Equivalence relation

Source: Wikipedia, the free encyclopedia.
Transitive binary relations
Strict total order
Green tickY Green tickY Green tickY Green tickY
Symmetric Antisymmetric Connected Well-founded Has joins Has meets Reflexive Irreflexive Asymmetric
Definitions, for all and
Green tickY indicates that the column's property is always true the row's term (at the very left), while indicates that the property is not guaranteed in general (it might, or might not, hold). For example, that every equivalence relation is symmetric, but not necessarily antisymmetric, is indicated by Green tickY in the "Symmetric" column and in the "Antisymmetric" column, respectively.

All definitions tacitly require the homogeneous relation be transitive: for all if and then
A term's definition may require additional properties that are not listed in this table.

The 52 equivalence relations on a 5-element set depicted as logical matrices (colored fields, including those in light gray, stand for ones; white fields for zeros). The row and column indices of nonwhite cells are the related elements, while the different colors, other than light gray, indicate the equivalence classes (each light gray cell is its own equivalence class).

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).

Each equivalence relation provides a partition of the underlying set into disjoint equivalence classes. Two elements of the given set are equivalent to each other if and only if they belong to the same equivalence class.

Notation

Various notations are used in the literature to denote that two elements and of a set are equivalent with respect to an equivalence relation the most common are "" and "ab", which are used when is implicit, and variations of "", "aR b", or "" to specify explicitly. Non-equivalence may be written "ab" or "".

Definition

A binary relation on a set is said to be an equivalence relation, if and only if it is reflexive, symmetric and transitive. That is, for all and in

  • (reflexivity).
  • if and only if (symmetry).
  • If and then (transitivity).

together with the relation is called a setoid. The equivalence class of under denoted is defined as [1][2]

Alternative definition using relational algebra

In relational algebra, if and are relations, then the composite relation is defined so that if and only if there is a such that and .[note 1] This definition is a generalisation of the definition of functional composition. The defining properties of an equivalence relation on a set can then be reformulated as follows:

  • . (reflexivity). (Here, denotes the identity function on .)
  • (symmetry).
  • (transitivity).[3]

Examples

Simple example

On the set , the relation is an equivalence relation. The following sets are equivalence classes of this relation:

The set of all equivalence classes for is This set is a partition of the set with respect to .

Equivalence relations

The following relations are all equivalence relations:

  • "Is equal to" on the set of numbers. For example, is equal to [2]
  • "Has the same birthday as" on the set of all people.
  • "Is
    triangles
    .
  • "Is
    triangles
    .
  • Given a natural number , "is congruent to, modulo " on the
    integers.[2]
  • Given a function , "has the same image under as" on the elements of 's domain . For example, and have the same image under , viz. .
  • "Has the same absolute value as" on the set of real numbers
  • "Has the same cosine as" on the set of all angles.

Relations that are not equivalences

  • The relation "≥" between real numbers is reflexive and transitive, but not symmetric. For example, 7 ≥ 5 but not 5 ≥ 7.
  • The relation "has a
    natural numbers
    greater than 1, is reflexive and symmetric, but not transitive. For example, the natural numbers 2 and 6 have a common factor greater than 1, and 6 and 3 have a common factor greater than 1, but 2 and 3 do not have a common factor greater than 1.
  • The
    vacuously
    symmetric and transitive; however, it is not reflexive (unless X itself is empty).
  • The relation "is approximately equal to" between real numbers, even if more precisely defined, is not an equivalence relation, because although reflexive and symmetric, it is not transitive, since multiple small changes can accumulate to become a big change. However, if the approximation is defined asymptotically, for example by saying that two functions f and g are approximately equal near some point if the limit of f − g is 0 at that point, then this defines an equivalence relation.

Connections to other relations

Well-definedness under an equivalence relation

If is an equivalence relation on and is a property of elements of such that whenever is true if is true, then the property is said to be

well-defined
or a class invariant under the relation

A frequent particular case occurs when is a function from to another set if implies then is said to be a morphism for a class invariant under or simply invariant under This occurs, e.g. in the character theory of finite groups. The latter case with the function can be expressed by a commutative triangle. See also invariant. Some authors use "compatible with " or just "respects " instead of "invariant under ".

More generally, a function may map equivalent arguments (under an equivalence relation ) to equivalent values (under an equivalence relation ). Such a function is known as a morphism from to

Related important definitions

Let , and be an equivalence relation. Some key definitions and terminology follow:

Equivalence class

A subset of such that holds for all and in , and never for in and outside , is called an equivalence class of by . Let denote the equivalence class to which belongs. All elements of equivalent to each other are also elements of the same equivalence class.

