This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

On algebraic and non-algebraic neighborhoods of rational curves

Serge Lvovski National Research University Higher School of Economics, Moscow, Russia Federal State Institution ”Scientific-Research Institute for System Analysis of the Russian Academy of Sciences” (SRISA) [email protected]
Abstract.

We prove that for any d>0d>0 there exists an embedding of the Riemann sphere 1\mathbb{P}^{1} in a smooth complex surface, with self-intersection dd, such that the germ of this embedding cannot be extended to an embedding in an algebraic surface but the field of germs of meromorphic functions along CC has transcendence degree 22 over \mathbb{C}. We give two different constructions of such neighborhoods, either as blowdowns of a neighborhood of the smooth plane conic, or as ramified coverings of a neighborhood of a hyperplane section of a surface of minimal degree.

The proofs of non-algebraicity of these neighborhoods are based on a classification, up to isomorphism, of algebraic germs of embeddings of 1\mathbb{P}^{1}, which is also obtained in the paper.

Key words and phrases:
Neighborhoods of rational curves, surfaces of minimal degree, blowup
1991 Mathematics Subject Classification:
32H99, 14J26
This study was partially supported by the HSE University Basic Research Program and by SRISA research project FNEF-2022-0007 (Reg. No 1021060909180-7-1.2.1).

1. Introduction

In this paper we study germs of embeddings of the Riemann sphere aka 1\mathbb{P}^{1} in smooth complex surfaces (see precise definitions in Section 1.1 below). The structure of such germs for which the degree of the normal bundle, which is equal to the self-intersection index of the curve is question, is non-positive, is well known and simple: if (CC)=d0(C\cdot C)=d\leq 0, then for any such dd there exists only one germ up to isomorphism, and these germs are algebraic (see [6] for the negative degree case and [11] for the zero degree case). Germs of embeddings (C,U)(C,U), where C1C\cong\mathbb{P}^{1}, UU is a smooth complex surface, and (CC)>0(C\cdot C)>0, are way more diverse.

Let us say that a germ of neighborhood of C1C\cong\mathbb{P}^{1} is algebraic if it is isomorphic to the germ of embedding of CC in a smooth algebraic surface. M. Mishustin, in his paper [9], showed that the space of isomorphism classes of germs of embeddings of C1C\cong\mathbb{P}^{1}, (CC)>0(C\cdot C)>0, in smooth surfaces, is infinite-dimensional, so one would expect that “most” germs of such embeddings are not algebraic. In this paper we will construct explicit examples of non-algebraic germs of embeddings of 1\mathbb{P}^{1}.

An interesting series of such examples was constructed in a recent paper by M. Falla Luza and F. Loray [4]. For each d>0d>0 they construct an embedding of C1C\cong\mathbb{P}^{1} is a smooth surface such that (CC)=d(C\cdot C)=d and the field of germs of meromorphic functions along CC consists only of constants. For curves on an algebraic surface this is impossible, so the germs of neighborhoods constructed in [4] are examples of non-algebraic germs.

In Section 5.3 of the paper [5] the same authors give an example of a non-algebraic germ of neighborhood of 1\mathbb{P}^{1}, with self-intersection 11, for which the field of germs of meromorphic functions is as big as possible: it has transcendence degree 22 over \mathbb{C}.

The aim of this paper is to construct two series of explicit examples of non-algebraic germs of neighborhoods of 1\mathbb{P}^{1} for which the field of germs of meromorphic functions is as big as possible.

The first of these series is constructed in Section 4. Specifically, for any integer m5m\geq 5 we construct a non-algebraic germ of neighborhood (C,U)(C,U) such that C1C\cong\mathbb{P}^{1}, (CC)=m(C\cdot C)=m, and one can blow up m4m-4 points on CC to obtain the germ of neighborhood of the conic in the plane. The field of germs of meromorphic functions along CC has transcendence degree 22 over \mathbb{C} (Construction 4.10 and Proposition 4.11).

The second series of examples is constructed in Section 5. For any positive integer nn we construct a non-algebraic germ of neighborhood (C,U)(C,U) such that C1C\cong\mathbb{P}^{1}, (CC)=n(C\cdot C)=n, and the field of germs of meromorphic functions along CC has transcendence degree 22 over \mathbb{C}, as a ramified two-sheeted covering of the germ of a neighborhood of 1\mathbb{P}^{1} with self-intersection 2n2n. This construction is a generalization of that from [5, Section 5.3] (and coincides with the latter for n=1n=1), but the method of proof if non-algebraicity is different. See Construction 5.3 and Proposition 5.4.

