Fact-checked by Grok 2 weeks ago

Canonical map

In , a , also known as a map, is a between objects that arises directly from their structural definitions or properties, without involving arbitrary choices or decisions. This ensures the map is uniquely determined by the inherent , making it a standard and intrinsic construction in various mathematical domains. Canonical maps are particularly prominent in , where they often emerge from universal properties, such as the s defining products or the unique factoring through a object. For example, in the , the canonical s from a product set X \times Y to its factors X and Y are determined solely by the universal property of the product. These maps facilitate proofs and constructions by providing a canonical way to relate objects, highlighting the systematic structure underlying . Specific instances abound across fields; in , the canonical map from a set X to its quotient X/R by an R sends each element to its , a surjection central to partitioning concepts. In group theory, the canonical map \pi: G \to G/H from a group G to its quotient by a normal subgroup H assigns elements to their left cosets, preserving the group operation on the quotient. Such maps underscore the role of canonical constructions in simplifying complex structures and enabling modular algebraic reasoning.

Definition and Principles

Formal Definition

In mathematics, a canonical map is a morphism \phi: A \to B between mathematical objects A and B that arises inherently from the definitions or constructions of A and B, preserving as much of the inherent structure as possible and often being unique up to isomorphism. This morphism is intrinsically tied to the properties defining the objects, ensuring it emerges without reliance on extraneous data or conventions. The naturality of a canonical map stems from its determination solely by the intrinsic features of A and B, independent of external selections such as bases, orderings, or other arbitrary choices. For instance, it avoids the introduction of non-essential elements that could vary between equivalent constructions of the same objects. In contrast to arbitrary maps, which may depend on specific, decisions, a canonical map is distinguished by being "the obvious one"—the standard, structure-preserving choice that any consistent construction would yield. The term "canonical map" gained prominence in mid-20th-century abstract mathematics, particularly through the works of the French school such as Bourbaki and Grothendieck in and .

Key Properties

Canonical maps are distinguished by their uniqueness, arising naturally from universal mapping properties (UMPs) in the relevant category, such that any two maps satisfying the same UMP are equal or via a unique isomorphism. This ensures that the map is determined solely by the structure of the objects involved, without reliance on extraneous choices. A core attribute is their preservation of structure, where canonical maps maintain the algebraic, topological, or categorical properties of the domain and codomain to the maximal extent possible, often as homomorphisms or natural transformations that commute with the given operations or morphisms. For instance, in categories like sets or vector spaces, they respect inclusions, projections, or dualities inherent to the objects. While typically , some maps incorporate standardized conventions, such as sign choices in oriented structures or ordering in lattices, to resolve ambiguities while remaining field-specific standards. These conventions ensure consistency across applications without introducing arbitrariness. When a canonical map is bijective with a two-sided that is also canonical, it is termed a canonical , exemplified by the on objects or the Yoneda embedding, which is full and faithful. In proofs, canonical maps facilitate establishing equivalences between categories or reducing complex problems to simpler canonical forms, leveraging their to verify structural identities or induct on constructions. They often serve as the mediating morphisms in limits and colimits, confirming universality.

Canonical Maps in Set Theory

Quotient Projections

In set theory, given a set X and an equivalence relation \sim on X, the canonical projection, denoted \pi: X \to X/\sim, maps each element x \in X to its equivalence class .[](http://people.whitman.edu/~guichard/260/halmos__naive_set_theory.pdf) This map is defined such that $\pi(x) = \{ y \in X \mid y \sim x \}$, where the equivalence class consists of all elements in X related to x under \sim. The fibers of \pi, or the preimages \pi^{-1}(\{\}), are precisely the equivalence classes, so the kernel of \pi corresponds exactly to the partition induced by \sim. The canonical projection \pi is always surjective by construction, as every equivalence class $$ in the quotient set X/\sim is the image of at least the element x. This surjectivity ensures that X/\sim fully captures the distinct classes formed by the relation. A key characterizing feature of \pi is its universal property: for any set Y and any function f: X \to Y that is constant on equivalence classes (i.e., f(x_1) = f(x_2) whenever x_1 \sim x_2), there exists a unique function \overline{f}: X/\sim \to Y such that f = \overline{f} \circ \pi. This property establishes \pi as the universal morphism from X to any set respecting the equivalence relation, allowing maps from X to factor through the quotient in a unique way. Through this construction, the canonical projection plays a central role in partitioning X, as the quotient set X/\sim is precisely the set of all equivalence classes, forming a partition of X into disjoint, non-empty subsets whose union is X. For example, considering the integers \mathbb{Z} with the equivalence relation m \sim n if m - n is divisible by 2, the projection \pi: \mathbb{Z} \to \mathbb{Z}/\sim maps even integers to one class and odd integers to another, yielding the quotient \{ [\text{even}], [\text{odd}] \}.

