Fact-checked by Grok 2 weeks ago

Commutative diagram

A commutative diagram in is a consisting of objects (vertices) and s (arrows) such that, for any two objects, the composition of s along every pair of paths connecting them results in the same . This commutativity condition ensures that the diagram captures consistent relational structures without ambiguity in how elements are mapped between objects. Commutative diagrams form a foundational tool in , serving as a visual and precise language for expressing abstract relationships, much like equations in algebraic contexts. They are indispensable for defining key concepts such as limits, colimits, products, coproducts, pullbacks, and pushouts, where the universal mapping properties are verified through commuting triangles or squares. Additionally, they illustrate the behavior of functors between categories and the naturality of transformations, enabling diagram chasing techniques to prove isomorphisms and equivalences. Beyond pure mathematics, commutative diagrams find applications in fields like , , and , where they model type systems, rules, and structural proofs in programming languages. Their graphical nature facilitates intuition and rigor, allowing mathematicians and computer scientists to abstract and unify diverse constructions across disciplines.

Fundamentals

Definition

In , a commutative diagram is a visual representation of objects and s within a such that the composition of morphisms along any two distinct paths connecting the same pair of objects yields the same resulting . This ensures that the diagram encodes equalities of composite morphisms independently of the chosen path, providing a concise way to express structural relationships without explicit computation of each . Such diagrams typically depict objects as vertices and morphisms as directed arrows between them. Commutativity implies path independence: for any two paths from an object A to an object B, the composite morphisms must be identical in the . This property captures the essence of relational equalities in categorical structures, where the diagram serves as a for verifying that multiple ways of mapping between objects are consistent. Formally, a commutative diagram is defined via a D: J \to C, where J is a small index (often a finite viewed as a free category) and C is the ambient . The diagram commutes if, for every pair of parallel arrows (morphisms with the same domain and codomain) in J, their images under D are equal in C; more generally, all composite morphisms along parallel paths in J map to equal composites in C. Equivalently, in the elementary sense, if J consists of vertices V and edges E, then D assigns objects of C to vertices and morphisms of C to edges (preserving sources and targets), and the diagram is commutative precisely when the composite D(p) = D(p') for any two paths p, p' in J from a vertex x to a vertex y. The concept of commutative diagrams was popularized in the category theory literature of the 1940s and 1950s through the foundational works of and , with early documented uses appearing in the 1945 paper by and titled "A General Theory of Natural Equivalences," and further developed in Mac Lane's 1963 monograph "".

Notation and Symbols

In commutative diagrams, morphisms between objects are typically represented by solid arrows, such as \to or \rightarrow, with labels indicating the specific map, for example, f: A \to B. Dashed arrows, denoted by \dashedrightarrow or similar, are conventionally used to indicate implied, hypothetical, or uniquely determined morphisms, often in the context of universal properties where existence is asserted but not explicitly constructed. These arrows ensure visual distinction between given structure and derived elements, promoting clarity in diagram interpretation. Common diagram shapes include squares, which depict two parallel pairs of morphisms whose compositions must equalize; triangles, representing triangular identities or cone structures; and polygons, such as pentagons or hexagons, for more intricate relations involving multiple paths. These geometric configurations are arranged to visually encode the requirement that all paths between two objects yield the same composite , with objects positioned at vertices and arrows along edges. For rendering these diagrams, LaTeX-based tools like TikZ with its tikz-cd extension provide precise control over arrow styles, labels, and layouts, generating high-quality output suitable for mathematical publications. Similarly, the xy-pic package specializes in category-theoretic diagrams, supporting complex arrow decorations and positional logic for automated alignment. Graphviz offers an alternative for automated rendering of directed graphs, including commutative structures, through its language, which excels in handling larger or algorithmic diagram generation. A notable variation in notation involves double arrows, such as \Rightarrow, to denote transformations between functors, briefly introducing a layer above morphisms without delving into higher categorical .

