Sesquilinear pairings on elliptic curves
Abstract.
Let be an elliptic curve with complex multiplication by a ring , where is an order in an imaginary quadratic field or quaternion algebra. We define sesquilinear pairings (-linear in one variable and -conjugate linear in the other), taking values in an -module, generalizing the Weil and Tate-Lichtenbaum pairings.
Key words and phrases:
Elliptic curves, Weil pairing, Tate-Lichtenbaum pairing, complex multiplication2020 Mathematics Subject Classification:
Primary: 11G05, 14H521. Introduction
The Weil and Tate-Lichtenbaum pairings are bilinear pairings on an elliptic curve with values in the multiplicative group . In the situation of complex multiplication, the points of the elliptic curve form more than just a -module, but also an -module, for some ring 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 -linearity. In this paper, we generalize these classical pairings to take values in an -module, so that the pairings can become sesquilinear, or conjugate linear in the following sense. If is commutative, an -sesquilinear pairing is a bilinear pairing on a pair of -modules, taking values in another -module, that satisfies
In the case that is non-commutative, we also consider a twisted version; see Section 4. For the remainder of the introduction, we assume 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 . By considering instead , we have an -module structure on divisor classes. To accommodate the values of the pairing, considering as a -module in multiplicative notation, we can extend scalars to , writing . (This multiplicative tensor notation is not without its pitfalls; see the end of the introduction for further discussion.) Write for the -torsion in an -module . For each , we obtain Galois invariant sesquilinear pairings
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 is also conjugate skew-Hermitian in the sense that
These are defined by essentially imitating the definition of the classical pairings, including extending Weil reciprocity to -divisors.
However, this formal exercise is most interesting when applied to a curve with endomorphism ring containing a copy of . Consider an exact sequence
given by
By restricting the pairing to the left-hand in the exact sequence, we obtain Galois invariant pairings
which are sesquilinear, if is commutative, in the sense that for all and , ,
and similarly for . When is non-commutative, a similar construction is possible, but sesquilinearity in one entry is twisted by an action of (Section 4).
In the case that , these pairings can be interpreted as a ‘sesquilinearization’ of the usual Weil and Tate-Lichtenbaum pairings. For example if
represents the usual Tate-Lichtenbaum pairing, and , then
In the general case, one can only express in terms of 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 () generally refer to elements of the ring , with the exception of , which is an element of a Galois group, and and , which are maps in Section 5. Roman letters in lower case () will generally refer to elements of and capital roman letters (besides and ) typically refer to points of an elliptic curve . We use the exponent for the extension of scalars from to when viewing an abelian group in multiplicative notation as a -module, as in . Simple tensors are written , but we will suppress the , writing . Note, however, that we will continue to view this as a left -module. Regular exponents will be reserved for the module action of and 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 .
For this reason we write for the image of the multiplicative left -module under the action of the -submodule , or equivalently, under . We refer to this as the set of -powers of . (If , or more generally the centre of , we can simplify the notation from to .)
We denote the algebraic closure of a field by . We denote the action of an endomorphism on by . For an -module , write . When is commutative, this is again an -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 be an integer. Let be an elliptic curve defined over a field which contains the field of definition of , and with characteristic coprime to in the case of positive characteristic. Suppose that . Choose divisors and of disjoint support such that
Then , hence there are functions and such that
The Weil pairing
is defined by
For example, we can choose and disjoint as follows: first choose some such that . Then set and . Set the notation for the rational function with divisor . Then,
Definition 2.2 (Weil pairing: second definition).
Let be an isogeny between elliptic curves defined over a perfect field which contains the field of definition of and , and with characteristic coprime to in the case of positive characteristic. Suppose that , and . Let be a rational function with principal divisor
(In the case that , this implies .) The Weil pairing
where is any positive integer with , and denotes the -th roots of unity, is defined by
where is any auxiliary point chosen disjoint from the supports of and (the function precomposed with translation by ).
The above definition generalizes naturally to a pairing associated to an isogeny; taking the isogeny to be the multiplication-by-m map recovers the -Weil pairing.
The standard properties are as follows.
Proposition 2.3.
