Probability Theory

Total Page:16

File Type:pdf, Size:1020Kb

Probability Theory Probability Theory "A random variable is neither random nor variable." Gian-Carlo Rota, M.I.T.. Florian Herzog 2013 Probability space Probability space A probability space W is a unique triple W = fΩ; F;P g: • Ω is its sample space •F its σ-algebra of events • P its probability measure Remarks: (1) The sample space Ω is the set of all possible samples or elementary events !: Ω = f! j ! 2 Ωg. (2)The σ-algebra F is the set of all of the considered events A, i.e., subsets of Ω: F = fA j A ⊆ Ω;A 2 Fg. (3) The probability measure P assigns a probability P (A) to every event A 2 F: P : F! [0; 1]. Stochastic Systems, 2013 2 Sample space The sample space Ω is sometimes called the universe of all samples or possible outcomes !. Example 1. Sample space • Toss of a coin (with head and tail): Ω = fH;T g. • Two tosses of a coin: Ω = fHH;HT;TH;TT g. • A cubic die: Ω = f!1;!2;!3;!4;!5;!6g. • The positive integers: Ω = f1; 2; 3;::: g. • The reals: Ω = f! j ! 2 Rg. Note that the !s are a mathematical construct and have per se no real or scientific meaning. The !s in the die example refer to the numbers of dots observed when the die is thrown. Stochastic Systems, 2013 3 Event An event A is a subset of Ω. If the outcome ! of the experiment is in the subset A, then the event A is said to have occurred. The set of all subsets of the sample space are denoted by 2Ω. Example 2. Events • Head in the coin toss: A = fHg. • Odd number in the roll of a die: A = f!1;!3;!5g. • An integer smaller than 5: A = f1; 2; 3; 4g, where Ω = f1; 2; 3;::: g: • A real number between 0 and 1: A = [0; 1], where Ω = f! j ! 2 Rg. We denote the complementary event of A by Ac = ΩnA. When it is possible to determine whether an event A has occurred or not, we must also be able to determine whether Ac has occurred or not. Stochastic Systems, 2013 4 Probability Measure I Definition 1. Probability measure A probability measure P on the countable sample space Ω is a set function P : F! [0; 1]; satisfying the following conditions • P(Ω) = 1. • P(!i) = pi. • If A1;A2;A3; ::: 2 F are mutually disjoint, then 1 1 [ X P Ai = P (Ai): i=1 i=1 Stochastic Systems, 2013 5 Probability The story so far: • Sample space: Ω = f!1;:::;!ng, finite! • Events: F = 2Ω: All subsets of Ω P • Probability: P (!i) = pi ) P (A 2 Ω) = pi wi2A Probability axioms of Kolmogorov (1931) for elementary probability: • P(Ω) = 1. • If A 2 Ω then P (A) ≥ 0. • If A1;A2;A3; ::: 2 Ω are mutually disjoint, then 1 1 [ X P Ai = P (Ai): i=1 i=1 Stochastic Systems, 2013 6 Uncountable sample spaces n Most important uncountable sample space for engineering: R, resp. R . Consider the example Ω = [0; 1], every ! is equally "likely". • Obviously, P(!) = 0. • Intuitively, P([0; a]) = a, basic concept: length! Question: Has every subset of [0; 1] a determinable length? Answer: No! (e.g. Vitali sets, Banach-Tarski paradox) Question: Is this of importance in practice? Answer: No! Question: Does it matter for the underlying theory? Answer: A lot! Stochastic Systems, 2013 7 Fundamental mathematical tools Not every subset of [0; 1] has a determinable length ) collect the ones with a determinable length in F. Such a mathematical construct, which has additional, desirable properties, is called σ-algebra. Definition 2. σ-algebra A collection F of subsets of Ω is called a σ-algebra on Ω if • Ω 2 F and ; 2 F (; denotes the empty set) • If A 2 F then ΩnA = Ac 2 F: The complementary subset of A is also in Ω S1 • For all Ai 2 F: i=1 Ai 2 F The intuition behind it: collect all events in the σ-algebra F, make sure that by performing countably many elementary set operation ([; \;c ) on elements of F yields again an element in F (closeness). The pair fΩ; Fg is called measure space. Stochastic Systems, 2013 8 Example of