Projective variety






An elliptic curve is a smooth projective curve of genus one.


In algebraic geometry, a projective variety over an algebraically closed field k is a subset of some projective n-space Pn{displaystyle mathbb {P} ^{n}}mathbb {P} ^{n} over k that is the zero-locus of some finite family of homogeneous polynomials of n + 1 variables with coefficients in k, that generate a prime ideal, the defining ideal of the variety. Equivalently, an algebraic variety is projective if it can be embedded as a Zariski closed subvariety of Pn{displaystyle mathbb {P} ^{n}}mathbb {P} ^{n}.


A projective variety is a projective curve if its dimension is one; it is a projective surface if its dimension is two; it is a projective hypersurface if its dimension is one less than the dimension of the containing projective space; in this case it is the set of zeros of a single homogeneous polynomial.


If X is a projective variety defined by a homogeneous prime ideal I, then the quotient ring


k[x0,…,xn]/I{displaystyle k[x_{0},ldots ,x_{n}]/I}k[x_{0},ldots ,x_{n}]/I

is called the homogeneous coordinate ring of X. Basic invariants of X such as the degree and the dimension can be read off the Hilbert polynomial of this graded ring.


Projective varieties arise in many ways. They are complete, which roughly can be expressed by saying that there are no points "missing". The converse is not true in general, but Chow's lemma describes the close relation of these two notions. Showing that a variety is projective is done by studying line bundles or divisors on X.


A salient feature of projective varieties are the finiteness constraints on sheaf cohomology. For smooth projective varieties, Serre duality can be viewed as an analog of Poincaré duality. It also leads to the Riemann-Roch theorem for projective curves, i.e., projective varieties of dimension 1. The theory of projective curves is particularly rich, including a classification by the genus of the curve. The classification program for higher-dimensional projective varieties naturally leads to the construction of moduli of projective varieties.[1]Hilbert schemes parametrize closed subschemes of Pn{displaystyle mathbb {P} ^{n}}mathbb {P} ^{n} with prescribed Hilbert polynomial. Hilbert schemes, of which Grassmannians are special cases, are also projective schemes in their own right. Geometric invariant theory offers another approach. The classical approaches include the Teichmüller space and Chow varieties.


A particularly rich theory, reaching back to the classics, is available for complex projective varieties, i.e., when the polynomials defining X have complex coefficients. Broadly, the GAGA principle says that the geometry of projective complex analytic spaces (or manifolds) is equivalent to the geometry of projective complex varieties. For example, the theory of holomorphic vector bundles (more generally coherent analytic sheaves) on X coincide with that of algebraic vector bundles. Chow's theorem says that a subset of projective space is the zero-locus of a family of holomorphic functions if and only if it is the zero-locus of homogeneous polynomials. The combination of analytic and algebraic methods for complex projective varieties lead to areas such as Hodge theory.




Contents






  • 1 Variety and scheme structure


    • 1.1 Variety structure


    • 1.2 Projective schemes




  • 2 Relation to complete varieties


  • 3 Examples and basic invariants


    • 3.1 Homogeneous coordinate ring and Hilbert polynomial


    • 3.2 Degree


    • 3.3 The ring of sections


    • 3.4 Projective curves


    • 3.5 Projective hypersurfaces


    • 3.6 Abelian varieties




  • 4 Projections


  • 5 Duality and linear system


  • 6 Cohomology of coherent sheaves


  • 7 Smooth projective varieties


    • 7.1 Serre duality


    • 7.2 Riemann-Roch theorem




  • 8 Hilbert schemes


  • 9 Complex projective varieties


    • 9.1 Relation to complex Kähler manifolds


    • 9.2 GAGA and Chow's theorem


    • 9.3 Complex tori vs. complex abelian varieties


    • 9.4 Kodaira vanishing


    • 9.5 Further topics




  • 10 Related notions


  • 11 See also


  • 12 Notes


  • 13 References


  • 14 External links





Variety and scheme structure



Variety structure


Let k be an algebraically closed field. The basis of the definition of projective varieties is projective space Pn{displaystyle mathbb {P} ^{n}}mathbb {P} ^{n}, which can be defined in different, but equivalent ways:


  • as the set of all lines through the origin in kn+1{displaystyle k^{n+1}}{displaystyle k^{n+1}} (i.e., one-dimensional sub-vector spaces of kn+1{displaystyle k^{n+1}}{displaystyle k^{n+1}})

  • as the set of tuples (x0,…,xn)∈kn+1{displaystyle (x_{0},dots ,x_{n})in k^{n+1}}{displaystyle (x_{0},dots ,x_{n})in k^{n+1}}, modulo the equivalence relation


(x0,…,xn)∼λ(x0,…,xn){displaystyle (x_{0},dots ,x_{n})sim lambda (x_{0},dots ,x_{n})}{displaystyle (x_{0},dots ,x_{n})sim lambda (x_{0},dots ,x_{n})}

for any λk∖{0}{displaystyle lambda in kbackslash {0}}{displaystyle lambda in kbackslash {0}}. The equivalence class of such a tuple is denoted by
[x0:⋯:xn]{displaystyle [x_{0}:dots :x_{n}]}{displaystyle [x_{0}:dots :x_{n}]}


and referred to as a homogeneous coordinate.


A projective variety is, by definition, a closed subvariety of Pn{displaystyle mathbb {P} ^{n}}mathbb {P} ^{n}, where closed refers to the Zariski topology.[2] In general, closed subsets of the Zariski topology are defined to be the zero-locus of polynomial functions. Given a polynomial f∈k[x0,…,xn]{displaystyle fin k[x_{0},dots ,x_{n}]}{displaystyle fin k[x_{0},dots ,x_{n}]}, the condition


f([x0:⋯:xn])=0{displaystyle f([x_{0}:dots :x_{n}])=0}{displaystyle f([x_{0}:dots :x_{n}])=0}

does not make sense for arbitrary polynomials, but only if f is homogeneous, i.e., the total degree of all the monomials (whose sum is f) is the same. In this case, the vanishing of


f(λx0,…xn)=λdeg⁡ff(x0,…,xn){displaystyle f(lambda x_{0},dots ,lambda x_{n})=lambda ^{deg f}f(x_{0},dots ,x_{n})}{displaystyle f(lambda x_{0},dots ,lambda x_{n})=lambda ^{deg f}f(x_{0},dots ,x_{n})}

is independent of the choice of λ(≠0){displaystyle lambda (neq 0)}{displaystyle lambda (neq 0)}.


Therefore, projective varieties arise from homogeneous prime ideals I of k[x0,...,xn]{displaystyle k[x_{0},...,x_{n}]}k[x_{0},...,x_{n}], and setting



X={[x0:⋯:xn]∈Pn,f([x0:⋯:xn])=0 for all f∈I}.{displaystyle X={[x_{0}:dots :x_{n}]in mathbb {P} ^{n},f([x_{0}:dots :x_{n}])=0{text{ for all }}fin I}.}{displaystyle X={[x_{0}:dots :x_{n}]in mathbb {P} ^{n},f([x_{0}:dots :x_{n}])=0{text{ for all }}fin I}.}.

