00 Introduction: Homological and Commutative Algebra 1. Some

Total Page:16

File Type:pdf, Size:1020Kb

00 Introduction: Homological and Commutative Algebra 1. Some (October 16, 2009) 00 Introduction: homological and commutative algebra Paul Garrett [email protected] http://www.math.umn.edu/˜garrett/ 1. Some recurring themes 2. Example: what is an indeterminate? 3. Foundational dangers: set theory and universes 4. Resolutions, complexes, snake lemma 5. Pictorial interlude, tensors 6. Adjoint pairs of functors 7. Counting holes: algebraic topology [iou] 8. Appendix: set theory axioms A fundamental idea emphasized by category theory is characterization of objects by their interactions with other objects, rather than constructions and internal details. [1] More generally: Sufficiently many good examples should make formal definitions nearly unnecessary, and should make most basic results obvious. A good methodological and technical viewpoint should enable not merely verification of interesting results, but should assist discovery. Phenomena are more interesting than symbols. 1. Some recurring themes There are many threads here, touched at various levels of sophistication, interacting in many ways. The following are some of the recurring ideas: Naive category theory has the immediate feature of characterizing or describing objects by their mappings to other objects, rather than by their internal structure or construction. The modifier naive indicates that we do not aggressively axiomatize, and do not necessarily worry about reconciling categorical ideas with set theory. Naive set theory is the typical background or foundation for constructing more complicated structures, in a century-old tradition. The modifier naive denotes an informal and non-axiomatic treatment, in contrast to versions of set theory specifically aimed at preventing paradoxes by limiting constructions. Part of the set theory tradition demands that any mathematical discussion be compatible with the explicit or implicit prohibitions of set theory. Linear algebra in the broadest sense takes vector spaces over fields as prototypes for modules over general rings. Decomposition into eigenvectors or more complicated atomic pieces appears everywhere. Collections of modules over rings are the prototypes for abelian categories, including important examples that are not literal categories of modules, such as categories of sheaves or complexes of sheaves. Functors arise immediately in category theory, being maps from one category to another, converting one type of object to another, and converting corresponding maps-of-objects from one type to the other. Homology [1] The occasional definition of new objects as symbols behaving a certain way is a pale shadow of the idea of characterization by interactions, but made clumsy and opaque by entangling half-hearted constructions with under- specification. 1 Paul Garrett: 00 Introduction: homological and commutative algebra (October 16, 2009) and homotopy groups of topological spaces are dramatic first examples. Once we realize that most systematic procedures describe functors, we see that functors are everywhere. Adjoint functors and adjunction relations: many important functors occur in adjoint pairs, satisfying an adjunction relation, in a sense a sort of inversion, but significantly different. Many forgetful or lossy functors, while boring themselves, have interesting adjoints. Yoneda’s small lemma yields left exactness of right adjoints, and right exactness of left adjoints, once and for all. Natural transformations are maps of functors, and allow formulation of the notion of equivalence of functors, the best version of isomorphism of functors. Homology and cohomology arose as counting holes in geometric objects. Geometric interpretations are still important, but the underlying algebra is universally useful. Extensions of rings appear in algebraic number theory and in algebraic geometry, with arithmetic or geometric motivations. Many important features are best described in categorical and homological terms. Complexes eventually appear as more appropriate fundamental objects than individual modules. The structure theorem for finitely-generated modules over principal ideal domains is a prototype. Projective and injective resolutions are the fundamental acyclic resolutions. These and other themes are profitably illustrated with examples, as follow. 