Fact-checked by Grok 2 weeks ago

Homotopy group

In , homotopy groups are algebraic invariants associated to a pointed (X, x_0), generalizing the \pi_1(X, x_0) to higher dimensions by classifying continuous maps from the n- S^n to X that send a basepoint of S^n to x_0, up to basepoint-preserving equivalence, with the group operation defined by concatenating maps along the equator of the sphere. The nth homotopy group, denoted \pi_n(X, x_0), captures obstructions to contracting n-dimensional spheres in X, thereby detecting higher-dimensional "holes" in the space. For n = 1, \pi_1(X, x_0) is the , which may be non-abelian and classifies loops based at x_0 up to . For n \geq 2, \pi_n(X, x_0) is always abelian, and the groups are independent of the choice of basepoint if X is path-connected. These groups are homotopy invariants, meaning they remain unchanged under weak homotopy equivalences, making them powerful tools for classifying topological spaces up to homotopy type. The concept was first proposed by Eduard Čech in 1932 but was rigorously defined and developed by in 1935, who showed that \pi_n(X) vanishes if and only if X is n-connected and established their abelian nature for n \geq 2. also introduced the Hurewicz homomorphism relating to , which is an under certain connectivity conditions, linking to . Homotopy groups play a central role in modern , appearing in the study of via long exact sequences—such as \dots \to \pi_n(F) \to \pi_n(E) \to \pi_n(B) \to \pi_{n-1}(F) \to \dots for a F \to E \to B—and in theorems like Whitehead's, which characterizes weak equivalences between simply connected spaces. groups is notoriously difficult, especially for spheres, where \pi_n(S^k) is trivial for n < k but highly nontrivial and finite for n > k, with ongoing research revealing their structure through spectral sequences and .

Fundamentals

Introduction

Homotopy groups form a sequence of algebraic structures associated to a topological space, serving as invariants that capture essential features of its topology, particularly the presence of "holes" in various dimensions. These groups generalize the intuitive notion of connectivity, providing a framework to distinguish spaces that cannot be continuously deformed into one another—known as homotopy equivalence—and thus play a central role in the classification of topological spaces within algebraic topology. By quantifying obstructions to contracting maps from spheres into the space, homotopy groups reveal deformation properties that simpler invariants, like homology groups, may overlook. The development of homotopy groups traces back to the early 20th century, building on Henri Poincaré's introduction of the in his 1895 paper "Analysis Situs," where he analyzed 1-dimensional loops to detect basic holes in manifolds. This laid the groundwork for , but it was who, building on Eduard Čech's 1932 proposal, rigorously defined higher-dimensional homotopy groups in his 1935 paper "Beiträge zur Topologie der Deformationen I. Höherdimensionale Homotopiegruppen," extending the concept to classify more intricate topological features beyond mere path-connectedness. Hurewicz's innovation shifted the focus from qualitative descriptions to algebraic invariants, profoundly influencing subsequent advances in the field. A key distinction in homotopy groups is their algebraic behavior: the first homotopy group π₁ is non-abelian, reflecting the non-commutative of loop compositions in a , whereas higher homotopy groups πₙ for n ≥ 2 are abelian, allowing for simpler group-theoretic analysis. In many significant examples, such as the , these invariants are finitely generated abelian groups, highlighting their computational tractability while underscoring their importance in probing the subtle, higher-dimensional structures that define a 's homotopy type.

Definition

A homotopy group is defined in the context of pointed topological spaces. A pointed topological space is a pair (X, x_0), where X is a and x_0 \in X is a designated basepoint. The higher homotopy groups were first proposed by Eduard Čech in 1932 and rigorously defined by in 1935, building on the . The loop space \Omega X of a pointed space (X, x_0) consists of all based loops in X, defined as the set of continuous maps f: (I, \partial I) \to (X, x_0), where I = [0,1] is the unit interval and \partial I = \{0,1\}. Two based loops f, g: I \to X are , denoted f \simeq g, if there exists a continuous H: I \times I \to X such that H(s, 0) = f(s), H(s, 1) = g(s), and H(0, t) = H(1, t) = x_0 for all s, t \in I. The first group, or , is \pi_1(X, x_0) = \pi_0(\Omega X), the set of path components ( classes) of the space \Omega X, equipped with the group operation induced by of loops: for homotopic classes and, the product is [f \cdot g], where (f \cdot g)(s) = f(2s) for s \in [0, 1/2] and (f \cdot g)(s) = g(2s - 1) for s \in [1/2, 1]. This group structure arises naturally from the topological of , with at x_0 serving as the identity. For n \geq 2, the nth homotopy group \pi_n(X, x_0) is defined as the set of pointed homotopy classes of continuous maps [S^n, X]_*, where S^n is the n-sphere with basepoint (the north pole, say), and maps send the basepoint of S^n to x_0. Equivalently, \pi_n(X, x_0) = \pi_0(\Omega^n X), the path components of the n-fold iterated loop space, where \Omega^1 X = \Omega X and \Omega^{k+1} X = \Omega(\Omega^k X) for k \geq 1. The group operation on \pi_n(X, x_0) for n \geq 2 is induced by the cogroup structure on S^n, given by the pinch map that combines two maps via the equatorial decomposition of S^n, resulting in an abelian group. Hurewicz showed that these groups are abelian for n \geq 2. If X is path-connected, the homotopy groups \pi_n(X, x_0) are independent of the choice of basepoint x_0, up to canonical isomorphism induced by paths connecting basepoints. More generally, if pointed spaces (X, x_0) and (Y, y_0) are equivalent via a pointed map f: (X, x_0) \to (Y, y_0) with pointed inverse, then f induces group isomorphisms \pi_n(X, x_0) \cong \pi_n(Y, y_0) for all n \geq 1.

