Simple Algebras and Involutions

Total Page:16

File Type:pdf, Size:1020Kb

Simple Algebras and Involutions Chapter 8 Simple algebras and involutions In this chapter, we examine further connections between quaternion algebras, simple algebras, and involutions. 8.1 The Brauer group and involutions ∼ An involution on an F-algebra B induces an isomorphism : B −→ Bop, for example such an isomorphism is furnished by the standard involution on a quaternion algebra B. More generally, if B1, B2 are quaternion algebras, then the tensor product B1 ⊗F B2 has an involution provided by the standard involution on each factor giving an isomorphism B ⊗ B op Bop ⊗ Bop to ( 1 F 2) 1 F 2 —but this involution is no longer a standard involution (Exercise 8.1). The algebra B1 ⊗F B2 is a central simple algebra over F called a biquaternion algebra. In some circumstances, we may have B1 ⊗F B2 M2(B3) (8.1.1) where B3 is again a quaternion algebra, and in other circumstances, we may not; following Albert, we begin this chapter by studying (8.1.1) and biquaternion algebras in detail. To this end, we look at the set of isomorphism classes of central simple algebras over F, which is closed under tensor product; if we think that the matrix ring is something that is ‘no more complicated than its base ring’, it is natural to introduce an equivalence relation on central simple algebras that identifies a division ring with the matrix ring (of any rank) over this division ring. More precisely, if A, A are central simple algebras over F we say A, A are Brauer equivalent if there exist n, n ≥ 1such that Mn(A) Mn (A ). In this way, (8.1.1) reads B1 ⊗F B2 ∼ B3. The set of Brauer equivalence classes [A] has the structure of a group under tensor product, known as the Brauer group Br(F)ofF, with identity element [F] and inverse [A]−1 = [Aop]. The class [B] ∈ Br(F) of a quaternion algebra B is a 2-torsion element, and therefore so is a biquaternion algebra. In fact, by a striking theorem of Merkurjev, when char F 2, all 2-torsion elements in Br(F) are represented by a tensor product of quaternion algebras (see section 8.3). © The Author(s) 2021 123 J. Voight, Quaternion Algebras, Graduate Texts in Mathematics 288, https://doi.org/10.1007/978-3-030-56694-4_8 124 CHAPTER 8. SIMPLE ALGEBRAS AND INVOLUTIONS Finally, our interest in involutions in Chapter 3 began with an observation of Hamilton: the product of a nonzero element with its involute in H is a positive real number (its norm, or square length). We then proved that the existence of such an involution characterizes quaternion algebras in an essential way. However, one may want to relax this setup and instead consider when the product of a nonzero element with its involute merely has positive trace. Such involutions are called positive involutions and they arise naturally in algebraic geometry: the Rosati involution is a positive involution on the endomorphism algebra of an abelian variety, and it is a consequence that this algebra (over Q) is semisimple, and unsurprisingly quaternion algebras once again feature prominently (see sections 8.4–8.5). 8.2 Biquaternion algebras Let F be a field. All tensor products in this section will be taken over F. 8.2.1. Let B1, B2 be quaternion algebras over F. The tensor product B1 ⊗ B2 is a central simple algebra over F of dimension 42 = 16 called a biquaternion algebra.A biquaternion algebra may be written as a tensor product of two quaternion algebras in different ways, so the pair is not intrinsic to the biquaternion algebra. By the Wedderburn–Artin theorem (Main Theorem 7.3.10), we have exactly one of the three following possibilities for this algebra: • B1 ⊗ B2 is a division algebra; • B1 ⊗ B2 M2(B3) where B3 is a quaternion division algebra over F;or • B1 ⊗ B2 M4(F). We could combine the latter two and just say that B1 ⊗ B2 M2(B3) where B3 is a quaternion algebra over F, since M2(M2(F)) M4(F)asF-algebras. Example 8.2.2. By Exercise 8.2, when char F 2wehave a, b a, b 1 ⊗ 2 B F F M2( 3) a, b b a, b a, b a, b2 B = 1 2 ⊗ F F where 3 F . In particular, F F M4( ), since F M2( ). Example 8.2.2 is no accident, as the following proposition indicates. Proposition 8.2.3. (Albert). The following are equivalent: (i) There exists a quadratic field extension K ⊃ F that can be embedded as an F-algebra in both B1 and B2; (ii) B1 and B2 have a common quadratic splitting field; and (iii) B1 ⊗ B2 is not a division algebra. Proof. The equivalence (i) ⇔ (ii) follows from Lemma 5.4.7. 