A → B Be an Injective Function with a Non-Empty. Prove Or Disprove

Total Page:16

File Type:pdf, Size:1020Kb

A → B Be an Injective Function with a Non-Empty. Prove Or Disprove 1. Let f : A ! B be an injective function with A non-empty. Prove or disprove: there is always a function g : B ! A so that f ◦ g = idB (where idB : B ! B is the function idB(b) = b)? A disproof: suppose f is injective but not surjective, for instance A = f1; 2g, B = f3; 4; 5g, f(1) = 4, f(2) = 3. No matter what function g : B ! A we use, we would need (f ◦ g)(5) = f(g(5)) 2 f(A) = f3; 4g, so f ◦ g 6= idB. Because there's a good guess as to what a hypothetical function g would have to be (g inverts f where possible, and picks an arbitrary element of A where that isn't possible), a lot of people made that guess and analyzed only that example. But that's not quite enough: we have to show that there's no g which works, not just that our best guess doesn't work. (Also, some people talked themselves into believing that their guess did work, usually by forgetting that they had to consider the case where g(b) was an arbitrary element of A.) 2. Prove or disprove the following statement: if f : A ! A is a function so that f ◦ f is bijective then f is bijective. A proof: First we show that f is injective. If f(a) = f(b) then f(f(a)) = f(f(b)), so (f ◦f)(a) = (f ◦ f)(b), so by the injectivity of f ◦ f, a = b. So f is injective. Next we show that f is surjective. For any a 2 A, since f ◦ f is surjective, there is a b so that (f ◦ f)(b) = a. Therefore f(f(b)) = a, and therefore c = f(b) is a value so that f(c) = a. So f is surjective. Therefore f is bijective. By far the most common problem was forgetting how to prove injectivity. A lot of people assumed f(f(a)) = f(f(b)) and concluded that f(a) = f(b), which, while true, doesn't tell you anything useful. 3. Prove that, for every natural number n, there exist positive integers x; y; and z so that x2 + y2 = z2n. (Hint: Make sure you understand the statement before trying to prove it. Use induction on n. Recall that 32 + 42 = 52.) By induction on n. Base case: n = 1, we have 32 + 42 = 52. Inductive case: Suppose there exist x; y; z so that x2 + y2 = z2n. We need to show that there exist a; b; c so that a2 + b2 = c2n+2. We take a = xz, b = yz, c = z. Then a2 + b2 = (xz)2 + (yz)2 = (x2 + y2)z2 = z2nz2 = z2n+2. I expected fewer people to actually solve this one all the way (and was glad to see that many people got it completely right). Most of the points were for setting up the problem correctly; in particular, stating the inductive hypothesis clearly (and understanding it). The most common problem was forgetting that there are quantifiers in the statement \there exist x; y; z so that x2 + y2 = z2n", and in particular mixing up the values x; y; z in the inductive hypothesis for P (n) and the ones we need to find for P (n + 1), which often led to people assuming they had to be the same values. (Here I used different names, which I think is good practice, but not necessary.) 1 of 2 4. One important fact about rational numbers is that every rational numbers has an additive inverse|that is, for every rational q, there is an r so that q+r = 0. Convert this into a statement about integers (You do not need to prove this statement.) a c ad+bc A rational number is a pair a=b with b 6= 0, and addition is given by b + d = bd , so this says (8a; b 2 Z)(b 6= 0 ) (9c; d 2 Z)[d 6= 0 ^ ad + bc = 0]): (If we prefer, we could be more explicit and say that 0 is the rational 0=1, so the last equality is (ad + bc) · 1 = bd · 0.) The most common problem here was losing track of the quantifers, or the conditions that b 6= 0 and d 6= 0, and just writing ad + bc = 0. 2 5. (a) Let S ⊆ R be the set of positive real numbers x such that x 2 Q. Show that S is countable. 2 (b) Let T ⊆ R be the set of all real numbers x such that x 2 Q. Show that T is countable. 2 For the first part, the function f : S ! Q given by f(x) = x is injective, so jSj ≤ jQj. Since N ⊆ S, we have jQj = jNj ≤ jSj, so by Schroeder-Bernstein, jSj = jQj. For the second part, we have T = S [U [f0g where U is the set of negative real numbers x such 2 that x 2 Q. Since S is countable and S ⊆ T , certainly jSj ≤ jT j. Also x 7! −x is a bijection between S and U, so T is the union of two countably infinite sets and a singleton 0, and we know that the union of finitely many countable sets is countable. If we want to be more explicit, we could proceed as follows: since S is countable, let h : N ! S be a bijection. We define k : Z ! T by 8 < h(z) if z > 0 k(z) = 0 if z = 0 : h(−z) if z < 0 Then k is a bijection. 2 of 2.
