Natural transformation


In category theory, a branch of mathematics, a natural transformation provides a way of transforming one functor into another while respecting the internal structure of the categories involved. Hence, a natural transformation can be considered to be a "morphism of functors". Indeed, this intuition can be formalized to define so-called functor categories. Natural transformations are, after categories and functors, one of the most fundamental notions of category theory and consequently appear in the majority of its applications.

Definition

If and are functors between the categories and, then a natural transformation from to is a family of morphisms that satisfies two requirements.
  1. The natural transformation must associate, to every object in, a morphism between objects of. The morphism is called the component of at.
  2. Components must be such that for every morphism in we have:
The last equation can conveniently be expressed by the commutative diagram
If both and are contravariant, the vertical arrows in this diagram are reversed. If is a natural transformation from to, we also write or. This is also expressed by saying the family of morphisms is natural in.
If, for every object in, the morphism is an isomorphism in, then is said to be a . Two functors and are called naturally isomorphic or simply isomorphic if there exists a natural isomorphism from to.
An infranatural transformation from to is simply a family of morphisms, for all in. Thus a natural transformation is an infranatural transformation for which for every morphism. The naturalizer of, nat, is the largest subcategory of containing all the objects of on which restricts to a natural transformation.

Examples

Opposite group

Statements such as
abound in modern mathematics. We will now give the precise meaning of this statement as well as its proof. Consider the category
of all groups with group homomorphisms as morphisms. If is a group, we define
its opposite group as follows: is the same set as, and the operation is defined
by. All multiplications in are thus "turned around". Forming the opposite group becomes
a functor from to if we define for any group homomorphism. Note that
is indeed a group homomorphism from to :
The content of the above statement is:
To prove this, we need to provide isomorphisms for every group, such that the above diagram commutes.
Set.
The formulas and
show that is a group homomorphism with inverse. To prove the naturality, we start with a group homomorphism
and show, i.e.
for all in. This is true since
and every group homomorphism has the property.

Abelianization

Given a group, we can define its abelianization Commutator subgroup#Definition|. Let
denote the projection map onto the cosets of. This homomorphism is "natural in
", i.e., it defines a natural transformation, which we now check. Let be a group. For any homomorphism, we have that
is contained in the kernel of, because any homomorphism into an abelian group kills the commutator subgroup. Then
factors through as for the unique homomorphism
. This makes a functor and
a natural transformation, but not a natural isomorphism, from the identity functor to.

Hurewicz homomorphism

Functors and natural transformations abound in algebraic topology, with the Hurewicz homomorphisms serving as examples. For any pointed topological space and positive integer there exists a group homomorphism
from the -th homotopy group of to the -th homology group of. Both and are functors from the category Top* of pointed topological spaces to the category Grp of groups, and is a natural transformation from to.

Determinant

Given commutative rings and with a ring homomorphism, the respective groups of invertible matrices and inherit a homomorphism which we denote by, obtained by applying
to each matrix entry. Similarly, restricts to a group homomorphism, where denotes the group of units of. In fact, and are functors from the category of commutative rings to.
The determinant on the group, denoted by, is a group homomorphism
which is natural in : because the determinant is defined by the same formula for every ring, holds. This makes the determinant a natural transformation from to.

Double dual of a vector space

If is a field, then for every vector space over we have a "natural" injective linear map from the vector space into its double dual. These maps are "natural" in the following sense: the double dual operation is a functor, and the maps are the components of a natural transformation from the identity functor to the double dual functor.

Finite calculus

For every abelian group, the set of functions from the integers to the underlying set of
forms an abelian group under pointwise addition.
Given an morphism, the map given by left composing with the elements of the former is itself a homomorphism of abelian groups; in this way we
obtain a functor. The finite difference operator taking each function
to is a map from to itself, and the collection of such maps gives a natural transformation.

Tensor-hom adjunction

Consider the category of abelian groups and group homomorphisms. For all abelian groups, and we have a group isomorphism
These isomorphisms are "natural" in the sense that they define a natural transformation between the two involved functors.
This is formally the tensor-hom adjunction, and is an archetypal example of a pair of adjoint functors. Natural transformations arise frequently in conjunction with adjoint functors, and indeed, adjoint functors are defined by a certain natural isomorphism. Additionally, every pair of adjoint functors comes equipped with two natural transformations called the unit and counit.

Unnatural isomorphism

The notion of a natural transformation is categorical, and states that a particular map between functors can be done consistently over an entire category. Informally, a particular map between individual objects is referred to as a "natural isomorphism", meaning implicitly that it is actually defined on the entire category, and defines a natural transformation of functors; formalizing this intuition was a motivating factor in the development of category theory. Conversely, a particular map between particular objects may be called an unnatural isomorphism if the map cannot be extended to a natural transformation on the entire category. Given an object a functor and an isomorphism proof of unnaturality is most easily shown by giving an automorphism that does not commute with this isomorphism. More strongly, if one wishes to prove that and are not naturally isomorphic, without reference to a particular isomorphism, this requires showing that for any isomorphism, there is some with which it does not commute; in some cases a single automorphism works for all candidate isomorphisms while in other cases one must show how to construct a different for each isomorphism. The maps of the category play a crucial role – any infranatural transform is natural if the only maps are the identity map, for instance.
This is similar to concepts in group theory or module theory, where a given decomposition of an object into a direct sum is "not natural", or rather "not unique", as automorphisms exist that do not preserve the direct sum decomposition – see for example.
Some authors distinguish notationally, using for a natural isomorphism and for an unnatural isomorphism, reserving for equality.

