Skip to main content
Springer logoLink to Springer
. 2024 Dec 18;11(1):16. doi: 10.1007/s40993-024-00580-z

Transcendental Brauer–Manin obstructions on singular K3 surfaces

Mohamed Alaa Tawfik 1, Rachel Newton 1,
PMCID: PMC11655618  PMID: 39713084

Abstract

Let E and E be elliptic curves over Q with complex multiplication by the ring of integers of an imaginary quadratic field K and let Y=Kum(E×E) be the minimal desingularisation of the quotient of E×E by the action of -1. We study the Brauer groups of such surfaces Y and use them to furnish new examples of transcendental Brauer–Manin obstructions to weak approximation.

Mathematics Subject Classification: Primary: 14G05, Secondary: 14F22, 11G05, 14J28

Introduction

Let k be a number field and let Ak denote the adèles of k. Let X/k be a smooth, projective, geometrically irreducible algebraic variety, let X¯ denote its base change to an algebraic closure of k, and let Br(X)=Ht2(X,Gm) denote the Brauer group of X. For v a place of k and ABr(X), functoriality yields an evaluation map

evA,v:X(kv)Br(kv)xvA(xv).

The Hasse invariant invv:Br(kv)Q/Z is an isomorphism for v finite, and has image 12Z/Z for v real and zero for v complex. In [17], Manin defined what became known as the Brauer–Manin pairing

X(Ak)×Br(X)Q/Z((xv)v,A)vinvv(A(xv)) 1

where the sum is over all places v of k. For BBr(X), the subset of X(Ak) consisting of all elements that are orthogonal to B under the pairing (1) is denoted X(Ak)B. The Brauer–Manin set is X(Ak)Br(X). Global class field theory (the Albert–Brauer–Hasse–Noether Theorem), and continuity of evaluation maps, shows that X(Ak)Br(X) contains the closure of X(k) in X(Ak)=vX(kv) with respect to the product of the v-adic topologies. This shows that, in some cases, the emptiness of X(k) despite X having points in all completions can be explained by the emptiness of X(Ak)Br(X). This is known as a Brauer–Manin obstruction to the Hasse principle. In cases where X(k) is non-empty, one would like to understand more about the rational points on X: for example, does weak approximation hold, i.e. is X(k) dense in X(Ak)? If X(Ak)B is not equal to X(Ak) for some BBr(X), we say that B obstructs weak approximation on X.

Manin’s work initiated a great deal of activity, see [29] for a recent summary. Initially, most research focused on the algebraic part of the Brauer group, which by definition is Br1(X)=ker(Br(X)Br(X¯)), and the more mysterious transcendental part Br(X)/Br1(X) was rarely computed. In [12, 14], the authors computed the odd order torsion in the transcendental Brauer groups of diagonal quartic surfaces by relating these surfaces to Kummer surfaces of products of elliptic curves over Q with complex multiplication by Z[i]. In [12, 13], Ieronymou and Skorobogatov went on to study the evaluation maps for these elements of odd order and thus gave new examples of Brauer–Manin obstructions to weak approximation coming from transcendental Brauer group elements.

In this paper, we replace Z[i] by the ring of integers OK of an imaginary quadratic field K and study Brauer groups and Brauer–Manin obstructions to weak approximation for Kummer surfaces of products of elliptic curves E,E over Q with complex multiplication by OK. Note that the assumption that OK is the endomorphism ring of an elliptic curve over Q implies that K is an imaginary quadratic field of class number one (see e.g. [25, Theorem II.4.1]), but this is the only restriction on K. Moreover, our assumptions also imply that the elliptic curves E and E are geometrically isomorphic, see e.g. [25, Proposition II.2.1].

Theorem 1.1

Let K be an imaginary quadratic field and let Y=Kum(E×E) for elliptic curves E,E over Q with EndE¯=EndE¯=OK. Suppose that Br(Y)/Br1(Y) contains an element of order n>1. Then K{Q(ζ3),Q(i),Q(-7),Q(-2),Q(-11)} and n10.

Remark 1.2

Similar results can be obtained in the more general setting where the elliptic curves can have CM by non-maximal orders in OK, see Remark 3.2 below.

In Theorem 1.1 and throughout the paper, we write Kum(E×E) to mean the minimal desingularisation of the quotient of E×E by the action of -1, which sends (PQ) to (-P,-Q). Such Kummer surfaces are examples of so-called singular K3 surfaces, which are defined to be K3 surfaces of maximal Picard rank.

The cases of Theorem 1.1 where K is Q(i) or Q(ζ3) follow from work of Valloni on Brauer groups of principal K3 surfaces with CM in [28]. In the case where K=Q(i), the odd order torsion in the Brauer group was computed by Ieronymou–Skorobogatov–Zarhin [14] and Ieronymou–Skorobogatov [12] in their study of Brauer groups of diagonal quartic surfaces. In particular, [12, Theorem 1.1] also applies to the Kummer surface Y=Kum(Em1×Em2) where Em has affine equation y2=x3-mx for mQ×. It shows that

(Br(Y)/Br(Q))odd=(Br(Y)/Br1(Y))oddZ/3Zif-3m1m2-4Q×4,Z/5Zif53m1m2-4Q×4,0otherwise. 2

Our next result handles all cases where OK×={±1}.

Theorem 1.3

Let K be an imaginary quadratic field with OK×={±1} and let Y=Kum(E×E) for elliptic curves E,E over Q with EndE¯=EndE¯=OK. Suppose that Br(Y)\Br(Q) contains an element of odd order. Then

  1. K{Q(-2),Q(-11)};

  2. Br(Y)/Br1(Y)=Br(Y)3/Br1(Y)3Z/3Z;

  3. Br1(Y)/Br(Q)(Z/2Z)2ifK=Q(-2),Z/2ZifK=Q(-11);

  4. Y is the minimal desingularisation of the projective surface with affine equation u2=af(x)f(t) where
    f(x)=x3+4x2+2xanda{-3,6}ifK=Q(-2),f(x)=x3-25·33·11x+24·33·7·112anda{-3,33}ifK=Q(-11).

The remaining, and most interesting, case is K=Q(ζ3). Any elliptic curve E over Q with EndE¯=Z[ζ3] has an affine equation of the form Ea:y2=x3+a for some aQ×. Let c,dQ× and let Y=Kum(Ec×Ed). Then Br(Y) can contain transcendental elements of odd order n for n9. Cases involving elements of order divisible by 3 require a more delicate analysis, essentially because 3 ramifies in the CM field Q(ζ3), and will be explored in future work. For elements of order 5 or 7, we have the following:

Theorem 1.4

For aQ×, let Ea be the elliptic curve over Q with affine equation y2=x3+a. Let c,dQ× and let Y=Kum(Ec×Ed). Let {5,7} and let ε()=(-1)(-1)/2. Then

(Br(Y)/Br1(Y))Z/Zifε()·24·ε()·cd-33Q×6,0otherwise.

Furthermore, if ε()·24·ε()·cd-33Q×6 then

Br(Y)/Br1(Y)=Br(Y)/Br1(Y).

The reason for the focus on odd order torsion in Br(Y)/Br1(Y) is a result of Skorobogatov and Zarhin (Theorem 2.1 below), which shows that odd order torsion in Br(E×E)/Br1(E×E) descends to the transcendental Brauer group of Kum(E×E). This means that one can transfer many calculations concerning transcendental Brauer classes to the realm of abelian varieties. In particular, the method of Skorobogatov and Zarhin described in Sect. 2.2 enables one to compute Brauer–Manin pairings for transcendental Brauer group elements of odd order without the need to find explicit Azumaya algebras representing them.

However, it is important that we consider Kummer surfaces and not just abelian surfaces, for the following reason. For torsors under abelian varieties over number fields, any Brauer–Manin obstruction to the Hasse principle or weak approximation can already be explained by an algebraic element in the Brauer group, see [8, 17]. In contrast, for K3 surfaces, it can happen that the algebraic part of the Brauer group consists only of constant elements (and so does not contribute to any Brauer–Manin obstruction), while there is an obstruction coming from a transcendental element in the Brauer group. Examples of this phenomenon were given in [11, 19, 21]. Our next two results yield a new source of examples.

Theorem 1.5

Let K be an imaginary quadratic field and let Y=Kum(E×E) for elliptic curves E,E over Q with EndE¯=EndE¯=OK. Let be an odd prime and if K=Q(ζ3) assume that >3. Suppose that ABr(Y)\Br(Q). Then the evaluation map evA,:Y(Q)Br(Q) is surjective and hence

Y(AQ)AY(AQ),

i.e. A obstructs weak approximation on Y.

Theorem 1.6

Reinstate the notation and assumptions of Theorem 1.4. Suppose that ε()·24·ε()·cd-33Q×6 and c2·ε()·Q×3. Then

Br1(Y)=Br(Q)

and hence the failure of weak approximation in Theorem 1.5 cannot be explained by any algebraic element in the Brauer group of Y.

Remark 1.7

  1. In Theorem 6.1 we prove a complement to Theorem 1.5, showing in many cases that the evaluation maps evA,v for places v are constant, cf. [12, Theorem 1.2(i)].

  2. Creutz and Viray showed in [9, Theorem 1.7] that if there is a Brauer–Manin obstruction to the Hasse principle on a Kummer variety, then there is an obstruction coming from an element in the 2-primary part of the Brauer group. It was already known (see e.g. [21]) that the analogous statement for Brauer–Manin obstructions to weak approximation does not hold. Theorem 1.6 gives further illustration of this fact, generalising the example given in [21, Theorem 1.3]. We note that the statement of [21, Theorem 1.3] needs correcting: for ABr(X)3\Br(Q), the evaluation map evA,v:X(Qv)Br(Qv)3 is constant, but not necessarily zero, for all v3. However, any such A does give an obstruction to weak approximation and there is a choice of A for which the theorem holds as stated – one takes A=B, where the notation is as in the proof of Theorem 6.1. This choice of A should be in force throughout [21, Sect. 5], yielding corrections to the statements of [21, Proposition 5.1, Theorems 5.2, 5.3], see [22].

  3. In the same way that Ieronymou, Skorobogatov and Zarhin used Kummer surfaces of products of elliptic curves with CM by Z[i] to study Brauer groups and Brauer–Manin obstructions on diagonal quartic surfaces in [1214], the results of this paper could be applied to the study of other families of quartic surfaces, e.g. those of the form ax4+cxy3=bz4+dzw3 with abcdQ×, in other words the family of surfaces geometrically isomorphic to Schur’s quartic surface. Note that these surfaces contain lines given by x=z=0 so any Brauer–Manin obstructions arising would be obstructions to weak approximation.

Outline of the paper

We begin by gathering some preliminary results on transcendental elements of Brauer groups and their evaluations at local points in Sect. 2. In Sect. 3, for K an imaginary quadratic field with OK×={±1}, we compute transcendental Brauer groups of products of elliptic curves over Q with CM by OK and prove Theorem 1.1. In Sect. 4 we perform the same calculation in the case where K=Q(ζ3) and prove Theorem 1.4. In Sect. 5 we compute the algebraic part of the Brauer group for each of the Kummer surfaces under consideration and prove Theorem 1.6. Combining the results for the transcendental and algebraic parts of the Brauer group allows us to prove Theorem 1.3. In Sect. 6 we consider the evaluation of a Brauer group element of prime order at p-adic points with p and show in many cases that these evaluation maps are constant, providing a complement to Theorem 1.5. Section 7 is devoted to the proof of Theorem 1.5.

Notation

If A is an abelian group and n a positive integer, then An and A/n denote the kernel and cokernel, respectively, of multiplication by n on A. If is prime, then A denotes the -power torsion subgroup of A.

If k is a field of characteristic zero, then k¯ denotes an algebraic closure of k and Γk denotes the absolute Galois group Gal(k¯/k). If X is an algebraic variety over k and l/k is a field extension then Xl denotes the base change X×kl. The base change X×kk¯ is denoted by X¯.

The Brauer group of X/k is denoted by Br(X) and its algebraic part ker(Br(X)Br(X¯)) is denoted by Br1(X). The quotient Br(X)/Br1(X) is called the transcendental part of Br(X), or the transcendental Brauer group of X.

For an elliptic curve E/k we denote by EndE¯ the full ring of endomorphisms defined over k¯.

Preliminaries

Transcendental Brauer groups

The following result of Skorobogatov and Zarhin allows us to move between the transcendental Brauer group of an abelian surface and that of the associated Kummer surface.

