Fact-checked by Grok 2 weeks ago

Homotopy theory

Homotopy theory is a branch of that studies topological spaces and continuous maps between them up to homotopy equivalence, where two maps f, g: X \to Y are considered equivalent if one can be continuously deformed into the other via a , a continuous map H: X \times I \to Y with I = [0,1] such that H(x,0) = f(x) and H(x,1) = g(x). This captures the idea of spaces being deformable into each other without tearing or gluing, focusing on invariants like and holes preserved under such deformations. Central to homotopy theory are the homotopy groups \pi_n(X, x_0), which generalize the \pi_1(X, x_0) to higher dimensions; \pi_1 measures one-dimensional holes via homotopy classes of loops based at x_0 \in X, while higher \pi_n for n \geq 2 use maps from the n- S^n and are abelian and harder to compute. For example, \pi_1(S^1) \cong \mathbb{Z}, \pi_1(S^n) = 0 for n \geq 2, and \pi_n(S^n) \cong \mathbb{Z}, reflecting the sphere's simplest higher-dimensional structure. equivalences induce isomorphisms on these groups, and tools like CW complexes allow approximation of maps by cellular ones, enabling computations via skeletons. Historically, homotopy theory emerged from Henri Poincaré's early work on in the late 19th century, with key developments including Luitzen Brouwer's fixed-point theorems around 1910 and Witold Hurewicz's introduction of higher groups in the 1930s. Post-World War II advancements by , , and formalized aspects like fibrations, Eilenberg-MacLane spaces, and the stable category, linking to and for broader applications in and physics. Modern extensions include rational theory, which localizes at the rationals to simplify computations, and , integrating homotopical ideas into foundational mathematics.

Foundational Concepts

Topological Spaces and Continuous Maps

A is a set X together with a collection \mathcal{T} of subsets of X, called open sets, that satisfies three axioms: the \emptyset and X itself are open; arbitrary unions of open sets are open; and finite intersections of open sets are open. This structure generalizes metric spaces like the real line, allowing the study of and deformation without relying on distances. Common examples include Euclidean spaces \mathbb{R}^n, equipped with the standard topology generated by open ; the n-spheres S^n, defined as the of the (n+1)- in \mathbb{R}^{n+1} with the ; and manifolds, which are spaces locally homeomorphic to \mathbb{R}^n, such as the obtained by identifying opposite sides of . These examples illustrate how topologies capture intuitive notions of nearness and openness in geometric objects. A continuous map between topological spaces X and Y is a f: X \to Y such that the preimage f^{-1}(U) of every U \subseteq Y is open in X. This definition ensures that continuous maps preserve the topological structure, making them the natural morphisms in the . Properties like —where every open cover has a finite subcover—and connectedness—where the space cannot be written as a of two nonempty open sets—are preserved under continuous maps and play key roles in homotopy theory by controlling the behavior of deformations on bounded or indivisible spaces. For instance, spheres S^n are compact and connected, while \mathbb{R}^n is connected but not compact. Homotopy theory motivates the study of topological spaces and continuous maps up to continuous deformation, viewing homotopies as paths in the space of continuous maps between spaces. Basic examples highlight this: \mathbb{R}^n is contractible, meaning it admits a continuous deformation to a single point, whereas S^1 is non-contractible, resisting such a deformation due to its looped structure. These distinctions underscore why homotopy theory seeks invariants that remain unchanged under deformation.

Homotopy of Maps

In homotopy theory, the central notion of homotopy provides a way to deform one continuous map into another while preserving topological structure. Given two continuous maps f, g: X \to Y between topological spaces X and Y, a homotopy from f to g is a continuous map H: X \times I \to Y, where I = [0, 1] denotes the unit interval, such that H(x, 0) = f(x) and H(x, 1) = g(x) for all x \in X. This construction, often visualized as "stretching a rubber sheet," captures deformations that ignore rigid geometric constraints. Equivalently, a homotopy can be viewed as a continuous path in the function space Y^X connecting f and g, parametrized by t \in I. The domain X \times I, known as the cylinder over X, geometrically represents the product of the source space with the interval, where the "bottom" X \times \{0\} maps via f and the "top" X \times \{1\} via g. Homotopy induces an on the set of all continuous maps \mathrm{Map}(X, Y). Specifically, reflexivity holds via the constant H(x, t) = f(x); via reversal of the t \mapsto 1 - t; and transitivity by concatenating homotopies along subdivided intervals. The resulting equivalence classes, denoted , partition \mathrm{Map}(X, Y) into homotopy classes [X, Y], which ignore distinctions arising from continuous deformations. This relation generates the homotopy category \mathbf{Ho}(\mathbf{Top}), where objects are topological spaces and morphisms are classes of maps, foundational for studying shape up to deformation. Illustrative examples highlight the concept's role in distinguishing essential features. For paths in Euclidean space \mathbb{R}^n sharing endpoints, the linear homotopy H(s, t) = (1 - t) f(s) + t g(s), \quad s, t \in I, continuously interpolates between them. On spheres, consider the equatorial inclusion i: S^1 \hookrightarrow S^2, which embeds the equator; this map is homotopic to a constant map via radial contraction from the north pole, shrinking the equator to a point within the higher-dimensional sphere. More generally, any continuous map S^m \to S^n with m < n is homotopic to a constant map, as the source can be embedded and contracted in the target's extra dimensions. These examples underscore how homotopy detects "nullhomotopic" maps, those deformable to constants, revealing trivial deformations. Homotopies are classified as free or relative depending on constraints. A free homotopy imposes no restrictions beyond continuity, allowing full deformation of the entire map. In contrast, a relative homotopy with respect to a subspace A \subset X requires H(a, t) = f(a) = g(a) for all a \in A and t \in I, fixing the behavior on A during deformation; this is crucial for basepoint-preserving maps. The cylinder construction facilitates both, as relative homotopies correspond to maps constant on A \times I.

Homotopy Equivalence and Deformation Retracts

In algebraic topology, two topological spaces X and Y are considered homotopy equivalent if there exist continuous maps f: X \to Y and g: Y \to X such that the compositions f \circ g and g \circ f are each homotopic to the respective identity maps \mathrm{id}_Y and \mathrm{id}_X. This notion captures the idea that X and Y have the same "shape" up to continuous deformation, preserving essential topological features while allowing for stretching and bending without tearing. Homotopy equivalence is an equivalence relation on the category of topological spaces, partitioning them into classes of spaces with identical homotopy types. A related but weaker concept is that of a weak homotopy equivalence, which is a continuous map f: X \to Y that induces isomorphisms on the homotopy groups \pi_n(X, x_0) \cong \pi_n(Y, f(x_0)) for all basepoints x_0 \in X and all n \geq 0. While every homotopy equivalence is a weak homotopy equivalence, the converse does not hold in general; stronger conditions, such as the spaces being , are often needed to ensure that weak equivalences imply full homotopy equivalences. This distinction highlights the role of weak equivalences in and simplicial homotopy, where they serve as the primary notion of equivalence without requiring strict deformation. Deformation retracts provide a concrete way to understand homotopy equivalences within a single space. A subspace A \subseteq X is a deformation retract of X if there exists a continuous retraction r: X \to A (a map fixing A pointwise) such that the identity map \mathrm{id}_X is homotopic to r via a homotopy H: X \times I \to X that restricts to the identity on A \times I. In this case, the inclusion i: A \hookrightarrow X is a homotopy equivalence with homotopy inverse r, and X and A share the same homotopy type. This structure allows for simplifying spaces by retracting "inessential" parts while preserving homotopy invariants. A classic example is the annulus A = \{ (x,y) \in \mathbb{R}^2 \mid 1 \leq x^2 + y^2 \leq 4 \} deformation retracting onto its inner boundary circle S^1 = \{ (x,y) \in \mathbb{R}^2 \mid x^2 + y^2 = 1 \}. The radial retraction r: A \to S^1 given by r(x,y) = (x,y)/\sqrt{x^2 + y^2} (with r fixing S^1) is homotopic to \mathrm{id}_A via the straight-line homotopy H((x,y), t) = (1-t)(x,y) + t \cdot r(x,y), which remains fixed on S^1. This illustrates how a space like the annulus A, homotopy equivalent to S^1, can deformation retract onto its inner boundary circle, simplifying the space while preserving its homotopy type. In the context of cell complexes, deformation retracts facilitate computing homotopy types by successively retracting cells onto their boundaries or skeletons, simplifying the space while maintaining equivalence. For instance, attaching a 2-cell to S^1 along a constant map yields a space homotopy equivalent to S^1 \vee S^2, where the deformation retract identifies the cell's interior as collapsible without altering the overall type. Such computations underpin the study of homotopy in more complex structures.

Core Constructions

CW Complexes

CW complexes are topological spaces constructed inductively by attaching open cells of successively increasing dimensions, providing a cellular filtration that controls their homotopy type. Introduced by , a CW complex X consists of a collection of cells e_\alpha^n for \alpha in some index set and n \geq 0, where each n-cell is the open n-disk D^n, attached to the (n-1)-skeleton via a characteristic map \phi_\alpha: S^{n-1} \to X^{n-1} that is continuous. The topology on X is the quotient topology obtained from the disjoint union of all cells, with the weak topology condition ensuring that the closure of each cell intersects the previous skeleton in a closed set. The n-skeleton X^n of a CW complex X is the union of all cells of dimension at most n, forming a subspace that inherits the CW structure up to dimension n. This decomposition allows X to be built as the direct limit \varinjlim X^n, where each X^n is compact if finitely many cells are attached at each stage. The cellular chain complex derived from this skeleton computes the homology of X, with boundary maps induced by the degrees of the attaching maps on the boundaries of cells. A fundamental result is the cellular approximation theorem, which states that any continuous map f: X \to Y between CW complexes is homotopic to a cellular map g: X \to Y, meaning g restricts to a map X^n \to Y^n for each n. This theorem implies that homotopies between maps of CW complexes can also be approximated by cellular homotopies, facilitating computations by reducing to skeletal maps. Moreover, every space is weakly homotopy equivalent to a CW complex, allowing the approximation of arbitrary spaces by cellular ones for homotopy purposes. Classic examples of CW complexes include spheres, where S^n has a minimal CW structure consisting of one 0-cell and one n-cell, with the attaching map being the constant map to the base point. Real projective spaces \mathbb{RP}^n admit a CW structure with exactly one cell in each dimension from 0 to n, where the k-cell attaches via the double-covering map S^{k-1} \to \mathbb{RP}^{k-1}. Moore spaces M(\mathbb{Z}/m\mathbb{Z}, n), which realize a single nontrivial homology group H_n \cong \mathbb{Z}/m\mathbb{Z}, are constructed as S^n with an (n+1)-cell attached by a map of degree m. These constructions highlight how CW complexes enable the realization of specific algebraic invariants through controlled cellular attachments. Homotopy equivalences between CW complexes often simplify their cellular structures, allowing minimal models with fewer cells while preserving homotopy type.

