adjunction formula

{{Short description|Concept in algebraic geometry}}

In mathematics, especially in algebraic geometry and the theory of complex manifolds, the adjunction formula relates the canonical bundle of a variety and a hypersurface inside that variety. It is often used to deduce facts about varieties embedded in well-behaved spaces such as projective space or to prove theorems by induction.

Adjunction for smooth varieties

=Formula for a smooth subvariety=

Let X be a smooth algebraic variety or smooth complex manifold and Y be a smooth subvariety of X. Denote the inclusion map {{nowrap|YX}} by i and the ideal sheaf of Y in X by \mathcal{I}. The conormal exact sequence for i is

:0 \to \mathcal{I}/\mathcal{I}^2 \to i^*\Omega_X \to \Omega_Y \to 0,

where Ω denotes a cotangent bundle. The determinant of this exact sequence is a natural isomorphism

:\omega_Y = i^*\omega_X \otimes \operatorname{det}(\mathcal{I}/\mathcal{I}^2)^\vee,

where \vee denotes the dual of a line bundle.

=The particular case of a smooth divisor=

Suppose that D is a smooth divisor on X. Its normal bundle extends to a line bundle \mathcal{O}(D) on X, and the ideal sheaf of D corresponds to its dual \mathcal{O}(-D). The conormal bundle \mathcal{I}/\mathcal{I}^2 is i^*\mathcal{O}(-D), which, combined with the formula above, gives

:\omega_D = i^*(\omega_X \otimes \mathcal{O}(D)).

In terms of canonical classes, this says that

:K_D = (K_X + D)|_D.

Both of these two formulas are called the adjunction formula.

Examples

= Degree d hypersurfaces =

Given a smooth degree d hypersurface i: X \hookrightarrow \mathbb{P}^n_S we can compute its canonical and anti-canonical bundles using the adjunction formula. This reads as

\omega_X \cong i^*\omega_{\mathbb{P}^n}\otimes \mathcal{O}_X(d)
which is isomorphic to \mathcal{O}_X(-n{-}1{+}d).

= Complete intersections =

For a smooth complete intersection i: X \hookrightarrow \mathbb{P}^n_S of degrees (d_1, d_2), the conormal bundle \mathcal{I}/\mathcal{I}^2 is isomorphic to \mathcal{O}(-d_1)\oplus \mathcal{O}(-d_2), so the determinant bundle is \mathcal{O}(-d_1{-}d_2) and its dual is \mathcal{O}(d_1{+}d_2), showing

\omega_X \,\cong\, \mathcal{O}_X(-n{-}1)\otimes \mathcal{O}_X(d_1{+}d_2)

\,\cong\, \mathcal{O}_X(-n{-}1 {+} d_1 {+} d_2).

This generalizes in the same fashion for all complete intersections.

= Curves in a quadric surface =

\mathbb{P}^1\times\mathbb{P}^1 embeds into \mathbb{P}^3 as a quadric surface given by the vanishing locus of a quadratic polynomial coming from a non-singular symmetric matrix.{{cite web |last1=Zhang |first1=Ziyu |title=10. Algebraic Surfaces |url=https://ziyuzhang.github.io/ma40188/Lecture19.pdf |archiveurl=https://web.archive.org/web/20200211004951/https://ziyuzhang.github.io/ma40188/Lecture19.pdf|archive-date=2020-02-11 }} We can then restrict our attention to curves on Y= \mathbb{P}^1\times\mathbb{P}^1. We can compute the cotangent bundle of Y using the direct sum of the cotangent bundles on each \mathbb{P}^1, so it is \mathcal{O}(-2,0)\oplus\mathcal{O}(0,-2). Then, the canonical sheaf is given by \mathcal{O}(-2,-2), which can be found using the decomposition of wedges of direct sums of vector bundles. Then, using the adjunction formula, a curve defined by the vanishing locus of a section f \in \Gamma(\mathcal{O}(a,b)), can be computed as

:

\omega_C \,\cong\, \mathcal{O}(-2,-2)\otimes \mathcal{O}_C(a,b) \,\cong\, \mathcal{O}_C(a{-}2, b{-}2).

