear, skew-Hermitian, coherent and compatible, since η\eta is a twisted RR-module homomorphism. Recalling that η([α]P)=α¯η(P)\eta([\alpha]P)=\overline{\alpha}\cdot\eta(P), we have to place the vincula carefully. Galois invariance of W^α\widehat{W}_{\alpha} follows from Galois invariance of WαW_{\alpha}, with reference to the discussion at the end of the last section. ∎

Theorem 5.2.

Let αR\alpha\in R. Let KK be a finite field with algebraic closure K¯\overline{K} and characteristic coprime to N(α)N(\alpha). Suppose also that n=N(α)n=N(\alpha) is coprime to the discriminant of RR. The pairing

W^α:E[α¯](K¯)×E[α](K¯)(R/nR)[α],W^α(P,Q)=Wα¯(η(P),η(Q)).\widehat{W}_{\alpha}:E[\overline{\alpha}](\overline{K})\times E[\alpha](\overline{K})\rightarrow(R/nR)[\alpha],\quad\widehat{W}_{\alpha}(P,Q)=W_{\overline{\alpha}}({\eta}(P),\eta(Q)).

is non-degenerate.

Proof.

Note that (K¯)R[α]R/nR[α](\overline{K}^{*})^{\otimes_{\mathbb{Z}}R}[\alpha]\cong R/nR[\alpha], as in the proof of Theorem 4.13. Using the alternate definition of W^α\widehat{W}_{\alpha} in Theorem 4.10, non-degeneracy is a consequence of the fact that the map

E[α]Aut[K¯(E)/[α]K¯(E)],StSE[\alpha]\rightarrow\operatorname{Aut}[\overline{K}(E)/[\alpha]^{*}\overline{K}(E)],\qquad S\mapsto t_{S}^{*}

is an isomorphism [17, Thm III.4.10(b)] (tSt_{S} denoting translation-by-SS).

In particular, fix PE[α¯](K¯)P\in E[\overline{\alpha}](\overline{K}) and assume that W^α(P,Q)=1\widehat{W}_{\alpha}(P,Q)=1 for all QE[α](K¯)Q\in E[\alpha](\overline{K}). Then, using the notation of Theorem 4.10 and its proof, gP(η(X+Q))=gP(η(X))g_{P}(\eta(X+Q))=g_{P}(\eta(X)) for all QE[α](K¯)Q\in E[\alpha](\overline{K}), where XE(K¯)X\in E(\overline{K}) need only satisfy appropriate conditions on supports. So tQt_{Q}^{*} fixes gPη(K¯(E))Rg_{P}\circ\eta\in(\overline{K}(E)^{*})^{\otimes_{\mathbb{Z}}R}. Therefore, gPη=h[α]g_{P}\circ\eta=h\circ[{\alpha}] for some h(K¯(E))Rh\in(\overline{K}(E)^{*})^{\otimes_{\mathbb{Z}}R}. Hence

hα¯[α]=(h[α])α¯=gPα¯η=fP[α]η=fPη[α],h^{\overline{\alpha}}\circ[\alpha]=(h\circ[{\alpha}])^{\overline{\alpha}}=g_{P}^{\overline{\alpha}}\circ\eta=f_{P}\circ[{\alpha}]\circ\eta=f_{P}\circ\eta\circ[\alpha],

implying that fPη=hα¯f_{P}\circ\eta=h^{\overline{\alpha}}. Taking divisors,

α¯div(h)=div(fPη)=div((fP[τ])fPτ¯)=[τ]div(fP)+τ¯div(fP)=α¯([τ]DP+τ¯DP).\overline{\alpha}\cdot\operatorname{div}(h)=\operatorname{div}(f_{P}\circ\eta)=\operatorname{div}((f_{P}\circ[-\tau])f_{P}^{\overline{\tau}})=[-\tau]^{*}\operatorname{div}(f_{P})+\overline{\tau}\cdot\operatorname{div}(f_{P})=\overline{\alpha}\cdot\left([-\tau]^{*}D_{P}+\overline{\tau}D_{P}\right).

