Jump to content

Draft:Complete algebraic curve

From Wikipedia, the free encyclopedia

In algebraic geometry, a complete algebraic curve is an algebraic curve that is complete as an algebraic variety.

A projective curve, a dimension-one projective variety, is a complete curve. A complete curve (over an algebraically closed field) is projective.[1] Because of this, over an algebraically closed field, the terms "projective curve" and "complete curve" are usually used interchangeably. Over a more general base scheme, the distinction still matters.

A curve in is called an (algebraic) space curve, while a curve in is called a plane curve. By means of a projection from a point, any smooth complete or projective curve can be embedded into ;[2] thus, up to a projection, every (smooth) curve is a space curve. Up to a birational morphism, every curve can be embedded into as a nodal curve.[3]

Riemann's existence theorem says that the category of compact Riemann surfaces is equivalent to that of smooth projective curves over the complex numbers.

Throughout the article, a curve mean a complete curve (but not necessarily smooth).

Abstract complete curve[edit]

Let k be an algebrically closed field. By a function field K over k, we mean a finitely generated field extension of k that is typically not algebraic (i.e., a transcendental extension). The function field of an algebraic variety is a basic example. For a function field of transcendence degree one, the converse holds by the following construction.[4] Let denote the set of all discrete valuation rings of . We put the topology on so that the closed subsets are either finite subsets or the whole space. We then make it a locally ringed space by taking to be the intersection . Then the for various function fields K of transcendence degree one form a category that is equivalent to the category of smooth projective curves.[5]

One consequence of the above construction is that a complete smooth curve is projective (since a complete smooth curve of C corresponds to , which corresponds to a projective smooth curve.)

Smooth completion of an affine curve[edit]

Let be a smooth affine curve given by a polynomial f in two variables. The closure in , the projective completion of it, may or may not be smooth. The normalization C of is smooth and contains as an open dense subset. Then the curve is called the smooth completion of .[6] (Note the smooth completion of is unique up to isomorphism since two smooth curves are isomorphic if they are birational to each other.)

For example, if , then is given by , which is smooth (by a Jacobian computation). On the other hand, consider . Then, by a Jacobian computation, is not smooth. In fact, is an (affine) hyperelliptic curve and a hyperelliptic curve is not a plane curve (since a hyperelliptic curve is never a complete intersection in a projective space).

Over the complex numbers, C is a compact Riemann surface that is classically called the Riemann surface associated to the algebraic function when .[6]. Conversely, each compact Riemann surface is of that form;[citation needed] this is known as the Riemann existence theorem.

A map from a curve to a projective space[edit]

To give a rational map from a (projective) curve C to a projective space is to give a linear system of divisors V on C, up to the fixed part of the system? (need to be clarified); namely, when B is the base locus (the common zero sets of the nonzero sections in V), there is:

that maps each point in to the hyperplane . Conversely, given a rational map f from C to a projective space,

In particular, one can take the linear system to be the canonical linear system and the corresponding map is called the canonical map.

Let be the genus of a smooth curve C. If , then is empty while if , then . If , then the canonical linear system can be shown to have no base point and thus determines the morphism . If the degree of f or equivalently the degree of the linear system is 2, then C is called a hyperelliptic curve.

Max Noether's theorem implies that a non-hyperelliptic curve is projectively normal when it is embedded into a projective space by the canonical divisor.

Classification of smooth algebraic curves in [edit]

The classification of a smooth projective curve begins with specifying a genus. For genus zero, there is only one: the projective line (up to isomorphism). A genus-one curve is precisely an elliptic curve and isomorphism classes of elliptic curves are specified by a j-invariant (which is an element of the base field). The classification of genus-2 curves is much more complicated; here is some partial result over an algebraically closed field of characteristic not two:[7]

  • Each genus-two curve X comes with the map determined by the canonical divisor; called the canonical map. The canonical map has exactly 6 ramified points of index 2.
  • Conversely, given 6 points,

For genus , the following terminology is used;

  • Given a smooth curve C, a divisor D on it and a vector subspace , one says the linear system is a grd if V has dimension r+1 and D has degree d. One says C has a grd if there is such a linear system.

Specific curves[edit]

Canonical curve[edit]

Stable curve[edit]

A stable curve is a connected nodal curve with finite automorphism group.

