Fact-checked by Grok 2 weeks ago

Isomorphism of categories

In , an isomorphism of categories is a F: \mathcal{C} \to \mathcal{D} between two categories \mathcal{C} and \mathcal{D} that induces a on the collections of objects and a on the hom-sets for every pair of objects, making F fully faithful and bijective on objects, with an inverse G: \mathcal{D} \to \mathcal{C} such that the compositions G \circ F = \mathrm{id}_{\mathcal{C}} and F \circ G = \mathrm{id}_{\mathcal{D}} are the respective identity functors. This strict invertibility ensures that \mathcal{C} and \mathcal{D} are indistinguishable in structure, preserving compositions, identities, and all categorical relations exactly. Unlike the weaker notion of an equivalence of categories, which requires only that F be fully faithful and essentially surjective on objects (meaning every object in \mathcal{D} is naturally isomorphic to the image of some object in \mathcal{C}), an isomorphism demands precise bijections without allowance for natural isomorphisms between functors. Equivalences capture "sameness up to isomorphism," which is often more useful in practice for comparing categories like the category of sets and the category of complete atomic Boolean algebras, but isomorphisms are rarer and more rigid, typically occurring in contrived or specially constructed settings. For instance, the categories of abelian groups (\mathbf{Ab}) and \mathbb{Z}-modules (\mathbb{Z}\mathbf{-Mod}) are isomorphic via the functor that equips each abelian group with its canonical \mathbb{Z}-module structure (with the forgetful functor as inverse, though the structures are identical). Isomorphisms of categories play a foundational role in formalizing when two categories are , facilitating proofs by transporting across them and underpinning advanced constructions like the category of categories \mathbf{[CAT](/page/Cat)}, where objects are all categories and morphisms include isomorphisms. They preserve and reflect all categorical features, such as limits, colimits, adjunctions, and monoidal structures, ensuring that any diagram or in one category corresponds exactly to its counterpart in the other. However, their strictness limits their prevalence; most "equivalent" categories in , such as finite-dimensional vector spaces over a and matrices, are not strictly isomorphic due to non-bijective choices of objects. This distinction highlights category theory's emphasis on , where isomorphisms provide the strongest form of structural .

Definition

Formal Definition

In , a consists of a collection of objects, a collection of morphisms between those objects, a operation for compatible morphisms, and morphisms for each object, all satisfying certain axioms such as associativity of and the laws. A between two categories C and D is a structure-preserving map that assigns objects in C to objects in D and morphisms in C to morphisms in D, while preserving and . A functor F: C → D is an isomorphism of categories if there exists a functor G: D → C such that the compositions F ∘ G and G ∘ F are equal to the identity functors on D and C, respectively; that is, F \circ G = \mathrm{id}_D, \quad G \circ F = \mathrm{id}_C. This equality must hold strictly, meaning the functors are identical on the nose, without the intervention of natural isomorphisms, which underscores the rigid invertibility required for such an isomorphism. The concept of isomorphism of categories originates from the foundational work in category theory developed in the 1940s by Samuel Eilenberg and Saunders Mac Lane, extending the notion of isomorphisms from algebraic structures like groups and rings to the broader framework of categories.

Role of the Inverse Functor

In the context of category theory, the inverse functor plays a pivotal role in establishing an isomorphism between categories \mathcal{C} and \mathcal{D}. Given a functor F: \mathcal{C} \to \mathcal{D} that is an isomorphism, there exists a functor G: \mathcal{D} \to \mathcal{C} such that the compositions satisfy F \circ G = \mathrm{id}_{\mathcal{D}} and G \circ F = \mathrm{id}_{\mathcal{C}}, where \mathrm{id}_{\mathcal{D}} and \mathrm{id}_{\mathcal{C}} are the respective identity functors. This construction ensures that G reverses the action of F precisely, mapping objects and morphisms back in a way that restores the original structure without deviation. The functoriality of the inverse G is essential, as it must preserve the composition of morphisms and the identities in \mathcal{D} exactly as they correspond to those in \mathcal{C}. Specifically, for any morphisms f: d_1 \to d_2 and g: d_2 \to d_3 in \mathcal{D}, G(f \circ g) = G(f) \circ G(g), and for each object d in \mathcal{D}, G(\mathrm{id}_d) = \mathrm{id}_{G(d)}. This strict preservation distinguishes the inverse from weaker notions, enforcing that G acts as a faithful reversal on the entire categorical framework. If such an inverse G exists for F, it is unique, as it can be explicitly derived by inverting the bijective actions of F on objects and morphisms. For instance, on objects, G assigns to each F(c) the unique preimage c, and on morphisms, it applies the inverse bijection induced by F. This uniqueness follows from the requirement that G satisfies the identity composition conditions, leaving no room for alternative constructions. The strict invertibility via G implies that \mathcal{C} and \mathcal{D} are essentially the same , differing only by a relabeling of objects and morphisms that preserves all internal verbatim. This in underscores the as a rigid notion of sameness, where every categorical feature—such as compositions and identities—is mirrored exactly through the .

