Linear Algebra Over Fields

Total Page:16

File Type:pdf, Size:1020Kb

Linear Algebra Over Fields Some notes on linear algebra Throughout these notes, k denotes a field (often called the scalars in this context). Recall that this means that there are two binary operations on k, denoted + and ·, that (k; +) is an abelian group, · is commutative and associative and distributes over addition, there exists a multiplicative identity 1, and, for all t 2 k, t 6= 0, there exists a multiplicative inverse for −1 t, denoted t . The main example in this course will be k = C, but we shall also consider the cases k = R and k = Q. Another important case is k = Fq, a finite field with q elements, where q = pn is necessarily a prime power (and p is a prime number). For example, for q = p, Fp = Z=pZ. If we drop the requirement that multiplication be commutative, then such a k is called a division algebra or skew field. It is hard to find noncommutative examples of division rings, though. For example, by a theorem of Wedderburn, a finite division ring is a field. One famous example is the quaternions H = R + R · i + R · j + R · k: A typical element of H is then of the form a0 + a1i + a2j + a3k. Here i2 = j2 = k2 = −1, and ij = k = −ji; jk = i = −kj; ki = j = −ik: These rules then force the rules for multiplication for H, and a somewhat lengthy check shows that H is a (noncommutative) division algebra. We shall occasionally relax the condition that every nonzero element has a multiplicative inverse. Thus a ring R is a set with two binary operations + and ·, that (R; +) is an abelian group, · is associative and distributes over addition, and there exists a multiplicative identity 1. If · is commutative as well, we call R a commutative ring. A standard example of a noncommuta- tive ring is the ring Mn(R) of n × n matrices with coefficients in R with the usual operations of matrix addition and multiplication, for n > 1, and with multiplicative identity the n × n identity matrix I. (For n = 1, it is still a ring but in fact it is isomorphic to R and hence is commutative.) More 1 generally, if k is any field, for example k = C and n > 1, then the set Mn(k) of n × n matrices with coefficients in k with the usual operations of matrix addition and multiplication is a noncommutative ring. 1 Definition of a vector space Definition 1.1. A k-vector space or simply a vector space is a triple (V; +; ·), where (V; +) is an abelian group (the vectors), and there is a func- tion k × V ! V (scalar multiplication), whose value at (t; v) is just denoted t · v or tv, such that 1. For all s; t 2 k and v 2 V , s(tv) = (st)v. 2. For all s; t 2 k and v 2 V ,(s + t)v = sv + tv. 3. For all t 2 k and v; w 2 V , t(v + w) = tv + tw. 4. For all v 2 V , 1 · v = v. A vector subspace W of a k-vector space V is a subgroup W of (V; +) such that, for all t 2 k and w 2 W , tw 2 W , i.e. W is a subgroup closed under scalar multiplication. It is then straightforward to check that W is again a k-vector space. For example, f0g and V are always vector subspaces of V . It is a straightforward consequence of the axioms that 0v = 0 for all v 2 V , where the first 0 is the element 0 2 F and the second is 0 2 V , that, for all t 2 F , t0 = 0 (both 0's here are the zero vector), and that, for all v 2 V ,(−1)v = −v. Hence, W is a vector subspace of V () W is nonempty and is closed under addition and scalar multiplication. Example 1.2. (0) The set f0g is a k-vector space. (1) The n-fold Cartesian product kn is a k-vector space, where kn is a group under componentwise addition and scalar multiplication, i.e. (a1; : : : ; an) + (b1; : : : ; bn) = (a1 + b1; : : : ; an + bn) t(a1; : : : ; an) = (ta1; : : : ; tan): (2) If X is a set and kX denotes the group of functions from X to k, then kX is a k-vector space. Here addition is defined pointwise: (f + g)(x) = f(x) + g(x), and similarly for scalar multiplication: (tf)(x) = tf(x). 2 (3) With X a set as in (2), we consider the subset k[X] ⊆ kX of functions f : X ! k such that the set fx 2 X : f(x) 6= 0g is a finite subset of X, then k[X] is easily seen to be a vector space under pointwise addition and scalar multiplication of functions, so that k[X] is a vector subspace of kX . Of course, k[X] = kX if and only if X is finite. In general, for a function f : X ! k, we define the support of f or Supp f by: Supp f = fx 2 X : f(x) 6= 0g: Then k[X] is the subset of kX consisting of all functions whose support is finite. Given x 2 X, let δx 2 k[X] be the function defined by ( 1; if y = x; δx(y) = 0; otherwise: The function δx is often called the delta function at x. Then clearly, for every f 2 k[X], X f(x) = f(x)δx; x2X where in fact the above sum is finite (and therefore meaningful). We will often identify the function δx with the element x 2 X. In this case, we can view X as a