Equivalence Class Maps

In , a inclusion map arises naturally when considering an as a of the original set. For an \sim on a set X, each _{\sim} = \{ y \in X \mid y \sim x \} is a of X, and the i_{}: _{\sim} \to X defined by i_{}(y) = y for all y \in _{\sim} embeds the class into X in the standard way. This map is as it is uniquely determined by the relation _{\sim} \subseteq X and preserves the set-theoretic structure without additional choices. In the special case where \sim is the equality relation on X, the are singletons \{x\}, and the inclusion \{x\} \to X similarly provides a embedding of points into the set. More generally, such inclusions can embed a X/\sim into a larger structure when the quotient is induced as a of a or product space, maintaining the natural identification of classes with their elements. Beyond basic projections, maps between sets emerge when comparing relations on the same set X. Suppose \sim_1 and \sim_2 are relations on X such that \sim_1 refines \sim_2, meaning every \sim_1- is contained in some \sim_2- (or equivalently, x \sim_2 y whenever x \sim_1 y). Then, there is a surjection \phi: X/\sim_1 \to X/\sim_2 defined by \phi(_{\sim_1}) = _{\sim_2}, which maps each finer to the coarser containing it. This construction generalizes the projection discussed earlier, as it induces a natural between the resulting structures. In terms of logic, where relations correspond to ordered by refinement, this surjection is the unique map preserving the distinctions (or atoms) of the . Such maps are well-defined because if _{\sim_1} = _{\sim_1}, then x \sim_1 z, implying x \sim_2 z by refinement, so _{\sim_2} = _{\sim_2}; thus, \phi depends only on the \sim_1-classes. Moreover, these maps are structure-preserving with respect to any partial order induced on the quotient sets—for instance, if \preceq is a partial order on X compatible with the equivalences, the quotients inherit a quotient order _{\sim} \preceq _{\sim} if there exist representatives satisfying x' \preceq y', and \phi respects this order by sending ordered finer classes to ordered coarser ones. These properties follow directly from the refinement relation and the universal mapping property of the surjections in the category of sets. In logic, these canonical maps between quotients and inclusions of classes find applications in , particularly for constructing canonical extensions of partial relations. For example, partial equivalence relations on a model can be extended to full relations via refinement, with the induced surjections providing canonical ways to relate the extended structures while preserving logical properties like definability and . This technique aids in building ultrapowers or other generic extensions where equivalence classes represent indistinguishable elements. The uniqueness of these canonical maps stems from their determination solely by the equivalence relations involved: the inclusion i_{} is fixed by the subset membership, and the surjection \phi is fixed by the refinement order, with no further choices required beyond the relations themselves. This canonicity ensures they are the standard tools for comparing quotient structures in set-theoretic constructions.