Recommended publications
  • The Cardinality of a Finite Set
    1 Math 300 Introduction to Mathematical Reasoning Fall 2018 Handout 12: More About Finite Sets Please read this handout after Section 9.1 in the textbook. The Cardinality of a Finite Set Our textbook defines a set A to be finite if either A is empty or A ≈ Nk for some natural number k, where Nk = {1,...,k} (see page 455). It then goes on to say that A has cardinality k if A ≈ Nk, and the empty set has cardinality 0. These are standard definitions. But there is one important point that the book left out: Before we can say that the cardinality of a finite set is a well-defined number, we have to ensure that it is not possible for the same set A to be equivalent to Nn and Nm for two different natural numbers m and n. This may seem obvious, but it turns out to be a little trickier to prove than you might expect. The purpose of this handout is to prove that fact. The crux of the proof is the following lemma about subsets of the natural numbers. Lemma 1. Suppose m and n are natural numbers. If there exists an injective function from Nm to Nn, then m ≤ n. Proof. For each natural number n, let P (n) be the following statement: For every m ∈ N, if there is an injective function from Nm to Nn, then m ≤ n. We will prove by induction on n that P (n) is true for every natural number n. We begin with the base case, n = 1.
    [Show full text]
  • Monomorphism - Wikipedia, the Free Encyclopedia
    Monomorphism - Wikipedia, the free encyclopedia http://en.wikipedia.org/wiki/Monomorphism Monomorphism From Wikipedia, the free encyclopedia In the context of abstract algebra or universal algebra, a monomorphism is an injective homomorphism. A monomorphism from X to Y is often denoted with the notation . In the more general setting of category theory, a monomorphism (also called a monic morphism or a mono) is a left-cancellative morphism, that is, an arrow f : X → Y such that, for all morphisms g1, g2 : Z → X, Monomorphisms are a categorical generalization of injective functions (also called "one-to-one functions"); in some categories the notions coincide, but monomorphisms are more general, as in the examples below. The categorical dual of a monomorphism is an epimorphism, i.e. a monomorphism in a category C is an epimorphism in the dual category Cop. Every section is a monomorphism, and every retraction is an epimorphism. Contents 1 Relation to invertibility 2 Examples 3 Properties 4 Related concepts 5 Terminology 6 See also 7 References Relation to invertibility Left invertible morphisms are necessarily monic: if l is a left inverse for f (meaning l is a morphism and ), then f is monic, as A left invertible morphism is called a split mono. However, a monomorphism need not be left-invertible. For example, in the category Group of all groups and group morphisms among them, if H is a subgroup of G then the inclusion f : H → G is always a monomorphism; but f has a left inverse in the category if and only if H has a normal complement in G.
    [Show full text]
  • A New Characterization of Injective and Surjective Functions and Group Homomorphisms
    A new characterization of injective and surjective functions and group homomorphisms ANDREI TIBERIU PANTEA* Abstract. A model of a function f between two non-empty sets is defined to be a factorization f ¼ i, where ¼ is a surjective function and i is an injective Æ ± function. In this note we shall prove that a function f is injective (respectively surjective) if and only if it has a final (respectively initial) model. A similar result, for groups, is also proven. Keywords: factorization of morphisms, injective/surjective maps, initial/final objects MSC: Primary 03E20; Secondary 20A99 1 Introduction and preliminary remarks In this paper we consider the sets taken into account to be non-empty and groups denoted differently to be disjoint. For basic concepts on sets and functions see [1]. It is a well known fact that any function f : A B can be written as the composition of a surjective function ! followed by an injective function. Indeed, f f 0 id , where f 0 : A Im(f ) is the restriction Æ ± B ! of f to Im(f ) is one such factorization. Moreover, it is an elementary exercise to prove that this factorization is unique up to an isomorphism: i.e. if f i ¼, i : A X , ¼: X B is 1 ¡ Æ ±¢ ! 1 ! another factorization for f , then we easily see that X i ¡ Im(f ) and that ¼ i ¡ f 0. The Æ Æ ± purpose of this note is to study the dual problem. Let f : A B be a function. We shall define ! a model of the function f to be a triple (X ,i,¼) made up of a set X , an injective function i : A X and a surjective function ¼: X B such that f ¼ i, that is, the following diagram ! ! Æ ± is commutative.