σ-algebra Example 3. σ-algebra of two coin-tosses • Ω = fHH;HT;TH;TT g = f!1;!2;!3;!4g •F min = f;; Ωg = f;; f!1;!2;!3;!4gg. •F 1 = f;; f!1;!2g; f!3;!4g; f!1;!2;!3;!4gg. •F max = f;; f!1g; f!2g; f!3g; f!4g; f!1;!2g; f!1;!3g; f!1;!4g; f!2;!3g; f!2;!4g; f!3;!4g; f!1;!2;!3g; f!1;!2;!4g; f!1;!3;!4g; f!2;!3;!4g; Ωg: Stochastic Systems, 2013 9 Generated σ-algebras The concept of generated σ-algebras is important in probability theory. Definition 3. σ(C): σ-algebra generated by a class C of subsets Let C be a class of subsets of Ω. The σ-algebra generated by C, denoted by σ(C), is the smallest σ-algebra F which includes all elements of C, i.e., C 2 F. Identify the different events we can measure of an experiment (denoted by A), we then just work with the σ-algebra generated by A and have avoided all the measure theoretic technicalities. Stochastic Systems, 2013 10 Borel σ-algebra The Borel σ-algebra includes all subsets of R which are of interest in practical applications (scientific or engineering). Definition 4. Borel σ-algebra B(R)) The Borel σ-algebra B(R) is the smallest σ-algebra containing all open intervals in R. The sets in B(R) are called Borel sets. The extension to the n multi-dimensional case, B(R ), is straightforward. • (−∞; a), (b; 1), (−∞; a) [ (b; 1) • [a; b] = (−∞; a) [ (b; 1), S1 S1 • (−∞; a] = n=1[a − n; a] and [b; 1) = n=1[b; b + n], • (a; b] = (−∞; b] \ (a; 1), T1 1 1 •f ag = n=1(a − n; a + n), Sn •f a1; ··· ; ang = k=1 ak. Stochastic Systems, 2013 11 Measure Definition 5. Measure Let F be the σ-algebra of Ω and therefore (Ω; F) be a measurable space. The map µ : F! [0; 1] is called a measure on (Ω; F) if µ is countably additive. The measure µ is countably additive (or σ-additive) if µ(;) = 0 and for every sequence of disjoint sets (F : i 2 ) in F with F = S F we have i N i2N i X µ(F ) = µ(Fi): i2N If µ is countably additive, it is also additive, meaning for every F; G 2 F we have µ(F [ G) = µ(F ) + µ(G) if and only if F \ G = ; The triple (Ω; F; µ) is called a measure space. Stochastic Systems, 2013 12 Lebesgue Measure The measure of length on the straight line is known as the Lebesgue measure. Definition 6. Lebesgue measure on B(R) The Lebesgue measure on B(R), denoted by λ, is defined as the measure on (R; B(R)) which assigns the measure of each interval to be its length. Examples: • Lebesgue measure of one point: λ(fag) = 0. P1 • Lebesgue measure of countably many points: λ(A) = i=1 λ(faig) = 0. • The Lebesgue measure of a set containing uncountably many points: { zero { positive and finite { infinite Stochastic Systems, 2013 13 Probability Measure Definition 7. Probability measure A probability measure P on the sample space Ω with σ-algebra F is a set function P : F! [0; 1]; satisfying the following conditions • P(Ω) = 1. • If A 2 F then P (A) ≥ 0. • If A1;A2;A3; ::: 2 F are mutually disjoint, then 1 1 [ X P Ai = P (Ai): i=1 i=1 The triple (Ω; F;P ) is called a probability space. Stochastic Systems, 2013 14 F-measurable functions Definition 8. F-measurable function The function f :Ω ! R defined on (Ω; F;P ) is called F-measurable if −1 f (B) = f! 2 Ω: f(!) 2 Bg 2 F for all B 2 B(R); −1 i.e., the inverse f maps all of the Borel sets B ⊂ R to F. Sometimes it is easier to work with following equivalent condition: y 2 R ) f! 2 Ω: f(!) ≤ yg 2 F This means that once we know the (random) value X(!) we know which of the events in F have happened. •F = f;; Ωg: only constant functions are measurable •F = 2Ω: all functions are measurable Stochastic Systems, 2013 15 F-measurable functions Ω: Sample space, Ai: Event, F = σ(A1;:::;An): σ-algebra of events X(!):Ω 7! R R 6 j X(!) 2 ! R A2 A1 t (−∞; a] 2 B Y A3 −1 Ω A4 X : B 7! F X: random variable, B: Borel σ-algebra Stochastic Systems, 2013 16 F-measurable functions - Example Roll of a die: Ω = f!1;!2;!3;!4;!5;!6g We only know, whether an even or and odd number has shown up: F = f;; f!1;!3;!5g; f!2;!4;!6g; Ωg = σ(f!1;!3;!5g) Consider the following random variable: 1; if ! = ! ;! ;! ; f(!) = 1 2 3 −1; if ! = !4;!5;!6. Check measurability with the condition y 2 R ) f! 2 Ω: f(!) ≤ yg 2 F f! 2 Ω: f(!) ≤ 0g = f!4;!5;!6g 2= F ) f is not F-measurable. Stochastic Systems, 2013 17 Lebesgue integral I Definition 9.