I do not know whether the germs of neighborhoods constructed in Sections 4 and  5 are isomorphic.

The proofs of non-algebraicity of the neighborhoods constructed in Sections 4 and  5 are based on the classification of algebraic germs of neighborhoods of 1\mathbb{P}^{1}. It turns out that any such germ with self-intersection d>0d>0 is isomorphic to the germ of a hyperplane section of a surface of degree dd in d+1\mathbb{P}^{d+1}; moreover, any isomorphism of germs of such embeddings (C1,F1)(C_{1},F_{1}) and (C2,F2)(C_{2},F_{2}) (where FjF_{j}, j=1,2j=1,2, are surfaces of degree dd in d+1\mathbb{P}^{d+1} and CjC_{j} a hyperplane section of FjF_{j}) is induced by a linear isomorphism between the surfaces F1,F2d+1F_{1},F_{2}\subset\mathbb{P}^{d+1} (Propositions 3.3 and 3.4). The key role in the proofs is played by Lemma 3.6.

The paper is organized as follows. In Section 2 we recall the properties of surfaces of degree dd in d+1\mathbb{P}^{d+1}. In Section 3 we obtain a classification of algebraic neighborhoods of 1\mathbb{P}^{1}. Finally, in Sections 4 and 5 we construct two series of examples of non-algebraic neighborhoods.

Acknowledgements

I am grateful to Frank Loray and Grigory Merzon for useful discussions.

1.1. Notation and conventions

All algebraic varieties are varieties over \mathbb{C}. All topological terminology pertains to the classical (complex) topology.

Suppose we are given the pairs (C1,S1)(C_{1},S_{1}) and (C2,S2)(C_{2},S_{2}), where CjC_{j}, j=1,2j=1,2, are projective algebraic curves contained in smooth complex analytic surfaces SjS_{j}. We will say that these two pairs are isomorphic as germs of neighborhoods if there exists an isomorphism φ:V1V2\varphi\colon V_{1}\to V_{2}, where C1V1S1C_{1}\subset V_{1}\subset S_{1}, C2V2S2C_{2}\subset V_{2}\subset S_{2}, V1V_{1} and V2V_{2} are open, such that φ(C1)=C2\varphi(C_{1})=C_{2}. In this paper we are mostly concerned with germs of neighborhoods, but sometimes, abusing the language, we will write “neighborhood” instead of “germ of neighborhoods”; this should not lead to confusion.

If CC is a projective algebraic curve on a smooth complex analytic surface SS, then a germ of holomorphic (resp. meromorphic) function along SS is an equivalence class of pairs (U,f)(U,f), where UU is a neighborhood of CC in SS, ff is a holomorphic (resp. meromorphic) function on UU, and (U1,f1)(U2,f2)(U_{1},f_{1})\sim(U_{2},f_{2}) if there exists a neighborhood VCV\supset C, VU1U2V\subset U_{1}\cap U_{2}, on which f1f_{1} and f2f_{2} agree. Germs of holomorphic (resp. meromorphic) functions along CSC\subset S form a ring (resp. a field). The field of germs of meromorphic functions along CSC\subset S will be denoted, following the paper [4], by (S,C)\mathcal{M}(S,C). If the self-intersection index (CC)(C\cdot C) is positive, then, according to Theorem 2.1 from [10], the curve CSC\subset S has a fundamental system of pseudoconcave neighborhoods, and it follows from [1, Théorème 5] that tr.deg(S,C)2\operatorname{tr.deg}_{\mathbb{C}}\mathcal{M}(S,C)\leq 2 (here and below, tr.deg\operatorname{tr.deg}_{\mathbb{C}} means “transcendence degree over \mathbb{C}”).

A germ of neighborhood of a projective curve CC will be called algebraic if it is isomorphic, as a germ, to the germ of neighborhood of CC in XX, where XCX\supset C is a smooth projective algebraic surface; since desingularization of algebraic surfaces over \mathbb{C} exists, one may as well say that a germ of neighborhood of CC is algebraic if it is isomorphic to the germ of a neighborhood CXC\subset X, where XX is a an arbitrary smooth algebraic surface, not necessarily projective. In this case, tr.deg(X,C)=2\operatorname{tr.deg}_{\mathbb{C}}\mathcal{M}(X,C)=2 since this field contains the field of rational functions on XX.