Fibrations and Cofibrations

In homotopy theory, fibrations and cofibrations are classes of maps between topological spaces characterized by specific lifting and extension properties that facilitate inductive arguments and the construction of homotopy invariants. These notions, dual to each other, play a central role in organizing the category of spaces up to homotopy equivalence. A Serre fibration is a continuous map p: E \to B that satisfies the homotopy lifting property with respect to the inclusions of the form i_0: X \to X \times I for any topological space X, where I = [0,1] is the unit interval. Specifically, given a map f: X \to E and a homotopy H: X \times I \to B such that H(x,0) = p(f(x)) for all x \in X, there exists a lift \tilde{H}: X \times I \to E with \tilde{H}(x,0) = f(x) and p \circ \tilde{H} = H. This property was introduced by in his work on singular homology of fibrations. An equivalent characterization is that p has the right lifting property against the maps D^n \to S^{n-1} \times I \cup D^n \times \{0\} for disks D^n and spheres S^{n-1}, ensuring compatibility with cellular decompositions. Serre fibrations also admit a path-lifting property: for any path \gamma: I \to B and point e \in p^{-1}(\gamma(0)), there exists a lift \tilde{\gamma}: I \to E with \tilde{\gamma}(0) = e and p \circ \tilde{\gamma} = \gamma. Dually, a cofibration is a continuous map i: A \to X that satisfies the homotopy extension property with respect to all topological spaces Y. That is, given a map g: X \to Y and a homotopy H: A \times I \to Y such that H(a,0) = g(i(a)) for all a \in A, there exists an extension \tilde{H}: X \times I \to Y with \tilde{H}|_{A \times I} = H and \tilde{H}(x,0) = g(x) for all x \in X. This definition ensures that homotopies defined on the subspace A can be extended over the whole space X while preserving the initial conditions. Cofibrations often arise as closed inclusions of retracts, and they are preserved under pushouts and composition. Prominent examples of Serre fibrations include principal G-bundles for a topological group G, which are locally trivial fiber bundles with fiber G and projection map satisfying the homotopy lifting property due to their local product structure. The canonical example is the Hopf fibration S^1 \to S^3 \to S^2, a principal S^1-bundle that is a Serre fibration. For cofibrations, cell attachments in CW complexes provide concrete instances: the inclusion of the attaching sphere S^{n-1} \to D^n is a cofibration, and more generally, the inclusion of a CW subcomplex into the full complex inherits the homotopy extension property from the cellular structure. A simple case is the inclusion of a point into any well-pointed space, which extends homotopies trivially. Fibrations and cofibrations, together with weak homotopy equivalences (maps inducing isomorphisms on all homotopy groups), form the structure of a model category on the category of topological spaces, enabling the localization to the homotopy category where morphisms are homotopy classes of maps. In this framework, every map factors as a cofibration followed by a weak equivalence followed by a fibration, allowing systematic computation of homotopy types via resolutions.

Lifting Properties

In category theory, particularly within the context of homotopy theory, the lifting properties formalize the conditions under which maps can be extended or lifted in commutative diagrams, providing an axiomatic foundation for homotopy extensions and the structure of model categories. These properties are dual in nature and are essential for defining classes of morphisms such as fibrations and cofibrations. These properties are dual in nature and are essential for defining classes of morphisms such as fibrations and cofibrations. The right lifting property (RLP) for a morphism p \colon E \to B with respect to another morphism i \colon A \to X asserts the existence of a lift in any commutative square formed by these maps. Specifically, given maps f \colon A \to E and g \colon X \to B such that p \circ f = g \circ i, there exists a map h \colon X \to E making both triangles commute, i.e., h \circ i = f and p \circ h = g. In homotopy theory, this is often visualized in a pushout square diagram where the domain of i involves an interval, such as the inclusion \{0,1\} \hookrightarrow I (the two endpoints into the unit interval), representing a basic ; fibrations prototypically satisfy the RLP with respect to such acyclic , enabling path lifting. For instance, in the category of topological spaces, exhibit this property, allowing homotopies in the base to lift to the total space. Dually, the left lifting property (LLP) for a morphism i \colon A \to X with respect to p \colon E \to B requires a lift h \colon X \to E in the same commutative square setup, ensuring h \circ i = f and p \circ h = g. This dual lifting criterion underpins the definition of , which satisfy the LLP against ; a topological example is the closed inclusion of a subspace, which extends maps over contractible spaces like the interval projection I \twoheadrightarrow \{0,1\}. These properties together facilitate the small object argument and factorization in . Trivial, or acyclic, fibrations are those morphisms that satisfy the RLP with respect to all cofibrations, while trivial cofibrations satisfy the LLP with respect to all fibrations. In the Quillen model structure on topological spaces, for example, acyclic fibrations include homotopy equivalences that are also Serre fibrations, lifting over any cofibration to preserve weak equivalences. Similarly, acyclic cofibrations are weak equivalences that are cofibrations, ensuring the LLP against fibrations. These notions extend the basic lifting criteria to capture homotopy-invariant extensions essential for deriving the homotopy category.

Homotopy Invariants

Fundamental Group

The fundamental group provides an algebraic invariant that detects one-dimensional holes in a topological space, encoding information about loops that cannot be continuously contracted to a point. For a pointed topological space (X, x_0), the fundamental group \pi_1(X, x_0) is defined as the set of homotopy classes of based loops in X starting and ending at x_0, where a based loop is a continuous map \gamma: S^1 \to X with \gamma(1) = x_0. This concept was introduced by in his seminal 1895 paper "Analysis Situs," where he recognized its role in classifying multiply connected surfaces. The group operation on \pi_1(X, x_0) arises from the concatenation of loops: for two homotopy classes [\gamma] and [\delta], the product [\gamma] \cdot [\delta] is represented by the loop that traverses \gamma followed by \delta, adjusted at the basepoint via a path reparameterization. This operation is associative, with the constant loop serving as the identity and inverses given by loops traversed in the opposite direction, making \pi_1(X, x_0) a group (non-abelian in general). If X is path-connected, then \pi_1(X, x_0) is isomorphic to \pi_1(X, x_1) for any other basepoint x_1 \in X, via conjugation by a path from x_0 to x_1. A key example is the circle S^1, where \pi_1(S^1, 1) \cong \mathbb{Z}, generated by the class of the identity map S^1 \to S^1, which winds once around the circle; integer powers correspond to multiple windings, capturing the single one-dimensional hole in S^1. More generally, the wedge sum of n circles (a space formed by identifying the basepoints of n disjoint copies of S^1) has fundamental group isomorphic to the free group on n generators, reflecting n independent one-dimensional holes that can be traversed without algebraic relations. These examples illustrate how \pi_1 distinguishes spaces with different loop structures, such as the torus T^2, whose \pi_1(T^2) \cong \mathbb{Z} \oplus \mathbb{Z} arises from two commuting generators corresponding to meridional and longitudinal loops. Covering space theory offers a geometric realization of the fundamental group. For a path-connected, locally path-connected space X, the universal covering space \tilde{X} \to X is a simply connected covering (i.e., \pi_1(\tilde{X}) = 0) that is unique up to covering isomorphism, and the fundamental group \pi_1(X, x_0) is isomorphic to the group of deck transformations of this cover—homeomorphisms of \tilde{X} that commute with the projection to X. In general, every covering space p: Y \to X corresponds to a subgroup N = p_*(\pi_1(Y, y_0)) of \pi_1(X, x_0). If N is normal (i.e., the covering is regular), the deck transformation group is isomorphic to the quotient \pi_1(X, x_0)/N; trivial fundamental group implies the space is simply connected and admits a universal cover homeomorphic to itself. The Seifert–van Kampen theorem facilitates explicit computations of fundamental groups for spaces constructed as pushouts. For a pushout X = U \cup V where U and V are path-connected open subsets of X with path-connected intersection U \cap V, the theorem states that \pi_1(X, x_0) is the free product \pi_1(U, u_0) * \pi_1(V, v_0) amalgamated over the image of \pi_1(U \cap V, w_0) under the inclusions into U and V, assuming x_0 \in U \cap V. This result, originally proved by Egbert van Kampen in 1933, allows decomposition of complex spaces into simpler pieces while preserving the algebraic structure of loops across gluings.

Higher Homotopy Groups

Higher homotopy groups generalize the fundamental group to higher dimensions, capturing information about the "holes" in a topological space X that maps from spheres of dimension n \geq 2 can detect. For a pointed topological space (X, x_0), the nth homotopy group \pi_n(X, x_0) is defined as the set of based homotopy classes of continuous maps f: (S^n, *) \to (X, x_0), where S^n is the n-sphere with a distinguished basepoint *, and two maps are homotopic if there exists a continuous homotopy between them that preserves the basepoint throughout. Equivalently, \pi_n(X, x_0) can be described using the n-cube: it consists of homotopy classes of maps (I^n, \partial I^n) \to (X, x_0), where I^n = [0,1]^n and \partial I^n is its boundary, with the homotopy fixing the boundary point to x_0. This group operation arises from concatenating maps along the equator of the sphere (or face of the cube), and for n \geq 2, \pi_n(X, x_0) is always abelian, unlike the potentially non-abelian \pi_1(X, x_0). The based homotopy groups \pi_n(X, x_0) depend on the choice of basepoint, but for path-connected spaces, changing the basepoint along a path \gamma induces an isomorphism \beta_\gamma: \pi_n(X, x_0) \to \pi_n(X, \gamma(1)). In contrast, free homotopy groups, denoted \pi_n(X), consist of unbased homotopy classes [S^n, X] without fixing a basepoint, which generally differ from the based versions unless X is simply connected. Moreover, the fundamental group \pi_1(X, x_0) acts on each higher homotopy group \pi_n(X, x_0) for n \geq 2 by conjugation: for a loop [\gamma] \in \pi_1(X, x_0) and \in \pi_n(X, x_0), the action is [\gamma] \cdot = [\tilde{\gamma} \circ f], where \tilde{\gamma} is a homotopy extending \gamma to the domain of f. This makes \pi_n(X, x_0) into a module over the group ring \mathbb{Z}[\pi_1(X, x_0)], encoding how lower-dimensional loops twist higher-dimensional spheres. A representative example is the homotopy groups of spheres: \pi_n(S^n) \cong \mathbb{Z}, generated by the identity map id: S^n \to S^n, which reflects the single n-dimensional hole in the n-sphere. More generally, the higher homotopy groups \pi_k(S^n) for k > n (unstable homotopy groups) are nontrivial and finitely generated abelian groups, but their computation is notoriously difficult, often requiring advanced tools like sequences; for instance, \pi_3(S^2) \cong \mathbb{Z}, generated by the map. One primary method for computing \pi_n(X) involves the long exact sequence of a Serre F \to E \to B, which states that there is an \cdots \to \pi_n(F) \to \pi_n(E) \to \pi_n(B) \to \pi_{n-1}(F) \to \pi_{n-1}(E) \to \pi_{n-1}(B) \to \cdots, allowing relative computations by relating the homotopy of the total space E to that of the base B and fiber F, for Serre fibrations. This sequence is exact at each term, meaning the image of one map equals the kernel of the next, and it terminates appropriately at low dimensions.