From this, we determine that [τ]DP+τ¯DP[-\tau]^{*}D_{P}+\overline{\tau}D_{P} is principal. Recall that DP=η(P)=([τ]P)(𝒪)+τ((P)(𝒪))D_{P}=\eta(P)=([-\tau]P)-(\mathcal{O})+\tau\left((P)-(\mathcal{O})\right). Thus, momentarily writing D=(P)(𝒪)D^{\prime}=(P)-(\mathcal{O}),

[τ]DP+τ¯DP\displaystyle[-\tau]^{*}D_{P}+\overline{\tau}D_{P} =[τ][τ]D+N(τ)D+Tr(τ)[τ]D+τ([τ]D[τ]D).\displaystyle=[-\tau]^{*}[-\tau]_{*}D^{\prime}+N(\tau)D^{\prime}+Tr(\tau)[-\tau]_{*}D^{\prime}+\tau\left([-\tau]^{*}D^{\prime}-[-\tau]_{*}D^{\prime}\right).

From principality, we conclude that, in particular,

[2N(τ)Tr(τ)τ]P=𝒪,[ττ¯]P=𝒪.[2N(\tau)-Tr(\tau)\tau]P=\mathcal{O},\quad[\tau-\overline{\tau}]P=\mathcal{O}.

The norms of these coefficients are N(τ)ΔR-N(\tau)\Delta_{R} and ΔR\Delta_{R}. Recalling that PE[α¯]P\in E[\overline{\alpha}], and that α¯\overline{\alpha} and ΔR\Delta_{R} are coprime, we can conclude that P=𝒪P=\mathcal{O}. ∎

We can describe W^α\widehat{W}_{\alpha} in terms of the usual α\alpha-Weil pairing, following immediately from Proposition 4.11.

Theorem 5.3.

Let eαe_{\alpha} be the α\alpha-Weil pairing as described in Section 2.1. Then

W^α(P,Q)=(eα¯(P,Q)2N(τ)eα¯([τ¯]P,Q)Tr(τ))(eα¯([τ¯τ]P,Q))τ.\widehat{W}_{\alpha}(P,Q)=\left(e_{\overline{\alpha}}(P,Q)^{2N(\tau)}e_{\overline{\alpha}}([-\overline{\tau}]P,Q)^{Tr(\tau)}\right)\left(e_{\overline{\alpha}}([\overline{\tau}-{\tau}]P,Q)\right)^{\tau}.

Furthermore, when both of the following quantities are defined,

W^N(α)(P,Q)=W^α(P,Q)α¯.\widehat{W}_{N(\alpha)}(P,Q)=\widehat{W}_{\alpha}(P,Q)^{\overline{\alpha}}.

Using the notation of the last subsection, define

T^α:E[α¯]×E/[α]E𝔾mR/(𝔾mR)α,T^α(P,Q)=Tα¯(η(P),η(Q)).\widehat{T}_{\alpha}:E[\overline{\alpha}]\times E/[\alpha]E\rightarrow\mathbb{G}_{m}^{\otimes_{\mathbb{Z}}R}/(\mathbb{G}_{m}^{\otimes_{\mathbb{Z}}R})^{\alpha},\quad\widehat{T}_{\alpha}(P,Q)=T_{\overline{\alpha}}({\eta}(P),\eta(Q)).
Theorem 5.4.

The pairing defined above is well-defined, bilinear, and satisfies

  1. (1)

    Sesquilinearity: For PE[α¯]P\in E[\overline{\alpha}] and QEQ\in E,

    T^α([γ]P,[δ]Q)=T^α(P,Q)γ¯δ.\widehat{T}_{\alpha}([\gamma]P,[\delta]Q)=\widehat{T}_{\alpha}(P,Q)^{{\overline{\gamma}}{\delta}}.
  2. (2)

    Compatibility: Let ϕ:EE\phi:E\rightarrow E^{\prime} be an isogeny between curves with CM by RR and satisfy [α]ϕ=ϕ[α][\alpha]\circ\phi=\phi\circ[\alpha]. Then for PE[α¯]P\in E[\overline{\alpha}] and QEQ\in E,

