Fact-checked by Grok 2 weeks ago

Vector bundle

In mathematics, particularly in and , a vector bundle is a topological construction consisting of a total space E, a base space B, and a continuous surjection \pi: E \to B such that each \pi^{-1}(b) over a point b \in B is a (typically over \mathbb{R} or \mathbb{C}), and the bundle is locally trivial: the base B admits an open cover \{U_\alpha\} with homeomorphisms \phi_\alpha: \pi^{-1}(U_\alpha) \to U_\alpha \times \mathbb{R}^k (or \mathbb{C}^k) that restrict to linear isomorphisms on each , for some fixed k. This local product structure ensures that vector and are defined continuously across the total space, making vector bundles a natural way to "parametrize families of vector spaces" over a topological base. Vector bundles generalize tangent bundles on manifolds, where the fiber over each point is the tangent space, providing a framework to study how linear structures vary smoothly or continuously over a space. Key properties include the existence of a zero section (mapping each base point to the zero vector in its fiber), the ability to form direct sums and tensor products of bundles, and the notion of stable isomorphism, where two bundles E_1 and E_2 are stably isomorphic if E_1 \oplus \epsilon^n \cong E_2 \oplus \epsilon^n for some trivial bundle \epsilon^n and n. Isomorphisms between vector bundles preserve the projection and induce linear isomorphisms on fibers, while sections—continuous maps s: B \to E with \pi \circ s = \mathrm{id}_B—allow global selections of vectors from the fibers. The rank of a vector bundle, the dimension of its fibers, is locally constant, reflecting the bundle's uniformity. Classic examples illustrate the nontriviality of vector bundles: the trivial bundle B \times \mathbb{R}^k over any base B, which admits k nowhere-zero sections; the bundle, a non-orientable over S^1 obtained as a quotient of [0,1] \times \mathbb{R}; and the TS^n of the n-, which is trivial for n=1,3,7 but nontrivial otherwise, such as TS^2. Complex examples include the canonical over \mathbb{CP}^1 \cong S^2, whose dual is the bundle, highlighting connections. These constructions often arise via clutching functions on sphere products or pullbacks along maps to classifying spaces like the Grassmannians G_k(\mathbb{R}^\infty). Vector bundles play a central role in and by encoding obstructions to triviality through characteristic classes, such as Stiefel-Whitney classes for real bundles (detecting and properties) and Chern classes for complex bundles (related to curvature in ). They underpin , where the of isomorphism classes under direct sum yields invariants like the topological K-group K(X), linking to Bott periodicity and . Applications extend to principal bundles (via associated constructions), in physics, and the study of manifolds' embeddability, making vector bundles indispensable for understanding global linear phenomena over nonlinear spaces.

Basic Definitions and Properties

Formal Definition

A vector bundle is formally defined as a triple (E, \pi, B), where B is a serving as the base (typically assumed to be Hausdorff, second-countable, and paracompact to ensure desirable properties like the existence of partitions of unity), E is the total space (also a topological space), and \pi: E \to B is a continuous surjective known as the bundle projection. For each point b \in B, the \pi^{-1}(b) is a over \mathbb{R} or \mathbb{C} that is isomorphic to a fixed V of finite n (called the of the bundle), and these isomorphisms preserve the vector space structure. The defining property of local triviality requires that there exists an open cover \{U_i\}_{i \in I} of B such that for each i, the restriction of the bundle over U_i is trivialized by a homeomorphism \phi_i: \pi^{-1}(U_i) \to U_i \times V satisfying \pi = \mathrm{pr}_1 \circ \phi_i (where \mathrm{pr}_1 is the projection onto the first factor), and such that the restriction of \phi_i to each fiber \pi^{-1}(b) for b \in U_i is a linear isomorphism \pi^{-1}(b) \to \{b\} \times V. These local trivializations ensure consistency across overlaps via transition functions taking values in the general linear group \mathrm{GL}(n, \mathbb{R}) or \mathrm{GL}(n, \mathbb{C}), though the details of such gluing are addressed separately. When B is a smooth manifold, the total space E and the trivializations are required to be smooth, making \pi a submersion and \phi_i diffeomorphisms. Vector space operations on the fibers—addition (v, w) \mapsto v + w (where v, w \in \pi^{-1}(b) for the same b) and \lambda \cdot v for \lambda \in \mathbb{R} or \mathbb{C}—are defined fiberwise and extend to continuous (or , in the manifold case) maps on the total space E, ensuring the bundle structure respects the linear algebra on each fiber. Classic examples illustrate this structure: the trivial bundle B \times V \to B with onto the first factor, where the total space is simply the product and global trivializations exist; the TM \to M of a n-manifold M, where fibers are the spaces T_b M \cong \mathbb{R}^n and local trivializations arise from charts on M; and the Möbius line bundle () over S^1, whose total space is homeomorphic to an open , demonstrating a non-trivial despite local triviality over suitable covers.

Transition Functions and Local Triviality

A vector bundle E \to B of rank n over a B is locally trivial if there exists an open cover \{U_i\}_{i \in I} of B such that for each i, there is a \phi_i: \pi^{-1}(U_i) \to U_i \times \mathbb{R}^n satisfying \pi \circ \phi_i^{-1}(p, v) = p for all (p, v) \in U_i \times \mathbb{R}^n, and such that the restriction of \phi_i to each \pi^{-1}(p) is a linear onto \{p\} \times \mathbb{R}^n. These local trivializations allow the bundle to be pieced together from trivial bundles over the open sets U_i, with the linear structure preserved on fibers. On overlaps U_i \cap U_j, the transition functions g_{ij}: U_i \cap U_j \to \mathrm{GL}(n, \mathbb{R}) encode the compatibility between trivializations, defined such that \phi_j = (\mathrm{id}_{U_i \cap U_j}, g_{ij}) \circ \phi_i. These are smooth (or continuous, depending on the bundle's category) maps taking values in the general linear group, ensuring that the gluing respects the vector space structure. The collection \{g_{ij}\} must satisfy the cocycle condition: g_{ik} = g_{ij} g_{jk} on triple overlaps U_i \cap U_j \cap U_k, along with g_{ii} = \mathrm{id} and g_{ij} = g_{ji}^{-1}. This condition guarantees a well-defined global bundle without inconsistencies in the fiber identifications. Given such a cocycle \{g_{ij}\}, the total space E can be constructed explicitly as the quotient of the \coprod_i U_i \times \mathbb{R}^n by the (p, v)_i \sim (p, g_{ij}(p) v)_j for p \in U_i \cap U_j. Conversely, any vector bundle admits such a via functions on a suitable open cover. For paracompact bases B, the classes of -n real vector bundles are in bijection with the first group H^1(B, \mathrm{GL}(n, \mathbb{R})), where the cohomology class of the cocycle \{g_{ij}\} determines the bundle up to . Two cocycles represent isomorphic bundles if they differ by a coboundary, corresponding to a change of trivializations. Local trivializations are unique up to in the sense that homotopies between cocycles yield isomorphic bundles, reflecting the topological invariance captured by the group.