Commutativity Verification

To verify the commutativity of a in a category, one follows a systematic process that begins with identifying all directed paths between any pair of objects in the diagram. For each such pair, the composite morphisms along these paths must be computed and shown to be equal; for instance, in a triangular diagram with morphisms f: A \to B, g: B \to C, and h: A \to C, commutativity requires g \circ f = h. This path-by-path equality check ensures that the diagram factors through a preorder where parallel paths are identified, as formalized in the elementary definition of a commutative . Algebraic verification often leverages properties inherent to categorical structures, such as functoriality, to establish equalities without explicit computations. Functors preserve the commutativity of by objects and morphisms while maintaining and identities, so if a commutes in the source , its under a commutes in the target ; this is particularly useful for naturality squares of natural transformations, where commutativity follows directly from the . Associativity of further aids in rearranging paths to match known equalities, reducing the need for direct calculation in larger . Common pitfalls in verification arise when the underlying lacks full categorical , such as in mere sets without specified , where paths may not yield well-defined composites, leading to non-commutativity unless additional relations are imposed. In diagrams with cycles, overlooking the requirement for morphisms along loops can cause inconsistencies, as compositions must return to the starting object unchanged. Similarly, fork structures (e.g., two morphisms from a ) demand outright of the morphisms, not merely post-composition with a common target, to avoid subtle failures in factorization. Detecting such issues involves checking for unresolved parallel paths or violations of conditions. For practical verification, especially in finite or concrete categories, matrix representations provide a computational approach: in finite-dimensional vector spaces, morphisms are represented as matrices, and commutativity reduces to equality of matrix products along paths, verifiable via linear algebra software. In formal settings, proof assistants like Coq enable rigorous checks through libraries that automate diagram chasing and commutativity proofs; for example, a Coq-based system uses tactics to synthesize and verify equalities in 1-categories by embedding diagrams as quivers and resolving the "commerge" problem for acyclic cases. These tools ensure decidability and reliability for complex verifications.

Examples

Basic Algebraic Examples

One foundational example of a commutative diagram arises in the context of group homomorphisms, particularly through short exact sequences. Consider the sequence $0 \to A \xrightarrow{i} B \xrightarrow{p} C \to 0, where A is a of the group B, i is the , p: B \to C = B/A is the , and the maps are group homomorphisms. This diagram is commutative because the composition p \circ i is the zero map, and exactness ensures \operatorname{im} i = \ker p with p surjective. In , a commutative square demonstrates the universal property of quotient rings. The canonical projection \pi: \mathbb{Z} \to \mathbb{Z}/n\mathbb{Z} and the inclusion \iota: \mathbb{Z} \to \mathbb{Q} form two paths from \mathbb{Z}; for any \phi: \mathbb{Z} \to R such that n annihilates the image (as in the case where R has dividing n), there exists a unique induced map \overline{\phi}: \mathbb{Z}/n\mathbb{Z} \to R making the square \mathbb{Z} \to \mathbb{Z}/n\mathbb{Z} \to R and \mathbb{Z} \to R commute, i.e., \phi = \overline{\phi} \circ \pi. While no nontrivial map \mathbb{Z}/n\mathbb{Z} \to \mathbb{Q} exists for n > 1 due to characteristic differences, the illustrates the factoring property when applicable to other targets like finite fields. For vector spaces, commutative diagrams often appear in the representation of linear maps under change of bases. Suppose V and W are finite-dimensional vector spaces over a field k, with bases \{v_1, \dots, v_n\} for V and \{w_1, \dots, w_m\} for W, and a linear map T: V \to W. A change of basis in V via invertible matrix P yields a new basis, transforming T to T' = Q^{-1} T P in the new coordinates, where Q is the change-of-basis matrix for W. The square diagram V \xrightarrow{T} W, V \xrightarrow{P} V' \xrightarrow{T'} W' \xleftarrow{Q} W commutes, as the representations align under basis transformations preserving linearity. A specific commutative triangle exemplifies the first isomorphism theorem for groups. For a group homomorphism f: G \to H with kernel K = \ker f, the diagram consists of the inclusion \iota: K \to G, the projection \pi: G \to G/K, and the induced isomorphism \overline{f}: G/K \to \operatorname{im} f \subseteq H, such that f = \overline{f} \circ \pi and the triangle K \to G \to \operatorname{im} f with K \to 0 \to \operatorname{im} f commutes, establishing G/K \cong \operatorname{im} f.

