Jump to content

Equivalence relation: Difference between revisions

From Wikipedia, the free encyclopedia
Content deleted Content added
→‎See also: add Congruence relation
→‎From order to equivalence via symmetry: reorganise, add partial equivalence relations
Line 38: Line 38:
Just as [[order relation]]s are grounded in [[ordered set]]s, sets closed under pairwise [[supremum]] and [[infimum]], equivalence relations are grounded in [[partition of a set|partitioned sets]], sets closed under [[bijection]]s preserving partition structure. Since all such bijections map an equivalence class onto itself, such bijections are also known as [[permutation]]s.
Just as [[order relation]]s are grounded in [[ordered set]]s, sets closed under pairwise [[supremum]] and [[infimum]], equivalence relations are grounded in [[partition of a set|partitioned sets]], sets closed under [[bijection]]s preserving partition structure. Since all such bijections map an equivalence class onto itself, such bijections are also known as [[permutation]]s.


[[order relation|Order]] and equivalence relations are both [[transitive]], but only equivalence relations are symmetric as well. If [[symmetry]]:
[[order relation|Order]] and equivalence relations are both [[transitive]], but only equivalence relations are symmetric as well. If [[symmetry]] is weakened to [[Antisymmetric relation|antisymmetry]], the result is a [[partial order]].
*Is weakened to [[Antisymmetric relation|antisymmetry]], the result is a [[partial order]];
*Is discarded entirely, the result is a [[preorder]];
*And reflexivity are both discarded, the result is a [[strict partial order]].
Hence equivalence relations can be seen as the culmination of a hierarchy of order relations.


A [[dependency relation]] is reflexive and symmetric, but not transitive.
A [[partial equivalence relation]] is transitive and symmetric, but not reflexive.<br>
A [[dependency relation]] is reflexive and symmetric, but not transitive.<br>
A [[preorder]] is reflexive and transitive, but neither symmetric nor antisymmetric.<br>
A [[strict partial order]] is transitive alone.

Hence equivalence relations can be seen as the culmination of a hierarchy of order relations.


== Equivalence class, quotient set, partition ==
== Equivalence class, quotient set, partition ==

Revision as of 10:51, 3 November 2007

In mathematics, an equivalence relation is a binary relation between two elements of a set which groups them together as being "equivalent" in some way. That a is equivalent to b is denoted as "a ~ b" or "ab". An equivalence relation is reflexive, symmetric, and transitive. In other words, for all elements a, b, and c of the set X, the following must hold for "~" to be an equivalence relation on X:

An equivalence relation partitions a set into several disjoint subsets, called equivalence classes. All the elements in a given equivalence class are equivalent among themselves, and no element is equivalent with any element from a different class.

A set X together with an equivalence relation on X is called a setoid.

Examples of equivalence relations

A ubiquitous equivalence relation is the equality ("=") relation between elements of any set. Other examples include:

Examples of relations that are not equivalences

  • The relation "≥" between real numbers is reflexive and transitive, but not symmetric. For example, 7 ≥ 5 does not imply that 5 ≥ 7. It is, however, a partial order.
  • The relation "has a common factor greater than 1 with" between natural numbers greater than 1, is reflexive and symmetric, but not transitive. (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 empty relation R on a non-empty set X (i.e. aRb is never true) is vacuously symmetric and transitive, but not reflexive. (If X is also empty then R is reflexive.)
  • The relation "is approximately equal to" between real numbers or other things, 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.
  • The relation "is a sibling of" on the set of all human beings is not an equivalence relation. Although siblinghood is symmetric (if A is a sibling of B, then B is a sibling of A) it is neither reflexive (no one is a sibling of himself), nor transitive (since if A is a sibling of B, then B is a sibling of A, but A is not a sibling of A). Instead of being transitive, siblinghood is "almost transitive", meaning that if A ~ B, and B ~ C, and AC, then A ~ C.
  • Symmetry and transitivity do not imply reflexivity, unless every A is related to some B.

From order to equivalence via symmetry

Just as order relations are grounded in ordered sets, sets closed under pairwise supremum and infimum, equivalence relations are grounded in partitioned sets, sets closed under bijections preserving partition structure. Since all such bijections map an equivalence class onto itself, such bijections are also known as permutations.