Moreover, the projective variety X is an algebraic variety, meaning that it is covered by open affine subvarieties and satisfies the separation axiom. Thus, the local study of X (e.g., singularity) reduces to that of an affine variety. The explicit structure is as follows. The projective space Pn{displaystyle mathbb {P} ^{n}}mathbb {P} ^{n} is covered by the standard open affine charts


Ui={[x0:⋯:xn],xi≠0},{displaystyle U_{i}={[x_{0}:dots :x_{n}],x_{i}neq 0},}{displaystyle U_{i}={[x_{0}:dots :x_{n}],x_{i}neq 0},}

which themselves are affine n-spaces with the coordinate ring


k[y1(i),…,yn(i)],yj(i)=xj/xi.{displaystyle kleft[y_{1}^{(i)},dots ,y_{n}^{(i)}right],quad y_{j}^{(i)}=x_{j}/x_{i}.}{displaystyle kleft[y_{1}^{(i)},dots ,y_{n}^{(i)}right],quad y_{j}^{(i)}=x_{j}/x_{i}.}

Say i = 0 for the notational simplicity and drop the superscript (0). Then X∩U0{displaystyle Xcap U_{0}}{displaystyle Xcap U_{0}} is a closed subvariety of U0≃An{displaystyle U_{0}simeq mathbb {A} ^{n}}{displaystyle U_{0}simeq mathbb {A} ^{n}} defined by the ideal of k[y1,…,yn]{displaystyle k[y_{1},dots ,y_{n}]}{displaystyle k[y_{1},dots ,y_{n}]} generated by


f(1,y1,…,yn){displaystyle f(1,y_{1},dots ,y_{n})}{displaystyle f(1,y_{1},dots ,y_{n})}

for all f in I. Thus, X is an algebraic variety covered by (n+1) open affine charts X∩Ui{displaystyle Xcap U_{i}}{displaystyle Xcap U_{i}}.


Note that X is the closure of the affine variety X∩U0{displaystyle Xcap U_{0}}{displaystyle Xcap U_{0}} in Pn{displaystyle mathbb {P} ^{n}}mathbb {P} ^{n}. Conversely, starting from some closed (affine) variety V⊂U0≃An{displaystyle Vsubset U_{0}simeq mathbb {A} ^{n}}{displaystyle Vsubset U_{0}simeq mathbb {A} ^{n}}, the closure of V in Pn{displaystyle mathbb {P} ^{n}}mathbb {P} ^{n} is the projective variety called the projective completion of V. If I⊂k[y1,…,yn]{displaystyle Isubset k[y_{1},dots ,y_{n}]}{displaystyle Isubset k[y_{1},dots ,y_{n}]} defines V, then the defining ideal of this closure is the homogeneous ideal[3] of k[x0,…,xn]{displaystyle k[x_{0},dots ,x_{n}]}k[x_{0},dots ,x_{n}] generated by


x0deg⁡(f)f(x1/x0,…,xn/x0){displaystyle x_{0}^{deg(f)}f(x_{1}/x_{0},dots ,x_{n}/x_{0})}{displaystyle x_{0}^{deg(f)}f(x_{1}/x_{0},dots ,x_{n}/x_{0})}

for all f in I.


For example, if V is an affine curve given by, say, y2=x3+ax+b{displaystyle y^{2}=x^{3}+ax+b}y^{2}=x^{3}+ax+b in the affine plane, then its projective completion in the projective plane is given by y2z=x3+axz2+bz3.{displaystyle y^{2}z=x^{3}+axz^{2}+bz^{3}.}{displaystyle y^{2}z=x^{3}+axz^{2}+bz^{3}.}



Projective schemes


For various applications, it is necessary to consider more general algebro-geometric objects than projective varieties, namely projective schemes. The first step towards projective schemes is to endow projective space with a scheme structure, in a way refining the above description of projective space as an algebraic variety, i.e., Pn(k){displaystyle mathbb {P} ^{n}(k)}{displaystyle mathbb {P} ^{n}(k)} is a scheme which it is a union of (n + 1) copies of the affine n-space kn. More generally,[4] projective space over a ring A is the union of the affine schemes


Ui=Spec⁡A[x1/xi,…,xn/xi],0≤i≤n,{displaystyle U_{i}=operatorname {Spec} A[x_{1}/x_{i},dots ,x_{n}/x_{i}],quad 0leq ileq n,}U_{i}=operatorname {Spec} A[x_{1}/x_{i},dots ,x_{n}/x_{i}],quad 0leq ileq n,

in such a way the variables match up as expected. The set of closed points of Pkn{displaystyle mathbb {P} _{k}^{n}}{mathbb  {P}}_{k}^{n}, for algebraically closed fields k, is then the projective space Pn(k){displaystyle mathbb {P} ^{n}(k)}{displaystyle mathbb {P} ^{n}(k)} in the usual sense.


An equivalent but streamlined construction is given by the Proj construction, which is an analog of the spectrum of a ring, denoted "Spec", which defines an affine scheme.[5] For example, if A is a ring, then


PAn=Proj⁡A[x0,…,xn].{displaystyle mathbb {P} _{A}^{n}=operatorname {Proj} A[x_{0},ldots ,x_{n}].}{displaystyle mathbb {P} _{A}^{n}=operatorname {Proj} A[x_{0},ldots ,x_{n}].}

If R is a quotient of k[x0,…,xn]{displaystyle k[x_{0},ldots ,x_{n}]}k[x_{0},ldots ,x_{n}] by a homogeneous ideal I, then the canonical surjection induces the closed immersion


Proj⁡R↪Pkn.{displaystyle operatorname {Proj} Rhookrightarrow mathbb {P} _{k}^{n}.}{displaystyle operatorname {Proj} Rhookrightarrow mathbb {P} _{k}^{n}.}

Compared to projective varieties, the condition that the ideal I be a prime ideal was dropped. This leads to a much more flexible notion: on the one hand the topological space X=Proj⁡R{displaystyle X=operatorname {Proj} R}{displaystyle X=operatorname {Proj} R} may have multiple irreducible components. Moreover, there may be nilpotent functions on X.


Closed subschemes of Pkn{displaystyle mathbb {P} _{k}^{n}}{mathbb  {P}}_{k}^{n} correspond bijectively to the homogeneous ideals I of k[x0,…,xn]{displaystyle k[x_{0},ldots ,x_{n}]}k[x_{0},ldots ,x_{n}] that are saturated; i.e., I:(x0,…,xn)=I.{displaystyle I:(x_{0},dots ,x_{n})=I.}{displaystyle I:(x_{0},dots ,x_{n})=I.}[6] This fact may be considered as a refined version of projective Nullstellensatz.


We can give a coordinate-free analog of the above. Namely, given a finite-dimensional vector space V over k, we let


P(V)=Proj⁡k[V]{displaystyle mathbb {P} (V)=operatorname {Proj} k[V]}{displaystyle mathbb {P} (V)=operatorname {Proj} k[V]}

where k[V]=Sym⁡(V∗){displaystyle k[V]=operatorname {Sym} (V^{*})}k[V]=operatorname {Sym} (V^{*}) is the symmetric algebra of V∗{displaystyle V^{*}}V^{*}.[7] It is the projectivization of V; i.e., it parametrizes lines in V. There is a canonical surjective map π:V∖{0}→P(V){displaystyle pi :Vsetminus {0}to mathbb {P} (V)}{displaystyle pi :Vsetminus {0}to mathbb {P} (V)}, which is defined using the chart described above.[8] One important use of the construction is this (cf., § Duality and linear system). A divisor D on a projective variety X corresponds to a line bundle L. One then set



|D|=P(Γ(X,L)){displaystyle |D|=mathbb {P} (Gamma (X,L))}{displaystyle |D|=mathbb {P} (Gamma (X,L))};

it is called the complete linear system of D.


