15 Generated Subgroups. Direct Product of Groups

Total Page:16

File Type:pdf, Size:1020Kb

15 Generated Subgroups. Direct Product of Groups Arkansas Tech University MATH 4033: Elementary Modern Algebra Dr. Marcel B. Finan 15 Generated Groups. Direct Product In this section, we discuss two procedures of building (sub)groups, namely, the (sub)groups generated by a subset of a given group and the direct product of two or more groups. 15.1 Finitely and Infinitely Generated Groups The concept of generators can be extended from cyclic groups < a > to more complicated situations where a subgroup is generated by more than one element. It is easy to see that a cyclic group with generator a is the smallest subgroup containing the set S = {a}. Can we extend groups generated by sets with more than one element? The answer is affirmative as suggested by the following theorem. Theorem 15.1 Let G be a group and C be the collection all subgroups of G. Then ∩H∈CH is a subgroup of G. Proof. Let K = ∩H∈CH. Since e belongs to all the subgroups of G then e ∈ K so that K 6= ∅. Now, let a, b ∈ K. Then a ∈ H and b ∈ H for all subgroups H of G. But then a ∈ H and b−1 ∈ H, ∀H ∈ C. Since every H in C is closed under multiplication then ab−1 ∈ H, ∀H ∈ C. That is, ab−1 ∈ K. By Theorem 7.5, K is a subgroup of G. Theorem 15.2 Let G be a group and S ⊆ G. Let C be the collection of all subgroups of G containing S. Then the set < S >= ∩H∈CH satisfies the following: (i) < S > is a subgroup of G containing S. (ii) For every H ∈ C, < S >⊆ H. Thus, < S > is the smallest subgroup of G containing S. 1 Proof. (i) Note that since S ⊆ G then G ∈ C. Thus, C 6= ∅. The fact that < S > is a subgroup of G follows from Theorem 15.1. Since S ⊆ H for all H ∈ C then S ⊆ ∩H∈CH. That is, S ⊆< S > . (ii) Let H ∈ C. If x ∈< S > then x ∈ ∩H∈CH and in particular, x ∈ H. Hence, < S >⊆ H. Theorem 15.3 < S > is the only subgroup of G that satisfies conditions (i) and (ii) of Theorem 15.2 Proof. Suppose that K is a subgroup of G satisfying conditions (i) and (ii) of the previous theorem. We will show that K =< S > . Since < S > is a subgroup containing S then by (ii), we have < S >⊆ K. Also, since K is a subgroup of G containing S then by condition (ii) again, we have K ⊆< S > . Thus, < S >= K. Definition 15.1 If G is a group and S ⊆ G then the subgroup < S > is called the subgroup of G generated by S. If G =< S >, then we say S generates G; and the elements in S are called generators. For S = {a1, a2, ··· , an} we write < S >=< a1, a2, ··· , an > . Example 15.1 Consider the subgroup of S4 generated by S = {(1432), (24)}. Then < S >= {(1), (1432), (24), (14)(23), (1234), (12)(34), (13)(24), (13)}. The following theorem characterizes the elements of < S > . Theorem 15.4 Let G be a group and S ⊆ G. Then < S > consists of finite products of ele- s1 s2 sk ments of S and inverses of elements of S. That is, if K = {a1 a2 ··· ak , ai ∈ S, si = ±1} then < S >= K. Proof. The proof is by double inclusions. By closure, K ⊆< S > . It remains to show that < S >⊆ K. Since each element of S is in K then S ⊆ K. 2 s1 s2 sk We will show that K is a subgroup of G. Indeed, if x = a1 a2 ··· ak and t1 t2 tm y = b1 b2 ··· bm are two elements in K then −1 s1 s2 sk t1 t2 tm −1 xy = (a1 a2 ··· ak )(b1 b2 ··· bm ) s1 s2 sk −tm −t2 −t1 = (a1 a2 ··· ak )(bm ··· b2 b1 ) s1 s2 sk −tm −t2 −t1 = a1 a2 ··· ak bm ··· b2 b1 Thus, xy−1 is a product of elements of S and/or inverses of elements of S. Hence, xy−1 ∈ K. By Theorem 7.5, K is a subgroup of G. By Theo- rem 15.2(ii), < S >⊆ K. Hence, < S >= K. The next theorem provides a condition under which two groups generated by different sets are equal. Theorem 15.5 Let T1 and T2 be subsets of a group G. Then < T1 >=< T2 > if and only if T1 ⊆< T2 > and T2 ⊆< T1 > Proof. Suppose first that < T1 >=< T2 > . Since T1 ⊆< T1 > and < T1 >=< T2 > then T1 ⊆< T2 > . Similarly, T2 ⊆< T1 > . Conversely, suppose that T1 ⊆< T2 > and T2 ⊆< T1 > . By Theorem 15.2(ii), we have < T1 >⊆< T2 > and < T2 >⊆< T1 > . Hence, < T1 >=< T2 > . Example 15.2 Consider (Z, +). Since {3} ⊆< 9, 12 > (3 = 12 + (−9)) and {9, 12} ⊆< 3 > then by the previous theorem we have < 9, 12 >=< 3 > . Example 15.3 In S4 we have that (124) = (123)(12)(34)(123) (234) = (132)(12)(34) = (123)−1(12)(34) (123) = (124)(234) (12)(34) = (234)(124)−1 Thus, we conclude that {(124), (134)} ⊆< (123), (12)(34) > and {(123), (12)(34)} ⊆< (124), (134) > . Hence, by Theorem 15.5, we have < (124), (234) >=< (123), (12)(34) > . 3 15.2 Direct Product of Groups. In this section we keep building examples of groups. By defining a binary operation on the Cartesian product of two groups we obtain the group known as the direct product of the two groups. Let H and K be two arbitrary groups and let H × K be the Cartesian product of H and K. In set-builder notation H × K = {(h, k): h ∈ H and k ∈ K}. Equality in H × K is defined by (h, k) = (h0, k0) if and only if h = h0 and k = k0. Define multiplication on H × K as follows: (h1, k1)(h2, k2) = (h1h2, k1k2) The set H × B is a group under the above multiplication as shown in the next theorem. Theorem 15.6 Multiplication on H × K satisfies the following properties: (i) H × K is closed under multiplication. (ii) Multiplication is associative. (iii) (eH , eK ) is the identity element. (iv) For each (h, k) ∈ H × K we have (h, k)−1 = (h−1, k−1). Proof. (i) We must show that the rule on (H ×K)×(H ×K) −→ H ×K defined by 0 0 0 0 0 0 ((h, k), (h , k )) −→ (hh , kk ) is a mapping. Indeed, if ((h1, k1), (h1, k1)) = 0 0 0 0 0 0 ((h2, k2), (h2, k2)) then (h1, k1) = (h2, k2) and (h1, k1) = (h2, k2). Thus, h1 = 0 0 0 0 0 0 0 0 h2, k1 = k2, h1 = h2, and k1 = k2. Therefore, h1h1 = h2h2 and k1k1 = k2k2. 0 0 0 0 Hence, (h1h1, k1k1) = (h2h2, k2k2). So multiplication is a binary operation. (ii) Multiplication on H × K is associative since multiplication is associative as operation on H and K. (h1, k1)[(h2, k2)(h3, k3)] = (h1, k1)(h2h3, k2k3) = (h1(h2h3), k1(k2k3)) = ((h1h2)h3, (k1k2)k3) = (h1h2, k1k2)(h3, k3) = [(h1, k1)(h2, k2)](h3, k3) 4 (iii) Let (h, k) ∈ H × K. Since heH = eH h = h and keK = eK k = k then (heH , keK ) = (eH h, eK k) = (h, k). Hence, (h, k)(eH , eK ) = (eH , eK )(h, k) = (h, k). That is, (eH , eK ) is the identity element of H × K. −1 −1 −1 −1 (iv) Let (h, k) ∈ H × K. Since hh = h h = eH and kk = k k = eK −1 −1 −1 −1 then (hh , kk ) = (eH , eK ). That is, (h, k)(h , k ) = (eH , eK ). Similarly, −1 −1 −1 −1 (h , k )(h, k) = (eH , eK ) so that (h, k) is invertible with inverse (h , k ) Definition 15.2 By the above theorem, H × K is a group, called the direct product of the groups H and K. Example 15.4 If Z3 = {[0], [1], [2]} and S2 = {(1), (12)} then Z3 × S2 = {([0], (1)), ([1], (1)), ([2], (1)), ([0], (12)), ([1], (12)), ([2], (12))} For example, ([1], (12))([2], (1)) = ([1] ⊕ [2], (12)(1)) = ([0], (12)). Some of the subgroups of H × K can be constructed as follows. Theorem 15.7 The sets H × {eK } = {(h, eK ): h ∈ H} and {eH } × K = {(eH , k)) : k ∈ K} are subgroups of H × K. Proof. We prove that H × {eK } is a subgroup of H × K and we leave the second part of the theorem as an exercise for the reader. See Exercise ?? Since (eH , eK ) ∈ H × {eK } then H × {eK } 6= ∅. Let (h1, eK ) and (h2, eK ) be two elements of H × {eK }. Then −1 −1 −1 (h1, eK )(h2, eK ) = (h1, eK )(h2 , eK ) = (h1h2 , eK ) ∈ H × {eK } −1 since h1h2 ∈ H. Hence, by Theorem 7.5, H × {eK } is a subgroup of H × K Remark 15.1 By the Principle of Counting, if H and K are finite sets then |H × K| = |H| · |K|. 5.
