Sorry, your browser cannot access this site
This page requires browser support (enable) JavaScript
Learn more >

双有理几何基础之linear system 上的positivity 以及 Kodaira vanishing theorem 的推论们. 全文抄写自Lazarsfeld 的Positivity in Algebraic Geometry, 不含任何一点原创. 属于是假期闲着无聊想了解下Minimal Model Program 时的一些前置知识. 最重要的部分还是最后一节关于一些有效消失定理们.

All schemes are assumed to be over C\mathbb{C}.

Some Homological Algebra

We first recall some basic facts in homological algebra:

Let XX be a variety, F\mathcal{F}^{\bullet} be the complex of coherent OX\mathcal{O}_{X}-module. The cohomology is defined by Hi(F)=kerdiimdi1\mathcal{H}^{i}(\mathcal{F}^{\bullet})=\frac{\mathop{\mathrm{ker}}d_{i}}{\mathop{\mathrm{im}}d_{i-1}}, where dd is the differential.

A spectral sequence is a colletion of data: (Erp,q,En)(E_{r}^{p,q}, E^{n}) and differential drp,q:Erp,qErp+r,qr+1d_{r}^{p,q}:E_{r}^{p,q}\to E_{r}^{p+r,q-r+1} such that drp+r,qr+1drp,q=0d_{r}^{p+r,q-r+1}\circ d_{r}^{p,q}=0. So drd_{r} actually defines complexes. Moreover, we require Er+1p,q=H0(Erp+r,qr+1)E_{r+1}^{p,q}=H^{0}(E_{r}^{p+\bullet r,q-\bullet r+1}). For fixed p,qp,q, if there is a number r0r_{0} such that drp,q=drpr,q+r1d_{r}^{p,q}=d_{r}^{p-r,q+r-1} for all rr0r\geq r_{0}, then the Erp,qE_{r}^{p,q} finally stabilizes, we say Ep,q=Erp,qE_{\infty}^{p,q}=E_{r}^{p,q} the limiting term. There exists a decreasing filtration Fp+1EnFpEnEn\cdots F^{p+1}E^{n}\subseteq F^{p}E^{n}\subseteq \cdots\subseteq E^{n} (in the context we may assume FpEn=0\cap F^{p}E^{n}=0 and FpEn=En\cup F^{p}E^{n}=E^{n}) such that Ep,q=FpEp+q/Fp+1Ep+qE_{\infty}^{p,q}=F^{p}E^{p+q} /F^{p+1}E^{p+q}. Write

Erp,qEp+q.E_{r}^{p,q}\Rightarrow E^{p+q}.

A filtered complex (C,d)(C^{\bullet}, d) is a collection of subobjects FpCqF^{p}C^{q} such that Fp+qCqFpCq+1F^{p+q}C^{q}\subseteq F^{p}C^{q+1} and dFpCqFpCq+1dF^{p}C^{q}\subseteq F^{p}C^{q+1} (n the context we may assume FpCq=0\cap F^{p}C^{q}=0 and FpCq=Cq\cup F^{p}C^{q}=C^{q}). Let

Zrp,q=d1(Fp+rCp+q+1)FpCp+qd1(Fp+rCp+q+1)Fp+1Cp+q;Z_{r}^{p,q}=\frac{d^{-1}(F^{p+r}C^{p+q+1})\cap F^{p}C^{p+q}}{d^{-1}(F^{p+r}C^{p+q+1})\cap F^{p+1}C^{p+q}};

Brp,q=d(Fpr+1Cp+q1)FpCp+qd(Fpr+qCp+q)Fp+qCp+q,B_{r}^{p,q}=\frac{d(F^{p-r+1}C^{p+q-1})\cap F^{p}C^{p+q}}{d(F^{p-r+q}C^{p+q})\cap F^{p+q}C^{p+q}},

Erp,q=Zrp,qBrp,q.E_{r}^{p,q}=\frac{Z_{r}^{p,q}}{B_{r}^{p,q}}.

The differential is given by

Erp,qZrp,qZr1p,qBr+1p+r,qr+1Brp+r,qr+1Erp+r,qr+1,E_{r}^{p,q}\to \frac{Z_{r}^{p,q}}{Z_{r-1}^{p,q}}\xrightarrow{\sim} \frac{B_{r+1}^{p+r,q-r+1}}{B_{r}^{p+r,q-r+1}}\hookrightarrow E_{r}^{p+r,q-r+1},

then Erp,qE_{r}^{p,q} forms a spectral sequence. The objects E0p,q=FpCp+1/Fp+1Cp+qE_{0}^{p,q}=F^{p}C^{p+1}/F^{p+1}C^{p+q} and

E1p,q=hp+q(grpC)hp+q(C).E_{1}^{p,q}=h^{p+q}(gr^{p}C^{\bullet})\Rightarrow h^{p+q}(C^{\bullet}).

For a double complex we mean (C,,dh,dv)(C^{\bullet,\bullet}, d_{h},d_{v}) such that dh2=dv2=0d_{h}^{2}=d_{v}^{2}=0 andn dhdv+dvdh=0d_{h}d_{v}+d_{v}d_{h}=0. The total complex associated to C,C^{\bullet,\bullet} is

Tot(C)n=p+q=nCp,qTot(C)^{n}=\bigoplus_{p+q=n}C^{p,q}

with differential d=dh+dvd=d_{h}+d_{v}. Then we can define filtered complex by FpCp+q=rpCp+qr,rF^{p}C^{p+q}=\bigoplus_{r\geq p}C^{p+q-r,r}. Define the spectral sequence associated to C,C^{\bullet, \bullet} with E0p,q=Cq,pE_{0}^{p,q}=C^{q,p}, d0=dvd_{0}=d_{v}, E1p,q=Hvq(C,p)E_{1}^{p,q}=H_{v}^{q}(C^{\bullet,p}) and d1d_{1} is induced by dhd_{h}, E2p,q=Hhp(Hvq(C,))E_{2}^{p,q}=H_{h}^{p}(H_{v}^{q}(C^{\bullet,\bullet})). Then

E2p,q=Hhp(Hvq(C,))Hp+q(Tot(C)).E_{2}^{p,q}=H_{h}^{p}(H_{v}^{q}(C^{\bullet,\bullet}))\Rightarrow H^{p+q}(Tot(C)).

The Cartan-Eilenberg resolution of AA^{\bullet} is a double complex C,C^{\bullet,\bullet} such that ApCp,A^{p}\to C^{p,\bullet}, ker(ApAp+1)ker(Cp,Cp+1,)\mathop{\mathrm{ker}}(A^{p}\to A^{p+1})\to \mathop{\mathrm{ker}}(C^{p,\bullet}\to C^{p+1,\bullet}), im(ApAp+1)im(Cp,Cp+1,bullet)\mathop{\mathrm{im}}(A^{p}\to A^{p+1})\to \mathop{\mathrm{im}}(C^{p,\bullet}\to C^{p+1,bullet}) and Hp(A)Hhp(C,)H^{p}(A^{\bullet})\to H_{h}^{p}(C^{\bullet, \bullet}) give injective resolutions. Moreover, we sometimes require 0kerdhi,jCi,jimdhi,j00\to \mathop{\mathrm{ker}}d_{h}^{i,j}\to C^{i,j}\to \mathop{\mathrm{im}}d_{h}^{i,j}\to 0 splits.

Every complex admits a Cartan-Eilenberg resolution. Using this, we can show

Theorem 1 (Grothendieck Spectral Sequence). Consider A\mathcal{A}, B\mathcal{B}, C\mathcal{C} which all all category of abelian sheaves on some projective variety. F:ABF:\mathcal{A}\to \mathcal{B} and G:BCG:\mathcal{B}\to \mathcal{C} are all additive functors such that FF maps injective elements into GG-acyclic part. Then there is a spectral sequence Erp,qE_{r}^{p,q} such that

E2p,q=RpG(RqF(A))Rp+q(FG)(A).E_{2}^{p,q}=R^{p}G(R^{q}F(A))\Rightarrow R^{p+q}(F\circ G)(A).

Divisors, line bundles and linear systems

Definition 2 (Cartier Divisor). A Cartier divisor on XX is a global section of the sheaf KX/OX\mathcal{K}_{X}^{*} /\mathcal{O}_{X}^{*}, where KX\mathcal{K}_{X} of the total quotients. Denote the divisor group by

Div(X)=H0(KX/OX).Div(X)=H^{0}(\mathcal{K}_{X}^{*}/\mathcal{O}_{X}^{*}).

We represent a Cartier divisor by {(Ui,fi)}\{(U_{i},f_{i})\} where fiH0(Ui,KX)f_{i}\in H^{0}(U_{i}, \mathcal{K}_{X}^{*}) and satisfies the cocycle condition fi=gijfjf_{i}=g_{ij}f_{j}, gijH0(UiUj,OX)g_{ij}\in H^{0}(U_{i}\cap U_{j}, \mathcal{O}_{X}^{*}).

Remark 3. The support of divisor D={(Ui,fi)}D=\{(U_{i}, f_{i})\} is the set

Supp(D)={xXxiUi and fi is not unit in OX,x}.\mathop{\mathrm{Supp}}(D)=\{x\in X|x_{i}\in U_{i} \text{ and } f_{i} \text{ is not unit in }\mathcal{O}_{X,x}\}.

If we regard DD as an Weil divisor D=ai[Di]D=\sum a_{i}[D_{i}], the support Supp(D)=ai0Di\mathop{\mathrm{Supp}}(D)=\coprod_{a_{i}\neq 0}\left| D_{i} \right|.

Definition 4. A kk-cycle on variety XX is a Z\mathbb{Z}-linear combination of irreducible subvariety of dimension kk. We denote the group of all kk-cycles by Zk(X)Z_{k}(X).

For further definition of property of divisors, e.g. linear equivalence and effective divisors e.t.c. we refer to Hartshorne.

Example 5. For a morphism f:YXf:Y\to X between schemes, define fDf^{*}D by the pullback of all local section {f1Ui,ffi}\{f^{-1}U_{i}, f^{*}f_{i}\}. To make {f1Ui,ffi}\{f^{-1}U_{i}, f^{*}f_{i}\} really define a Cartier divisor, the section must be nontrivial, which is true if we assume yY\exists y\in Y such that f(y)Supp(D)f(y)\notin Supp(D). (Note here we can assume DD effective and fiH0(Ui,OUi)f_{i}\in H^{0}(U_{i},\mathcal{O}_{U_{i}}^{*}), and use linear extension to all divisors.) If YY is reduced, the requirement is just that no components of YY map into the support of DD.

Definition 6 (Canonical Divisor). Let XX be a nonsingular complete variety of dimension nn. Denote ωX=nΩX\omega_{X}=\wedge^{n}\Omega_{X} to be the canonical bundle on XX, and by KXK_{X} we denote the canonical divisor. Then O(KX)=ωX\mathcal{O}(K_{X})=\omega_{X}.

Definition 7. Let L\mathcal{L} be a line bundle, VH0(X,L)V\subseteq H^{0}(X,\mathcal{L}) is a finite dimensional subspace. Denote V=P(V)\left| V \right| =\mathbb{P}(V) be the projective space of one dimensional subspace of VV. We say V\left| V \right| is complete if V=H0(X,L)V=H^{0}(X,\mathcal{L}).

Definition 8. Let eval:VOXLeval:V\otimes \mathcal{O}_{X}\to\mathcal{L} be the evaluation map. Let b(V)\mathfrak{b}(\left| V \right| ) be the image of VLOXV\otimes \mathcal{L}^{*}\to \mathcal{O}_{X} given by the evaleval. The base point is the locus cutting down by the ideal b(V)\mathfrak{b}(\left| V \right|), denoted by Bs(V)Bs(\left| V \right|). We say L\mathcal{L} is base point free if Bs(V)=Bs(\left| V \right| )=\varnothing.

Example 9. Assume XX is projective variety, DD is a divisor on XX, then for m,n1m,n\geq 1, the natural homomorphism

H0(O(mD))H0(O(nD))H0(O((m+n)D))H^{0}(\mathcal{O}(mD))\otimes H^{0}(\mathcal{O}(nD))\to H^{0}(\mathcal{O}((m+n)D))

defines inclusion b(mD)b(nD)b((m+n)D)\mathfrak{b}(\left| mD \right| )\cdot \mathfrak{b}(\left| nD \right|) \subseteq \mathfrak{b}(\left| (m+n)D \right| ).

Example 10. The linear system V\left| V \right| defines a morphsm

φ:XBs(V)P(V),\varphi:X-Bs(\left| V \right| ) \to \mathbb{P}(\left| V \right| ),

defined by

φ:x(s1(x)::sn(x)),\varphi:x\mapsto (s_{1}(x):\cdots: s_{n}(x)),

where s1,,sns_{1},\dots, s_{n} is a set of basis of VV.

Note if we choose different basis of VV, then the corresponding morphism will differ by an automorphism of P(V)\mathbb{P}(V). If V\left| V \right| is base point free then φ\varphi defines a morphism XPnX\to \mathbb{P}^{n}.

Example 11. Suppose WVW\subseteq V is a subspace, then Bs(V)Bs(W)Bs(\left| V \right| )\subseteq Bs(\left| W \right| ) and on XBs(W)X-Bs(\left| W \right| ), the morphism φW=πφV\varphi_{W}=\pi\circ \varphi_{V}, where π:P(V)P(V/W)P(W)\pi:\mathbb{P}(V)-\mathbb{P}(V /W)\to \mathbb{P}(W) is the linear projection.

Definition 12. We say two Cartier divisors D1,D2Div(X)D_{1}, D_{2}\in Div(X) are numberical equivalent if D1C=D2CD_{1}\cdot C=D_{2}\cdot C for every irreducible curve CXC\subseteq X. A divisor DD is called numerically trivial if it is numerical equivalent to zero.

Definition 13. Denote the set of all numerically trivial divisor by Num(X)Num(X). The Neron Severi group of XX is defined by N1(X)=Div(X)/Num(X)N^{1}(X)=Div(X) /Num(X).

Remark 14. This definition is slightly different with general definition, which is Pic(X)/Pic0(X)Pic(X) /Pic^{0}(X). Using intersection theory on Picard scheme one can prove that they are the same with respect to Z,Q,R\mathbb{Z}, \mathbb{Q},\mathbb{R} coefficients.

Theorem 15. Neron Severi group is torsion free group of finite rank. Its rank is called Picard number and denoted by ρ(X)\rho(X).

Proof

Proof. Recall the intersection number DCD\cdot C is given by Cc1(O(D))\int_{C}c_{1}(\mathcal{O}(D)). Then divisors with trivial first chern class are in Num(X)Num(X). Consider the φ:Div(X)H2(X,Z)\varphi:Div(X)\to H^{2}(X,\mathbb{Z}) arising from the exponential sequence and Div(X)Pic(X)Div(X)\to \mathop{\mathrm{Pic}}(X). Then kerφNum(X)\mathop{\mathrm{ker}}\varphi\subseteq Num(X) and therefore N1(X)N^{1}(X) is the quotient of imφ\mathop{\mathrm{im}}\varphi by some subgroup, which is evidently finite generated and torsion free. ◻

From above proof we also see the intersection number only depends on numerical equivalence, i.e. for D1inumD2jD_{1}^{i}\equiv_{num}D_{2}^{j}, D11D1dimXdimVV=D21D2dimXdimVVD_{1}^{1}\cdots D_{1}^{\dim X-\dim V}\cdot V=D_{2}^{1}\cdots D_{2}^{\dim X-\dim V}\cdot V.

Definition 16 (Rank of coherent sheaf). Let XX be variety of dimension nn, F\mathcal{F} is coherent sheaf on XX. The rank rk(F)\mathop{\mathrm{rk}}(\mathcal{F}) is defined to be lengthOX,ξFξlength_{\mathcal{O}_{X,\xi}}\mathcal{F}_{\xi}, where ξ\xi is the generic point of XX.

Theorem 17 (Asymptotic Riemann Roch). Let XX be projective variety of dimension nn. DD is a divisor on XX, then χ(O(mD))\chi(\mathcal{O}(mD)) is a polynomial of degree n\leq n.

χ(O(mD))=Dnn!mn+O(mn1).\chi(\mathcal{O}(mD))=\frac{D^{n}}{n!}m^{n}+O(m^{n-1}).

More generally, for coherent sheaf F\mathcal{F} on XX,

χ(F(mD))=rk(F)Dnn!mn+O(mn1).\chi(\mathcal{F}(mD))=\mathop{\mathrm{rk}}(\mathcal{F})\frac{D^{n}}{n!}m^{n}+O(m^{n-1}).

Proof

Proof. This is direct corollary of Hirzebruch Riemann Roch Theorem. ◻

Proposition 18. Let XX be a projective variety of dimension nn. Let DD be a divisor on XX with the property that
hi(O(mD))=O(mn1)h^{i}(\mathcal{O}(mD))=O(m^{n-1}) for i>0i>0. Fix a positive rational number α\alpha with 0<αn<Dn0<\alpha^{n}<D^{n}. Then for m0m\gg 0, there exists for any smooth point xXx\in X a divisor E=ExmDE=E_{x}\in \left| mD \right| with multx(D)>mαmult_{x}(D)>m\alpha.

Proof

Proof. Omitted. ◻

Amplitude

Definition 19. Let XX be a proper scheme, L\mathcal{L} is a line bundle on XX,

  1. L\mathcal{L} is very ample if the corresponding morphism φ:XPn\varphi:X\to \mathbb{P}^{n} is closed immersion, and then L=φOPn(1)\mathcal{L}=\varphi^{*}\mathcal{O}_{\mathbb{P}^{n}}(1).

  2. L\mathcal{L} is ample if Lm\mathcal{L}^{m} is very ample for some m>0m>0.

Note this definition applys for proper scheme, therefore every immersion XPnX\to \mathbb{P}^{n} has closed image, and thus a closed immersion. The definition here is compactible with definition in Hartshorne.

Example 20. If D1,,DnD_{1},\dots, D_{n} are ample divisors on dimension nn scheme XX, then miDim_{i}D_{i} are very ample for some mi>0m_{i}>0. Then

m1D1mnDn=m1fH1mnfHn=degφi=1nmiHi>0,m_{1}D_{1}\cdots m_{n}D_{n}=m_{1}f^{*}H_{1}\cdots m_{n}f^{*}H_{n}=\deg \varphi \prod_{i=1}^{n}m_{i}H_{i}>0,

and thus D1Dn>0D_{1}\cdots D_{n}>0.

Theorem 21. Let L\mathcal{L} be a line bundle on proper scheme XX. Then the followings are equivalent:

  1. L\mathcal{L} is ample;

  2. For any coherent sheaf F\mathcal{F} on XX, there exists positive number m1=m1(F)m_{1}=m_{1}(\mathcal{F}) such that Hi(FLm)=0H^{i}(\mathcal{F}\otimes \mathcal{L}^{m})=0 for all mm1m\geq m_{1};

  3. For any coherent sheaf F\mathcal{F} on XX, there exists positive number m2=m2(F)m_{2}=m_{2}(\mathcal{F}) such that FLm\mathcal{F}\otimes \mathcal{L}^{m} is generated by global sections for all mm2m\geq m_{2};

There exists a positive integer m3>0m_{3}>0 such that Lm\mathcal{L}^{m} is very ample for all mm3m\geq m_{3}.

Proof

Proof. See Hartshorne III 5.3 and II 7.5. ◻

Example 22. For L\mathcal{L} a globally generated line bundle on a proper scheme XX, the set

U={yXLmy is globally generated}U=\{y\in X|\mathcal{L}\otimes m_{y}\text{ is globally generated}\}

is open.

To see the assertion, let

V={yXH0(L)LOX/my2 is surjective}.V=\{y\in X|H^{0}(\mathcal{L})\to \mathcal{L}\otimes \mathcal{O}_{X}/m_{y}^{2}\text{ is surjective}\}.

We first check U=VU=V. Since k(y)=Ck(y)=\mathbb{C}, H0(Lmy)LmyH^{0}(\mathcal{L}\otimes m_{y})\to \mathcal{L}\otimes m_{y} is surjective if and only if H0(L)Lymy/my2H^{0}(\mathcal{L})\to \mathcal{L}_{y}\otimes m_{y}/m_{y}^{2} is surjective by Nakayama lemma. Also consider the diagram:

0H0(L¬my)H0(L)Ly00Ly¬my=m2yLy¬OX=m2yLy¬OX=my=Ly0®¯

first row is exact since L\mathcal{L} is generated by global sections. α\alpha is surjective if and only if β\beta is. So U=VU=V.