Topological and Geometric Examples

In , commutative diagrams arise naturally in the study of and their relation to the . Consider a covering space p: \tilde{X} \to X of a path-connected, locally path-connected space X, with a H \subseteq \pi_1(X, x_0) corresponding to the covering via the Galois correspondence. The i: H \to \pi_1(X, x_0) induces a commutative diagram where the p_*: \pi_1(\tilde{X}, \tilde{x}_0) \to H commutes with the inclusion, reflecting how loops in the covering space project to loops in the base fixed by the action. Deck transformations, which are automorphisms of the covering space commuting with the projection p, further illustrate commutativity: for a deck transformation \tau: \tilde{X} \to \tilde{X}, the diagram \tilde{X} \xrightarrow{\tau} \tilde{X} \xrightarrow{p} X and \tilde{X} \xrightarrow{p} X \xrightarrow{\mathrm{id}} X commutes, ensuring that \tau preserves fibers and induces the identity on the base. This structure underpins the isomorphism between the deck transformation group and the quotient \pi_1(X)/H, highlighting how topological coverings encode group actions via commuting projections. Homotopy commutative squares extend this to higher-dimensional invariants, particularly in the context of groups. A homotopy commutative square consists of maps f: X \to Y, g: X \to A, h: Y \to B, and k: A \to B such that h \circ f \simeq k \circ g up to , meaning there exists a H: X \times I \to B with H(-,0) = h \circ f and H(-,1) = k \circ g. In the definition of higher groups \pi_n(X, x_0), such squares appear in the relative long exact sequence for a pair (X, A), where the boundary map \partial: \pi_n(X, A) \to \pi_{n-1}(A) is induced by a commutative diagram involving the inclusion i: A \to X and q: X \to X/A, ensuring that spheres in the lift appropriately up to . For instance, in computing \pi_n(S^n), the square relating the attaching map of cells in the CW-complex structure of the sphere commutes up to , confirming \pi_n(S^n) \cong \mathbb{Z} for the . This commutativity captures the invariance of classes under continuous deformations, distinguishing topological from strict algebraic commutativity. In , commutative diagrams manifest in , such as the fiber product of manifolds. Given smooth maps f: M \to N and g: P \to N between manifolds, the fiber product M \times_N P = \{(m, p) \in M \times P \mid f(m) = g(p)\} forms the vertex of a commutative square with projections \pi_M: M \times_N P \to M and \pi_P: M \times_N P \to P, satisfying f \circ \pi_M = g \circ \pi_P. If f and g are submersions, the fiber product inherits a manifold , with the diagram's commutativity ensuring it is the categorical pullback in the category of manifolds. This construction is pivotal in theory, where pulling back a bundle E \to N along f: M \to N yields f^*E \to M via the commutative square involving the bundle projection, preserving local triviality and transition functions. An illustrative example from differential geometry involves connection forms on principal bundles and their curvature. For a principal G-bundle \pi: P \to M with connection 1-form \omega \in \Omega^1(P, \mathfrak{g}), the curvature 2-form \Omega = d\omega + \frac{1}{2}[\omega, \omega] satisfies the structure equation, which underlies commutativity in the diagram for horizontal lifts: the infinitesimal parallel transport along vector fields commutes with the right G-action via the relation R_g^* \omega = \mathrm{Ad}_{g^{-1}} \omega, ensuring \Omega is horizontal and G-invariant. In the Maurer-Cartan form context, this commutativity via curvature equations measures the failure of flatness, as seen in the Bianchi identity d_\omega \Omega = 0, where the covariant derivative d_\omega \Omega = d\Omega + [\omega, \Omega] vanishes, forming a closed loop in the de Rham complex of Lie algebra-valued forms. Such diagrams connect local connection data to global topological invariants like Chern classes.

