ebook img

The Abel–Jacobi map for complete intersections PDF

26 Pages·2008·0.19 MB·English
by  
Save to my drive
Quick download
Download
Most books are stored in the elastic cloud where traffic is expensive. For this reason, we have a limit on daily download.

Preview The Abel–Jacobi map for complete intersections

The Abel–Jacobi map for complete intersections J. Nagel 1 Introduction In an attempt to generalize the classical Noether–Lefschetz theorem for sur- faces of degree d ≥ 4 in P3, Griffiths and Harris [G-H] raised a number of questions concerning the behaviour of curves on a very general threefold X of degree d ≥ 6 in P4. One of their questions is whether the image of the Abel–Jacobi map ψ : CH2 (X) → J2(X) is zero. Green [G2] and Voisin X hom [V1] partially solved this problem; they showed that the image of ψ is X contained in the torsion points of J2(X). A similar statement holds for odd- dimensional hypersurfaces in projective space: if X = V(d) ⊂ P2m (m ≥ 2) is a very general hypersurface of degree d ≥ 4+2/(m−1), then the image of the Abel–Jacobi map ψ is contained in the torsion points of Jm(X); see X [G2]. We extend the result of Green and Voisin to smooth complete intersec- tions of odd dimension in projective space (Theorem 4.1). In all but one of the cases where the conditions of Theorem 4.1 are not satisfied, it is known that the image of the Abel–Jacobi map is indeed non–torsion for a very gen- eralmember of the family of complete intersections under consideration. The remaining exceptional case will be dealt with later. To extend the result of Green and Voisin, we have to find an efficient algebraic description of the variable cohomology of complete intersections, analogous to the Jacobi ring description in the case of projective hypersur- faces. This problem has been solved through the work of various people, 1 including Terasoma [Te], Konno [Ko], Libgober–Teitelbaum [L], [L-T] and Dimca [Di2]. The starting point is the following observation, due to Tera- soma: if X = V(f ,...,f ) is a smooth complete intersection of multidegree 0 r (d ,...,d ) in Pn+r+1 with d = ... = d = d, then the variable cohomology 0 r 0 r of X is isomorphic (up to a shift in the Hodge filtration) to the variable co- homology of the hypersurface X = V(F) ⊂ Pr×Pn+r+1 of type (1,d) defined by the bihomogeneous polynomial F(x,y) = y f (x)+...+y f (x). 0 0 r r Konnoextendedthisapproachtothecaseofarbitrarymultidegreebyviewing (f ,...,f ) as a section of the vector bundle E = O (d )⊕...⊕O (d ). The 0 r P 0 P r product of projective spaces is replaced by the projective bundle P(E∨), and X is replaced by the zero locus of the associated section of the tautological line bundle ξ on P(E∨). The variable cohomology is then described via the E pseudo–Jacobi ring introduced by Green [G1]. Working with an additional hypothesis, Libgober obtained a description of the variable cohomology via residues of differential forms defined on Pn+r+1, in the spirit of the