Spectral curve[edit]

Vector bundles on a curve[edit]

Line bundles and dual graph[edit]

Let X be a possibly singular curve. Then

where r is the number of irreducible components of X, is the normalization and . (To get this use the fact and )

Taking the long exact sequence of the exponential sheaf sequence gives the degree map:

By definition, the Jacobian variety J(X) of X is the identity component of the kernel of this map. Then the previous exact sequence gives:

We next define the dual graph of X; a one-dimensional CW complex defined as follows. (related to whether a curve is of compact type or not)

The Jacobian of a curve[edit]

Let C be a smooth connected curve. Given an integer d, let denote the set of isomorphism classes of line bundles on C of degree d. It can be shown to have a structure of an algebraic variety.

For each integer d > 0, let denote respectively the d-th fold Cartesian and symmetric product of C; by definition, is the quotient of by the symmetric group permuting the factors.

Fix a base point of C. Then there is the map

given by .

Stable bundles on a curve[edit]

The Jacobian of a curve can be generalized to higher-rank vector bundles; a key notion introduced by Mumford that allows for a moduli construction is that of stability.

Let C be a connected smooth curve. A rank-2 vector bundle E on C is said to be stable if for every line subbundle L of E,

.

Given some line bundle L on C, let denote the set of isomorphism classes of rank-2 stable bundles E on C whose determinants are isomorphic to L.

Generalization: [edit]

The osculating behavior of a curve[edit]

Vanishing sequence[edit]

Given a linear series V on a curve X, the image of it under is a finite set and following the tradition we write it as

This sequence is called the vanishing sequence. For example, is the multiplicity of a base point p. We think of higher as encoding information about inflection of the Kodaira map . The ramification sequence is then

Their sum is called the ramification index of p. The global ramification is given by the following formula:

Plücker formula — 

Bundle of principal parts[edit]

Uniformization[edit]

An elliptic curve Xover the complex numbers has a uniformization given by taking the quotient by a lattice.

Relative curve[edit]

A relative curve or a curve over a scheme S or a relative curve is a flat morphism of schemes such that each geometric fiber is an algebraic curve; in other words, it is a family of curves parametrized by the base scheme S.

See also Semistable reduction theorem.

The Mumford–Tate uniformization[edit]

This generalizes the classical construction due to Tate (cf. Tate curve)[8] In Mumford 1972, Mumford showed: given a smooth projective curve of genus at least two and has a split degeneration,

See also[edit]

Notes[edit]

  1. ^ Hartshorne, Ch. III., Exercise 5.8.
  2. ^ Hartshorne, Ch. IV., Corollay 3.6.
  3. ^ Hartshorne, Ch. IV., Theorem 3.10.
  4. ^ Hartshorne 1977, Ch. I, § 6.
  5. ^ Hartshorne 1977, Ch. I, § 6. Corollary 6.12.
  6. ^ a b ACGH 1985, Ch I, Exercise A.
  7. ^ Hartshorne, Ch. IV., Exercise 2.2.
  8. ^ Gerritzen, L.; Van Der Put, M. (14 November 2006). Schottky Groups and Mumford Curves. Springer. ISBN 9783540383048.

References[edit]

  • E. Arbarello, M. Cornalba, P.A. Griffiths, and J. Harris, Geometry of algebraic curves. Vol. I, Grundlehren der Mathematischen Wissenschaften, vol. 267, Springer-Verlag, New York, 1985. MR0770932
  • E. Arbarello, M. Cornalba, and P.A. Griffiths, Geometry of algebraic curves. Vol. II, with a contribution by Joseph Daniel Harris, Grundlehren der Mathematischen Wissenschaften, vol. 268, Springer, Heidelberg, 2011. MR-2807457
  • Hartshorne, Robin (1977), Algebraic Geometry, Graduate Texts in Mathematics, vol. 52, New York: Springer-Verlag, ISBN 978-0-387-90244-9, MR 0463157
  • Mukai, S. (2002). An introduction to invariants and moduli. Cambridge Studies in Advanced Mathematics. Vol. 81. ISBN 978-0-521-80906-1.
  • Mumford, D.: An analytic construction of degenerating curves over complete local rings. Compos. Math. 24, 129–174 (1972)
  • McMcallum, W.; Poonen, B. (2017). "THE METHOD OF CHABAUTY AND COLEMAN". S2CID 41964616. {{cite journal}}: Cite journal requires |journal= (help)
  • Voight, John; Zureick-Brown, David (16 March 2022). "The canonical ring of a stacky curve". arXiv:1501.04657 [math.AG].

Further reading[edit]