Poincaré residue

{{see also|Poincaré residue}}

The restriction map \omega_X \otimes \mathcal{O}(D) \to \omega_D is called the Poincaré residue. Suppose that X is a complex manifold. Then on sections, the Poincaré residue can be expressed as follows. Fix an open set U on which D is given by the vanishing of a function f. Any section over U of \mathcal{O}(D) can be written as s/f, where s is a holomorphic function on U. Let η be a section over U of ωX. The Poincaré residue is the map

:\eta \otimes \frac{s}{f} \mapsto s\frac{\partial\eta}{\partial f}\bigg|_{f = 0},

that is, it is formed by applying the vector field ∂/∂f to the volume form η, then multiplying by the holomorphic function s. If U admits local coordinates z1, ..., zn such that for some i, {{nowrap begin}}∂f/∂zi ≠ 0{{nowrap end}}, then this can also be expressed as

:\frac{g(z)\,dz_1 \wedge \dotsb \wedge dz_n}{f(z)} \mapsto (-1)^{i-1}\frac{g(z)\,dz_1 \wedge \dotsb \wedge \widehat{dz_i} \wedge \dotsb \wedge dz_n}{\partial f/\partial z_i}\bigg|_{f = 0}.

Another way of viewing Poincaré residue first reinterprets the adjunction formula as an isomorphism

:\omega_D \otimes i^*\mathcal{O}(-D) = i^*\omega_X.

On an open set U as before, a section of i^*\mathcal{O}(-D) is the product of a holomorphic function s with the form {{nowrap|df/f}}. The Poincaré residue is the map that takes the wedge product of a section of ωD and a section of i^*\mathcal{O}(-D).

Inversion of adjunction

The adjunction formula is false when the conormal exact sequence is not a short exact sequence. However, it is possible to use this failure to relate the singularities of X with the singularities of D. Theorems of this type are called inversion of adjunction. They are an important tool in modern birational geometry.

The Canonical Divisor of a Plane Curve

Let C \subset \mathbf{P}^2 be a smooth plane curve cut out by a degree d homogeneous polynomial F(X, Y, Z). We claim that the canonical divisor is K = (d-3)[C \cap H] where H is the hyperplane divisor.

First work in the affine chart Z \neq 0. The equation becomes f(x, y) = F(x, y, 1) = 0 where x = X/Z and y = Y/Z.

We will explicitly compute the divisor of the differential

:\omega := \frac{dx}{\partial f / \partial y} = \frac{-dy}{\partial f / \partial x}.

At any point (x_0, y_0) either \partial f / \partial y \neq 0 so x - x_0 is a local parameter or

\partial f / \partial x \neq 0 so y - y_0 is a local parameter.

In both cases the order of vanishing of \omega at the point is zero. Thus all contributions to the divisor \text{div}(\omega) are at the line at infinity, Z = 0.

Now look on the line {Z = 0}. Assume that [1, 0, 0] \not\in C so it suffices to look in the chart Y \neq 0 with coordinates u = 1/y and v = x/y. The equation of the curve becomes

:g(u, v) = F(v, 1, u) = F(x/y, 1, 1/y) = y^{-d}F(x, y, 1) = y^{-d}f(x, y).

Hence

: \partial f/\partial x = y^d \frac{\partial g}{\partial v} \frac{\partial v}{\partial x} = y^{d-1}\frac{\partial g}{\partial v}

so

:\omega = \frac{-dy}{\partial f / \partial x} = \frac{1}{u^2} \frac{du}{y^{d-1}\partial g/ \partial v} = u^{d-3} \frac{du}{\partial g / \partial v}

with order of vanishing \nu_p(\omega) = (d-3)\nu_p(u). Hence \text{div}(\omega) = (d-3)[C \cap \{Z = 0\}] which agrees with the adjunction formula.