Recommended publications
  • Measure-Theoretic Probability I
    Measure-Theoretic Probability I Steven P.Lalley Winter 2017 1 1 Measure Theory 1.1 Why Measure Theory? There are two different views – not necessarily exclusive – on what “probability” means: the subjectivist view and the frequentist view. To the subjectivist, probability is a system of laws that should govern a rational person’s behavior in situations where a bet must be placed (not necessarily just in a casino, but in situations where a decision must be made about how to proceed when only imperfect information about the outcome of the decision is available, for instance, should I allow Dr. Scissorhands to replace my arthritic knee by a plastic joint?). To the frequentist, the laws of probability describe the long- run relative frequencies of different events in “experiments” that can be repeated under roughly identical conditions, for instance, rolling a pair of dice. For the frequentist inter- pretation, it is imperative that probability spaces be large enough to allow a description of an experiment, like dice-rolling, that is repeated infinitely many times, and that the mathematical laws should permit easy handling of limits, so that one can make sense of things like “the probability that the long-run fraction of dice rolls where the two dice sum to 7 is 1/6”. But even for the subjectivist, the laws of probability should allow for description of situations where there might be a continuum of possible outcomes, or pos- sible actions to be taken. Once one is reconciled to the need for such flexibility, it soon becomes apparent that measure theory (the theory of countably additive, as opposed to merely finitely additive measures) is the only way to go.
    [Show full text]
  • Effective Program Reasoning Using Bayesian Inference
    EFFECTIVE PROGRAM REASONING USING BAYESIAN INFERENCE Sulekha Kulkarni A DISSERTATION in Computer and Information Science Presented to the Faculties of the University of Pennsylvania in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy 2020 Supervisor of Dissertation Mayur Naik, Professor, Computer and Information Science Graduate Group Chairperson Mayur Naik, Professor, Computer and Information Science Dissertation Committee: Rajeev Alur, Zisman Family Professor of Computer and Information Science Val Tannen, Professor of Computer and Information Science Osbert Bastani, Research Assistant Professor of Computer and Information Science Suman Nath, Partner Research Manager, Microsoft Research, Redmond To my father, who set me on this path, to my mother, who leads by example, and to my husband, who is an infinite source of courage. ii Acknowledgments I want to thank my advisor Prof. Mayur Naik for giving me the invaluable opportunity to learn and experiment with different ideas at my own pace. He supported me through the ups and downs of research, and helped me make the Ph.D. a reality. I also want to thank Prof. Rajeev Alur, Prof. Val Tannen, Prof. Osbert Bastani, and Dr. Suman Nath for serving on my dissertation committee and for providing valuable feedback. I am deeply grateful to Prof. Alur and Prof. Tannen for their sound advice and support, and for going out of their way to help me through challenging times. I am also very grateful for Dr. Nath's able and inspiring mentorship during my internship at Microsoft Research, and during the collaboration that followed. Dr. Aditya Nori helped me start my Ph.D.