Remark 1.1.

A germ of neighborhood of a projective curve CC with positive self-intersection is algebraic if and only if it is isomorphic to the germ of embedding of CC into a compact complex surface. Indeed, any smooth complex surface containing a projective curve with positive self-intersection, can be embedded in N\mathbb{P}^{N} for some NN (see [2, Chapter IV, Theorem 6.2]).

A projective subvariety XNX\subset\mathbb{P}^{N} is called non-degenerate if it does not lie in a hyperplane, and linearly normal if the natural homomorphism from H0(N,𝒪N(1))H^{0}(\mathbb{P}^{N},\mathcal{O}_{\mathbb{P}^{N}}(1)) to H0(X,𝒪X(1))H^{0}(X,\mathcal{O}_{X}(1)) is surjective. If XX is non-degenerate, the latter condition holds if and only if XX is not an isomorphic projection of a non-degenerate subvariety of N+1\mathbb{P}^{N+1}.

Non-degenerate projective subvarieties X1N1X_{1}\subset\mathbb{P}^{N_{1}} and X2N2X_{2}\subset\mathbb{P}^{N_{2}} will be called projectively isomorphic if N1=N2N_{1}=N_{2} and there exists a linear isomorphism f:N1N2f\colon\mathbb{P}^{N_{1}}\to\mathbb{P}^{N_{2}} such that f(X1)=X2f(X_{1})=X_{2} .

If C1C_{1} and C2C_{2} are two projective algebraic curves on a smooth complex surface, then (C1C2)(C_{1}\cdot C_{2}) is their intersection index.

The terms “vector bundle” and “locally free sheaf” will be used interchangeably.

If CC is a smooth curve on a smooth complex surface XX, then 𝒩X|C\mathcal{N}_{X|C} is the normal bundle to CC in XX.

If \mathcal{E} is a vector bundle on an algebraic variety, then ()\mathbb{P}(\mathcal{E}) (the projectivisation of \mathcal{E}) is the algebraic variety such that its points are lines in the fibers of \mathcal{E}.

If \mathcal{F} is a coherent sheaf on a complex space XX, we will sometimes write hi()h^{i}(\mathcal{F}) instead of dimHi(X,)\dim H^{i}(X,\mathcal{F}).

By definition, a projective surface XX has rational singularities if p𝒪X¯=𝒪Xp_{*}\mathcal{O}_{\bar{X}}=\mathcal{O}_{X} and R1p𝒪X¯=0R^{1}p_{*}\mathcal{O}_{\bar{X}}=0 for some (hence, any) desingularization p:X¯Xp\colon\bar{X}\to X.

If f:STf\colon S\to T is a dominant morphism of smooth projective algebraic surfaces, we will say that its critical locus is the set

R=f({xS:dfx is degenerate})T,R=f(\{x\in S\colon\text{$df_{x}$ is degenerate}\})\subset T,

and that its branch divisor BTB\subset T is the union of one-dimensional components of RR.

2. Recap on surfaces of minimal degree

In this section we will recall, without proofs, some well-known results. Most of the details can be found in [3].

If XNX\subset\mathbb{P}^{N} is a non-degenerate irreducible projective variety, then

(1) degXcodimX+1,\deg X\geq\operatorname{codim}X+1,

and there exists a classification of the varieties for which the lower bound (1) is attained. We reproduce this classification for the case dimX=2\dim X=2.

Notation 2.1.

For any integer n0n\geq 0, put 𝔽n=(𝒪1𝒪1(n))\mathbb{F}_{n}=\mathbb{P}(\mathcal{O}_{\mathbb{P}^{1}}\oplus\mathcal{O}_{\mathbb{P}^{1}}(n)).