Next we construct coherent sheaf P\mathcal{P} on XX whose fibre at yy is LOX/my2\mathcal{L}\otimes \mathcal{O}_{X}/m_{y}^{2} and a map u:H0(L)Pu:H^{0}(\mathcal{L})\to \mathcal{P} that on each fibre given by evaluations. If such P\mathcal{P} and uu exist, since coker(u)\mathop{\mathrm{coker}}(u) is coherent, the set V=XSupp(coker(u))V=X-\mathop{\mathrm{Supp}}(\mathop{\mathrm{coker}}(u)) is open. For such P\mathcal{P}, consider the natural projections XpX×XqXX\xleftarrow{p}X\times X\xrightarrow{q}X, take

P=p(qL(OX×X/IΔ2)),\mathcal{P}=p_{*}(q^{*}\mathcal{L}\otimes (\mathcal{O}_{X\times X}/\mathcal{I}_{\Delta}^{2})),

where Δ\Delta is the diagonal. It’s easy to verify P\mathcal{P} has desired properties.

Example 23. If DD is ample divisor on XX and EE is any divisor, then mD+EmD+E is very ample for m0m\gg 0. This is an easy corollary of the fact that tensor product of globally generated sheaves is still globally generated.

Example 24. If L\mathcal{L} and M\mathcal{M} are ample line bundles on projective schemes XX and YY respectively, using Segre embedding, LM\mathcal{L} \boxtimes \mathcal{M} is ample on X×YX\times Y.

Proposition 25. Let f:YXf:Y\to X be a finite morphism of proper schemes, and L\mathcal{L} an ample line bundle on XX. Then fLf^{*}\mathcal{L} is an ample line bundle on YY. In particular, if YXY\subseteq X is a subscheme of XX, then the restriction LY\mathcal{L}|_{Y} is ample.

Proof

Proof. Let F\mathcal{F} be a coherent sheaf on YY. Then f(FfLm)=fFLmf_{*}(\mathcal{F}\otimes f^{*}\mathcal{L}^{m})=f_{*}\mathcal{F}\otimes \mathcal{L}^{m}. Since ff is finite, RifFfLm=0R^{i}f_{*}\mathcal{F}\otimes f^{*}\mathcal{L}^{m}=0 for all i>0i>0, and fFf_{*}\mathcal{F} is coherent. Then Hi(Y,FfLm)=Hi(X,fFfLm)H^{i}(Y,\mathcal{F}\otimes f^{*}\mathcal{L}^{m})=H^{i}(X,f_{*}\mathcal{F}\otimes f^{*}\mathcal{L}^{m}). (This is Hartshorne exercise III.8.1, but we will check it here:

Let 0MI0\to \mathcal{M}\to \mathcal{I}^{\bullet} be an injective resolution of M\mathcal{M}, then 0fMfI0\to f_{*}\mathcal{M}\to f_{*}\mathcal{I} is acyclic since RifM=0R^{i}f_{*}\mathcal{M}=0 for all i>0i>0, and

Hi(X,fM)=Hi(Γ(X,fI))=Hi(Γ(Y,I))=Hi(Y,M).)H^{i}(X,f_{*}\mathcal{M})=H^{i}(\Gamma(X,f_{*}\mathcal{I}^{\bullet}))=H^{i}(\Gamma(Y,\mathcal{I}^{\bullet}))=H^{i}(Y,\mathcal{M}).)

Corollary 26. Suppose L\mathcal{L} is generated by global sections, let ϕ:XP(H0(L))\phi: X\to \mathbb{P}(H^{0}(\mathcal{L})) be the morphism corresponding to L\mathcal{L}. Then the followings are equivalent:

  1. L\mathcal{L} is ample.

  2. ϕ\phi is finite.

  3. For every irreducible curve CXC\subseteq X, Cc1(L)>0\int_{C}c_{1}(\mathcal{L})>0.

Proof

Proof. We have shown that if ϕ\phi is finite, then L\mathcal{L} is ample and ample implies Cc1(L)>0\int_{C}c_{1}(\mathcal{L})>0.

Suppose Cc1(L)>0\int_{C}c_{1}(\mathcal{L})>0 for every curve CC and ϕ\phi is not finite. Then there is a subvariety ZXZ\subseteq X of dimension 1\geq 1 that contracts to a point by ϕ\phi. Pick any irreducible curve CZC\subseteq Z, L=ϕO(1)\mathcal{L}=\phi^{*}\mathcal{O}(1) pullbacks to trivial sheaf on CC, since CC maps to a point. Thus degLCCc1(L)=0\deg\mathcal{L}|_{C}-\int_{C}c_{1}(\mathcal{L})=0. ◻

Proposition 27. Let XX be proper scheme and L\mathcal{L} is a line bundle on XX,

  1. L\mathcal{L} is ample if and only if LOXred\mathcal{L}\otimes \mathcal{O}_{X_{red}} is ample on XredX_{red}.

  2. L\mathcal{L} is ample if and only if L\mathcal{L} restricts to every irreducible component is ample.

Proof

Proof. We only prove the “if” statement for (1), (2) is similar.

Consider any coherent sheaf F\mathcal{F} on XX and N\mathcal{N} is the nilradical sheaf of OX\mathcal{O}_{X}, then we can find a filtration:

FNFNrF=0.\mathcal{F}\supseteq \mathcal{N}\mathcal{F}\supseteq \cdots\supseteq \mathcal{N}^{r}\mathcal{F}=0.

The quotients NiF/Ni+1F\mathcal{N}^{i}\mathcal{F}/\mathcal{N}^{i+1}\mathcal{F} are all coherent OXred=OX/N\mathcal{O}_{X_{red}}=\mathcal{O}_{X} /\mathcal{N}-modules. Thus Hj(X,NiF/Ni+1FLm)=0H^{j}(X,\mathcal{N}^{i}\mathcal{F}/\mathcal{N}^{i+1}\mathcal{F}\otimes \mathcal{L}^{m})=0 for j>0j>0 for all m0m\gg 0 by the assumption.

By induction and taking cohomology to the exact sequence

0Ni+1FNiFNiF/Ni+1F0,0\to \mathcal{N}^{i+1}\mathcal{F}\to \mathcal{N}^{i}\mathcal{F}\to \mathcal{N}^{i}\mathcal{F} /\mathcal{N}^{i+1}\mathcal{F}\to 0,

we know Hj(X,NiFLm)=0H^{j}(X,\mathcal{N}^{i}\mathcal{F}\otimes \mathcal{L}^{m})=0 for all j>0j>0 and m0m\gg 0. ◻

Theorem 28 (Amplitude families). Let f:XTf:X\to T be a proper morphism of schemes, L\mathcal{L} is a line bundle on XX and for tTt\in T, Lt=LXt\mathcal{L}_{t}=\mathcal{L}|_{X_{t}}. If L0\mathcal{L}_{0} is ample on X0X_{0} for some point 0T0\in T, then there exists an open neighborhood UU of 00 such that Lt\mathcal{L}_{t} is ample for every tUt\in U.

Proof

Proof. The statement is local on TT so we may assume T=SpecAT=\mathop{\mathrm{Spec}}A.

We first proof for any coherent sheaf F\mathcal{F} on XX there is an integer m(F)m(\mathcal{F}) such that Rif(FLm)=0R^{i}f_{*}(\mathcal{F}\otimes \mathcal{L}^{m})=0 in UmTU_{m}\subseteq T for all i1i\geq 1 and mm(F)m\geq m(\mathcal{F}). We proceed by decreasing induction on ii, as Rif(FLm)=0R^{i}f_{*}(\mathcal{F}\otimes \mathcal{L}^{m})=0 for idimXi\geq \dim X. Consider the maximal ideal m0m_{0} corresponding to 0T0\in T and pick up generators u1,,upm0u_{1},\dots, u_{p}\in m_{0}. Pulling back the exact sequence $$A^{p}\to A\to A/m_{0}\to 0$$ via ff and tensoring with F\mathcal{F} give rise to exact sequence

0ker(fu1)OXpFfu1FFOX00.0\to \mathop{\mathrm{ker}}(f^{*}u\otimes 1)\to \mathcal{O}_{X}^{p}\otimes\mathcal{F}\xrightarrow{f^{*}u\otimes 1}\mathcal{F}\to \mathcal{F}\otimes \mathcal{O}_{X_{0}}\to 0.

Apply induction hypothesis to ker(fu1)\mathop{\mathrm{ker}}(f^{*}u\otimes 1) one has

Rif(ker(fu1)Lm)=0 for m0.R^{i}f_{*}(\mathop{\mathrm{ker}}(f_{*}u\otimes 1)\otimes \mathcal{L}^{m})=0 \text{ for }m\gg 0.

Since L0\mathcal{L}_{0} is ample, Rif(FOX0Lm)=Hi(X0,(FLm)X0)=0R^{i}f_{*}(\mathcal{F}\otimes \mathcal{O}_{X_{0}}\otimes \mathcal{L}^{m})=H^{i}(X_{0},(\mathcal{F}\otimes \mathcal{L}^{m})|_{X_{0}})=0 for mm large. Chasing the diagram one easily get

OTRi1f(FLm)Ri1f(FLm)\mathcal{O}_{T}\otimes R^{i-1}f_{*}(\mathcal{F}\otimes \mathcal{L}^{m})\to R^{i-1}f_{*}(\mathcal{F}\otimes \mathcal{L}^{m})

is surjective on a neighborhood UmU_{m}' for mm large. Since u1u\otimes 1 factors through m0Ri1f(FLm)m_{0}R^{i-1}f_{*}(\mathcal{F}\otimes \mathcal{L}^{m}), for m0m\gg 0, m0Ri1f(FLm)=Ri1f(FLm)m_{0}\cdot R^{i-1}f_{*}(\mathcal{F}\otimes \mathcal{L}^{m})=R^{i-1}f_{*}(\mathcal{F}\otimes \mathcal{L}^{m}) in UmU_{m}'. By Nakayama lemma Ri1f(FLm)=0R^{i-1}f_{*}(\mathcal{F}\otimes \mathcal{L}^{m})=0. This finishes the induction.

Next we show that ρm:ffLmLm\rho_{m}:f^{*}f_{*}\mathcal{L}^{m}\to\mathcal{L}^{m} is surjective along XtX_{t} for tUmt\in U_{m}'', m0m\gg 0. Since Rif(IX0Lm)=0R^{i}f_{*}(\mathcal{I}_{X_{0}}\otimes \mathcal{L}^{m})=0 in UmU_{m} around 0, we have fLmfOX0Lm=H0(X0,L0m)f_{*}\mathcal{L}^{m}\to f_{*}\mathcal{O}_{X_{0}}\otimes \mathcal{L}^{m}= H^{0}(X_{0},\mathcal{L}_{0}^{m}) surjective for m0m\gg 0. We may take large mm such that L0m\mathcal{L}_{0}^{m} is generated by global sections. Thus we get a surjection fL0mOX0L0mf_{*}\mathcal{L}_{0}^{m}\otimes \mathcal{O}_{X_{0}}\to \mathcal{L}_{0}^{m} surjective, which shows ρm\rho_{m} is surjective along X0X_{0}. By the coherence of the cokernel, we find such UmU_{m}''.

Shrinking TT to UmU_{m}'' and make it affine we amy assume ρm\rho_{m} is surjection for fixed large integer mm. Since TT is affine, assume fLmf_{*}\mathcal{L}^{m} is generated by rr sections. Pull it back to XX, we get surjection fOTrLmf^{*}\mathcal{O}_{T}^{r}\to \mathcal{L}^{m} and this defines a morphism ϕ:XPTr1\phi:X\to \mathbb{P}_{T}^{r-1}. ϕ\phi is finite on X0X_{0}, and hence finite on a open neighborhood UU of 0. We conclude Ltm\mathcal{L}_{t}^{m} is ample when tUt\in U. ◻

Theorem 29 (Asymptotic Riemann Roch). Same set up as Theorem 17, but assume DD is ample. Then

h0(O(mD))=Dnn!mn+O(mn1).h^{0}(\mathcal{O}(mD))=\frac{D^{n}}{n!}m^{n}+O(m^{n-1}).

More generally, for coherent sheaf F\mathcal{F} on XX,

h0(F(mD))=rk(F)Dnn!mn+O(mn1).h^{0}(\mathcal{F}(mD))=\mathop{\mathrm{rk}}(\mathcal{F})\frac{D^{n}}{n!}m^{n}+O(m^{n-1}).

Proof

Proof. Clear by Serre vanishing. ◻

Example 30. Let XX be a projective scheme and DD, EE ample divisors on XX. Then there is an integer m=m(D,E)m=m(D,E) such that

H0(OX(aD))H0(OX(bE))H0(OX(aD+bE))H^{0}(\mathcal{O}_{X}(aD))\otimes H^{0}(\mathcal{O}_{X}(bE))\to H^{0}(\mathcal{O}_{X}(aD+bE))

is surjective for a,bma,b\geq m.

To see the assertion, consider the diagram XpX×XqXX\xleftarrow{p}X\times X\xrightarrow{q} X and the exact sequence

0IΔOX×XOΔ0.0\to \mathcal{I}_{\Delta}\to \mathcal{O}_{X\times X}\to \mathcal{O}_{\Delta}\to 0.

We write (aD,bE)(aD,bE) for the divisor paD+qbEp^{*}aD+q^{*}bE. By Kunneth formula, it suffices to show H1(X×X,IΔ(aD,bE))=0H^{1}(X\times X, \mathcal{I}_{\Delta}(aD,bE))=0 for all a,bm0a,b\geq m_{0}. Pick a resolution of IΔ\mathcal{I}_{\Delta} by

OX×X(p1D,p1E)OX×X(p0D,p0E)IΔ0,\cdots \to \oplus\mathcal{O}_{X\times X}(-p_{1}D,-p_{1}E)\to \oplus \mathcal{O}_{X\times X}(-p_{0}D,-p_{0}E)\to \mathcal{I}_{\Delta}\to 0,

we only need to show Hi(X×X,OX×X(api1D,bpi1E))=0H^{i}(X\times X,\mathcal{O}_{X\times X}(a-p_{i-1}D, b-p_{i-1}E))=0 for a,bma,b\geq m. But there are only finitely many nonzero terms, we can pick mm large enough to fulfill the requirement.

Remark 31. Note in previous proof, we use a subtle fact:

Let FM0\mathcal{F}_{\bullet}\to \mathcal{M}\to 0 be a resolution of sheaves on dimension nn scheme XX, and Hk+i(Fi)=0H^{k+i}(\mathcal{F}_{i})=0 for all ii. Then Hk(M)=0H^{k}(\mathcal{M})=0.

The proof is simply split the resolution into short exact sequences and embed Hk(M)H^{k}(\mathcal{M}) into Hk+i+1(im(Fi+1Fi))H^{k+i+1}(\mathop{\mathrm{im}}(\mathcal{F}_{i+1}\to \mathcal{F}_{i})).

Definition 32 (Relative Amplitude). Let f:XTf:X\to T be a proper morphism of schemes, let L\mathcal{L} be a line bundle on XX,

  1. L\mathcal{L} is very ample relative to ff if the canonical map ρ:ffLL\rho: f^{*}f_{*}\mathcal{L}\to \mathcal{L} is surjective and defines an immersion

XP(f¤L)Ti

  1. L\mathcal{L} is ample relative to ff if Lm\mathcal{L}^{m} is very ample for some m>0m>0.

Clearly relative ampleness is local on the target.

Corollary 33. Let f:XTf:X\to T be a proper morphism of schemes, and L\mathcal{L} a line bundle on XX. Then the following are equivalent:

  1. L\mathcal{L} is ff-ample.

  2. Given any coherent sheaf F\mathcal{F} on XX, there exists a positive integer m1=m1(F)m_{1} = m_{1}(\mathcal{F}) such that

Rif(FLm)=0 for all i>0,mm1.R^{i}f_{*}(\mathcal{F}\otimes\mathcal{L}^{m})=0 \text{ for all }i>0, m\geq m_{1}.

  1. Given any coherent sheaf F\mathcal{F} on XX, there is a positive integer m2=m2(F)m_{2} = m_{2}(\mathcal{F}) such that the canonical mapping

ff(FLm)FLmf^{*}f_{*}(\mathcal{F} \otimes\mathcal{L}^{m})\to \mathcal{F}\otimes\mathcal{L}^{m}

is surjective whenever mm2m \geq m_{2}.

  1. There is a positive integer m3>0m_{3} > 0 such that Lm\mathcal{L}^{m} is ff-very ample for every mm3m \geq m_{3}.
Proof

Proof. This follows from the local property of ff-ampleness and Theorem 21. ◻

Example 34. Very ample relative to ff is equivalent to the existence of coherent sheaf F\mathcal{F} and an immersion i:XP(F)i:X\to \mathbb{P}(\mathcal{F}) such that L=iOP(F)(1)\mathcal{L}=i^{*}\mathcal{O}_{\mathbb{P}(\mathcal{F})}(1).

One direction is clear, just take F=fL\mathcal{F}=f_{*}\mathcal{L}. For the converse, by the universal property, there is a surjection fFLf^{*}\mathcal{F}\to \mathcal{L}, which factor through the pullback of σ:FfL\sigma: \mathcal{F}\to f_{*}\mathcal{L} via ff, i.e.

fFfσffLρL.f^{*}\mathcal{F}\xrightarrow{f^{*}\sigma}f^{*}f_{*}\mathcal{L}\xrightarrow{\rho}\mathcal{L}.

Clearly ρ\rho is surjective and defines a morphism j:XP(fL)j:X\to \mathbb{P}(f_{*}\mathcal{L}) and therefore factors through P(fL)P(cokerσ)P(F)\mathbb{P}(f_{*}\mathcal{L})-\mathbb{P}(\mathop{\mathrm{coker}}\sigma)\to \mathbb{P}(\mathcal{F}) and jj. The image of jj lies in the open set P(fL)P(cokerσ)\mathbb{P}(f_{*}\mathcal{L})-\mathbb{P}(\mathop{\mathrm{coker}}\sigma). Since the projection P(fL)P(cokerσ)P(F)\mathbb{P}(f_{*}\mathcal{L})-\mathbb{P}(\mathop{\mathrm{coker}}\sigma)\to \mathbb{P}(\mathcal{F}) is locally of finite type, we get jj is also an immersion.

Example 35. Suppose we have a diagram

YXT¹gf

μ\mu is finite TT-morphism. If L\mathcal{L} is ff-ample, then μL\mu^{*}\mathcal{L} is gg-ample.

Note that for any coherent sheaf F\mathcal{F} on YY, RiμF=0R^{i}\mu_{*}\mathcal{F}=0, so by Grothendieck spectral sequence

E2i,j=Rif(Rjμ(FμLm))Ri+jg(FμLm)E_{2}^{i,j}=R^{i}f_{*}(R^{j}\mu_{*}(\mathcal{F}\otimes \mu^{*}\mathcal{L}^{m}))\Rightarrow R^{i+j}g_{*}(\mathcal{F}\otimes \mu^{*}\mathcal{L}^{m})

vanishes for i+j>0i+j>0.

Theorem 36. Let f:XTf:X\to T be a proper morphism of schemes. L\mathcal{L} is a line bundle on XX. Then L\mathcal{L} is ff-ample if and only if Lt\mathcal{L}_{t} is ample on XtX_{t} for every tTt\in T.

Proof

Proof. By above we know ff-ampleness of L\mathcal{L} implies ampleness of Lt\mathcal{L}_{t}. Conversely, if Lt\mathcal{L}_{t} is ample, by Theorem 28, there is an open neighborhood UU of tt such that LU\mathcal{L}|_{U} is ff-ample. Since ff-ampleness is local on targert, L\mathcal{L} itself is ample. ◻

Theorem 37 (Nakai-Moishezon-Kleiman criterion). Let L\mathcal{L} be a line bundle on a projective scheme XX. Then L\mathcal{L} is ample if and only if $$\int_{V} c_{1}(L)^{\dim(V)} > 0$$ for every positive-dimensional irreducible subvariety VXV\subseteq X (including the irreducible components of XX).

Proof

Proof. If we have L\mathcal{L} ample, then Lm\mathcal{L}^{m} is very ample, therefore

mdimVVc1(L)dimV=Vc1(Lm)dimV>0.m^{\dim V}\int_{V}c_{1}(\mathcal{L})^{\dim V}=\int_{V}c_{1}(\mathcal{L}^{m})^{\dim V}>0.

