Almost all

Last updated

In mathematics, the term "almost all" means "all but a negligible quantity". More precisely, if is a set, "almost all elements of " means "all elements of but those in a negligible subset of ". The meaning of "negligible" depends on the mathematical context; for instance, it can mean finite, countable, or null.

Contents

In contrast, "almost no" means "a negligible quantity"; that is, "almost no elements of " means "a negligible quantity of elements of ".

Meanings in different areas of mathematics

Prevalent meaning

Throughout mathematics, "almost all" is sometimes used to mean "all (elements of an infinite set) except for finitely many". [1] [2] This use occurs in philosophy as well. [3] Similarly, "almost all" can mean "all (elements of an uncountable set) except for countably many". [sec 1]

Examples:

Meaning in measure theory

The Cantor function as a function that has zero derivative almost everywhere CantorEscalier.svg
The Cantor function as a function that has zero derivative almost everywhere

When speaking about the reals, sometimes "almost all" can mean "all reals except for a null set". [6] [7] [sec 2] Similarly, if S is some set of reals, "almost all numbers in S" can mean "all numbers in S except for those in a null set". [8] The real line can be thought of as a one-dimensional Euclidean space. In the more general case of an n-dimensional space (where n is a positive integer), these definitions can be generalised to "all points except for those in a null set" [sec 3] or "all points in S except for those in a null set" (this time, S is a set of points in the space). [9] Even more generally, "almost all" is sometimes used in the sense of "almost everywhere" in measure theory, [10] [11] [sec 4] or in the closely related sense of "almost surely" in probability theory. [11] [sec 5]

Examples:

Meaning in number theory

In number theory, "almost all positive integers" can mean "the positive integers in a set whose natural density is 1". That is, if A is a set of positive integers, and if the proportion of positive integers in A below n (out of all positive integers below n) tends to 1 as n tends to infinity, then almost all positive integers are in A. [16] [17] [sec 7]

More generally, let S be an infinite set of positive integers, such as the set of even positive numbers or the set of primes, if A is a subset of S, and if the proportion of elements of S below n that are in A (out of all elements of S below n) tends to 1 as n tends to infinity, then it can be said that almost all elements of S are in A.

Examples:

Meaning in graph theory

In graph theory, if A is a set of (finite labelled) graphs, it can be said to contain almost all graphs, if the proportion of graphs with n vertices that are in A tends to 1 as n tends to infinity. [19] However, it is sometimes easier to work with probabilities, [20] so the definition is reformulated as follows. The proportion of graphs with n vertices that are in A equals the probability that a random graph with n vertices (chosen with the uniform distribution) is in A, and choosing a graph in this way has the same outcome as generating a graph by flipping a coin for each pair of vertices to decide whether to connect them. [21] Therefore, equivalently to the preceding definition, the set A contains almost all graphs if the probability that a coin-flip–generated graph with n vertices is in A tends to 1 as n tends to infinity. [20] [22] Sometimes, the latter definition is modified so that the graph is chosen randomly in some other way, where not all graphs with n vertices have the same probability, [21] and those modified definitions are not always equivalent to the main one.

The use of the term "almost all" in graph theory is not standard; the term "asymptotically almost surely" is more commonly used for this concept. [20]

Example:

Meaning in topology

In topology [24] and especially dynamical systems theory [25] [26] [27] (including applications in economics), [28] "almost all" of a topological space's points can mean "all of the space's points except for those in a meagre set". Some use a more limited definition, where a subset contains almost all of the space's points only if it contains some open dense set. [26] [29] [30]

Example:

Meaning in algebra

In abstract algebra and mathematical logic, if U is an ultrafilter on a set X, "almost all elements of X" sometimes means "the elements of some element of U". [31] [32] [33] [34] For any partition of X into two disjoint sets, one of them will necessarily contain almost all elements of X. It is possible to think of the elements of a filter on X as containing almost all elements of X, even if it isn't an ultrafilter. [34]

