Fact-checked by Grok 2 weeks ago

Holomorphic vector bundle

A holomorphic vector bundle is a vector bundle E \to X over a X, equipped with a holomorphic structure such that the total space E is a and the projection \pi: E \to X is a holomorphic map. Equivalently, it consists of an open cover \{U_i\} of X with holomorphic trivializations \phi_i: \pi^{-1}(U_i) \to U_i \times \mathbb{C}^r (for rank r), where the transition functions g_{ij}: U_i \cap U_j \to \mathrm{GL}(r, \mathbb{C}) are holomorphic. This structure ensures that the fibers E_x \cong \mathbb{C}^r vary holomorphically with the base point x \in X. Holomorphic vector bundles form a of , providing the framework for analyzing families of complex vector spaces parametrized by complex manifolds. They generalize bundles and line bundles, enabling the study of geometric invariants like Chern classes, which measure topological and analytic properties, and support the definition of holomorphic sections—maps s: X \to E that are holomorphic when composed with local trivializations. In , they underpin the theory of coherent sheaves and projective embeddings via ample bundles, as in Kodaira's theorem. Key developments include Grothendieck's classification of holomorphic bundles over the as direct sums of line bundles of specific degrees, and the notion of , which characterizes bundles admitting Hermitian-Einstein metrics that minimize the Yang-Mills functional and is vital for moduli problems. These objects also admit compatible connections, such as the Chern connection, facilitating the integration of with .

Definition and Structure

Local trivializations

A holomorphic vector bundle over a X is a complex vector bundle equipped with a holomorphic structure, meaning that it admits an atlas of local trivializations related by holomorphic transition maps. More precisely, let E \to X be a complex vector bundle of r, with total space E a manifold and projection \pi: E \to X a holomorphic submersion such that each fiber E_x = \pi^{-1}(x) is a complex vector space isomorphic to \mathbb{C}^r. The holomorphic structure is induced by requiring that E is locally holomorphically trivial. To elaborate, consider an open cover \{U_i\}_{i \in I} of X. For each i, the restriction of the bundle over U_i, denoted E|_{U_i}, is holomorphically trivial, meaning there exists a biholomorphic map \phi_i: \pi^{-1}(U_i) \to U_i \times \mathbb{C}^r such that the following commutes: \begin{CD} \pi^{-1}(U_i) @>{\phi_i}>> U_i \times \mathbb{C}^r \\ @V{\pi}VV @VV{\mathrm{pr}_1}V \\ U_i @= U_i, \end{CD} where \mathrm{pr}_1 is the onto the first . The collection \{(U_i, \phi_i)\}_{i \in I} forms a holomorphic atlas of trivializations, provided the transition maps are holomorphic. On overlaps U_i \cap U_j \neq \emptyset, the transition maps g_{ij}: U_i \cap U_j \to \mathrm{GL}(r, \mathbb{C}) are defined by \phi_j \circ \phi_i^{-1}(u, v) = (u, g_{ij}(u) \cdot v) for u \in U_i \cap U_j and v \in \mathbb{C}^r. These maps must be holomorphic functions, satisfying the cocycle condition g_{ij}(u) \cdot g_{jk}(u) = g_{ik}(u) on triple overlaps U_i \cap U_j \cap U_k, and g_{ji} = g_{ij}^{-1}. This ensures the bundle is consistently glued across the cover while preserving the holomorphic structure. The holomorphic atlas implicitly defines the sheaf of holomorphic sections of the bundle, denoted \mathcal{O}_E, where over each U_i, sections correspond to holomorphic maps s_i: U_i \to \mathbb{C}^r via \phi_i, and compatibility on overlaps requires s_j(u) = g_{ij}(u) \cdot s_i(u). This sheaf structure captures the global holomorphic sections as those that glue holomorphically from the local frames.

Transition functions and holomorphic structure

A holomorphic vector bundle over a M is constructed by endowing a smooth complex vector bundle with a holomorphic structure. This is achieved by selecting an open cover \{U_i\}_{i \in I} of M and associating to each pair i, j a holomorphic transition function g_{ij}: U_i \cap U_j \to \mathrm{GL}(r, \mathbb{C}), where r is the of the bundle. These functions must satisfy the cocycle g_{ik} = g_{ij} g_{jk} on triple overlaps U_i \cap U_j \cap U_k, ensuring consistent gluing of the local trivializations \pi^{-1}(U_i) \cong U_i \times \mathbb{C}^r. The total space E then inherits a complex manifold structure from these biholomorphic trivializations, with the projection \pi: E \to M being holomorphic. The transition functions uniquely determine the holomorphic structure up to isomorphism. Specifically, two collections of transition functions \{g_{ij}\} and \{g'_{ij}\} define isomorphic bundles if there exist holomorphic maps h_i: U_i \to \mathrm{GL}(r, \mathbb{C}) such that g'_{ij} = h_i g_{ij} h_j^{-1} on U_i \cap U_j, meaning the collections differ by a holomorphic coboundary. This equivalence relation classifies holomorphic vector bundles in terms of non-abelian Čech cohomology, where the isomorphism classes correspond to elements in H^1(M, \mathrm{GL}(r, \mathcal{O}_M)), with \mathcal{O}_M the sheaf of holomorphic functions on M; though the explicit cocycle description provides the concrete gluing data. With this structure in place, sections of the bundle acquire a notion of holomorphy. A s: U \to E over an open subset U \subset M is holomorphic if it is a holomorphic map satisfying \pi \circ s = \mathrm{id}_U, and locally, in a trivialization over U_i, it takes the form s(z) = \sum_{k=1}^r f_k(z) \, e_k, where \{e_k\}_{k=1}^r is a local holomorphic frame and each f_k: U_i \to \mathbb{C} is a . This local expression ensures compatibility across overlaps via the functions, as s on U_j transforms accordingly to maintain holomorphy. In contrast to algebraic vector bundles, which are defined on algebraic varieties using regular (e.g., ) transition functions within the framework, holomorphic vector bundles reside in the complex analytic category. Here, the transition functions are merely holomorphic, enabling constructions on general complex manifolds that may not admit algebraic structures.