work of Griffiths [Gr]; he observes that the variable cohomology is related to a quo- tient of the ring S = C[x ,...,x ,y ,...,y ], where S carries a suitable 0 n+r+1 0 r bigrading. These different approaches were elegantly combined in the recent work of Dimca. He observes that P(E∨), being a smooth and compact toric variety, can be constructed as a geometric quotient. This explains the bigrading on S and shows that P(E∨) behaves like the ordinary projective space in many ways. Giventhis, Terasoma’soriginalapproachgoesthroughwithonlyminor modifications. This paper is oranized as follows: in Section 2, Dimca’s method is used to give a description of the variable cohomology in terms of the Jacobi ring of X in P(E∨). Next we discuss the so–called symmetrizer lemma in Section 3, using which we prove our main result in Section 4. 2 Description of the Jacobi ring Let X = V(d ,...,d ) ⊂ Pn+r+1 be a smooth complete intersection of di- 0 r mension n ≥ 3, where d ≥ 2 for i = 0,...,r. We assume for the moment i 2 that r > 0, i.e., X is not a hypersurface. The (r + 1)–tuple (f ,...,f ) 0 r of equations that define X represents a global section of the vector bundle E = O (d )⊕...O (d ). LetP = P(E∨)betheprojective bundlewhosefiber P 0 P r over a point x ∈ Pn+r+1 is the projective space of hyperplanes in E . Using x theresults in [Cox], we can associateto thesmoothandcompact toricvariety P its ’homogeneous coordinate ring’ S = C[x ,...,x ,y ,...,y ]. This 0 n+r+1 0 r ring carries a natural grading by elements of Pic(P) = Pic(Pn+r+1)×Z ∼= Z2. The variables x (i = 0,...,n+r +1) have bidegree (0,1); the variables y i j (j = 0,...,r) have bidegree (1,−d ). Set j F(x,y) = y f (x)+...+y f (x), 0 0 r r and let X ⊂ P be the divisor defined by F(x,y) ∈ S . 1,0 Remark 2.1. Set N = n+ r + 1. We denote the open subset CN \ {0}× Cr+1 \ {0} ⊂ CN+r+1 by U. The presence of a bigrading on the ring S is a consequence of the construction of P as a geometric quotient U/G, where G = C∗ ×C∗ acts on U by (t ,t ).(x ,...,x ,y ,...,y ) = (t x ,...,t x ,t−d0t y ,...,t−drt y ). 1 2 0 N 0 r 2 0 2 N 2 1 0 2 1 r Good references for this construction are [Cox], [Bat] and [Bat-Cox]. Lemma 2.2 X is a very ample divisor on P if and only if d > 0 for all i i = 0,...,r. Proof: Note that F(x,y) ∈ S represents the global section of the tau- 1,0 tological line bundle ξ = O (1) that corresponds to (f ,...,f ) under the E P 0 r canonical isomorphism H0(P,ξ ) = H0(Pn,E). It is readily verified that ξ E E is very ample if and only if the line bundles O (d ) are very ample for all P i i = 0,...,r, cf. [B-S, (3.2.3)]. (cid:3) Remark 2.3 To prove the above lemma from a toric point of view, one chooses a suitable divisor D of bidegree (1,0) and shows that the support function ψ of the corresponding line bundle O(D) is strictly upper convex D if and only if all the degrees d are positive. i Lemma 2.4 X is non–singular ⇐⇒ X is non–singular. Proof: Since ∂F = f (x) and ∂F ver∂x = r y ∂fj, a point (x,y) ∈ Y is ∂yi i o i j=0 j∂xi singular if and only if f (x) = ... = f (x) =P0 and the matrix (∂fj(x)) jas rank at most r, i.e., if a0nd only if x ∈rX is singular. ∂xi i,j (cid:3) 3 Definition 2.5 Let i (resp. j) be the inclusion of X in Pn+r+1 (resp. the inclusion of X in P). We define the variable cohomology of X and X as Hn (X,Q) = coker(Hn(Pn+r+1,Q))−→i∗ Hn(X,Q)) var Hn+2r(X,Q) = coker(Hn+2r(P,Q)−j→∗ Hn+2r(X,Q)). var Remark 2.6 The notions of variable and primitive cohomology are strongly related: if V is a msooth projective variety and D ⊂ V is a smooth com- plete intersection of ample divisors in V of dimension d, one can show that Hd (D) = i∗Hd (V)⊕Hd (D). pr pr var Let π : P → Pn+r+1 be the projection, and write X˜ = π−1(X) = X ×Pr. Let ι : X → P be the inclusion map, and put ϕ = π◦ι : X → Pn+r+1. Lemma 2.7 The inclusion X˜ ֒→ X induces an isomorphism of Hodge struc- tures ∼ Hn+2r(X,C) −→ Hn (X,C)⊗H2r(Pr,C). var var Proof: Consider the Leray spectral sequences E2p,q = Hp(Pn+r+1,Rqπ∗C) ⇒ Hp+q(P,C) ′Ep,q = Hp(Pn+r+1,Rqϕ∗C) ⇒ Hp+q(X,C) 2 associated to the maps π : P → Pn+r+1 and ϕ : X → Pn+r+1. Consider also the Ku¨nneth spectral sequence ′′Ep,q = Hp(X,C)⊗Hq(Pr,C) ⇒ Hp+q(X˜,C). 2 The local systems Rqπ∗C and Rqϕ∗C can be described as follows: C(−q) if q ≤ 2r, q even Rqπ∗C = 2 (cid:26) 0 otherwise C(−q) if q < 2r, q even 2 Rqϕ∗C =  CX(−r) if q = 2r  0 otherwise.  All the spectral sequences degenerate at E . The Lefschetz hyperplane 2 theorem shows that Ep,q−→∼ ′Ep,q−→∼ ′′Ep,q 2 2 2 4 for (p,q) 6= (n,2r) and En,2r ֒→ ′En,2r−→∼ ′′En,2r. 2 2 2 Hence Hn+2r(X,C) ∼= coker(En,2r → ′En,2r) var 2 2 ∼= coker(En,2r ֒→ ′′En,2r) 2 2 ∼= Hn (X,C)⊗H2r(Pr,C). var (cid:3) For details, see [Ko] or [Te]. The description of the variable cohomology of X strongly resembles the description of the primitive cohomology of a hypersurface in Pn+1. As most of the results are similar to those in [Di1], [Do] and [CGGH], we shall omit their proofs. The Euler vector fields r ∂ e = y 1 i ∂y Xi=0 i and n+r+1 r ∂ ∂ e = x − d y 2 i i i ∂x ∂y Xi=0 i Xi=0 i generate the action of G = C∗ ×C∗ on U = Cn+r+1 \{0}×Cr+1 \{0}. The orbits of this action are the fibers of π : U → P. Definition 2.8. (i) For a monomial f = xα0 ...xαNyβ0...yβr we define |f| = r β , 0 N 0 r 1 i=0 i |f|2,x = in=+0r+1αi, |f|2,y = − rj=0djβj and |f|2 = |f|2,x +|f|P2,y. P P (ii) For a differential form ω = f.dx ∧ ... ∧ dx ∧ dy ∧ ... ∧ dy we s1 si t1 tj set |ω| = |f| + j, |ω| = |f| + i, |ω| = |f| − j d and 1 1 2,x 2,x 2,y 2,y k=1 tk |ω|2 = |ω|2,x +|ω|2,y. P In the statement of the following Lemma, the differential d is written in the form d = d +d . We write Ak = H0(Cn+2r+2,Ωk ), and denote the x y Cn+2r+2 contraction with a vector field e by i . e 5 Lemma 2.9. If ω ∈ Ai and ω′ ∈ Aj, then (i) i (ω ∧ω′) = i (ω)∧ω′+(−1)iω ∧i (ω′) e e e (ii) i (df) = |f| f, i (df) = |f| f e1 1 e2 2 (iii) d (i ω)+i (d ω) = |ω| ω y e1 e1 y 1 (iv) d (i ω)+i (d ω) = |ω| ω x e2 e2 x 2,x (v) d (i ω)+i (d ω) = |ω| ω y e2 e2 y 2,y (vi) |i (ω)| = |i (ω)| = |ω| , k = 1,2. e1 k e2 k k Lemma 2.10. A rational k–form ϕ on U given by 1 ϕ = A (x,y)dx ∧dy I,J I J H(x,y) XI,J |I|+|J|=k ∗ satisfies ϕ = π ω for a rational k–form ω on P if and only if (i) ϕ is G–invariant, i.e., |ϕ| = |ϕ| = 0. 