Proofs

  1. The prime number theorem shows that the number of primes less than or equal to n is asymptotically equal to n/ln(n). Therefore, the proportion of primes is roughly ln(n)/n, which tends to 0 as n tends to infinity, so the proportion of composite numbers less than or equal to n tends to 1 as n tends to infinity. [17]

See also

Related Research Articles

<span class="mw-page-title-main">Integer</span> Number in {..., –2, –1, 0, 1, 2, ...}

An integer is the number zero (0), a positive natural number or a negative integer. The negative numbers are the additive inverses of the corresponding positive numbers. The set of all integers is often denoted by the boldface Z or blackboard bold .

<span class="mw-page-title-main">Logarithm</span> Mathematical function, inverse of an exponential function

In mathematics, the logarithm is the inverse function to exponentiation. That means that the logarithm of a number x to the base b is the exponent to which b must be raised to produce x. For example, since 1000 = 103, the logarithm base 10 of 1000 is 3, or log10 (1000) = 3. The logarithm of x to base b is denoted as logb (x), or without parentheses, logbx, or even without the explicit base, log x, when no confusion is possible, or when the base does not matter such as in big O notation.

<span class="mw-page-title-main">Power set</span> Mathematical set containing all subsets of a given set

In mathematics, the power set (or powerset) of a set S is the set of all subsets of S, including the empty set and S itself. In axiomatic set theory (as developed, for example, in the ZFC axioms), the existence of the power set of any set is postulated by the axiom of power set. The powerset of S is variously denoted as P(S), 𝒫(S), P(S), , , or 2S. Any subset of P(S) is called a family of sets over S.

<span class="mw-page-title-main">Set (mathematics)</span> Collection of mathematical objects

In mathematics, a set is a collection of different things; these things are called elements or members of the set and are typically mathematical objects of any kind: numbers, symbols, points in space, lines, other geometrical shapes, variables, or even other sets. A set may have a finite number of elements or be an infinite set. There is a unique set with no elements, called the empty set; a set with a single element is a singleton.

0 (zero) is a number representing an empty quantity. Adding 0 to any number leaves that number unchanged. In mathematical terminology, 0 is the additive identity of the integers, rational numbers, real numbers, and complex numbers, as well as other algebraic structures. Multiplying any number by 0 has the result 0, and consequently, division by zero has no meaning in arithmetic.

<span class="mw-page-title-main">Cyclic group</span> Mathematical group that can be generated as the set of powers of a single element

In abstract algebra, a cyclic group or monogenous group is a group, denoted Cn, that is generated by a single element. That is, it is a set of invertible elements with a single associative binary operation, and it contains an element g such that every other element of the group may be obtained by repeatedly applying the group operation to g or its inverse. Each element can be written as an integer power of g in multiplicative notation, or as an integer multiple of g in additive notation. This element g is called a generator of the group.

In mathematics, the probabilistic method is a nonconstructive method, primarily used in combinatorics and pioneered by Paul Erdős, for proving the existence of a prescribed kind of mathematical object. It works by showing that if one randomly chooses objects from a specified class, the probability that the result is of the prescribed kind is strictly greater than zero. Although the proof uses probability, the final conclusion is determined for certain, without any possible error.

In combinatorics, Ramsey's theorem, in one of its graph-theoretic forms, states that one will find monochromatic cliques in any edge labelling (with colours) of a sufficiently large complete graph. To demonstrate the theorem for two colours (say, blue and red), let r and s be any two positive integers. Ramsey's theorem states that there exists a least positive integer R(r, s) for which every blue-red edge colouring of the complete graph on R(r, s) vertices contains a blue clique on r vertices or a red clique on s vertices. (Here R(r, s) signifies an integer that depends on both r and s.)

In graph theory and computer science, an adjacency matrix is a square matrix used to represent a finite graph. The elements of the matrix indicate whether pairs of vertices are adjacent or not in the graph.