Recommended publications
  • Direct Products and Homomorphisms
    Commuting properties Direct products and homomorphisms Simion Breaz logo Simion Breaz Direct products and homomorphisms Products and coproducts Commuting properties Contravariant functors Covariant functors Outline 1 Commuting properties Products and coproducts Contravariant functors Covariant functors logo Simion Breaz Direct products and homomorphisms Products and coproducts Commuting properties Contravariant functors Covariant functors Introduction Important properties of objects in particular categories (e.g. varieties of universal algebras) can be described using commuting properties of some canonical functors. For instance, in [Ad´amek and Rosicki: Locally presentable categories] there are the following examples: If V is a variety of finitary algebras and A ∈ V then A is finitely generated iff the functor Hom(A, −): V → Set preserves direct unions (i.e. directed colimits of monomorphisms); A is finitely presented (i.e. it is generated by finitely many generators modulo finitely many relations) iff the functor Hom(A, −): V → Set preserves directed colimits. logo Simion Breaz Direct products and homomorphisms Products and coproducts Commuting properties Contravariant functors Covariant functors Introduction Important properties of objects in particular categories (e.g. varieties of universal algebras) can be described using commuting properties of some canonical functors. For instance, in [Ad´amek and Rosicki: Locally presentable categories] there are the following examples: If V is a variety of finitary algebras and A ∈ V then A is finitely generated iff the functor Hom(A, −): V → Set preserves direct unions (i.e. directed colimits of monomorphisms); A is finitely presented (i.e. it is generated by finitely many generators modulo finitely many relations) iff the functor Hom(A, −): V → Set preserves directed colimits.
    [Show full text]
  • Lecture 1.3: Direct Products and Sums
    Lecture 1.3: Direct products and sums Matthew Macauley Department of Mathematical Sciences Clemson University http://www.math.clemson.edu/~macaule/ Math 8530, Advanced Linear Algebra M. Macauley (Clemson) Lecture 1.3: Direct products and sums Math 8530, Advanced Linear Algebra 1 / 5 Overview In previous lectures, we learned about vectors spaces and subspaces. We learned about what it meant for a subset to span, to be linearly independent, and to be a basis. In this lecture, we will see how to create new vector spaces from old ones. We will see several ways to \multiply" vector spaces together, and will learn how to construct: the complement of a subspace the direct sum of two subspaces the direct product of two vector spaces M. Macauley (Clemson) Lecture 1.3: Direct products and sums Math 8530, Advanced Linear Algebra 2 / 5 Complements and direct sums Theorem 1.5 (a) Every subspace Y of a finite-dimensional vector space X is finite-dimensional. (b) Every subspace Y has a complement in X : another subspace Z such that every vector x 2 X can be written uniquely as x = y + z; y 2 Y ; z 2 Z; dim X = dim Y + dim Z: Proof Definition X is the direct sum of subspaces Y and Z that are complements of each other. More generally, X is the direct sum of subspaces Y1;:::; Ym if every x 2 X can be expressed uniquely as x = y1 + ··· + ym; yi 2 Yi : We denote this as X = Y1 ⊕ · · · ⊕ Ym. M. Macauley (Clemson) Lecture 1.3: Direct products and sums Math 8530, Advanced Linear Algebra 3 / 5 Direct products Definition The direct product of X1 and X2 is the vector space X1 × X2 := (x1; x2) j x1 2 X1; x2 2 X2 ; with addition and multiplication defined component-wise.