Subbundles and Quotient Bundles

A subbundle of a vector bundle \pi: E \to X of n is a vector subbundle F \subset E of k \leq n, where F is itself a vector bundle over X via the restricted projection \pi|_F: F \to X, and the i: F \hookrightarrow E is a bundle that is injective on each F_x \hookrightarrow E_x. The k of the subbundle is the constant of its fibers, which must be locally constant over X. Locally, over an open cover \{U_i\} of X, F|_{U_i} is a direct summand of E|_{U_i}, meaning E|_{U_i} \cong F|_{U_i} \oplus Q_i for some bundle Q_i over U_i, ensuring the subbundle inherits local triviality from E. For a subbundle F \subset E of constant rank, the quotient bundle E/F is the vector bundle of rank n - k with fibers E_x / F_x, and there is a short exact sequence of vector bundles $0 \to F \to E \to E/F \to 0, where the maps are the and the , respectively. This sequence splits locally over the cover \{U_i\}, reflecting the direct sum decomposition; over paracompact bases, a global splitting exists. Trivial examples include the zero subbundle \{0\} \subset E, which has rank 0 and quotient isomorphic to E, and the full bundle E \subset E, with rank n and quotient the zero bundle. A nontrivial example arises in differential geometry: for a submanifold M \subset N of manifolds, the tangent bundle TM \subset TN|_M is a subbundle of rank \dim M, and the quotient (TN|_M)/TM is the normal bundle to M in N. More generally, integrable distributions on a manifold correspond to subbundles of the tangent bundle. The space of all rank-k subbundles of a fixed vector bundle E of rank n over X is parameterized by the Grassmannian bundle \mathrm{Gr}_k(E) \to X, whose fibers are Grassmannians \mathrm{Gr}_k(\mathbb{R}^n).

Morphisms and Equivalences

Vector Bundle Homomorphisms

A vector bundle homomorphism, also known as a of vector bundles, between two vector bundles \xi = (E, \pi, B) and \xi' = (E', \pi', B') is a pair of continuous maps f: E \to E' and g: B \to B' such that \pi' \circ f = g \circ \pi, and for every b \in B, the restriction f_b: \pi^{-1}(b) \to \pi'^{-1}(g(b)) is a linear between the corresponding fibers, which are s. This ensures that the map preserves the vector space structure fiberwise while commuting with the projections to the base spaces. In the smooth category, where the bundles are smooth over smooth manifolds, the maps f and g are required to be . The base map g: B \to B' is uniquely determined by the total space map f, as g = \pi' \circ f \circ \pi^{-1} on the base, reflecting the induced action on the base space. Fiberwise linearity means that each f_b is a , preserving addition and within the fibers, which aligns with the local trivializations of the bundles where functions are linear isomorphisms. For a homomorphism f: E \to E' over the identity on the base (i.e., g = \mathrm{id}_B), the kernel \ker f at each point b \in B is defined as \{ v \in \pi^{-1}(b) \mid f(v) = 0 \}, forming a of the fiber, and similarly the \mathrm{im} f_b = f(\pi^{-1}(b)) is a of \pi'^{-1}(b). If the of f—the of the on each fiber—is constant across the base, then by the constant rank theorem, \ker f and \mathrm{im} f assemble into vector subbundles of \xi and \xi', respectively. In this case, the homomorphism splits locally as a of vector bundles $0 \to \ker f \to \xi \to \mathrm{im} f \to 0. For surjective homomorphisms of constant equal to the of the target bundle, the map is locally an onto its , and similarly for injective maps. Examples of vector bundle homomorphisms include the of a subbundle \eta \hookrightarrow \xi over the identity base , where each inclusion is linear by definition of a subbundle; the zero , which sends every to the zero and has constant zero; and the identity \mathrm{id}_\xi, which is linear on each with constant full .