Conversely, assume the positivity of intersection numbers. We may assume XX is reduced and irreducible. We may apply the induction on dimension. For curve case, we have already known the result. Assume for dimension n1\leq n-1. Note that the group Div(X)Div(X) is generated by very ample divisors (by Serre vanishing, any divisor DD and ample divisor AA', D+mAD+mA' is globally generated and D+mA+AD+mA'+A' is very ample.) We may write D=ABD=A-B, we have an exact sequence

0OX(mDB)OX((m+1)D)OA((m+1)D)00\to \mathcal{O}_{X}(mD-B)\to \mathcal{O}_{X}((m+1)D)\to \mathcal{O}_{A}((m+1)D)\to 0

and

0OX(mDB)OX(mD)OB(mD)0.0\to \mathcal{O}_{X}(mD-B)\to \mathcal{O}_{X}(mD)\to \mathcal{O}_{B}(mD)\to 0.

By induction hypothesis, OA(D)\mathcal{O}_{A}(D) and OB(D)\mathcal{O}_{B}(D) are all ample. For mm large,

Hi(OX(mD))Hi(OX(mDB))=Hi(OX((m+1)D))H^{i}(\mathcal{O}_{X}(mD))\to H^{i}(\mathcal{O}_{X}(mD-B))=H^{i}(\mathcal{O}_{X}((m+1)D))

for i2i\geq 2. So χ(OX(mD))=h0(OX(mD))h1(OX(mD))+C\chi(\mathcal{O}_{X}(mD))=h^{0}(\mathcal{O}_{X}(mD))-h^{1}(\mathcal{O}_{X}(mD))+C for some constant CC when m0m\gg 0. By Theorem 17, for large mm we have H0(OX(mD))0H^{0}(\mathcal{O}_{X}(mD))\neq 0. Replacing DD by mDmD, we may assume DD is effective.

Consider the exact sequence

0OX((m1)D)OX(mD)OD(mD)0,0\to \mathcal{O}_{X}((m-1)D)\to \mathcal{O}_{X}(mD)\to \mathcal{O}_{D}(mD)\to 0,

OX(D)\mathcal{O}_{X}(D) is ample by induction. It follows that for m0m\gg 0, H1(OX((m1)D))H1(OX(mD))H^{1}(\mathcal{O}_{X}((m-1)D))\to H^{1}(\mathcal{O}_{X}(mD)) is surjective and will eventually become isomorphism due to the dimension. Therefore the map H0(OX(mD))H0(OD(mD))H^{0}(\mathcal{O}_{X}(mD))\to H^{0}(\mathcal{O}_{D}(mD)) is surjective for mm large. Since OD(mD)\mathcal{O}_{D}(mD) is base point free, the restriction of elements in mD\left| mD \right| to SuppD\mathop{\mathrm{Supp}}D do not share a base point. Since OX(mD)\mathcal{O}_{X}(mD) is base point free from SuppD\mathop{\mathrm{Supp}}D, we conclude that OX(mD)\mathcal{O}_{X}(mD) is globally generated. Apply Corollary 26, we get
OX(mD)\mathcal{O}_{X}(mD) is ample. ◻

Example 38. Let XX be projective variety and DD an effective divisor on XX whose normal bundle OD(D)\mathcal{O}_{D}(D) is ample, then

  1. For m0m\gg 0, OX(mD)\mathcal{O}_{X}(mD) is globally generated.

  2. For m0m\gg 0, the restriction map H0(OX(mD))H0(OD(mD))H^{0}(\mathcal{O}_{X}(mD))\to H^{0}(\mathcal{O}_{D}(mD)) is surjective.

There is a proper birational morphism f:XXˉf:X\to \bar{X} from XX to a projective variety Xˉ\bar{X} such that ff is an isomorphism on neighborhood of DD and Dˉ=f(D)\bar{D}=f(D) is ample effective divisor on Xˉ\bar{X}.

1. and 2. are from above proof. We only show for 3. Take mm large to make the first two assertions hold. Take Xˉ\bar{X} to be the image of Stein factorization ff of ϕ:XP(H0(OX(mD)))\phi:X\to \mathbb{P}(H^{0}(\mathcal{O}_{X}(mD))). Since OD(mD)\mathcal{O}_{D}(mD) is ample, there is an open neighborhood of f(D)f(D) such that ff is finite. Since fOX=OXˉf_{*}\mathcal{O}_{X}=\mathcal{O}_{\bar{X}}, we know the neighborhood of DD is isomorphic to its image.

Example 39. Using similar skills as in the proof, we get

hi(F(mD))=O(mn)h^{i}(\mathcal{F}(mD))=O(m^{n})

for coherent sheaf F\mathcal{F} and divisor DD on XX.

Example 40. We can also get the converse of Proposition 26, i.e. Let f:YXf:Y\to X be a finite morphism of proper schemes, L\mathcal{L} be a line bundle on XX. If fLf^{*}\mathcal{L} is ample, then L\mathcal{L} is ample.

To see so, let VXV\subseteq X be an variety on XX. Since ff is finite, there is an irreducible variety WW\subseteq maps to VV with the same dimension. Then by projection formula,

deg(fW)Vc1(L)dimV=Wc1(fL)>0.\deg(f|_{W}) \int_{V}c_{1}(\mathcal{L})^{\dim V}=\int_{W}c_{1}(f^{*}\mathcal{L})>0.

Definition 41 (Q\mathbb{Q} and R\mathbb{R}-divisors). We say DD is a Q\mathbb{Q}-divisor if DD is a finite sum aiAi\sum a_{i}A_{i}, where aiQa_{i}\in \mathbb{Q} and AiDiv(X)A_{i}\in Div(X). The set of Q\mathbb{Q}-divisors is denoted by Div(X)Q=Div(X)QDiv(X)_{\mathbb{Q}}=Div(X)\otimes\mathbb{Q}.

We say DD is an R\mathbb{R}-divisor if DD is a finite sum aiAi\sum a_{i}A_{i}, where aiRa_{i}\in \mathbb{R} and AiDiv(X)A_{i}\in Div(X). The set of R\mathbb{R}-divisors is denoted by Div(X)R=Div(X)RDiv(X)_{\mathbb{R}}=Div(X)\otimes\mathbb{R}.

Similarly we can define the effective Q\mathbb{Q} and R\mathbb{R}-divisors and NQ1(X)N^{1}_{\mathbb{Q}}(X), NR1(X)N^{1}_{\mathbb{R}}(X).

Remark 42. Q\mathbb{Q}-divisors D1D_{1}, D2D_{2} are linearly equivalent if there is an integer rr such that rD1linrD2rD_{1}\equiv_{lin}rD_{2}, i.e. r(D1D2)r(D_{1}-D_{2}) is in the image of principal divisors in Div(X)Div(X).

Definition 43. A Q\mathbb{Q}-divisor DD is ample if one of the following equivalent conditions is satisfied:

  1. D=aiA1D=\sum_{a_{i}}A_{1} where ai>0a_{i}>0 and AiA_{i} are ample Cartier divisors.

  2. There’s a positive integer r>0r>0 such that rDrD is ample Cartier divisor.

  3. For any variety VXV\subseteq X, Vc1(OX(D))dimV>0\int_{V}c_{1}(\mathcal{O}_{X}(D))^{\dim V}>0.

Definition 44. An R\mathbb{R}-divisor DD is ample if D=aiA1D=\sum_{a_{i}}A_{1} where ai>0a_{i}>0 and AiA_{i} are ample Cartier divisors.

Remark 45. Nakai’s criterion for ampleness does not hold for R\mathbb{R}-divisors.

Proposition 46. *Let XX be a projective variety and HH is an ample R\mathbb{R}-divisor. For finitely many R\mathbb{R}-divisors E1,,ErE_{1},\dots,E_{r} the R\mathbb{R}-divisor

H+ϵ1E1++ϵrErH+\epsilon_{1}E_{1}+\cdots+\epsilon_{r}E_{r}

is ample for sufficiently small 0ϵi10\leq \left| \epsilon_{i} \right| \ll 1.*

Proof

Proof. First prove for Q\mathbb{Q}-divisors. Omitted. ◻

Remark 47. Amplitude of Q\mathbb{Q} and R\mathbb{R}-divisors only depends on numerical equivalence.

Definition 48. Let D=aiDiD=\sum a_{i} D_{i} be a Q\mathbb{Q}(resp. R\mathbb{R})-divisor on XX. The round up D\lceil D \rceil and D=[D]\lfloor D \rfloor=[D] are integer divisors:

D=aiDi;\lceil D \rceil=\sum \lceil a_{i}\rceil D_{i};

D=aiDi.\lfloor D\rfloor =\sum \lfloor a_{i}\rfloor D_{i}.

The round operators do not in general commutes with pullback and compatible with numerical equivalence.

Nef Divisors

Definition 49 (Nef Line Bundles and Divisors). Let XX be a proper scheme. A line bundle L\mathcal{L} is nef if for every irreducible curve CXC\subseteq X, Cc1(L)0\int_{C}c_{1}(\mathcal{L})\geq 0. We say a divisor DD is nef if OX(D)\mathcal{O}_{X}(D) is nef.

Remark 50. By Chow’s lemma, every proper variety XX admits a projection model, i.e. a surjective birational morphism μ:XX\mu:X'\to X such that XX' is projective, By the projection formula on intersection products, nefness of divisor DD is equivalent to the nefness of divisor μD\mu^{*}D.

By the similar virtue, we can show:

  1. If f:XYf:X\to Y is a proper morphism. L\mathcal{L} is nef line bundle on YY, then fLf^{*}\mathcal{L} is nef on XX.

  2. If f:XYf:X\to Y is a proper surjective morphism. fMf^{*}\mathcal{M} is nef bundle on XX, then M\mathcal{M} is also nef on YY.

  3. L\mathcal{L} is nef if and only if Lred\mathcal{L}_{red} is nef on XredX_{red}.

  4. L\mathcal{L} is nef if and only if the restriction to any irreducible components is nef.

Example 51. Let XX be a proper variety, DXD\subseteq X be a effective divisor. If ND/X=OD(D)\mathcal{N}_{D/X}=\mathcal{O}_{D}(D) is nef, then DD is nef.

Theorem 52 (Kleiman). Let XX be a projective scheme. If DD is nef R\mathbb{R}-divisor on XX, then DkV0D^{k}\cdot V\geq 0 for all VXV\subseteq X of dimension kk.

Proof

Proof. We will do it by induction on dimension of XX. The base case for XX which is a curve is clear. So we may assume DkV0D^{k}\cdot V\geq 0 of rall variety of dimension n1\leq n-1 and show Dn0D^{n}\geq 0.

We first show for Q\mathbb{Q}-divisor DD. Take ample divisor HH on XX and define $$P(t)=(D+tH)^{n}\in\mathbb{R}.$$ Suppose P(0)<0P(0)<0. Since HnkH^{n-k} for k<nk<n is effective kk-cycle, DnHnk0D^{n}\cdot H^{n-k}\geq 0. So the coefficients of P(t)P(t) is positive except for the last term. Thus P(t)P(t) has a single root t0>0t_{0}>0. For any t>t0t>t_{0}, (D+tH)kV>0(D+tH)^{k}\cdot V>0 by expanding into intersection terms. So D+tHD+tH is ample. Let Q(t)=D(D+tH)n1Q(t)=D\cdot (D+tH)^{n-1} and R(t)=tH(D+tH)n1R(t)=tH\cdot (D+tH)^{n-1}. For t>t0t>t_{0}, D(D+tH)n10D\cdot (D+tH)^{n-1}\geq 0 by nefness. Thus by continuity of Q(t)Q(t), Q(t0)0Q(t_{0})\geq 0. But R(t0)>0R(t_{0})>0 by Nakai criterion for ampleness. Then P(t0)>0P(t_{0})>0, contradiction!

Now for any R\mathbb{R}-divisor DD, one may choose ample divisor H1,,HrH_{1},\dots, H_{r} spanning N1(X)RN^{1}(X)_{\mathbb{R}}. Set D(ϵ1,,ϵr)=D+ϵ1H1++ϵrHrD(\epsilon_{1},\dots, \epsilon_{r})=D+\epsilon_{1}H_{1}+\cdots+\epsilon_{r}H_{r}. Then D(ϵ1,,ϵr)D(\epsilon_{1},\dots, \epsilon_{r}) is clearly nef by the definition of nefness. We can choose (ϵi)(\epsilon_{i}) such that D(ϵ1,,ϵr)D(\epsilon_{1},\dots, \epsilon_{r}) is Q\mathbb{Q}-divisor which approaches to DD. ◻

Corollary 53. Let XX be a projective variety and DD be a nef R\mathbb{R}-divisor on XX, HH be an ample R\mathbb{R}-divisor on XX. Then D+ϵHD+\epsilon H is ample for all ϵ>0\epsilon>0. Conversely, if DD and HH are any two divisors such that D+ϵHD+\epsilon H is ample for all sufficiently small 0<ϵ00<\epsilon \ll 0, then DD is nef.

Proof

Proof. If D+ϵHD+\epsilon H is ample then for any effective curve CC we have (D+ϵH)C>0(D+\epsilon H)\cdot C>0. Take ϵ0\epsilon \to 0 we get DC0D\cdot C\geq 0, so DD is nef.

Assume DD is nef and HH is ample, it suffices to show D+HD+H is ample. For any subvariety VXV\subseteq X of dimension kk,

(D+H)kV=s=0k(ks)(HsDksV),(D+H)^{k}\cdot V=\sum_{s=0}^{k}\binom{k}{s}(H^{s}\cdot D^{k-s}\cdot V),

here HsVH^{s}\cdot V is represented by effective (ks)(k-s)-cycle. So HsDksV0H^{s}\cdot D^{k-s}\cdot V\geq 0. However, HkV>0H^{k}\cdot V>0, thus (D+H)kV>0(D+H)^{k}\cdot V>0. Then we have done the proof for Q\mathbb{Q}-divisors by Nakai criterion. For R\mathbb{R}-divisors, one can argue similar to the proof of Kleiman’s theorem. ◻

Example 54. Let δ1,,δnN1(X)R\delta_{1},\dots,\delta_{n}\in N^{1}(X)_{\mathbb{R}} be nef classes on projective variety XX. Then δ1δn0\delta_{1}\cdots \delta_{n}\geq 0.

One may replace δ1,,δn\delta_{1},\dots, \delta_{n} by ample divisors D1+ϵH,,Dn+ϵHD_{1}+\epsilon H,\dots, D_{n}+\epsilon H, where HH is ample. Then (D1+ϵH)(Dn+ϵH)>0(D_{1}+\epsilon H)\cdots (D_{n}+\epsilon H)>0 and take ϵ0\epsilon\to 0.

Corollary 55. Let XX be projective variety and HHbe an ample R\mathbb{R}-divisor on XX. Fix an R\mathbb{R}-divisor DD on XX. Then DD is ample if and only if there exists ϵ>0\epsilon >0 such that for every irreducible curve CXC\subseteq X,

DCHCϵ.\frac{D\cdot C}{H\cdot C}\geq \epsilon.

Proof

Proof. DCHCϵ\frac{D\cdot C}{H\cdot C}\geq \epsilon is equivalent to say DϵHD-\epsilon H is nef. The remaining proof is direct from above corollary. ◻

Example 56. If H1H_{1}, H2H_{2} are ample divisors on projective variety XX. Then there are rational number M,m>0M,m>0 such that

mH1CH2CMH1CmH_{1}\cdot C\leq H_{2}\cdot C\leq MH_{1}\cdot C

for all irreducible curve CXC\subseteq X.

This is archieved by choose appropriate MM and mm such that MH1H2MH_{1}-H_{2} and H1mH2H_{1}-mH_{2} are both ample.

Theorem 57. Let XX be projective variety and DD be a divisor on XX. Then DD is ample if and only if there exists a positive number ϵ>0\epsilon>0 such that DCmultxCϵ\frac{D\cdot C}{mult_{x}C}\geq \epsilon for every point xXx\in X and every irreducible curve CXC\subseteq X pass through XX.

Proof

Proof. Omitted. ◻

Proposition 58. Let f:XTf:X\to T be a surjective proper morphism of varieties. L\mathcal{L} is a line bundle on XX. If L0\mathcal{L}_{0} is nef for some 0T0\in T. Then there is a countable union BTB\subseteq T of proper subvarieties of TT noting containing 00 such that Lt\mathcal{L}_{t} is nef for all tTBt\in T-B.

Proof

Proof. One can assume ff is projective by Chow’s lemma. After shrinking TT we may write L=OX(D)\mathcal{L}=\mathcal{O}_{X}(D) where DD’s support does not contain any fibre XtX_{t}. Fix a divisor AA on XX such that At=AXtA_{t}=A|_{X_{t}} is ample for all tt. Then DtD_{t} is nef if and only if Dt+1mAtD_{t}+\frac{1}{m}A_{t} is ample for every integer m>0m>0. By assumption this holds for t=0t=0, then apply Theorem 28 we get BmB_{m} such that Dt+1mAtD_{t}+\frac{1}{m}A_{t} is ample on XBmX-B_{m}. Then take the union of all those excisions. ◻

Example 59. We say a surface XX is minimal if it does not contain any rational curve CC whose self intersection C2=1C^{2}=-1. Assume XX is smooth projective variety with κ(X)0\kappa(X)\geq 0. Then XX is minimal if and only if KXK_{X} is nef.

To see so, fix any divisor DmKXD\in\left| mK_{X} \right| and write D=aiCiD=\sum a_{i}C_{i} with ai>0a_{i}>0 and CiC_{i} irreducible curves. If CXC\subseteq X is any irreducible curve such that KXC<0K_{X}\cdot C<0, then DC<0D\cdot C<0, but CiC>0C_{i}\cdot C>0 unless C=CiC=C_{i}. So we may assume C=C1C=C_{1}. Then a1C1C1DC1<0a_{1}C_{1}\cdot C_{1}\leq D\cdot C_{1}<0. By adjunction formula, the genus g(C)=0g(C)=0 and C2=1C^{2}=-1.

Conversely, if CC is a rational (1)(-1) curve, then the adjunction formula shows KXC=1K_{X}\cdot C=-1.

Definition 60. The ample cone Amp(X)N1(X)RAmp(X)\subseteq N^{1}(X)_{\mathbb{R}} is the convex cone of all ample R\mathbb{R}-divisor classes of XX. The nef cone Nef(X)N1(X)RNef(X)\subseteq N^{1}(X)_{\mathbb{R}} is the convex cone of all nef R\mathbb{R}-divisor classes.

Theorem 61. Nef(X)Amp(X)Nef(X)\subseteq \overline{Amp}(X) and Amp(X)=int(Nef(X))Amp(X)=int(Nef(X)).

Proof

Proof. Clearly Amp(X)Amp(X) is open and Nef(X)Nef(X) is closed. We have already known that Amp(X)Nef(X)\overline{Amp}(X)\subseteq Nef(X) and Amp(X)int(Nef(X))Amp(X)\subseteq int(Nef(X)).

Let HH be an ample divisor and DD be a nef divisor on XX. Then D+ϵHD+\epsilon H is ample for all ϵ>0\epsilon>0. Thus DD is limit of ample divisors and Nef(X)Amp(X)Nef(X)\subseteq \overline{Amp}(X).

For DD lies in int(Nef(X))int(Nef(X)), DϵHD-\epsilon H is still nef for ϵ1\epsilon \ll 1. So D=DϵH+ϵHD=D-\epsilon H+\epsilon H is ample and int(Nef(X))Amp(X)int(Nef(X))\subseteq Amp(X). ◻

Definition 62. Let XX be a proper variety. Let Z1(X)RZ_{1}(X)_{\mathbb{R}} be the R\mathbb{R}-vector space of all real one-cycles on XX. An element γZ1(X)R\gamma\in Z_{1}(X)_{\mathbb{R}} can be written as γ=aiCi\gamma=\sum a_{i}C_{i} where CiXC_{i}\subseteq X are irreducible curves. Two cycles γ1\gamma_{1}, γ2\gamma_{2} are numerically equivalent if Dγ1=Dγ2D\cdot \gamma_{1}=D\cdot \gamma_{2} for all divisor DD. We write the space modulo numerical equivalence of one-cycles by N1(X)RN_{1}(X)_{\mathbb{R}}.

By construction, there is a perfect pairing N1(X)R×N1(X)RRN_{1}(X)_{\mathbb{R}}\times N^{1}(X)_{\mathbb{R}}\to \mathbb{R}.

Definition 63. Let XX be a proper variety. The cone of curves NE(X)N1(X)RNE(X)\subseteq N_{1}(X)_{\mathbb{R}} is the cone spanned by all effective one-cycles on XX.

