• Nie Znaleziono Wyników

An explicit version of Faltings’ Product Theorem and an improvement of Roth’s lemma

N/A
N/A
Protected

Academic year: 2021

Share "An explicit version of Faltings’ Product Theorem and an improvement of Roth’s lemma"

Copied!
34
0
0

Pełen tekst

(1)

LXXIII.3 (1995)

An explicit version of Faltings’ Product Theorem and an improvement of Roth’s lemma

by

Jan-Hendrik Evertse (Leiden)

To Professor Wolfgang Schmidt on his 60th birthday

Introduction. In his paper Diophantine approximation on abelian va- rieties [1], Faltings proved, among other things, the following conjecture of Weil and Lang: if A is an abelian variety over a number field k and X a subvariety of A not containing a translate of a positive dimensional abelian subvariety of A, then X contains only finitely many k-rational points. One of Faltings’ basic tools was a new non-vanishing result of his, also proved in [1], the so-called (arithmetic version of the) Product Theorem. It has turned out that this Product Theorem has a much wider range of applicability in Diophantine approximation. For instance, recently Faltings and W¨ ustholz gave an entirely new proof [2] of Schmidt’s Subspace Theorem [15] based on the Product Theorem.

Faltings’ Product Theorem is not only very powerful for deriving new qualitative finiteness results in Diophantine approximation but, in an ex- plicit form, it can be used also to derive significant improvements of existing quantitative results. In the present paper, we work out an explicit version of the arithmetic version of the Product Theorem; except for making explicit some of Faltings’ arguments from [1] this did not involve anything new.

By using the same techniques we improve Roth’s lemma from [12]. Roth’s lemma was used by Roth in his theorem on the approximation of algebraic numbers by rationals [12] and later by Schmidt in his proof of the Subspace Theorem [15].

In two subsequent papers we shall apply our improvement of Roth’s lemma to derive significant improvements on existing explicit upper bounds for the number of subspaces in the Subspace Theorem, due to Schmidt [16]

and Schlickewei [14] and for the number of solutions of norm form equations [17] and S-unit equations [13].

[215]

(2)

At the conference on Diophantine problems in Boulder in honour of W. M. Schmidt (26 June – 1 July, 1994), W¨ ustholz announced that his stu- dent R. Ferretti had independently obtained results similar to our Theorems 1 and 2. These results have been published in [3]. Part of the arguments used in the proof of Theorem 1 had already been worked out by van der Put [11]

in his lecture at the conference “Diophantine approximation and Abelian varieties”, Soesterberg, The Netherlands, 12–15 April 1992.

As the Product Theorem appears to have applications outside arithmetic algebraic geometry, we have tried to make this paper accessible to non- geometers with a modest knowledge of algebraic geometry.

1. Statement of the results. Let n = (n

1

, . . . , n

m

) be a tuple of pos- itive integers. For h = 1, . . . , m, denote by X

h

the block of n

h

+ 1 variables X

h0

, . . . , X

h,nh

. For a ring R, denote by R[X] or R[X

1

, . . . , X

m

] the polyno- mial ring in the (n

1

+ 1) + . . . + (n

m

+ 1) variables X

hj

(h = 1, . . . , m, j = 0, . . . , n

h

). For a tuple of non-negative integers d = (d

1

, . . . , d

m

), denote by Γ

Rn

(d) the R-module of polynomials in R[X] which are homogeneous of degree d

1

in the block X

1

, . . . , homogeneous of degree d

m

in X

m

, i.e. the R-module generated by the monomials

X

i

:=

Y

m h=1

nh

Y

j=0

X

hjihj

with

nh

X

j=0

i

hj

= d

h

for h = 1, . . . , m.

Let Γ

Rn

:= S

d∈(Z≥0)m

Γ

Rn

(d) be the set of polynomials which are homoge- neous in each block X

h

for h = 1, . . . , m. An n-ideal of R[X] is an ideal generated by polynomials from Γ

Rn

. An essential n-prime ideal of R[X] is an n-ideal which is a prime ideal and which does not contain any of the ideals (X

h0

, . . . , X

h,nh

) (h = 1, . . . , m).

Let k be an algebraically closed field and denote by P

n

(k) the n-dimen- sional projective space over k. Every point P ∈ P

n

(k) can be represented by a unique (up to a scalar multiple) non-zero vector x = (x

0

, . . . , x

n

) ∈ k

n+1

of homogeneous coordinates. Let again n = (n

1

, . . . , n

m

) be a tuple of positive integers. Define the multi-projective space P

n

(k) as the cartesian product

P

n

(k) := P

n1

(k) × . . . × P

nm

(k).

