Izvestiya: Mathematics
RUS  ENG    JOURNALS   PEOPLE   ORGANISATIONS   CONFERENCES   SEMINARS   VIDEO LIBRARY   PACKAGE AMSBIB  
General information
Latest issue
Forthcoming papers
Archive
Impact factor
Guidelines for authors
Submit a manuscript

Search papers
Search references

RSS
Latest issue
Current issues
Archive issues
What is RSS



Izv. RAN. Ser. Mat.:
Year:
Volume:
Issue:
Page:
Find






Personal entry:
Login:
Password:
Save password
Enter
Forgotten password?
Register


Izvestiya: Mathematics, 2023, Volume 87, Issue 3, Pages 469–487
DOI: https://doi.org/10.4213/im9332e
(Mi im9332)
 

Isogeny classes and endomorphism algebras of abelian varieties over finite fields

Yu. G. Zarhin

Pennsylvania State University, Department of Mathematics, PA, USA
References:
Abstract: We construct nonisogenous simple ordinary abelian varieties over an algebraic closure of a finite field with isomorphic endomorphism algebras.
Keywords: abelian varieties, endomorphism algebras, isogenies, finite fields.
Funding agency Grant number
Simons Foundation 585711
Max-Planck-Institut für Mathematik
The author was partially supported by Simons Foundation Collaboration grant no. 585711. Part of this work was done during his stay in January–May 2022 at the Max-Planck Institut für Mathematik (Bonn, Germany), whose hospitality and support are gratefully acknowledged.
Received: 15.03.2022
Revised: 10.05.2022
Russian version:
Izvestiya Rossiiskoi Akademii Nauk. Seriya Matematicheskaya, 2023, Volume 87, Issue 3, Pages 56–74
DOI: https://doi.org/10.4213/im9332
Bibliographic databases:
Document Type: Article
UDC: 512.742.2+512.742.7
MSC: 11G10, 11J95, 11G15
Language: English
Original paper language: English

§ 1. Introduction

1.1. If $K$ is a number field, then we write $\mathrm{Cl}(K)$ for the (finite commutative) ideal class group of $K$, $\operatorname{cl}(K)$ for the class number of $K$ (i. e., the cardinality of $\mathrm{Cl}(K)$) and $\exp(K)$ for the exponent of $\mathrm{Cl}(K)$. Clearly, $\exp(K)$ divides $\operatorname{cl}(K)$. (The equality holds if and only if $\mathrm{Cl}(K)$ is cyclic, which is not always the case, see [1], Tables.) In addition, $\exp(K)$ is odd if and only if $\operatorname{cl}(K)$ is odd. We write $\mathcal{O}_K$ for the ring of integers in $K$ and $\operatorname{U}_K$ for the group of units, i. e., the multiplicative group of invertible elements in $\mathcal{O}_K$. As usual, an element of $\operatorname{U}_K$ is called a unit in $K$ or a $K$-unit. It is well known (and can be easily checked) that if a unit $u$ in $K$ is a square in $K$, then it is also a square in $\operatorname{U}_K$.

Let $p$ be a prime and $q$ a positive integer that is a power of $p$. We write $\mathbb{F}_p$ for the $p$-element finite field and $\mathbb{F}_q$ for its $q$-element overfield. As usual, $\overline{\mathbb{F}}_p$ denotes an algebraic closure of $\mathbb{F}_p$, which is also an algebraic closure of $\mathbb{F}_q$. We have

$$ \begin{equation*} \mathbb{F}_p\subset \mathbb{F}_q\subset \overline{\mathbb{F}}_p. \end{equation*} \notag $$
If $X$ is an abelian variety over $\overline{\mathbb{F}}_p$, then we write $\operatorname{End}^{0}(X)$ for its endomorphism algebra $\operatorname{End}(X)\otimes\mathbb{Q}$, which is a finite-dimensional semisimple algebra over the field $\mathbb{Q}$ of rational numbers. If $X$ is defined over $k=\mathbb{F}_q$, then we write $\operatorname{End}_k(X)$ for its ring of $k$-endomorphisms and $\operatorname{End}_k^{0}(X)$ for the $\mathbb{Q}$-algebra $\operatorname{End}_k(X)\otimes\mathbb{Q}$; one may view $\operatorname{End}_k^{0}(X)$ as the $\mathbb{Q}$-subalgebra of $\operatorname{End}^{0}(X)$ with the same $1$.

It is well known that isogenous abelian varieties have isomorphic endomorphism algebras and the same dimension (and $p$-adic Newton polygon). In addition, an abelian variety is simple if and only if its endomorphism algebra is a division algebra over $\mathbb{Q}$. It is also known (Grothendieck–Tate) that $\operatorname{End}^{0}(X)$ uniquely determines the dimension of $X$ [2]. Namely, $2\dim(X)$ is the maximal $\mathbb{Q}$-dimension of a semisimple commutative $\mathbb{Q}$-subalgebra of $\operatorname{End}^0(X)$. However, it turns out that there are nonisogenous abelian varieties over $\overline{\mathbb{F}}_p$ with isomorphic endomorphism algebras.

The aim of this note is to provide explicit examples of such varieties.

Let me start with a classical result of M. Deuring about elliptic curves [3], [4], Chapter 4.

I did not find in the literature the following assertion that complements Proposition 1.2.

Proposition 1.3. Let $K$ be an imaginary quadratic field and $p$ a prime that splits in $K$. Let us put $q=p^{\exp(K)}$.

Then there exists an elliptic curve $E$ that is defined with all its endomorphisms over $\mathbb{F}_{q}$ and such that $\operatorname{End}^0(E) \cong K$.

Remark 1.4. One may deduce from ([5], Satz 3, [6], § 6, Corollary 1, p. 507) that if we put $q_1=p^{\operatorname{cl}(K)}$, then there exists an elliptic curve $E$ that is defined with all its endomorphisms over $\mathbb{F}_{q_1}$ and such that $\operatorname{End}(E)\,{\cong}\, \mathcal{O}_K$ (and therefore $\operatorname{End}^0(E)\,{\cong}\, K$).

The next result is an analogue of Proposition 1.2 for abelian surfaces and quartic fields.

Our main result is the following assertion.

Theorem 1.6. Let $n$ be a positive integer and $K$ be a $\mathrm{CM}$ field that is a cyclic degree $2^n$ extension of $\mathbb{Q}$. Let $K_0$ be the only degree $2^{n-1}$ subfield of $K$, which is the maximal totally real subfield of $K$. Let us put $c=\exp(K)$.

(i) Let $p$ be a prime and $A$ an abelian variety over $\overline{\mathbb{F}}_p$ such that $\operatorname{End}^0(A)$ is isomorphic to $K$.

Then $p$ splits completely in $K$ and $A$ is an ordinary simple abelian variety of dimension $2^{n-1}$.

(ii) Let $p$ be a prime that splits completely in $K$. Let us put $q=p^c$.

Remark 1.7. (a) If $n=1$, then $K$ is an imaginary quadratic field and therefore $K_0=\mathbb{Q}$ and $\operatorname{U}_{\mathbb{Q}}=\{\pm 1\}$. The only (totally) positive unit in $\mathbb{Q}$ is $1$, which is obviously a square in $\mathbb{Q}$. Hence, Propositions 1.2 and 1.3 are the special cases of Theorem 1.6 with $n=1$. On the other hand, Proposition 1.5 follows readily from the special case of Theorem 1.6 with $n=2$.

(b) If $n \geqslant 3$, then the number $2^{2^{n-1}-n}$ of the corresponding isogeny classes is strictly greater than $1$. This gives us examples of nonisogenous abelian varieties over $\overline{\mathbb{F}}_p$, whose endomorphism algebras are isomorphic to $K$ and therefore are mutually isomorphic.

(c) Now let $n$ be an arbitrary positive integer. By Chebotarev’s density theorem, the set of primes that split completely in $K$ is infinite (and even has a positive density $1/2^n$).

Corollary 1.8. Let $r$ be a Fermat prime (e. g., $r=3,5,17,257, 65537$). Let $p$ be a prime that is congruent to $1$ modulo $r$. Let us put

$$ \begin{equation} \operatorname{isg}(r)=\frac{2^{(r-1)/2}}{(r-1)}. \end{equation} \tag{1} $$
Then there are precisely $\operatorname{isg}(r)$ isogeny classes of simple ($(r-1)/2$)-dimensional ordinary abelian varieties $A$ over $\overline{\mathbb{F}}_p$, whose endomorphism algebra
$$ \begin{equation*} \operatorname{End}^{0}(A)=\operatorname{End}(A)\otimes\mathbb{Q} \end{equation*} \notag $$
is isomorphic to the $r$th cyclotomic field $\mathbb{Q}(\zeta_r)$. In addition, if $c=\exp(\mathbb{Q}(\zeta_r))$ and $q=p^c$, then each of these isogeny classes contains an abelian variety that is defined with all its endomorphisms over $\mathbb{F}_{q}$.

Remark 1.9. The congruence condition on $p$ means that $p$ splits completely in $\mathbb{Q}(\zeta_r)$. There are infinitely many such $p$ thanks to Dirichlet’s theorem on primes in an arithmetic progression. More precisely, the set of such primes has density $1/(r- 1)$.

Remark 1.10. It is well known that the property of being simple (respectively, ordinary) is invariant under isogenies.

Remark 1.11. Let $r$ be a Fermat prime. Clearly, $\operatorname{isg}(r)=1$ if and only if $r \leqslant 5$.