8.2. BIQUATERNION ALGEBRAS 125 ⇒ i = , α ∈ B K α2 = tα − n For the implication (i) (iii), for 1 2let i i generate so i i with t, n ∈ F.Let β := α1 ⊗ 1 − 1 ⊗ α2. Then 2 2 β(α1 ⊗ 1 + 1 ⊗ α2 − t) = α ⊗ 1 − 1 ⊗ α − tβ 1 2 (8.2.4) = (tα1 − n) ⊗ 1 − 1 ⊗ (tα2 − n) − tβ = 0. Therefore β is a zerodivisor and B1 ⊗ B2 is not a division algebra. To finish, we prove (iii) ⇒ (i). We have an embedding B1 → B1 ⊗ B2 α → α ⊗ 1 and similarly B2; the images of B1 and B2 in B1 ⊗ B2 commute. Write B2 = (K, b2 | F). Consider (B1)K = B1 ⊗ K ⊂ B1 ⊗ B2; then (B1)K is a quaternion algebra over K (with dimF (B1)K = 8). If (B1)K is not a division algebra, then K splits B1 and K → B1 and we are done. So suppose that (B1)K is a division algebra. Then B1 ⊗ B2 = (B1)K + (B1)K j is free of rank 2 as a left (B1)K -module. Since B1 ⊗ B2 M2(B3) is not a division algebra, there exists ∈ B1 ⊗ B2 nonzero 2 such that = 0. Without loss of generality, we can write = α1 ⊗ z + j where α1 ∈ B1 and z ∈ K. Then = 2 = α2 ⊗ z2 + α ⊗ z j + α ⊗ z j + b . 0 1 ( 1 ) ( 1 ) 2 (8.2.5) From the basis 1, j over (B1)K ,ifz = t − z with t ∈ F, we conclude that α1 ⊗ z + α1 ⊗ (t − z) = α1 ⊗ t = 0. t = z2 = c c ∈ F× cα2 + b = Therefore 0, and for some . Then√ from (8.2.5) 1 2 0 α2 = −b /c B F −b c B so 1 2 and 1 contains the quadratic field ( 2 ). But so does 2,as 2 (zj) = −b2c as well. (For an alternate proof, see Jacobson [Jacn2009, Theorem 2.10.3].) Remark 8.2.6. In view of Proposition 8.2.3, we say that two quaternion algebras B1, B2 over F are linked if they contain a common quadratic field extension K ⊇ F. For further discussion of biquaternion algebras and linkage in characteristic 2 (where one must treat separable and inseparable extensions differently), see Knus [Knu93], Lam [Lam2002], or Sah [Sah72]. Garibaldi–Saltman [GS2010] study the subfields of quaternion algebra over fields with char F 2. From now on, we suppose that char F 2. (For the case char F = 2, see Chapman– Dolphin–Laghbribi [CDL2015, §6].) 126 CHAPTER 8. SIMPLE ALGEBRAS AND INVOLUTIONS 8.2.7. Motivated by Proposition 8.2.3, we consider the quadratic extensions repre- sented by B1 and B2 encoded in the language of quadratic forms (recalling Lemma 5.5.4). Let V = {α1 ⊗ 1 − 1 ⊗ α2 ∈ B1 ⊗ B2 :trd(α1) = trd(α2)}. V = V = B0 ⊗ − ⊗ B0 Then dimF 6, and we may identify 1 1 1 2. The reduced norm on each factor separately defines a quadratic form on V by taking the difference: explicitly, if B1 = (a1, b1 | F) and B2 = (a2, b2 | F), then taking the standard bases for B1, B2 Q(B1, B2) −a1, −b1, a1b1 −−a2, −b2, a2b2 −a1, −b1, a1b1, a2, b2, −a2b2. The quadratic form Q(B1, B2):V → F is called the Albert form of the biquaternion algebra B1 ⊗ B2. We then add onto Proposition 8.2.3 as follows. Proposition 8.2.8. (Albert).LetB1 ⊗ B2 be a biquaternion algebra over F (with char F 2) with Albert form Q(B1, B2). Then the following are equivalent: (i) B1, B2 have a common quadratic splitting field; (iv) Q(B1, B2) is isotropic. Proof. The implication (ii) ⇒ (iv) follows by construction 8.2.7. To prove (iv) ⇒ (ii), without loss of generality, we may suppose B1, B2 are division algebras; then Q α ∈ B α ∈ B an isotropic vector of corresponds√ to elements 1 1 and 2 2 such that α2 = α2 = c ∈ F× K = F c 1 2 . Therefore ( ) is a common quadratic splitting field. Remark 8.2.9. Albert’s book [Alb39] on algebras still reads well today. The proof of Proposition 8.2.3 is due to him [Alb72]. (“I discovered this theorem some time ago. There appears to be some continuing interest in it, and I am therefore publishing it now.”) Albert used Proposition 8.2.8 to show that − , − x, B = 1 1 B = y 1 F and 2 F over F = R(x, y) have tensor product B1 ⊗F B2 a division algebra by verifying that the Albert form Q(B1, B2) is anisotropic over F. See Lam [Lam2005, Example VI.1] for more details. For the fields of interest in this book (local fields and global fields), a biquaternion algebra will never be a division algebra—the proof of this fact rests on classification results for quaternion algebras over these fields, which we will take up in earnest in Part II.