Sections and Sheaves

Sheaf of holomorphic sections

Given a holomorphic vector bundle E \to X over a complex manifold X, the sheaf \mathcal{E} of holomorphic sections is constructed by assigning to each open subset U \subset X the \mathbb{C}-vector space \Gamma(U, \mathcal{E}) consisting of all holomorphic maps s: U \to E|_U such that the projection \pi \circ s = \mathrm{id}_U and, locally over trivializing open sets U_\alpha \subset U, s corresponds to a holomorphic function with values in \mathbb{C}^r via the trivialization \psi_\alpha: \pi^{-1}(U_\alpha) \to U_\alpha \times \mathbb{C}^r. The sheaf \mathcal{E} satisfies the standard sheaf axioms: for any open cover \{U_i\}_{i \in I} of an U \subset X, the restriction maps \rho_{U, U_i}: \Gamma(U, \mathcal{E}) \to \Gamma(U_i, \mathcal{E}), defined by projecting sections along the bundle fibers, are holomorphic linear maps satisfying \rho_{U, U_i} \circ \rho_{V, U} = \rho_{V, U_i} for U_i \subset U \subset V; moreover, sections over the U_i glue uniquely to a section over U if they agree on pairwise intersections U_i \cap U_j, where agreement is determined by the holomorphic transition functions g_{\alpha\beta}: U_\alpha \cap U_\beta \to \mathrm{GL}(r, \mathbb{C}) of the bundle such that the local representations match via s_\alpha = g_{\alpha\beta} s_\beta. The sheaf \mathcal{E} is naturally a sheaf of \mathcal{O}_X-modules, where \mathcal{O}_X is the structure sheaf of holomorphic functions on X, with the module structure given by pointwise multiplication: for f \in \Gamma(U, \mathcal{O}_X) and s \in \Gamma(U, \mathcal{E}), (f \cdot s)(x) = f(x) \cdot s(x) in the fiber E_x. Since E is locally trivial of finite rank r, \mathcal{E} is locally free of rank r over \mathcal{O}_X and hence a coherent \mathcal{O}_X-module sheaf. For a holomorphic f: Y \to X of complex manifolds, the direct image sheaf f_* \mathcal{E} (or pushforward sheaf) is defined by (f_* \mathcal{E})(U) = \Gamma(f^{-1}(U), \mathcal{E}) for open U \subset X, equipped with the natural \mathcal{O}_X- structure induced by composition with f, providing an algebraic framework to study sections of \mathcal{E} pulled back to Y. This construction preserves coherence and is functorial in the category of holomorphic vector bundles and s.

Global holomorphic sections

The space of global holomorphic sections of a holomorphic vector bundle E \to X over a complex manifold X, denoted H^0(X, E) or \Gamma(X, E), consists of all holomorphic maps s: X \to E satisfying \pi \circ s = \mathrm{id}_X, where \pi: E \to X is the bundle projection. Locally, over a trivializing open cover \{U_\alpha\} with transition functions g_{\alpha\beta}: U_\alpha \cap U_\beta \to \mathrm{GL}_r(\mathbb{C}), a section s is represented by holomorphic functions s_\alpha: U_\alpha \to \mathbb{C}^r such that s_\alpha = g_{\alpha\beta} s_\beta on overlaps, and holomorphicity requires that each s_\alpha lies in the kernel of the \bar{\partial}-operator, defined componentwise as \bar{\partial} s_\alpha = (\bar{\partial} s_{\alpha,1}, \dots, \bar{\partial} s_{\alpha,r}). On a compact X, the space \Gamma(X, \mathcal{E}) of global sections of the sheaf \mathcal{E} of holomorphic sections of E is finite-dimensional as a vector space, since \mathcal{E} is coherent and the groups of coherent sheaves on compact manifolds are finite-dimensional by the Cartan-Serre . This finiteness follows from the coherence of the structure sheaf \mathcal{O}_X, established by Oka's , which ensures that \mathcal{E} admits finite locally resolutions, implying bounded dimensionality of global sections. For the trivial bundle E = X \times \mathbb{C}^r, the global holomorphic sections are precisely the r-tuples of global holomorphic functions on X. In contrast, for an ample line bundle L on a projective manifold X of dimension n, the space of sections \Gamma(X, L^k) grows richly with k, satisfying \dim \Gamma(X, L^k) \sim C k^n for some constant C > 0 independent of k, by the asymptotic Hirzebruch-Riemann-Roch theorem, reflecting the positivity that allows embedding X via high powers. Geometrically, for a line bundle L, the zero locus of a non-zero global holomorphic section s \in \Gamma(X, L) defines a on X, whose associated line bundle is isomorphic to L, providing a correspondence between effective divisors and sections up to scaling.