Let $p$ be a prime that is congruent to $1\bmod r$. It follows from Theorem 1.6 that $r \leqslant 5$ if and only if there is a precisely one isogeny class of simple ordinary ($(r-1)/2$)-dimensional abelelian varieties over $\overline{\mathbb{F}}_p$, whose endomorphiam algebra is isomorphic to $\mathbb{Q}(\zeta_r)$. In other words, all such abelian varieties are mutually isogenous over $\overline{\mathbb{F}}_p$, if and only if $r \in \{3,5\}$.

Example 1.12. (i) Take $r=3$. We have $\operatorname{isg}(3)=1$. It follows from Remark 1.11 that if $p \equiv 1 \bmod 3$, then all ordinary elliptic curves over $\overline{\mathbb{F}}_p$ with endomorphism algebra $\mathbb{Q}(\zeta_3)$ are isogenous. (This assertion seems to be well known.) This implies that each such elliptic curve is isogenous over $\overline{\mathbb{F}}_p$ to $y^2=x^3-1$.

(ii) Take $r=5$. We have $\operatorname{isg}(5)=1$. It follows from Remark 1.11 that if $p \equiv 1 \bmod 5$, then all abelian varieties over $\overline{\mathbb{F}}_p$ with endomorphism algebra $\mathbb{Q}(\zeta_5)$ are two-dimensional simple ordinary and mutually isogenous. This implies that each such abelian variety is isogenous to the jacobian of the genus $2$ curve $y^2=x^5-1$.

Example 1.13. Let us take $r=17$. Then $\operatorname{cl}(\mathbb{Q}(\zeta_{17}))=1$ [7]. Let us choose a prime $p$ that is congruent to $1$ modulo $17$ (e. g., $p=103$). We have

$$ \begin{equation*} \operatorname{isg}(17)=\frac{2^{8}}{16}=16. \end{equation*} \notag $$
By Theorem 1.6, there are precisely $16$ isogeny classes of simple ordinary $\frac{16}{2}=8$-dimensional abelian varieties over $\overline{\mathbb{F}}_p$ with endomorphism algebras $\mathbb{Q}(\zeta_{17})$. In addition, each of these isogeny classes contains an abelian eightfold that is defined with all its endomorphisms over $\mathbb{F}_{p}$.

This implies that there exist sixteen $8$-dimensional ordinary simple abelian varieties $A_1, \dots, A_{16}$ over $\overline{\mathbb{F}}_p$ that are mutually non-isogenous but each endomorphism algebra $\operatorname{End}^0(A_i)$ is isomorphic to $\mathbb{Q}(\zeta_{17})$ (for all $i$ with $1 \leqslant i \leqslant 16$). In particular,

$$ \begin{equation*} \operatorname{End}^0(A_i) \cong \operatorname{End}^0(A_j) \quad \forall\, i,j,\qquad 1 \leqslant i<j \leqslant 16. \end{equation*} \notag $$
In addition, each $A_i$ and all its endomorphisms are defined over $\mathbb{F}_{p}$. This gives an answer to a question of L. Watson [8].

The following assertion is a natural generalization of Corollary 1.8.

Corollary 1.14. Let $r$ be an odd prime and $r-1=2^n \cdot m$ where $n$ is a positive integer and $m$ is a positive odd integer. Let $\mathbf{H}$ be the only order $m$ subgroup of the cyclic Galois group

$$ \begin{equation*} \operatorname{Gal}(\mathbb{Q}(\zeta_r)/\mathbb{Q})=(\mathbb{Z}/r\mathbb{Z})^* \end{equation*} \notag $$
of order $r-1$. Let
$$ \begin{equation} K=K^{(r)}:=\mathbb{Q}(\zeta_r)^{\mathbf{H}} \end{equation} \tag{2} $$
be the subfield of $\mathbf{H}$-invariants in $\mathbb{Q}(\zeta_r)$.

Then

Remark 1.15. Let $K=K^{(r)}$. It is well known that $r$ is totally ramified in $\mathbb{Q}(\zeta_r)$ and therefore in its subfield $K$ as well. This implies that if $K_0$ is the only degree $2^{n-1}$ subfield of $K$, which is the maximal totally real subfield of $K$, then the quadratic extension $K/K_0$ is ramified. On the other hand, it is known ([9], § 38, [10], § 13, pp. 77, 78) that $\operatorname{cl}(K^{(r)})$ is odd (and therefore $c=\exp(K^{(r)})$ is also odd). It follows from [9], § 37, Satz 42 (see also [10], Corollary 13.10, p. 76) that $K_0$ has units with independent signs (there are units of $K_0$ of every possible signature), which implies (thanks to [10], Lemma 12.2, p. 55) that every totally positive unit in $K_0$ is a square in $K_0$ and therefore is a square in $\operatorname{U}_{K_0}$.

Example 1.16. Let us fix an integer $n \geqslant 2$. Here is a construction of infinitely many mutually nonisomorphic $\mathrm{CM}$ fields that are cyclic degree $2^n$ extensions of $\mathbb{Q}$. Let us consider the infinite (thanks to Dirichlet’s theorem) set of primes $r$ that are congruent to $1+2^n$ modulo $2^{n+1}$. Then $r-1=2^n \cdot m$, where $m$ is an odd positive integer. In light of Corollary 1.14, the corresponding subfield $K^{(r)}$ of $\mathbb{Q}(\zeta_r)$ defined by (2) enjoys the desired properties. Since $K^{(r)}$ is a subfield of $\mathbb{Q}(\zeta_r)$, the field extension $K^{(r)}/\mathbb{Q}$ is ramified precisely at $r$. This implies that the fields $K^{(r)}$ are mutually nonisomorphic (and even linearly disjoint) for distinct $r$.

The paper is organized as follows. In § 2 we review basic results about maximal ideals of $\mathcal{O}_K$. In § 3 we concentrate on the so called ordinary Weil’s $q$-numbers in $K$. In § 4 we discuss simple abelian varieties over $\mathbb{F}_q$, whose Weil’s numbers lie in $K$. In § 5 we discuss some basic facts of Honda–Tate theory (see [11]–[13]). The main results will be proved in § 6.

In what follows, we will freely use the following elementary well-known observation. Any $\mathbb{Q}$-subalgebra with $1$ of a number field $K$ is actually a subfield of $K$; in particular, it is also a number field. For example, if $u$ is an element of $K$, then the subfield $\mathbb{Q}(u)$ generated by $u$ coincides with the $\mathbb{Q}$-subalgebra $\mathbb{Q}[u]$ generated by $u$.

§ 2. Preliminaries

2.1. We keep the notation and assumptions of § 1.1 and Theorem 1.6. As usual, $\mathbb{Q}$, $\mathbb{R}$, $\mathbb{C}$ stand for the fields of rational, real and complex numbers and $\overline{\mathbb{Q}}$ for the (algebraically closed) subfield of all algebraic numbers in $\mathbb{C}$. We write $\mathbb{Z}$ (respectively, $\mathbb{Z}_+$) for the ring of integers (respectively, for the additive semigroup of nonnegative integers). If $T$ is a finite set, then we write $\#(T)$ for the number of its elements.

Recall (see [12], [13]) that an algebraic integer $\pi\in \overline{\mathbb{Q}}$ is called a Weil’s $q$-number if all its Galois-conjugates have the archimedean absolute value $\sqrt{q}$.

Throughout this paper, $n$ is a positive integer and $K$ is a $\mathrm{CM}$ field that is a degree $2^n$ cyclic extension of $\mathbb{Q}$. We view $K$ as a subfield of $\mathbb{C}$; in particular, $K$ is a subfield of $\overline{\mathbb{Q}}$ that is stable under the complex conjugation. We denote by

$$ \begin{equation*} \rho\colon K \to K \end{equation*} \notag $$
the restriction of the complex conjugation to $K$; one may view $\rho$ as an element of order $2$ in the Galois group
$$ \begin{equation*} G:=\operatorname{Gal}(K/\mathbb{Q}), \end{equation*} \notag $$
where $G$ is a cyclic group of order $2^n$.

Remark 2.2. Let $\pi \in K\subset \mathbb{C}$.

We write $W(q,K)$ for the set of Weil’s $q$-numbers in $K$ and $\mu_K$ for the (finite cyclic) multiplicative group of roots of unity in $K$. Clearly, $W(q,K)$ is a finite $G$-stable subset of $\mathcal{O}_K$, which is also stable under multiplication by elements of $\mu_K$. The latter gives rise to the free action of $\mu_K$ on $W(q,K)$ defined by

$$ \begin{equation*} \mu_K \times W(q,K) \to W(q,K), \qquad \zeta, \pi \mapsto \zeta \pi \quad \forall\, \zeta\in \mu_K,\quad \pi \in W(q,K). \end{equation*} \notag $$

Remark 2.3. It is well known (and follows easily from a theorem of Kronecker [14], Chapter IV, § 4, Theorem 8) that $\pi_1,\pi_2 \in W(q,K)$ lie in the same $\mu_K$-orbit (i. e., $\pi_2/\pi_1$ is a root of unity) if and only if the ideals $\pi_1\mathcal{O}_K$ and $\pi_2\mathcal{O}_K$ of $\mathcal{O}_K$ coincide.

