4 Direct Products

Total Page:16

File Type:pdf, Size:1020Kb

4 Direct Products 4 Direct products 4.1 Definition and Examples Direct products are a straightforward way to define new groups from old. The simplest example is the Cartesian product Z2 × Z2. Writing Z2 = f0, 1g, we see that Z2 × Z2 = f(0, 0), (0, 1), (1, 0), (1, 1)g has four elements. This set can be given a group structure in an obvious way. Simply define (p, q) + (r, s) := (p + r, q + s) where p + r and q + s are computed in Z2. This is clearly a binary operation on Z2 × Z2. It is straightforward to check that this operation defines a group. Indeed, we obtain the following Cayley table: + (0, 0) (0, 1) (1, 0) (1, 1) (0, 0) (0, 0) (0, 1) (1, 0) (1, 1) (0, 1) (0, 1) (0, 0) (1, 1) (1, 0) (1, 0) (1, 0) (1, 1) (0, 0) (0, 1) (1, 1) (1, 1) (1, 0) (0, 1) (0, 0) ∼ This should look familiar: it is exactly the Cayley table for the Klein 4-group: Z2 × Z2 = V. This construction works in general. Theorem 4.1. Let G1,..., Gn be multiplicative groups. The binary operation (a1,..., an) · (b1,..., bn) := (a1b1,..., anbn) induces a group structure on the set G1 × · · · × Gn. Definition 4.2. The group G1 × · · · × Gn is the direct product of the groups G1,..., Gn. If the groups G1,..., Gn are additive, then the operation will also be written additively. It is immediate that a direct product has cardinality equal to the product of the cardinalities of G1,..., Gn. Proof of Theorem. Simply check the group axioms. Closure Since each aibi 2 Gi this is immediate. Associativity Each Gi is associative: it is a short calculation to see that the associativity condition on the whole direct product is equivalent to each individual Gi being associative. Identity If ei is the identity in Gi, then (e1,..., en) is the identity in the direct product. −1 −1 −1 Inverse (a1,..., an) = (a1 ,..., an ). The proof uses nothing beyond the fact that each factor Gi is a group in its own right—this should not be surprising as we have no other information to work with! In particular you should note that the individual groups Gi do not interact with each other. 1 Examples 1. Z2 × Z3 = f(0, 0), (0, 1), (0, 2), (1, 0), (1, 1), (1, 2)g under addition (+2, +3). Observe that if we choose a = (1, 1), then the group can be written (in the above order) as fe, 4a, 2a, 3a, a, 5ag. Thus Z2 × Z3 is generated by a and is therefore cyclic. Being a cyclic group of order 6, we necessarily ∼ have Z2 × Z3 = Z6. 2. The direct sum of vector spaces W = U ⊕ V is a more general example. Indeed in linear algebra it is typical to use direct sum notation rather than Cartesian products. For example the direct sum of n copies of the real line R is the familiar vector space n Rn = M R = R ⊕ · · · ⊕ R i=1 4.2 Orders of elements in direct products Z = 12 Z In 12 the element 10 has order 6 gcd(10,12) , while in 8 the element 2 has order 4. What is the order of the element (10, 2) 2 Z12 × Z8? Simply compute the cyclic group generated by this element: n o h(10, 2)i = (10, 2), (8, 4), (6, 6), (4, 0), (2, 2), (0, 4), (10, 6), (8, 0), (6, 2), (4, 4), (2, 6), (0, 0) The order is therefore 12. If you consider what happens to each of the entries of each pair, it should be obvious that the order is the least common multiple of the orders of the originals: 12 = lcm(6, 4). In general we have the following. Theorem 4.3. Suppose that ai 2 Gi has order ri for each i. Then (a1,..., an) 2 G1 × · · · × Gn has order lcm(r1,..., rn). Proof. We simply compute the order: k k k k (a1,..., an) = (a1,..., an) = (e1, e2,..., en) () 8i, ai = ei () 8i, ri j k The order is the minimal positive integer k satisfying the above, which is precisely k = lcm(r1,..., rn) 1 Example What is the order of (1, 4, 3, 2) 2 Z4 × Z7 × Z5 × Z20? 2 Z n Recalling that the order of x n is gcd(x,n) we see that the above elements have orders 4, 7, 5 and 10 respectively. Thus the order of (1, 4, 3, 2) is lcm(4, 7, 5, 10) = 140. 4.3 Finite(ly generated) abelian groups We have already seen that all finite cyclic groups are isomorphic to some Zn. As we shall see in a moment, all finite abelian groups are isomorphic to direct products of these. ∼ We have already seen that Z2 × Z3 = Z6 but that Z2 × Z2 Z4. What is the pattern here? Your gut should suspect that it has something to do with the fact that 2 and 3 are relatively prime. 