Fact-checked by Grok 2 weeks ago

Intersection theory

Intersection theory is a branch of that studies the intersections of subvarieties within an ambient , providing tools to define and compute intersection multiplicities, cycles, and products in structures such as Chow groups or . It formalizes the notion of how two subvarieties "meet," accounting for both transverse and non-transverse cases through refined invariants like the intersection product, which endows the Chow with a . The theory originated in classical during the , with foundational results such as , which states that two curves of degrees d_1 and d_2 intersect in exactly d_1 d_2 points, counting multiplicity and points at , under conditions in . Pioneering work by mathematicians like Chasles, Schubert, and Zeuthen addressed enumerative problems, such as determining the number of conics tangent to five given conics (3,264) or cubics tangent to nine lines (33,616), laying the groundwork for rigorous intersection calculus. In the , the theory was modernized through the development of algebraic cycles and rational equivalence by Chow in the 1940s, followed by Fulton's comprehensive scheme-theoretic framework in the 1980s, which eliminated reliance on classical tools like the moving lemma and extended the theory to singular varieties. Key concepts include Chow groups, which classify cycles modulo rational equivalence, and the intersection product, enabling computations via pushforwards and pullbacks under morphisms; these structures connect to through isomorphisms with rings on smooth projective varieties. Notable applications span —such as counting lines on cubic surfaces (27 lines) or conics through points—moduli spaces of curves, theory, and even via Gromov-Witten invariants. The theory's influence persists in contemporary research, including equivariant and motivic extensions, underscoring its role in bridging classical and modern .

Overview

Definition and motivation

Intersection theory provides a mathematical framework for quantifying the intersections of geometric objects, such as subvarieties in or submanifolds in , by assigning invariants like intersection numbers or classes in associated rings. This approach addresses the need to measure how these objects "meet" in a rigorous way, extending beyond simple point counting to account for geometric complexities like tangencies and higher-dimensional overlaps. The motivation for intersection theory stems from longstanding problems in , where one seeks to count solutions to geometric conditions—such as the number of plane curves of a given passing through specified points—but encounters issues with multiplicities when intersections are not transverse. In , similar challenges arise in defining linking numbers or self-intersections of cycles in manifolds, necessitating invariants that capture topological intersections independently of embeddings. These invariants enable the computation of global properties, such as degrees of maps or Euler characteristics, through local intersection data. A basic prerequisite for understanding intersection theory includes knowledge of smooth manifolds from , singular homology from , and algebraic varieties from , as intersections are often analyzed within these structures. For a simple illustration, consider two distinct lines in the , which intersect transversely at a single point; their is defined to be 1, reflecting the minimal transverse case without multiplicity. In specific contexts, realizations like the topological intersection form on groups or the Chow ring for algebraic cycles provide concrete tools for these computations.

Historical development

Intersection theory originated in 19th-century , addressing problems of counting intersections of algebraic curves and surfaces, with foundational results such as and pioneering work by Chasles, Schubert, and Zeuthen on problems like the number of conics tangent to five given conics. Topological aspects were pioneered by in his 1895 paper "Analysis Situs," where he introduced the concept of intersection numbers for cycles on manifolds, providing a topological measure of how submanifolds intersect, which laid the groundwork for understanding duality in . This was further developed in his 1904 supplement, where he refined these ideas to establish , linking intersection numbers to the topology of closed orientable manifolds. In the 1920s and 1930s, advanced the topological aspects of intersection theory through his work on and s. Lefschetz's 1923 generalized Poincaré's ideas, using intersection numbers to count fixed points of maps on manifolds via traces on groups. He also introduced the intersection form on the middle of 4-manifolds in the 1930s, a capturing self-intersection properties that became central to topological intersection theory. Building on classical foundations, the mid-20th century saw further developments in algebraic intersection theory, beginning with André Weil's contributions in the 1940s. In his 1946 "Foundations of Algebraic Geometry," Weil developed intersection multiplicities and reciprocity laws for divisors on algebraic varieties over arbitrary fields, bridging classical with modern algebraic structures. built on this in the 1950s with his introduction of coherent sheaves in "Faisceaux algébriques cohérents" (1955), enabling intersection products via sheaf cohomology and refining multiplicity computations for non-transverse intersections. Around the same time, Wei-Liang Chow and others developed the Chow ring in the early 1950s, a structure formalizing intersections of cycles modulo rational equivalence on projective varieties. Alexander Grothendieck's work in the 1960s revolutionized the field by reformulating intersection theory in the language of schemes and . Starting in the late 1950s, Grothendieck introduced algebraic to study coherent sheaves and their intersections, culminating in the Grothendieck-Riemann-Roch theorem (1958), which provided a universal transformation for pushforwards in intersection settings. His scheme-theoretic approach, detailed in foundational texts from the early 1960s, unified arithmetic and geometric intersections, while his 1966 talk outlined the standard conjectures on algebraic cycles, positing homological equivalence as sufficient for intersection properties and motivating the theory of motives. The modern synthesis of intersection theory emerged in the , with William Fulton's 1984 book "Intersection Theory" serving as a comprehensive reference that integrated topological and algebraic developments into a general framework for arbitrary schemes. Fulton's treatment encompassed refined intersection products, excess intersection formulas, and applications to blow-ups, drawing on Grothendieck's foundations. Subsequent advances in the and 2000s incorporated motives and derived categories, with Grothendieck's standard conjectures influencing higher and , though many remain open.