Geometric and Algebraic Properties

Relation to holes

Homotopy groups provide a geometric means to detect and classify "holes" in topological spaces, where a hole in dimension n corresponds to the inability to continuously deform an n-dimensional sphere embedded in the space to a point. The nth homotopy group \pi_n(X, x_0) at basepoint x_0 \in X captures these features by classifying pointed maps from the n-sphere S^n into X up to homotopy, with non-trivial elements indicating the presence of such holes. The 0th homotopy set \pi_0(X), which is the set of path components of X, detects 0-dimensional holes, manifesting as disconnectedness in the space. Each component represents a distinct "piece" that cannot be connected by continuous paths, thus quantifying the space's overall fragmentation. In dimension 1, the \pi_1(X, x_0) identifies 1-dimensional holes through loops based at x_0 that cannot be contracted to a point. For instance, S^1 has \pi_1(S^1) \cong \mathbb{Z}, generated by the loop winding once around the circle, reflecting the single 1-dimensional hole inherent to its . For n \geq 2, the higher \pi_n(X, x_0) reveal n-dimensional holes via maps S^n \to X that resist to a . These groups are always abelian and detect more subtle voids; notably, the 2- S^2 has trivial \pi_1(S^2) but \pi_2(S^2) \cong \mathbb{Z}, indicating no 1-dimensional holes yet a fundamental 2-dimensional hole filled by the on S^2. In general, the n- S^n exhibits \pi_n(S^n) \cong \mathbb{Z}, capturing its own n-dimensional hole, while lower homotopy groups vanish for i < n. A concrete example is the torus T^2, which has \pi_1(T^2) \cong \mathbb{Z} \oplus \mathbb{Z}, corresponding to two independent 1-dimensional holes (one along each generating loop), while all higher homotopy groups \pi_n(T^2) = 0 for n \geq 2, indicating no higher-dimensional holes. Vanishing theorems further illuminate this detection: a space X is simply connected if it is path-connected and \pi_1(X) = 0, meaning no 1-dimensional holes obstruct loop contractions. More generally, X is k-connected if \pi_i(X) = 0 for all i \leq k, signifying the absence of holes in dimensions up to k.

Basic algebraic structure

Homotopy groups possess a rich algebraic structure, beginning with their functorial nature. A continuous map f: (X, x_0) \to (Y, y_0) between pointed topological spaces induces a group homomorphism f_*: \pi_n(X, x_0) \to \pi_n(Y, y_0) for each n \geq 1, preserving the group operation and satisfying (f \circ g)_* = f_* \circ g_* and \mathrm{id}_* = \mathrm{id}. This makes the assignment (X, x_0) \mapsto \pi_n(X, x_0) a functor from the category of pointed spaces to the category of groups, with composition and identities preserved. Homotopy invariance ensures that the algebraic structure is robust under deformation. If two maps f, g: (X, x_0) \to (Y, y_0) are homotopic relative to the basepoint, then f_* = g_* on all homotopy groups \pi_n. Consequently, homotopy equivalences between pointed spaces induce isomorphisms on all \pi_n, providing a way to classify spaces up to homotopy via their homotopy groups. In particular, the fundamental group \pi_1 may be non-abelian, reflecting complex looping behaviors, whereas higher homotopy groups \pi_n for n \geq 2 are always abelian. Exact sequences arise in specific contexts, such as covering spaces, illustrating the interplay between homotopy and group extensions. For a path-connected regular covering space projection p: (\tilde{X}, \tilde{x}_0) \to (X, x_0) with discrete fiber, the induced map p_*: \pi_1(\tilde{X}, \tilde{x}_0) \to \pi_1(X, x_0) is injective, and there is a short exact sequence $1 \to \pi_1(\tilde{X}, \tilde{x}_0) \to \pi_1(X, x_0) \to G \to 1, where G is the deck transformation group acting freely and properly discontinuously on \tilde{X}. For n \geq 2, p_* induces isomorphisms \pi_n(\tilde{X}, \tilde{x}_0) \cong \pi_n(X, x_0), showing that higher homotopy detects the same "holes" in base and cover. The Hurewicz homomorphism provides a bridge to homology theory, encoding homotopy information algebraically. For a path-connected pointed space X, the n-th Hurewicz map h_n: \pi_n(X, x_0) \to H_n(X; \mathbb{Z}) sends a homotopy class represented by a map f: (S^n, *) \to (X, x_0) to the homology class of its image, using the fundamental class of the sphere. This map is surjective for all n \geq 1, and if X is simply connected, it is an isomorphism for n = 2; more generally, for an (n-1)-connected space, h_n is an isomorphism onto the first nonzero homology group. The Whitehead theorem ties homotopy groups to the global topology of CW-complexes. A map f: X \to Y between connected CW-complexes that induces isomorphisms f_*: \pi_n(X, x_0) \to \pi_n(Y, y_0) for all n \geq 1 is a homotopy equivalence. Equivalently, a weak homotopy equivalence between such spaces is a genuine homotopy equivalence, emphasizing the role of CW-structure in detecting homotopy types through algebraic invariants.

Exact Sequences and Fibrations

Long exact sequence of a fibration

In algebraic topology, a Serre fibration is a continuous map p: E \to B between topological spaces that satisfies the homotopy lifting property with respect to all inclusions \partial I^n \hookrightarrow I^n, where I^n denotes the n-dimensional cube (or equivalently, the n-disk D^n) and \partial I^n its boundary, for all n \geq 0. Specifically, for any map f: \partial I^n \to E and homotopy H: I^n \times I \to B such that p \circ f is homotopic to the restriction of H to \partial I^n \times \{0\} relative to \partial I^n \times \{0\}, there exists a lift \tilde{H}: I^n \times I \to E extending f on \partial I^n \times \{0\} and projecting to H via p. This property ensures that fibrations behave well with respect to homotopy, allowing local properties of the fiber to relate to global homotopy invariants of the total space and base. Given a Serre fibration p: E \to B with basepoint b_0 \in B and fiber F = p^{-1}(b_0), assuming all spaces are path-connected and pointed appropriately, the homotopy groups of E, B, and F are connected by a long exact sequence. This sequence takes the form \cdots \to \pi_{n+1}(B) \xrightarrow{\partial} \pi_n(F) \to \pi_n(E) \to \pi_n(B) \to \pi_{n-1}(F) \to \cdots \to \pi_1(F) \to \pi_1(E) \to \pi_1(B) \to \pi_0(F) \to \pi_0(E) \to \pi_0(B) \to 0, which is exact at each term for n \geq 1, where exactness means the image of each map equals the kernel of the next. For n=0, the sequence terminates with \pi_0(B) \to 0, but \pi_0 groups are pointed sets rather than groups, so exactness is interpreted set-theoretically: the map \pi_0(F) \to \pi_0(E) is surjective onto the preimage under \pi_0(E) \to \pi_0(B) of the basepoint component. The maps \pi_n(F) \to \pi_n(E) and \pi_n(E) \to \pi_n(B) are induced by the inclusion F \hookrightarrow E and p: E \to B, respectively. The boundary map \partial: \pi_{n+1}(B) \to \pi_n(F) is the connecting homomorphism, which captures how homotopy classes in the base relate to those in the fiber via the fibration structure. A sketch of its derivation relies on the path-lifting property of : given a representative [ \alpha ] \in \pi_{n+1}(B) based at b_0, lift the corresponding (n+1)-sphere in B (or its cell decomposition) to a homotopy in E whose boundary traces a loop in F, yielding an element in \pi_n(F); the exactness follows from composing lifts and homotopies uniquely up to homotopy in E, B, and F. This construction ensures the sequence is long and exact, providing a tool to compute recursively from fiber, total space, and base. A special case arises for the trivial fibration p: F \times B \to B given by projection onto the second factor, where the fiber is F \times \{b_0\} \cong F. Here, the long exact sequence splits naturally, with \pi_n(E) \cong \pi_n(F) \oplus \pi_n(B) for all n \geq 1 via the product structure, and the boundary map \partial vanishes, reflecting the direct product decomposition. This splitting highlights how non-trivial fibrations introduce interactions between the homotopy of F and B through \partial.