    [Show full text]
  • Set Theory, Including the Axiom of Choice) Plus the Negation of CH
    ANNALS OF ~,IATltEMATICAL LOGIC - Volume 2, No. 2 (1970) pp. 143--178 INTERNAL COHEN EXTENSIONS D.A.MARTIN and R.M.SOLOVAY ;Tte RockeJ~'ller University and University (.~t CatiJbrnia. Berkeley Received 2 l)ecemt)er 1969 Introduction Cohen [ !, 2] has shown that tile continuum hypothesis (CH) cannot be proved in Zermelo-Fraenkel set theory. Levy and Solovay [9] have subsequently shown that CH cannot be proved even if one assumes the existence of a measurable cardinal. Their argument in tact shows that no large cardinal axiom of the kind present;y being considered by set theorists can yield a proof of CH (or of its negation, of course). Indeed, many set theorists - including the authors - suspect that C1t is false. But if we reject CH we admit Gurselves to be in a state of ignorance about a great many questions which CH resolves. While CH is a power- full assertion, its negation is in many ways quite weak. Sierpinski [ 1 5 ] deduces propcsitions there called C l - C82 from CH. We know of none of these propositions which is decided by the negation of CH and only one of them (C78) which is decided if one assumes in addition that a measurable cardinal exists. Among the many simple questions easily decided by CH and which cannot be decided in ZF (Zerme!o-Fraenkel set theory, including the axiom of choice) plus the negation of CH are tile following: Is every set of real numbers of cardinality less than tha't of the continuum of Lebesgue measure zero'? Is 2 ~0 < 2 ~ 1 ? Is there a non-trivial measure defined on all sets of real numbers? CIhis third question could be decided in ZF + not CH only in the unlikely event t Tile second author received support from a Sloan Foundation fellowship and tile National Science Foundation Grant (GP-8746).
    [Show full text]
  • A Convenient Category for Higher-Order Probability Theory
    A Convenient Category for Higher-Order Probability Theory Chris Heunen Ohad Kammar Sam Staton Hongseok Yang University of Edinburgh, UK University of Oxford, UK University of Oxford, UK University of Oxford, UK Abstract—Higher-order probabilistic programming languages 1 (defquery Bayesian-linear-regression allow programmers to write sophisticated models in machine 2 let let sample normal learning and statistics in a succinct and structured way, but step ( [f ( [s ( ( 0.0 3.0)) 3 sample normal outside the standard measure-theoretic formalization of proba- b ( ( 0.0 3.0))] 4 fn + * bility theory. Programs may use both higher-order functions and ( [x] ( ( s x) b)))] continuous distributions, or even define a probability distribution 5 (observe (normal (f 1.0) 0.5) 2.5) on functions. But standard probability theory does not handle 6 (observe (normal (f 2.0) 0.5) 3.8) higher-order functions well: the category of measurable spaces 7 (observe (normal (f 3.0) 0.5) 4.5) is not cartesian closed. 8 (observe (normal (f 4.0) 0.5) 6.2) Here we introduce quasi-Borel spaces. We show that these 9 (observe (normal (f 5.0) 0.5) 8.0) spaces: form a new formalization of probability theory replacing 10 (predict :f f))) measurable spaces; form a cartesian closed category and so support higher-order functions; form a well-pointed category and so support good proof principles for equational reasoning; and support continuous probability distributions. We demonstrate the use of quasi-Borel spaces for higher-order functions and proba- bility by: showing that a well-known construction of probability theory involving random functions gains a cleaner expression; and generalizing de Finetti’s theorem, that is a crucial theorem in probability theory, to quasi-Borel spaces.