Topological intersection theory

Basic intersection numbers

In topological intersection theory, the basic provides a way to quantify the algebraic of cycles in a compact oriented manifold. To define it, one first requires the cycles to intersect transversally. Two submanifolds \alpha and \beta of a manifold M intersect transversally if, at every point p \in \alpha \cap \beta, the spaces satisfy T_p \alpha + T_p \beta = T_p M. This ensures that the is a of the expected dimension \dim \alpha + \dim \beta - \dim M. For cycles representing homology classes in , transversality can be achieved by perturbing one cycle slightly via a , as generic perturbations place intersections in , resulting in a of isolated points when \dim \alpha + \dim \beta = \dim M. For two transverse cycles \alpha and \beta in a compact oriented n-manifold M with \dim \alpha + \dim \beta = n, the [\alpha] \cdot [\beta] is defined as the algebraic count of their points: [\alpha] \cdot [\beta] = \sum_{p \in \alpha \cap \beta} \varepsilon_p, where \varepsilon_p = \pm 1 is the local intersection sign at p, determined by whether the orientation induced on T_p M by the direct sum T_p \alpha \oplus T_p \beta matches the of M. If the orientations agree, \varepsilon_p = +1; otherwise, \varepsilon_p = -1. This number is well-defined on classes since it is invariant under deformations that preserve transversality. The intersection number satisfies several key properties. It is bilinear over \mathbb{Z}, meaning [k\alpha + \alpha'] \cdot [\beta] = k [\alpha] \cdot [\beta] + [\alpha'] \cdot [\beta] and similarly for the second argument. It is skew-commutative: [\alpha] \cdot [\beta] = (-1)^{\dim \alpha \cdot \dim \beta} [\beta] \cdot [\alpha]. Additionally, it is invariant under homotopies of the cycles that maintain transversality, making it a topological . These properties arise from the local nature of the signs and the additivity of the sum over intersections. enables the extension of these numbers to pairings between and groups. In three-dimensional manifolds, the linking number provides a special case of the intersection number. For two disjoint oriented closed curves [\alpha, \beta](/page/Alpha_Beta) \subset M^3, the linking number \mathrm{lk}([\alpha, \beta](/page/Alpha_Beta)) is defined as the intersection number [\alpha] \cdot [\Sigma], where \Sigma is a compact oriented surface (Seifert surface) bounded by \beta. This counts the signed crossings of \alpha through \Sigma, and it is symmetric up to sign: \mathrm{lk}(\alpha, \beta) = -\mathrm{lk}(\beta, \alpha). A representative example occurs with two closed curves on the T^2 = S^1 \times S^1. Consider the meridional curve \alpha (wrapping once around the first factor) and the longitudinal curve \beta (around the second factor); they intersect transversally at one point with positive local sign, yielding [\alpha] \cdot [\beta] = 1. More generally, for curves of types (p, q) and (r, s) (winding p times meridionally and q longitudinally, etc.), the equals the ps - qr, which coincides with the of the induced map from the curves' parametrizations to S^1 \times S^1. This illustrates how intersection numbers capture winding relations on surfaces.

The intersection form on manifolds

In the context of topological intersection theory, the intersection form on a closed oriented manifold provides a fundamental bilinear pairing that encodes global topological information through duality. For a closed oriented $4k-manifold M, the intersection form Q: H^{2k}(M; \mathbb{Z}) \times H^{2k}(M; \mathbb{Z}) \to \mathbb{Z}is defined byQ(\alpha, \beta) = \langle \alpha \cup \beta, [M] \rangle, where \cupdenotes the [cup product](/page/Cup_product) in [cohomology](/page/Cohomology),\langle -, - \rangleis the pairing with the fundamental class[M] \in H_{4k}(M; \mathbb{Z}), and the form arises naturally from [Poincaré duality](/page/Poincaré_duality), which identifies H^{2k}(M; \mathbb{Z})with the dual ofH_{2k}(M; \mathbb{Z})via cap products with[M]$. This pairing extends the basic notion of intersection numbers between homology classes to a symmetric bilinear structure on the middle-dimensional , reflecting the manifold's self-intersection properties. The intersection form possesses several key properties derived from the underlying duality theorems. It is unimodular, meaning its matrix representation with respect to a basis has determinant \pm 1, which follows directly from Poincaré duality ensuring the pairing is non-degenerate. For even dimensions like $4k, the form is symmetric, Q(\alpha, \beta) = Q(\beta, \alpha), whereas in odd middle dimensions it is skew-symmetric; classification of such forms over \mathbb{Z} for $4-manifolds (the k=1 case) is determined by invariants including the rank (dimension of the free part of H^{2k}(M; \mathbb{Z})), the signature (difference between positive and negative eigenvalues of the real extension), and parity (even if all self-intersections Q(\alpha, \alpha) are even integers, odd otherwise). Poincaré-Lefschetz duality underpins the intersection form by relating and via and products. Specifically, for cohomology classes \alpha \in H^p(M; \mathbb{Z}) and \beta \in H^q(M; \mathbb{Z}) with p + q = n = \dim M, the \alpha \cap \beta \in H_{n-p}(M; \mathbb{Z}) satisfies \langle \alpha \cap \beta, [M] \rangle = \langle \alpha, \beta \cup \mathrm{PD}([M]) \rangle, where \mathrm{PD}: H_n(M; \mathbb{Z}) \to H^n(M; \mathbb{Z}) is the Poincaré dual; in the middle dimension, this yields the intersection form as Q(\alpha, \beta) = \langle \alpha \cap \mathrm{PD}(\beta), [M] \rangle, dualizing geometric intersections of submanifolds to algebraic operations on chains. This formulation highlights how the form captures transversal intersections modulo boundaries, with the providing a computable algebraic representative. Representative examples illustrate the diversity of intersection forms. For the complex projective plane \mathbb{CP}^2, a simply connected $4-manifold, the second cohomology H^2(\mathbb{CP}^2; \mathbb{Z}) \cong \mathbb{Z}is generated by the class of a line, yielding the odd unimodular formQ = \langle 1 \ranglewith self-intersection+1.[](https://math.uchicago.edu/~dannyc/courses/4manifolds_2018/4_manifolds_notes.pdf) In contrast, the K3 surface, a simply connected compact $4-manifold with trivial canonical bundle, has H^2(K3; \mathbb{Z}) \cong \mathbb{Z}^{22} and an even unimodular intersection form Q = E_8(-1)^{\oplus 2} \oplus H^{\oplus 3}, where E_8(-1) is the negative definite E_8 lattice and H is the hyperbolic plane \begin{pmatrix} 0 & 1 \\ 1 & 0 \end{pmatrix}, giving signature -16 and rank $22$. A significant application of the intersection form is Rokhlin's theorem, which imposes constraints on structures. For a closed smooth $4-manifold M, the signature \sigma(Q)of the intersection form satisfies\sigma(Q) \equiv 0 \pmod{}, linking the form's algebraic invariants to differential-topological obstructions via the \hat{A}-genus and index theory. This result, sharp as seen in the K3 surface example where \sigma = -, underscores the form's role in classifying $4-manifolds up to .