Quotient set

The set of all equivalence classes of by denoted is the quotient set of by If is a topological space, there is a natural way of transforming into a topological space; see Quotient space for the details.

Projection

The projection of is the function defined by which maps elements of into their respective equivalence classes by

Theorem on projections:[4] Let the function be such that if then Then there is a unique function such that If is a
surjection
and then is a bijection.

Equivalence kernel

The equivalence kernel of a function is the equivalence relation ~ defined by The equivalence kernel of an

identity relation
.

Partition

A partition of X is a set P of nonempty subsets of X, such that every element of X is an element of a single element of P. Each element of P is a cell of the partition. Moreover, the elements of P are

pairwise disjoint and their union
is X.

Counting partitions

Let X be a finite set with n elements. Since every equivalence relation over X corresponds to a partition of X, and vice versa, the number of equivalence relations on X equals the number of distinct partitions of X, which is the nth Bell number Bn:

(
Dobinski's formula
).

Fundamental theorem of equivalence relations

A key result links equivalence relations and partitions:[5][6][7]

  • An equivalence relation ~ on a set X partitions X.
  • Conversely, corresponding to any partition of X, there exists an equivalence relation ~ on X.

In both cases, the cells of the partition of X are the equivalence classes of X by ~. Since each element of X belongs to a unique cell of any partition of X, and since each cell of the partition is identical to an equivalence class of X by ~, each element of X belongs to a unique equivalence class of X by ~. Thus there is a natural bijection between the set of all equivalence relations on X and the set of all partitions of X.

Comparing equivalence relations

If and are two equivalence relations on the same set , and implies for all then is said to be a coarser relation than , and is a finer relation than . Equivalently,

  • is finer than if every equivalence class of is a subset of an equivalence class of , and thus every equivalence class of is a union of equivalence classes of .
  • is finer than if the partition created by is a refinement of the partition created by .

The equality equivalence relation is the finest equivalence relation on any set, while the universal relation, which relates all pairs of elements, is the coarsest.

The relation " is finer than " on the collection of all equivalence relations on a fixed set is itself a partial order relation, which makes the collection a geometric lattice.[8]

Generating equivalence relations

if there exists a natural number and elements such that , , and or , for
The equivalence relation generated in this manner can be trivial. For instance, the equivalence relation generated by any total order on X has exactly one equivalence class, X itself.
  • Equivalence relations can construct new spaces by "gluing things together." Let X be the unit
    Cartesian square
    and let ~ be the equivalence relation on X defined by for all and for all Then the quotient space can be naturally identified (homeomorphism) with a torus: take a square piece of paper, bend and glue together the upper and lower edge to form a cylinder, then bend the resulting cylinder so as to glue together its two open ends, resulting in a torus.

Algebraic structure

Much of mathematics is grounded in the study of equivalences, and

order relations. Lattice theory captures the mathematical structure of order relations. Even though equivalence relations are as ubiquitous in mathematics as order relations, the algebraic structure of equivalences is not as well known as that of orders. The former structure draws primarily on group theory and, to a lesser extent, on the theory of lattices, categories, and groupoids
.

Group theory

Just as

orbit
shed light on the mathematical structure of equivalence relations.

Let '~' denote an equivalence relation over some nonempty set A, called the universe or underlying set. Let G denote the set of bijective functions over A that preserve the partition structure of A, meaning that for all and Then the following three connected theorems hold:[10]

In sum, given an equivalence relation ~ over A, there exists a

transformation group
G over A whose orbits are the equivalence classes of A under ~.

This transformation group characterisation of equivalence relations differs fundamentally from the way

bijections
, AA.

Moving to groups in general, let H be a subgroup of some group G. Let ~ be an equivalence relation on G, such that The equivalence classes of ~—also called the orbits of the

action of H on G—are the right cosets
of H in G. Interchanging a and b yields the left cosets.

Related thinking can be found in Rosen (2008: chpt. 10).

Categories and groupoids

Let G be a set and let "~" denote an equivalence relation over G. Then we can form a groupoid representing this equivalence relation as follows. The objects are the elements of G, and for any two elements x and y of G, there exists a unique morphism from x to y if and only if

The advantages of regarding an equivalence relation as a special case of a groupoid include:

  • Whereas the notion of "free equivalence relation" does not exist, that of a free groupoid on a directed graph does. Thus it is meaningful to speak of a "presentation of an equivalence relation," i.e., a presentation of the corresponding groupoid;
  • Bundles of groups,
    group actions
    , sets, and equivalence relations can be regarded as special cases of the notion of groupoid, a point of view that suggests a number of analogies;
  • In many contexts "quotienting," and hence the appropriate equivalence relations often called congruences, are important. This leads to the notion of an internal groupoid in a category.[17]

Lattices

The equivalence relations on any set X, when ordered by

injective. Less formally, the equivalence relation ker on X, takes each function f : XX to its kernel
ker f. Likewise, ker(ker) is an equivalence relation on X^X.