subset of k[X], and every element of k[X] can be uniquely P written as x2X tx · x, where the tx 2 k and tx = 0 for all but finitely n many x. The case X = f1; : : : ; ng corresponds to k , and the element δi th to the vector ei = (0;:::; 1;:::; 0) (the i component is 1 and all other components are 0). For a general ring R, the analogue of a vector space is an R-module, with some care if R is not commutative: Definition 1.3. A left R-module M is is a triple (M; +; ·), where (M; +) is an abelian group and there is a function R × M ! M, whose value at (r; m) is just denoted r · m or rm, such that 1. For all r; s 2 R and m 2 M, r(sm) = (rs)m. 2. For all r; s 2 R and m 2 M,(r + s)m = rm + sm. 3. For all r 2 R and m; n 2 M, r(m + n) = rm + rn. 4. For all m 2 M, 1 · m = m. 3 Left submodules of an R-module M are defined in the obvious way. A right R-module is an abelian group M with a multiplication by R on the right, i.e. a function R × M ! M whose value at (r; m) is denoted mr, satisfying (ms)r = m(sr) and the remaining analogues of (2), (3), (4) above. Thus, if R is commutative, there is no difference between left and right R-modules. Example 1.4. (1) The n-fold Cartesian product Rn is a left R-module, where Rn is a group under componentwise addition and scalar multiplica- tion, i.e. (a1; : : : ; an) + (b1; : : : ; bn) = (a1 + b1; : : : ; an + bn) r(a1; : : : ; an) = (ra1; : : : ; ran): Of course, it can also be made into a right R-module. (2) If X is a set and RX denotes the group of functions from X to R, then RX is a left R-module using pointwise addition and scalar multiplication. Similarly the subset R[X], defined in the obvious way, is a submodule. (3) Once we replace fields by rings, there are many more possibilities for R-modules. For example, if I is an ideal in R (or a left ideal if R is not commutative), then I is an R-submodule of R and the quotient ring R=I (in the commutative case, say) is also an R-module. For example, Z=nZ n is a Z-module, but it looks very different from a Z-module of the form Z . For another example, which will be more relevant to us, if R = Mn(k) is the (noncommutative) ring of n × n matrices with coefficients in the field k, then kn is a left R-module, by defining A · v to be the usual multiplication of the matrix A with the vector v (viewed as a column vector). 2 Linear maps Linear maps are the analogue of homomorphisms and isomorphisms: Definition 2.1. Let V1 and V2 be two k-vector spaces and let F : V1 ! V2 be a function (= map). Then F is linear or a linear map if it is a group homomorphism from (V1; +) to (V2; +), i.e. is additive, and satisfies: For all t 2 k and v 2 V1, F (tv) = tF (v). The composition of two linear maps is again linear. For example, for all V1 and V2, the constant function F = 0 is linear. If V1 = V2 = V , then the identity function Id: V ! V is linear. The function F is a linear isomorphism or briefly an isomorphism if it is both linear and a bijection; in this case, it is easy to check that F −1 is 4 ∼ also linear. We say that V1 and V2 are isomorphic, written V1 = V2, if there exists an isomorphism F from V1 to V2.
Recommended publications
  • Equivalence of A-Approximate Continuity for Self-Adjoint
    Equivalence of A-Approximate Continuity for Self-Adjoint Expansive Linear Maps a,1 b, ,2 Sz. Gy. R´ev´esz , A. San Antol´ın ∗ aA. R´enyi Institute of Mathematics, Hungarian Academy of Sciences, Budapest, P.O.B. 127, 1364 Hungary bDepartamento de Matem´aticas, Universidad Aut´onoma de Madrid, 28049 Madrid, Spain Abstract Let A : Rd Rd, d 1, be an expansive linear map. The notion of A-approximate −→ ≥ continuity was recently used to give a characterization of scaling functions in a multiresolution analysis (MRA). The definition of A-approximate continuity at a point x – or, equivalently, the definition of the family of sets having x as point of A-density – depend on the expansive linear map A. The aim of the present paper is to characterize those self-adjoint expansive linear maps A , A : Rd Rd for which 1 2 → the respective concepts of Aµ-approximate continuity (µ = 1, 2) coincide. These we apply to analyze the equivalence among dilation matrices for a construction of systems of MRA. In particular, we give a full description for the equivalence class of the dyadic dilation matrix among all self-adjoint expansive maps. If the so-called “four exponentials conjecture” of algebraic number theory holds true, then a similar full description follows even for general self-adjoint expansive linear maps, too. arXiv:math/0703349v2 [math.CA] 7 Sep 2007 Key words: A-approximate continuity, multiresolution analysis, point of A-density, self-adjoint expansive linear map. 1 Supported in part in the framework of the Hungarian-Spanish Scientific and Technological Governmental Cooperation, Project # E-38/04.
