Fact-checked by Grok 2 weeks ago

Complete intersection

In and , a complete intersection refers to an or a subvariety defined by the minimal number of equations necessary to achieve its , ensuring a particularly simple and well-behaved . Specifically, in a commutative Noetherian ring R, an I \neq R is a complete intersection if its height h(I) equals the minimal number of generators \mu(I), meaning I can be generated by a regular sequence of length equal to its height. Geometrically, for a subvariety Y \subset X in an algebraic variety X, Y is a complete intersection if the ideal sheaf \mathcal{I}(Y) in the structure sheaf \mathcal{O}(X) is an ideal-theoretic complete intersection, corresponding to Y being the common zero set of codimension-many hypersurfaces. The concept distinguishes between local and global complete intersections: an ideal I is a local complete intersection if, for every m of R, the localization I_m is a complete intersection in R_m. In the geometric setting, a is a local complete intersection if it is locally defined this way at every point, which includes all subvarieties but excludes more singular or complicated objects. A related but weaker notion is the set-theoretic complete intersection, where the radical of I is generated by codimension-many elements, allowing for non-ideal-theoretic definitions but still capturing the "minimal intersection" idea geometrically. Complete intersections are fundamental due to their strong homological properties; for instance, quotient rings by complete intersection ideals are Cohen-Macaulay, with depth equal to dimension, and they exhibit finite projective dimension over regular rings. In , they arise naturally in , deformation theory, and enumerative problems, where their simplicity facilitates computations of invariants like Hilbert series or ; for example, the normal bundle of a complete intersection subvariety in is a of line bundles, providing a simple structure that aids in such calculations. These objects contrast with Gorenstein rings or other classes, highlighting their role in classifying rings and varieties with "nice" singularities.

Fundamentals

Definition

In , a closed subvariety X of an Y is a local complete intersection if the ideal sheaf \mathcal{I}_X of X in Y is locally generated by a whose length equals the of X in Y. A sequence of elements f_1, \dots, f_k in a A (such as the stalk of the structure sheaf at a point) forms a if f_1 is a non-zerodivisor in A, f_2 is a non-zerodivisor in A/(f_1), and inductively, each f_i is a non-zerodivisor in A/(f_1, \dots, f_{i-1}) for i = 2, \dots, k. This condition implies that the ideal (f_1, \dots, f_k) has height exactly k, ensuring the has the expected dimension. The local complete intersection property is defined pointwise via local rings, but a global complete intersection arises when the ideal is generated globally by such a , often as the intersection of hypersurfaces in the case of projective varieties. In \mathbb{P}^n, for instance, a subvariety is a global complete intersection if its homogeneous ideal is generated by c homogeneous polynomials forming a , where c is the . The equality between the length of the and the guarantees that X is equidimensional, as the minimal number of generators matches the height of the ideal, preventing embedded components or unexpected drops. Local complete intersections are locally Cohen-Macaulay.

Basic properties

A complete intersection , defined as the of a by an ideal generated by a , exhibits the Cohen-Macaulay property, wherein the depth equals the . This equality ensures that the behaves well homologically, with no discrepancies between local dimensions and the ring's geometric . The ideal of a complete intersection admits a minimal free resolution given by the on its generators. This resolution is exact because the generators form a , providing a concrete and efficient way to compute projective dimensions and other homological invariants. For schemes that are local complete intersections, Serre duality applies directly, pairing groups via the dualizing sheaf, which simplifies due to the regular sequence defining the embedding. This duality manifests straightforwardly without additional complications from non-regular elements, facilitating computations in sheaf . Complete intersection rings are equidimensional, meaning all prime ideals have the same height relative to the , and they contain no embedded points, as the Cohen-Macaulay condition precludes associated primes of lower . This purity of support ensures that the scheme's components are uniformly dimensioned without extraneous lower-dimensional loci.