<span class="mw-page-title-main">Graph (discrete mathematics)</span> Vertices connected in pairs by edges

In discrete mathematics, and more specifically in graph theory, a graph is a structure amounting to a set of objects in which some pairs of the objects are in some sense "related". The objects are represented by abstractions called vertices and each of the related pairs of vertices is called an edge. Typically, a graph is depicted in diagrammatic form as a set of dots or circles for the vertices, joined by lines or curves for the edges.

In probability theory, an event is said to happen almost surely if it happens with probability 1. In other words, the set of outcomes on which the event does not occur has probability 0, even though the set might not be empty. The concept is analogous to the concept of "almost everywhere" in measure theory. In probability experiments on a finite sample space with a non-zero probability for each outcome, there is no difference between almost surely and surely ; however, this distinction becomes important when the sample space is an infinite set, because an infinite set can have non-empty subsets of probability 0.

In mathematics, an incidence matrix is a logical matrix that shows the relationship between two classes of objects, usually called an incidence relation. If the first class is X and the second is Y, the matrix has one row for each element of X and one column for each element of Y. The entry in row x and column y is 1 if x and y are related and 0 if they are not. There are variations; see below.

<span class="mw-page-title-main">Perfect graph</span> Graph with tight clique-coloring relation

In graph theory, a perfect graph is a graph in which the chromatic number equals the size of the maximum clique, both in the graph itself and in every induced subgraph. In all graphs, the chromatic number is greater than or equal to the size of the maximum clique, but they can be far apart. A graph is perfect when these numbers are equal, and remain equal after the deletion of arbitrary subsets of vertices.

<span class="mw-page-title-main">Chromatic polynomial</span> Function in algebraic graph theory

The chromatic polynomial is a graph polynomial studied in algebraic graph theory, a branch of mathematics. It counts the number of graph colorings as a function of the number of colors and was originally defined by George David Birkhoff to study the four color problem. It was generalised to the Tutte polynomial by Hassler Whitney and W. T. Tutte, linking it to the Potts model of statistical physics.

In the mathematical discipline of graph theory, a graph labeling is the assignment of labels, traditionally represented by integers, to edges and/or vertices of a graph.

In number theory, natural density, also referred to as asymptotic density or arithmetic density, is one method to measure how "large" a subset of the set of natural numbers is. It relies chiefly on the probability of encountering members of the desired subset when combing through the interval [1, n] as n grows large.

5 (five) is a number, numeral and digit. It is the natural number, and cardinal number, following 4 and preceding 6, and is a prime number. It has garnered attention throughout history in part because distal extremities in humans typically contain five digits.

<span class="mw-page-title-main">Rado graph</span> Infinite graph containing all countable graphs

In the mathematical field of graph theory, the Rado graph, Erdős–Rényi graph, or random graph is a countably infinite graph that can be constructed by choosing independently at random for each pair of its vertices whether to connect the vertices by an edge. The names of this graph honor Richard Rado, Paul Erdős, and Alfréd Rényi, mathematicians who studied it in the early 1960s; it appears even earlier in the work of Wilhelm Ackermann. The Rado graph can also be constructed non-randomly, by symmetrizing the membership relation of the hereditarily finite sets, by applying the BIT predicate to the binary representations of the natural numbers, or as an infinite Paley graph that has edges connecting pairs of prime numbers congruent to 1 mod 4 that are quadratic residues modulo each other.

In the mathematical fields of graph theory and finite model theory, the logic of graphs deals with formal specifications of graph properties using sentences of mathematical logic. There are several variations in the types of logical operation that can be used in these sentences. The first-order logic of graphs concerns sentences in which the variables and predicates concern individual vertices and edges of a graph, while monadic second-order graph logic allows quantification over sets of vertices or edges. Logics based on least fixed point operators allow more general predicates over tuples of vertices, but these predicates can only be constructed through fixed-point operators, restricting their power.