Techniques and Applications

Diagram Chasing

Diagram chasing is a proof technique in that involves systematically tracing elements or morphisms through the paths of a commutative to establish equalities, injectivity, surjectivity, or exactness. This method exploits the commutativity of the and properties such as exactness of sequences, kernels, and cokernels to "" an element forward along one path and backward along another equivalent path, often revealing relationships between homology groups or modules. A prominent example is the , which derives a long exact from a commutative diagram of two short exact sequences of abelian groups or modules. Consider the following commutative diagram with exact rows:
0 → A ──→ B ──→ C ──→ 0
    │     │     │
    ↓     ↓     ↓
0 → A'──→ B'──→ C'──→ 0
Here, the vertical maps are f: A \to A', g: B \to B', and h: C \to C'. The snake lemma asserts the existence of a connecting \delta: \ker h \to \coker f such that the sequence $0 \to \ker f \to \ker g \to \ker h \xrightarrow{\delta} \coker f \to \coker g \to \coker h \to 0 is exact. To prove this via diagram chasing, first define \delta on an element c \in \ker h: since the bottom row is exact, there exists b' \in B' such that the map B' \to C' sends b' to c; by commutativity, h(c) = 0 implies the preimage under the top row map B \to C lifts appropriately, but chasing backward, the element in \ker g maps to zero in B', so its image under the cokernel map from A yields the class in \coker f. Exactness at \ker h follows by chasing an element in \im(\ker g \to \ker h) forward to show it maps to zero in \coker f, using surjectivity of the bottom row; similarly, exactness at \coker f is shown by lifting elements from \ker h and verifying they cover the kernel of \coker f \to \coker g, leveraging injectivity of the top row. This process connects the kernel of the rightmost vertical map to the cokernel of the leftmost, "snaking" through the . The zig-zag lemma, often synonymous with the in this context, emphasizes the alternating forward and backward paths used in to prove properties like injectivity or surjectivity in s of chain complexes. For instance, in a short of chain complexes $0 \to C \xrightarrow{\phi} D \xrightarrow{\psi} E \to 0, the induced long exact sequence in arises by chasing elements zig-zag through the of differentials and maps, establishing exactness by alternating applications of and constructions. Diagram chasing finds key applications in proving the five lemma and nine lemma in homological algebra. The five lemma states that in a commutative diagram of abelian groups with exact rows:
A ──→ B ──→ C ──→ D ──→ E
│     │     │     │     │
A'──→ B'──→ C'──→ D'──→ E'
if the vertical maps on A, B, D, E are isomorphisms and the rows are exact, then the map on C is an isomorphism; this is established by chasing elements to show injectivity and surjectivity at C. The nine lemma extends this to a $3 \times 3 grid diagram, concluding that if the boundary maps (first two columns and rows) are isomorphisms, then the central map is an isomorphism, again via systematic chasing to relate kernels and cokernels across the grid. These lemmas underpin many results in algebraic topology and module theory.