Example: fundamental group of torus

As an example of the distinction between the functorial statement and individual objects, consider homotopy groups of a product space, specifically the fundamental group of the torus.
The homotopy groups of a product space are naturally the product of the homotopy groups of the components, with the isomorphism given by projection onto the two factors, fundamentally because maps into a product space are exactly products of maps into the components – this is a functorial statement.
However, the torus has fundamental group isomorphic to, but the splitting is not natural. Note the use of,, and :
This abstract isomorphism with a product is not natural, as some isomorphisms of do not preserve the product: the self-homeomorphism of given by acts as this matrix on , which does not preserve the decomposition as a product because it is not diagonal. However, if one is given the torus as a product – equivalently, given a decomposition of the space – then the splitting of the group follows from the general statement earlier. In categorical terms, the relevant category is "maps of product spaces, namely a pair of maps between the respective components".
Naturality is a categorical notion, and requires being very precise about exactly what data is given – the torus as a space that happens to be a product is different from the torus presented as a product.

Example: dual of a finite-dimensional vector space

Every finite-dimensional vector space is isomorphic to its dual space, but there may be many different isomorphisms between the two spaces. There is in general no natural isomorphism between a finite-dimensional vector space and its dual space. However, related categories do have a natural isomorphism, as described below.
The dual space of a finite-dimensional vector space is again a finite-dimensional vector space of the same dimension, and these are thus isomorphic, since dimension is the only invariant of finite-dimensional vector spaces over a given field. However, in the absence of additional constraints, the map from a space to its dual is not unique, and thus such an isomorphism requires a choice, and is "not natural". On the category of finite-dimensional vector spaces and linear maps, one can define an infranatural isomorphism from vector spaces to their dual by choosing an isomorphism for each space, but this will not define a natural transformation. Intuitively this is because it required a choice, rigorously because any such choice of isomorphisms will not commute with, say, the zero map; see for detailed discussion.
Starting from finite-dimensional vector spaces and the identity and dual functors, one can define a natural isomorphism, but this requires first adding additional structure, then restricting the maps from "all linear maps" to "linear maps that respect this structure". Explicitly, for each vector space, require that it comes with the data of an isomorphism to its dual,. In other words, take as objects vector spaces with a nondegenerate bilinear form. This defines an infranatural isomorphism. One then restricts the maps to only those maps that commute with the isomorphisms: or in other words, preserve the bilinear form:. The resulting category, with objects finite-dimensional vector spaces with a nondegenerate bilinear form, and maps linear transforms that respect the bilinear form, by construction has a natural isomorphism from the identity to the dual. Viewed in this light, this construction is completely general, and does not depend on any particular properties of vector spaces.
In this category, the dual of a map between vector spaces can be identified as a transpose. Often for reasons of geometric interest this is specialized to a subcategory, by requiring that the nondegenerate bilinear forms have additional properties, such as being symmetric, symmetric and positive definite, symmetric sesquilinear, skew-symmetric and totally isotropic, etc. – in all these categories a vector space is naturally identified with its dual, by the nondegenerate bilinear form.

Operations with natural transformations

If and
are natural transformations between functors, then we can compose them to get a natural transformation.
This is done componentwise:. This "vertical composition" of natural transformation is associative and has an identity, and
allows one to consider the collection of all functors itself as a category.
Natural transformations also have a "horizontal composition". If
is a natural transformation between functors and
is a natural transformation between functors, then the composition of functors allows a composition of natural transformations.
This operation is also associative with identity, and the identity coincides with that for vertical composition. The two operations are related by an identity which exchanges vertical composition with horizontal composition.
If is a natural transformation between functors, and is another functor, then we can form the natural transformation by defining
If on the other hand is a functor, the natural transformation is defined by

Functor categories

If is any category and is a small category, we can form the functor category having as objects all functors from to and as morphisms the natural transformations between those functors. This forms a category since for any functor there is an identity natural transformation and the composition of two natural transformations is again a natural transformation.
The isomorphisms in are precisely the natural isomorphisms. That is, a natural transformation is a natural isomorphism if and only if there exists a natural transformation such that and.
The functor category is especially useful if arises from a directed graph. For instance, if is the category of the directed graph, then has as objects the morphisms of, and a morphism between and in is a pair of morphisms and in such that the "square commutes", i.e..
More generally, one can build the 2-category whose
The horizontal and vertical compositions are the compositions between natural transformations described previously. A functor category is then simply a hom-category in this category.

More examples

Every limit and colimit provides an example for a simple natural transformation, as a cone amounts to a natural transformation with the diagonal functor as domain. Indeed, if limits and colimits are defined directly in terms of their universal property, they are universal morphisms in a functor category.

Yoneda lemma

If is an object of a locally small category, then the assignment defines a covariant functor. This functor is called representable. The natural transformations from a representable functor to an arbitrary functor are completely known and easy to describe; this is the content of the Yoneda lemma.

Historical notes

, one of the founders of category theory, is said to have remarked, "I didn't invent categories to study functors; I invented them to study natural transformations." Just as the study of groups is not complete without a study of homomorphisms, so the study of categories is not complete without the study of functors. The reason for Mac Lane's comment is that the study of functors is itself not complete without the study of natural transformations.
The context of Mac Lane's remark was the axiomatic theory of homology. Different ways of constructing homology could be shown to coincide: for example in the case of a simplicial complex the groups defined directly would be isomorphic to those of the singular theory. What cannot easily be expressed without the language of natural transformations is how homology groups are compatible with morphisms between objects, and how two equivalent homology theories not only have the same homology groups, but also the same morphisms between those groups.