References

Primary sources

  1. Cahen, Paul-Jean; Chabert, Jean-Luc (3 December 1996). Integer-Valued Polynomials. Mathematical Surveys and Monographs. Vol. 48. American Mathematical Society. p. xix. ISBN   978-0-8218-0388-2. ISSN   0076-5376.
  2. Cahen, Paul-Jean; Chabert, Jean-Luc (7 December 2010) [First published 2000]. "Chapter 4: What's New About Integer-Valued Polynomials on a Subset?". In Hazewinkel, Michiel (ed.). Non-Noetherian Commutative Ring Theory. Mathematics and Its Applications. Vol. 520. Springer. p. 85. doi:10.1007/978-1-4757-3180-4. ISBN   978-1-4419-4835-9.
  3. Gärdenfors, Peter (22 August 2005). The Dynamics of Thought. Synthese Library. Vol. 300. Springer. pp. 190–191. ISBN   978-1-4020-3398-8.
  4. 1 2 Courant, Richard; Robbins, Herbert; Stewart, Ian (18 July 1996). What is Mathematics? An Elementary Approach to Ideas and Methods (2nd ed.). Oxford University Press. ISBN   978-0-19-510519-3.
  5. Movshovitz-hadar, Nitsa; Shriki, Atara (2018-10-08). Logic In Wonderland: An Introduction To Logic Through Reading Alice's Adventures In Wonderland - Teacher's Guidebook. World Scientific. p. 38. ISBN   978-981-320-864-3. This can also be expressed in the statement: 'Almost all prime numbers are odd.'
  6. 1 2 Korevaar, Jacob (1 January 1968). Mathematical Methods: Linear Algebra / Normed Spaces / Distributions / Integration. Vol. 1. New York: Academic Press. pp. 359–360. ISBN   978-1-4832-2813-6.
  7. Natanson, Isidor P. (June 1961). Theory of Functions of a Real Variable. Vol. 1. Translated by Boron, Leo F. (revised ed.). New York: Frederick Ungar Publishing. p. 90. ISBN   978-0-8044-7020-9.
  8. Sohrab, Houshang H. (15 November 2014). Basic Real Analysis (2 ed.). Birkhäuser. p. 307. doi:10.1007/978-1-4939-1841-6. ISBN   978-1-4939-1841-6.
  9. Helmberg, Gilbert (December 1969). Introduction to Spectral Theory in Hilbert Space. North-Holland Series in Applied Mathematics and Mechanics. Vol. 6 (1st ed.). Amsterdam: North-Holland Publishing Company. p. 320. ISBN   978-0-7204-2356-3.
  10. Vestrup, Eric M. (18 September 2003). The Theory of Measures and Integration. Wiley Series in Probability and Statistics. United States: Wiley-Interscience. p. 182. ISBN   978-0-471-24977-1.
  11. 1 2 Billingsley, Patrick (1 May 1995). Probability and Measure (PDF). Wiley Series in Probability and Statistics (3rd ed.). United States: Wiley-Interscience. p. 60. ISBN   978-0-471-00710-4. Archived from the original (PDF) on 23 May 2018.
  12. Niven, Ivan (1 June 1956). Irrational Numbers. Carus Mathematical Monographs. Vol. 11. Rahway: Mathematical Association of America. pp. 2–5. ISBN   978-0-88385-011-4.
  13. Baker, Alan (1984). A concise introduction to the theory of numbers. Cambridge University Press. p.  53. ISBN   978-0-521-24383-4.
  14. Granville, Andrew; Rudnick, Zeev (7 January 2007). Equidistribution in Number Theory, An Introduction. Nato Science Series II. Vol. 237. Springer. p. 11. ISBN   978-1-4020-5404-4.
  15. Burk, Frank (3 November 1997). Lebesgue Measure and Integration: An Introduction. A Wiley-Interscience Series of Texts, Monographs, and Tracts. United States: Wiley-Interscience. p. 260. ISBN   978-0-471-17978-8.
  16. Hardy, G. H. (1940). Ramanujan: Twelve Lectures on Subjects Suggested by His Life and Work. Cambridge University Press. p. 50.
  17. 1 2 Hardy, G. H.; Wright, E. M. (December 1960). An Introduction to the Theory of Numbers (4th ed.). Oxford University Press. pp. 8–9. ISBN   978-0-19-853310-8.
  18. Prachar, Karl (1957). Primzahlverteilung. Grundlehren der mathematischen Wissenschaften (in German). Vol. 91. Berlin: Springer. p. 164. Cited in Grosswald, Emil (1 January 1984). Topics from the Theory of Numbers (2nd ed.). Boston: Birkhäuser. p. 30. ISBN   978-0-8176-3044-7.
  19. 1 2 Babai, László (25 December 1995). "Automorphism Groups, Isomorphism, Reconstruction". In Graham, Ronald; Grötschel, Martin; Lovász, László (eds.). Handbook of Combinatorics. Vol. 2. Netherlands: North-Holland Publishing Company. p. 1462. ISBN   978-0-444-82351-9.