Canonical Maps in Abstract Algebra

Group Homomorphisms

In group theory, a canonical projection arises when N is a of a group G, defining the map \pi: G \to G/N by \pi(g) = gN, where gN denotes the left of N in G. This is a surjective , as it maps onto all cosets and preserves the group operation: \pi(gh) = (gh)N = (gN)(hN) = \pi(g)\pi(h) for all g, h \in G. The of \pi is precisely N, since \ker(\pi) = \{g \in G \mid \pi(g) = N\} = \{g \in G \mid g \in N\} = N. The canonical projection satisfies a universal property: for any group homomorphism \phi: G \to H such that N \subseteq \ker(\phi), there exists a homomorphism \psi: G/N \to H with \phi = \psi \circ \pi. This factorization ensures that the G/N captures the structure of G modulo N, making \pi the natural map through which such homomorphisms factor ly. Another canonical map is the i: H \to G for a H \leq G, defined by i(h) = h for all h \in H. This map is a , as i(h_1 h_2) = h_1 h_2 = i(h_1) i(h_2) for h_1, h_2 \in H, and it is injective since distinct elements in H remain distinct in G. As a natural , it preserves the group operation without altering the elements. These maps connect via the first isomorphism theorem, which states that for any group homomorphism \phi: G \to H, the quotient G / \ker(\phi) is isomorphic to the image \operatorname{im}(\phi). In particular, when N = \ker(\phi), the canonical projection \pi: G \to G/N induces an isomorphism G/N \cong \operatorname{im}(\phi), establishing the structural equivalence between the quotient and the homomorphism's range.

Module Projections

In module theory, the canonical projection arises in the construction of quotient modules. Given a ring R and an R-module M with submodule N \subseteq M, the quotient module M/N consists of cosets m + N for m \in M, equipped with the induced R-module structure (r \cdot (m + N)) = rm + N. The canonical projection \pi: M \to M/N is defined by \pi(m) = m + N, analogous to the set-theoretic quotient map but preserving the module operations. This map \pi is an R-, meaning it respects both and : \pi(m_1 + m_2) = \pi(m_1) + \pi(m_2) and \pi(rm) = r \pi(m) for r \in R. It is surjective by construction, as every is hit, and its is precisely N, since \pi(m) = 0 if and only if m \in N. The universal property of the quotient module characterizes \pi up to : for any R-module P and R- \phi: M \to P such that \phi(N) = 0, there exists a unique R- \overline{\phi}: M/N \to P with \phi = \overline{\phi} \circ \pi, defined by \overline{\phi}(m + N) = \phi(m). This property ensures that homomorphisms factoring through submodules are uniquely determined by their action on the quotient. In the context of rings viewed as modules over themselves, the canonical projection extends to a ring epimorphism. For a ring R with two-sided ideal I \trianglelefteq R, the quotient ring R/I inherits a ring structure via (r + I)(s + I) = rs + I, and the map \pi: R \to R/I given by \pi(r) = r + I is a surjective ring homomorphism that preserves both addition and multiplication. A key consequence is the third isomorphism theorem for modules, which relates nested quotients: if N \subseteq K \subseteq M are submodules, then (M/N)/(K/N) \cong M/K as R-modules, via the induced map sending (m + N) + (K/N) to m + K. This theorem facilitates computations of successive quotients by collapsing submodules stepwise. For free modules, a natural inclusion map embeds the direct sum into the direct product. Consider a family of R-modules \{M_i\}_{i \in I}; the direct sum \bigoplus_{i \in I} M_i, consisting of tuples with finitely many nonzero entries, maps canonically into the direct product \prod_{i \in I} M_i by extending finite-support tuples with zeros, forming a submodule when I is infinite. This map is an isomorphism when I is finite.