    T^α(ϕP,ϕQ)=T^α(P,Q)degϕ.\widehat{T}_{\alpha}(\phi P,\phi Q)=\widehat{T}_{\alpha}(P,Q)^{\deg\phi}.
  3. (3)

    Coherence: Suppose PE[αβ¯]P\in E[\overline{\alpha\beta}], and QE/[αβ]EQ\in E/[\alpha\beta]E. Then

    T^αβ(P,Q)mod(𝔾mR)α=T^α([β¯]P,Qmod[α]E).\widehat{T}_{\alpha\beta}(P,Q)\bmod{(\mathbb{G}_{m}^{\otimes_{\mathbb{Z}}R})^{\alpha}}=\widehat{T}_{\alpha}([\overline{\beta}]P,Q\bmod[{\alpha}]E).

    Suppose PE[β¯]P\in E[\overline{\beta}], and QE/[αβ]EQ\in E/[\alpha\beta]E. Then

    T^αβ(P,Q)mod(𝔾mR)β=T^β(P,[α]Qmod[β]E).\widehat{T}_{\alpha\beta}(P,Q)\bmod{(\mathbb{G}_{m}^{\otimes_{\mathbb{Z}}R})^{\beta}}=\widehat{T}_{\beta}(P,[{\alpha}]Q\bmod[{\beta}]E).
  4. (4)

    Galois invariance: Suppose EE is defined over a field KK, and suppose there is an injection ι:RK¯\iota:R\rightarrow\overline{K}; indicate this in the notation for the pairing as discussed above. For σGal(K¯/K)\sigma\in\operatorname{Gal}(\overline{K}/K),

    T^αι(P,Q)σ=T^αισ(Pσ,Qσ).\widehat{T}^{\iota}_{\alpha}(P,Q)^{\sigma}=\widehat{T}^{\iota\circ\sigma}_{\alpha}(P^{\sigma},Q^{\sigma}).
Proof.

The proof is as for Theorem 5.1. ∎

We can describe T^n\widehat{T}_{n} in terms of the usual nn-Tate-Lichtenbaum pairing by Proposition 4.4.

Theorem 5.5.

Let tnt_{n} be the nn-Tate-Lichtenbaum pairing as described in Section 2.2.

T^n(P,Q)=(tn(P,Q)2N(τ)tn([τ¯]P,Q)Tr(τ))(tn([τ¯τ]P,Q))τ.\widehat{T}_{n}(P,Q)=\left(t_{n}(P,Q)^{2N(\tau)}t_{n}([-\overline{\tau}]P,Q)^{Tr(\tau)}\right)\left(t_{n}([\overline{\tau}-{\tau}]P,Q)\right)^{\tau}.

Furthermore, provided both of the following quantities are defined,

T^N(α)(P,Q)=T^α(P,Q)α¯(mod(𝔾mR)α)\widehat{T}_{N(\alpha)}(P,Q)=\widehat{T}_{\alpha}(P,Q)^{\overline{\alpha}}\pmod{(\mathbb{G}_{m}^{\otimes_{\mathbb{Z}}R})^{\alpha}}

Our final result is about non-degeneracy.

Proposition 5.6.

Let KK be a finite field, and let EE be an elliptic curve defined over KK. Let αR\alpha\in R be coprime to char(K)char(K) and the discriminant of RR. Let N=N(α)N=N(\alpha). Suppose KK contains the NN-th roots of unity, and E[N]=E[N](K)E[N]=E[N](K). Then

T^α:E[α¯](K)×E(K)/[α]E(K)(K)R/((K)R)α,\widehat{T}_{\alpha}:E[\overline{\alpha}](K)\times E(K)/[\alpha]E(K)\rightarrow(K^{*})^{\otimes_{\mathbb{Z}}R}/((K^{*})^{\otimes_{\mathbb{Z}}R})^{\alpha},

is non-degenerate. Furthermore, if PP has annihilator α¯R\overline{\alpha}R, then Tα(P,)T_{\alpha}(P,\cdot) is surjective; and if QQ has annihilator αR\alpha R, then Tα(,Q)T_{\alpha}(\cdot,Q) is surjective.

