algorithmic
Efficient computations in central simple algebras using Amitsur cohomology
Abstract
We introduce a presentation for central simple algebras over a field using Amitsur cohomology. We provide efficient algorithms for computing a cocycle corresponding to any such algebra given by structure constants. If 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 -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 and a -algebra isomorphic to , constructing an explicit isomorphism . The explicit isomorphism problem may be thought of as a natural problem in computational representation theory. Given a -algebra , one may wish to assay it. That is, compute the Jacobson radical of , and the decomposition of the semi-simple quotient of as a sum of simple -algebras, themselves identified to some , for a division -algebra. In general, the hard part of this task is to find an isomorphism when is simple. A general recipe for solving this problem is to identify the Brauer class of over its centre , find structure constants for and then compute an explicit isomorphism [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 is also relevant to error correcting codes [32].
It is well known that the Brauer group of equivalence classes of central simple -algebras may be described as the second Galois cohomology group . 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 -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 (where 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 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 -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 -cochain whose differential is a given -coboundary may be found within a certain group of -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 -algebras isomorphic to were first treated separately for small values of . When , 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 by finding a cyclic presentation and solving a cubic norm equation. The case is tackled in [43] by reducing to the case of quaternion algebras over and and solving a norm equation.
In [18] an algorithm was given and studied mostly for the cases and . It was then generalised in [37, 36] to a -algebra isomorphic to , where is a natural number and 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 , the degree of and the size of the discriminant of .
In 2018, [34] exhibited a polynomial-time algorithm for algebras isomorphic to .
2 Preliminaries
2.1 Algorithmic preliminaries
2.1.1 Computational model
The theoretical results of Section 3 are valid for any base field . In order to translate them into efficient algorithms, it is necessary to have an encoding of the elements of , 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 -algebra may be given as input to a program in the form of a table of structure constants. Let be a finite dimensional -algebra of dimension , let be the underlying vector space of and let be a basis of . Multiplication in is a bilinear map which may be represented as a tensor . The table of structure constants of for the basis is then the coordinates of in the basis of , where is the basis of dual to .
An -algebra 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 -algebra of degree is equivalent to knowing an embedding , where is a cyclic (resp. Galois) extension of of degree .
Field extensions and étale algebras admit a presentation as a quotient ring , for some . 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 -units in number fields
The class group of , denoted by is the group of fractional ideals modulo principal ideals. The units in form a group which is called the unit group and is denoted by . All these are important objects in number theory and they motivate the following three algorithmic problems:
-
1.
Compute generators of and their relations,
-
2.
Compute generators of ,
-
3.
Given a principal ideal , find a generator element of .
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 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 where is the discriminant of the number field. We summarise this discussion into the following facts:
Fact 2.3 ([9, Fact 4.1]).
Let be a number field and its discriminant. Under GRH, there exists a polynomial-time algorithm computing a generator set of of size polynomial in .
Proof.
Using the Bach bound one simply enumerates all prime ideals of norm bounded by . The number of such prime ideals is polynomial in . ∎
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 be a set of prime ideals of a number field . There exists a polynomial-time quantum algorithm for computing generators of the group of -units (polynomial in the size of ).
One application of -unit computation (mentioned in [9] and [46]) is that one can compute solutions to norm equations in Galois extensions. Let be a Galois extension of number field and let . Then one would like to find a solution to (or show that it does not exist). The main result by Simon is the following:
Fact 2.6.
([46, Theorem 4.2]) Let be a Galois extension of number fields. Let be a set of primes generating the relative class group . Let be a set of primes containing and let us denote the set of -units in by and the set of -units in by . Then one has that:
Informally, this means that for Galois extensions one can search for in the group of -units. Then by factoring and computing the class group of one can turn finding into solving a system of linear equations. Thus, computing -units implies a polynomial quantum algorithm for solving norm equations in Galois extensions. One important issue though is that the size of has to polynomially bounded as the complexity could explode if was too large. If one assumes GRH, then [5] implies that the size of is not too large.
We summarise this discussion into the following theorem:
Theorem 2.7.
Let be a Galois extension of number fields and let be a set of primes of . Then there exists a polynomial-time quantum algorithm that computes a set of generators for the group of -units of . Furthermore, let . Then there exists a polynomial-time quantum algorithm (assuming GRH) for solving the norm equation .
2.2 Brauer factor sets and central simple algebras
For the remainder of this section, we fix a field , a separable polynomial of degree and an étale algebra over . We let be the smallest Galois extension of which splits , and set
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 -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 be the set of -algebra morphisms from to .