In what follows, P

n

(k) with a non-bold face superscript denotes the n- dimensional (single-) projective space, and P

n

(k) with a bold-face super- script a multi-projective space. For f ∈ Γ

kn

and for P = (P

1

, . . . , P

m

) ∈ P

n

(k) with P

h

∈ P

nh

(k) for h = 1, . . . , m we say that f (P ) = 0 (or 6= 0) if f (x

1

, . . . , x

m

) = 0 (or 6= 0) for any vectors of homogeneous coordinates x

1

, . . . , x

m

, representing P

1

, . . . , P

m

respectively. This is well defined. A (Zariski-) closed subset of P

n

(k) is a set

{P ∈ P

n

(k) : f

1

(P ) = 0, . . . , f

r

(P ) = 0}

(3)

(abbreviated {f

1

= 0, . . . , f

r

= 0}), where f

1

, . . . , f

r

∈ Γ

kn

\{0}. A closed subset X of P

n

(k) is called reducible if it is the union of two closed subsets A, B of P

n

(k) with A ( X, B ( X, and irreducible otherwise. Every closed subset X of P

n

(k) can be expressed uniquely as

X = Z

1

∪ . . . ∪ Z

r

,

where Z

1

, . . . , Z

r

are irreducible closed subsets of P

n

(k) such that Z

i

* Z

j

for i, j ∈ {1, . . . , r}, i 6= j (cf. [18], p. 23). Z

1

, . . . , Z

r

are called the irreducible components of X. We use the term “subvariety” exclusively for a projective subvariety, i.e. a closed irreducible subset.

There is a one-to-one correspondence between subvarieties of P

n

(k) and essential n-prime ideals I of k[X]:

I ↔ V (I) = {P ∈ P

n

(k) : f (P ) = 0 for all f ∈ I}.

We say that the subvariety V of P

n

(k) is defined over a subfield k

1

of k if its corresponding prime ideal can be generated by polynomials with coefficients from k

1

. An important class of subvarieties of P

n

(k) we will encounter are the product varieties

Z

1

× . . . × Z

m

= {(P

1

, . . . , P

m

) : P

h

∈ Z

h

for h = 1, . . . , m},

where Z

h

is a subvariety of P

nh

(k) for h = 1, . . . , m. It is a theorem (cf. [18], pp. 61–62) that the cartesian product of subvarieties of P

n1

(k), . . . , P

nm

(k), respectively, is a subvariety of P

n

(k).

Let F ∈ Γ

kn

. For a tuple of non-negative integers i = (i

hj

: h = 1, . . . , m, j = 0, . . . , n

h

) define the partial derivative of F :

F

i

:=

 Y

m

h=0 nh

Y

j=0

ihj

∂X

hjihj

 F.

Let d = (d

1

, . . . , d

m

) be a tuple of positive integers. For a tuple i as above, put

(i/d) :=

X

m h=1

1 d

h

(i

h0

+ . . . + i

h,nh

).

The index of F with respect to P ∈ P

n

(k) and d, notation i

d

(F, P ), is the largest number σ such that

F

i

(P ) = 0 for all i with (i/d) ≤ σ.

The index of F at P is some kind of weighted multiplicity of F at P . The

index is independent of the choice of homogeneous coordinates on P

nh

for

h = 1, . . . , m. Namely, if for h = 1, . . . , m, Y

h0

, . . . , Y

h,nh

are linearly inde-

pendent linear forms in X

h

, then the differential operators ∂/∂Y

hj

are linear

combinations of the ∂/∂X

hj

and vice versa, hence the index does not change

(4)

when in its definition the operators ∂/∂X

hj

(j = 0, . . . , n

h

) are replaced by

∂/∂Y

hj

(j = 0, . . . , n

h

) for h = 1, . . . , m.

For σ ≥ 0, define the closed subset of P

n

(k),

Z

σ

= Z

σ

(F, d) := {P ∈ P

n

(k) : i

d

(F, P ) ≥ σ}

= {P ∈ P

n

(k) : F

i

(P ) = 0 for all i with (i/d) ≤ σ}.

Z

σ

need not be irreducible. The Product Theorem of Faltings [1], Thm. 3.1, states that if Z is an irreducible component of Z

σ

and also of Z

σ+ε

for some ε > 0, and if the quotients d

1

/d

2

, . . . , d

m−1

/d

m

are sufficiently large in terms of ε and m, then Z is a product variety. Below we have stated this result in an explicit form. The degree deg Z of a subvariety Z of P

n

is the number of points in the intersection of Z with a generic linear projective subspace L of P

n

such that dim Z + dim L = n. The codimension of Z is n − dim Z.

Theorem 1. Let k be an algebraically closed field of characteristic 0.