Recommended publications
  • Nilpotent Elements Control the Structure of a Module
    Nilpotent elements control the structure of a module David Ssevviiri Department of Mathematics Makerere University, P.O BOX 7062, Kampala Uganda E-mail: [email protected], [email protected] Abstract A relationship between nilpotency and primeness in a module is investigated. Reduced modules are expressed as sums of prime modules. It is shown that presence of nilpotent module elements inhibits a module from possessing good structural properties. A general form is given of an example used in literature to distinguish: 1) completely prime modules from prime modules, 2) classical prime modules from classical completely prime modules, and 3) a module which satisfies the complete radical formula from one which is neither 2-primal nor satisfies the radical formula. Keywords: Semisimple module; Reduced module; Nil module; K¨othe conjecture; Com- pletely prime module; Prime module; and Reduced ring. MSC 2010 Mathematics Subject Classification: 16D70, 16D60, 16S90 1 Introduction Primeness and nilpotency are closely related and well studied notions for rings. We give instances that highlight this relationship. In a commutative ring, the set of all nilpotent elements coincides with the intersection of all its prime ideals - henceforth called the prime radical. A popular class of rings, called 2-primal rings (first defined in [8] and later studied in [23, 26, 27, 28] among others), is defined by requiring that in a not necessarily commutative ring, the set of all nilpotent elements coincides with the prime radical. In an arbitrary ring, Levitzki showed that the set of all strongly nilpotent elements coincides arXiv:1812.04320v1 [math.RA] 11 Dec 2018 with the intersection of all prime ideals, [29, Theorem 2.6].
    [Show full text]
  • Classifying the Representation Type of Infinitesimal Blocks of Category
    Classifying the Representation Type of Infinitesimal Blocks of Category OS by Kenyon J. Platt (Under the Direction of Brian D. Boe) Abstract Let g be a simple Lie algebra over the field C of complex numbers, with root system Φ relative to a fixed maximal toral subalgebra h. Let S be a subset of the simple roots ∆ of Φ, which determines a standard parabolic subalgebra of g. Fix an integral weight ∗ µ µ ∈ h , with singular set J ⊆ ∆. We determine when an infinitesimal block O(g,S,J) := OS of parabolic category OS is nonzero using the theory of nilpotent orbits. We extend work of Futorny-Nakano-Pollack, Br¨ustle-K¨onig-Mazorchuk, and Boe-Nakano toward classifying the representation type of the nonzero infinitesimal blocks of category OS by considering arbitrary sets S and J, and observe a strong connection between the theory of nilpotent orbits and the representation type of the infinitesimal blocks. We classify certain infinitesimal blocks of category OS including all the semisimple infinitesimal blocks in type An, and all of the infinitesimal blocks for F4 and G2. Index words: Category O; Representation type; Verma modules Classifying the Representation Type of Infinitesimal Blocks of Category OS by Kenyon J. Platt B.A., Utah State University, 1999 M.S., Utah State University, 2001 M.A, The University of Georgia, 2006 A Thesis Submitted to the Graduate Faculty of The University of Georgia in Partial Fulfillment of the Requirements for the Degree Doctor of Philosophy Athens, Georgia 2008 c 2008 Kenyon J. Platt All Rights Reserved Classifying the Representation Type of Infinitesimal Blocks of Category OS by Kenyon J.