Intersection theory in algebraic geometry

Cycles and rational equivalence

In , an algebraic of k on a X over an is defined as a formal \mathbb{Z}- \sum n_i V_i, where each n_i \in \mathbb{Z} and each V_i is an irreducible sub of X of k. The set of all such forms a Z^k(X), generated by the irreducible subvarieties of k. This provides the basic building blocks for intersection theory, allowing the study of subvarieties through linear combinations that capture geometric relations. Rational equivalence is an on the groups Z^k(X) introduced to identify cycles that can be "deformed" into one another via rational families. Specifically, two cycles Z, Z' \in Z^k(X) are rationally equivalent if there exists a cycle \Gamma \in Z^k(X \times \mathbb{P}^1) such that Z - Z' = [ \Gamma_0 ] - [ \Gamma_1 ], where \Gamma_t denotes the of \Gamma over the point t \in \mathbb{P}^1 (with points 0 and 1). This relation is generated by differences of divisors of rational functions on subvarieties of X, and it preserves while being compatible with the of X. The Chow groups are the quotients CH^k(X) = Z^k(X) / \sim_{\mathrm{rat}}, where \sim_{\mathrm{rat}} denotes rational equivalence; these groups are graded by , yielding the Chow ring \mathrm{CH}(X) = \bigoplus_k \mathrm{CH}^k(X). The Chow groups admit a filtration by dimension (since k corresponds to dimension \dim X - k) and exhibit functoriality under proper morphisms f: Y \to X, via a f_*: \mathrm{CH}^k(Y) \to \mathrm{CH}^k(X) that preserves the equivalence relation. For example, on \mathbb{P}^n, the Chow groups are \mathrm{CH}^k(\mathbb{P}^n) \cong \mathbb{Z} [H]^k, where H is the class of a , generated by powers of the hyperplane class up to n. Algebraic cycles map to homology classes via the cycle class map \mathrm{cl}: Z^k(X) \to H_{2(\dim X - k)}(X(\mathbb{C}), \mathbb{Z}) (assuming X is defined over \mathbb{C}), which factors through rational equivalence to induce a map on Chow groups. However, rational equivalence is a finer relation than homological equivalence, as rationally equivalent cycles are homologically equivalent, but the converse does not hold in general, allowing Chow groups to distinguish more refined geometric structures than .

The moving lemma

The moving lemma is a cornerstone of intersection theory in , enabling the deformation of within their rational equivalence classes to achieve proper intersections. Specifically, for a X over an and \alpha \in Z_r(X), \beta \in Z_s(X), there exists a cycle \alpha' \in Z_r(X) rationally equivalent to \alpha such that \alpha' intersects \beta properly, meaning that for every irreducible component Z of \alpha' and W of \beta, either \dim(Z \cap W) = \dim Z + \dim W - \dim X or Z \cap W = \emptyset. This formulation extends naturally to the setting of A on X and B on Y by considering the product space X \times Y, which is if X and Y are; here, one can find A' rationally equivalent to A and B' rationally equivalent to B such that A' \times B' intersects a given cycle on X \times Y properly, with codimensions adding appropriately. A proof sketch relies on embedding X into projective space \mathbb{P}^N and using generic linear projections. One projects \alpha via a general linear subspace \Lambda of complementary dimension to deform it rationally; by induction on dimension, this ensures the deformed cycle intersects \beta properly, leveraging the projection formula for cycles. Alternatively, for quasi-projective varieties, the proof employs Bertini's theorem applied to sections of an ample line bundle on the support of \alpha, deforming components to avoid unwanted intersections with \beta. The primary consequence of the moving lemma is that it allows the definition of an intersection product on the Chow groups A_*(X) without requiring initial transversality: given [\alpha], [\beta] \in A_*(X), choose \alpha' as above, define [\alpha] \cdot [\beta] = [\alpha' \cap \beta], and verify independence from the choice of representative via rational equivalence. This extends to relative settings over curves, where cycles can be moved properly in families, facilitating computations in enumerative geometry. A representative example occurs on a smooth projective surface S, where a cycle C \subset S (dimension 1) may initially pass through a point p \in S with multiplicity greater than 1, failing proper . By the moving lemma, there exists C' \sim_{\mathrm{rat}} C such that C' intersects the 0-cycle $$ transversely at distinct points, yielding the \deg(C' \cap ) = C \cdot without singularities. While effective over algebraically closed fields of characteristic zero, the moving lemma fails in positive characteristic without modifications, as Bertini's theorem does not hold in general for hypersurface sections, requiring alternative approaches like generic projections avoiding characteristic divisors or Frobenius techniques to ensure proper deformations.