Projective space over any scheme S can be defined as a fiber product of schemes


PSn=PZn×Spec⁡ZS.{displaystyle mathbb {P} _{S}^{n}=mathbb {P} _{mathbb {Z} }^{n}times _{operatorname {Spec} mathbb {Z} }S.}{displaystyle mathbb {P} _{S}^{n}=mathbb {P} _{mathbb {Z} }^{n}times _{operatorname {Spec} mathbb {Z} }S.}

If O(1){displaystyle {mathcal {O}}(1)}{mathcal {O}}(1) is the twisting sheaf of Serre on PZn{displaystyle mathbb {P} _{mathbb {Z} }^{n}}{displaystyle mathbb {P} _{mathbb {Z} }^{n}}, we let O(1){displaystyle {mathcal {O}}(1)}{mathcal {O}}(1) denote the pullback of O(1){displaystyle {mathcal {O}}(1)}{mathcal {O}}(1) to PSn{displaystyle mathbb {P} _{S}^{n}}{displaystyle mathbb {P} _{S}^{n}}; that is, O(1)=g∗(O(1)){displaystyle {mathcal {O}}(1)=g^{*}({mathcal {O}}(1))}{mathcal {O}}(1)=g^{*}({mathcal {O}}(1)) for the canonical map g:PSn→PZn.{displaystyle g:mathbb {P} _{S}^{n}to mathbb {P} _{mathbb {Z} }^{n}.}{displaystyle g:mathbb {P} _{S}^{n}to mathbb {P} _{mathbb {Z} }^{n}.}


A scheme XS is called projective over S if it factors as a closed immersion


X→PSn{displaystyle Xto mathbb {P} _{S}^{n}}{displaystyle Xto mathbb {P} _{S}^{n}}

followed by the projection to S.


A line bundle (or invertible sheaf) L{displaystyle {mathcal {L}}}{mathcal {L}} on a scheme X over S is said to be very ample relative to S if there is an immersion (i.e., an open immersion followed by a closed immersion)


i:X→PSn{displaystyle i:Xto mathbb {P} _{S}^{n}}{displaystyle i:Xto mathbb {P} _{S}^{n}}

for some n so that O(1){displaystyle {mathcal {O}}(1)}{mathcal {O}}(1) pullbacks to L.{displaystyle {mathcal {L}}.}{mathcal {L}}. Then a S-scheme X is projective if and only if it is proper and there exists a very ample sheaf on X relative to S. Indeed, if X is proper, then an immersion corresponding to the very ample line bundle is necessarily closed. Conversely, if X is projective, then the pullback of O(1){displaystyle {mathcal {O}}(1)}{mathcal {O}}(1) under the closed immersion of X into a projective space is very ample. That "projective" implies "proper" is deeper: the main theorem of elimination theory.



Relation to complete varieties


By definition, a variety is complete, if it is proper over k. The valuative criterion of properness expresses the intuition that in a proper variety, there are no points "missing".


There is a close relation between complete and projective varieties: on the one hand, projective space and therefore any projective variety is complete. The converse is not true in general. However:


  • A smooth curve C is projective if and only if it is complete. This is proved by identifying C with the set of discrete valuation rings of the function field k(C) over k. This set has a natural Zariski topology called the Zariski–Riemann space.


  • Chow's lemma states that for any complete variety X, there is a projective variety Z and a birational morphism ZX.[9] (Moreover, through normalization, one can assume this projective variety is normal.)

Some properties of a projective variety follow from completeness. For example,


Γ(X,OX)=k{displaystyle Gamma (X,{mathcal {O}}_{X})=k}Gamma (X,{mathcal {O}}_{X})=k

for any projective variety X over k.[10] This fact is an algebraic analogue of Liouville's theorem (any holomorphic function on a connected compact complex manifold is constant). In fact, the similarity between complex analytic geometry and algebraic geometry on complex projective varieties goes much further than this, as is explained below.


Quasi-projective varieties are, by definition, those which are open subvarieties of projective varieties. This class of varieties includes affine varieties. Affine varieties are almost never complete (or projective). In fact, a projective subvariety of an affine variety must have dimension zero. This is because only the constants are globally regular functions on a projective variety.



Examples and basic invariants


By definition, any homogeneous ideal in a polynomial ring yields a projective scheme (required to be prime ideal to give a variety). In this sense, examples of projective varieties abound. The following list mentions various classes of projective varieties which are noteworthy since they have been studied particularly intensely. The important class of complex projective varieties, i.e., the case k=C,{displaystyle k=mathbb {C} ,}{displaystyle k=mathbb {C} ,} is discussed further below.


The product of two projective spaces is projective. In fact, there is the explicit immersion (called Segre embedding)


{Pn×Pm→P(n+1)(m+1)−1(xi,yj)↦xiyj{displaystyle {begin{cases}mathbb {P} ^{n}times mathbb {P} ^{m}to mathbb {P} ^{(n+1)(m+1)-1}\(x_{i},y_{j})mapsto x_{i}y_{j}end{cases}}}{displaystyle {begin{cases}mathbb {P} ^{n}times mathbb {P} ^{m}to mathbb {P} ^{(n+1)(m+1)-1}\(x_{i},y_{j})mapsto x_{i}y_{j}end{cases}}}

As a consequence, the product of projective varieties over k is again projective. The Plücker embedding exhibits a Grassmannian as a projective variety. Flag varieties such as the quotient of the general linear group GLn(k){displaystyle GL_{n}(k)}{displaystyle GL_{n}(k)} modulo the subgroup of upper triangular matrices, are also projective, which is an important fact in the theory of algebraic groups.[11]



Homogeneous coordinate ring and Hilbert polynomial



As the prime ideal P defining a projective variety X is homogeneous, the homogeneous coordinate ring


R=k[x0,…,xn]/P{displaystyle R=k[x_{0},dots ,x_{n}]/P}{displaystyle R=k[x_{0},dots ,x_{n}]/P}

is a graded ring, i.e., can be expressed as the direct sum of its graded components:


R=⨁n∈NRn.{displaystyle R=bigoplus _{nin mathbb {N} }R_{n}.}{displaystyle R=bigoplus _{nin mathbb {N} }R_{n}.}

There exists a polynomial P such that dim⁡Rn=P(n){displaystyle dim R_{n}=P(n)}{displaystyle dim R_{n}=P(n)} for all sufficiently large n; it is called the Hilbert polynomial of X. It is a numerical invariant encoding some extrinsic geometry of X. The degree of P is the dimension r of X and its leading coefficient times r! is the degree of the variety X. The arithmetic genus of X is (−1)r (P(0) − 1) when X is smooth.


For example, the homogeneous coordinate ring of Pn{displaystyle mathbb {P} ^{n}}mathbb {P} ^{n} is k[x0,…,xn]{displaystyle k[x_{0},ldots ,x_{n}]}k[x_{0},ldots ,x_{n}] and its Hilbert polynomial is P(z)=(z+nn){displaystyle P(z)={binom {z+n}{n}}}P(z)={binom {z+n}{n}}; its arithmetic genus is zero.


If the homogeneous coordinate ring R is an integrally closed domain, then the projective variety X is said to be projectively normal. Note, unlike normality, projective normality depends on R, the embedding of X into a projective space. The normalization of a projective variety is projective; in fact, it's the Proj of the integral closure of some homogeneous coordinate ring of X.



Degree



Let X⊂PN{displaystyle Xsubset mathbb {P} ^{N}}{displaystyle Xsubset mathbb {P} ^{N}} be a projective variety. There are at least two equivalent ways to define the degree of X relative to its embedding. The first way is to define it as the cardinality of the finite set


#(X∩H1∩Hd){displaystyle #(Xcap H_{1}cap cdots cap H_{d})}{displaystyle #(Xcap H_{1}cap cdots cap H_{d})}

where d is the dimension of X and Hi's are hyperplanes in "general positions". This definition corresponds to an intuitive idea of a degree. Indeed, if X is a hypersurface, then the degree of X is the degree of the homogeneous polynomial defining X. The "general positions" can be made precise, for example, by intersection theory; one requires that the intersection is proper and that the multiplicities of irreducible components are all one.


