MAT301H1S Lec5101 Burbulla

Total Page:16

File Type:pdf, Size:1020Kb

MAT301H1S Lec5101 Burbulla Chapter 10: Group Homomorphisms Chapter 6: Isomorphisms MAT301H1S Lec5101 Burbulla Week 5 Lecture Notes Winter 2020 Week 5 Lecture Notes MAT301H1S Lec5101 Burbulla Chapter 10: Group Homomorphisms Chapter 6: Isomorphisms Chapter 10: Group Homomorphisms Chapter 6: Isomorphisms Week 5 Lecture Notes MAT301H1S Lec5101 Burbulla Chapter 10: Group Homomorphisms Chapter 6: Isomorphisms What is a Homomorphism? Definition: Let G and H be groups and suppose f : G −! H is a function such that for all x; y 2 G; f (xy) = f (x)f (y): Then f is called a group homomorphism, or just homomorphism for short. In words, we say \f preserves products." Note: to calculate xy we are using the operation in G; but to calculate f (x)f (y) we are using the operation in H: Week 5 Lecture Notes MAT301H1S Lec5101 Burbulla Chapter 10: Group Homomorphisms Chapter 6: Isomorphisms Examples ∗ ∗ 1. f : R −! R by f (x) = jxj f (xy) = jxyj = jxjjyj = f (x)f (y): ∗ ∗ 2 2. f : R −! R by f (x) = x f (xy) = (xy)2 = x2y 2 = f (x)f (y): 3. f : Z −! Zn by f (m) = m mod n; with addition. Use the division algorithm to write m1 = p1n + q1; m2 = p2n + q2: Then m1 + m2 = (p1 + p2)n + q1 + q2; so f (m1+m2) = (q1+q2) mod n ≡ q1 mod n+q2 mod n = f (m1)+f (m2) Week 5 Lecture Notes MAT301H1S Lec5101 Burbulla Chapter 10: Group Homomorphisms Chapter 6: Isomorphisms Not All Functions are Homomorphisms 2 Let f : R −! R by f (x) = x : The operation is addition. Then f (x + y) = (x + y)2 = x2 + 2xy + y 2 but f (x) + f (y) = x2 + y 2: So it is not true that f (x + y) = f (x) + f (y) for all x; y 2 R; so f is not a homomorphism. Week 5 Lecture Notes MAT301H1S Lec5101 Burbulla Chapter 10: Group Homomorphisms Chapter 6: Isomorphisms Some More Interesting Examples ∗ The determinant function, det : GL(n; R) −! R ; is a homomorphism, since det(AB) = det(A) det(B): Define sgn : Sn −! f1; −1g by 1 if σ is an even permutaton sgn (σ) = −1 if σ is an odd permutaton You have to check that for all α; β 2 Sn; sgn (αβ) = sgn (α) sgn (β); which boils down to checking four cases, namely even plus even is even, even plus odd is odd, odd plus even is odd, and odd plus odd is even. Week 5 Lecture Notes MAT301H1S Lec5101 Burbulla Chapter 10: Group Homomorphisms Chapter 6: Isomorphisms Properties of Homomorphisms: Th 10.1 First Three Parts Let f : G −! H be a homomorphism. Then 1. f (eG ) = eH Proof: x = xeG ) f (x) = f (xeG ) = f (x)f (eG ) ) eH = f (eG ): 2. f (xn) = (f (x))n; for all x 2 G and all n 2 Z Proof: f (x2) = f (x x) = f (x) f (x) = (f (x))2; etc for positive n: If n = −1; then −1 −1 −1 xx = eG ) f (xx ) = f (eG ) ) f (x)f (x ) = eH ) f (x−1) = (f (x))−1: 3. If jxj is finite, then jf (x)j divides jxj Proof: let jxj = n: Then n n n x = eG ) f (x ) = f (eG ) ) (f (x)) = eH ) jf (x)j divides n: Week 5 Lecture Notes MAT301H1S Lec5101 Burbulla Chapter 10: Group Homomorphisms Chapter 6: Isomorphisms Kernels and Images If f : G −! H is a homomorphism, define 1. ker(f ) = fx 2 G j f (x) = eH g 2. im (f ) = ff (x) j x 2 Gg = f (G) Then ker(f ) ≤ G and im (f ) ≤ H: Proof: (for kernel) I f (eG ) = eH ) eG 2 ker(f ): I x; y 2 ker(f ) −1 −1 −1 ) f (xy ) = f (x)f (y ) = f (x)(f (y)) = eH eH = eH ; so xy −1 2 ker(f ): I Proof that im (f ) ≤ H is left as an exercise. Week 5 Lecture Notes MAT301H1S Lec5101 Burbulla Chapter 10: Group Homomorphisms Chapter 6: Isomorphisms Examples ∗ 1. For det : GL(n; R) −! R ; ker(det) = fA 2 GL(n; R) j det(A) = 1g = SL(n; R): 2. For sgn : Sn −! f1; −1g ker(sgn ) = fσ 2 Sn j sgn (σ) = 1g = An: 3. For f : Z −! Zn by f (m) = m mod n; with