Isomorphisms and Stable Equivalence

An isomorphism between two vector bundles E \to B and F \to B over the same base space B is a bundle homomorphism \phi: E \to F that is bijective and admits an inverse bundle homomorphism \phi^{-1}: F \to E, such that both \phi and \phi^{-1} are linear isomorphisms on each fiber. Equivalently, \phi is a of total spaces that restricts to linear isomorphisms on fibers over corresponding base points. Two vector bundles over the same base are isomorphic if there exist local trivializations such that their transition functions g_{ij} and g'_{ij} are related by a coboundary, i.e., g'_{ij} = h_i g_{ij} h_j^{-1} for continuous maps h_i: U_i \to \mathrm{GL}(n, \mathbb{R}). Equivalently, this holds if their classifying maps to the \mathrm{Gr}(n, \mathbb{R}^\infty) are homotopic, preserving the vector space structure. A vector bundle E \to B of rank n is trivial, meaning isomorphic to the product bundle B \times \mathbb{R}^n \to B, if and only if it admits n global sections that are linearly independent on every fiber. The obstructions to triviality are captured by characteristic classes, such as the Stiefel-Whitney classes w_i(E) \in H^i(B; \mathbb{Z}/2\mathbb{Z}), where the bundle is orientable if w_1(E) = 0 (for line bundles over simply connected bases, this also implies triviality), and the bundle is trivial if all w_i(E) = 0 for i \geq 1. For example, the tangent bundle of the sphere S^n is trivial for n=1,3,7 but nontrivial otherwise. For even n, the nonzero Euler class provides an obstruction to triviality, while for odd n>7, other characteristic classes (such as certain Stiefel-Whitney classes) obstruct triviality. Stable equivalence provides a coarser notion of similarity between vector bundles, where two bundles E and F over B are stably equivalent if there exist integers k, m \geq 0 such that E \oplus \epsilon^k \cong F \oplus \epsilon^m, with \epsilon^1 = B \times \mathbb{R} denoting the trivial line bundle and \oplus the direct sum. The stable equivalence classes form the reduced K-group \tilde{K}(B), an abelian group under direct sum, where the operation is well-defined since adding trivial bundles does not change the isomorphism class in the stable range. Stiefel-Whitney classes are invariant under stable equivalence, providing topological invariants for these classes. The clutching construction illustrates stable equivalence particularly for bundles over spheres: a rank-n vector bundle over S^r is determined up to isomorphism by a clutching function f: S^{r-1} \to \mathrm{GL}(n, \mathbb{R}), obtained by gluing trivial bundles over the upper and lower hemispheres via f, and two such bundles are stably equivalent if their clutching functions become homotopic after stabilizing with trivial bundles of sufficiently high rank. For instance, over S^2, clutching functions in [\pi_1(\mathrm{U}(n))] classify complex bundles, and stable equivalence corresponds to maps into the stable unitary group, linking to Bott periodicity in K-theory.

Sections and Sheaf Perspectives

Global and Local Sections

A section of a vector bundle \pi: E \to B is a continuous s: B \to E such that \pi \circ s = \mathrm{id}_B, meaning that for every point b \in B, s(b) lies in the \pi^{-1}(b). Such a map assigns to each point an element of the corresponding in a continuous manner. A global is one defined over the entire space B, while a local is defined over an open subset U \subseteq B, with its image contained in \pi^{-1}(U). The set of sections over a fixed inherits a vector space structure from the fibers: for two sections s_1, s_2 over U and a scalar , the sum (s_1 + s_2)(b) = s_1(b) + s_2(b) and scalar multiple (\lambda s_1)(b) = \lambda s_1(b) are well-defined since all fibers are isomorphic to the same . The space of all global sections, denoted \Gamma(E), forms a over the C^0(B) of continuous real-valued functions on B, with the module action given by (\phi \cdot s)(b) = \phi(b) s(b) for \phi \in C^0(B) and s \in \Gamma(E). Every vector bundle admits a zero section, the global section z: B \to E defined by z(b) = 0 in each \pi^{-1}(b). In a trivial bundle E \cong B \times V, constant sections exist, corresponding to maps s_v(b) = v for a fixed v \in V. Non-trivial bundles may lack non-zero global sections, illustrating topological obstructions. For example, the tautological line bundle over \mathbb{CP}^1 \cong S^2, which is the associated vector bundle to the S^3 \to S^2, admits no global non-vanishing sections. This follows from the fact that a is trivial if and only if it has a global non-zero section. Thus, it admits only the zero global section, and the space of global sections is the zero (dimension 0).

Locally Free Sheaves

In algebraic geometry, there is a natural equivalence between vector bundles over a scheme B and certain sheaves of \mathcal{O}_B-modules on B. Specifically, given a vector bundle E \to B of rank n, one associates the sheaf \tilde{E} of \mathcal{O}_B-modules defined by \tilde{E}(U) = \Gamma(\pi^{-1}(U), E) for each open subset U \subseteq B, where \pi: E \to B is the projection morphism and \Gamma denotes the module of sections of E over the preimage \pi^{-1}(U). This construction yields a coherent sheaf \tilde{E} that is locally free of constant rank n, and the assignment E \mapsto \tilde{E} establishes an equivalence of categories between vector bundles over B and finite locally free sheaves of \mathcal{O}_B-modules. A sheaf \mathcal{F} of \mathcal{O}_X-modules on a scheme X is called locally free of rank n if, for every point x \in X, there exists an open neighborhood U \ni x such that the restriction \mathcal{F}|_U is isomorphic to the direct sum of n copies of the structure sheaf \mathcal{O}_U, i.e., \mathcal{F}|_U \cong \mathcal{O}_U^{\oplus n}. Equivalently, each stalk \mathcal{F}_x is a free \mathcal{O}_{X,x}-module of rank n, generated by n elements that form a local frame, mirroring the local trivializations of the corresponding vector bundle. The rank function \rho(\mathcal{F}): X \to \mathbb{Z}_{\geq 0}, defined by \rho(\mathcal{F})(x) = \rank_{\mathcal{O}_{X,x}}(\mathcal{F}_x), is locally constant on X, and for sheaves arising from vector bundles, it takes a constant value equal to the bundle's rank. Representative examples illustrate this correspondence. The trivial line bundle over B, which is isomorphic to B \times k \to B for a field k, corresponds to the structure sheaf \mathcal{O}_B, a locally free sheaf of rank 1 whose stalks are free of rank 1 everywhere. On a smooth variety X over a field, the tangent sheaf T_X = \Hom_{\mathcal{O}_X}(\Omega_{X/k}, \mathcal{O}_X), dual to the sheaf of Kähler differentials, is locally free of rank equal to \dim X, reflecting the local triviality of the tangent bundle. This sheaf perspective preserves functoriality with respect to base change. For a of schemes g: B' \to B, the sheaf g^* \tilde{E} is defined by (g^* \tilde{E})(V) = \tilde{E}(g(V)) for open V \subseteq B', and it corresponds to the vector bundle E' = E \times_B B' \to B', ensuring that the equivalence respects . Global sections of the sheaf recover those of the bundle, with \Gamma(B, \tilde{E}) = H^0(B, \tilde{E}) consisting of sections over the total space.