Canonical Constructions

In , canonical constructions of commutative diagrams arise from universal properties of limits and colimits, providing standard building blocks that ensure commutativity through unique factorizations of morphisms. These constructions, such as pullbacks, pushouts, products, coproducts, equalizers, and coequalizers, define objects and arrows that mediate between given diagrams in a way that preserves equality of composite paths. Pullback diagrams are limits of cospans, consisting of two morphisms with a common codomain, and form a commutative square via the universal property. For objects A, B, and C with morphisms f: A \to C and g: B \to C, the pullback is an object P equipped with morphisms p_1: P \to A and p_2: P \to B such that f \circ p_1 = g \circ p_2, and for any object X with morphisms x_1: X \to A and x_2: X \to B satisfying f \circ x_1 = g \circ x_2, there exists a unique morphism u: X \to P making the diagram commute, i.e., p_1 \circ u = x_1 and p_2 \circ u = x_2. This ensures commutativity in the square, as the universal property guarantees the equality of composites. In the category of sets, the pullback corresponds to the fiber product, where P = \{(a, b) \mid f(a) = g(b)\}. Dually, pushout diagrams are colimits of spans, with two morphisms sharing a common domain, forming a commutative square. For morphisms i: A \to B and j: A \to C, the pushout is an object D with morphisms k: B \to D and l: C \to D such that k \circ i = l \circ j, and for any E with m: B \to E and n: C \to E satisfying m \circ i = n \circ j, there is a unique v: D \to E with v \circ k = m and v \circ l = n. This universal property enforces commutativity, and in sets, it yields the disjoint union quotiented by identifying images of A. Product diagrams arise as limits of discrete diagrams with multiple objects, typically forming commutative triangles via projection morphisms. For objects A and B, the product A \times B comes with projections \pi_A: A \times B \to A and \pi_B: A \times B \to B, such that for any X with morphisms f: X \to A and g: X \to B, there is a unique \langle f, g \rangle: X \to A \times B satisfying \pi_A \circ \langle f, g \rangle = f and \pi_B \circ \langle f, g \rangle = g, ensuring the triangle commutes. This construction generalizes to finite or infinite products in categories with them, like the in sets. Coproducts, as colimits of diagrams, form dual commutative triangles using inclusions. For A and B, the coproduct A \coprod B has injections i_A: A \to A \coprod B and i_B: B \to A \coprod B, such that for any Z with u: A \to Z and v: B \to Z, there exists a unique [u, v]: A \coprod B \to Z with [u, v] \circ i_A = u and [u, v] \circ i_B = v, guaranteeing commutativity. In sets, this is the . Equalizer diagrams are limits of fork diagrams with two parallel arrows, yielding a commutative setup with unique . Given parallel morphisms f, g: A \to B, the is an object E with morphism e: E \to A such that f \circ e = g \circ e, and for any X with h: X \to A where f \circ h = g \circ h, there is a unique k: X \to E with e \circ k = h. This ensures the diagram commutes by equalizing the paths through f and g. In abelian categories, equalizers coincide with kernels. Dually, coequalizer diagrams are colimits of forks, with a morphism q: B \to C such that for parallel f, g: A \to B, q \circ f = q \circ g, and for any D with r: B \to D where r \circ f = r \circ g, there is a unique s: C \to D with s \circ q = r. Commutativity follows from the coequalizing property, and in abelian categories, coequalizers are cokernels. In abelian categories, short exact sequences provide canonical commutative diagrams embodying these constructions. A sequence $0 \to A \xrightarrow{i} B \xrightarrow{p} C \to 0 is short exact if i is the kernel of p (an equalizer of p and the zero map) and p is the cokernel of i (a coequalizer of i and zero), ensuring exactness at A, B, and C. This forms a commutative diagram with zero morphisms at the ends, where the universal properties of kernel and cokernel guarantee the compositions p \circ i = 0 and the exactness conditions. Such sequences are foundational for , representing extensions where B is an extension of C by A, and morphisms between them induce commutative ladders.