    [Show full text]
  • Classification of Finite Abelian Groups
    Math 317 C1 John Sullivan Spring 2003 Classification of Finite Abelian Groups (Notes based on an article by Navarro in the Amer. Math. Monthly, February 2003.) The fundamental theorem of finite abelian groups expresses any such group as a product of cyclic groups: Theorem. Suppose G is a finite abelian group. Then G is (in a unique way) a direct product of cyclic groups of order pk with p prime. Our first step will be a special case of Cauchy’s Theorem, which we will prove later for arbitrary groups: whenever p |G| then G has an element of order p. Theorem (Cauchy). If G is a finite group, and p |G| is a prime, then G has an element of order p (or, equivalently, a subgroup of order p). ∼ Proof when G is abelian. First note that if |G| is prime, then G = Zp and we are done. In general, we work by induction. If G has no nontrivial proper subgroups, it must be a prime cyclic group, the case we’ve already handled. So we can suppose there is a nontrivial subgroup H smaller than G. Either p |H| or p |G/H|. In the first case, by induction, H has an element of order p which is also order p in G so we’re done. In the second case, if ∼ g + H has order p in G/H then |g + H| |g|, so hgi = Zkp for some k, and then kg ∈ G has order p. Note that we write our abelian groups additively. Definition. Given a prime p, a p-group is a group in which every element has order pk for some k.
    [Show full text]
  • LECTURES on PURE SPINORS and MOMENT MAPS Contents 1
    LECTURES ON PURE SPINORS AND MOMENT MAPS E. MEINRENKEN Contents 1. Introduction 1 2. Volume forms on conjugacy classes 1 3. Clifford algebras and spinors 3 4. Linear Dirac geometry 8 5. The Cartan-Dirac structure 11 6. Dirac structures 13 7. Group-valued moment maps 16 References 20 1. Introduction This article is an expanded version of notes for my lectures at the summer school on `Poisson geometry in mathematics and physics' at Keio University, Yokohama, June 5{9 2006. The plan of these lectures was to give an elementary introduction to the theory of Dirac structures, with applications to Lie group valued moment maps. Special emphasis was given to the pure spinor approach to Dirac structures, developed in Alekseev-Xu [7] and Gualtieri [20]. (See [11, 12, 16] for the more standard approach.) The connection to moment maps was made in the work of Bursztyn-Crainic [10]. Parts of these lecture notes are based on a forthcoming joint paper [1] with Anton Alekseev and Henrique Bursztyn. I would like to thank the organizers of the school, Yoshi Maeda and Guiseppe Dito, for the opportunity to deliver these lectures, and for a greatly enjoyable meeting. I also thank Yvette Kosmann-Schwarzbach and the referee for a number of helpful comments. 2. Volume forms on conjugacy classes We will begin with the following FACT, which at first sight may seem quite unrelated to the theme of these lectures: FACT. Let G be a simply connected semi-simple real Lie group. Then every conjugacy class in G carries a canonical invariant volume form.
    [Show full text]
  • NOTES on DIFFERENTIAL FORMS. PART 3: TENSORS 1. What Is A
    NOTES ON DIFFERENTIAL FORMS. PART 3: TENSORS 1. What is a tensor? 1 n Let V be a finite-dimensional vector space. It could be R , it could be the tangent space to a manifold at a point, or it could just be an abstract vector space. A k-tensor is a map T : V × · · · × V ! R 2 (where there are k factors of V ) that is linear in each factor. That is, for fixed ~v2; : : : ;~vk, T (~v1;~v2; : : : ;~vk−1;~vk) is a linear function of ~v1, and for fixed ~v1;~v3; : : : ;~vk, T (~v1; : : : ;~vk) is a k ∗ linear function of ~v2, and so on. The space of k-tensors on V is denoted T (V ). Examples: n • If V = R , then the inner product P (~v; ~w) = ~v · ~w is a 2-tensor. For fixed ~v it's linear in ~w, and for fixed ~w it's linear in ~v. n • If V = R , D(~v1; : : : ;~vn) = det ~v1 ··· ~vn is an n-tensor. n • If V = R , T hree(~v) = \the 3rd entry of ~v" is a 1-tensor. • A 0-tensor is just a number. It requires no inputs at all to generate an output. Note that the definition of tensor says nothing about how things behave when you rotate vectors or permute their order. The inner product P stays the same when you swap the two vectors, but the determinant D changes sign when you swap two vectors. Both are tensors. For a 1-tensor like T hree, permuting the order of entries doesn't even make sense! ~ ~ Let fb1;:::; bng be a basis for V .