Examples

Trivial and direct sum bundles

A trivial holomorphic vector bundle of rank r over a X is the product bundle E = X \times \mathbb{C}^r equipped with the projection map \mathrm{pr}_1: (x, v) \mapsto x. The transition functions with respect to any trivializing cover are the identity matrices in \mathrm{GL}(r, \mathbb{C}), ensuring the holomorphic structure is induced directly from that of X. The holomorphic sections of such a bundle are precisely the holomorphic maps s: X \to \mathbb{C}^r, which assign to each point an element of the fiber in a holomorphic manner. The direct sum of two holomorphic vector bundles E and F over the same base X, denoted E \oplus F, has total space consisting of pairs (e, f) with e \in E and f \in F such that \pi_E(e) = \pi_F(f), and projection onto X given by this common base point. If E and F have ranks r and s, then E \oplus F has rank r + s, with fibers E_x \oplus F_x at each x \in X. The holomorphic structure on E \oplus F is defined using transition functions that are block-diagonal matrices, combining the transition functions of E and F along the diagonal. Holomorphic sections of E \oplus F are pairs (s_E, s_F) where s_E and s_F are holomorphic sections of E and F, respectively. The Hom bundle \mathrm{Hom}(E, F) between holomorphic vector bundles E and F over X has fibers consisting of linear maps between E_x and F_x, and in particular the endomorphism bundle \mathrm{End}(E) = \mathrm{Hom}(E, E) can be identified with the tensor product E^* \otimes E, where E^* is the dual bundle. The holomorphic structure on E^* uses transition functions that are the adjoints (inverse transposes) of those for E, ensuring the tensor product inherits a natural holomorphic structure via the product of transition functions. A holomorphic vector bundle is trivial if and only if it admits a global holomorphic , that is, r global holomorphic sections that are linearly independent at every point and span the . Necessary conditions for triviality include the base manifold being parallelizable (for the case) or all Chern classes of the bundle vanishing.

Tangent and cotangent bundles

The holomorphic T^{1,0}X of a X of dimension n provides a fundamental example of a holomorphic vector bundle of n. Given a holomorphic atlas \{ (U_\alpha, z^\alpha) \} for X, where z^\alpha = (z^1_\alpha, \dots, z^n_\alpha) are local holomorphic coordinates on U_\alpha, the bundle admits a local trivialization \phi_\alpha: \pi^{-1}(U_\alpha) \to U_\alpha \times \mathbb{C}^n over each chart, with the fiber at p \in U_\alpha spanned by the basis \left\{ \frac{\partial}{\partial z^i_\alpha} \big|_p \right\}_{i=1}^n. On overlapping charts U_\alpha \cap U_\beta, with coordinate change z^j_\alpha = h^j_{\alpha\beta}(z^\beta) given by a h_{\alpha\beta}: z^\beta(U_\alpha \cap U_\beta) \to z^\alpha(U_\alpha \cap U_\beta), the transition functions for T^{1,0}X are the entries of the g_{\alpha\beta}(z^\beta) = \left( \frac{\partial z^j_\alpha}{\partial z^k_\beta} \right)_{j,k=1}^n, valued in \mathrm{GL}_n(\mathbb{C}). These functions are holomorphic because h_{\alpha\beta} is biholomorphic, so each \frac{\partial z^j_\alpha}{\partial z^k_\beta} is a of z^\beta, as ensured by the chain rule applied to the composition of holomorphic maps. The holomorphic cotangent bundle \Omega^1_X, also of rank n, is the dual vector bundle (T^{1,0}X)^*. Its local trivializations are induced by the dual basis \{ dz^i_\alpha \} over U_\alpha, and the transition functions are the inverse transpose matrices g_{\alpha\beta}^{*}(z^\beta) = \left( g_{\beta\alpha}(z^\alpha) \right)^{-T}, which remain holomorphic since the original Jacobians are invertible holomorphic matrices. Global holomorphic sections of T^{1,0}X are precisely the holomorphic vector fields on X, which in local coordinates take the form \sum_{i=1}^n f_i(z) \frac{\partial}{\partial z^i} with each f_i holomorphic on the . Similarly, holomorphic sections of \Omega^1_X are the holomorphic 1-forms, locally expressed as \sum_{j=1}^n g_j(z) \, dz^j with g_j holomorphic. The higher exterior powers \Omega^p_X = \bigwedge^p \Omega^1_X for $1 \leq p \leq n form holomorphic vector bundles of rank \binom{n}{p}, whose fibers consist of decomposable p-vectors in the alternating algebra on the cotangent fibers. The transition functions for \Omega^p_X are induced by the action of the cotangent transition matrices on the exterior power, yielding expressions involving pluri-differentials such as dz^{j_1} \wedge \cdots \wedge dz^{j_p} = \det\left( \frac{\partial z^{j_k}_\alpha}{\partial z^{\ell_m}_\beta} \right)_{k,m} \, dw^{\ell_1} \wedge \cdots \wedge dw^{\ell_p} over suitable minors, with holomorphicity preserved by the determinant of holomorphic matrices. In particular, for p = n, \Omega^n_X is the canonical line bundle, with transition functions given by the determinant \det(g_{\alpha\beta}^{*}). On \mathbb{C}^n, both T^{1,0}\mathbb{C}^n and \Omega^1_{\mathbb{C}^n} are holomorphically trivial bundles, isomorphic to \mathbb{C}^n \times \mathbb{C}^n, with global sections comprising entire vector fields \sum f_i(z) \frac{\partial}{\partial z^i} and entire 1-forms \sum g_j(z) \, dz^j where the f_i, g_j are entire functions. On a (complex dimension 1), T^{1,0}X reduces to the holomorphic tangent , while \Omega^1_X is the whose global sections are the holomorphic differentials, forming a of dimension equal to the of X.