Theoretical Context

Role in Category Theory

Commutative diagrams play a foundational role in category theory, emerging as essential tools in the axiomatization of categories by Samuel Eilenberg and Saunders Mac Lane in the 1940s. In their seminal 1945 paper, they introduced the concepts of categories, functors, and natural transformations, where commutative diagrams served to formalize the conditions under which transformations between functors preserve the structure of morphisms across objects. This axiomatic framework, motivated by problems in algebraic topology and homology, used diagrams to ensure that compositions of morphisms yield consistent results regardless of the path taken, thereby establishing commutativity as a core principle for relational consistency in abstract mathematical structures. Central to , commutative diagrams define s as families of s that make all relevant squares commute, meaning that for functors F, G: \mathcal{C} \to \mathcal{D}, a \eta: F \Rightarrow G consists of components \eta_X: F(X) \to G(X) for each object X in \mathcal{C}, such that the involving F(f) and G(f) for any f: X \to Y commutes. This diagrammatic condition ensures the transformation is "natural," independent of choices in the , and underpins the theory's ability to abstract equivalences across diverse mathematical domains. Similarly, commutative diagrams are instrumental in defining s and colimits: a of a D: J \to \mathcal{C} is an object L with projections p_i: L \to D(i) such that all triangles involving maps to other potential limits commute, guaranteeing universality. Colimits operate dually, with commutative conditions ensuring the universal property for coprojections. Adjunctions, another cornerstone, rely on commutative diagrams to characterize pairs of functors F: \mathcal{C} \to \mathcal{D} and G: \mathcal{D} \to \mathcal{C} as , where the unit \eta: 1_\mathcal{C} \to GF and counit \epsilon: FG \to 1_\mathcal{D} satisfy the triangular identities—commutativity of the diagrams linking \eta, \epsilon, F, and G. These conditions encapsulate reciprocal relationships, such as groups or tensor-hom adjunctions, enabling powerful abstractions in and . Furthermore, commutative diagrams underpin universal mapping properties, where commutativity ensures the uniqueness of mediating morphisms; this is exemplified in the , which states that for a \mathcal{C}, the hom-functor \mathcal{C}(C, -) is fully faithful, with natural isomorphisms determined by commuting squares that preserve the action on morphisms, thus objects via their mappings without loss of information.

Functorial Interpretation

In , a commutative can be rigorously interpreted as a D: \mathcal{J} \to \mathcal{C}, where \mathcal{J} is a small index category and \mathcal{C} is the target category. The objects of \mathcal{J} correspond to the vertices (or nodes) of the diagram, while its morphisms represent the directed edges (or arrows) them. This formalizes the diagram as a structured that embeds the combinatorial skeleton provided by \mathcal{J} into the morphisms and objects of \mathcal{C}, enabling precise statements about preservation of relations. The commutativity of the diagram is a direct consequence of the functoriality of D. Specifically, for any pair of composable morphisms f: j \to k and g: k \to l in \mathcal{J}, the functor ensures that D(g) \circ D(f) = D(g \circ f) in \mathcal{C}. This equality guarantees that all paths from an object D(j) to D(l) in the diagram yield the same composite morphism, eliminating ambiguities in the structure and allowing diagrams to serve as concise proofs of equalities between morphisms. Particularly in the context of universal constructions, and colimit diagrams are functors D: \mathcal{J} \to \mathcal{C} endowed with additional structure via s. A of D is an object L in \mathcal{C} together with a cone—a \pi: \Delta L \to D (where \Delta is the diagonal \mathcal{C} \to \mathcal{C}^{\mathcal{J}}) such that for any other cone \sigma: \Delta X \to D, there exists a unique u: X \to L making the evident triangle commute. This inherently enforces commutativity throughout the , as any mediating must respect all paths defined by D. Colimits are , represented by cocones \iota: D \to \Delta C with a co-mediating , again preserving commutativity via the 's action on compositions. The functorial interpretation also underpins the pasting lemma for composite diagrams, which asserts that commutativity extends across amalgamated index categories. If two diagrams are given by functors D_1: \mathcal{J}_1 \to \mathcal{C} and D_2: \mathcal{J}_2 \to \mathcal{C} sharing a subdiagram along an embedding of index categories, the composite diagram defined by the pushout (or pullback) in the category of index categories induces a functor D: \mathcal{J}_1 \coprod_{\mathcal{J}} \mathcal{J}_2 \to \mathcal{C} whose commutativity follows from the preservation of colimits (or limits) by D, ensuring that pasted squares or rectangles commute when the individual components do. This lemma is essential for verifying larger diagrams by local checks, leveraging the global consistency enforced by functoriality.

