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

\AtBeginEnvironment

algorithmic

Efficient computations in central simple algebras using Amitsur cohomology

Péter Kutas Faculty of Informatics, Eötvös Loránd University and School of Computer Science, University of Birmingham (https://sites.google.com/view/peterkutas89). The first author is supported by the Hungarian Ministry of Innovation and Technology NRDI Office within the framework of the Quantum Information National Laboratory Program, the János Bolyai Research Scholarship of the Hungarian Academy of Sciences and by the UNKP-22-5 New National Excellence Program. The first author is also partly supported by EPSRC through grant number EP/V011324/1.    Mickaël Montessinos Institute of Mathematics, Faculty of Mathematics and Informatics, Vilnius University (http://mickael.montessinos.fr).
Abstract

We introduce a presentation for central simple algebras over a field kk using Amitsur cohomology. We provide efficient algorithms for computing a cocycle corresponding to any such algebra given by structure constants. If kk is a number field, we use this presentation to prove that the explicit isomorphism problem (i.e., finding an isomorphism between central simple algebras given by structure constants) reduces to SS-unit group computation and other related number theoretical computational problems. This also yields, conditionally on the generalised Riemann hypothesis, the first polynomial quantum algorithm for the explicit isomorphism problem over number fields.

1 Introduction

The explicit isomorphism problem is the algorithmic problem of, given a field kk and a kk-algebra AA isomorphic to Md(k)M_{d}(k), constructing an explicit isomorphism φ:AMd(k)\varphi\colon A\to M_{d}(k). The explicit isomorphism problem may be thought of as a natural problem in computational representation theory. Given a kk-algebra AA, one may wish to assay it. That is, compute the Jacobson radical of AA, and the decomposition of the semi-simple quotient of AA as a sum of simple kk-algebras, themselves identified to some Md(D)M_{d}(D), for DD a division kk-algebra. In general, the hard part of this task is to find an isomorphism AMn(D)A\to M_{n}(D) when AA is simple. A general recipe for solving this problem is to identify the Brauer class of DD over its centre K/FK/F, find structure constants for Mn(Dop)M_{n}(D^{op}) and then compute an explicit isomorphism AMn(Dop)Mm(K)A\otimes M_{n}(D^{op})\simeq M_{m}(K) [37, 32, 20].

Applications of the explicit isomorphism problem go beyond the mere computational theory of finite-dimensional associative algebras. In arithmetic geometry, the problem is relevant for trivialising obstruction algebras in explicit descent over elliptic curves [16, 17, 18, 26] and computation of Cassel-Tate pairings [28, 49]. The problem is also connected to the parametrisation of Severi-Brauer surfaces [21]. Recent work in algebraic complexity theory reduced the determinant equivalence test to the explicit isomorphism problem [30]. Finally, the explicit isomorphism problem over the function field 𝔽q(T)\mathbb{F}_{q}(T) is also relevant to error correcting codes [32].

It is well known that the Brauer group of equivalence classes of central simple kk-algebras may be described as the second Galois cohomology group H2(k,ksep×)H^{2}(k,k_{sep}^{\times}). The remarkable fact that these two groups are isomorphic may be seen as a consequence of the crossed-product presentation of central simple algebras containing a maximal commutative subalgebra which is a Galois extension of the base field.

When a cohomological presentation is known for a certain algebra, the explicit isomorphism problem translates into a multiplicative algebraic equation. Indeed, the algebra is represented by a cocycle, and since the algebra is split, the cocycle is a coboundary. Then, an explicit isomorphism to a matrix algebra may be computed using a cochain whose differential is this coboundary. When the base field is a number field and a presentation of the algebra (either a cyclic or a crossed product presentation) is known, the resulting equation may be solved by computing a certain group of SS-units [46, 25].

Before Noether introduced crossed-product algebras, Brauer had already introduced a factor set presentation which allowed to describe an algebra with the knowledge of any maximal commutative subalgebra, not necessarily a Galois extension of the base field. We refer the reader to [38] for a modern description of presentations introduced in the works of Brauer and Noether. It is interesting to note that Adamson developed a cohomology theory for non-Galois field extensions which describes Brauer factor sets as cocycles [1], much like classical Galois cohomology describes Noether factor sets as cocycles.

A line of work introduced Azumaya algebras, which generalise the notion of central simple algebra (and therefore of the Brauer group) over arbitrary rings [4, 3], and then later over schemes (see [15] for a modern treatment). While trying to connect the usual Galois cohomological description of the Brauer group and another description introduced by Hochschild, Amitsur came upon an exact complex which yields a cohomology theory suitable to classify central simple algebras, the so-called Amitsur cohomology. As the complex he defined may be constructed over arbitrary rings, a line of work generalised his classification result to Azumaya algebras over increasingly general classes of ring extensions [2, 45, 12].

Cohomological presentations have the potential of being used both for a computational representation of a central simple algebra and for solving the explicit isomorphism problem. For instance, in the computer algebra system PARI/GP [47], central simple algebras over a number field may be represented as cyclic algebras. While the already existing constructions relying on Amitsur cohomology are abstract and do not lend themselves easily to a practical implementation, the crossed product and cyclic presentations have the inconvenient that one is required to know a Galois, or even cyclic, maximal subfield in order to convert a given algebra into these presentations. While such a subfield always exists for algebras over global fields (as per the Brauer-Hasse-Noether theorem), there is no known efficient algorithm to find one (except for quaternion and degree 3 central simple algebras) to the best of our knowledge. The presentation using Brauer factor sets may be constructed using an arbitrary maximal commutative subalgebra, which is easy to find, but factor sets take value in a normal splitting field of the subalgebra used. Results in arithmetic statistics [23] suggest that a random element of such an algebra will generate a maximal subfield whose Galois group is the full symmetric group, and hence will have a Galois closure of degree n!n! (where nn is the degree of the algebra), preventing it to be used efficiently in a computation.

1.1 Our contributions

In Section 3, we give a presentation for central simple algebras over a field using Amitsur cocycles. Our presentation only requires knowledge of a separable maximal commutative subalgebra, which may be found efficiently in deterministic polynomial time [22]. Furthermore, multiplication in our representation may be computed using O(n3)O(n^{3}) base field multiplications, making it as efficient as the naive representation using structure constants. In Section 3.1 we construct a central simple algebra from an Amitsur 22-cocycle. In Section 3.2, we discuss the computational representation of Amitsur cocycles and related algebras. Finally, in Section 3.3 we give an efficient algorithm for computing a cocycle representing a given central simple algebra.

In Section 4, we prove that over a number field, an Amitsur 11-cochain whose differential is a given 22-coboundary may be found within a certain group of SS-units. We then prove that, assuming the generalised Riemann hypothesis (henceforth denoted by GRH), the explicit isomorphism problem is solved by a polynomial quantum algorithm.

1.2 Related works

Below, we review existing classical algorithms for solving the explicit isomorphism problem over various fields.

In the case of a finite base field, a polynomial-time algorithm was introduced by Ronyai in [44].

Instances of the problem for \mathbb{Q}-algebras isomorphic to Mn()M_{n}(\mathbb{Q}) were first treated separately for small values of nn. When n=2n=2, the problem reduces to finding a rational point on a projective conic [48, Theorem 5.5.4], which is solved for instance in [19]. Then, [21] presented a subexponential algorithm when n=3n=3 by finding a cyclic presentation and solving a cubic norm equation. The case n=4n=4 is tackled in [43] by reducing to the case of quaternion algebras over QQ and (d)\mathbb{Q}(\sqrt{d}) and solving a norm equation.

In [18] an algorithm was given and studied mostly for the cases n=3n=3 and n=5n=5. It was then generalised in [37, 36] to a KK-algebra isomorphic to Mn(K)M_{n}(K), where nn is a natural number and KK is a number field. The complexity of this last algorithm is polynomial in the size of the structure constants of the input algebra, but depends exponentially on nn, the degree of KK and the size of the discriminant of KK.

In 2018, [34] exhibited a polynomial-time algorithm for algebras isomorphic to Mn(𝔽q(T))M_{n}(\mathbb{F}_{q}(T)).

For the case of fixed nn and varying base field, [27, 39] independently gave an algorithm for an algebra isomorphic to M2((d))M_{2}(\mathbb{Q}(\sqrt{d})) with complexity polynomial in log(d)\log(d). A similar algorithm for quadratic extensions of 𝔽q(T)\mathbb{F}_{q}(T) was given in [35] for the case of odd qq. The case of even qq was then treated in [40].

2 Preliminaries

2.1 Algorithmic preliminaries

2.1.1 Computational model

The theoretical results of Section 3 are valid for any base field kk. In order to translate them into efficient algorithms, it is necessary to have an encoding of the elements of kk, for which the usual field operations and linear algebraic tasks may be performed efficiently. Following [22], we say that such a field admits efficient linear algebra. It is well known that finite fields and global fields, among others, admit efficient linear algebra.

2.1.2 Computational representations of algebras

A kk-algebra may be given as input to a program in the form of a table of structure constants. Let AA be a finite dimensional kk-algebra of dimension nn, let VV be the underlying vector space of AA and let (e1,,en)(e_{1},\ldots,e_{n}) be a basis of VV. Multiplication in AA is a bilinear map which may be represented as a tensor λVVV\lambda\in V^{\vee}\otimes V^{\vee}\otimes V. The table of structure constants of AA for the basis (e1,,en)(e_{1},\ldots,e_{n}) is then the coordinates of λ\lambda in the basis (eiejek)\left(e_{i}^{\vee}\otimes e_{j}^{\vee}\otimes e_{k}\right) of VVVV^{\vee}\otimes V^{\vee}\otimes V, where (ei)\left(e_{i}^{\vee}\right) is the basis of VV^{\vee} dual to (ei)(e_{i}).

An FF-algebra AMd(k)A\cong M_{d}(k) may be presented as a cyclic algebra [31, Section 2.5] or as a crossed-product algebra [38, Section 2.6]. It follows readily from the algebraic constructions that, computationally, knowing a cyclic (resp. crossed-product) presentation of a central simple FF-algebra AA of degree dd is equivalent to knowing an embedding FAF\to A, where F/kF/k is a cyclic (resp. Galois) extension of kk of degree dd.

Field extensions and étale algebras admit a presentation as a quotient ring k[x]/(P(x))k[x]/(P(x)), for some Pk[x]P\in k[x]. As a result, their elements are represented as residue classes of polynomials. The question of tensor products of such algebras is discussed below in Section 3.2.

2.1.3 Computing rings of integers and factoring ideals

The ability to factor integers in polynomial time yields polynomial quantum algorithms for several tasks of computational number theory. We list some auxiliary results that will be needed later on:

Fact 2.1.

There exists a polynomial-time quantum algorithm for computing the maximal order in number fields.

Proof.

Zassenhaus’ round 2 algorithm [13, Algorithm 6.1.8] computes the maximal order of a number field in polynomial time, provided an oracle for factoring. Thus, one may deduce a polynomial quantum algorithm for computing maximal orders in number fields.

Fact 2.2.

There exists a polynomial-time quantum algorithm [14, Algorithm 2.3.22] for factoring ideals in number fields.

2.1.4 A polynomial quantum algorithm for computing class groups and groups of SS-units in number fields

The class group of LL, denoted by Cl(L)Cl(L) is the group of fractional ideals modulo principal ideals. The units in L\mathbb{Z}_{L} form a group which is called the unit group and is denoted by ULU_{L}. All these are important objects in number theory and they motivate the following three algorithmic problems:

  1. 1.

    Compute generators of Cl(L)Cl(L) and their relations,

  2. 2.

    Compute generators of ULU_{L},

  3. 3.

    Given a principal ideal II, find a generator element of II.

To the best of our knowledge, there only exist polynomial-time classical algorithms for these problems in very special cases. In [7] the authors propose a subexponential heuristic algorithm for computing class groups and unit groups for arbitrary degree number fields. In [6] and [8] the authors propose subexponential heuristic algorithms for the principal ideal problem for a certain subclass of number fields (multiquadratic extensions of \mathbb{Q} and power-of-two cyclotomic fields).

Hallgren [33] first proposed a polynomial-time quantum algorithm for computing the class group and unit group for bounded degree number fields, where GRH is assumed for computing generators of the class group. Later, polynomial-time algorithms were proposed (under GRH for the computation of the class group) for these tasks for arbitrary number fields [24],[9]. Furthermore, [9] also provides a polynomial-time quantum algorithm for computing generators of principal ideals (dubbed the principal ideal problem). In the class group computation, GRH is required to provide a polynomially sized generating set of the class group. Bach [5] proved that, assuming GRH, the class group can be generated by prime ideals whose norm is at most 48log2|ΔK|48\log^{2}|\Delta_{K}| where ΔK\Delta_{K} is the discriminant of the number field. We summarise this discussion into the following facts:

Fact 2.3 ([9, Fact 4.1]).

Let KK be a number field and ΔK\Delta_{K} its discriminant. Under GRH, there exists a polynomial-time algorithm computing a generator set of Cl(K)Cl(K) of size polynomial in log(|ΔK|)\log(|\Delta_{K}|).

Proof.

Using the Bach bound one simply enumerates all prime ideals of norm bounded by 12log2(|ΔK|)12\log^{2}(|\Delta_{K}|). The number of such prime ideals is polynomial in log(|ΔK|)\log(|\Delta_{K}|). ∎

Fact 2.4.

Assuming GRH, there exists a polynomial-time quantum algorithm [9] for computing the class group of a number field.

Fact 2.5.

[[9, Theorem 1.1]] Let SS be a set of prime ideals of a number field KK. There exists a polynomial-time quantum algorithm for computing generators of the group of SS-units (polynomial in the size of SS).