Algebraic Operations

Direct Sums and Tensor Products

The direct sum of two vector bundles E \to B and F \to B over the same base space B is the vector bundle E \oplus F \to B defined by the total space \{(e, f) \in E \times F \mid \pi_E(e) = \pi_F(f)\}, where \pi_E and \pi_F are the respective projections to B, and the projection \pi: E \oplus F \to B given by \pi(e, f) = \pi_E(e). The fiber over each b \in B is the direct sum of the fibers E_b \oplus F_b. This construction extends associatively and commutatively to finite direct sums of bundles, with the zero bundle serving as the identity element. To ensure compatibility with the bundle structure, the is constructed via local trivializations: if E|_U \cong U \times V and F|_U \cong U \times W over an open cover \{U_i\} of B, then (E \oplus F)|_U \cong U \times (V \oplus W), with transition functions forming block-diagonal matrices using those of E and F. The of E \oplus F is the sum of the ranks, \mathrm{rank}(E \oplus F) = \mathrm{rank}(E) + \mathrm{rank}(F). A representative example is the Whitney sum of the TM and NM to the n-sphere S^n \subset \mathbb{R}^{n+1}, which yields the trivial bundle S^n \times \mathbb{R}^{n+1}. The tensor product E \otimes F \to B of vector bundles E \to B and F \to B has total space \bigsqcup_{b \in B} (E_b \otimes F_b), equipped with the finest making the s and inclusions continuous, and \pi: E \otimes F \to B defined fiberwise. The over b \in B is E_b \otimes F_b. Locally, if E|_U \cong U \times V and F|_U \cong U \times W, then (E \otimes F)|_U \cong U \times (V \otimes W), with transition functions given by the tensor products g^E_{ij} \otimes g^F_{ij} of those for E and F. The is the product \mathrm{rank}(E \otimes F) = \mathrm{rank}(E) \cdot \mathrm{rank}(F). This operation is associative, commutative, and distributive over direct sums. Examples include the exterior powers \Lambda^k E, obtained as quotients of iterated tensor products by the alternating relations, which form bundles of \binom{\mathrm{rank}(E)}{k}. The satisfies the universal property that vector bundle homomorphisms from E \oplus F to another bundle G correspond bijectively to pairs of homomorphisms from E to G and F to G. The satisfies the universal property for bilinear maps: for any vector bundle G \to B, the bundle homomorphisms E \otimes F \to G correspond to bilinear maps of bundles E \times F \to G that are linear in each factor over B. These properties mirror those in the category of vector spaces and ensure the operations are well-behaved in the category of vector bundles.

Dual Bundles

The dual bundle of a vector bundle E \to B over a base space B (with real or complex fibers) is the vector bundle E^* \to B whose fiber over each point x \in B is the dual vector space (E_x)^* = \Hom(V_x, \mathbb{R}) (or \Hom(V_x, \mathbb{C})), consisting of all linear functionals on the fiber E_x. The total space E^* is equipped with the finest topology making the projection \pi_{E^*}: E^* \to B continuous and ensuring local trivializations are homeomorphisms (or diffeomorphisms in the smooth case). If \{U_i\} is an open cover of B with local trivializations \psi_i: \pi_E^{-1}(U_i) \to U_i \times \mathbb{R}^k for E of rank k, then the transition functions for E^* are given by g_{ij}^{E^*}(x) = (g_{ij}^E(x))^{-T}, the negative transpose of those for E, ensuring compatibility on overlaps U_{ij} = U_i \cap U_j. For vector bundles E \to B and F \to B, the Hom bundle \Hom(E, F) \to B has fibers \Hom(E_x, F_x) over each x \in B, comprising linear maps between fibers, and its sections over open sets correspond to vector bundle homomorphisms from E to F that are fiberwise linear. This bundle is isomorphic to E^* \otimes F, where the tensor product is the algebraic operation on bundles (built fiberwise and glued via transition functions g_{ij}^{E^* \otimes F} = g_{ij}^{E^*} \otimes g_{ij}^F). A \ev: E \times_B E^* \to \underline{\mathbb{R}}_B exists, where \underline{\mathbb{R}}_B denotes the trivial over B with \mathbb{R} (or \mathbb{C}), defined fiberwise by \ev_x(v_x, \phi_x) = \phi_x(v_x) for v_x \in E_x and \phi_x \in (E_x)^*. This bilinear induces a nondegenerate duality between E and E^*. Prominent examples include the T^*M of a smooth manifold M, which is the dual bundle (TM)^* to the TM \to M, with sections being differential 1-forms. Another is the \det E = \bigwedge^k E for a rank-k bundle E, the top exterior power, which is a whose dual is \det E^* = \bigwedge^k E^* \cong (\bigwedge^k E)^*. The duality yields a canonical pairing on global sections \langle \cdot, \cdot \rangle: \Gamma(E) \times \Gamma(E^*) \to C^0(B), mapping continuous sections s \in \Gamma(E) and \sigma \in \Gamma(E^*) to the x \mapsto \sigma_x(s_x) on the base B. This pairing is bilinear and extends the fiberwise , facilitating integrations and other operations in and .