Loop Spaces and Suspension

The loop space \Omega X of a pointed topological space (X, x_0) consists of the set of all based loops in X, that is, continuous maps \gamma: (S^1, 1) \to (X, x_0), topologized by the compact-open topology. This construction, introduced by Hurewicz in his foundational work on higher homotopy groups, shifts the homotopy information of X to lower dimensions: specifically, \pi_n(\Omega X, \text{const}_{x_0}) \cong \pi_{n+1}(X, x_0) for all n \geq 0. Iterating the loop space functor yields models for higher homotopy groups, realizing \pi_{k+1}(X) as \pi_0 of the k-fold loop space \Omega^k X. The reduced suspension \Sigma X of a pointed space (X, x_0) is defined as the smash product S^1 \wedge X = (S^1 \times X)/(S^1 \vee X), where S^1 is with basepoint $1. This operation raises the dimension of homotopy classes, with the suspension inducing a homomorphism \pi_{n-1}(X, x_0) \to \pi_n(\Sigma X, \Sigma x_0), which is an isomorphism for nin certain ranges depending on the connectivity ofX$ (see Freudenthal suspension theorem). This reflects the dimension-shifting duality between looping and suspending via the adjunction. Loop spaces and suspensions are on the category of pointed topological spaces: the suspension \Sigma \dashv \Omega, meaning there is a natural of pointed classes [\Sigma X, Y]_* \cong [X, \Omega Y]_* for pointed spaces X and Y. This adjunction, a of unstable theory, preserves essential homotopy data and underpins many computations, such as relating maps into loop spaces to suspended sources. A representative example is the loop space of the circle: \Omega S^1 is equivalent to the \mathbb{Z}, where components correspond to winding numbers of loops around the basepoint. A related result is the double suspension theorem, which states that the double suspension of a homology sphere is a sphere, illustrating stabilization in homotopy types. For finite CW complexes, iterated looping and suspending relate to stable homotopy, but the canonical map X \to \Omega^2 \Sigma^2 X is not generally a weak equivalence.

Key Theorems

Hurewicz Theorem

The Hurewicz theorem establishes a fundamental connection between homotopy groups and homology groups in algebraic topology, providing isomorphisms under suitable connectivity assumptions. For a path-connected topological space X, the Hurewicz homomorphism h: \pi_1(X, x_0) \to H_1(X) induces an isomorphism \pi_1(X, x_0)^{\mathrm{ab}} \cong H_1(X) from the abelianization of the fundamental group to the first homology group, where the abelianization is the quotient by the commutator subgroup. This result, originally due to Witold Hurewicz, reflects how loops in X generate homology classes modulo higher-order relations captured by the non-abelian structure of \pi_1. For higher dimensions, if X is path-connected and \pi_k(X) = 0 for all $1 \leq k < n with n \geq 2, then the Hurewicz homomorphism h: \pi_n(X, x_0) \to H_n(X) is an isomorphism, and moreover, H_k(X) = 0 for all k < n. This absolute version of the theorem shows that the first non-vanishing homotopy group determines the corresponding homology group exactly when lower homotopy vanishes. A relative version holds for a pair (X, A) where A is path-connected and \pi_k(X, A) = 0 for $1 \leq k < n with n \geq 2: the relative Hurewicz homomorphism h: \pi_n(X, A, x_0) \to H_n(X, A) is an isomorphism, with H_k(X, A) = 0 for k < n. These statements extend the bridge between homotopy and homology, allowing computations in one theory to inform the other under connectivity conditions. The Hurewicz homomorphism is defined by sending a homotopy class represented by a map f: (S^n, s_0) \to (X, x_0) to the homology class of its image under the fundamental class of S^n in singular or cellular homology chains. A proof sketch proceeds via cellular approximation for CW complexes: any map from an n-sphere to X is homotopic to a cellular map, and the induced chain map on cellular homology aligns with the homotopy class, yielding the homomorphism. To establish the isomorphism, one uses the long exact sequences of the pair (X^{(n)}, X^{(n-1)}) where X^{(k)} denotes the k-skeleton; the relative homotopy and homology groups simplify under the connectivity hypothesis, and induction on n shows surjectivity and injectivity via exactness and vanishing lower terms. For n=1, the kernel is precisely the commutator subgroup, while for n \geq 2, the groups are already abelian, so the kernel is trivial under the assumptions; the image fills H_n completely due to generation by n-cells. In higher dimensions without full connectivity, the homomorphism remains well-defined, but its kernel may contain elements arising from actions of lower homotopy groups, and the image is a subgroup of H_n(X). An important application arises in computing homotopy groups of Eilenberg-MacLane spaces. For the infinite complex projective space \mathbb{CP}^\infty, which has cellular structure with one cell in each even dimension, the homology groups satisfy H_{2k}(\mathbb{CP}^\infty; \mathbb{Z}) \cong \mathbb{Z} for k \geq 1 and vanish otherwise. Since \mathbb{CP}^\infty is simply connected (\pi_1(\mathbb{CP}^\infty) = 0), the Hurewicz theorem implies \pi_2(\mathbb{CP}^\infty) \cong H_2(\mathbb{CP}^\infty; \mathbb{Z}) \cong \mathbb{Z}, generated by the fundamental class of the 2-skeleton \mathbb{CP}^1 \cong S^2. This isomorphism facilitates further computations, as higher homotopy groups of \mathbb{CP}^\infty align with its role as K(\mathbb{Z}, 2).

Freudenthal Suspension Theorem

The Freudenthal suspension theorem provides a precise description of the behavior of the suspension map on of spaces, establishing isomorphisms and surjections in specific dimensional ranges known as the metastable range. For a pair of pointed spaces (X, A) where the inclusion A \to X is obtained by attaching n-cells (with n \geq 2), making the map (n-1)-connected, the induced suspension homomorphism \pi_k(X, A) \to \pi_{k+1}(\Sigma X, \Sigma A) is an isomorphism for k < 2n - 1 and a surjection for k = 2n - 1. This result, originally proved by in 1937 for spheres and later generalized, relies on the connectivity properties of the suspension construction and the exact sequence of a cofibration. The theorem highlights the increasing stability of homotopy groups under suspension, particularly in the context of CW complexes. For an (n-1)-connected CW pair (X, A), the suspension map preserves homotopy information up to the metastable range, where the dimension of the attaching cells limits the exactness. A key tool in proofs is the homotopy excision theorem, which allows decomposition of relative homotopy groups via pushouts, ensuring that maps into the suspension can be analyzed through cellular approximations. Additionally, the suspension-loop adjunction provides a dual perspective, relating the connectivity of the suspension map to that of the loop space fibration. In applications, the theorem defines the stable range for homotopy groups of spheres: for fixed m, the groups \pi_{n+m}(S^n) stabilize as n increases, with the suspension map \pi_{n+m}(S^n) \to \pi_{n+m+1}(S^{n+1}) becoming an isomorphism once n > m + 1. This stabilization underpins the theory of stable homotopy groups \pi_k^s = \lim_{n \to \infty} \pi_{n+k}(S^n). Another significant application is the EHP exact sequence, which decomposes \pi_k(S^n) via maps E: \pi_k(S^n) \to \pi_{k+1}(S^{n+1}) (suspension), H: \pi_{k+1}(S^{n+1}) \to \pi_{k+1}(S^{2n+1}) (Hopf invariant), and P: \pi_{k+1}(S^{2n+1}) \to \pi_{k-1}(S^n) (projection), facilitating computations of unstable homotopy groups. For example, iterating the suspension on \pi_{n+m}(S^n) yields the stable value \pi_m^s for m < n.

Whitehead Theorem

The Whitehead theorem provides a characterization of homotopy equivalences between CW complexes in terms of their induced maps on homotopy groups. Specifically, a map f: X \to Y between path-connected CW complexes X and Y is a homotopy equivalence if and only if f_*: \pi_n(X, x_0) \to \pi_n(Y, f(x_0)) is an isomorphism for every n \geq 0 and every basepoint x_0 \in X. This result, originally established by J. H. C. Whitehead, highlights the sufficiency of CW structure for aligning algebraic invariants like homotopy groups with topological equivalences up to homotopy. A brief sketch of the proof relies on cellular approximation, which allows any map between CW complexes to be homotoped to a cellular map that sends the n-skeleton of X into the n-skeleton of Y. For such cellular maps inducing isomorphisms on all homotopy groups, induction over skeleta shows that f admits a homotopy inverse: the relative homotopy groups \pi_k(Y, f(X^{(n)})) vanish for k \leq n due to the isomorphisms on \pi_*, enabling the construction of homotopy extensions step by step. Alternatively, using Postnikov towers, the map f induces weak equivalences on each stage of the towers for X and Y, implying a global homotopy equivalence since CW complexes are fibrant in the classical model structure. This approach underscores the role of obstruction theory, where higher obstructions to extending lifts vanish precisely because of the homotopy group isomorphisms. The theorem extends naturally to cellular maps between CW complexes, where the cellular structure simplifies the verification of homotopy inverses via the long exact sequence of the mapping cone. For simply connected CW complexes, the result is particularly powerful, as the vanishing fundamental group eliminates non-abelian complications in low dimensions, allowing homotopy groups to fully determine the homotopy type. In this setting, maps inducing isomorphisms on \pi_n for n \geq 2 suffice, with \pi_1 trivially handled. However, the converse fails without the CW hypothesis: there exist non-CW spaces where maps induce isomorphisms on all homotopy groups but are not homotopy equivalences. A classic counterexample is the Hawaiian earring, a compact subspace of \mathbb{R}^2 consisting of circles of radius $1/n tangent at the origin; its fundamental group is uncountable and not free, so inclusions or certain projections can match homotopy groups with the wedge of circles (a CW complex) without being homotopy equivalences. Such pathologies arise because non-CW spaces lack the skeletal control that enables cellular approximation and inductive arguments.