Recall (§ 2.1) that $K$ is a subfield of the field $\mathbb{C}$ of complex numbers that is stable under the complex conjugation. Then

$$ \begin{equation*} K_0:=K \cap\mathbb{R} \end{equation*} \notag $$
is a (maximal) totally real number (sub)field, whose degree $[K_0:\mathbb{Q}]$ is
$$ \begin{equation*} \frac{[K:\mathbb{Q}]}{2}=\frac{2^n}{2}=2^{n-1}. \end{equation*} \notag $$

2.4. Recall that the Galois group $G=\operatorname{Gal}(K/\mathbb{Q})$ is a cyclic group of order $2^n$. Hence, it has precisely one element of order $2$ and therefore this element must coincide with the complex conjugation

$$ \begin{equation*} \rho\colon K \to K. \end{equation*} \notag $$
The properties of $G$ imply that every nontrivial subgroup $H$ of $G$ contains $\rho$. It follows that every proper subfield of $K$ is totally real. Indeed, each such subfield is the subfield $K ^H$ of $H$-invariants for a certain nontrivial subgroup $H$ of $G$. Since $H$ contains $\rho$, the subfield $K^H$ consists of $\rho$-invaraiants and therefore is totally real; in particular,
$$ \begin{equation*} K^H \subset \mathbb{R}. \end{equation*} \notag $$

2.5. Let $\ell$ be a prime and $S(\ell)$ be the set of maximal ideals $\mathfrak{P}$ of $\mathcal{O}_K$ that divide $\ell$. Since $K/\mathbb{Q}$ is a Galois extension, $G$ acts transitively on $S(\ell)$. In particular, $\#(S(\ell))$ divides $\#(G)=2^n$. This implies that if $\ell$ splits completely in $K$, i. e.,

$$ \begin{equation*} \#(S(\ell))=2^n=\#(G), \end{equation*} \notag $$
then the action of $G$ on $S(\ell)$ is free.

On the other hand, if a prime $\ell$ does not split completely in $K$, i. e.,

$$ \begin{equation*} \#(S(\ell))<2^n=\#(G), \end{equation*} \notag $$
then the action of $G$ on $S(\ell)$ is not free. Let $H(\ell)$ be the stabilizer of any $\mathfrak{P}\in S(\ell)$, which does not depend on a choice of $\mathfrak{P}$, because $G$ is commutative. Then $H(\ell)$ is a nontrivial subgroup of $G$ and therefore contains $\rho$, i. e.,
$$ \begin{equation*} \rho(\mathfrak{P})=\mathfrak{P} \quad \forall\, \mathfrak{P} \in S(\ell) \end{equation*} \notag $$
if $\ell$ does not split completely in $K$.

Let $e(\ell)$ be the ramification index in $K/\mathbb{Q}$ of $\mathfrak{P}\in S(\ell)$, which does not depend on $\mathfrak{P}$, because $K/\mathbb{Q}$ is Galois. We have the equality of ideals

$$ \begin{equation} \ell \mathcal{O}_K=\prod_{\mathfrak{P}\in S(\ell)}\mathfrak{P}^{e(\ell)}. \end{equation} \tag{4} $$
It follows that $K/\mathbb{Q}$ is unramified at $\ell$ if and only if $e(\ell)=1$. We write
$$ \begin{equation} \operatorname{ord}_{\mathfrak{P}}\colon K^* \twoheadrightarrow \mathbb{Z} \end{equation} \tag{5} $$
for the discrete valuation map attached to $\mathfrak{P}$. We have
$$ \begin{equation} \operatorname{ord}_{\mathfrak{P}}(\ell)=e(\ell) \quad \forall\, \mathfrak{P}\in S(\ell), \end{equation} \tag{6} $$
$$ \begin{equation} \operatorname{ord}_{\mathfrak{P}}(u)\geqslant 0 \quad \forall\, u \in \mathcal{O}_R\setminus\{0\},\quad \mathfrak{P}\in S(\ell), \end{equation} \tag{7} $$
$$ \begin{equation} \operatorname{ord}_{\mathfrak{P}}(\rho(u))=\operatorname{ord}_{\rho(\mathfrak{P})}(u) \quad \forall\, u \in K^*,\quad \mathfrak{P}\in S(\ell). \end{equation} \tag{8} $$

2.6. Let $p$ be a prime, $j$ a positive integer, and $q=p^j$.

Let $\pi \in O_K$ be a Weil’s $q=p^j$-number. Let us consider the ideal $\pi\mathcal{O}_K$ in $\mathcal{O}_K$. Then there is a nonnegative integer-valued function

$$ \begin{equation} d_{\pi}\colon S(p) \to \mathbb{Z}_+, \qquad \mathfrak{P} \mapsto d_{\pi}(\mathfrak{P}):=\operatorname{ord}_{\mathfrak{P}}(\pi) \end{equation} \tag{9} $$
such that
$$ \begin{equation} \pi\mathcal{O}_K=\prod_{\mathfrak{P}\in S(p)}\mathfrak{P}^{d_{\pi}(\mathfrak{P})}. \end{equation} \tag{10} $$
It follows from (3) that
$$ \begin{equation} d_{\pi}(\mathfrak{P})+d_{\pi}(\rho(\mathfrak{P}))=\operatorname{ord}_{\mathfrak{P}}(q)=j \cdot e(\ell) \quad \forall\, \mathfrak{P}\in S(p). \end{equation} \tag{11} $$

Lemma 2.7. Let $\pi \in O_K$ be a Weil’s $q=p^j$-number. If $p$ does not split completely in $K$, then $\pi^2/q$ is a root of unity.

Proof. Since $p$ does not split completely in $K$, it follows from arguments of § 2.4 that
$$ \begin{equation*} \rho(\mathfrak{P})=\mathfrak{P} \quad \forall\, \mathfrak{P}\in S(p). \end{equation*} \notag $$
It follows from (11) that
$$ \begin{equation*} d_{\pi}(\mathfrak{P})=\frac{j\cdot e(p)}{2} \quad \forall\, \mathfrak{P}\in S(p); \end{equation*} \notag $$
in particular, $j$ is even if $e(p)=1$ (i. e., if $K/\mathbb{Q}$ is unramified at $p$). This implies that $\pi^2/q$ is a $\mathfrak{P}$-adic unit for all $\mathfrak{P}\in S(p)$. On the other hand, it follows from (3) that $\pi^2/q$ is an $\ell$-adic unit for all primes $\ell \ne p$. It follows from the very definition of Weil’s numbers that
$$ \begin{equation*} \biggl|\sigma\biggl(\frac{\pi^2}{q}\biggr)\biggr|_{\infty}=1 \quad \forall\, \sigma \in G. \end{equation*} \notag $$
(Here $|\,{\cdot}\,|_{\infty}\colon \mathbb{C} \to \mathbb{R}_+$ is the standard archimedean value on $\mathbb{C}$.) Now it follows from a classical theorem of Kronecker [14], Chapter IV, § 4, Theorem 8, that $\pi^2/q$ is a root of unity. $\Box$

Lemma 2.8. Suppose that a prime $p$ completely splits in $K$. (In particular, $K/\mathbb{Q}$ is unramified at $p$.) Let $\pi \in O_K$ be a Weil’s $q=p^j$-number.

Then either $\mathbb{Q}(\pi)=K$ or $j$ is even and $\pi=\pm p^{j/2}$ .

Proof. So, $K/\mathbb{Q}$ is unramified at $p$, i. e., $e(p)=1$ and
$$ \begin{equation} p\mathcal{O}_K=\prod_{\mathfrak{P}\in S(p)}\mathfrak{P}. \end{equation} \tag{12} $$
This implies that
$$ \begin{equation} q\mathcal{O}_K=\prod_{\mathfrak{P}\in S(p)}\mathfrak{P}^{j}. \end{equation} \tag{13} $$
Since $p$ splits completely in $K$, the group $G$ acts freely on $S(p)$, in light of § 2.5. In particular,
$$ \begin{equation} \mathfrak{P} \ne \rho(\mathfrak{P}) \quad \forall\, \mathfrak{P}\in S(p). \end{equation} \tag{14} $$

If the subfield $\mathbb{Q}(\pi)$ of $K$ does not coincide with $K$, then it is totally real, thanks to arguments of § 2.4. This implies that $\rho(\pi)=\pi$. It follows from (3) that $\pi^2=q$, i. e., $\pi=\pm p^{j/2}$. This implies that the ideal $q\mathcal{O}_K$ is a square. It follows from (13) that $j$ is even. $\Box$

2.9. Suppose that a prime $p$ completely splits in $K$.

Definition 2.10. Let $\pi \in O_K$ be a Weil’s $q=p^j$-number. We say that $\pi$ is an ordinary Weil’s $q$-number if the “slope” $\operatorname{ord}_{\mathfrak{P}}(\alpha)/\operatorname{ord}_{\mathfrak{P}}(q)$ is an integer for all $\mathfrak{P}\in S(p)$.

It (is well known and) follows from (3), (7) and (8) that if $\pi$ is an ordinary Weil’s $q$-number, then

$$ \begin{equation} \frac{\operatorname{ord}_{\mathfrak{P}}(\pi)}{\operatorname{ord}_{\mathfrak{P}}(q)}=0 \quad \text{or}\quad 1. \end{equation} \tag{15} $$

§ 3. Equivalence classes of ordinary Weil’s $q$-numbers

Let $p$ be a prime that splits completely in $K$. Throughout this section, by Weil’s numbers we mean Weil’s $q$-numbers where $q$ is a power of $p$. We write $W(q,K)$ for the set of Weil’s $q$-numbers in $K$. We write $\mu_K$ for the (finite cyclic) multiplicative group of roots of unity in $K$.

Definition 3.1. Let $q$ and $q'$ be integers $>1$ that are integral powers of $p$. Let $\pi \in K$ (respectively, $\pi'\in K$) be a Weil’s $q$-number (respectively, Weil’s $q'$-number). Following Honda [12], we say that $\pi$ and $\pi'$ are equivalent, if there are positive integers $a$ and $b$ such that $\pi^a$ is Galois-conjugate to ${\pi'}^b$.