1 2 (ii) i (ϕ) = i (ϕ) = 0. e1 e2 Proof: One easily checks that ϕ = π∗ω for a rational k–form ω on P if and only if ϕ and dϕ are horizontal, i.e., i (ϕ) = i (dϕ) = 0 for all vertical vector e e fields e. This is equivalent to i (ϕ) = i (ϕ) = i (dϕ) = i (dϕ) = 0, hence e1 e2 e1 e2 (cid:3) the assertion follows from the previous Lemma. From now on we shall identify rational differential forms on P with their pullbacks to U. Lemma 2.11. Suppose that ψ ∈ Ak satisfies the following conditions (i) i (ψ) = i (ψ) = 0 e1 e2 (ii) |ψ| 6= 0 and at least one of |ψ| , |ψ| is nonzero. 1 2 2,x Then ψ = i i (ϕ) for some ϕ ∈ Ak+2. e2 e1 6 Proof: If |ψ| = α 6= 0 and |ψ| = β 6= 0, we can write 1 2,x αβψ = α(i (d ψ)+d (i ψ)) e2 x x e2 = αi d ψ = i (d (αψ)) e2 x e2 x = i (d (i (d ψ)+d (i ψ))) e2 x e1 y y e1 = i (d (i (d ψ))) = −i i (d d ψ). e2 x e1 y e2 e1 x y (cid:3) Corollary 2.12. i i (ϕ) ψ ∈ H0(P,Ωk(qX)) ⇐⇒ ψ = e2 e1 P F(x,y)q for some ϕ ∈ Ak+2 with |ϕ| = q and |ϕ| = 0. 1 2 The following Lemma shows how to express dψ in a similar form: Lemma 2.13. If i i (ϕ) ψ = e2 e1 , Fq then i i (Fdϕ−qdF ∧ϕ) dψ = e2 e1 . Fq+1 Lemma 2.14. (i) P(x,y)Ω ψ ∈ H0(P,Ωn+2r+1((q+1)X)) ⇐⇒ ψ = P Fq+1 where Ω = i i (dx ∧...∧dx ∧dy ∧...∧dy ). e2 e1 0 n+r+1 0 r (ii) i i (ϕ) ψ˜ ∈ H0(P,Ωn+2r(qX) ⇐⇒ ψ˜ = e2 e1 P Fq 7 where n+r+1 r ϕ = Q (x,y)Ω ∧dy ∧...∧dy + R (x,y)dx ∧...∧dx ∧Ω , i i 0 r α 0 n α Xi=0 Xα=0 Ω = dx ∧...∧dx ∧...∧dx , i 0 i n+r+1 Ω = dy ∧...∧cdy ∧...∧dy . α 0 α r d Definition 2.15. The Jacobi ideal J(F) ⊂ S is the ideal in S generated by the partial derivatives ∂F ∂F ∂F ∂F ,..., , ,..., . ∂x ∂x ∂y ∂y 0 n+r+1 0 r The Jacobi ring R is the quotient ring S/J(F). The bigrading on S induces a bigrading on R. Proposition 2.16. There is a natural isomorphism Hn−p,p(X) ∼= R , var p,d(X) where n = dim X and d(X) = r d −n−r −2. i=0 i P Proof: We have already seen that Hn−p,p(X) ∼= Hn−p+r,p+r(X). There is an var var exact sequence 0 → Hn−p+r+1,p+r(P) → Hp+r(Ωn−p+r+1(logX)) → Hn−p+r,p+r(X) → 0. pr P var The Leray–Hirsch theorem shows that Hn+2r(P) ∼= Hi(Pn+r+1)⊗Hj(P2r. i+j=n+2r L Write bi(P) = dim Hi(P,C). The above formula implies that bn+2r−2(P) = b (P), so Hn+2r(P) = 0. Hence n+2r pr Hn−p+r,p+r(X) ∼= Hp+r(Ωn−p+r+1(logX)). var P 8 Since X ⊂ P is an ample divisor by Lemma 2.2, we can apply the Bott vanishing theorem on P (see [Bat-Cox, Theorem 7.1]) to obtain Hi(P,Ωj (kX)) = 0 for all i > 0,k > 0 and j ≥ 0. P A spectral sequence argument shows that H0(Ωd((p+r +1)X)) Hp+r(Ωn−p+r+1(logX)) ∼= P P H0(Ωd((p+r)X))+dH0(Ωd−1((p+r)X)), P P where d = dim P = n+2r +1. By Lemma 2.14, an element of H0(P,Ωd((p+r +1)X)) can be written P in the form P(x,y)Ω ψ = P Fp+r+1 where degΩ = (r +1,−d(X)) and deg(P(x,y)) = (p,d(X)). What we have shown so far is that the