Notes on the Symmetric Group

Total Page:16

File Type:pdf, Size:1020Kb

Notes on the Symmetric Group Notes on the symmetric group 1 Computations in the symmetric group Recall that, given a set X, the set SX of all bijections from X to itself (or, more briefly, permutations of X) is group under function composition. In particular, for each n 2 N, the symmetric group Sn is the group of per- mutations of the set f1; : : : ; ng, with the group operation equal to function composition. Thus Sn is a group with n! elements, and it is not abelian if n ≥ 3. If X is a finite set with #(X) = n, then any labeling of the elements of X as x1; : : : ; xn defines an isomorphism from SX to Sn (see the homework for a more precise statement). We will write elements of Sn using Greek let- ters σ; τ; ρ, : : : , we denote the identity function by 1, and we just use the multiplication symbol · or juxtaposition to denote composition (instead of the symbol ◦). Recall however that functions work from right to left: thus στ(i) = σ(τ(i)), in other words evaluate first τ at i, and then evaluate σ on this. We say that σ moves i if σ(i) 6= i, and that σ fixes i if σ(i) = i. There are many interesting subgroups of Sn. For example, the subset Hn defined by Hn = fσ 2 Sn : σ(n) = ng is easily checked to be a subgroup of Sn isomorphic to Sn−1 (see the home- work for a generalization of this). If n = n1 + n2 for two positive integers n1; n2 then the subset H = fσ 2 Sn : σ(f1; : : : ; n1g) = f1; : : : ; n1gg is also a subgroup of Sn. Note that, if σ 2 H, then automatically σ(fn1 + 1; : : : ; n2g) = fn1 + 1; : : : ; n2g; and in fact it is easy to check that H is isomorphic to Sn1 × Sn2 . There are many other subgroups of Sn. For example, the dihedral group Dn, the group of symmetries of a regular n-gon in the plane, is (isomorphic to) a subgroup of Sn by looking at the permutation of the vertices of the n-gon. 1 Note that #(Dn) = 2n, and hence that Dn is a proper subgroup of Sn if n ≥ 4. We shall see (Cayley's theorem) that, if G is a finite group, then there exists an n such that G is isomorphic to a subgroup of Sn (and in fact one can take n = #(G)). Thus the groups Sn are as complicated as all possible finite groups. To describe a function σ : f1; : : : ; ng ! f1; : : : ; ng, not necessarily a per- mutation, we can give a table of its values, recording i and then σ(i), as follows: i 1 2 ... n σ(i) σ(1) σ(2) ... σ(n) Of course, we could describe the same information by a 2 × n matrix: 1 2 : : : n : σ(1) σ(2) : : : σ(n) The condition that σ is a permutation is then the statement that the integers 1; : : : ; n each occur exactly once in the second row of the matrix. For example, if σ is given by the matrix 1 2 3 4 5 6 7 8 ; 2 4 8 1 5 7 3 6 then σ(1) = 2 and σ(5) = 5. This notation is however cumbersome and not well suited to calculation. We describe a much more efficient way to write down elements of Sn by first writing down some special ones, called cycles, and then showing that very element of Sn can be factored into a product of cycles, in an essentially unique way if we are careful. Definition 1.1. Let fa1; : : : ; akg be a subset of f1; : : : ; ng with exactly k elements; equivalently, a1; : : : ; ak are distinct. We denote by (a1; : : : ; ak) the following element of Sn: (a1; : : : ; ak)(ai) = ai+1; if i < k; (a1; : : : ; ak)(ak) = a1; (a1; : : : ; ak)(j) = j; if j 6= ai for any i. We call (a1; : : : ; ak) a k-cycle or cycle of length k. For k > 1, we define the support Supp(a1; : : : ; ak) to be the set fa1; : : : ; akg. Note that, again for k > 1, the k-cycle (a1; : : : ; ak) moves i () i 2 Supp(a1; : : : ; ak). Two cycles (a1; : : : ; ak) and (b1; : : : ; b`) are disjoint if Supp(a1; : : : ; ak) \ Supp(b1; : : : ; b`) = ;; i.e. the sets fa1; : : : ; akg and fb1; : : : ; b`g are disjoint. 