    [Show full text]
  • Generic Splitting Fields of Central Simple Algebras: Galois Cohomology and Non-Excellence
    GENERIC SPLITTING FIELDS OF CENTRAL SIMPLE ALGEBRAS: GALOIS COHOMOLOGY AND NON-EXCELLENCE OLEG T. IZHBOLDIN AND NIKITA A. KARPENKO Abstract. A field extension L=F is called excellent, if for any quadratic form ϕ over F the anisotropic part (ϕL)an of ϕ over L is defined over F ; L=F is called universally excellent, if L · E=E is excellent for any field ex- tension E=F . We study the excellence property for a generic splitting field of a central simple F -algebra. In particular, we show that it is universally excellent if and only if the Schur index of the algebra is not divisible by 4. We begin by studying the torsion in the second Chow group of products of Severi-Brauer varieties and its relationship with the relative Galois co- homology group H3(L=F ) for a generic (common) splitting field L of the corresponding central simple F -algebras. Let F be a field and let A1;:::;An be central simple F -algebras. A splitting field of this collection of algebras is a field extension L of F such that all the def algebras (Ai)L = Ai ⊗F L (i = 1; : : : ; n) are split. A splitting field L=F of 0 A1;:::;An is called generic (cf. [41, Def. C9]), if for any splitting field L =F 0 of A1;:::;An, there exists an F -place of L to L . Clearly, generic splitting field of A1;:::;An is not unique, however it is uniquely defined up to equivalence over F in the sense of [24, x3]. Therefore, working on problems (Q1) and (Q2) discussed below, we may choose any par- ticular generic splitting field L of the algebras A1;:::;An (cf.
    [Show full text]
  • Hyperbolicity of Hermitian Forms Over Biquaternion Algebras
    HYPERBOLICITY OF HERMITIAN FORMS OVER BIQUATERNION ALGEBRAS NIKITA A. KARPENKO Abstract. We show that a non-hyperbolic hermitian form over a biquaternion algebra over a field of characteristic 6= 2 remains non-hyperbolic over a generic splitting field of the algebra. Contents 1. Introduction 1 2. Notation 2 3. Krull-Schmidt principle 3 4. Splitting off a motivic summand 5 5. Rost correspondences 7 6. Motivic decompositions of some isotropic varieties 12 7. Proof of Main Theorem 14 References 16 1. Introduction Throughout this note (besides of x3 and x4) F is a field of characteristic 6= 2. The basic reference for the staff related to involutions on central simple algebras is [12].p The degree deg A of a (finite dimensional) central simple F -algebra A is the integer dimF A; the index ind A of A is the degree of a central division algebra Brauer-equivalent to A. Conjecture 1.1. Let A be a central simple F -algebra endowed with an orthogonal invo- lution σ. If σ becomes hyperbolic over the function field of the Severi-Brauer variety of A, then σ is hyperbolic (over F ). In a stronger version of Conjecture 1.1, each of two words \hyperbolic" is replaced by \isotropic", cf. [10, Conjecture 5.2]. Here is the complete list of indices ind A and coindices coind A = deg A= ind A of A for which Conjecture 1.1 is known (over an arbitrary field of characteristic 6= 2), given in the chronological order: • ind A = 1 | trivial; Date: January 2008. Key words and phrases.