Canonical Maps in Linear Algebra

Vector Space Quotients

In the context of finite-dimensional vector spaces over a field F, the canonical map associated with a quotient is the projection from a vector space V onto the quotient space V/U, where U \subseteq V is a subspace. Specifically, the canonical projection \pi: V \to V/U is the linear surjection defined by \pi(v) = v + U for each v \in V, which identifies elements differing by vectors in U. This map is surjective by construction, as every coset v + U is the image of v, and its kernel is precisely U. A fundamental consequence is the dimension theorem for quotient spaces, which states that \dim(V/U) = \dim(V) - \dim(U). This follows from the rank-nullity theorem applied to \pi, where \dim(\ker \pi) = \dim(U) and \dim(\operatorname{im} \pi) = \dim(V/U), so \dim(V) = \dim(U) + \dim(V/U). To construct a basis for V/U, suppose \{u_1, \dots, u_k\} is a basis for U; extend it to a basis \{u_1, \dots, u_k, v_1, \dots, v_m\} for V. Then \{v_1 + U, \dots, v_m + U\} forms a basis for V/U, as these cosets are linearly independent (any \sum \lambda_j (v_j + U) = U implies \sum \lambda_j v_j \in U, hence zero by basis extension) and V/U (every coset is a of the v_j + U). This aligns with the dimension formula, confirming m = \dim(V) - \dim(U). The canonical projection satisfies a universal property in the category of vector spaces: for any linear map f: V \to W such that U \subseteq \ker f, there exists a unique linear map \overline{f}: V/U \to W with f = \overline{f} \circ \pi. This means linear maps from V that factor through the subspace U correspond uniquely to linear maps from the quotient V/U, providing a categorical characterization of the construction. As a special case, this specializes the module projections discussed earlier when the scalar ring is a field. For a concrete example, consider V = \mathbb{R}^2 and U = \operatorname{[span](/page/Span)}\{(1,0)\}, the x-axis. The canonical projection \pi: \mathbb{R}^2 \to \mathbb{R}^2 / U sends (x, y) \mapsto (x, y) + U, which identifies points differing only in the x-coordinate, effectively yielding cosets represented by (0, y). Thus, \mathbb{R}^2 / U \cong \mathbb{R} with basis \{(0,1) + U\}, and \dim(\mathbb{R}^2 / U) = 2 - 1 = 1, illustrating the surjection onto the y-direction.

Dual Space Evaluations

In the context of dual vector spaces, the evaluation map provides a fundamental canonical construction. For a vector space V over a field F, the evaluation map \mathrm{ev}: V \times V^* \to F is defined by \mathrm{ev}(v, \phi) = \phi(v) for all v \in V and \phi \in V^*, where V^* denotes the dual space of all linear functionals from V to F. This bilinear map induces a canonical linear map \iota: V \to V^{**}, where V^{**} is the double dual space, given by \iota(v) = \mathrm{ev}_v and \mathrm{ev}_v(\phi) = \phi(v) for \phi \in V^*. The map \iota is injective for any vector space V and serves as the standard embedding of V into its double dual. When V is finite-dimensional, the canonical map \iota: V \to V^{**} becomes an , establishing V \cong V^{**}. Specifically, for \dim V = n < \infty, the inverse map sends an element of V^{**} back to the unique vector in V via the evaluation structure, preserving the vector space operations without requiring a choice of basis. This isomorphism highlights the reflexive nature of finite-dimensional spaces under duality. Another key canonical map arises in the context of linear transformations between dual spaces: the transpose map. Given a linear map T: V \to W between vector spaces over F, the transpose T^*: W^* \to V^* is defined by (T^* \psi)(v) = \psi(T v) for all \psi \in W^* and v \in V. This construction is linear and, in the finite-dimensional case with matrix representations, corresponds to the transpose of the matrix of T. The transpose map satisfies a universal property through its naturality: for composable linear maps S: W \to U and T: V \to W, the diagram commutes via (S \circ T)^* = T^* \circ S^*, ensuring the construction is functorial and independent of bases. Additionally, for any finite-dimensional vector space V, the dimension of the dual space equals that of the original: \dim V^* = \dim V. This equality extends to the double dual, reinforcing the isomorphism V \cong V^{**}.