Geometric and Topological Structures

Pullbacks and Pushforwards

In the context of vector bundles, the pullback operation allows one to induce a new vector bundle over a different base space via a continuous between bases. Given a vector bundle \pi: E \to B over a B and a continuous f: B' \to B, the f^* E \to B' is defined with total space f^* E = \{ (b', e) \in B' \times E \mid f(b') = \pi(e) \}, equipped with the \pi_{f^* E}: f^* E \to B' given by (b', e) \mapsto b'. The fiber over a point b' \in B' is (f^* E)_{b'} = E_{f(b')}, which is naturally isomorphic to the E_{f(b')} of the original bundle as vector spaces. To verify that f^* E forms a vector bundle, local trivializations are constructed from those of E. If \phi_i: \pi^{-1}(U_i) \to U_i \times \mathbb{R}^n are local trivializations over an open cover \{U_i\} of B with transition functions g_{ij}: U_i \cap U_j \to \mathrm{GL}_n(\mathbb{R}), then for open sets V_k = f^{-1}(U_k) covering B', the pullback trivializations are \phi'_k: (\pi_{f^* E})^{-1}(V_k) \to V_k \times \mathbb{R}^n defined by \phi'_k(b', e) = (b', \phi_k(e)), and the induced transition functions are (f^* g_{ij})(b') = g_{ij}(f(b')) for b' \in V_i \cap V_j. These transition functions ensure the required for a vector bundle structure, and the construction is unique up to . The pullback functor f^* is and preserves algebraic operations on vector bundles. Specifically, for vector bundles E_1, E_2 over B, it satisfies f^*(E_1 \oplus E_2) \cong f^* E_1 \oplus f^* E_2 and f^*(E_1 \otimes E_2) \cong f^* E_1 \otimes f^* E_2, with the isomorphisms induced on fibers. Moreover, for composable maps g: B'' \to B' and f: B' \to B, the functoriality holds: (f \circ g)^* E \cong g^* (f^* E). The pushforward, or direct image, f_* E of a vector bundle \pi: E \to B' along a continuous map f: B' \to B is primarily defined at the level of sheaves of sections rather than as a bundle itself. For an open set V \subset B, the sections over V are \Gamma(f_* E, V) = \Gamma(E, f^{-1}(V)), where sections of E over f^{-1}(V) are those of the restriction E|_{f^{-1}(V)} \to f^{-1}(V). This sheaf f_* E is a sheaf of \mathcal{O}_B-modules if E corresponds to a locally free sheaf, but it is locally free (hence represents a vector bundle) only under additional conditions on f, such as when f is proper with finite fibers. In such cases, particularly for finite covering maps, the fiber over b \in B can be taken as the direct sum \bigoplus_{b' \in f^{-1}(b)} E_{b'}, yielding a vector bundle structure. Examples of pullbacks include the induced bundle on a : if i: M' \hookrightarrow M is the inclusion of a submanifold, then i^* (TM) = TM', the restricted to M'. Another instance is the restriction to fibers: for a bundle \pi: E \to B, the pullback \pi^* E over E has fibers over e \in E isomorphic to the fiber E_{\pi(e)}. For pushforwards, consider a finite covering p: \tilde{B} \to B; then p_* \mathcal{O}_{\tilde{B}} is locally free of rank equal to the degree of the cover, corresponding to a vector bundle whose sections over an open set V \subset B are the continuous functions on p^{-1}(V) \subset \tilde{B}.

Principal Bundles and Associated Vector Bundles

A principal G-bundle over a base space B is a fiber bundle \pi: P \to B equipped with a continuous free right action of a topological group G on P satisfying \pi(p g) = \pi(p) for all p \in P and g \in G, such that the orbit map P \to P/G identifies B with the quotient space P/G. This action ensures that each fiber \pi^{-1}(b) is equivariantly homeomorphic to G, with G acting on itself by right multiplication. Locally, principal G-bundles are trivial: there exists an open cover \{U_\alpha\} of B with equivariant homeomorphisms \phi_\alpha: \pi^{-1}(U_\alpha) \to U_\alpha \times G satisfying \phi_\alpha(p g) = (\pi(p), g') where g' \in G aligns with the action, and these trivializations are compatible on overlaps via transition functions g_{\alpha\beta}: U_\alpha \cap U_\beta \to G. Given a principal G-bundle \pi: P \to B and a continuous representation \rho: G \to \mathrm{GL}(V) of G on a vector space V, the associated vector bundle is constructed as the quotient space E = (P \times V)/G \to B, where G acts diagonally via (p, v) \cdot g = (p g, \rho(g^{-1}) v) for p \in P, v \in V, and g \in G. The projection map sends the equivalence class [p, v] to \pi(p) \in B, yielding fibers homeomorphic to V with a natural vector space structure induced by the representation. The transition functions of E over overlaps U_\alpha \cap U_\beta are then given by \rho(g_{\alpha\beta}^P), where g_{\alpha\beta}^P are the transition functions of the principal bundle P, ensuring that local trivializations of E over U_\alpha are U_\alpha \times V glued compatibly via these linear maps. A canonical example arises from any rank-n vector bundle \xi: E \to B, whose frame bundle \mathrm{Fr}(\xi) is the principal \mathrm{GL}(n, \mathbb{R})-bundle over B with fiber over b \in B consisting of all ordered bases (frames) of the fiber E_b \cong \mathbb{R}^n. The right \mathrm{GL}(n, \mathbb{R})-action changes bases via matrix multiplication, and local trivializations follow from those of \xi by mapping frames to standard bases in \mathbb{R}^n. If \xi is equipped with a Riemannian metric, the orthonormal frame bundle reduces to a principal O(n)-bundle, where fibers consist of orthonormal bases, obtained by restricting the structure group from \mathrm{GL}(n, \mathbb{R}) to its subgroup O(n) via the standard orthogonal representation. Conversely, every rank-n vector bundle \xi: E \to B determines a unique principal \mathrm{GL}(n, \mathbb{R})-bundle, namely its \mathrm{Fr}(\xi), and \xi recovers as the associated vector bundle \mathrm{Fr}(\xi) \times_{\mathrm{GL}(n, \mathbb{R})} \mathbb{R}^n under the standard representation of \mathrm{GL}(n, \mathbb{R}) on \mathbb{R}^n. This equivalence shows that vector bundles and principal \mathrm{GL}(n, \mathbb{R})-bundles over the same base are in bijective correspondence, with the association preserving bundle isomorphisms.