The surface 𝔽0\mathbb{F}_{0} is just the quadric 1×1\mathbb{P}^{1}\times\mathbb{P}^{1}; if n>0n>0, then the natural projection 𝔽n1\mathbb{F}_{n}\to\mathbb{P}^{1} has a unique section E𝔽nE\subset\mathbb{F}_{n} such that (EE)=n(E\cdot E)=-n. The section EE will be called the exceptional section of 𝔽n\mathbb{F}_{n}. The divisor class group of 𝔽n\mathbb{F}_{n} is generated by the class ee of the exceptional section and the class ff of the fiber (if n=0n=0, we denote by ee and ff the classes of lines of two rulings on 1×1\mathbb{P}^{1}\times\mathbb{P}^{1}). One has

(2) (ee)=n,(ef)=1,(ff)=0.(e\cdot e)=-n,\quad(e\cdot f)=1,\quad(f\cdot f)=0.

If r0r\geqslant 0 is an integer and (n,r)(0,0)(n,r)\neq(0,0), then the complete linear system |e+(n+r)f||e+(n+r)f| on 𝔽n\mathbb{F}_{n} has no basepoints and defines a mapping 𝔽nn+2r+1\mathbb{F}_{n}\to\mathbb{P}^{n+2r+1}.

Notation 2.2.

If n,rn,r are non-negative integers and (n,r)(0,0)(n,r)\neq(0,0), then by Fn,rn+2r+1F_{n,r}\subset\mathbb{P}^{n+2r+1} we will denote the image of the mapping 𝔽nn+2r+1\mathbb{F}_{n}\to\mathbb{P}^{n+2r+1} defined by the complete linear system |e+(n+r)f||e+(n+r)f|.

It follows from (2) that the variety Fn,rn+2r+1F_{n,r}\subset\mathbb{P}^{n+2r+1} is a surface of degree n+2rn+2r. If r>0r>0, the surface Fn,rF_{n,r} is smooth and isomorphic to 𝔽n\mathbb{F}_{n}. The surface Fn,0F_{n,0} with n2n\geq 2 is the cone over the normal rational curve of degree nn in n\mathbb{P}^{n} (this cone is obtained by contracting the exceptional section on 𝔽n\mathbb{F}_{n}), and the surface F1,0F_{1,0} is just the plane; the surfaces Fn,0F_{n,0} are normal and, moreover, have rational singularities. If n>0n>0 and r>0r>0, then the exceptional section on Fn,rF_{n,r} is a rational curve of degree rr. Finally, the surface F0,13F_{0,1}\subset\mathbb{P}^{3} is the smooth quadric.

Recall that the quadratic Veronese surface V5V\subset\mathbb{P}^{5} is the image of the mapping v:25v\colon\mathbb{P}^{2}\to\mathbb{P}^{5} defined by the formula

(3) v:(x0:x1:x2)(x02:x12:x22:x0x1:x0x2:x1x2);v\colon(x_{0}:x_{1}:x_{2})\mapsto(x_{0}^{2}:x_{1}^{2}:x_{2}^{2}:x_{0}x_{1}:x_{0}x_{2}:x_{1}x_{2});

one has degV=4\deg V=4. The mapping vv induces an isomorphism from 2\mathbb{P}^{2} to VV; hyperplane sections of VV are images of conics in 2\mathbb{P}^{2}.

Proposition 2.3.

If XNX\subset\mathbb{P}^{N} is a non-degenerate irreducible projective surface, then degXN1\deg X\geq N-1 and the bound is attained if and only if either X=Fn,rX=F_{n,r}, where (n,r)(0,0)(n,r)\neq(0,0), or X=V5X=V\subset\mathbb{P}^{5}, where VV is the quadratic Veronese surface.

If (n,r)(n,r)(n,r)\neq(n^{\prime},r^{\prime}), then the surfaces Fn,rF_{n,r} and Fn,rF_{n^{\prime},r^{\prime}} are not projectively isomorphic, and none of them is projectively isomorphic to the Veronese surface VV. (Indeed, if nnn\neq n^{\prime} then Fn,rF_{n,r} and Fn,rF_{n^{\prime},r^{\prime}} are not isomorphic even as abstract varieties, and if n=nn=n^{\prime} but rrr\neq r^{\prime} then their degrees are different; speaking of the Veronese surface, it does not contain lines while each Fn,rF_{n,r} is swept by lines.)

