EQUIVALENCE RELATIONS and FUNCTIONS. 1. Definition. a Function F

Total Page:16

File Type:pdf, Size:1020Kb

EQUIVALENCE RELATIONS and FUNCTIONS. 1. Definition. a Function F EQUIVALENCE RELATIONS AND FUNCTIONS. 1. Definition. A function f : A ! B is said to be compatible with an equivalence relation R on A if: (8x; y 2 A)(xRy ! f(x) = f(y)): Given an equivalence relation R on A, we always have the function (called the \quotient projection") π : A ! A=R defined by π(x) = [x], the equiva- lence class of x (recall A=R is the set of all equivalence classes, so A=R is a \family of sets".) Problem 1. If f : A ! B is compatible with an equivalence relation R on A, show that the function (or \map") h : A=R ! B given by h([x]) = f(x) is well-defined, and satisfies h ◦ π = f. (\Well-defined” here means we get the same element h([x]) of B, regard- less of which element of A we pick as representative of the equivalence class [x]). Going in the other direction: any function f : A ! B defines an equiva- lence relation Rf in its domain A by the rule: xRf y $ f(x) = f(y); for x; y 2 A: Problem 2. Prove that this is an equivalence relation, and that f is compat- ible with Rf . Example 1. Two different functions f; g with the same domain A may induce the same equivalence relation Rf = Rg on A. For example, 2 f : R ! R; f(x) = x and g : R ! R; g(x) = jxj , induce the same equivalence relation on R, since: f(x) = f(¯x) $ x2 =x ¯2 $ jxj = jx¯j $ g(x) = g(¯x): Problem 3. Show that the functions sin x and cos x (both from R to R) define the same equivalence relation on R; describe a typical equivalence class (say, [π=3].) It is a common situation in Mathematics that one needs to idenfity the abstractly defined set A=R in terms of a more \concrete" set B, ideally by exhibiting a bijection between them. If we can find a set B and a function f : A ! B in such a way that R = Rf , from the preceding we'll have a function h : A=R ! B; so we need to find conditions for the injectivity and surjectivity of h. 1 Problem 4. Let f : A ! B be compatible with an equivalence relation R on A. Let h : A=R ! B be defined as above: h([x]) = f(x). Prove that: h is injective $ (8x; y 2 A)(f(x) = f(y) ! xRy): In particular, this shows that for any function f : A ! B, the function h : A=Rf ! B (defined by the equivalence relation Rf ) is always injective. Problem 5. Prove that if f : A ! B is surjective, then h : A=Rf ! B is also surjective. (From Problem 4 we already know it is injective, hence it is always bijective.) Conclusion: if R is an equivalence relation on A and we can find a set B and a function f : A ! B so that R = Rf (that is, f induces the same partition of A as R does) and f is surjective, then the function h : A=R ! B given by h([x]) = f(x) is well-defined, and a bijection. Example 2. Consider the equivalence relation ∼ on R defined by: n x ∼ y $ x = 10 y; for some n 2 Z: Let B ⊂ R be the set: B = (−10; −1] [ f0g [ [1; 10): We can use the results above to construct a bijection h : R= ∼→ B. For this it suffices to find a surjective function f : R ! B so that Rf and ∼ are the same equivalence relation on R. (The latter means: f(x) = f(y) $ x ∼ y, for all x; y 2 R.) Let f : R ! B be defined as follows; (a) f(0) = 0; (b) If x > 0, there is a n n+1 −n unique n 2 Z so that 10 ≤ x < 10 , or 1 ≤ 10 x < 10. In this case, set: f(x) = 10−nx 2 B. (c) If x < 0, we have −10n+1 < x ≤ −10n for a unique −n −n n 2 Z, or −10 < 10 x ≤ −1. In this case we set f(x) = 10 x 2 B. Problem 6. Show that f : R ! B is surjective. Problem 7. Show that Rf and ∼ are the same equivalence relation on R, that is: 8x; y 2 R x ∼ y $ f(x) = f(y): Hint: consider separately the cases (i)x = y = 0; (ii) x > 0; y > 0; (iii) x < 0; y < 0. This is possible since all the real numbers in a given equivalence class (other than [0] = f0g) are positive, or all are negative, and f(x) has the same sign as x, for any x (including f(0) = 0.) Thus the corresponding function h : R= ∼→ B given by h([x]) = f(x) is well-defined, and a bijection. 