Order and equivalence relations are both transitive, but only equivalence relations are symmetric as well. If symmetry is weakened to antisymmetry, the result is a partial order.

A partial equivalence relation is transitive and symmetric, but not reflexive.
A dependency relation is reflexive and symmetric, but not transitive.
A preorder is reflexive and transitive, but neither symmetric nor antisymmetric.
A strict partial order is transitive alone.

Hence equivalence relations can be seen as the culmination of a hierarchy of order relations.

Equivalence class, quotient set, partition

Let X be a nonempty set with typical members a and b. Some definitions:

  • The set of all a and b for which a ~ b holds make up an equivalence class of X by ~. Let [a ] =: {xX : x ~ a} denote the equivalence class to which a belongs. Then all members of X equivalent to each other are also members of the same equivalence class: ∀a, bX (a ~ b ↔ [a ] = [b ]).
  • The set of all possible equivalence classes of X by ~, denoted X/~ =: {[x] : xX}, is the quotient set of X by ~. If X is a topological space, there is a natural way of transforming X/~ into a topological space; see quotient space for the details.
  • The projection of ~ is the function π : XX/~, defined by π(x) = [x ], mapping members of X into their respective equivalence classes by ~.
  • The equivalence kernel of a function f is the equivalence relation, denoted Ef, such that xEfyf(x) = f(y). The equivalence kernel of an injection is the identity relation.
  • A partition of X is a set P of subsets of X, such that every member of X is a member of a unique member of P. Each member of P is a cell of the partition. Moreover, the members of P are pairwise disjoint and their union is X.

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 possible equivalence relations on X equals the number of distinct partitions of X, which is the n th Bell number Bn:

Theorem ("Fundamental Theorem of Equivalence Relations": Wallace 1998: 31, Th. 8; Dummit and Foote 2004: 3, Prop. 2):

  • An equivalence relation ~ 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 member 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 member of X belongs to a unique equivalence class of X by ~. Thus there is a natural bijection from the set of all possible equivalence relations on X and the set of all partitions of X.

Theorem on projections (Birkhoff and Mac Lane 1999: 35, Th. 19): Let the function f: XB be such that a ~ bf(a) = f(b). Then there is a unique function g : X/~B, such that f = g π. If f is a surjection and a ~ bf(a) = f(b), then g is a bijection.

Generating equivalence relations

  • An equivalence relation ~ on X is the equivalence kernel of its surjective projection π : XX/~. (Birkhoff and Mac Lane 1999: 33 Th. 18). Conversely, any surjection between sets determines a partition on its domain, the set of preimages of singletons in the codomain. Thus an equivalence relation over X, a partition of X, and a projection whose domain is X, are three equivalent ways of specifying the same thing.
  • The intersection of any two equivalence relations over X (viewed as a subset of X × X) is also an equivalence relation. This yields a convenient way of generating an equivalence relation: given any binary relation R on X, the equivalence relation generated by R is the smallest equivalence relation containing R. Concretely, R generates the equivalence relation if:
a ~ b if and only if there exist elements x1, x2, ..., xn in X such that x1 = a, xn = b, and such that (xi , xi + 1) or (xi + 1, xi ) is in R for every i = 1, ..., n − 1.
Note that the generated equivalence relation generated in this manner can be trivial. For instance, the equivalence relation ~ generated by:
  • The binary relation has exactly one equivalence class, X itself, because x ~ y for all x and y;
  • An antisymmetric relation has equivalence classes that are the singletons of X.
  • Let r be any sort of relation on X. Then rr−1 is a symmetric relation. The transitive closure s of rr−1 assures that s is transitive and reflexive. Moreover, s is the "smallest" equivalence relation containing r, and r/s partially orders X/s.
  • Equivalence relations can construct new spaces by "gluing things together." Let X be the unit Cartesian square [0,1] × [0,1], and let ~ be the equivalence relation on X defined by ∀a, b ∈ [0,1] ((a, 0) ~ (a, 1) ∧ (0, b) ~ (1, b)). Then the quotient space X/~ can be naturally identified 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.
  • Let G be a group and H a subgroup of G. Define an equivalence relation ~ on G such that a ~ b ↔ (ab−1H ). 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.