Equivalence relations and mathematical logic

Equivalence relations are a ready source of examples or counterexamples. For example, an equivalence relation with exactly two infinite equivalence classes is an easy example of a theory which is ω-

categorical, but not categorical for any larger cardinal number
.

An implication of model theory is that the properties defining a relation can be proved independent of each other (and hence necessary parts of the definition) if and only if, for each property, examples can be found of relations not satisfying the given property while satisfying all the other properties. Hence the three defining properties of equivalence relations can be proved mutually independent by the following three examples:

  • Reflexive and transitive: The relation ≤ on N. Or any preorder;
  • Symmetric and transitive: The relation R on N, defined as aRbab ≠ 0. Or any partial equivalence relation;
  • Reflexive and symmetric: The relation R on Z, defined as aRb ↔ "ab is divisible by at least one of 2 or 3." Or any dependency relation.

Properties definable in first-order logic that an equivalence relation may or may not possess include:

  • The number of equivalence classes is finite or infinite;
  • The number of equivalence classes equals the (finite) natural number n;
  • All equivalence classes have infinite cardinality;
  • The number of elements in each equivalence class is the natural number n.

See also

Notes

  1. ^ Sometimes the composition is instead written as , or as ; in both cases, is the first relation that is applied. See the article on Composition of relations for more information.
  1. ^ If: Given let hold using totality, then by symmetry, hence by transitivity. — Only if: Given choose then by reflexivity.
  1. ^ Weisstein, Eric W. "Equivalence Class". mathworld.wolfram.com. Retrieved 2020-08-30.
  2. ^ a b c "7.3: Equivalence Classes". Mathematics LibreTexts. 2017-09-20. Retrieved 2020-08-30.
  3. .
  4. ^ Garrett Birkhoff and Saunders Mac Lane, 1999 (1967). Algebra, 3rd ed. p. 35, Th. 19. Chelsea.
  5. ^ Wallace, D. A. R., 1998. Groups, Rings and Fields. p. 31, Th. 8. Springer-Verlag.
  6. ^ Dummit, D. S., and Foote, R. M., 2004. Abstract Algebra, 3rd ed. p. 3, Prop. 2. John Wiley & Sons.
  7. Karel Hrbacek & Thomas Jech (1999) Introduction to Set Theory, 3rd edition, pages 29–32, Marcel Dekker
  8. . Sect. IV.9, Theorem 12, page 95
  9. ^ Garrett Birkhoff and Saunders Mac Lane, 1999 (1967). Algebra, 3rd ed. p. 33, Th. 18. Chelsea.
  10. ^ Rosen (2008), pp. 243–45. Less clear is §10.3 of Bas van Fraassen, 1989. Laws and Symmetry. Oxford Univ. Press.
  11. ^ Bas van Fraassen, 1989. Laws and Symmetry. Oxford Univ. Press: 246.
  12. ^ Wallace, D. A. R., 1998. Groups, Rings and Fields. Springer-Verlag: 22, Th. 6.
  13. ^ Wallace, D. A. R., 1998. Groups, Rings and Fields. Springer-Verlag: 24, Th. 7.
  14. ^ Proof.[11] Let function composition interpret group multiplication, and function inverse interpret group inverse. Then G is a group under composition, meaning that and because G satisfies the following four conditions:
    • G is closed under composition. The composition of any two elements of G exists, because the
      bijective;[12]
    • Existence of identity function. The identity function, I(x) = x, is an obvious element of G;
    • Existence of
      bijective function
      g has an inverse g−1, such that gg−1 = I;
    • Composition
      associates. f(gh) = (fg)h. This holds for all functions over all domains.[13]
    Let f and g be any two elements of G. By virtue of the definition of G, [g(f(x))] = [f(x)] and [f(x)] = [x], so that [g(f(x))] = [x]. Hence G is also a transformation group (and an automorphism group) because function composition preserves the partitioning of
  15. ^ Wallace, D. A. R., 1998. Groups, Rings and Fields. Springer-Verlag: 202, Th. 6.
  16. ^ Dummit, D. S., and Foote, R. M., 2004. Abstract Algebra, 3rd ed. John Wiley & Sons: 114, Prop. 2.

References

External links