Further , let m be an integer ≥ 2, n = (n

1

, . . . , n

m

), d = (d

1

, . . . , d

m

) tuples of positive integers and σ, ε reals such that σ ≥ 0, 0 < ε ≤ 1 and

(1.1) d

h

d

h+1

 mM ε



M

for h = 1, . . . , m − 1,

where M := n

1

+ . . . + n

m

. Finally, let F ∈ Γ

kn

(d)\{0}, and let Z be an irreducible component of both Z

σ

(F, d) and Z

σ+ε

(F, d). Then Z is a product variety

(1.2) Z = Z

1

× . . . × Z

m

,

where Z

h

is a subvariety of P

nh

(k) for h = 1, . . . , m. Further , if F has its coefficients in a subfield k

0

of k, then Z

1

, . . . , Z

m

are defined over an extension k

1

of k

0

with

(1.3) [k

1

: k

0

] deg Z

1

. . . deg Z

m

 ms ε



s

, where s = P

m

i=1

codim Z

i

.

The idea behind the proof of Theorem 1 is roughly as follows. Any irre- ducible component Z of both Z

σ

and Z

σ+ε

must have in some sense large multiplicity (analogously, if for a polynomial f in one variable all derivatives of f up to some order vanish at P then P has large multiplicity). On the other hand, using intersection theory one shows that the multiplicity of Z

σ

can be that large only if this component is a product variety.

Now let k = Q be the field of algebraic numbers. We need estimates for

the heights of Z

1

, . . . , Z

m

in terms of the height of F . First we define the

height of x = (x

0

, . . . , x

n

) ∈ Q

n+1

\{0}. Take any number field K containing

x

0

, . . . , x

n

. Denote by O

K

the ring of integers of K and let σ

1

, . . . , σ

f

, f =

[K : Q], be the embeddings of K into C. Choose α ∈ O

K

\{0} such that

(5)

αx

0

, . . . , αx

n

∈ O

K

, let a = αx

0

O

K

+ . . . + αx

n

O

K

be the ideal generated by αx

0

, . . . , αx

n

, and N a = #(O

K

/a) the norm of a. Then the height of x is defined by

(1.4) H(x) :=

 1 N a

Y

f j=1

 X

n

i=0

j

(αx

i

)|

2



1/2



1/f

.

It is easy to show that this does not depend on the choices of α and K. The height of a non-zero polynomial F ∈ Q[X

1

, . . . , X

n

] is defined by H(F ) = H(x), where x is the vector of non-zero coefficients of F .

It is obvious that H(λx) = H(x) for every λ ∈ Q

. Hence we can define a height on P

n

(Q) by H(P ) = H(x), where x ∈ Q

n+1

\{0} is any vector rep- resenting P . By using the arithmetic intersection theory of Gillet and Soul´e [5] for schemes over Spec Z, Faltings defined a height h(Z) for subvarieties Z of P

n

(Q) (cf. [1], pp. 552–553 and [7] for more details). This height is always ≥ 0. Further, for points P ∈ P

n

(Q) we have

(1.5) h(P ) = log H(P ).

Philippon [10] and Soul´e [19] gave an explicit expression for the Faltings height of Z in terms of the Chow form of Z. This is the unique (up to a constant) polynomial F

Z

in the r + 1 blocks of n + 1 variables ζ

0

=

00

, . . . , ζ

0n

), . . . , ζ

r

= (ζ

r0

, . . . , ζ

rn

), where r = dim Z such that F

Z

has degree deg Z in each block ζ

i

(i = 0, . . . , r) and such that F

Z

0

, . . . , ζ

r

) = 0 if and only if Z and the r + 1 linear hyperplanes (ζ

i

, X) = 0 (i = 0, . . . , r) have a point in common (cf. [18], pp. 65–66). From the investigations of Philippon and Soul´e it follows that

(1.6) |h(Z) − log H(F

Z

)| ≤ c(n) deg Z, where c(n) is effectively computable in terms of n.

Below we give an explicit version of [1], Theorem 3.3.

Theorem 2. Let m, n, d, σ, ε, F, Z, Z

1

, . . . , Z

m

, k

0

, k

1

, s = P

m

h=1

codim Z

h

be as in Theorem 1, except that k = Q. Then

(1.7) [k

1

: k

0

] deg Z

1

. . . deg Z

m

 X

m

h=1

1

deg Z

h

d

h

h(Z

h

)



≤ 2(s/ε)

s

m

M

M

2

(d

1

+ . . . + d

m

+ log H(F )).

As mentioned in the introduction, results similar to our Theorems 1 and 2 were obtained independently by Ferretti [3].

The following corollary of Theorems 1 and 2 is useful.