Lemma 2.8.
The map is a bijection between and the set of roots of in . In particular, the set has elements.
Proof.
If , it is clear that is a root of in . Now, the morphism is entirely determined by so the map is injective.
Let be a root of . Then, let be the unique irreducible factor of such that . Then, and we have a map from to sending to . This maps precomposed with the isomorphism defined above sends to . ∎
We may then index the maps from to , so that . If and , we write for the index of .
Definition 2.9.
Let . An -cochain is a map
which satisfies the following homogeneity condition: for , for ,
(1) |
A cochain may be denoted as the family of values that it takes over .
A cochain is said to be reduced if for all . Under pointwise multiplication, the -cochains form a group which we denote by . Then, the reduced cochains form a subgroup denoted by .
If and , we set
For , there is a group homomorphism defined as follows: if ,
We set and call elements of this subgroup -cocycles. We write for the group of reduced cocycles. For positive, we also set and call its elements -coboundaries. We note that is a subgroup of . For completeness, we let be the trivial subgroup of . Finally, we set .
Two cocycles are called associated, denoted by , if they have the same class in . If , a preimage of by is called a trivialisation of .
Remark 2.10.
The cochain is always a cocycle and is a coboundary for . It is called the trivial cocycle.
Proposition 2.11.
For positive, every -cocycle is associated to a reduced cocycle.
Proof.
Let , and consider the cochain defined by and if the are not all equal to one another. Then, the cocycle is reduced and associated to . ∎
We may now recall the explicit description of the isomorphism (that is, the group of Brauer classes of central simple -algebras that are split by ), as exposed in [38, Chapter 2]. Using Lemma 2.8, we observe that what is called a Brauer factor set there is a -cocycle in our notation.
Definition 2.12.
Let . Let be the -vector space of -homogeneous maps from to . We turn into a -algebra using the following definition. If , we set such that for all ,
Fact 2.13.
If is a -cocycle, the -algebra is a central simple -algebra. Furthermore, if and are associated cocycles, and is a -cochain such that , then the map
is an isomorphism. Conversely, if is a central simple -algebra split by , then there exists a cocycle such that If are such that and are isomorphic algebras, then the cocycles and 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 are associated and both reduced, then a cochain such that 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 be a central simple -algebra, and let be a separable maximal commutative subalgebra of . Then, there exists such that . Furthermore, letting be a Galois extension of which splits , there is a map such that every is mapped to . Let be the image of in . Then, every element of becomes a matrix of the form , where . The map gives an isomorphism from to , where .
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 , and by we mean the tensor product of copies of the -algebra .
Definition 3.1.
An -cochain in the sense of Amitsur is an invertible element in the -algebra , and we write for the group of -cochains.
For and , we define a -algebra homomorphism from to . The map is defined on the simple tensor elements as follows:
We then define the group homomorphism
Define the subgroup of , and its elements are called -cocycles in the sense of Amitsur.
If is positive, we also define the subgroup of and its elements are -coboundaries in the sense of Amitsur. Then, the th Amitsur cohomology group is defined as the quotient group .
Two -cocycles and are called associated if they have the same class in . If , a cochain such that is called a trivialisation of .
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 from to the algebra of -homogeneous maps from to , which is defined over simple tensors by
The map is an isomorphism of -algebras, which also sends the unit group onto the unit group . Furthermore, , and it follows that cocycle and coboundaries subgroup are also isomorphic to one another, and so are cohomology groups.
Proof.
If is a separable field extension of , this is the content of [45, Section 2]. However, the proof given there readily generalises to the case that is an étale -algebra. ∎
If , we get a homomorphism of -algebras
By the construction of , if and , then we get
(2) |
Adapting Definition 2.12 to Amitsur cohomology, we get:
Definition 3.3.
Let be a cocycle. The -algebra is defined as follows:
As a -vector space, is . We see as a -algebra via the embedding . Then, multiplication in is defined by
In the sequel, we use the same embedding whenever we write .
Observe that the algebra may be seen as the extension of scalars to of the -algebra . Since we make the choice of seeing as a -algebra via the mapping , the relevant injection sends an element to . Then, by [10, Section III.9.3], for , we have
By linearity of the trace, if , we get
Proposition 3.4.
Let . The map is a -algebra isomorphism from to .
Proof.
The map is already known to be an isomorphism of vector space between and , as the underlying vector spaces of these algebras are respectively and the space of -homogeneous maps from to . We therefore only need to check that the map is a homomorphism of -algebras.