    [Show full text]
  • Geometric Integration Theory Contents
    Steven G. Krantz Harold R. Parks Geometric Integration Theory Contents Preface v 1 Basics 1 1.1 Smooth Functions . 1 1.2Measures.............................. 6 1.2.1 Lebesgue Measure . 11 1.3Integration............................. 14 1.3.1 Measurable Functions . 14 1.3.2 The Integral . 17 1.3.3 Lebesgue Spaces . 23 1.3.4 Product Measures and the Fubini–Tonelli Theorem . 25 1.4 The Exterior Algebra . 27 1.5 The Hausdorff Distance and Steiner Symmetrization . 30 1.6 Borel and Suslin Sets . 41 2 Carath´eodory’s Construction and Lower-Dimensional Mea- sures 53 2.1 The Basic Definition . 53 2.1.1 Hausdorff Measure and Spherical Measure . 55 2.1.2 A Measure Based on Parallelepipeds . 57 2.1.3 Projections and Convexity . 57 2.1.4 Other Geometric Measures . 59 2.1.5 Summary . 61 2.2 The Densities of a Measure . 64 2.3 A One-Dimensional Example . 66 2.4 Carath´eodory’s Construction and Mappings . 67 2.5 The Concept of Hausdorff Dimension . 70 2.6 Some Cantor Set Examples . 73 i ii CONTENTS 2.6.1 Basic Examples . 73 2.6.2 Some Generalized Cantor Sets . 76 2.6.3 Cantor Sets in Higher Dimensions . 78 3 Invariant Measures and the Construction of Haar Measure 81 3.1 The Fundamental Theorem . 82 3.2 Haar Measure for the Orthogonal Group and the Grassmanian 90 3.2.1 Remarks on the Manifold Structure of G(N,M).... 94 4 Covering Theorems and the Differentiation of Integrals 97 4.1 Wiener’s Covering Lemma and its Variants .
    [Show full text]
  • Jointly Measurable and Progressively Measurable Stochastic Processes
    Jointly measurable and progressively measurable stochastic processes Jordan Bell [email protected] Department of Mathematics, University of Toronto June 18, 2015 1 Jointly measurable stochastic processes d Let E = R with Borel E , let I = R≥0, which is a topological space with the subspace topology inherited from R, and let (Ω; F ;P ) be a probability space. For a stochastic process (Xt)t2I with state space E, we say that X is jointly measurable if the map (t; !) 7! Xt(!) is measurable BI ⊗ F ! E . For ! 2 Ω, the path t 7! Xt(!) is called left-continuous if for each t 2 I, Xs(!) ! Xt(!); s " t: We prove that if the paths of a stochastic process are left-continuous then the stochastic process is jointly measurable.1 Theorem 1. If X is a stochastic process with state space E and all the paths of X are left-continuous, then X is jointly measurable. Proof. For n ≥ 1, t 2 I, and ! 2 Ω, let 1 n X Xt (!) = 1[k2−n;(k+1)2−n)(t)Xk2−n (!): k=0 n Each X is measurable BI ⊗ F ! E : for B 2 E , 1 n [ −n −n f(t; !) 2 I × Ω: Xt (!) 2 Bg = [k2 ; (k + 1)2 ) × fXk2−n 2 Bg: k=0 −n −n Let t 2 I. For each n there is a unique kn for which t 2 [kn2 ; (kn +1)2 ), and n −n −n thus Xt (!) = Xkn2 (!). Furthermore, kn2 " t, and because s 7! Xs(!) is n −n left-continuous, Xkn2 (!) ! Xt(!).