Dolbeault Apparatus

Dolbeault operators

In the context of a holomorphic vector bundle E \to X over a X, the Dolbeault operator \bar{\partial}_E is defined as the (0,1)-part of the d acting on the space A^{0,q}(X, E) of smooth sections of \Lambda^{0,q} T^*X \otimes E. This operator is a \mathbb{C}-linear differential \bar{\partial}_E: A^{0,q}(X, E) \to A^{0,q+1}(X, E) of bidegree (0,1) that satisfies the nilpotency condition \bar{\partial}_E^2 = 0 and the graded Leibniz rule \bar{\partial}_E(f \sigma) = \bar{\partial} f \wedge \sigma + (-1)^q f \bar{\partial}_E \sigma for a smooth function f on X and a section \sigma \in A^{0,q}(X, E). When f is holomorphic, \bar{\partial} f = 0, so the rule simplifies to \bar{\partial}_E(f \sigma) = f \bar{\partial}_E \sigma. Locally, over a trivializing U \subset X where E|_U \cong U \times \mathbb{C}^r, a \sigma \in A^{0,q}(U, E) is represented by an \mathbb{C}^r-valued (0,q)-form s, and the action of \bar{\partial}_E takes the form \bar{\partial}_E \sigma = \bar{\partial} s + \bar{\theta} \wedge s, where \bar{\theta} \in A^{0,1}(U, \mathfrak{gl}(r, \mathbb{C})) is the (0,1)- determined by the holomorphic transition functions of E. On overlaps U \cap V, the forms \bar{\theta}_U and \bar{\theta}_V are related by the transformation law \bar{\theta}_U = g_{UV} \bar{\theta}_V g_{UV}^{-1} + \bar{\partial} g_{UV} \cdot g_{UV}^{-1}, ensuring the operator is globally well-defined without invoking a full on E. The space of holomorphic sections of E, denoted \mathcal{O}(X, E), is precisely the kernel of \bar{\partial}_E acting on A^{0,0}(X, E), i.e., \mathcal{O}(X, E) = \ker(\bar{\partial}_E: A^{0,0}(X, E) \to A^{0,1}(X, E)). This identifies the sheaf of holomorphic sections with the zeroth cohomology of the Dolbeault complex (A^{0,\bullet}(X, E), \bar{\partial}_E), which resolves the structure sheaf \mathcal{O}_E. The holomorphic structure on E is integrable, meaning \bar{\partial}_E^2 = 0 holds globally, if and only if the local (0,1)-connection forms satisfy the \bar{\partial} \bar{\theta} + \bar{\theta} \wedge \bar{\theta} = 0 on each trivializing set; this ensures consistency under the Newlander-Nirenberg theorem that the distribution defined by the kernel of \bar{\theta} is integrable.

Cohomology of holomorphic vector bundles

The Dolbeault cohomology groups of a holomorphic vector bundle E over a complex manifold X are defined as H^{p,q}(X, E) = \ker(\bar{\partial}_E : A^{p,q}(X, E) \to A^{p,q+1}(X, E)) / \operatorname{im}(\bar{\partial}_E : A^{p,q-1}(X, E) \to A^{p,q}(X, E)), where A^{p,q}(X, E) denotes the space of smooth global sections of the bundle \Lambda^{p,0}T^*X \otimes \overline{\Lambda^{0,q}T^*X} \otimes E and \bar{\partial}_E is the extension of the Dolbeault operator to E-valued forms. These groups can also be computed as the sheaf cohomology \check{H}^q(X, \mathcal{A}^{p,E}), where \mathcal{A}^{p,E} is the sheaf of smooth E-valued (p,q)-forms on X. By Dolbeault's theorem, for a X, there is an of vector spaces H^{p,q}(X, E) \cong H^q(X, \Omega^p \otimes E), where \Omega^p is the sheaf of holomorphic p-forms on X. Additionally, Serre duality establishes a natural between H^q(X, E) and H^{n-q}(X, E^* \otimes K_X), where n = \dim X, E^* is the dual bundle, and K_X is the , yielding \dim H^q(X, E) = \dim H^{n-q}(X, E^* \otimes K_X). For line bundles L on a compact X of g, the provides a formula \dim H^0(X, L) - \dim H^1(X, L) = \deg L + 1 - g, which computes the space of global holomorphic sections when higher vanishes, such as for sufficiently positive degree.

