Exercise. Describe All Homomorphisms from Z24 to Z18. Solution

Total Page:16

File Type:pdf, Size:1020Kb

Exercise. Describe All Homomorphisms from Z24 to Z18. Solution 1. Exercise. Describe all homomorphisms from Z24 to Z18. Solution. We need some preliminary discussion before we try to define the homomorphisms to try to figure out how they might look like. Let f : Z24 → Z18 be a homomorphism. For any integer k ≥ 0 notice that f(k mod 24) = f((1 mod 24)+ ··· +(1mod 24)) (added k times) = f(1 mod 24)+ ··· + f(1 mod 24) = k · f(1 mod 24). It follows that a homomorphism f is completely determined by the value f(1 mod 24). Write f(1 mod 24) = n mod 18 where n is an integer such that 0 ≤ n ≤ 17. Notice that 24n mod 18 = 24f(1 mod 24) = f(24 mod 24) = f(0 mod 24) = 0 mod 18. since any homomorphism maps the identity to identity. So 18 must divide 24n, so 3 must divide 4n, hence 3 must divide n. So n must be among 0, 3, 6, 9, 12, 15. The above discussion suggests how we might want to define our homomorphisms. We find that any such homomorphism must send (1 mod 24) to (3k mod 18) for some k =0, 1, ··· , 5. It follows that there are at most 6 possible homomorphisms from Z24 → Z18, given by f0, f1, ··· , f5 where fk(r mod 24) = 3kr mod18 (1) for k =0, 1, ··· , 5. First we need to check that for each of this k, equation (1) gives a well ′ ′ defined function from Z24 to Z18. To this end, let r, r ∈ Z such that r mod 24 = r mod 24. Then (r − r′) is a multiple of 24. So (3kr − 3kr′)=3k(r − r′) is a multiple of 3 × 24 = 72. In particular, (3kr − 3kr′) is a multiple of 18. It follows that (3kr mod 18) only depends on the congruence class of r modulo 24. So indeed equation (1) gives us a well defined function from Z24 to Z18. Now we verify that for any r, s ∈ Z, we have fk(r mod 24+ s mod 24) = fk((r + s)mod 24)= 3k(r + s) mod 18 =3kr mod 18+3ks mod 18 = fk(r mod 18) + fk(s mod 18). So each fk is indeed a homomorphism from Z24 to Z18. At the risk of repeating ourselves, we recall that if f : Z24 → Z18 is any homomorphism, then f(1 mod 18) must be of the form (3k mod 18) for some k = 0, 1, ··· , 5 and that any homomorphism f is determined completely by the value of f(1 mod 18). For k =0, 1, ··· , 5, we have constructed a homomorphism fk such that fk(1mod18) = 3k mod 18. It follows that there are six homomorphisms from Z24 to Z18 and these are given by f0, f1, ··· , f5 as defined in equation (1). 1 2. If m is an integer, then mZ denotes the set of all integer multiples of m. 2.1. Lemma. Let m, n be nonzero integers and r = lcm(m, n). Then mZ ∩ nZ = rZ. Proof. Let x ∈ mZ ∩ nZ. Then x is a multiple of m and x is a multiple of n. Hence x is a multiple of r. So x ∈ rZ. So mZ ∩ nZ ⊆ rZ. On the other hand r is a multiple of m and n, so r ∈ mZ ∩ nZ. So rZ ⊆ mZ ∩ nZ since mZ ∩ nZ is a subgroup of Z. 2.2. Lemma. Let G be a group and a be an element of G. Then hami = {ak : k ∈ mZ}. Proof. If k ∈ mZ, then k has the form k = mt for some integer t, so ak = amt ∈hami. Thus {ak : k ∈ mZ}⊆hami. On the other hand, if g ∈ hami, then g has the form g = ams for some integer s. Let k = ms. Then g = ak and k ∈ mZ. So g ∈{ak : k ∈ mZ}. 2.3. Lemma. Let G be a group and a ∈ G. Then show that hami∩hani = halcm(m,n)i. Proof. Write r = lcm(m, n). From the previous lemma, we have hami∩hani = {ak : k ∈ mZ}∩{ak : k ∈ nZ} = {ak : k ∈ mZ ∩ nZ} = {ak : k ∈ rZ} = hari. Here the first and the final equality follows from lemma 2.2 and the third equality follows from lemma 2.1. The above is a rather straight forward proof. On the next page you will find a more elaborate argument which also ends up giving a proof of the same statement in a much more long winded manner. But along the way we prove a bunch of things which are of some importance in themselves. 