Suppose now that XN+1X\subset\mathbb{P}^{N+1}, N2N\geq 2, is a surface of minimal degree (i.e., of degree NN) and pFp\in F is a smooth point (i.e., either XFn,0X\neq F_{n,0} and pp is arbitrary or X=Fn,0X=F_{n,0} and pp is not the vertex of the cone). Let πp\pi_{p} denote the projection from N+1\mathbb{P}^{N+1} to N\mathbb{P}^{N} with the center pp.

Proposition 2.4.

In the above setting, the projection πp\pi_{p} induces a birational mapping from XX onto its image. If XNX^{\prime}\subset\mathbb{P}^{N} is (the closure of) the image of πp:XN\pi_{p}\colon X\dasharrow\mathbb{P}^{N}, then XX^{\prime} is also a surface of minimal degree.

If X=Fn,rX=F_{n,r}, where n>0n>0 and r>0r>0, then X=Fn+1,r1X^{\prime}=F_{n+1,r-1} if pp lies on the exceptional section and X=Fn1,rX^{\prime}=F_{n-1,r} otherwise.

If X=F0,rX=F_{0,r}, r>0r>0, then X=F1,r1X^{\prime}=F_{1,r-1}.

If X=Fn,0X=F_{n,0}, n2n\geq 2, then X=Fn1,0X^{\prime}=F_{n-1,0}.

Finally, if X=V5X=V\subset\mathbb{P}^{5}, then X=F1,14X^{\prime}=F_{1,1}\subset\mathbb{P}^{4}.

These projections for surfaces of minimal degree 6\leq 6 are depicted in Figure 1. Observe that no arrow points at the surface VV; this fact will play a crucial role in the sequel.

degree6\textstyle{6}F6,0\textstyle{F_{6,0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}F4,1\textstyle{F_{4,1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}F2,2\textstyle{F_{2,2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}F0,3\textstyle{F_{0,3}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}5\textstyle{5}F5,0\textstyle{F_{5,0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}F3,1\textstyle{F_{3,1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}F1,2\textstyle{F_{1,2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}4\textstyle{4}F4,0\textstyle{F_{4,0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}F2,1\textstyle{F_{2,1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}F0,2\textstyle{F_{0,2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}V\textstyle{V\ignorespaces\ignorespaces\ignorespaces\ignorespaces}3\textstyle{3}F3,0\textstyle{F_{3,0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}F1,1\textstyle{F_{1,1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}2\textstyle{2}F2,0\textstyle{F_{2,0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}F0,1\textstyle{F_{0,1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1\textstyle{1}F1,0\textstyle{F_{1,0}}
Figure 1. Projections of surfaces of minimal degree

The birational transformations induced by the projections form Proposition 2.4 can be described explicitly.

Proposition 2.5.

Suppose that FNF\subset\mathbb{P}^{N} is a surface of minimal degree, pFp\in F is a non-singular point, and πp:FN1\pi_{p}\colon F\dasharrow\mathbb{P}^{N-1} is the projection from pp. The projection πp\pi_{p} acts on the surface FF as follows.

  1. (1)

    If F=Fn,rn+2r+1F=F_{n,r}\subset\mathbb{P}^{n+2r+1}, where r>1r>1, then the projection πp\pi_{p} blows up the point pp and blows down the strict transform of the only line on FF passing through pp.

  2. (2)

    If F=Fn,1n+3F=F_{n,1}\subset\mathbb{P}^{n+3}, where n>0n>0 (in this case the exceptional section of F𝔽nF\cong\mathbb{F}_{n} is a line), then two cases are possible.

    If the point pp does not lie on the exceptional section, then the projection πp\pi_{p} blows up the point pp and blows down the strict transform of the only line passing through pp, and the image of the projection is the surface Fn1,1n+2F_{n-1,1}\subset\mathbb{P}^{n+2}. If, on the other hand, pp lies on the exceptional section, then the projection πp\pi_{p} blows up the point pp and blows down both the strict transform of the fiber passing through pp and the strict transform of the exceptional section; in this latter case the image of the projection is the cone Fn+1,0n+2F_{n+1,0}\subset\mathbb{P}^{n+2} (the strict transform of the exceptional section is blown down to the vertex of the cone).

  3. (3)

    If F=F0,13F=F_{0,1}\subset\mathbb{P}^{3} (that is, if F3F\subset\mathbb{P}^{3} is the smooth quadric), then the projection πp\pi_{p} blows up the point pp and blows down strict transforms of the two lines passing through pp; the image of the projection is, of course, just the plane.