2. Example: what is an indeterminate? What is an indeterminatex? We are not asking about a merely unknown literal number, as in the earliest use of algebra, but about the x that appears in polynomials and other functions in calculus. The x is sometimes alleged to be a variable number, or an arbitrary number. It’s not clear what variable means, but the vague idea of motion is generally compatible with the motivating issues of calculus. These are not bad pseudo-definitions to give calculus students. However, the idea that x is merely a variable number does not match our use of it. For example, in the Cayley-Hamilton theorem [2] the indeterminate x is replaced by a linear transformation. This suggests that x is something that can be substituted for. Elementary mathematics has no way to say this with any precision, but a categorical viewpoint succeeds, as follows. Let R be a commutative ring, probably [3] with 1. An R-algebra A is a ring A that is also an R-module, and so that the R-module structure is compatible with the multiplication in A: (r a)b = r (ab) (for r R, a,b A) · · ∈ ∈ For example, R actually sitting inside the center of A gives A an R-algebra structure. But we do also want to allow some collapsing of R in the way that it acts on A. For example, Z/n is a Z-algebra in the obvious fashion. [2] Recall the assertion of the Cayley-Hamilton theorem: let T be an endomorphism of a finite-dimensional vector space V , let χT (x) be the characteristic polynomial det(x · I − T ), with I the identity map on V . Then χT (T ) = 0. The non-proof of this by (falsely) claiming that χT (T ) = det(T − T ) = det 0 = 0 is of interest, if only to provoke us to see why it’s not a proof. [3] Rings with 1 are more intuitive, and we hope that rings have 1, but occasionally there are good reasons to consider rings without units. But at the moment there’s insufficient motivation to get embroiled with issues about lack of units. 2 Paul Garrett: 00 Introduction: homological and commutative algebra (October 16, 2009) An R-algebra homomorphism or map f : A B is a ring homomorphism f : A B that respects the R-algebra structure, in the reasonable sense that→ → f(r a) = r f(a) (for r R and a A) · · ∈ ∈ Given the idea of R-algebra, we can say what x is, albeit indirectly. A free R-algebra on x is an R-algebra F (thinking that F = R[x], but don’t want to prejudice ourselves by notation!) with distinguished element x F , having the universal property that, for any R-algebra A and choice of element ao A, there exists a unique∈ R-algebra map f : F A such that fx = a . ∈ → o This universal property characterizes the free algebra R[x] not by telling what x is, but by telling how it interacts with other R-algebras. The main idea is that we have substituted ao for x, and that the rest of F accompanies x suitably. We do not exclude the possibility that something happens to the ring R, as in quotient maps Z Z/n. → Grant for a moment that the more elementary notion of polynomial ring R[x] really is a free R-algebra on the generator x. Then for a field k we can certainly map x to an endomorphism T of a finite-dimensional k-vectorspace V , mapping the free k-algebra k[x] to the k-algebra k[T ] of k-linear endomorphisms of V [4] generated by T . The more mundane case of replacing x by an element ro of the ring R gives an R-algebra homomorphism R[x] R mapping x ro, called an evaluation homomorphism. A slightly subtler example is Q[x] Q(√→2) by mapping x→ √2. → → We might worry that the mapping characterization of the free algebra F = R[x] accidently under-specified it. After all, we’ve not directly said anything about the internal structure of it. However, mildly amazingly, this seemingly vague and indirect characterization completely determines F = R[x]. Further, the style of the proof uses almost nothing about the specifics of the situation. That is, suppose we had two free R-algebras on single generators: suppose A a free R-algebra on ao, and B a free R-algebra on b . Then we claim that there is a unique isomorphism j : A B of R-algebras such o → that jao = bo. Indeed, using the characterization of A, a as free R-algebra, there is a unique α : A B with αa = b . o → o o On the other hand, using the characterization of B,bo as free R-algebra, there is a unique β : B A with βb = a . Then α β is an R-algebra map of B to itself sending b to itself, and β α is an R-algebra→ map of o o ◦ o ◦ A to itself sending ao to itself. Obviously we imagine that both these composites are the respective identity maps, which would prove the claim. But one small point is missing, as follows. Universal objects have no proper automorphisms: we claim that the only R-algebra map of A to itself sending a to itself is the identity map on A. This and the corresponding fact for B and b would prove that α β o o ◦ and β α are the respective identities, and, thus, that A, ao and B,bo are isomorphic. Using the mapping characterization◦ of A and a in a seemingly silly way, there is a unique R-algebra map A A sending a o → o to ao. On the other hand, the identity map on A certainly has this property. The uniqueness part of the characterization then says that there is no other such map than the identity.