    [Show full text]
  • Lie Groups and Lie Algebras
    2 LIE GROUPS AND LIE ALGEBRAS 2.1 Lie groups The most general definition of a Lie group G is, that it is a differentiable manifold with group structure, such that the multiplication map G G G, (g,h) gh, and the inversion map G G, g g−1, are differentiable.× → But we shall7→ not need this concept in full generality.→ 7→ In order to avoid more elaborate differential geometry, we will restrict attention to matrix groups. Consider the set of all invertible n n matrices with entries in F, where F stands either for the real (R) or complex× (C) numbers. It is easily verified to form a group which we denote by GL(n, F), called the general linear group in n dimensions over the number field F. The space of all n n matrices, including non- invertible ones, with entries in F is denoted by M(n,×F). It is an n2-dimensional 2 vector space over F, isomorphic to Fn . The determinant is a continuous function det : M(n, F) F and GL(n, F) = det−1(F 0 ), since a matrix is invertible → −{ } 2 iff its determinant is non zero. Hence GL(n, F) is an open subset of Fn . Group multiplication and inversion are just given by the corresponding familiar matrix 2 2 2 2 2 operations, which are differentiable functions Fn Fn Fn and Fn Fn respectively. × → → 2.1.1 Examples of Lie groups GL(n, F) is our main example of a (matrix) Lie group. Any Lie group that we encounter will be a subgroups of some GL(n, F).
    [Show full text]
  • Forcing with Copies of Countable Ordinals
    PROCEEDINGS OF THE AMERICAN MATHEMATICAL SOCIETY Volume 143, Number 4, April 2015, Pages 1771–1784 S 0002-9939(2014)12360-4 Article electronically published on December 4, 2014 FORCING WITH COPIES OF COUNTABLE ORDINALS MILOSˇ S. KURILIC´ (Communicated by Mirna Dˇzamonja) Abstract. Let α be a countable ordinal and P(α) the collection of its subsets isomorphic to α. We show that the separative quotient of the poset P(α), ⊂ is isomorphic to a forcing product of iterated reduced products of Boolean γ algebras of the form P (ω )/Iωγ ,whereγ ∈ Lim ∪{1} and Iωγ is the corre- sponding ordinal ideal. Moreover, the poset P(α), ⊂ is forcing equivalent to + a two-step iteration of the form (P (ω)/Fin) ∗ π,where[ω] “π is an ω1- + closed separative pre-order” and, if h = ω1,to(P (ω)/Fin) . Also we analyze δ I the quotients over ordinal ideals P (ω )/ ωδ and the corresponding cardinal invariants hωδ and tωδ . 1. Introduction The posets of the form P(X), ⊂,whereX is a relational structure and P(X) the set of (the domains of) its isomorphic substructures, were considered in [7], where a classification of the relations on countable sets related to the forcing-related properties of the corresponding posets of copies is described. So, defining two structures to be equivalent if the corresponding posets of copies produce the same generic extensions, we obtain a rough classification of structures which, in general, depends on the properties of the model of set theory in which we work. For example, under CH all countable linear orders are partitioned in only two classes.
    [Show full text]
  • Ring (Mathematics) 1 Ring (Mathematics)
    Ring (mathematics) 1 Ring (mathematics) In mathematics, a ring is an algebraic structure consisting of a set together with two binary operations usually called addition and multiplication, where the set is an abelian group under addition (called the additive group of the ring) and a monoid under multiplication such that multiplication distributes over addition.a[›] In other words the ring axioms require that addition is commutative, addition and multiplication are associative, multiplication distributes over addition, each element in the set has an additive inverse, and there exists an additive identity. One of the most common examples of a ring is the set of integers endowed with its natural operations of addition and multiplication. Certain variations of the definition of a ring are sometimes employed, and these are outlined later in the article. Polynomials, represented here by curves, form a ring under addition The branch of mathematics that studies rings is known and multiplication. as ring theory. Ring theorists study properties common to both familiar mathematical structures such as integers and polynomials, and to the many less well-known mathematical structures that also satisfy the axioms of ring theory. The ubiquity of rings makes them a central organizing principle of contemporary mathematics.[1] Ring theory may be used to understand fundamental physical laws, such as those underlying special relativity and symmetry phenomena in molecular chemistry. The concept of a ring first arose from attempts to prove Fermat's last theorem, starting with Richard Dedekind in the 1880s. After contributions from other fields, mainly number theory, the ring notion was generalized and firmly established during the 1920s by Emmy Noether and Wolfgang Krull.[2] Modern ring theory—a very active mathematical discipline—studies rings in their own right.