One application of SS-unit computation (mentioned in [9] and [46]) is that one can compute solutions to norm equations in Galois extensions. Let L|KL|K be a Galois extension of number field and let aKa\in K. Then one would like to find a solution to NL|K(x)=aN_{L|K}(x)=a (or show that it does not exist). The main result by Simon is the following:

Fact 2.6.

([46, Theorem 4.2]) Let L|KL|K be a Galois extension of number fields. Let S0S_{0} be a set of primes generating the relative class group Cli(L|K)Cl_{i}(L|K). Let SS be a set of primes containing S0S_{0} and let us denote the set of SS-units in LL by UL,SU_{L,S} and the set of SS-units in KK by UK,SU_{K,S}. Then one has that:

NL|K(UL,S)=NL|K(L)UK,SN_{L|K}(U_{L,S})=N_{L|K}(L^{*})\cap U_{K,S}

Informally, this means that for Galois extensions one can search for xx in the group of SS-units. Then by factoring SS and computing the class group of LL one can turn finding xx into solving a system of linear equations. Thus, computing SS-units implies a polynomial quantum algorithm for solving norm equations in Galois extensions. One important issue though is that the size of S0S_{0} has to polynomially bounded as the complexity could explode if S0S_{0} was too large. If one assumes GRH, then [5] implies that the size of S0S_{0} is not too large.

We summarise this discussion into the following theorem:

Theorem 2.7.

Let L|KL|K be a Galois extension of number fields and let SS be a set of primes of LL. Then there exists a polynomial-time quantum algorithm that computes a set of generators for the group of SS-units of LL. Furthermore, let aKa\in K. Then there exists a polynomial-time quantum algorithm (assuming GRH) for solving the norm equation NL|K(x)=aN_{L|K}(x)=a.

2.2 Brauer factor sets and central simple algebras

For the remainder of this section, we fix a field kk, a separable polynomial Pk[X]P\in k[X] of degree dd and an étale algebra F=k[X]/PF=k[X]/P over kk. We let KK be the smallest Galois extension of kk which splits FF, and set G=Gal(K/k)G=\operatorname{Gal}(K/k)

We will introduce definitions that are equivalent to those of Adamson cohomology [1] with values in the multiplicative group. However, while Adamson cohomology is normally defined for separable field extensions, we extend our definitions to étale algebras. One difference is that cochains are indexed by kk-algebra morphisms which are not necessarily injective. We may then use well-known results on Brauer factor sets to establish a cohomological description of the Brauer group (see [38, Chapter 3] for a modern reference).

Let Φ\Phi be the set of kk-algebra morphisms from FF to KK.

Lemma 2.8.

The map φφ(X)\varphi\mapsto\varphi(X) is a bijection between Φ\Phi and the set of roots of PP in KK. In particular, the set Φ\Phi has dd elements.

Proof.

If φΦ\varphi\in\Phi, it is clear that φ(X)\varphi(X) is a root of PP in KK. Now, the morphism φ\varphi is entirely determined by φ(X)\varphi(X) so the map φφ(X)\varphi\mapsto\varphi(X) is injective.

Let rKr\in K be a root of PP. Then, let QQ be the unique irreducible factor of PP such that Q(r)=0Q(r)=0. Then, Fk[X]/Qk[X]/(P/Q)F\simeq k[X]/Q\oplus k[X]/(P/Q) and we have a map from k[X]/Qk[X]/(P/Q)k[X]/Q\oplus k[X]/(P/Q) to k[X]/Qk[X]/Q sending (A(X),B(X))(A(X),B(X)) to A(r)A(r). This maps precomposed with the isomorphism defined above sends XX to rr. ∎

We may then index the maps from FF to KK, so that Φ={φ1,,φd}\Phi=\{\varphi_{1},\ldots,\varphi_{d}\}. If i[d]i\in[d] and gGg\in G, we write gigi for the index of gφig\varphi_{i}.

Definition 2.9.

Let nn\in\mathbb{N}. An nn-cochain is a map

a:Φn+1K×(φi0,,φin)ai0,,ina:\begin{array}[]{ccl}\Phi^{n+1}&\to&K^{\times}\\ (\varphi_{i_{0}},\ldots,\varphi_{i_{n}})&\mapsto&a_{i_{0},\ldots,i_{n}}\end{array}

which satisfies the following homogeneity condition: for i0,,in[d]i_{0},\ldots,i_{n}\in[d], for gGg\in G,

gai0,,in=agi0,,ginga_{i_{0},\ldots,i_{n}}=a_{gi_{0},\ldots,gi_{n}} (1)

A cochain aa may be denoted as the family (ai0,,in)i0,,in[d](a_{i_{0},\ldots,i_{n}})_{i_{0},\ldots,i_{n}\in[d]} of values that it takes over Φn+1\Phi^{n+1}.

A cochain is said to be reduced if ai,,i=1a_{i,\ldots,i}=1 for all i[d]i\in[d]. Under pointwise multiplication, the nn-cochains form a group which we denote by Cn(k,F)C^{n}(k,F). Then, the reduced cochains form a subgroup denoted by Cn(k,F)C^{\prime n}(k,F).

If aCn(k,F)a\in C^{n}(k,F) and i0,,in+1[d]i_{0},\ldots,i_{n+1}\in[d], we set

ai0,,ik^,,in+1ai0,,ik1,ik+1,,in+1.a_{i_{0},\ldots,\widehat{i_{k}},\ldots,i_{n+1}}\coloneqq a_{i_{0},\ldots,i_{k-1},i_{k+1},\ldots,i_{n+1}}.

For nn\in\mathbb{N}, there is a group homomorphism Δn:Cn(k,F)Cn+1(k,F)\Delta^{n}\colon C^{n}(k,F)\to C^{n+1}(k,F) defined as follows: if aCn(k,F)a\in C^{n}(k,F),

Δn(a)i0,,in+1=k=0n+1ai0,,ik^,,in(1)k.\Delta^{n}(a)_{i_{0},\ldots,i_{n+1}}=\prod_{k=0}^{n+1}a_{i_{0},\ldots,\widehat{i_{k}},\ldots,i_{n}}^{(-1)^{k}}.

We set Zn(k,F)=Ker(Δn)Z^{n}(k,F)=\operatorname{Ker}(\Delta^{n}) and call elements of this subgroup nn-cocycles. We write Zn(k,F)Z^{\prime n}(k,F) for the group of reduced cocycles. For nn positive, we also set Bn(k,F)=Im(Δn1)B^{n}(k,F)=\operatorname{Im}(\Delta^{n-1}) and call its elements nn-coboundaries. We note that Bn(k,F)B^{n}(k,F) is a subgroup of Zn(k,F)Z^{n}(k,F). For completeness, we let B0(k,F)B^{0}(k,F) be the trivial subgroup of Z0(k,F)Z^{0}(k,F). Finally, we set Hn(k,F)=Zn(k,F)/Bn(k,F)H^{n}(k,F)=Z^{n}(k,F)/B^{n}(k,F).

Two cocycles c1,c2Zn(k,F)c_{1},c_{2}\in Z^{n}(k,F) are called associated, denoted by c1c2c_{1}\sim c_{2}, if they have the same class in Hn(k,F)H^{n}(k,F). If bBn+1(k,F)b\in B^{n+1}(k,F), a preimage of bb by Δn\Delta^{n} is called a trivialisation of bb.

Remark 2.10.

The cochain (1)i,j,k[d](1)_{i,j,k\in[d]} is always a cocycle and is a coboundary for n1n\geq 1. It is called the trivial cocycle.

Proposition 2.11.

For nn positive, every nn-cocycle is associated to a reduced cocycle.

Proof.

Let cZn(k,F)c\in Z^{n}(k,F), and consider the cochain aCn1(k,F)a\in C^{n-1}(k,F) defined by ai,,i=ci,,i1a_{i,\ldots,i}=c_{i,\ldots,i}^{-1} and ai0,,in1=1a_{i_{0},\ldots,i_{n-1}}=1 if the iki_{k} are not all equal to one another. Then, the cocycle cΔn1(a)c\Delta^{n-1}(a) is reduced and associated to cc. ∎

We may now recall the explicit description of the isomorphism H2(k,F)Br(F/k)H^{2}(k,F)\simeq Br(F/k) (that is, the group of Brauer classes of central simple kk-algebras that are split by FF), as exposed in [38, Chapter 2]. Using Lemma 2.8, we observe that what is called a Brauer factor set there is a 22-cocycle in our notation.

Definition 2.12.

Let cZ2(k,F)c\in Z^{\prime 2}(k,F). Let B(F,c)B(F,c) be the kk-vector space of GG-homogeneous maps from Φ2\Phi^{2} to KK. We turn B(F,c)B(F,c) into a kk-algebra using the following definition. If ,B(F,c)\ell,\ell^{\prime}\in B(F,c), we set =′′\ell\ell^{\prime}=\ell^{\prime\prime} such that for all i,j[d]i,j\in[d],

i,j′′=k=1di,kci,k,jk,j.\ell^{\prime\prime}_{i,j}=\sum_{k=1}^{d}\ell_{i,k}c_{i,k,j}\ell^{\prime}_{k,j}.
Fact 2.13.

If cc is a 22-cocycle, the kk-algebra B(F,c)B(F,c) is a central simple kk-algebra. Furthermore, if c1c_{1} and c2c_{2} are associated cocycles, and aa is a 11-cochain such that c2=Δ1(a)c1c_{2}=\Delta^{1}(a)c_{1}, then the map

B(F,c1)B(F,c2)(mi,j)i,j[d](mi,jai,j1)i,j[d]\begin{array}[]{ccl}B(F,c_{1})&\to&B(F,c_{2})\\ (m_{i,j})_{i,j\in[d]}&\mapsto&(m_{i,j}a_{i,j}^{-1})_{i,j\in[d]}\end{array}

is an isomorphism. Conversely, if AA is a central simple kk-algebra split by FF, then there exists a cocycle cZ2(k,F)c\in Z^{2}(k,F) such that AB(F,c)A\simeq B(F,c) If c1,c2Z2(k,F)c_{1},c_{2}\in Z^{\prime 2}(k,F) are such that B(F,c1)B(F,c_{1}) and B(F,c2)B(F,c_{2}) are isomorphic algebras, then the cocycles c1c_{1} and c2c_{2} are associated.

Proof.

This is essentially the content of [38, Theorem 2.5.6]. While the result is stated in the source for reduced factor sets, we note that the isomorphism constructed in [38, Theorem 2.3.21] is valid even for non-reduced associated cocycles. Then, the results stated here extend to non-reduced cocycle by 2.11. ∎

Remark 2.14.

We note that if c1,c2Zn(k,F)c_{1},c_{2}\in Z^{\prime n}(k,F) are associated and both reduced, then a cochain aCn1(k,F)a\in C^{n-1}(k,F) such that c1=Δn1(a)c2c_{1}=\Delta^{n-1}(a)c_{2} must be reduced as well.

We refer the reader to [38, Section 2.3] for the detailed construction of a Brauer factor-set associated to a given central simple algebra, and we give a brief description below: Let AA be a central simple kk-algebra, and let FF be a separable maximal commutative subalgebra of AA. Then, there exists vAv\in A such that A=FvFA=FvF. Furthermore, letting KK be a Galois extension of kk which splits FF, there is a map AMd(K)A\to M_{d}(K) such that every uFu\in F is mapped to Diag(ϕ1(x),,ϕd(x))\operatorname{Diag}(\phi_{1}(x),\ldots,\phi_{d}(x)). Let (vij)i,j[d](v_{ij})_{i,j\in[d]} be the image of vv in Md(K)M_{d}(K). Then, every element uvuuvu^{\prime} of AA becomes a matrix of the form (ijvij)i,j[d](\ell_{ij}v_{ij})_{i,j\in[d]}, where ij=ϕi(u)ϕj(u)\ell_{ij}=\phi_{i}(u)\phi_{j}(u^{\prime}). The map uvu(ij)i,j[d]uvu^{\prime}\mapsto(\ell_{ij})_{i,j\in[d]} gives an isomorphism from AA to B(F,c)B(F,c), where cijkvikvkjvij1c_{ijk}\coloneqq v_{ik}v_{kj}v_{ij}^{-1}.

3 Explicit computations with Amitsur cohomology

In this section, we recall the basic definitions of Amitsur cohomology and then use them to get efficient representations of cocycles as introduced in Section 2.2. We also keep notations as they were in Section 2.2.

Although the Amitsur complex and its cohomology are defined for general commutative rings, we will only need to state the definitions over fields and étale algebras. We refer to [29, Chapter 5] for more general definitions and results.

Unless specified otherwise, all tensor products are taken over kk, and by FnF^{\otimes n} we mean the tensor product of nn copies of the kk-algebra FF.

Definition 3.1.

An nn-cochain in the sense of Amitsur is an invertible element in the kk-algebra Fn+1F^{\otimes n+1}, and we write CAmn(k,F)C_{Am}^{n}(k,F) for the group of nn-cochains.

For nn\in\mathbb{N} and 0in+10\leq i\leq n+1, we define a kk-algebra homomorphism εin\varepsilon_{i}^{n} from Fn+1F^{\otimes n+1} to Fn+2F^{\otimes n+2}. The map εin\varepsilon_{i}^{n} is defined on the simple tensor elements as follows:

εin(f0fn)=f0fi11fifn.\varepsilon_{i}^{n}(f_{0}\otimes\ldots\otimes f_{n})=f_{0}\otimes\ldots\otimes f_{i-1}\otimes 1\otimes f_{i}\otimes\ldots\otimes f_{n}.