Forms and Cohomology

Sheaves of bundle-valued forms

In the context of a holomorphic E over a X, the sheaf \mathcal{A}^{p,q}_E is defined as the sheaf of smooth sections of the \Lambda^p [T^*X](/page/T-X) \otimes \Lambda^q \overline{T}^*X \otimes E. This sheaf generalizes the structure sheaf of smooth functions on X by incorporating the bundle E and the decomposition of the into holomorphic and anti-holomorphic parts according to the complex structure. If E is holomorphic, there exists a subsheaf consisting of holomorphic E-valued (p,q)-forms, which are the sections annihilated by the \overline{\partial} operator in the appropriate sense. Locally, on a coordinate U \subset X where E is trivialized as U \times \mathbb{C}^r, the sections of \mathcal{A}^{p,q}_E over U take the form \sum_{I,J} dz^I \wedge d\overline{z}^J \otimes \sigma_{I,J}, where I and J are multi-indices of lengths p and q respectively, and each \sigma_{I,J} is a smooth local section of the trivial bundle over U. Under functions between overlapping trivializations, the representation transforms via the holomorphic maps on the E-valued coefficients \sigma_{I,J}, combined with the change of coordinates affecting the differentials dz^I and d\overline{z}^J. This ensures that \mathcal{A}^{p,q}_E is well-defined globally as a sheaf on X. The sheaves \mathcal{A}^{0,q}_E for q \geq 0 form part of the Dolbeault complex $0 \to \mathcal{O}_X \otimes E \to \mathcal{A}^{0,1}_E \to \mathcal{A}^{0,2}_E \to \cdots, which provides a fine of the sheaf of holomorphic sections \mathcal{O}_X \otimes E (also denoted \mathcal{O}(E)). This is exact on the level of sheaves, allowing the computation of sheaf cohomology via the hypercohomology of the complex. Note that the degree (0,0) case corresponds to the sheaf of E-valued functions, which contains the sheaf of holomorphic sections as a subsheaf. At the level of germs and stalks, for a point x \in X, the stalk (\mathcal{A}^{p,q}_E)_x is isomorphic to the tensor product of the stalk of (p,q)-forms at x with the E_x. This reflects the local triviality of both the form bundles and E, where sections near x behave like matrix-valued forms tensored with the E_x.

de Rham and Dolbeault cohomology comparison

The of a holomorphic vector bundle E over a X is defined using the complex of E-valued differential forms \Omega^\bullet(X, E), equipped with the total d = \partial + \bar{\partial}. The resulting groups H^k_{\mathrm{dR}}(X, E) are isomorphic to H^k(X, \mathbb{C}) \otimes_{\mathbb{C}} \mathbb{C}^r, where r = \mathrm{rank}(E), capturing topological invariants of the base X scaled by the of E. In contrast, the of E arises from the \bar{\partial}-complex of smooth E-valued (p,q)-forms, yielding groups H^{p,q}_{\bar{\partial}}(X, E) that are isomorphic to the sheaf cohomology H^q(X, \Omega^p \otimes E), reflecting the holomorphic structure of E. On a compact , the \partial \bar{\partial}-lemma provides a precise link between these theories: for any d-closed E-valued form of bidegree (p,q), exactness with respect to d, \partial, or \bar{\partial} is equivalent to exactness with respect to \partial \bar{\partial}. This implies the Hodge decomposition H^k_{\mathrm{dR}}(X, E) \cong \bigoplus_{p+q=k} H^{p,q}_{\bar{\partial}}(X, E), decomposing the topological into purely holomorphic components. Bott-Chern cohomology offers an intermediate refinement, defined via the double complex (\Omega^{p,q}(X, E), \partial, \bar{\partial}) as H^{p,q}_{\mathrm{BC}}(X, E) = \ker(d \colon \Omega^{p,q}(X, E) \to \Omega^{p+q+1}(X, E)) / \operatorname{im}(\partial \bar{\partial} \colon \Omega^{p-1,q-1}(X, E) \to \Omega^{p,q}(X, E)). Unlike de Rham or , Bott-Chern groups need not vanish even when both H^\bullet_{\mathrm{dR}}(X, E) and H^{\bullet,\bullet}_{\bar{\partial}}(X, E) do, highlighting obstructions that blend analytic and topological features of the holomorphic structure. On compact Kähler manifolds, however, Bott-Chern cohomology is isomorphic to both de Rham and Dolbeault cohomologies. On compact complex manifolds without a Kähler structure, primarily detects the of the base manifold X (such as its Betti numbers), scaled by the of E, while probes the existence of holomorphic sections or extensions, often vanishing in higher degrees by Kodaira vanishing theorems under suitable metric conditions. For instance, if E is ample on a projective manifold, H^q(X, \Omega^p \otimes E) = 0 for p+q > \dim X and suitable positivity, whereas groups remain non-trivial if the demands it.

Line Bundles and Classification

The Picard group

A holomorphic over a X is a rank-one holomorphic vector bundle, locally trivialized by holomorphic sections with transition functions g_{ij}: U_i \cap U_j \to \mathbb{C}^* satisfying the Čech cocycle condition g_{ij} g_{jk} = g_{ik} on triple overlaps U_i \cap U_j \cap U_k, where \{U_i\} is an open cover of X. These transition functions ensure the bundle's sections glue holomorphically across the cover. Two holomorphic line bundles are isomorphic if their transition functions differ by a coboundary, meaning there exist holomorphic functions h_i: U_i \to \mathbb{C}^* such that the new transition functions satisfy g'_{ij} = h_i g_{ij} h_j^{-1}, or equivalently, g'_{ij} g_{ij}^{-1} = h_i h_j^{-1}. The set of isomorphism classes of holomorphic line bundles on X, equipped with the operation, forms an known as the \operatorname{Pic}(X). This group is isomorphic to the first group H^1(X, \mathcal{O}_X^*), where \mathcal{O}_X^* is the sheaf of nowhere-vanishing holomorphic functions on X. The captures the obstruction to trivializing the bundle globally, with cocycles corresponding to functions coboundaries. For example, on the \mathbb{CP}^1, the is \operatorname{Pic}(\mathbb{CP}^1) \cong \mathbb{Z}, generated by the isomorphism class of the tautological \mathcal{O}(1), whose powers \mathcal{O}(d) correspond to line bundles of degree d. More generally, \operatorname{Pic}(X) is isomorphic to the quotient of the group of divisors \operatorname{Div}(X) by the subgroup of principal divisors \operatorname{Prin}(X), linking line bundles to meromorphic divisors on X.