References

  1. [1]
    [PDF] Intro to Category Theory - cs.wisc.edu
    Page 3. Definition 2.2. A commutative diagram is a directed graph where any two paths that both start. at the same point and end at the same point, yield the ...
  2. [2]
    [PDF] limits in category theory - UChicago Math
    Aug 17, 2007 · A commutative diagram is one where, between any two given objects, composi- tion along every (directed) path of arrows yields the same morphism ...
  3. [3]
    [PDF] An Introduction to Category Theory - eCommons
    Dec 31, 2023 · Commutative diagrams are an essential tool in category theory, often described as analogous to the role equations play in algebra. 4.3 ...
  4. [4]
    [PDF] Category Theory
    (To say that a diagram such as this is commutative is to say that the com- positions of morphisms along every path from one object to another yield the same ...
  5. [5]
    [PDF] an introduction to category theory and the yoneda lemma
    Aug 28, 2020 · Commutative diagrams are used very frequently in category theory for diagram chasing and provide a visualization in cases where one may ...
  6. [6]
    [PDF] Chapter 4 - Basic category theory - MIT OpenCourseWare
    Our commutative square is precisely what's needed for a graph homomorphism, as shown in Exercise 3.3.3.7. Thus, we have defined. F on objects and on morphisms.
  7. [7]
    [PDF] Category Theory for Linear Logicians
    Sep 10, 2003 · Equations in categories are also typically represented graphically and are called commutative diagrams. Many familiar classes of structures in ...
  8. [8]
    Commutative Diagram -- from Wolfram MathWorld
    A commutative diagram is a collection of maps in which all map compositions starting from the same set and ending with the same set give the same result. In ...
  9. [9]
    commutative diagram in nLab
    ### Precise Definition of a Commutative Diagram
  10. [10]
    whence commutative diagrams? - MathOverflow
    Mar 24, 2011 · It seems that commutative diagrams appeared sometime in the late 1940s -- for example, Eilenberg-McLane (1943) group cohomology paper does not have any.
  11. [11]
    [PDF] HOMOLOGY
    SAUNDERS MAC LANE. MAX MASON DISTINGUISHED SERVICE PROFESSOR OF ... EILENBERG-MAC LANE 1954, Thm. 26.51. 7. For p a prime, P the additive group ...
  12. [12]
  13. [13]
    [PDF] tikzcd: Commutative diagrams with TikZ - CTAN
    May 4, 2021 · The general-purpose drawing package TikZ can be used to typeset commutative diagrams and other kinds of mathematical pictures, generating ...
  14. [14]
    Xy -pic — Typesetting Graphs and Diagrams in TeX
    Xy-pic is a package for typesetting graphs and diagrams using the principle of logical composition of visual components.
  15. [15]
    tex-archive/macros/generic/diagrams/xypic - CTAN
    May 6, 2010 · ... Xy-pic is a package for typesetting graphs and diagrams ... category theory, automata theory, algebra, neural networks, and database theory.
  16. [16]
    Graphviz
    Graphviz is open source graph visualization software that represents structural information as diagrams of abstract graphs and networks.DOT Language · Download · Graphviz · AttributesMissing: commutative | Show results with:commutative
  17. [17]
    [PDF] Intro to Category Theory: Functors 1 Functors - cs.wisc.edu
    One of the nice things about the definition of a functor is that they preserve commutative diagrams. Theorem 1.1. Functors preserve commutative diagrams. That ...<|separator|>
  18. [18]
    [PDF] Notes on Representations of Finite Groups
    The moral of this example is that a representation is an assignment of a matrix to each element of a group, in a way that respects composition of group elements ...
  19. [19]
    [PDF] WOMP 2004: HOMOLOGICAL ALGEBRA Kate Ponto
    Sep 24, 2004 · An exact sequence 0 → A → B → C → 0 is called a short exact sequence. ... If a sequence 0 → C → 0 of Abelian groups is exact then C = 0.
  20. [20]
    [PDF] Homological Algebra - Department of Mathematics
    Suppose we have a commutative diagram of R-modules and R-homomorphisms as below, denoted by solid morphisms, where the rows are exact. A. B. C. 0. 0. A0. B0. C0.<|separator|>