Recommended publications
  • Math 651 Homework 1 - Algebras and Groups Due 2/22/2013
    Math 651 Homework 1 - Algebras and Groups Due 2/22/2013 1) Consider the Lie Group SU(2), the group of 2 × 2 complex matrices A T with A A = I and det(A) = 1. The underlying set is z −w jzj2 + jwj2 = 1 (1) w z with the standard S3 topology. The usual basis for su(2) is 0 i 0 −1 i 0 X = Y = Z = (2) i 0 1 0 0 −i (which are each i times the Pauli matrices: X = iσx, etc.). a) Show this is the algebra of purely imaginary quaternions, under the commutator bracket. b) Extend X to a left-invariant field and Y to a right-invariant field, and show by computation that the Lie bracket between them is zero. c) Extending X, Y , Z to global left-invariant vector fields, give SU(2) the metric g(X; X) = g(Y; Y ) = g(Z; Z) = 1 and all other inner products zero. Show this is a bi-invariant metric. d) Pick > 0 and set g(X; X) = 2, leaving g(Y; Y ) = g(Z; Z) = 1. Show this is left-invariant but not bi-invariant. p 2) The realification of an n × n complex matrix A + −1B is its assignment it to the 2n × 2n matrix A −B (3) BA Any n × n quaternionic matrix can be written A + Bk where A and B are complex matrices. Its complexification is the 2n × 2n complex matrix A −B (4) B A a) Show that the realification of complex matrices and complexifica- tion of quaternionic matrices are algebra homomorphisms.
    [Show full text]
  • CLIFFORD ALGEBRAS and THEIR REPRESENTATIONS Introduction
    CLIFFORD ALGEBRAS AND THEIR REPRESENTATIONS Andrzej Trautman, Uniwersytet Warszawski, Warszawa, Poland Article accepted for publication in the Encyclopedia of Mathematical Physics c 2005 Elsevier Science Ltd ! Introduction Introductory and historical remarks Clifford (1878) introduced his ‘geometric algebras’ as a generalization of Grassmann alge- bras, complex numbers and quaternions. Lipschitz (1886) was the first to define groups constructed from ‘Clifford numbers’ and use them to represent rotations in a Euclidean ´ space. E. Cartan discovered representations of the Lie algebras son(C) and son(R), n > 2, that do not lift to representations of the orthogonal groups. In physics, Clifford algebras and spinors appear for the first time in Pauli’s nonrelativistic theory of the ‘magnetic elec- tron’. Dirac (1928), in his work on the relativistic wave equation of the electron, introduced matrices that provide a representation of the Clifford algebra of Minkowski space. Brauer and Weyl (1935) connected the Clifford and Dirac ideas with Cartan’s spinorial represen- tations of Lie algebras; they found, in any number of dimensions, the spinorial, projective representations of the orthogonal groups. Clifford algebras and spinors are implicit in Euclid’s solution of the Pythagorean equation x2 y2 + z2 = 0 which is equivalent to − y x z p = 2 p q (1) −z y + x q ! " ! " # $ so that x = q2 p2, y = p2 + q2, z = 2pq. If the numbers appearing in (1) are real, then − this equation can be interpreted as providing a representation of a vector (x, y, z) R3, null ∈ with respect to a quadratic form of signature (1, 2), as the ‘square’ of a spinor (p, q) R2.