Intersection products and multiplicities

In , the intersection product on the Chow groups of a smooth projective variety X provides a way to multiply cycle classes. For cycle classes \alpha \in \mathrm{CH}^k(X) and \beta \in \mathrm{CH}^l(X), their product \alpha \cdot \beta \in \mathrm{CH}^{k+l}(X) is defined by first using the moving lemma to represent \alpha and \beta by cycles whose supports intersect properly, meaning the dimension of the intersection equals the \dim X - k - l. The product is then the class of the sum over the components of the intersection, each weighted by its local intersection multiplicity. The local intersection multiplicity at a point p \in Z \cap W, where Z and W are subvarieties of X intersecting properly at p, quantifies the "order of contact" and is given by \mathrm{mult}_p(Z, W) = \sum_{i \geq 0} (-1)^i \length(\Tor_i^{\mathcal{O}_{X,p}}(\mathcal{O}_{Z,p}, \mathcal{O}_{W,p})), where \length denotes the length as an \mathcal{O}_{X,p}-module. Equivalently, in local coordinates where Z and W are defined by ideals generated by regular sequences f_1, \dots, f_k and g_1, \dots, g_l respectively, with k + l = \dim X - \dim(Z \cap W), the multiplicity is \dim_{\mathbb{C}} \mathbb{C}[x_1, \dots, x_n]_p / (f_1, \dots, f_k, g_1, \dots, g_l). This definition satisfies properties such as positivity and invariance under rational equivalence. When the intersection is not proper, meaning \dim(Z \cap W) > \dim X - \codim Z - \codim W, the excess intersection formula refines the naive intersection class using sheaf theory. Specifically, if i: Z \to X and j: W \to X are regular embeddings, the refined product [Z] \cdot [W] in \mathrm{CH}_*(Z \cap W) is given by [Z \cap W] + higher terms involving the Chern classes of the excess bundle, which is the quotient of the conormal sheaf of Z \cap W in W by the pullback of the conormal sheaf of Z in X. More precisely, if q: C_Z W \to Z \cap W is the projection from the normal cone of Z in W, then the class is c(E) \cap [Z \cap W], where E is the excess normal bundle \ker(N_{Z \cap W / W} \to i^* N_{Z/X}). This allows computation even when dimensions exceed expectations, such as in self-intersections or when subvarieties share components. A concrete example occurs with two curves in \mathbb{P}^3: consider the curve C_1 parametrized by [s^3 : s^2 t : s t^2 : t^3] and a line C_2 that is to it at the point [1:0:0:0], such as the line joining [1:0:0:0] and a point where the matches. Their at this point has multiplicity 2, computed via the of the defining equations after dehomogenization or by evaluating the Tor length in the local ring, reflecting the shared . The intersection product satisfies key properties that make the Chow groups into a . It is associative: (\alpha \cdot \beta) \cdot \gamma = \alpha \cdot (\beta \cdot \gamma) for \alpha \in \mathrm{CH}^k(X), \beta \in \mathrm{CH}^l(X), \gamma \in \mathrm{CH}^m(X). It is graded commutative up to sign: \alpha \cdot \beta = (-1)^{kl} \beta \cdot \alpha. Additionally, the projection formula holds: for a f: Y \to X and \alpha \in \mathrm{CH}_*(Y), \beta \in \mathrm{CH}_*(X), we have f_*(\alpha \cdot f^*\beta) = f_*(\alpha) \cdot \beta. These ensure compatibility with pullbacks and pushforwards, enabling applications in enumerative problems.

The Chow ring

The Chow ring of an X, denoted A(X) or CH^*(X), is the formed by the \bigoplus_k CH^k(X), where CH^k(X) is the Chow group of -k algebraic cycles on X modulo rational equivalence, and the multiplication is given by the product of cycles. The ring is , with the product \alpha \cdot \beta = (-1)^{kl} \beta \cdot \alpha for \alpha \in CH^k(X) and \beta \in CH^l(X), and the unit element is the class of X itself in CH^0(X). This structure captures the of intersections on X, extending the additive groups of Chow classes to a multiplicative framework. Fulton's axiomatic approach defines the intersection product via three key properties: the moving axiom, which allows deforming cycles to achieve proper intersections while preserving rational equivalence; the excess intersection axiom, which provides a refined formula for intersections where the expected dimension is not met, involving normal cone deformations; and the specialization axiom, which ensures compatibility under flat families or deformations. These axioms guarantee that the resulting ring structure on the Chow groups is unique up to for any , independent of choices in the construction. For the projective space \mathbb{P}^n, the Chow ring is A(\mathbb{P}^n) \cong \mathbb{Z} / (h^{n+1}), where h is the class of a in codimension 1, and the grading is by \deg h = 1. More generally, for a X over \mathbb{Q}, the Chow ring A(X) \otimes \mathbb{Q} is generated by the Chern classes of vector bundles on X. The map in Chow groups induces a \deg: A_{\dim X}(X) \to \mathbb{Z}, obtained by pushing cycles of dimension \dim X forward to a point, which corresponds to the classical topological in the analogy. The Chow ring relates to the K_0(X) of vector bundles on X through the \gamma-filtration, where the associated graded pieces recover the Chow groups via the Chern character. Fulton's Riemann-Roch theorem without denominators expresses the in K_0 in terms of the Todd class and intersection products in the Chow ring, avoiding holomorphic forms and providing a purely algebraic statement.