Exact sequences for pairs and cofibrations

In algebraic topology, relative homotopy groups provide a way to capture the homotopy information of a space X relative to a subspace A \subset X. For a pair of pointed spaces (X, A) with basepoint in A, the nth relative homotopy group \pi_n(X, A) is defined as the set of homotopy classes of pointed maps (I^n, \partial I^n, J^{n-1}) \to (X, A, *), where I^n is the n-dimensional cube, \partial I^n is its boundary, and J^{n-1} = \partial I^{n-1} \times I \cup I^{n-1} \times \{0\} is the relevant face for the basepoint; these classes form a group under pointwise concatenation for n \geq 2, which is abelian for n \geq 3. Equivalently, \pi_n(X, A) can be viewed as the (n-1)th homotopy group of the path space of maps from the basepoint to A. A key tool for relating absolute and relative homotopy groups is the notion of a cofibration, which ensures that homotopies defined on a subspace extend well-behavedly to the whole space. A continuous map i: A \to X is a cofibration if it has the homotopy extension property: for every topological space Y, every continuous map u: X \to Y, and every continuous homotopy H: A \times I \to Y such that H(a, 0) = u(i(a)) for all a \in A, there exists a continuous homotopy \tilde{H}: X \times I \to Y such that \tilde{H}|_{A \times I} = H and \tilde{H}(x, 0) = u(x) for all x \in X. This property holds, for example, for cell inclusions in CW-complexes. In categories such as compactly generated Hausdorff spaces, cofibrations are equivalent to the pair (X, A) being a neighborhood deformation retract pair, where there exists an open neighborhood V of A in X and a deformation retraction of V onto A. When i: A \to X is a cofibration, there is a long exact sequence in : \cdots \to \pi_n(A) \xrightarrow{i_*} \pi_n(X) \xrightarrow{j_*} \pi_n(X, A) \xrightarrow{\partial} \pi_{n-1}(A) \to \cdots, where i_* and j_* are induced by the inclusion and quotient maps, respectively, and the boundary map \partial is defined by composing a relative class with the quotient map X/A \simeq X \cup_A CA (the mapping cone on A) and restricting to the top face of the cube. This sequence is exact at each term, meaning the image of each map equals the kernel of the next, and it arises from the cofiber sequence A \to X \to X/A. The exactness implies that relative homotopy groups measure the "new" homotopy classes introduced by adjoining X to A. The cofiber sequence provides a dual perspective: for a cofibration i: A \to X, the cofiber C_i = X \cup_A CA, where CA = A \times I / A \times \{1\} is the cone on A, fits into the exact sequence of homotopy groups \pi_*(A) \to \pi_*(X) \to \pi_*(C_i) \to \pi_{*-1}(A) \to \cdots, which is a segment of the long exact sequence for the pair (C_i, X), with C_i / X \simeq \Sigma A (the suspension of A). This connects relative homotopy to absolute homotopy of quotient spaces and is particularly useful for computations involving cell attachments. To preserve exactness under natural transformations or maps between such sequences, the five lemma applies: given a commutative diagram of abelian groups with exact rows \begin{CD} 0 @>>> A_1 @>f_1>> B_1 @>g_1>> C_1 @>>> 0 \\ @. @VVV @VVf_2V @VVg_2V @. \\ 0 @>>> A_2 @>>h_2> B_2 @>>k_2> C_2 @>>> 0, \end{CD} if f_1 and g_2 are isomorphisms (or more generally, if the outer maps are iso and middles mono/epi in certain combinations), then f_2 and g_1 are isomorphisms; this lemma extends to the ends of long exact sequences in , ensuring isomorphisms in relative settings like excisions.