We then define the group homomorphism

ΔAmn:CAmn(k,F)CAmn+1(k,F)ai=0n+1εin(a)(1)i.\Delta_{Am}^{n}\colon\begin{array}[]{ccl}C_{Am}^{n}(k,F)&\to&C_{Am}^{n+1}(k,F)\\ a&\mapsto&\prod_{i=0}^{n+1}\varepsilon_{i}^{n}(a)^{(-1)^{i}}\end{array}.

Define the subgroup ZAmn(k,F)=Ker(ΔAmn)Z_{Am}^{n}(k,F)=\operatorname{Ker}(\Delta_{Am}^{n}) of CAmn(k,F)C_{Am}^{n}(k,F), and its elements are called nn-cocycles in the sense of Amitsur.

If nn is positive, we also define the subgroup BAmn(k,F)=Im(ΔAmn1)B_{Am}^{n}(k,F)=\operatorname{Im}(\Delta_{Am}^{n-1}) of ZAmn(k,F)Z_{Am}^{n}(k,F) and its elements are nn-coboundaries in the sense of Amitsur. Then, the nnth Amitsur cohomology group HAmn(k,F)H_{Am}^{n}(k,F) is defined as the quotient group ZAmn(k,F)/BAmn(k,F)Z_{Am}^{n}(k,F)/B_{Am}^{n}(k,F).

Two nn-cocycles c1c_{1} and c2c_{2} are called associated if they have the same class in HAmn(k,F)H_{Am}^{n}(k,F). If bBAmn(k,F)b\in B^{n}_{Am}(k,F), a cochain aCAmn1(k,F)a\in C^{n-1}_{Am}(k,F) such that ΔAmn1(a)=c\Delta_{Am}^{n-1}(a)=c is called a trivialisation of cc.

3.1 Amitsur cohomology and central simple algebras

The first result we need to leverage Amitsur cohomology in representing Brauer factor sets is the following

Lemma 3.2.

Consider the map Ψn\Psi_{n} from Fn+1F^{\otimes n+1} to the algebra of GG-homogeneous maps from Φn+1\Phi^{n+1} to KK, which is defined over simple tensors by

f0fn(Φn+1K(φi0,,φin)j=0nφij(fj)).f_{0}\otimes\ldots\otimes f_{n}\mapsto\left(\begin{array}[]{ccl}\Phi^{n+1}&\to&K\\ (\varphi_{i_{0}},\ldots,\varphi_{i_{n}})&\mapsto&\prod_{j=0}^{n}\varphi_{i_{j}}(f_{j})\end{array}\right).

The map Ψ\Psi is an isomorphism of kk-algebras, which also sends the unit group CAmn(k,F)C_{Am}^{n}(k,F) onto the unit group Cn(k,F)C^{n}(k,F). Furthermore, Ψn+1ΔAmn=ΔnΨn\Psi_{n+1}\circ\Delta_{Am}^{n}=\Delta^{n}\circ\Psi_{n}, and it follows that cocycle and coboundaries subgroup are also isomorphic to one another, and so are cohomology groups.

Proof.

If FF is a separable field extension of kk, this is the content of [45, Section 2]. However, the proof given there readily generalises to the case that FF is an étale kk-algebra. ∎

If (φi0,,φin)Φn+1(\varphi_{i_{0}},\ldots,\varphi_{i_{n}})\in\Phi^{n+1}, we get a homomorphism of kk-algebras

Fn+1Ka0anj=0nφij(aj)\begin{array}[]{ccc}F^{\otimes n+1}&\to&K\\ a_{0}\otimes\ldots\otimes a_{n}&\mapsto&\prod_{j=0}^{n}\varphi_{i_{j}}(a_{j})\end{array}

By the construction of Ψ\Psi, if aFn+1a\in F^{\otimes n+1} and (φi0,,φin)Φn+1(\varphi_{i_{0}},\ldots,\varphi_{i_{n}})\in\Phi^{n+1}, then we get

Ψn(a)i0,,in=(φi0,,φin)(a).\Psi_{n}(a)_{i_{0},\ldots,i_{n}}=(\varphi_{i_{0}},\ldots,\varphi_{i_{n}})(a). (2)

Adapting Definition 2.12 to Amitsur cohomology, we get:

Definition 3.3.

Let cZAm2(k,F)c\in Z_{Am}^{2}(k,F) be a cocycle. The kk-algebra A(F,c)A(F,c) is defined as follows:

As a kk-vector space, A(F,c)A(F,c) is F2F^{\otimes 2}. We see F3F^{\otimes 3} as a F2F^{\otimes 2}-algebra via the embedding ε11:F2F3\varepsilon_{1}^{1}\colon F^{\otimes 2}\to F^{\otimes 3}. Then, multiplication in A(F,c)A(F,c) is defined by

aa=TrF3/F2(ε21(a)cε01(a)).aa^{\prime}=\operatorname{Tr}_{F^{\otimes 3}/F^{\otimes 2}}\left(\varepsilon_{2}^{1}(a)c\varepsilon_{0}^{1}(a^{\prime})\right).

In the sequel, we use the same embedding whenever we write TrF3/F2\operatorname{Tr}_{F^{\otimes 3}/F^{\otimes 2}}.

Observe that the algebra F3/F2F^{\otimes 3}/F^{\otimes 2} may be seen as the extension of scalars to F2F^{\otimes 2} of the kk-algebra FF. Since we make the choice of seeing F3F^{\otimes 3} as a F2F^{\otimes 2}-algebra via the mapping ε12\varepsilon_{1}^{2}, the relevant injection FF3F\to F^{\otimes 3} sends an element aFa\in F to 1a11\otimes a\otimes 1. Then, by [10, Section III.9.3], for aFa\in F, we have

TrF3/F2(1a1)=TrF/k(a)(11).\operatorname{Tr}_{F^{\otimes 3}/F^{\otimes 2}}(1\otimes a\otimes 1)=\operatorname{Tr}_{F/k}(a)(1\otimes 1).

By linearity of the trace, if a=iIai0ai1ai2F3a=\sum_{i\in I}a_{i0}\otimes a_{i1}\otimes a_{i2}\in F^{\otimes 3}, we get

TrF3/F2(a)=iITrF/k(ai1)(ai0ai2).\operatorname{Tr}_{F^{\otimes 3}/F^{\otimes 2}}(a)=\sum_{i\in I}\operatorname{Tr}_{F/k}(a_{i1})(a_{i0}\otimes a_{i2}).
Proposition 3.4.

Let cZAm2(k,F)c\in Z_{Am}^{2}(k,F). The map Ψ1\Psi_{1} is a kk-algebra isomorphism from A(F,c)A(F,c) to B(F,Ψ2(c))B(F,\Psi_{2}(c)).

Proof.

The map Ψ1\Psi_{1} is already known to be an isomorphism of vector space between A(F,c)A(F,c) and B(F,Ψ2(c))B(F,\Psi_{2}(c)), as the underlying vector spaces of these algebras are respectively F2F^{\otimes 2} and the space of GG-homogeneous maps from Φ2\Phi^{2} to KK. We therefore only need to check that the map Ψ1\Psi_{1} is a homomorphism of kk-algebras.

It is well known that if aFa\in F, TrF/k(a)=[d]φ(a)\operatorname{Tr}_{F/k}(a)=\sum_{\ell\in[d]}\varphi_{\ell}(a). Then, if a=a0a1a2F3a=a_{0}\otimes a_{1}\otimes a_{2}\in F^{\otimes 3} and i,j[d]i,j\in[d], we get:

(φi,φj)(TrF3/F2(a))\displaystyle(\varphi_{i},\varphi_{j})\left(\operatorname{Tr}_{F^{\otimes 3}/F^{\otimes 2}}(a)\right) =(φi,φj)(TrF/k(a1)(a0a2))\displaystyle=(\varphi_{i},\varphi_{j})\left(\operatorname{Tr}_{F/k}(a_{1})(a_{0}\otimes a_{2})\right)
=(φi,φj)([d]φ(a1)(a0a2))\displaystyle=(\varphi_{i},\varphi_{j})\left(\sum_{\ell\in[d]}\varphi_{\ell}(a_{1})(a_{0}\otimes a_{2})\right)
=[d]φi(a0)φ(a1)φj(a2)\displaystyle=\sum_{\ell\in[d]}\varphi_{i}(a_{0})\varphi_{\ell}(a_{1})\varphi_{j}(a_{2})
=[d](φi,φ,φj)(a0a1a2).\displaystyle=\sum_{\ell\in[d]}(\varphi_{i},\varphi_{\ell},\varphi_{j})(a_{0}\otimes a_{1}\otimes a_{2}).

Extending this result by linearity, we get that

(φi,φj)TrF3/F2=[d](φi,φ,φj).(\varphi_{i},\varphi_{j})\circ\operatorname{Tr}_{F^{\otimes 3}/F^{\otimes 2}}=\sum_{\ell\in[d]}(\varphi_{i},\varphi_{\ell},\varphi_{j}).

We also note that if aF2a\in F^{\otimes 2}, then by Equation 2 we get

(φi,φ,φj)(ε21(a))=(φi,φ)(a)=Ψ1(a)i,.(\varphi_{i},\varphi_{\ell},\varphi_{j})(\varepsilon_{2}^{1}(a))=(\varphi_{i},\varphi_{\ell})(a)=\Psi_{1}(a)_{i,\ell}.

Likewise,

(φi,φ,φj)(ε01(a))=Ψ1(a),j.(\varphi_{i},\varphi_{\ell},\varphi_{j})(\varepsilon_{0}^{1}(a))=\Psi_{1}(a)_{\ell,j}.

Now, using Equation 2 again, we get that if a,aA(F,c)a,a^{\prime}\in A(F,c) and i,j[d]i,j\in[d],

Ψ1(aa)i,j\displaystyle\Psi_{1}(aa^{\prime})_{i,j} =(φi,φj)(TrF3/F2(ε21(a)cε01(a)))\displaystyle=(\varphi_{i},\varphi_{j})\left(\operatorname{Tr}_{F^{\otimes 3}/F^{\otimes 2}}\left(\varepsilon_{2}^{1}(a)c\varepsilon_{0}^{1}(a^{\prime})\right)\right)
=[d](φi,φ,φj)(ε21(a)cε01(a))\displaystyle=\sum_{\ell\in[d]}(\varphi_{i},\varphi_{\ell},\varphi_{j})\left(\varepsilon_{2}^{1}(a)c\varepsilon_{0}^{1}(a^{\prime})\right)
=[d](φi,φ,φj)(ε21(a))(φi,φ,φj)(c)(φi,φ,φj)(ε01(a))\displaystyle=\sum_{\ell\in[d]}(\varphi_{i},\varphi_{\ell},\varphi_{j})(\varepsilon_{2}^{1}(a))(\varphi_{i},\varphi_{\ell},\varphi_{j})(c)(\varphi_{i},\varphi_{\ell},\varphi_{j})(\varepsilon_{0}^{1}(a^{\prime}))
=[d]Ψ1(a)i,Ψ2(c)i,,jΨ1(a),j\displaystyle=\sum_{\ell\in[d]}\Psi_{1}(a)_{i,\ell}\Psi_{2}(c)_{i,\ell,j}\Psi_{1}(a^{\prime})_{\ell,j}
=Ψ1(a)Ψ1(a),\displaystyle=\Psi_{1}(a)\Psi_{1}(a^{\prime}),

where the multiplication in the last line is meant to be in B(F,Ψ2(c))B(F,\Psi_{2}(c)). ∎

Remark 3.5.

Let 𝟏F3\mathbf{1}\in F^{\otimes 3} be the trivial cocycle. There is an explicit isomorphism between A(F,𝟏)A(F,\mathbf{1}) and Md(k)M_{d}(k). Indeed, the algebra Md(k)M_{d}(k) is isomorphic to the algebra FFF\otimes F^{\vee} of kk-linear endomorphisms of FF, where FF^{\vee} is the dual kk-vector space of FF. In fact, the map aTrF/k(a):-(bTrF/k(ab))a\mapsto\operatorname{Tr}_{F/k}(a\cdot)\coloneq\left(b\mapsto\operatorname{Tr}_{F/k}(ab)\right) gives a kk-linear isomorphism from FF to its dual space FF^{\vee}, so we may see Md(k)M_{d}(k) as the algebra FFF\otimes F with the following multiplication law:

(aa)(bb)=TrF/k(ab)ab.(a\otimes a^{\prime})(b\otimes b^{\prime})=\operatorname{Tr}_{F/k}(a^{\prime}b)a\otimes b^{\prime}.

Let a,a,b,bFa,a^{\prime},b,b^{\prime}\in F, we compute the following multiplication in A(F,𝟏)A(F,\mathbf{1}):

(aa)(bb)\displaystyle(a\otimes a^{\prime})(b\otimes b^{\prime}) =TrF3/F2(ε21(aa)ε01(bb))\displaystyle=\operatorname{Tr}_{F^{\otimes 3}/F^{\otimes 2}}\left(\varepsilon_{2}^{1}(a\otimes a^{\prime})\varepsilon_{0}^{1}(b\otimes b^{\prime})\right)
=TrF3/F2(aabb)\displaystyle=\operatorname{Tr}_{F^{\otimes 3}/F^{\otimes 2}}(a\otimes a^{\prime}b\otimes b^{\prime})
=TrF/k(ab)ab.\displaystyle=\operatorname{Tr}_{F/k}(a^{\prime}b)a\otimes b^{\prime}.