    [Show full text]
  • Conditional Generalizations of Strong Laws Which Conclude the Partial Sums Converge Almost Surely1
    The Annals ofProbability 1982. Vol. 10, No.3. 828-830 CONDITIONAL GENERALIZATIONS OF STRONG LAWS WHICH CONCLUDE THE PARTIAL SUMS CONVERGE ALMOST SURELY1 By T. P. HILL Georgia Institute of Technology Suppose that for every independent sequence of random variables satis­ fying some hypothesis condition H, it follows that the partial sums converge almost surely. Then it is shown that for every arbitrarily-dependent sequence of random variables, the partial sums converge almost surely on the event where the conditional distributions (given the past) satisfy precisely the same condition H. Thus many strong laws for independent sequences may be immediately generalized into conditional results for arbitrarily-dependent sequences. 1. Introduction. Ifevery sequence ofindependent random variables having property A has property B almost surely, does every arbitrarily-dependent sequence of random variables have property B almost surely on the set where the conditional distributions have property A? Not in general, but comparisons of the conditional Borel-Cantelli Lemmas, the condi­ tional three-series theorem, and many martingale results with their independent counter­ parts suggest that the answer is affirmative in fairly general situations. The purpose ofthis note is to prove Theorem 1, which states, in part, that if "property B" is "the partial sums converge," then the answer is always affIrmative, regardless of "property A." Thus many strong laws for independent sequences (even laws yet undiscovered) may be immediately generalized into conditional results for arbitrarily-dependent sequences. 2. Main Theorem. In this note, IY = (Y1 , Y2 , ••• ) is a sequence ofrandom variables on a probability triple (n, d, !?I'), Sn = Y 1 + Y 2 + ..
    [Show full text]
  • Probability and Statistics Lecture Notes
    Probability and Statistics Lecture Notes Antonio Jiménez-Martínez Chapter 1 Probability spaces In this chapter we introduce the theoretical structures that will allow us to assign proba- bilities in a wide range of probability problems. 1.1. Examples of random phenomena Science attempts to formulate general laws on the basis of observation and experiment. The simplest and most used scheme of such laws is: if a set of conditions B is satisfied =) event A occurs. Examples of such laws are the law of gravity, the law of conservation of mass, and many other instances in chemistry, physics, biology... If event A occurs inevitably whenever the set of conditions B is satisfied, we say that A is certain or sure (under the set of conditions B). If A can never occur whenever B is satisfied, we say that A is impossible (under the set of conditions B). If A may or may not occur whenever B is satisfied, then A is said to be a random phenomenon. Random phenomena is our subject matter. Unlike certain and impossible events, the presence of randomness implies that the set of conditions B do not reflect all the necessary and sufficient conditions for the event A to occur. It might seem them impossible to make any worthwhile statements about random phenomena. However, experience has shown that many random phenomena exhibit a statistical regularity that makes them subject to study. For such random phenomena it is possible to estimate the chance of occurrence of the random event. This estimate can be obtained from laws, called probabilistic or stochastic, with the form: if a set of conditions B is satisfied event A occurs m times =) repeatedly n times out of the n repetitions.
    [Show full text]
  • 1 Dependent and Independent Events 2 Complementary Events 3 Mutually Exclusive Events 4 Probability of Intersection of Events
    1 Dependent and Independent Events Let A and B be events. We say that A is independent of B if P (AjB) = P (A). That is, the marginal probability of A is the same as the conditional probability of A, given B. This means that the probability of A occurring is not affected by B occurring. It turns out that, in this case, B is independent of A as well. So, we just say that A and B are independent. We say that A depends on B if P (AjB) 6= P (A). That is, the marginal probability of A is not the same as the conditional probability of A, given B. This means that the probability of A occurring is affected by B occurring. It turns out that, in this case, B depends on A as well. So, we just say that A and B are dependent. Consider these events from the card draw: A = drawing a king, B = drawing a spade, C = drawing a face card. Events A and B are independent. If you know that you have drawn a spade, this does not change the likelihood that you have actually drawn a king. Formally, the marginal probability of drawing a king is P (A) = 4=52. The conditional probability that your card is a king, given that it a spade, is P (AjB) = 1=13, which is the same as 4=52. Events A and C are dependent. If you know that you have drawn a face card, it is much more likely that you have actually drawn a king than it would be ordinarily.