2 Remark 1.2. 1) A 1-cycle (a1) is the identity function 1, no matter what a1 is. For this reason, we will generally only consider cycles of length at least 2. If σ = (a1; : : : ; ak) with k ≥ 2, then σ is never the identity, since σ(a1) = a2 6= a1. 2) A 2-cycle (a1; a2) is also called a transposition. It is the unique permu- tation of f1; : : : ; ng which switches a1 and a2 and leaves all other elements alone. 3) From the description in 2), it is clear that (a1; a2) = (a2; a1). But for k ≥ 3, the order of the elements a1; : : : ; ak is important: for example σ1 = (1; 3; 2) 6= σ2 = (1; 2; 3), because σ1(1) = 3 but σ2(1) = 2. 4) However, there are a few ways to change the order without changing the element of Sn: clearly (a1; a2; : : : ; ak) = (a2; a3; : : : ; ak; a1) = (a3; a4; : : : ; ak; a1; a2) = ··· = (ak; a1; : : : ; ak−2; ak−1): In other words, you can start the cycle anywhere, at ai, say, but then you have to list the elements in order: the next one must be ai+1, and so on, with the understanding that once you reach ak, the next one has to be a1, then a2, and then so on up to ai−1. Clearly, this the only way you can change the order. By convention, we often start with the smallest ai. Of course, after that, there is no constraint on the sizes of the consecutive members of the cycle. −1 5) It is easy to see that the inverse (a1; a2; : : : ; ak) = (ak; ak−1; : : : ; a1). In other words, the inverse of the k-cycle (a1; a2; : : : ; ak) is the k-cycle where the elements are written in the opposite order. In particular, for a transposition −1 (a1; a2), (a1; a2) = (a2; a1) = (a1; a2), i.e. a transposition has order 2. 6) Generalizing the last line of 5), it is easy to see that the order of a k-cycle σ = (a1; a2; : : : ; ak) is exactly k. In fact, σ(ai) = ai+1, so that r σ (ai) = ai+r, with the understanding that the addition i + r is to be taken mod k, but using the representatives 1; : : : ; k for addition mod k instead of the more usual ones (in this course) of 0; : : : ; k − 1. In particular, we see r r that σ (ai) = ai for all i () r is a multiple of k, and since σ (j) = j for r j 6= ai, we see that k is the smallest positive integer r such that σ = 1. Note however that, if σ is a k-cycle, its powers σr need not be k-cycles. For example, (1; 2; 3; 4)2 = (1; 3)(2; 4); and (1; 3)(2; 4) is not a k-cycle for any k. 3 7) Suppose that σ1 = (a1; : : : ; ak) and σ2 = (b1; : : : ; b`) are two disjoint cy- cles. Then it is easy to see that σ1σ2 = σ2σ1, i.e. \disjoint cycles commute." To check this we have to check that, for all j 2 f1; : : : ; ng, σ1σ2(j) = σ2σ1(j). First, if j = ai for some i, then σ1(ai) = ai+1 (with our usual conventions on adding mod k) but σ2(ai) = ai since ai 6= br for any r. For the same reason, σ2(ai+1) = ai+1. Thus, for all i with 1 ≤ i ≤ k, σ1σ2(ai) = σ1(σ2(ai)) = σ1(ai) = ai+1; whereas