  4. (4)

    If F=Fn,0n+1F=F_{n,0}\subset\mathbb{P}^{n+1}, where n>1n>1 (that is, if FF is a cone over the rational normal curve of degree nn in n\mathbb{P}^{n}), then the projection πp\pi_{p} blows up the point pp and blows down the strict transform of the generatrix of the cone passing through pp; the image of this projection is the cone Fn1,0nF_{n-1,0}\subset\mathbb{P}^{n} (if n=2n=2, this cone is just the plane).

  5. (5)

    Finally, if F=V5F=V\subset\mathbb{P}^{5} is the Veronese surface, then the projection πp\pi_{p} just blows up the point pp, and the image of this projection is F1,13F_{1,1}\subset\mathbb{P}^{3}.

3. Rational curves with positive self-intersection and algebraic germs

Proposition 3.1.

Suppose that XX is a smooth projective surface and CXC\subset X is a curve isomorphic to 1\mathbb{P}^{1}. If (CC)=m>0(C\cdot C)=m>0, then the complete linear system |C||C| has no basepoints, dim|C|=m+1\dim|C|=m+1, the morphism φ|C|\varphi_{|C|} is a birational isomorphism between XX and φ(X)m+1\varphi(X)\subset\mathbb{P}^{m+1}, and φ(X)\varphi(X) is a surface in m+1\mathbb{P}^{m+1} of minimal degree mm.

Proof.

Since the normal bundle 𝒩X|C\mathcal{N}_{X|C} is isomorphic to 𝒪1(m)\mathcal{O}_{\mathbb{P}^{1}}(m), where m>0m>0, one has h0(𝒩X|C)=m+1h^{0}(\mathcal{N}_{X|C})=m+1, h1(𝒩X|C)=0h^{1}(\mathcal{N}_{X|C})=0, so the Hilbert scheme of the curve CXC\subset X is smooth and has dimension m+1m+1 at the point corresponding to CC. A general curve from this (m+1)(m+1)-dimensional family is isomorphic to 1\mathbb{P}^{1}; since m>0m>0, through a general point pXp\in X there passes a positive-dimensional family of rational curves. Therefore, the Albanese mapping of XX is constant, whence H1(X,𝒪X)=0H^{1}(X,\mathcal{O}_{X})=0. Now it follows from the exact sequence

(4) 0𝒪X𝒪X(C)𝛼𝒪C(C)00\to\mathcal{O}_{X}\to\mathcal{O}_{X}(C)\xrightarrow{\alpha}\mathcal{O}_{C}(C)\to 0

that the homomorphism α:H0(X,𝒪X(C))H0(C,𝒪C(C))\alpha_{*}\colon H^{0}(X,\mathcal{O}_{X}(C))\to H^{0}(C,\mathcal{O}_{C}(C)) is surjective. Since the linear system |𝒪C(C)|=|𝒪1(m)||\mathcal{O}_{C}(C)|=|\mathcal{O}_{\mathbb{P}^{1}}(m)| has no basepoints, the linear system |C||C| has no basepoints either, and it follows from (4) and the vanishing of H1(𝒪X)H^{1}(\mathcal{O}_{X}) that dim|C|=m+1\dim|C|=m+1. If φ=φ|C|:Xm+1\varphi=\varphi_{|C|}\colon X\to\mathbb{P}^{m+1} and Y=φ(X)Y=\varphi(X), then dimY=2\dim Y=2 and degYdegφ=(CC)=m\deg Y\cdot\deg\varphi=(C\cdot C)=m. Since Ym+1Y\subset\mathbb{P}^{m+1} is non-degenerate, one has degYm\deg Y\geq m (see (1)), whence degY=m\deg Y=m and degφ=1\deg\varphi=1, so φ\varphi is birational onto its image. ∎

Corollary 3.2.

If XX is a projective surface with rational singularities and CXsmC\subset X_{\mathrm{sm}} is a curve that is isomorphic to 1\mathbb{P}^{1} and such that (CC)>0(C\cdot C)>0, then h1(𝒪X)=0h^{1}(\mathcal{O}_{X})=0 and h0(𝒪X(C))=(CC)+1h^{0}(\mathcal{O}_{X}(C))=(C\cdot C)+1.

Proof.