2 Example. It is geometrically clear that \oriented directions in the plane are parametrized by the unit circle", but it is useful to make this more 2 precise. Two nonzero vectors in R define the same oriented direction if they one can be obtained from the other by multiplying by a positive scalar. 2 This suggests setting up the equivalence relation on R n f0g: v ∼ w $ (9c > 0)(v = cw): (Exercise: this is an equivalence relation.) We now want to define a bijection 2 2 h :(R n f0g)= ∼→ C, where C = fv 2 R ; jvj = 1g is the unit circle, and p 2 2 jvj = x1 + x2 if v = (x1; x2). From the discussion above, it is enough to 2 find a function f : R ! C such that (i) f is surjective and (ii) f and ∼ 2 define the same equivalence relation on R n f0g. Then h([v]) = f(v) defines 2 the sought-after bijection h :(R n f0g)= ∼→ C. It is natural to consider the function: v f : 2 n f0g ! C; f(v) = : R jvj (i) f is surjective. Note that (8v 2 C)(f(v) = v) (since jvj = 1 for v 2 C). 2 Hence any vector v 2 C is the image of under f of a vector in R n f0g (namely v itself), so f is surjective. Next we need to show: 2 (8v; w 2 R n f0g)(v ∼ w $ f(v) = f(w)): 2 Proof. Let v; w 2 R n f0g be arbitrary. First, assume f(v) = f(w), or v w jvj = jwj . Then letting c = jvj=jwj we have v = cw with c > 0, so v ∼ w. For the converse, assume v ∼ w. Then v = cw for some c > 0, so v w jvj = cjwj and dividing these two equalities we see that: jvj = jwj , or f(v) = f(w): This ends the proof. Exercise. Now try to repeat the above to parametrize the set of unori- ented directions in the plane (any family of parallel lines defines the same 2 unoriented direction.) The equivalence relation in R n f0g is: v ∼ w $ (9c 6= 0)(v = cw): (We no longer require c to be positive.) Show that this is an equivalence rela- tion. It is geometrically natural to try to parametrize the set of equivalence 3 classes (which is the set of unoriented directions| by the upper semicircle, including its right endpoint: 2 C+ = fv 2 R n f0gjv = (x1; x2); x2 ≥ 0g n f(−1; 0)g: 2 Find a function f : R n f0g ! C+ with the properties: f(v) = v for v 2 C+ (so f is onto) (ii) f defines the same equivalence relation as ∼: v ∼ w $ f(v) = f(w). 4.
Recommended publications
  • I0 and Rank-Into-Rank Axioms
    I0 and rank-into-rank axioms Vincenzo Dimonte∗ July 11, 2017 Abstract Just a survey on I0. Keywords: Large cardinals; Axiom I0; rank-into-rank axioms; elemen- tary embeddings; relative constructibility; embedding lifting; singular car- dinals combinatorics. 2010 Mathematics Subject Classifications: 03E55 (03E05 03E35 03E45) 1 Informal Introduction to the Introduction Ok, that’s it. People who know me also know that it is years that I am ranting about a book about I0, how it is important to write it and publish it, etc. I realized that this particular moment is still right for me to write it, for two reasons: time is scarce, and probably there are still not enough results (or anyway not enough different lines of research). This has the potential of being a very nasty vicious circle: there are not enough results about I0 to write a book, but if nobody writes a book the diffusion of I0 is too limited, and so the number of results does not increase much. To avoid this deadlock, I decided to divulge a first draft of such a hypothetical book, hoping to start a “conversation” with that. It is literally a first draft, so it is full of errors and in a liquid state. For example, I still haven’t decided whether to call it I0 or I0, both notations are used in literature. I feel like it still lacks a vision of the future, a map on where the research could and should going about I0. Many proofs are old but never published, and therefore reconstructed by me, so maybe they are wrong.