σ2σ1(ai) = σ2(σ1(ai)) = σ2(ai+1) = ai+1 = σ1σ2(ai): Similarly, σ1σ2(br) = br+1 = σ2σ1(br) for all r, 1 ≤ r ≤ `. Finally, if j is not an ai or a br for any i; r, then σ1σ2(j) = σ1(σ2(j)) = σ1(j) = j; and similarly σ2σ1(j) = j. Thus, for all possible j 2 f1; : : : ; ng, σ1σ2(j) = σ2σ1(j), hence σ1σ2 = σ2σ1. Note that non-disjoint cycles might or might not commute. For example, (1; 2)(1; 3) = (1; 3; 2) 6= (1; 2; 3) = (1; 3)(1; 2); whereas (1; 2; 3; 4; 5)(1; 3; 5; 2; 4) = (1; 4; 2; 5; 3) = (1; 3; 5; 2; 4)(1; 2; 3; 4; 5): 8) Finally, we have the beautiful formula: if σ 2 Sn is an arbitrary element (not necessarily a k-cycle), and (a1; : : : ; ak) is a k-cycle, then −1 σ · (a1; : : : ; ak) · σ = (σ(a1); : : : ; σ(ak)): −1 In other words, σ · (a1; : : : ; ak) · σ is again a k-cycle, but it is the k-cycle where the elements a1; : : : ; ak have been \renamed" by σ. To prove this formula, it suffices to check that −1 σ · (a1; : : : ; ak) · σ (j) = (σ(a1); : : : ; σ(ak))(j) for every j 2 f1; : : : ; ng. First, if j is of the form σ(ai) for some i, then by definition (σ(a1); : : : ; σ(ak))(σ(ai)) = σ(ai+1), with the usual remark that if i = k then we interpret k + 1 as 1. On the other hand, −1 −1 σ · (a1; : : : ; ak) · σ (σ(ai)) = σ · (a1; : : : ; ak)(σ (σ(ai)) = σ · (a1; : : : ; ak)(ai) = σ((a1; : : : ; ak)(ai)) = σ(ai+1): 4 Thus both sides agree if j = σ(ai) for some i.
Recommended publications
  • 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]
  • On Some Generation Methods of Finite Simple Groups
    Introduction Preliminaries Special Kind of Generation of Finite Simple Groups The Bibliography On Some Generation Methods of Finite Simple Groups Ayoub B. M. Basheer Department of Mathematical Sciences, North-West University (Mafikeng), P Bag X2046, Mmabatho 2735, South Africa Groups St Andrews 2017 in Birmingham, School of Mathematics, University of Birmingham, United Kingdom 11th of August 2017 Ayoub Basheer, North-West University, South Africa Groups St Andrews 2017 Talk in Birmingham Introduction Preliminaries Special Kind of Generation of Finite Simple Groups The Bibliography Abstract In this talk we consider some methods of generating finite simple groups with the focus on ranks of classes, (p; q; r)-generation and spread (exact) of finite simple groups. We show some examples of results that were established by the author and his supervisor, Professor J. Moori on generations of some finite simple groups. Ayoub Basheer, North-West University, South Africa Groups St Andrews 2017 Talk in Birmingham Introduction Preliminaries Special Kind of Generation of Finite Simple Groups The Bibliography Introduction Generation of finite groups by suitable subsets is of great interest and has many applications to groups and their representations. For example, Di Martino and et al. [39] established a useful connection between generation of groups by conjugate elements and the existence of elements representable by almost cyclic matrices. Their motivation was to study irreducible projective representations of the sporadic simple groups. In view of applications, it is often important to exhibit generating pairs of some special kind, such as generators carrying a geometric meaning, generators of some prescribed order, generators that offer an economical presentation of the group.