It follows that if \mathcal{B} is a basis of FF and we identify elements of FF with the column vectors of their components in basis \mathcal{B}, and for aFa\in F, we set φ(a)\varphi(a) to be the row matrix of the linear form bTrF/k(ab)b\mapsto\operatorname{Tr}_{F/k}(ab), then an isomorphism from A(F,𝟏)A(F,\mathbf{1}) to Md(k)M_{d}(k) is given by

aaaφ(a).a\otimes a^{\prime}\mapsto a\varphi(a^{\prime}).
Theorem 3.6.

Let AA be a central simple kk-algebra, let FAF\subset A be a separable maximal commutative subalgebra, and let vAv\in A be such that A=FvFA=FvF. Then, there is a cocycle cZAm2(k,F)c\in Z^{2}_{Am}(k,F) such that the map

Φv:F2Ar1r2r1vr2\begin{array}[]{rlcl}\Phi_{v}\colon&F^{\otimes 2}&\to&A\\ &r_{1}\otimes r_{2}&\mapsto&r_{1}vr_{2}\end{array}

is an isomorphism of kk-algebras from A(F,c)A(F,c) to AA (which makes sense as F2F^{\otimes 2} is the underlying kk-vector space of A(F,c)A(F,c)). Furthermore, cc is the unique element of F3F^{\otimes 3} with this property (if we extend the definition of the multiplication in A(F,c)A(F,c) to non-cocycle elements).

Proof.

Let cB2(k,F)c^{\prime}\in B^{2}(k,F) be the Brauer factor set constructed from the data of F,AF,A and vv as in the discussion at the end of Section 2.2, and let θv\theta_{v} be the isomorphism from AA to B(F,c)B(F,c^{\prime}) constructed there. We also let c=Ψ21(c)c=\Psi_{2}^{-1}(c^{\prime}). The fact that Φv\Phi_{v} is an isomorphism of kk-algebras follows from the observation that θvΦv=Ψ1\theta_{v}\circ\Phi_{v}=\Psi_{1} and from 3.4.

The cocycle cc is a solution to the system of F2F^{\otimes 2}-linear equations

TrF3/F2(ε21(u)cε01(u))=Φv1(Φv(u)Φv(u))\operatorname{Tr}_{F^{\otimes 3}/F^{\otimes 2}}\left(\varepsilon_{2}^{1}(u)c\varepsilon_{0}^{1}(u^{\prime})\right)=\Phi_{v}^{-1}\left(\Phi_{v}(u)\Phi_{v}(u^{\prime})\right)

for all u,uF2u,u^{\prime}\in F^{\otimes 2}. Observe that ε21(u)ε01(u)\varepsilon_{2}^{1}(u)\varepsilon_{0}^{1}(u^{\prime}) ranges over F3F^{\otimes 3} as u,uu,u^{\prime} range over F2F^{\otimes 2}. Then, the uniqueness of cc follows from the fact that the map cTrF3/F2(c)c\mapsto\operatorname{Tr}_{F^{\otimes 3}/F^{\otimes 2}}(c\cdot) is an isomorphism from F3F^{\otimes 3} to its dual as a F2F^{\otimes 2}-module.

This last fact, in turn, follows from [29, Corollary 4.6.8 and Proposition 4.6.1]. Indeed, F3F2[X]/PF^{\otimes 3}\simeq F^{\otimes 2}[X]/P, and Pk[X]F2[X]P\in k[X]\subset F^{\otimes 2}[X] is a separable polynomial, so F3F^{\otimes 3} is a separable F2F^{\otimes 2}-algebra. ∎

Corollary 3.7.

Let c,cZAm2(k,F)c,c^{\prime}\in Z_{Am}^{2}(k,F) be associated cocycles, and let aCam2(k,F)a\in C_{am}^{2}(k,F) be such that c=cΔAm1(a)c^{\prime}=c\Delta_{Am}^{1}(a). Then, the map mma1m\mapsto ma^{-1}, where the multiplication used is the natural one in F2F^{\otimes 2}, is an isomorphism from A(F,c)A(F,c) to A(F,c)A(F,c^{\prime}).

Proof.

This is just a consequence of the analogous property for Brauer factor sets (2.13), pulled back through the isomorphisms Ψi\Psi_{i}. ∎

Remark 3.8.

For the sake of efficiency, we use already proven results on Brauer factor-sets and the isomorphisms Ψ1\Psi_{1} and Ψ2\Psi_{2} to describe the isomorphism between the groups HAm2(k,F)H^{2}_{Am}(k,F) and Br(F/k)\operatorname{Br}(F/k). However, the simplicity of the isomorphism Φc\Phi_{c} described in Theorem 3.6 suggests that Amitsur cohomology is a natural setting for giving an algebraic description of Br(F/k)\operatorname{Br}(F/k) and that direct proofs of these facts may be given without using references to Brauer factor-sets.

3.2 Computational representation of Amitsur cohomology

Once a field kk and an étale kk-algebra F=k[X]/PF=k[X]/P are fixed, for nn\in\mathbb{N}, we have an isomorphism

Ξn:Fn+1Rnk[X0,,Xn]/In,\Xi_{n}\colon F^{\otimes n+1}\to R_{n}\coloneqq k[X_{0},\ldots,X_{n}]/I_{n},

where InI_{n} is the ideal generated by the polynomials P(Xi)P(X_{i}) for 0in0\leq i\leq n. Furthermore, an element of RnR_{n} admits a unique lift in k[X0,,Xn]k[X_{0},\ldots,X_{n}] such that the individual degrees in X0,X1,,XnX_{0},X_{1},\ldots,X_{n} are all lesser or equal to d1d-1. If Qk[X0,,Xn]Q\in k[X_{0},\ldots,X_{n}], we denote by Q~\widetilde{Q} the lift of QmodIQ\mod I as described above. There is an obvious polynomial algorithm which, given some Qk[X0,,Xn]Q\in k[X_{0},\ldots,X_{n}], computes Q~\widetilde{Q}. Indeed, one may simply reduce QQ modulo P(X0)P(X_{0}), then modulo P(X1)P(X_{1}), and so on… This allows us to represent an element QRnQ\in R_{n} as its lift Q~\widetilde{Q}, and in general we get Q1+Q2=Q1~+Q2~modInQ_{1}+Q_{2}=\widetilde{Q_{1}}+\widetilde{Q_{2}}\mod I_{n} and Q1Q2=Q1~Q2~modInQ_{1}Q_{2}=\widetilde{Q_{1}}\widetilde{Q_{2}}\mod I_{n}.

In this setting, the map εin\varepsilon_{i}^{n} becomes the map

QmodInQ~(X0,,Xi1,Xi+1,,Xn)modIn+1.Q\mod I_{n}\mapsto\widetilde{Q}(X_{0},\ldots,X_{i-1},X_{i+1},\ldots,X_{n})\mod I_{n+1}.

That is, we define a map

νin(Xj)={Xjif j<iXj+1otherwise.\nu_{i}^{n}(X_{j})=\begin{cases}X_{j}\quad\text{if }j<i\\ X_{j+1}\quad\text{otherwise.}\end{cases}

from k[X0,,Xn]k[X_{0},\ldots,X_{n}] to k[X0,,Xn+1]k[X_{0},\ldots,X_{n+1}], and we define

εin:RnRn+1QmodInνin(Q~)modIn+1.\begin{array}[]{rccc}\varepsilon_{i}^{n}\colon&R_{n}&\to&R_{n+1}\\ &Q\mod I_{n}&\mapsto&\nu_{i}^{n}(\widetilde{Q})\mod I_{n+1}.\end{array}

The maps defined above are compatible with the eponymous maps from the Amitsur complex. That is, we have εinΞn=Ξn+1εin\varepsilon_{i}^{n}\circ\Xi_{n}=\Xi_{n+1}\circ\varepsilon_{i}^{n}.

Now, the trace map of R2R_{2} as an R1R_{1} algebra (via the morphism ε11\varepsilon_{1}^{1}) may easily be computed in the R1R_{1}-basis (X1i)0i<d(X_{1}^{i})_{0\leq i<d} of R2R_{2}. Let cZAm2(k,F)c\in Z^{2}_{Am}(k,F), Q=Ξ2(c)Q=\Xi_{2}(c), a,aA(F,c)a,a^{\prime}\in A(F,c), T=Ξ1(a)T=\Xi_{1}(a) and T=Ξ1(a)T^{\prime}=\Xi_{1}(a^{\prime}). We set T′′=Ξ11(aa)T^{\prime\prime}=\Xi_{1}^{-1}(aa^{\prime}), and we have

T′′=TrR2/R1(ε21(T)Qε01(T)).T^{\prime\prime}=\operatorname{Tr}_{R_{2}/R_{1}}\left(\varepsilon_{2}^{1}(T)Q\varepsilon_{0}^{1}(T^{\prime})\right).

That is,

Ξ1(aa)=TrR2/R1(ε21(Ξ1(a))Ξ2(c)ε01(Ξ1(a))).\Xi_{1}(aa^{\prime})=\operatorname{Tr}_{R_{2}/R_{1}}\left(\varepsilon_{2}^{1}(\Xi_{1}(a))\Xi_{2}(c)\varepsilon_{0}^{1}(\Xi_{1}(a^{\prime}))\right).

For the remainder of this work, elements of algebras of the form Fn+1F^{\otimes n+1}, when Fk[X]/(P(X))F\simeq k[X]/(P(X)) is an étale kk-algebra, are represented algorithmically as their images by Ξn\Xi_{n} in RnR_{n}.

3.3 Representing central simple algebras as Amitsur algebras

We may now give an algorithm for finding a representation of a central simple algebra AA over any field of sufficiently large size where linear algebra tasks may be performed efficiently.

Input: A field kk
Input: A central simple kk-algebra AA such that |k|>dimkA|k|>\dim_{k}A
Output: uAu\in A such that F=k(u)F=k(u) is a maximal commutative subalgebra of AA
Output: Pk[X]P\in k[X], the minimal polynomial of uu
Output: cZAm2(k,F)c\in Z_{Am}^{2}(k,F)
Output: A linear map e:F2Ae\colon F^{\otimes 2}\to A that is an isomorphism from A(F,c)A(F,c) to AA
1 Find uAu\in A such that Fk[u]F\coloneqq k[u] is a maximal separable commutative subalgebra of AA ;
2 Compute PP, the minimal polynomial of uu;
3 Find vAv\in A such that A=FvFA=FvF;
4 Compute the kk-vector space isomorphism e:F2Ae:F^{\otimes 2}\to A sending uiuju^{i}\otimes u^{j} to uivuju^{i}vu^{j};
5 Compute cF3c\in F^{\otimes 3} such that for 0i,j,i,jd10\leq i,j,i^{\prime},j^{\prime}\leq d-1, e(uiuj)e(uiuj)=TrF3/F2(ε21(uiuj)cε01(uiuj))e(u^{i}\otimes u^{j})e(u^{i^{\prime}}\otimes u^{j^{\prime}})=\operatorname{Tr}_{F^{\otimes 3}/F^{\otimes 2}}(\varepsilon_{2}^{1}(u^{i}\otimes u^{j})c\varepsilon_{0}^{1}(u^{i^{\prime}}\otimes u^{j^{\prime}}));
return (u,P,c,e)(u,P,c,e)
Algorithm 1 Computing a presentation of a central simple algebra as an Amitsur algebra

Before we prove the correctness and efficiency of Algorithm 1, we need a lemma:

Lemma 3.9.

Let kk be a field and let AA be a central simple kk-algebra. Assume that |k|>dimkA|k|>\dim_{k}A. Let uAu\in A be such that Fk[u]F\coloneqq k[u] is a maximal commutative subalgebra of AA. Then an element vAv\in A such that A=FvFA=FvF may be found in probabilistic polynomial time.

Proof.

For vv in AA, by an argument of dimensions over kk, we observe that A=FvFA=FvF if and only if the map

e:FFAf1f2f1vf2e\colon\begin{array}[]{ccl}F\otimes F&\to&A\\ f_{1}\otimes f_{2}&\mapsto&f_{1}vf_{2}\end{array}

is injective.

The family (uiuj)0i,jdegA1(u^{i}\otimes u^{j})_{0\leq i,j\leq\deg A-1} is a basis of F2F^{\otimes 2}. Let B=(b1,,bdimA)B=(b_{1},\ldots,b_{\dim A}) be the input basis of AA (that is, the basis for which the structure constants of AA are expressed). Then, identifying ee with its matrix in the bases written above, the determinant of ee is a homogeneous polynomial on the coordinates of vv in basis BB of degree bounded by dimA\dim A. Furthermore, A=FvFA=FvF if and only if vv is not a zero of this polynomial. We note that by [38, Theorem 2.2.2], this polynomial is nonzero.

Letting SS be a finite subset of kk, the Schwartz-Zippel lemma ensures that a random vv in Sb1SbdimASb_{1}\oplus\ldots\oplus Sb_{\dim A} satisfies this condition with probability larger than 1dimA|S|1-\frac{\dim A}{|S|}.

Therefore, if |k|>dimA|k|>\dim A, we may pick SS large enough that vv has the desired property with positive probability, and small enough that we may draw a random element of SS efficiently. For instance, take |S|=dimA+1|S|=\dim A+1. Then, vv has the desired property with probability larger than 1dimA+1\frac{1}{\dim A+1}. In this case, it is expected to take less than dimA+1\dim A+1 attempts to find a suitable element vAv\in A. ∎

Theorem 3.10.

If kk is a field over which linear algebra may be performed efficiently, and AA is a central simple kk-algebra such that |k|>dimkA|k|>\dim_{k}A, then Algorithm 1 is correct and runs in probabilistic polynomial time.

Proof.