    [Show full text]
  • Chapter 1 Logic and Set Theory
    Chapter 1 Logic and Set Theory To criticize mathematics for its abstraction is to miss the point entirely. Abstraction is what makes mathematics work. If you concentrate too closely on too limited an application of a mathematical idea, you rob the mathematician of his most important tools: analogy, generality, and simplicity. – Ian Stewart Does God play dice? The mathematics of chaos In mathematics, a proof is a demonstration that, assuming certain axioms, some statement is necessarily true. That is, a proof is a logical argument, not an empir- ical one. One must demonstrate that a proposition is true in all cases before it is considered a theorem of mathematics. An unproven proposition for which there is some sort of empirical evidence is known as a conjecture. Mathematical logic is the framework upon which rigorous proofs are built. It is the study of the principles and criteria of valid inference and demonstrations. Logicians have analyzed set theory in great details, formulating a collection of axioms that affords a broad enough and strong enough foundation to mathematical reasoning. The standard form of axiomatic set theory is denoted ZFC and it consists of the Zermelo-Fraenkel (ZF) axioms combined with the axiom of choice (C). Each of the axioms included in this theory expresses a property of sets that is widely accepted by mathematicians. It is unfortunately true that careless use of set theory can lead to contradictions. Avoiding such contradictions was one of the original motivations for the axiomatization of set theory. 1 2 CHAPTER 1. LOGIC AND SET THEORY A rigorous analysis of set theory belongs to the foundations of mathematics and mathematical logic.
    [Show full text]
  • Elementary Functions Part 1, Functions Lecture 1.6D, Function Inverses: One-To-One and Onto Functions
    Elementary Functions Part 1, Functions Lecture 1.6d, Function Inverses: One-to-one and onto functions Dr. Ken W. Smith Sam Houston State University 2013 Smith (SHSU) Elementary Functions 2013 26 / 33 Function Inverses When does a function f have an inverse? It turns out that there are two critical properties necessary for a function f to be invertible. The function needs to be \one-to-one" and \onto". Smith (SHSU) Elementary Functions 2013 27 / 33 Function Inverses When does a function f have an inverse? It turns out that there are two critical properties necessary for a function f to be invertible. The function needs to be \one-to-one" and \onto". Smith (SHSU) Elementary Functions 2013 27 / 33 Function Inverses When does a function f have an inverse? It turns out that there are two critical properties necessary for a function f to be invertible. The function needs to be \one-to-one" and \onto". Smith (SHSU) Elementary Functions 2013 27 / 33 Function Inverses When does a function f have an inverse? It turns out that there are two critical properties necessary for a function f to be invertible. The function needs to be \one-to-one" and \onto". Smith (SHSU) Elementary Functions 2013 27 / 33 One-to-one functions* A function like f(x) = x2 maps two different inputs, x = −5 and x = 5, to the same output, y = 25. But with a one-to-one function no pair of inputs give the same output. Here is a function we saw earlier. This function is not one-to-one since both the inputs x = 2 and x = 3 give the output y = C: Smith (SHSU) Elementary Functions 2013 28 / 33 One-to-one functions* A function like f(x) = x2 maps two different inputs, x = −5 and x = 5, to the same output, y = 25.
    [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]
  • Mereology Then and Now
    Logic and Logical Philosophy Volume 24 (2015), 409–427 DOI: 10.12775/LLP.2015.024 Rafał Gruszczyński Achille C. Varzi MEREOLOGY THEN AND NOW Abstract. This paper offers a critical reconstruction of the motivations that led to the development of mereology as we know it today, along with a brief description of some questions that define current research in the field. Keywords: mereology; parthood; formal ontology; foundations of mathe- matics 1. Introduction Understood as a general theory of parts and wholes, mereology has a long history that can be traced back to the early days of philosophy. As a formal theory of the part-whole relation or rather, as a theory of the relations of part to whole and of part to part within a whole it is relatively recent and came to us mainly through the writings of Edmund Husserl and Stanisław Leśniewski. The former were part of a larger project aimed at the development of a general framework for formal ontology; the latter were inspired by a desire to provide a nominalistically acceptable alternative to set theory as a foundation for mathematics. (The name itself, ‘mereology’ after the Greek word ‘µρoς’, ‘part’ was coined by Leśniewski [31].) As it turns out, both sorts of motivation failed to quite live up to expectations. Yet mereology survived as a theory in its own right and continued to flourish, often in unexpected ways. Indeed, it is not an exaggeration to say that today mereology is a central and powerful area of research in philosophy and philosophical logic. It may be helpful, therefore, to take stock and reconsider its origins.