Properties

Bijection on Objects and Morphisms

An isomorphism of categories F: \mathcal{C} \to \mathcal{D} is equipped with an functor G: \mathcal{D} \to \mathcal{C} such that F \circ G = \mathrm{Id}_{\mathcal{D}} and G \circ F = \mathrm{Id}_{\mathcal{C}}. This ensures that F induces a strict between the class of objects of \mathcal{C}, denoted \mathrm{Ob}(\mathcal{C}), and \mathrm{Ob}(\mathcal{D}). For injectivity, suppose F(c_1) = F(c_2) for objects c_1, c_2 \in \mathrm{Ob}(\mathcal{C}); then c_1 = G(F(c_1)) = G(F(c_2)) = c_2. For surjectivity, given any d \in \mathrm{Ob}(\mathcal{D}), there exists a unique c = G(d) \in \mathrm{Ob}(\mathcal{C}) such that F(c) = F(G(d)) = d. On morphisms, F induces a bijection F_*: \mathrm{Hom}_{\mathcal{C}}(a, b) \to \mathrm{Hom}_{\mathcal{D}}(F(a), F(b)) for any objects a, b \in \mathrm{Ob}(\mathcal{C}), with G providing the inverse mapping G_*: \mathrm{Hom}_{\mathcal{D}}(F(a), F(b)) \to \mathrm{Hom}_{\mathcal{C}}(a, b). To see injectivity, if F(f) = F(f') for f, f': a \to b, then f = G(F(f)) = G(F(f')) = f'. For surjectivity, given any g: F(a) \to F(b) in \mathcal{D}, define f = G(g): a \to b in \mathcal{C}; then F(f) = F(G(g)) = g. These bijections hold uniformly for all pairs of objects, as the functoriality of F and G preserves the necessary compositions. As a consequence, an F effectively relabels the objects and morphisms of \mathcal{C} to those of \mathcal{D} while preserving the category's table exactly, without any alteration to the among them. The inverse G ensures that compositions are mapped exactly, maintaining this correspondence.

Preservation of Categorical Structure

An isomorphism of categories F: \mathcal{C} \to \mathcal{D} with strict G: \mathcal{D} \to \mathcal{C} preserves exactly: if a in \mathcal{C} admits a \lim_D with projections \pi_i: \lim_D \to D_i, then the image F \circ D in \mathcal{D} admits F(\lim_D) with projections F(\pi_i): F(\lim_D) \to F(D_i), and the universal property holds strictly via the bijective correspondence on morphisms induced by F and G. Similarly, colimits are preserved: a colimit \colim_D in \mathcal{C} with inclusions \iota_i: D_i \to \colim_D maps to the colimit F(\colim_D) in \mathcal{D} with F(\iota_i), ensuring exact cocone universality. This strict preservation follows from the bijective action of F on objects and morphisms, which maps limiting or colimiting cones directly without alteration. Isomorphisms of categories preserve adjunctions strictly. The defining hom-set of an adjunction transfers exactly to the corresponding in the image category via the bijective on hom-sets induced by F, with units and counits mapped directly, preserving the triangular identities. Isomorphisms also preserve other categorical structures precisely. Monomorphisms (injective morphisms) in \mathcal{C} map to monomorphisms in \mathcal{D} under F, as the full and faithful nature ensures that the cancellation property defining monicity holds exactly in the image. Epimorphisms (surjective morphisms) are preserved analogously, with the dual cancellation property transferred bijectively. Products and coproducts follow suit: a product A \times B in \mathcal{C} with projections \pi_A, \pi_B becomes the product F(A) \times F(B) in \mathcal{D} with F(\pi_A), F(\pi_B), satisfying the universal pairing property strictly. In enriched or cartesian closed settings, exponential objects B^A (internal homs) are mapped exactly, preserving the evaluation and isomorphisms via the bijective action. This exact preservation underscores the rigidity of category isomorphisms compared to equivalences of categories, which only guarantee structures up to natural isomorphism rather than strict . Isomorphisms maintain not only the existence but also the precise "size" ( of objects and morphisms) and diagram shapes, rendering them particularly strict in concrete categories like \mathbf{Set} or \mathbf{[Ab](/page/AB)}, where set-theoretic bijections align directly. While isomorphisms apply uniformly to small and large categories, their formulation in large settings demands caution with set-theoretic foundations, as the bijection on object classes may involve proper classes rather than sets, potentially invoking axioms like the or global choice principles to ensure well-definedness.