    [Show full text]
  • Arxiv:Hep-Th/0510191V1 22 Oct 2005
    Maximal Subgroups of the Coxeter Group W (H4) and Quaternions Mehmet Koca,∗ Muataz Al-Barwani,† and Shadia Al-Farsi‡ Department of Physics, College of Science, Sultan Qaboos University, PO Box 36, Al-Khod 123, Muscat, Sultanate of Oman Ramazan Ko¸c§ Department of Physics, Faculty of Engineering University of Gaziantep, 27310 Gaziantep, Turkey (Dated: September 8, 2018) The largest finite subgroup of O(4) is the noncrystallographic Coxeter group W (H4) of order 14400. Its derived subgroup is the largest finite subgroup W (H4)/Z2 of SO(4) of order 7200. Moreover, up to conjugacy, it has five non-normal maximal subgroups of orders 144, two 240, 400 and 576. Two groups [W (H2) × W (H2)]×Z4 and W (H3)×Z2 possess noncrystallographic structures with orders 400 and 240 respectively. The groups of orders 144, 240 and 576 are the extensions of the Weyl groups of the root systems of SU(3)×SU(3), SU(5) and SO(8) respectively. We represent the maximal subgroups of W (H4) with sets of quaternion pairs acting on the quaternionic root systems. PACS numbers: 02.20.Bb INTRODUCTION The noncrystallographic Coxeter group W (H4) of order 14400 generates some interests [1] for its relevance to the quasicrystallographic structures in condensed matter physics [2, 3] as well as its unique relation with the E8 gauge symmetry associated with the heterotic superstring theory [4]. The Coxeter group W (H4) [5] is the maximal finite subgroup of O(4), the finite subgroups of which have been classified by du Val [6] and by Conway and Smith [7].
    [Show full text]
  • Orders on Computable Torsion-Free Abelian Groups
    Orders on Computable Torsion-Free Abelian Groups Asher M. Kach (Joint Work with Karen Lange and Reed Solomon) University of Chicago 12th Asian Logic Conference Victoria University of Wellington December 2011 Asher M. Kach (U of C) Orders on Computable TFAGs ALC 2011 1 / 24 Outline 1 Classical Algebra Background 2 Computing a Basis 3 Computing an Order With A Basis Without A Basis 4 Open Questions Asher M. Kach (U of C) Orders on Computable TFAGs ALC 2011 2 / 24 Torsion-Free Abelian Groups Remark Disclaimer: Hereout, the word group will always refer to a countable torsion-free abelian group. The words computable group will always refer to a (fixed) computable presentation. Definition A group G = (G : +; 0) is torsion-free if non-zero multiples of non-zero elements are non-zero, i.e., if (8x 2 G)(8n 2 !)[x 6= 0 ^ n 6= 0 =) nx 6= 0] : Asher M. Kach (U of C) Orders on Computable TFAGs ALC 2011 3 / 24 Rank Theorem A countable abelian group is torsion-free if and only if it is a subgroup ! of Q . Definition The rank of a countable torsion-free abelian group G is the least κ cardinal κ such that G is a subgroup of Q . Asher M. Kach (U of C) Orders on Computable TFAGs ALC 2011 4 / 24 Example The subgroup H of Q ⊕ Q (viewed as having generators b1 and b2) b1+b2 generated by b1, b2, and 2 b1+b2 So elements of H look like β1b1 + β2b2 + α 2 for β1; β2; α 2 Z.
    [Show full text]
  • 14.2 Covers of Symmetric and Alternating Groups
    Remark 206 In terms of Galois cohomology, there an exact sequence of alge- braic groups (over an algebrically closed field) 1 → GL1 → ΓV → OV → 1 We do not necessarily get an exact sequence when taking values in some subfield. If 1 → A → B → C → 1 is exact, 1 → A(K) → B(K) → C(K) is exact, but the map on the right need not be surjective. Instead what we get is 1 → H0(Gal(K¯ /K), A) → H0(Gal(K¯ /K), B) → H0(Gal(K¯ /K), C) → → H1(Gal(K¯ /K), A) → ··· 1 1 It turns out that H (Gal(K¯ /K), GL1) = 1. However, H (Gal(K¯ /K), ±1) = K×/(K×)2. So from 1 → GL1 → ΓV → OV → 1 we get × 1 1 → K → ΓV (K) → OV (K) → 1 = H (Gal(K¯ /K), GL1) However, taking 1 → µ2 → SpinV → SOV → 1 we get N × × 2 1 ¯ 1 → ±1 → SpinV (K) → SOV (K) −→ K /(K ) = H (K/K, µ2) so the non-surjectivity of N is some kind of higher Galois cohomology. Warning 207 SpinV → SOV is onto as a map of ALGEBRAIC GROUPS, but SpinV (K) → SOV (K) need NOT be onto. Example 208 Take O3(R) =∼ SO3(R)×{±1} as 3 is odd (in general O2n+1(R) =∼ SO2n+1(R) × {±1}). So we have a sequence 1 → ±1 → Spin3(R) → SO3(R) → 1. 0 × Notice that Spin3(R) ⊆ C3 (R) =∼ H, so Spin3(R) ⊆ H , and in fact we saw that it is S3. 14.2 Covers of symmetric and alternating groups The symmetric group on n letter can be embedded in the obvious way in On(R) as permutations of coordinates.