Theorem 2.1

([26, Theorem 2.4]) Let A be an abelian surface, let Y=Kum(A) and let nZ>0. There is a natural embedding

Br(Y)n/Br1(Y)nBr(A)n/Br1(A)n 3

which is an isomorphism if n is odd. The subgroups of elements of odd order of the transcendental Brauer groups Br(Y)/Br1(Y) and Br(A)/Br1(A) are isomorphic.

To calculate the transcendental part of the Brauer group for products of elliptic curves, we use another result of Skorobogatov and Zarhin.

Proposition 2.2

([26, Proposition 3.3]) Let E and E be elliptic curves over a field k of characteristic zero. For nZ>0, we have a canonical isomorphism of abelian groups

Br(E×E)n/Br1(E×E)n=HomΓk(En,En)/(Hom(E¯,E¯)/n)Γk.

When the elliptic curves have complex multiplication, we can say more.

Definition 2.3

Let k be a field of characteristic zero and let E,E be elliptic curves over k with E¯E¯ and EndE¯=EndE¯=OK for an imaginary quadratic field K. For nZ>0, we define Γk-submodules of Hom(En,En), as follows:

Hom(En,En)+={φHom(En,En)φβ=βφβOK};Hom(En,En)-={φHom(En,En)φβ=β¯φβOK},

where β¯ denotes the complex conjugate of β.

The following lemma and corollary are slight generalisations of some results from [12, §3].

Lemma 2.4

Let k be a field of characteristic zero. Let dZ>0 be squarefree and let K=Q(-d). Let E,E be elliptic curves over k with E¯E¯ and EndE¯=EndE¯=OK. For all nZ>0 coprime to 2d, we have equalities of Γk-modules

Hom(En,En)=Hom(En,En)+Hom(En,En)- 4

and

Hom(E¯,E¯)/n=Hom(En,En)+. 5

In the case d=3, (4) and (5) hold for all nZ>0 coprime to 3.

Proof

First suppose that n is coprime to 2d. Then multiplication by -d is invertible on En and En and we have

Hom(En,En)+={φHom(En,En)-dφ-d-1=φ};Hom(En,En)-={φHom(En,En)-dφ-d-1=-φ}.

Since n is odd, we can write any φHom(En,En) as

φ=12φ+-dφ-d-1+12φ--dφ-d-1

and thus prove (4).

The case d=3 and n coprime to 3 is similar except that we consider conjugation by ζ3 instead of -d. Since ζ33=1, we have

Hom(En,En)=Hom(En,En)0Hom(En,En)1Hom(En,En)2

where Hom(En,En)j={φHom(En,En)ζ3φζ3-1=ζ3jφ}. But if φHom(En,En)1 then 0=φ(ζ3-1-1), which implies that φ3=0 and hence φ=0, as n is coprime to 3. Therefore, Hom(En,En)1=0. Now observe that Hom(En,En)0=Hom(En,En)+ and Hom(En,En)2=Hom(En,En)- to complete the proof of (4).

For (5), view End(E¯) and End(En) as OK×-modules via the action of OK×=AutE¯ on the second factor in each case, so that αOK× sends an endomorphism φ to αφ. Since E¯E¯, the elliptic curve E is a twist of E by an element in H1(k,AutE¯)=H1(k,OK×). Thus, the Γk-modules Hom(E¯,E¯) and Hom(En,En) are twists of End(E¯) and End(En), respectively, by the same element of H1(k,OK×), acting on the second factor. Therefore, it is enough to prove that End(E¯)/n=End(En)+. Since End(E¯)=OK, it is clear that End(E¯)/nEnd(En)+. Equality follows from the fact that both are isomorphic to (Z/nZ)2 as abelian groups.

Corollary 2.5

Under the assumptions of Lemma 2.4, we have a canonical isomorphism of abelian groups

Br(E×E)n/Br1(E×E)n=HomΓk(En,En)-.

Proof

Follows immediately from Proposition 2.2 and Lemma 2.4.

In order to use Proposition 2.2 to calculate the whole of the transcendental part of the Brauer group, we will need the following material from [21].

Definition 2.6

Fix a number field L, an imaginary quadratic field K and a prime number Z. Define n() to be the largest integer t such that the ring class field Kt corresponding to the order Z+tOK embeds into KL.

Theorem 2.7

Let L be a number field and let E/L be an elliptic curve such that EndE¯=OK for an imaginary quadratic field K. Then for all prime numbers Z,

Br(E×E)Br1(E×E)=Br(E×E)n()Br1(E×E)n().

Proof

This is an immediate consequence of [21, Proposition 2.2, Theorems 2.5 and 2.9].

To aid us in our applications of Theorem 2.7, we will need the following well-known formula for the degree of a ring class field (see [7, Theorem 7.24], for example). Let K be an imaginary quadratic field with discriminant ΔK and class number hK, let cZ>0 and let Oc=Z+cOK be the order of conductor c in OK. Then

[Kc:K]=hK·c[OK×:Oc×]pc1-ΔKp1p. 6

The symbol (·p) denotes the Legendre symbol for odd primes. For the prime 2, the Legendre symbol is replaced by the Kronecker symbol (·2), with

ΔK2=0if2ΔK,1ifΔK1(mod8),-1ifΔK5(mod8).

Evaluation maps

Let k be a number field. Let dZ>0 be squarefree and let K=Q(-d). Let E,E be elliptic curves over k with E¯E¯ and EndE¯=EndE¯=OK and let Y=Kum(E×E). Let nZ>0 be coprime to 2d and let φHomΓk(En,En)-. For the reader’s convenience, following [12, §5.1], [26, §3], we summarise here the construction of an element of Br(Y)n from φ, and describe its evaluation at a p-adic point of Y in terms of a cup-product map.

Multiplication by n on E turns E into an E-torsor with structure group En. Denote this torsor by T and let [T] denote its class in Ht1(E,En). Similarly, let T denote E considered as an E-torsor with structure group En, and let [T] denote its class in Ht1(E,En). The homomorphism φ:EnEn gives rise to the E-torsor φT with structure group En, with class [φT]Ht1(E,En).

Composing the cup-product map with the Weil pairing En×Enμn yields a pairing

Ht1(E×E,En)×Ht1(E×E,En)Br(E×E)n. 7

Let p:E×EE and p:E×EE be the natural projection maps. The pullbacks pT1 and pφT are E×E-torsors with structure group En; let CBr(E×E)n denote the pairing of their classes in Ht1(E×E,En) via (7). By [26, Lemma 3.1, Proposition 3.3], the natural map

Br(E×E)nBr(E×E)n/Br1(E×E)n=HomΓk(En,En)-

sends C to φ. Let ι denote the involution on Br(E×E) induced by (P,Q)(-P,-Q) on E×E. The proof of [26, Theorem 2.4] identifies Br(Y) with the subgroup of Br(E×E) consisting of elements fixed by ι. By the functoriality and bilinearity of the cup product, we find that ι(C)=C. Let B be the element of Br(Y) corresponding to C.

Note that if we take quadratic twists of E and E by the same element ak× then there is a natural isomorphism Kum(Ea×Ea)Y. Applying the construction described above to the homomorphism of Γk-modules EnaEna coming from φ, we obtain an element of the Brauer group of Kum(Ea×Ea) that is identified with BBr(Y) under the isomorphism Kum(Ea×Ea)Y.

Let F be a field containing k (e.g. F could be the completion of k at some place v of k). Let PE(F)\E2, let QE(F)\E2 and let [P,Q]Y(F) denote the corresponding point on the Kummer surface. Then

B([P,Q])=C((P,Q)).

Let χP denote the image of P under the natural map χ:E(F)H1(F,En) and let χQ denote the image of Q under the natural map χ:E(F)H1(F,En). The cup product and the Weil pairing En×Enμn give a pairing

:H1(F,En)×H1(F,En)Br(F)n. 8

Now the construction of C and the functoriality of the cup product show that

B([P,Q])=C((P,Q))=χPφ(χQ)Br(F)n. 9

This description of the evaluation map, due to Skorobogatov and Zarhin in [26], is very powerful because it enables one to evaluate transcendental elements of Br(Y) at local points (and thus compute the Brauer–Manin pairing) without the need to obtain explicit Azumaya algebras representing these elements of the Brauer group.

Lemma 2.8

Let k be a number field and let kv be its completion at a place v. Let dZ>0 be squarefree and let K=Q(-d). Let E,E be elliptic curves over k with E¯E¯ and EndE¯=EndE¯=OK and let Y=Kum(E×E). Let nZ>0 be coprime to 2d and let φHomΓk(En,En)-. Let BBr(Y) be constructed from φ as described above. If the evaluation map

evB,v:Y(kv)Br(kv)nyB(y)

is constant then it is zero.

Proof

Let PE(kv)\E4 (so that 2PE2) and let QE(kv)\E2. If evB is constant then B([2P,Q])=B([P,Q]) and (9) gives

B([2P,Q])=χ2Pφ(χQ)=2·(χPφ(χQ))=2B([P,Q]).

Hence B([P,Q])=0, which suffices to prove the lemma.

CM by OK with OK×={±1}: transcendental Brauer groups

If EndE¯=OK for an imaginary quadratic field K with OK×={±1} (i.e. K is not Q(i) or Q(ζ3)), then the only twists of E are quadratic twists. Theorem 3.1 below shows that in this case the transcendental part of the Brauer group of E×E has exponent at most 6, where E denotes a quadratic twist of E. The theorem is stated for elliptic curves over K, but the conclusion also holds for elliptic curves over Q, by definition of the transcendental part of the Brauer group (cf. (10) below).

Theorem 3.1

Let K be an imaginary quadratic field with OK×={±1}. Let E/K be an elliptic curve with EndE¯=OK, let E be a quadratic twist of E and let T=Br(E×E)/Br1(E×E).

  • (i)

    If K=Q(-2) then T is killed by 6.

  • (ii)

    If K=Q(-7) then T is killed by 4.

  • (iii)

    If K=Q(-11) then T is killed by 3.

  • (iv)

    In all other cases, T=0.

Proof

Let E be the quadratic twist of E by some aK× and let L=K(a). Write A=E×E=E×Ea. Observe that AL is isomorphic to E×E over L and by definition we have

Br(A)/Br1(A)Br(AL)/Br1(AL). 10

Applying [21, Theorem 1.1 and Proposition 2.2], we see that for any prime number ,

(Br(AL)/Br1(AL))=Br(AL)n()/Br1(AL)n() 11

where n() is the largest integer t such that the ring class field Kt corresponding to the order Z+tOK embeds into L. Bounds on n() are easily obtained by noting that if Kt embeds into L then [Kt:K] divides [L:K]2. Furthermore, we can use the formula (6) to calculate [Kt:K], noting that the theory of complex multiplication shows that hK=1, since the Hilbert class field HK is equal to K(j(E)) and E is defined over K. In this way, we find that n()=0 for all 5. For =3, we find that n(3)1, and n(3)=0 unless K is an imaginary quadratic field of class number one with ΔK3=1, i.e. unless K{Q(-2),Q(-11)}. Similarly, we find that n(2)=0 unless K{Q(-7),Q(-2)}, and furthermore n(2)1 if K=Q(-2) and n(2)2 if K=Q(-7).

Proof of Theorem 1.1

As noted in the introduction, the assumptions of Theorem 1.1 imply that the CM field K has class number one and the elliptic curves E and E are geometrically isomorphic. If OK×={±1}, then this means that E is a quadratic twist of E and the result follows from Theorems 3.1 and 2.1. The remaining cases, where K{Q(i),Q(ζ3)}, follow from [28, Examples 1 and 2, pp. 48–51] and Theorem 2.1.

Remark 3.2

If we relax the assumptions of Theorem 1.1 to allow E and E to have CM by orders Of and Of in OK of conductors f and f, respectively, we can obtain similar results using the existence of isogenies of degrees f and f from E and E, respectively, to elliptic curves with CM by OK. Since Kf=K(j(E))=K=K(j(E))=Kf (see e.g. [7, Theorem 11.1]), the formula (6) shows that f,f3 and K{Q(ζ3),Q(i),Q(-7),Q(-2),Q(-11)}. Bounds on the order of a non-trivial class in the transcendental Brauer group then follow from [2, Theorem 5.13] and Theorem 2.1.

Thus, in the setting of Theorem 3.1, an element of odd order in Br(E×E)/Br1(E×E) has order dividing 3. In Sect. 4, we will see that the situation is more interesting in the case of elliptic curves with complex multiplication by Z[ζ3], where sextic twists can occur. But first we will investigate the cases in Theorem 3.1 where non-trivial elements of odd order can occur in the transcendental part of the Brauer group, namely when K{Q(-2),Q(-11)}. Elements of odd order are of particular interest to us because Theorem 2.1 shows that they descend to the transcendental part of the Brauer group of the relevant Kummer surface, where there is a chance they may give obstructions to weak approximation that cannot be explained by any algebraic element in the Brauer group.