Proposition 64. Let NE(X)\overline{NE}(X) be the closure of NE(X)NE(X) in N1(X)RN_{1}(X)_{\mathbb{R}}. Then NE(X)\overline{NE}(X) it the dual of Nef(X)Nef(X), i.e.

NE(X)={γN1(X)Rγδ0 for all δNef(X)}.\overline{NE}(X)=\{\gamma\in N_{1}(X)_{\mathbb{R}}|\gamma\cdot \delta \geq 0 \text{ for all } \delta \in Nef(X)\}.

Proof

Proof. Let KVK\subseteq V be a closed convex cone of finite dimensional real vector space. The dual K={ϕVϕ(x)0xK}K^{*}=\{\phi\in V^{*}|\phi(x)\geq 0 \forall x\in K\}. Then take V=N1(X)RV=N_{1}(X)_{\mathbb{R}} and K=NE(X)K=\overline{NE}(X) we get the desired proposition. ◻

Fix divisor DD that is no numerically trivial. Let ϕD:N1(X)RR\phi_{D}:N_{1}(X)_{\mathbb{R}}\to \mathbb{R} given by ϕD:γγD\phi_{D}:\gamma\to \gamma\cdot D. Set D=kerϕDD^{\perp}=\mathop{\mathrm{ker}}\phi_{D} and D>0={γN1(X)RDγ>0}D_{>0}=\{\gamma\in N_{1}(X)_{\mathbb{R}}|D\cdot \gamma>0\}.

Theorem 65 (Kleiman Criterion for Amplitude). Let XX be a projective variety and DD is an R\mathbb{R}-divisor on XX. Then DD is ample if and only if NE(X){0}D>0\overline{NE}(X)-\{0\}\subseteq D_{>0}. Equivalently, choose any norm on N1(X)RN_{1}(X)_{\mathbb{R}}, let S={γN1(X)Rγ=1}S=\{\gamma\in N_{1}(X)_{\mathbb{R}}| || \gamma ||=1\}. Then DD is ample if and only if NE(X)SD>0S\overline{NE}(X)\cap S\subseteq D_{>0}\cap S.

Proof

Proof. Assume NE(X)SD>0S\overline{NE}(X)\cap S\subseteq D_{>0}\cap S. ϕD(γ)>0\phi_{D}(\gamma)>0 for all γNE(X)S\gamma\in \overline{NE}(X)\cap S. Since NE(X)S\overline{NE}(X)\cap S is compact, there is positive ϵ\epsilon such that ϕD(γ)ϵ\phi_{D}(\gamma)\geq \epsilon for all γNE(X)S\gamma\in \overline{NE}(X)\cap S. Then DCϵCD\cdot C\geq \epsilon || C || for all irreducible curve CXC\subseteq X. Take ample basis H1,,HrH_{1},\dots, H_{r} of N1(X)RN^{1}(X)_{\mathbb{R}}. Since N1(X)RN_{1}(X)_{\mathbb{R}} is finite dimensional, norms on N1(X)RN_{1}(X)_{\mathbb{R}} are all equivalent. So we can choose norm by x=Hix|| x ||=\sum | H_{i}\cdot x|. Then for some suitable ϵ>0\epsilon'>0, DCϵHCD\cdot C\geq \epsilon'H\cdot C. So DD is ample by Corollary 54, DD is ample.

The proof for the converse part is just reverse the argument. ◻

Example 66. Let XX be a projective variety and HH be an ample divisor on XX. Let N1(X)ZN_{1}(X)_{\mathbb{Z}} be the Z\mathbb{Z}-coefficient group of one-cycles. NE(X)Z=NE(X)N1(X)Z\overline{NE}(X)_{\mathbb{Z}}=\overline{NE}(X)\cap N_{1}(X)_{\mathbb{Z}}. Then for any M>0M>0, {γNE(X)ZHγM}\{\gamma\in \overline{NE}(X)_{\mathbb{Z}}|H\cdot \gamma\leq M\} is a finite set.

One can choose ample R\mathbb{R}-basis H1,,HrH_{1},\dots, H_{r} such that H=HiH=\sum H_{i}. THen for γNE(X)\gamma \in \overline{NE}(X), Hγ=HiγH\cdot \gamma=\sum \left| H_{i}\cdot \gamma \right| is a norm of NE(X)Z\overline{NE}(X)_{\mathbb{Z}}. The set is the ball of radius MM in the norm.

Definition 67. Let KVK\subseteq V be a closed convex cone in a finite dimensional real vector space. An extremal ray rr is one dimensional subcone such that if v+wrv+w\in r, then v,wrv,w\in r.

Theorem 68. Let XX be a projective variety of dimension nn. Let δ1,,δnN1(X)R\delta_{1},\dots, \delta_{n}\in N^{1}(X)_{\mathbb{R}} be nef classed. Then

(δ1δn)n(δ1n)(δnn).(\delta_{1}\cdots \delta_{n})^{n}\geq (\delta_{1}^{n})\cdots (\delta_{n}^{n}).

Proof

Proof. It suffices to prove for δ1,,δn\delta_{1},\dots, \delta_{n} ample classes. Pass to the resolution of singularities, we may assume XX is smooth. We use induction on dimension of XX. The surface case is done in Hartshorne Ex V.1.9. Assume the results for dimension n1\leq n-1. For any given ample classes B1,Bn1,HN1(X)RB_{1},\dots B_{n-1}, H\in N^{1}(X)_{\mathbb{R}}, we first show the inequality

(B1Bn1H)n1(B1n1H)(Bn1n1H).(B_{1}\cdots B_{n-1}H)^{n-1}\geq (B_{1}^{n-1}\cdot H)\cdots (B_{n-1}^{n-1}\cdot H).

By continuity it suffices to show for B1,,Bn1,HN1(X)RB_{1},\dots, B_{n-1}, H\in N^{1}(X)_{\mathbb{R}} are all very ample integer divisors. By moving lemma, we can assume HH intersect with BiB_{i} transversally. Let Bi\overline{B_{i}} be the restriction of BiB_{i} to HH, then the inequality becomes

(B1Bn1)n1(B1n1)(Bn1n1)(\overline{B_{1}}\cdots \overline{B_{n-1}})^{n-1}\geq (\overline{B_{1}}^{n-1})\cdots (\overline{B_{n-1}}^{n-1})

on HH, which is true by induction hypothesis.

Next we show the original inequality. Let δ1,,δnN1(X)R\delta_{1},\dots, \delta_{n}\in N^{1}(X)_{\mathbb{R}} be ample classes. Fix index j{1,,n}j\in\{1,\dots, n\} and apply above inequality, we have

(δ1δn)n1ij(δin1δj).(\delta_{1}\cdots \delta_{n})^{n-1}\geq \prod_{i\neq j}(\delta_{i}^{n-1}\cdot \delta_{j}).

By above inequality again (with H=B1==Bn2=δiH=B_{1}=\cdots =B_{n-2}=\delta_{i} and Bn1=δjB_{n-1}=\delta_{j}),

(δ1n1δj)n1(δin)n2(δiδjn1).(\delta_{1}^{n-1}\cdot \delta_{j})^{n-1}\geq (\delta_{i}^{n})^{n-2}(\delta_{i}\cdot \delta_{j}^{n-1}).

Thus

jij(δin1δj)n1jij(δin)n2(δiδjn1)=(i(δin)(n1)(n2))(jij(δjδin1)),\prod_{j}\prod_{i\neq j}(\delta_{i}^{n-1}\delta_{j})^{n-1}\geq \prod_{j}\prod_{i\neq j}(\delta_{i}^{n})^{n-2}(\delta_{i}\cdot \delta_{j}^{n-1})=(\prod_{i}(\delta_{i}^{n})^{(n-1)(n-2)})(\prod_{j}\prod_{i\neq j}(\delta_{j}\cdot \delta_{i}^{n-1})),

cancel jij(δjδin1)\prod\limits_{j}\prod\limits_{i\neq j}(\delta_{j}\cdot \delta_{i}^{n-1}) and take the n2n-2-th root we get the desired inequality. ◻

Corollary 69. Let XX be an projective variety of dimension nn, and fix an integer 0pn0 \leq p \leq n. Let α1,,αp,β1,,βnpN1(X)R\alpha_{1},\dots, \alpha_{p},\beta_{1},\dots, \beta_{n-p}\in N^{1}(X)_{\mathbb{R}} be nef classes. Then

(α1αpβ1βnp)p(α1pβ1βnp)(αppβ1βnp).(\alpha_{1}\cdots \alpha_{p}\cdot \beta_{1}\cdots \beta_{n-p})^{p}\geq (\alpha_{1}^{p}\cdot \beta_{1}\cdots\beta_{n-p})\cdots (\alpha_{p}^{p}\cdot \beta_{1}\cdots\beta_{n-p}).

Corollary 70. Let XX be a projective variety of dimension nn, and let α,βN1(X)R\alpha,\beta\in N^{1}(X)_{\mathbb{R}} be nef classes on X. Then the following inequalities are satisfied:

  1. For any integers 0qpn0\leq q\leq p\leq n,

(αqβnq)p(αpβnp)q(βp)pq.(\alpha^{q}\cdot \beta^{n-q})^{p}\geq (\alpha^{p}\cdot \beta^{n-p})^{q}\cdot (\beta^{p})^{p-q}.

  1. For any 0in0\leq i\leq n,

(αiβni)n(αn)i(βn)ni.(\alpha^{i}\cdot \beta^{n-i})^{n}\geq (\alpha^{n})^{i}\cdot (\beta^{n})^{n-i}.

  1. ((α+β)n)1/n(αn)1/n+(βn)1/n.((\alpha+\beta)^{n})^{1/n}\geq (\alpha^{n})^{1/n}+(\beta^{n})^{1/n}.

Asymptotic Theory

Consider the line bundle L\mathcal{L} on a projective variety XX if H0(L)0H^{0}(\mathcal{L})\neq 0, then there is an effective divisor mDLmmD\in \left| \mathcal{L}^{m} \right|, where DLD\in \left| \mathcal{L} \right|. We thus know that H0(Lm)0H^{0}(\mathcal{L}^{m})\neq 0 for all m>0m>0. For general case, we define $$N(L):={m\geq 0|H{0}(\mathcal{L}{m})\neq 0}$$ and e(L)e(\mathcal{L}) be the largest number such that me(L)me(\mathcal{L}) lies in N(L)N(\mathcal{L}) for m0m\gg 0. We also write N(D)N(D) and e(D)e(D) to denote N(OX(D))N(\mathcal{O}_{X}(D)) and e(OX(D))e(\mathcal{O}_{X}(D)).

Example 71. Let XX be a projective variety and L\mathcal{L} is a torsion element in Pic(X)Pic(X). Assume the order of L\mathcal{L} to be mm, then N(L)=mNN(\mathcal{L})=m\mathbb{N}.

Clearly Lm=OX\mathcal{L}^{m}=\mathcal{O}_{X} has global sections. For mnm\neq n, suppose Ln\mathcal{L}^{n} has a global section, say it corresponding to effective divisor DD. Then mD=0mD=0. Since the Neron Severi group of XX is torsion free, D=num0D=_{num}0. However, there is a very ample divisor HH on XX, thus XHdimX1D>0\int_{X}H^{\dim X-1}\cdot D>0 by Nakai’s criterion, contradiction.

Definition 72 (Iitaka Dimension). Assume that XX is normal. Then the Iitaka dimension of L\mathcal{L} is defined to be

κ(L)=maxmN(L){dimϕm(X)},\kappa(\mathcal{L})=\max_{m\in N(\mathcal{L})}\{\dim \phi_{m}(X)\},

where ϕm:XP(H0(Lm))\phi_{m}:X\to \mathbb{P}(H^{0}(\mathcal{L}^{m})) is the morphism defined by the global sections of Lm\mathcal{L}^{m}, provided N(L)0N(\mathcal{L})\neq 0;

If H0(Lm)0H^{0}(\mathcal{L}^{m})\neq 0 for m>0m>0, set κ(L)=\kappa(\mathcal{L})=-\infty.

If XX is not normal, pass to it to the normalization π:XX\pi:X'\to X. and set κ(L)=κ(πL)\kappa(\mathcal{L})=\kappa(\pi^{*}\mathcal{L}). By our definition, either κ(L)=\kappa(\mathcal{L})=-\infty or 0κ(L)dimX0\leq \kappa(\mathcal{L})\leq \dim X.

Definition 73. Let XX be a smooth projective variety. The Kodaira dimension κ(X)=κ(KX)\kappa(X)=\kappa(K_{X}) is defined to be the Iitaka dimension of canonical class.

Example 74. Assume XX is normal, then dimϕm(X)=κ(L)\dim \phi_{m}(X)=\kappa(\mathcal{L}) for all sufficiently large mN(L)m\in N(\mathcal{L}).

Without losing of generality assume H0(L)0H^{0}(\mathcal{L})\neq 0 and there exists k>0k>0 such that dimϕk(X)=κ(L)\dim \phi_{k}(X)=\kappa(\mathcal{L}). Consider the exact sequence 0LkLk+n0\to \mathcal{L}^{k}\to \mathcal{L}^{k+n}, taking H0H^{0} one can get

0H0(Lk)uH0(Lk+n).0\to H^{0}(\mathcal{L}^{k})\xrightarrow{u}H^{0}(\mathcal{L}^{k+n}).

This in turn give rise to ϕk=πϕk+n\phi_{k}=\pi\circ \phi_{k+n}, where π:P(H0(Lk+n))P(cokeru)P(H0(Lk))\pi: \mathbb{P}(H^{0}(\mathcal{L}^{k+n}))-\mathbb{P}(\mathop{\mathrm{coker}}u)\to \mathbb{P}(H^{0}(\mathcal{L}^{k})) is the linear projection with center P(cokeru)\mathbb{P}(\mathop{\mathrm{coker}}u)/ Therefore dimϕk+n(X)dimϕk(X)\dim \phi_{k+n}(X)\geq \dim \phi_{k}(X) for all m1m\geq 1. The reverse inequality if direct from the definition.

Definition 75. An algebraic fibre space is a surjective projective morphism f:XYf:X\to Y such that fOX=OYf_{*}\mathcal{O}_{X}=\mathcal{O}_{Y}.

Example 76. Let f:XYf:X\to Y be a projective surjective morphism of normal varieties and the function field extension K(Y)K(X)K(Y)\subseteq K(X). K(Y)K(Y) is algebraically closed in K(X)K(X) if and only if ff defines an algebraic fibre space.

We only prove for the only if. Consider the Stein factorization XgXhYX\xrightarrow{g}X'\xrightarrow{h}Y of ff. hh induces a finite algebraic extension K(Y)K(X)K(Y)\subseteq K(X') which is actually an isomorphism. Since YY is normal and locally hOXh_{*}\mathcal{O}_{X'} is a finite OY\mathcal{O}_{Y}-algebra, hOXOYh_{*}\mathcal{O}_{X'}\cong \mathcal{O}_{Y}. Furthermore, since hh is affine, we actually get hh is an isomorphism.

Lemma 77. Let f:XYf:X\to Y be a fibre space. L\mathcal{L} is a line bundle on YY. Then H0(X,fLm)=H0(Y,Lm)H^{0}(X,f^{*}\mathcal{L}^{m})=H^{0}(Y,\mathcal{L}^{m}) for all m>0m>0.

Proof

Proof. Direct from projection formula. ◻

Example 78. Let XX, YY be projective varieties, f:XYf:X\to Y is a fibre space. Then the morphism f:PicYPicXf^{*}:\mathop{\mathrm{Pic}}Y\to \mathop{\mathrm{Pic}}X is injective.

For any line bundle L\mathcal{L} on YY, L\mathcal{L} and L\mathcal{L}^{*} cannot both have global sections unless L=OX\mathcal{L}=\mathcal{O}_{X}. (Consider the section DLD\in \left| \mathcal{L} \right| and exact sequence 0LDOXOD00\to \mathcal{L}^{*}\xrightarrow{\cdot D} \mathcal{O}_{X}\to \mathcal{O}_{D}\to 0) If fLOXf^{*}\mathcal{L}\cong \mathcal{O}_{X}, then

H0(Y,L)=H0(X,fL)=H0(X,OX)=H0(X,fL)=H0(Y,L)=C.H^{0}(Y,\mathcal{L})=H^{0}(X,f^{*}\mathcal{L})=H^{0}(X,\mathcal{O}_{X})=H^{0}(X,f^{*}\mathcal{L})=H^{0}(Y,\mathcal{L})^{*}=\mathbb{C}.

Example 79. For a birational morphism f:YXf:Y\to X between normal varieties, let L\mathcal{L} be a line bundle on XX and EE be the exceptional divisor of ff on YY. Then there is an isomorphism

H0(fL)H0(fLOY(E)).H^{0}(f^{*}\mathcal{L})\cong H^{0}(f^{*}\mathcal{L}\otimes \mathcal{O}_{Y}(E)).

Note fOY(E)(U)={gK(Y)=K(X)div(g)+E0 on E}f_{*}\mathcal{O}_{Y}(E)(U)=\{g\in K(Y)=K(X)|div(g)+E\geq 0 \text{ on }E\}. Since EE maps to a set of codimension 2\geq 2, div(g)+E0    div(g)=0div(g)+E\geq 0 \iff div(g)=0. Thus fOY(E)=OXf_{*}\mathcal{O}_{Y}(E)=\mathcal{O}_{X} and then all the equalities are followed by projection formula.

Definition 80. For a line bundle L\mathcal{L} on projective variety XX, the section ring is the graded C\mathbb{C}-algebra

R(L)=m0H0(X,Lm).R(\mathcal{L})= \bigoplus_{m\geq 0}H^{0}(X,\mathcal{L}^{m}).

We say L\mathcal{L} is globally generated if R(L)R(\mathcal{L}) is a finitely generated C\mathbb{C}-algebra.

Definition 81. The stable base locus is the set B(D)=m1Bs(mD)B(D)=\bigcap\limits_{m\geq 1}Bs(\left| mD \right| ).

Proposition 82. The stable base locus B(D)B(D) is the unique minimal element in the family of algebraic sets {Bs(mD)}m1\{Bs(\left| mD \right| )\}_{m\geq 1}. Moreover, there exist an integer m0m_{0} such that B(D)=Bs(km0D)B(D)=Bs(\left| km_{0}D \right| ) for all k0k\gg 0.

Proof

Proof. For m,l1m,l\geq 1, we have inclusion Bs(lmD)Bs(md)Bs(lmD)\subseteq Bs(md). Thus the set {Bs(mD)}m1\{Bs(mD)\}_{m\geq 1} have a unique minimal element. Then by the decreasing chain condition we know that such m0m_{0} exists. ◻

As an easy corollary, B(mD)=B(D)B(mD)=B(D) for all m1m\geq 1.

Let XX be a normal projective variety and L\mathcal{L} be an semiample line bundle on XX. Suppose Lm\mathcal{L}^{m} is globally generated, then SkH0(Lm)S^{k}H^{0}(\mathcal{L}^{m}) is a globally generated subsystem of Lm\left| \mathcal{L}^{m} \right|, corresponding to the kk-th Veronese embedding of Ym:=ϕm(X)Y_{m}:=\phi_{m}(X). By the inclusion of global sections, the morphism ϕm\phi_{m} factors as a composition of ϕkm\phi_{km} and a linear projection pik:YkmYmpi_{k}:Y_{km}\to Y_{m}, i.e. ϕm=πkϕkm\phi_{m}=\pi_{k}\circ \phi_{km}. Since πkYkm\pi_{k}|_{Y_{km}} is affine and proper morphism between noetherian schemes, πkYkm\pi_{k}|Y_{km} is proper (Affineness implies the morphism is equivalent to a coherent OYm\mathcal{O}_{Y_{m}}-algebra, which is finite by properness.) Let Mm\mathcal{M}_{m} be the very ample line bundle given by the embedding into P(H0(Lm))\mathbb{P}(H^{0}(\mathcal{L}^{m})), then ϕmMm=Lm\phi_{m}^{*}\mathcal{M}_{m}=\mathcal{L}^{m}.

Lemma 83. *For mM(L)m\in M(\mathcal{L}) and k0k\gg 0, the composition

XϕkmYkmπkYmX\xrightarrow{\phi_{km}}Y_{km}\xrightarrow{\pi_{k}}Y_{m}

gives the Stein factorization of ϕm\phi_{m}. Furthermore ϕkm\phi_{km} defines a fibre space and YkmY_{km}, ϕkm\phi_{km} are independent of kk for k0k\gg 0.*

Proof

Proof. Let XfVgYmX\xrightarrow{f} V\xrightarrow{g}Y_{m} be the Stein factorization of ϕm\phi_{m}. Let Fm\mathcal{F}_{m} be the ample line bundle on YmY_{m} that pulls back to Lm\mathcal{L}^{m}. Since gg is finite, G=gFm\mathcal{G}=g^{*}\mathcal{F}_{m} is also ample. So Gk\mathcal{G}^{k} is very ample for some k0k\gg 0.