Key Examples and Applications

Homotopy groups of spheres

The homotopy groups of spheres, denoted \pi_n(S^k), vanish for n < k, are isomorphic to \mathbb{Z} when n = k, and become increasingly complex for n > k, capturing essential features of higher-dimensional . These groups are fundamental in , as spheres serve as building blocks for more general spaces, and their computation reveals patterns of stability under . In low dimensions, explicit computations yield \pi_1(S^1) \cong \mathbb{Z}, \pi_2(S^2) \cong \mathbb{Z}, \pi_3(S^2) \cong \mathbb{Z}, \pi_3(S^3) \cong \mathbb{Z}, and \pi_4(S^3) \cong \mathbb{Z}_2. The S^1 \to S^3 \to S^2 provides a key example, where the long of the implies that the generator of \pi_3(S^2) corresponds to the Hopf map, establishing \pi_3(S^2) \cong \mathbb{Z}. The suspension map \Sigma: \pi_n(S^k) \to \pi_{n+1}(S^{k+1}) induces in a range determined by the Freudenthal suspension theorem, which states that it is an isomorphism for n < 2k - 1 and surjective for n = 2k - 1, assuming k \geq 2. This theorem highlights the stabilization phenomenon, where homotopy groups become independent of the base dimension in sufficiently high ranges. The stable homotopy groups of spheres are defined as \pi_n^s = \lim_{k \to \infty} \pi_{n+k}(S^k), capturing the behavior for n \geq k + 1. These groups are finite abelian for n > 0 (by the Serre finiteness theorem) and exhibit intricate torsion structures; for instance, \pi_1^s \cong \mathbb{Z}_2, \pi_2^s \cong \mathbb{Z}_2, and \pi_3^s \cong \mathbb{Z}_{24}. Computations of these groups up to dimension 90 rely on spectral sequences and motivic methods, underscoring their beyond low dimensions. The following table lists selected low-dimensional homotopy groups \pi_n(S^m) for m = 1 to $7 and n \leq 10, focusing on non-trivial cases (trivial groups are indicated as ); values are drawn from classical computations, with infinite cyclic groups denoted \mathbb{Z} and finite ones by their standard decompositions.
n \setminus m
\mathbb{Z}
\mathbb{Z}
\mathbb{Z}\mathbb{Z}
\mathbb{Z}_2\mathbb{Z}_2\mathbb{Z}
\mathbb{Z}_2\mathbb{Z}_2\mathbb{Z}_2\mathbb{Z}
\mathbb{Z}_{12}\mathbb{Z}_{12}\mathbb{Z}_2\mathbb{Z}_2\mathbb{Z}
\mathbb{Z}_2\mathbb{Z}_2\mathbb{Z} \oplus \mathbb{Z}_{12}\mathbb{Z}_2\mathbb{Z}_2\mathbb{Z}
\mathbb{Z}_2\mathbb{Z}_2\mathbb{Z}_2 \oplus \mathbb{Z}_2\mathbb{Z}_{24}\mathbb{Z}_2\mathbb{Z}_2
\mathbb{Z}_3\mathbb{Z}_3\mathbb{Z}_{24}\mathbb{Z}_2
\mathbb{Z}_{15} \oplus \mathbb{Z}_2\mathbb{Z}_{15} \oplus \mathbb{Z}_2\mathbb{Z}_2\mathbb{Z}_{24}

Homogeneous spaces and Lie groups

Homogeneous spaces arise as quotients G/H, where G is a and H is a closed , providing a natural setting for applying through associated fibrations. The projection map p: G \to G/H defines a locally trivial with H, inducing a long in : \cdots \to \pi_n(H) \to \pi_n(G) \to \pi_n(G/H) \to \pi_{n-1}(H) \to \cdots. This allows computation of of the base G/H using known groups of G and H, particularly when G and H are classical Lie groups. For the special orthogonal group \mathrm{SO}(n), which is a Lie group parametrizing rotations in n-dimensions, the fundamental group is \pi_1(\mathrm{SO}(n)) \cong \mathbb{Z}/2\mathbb{Z} for n \geq 3. Higher homotopy groups of \mathrm{SO}(n) stabilize for large n and relate to those of spheres via the double cover by the spin group \mathrm{Spin}(n), with \pi_3(\mathrm{SO}(n)) \cong \mathbb{Z} for n \geq 3. A concrete example is \mathrm{SO}(3), which is homotopy equivalent to the real projective space \mathbb{RP}^3, yielding \pi_1(\mathrm{SO}(3)) \cong \mathbb{Z}/2\mathbb{Z} and \pi_3(\mathrm{SO}(3)) \cong \mathbb{Z}. The U(n), consisting of unitary matrices preserving the Hermitian inner product, exhibits Bott periodicity in its groups. The stable unitary group U = \lim_{n \to \infty} U(n) is equivalent to the \Omega U, with odd-dimensional groups \pi_{2k-1}(U) \cong \mathbb{Z} for k \geq 1 and even-dimensional groups trivial. This 2-periodicity arises from the U(n) \to U(n+1) \to S^{2n+1} and stabilizes the groups as n increases. Flag manifolds and , as homogeneous spaces like the real \mathrm{Gr}_k(\mathbb{R}^n) \cong O(n)/(O(k) \times O(n-k)), have homotopy groups computed via successive fibrations, such as sphere bundles over lower Grassmannians. For instance, the fibration S^{n-k-1} \to \mathrm{Gr}_k(\mathbb{R}^n) \to \mathrm{Gr}_k(\mathbb{R}^{n-1}) induces exact sequences that reveal the groups, often linking to stable ranges of orthogonal or unitary groups. These computations highlight how homogeneous spaces encode geometric structures through their homotopy, with applications in vector bundle classification.