addition, ker(f ) = fm j m ≡ 0 mod ng = < n > and im (f ) = fm mod n j m 2 Zg = Zn: Week 5 Lecture Notes MAT301H1S Lec5101 Burbulla Chapter 10: Group Homomorphisms Chapter 6: Isomorphisms Proper Definition of Determinant Let A = (aij ) be an n × n matrix. Here is the \proper" definition of det(A): X det(A) = sgn (σ) a1σ(1)a2σ(2) ····· anσ(n): σ2Sn For example, if n = 3; the even permutations are (1); (123); (132) and the odd permutations are (12); (23); (31): Then 2 3 a11 a12 a13 det 4 a21 a22 a23 5 = a31 a32 a33 a11a22a33+a12a23a31+a13a21a32−a12a21a33−a11a23a32−a13a22a31: Week 5 Lecture Notes MAT301H1S Lec5101 Burbulla Chapter 10: Group Homomorphisms Chapter 6: Isomorphisms Properties of Subgroups Under Homomorphisms First three parts of Theorem 10.2: let f : G −! H be a homomorphism and let K be a subgroup of G: Then 1. f (K) = ff (k) j k 2 Kg is a subgroup of H: 2. If K is cyclic then f (K) is cyclic. 3. If K is Abelian then f (K) is Abelian. Proof: 1. left as an exercise. 2. Suppose K = hki: Then x 2 K ) x = kn: Then f (x) = f (kn) = (f (k))n; which means that f (K) = hf (k)i: 3. Suppose xy = yx for all x; y 2 K: Then f (xy) = f (yx) , f (x)f (y) = f (y)f (x); which means all elements in f (K) commute. Week 5 Lecture Notes MAT301H1S Lec5101 Burbulla Chapter 10: Group Homomorphisms Chapter 6: Isomorphisms Example cos θ − sin θ Let f : −! GL(2; ) by f (θ) = : f is a R R sin θ cos θ homomorphism because f (α)f (β) = f (α + β): cos α − sin α cos β − sin β cos(α + β) − sin(α + β) = ; sin α cos α sin β cos β sin(α + β) cos(α + β using appropriate trig identities. Check that ker(f ) = h2πi and im (f ) is the group of 2 × 2 rotation matrices. In particular, R is Abelian, so im (f ) is Abelian and im (f ) is an Abelian subgroup of GL(2; R): If the positive integer n is fixed and 2π 2πk K = = j k 2 ; n n Z then f (K) = hf (2π=n)i ≤ Dn; consisting of all rotations of a regular n-gon. Week 5 Lecture Notes MAT301H1S Lec5101 Burbulla Chapter 10: Group Homomorphisms Chapter 6: Isomorphisms One-to-one and Onto Homomorphisms Let f : G −! H be a group homomorphism. Definition: f is called one-to-one if f (x1) = f (x2) ) x1 = x2: Defintion: f is called onto if every element in H is in im (f ): Theorem: let f : G −! H be a group homomorphism. I f is one-to-one if and only if ker(f ) = feG g I f is onto if and only if im (f ) = H Proof: (for one-to-one) suppose f is one-to-one: Let x 2 ker(f ): Then f (x) = eH and f (eG ) = eH : Therefore x = eG : Now suppose ker(f ) = feG g, and let f (x1) = f (x2): Then −1 −1 −1 −1 f (x1x2 ) = f (x1)f (x2 ) = f (x1)(f (x2)) = f (x1)(f (x1)) = eH : −1 −1 Thus x1x2 2 ker(f ) ) x1x2 = eG ) x1 = x2: Week 5 Lecture Notes MAT301H1S Lec5101 Burbulla Chapter 10: Group Homomorphisms Chapter 6: Isomorphisms What is an Isomorphism? Let f : G −! H be a group homomorphism. Definition: f is called an isomorphism if f is one-to-one and onto. Thus f is an isomorphism if ker(f ) = feG g and im (f ) = H: Note: every isomorphism f : G −! H has an inverse f −1 : H −! G; defined by f −1(x) = y , x = f (y); which is also an isomorphism. Definition: if there is an isomorphism f : G −! H; then we say the groups G and H are isomorphic, and we write G ≈ H: Week 5 Lecture Notes MAT301H1S Lec5101 Burbulla Chapter 10: Group Homomorphisms Chapter 6: Isomorphisms Example 1 1 m Let H = 2 GL(2; ) j m 2 and define f : −! H 0 1 R Z Z by 1 m f (m) = : 0 1 Then f is an isomorphism: 1. f is a homomorphism since 1 m + n 1 m 1 n f (m+n) = = = f (m)f (n) 0 1 0 1 0 1 1 0 2. f is one-to-one since f (m) = ) m = 0 0 1 3. f is onto since obviously im (f ) = H: Week 5 Lecture Notes MAT301H1S Lec5101 