    [Show full text]
  • Solutions to Homework Set 5
    MAT320/640, Fall 2020 Renato Ghini Bettiol Solutions to Homework Set 5 1. Decide if each of the statements below is true or false. If it is true, give a complete proof; if it is false, give an explicit counter-example. (Recall that a function f : X ! Y is called surjective, or onto, if every point of Y belongs to its image, that is, if f(X) = Y .) (a) There exists a continuous function f : R n f0g ! R. (b) There exists a continuous surjective function f : R n f0g ! R. (c) There exists a continuous function f : [0; 1] ! R. (d) There exists a continuous surjective function f : [0; 1] ! R. (e) The function f : R ! R, f(x) = x2, is uniformly continuous. (f) The function f : [0; 1] ! R, f(x) = x2, uniformly continuous. Solution: (a) TRUE: For example, any constant function f : R n f0g ! R is continuous; e.g., f(x) = 0 for all x 2 R n f0g. (b) TRUE: For example, consider the function f : R n f0g ! R given by ( x; if x < 0; f(x) = x − 1; if x > 0: Since the domain of f is R n f0g, it is continuous on its domain. Moreover, f is clearly surjective (draw its graph). (c) TRUE: For example, any constant function f : [0; 1] ! R is continuous; e.g., f(x) = 0 for all x 2 [0; 1]. (d) FALSE: Suppose, by contradiction, that such a function f : [0; 1] ! R exists. Since [0; 1] is compact and f : [0; 1] ! R is continuous, its image f([0; 1]) is compact.
    [Show full text]
  • 06. Naive Set Theory I
    Topics 06. Naive Set Theory I. Sets and Paradoxes of the Infinitely Big II. Naive Set Theory I. Sets and Paradoxes of the Infinitely Big III. Cantor and Diagonal Arguments Recall: Paradox of the Even Numbers Claim: There are just as many even natural numbers as natural numbers natural numbers = non-negative whole numbers (0, 1, 2, ..) Proof: { 0 , 1 , 2 , 3 , 4 , ............ , n , ..........} { 0 , 2 , 4 , 6 , 8 , ............ , 2n , ..........} In general: 2 criteria for comparing sizes of sets: (1) Correlation criterion: Can members of one set be paired with members of the other? (2) Subset criterion: Do members of one set belong to the other? Can now say: (a) There are as many even naturals as naturals in the correlation sense. (b) There are less even naturals than naturals in the subset sense. To talk about the infinitely Moral: notion of sets big, just need to be clear makes this clear about what’s meant by size Bolzano (1781-1848) Promoted idea that notion of infinity was fundamentally set-theoretic: To say something is infinite is “God is infinite in knowledge” just to say there is some set means with infinite members “The set of truths known by God has infinitely many members” SO: Are there infinite sets? “a many thought of as a one” -Cantor Bolzano: Claim: The set of truths is infinite. Proof: Let p1 be a truth (ex: “Plato was Greek”) Let p2 be the truth “p1 is a truth”. Let p3 be the truth “p3 is a truth”. In general, let pn be the truth “pn-1 is a truth”, for any natural number n.
    [Show full text]
  • Logic and Proof Release 3.18.4
    Logic and Proof Release 3.18.4 Jeremy Avigad, Robert Y. Lewis, and Floris van Doorn Sep 10, 2021 CONTENTS 1 Introduction 1 1.1 Mathematical Proof ............................................ 1 1.2 Symbolic Logic .............................................. 2 1.3 Interactive Theorem Proving ....................................... 4 1.4 The Semantic Point of View ....................................... 5 1.5 Goals Summarized ............................................ 6 1.6 About this Textbook ........................................... 6 2 Propositional Logic 7 2.1 A Puzzle ................................................. 7 2.2 A Solution ................................................ 7 2.3 Rules of Inference ............................................ 8 2.4 The Language of Propositional Logic ................................... 15 2.5 Exercises ................................................. 16 3 Natural Deduction for Propositional Logic 17 3.1 Derivations in Natural Deduction ..................................... 17 3.2 Examples ................................................. 19 3.3 Forward and Backward Reasoning .................................... 20 3.4 Reasoning by Cases ............................................ 22 3.5 Some Logical Identities .......................................... 23 3.6 Exercises ................................................. 24 4 Propositional Logic in Lean 25 4.1 Expressions for Propositions and Proofs ................................. 25 4.2 More commands ............................................