The other definition, which is mentioned in the previous section, is that the degree of X is the leading coefficient of the Hilbert polynomial of X times (dim X)!. Geometrically, this definition means that the degree of X is the multiplicity of the vertex of the affine cone over X.[12]


Let V1,…,Vr⊂PN{displaystyle V_{1},dots ,V_{r}subset mathbb {P} ^{N}}{displaystyle V_{1},dots ,V_{r}subset mathbb {P} ^{N}} be closed subschemes of pure dimensions that intersect properly (they are in general position). If mi denotes the multiplicity of an irreducible component Zi in the intersection (i.e., intersection multiplicity), then the generalization of Bézout's theorem says:[13]


1smideg⁡Zi=∏1rdeg⁡Vi.{displaystyle sum _{1}^{s}m_{i}deg Z_{i}=prod _{1}^{r}deg V_{i}.}{displaystyle sum _{1}^{s}m_{i}deg Z_{i}=prod _{1}^{r}deg V_{i}.}

The intersection multiplicity mi can be defined as the coefficient of Zi in the intersection product V1⋅Vr{displaystyle V_{1}cdot cdots cdot V_{r}}{displaystyle V_{1}cdot cdots cdot V_{r}} in the Chow ring of PN{displaystyle mathbb {P} ^{N}}{mathbb  {P}}^{N}.


In particular, if H⊂PN{displaystyle Hsubset mathbb {P} ^{N}}{displaystyle Hsubset mathbb {P} ^{N}} is a hypersurface not containing X, then


1smideg⁡Zi=deg⁡(X)deg⁡(H){displaystyle sum _{1}^{s}m_{i}deg Z_{i}=deg(X)deg(H)}{displaystyle sum _{1}^{s}m_{i}deg Z_{i}=deg(X)deg(H)}

where Zi are the irreducible components of the scheme-theoretic intersection of X and H with multiplicity (length of the local ring) mi.



The ring of sections


Let X be a projective variety and L a line bundle on it. Then the graded ring


R(X,L)=⨁n=0∞H0(X,L⊗n){displaystyle R(X,L)=bigoplus _{n=0}^{infty }H^{0}(X,L^{otimes n})}{displaystyle R(X,L)=bigoplus _{n=0}^{infty }H^{0}(X,L^{otimes n})}

is called the ring of sections of L. If L is ample, then Proj of this ring is X. Moreover, if X is normal and L is very ample, then R(X,L){displaystyle R(X,L)}{displaystyle R(X,L)} is the integral closure of the homogeneous coordinate ring of X determined by L; i.e., X↪PN{displaystyle Xhookrightarrow mathbb {P} ^{N}}{displaystyle Xhookrightarrow mathbb {P} ^{N}} so that OPN(1){displaystyle {mathcal {O}}_{mathbb {P} ^{N}}(1)}{displaystyle {mathcal {O}}_{mathbb {P} ^{N}}(1)} pulls-back to L.[14]


For applications, it is useful to allow for divisors (or Q{displaystyle mathbb {Q} }mathbb {Q} -divisors) not just line bundles; assuming X is normal, the resulting ring is then called a generalized ring of sections. If KX{displaystyle K_{X}}K_{X} is a canonical divisor on X, then the generalized ring of sections


R(X,KX){displaystyle R(X,K_{X})}{displaystyle R(X,K_{X})}

is called the canonical ring of X. If the canonical ring is finitely generated, then Proj of the ring is called the canonical model of X. The canonical ring or model can then be used to define the Kodaira dimension of X.



Projective curves



Projective schemes of dimension one are called projective curves. Much of the theory of projective curves is about smooth projective curves, since the singularities of curves can be resolved by normalization, which consists in taking locally the integral closure of the ring of regular functions. Smooth projective curves are isomorphic if and only if their function fields are isomorphic. The study of finite extensions of


Fp(t),{displaystyle mathbb {F} _{p}(t),}{displaystyle mathbb {F} _{p}(t),}

or equivalently smooth projective curves over Fp{displaystyle mathbb {F} _{p}}mathbb {F} _{p} is an important branch in algebraic number theory.[15]


A smooth projective curve of genus one is called an elliptic curve. As a consequence of the Riemann-Roch theorem, such a curve can be embedded as a closed subvariety in P2{displaystyle mathbb {P} ^{2}}mathbb {P} ^{2}. In general, any (smooth) projective curve can be embedded in P3{displaystyle mathbb {P} ^{3}}{mathbb  {P}}^{3} (for a proof, see Secant variety#Examples). Conversely, any smooth closed curve in P2{displaystyle mathbb {P} ^{2}}mathbb {P} ^{2} of degree three has genus one by the genus formula and is thus an elliptic curve.


A smooth complete curve of genus greater than or equal to two is called a hyperelliptic curve if there is a finite morphism C→P1{displaystyle Cto mathbb {P} ^{1}}C to mathbb{P}^1 of degree two.[16]



Projective hypersurfaces


Every irreducible closed subset of Pn{displaystyle mathbb {P} ^{n}}mathbb {P} ^{n} of codimension one is a hypersurface; i.e., the zero set of some homogeneous irreducible polynomial.[17]



Abelian varieties


Another important invariant of a projective variety X is the Picard group Pic⁡(X){displaystyle operatorname {Pic} (X)}operatorname {Pic} (X) of X, the set of isomorphism classes of line bundles on X. It is isomorphic to H1(X,OX∗){displaystyle H^{1}(X,{mathcal {O}}_{X}^{*})}H^{1}(X,{mathcal  O}_{X}^{*}) and therefore an intrinsic notion (independent of embedding). For example, the Picard group of Pn{displaystyle mathbb {P} ^{n}}mathbb {P} ^{n} is isomorphic to Z{displaystyle mathbb {Z} }mathbb {Z} via the degree map. The kernel of deg:Pic⁡(X)→Z{displaystyle deg :operatorname {Pic} (X)to mathbb {Z} }{displaystyle deg :operatorname {Pic} (X)to mathbb {Z} } is not only an abstract abelian group, but there is a variety called the Jacobian variety of X, Jac(X), whose points equal this group. The Jacobian of a (smooth) curve plays an important role in the study of the curve. For example, the Jacobian of an elliptic curve E is E itself. For a curve X of genus g, Jac(X) has dimension g.


Varieties, such as the Jacobian variety, which are complete and have a group structure are known as abelian varieties, in honor of Niels Abel. In marked contrast to affine algebraic groups such as GLn(k){displaystyle GL_{n}(k)}{displaystyle GL_{n}(k)}, such groups are always commutative, whence the name. Moreover, they admit an ample line bundle and are thus projective. On the other hand, an abelian scheme may not be projective. Examples of abelian varieties are elliptic curves, Jacobian varieties and K3 surfaces.