Burbulla Chapter 10: Group Homomorphisms Chapter 6: Isomorphisms Example 2 Let G = hai be a cyclic group of order n: Then G ≈ Zn: k Proof: for k 2 Z; define f : G −! Zn by f (a ) = k mod n: j k I f is well-defined: a = a ) n divides j − k ) j ≡ k mod n: j k j+k I f is a homomorphism: f (a a ) = f (a ) = (j + k) mod n; and f (aj ) + f (ak ) = j mod n + k mod n = (j + k) mod n: I f is one-to-one: k k n q f (a ) = 0 ) k ≡ 0 mod n ) k = qn ) a = (a ) = eG : I f is onto: this is obvious since k 2 Z; so im (f ) = Zn: m Alternate Proof: use g : Zn −! G defined by g(m) = a : (g = f −1:) This makes the algebra easier since g is obviously well-defined.
Recommended publications
  • Group Homomorphisms
    1-17-2018 Group Homomorphisms Here are the operation tables for two groups of order 4: · 1 a a2 + 0 1 2 1 1 a a2 0 0 1 2 a a a2 1 1 1 2 0 a2 a2 1 a 2 2 0 1 There is an obvious sense in which these two groups are “the same”: You can get the second table from the first by replacing 0 with 1, 1 with a, and 2 with a2. When are two groups the same? You might think of saying that two groups are the same if you can get one group’s table from the other by substitution, as above. However, there are problems with this. In the first place, it might be very difficult to check — imagine having to write down a multiplication table for a group of order 256! In the second place, it’s not clear what a “multiplication table” is if a group is infinite. One way to implement a substitution is to use a function. In a sense, a function is a thing which “substitutes” its output for its input. I’ll define what it means for two groups to be “the same” by using certain kinds of functions between groups. These functions are called group homomorphisms; a special kind of homomorphism, called an isomorphism, will be used to define “sameness” for groups. Definition. Let G and H be groups. A homomorphism from G to H is a function f : G → H such that f(x · y)= f(x) · f(y) forall x,y ∈ G.
    [Show full text]
  • Orbits of Automorphism Groups of Fields
    ORBITS OF AUTOMORPHISM GROUPS OF FIELDS KIRAN S. KEDLAYA AND BJORN POONEN Abstract. We address several specific aspects of the following general question: can a field K have so many automorphisms that the action of the automorphism group on the elements of K has relatively few orbits? We prove that any field which has only finitely many orbits under its automorphism group is finite. We extend the techniques of that proof to approach a broader conjecture, which asks whether the automorphism group of one field over a subfield can have only finitely many orbits on the complement of the subfield. Finally, we apply similar methods to analyze the field of Mal'cev-Neumann \generalized power series" over a base field; these form near-counterexamples to our conjecture when the base field has characteristic zero, but often fall surprisingly far short in positive characteristic. Can an infinite field K have so many automorphisms that the action of the automorphism group on the elements of K has only finitely many orbits? In Section 1, we prove that the answer is \no" (Theorem 1.1), even though the corresponding answer for division rings is \yes" (see Remark 1.2). Our proof constructs a \trace map" from the given field to a finite field, and exploits the peculiar combination of additive and multiplicative properties of this map. Section 2 attempts to prove a relative version of Theorem 1.1, by considering, for a non- trivial extension of fields k ⊂ K, the action of Aut(K=k) on K. In this situation each element of k forms an orbit, so we study only the orbits of Aut(K=k) on K − k.