Lemma 3.3

Let K{Q(-2),Q(-11)}, let F{Q,K}, let E/F be an elliptic curve such that EndE¯=OK and let aF×. Then

Br(E×Ea)Br1(E×Ea)3=Br(E×Ea)3Br1(E×Ea)3.

Proof

Let A=E×Ea. Since Br(A) is torsion (see [5, Lemma 3.5.3]),

(Br(A)/Br1(A))3=Br(A)3/Br1(A)3.

Let ABr(A)3 and let L=K(a). Then (10), (11) and the fact that n(3)1 (see the proof of Theorem 3.1) give

(Br(A)/Br1(A))3Br(AL)3/Br1(AL)3. 12

Hence, ResL/FA=B+C where BBr(AL)3 and CBr1(AL)3. Applying corestriction yields

·A=CorL/FB+CorL/FC,

where CorL/FBBr(A)3 and CorL/FCBr1(A)3, by [4, Lemme 1.4]. Since [L : F] is coprime to 3, we can invert [L : F] modulo the order of A to see that the class of A in Br(A)3/Br1(A)3 lies in Br(A)3/Br1(A)3.

Proposition 3.4

Let K{Q(-2),Q(-11)}, let E/K be an elliptic curve with CM by OK, let aK× and let Ea denote the quadratic twist of E by a. Suppose that Br(E×Ea)/Br1(E×Ea) contains an element of order 3. Then K(a)=K(-3) and hence a-3·K×2.

Proof

The proof of Theorem 3.1 shows that if the 3-primary part of Br(E×Ea)/Br1(E×Ea) is non-trivial (so necessarily n(3)>0) then K(a)=K3. It is easily checked that K3=K(-3).

Theorem 3.5

Let K{Q(-2),Q(-11)}, let F{Q,K} and let E/F be an elliptic curve such that EndE¯=OK. Furthermore, let aF×-3·K×2. Then

Br(E×Ea)/Br1(E×Ea)=Br(E×Ea)3Br1(E×Ea)3=HomΓF(E3,E3a)-(Z/3Z)2ifF=K;Z/3ZifF=Q.

Remark 3.6

In Theorem 3.5, if F=K then one may assume a=-3 since multiplying a by an element of F×2 does not change Ea. Likewise, if F=Q then one may assume a{-3,6} if K=Q(-2), and a{-3,33} if K=Q(-11). Proposition 3.4 shows that these are the only quadratic twists for which Br(E×Ea)/Br1(E×Ea) contains non-trivial elements of odd order.

Proof of Theorem 3.5

We begin by computing HomΓF(E3,E3a)-. Since K has class number one, the theory of complex multiplication shows that any two elliptic curves over F with CM by OK are geometrically isomorphic and hence quadratic twists of each other, in this case. Therefore, we can select a chosen elliptic curve E/F with CM by OK and write E=Eδ for some δF×. Thus, Ea=Eδa and Hom(E3,E3a)=Hom(E3δ,E3aδ). Now ΓF acts on Hom(E3δ,E3aδ) by conjugation and the two quadratic twists by δ cancel each other out so that Hom(E3δ,E3aδ)=Hom(E3,E3a) as a ΓF-module. Furthermore, the ΓF-module Hom(E3,E3a)- is the quadratic twist of End(E3)- by the quadratic character corresponding to F(a)/F. In other words, we identify Hom(E3,E3a)- with the group End(E3)- equipped with an action of ΓF such that σΓF sends φEnd(E3)- to σ(a)aσφσ-1.

K=Q(-2): We take E to be the elliptic curve with affine equation y2=x3+4x2+2x, which has complex multiplication by Z[-2] by [25, Proposition II.2.3.1(ii)]. One computes that HomΓK(E3,E3-3)-(Z/3Z)2 and HomΓQ(E3,E3-3)-HomΓQ(E3,E36)-Z/3Z. This can be done by calculating the 3-torsion points of E and using the fact that the ΓF-module Hom(E3,E3-3)- is the quadratic twist of End(E3)- by the quadratic character corresponding to F(-3)/F. Alternatively, for an explicit computation of the ΓQ-modules Hom(E3,E3a) for a{-3,6}, see the proof of [1, Lemma 2.3.3].

Now Lemma 3.3 and Corollary 2.5 show that

Br(E×Ea)Br1(E×Ea)3=Br(E×Ea)3Br1(E×Ea)3=HomΓF(E3,E3a)-.

By Theorem 3.1(i), it only remains to show that the 2-primary part of Br(E×Ea)/Br1(E×Ea) is trivial. Write A=E×Ea and L=K(a). By (10), (11) and the computation of n(2) in the proof of Theorem 3.1, it is enough to show that Br(AL)2/Br1(AL)2=0. Now Proposition 2.2 shows that

Br(AL)2/Br1(AL)2=EndΓL(E2)/(EndE¯/2)ΓL.

One computes that this quotient is trivial. As before, one can take E=E and compute the 2-torsion explicitly. For the details, see the proof of [1, Lemma 2.3.6].

K=Q(-11): We take E to be the elliptic curve with LMFDB label 121.b2, which has affine equation y2+y=x3-x2-7x+10. This is the modular curve Xns+(11). It has complex multiplication by Z[1+-112] by [27]. One computes that HomΓK(E3,E3-3)-(Z/3Z)2 and HomΓQ(E3,E3-3)-HomΓQ(E3,E333)-Z/3Z. Explicit calculations can be found in the proof of [1, Theorem 2.4.1]. Now Theorem 3.1(ii), Lemma 3.3 and Corollary 2.5 show that

Br(E×Ea)Br1(E×Ea)=Br(E×Ea)3Br1(E×Ea)3=HomΓF(E3,E3a)-.

CM by Z[ζ3]: transcendental Brauer groups

Throughout this section, for cQ×, let Ec denote the elliptic curve over Q with affine equation

Ec:y2=x3+c.

The curve Ec has complex multiplication by Z[ζ3], where ζ3 denotes a primitive 3rd root of unity. Multiplication by ζ3 sends

(x,y)(ζ3x,y).

The curve Ec is the sextic twist of y2=x3+1 by the class of c-1 in H1(Q,μ6). Since Q(ζ3) has class number one, any elliptic curve over Q with complex multiplication by Z[ζ3] is of the form Ec for some cQ×.

In this section, we study the transcendental Brauer groups of Ec×Ed and Kum(Ec×Ed) for c,dQ×.

Lemma 4.1

Let c,dQ×. For every prime number >7,

(Br(Ec×Ed)/Br1(Ec×Ed))=0.

For {5,7},

(Br(Ec×Ed)/Br1(Ec×Ed))=Br(Ec×Ed)/Br1(Ec×Ed).

For {2,3},

(Br(Ec×Ed)/Br1(Ec×Ed))=(Br(Ec×Ed)/Br1(Ec×Ed))2.

Proof

Let A=Ec×Ed and let Y=KumA. By definition of Br1(Y), we have an injection

Br(Y)/Br1(Y)Br(Y¯)ΓQ(ζ3).

Since Y is a K3 surface with CM by Z[ζ3], we can apply [28, Example 2, pp. 50–51] to Y to see that, for all primes >7, Br(Y¯)Q(ζ3)=0 and hence (Br(Y)/Br1(Y))=0. Since Br(Y) is a torsion group, (Br(Y)/Br1(Y))=Br(Y)/Br1(Y). The first statement now follows easily from Theorem 2.1.

Let L=Q(ζ3,c/d6) so that ALEc×Ec. By definition,

Br(A)n/Br1(A)nBr(AL)n/Br1(AL)n 13

for all nZ>0. By Theorem 2.7, for any prime number ,

(Br(AL)/Br1(AL))=Br(AL)n()/Br1(AL)n() 14

where n() is as defined in Definition 2.6.

Observe that [L:Q(ζ3)]6. By (6), we have [K8:Q(ζ3)]=4 and [K27:Q(ζ3)]=9, whence n(2),n(3)2. This, together with (13) and (14), proves the statement for {2,3}.

Now suppose that {5,7}. By (6), [K2:Q(ζ3)]=2>6, so n()1. Our discussion above shows that

(Br(AL)/Br1(AL))=(Br(AL))/(Br1(AL)). 15

Now an argument using restriction and corestriction similar to the one used in the proof of Lemma 3.3 shows that (Br(A)/Br1(A))=Br(A)/Br1(A).

In this paper, we will focus on the cases with CM by Z[ζ3] where the transcendental part of the Brauer group contains an element of order 5 or 7. The other cases will be discussed in future work.

Lemma 4.2

Let c,dQ× and let Y=Kum(Ec×Ed). Let {5,7} and suppose that (Br(Y)/Br1(Y))0. Then (3) yields an isomorphism

Br(Y)/Br1(Y)Br(Ec×Ed)/Br1(Ec×Ed).

Proof

This follows from [28, Example 2, pp. 50–51] and Lemma 4.1.

We now calculate Br(Ec×Ed)/Br1(Ec×Ed) for {5,7}, using Corollary 2.5. To compute HomΓQ(Ed,Ec)- for {5,7}, we will use Eisenstein’s sextic reciprocity law, as stated in [16, Theorem 7.10].

Definition 4.3

An element a+bζ3Z[ζ3] is called E-primary if b0mod3 and

a+b1mod4,if2b,b1mod4,if2a,a3mod4,if2ab.

Let N denote the norm map NQ(ζ3)/Q:Q(ζ3)Q. Recall the definition of the sextic residue symbol: for λ,πZ[ζ3] with π prime, λπ6 is the unique 6th root of unity satisfying

λN(π)-16λπ6(modπ).

Theorem 4.4

(Eisenstein) If β,γZ[ζ3] are E-primary and relatively prime, then

βγ6=(-1)N(β)-12N(γ)-12γβ6.

Definition 4.5

Let α=c/d6 and let ϕα:E¯dE¯c be the isomorphism defined over Q(α) given by (x,y)(α2x,α3y).

We use -33Q×6 to denote Q×6-33·Q×6.

Proposition 4.6

View ΓQ(α) and ΓQ(ζ3) as subgroups of ΓQ, so that the set difference ΓQ(α)\ΓQ(ζ3) is defined.

  • (i)
    If 24·5·cd-33Q×6 then, for τΓQ(α)\ΓQ(ζ3), abusing notation and viewing τϕα as an element of Hom(E5d,E5c),
    Br(Ec×Ed)5/Br1(Ec×Ed)5=HomΓQ(E5d,E5c)-=(Z/5Z)·τϕαZ/5Z.
    Otherwise, Br(Ec×Ed)5/Br1(Ec×Ed)5=0.
  • (ii)
    If -24·7-1·cd-33Q×6 then, for τΓQ(α)\ΓQ(ζ3), abusing notation and viewing τϕα as an element of Hom(E7d,E7c),
    Br(Ec×Ed)7/Br1(Ec×Ed)7=HomΓQ(E7d,E7c)-=(Z/7Z)·τϕαZ/7Z.
    Otherwise, Br(Ec×Ed)7/Br1(Ec×Ed)7=0.

Proof

Multiplying by 6th powers if necessary, we may assume that c,dZ. By Corollary 2.5 it suffices to compute HomΓQ(Ed,Ec)- for {5,7}. Let ε()=(-1)(-1)/2. First, we will show that

HomΓQ(ζ3)(Ed,Ec)-=Hom(Ed,Ec)-ifε()·24ε()·cd-33Q×6;0otherwise.

To prove this claim, we will determine the action of ΓQ(ζ3) on Hom(Ed,Ec)- via a study of the actions of Frobenius elements for sufficiently many primes in Z[ζ3]. The action of ΓQ(ζ3) factors through Gal(Q(ζ3,Ec,Ed)/Q(ζ3)). Let πZ[ζ3] be an E-primary prime that is coprime to 6cd and unramified in Q(ζ3,Ec,Ed)/Q(ζ3). The prime ideals generated by such π comprise all but finitely many prime ideals of Z[ζ3], as every prime ideal of Z[ζ3] that is coprime to 6 has an E-primary generator. Furthermore, Chebotarev’s density theorem shows that Gal(Q(ζ3,Ec,Ed)/Q(ζ3)) is generated by Frobenius elements associated to such primes π. We require π to be E-primary so that we can apply sextic reciprocity later on in the proof. In particular, we have π±1(mod3).