Projective spaces

The real projective space \mathbb{RP}^n admits a CW complex structure consisting of one open cell in each dimension from 0 to n. The 0-cell is a point, the 1-cell is attached via the constant map to form \mathbb{RP}^1 \cong S^1, and for k \geq 2, the k-cell is attached to the (k-1)-skeleton \mathbb{RP}^{k-1} via the double covering map S^{k-1} \to \mathbb{RP}^{k-1}, which has degree 2. This fibration perspective arises from viewing \mathbb{RP}^n as the quotient of the n-sphere by the antipodal action, yielding the principal \mathbb{Z}/2\mathbb{Z}-bundle S^n \to \mathbb{RP}^n with fiber \mathbb{Z}/2\mathbb{Z} \cong S^0. The associated long exact sequence in homotopy groups is \cdots \to \pi_k(S^0) \to \pi_k(S^n) \to \pi_k(\mathbb{RP}^n) \to \pi_{k-1}(S^0) \to \cdots. Since \pi_k(S^0) = 0 for all k \geq 1, the sequence simplifies for k \geq 2 to \pi_k(\mathbb{RP}^n) \cong \pi_k(S^n). In particular, for n \geq 2, the sequence at low dimensions gives \pi_2(\mathbb{RP}^n) \cong \mathbb{Z} and \pi_1(\mathbb{RP}^n) \cong \mathbb{Z}/2\mathbb{Z}. The complex projective space \mathbb{CP}^n possesses a CW complex structure with one cell in each even dimension from 0 to $2n. Specifically, \mathbb{CP}^0 is a point (the 0-cell), \mathbb{CP}^1 \cong S^2 (attaching a 2-cell to the point via the constant map), and for k \geq 2, the $2k-cell is attached to the skeleton \mathbb{CP}^{k-1} via a map S^{2k-1} \to \mathbb{CP}^{k-1} that represents the generator of H_2(\mathbb{CP}^{k-1}; \mathbb{Z}) \cong \mathbb{Z} under the Hurewicz homomorphism. This structure reflects the quotient construction \mathbb{CP}^n = S^{2n+1} / S^1, where S^1 acts by complex multiplication on coordinates. The homotopy groups of \mathbb{CP}^n are computed using the Hopf fibration S^1 \to S^{2n+1} \to \mathbb{CP}^n, a principal S^1-bundle. The long exact sequence in homotopy groups is \cdots \to \pi_{k+1}(\mathbb{CP}^n) \to \pi_k(S^1) \to \pi_k(S^{2n+1}) \to \pi_k(\mathbb{CP}^n) \to \pi_{k-1}(S^1) \to \cdots. Since \pi_k(S^1) = 0 for k \geq 2 and \pi_1(S^1) \cong \mathbb{Z}, while S^{2n+1} is $2n-connected, the sequence yields \pi_1(\mathbb{CP}^n) = 0, \pi_2(\mathbb{CP}^n) \cong \mathbb{Z}, \pi_k(\mathbb{CP}^n) = 0 for $3 \leq k \leq 2n, \pi_{2n+1}(\mathbb{CP}^n) \cong \mathbb{Z}, and \pi_k(\mathbb{CP}^n) \cong \pi_k(S^{2n+1}) for k > 2n+1. In the stable range k > 2n+1, these are the stable homotopy groups \pi_{k-(2n+1)}^s. The infinite projective spaces serve as classifying spaces in algebraic topology. The infinite real projective space \mathbb{RP}^\infty = \varinjlim \mathbb{RP}^n is the classifying space BO(1) for the group O(1) \cong \mathbb{Z}/2\mathbb{Z}, hence an K(\mathbb{Z}/2\mathbb{Z}, 1) with \pi_1(\mathbb{RP}^\infty) \cong \mathbb{Z}/2\mathbb{Z} and \pi_k(\mathbb{RP}^\infty) = 0 for k \geq 2. Likewise, \mathbb{CP}^\infty = \varinjlim \mathbb{CP}^n is the classifying space BU(1) for U(1) \cong S^1, hence K(\mathbb{Z}, 2) with \pi_2(\mathbb{CP}^\infty) \cong \mathbb{Z} and \pi_k(\mathbb{CP}^\infty) = 0 for k \neq 2. The inclusions \mathbb{RP}^n \to \mathbb{RP}^\infty and \mathbb{CP}^n \to \mathbb{CP}^\infty induce isomorphisms on homotopy groups up to dimension n and $2n, respectively, with higher groups mapping to the trivial groups of the limits.

Computation Techniques

General methods for calculation

One foundational approach to homotopy groups involves the Postnikov tower of a topological space X, which decomposes X into a sequence of stages based on its homotopy groups. The tower consists of fibrations X_n \to X_{n-1} for n \geq 2, where each X_n is an n-stage Postnikov approximation with \pi_k(X_n) \cong \pi_k(X) for k \leq n and \pi_k(X_n) = 0 for k > n, connected by k-invariants in H^{n+1}(X_{n-1}; \pi_n(X)). This allows step-by-step starting from the \pi_1(X), building higher stages inductively by attaching cells or using principal fibrations over Eilenberg-MacLane spaces. Eilenberg-MacLane spaces K(\pi, n), where \pi is an for n \geq 2, are connected CW-complexes characterized by \pi_n(K(\pi, n)) \cong \pi and \pi_k(K(\pi, n)) = 0 for k \neq n. These spaces classify groups, as [X, K(\pi, n)] \cong H^n(X; \pi) for suitable X, and serve as building blocks in Postnikov towers: the fiber of the map X_n \to X_{n-1} is homotopy equivalent to K(\pi_n(X), n), with the connecting map given by the k-invariant. Constructions include classifying spaces for discrete groups (K(G,1) = BG) or infinite projective spaces for \mathbb{Z}/2 coefficients. The provides a tool for computing the groups of the total space in a F \to E \to B, where F is path-connected and the fibration satisfies conditions such as the base B having finitely generated or the coefficients being a . The E_2-page is given by E_2^{p,q} = H_p(B; H_q(F; \mathbb{Z})), converging to H_{p+q}(E; \mathbb{Z}), which can then inform groups via the or further spectral sequences under connectivity assumptions on F and B. For simply connected fibrations, the sequence simplifies and often collapses to yield explicit isomorphisms in low degrees. Obstruction theory addresses the problem of lifting s or extending partial s between spaces by analyzing cohomological barriers in Postnikov towers. Given a f: X^{(n-1)} \to Y defined on the (n-1)- of X, extension to the n- exists up to if and only if the obstruction class in H^{n+1}(X^{(n)}; \pi_n(Y)) vanishes; for s between Postnikov stages, lifting through Y_k \to Y_{k-1} is obstructed by classes in H^{k+1}(X; \pi_k(Y)). This framework, applicable to classifying s into Eilenberg-MacLane spaces, reduces set computations to calculations and is particularly effective for simply connected targets. For triads (X; A, B) where A and B are subspaces with X = A \cup B, the Blakers-Massey exact sequence provides a long exact sequence relating relative homotopy groups, valid in a range determined by the connectivities of A, B, and their intersection. This excision theorem generalizes the exact sequence of a pair and is crucial for computations in pushouts or cofiber sequences, such as suspensions, by identifying relative groups in terms of absolute ones: under suitable connectivity assumptions, \pi_n(A \cap B, A) \to \pi_n(X, B) is an isomorphism for n below the connectivity threshold.