Corollary. Let m be an integer ≥ 2, n = (n

1

, . . . , n

m

), d = (d

1

, . . . , d

m

)

tuples of positive integers and ε a real such that 0 ≤ ε ≤ M + 1 and

(6)

(1.8) d

h

d

h+1

 mM (M + 1) ε



M

for h = 1, . . . , m − 1, where again M := n

1

+ . . . + n

m

. Further , let F ∈ Γ

n

Q

(d)\{0}. Then each irreducible component of Z

ε

is contained in a product variety

Z

1

× . . . × Z

m

P

n

(Q),

where for h = 1, . . . , m, Z

h

is a subvariety of P

nh

(Q). Further , if F has its coefficients in an algebraic number field k

0

, then Z

1

, . . . , Z

m

are defined over an extension k

1

of k

0

with

(1.9) [k

1

: k

0

] deg Z

1

. . . deg Z

m

 m(M + 1)s ε



s

, where s = P

m

h=1

codim Z

h

, and (1.10) [k

1

: k

0

] deg Z

1

. . . deg Z

m

 X

m

h=1

1

deg Z

h

d

h

h(Z

h

)



≤ 2

 (M + 1)s ε



s

m

M

M

2

(d

1

+ . . . + d

m

+ log H(F )).

P r o o f. Put ε

0

:= ε/(M + 1). Consider the sequence of closed subsets of P

n

(Q):

P

n

(Q) = Z

0

⊇ Z

ε0

⊇ Z

0

⊇ . . . ⊇ Z

(M +1)ε0

= Z

ε

.

For i = 0, . . . , M + 1, choose an irreducible component W

i

of Z

0

such that P

n

(Q) = W

0

⊇ W

1

⊇ . . . ⊇ W

M +1

= Z.

By [18], p. 54, P

n

(Q) has dimension n

1

+ . . . + n

m

= M and if V

1

, V

2

are two subvarieties of P

n

(Q) with V

1

V

2

then dim V

1

< dim V

2

. It follows that there is an i ∈ {0, . . . , M } with W

i

= W

i+1

. Clearly, W := W

i

= W

i+1

P

n

(Q) as it is contained in {F = 0}. Further, W is an irreducible component of both Z

0

and Z

00

. By (1.8), the conditions of Theorems 1 and 2 are satisfied with iε

0

, ε

0

replacing σ, ε. Hence W = Z

1

× . . . × Z

m

, where Z

h

is a subvariety of P

nh

(Q), for h = 1, . . . , m. Inequalities (1.9), (1.10) follow by replacing ε by ε/(M + 1) in (1.3), (1.7), respectively.

Using the techniques of the proofs of Theorems 1 and 2 one can prove the

following sharpening of a non-vanishing result of Roth from 1955 [12], now

known as Roth’s lemma. Roth used this in his proof of his famous theorem,

also in [12], that for every algebraic number α and every κ > 2 there are only

finitely many rationals x/y with x, y ∈ Z, y > 0 and |α−x/y| < y

−κ

. In fact,

from the Corollary with n

1

= . . . = n

m

= 1 one can derive Theorem 3 below

with instead of (1.11) the more restrictive condition d

h

/d

h+1

≥ (2m

3

/ε)

m

for h = 1, . . . , m − 1.

(7)

Theorem 3 (Roth’s lemma). Let m be an integer ≥ 2, let d = (d

1

, . . . . . . , d

m

) be a tuple of positive integers, let F ∈ Q[X

10

, X

11

; . . . ; X

m0

, X

m1

] be a non-zero polynomial which is homogeneous of degree d

h

in the pair of variables (X

h0

, X

h1

) for h = 1, . . . , m and let ε be a real with 0 < ε ≤ m + 1 such that

(1.11) d

h

/d

h+1

≥ 2m

3

for h = 1, . . . , m − 1.

Further , let P = (P

1

, . . . , P

m

), where P

1

, . . . , P

m

are points in P

1

(Q) with (1.12) H(P

h

)

dh

> {e

d1+...+dm

H(F )}

(3m3/ε)m

for h = 1, . . . , m, where e = 2.7182 . . . Then i

d

(F, P ) < ε.

The original lemma proved by Roth in 1955 [12] differs from Theorem 3 in that instead of (1.11) it has the more restrictive condition

(1.13) d

h

/d

h+1

≥ (10

m

/ε)

2m

for h = 1, . . . , m − 1.

Roth’s lemma with (1.13) was also used by Schmidt in his proof of the Sub- space Theorem and by Schmidt and Schlickewei in their proofs of quantita- tive versions of the Subspace Theorem. In our improvements of the results of Schmidt and Schlickewei mentioned in the introduction, it was crucial that (1.13) could be replaced by (1.11).

R e m a r k (inspired by a suggestion of the referee). We have formulated the Product Theorem and its consequences for multi-homogeneous poly- nomials. There are affine analogues for polynomials which are not multi- homogeneous. For instance, for h = 1, . . . , m, let Y

h

= (Y

h1

, . . . , Y

h,nh

) be a block of affine variables, and let f ∈ k[Y

1

, . . . , Y

m

] be a polynomial whose total degree in the block Y

h

is at most d

h

, for h = 1, . . . , m. De- note by i and k tuples (i

hj

: h = 1, . . . , m, j = 0, . . . , n

h

) and (k

hj

: h = 1, . . . , m, j = 1, . . . , n

h

), respectively. Define the index of f at a point p as the largest number σ such that f

k

(p) = 0 for all tuples k with P

m

h=1

d

−1h

(k

h1

+ . . . + k

h,nh

) ≤ σ, where f

k

= ( Q

m

h=1

Q

nh

j=1

khj

/∂Y

hjkhj

)f . For h = 1, . . . , m, define a block of variables X

h

= (X

h0

, . . . , X

hm

) such that Y

hj

= X

hj

/X

h0

for j = 1, . . . , n

h

. Let F = Q

m

h=1

X

h0dh

f be the multi- homogeneous polynomial in X

1

, . . . , X

m

corresponding to f . One obtains an analogue of Theorem 1 for f (the same statement with everywhere “affine varieties” replacing “projective varieties”) simply by applying Theorem 1 to F .