Classification via cohomology

The isomorphism classes of rank-r holomorphic vector bundles over a complex manifold X form a pointed set in bijection with the non-abelian sheaf group H^1(X, \mathrm{GL}(r, \mathcal{O}_X)), where \mathcal{O}_X denotes the structure sheaf of holomorphic functions and the base point corresponds to the trivial bundle. This classification stems from describing such bundles via open covers with holomorphic transition functions valued in \mathrm{GL}(r, \mathbb{C}), where two cocycles define isomorphic bundles if they differ by a coboundary (holomorphic transformations). For , this specializes to the \mathrm{Pic}(X) \cong H^1(X, \mathcal{O}_X^\times), parametrizing line bundles up to . Obstructions to the triviality of a holomorphic vector bundle E lie in H^1(X, \mathrm{GL}(r, \mathcal{O}_X)); specifically, E is holomorphically trivial its classifying cocycle is cohomologous to the constant , allowing a global frame of r nowhere-vanishing holomorphic sections. More generally, extensions of holomorphic vector bundles, captured in short exact sequences $0 \to F \to E \to Q \to 0, are classified by elements of \mathrm{Ext}^1(Q, F) \cong H^1(X, \mathrm{Hom}(Q, F)), measuring how E fails to split holomorphically. These groups provide analytic invariants distinguishing non-isomorphic bundles, with vanishing higher H^i(X, \mathrm{GL}(r, \mathcal{O}_X)) = 0 for i \geq 2 on Stein manifolds ensuring no further obstructions. On projective varieties, where the holomorphic and algebraic categories coincide, stable isomorphism classes of holomorphic vector bundles are parametrized by moduli spaces, which are quasi-projective algebraic varieties encoding bundles up to S-equivalence (semistable limits). For the projective line \mathbb{P}^1, the Birkhoff–Grothendieck theorem asserts that every holomorphic vector bundle decomposes uniquely (up to ordering) as a E \cong \bigoplus_{i=1}^r \mathcal{O}_{\mathbb{P}^1}(d_i) of line bundles, with integers d_i determined by the degrees; this provides a complete algebraic extensible to the analytic setting. In cases of fixed topological type (e.g., trivial underlying smooth bundle), the space of holomorphic structures is an open subset of a parametrizing compatible complex structures, though global relies on these obstructions rather than finite-dimensional Grassmannians alone. For compact Riemann surfaces, a key analytic relation links holomorphic vector bundles to unitary representations: by the Narasimhan–Seshadri theorem, stable holomorphic bundles of degree zero correspond precisely to irreducible projective unitary representations of the \pi_1(X, x_0) up to conjugacy, realized via flat unitary connections on the underlying smooth bundle. This highlights how classifies the holomorphic structures, while stability ensures projectivity of the , bridging algebraic and representation-theoretic viewpoints without relying on metrics.

Metrics and Connections

Hermitian metrics

A Hermitian metric on a holomorphic vector bundle E \to X, where X is a , is a assignment to each point x \in X of a positive definite Hermitian inner product h_x: E_x \times E_x \to \mathbb{C} on the fiber E_x \cong \mathbb{C}^r, such that the assignment x \mapsto h_x varies with x. This metric provides a way to measure lengths and angles within each fiber, respecting the structure in the sense that it is defined on the underlying vector bundle. The inner product h_x is sesquilinear, meaning it is \mathbb{C}-linear in the first argument and conjugate-linear in the second, with h_x(\sigma, \tau) = \overline{h_x(\tau, \sigma)} for sections \sigma, \tau, and positive definite, satisfying h_x(\sigma, \sigma) > 0 for all nonzero \sigma \in E_x. In a local trivialization of E over an open set U \subset X, with holomorphic frame \{e_k\}_{k=1}^r, any sections \sigma = \sum f_k e_k and \tau = \sum g_l e_l satisfy h(\sigma, \tau) = \sum_{k,l} \overline{f_k} \, h_{kl} \, g_l, where (h_{kl}) is a smooth positive definite Hermitian matrix on U with h_{kl} = \overline{h_{lk}}. Such a metric induces a smooth unitary frame locally, orthonormal with respect to h, though this frame need not align with the holomorphic structure beyond the smooth variation. On a paracompact X, every holomorphic vector bundle admits a Hermitian , constructed via a subordinate to a locally finite open cover by trivializing sets. Specifically, local Hermitian on each trivialization are glued globally by weighting with the partition functions \{\lambda_i\}, yielding h = \sum_i \lambda_i h_i, which is smooth and positive definite everywhere. This existence relies on the paracompactness of X, ensuring the availability of such .