For a{c,d}, let ψEa/Q(ζ3) be the Grössencharakter attached to Ea/Q(ζ3). We write ψEa/Q(ζ3)(π) for the image under ψEa/Q(ζ3) of the idèle (1,,1,π,1,1,) with entry π at the place (π) and entry 1 at every other place. Then FrobπGal(Q(ζ3,Ec,Ed)/Q(ζ3)) acts on Ea as multiplication by ψEa/Q(ζ3)(π)Z[ζ3], see [15, Corollary 4.1.3], for example. By [25, Example II.10.6],

ψEa/Q(ζ3)(π)=±4aπ¯6π=±4aπ6-1π. 16

The ±1 here comes from the fact that [25, Example II.10.6] is stated for primes that are congruent to -1(mod3), whereas our E-primary prime π maybe congruent to either 1 or -1(mod3). In any case, the ±1 in (16) is independent of a and is therefore of no consequence for the action on Hom(Ed,Ec)- by conjugation, as the ±1 for the actions on Ec and Ed cancel out. Thus, for φHom(Ed,Ec)-, we have

Frobπ·φ=4cπ6-1πφ4dπ6π-1=24cdπ6-1ππ¯-1φ. 17

For =5, sextic reciprocity gives

5π6=π56π4(mod5).

Furthermore, π¯π5(mod5), whereby ππ¯-15π6-1(mod5). Substituting this into (17) gives

Frobπ·φ=24·5·cdπ6-1φ 18

for all E-primary primes πZ[ζ3] that are coprime to 30cd and unramified in Q(ζ3,E5c,E5d)/Q(ζ3). Since z-1 is invertible modulo 5 for all zμ6\{1}, we deduce that Frobπ·φ=φ if and only if either φ=0 or 24·5·cdπ6=1 for all E-primary primes πZ[ζ3] that are coprime to 30cd and unramified in Q(ζ3,E5c,E5d)/Q(ζ3). The latter condition holds if and only if 24·5·cdQ×Q(ζ3)×6=-33Q×6 (see [20, Theorem 9.1.11]). Hence, if HomΓQ(ζ3)(E5d,E5c)-0 then 24·5·cdQ×Q(ζ3)×6, whereby (18) shows that HomΓQ(ζ3)(E5d,E5c)-=Hom(E5d,E5c)-, as required.

For =7, factorise 7 in Z[ζ3] as 7=ϖϖ¯ with ϖ=-1-3ζ3. Then ϖ and -ϖ¯ are both E-primary and sextic reciprocity gives

-7π6=ϖπ6-ϖ¯π6=πϖ6πϖ¯6ππ¯-1(modϖ). 19

Taking complex conjugates and then inverting both sides of (19) gives -7π6ππ¯-1(modϖ¯) and hence -7π6ππ¯-1(mod7). Substituting this into (17) gives

Frobπ·φ=-24·7-1·cdπ6-1φ 20

for all E-primary primes πZ[ζ3] that are coprime to 42cd and unramified in Q(ζ3,E7c,E7d)/Q(ζ3). As before, we deduce that if HomΓQ(ζ3)(E7d,E7c)-0 then -24·7-1·cdQ×Q(ζ3)×6=-33Q×6 and HomΓQ(ζ3)(E7d,E7c)-=Hom(E7d,E7c)-, completing the proof of our claim.

To complete the proof of Proposition 4.6, it remains to compute HomΓQ(Ed,Ec)- for {5,7} in the case where ε()·24ε()·cd-33Q×6. It is easy to see that the conditions on cd ensure that Q(ζ3)Q(α) and hence there exists some τΓQ(α)\ΓQ(ζ3). Now observe that

HomΓQ(ζ3)(Ed,Ec)-=Hom(Ed,Ec)-={(a+bζ3)τϕαa,bZ/Z}.

Indeed, it is clear that {(a+bζ3)τϕαa,bZ/Z}Hom(Ed,Ec)- and both are isomorphic to (Z/Z)2 as abelian groups.

Furthermore, since the image of τ generates Gal(Q(ζ3)/Q), an element of HomΓQ(ζ3)(Ed,Ec)- is fixed by the action of ΓQ if and only if it commutes with τ. Therefore, HomΓQ(Ed,Ec)-=(Z/Z)·τϕα, as claimed.

Proof of Theorem 1.4

This now follows from Theorem 2.1, Proposition 4.6 and Lemma 4.2.

Algebraic Brauer groups

Let E and E be elliptic curves over Q and let Y=Kum(E×E). Since Y(Q), the Hochschild–Serre spectral sequence gives a short exact sequence

0Br(Q)Br1(Y)H1(Q,Pic(Y¯))0. 21

Since Y is a K3 surface, Pic(Y¯)=NS(Y¯). Furthermore, [26, Proposition 1.4(i)] gives a short exact sequence

0NΛNΣNS(Y¯)Hom(E¯,E¯)0, 22

where NΛ and NΣ are permutation ΓQ modules and hence H1(Q,NΛ)=H1(Q,NΣ)=0. Therefore, the long exact sequence of Galois cohomology attached to (22) can be combined with (21) to yield

Br1(Y)/Br(Q)=H1(Q,Pic(Y¯))=H1(Q,Hom(E¯,E¯)). 23

Now suppose that EndE¯=OK for an imaginary quadratic field K and suppose that there exists a finite extension L/K such that L/Q is Galois and ELEL. Then inflation-restriction gives an exact sequence

0H1(Gal(L/Q),Hom(E¯,E¯))H1(Q,Hom(E¯,E¯))H1(L,Hom(E¯,E¯)), 24

where we view Hom(E¯,E¯) as a twist of OK defined over L. But then H1(L,Hom(E¯,E¯))Homcts(ΓL,Z2)=0. Thus, (24) gives a canonical isomorphism from H1(Gal(L/Q),Hom(E¯,E¯)) to H1(Q,Hom(E¯,E¯)). Plugging this into (23) gives

Br1(Y)/Br(Q)=H1(Q,Hom(E¯,E¯))=H1(Gal(L/Q),Hom(E¯,E¯)). 25

Theorem 5.1

Let dZ>0 be squarefree so that K=Q(-d) is an imaginary quadratic field. Let E/Q be an elliptic curve with EndE¯=OK, let E/Q be the quadratic twist of E by aQ× and let Y=Kum(E×E). Then

Br1(Y)/Br(Q)0ifQ(a)Kand-d1mod4;(Z/2Z)2ifQ(a)Kand-d2,3mod4;Z/2Zotherwise.

Proof

Let L=K(a). Then (25) gives

Br1(Y)/Br(Q)=H1(Gal(L/Q),Hom(E¯,E¯))

where the ΓQ-module Hom(E¯,E¯) is the twist of EndE¯=OK by the quadratic character corresponding to Q(a)/Q.

If Q(a)K then Gal(L/Q)=Gal(K/Q) is cyclic and a Tate cohomology calculation gives

H1(Gal(L/Q),Hom(E¯,E¯))0if-d1mod4;Z/2Zif-d2,3mod4.

If Q(a)K then letting G=Gal(L/Q), N=Gal(L/K) and M=Hom(E¯,E¯), inflation-restriction gives an exact sequence

0H1(G/N,MN)H1(G,M)H1(N,M)G/NH2(G/N,MN).

Since M is the twist of OK by the character corresponding to Q(a)/Q, the generator of N acts as multiplication by -1 on M, whereby MN=0 and the inflation-restriction sequence yields an isomorphism from H1(G,M) to H1(N,M)G/N. Now a Tate cohomology calculation gives H1(N,M)=M/2M. One checks that

(M/2M)G/NZ/2Zif-d1mod4(Z/2Z)2if-d2,3mod4.

We now have all the necessary ingredients for the proof of Theorem 1.3.

Proof of Theorem 1.3

By Theorem 5.1, Br1(Y)\Br(Q) contains no elements of odd order. Therefore, the assumption that Br(Y)\Br(Q) contains an element of odd order implies that Br(Y)/Br1(Y)0. Therefore, by Theorems 2.1 and 3.1, Br(Y)/Br1(Y) contains an element of order 3 and K{Q(-2),Q(-11)}, proving (1). Hence, Y=Kum(E×Ea) with a-3K×2Q×, by Proposition 3.4. Now (2) follows from Theorems 2.1 and 3.5, and (3) follows from Theorem 5.1. To prove (4), note that K has class number one and therefore E is a quadratic twist of any chosen elliptic curve E/Q with EndE¯=OK. We take E with affine equation y2=f(x), where f(x) is as stated in Theorem 1.3. Thus, E has equation λy2=f(x) for some λQ× and Kum(E×Ea) is the minimal desingularisation of the projective surface with affine equation λ2u2=af(x)f(t). Replacing u by λu and computing (-3K×2Q×)/Q×2 completes the proof.

Theorem 5.2

For aQ, let Ea/Q be the elliptic curve with affine equation y2=x3+a. Let c,dZ\{0}, let Y=Kum(Ec×Ed) and let α=c/d6. We have

Br1(Y)/Br(Q)Z/3ZifQ(ζ3)Q(α)and[Q(α):Q(ζ3)]=3;Z/2Zif[Q(α):Q]=2andQ(α)Q(ζ3);0otherwise.

Proof

If [Q(α):Q]2 then Ed is a quadratic twist of Ec and the result follows from Theorem 5.1. So henceforth we may assume that [Q(α):Q]{3,6}.

Let ϕα:E¯dE¯c be the isomorphism defined over Q(α) given by (x,y)(α2x,α3y), whereby

Hom(E¯d,E¯c)=End(E¯c)ϕα=Z[ζ3]ϕα

so (25) gives

Br1(Y)/Br(Q)=H1(Gal(L/Q),Z[ζ3]ϕα) 26

where L=Q(ζ3,α). It remains to calculate H1(Gal(L/Q),Z[ζ3]ϕα).

Inflation-restriction gives an exact sequence

0H1(Gal(Q(ζ3)/Q),(Z[ζ3]ϕα)Gal(L/Q(ζ3)))H1(Gal(L/Q),Z[ζ3]ϕα)H1(Gal(L/Q(ζ3)),Z[ζ3]ϕα)Gal(Q(ζ3)/Q)H2(Gal(Q(ζ3)/Q),(Z[ζ3]ϕα)Gal(L/Q(ζ3))). 27

Since Q(α)Q(ζ3), the Galois group Gal(L/Q(ζ3)) acts non-trivially on α and therefore on ϕα, and we have (Z[ζ3]ϕα)Gal(L/Q(ζ3))=0. Thus, (27) yields

H1(Gal(L/Q),Z[ζ3]ϕα)=H1(Gal(L/Q(ζ3)),Z[ζ3]ϕα)Gal(Q(ζ3)/Q). 28

Since Gal(L/Q(ζ3)) is cyclic, we can use Tate cohomology to compute H1(Gal(L/Q(ζ3)),Z[ζ3]ϕα).

First suppose that Gal(L/Q(ζ3))Z/3Z. Then we can choose a generator σ of Gal(L/Q(ζ3)) that sends ϕα to ζ3ϕα. Therefore,

H1(Gal(L/Q(ζ3)),Z[ζ3]ϕα)(Z[ζ3]ϕα)/(ζ3-1)(Z[ζ3]ϕα)(Z/3Z)·ϕα. 29

The first isomorphism in (29) is induced by sending a 1-cocycle to its value at σ. Let f:Gal(L/Q(ζ3))Z[ζ3]ϕα be the 1-cocycle sending σ to ϕα. To determine whether H1(Gal(L/Q(ζ3)),Z[ζ3]ϕα)Gal(Q(ζ3)/Q) is trivial or isomorphic to Z/3Z, we just have to check whether the class of f in H1(Gal(L/Q(ζ3)),Z[ζ3]ϕα) is fixed by the action of Gal(Q(ζ3)/Q). Let τGal(L/Q) be such that its image in Gal(Q(ζ3)/Q) is non-trivial. Then τστ-1=σ-1 and the 1-cocycle property gives

f(τ-1στ)=f(σ-1)=-σ-1(f(σ))=-ζ32ϕα.

Therefore,

fτ(σ)=τ·f(τ-1στ)=τ·(-ζ32ϕα)=-ζ3τ·ϕα, 30

by definition of the action of τ on H1(Gal(L/Q(ζ3)),Z[ζ3]ϕα) in terms of its actions on Gal(L/Q(ζ3)) and on Z[ζ3]ϕα=Hom(E¯d,E¯c).