2 All through the following discussion, let G be a group and let a be an element of G of finite order |a| = d. If m is an integer, we shall write m′ = gcd(m, d). Recall from the text that |am| = d/m′. 2.4. Lemma. If m and n are relatively prime nonzero integers, then (mn)′ = m′n′. Proof. Note that m′ is a factor of m and n′ is a factor of n. Since m and n are relatively prime, so are m′ and n′. Since m′ and n′ each divide d, it follows that m′n′ divides d. Since (m/m′) and (n/n′) are each relatively prime to d, so is the product (m/m′)(n/n′). Writing mn =(m′n′)((m/m′)(n/n′)), we find that gcd(mn, d)= m′n′. 2.5. Lemma. Let k, n be nonzero integers. If k is a multiple of n, then haki⊆hani. Proof. We can write k = nt for some t ∈ Z. So ak = ant ∈hani. So haki⊆hani. ′ 2.6. Lemma. Let m is a nonzero integer, then hami = ham i. ′ Proof. Note that m is a multiple of m′, so lemma 2.5 implies am ∈ham i. Conversely, we can write m′ = αm + β|a| for some integers α, β ∈ Z. So ′ am = aαm+β|a| = aαm ∈hami. 2.7. Lemma. (a) If b ∈hai, then |b| divides |a|. (b)If k, m are nonzero integers such that ak ∈hami, then m′ | k′. Proof. (a) We can write b = at for some nonzero integer t. So |b| = |a|/t′ is a factor of |a|. (b) By part (a), we have |ak| divides |am|, that is, |a|/k′ divides |a|/m′. So m′ divides k′. 2.8. Lemma. Let m, n be relatively prime nonzero integers. Then hami∩hani = hamni. Proof. Write H = hami∩hani. By lemma 2.5, we have hamni⊆hami and hamni⊆hani. So hamni ⊆ H. Notice that H is a subgroup of the cyclic group hami, so H is cyclic. So H = haki for some nonzero integer k. Since ak ∈ hami and ak ∈ hani, lemma 2.7(b) implies that m′ | k′ and that n′ | k′. Since m′ and n′ are relatively prime, we have m′n′ | k′. So k′ is a multiple ′ ′ of m′n′ =(mn)′ (by lemma 2.4) and lemma 2.5 implies that hak i⊆ha(mn) i. Now we obtain ′ ′ H = haki = hak i⊆ha(mn) i = hamni where the second and the final equality follows from lemma 2.6. 2.9. Lemma. Let m and n be two nonzero integers and l = lcm(m, n). Then hami∩hani = hali. Proof. Since a−m and am generate the same subgroup, without loss, we may assume that m is a positive integer and similarly, that n is a positive integer. Let d = gcd(m, n). Let b = ad. Then am = bm/d and an = bn/d. Now we have 2 hami∩hani = hbm/di∩hbn/di = hbmn/d i = hamn/di = hali where the second equality follows from 2.8 since (m/d) and (n/d) are relatively prime and the last equality follows since mn = dl. 3 3. Sign of a permutation, even and odd permutations. 3.1. Definition. We want to define the “sign” of a permutation. We start with the polyno- mial P (x1, ··· , xn) in n variables x1, ··· , xn with n(n − 1)/2 distinct linear factors (xi − xj) for 1 ≤ i < j ≤ n. For example, if n = 4, then P (x1, x2, x3, x4)=(x1 − x2)(x1 − x3)(x1 − x4)(x2 − x3)(x2 − x4)(x3 − x4). Let α ∈ Sn be any permutation. Say that a pair (i, j) is a crossing pair for α if (i < j) −1 −1 1 2 ··· n and (α >α (j)), that is, when we represent α in the form α(1) α(2) ··· ··· α(n) then in the second row “i is sitting under a bigger number than j”, that is, “i is sitting to the right of j”. Let N(α) be set of crossing pairs for α N(α)= {(i, j): 1 ≤ i < j ≤ n, α−1(i) >α−1(j)}. The number |N(α)| is called the number of crossings of α. Claim: The polynomial P (xα(1), ··· , xα(n)) also has n(n − 1)/2 distinct linear factors. Upto a plus or minus sign, these factors are same as the factors of P (x1, ··· , xn). More precisely, for each 1 ≤ i < j ≤ n, either (xi − xj) or (xj − xi) is a factor of P (xα(1), ··· , xα(n)), but not both. If (i, j) ∈/ N(α), then (xi − xj) is a factor, otherwise (xj − xi) is a factor. proof of claim: Pick a pair i, j such that 1 ≤ i < j ≤ n. Suppose (i, j) ∈/ N(α). Then −1 −1 α (i) <α (j), so (xα−1(i) −xα−1(j)) is a factor of P (x1, ··· , xn). So(xα(α−1(i)) −xα(α−1 (j)))= (xi − xj) is a factor of P (xα(1), ··· , xα(n)).