Chern connections and curvature

Given a holomorphic vector bundle E \to X over a X, equipped with a h, there exists a unique \nabla, called the Chern connection, that is compatible with both the holomorphic structure and the . This means \nabla^{0,1} = \bar{\partial}_E and \nabla h = 0, ensuring that the connection preserves the in the sense that h(\nabla s, t) + h(s, \nabla t) = 0 for sections s, t \in \Gamma(E). Locally, in a holomorphic where the is represented by the matrix h = (h_{\alpha \bar{\beta}}), the (1,0)-part of the is \theta = h^{-1} \partial h, so \nabla = d + \theta with the (0,1)-part being \bar{\partial}. The curvature form \Omega of the Chern connection is defined by \Omega = \nabla^2, which acts on sections as \Omega(s) = \nabla^2 s. In terms of the connection forms, it computes locally as \Omega = \bar{\partial} \theta + \theta \wedge \theta, and due to the holomorphic compatibility, \Omega is a (1,1)-form taking values in \operatorname{End}(E). More explicitly, in a holomorphic frame, \Omega = \bar{\partial} (h^{-1} \partial h), reflecting the metric's influence on the bundle's geometry. This curvature satisfies the Bianchi identity d_\nabla \Omega = 0, ensuring its closedness in de Rham cohomology. The Chern classes of E, which are topological invariants in H^{2k}(X, \mathbb{Z}), arise from the curvature via Chern-Weil theory. Specifically, the k-th Chern class is represented by the cohomology class c_k(E) = \left[ \left( \frac{i}{2\pi} \right)^k \operatorname{Tr}(\wedge^k \Omega) \right], where \wedge^k \Omega denotes the k-th exterior power in the Chern character or via the total class \det\left( \operatorname{Id} + \frac{i}{2\pi} \Omega \right) = \sum c_k(E). These classes are independent of the choice of Hermitian metric and , capturing essential topological features of the bundle. A special class of Hermitian metrics on E are the Hermitian-Einstein metrics, for which the curvature satisfies the condition \operatorname{Ric}(\Omega) = \lambda \operatorname{Id} for some constant \lambda, where \operatorname{Ric}(\Omega) is the trace of \Omega contracted with a Kähler form on X. Such metrics are linked to the stability of the holomorphic vector bundle, as established in foundational results on the existence for stable bundles over compact Kähler manifolds.

