Disjoint sets

Last updated
Two disjoint sets Disjunkte Mengen.svg
Two disjoint sets

In set theory in mathematics and formal logic, two sets are said to be disjoint sets if they have no element in common. Equivalently, two disjoint sets are sets whose intersection is the empty set. [1] For example, {1, 2, 3} and {4, 5, 6} are disjoint sets, while {1, 2, 3} and {3, 4, 5} are not disjoint. A collection of two or more sets is called disjoint if any two distinct sets of the collection are disjoint.

Contents

Generalizations

A disjoint collection of sets Disjoint sets.svg
A disjoint collection of sets

This definition of disjoint sets can be extended to families of sets and to indexed families of sets. By definition, a collection of sets is called a family of sets (such as the power set, for example). In some sources this is a set of sets, while other sources allow it to be a multiset of sets, with some sets repeated. An indexed family of sets is by definition a set-valued function (that is, it is a function that assigns a set to every element in its domain) whose domain is called its index set (and elements of its domain are called indices).

There are two subtly different definitions for when a family of sets is called pairwise disjoint. According to one such definition, the family is disjoint if each two sets in the family are either identical or disjoint. This definition would allow pairwise disjoint families of sets to have repeated copies of the same set. According to an alternative definition, each two sets in the family must be disjoint; repeated copies are not allowed. The same two definitions can be applied to an indexed family of sets: according to the first definition, every two distinct indices in the family must name sets that are disjoint or identical, while according to the second, every two distinct indices must name disjoint sets. [2] For example, the family of sets { {0, 1, 2}, {3, 4, 5}, {6, 7, 8}, ... } is disjoint according to both definitions, as is the family { {..., −2, 0, 2, 4, ...}, {..., −3, −1, 1, 3, 5} } of the two parity classes of integers. However, the family with 10 members has five repetitions each of two disjoint sets, so it is pairwise disjoint under the first definition but not under the second.

Two sets are said to be almost disjoint sets if their intersection is small in some sense. For instance, two infinite sets whose intersection is a finite set may be said to be almost disjoint. [3]

In topology, there are various notions of separated sets with more strict conditions than disjointness. For instance, two sets may be considered to be separated when they have disjoint closures or disjoint neighborhoods. Similarly, in a metric space, positively separated sets are sets separated by a nonzero distance. [4]

Intersections

Disjointness of two sets, or of a family of sets, may be expressed in terms of intersections of pairs of them.

Two sets A and B are disjoint if and only if their intersection is the empty set. [1] It follows from this definition that every set is disjoint from the empty set, and that the empty set is the only set that is disjoint from itself. [5]

If a collection contains at least two sets, the condition that the collection is disjoint implies that the intersection of the whole collection is empty. However, a collection of sets may have an empty intersection without being disjoint. Additionally, while a collection of less than two sets is trivially disjoint, as there are no pairs to compare, the intersection of a collection of one set is equal to that set, which may be non-empty. [2] For instance, the three sets { {1, 2}, {2, 3}, {1, 3} } have an empty intersection but are not disjoint. In fact, there are no two disjoint sets in this collection. Also the empty family of sets is pairwise disjoint. [6]

A Helly family is a system of sets within which the only subfamilies with empty intersections are the ones that are pairwise disjoint. For instance, the closed intervals of the real numbers form a Helly family: if a family of closed intervals has an empty intersection and is minimal (i.e. no subfamily of the family has an empty intersection), it must be pairwise disjoint. [7]

Disjoint unions and partitions

A partition of a set X is any collection of mutually disjoint non-empty sets whose union is X. [8] Every partition can equivalently be described by an equivalence relation, a binary relation that describes whether two elements belong to the same set in the partition. [8] Disjoint-set data structures [9] and partition refinement [10] are two techniques in computer science for efficiently maintaining partitions of a set subject to, respectively, union operations that merge two sets or refinement operations that split one set into two.