Examples

Trivial Isomorphisms

In , the identity functor provides the simplest example of a categorical isomorphism. For any \mathcal{C}, the identity functor \mathrm{id}_\mathcal{C}: \mathcal{C} \to \mathcal{C} maps each object and to itself and serves as its own , satisfying \mathrm{id}_\mathcal{C} \circ \mathrm{id}_\mathcal{C} = \mathrm{id}_\mathcal{C}. This construction aligns with the formal definition of an isomorphism as a pair of functors that are strictly inverse to each other. Another straightforward case arises from relabeling the objects of a via a while preserving the morphisms. Given a \sigma on the objects of \mathcal{C}, one can define a F: \mathcal{C} \to \mathcal{C} by setting F(A) = \sigma(A) for objects A and F(f: A \to B) = f: \sigma(A) \to \sigma(B) for morphisms f, assuming the hom-sets remain unchanged under this renaming. The employs \sigma^{-1} similarly, yielding an that merely reindexes the objects without altering the categorical . Every is isomorphic to itself through any , which is an from \mathcal{C} to \mathcal{C}; the trivial such automorphisms are precisely the functors. These self-isomorphisms highlight the inherent symmetry in categorical presentations but offer no new structural insights. In discrete categories, where the only morphisms are identity arrows, any two such categories with the same of objects are isomorphic via any between their object sets. This extends uniquely to a on morphisms, as each must map identities to identities, and its provides the required strict inverse. Trivial isomorphisms of these forms are ubiquitous in , existing for every and often serving to standardize or normalize presentations without introducing complexity. Despite their prevalence, they are generally uninteresting for deeper , as they do not reveal non-obvious equivalences.

Non-Trivial Isomorphisms

Non-trivial isomorphisms of categories, distinct from mere relabelings of objects and s, arise infrequently and typically demand a precise matching of complex internal structures, such as compositions that align exactly beyond superficial bijections. These examples underscore the rigidity of categorical isomorphisms, where the functors must not only be bijective on objects and s but also strictly preserve the entire compositional framework without relying on equivalences that allow for natural transformations. In practice, such isomorphisms are challenging to construct outside of finite or artificially simplified settings, as discrepancies in how s compose often prevent strict equality. A notable example occurs in the category of posets, where objects are partially ordered sets and morphisms are order-preserving maps; a non-trivial isomorphism exists between the category of all posets and the category of Alexandroff T0 topological spaces (with continuous maps). This isomorphism is induced by the specialization on Alexandroff T0-spaces, which turns closure operators into partial orders, and conversely by equipping posets with the Alexandroff topology where open sets are downward-closed; the functors are inverses, yielding a strict categorical that reveals posets and Alexandroff T0-spaces as structurally identical despite their topological versus order-theoretic presentations. For finite cases, the categories of finite posets and finite T0-spaces are likewise isomorphic via the same mechanism, providing a concrete, non-relabeling match where the finite restrictions preserve the bijections on objects and morphisms exactly. A classic example from algebra is the from the of (\mathbf{Ab}) to the of \mathbb{Z}-modules (\mathbb{Z}\mathbf{-Mod}), which is an . It bijectively maps to \mathbb{Z}-modules (since every is a \mathbb{Z}-module) and group homomorphisms to module homomorphisms, with a strict inverse given by the that recovers the additive structure exactly. In the context of groups, consider the one-object category BG associated to a group G, where the single object represents the group and morphisms are the elements of G with composition given by group multiplication. This category is isomorphic to itself via any automorphism of G, but inner automorphisms—conjugations by elements of G—provide non-trivial isomorphisms for non-abelian groups, as they induce bijections on morphisms that rearrange the multiplication table non-trivially yet preserve composition strictly (e.g., for G = S_3, conjugation by a transposition yields a distinct but isomorphic labeling of permutations). Such isomorphisms highlight how group symmetries can permute the morphism set without altering the categorical structure, though they remain rare for categories beyond these monoidal presentations. Discussions on platforms like MathOverflow illustrate further instances, such as isomorphisms between categories sharing the same objects but with permuted morphisms that preserve fullness and skeletonicity, where equivalences coincide with strict isomorphisms due to the categories being skeletal (e.g., permuting morphisms in a small category with rigid composition yields an isomorphism if the permutation is an automorphism of the morphism poset). These examples, often drawn from synthetic or small categories, reinforce the scarcity of non-trivial isomorphisms, as constructing them typically necessitates identical composition tables across potentially disparate presentations, limiting their occurrence to specialized algebraic or order-theoretic domains.