If [Q(α):Q]=3 then we may assume that τ acts trivially on ϕα, whereby (30) gives fτ(σ)=-ζ3ϕα. Hence, fτ and f are not cohomologous (as can be seen using (29), for example). Therefore, H1(Gal(L/Q(ζ3)),Z[ζ3]ϕα)Gal(Q(ζ3)/Q)=0 and (26) and (28) give

Br1(Y)/Br(Q)=H1(Gal(L/Q(ζ3)),Z[ζ3]ϕα)Gal(Q(ζ3)/Q)=0.

On the other hand, if [Q(α):Q]=6 then Q(ζ3)Q(α)=L and in fact Q(ζ3)=Q(α3). In this case we may assume that τ(α)=-α and hence τ·ϕα=-ϕα. Therefore, (30) gives fτ(σ)=ζ3ϕα and hence fτ is cohomologous to f (as can be seen using (29), for example). Therefore, H1(Gal(L/Q(ζ3)),Z[ζ3]ϕα)Gal(Q(ζ3)/Q)(Z/3Z)·ϕα and the result follows from (26) and (28).

Finally, suppose that Gal(L/Q(ζ3))Z/6Z. Then we can choose a generator of Gal(L/Q(ζ3)) that sends ϕα to -ζ3ϕα. Therefore,

H1(Gal(L/Q(ζ3)),Z[ζ3]ϕα)(Z[ζ3]ϕα)/(ζ3+1)(Z[ζ3]ϕα)=0.

By (26) and (28), our proof is complete.

Theorem 1.6 now follows easily from Theorem 5.2.

Proof of Theorem 1.6

We have ε()·24·ε()·cd=(-3)3n·t6 for some n{0,1} and tQ× so

cd=ε()·24·ε()·c2(-3)3n·t6.

Suppose for contradiction that Q(ζ3)Q(c/d6). Then Q(ζ3)=Q(c/d) and hence c/d-3·Q×2, which is evidently not the case. Furthermore, [Q(c/d6):Q]=2 if and only if c/d lies in Q×3, if and only if c2·ε()·Q×3. Now the result follows from Theorem 5.2.

For completeness, we also include the calculation of the algebraic part of the Brauer group in the case of CM by Z[i].

Theorem 5.3

For aQ, let Ea/Q be the elliptic curve with affine equation y2=x3-ax. Let c,dZ\{0}, let Y=Kum(Ec×Ed) and let α=c/d4. We have

Br1(Y)/Br(Q)(Z/2Z)2if[Q(α):Q]=2andQ(α)Q(i);Z/2Zotherwise.

Proof

The proof is very similar to that of Theorem 5.2 so we shall be brief. Given Theorem 5.1, our concern is the case [Q(α):Q]=4. Let ϕα:E¯dE¯c be the isomorphism defined over Q(α) given by (x,y)(α2x,α3y), whereby

Br1(Y)/Br(Q)=H1(Gal(L/Q),Z[i]ϕα)

where L=Q(i,α). Inflation-restriction gives

H1(Gal(L/Q),Z[i]ϕα)=H1(Gal(L/Q(i)),Z[i]ϕα)Gal(Q(i)/Q)Z/2Z.

Constant evaluation maps

In this section, we prove the following analogue of [12, Theorem 1.2(i)].

Theorem 6.1

Let KQ(-11) be an imaginary quadratic field and let Y=Kum(E×E) for elliptic curves E,E over Q with EndE¯=EndE¯=OK. Let be an odd prime and if K=Q(ζ3) assume that >3. Suppose that ABr(Y) and let v be a place of Q. Then the evaluation map evA,v:Y(Qv)Br(Qv) is constant.

The statement for odd primes of good reduction for Y follows immediately from [3, Theorem D, Remark 1.6.2], and in the remaining cases the proof boils down to showing -divisibility for the sets of local points of certain elliptic curves with bad reduction. We will need the following elementary lemmas.

Lemma 6.2

Let K be an imaginary quadratic field and let Y=Kum(E×E) for elliptic curves E,E over Q with EndE¯=EndE¯=OK. Let be an odd prime and if K=Q(ζ3) assume that >3. Suppose that ABr(Y)\Br(Q). Then 7, K is one of Q(ζ3),Q(i),Q(-2),Q(-11) and

Br(Y)/Br(Q)=Br(Y)/Br1(Y)=HomΓQ(E,E)-Z/Z. 31

Furthermore, if φHomΓQ(E,E)- is non-zero then it is an isomorphism.

Proof

Suppose for contradiction that ABr1(Y). Then Theorems 5.15.2 and 5.3 show that K=Q(ζ3) and =3, contradicting our assumptions. Consequently, ABr(Y)\Br1(Y) and Theorem 1.1 shows that K{Q(ζ3),Q(i),Q(-7),Q(-2),Q(-11)} and 7. Since is odd, Theorem 1.3 shows that KQ(-7).

For K=Q(ζ3), (31) follows from Theorems 5.2, 2.1 and Proposition 4.6. For K=Q(i), (31) follows from (2) and [12, Sect. 4]. For K equal to Q(-2) or Q(-11), (31) follows from Theorems 1.3, 2.1 and Corollary 2.5 (note that must equal 3 in this case).

Now let 0φHomΓQ(E,E)-. We claim that φ is an isomorphism. For K=Q(ζ3), this follows from the explicit generator given in Proposition 4.6. For K=Q(i), it is proved in [12, Sect. 5.1], and we adapt the argument therein for K equal to Q(-2) or Q(-11), observing that 3 splits in K/Q as 3=λλ¯ and we can write E3=EλEλ¯ but neither factor is a ΓQ-submodule of E3 and therefore neither factor can be the kernel of φ. Thus, the restrictions of φ to Eλ and Eλ¯ are isomorphisms EλEλ¯ and Eλ¯Eλ. Hence, φ is an isomorphism.

Lemma 6.3

Let E/Q be an elliptic curve with EndE¯=Z[-2]. Then 2 is a prime of bad reduction for E. Furthermore, if p3 is a prime of bad reduction for E then E(Qp) is 3-divisible.

Proof

Since Q(-2) has class number one, E is a quadratic twist of y2=x3+4x2+2x by some squarefree aZ\{0}. Hence E has an affine equation y2=x3+4ax2+2a2x. One checks that this equation is minimal and that 2 is a prime of bad reduction. Running Tate’s algorithm (see [25, IV.9], for example) shows that if p is a prime of bad reduction for E then the reduction type is additive and the Tamagawa number |E(Qp)/E0(Qp)| is either 2 or 4, so it suffices to show that E0(Qp) is 3-divisible. This follows from the description of E0(Qp) given in [23, Theorem 1].

Remark 6.4

The analogue of Lemma 6.3 for elliptic curves with complex multiplication by Z[1+-112] is false. For example, the elliptic curve E/Q with LMFDB label 121.b2, which has affine equation y2+y=x3-x2-7x+10, has good (supersingular) reduction at 2 and E(Q2)/E1(Q2)Z/3Z, so E(Q2) is not 3-divisible. This is why Q(-11) was excluded from Theorem 6.1; further investigation would be needed to determine whether the statement still holds in that case.

Lemma 6.5

Let p and be distinct primes with >3. Let aQp×, let Ea denote the elliptic curve with affine equation y2=x3+a and if p is odd suppose that Ea/Qp has bad reduction. Then Ea(Qp) is -divisible.

Proof

First suppose that Ea/Qp has bad reduction. An examination of Tate’s algorithm (see [25, IV.9], for example) shows that Ea has additive reduction at p and the Tamagawa number [Ea(Qp):E0a(Qp)] is at most 4. In particular, the Tamagawa number is coprime to and thus the claim is proved once we have shown that E0a(Qp) is -divisible. The -divisibility of E0a(Qp) follows from the description of this group given in [23, Theorem 1].

Now suppose that p=2 and Ea/Q2 has good reduction. Tate’s algorithm shows that this can only happen if ord2(a)=4 and Ea has a minimal Weierstrass equation of the form y2+y=x3+b for some bZ2. The standard filtration on the Q2 points of Ea is

Ea(Q2)E1a(Q2)E2a(Q2)

where E1a(Q2) denotes the kernel of the reduction map. The theory of formal groups (see [24, IV, VII], for example) shows that E2a(Q2)4Z2, which is -divisible, and E1a(Q2)/E2a(Q2)Z/2Z. Therefore, E1a(Q2) is -divisible. Finally, Ea(Q2)/E1a(Q2)E~(F2)Z/3Z, whence it follows that Ea(Q2) is -divisible, as required.

Proof of Theorem 6.1

The statement for the infinite place is clear, since Br(R)=Z/2 has trivial -torsion. The statement for odd primes of good reduction for Y follows from [3, Theorem D, Remark 1.6.2].

If ABr(Q) then there is nothing to prove, so henceforth we will assume that ABr(Y)\Br(Q). Thus, by Lemma 6.2, K is one of Q(ζ3),Q(i),Q(-2) and 7.

For K=Q(i), we have (Br(Y)/Br1(Y))oddZ/Z by (2). For K=Q(-2), Theorem 1.3 shows that =3 and Br(Y)/Br1(Y)Z/3Z. For K=Q(ζ3), we are assuming that >3, and Theorem 1.4 gives Br(Y)/Br1(Y)Z/Z. So from now on let p be a finite prime and if p is odd assume that Y has bad reduction at p. Our task is to show that evA,p is constant.

Let 0φHomΓQ(E,E)- and let B be the element of Br(Y) constructed from φ as in Sect. 2.2. By construction, and Lemma 6.2, B generates Br(Y)/Br1(Y)=Br(Y)/Br(Q). Therefore, it suffices to prove that evB:Y(Qp)Br(Qp) is the zero map.

Let E and E have affine equations E:y2=f(x) and E:y2=g(x), respectively. Then Y is the minimal desingularisation of the projective surface with affine equation

u2=f(x)g(t). 32

Note that if E and E are both replaced by their quadratic twists by some λQ×, with affine equations Eλ:λy2=f(x) and Eλ:λy2=g(x), respectively, then the resulting Kummer surface has affine equation

λ2u2=f(x)g(t)

and the map (x,t,u)(x,t,λu) gives an isomorphism back to the original model, showing that the Kummer surface remains the same.

We adapt the arguments given in [12, Sect. 5] to our setting. Since the evaluation map evB,p:Y(Qp)Br(Qp) is locally constant, it is enough to show that it is zero on all Qp-points R=(x0,t0,u0) satisfying (32). Let δR=g(t0). Again by local constancy, we are free to use the implicit function theorem to replace R by a point R=(x1,t1,u1) satisfying (32), sufficiently close to R, such that δ=δRQ× and u10. Now R gives rise to points P=(x1,u1/δ)Eδ(Qp)\E2δ and Q=(t1,1)Eδ(Qp)\E2δ. Note that Kum(Eδ×Eδ) is the minimal desingularisation of the projective surface with affine equation δ2u2=f(x)g(t) and the map (x,t,u)(x,t,δu) gives a Q-isomorphism from Kum(Eδ×Eδ) to Y that sends the point corresponding to (P,Q)Eδ(Qp)×Eδ(Qp) to R. Let φδHomΓQ(Eδ,Eδ)- denote the isomorphism EδEδ coming from φ. Now (9) shows that

B(R)=χPφδ(χQ)Br(Qp)-1Z/Z 33

where χP is the image of P under χ:Eδ(Qp)H1(Qp,Eδ) and χQ is the image of Q under χ:Eδ(Qp)H1(Qp,Eδ). The maps denoted by χ factor through the quotients Eδ(Qp)/ and Eδ(Qp)/, respectively.

Recall that p is either equal to 2 or a prime of bad reduction for Y. By [18, Lemma 4.2], odd primes of good reduction for an abelian surface are primes of good reduction for the corresponding Kummer surface. Therefore, switching Eδ and Eδ if necessary, if p is odd then we may assume that p is a prime of bad reduction for Eδ. Now Lemmas 6.36.5 and [12, Sect. 5.2] show that Eδ(Qp)/=0 and hence (33) shows that B(R)=0, as required.

Surjective evaluation maps

In this section, we prove Theorem 1.5. The notation and assumptions of that theorem will be in force throughout this section. Furthermore, henceforth let φ be a non-zero element of HomΓQ(E,E)-.

Proposition 7.1

To prove Theorem 1.5, it suffices to prove the existence of PE(Q) and QE(Q) such that

χPφ(χQ)0

where χP is the image of P under χ:E(Q)H1(Q,E) and χQ is the image of Q under χ:E(Q)H1(Q,E).