Self-intersection classes

In intersection theory, the self-intersection class of a subvariety Z \subset X of k in a variety X is defined as the product [Z] \cdot [Z] \in CH^{2k}(X) in the Chow ring, where this product is understood via the refined intersection theory on the normal cone C_Z X or, when applicable, deformation to the normal bundle N_{Z/X}. For a subvariety Z, the self-intersection class admits an explicit formula: [Z]^2 = c_k(N_{Z/X}) \cap [Z], where c_k denotes the top of the normal bundle. This expression arises from the self-intersection axiom in Fulton's refined theory, where the intersection [Z] \cdot [Z] is refined to a class on Z via the Gysin map from the normal cone, coinciding with the normal bundle in the smooth case. Geometrically, the self-intersection class encodes the "obstruction" or intrinsic intersection behavior of Z with itself, measuring how points of Z are counted with multiplicity under infinitesimal deformations within X; it vanishes in exceeding \dim X, but in general captures the embedding's rigidity or thickness. A representative example occurs for a line L \subset \mathbb{P}^2, where the normal bundle N_{L/\mathbb{P}^2} \cong \mathcal{O}_L(1) has top of 1, yielding self-intersection [L]^2 = 1. Similarly, for a projective curve C of g, the self-intersection of its canonical class [K_C] \in CH^1(C) is the \deg K_C = 2g - 2, obtained via the relating the to the genus. A focal case is the self-intersection of the diagonal subvariety \Delta \subset X \times X, which has codimension \dim X and normal bundle isomorphic to the pullback of the tangent bundle TX; its refined self-intersection class [\Delta] \cdot [\Delta] in CH^{2 \dim X}(X \times X) is represented by the structure sheaf of the scheme-theoretic intersection, involving Tor terms from the resolution of \mathcal{O}_\Delta. The geometric significance emerges in the Grothendieck-Riemann-Roch theorem, where the pushforward under the projection X \times X \to X of this self-intersection relates directly to the Todd class td(X), providing the multiplicative characteristic class that generalizes the Hirzebruch-Riemann-Roch formula \chi(X, F) = \int_X ch(F) \cdot td(X) for coherent sheaves F. This connection underscores how self-intersections of diagonals encode global topological invariants like Euler characteristics through algebraic cycle classes.

Blow-ups and resolution

In algebraic geometry, the blow-up of a scheme X along a coherent ideal sheaf \mathcal{I} \subset \mathcal{O}_X is defined as the scheme \mathrm{Bl}_\mathcal{I} X = \mathrm{Proj}_X \bigoplus_{d \geq 0} \mathcal{I}^d, equipped with the natural projection morphism \pi: \mathrm{Bl}_\mathcal{I} X \to X. This morphism is proper and birational, and the inverse image \pi^{-1}(\mathcal{I}) \cdot \mathcal{O}_{\mathrm{Bl}_\mathcal{I} X} is an invertible sheaf corresponding to the exceptional divisor E, which is an effective Cartier divisor. The support Z = V(\mathcal{I}) of the ideal sheaf is the closed subscheme along which the blow-up occurs, and the exceptional divisor is given by E = \pi^{-1}(Z), with fibers over points of Z forming the projectivization \mathbb{P}(\mathrm{Sym}(\mathcal{I}/\mathcal{I}^2)^\vee) of the conormal sheaf when the embedding is regular. The \pi^* induced by morphism preserves products rationally in the Chow groups. Specifically, for cycles \alpha and \beta on X, the pullback satisfies \pi^*(\alpha \cdot \beta) = \pi^*\alpha \cdot \pi^*\beta up to rational equivalence. On the blow-up \mathrm{Bl}_\mathcal{I} X, which is a \mathbb{P}^r-bundle over Z (where r is the ), the self- of the exceptional divisor [E] relates to the tautological bundle via [E]^2 = -[\mathrm{PD}(\pi^* \mathcal{O}(1))] \cdot [E], where \mathrm{PD} denotes the Poincaré dual class and \mathcal{O}(1) is the relative on the projective bundle. This reflects the negative self-intersection arising from the geometry of the projective fibers. Blow-ups are a key tool for resolving singularities in algebraic varieties. By iteratively blowing up along suitable centers (such as singular loci or non-normal-crossing components), one obtains a : a proper birational \rho: \tilde{X} \to X where \tilde{X} is and the exceptional locus has crossings. In characteristic zero, Hironaka's theorem guarantees the existence of such a for any variety over a of characteristic zero, achievable through a finite sequence of blow-ups along centers. A concrete example is of \mathbb{P}^2 at a point p. The resulting surface \mathrm{Bl}_p \mathbb{P}^2 is a of degree 8, with exceptional divisor E \cong \mathbb{P}^1 having self-intersection [E]^2 = -1. This negative self-intersection ensures that E is a rigid exceptional , and the anticanonical class -K_{\mathrm{Bl}_p \mathbb{P}^2} = \pi^*(-K_{\mathbb{P}^2}) - E remains ample, highlighting the variety's property post-blow-up. For proper morphisms \pi: Y \to X, the and operations in intersection theory satisfy the projection formula: \pi_*(\pi^* \alpha \cdot \beta) = \alpha \cdot \pi_* \beta for \alpha \in A^k(X) and \beta \in A_*(Y), where A_* denotes the Chow group. This compatibility allows intersections on to be pushed forward to compute refined numbers on the base, preserving rational equivalence and facilitating computations in singular settings after .