Proof.

First, the target is isomorphic to the finite RR-module R/αRR/{\alpha}R, which is a principal ideal ring (using the coprimality to the discriminant). So we can apply Lemma 4.6, and need only show the non-degeneracy.

Recall that R=[τ]R=\mathbb{Z}[\tau] for some τ\tau and by the hypotheses on α\alpha, NN is coprime to ττ¯\tau-\overline{\tau}. First we prove an auxiliary result about T^N\widehat{T}_{N}. Let PE[N](K)P\in E[N](K). Choose QE(K)Q\in E(K) so that tN([τ¯τ]P,Q)t_{N}([\overline{\tau}-\tau]P,Q) has order NN (this must exist since PP has order NN, and NN is coprime to τ¯τ\overline{\tau}-\tau). Then by Theorem 5.5,

T^N(P,Q)\displaystyle\widehat{T}_{N}(P,Q) =(tN(P,Q)2N(τ)tN([τ¯]P,P)Tr(τ))(tN([τ¯τ]P,Q))τ.\displaystyle=\left(t_{N}(P,Q)^{2N(\tau)}t_{N}([-\overline{\tau}]P,P)^{Tr(\tau)}\right)(t_{N}([\overline{\tau}-\tau]P,Q))^{\tau}.

Thus T^N\widehat{T}_{N} is non-degenerate on the left. On the other hand, choosing QQ first, then since ττ¯\tau-\overline{\tau} is coprime to NN, there exists PP making this non-trivial also. Hence we have both left and right non-degeneracy.

Next, we consider general α\alpha. Let PE[α¯](K)P\in E[\overline{\alpha}](K). Then we can let div(fα,P)=αη(P)\operatorname{div}(f_{\alpha,P})={\alpha}\cdot\eta(P). Let div(fN,P)=Nη(P)=α¯αη(P)\operatorname{div}(f_{N,P})=N\cdot\eta(P)=\overline{\alpha}\alpha\cdot\eta(P). Then

fN,P(η(Q))=fα,P(η(Q))α¯.f_{N,P}(\eta(Q))=f_{\alpha,P}(\eta(Q))^{\overline{\alpha}}.

This is a representative of T^N(P,Q)\widehat{T}_{N}(P,Q), and for an appropriate choice of QQ modulo [N]E(K)[N]E(K), is not an NN-th power (by the first case above). Taking this QQ modulo [α]E(K)[\alpha]E(K), fα,P(η(Q))f_{\alpha,P}(\eta(Q)), a representative of T^α(P,Q)\widehat{T}_{\alpha}(P,Q), is not an α\alpha power, i.e. non-trivial.

On the other hand, choose βR\beta\in R coprime to α\alpha with m:=αβm:=\alpha\beta\in\mathbb{Z} and mm divides NN. Fix non-trivial QE(K)Q\in E(K) modulo [α]E(K)[\alpha]E(K). We can choose a lift of the form [β]Q[\beta]Q^{\prime} modulo [m]E(K)[m]E(K) for some QE(K)Q^{\prime}\in E(K). Consider the quantity

fm,P(η(Q)),div(fm,P)=mη(P).f_{m,P}(\eta(Q^{\prime})),\quad\operatorname{div}(f_{m,P})=m\eta(P).

Then there is some PE[m](K)P\in E[m](K) so that the quantity above, as a representative of T^m(P,Q)\widehat{T}_{m}(P,Q^{\prime}), is not an mm-th power (as mm divides NN, this follows from the first part of the proof). But the quantity is also a representative of T^α(P,Q)=T^α(P,Q)β\widehat{T}_{\alpha}(P,Q)=\widehat{T}_{\alpha}(P,Q^{\prime})^{\beta}, which is still not an mm-th power. So T^α(P,Q)\widehat{T}_{\alpha}(P,Q^{\prime}) is not an α\alpha power. And so T^α(P,Q)\widehat{T}_{\alpha}(P,Q) is not an α\alpha power. ∎

5.4. Computation.

We end by giving an explicit formula for T^α(P,Q)\widehat{T}_{\alpha}(P,Q) amenable to computation. This algorithm can be adapted to compute W^α(P,Q)\widehat{W}_{\alpha}(P,Q) also.