Canonical Maps in Category Theory

Natural Transformations

In category theory, canonical maps frequently appear as the components of natural transformations between functors. A natural transformation \eta: F \Rightarrow G between parallel functors F, G: \mathcal{C} \to \mathcal{D} assigns to each object X in the category \mathcal{C} a morphism \eta_X: F(X) \to G(X) in \mathcal{D}, such that these components are natural in X. This naturality requires that for every morphism f: X \to Y in \mathcal{C}, the following square commutes: \begin{CD} F(X) @>{F(f)}>> F(Y)\\ @V{\eta_X}VV @VV{\eta_Y}V \\ G(X) @>>{G(f)}> G(Y) \end{CD} Equivalently, the components satisfy the condition G(f) \circ \eta_X = \eta_Y \circ F(f) for all such f. This structure ensures that the maps \eta_X respect the morphisms in \mathcal{C}, making them canonical in the sense that they are determined uniformly across the category without arbitrary choices. Prominent canonical examples of natural transformations include the identity transformation \mathrm{Id}_F: F \Rightarrow F, where each component is \eta_X = \mathrm{id}_{F(X)}, which satisfies naturality by direct substitution into the condition. Another standard example arises from the composition of functors: if H: \mathcal{D} \to \mathcal{E} is a functor, then the transformation H \circ F \Rightarrow H \circ G induced by \eta via H(\eta_X): H(F(X)) \to H(G(X)) is again natural, preserving the categorical structure through functoriality. These examples illustrate how natural transformations provide a canonical way to relate functors, often serving as the mediating morphisms in larger diagrams. The establishes a deep link between s and the data they encode, asserting a natural bijection \mathrm{Nat}(y_X, F) \cong F(X), where y_X = \hom_{\mathcal{C}}(-, X): \mathcal{C}^\mathrm{op} \to \mathrm{Set} is the representable functor and F: \mathcal{C}^\mathrm{op} \to \mathrm{Set} is any presheaf on \mathcal{C}. Under this isomorphism, each \eta: y_X \Rightarrow F corresponds uniquely to the element \eta_X(\mathrm{id}_X) \in F(X), showing that the components of such canonical transformations fully determine the functor F. This bijection underscores the representational power of representable functors, where the maps \eta_X canonically capture the object's "universal" properties through naturality. In many categorical contexts, the components \eta_X of a natural transformation are the unique morphisms that render specified diagrams commutative, arising from the requirement that the transformation be natural. This uniqueness often stems from the functorial action on identities and compositions, ensuring that no other family of maps can satisfy the naturality squares without violating the categorical axioms. Such canonical maps thus embody the , structure-preserving essence of transformations between functors.

Universal Property Maps

In , a universal arises from the of an object U equipped with s to or from it, where the map is the unique induced by this property that factors through any compatible in the category. For instance, consider the product object X \times Y in a category with products; the projections \pi_X: X \times Y \to X and \pi_Y: X \times Y \to Y satisfy the universal property that for any object Z and s f: Z \to X, g: Z \to Y, there exists a unique h: Z \to X \times Y such that \pi_X \circ h = f and \pi_Y \circ h = g. This uniqueness ensures that the projections are , characterizing the product up to . A prominent example of such a canonical map is the unique from an initial object. An initial object $0 in a is one from which there exists exactly one ! : 0 \to A to every object A; this ! is thus the canonical map induced by the initiality , serving as the starting point for all constructions in the . For colimits, canonical maps often take the form of inclusions into the colimit object. In the case of a A \sqcup B, the canonical injections i_A: A \to A \sqcup B and i_B: B \to A \sqcup B embody the universal property that for any object C and s f: A \to C, g: B \to C, there is a unique h: A \sqcup B \to C satisfying h \circ i_A = f and h \circ i_B = g. Similarly, for a of parallel s f, g: A \rightrightarrows B, the canonical map q: B \to B / \sim (where \sim is the generated by f(a) \sim g(a)) is universal in that any k: B \to C with k \circ f = k \circ g factors uniquely through q. Canonical maps also emerge from pullbacks and pushouts, which are for specific commutative diagrams. In a , given morphisms p: E \to B and f: A \to B, the object P comes with projections \pi_A: P \to A and \pi_E: P \to E such that the square P \to A \to B \leftarrow E \leftarrow P commutes, and any other object mediating compatible morphisms factors uniquely through these projections. Dually, in a pushout of i: A \to B and j: A \to C, the inclusions k_B: B \to B \sqcup_A C and k_C: C \to B \sqcup_A C ensure that any pair of morphisms from B and C compatible on A factors uniquely through the pushout object. In the context of adjoint functors, the unit and counit provide canonical natural transformations defined by the adjunction's . For functors F: \mathcal{C} \to \mathcal{D} and G: \mathcal{D} \to \mathcal{C} with F \dashv G, the unit \eta: \mathrm{Id}_\mathcal{C} \Rightarrow G F assigns to each object C in \mathcal{C} a \eta_C: C \to G F(C), while the counit \varepsilon: F G \Rightarrow \mathrm{Id}_\mathcal{D} assigns \varepsilon_D: F G(D) \to D for each D in \mathcal{D}; these satisfy the triangular identities and induce the natural \mathrm{Hom}_\mathcal{D}(F(C), D) \cong \mathrm{Hom}_\mathcal{C}(C, G(D)).