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]
  • Categories of Sets with a Group Action
    Categories of sets with a group action Bachelor Thesis of Joris Weimar under supervision of Professor S.J. Edixhoven Mathematisch Instituut, Universiteit Leiden Leiden, 13 June 2008 Contents 1 Introduction 1 1.1 Abstract . .1 1.2 Working method . .1 1.2.1 Notation . .1 2 Categories 3 2.1 Basics . .3 2.1.1 Functors . .4 2.1.2 Natural transformations . .5 2.2 Categorical constructions . .6 2.2.1 Products and coproducts . .6 2.2.2 Fibered products and fibered coproducts . .9 3 An equivalence of categories 13 3.1 G-sets . 13 3.2 Covering spaces . 15 3.2.1 The fundamental group . 15 3.2.2 Covering spaces and the homotopy lifting property . 16 3.2.3 Induced homomorphisms . 18 3.2.4 Classifying covering spaces through the fundamental group . 19 3.3 The equivalence . 24 3.3.1 The functors . 25 4 Applications and examples 31 4.1 Automorphisms and recovering the fundamental group . 31 4.2 The Seifert-van Kampen theorem . 32 4.2.1 The categories C1, C2, and πP -Set ................... 33 4.2.2 The functors . 34 4.2.3 Example . 36 Bibliography 38 Index 40 iii 1 Introduction 1.1 Abstract In the 40s, Mac Lane and Eilenberg introduced categories. Although by some referred to as abstract nonsense, the idea of categories allows one to talk about mathematical objects and their relationions in a general setting. Its origins lie in the field of algebraic topology, one of the topics that will be explored in this thesis. First, a concise introduction to categories will be given.
    [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]
  • Självständiga Arbeten I Matematik
    SJÄLVSTÄNDIGA ARBETEN I MATEMATIK MATEMATISKA INSTITUTIONEN, STOCKHOLMS UNIVERSITET Geometric interpretation of non-associative composition algebras av Fredrik Cumlin 2020 - No K13 MATEMATISKA INSTITUTIONEN, STOCKHOLMS UNIVERSITET, 106 91 STOCKHOLM Geometric interpretation of non-associative composition algebras Fredrik Cumlin Självständigt arbete i matematik 15 högskolepoäng, grundnivå Handledare: Wushi Goldring 2020 Abstract This paper aims to discuss the connection between non-associative composition algebra and geometry. It will first recall the notion of an algebra, and investigate the properties of an algebra together with a com- position norm. The composition norm will induce a law on the algebra, which is stated as the composition law. This law is then used to derive the multiplication and conjugation laws, where the last is also known as convolution. These laws are then used to prove Hurwitz’s celebrated the- orem concerning the different finite composition algebras. More properties of composition algebras will be covered, in order to look at the structure of the quaternions H and octonions O. The famous Fano plane will be the finishing touch of the relationship between the standard orthogonal vectors which construct the octonions. Lastly, the notion of invertible maps in relation to invertible loops will be covered, to later show the connection between 8 dimensional rotations − and multiplication of unit octonions. 2 Contents 1 Algebra 4 1.1 The multiplication laws . .6 1.2 The conjugation laws . .7 1.3 Dickson double . .8 1.4 Hurwitz’s theorem . 11 2 Properties of composition algebras 14 2.1 The left-, right- and bi-multiplication maps . 16 2.2 Basic properties of quaternions and octonions .