Suppose is coprime to 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 for the first definition):
-
(1)
Bilinearity: for and ,
-
(2)
Alternating: for ,
-
(3)
Skew-symmetry: for and ,
-
(4)
Non-degeneracy: for nonzero , there exists such that
-
(5)
Coherence: for , and ,
and for , and ,
-
(6)
Compatibility: For -torsion points and ,
-
(7)
Galois invariance: for , and ,
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 -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 be an integer. Let be an elliptic curve defined over a field . Suppose that . Choose divisors and of disjoint support such that
Then , hence there is a function such that
The Tate-Lichtenbaum pairing
is defined by
Proposition 2.5.
Definition 2.4 is well-defined, and has the following properties:
-
(1)
Bilinearity: for and
-
(2)
Non-degeneracy: Let be a finite field containing the -th roots of unity . For nonzero , there exists such that
Furthermore, for , there exists such that
-
(3)
Compatibility: For an -torsion point , an isogeny , and a point ,
-
(4)
Galois invariance: for , and ,
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 in order to obtain values in .
Including this final exponentiation, there is a more general notion of Tate pairing associated to a -rational isogeny when , that is,
where is any positive integer so that . This generalizes the definition above when , and can be given by
where is an arbitrarily chosen -preimage of , is the -power Frobenius, and is the Weil pairing. It has the property that its values agree with those of 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 -divisors
Let be an order in an imaginary quadratic field or quaternion algebra. Such a ring comes equipped with an involution which we term conjugation, denoted . In the quaternion algebra case, this is order reversing: .
Let be an elliptic curve with divisor group . We extend common notions from to . We emphasize that in this section we make no assumption that has complex multiplication.
In what follows we choose an integral basis: write , where and we let range in or according to the rank of . When we sum over the range will be understood in context.
3.1. -divisors
We define to be the -module generated by all symbols , where is a point of , i.e. finite formal -linear combinations of such symbols, which we call -divisors. (We will frequently suppress the for notational simplicity.) Then is an -module under the action . A divisor is of degree if in ; these form a sub--module .
In the presence of a preferred integral basis for , we can write a sum over :
We say that a divisor of degree zero
is principal if , , are all principal in . We see that the principal divisors form a sub--module and we define and to be the -module quotient of and by the principal divisors. Observe that being principal in is independent of basis, and that , .
If , we define
Thus principal divisors are those which are divisors of .
We define the usual push-foward and pull-back operations on divisors by extending -linearly. Suppose . Then
These inherit the usual desired properties:
-
(1)
-
(2)
,
-
(3)
,
where we define and .
We also have a Galois action: for .
For a divisor , , we define
Viewing as a -module, we obtain an -module . Then we have an -module isomorphism
To show this is an isomorphism, we need to check that it is injective (surjectivity is clear). If then for all , so is principal. In fact, an inverse is given by
3.2. Evaluation of functions at divisors
Let be the multiplicative group. Then is an -module whose action is written multiplicatively as . As a reminder, the action is still a left action, so
It also has a conjugation which will be useful:
Similarly, has a left -module structure and conjugation. Therefore, for , we let
(1) |
so that becomes an -module homomorphism.
We define evaluation of at as , and extend to by defining
This definition requires that the supports of and are disjoint. Observe the vinculum222Thank you to my brother and Wikipedia for teaching me this term for an ., which reflects the duality between and . Among other things, it allows for the two left -actions to communicate in the non-commutative setting:
3.3. Weil reciprocity
A variation of Weil reciprocity ([9, Chapter VI, Corollary to Theorem 10]) holds for us:
Theorem 3.1.
Let . Then
Proof.
The proof relies on Weil reciprocity for . Suppose and . We have
∎
4. Sesquilinear pairings
If is commutative, an -sesquilinear pairing is a bilinear pairing on a pair of -modules, taking values in another -module, that satisfies
For the non-commutative case, we need to add a type of twisting. Recall that is a maximal order in a division algebra. Thus we can set the notation , a subring of . For and , let be defined as that element of which satisfies . For us, a -twisted -sesquilinear pairing is a bilinear pairing on a pair of modules, the first an -module and the second an -module, taking values in another -module, that satisfies
Observe that for rank , commutativity implies and , so the -twisting is vacuous, and we recover sesquilinear pairings in the traditional sense.
4.1. Generalization of Tate-Lichtenbaum pairing
For each , we define an -twisted -sesquilinear pairing generalizing the Tate-Lichtenbaum pairing:
by
where and are chosen to have disjoint support. Observe that is an -module when is commutative, but in general we can only assume it is an -module. Also, we use since may not be an -module in the non-commutative case.