Illustrative Examples

Projection and Inclusion Variants

In , the canonical projection map \pi: X \times Y \to X is defined by \pi(x, y) = x for all (x, y) \in X \times Y. Together with the projection \pi_Y: X \times Y \to Y defined by \pi_Y(x, y) = y, these maps satisfy the universal property of the product: for any set Z and any functions f: Z \to X, h: Z \to Y, there exists a unique function g: Z \to X \times Y such that \pi \circ g = f and \pi_Y \circ g = h, given by g(z) = (f(z), h(z)). In group theory, the canonical inclusion map i: \{e\} \to G from the \{e\} (where e is the ) to a group G sends the e to the identity of G. This map is the unique from the to G, as the trivial group serves as the initial object in the category of groups. For modules over a R, consider the \bigoplus_{i \in I} M_i. The canonical inclusions \iota_j: M_j \to \bigoplus_{i \in I} M_i embed M_j by placing the element in the j-th position with zeros elsewhere. These satisfy the universal of the : for any R- N and family of R-s f_i: M_i \to N, there is a unique R- f: \bigoplus_{i \in I} M_i \to N such that f \circ \iota_k = f_k for all k, defined by on finite sums. The canonical projections \pi_j: \bigoplus_{i \in I} M_i \to M_j are defined by \pi_j((m_i)_{i \in I}) = m_j, and satisfy \pi_j \circ \iota_k = \delta_{jk} \mathrm{id}_{M_j}. For finite I, the direct sum is a , with joint universal properties for both inclusions and projections. Each of these maps verifies its canonicity through a that ensures it factors through appropriate morphisms while being constant on : for the set , are singletons times Y; for the group inclusion, the "fiber" is the single point \{e\}; and for the , consist of elements with zero in the j-th slot. As a simple computation, take X = \{1, 2\} and Y = \{a, b\}. Then \pi(1, a) = 1, \pi(1, b) = 1, \pi(2, a) = 2, and \pi(2, b) = 2, with image \{1, 2\} = X.

Functorial Isomorphisms

In , a functorial isomorphism, more precisely termed a natural isomorphism, consists of a family of isomorphisms between the values of two functors F, G: \mathcal{C} \to \mathcal{D} on objects of \mathcal{C}, such that the components form a , meaning they commute with all morphisms in \mathcal{C}. This structure ensures the isomorphisms are canonical, as they arise uniquely from the definitions of the functors without arbitrary choices, respecting the categorical morphisms via commutative diagrams. Such isomorphisms exemplify canonical maps because their naturality guarantees independence from basis selections or other non-functorial constructions, providing a "" identification between functorial constructions. For instance, in the of finite-dimensional spaces over a , there is a natural V \cong V^{**} between a space and its double dual, given by the evaluation map v \mapsto ( \phi \mapsto \phi(v) ), which is functorial and thus canonical, commuting with all linear transformations. This contrasts with infinite-dimensional cases, where only a canonical injection exists. Another illustrative example occurs in abelian categories, where the direct sum functor induces a natural \mathrm{Hom}(A \oplus B, C) \cong \mathrm{Hom}(A, C) \oplus \mathrm{Hom}(B, C), with components defined by the universal property of the direct sum: the map sends a f: A \oplus B \to C to (f \circ i_A, f \circ i_B), where i_A, i_B are the inclusions. This is functorial in all variables A, B, C, making it a canonical tool for decomposing hom-spaces without selecting bases. These functorial isomorphisms underpin many canonical maps in higher structures, such as equivalences of categories, where pairs of functors are up to isomorphism, ensuring the categories are "essentially the same" in a rigorous, choice-free manner. Seminal work formalized this framework, emphasizing that equivalences preserve all structural relations inherently.