    [Show full text]
  • The General Linear Group
    18.704 Gabe Cunningham 2/18/05 [email protected] The General Linear Group Definition: Let F be a field. Then the general linear group GLn(F ) is the group of invert- ible n × n matrices with entries in F under matrix multiplication. It is easy to see that GLn(F ) is, in fact, a group: matrix multiplication is associative; the identity element is In, the n × n matrix with 1’s along the main diagonal and 0’s everywhere else; and the matrices are invertible by choice. It’s not immediately clear whether GLn(F ) has infinitely many elements when F does. However, such is the case. Let a ∈ F , a 6= 0. −1 Then a · In is an invertible n × n matrix with inverse a · In. In fact, the set of all such × matrices forms a subgroup of GLn(F ) that is isomorphic to F = F \{0}. It is clear that if F is a finite field, then GLn(F ) has only finitely many elements. An interesting question to ask is how many elements it has. Before addressing that question fully, let’s look at some examples. ∼ × Example 1: Let n = 1. Then GLn(Fq) = Fq , which has q − 1 elements. a b Example 2: Let n = 2; let M = ( c d ). Then for M to be invertible, it is necessary and sufficient that ad 6= bc. If a, b, c, and d are all nonzero, then we can fix a, b, and c arbitrarily, and d can be anything but a−1bc. This gives us (q − 1)3(q − 2) matrices.
    [Show full text]
  • The Matrix Calculus You Need for Deep Learning
    The Matrix Calculus You Need For Deep Learning Terence Parr and Jeremy Howard July 3, 2018 (We teach in University of San Francisco's MS in Data Science program and have other nefarious projects underway. You might know Terence as the creator of the ANTLR parser generator. For more material, see Jeremy's fast.ai courses and University of San Francisco's Data Institute in- person version of the deep learning course.) HTML version (The PDF and HTML were generated from markup using bookish) Abstract This paper is an attempt to explain all the matrix calculus you need in order to understand the training of deep neural networks. We assume no math knowledge beyond what you learned in calculus 1, and provide links to help you refresh the necessary math where needed. Note that you do not need to understand this material before you start learning to train and use deep learning in practice; rather, this material is for those who are already familiar with the basics of neural networks, and wish to deepen their understanding of the underlying math. Don't worry if you get stuck at some point along the way|just go back and reread the previous section, and try writing down and working through some examples. And if you're still stuck, we're happy to answer your questions in the Theory category at forums.fast.ai. Note: There is a reference section at the end of the paper summarizing all the key matrix calculus rules and terminology discussed here. arXiv:1802.01528v3 [cs.LG] 2 Jul 2018 1 Contents 1 Introduction 3 2 Review: Scalar derivative rules4 3 Introduction to vector calculus and partial derivatives5 4 Matrix calculus 6 4.1 Generalization of the Jacobian .