    [Show full text]
  • Notes and Solutions to Exercises for Mac Lane's Categories for The
    Stefan Dawydiak Version 0.3 July 2, 2020 Notes and Exercises from Categories for the Working Mathematician Contents 0 Preface 2 1 Categories, Functors, and Natural Transformations 2 1.1 Functors . .2 1.2 Natural Transformations . .4 1.3 Monics, Epis, and Zeros . .5 2 Constructions on Categories 6 2.1 Products of Categories . .6 2.2 Functor categories . .6 2.2.1 The Interchange Law . .8 2.3 The Category of All Categories . .8 2.4 Comma Categories . 11 2.5 Graphs and Free Categories . 12 2.6 Quotient Categories . 13 3 Universals and Limits 13 3.1 Universal Arrows . 13 3.2 The Yoneda Lemma . 14 3.2.1 Proof of the Yoneda Lemma . 14 3.3 Coproducts and Colimits . 16 3.4 Products and Limits . 18 3.4.1 The p-adic integers . 20 3.5 Categories with Finite Products . 21 3.6 Groups in Categories . 22 4 Adjoints 23 4.1 Adjunctions . 23 4.2 Examples of Adjoints . 24 4.3 Reflective Subcategories . 28 4.4 Equivalence of Categories . 30 4.5 Adjoints for Preorders . 32 4.5.1 Examples of Galois Connections . 32 4.6 Cartesian Closed Categories . 33 5 Limits 33 5.1 Creation of Limits . 33 5.2 Limits by Products and Equalizers . 34 5.3 Preservation of Limits . 35 5.4 Adjoints on Limits . 35 5.5 Freyd's adjoint functor theorem . 36 1 6 Chapter 6 38 7 Chapter 7 38 8 Abelian Categories 38 8.1 Additive Categories . 38 8.2 Abelian Categories . 38 8.3 Diagram Lemmas . 39 9 Special Limits 41 9.1 Interchange of Limits .
    [Show full text]
  • On the Complexification of the Classical Geometries And
    On the complexication of the classical geometries and exceptional numb ers April Intro duction The classical groups On R Spn R and GLn C app ear as the isometry groups of sp ecial geome tries on a real vector space In fact the orthogonal group On R represents the linear isomorphisms of arealvector space V of dimension nleaving invariant a p ositive denite and symmetric bilinear form g on V The symplectic group Spn R represents the isometry group of a real vector space of dimension n leaving invariant a nondegenerate skewsymmetric bilinear form on V Finally GLn C represents the linear isomorphisms of a real vector space V of dimension nleaving invariant a complex structure on V ie an endomorphism J V V satisfying J These three geometries On R Spn R and GLn C in GLn Rintersect even pairwise in the unitary group Un C Considering now the relativeversions of these geometries on a real manifold of dimension n leads to the notions of a Riemannian manifold an almostsymplectic manifold and an almostcomplex manifold A symplectic manifold however is an almostsymplectic manifold X ie X is a manifold and is a nondegenerate form on X so that the form is closed Similarly an almostcomplex manifold X J is called complex if the torsion tensor N J vanishes These three geometries intersect in the notion of a Kahler manifold In view of the imp ortance of the complexication of the real Lie groupsfor instance in the structure theory and representation theory of semisimple real Lie groupswe consider here the question on the underlying geometrical
    [Show full text]
  • Patterns of Maximally Entangled States Within the Algebra of Biquaternions
    J. Phys. Commun. 4 (2020) 055018 https://doi.org/10.1088/2399-6528/ab9506 PAPER Patterns of maximally entangled states within the algebra of OPEN ACCESS biquaternions RECEIVED 21 March 2020 Lidia Obojska REVISED 16 May 2020 Siedlce University of Natural Sciences and Humanities, ul. 3 Maja 54, 08-110 Siedlce, Poland ACCEPTED FOR PUBLICATION E-mail: [email protected] 20 May 2020 Keywords: bipartite entanglement, biquaternions, density matrix, mereology PUBLISHED 28 May 2020 Original content from this Abstract work may be used under This paper proposes a description of maximally entangled bipartite states within the algebra of the terms of the Creative Commons Attribution 4.0 biquaternions–Ä . We assume that a bipartite entanglement is created in a process of splitting licence. one particle into a pair of two indiscernible, yet not identical particles. As a result, we describe an Any further distribution of this work must maintain entangled correlation by the use of a division relation and obtain twelve forms of biquaternions, attribution to the author(s) and the title of representing pure maximally entangled states. Additionally, we obtain other patterns, describing the work, journal citation mixed entangled states. Finally, we show that there are no other maximally entangled states in Ä and DOI. than those presented in this work. 1. Introduction In 1935, Einstein, Podolski and Rosen described a strange phenomenon, in which in its pure state, one particle behaves like a pair of two indistinguishable, yet not identical, particles [1]. This phenomenon, called by Einstein ’a spooky action at a distance’, has been investigated by many authors, who tried to explain this strange correlation [2–9].