Advanced methods and known results

One of the most powerful tools for computing the p-primary components of the stable homotopy groups of spheres is the Adams spectral sequence, introduced by J. Frank Adams in 1959. This spectral sequence converges to the p-local stable stems \pi_*(S)^{\wedge}_p, with its E_2-page given by the Ext groups \operatorname{Ext}_{\mathcal{A}_p}^{s,t}(\mathbb{F}_p, \mathbb{F}_p) over the Steenrod algebra \mathcal{A}_p, where the grading corresponds to the dimension t - s and filtration s. The sequence arises from a minimal resolution of the trivial module in the category of comodules over the dual Steenrod algebra, providing a systematic way to detect elements in homotopy via cohomology operations. Extensive computations using this sequence, often aided by computer algorithms, have resolved many differentials and permanent cycles in low dimensions. Bott periodicity provides a foundational result on the stable homotopy groups of the classical Lie groups, particularly the U and O. For the unitary group, the states that the stable homotopy groups \pi_k(U) are periodic with period 2 for k \geq 1, specifically \pi_{2m}(U) \cong \mathbb{Z} and \pi_{2m+1}(U) = 0. For the orthogonal group, the periodicity is 8, and the ring structure of the stable homotopy is \pi_*(O) \cong \mathbb{Z}_2[\theta_1, \theta_2, \dots] \oplus \mathbb{Z}[\alpha_1, \alpha_2, \dots], where the \theta_i generate the 2-torsion in even degrees and the \alpha_i generate the infinite cyclic groups in odd degrees starting from degree 3. This periodicity, proved using index theory and on loop spaces, underpins much of and has implications for computing homotopy groups of related spaces like Grassmannians. In the 1980s, Douglas Ravenel formulated a series of conjectures, known as the X(n) and Y(m), that impose bounds on the p-torsion in the stable , motivated by chromatic and the Adams-Novikov . The X(n) conjecture asserts that certain v_n-periodic homotopy classes vanish in stems up to specific dimensions, while Y(m) provides similar constraints on the image of the J-homomorphism in higher chromatic layers; several cases, such as X(1) through X(4) at odd primes, have been resolved affirmatively using techniques from elliptic cohomology and synthetic spectra. These conjectures highlight the finite nature of most torsion in stable stems and guide ongoing computations by predicting the chromatic filtration of elements. Mark Mahowald's contributions to the image of the J-homomorphism describe its p-primary component in the stable as arising from orthogonal representations, providing an exact determination of Im J up to high dimensions via the . Specifically, at odd primes p, the image consists of elements detected by certain monomials in the , excluding higher v_1-periodic families, while at p=2, it includes the η_j family but misses certain Toda brackets. This work, building on Adams' original computations, resolves the connective cover of the orthogonal spectrum and informs the structure of 2-primary stems. Snaith's theorem relates the stable homotopy groups of the sphere to complex cobordism and through a result identifying periodic complex with the sphere localized at the Bott element \beta \in \pi_2([BU](/page/BU)). It implies that the homotopy groups of the localized sphere \pi_*(S^0[\beta^{-1}]) are isomorphic to the stable stems tensored with \mathbb{Z}[\beta, \beta^{-1}]. This result facilitates computations in periodic settings and connects stable to via formal group laws. Known results for unstable homotopy groups of spheres include the fact that \pi_n(S^k) is finite for all n > k, except for the \pi_k(S^k) \cong \mathbb{Z}, with explicit generators given by the Hopf maps in low dimensions. In the stable regime, all \pi_n(S^k) for fixed k and large n have been computed up to dimension 90 using the over the complex numbers, revealing intricate patterns of torsion and infinite order elements like the Hopf invariant one classes. These computations, performed algorithmically, confirm the rarity of infinite cyclic summands beyond the stable stems and support conjectures on the growth of ranks.