Proof

Lemma 6.2 shows that φ is an isomorphism. Let BBr(Y) be the element constructed from φ as in Sect. 2.2. By construction, and Lemma 6.2, B generates Br(Y)/Br1(Y)=Br(Y)/Br(Q). Therefore, it suffices to prove Theorem 1.5 with A=B. By continuity, we may assume that P and Q are not 2-torsion points (replacing them with nearby points if necessary). Let [P,Q]Y(Q) denote the point of Y coming from (PQ). Then (9) shows that

0χPφ(χQ)=B([P,Q])Br(Q)-1Z/Z.

Furthermore, for all mZ such that mPE2, we have

B([mP,Q])=χmPφ(χQ)=m(χPφ(χQ)).

Therefore, taking scalar multiples of P gives the desired surjectivity. (Again, we can always substitute a sufficiently close point to avoid any issues with 2-torsion points.)

Finally, it is well known and easy to verify that having a non-constant evaluation map evB,:Y(Q)Br(Q) implies the existence of an adelic point on Y that does not pair to zero with B under the Brauer–Manin pairing (1).

Therefore, our task is to find PE(Q) and QE(Q) such that χPφ(χQ)0, as in Proposition 7.1. We begin by treating the easier case where the order of our Brauer group element splits in the CM field K.

The case of split in K/Q

This section mimics [12, §5.3]. Let be an odd prime number that splits in K/Q, so =λλ¯ for some λOK. This splitting will allow us to replace the cup-product pairing that gives the evaluation map (see (9)) with a non-degenerate pairing H1(Q,Eλ)×H1(Q,Eλ)Br(Q) (see (35) below). The proof of Theorem 1.5 in this setting will then come down to showing that the images of E(Q) and E(Q) in H1(Q,Eλ) and H1(Q,Eλ), respectively, are sufficiently large.

Choose an embedding of K into Q such that λ is a uniformiser of Z and λ¯Z×. Now E=EλEλ¯ as ΓQ-modules and therefore H1(Q,E)=H1(Q,Eλ)H1(Q,Eλ¯). Since the restriction of the skew-symmetric Weil pairing to each of the one-dimensional F-subspaces Eλ and Eλ¯ is trivial, each of the subspaces H1(Q,Eλ) and H1(Q,Eλ¯) is isotropic for the pairing

:H1(Q,E)×H1(Q,E)Br(Q)-1Z/Z

described in (8). By the non-degeneracy of the cup product, these subspaces are maximal isotropic subspaces of H1(Q,E), each of dimension 12dimH1(Q,E). Thus, (8) induces a non-degenerate pairing

:H1(Q,Eλ)×H1(Q,Eλ¯)Br(Q)-1Z/Z. 34

Recall that we have

0φHomΓQ(E,E)-={ψHomΓQ(E,E)ψγ=γ¯ψγOK}.

Then φ=φ+φ where φ is an isomorphism of ΓQ-modules EλEλ¯ and φ is an isomorphism of ΓQ-modules Eλ¯Eλ. Consequently, the induced isomorphism φ:H1(Q,E)H1(Q,E) is a sum of φ:H1(Q,Eλ)H1(Q,Eλ¯) and φ:H1(Q,Eλ¯)H1(Q,Eλ). In conclusion, (34) together with φ induces a non-degenerate pairing

H1(Q,Eλ)×H1(Q,Eλ)Br(Q)-1Z/Z. 35

We will use the non-degeneracy of the pairing (35) to prove Theorem 1.5 via Proposition 7.1 in the case where splits in K/Q. For this to work, we will need to show that the images of E(Q) and E(Q) in H1(Q,Eλ) and H1(Q,Eλ), respectively, are sufficiently large. We will frequently use that H1(Q,Eλ) and E(Q)/ are both maximal isotropic subspaces of H1(Q,E). Moreover,

E(Q)/λH1(Q,Eλ)

and E(Q)/=E(Q)/λE(Q)/λ¯.

We need to treat four cases as delineated in Lemma 6.2: the CM field K is one of Q(ζ3),Q(i),Q(-2),Q(-11).

We begin with the case K=Q(-2). In this case, Theorem 1.3 shows that =3 and Y=Kum(E×Ea) where E/Q is the elliptic curve with affine equation y2=x3+4x2+2x, a{-3,6} and Ea denotes the quadratic twist of E by a.

Proposition 7.2

Let E/Q have affine equation y2=x3+4x2+2x and let Ea denote its quadratic twist by aQ×. Choose an embedding of Q(-2) in Q3 such that -21mod3, whereby λ=1--2 is a uniformiser for Z3 and λ¯=1+-2Z3×. Then

E(Q3)/λE(Q3)/λ¯Z/3Z

and for a{-3,6} we have Ea(Q3)/λ¯=0 and

H1(Q3,Eλa)=Ea(Q3)/λ=Ea(Q3)/3(Z/3Z)2.

Proof

First, we prove the statement for E. This elliptic curve has good reduction at 3 and we have

E(Q3)/E1(Q3)E~(F3)={O,(0,0),(1,1),(1,-1),(-1,1),(-1,-1)}.

Let E^ denote the formal group associated to E. We have isomorphisms of topological groups

E1(Q3)(x,y)-xyE^(3Z3)log3Z3 36

where log denotes the formal logarithm, which is given by a power series T+n2bnnTn for some bnZ, see [24, Proposition IV.5.5]. These isomorphisms respect the action of -2. This means that λ¯=1+-2 acts as a multiplication by a unit in Z3 and λ=1--2 acts as multiplication by a uniformiser and hence

E1(Q3)/λ=E1(Q3)/3Z/3. 37

Now one checks that λ induces an automorphism of E(Q3)/E1(Q3)E~(F3). Thus,

E(Q3)=λE(Q3)+E1(Q3) 38

and

E(Q3)/λ=E1(Q3)/λZ/3

by (37). Furthermore, (38) gives

λ¯E(Q3)=3E(Q3)+λ¯E1(Q3)=3E(Q3)+E1(Q3)

since λ¯ acts as an automorphism on E1(Q3) by our discussion above. Therefore,

E(Q3)/λ¯=E(Q3)/(3E(Q3)+E1(Q3))E~(F3)/3Z/3.

Now we deal with Ea for a{-3,6}. In fact, since -2Q3×2, E-3 and E6 are isomorphic over Q3 and we may take a=-3 in what follows. The elliptic curve E-3 has additive reduction at 3 and the Tamagawa number |E-3(Q3)/E0-3(Q3)| is 2, so we can replace E-3(Q3) by E0-3(Q3) in our calculations. As above, we find that

E1-3(Q3)/λ=E1-3(Q3)/3Z/3 39

and

E1-3(Q3)/λ¯=0. 40

Furthermore, E0-3(Q3)/E1-3(Q3)E~ns-3(F3)={O,(1,1),(1,-1)} and -2 acts as the identity on E~ns-3(F3), whereby

E0-3(Q3)=λ¯E0-3(Q3)+E1-3(Q3) 41

and hence

E0-3(Q3)/λ¯=E1-3(Q3)/λ¯=0

by (40). Moreover, (41) and (39) give

λE0-3(Q3)=3E0-3(Q3)+λE1-3(Q3)=3E0-3(Q3).

Now [23, Theorem 1] shows that E0-3(Q3)3Z3×Z/3Z and hence

E0-3(Q3)/λ=E0-3(Q3)/3(Z/3Z)2.

Now since E-3(Q3)/3=E-3(Q3)/λH1(Q3,Eλ-3), and E-3(Q3)/3 and H1(Q3,Eλ-3) are both maximal isotropic subspaces of H1(Q3,E3-3), they must be equal.

Corollary 7.3

Let E/Q have affine equation y2=x3+4x2+2x and let Ea denote its quadratic twist by aQ×. Then for a{-3,6}, there exist PE(Q3) and QEa(Q3) such that χPφ(χQ) is non-zero.

Proof

This follows from the non-degeneracy of (35), with =3 and λ=1--2, and Proposition 7.2.

Next, we treat the case where K=Q(-11). In this case, Theorem 1.3 shows that =3 and Y=Kum(E×Ea) where E/Q has affine equation y2=x3-25·33·11x+24·33·7·112, a{-3,33} and Ea denotes the quadratic twist of E by a. A minimal Weierstrass equation for E is given by y2+y=x3-x2-7x+10, see [27].

Proposition 7.4

Let E/Q have affine equation y2+y=x3-x2-7x+10 and let Ea denote its quadratic twist by aQ×. Choose an embedding of Q(-11) in Q3 such that -111mod3, whereby λ=(1--11)/2 is a uniformiser for Z3 and λ¯=(1+-11)/2Z3×. Then for a{1,-3,33} we have Ea(Q3)/λ¯=0 and

H1(Q3,Eλa)=Ea(Q3)/λ=Ea(Q3)/3Z/3Z.

Proof

Standard calculations similar to those in the proof of Proposition 7.2 show that

E(Q3)/λ¯=0andE(Q3)/λ=E(Q3)/3Z/3.

Now since E(Q3)/3=E(Q3)/λH1(Q3,Eλ) and both are maximal isotropic subspaces of H1(Q3,E3), they must be equal. This completes the proof of the statement for E.

Now we deal with Ea for a{-3,33}. Since -11Q3×2, it suffices to take a=-3. Using [6, §4.3] and Tate’s algorithm, we find that E-3(Q3)=E0-3(Q3). Moreover, the isomorphism (36) respects complex multiplication. Therefore,

E1-3(Q3)/λ¯=0andE1-3(Q3)/λ=E1-3(Q3)/3Z/3.

By [23, Definition 10 and Proposition 11], the map (x,y)-x/y extends to an isomorphism of topological groups E0-3(Q3)E^-3(Z3). Furthermore, [23, Proposition 18] gives an isomorphism of topological groups E^-3(Z3)Z3 extending the isomorphism E^-3(3Z3)3Z3 given by the formal logarithm. Transporting the action of complex multiplication by λ along these isomorphisms gives an endomorphism of Z3 that coincides with multiplication by λ on 3Z3. But this endomorphism must then be multiplication by λ. The same argument applies to λ¯ and hence we have

E0-3(Q3)/λ¯=0andE0-3(Q3)/λ=E0-3(Q3)/3Z/3. 42

The usual argument about maximal isotropic subspaces of H1(Q3,E3-3) completes the proof.

Corollary 7.5

Let E/Q have affine equation y2=x3-x2-7x+10 and let Ea denote its quadratic twist by aQ×. Then for a{-3,33}, there exist PE(Q3) and QEa(Q3) such that χPφ(χQ) is non-zero.

Proof

This follows from the non-degeneracy of (35), with =3 and λ=(1--11)/2, and Proposition 7.4.

Now we treat the case where K=Q(ζ3). By Lemma 6.2, we have 7. In this section, our focus is on the case where splits in K/Q, so we take =7. For aQ×, let Ea denote the ellliptic curve over Q with affine equation y2=x3+a. Now Theorems 1.4 and 5.2 give Y=Kum(Ec×Ed) where c,dQ× satisfy -24·7-1·cd-33Q×6.

Proposition 7.6

For any aQ7×, let E be the elliptic curve with affine equation y2=x3+a. Choose an embedding of Q(ζ3) in Q3 such that ζ32mod7, whereby λ=1-2ζ32=3+2ζ3 is a uniformiser for Z7 and λ¯=1-2ζ3Z7×.

  1. If a2·7·Q7×6 then Ea(Q7)/λ¯=0 and
    H1(Q7,Eλa)=Ea(Q7)/λ=Ea(Q7)/7(Z/7Z)2.
  2. If a-2·Q7×6 then Ea(Q7)/λEa(Q7)/λ¯Z/7Z.

  3. In all other cases, Ea(Q7)/λ¯=0 and
    H1(Q7,Eλa)=Ea(Q7)/λ=Ea(Q7)/7Z/7Z.

Proof

Recall that Ea(Q7)/7 and H1(Q7,Eλa) are both maximal isotropic subspaces of H1(Q7,E7a), and Ea(Q7)/λH1(Q7,Eλa). Therefore, in cases where we show that Ea(Q7)/7=Ea(Q7)/λ, it will follow that this group is also equal to H1(Q7,Eλa).