A disjoint union may mean one of two things. Most simply, it may mean the union of sets that are disjoint. [11] But if two or more sets are not already disjoint, their disjoint union may be formed by modifying the sets to make them disjoint before forming the union of the modified sets. [12] For instance two sets may be made disjoint by replacing each element by an ordered pair of the element and a binary value indicating whether it belongs to the first or second set. [13] For families of more than two sets, one may similarly replace each element by an ordered pair of the element and the index of the set that contains it. [14]

See also

Related Research Articles

<span class="mw-page-title-main">Axiom of choice</span> Axiom of set theory

In mathematics, the axiom of choice, abbreviated AC or AoC, is an axiom of set theory equivalent to the statement that a Cartesian product of a collection of non-empty sets is non-empty. Informally put, the axiom of choice says that given any collection of sets, each containing at least one element, it is possible to construct a new set by choosing one element from each set, even if the collection is infinite. Formally, it states that for every indexed family of nonempty sets, there exists an indexed set such that for every . The axiom of choice was formulated in 1904 by Ernst Zermelo in order to formalize his proof of the well-ordering theorem.

Naive set theory is any of several theories of sets used in the discussion of the foundations of mathematics. Unlike axiomatic set theories, which are defined using formal logic, naive set theory is defined informally, in natural language. It describes the aspects of mathematical sets familiar in discrete mathematics, and suffices for the everyday use of set theory concepts in contemporary mathematics.

In mathematics, a set is countable if either it is finite or it can be made in one to one correspondence with the set of natural numbers. Equivalently, a set is countable if there exists an injective function from it into the natural numbers; this means that each element in the set may be associated to a unique natural number, or that the elements of the set can be counted one at a time, although the counting may never finish due to an infinite number of elements.

<span class="mw-page-title-main">Connected space</span> Topological space that is connected

In topology and related branches of mathematics, a connected space is a topological space that cannot be represented as the union of two or more disjoint non-empty open subsets. Connectedness is one of the principal topological properties that are used to distinguish topological spaces.

<span class="mw-page-title-main">Equivalence relation</span> Mathematical concept for comparing objects

In mathematics, an equivalence relation is a binary relation that is reflexive, symmetric and transitive. The equipollence relation between line segments in geometry is a common example of an equivalence relation. A simpler example is equality. Any number is equal to itself (reflexive). If , then (symmetric). If and , then (transitive).

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

A set is the mathematical model for a collection of different things; a set contains elements or members, which can be mathematical objects of any kind: numbers, symbols, points in space, lines, other geometrical shapes, variables, or even other sets. The set with no element is the empty set; a set with a single element is a singleton. A set may have a finite number of elements or be an infinite set.

This is a glossary of some terms used in the branch of mathematics known as topology. Although there is no absolute distinction between different areas of topology, the focus here is on general topology. The following definitions are also fundamental to algebraic topology, differential topology and geometric topology.

<span class="mw-page-title-main">Union (set theory)</span> Set of elements in any of some sets

In set theory, the union of a collection of sets is the set of all elements in the collection. It is one of the fundamental operations through which sets can be combined and related to each other. A nullary union refers to a union of zero sets and it is by definition equal to the empty set.

<span class="mw-page-title-main">Disjoint union</span> In mathematics, operation on sets

In mathematics, the disjoint union of the sets A and B is the set formed from the elements of A and B labelled (indexed) with the name of the set from which they come. So, an element belonging to both A and B appears twice in the disjoint union, with two different labels.

<span class="mw-page-title-main">Symmetric difference</span> Elements in exactly one of two sets

In mathematics, the symmetric difference of two sets, also known as the disjunctive union and set sum, is the set of elements which are in either of the sets, but not in their intersection. For example, the symmetric difference of the sets and is .

<span class="mw-page-title-main">Partition of a set</span> Mathematical ways to group elements of a set

In mathematics, a partition of a set is a grouping of its elements into non-empty subsets, in such a way that every element is included in exactly one subset.

In mathematics, the Lebesgue covering dimension or topological dimension of a topological space is one of several different ways of defining the dimension of the space in a topologically invariant way.

In mathematics, two sets are almost disjoint if their intersection is small in some sense; different definitions of "small" will result in different definitions of "almost disjoint".

In the mathematical field of measure theory, an outer measure or exterior measure is a function defined on all subsets of a given set with values in the extended real numbers satisfying some additional technical conditions. The theory of outer measures was first introduced by Constantin Carathéodory to provide an abstract basis for the theory of measurable sets and countably additive measures. Carathéodory's work on outer measures found many applications in measure-theoretic set theory, and was used in an essential way by Hausdorff to define a dimension-like metric invariant now called Hausdorff dimension. Outer measures are commonly used in the field of geometric measure theory.