We have to check that the index of f at p = (p

11

, . . . , p

1,n1

; . . . ; p

m1

, . . . ,

p

m,nm

), defined using the variables Y

hj

, is equal to the index of F at P =

(1, p

11

, . . . , p

1,n1

; . . . ; 1, . . . , p

m,nm

) defined using the variables X

hj

. This fol-

lows by observing first that f

k

= H

−1

F

i

, where H is a product of pow-

ers of X

h0

(h = 1, . . . , m) and i is the same tuple as k augmented with

i

h0

:= 0 for h = 1, . . . , m, and second, in view of Euler’s identity ∂H/∂X

h0

=

(8)

X

h0−1

(e

h

H − P

nh

j=1

X

hj

∂H/∂X

hj

) for polynomials H homogeneous of degree e

h

in X

h

, that for each tuple i, F

i

is a linear combination of f

k

over tuples k with k

hj

≤ i

hj

for h = 1, . . . , m, j = 1, . . . , n

h

, the coefficients being rational functions whose denominators are products of powers of X

h0

(h = 1, . . . , m).

2. Intersection theory. Most of the results from intersection theory we need can be found in [4], Chaps. 1, 2 and in [9]. As in Section 1, k denotes an algebraically closed field and n = (n

1

, . . . , n

m

) a tuple of positive integers.

The block X

h

of n

h

+ 1 variables, the ring k[X] = k[X

1

, . . . , X

m

] and the sets Γ

kn

(d) will have the meaning of Section 1. We write P

n

, Γ

n

, Γ

n

(d) for P

n

(k), Γ

kn

, Γ

kn

(d).

For every subvariety Z of P

n

there is a unique essential n-prime ideal I of k[X] such that Z = V (I) = {P ∈ P

n

: f (P ) = 0 for every f ∈ I}. The local ring of Z is defined by

(2.1) O

Z

:= {f /g : ∃d ∈ (Z

≥0

)

m

with f, g ∈ Γ

n

(d), g 6∈ I}.

For any n-ideal J of k[X] we put JO

Z

:= {f /g : ∃d ∈ (Z

≥0

)

m

with f, g ∈ Γ

n

(d), f ∈ J, g 6∈ I}. Then M

Z

:= IO

Z

is the maximal ideal of O

Z

. The residue field k(Z) := O

Z

/M

Z

is called the function field of Z. The dimension of Z is dim Z := trdeg

k

k(Z). In particular, dim P

n

= M := n

1

+ . . . + n

m

. The codimension of Z is codim Z := M − dim Z; if W is a subvariety of Z then the codimension of W in Z is codim(W, Z) = dim Z − dim W .

A cycle in P

n

is a finite formal linear combination with integer coefficients of subvarieties V of P

n

, Z = P

n

V

V , say. The components of Z are the subvarieties V for which n

V

6= 0, and n

V

is called the multiplicity of V in Z. Z is called effective if all n

V

≥ 0. Denote by Z

k

= Z

k

(P

n

) the abelian group of cycles in P

n

all of whose components have dimension k and put Z

k

:= (0) for k < 0. We denote by Z cycles as well as varieties.

For a ring A and an A-module M , we define the length l

A

(M ) to be the integer l for which there exists a sequence of A-modules

M = M

0

! M

1

! . . . ! M

l

= (0)

such that M

i−1

/M

i

= A/p

i

for i = 1, . . . , l, where p

i

is a maximal ideal of A (cf. [4], p. 406); l

A

(M ) is independent of the choice of M

0

, . . . , M

l

. Now let Z = V (I) be a subvariety of P

n

and f ∈ Γ

n

\{0} such that f does not vanish identically on Z, i.e. f 6∈ I. We define the divisor of f restricted to Z by attaching certain multiplicities to the irreducible components of Z ∩ {f = 0}. These irreducible components are all of codimension 1 in V (cf. [21], p. 196). For each subvariety W of Z with codim(W, Z) = 1, the number

(2.2) ord

W

(f |Z) := l

OW

(O

W

/(I + (f ))O

W

)

(9)

is a finite, non-negative integer and ord

W

(f |Z) > 0 if and only if I + (f ) is contained in the prime ideal of W , i.e. if W is an irreducible component of Z ∩ {f = 0}. Now define

(2.3) div(f |Z) = X

W

ord

W

(f |Z) · W,

where the sum is taken over all subvarieties W of codimension 1 in Z. By [3], App. A3, ord

W

(f g|Z) = ord

W

(f |Z) + ord

W

(g|Z) and hence div(f g|Z) = div(f |Z) + div(g|Z) whenever f, g do not identically vanish on Z. By abuse of terminology, we say that f does not identically vanish on a cycle Z = P n

V

V if for each component V of Z, f does not identically vanish on V . In that case we define div(f |Z) = P

n

V

div(f |V ). Note that div(f |Z) is effective if Z is effective. We write div(f ) if Z = P

n

.