    [Show full text]
  • The Probability Set-Up.Pdf
    CHAPTER 2 The probability set-up 2.1. Basic theory of probability We will have a sample space, denoted by S (sometimes Ω) that consists of all possible outcomes. For example, if we roll two dice, the sample space would be all possible pairs made up of the numbers one through six. An event is a subset of S. Another example is to toss a coin 2 times, and let S = fHH;HT;TH;TT g; or to let S be the possible orders in which 5 horses nish in a horse race; or S the possible prices of some stock at closing time today; or S = [0; 1); the age at which someone dies; or S the points in a circle, the possible places a dart can hit. We should also keep in mind that the same setting can be described using dierent sample set. For example, in two solutions in Example 1.30 we used two dierent sample sets. 2.1.1. Sets. We start by describing elementary operations on sets. By a set we mean a collection of distinct objects called elements of the set, and we consider a set as an object in its own right. Set operations Suppose S is a set. We say that A ⊂ S, that is, A is a subset of S if every element in A is contained in S; A [ B is the union of sets A ⊂ S and B ⊂ S and denotes the points of S that are in A or B or both; A \ B is the intersection of sets A ⊂ S and B ⊂ S and is the set of points that are in both A and B; ; denotes the empty set; Ac is the complement of A, that is, the points in S that are not in A.
    [Show full text]
  • Random Variable = a Real-Valued Function of an Outcome X = F(Outcome)
    Random Variables (Chapter 2) Random variable = A real-valued function of an outcome X = f(outcome) Domain of X: Sample space of the experiment. Ex: Consider an experiment consisting of 3 Bernoulli trials. Bernoulli trial = Only two possible outcomes – success (S) or failure (F). • “IF” statement: if … then “S” else “F” • Examine each component. S = “acceptable”, F = “defective”. • Transmit binary digits through a communication channel. S = “digit received correctly”, F = “digit received incorrectly”. Suppose the trials are independent and each trial has a probability ½ of success. X = # successes observed in the experiment. Possible values: Outcome Value of X (SSS) (SSF) (SFS) … … (FFF) Random variable: • Assigns a real number to each outcome in S. • Denoted by X, Y, Z, etc., and its values by x, y, z, etc. • Its value depends on chance. • Its value becomes available once the experiment is completed and the outcome is known. • Probabilities of its values are determined by the probabilities of the outcomes in the sample space. Probability distribution of X = A table, formula or a graph that summarizes how the total probability of one is distributed over all the possible values of X. In the Bernoulli trials example, what is the distribution of X? 1 Two types of random variables: Discrete rv = Takes finite or countable number of values • Number of jobs in a queue • Number of errors • Number of successes, etc. Continuous rv = Takes all values in an interval – i.e., it has uncountable number of values. • Execution time • Waiting time • Miles per gallon • Distance traveled, etc. Discrete random variables X = A discrete rv.
    [Show full text]
  • Lecture 2: Modeling Random Experiments
    Department of Mathematics Ma 3/103 KC Border Introduction to Probability and Statistics Winter 2021 Lecture 2: Modeling Random Experiments Relevant textbook passages: Pitman [5]: Sections 1.3–1.4., pp. 26–46. Larsen–Marx [4]: Sections 2.2–2.5, pp. 18–66. 2.1 Axioms for probability measures Recall from last time that a random experiment is an experiment that may be conducted under seemingly identical conditions, yet give different results. Coin tossing is everyone’s go-to example of a random experiment. The way we model random experiments is through the use of probabilities. We start with the sample space Ω, the set of possible outcomes of the experiment, and consider events, which are subsets E of the sample space. (We let F denote the collection of events.) 2.1.1 Definition A probability measure P or probability distribution attaches to each event E a number between 0 and 1 (inclusive) so as to obey the following axioms of probability: Normalization: P (?) = 0; and P (Ω) = 1. Nonnegativity: For each event E, we have P (E) > 0. Additivity: If EF = ?, then P (∪ F ) = P (E) + P (F ). Note that while the domain of P is technically F, the set of events, that is P : F → [0, 1], we may also refer to P as a probability (measure) on Ω, the set of realizations. 2.1.2 Remark To reduce the visual clutter created by layers of delimiters in our notation, we may omit some of them simply write something like P (f(ω) = 1) orP {ω ∈ Ω: f(ω) = 1} instead of P {ω ∈ Ω: f(ω) = 1} and we may write P (ω) instead of P {ω} .
    [Show full text]