  20. 1 2 3 Spencer, Joel (9 August 2001). The Strange Logic of Random Graphs. Algorithms and Combinatorics. Vol. 22. Springer. pp. 3–4. ISBN   978-3-540-41654-8.
  21. 1 2 Bollobás, Béla (8 October 2001). Random Graphs. Cambridge Studies in Advanced Mathematics. Vol. 73 (2nd ed.). Cambridge University Press. pp. 34–36. ISBN   978-0-521-79722-1.
  22. Grädel, Eric; Kolaitis, Phokion G.; Libkin, Leonid; Marx, Maarten; Spencer, Joel; Vardi, Moshe Y.; Venema, Yde; Weinstein, Scott (11 June 2007). Finite Model Theory and Its Applications. Texts in Theoretical Computer Science (An EATCS Series). Springer. p. 298. ISBN   978-3-540-00428-8.
  23. Buckley, Fred; Harary, Frank (21 January 1990). Distance in Graphs. Addison-Wesley. p. 109. ISBN   978-0-201-09591-3.
  24. Oxtoby, John C. (1980). Measure and Category. Graduate Texts in Mathematics. Vol. 2 (2nd ed.). United States: Springer. pp. 59, 68. ISBN   978-0-387-90508-2. While Oxtoby does not explicitly define the term there, Babai has borrowed it from Measure and Category in his chapter "Automorphism Groups, Isomorphism, Reconstruction" of Graham, Grötschel and Lovász's Handbook of Combinatorics (vol. 2), and Broer and Takens note in their book Dynamical Systems and Chaos that Measure and Category compares this meaning of "almost all" to the measure theoretic one in the real line (though Oxtoby's book discusses meagre sets in general topological spaces as well).
  25. Baratchart, Laurent (1987). "Recent and New Results in Rational L2 Approximation". In Curtain, Ruth F. (ed.). Modelling, Robustness and Sensitivity Reduction in Control Systems. NATO ASI Series F. Vol. 34. Springer. p. 123. doi:10.1007/978-3-642-87516-8. ISBN   978-3-642-87516-8.
  26. 1 2 Broer, Henk; Takens, Floris (28 October 2010). Dynamical Systems and Chaos. Applied Mathematical Sciences. Vol. 172. Springer. p. 245. doi:10.1007/978-1-4419-6870-8. ISBN   978-1-4419-6870-8.
  27. Sharkovsky, A. N.; Kolyada, S. F.; Sivak, A. G.; Fedorenko, V. V. (30 April 1997). Dynamics of One-Dimensional Maps. Mathematics and Its Applications. Vol. 407. Springer. p. 33. doi:10.1007/978-94-015-8897-3. ISBN   978-94-015-8897-3.
  28. Yuan, George Xian-Zhi (9 February 1999). KKM Theory and Applications in Nonlinear Analysis. Pure and Applied Mathematics; A Series of Monographs and Textbooks. Marcel Dekker. p. 21. ISBN   978-0-8247-0031-7.
  29. Albertini, Francesca; Sontag, Eduardo D. (1 September 1991). "Transitivity and Forward Accessibility of Discrete-Time Nonlinear Systems". In Bonnard, Bernard; Bride, Bernard; Gauthier, Jean-Paul; Kupka, Ivan (eds.). Analysis of Controlled Dynamical Systems. Progress in Systems and Control Theory. Vol. 8. Birkhäuser. p. 29. doi:10.1007/978-1-4612-3214-8. ISBN   978-1-4612-3214-8.
  30. De la Fuente, Angel (28 January 2000). Mathematical Models and Methods for Economists. Cambridge University Press. p. 217. ISBN   978-0-521-58529-3.
  31. Komjáth, Péter; Totik, Vilmos (2 May 2006). Problems and Theorems in Classical Set Theory. Problem Books in Mathematics. United States: Springer. p. 75. ISBN   978-0387-30293-5.
  32. Salzmann, Helmut; Grundhöfer, Theo; Hähl, Hermann; Löwen, Rainer (24 September 2007). The Classical Fields: Structural Features of the Real and Rational Numbers. Encyclopedia of Mathematics and Its Applications. Vol. 112. Cambridge University Press. p.  155. ISBN   978-0-521-86516-6.
  33. Schoutens, Hans (2 August 2010). The Use of Ultraproducts in Commutative Algebra. Lecture Notes in Mathematics. Vol. 1999. Springer. p. 8. doi:10.1007/978-3-642-13368-8. ISBN   978-3-642-13367-1.
  34. 1 2 Rautenberg, Wolfgang (17 December 2009). A Concise to Mathematical Logic. Universitext (3rd ed.). Springer. pp. 210–212. doi:10.1007/978-1-4419-1221-3. ISBN   978-1-4419-1221-3.