Also fGk=Lkmf^{*}\mathcal{G}^{k}=\mathcal{L}^{km} and H0(X,Lkm)=H0(V,Gk)H^{0}(X,\mathcal{L}^{km})=H^{0}(V,\mathcal{G}^{k}). Then one knows ϕkm\phi_{km} factors through VV by the universal property of projective space. Since VPnV\to \mathbb{P}^{n} is closed immersion, VV is the image of XX under ϕkm\phi_{km}. Thus Ykm=YY_{km}=Y and ϕkm=f\phi_{km}=f for k0k\gg 0. ◻

Theorem 84 (Semiample Fibration). There is a fibre space ϕ:XY\phi:X\to Y such that for any sufficiently large mM(L)m\in M(\mathcal{L}), Ym=YY_{m}=Y and ϕm=ϕ\phi_{m}=\phi. Furthermore there’s an ample line bundle M\mathcal{M} on YY such that ϕM=Lf\phi^{*}\mathcal{M}=\mathcal{L}^{f}, where f=f(L)f=f(\mathcal{L}) is the exponent of M(L)M(\mathcal{L}).

Proof

Proof. We may assume f=1f=1 by replacing L\mathcal{L} with Lf\mathcal{L}^{f}, so every large power of L\mathcal{L} is globally generated. Take relative prime number pp and qq such that ϕp\phi_{p} and ϕq\phi_{q} satisfies Lemma 84, i.e Ykp=YpY_{kp}=Y_{p}, Ykq=YqY_{kq}=Y_{q} and ϕkp=ϕp\phi_{kp}=\phi_{p}, ϕkq=ϕq\phi_{kq}=\phi_{q} for k1k\geq 1. Then Yp=Ypq=YqY_{p}=Y_{pq}=Y_{q} and ϕp=ϕq\phi_{p}=\phi_{q}, denoting the morphism to be ϕ:XY=Ypq\phi:X\to Y=Y_{pq}.

There are line bundles Mp\mathcal{M}_{p}, Mq\mathcal{M}_{q} on YY such that ϕMp=Lp\phi^{*}\mathcal{M}_{p}=\mathcal{L}^{p} and ϕMq=Lq\phi^{*}\mathcal{M}_{q}=\mathcal{L}^{q}. Since pp, qq are relatively prime, one may take M\mathcal{M} such that M=MprMqs\mathcal{M}=\mathcal{M}_{p}^{r}\otimes \mathcal{M}_{q}^{s} where rp+sq=1rp+sq=1. Since ϕ:PicYPicX\phi^{*}:\mathop{\mathrm{Pic}}Y \to \mathop{\mathrm{Pic}}X is injective, Mp=Mp\mathcal{M}^{p}=\mathcal{M}_{p} and M\mathcal{M} is ample.

Now we show Ym=YY_{m}=Y and ϕm=ϕ\phi_{m}=\phi for large mm. Write m=cp+dqm=cp+dq for c,d1c,d\geq 1, this can be done when m0m\gg 0. Then ScH0(Y,Mp)SdH0(Y,Mq)S^{c}H^{0}(Y,\mathcal{M}^{p})\otimes S^{d}H^{0}(Y,\mathcal{M}^{q}) determines a globally generated subspace of H0(Y,Mcp+dq)=H0(X,Lcp+dq)H^{0}(Y,\mathcal{M}^{cp+dq})=H^{0}(X,\mathcal{L}^{cp+dq}). Then by Lemma 84 again, ϕ\phi factor through ϕcp+dq\phi_{cp+dq} and a finite morphism. Since ϕ\phi is a fibre space, this implies ϕ=ϕm\phi=\phi_{m}. ◻

Lemma 85. *Let L\mathcal{L} be a line bundle generated by global sections on a normal projective variety XX. For any coherent sheaf F\mathcal{F} on XX, there exists a number m0>0m_{0}>0 such that

H0(FLa)H0(Lb)H0(FLa+b)H^{0}(\mathcal{F}\otimes \mathcal{L}^{a})\otimes H^{0}(\mathcal{L}^{b})\to H^{0}(\mathcal{F}\otimes \mathcal{L}^{a+b})

is surjective for a,bm0a,b\geq m_{0}.*

Proof

Proof. Consider the fibre space ϕ:XY\phi:X\to Y given by L\mathcal{L} and L=ϕM\mathcal{L}=\phi^{*}\mathcal{M} for some ample line bundle M\mathcal{M} on YY. Then LHS=H0(Y,fFMa+b)H0(Y,Mb)LHS=H^{0}(Y,f_{*}\mathcal{F}\otimes \mathcal{M}^{a+b})\otimes H^{0}(Y,\mathcal{M}^{b}) clearly surjects to H0(Y,fFMa+b)H^{0}(Y,f_{*}\mathcal{F}\otimes \mathcal{M}^{a+b}). ◻

Theorem 86. Let XX be a normal projective variety and L\mathcal{L} be a semiample line bundle on XX. Then L\mathcal{L} is finitely generated.

Proof

Proof. Suppose Lk\mathcal{L}^{k} is generated by global sections. Then previous lemma shows that R(Lk)R(\mathcal{L}^{k}) is finitely generated. Apply the previous lemma to F=L,,Lk1\mathcal{F}=\mathcal{L},\dots,\mathcal{L}^{k-1} we get the desired result. ◻

More generally, we want to sudy the line bundle which is not globally generate, i.e. whose associated rational map ϕk:XYk\phi_{k}:X\dashrightarrow Y_{k} is not a morphism. Assume XX is a normal projective variety.

Theorem 87 (Iitaka Fibration). Let L\mathcal{L} be a line bundle with κ(L)>0\kappa(\mathcal{L})>0. Then for all sufficiently large kN(L)k\in N(\mathcal{L}), the rational maps ϕk:XYk\phi_{k}:X\dashrightarrow Y_{k} are birationally equivalent to a fixed fibre space ϕ:XY\phi_{\infty}:X_{\infty}\to Y_{\infty} of normal varieties. The restriction to a very general fibre of ϕ\phi_{\infty} has Iitaka dimension 0. More specifically, for large kN(L)k\in N(\mathcal{L}), there exists a commutative diagram

XX1YkY1Áku1Á1½k

of rational maps and morphisms. One has dimY=κ(L)\dim Y_{\infty}=\kappa(\mathcal{L}). Moreover, if we set L=uL\mathcal{L}_{\infty}=u_{\infty}^{*}\mathcal{L}, take FXF\subseteq X_{\infty} be a very general fibre of ϕ\phi_{\infty}, then κ(LF)=0\kappa(\mathcal{L}_{\infty}|_{F})=0.

Proof

Proof. Fix mN(L)m\in N(\mathcal{L}) such that dimYm=κ(L)\dim Y_{m}=\kappa(\mathcal{L}). We first check that for k0k\gg 0, ϕkm:XYkm\phi_{km}:X\dashrightarrow Y_{km} are birational to a fixed fibre space ψ(m):X(m)Y(m)\psi_{(m)}:X_{(m)}\to Y_{(m)} of normal varieties. Let um:X(m)Xu_{m}:X_{(m)}\to X be a resolution of indeterminacies of ϕm\phi_{m}, i.e. a birational morphism with umLm=Mm+Fmu_{m}^{*}|\mathcal{L}^{m}=|M_{m}|+F_{m}, where Mm\mathcal{M}_{m} is globally generated line bundle and FmF_{m} is a fixed divisor in umLm\left| u_{m}^{*}\mathcal{L}^{m} \right| and ψm:X(m)YmP(Hm(X(m),Mm))=P(H0(X,Lm))\psi_{m}:X_{(m)}\to Y_{m}\subseteq \mathbb{P}(H^{m}(X_{(m)},\mathcal{M}_{m}))=\mathbb{P}(H^{0}(X,\mathcal{L}^{m})) is the morphism defined by Mm\left| \mathcal{M}_{m} \right| (Hartshorne II 7.17.3). Consider the morphism ψkm:X(m)Ykm\psi_{km}:X_{(m)}\to Y_{km}' determined by Mmk\mathcal{M}_{m}^{k}, then ψm\psi_{m} factors as X(m)ψkmYkmλkYmX_{(m)}\xrightarrow{\psi_{km}}Y_{km}'\xrightarrow{\lambda_{k}}Y_{m} with λk\lambda_{k} finite. By Lemma 84, for k0k\gg 0, ψkm\psi_{km} stablize to a fixed fibre space ψ(m):X(m)Y(m)\psi_{(m)}:X_{(m)}\to Y_{(m)}. On the other hand, Mmk\left| \mathcal{M}_{m}^{k} \right| is a subsystem of umLkm\left| u_{m}^{*}\mathcal{L}^{km} \right|. Since X(m)X_{(m)} is birational to XX, we can view ψ(m)\psi_{(m)} as a rational map from XX to Y(m)Y_{(m)}. Thus we get a factorization

X(m)»XY(m)YkmÃ(m)Ákm¹k

where μk\mu_{k} is generically finite. Since K(Y(m))K(Y_{(m)}) is algebraically closed in K(X)K(X), μk\mu_{k} is birational. So ϕkm\phi_{km} is birationally equivalent to ψ(m):X(m)Y(m)\psi_{(m)}:X_{(m)}\to Y_{(m)}.

Replacing L\mathcal{L} with Le(L)\mathcal{L}^{e(\mathcal{L})} and we may assume e(L)=1e(\mathcal{L})=1. Fix relative prime integers pp, qq such that dimYp=dimYq=κ(L)\dim Y_{p}=\dim Y_{q}=\kappa(\mathcal{L}). By construction there exists m0m\gg 0 such that Y(p)Y_{(p)} and Y(q)Y_{(q)} are image of X(p)X_{(p)} and X(q)X_{(q)} under the morphism determined by Mppm1\mathcal{M}_{p}^{p^{m-1}} and Mqqm1\mathcal{M}_{q}^{q^{m-1}}. Fix a normal variety XX_{\infty} together with birational morphisms vp:XX(p)v_{p}:X\to X_{(p)} and vq:XX(q)v_{q}:X_{\infty}\to X_{(q)}. such that the diagram commutes:

XX(q)X(p)X1uqupvqvp

write u:XXu_{\infty}:X_{\infty}\to X be the birational morphism. Consider on XX_{\infty} the globally generated line bundle

Mp,q=vpMppm1vqMqpm1.\mathcal{M}_{p,q}=v_{p}^{*}\mathcal{M}_{p}^{p^{m-1}}\otimes v_{q}^{*}\mathcal{M}_{q}^{p^{m-1}}.

Denote YY_{\infty} be the normalization of the image of XX_{\infty} under the morphism defined by Mp,q\mathcal{M}_{p,q}, by ϕ:XY\phi_{\infty}:X_{\infty}\to Y_{\infty} the corresponding morphism. Then YY_{\infty} maps finitely to the Segre embedding of Y(p)×Y(q)Y_{(p)}\times Y_{(q)} and so one has morphisms wp:YY(p)w_{p}:Y_{\infty}\to Y_{(p)} and wq:YY(q)w_{q}:Y_{\infty}\to Y_{(q)} such that the diagram commutes:

X(p)X1X(q)Y(p)Y1Y(q)Ã(p)v(p)v(q)Á1Ã(q)wpwq

Then dimY=κ(L)\dim Y_{\infty}=\kappa(\mathcal{L}) and wpw_{p}, wqw_{q} are generically finite. Since wpw_{p}, wqw_{q} factor through fibre space ψ(p)vp\psi_{(p)}\circ v_{p} and ψ(q)vq\psi_{(q)}\circ v_{q} respectively, they are actually birational. Since wpϕw_{p}\circ \phi_{\infty} is fibre space, so is ϕ\phi_{\infty}. By construction YY_{\infty} carries ample and globally generated line bundle Ap,q\mathcal{A}_{p,q} such that ϕAp,q=Mp,q\phi_{\infty}^{*}\mathcal{A}_{p,q}=\mathcal{M}_{p,q}.

Now fix positive integers c,d1c,d \geq 1. Then one has

H0(X,Mp,q)H0(X,upMpcpm1uqMqdqm1)H0(X,uLcpm+dqm).H^{0}(X_{\infty},\mathcal{M}_{p,q})\subseteq H^{0}(X_{\infty},u_{p}^{*}\mathcal{M}_{p}^{cp^{m-1}}\otimes u_{q}^{*}\mathcal{M}_{q}^{dq^{m-1}})\subseteq H^{0}(X_{\infty},u_{\infty}^{*}\mathcal{L}^{cp^{m}+dq^{m}}).

This give rise to factorization

X1XY1Ycpm+dqmu1Á1Ácpm+dqm¹cpm+dqm

with μcpm+dqm\mu_{cp^{m}+dq^{m}} generically finite. By considering the extension of functional field, μcpm+dqm\mu_{cp^{m}+dq^{m}} is birational. Since every k0k\gg 0 is of the form cpm+dqmcp^{m}+dq^{m}, taking ρk=μk1\rho_{k}=\mu_{k}^{-1} gives the diagram in the theorem.

It remains to show that for very general fibre FF of ϕ\phi_{\infty} one has κ(LF)=0\kappa(\mathcal{L}_{\infty}|_{F})=0. Set L=uL\mathcal{L}_{\infty}=u_{\infty}^{*}\mathcal{L}. Clearly κ(LF)0\kappa(\mathcal{L}_{\infty}|_{F})\geq 0, so we only need to sow κ(LF)0\kappa(\mathcal{L}_{\infty}|_{F})\leq 0. For general yYy\in Y_{\infty}, let F=Fy=ϕ1YF=F_{y}=\phi_{\infty}^{-1}\subseteq Y_{\infty}. Assume ρk\rho_{k} is defined and regular at yy for k0k\gg 0 and u(F)u_{\infty}(F) is not contained in the indeterminacy of ϕk\phi_{k}. Then ϕku\phi_{k}\circ u_{\infty} maps FF to a point and the restriction αk:H0(X,L)kH0(F,LkF)\alpha_{k}:H^{0}(X_{\infty},\mathcal{L}_{\infty})^{k}\to H_{0}(F,\mathcal{L}_{\infty}^{k}|_{F}) has rank one for k0k\gg 0. Fix a very ample line bundle N\mathcal{N} on YY_{\infty}, we asserts there is a large positive integer m0>0m_{0}>0 such that H0(X,Lm0ϕ(N))0H^{0}(X_{\infty},\mathcal{L}_{\infty}^{m_{0}}\otimes \phi_{\infty}^{*}(\mathcal{N}^{*}))\neq 0. Since A=Ap,q\mathcal{A}=\mathcal{A}_{p,q} is ample and globally generated on YY_{\infty}, which pulls back to Mp,q\mathcal{M}_{p,q}. Then Am1N\mathcal{A}^{m_{1}}\otimes \mathcal{N}^{*} has nonzero section for m10m_{1}\gg 0. On the other hand, Mp,qm1\mathcal{M}_{p,q}^{m_{1}} is a subsheaf of L(pm+qm)m1\mathcal{L}_{\infty}^{(p^{m}+q^{m})m_{1}} by construction, we get H0(X,Lm0ϕN)0H^{0}(X_{\infty},\mathcal{L}_{\infty}^{m_{0}}\otimes \phi_{\infty}^{*}\mathcal{N}^{*})\neq 0.

For a fixed kk and any r>0r>0, we have diagram

H0(X1;Lk1¬Á¤1Nr)H0(X1;Lk+rm01)H0(F;(Lk1¬Á¤1Nr)jF)H0(F;Lk+rm01jF)¯k;r®k+rm0

the vertical maps arising via th restriction to FF. For general F=FyF=F_{y}, βk,r\beta_{k,r} is indentified with the map

βk,r:H0(Y,(ϕ)LkNr)(ϕ)LkNrk(y)\beta_{k,r}:H^{0}(Y_{\infty},(\phi_{\infty})_{*}\mathcal{L}_{\infty}^{k}\otimes \mathcal{N}^{r})\to (\phi_{\infty})_{*}\mathcal{L}_{\infty}^{k}\otimes \mathcal{N}^{r}\otimes k(y)

obtained by evaluating sections of direct image at yYy\in Y. But since N\mathcal{N} is ample, for fixed kk, (ϕ)LkNr(\phi_{\infty})_{*}\mathcal{L}_{\infty}^{k}\otimes \mathcal{N}^{r} is globally generated for r0r\gg 0. Thus βk,r\beta_{k,r} is surjective for r0r\gg 0. On the other hand, αk+rm0\alpha_{k+rm_{0}} has rank one. So βk,r\beta_{k,r} also has rank one and hence h0(LkϕNr)=1h^{0}(\mathcal{L}_{\infty}^{k}\otimes \phi_{\infty}^{*}\mathcal{N}^{r})=1. Since ϕN\phi_{\infty}^{*}\mathcal{N} is trivial on FF, h0(F,LkF)=1h^{0}(F,\mathcal{L}_{\infty}^{k}|_{F})=1. ◻

Remark 88. If λ:XW\lambda:X\dashrightarrow W is a rational fibre space of normal varieties and if the restriction of L\mathcal{L} to a very general fibre FF of λ\lambda has Iitaka dimension 0, then λ\lambda factor through Iitaka fibration of L\mathcal{L}.

Corollary 89. *Let L\mathcal{L} be a line bundle on normal projective variety XX and set κ=κ(L)\kappa =\kappa(\mathcal{L}). Then there are constants a,A>0a, A>0 such that

amκh0(X,Lm)Amκa\cdot m^{\kappa}\leq h^{0}(X,\mathcal{L}^{m})\leq A\cdot m^{\kappa}

for all sufficiently large mN(L)m\in N(\mathcal{L}).*

Proof

Proof. We may repalce XX be XX_{\infty} and consider the Iitaka fibration ϕ:XY\phi:X\to Y associate to L\mathcal{L}. Also, we can reduce to the case that e(L)=1e(\mathcal{L})=1. By resolving the singularities, assume XX is smooth. By the proof of Theorem 87 we know there is an ample bundle N\mathcal{N} on YY and a large positive integer m0m_{0} such that H0(X,Lm0ϕ(N))0H^{0}(X_{\infty},\mathcal{L}_{\infty}^{m_{0}}\otimes \phi_{\infty}^{*}(\mathcal{N}^{*}))\neq 0. This implies

h0(X,Llm0)h0(Y,Nl)=lκκ!Yc1(N)κ+O(lκ1).h^{0}(X,\mathcal{L}^{lm_{0}})\geq h^{0}(Y,\mathcal{N}^{l})=\frac{l^{\kappa}}{\kappa!}\int_{Y}c_{1}(\mathcal{N})^{\kappa}+O(l^{\kappa-1}).

Hence for sufficiently large mm0Nm\in m_{0}\mathbb{N} we get b>0b>0 such that

bmκh0(X,Lm).bm^{\kappa}\leq h^{0}(X,\mathcal{L}^{m}).

For general mm, assume km0m<(k+1)m0km_{0}\leq m<(k+1)m_{0}. Since h0(Lm)h^{0}(\mathcal{L}^{m}) is non-decreasing in mm, we have

b(mm0)κh0(Lm).b(m-m_{0})^{\kappa}\leq h^{0}(\mathcal{L}^{m}).

for large mm, we can assume mm0>m2m-m_{0}>\frac{m}{2}. So a=b2κa=\frac{b}{2^{\kappa}} satisfies the requirement. We get the lower bound for h0(Lm)h^{0}(\mathcal{L}^{m}).

For the converse, take DD tobe the effective divisor in L\left| \mathcal{L} \right| and set D=D1+D2D=D_{1}+D_{2}, where D1D_{1} consists of those components of DD that maps to a proper subvariety of YY via ϕ\phi, and D2D_{2} maps onto YY. Thus every fibre FF of ϕ\phi meets D2D_{2} but general fibre is disjoint from D1D_{1}. We claim that

H0(OX(mD1))=H0(OX(mD)).H^{0}(\mathcal{O}_{X}(mD_{1}))=H^{0}(\mathcal{O}_{X}(mD)).