    [Show full text]
  • IVC Factsheet Functions Comp Inverse
    Imperial Valley College Math Lab Functions: Composition and Inverse Functions FUNCTION COMPOSITION In order to perform a composition of functions, it is essential to be familiar with function notation. If you see something of the form “푓(푥) = [expression in terms of x]”, this means that whatever you see in the parentheses following f should be substituted for x in the expression. This can include numbers, variables, other expressions, and even other functions. EXAMPLE: 푓(푥) = 4푥2 − 13푥 푓(2) = 4 ∙ 22 − 13(2) 푓(−9) = 4(−9)2 − 13(−9) 푓(푎) = 4푎2 − 13푎 푓(푐3) = 4(푐3)2 − 13푐3 푓(ℎ + 5) = 4(ℎ + 5)2 − 13(ℎ + 5) Etc. A composition of functions occurs when one function is “plugged into” another function. The notation (푓 ○푔)(푥) is pronounced “푓 of 푔 of 푥”, and it literally means 푓(푔(푥)). In other words, you “plug” the 푔(푥) function into the 푓(푥) function. Similarly, (푔 ○푓)(푥) is pronounced “푔 of 푓 of 푥”, and it literally means 푔(푓(푥)). In other words, you “plug” the 푓(푥) function into the 푔(푥) function. WARNING: Be careful not to confuse (푓 ○푔)(푥) with (푓 ∙ 푔)(푥), which means 푓(푥) ∙ 푔(푥) . EXAMPLES: 푓(푥) = 4푥2 − 13푥 푔(푥) = 2푥 + 1 a. (푓 ○푔)(푥) = 푓(푔(푥)) = 4[푔(푥)]2 − 13 ∙ 푔(푥) = 4(2푥 + 1)2 − 13(2푥 + 1) = [푠푚푝푙푓푦] … = 16푥2 − 10푥 − 9 b. (푔 ○푓)(푥) = 푔(푓(푥)) = 2 ∙ 푓(푥) + 1 = 2(4푥2 − 13푥) + 1 = 8푥2 − 26푥 + 1 A function can even be “composed” with itself: c.
    [Show full text]
  • The Pure Symmetric Automorphisms of a Free Group Form a Duality Group
    THE PURE SYMMETRIC AUTOMORPHISMS OF A FREE GROUP FORM A DUALITY GROUP NOEL BRADY, JON MCCAMMOND, JOHN MEIER, AND ANDY MILLER Abstract. The pure symmetric automorphism group of a finitely generated free group consists of those automorphisms which send each standard generator to a conjugate of itself. We prove that these groups are duality groups. 1. Introduction Let Fn be a finite rank free group with fixed free basis X = fx1; : : : ; xng. The symmetric automorphism group of Fn, hereafter denoted Σn, consists of those au- tomorphisms that send each xi 2 X to a conjugate of some xj 2 X. The pure symmetric automorphism group, denoted PΣn, is the index n! subgroup of Σn of symmetric automorphisms that send each xi 2 X to a conjugate of itself. The quotient of PΣn by the inner automorphisms of Fn will be denoted OPΣn. In this note we prove: Theorem 1.1. The group OPΣn is a duality group of dimension n − 2. Corollary 1.2. The group PΣn is a duality group of dimension n − 1, hence Σn is a virtual duality group of dimension n − 1. (In fact we establish slightly more: the dualizing module in both cases is -free.) Corollary 1.2 follows immediately from Theorem 1.1 since Fn is a 1-dimensional duality group, there is a short exact sequence 1 ! Fn ! PΣn ! OPΣn ! 1 and any duality-by-duality group is a duality group whose dimension is the sum of the dimensions of its constituents (see Theorem 9.10 in [2]). That the virtual cohomological dimension of Σn is n − 1 was previously estab- lished by Collins in [9].
    [Show full text]
  • Generalized Quaternions
    GENERALIZED QUATERNIONS KEITH CONRAD 1. introduction The quaternion group Q8 is one of the two non-abelian groups of size 8 (up to isomor- phism). The other one, D4, can be constructed as a semi-direct product: ∼ ∼ × ∼ D4 = Aff(Z=(4)) = Z=(4) o (Z=(4)) = Z=(4) o Z=(2); where the elements of Z=(2) act on Z=(4) as the identity and negation. While Q8 is not a semi-direct product, it can be constructed as the quotient group of a semi-direct product. We will see how this is done in Section2 and then jazz up the construction in Section3 to make an infinite family of similar groups with Q8 as the simplest member. In Section4 we will compare this family with the dihedral groups and see how it fits into a bigger picture. 2. The quaternion group from a semi-direct product The group Q8 is built out of its subgroups hii and hji with the overlapping condition i2 = j2 = −1 and the conjugacy relation jij−1 = −i = i−1. More generally, for odd a we have jaij−a = −i = i−1, while for even a we have jaij−a = i. We can combine these into the single formula a (2.1) jaij−a = i(−1) for all a 2 Z. These relations suggest the following way to construct the group Q8. Theorem 2.1. Let H = Z=(4) o Z=(4), where (a; b)(c; d) = (a + (−1)bc; b + d); ∼ The element (2; 2) in H has order 2, lies in the center, and H=h(2; 2)i = Q8.