We first prove the correctness of the algorithm. The first 3 steps compute u,vAu,v\in A, as well as Pk[X]P\in k[X], the minimal polynomial of uu. We obtain a subalgebra F=k[u]k[X]/PF=k[u]\simeq k[X]/P of AA and vAv\in A such that A=FvFA=FvF. Then, by Theorem 3.6, there exists a cocycle cZAm2(k,F)c\in Z_{Am}^{2}(k,F) such that e:abavbe\colon a\otimes b\mapsto avb is an isomorphism from A(k,F)A(k,F) to AA, and such a cc is unique in F3F^{\otimes 3}. The equation solved in Algorithm 1 finds a representation of this cc.

Now we consider the complexity of the algorithm:

The element uAu\in A in Algorithm 1 may be found using the polynomial algorithm given in [22].

Then, vv in Algorithm 1 may be found in probabilistic polynomial time using the algorithm of Lemma 3.9.

Then, the remaining lines involve arithmetic in the kk-algebras AA, R1R_{1} and R2R_{2}, as well as the computation of the solution of a system of linear equations.

All in all, this makes Algorithm 1 a polynomial probabilistic algorithm. ∎

4 Trivialisation of Amitsur cocycles

In this section, we present an algorithm for computing a trivialisation of a coboundary using SS-unit group computation. For this purpose, we prove Theorem 4.11, which states that a trivialisation of a coboundary may be found in an appropriate group of SS-units. This result is to be compared with 2.6 and [25, Theorem 7]. Then, we may state Algorithm 2 which computes a trivialisation of a given coboundary by solving the related equation in an appropriate group of SS-units. Furthermore, we get Theorem 4.13 which, assuming GRH, solves the explicit isomorphism problem in quantum polynomial time.

The technical part of this section is devoted to proving Theorem 4.11. The strategy we employ is similar to the proof of [25, Theorem 7]. That is, we prove Lemma 4.10, which is a generalisation to divisors of Hilbert’s Theorem 90 on the vanishing of the first cohomology group. While [25, Lemma 9] is stated for fractional ideals of a number field, our setting requires us bring this machinery to the setting of étale algebras. We therefore first introduce definitions and basic results for the divisors of an étale algebra over a global field in Section 4.1.

4.1 Divisors of étale algebras

Let kk be a global field. Then, recall that a commutative kk-algebra FF is étale if and only if it satisfies the following equivalent conditions:

  1. 1.

    The algebra FF factors as a direct product FF1××FrF\simeq F_{1}\times\ldots\times F_{r} of finite separable extensions of kk.

  2. 2.

    The algebra FF is isomorphic to an algebra of the form k[X]/(P(X))k[X]/(P(X)), where Pk[X]P\in k[X] is a separable polynomial (that is, PP has no multiple root in an algebraic closure of kk).

Remark 4.1.

If FF is an étale kk-algebra, the isomorphism FF1××FrF\simeq F_{1}\times\ldots\times F_{r} discussed above is unique up to reindexing and automorphism of the factors. In fact, the factors identify with the minimal ideals of FF. Therefore, we usually identify an étale algebra FF with the product F1××FrF_{1}\times\ldots\times F_{r}. We will also denote by πi\pi_{i} the projection map FFiF\to F_{i}.

If FF is a global field, we write Pl(F)\operatorname{Pl}(F) for the set of non-archimedean places of FF. Then, recall that the divisor group 𝒟(F)\operatorname{\mathcal{D}}(F) is the free abelian group on the set Pl(F)\operatorname{Pl}(F), that an element aFa\in F has an associated divisor 𝒟(a)=PPl(F)ordP(a)P\operatorname{\mathcal{D}}(a)=\sum_{P\in\operatorname{Pl}(F)}\operatorname{ord}_{P}(a)P, and that if 𝒫(F)={𝒟(a):aF}\operatorname{\mathcal{P}}(F)=\{\operatorname{\mathcal{D}}(a)\colon a\in F\} is the group of principal divisors, then the class group Cl(F)=𝒟(F)/𝒫(F)\operatorname{Cl}(F)=\operatorname{\mathcal{D}}(F)/\operatorname{\mathcal{P}}(F) of FF is finite. When FF is a number field, the divisor group is isomorphic to the group of fractional ideals of FF, and the class group in terms of divisors is then isomorphic to the ideal class group.

Definition 4.2.

Let F=F1××FrF=F_{1}\times\ldots\times F_{r} be an étale kk-algebra factored as above. The set of non-archimedean places of FF is the disjoint union Pl(F)=i[r]Pl(Fi)\operatorname{Pl}(F)=\bigsqcup_{i\in[r]}\operatorname{Pl}(F_{i}). The divisor group 𝒟(F)\operatorname{\mathcal{D}}(F) of FF is the free abelian group on Pl(F)\operatorname{Pl}(F). The divisor 𝒟(a)\operatorname{\mathcal{D}}(a) of an invertible element aF×a\in F^{\times} is defined as the sum of the divisors of its projections in the factors FiF_{i} of FF, and the groups 𝒫(F)\operatorname{\mathcal{P}}(F) and Cl(F)\operatorname{Cl}(F) are defined by analogy with their definitions for global fields. If D𝒟(F)=PPl(F)nPPD\in\operatorname{\mathcal{D}}(F)=\sum_{P\in\operatorname{Pl}(F)}n_{P}P, we let Supp(D)={PPl(F):nP0}\operatorname{Supp}(D)=\{P\in\operatorname{Pl}(F)\colon n_{P}\neq 0\} be the support of DD.

An immediate consequence of this definition is that the class group of an étale kk-algebra is finite.

If F=F1××FrF=F_{1}\times\ldots\times F_{r} and F=F1××FrF^{\prime}=F^{\prime}_{1}\times\ldots\times F^{\prime}_{r^{\prime}} are two étale kk-algebras, and f:FFf\colon F\to F^{\prime} is a homomorphism of kk-algebras, it decomposes as a matrix (fij)i[r]j[r](f_{ij})_{\begin{subarray}{c}i\in[r]\\ j\in[r^{\prime}]\end{subarray}}, where fij:FiFjf_{ij}\colon F_{i}\to F_{j}^{\prime}, and each fijf_{ij} is either the zero map or an embedding of field extensions of kk (depending on whether the unit of FiF_{i} is mapped to 0 or to the unit of FjF_{j}). That is, if aFa\in F and j[r]j\in[r^{\prime}],

πj(f(a))=i[r]fij(πi(a)),\pi^{\prime}_{j}(f(a))=\sum_{i\in[r]}f_{ij}(\pi_{i}(a)),

where the πi\pi_{i} are the projections FFiF\to F_{i} and the πj\pi^{\prime}_{j} are the projections FFjF^{\prime}\to F^{\prime}_{j}.

We may then define a map 𝒟(f):𝒟(F)𝒟(F)\operatorname{\mathcal{D}}(f)\colon\operatorname{\mathcal{D}}(F)\to\operatorname{\mathcal{D}}(F^{\prime}) the following way: let i[r]i\in[r] and PPl(Fi)P\in\operatorname{Pl}(F_{i}), we set 𝒟(f)(P)=j[r]𝒟(fij)(P)\operatorname{\mathcal{D}}(f)(P)=\sum_{j\in[r^{\prime}]}\operatorname{\mathcal{D}}(f_{ij})(P), where 𝒟(fij)\operatorname{\mathcal{D}}(f_{ij}) is the usual map on divisors of global fields if fijf_{ij} is an embedding of fields and is zero if fij=0f_{ij}=0. One may check that this definition makes 𝒟\operatorname{\mathcal{D}} into a functor from the category of étale kk-algebras to the category of abelian groups. Furthermore, the map 𝒟:F×𝒟(F)\operatorname{\mathcal{D}}\colon F^{\times}\to\operatorname{\mathcal{D}}(F) is a natural transformation of the multiplicative group functor into the functor 𝒟\operatorname{\mathcal{D}} sending an étale algebra to its divisor group. That is, if aF×a\in F^{\times}, then 𝒟(f)(𝒟(a))=𝒟(f(a))\operatorname{\mathcal{D}}(f)(\operatorname{\mathcal{D}}(a))=\operatorname{\mathcal{D}}(f(a)).

As is the case for global fields, if QMFQ\in M_{F^{\prime}}, there is at most one place PPl(F)P\in\operatorname{Pl}(F) such that QSupp(𝒟(f)(P))Q\in\operatorname{Supp}(\operatorname{\mathcal{D}}(f)(P)). To prove this, we first need a lemma:

Lemma 4.3.

Let notations be as above and let j[r]j\in[r^{\prime}]. Then there is at most one i[r]i\in[r] such that fijf_{ij} is nonzero.

Proof.

Let i1,i2[r]i_{1},i_{2}\in[r] and j[r]j\in[r^{\prime}], and let e1,e2e_{1},e_{2} be the units respectively of Fi1F_{i_{1}} and Fi2F_{i_{2}}. We then have πj(f(e1))=fi1j(1)Fj\pi^{\prime}_{j}(f(e_{1}))=f_{i_{1}j}(1)\in F_{j} and πj(f(e2))=fi2j(1)Fj\pi^{\prime}_{j}(f(e_{2}))=f_{i_{2}j}(1)\in F_{j}. Now, since e1e2=0e_{1}e_{2}=0, we have fi1j(1)fi2j(1)=0f_{i_{1}j}(1)f_{i_{2}j}(1)=0. That is, either fi1j(1)=0f_{i_{1}j}(1)=0 or fi2j(1)=0f_{i_{2}j}(1)=0. It follows that either fi1jf_{i_{1}j} or fi2jf_{i_{2}j} is the zero map. ∎

Now, we prove the following:

Proposition 4.4.

Let notations be as above, and let QPl(F)Q\in\operatorname{Pl}(F^{\prime}). Then there is at most one place PPl(F)P\in\operatorname{Pl}(F) such that QSupp(𝒟(f)(P))Q\in\operatorname{Supp}(\operatorname{\mathcal{D}}(f)(P)).

Proof.

Let j[r]j\in[r^{\prime}] be such that QPl(Fj)Q\in\operatorname{Pl}(F_{j}). Then, by Lemma 4.3, there is at most one value i[r]i\in[r] such that fij0f_{ij}\neq 0. If there is no such ii, then no divisor of FF has an image by 𝒟(f)\operatorname{\mathcal{D}}(f) with QQ in its support. Otherwise, if QSupp(𝒟(f)(P))Q\in\operatorname{Supp}(\operatorname{\mathcal{D}}(f)(P)), it follows that PPl(Fi)P\in\operatorname{Pl}(F_{i}). Then, the usual theory of divisors of global fields ensures that there is only one PPl(Fi)P\in\operatorname{Pl}(F_{i}) such that QSupp(𝒟(f)(P))Q\in\operatorname{Supp}(\operatorname{\mathcal{D}}(f)(P)). ∎

If j[r]j\in[r^{\prime}] is such that fijf_{ij} is nonzero for some i[r]i\in[r], and if QPl(Fj)Q\in\operatorname{Pl}(F^{\prime}_{j}), we write QfQ_{f} for the unique PPl(F)P\in\operatorname{Pl}(F) such that QSupp(𝒟(f)(P))Q\in\operatorname{Supp}(\operatorname{\mathcal{D}}(f)(P)). We note that if ff is an inclusion of the form FFkFF\hookrightarrow F\otimes_{k}F^{\prime}, aa1a\mapsto a\otimes 1, then QfQ_{f} exists for all QPl(FkF)Q\in\operatorname{Pl}(F\otimes_{k}F^{\prime}).

Let QPl(F)Q\in\operatorname{Pl}(F^{\prime}) such that QfQ_{f} is well defined. We let eQ,fe_{Q,f} be the coefficient of QQ in the expansion of the divisor 𝒟(f)(Qf)\operatorname{\mathcal{D}}(f)(Q_{f}). Let PPl(F)P\in\operatorname{Pl}(F), if eQ,f1e_{Q,f}\neq 1 for some QSupp(𝒟(f)(P))Q\in\operatorname{Supp}(\operatorname{\mathcal{D}}(f)(P)), we say that the place PP ramifies through ff. This is equivalent to saying that, if i[r]i\in[r] is such that PPl(Fi)P\in\operatorname{Pl}(F_{i}), then for some j[r]j\in[r^{\prime}] such that fijf_{ij} is nonzero, the place PP ramifies in FjF_{j}.

In order to prove some facts about the behaviour of places in étale algebras, we will need to use local fields. For this purpose, we make the following definition:

Definition 4.5.

Let FF be as above. Let i[r]i\in[r] and let PPl(Fi)Pl(F)P\in\operatorname{Pl}(F_{i})\subset\operatorname{Pl}(F). Then the local completion FPF_{P} of FF at PP is the completion of the global field FiF_{i} at the place PP. This completion is an FF-algebra via the composite map

FπiFiFP.F\xrightarrow{\pi_{i}}F_{i}\rightarrow F_{P}.

As in the case of extensions of global fields, the places of FF^{\prime} lying above a place of FF may be classified as the direct factors of a tensor product.

Proposition 4.6.

Let notations be as above, let i[r]i\in[r] and PPl(Fi)Pl(F)P\in\operatorname{Pl}(F_{i})\subset\operatorname{Pl}(F). Let J={j[r]:fij0}J=\{j\in[r^{\prime}]\colon f_{ij}\neq 0\}. Then, we have

FPFFFPFijJFj.F_{P}\otimes_{F}F^{\prime}\simeq F_{P}\otimes_{F_{i}}\prod_{j\in J}F^{\prime}_{j}.