Two cycles Z

1

, Z

2

∈ Z

t

(P

n

) are called rationally equivalent if Z

1

− Z

2

is a linear combination of cycles div(f |V ) − div(g|V ), where V is a (t + 1)-dimensional subvariety of P

n

and f, g ∈ Γ

n

(d) for some d ∈ (Z

≥0

)

n

. Addition of cycles induces addition of rational equivalence classes. Note that all divisors div(f ) with f ∈ Γ

n

(d) (d ∈ (Z

≥0

)

m

) are rationally equivalent;

denote by O(d) the rational equivalence class of div(f ), f ∈ Γ

n

(d). Clearly, O(d

1

) + O(d

2

) = O(d

1

+ d

2

). We define O(d) for d ∈ Z

m

by additivity.

Put Pic(n) = {O(d) : d ∈ Z

m

}, Pic

+

(n) = {O(d) : d ∈ (Z

≥0

)

m

}. If M = O(d) ∈ Pic

+

(n), then write Γ (M) or Γ

k

(M) for Γ

kn

(d).

For a zero-dimensional cycle Z = P

P

n

P

P we define its degree:

deg Z := X

P

n

P

. Then we have:

Lemma 1. For t = 0, . . . , M there is a unique function (intersection number ) from Z

t

(P

n

) × Pic(n)

t

to Z : (Z, M

1

, . . . , M

t

) 7→ (Z · M

1

. . . M

t

) with the following properties:

(i) (Z · M

1

. . . M

t

) is additive in Z, M

1

, . . . , M

t

and invariant under permutations of M

1

, . . . , M

t

;

(ii) (Z · M

1

. . . M

t

) = 0 if Z is rationally equivalent to 0;

(iii) if Z ∈ Z

0

(P

n

) then (Z) = deg Z;

(iv) if M

1

∈ Pic

+

(n) then there is an f ∈ Γ (M

1

) not identically van- ishing on Z and for every such f we have

(Z · M

1

. . . M

t

) = (div(f |Z) · M

2

. . . M

t

).

P r o o f. This comprises some of the results from [4], Chaps. 1, 2. Ra-

tionally equivalent cycles in Z

0

have the same degree and if Z, Z

0

∈ Z

t

are

rationally equivalent and f, f

0

∈ Γ (M

1

), then div(f |Z), div(f |Z

0

) are ratio-

nally equivalent. Hence the intersection number can be defined inductively

by (iii), (iv).

(10)

We write (M

1

. . . M

m

) for (P

n

· M

1

. . . M

m

). Further, for divisor classes N

1

, . . . , N

s

and non-negative integers e

1

, . . . , e

s

with e

1

+ . . . + e

s

= t, we write (Z · N

1e1

. . . N

ses

) for (Z · N |

1

. . . N {z

1

}

e1times

. . . N |

s

. . . N {z }

s

es times

). The degree of Z ∈ Z

t

is defined by deg Z := (Z · O(1)

t

).

R e m a r k s . (i) By induction on the dimension it follows easily that if Z ∈ Z

t

is effective and M

1

, . . . , M

t

∈ Pic

+

(n) then (Z · M

1

. . . M

t

) ≥ 0.

Moreover, if Z is a subvariety of P

n

and f

1

∈ Γ (M

1

), . . . , f

t

∈ Γ (M

t

) are

“generic”, then (Z · M

1

. . . M

t

) is precisely the cardinality of the set of points V ∩ {f

1

= . . . = f

t

= 0}.

(ii) Let k

0

be a perfect subfield of k, i.e. every finite extension of k

0

is separable. A k

0

-subvariety of P

n

is a set {P ∈ P

n

: f (P ) = 0 for every f ∈ I}, where I is an essential n-prime ideal of k

0

[X]. Every such k

0

-subvariety Z is a union of equal dimensional subvarieties of P

n

, Z = Z

1

∪ . . . ∪ Z

q

, and we put dim Z := dim Z

1

; now if dim Z = k and M

1

, . . . , M

t

∈ Pic

+

(n) then we define

(2.4) (Z · M

1

. . . M

t

) :=

X

q i=1

(Z

i

· M

1

. . . M

t

).