map Res : S → Fd−p+rHn+2r+1(X) p,d(X) var P(x,y) 7→ [Res(ψ )] P is surjective. If ψ˜ ∈ H0(P,Ωn+2r((p+r)X)), then ψ˜ = ie2ie1(ϕ) and P Fp+r {F( N ∂Qi + r ∂Rj)−(p+r)( N ∂F Q + r ∂F R }Ω dψ˜ = i=0 ∂xi j=0 ∂yj i=0 ∂xi i j=0 ∂yj j , P P Fp+r+1 P P where N = n+r +1. Hence ψ ≡ dψ˜ modH0(Ωn+2r+1((p+r)X) ⇐⇒ P ∈ J(F). P P This shows that Res induces an isomorphism R ∼= Hn−p+r,p+r(X), as p,d(X) var (cid:3) desired. Remark 2.17. (i) If r = 0, then P(E∨) ∼= Pn+1, S = C[x ,...,x ,y ] and F(x,y) = 0 n+1 0 y f (x). Clearly the ring R(F) is different from the Jacobi ring R(f ) 0 0 0 of thehypersurface V(f ) ⊂ Pn+1, but the map α : S → C[x ,...,x ] 0 0 n+1 that sends G(x ,...,x ,y ) to G(x ,...,x ,1) induces an isomor- 0 n+1 0 0 n+1 ∼ phism R(F)p,d(X) −→ R(f0)(p+1)d0−n−1 between the graded pieces of these rings that describe Hn−p,p(X). var 9 (ii) The toric description of P shows that the bidegree of the canonical bundle K is (−r −1,d(X)) and S ∼= H0(P,K ⊗ξp+r+1). P p,d(X) P E (iii) The description of the variable cohomology Hn (X) for a complete var intersection X in an arbitrary smooth and compact toric variety P Σ proceeds along the same lines. The number of Euler vector fields equals the rank of Pic(P(E∨)) = Pic(P )×Z. Σ 3 Symmetrizer lemma Using a version of the symmetrizer lemma, we prove that the infinitesimal invariants associated to certain normal functions are zero. Consequently these normal functions are torsion sections of the fiber space of intermediate Jacobians. We keep the notation of Section 2, but from now on we consider the case where X is a smooth complete intersection in P2m+r of odd dimension n = 2m − 1. In this case we have H2m−1(X) = H2m−1(X) = H2m−1(X). var pr Let U ⊂ PH0(P2m+r,E) be the open subset parametrizing smooth complete intersections, and let f : X → U be the universal family. The cohomology U groups of the fibers of f give rise to a local system HZ = R2m−1f∗Z. Let H2m−1 = H ⊗ O be the associated Hodge bundle; it is filtered by holomor- Z Z U phic subbundles Fp (0 ≤ p ≤ 2m−1). The Hodge bundle comes equipped with a flat connection ∇, the Gauss–Manin connection, whose flat sections • are the sections of the local system H . The filtration of subbundles F is Z shifted by ∇ according to the Griffiths transversality rule ∇Fp ⊂ Ω1 ⊗Fp−1. U Let Jm = H2m−1/(Fm +H ) Z be the sheaf of intermediate Jacobians over U. The Gauss–Manin connection induces a map ∇ : Jm → Ω1 ⊗H2m−1/Fm−1, U whose kernel is denoted by Jm. By abuse of language, a global section of h Jm is called a normal function. h To study the image of the Abel–Jacobi map using normal functions, we ’spread out’ cycles on a very general fiber to relative cycles. 10

Description:
remaining exceptional case will be dealt with later. If ω ∈ Ai and ω′ ∈ Aj, then C. Voisin, Une approche infinitésimale du théor`eme de H.
See more

The list of books you might like

Most books are stored in the elastic cloud where traffic is expensive. For this reason, we have a limit on daily download.