Equivalence of Categories

In , an equivalence of categories offers a more flexible alternative to strict isomorphism, establishing that two categories \mathcal{C} and \mathcal{D} are essentially the same despite potential differences in their concrete presentations. A F: \mathcal{C} \to \mathcal{D} is an equivalence if there exists a functor G: \mathcal{D} \to \mathcal{C} and natural isomorphisms \eta: \mathrm{id}_\mathcal{C} \Rightarrow G \circ F (the unit) and \varepsilon: F \circ G \Rightarrow \mathrm{id}_\mathcal{D} (the counit), demonstrating that F and G act as mutual inverses up to isomorphism. This structure ensures that structures and relationships in \mathcal{C} correspond precisely to those in \mathcal{D}, but without requiring exact matching of objects or morphisms. A functor F forms an equivalence precisely when it is full, faithful, and essentially surjective on objects; full and faithful guarantee bijective hom-sets, while essential surjectivity ensures every object in \mathcal{D} is isomorphic to the image of some object in \mathcal{C}. This characterization highlights how equivalences relax the rigidity of strict isomorphisms, which demand literal bijections on objects and morphisms without intermediary isomorphisms. Equivalences tolerate such "witnessing" isomorphisms, accommodating non-strict alignments; for example, the skeleton of a category—which collapses isomorphic objects into unique representatives—is equivalent to the original but rarely isomorphic due to the loss of redundant objects. Equivalences are far more prevalent in mathematical applications than strict isomorphisms, as most intuitively "isomorphic" categories differ only in representational choices rather than intrinsic structure. For instance, the category of finite-dimensional vector spaces over a k (with linear maps as morphisms) is equivalent to the category of finite matrices over k (with ), but not isomorphic, because bases are not canonically fixed and isomorphisms depend on basis selections. Saunders Mac Lane played a pivotal role in promoting equivalences, emphasizing them as the appropriate notion of categorical sameness to sidestep overly rigid set-theoretic constraints and better reflect mathematical equivalence in practice.