To satisfy the condition on supports, observe that for any divisor , there exist points so that
(2) |
for any auxiliary point . In particular, if are such that , and
then we can take , where
(3) |
and then by a judicious choice of (choosing in the linearly equivalent form (2)), we can satisfy the condition on disjoint supports.
Remark 4.1.
Theorem 4.2.
The pairing defined above is well-defined, bilinear, and satisfies
-
(1)
Twisted sesquilinearity: For and ,
-
(2)
Compatibility: Let . Then
-
(3)
Coherence: Suppose , and . Then
Suppose , and . Then
-
(4)
Galois invariance: Suppose is defined over a field . Let . Then
Proof.
Choice of representative in the divisor class: Suppose . Then for some , having divisor , and using Weil reciprocity333There’s a subtlety here. Observe that , so that it is only in the case that is commutative that . However, it is still true that . (Theorem 3.1),
Choice of modulo :
Choice of representative in the divisor class: Suppose . Notice that if we let and , then
Hence where , which is principal by assumption. Then
Choice of : Any two choices of differ by a constant scalar, but has degree by assumption, so the constant cancels in the formula .
Bilinearity: Let , and , . Then
In the other factor,
Twisted sesquilinearity: Suppose has divisor . In evaluating , we evaluate the function with divisor at the divisor . Since by (1), this becomes
Compatibility: Observe that . Therefore, in the computation of , we evaluate at . We have
where the last equality depends upon the fact that for .
Coherence: Both statements follow immediately from the definitions.
Galois invariance: This is immediate, since by our definition of the actions of on the various entities involved, we have for any . ∎
Remark 4.3.
In cryptographic applications, we typically restrict to inputs defined over a field . If is commutative, to obtain canonical representatives of the codomain, it may be useful to post-compose with a map
given by
Proposition 4.4.
Let . For positive integers , let
denote the usual Tate-Lichtenbaum pairing as in Section 2.2. Let and . Suppose
Then
Furthermore, when both of the following quantities are defined, we have
Proof.
By a linear equivalence, assume that
where is chosen to avoid intersections of supports. We have from (3), we have where
We obtain
That shows the first statement. For the second, suppose . Then for any divisor with sufficiently disjoint support,
On the left, we see this is by definition a representative of in , since . However, looking at the right, this is also a representative of in . ∎
In particular, in the rank case,
which gives
(5) |
Let be a bilinear pairing on . Then
defines a sesquilinear pairing (conjugate linear in second entry). This explains the formula (5), and in fact we could define the pairing from directly by using Proposition 4.4 as a definition.
Remark 4.5.
There does not seem to be an analogous construction for in terms of . The best we can do requires computing some preimages under multiplication maps. Specifically, by coherence,
To use this for calculation, letting for simplicity, suppose . Then suppose , . Then
Thus, we can give an expression for in terms of the classical Tate-Lichtenbaum pairing applied to combinations of provided the solve
A principal ideal ring is one in which all right and left ideals are principal.
Lemma 4.6.
Let be a ring with an involution called conjugation, be a principal two-sided ideal of , and suppose that is a finite principal ideal ring. Let be a sesquilinear form on -modules (conjugate linear in one variable). Suppose that is non-degenerate. Then if has annihilator , then is surjective. Furthermore, if has annihilator , then is surjective.
Proof.
Since is a principal ideal ring, we claim that there is no proper -submodule of with annihilator . Indeed, every submodule of is cyclic as an module, hence of the form for some ideal which is the annihilator of . By a cardinality argument, if is a proper submodule of , then is non-trivial and the annihilator of as an -module is strictly larger than .
Now let have annihilator . Then is an -module with annihilator equal to the intersection of the annihilators of all elements , . If this intersection is equal to , then we have surjectivity, by the preceding argument. If not, then there exists some element which does not annihilate , but does annihilate . These two properties, respectively, have the consequences that there exists such that by non-degeneracy, but simultaneously that . This contradiction completes the argument that is surjective. The argument that is surjective is similar. ∎
Theorem 4.7.
Let be a finite field over which the endomorphisms of are defined. Let be coprime to and the discriminant of . Let . Suppose contains the -th roots of unity. Then
is non-degenerate. Furthermore, if has annihilator , then is surjective; and if has annihilator , then is surjective.