1 Note that (1, 4, 3, 2) is really a quadruple of equivalence classes ([1]4, [4]7, [3]5, [2]20). 2 Theorem 4.4. Zm × Zn is cyclic () gcd(m, n) = 1, specifically, ∼ Zm × Zn = Zmn () gcd(m, n) = 1 More generally, ∼ Zm1 × · · · × Zmk = Zm1···mk () gcd(mi, mj) = 1, 8i 6= j r1 rk Moreover if n = p1 ··· pk is the unique prime factorization of an integer n, then ∼ Zn = Z r1 × · · · × Z rk p1 pk Proof. This is merely a corollary of Theorem 4.3. For brevity we just prove the first part: the remainder follows by induction. If gcd(m, n) = 1, then the element (1, 1) 2 Zm × Zn has order mn lcm(m, n) = gcd(m, n) Hence (1, 1) is a generator of Zm × Zn, which is then cyclic. Conversely, suppose gcd(m, n) = d ≥ 2. Then the maximum order of an element (p, q) 2 Zm × Zn is mn mn lcm(m, n) = = < mn gcd(m, n) d It follows that Zm × Zn is not cyclic. Example Is Z5 × Z12 × Z43 cyclic? The Theorem says yes, since no pairs of the numbers 5, 12, 43 have any common factors. It is ghastly to write out, but there are 15 different ways (up to reordering) of expressing this group! ∼ ∼ ∼ ∼ Z2580 = Z3 × Z860 = Z4 × Z645 = Z5 × Z516 = Z43 × Z60 ∼ ∼ ∼ = Z12 × Z215 = Z15 × Z172 = Z20 × Z129 ∼ ∼ ∼ = Z3 × Z4 × Z215 = Z3 × Z5 × Z172 = Z3 × Z20 × Z43 ∼ ∼ ∼ = Z4 × Z5 × Z129 = Z4 × Z15 × Z43 = Z5 × Z12 × Z43 ∼ = Z3 × Z4 × Z5 × Z43 Direct products of cyclic groups have a universal application here. As the next Theorem shows, every finitely generated abelian group is isomorphic to a direct product of cyclic groups. Theorem 4.5 (Fundamental Theorem of finitely generated Abelian groups). Every finitely generated abelian group is isomorphic to a group of the form Z r1 × · · · × Zprn × Z × · · · × Z p1 n where the pi are (not necessarily distinct) primes, the ri are positive integers and there are finitely many factors of Z. The proof is far too difficult for this course. Its purpose here is to allow us to classify finite abelian groups up to isomorphism. Recall the optional section on generating sets to remind yourself of what finitely generated means: the above direct product is only a finite group if it has no factors of Z. 3 Example Find, up to isomorphism, all abelian groups of order 450. First note that 450 = 2 · 32 · 52. Now apply the fundamental theorem to see that the complete list is ∼ 1. Z450 = Z2 × Z32 × Z52 2. Z2 × Z3 × Z3 × Z52 3. Z2 × Z32 × Z5 × Z5 4. Z2 × Z3 × Z3 × Z5 × Z5 2 Theorem 4.6. If m is a square free integer (@k 2 Z≥2 such that k j m) then there is only one abelian group of order m (up to isomorphism). Proof. By the Fundamental Theorem such a group G must be isomorphic to some Z r1 × · · · × Zprn p1 n r1 rn with m = p ··· pn . But m being square free implies that every exponent si is equal to 1 and all the 1 ∼ primes pi are distinct. By Theorem 4.4 we have G = Zp1···pn = Zm. All groups of small order As examples of the above we list all the groups of orders 1 through 15 and the abelian groups of order 16 up to isomorphism. The Fundamental Theorem gives us all abelian groups. In particular observe where Theorem 4.6 applies. There are three groups we haven’t previously encountered: the Quaternion group Q8, the Alternating group A4, and the Dicyclic group Dic3. We will consider the alternating group properly later, the others you can look up if you’re interested. There are nine non- abelian groups of order 16 up to isomorphism, six of which we have no notation for in this class! You may be suspicious from looking at the table that there are no non-abelian groups of any odd order. This is not so, but you need to go to order 21 before you find one. Order Abelian Non-Abelian 1 Z1 2 Z2 3 Z3 ∼ 4 Z4, V = Z2 × Z2 5 Z5 ∼ ∼ 6 Z6 = Z2 × Z3 D3 = S3 7 Z7 ∼ 8 Z8, Z2 × Z4, Z2 × Z2 × Z2 D4, Q8 = Dic2 9 Z9, Z3 × Z3 ∼ 10 Z10 = Z2 × Z5 D5 11 Z11 ∼ ∼ 12 Z12 = Z3 × Z4, Z2 × Z6 = Z2 × Z2 × Z3 D6, A4, Dic3 13 Z13 ∼ 14 Z14 = Z2 × Z7 D7 ∼ 15 Z15 = Z3 × Z5 16 Z16, Z4 × Z4, Z2 × Z8, Z2 × Z2 × Z4, Z2 × Z2 × Z2 × Z2 Many 4.