Generalizations and Extensions

Relative homotopy groups

Relative homotopy groups generalize the absolute homotopy groups by considering maps that are fixed on a . For a pair of topological spaces (X, A) with A \subset X and a basepoint x_0 \in A, the nth relative group \pi_n(X, A, x_0) for n \geq 1 is defined as the set of homotopy classes of continuous maps f: (D^n, S^{n-1}) \to (X, A) such that f(s_0) = x_0 for some fixed basepoint s_0 \in S^{n-1}, where D^n is the n-dimensional disk and S^{n-1} = \partial D^n is its sphere. These classes form a group under the operation induced by pinching the equator of D^n \vee D^n to a point, with the constant map serving as the identity; the group is abelian for n \geq 2. A key feature of relative homotopy groups is their isomorphism with absolute homotopy groups of a quotient space: \pi_n(X, A, x_0) \cong \pi_n(X/A, *, [x_0]), where X/A is the quotient space obtained by collapsing A to a single basepoint *. This identification arises because maps from (D^n, S^{n-1}) to (X, A) correspond to maps from (D^n / S^{n-1}, *) to (X/A, *), up to . The long exact sequence of the pair (X, A) is given by \cdots \to \pi_n(A, x_0) \xrightarrow{i_*} \pi_n(X, x_0) \xrightarrow{j_*} \pi_n(X, A, x_0) \xrightarrow{\partial} \pi_{n-1}(A, x_0) \to \cdots, where i_* is induced by the inclusion A \hookrightarrow X, j_* by the projection X \twoheadrightarrow X/A, and the boundary map \partial sends a relative class $$ to the class of the restriction f|_{S^{n-1}}: (S^{n-1}, s_0) \to (A, x_0); this sequence is exact at every term. The excision property provides a powerful tool for computing relative groups. Excision in relative holds under suitable conditions; for example, if U is open in X such that \overline{U} \subset \operatorname{[int](/page/INT)}(A), then the inclusion (X \setminus U, A \setminus U) \to (X, A) induces isomorphisms \pi_n(X \setminus U, A \setminus U, x_0) \cong \pi_n(X, A, x_0) for all n \geq 1. More precisely, in the context of CW complexes, if X = A \cup C with A \cap C a subcomplex and the pair (A, A \cap C) m-connected while (C, A \cap C) k-connected, then the map \pi_i(A, A \cap C) \to \pi_i(X, C) is an isomorphism for i < m + k and surjective for i = m + k. From excision, one derives the Mayer-Vietoris sequence in . For a space X = U \cup V with U, V \subset X open and W = U \cap V, there is a long \cdots \to \pi_n(W, x_0) \xrightarrow{(i_*, j_*)} \pi_n(U, x_0) \oplus \pi_n(V, x_0) \xrightarrow{k_* - l_*} \pi_n(X, x_0) \to \pi_{n-1}(W, x_0) \to \cdots, where i_*: W \hookrightarrow U and j_*: W \hookrightarrow V induce the first map, and k_*: U \hookrightarrow X, l_*: V \hookrightarrow X the second; this sequence aids in inductive computations by relating local and global homotopy data. The boundary map \partial in the long exact sequence interprets relative classes in terms of absolute ones on the subspace A, effectively capturing how extensions from A to X fail; via the quotient isomorphism, it connects to the homotopy fiber of the projection X \to X/A. Additionally, the relative suspension theorem, a version of Freudenthal's result, states that if the pair (X, A) is n-connected (meaning \pi_k(X, A) = 0 for k \leq n), then the suspension homomorphism \Sigma: \pi_k(X, A, x_0) \to \pi_{k+1}(\Sigma X, \Sigma A, \Sigma x_0) is an isomorphism for k < 2n and surjective for k = 2n, where \Sigma denotes the reduced suspension. This stability range facilitates computations of higher homotopy groups from lower-dimensional data. Homotopy groups are connected to groups H_*(X; G), which are s measuring the number of n-dimensional holes in a X with coefficients in an G, via the Hurewicz homomorphism, though the precise relation is detailed elsewhere. In contrast, groups provide a coarser for compact Hausdorff spaces, capturing through inverse limits over open covers, and differ from singular cohomology by being insensitive to certain pathological features. The universal coefficient theorem relates and groups by splitting short exact sequences involving Ext and tensor products, providing a bridge to homotopy invariants without directly computing them. Topological K-theory introduces the group K^0(X) = [X, BU \times \mathbb{Z}], the of stable vector bundles over X, which represents homotopy classes of maps to the of the . This invariant connects to homotopy groups through Bott periodicity, which asserts that \pi_{2k}(U(n)) \cong \mathbb{Z} for large n and even dimensions, establishing a period-2 in the stable range K^{n}(X) \cong K^{n+2}(X). Shape theory extends homotopy invariants to non-locally nice spaces via approximations by polyhedra, where two compacta are shape equivalent if their fundamental pro-groups are pro-isomorphic, even if not homotopy equivalent. For instance, the Warsaw circle, a compact formed by adjoining the to the unit along its limit points, is not homotopy equivalent to the S^1—as it lacks a non-constant based at certain points—but is shape equivalent to S^1, with matching Čech homotopy pro-groups \check{\pi}_1(W) \simeq \mathbb{Z}. Pro-homotopy theory refines this for compacta by considering inverse systems of homotopy groups over nerves of open covers, yielding pro-groups that classify homotopy types up to proper homotopy equivalence, particularly useful for metric compacta where shape and pro-homotopy coincide. These systems capture the "net homotopy" of spaces without arcwise connectedness, extending classical homotopy to broader classes like continua. Bordism groups \Omega_*(X) classify oriented manifolds up to , represented as homotopy classes [M, X] of maps from closed n-manifolds M to X, and relate to via the Pontryagin-Thom construction, where \pi_*^s \cong \lim_n \pi_{n+k}(V_{n,k}) identifies bordism classes with stable maps. groups, dual in some senses, further link to generalized theories like , the complex bordism , whose homotopy groups encode stable stems. Key differences distinguish homotopy groups from these invariants: unlike homology, which satisfies H_n(X \sqcup Y) \cong H_n(X) \oplus H_n(Y), the homotopy groups of a disjoint union with basepoint in one component are isomorphic to those of that component: \pi_n(X \sqcup Y, x_0) \cong \pi_n(X, x_0) for n \geq 1. This highlights their sensitivity to path components and basepoint choice. Homotopy groups detect weak homotopy equivalences, preserving all \pi_n, while homology detects only acyclic maps, and cohomology often serves for obstruction theory in lifting problems, providing primary obstructions in Postnikov towers.