    [Show full text]
  • Automorphisms of Direct Products of Groups and Their Geometric Realisations
    Math. Ann. 263, 343 364 ~1983) Am Springer-Verlag 1983 Automorphisms of Direct Products of Groups and Their Geometric Realisations Francis E. A. Johnson Department of Mathematics, University College, Gower Street, London WC1E 6BT, UK O. Introduction An outstanding problem in the theory of Poincar6 Duality (=PD) groups is to decide whether each PD group can be realised as the fundamental group of a smooth closed aspherical manifold. See [2, 15-17] for previous work on this problem. There are two main ways of constructing new PD groups from old [18] ; (i) by extension of one PD group by another, and (ii) by torsion free extension of a PD group by a finite group. One approach to the above problem is to apply the constructions (i) and (ii) to known realisable PD groups and try to decide whether the resulting PD groups are realisable. Though other examples are known, for instance the examples con- structed by Mostow and Siu in [25], the most convenient PD groups with which to start are discrete uniform subgroups of noncompact Lie groups. If F is a torsion free discrete uniform subgroup of a connected noncompact Lie group G, and if K is a maximal compact subgroup of G, then F\G/K is a smooth closed manifold of type K(F, 1). The present paper is a continuation of both [15] and [16]. Whereas in [16] we allowed only Surface groups, that is, torsion free discrete uniform subgroups of PGLz(IR), and in [15] allowed only discrete uniform subgroups of Lie groups with no PSLz(IR) factors as building blocks, here we go some way towards combining the two.
    [Show full text]
  • A STUDY on the ALGEBRAIC STRUCTURE of SL 2(Zpz)
    A STUDY ON THE ALGEBRAIC STRUCTURE OF SL2 Z pZ ( ~ ) A Thesis Presented to The Honors Tutorial College Ohio University In Partial Fulfillment of the Requirements for Graduation from the Honors Tutorial College with the degree of Bachelor of Science in Mathematics by Evan North April 2015 Contents 1 Introduction 1 2 Background 5 2.1 Group Theory . 5 2.2 Linear Algebra . 14 2.3 Matrix Group SL2 R Over a Ring . 22 ( ) 3 Conjugacy Classes of Matrix Groups 26 3.1 Order of the Matrix Groups . 26 3.2 Conjugacy Classes of GL2 Fp ....................... 28 3.2.1 Linear Case . .( . .) . 29 3.2.2 First Quadratic Case . 29 3.2.3 Second Quadratic Case . 30 3.2.4 Third Quadratic Case . 31 3.2.5 Classes in SL2 Fp ......................... 33 3.3 Splitting of Classes of(SL)2 Fp ....................... 35 3.4 Results of SL2 Fp ..............................( ) 40 ( ) 2 4 Toward Lifting to SL2 Z p Z 41 4.1 Reduction mod p ...............................( ~ ) 42 4.2 Exploring the Kernel . 43 i 4.3 Generalizing to SL2 Z p Z ........................ 46 ( ~ ) 5 Closing Remarks 48 5.1 Future Work . 48 5.2 Conclusion . 48 1 Introduction Symmetries are one of the most widely-known examples of pure mathematics. Symmetry is when an object can be rotated, flipped, or otherwise transformed in such a way that its appearance remains the same. Basic geometric figures should create familiar examples, take for instance the triangle. Figure 1: The symmetries of a triangle: 3 reflections, 2 rotations. The red lines represent the reflection symmetries, where the trianlge is flipped over, while the arrows represent the rotational symmetry of the triangle.