Applications to curves

The genus-degree formula for plane curves can be deduced from the adjunction formula.Hartshorne, chapter V, example 1.5.1 Let C ⊂ P2 be a smooth plane curve of degree d and genus g. Let H be the class of a hyperplane in P2, that is, the class of a line. The canonical class of P2 is −3H. Consequently, the adjunction formula says that the restriction of {{nowrap|(d − 3)H}} to C equals the canonical class of C. This restriction is the same as the intersection product {{nowrap|(d − 3)HdH}} restricted to C, and so the degree of the canonical class of C is {{nowrap|d(d−3)}}. By the Riemann–Roch theorem, {{nowrap begin}}g − 1 = (d−3)dg + 1{{nowrap end}}, which implies the formula

:g = \tfrac12(d{-} 1)(d {-} 2).

Similarly,Hartshorne, chapter V, example 1.5.2 if C is a smooth curve on the quadric surface P1×P1 with bidegree (d1,d2) (meaning d1,d2 are its intersection degrees with a fiber of each projection to P1), since the canonical class of P1×P1 has bidegree (−2,−2), the adjunction formula shows that the canonical class of C is the intersection product of divisors of bidegrees (d1,d2) and (d1−2,d2−2). The intersection form on P1×P1 is ((d_1,d_2),(e_1,e_2))\mapsto d_1 e_2 + d_2 e_1 by definition of the bidegree and by bilinearity, so applying Riemann–Roch gives 2g-2 = d_1(d_2{-}2) + d_2(d_1{-}2) or

:g = (d_1 {-} 1)(d_2 {-} 1) \,=\, d_1 d_2 - d_1 - d_2 + 1.

The genus of a curve C which is the complete intersection of two surfaces D and E in P3 can also be computed using the adjunction formula. Suppose that d and e are the degrees of D and E, respectively. Applying the adjunction formula to D shows that its canonical divisor is {{math|(d − 4)H|D}}, which is the intersection product of {{nowrap|(d − 4)H}} and D. Doing this again with E, which is possible because C is a complete intersection, shows that the canonical divisor C is the product {{math|(d + e − 4)HdHeH}}, that is, it has degree {{math|de(d + e − 4)}}. By the Riemann–Roch theorem, this implies that the genus of C is

:g = de(d + e - 4) / 2 + 1.

More generally, if C is the complete intersection of {{math|n − 1}} hypersurfaces {{math|D1, ..., Dn − 1}} of degrees {{math|d1, ..., dn − 1}} in Pn, then an inductive computation shows that the canonical class of C is (d_1 + \cdots + d_{n-1} - n - 1)d_1 \cdots d_{n-1} H^{n-1}. The Riemann–Roch theorem implies that the genus of this curve is

:g = 1 + \tfrac{1}{2}(d_1 + \cdots + d_{n-1} - n - 1)d_1 \cdots d_{n-1}.

In low dimensional topology

Let S be a complex surface (in particular a 4-dimensional manifold) and let C\to S be a smooth (non-singular) connected complex curve. ThenGompf, Stipsicz, Theorem 1.4.17

2g(C)-2=[C]^2-c_1(S)[C]

where g(C) is the genus of C, [C]^2 denotes the self-intersections and c_1(S)[C] denotes the Kronecker pairing .

See also

References

{{Reflist}}

  • Intersection theory 2nd edition, William Fulton, Springer, {{ISBN|0-387-98549-2}}, Example 3.2.12.
  • Principles of algebraic geometry, Griffiths and Harris, Wiley classics library, {{ISBN|0-471-05059-8}} pp 146–147.
  • Algebraic geometry, Robin Hartshorne, Springer GTM 52, {{ISBN|0-387-90244-9}}, Proposition II.8.20.

{{Algebraic curves navbox}}

{{DEFAULTSORT:Adjunction Formula (Algebraic Geometry)}}

Category:Algebraic geometry