This is extended by linearity to k

0

-cycles, i.e. finite formal sums of k

0

- subvarieties.

We need some further properties of the intersection number. Let e

1

= (1, 0, . . . , 0), e

2

= (0, 1, . . . , 0), . . . , e

m

= (0, . . . , 0, 1) and put L

h

= O(e

h

) for h = 1, . . . , m. Further, fix d = (d

1

, . . . , d

m

) ∈ (Z

>0

)

m

and put L :=

O(d) = d

1

L

1

+ . . . + d

m

L

m

. If Z

h

= P

Vh

n

Vh

V

h

(h = 1, . . . , m) is a cycle in P

nh

then of course we define

Z

1

× . . . × Z

m

= X

n

V1

n

V2

. . . n

Vm

V

1

× . . . × V

m

.

Further, we denote by π

h

the projection to the hth factor P

n

→ P

nh

and by π

h

the inclusion (“pull back”) k[X

h

] ,→ k[X

1

, . . . , X

m

] = k[X].

Lemma 2. Let Z

h

∈ Z

δh

(P

nh

) (h = 1, . . . , m) and Z = Z

1

× . . . × Z

m

. Put δ = δ

1

+ . . . + δ

m

.

(i) Suppose that f ∈ Γ

n1

does not vanish identically on Z

1

. Then π

1

f does not vanish identically on Z and

(2.5) div(π

1

f |Z) = div(f |Z

1

) × Z

2

× . . . × Z

m

.

(ii) Let e

1

, . . . , e

m

be non-negative integers with e

1

+ . . . + e

m

= δ. Then (Z · L

e11

. . . L

emm

) =

n deg Z

1

. . . deg Z

m

if (e

1

, . . . , e

m

) = (δ

1

, . . . , δ

m

),

0 otherwise.

(11)

(iii) (Z ·L

δ

) = (δ!/(δ

1

! . . . δ

m

!))d

δ11

. . . d

δmm

deg Z

1

. . . deg Z

m

. In particular (L

M

) = C := (M !/(n

1

! . . . n

m

!))d

n11

. . . d

nmm

.

P r o o f. (i) ([4], p. 35, Ex. 2.3.1). This is analogous to the set-theoretic statement that if Z

1

, . . . , Z

m

are varieties then Z ∩ {π

1

f = 0} = (Z

1

∩ {f = 0}) × Z

2

× . . . × Z

m

.

(ii) This follows easily from (i) by induction on δ. Another way is as follows. For h = 1, . . . , m assume that Z

h

is a subvariety of P

nh

, take generic linear forms f

hj

∈ k[X

h

] for j = 1, . . . , e

h

and put W

h

= Z

h

∩ {f

h1

= 0, . . . , f

h,eh

= 0}. Then by remark (i) above (Z·L

e11

. . . L

emm

) is the cardinality of the set W = W

1

× . . . × W

m

. This cardinality is zero if (e

1

, . . . , e

m

) 6=

1

, . . . , δ

m

) since then one of the sets W

h

is empty; while otherwise this cardinality is Q

m

h=1

#W

h

= Q

m

h=1

deg Z

h

. (iii) By additivity we have

(Z · L

δ

) = (Z · (d

1

L

1

+ . . . + d

m

L

m

)

δ

)

= X

e1+...+em=d

δ!

e

1

! . . . e

m

! d

e11

. . . d

emm

(Z · L

e11

. . . L

emm

)

= δ!

δ

1

! . . . δ

m

! d

δ11

. . . d

δmm

deg Z

1

. . . deg Z

m

.

Lemma 3. Suppose that m ≥ 2. Let Z be a δ-dimensional subvariety of P

n

that cannot be expressed as a product Z = Z

1

× . . . × Z

m

with Z

h

⊆ P

nh

for h = 1, . . . , m. Then there are at least two tuples of non-negative integers (e

1

, . . . , e

m

) with e

1

+ . . . + e

m

= δ and (Z · L

e11

. . . L

emm

) > 0.

P r o o f (cf. [11], p. 79). The idea is as follows. By [18], p. 45, Thm. 2, if X is a closed subset of P

n

and f : X → P

n

a morphism, then f (X) is closed, and f maps subvarieties of X to subvarieties of f (X). We apply this with the projections π

h

: P

n

→ P

nh

. Put Z

h

:= π

h

(Z), δ

h

:= dim Z

h

for h = 1, . . . , m. Since Z is not a product, Z is a proper subvariety of Z

1

× . . . × Z

m

and therefore, δ = dim Z < dim Z

1

× . . . × Z

m

= δ

1

+ . . . + δ

m

. We prove by induction on m the following assertion: for each h ∈ {1, . . . , m}

there is a tuple (e

1

, . . . , e

m

) as in the statement of Lemma 3 with e

h

= δ

h

. This implies Lemma 3 since δ

1

+ . . . + δ

m

> δ.