References

  1. [1]
    [PDF] Canonical maps - CORE
    In the mathematical vernacular, a canonical map f is usually said to be a map defined without any arbitrary decision. What can this possibly mean? It suggests ...
  2. [2]
    [PDF] Category theory Supplemental notes 1
    Feb 10, 2017 · A universal property is a uniform description of morphisms into or out of an object, and it characterizes an object uniquely up to isomorphism.
  3. [3]
    Canonical Map -- from Wolfram MathWorld
    A function f mapping a set X->X/R (X modulo R), where R is an equivalence relation in X, is called a canonical map.
  4. [4]
    Canonical Map - an overview | ScienceDirect Topics
    A canonical map is defined as a function that assigns each element of a group to its corresponding left coset in a quotient structure, facilitating the ...
  5. [5]
    [PDF] The Logical Theory of Canonical Maps: The Elements & Distinctions ...
    Oct 4, 2024 · For instance, given a cone f : Z → X and g : Z → Y , the canonical map Z → X × Y is only canonical relative to the data f and g. The ...
  6. [6]
    [PDF] Basic Category Theory
    Feb 5, 2023 · a limit, and the canonical map. F lim. ←I. D. → lim. ←I. (F ◦ D) is an isomorphism. Here the 'canonical map' has I-component. F lim. ←I. D. F(pI ).
  7. [7]
    soft question - What is the definition of "canonical"? - MathOverflow
    Mar 28, 2010 · E.g. the Normal n by n matrices over C is sufficiently structured such that there is a canonical map to Cn modulo the action of the ...The meaning and purpose of "canonical'' - MathOverflowWhat, precisely, do we mean when we say that a f.d. vector space is ...More results from mathoverflow.net
  8. [8]
    [PDF] Naive set theory. - Whitman People
    then the induced relation is an equivalence relation whose set of equivalence classes is exactly 6. For the proof, let us start with an equivalence relation R.Missing: quotient | Show results with:quotient
  9. [9]
    Definition: Canonical Projection - BookOfProofs
    Definition: Canonical Projection. Let V be a set, R\subseteq V\times V an equivalence relation and let V/_R be the corresponding quotient set. The function q ...
  10. [10]
    None
    ### Definitions and Properties of Canonical Projection for Quotient Sets
  11. [11]
    [PDF] A rough guide to linear algbera - Stanford University
    Exercise 1.4.B (Universal property for quotients). Let X be a set with an equivalence relation R, and let f : X → Y be a function satisfying f(x1) = f(x2).
  12. [12]
    [PDF] Notes on Sets, Mappings, and Cardinality - Berkeley Math
    Oct 18, 2010 · Mapping (6) is called the canonical quotient mapping. 1.2.4 Example: the evaluation map. For any sets X and Y, mappings f : X−→Y form a set, ...
  13. [13]
  14. [14]
    None
    Below is a merged summary of quotient groups in group theory based on the provided segments from J.S. Milne's *Group Theory* (v4.01), sourced from https://www.jmilne.org/math/CourseNotes/GT.pdf. The information is consolidated into a dense, structured format using a table to retain all details efficiently, followed by a narrative summary for clarity. The table captures key concepts, definitions, properties, and page references across the segments, while the narrative integrates the universal property and theorems with additional context.
  15. [15]
    [PDF] "Abstract Algebra: Theory and Applications"
    Aug 11, 2012 · This text is intended for a one- or two-semester undergraduate course in abstract algebra. Traditionally, these courses have covered the ...
  16. [16]
    [PDF] Module Theory - The University of Memphis
    If N ≤ M are R-modules, the quotient module M/N is an R-module such that. (M/N, +) is the usual quotient group of (M,+) by (N,+) (since M is abelian, N is.
  17. [17]
    [PDF] Modules