    [Show full text]
  • Commutative Algebra
    Commutative Algebra Andrew Kobin Spring 2016 / 2019 Contents Contents Contents 1 Preliminaries 1 1.1 Radicals . .1 1.2 Nakayama's Lemma and Consequences . .4 1.3 Localization . .5 1.4 Transcendence Degree . 10 2 Integral Dependence 14 2.1 Integral Extensions of Rings . 14 2.2 Integrality and Field Extensions . 18 2.3 Integrality, Ideals and Localization . 21 2.4 Normalization . 28 2.5 Valuation Rings . 32 2.6 Dimension and Transcendence Degree . 33 3 Noetherian and Artinian Rings 37 3.1 Ascending and Descending Chains . 37 3.2 Composition Series . 40 3.3 Noetherian Rings . 42 3.4 Primary Decomposition . 46 3.5 Artinian Rings . 53 3.6 Associated Primes . 56 4 Discrete Valuations and Dedekind Domains 60 4.1 Discrete Valuation Rings . 60 4.2 Dedekind Domains . 64 4.3 Fractional and Invertible Ideals . 65 4.4 The Class Group . 70 4.5 Dedekind Domains in Extensions . 72 5 Completion and Filtration 76 5.1 Topological Abelian Groups and Completion . 76 5.2 Inverse Limits . 78 5.3 Topological Rings and Module Filtrations . 82 5.4 Graded Rings and Modules . 84 6 Dimension Theory 89 6.1 Hilbert Functions . 89 6.2 Local Noetherian Rings . 94 6.3 Complete Local Rings . 98 7 Singularities 106 7.1 Derived Functors . 106 7.2 Regular Sequences and the Koszul Complex . 109 7.3 Projective Dimension . 114 i Contents Contents 7.4 Depth and Cohen-Macauley Rings . 118 7.5 Gorenstein Rings . 127 8 Algebraic Geometry 133 8.1 Affine Algebraic Varieties . 133 8.2 Morphisms of Affine Varieties . 142 8.3 Sheaves of Functions .
    [Show full text]
  • Nilpotent Ideals in Polynomial and Power Series Rings 1609
    PROCEEDINGS OF THE AMERICAN MATHEMATICAL SOCIETY Volume 138, Number 5, May 2010, Pages 1607–1619 S 0002-9939(10)10252-4 Article electronically published on January 13, 2010 NILPOTENT IDEALS IN POLYNOMIAL AND POWER SERIES RINGS VICTOR CAMILLO, CHAN YONG HONG, NAM KYUN KIM, YANG LEE, AND PACE P. NIELSEN (Communicated by Birge Huisgen-Zimmermann) Abstract. Given a ring R and polynomials f(x),g(x) ∈ R[x] satisfying f(x)Rg(x) = 0, we prove that the ideal generated by products of the coef- ficients of f(x)andg(x) is nilpotent. This result is generalized, and many well known facts, along with new ones, concerning nilpotent polynomials and power series are obtained. We also classify which of the standard nilpotence properties on ideals pass to polynomial rings or from ideals in polynomial rings to ideals of coefficients in base rings. In particular, we prove that if I ≤ R[x]is aleftT -nilpotent ideal, then the ideal formed by the coefficients of polynomials in I is also left T -nilpotent. 1. Introduction Throughout this paper, all rings are associative rings with 1. It is well known that a polynomial f(x) over a commutative ring is nilpotent if and only if each coefficient of f(x) is nilpotent. But this result is not true in general for noncommutative rings. For example, let R = Mn(k), the n × n full matrix ring over some ring k =0. 2 Consider the polynomial f(x)=e12 +(e11 − e22)x − e21x ,wheretheeij’s are 2 the matrix units. In this case f(x) = 0, but e11 − e22 is not nilpotent.
    [Show full text]
  • Problem 1. an Element a of a Ring R Is Called Nilpotent If a M = 0 for Some M > 0. A) Prove That in a Commutative Ring R
    Problem 1. An element a of a ring R is called nilpotent if am = 0 for some m> 0. a) Prove that in a commutative ring R the set N of all nilpotent elements of R is an ideal. This ideal is called the nilradical of R. Prove that 0 is the only nilpotent element of R/N. b) Let R be a commutative ring and let a1, ..., an R be nilpotent. Set I for the ∈ ideal < a1, ..., an > generated by a1, ..., an. Prove that there is a positive integer N N such that x1x2...xN = 0 for any x1, ..., xN in I (i.e. that I = 0). c) Prove that the set of all nilpotent elements in the ring M2(R) is not an ideal. d) Prove that if p is a prime and m > 0 then every element of Z/pmZ is either nilpotent or invertible. e) Find the nilradical of Z/36Z (by correspondence theorem, it is equal to nZ/36Z for some n). Solution:a) Suppose that a, b N and r R. Thus an = 0 and bm = 0 for ∈ ∈ some positive integers m, n. By the Newton’s binomial formula we have m+n m + n (a b)m+n = ai( b)m+n−i. (1) − i − i=0 X Note that, for every i 1, 2, ..., m+n , either i n or m+n i m and therefore ∈{ } ≥ − ≥ either ai =0or( b)m+n−i = 0. It follows that every summand in the sum (1) is 0, − so (a b)m+n = 0. Thus a b N.