Proof.
First, a few preliminaries. Using the fact that is cyclic of order divisible by , the target as -modules, and this is finite. We wish to apply Lemma 4.6.
If is an imaginary quadratic order, then its quotient is a principal ideal ring (since is coprime to the discriminant).
If is an order in a quaternion algebra, then for not dividing the discriminant of . This implies, in particular, that , which is a principal ideal ring. By assumption, is coprime to the discriminant. For any prime , the ring is a quotient of such a ring, hence a principal ideal ring. In general, 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 , . Let be given. We show non-degeneracy on the left by finding so that is non-trivial. By Proposition 4.4, and the non-degeneracy of the traditional Tate pairing , we can choose so that is non-trivial (e.g., provided , choose , to be to simplify the condition). This depends upon the following fact: the image of is taken modulo -th powers, hence a non--th power entry in one position of implies the element represents a non-trivial coset. Hence is left-non-degenerate. An exactly similar argument shows is right-non-degenerate.
Now we consider general , with . Suppose . Then for any divisor with sufficiently disjoint support, as observed in the proof of Proposition 4.4,
(6) |
By non-degeneracy of , fixing non-trivial , one may choose so that is not an -th power. The expression (6) is a representative of , so is not an -th power. Therefore cannot be an -power in . However, this is a representative of . Therefore we have shown left non-degeneracy.
On the right, fix a non-trivial . Choose coprime to such that and divides . By coprimality, we may choose a lift of . We know there exists some so that is non-trivial, using the earlier case (since divides ). Consider the two quantities
Suppose . Then the quantity is a representative of both of the two quantities just displayed, in their respective domains. Since is not an -th power in , we observe that is not a -th power, so is not an power. By coprimality, is not an power. ∎
4.2. Generalization of Weil pairing
Let , which444Keep in mind the multiplicative nature of our notation: , all representing the identity element of the -module. we might call the -th roots of unity in . We can define a generalization of the Weil pairing
where and , where the pairs (, ) and (, ) have disjoint support; we reuse the notation from the definition of (Section 4.1).
Remark 4.8.
Comparing to , we may wish to write
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 or .
Theorem 4.9.
The definition above is well-defined, bilinear, and satisfies:
-
(1)
Restricted Sesquilinearity: For such that and , we have
-
(2)
Conjugate skew-Hermitianity:
-
(3)
Compatibility: Let . Then
-
(4)
Coherence: For , ,
-
(5)
Galois invariance: Suppose is defined over a field . Let ; then
Proof.
We begin with well-definition. Suppose and , and let and . From Weil reciprocity,
Therefore, . By a symmetrical argument, . Note that a scalar change of or will cancel. Thus is well-defined taking values in . The proof of bilinearity is as for in Theorem 4.2. From the definition, observe that . In particular, bilinearity implies the image is in .
Theorem 4.10.
Suppose has CM by . Suppose is an imaginary quadratic order. Let . Fixing as a representative in its class, let be a function with divisor . Suppose . It is possible to choose the representative so that ; do so. Then
where is any element of such that and are not in the support of .
Proof.
In the case has CM by , is a -module in two ways. To distinguish them, write versus . Fix to have divisor . Then the condition on a function that up to scaling is equivalent to up to scaling, which is equivalent to because
We now give a formula for a function and show it has the equivalent properties above, so it must be an elliptic function with divisor . Choose an auxiliary point with support disjoint from that of but such that . Define for all divisors ,
This is principal by construction, so we write . In order to specify up to scaling, it suffices to give its values on . Let be an arbitrary divisor. Set
Then
Replace with a scalar multiple so that we obtain . This provides us with a formula for .
Now, we observe that since is principal, is principal. But then we can translate by a principal divisor (every principal divisor in the image of a pushforward has a principal preimage, since such principal divisors are supported on points in the image of ) so that . This allows us to make the choice stipulated by the theorem statement. Then, using , the divisor
is the divisor of a function . We may now compute
∎
Analogously to Proposition 4.4, for , we can give an expression in terms of the classical Weil pairing.
Proposition 4.11.
The following hold.
- (1)
-
(2)
Now suppose is an imaginary quadratic order, and . Suppose
denote the usual Weil pairing as in Section 2.1. Let , . Suppose
Then
-
(3)
Finally, when both of the following quantities are defined, and when is an imaginary quadratic order, with , then
Proof.