    [Show full text]
  • Homomorphism Problems for First-Order Definable Structures
    Homomorphism Problems for First-Order Definable Structures∗ Bartek Klin1, Sławomir Lasota1, Joanna Ochremiak2, and Szymon Toruńczyk1 1 University of Warsaw 2 Universitat Politécnica de Catalunya Abstract We investigate several variants of the homomorphism problem: given two relational structures, is there a homomorphism from one to the other? The input structures are possibly infinite, but definable by first-order interpretations in a fixed structure. Their signatures can be either finite or infinite but definable. The homomorphisms can be either arbitrary, or definable with parameters, or definable without parameters. For each of these variants, we determine its decidability status. 1998 ACM Subject Classification F.4.1 Mathematical Logic–Model theory, F.4.3 Formal Lan- guages–Decision problems Keywords and phrases Sets with atoms, first-order interpretations, homomorphism problem Digital Object Identifier 10.4230/LIPIcs.FSTTCS.2016.? 1 Introduction First-order definable sets, although usually infinite, can be finitely described and are therefore amenable to algorithmic manipulation. Definable sets (we drop the qualifier first-order in what follows) are parametrized by a fixed underlying relational structure A whose elements are called atoms. We shall assume that the first-order theory of A is decidable. To simplify the presentation, unless stated otherwise, let A be a countable set {1, 2, 3,...} equipped with the equality relation only; we shall call this the pure set. I Example 1. Let V = {{a, b} | a, b ∈ A, a 6= b } , E = { ({a, b}, {c, d}) | a, b, c, d ∈ A, a 6= b ∧ a 6= c ∧ a 6= d ∧ b 6= c ∧ b 6= d ∧ c 6= d } .
    [Show full text]
  • Almost Recursively Enumerable Sets
    TRANSACTIONS OF THE AMERICAN MATHEMATICAL SOCIETY Volume 164, February 1972 ALMOST RECURSIVELY ENUMERABLE SETS BY JOHN W. BERRYO) Abstract. An injective function on N, the nonnegative integers, taking values in TV, is called almost recursive (abbreviated a.r.) if its inverse has a partial recursive exten- sion. The range of an a.r. function /is called an almost recursively enumerable set in general; an almost recursive set if in addition/is strictly increasing. These are natural generalizations of regressive and retraceable sets respectively. We show that an infinite set is almost recursively enumerable iff it is point decomposable in the sense of McLaughlin. This leads us to new characterizations of certain classes of immune sets. Finally, in contrast to the regressive case, we show that a.r. functions and sets are rather badly behaved with respect to recursive equivalence. 1. Introduction and notation. Almost recursive functions, almost recursive sets and almost recursively enumerable sets were introduced by Vuckovic in [7]. For basic facts about retraceable sets the reader is referred to [3]; for regressive sets to [2] and [1]. Throughout this paper, N denotes the set of nonnegative integers. The word "set" will usually mean "subset of TV," and the complement TV—Aof a set A is denoted by A. Iff is a partial function on TVwe denote its domain and range by S(/) and p{f) respectively. Tre denotes the eth partial recursive function in the Kleene enumeration, i.e. ■ne{x)^U{p.yTx(e, x, y)). we denotes 8(TTe).As a pairing function we use j(x, y) = ^[(x+y)2 + 3y + x], and k(x), l(x) are recursive functions which satisfy x =j{k(x), l(x)) for all x, and k(j(x, y)) = x, l(j(x, y))=y for all x and y.
    [Show full text]
  • Minicourse on the Axioms of Zermelo and Fraenkel
    Minicourse on The Axioms of Zermelo and Fraenkel Heike Mildenberger January 15, 2021 Abstract These are the notes for a minicourse of four hours in the DAAD-summerschool in mathematical logic 2020. To be held virtually in G¨ottingen,October 10-16, 2020. 1 The Zermelo Fraenkel Axioms Axiom: a plausible postulate, seen as evident truth that needs not be proved. Logical axioms, e.g. A∧B → A for any statements A, B. We assume the Hilbert axioms, see below. Now we focus on the non-logical axioms, the ones about the underlying set the- oretic structure. Most of nowadays mathematics can be justified as built within the set theoretic universe or one of its extensions by classes or even a hierarchy of classes. In my lecture I will report on the Zermelo Fraenkel Axioms. If no axioms are mentioned, usually the proof is based on these axioms. However, be careful, for example nowaday’s known and accepted proof of Fermat’s last theorem by Wiles uses “ZFC and there are infinitely many strongly inaccessible cardinals”. It is conjectured that ZFC or even the Peano Arithmetic will do. The language of the Zermelo Fraenkel axioms is the first order logic L (τ) with signature τ = {∈}. The symbol ∈ stands for a binary relation. However, we will work in English and thus expand the language by many defined notions. If you are worried about how this is compatible with the statement “the language of set theory is L({∈})” you can study the theory of language expansions by defined symbols. E.g., you may read page 85ff in [15].