  21. [21]
    [PDF] 1. Introduction to rings, ideals and homomorphisms
    Introduction to rings, ideals and homomorphisms. Commutative algebra is the study of commutative, associative rings with.
  22. [22]
    [PDF] mas 439 commutative algebra and algebraic geometry
    The quotient map has the following universal property. Proposition 8.1 (Universal property of quotient rings). Suppose that f : R → S is a ring homomorphism ...
  23. [23]
    Linear Maps - I - Euclidean Tensor Analysis
    Notice how the commutative diagram neatly summarizes the change in the representation of a linear map upon change of bases in the domain and codomain vector ...
  24. [24]
    [PDF] ISOMORPHISMS 1. Introduction Groups that are not literally the ...
    A visual way to describe the first isomorphism theorem is through a commutative diagram. Starting with the homomorphism f : G → CG having kernel K, the ...
  25. [25]
    [PDF] Algebraic Topology - Cornell Mathematics
    This book was written to be a readable introduction to algebraic topology with rather broad coverage of the subject. The viewpoint is quite classical in ...
  26. [26]
    [PDF] Review 1: Mathematics 432/532 - University of Oregon
    Feb 9, 2006 · X → X be a covering space. A deck transformation is a homeomorphism d : ˜X → X such that the diagram below commutes. The set of all deck ...
  27. [27]
    [PDF] The Topology of Fiber Bundles Lecture Notes Ralph L. Cohen Dept ...
    A fiber bundle is a parameterized family of objects, each isomorphic to a fiber, where the family is parameterized by points in the base space.
  28. [28]
    [PDF] 240AB Differential Geometry - UCI Mathematics
    We next claim that this diagram commutes up to sign, so by changing some of the vertical maps to their negatives if necessary, we obtain a commutative diagram.<|separator|>
  29. [29]
    [PDF] Characteristic Classes, Principal Bundles, and Curvature
    Jul 26, 2024 · We can apply this process to the curvature form of a connection Ω in order to yield a well defined form f( ¯Ω) ∈ Ω2k(B, R). The following ...Missing: commutative | Show results with:commutative
  30. [30]
    diagram chasing in nLab
    ### Definition and Role of Diagram Chasing in Homological Algebra
  31. [31]
    Diagram Chasing -- from Wolfram MathWorld
    The construction of the connecting homomorphism in the proof of the snake lemma is an example of diagram chasing. See also. Commutative Diagram, Connecting ...
  32. [32]
    [PDF] A crash course in homological algebra - Purdue Math
    A useful skill in this business is to be able to prove things by diagram chasing. ... Theorem 2.0.8 (Snake Lemma). If. 0. // A. // f. ✏✏. B. // g. ✏✏. C. // h. ✏✏.
  33. [33]
    [PDF] A Primer on Homological Algebra - UChicago Math
    Jul 12, 2013 · It is called Snake Lemma because the induced exact sequence zig zags through the original diagram. 0. 0. 0 kerα. // kerβ. // kerγ. A. //. B. //.
  34. [34]
    Zig-Zag Lemma -- from Wolfram MathWorld
    A diagram lemma which states that every short exact sequence of chain complexes and chain homomorphisms 0-->C-->^phiD-->^psiE-->0 gives rise to a long exact ...
  35. [35]
    [PDF] Homological Algebra - Stacks Project
    This document explains basic homological algebra, covering topics such as preadditive and additive categories, Karoubian categories, and Abelian categories.
  36. [36]
    [PDF] maclane-categories.pdf - MIT Mathematics
    ... Commutative Algebra. Analysis, and Zeta-Functions. 2nd ed. Vol.!. 59 LANG ... Saunders Mac Lane. Categories for the. Working Mathematician. Second Edition.
  37. [37]
    [PDF] AN INTRODUCTION TO HOMOLOGICAL ALGEBRA
    ... Commutative ring theory. 9 C. B. Thomas Characteristic classes and the ... Weibel An introduction to homological algebra. 39 W. Bruns and J. Herzog Cohen ...
  38. [38]
    Categories for the Working Mathematician - SpringerLink
    Saunders Mac Lane. Pages 1-5. Categories, Functors, and Natural ... Series Title: Graduate Texts in Mathematics. DOI: https://doi.org/10.1007/978-1-4757 ...