In fact, it is equivalent to show that every divisor in the linear system OX(mD)\left| \mathcal{O}_{X}(mD) \right| contains a base divisor mD2mD_{2}. Suppose on the contrary, then DF=D2FD|_{F}=D_{2}|_{F} and there are at least two linearly independent global sections in H0(F,OF(mD2))H^{0}(F,\mathcal{O}_{F}(mD_{2})) (one is 1 and the other defines mD2mD_{2}), then h0(F,OF(mD))=h0(F,OF(mD2))2h^{0}(F,\mathcal{O}_{F}(mD))=h^{0}(F,\mathcal{O}_{F}(mD_{2}))\geq 2. But this contradicts to the fact that κ(LF)=0\kappa(\mathcal{L}|_{F})=0 for general FF.

We can find ample divisor HH such that HH contains schematic image of D1D_{1}. Then one have

h0(X,OX(mD))=h0(X,OX(mD1))h0(X,ϕOY(mH))=h0(Y,OY(mH)).h^{0}(X,\mathcal{O}_{X}(mD))=h^{0}(X,\mathcal{O}_{X}(mD_{1}))\leq h^{0}(X,\phi^{*}\mathcal{O}_{Y}(mH))=h^{0}(Y,\mathcal{O}_{Y}(mH)).

The result follows from Theorem 17 on YY. ◻

Theorem 90. Let L\mathcal{L} be a line bundle on a normal projective variety XX such that κ(L)0\kappa(\mathcal{L})\geq 0. Then κ(L)=dimR(L)1\kappa(\mathcal{L})=\dim R(\mathcal{L})-1.

Proof

Proof. We may replace L\mathcal{L} by Lme(L)\mathcal{L}^{me(\mathcal{L})} with mm large enough and assume L\mathcal{L} has global sections. Pick any nontrivial section sH0(L)s\in H^{0}(\mathcal{L}), since Lm\mathcal{L}^{m} is a subsheaf of K(X)K(X), tt/smt\mapsto t/s^{m} gives an inclusion H0(Lm)K(X)H^{0}(\mathcal{L}^{m})\to K(X). Therefore one can embed the fractional field FracR(L)\mathop{\mathrm{Frac}}R(\mathcal{L}) into K(X)(s~)K(X)(\tilde{s}), where s~\tilde{s} is the image of ss. Thus R(L)R(\mathcal{L}) is finite dimensional.

Assume FracR(L)\mathop{\mathrm{Frac}}R(\mathcal{L}) is generated by sections siH0(Lmi)s_{i}\in H^{0}(\mathcal{L}^{m_{i}}), take nn to be the least common multiple of mim_{i}s, then R(L)R(\mathcal{L}) is finite over the subring SnS^{n} generated by H0(Ln)H^{0}(\mathcal{L}^{n}). So dimR(L)\dim R(\mathcal{L}) is equal to maixmal dimension of subring generated by H0(Lm)H^{0}(\mathcal{L}^{m}). Consider the Veronese subring R(m)=kH0(Lmk)R^{(m)}=\oplus_{k}H^{0}(\mathcal{L}^{mk}), then the ideal sheaf defining the schematic image of ϕm\phi_{m} is Im=ker(SymH0(Lm)R(m))\mathcal{I}_{m}=\mathop{\mathrm{ker}}(\mathop{\mathrm{Sym}}H^{0}(\mathcal{L}^{m})\to R^{(m)}). Then SymH0(Lm)/ImSm\mathop{\mathrm{Sym}}H^{0}(\mathcal{L}^{m})/\mathcal{I}_{m}\cong S^{m} and therefore dimimϕm(X)=dimSm1\dim \overline{\mathop{\mathrm{im}}\phi_{m}(X)}=\dim S^{m}-1 and κ(L)=dimR(L)1\kappa(\mathcal{L})=\dim R(\mathcal{L})-1. ◻

Corollary 91. Kodaira dimension of a smooth projective variety is birational invariant (in the category of smooth varieties).

Proof

Proof. An easy way is using Weak Factorization Theorem. We omit it here. ◻

Big Line Bundles and Divisors

Definition 92 (Big Line Bundles). A line bundle L\mathcal{L} on projective variety XX is big if κ(L)=dimX\kappa(\mathcal{L})=\dim X.

Example 93. Let f:XYf:X\to Y be a generically finite morphism of normal projective varieties. L\mathcal{L} is an ample bundle on YY. Then fLf^{*}\mathcal{L} is a big bundle.

Restricting L\mathcal{L} to an open set UU is also ample. Shrink UU if necessary to make ff1(U)f|_{f^{-1}(U)} finite, then fLUf^{*}\mathcal{L}|_{U} is ample so

κ(fLU)=dimf1(U)=dimX.\kappa(f^{*}\mathcal{L}|_{U})=\dim f^{-1}(U)=\dim X.

Lemma 94. Assume XX is projective variety of dimension nn (unnecessarily normal). A divisor DD on XX is big if and only if there exists C>0C>0 such that h0(OX(mD))Cmnh^{0}(\mathcal{O}_{X}(mD))\geq C\cdot m^{n} for all large mN(D)m\in N(D).

Proof

Proof. Pass to its normalization. ◻

Proposition 95 (Kodaira lemma). Let DD be a big divisor and EE an arbitary effective divisor on XX. Then H0(OX(mDE))0H^{0}(\mathcal{O}_{X}(mD-E))\neq 0 for all large mN(D)m\in N(D).

Proof

Proof. Assume dimX=n\dim X=n. Consider the exact sequence

0OX(mDE)OX(mD)OE(mD)00\to \mathcal{O}_{X}(mD-E)\to \mathcal{O}_{X}(mD)\to \mathcal{O}_{E}(mD)\to 0

we have h0(OE(mD))O(mdimE)h^{0}(\mathcal{O}_{E}(mD))\leq O(m^{\dim E}). Since h0(OX(mD))h0(OE(mD))+h0(OX(mDE))h^{0}(\mathcal{O}_{X}(mD))\leq h^{0}(\mathcal{O}_{E}(mD))+h^{0}(\mathcal{O}_{X}(mD-E)), we know that H0(OX(mDE))0H^{0}(\mathcal{O}_{X}(mD-E))\neq 0. ◻

Corollary 96. Let DD be a divisor on a projective variety XX. Then the followings are equivalent.

  1. DD is big.

  2. For any ample Cartier divisor AA on XX, there exists a positive integer m>0m>0 and an effective divisor NN on XX such that mD=linA+NmD=_{lin}A+N.

  3. There exists a ample Cartier divisor AA, a positive integer m>0m>0 and an effective divisor NN on XX such that mD=linA+NmD=_{lin}A+N.

  4. There exists a ample Cartier divisor AA, a positive integer m>0m>0 and an effective divisor NN on XX such that mD=numA+NmD=_{num}A+N.

Proof

Proof. For 1. implies 2. One may take rr large enough such that rArA and (r1)A(r-1)A are both effective. Then apply above proposition there is NN' such that mD=(r+1)A+NmD=(r+1)A+N' for m0m\gg0.

For 4. implies 1. Since mDN=numAmD-N=_{num}A, mDNmD-N is ample, some multiple of mDNmD-N is very ample. So we may assume mD=H+NmD=H+N' where HH is very ample. Thus κ(D)κ(H)=dimX\kappa(D)\geq \kappa(H)=\dim X. So DD is big.

The other implications are obvious. ◻

Corollary 97. If DD is big then e(D)=1e(D)=1, i.e. every sufficiently large mm will make mDmD effective.

Proof

Proof. Consider a very ample divisor HH such that HD=linEH-D=_{lin}E with EE effective. There exist mm such that mD=linH+NmD=_{lin}H+N for some effective divisor NN. Then (m1)D=E+N(m-1)D=E+N is also effective.mDmD and (m1)D(m-1)D are both effective implies e(D)=1e(D)=1. ◻

Corollary 98. Let L\mathcal{L} be a big bundle on XX. Then there exists a proper closed subset XXX\subseteq X such that: if YXY\subseteq X is any subvariety not contained in VV, then LY\mathcal{L}|_{Y} is a big line bundle on XX.

Proof

Proof. Let L=O(D)\mathcal{L}=\mathcal{O}(D) and mD=linH+NmD=_{lin}H+N where HH is very ample and NN is effective. Take VV to be the support of NN. If Y⊈VY\not\subseteq V, then mDY=HY+NYmD|_{Y}=H|_{Y}+N|_{Y} is a sum of ample divisor and a effective divisor. ◻

Definition 99. A Q\mathbb{Q}-divisor DD is big if m>0\exists m>0 such that mDmD is big Cartier divisor.

Theorem 100. Let XX be a projective variety of dimension nn and DD, EE be nef Q\mathbb{Q}-divisors on XX. If Dn>nDn1ED^{n}>nD^{n-1}E, then DED-E is big.

Proof

Proof. We fix an ample divisor HH and replace DD and EE by D+ϵHD+\epsilon H and E+ϵHE+\epsilon H for small ϵ\epsilon. Then Dn>nDn1ED^{n}>nD^{n-1}E still holds. So we may assume DD and EE ample, and therefore very ample.

Choose E1,E2,EE_{1},E_{2},\dots \in \left| E \right| of genenal divisors linearly equivalent to EE. Fix m1m\geq 1, then

H0(OX(m(DE)))=H0(OX(mDEi))H^{0}(\mathcal{O}_{X}(m(D-E)))=H^{0}(\mathcal{O}_{X}(mD-\sum E_{i}))

is the sections vanishing at each EiE_{i} for all ii. So we get an exact sequence

0H0(X,OX(m(DE)))H0(X,OX(mD))i=1mH0(Ei,OEi(mD)).0\to H^{0}(X,\mathcal{O}_{X}(m(D-E)))\to H^{0}(X,\mathcal{O}_{X}(mD))\to \bigoplus_{i=1}^{m}H^{0}(E_{i},\mathcal{O}_{E_{i}}(mD)).

For large mm, h0(Ei,OEi(mD))h^{0}(E_{i},\mathcal{O}_{E_{i}}(mD)) are independent of ii by Hirzebruch Riemann Roch. Apply Theorem 17 on XX and each EiE_{i}, one have

h0(X,OX(mDE))h0(X,OX(mD))i=1mh0(Ei,OEi(mD))=Dnn!mnnDn1En!mn+O(mn1).h^{0}(X,\mathcal{O}_{X}(mD-E))\geq h^{0}(X,\mathcal{O}_{X}(mD))-\sum_{i=1}^{m}h^{0}(E_{i},\mathcal{O}_{E_{i}}(mD))=\frac{D^{n}}{n!}m^{n}-n\frac{D^{n-1}E}{n!}m^{n}+O(m^{n-1}).

Thus DED-E is big. ◻

Theorem 101. Let DD be a nef divisor on a projective variety XX of dimension nn. Then DD is big if and only if Dn>0D^{n}>0.

Proof

Proof. Dn>0D^{n}>0 implies DD os big is a direct corollary from above (E=0E=0). Conversely, we may write mD=linH+NmD=_{lin}H+N where HH is very ample and NN is effective. Since DD is nef, Dn1N0D^{n-1}N\geq 0, so mDnHDn1mD^{n}\geq HD^{n-1}. Note that we can also assume DHD|_{H} big by Corollary 98, then by induction HDn1>0HD^{n-1}>0. ◻

Theorem 102. Let DD be a divisor on projective variety XX. Then DD is nef and big if and only if there exists an effective divisor NN such that D1kND-\frac{1}{k}N is ample for k0k\gg 0.

Definition 103. An R\mathbb{R}-divisor DD is big if it can be written into aiDi\sum a_{i}D_{i} where ai>0a_{i}>0 and DiD_{i} are big Cartier divisors.

Similar to nefness, bigness also only depends on numerical equivalence.

Proposition 104. Let DD be an R\mathbb{R}-divisor on XX. Then DD is big if and only if D=numA+ND=_{num}A+N where AA is ample and NN is effective R\mathbb{R}-divisor.

Proof

Proof. Similar to nefness, first argue for Q\mathbb{Q}-divisors. ◻

Example 105. Let DD be nef and big R\mathbb{R}-divisor, then there exists an effective R\mathbb{R}-divisor NN such that D1kND-\frac{1}{k}N is ample R\mathbb{R}-divisor for large kNk\in \mathbb{N}.

Corollary 106. Let DDivR(X)D\in Div_{\mathbb{R}}(X) be a big divisor, let E1,,EtDivR(X)E_{1},\dots, E_{t}\in Div_{\mathbb{R}}(X). Then D+ϵ1E1++ϵtEtD+\epsilon_{1}E_{1}+\cdots +\epsilon_{t}E_{t} is big for all small real number 0ϵi10\leq \left| \epsilon_{i} \right| \ll 1.

Definition 107. The big cone Big(X)N1(X)RBig(X)\subseteq N^{1}(X)_{\mathbb{R}} is the cone of all big R\mathbb{R}-divisor of XX. The pseudoeffective cone Eff(X)N1(X)R\overline{Eff}(X)\subseteq N^{1}(X)_{\mathbb{R}} is the closure of the cone of effective divisors.

Theorem 108. Big(X)=int(Eff(X))Big(X)=int(\overline{Eff}(X)) and Eff(X)=Big(X)\overline{Eff}(X)=\overline{Big}(X).

Proof

Proof. Big(X)Eff(X)Big(X)\subseteq \overline{Eff}(X) by the previous corollary.

For ηEff(X)\eta\in \overline{Eff}(X) which the limit of sequence of effective divisors ηk\eta_{k}. Fix an ample class αN1(X)R\alpha\in N^{1}(X)_{\mathbb{R}}, then η=limkηk+1kα\eta=\lim\limits_{k\to \infty}\eta_{k}+\frac{1}{k}\alpha and each ηk+1kα\eta_{k}+\frac{1}{k}\alpha for kk large. So ηBig(X)\eta\in \overline{Big}(X).

For any ηint(Eff(X))\eta\in int(\overline{Eff}(X)), there is an ϵ>0\epsilon >0 such that ηϵα=σ\eta-\epsilon \alpha=\sigma, where α\alpha is ample and σ\sigma is pseudoeffective. Note ϵ2α+σ\frac{\epsilon}{2}\alpha+\sigma is effective, we have η\eta big. ◻

Definition 109 (Volume of Line Bundles). Let XX be a projective variety of dimension nn, L\mathcal{L} be an line bundle on XX. The volume of L\mathcal{L} is

vol(L)=lim supmh0(Lm)mn/n!.vol(\mathcal{L})=\limsup_{m\to \infty}\frac{h^{0}(\mathcal{L}^{m})}{m^{n}/n!}.

vol(L)>0vol(\mathcal{L})>0 if and only if L\mathcal{L} is big. If L\mathcal{L} is nef, by Asymptotic Riemann Roch, vol(L)=Xc1(L)nvol(\mathcal{L})=\int_{X}c_{1}(\mathcal{L})^{n}.

Lemma 110. Let L\mathcal{L} be a big line bundle and L\mathcal{L} is very ample divisor on XX and E,EAE, E'\in \left| A \right| are general divisors. Then volE(LE)=volE(LE)vol_{E}(\mathcal{L}|_{E})=vol_{E'}(\mathcal{L}|_{E'}).

Proof

Proof. Similar to the proof of Corollary 98. Omitted. ◻

Lemma 111. Let DD be any divisor on XX. aa is a fixed integer. Then

lim supmh0(OX(mD))mn/n!=lim supkh0(OX(akD))(ak)n/n!.\limsup_{m}\frac{h^{0}(\mathcal{O}_{X}(mD))}{m^{n}/n!}=\limsup_{k}\frac{h^{0}(\mathcal{O}_{X}(akD))}{(ak)^{n}/n!}.

Proof

Proof. It is enough to show for DD big. Set vr=lim supkh0(OX((ak+r)D))(ak+r)n/n!v_{r}=\limsup\limits_{k}\frac{h^{0}(\mathcal{O}_{X}((ak+r)D))}{(ak+r)^{n}/n!}. For fixed r0r_{0}, by the definition of volume, we have vol(D)=max{vr0+1,,vr0+a}vol(D)=\max\{v_{r_{0}+1},\dots, v_{r_{0}+a}\}. So we only need to show v0=vrv_{0}=v_{r} for all r[r0+1,r0+a]r\in [r_{0}+1,r_{0}+a]. Fix r00r_{0}\gg 0 such that h0(OX(rD))0h^{0}(\mathcal{O}_{X}(rD))\neq 0 for rr0r\geq r_{0}, take divisor DrrDD_{r}\in \left| rD \right| and Dr(qar)DD_{r}'\in \left| (qa-r)D \right|. Then

h0(OX(kaD))h0(OX((ka+r)D))h0(OX((k+q)aD)),h^{0}(\mathcal{O}_{X}(kaD))\leq h^{0}(\mathcal{O}_{X}((ka+r)D))\leq h^{0}(\mathcal{O}_{X}((k+q)aD)),

also

lim supkh0(OX((ka+r)D))(ka)n/n!=lim supkh0(OX((ka+r)D))(ka+r)n/n!.\limsup_{k}\frac{h^{0}(\mathcal{O}_{X}((ka+r)D))}{(ka)^{n}/n!}=\limsup_{k}\frac{h^{0}(\mathcal{O}_{X}((ka+r)D))}{(ka+r)^{n}/n!}.

Thus v0=vrv_{0}=v_{r} for r[r0+1,r0+a]r\in [r_{0}+1,r_{0}+a]. ◻

Proposition 112. Let DD be a big divisor on projective variety XX of dimension nn. Then

  1. For aN+a\in \mathbb{N}_{+}, vol(aD)=anvol(D)vol(aD)=a^{n}vol(D);

  2. Fix any divisor NN on XX and ϵ>0\epsilon>0. Then there exists an integer p0p_{0} such that

1pnvol(pDN)vol(pD)<ϵ\frac{1}{p^{n}}\left| vol(pD-N)-vol(pD) \right|<\epsilon

for every p>p0p>p_{0}.

Proof

Proof. 1. is direct corollary of lemma above.

For 2. Let N=linABN=_{lin}A-B with AA, BB effective. Since DD is big, write rDB=linBrD-B=_{lin}B' for some effective divisor BB'. Then pDN=lin(p+r)D(A+B)pD-N=_{lin}(p+r)D-(A+B'). We may replace pDpD by (p+r)D(p+r)D and NN by A+BA+B' to assume NN effective. If NN' is another effective divisor, then vol(pD(N+N))vol(pDN)vol(pD)vol(pD-(N+N'))\leq vol(pD-N)\leq vol(pD). So we can replace NN by N+mNN+mN' and choose mm large, NN' ample to assume NN very ample.

Now by the proof of Theorem 100,

h0(X,OX(m(pDN)))h0(X,OX(mpD))mh0(E,OE(mpD)),h^{0}(X,\mathcal{O}_{X}(m(pD-N)))\geq h^{0}(X,\mathcal{O}_{X}(mpD))-m\cdot h^{0}(E,\mathcal{O}_{E}(mpD)),

thus

volX(pDN)volX(pD)=nvolE(pDE)=volX(pD)pn1nvolE(DE).vol_{X}(pD-N)\geq vol_{X}(pD)=n\cdot vol_{E}(pD|_{E})=vol_{X}(pD)-p^{n-1}n\cdot vol_{E}(D|_{E}).

Lemma 113. There exists a fixed divisor NN having the property that H0(OX(N+P))0H^{0}(\mathcal{O}_{X}(N+P))\neq 0 for every P=num0P=_{num}0.

Proof

Proof. Omitted. ◻

Proposition 114. If D=numDD=_{num}D', then vol(D)=vol(D)vol(D)=vol(D').

Proof

Proof. Assume DD and DD' are big. For any P=num0P=_{num}0, fix any integer p>0p>0 we can find NN such that H0(OX(NpP))0H^{0}(\mathcal{O}_{X}(N-pP))\neq 0. Then we have

h0(OX(p(P+D)N))h0(OX(mpD))h^{0}(\mathcal{O}_{X}(p(P+D)-N))\leq h^{0}(\mathcal{O}_{X}(mpD))

Therefore vol(p(D+P)N)vol(pD)=pnvol(D)vol(p(D+P)-N)\leq vol(pD)=p^{n}vol(D). Since 1pnvol(p(D+P)N)vol(D+P)\frac{1}{p^{n}}vol(p(D+P)-N)\to vol(D+P) as pp\to \infty, we have vol(D+P)vol(D)vol(D+P)\leq vol(D). Also, take P=PP=-P and apply the same argument we will have vol(D+P)=vol(D)vol(D+P)=vol(D). ◻

Proposition 115. Let f:XXf:X'\to X be a birational projective morphism between varieties of dimension nn. For any Q\mathbb{Q}-divisor DD on XX, we have volX(fD)=volX(D)vol_{X'}(f^{*}D)=vol_{X}(D).