Algorithm 5.7.

Recall Remark 4.1. Suppose a+cτ=αa+c\tau=\alpha, b+dτ=ατb+d\tau=\alpha\tau, a,b,c,da,b,c,d\in\mathbb{Z}, which implies dcτ=α¯d-c\tau=\overline{\alpha}, b+aτ=α¯τ-b+a\tau=\overline{\alpha}\tau. We take PE[α¯]P\in E[\overline{\alpha}], DP=η(P)D_{P}=\eta(P), div(fP)=αDP\operatorname{div}(f_{P})=\alpha\cdot D_{P}, fP=fP,1fP,2τf_{P}=f_{P,1}f_{P,2}^{\tau}. The following divisors are principal:

div(fP,1)=a([τ]P)+b(P)(a+b)(𝒪),div(fP,2)=c([τ]P)+d(P)(c+d)(𝒪).\operatorname{div}(f_{P,1})=a([-\tau]P)+b(P)-(a+b)(\mathcal{O}),\quad\operatorname{div}(f_{P,2})=c([-\tau]P)+d(P)-(c+d)(\mathcal{O}).

Choose an auxiliary point SS and define DQ=DQ,1+τDQ,2D_{Q}=D_{Q,1}+\tau\cdot D_{Q,2} where

DQ,1=([τ]Q+[τ]S)([τ]S),DQ,2=(Q+S)(S).D_{Q,1}=([-\tau]Q+[-\tau]S)-([-\tau]S),\quad D_{Q,2}=(Q+S)-(S).

Note that DQη(Q)D_{Q}\sim\eta(Q). Then, choosing SS so that the necessary supports are disjoint (i.e. the support of div(fP,i)\operatorname{div}(f_{P,i}) and DQ,jD_{Q,j} are disjoint for each pair ii, jj), the pairing is defined as

T^α(P,Q):=fP(DQ)=fP,1(DQ,1)fP,2(DQ,1)τ(fP,1(DQ,2)fP,2(DQ,2)τ)τ¯\widehat{T}_{\alpha}(P,Q):=f_{P}(D_{Q})=f_{P,1}(D_{Q,1})f_{P,2}(D_{Q,1})^{\tau}\left(f_{P,1}(D_{Q,2})f_{P,2}(D_{Q,2})^{\tau}\right)^{\overline{\tau}}

which can also be expressed as

(fP,1(DQ,1)fP,1(DQ,2)Tr(τ)fP,2(DQ,2)N(τ))(fP,2(DQ,1)fP,1(DQ,2)1)τ.\left(f_{P,1}(D_{Q,1})f_{P,1}(D_{Q,2})^{Tr(\tau)}f_{P,2}(D_{Q,2})^{N(\tau)}\right)\left(f_{P,2}(D_{Q,1})f_{P,1}(D_{Q,2})^{-1}\right)^{\tau}.

To turn this into an efficient algorithm, observe that we can compute fP,i(D)f_{P,i}(D) for any divisor DD supported on a constant number of points, in O(logmax{a,b,c,d})O(\log\max\{a,b,c,d\}) steps, as follows. Define

div(hP,n)=n(P)([n]P)(n1)(𝒪).\operatorname{div}(h_{P,n})=n(P)-([n]P)-(n-1)(\mathcal{O}).

We can compute hP,n(D)h_{P,n}(D) using a double-and-add algorithm [11] [7, §26.3.1], evaluating at DD at each step. Then observe that

div(fP,1)=div(h[τ]P,a)+div(hP,b)+div(g),div(g)=([aτ]P)+([b]P)2(𝒪)\operatorname{div}(f_{P,1})=\operatorname{div}(h_{[-\tau]P,a})+\operatorname{div}(h_{P,b})+\operatorname{div}(g),\quad\operatorname{div}(g)=([-a\tau]P)+([b]P)-2(\mathcal{O})

Thus, compute g(D)g(D) (the straight line through [aτ]P[-a\tau]P and [b]P[b]P in Weierstrass coordinates), and multiply together to compute fP,1(D)=h[τ]P,a(D)hP,b(D)g(D)f_{P,1}(D)=h_{[-\tau]P,a}(D)h_{P,b}(D)g(D). Computing fP,2(D)f_{P,2}(D) is similar.