Illustrative cases

Examples

A in affine or , defined as the zero locus of a single equation, is a complete intersection of 1, as it satisfies the condition with a single generator. For instance, defined by x^2 + y^2 = [1](/page/1) in \mathbb{A}^2 over the complex numbers is such a hypersurface. The intersection of two hypersurfaces provides another fundamental example, particularly when the defining polynomials form a of length equal to the . In \mathbb{P}^3, the zero locus of two homogeneous quadratic polynomials, assuming they intersect transversely, yields a smooth of degree 4 and 1, known as an , which is a complete intersection of 2. This embedding realizes the elliptic curve as the common zeros of the two quadrics, illustrating how complete intersections can capture low-genus curves in . Certain determinantal varieties, which are loci where a generic has at most a fixed value, arise as complete intersections in low-dimensional cases. For example, the of \mathbb{P}^1 \times \mathbb{P}^1 into \mathbb{P}^3 defines a surface as the zero set of a single , corresponding to the 2-by-2 s of a $2 \times 2 of coordinates, making it a determinantal complete intersection of 1. This variety exemplifies how bilinear conditions can produce complete intersections via minor determinants. Over finite fields, complete intersections retain their , with s serving as straightforward instances. The Fermat x^d + y^d + z^d = 0 in \mathbb{P}^2 over \mathbb{F}_q is a complete intersection whose rational points can be enumerated using character sums, highlighting arithmetic applications. Such examples demonstrate the uniformity of the complete intersection property across base fields.

Non-examples

The twisted cubic curve in \mathbb{P}^3, parametrized by (1 : t : t^2 : t^3), provides a classical example of a curve that is not a complete intersection. This curve has codimension 2 in \mathbb{P}^3, so it would be a complete intersection if its homogeneous ideal were generated by exactly two elements forming a regular sequence. However, the ideal is minimally generated by three independent quadrics, specifically the $2 \times 2 minors of the matrix \begin{pmatrix} x_0 & x_1 & x_2 \\ x_1 & x_2 & x_3 \end{pmatrix}, which are x_1^2 - x_0 x_2, x_2^2 - x_1 x_3, and x_0 x_3 - x_1 x_2. This excess of generators over the codimension prevents the ideal from being generated by a regular sequence of length 2. A related non-example is the rational normal curve of degree 4 in \mathbb{P}^4, parametrized by (1 : t : t^2 : t^3 : t^4). This curve also has 3, but its homogeneous ideal requires six minimal generators, consisting of the $2 \times 2 minors of the catalecticant matrix \begin{pmatrix} x_0 & x_1 & x_2 & x_3 \\ x_1 & x_2 & x_3 & x_4 \end{pmatrix}. The number of these minors is \binom{4}{2} = 6, exceeding the codimension and thus violating the condition for being a complete intersection, as the generators do not form a of 3. Another illustration of failure arises from schemes that are not equidimensional, such as the union of a and an in \mathbb{P}^3. Suppose C is an irreducible curve of 1 and p is a point not on C; their union V = C \cup \{p\} has components of 1 and 0. A complete intersection in projective space must be equidimensional, with all components having the expected equal to the ambient minus the length of the defining . Here, the of V cannot be generated by a of length 2 (to achieve 2 for the curve part), as the embedded point disrupts the purity of , making the quotient ring non-Cohen--Macaulay in a way incompatible with a complete intersection structure. Rational scroll surfaces offer further examples where the complete intersection condition fails, particularly due to the structure of their ideals and, in singular cases, non-Cohen--Macaulay local rings at certain points. Consider the rational normal S(1,2) \subset \mathbb{P}^4, a smooth of dimension 2 and 2, obtained as the union of lines joining corresponding points on rational normal curves of degrees 1 and 2 in disjoint projective spaces. Its homogeneous ideal is generated by the three $2 \times 2 minors of a $2 \times 3 matrix, exceeding the codimension and preventing generation by a regular sequence of length 2. For singular scrolls like S(0,3) \subset \mathbb{P}^3, a over the with vertex at the origin, the local ring at the singular vertex fails to be Cohen--Macaulay, as the depth does not equal the dimension there, further ruling out a complete intersection structure despite the global matching the number of potential generators.