It is well known that if , . Then, if and , we get:
Extending this result by linearity, we get that
Now, using Equation 2 again, we get that if and ,
where the multiplication in the last line is meant to be in . ∎
Remark 3.5.
Let be the trivial cocycle. There is an explicit isomorphism between and . Indeed, the algebra is isomorphic to the algebra of -linear endomorphisms of , where is the dual -vector space of . In fact, the map gives a -linear isomorphism from to its dual space , so we may see as the algebra with the following multiplication law:
Let , we compute the following multiplication in :
It follows that if is a basis of and we identify elements of with the column vectors of their components in basis , and for , we set to be the row matrix of the linear form , then an isomorphism from to is given by
Theorem 3.6.
Let be a central simple -algebra, let be a separable maximal commutative subalgebra, and let be such that . Then, there is a cocycle such that the map
is an isomorphism of -algebras from to (which makes sense as is the underlying -vector space of ). Furthermore, is the unique element of with this property (if we extend the definition of the multiplication in to non-cocycle elements).
Proof.
Let be the Brauer factor set constructed from the data of and as in the discussion at the end of Section 2.2, and let be the isomorphism from to constructed there. We also let . The fact that is an isomorphism of -algebras follows from the observation that and from 3.4.
The cocycle is a solution to the system of -linear equations
for all . Observe that ranges over as range over . Then, the uniqueness of follows from the fact that the map is an isomorphism from to its dual as a -module.
This last fact, in turn, follows from [29, Corollary 4.6.8 and Proposition 4.6.1]. Indeed, , and is a separable polynomial, so is a separable -algebra. ∎
Corollary 3.7.
Let be associated cocycles, and let be such that . Then, the map , where the multiplication used is the natural one in , is an isomorphism from to .
Proof.
This is just a consequence of the analogous property for Brauer factor sets (2.13), pulled back through the isomorphisms . ∎
Remark 3.8.
For the sake of efficiency, we use already proven results on Brauer factor-sets and the isomorphisms and to describe the isomorphism between the groups and . However, the simplicity of the isomorphism described in Theorem 3.6 suggests that Amitsur cohomology is a natural setting for giving an algebraic description of 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 and an étale -algebra are fixed, for , we have an isomorphism
where is the ideal generated by the polynomials for . Furthermore, an element of admits a unique lift in such that the individual degrees in are all lesser or equal to . If , we denote by the lift of as described above. There is an obvious polynomial algorithm which, given some , computes . Indeed, one may simply reduce modulo , then modulo , and so on… This allows us to represent an element as its lift , and in general we get and .
In this setting, the map becomes the map
That is, we define a map
from to , and we define
The maps defined above are compatible with the eponymous maps from the Amitsur complex. That is, we have .
Now, the trace map of as an algebra (via the morphism ) may easily be computed in the -basis of . Let , , , and . We set , and we have
That is,
For the remainder of this work, elements of algebras of the form , when is an étale -algebra, are represented algorithmically as their images by in .
3.3 Representing central simple algebras as Amitsur algebras
We may now give an algorithm for finding a representation of a central simple algebra over any field of sufficiently large size where linear algebra tasks may be performed efficiently.
Before we prove the correctness and efficiency of Algorithm 1, we need a lemma:
Lemma 3.9.
Let be a field and let be a central simple -algebra. Assume that . Let be such that is a maximal commutative subalgebra of . Then an element such that may be found in probabilistic polynomial time.
Proof.
For in , by an argument of dimensions over , we observe that if and only if the map
is injective.
The family is a basis of . Let be the input basis of (that is, the basis for which the structure constants of are expressed). Then, identifying with its matrix in the bases written above, the determinant of is a homogeneous polynomial on the coordinates of in basis of degree bounded by . Furthermore, if and only if is not a zero of this polynomial. We note that by [38, Theorem 2.2.2], this polynomial is nonzero.
Letting be a finite subset of , the Schwartz-Zippel lemma ensures that a random in satisfies this condition with probability larger than .
Therefore, if , we may pick large enough that has the desired property with positive probability, and small enough that we may draw a random element of efficiently. For instance, take . Then, has the desired property with probability larger than . In this case, it is expected to take less than attempts to find a suitable element . ∎
Theorem 3.10.
If is a field over which linear algebra may be performed efficiently, and is a central simple -algebra such that , 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 , as well as , the minimal polynomial of . We obtain a subalgebra of and such that . Then, by Theorem 3.6, there exists a cocycle such that is an isomorphism from to , and such a is unique in . The equation solved in Algorithm 1 finds a representation of this .