Differentiable Vector Bundles

Smooth Vector Bundles

A vector bundle over a manifold M refines the topological notion by requiring that the transition functions g_{ij}: U_i \cap U_j \to \mathrm{GL}(n, \mathbb{R}), where \{U_i\} is an open cover of M and the g_{ij} satisfy the cocycle condition g_{ij} g_{jk} = g_{ik}, are maps compatible with the atlas of M. This compatibility ensures that the total space E inherits a smooth manifold structure of \dim M + n, the projection \pi: E \to M is a smooth submersion, and local trivializations \phi_i: \pi^{-1}(U_i) \to U_i \times \mathbb{R}^n are smooth diffeomorphisms preserving the vector space structure on each fiber. Such bundles capture families of vector spaces that vary smoothly along M, enabling differential geometric constructions on the total space. The space of smooth sections \Gamma^\infty(E) comprises all smooth maps s: M \to E satisfying \pi \circ s = \mathrm{id}_M, which locally correspond to smooth functions M \to \mathbb{R}^n via the trivializations. This space forms a over C^\infty(M) and carries a natural Fréchet induced by seminorms measuring suprema of derivatives of local representatives, rendering \Gamma^\infty(E) a Fréchet manifold of infinite dimension. Smooth sections thus provide a framework for defining differential operators and integrals on the bundle. Representative examples illustrate these features. The TM of a smooth manifold M is a smooth vector bundle of \dim M, with transition functions given by the Jacobians of coordinate change maps, which are smooth by definition of the manifold structure. Smooth line bundles over surfaces, such as the hyperplane bundle over \mathbb{CP}^1 (or S^2), arise from clutching constructions with smooth transition functions like g(z) = z/|z| on the , yielding non-trivial while preserving smoothness. Partitions of unity subordinate to the cover \{U_i\} on paracompact manifolds like M enable global integration of compactly supported sections over the fibers: locally, a section s integrates fiberwise via the on \mathbb{R}^n, and these are glued smoothly using the to define \int_E |s|^2 \, d\mathrm{vol} or similar fiber densities. For smooth classification, the embeds M into , allowing reduction of bundle structures to those over \mathbb{R}^k via smooth extensions, which classifies classes up to stable equivalence through clutching functions on spheres.

Connections on Vector Bundles

In the context of smooth vector bundles over a smooth manifold M, a connection provides a means to differentiate sections along vector fields, enabling the definition of parallel transport and covariant derivatives that respect the bundle's geometry. Formally, a connection \nabla on a smooth vector bundle E \to M is a bilinear map \nabla: \Gamma(TM) \times \Gamma(E) \to \Gamma(E), written (X, s) \mapsto \nabla_X s, where X is a smooth vector field on M and s is a smooth section of E. It satisfies the Leibniz rule: for any smooth function f \in C^\infty(M), \nabla_X (f s) = (X f) s + f \nabla_X s. This rule ensures that the connection acts linearly over the structure sheaf while accounting for the directional derivative of scalar coefficients. Locally, over a coordinate U \subset M with a local \{e_1, \dots, e_r\} for E|_U, the connection takes the form \nabla = d + A, where d is the on forms with values in E|_U, and A is a 1-form taking values in \Omega^1(U, \End(E)), the space of endomorphism-valued 1-forms. For a section s = \sum_i f^i e_i with f^i \in C^\infty(U), the local is \nabla s = \sum_i e_i \otimes df^i + \sum_{i,j} f^i A^j_i \otimes e_j, where A = (A^j_i) is the matrix of 1-forms. Under a change of via a GL(r, \mathbb{R})-valued transition function g: U \to GL(r, \mathbb{R}), the transforms as \tilde{A} = g^{-1} A g + g^{-1} dg, preserving the global structure. Parallel transport induced by \nabla allows the of fibers along curves \gamma: [0,1] \to M. For a v \in E_{\gamma(0)}, the parallel transport to E_{\gamma(1)} is the solution to the (ODE) \frac{d}{dt} s(t) + A_{\gamma(t)}(s(t)) \cdot \dot{\gamma}(t) = 0, where s(t) \in E_{\gamma(t)} with s(0) = v. This defines a linear P^\nabla_\gamma: E_{\gamma(0)} \to E_{\gamma(1)}, which is independent of parametrization and satisfies the group law for concatenated curves, forming the holonomy representation. The curvature of a connection \nabla measures its deviation from flatness and is given by the 2-form \Omega^\nabla = dA + A \wedge A \in \Omega^2(M, \End(E)), where the wedge product incorporates the bracket in \End(E). Locally, for vector fields X, Y, \Omega^\nabla(X, Y) s = \nabla_X \nabla_Y s - \nabla_Y \nabla_X s - \nabla_{[X,Y]} s, acting as an on sections. The second Bianchi identity holds: d\Omega + [A, \Omega] = 0, where [ \cdot, \cdot ] denotes the graded , ensuring the curvature satisfies a consistency condition under . Examples illustrate these concepts concretely. On the TM \to M of a (M, g), the is the unique torsion-free satisfying \nabla_X Y \cdot Z = X(Y \cdot Z) for all vector fields X, Y, Z, preserving the metric and enabling geodesic flow. For a trivial bundle E = M \times V \to M with global frame \{e_i\}, the trivial is \nabla_X s = X(s), where s is viewed as a V-valued function, yielding zero \Omega = 0.

Applications in Topology and Geometry

Characteristic Classes

Characteristic classes are topological invariants associated to vector bundles that capture obstructions to the existence of sections or provide classes measuring the bundle's twisting relative to the trivial bundle. For a vector bundle E \to B over a paracompact base space B, these classes live in the ring of B and are natural under bundle maps, making them useful for classifying bundles up to . They were developed in and 1940s through axiomatic approaches that emphasize functoriality and additivity under direct sums. For complex vector bundles, the Chern classes c_k(E) \in H^{2k}(B; \mathbb{Z}) form a sequence of cohomology classes, with the total Chern class defined as c(E) = 1 + c_1(E) + c_2(E) + \cdots + c_r(E), where r is the rank of E, and higher classes vanish. These classes satisfy the axioms of naturality, meaning that for a continuous map f: X \to B, f^* c_k(E) = c_k(f^* E), and the Whitney sum formula c(E \oplus F) = c(E) \cup c(F) for bundles E, F \to B. The Chern classes originate from Shiing-Shen Chern's work on Hermitian manifolds and were axiomatized to apply topologically to all complex bundles. For real vector bundles, the Stiefel-Whitney classes w_k(E) \in H^k(B; \mathbb{Z}/2) provide analogous invariants, with the total class w(E) = 1 + w_1(E) + \cdots + w_r(E). They satisfy the same naturality axiom and Whitney sum formula w(E \oplus F) = w(E) \cup w(F), and were introduced independently by Eduard Stiefel and as obstructions to extending frames over skeleta. On manifolds, the Stiefel-Whitney classes of the obey Wu's formula, relating them to the action of Steenrod squares on the Wu class \nu, via w(TM) = \mathrm{Sq}(\nu), which determines the and other properties. The Euler class e(E) \in H^n(B; \mathbb{Z}) is defined for an oriented real vector bundle E of n, serving as the top and vanishing if E admits a nowhere-zero section. For oriented rank-2 bundles, it coincides with the first of the . The Pontryagin classes p_k(E) \in H^{4k}(B; \mathbb{Z}) for real bundles of rank at least $4k are derived from Chern classes via p_k(E) = (-1)^k c_{2k}(E \otimes \mathbb{C}), linking real and complex invariants and satisfying analogous axioms. Examples illustrate these classes' utility: for the tangent bundle TS^2 of the 2-sphere, viewed as a complex line bundle, c_1(TS^2) = 2g, where g \in H^2(S^2; \mathbb{Z}) is the positive generator, reflecting the bundle's non-triviality. The hairy ball theorem follows from the Euler class, as e(TS^{2m}) \neq 0 for even-dimensional spheres, implying no nowhere-vanishing vector field.