Secondary sources

  1. Schwartzman, Steven (1 May 1994). The Words of Mathematics: An Etymological Dictionary of Mathematical Terms Used in English . Spectrum Series. Mathematical Association of America. p.  22. ISBN   978-0-88385-511-9.
  2. Clapham, Christopher; Nicholson, James (7 June 2009). The Concise Oxford Dictionary of mathematics. Oxford Paperback References (4th ed.). Oxford University Press. p. 38. ISBN   978-0-19-923594-0.
  3. James, Robert C. (31 July 1992). Mathematics Dictionary (5th ed.). Chapman & Hall. p. 269. ISBN   978-0-412-99031-1.
  4. Bityutskov, Vadim I. (30 November 1987). "Almost-everywhere". In Hazewinkel, Michiel (ed.). Encyclopaedia of Mathematics . Vol. 1. Kluwer Academic Publishers. p. 153. doi:10.1007/978-94-015-1239-8. ISBN   978-94-015-1239-8.
  5. Itô, Kiyosi, ed. (4 June 1993). Encyclopedic Dictionary of Mathematics . Vol. 2 (2nd ed.). Kingsport: MIT Press. p. 1267. ISBN   978-0-262-09026-1.
  6. "Almost All Real Numbers are Transcendental - ProofWiki". proofwiki.org. Retrieved 2019-11-11.
  7. 1 2 Weisstein, Eric W. "Almost All". MathWorld . See also Weisstein, Eric W. (25 November 1988). CRC Concise Encyclopedia of Mathematics (1st ed.). CRC Press. p. 41. ISBN   978-0-8493-9640-3.
  8. Itô, Kiyosi, ed. (4 June 1993). Encyclopedic Dictionary of Mathematics. Vol. 1 (2nd ed.). Kingsport: MIT Press. p. 67. ISBN   978-0-262-09026-1.