    [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]
  • LECTURE 12: LIE GROUPS and THEIR LIE ALGEBRAS 1. Lie
    LECTURE 12: LIE GROUPS AND THEIR LIE ALGEBRAS 1. Lie groups Definition 1.1. A Lie group G is a smooth manifold equipped with a group structure so that the group multiplication µ : G × G ! G; (g1; g2) 7! g1 · g2 is a smooth map. Example. Here are some basic examples: • Rn, considered as a group under addition. • R∗ = R − f0g, considered as a group under multiplication. • S1, Considered as a group under multiplication. • Linear Lie groups GL(n; R), SL(n; R), O(n) etc. • If M and N are Lie groups, so is their product M × N. Remarks. (1) (Hilbert's 5th problem, [Gleason and Montgomery-Zippin, 1950's]) Any topological group whose underlying space is a topological manifold is a Lie group. (2) Not every smooth manifold admits a Lie group structure. For example, the only spheres that admit a Lie group structure are S0, S1 and S3; among all the compact 2 dimensional surfaces the only one that admits a Lie group structure is T 2 = S1 × S1. (3) Here are two simple topological constraints for a manifold to be a Lie group: • If G is a Lie group, then TG is a trivial bundle. n { Proof: We identify TeG = R . The vector bundle isomorphism is given by φ : G × TeG ! T G; φ(x; ξ) = (x; dLx(ξ)) • If G is a Lie group, then π1(G) is an abelian group. { Proof: Suppose α1, α2 2 π1(G). Define α : [0; 1] × [0; 1] ! G by α(t1; t2) = α1(t1) · α2(t2). Then along the bottom edge followed by the right edge we have the composition α1 ◦ α2, where ◦ is the product of loops in the fundamental group, while along the left edge followed by the top edge we get α2 ◦ α1.
    [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]
  • Math 412. Adventure Sheet on Quotient Groups
    Math 412. Adventure sheet on Quotient Groups Fix an arbitrary group (G; ◦). DEFINITION: A subgroup N of G is normal if for all g 2 G, the left and right N-cosets gN and Ng are the same subsets of G. NOTATION: If H ⊆ G is any subgroup, then G=H denotes the set of left cosets of H in G. Its elements are sets denoted gH where g 2 G. The cardinality of G=H is called the index of H in G. DEFINITION/THEOREM 8.13: Let N be a normal subgroup of G. Then there is a well-defined binary operation on the set G=N defined as follows: G=N × G=N ! G=N g1N ? g2N = (g1 ◦ g2)N making G=N into a group. We call this the quotient group “G modulo N”. A. WARMUP: Define the sign map: Sn ! {±1g σ 7! 1 if σ is even; σ 7! −1 if σ is odd. (1) Prove that sign map is a group homomorphism. (2) Use the sign map to give a different proof that An is a normal subgroup of Sn for all n. (3) Describe the An-cosets of Sn. Make a table to describe the quotient group structure Sn=An. What is the identity element? B. OPERATIONS ON COSETS: Let (G; ◦) be a group and let N ⊆ G be a normal subgroup. (1) Take arbitrary ng 2 Ng. Prove that there exists n0 2 N such that ng = gn0. (2) Take any x 2 g1N and any y 2 g2N. Prove that xy 2 g1g2N.
    [Show full text]
  • 6. Localization
    52 Andreas Gathmann 6. Localization Localization is a very powerful technique in commutative algebra that often allows to reduce ques- tions on rings and modules to a union of smaller “local” problems. It can easily be motivated both from an algebraic and a geometric point of view, so let us start by explaining the idea behind it in these two settings. Remark 6.1 (Motivation for localization). (a) Algebraic motivation: Let R be a ring which is not a field, i. e. in which not all non-zero elements are units. The algebraic idea of localization is then to make more (or even all) non-zero elements invertible by introducing fractions, in the same way as one passes from the integers Z to the rational numbers Q. Let us have a more precise look at this particular example: in order to construct the rational numbers from the integers we start with R = Z, and let S = Znf0g be the subset of the elements of R that we would like to become invertible. On the set R×S we then consider the equivalence relation (a;s) ∼ (a0;s0) , as0 − a0s = 0 a and denote the equivalence class of a pair (a;s) by s . The set of these “fractions” is then obviously Q, and we can define addition and multiplication on it in the expected way by a a0 as0+a0s a a0 aa0 s + s0 := ss0 and s · s0 := ss0 . (b) Geometric motivation: Now let R = A(X) be the ring of polynomial functions on a variety X. In the same way as in (a) we can ask if it makes sense to consider fractions of such polynomials, i.