Projections


Let E⊂Pn{displaystyle Esubset mathbb {P} ^{n}}{displaystyle Esubset mathbb {P} ^{n}} be a linear subspace; i.e., E={s0=s1=⋯=sr=0}{displaystyle E={s_{0}=s_{1}=cdots =s_{r}=0}}{displaystyle E={s_{0}=s_{1}=cdots =s_{r}=0}} for some linearly independent linear functionals si. Then the projection from E is the (well-defined) morphism


:Pn−E→Prx↦[s0(x):⋯:sr(x)]{displaystyle {begin{cases}phi :mathbb {P} ^{n}-Eto mathbb {P} ^{r}\xmapsto [s_{0}(x):cdots :s_{r}(x)]end{cases}}}{displaystyle {begin{cases}phi :mathbb {P} ^{n}-Eto mathbb {P} ^{r}\xmapsto [s_{0}(x):cdots :s_{r}(x)]end{cases}}}

The geometric description of this map is as follows:[18]


  • We view Pr⊂Pn{displaystyle mathbb {P} ^{r}subset mathbb {P} ^{n}}{displaystyle mathbb {P} ^{r}subset mathbb {P} ^{n}} so that it is disjoint from E. Then, for any x∈Pn∖E,{displaystyle xin mathbb {P} ^{n}setminus E,}{displaystyle xin mathbb {P} ^{n}setminus E,}


ϕ(x)=Wx∩Pr,{displaystyle phi (x)=W_{x}cap mathbb {P} ^{r},}{displaystyle phi (x)=W_{x}cap mathbb {P} ^{r},}

where Wx{displaystyle W_{x}}W_x denotes the smallest linear space containing E and x (called the join of E and x.)



  • ϕ1({yi≠0})={si≠0},{displaystyle phi ^{-1}({y_{i}neq 0})={s_{i}neq 0},}{displaystyle phi ^{-1}({y_{i}neq 0})={s_{i}neq 0},} where yi{displaystyle y_{i}}y_{i} are the homogeneous coordinates on Pr.{displaystyle mathbb {P} ^{r}.}{displaystyle mathbb {P} ^{r}.}

  • For any closed subscheme Z⊂Pn{displaystyle Zsubset mathbb {P} ^{n}}{displaystyle Zsubset mathbb {P} ^{n}} disjoint from E, the restriction ϕ:Z→Pr{displaystyle phi :Zto mathbb {P} ^{r}}{displaystyle phi :Zto mathbb {P} ^{r}} is a finite morphism.[19]

Projections can be used to cut down the dimension in which a projective variety is embedded, up to finite morphisms. Start with some projective variety X⊂Pn.{displaystyle Xsubset mathbb {P} ^{n}.}{displaystyle Xsubset mathbb {P} ^{n}.} If n>dim⁡X,{displaystyle n>dim X,}{displaystyle n>dim X,} the projection from a point not on X gives ϕ:X→Pn−1.{displaystyle phi :Xto mathbb {P} ^{n-1}.}{displaystyle phi :Xto mathbb {P} ^{n-1}.} Moreover, ϕ{displaystyle phi }phi is a finite map to its image. Thus, iterating the procedure, one sees there is a finite map


X→Pd,d=dim⁡X.{displaystyle Xto mathbb {P} ^{d},quad d=dim X.}{displaystyle Xto mathbb {P} ^{d},quad d=dim X.}

This result is the projective analog of Noether's normalization lemma. (In fact, it yields a geometric proof of the normalization lemma.)


The same procedure can be used to show the following slightly more precise result: given a projective variety X over a perfect field, there is a finite birational morphism from X to a hypersurface H in Pd+1.{displaystyle mathbb {P} ^{d+1}.}{displaystyle mathbb {P} ^{d+1}.}[20] In particular, if X is normal, then it is the normalization of H.



Duality and linear system


While a projective n-space Pn{displaystyle mathbb {P} ^{n}}mathbb {P} ^{n} parameterizes the lines in an affine n-space, the dual of it parametrizes the hyperplanes on the projective space, as follows. Fix a field k. By kn{displaystyle {breve {mathbb {P} }}_{k}^{n}}{displaystyle {breve {mathbb {P} }}_{k}^{n}}, we mean a projective n-space


kn=Proj⁡(k[u0,…,un]){displaystyle {breve {mathbb {P} }}_{k}^{n}=operatorname {Proj} (k[u_{0},dots ,u_{n}])}{displaystyle {breve {mathbb {P} }}_{k}^{n}=operatorname {Proj} (k[u_{0},dots ,u_{n}])}

equipped with the construction:



f↦Hf={α0x0+⋯nxn=0}{displaystyle fmapsto H_{f}={alpha _{0}x_{0}+cdots +alpha _{n}x_{n}=0}}{displaystyle fmapsto H_{f}={alpha _{0}x_{0}+cdots +alpha _{n}x_{n}=0}}, a hyperplane on PLn{displaystyle mathbb {P} _{L}^{n}}{displaystyle mathbb {P} _{L}^{n}}

where f:Spec⁡L→kn{displaystyle f:operatorname {Spec} Lto {breve {mathbb {P} }}_{k}^{n}}{displaystyle f:operatorname {Spec} Lto {breve {mathbb {P} }}_{k}^{n}} is an L-point of kn{displaystyle {breve {mathbb {P} }}_{k}^{n}}{displaystyle {breve {mathbb {P} }}_{k}^{n}} for a field extension L of k and αi=f∗(ui)∈L.{displaystyle alpha _{i}=f^{*}(u_{i})in L.}{displaystyle alpha _{i}=f^{*}(u_{i})in L.}


For each L, the construction is a bijection between the set of L-points of kn{displaystyle {breve {mathbb {P} }}_{k}^{n}}{displaystyle {breve {mathbb {P} }}_{k}^{n}} and the set of hyperplanes on PLn{displaystyle mathbb {P} _{L}^{n}}{displaystyle mathbb {P} _{L}^{n}}. Because of this, the dual projective space kn{displaystyle {breve {mathbb {P} }}_{k}^{n}}{displaystyle {breve {mathbb {P} }}_{k}^{n}} is said to be the moduli space of hyperplanes on Pkn{displaystyle mathbb {P} _{k}^{n}}{mathbb  {P}}_{k}^{n}.


A line in kn{displaystyle {breve {mathbb {P} }}_{k}^{n}}{displaystyle {breve {mathbb {P} }}_{k}^{n}} is called a pencil: it is a family of hyperplanes on Pkn{displaystyle mathbb {P} _{k}^{n}}{mathbb  {P}}_{k}^{n} parametrized by Pk1{displaystyle mathbb {P} _{k}^{1}}{mathbb  {P}}_{k}^{1}.


If V is a finite-dimensional vector space over k, then, for the same reason as above, P(V∗)=Proj⁡(Sym⁡(V)){displaystyle mathbb {P} (V^{*})=operatorname {Proj} (operatorname {Sym} (V))}{displaystyle mathbb {P} (V^{*})=operatorname {Proj} (operatorname {Sym} (V))} is the space of hyperplanes on P(V){displaystyle mathbb {P} (V)}mathbb {P} (V). An important case is when V consists of sections of a line bundle. Namely, let X be an algebraic variety, L a line bundle on X and V⊂Γ(X,L){displaystyle Vsubset Gamma (X,L)}{displaystyle Vsubset Gamma (X,L)} a vector subspace of finite positive dimension. Then there is a map:[21]


V:X∖B→P(V∗)x↦Hx={s∈V|s(x)=0}{displaystyle {begin{cases}varphi _{V}:Xsetminus Bto mathbb {P} (V^{*})\xmapsto H_{x}={sin V|s(x)=0}end{cases}}}{displaystyle {begin{cases}varphi _{V}:Xsetminus Bto mathbb {P} (V^{*})\xmapsto H_{x}={sin V|s(x)=0}end{cases}}}

determined by the linear system V, where B, called the base locus, is the intersection of the divisors of zero of nonzero sections in V.