    [Show full text]
  • 324 COHOMOLOGY of DIHEDRAL GROUPS of ORDER 2P Let D Be A
    324 PHYLLIS GRAHAM [March 2. P. M. Cohn, Unique factorization in noncommutative power series rings, Proc. Cambridge Philos. Soc 58 (1962), 452-464. 3. Y. Kawada, Cohomology of group extensions, J. Fac. Sci. Univ. Tokyo Sect. I 9 (1963), 417-431. 4. J.-P. Serre, Cohomologie Galoisienne, Lecture Notes in Mathematics No. 5, Springer, Berlin, 1964, 5. , Corps locaux et isogênies, Séminaire Bourbaki, 1958/1959, Exposé 185 6. , Corps locaux, Hermann, Paris, 1962. UNIVERSITY OF MICHIGAN COHOMOLOGY OF DIHEDRAL GROUPS OF ORDER 2p BY PHYLLIS GRAHAM Communicated by G. Whaples, November 29, 1965 Let D be a dihedral group of order 2p, where p is an odd prime. D is generated by the elements a and ]3 with the relations ap~fi2 = 1 and (ial3 = or1. Let A be the subgroup of D generated by a, and let p-1 Aoy Ai, • • • , Ap-\ be the subgroups generated by j8, aft, • * • , a /3, respectively. Let M be any D-module. Then the cohomology groups n n H (A0, M) and H (Aif M), i=l, 2, • • • , p — 1 are isomorphic for 1 l every integer n, so the eight groups H~ (Df Af), H°(D, M), H (D, jfef), 2 1 Q # (A Af), ff" ^. M), H (A, M), H-Wo, M), and H»(A0, M) deter­ mine all cohomology groups of M with respect to D and to all of its subgroups. We have found what values this array takes on as M runs through all finitely generated -D-modules. All possibilities for the first four members of this array are deter­ mined as a special case of the results of Yang [4].
    [Show full text]
  • Dihedral Groups Ii
    DIHEDRAL GROUPS II KEITH CONRAD We will characterize dihedral groups in terms of generators and relations, and describe the subgroups of Dn, including the normal subgroups. We will also introduce an infinite group that resembles the dihedral groups and has all of them as quotient groups. 1. Abstract characterization of Dn −1 −1 The group Dn has two generators r and s with orders n and 2 such that srs = r . We will show every group with a pair of generators having properties similar to r and s admits a homomorphism onto it from Dn, and is isomorphic to Dn if it has the same size as Dn. Theorem 1.1. Let G be generated by elements x and y where xn = 1 for some n ≥ 3, 2 −1 −1 y = 1, and yxy = x . There is a surjective homomorphism Dn ! G, and if G has order 2n then this homomorphism is an isomorphism. The hypotheses xn = 1 and y2 = 1 do not mean x has order n and y has order 2, but only that their orders divide n and divide 2. For instance, the trivial group has the form hx; yi where xn = 1, y2 = 1, and yxy−1 = x−1 (take x and y to be the identity). Proof. The equation yxy−1 = x−1 implies yxjy−1 = x−j for all j 2 Z (raise both sides to the jth power). Since y2 = 1, we have for all k 2 Z k ykxjy−k = x(−1) j by considering even and odd k separately. Thus k (1.1) ykxj = x(−1) jyk: This shows each product of x's and y's can have all the x's brought to the left and all the y's brought to the right.