By a linear equivalence, assume that
where is chosen to avoid intersections of supports. We have from (3), we have , where
We obtain555In counterpoint to the footnote in the proof of Theorem 4.2, we do have when .
That shows the first statement. For the second and third, suppose and .
For the second statement, we use the alternate definitions of and in terms of (Definition 2.2 and Theorem 4.10). Using the notation and from Theorem 4.10, we write , (so that ), and . Then, much as in the computation above,
For the final (third) statement, observe that for any divisor with sufficiently disjoint support,
On the left, this is a representative of in , since and . However, looking at the right, this is also a representative of in . ∎
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 is a quaternion algebra. Furthermore, if one is interested in the second statement of the theorem, in the case of 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 has CM by , and 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 containing the -th roots of unity, where , we have .
Theorem 4.13.
Let have norm . Let be an algebraically closed field with characteristic coprime to . Suppose is also coprime to the discriminant of . The pairing
is non-degenerate.
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 , and see that the pairings interact with the endomorphisms.
If we have an -module homomorphism into , this transports a pairing and its properties from the target to the source.
5.1. Transport via CM subrings
Suppose is a subring, and suppose that has CM by . Fix a map , .
Then for , acts on . Then there is a surjective -module homomorphism
which in particular takes
for all . This gives rise to an exact sequence of -modules defining as follows:
(7) |
Thus we can transport pairings to . When , we can identify with via
(This is not canonical; there’s a choice of automorphism of .) Thus we obtain pairings on .
5.2. Imaginary quadratic case
Suppose defined over has CM by , an order in an imaginary quadratic field. To fix a map , denoted , we first fix an injection , and then we can take that which is normalized as in [18, II.1.1], i.e. for the invariant differential of and . The situation of the last subsection becomes
(8) |
given by -module homomorphism
The kernel is an -module, identified with via
but note that the -module action on this is twisted:
(9) |
because if and , then and , so
Observe that is not actually dependent on the choice of ; a map fitting the exact sequence is unique up to automorphism of . Notice respects the action of any isogeny which itself respects CM by , i.e., if , then
Finally, we discuss the Galois action. Let . Recall that the exact sequence (8) depends upon the normalized choice of map and the injection . Write and to distinguish. When we conjugate to , making these normalized choices, there is an isomorphism given by (this follows as in [18, II.2.2(a)]). Then the following commutes:
(10) |
where the notation indicates that we use the injection in defining , i.e. we initially replace with so that
This preserves the Galois action on as given before:
5.3. Imaginary quadratic pairings
Define
where is as in the previous section.
Theorem 5.1.
The pairing defined above is well-defined, bilinear, and satisfies
-
(1)
Restricted Sesquilinearity: For such that and , we have
-
(2)
Conjugate skew-Hermitianity:
-
(3)
Compatibility: Let be an isogeny between curves with CM by and satisfy . Then for and ,
-
(4)
Coherence: For , ,
-
(5)
Galois invariance: Suppose is defined over a field , and suppose there is an injection ; indicate this in the notation for the pairing as discussed above. For ,
Proof.
We see immediately that this pairing is sesquilinear, skew-Hermitian, coherent and compatible, since is a twisted -module homomorphism. Recalling that , we have to place the vincula carefully. Galois invariance of follows from Galois invariance of , with reference to the discussion at the end of the last section. ∎
Theorem 5.2.
Let . Let be a finite field with algebraic closure and characteristic coprime to . Suppose also that is coprime to the discriminant of . The pairing
is non-degenerate.
Proof.
Note that , as in the proof of Theorem 4.13. Using the alternate definition of in Theorem 4.10, non-degeneracy is a consequence of the fact that the map
is an isomorphism [17, Thm III.4.10(b)] ( denoting translation-by-).
In particular, fix and assume that for all . Then, using the notation of Theorem 4.10 and its proof, for all , where need only satisfy appropriate conditions on supports. So fixes . Therefore, for some . Hence
implying that . Taking divisors,
From this, we determine that is principal. Recall that . Thus, momentarily writing ,
From principality, we conclude that, in particular,
The norms of these coefficients are and . Recalling that , and that and are coprime, we can conclude that . ∎
We can describe in terms of the usual -Weil pairing, following immediately from Proposition 4.11.