Cohomology of coherent sheaves



Let X be a projective scheme over a field (or, more generally over a Noetherian ring A). Cohomology of coherent sheaves F{displaystyle {mathcal {F}}}{mathcal {F}} on X satisfies the following important theorems due to Serre:




  1. Hp(X,F){displaystyle H^{p}(X,{mathcal {F}})}H^{p}(X,{mathcal {F}}) is a finite-dimensional k-vector space for any p.

  2. There exists an integer n0{displaystyle n_{0}}n_{0} (depending on F{displaystyle {mathcal {F}}}{mathcal {F}}; see also Castelnuovo–Mumford regularity) such that



Hp(X,F(n))=0{displaystyle H^{p}(X,{mathcal {F}}(n))=0}H^{p}(X,{mathcal {F}}(n))=0

for all n≥n0{displaystyle ngeq n_{0}}ngeq n_{0} and p > 0, where F(n)=F⊗O(n){displaystyle {mathcal {F}}(n)={mathcal {F}}otimes {mathcal {O}}(n)}{displaystyle {mathcal {F}}(n)={mathcal {F}}otimes {mathcal {O}}(n)} is the twisting with a power of a very ample line bundle O(1).{displaystyle {mathcal {O}}(1).}{displaystyle {mathcal {O}}(1).}


These results are proven reducing to the case X=Pn{displaystyle X=mathbb {P} ^{n}}{displaystyle X=mathbb {P} ^{n}} using the isomorphism


Hp(X,F)=Hp(Pr,F),p≥0{displaystyle H^{p}(X,{mathcal {F}})=H^{p}(mathbb {P} ^{r},{mathcal {F}}),pgeq 0}{displaystyle H^{p}(X,{mathcal {F}})=H^{p}(mathbb {P} ^{r},{mathcal {F}}),pgeq 0}

where in the right-hand side F{displaystyle {mathcal {F}}}{mathcal {F}} is viewed as a sheaf on the projective space by extension by zero.[22] The result then follows by a direct computation for F=OPr(n),{displaystyle {mathcal {F}}={mathcal {O}}_{mathbb {P} ^{r}}(n),}{displaystyle {mathcal {F}}={mathcal {O}}_{mathbb {P} ^{r}}(n),} n any integer, and for arbitrary F{displaystyle {mathcal {F}}}{mathcal {F}} reduces to this case without much difficulty.[23]


As a corollay to 1. above, if f is a projective morphism from a noetherian scheme to a noetherian ring, then the higher direct image Rpf∗F{displaystyle R^{p}f_{*}{mathcal {F}}}R^{p}f_{*}{mathcal {F}} is coherent. The same result holds for proper morphisms f, as can be shown with the aid of Chow's lemma.


Sheaf cohomology groups Hi on a noetherian topological space vanish for i strictly greater than the dimension of the space. Thus the quantity, called the Euler characteristic of F{displaystyle {mathcal {F}}}{mathcal {F}},


χ(F)=∑i=0∞(−1)idim⁡Hi(X,F){displaystyle chi ({mathcal {F}})=sum _{i=0}^{infty }(-1)^{i}dim H^{i}(X,{mathcal {F}})}{displaystyle chi ({mathcal {F}})=sum _{i=0}^{infty }(-1)^{i}dim H^{i}(X,{mathcal {F}})}

is a well-defined integer (for X projective). One can then show χ(F(n))=P(n){displaystyle chi ({mathcal {F}}(n))=P(n)}chi ({mathcal {F}}(n))=P(n) for some polynomial P over rational numbers.[24] Applying this procedure to the structure sheaf OX{displaystyle {mathcal {O}}_{X}}{mathcal {O}}_{X}, one recovers the Hilbert polynomial of X. In particular, if X is irreducible and has dimension r, the arithmetic genus of X is given by


(−1)r(χ(OX)−1),{displaystyle (-1)^{r}(chi ({mathcal {O}}_{X})-1),}(-1)^{r}(chi ({mathcal {O}}_{X})-1),

which is manifestly intrinsic; i.e., independent of the embedding.


The arithmetic genus of a hypersurface of degree d is (d−1n){displaystyle {binom {d-1}{n}}}{binom {d-1}{n}} in Pn{displaystyle mathbb {P} ^{n}}mathbb {P} ^{n}. In particular, a smooth curve of degree d in P2{displaystyle mathbb {P} ^{2}}mathbb {P} ^{2} has arithmetic genus (d−1)(d−2)/2{displaystyle (d-1)(d-2)/2}(d-1)(d-2)/2. This is the genus formula.



Smooth projective varieties


Let X be a smooth projective variety where all of its irreducible components have dimension n. In this situation, the canonical sheaf ωX, defined as the sheaf of Kähler differentials of top degree (i.e., algebraic n-forms), is a line bundle.



Serre duality


Serre duality states that for any locally free sheaf F{displaystyle {mathcal {F}}}{mathcal {F}} on X,


Hi(X,F)≃Hn−i(X,F∨ωX)′{displaystyle H^{i}(X,{mathcal {F}})simeq H^{n-i}(X,{mathcal {F}}^{vee }otimes omega _{X})'}H^{i}(X,{mathcal {F}})simeq H^{n-i}(X,{mathcal {F}}^{vee }otimes omega _{X})'

where the superscript prime refers to the dual space and F∨{displaystyle {mathcal {F}}^{vee }}{mathcal {F}}^{vee } is the dual sheaf of F{displaystyle {mathcal {F}}}{mathcal {F}}.
A generalization to projective, but not necessarily smooth schemes is known as Verdier duality.



Riemann-Roch theorem


For a (smooth projective) curve X, H2 and higher vanish for dimensional reason and the space of the global sections of the structure sheaf is one-dimensional. Thus the arithmetic genus of X is the dimension of H1(X,OX){displaystyle H^{1}(X,{mathcal {O}}_{X})}H^{1}(X,{mathcal {O}}_{X}). By definition, the geometric genus of X is the dimension of H0(X, ωX). Serre duality thus implies that the arithmetic genus and the geometric genus coincide. They will simply be called the genus of X.


Serre duality is also a key ingredient in the proof of the Riemann–Roch theorem. Since X is smooth, there is an isomorphism of groups


{Cl⁡(X)→Pic⁡(X)D↦O(D){displaystyle {begin{cases}operatorname {Cl} (X)to operatorname {Pic} (X)\Dmapsto {mathcal {O}}(D)end{cases}}}{displaystyle {begin{cases}operatorname {Cl} (X)to operatorname {Pic} (X)\Dmapsto {mathcal {O}}(D)end{cases}}}

from the group of (Weil) divisors modulo principal divisors to the group of isomorphism classes of line bundles. A divisor corresponding to ωX is called the canonical divisor and is denoted by K. Let l(D) be the dimension of H0(X,O(D)){displaystyle H^{0}(X,{mathcal {O}}(D))}H^{0}(X,{mathcal {O}}(D)). Then the Riemann–Roch theorem states: if g is a genus of X,


l(D)−l(K−D)=deg⁡D+1−g,{displaystyle l(D)-l(K-D)=deg D+1-g,}{displaystyle l(D)-l(K-D)=deg D+1-g,}

for any divisor D on X. By the Serre duality, this is the same as:


χ(O(D))=deg⁡D+1−g,{displaystyle chi ({mathcal {O}}(D))=deg D+1-g,}{displaystyle chi ({mathcal {O}}(D))=deg D+1-g,}

which can be readily proved.[25] A generalization of the Riemann-Roch theorem to higher dimension is the Hirzebruch-Riemann-Roch theorem, as well as the far-reaching Grothendieck-Riemann-Roch theorem.