    [Show full text]
  • BLECHER • All Numbers Below Are Integers, Usually Positive. • Read Up
    MATH 4389{ALGEBRA `CHEATSHEET'{BLECHER • All numbers below are integers, usually positive. • Read up on e.g. wikipedia on the Euclidean algorithm. • Diophantine equation ax + by = c has solutions iff gcd(a; b) doesn't divide c. One may find the solutions using the Euclidean algorithm (google this). • Euclid's lemma: prime pjab implies pja or pjb. • gcd(m; n) · lcm(m; n) = mn. • a ≡ b (mod m) means mj(a − b). An equivalence relation (reflexive, symmetric, transitive) on the integers. • ax ≡ b (mod m) has a solution iff gcd(m; a)jb. • Division Algorithm: there exist unique integers q and r with a = bq + r and 0 ≤ r < b; so a ≡ r (mod b). • Fermat's theorem: If p is prime then ap ≡ a (mod p) Also, ap−1 ≡ 1 (mod p) if p does not divide a. • For definitions for groups, rings, fields, subgroups, order of element, etc see Algebra fact sheet" below. • Zm = f0; 1; ··· ; m−1g with addition and multiplication mod m, is a group and commutative ring (= Z =m Z) • Order of n in Zm (i.e. smallest k s.t. kn ≡ 0 (mod m)) is m=gcd(m; n). • hai denotes the subgroup generated by a (the smallest subgroup containing a). For example h2i = f0; 2; 2+2 = 4; 2 + 2 + 2 = 6g in Z8. (In ring theory it may be the subring generated by a). • A cyclic group is a group that can be generated by a single element (the group generator). They are abelian. • A finite group is cyclic iff G = fg; g + g; ··· ; g + g + ··· + g = 0g; so G = hgi.
    [Show full text]
  • Direct Products and Homomorphisms
    Commuting properties Direct products and homomorphisms Simion Breaz logo Simion Breaz Direct products and homomorphisms Products and coproducts Commuting properties Contravariant functors Covariant functors Outline 1 Commuting properties Products and coproducts Contravariant functors Covariant functors logo Simion Breaz Direct products and homomorphisms Products and coproducts Commuting properties Contravariant functors Covariant functors Introduction Important properties of objects in particular categories (e.g. varieties of universal algebras) can be described using commuting properties of some canonical functors. For instance, in [Ad´amek and Rosicki: Locally presentable categories] there are the following examples: If V is a variety of finitary algebras and A ∈ V then A is finitely generated iff the functor Hom(A, −): V → Set preserves direct unions (i.e. directed colimits of monomorphisms); A is finitely presented (i.e. it is generated by finitely many generators modulo finitely many relations) iff the functor Hom(A, −): V → Set preserves directed colimits. logo Simion Breaz Direct products and homomorphisms Products and coproducts Commuting properties Contravariant functors Covariant functors Introduction Important properties of objects in particular categories (e.g. varieties of universal algebras) can be described using commuting properties of some canonical functors. For instance, in [Ad´amek and Rosicki: Locally presentable categories] there are the following examples: If V is a variety of finitary algebras and A ∈ V then A is finitely generated iff the functor Hom(A, −): V → Set preserves direct unions (i.e. directed colimits of monomorphisms); A is finitely presented (i.e. it is generated by finitely many generators modulo finitely many relations) iff the functor Hom(A, −): V → Set preserves directed colimits.