Vector Bundles in K-Theory

Topological K-theory associates to a compact Hausdorff space B the abelian group K^0(B), defined as the Grothendieck group of the semigroup of isomorphism classes of complex vector bundles over B under direct sum, where elements are formal differences [E] - [F] of bundle classes and relations arise from short exact sequences $0 \to F \to E \oplus H \to G \to 0 yielding [E] = [F] + [G]. This construction incorporates stable equivalence, where bundles E and F are stably isomorphic if E \oplus H \cong F \oplus H for some bundle H. The group K^0(B) acquires a ring structure from the tensor product of bundles, with the trivial line bundle as the unit. For connected B, the reduced group \tilde{K}^0(B) is the kernel of the rank map K^0(B) \to \mathbb{Z} = K^0(\mathrm{pt}), capturing the non-trivial stable classes orthogonal to the trivial bundle. Bott periodicity asserts that the higher K-groups, defined via suspension \Sigma B = S^1 \wedge B, satisfy K^{n+2}(B) \cong K^n(B) for all n \in \mathbb{Z}, endowing K-theory with a \mathbb{Z}/2\mathbb{Z}-periodic cohomology theory. K-theory is contravariant: for a continuous map f: B' \to B, the pullback induces a ring homomorphism f^*: K^0(B) \to K^0(B') by f^*[E] = [f^*E]. For a CW pair (X, A) with inclusion i: A \hookrightarrow X and quotient p: X \to X/A, the long exact sequence is \cdots \to \tilde{K}^0(A) \xrightarrow{i^*} \tilde{K}^0(X) \xrightarrow{p^*} \tilde{K}^0(X/A) \xrightarrow{\partial} \tilde{K}^{-1}(A) \to \cdots, where the boundary map \partial detects extensions of bundles. The Chern character provides a natural ring homomorphism \mathrm{ch}: K^0(B) \to H^*(B; \mathbb{Q}) from K-theory to rational cohomology, expressing bundle classes in terms of their Chern classes via the formula \mathrm{ch}(E) = \mathrm{rank}(E) + c_1(E) + \frac{c_1(E)^2 - 2c_2(E)}{2!} + \cdots + \frac{1}{k!} \sum \sigma_k(c_1(E), \dots, c_k(E)) + \cdots, where \sigma_k denotes the k-th power sum of the formal Chern roots, thus linking stable bundle invariants to topological cohomology. A representative example is K^0(S^2) \cong \mathbb{Z} \oplus \mathbb{Z}, where the first factor is generated by the trivial bundle (via ) and the reduced group \tilde{K}^0(S^2) \cong \mathbb{Z} by the Hopf line bundle \mathcal{O}(1) over \mathbb{CP}^1 \cong S^2. Vector bundles over S^2 are classified up to stable isomorphism by clutching constructions: a bundle is obtained by gluing trivial bundles over the hemispheres via a transition function S^1 \to \mathrm{U}(n), and the stable classes correspond to classes [S^1, U(n)], yielding \mathbb{Z} from the line bundle's .