Algebraic invariants

Multidegree

In , a complete intersection X in a smooth variety Y is often defined as the common zero locus of global sections s_1, \dots, s_k of line bundles L_1, \dots, L_k on Y, where the sections form a and k = \codim_Y X. The cycle class [X] \in A_k(Y) is given by the of the ambient fundamental class [Y] with the top Chern class of the virtual , which for this setup equals the product \prod_{i=1}^k c_1(L_i) \cap [Y]. The multidegree of X refers to the of coefficients of this class when expressed in a basis for the Chow group A_k(Y). In the specific case of hypersurfaces of degrees d_1, \dots, d_k in \mathbb{P}^n, the line bundles are powers of the bundle \mathcal{O}(1), so c_1(L_i) = d_i h where h is the class of a . The Chow ring A^*(\mathbb{P}^n) is \mathbb{Z}/(h^{n+1}), and the class simplifies to [X] = \left( \prod_{i=1}^k d_i \right) h^k \cap [\mathbb{P}^n]. Thus, the multidegree is the scalar coefficient \prod_{i=1}^k d_i, which equals the degree of X as a subvariety of \mathbb{P}^n. For example, consider a C in \mathbb{P}^3 arising as the complete intersection of two surfaces of degrees a and b. The class is [C] = ab \, h^2 \cap [\mathbb{P}^3], so the multidegree (degree of C) is ab. This notion generalizes from classical , where the intersection number of k = n hypersurfaces in in \mathbb{P}^n is \prod_{i=1}^n d_i; for a positive-dimensional complete intersection, the degree \prod_{i=1}^k d_i plays an analogous role as the intersection multiplicity with a general linear subspace complementary to the dimension of X.

General position

In , a collection of hypersurfaces is said to be in if their scheme-theoretic has the expected and is transverse, meaning that at every point of the , the differentials of the defining equations are linearly independent, resulting in a reduced scheme with no multiple components. This condition ensures that the is a complete when the ambient is . Bertini's theorem plays a central role in establishing this genericity: for a over an of characteristic zero, a section is and thus transverse to the . Iterating this result, the of multiple hypersurfaces of prescribed degrees in yields a complete intersection of the expected , as the defining sections form a and the remains reduced away from any base locus. Specifically, in \mathbb{P}^n, the of r hypersurfaces of degrees d_1, \dots, d_r (with r \leq n) is an irreducible complete intersection of r. When hypersurfaces are not in , their may still be a complete intersection if the defining equations generate an of the correct and form a , but the can have positive-dimensional components with higher multiplicity due to tangencies or singularities. In such cases, the multiplicity at points exceeds one, reflecting non-transversality, though the overall is preserved. This contrasts with the reduced structure guaranteed in the general position scenario.

Topological features

Homology

For a smooth complete intersection X \subset \mathbb{P}^N of dimension n, the sheaf groups H^i(X, \mathcal{O}_X) vanish for $0 < i < n. This intermediate vanishing follows from the Koszul resolution of \mathcal{O}_X as an \mathcal{O}_{\mathbb{P}^N}-module, which is a finite complex of locally free sheaves whose is known from Bott's theorem on projective space. An adapted form of Kodaira's vanishing theorem applies to smooth complete intersections in projective space. If the anticanonical bundle -\omega_X is ample (which holds when the multidegree satisfies \sum d_j \le N), then Kodaira's theorem implies H^i(X, \mathcal{O}_X \otimes L) = 0 for i > 0 and ample line bundles L on X, providing vanishing for twists of the structure sheaf. The topological Betti numbers of smooth complete intersections can be computed using the Koszul complex associated to the defining equations. For a (codimension 1 complete intersection), the states that the in H^i(\mathbb{P}^n, \mathbb{Z}) \to H^i(X, \mathbb{Z}) is an for i < n-1 and surjective for i = n-1, where \dim X = n-1. Thus, the Betti numbers agree with those of \mathbb{P}^n outside the middle dimension, and the dimension of the primitive cohomology H^{n-1}(X)_{\mathrm{prim}} is given by (-1)^{n-1} (\chi(X) - n), with the \chi(X) computed via the . For general , iterative applications of the weak Lefschetz theorem yield isomorphisms in up to degrees i < n, where n is the dimension of X, and the provides a converging to the , allowing explicit computation of primitive Betti numbers in terms of the degrees d_1, \dots, d_c. Regarding , the complex points of a smooth complete intersection X of dimension n have H_{2n}(X(\mathbb{C}), \mathbb{Z}) \cong \mathbb{Z}, generated by the fundamental class. For the real points X(\mathbb{R}), even-dimensional complete intersections often exhibit non-trivial top-dimensional . For example, a smooth complete intersection of two quadrics in \mathbb{CP}^{2m+2} (dimension $2m) has H_{2m}(X(\mathbb{R}), \mathbb{Z}) \cong \mathbb{Z}^k for some k \geq 1, depending on the real topology, such as the number of connected components of the real locus. In singular cases with isolated singularities, the top adjusts by the Milnor numbers at singular points, yielding b_n(V) = b_n(V_{\text{smooth}}) - \sum \mu(V, a_i), where \mu is the Milnor number.