Now we consider the complexity of the algorithm:
The element in Algorithm 1 may be found using the polynomial algorithm given in [22].
Then, 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 -algebras , and , 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 -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 -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 -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 be a global field. Then, recall that a commutative -algebra is étale if and only if it satisfies the following equivalent conditions:
-
1.
The algebra factors as a direct product of finite separable extensions of .
-
2.
The algebra is isomorphic to an algebra of the form , where is a separable polynomial (that is, has no multiple root in an algebraic closure of ).
Remark 4.1.
If is an étale -algebra, the isomorphism discussed above is unique up to reindexing and automorphism of the factors. In fact, the factors identify with the minimal ideals of . Therefore, we usually identify an étale algebra with the product . We will also denote by the projection map .
If is a global field, we write for the set of non-archimedean places of . Then, recall that the divisor group is the free abelian group on the set , that an element has an associated divisor , and that if is the group of principal divisors, then the class group of is finite. When is a number field, the divisor group is isomorphic to the group of fractional ideals of , and the class group in terms of divisors is then isomorphic to the ideal class group.
Definition 4.2.
Let be an étale -algebra factored as above. The set of non-archimedean places of is the disjoint union . The divisor group of is the free abelian group on . The divisor of an invertible element is defined as the sum of the divisors of its projections in the factors of , and the groups and are defined by analogy with their definitions for global fields. If , we let be the support of .
An immediate consequence of this definition is that the class group of an étale -algebra is finite.
If and are two étale -algebras, and is a homomorphism of -algebras, it decomposes as a matrix , where , and each is either the zero map or an embedding of field extensions of (depending on whether the unit of is mapped to or to the unit of ). That is, if and ,
where the are the projections and the are the projections .
We may then define a map the following way: let and , we set , where is the usual map on divisors of global fields if is an embedding of fields and is zero if . One may check that this definition makes into a functor from the category of étale -algebras to the category of abelian groups. Furthermore, the map is a natural transformation of the multiplicative group functor into the functor sending an étale algebra to its divisor group. That is, if , then .
As is the case for global fields, if , there is at most one place such that . To prove this, we first need a lemma:
Lemma 4.3.
Let notations be as above and let . Then there is at most one such that is nonzero.
Proof.
Let and , and let be the units respectively of and . We then have and . Now, since , we have . That is, either or . It follows that either or is the zero map. ∎
Now, we prove the following:
Proposition 4.4.
Let notations be as above, and let . Then there is at most one place such that .
Proof.
Let be such that . Then, by Lemma 4.3, there is at most one value such that . If there is no such , then no divisor of has an image by with in its support. Otherwise, if , it follows that . Then, the usual theory of divisors of global fields ensures that there is only one such that . ∎
If is such that is nonzero for some , and if , we write for the unique such that . We note that if is an inclusion of the form , , then exists for all .
Let such that is well defined. We let be the coefficient of in the expansion of the divisor . Let , if for some , we say that the place ramifies through . This is equivalent to saying that, if is such that , then for some such that is nonzero, the place ramifies in .
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 be as above. Let and let . Then the local completion of at is the completion of the global field at the place . This completion is an -algebra via the composite map
As in the case of extensions of global fields, the places of lying above a place of may be classified as the direct factors of a tensor product.
Proposition 4.6.
Let notations be as above, let and . Let . Then, we have
The usual bijections between the places of above and the direct factors of give a bijection between the direct factors of and . Furthermore, for , we have if and only if the corresponding factor is a ramified extension of .
Proof.
We have
Then, the rest of the results follows directly from the definition of 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 and be étale -algebras, and let be a place that ramifies neither in nor in . Then, does not ramify in either.
Proof.
Let be the completion of at . By 4.6, it is enough to prove that the direct factors of are unramified extensions of . Now, since
the result will follow from the fact that if are unramified finite extensions of , then the direct factors of are themselves unramified extensions of .
This, in turns, follows from [11, Proposition I.7.1]. Indeed, by the proposition, we know that there are polynomials (where is the valuation ring of ) such that , and whose residues and are irreducible and separable in , where is the residue field of . Now, we have . A direct factor of this algebra is of the form , where is an irreducible factor of in , where is the valuation ring of . Since is a factor of the separable polynomial , it is itself separable as a polynomial in , where is the residue field of . Since the polynomial is irreducible in and its residue is separable, this residue is irreducible in by Hensel’s lemma. By the proposition cited above, the field extension is unramified over , and therefore over . It follows that the direct factors of are unramified over . ∎
4.2 An algorithm for trivialising Amitsur coboundaries over a number field
Let be a number field, and let be an étale -algebra. The Amitsur complex for yields a complex
of abelian groups, where we define as .