    [Show full text]
  • Representations of the Infinite Symmetric Group
    Representations of the infinite symmetric group Alexei Borodin1 Grigori Olshanski2 Version of February 26, 2016 1 Department of Mathematics, Massachusetts Institute of Technology, Cambridge, MA, USA Institute for Information Transmission Problems of the Russian Academy of Sciences, Moscow, Russia 2 Institute for Information Transmission Problems of the Russian Academy of Sciences, Moscow, Russia National Research University Higher School of Economics, Moscow, Russia Contents 0 Introduction 4 I Symmetric functions and Thoma's theorem 12 1 Symmetric group representations 12 Exercises . 19 2 Theory of symmetric functions 20 Exercises . 31 3 Coherent systems on the Young graph 37 The infinite symmetric group and the Young graph . 37 Coherent systems . 39 The Thoma simplex . 41 1 CONTENTS 2 Integral representation of coherent systems and characters . 44 Exercises . 46 4 Extreme characters and Thoma's theorem 47 Thoma's theorem . 47 Multiplicativity . 49 Exercises . 52 5 Pascal graph and de Finetti's theorem 55 Exercises . 60 6 Relative dimension in Y 61 Relative dimension and shifted Schur polynomials . 61 The algebra of shifted symmetric functions . 65 Modified Frobenius coordinates . 66 The embedding Yn ! Ω and asymptotic bounds . 68 Integral representation of coherent systems: proof . 71 The Vershik{Kerov theorem . 74 Exercises . 76 7 Boundaries and Gibbs measures on paths 82 The category B ............................. 82 Projective chains . 83 Graded graphs . 86 Gibbs measures . 88 Examples of path spaces for branching graphs . 91 The Martin boundary and Vershik{Kerov's ergodic theorem . 92 Exercises . 94 II Unitary representations 98 8 Preliminaries and Gelfand pairs 98 Exercises . 108 9 Spherical type representations 111 10 Realization of spherical representations 118 Exercises .
    [Show full text]
  • The Classical Groups and Domains 1. the Disk, Upper Half-Plane, SL 2(R
    (June 8, 2018) The Classical Groups and Domains Paul Garrett [email protected] http:=/www.math.umn.edu/egarrett/ The complex unit disk D = fz 2 C : jzj < 1g has four families of generalizations to bounded open subsets in Cn with groups acting transitively upon them. Such domains, defined more precisely below, are bounded symmetric domains. First, we recall some standard facts about the unit disk, the upper half-plane, the ambient complex projective line, and corresponding groups acting by linear fractional (M¨obius)transformations. Happily, many of the higher- dimensional bounded symmetric domains behave in a manner that is a simple extension of this simplest case. 1. The disk, upper half-plane, SL2(R), and U(1; 1) 2. Classical groups over C and over R 3. The four families of self-adjoint cones 4. The four families of classical domains 5. Harish-Chandra's and Borel's realization of domains 1. The disk, upper half-plane, SL2(R), and U(1; 1) The group a b GL ( ) = f : a; b; c; d 2 ; ad − bc 6= 0g 2 C c d C acts on the extended complex plane C [ 1 by linear fractional transformations a b az + b (z) = c d cz + d with the traditional natural convention about arithmetic with 1. But we can be more precise, in a form helpful for higher-dimensional cases: introduce homogeneous coordinates for the complex projective line P1, by defining P1 to be a set of cosets u 1 = f : not both u; v are 0g= × = 2 − f0g = × P v C C C where C× acts by scalar multiplication.