    [Show full text]
  • Majorana Spinors
    MAJORANA SPINORS JOSE´ FIGUEROA-O'FARRILL Contents 1. Complex, real and quaternionic representations 2 2. Some basis-dependent formulae 5 3. Clifford algebras and their spinors 6 4. Complex Clifford algebras and the Majorana condition 10 5. Examples 13 One dimension 13 Two dimensions 13 Three dimensions 14 Four dimensions 14 Six dimensions 15 Ten dimensions 16 Eleven dimensions 16 Twelve dimensions 16 ...and back! 16 Summary 19 References 19 These notes arose as an attempt to conceptualise the `symplectic Majorana{Weyl condition' in 5+1 dimensions; but have turned into a general discussion of spinors. Spinors play a crucial role in supersymmetry. Part of their versatility is that they come in many guises: `Dirac', `Majorana', `Weyl', `Majorana{Weyl', `symplectic Majorana', `symplectic Majorana{Weyl', and their `pseudo' counterparts. The tra- ditional physics approach to this topic is a mixed bag of tricks using disparate aspects of representation theory of finite groups. In these notes we will attempt to provide a uniform treatment based on the classification of Clifford algebras, a work dating back to the early 60s and all but ignored by the theoretical physics com- munity. Recent developments in superstring theory have made us re-examine the conditions for the existence of different kinds of spinors in spacetimes of arbitrary signature, and we believe that a discussion of this more uniform approach is timely and could be useful to the student meeting this topic for the first time or to the practitioner who has difficulty remembering the answer to questions like \when do symplectic Majorana{Weyl spinors exist?" The notes are organised as follows.
    [Show full text]
  • A Review of Commutative Ring Theory Mathematics Undergraduate Seminar: Toric Varieties
    A REVIEW OF COMMUTATIVE RING THEORY MATHEMATICS UNDERGRADUATE SEMINAR: TORIC VARIETIES ADRIANO FERNANDES Contents 1. Basic Definitions and Examples 1 2. Ideals and Quotient Rings 3 3. Properties and Types of Ideals 5 4. C-algebras 7 References 7 1. Basic Definitions and Examples In this first section, I define a ring and give some relevant examples of rings we have encountered before (and might have not thought of as abstract algebraic structures.) I will not cover many of the intermediate structures arising between rings and fields (e.g. integral domains, unique factorization domains, etc.) The interested reader is referred to Dummit and Foote. Definition 1.1 (Rings). The algebraic structure “ring” R is a set with two binary opera- tions + and , respectively named addition and multiplication, satisfying · (R, +) is an abelian group (i.e. a group with commutative addition), • is associative (i.e. a, b, c R, (a b) c = a (b c)) , • and the distributive8 law holds2 (i.e.· a,· b, c ·R, (·a + b) c = a c + b c, a (b + c)= • a b + a c.) 8 2 · · · · · · Moreover, the ring is commutative if multiplication is commutative. The ring has an identity, conventionally denoted 1, if there exists an element 1 R s.t. a R, 1 a = a 1=a. 2 8 2 · ·From now on, all rings considered will be commutative rings (after all, this is a review of commutative ring theory...) Since we will be talking substantially about the complex field C, let us recall the definition of such structure. Definition 1.2 (Fields).