    [Show full text]
  • Logic of Paradoxes in Classical Set Theories
    Logic of Paradoxes in Classical Set Theories Boris Culinaˇ Department of Mathematics University of Applied Sciences Velika Gorica Zagrebaˇcka cesta 5, 10410 Velika Gorica, Croatia e-mail: [email protected] Logistic is not sterile; it engenders antinomies. H. Poincar´e Abstract According to Georg Cantor [Can83, Can99] a set is any multitude which can be thought of as one ("jedes Viele, welches sich als Eines denken l¨aßt”)without contradiction { a consistent multitude. Other multitudes are inconsistent or paradoxical. Set theoretical paradoxes have common root { lack of understanding why some multitudes are not sets. Why some multitudes of objects of thought cannot themselves be objects of thought? Moreover, it is a logical truth that such multitudes do exist. However we do not understand this logical truth so well as we understand, for example, the logical truth 8x x = x. In this paper we formulate a logical truth which we call the productivity principle. Bertrand Rusell [Rus06] was the first one to formulate this principle, but in a restricted form and with a different purpose. The principle explicates a logical mechanism that lies behind paradoxical multitudes, and is understandable as well as any simple logical truth. However, it does not explain the concept of set. It only sets logical bounds of the concept within the framework of the classical two valued 2 - language. The principle behaves as a logical regulator of any theory we formulate to explain and describe sets. It provides tools to identify paradoxical classes inside the theory. We show how the known paradoxical classes follow from the productivity principle and how the principle gives us a uniform way to generate new paradoxical classes.
    [Show full text]
  • Solutions to Tutorial for Week 4
    The University of Sydney School of Mathematics and Statistics Solutions to Tutorial for Week 4 MATH1921/1931: Calculus of One Variable (Advanced) Semester 1, 2018 Web Page: sydney.edu.au/science/maths/u/UG/JM/MATH1921/ Lecturer: Daniel Daners Material covered Definition of a function f : A ! B and composites, domain, codomain and image/range; Injective, surjective, and bijective functions; inverse functions. The concept of natural domain of a real valued function of a real variable. The hyperbolic sine and cosine functions sinh x and cosh x. Outcomes After completing this tutorial you should understand the concepts of domain, codomain and image/range of functions; be able to calculate the image/range of various functions; be able to prove whether given functions are injective, surjective or bijective and compute inverse functions; identify the natural domain of real valued functions of a real variable; work with the hyperbolic cosine and sine functions, and prove identities involving them. Summary of essential material The hyperbolic sine and cosine. The hyper- y bolic cosine and hyperbolic sine functions are de- cosh¹xº fined by ex + e−x cosh x = sinh¹xº 2 x −x e − e −x sinh x = ex 1 e 2 2 2 for all x 2 R. They share many properties with the cosine and sine functions as shown in some 0 x e−x questions below. − The graph of the hyperbolic cosine function is 2 the shape of a hanging cable or chain attached at two ends. Functions. Let A and B be sets. A function f : A ! B is a rule which assigns exactly one element of B to each element of A.
    [Show full text]
  • Lecture 10: Injection, Surjection, and Linear Maps 1 Injection, Surjection, Bijection and Size
    Math/CS 103 Professor: Padraic Bartlett Lecture 10: Injection, Surjection, and Linear Maps Week 5 UCSB 2013 This talk is designed to go over some of the concepts we've been exploring recently with injections, surjections, and linear maps. 1 Injection, Surjection, Bijection and Size We've been dealing with injective and surjective maps for a while now. Something you might have noticed, when looking at injective and surjective maps on finite sets, is the following triple of observations: Observation. If f : A ! B is an injective function and A; B are finite sets , then size(A) ≤ size(B). The reasoning for this, in the finite case, is relatively simple: 1. If f is injective, then each element in A is being sent to a different element in B. 2. Thus, you'll need B to have at least size(A)-many elements, in order to provide that many targets. Observation. If f : A ! B is an surjective function and A; B are finite sets , then jAj ≥ jBj. Basically, this holds true because 1. Thinking about f as a collection of arrows from A to B, it has precisely jAj-many arrows by definition, as each element in A gets to go to precisely one place in B. 2. Thus, if we have to hit every element in B, and we start with only jAj-many arrows, we need to have jAj ≥ jBj in order to hit everything. Observation. If f : A ! B is an bijective function and A; B are finite sets, then jAj = jBj. Proof. A bijection is a map that is both injective and surjective.