    [Show full text]
  • Gauging the Octonion Algebra
    UM-P-92/60_» Gauging the octonion algebra A.K. Waldron and G.C. Joshi Research Centre for High Energy Physics, University of Melbourne, Parkville, Victoria 8052, Australia By considering representation theory for non-associative algebras we construct the fundamental and adjoint representations of the octonion algebra. We then show how these representations by associative matrices allow a consistent octonionic gauge theory to be realized. We find that non-associativity implies the existence of new terms in the transformation laws of fields and the kinetic term of an octonionic Lagrangian. PACS numbers: 11.30.Ly, 12.10.Dm, 12.40.-y. Typeset Using REVTEX 1 L INTRODUCTION The aim of this work is to genuinely gauge the octonion algebra as opposed to relating properties of this algebra back to the well known theory of Lie Groups and fibre bundles. Typically most attempts to utilise the octonion symmetry in physics have revolved around considerations of the automorphism group G2 of the octonions and Jordan matrix representations of the octonions [1]. Our approach is more simple since we provide a spinorial approach to the octonion symmetry. Previous to this work there were already several indications that this should be possible. To begin with the statement of the gauge principle itself uno theory shall depend on the labelling of the internal symmetry space coordinates" seems to be independent of the exact nature of the gauge algebra and so should apply equally to non-associative algebras. The octonion algebra is an alternative algebra (the associator {x-1,y,i} = 0 always) X -1 so that the transformation law for a gauge field TM —• T^, = UY^U~ — ^(c^C/)(/ is well defined for octonionic transformations U.
    [Show full text]
  • Cohomology Theory of Lie Groups and Lie Algebras
    COHOMOLOGY THEORY OF LIE GROUPS AND LIE ALGEBRAS BY CLAUDE CHEVALLEY AND SAMUEL EILENBERG Introduction The present paper lays no claim to deep originality. Its main purpose is to give a systematic treatment of the methods by which topological questions concerning compact Lie groups may be reduced to algebraic questions con- cerning Lie algebras^). This reduction proceeds in three steps: (1) replacing questions on homology groups by questions on differential forms. This is accomplished by de Rham's theorems(2) (which, incidentally, seem to have been conjectured by Cartan for this very purpose); (2) replacing the con- sideration of arbitrary differential forms by that of invariant differential forms: this is accomplished by using invariant integration on the group manifold; (3) replacing the consideration of invariant differential forms by that of alternating multilinear forms on the Lie algebra of the group. We study here the question not only of the topological nature of the whole group, but also of the manifolds on which the group operates. Chapter I is concerned essentially with step 2 of the list above (step 1 depending here, as in the case of the whole group, on de Rham's theorems). Besides consider- ing invariant forms, we also introduce "equivariant" forms, defined in terms of a suitable linear representation of the group; Theorem 2.2 states that, when this representation does not contain the trivial representation, equi- variant forms are of no use for topology; however, it states this negative result in the form of a positive property of equivariant forms which is of interest by itself, since it is the key to Levi's theorem (cf.
    [Show full text]
  • Lecture 1.3: Direct Products and Sums
    Lecture 1.3: Direct products and sums Matthew Macauley Department of Mathematical Sciences Clemson University http://www.math.clemson.edu/~macaule/ Math 8530, Advanced Linear Algebra M. Macauley (Clemson) Lecture 1.3: Direct products and sums Math 8530, Advanced Linear Algebra 1 / 5 Overview In previous lectures, we learned about vectors spaces and subspaces. We learned about what it meant for a subset to span, to be linearly independent, and to be a basis. In this lecture, we will see how to create new vector spaces from old ones. We will see several ways to \multiply" vector spaces together, and will learn how to construct: the complement of a subspace the direct sum of two subspaces the direct product of two vector spaces M. Macauley (Clemson) Lecture 1.3: Direct products and sums Math 8530, Advanced Linear Algebra 2 / 5 Complements and direct sums Theorem 1.5 (a) Every subspace Y of a finite-dimensional vector space X is finite-dimensional. (b) Every subspace Y has a complement in X : another subspace Z such that every vector x 2 X can be written uniquely as x = y + z; y 2 Y ; z 2 Z; dim X = dim Y + dim Z: Proof Definition X is the direct sum of subspaces Y and Z that are complements of each other. More generally, X is the direct sum of subspaces Y1;:::; Ym if every x 2 X can be expressed uniquely as x = y1 + ··· + ym; yi 2 Yi : We denote this as X = Y1 ⊕ · · · ⊕ Ym. M. Macauley (Clemson) Lecture 1.3: Direct products and sums Math 8530, Advanced Linear Algebra 3 / 5 Direct products Definition The direct product of X1 and X2 is the vector space X1 × X2 := (x1; x2) j x1 2 X1; x2 2 X2 ; with addition and multiplication defined component-wise.