    [Show full text]
  • Semigroups and Monoids 
    S Luis Alonso-Ovalle // Contents Subgroups Semigroups and monoids Subgroups Groups. A group G is an algebra consisting of a set G and a single binary operation ◦ satisfying the following axioms: . ◦ is completely defined and G is closed under ◦. ◦ is associative. G contains an identity element. Each element in G has an inverse element. Subgroups. We define a subgroup G0 as a subalgebra of G which is itself a group. Examples: . The group of even integers with addition is a proper subgroup of the group of all integers with addition. The group of all rotations of the square h{I, R, R0, R00}, ◦i, where ◦ is the composition of the operations is a subgroup of the group of all symmetries of the square. Some non-subgroups: SEMIGROUPS AND MONOIDS . The system h{I, R, R0}, ◦i is not a subgroup (and not even a subalgebra) of the original group. Why? (Hint: ◦ closure). The set of all non-negative integers with addition is a subalgebra of the group of all integers with addition, because the non-negative integers are closed under addition. But it is not a subgroup because it is not itself a group: it is associative and has a zero, but . does any member (except for ) have an inverse? Order. The order of any group G is the number of members in the set G. The order of any subgroup exactly divides the order of the parental group. E.g.: only subgroups of order , , and are possible for a -member group. (The theorem does not guarantee that every subset having the proper number of members will give rise to a subgroup.
    [Show full text]
  • 6 Permutation Groups
    Arkansas Tech University MATH 4033: Elementary Modern Algebra Dr. Marcel B. Finan 6 Permutation Groups Let S be a nonempty set and M(S) be the collection of all mappings from S into S. In this section, we will emphasize on the collection of all invertible mappings from S into S. The elements of this set will be called permutations because of Theorem 2.4 and the next definition. Definition 6.1 Let S be a nonempty set. A one-to-one mapping from S onto S is called a permutation. Consider the collection of all permutations on S. Then this set is a group with respect to composition. Theorem 6.1 The set of all permutations of a nonempty set S is a group with respect to composition. This group is called the symmetric group on S and will be denoted by Sym(S). Proof. By Theorem 2.4, the set of all permutations on S is just the set I(S) of all invertible mappings from S to S. According to Theorem 4.3, this set is a group with respect to composition. Definition 6.2 A group of permutations , with composition as the operation, is called a permutation group on S. Example 6.1 1. Sym(S) is a permutation group. 2. The collection L of all invertible linear functions from R to R is a permu- tation group with respect to composition.(See Example 4.4.) Note that L is 1 a proper subset of Sym(R) since we can find a function in Sym(R) which is not in L, namely, the function f(x) = x3.
    [Show full text]
  • Countable Torsion-Free Abelian Groups
    COUNTABLE TORSION-FREE ABELIAN GROUPS By M. O'N. CAMPBELL [Received 27 January 1959.—Read 19 February 1959] Introduction IN this paper we present a new classification of the countable torsion-free abelian groups. Since any abelian group 0 may be regarded as an extension of a periodic group (namely the torsion part of G) by a torsion-free group, and since there exists the well-known complete classification of the count- able periodic abelian groups due to Ulm, it is clear that the classification problem studied here is of great interest. By a group we shall always mean an abelian group, and the additive notation will be employed throughout. It is well known that all the maximal linearly independent sets of elements of a given group are car- dinally equivalent; the corresponding cardinal number is the rank of the group. Derry (2) and Mal'cev (4) have studied the torsion-free groups of finite rank only, and have given classifications of these groups in terms of certain equivalence classes of systems of matrices with £>-adic elements. Szekeres (5) attempted to abolish the finiteness restriction on rank; he extracted certain invariants, but did not derive a complete classification. By a basis of a torsion-free group 0 we mean any maximal linearly independent subset of G. A subgroup generated by a basis of G will be called a basal subgroup or described as basal in G. Thus U is a basal sub- group of G if and only if (i) U is free abelian, and (ii) the factor group G/U is periodic.
    [Show full text]