    [Show full text]
  • CLIFFORD ALGEBRAS Property, Then There Is a Unique Isomorphism (V ) (V ) Intertwining the Two Inclusions of V
    CHAPTER 2 Clifford algebras 1. Exterior algebras 1.1. Definition. For any vector space V over a field K, let T (V ) = k k k Z T (V ) be the tensor algebra, with T (V ) = V V the k-fold tensor∈ product. The quotient of T (V ) by the two-sided⊗···⊗ ideal (V ) generated byL all v w + w v is the exterior algebra, denoted (V ).I The product in (V ) is usually⊗ denoted⊗ α α , although we will frequently∧ omit the wedge ∧ 1 ∧ 2 sign and just write α1α2. Since (V ) is a graded ideal, the exterior algebra inherits a grading I (V )= k(V ) ∧ ∧ k Z M∈ where k(V ) is the image of T k(V ) under the quotient map. Clearly, 0(V )∧ = K and 1(V ) = V so that we can think of V as a subspace of ∧(V ). We may thus∧ think of (V ) as the associative algebra linearly gener- ated∧ by V , subject to the relations∧ vw + wv = 0. We will write φ = k if φ k(V ). The exterior algebra is commutative | | ∈∧ (in the graded sense). That is, for φ k1 (V ) and φ k2 (V ), 1 ∈∧ 2 ∈∧ [φ , φ ] := φ φ + ( 1)k1k2 φ φ = 0. 1 2 1 2 − 2 1 k If V has finite dimension, with basis e1,...,en, the space (V ) has basis ∧ e = e e I i1 · · · ik for all ordered subsets I = i1,...,ik of 1,...,n . (If k = 0, we put { } k { n } e = 1.) In particular, we see that dim (V )= k , and ∅ ∧ n n dim (V )= = 2n.
    [Show full text]
  • A Natural Transformation of the Spec Functor
    View metadata, citation and similar papers at core.ac.uk brought to you by CORE provided by Elsevier - Publisher Connector JOURNAL OF ALGEBRA 10, 69-91 (1968) A Natural Transformation of the Spec Functor IAN G. CONNELL McGill University, Montreal, Canada Communicated by P. Cohn Received November 28, 1967 The basic functor underlying the Grothendieck algebraic geometry is Spec which assigns a ringed space to each ring1 A; we denote the topological space by spec A and the sheaf of rings by Spec A. We shall define a functor which is both a generalization and a natural transformation of Spec, in a sense made precise below. Our setting is the category Proj, of commutative K-algebras, where k is a fixed ring, and specializations over K which are defined below. A k-algebra is, strictly speaking, a ring homomorphism pA : K -+ A (pa is called the representation or structural homomorphism) and a K-algebra homomorphism is a commutative triangle k of ring homomorphisms. As is customary we simplify this to “A is a K-algebra” and “4 : A -+ B is a K-algebra homomorphism”, when there can be no ambiguity about the p’s. When we refer to K as a K-algebra we mean of course the identity representation. 1. PRIMES AND PLACES Recall the definition of a place $ on the field A with values in the field B (cf. [6], p. 3). It is a ring homomorphism 4 : A, -+ B defined on a subring A, of A with the property that for all X, y E A, xy E A, , x 6 A, =Py E Ker 4.
    [Show full text]
  • Arxiv:1005.0156V3
    FOUR PROBLEMS REGARDING REPRESENTABLE FUNCTORS G. MILITARU C Abstract. Let R, S be two rings, C an R-coring and RM the category of left C- C C comodules. The category Rep (RM, SM) of all representable functors RM→ S M is C shown to be equivalent to the opposite of the category RMS . For U an (S, R)-bimodule C we give necessary and sufficient conditions for the induction functor U ⊗R − : RM→ SM to be: a representable functor, an equivalence of categories, a separable or a Frobenius functor. The latter results generalize and unify the classical theorems of Morita for categories of modules over rings and the more recent theorems obtained by Brezinski, Caenepeel et al. for categories of comodules over corings. Introduction Let C be a category and V a variety of algebras in the sense of universal algebras. A functor F : C →V is called representable [1] if γ ◦ F : C → Set is representable in the classical sense, where γ : V → Set is the forgetful functor. Four general problems concerning representable functors have been identified: Problem A: Describe the category Rep (C, V) of all representable functors F : C→V. Problem B: Give a necessary and sufficient condition for a given functor F : C→V to be representable (possibly predefining the object of representability). Problem C: When is a composition of two representable functors a representable functor? Problem D: Give a necessary and sufficient condition for a representable functor F : C → V and for its left adjoint to be separable or Frobenius. The pioneer of studying problem A was Kan [10] who described all representable functors from semigroups to semigroups.