Euler characteristic

The topological Euler characteristic \chi(V) of a complete intersection variety V is defined as the alternating sum \sum_{i} (-1)^i b_i(V), where b_i(V) are the Betti numbers of V. For smooth projective complete intersections over \mathbb{C}, this invariant is determined solely by the dimension of the ambient and the degrees of the defining hypersurfaces, reflecting the rigidity of their topology via the . This theorem implies that the of V agrees with that of the ambient space \mathbb{P}^n outside the middle dimension, with the difference arising from the primitive cohomology in the middle degree, allowing explicit computation of \chi(V). For a smooth V \subset \mathbb{P}^{n+1} of degree d \geq 1 (so \dim V = n), the admits the closed-form expression \chi(V) = n + 2 - \frac{1}{d} \left[ 1 + (-1)^{n+1} (d-1)^{n+2} \right]. This formula arises from integrating the top c_n(TV) over V, using the for the TV = TP^{n+1}|_V \otimes N_{V/\mathbb{P}^{n+1}}^\vee and the known Chern classes of \mathbb{P}^{n+1} and the normal bundle N_{V/\mathbb{P}^{n+1}} = \mathcal{O}_V(d). For instance, when n=1 (plane curves), it recovers \chi(V) = d(3-d), consistent with the arithmetic genus formula. When n=2 (surfaces in \mathbb{P}^3), a degree-3 hypersurface yields \chi(V) = 9, reflecting its rank and 27 lines. For general smooth complete intersections defined by hypersurfaces of degrees d_1, \dots, d_r in \mathbb{P}^{n+r} (with \dim V = n), no simple exists for arbitrary r, but \chi(V) can be computed recursively by viewing V as a in the complete intersection of the first r-1 equations, applying the formula iteratively along with the Künneth formula for the Betti numbers. Explicit algorithms for this recursion, based on Segre classes and projective degrees, are available and implementable for computational verification, even extending to mildly singular cases via the Chern-Schwartz-MacPherson class. These computations highlight how \chi(V) grows polynomially with the degrees, providing key data for and mirror symmetry applications.