We give a precise description of the map . Let . Then, for any , we set and . Then, if , we get
and it follows that
and
We first need two lemmas:
Lemma 4.8.
Let be places of such that . Then there exists a place such that and .
Proof.
We let be a defining polynomial of . That is, , with separable. We may then identify with , where maps into the ring of scalars in and is the map
Likewise, we identify with . Then, the descriptions of the maps and become:
These identifications are coherent with the maps of the Amitsur complex. In particular, we note that . With this in place, let . By 4.6, the support of is in bijection with the direct factors of . That is, it is in bijection with the irreducible factors of in . Likewise, the support of is in bijection with the irreducible factors of . That is, it is in bijection with the maximal ideals that contain the ideal .
Now, if , we let be the irreducible factor of corresponding to . Likewise, if , we let be the maximal ideal of corresponding to . Observe that if and only if and if and only if . Then, the result simply follows from the observation that the ideal contains and is contained in some maximal ideal of . Such a maximal ideal then corresponds to a place such that and . ∎
For and , we let be the subset of of places lying above some . We let be the set of places of that ramify in . Applying 4.7 by induction, we get the following:
Lemma 4.9.
Let . Then for , we have .
We may now prove the following generalisation of Hilbert’s Theorem 90:
Lemma 4.10.
Let be supported by places outside of . Then, there exists such that .
Proof.
It follows that
We introduce the following automorphisms:
Observe that we have the following equalities:
(3) | ||||
(4) | ||||
(5) |
If (resp. ), we write (resp. ) for the unique place in the support of (resp. ).
Now, let such that . By Equation 5, we have . Applying Lemma 4.8 to and , we get such that and . By Equation 4, we have and by Equation 3, we get . We set . Now, the coefficient of in is , and since this divisor is zero by hypothesis, we get (note that , so for ). Observe that we have
and
By 4.4, it follows that . That is, we proved that there exists such that and .
Conversely, let such that . Then by Lemma 4.8, there is such that and . Set and, as above, we get . As above, we use the fact that to prove that .
Putting things together, we have proved that for ,
It follows directly that . ∎
We now get our main theorem for this section:
Theorem 4.11.
Let be a coboundary. Let be a set of non-archimedean places of such that:
-
•
. That is, contains the places of that ramify in .
-
•
The finite places of generate the class group .
-
•
.
Then there exists a cochain in the group of -units of such that
Proof.
Let be such that . We consider the divisor of . We set and . Now we get and therefore it is supported by . Observe that if , then and conversely, if and , then . It follows that .
The support of is disjoint from . We may therefore apply Lemma 4.10 and get a divisor such that . Now, as generates the class group of , there exists with support in and such that . Then, we get that
and therefore
Now, since has support in , this shows that is a -unit. Furthermore,
and is a cochain with the required properties. ∎
From Theorem 4.11 we directly get an algorithm for computing a trivialisation of a -coboundary:
Theorem 4.12.
Given a number field , a separable polynomial and a coboundary , where , Algorithm 2 outputs the representation of a cochain such that . 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 , and as direct products of number fields. Set .
The set may be computed by factoring the discriminants of the extensions , 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 that generate the class group , and the set of places of lying below them. In practice, one may set to be the set of prime ideals of such that . We may then take the image of in the product and factor the principal ideals in into a product , which may be done in polynomial time by 2.2. We may then set .
Isomorphisms and 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 .
The correctness of the algorithm relies on the fact that a cochain such that exists and may be found in the group of -units of , which is the content of Theorem 4.11. ∎
Theorem 4.13.
Assuming GRH, there exists a polynomial quantum algorithm which, give a number field and an algebra , computes an explicit algorithm from to .
Proof.
This is simply a combination of Theorems 3.10 and 4.12. Indeed, using Algorithm 1, one may compute an étale algebra , a cocycle and an isomorphism . Since is isomorphic to , the cocycle is in fact a coboundary. Then, a trivialisation of may be computed using Algorithm 2. Applying 3.7, we obtain an explicit isomorphism . Finally, an isomorphism 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 . Found. Comput. Math. 18 (2018), 381–397.
- [35] Ivanyos, G., Kutas, P., and Rónyai, L. Explicit equivalence of quadratic forms over . 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 -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.