Recommended publications
  • Complex Algebras of Semigroups Pamela Jo Reich Iowa State University
    Iowa State University Capstones, Theses and Retrospective Theses and Dissertations Dissertations 1996 Complex algebras of semigroups Pamela Jo Reich Iowa State University Follow this and additional works at: https://lib.dr.iastate.edu/rtd Part of the Mathematics Commons Recommended Citation Reich, Pamela Jo, "Complex algebras of semigroups " (1996). Retrospective Theses and Dissertations. 11765. https://lib.dr.iastate.edu/rtd/11765 This Dissertation is brought to you for free and open access by the Iowa State University Capstones, Theses and Dissertations at Iowa State University Digital Repository. It has been accepted for inclusion in Retrospective Theses and Dissertations by an authorized administrator of Iowa State University Digital Repository. For more information, please contact [email protected]. INFORMATION TO USERS This manuscript has been reproduced from the microfihn master. UMI fibns the text du-ectly from the original or copy submitted. Thus, some thesis and dissertation copies are in typewriter face, while others may be from any type of computer printer. The quality of this reproductioii is dependent upon the quality of the copy submitted. Broken or indistinct print, colored or poor quality illustrations and photographs, print bleedthrough, substandard margins, and unproper alignment can adversely affect reproduction. In the unlikely event that the author did not send UMI a complete manuscript and there are missing pages, these will be noted. Also, if unauthorized copyright material had to be removed, a note will indicate the deletion. Oversize materials (e.g., m^s, drawings, charts) are reproduced by sectioning the original, beginning at the upper left-hand comer and continuing from left to right in equal sections with small overiaps.
    [Show full text]
  • Group Isomorphisms MME 529 Worksheet for May 23, 2017 William J
    Group Isomorphisms MME 529 Worksheet for May 23, 2017 William J. Martin, WPI Goal: Illustrate the power of abstraction by seeing how groups arising in different contexts are really the same. There are many different kinds of groups, arising in a dizzying variety of contexts. Even on this worksheet, there are too many groups for any one of us to absorb. But, with different teams exploring different examples, we should { as a class { discover some justification for the study of groups in the abstract. The Integers Modulo n: With John Goulet, you explored the additive structure of Zn. Write down the addition table for Z5 and Z6. These groups are called cyclic groups: they are generated by a single element, the element 1, in this case. That means that every element can be found by adding 1 to itself an appropriate number of times. The Group of Units Modulo n: Now when we look at Zn using multiplication as our operation, we no longer have a group. (Why not?) The group U(n) = fa 2 Zn j gcd(a; n) = 1g ∗ is sometimes written Zn and is called the group of units modulo n. An element in a number system (or ring) is a \unit" if it has a multiplicative inverse. Write down the multiplication tables for U(6), U(7), U(8) and U(12). The Group of Rotations of a Regular n-Gon: Imagine a regular polygon with n sides centered at the origin O. Let e denote the identity transformation, which leaves the poly- gon entirely fixed and let a denote a rotation about O in the counterclockwise direction by exactly 360=n degrees (2π=n radians).
    [Show full text]
  • 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]
  • Arxiv:2001.06557V1 [Math.GR] 17 Jan 2020 Magic Cayley-Sudoku Tables
    Magic Cayley-Sudoku Tables∗ Rosanna Mersereau Michael B. Ward Columbus, OH Western Oregon University 1 Introduction Inspired by the popularity of sudoku puzzles along with the well-known fact that the body of the Cayley table1 of any finite group already has 2/3 of the properties of a sudoku table in that each element appears exactly once in each row and in each column, Carmichael, Schloeman, and Ward [1] in- troduced Cayley-sudoku tables. A Cayley-sudoku table of a finite group G is a Cayley table for G the body of which is partitioned into uniformly sized rectangular blocks, in such a way that each group element appears exactly once in each block. For example, Table 1 is a Cayley-sudoku ta- ble for Z9 := {0, 1, 2, 3, 4, 5, 6, 7, 8} under addition mod 9 and Table 3 is a Cayley-sudoku table for Z3 × Z3 where the operation is componentwise ad- dition mod 3 (and ordered pairs (a, b) are abbreviated ab). In each case, we see that we have a Cayley table of the group partitioned into 3 × 3 blocks that contain each group element exactly once. Lorch and Weld [3] defined a modular magic sudoku table as an ordinary sudoku table (with 0 in place of the usual 9) in which the row, column, arXiv:2001.06557v1 [math.GR] 17 Jan 2020 diagonal, and antidiagonal sums in each 3 × 3 block in the table are zero mod 9. “Magic” refers, of course, to magic Latin squares which have a rich history dating to ancient times.