References

  1. [1]
    Homotopy Group -- from Wolfram MathWorld
    The homotopy groups generalize the fundamental group to maps from higher dimensional spheres, instead of from the circle.
  2. [2]
    homotopy group in nLab
    Sep 16, 2025 · A sequence of groups that generalise the fundamental group π 1 ( X , x ) \pi_1(X,x) to higher homotopies.Idea · Definition · Truncated and connected... · History
  3. [3]
    None
    ### Summary of Introduction and Definition of Higher Homotopy Groups by Hurewicz
  4. [4]
    Cinquième complément à l'Analysis situs - Zenodo
    Sep 21, 2018 · Cinquième complément à l'Analysis situs. Creators. Poincaré, M. H.. Description. n/a. Files. article.pdf. Files (2.9 MB). Name, Size, Download ...
  5. [5]
    None
    ### Summary of Homotopy Groups Definition from Hurewicz (1935)
  6. [6]
    [PDF] Groupes D'Homotopie Et Classes De Groupes Abeliens
    ANNALS OF MATHEMATICS. Vol. 58, No. 2, September, 1953. Printed in U.S.A.. GROUPES D'HOMOTOPIE ET CLASSES DE GROUPES ABPLIENS.
  7. [7]
    [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 ...
  8. [8]
    [PDF] Homotopy theory begins with the homotopy groups πn(X ... - UiO
    Homotopy Theory. Taking n = n′ in the proposition, we obtain also a functoriality property for n connected CW models. For example, a map X→X′ induces a map ...
  9. [9]
    [PDF] A Concise Course in Algebraic Topology J. P. May - UChicago Math
    Higher homotopy groups. The most basic invariants in algebraic topology are the homotopy groups. They are very easy to define, but very hard to compute. We ...
  10. [10]
    five lemma in nLab
    Sep 22, 2021 · The five lemma is one of the basic lemmas of homological algebra, useful for example in the construction of the connecting homomorphism in the homology long ...
  11. [11]
    [PDF] Homotopy Groups of spheres - UChicago Math
    Aug 11, 2011 · Abstract. We consider the elementary theory of higher homotopy groups through its application to the homotopy groups of spheres.
  12. [12]
    [PDF] Homotopy groups of spheres - MIT
    May 6, 2025 · Spheres are among the most fundamental topological spaces, and their homotopy groups help characterize the maps between spheres.
  13. [13]
    [PDF] Introduction to Homotopy Groups of Spheres - Derek Sorensen
    In this paper we will define a seqeuence of positively graded groups associated to any topological space X, and called the homotopy groups of X. These come ...
  14. [14]
    [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 ...
  15. [15]
    [PDF] Toward a Global Understanding of the Homotopy Groups of Spheres
    Even the stable groups are quite mysterious. The following table gives their values for small k. k. πS k k.
  16. [16]
    None
    ### Summary of Stable Homotopy Groups of Spheres (arXiv:2001.04247)
  17. [17]
    Stable homotopy groups of spheres - PNAS
    We discuss the current state of knowledge of stable homotopy groups of spheres. We describe a computational method using motivic homotopy theory.
  18. [18]
    [PDF] Homotopy Groups of Spheres
    See Composition methods in homotopy groups of spheres by H. Toda, Annals of Math. Studies 49 (Princeton. Press, 1962) for this table and much more. n = 1 n ...
  19. [19]
    [PDF] 4. Homogeneous spaces, Lie group actions - MIT OpenCourseWare
    Theorem 4.1. (i) Let G be a Lie group of dimension n and H ⊂ G a closed Lie subgroup of dimension k. Then the homogeneous space.
  20. [20]
  21. [21]
    [PDF] The Stable Homotopy of the Classical Groups
    HOMOTOPY OF CLASSICAL GROUPS. 315. The groups 7k(U) are 0, Z for k = 0, 1 ... , An application of the Morse theory to the topology of Lie groups, Bull.
  22. [22]
    [PDF] Relations Between Homology and Homotopy Groups of Spaces
    A. This paper is a continuation of an investigation, started by H. Hopf. [5][6], studying the influence of the fundamental group 7r,(X) on the homology.<|control11|><|separator|>
  23. [23]
    [PDF] Spectral Sequences
    There are many situations in algebraic topology where the relationship between certain homotopy, homology, or cohomology groups is expressed perfectly by an ...
  24. [24]
    [PDF] the homotopy groups of a triad i
    THEOREM 3.5-4. Each homotopy sequence of a triad is exact. (ie., the kernel of any homomorphism is precisely the image of the preceding homomorphism.).
  25. [25]
    [1801.07530] A Guide for Computing Stable Homotopy Groups - arXiv
    Jan 23, 2018 · We provide a working guide to the stable homotopy category, to the Steenrod algebra and to computations using the Adams spectral sequence.<|separator|>
  26. [26]
    [PDF] Glimpses of Bott periodicity
    While it was intended to understand geodesic motion, Bott employed this theory to show that the homotopy groups of the unitary group U, the orthogonal group O, ...
  27. [27]
    THE MATHEMATICAL WORK OF DOUGLAS C. RAVENEL
    So you can make a chart of the homotopy groups of spheres, and look at the highest filtration in the n-stem. Call it g(n), and you get a curve, which I like to ...
  28. [28]
    [PDF] Mark Mahowald's work on the homotopy groups of spheres
    Mahowald used the iterated loop structure to blow the low-dimensional class o up to high-dimensional homotopy on the cells of the infinite complex 259. The ...
  29. [29]
    Stable homotopy groups of spheres: From dimension 0 to 90 - arXiv
    Jan 13, 2020 · This paper computes classical and motivic stable homotopy groups of spheres from dimension 0 to 90, using motivic homotopy theory.
  30. [30]
    [PDF] Math 527 - Homotopy Theory Hurewicz theorem - University of Regina
    Mar 25, 2013 · The relative Hurewicz morphism is natural, and is a group homomorphism. Moreover, it is compatible with the long exact sequences in homotopy and ...
  31. [31]
    BOR SUKS SHAPE, THEORY - Project Euclid
    In 1968, K. Borsuk introduced a new branch of topology called shape theory. Like homotopy theory, shape theory is devoted to the study of global properties.
  32. [32]
    [PDF] A quick proof of the rational Hurewicz theorem and a computation of ...
    Apr 21, 2004 · Abstract. We give an elementary proof of the rational Hurewicz theorem and compute the rational cohomology groups of Eilenberg–MacLane ...<|control11|><|separator|>
  33. [33]
    [PDF] topological k-theory and bott periodicity
    An important fact about pullbacks is that the respect homotopy equivalences. That is, if f,g : Y → X are homotopic, then f∗E and g∗E are isomorphic as vector ...
  34. [34]
    [PDF] Bott periodicity - Columbia Math Department
    Originally, Bott computed the homotopy groups. πkU = πk colimn U(n) ... The Stable Homotopy of the Classical Groups. Annals of Math., Vol. 70, No. 2 ...
  35. [35]
    [PDF] arXiv:2307.12899v1 [math.GN] 24 Jul 2023
    Jul 24, 2023 · The Warsaw circle and the circle. S1 are examples of compact metrizable spaces which have different homotopy types but are shape equivalent. In ...
  36. [36]
    net homotopy for compacta(1) - jstor
    The important Hurewicz theory of homotopy groups(2) is applicable only to arc-wise connected spaces. If these groups are defined for a space which is.
  37. [37]
    The Hurewicz Isomorphism Theorem on Homotopy - Project Euclid
    We call this inverse system the Cech system o (X,A, Xo). The Cech system o (X, A) is defined similarly by using all locally finite normal open covers of X.
  38. [38]
    [PDF] Complex Cobordism and Stable Homotopy Groups of Spheres
    May 1, 2025 · Tables in low dimensions for p = 3. 3. The Adams–Novikov E2-term, Formal Group Laws, and the Greek. Letter Construction. 12. Formal group laws ...
  39. [39]
    [PDF] Notes on Cobordism Haynes Miller - MIT Mathematics
    We generalize the bordism relation to any fibration π : B → BO. ... [19] D.C. Ravenel, Complex Cobordism and Stable Homotopy Groups of Spheres, Academic.<|separator|>
  40. [40]
    Why do the homology groups capture holes in a space better than ...
    May 14, 2010 · Homology groups are better at capturing holes because higher homotopy groups are complicated, and homology is better at detecting cycles, which ...Spaces with same homotopy and homology groups that are not ...What is the difference between homology and cohomology?More results from mathoverflow.net