Hilbert schemes


Hilbert schemes parametrize all closed subvarieties of a projective scheme X in the sense that the points (in the functorial sense) of H correspond to the closed subschemes of X. As such, the Hilbert scheme is an example of a moduli space, i.e., a geometric object whose points parametrize other geometric objects. More precisely, the Hilbert scheme parametrizes closed subvarieties whose Hilbert polynomial equals a prescribed polynomial P.[26] It is a deep theorem of Grothendieck that there is a scheme[27]HXP{displaystyle H_{X}^{P}}H_{X}^{P} over k such that, for any k-scheme T, there is a bijection


{morphisms T→HXP}  ⟷  {closed subschemes of X×kT flat over T, with Hilbert polynomial P.}{displaystyle {{text{morphisms }}Tto H_{X}^{P}} longleftrightarrow {{text{closed subschemes of }}Xtimes _{k}T{text{ flat over }}T,{text{ with Hilbert polynomial }}P.}}{displaystyle {{text{morphisms }}Tto H_{X}^{P}}  longleftrightarrow   {{text{closed subschemes of }}Xtimes _{k}T{text{ flat over }}T,{text{ with Hilbert polynomial }}P.}}

The closed subscheme of HXP{displaystyle Xtimes H_{X}^{P}}{displaystyle Xtimes H_{X}^{P}} that corresponds to the identity map HXP→HXP{displaystyle H_{X}^{P}to H_{X}^{P}}H_{X}^{P}to H_{X}^{P} is called the universal family.


For P(z)=(z+rr){displaystyle P(z)={binom {z+r}{r}}}{displaystyle P(z)={binom {z+r}{r}}}, the Hilbert scheme HPnP{displaystyle H_{mathbb {P} ^{n}}^{P}}{displaystyle H_{mathbb {P} ^{n}}^{P}} is called the Grassmannian of r-planes in Pn{displaystyle mathbb {P} ^{n}}mathbb {P} ^{n} and, if X is a projective scheme, HXP{displaystyle H_{X}^{P}}H_{X}^{P} is called the Fano scheme of r-planes on X.[28]



Complex projective varieties



In this section, all algebraic varieties are complex algebraic varieties. A key feature of the theory of complex projective varieties is the combination of algebraic and analytic methods. The transition between these theories is provided by the following link: since any complex polynomial is also a holomorphic function, any complex variety X yields a complex analytic space, denoted X(C).{displaystyle X(mathbb {C} ).}{displaystyle X(mathbb {C} ).} Moreover, geometric properties of X are reflected by the ones of X(C).{displaystyle X(mathbb {C} ).}{displaystyle X(mathbb {C} ).} For example, the latter is a complex manifold iff X is smooth; it is compact iff X is proper over C.{displaystyle mathbb {C} .}{displaystyle mathbb {C} .}



Relation to complex Kähler manifolds


Complex projective space is a Kähler manifold. This implies that, for any projective algebraic variety X, X(C){displaystyle X(mathbb {C} )}{displaystyle X(mathbb {C} )} is a compact Kähler manifold. The converse is not in general true, but the Kodaira embedding theorem gives a criterion for a Kähler manifold to be projective.


In low dimensions, there are the following results:


  • (Riemann) A compact Riemann surface (i.e., compact complex manifold of dimension one) is a projective variety. By the Torelli theorem, it is uniquely determined by its Jacobian.

  • (Chow-Kodaira) A compact complex manifold of dimension two with two algebraically independent meromorphic functions is a projective variety.[29]


GAGA and Chow's theorem


Chow's theorem provides a striking way to go the other way, from analytic to algebraic geometry. It states that every analytic subvariety of a complex projective space is algebraic. The theorem may be interpreted to saying that a holomorphic function satisfying certain growth condition is necessarily algebraic: "projective" provides this growth condition. One can deduce from the theorem the following:


  • Meromorphic functions on the complex projective space are rational.

  • If an algebraic map between algebraic varieties is an analytic isomorphism, then it is an (algebraic) isomorphism. (This part is a basic fact in complex analysis.) In particular, Chow's theorem implies that a holomorphic map between projective varieties is algebraic. (consider the graph of such a map.)

  • Every holomorphic vector bundle on a projective variety is induced by a unique algebraic vector bundle.[30]

  • Every holomorphic line bundle on a projective variety is a line bundle of a divisor.[31]

Chow's theorem can be shown via Serre's GAGA principle. Its main theorem states:



Let X be a projective scheme over C{displaystyle mathbb {C} }mathbb {C} . Then the functor associating the coherent sheaves on X to the coherent sheaves on the corresponding complex analytic space Xan is an equivalence of categories. Furthermore, the natural maps
Hi(X,F)→Hi(Xan,F){displaystyle H^{i}(X,{mathcal {F}})to H^{i}(X^{text{an}},{mathcal {F}})}H^{i}(X,{mathcal {F}})to H^{i}(X^{text{an}},{mathcal {F}})


are isomorphisms for all i and all coherent sheaves F{displaystyle {mathcal {F}}}{mathcal {F}} on X.[32]



Complex tori vs. complex abelian varieties


The complex manifold associated to an abelian variety A over C{displaystyle mathbb {C} }mathbb {C} is a compact complex Lie group. These can be shown to be of the form


Cg/L{displaystyle mathbb {C} ^{g}/L}{displaystyle mathbb {C} ^{g}/L}

and are also referred to as complex tori. Here, g is the dimension of the torus and L is a lattice (also referred to as period lattice).


According to the uniformization theorem already mentioned above, any torus of dimension 1 arises from an abelian variety of dimension 1, i.e., from an elliptic curve. In fact, the Weierstrass's elliptic function {displaystyle wp }wp attached to L satisfies a certain differential equation and as a consequence it defines a closed immersion:[33]


{C/L→P2L↦(0:0:1)z↦(1:℘(z):℘′(z)){displaystyle {begin{cases}mathbb {C} /Lto mathbb {P} ^{2}\Lmapsto (0:0:1)\zmapsto (1:wp (z):wp '(z))end{cases}}}{displaystyle {begin{cases}mathbb {C} /Lto mathbb {P} ^{2}\Lmapsto (0:0:1)\zmapsto (1:wp (z):wp '(z))end{cases}}}

There is a p-adic analog, the p-adic uniformization theorem.


For higher dimensions, the notions of complex abelian varieties and complex tori differ: only polarized complex tori come from abelian varieties.



Kodaira vanishing


The fundamental Kodaira vanishing theorem states that for an ample line bundle L{displaystyle {mathcal {L}}}{mathcal {L}} on a smooth projective variety X over a field of characteristic zero,


Hi(X,L⊗ωX)=0{displaystyle H^{i}(X,{mathcal {L}}otimes omega _{X})=0}H^{i}(X,{mathcal {L}}otimes omega _{X})=0

for i > 0, or, equivalently by Serre duality Hi(X,L−1)=0{displaystyle H^{i}(X,{mathcal {L}}^{-1})=0}{displaystyle H^{i}(X,{mathcal {L}}^{-1})=0} for i < n.[34] The first proof of this theorem used analytic methods of Kähler geometry, but a purely algebraic proof was found later. The Kodaira vanishing in general fails for a smooth projective variety in positive characteristic. Kodaira's theorem is one of various vanishing theorems, which give criteria for higher sheaf cohomologies to vanish. Since the Euler characteristic of a sheaf (see above) is often more manageable than individual cohomology groups, this often has important consequences about the geometry of projective varieties.[35]



Further topics


Hodge decomposition, Hodge conjecture, Tate conjecture



Related notions



  • Multi-projective variety


  • Weighted projective variety, a closed subvariety of a weighted projective space[36]



See also



  • Algebraic geometry of projective spaces

  • Hilbert scheme

  • Lefschetz hyperplane theorem

  • Minimal model program



Notes





  1. ^ Kollár & Moduli, Ch I.