6. Examples

Consider the curve y2=x3xy^{2}=x^{3}-x over the prime field 𝔽q\mathbb{F}_{q}, q=401q=401. This curve has complex multiplication by [i]\mathbb{Z}[i]. Let α=12i\alpha=1-2i. A basis for the 55-torsion is P=(204,283)E[α¯]P=(204,283)\in E[\overline{\alpha}], Q=(56,137)E[α]Q=(56,137)\in E[\alpha]. Also, [i]P=(197,46)[i]P=(197,46), [i]Q=(345,334)[i]Q=(345,334). Note that QQ generates E/[α]EE/[\alpha]E and PP generates E[α¯]E[\overline{\alpha}], each of size 55. We will compute T^α(P,Q)\widehat{T}_{\alpha}(P,Q) in a variety of ways.

Method 1. Let us compute the pairing using Algorithm 5.7. We have, for a=d=1a=d=1, b=2b=2, c=2c=-2, that

a+ci=α,b+di=ατ,dci=α¯,b+ai=α¯τ.a+ci=\alpha,\quad b+di=\alpha\tau,\quad d-ci=\overline{\alpha},\quad-b+ai=\overline{\alpha}\tau.

Therefore we define

div(fP,1)=([i]P)+2(P)3(𝒪),div(fP,2)=2([i]P)+(P)+(𝒪).\operatorname{div}(f_{P,1})=([-i]P)+2(P)-3(\mathcal{O}),\quad\operatorname{div}(f_{P,2})=-2([-i]P)+(P)+(\mathcal{O}).

Recall that [2]P=[i]P[2]P=[i]P, since [α¯]P=𝒪[\overline{\alpha}]P=\mathcal{O}. Using the notation L(T,U)L(T,U) for the line through TT and UU, having divisor (T)+(U)(T+U)(𝒪)(T)+(U)-(T+U)-(\mathcal{O}) and V(T)V(T) for the vertical line through TT, having divisor (T)+(T)2(𝒪)(T)+(-T)-2(\mathcal{O}), we have from the expression above that

fP,1=L(P,P).f_{P,1}=L(P,P).

Therefore, using the standard Weierstrass model and its addition formulæ,

fP,1(X,Y)=(Yλ1X+λ1x(P)y(P))(Xx(2P)),λ1=3x(P)212y(P).f_{P,1}(X,Y)=(Y-\lambda_{1}X+\lambda_{1}x(P)-y(P))(X-x(2P)),\quad\lambda_{1}=\frac{3x(P)^{2}-1}{2y(P)}.

This becomes

fP,1(X,Y)=47X+Y+82.f_{P,1}(X,Y)=-47X+Y+82.

Now for the second function

div(fP,2)=2([i]P)+(P)+𝒪\operatorname{div}(f_{P,2})=-2([-i]P)+(P)+\mathcal{O}

we have

fP,2=(L(iP,iP)V(2iP))1=V(2iP)L(iP,iP).f_{P,2}=\left(\frac{L(-iP,-iP)}{V(-2iP)}\right)^{-1}=\frac{V(-2iP)}{L(-iP,-iP)}.

That is,

fP,2(X,Y)=(Xx(2iP)(Yλ2X+λ2x(iP)y(iP)),λ2=3x(iP)212y(iP).f_{P,2}(X,Y)=\frac{(X-x(-2iP)}{(Y-\lambda_{2}X+\lambda_{2}x(-iP)-y(-iP))},\quad\lambda_{2}=\frac{3x(-iP)^{2}-1}{2y(-iP)}.

This becomes

fP,2(X,Y)=X+197138X+Y36.f_{P,2}(X,Y)=\frac{X+197}{-138X+Y-36}.