    [Show full text]
  • Semisimple Cyclic Elements in Semisimple Lie Algebras
    Semisimple cyclic elements in semisimple Lie algebras A. G. Elashvili1 , M. Jibladze1 , and V. G. Kac2 1 Razmadze Mathematical Institute, TSU, Tbilisi 0186, Georgia 2 Department of Mathematics, MIT, Cambridge MA 02139, USA Abstract This paper is a continuation of the theory of cyclic elements in semisimple Lie algebras, developed by Elashvili, Kac and Vinberg. Its main result is the classification of semisimple cyclic elements in semisimple Lie algebras. The importance of this classification stems from the fact that each such element gives rise to an integrable hierarchy of Hamiltonian PDE of Drinfeld-Sokolov type. 1 Introduction Let g be a semisimple finite-dimensional Lie algebra over an algebraically closed field F of characteristic 0 and let e be a non-zero nilpotent element of g. By the Morozov- Jacobson theorem, the element e can be included in an sl(2)-triple s = fe; h; fg (unique, up to conjugacy [Ko]), so that [e; f] = h,[h; e] = 2e,[h; f] = −2f. Then the eigenspace decomposition of g with respect to ad h is a Z-grading of g: d M (1.1) g = gj; where g±d 6= 0. j=−d The positive integer d is called the depth of the nilpotent element e. An element of g of the form e + F , where F is a non-zero element of g−d, is called a cyclic element, associated to e. In [Ko] Kostant proved that any cyclic element, associated to a principal (= regular) nilpotent element e, is regular semisimple, and in [Sp] Springer proved that any cyclic element, associated to a subregular nilpotent element of a simple exceptional Lie algebra, is regular semisimple as well, and, moreover, found two more distinguished nilpotent elements in E8 with the same property.
    [Show full text]
  • Are Octonions Necessary to the Standard Model?
    Vigier IOP Publishing IOP Conf. Series: Journal of Physics: Conf. Series 1251 (2019) 012044 doi:10.1088/1742-6596/1251/1/012044 Are octonions necessary to the Standard Model? Peter Rowlands, Sydney Rowlands Physics Department, University of Liverpool, Oliver Lodge Laboratory, Oxford St, Liverpool. L69 7ZE, UK [email protected] Abstract. There have been a number of claims, going back to the 1970s, that the Standard Model of particle physics, based on fermions and antifermions, might be derived from an octonion algebra. The emergence of SU(3), SU(2) and U(1) groups in octonion-based structures is suggestive of the symmetries of the Standard Model, but octonions themselves are an unsatisfactory model for physical application because they are antiassociative and consequently not a group. Instead, the ‘octonion’ models have to be based on adjoint algebras, such as left- or right-multiplied octonions, which can be seen to have group-like properties. The most promising of these candidates is the complexified left-multiplied octonion algebra, because it reduces, in effect, to Cl(6), which has been identified by one of us (PR) in a number of previous publications as the basic structure for the entire foundation of physics, as well as the algebra required for the Standard Model and the Dirac equation. Though this algebra has long been shown by PR as equivalent to using a complexified left-multiplied or ‘broken’ octonion, it doesn’t need to be derived in this way, as its real origins are in the respective real, complex, quaternion and complexified quaternion algebras of the fundamental parameters of mass, time, charge and space.