  2. ^ Shafarevich, Igor R. (1994), Basic Algebraic Geometry 1: Varieties in Projective Space, Springer.mw-parser-output cite.citation{font-style:inherit}.mw-parser-output .citation q{quotes:"""""""'""'"}.mw-parser-output .citation .cs1-lock-free a{background:url("//upload.wikimedia.org/wikipedia/commons/thumb/6/65/Lock-green.svg/9px-Lock-green.svg.png")no-repeat;background-position:right .1em center}.mw-parser-output .citation .cs1-lock-limited a,.mw-parser-output .citation .cs1-lock-registration a{background:url("//upload.wikimedia.org/wikipedia/commons/thumb/d/d6/Lock-gray-alt-2.svg/9px-Lock-gray-alt-2.svg.png")no-repeat;background-position:right .1em center}.mw-parser-output .citation .cs1-lock-subscription a{background:url("//upload.wikimedia.org/wikipedia/commons/thumb/a/aa/Lock-red-alt-2.svg/9px-Lock-red-alt-2.svg.png")no-repeat;background-position:right .1em center}.mw-parser-output .cs1-subscription,.mw-parser-output .cs1-registration{color:#555}.mw-parser-output .cs1-subscription span,.mw-parser-output .cs1-registration span{border-bottom:1px dotted;cursor:help}.mw-parser-output .cs1-ws-icon a{background:url("//upload.wikimedia.org/wikipedia/commons/thumb/4/4c/Wikisource-logo.svg/12px-Wikisource-logo.svg.png")no-repeat;background-position:right .1em center}.mw-parser-output code.cs1-code{color:inherit;background:inherit;border:inherit;padding:inherit}.mw-parser-output .cs1-hidden-error{display:none;font-size:100%}.mw-parser-output .cs1-visible-error{font-size:100%}.mw-parser-output .cs1-maint{display:none;color:#33aa33;margin-left:0.3em}.mw-parser-output .cs1-subscription,.mw-parser-output .cs1-registration,.mw-parser-output .cs1-format{font-size:95%}.mw-parser-output .cs1-kern-left,.mw-parser-output .cs1-kern-wl-left{padding-left:0.2em}.mw-parser-output .cs1-kern-right,.mw-parser-output .cs1-kern-wl-right{padding-right:0.2em}


  3. ^ This homogeneous ideal is sometimes called the homogenization of I.


  4. ^ Mumford 1999, pg. 82


  5. ^ Hartshorne 1977, Section II.5


  6. ^ Mumford 1999, pg. 111


  7. ^ This definition differs from Eisenbud–Harris 2000, III.2.3 but is consistent with the other parts of Wikipedia.


  8. ^ cf. the proof of Hartshorne 1977, Ch II, Theorem 7.1


  9. ^ Grothendieck & Dieudonné 1961, 5.6


  10. ^ Hartshorne 1977, Ch II. Exercise 4.5


  11. ^ Humphreys, James (1981), Linear algebraic groups, Springer, Theorem 21.3


  12. ^ Hartshorne, Ch. V, Exercise 3.4. (e).


  13. ^ Fulton 1998, Proposition 8.4.


  14. ^ Hartshorne, Ch. II, Exercise 5.14. (a)


  15. ^ Rosen, Michael (2002), Number theory in Function Fields, Springer


  16. ^ Hartshorne, 1977 & Ch IV, Exercise 1.7.


  17. ^ Hartshorne 1977, Ch I, Exercise 2.8; this is because the homogeneous coordinate ring of Pn{displaystyle mathbb {P} ^{n}}mathbb {P} ^{n} is a unique factorization domain and in a UFD every prime ideal of height 1 is principal.


  18. ^ Shafarevich 1994, Ch. I. § 4.4. Example 1.


  19. ^ Mumford, Ch. II, § 7. Proposition 6.


  20. ^ Hartshorne, Ch. I, Exercise 4.9.


  21. ^ Fulton, § 4.4.


  22. ^ This is not difficult:(Hartshorne 1977, Ch III. Lemma 2.10) consider a flasque resolution of F{displaystyle {mathcal {F}}}{mathcal {F}} and its zero-extension to the whole projective space.


  23. ^ Hartshorne 1977, Ch III. Theorem 5.2


  24. ^ Hartshorne 1977, Ch III. Exercise 5.2


  25. ^ Hartshorne 1977, Ch IV. Theorem 1.3


  26. ^ Kollár 1996, Ch I 1.4


  27. ^ To make the construction work, one needs to allow for a non-variety.


  28. ^ Eisenbud & Harris 2000, VI 2.2


  29. ^ Hartshorne 1977, Appendix B. Theorem 3.4.


  30. ^ Griffiths-Adams, IV. 1. 10. Corollary H


  31. ^ Griffiths-Adams, IV. 1. 10. Corollary I


  32. ^ Hartshorne 1977, Appendix B. Theorem 2.1


  33. ^ Mumford 1970, pg. 36


  34. ^ Hartshorne 1977, Ch III. Remark 7.15.


  35. ^ Esnault, Hélène; Viehweg, Eckart (1992), Lectures on vanishing theorems, Birkhäuser


  36. ^ Dolgachev, Igor (1982), "Weighted projective varieties", Group actions and vector fields (Vancouver, B.C., 1981), Lecture Notes in Math., 956, Berlin: Springer, pp. 34–71, CiteSeerX 10.1.1.169.5185, doi:10.1007/BFb0101508, ISBN 978-3-540-11946-3, MR 0704986




References




  • Eisenbud, David; Harris, Joe (2000), The geometry of schemes


  • William Fulton. (1998), Intersection theory, Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge., 2 (2nd ed.), Berlin, New York: Springer-Verlag, ISBN 978-3-540-62046-4, MR 1644323

  • P. Griffiths and J. Adams, Topics in algebraic and analytic geometry, Princeton University Press, Princeton, N.J., 1974.


  • Hartshorne, Robin (1977), Algebraic Geometry, Graduate Texts in Mathematics, 52, New York: Springer-Verlag, ISBN 978-0-387-90244-9, MR 0463157


  • Huybrechts, Daniel (2005). Complex Geometry: An Introduction. Springer. ISBN 978-3-540-21290-4.


  • Grothendieck, Alexandre; Dieudonné, Jean (1961). "Éléments de géométrie algébrique: II. Étude globale élémentaire de quelques classes de morphismes". Publications Mathématiques de l'IHÉS. 8. doi:10.1007/bf02699291. MR 0217084.


  • Kollár, János, Book on Moduli of Surfaces


  • Kollár, János (1996), Rational curves on algebraic varieties


  • Mumford, David (1970), Abelian Varieties


  • Mumford, David (1995), Algebraic Geometry I: Complex Projective Varieties


  • Mumford, David (1999), The Red Book of Varieties and Schemes: Includes the Michigan Lectures (1974) on Curves and Their Jacobians, Lecture Notes in Mathematics, 1358 (2nd ed.), Springer-Verlag, doi:10.1007/b62130, ISBN 978-3540632931

  • Mumfords's "Algebraic Geometry II", coauthored with Tadao Oda: available at [1]


  • Igor Shafarevich (1995). Basic Algebraic Geometry I: Varieties in Projective Space (2nd ed.). Springer-Verlag. ISBN 978-0-387-54812-8.

  • R. Vakil, Foundations Of Algebraic Geometry



External links




  • The Hilbert Scheme by Charles Siegel - a blog post

  • Projective varieties Ch. 1




這個網誌中的熱門文章

Tangent Lines Diagram Along Smooth Curve

Yusuf al-Mu'taman ibn Hud

Zucchini