    [Show full text]
  • MATH 436 Notes: Functions and Inverses
    MATH 436 Notes: Functions and Inverses. Jonathan Pakianathan September 12, 2003 1 Functions Definition 1.1. Formally, a function f : A → B is a subset f of A × B with the property that for every a ∈ A, there is a unique element b ∈ B such that (a, b) ∈ f. The set A is called the domain of f and the set B is the codomain of f. While the above definition provides a definition of a function purely in terms of set theory, it is usually not a useful picture to work with. However it does emphasize the important point that the domain and codomain of a function are an intrinsic part of any function f. Less formally, we usually think of a function f : A → B as a rule of assignment which assigns a unique output f(a) ∈ B for every input a ∈ A. The graph of f, denoted Graph(f) = {(a, f(a))|a ∈ A} ⊂ A × B then recovers the more formal set theoretic definition of the function. Definition 1.2. Let f : A → B be a function. Given S ⊂ A we define f(S)= {f(s)|s ∈ S}. Note that f(S) ⊆ B. f(S) is called the image of the set S under f. f(A) is called the image of f, and is denoted Im(f). Given T ⊂ B we define f −1(T )= {a ∈ A|f(a) ∈ T }. Note that f −1(T ) ⊆ A. f −1(T ) is called the preimage of the set T under f. Fix a function f : A → B, then it is easy to see that for all S ⊂ A, S ⊂ f −1(f(S)) and for all T ⊆ B, we have f(f −1(T )) ⊆ T .
    [Show full text]
  • WEEK 6: MAPS on GROUPS A. Homomorphisms A.1. Definitions. Naturally We Can Consider Functions Mapping One Group to Another in Ex
    WEEK 6: MAPS ON GROUPS A. Homomorphisms A.1. Definitions. Naturally we can consider functions mapping one group to another in exactly the same way that we may look at functions relating any two sets. But if we gain nothing by limiting our discussion to groups unless the functions interact with the group structure in some way. Let G and H be groups. A function f : G ! H is a homomorphism if f(ab) = f(a)f(b): Note that the product on the left is with the operation on G, while that on the right is the operation on H. We might say that the map f preserves the operation. These maps need not be injective or surjective. For example, consider the following map from Z to D3: 8 r ; if n = 3q; <> 0 f(n) = r1; if n = 3q + 1; and > :r2; if n = 3q + 2; where q is some integer. Let's check to see that f : Z ! D3 is a homomorphism. If n = 3q + j and m = 3s + t, with 0 ≤ j; t < 3, we have f(n + m) = f ((3q + j) + (3s + t)) = f (3(q + s) + (j + t)) : The image of this element depends entirely on j +t: if j +t is divisible by 3, then f(n+m) = r0, if j + t − 1 is divisible by 3, then f(n + m) = r1, and if j + t − 2 is divisible by 3, then f(n + m) = r2. Thus, f(n+m) = ri if j +t = i mod 3. But the operation on D3 is such that ri = rjrt if j +t = i mod 3.
    [Show full text]
  • MATH 220 (All Sections)—Homework #12 Not to Be Turned in Posted Friday, November 24, 2017
    MATH 220 (all sections)—Homework #12 not to be turned in posted Friday, November 24, 2017 Definition: A set A is finite if there exists a nonnegative integer c such that there exists a bijection from A to fn 2 N: n ≤ cg. (The integer c is called the cardinality of A.) I. (a) Let A be a finite set, and let B be a subset of A. Prove that B is finite. (Hint: induction on jAj. Note that our proof can’t use induction on jBj, or indeed refer to “the number of elements in B” at all, because we don’t yet know that B is finite!) (b) Prove that the union of two disjoint finite sets is finite. (c) Prove that the union of any two finite sets is finite. (Hint: A [ B = A [ (B − A).) (a) We proceed by induction on the nonnegative integer c in the definition that A is finite (the cardinality of c). Basis step: c = 0. Then there is a bijection from A to fn 2 N: n ≤ 0g = ;, and thus A = ; (and, for that matter, the bijection is also the empty function). This forces B = ; as well (since that is the only subset of ;), which is certainly finite. Induction step: Let c ≥ 0. The induction hypothesis is: if A has cardinality c, then any subset of A is finite. Now suppose A has cardinality c + 1, and let f : A ! fn 2 −1 N: n ≤ c + 1g be a bijection. Define A1 = f (f1; : : : ; cg), and note that f restricted to A1 is a bijection from A1 to f1; : : : ; cg, so that A1 has cardinality c.
    [Show full text]