Clearly, if $\pi$ and $\pi'$ are equivalent, then $\pi$ is ordinary if and only if $\pi'$ is ordinary. In order to classify ordinary Weil’s numbers in $K$ up to equivalence, we introduce the following notion that is inspired by the notion of $\mathrm{CM}$ type for complex abelian varieties [15] (see also [12], § 1, Theorem 2, and [13], § 5).

Definition 3.2. We call a subset $\Phi\subset S(p)$ a $p$-type if $S(p)$ is a disjoint union of $\Phi$ and $\rho(\Phi)$.

Clearly, $\Phi\subset S(p)$ is a $p$-type if and only if the following two conditions hold (recall that $[K:\mathbb{Q}]=2^n$).

(i) $\#(\Phi)=2^{n-1}$.

(ii) If $\mathfrak{P} \in \Phi$, then $\rho(\mathfrak{P}) \notin \Phi$.

It is also clear that $\Phi\subset S(p)$ is a $p$-type if and only if $\rho(\Phi)$ is a $p$-type.

Let $H(p)$ be the set of nonzero ideals $\mathfrak{B}$ of $\mathcal{O}_K$ such that

$$ \begin{equation*} \mathfrak{B}\cdot \rho(\mathfrak{B})=p \cdot \mathcal{O}_K. \end{equation*} \notag $$
In light of (12) and (14), an ideal $\mathfrak{B}$ of $\mathcal{O}_K$ lies in $H(p)$ if and only if there exists a $2^{n-1}$-element subset $\Phi=\Phi(\mathfrak{B})$ of $H(p)$ that meets every $\rho$-orbit of $S(p)$ at exactly one place and
$$ \begin{equation} \mathfrak{B}=\prod_{\mathfrak{P}\in \Phi(\mathfrak{B})}\mathfrak{P}. \end{equation} \tag{16} $$
Such a $\Phi(\mathfrak{B})$ is uniquely determined by $\mathfrak{B}\in H(p)$: namely, it coincides with the set of maximal ideals in $\mathcal{O}_K$ that contain $\mathfrak{B}$. This implies that
$$ \begin{equation} \#(H(p))=2^{2^{n-1}}. \end{equation} \tag{17} $$
Clearly,
$$ \begin{equation} \Phi(\sigma(\mathfrak{B}))=\sigma(\Phi(\mathfrak{B})) \quad \forall \, \sigma \in G. \end{equation} \tag{18} $$

Lemma 3.3. Let $m$ be a positive integer and $\pi$ be a Weil’s $q=p^m$-number in $K$. Then the following conditions are equivalent:

If these equivalent conditions hold, then such an ideal $\mathfrak{B}$ is unique and

$$ \begin{equation*} \Phi(\mathfrak{B})=\Psi(\pi). \end{equation*} \notag $$

Proof. We have
$$ \begin{equation} \pi\mathcal{O}_K=\prod_{\mathfrak{P}\in S(p)}\mathfrak{P}^{d(\mathfrak{P})}, \end{equation} \tag{21} $$
for some $d(\mathfrak{P}) \in \mathbb{Z}_+$ such that
$$ \begin{equation} d(\mathfrak{P})+d(\rho(\mathfrak{P}))=m, \end{equation} \tag{22} $$
$$ \begin{equation} \frac{\operatorname{ord}_{\mathfrak{P}}(\pi)}{\operatorname{ord}_{\mathfrak{P}}(q)} =\frac{d(\mathfrak{P})}{m} \quad \forall \, \mathfrak{P}\in S(p). \end{equation} \tag{23} $$
This implies that
$$ \begin{equation} \Psi(\pi):=\{\mathfrak{P}\in S(p)\mid d(\mathfrak{P})=m\}\subset S(p). \end{equation} \tag{24} $$
Combining (24) with (22), we obtain that
$$ \begin{equation} \rho(\Psi(\pi)):=\{\mathfrak{P}\in S(p)\mid d(\mathfrak{P})=0\}=\biggl\{\mathfrak{P}\in S(p)\biggm| \frac{\operatorname{ord}_{\mathfrak{P}}(\pi)}{\operatorname{ord}_{\mathfrak{P}}(q)}=0 \biggr\}\subset S(p); \end{equation} \tag{25} $$
in particular, the subsets $\Psi(\pi)$ and $\rho(\Psi(\pi))$ do not meet each other. In light of (20) and (25) combined with (15), $\pi$ is ordinary if and only if $S(p)$ is a disjoint union of $\Psi(\pi)$ and $\rho(\Psi(\pi))$, i. e., $\Psi(\pi)$ is a $p$-type. This proves the equivalence of (i) and (iii). If (i) and (iii) hold, then it follows from (21) that
$$ \begin{equation*} \pi \mathcal{O}_K=\prod_{\mathfrak{P}\in \Psi(\pi)}\mathfrak{P}^m=\mathfrak{B}^m, \quad \text{where } \mathfrak{B}:=\prod_{\mathfrak{P}\in \Psi(\pi)}\mathfrak{P}. \end{equation*} \notag $$
Since $\Psi(\pi)$ is a $p$-type, $\mathfrak{B}\in H(p)$ and, obviously, $\Phi(\mathfrak{B})=\Psi(\pi)$. This proves that equivalent (i) and (iii) imply (ii).

Let us assume that (ii) holds. This means that there is $\mathfrak{B} \in H(p)$ that satisfies (19). This implies that

$$ \begin{equation*} \mathfrak{B} =\prod_{\mathfrak{P}\in \Phi(\mathfrak{B})}\mathfrak{P}, \qquad \pi\mathcal{O}_K=\mathfrak{B}^m=\prod_{\mathfrak{P}\in \Phi(\mathfrak{B})}\mathfrak{P}^m. \end{equation*} \notag $$
It follows that
$$ \begin{equation*} \begin{alignedat}{2} \frac{\operatorname{ord}_{\mathfrak{P}}(\pi)}{\operatorname{ord}_{\mathfrak{P}}(q)} &=1 &\quad \forall \, \mathfrak{P} &\in \Phi(\mathfrak{B}), \\ \frac{\operatorname{ord}_{\mathfrak{P}}(\pi)}{\operatorname{ord}_{\mathfrak{P}}(q)}&=0 &\quad \forall \, \mathfrak{P}&\notin \Phi(\mathfrak{B}). \end{alignedat} \end{equation*} \notag $$
This implies that $\pi$ is ordinary and therefore (ii) implies (i). So, we have proved the equivalence of (i), (ii), (iii). The uniqueness of such $\mathfrak{B} $ is obvious. $\Box$

Lemma 3.4. The natural action of $G$ on $H(p)$ is free. In particular, $H(p)$ partitions into a disjoint union of $2^{2^{n-1}-n}$ orbits of $G$, each of which consists of $2^n$ elements.

Proof. Suppose that there exists $\mathfrak{B}\in H(p)$ such that its stabilizer
$$ \begin{equation*} G_{\mathfrak{B}}=\{\sigma \in G\mid \sigma(\mathfrak{B})=\mathfrak{B}\} \end{equation*} \notag $$
is a nontrivial subgroup. Then $G_{\mathfrak{B}}$ must contain $\rho$, thanks to the arguments of § 2.4. This means that $\rho(\mathfrak{B})=\mathfrak{B}$ and therefore
$$ \begin{equation*} p\cdot \mathcal{O}_K=\mathfrak{B}\cdot \rho(\mathfrak{B})=\mathfrak{B}^2, \end{equation*} \notag $$
which is not true, since $p$ is unramified in $K$. The obtained contradiction proves that the action of $G$ on $H(p)$ is free. Hence, every $G$-orbit in $H(p)$ consists of $\#(G)=2^n$ elements and the number of such orbits is
$$ \begin{equation*} \frac{\#(H(p))}{\#(G)}=\frac{2^{2^{n-1}}}{2^n}=2^{2^{n-1}-n}. \end{equation*} \notag $$
Lemma is proved. $\Box$

In what follows we define (noncanonically) certain $G$-equivariant injective maps $\mathcal{Z}$, $\Pi$ and $\Pi_1$ from $H(p)$ to $K$; they will play a crucial role in the classification of ordinary Weil’s numbers in $K$ up to equivalence.

Corollary 3.5. Let $c=\exp(K)$. Then there exists a $G$-equivariant map

$$ \begin{equation} \mathcal{Z}\colon H(p) \hookrightarrow \mathcal{O}_K \setminus\{0\}\subset \mathcal{O}_K\subset K \end{equation} \tag{26} $$
such that $\mathcal{Z}(\mathfrak{B})$ is a generator of $\mathfrak{B}^c$ for all $\mathfrak{B}\in H(p)$.

Proof. We define $\mathcal{Z}$ separately for each $G$-orbit $O\subset H(p)$. Pick $\mathfrak{B}_O \in O$ and choose a generator $z_O$ of the principal ideal $\mathfrak{B}_O^c$. In light of Lemma 3.4, if $\mathfrak{B} \in O$, then there is precisely one $\sigma \in G$ such that $\mathfrak{B}=\sigma(\mathfrak{B}_O)$. This implies that
$$ \begin{equation*} \mathfrak{B}^c=\sigma(\mathfrak{B}_O)^c=\sigma(\mathfrak{B}_O^c)=\sigma(z_O)\mathcal{O}_K, \end{equation*} \notag $$
i. e., $\sigma(z_O)$ is a generator of $\mathfrak{B}^c$. It remains to put
$$ \begin{equation*} \mathcal{Z}(\mathfrak{B}):=\sigma(z_O). \end{equation*} \notag $$
Corollary is proved. $\Box$

Theorem 3.6. Let us put

$$ \begin{equation*} c:=\exp(K), \qquad q:=p^c. \end{equation*} \notag $$
Let $K_0=K^{\rho}$ be the maximal totally real subfield of $K$.