Extensions and applications

Enumerative geometry

Enumerative geometry employs to solve problems of counting the number of geometric objects, such as curves or surfaces, that satisfy specified incidence conditions like passing through given points or intersecting given subvarieties in . Classical results often rely on the intersection product in the Chow ring to compute these numbers rigorously, assuming of the conditions. For instance, the number of conics passing through 5 general points in the \mathbb{P}^2 is 1; this follows from the fact that the parameter space of conics is \mathbb{P}^5, and the 5 point conditions correspond to 5 general sections whose has 1. Similarly, in \mathbb{P}^3, the number of lines intersecting 4 general lines is 2, computed as the of the product of 4 copies of the Schubert class \sigma_{1,0} in the Chow ring of the \mathrm{Gr}(2,4). The assumption of is justified by Bertini's theorem, which states that for a projective with irreducible fibers, the general member of the family intersects a fixed ample properly and with the expected . In the context of intersection theory, this ensures that the imposed conditions lead to transverse intersections of the expected . A key generalization is the Kleiman-Bertini theorem, which extends Bertini to the intersection of two families of subvarieties: for general members of ample families over a projective , their scheme-theoretic intersection has the expected and is equidimensional. This theorem underpins many enumerative computations by guaranteeing that generic choices yield the predicted counts without excess components or multiplicities. Modern enumerative geometry extends these classical counts using Gromov-Witten invariants, which are defined as virtual intersection numbers on the moduli space \overline{\mathcal{M}}_{g,k}(X,\beta) of stable maps from genus-g curves with k marked points to a projective variety X of class \beta. These invariants generalize classical intersection numbers by incorporating virtual fundamental classes to handle non-transverse or obstructed situations, particularly for pseudoholomorphic curves in symplectic geometry and algebraic stable maps. For example, the Gromov-Witten invariant \langle \mathrm{pt}^{3d-1} \rangle_{0,d}(\mathbb{P}^2) counts the number of rational degree-d curves in \mathbb{P}^2 passing through $3d-1 general points, yielding the Kontsevich-Manin numbers N_d, which satisfy recursive relations derived from degeneration to nodal curves. In \mathbb{P}^3, the analogous invariant \langle \mathrm{pt}^{2d} \rangle_{0,d}(\mathbb{P}^3) counts rational degree-d curves through $2d general points, with computations often involving localization on toric varieties or relations to quantum cohomology. Such invariants refine classical enumerative problems through quantum corrections, particularly in the context of mirror symmetry and , where higher-genus contributions account for worldsheet instantons of genus greater than zero. These higher-genus Gromov-Witten invariants modify the classical counts by including terms that reflect multiple covers and higher topology, providing a more complete picture of curve enumerations in Calabi-Yau manifolds used in physical models. For instance, on quintic threefolds, genus-zero invariants count rational curves through points, while genus-one terms offer corrections essential for matching mirror symmetry predictions. Classical Schubert calculus on Grassmannians also facilitates enumerative counts involving higher-degree objects via the Chow ring structure, which is generated by Schubert classes with relations given by the Pieri and Giambelli formulas. An illustrative computation arises in determining the number of lines in \mathbb{P}^3 intersecting 4 general degree-d curves, which is $2d^4; this is the intersection number of the classes d \sigma_{1,0} (the class for intersecting a degree-d curve) raised to the fourth power in A^*(\mathrm{Gr}(2,4)). This approach extends to more complex incidences, linking linear spaces to higher-degree curves through pullbacks and pushforwards in intersection theory.

Connections to other fields

Intersection theory, originally developed in algebraic geometry and topology, finds significant extensions in symplectic geometry through the theory of Gromov-Witten invariants. These invariants are defined by counting pseudoholomorphic curves in a , providing a refinement of the algebraic intersection numbers when the manifold admits a compatible Kähler structure. In particular, J-holomorphic curves, which are holomorphic with respect to an almost complex structure compatible with the form, allow for the computation of symplectic invariants that generalize and sometimes correct the enumerative counts from . In arithmetic geometry, intersection theory is generalized to arithmetic varieties over number fields, incorporating both Archimedean and non-Archimedean places through arithmetic Chow groups. These groups combine algebraic cycles with metrics at infinite places and measures at finite places, enabling the definition of intersection products that yield arithmetic intersection numbers. A key application arises on abelian varieties, where the Néron-Tate height pairing serves as an intersection form on the Mordell-Weil group, pairing points via a quadratic form that regulates their arithmetic height and connects to the Birch and Swinnerton-Dyer conjecture. This pairing extends classical intersection theory by accounting for the arithmetic geometry over the ring of integers. The motivic perspective further unifies these developments by embedding intersection theory into Voevodsky's triangulated categories of motives over a , where motives of varieties are constructed using correspondences and transfers. In this , intersection products correspond to tensor products of motive objects, preserving homological and numerical equivalences. The standard conjectures on algebraic cycles posit that numerical equivalence coincides with homological equivalence in these categories, implying deep relations between intersection multiplicities and groups. Connections to algebraic are established via Adams operations on the K_0(X) and the Chern character, which maps K_0(X) \otimes \mathbb{Q} isomorphically to the rational Chow ring A^*(X) \otimes \mathbb{Q}. The Adams operations \psi^k act as power operations compatible with the Chern character, allowing intersection-theoretic data to inform K-theoretic invariants and vice versa. An illustrative example occurs in of arithmetic surfaces, where l-adic intersection numbers provide finite-place contributions to arithmetic intersection pairings, complementing the Archimedean components and facilitating computations in the arithmetic Chow groups.