    The map π : A → A/B given by a 7→ a + B is an R-module epimorphism with kernel B (called canonical epimorphism or projection). Page 4. 30. CHAPTER 2. MODULES.
  18. [18]
    [PDF] representation theory, chapter 0. basics - Yale Math
    The following lemma states the universal property of a quotient module. Note ... Every irreducible A-module is isomorphic to a quotient module of the regular.
  19. [19]
    [PDF] ring theory - Northwestern University
    Examples: The canonical epimorphism Z → Z/nZ is a ring homomorphism. ... A ring is called a principal ideal ring if it is a commutative ring and every ideal is ...
  20. [20]
    [PDF] Module Fundamentals
    The canonical map π: M → M/N is a module homomorphism with kernel N. Just as with groups and rings, we can establish the basic isomorphism theorems for modules.
  21. [21]
    [PDF] 1. Cosets and the Quotient Space - Christopher Heil
    QUOTIENT SPACES. 2. The Canonical Projection. Definition 2.1. If M is a subspace of a vector space X, then the canonical projection or the canonical mapping ...
  22. [22]
    [PDF] Quotient Spaces
    Given a subspace S of a vector space V , the linear transformation pS : V → V/S defined above is often referrred to as the canonical projection (of V onto V/S).
  23. [23]
    [PDF] NOTES ON QUOTIENT SPACES Let V be a vector ... - Academic Web
    In particular, since we know what the dimension of a quotient space is, one can use this theorem (or more precisely the consequence that dim(V/nullT) = ...
  24. [24]
    [PDF] Quotient Spaces - Cornell University
    Sep 6, 2019 · We finally consider the special role that this linear transformation p : V −→ V/W plays for quotient spaces. Theorem 2 (Universal Mapping ...
  25. [25]
    [PDF] Math 396. Quotient spaces 1. Definition Let F be a field, V a vector ...
    Definition. Let F be a field, V a vector space over F and W ⊆ V a subspace of V . For v1,v2 ∈ V , we say that v1 ≡ v2 mod W if and only if v1 − v2 ∈ W. One ...
  26. [26]
    [PDF] Notes on Dual Spaces
    Now we look at maps between dual spaces. Definition 3. Let T : V → W be linear. The dual map (or transpose) of T is the map. T∗ : W∗ → V ∗ defined by. T.
  27. [27]
    [PDF] Chapter 8 The Dual Space, Duality - UPenn CIS
    Given a vector space E, the vector space Hom(E,K) of linear maps from E to K is called the dual space (or dual) of E. The space Hom(E,K) is also denoted by E⇤, ...Missing: inclusion | Show results with:inclusion
  28. [28]
    [PDF] Chapter 4 Dual Spaces
    Let V be a vector space over a field F . Then the canonical map τ is a monomorphism. Furthermore, if V is finite-dimensional, then τ is an isomorphism.
  29. [29]
    [PDF] Lecture 2.5: The transpose of a linear map
    Given a linear map T : X → U, its transpose is a certain induced linear map T0 : U0 → X0 between the dual spaces. In the next lecture, we will learn how ...
  30. [30]
    Yoneda lemma in nLab
    ### Summary of Yoneda Lemma Statement and References
  31. [31]
    [PDF] maclane-categories.pdf - MIT Mathematics
    ... Categories for the working mathematician/Saunders Mac Lane. -. 2nd ed. p. cm ... definition of a group by arrows J1, 1'/, and ( in such commutative.
  32. [32]
    [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.
  33. [33]
    [PDF] Category Theory - GitHub Pages
    In this case, we say α is a natural isomorphism. We can then say ... is exact at An if fn+1◦fn = 0 and the canonical map Im fn → Kerfn+1 is an isomorphism.
  34. [34]
    [PDF] Universal mapping properties - Keith Conrad
    Universal mapping property means a new object is related to all similar objects by mappings, like a cyclic n-marked group (C, c) to every n-marked group (G, γ) ...
  35. [35]
    GENERAL THEORY OF NATURAL EQUIVALENCES
    GENERAL THEORY OF NATURAL EQUIVALENCES. 233. A discussion of the "simultaneous ... homology theory (Eilenberg and MacLane, Group extensions and homology, Ann.