Theorem 5.3.
Let be the -Weil pairing as described in Section 2.1. Then
Furthermore, when both of the following quantities are defined,
Using the notation of the last subsection, define
Theorem 5.4.
The pairing defined above is well-defined, bilinear, and satisfies
-
(1)
Sesquilinearity: For and ,
-
(2)
Compatibility: Let be an isogeny between curves with CM by and satisfy . Then for and ,
-
(3)
Coherence: Suppose , and . Then
Suppose , and . Then
-
(4)
Galois invariance: Suppose is defined over a field , and suppose there is an injection ; indicate this in the notation for the pairing as discussed above. For ,
Proof.
The proof is as for Theorem 5.1. ∎
We can describe in terms of the usual -Tate-Lichtenbaum pairing by Proposition 4.4.
Theorem 5.5.
Let be the -Tate-Lichtenbaum pairing as described in Section 2.2.
Furthermore, provided both of the following quantities are defined,
Our final result is about non-degeneracy.
Proposition 5.6.
Let be a finite field, and let be an elliptic curve defined over . Let be coprime to and the discriminant of . Let . Suppose contains the -th roots of unity, and . Then
is non-degenerate. Furthermore, if has annihilator , then is surjective; and if has annihilator , then is surjective.
Proof.
First, the target is isomorphic to the finite -module , 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 for some and by the hypotheses on , is coprime to . First we prove an auxiliary result about . Let . Choose so that has order (this must exist since has order , and is coprime to ). Then by Theorem 5.5,
Thus is non-degenerate on the left. On the other hand, choosing first, then since is coprime to , there exists making this non-trivial also. Hence we have both left and right non-degeneracy.
Next, we consider general . Let . Then we can let . Let . Then
This is a representative of , and for an appropriate choice of modulo , is not an -th power (by the first case above). Taking this modulo , , a representative of , is not an power, i.e. non-trivial.
On the other hand, choose coprime to with and divides . Fix non-trivial modulo . We can choose a lift of the form modulo for some . Consider the quantity
Then there is some so that the quantity above, as a representative of , is not an -th power (as divides , this follows from the first part of the proof). But the quantity is also a representative of , which is still not an -th power. So is not an power. And so is not an power. ∎
5.4. Computation.
We end by giving an explicit formula for amenable to computation. This algorithm can be adapted to compute also.
Algorithm 5.7.
Recall Remark 4.1. Suppose , , , which implies , . We take , , , . The following divisors are principal:
Choose an auxiliary point and define where
Note that . Then, choosing so that the necessary supports are disjoint (i.e. the support of and are disjoint for each pair , ), the pairing is defined as
which can also be expressed as
To turn this into an efficient algorithm, observe that we can compute for any divisor supported on a constant number of points, in steps, as follows. Define
We can compute using a double-and-add algorithm [11] [7, §26.3.1], evaluating at at each step. Then observe that
Thus, compute (the straight line through and in Weierstrass coordinates), and multiply together to compute . Computing is similar.
6. Examples
Consider the curve over the prime field , . This curve has complex multiplication by . Let . A basis for the -torsion is , . Also, , . Note that generates and generates , each of size . We will compute in a variety of ways.
Method 1. Let us compute the pairing using Algorithm 5.7. We have, for , , , that
Therefore we define
Recall that , since . Using the notation for the line through and , having divisor and for the vertical line through , having divisor , we have from the expression above that
Therefore, using the standard Weierstrass model and its addition formulæ,
This becomes
Now for the second function
we have
That is,
This becomes
Let , a multiplicative generator for . Using an auxiliary point such as and the formula from Algorithm 5.7, we obtain
Using instead an auxiliary point such as , we obtain
This illustrates the independence of the choice of .
To take this into , for the purposes of comparing with the next method, we raise to the . Let , a generator for . We obtain a type of reduced pairing (albeit slightly different than that of Remark 2.6):
Method 2. Now we will compute it by using both parts of Theorem 5.5, relating it to . We have the reduced Tate-Lichtenbaum pairing as implemented in many mathematical software systems,
Therefore,
Since is an -multiple, we expect to be powers. Note that . Therefore, modulo , we have
Finally, we repeat the first part of the computation above using a single generator for the -module . Observe that , where . In particular, and . We have
We can verify that in fact
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 -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 . 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 -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. -groups over -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.