Advanced Structures

Homotopy Limits and Colimits

In homotopy theory, homotopy colimits provide a homotopically invariant replacement for ordinary colimits of diagrams of topological spaces or simplicial sets. For a diagram D: \mathcal{I} \to \mathbf{Top}, where \mathcal{I} is a small category, the homotopy colimit is constructed as the geometric realization of the simplicial replacement of D, denoted |\operatorname{srep}(D)|, where the n-simplices are given by the disjoint union \coprod_{i_0 \to \cdots \to i_n} D(i_n). This construction ensures that the homotopy colimit is functorial and preserves weak equivalences up to homotopy, making it suitable for computations in the homotopy category. A prominent example of a homotopy colimit is the mapping telescope, which computes the homotopy colimit of a sequential diagram A_0 \to A_1 \to A_2 \to \cdots. The mapping telescope \operatorname{Tel}(A_\bullet) is formed by iteratively attaching mapping cylinders, yielding a space whose homotopy type captures the "union at infinity" of the sequence in a homotopical sense. Another example is the homotopy pushout of a diagram A \to B \leftarrow C, realized as the double mapping cylinder (B \sqcup (A \times I) \sqcup C)/{\sim}, where identifications glue the endpoints of the cylinders along the maps from A. For diagrams arising from group actions, the Borel construction computes the homotopy colimit over the category of simplices in the classifying space BG, yielding (EG \times_G X) for a G-space X, which is homotopy equivalent to the homotopy quotient. Dually, homotopy limits serve as the right derived functors of ordinary limits, providing a way to take limits that respect the homotopy structure of diagrams. For a diagram D: \mathcal{I}^{op} \to \mathbf{sSet}, the homotopy limit is the end \int_{i \in \mathcal{I}} \operatorname{Map}(N(\mathcal{I}/i), D(i)), or equivalently, the totalization of the cosimplicial replacement \operatorname{Tot}(\operatorname{crep}(D)). This can be computed as a function complex over the simplicial diagram, ensuring invariance under weak equivalences. A basic example is the homotopy pullback for fiber products, which for a cospan B \to A \leftarrow C is the pullback of the path space fibration, yielding a space whose homotopy type encodes the homotopy fiber. In the homotopy category \mathbf{Ho}(\mathbf{Top}), homotopy limits and colimits correspond to the derived functors R \lim and L \operatorname{colim} of the underlying limit and colimit functors, respectively, allowing diagrams to be inverted up to homotopy when they induce quasi-isomorphisms on homotopy groups. These constructions are essential for handling infinite diagrams and ensuring that categorical operations remain well-defined in homotopical settings, such as in the study of fibrations where homotopy limits can be computed via pullbacks along fibrations.

Classifying Spaces and Homotopy Operations

In algebraic topology, the classifying space BG of a G is a space whose fundamental group is isomorphic to G and whose higher homotopy groups vanish, i.e., \pi_1(BG) \cong G and \pi_i(BG) = 0 for i > 1, when G is . More generally, for any topological group G, BG is defined up to homotopy equivalence as the base space of a principal G-bundle EG \to BG, where EG is contractible and thus serves as a universal cover. The existence of such a classifying space is guaranteed by the Milnor construction, which builds BG as the quotient of an infinite join of copies of G. Principal G-bundles over a paracompact base space X are classified up to isomorphism by the homotopy classes of maps [X, BG], the set of homotopy classes from X to BG. Specifically, any principal G-bundle P \to X admits a classifying map f: X \to BG such that P is pullback equivalent to the universal bundle EG \to BG along f, and two bundles are isomorphic if and only if their classifying maps are homotopic. This classification reduces the study of bundle geometry to homotopy theory on BG. For example, the classifying space for the circle group S^1 is the infinite complex projective space \mathbb{CP}^\infty = BS^1, which classifies complex line bundles over X via [X, \mathbb{CP}^\infty]. Eilenberg-MacLane spaces K(G, n) generalize classifying spaces to higher dimensions, defined as connected spaces with \pi_n(K(G, n)) \cong G (abelian for n \geq 2) and \pi_i(K(G, n)) = 0 for i \neq n. These spaces were introduced by Eilenberg and Mac Lane to model groups, as the H^n(X; G) of a space X is isomorphic to the classes [X, K(G, n)]. For discrete groups, BG = K(G, 1), linking the two concepts. Constructions of K(G, n) often use Postnikov towers or simplicial methods, ensuring uniqueness up to equivalence. Homotopy operations, or cohomology operations, arise naturally from the homotopy classes of maps between Eilenberg-MacLane spaces, providing tools to extract structural information from cohomology rings. Primary cohomology operations are induced by fixed maps \phi: K(G, n) \to K(H, m), defining natural transformations \phi^*: H^m(-; H) \to H^n(-; G) on the cohomology of arbitrary spaces. For mod-2 cohomology, the Steenrod squares Sq^k: H^n(X; \mathbb{Z}/2) \to H^{n+k}(X; \mathbb{Z}/2) form a family of primary operations, satisfying the and generating the \mathcal{A}, the ring of all stable cohomology operations under mod-2 coefficients. This algebra, first systematically described by Steenrod, acts on the cohomology ring and detects topological features like . Secondary cohomology operations refine primary ones, measuring obstructions or differences when a primary operation is undefined on certain classes. They are defined via the difference between two extensions in a Postnikov tower or as elements in the cohomology of the primary operation's domain, often living in groups like H^{m+1}(K(G, n); H). For instance, secondary operations associated to the Steenrod square detect violations of the Wu formula in manifold cohomology. These operations highlight the richness of unstable homotopy theory before stabilization to spectra.

Spectra and Generalized Cohomology Theories

In homotopy theory, spectra formalize the stable aspects of groups and enable the construction of generalized theories that satisfy the Eilenberg-Steenrod axioms except the dimension axiom. A E consists of a sequence of pointed connected topological spaces \{E_n\}_{n \in \mathbb{Z}} equipped with structure maps \sigma_n: \Sigma E_n \to E_{n+1} for each n, where \Sigma denotes the reduced , and the maps are such that the induced maps on homotopy groups are isomorphisms in sufficiently high dimensions. This construction stabilizes the homotopy information, as the adjoint of \sigma_n induces maps E_n \to \Omega E_{n+1}, and in the \Omega- case (a refinement), these are weak equivalences. The groups of a spectrum E are defined by \pi_k(E) = \colim_n \pi_{k+n}(E_n), where the colimit is taken over the system induced by the structure maps, capturing the stable in each degree k. Generalized theories represented by E are given by h^n(X) = [X, E_n], the pointed classes of maps from a X to E_n, with suspension isomorphisms h^n(X) \cong h^{n+1}(\Sigma X). These functors satisfy exactness for cofiber sequences in spaces, yielding long exact sequences in . The Atiyah-Hirzebruch relates h^*(X) to ordinary via an exact couple derived from the skeletal of a CW-complex X, with E_2^{p,q} = H^p(X; \pi_{-q}(E)) converging to h^{p+q}(X). Prominent examples include the Eilenberg-MacLane spectrum HG for an G, where (HG)_n = K(G, n) (the Eilenberg-MacLane space) and the structure maps \sigma_n: \Sigma K(G, n) \to K(G, n+1) are the canonical homeomorphisms, representing ordinary H^n(X; G) = [X, K(G, n)]. The spectrum S, defined by S_n = S^n (the n-) with identity structure maps, represents stable , where the groups \pi_k(S) are the stable stems \pi_k^s. Thom spectra arise in the context of oriented generalized cohomology theories, where orientation provides a Thom class in the cohomology of Thom spaces of vector bundles. For complex-oriented theories, the Thom spectrum MU is constructed as the spectrum associated to the Thom spaces of the universal complex vector bundles over BU(n), serving as the universal such theory and representing complex cobordism. More generally, Thom spectra like MO (for real-oriented theories over BO) or MSO (for oriented theories over BSO) encode bordism theories via their homotopy groups. Many spectra carry additional structure, particularly ring structures, making them commutative monoids in the category of spectra under the smash product. A ring spectrum E has maps E \wedge E \to E compatible with the unit map S \to E and associative up to coherent homotopy, inducing ring structures on \pi_*(E) and multiplicative pairings in the represented cohomology. The spectrum MU exemplifies this, as an E_\infty-ring spectrum whose associated formal group law governs complex-oriented theories.

Obstruction and Extension Theory

Primary and Secondary Obstructions

Obstruction theory provides a cohomological framework for determining whether a continuous map defined on a subspace of a space can be extended to the entire space, or whether a section of a fibration can be lifted over successive skeleta. In the context of a fibration p: E \to B with fiber F, consider a map f: X \to B defined on the (n-1)-skeleton X^{(n-1)} of a CW complex X, and attempt to extend it to X^{(n)} while lifting through p to obtain a map \tilde{f}: X^{(n)} \to E such that p \circ \tilde{f} = f on X^{(n)}. The primary obstruction to this extension is a well-defined cohomology class in the group H^{n}(X; \pi_{n-1}(F)), where the coefficients are the (n-1)th homotopy group of the fiber F. This class arises from the difference in homotopy classes of maps on n-cells, measured relative to the fiber's homotopy, and vanishes if and only if the extension exists up to homotopy. If the primary obstruction vanishes, the lift exists over the n-skeleton, allowing inductive construction over the CW structure of X. When the primary obstruction is zero, secondary obstructions may still prevent a global or extension. These are captured by classes in H^{n+1}(X; \pi_n(F)), representing ambiguities or differences between possible refinements in the Postnikov tower of the target space or fibration. Specifically, secondary obstructions measure inconsistencies in choosing compatible lifts over higher skeleta, arising from the action of higher groups on the primary data. In the Postnikov tower decomposition, where the space is built as a sequence of fibrations with Eilenberg-MacLane spaces as fibers, these obstructions correspond to the failure of sections to refine consistently across tower stages. Computations of such obstructions often employ the Eilenberg-Moore , which converges to the of the total space and aids in evaluating the relevant coefficient groups \pi_n(F) in terms of terms over the base's ring. A classic example of primary obstructions occurs in the problem of finding sections of vector bundles. For an oriented bundle S^k \to E \to B over a B, the primary obstruction to a global is the e(E) \in H^{k+1}(B; \mathbb{Z}), which lies in the detected by \pi_k(S^k) \cong \mathbb{Z}. This class vanishes over the (k+1)-skeleton if a exists there, but nonzero values, such as in the S^1 \to S^3 \to S^2 where e = 1 \in H^2(S^2; \mathbb{Z}), obstruct a entirely. For map extensions over spheres, consider extending a S^{n-1} \to Y to the disk D^n; the primary obstruction is the class in \pi_{n-1}(Y), which detects when Y is an Eilenberg-MacLane space. Secondary obstructions appear in higher dimensions, for instance, in non-orientable bundles where \pi_{k+1}(SO(k+1)/SO(k)) \cong \mathbb{Z}/2\mathbb{Z} yields torsion classes in H^{k+2}(B; \mathbb{Z}/2\mathbb{Z}) that block sections even if the primary vanishes.