    [Show full text]
  • Right-Angled Artin Groups in the C Diffeomorphism Group of the Real Line
    Right-angled Artin groups in the C∞ diffeomorphism group of the real line HYUNGRYUL BAIK, SANG-HYUN KIM, AND THOMAS KOBERDA Abstract. We prove that every right-angled Artin group embeds into the C∞ diffeomorphism group of the real line. As a corollary, we show every limit group, and more generally every countable residually RAAG ∞ group, embeds into the C diffeomorphism group of the real line. 1. Introduction The right-angled Artin group on a finite simplicial graph Γ is the following group presentation: A(Γ) = hV (Γ) | [vi, vj] = 1 if and only if {vi, vj}∈ E(Γ)i. Here, V (Γ) and E(Γ) denote the vertex set and the edge set of Γ, respec- tively. ∞ For a smooth oriented manifold X, we let Diff+ (X) denote the group of orientation preserving C∞ diffeomorphisms on X. A group G is said to embed into another group H if there is an injective group homormophism G → H. Our main result is the following. ∞ R Theorem 1. Every right-angled Artin group embeds into Diff+ ( ). Recall that a finitely generated group G is a limit group (or a fully resid- ually free group) if for each finite set F ⊂ G, there exists a homomorphism φF from G to a nonabelian free group such that φF is an injection when restricted to F . The class of limit groups fits into a larger class of groups, which we call residually RAAG groups. A group G is in this class if for each arXiv:1404.5559v4 [math.GT] 16 Feb 2015 1 6= g ∈ G, there exists a graph Γg and a homomorphism φg : G → A(Γg) such that φg(g) 6= 1.
    [Show full text]
  • MA4E0 Lie Groups
    MA4E0 Lie Groups December 5, 2012 Contents 0 Topological Groups 1 1 Manifolds 2 2 Lie Groups 3 3 The Lie Algebra 6 4 The Tangent space 8 5 The Lie Algebra of a Lie Group 10 6 Integrating Vector fields and the exponential map 13 7 Lie Subgroups 17 8 Continuous homomorphisms are smooth 22 9 Distributions and Frobenius Theorem 23 10 Lie Group topics 24 These notes are based on the 2012 MA4E0 Lie Groups course, taught by John Rawnsley, typeset by Matthew Egginton. No guarantee is given that they are accurate or applicable, but hopefully they will assist your study. Please report any errors, factual or typographical, to [email protected] i MA4E0 Lie Groups Lecture Notes Autumn 2012 0 Topological Groups Let G be a group and then write the group structure in terms of maps; multiplication becomes m : G × G ! G −1 defined by m(g1; g2) = g1g2 and inversion becomes i : G ! G defined by i(g) = g . If we suppose that there is a topology on G as a set given by a subset T ⊂ P (G) with the usual rules. We give G × G the product topology. Then we require that m and i are continuous maps. Then G with this topology is a topological group Examples include 1. G any group with the discrete topology 2. Rn with the Euclidean topology and m(x; y) = x + y and i(x) = −x 1 2 2 2 3. S = f(x; y) 2 R : x + y = 1g = fz 2 C : jzj = 1g with m(z1; z2) = z1z2 and i(z) =z ¯.
    [Show full text]
  • Monoids: Theme and Variations (Functional Pearl)
    University of Pennsylvania ScholarlyCommons Departmental Papers (CIS) Department of Computer & Information Science 7-2012 Monoids: Theme and Variations (Functional Pearl) Brent A. Yorgey University of Pennsylvania, [email protected] Follow this and additional works at: https://repository.upenn.edu/cis_papers Part of the Programming Languages and Compilers Commons, Software Engineering Commons, and the Theory and Algorithms Commons Recommended Citation Brent A. Yorgey, "Monoids: Theme and Variations (Functional Pearl)", . July 2012. This paper is posted at ScholarlyCommons. https://repository.upenn.edu/cis_papers/762 For more information, please contact [email protected]. Monoids: Theme and Variations (Functional Pearl) Abstract The monoid is a humble algebraic structure, at first glance ve en downright boring. However, there’s much more to monoids than meets the eye. Using examples taken from the diagrams vector graphics framework as a case study, I demonstrate the power and beauty of monoids for library design. The paper begins with an extremely simple model of diagrams and proceeds through a series of incremental variations, all related somehow to the central theme of monoids. Along the way, I illustrate the power of compositional semantics; why you should also pay attention to the monoid’s even humbler cousin, the semigroup; monoid homomorphisms; and monoid actions. Keywords monoid, homomorphism, monoid action, EDSL Disciplines Programming Languages and Compilers | Software Engineering | Theory and Algorithms This conference paper is available at ScholarlyCommons: https://repository.upenn.edu/cis_papers/762 Monoids: Theme and Variations (Functional Pearl) Brent A. Yorgey University of Pennsylvania [email protected] Abstract The monoid is a humble algebraic structure, at first glance even downright boring.
    [Show full text]