    [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]
  • Cosets and Cayley-Sudoku Tables
    Cosets and Cayley-Sudoku Tables Jennifer Carmichael Chemeketa Community College Salem, OR 97309 [email protected] Keith Schloeman Oregon State University Corvallis, OR 97331 [email protected] Michael B. Ward Western Oregon University Monmouth, OR 97361 [email protected] The wildly popular Sudoku puzzles [2] are 9 £ 9 arrays divided into nine 3 £ 3 sub-arrays or blocks. Digits 1 through 9 appear in some of the entries. Other entries are blank. The goal is to ¯ll the blank entries with the digits 1 through 9 in such a way that each digit appears exactly once in each row and in each column, and in each block. Table 1 gives an example of a completed Sudoku puzzle. One proves in introductory group theory that every element of any group appears exactly once in each row and once in each column of the group's op- eration or Cayley table. (In other words, any Cayley table is a Latin square.) Thus, every Cayley table has two-thirds of the properties of a Sudoku table; only the subdivision of the table into blocks that contain each element ex- actly once is in doubt. A question naturally leaps to mind: When and how can a Cayley table be arranged in such a way as to satisfy the additional requirements of being a Sudoku table? To be more speci¯c, group elements labeling the rows and the columns of a Cayley table may be arranged in any order. Moreover, in de¯ance of convention, row labels and column labels need not be in the same order.
    [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]
  • Midterm # 1 Solutions
    Midterm # 1 Solutions The Nintendo game “Baseball Stars” February 12, 2008 Hi baseball fans! We’re coming to you from video game land to give you the solutions to the first test. We’re lazy aging video game superstars and don’t feel the need to type out something that has already been typed out, or can be found verbatim from the book. We hope you enjoy them! Rock on! 1. All of these first ones are definition and can be found in the book. 2. (a) (i) U(12) is the set of all elements mod 12 that are relatively prime to 12. This would be the set containing the integer representatives {1, 5, 7, 11}. The multiplication table is as follows: 1 5 7 11 1 1 5 7 11 5 5 1 11 7 7 7 11 1 5 11 11 7 5 1 (ii) This group is not cyclic: you can see this from the multiplication table that < 1 >= {1}, < 5 >= {1, 5}, < 7 >= {1, 7}, < 11 >= {1, 11}. None of these elements generate U(12), so U(12) is not cyclic. 1 (b) (i) I’m sure at some point we produced a Cayley table for D3. We computed in class (ii) the the center of D3 is Z(D3) = {e}. Should your class notes be incomplete on this matter or we never did a Cayley table of D3, then speak with Corey. Nobody chose this problem to do, so we Baseball Stars feel okay about leaving it at that. (c) (i) The group Gl(n, R) = {A ∈ Mn×n(R)|det(A) 6= 0}.
    [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]
  • Dihedral Group of Order 8 (The Number of Elements in the Group) Or the Group of Symmetries of a Square
    CHAPTER 1 Introduction to Groups Symmetries of a Square A plane symmetry of a square (or any plane figure F ) is a function from the square to itself that preserves distances, i.e., the distance between the images of points P and Q equals the distance between P and Q. The eight symmetries of a square: 22 1. INTRODUCTION TO GROUPS 23 These are all the symmetries of a square. Think of a square cut from a piece of glass with dots of di↵erent colors painted on top in the four corners. Then a symmetry takes a corner to one of four corners with dot up or down — 8 possibilities. A succession of symmetries is a symmetry: Since this is a composition of functions, we write HR90 = D, allowing us to view composition of functions as a type of multiplication. Since composition of functions is associative, we have an operation that is both closed and associative. R0 is the identity: for any symmetry S, R0S = SR0 = S. Each of our 1 1 1 symmetries S has an inverse S− such that SS− = S− S = R0, our identity. R0R0 = R0, R90R270 = R0, R180R180 = R0, R270R90 = R0 HH = R0, V V = R0, DD = R0, D0D0 = R0 We have a group. This group is denoted D4, and is called the dihedral group of order 8 (the number of elements in the group) or the group of symmetries of a square. 24 1. INTRODUCTION TO GROUPS We view the Cayley table or operation table for D4: For HR90 = D (circled), we find H along the left and R90 on top.
    [Show full text]