The usual bijections between the places of FjF^{\prime}_{j} above PP and the direct factors of FPFjF_{P}\otimes F^{\prime}_{j} give a bijection between the direct factors of FPFF_{P}\otimes F^{\prime} and Supp(𝒟(f)(P))\operatorname{Supp}(\operatorname{\mathcal{D}}(f)(P)). Furthermore, for QSupp(𝒟(f)(P)Q\in\operatorname{Supp}(\operatorname{\mathcal{D}}(f)(P), we have eQ,f1e_{Q,f}\neq 1 if and only if the corresponding factor is a ramified extension of FPF_{P}.

Proof.

We have

FPFijJFjFPFi(FiFF)FPFF.F_{P}\otimes_{F_{i}}\prod_{j\in J}F^{\prime}_{j}\simeq F_{P}\otimes_{F_{i}}(F_{i}\otimes_{F}F^{\prime})\simeq F_{P}\otimes_{F}F^{\prime}.

Then, the rest of the results follows directly from the definition of 𝒟(f)\operatorname{\mathcal{D}}(f) and the equivalent results for extensions of global fields (see e.g [11, Sections II.10 and II.19]). ∎

Finally, we also prove a result about ramification and tensor products:

Proposition 4.7.

Let FF and GG be étale kk-algebras, and let PPl(k)P\in\operatorname{Pl}(k) be a place that ramifies neither in FF nor in GG. Then, PP does not ramify in FkGF\otimes_{k}G either.

Proof.

Let kPk_{P} be the completion of kk at PP. By 4.6, it is enough to prove that the direct factors of kPkFkFk_{P}\otimes_{k}F\otimes_{k}F^{\prime} are unramified extensions of kPk_{P}. Now, since

kPkFkF(kPkF)kP(kPkF),k_{P}\otimes_{k}F\otimes_{k}F^{\prime}\simeq(k_{P}\otimes_{k}F)\otimes_{k_{P}}(k_{P}\otimes_{k}F^{\prime}),

the result will follow from the fact that if K,KK,K^{\prime} are unramified finite extensions of kPk_{P}, then the direct factors of KkPKK\otimes_{k_{P}}K^{\prime} are themselves unramified extensions of kPk_{P}.

This, in turns, follows from [11, Proposition I.7.1]. Indeed, by the proposition, we know that there are polynomials g,gR[X]g,g^{\prime}\in R[X] (where RR is the valuation ring of kPk_{P}) such that KkP[X]/(g(X))K\simeq k_{P}[X]/(g(X)), KkP[X]/g(X)K^{\prime}\simeq k_{P}[X]/g^{\prime}(X) and whose residues g¯\overline{g} and g¯\overline{g^{\prime}} are irreducible and separable in κP[X]\kappa_{P}[X], where κP\kappa_{P} is the residue field of kPk_{P}. Now, we have KkPKK[X]/(g(X))K\otimes_{k_{P}}K^{\prime}\simeq K[X]/(g^{\prime}(X)). A direct factor of this algebra is of the form K[X]/(h(X))K[X]/(h(X)), where hh is an irreducible factor of gg^{\prime} in 𝒪[X]\mathcal{O}[X], where 𝒪\mathcal{O} is the valuation ring of KK. Since h¯\overline{h} is a factor of the separable polynomial g¯\overline{g^{\prime}}, it is itself separable as a polynomial in 𝔨[X]\mathfrak{k}[X], where 𝔨\mathfrak{k} is the residue field of KK. Since the polynomial hh is irreducible in 𝒪[X]\mathcal{O}[X] and its residue is separable, this residue h¯\overline{h} is irreducible in 𝔨[X]\mathfrak{k}[X] by Hensel’s lemma. By the proposition cited above, the field extension K[X]/(h(X))K[X]/(h(X)) is unramified over KK, and therefore over kPk_{P}. It follows that the direct factors of KkPKK\otimes_{k_{P}}K^{\prime} are unramified over kPk_{P}. ∎

4.2 An algorithm for trivialising Amitsur coboundaries over a number field

Let kk be a number field, and let FF be an étale kk-algebra. The Amitsur complex for FF yields a complex

𝒟(Fn+1)𝒟(ΔAmn)𝒟(Fn+2).\ldots\to\operatorname{\mathcal{D}}(F^{\otimes n+1})\xrightarrow{\operatorname{\mathcal{D}}(\Delta_{Am}^{n})}\operatorname{\mathcal{D}}(F^{\otimes n+2})\to\ldots.

of abelian groups, where we define 𝒟(ΔAmn)\operatorname{\mathcal{D}}(\Delta_{Am}^{n}) as 0in(1)i𝒟(εin)\sum_{0\leq i\leq n}(-1)^{i}\operatorname{\mathcal{D}}(\varepsilon^{n}_{i}).

We give a precise description of the map 𝒟(ΔAmn)\operatorname{\mathcal{D}}(\Delta_{Am}^{n}). Let QPl(Fn+2)Q\in\operatorname{Pl}(F^{\otimes n+2}). Then, for any 0in+10\leq i\leq n+1, we set Qi=QεinQ_{i}=Q_{\varepsilon_{i}^{n}} and eQ,i=eQ,εine_{Q,i}=e_{Q,\varepsilon_{i}^{n}}. Then, if PPl(Fn+1)P\in\operatorname{Pl}(F^{\otimes n+1}), we get

𝒟(εin)(P)=QPl(Fn+2)Qi=PeQ,iQ,\operatorname{\mathcal{D}}(\varepsilon_{i}^{n})(P)=\sum_{\begin{subarray}{c}Q\in\operatorname{Pl}(F^{\otimes n+2})\\ Q_{i}=P\end{subarray}}e_{Q,i}Q,

and it follows that

𝒟(εin)(PPl(Fn+1)nPP)=QPl(Fn+2)eQ,inQiQ\operatorname{\mathcal{D}}(\varepsilon_{i}^{n})\left(\sum_{P\in\operatorname{Pl}(F^{\otimes n+1})}n_{P}P\right)=\sum_{Q\in\operatorname{Pl}(F^{\otimes n+2})}e_{Q,i}n_{Q_{i}}Q

and

𝒟(ΔAmn)(PPl(Fn+1)nPP)=QPl(Fn+2)(0in+1(1)ieQ,inQi)Q.\operatorname{\mathcal{D}}(\Delta_{Am}^{n})\left(\sum_{P\in\operatorname{Pl}(F^{\otimes n+1})}n_{P}P\right)=\sum_{Q\in\operatorname{Pl}(F^{\otimes n+2})}\left(\sum_{0\leq i\leq{n+1}}(-1)^{i}e_{Q,i}n_{Q_{i}}\right)Q.

We first need two lemmas:

Lemma 4.8.

Let Q,QQ,Q^{\prime} be places of F2F^{\otimes 2} such that Q0=Q0=PQ_{0}=Q^{\prime}_{0}=P. Then there exists a place RPl(F3)R\in\operatorname{Pl}(F^{\otimes 3}) such that R1=QR_{1}=Q and R0=QR_{0}=Q^{\prime}.

Proof.

We let α\alpha be a defining polynomial of FF. That is, Fk[X]/(α(X))F\simeq k[X]/(\alpha(X)), with αk[X]\alpha\in k[X] separable. We may then identify F2F^{\otimes 2} with F[X]/(α(X))F[X]/(\alpha(X)), where ε00\varepsilon_{0}^{0} maps FF into the ring of scalars in F[X]/(α(X))F[X]/(\alpha(X)) and ε10\varepsilon_{1}^{0} is the map

k[X]/(α(X))F[X]/(α(X))XX.\begin{array}[]{ccc}k[X]/(\alpha(X))&\to&F[X]/(\alpha(X))\\ X&\mapsto&X.\end{array}

Likewise, we identify F3F^{\otimes 3} with F[X,Y]/(α(X),α(Y))F[X,Y]/(\alpha(X),\alpha(Y)). Then, the descriptions of the maps ε01\varepsilon_{0}^{1} and ε11\varepsilon_{1}^{1} become:

ε01:F[X]/(α(X))F[X,Y]/(α(X),α(Y))XYε11:F[X]/(α(X))F[X,Y]/(α(X),α(Y))XX\begin{array}[]{rccc}\varepsilon_{0}^{1}\colon&F[X]/(\alpha(X))&\to&F[X,Y]/(\alpha(X),\alpha(Y))\\ &X&\mapsto&Y\\ \varepsilon_{1}^{1}\colon&F[X]/(\alpha(X))&\to&F[X,Y]/(\alpha(X),\alpha(Y))\\ &X&\mapsto&X\end{array}

These identifications are coherent with the maps of the Amitsur complex. In particular, we note that ε01ε00=ε11ε00\varepsilon_{0}^{1}\circ\varepsilon_{0}^{0}=\varepsilon_{1}^{1}\circ\varepsilon_{0}^{0}. With this in place, let PPl(F)P\in\operatorname{Pl}(F). By 4.6, the support of 𝒟(ε00)(P)\operatorname{\mathcal{D}}(\varepsilon_{0}^{0})(P) is in bijection with the direct factors of FP[X]/(α(X))F_{P}[X]/(\alpha(X)). That is, it is in bijection with the irreducible factors of α(X)\alpha(X) in FP[X]F_{P}[X]. Likewise, the support of 𝒟(ε01ε00)(P)\operatorname{\mathcal{D}}(\varepsilon_{0}^{1}\circ\varepsilon_{0}^{0})(P) is in bijection with the irreducible factors of FP[X,Y]/(α(X),α(Y))F_{P}[X,Y]/(\alpha(X),\alpha(Y)). That is, it is in bijection with the maximal ideals that contain the ideal (α(X),α(Y))(\alpha(X),\alpha(Y)).

Now, if QSupp(𝒟(ε00)(P))Q\in\operatorname{Supp}(\operatorname{\mathcal{D}}(\varepsilon_{0}^{0})(P)), we let αQ(X)\alpha_{Q}(X) be the irreducible factor of α(X)\alpha(X) corresponding to QQ. Likewise, if RSupp(𝒟(ε01ε00)(P)R\in\operatorname{Supp}(\operatorname{\mathcal{D}}(\varepsilon_{0}^{1}\circ\varepsilon_{0}^{0})(P), we let 𝔪R\mathfrak{m}_{R} be the maximal ideal of FP[X,Y]F_{P}[X,Y] corresponding to RR. Observe that RSupp(ε01)(Q)R\in\operatorname{Supp}(\varepsilon_{0}^{1})(Q) if and only if αQ(Y)𝔪R\alpha_{Q}(Y)\in\mathfrak{m}_{R} and RSupp(ε11)(Q)R\in\operatorname{Supp}(\varepsilon_{1}^{1})(Q) if and only if αQ(X)𝔪R\alpha_{Q}(X)\in\mathfrak{m}_{R}. Then, the result simply follows from the observation that the ideal (αQ(X),αQ(Y))(\alpha_{Q^{\prime}}(X),\alpha_{Q}(Y)) contains (α(X),α(Y))(\alpha(X),\alpha(Y)) and is contained in some maximal ideal of FP[X,Y]F_{P}[X,Y]. Such a maximal ideal then corresponds to a place RMF3R\in M_{F^{\otimes 3}} such that R1=QR_{1}=Q and R0=QR_{0}=Q^{\prime}. ∎

For n0n\geq 0 and SPl(k)S\subset\operatorname{Pl}(k), we let S(n)S^{(n)} be the subset of Pl(Fn+1)\operatorname{Pl}(F^{\otimes n+1}) of places lying above some PSP\in S. We let SrS_{r} be the set of places of kk that ramify in FF. Applying 4.7 by induction, we get the following:

Lemma 4.9.

Let QPl(Fn+1)Sr(n)Q\in\operatorname{Pl}(F^{\otimes n+1})\setminus S_{r}^{(n)}. Then for 0in+10\leq i\leq n+1, we have eQ,i=1e_{Q,i}=1.

We may now prove the following generalisation of Hilbert’s Theorem 90:

Lemma 4.10.

Let D=QPl(F2)nQQKer𝒟(ΔAm1)D=\sum_{Q\in\operatorname{Pl}(F^{\otimes 2})}n_{Q}Q\in\operatorname{Ker}\operatorname{\mathcal{D}}(\Delta_{Am}^{1}) be supported by places outside of Sr(1)S_{r}^{(1)}. Then, there exists E𝒟(F)E\in\operatorname{\mathcal{D}}(F) such that D=𝒟(ΔAm0)(E)D=\operatorname{\mathcal{D}}(\Delta_{Am}^{0})(E).

Proof.

We set

E=PPl(F)(minQPl(F2)Q0=PnQ)P.E=\sum_{P\in\operatorname{Pl}(F)}\left(\min_{\begin{subarray}{c}Q\in\operatorname{Pl}(F^{\otimes 2})\\ Q_{0}=P\end{subarray}}n_{Q}\right)P.

Then, by Lemma 4.9, we get

𝒟(ε00)(E)=QPl(F2)(minQ′′Pl(F2)Q0′′=Q0nQ′′)Q\operatorname{\mathcal{D}}(\varepsilon_{0}^{0})(E)=\sum_{Q\in\operatorname{Pl}(F^{\otimes 2})}\left(\min_{\begin{subarray}{c}Q^{\prime\prime}\in\operatorname{Pl}(F^{\otimes 2})\\ Q^{\prime\prime}_{0}=Q_{0}\end{subarray}}n_{Q^{\prime\prime}}\right)Q

and

𝒟(ε10)(E)=QPl(F2)(minQPl(F2)Q0=Q1nQ)Q\operatorname{\mathcal{D}}(\varepsilon_{1}^{0})(E)=\sum_{Q\in\operatorname{Pl}(F^{\otimes 2})}\left(\min_{\begin{subarray}{c}Q^{\prime}\in\operatorname{Pl}(F^{\otimes 2})\\ Q^{\prime}_{0}=Q_{1}\end{subarray}}n_{Q^{\prime}}\right)Q

It follows that

D+𝒟(ε10)(E)=QPl(F2)(minQPl(F2)Q0=Q1nQ+nQ)Q.D+\operatorname{\mathcal{D}}(\varepsilon_{1}^{0})(E)=\sum_{Q\in\operatorname{Pl}(F^{\otimes 2})}\left(\min_{\begin{subarray}{c}Q^{\prime}\in\operatorname{Pl}(F^{\otimes 2})\\ Q^{\prime}_{0}=Q_{1}\end{subarray}}n_{Q}+n_{Q^{\prime}}\right)Q.