    [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]
  • Math 127: Equivalence Relations
    Math 127: Equivalence Relations Mary Radcliffe 1 Equivalence Relations Relations can take many forms in mathematics. In these notes, we focus especially on equivalence relations, but there are many other types of relations (such as order relations) that exist. Definition 1. Let X; Y be sets. A relation R = R(x; y) is a logical formula for which x takes the range of X and y takes the range of Y , sometimes called a relation from X to Y . If R(x; y) is true, we say that x is related to y by R, and we write xRy to indicate that x is related to y by R. Example 1. Let f : X ! Y be a function. We can define a relation R by R(x; y) ≡ (f(x) = y). Example 2. Let X = Y = Z. We can define a relation R by R(a; b) ≡ ajb. There are many other examples at hand, such as ordering on R, multiples in Z, coprimality relationships, etc. The definition we have here is simply that a relation gives some way to connect two elements to each other, that can either be true or false. Of course, that's not a very useful thing, so let's add some conditions to make the relation carry more meaning. For this, we shall focus on relations from X to X, also called relations on X. There are several properties that will be interesting in considering relations: Definition 2. Let X be a set, and let ∼ be a relation on X. • We say that ∼ is reflexive if x ∼ x 8x 2 X.
    [Show full text]
  • Well-Orderings, Ordinals and Well-Founded Relations. an Ancient
    Well-orderings, ordinals and well-founded relations. An ancient principle of arithmetic, that if there is a non-negative integer with some property then there is a least such, is useful in two ways: as a source of proofs, which are then said to be \by induction" and as as a source of definitions, then said to be \by recursion." An early use of induction is in Euclid's proof that every integer > 2 is a product of primes, (where we take \prime" to mean \having no divisors other than itself and 1"): if some number is not, then let n¯ be the least counter-example. If not itself prime, it can be written as a product m1m2 of two strictly smaller numbers each > 2; but then each of those is a product of primes, by the minimality of n¯; putting those two products together expresses n¯ as a product of primes. Contradiction ! An example of definition by recursion: we set 0! = 1; (n + 1)! = n! × (n + 1): A function defined for all non-negative integers is thereby uniquely specified; in detail, we consider an attempt to be a function, defined on a finite initial segment of the non-negative integers, which agrees with the given definition as far as it goes; if some integer is not in the domain of any attempt, there will be a least such; it cannot be 0; if it is n + 1, the recursion equation tells us how to extend an attempt defined at n to one defined at n + 1. So no such failure exists; we check that if f and g are two attempts and both f(n) and g(n) are defined, then f(n) = g(n), by considering the least n where that might fail, and again reaching a contradiction; and so, there being no disagreement between any two attempts, the union of all attempts will be a well-defined function, which is familiar to us as the factorial function.
    [Show full text]
  • Relation Liftings on Preorders and Posets
    Relation Liftings on Preorders and Posets Marta B´ılkov´a1, Alexander Kurz2,, Daniela Petri¸san2,andJiˇr´ı Velebil3, 1 Faculty of Philosophy, Charles University, Prague, Czech Republic [email protected] 2 Department of Computer Science, University of Leicester, United Kingdom {kurz,petrisan}@mcs.le.ac.uk 3 Faculty of Electrical Engineering, Czech Technical University in Prague, Czech Republic [email protected] Abstract. The category Rel(Set) of sets and relations can be described as a category of spans and as the Kleisli category for the powerset monad. A set-functor can be lifted to a functor on Rel(Set) iff it preserves weak pullbacks. We show that these results extend to the enriched setting, if we replace sets by posets or preorders. Preservation of weak pullbacks becomes preservation of exact lax squares. As an application we present Moss’s coalgebraic over posets. 1 Introduction Relation lifting [Ba, CKW, HeJ] plays a crucial role in coalgebraic logic, see eg [Mo, Bal, V]. On the one hand, it is used to explain bisimulation: If T : Set −→ Set is a functor, then the largest bisimulation on a coalgebra ξ : X −→ TX is the largest fixed point of the operator (ξ × ξ)−1 ◦ T on relations on X,whereT is the lifting of T to Rel(Set) −→ Rel(Set). (The precise meaning of ‘lifting’ will be given in the Extension Theorem 5.3.) On the other hand, Moss’s coalgebraic logic [Mo] is given by adding to propo- sitional logic a modal operator ∇, the semantics of which is given by applying T to the forcing relation ⊆ X ×L,whereL is the set of formulas: If α ∈ T (L), then x ∇α ⇔ ξ(x) T () α.
    [Show full text]