There exists an injective map

$$ \begin{equation} \Pi\colon H(p) \hookrightarrow W(q^2,K), \qquad \mathfrak{B} \mapsto \Pi(\mathfrak{B}) \end{equation} \tag{27} $$
that enjoys the following properties.

(0) $\Pi$ is $G$-equivariant, i. e.,

$$ \begin{equation*} \Pi(\sigma(\mathfrak{B}))=\sigma(\Pi(\mathfrak{B})) \quad \forall \, \sigma\in G, \quad \mathfrak{B}\in H(p). \end{equation*} \notag $$

(i) For all $\mathfrak{B} \in H(p)$ the ideal $\Pi(\mathfrak{B})\mathcal{O}_K$ coincides with $\mathfrak{B}^{2 c}$.

(ii) The image $\Pi(H(p))$ consists of ordinary Weil’s $q^2$-numbers.

(iii) If $\pi'$ is an ordinary Weil’s $p^m$-number in $K$, then there exists precisely one $ \mathfrak{B} \in H(p)$ such that the ratio $(\pi')^{2c}/\Pi(\mathfrak{B})^m$ is a root of unity.

(iv) Let $\mathfrak{B}_1, \mathfrak{B}_2 \in H(p)$. Then Weil’s $q^2$-numbers $\Pi(\mathfrak{B}_1)$ and $\Pi(\mathfrak{B}_2)$ are equivalent if and only if $\mathfrak{B}_1$ and $\mathfrak{B}_2$ lie in the same $G$-orbit.

(v) If $h$ is a positive integer, then the subfield $\mathbb{Q}(\Pi(\mathfrak{B})^h)$ of $K$ generated by $\Pi(\mathfrak{B})^h$ coincides with $K$.

(vi) Suppose that every totally positive unit in $\operatorname{U}_{K_0}$ is a square in $K_0$ (and therefore in $\operatorname{U}_{K_0}$). Then there exists a map

$$ \begin{equation*} \Pi_0\colon H(p) \to W(q,K) \end{equation*} \notag $$
that enjoys the following properties:

Proof. Let us choose $\mathcal{Z}\colon H(p) \to \mathcal{O}_E \setminus\{0\}$ that enjoys the properties described in Corollary 3.5. Let $\mathfrak{B}\in H(p)$. In order to define $\Pi(\mathfrak{B})$, notice that
$$ \begin{equation*} \mathfrak{B}\cdot \rho(\mathfrak{B})=p\mathcal{O}_K, \qquad \mathfrak{B}^c=z \mathcal{O}_K, \end{equation*} \notag $$
where
$$ \begin{equation} z=\mathcal{Z}(\mathfrak{B}) \in \mathcal{O}_K\setminus \{0\}. \end{equation} \tag{28} $$
Then $z\rho(z)$ is a generator of the ideal
$$ \begin{equation*} \mathfrak{B}^c\cdot \rho(\mathfrak{B}^c)=\left(\mathfrak{B}\cdot \rho(\mathfrak{B})\right)^c=p^c\cdot\mathcal{O}_K=q\mathcal{O}_K. \end{equation*} \notag $$
Since $\rho$ is the complex conjugation, $z\rho(z)$ is a real (i. e., $\rho$-invariant) totally positive element of $\mathcal{O}_K$. Clearly,
$$ \begin{equation*} u:=\frac{z\rho(z)}{q} \end{equation*} \notag $$
is an invertible element of $\mathcal{O}_K$ that is also $\rho$-invariant and totally positive unit in $\operatorname{U}_{K_0}$. Obviously,
$$ \begin{equation*} q=\frac{z\cdot \rho(z)}{u}. \end{equation*} \notag $$
Now let us put
$$ \begin{equation} \Pi(\mathfrak{B}):=q \cdot \frac{z}{\rho(z)}=\frac{z^2}{z\rho(z)/q}=\frac{z^2}{u} \in \mathcal{O}_K. \end{equation} \tag{29} $$
If $u$ is a square in $K_0$, then there is a unit $u_0$ in $K_0$ such that $u=u_0^2$. If this is the case, then let us put
$$ \begin{equation} \Pi_0(\mathfrak{B}):=\frac{z}{u_0}\in \mathcal{O}_K \quad \text{and get } \Pi_0(\mathfrak{B})^2=\biggl(\frac{z}{u_0}\biggr)^2=\frac{z^2}{u}=\Pi(\mathfrak{B}). \end{equation} \tag{30} $$
Clearly,
$$ \begin{equation} \Pi(\mathfrak{B})\cdot \mathcal{O}_K=z^2\cdot \mathcal{O}_K =(z\cdot\mathcal{O}_K)^2=(\mathfrak{B}^{c})^2=\mathfrak{B}^{2c}, \end{equation} \tag{31} $$
which proves (i). In order to check that $\Pi(\mathfrak{B})$ is a Weil’s $q^2$-number, notice that
$$ \begin{equation*} \Pi(\mathfrak{B})\cdot \rho(\Pi(\mathfrak{B}))=q \cdot \frac{z}{\rho(z)}\cdot \rho\biggl(q \cdot \frac{z}{\rho(z)}\biggr)=q^2\cdot \frac{z}{\rho(z)} \cdot \frac{\rho(z)}{z}=q^2. \end{equation*} \notag $$
In light of Remark 2.2, this proves that $\Pi(\mathfrak{B})$ is a Weil’s $q^2$-number. It follows from (30) that if $\Pi_0(\mathfrak{B})$ is defined, then it is a Weil’s $q$-number. By construction,
$$ \begin{equation*} \Pi(\mathfrak{B})\mathcal{O}_K=\mathfrak{B}^{2c}, \end{equation*} \notag $$
which also implies that $\Pi(\mathfrak{B})$ is $p^{2c}=q^2$-ordinary Weil’s number. The $G$-invariance of $\mathcal{Z}$ (see Corollary 3.5) combined with (28) and (29) implies the $G$-equivariance of $\Pi$, which proves (0). The injectiveness of $\Pi$ follows from (31). This proves (i) and (ii).

In order to prove (v), notice that if $\mathbb{Q}(\Pi(\mathfrak{B})^h)$ does not coincide with $K$, then it consists of $\rho$-invariants (§ 2.4). In particular, the ideal $\Pi(\mathfrak{B})^h\mathcal{O}_K=\mathfrak{B}^{2ch}$ coincides with its complex-conjugate

$$ \begin{equation*} \rho\bigl(\Pi(\mathfrak{B})^h\mathcal{O}_K\bigr)=\rho(\mathfrak{B}^{2ch}) =\rho(\mathfrak{B})^{2ch}. \end{equation*} \notag $$
This implies that $\mathfrak{B}=\rho(\mathfrak{B})$, which is not the case, since $\mathfrak{B}\in H(p)$. The obtained contradiction proves (v).

In order to prove (iii), we need to check that if $\pi'$ is an ordinary Weil’s $p^m$-number in $K$, then it is equivalent to $\Pi(\mathfrak{B})$ for some $\mathfrak{B}\in H(p)$. In order to do that, let us consider the ideal $\mathfrak{M}:=\pi'\mathcal{O}_K$ in $\mathcal{O}_K$. Since $\pi'\cdot\rho(\pi')=p^m$, we get $\mathfrak{M}\cdot \rho(\mathfrak{M})=p^m \mathcal{O}_K$. It follows that

$$ \begin{equation*} \mathfrak{M}=\prod_{\mathfrak{P}\in S(p)}\mathfrak{P}^{d(\mathfrak{P})}, \qquad d(\mathfrak{P})+d(\mathfrak{\rho(P)})=m \quad \forall \, \mathfrak{P}\in S(p). \end{equation*} \notag $$
The ordinarity of $\mathfrak{M}$ implies that
$$ \begin{equation*} d(\mathfrak{P})=0\text{ or } m \quad \forall \, \mathfrak{P}\in S(p). \end{equation*} \notag $$
This implies that if we put
$$ \begin{equation*} \Phi=\{\mathfrak{P}\in S(p)\mid d(\mathfrak{P})=m\} \subset S(p), \end{equation*} \notag $$
then $\Phi$ is a $p$-type and
$$ \begin{equation*} \mathfrak{M}=\prod_{\mathfrak{P}\in \Phi}\mathfrak{P}^m=\biggl(\prod_{\mathfrak{P}\in \Phi}\mathfrak{P}\biggr)^m. \end{equation*} \notag $$
It is also clear that
$$ \begin{equation*} \mathfrak{B}:=\prod_{\mathfrak{P}\in \Phi}\mathfrak{P}\in H(p), \end{equation*} \notag $$
and
$$ \begin{equation*} (\pi')^{2c}\mathcal{O}_K=\mathfrak{M}^{2c}=\mathfrak{B}^{2cm}=(\mathfrak{B}^{2c})^m =\bigl(\Pi(\mathfrak{B})\mathcal{O}_K\bigr)^m=\Pi(\mathfrak{B}^m)\mathcal{O}_K. \end{equation*} \notag $$
It follows from Remark 2.3 that the ratio $\Pi(\mathfrak{B})^m /(\pi')^{2c}$ is a root of unity. The uniqueness follows from the already proved (i).