    [Show full text]
  • Math 395: Category Theory Northwestern University, Lecture Notes
    Math 395: Category Theory Northwestern University, Lecture Notes Written by Santiago Can˜ez These are lecture notes for an undergraduate seminar covering Category Theory, taught by the author at Northwestern University. The book we roughly follow is “Category Theory in Context” by Emily Riehl. These notes outline the specific approach we’re taking in terms the order in which topics are presented and what from the book we actually emphasize. We also include things we look at in class which aren’t in the book, but otherwise various standard definitions and examples are left to the book. Watch out for typos! Comments and suggestions are welcome. Contents Introduction to Categories 1 Special Morphisms, Products 3 Coproducts, Opposite Categories 7 Functors, Fullness and Faithfulness 9 Coproduct Examples, Concreteness 12 Natural Isomorphisms, Representability 14 More Representable Examples 17 Equivalences between Categories 19 Yoneda Lemma, Functors as Objects 21 Equalizers and Coequalizers 25 Some Functor Properties, An Equivalence Example 28 Segal’s Category, Coequalizer Examples 29 Limits and Colimits 29 More on Limits/Colimits 29 More Limit/Colimit Examples 30 Continuous Functors, Adjoints 30 Limits as Equalizers, Sheaves 30 Fun with Squares, Pullback Examples 30 More Adjoint Examples 30 Stone-Cech 30 Group and Monoid Objects 30 Monads 30 Algebras 30 Ultrafilters 30 Introduction to Categories Category theory provides a framework through which we can relate a construction/fact in one area of mathematics to a construction/fact in another. The goal is an ultimate form of abstraction, where we can truly single out what about a given problem is specific to that problem, and what is a reflection of a more general phenomenom which appears elsewhere.
    [Show full text]
  • 1.7 Categories: Products, Coproducts, and Free Objects
    22 CHAPTER 1. GROUPS 1.7 Categories: Products, Coproducts, and Free Objects Categories deal with objects and morphisms between objects. For examples: objects sets groups rings module morphisms functions homomorphisms homomorphisms homomorphisms Category theory studies some common properties for them. Def. A category is a class C of objects (denoted A, B, C,. ) together with the following things: 1. A set Hom (A; B) for every pair (A; B) 2 C × C. An element of Hom (A; B) is called a morphism from A to B and is denoted f : A ! B. 2. A function Hom (B; C) × Hom (A; B) ! Hom (A; C) for every triple (A; B; C) 2 C × C × C. For morphisms f : A ! B and g : B ! C, this function is written (g; f) 7! g ◦ f and g ◦ f : A ! C is called the composite of f and g. All subject to the two axioms: (I) Associativity: If f : A ! B, g : B ! C and h : C ! D are morphisms of C, then h ◦ (g ◦ f) = (h ◦ g) ◦ f. (II) Identity: For each object B of C there exists a morphism 1B : B ! B such that 1B ◦ f = f for any f : A ! B, and g ◦ 1B = g for any g : B ! C. In a category C a morphism f : A ! B is called an equivalence, and A and B are said to be equivalent, if there is a morphism g : B ! A in C such that g ◦ f = 1A and f ◦ g = 1B. Ex. Objects Morphisms f : A ! B is an equivalence A is equivalent to B sets functions f is a bijection jAj = jBj groups group homomorphisms f is an isomorphism A is isomorphic to B partial order sets f : A ! B such that f is an isomorphism between A is isomorphic to B \x ≤ y in (A; ≤) ) partial order sets A and B f(x) ≤ f(y) in (B; ≤)" Ex.