We introduce the following automorphisms:

σ:F2F2abba;τ:F3F3abcacb.\begin{array}[]{rccc}\sigma\colon&F^{\otimes 2}&\to&F^{\otimes 2}\\ &a\otimes b&\mapsto&b\otimes a;\\ \tau\colon&F^{\otimes 3}&\to&F^{\otimes 3}\\ &a\otimes b\otimes c&\mapsto&a\otimes c\otimes b.\end{array}

Observe that we have the following equalities:

τε01\displaystyle\tau\circ\varepsilon_{0}^{1} =ε01σ\displaystyle=\varepsilon_{0}^{1}\circ\sigma (3)
τε21\displaystyle\tau\circ\varepsilon_{2}^{1} =ε11\displaystyle=\varepsilon_{1}^{1} (4)
ε10\displaystyle\varepsilon_{1}^{0} =σε00.\displaystyle=\sigma\circ\varepsilon_{0}^{0}. (5)

If QPl(F2)Q\in\operatorname{Pl}(F^{\otimes 2}) (resp. RPl(F3)R\in\operatorname{Pl}(F^{\otimes 3})), we write QσQ^{\sigma} (resp. RτR^{\tau}) for the unique place in the support of 𝒟(σ)(Q)\operatorname{\mathcal{D}}(\sigma)(Q) (resp. 𝒟(τ)(R)\operatorname{\mathcal{D}}(\tau)(R)).

Now, let Q,QPl(F2)Q,Q^{\prime}\in\operatorname{Pl}(F^{\otimes 2}) such that Q0=Q1Q^{\prime}_{0}=Q_{1}. By Equation 5, we have Q0=Q0σQ^{\prime}_{0}=Q_{0}^{\sigma}. Applying Lemma 4.8 to QσQ^{\sigma} and QQ^{\prime}, we get RPl(F3)R\in\operatorname{Pl}(F^{\otimes 3}) such that R1=QR_{1}=Q^{\prime} and R0=QσR_{0}=Q^{\sigma}. By Equation 4, we have R2τ=R1=QR^{\tau}_{2}=R_{1}=Q^{\prime} and by Equation 3, we get R0τ=(R0)σ=QR^{\tau}_{0}=(R_{0})^{\sigma}=Q. We set Q′′=R1τQ^{\prime\prime}=R_{1}^{\tau}. Now, the coefficient of RτR^{\tau} in 𝒟(ΔAm1)(D)\operatorname{\mathcal{D}}(\Delta^{1}_{Am})(D) is nQ+nQnQ′′n_{Q}+n_{Q^{\prime}}-n_{Q^{\prime\prime}}, and since this divisor is zero by hypothesis, we get nQ′′=nQ+nQn_{Q^{\prime\prime}}=n_{Q}+n_{Q^{\prime}} (note that RSr(2)R\notin S_{r}^{(2)}, so eR,i=1e_{R,i}=1 for 0i20\leq i\leq 2). Observe that we have

R\displaystyle R Supp(𝒟(ε01)(Q))\displaystyle\in\operatorname{Supp}(\operatorname{\mathcal{D}}(\varepsilon_{0}^{1})(Q))
Supp(𝒟(ε01ε00)(Q0))\displaystyle\subset\operatorname{Supp}(\operatorname{\mathcal{D}}(\varepsilon_{0}^{1}\circ\varepsilon_{0}^{0})(Q_{0}))
=Supp(𝒟(ε11ε00)(Q0))\displaystyle=\operatorname{Supp}(\operatorname{\mathcal{D}}(\varepsilon_{1}^{1}\circ\varepsilon_{0}^{0})(Q_{0}))

and

RSupp(𝒟(ε11)(Q′′))Supp(𝒟(ε11ε00)(Q0′′)).R\in\operatorname{Supp}(\operatorname{\mathcal{D}}(\varepsilon_{1}^{1})(Q^{\prime\prime}))\subset\operatorname{Supp}(\operatorname{\mathcal{D}}(\varepsilon_{1}^{1}\circ\varepsilon_{0}^{0})(Q^{\prime\prime}_{0})).

By 4.4, it follows that Q0=Q0′′Q_{0}=Q^{\prime\prime}_{0}. That is, we proved that there exists Q′′Pl(F2)Q^{\prime\prime}\in\operatorname{Pl}(F^{\otimes 2}) such that Q0′′=Q0Q^{\prime\prime}_{0}=Q_{0} and nQ+nQ=nQ′′n_{Q}+n_{Q^{\prime}}=n_{Q^{\prime\prime}}.

Conversely, let Q,Q′′Pl(F2)Q,Q^{\prime\prime}\in\operatorname{Pl}(F^{\otimes 2}) such that Q0=Q0′′Q_{0}=Q^{\prime\prime}_{0}. Then by Lemma 4.8, there is RPl(F3)R\in\operatorname{Pl}(F^{\otimes 3}) such that R0=QR_{0}=Q and R1=Q′′R_{1}=Q^{\prime\prime}. Set Q=R2Q^{\prime}=R_{2} and, as above, we get nQ+nQ=nQ′′n_{Q}+n_{Q^{\prime}}=n_{Q^{\prime\prime}}. As above, we use the fact that ε01ε10=ε21ε00\varepsilon_{0}^{1}\circ\varepsilon_{1}^{0}=\varepsilon_{2}^{1}\circ\varepsilon_{0}^{0} to prove that Q1=Q0Q_{1}=Q^{\prime}_{0}.

Putting things together, we have proved that for QPl(F2)Sr(1)Q\in\operatorname{Pl}(F^{\otimes 2})\setminus S_{r}^{(1)},

minQPl(F2)Q0=Q1nQ+nQ=minQ′′Pl(F2)Q0′′=Q0nQ′′.\min_{\begin{subarray}{c}Q^{\prime}\in\operatorname{Pl}(F^{\otimes 2})\\ Q^{\prime}_{0}=Q_{1}\end{subarray}}n_{Q}+n_{Q^{\prime}}=\min_{\begin{subarray}{c}Q^{\prime\prime}\in\operatorname{Pl}(F^{\otimes 2})\\ Q^{\prime\prime}_{0}=Q_{0}\end{subarray}}n_{Q^{\prime\prime}}.

It follows directly that D+𝒟(ε10)(E)=𝒟(ε00)(E)D+\operatorname{\mathcal{D}}(\varepsilon_{1}^{0})(E)=\operatorname{\mathcal{D}}(\varepsilon_{0}^{0})(E). ∎

We now get our main theorem for this section:

Theorem 4.11.

Let bBAm2(k,F)b\in B_{Am}^{2}(k,F) be a coboundary. Let SS be a set of non-archimedean places of kk such that:

  • SrSS_{r}\subset S. That is, SS contains the places of kk that ramify in FF.

  • The finite places of S(0)S^{(0)} generate the class group Cl(F)\operatorname{Cl}(F).

  • Supp(𝒟(b))S(2)\operatorname{Supp}(\operatorname{\mathcal{D}}(b))\subset S^{(2)}.

Then there exists a cochain aa in the group of S(2)S^{(2)}-units of F2F^{\otimes 2} such that b=ΔAm1(a)b=\Delta_{Am}^{1}(a)

Proof.

Let α(F2)×\alpha\in(F^{\otimes 2})^{\times} be such that ΔAm1(α)=b\Delta_{Am}^{1}(\alpha)=b. We consider the divisor D=𝒟(α)=QPl(F2)nQQD=\operatorname{\mathcal{D}}(\alpha)=\sum_{Q\in\operatorname{Pl}(F^{\otimes 2})}n_{Q}Q of α\alpha. We set DS=QS(1)nQQD_{S}=\sum_{Q\in S^{(1)}}n_{Q}Q and DS=QS(1)nQQD^{\prime}_{S}=\sum_{Q\notin S^{(1)}}n_{Q}Q. Now we get 𝒟(ΔAm1)(D)=𝒟(b)\operatorname{\mathcal{D}}(\Delta_{Am}^{1})(D)=\operatorname{\mathcal{D}}(b) and therefore it is supported by S(2)S^{(2)}. Observe that if QS(1)Q\in S^{(1)}, then Supp(𝒟(ΔAm1)(Q))S(2)\operatorname{Supp}(\operatorname{\mathcal{D}}(\Delta_{Am}^{1})(Q))\subset S^{(2)} and conversely, if Supp(𝒟(ΔAm1)(Q))S(2)\operatorname{Supp}(\operatorname{\mathcal{D}}(\Delta_{Am}^{1})(Q))\subset S^{(2)} and 𝒟(ΔAm1)1(Q)0\operatorname{\mathcal{D}}(\Delta_{Am}^{1})^{1}(Q)\neq 0, then QS(1)Q\in S^{(1)}. It follows that 𝒟(ΔAm1)(DS)=𝒟(ΔAm1)(D)𝒟(ΔAm1)(DS)=0\operatorname{\mathcal{D}}(\Delta_{Am}^{1})(D^{\prime}_{S})=\operatorname{\mathcal{D}}(\Delta_{Am}^{1})(D)-\operatorname{\mathcal{D}}(\Delta_{Am}^{1})(D_{S})=0.

The support of DSD^{\prime}_{S} is disjoint from Sr(1)S_{r}^{(1)}. We may therefore apply Lemma 4.10 and get a divisor E𝒟(F)E\in\operatorname{\mathcal{D}}(F) such that DS=ΔAm0(E)D^{\prime}_{S}=\Delta_{Am}^{0}(E). Now, as S(0)S^{(0)} generates the class group of FF, there exists E𝒟(F)E^{\prime}\in\operatorname{\mathcal{D}}(F) with support in S(0)S^{(0)} and γF×\gamma\in F^{\times} such that E=𝒟(γ)+EE=\operatorname{\mathcal{D}}(\gamma)+E^{\prime}. Then, we get that

𝒟(ΔAm0)(𝒟(γ))+𝒟(ΔAm0)(E)=DDS\operatorname{\mathcal{D}}(\Delta_{Am}^{0})(\operatorname{\mathcal{D}}(\gamma))+\operatorname{\mathcal{D}}(\Delta_{Am}^{0})(E^{\prime})=D-D_{S}

and therefore

𝒟(ΔAm0)(E)+DS=𝒟(αΔAm0(γ1)).\operatorname{\mathcal{D}}(\Delta_{Am}^{0})(E^{\prime})+D_{S}=\operatorname{\mathcal{D}}(\alpha\Delta_{Am}^{0}(\gamma^{-1})).

Now, since 𝒟(ΔAm0)(E)+DS\operatorname{\mathcal{D}}(\Delta_{Am}^{0})(E^{\prime})+D_{S} has support in S(1)S^{(1)}, this shows that αΔAm0(γ1)\alpha\Delta_{Am}^{0}(\gamma^{-1}) is a S(1)S^{(1)}-unit. Furthermore,

ΔAm1(αΔAm0(γ1))=ΔAm1(α)=b,\Delta_{Am}^{1}(\alpha\Delta_{Am}^{0}(\gamma^{-1}))=\Delta_{Am}^{1}(\alpha)=b,

and αΔAm0(γ1)\alpha\Delta_{Am}^{0}(\gamma^{-1}) is a cochain with the required properties. ∎

From Theorem 4.11 we directly get an algorithm for computing a trivialisation of a 22-coboundary:

Input: A number field kk
Input: An étale kk-algebra FF defined by a separable polynomial Pk[X]P\in k[X]
Input: A coboundary bBAm2(k,F)b\in B_{Am}^{2}(k,F)
Output: A cochain aCAm0(k,F)a\in C_{Am}^{0}(k,F) such that ΔAm1(a)=c\Delta_{Am}^{1}(a)=c
1 Compute S1S_{1}, the set of places of kk that ramify in FF;
2 Compute S2S_{2}, a set of places of kk such that S2(0)S_{2}^{(0)} generates the class group Cl(F)\operatorname{Cl}(F);
3 Compute 𝒟(b)\operatorname{\mathcal{D}}(b). Let S3S_{3} be the set of places of kk below the places in the support of 𝒟(b)\operatorname{\mathcal{D}}(b);
4 Set S=S1S2S3S=S_{1}\cup S_{2}\cup S_{3};
5 Compute the sets S(2)S^{(2)} and S(3)S^{(3)};
6 Compute an isomorphism ϕ\phi from the group of S(2)S^{(2)}-units of F2F^{\otimes 2} to r/m\mathbb{Z}^{r}\oplus\mathbb{Z}/m\mathbb{Z};
7 Compute an isomorphism ψ\psi from the group of S(3)S^{(3)}-units of F3F^{\otimes 3} to r/m\mathbb{Z}^{r^{\prime}}\oplus\mathbb{Z}/m^{\prime}\mathbb{Z};
8 Solve the linear equation (ψΔAm1ϕ1)(a)=ψ(b)(\psi\circ\Delta_{Am}^{1}\circ\phi^{-1})(a)=\psi(b);
return aa
Algorithm 2 Computing a trivialisation of a 22-coboundary
Theorem 4.12.