References

  1. [1]
    [PDF] Intersection Theory in Algebraic Geometry and Applications
    Jul 13, 2018 · This three part series will focus on the basics of intersection theory in algebraic geometry. An emphasis.
  2. [2]
    [PDF] INTERSECTION THEORY IN ALGEBRAIC GEOMETRY 1. 1/27/20
    Jan 27, 2020 · Definition. A projective manifold is a closed complex submanifold of CPN . Theorem. (Chow) Any projective manifold is cut out by polynomial ...
  3. [3]
    [PDF] Intersection theory, characteristic classes, and algebro-geometric ...
    Dec 31, 2020 · Intersection Theory is a venerable branch of Algebraic Geometry, with roots grounded in the very origin of the subject and many ramifications ...
  4. [4]
    Intersection Theory - SpringerLink
    The aim of this book is to develop the foundations of this theory, and to indicate the range of classical and modern applications.
  5. [5]
    Intersection Theory | SpringerLink
    Jun 29, 2013 · The aim of this book is to develop the foundations of intersection theory, and to indicate the range of classical and modern applications.
  6. [6]
    [PDF] 3264 & All That Intersection Theory in Algebraic Geometry
    Page 1. 3264 & All That. Intersection Theory in Algebraic Geometry c David Eisenbud and Joe Harris. Page 2. Page 3. Contents. Preface xv. Chapter 0 Introduction.
  7. [7]
    [PDF] Motivating Intersection Theory via Enumerative Geometry
    Apr 27, 2015 · How many points lie on two distinct conics? Each conic is a degree 2 curve in projective space so they intersect in. 2 · 2 = 4 points. Traves.
  8. [8]
    [PDF] Victor Guillemin - Alan Pollack - UCSB Math
    In Chapter 3 we reconstruct intersection theory to include orientations. The Euler characteristic is defined as a self-intersection number and shown to vanish ...
  9. [9]
    [PDF] Intersection theory - University of Notre Dame
    Oct 31, 2011 · INTERSECTION THEORY. 3. A co-orientation on a closed submanifold S ⊂ M is by definition an orientation on the normal bundle. TSM. Thus, an ...
  10. [10]
    [PDF] 2.4 Intersection theory
    Intersection theory also allows us to define the degree of a map modulo 2. ... For an n-manifold M, the set TM has a natural topology and smooth structure which.
  11. [11]
    [PDF] Poincar´e and Analysis Situs, the beginning of algebraic topology
    Sep 3, 2012 · In 1895 Henri Poincaré published his topological work 'Analysis Situs'. ... cal intersection numbers. He also claimed the independence under ...
  12. [12]
    [PDF] INTERSECTION HOMOLOGY THEORY
    Poincart, in his 1895 paper which founded modern algebraic topology ([18], p. 218; corrected in [19]), studied the intersection of an i-cycle V and a j-cycle W ...
  13. [13]
    Solomon Lefschetz | Biographical Memoirs: Volume 61
    In his first proof of the fixed-point theorem in 1923 (1923, 1), Lefschetz made the additional assumption that X is an orientable closed n-manifold. One can ...
  14. [14]
    [PDF] The Riemann Hypothesis over Finite Fields - James Milne
    Sep 14, 2015 · His proof makes use of the general theory of intersection multiplicities in Weil's Foundations. Intersection multiplicities in which one of ...
  15. [15]
    Coherent sheaf - Wikipedia
    The fundamental technical tool in algebraic geometry is the cohomology theory of coherent sheaves. Although it was introduced only in the 1950s, many ...
  16. [16]
    Algebraic K-theory - Wikipedia
    K-theory was discovered in the late 1950s by Alexander Grothendieck in his study of intersection theory on algebraic varieties. In the modern language, ...
  17. [17]
    [PDF] Proceedings 3 - International Mathematical Union
    It was in- troduced by Grothendieck in order to obtain a cohomology theory which one could use to state Weil's analogues to the Riemann Hypo- thesis for higher ...
  18. [18]
    [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 ...
  19. [19]
    [PDF] Notes on 4-manifolds - UChicago Math
    Jan 30, 2019 · Intersection forms. If W is a closed, oriented 4-manifold, the intersection pairing on H2(W;Z) is the quadratic form defined by cup product:.
  20. [20]
    [PDF] The stable classification of 4-manifolds Notes from the lectures of ...
    4 The Intersection form of a 2k-manifold. 7. The intersection form on a closed oriented 4k-dimensional manifold M is defined to be: Free(H2k(M)) ⊗ Free(H2k ...
  21. [21]
    [PDF] Algorithms in 4-manifold topology - School of Mathematics & Statistics
    note that it follows from Poincaré duality that the intersection form is unimodular if and only if. X is closed or if ∂X is an integral homology 3-sphere.
  22. [22]
    [PDF] Four Manifold Topology
    Feb 13, 2019 · A closed smooth 4-manifold X is called irreducible if in any smooth splitting X = X1#X2 one. Xi is a homotopy sphere. For d ≥ 4, the surface V ( ...
  23. [23]
    [PDF] Cup product and intersections - Berkeley Math
    Mar 15, 2011 · This requires the following lemmas about intersection theory in X × X. Recall first that for any topological spaces X and Y there is a homology.
  24. [24]
    [PDF] Cup product and intersection - Brandeis
    Mar 28, 2005 · The theorem says roughly that on a manifold, cup product is Poincaré dual to intersection of submanifolds. In my opinion, this is the most ...