Proof

Proof. We only consider Cartier divisor DD. Consider the exact
sequence

0OXfOXE00\to \mathcal{O}_{X}\to f_{*}\mathcal{O}_{X'}\to \mathcal{E}\to 0

where E\mathcal{E} is supported on a scheme of dimension n1\leq n-1. Then

h0(X,OX(mD))h0(X,OX(mfD))h0(X,OX(mD))+h0(E(mD))h0(X,OX(mD))+O(mn1)\begin{aligned} h^{0}(X,\mathcal{O}_{X}(mD))&\leq h^{0}(X',\mathcal{O}_{X'}(mf^{*}D))\\&\leq h^{0}(X,\mathcal{O}_{X}(mD))+h^{0}(\mathcal{E}(mD))\\&\leq h^{0}(X,\mathcal{O}_{X}(mD))+O(m^{n-1})\end{aligned}

by projection formula. Thus volX(fD)=vol(D)vol_{X'}(f^{*}D)=vol(D). ◻

Theorem 116 (Continuity of Volume). Let XX be a projective variety of dimension nn. Let ||\cdot|| be the usual norm on N1(X)RN^{1}(X)_{\mathbb{R}}. Then there exists a positive number C>0C>0 such that for all a,bN1(X)Qa,b\in N^{1}(X)_{\mathbb{Q}},

vol(a)vol(b)C(max(a,b))n1ab.\left| vol(a)-vol(b) \right| \leq C\cdot (\max(||a||, ||b||))^{n-1}\cdot ||a-b||.

There is a natural way to define volume on N1(X)RN^{1}(X)_{\mathbb{R}} by the continuity of volume.

Kodaira Vanishing Theorem

We begin with some theorems from complex geometry. We will assume GAGA.

Theorem 117 (Lefschetz hyperplane). Let XX be a smooth complex projective variety of dimension nn, DD is an effective divisor on XX. Then Hi(X,D,Z)=0H^{i}(X,D,\mathbb{Z})=0 for i<ni<n (in usual topology).

Theorem 118 (Hard Lefschetz). For any Kahler form ω\omega on XX, the kk-fold iterate Lk:Hnk(X,C)Hn+k(X,C)L^{k}:H^{n-k}(X,\mathbb{C})\to H^{n+k}(X,\mathbb{C}) of L=LωL=L_{\omega} is an isomorphism.

Corollary 119. Let XX be a smooth projective variety of dimension nn, DD is smooth effective ample divisor. Let rp,q:Hq(X,ΩXp)Hq(D,ΩDp)r_{p,q}:H^{q}(X,\Omega_{X}^{p})\to H^{q}(D,\Omega_{D}^{p}) be the restriction. Then rp,qr_{p,q} is bijection if p+qn2p+q\leq n-2 and is injection if p+q=n1p+q=n-1.

Definition 120 (Simple Normal Crossings). An effective divisor DD is simple normal corssing if it is locally defined by n1n-1 coordinate sections.

Example 121. Let XX be a smooth projective variety and D=aiDiD=\sum a_{i}D_{i} is a snc Q\mathbb{Q}-divisor. Assume EXE\subseteq X is a smooth very ample divisor and E+DE+D is snc. Then [D]E=[DE][D]|_{E}=[D|_{E}].

To see so, it suffices to show for a prime divisor DD such that D+ED+E is nsc, then DED|_{E} is reduced. This is local so assume we are working in affine ambient space. Then by definition the defining section x1x_{1}, x2x_{2} of DD, EE is a regular sequence. DED\cap E is smooth and the local ring is regular, and thus a domain.

Theorem 122 (Resolution of Singularities). Let XX be a variety and DD be an effective Cartier divisor on XX.

  1. There is a projective birational morphism μ:XX\mu:X'\to X where XX is nonsingular and μ\mu has exceptional locus except(μ)except(\mu) and μD+except(μ)\mu^{*}D+except(\mu) is a snc.

  2. One can construct XX' via sequence of blow-ups along smooth centers supported in singular locus of DD and XX. In particular, one can assume μ\mu is an isomorphism on X(Sing(X)Sing(D))X-(\mathop{\mathrm{Sing}}(X)\cup \mathop{\mathrm{Sing}}(D)).

Definition 123. A log resolution of linear system VH0(X,OX(L))\left| V \right| \subseteq H^{0}(X,\mathcal{O}_{X}(L)) is a projective birational morphism μ:XX\mu:X'\to X such that XX' is nonsingular and

μV=W+F\mu^{*}\left| V \right| =\left| W \right| +F

where F+except(μ)F+except(\mu) is snc and WH0(X,OX(μLF))\left| W \right| \subseteq H^{0}(X',\mathcal{O}_{X'}(\mu^{*}L-F)) is base point free.

Remark 124. A log resolution of V\left| V \right| is the same as log resolution of Bs(V)Bs(\left| V \right| ).

As the resolution can be constructed using blow-ups, we have the followings:

Corollary 125. Assume XX is smooth projective variety and μ:XX\mu:X'\to X is a resolution of effective divisor DXD\subseteq X.

  1. Let KXK_{X}, KXK_{X'} be the canonical divisors, then μOX(KX)=OX(KX)\mu_{*}\mathcal{O}_{X'}(K_{X'})=\mathcal{O}_{X}(K_{X}) and RjμOX(KX)=0R^{j}\mu_{*}\mathcal{O}_{X'}(K_{X'})=0 for j>0j>0;

  2. Assume HH is ample on XX. Then for some integer pp sufficiently large there exist bj0b_{j}\geq 0, μ(pH)bjEj\mu^{*}(pH)-\sum b_{j}E_{j} is ample on XX. Here EjE_{j} is the irreducible components of except(μ)except(\mu).

Proof

Proof. By induction it suffices to do it for single blow up f:XXf:X'\to X with smooth center. For 1. Let EE be the exceptional divisor and mm is the codimension of blow up center ZZ. Then

KX=fKX+(m1)E.K_{X'}=f^{*}K_{X}+(m-1)E.

Take higher direct image to the exact sequence

0OXOX(E)OE(E)0,0\to \mathcal{O}_{X'}\to \mathcal{O}_{X'}(E)\to \mathcal{O}_{E}(E)\to 0,

we get

0fOXfOX(E)fOE(E)R1fOX0\to f_{*}\mathcal{O}_{X'}\to f_{*}\mathcal{O}_{X'}(E)\to f_{*}\mathcal{O}_{E}(E)\to R^{1}f_{*}\mathcal{O}_{X'}\to \cdots

where fOX=OXf_{*}\mathcal{O}_{X'}=\mathcal{O}_{X}. We consider the diagram

X0XEZfifjEj

where ii and jj are closed immersions so RpiOE(E)=0R^{p}i_{*}\mathcal{O}_{E}(E)=0. By Grothendieck spectral sequence

E2p,q=Rpf(RqiOE(kE))Rp+q(fi)OE(kE),E_{2}^{p,q}=R^{p}f_{*}(R^{q}i_{*}\mathcal{O}_{E}(kE))\Rightarrow R^{p+q}(f\circ i)_{*}\mathcal{O}_{E}(kE),

so

Rpf(iOE(kE))=Rp(fi)OE(kE)=Rp(jfE)OE(kE).R^{p}f_{*}(i_{*}\mathcal{O}_{E}(kE))=R^{p}(f\circ i)_{*}\mathcal{O}_{E}(kE)=R^{p}(j\circ f|_{E})_{*}\mathcal{O}_{E}(kE).

Similarly, we can derive jRp(fE)OE(kE)=Rp(jfE)OE(kE)j_{*}R^{p}(f|_{E})_{*}\mathcal{O}_{E}(kE)=R^{p}(j\circ f|_{E})_{*}\mathcal{O}_{E}(kE),
but OE(kE)=OP(NZ/X)(k)\mathcal{O}_{E}(kE)=\mathcal{O}_{\mathbb{P}(\mathcal{N}_{Z /X})}(-k) so Rp(fE)OE(kE)=0R^{p}(f|_{E})_{*}\mathcal{O}_{E}(kE)=0 for p0p\geq 0.

Thus RpfOXRpfOX(E)R^{p}f_{*}\mathcal{O}_{X'}\cong R^{p}f_{*}\mathcal{O}_{X'}(E) for i0i\geq 0. Apply the same argument to the exact sequence

0OX((n1)E)OX(nE)OE(nE)0,nm1,0\to \mathcal{O}_{X'}((n-1)E)\to \mathcal{O}_{X'}(nE)\to \mathcal{O}_{E}(nE)\to 0, \quad n\leq m-1,

we conclude that RpfOXRpfOX(nE)R^{p}f_{*}\mathcal{O}_{X'}\cong R^{p}f_{*}\mathcal{O}_{X'}(nE). RpfOX=0R^{p}f_{*}\mathcal{O}_{X'}=0 for p>0p>0 by Theorem of formal functions (similar to Hartshorne V 3.4) The remaining part is an easy application of projection formula.

For 2. Consider a very ample divisor PP on XX'. Write P=aiDi+bEP=\sum a_{i}D_{i}+bE where DiD_{i} are divisors distinct from DiD_{i}. Then f(Di)f(D_{i}) are divisors on XX. Consider Q=Pfaif(Di)Q=P-f^{*}\sum a_{i}f(D_{i}), clearly QQ is supported on the exceptional locus. Take any curve CC contained in a general fibre FF in the exceptional divisor EE. We can choose CC general such that CC does not intersect with any divisor DiD_{i}. So DiC=0D_{i}\cdot C=0 but HC>0H\cdot C>0, so bEC=0bE\cdot C=0. Since EC=degCOP(NZ/X)(1)<0E\cdot C=\deg_{C}\mathcal{O}_{\mathbb{P}(\mathcal{N}_{Z /X})}(-1)<0, b<0b<0. Then we can choose PP large such that pHaif(Di)pH-\sum a_{i}f(D_{i}) ample so fpHaif(Di)f^{*}pH-\sum a_{i}f(D_{i}) is nef. Thus fpHaif(Di)+Hf^{*}pH-\sum a_{i}f(D_{i})+H is ample and satisfies the requirements. ◻

Proposition 126 (Cyclic Covering). Let XX be a variety and L\mathcal{L} is a line bundle on XX. For an integer m1m\geq 1, the section sH0(Lm)s\in H^{0}(\mathcal{L}^{m}) defines DXD\subseteq X. Then there exists a finite flat morphism π:YX\pi:Y\to X such that there is a section sH0(πL)s'\in H^{0}(\pi^{*}\mathcal{L}) with (s)m=πs(s')^{m}=\pi^{*}s. The divisor D=div(s)D=div(s') maps isomorphically to DD. If DD is snc and XX is smooth, then the singularity of YY lies in the singularity of DD. In particular, if XX, DD is smooth, then we may take YY and DD' so.

Proof

Proof. We only give construction of π:YX\pi:Y\to X, the others are local computations. Consider the bundle map p:LXp:L\to X where L=SpecSym(L)L=\mathop{\mathrm{Spec}}\mathop{\mathrm{Sym}}(\mathcal{L}^{*}). Let TT be the tautological section of pLp^{*}\mathcal{L}. Then we take YLY\subseteq L be to the divisor corresponding to the section TmpsH0(pLm)T^{m}-p^{*}s\in H^{0}(p^{*}\mathcal{L}^{m}). ◻

Corollary 127. In the same setting as above construction,

πOY=OXL1L1m.\pi_{*}\mathcal{O}_{Y}=\mathcal{O}_{X}\oplus \mathcal{L}^{-1}\oplus \mathcal{L}^{1-m}.

Proof

Proof. By the definition of LL, πOL=i=10Li\pi_{*}\mathcal{O}_{L}=\oplus_{i=10}^{\infty} \mathcal{L}^{-i}. In fact, TmpsH0(L,pLm)T^{m}-p^{*}s\in H^{0}(L,p^{*}\mathcal{L}^{m}) implies p(Tmps)=ppLm=i=mLip_{*}(T^{m}-p^{*}s)=p_{*}p^{*}\mathcal{L}^{-m}=\oplus_{i=m}^{\infty}\mathcal{L}^{-i}. Thus πOY=OXL1L1m\pi_{*}\mathcal{O}_{Y}=\mathcal{O}_{X}\oplus \mathcal{L}^{-1}\oplus \mathcal{L}^{1-m}. ◻

Remark 128. If DD is smooth and Di\sum D_{i} is reduced effective divisor on XX such that D+DiD+\sum D_{i} is snc. Then D+πDiD'+\sum \pi^{*}D_{i} is also snc.

Theorem 129 (Kawamata Covering). Let XX be smooth projective variety and D=i=1tDiD=\sum\limits_{i=1}^{t}D_{i} be snc on XX. Fix integers m1,,mt>0m_{1}, \dots, m_{t}>0, then there exists a smooth variety YY and a finite flat covering f:YXf:Y\to X such that fD=miDif^{*}D=m_{i}D_{i}' for some smooth DiD_{i}' on YY. D=i=1tDiD'=\sum\limits_{i=1}^{t}D_{i} is also snc.

Proof

Proof. By induction on the components, we may assume m2=mt=1m_{2}=\cdots m_{t}=1. To get the smoothness, take an ample divisor HH such that m1HD1m_{1}H-D_{1} is globally generated. Assume dimX=n\dim X=n, consider H1,,Hn+1m1HD1H_{1},\dots,H_{n+1}\in \left| m_{1}H-D_{1} \right|, by Bertini Theorem, we may assume D+i=1n+1HiD+\sum\limits_{i=1}^{n+1}H_{i} is snc. We shall construct sequence of cyclic coverings

Yn+1fn+1Y1f1XY_{n+1}\xrightarrow{f_{n+1}}\cdots Y_{1}\xrightarrow{f_{1}}X

inductively. Assume Yi1Y_{i-1} has been constructed, denote πi:YiX\pi_{i}:Y_{i}\to X the composition of morphisms. Let fif_{i} be the m1m_{1}-fold cover of YiY_{i} with center πi1(Hi+D)=πi1Hi+m(πi1D1)red\pi_{i-1}^{*}(H_{i}+D)=\pi_{i-1}^{*}H_{i}+m(\pi_{i-1}^{*}D_{1})_{red}, which is, by our construction, same as m1m_{1}-fold branched at pii1Hipi_{i-1}^{*}H_{i} (if locally πi1Hi=div(h)\pi_{i-1}^{*}H_{i}=div(h) and πi1D1=div(sm)\pi_{i-1}^{*}D_{1}=div(s^{m}), then the relation tm=smht^{m}=s^{m}h is equivalent to (t/s)m=h(t/s)^{m}=h). Thus YiY_{i} is smooth outside the singular locus of HiHi1H1DH_{i}\cap H_{i-1}\cap \cdots\cap H_{1}\cap D. So Y=Yn+1Y=Y_{n+1} is smooth. ◻

Theorem 130 (Bloch-Gieseker Covering). Let XX be a smooth quasiprojective variety, let M\mathcal{M} be a line bundle on XX, and fix a positive integer mm. Then there exist a smooth variety YY , a finite surjective morphism f:YXf:Y\to X, and a line bundle L\mathcal{L} on YY such that fM=Lmf_{*}\mathcal{M} = \mathcal{L}^{m}. Further, given a snc divisor DD on XX, we can arrange that its pullback fDf ^{*}D is a snc on YY.

Proof

Proof. We may write M\mathcal{M} be the difference of two globally generated sheaves to assume M\mathcal{M} is globally generated. Then the theorem is a direct corollary from Theorem 129. ◻

Lemma 131 (Injectivity). Let f:YXf:Y\to X be a finite surjective morphism of projective varieties. Assume XX is normal and E\mathcal{E} is a vector bundle on XX. THen the natrual homomorphism Hj(X,E)Hj(Y,fE)H^{j}(X,\mathcal{E})\to H^{j}(Y,f^{*}\mathcal{E}) is injective.

Proof

Proof. By passing to normalization of YY we may assume YY is normal. By projection formula, Hj(Y,fE)=Hj(X,fOYE)H^{j}(Y,f^{*}\mathcal{E})=H^{j}(X,f_{*}\mathcal{O}_{Y}\otimes \mathcal{E}). Consider the trace map TrK(Y)/K(X)Tr_{K(Y)/K(X)}, it gives rise to a trace map TrY/X:fOYOXTr_{Y/X}:f_{*}\mathcal{O}_{Y}\to \mathcal{O}_{X}, which splits the natural inclusion OXfOY\mathcal{O}_{X}\to f_{*}\mathcal{O}_{Y} by normality of fOYf_{*}\mathcal{O}_{Y}. Therefore the map EEfOY\mathcal{E}\to \mathcal{E}\otimes f_{*}\mathcal{O}_{Y} also splits, and thus Hj(E)H^{j}(\mathcal{E}) embeds into Hj(X,fOYE)H^{j}(X,f_{*}\mathcal{O}_{Y}\otimes \mathcal{E}). ◻

Theorem 132 (Kodaira Vanishing). Let XX be a smooth projective variety of dimension nn. Let AA be an ample divisor on XX. Then Hi(OX(KX+A))=0H^{i}(\mathcal{O}_{X}(K_{X}+A))=0 for i>0i>0, or by Serre duality, Hj(OX(A))=0H^{j}(\mathcal{O}_{X}(-A))=0 for j<nj<n.

Proof

Proof. Since AA is ample, we can choose smooth DmAD\in \left| mA \right| for m0m\gg 0. Let p:YXp:Y\to X be the mm-fold cyclic covering branched along DD. Then there is a smooth ample divisor DfOX(A)D'\in f^{*}\mathcal{O}_{X}(A) such that mD=fDmD'= f^{*}D. By Lemma 131, it suffices to show Hj(Y,OY(D))=0H^{j}(Y,\mathcal{O}_{Y}(-D'))=0 for j<nj<n. By the Hodge decomposition we have Hj(X,OX)Hj(D,OD)H^{j}(X,\mathcal{O}_{X})\to H^{j}(D,\mathcal{O}_{D}) is isomorphism for jn2j\leq n-2 and is injective if j=n1j=n-1. But then taking cohomology to the exact sequence

0OY(D)OYOD00\to \mathcal{O}_{Y}(-D')\to \mathcal{O}_{Y}\to \mathcal{O}_{D'}\to 0

shows that Hj(Y,O(D))=0H^{j}(Y,\mathcal{O}(-D'))=0 when jn1j\leq n-1. ◻

Definition 133. Suppose XX is smooth variety and DXD\subseteq X is a smooth divisor (or snc). ΩX1(logD)\Omega_{X}^{1}(\log D) is the sheaf of 1-forms on XX with logiarithmic poles along DD, i.e. if z1,,znz_{1},\dots, z_{n} are local coordinates on XX, with DD defined by zn=0z_{n}=0, then Ω1(logD)\Omega^{1}(\log D) is locally generated by dz1,,dzn1,dznzndz_{1},\dots, dz_{n-1},\frac{dz_{n}}{z_{n}}. Denote ΩXp(logD)=p(ΩX1(logD))\Omega_{X}^{p}(\log D)=\wedge^{p}(\Omega_{X}^{1}(\log D)).

Lemma 134. Assume DXD\subseteq X is a smooth divisor.

We have exact sequence

0ΩXpΩXp(logD)resΩDp100\to \Omega_{X}^{p}\to \Omega_{X}^{p}(\log D)\xrightarrow{res}\Omega_{D}^{p-1}\to 0

and

0ΩXp(logD)OX(D)ΩXpΩDp0.0\to \Omega_{X}^{p}(\log D)\otimes \mathcal{O}_{X}(-D)\to \Omega_{X}^{p}\to \Omega_{D}^{p}\to 0.

π:YX\pi:Y\to X be mm-fold cyclic covering branched along DD and DYD'\subseteq Y is the divisor such that πD=mD\pi^{*}D=mD', then π(ΩXp(logD))=ΩYp(logD)\pi^{*}(\Omega_{X}^{p}(\log D))=\Omega_{Y}^{p}(\log D'). Here the residue map resres is defined as follows: For any section φ\varphi in (ΩXp(logD))(\Omega_{X}^{p}(\log D)), we can write φ=φ1+φ2dznzn\varphi=\varphi_{1}+\varphi_{2}\wedge \frac{d z_{n}}{z_{n}} where φ1\varphi_{1} consists of pp-froms without dznzn\frac{d z_{n}}{z_{n}} and φ2\varphi_{2} is the p1p-1 forms without dznzn\frac{d z_{n}}{z_{n}}. Then res:φφ2Dres:\varphi\mapsto \varphi_{2}|_{D}.