Let us prove (iv). The already proved (0) tells us that if $\mathfrak{B}_2 =\sigma(\mathfrak{B}_1)$ for $\sigma \in G$, then $\Pi(\mathfrak{B}_2)=\sigma(\Pi(\mathfrak{B}_1))$ and therefore Weil’s numbers $\Pi(\mathfrak{B}_1)$ and $\Pi(\mathfrak{B}_2)$ are equivalent.

Conversely, suppose that $\Pi(\mathfrak{B}_1)$ and $\Pi(\mathfrak{B}_2)$ are equivalent. This means that there are positive integers $a,b$, a Galois automorphism $\sigma \in G$, and a root of unity $\zeta \in \mu_K$ such that

$$ \begin{equation*} \Pi(\mathfrak{B}_2)^a=\zeta \cdot \sigma(\Pi(\mathfrak{B}_1))^b. \end{equation*} \notag $$
This implies the equality of the corresponding ideals in $\mathcal{O}_K$:
$$ \begin{equation*} \Pi(\mathfrak{B}_2)^a\mathcal{O}_K=\sigma(\Pi(\mathfrak{B}_1))^b\mathcal{O}_K =\Pi(\sigma(\mathfrak{B}_1))^b. \end{equation*} \notag $$
This means (in light of already proved (i)) that
$$ \begin{equation*} \mathfrak{B}_2^{2 ca}=(\sigma(\mathfrak{B}_1))^{2cb}, \end{equation*} \notag $$
which implies $\mathfrak{B}_2=\sigma(\mathfrak{B}_1)$. Hence $\mathfrak{B}_1$ and $\mathfrak{B}_2$ lie in the same $G$-orbit.

Let us prove (vi). Actually, we have already constructed the map $\Pi_0\colon H(p) \to \mathcal{O}_K$, checked that its image lies in $W(q,K)$; we have also proved property (vi-a). As for (vi-b), it follows readily from (30) combined with the $G$-equivariance of $\Pi$. As for (vi-c), it follows readily from (v) combined with (30). In order to prove (vi-d), it suffices to recall that $\Pi(\mathfrak{B})$ is an ordinary Weil’s $q^2$-number and notice that in light of (30), the integer

$$ \begin{equation*} \frac{\operatorname{ord}_{\mathfrak{P}}(\Pi(\mathfrak{B}))}{\operatorname{ord}_{\mathfrak{P}}(q^2)} =\frac{2\operatorname{ord}_{\mathfrak{P}}(\Pi_0(\mathfrak{B}))} {2\operatorname{ord}_{\mathfrak{P}}(q)} =\frac{\operatorname{ord}_{\mathfrak{P}}(\Pi_0(\mathfrak{B}))} {\operatorname{ord}_{\mathfrak{P}}(q)}. \end{equation*} \notag $$
Theorem 3.6 is proved. $\Box$

§ 4. Abelian varieties with Weil’s numbers in $K$

As above, $p$ is a prime, $m$ a positive integer, and $q=p^m$.

Theorem 4.1. Let $A$ be a simple abelian variety over $k=\mathbb{F}_q$ such that the corresponding Weil’s $q$-number

$$ \begin{equation*} \pi_A \in K. \end{equation*} \notag $$
Let $\mathbb{Q}(\pi_A)$ be the subfield of $K$ generated by $\pi_A$.

(i) Suppose that either $\mathbb{Q}(\pi_A)\ne K$ or $p$ does not split completely in $K$.

Then $A$ is supersingular.

(ii) If $p$ splits completely in $K$, $\mathbb{Q}(\pi_A)=K$, and $\pi_A$ is not ordinary, then the division $\mathbb{Q}$-algebra $\operatorname{End}_k^{0}(A)$ is not commutative.

(iii) If $\pi_A$ is ordinary, then $K=\mathbb{Q}(\pi_A)$ and $\operatorname{End}_k^{0}(A) \cong K$; in particular, $\operatorname{End}_k^{0}(A)$ is commutative.

Proof. (i) It follows from Lemmas 2.7 and 2.8 that $\pi_A^2/q$ is a root of unity. This means that $A$ is supersingular.

(ii), (iii) Recall (see [11], [13]) that $E:=\operatorname{End}_k^{0}(A)$ is a central division algebra over the field $\mathbb{Q}(\pi_A)=K$. Since $p$ splits completely in $K$, the $\mathfrak{P}$-adic completion $K_{\mathfrak{P}}$ of $K$ coincides with $\mathbb{Q}_p$, i. e.,

$$ \begin{equation*} [K_{\mathfrak{P}}:\mathbb{Q}_p]=1 \quad \forall \, \mathfrak{P} \in S(p). \end{equation*} \notag $$
By [13], Theorem 1, the local $\mathfrak{P}$-adic invariant
$$ \begin{equation*} \operatorname{inv}_{\mathfrak{P}}(E)\in \mathbb{Q}/\mathbb{Z} \end{equation*} \notag $$
of the central division $K$-algebra $E$ is given by the formula
$$ \begin{equation} \operatorname{inv}_{\mathfrak{P}}(E)= \frac{\operatorname{ord}_{\mathfrak{P}}(\pi_A)}{\operatorname{ord}_{\mathfrak{P}}(q)} [K_{\mathfrak{P}}:\mathbb{Q}_p]\bmod \mathbb{Z}= \frac{\operatorname{ord}_{\mathfrak{P}}(\pi_A)}{\operatorname{ord}_{\mathfrak{P}}(q)}\bmod \mathbb{Z}\in \mathbb{Q}/\mathbb{Z}. \end{equation} \tag{32} $$
All other local invariants of $E$ (outside $S(p)$) are $0$ (ibid).

Suppose that $\pi_A$ is ordinary. Then $\mathbb{Q}(\pi_A)=K$, because otherwise $\mathbb{Q}(\pi_A)\subset \mathbb{R}$ and therefore $\pi_A$ is real, i. e., $A$ is supersingular [13], Examples, which is not the case. Since $\pi_A$ is ordinary, all the slopes $\operatorname{ord}_{\mathfrak{P}}(\pi_A)/\operatorname{ord}_{\mathfrak{P}}(q)$ are integers and therefore $\operatorname{inv}_{\mathfrak{P}}(E)=0$ for all $\mathfrak{P}\in S(p)$. This implies that the division algebra $E=\operatorname{End}_k^{0}(A)$ is actually a field, i. e., is isomorphic to $K$. This proves (iii).

In order to prove (ii), assume that $\pi_A$ is not ordinary. Then there is a maximal ideal $\mathfrak{P} \in S(p)$ such that the ratio $\operatorname{ord}_{\mathfrak{P}}(\pi_A)/\operatorname{ord}_{\mathfrak{P}}(q)$ is not an integer, i. e.,

$$ \begin{equation} \frac{\operatorname{ord}_{\mathfrak{P}}(\pi_A)}{\operatorname{ord}_{\mathfrak{P}}(q)}\bmod \mathbb{Z} \ne 0 \quad \text{in } \mathbb{Q}/\mathbb{Z}. \end{equation} \tag{33} $$
Combining (33) with (32), we obtain that $\operatorname{inv}_{\mathfrak{P}}(E) \ne 0.$ It follows that $E=\operatorname{End}_k^{0}(A)$ does not coincide with its center, i. e., is noncommutative. This proves (ii). $\Box$

Remark 4.2. Let $A$ be a simple abelian variety over $\mathbb{F}_q$ such that $\pi_A \in K$. Obviously, $A$ is ordinary if and only if $\pi_A$ is ordinary.

§ 5. Honda–Tate theory for ordinary abelian varieties

As above, $p$ is a prime that splits completely in $K$, $m$ a positive integer, and $q=p^m$.

Let $\pi \in K$ be a Weil’s $q$-number. The Honda–Tate theory (see [11]–[13]) attaches to $\pi$ a simple abelian variety $\mathcal{A}$ over $\mathbb{F}_{q}$ that is defined up to an $\mathbb{F}_{q}$-isogeny and enjoys the following properties.

Let $\mathrm{Fr}_{\mathcal{A}}\colon \mathcal{A}\to \mathcal{A}$ be the Frobenius endomorphism of $\mathcal{A}$ and $F:=\mathbb{Q}[\mathrm{Fr}_{\mathcal{A}}]$ be the $\mathbb{Q}$-subalgebra of the division $\mathbb{Q}$-algebra $E:=\operatorname{End}_{\mathbb{F}_{q}}^{0}(\mathcal{A})$ (which is actually a subfield). Then $F$ is the center of $E$ and there is a field embedding

$$ \begin{equation*} i\colon F \hookrightarrow \mathbb{C} \quad \text{such that} \quad i(\mathrm{Fr}_{\mathcal{A}})=\pi. \end{equation*} \notag $$

Lemma 5.1. Suppose $\pi$ is ordinary and $\mathbb{Q}(\pi^h)=K$ for all positive integers $h$. Then $\mathcal{A}$ is an absolutely simple $2^{n-1}$-dimensional ordinary abelian variety, $\operatorname{End}^0(A) \cong K$, and all endomorphisms of $\mathcal{A}$ are defined over $\mathbb{F}_{q}$.