Since the Q7-isomorphism class of Ea only depends on the class of a in Q7×/Q7×6, we may assume that 0ord7(a)5. The reduction type of Ea is either good or additive. The Tamagawa number |Ea(Q7)/E0a(Q7)| is coprime to 7, so we can replace Ea(Q7) by E0a(Q7) in our calculations. Standard calculations give

λ¯E1a(Q7)=E1a(Q7)andλE1a(Q7)=7E1a(Q7). 43

If a14(mod49), then [23, Proposition 18] shows that the extension 0E1a(Q7)E0a(Q7)E~nsa(F7)0 is split. Computing the action of ζ3 on E~nsa(F7) shows that it coincides with multiplication by 2. Therefore, λ¯E~nsa(F7)=E~nsa(F7) and λE~nsa(F7)=7E~nsa(F7)=0. When combined with  (43), this proves part (1) of the proposition.

In the case where Ea has additive reduction and a14(mod49), we use [23, Definition 10, Propositions 11 and 18] to obtain an isomorphism E0a(Q7)Z7 respecting the action of ζ3, and hence show that λ¯E0a(Q7)=E0a(Q7) and λE0a(Q7)=7E0a(Q7), proving the proposition in this case.

Now suppose that Ea has good reduction so Ea(Q7)/E1a(Q7)E~a(F7). Elementary calculations show that |E~a(F7)| is coprime to 7, unless a-2·Q7×6 when E~a(F7)=E~-2(F7)Z/7Z. Thus, (43) proves the proposition for a-2·Q7×6.

Our final task is to prove part (2). Taking a=-2, we have

E~-2(F7)={O,(3,2),(3,-2),(-2,2),(-2,-2),(-1,2),(-1,-2)}

and multiplication by ζ3 sends (xy) to (2xy), which coincides with multiplication by 4 on E~-2(F7)Z/7Z. Therefore, λ¯E~-2(F7)=0 and hence λ¯E-2(Q7)E1-2(Q7). By (43), this containment is an equality and hence

E-2(Q7)/λ¯=E-2(Q7)/E1-2(Q7)E~-2(F7)Z/7Z.

Moreover, λE~-2(F7)=E~-2(F7) and hence

E-2(Q7)=λE-2(Q7)+E1-2(Q7),

whereby (43) gives

E-2(Q7)/λ=E1-2(Q7)/λZ/7Z.

Corollary 7.7

For aQ×, let Ea denote the ellliptic curve over Q with affine equation y2=x3+a. Let c,dQ× satisfy -24·7-1·cd-33Q×6. Then there exist PEc(Q7) and QEd(Q7) such that χPφ(χQ) is non-zero.

Proof

First note that the relation -24·7-1·cd-33Q×6 implies that c2·7·Q7×6 if and only if d-2·Q7×6. Suppose that c2·7·Q7×6. Then d-2·Q7×6 and the result follows from the non-degeneracy of (35), with =7 and λ=1-2ζ32, and Proposition 7.6(1) and (2). The proof in the case where c-2·Q7×6 follows by symmetry in c and d. In the remaining case, where c and d are in neither 2·7·Q7×6 nor -2·Q7×6, the result follows from the non-degeneracy of (35) and Proposition 7.6(3).

The remaining case is where K=Q(i) and (2) together with the splitting condition gives =5 and Y=Kum(Em1×Em2) where m1,m2Q× satisfy 53m1m2-4Q×4 and Em/Q is the elliptic curve with affine equation y2=x3-mx. This case was tackled by Ieronymou and Skorobogatov in [12], en route to their treatment of diagonal quartic surfaces. In particular, [12, Proposition 5.7 and Corollary 5.8] shows that if m1 and m2 are not in (1±2i)Q5×4 then there exist PEm1(Q5) and QEm2(Q5) such that χPφ(χQ) is non-zero. If m1 or m2 is in (1±2i)Q5×4 then replace m1 and m2 by 52m1 and 52m2, respectively. Now the relation 53m1m2-4Q×4 implies that 52m1 and 52m2 are not in (1±2i)Q5×4 and one can apply [12, Proposition 5.7 and Corollary 5.8] as before. The Kummer surface Y is unchanged when we replace m1 and m2 by 52m1 and 52m2, respectively, because this simply amounts to replacing each of Em1 and Em2 by their quadratic twists by 5Q×/Q×2. We have seen previously (in the proof of Theorem 6.1) that this does not change the Kummer surface.

At this stage, thanks to Proposition 7.1, Corollaries 7.37.57.7 and the above discussion for K=Q(i) and =5, we have proved Theorem 1.5 in all cases where splits in K.

The case of inert in K/Q

This is the hardest case. We will follow the work of Ieronymou and Skorobogatov in [13] and use a result of Harpaz and Skorobogatov (Proposition 7.10 below) to reduce the proof of Theorem 1.5 in this case to the task of showing that the function fields of torsors associated to certain elements of H1(Q5,E5) are not isomorphic. We will avoid difficult calculations with these totally wildly ramified extensions of degree 25 by using quadratic twists to obtain function fields with distinct discriminants.

By Lemma 6.2 and our assumptions in the statement of Theorem 1.5 that is odd, and if K=Q(ζ3) then >3, we have excluded all cases where ramifies in K/Q. For K=Q(ζ3), the evaluations at 3-adic points of Brauer group elements of orders 3 and 9 will be studied in future work.

By Lemma 6.2 and Theorems 1.31.4 and (2), the only cases where the order of our Brauer group element is inert in the CM field K are when =5 and K=Q(ζ3), and when =3 and K=Q(i).

The case where K=Q(i) and =3 was tackled by Ieronymou and Skorobogatov in [13]. In this case, (2) gives Y=Kum(Em1×Em2) where m1,m2Q× satisfy -3m1m2-4Q×4 and Em/Q is the elliptic curve with affine equation y2=x3-mx. By [13, Proposition 2.2], if the 3-adic valuations of m1 and m2 are both non-zero modulo 4, then there exist PEm1(Q3) and QEm2(Q3) such that χPφ(χQ)0. If the 3-adic valuation of m1 or m2 is zero modulo 4, then we can replace m1 and m2 by 32m1 and 32m2, respectively, which does not change the Kummer surface Y, and then apply [13, Proposition 2.2]. Now applying Proposition 7.1 completes the proof of Theorem 1.5 in this case.

Note the qualitative difference in behaviour between the Kummer surfaces Y and the closely related diagonal quartic surfaces studied by Ieronymou and Skorobogatov. In [13, Theorem 2.3], the authors show that for some diagonal quartic surfaces, a Brauer group element of order 3 has constant evaluation on 3-adic points, while for others it attains all three possible values.

The remainder of this section is devoted to proving the last remaining case of Theorem 1.5, wherein K=Q(ζ3) and =5. Henceforth, for aQ×, we use Ea to denote the elliptic curve over Q with affine equation y2=x3+a. Theorems 1.4 and 5.2 give Y=Kum(Ec×Ed) where c,dQ× satisfy 24·5·cd-33Q×6. The work below should be compared to [13].

We begin by showing that the image of Ea(Q5) in H1(Q5,E5a) is a one-dimensional vector space over F5.

Lemma 7.8

Let aQ5×. Then Ea(Q5)/5Z/5Z as abelian groups.

Proof

An inspection of Tate’s algorithm shows that Ea/Q5 has either good or additive reduction and in all cases the Tamagawa number |Ea(Q5)/E0a(Q5)| is coprime to 5. So it is enough to show that E0a(Q5)/5Z/5. In the case of additive reduction, this follows from [23, Theorem 1]. Now suppose that Ea/Q5 has good reduction. Then we compute Ea(Q5)/E1a(Q5)Ea~(F5)Z/6Z, so it suffices to show that E1a(Q5)/5Z/5Z. The theory of formal groups gives E1a(Q5)5Z5.

Proposition 7.9

Let c,dQ× be such that 24·5·cd-33Q×6. Let P and Q generate Ec(Q5)/5 and Ed(Q5)/5, and let χP and χQ denote their images in H1(Q5,E5c) and H1(Q5,E5d), respectively. To prove Theorem 1.5 for Y=Kum(Ec×Ed) and =5, it suffices to show that φ(χQ) is not a scalar multiple of χP.

Proof

This follows from Proposition 7.1, Lemma 7.8 and the fact that the image of Ea(Q5)/5 in H1(Q5,E5a) is a maximal isotropic subspace for the pairing (8).

To prove that φ(χQ) is not a scalar multiple of χP (with the notation of Proposition 7.9), we wish to apply the following special case of a result of Harpaz and Skorobogatov.

Proposition 7.10

([10, Corollary 3.7]) Let k be a field of characteristic zero and let M be a finite simple Γk-module, identified with the group of k¯-points of a finite étale commutative group k-scheme GM. Let K be the smallest extension of k such that ΓK acts trivially on M, let G=Gal(K/k) and suppose that H1(G,M)=0. Let α,βH1(k,M) be non-zero. Then the associated torsors Zα and Zβ for GM are integral k-schemes. Furthermore, the following conditions are equivalent:

  1. there exists rR:=EndG(M) such that rα=β;

  2. Rα=RβH1(k,M);

  3. Zα and Zβ are isomorphic as abstract k-schemes.

Lemmas 7.11 and 7.12 below are used to show that the hypotheses relevant for our application of Proposition 7.10 are satisfied.

Lemma 7.11

Let aQ× and let G and H be the images of ΓQ and ΓQ5 in Aut(E5a), respectively. Then [G : H] divides 3.

Proof

We adapt the strategy of the proof of [13, Lemma 2.1]. Multiplying by a 6th power, we may assume that aZ\{0}. Let L=Q(E5a), so G=Gal(L/Q). For any E-primary prime πZ[ζ3] that is coprime to 2·3·5·a and unramified in L/Q(ζ3), the action of FrobπΓQ(ζ3) on E5a is given by multiplication by (the reduction modulo 5 of)

±4aπ¯6π, 44

see [25, Example II.10.6]. Thus, the action of ΓQ(ζ3) on E5a factors through a homomorphism to F5[ζ3]×=F25×.

Let FL be the fixed field of the kernel of the action of ΓQ on End(E5a). The natural map GL(E5a)PGL(E5a) restricts to a surjective homomorphism Φ:GGal(F/Q) with kernel F5×. By Lemma 2.4, End(E5a)=(Z[ζ3]/5)End(E5a)-. Therefore, Q(ζ3)F.

The proof of Proposition 4.6 shows that FrobπGal(L/Q(ζ3)) acts on End(E5a)- as left-multiplication by 24·5·a2π6-1. By the definition of the sextic residue symbol and the fact that these Frobenius elements generate Gal(L/Q(ζ3)), this means that the action of Gal(L/Q(ζ3)) on End(E5a)- is given by the sextic character attached to (24·5·a2)-1, which sends σΓQ(ζ3) to 24·5·a26σ(24·5·a26). Therefore, F=Q(ζ3,24·5·a26).

Now fix an inclusion ΓQ5ΓQ. Let K=Q5(E5a), so H=Gal(K/Q5). Let F5=Q5(ζ3,24·5·a26) be the fixed field of the kernel of the action of ΓQ5 on End(E5a). Write a=5ru for uZ5× and use that every unit in Z5× is a cube to see that F5=Q5(ζ3,52r+16). Therefore, [F5:Q5(ζ3)] is either 2 or 6 depending on whether r1mod3. Thus, Φ(H)=Gal(F5/Q5) has index dividing 3 in Φ(G)=Gal(F/Q). To complete the proof, it remains to show that F5×=ker(Φ)H.

Consider the restriction of Φ to G:=Gal(L/Q(ζ3)). This is a cyclic group of order dividing |F25×|=24, which contains F5× as its unique subgroup of order 4, and Φ:GGal(F/Q(ζ3)) is the quotient by F5×. It suffices to show that F5×H:=Gal(K/Q5(ζ3)). We have

H/(HF5×)Φ(H)=Gal(F5/Q5(ζ3)).

Therefore, [F5:Q5(ζ3)] divides |H| (whereby H has even order) and HF5×. Since H is a subgroup of the cyclic group G of order dividing 24, we can now conclude that F5×H unless H has order 6. Suppose for contradiction that |H|=6. Then |HF5×|=2 and [F5:Q5(ζ3)]=|H/(HF5×)|=3, a contradiction.

Lemma 7.12

Let aQ5× and let H denote the image of ΓQ5 in Aut(E5a). Then

  1. H1(H,E5a)=0;

  2. E5a is a simple ΓQ5-module;

  3. EndH(E5a)=F5.

Proof

Multiplying by a 6th power in Q5×, we may assume that aZ\{0}. Let G denote the image of ΓQ in Aut(E5a). Recall from the proof of Lemma 7.11 that the action of ΓQ(ζ3) on E5a factors through a homomorphism to F5[ζ3]×=F25×. Therefore, G has order dividing 2·24. Now (1) is immediate since HG is a finite group of order coprime to 5 and H1(H,E5a) is killed by |H| and by 5.