Characteristic Classes in Homotopy

In homotopy theory, characteristic classes for vector bundles arise from the homotopy classes of maps from the base space to the classifying spaces of the bundles. For a complex vector bundle of rank n over a space P, the isomorphism classes correspond to elements of [P, BU(n)], the set of homotopy classes of maps from P to the classifying space BU(n). The Chern classes c_k are then the pullbacks of the universal Chern classes in H^{2k}(BU(n); \mathbb{Z}), providing invariants that detect the topological twisting of the bundle. Similarly, for real vector bundles, the Stiefel-Whitney classes w_k are defined via [P, BO(n)], pulling back universal classes from H^k(BO(n); \mathbb{Z}/2), and serve as mod 2 invariants of the bundle's structure. A homotopy-theoretic definition of these classes employs the Thom isomorphism and Steenrod operations. The Thom isomorphism relates the cohomology of the base P to that of the Thom space of the bundle, mapping H^k(P; \mathbb{Z}/2) isomorphically to H^{k+n}(Th(\xi); \mathbb{Z}/2) via the Thom class U \in H^n(Th(\xi); \mathbb{Z}/2). For Stiefel-Whitney classes, Thom's construction applies the Steenrod square Sq^k to the Thom class: w_k(\xi) = \phi^{-1}(Sq^k U), where \phi is the Thom isomorphism, yielding the classes as the unique natural, multiplicative mod 2 cohomology classes satisfying this relation. For Chern classes, an analogous construction uses the oriented Thom isomorphism and higher cohomology operations, though the primary axiomatic approach via classifying spaces predominates. These classes are intimately related to homotopy obstructions in . The first c_1 serves as the primary obstruction to the triviality of a line bundle, as the classifying map to BU(1) \simeq \mathbb{CP}^\infty induces c_1 \in H^2(P; \mathbb{Z}), and vanishing of c_1 is necessary and sufficient for the bundle to be trivial over simply connected bases. More generally, higher characteristic classes detect obstructions to sections or reductions of structure groups, with the k-th class lying in the appropriate group corresponding to the Postnikov tower of the . A representative example is the Euler class e(\xi) for an oriented real vector bundle \xi of rank n, defined in H^n(P; \mathbb{Z}) as the primary obstruction to the existence of a nowhere-zero section, via the restriction of the fundamental class of the Thom space to the zero section. For complex bundles, the top Chern class c_n coincides with the Euler class of the underlying oriented real bundle. Computations often proceed via clutching functions: for a bundle over S^m, the clutching map in \pi_{m-1}(O(n)) or U(n) determines the classes; for instance, the Hopf line bundle over S^2 has clutching function the Hopf fibration, yielding c_1 as the generator of H^2(S^2; \mathbb{Z}).

Localization and Completion

Bousfield Localization of Spaces

Bousfield localization provides a method to refine the homotopy type of spaces by formally inverting a specified class of morphisms, analogous to localization in algebra but adapted to the homotopy category of spaces. For a map f: A \to B between pointed connected spaces, the f-localization functor L_f assigns to each space X a space L_f X equipped with a natural f-equivalence X \to L_f X, such that L_f X is f-local, meaning that the induced map \operatorname{Map}(B, L_f X) \to \operatorname{Map}(A, L_f X) is a weak homotopy equivalence. This construction ensures that f becomes an equivalence in the localized homotopy category, where morphisms are inverted if they induce weak equivalences after composition with maps to local objects. The f-localization L_f X can be realized as a homotopy colimit of the cosimplicial space obtained by resolving X with respect to f, specifically the homotopy colimit \operatorname{hocolim}^n \operatorname{[Map](/page/Map)}(( \Delta^n, \partial \Delta^n ) \times (B, A), X), where the diagram is built using simplicial replacements and pushouts to approximate the universal property. Local objects are precisely those spaces Y for which every f-equivalence X \to X' induces a weak equivalence \operatorname{[Map](/page/Map)}(Y, X) \to \operatorname{[Map](/page/Map)}(Y, X'), or equivalently, \operatorname{[Map](/page/Map)}(B, Y) \to \operatorname{[Map](/page/Map)}(A, Y) is a weak equivalence. The Bousfield class \langle f \rangle associated to f consists of all maps that yield the same class of local objects, forming a under inclusion that classifies distinct localizations. Prominent examples include rational homotopy localization L_\mathbb{Q} X, which inverts the rational equivalences (maps inducing isomorphisms on rational groups \pi_* \otimes \mathbb{Q}), resulting in \pi_*(L_\mathbb{Q} X) \cong \pi_*(X) \otimes \mathbb{Q} as graded vector spaces over \mathbb{Q}. Another is p-localization L_p X for a prime p, which inverts the p-local equivalences (maps inducing isomorphisms on \pi_* \otimes \mathbb{Z}_{(p)}), making \pi_*(L_p X) into modules over the p-local integers \mathbb{Z}_{(p)}. To compute the homotopy groups \pi_*(L_f X), one employs the Bousfield-Kan arising from the cosimplicial resolution used in the construction of L_f X; this has E_2^{s,t} = \pi_s \operatorname{[Map](/page/Map)}(( \Delta^t, \partial \Delta^t ) \times (B, A), X) and converges strongly to \pi_{s+t} (L_f X) under suitable connectivity assumptions on f and X.

p-Completion and Rationalization

In theory, p-completion provides a way to extract and approximate the p-adic information from the homotopy groups of a space. For a simply connected pointed X, the p-completion \hat{X}_p is a p-complete space whose homotopy groups are the p-adic completions of those of X, satisfying \pi_*(\hat{X}_p) \cong \pi_*(X) \hat{\otimes} \mathbb{Z}_p. This construction, developed by Bousfield and Kan, proceeds via p-adic approximations using a cosimplicial of X that builds a tower of fibrations, with the p-completion obtained as the homotopy limit of this tower. The resulting \hat{X}_p preserves p-local homotopy equivalences and is useful for studying torsion-free aspects modulo other primes. Rationalization, on the other hand, localizes spaces at the rationals to capture the rational vector space structure of their homotopy groups. For a simply connected space X, the rationalization X_\mathbb{Q} is a rational space such that \pi_*(X_\mathbb{Q}) \otimes \mathbb{Q} \cong \pi_*(X) \otimes \mathbb{Q}. In Sullivan's approach, this is modeled using minimal models, which are free commutative differential graded algebras minimally extending the cohomology algebra of X with generators corresponding to rational homotopy classes. These models facilitate computations by translating rational homotopy problems into algebraic questions about derivations and quasi-isomorphisms. Rationalization often simplifies the homotopy type, as rational spaces behave like products of Eilenberg-MacLane spaces and odd-dimensional spheres. A representative example of p-completion is the 3-sphere S^3. Its p-completion \widehat{S^3}_p has fundamental group \pi_3(\widehat{S^3}_p) \cong \mathbb{Z}_p, with higher homotopy groups given by the p-adic completions of the unstable homotopy groups of S^3. In contrast, rationalization of spheres illustrates the sparsity of rational homotopy: for n \geq 2, the rational homotopy groups of S^n are trivial above dimension n, with \pi_n(S^n_\mathbb{Q}) \cong \mathbb{Q} and \pi_k(S^n_\mathbb{Q}) = 0 for k > n. This follows from the Sullivan minimal model of S^n, which is the free commutative dga on a single generator in degree n with trivial differential. In the stable regime, the Adams spectral sequence computes p-local stable homotopy groups, providing essential data for p-local and p-complete approximations of stable stems. Originating from Adams' work on the , the spectral sequence converges to the p-local homotopy groups of the sphere spectrum, \pi_*^s_{(p)}, with E_2-term given by Ext groups in the category of modules over the . This tool is particularly powerful for determining the p-local structure of stable , including images of J and v_1-periodic phenomena, and underpins computations in p-complete stable .