    [Show full text]
  • On the Group-Theoretic Properties of the Automorphism Groups of Various Graphs
    ON THE GROUP-THEORETIC PROPERTIES OF THE AUTOMORPHISM GROUPS OF VARIOUS GRAPHS CHARLES HOMANS Abstract. In this paper we provide an introduction to the properties of one important connection between the theories of groups and graphs, that of the group formed by the automorphisms of a given graph. We provide examples of important results in graph theory that can be understood through group theory and vice versa, and conclude with a treatment of Frucht's theorem. Contents 1. Introduction 1 2. Fundamental Definitions, Concepts, and Theorems 2 3. Example 1: The Orbit-Stabilizer Theorem and its Application to Graph Automorphisms 4 4. Example 2: On the Automorphism Groups of the Platonic Solid Skeleton Graphs 4 5. Example 3: A Tight Bound on the Product of the Chromatic Number and Independence Number of Vertex-Transitive Graphs 6 6. Frucht's Theorem 7 7. Acknowledgements 9 8. References 9 1. Introduction Groups and graphs are two highly important kinds of structures studied in math- ematics. Interestingly, the theory of groups and the theory of graphs are deeply connected. In this paper, we examine one particular such connection: that which emerges from the observation that the automorphisms of any given graph form a group under composition. In section 2, we provide a framework for understanding the material discussed in the paper. In sections 3, 4, and 5, we demonstrate how important results in group theory illuminate some properties of automorphism groups, how the geo- metric properties of particular embeddings of graphs can be used to determine the structure of the automorphism groups of all embeddings of those graphs, and how the automorphism group can be used to determine fundamental truths about the structure of the graph.
    [Show full text]
  • The Fundamental Homomorphism Theorem
    Lecture 4.3: The fundamental homomorphism theorem Matthew Macauley Department of Mathematical Sciences Clemson University http://www.math.clemson.edu/~macaule/ Math 4120, Modern Algebra M. Macauley (Clemson) Lecture 4.3: The fundamental homomorphism theorem Math 4120, Modern Algebra 1 / 10 Motivating example (from the previous lecture) Define the homomorphism φ : Q4 ! V4 via φ(i) = v and φ(j) = h. Since Q4 = hi; ji: φ(1) = e ; φ(−1) = φ(i 2) = φ(i)2 = v 2 = e ; φ(k) = φ(ij) = φ(i)φ(j) = vh = r ; φ(−k) = φ(ji) = φ(j)φ(i) = hv = r ; φ(−i) = φ(−1)φ(i) = ev = v ; φ(−j) = φ(−1)φ(j) = eh = h : Let's quotient out by Ker φ = {−1; 1g: 1 i K 1 i iK K iK −1 −i −1 −i Q4 Q4 Q4=K −j −k −j −k jK kK j k jK j k kK Q4 organized by the left cosets of K collapse cosets subgroup K = h−1i are near each other into single nodes Key observation Q4= Ker(φ) =∼ Im(φ). M. Macauley (Clemson) Lecture 4.3: The fundamental homomorphism theorem Math 4120, Modern Algebra 2 / 10 The Fundamental Homomorphism Theorem The following result is one of the central results in group theory. Fundamental homomorphism theorem (FHT) If φ: G ! H is a homomorphism, then Im(φ) =∼ G= Ker(φ). The FHT says that every homomorphism can be decomposed into two steps: (i) quotient out by the kernel, and then (ii) relabel the nodes via φ. G φ Im φ (Ker φ C G) any homomorphism q i quotient remaining isomorphism process (\relabeling") G Ker φ group of cosets M.