The strategy of the rest of the proof is to use Lemma 7.11 to move from the action of ΓQ5 on E5a to that of ΓQ, and from there to the action of ΓQ(ζ3), where we have the explicit action of Frobenius elements on E5a given by (44). Henceforth, let πZ[ζ3] be an E-primary prime lying above a rational prime p that is coprime to 2·3·5·a and completely split in Q(ζ3,4a3)/Q. Then (44) shows that the action of Frobπ on E5a is given by multiplication by the reduction modulo 5 of ±π=x+yζ3 for some x,yZ. Assume in addition that p is inert in Q(5)/Q. Then, writing p=ππ¯=x2-xy+y2, we see that if y0(mod5) then px2mod5 and hence quadratic reciprocity gives 5p=p5=1, contradicting the fact that p is inert in Q(5)/Q. Therefore, y0(mod5). Similarly, x(x-y)0(mod5). By Lemma 7.11, the image of Frobπ3 in Aut(E5a) is contained in H. It acts on E5a as multiplication by the reduction modulo 5 of (x+yζ3)3=x3+y3-3xy2+3xy(x-y)ζ3.

Let M be a ΓQ5-submodule of E5a and suppose TM is non-zero. We have Frobπ3·T=(x3+y3-3xy2+3xy(x-y)ζ3)·TM. Since 3xy(x-y)0(mod5), T and (x3+y3-3xy2+3xy(x-y)ζ3)·T form an F5-basis of E5a and hence M=E5a, proving (2).

For (3), Propositions 2.2 and 4.6 show that

EndΓQ(E5a)=(End(E¯a)/5)ΓQ=(Z[ζ3]/5)ΓQ=F5.

Thus, it suffices to show that EndH(E5a)=EndΓQ(E5a). Clearly,

EndΓQ(E5a)=EndG(E5a)EndH(E5a)

so it suffices to show the reverse inclusion. Let φEndH(E5a). Then for π as above, φ commutes with the image of Frobπ3 in Aut(E5a), which implies that φ commutes with the image of ζ3 in Aut(E5a). Therefore, φ commutes with the image of ΓQ(ζ3) in Aut(E5a). Let τΓQ denote complex conjugation, which generates ΓQ/ΓQ(ζ3). Since τ has order 2, Lemma 7.11 shows that the image of τ in Aut(E5a) lies in H. Therefore, φ commutes with τ and hence with the whole image of ΓQ in Aut(E5a), as required.

For χH1(Q5,E5c), write K(χ) for the function field of the Q5-torsor for the group Q5-scheme E5c corresponding to χ. Note that K(χ) is a finite extension of Q5.

Proposition 7.13

Let c,d,χP,χQ be as in Proposition 7.9. To prove Theorem 1.5 for Y=Kum(Ec×Ed) and =5, it suffices to show that K(χP) is not isomorphic to K(χQ). In particular, it suffices to show that disc(K(χP)) is not equal to disc(K(χQ)).

Proof

Proposition 7.9 tells us that it suffices to show that φ(χQ) is not a scalar multiple of χP. By Lemma 7.12, we can apply Proposition 7.10 to see that this is equivalent to showing that the Q5-torsor for the group Q5-scheme E5c determined by φ(χQ) is not isomorphic (as an abstract Q5-scheme) to that determined by χP. For this, it suffices to show that the function fields of the torsors are not isomorphic as extensions of Q5. Since φ is an isomorphism (by Lemma 6.2), it follows from the construction of the pushforward that χQ and ϕ(χQ) are isomorphic as Q5-schemes.

Note that if we prove Theorem 1.5 with =5 and Y=Kum(Ec×Ed) for a given pair of rational numbers c and d, then it also holds for all multiples of c and d by elements in Q×Q5×6, since the relevant elliptic curves are isomorphic over Q5. Hence, we can reduce to considering c,d{2i·5j0i1,0j5}, since these elements represent all cosets in Q5×/Q5×6. For a{2i·5j0i1,0j5}, Table 1 records a generator Pa for Ea(Q5)/5 and the discriminant of the function field of the torsor χPa, which we denote by disc(K(χPa)).

Table 1.

.

a Pa disc(K(χPa))
1 (152,1+5653) 525
5 (1,1+5) 545
52 (1,1+52) 541
53 (1,1+53) 537
54 (1,1+54) 533
55 (1,1+55) 525
2 (152,1+2·5653) 525
2·5 (1,1+2·5) 545
2·52 (1,1+2·52) 541
2·53 (1,1+2·53) 537
2·54 (1,1+2·54) 533
2·55 (1,1+2·55) 525

Note that in all cases disc(K(χPa))=disc(K(χP2a)).

We are now ready to prove the last remaining case of Theorem 1.5, in which Y=Kum(Ec×Ed) for c,dQ× with 24·5·cd-33Q×6, and =5.

Proof

(Completion of the proof of Theorem 1.5) Since 24·5·cd-33·Q×6, 2Q5×3 and -3·Q5×6=2·Q5×6, Table 1 and the preceding discussion show that disc(K(χPd))=disc(K(χP55c5)), where we can take 55c5 modulo Q5×6.

Let S=Q5×62·Q5×655·Q5×62·55·Q5×6. An inspection of Table 1 shows that disc(K(χPc)) is not equal to disc(K(χP55c5)) unless cS. Thus, by Proposition 7.13, we have proved Theorem 1.5 for cS. Now suppose that cS. Then 53cS so Theorem 1.5 holds for Kum(E53c×E53d). Now recall that Kum(Ec×Ed)Kum(E53c×E53d), so the proof of Theorem 1.5 is complete.

Acknowledgements

We are grateful to David Kurniadi Angdinata, Tudor Ciurca, Netan Dogra, Jack Shotton, Alexei Skorobogatov and Alex Torzewski for useful discussions. We thank the anonymous reviewers for their thorough reading of an earlier draft of this paper and for many insightful comments and suggestions that have significantly improved it, including pointing out an error in a previous version of Theorem 5.2. Mohamed Alaa Tawfik was supported by a University of Reading International Research Studentship and a King’s College London Faculty of Natural, Mathematical and Engineering Sciences Studentship. Rachel Newton was supported by EPSRC grant EP/S004696/1 and EP/S004696/2, and UKRI Future Leaders Fellowship MR/T041609/1 and MR/T041609/2.

Data availability

Data sharing is not applicable to this article as it has no associated datasets.

Footnotes

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

References

  • 1.Alaa Tawfik, M.: Brauer–Manin obstructions for Kummer surfaces of products of CM elliptic curves. PhD Thesis, King’s College London (2024)
  • 2.Balestrieri, F., Johnson, A., Newton, R.: Explicit uniform bounds for Brauer groups of singular K3 surfaces. Ann. Inst. Fourier 73(2), 567–607 (2023) [Google Scholar]
  • 3.Bright, M., Newton, R.: Evaluating the wild Brauer group. Invent. Math. 234, 819–891 (2023) [Google Scholar]
  • 4.Colliot-Thélène, J.-L., Skorobogatov, A.N.: Descente galoisienne sur le groupe de Brauer. J. reine angew. Math. 2013(682), 141–165 (2013) [Google Scholar]
  • 5.Colliot-Thélène, J.-L., Skorobogatov, A.N.: The Brauer-Grothendieck group. Volume 71 of Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge. A Series of Modern Surveys in Mathematics [Results in Mathematics and Related Areas. 3rd Series. A Series of Modern Surveys in Mathematics]. Springer, Cham (2021)
  • 6.Connell, I.: Elliptic Curve Handbook. Preprint https://webs.ucm.es/BUCM/mat/doc8354.pdf
  • 7.Cox, D.A.: Primes of the form Inline graphic. Fermat, Class Field Theory, and Complex Multiplication. Wiley, New York (1989) [Google Scholar]
  • 8.Creutz, B.M.: There are no transcendental Brauer-Manin obstructions on abelian varieties. Int. Math. Res. Not. 9, 2684–2697 (2020) [Google Scholar]
  • 9.Creutz, B., Viray, B.: Degree and the Brauer-Manin obstruction. Appendix by A. Skorobogatov. Algebra Number Theory 12(10), 2445–2470 (2018) [Google Scholar]
  • 10.Harpaz, Y., Skorobogatov, A.: Hasse principle for Kummer varieties. Algebra Number Theory 10(4), 813–841 (2016) [Google Scholar]
  • 11.Ieronymou, E.: Diagonal quartic surfaces and transcendental elements of the Brauer group. J. Inst. Math. Jussieu 4(9), 769–798 (2010) [Google Scholar]
  • 12.Ieronymou, E., Skorobogatov, A.N.: Odd order Brauer-Manin obstruction on diagonal quartic surfaces. Adv. Math. 270, 181–205 (2015) [Google Scholar]
  • 13.Ieronymou, E., Skorobogatov, A.N.: Corrigendum to “Odd order Brauer-Manin obstruction on diagonal quartic surfaces’’. Adv. Math. 307, 1372–1377 (2017) [Google Scholar]
  • 14.Ieronymou, E., Skorobogatov, A.N., Zarhin, Y.G.: On the Brauer group of diagonal quartic surfaces. (With an appendix by Sir Peter Swinnerton-Dyer.). J. Lond. Math. Soc. 83, 659–672 (2011) [Google Scholar]
  • 15.Lang, S.: Complex Multiplication. Grundlehren Math. Wiss., vol. 255. Springer, New York (1983) [Google Scholar]
  • 16.Lemmermeyer, F.: Reciprocity Laws: From Euler to Eisenstein. Springer Monographs in Mathematics. Springer, Berlin, Heidelberg (2000)
  • 17.Manin, Y.: Le groupe de Brauer–Grothendieck en géométrie diophantienne. In: Actes du Congrès International Mathématiciens (Nice, 1970), vol. 1, pp. 401–411. Gauthier-Villars, Paris (1971)
  • 18.Matsumoto, Y.: On good reduction of some K3 surfaces related to abelian surfaces. Tohoku Math J. 67, 83–104 (2015) [Google Scholar]
  • 19.McKinnie, K., Sawon, J., Tanimoto, S., Várilly-Alvarado, A.: Brauer groups on K3 surfaces and arithmetic applications. In: Auel, A., Hassett, B., Várilly-Alvarado, A., Viray, B. (eds.) Brauer Groups and Obstruction Problems. Progr. Math., vol. 320. Birkhäuser, Cham (2017) [Google Scholar]
  • 20.Neukirch, J., Schmidt, A., Wingberg, K.: Cohomology of Number Fields. Grundlehren der Mathematischen Wissenschaften, vol. 323, 2nd edn. Springer, Berlin, Heidelberg (2008)
  • 21.Newton, R.: Transcendental Brauer groups of products of CM elliptic curves. J. Lond. Math. Soc. 93(2), 397–419 (2016) [Google Scholar]
  • 22.Newton, R.: Corrigendum: Transcendental Brauer groups of products of CM elliptic curves. J. Lond. Math. Soc. 110, e12953 (2024). 10.1112/jlms.12953 [Google Scholar]
  • 23.Pannekoek, R.: On Inline graphic-torsion of Inline graphic-adic elliptic curves with additive reduction. arXiv:1211.5833v2
  • 24.Silverman, J.H.: The Arithmetic of Elliptic Curves. Graduate Texts in Mathematics, vol. 106, 2nd edn. Springer, New York (1986) [Google Scholar]
  • 25.Silverman, J.H.: Advanced Topics in the Arithmetic of Elliptic Curves. Graduate Texts in Mathematics, Springer, New York (1994) [Google Scholar]
  • 26.Skorobogatov, A.N., Zarhin, Yu.G.: The Brauer group of Kummer surfaces and torsion of elliptic curves. J. reine angew. Math. 666, 115–140 (2012) [Google Scholar]
  • 27.The LMFDB Collaboration. The L-functions and modular forms database (2023). https://www.lmfdb.org. Accessed 16 Aug 2023
  • 28.Valloni, D.: Complex multiplication and Brauer groups of K3 surfaces. Adv. Math. 385, 107772 (2021) [Google Scholar]
  • 29.Wittenberg, O.: Rational points and zero-cycles on rationally connected varieties over number fields. In: Algebraic Geometry: Salt Lake City 2015, Part 2. In: Proceedings of Symposia in Pure Mathematics vol. 97, 597–635, American Mathematical Society, Providence (2018)

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Data Availability Statement

Data sharing is not applicable to this article as it has no associated datasets.


Articles from Research in Number Theory are provided here courtesy of Springer

RESOURCES