Given a number field kk, a separable polynomial Pk[X]P\in k[X] and a coboundary bBAm2(k,F)b\in B^{2}_{Am}(k,F), where F=k[X]/PF=k[X]/P, Algorithm 2 outputs the representation of a cochain αC1(k,F)\alpha\in C^{1}(k,F) such that ΔAm1(α)=b\Delta_{Am}^{1}(\alpha)=b. Furthermore, assuming GRH, Algorithm 2 is a polynomial quantum algorithm.

Proof.

Using a polynomial-time algorithm for factoring polynomials over number fields [41], one may compute splittings of FF, F2F^{\otimes 2} and F3F^{\otimes 3} as direct products of number fields. Set Fn+1=F1(n)××Frn(n)F^{\otimes n+1}=F^{(n)}_{1}\times\ldots\times F_{r_{n}}^{(n)}.

The set S1S_{1} may be computed by factoring the discriminants of the extensions Fi(0)/kF^{(0)}_{i}/k, which may be done in polynomial time by 2.1. Then, by 2.3, one may compute in polynomial time a set of polynomial size of places of FF that generate the class group Cl(F)\operatorname{Cl}(F), and the set S2S_{2} of places of kk lying below them. In practice, one may set S2S_{2} to be the set of prime ideals 𝔭\mathfrak{p} of kk such that N(𝔭)12log2(maxi[r]|ΔFi|)N(\mathfrak{p})\leq 12\log^{2}(\max_{i\in[r]}|\Delta_{F_{i}}|). We may then take the image (b1,,br2)(b_{1},\ldots,b_{r_{2}}) of bb in the product F1(2)××Fr2(2)F^{(2)}_{1}\times\ldots\times F^{(2)}_{r_{2}} and factor the principal ideals (bi)(b_{i}) in Fi(2)F^{(2)}_{i} into a product 𝔭i,1𝔭i,si\mathfrak{p}_{i,1}\ldots\mathfrak{p}_{i,s_{i}}, which may be done in polynomial time by 2.2. We may then set S3={𝔭i,jk,i[r2],j[si]}Pl(k)S_{3}=\{\mathfrak{p}_{i,j}\cap k,i\in[r_{2}],j\in[s_{i}]\}\subset\operatorname{Pl}(k).

Isomorphisms ϕ\phi and ψ\psi are computed using 2.5 and [42, Algorithm 8.4]. Finally, the last step is the computation of a solution of a system of linear equations over \mathbb{Z}.

The correctness of the algorithm relies on the fact that a cochain aa such that b=ΔAm1(a)b=\Delta_{Am}^{1}(a) exists and may be found in the group of S(1)S^{(1)}-units of F2F^{\otimes 2}, which is the content of Theorem 4.11. ∎

Theorem 4.13.

Assuming GRH, there exists a polynomial quantum algorithm which, give a number field kk and an algebra AMn(k)A\simeq M_{n}(k), computes an explicit algorithm from AA to Mn(k)M_{n}(k).

Proof.

This is simply a combination of Theorems 3.10 and 4.12. Indeed, using Algorithm 1, one may compute an étale kk algebra F=k[X]/PF=k[X]/P, a cocycle c2(k,F)c\in\mathbb{Z}^{2}(k,F) and an isomorphism AA(F,c)A\simeq A(F,c). Since AA is isomorphic to Mn(k)M_{n}(k), the cocycle cc is in fact a coboundary. Then, a trivialisation of cc may be computed using Algorithm 2. Applying 3.7, we obtain an explicit isomorphism A(F,c)A(F,𝟏)A(F,c)\simeq A(F,\mathbf{1}). Finally, an isomorphism A(F,𝟏)Mn(k)A(F,\mathbf{1})\simeq M_{n}(k) may easily be computed using Remark 3.5. ∎

References

  • [1] Adamson, I. T. Cohomology theory for non-normal subgroups and non-normal fields. Glasg. Math. J. 2, 2 (1954), 66–76.
  • [2] Amitsur, S. A. Simple algebras and cohomology groups of arbitrary fields. Trans. Amer. Math. Soc. 90 (1959), 73–112.
  • [3] Auslander, M., and Goldman, O. The brauer group of a commutative ring. Trans. Amer. Math. Soc. 97, 3 (1960), 367–409.
  • [4] Azumaya, G. On maximally central algebras. Nagoya Math. J. 2 (1951), 119–150.
  • [5] Bach, E. Explicit bounds for primality testing and related problems. Math. Comp. 55, 191 (1990), 355–380.
  • [6] Bauch, J., Bernstein, D. J., de Valence, H., Lange, T., and Van Vredendaal, C. Short generators without quantum computers: the case of multiquadratics. In Advances in Cryptology–EUROCRYPT 2017: 36th Annual International Conference on the Theory and Applications of Cryptographic Techniques, Paris, France, April 30–May 4, 2017, Proceedings, Part I (2017), Springer, pp. 27–59.
  • [7] Biasse, J.-F. Subexponential time relations in the class group of large degree number fields. Adv. Math. Commun. 8, 4 (2014), 407–425.
  • [8] Biasse, J.-F., Espitau, T., Fouque, P.-A., Gélin, A., and Kirchner, P. Computing generator in cyclotomic integer rings. In Advances in Cryptology–EUROCRYPT 2017: 36th Annual International Conference on the Theory and Applications of Cryptographic Techniques, Paris, France, April 30–May 4, 2017, Proceedings, Part I (Cham, 2017), Springer, pp. 60–88.
  • [9] Biasse, J.-F., and Song, F. Efficient quantum algorithms for computing class groups and solving the principal ideal problem in arbitrary degree number fields. In Proceedings of the Twenty-Seventh Annual ACM-SIAM Symposium on Discrete Algorithms (USA, 2016), SODA ’16, Society for Industrial and Applied Mathematics, p. 893–902.
  • [10] Bourbaki, N. Algebra I. Chapters 1–3. Elements of Mathematics (Berlin). Springer-Verlag, Berlin, 1998.
  • [11] Cassels, J. W. S., and Fröhlich, A., Eds. Algebraic number theory (1967), Academic Press, London; Thompson Book Co., Inc., Washington, DC.
  • [12] Chase, S. U., and Rosenberg, A. Amitsur cohomology and the brauer group. In Galois theory and cohomology of commutative rings, vol. 1 of Memoirs of the Americam Mathematical Society. American Mathematical Society, 1965, pp. 20–65.
  • [13] Cohen, H. A course in computational algebraic number theory, vol. 138 of Graduate Texts in Mathematics. Springer-Verlag, Berlin, 1993.
  • [14] Cohen, H. Advanced topics in computational number theory, vol. 193. Springer, New-York, NY, 2012.
  • [15] Colliot-Thélène, J.-L., and Skorobogatov, A. N. The Brauer-Grothendieck Group. Springer Cham, 2021.
  • [16] Cremona, J., Fisher, T., O’Neil, C., Simon, D., and Stoll, M. Explicit n-descent on elliptic curves, i. algebra. J. Reine Angew. Math. 2008, 615 (2008), 121–155.
  • [17] Cremona, J., Fisher, T., O’Neil, C., Simon, D., and Stoll, M. Explicit n-descent on elliptic curves, ii. geometry. J. Reine Angew. Math. 2009, 632 (2009), 63–84.
  • [18] Cremona, J., Fisher, T., O’Neil, C., Simon, D., and Stoll, M. Explicit n-descent on elliptic curves iii. algorithms. Math. Comp. 84, 292 (2015), 895–922.
  • [19] Cremona, J., and Rusin, D. Efficient solution of rational conics. Math. Comp. 72, 243 (2003), 1417–1441.
  • [20] Csahók, T., Kutas, P., Montessinos, M., and Zábrádi, G. Explicit isomorphisms of quaternion algebras over quadratic global fields. Research in Number Theory 8, 4 (2022), 77.
  • [21] De Graaf, W. A., Harrison, M., Pílniková, J., and Schicho, J. A lie algebra method for rational parametrization of severi–brauer surfaces. J. Algebra 303, 2 (2006), 514–529.
  • [22] de Graaf, W. A., and Ivanyos, G. Finding splitting elements and maximal tori in matrix algebras. In Interactions Between Ring Theory and Representations of Algebras, F. van Oystaeyen and M. Saorin, Eds. CRC Press, 2000, ch. 7, pp. 95–105.
  • [23] Eberhard, S. The characteristic polynomial of a random matrix. Combinatorica 42, 4 (2022), 491–527.
  • [24] Eisenträger, K., Hallgren, S., Kitaev, A., and Song, F. A quantum algorithm for computing the unit group of an arbitrary degree number field. In Proceedings of the Forty-Sixth Annual ACM Symposium on Theory of Computing (New York, NY, USA, 2014), STOC ’14, Association for Computing Machinery, p. 293–302.
  • [25] Fieker, C. Minimizing representations over number fields ii: Computations in the brauer group. J. Algebra 322, 3 (2009), 752–765.
  • [26] Fisher, T. Explicit 5-descent on elliptic curves. Open Book Ser. 1, 1 (2013), 395–411.
  • [27] Fisher, T. Higher descents on an elliptic curve with a rational 2-torsion point. Math. Comp. 86, 307 (2017), 2493–2518.
  • [28] Fisher, T., and Newton, R. Computing the cassels-tate pairing on the 3-selmer group of an elliptic curve. Int. J. Number Theory 10, 07 (2014), 1881–1907.
  • [29] Ford, T. J. Separable Algebras, vol. 183. American Mathematical Society, 2017.
  • [30] Garg, A., Gupta, N., Kayal, N., and Saha, C. Determinant Equivalence Test over Finite Fields and over Q. In 46th International Colloquium on Automata, Languages, and Programming (ICALP 2019) (Dagstuhl, Germany, 2019), C. Baier, I. Chatzigiannakis, P. Flocchini, and S. Leonardi, Eds., vol. 132 of Leibniz International Proceedings in Informatics (LIPIcs), Schloss Dagstuhl–Leibniz-Zentrum fuer Informatik, pp. 62:1–62:15.
  • [31] Gille, P., and Szamuely, T. Central simple algebras and Galois cohomology, vol. 165. Cambridge University Press, Cambridge, 2017.
  • [32] Gómez-Torrecillas, J., Kutas, P., Lobillo, F. J., and Navarro, G. Primitive idempotents in central simple algebras over fq (t) with an application to coding theory. Finite Fields Appl. 77 (2022), 101935.
  • [33] Hallgren, S. Fast quantum algorithms for computing the unit group and class group of a number field. In Proceedings of the Thirty-Seventh Annual ACM Symposium on Theory of Computing (New York, NY, 2005), STOC ’05, Association for Computing Machinery, p. 468–474.
  • [34] Ivanyos, G., Kutas, P., and Rónyai, L. Computing explicit isomorphisms with full matrix algebras over 𝔽q(x)\mathbb{F}_{q}(x). Found. Comput. Math. 18 (2018), 381–397.
  • [35] Ivanyos, G., Kutas, P., and Rónyai, L. Explicit equivalence of quadratic forms over 𝔽q(t)\mathbb{F}_{q}(t). Finite Fields Appl. 55 (2019), 33–63.
  • [36] Ivanyos, G., Lelkes, Á., and Rónyai, L. Improved algorithms for splitting full matrix algebras. JP J Algebra, Number Theory Appl. 28, 2 (2013), 141–156.
  • [37] Ivanyos, G., Rónyai, L., and Schicho, J. Splitting full matrix algebras over algebraic number fields. J. Algebra 354 (2012), 211–223.
  • [38] Jacobson, N. Finite-dimensional division algebras over fields. Springer-Verlag, Berlin, 1996.
  • [39] Kutas, P. Splitting quaternion algebras over quadratic number fields. J. Symbolic Comput. 94 (2019), 173–182.
  • [40] Kutas, P., Montessinos, M., Zábrádi, G., and Csahók, T. Finding nontrivial zeros of quadratic forms over rational function fields of characteristic 2. In Proceedings of the 2022 International Symposium on Symbolic and Algebraic Computation (New York, NY, USA, 2022), ISSAC ’22, Association for Computing Machinery, p. 235–244.
  • [41] Lenstra, A. K. Factoring polynomials over algebraic number fields. In Computer Algebra: EUROCAL’83, European Computer Algebra Conference London, England, March 28–30, 1983 Proceedings (Berlin, Heidelberg, 1983), Springer, pp. 245–254.
  • [42] Lenstra, Jr., H. W., and Silverberg, A. Algorithms for commutative algebras over the rational numbers. Found. Comput. Math. 18, 1 (2018), 159–180.
  • [43] Pílniková, J. Trivializing a central simple algebra of degree 4 over the rational numbers. J. Symbolic Comput. 42, 6 (2007), 579–586.
  • [44] Rónyai, L. Computing the structure of finite algebras. J. Symbolic Comput. 9, 3 (1990), 355–373.
  • [45] Rosenberg, A., and Zelinsky, D. On amitsur’s complex. Trans. Amer. Math. Soc. 97 (1960), 327–356.
  • [46] Simon, D. Solving norm equations in relative number fields using SS-units. Math. Comp. 71, 239 (2002), 1287–1305.
  • [47] The PARI Group. PARI/GP version 2.15.5. Univ. Bordeaux, 2023. available from http://pari.math.u-bordeaux.fr/.
  • [48] Voight, J. Quaternion algebras. Springer, Cham, 2021.
  • [49] Yan, J. Computing the Cassels-Tate Pairing for Jacobian Varieties of Genus Two Curves. PhD thesis, Apollo - University of Cambridge Repository, 2021.