    [Show full text]
  • Math 250A: Groups, Rings, and Fields. H. W. Lenstra Jr. 1. Prerequisites
    Math 250A: Groups, rings, and fields. H. W. Lenstra jr. 1. Prerequisites This section consists of an enumeration of terms from elementary set theory and algebra. You are supposed to be familiar with their definitions and basic properties. Set theory. Sets, subsets, the empty set , operations on sets (union, intersection, ; product), maps, composition of maps, injective maps, surjective maps, bijective maps, the identity map 1X of a set X, inverses of maps. Relations, equivalence relations, equivalence classes, partial and total orderings, the cardinality #X of a set X. The principle of math- ematical induction. Zorn's lemma will be assumed in a number of exercises. Later in the course the terminology and a few basic results from point set topology may come in useful. Group theory. Groups, multiplicative and additive notation, the unit element 1 (or the zero element 0), abelian groups, cyclic groups, the order of a group or of an element, Fermat's little theorem, products of groups, subgroups, generators for subgroups, left cosets aH, right cosets, the coset spaces G=H and H G, the index (G : H), the theorem of n Lagrange, group homomorphisms, isomorphisms, automorphisms, normal subgroups, the factor group G=N and the canonical map G G=N, homomorphism theorems, the Jordan- ! H¨older theorem (see Exercise 1.4), the commutator subgroup [G; G], the center Z(G) (see Exercise 1.12), the group Aut G of automorphisms of G, inner automorphisms. Examples of groups: the group Sym X of permutations of a set X, the symmetric group S = Sym 1; 2; : : : ; n , cycles of permutations, even and odd permutations, the alternating n f g group A , the dihedral group D = (1 2 : : : n); (1 n 1)(2 n 2) : : : , the Klein four group n n h − − i V , the quaternion group Q = 1; i; j; ij (with ii = jj = 1, ji = ij) of order 4 8 { g − − 8, additive groups of rings, the group Gl(n; R) of invertible n n-matrices over a ring R.
    [Show full text]
  • Maxwell's Equations
    Maxwell’s equations Daniel Henry Gottlieb August 1, 2004 Abstract We express Maxwell’s equations as a single equation, first using the divergence of a special type of matrix field to obtain the four current, and then the divergence of a special matrix to obtain the Electromagnetic field. These two equations give rise to a remarkable dual set of equations in which the operators become the matrices and the vectors become the fields. The decoupling of the equations into the wave equation is very simple and natural. The divergence of the stress energy tensor gives the Lorentz Law in a very natural way. We compare this approach to the related descriptions of Maxwell’s equations by biquaternions and Clifford algebras. 1 Introduction Maxwell’s equations have been expressed in many forms in the century and a half since their dis- covery. The original equations were 16 in number. The vector forms, written below, consist of 4 equations. The differential form versions consists of two equations; see [Misner, Thorne and Wheeler(1973); see equations 4.10, 4.11]. See also [Steven Parrott(1987) page 98 -100 ] The ap- plication of quaternions, and their complexification, the biquaternions, results in a version of one equation. William E. Baylis (1999) equation 3.8, is an example. In this work, we obtain one Maxwell equation, (10), representing the electromagnetic field as a matrix and the divergence as a vector multiplying the field matrix. But we also obtain a remarkable dual formulation of Maxwell’s equation, (15), wherein the operator is now the matrix and the field is now the vector.
    [Show full text]