Let h=3h=3, a multiplicative generator for 𝔽q\mathbb{F}_{q}. Using an auxiliary point such as R=(0,0)R=(0,0) and the formula from Algorithm 5.7, we obtain

T^α(P,Q)=175(5)i=h158+248ih3+3i.\widehat{T}_{\alpha}(P,Q)=175(-5)^{i}=h^{158+248i}\equiv h^{3+3i}.

Using instead an auxiliary point such as R=(1,0)R=(1,0), we obtain

T^α(P,Q)=186144i=h134+106ih4+ih3+3i.\widehat{T}_{\alpha}(P,Q)=186\cdot 144^{i}=h^{134+106i}\equiv h^{4+i}\equiv h^{3+3i}.

This illustrates the independence of the choice of RR.

To take this into μ5R\mu_{5}^{\otimes R}, for the purposes of comparing with the next method, we raise to the (q1)/5=80(q-1)/5=80. Let g=72=h80g=72=h^{80}, a generator for μ5\mu_{5}. We obtain a type of reduced pairing (albeit slightly different than that of Remark 2.6):

T^αred(P,Q):=T^α(P,Q)q15=g3+3ig2.\widehat{T}^{red}_{\alpha}(P,Q):=\widehat{T}_{\alpha}(P,Q)^{\frac{q-1}{5}}=g^{3+3i}\equiv g^{2}.

Method 2. Now we will compute it by using both parts of Theorem 5.5, relating it to T^5\widehat{T}_{5}. We have the reduced Tate-Lichtenbaum pairing tnred=tn(q1)/nt_{n}^{red}=t_{n}^{(q-1)/n} as implemented in many mathematical software systems,

t5red(P,Q)=g,t5red([2i]P,Q)=g4,t5red(P,P)=1,t5red([2i]P,P)=1,t5red(Q,Q)=1,t5red([2i]Q,Q)=1.t^{red}_{5}(P,Q)=g,\quad t^{red}_{5}([2i]P,Q)=g^{4},\quad t^{red}_{5}(P,P)=1,\quad t^{red}_{5}([2i]P,P)=1,\quad t^{red}_{5}(Q,Q)=1,\quad t^{red}_{5}([2i]Q,Q)=1.

Therefore,

T^5red(P,Q)=g2i,T^5red(P,P)=g0,T^5red(Q,Q)=g0.\widehat{T}^{red}_{5}(P,Q)=g^{2-i},\quad\widehat{T}^{red}_{5}(P,P)=g^{0},\quad\widehat{T}^{red}_{5}(Q,Q)=g^{0}.

Since PP is an α\alpha-multiple, we expect T^5(P,)\widehat{T}_{5}(P,\cdot) to be α¯\overline{\alpha} powers. Note that α¯13(modα)\overline{\alpha}^{-1}\equiv 3\pmod{\alpha}. Therefore, modulo α\alpha, we have

T^αred(P,Q)=(g2i)3=g1+2i=g2.\widehat{T}^{red}_{\alpha}(P,Q)=(g^{2-i})^{3}=g^{1+2i}=g^{2}.

Finally, we repeat the first part of the computation above using a single generator for the RR-module E[5]E[5]. Observe that E[5]=𝒪SE[5]=\mathcal{O}S, where S=P+QS=P+Q. In particular, P=(3+4i)SP=(3+4i)S and Q=(3+i)SQ=(3+i)S. We have

T^5red(S,S)=g4,T^5red(S,P)=g24i,T^5red(S,Q)=g2i.\widehat{T}^{red}_{5}(S,S)=g^{4},\quad\widehat{T}^{red}_{5}(S,P)=g^{2-4i},\quad\widehat{T}^{red}_{5}(S,Q)=g^{2-i}.

We can verify that in fact

T^5red(P,Q)=T^5red([3+4i]S,[3+i]S)=T^5red(S,S)(34i)(3+i)=T^5red(S,S)8+6i=(g4)3+i=g2i.\widehat{T}^{red}_{5}(P,Q)=\widehat{T}^{red}_{5}([3+4i]S,[3+i]S)=\widehat{T}^{red}_{5}(S,S)^{(3-4i)(3+i)}=\widehat{T}^{red}_{5}(S,S)^{8+6i}=(g^{4})^{3+i}=g^{2-i}.

agreeing with the previous work.