    [Show full text]
  • Group Theory Group Theory
    12/16/2018 Group Theory Group Theory Properties of Real Topic No. 1 Numbers 1 2 Properties of Real Numbers Properties of Real Numbers Number Systems 0.131313…=0.13+ ℕ ={ 1, 2, 3, … } 0.0013+0.000013+… ℤ={…, -2, -1, 0, 1, 2, … } =13/100+13/10000+ ℚ={p/q | p, q ∊ ℤ and q≠0} 13/1000000+… ℚˊ= Set of Irrational =(13/100)(1+1/100+ Numbers 1/10000+…) ℝ=ℚ ∪ ℚˊ =(13/100)(100/99) =13/99 3 4 Properties of Real Numbers Properties of Real Numbers . e=2.718281828459045… ∊ ℚˊ . ∀ a, b, c ∊ ℝ, (ab)c=a(bc) . √2=1.414213562373095… ∊ ℚˊ . For instance, ((-2/3)4)√2=(-8/3) √2 =(-2/3)(4 √2) . √5=2.23606797749978… ∊ ℚˊ . For every a ∊ ℝ and 0 ∊ ℝ, a+0=a=0+a . ∀ a, b ∊ ℝ, a.b ∊ ℝ . For every a ∊ ℝ and 1 ∊ ℝ, a.1=a=1.a . ∀ a, b ∊ ℝ, a+b ∊ ℝ . For every a ∊ ℝ there exists -a ∊ ℝ such that . ∀ a, b, c ∊ ℝ, (a+b)+c=a+(b+c) a+(-a)=0=(-a)+a . For example, (1/4+3)+ √7=(13+4 √7)/4=1/4+(3+ √7) . For every a ∊ ℝ\{0} there exists 1/a ∊ ℝ\{0} such that a(1/a)=1=(1/a)a . ∀ a, b ∊ ℝ, a+b=b+a . ∀ a, b ∊ ℝ, a.b=b.a 5 6 1 12/16/2018 Group Theory Group Theory Properties of Complex Topic No. 2 Numbers 7 8 Properties of Complex Numbers Properties of Complex Numbers . ℂ={a+bi | a, b ∊ ℝ} .
    [Show full text]
  • RING THEORY 1. Ring Theory a Ring Is a Set a with Two Binary Operations
    CHAPTER IV RING THEORY 1. Ring Theory A ring is a set A with two binary operations satisfying the rules given below. Usually one binary operation is denoted `+' and called \addition," and the other is denoted by juxtaposition and is called \multiplication." The rules required of these operations are: 1) A is an abelian group under the operation + (identity denoted 0 and inverse of x denoted x); 2) A is a monoid under the operation of multiplication (i.e., multiplication is associative and there− is a two-sided identity usually denoted 1); 3) the distributive laws (x + y)z = xy + xz x(y + z)=xy + xz hold for all x, y,andz A. Sometimes one does∈ not require that a ring have a multiplicative identity. The word ring may also be used for a system satisfying just conditions (1) and (3) (i.e., where the associative law for multiplication may fail and for which there is no multiplicative identity.) Lie rings are examples of non-associative rings without identities. Almost all interesting associative rings do have identities. If 1 = 0, then the ring consists of one element 0; otherwise 1 = 0. In many theorems, it is necessary to specify that rings under consideration are not trivial, i.e. that 1 6= 0, but often that hypothesis will not be stated explicitly. 6 If the multiplicative operation is commutative, we call the ring commutative. Commutative Algebra is the study of commutative rings and related structures. It is closely related to algebraic number theory and algebraic geometry. If A is a ring, an element x A is called a unit if it has a two-sided inverse y, i.e.
    [Show full text]
  • Supplement. Direct Products and Semidirect Products
    Supplement: Direct Products and Semidirect Products 1 Supplement. Direct Products and Semidirect Products Note. In Section I.8 of Hungerford, we defined direct products and weak direct n products. Recall that when dealing with a finite collection of groups {Gi}i=1 then the direct product and weak direct product coincide (Hungerford, page 60). In this supplement we give results concerning recognizing when a group is a direct product of smaller groups. We also define the semidirect product and illustrate its use in classifying groups of small order. The content of this supplement is based on Sections 5.4 and 5.5 of Davis S. Dummitt and Richard M. Foote’s Abstract Algebra, 3rd Edition, John Wiley and Sons (2004). Note. Finitely generated abelian groups are classified in the Fundamental Theo- rem of Finitely Generated Abelian Groups (Theorem II.2.1). So when addressing direct products, we are mostly concerned with nonabelian groups. Notice that the following definition is “dull” if applied to an abelian group. Definition. Let G be a group, let x, y ∈ G, and let A, B be nonempty subsets of G. (1) Define [x, y]= x−1y−1xy. This is the commutator of x and y. (2) Define [A, B]= h[a,b] | a ∈ A,b ∈ Bi, the group generated by the commutators of elements from A and B where the binary operation is the same as that of group G. (3) Define G0 = {[x, y] | x, y ∈ Gi, the subgroup of G generated by the commuta- tors of elements from G under the same binary operation of G.
    [Show full text]