Proof. Since $\mathbb{Q}(\pi)=K$, we get $i(F)=K$. In particular, number fields $K$ and $F$ are isomorphic. In light of Theorem 4.1, $\mathcal{A}$ is an ordinary abelian variety with commutative endomorphism algebra $E=F\cong K$. By Theorem 2 (c) of [11], § 3,
$$ \begin{equation*} \dim(\mathcal{A})=\frac{[E:\mathbb{Q}]}{2}=\frac{[K:\mathbb{Q}]}{2}=2^{n-1}. \end{equation*} \notag $$
We are going to prove that $\mathcal{A}$ is absolutely simple and all its endomorphisms are defined over $\mathbb{F}_{q}$. Let $h$ be a positive integer and $k =\mathbb{F}_{q^{h}}$ a degree $h$ field extension of $\mathbb{F}_{q}$. Let $\mathcal{A}_k=\mathcal{A}\times_{\mathbb{F}_{q}}k$ be the abelian variety over $k$ obtained from $\mathcal{A}$ by the extension of scalars. There is the natural embedding (inclusion) of $\mathbb{Q}$-algebras
$$ \begin{equation*} \operatorname{End}_{\mathbb{F}_{q}}^{0}(\mathcal{A})\subset \operatorname{End}_k^0(\mathcal{A}_k) \end{equation*} \notag $$
such that the Frobenius endomorphism $\mathrm{Fr}_{\mathcal{A}_k}$ coincides with $\mathrm{Fr}_{\mathcal{A}}^h$. In particular,
$$ \begin{equation*} \mathbb{Q}[\mathrm{Fr}_{\mathcal{A}_k}]\subset \mathbb{Q}[\mathrm{Fr}_{\mathcal{A}}]=F. \end{equation*} \notag $$
In addition,
$$ \begin{equation*} i(\mathrm{Fr}_{\mathcal{A}_k})=i(\mathrm{Fr}_{\mathcal{A}}^h)= i(\mathrm{Fr}_{\mathcal{A}})^h=\pi^h. \end{equation*} \notag $$
Since $\mathbb{Q}[\pi^h]=K=\mathbb{Q}(\pi)$, we get
$$ \begin{equation*} i(\mathbb{Q}[\mathrm{Fr}_{\mathcal{A}_k}])=K=i(\mathbb{Q}[\mathrm{Fr}_{\mathcal{A}}]). \end{equation*} \notag $$
Hence, $\mathbb{Q}[\mathrm{Fr}_{\mathcal{A}_k}]=\mathbb{Q}[\mathrm{Fr}_{\mathcal{A}}]$ is a number field of degree $2\dim(\mathcal{A})=2\dim(\mathcal{A}_k)$. Applying again Theorem 2 (c) of [11], § 3, to $\mathcal{A}_k$, we conclude that
$$ \begin{equation*} \operatorname{End}^0(\mathcal{A}_k)=\mathbb{Q}[\mathrm{Fr}_{\mathcal{A}_k}] =\mathbb{Q}[\mathrm{Fr}_{\mathcal{A}}]=\operatorname{End}_{\mathbb{F}_{q}}^{0}(\mathcal{A}) \end{equation*} \notag $$
for all finite overfields $k$ of $\mathbb{F}_{q}$. This implies that
$$ \begin{equation*} \operatorname{End}^0(\mathcal{A}_k)=\operatorname{End}_{\mathbb{F}_{q}}^{0}(\mathcal{A}), \end{equation*} \notag $$
i. e., all the endomorphisms of $\mathcal{A}$ are defined over $\mathbb{F}_{q}$. In particular, $\mathcal{A}$ is absolutely simple and $\operatorname{End}^0(\mathcal{A}) \cong K$. $\Box$

§ 6. Proofs of main results

As above, $c=\exp(K)$, a prime $p$ splits completely in $K$ and $q=p^c$.

Proof of Theorem 1.6. Let $\Pi\colon H(p) \to W(q^2,K)$ be as in Theorem 3.6. Let $\mathcal{B}\in H(p)$ and let $\Pi(\mathcal{B})$ be the corresponding ordinary Weil’s $q^2$-number in $K$. In light of Theorem 3.6 (v), $\mathbb{Q}[\Pi(\mathfrak{B})^h]=K$ for all positive integers $h$. In light of Lemma 5.1 applied to $q^2$ and $\Pi(\mathfrak{B})$, the Honda–Tate theory [11]–[13] attaches to $\Pi(\mathcal{B})$ an absolutely simple $2^{n-1}$-dimensional abelian variety $\mathcal{A}=A(\mathcal{B})$ over $\mathbb{F}_{q^2}$ (that is defined up to an $\mathbb{F}_{q^2}$-isogeny) such that $\operatorname{End}^0(A(\mathcal{B}))\cong K$, and all endomorphisms of $A(\mathcal{B})$ are defined over $\mathbb{F}_{q^2}$.

By Theorem 3.6 (iv), if $\mathfrak{B}_1, \mathfrak{B}_2 \in H(p)$, then the Weil numbers $\Pi(\mathfrak{B}_1)$ and $\Pi(\mathfrak{B}_2)$ are equivalent if and only if $\mathfrak{B}_1$, and $\mathfrak{B}_2$ belong to the same $G$-orbit. In light of [11], Theorem 1, [12], p. 84, combined with Lemma 3.4, all the $A(\mathfrak{B})$ lie in precisely $2^{2^{n-1}-n}$ isogeny classes of aelian varieties over $\overline{\mathbb{F}}_p$. We also know that each of these varieties is ordinary, has dimension $2^{n-1}$ and their endomorphism algebras are isomorphic to $K$.

Now, let us prove that each abelian variety $\mathcal{B}$ over $\overline{\mathbb{F}}_p$, whose endomorphism algebra is isomorphic to $K$, is isogenous to one of $A(\mathfrak{B})$ over $\overline{\mathbb{F}}_p$.

In order to do that, first, notice that since $K$ is a field, $\mathcal{B}$ is simple over $\overline{\mathbb{F}}_p$. Second, $\mathcal{B}$ is defined with all its endomorphisms over a certain finite field $k=\mathbb{F}_{q^{2h}}$ (where $h$ is a certain positive integer), i. e., there is a simple abelian variety $\mathcal{B}_k$ over $k$ such that

$$ \begin{equation*} \mathcal{B}=\mathcal{B}_k\times_k \overline{\mathbb{F}}_p, \qquad \operatorname{End}_k^{0}(\mathcal{B}_k)=\operatorname{End}^0(\mathcal{B}) \cong K. \end{equation*} \notag $$
Applying Theorem 2 (c) of [11], § 3, to $\mathcal{B}_k$, we get
$$ \begin{equation*} K \cong \operatorname{End}^0(\mathcal{B})= \operatorname{End}_k^0(\mathcal{B}_k) =\mathbb{Q}[\mathrm{Fr}_{\mathcal{B}_k}], \end{equation*} \notag $$
where $\mathrm{Fr}_{\mathcal{B}_k}$ is the Frobenius endomorphism of $\mathcal{B}_k$. This gives us a field isomorphism $\mathbb{Q}[\mathrm{Fr}_{\mathcal{B}_k}] \to K$; let us denote by $\pi_{\mathcal{B}_k}$ the image of $\mathrm{Fr}_{\mathcal{B}_k}$ in $K$. Clearly, $\mathbb{Q}(\pi_{\mathcal{B}_k})\,{=}\,K$; according to a classical result of Weil [16], $\pi_{\mathcal{B}_k}$ is a Weil’s $q^{2h}$-number. By Theorem 4.1 (i) (applied to $q^{2h}$ instead of $q$), $\pi_{\mathcal{B}_k}$ is ordinary, since $\operatorname{End}_k^0(\mathcal{B}_k) \cong K$ is commutative. It follows from Theorem 3.6 (iii) that there is $\mathfrak{B} \in H(p)$ such that Weil’s numbers $\pi_{\mathcal{B}_k}$ and $\Pi(\mathfrak{B})$ are equivalent. This means (thanks to Theorem 1 of [11], see also [12], pp. 83, 84) that absolutely simple abelian varieties $\mathcal{B}_k$ and $A(\mathfrak{B})$ become isogenous over $\overline{\mathbb{F}}_p$. It follows that absolutely simple abelian varieties $\mathcal{B}=\mathcal{B}_k\times_k \overline{\mathbb{F}}_p$ and $A(\mathfrak{B})$ are isogenous over $\overline{\mathbb{F}}_p$.

This proves (i), (ii) (1), and (ii) (2). It remains to prove (ii) (3). It suffices to check that for each $\mathcal{B}\in H(p)$ there exists an abelian variety $A_0$ that is defined over $\mathbb{F}_q$ with all its endomorphisms and such that $A(\mathcal{B})$ is isogenous to $A_0$ over $\overline{\mathbb{F}}_p$.

Let $\Pi_0\colon H(p) \to W(q,K)$ be as in Theorem 3.6 (vi) and $\Pi_0(\mathcal{B})$ be the corresponding ordinary Weil’s $q$-number in $K$. In light of Theorem 3.6 (vi-c), $\mathbb{Q}[\Pi_0(\mathfrak{B})^h]=K$ for all positive integers $h$. In light of Lemma 5.1 applied to $q$ and $\Pi_0(\mathfrak{B})$, the Honda–Tate theory [11]–[13] attaches to Weil’s $q$-number $\Pi_0(\mathcal{B})$ an absolutely simple $2^{n-1}$-dimensional abelian variety $\mathcal{A}_0$ over $\mathbb{F}_{q}$ (that is defined up to an $\mathbb{F}_{q}$-isogeny) such that $\operatorname{End}^0(\mathcal{A}_0)\cong K$, and all endomorphisms of $\mathcal{A}_0$ are defined over $\mathbb{F}_{q}$.