Algebraic characterizations

Transformation groups

It is very well known that lattice theory captures the mathematical structure of order relations. It is much less known that transformation groups (some authors prefer permutation groups) and their orbits capture the mathematical structure of equivalence relations. See also Lucas (1973: §31).

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: ∀xAgG (g(x) ∈ [x ]). Then the following three connected theorems hold (Van Fraassen 1989: §10.3):

  • ~ partitions A into equivalence classes. (This is the Fundamental Theorem of Equivalence Relations, mentioned above);
  • Given a partition of A, G is a transformation group under composition, whose orbits are the cells of the partition‡;
  • Given a transformation group G over A, there exists an equivalence relation ~ over A, whose equivalence classes are the orbits of G. (Wallace 1998: 202, Th. 6; Dummit and Foote 2004: 114, Prop. 2).

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 lattices characterize order relations. The arguments of the lattice theory operations meet and join are elements of some universe A. Meanwhile, the arguments of the transformation group operations composition and inverse are elements of a set of bijections, AA.

Proof (adapted from Van Fraassen 1989: 246). Let function composition interpret group multiplication, and function inverse interpret group inverse. Then G is a group under composition, meaning that ∀xAgG ([g(x)] = [x]), because G satisfies the following four conditions:

  • G is closed under composition. The composition of any two members of G exists, because the domain and codomain of any member of G is A. Moreover, the composition of bijections is bijective (Wallace 1998: 22, Th. 6);
  • Existence of identity. The identity function, I(x)=x, is an obvious member of G;
  • Existence of inverse. Every 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 (Wallace 1998: 24, Th. 7).

Let f and g be any two members 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 A.

Category theory

The composition of morphisms central to category theory, denoted by concatenation, generalizes the composition of functions central to transformation groups. The two defining axioms of category theory assert that the composition of morphisms associates, and that the left and right identity morphisms exist. If all morphisms in a category were to have "inverses," the category would resemble a transformation group, whose close relation to equivalence relations has just been explained. A morphism f can be said to have an inverse when f is an automorphism, i.e., the domain and codomain of f are identical, and there exists a morphism g such that fg = gf = identity morphism. Hence the category-theoretic concept nearest to an equivalence relation is a category whose morphisms are all automorphisms.

Equivalence relations and mathematical logic

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

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.

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.

Equivalence relations are not all that difficult or interesting, but can be a source of easy examples or counterexamples.

Euclid anticipated equivalence

Euclid's The Elements includes the following "Common Notion 1":

Things which equal the same thing also equal one another.

Nowadays, the property described by Common Notion 1 is called Euclidean (replacing "equal" by "are in relation with"). The following theorem connects Euclidian relations and equivalence relations:

Theorem. If a relation is Euclidian and reflexive, it is also symmetric and transitive.

Proof:

  • (aRcbRc) → aRb [a/c] = (aRabRa) → aRb [reflexive; erase T∧] = bRaaRb. Hence R is symmetric.
  • (aRcbRc) → aRb [symmetry] = (aRccRb) → aRb. Hence R is transitive.

Hence an equivalence relation is a relation that is Euclidean and reflexive. The Elements mentions neither symmetry nor reflexivity, and Euclid probably would have deemed the reflexivity of equality too obvious to warrant explicit mention. If this (and considering "equality" as an abstract relation) is granted, a charitable reading of Common Notion 1 would credit Euclid with being the first to conceive of equivalence relations and their importance in deductive systems.

See also

References

  • Garrett Birkhoff and Saunders Mac Lane, 1999 (1967). Algebra, 3rd ed. Chelsea.
  • Castellani, E., 2003, "Symmetry and equivalence" in Katherine Brading and E. Castellani (eds.), Symmetries in Physics: Philosophical Reflections. Cambridge University Press: 422-433.
  • Robert Dilworth and Crawley, Peter, 1973. Algebraic Theory of Lattices. Prentice Hall. Chpt. 12 discusses how equivalence relations arise in lattice theory.
  • Dummit, D. S., and Foote, R. M., 2004. Abstract Algebra, 3rd ed. John Wiley & Sons.
  • John Randolph Lucas, 1973. A Treatise on Time and Space. London: Methuen. Section 31.
  • Bas van Fraassen, 1989. Laws and Symmetry. Oxford Univ. Press.
  • Wallace, D. A. R., 1998. Groups, Rings and Fields. Springer-Verlag.

External links