    [Show full text]
  • Right Ideals of a Ring and Sublanguages of Science
    RIGHT IDEALS OF A RING AND SUBLANGUAGES OF SCIENCE Javier Arias Navarro Ph.D. In General Linguistics and Spanish Language http://www.javierarias.info/ Abstract Among Zellig Harris’s numerous contributions to linguistics his theory of the sublanguages of science probably ranks among the most underrated. However, not only has this theory led to some exhaustive and meaningful applications in the study of the grammar of immunology language and its changes over time, but it also illustrates the nature of mathematical relations between chunks or subsets of a grammar and the language as a whole. This becomes most clear when dealing with the connection between metalanguage and language, as well as when reflecting on operators. This paper tries to justify the claim that the sublanguages of science stand in a particular algebraic relation to the rest of the language they are embedded in, namely, that of right ideals in a ring. Keywords: Zellig Sabbetai Harris, Information Structure of Language, Sublanguages of Science, Ideal Numbers, Ernst Kummer, Ideals, Richard Dedekind, Ring Theory, Right Ideals, Emmy Noether, Order Theory, Marshall Harvey Stone. §1. Preliminary Word In recent work (Arias 2015)1 a line of research has been outlined in which the basic tenets underpinning the algebraic treatment of language are explored. The claim was there made that the concept of ideal in a ring could account for the structure of so- called sublanguages of science in a very precise way. The present text is based on that work, by exploring in some detail the consequences of such statement. §2. Introduction Zellig Harris (1909-1992) contributions to the field of linguistics were manifold and in many respects of utmost significance.
    [Show full text]
  • Formal Introduction to Digital Image Processing
    MINISTÉRIO DA CIÊNCIA E TECNOLOGIA INSTITUTO NACIONAL DE PESQUISAS ESPACIAIS INPE–7682–PUD/43 Formal introduction to digital image processing Gerald Jean Francis Banon INPE São José dos Campos July 2000 Preface The objects like digital image, scanner, display, look–up–table, filter that we deal with in digital image processing can be defined in a precise way by using various algebraic concepts such as set, Cartesian product, binary relation, mapping, composition, opera- tion, operator and so on. The useful operations on digital images can be defined in terms of mathematical prop- erties like commutativity, associativity or distributivity, leading to well known algebraic structures like monoid, vector space or lattice. Furthermore, the useful transformations that we need to process the images can be defined in terms of mappings which preserve these algebraic structures. In this sense, they are called morphisms. The main objective of this book is to give all the basic details about the algebraic ap- proach of digital image processing and to cover in a unified way the linear and morpholog- ical aspects. With all the early definitions at hand, apparently difficult issues become accessible. Our feeling is that such a formal approach can help to build a unified theory of image processing which can benefit the specification task of image processing systems. The ultimate goal would be a precise characterization of any research contribution in this area. This book is the result of many years of works and lectures in signal processing and more specifically in digital image processing and mathematical morphology. Within this process, the years we have spent at the Brazilian Institute for Space Research (INPE) have been decisive.
    [Show full text]
  • Division Rings and V-Domains [4]
    PROCEEDINGS OF THE AMERICAN MATHEMATICAL SOCIETY Volume 99, Number 3, March 1987 DIVISION RINGS AND V-DOMAINS RICHARD RESCO ABSTRACT. Let D be a division ring with center k and let fc(i) denote the field of rational functions over k. A square matrix r 6 Mn(D) is said to be totally transcendental over fc if the evaluation map e : k[x] —>Mn(D), e(f) = }(r), can be extended to fc(i). In this note it is shown that the tensor product D®k k(x) is a V-domain which has, up to isomorphism, a unique simple module iff any two totally transcendental matrices of the same order over D are similar. The result applies to the class of existentially closed division algebras and gives a partial solution to a problem posed by Cozzens and Faith. An associative ring R is called a right V-ring if every simple right fi-module is injective. While it is easy to see that any prime Goldie V-ring is necessarily simple [5, Lemma 5.15], the first examples of noetherian V-domains which are not artinian were constructed by Cozzens [4]. Now, if D is a division ring with center k and if k(x) is the field of rational functions over k, then an observation of Jacobson [7, p. 241] implies that the simple noetherian domain D (g)/.k(x) is a division ring iff the matrix ring Mn(D) is algebraic over k for every positive integer n. In their monograph on simple noetherian rings, Cozzens and Faith raise the question of whether this tensor product need be a V-domain [5, Problem (13), p.
    [Show full text]