References

  1. [1]
    [PDF] Holomorphic Vector Bundles over Riemann Surfaces
    Aug 15, 2021 · Definition 2.3. A holomorphic vector bundle is a complex vector bundle p : E → X, where. E and X are complex manifolds and p is holomorphic, ...
  2. [2]
    [PDF] 0.1 Vector Bundles
    For example, if all φαβ are holomorphic, then E is a holomorphic vector bundle; if all φαβ are. C∞, then E is a C∞ vector bundle; and if all φαβ are continuous, ...
  3. [3]
    [PDF] Grothendieck's Classification of Holomorphic Bundles over ... - arXiv
    Sep 29, 2020 · We say that the rank of the vector bundle V (over R) is k. We can extend this definition to holomorphic vector bundles over a complex manifold.
  4. [4]
    [PDF] Mini course “Introduction to holomorphic vector bundles on compact ...
    The theory of holomorphic vector bundles on complex manifolds has be- come a central field of complex geometry. Classically, the theory of holomor- phic ...<|control11|><|separator|>
  5. [5]
    [PDF] Complex geometry Holomorphic vector bundles, elliptic operators ...
    Jun 25, 2022 · Let E → M be a holomorphic vector bundle. Then ∂. 2. : Ωq(M;E) → Ωq+1(M;E) satisfies ∂. 2. = 0. Definition 5.7. Given a holomorphic vector ...
  6. [6]
    [PDF] Introduction to Complex Geometry
    Aug 11, 2022 · Complex Geometry. 21 / 51. Page 31. Holomorphic vector bundle. Roughly speaking, a holomorphic vector bundle over a complex manifold is a family.
  7. [7]
    [PDF] Class 20. Holomorphic vector bundles and connections (November 7)
    Nov 7, 2024 · Given a holomorphic vector bundle ⇡: E ! M, we let A(U, E) denote the space of smooth sections of E over an open set U ✓ M. Likewise, Ap,q ...
  8. [8]
    [PDF] A Beginner's Guide to Holomorphic Manifolds
    A manifold M is “complex” if TM is a complex vector bundle, and is “holomorphic” if TM is a holomorphic vector bundle. This is in accord with the informal usage ...
  9. [9]
    Daniel Huybrechts - Complex Geometry
    Feb 3, 2013 · Definition 2.2.21 If E is a holomorphic vector bundle on a complex manifold. X, then Hg(X,E) denotes the q-th cohomology of its sheaf of ...<|control11|><|separator|>
  10. [10]
    None
    Summary of each segment:
  11. [11]
    [PDF] Study of algebraic and holomorphic vector bundles
    If Xho1 is not compact, usually Xho1 admits holomorphic vector bundles that are not algebrazable and non-isomorphic algebraic vector bundles which are ...
  12. [12]
    [PDF] MATH 217C NOTES Contents 1. Introduction, Complex Manifolds ...
    Aug 21, 2015 · ... holomorphic vector bundle. For now, if (z1,...,zk) are holomorphic ... sheaf of holomorphic sections of a holomorphic vector bundle E ...
  13. [13]
    None
    Summary of each segment:
  14. [14]
    [PDF] Complex Geometry - UT Math
    Aug 23, 2023 · The exterior powers ΛiE and symmetric powers SiE are the holomorphic vector bundles over X whose fibers are isomorphic to Λi(Ex) and Si(Ex), ...
  15. [15]
    FINITENESS AND CONSTRUCTIBILITY IN LOCAL ANALYTIC ...
    The proof of this theorem is a simple variant of the Cartan-Serre-Schwartz proof for the finiteness of coherent cohomology on a compact complex analytic.
  16. [16]
    [PDF] COMPLEX MANIFOLDS, FALL 2024 Class 1. Holomorphic functions ...
    Aug 27, 2025 · a holomorphic vector bundle E. It induces a mapping ∇: A 1(E) → A ... Likewise, we let Ip(L) denote the sheaf of holomorphic sections of L that.
  17. [17]
    [PDF] holomorophic sections of powers of a line bundle - Howard Jacobowitz
    Γ(X, Lk) is the space of global holomorphic sections X → Lk. • C is a constant, which depends on X and L but not on k. We start in Section 2 with an explicit ...
  18. [18]
    [PDF] Holomorphic line bundles - Joel Fine
    The proof is an exercise. 2.5 Linear systems and maps to projective space. Consider two independent sections s, s0 of a holomorphic line bundle L ! X. Given.
  19. [19]
    [PDF] Complex Analytic and Differential Geometry
    Page 1. Complex Analytic and. Differential Geometry. Jean-Pierre Demailly. Université de Grenoble I. Institut Fourier, UMR 5582 du CNRS. 38402 Saint-Martin d'H ...
  20. [20]
    [PDF] Class 6. The tangent bundle
    Sep 17, 2024 · , the holomorphic tangent bundle of M. To describe a set of transition functions for the tanget bundle, we continue to assume that dim M = n ...
  21. [21]
    [PDF] Kähler Geometry
    they are indeed transition functions. A section of TX is called a holomorphic vector field. The holomorphic cotangent bundle ΩX is the dual of TX. The ...
  22. [22]
    [PDF] Riemann Surfaces
    Dec 12, 2001 · If X is a Riemann surface, Theorem 9. 5 below says that a holomorphic line bundle of sufficiently high degree has no base point and that the ...
  23. [23]
  24. [24]
    [PDF] arXiv:1809.07277v2 [math.AG] 6 Dec 2018
    Dec 6, 2018 · In general, let V be a holomorphic vector bundle over a complex manifold X. Then we can define the Dolbeault cohomology groups of X with values ...
  25. [25]
    [PDF] Complex Manifolds
    We have isomorphisms between Dolbeault cohomology and Čech cohomology. H p,q. ∂. (X, E) ∼= Hq(X, Ωp ⊗ E) where Ωp is the sheaf of holomorphic p-forms. Proof.
  26. [26]
    [PDF] 1. Kähler manifolds - UChicago Math
    Nov 20, 2013 · Since the ordinary Laplacian is a real operator, it commutes with complex con- jugation, which therefore takes Hp,q isomorphically to Hq,p.
  27. [27]
    [PDF] serre duality and applications
    Sep 15, 2013 · Serre duality theory is developed, using spectral sequences and Yoneda pairing. Applications include the Riemann-Roch theorem for curves and ...
  28. [28]
    [PDF] 1 The Riemann-Roch theorem - Kiran S. Kedlaya
    For every line bundle L on C, there is a canonical perfect pairing. H0(C,L) × H1(C,Ω ⊗ L−1) → k. In particular, the two vector spaces have the same dimension.
  29. [29]
    [PDF] a dolbeault-hilbert complex for a variety with isolated singular points
    Apr 23, 2019 · In this section we construct a certain resolution of the sheaf of holomorphic sections of a holomorphic vector bundle V on X. To begin, we ...
  30. [30]
    [PDF] Principles of Algebraic Geometry - agorism.dev
    Algebraic geometry is among the oldest and most highly developed sub- jects in mathematics. It is intimately connected with projective geometry,.
  31. [31]
    [PDF] 1. Overview We look at complex line bundles from the topological ...
    139 Griffiths Harris.) Will show: H1(C) = H2(C) = 0. Cor: H1(C∗) = H2(X; Z) ... this as a vector in CN+1 take any local trivialization hα whose base Uα contains.
  32. [32]
    [PDF] Picard groups for line bundles with connections
    Abstract. We study analogues of the usual Picard group for complex manifolds or non- singular complex algebraic varieties but instead of line bundles we ...
  33. [33]
    Divisors, Picard group and Kodaira embedding theorem
    Dec 13, 2017 · Let X be a compact complex manifold with L a positive holomorphic line bundle on X . Then L is generated by finitely many of its global sections ...
  34. [34]
    [PDF] 13. Topological versus holomorphic classification - UCSD Math
    Of course every holomorphic vector bundle gives rise to a topological vector bundle. A priori, topo- logical vector bundle means that the transition functions ...
  35. [35]
    $$H^1(X, GL(n, \mathcal{O}_X))$ and Vector Bundles - MathOverflow
    Mar 17, 2021 · We know that line bundles on X up to isomorphisms are given by H1(X,O∗X). Can it be generalized to higher rankal vector bundles on X by ...Are all holomorphic vector bundles on a contractible complex ...Extension of a holomorphic vector bundle - MathOverflowMore results from mathoverflow.net
  36. [36]
    [PDF] The theorem of Narasimhan and Seshadri and generalizations
    Theorem of Narasimhan–Seshadri (1965). A holomorphic vector bundle over X is stable if and only if it arises from an irreducible projective unitary ...
  37. [37]
    [PDF] Complex Differential Geometry - Institut für Differentialgeometrie
    deep theorem, Cartan's theorem B, which states that on a Stein manifold (and even on a Stein space) ˇHp(M, F) = 0 for every coherent sheaf F and all p > 0 ...
  38. [38]
    Symplectic and Holomorphic Vector Bundles - MathOverflow
    Nov 4, 2014 · Any C∞-complex vector bundle over a paracompact smooth manifold admits a Hermitian metric, by employing a partition of unity. The imaginary part ...
  39. [39]
    [PDF] connections on vector bundles and characteristic classes
    Jun 6, 2011 · Therefore, the curvature of a holomorphic connection has its (0,2)–part defined by (∂E)2 = 0. It is possible to define a holomorphic structure ...
  40. [40]
    [PDF] 2. Chern connections and Chern curvatures1
    On the holomorphic vector bundle E π. −→ M, we can define the ∂-operator: ∂ : Ap(M,E) −→ Ap+1(M,E). If we take the holomorphic frame e = (e1,··· ,en) ...