In set theory and related branches of mathematics, a family can mean any of

In mathematics, particularly in combinatorics, given a family of sets, here called a collection C, a transversal (also called a cross-section) is a set containing exactly one element from each member of the collection. When the sets of the collection are mutually disjoint, each element of the transversal corresponds to exactly one member of C (the set it is a member of). If the original sets are not disjoint, there are two possibilities for the definition of a transversal:

In mathematics, a Grothendieck universe is a set U with the following properties:

  1. If x is an element of U and if y is an element of x, then y is also an element of U. (U is a transitive set.)
  2. If x and y are both elements of U, then is an element of U.
  3. If x is an element of U, then P(x), the power set of x, is also an element of U.
  4. If is a family of elements of U, and if I is an element of U, then the union is an element of U.
<span class="mw-page-title-main">Sunflower (mathematics)</span> Collection of sets in which every two sets have the same intersection

In the mathematical fields of set theory and extremal combinatorics, a sunflower or -system is a collection of sets in which all possible distinct pairs of sets share the same intersection. This common intersection is called the kernel of the sunflower.

<span class="mw-page-title-main">Intersection (set theory)</span> Set of elements common to all of some sets

In set theory, the intersection of two sets and denoted by is the set containing all elements of that also belong to or equivalently, all elements of that also belong to

<span class="mw-page-title-main">Ordinal number</span> Generalization of "n-th" to infinite cases

In set theory, an ordinal number, or ordinal, is a generalization of ordinal numerals aimed to extend enumeration to infinite sets.

References

  1. 1 2 Halmos, P. R. (1960), Naive Set Theory, Undergraduate Texts in Mathematics, Springer, p. 15, ISBN   9780387900926 .
  2. 1 2 Smith, Douglas; Eggen, Maurice; St. Andre, Richard (2010), A Transition to Advanced Mathematics, Cengage Learning, p. 95, ISBN   978-0-495-56202-3 .
  3. Halbeisen, Lorenz J. (2011), Combinatorial Set Theory: With a Gentle Introduction to Forcing, Springer monographs in mathematics, Springer, p. 184, ISBN   9781447121732 .
  4. Copson, Edward Thomas (1988), Metric Spaces, Cambridge Tracts in Mathematics, vol. 57, Cambridge University Press, p. 62, ISBN   9780521357326 .
  5. Oberste-Vorth, Ralph W.; Mouzakitis, Aristides; Lawrence, Bonita A. (2012), Bridge to Abstract Mathematics, MAA textbooks, Mathematical Association of America, p. 59, ISBN   9780883857793 .
  6. See answers to the question ″Is the empty family of sets pairwise disjoint?″
  7. Bollobás, Béla (1986), Combinatorics: Set Systems, Hypergraphs, Families of Vectors, and Combinatorial Probability, Cambridge University Press, p. 82, ISBN   9780521337038 .
  8. 1 2 Halmos (1960), p. 28.
  9. Cormen, Thomas H.; Leiserson, Charles E.; Rivest, Ronald L.; Stein, Clifford (2001), "Chapter 21: Data structures for Disjoint Sets", Introduction to Algorithms (Second ed.), MIT Press, pp. 498–524, ISBN   0-262-03293-7 .
  10. Paige, Robert; Tarjan, Robert E. (1987), "Three partition refinement algorithms", SIAM Journal on Computing, 16 (6): 973–989, doi:10.1137/0216062, MR   0917035, S2CID   33265037 .
  11. Ferland, Kevin (2008), Discrete Mathematics: An Introduction to Proofs and Combinatorics, Cengage Learning, p. 45, ISBN   9780618415380 .
  12. Arbib, Michael A.; Kfoury, A. J.; Moll, Robert N. (1981), A Basis for Theoretical Computer Science, The AKM series in Theoretical Computer Science: Texts and monographs in computer science, Springer-Verlag, p. 9, ISBN   9783540905738 .
  13. Monin, Jean François; Hinchey, Michael Gerard (2003), Understanding Formal Methods, Springer, p. 21, ISBN   9781852332471 .
  14. Lee, John M. (2010), Introduction to Topological Manifolds, Graduate Texts in Mathematics, vol. 202 (2nd ed.), Springer, p. 64, ISBN   9781441979407 .