Full and Faithful Functors

A functor F: \mathcal{C} \to \mathcal{D} between categories is faithful if, for every pair of objects A, B in \mathcal{C}, the induced map on hom-sets F: \hom_{\mathcal{C}}(A, B) \to \hom_{\mathcal{D}}(F(A), F(B)) is injective. This means that distinct morphisms in \mathcal{C} are mapped to distinct morphisms in \mathcal{D}, preserving the distinction between different arrows without collapsing them. A functor F: \mathcal{C} \to \mathcal{D} is full if the same induced map on hom-sets is surjective for every pair of objects A, B. In other words, every in \mathcal{D} between the images F(A) and F(B) arises as the image under F of some in \mathcal{C}. A that is both full and faithful thus induces a on hom-sets, establishing a one-to-one correspondence between the morphisms in \mathcal{C} and those in \mathcal{D} between the corresponding images of objects. While provide strong control over morphisms, they are insufficient for an isomorphism of categories without additional conditions, such as bijectivity on objects. For instance, the inclusion of a into a larger is typically , as it restricts to bijections on hom-sets between objects in the subcategory, but it fails to be bijective on the entire collection of objects, preventing a strict . An isomorphism of categories requires a that is also bijective on objects and admits a strict , ensuring the categories are identical up to relabeling. Without the strict invertibility, contribute to weaker notions like equivalences when paired with essential surjectivity on objects. A concrete example is the forgetful functor U: \mathbf{Grp} \to \mathbf{Set}, which sends a group to its underlying set and a group homomorphism to the corresponding function on sets. This functor is faithful, as distinct group homomorphisms induce distinct set functions, but it is not full, since not every set function between underlying sets preserves the group structure. Such examples illustrate how full and faithful properties lay groundwork for comparing categorical structures, even if they do not yield isomorphisms on their own.

References

  1. [1]
    [PDF] maclane-categories.pdf - MIT Mathematics
    ... Categories for the working mathematician/Saunders Mac Lane. -. 2nd ed. p. cm ... isomorphism between corresponding sets of arrows. All of these forms ...
  2. [2]
    [PDF] Category theory in context Emily Riehl - Johns Hopkins University
    The category of categories gives rise to a notion of an isomorphism of categories, defined by interpreting Definition 1.1.5 in CAT. Namely, an isomorphism ...
  3. [3]
    [PDF] Category Theory
    A category is an abstraction of objects and morphisms, focusing on morphisms between objects, and includes a class of objects and morphisms between them.<|control11|><|separator|>
  4. [4]
  5. [5]
    [PDF] Category theory - Jakob Scholbach
    Jan 27, 2022 · one still obtains an isomorphism of categories (which are no ... inverse functor. Adjunctions are in between the void and the perfect ...
  6. [6]
    isomorphism in nLab
    Dec 12, 2024 · Two objects of a category are said to be isomorphic if there exists an isomorphism between them. This means that they “are the same for all ...Idea · Definitions · Properties
  7. [7]
    [PDF] Saunders Mac Lane - Categories for the Working Mathematician
    Categories for the working mathematician/Saunders Mac Lane. -. 2nd ed. p. cm ... and inverse limit, and that of pairs of functors with a natural isomorphism.
  8. [8]
    [PDF] Category Theory in Context Emily Riehl
    Mar 1, 2014 · The aim of theory really is, to a great extent, that of systematically organizing past experience in such a way that the next generation, ...
  9. [9]
    automorphism in nLab
    Oct 5, 2025 · An automorphism of an object x x in a category C C is an isomorphism f : x → x f : x \to x . In other words, an automorphism is an endomorphism ...Missing: identity | Show results with:identity
  10. [10]
    nontrivial isomorphisms of categories - MathOverflow
    Feb 3, 2010 · Indeed, if you look in Categories for the Working Mathematician, you'll see that Mac Lane calls a functor monadic if K is an isomorphism. He ...
  11. [11]
    [PDF] Finite spaces and larger contexts J. P. May - The University of Chicago
    we restrict attention to T0-spaces, that is, to posets. The proof is based ... from the category of posets to the category of spaces. Remember that the ...<|control11|><|separator|>
  12. [12]
    What are the automorphisms of $BG$? - MathOverflow
    Apr 7, 2015 · We can only say that two automorphisms are isomorphic. An automorphism of BG might have nontrivial (2-)automorphisms. What I've Tried: ...at.algebraic topology - What's the relationship between B Aut(G) and ...Automorphisms of category of groups [duplicate] - MathOverflowMore results from mathoverflow.net
  13. [13]
    GENERAL THEORY OF NATURAL EQUIVALENCES
    SAMUEL EILENBERG AND SAUNDERS MacLANE. Contents. Page. Introduction. 231. I. Categories and functors. 237. 1. Definition of categories. 237. 2. Examples of ...Missing: citation | Show results with:citation
  14. [14]
    [PDF] Basic Category Theory - arXiv
    Jan 5, 2017 · egory, and in particular for isomorphism of categories (which are objects of ... Saunders Mac Lane, Categories for the Working Mathematician.