Specific Homotopy Theories

Stable Homotopy Theory

Stable homotopy theory emerges in the study of homotopy groups of spheres by considering the behavior under repeated suspension, where the suspension functor becomes an equivalence in sufficiently high dimensions. The stable homotopy groups of spheres, denoted \pi_*^s, are defined as the colimit \pi_k^s = \colim_n \pi_{n+k}(S^n), capturing the groups that stabilize after infinitely many suspensions. This colimit exists due to the Freudenthal suspension theorem, which ensures that the suspension maps are isomorphisms for k \leq n-2 (the stable range) and surjective for k = n-1, allowing the groups to stabilize for fixed k and sufficiently large n, so that the colimit exists. The sphere spectrum S, whose n-th space is the n-sphere S^n with structure maps given by the degree-one map on loop spaces, represents the initial object in the stable homotopy category, and its homotopy groups are precisely \pi_*^s. A primary tool for computing these groups is the , which converges to the p-local stable stems \pi_*^s \otimes \mathbb{Z}_{(p)}. The E_2-page is given by \operatorname{Ext}_{\mathcal{A}}(\mathbb{F}_p, H_*(S; \mathbb{F}_p)), where \mathcal{A} is the acting on the mod-p of the sphere spectrum. Introduced by , this spectral sequence arises from a minimal of the trivial module over \mathcal{A} and detects permanent cycles corresponding to elements in the stable stems, with differentials arising from higher operations in . Computations via this sequence have determined \pi_*^s up to high dimensions, revealing a rich structure including torsion and infinite families. As of 2025, these computations, enhanced by synthetic spectra and motivic homotopy theory, have determined the groups up to stems exceeding dimension 100, uncovering new periodic families and torsion elements. Key examples illustrate the depth of stable homotopy computations. The image of the J-homomorphism, J: \pi_*(O) \to \pi_*^s, embeds the homotopy groups of the into the stable stems and was fully described by Adams using secondary cohomology operations; at odd primes p, it is cyclic of order the denominator of the p-part of Bernoulli numbers B_{2m}/(4m), while at p=2 it involves numbers up to certain indices. Adams' solution to the Hopf invariant one problem, using the , shows that maps S^{2n-1} \to S^n of Hopf invariant \pm 1 exist only for n=2,4,8, corresponding to the classical division algebras \mathbb{R}, \mathbb{C}, \mathbb{H}, \mathbb{O}. This resolution highlighted the role of the in obstructing such maps beyond these dimensions. Adams' conjecture posits that the image of J forms a direct summand of \pi_*^s, which was proved by Quillen using algebraic and the trace. In the context of spectra, the provides a tool for of spectral sequences associated to filtered spectra, ensuring that the filtration on homotopy groups refines the Adams filtration. Connective covers of periodic spectra, such as the connective cover f_0 X of a spectrum X with \pi_k X = 0 for k < 0, truncate negative homotopy groups while preserving positive ones, facilitating computations in the connective stable homotopy category.

Equivariant Homotopy Theory

Equivariant homotopy theory extends classical homotopy theory to spaces equipped with a continuous of a G, known as G-spaces. A G-space X is a together with a continuous G \times X \to X satisfying the group axioms, and an f: X \to Y between G-spaces is a continuous such that f(g \cdot x) = g \cdot f(x) for all g \in G and x \in X. This framework captures symmetries in topological spaces, with applications in and . The orbit category \mathcal{O}_G, introduced by tom Dieck, provides a foundational structure for equivariant homotopy; its objects are the homogeneous G-spaces G/H for closed subgroups H \leq G, and its morphisms are G-equivariant continuous maps. The tom Dieck completion \widehat{\mathcal{O}}_G enhances this category by formally completing it with respect to certain colimits, enabling the construction of equivariant spectra and the study of fixed points; for finite G, it decomposes equivariant objects into sums over conjugacy classes of subgroups. Equivariant homotopy groups \pi_*^G(X) for a pointed G-space X are defined as the colimit over n of [S^{n\rho_G}, X]^G, where \rho_G is the of G and [-,-]^G denotes G-equivariant homotopy classes of maps; in the stable range, these groups are \mathrm{RO}(G)-graded, indexed by virtual real G-representations V \in \mathrm{RO}(G), so \pi_V^G(X) = [S^V, X]^G. This grading reflects the representation-theoretic nature of equivariant phenomena. A key example is equivariant cohomology, computed via the Borel construction: for a G-space X, the Borel space X_{hG} = EG \times_G X (where EG is the total space of the universal G-bundle) has ordinary H^*(X_{hG}; R) isomorphic to the equivariant H^*_G(X; R) for a coefficient R. The classifying space BG arises as the Borel construction on a point, parametrizing principal G-bundles. The Burnside A(G), the Grothendieck ring of isomorphism classes of finite G-sets under and product, is isomorphic as a to the degree-zero equivariant stable \pi_0^G(S) of the spectrum S, with basis given by conjugacy classes of subgroups of G; this maps the structure set [G/H] to the class of the induced [G/H]_+. Computations of A(G) for finite G reveal its rank equals the number of conjugacy classes, with further periodic structures in its chromatic explored by , Ravenel, and using Hopf methods. Transfer maps play a central role in relating homotopy groups across subgroups: for H \leq G, the \mathrm{tr}_H^G: \pi_*^H(Y) \to \pi_*^G(G \wedge_H Y) induces from H-equivariant data to G-equivariant, corresponding to of representations and compatible with the Burnside structure via multiplicative norms. Fixed points spectral sequences facilitate computations by relating the equivariant homotopy of a spectrum E to its fixed point spectra E^H; for instance, tom Dieck's splitting theorem decomposes the suspension spectrum \Sigma^\infty X_+ as a over conjugacy classes [(G/H)] \vee \Sigma^\infty (X^H)_{h(W_H)}, where W_H = N_G(H)/H is the , leading to sequences converging to \pi_*^G(X) from the fixed point data. These tools underpin the reduction of equivariant problems to non-equivariant ones via geometric fixed points.

Abstract Frameworks

Simplicial Sets and Models

Simplicial sets provide a combinatorial for modeling topological spaces in homotopy theory, defined as functors from the opposite category of the simplex category to the . The simplex category has objects = \{0, 1, \dots, n\} for n \geq 0, with morphisms being non-decreasing functions, and a simplicial set K assigns to each $$ a set K_n of n-simplices, together with face maps d_i: K_n \to K_{n-1} and degeneracy maps s_i: K_n \to K_{n+1} satisfying the simplicial identities. This functorial perspective allows simplicial sets to capture higher-dimensional structures discretely, avoiding the pathologies of point-set topology. The geometric realization functor |K| maps a simplicial set K to a by gluing standard simplices \Delta^n along their faces according to the simplicial structure. Formally, |K| is the quotient of the \coprod_{n \geq 0} K_n \times |\Delta^n| by an identifying faces and degeneracies appropriately, yielding a CW-complex whose cells correspond to the non-degenerate simplices of K. This realization preserves types, enabling computations in the combinatorial setting to inform topological properties. Kan fibrations introduce a notion of fibrations within simplicial sets, characterized by lifting properties against horn inclusions. A horn \Lambda^n_k is the simplicial subset of \Delta^n obtained by removing the interior of the face opposite vertex k, and a map p: E \to B is a Kan fibration if it has the right lifting property against all horn inclusions \Lambda^n_k \hookrightarrow \Delta^n. That is, given any commutative diagram consisting of the inclusion \Lambda^n_k \hookrightarrow \Delta^n on the left, a map \Delta^n \to B on the bottom, and a map \Lambda^n_k \to E on the top such that the square commutes via p, there exists a diagonal filler \Delta^n \to E. A Kan complex is a simplicial set fibrant over the terminal simplicial set \Delta^0, meaning it admits fillers for all horns, and such complexes model \infty-groupoids by encoding higher homotopies as simplices. The singular simplicial set \operatorname{Sing}(X) of a topological space X is defined by \operatorname{Sing}(X)_n = \operatorname{Hom}(|\Delta^n|, X), the set of continuous maps from the standard n-simplex to X, with faces and degeneracies induced by the simplicial operators on \Delta^n. The adjoint pair between geometric realization |-|: \mathbf{sSet} \dashv \operatorname{Sing}: \mathbf{Top} induces a Quillen equivalence between the Kan-Quillen model structure on simplicial sets and the Serre model structure on topological spaces, preserving weak homotopy equivalences and thus homotopy types. Prominent examples include the nerve construction, where for a small category \mathcal{C}, the nerve N\mathcal{C} has (N\mathcal{C})_n as the set of chains of n composable morphisms in \mathcal{C}, modeling the B\mathcal{C} = |N\mathcal{C}| up to . For a G, the BG is the nerve of the one-object with morphisms G, capturing the homotopy type of the Eilenberg-MacLane space K(G,1).

Model Categories and ∞-Categories

Model categories provide a general framework for doing homotopy theory in arbitrary categories, introduced by Daniel Quillen in his seminal work on homotopical algebra. A model category \mathcal{C} is equipped with three distinguished classes of morphisms: weak equivalences, fibrations, and cofibrations, satisfying the axioms of a complete and cocomplete category together with two-out-of-three properties for these classes, lifting axioms, and factorization axioms that allow every morphism to be factored into a cofibration followed by a trivial fibration or a trivial cofibration followed by a fibration. These axioms ensure that homotopy-theoretic constructions, such as colimits and limits, can be computed using resolutions by fibrant or cofibrant objects, enabling the abstract treatment of derived functors and homotopy limits and colimits in a wide variety of settings beyond topological spaces. Central to the theory is the \mathrm{Ho}(\mathcal{C}), obtained by localizing the \mathcal{C} at the class of weak equivalences, denoted \mathcal{C}[\mathrm{we}^{-1}], where morphisms are equivalence classes of zigzags of weak equivalences and original morphisms, up to relations defined using path objects or cylinder objects in the model . This localization functor \gamma: \mathcal{C} \to \mathrm{Ho}(\mathcal{C}) inverts weak equivalences and preserves -invariant information, allowing the definition of derived functors L_F and R_F for functors F: \mathcal{C} \to \mathcal{D} between model categories as the total derived functors on the homotopy categories, computed via fibrant-cofibrant replacements. Quillen equivalences between model categories induce equivalences on their homotopy categories, providing a way to compare different models of the same theory. While model categories capture 1-categorical aspects of , higher-dimensional phenomena require the more flexible notion of \infty-categories, developed by Jacob Lurie as a model for weak (\infty,1)-categories where composition is defined up to coherent higher homotopies. Lurie presents \infty-categories via quasicategories, which are simplicial sets satisfying the weak Kan condition, or equivalently through Segal categories and complete Segal spaces, providing a homotopy-coherent enrichment over spaces that generalizes ordinary categories to allow for invertible higher morphisms. The homotopy category of an \infty-category recovers the 1-truncated structure, and model categories can be regarded as presentations of \infty-categories via their homotopy coherent nerve, bridging classical homotopical algebra with higher categorical foundations. A key example is the Quillen model structure on the category \mathrm{sSet} of simplicial sets, where weak equivalences are weak equivalences, fibrations are Kan fibrations, and cofibrations are monomorphisms, enabling simplicial sets to model the \infty-category of spaces up to . In , the \infty-category of spectra \mathrm{Sp} forms a stable \infty-category, characterized by the property that the suspension functor is an equivalence and it admits finite colimits, serving as the universal stable \infty-category into which any pointed \infty-category with finite colimits admits a colimit-preserving . This structure underlies the stable homotopy category and facilitates computations in chromatic homotopy theory and derived algebraic geometry.