    [Show full text]
  • The Centralizer of a Group Automorphism
    View metadata, citation and similar papers at core.ac.uk brought to you by CORE provided by Elsevier - Publisher Connector JOURNAL OF ALGEBRA 54, 27-41 (1978) The Centralizer of a Group Automorphism CARLTON J. MAXSON AND KIRBY C. SWTH Texas A& M University, College Station, Texas 77843 Communicated by A. Fr6hlich Received May 31, 1971 Let G be a finite group. The structure of the near-ring C(A) of identity preserving functionsf : G + G, which commute with a given automorphism A of G, is investigated. The results are then applied to the case in which G is a finite vector space and A is an invertible linear transformation. Let A be a linear transformation on a finite-dimensional vector space V over a field F. The problem of determining the structure of the ring of linear trans- formations on V which commute with A has been studied extensively (e.g., [3, 5, 81). In this paper we consider a nonlinear analogue to this well-known problem which arises naturally in the study of automorphisms of a linear auto- maton [6]. Specifically let G be a finite group, written additively, and let A be an auto- morphism of G. If C(A) = {f: G + G j fA = Af and f (0) = 0 where 0 is the identity of G), then C(A) forms a near ring under the operations of pointwise addition and function composition. That is (C(A), +) is a group, (C(A), .) is a monoid, (f+-g)h=fh+gh for all f, g, hcC(A) and f.O=O.f=O, f~ C(A).
    [Show full text]
  • The Logic of Recursive Equations Author(S): A
    The Logic of Recursive Equations Author(s): A. J. C. Hurkens, Monica McArthur, Yiannis N. Moschovakis, Lawrence S. Moss, Glen T. Whitney Source: The Journal of Symbolic Logic, Vol. 63, No. 2 (Jun., 1998), pp. 451-478 Published by: Association for Symbolic Logic Stable URL: http://www.jstor.org/stable/2586843 . Accessed: 19/09/2011 22:53 Your use of the JSTOR archive indicates your acceptance of the Terms & Conditions of Use, available at . http://www.jstor.org/page/info/about/policies/terms.jsp JSTOR is a not-for-profit service that helps scholars, researchers, and students discover, use, and build upon a wide range of content in a trusted digital archive. We use information technology and tools to increase productivity and facilitate new forms of scholarship. For more information about JSTOR, please contact [email protected]. Association for Symbolic Logic is collaborating with JSTOR to digitize, preserve and extend access to The Journal of Symbolic Logic. http://www.jstor.org THE JOURNAL OF SYMBOLIC LOGIC Volume 63, Number 2, June 1998 THE LOGIC OF RECURSIVE EQUATIONS A. J. C. HURKENS, MONICA McARTHUR, YIANNIS N. MOSCHOVAKIS, LAWRENCE S. MOSS, AND GLEN T. WHITNEY Abstract. We study logical systems for reasoning about equations involving recursive definitions. In particular, we are interested in "propositional" fragments of the functional language of recursion FLR [18, 17], i.e., without the value passing or abstraction allowed in FLR. The 'pure," propositional fragment FLRo turns out to coincide with the iteration theories of [1]. Our main focus here concerns the sharp contrast between the simple class of valid identities and the very complex consequence relation over several natural classes of models.
    [Show full text]
  • Homomorphisms and Isomorphisms
    Lecture 4.1: Homomorphisms and isomorphisms Matthew Macauley Department of Mathematical Sciences Clemson University http://www.math.clemson.edu/~macaule/ Math 4120, Modern Algebra M. Macauley (Clemson) Lecture 4.1: Homomorphisms and isomorphisms Math 4120, Modern Algebra 1 / 13 Motivation Throughout the course, we've said things like: \This group has the same structure as that group." \This group is isomorphic to that group." However, we've never really spelled out the details about what this means. We will study a special type of function between groups, called a homomorphism. An isomorphism is a special type of homomorphism. The Greek roots \homo" and \morph" together mean \same shape." There are two situations where homomorphisms arise: when one group is a subgroup of another; when one group is a quotient of another. The corresponding homomorphisms are called embeddings and quotient maps. Also in this chapter, we will completely classify all finite abelian groups, and get a taste of a few more advanced topics, such as the the four \isomorphism theorems," commutators subgroups, and automorphisms. M. Macauley (Clemson) Lecture 4.1: Homomorphisms and isomorphisms Math 4120, Modern Algebra 2 / 13 A motivating example Consider the statement: Z3 < D3. Here is a visual: 0 e 0 7! e f 1 7! r 2 2 1 2 7! r r2f rf r2 r The group D3 contains a size-3 cyclic subgroup hri, which is identical to Z3 in structure only. None of the elements of Z3 (namely 0, 1, 2) are actually in D3. When we say Z3 < D3, we really mean is that the structure of Z3 shows up in D3.
    [Show full text]