This assertion is obviously true if m = 1. Suppose that the assertion holds for m = r − 1 where r > 1. We prove the assertion for m = r, h = 1, which clearly suffices. In the induction step we proceed by induction on δ

1

. If δ

1

= 0 then Z = Q × W where Q ∈ P

n1

and W is a subvariety of P

n2

×. . .×P

nm

and the assertion follows by applying the induction hypothesis to W . If δ

1

> 0 then choose a linear form f ∈ k[X

1

] that does not identically vanish on Z

1

. Then g := π

1

f does not identically vanish on Z. Clearly, π

1

maps the irreducible components of Z ∩ {g = 0} to those of Z

1

∩ {f = 0}

and the latter have dimension δ

1

− 1. By applying the second induction

(12)

hypothesis to the irreducible components of Z ∩ {g = 0}, we infer that there are non-negative integers e

1

, . . . , e

m

with e

1

+ . . . + e

m

= δ and e

1

= δ

1

such that (div(g|Z) · L

e11−1

. . . L

emm

) > 0. Hence (Z · L

e11

. . . L

emm

) > 0. This proves the assertion.

Lemma 4. Let A be a set of polynomials from Γ

n

(d)\{0} and I the ideal generated by A. Let Z

1

, . . . , Z

t

be irreducible components of codimension t of X := {P ∈ P

n

: f (P ) = 0 for f ∈ A}. Then for all tuples of non-negative integers (e

1

, . . . , e

m

) with e

1

+ . . . + e

m

= M − t one has

X

r i=1

m

Zi

(Z

i

· L

e11

. . . L

emm

) ≤ (L

e11

. . . L

emm

· L

t

), where m

Zi

:= l

OZi

(O

Zi

/IO

Zi

) for i = 1, . . . , r.

P r o o f. This is essentially Proposition 2.3 of [1] and Lemma 6.4, p. 76 of [9]. We give some details of the proof to which we have to refer later. For a subvariety Z of P

n

and f ∈ Γ

n

\{0} not vanishing identically on Z, define the truncated divisor

div

X

(f |Z) := X

W 6⊂X

ord

W

(f |Z)W,

where the sum is taken over all irreducible components of div(f |Z) which are not irreducible components of X. This is extended by linearity to cycles. Put Z

0

:= P

n

and choose inductively f

1

, . . . , f

t

∈ I and define cycles C

1

, . . . , C

t

as follows:

(2.6) for j = 1, . . . , t, f

j

does not vanish identically on C

j−1

, each Z

i

(i = 1, . . . , r) is a subvariety of one of the irreducible components of div(f

j

|C

j−1

), and C

j

:= div

X

(f

j

|C

j−1

);

in the next lemma we explicitly construct such f

j

. Clearly, the irreducible components of C

j

have codimension j. Therefore Z

1

, . . . , Z

r

are irreducible components of C

t

.

We need some more advanced results from intersection theory to estimate the multiplicity m

Zi,Ct

of Z

i

in C

t

from below. By [4], Ex. 7.1.10, p. 123, m

Zi,Ct

is equal to l

OZi

(O

Zi

/I

0

O

Zi

), where I

0

= (f

1

, . . . , f

t

). (Note that P

n

is smooth whence all local rings O

Zi

are Cohen–Macaulay rings.) Further, since I

0

⊆ I we have l

OZi

(O

Zi

/I

0

O

Zi

) ≥ l

OZi

(O

Zi

/IO

Zi

) = m

Zi

. Hence m

Zi,Ct

≥ m

Zi

. It follows that

(2.7) C

t

=

X

r i=1

m

Zi

Z

i

+ (effective cycle).

Further, by (2.6) we have

(2.8) div(f

j

|C

j−1

) = C

j

+ (effective cycle) for j = 1, . . . , t.

Cytaty

Powiązane dokumenty

Lemma 1.1 is the key point for Theorem 1.2, and the crucial point consists of showing that a weak limit of a sequence of solutions of approximate problems is in fact a solution of

It appears that it is enough to check that the mappings change the modulus of continuity of the argument function in a “regular way” (see Definition 1).. In our case

Using tbs relation between a sublinear nonoompactneas measure and the Hauadorff measure the authoress has proved an existence theorem for differential equations in Banach

Kuttner for pointing out an error in the first version of the author’s proof and for suggesting the present version.. (ii) Necessity of

ROCZNIKI POLSKIEGO TOWARZYSTWA MATEMATYCZNEGO Seria I: PRACE MATEMATYCZNE IX (1965).. C

Concerning the present problem - the problem clearly is the giving of a satisfactory account of the relation between Fact and Norm, if facts are facts and norms are norms how are

We construct the Doob–Meyer decomposition of a submartingale as a pointwise superior limit of decompositions of discrete submartingales suitably built upon discretizations of

Precise orbit determination (POD) of the Global Satellite Navigation System (GNSS) is especially important in light of the emerging of new global, and