Homotopy Hypothesis

The homotopy hypothesis, proposed by , asserts that the homotopy types of topological spaces are equivalent to ∞-groupoids up to ∞-equivalence. This conjecture posits that every homotopy type can be modeled by an ∞-groupoid, where higher-dimensional morphisms capture homotopies, higher homotopies, and so forth, providing a categorical framework for the weak equivalences and homotopy relations in spaces. formulated this idea in the context of his work on stacks and the Grothendieck-Teichmüller group, aiming to reinterpret through higher categorical structures in . Evidence for the hypothesis emerges from combinatorial models of ∞-groupoids, such as complexes, which are simplicial sets satisfying the Kan filler conditions and thus realize the homotopy types of spaces via their geometric realization functor. Similarly, quasi-categories, introduced by Boardman and Vogt and developed by , serve as another model where the ∞-category of quasi-categories equivalent to complexes provides a presentation of the hypothesis. These models demonstrate that the weak higher-categorical structure of ∞-groupoids aligns with the theory of spaces, supporting Grothendieck's vision. A key implication is the existence of the fundamental ∞-groupoid of a space X, denoted \Pi_\infty(X), which encodes the entire homotopy type of X through its objects (points of X) and morphisms (paths up to , higher homotopies, etc.). This structure extends Grothendieck's idea to weak n-categories modeling n-types for finite n. The Baez-Dolan further refines this by proposing that weak \omega-categories suffice to model all homotopy types, with their stabilization hypothesis ensuring that adding higher invertible morphisms preserves the homotopy theory. Modern resolutions of the homotopy hypothesis appear in Lurie's development of ∞-categories, where the ∞-category of spaces \mathcal{S} is equivalent to the ∞-category of ∞-groupoids, establishing a Quillen equivalence between models like simplicial sets and quasi-categories. This framework confirms that homotopy types are precisely the ∞-groupoids, providing a rigorous foundation for Grothendieck's conjecture in higher topos theory.

Applications

In Algebraic Topology

Homotopy theory plays a central role in by providing tools to compute fundamental invariants of topological spaces and to classify manifolds up to or . Through the study of groups and related structures, it enables the identification of obstructions to equivalence relations among manifolds, such as embeddings and smooth structures. In particular, it underpins the of high-dimensional manifolds by revealing how homotopy types determine geometric properties. A landmark application is the via , pioneered by Michel Kervaire and in their 1963 work on groups of . They demonstrated that smooth manifolds homotopy equivalent to spheres in dimensions greater than four can be classified using the stable , combined with surgery obstructions that measure deviations from the standard . Specifically, the group of homotopy spheres, denoted Θ_n, captures the exotic spheres—manifolds homeomorphic but not diffeomorphic to the standard —and provides an relating these to groups and homotopy invariants. This framework reduces the classification problem to algebraic computations, showing, for example, that there are 28 exotic 7-spheres. Homotopy theory also detects obstructions to embeddings and immersions of manifolds into spaces. In the work of Haefliger and Hirsch from the early , immersions are classified up to regular using the groups of the general linear groups, while embeddings face primary obstructions in the of the manifold with coefficients in stable normal bundles. For instance, a map from an m-manifold to ℝ^n admits an immersion if n ≥ 2m, but embedding obstructions arise in the metastable range (3m/2 < n < 2m) via Haefliger invariants in twisted groups, ensuring that certain classes cannot be realized without self-intersections. These results extend to higher dimensions, where secondary obstructions in groups further refine the classification. The resolution of the by in 2002–2003 exemplifies 's interplay with in classifying 3-manifolds. Perelman used with to prove that every simply connected closed is homeomorphic to the , linking the to the of spheres. His entropy functional and deform the metric to a , where the absence of nontrivial spheres in 3 (as established by earlier work) implies diffeomorphism to the standard after finite-time extinction. This not only verifies the but also geometrizes all 3-manifolds, with equivalence serving as the key invariant. Recent advances employ homotopy-theoretic methods to study diffeomorphism groups of manifolds, revealing their stable homotopy types and rational cohomology. For prime 3-manifolds, the diffeomorphism group deformation retracts onto the isometry group, implying finite presentation and controlled homotopy structure, which aids in understanding moduli spaces of geometric structures. These approaches, building on embedding calculus and stable homotopy, have classified rational homotopy groups for diffeomorphisms of odd-dimensional discs and extended to reducible manifolds, where the classifying space has finite type. Higher homotopy groups occasionally provide additional manifold invariants in these contexts. More recently, Bamler and Kleiner (2023) completed the proof of the generalized Smale conjecture for most 3-manifolds using Ricci flow, confirming that the diffeomorphism group is homotopy equivalent to the isometry group. Additionally, Bregman, Boyd, and Steinebrunner (2024) proved Kontsevich's conjecture, showing that the moduli spaces of orientable 3-manifolds with boundary are finite, providing further insights into the homotopy types of these groups.

To Other Mathematical Fields

Homotopy theory extends its principles to algebraic contexts through derived categories and differential graded (DG) algebras, where homotopical equivalences capture essential structures beyond strict isomorphisms. Derived categories, initially formalized by Verdier in the context of triangulated categories, incorporate to model complexes up to quasi-isomorphisms, enabling the study of modules and sheaves with homological precision. This framework allows DG-algebras to serve as models for associative algebras in a homotopical setting, where the homotopy category of DG-modules yields the derived category, facilitating derived Morita equivalences between DG-categories. A seminal development is Toën's construction of a model structure on DG-categories, which establishes a theory that aligns derived categories with applications, such as non-commutative motives. In , motivic homotopy theory applies homotopy-theoretic tools to algebraic varieties, treating them analogously to topological spaces but respecting the affine line \mathbb{A}^1 as a model for contractible intervals. Pioneered by Voevodsky in the 1990s, this theory replaces singular simplices with representable presheaves and defines weak equivalences via \mathbb{A}^1-invariance, leading to a stable homotopy category of motives that encodes arithmetic and geometric invariants like . Voevodsky's lectures formalized the unstable and stable versions, proving that aligns with classical cohomology theories for smooth varieties, thus bridging with topological methods. Homotopy theory influences physics through connections to topological quantum field theories (TQFTs), particularly via the Stolz-Teichner program, which links supersymmetric field theories to spectra in . This approach posits that 2-dimensional supersymmetric TQFTs correspond to elliptic cohomology theories represented by Thom spectra of oriented , providing a geometric realization of modular forms and invariants. Stolz and Teichner's framework extends categories to include Riemannian metrics and super-Minkowski spaces, yielding homotopy-invariant functors that classify field theories up to . A recent interdisciplinary impact is (HoTT), developed by Voevodsky and collaborators in the as univalent foundations for mathematics, where types are interpreted as types and as paths in the of types. This synthesis of Martin-Löf with theory via the univalence equates equivalent types, enabling formal proofs in proof assistants like that mirror homotopical reasoning. The foundational text outlines HoTT's syntax and semantics, demonstrating its use in synthetic theory and higher inductive types for modeling ∞-groupoids. Model categories serve as a brief bridge here, providing the categorical infrastructure to interpret type constructors homotopically.