  25. [25]
    [PDF] Basic Properties of K3 Surfaces
    Feb 5, 2020 · Thus, the intersection form has rank 22, even type, and the dimension of the positive and negative eigenspaces are 3 and 19 respectively. Thus, ...Missing: parity | Show results with:parity
  26. [26]
    [PDF] Lecture 8 Essential Example: the K3 Surface - SUSTech Topology
    Apr 29, 2021 · Let X be a K3 surface. Then. (a) H2(X;Z) = 22;. (b) QX = (−E8)⊕2 ⊕ H⊕3. Proof. We only prove (a), while the proof of (b) uses some ...
  27. [27]
    [PDF] Rochlin's theorem
    Let M be a closed, oriented smooth spin. 4-manifold. Then the signature of M is divisible by 16. The theorem is also sharp: an example of a spin 4-manifold with.
  28. [28]
    [PDF] Rochlin's theorem on signatures of spin 4-manifolds via algebraic ...
    If M4 is a smooth spin closed 4-manifold, then σ(M4) ≡ 0 (mod 16). Remark ... such that ∞ ∉ Mk (including the empty manifold) modulo the equivalence relation of.
  29. [29]
    On Equivalence Classes of Cycles in an Algebraic Variety - jstor
    3, November, 1956. Printed in U.S.A.. ON EQUIVALENCE CLASSES OF CYCLES IN AN ALGEBRAIC VARIETY*. BY WET-LIANG CHOW. (Received December 30, ...Missing: original | Show results with:original
  30. [30]
    Section 43.24 (0B0D): Moving Lemma—The Stacks project
    The moving lemma states that given an r-cycle \alpha and an s-cycle \beta there exists \alpha ', \alpha ' \sim _{rat} \alpha such that \alpha ' andMissing: XB YX × Y smooth
  31. [31]
  32. [32]
  33. [33]
    [PDF] Fulton.Laz.Positivity.Excess.Intersection.pdf
    We give several simple applications and related results, including a lower bound for the of a proper intersection, generalizing a classical result for curves on ...
  34. [34]
    Section 43.26 (0B0G): Chow rings—The Stacks project
    - **Definition of Chow Ring**: For a nonsingular projective variety \(X\), the Chow ring \(CH^*(X)\) is defined with groups \(CH^c(X) = CH_{\dim X - c}(X)\), where \(CH^c(X)\) is the Chow group of codimension \(c\) cycles, using rational equivalence.
  35. [35]
    [PDF] INTERSECTION THEORY CLASS 17
    Nov 17, 2004 · Answer: Excess intersection formula. For any α ∈ AkY00, i! (α) = ce(q∗E) ∩ (i0)!
  36. [36]
    Self-intersection number in Projective Space - Math Stack Exchange
    Feb 5, 2016 · Consider a line L in the projective plane CP2: it has self-intersection number 1 since all other lines cross it once: one can push L off to L′, ...Self-intersection number of a complex curve in complex projective ...Geometric motivation for negative self-intersectionMore results from math.stackexchange.com
  37. [37]
    [PDF] on the arithmetic self-intersection number of the dualizing sheaf for ...
    In this article we improve the upper bound for the arithmetic self-intersection number of the dualizing sheaf of the minimal regular model for the Fermat curves ...
  38. [38]
    [PDF] Intersection Theory
    More specifically, the basic goal is to understand intersections of subvarieties in a given ambient variety.
  39. [39]
    [PDF] 18.726 Algebraic Geometry - MIT OpenCourseWare
    Hirzebruch noticed that the Euler characteristic aspect of Riemann-Roch could be general ized to handle arbitrary vector bundles on arbitrary smooth varieties ...
  40. [40]
    [PDF] FOUNDATIONS OF ALGEBRAIC GEOMETRY CLASSES 49 AND 50
    May 23, 2006 · I'll start by defining the blow-up using the universal property. The disadvantage of starting here is that this definition won't obviously be ...
  41. [41]
    [PDF] INTERSECTION THEORY CLASS 8
    Oct 18, 2004 · INTERSECTION THEORY CLASS 8. RAVI VAKIL. CONTENTS. 1. Proof of key result of Chapter 2. 1. 1.1. Crash course in blowing up. 1. 1.2. Back to the ...
  42. [42]
  43. [43]
    A general homological Kleiman--Bertini theorem - MSP
    Speiser's result implies that the generic intersection gW ∩Y, for g ∈ U, is also smooth. We also give a more general homological version of this. For simplicity ...
  44. [44]
    Curve-Counting and Mirror Symmetry
    For example, someone wishing to study the number of conics passing through five points in the plane might form a moduli space ℳ in which each point corresponds ...
  45. [45]
    [PDF] Symplectic Gromov-Witten invariants
    Gromov-Witten (GW) invariants have a rather interesting and involved history, with connections to gauge theory, quantum field theory, symplectic geometry and.
  46. [46]
    [PDF] Arithmetic intersection theory - Numdam
    Feb 2, 2020 · This paper describes an intersection theory for arithmetic varieties which generalizes the work of Arakelov and others on arithmetic ...
  47. [47]
    [PDF] arXiv:1809.06791v2 [math.NT] 3 Apr 2019
    Apr 3, 2019 · EXPLICIT ARITHMETIC INTERSECTION THEORY AND COMPUTATION OF NÉRON-TATE HEIGHTS ... Tate pairing on polarized abelian varieties, Ann. of Math ...
  48. [48]
    [PDF] Triangulated categories of motives over a field.
    Triangulated categories of motives over a field. 1 Introduction. (k) generated by “motives” of smooth varieties.
  49. [49]
    [PDF] Lecture 14 Comparing K-theory and the Chow groups
    There is a “Chern character” map from K-theory to Chow. The Grothendieck–. Riemann–Roch theorem implies that it becomes an inverse after tensoring with. Q.