References

  • [1] Peter Bruin. The Tate pairing for Abelian varieties over finite fields. J. Théor. Nombres Bordeaux, 23(2):323–328, 2011.
  • [2] Wouter Castryck, Marc Houben, Simon-Philipp Merz, Marzio Mula, Sam van Buuren, and Frederik Vercauteren. Weak instances of class group action based cryptography via self-pairings. In Advances in cryptology—CRYPTO 2023. Part III, volume 14083 of Lecture Notes in Comput. Sci., pages 762–792. Springer, Cham, [2023] ©2023.
  • [3] Sylvain Duquesne and Gerhard Frey. Background on pairings. In Handbook of elliptic and hyperelliptic curve cryptography, Discrete Math. Appl. (Boca Raton), pages 115–124. Chapman & Hall/CRC, Boca Raton, FL, 2006.
  • [4] Gerhard Frey and Hans-Georg Rück. A remark concerning mm-divisibility and the discrete logarithm in the divisor class group of curves. Math. Comp., 62(206):865–874, 1994.
  • [5] Steven D. Galbraith. Pairings. In Advances in elliptic curve cryptography, volume 317 of London Math. Soc. Lecture Note Ser., pages 183–213. Cambridge Univ. Press, Cambridge, 2005.
  • [6] Steven D. Galbraith. The Weil pairing on elliptic curves over \mathbb{C}. 2005.
  • [7] Steven D. Galbraith. Mathematics of public key cryptography. Cambridge University Press, Cambridge, 2012.
  • [8] Theodoulos Garefalakis. The generalized Weil pairing and the discrete logarithm problem on elliptic curves. In LATIN 2002: Theoretical informatics (Cancun), volume 2286 of Lecture Notes in Comput. Sci., pages 118–130. Springer, Berlin, 2002.
  • [9] Serge Lang. Abelian varieties. Springer-Verlag, New York-Berlin, 1983. Reprint of the 1959 original.
  • [10] Stephen Lichtenbaum. Duality theorems for curves over pp-adic fields. Invent. Math., 7:120–136, 1969.
  • [11] Victor S. Miller. Short programs for functions on elliptic curves. Unpublished manuscript, 1986.
  • [12] Victor S. Miller. The Weil pairing, and its efficient calculation. J. Cryptology, 17(4):235–261, 2004.
  • [13] J. S. Milne. Abelian varieties. In Arithmetic geometry (Storrs, Conn., 1984), pages 103–150. Springer, New York, 1986.
  • [14] David Mumford. Abelian varieties. Tata Institute of Fundamental Research Studies in Mathematics, No. 5. Published for the Tata Institute of Fundamental Research, Bombay, 1970.
  • [15] Damien Robert. The geometric interpretation of the Tate pairing and its applications. Cryptology ePrint Archive, Paper 2023/177, 2023. https://eprint.iacr.org/2023/177.
  • [16] Damien Robert. Fast pairings via biextensions and cubical arithmetic. Cryptology ePrint Archive, Paper 2024/517, 2024. https://eprint.iacr.org/2024/517.
  • [17] Joseph H. Silverman. The arithmetic of elliptic curves, volume 106 of Graduate Texts in Mathematics. Springer-Verlag, New York, 1992. Corrected reprint of the 1986 original.
  • [18] Joseph H. Silverman. Advanced topics in the arithmetic of elliptic curves, volume 151 of Graduate Texts in Mathematics. Springer-Verlag, New York, 1994.
  • [19] Katherine E. Stange. Elliptic nets and elliptic curves. PhD thesis, Brown University, May 2008.
  • [20] J. Tate. WCWC-groups over 𝐩{\mathbf{p}}-adic fields, volume 13 of Séminaire Bourbaki; 10e année: 1957/1958. Textes des conférences; Exposés 152 à 168; 2e éd. corrigée, Exposé 156. Secrétariat mathématique, Paris, 1958.