References

  1. [1]
    [PDF] Version 2.2, November 2017 Allen Hatcher Copyright c 2003 by ...
    Page 8. To motivate the definition of a vector bundle let us consider tangent vectors to. the unit 2 sphere S2. in R.
  2. [2]
    [PDF] Notes on Vector Bundles
    Mar 16, 2010 · Definition 1.1. A real vector bundle of rank k is a tuple (M,V,π,·,+) such that. (1) M and V ...
  3. [3]
    [PDF] FIBER BUNDLES AND VECTOR BUNDLES These notes, written for ...
    A vector bundle is a fiber bundle as in Definition 3 for which the fibers π−1(p), p ∈. M are vector spaces, the manifolds F in the local trivialization are ...
  4. [4]
    [PDF] Fibre Bundles
    ... vector bundle is a pair consisting of a vector bundle and an orientation on the bundle. In other words, a vector bundle has an atlas of charts where the.
  5. [5]
    [PDF] FOUNDATIONS OF ALGEBRAIC GEOMETRY CLASS 24
    Jan 18, 2008 · Transition functions for the sheaf of sections. Suppose we have a vector bundle on. M, along with a trivialization over an open cover Ui.
  6. [6]
    [PDF] Math 396. Subbundles and quotient bundles 1. Motivation We want ...
    The preceding considerations suggest a couple of different ways to define the notion of subbundle of a vector bundle. We begin with the most naive definition, ...
  7. [7]
    Fibre Bundles | SpringerLink
    Download chapter PDF. Preliminaries on Homotopy Theory. Preliminaries on Homotopy Theory. Dale Husemoller. Pages 1-8. The General Theory of Fibre Bundles. Front ...
  8. [8]
    [PDF] 2 Vector bundles. - DPMMS
    Definition. A smooth map F : E → E0 is a vector bundle morphism covering f if. for any p ∈ B F restricts to a linear map between the fibres F : Ep → E0. f(p)
  9. [9]
    [PDF] Cohomology and Vector Bundles
    Vector bundles are a generalization of the cross product of a topo- logical space with a vector space. Characteristic classes assign to the each vector bundle a ...
  10. [10]
    [PDF] Lent Term 2023 O. Randal-Williams Part III Characteristic classes ...
    This question leads you through the proof of the Constant Rank Theorem: If πi : Ei → X, i = 1,2, are vector bundles and f : E1 → E2 is a morphism of vector ...
  11. [11]
    [PDF] Lecture 1: Introduction Overview Vector bundles arise in many parts ...
    Definition 1.12. A family of vector spaces π: E ر X parametrized by X is a vector bundle if it is locally trivial, i.e., if for each x P X there exists an open ...
  12. [12]
    [PDF] 1. Overview We look at complex line bundles from the topological ...
    A bundle is trivial if and only if it admits a global nonvanishing section. A bundle is holomorphically trivial if and only if it admits a global non-vanishing.<|control11|><|separator|>
  13. [13]
    Section 27.6 (01M1): Vector bundles—The Stacks project
    Every vector bundle over is a module scheme for the ring scheme , i.e. is a group scheme together with a scalar multiplication that makes into an -module for ...
  14. [14]
    Section 17.14 (01C5): Locally free sheaves—The Stacks project
    We say \mathcal{F} is locally free if for every point x \in X there exist a set I and an open neighbourhood x \in U \subset X such that \mathcal{F}|_ U is ...
  15. [15]
    [PDF] Math 396. Direct sums of vector bundles 1. Overview Let E1,...,En be ...
    The direct sum vector bundle E1 ⊕···⊕ En is defined as the vector bundle whose x-fiber is E1(x) ⊕···⊕ En(x) for all x ∈ X.
  16. [16]
    [PDF] Math 396. Linear algebra operations on vector bundles
    Let (X,O) be a Cp premanifold with corners, 0 ≤ p ≤ ∞. We have developed the notion of a Cp vector bundle over X as a certain kind of Cp mapping π : E → X that ...
  17. [17]
    [PDF] Differential geometry Lecture 10: Dual bundles, 1-forms, and the ...
    May 22, 2020 · The most important example of a dual vector bundle for this course is the dual to the tangent bundle of a smooth manifold: Definition. The ...
  18. [18]
    [PDF] 3.2 Vector bundles
    If E −→ M is a vector bundle, then E∗ −→ M is the dual vector bundle. If E,F are vector bundles then E ⊕ F is called the direct or “Whitney” sum, and ...Missing: mathematics | Show results with:mathematics
  19. [19]
    [PDF] 1. The definition of a vector bundle - KSU Math
    A vector bundle over a variety X is a variety Y with a map p : Y → X, plus the structure of a vector space on every p−1(x) for x ∈ X.Missing: mathematics | Show results with:mathematics
  20. [20]
    [PDF] Math 396. Bundle pullback and transition matrices 1. Motivation Let f
    A pullback bundle (f∗E → X0, e f) is a bundle that promotes bundle morphisms between different base spaces to bundle morphisms between a common base space.
  21. [21]
    [PDF] Vector Bundles and Pullbacks - Nikolai Nowaczyk
    In section 2 we define pullbacks of vector bundles and proof some algebraic properties. Then we are able to tackle the proof of the Homotopy Invariance ...
  22. [22]
    [PDF] Pullback of a vector bundle and connection
    Apr 1, 2021 · smooth map Φ : N → M, the pullback bundle EΦ over N is defined as follows: • The fiber at x ∈ N is EΦ x = EΦ(x). • Given any local frame ...
  23. [23]
    [PDF] 4 Sheaves of modules, vector bundles, and (quasi-)coherent sheaves
    We call it the direct image of F, or the pushforward of F by f. Also note that the inverse image sheaf f 1G is an f 1OY -module. Thanks to the adjoint.
  24. [24]
    Proper smooth pushforward of vector bundle is a vector bundle?
    Nov 11, 2024 · The answer is "no". For a counterexample, let Y=E be an elliptic curve and let X=E×E with f:E×E→E the first projection.Direct image of a vector bundle under birational morphismThe injection of direct image sheaf - MathOverflowMore results from mathoverflow.net
  25. [25]
    Pushforward and smooth vector pseudo-bundles - MSP
    We study a new operation named pushforward on diffeological vector pseudo- bundles, which is left adjoint to the pullback. We show how to pushforward.
  26. [26]
    [PDF] Principal G-Bundles
    Oct 28, 2019 · 2.1 Principal Bundles. Definition 2.1. A (right) principal G-bundle is a triple (P,B,π) where π : P → B is a map. There is a continuous, free ...Missing: mathematics | Show results with:mathematics
  27. [27]
    [PDF] Chapter 3 - Vector bundles and principal bundles
    Definition 16.2. A vector bundle over B is a vector space E over B that is locally trivial – that is, every point b ∈ B has a neighborhood ...
  28. [28]
    Associated Vector Bundle -- from Wolfram MathWorld
    This construction has many uses. For instance, any group representation of the orthogonal group gives rise to a bundle of tensors on a Riemannian manifold as ...
  29. [29]
    [PDF] Lecture Notes on Bundles and Connections
    Sep 26, 2008 · Once the basics about connections on vector bundles are well understood, it becomes fairly simple to discuss the most important concepts from ...
  30. [30]
    [PDF] Differential geometry Lecture 15: Connections in vector bundles
    Jun 15, 2020 · Connections in vector bundles. Definition (continuation). Thus we have ωi = n. X j=1 ωij ⊗ dxj for all 1 ≤ i ≤ `, where ωij ∈ Γ(E|U ) for ...
  31. [31]
    [PDF] VECTOR BU1\'DLES A.,\'D BO)IOGEl''EOUS SPACES
    M. F. ATIYAH A .. '\D F. HIRZEBRUCH. Dedicated to Professor llarston Morae. Introduction. In [1] we introduced for a space X the "ring of complex vector ...
  32. [32]
    [PDF] K-THEORY LECTURES BY NOTES BY M. F. A.TIYAH* D. W. ...
    Apr 1, 2011 · ON K-THEORY AND REALITY. The Grothendieck group of the category of real vector bundles over a real space X is denoted by KR(X). Restrictillg ...Missing: original | Show results with:original