Proof

Proof. All statements are local so we may assume XX is affine. Easily we can see that “surjective” part of both exact sequence are true. Sections φ\varphi restrict to 0 by the residue map if and only if it is in ΩXp\Omega_{X}^{p}. For sections in ker(ΩXpΩDp)\mathop{\mathrm{ker}}(\Omega_{X}^{p}\to \Omega_{D}^{p}), one easily see that it is of the form znΩXp+dznΩXp1z_{n}\cdot \Omega_{X}^{p}+dz_{n}\wedge\Omega_{X}^{p-1}, which is exactly ΩX(logD)OX(D)\Omega_{X}(\log D)\otimes \mathcal{O}_{X}(-D).

For 2. Assume the local section of DD' is defined by zn+1z_{n+1} with zn+1m=znz_{n+1}^{m}=z_{n}. Then

πdznzn=dzn+1mzn+1m=mdzn+1zn+1,\pi^{*}\frac{d z_{n}}{z_{n}}=\frac{d z_{n+1}^{m}}{z_{n+1}^{m}}=m\frac{dz_{n+1}}{z_{n+1}},

we get the desired isomorphism since the generators are the same. ◻

Theorem 135 (Nakano Vanishing). Let XX be a smooth projective variety of dimension nn. Let AA be an ample divisor on XX. Then

Hq(ΩXp(A))=0 for p+q>n,H^{q}(\Omega_{X}^{p}(A))=0 \text{ for }p+q>n,

or equivalently,

Hs(ΩXr(A))=0 for r+s<n.H^{s}(\Omega_{X}^{r}(-A))=0 \text{ for }r+s<n.

Proof

Proof. Since AA is ample there exists m0m\gg 0 and a smooth divisor DmAD\in \left| mA \right|. Use induction on dimension we may assume vanishing for DD ,i.e.

Hs(D,ΩDr1(A))=0 for r1+s<n1.H^{s}(D,\Omega_{D}^{r-1}(-A))=0 \text{ for }r-1+s<n-1.

Consider the exact sequence

0ΩXr(A)ΩXr(logD)OX(A)ΩDr1(A)0,0\to \Omega_{X}^{r}(-A)\to \Omega_{X}^{r}(\log D)\otimes \mathcal{O}_{X}(-A)\to \Omega_{D}^{r-1}(-A)\to 0,

taking the cohomology, we only need to show Hs(X,ΩXr(logD)O(A))=0H^{s}(X,\Omega_{X}^{r}(\log D)\otimes \mathcal{O}_{(-A)})=0 for r+s<nr+s<n. Consider the mm-fold covering π:YX\pi:Y\to X branched at DD with πD=mD\pi^{*}D=mD'. By Lemma 131, it suffices to show Hs(Y,π(ΩXr(logD)(A)))=Hs(Y,ΩYr(logD)OY(D))=0H^{s}(Y,\pi^{*}(\Omega_{X}^{r}(\log D)\otimes(-A)))=H^{s}(Y,\Omega_{Y}^{r}(\log D')\otimes \mathcal{O}_{Y}(-D'))=0. Take cohomology to the exact sequence

0ΩYr(logD)OY(D)ΩYrΩDr0,0\to \Omega_{Y}^{r}(\log D')\otimes \mathcal{O}_{Y}(-D')\to \Omega_{Y}^{r}\to \Omega_{D'}^{r}\to 0,

since Hs(Y,ΩYr)Hs(D,ΩDr)H^{s}(Y,\Omega_{Y}^{r})\to H^{s}(D',\Omega_{D'}^{r}) is bijection for r+s<nr+s<n and is injection for r+s=nr+s=n, we get the desired result. ◻

Lemma 136. Let XX be smooth projective variety of dimension nn. AA is an ample divisor and EE is snc on XX. Then Hi(OX(KX+A+E))=0H^{i}(\mathcal{O}_{X}(K_{X}+A+E))=0 for i>0i>0 or equivalently, Hj(OX(AE))=0H^{j}(\mathcal{O}_{X}(-A-E))=0 for j<nj<n.

Proof

Proof. Write E=i=1tEiE=\sum\limits_{i=1}^{t}E_{i}. We use induction on tt. Assume for tk1t\leq k-1 is true. Then consider the exact sequence

0OX(Ai=1tEi)OX(Ai=1k1Ei)OEk(Ai=1k1Ei)0.0\to\mathcal{O}_{X}(-A-\sum_{i=1}^{t}E_{i})\to \mathcal{O}_{X}(-A-\sum_{i=1}^{k-1}E_{i})\to \mathcal{O}_{E_{k}}(-A-\sum_{i=1}^{k-1}E_{i})\to 0.

Note the restriction of i=1k1Ei\sum\limits_{i=1}^{k-1}E_{i} to EkE_{k} is still snc, taking the cohomology we get the desired result. ◻

Theorem 137 (Kawamata Viehweg Vanishing). Let XX be a smooth projective variety of dimension nn. DD is nef and big divisor on XX. Then Hi(OX(KX+D))=0H^{i}(\mathcal{O}_{X}(K_{X}+D))=0 for i>0i>0 or equivalently, Hj(OX(D))=0H^{j}(\mathcal{O}_{X}(-D))=0 for j<nj<n.

Proof

Proof. We may write mD=linH+NmD=_{lin}H+N where HH is ample and NN is effective divisor. We first show that we can reduce to the case that NN has snc support. One can consider the log resolution f:XXf:X'\to X for any singular divisor NN. Let fN=aiFif^{*}N=\sum a_{i}F_{i} where FiF_{i} includes all exceptional divisors and ai0a_{i}\geq 0. Then for some p0p\gg 0, there exists bi0b_{i}\geq 0 such that f(pH)biFif^{*}(pH)-\sum b_{i}F_{i} is ample on XX'. Then

f(pmD)=linf(pH)biFi+(pai+bi)Fif^{*}(pmD)=_{lin}f^{*}(pH)-\sum b_{i}F_{i}+\sum(pa_{i}+b_{i})F_{i}

is a sum of ample divisor and an effective divisor with snc support. Since RjfOX(KX+fD)=0R^{j}f_{*}\mathcal{O}_{X'}(K_{X'}+f^{*}D)=0 by projection formula, we know Hj(X,OX(KX+D))=0H^{j}(X,\mathcal{O}_{X}(K_{X}+D))=0 if Hj(X,OX(KX+fD))=0H^{j}(X',\mathcal{O}_{X'}(K_{X'}+f^{*}D))=0 (see Corollary 125). Thus we can reduce to the case that NN has snc support.

Assume N=i=1teiEiN=\sum\limits_{i=1}^{t}e_{i}E_{i} with ei>0e_{i}>0. Set e=e1ete^{*}=e_{1} \cdots e_{t} and ei=eeie_{i}^{*}=\frac{e^{*}}{e_{i}}. Apply Theorem 129, we can construct h:YXh:Y\to X such that hEi=meiEih^{*}E_{i}=me_{i}^{*}E_{i}' for some EiE_{i}'. Sete E=EiE'=\sum E_{i}' , D=hDD'=h^{*}D and H=hHH'=h^{*}H, then mD=linH+meEmD'=_{lin}H'+me^{*}E'. DD' is nef and thus H+m(e1)D=linme(DE)H'+m(e^{*}-1)D'=_{lin}me^{*}(D'-E') is ample. We may write D=linA+ED'=_{lin}A'+E' for some ample divisor AA', so Hj(OY(D))=Hj(OY(AE))=0H^{j}(\mathcal{O}_{Y}(-D'))=H^{j}(\mathcal{O}_{Y}(-A'-E'))=0 by above lemma. By Lemma 131, we get desired results. ◻

Theorem 138. Let DD be a nef divisor on a smooth projective variety XX of dimension nn. Assume DnkHk>0D^{n-k}H^{k}>0 for some kk and ample divisor HH. Then Hi(OX(K+D))=0H^{i}(\mathcal{O}_{X}(K+D))=0 for i>ki>k.

Proof

Proof. We may assume HH is smooth very ample divisor and apply the induction on dimension nn. The base case is from Theorem 137. Take the cohomology to the exact sequence

0OX(KX+D)OX(KX+D+H)OH(KX+D)0.0\to \mathcal{O}_{X}(K_{X}+D)\to \mathcal{O}_{X}(K_{X}+D+H)\to \mathcal{O}_{H}(K_{X}+D)\to 0.

Then Hi(H,OH(KX+D))=0H^{i}(H,\mathcal{O}_{H}(K_{X}+D))=0 for i>k1i>k-1 by induction hypothesis and Hi(X,OX(KX+D+H))=0H^{i}(X,\mathcal{O}_{X}(K_{X}+D+H))=0 by Kodaira vanishing. ◻

Theorem 139 (Kawamata Vanishing Theorem for Q\mathbb{Q}-divisors). Let XX be a smooth projective variety of dimension nn. Let NN be an integral divisor on XX and N=numB+ΔN=_{num}B+\Delta, where BB is a big and nef Q\mathbb{Q}-divisor, Δ=aiΔi\Delta=\sum a_{i}\Delta_{i} is a Q\mathbb{Q}-divisor with snc support and 0ai<10\leq a_{i}<1 (we call divisors with such coefficients fractional). Then Hi(OX(KX+N))=0H^{i}(\mathcal{O}_{X}(K_{X}+N))=0 for all i>0i>0, or equivalently, Hj(OX(N))=0H^{j}(\mathcal{O}_{X}(-N))=0 for all j<nj<n.

Proof

Proof. We apply induction on the number of fractional terms in Δ\Delta. The base case followed from Theorem 137. Assume a1=cd0a_{1}=\frac{c}{d}\neq 0 for some 0<c<d0<c<d. By Lemma 131, we only need to construct a finite moprhism p:XXp:X'\to X such that XX' is smooth and Hj(X,pOX(N))=0H^{j}(X',p^{*}\mathcal{O}_{X}(-N))=0 for j<nj<n. By Theorem 129, we can construct p:XXp:X'\to X such that pΔi\sum p^{*}\Delta_{i} is snc and pΔ1=lindAp^{*}\Delta_{1}=_{lin}dA' for some divisor AA' on XX'. Denote Δi=pΔi\Delta_{i}'=p^{*}\Delta_{i}, N=pNN'=p^{*}N and B=pBB'=p^{*}B. Consider the cyclic covering q:XXq:X''\to X' branched at Δ1\Delta_{1}' and write Δi=qΔi\Delta_{i}''=q^{*}\Delta_{i}', A=qAA''=q^{*}A', B=qBB''=q^{*}B', N=qNN''=q^{*}N'. Then N=B+cA+i2aiΔiN''=B''+cA''+\sum\limits_{i\geq 2}a_{i} \Delta_{i}''. Then $$N’‘-cA’‘={num} B’’ +\sum{i\geq 2}a_{i}\Delta_{i}’'$$ and the number of fractional part is reduced. By induction hypothesis, we have Hj(X,OX(cAN))=Hj(X,qOX(cAN))H^{j}(X'',\mathcal{O}_{X''}(cA''-N''))=H^{j}(X',q_{*}\mathcal{O}_{X''}(cA''-N'')). Since qX=OXOX(A)OX((1d)A)q_{*}\mathcal{X''}=\mathcal{O}_{X'}\oplus\mathcal{O}_{X'}(-A')\oplus \cdots \oplus\mathcal{O}_{X'}(-(1-d)A'),

qOX(cAN)=(qOX)(cAN)=OX(cAN)OX((cd+1)AN),q_{*}\mathcal{O}_{X''}(cA'-N')=(q_{*}\mathcal{O}_{X''})(cA'-N')=\mathcal{O}_{X'}(cA'-N')\oplus \cdots\oplus \mathcal{O}_{X'}((c-d+1)A'-N'),

note cd1c\leq d-1, so OX(N)\mathcal{O}_{X'}(-N') is a direct summands of qOX(cAN)q_{*}\mathcal{O}_{X''}(cA''-N''). Thus Hj(X,OX)(N)=0H^{j}(X',\mathcal{O}_{X'})(-N')=0. ◻

Corollary 140. Let XX be a smooth projective variety and BB be a nef and big Q\mathbb{Q}-divisor whose fractional part is snc. Then Hi(X,OX(KX+B))=0H^{i}(X,\mathcal{O}_{X}(K_{X}+\lceil B\rceil ))=0 for all i>0i>0.

Definition 141. Let f:XYf:X\to Y be a surjective projective morphism of varieties. We say a Q\mathbb{Q}-divisor is ff-nef (resp. ff-big) if the restriction to any fibre of ff is nef (resp. big).

Theorem 142. Let BB be a Q\mathbb{Q}-divisor which is ff-nef and ff-big whose fractional part is snc. Then Hi(X,OX(KX+B))=0H^{i}(X,\mathcal{O}_{X}(K_{X}+\lceil B\rceil))=0 for all i>0i>0.

We also have genenalization to R\mathbb{R}-divisors:

Theorem 143. Let XX be a smooth projective variety and EE be a integral divisor. DD is effective R\mathbb{R}-divisor with snc support. Assume EDE-D is big and nef, then Hi(X,OX(KX+E[D]))=0H^{i}(X,\mathcal{O}_{X}(K_{X}+E-[D]))=0 for all i>0i>0.

Theorem 144 (Kollar Injectivity). Let f:XYf:X\to Y be a surjective morphism of projective varieties. Assume XX is smooth and YY is normal. Consider integral divisors NN and DD on XX such that DD is effective and f(D)Yf(D)\neq Y. Assume N=numfB+ΔN=_{num}f^{*}B+\Delta where BB is big and nef Q\mathbb{Q}-divisor on YY. Then for every i>0i>0 the morphism

Hi(OX(KX+N))DHi(OX(KX+N+D))H^{i}(\mathcal{O}_{X}(K_{X}+N))\xrightarrow{\cdot D}H^{i}(\mathcal{O}_{X}(K_{X}+N+D))

is injective.

Lemma 145. Let f:XYf:X\to Y be morphism of projective varieties. Let AA be ample divisor on YY and F\mathcal{F} is coherent on XX such that Hj(X,FOX()fmA)=0H^{j}(X,\mathcal{F}\otimes \mathcal{O}_{X}()f^{*}mA)=0 for all j>0j>0 and m0m \gg 0. Then RifF=0R^{i}f_{*}\mathcal{F}=0 for all j>0j>0.

Proof

Proof. Take mm large such that Hi(Y,RjfFOY(mA))=0H^{i}(Y,R^{j}f_{*}\mathcal{F}\otimes \mathcal{O}_{Y}(mA))=0 for all i,j>0i,j>0 and RifFOY(mA)R^{i}f_{*}\mathcal{F}\otimes \mathcal{O}_{Y}(mA) are all generated by global sections. Then the Leray spectral sequence gives

E2i,j=Hi(Y,RjfFOY(mA))Hi+j(X,FOX(fmA)).E_{2}^{i,j}=H^{i}(Y,R^{j}f_{*}\mathcal{F}\otimes \mathcal{O}_{Y}(mA))\Rightarrow H^{i+j}(X,\mathcal{F}\otimes \mathcal{O}_{X}(f^{*}mA)).

So

H0(Y,RjfFOY(mA))Hj(X,FOX(fmA)).H^{0}(Y,R^{j}f_{*}\mathcal{F}\otimes \mathcal{O}_{Y}(mA))\cong H^{j}(X,\mathcal{F}\otimes \mathcal{O}_{X}(f^{*}mA)).

If RjfF0R^{j}f_{*}\mathcal{F}\neq 0, then Hj(X,FOX(fmA))=0H^{j}(X,\mathcal{F}\otimes \mathcal{O}_{X}(f^{*}mA))=0 contradicts with assumption. ◻

Theorem 146 (Grauert-Riemenschneider Vanishing). Let f:XYf:X\to Y be generically finite and surjective projective morphism between varieties. Assume XX is smoooth. Then RifOX(KX)=0R^{i}f_{*}\mathcal{O}_{X}(K_{X})=0 for i>0i>0.

Proof

Proof. We may assume YY is projective. The general case comes from compatify YY and XX locally. Let AA be ample divisor on YY, then mfAmf^{*}A is nef and big on XX. So Hi(OX(KX+fmA))=0H^{i}(\mathcal{O}_{X}(K_{X}+f^{*}mA))=0 by Theorem 137 and by above lemma RifOX(KX)=0R^{i}f_{*}\mathcal{O}_{X}(K_{X})=0. ◻

Example 147. Let XX be a projective variety and μ:XX\mu:X'\to X be a resolution of singularities. Let KX=μOX(KX)\mathcal{K}_{X}=\mu_{*}\mathcal{O}_{X'}(K_{X'}). Then KX\mathcal{K}_{X} is independent of the choice of resolution. For big and nef divisor DD, we have Hi(KX(D))=0H^{i}(\mathcal{K}_{X}(D))=0 for all i>0i>0.

To see so, note that any two resolutions can be dominated by the third one. If ν:XX\nu:X''\to X' is a projective birational morphism of smooth varieties, then ν(OX(KX))=OX(KX)\nu_{*}(\mathcal{O}_{X''}(K_{X''}))=\mathcal{O}_{X'}(K_{X'}). So we can reduce to a further resolution XX''. Since higher direct image RiμOX(KX+fD)R^{i}\mu_{*}\mathcal{O}_{X'}(K_{X}+f^{*}D) vanishes, Hi(X,KX(D))=Hi(X,OX(KX+μD))=0H^{i}(X,\mathcal{K}_{X}(D))=H^{i}(X',\mathcal{O}_{X'}(K_{X'}+\mu^{*}D))=0 by Theorem 129.

Example 148. We say a normal variety XX has rational singularities if there is a resolution μ:XX\mu:X'\to X such that RiμOX=0R^{i}\mu_{*}\mathcal{O}_{X'}=0. If DD is big and nef divisor on XX then Hj(X,OX(D))=0H^{j}(X,\mathcal{O}_{X}(-D))=0 for j<nj<n.

Theorem 149 (Fujita Vanishing). *Let XX be a projective variety and HH is an integral divisor on XX. For any coherent sheaf F\mathcal{F} on XX, there is an integer m(F,H)>0m(\mathcal{F},H)>0 such that

Hi(X,FOX(mH+D))=0H^{i}(X,\mathcal{F}\otimes\mathcal{O}_{X}(mH+D))=0

for all i>0i>0, mm(F,H)m\geq m(\mathcal{F},H) and any nef divisor DD.*

Proof

Proof. For any coherent sheaf F\mathcal{F}, using the resolution one can reduce to show for OX(aH)\mathcal{O}_{X}(aH). In fact, it suffices to show for a specific aa. Let μ:XX\mu:X'\to X be a resolution of singularities and KX=μOX(KX)\mathcal{K}_{X}=\mu_{*}\mathcal{O}_{X'}(K_{X'}). For a sufficiently large aa there is an inejction u:KXOX(aH)u:\mathcal{K}_{X}\to \mathcal{O}_{X}(aH). The cokernel cokeru\mathop{\mathrm{coker}}u is supported on a proper subvariety of XX, by induction on dimension, we only need to show for KX\mathcal{K}_{X}.

Note RjμOX(KX)=0R^{j}\mu_{*}\mathcal{O}_{X'}(K_{X'})=0 for j>0j>0, so

Hi(X,KXOX(aH+D))=Hi(X,OX(KX+μ(aH+D)))=0H^{i}(X,\mathcal{K}_{X}\otimes \mathcal{O}_{X}(aH+D))=H^{i}(X',\mathcal{O}_{X'}(K_{X'}+\mu^{*}(aH+D)))=0

by Theorem 137. ◻

Theorem 150. Let XX be a projective variety of dimension nn. DD is a nef divisor on XX. Then for any coherent sheaf F\mathcal{F} on XX, hi(F(mD))=O(mni)h^{i}(\mathcal{F}(mD))=O(m^{n-i}).

Proof

Proof. Use the induction on dimension and consider the exact sequence

0F(mD)F(mD+H)F(mD+h)OH0.0\to \mathcal{F}(mD)\to \mathcal{F}(mD+H)\to \mathcal{F}(mD+h)\otimes \mathcal{O}_{H}\to 0.

Corollary 151. Let XX be a projective variety of dimension nn and DD be a nef divisor on XX. Then

h0(O(mD))=Dnn!mn+O(mn1).h^{0}(\mathcal{O}(mD))=\frac{D^{n}}{n!}m^{n}+O(m^{n-1}).

More generally, for coherent sheaf F\mathcal{F} on XX,

h0(F(mD))=rk(F)Dnn!mn+O(mn1).h^{0}(\mathcal{F}(mD))=\mathop{\mathrm{rk}}(\mathcal{F})\frac{D^{n}}{n!}m^{n}+O(m^{n-1}).

评论