References

  1. [1]
    [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 ...
  2. [2]
    [PDF] Rational Homotopy Theory
    Rational homotopy theory is the study of the rational homotopy category, that is the category obtained from the category of 1-connected pointed spaces.
  3. [3]
    [2212.11082] Introduction to Homotopy Type Theory - arXiv
    Dec 21, 2022 · This is an introductory textbook to univalent mathematics and homotopy type theory, a mathematical foundation that takes advantage of the structural nature of ...Missing: scholarly | Show results with:scholarly
  4. [4]
    [PDF] homotopy theory for beginners
    Definition 2.1. The maps f0 and f1 are homotopic, f0 ≃ f1, if there exists a map, a homotopy, F : X × I → Y such that f0(x) = F(x, 0) and f1(x) = F(x, 1) for ...<|control11|><|separator|>
  5. [5]
    [PDF] Algebraic Topology I: Lecture 5 Homotopy, Star-shaped Regions
    A space X is contractible if the map X → ∗ is a homotopy equivalence. ... and important class of homotopy equivalences is given by the following definition.
  6. [6]
    [PDF] algebraic topology - UChicago Math
    As its name suggests, a deformation retract is a continuous map that deforms a space into one of its subspaces. Definition 2.3. A deformation retract from a ...<|control11|><|separator|>
  7. [7]
    [PDF] HOMOTOPIES AND DEFORMATION RETRACTS - UNT Digital Library
    Definition 3.2: A retract D of a space A is a deformation retract if and only if a retraction r : A - D is homotopic to idA by some homotopy,6. The retraction r ...
  8. [8]
    [PDF] Notes on Algebraic Topology
    The relation of homotopy between maps f: X→Y is reflexive, symetric and transitive. PROOF: The relation is reflexive; for any f: Y is homotopic to itself under ...
  9. [9]
    Homotopical Algebra - SpringerLink
    Free delivery 14-day returnsDownload chapter PDF · Axiomatic homotopy theory. Daniel G. Quillen. Pages 1-64. Examples of simplicial homotopy theories. Daniel G. Quillen. Pages 65-155. Back ...
  10. [10]
    [PDF] the quillen model category of topological spaces - MIT Mathematics
    Sep 11, 2017 · ... cofibrations have the left lifting property with respect to all trivial fibra- tions, and. • fibrations have the right lifting property ...
  11. [11]
    [PDF] The Idea of the Fundamental Group - Cornell Mathematics
    This first chapter introduces one of the simplest and most important functors of algebraic topology, the fundamental group, which creates an algebraic image of ...
  12. [12]
    [PDF] Papers on Topology - School of Mathematics
    Jul 31, 2009 · ... Poincaré before topology. In the introduction to his first major topology paper, the Analysis situs, Poincaré. (1895) announced his goal of ...
  13. [13]
    [PDF] Beiträge zur Topologie der Deformationen. (II. Homotopie
    - Beiträge zur Topologie der Deformationen (11. ... HUREWICZ \Commu- nicated by Prof. L. E. J. BRouwER). (Communicated at the meeting of April 27. 1935).
  14. [14]
    [PDF] Homotopy theory begins with the homotopy groups πn(X ... - UiO
    Homotopy theory begins with the homotopy groups πn(X), which are the nat- ural higher-dimensional analogs of the fundamental group. These higher homotopy ...<|control11|><|separator|>
  15. [15]
    [PDF] freudenthal suspension theorem - UChicago Math
    In this paper, I will prove the Freudenthal suspension theorem, and use that to explain what stable homotopy groups are. All the results stated in this paper ...
  16. [16]
    [PDF] 1 Whitehead's theorem.
    Relative cellular approximation: Any map f : (X, A) → (Y,B) of CW pairs has a cellular approximation by a homotopy through such maps of pairs. Proof: First ...Missing: source | Show results with:source
  17. [17]
    [PDF] A primer on homotopy colimits - University of Oregon
    Part 1 of the paper (Sections 2–6) develops the basic definition of homotopy colimits and limits, as well as some foundational properties. Every- thing is done ...
  18. [18]
    [PDF] Homotopy (limits and) colimits - Emily Riehl
    We will define the homotopy colimit of an ordinary functor F : C → M, where C is an arbitrary small category.
  19. [19]
    [PDF] Notes on principal bundles and classifying spaces
    This elegant paper is independent of “Universal bundles I”. Very readable. [Steenrod] The topology of fibre bundles. A classic by one of the ancient masters ...<|control11|><|separator|>
  20. [20]
    Classifying Spaces Made Easy - UCR Math Department
    Here's some stuff which should help you understand some fundamental topics in algebraic topology: classifying spaces and their relation to Eilenberg-Mac Lane ...Missing: sources | Show results with:sources
  21. [21]
    [PDF] bundles, classifying spaces and characteristic classes
    In this section we will discuss the homotopy clas- sification of principal bundles. ... We say that BG is a classifying space for principal G-bundles.
  22. [22]
    [PDF] Stable Algebraic Topology, 1945-1966 - UChicago Math
    Hurewicz. Beiträge zur Topologie der Deformationen. Nederl. Akad. Wetensch. Proc. Ser. A 38(1935), 112-119, 521-528; 39(1936), 117-126, 213-224. [Ja55]. I.M. ...<|control11|><|separator|>
  23. [23]
    [PDF] The Steenrod algebra - Purdue Math
    Jan 25, 2016 · A natural question at this point is: is there a stable cohomology operation, HnX → Hn+kX for varying n, that is the cup product square on Hk?
  24. [24]
    [PDF] Handbook of Homotopy Theory
    We must determine primary operations in the MU-dual Steenrod algebra. Some ... tionship between homotopy operations and TAQ-cohomology operations. Give.
  25. [25]
    [PDF] Stable homotopy and generalised homology
    Stable Homotopy and Generalised Homology, by J.F. Adams (1974). The University of Chicago Press, Chicago 60637. The University of Chicago Press, Ltd., London.
  26. [26]
    [PDF] J. Peter May - E, Ring Spaces and E, Ring Spectra - UChicago Math
    Thus the usual notion of spectrum implicitly refers to a fixed chosen basis for RCO. Many very real difficulties in the homotopy theory of spectra, in.
  27. [27]
    [PDF] 1. Obstruction Theory
    INCOMPLETE AND UPDATING! 1. Obstruction Theory. A fundamental question in topology is how to compute the homotopy classes of maps between two spaces.
  28. [28]
    [PDF] Introduction to higher homotopy groups and obstruction theory
    Feb 17, 2011 · Higher homotopy groups (πk) are generalizations of the fundamental group, defined as homotopy classes of maps from (Ik,∂Ik) to (X, x0), where I ...
  29. [29]
    [PDF] Characteristic Classes
    Sep 28, 2021 · The text which follows is based mostly on lectures at Princeton University in. 1957. The senior author wishes to apologize for the delay in ...
  30. [30]
    [PDF] Homotopal localizations of spaces
    The main theorem of this paper shows that f-localization functors always preserve fiber sequences of connected H-spaces up to small error terms. For instance, ...
  31. [31]
    Homotopy Limits, Completions and Localizations - SpringerLink
    The main purpose of part I of these notes is to develop for a ring R a functional notion of R-completion of a space X.
  32. [32]
    [PDF] Lecture Notes in Mathematics - On-Line.It
    Chapter I. Chapter II. Contents. Completions and localizations. §O. Introduction ... Homotopy direct limits .....••.•••••••.•••••.• 327. §3. Examples and ...
  33. [33]
    [PDF] Infinitesimal computations in topology - Numdam
    This paper was written in the effort to understand the nature of the mathematical object presented by a diffeomorphism class of compact smooth manifolds.
  34. [34]
    [PDF] Complex Cobordism and Stable Homotopy Groups of Spheres
    May 1, 2025 · Page 1. Complex Cobordism and. Stable Homotopy Groups of Spheres. Douglas C. Ravenel. Department of Mathematics, University of Rochester ...
  35. [35]
    [PDF] arXiv:1801.07530v2 [math.AT] 9 May 2018
    May 9, 2018 · The aim of this paper is to introduce the reader to enough of the machinery of spectra, the Steenrod algebra and the Adams spectral sequence to ...
  36. [36]
    [PDF] arXiv:1205.6668v3 [math.AT] 7 Apr 2014
    Apr 7, 2014 · The most fundamental result in this direction is the Freudenthal Suspension Theorem, which gives a range of dimensions in which the homotopy ...
  37. [37]
    [PDF] arXiv:1706.06846v2 [math.KT] 4 May 2022
    May 4, 2022 · Boardman [7, 5.10] defines conditional convergence of spectral sequences in ... may be studied in the homotopy category of filtered spectra.
  38. [38]
    [PDF] An introduction to filtered and synthetic spectra - arXiv
    Sep 25, 2025 · Complex Cobordism and Stable Homotopy Groups of Spheres. 2nd ed. AMS Chelsea Publishing v. 347. AMS Chelsea Pub, 2004. 395 pp. DOI: 10.1090 ...
  39. [39]
    [PDF] Lectures on equivariant stable homotopy theory
    We review some foundations for equivariant stable homotopy theory in the context of orthogonal G- spectra. The main reference for this theory is the AMS ...
  40. [40]
    Transformation Groups and Representation Theory - SpringerLink
    Free delivery 14-day returnsBook Title: Transformation Groups and Representation Theory · Authors: Tammo Dieck · Series Title: Lecture Notes in Mathematics · Publisher: Springer Berlin, ...
  41. [41]
    [PDF] arXiv:2411.18561v1 [math.AT] 27 Nov 2024
    Nov 27, 2024 · Since the geometric realization of each of these simplicial sets results in a contractible space, the above functor induces a weak homotopy ...
  42. [42]
    [PDF] Homotopical Algebra - Aareyan Manzoor's website
    Homotopical algebra or non-linear homological algebra is the generalization of homological algebra to arbitrary categories which results by considering a.
  43. [43]
    [PDF] Higher topos theory / Jacob Lurie
    Let X be a nice topological space (for example, a CW complex). One goal of algebraic topology is to study the topology of X by means of algebraic.
  44. [44]
    [PDF] Quillen model categories
    Quillen generalized this by requiring that every fibration in a model category have the right lifting property with respect to every trivial cofibration.
  45. [45]
    [math/0608040] Higher Topos Theory - arXiv
    Aug 2, 2006 · This purpose of this book is twofold: to provide a general introduction to higher category theory (using the formalism of quasicategories or weak Kan complexes)Missing: ∞- | Show results with:∞-
  46. [46]
    [PDF] Derived Algebraic Geometry I: Stable ∞-Categories
    Oct 8, 2009 · The collection of spectra can be organized into a stable ∞-category Sp. Moreover, Sp is in some sense the universal example of a stable ∞- ...
  47. [47]
    [PDF] Higher Algebra
    Sep 18, 2017 · The ∞-category Sp of spectra is an example of a stable ∞-category. In fact, it is universal among stable ∞-categories: we will show that ...
  48. [48]
    [PDF] Pursuing Stacks - GitHub Pages
    Page 1. PURSUING STACKS. (À la poursuite des Champs). First episode: THE MODELIZING STORY. (histoire de modèles) by Alexander Grothendieck. 1983. Page 2. ii.
  49. [49]
    [PDF] Lectures on n-category and cohomology - Department of Mathematics
    Hypothesis 1 (Grothendieck's Dream, aka the Homotopy Hypothe- sis). n-Groupoids are the same as homotopy n-types. Here 'n-groupoids' means weak n-groupoids, in ...
  50. [50]
    [PDF] Weak Omega Categories I - arXiv
    We show that the Baez-Dolan stabilization hypothesis is a direct consequence of our definition of weak omega categories. We define a natural notion of a pseudo- ...
  51. [51]
    [math/9905211] Surgery and duality - arXiv
    May 1, 1999 · Surgery, as developed by Browder, Kervaire, Milnor, Novikov, Sullivan, Wall and others is a method for comparing homotopy types of topological ...
  52. [52]
    [PDF] ON THE EXISTENCE AND CLASSIFICATION OF DIFFERENTIABLE ...
    An embedding is an immersion which is l- 1. If f and g are immersions of X in X”, a regular homotopy connectingfto g is a differentiable homotopy. F: X ...
  53. [53]
    [PDF] Immersions of manifolds and homotopy theory - Stanford Mathematics
    Jun 30, 2022 · We begin in Section 1 by discussing how vector bundle theory can be used to find obstructions to the existence of embeddings and immersions.
  54. [54]
    [math/0303109] Ricci flow with surgery on three-manifolds - arXiv
    Mar 10, 2003 · This is a technical paper, which is a continuation of math.DG/0211159. Here we construct Ricci flow with surgeries and verify most of the assertions.<|separator|>
  55. [55]
    [2108.03302] Diffeomorphism groups of prime 3-manifolds - arXiv
    Aug 6, 2021 · We show that the diffeomorphism group \text{Diff}(X) deformation retracts to the isometry group \text{Isom}(X). Combining this with earlier work ...
  56. [56]
    On the finiteness of the classifying space of diffeomorphisms of ...
    Apr 28, 2025 · ... homotopy type of diffeomorphism group of prime manifolds. In the reducible case, the prime decomposition theorem cuts the manifold along ...
  57. [57]
    The homotopy theory of dg-categories and derived Morita theory
    Aug 24, 2004 · Abstract: The main purpose of this work is the study of the homotopy theory of dg-categories up to quasi-equivalences.
  58. [58]
    [PDF] Voevodsky's Nordfjordeid Lectures: Motivic Homotopy Theory
    This text is a report from Voevodsky's summer school lectures on motivic homotopy in Nordfjordeid. Its first part consists of a leisurely introduction to.
  59. [59]
    [PDF] Super symmetric field theories and integral modular functions
    In addition, we explain the relevance of this result for our program of relating 2-dimensional quantum field theories to the topological modular form spectrum ...
  60. [60]
    The HoTT Book | Homotopy Type Theory
    The present book is intended as a first systematic exposition of the basics of univalent foundations, and a collection of examples of this new style of ...