Since $\Pi_0(\mathcal{B})^2=\Pi(\mathcal{B})$, Weil’s numbers $\Pi_0(\mathcal{B})$ and $\Pi(\mathcal{B})$ are equivalent. As above, in light of Theorem 1 of [11] (see also [12], pp. 83, 84), the corresponding absolutely simple abelian varieties $\mathcal{A}_0$ and $A(\mathcal{B})$ are isogenous over $\overline{\mathbb{F}}_p$. $\Box$

Proof of Corollary 1.14. Recall that $r$ is an odd prime and $\zeta_r$ is a primitive $r$th root of unity. Clearly, $\mathbb{Q}(\zeta_r)$ is a $\mathrm{CM}$ field. Hence, its subfield $K$ is either $\mathrm{CM}$ or totally real. Since $\mathbf{H}$ has odd order $m$, it does not contain the complex conjugation $\rho\colon \mathbb{Q}(\zeta_r) \to \mathbb{Q}(\zeta_r)$, because $\rho$ has order $2$. Hence, $\rho$ acts nontrivially on $K=\mathbb{Q}(\zeta_r)^{\mathbf{H}}=K^{(r)}$, which implies that $K$ is a $\mathrm{CM}$ field. (See also [10], p. 78.) Its degree
$$ \begin{equation*} [K:\mathbb{Q}]=\frac{[\mathbb{Q}(\zeta_r):\mathbb{Q}]}{\#(\mathbf{H})}=\frac{m \cdot 2^n}{m}=2^n. \end{equation*} \notag $$
We also know (Remark 1.15) that every totally positive unit in $K_0$ is a square in $K_0$.

Clearly, $K/\mathbb{Q}$ is ramified at $r$ and unramified at every prime $p \ne r$. Let us find which $p\ne r$ split completely in $K$. Let

$$ \begin{equation*} f_p \in \operatorname{Gal}(\mathbb{Q}(\zeta_r)/\mathbb{Q})=(\mathbb{Z}/r\mathbb{Z})^* \end{equation*} \notag $$
be the Frobenius element attached to $p$, which is characterized by the property
$$ \begin{equation*} f_p(\zeta_r)=\zeta_r^p. \end{equation*} \notag $$
In other words,
$$ \begin{equation*} f_p=(p\ \operatorname{mod} r)\in (\mathbb{Z}/r\mathbb{Z})^* =\operatorname{Gal}(\mathbb{Q}(\zeta_r)/\mathbb{Q}). \end{equation*} \notag $$
Clearly, $p$ splits completely in $K$ if and only if $f_p\in \mathbf{H}$. So, we need to find when $f_p$ lies in $\mathbf{H}$. In order to do it, notice that
$$ \begin{equation*} \mathbf{H}=\{\sigma^{2^n}\mid \sigma \in \operatorname{Gal}(\mathbb{Q}(\zeta_r)/\mathbb{Q})= (\mathbb{Z}/r\mathbb{Z})^*\}. \end{equation*} \notag $$
This implies that $f_p$ lies in $\mathbf{H}$ if and only if $p \bmod r$ is a $2^n$th power in $\mathbb{Z}/r\mathbb{Z}=\mathbb{F}_r$ completing the proof of (0).

Assertions (i) and (ii) follow from Theorem 1.6 combined with (0). $\Box$

Proof of Corollary 1.8. In the notation of Corollary 1.14, this is the case when $m=1$ and $2^n=r-1$. By little Fermat’s theorem, every nonzero $a\in \mathbb{Z}/r\mathbb{Z}$ satisfies
$$ \begin{equation*} a^{2^n}=a^{r-1}=1. \end{equation*} \notag $$
Now the desired result follows readily from Corollary 1.14. $\Box$

Acknowledgements

I am grateful to Ley Watson for an interesting stimulating question [8].


Bibliography

1. Z. I. Borevich and I. R. Shafarevich, Number theory, 3rd ed., Nauka, Moscow, 1985  mathscinet  zmath; English transl. of 1st ed. Pure Appl. Math., 20, Academic Press, New York–London, 1966  mathscinet  zmath
2. F. Oort, “The isogeny class of a $\mathrm{CM}$-type abelian variety is defined over a finite extension of the prime field”, J. Pure Appl. Algebra, 3:4 (1973), 399–408  crossref  mathscinet  zmath
3. M. Deuring, “Die Typen der Multiplikatorenringe elliptischer Funktionenkörper”, Abh. Math. Sem. Hansischen Univ., 14 (1941), 197–272  crossref  mathscinet  zmath
4. W. C. Waterhouse, “Abelian varieties over finite fields”, Ann. Sci. École Norm. Sup. (4), 2:4 (1969), 521–560  crossref  mathscinet  zmath
5. M. Deuring, “Die Zetafunktion einer algebraischen Kurve vom Geschlechte Eins. III”, Nachr. Akad. Wiss. Göttingen. Math.-Phys. Kl. IIa, 1956 (1956), 37–76  mathscinet  zmath
6. J.-P. Serre and J. Tate, “Good reduction of abelian varieties”, Ann. of Math. (2), 88:3 (1968), 492–517  crossref  mathscinet  zmath
7. L. C. Washington, Introduction to cyclotomic fields, Grad. Texts in Math., 83, 2nd ed., Springer-Verlag, New York, 1997  crossref  mathscinet  zmath
8. L. Watson, Isomorphic endomorphism algebras implies isogenous (for abelian varieties over finite fields)? https://mathoverflow.net/questions/393506/isomorphic-endomorphism-algebras-implies-isogenous-for-abelian-varieties-over-f/
9. H. Hasse, Über die Klassenzahl abelscher Zahlkörper, Math. Lehrbücher und Monogr., 1, Akademie-Verlag, Berlin, 1952  crossref  mathscinet  zmath; reprint Springer-Verlag, Berlin, 1985  mathscinet  zmath; English transl. H. Hasse, On the class number of Abelian number fields, Extended with tables by Ken-ichi Yoshino and Mikihito Hirabayashi. Translated from the 1985 German reprint and with a preface by Hirabayashi. With a foreword by Franz Lemmermeyer, Springer, Cham, 2019  mathscinet
10. P. E. Conner and J. Hurrelbrink, Class number parity, Ser. Pure Math., 8, World Sci. Publ., Singapore, 1988  crossref  mathscinet  zmath
11. J. Tate, “Endomorphisms of abelian varieties over finite fields”, Invent. Math., 2 (1966), 134–144  crossref  mathscinet  zmath  adsnasa
12. T. Honda, “Isogeny classes of abelian varieties over finite fields”, J. Math. Soc. Japan, 20:1-2 (1968), 83–95  crossref  mathscinet  zmath
13. J. Tate, “Classes d'isogénie des variétés abéliennes sur un corps fini”, Matematika, 14:6 (1970), 129–137 (Russian)  mathnet  zmath; Transl. from the French J. Tate, “Classes d'isogénie des variétés abéliennes sur un corps fini (d'aprés T. Honda)”, Séminaire Bourbaki, Lecture Notes in Math., 179, Springer, Berlin, 1968, Exp. No. 352, 95–110  crossref  mathscinet  zmath
14. A. Weil, Basic number theory, Grundlehren Math. Wiss., 144, 3rd ed., Springer-Verlag, New York–Berlin, 1974  mathscinet  zmath; Russian transl. Mir, Moscow, 1972  mathscinet  zmath
15. G. Shimura, Abelian varieties with complex multiplication and modular functions, Princeton Math. Ser., 46, Princeton Univ. Press, Princeton, NJ, 1998  crossref  mathscinet  zmath
16. D. Mumford, Abelian varieties, Tata Inst. Fundam. Res. Stud. Math., 5, 2nd ed., Oxford Univ. Press, London, 1974  mathscinet  zmath; Russian transl. of 1st ed. Mir, Moscow, 1971  zmath

Citation: Yu. G. Zarhin, “Isogeny classes and endomorphism algebras of abelian varieties over finite fields”, Izv. RAN. Ser. Mat., 87:3 (2023), 56–74; Izv. Math., 87:3 (2023), 469–487
Citation in format AMSBIB
\Bibitem{Zar23}
\by Yu.~G.~Zarhin
\paper Isogeny classes and endomorphism algebras of abelian varieties over finite fields
\jour Izv. RAN. Ser. Mat.
\yr 2023
\vol 87
\issue 3
\pages 56--74
\mathnet{http://mi.mathnet.ru/im9332}
\crossref{https://doi.org/10.4213/im9332}
\mathscinet{http://mathscinet.ams.org/mathscinet-getitem?mr=4640915}
\adsnasa{https://adsabs.harvard.edu/cgi-bin/bib_query?2023IzMat..87..469Z}
\transl
\jour Izv. Math.
\yr 2023
\vol 87
\issue 3
\pages 469--487
\crossref{https://doi.org/10.4213/im9332e}
\isi{https://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcApp=Publons&SrcAuth=Publons_CEL&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=001063937600003}
\scopus{https://www.scopus.com/record/display.url?origin=inward&eid=2-s2.0-85171845368}
Linking options:
  • https://www.mathnet.ru/eng/im9332
  • https://doi.org/10.4213/im9332e
  • https://www.mathnet.ru/eng/im/v87/i3/p56
  • Citing articles in Google Scholar: Russian citations, English citations
    Related articles in Google Scholar: Russian articles, English articles
    Известия Российской академии наук. Серия математическая Izvestiya: Mathematics
    Statistics & downloads:
    Abstract page:236
    Russian version PDF:12
    English version PDF:47
    Russian version HTML:79
    English version HTML:103
    References:18
    First page:6
     
      Contact us:
     Terms of Use  Registration to the website  Logotypes © Steklov Mathematical Institute RAS, 2024