    [Show full text]
  • The Descent of Biquaternion Algebras in Characteristic
    The descent of biquaternion algebras in characteristic two Demba Barrya,b, Adam Chapmanc, Ahmed Laghribid aFacult´edes Sciences et Techniques de Bamako, BP: E3206 Bamako, Mali bDepartement Wiskunde–Informatica, Universiteit Antwerpen, Belgium cDepartment of Computer Science, Tel-Hai College, Upper Galilee, 12208 Israel dUniversit´ed’Artois, Facult´edes Sciences Jean Perrin, Laboratoire de math´ematiques de Lens EA 2462, rue Jean Souvraz - SP18, 62307 Lens, France Abstract In this paper we associate an invariant to a biquaternion algebra B over a field K with a subfield F such that K/F is a quadratic separable extension and char(F) = 2. We show that this invariant is trivial exactly when B B0 ⊗ K for some biquaternion algebra B0 over F. We also study the behavior of this invariant under certain field extensions and provide several interesting examples. Keywords: Kato-Milne Cohomology, Cohomological invariants, Algebras with involution, Biquaternion algebras 2010 MSC: primary 11E81; secondary 11E04, 16K20, 19D45 1. Introduction Given a central simple algebra C over a field K with a subfield F, we say that C has a descent to F if there exists a central simple algebra C0 over F such that C C0 ⊗ K. When [K : F] = exp(C), one necessary condition for C to have a descent to F is that corK/F(C) ∼Br F, because if C C0⊗K ⊗[K:F] then corK/F(C) = corK/F(C0 ⊗ K) = C0 ∼Br F, where ∼Br denotes the Brauer equivalence. If K/F is a separable quadratic extension and Q is a quaternion algebra over K, then corK/F(Q) ∼Br F is a necessary and sufficient condition for Q to have a descent to F by [1, Chapter X, Theorem 21].
    [Show full text]
  • Honors Algebra 4, MATH 371 Winter 2010 Assignment 4 Due Wednesday, February 17 at 08:35
    Honors Algebra 4, MATH 371 Winter 2010 Assignment 4 Due Wednesday, February 17 at 08:35 1. Let R be a commutative ring with 1 6= 0. (a) Prove that the nilradical of R is equal to the intersection of the prime ideals of R. Hint: it’s easy to show using the definition of prime that the nilradical is contained in every prime ideal. Conversely, suppose that f is not nilpotent and consider the set S of ideals I of R with the property that “n > 0 =⇒ f n 6∈ I.” Show that S has maximal elements and that any such maximal element must be a prime ideal. Solution: Suppose that f ∈ R is not nilpotent and let S be the set S of ideals I of R with the property that “n > 0 =⇒ f n 6∈ I.” Ordering S by inclusion, note that every chain is bounded above: if I1 ⊆ I2 ⊆ · · · is a chain, then I = ∪Ii is an upper bound which clearly lies in S. By Zorn’s lemma, S has a maximal element, say M, which we claim is prime. Indeed, suppose that uv ∈ M but that u 6∈ M and v 6∈ M. Then the ideals M + (u) and M + (v) strictly contain M so do not belong to S by maximality of M. Thus, there exist m and n such that f n ∈ M + (u) and f m ∈ M + (v). It follows that f m+n ∈ M + (uv) = M and hence that M is not in S, a contradiction. Thus, either u or v lies in M and M is prime.
    [Show full text]
  • Homework 4 Solutions
    Homework 4 Solutions p Problem 5.1.4. LetpR = fm + n 2 j m; n 2 Zg. 2 2 (a) Show that m +pn 2 is a unit in R if and only if m − 2n = ±1. (b) Show that 1 + 2 has infinite order in R×. (c) Show that 1 and −1 are the only units that have finite order in R×. p Proof. (a) Define a map N : R ! Z where N(r) = m2 − 2n2 where r = m + n 2. One can verify that N is multiplicative, i.e. N(rs) = N(r)N(s) for r; s 2 R. Thus if r 2 R× then rs = q for some s 2 R. This implies 2 2 that N(r)N(s) = N(rs) = N(1) = 1 andp hence N(rp);N(s) 2 Zpimplies that m − 2n = N(r) = ±p1. 2 2 2 2 × For the converse observep that (m + n 2)(m +pn 2)(m − n 2) = (m − 2np) = 1, i.e. m + n 2 2 R . (b) Note that 1 + 2 > 1 and therefore (1 + 2)k > 1k = 1. Therefore 1 + 2 has infinite order. (c) Since R is a subring of R and the only elements in R with finite order in R× are ±1 it follows that the only elements in R× with finite order are ±1. Problem 5.1.5. Let R be a subset of an integral domain D. Prove that if R 6= f0g is a ring under the operations of D, then R is a subring of D. Proof.
    [Show full text]