References

  1. [1]
  2. [2]
  3. [3]
  4. [4]
    [PDF] Cohen-Macaulay rings and schemes - Columbia Math Department
    This was essentially proven when we showed that Cohen-Macaulay rings have no embedded primes. This theorem can be used to show that a set of polynomials ...
  5. [5]
    [PDF] MINIMAL FREE RESOLUTIONS over COMPLETE INTERSECTIONS ...
    Eisenbud: Homological algebra on a complete intersection, with an application to group representations, Trans. Amer. Math. Soc. 260 (1980), 35–64. 24. D.
  6. [6]
    [PDF] Cohen-Macaulay schemes and Serre duality - Kiran S. Kedlaya
    Any regular local ring is Cohen-Macaulay, since we can use generators of the cotangent space as a regular sequence. But the Cohen-Macaulay condition is much ...
  7. [7]
    [PDF] Complete Intersections in Regular Local Rings - IITB Math
    Let f,g be a regular sequence and put I = (f,g). Then In has no embedded associated primes. Proof. Suppose m is an associated prime of In = J then there is an h ...
  8. [8]
    [PDF] Algebraic Geometry I (Math 6130) Utah/Fall 2020 5. More Projective ...
    If X is a variety and Z = X(f1, ..., fc) ⊂ X is irreducible of codimension c, then Z is a (set-theoretic) complete intersection in X. Example. The planes Z1,Z2 ...
  9. [9]
    [PDF] 3264 & All That Intersection Theory in Algebraic Geometry
    ... commutative algebra in work of van der Waerden, Zariski, Samuel, Weil and ... (Eisenbud and Harris [2000]). (c) An acquaintance with coherent sheaves ...
  10. [10]
    [PDF] FOUNDATIONS OF ALGEBRAIC GEOMETRY CLASS 48
    Apr 24, 2008 · Show that G ∩ C is a nonsingular curve. The complete intersection E of two quadric surfaces in P3 has genus 1. Choose any point of it, so E have ...
  11. [11]
    Biquadratic addition laws on elliptic curves in $\mathbb{P}^3 ... - arXiv
    Apr 24, 2019 · ... elliptic curves, embedded in \mathbb{P}^3 as complete intersection of two quadrics. Finally, we use this interesting relationship with the ...
  12. [12]
    Chapter 11 - Segre Varieties - EMS Press
    The variety which arises is a special case of a determinantal variety. ... If q D 0, i.e., ˛ D ˇ, L is the complete intersection of the quadric with another.
  13. [13]
    Complete Intersections on General Hypersurfaces - Project Euclid
    If. F = F1G1 +···+ FrGr and the forms Fi do not form a regular sequence, then. (F1, ..., Fr) is not the ideal of a complete intersection. To produce an ...<|control11|><|separator|>
  14. [14]
    [1608.04338] On generalizations of Fermat curves over finite fields ...
    Aug 15, 2016 · In this paper, we provide a complete classification of such curves, as well as a characterization of their full automorphism groups. Subjects: ...Missing: intersections | Show results with:intersections
  15. [15]
    Normal bundles of rational curves on complete intersections - arXiv
    May 23, 2017 · If at least one d_i is greater than 2, we show that X contains rational curves of degree e \leq n with balanced normal bundle.Missing: 4 | Show results with:4
  16. [16]
    [PDF] Cohomologies of Derived Intersections - -ORCA - Cardiff University
    instance, any subscheme which is not equidimensional is trivially not a local complete intersection. One needs at least as many equations at the codimension ...
  17. [17]
    [PDF] CHAPTER 2. Rational scrolls Summary 2.1. Reminder: P - mimuw
    This chapter describes scrolls, especially the rational normal scrolls. These va- rieties occur throughout projective and algebraic geometry, and the student ...
  18. [18]
    [PDF] Intersection Theory
    William Fulton. Intersection. Theory. Second Edition. Springer. Page 2. Contents. Introduction. Chapter 1. Rational Equivalence. 6. Summary. 6. 1.1 Notation and ...
  19. [19]
    [PDF] BERTINI1S THEOREM GENERIC SMOOTHNESS
    Bertini's theorem as- serts (in its weak form) that the generic hyperplane section of a smooth projective variety preserves the original smoothness; moreover, ...Missing: position | Show results with:position
  20. [20]
    [PDF] COMPLEX ALGEBRAIC SURFACES CLASS 5
    Oct 16, 2024 · How do you know that there are smooth complete intersections? Answer: Bertini's theorem (fact). A general hyperplane section of a smooth ...
  21. [21]
    [PDF] Intersection Theory
    ... Eisenbud-Harris):. Page 33. 28. I Chow groups. This leads to the following ... a complete intersection of hypersurfaces Hi = V+(fi) with fi ∈ H0(Pn,OP n ...
  22. [22]
    [PDF] notes on the kodaira vanishing theorem - UChicago Math
    Feb 11, 2012 · The Kodaira vanishing and embedding theorems help clarify the equation of “positivity” with ampleness. One consequence of them, for instance, is ...
  23. [23]
    [PDF] On the homology and cohomology of complete intersections with ...
    In the forth section we compute the homology of two classes of varieties: the cubic surfaces in p3 and some 2 m-dimensional complete intersections of two ...