Let X¯\bar{X} be a desingularization of XX. Arguing as in the proof of Proposition 3.1, we conclude that through a general point of X¯\bar{X} there passes a positive-dimensional family of rational curves, whence H1(X¯,𝒪X¯)=0H^{1}(\bar{X},\mathcal{O}_{\bar{X}})=0. Since the singularities of the surface XX are rational, H1(X¯,𝒪X¯)H1(X,𝒪X)H^{1}(\bar{X},\mathcal{O}_{\bar{X}})\cong H^{1}(X,\mathcal{O}_{X}), so H1(X,𝒪X)=0H^{1}(X,\mathcal{O}_{X})=0. Now the result follows from the exact sequence (4). ∎

Proposition 3.1 implies the following characterisation of algebraic neighborhoods of rational curves.

Proposition 3.3.

If (C,U)(C,U) is an algebraic neighborhood of the curve C1C\cong\mathbb{P}^{1} and if (CC)=d>0(C\cdot C)=d>0, then the germ of this neighborhood is isomorphic to the germ of neighborhood of a smooth hyperplane section in a surface of minimal degree dd in d+1\mathbb{P}^{d+1}.

Proof.

Passing to desingularization, one may without loss of generality assume that the neighborhood in question is a neighborhood of a curve CXC\subset X, where C1C\cong\mathbb{P}^{1}, XX is a smooth projective surface, and (CC)=d>0(C\cdot C)=d>0. If φ:XY\varphi\colon X\to Y, where YY is a surface of minimal degree, is the birational morphism the existence of which is asserted by Proposition 3.1, then φ\varphi is an isomorphism in a neighborhood of CC and φ(C)\varphi(C) is a hyperplane section of YY, whence the result. ∎

Proposition 3.3 may be regarded as a generalization of Proposition 4.7 from [8], which asserts that any algebraic germ of neighborhood of 1\mathbb{P}^{1} with self-intersection 11 is isomorphic to the germ of neighborhood of a line in 2\mathbb{P}^{2}.

Now we show that not only all germs of algebraic neighborhoods of 1\mathbb{P}^{1} can be obtained from surfaces of minimal degree, but that their isomorphisms are induced by isomorphisms of surfaces of minimal degree.

Proposition 3.4.

Suppose that X1N1X_{1}\subset\mathbb{P}^{N_{1}} and X2N2X_{2}\subset\mathbb{P}^{N_{2}} are linearly normal projective surfaces with rational singularities, C1X1C_{1}\subset X_{1} and C2X2C_{2}\subset X_{2} are their smooth hyperplane sections, and that C1C_{1} and C2C_{2} are isomorphic to 1\mathbb{P}^{1}.

If there exist analytic neighborhoods UjCjU_{j}\supset C_{j}, j=1,2j=1,2, and a holomorphic isomorphism φ:U1U2\varphi\colon U_{1}\to U_{2} such that φ(C1)=C2\varphi(C_{1})=C_{2}, then φ\varphi extends to a projective isomorphism Φ:X1X2\Phi\colon X_{1}\to X_{2}.

To prove this proposition we need two lemmas. The first of them is well known.

Lemma 3.5.

If XX is a projective surface with isolated singularities and CXsmC\subset X_{\mathrm{sm}} is an ample irreducible curve, then the ring of germs of holomorphic functions along CC coincides with \mathbb{C}.

Sketch of proof.

This follows immediately from the fact that H0(X^,𝒪X^)=H^{0}(\hat{X},\mathcal{O}_{\hat{X}})=\mathbb{C}, where X^\hat{X} is the formal completion of XX along CC (see [7, Chapter V, Proposition 1.1 and Corollary 2.3]).

Here is a more elementary argument. Since CC is an ample divisor in XX, there exists and embedding of XX in N\mathbb{P}^{N} such that rCrC, for some r>0r>0, is a hyperplane section of XX. Suppose that UCU\supset C, UXU\subset X is a connected neighborhood of CC. There exists a family of hyperplane sections {Hα}\{H_{\alpha}\}, close to the one corresponding to rCrC, such that HαUH_{\alpha}\subset U for each α\alpha and αHα\bigcup_{\alpha}H_{\alpha} contains a non-empty open subset VUV\subset U. If ff is a holomorphic function on UU, then ff is constant on each HαH_{\alpha}; since HαHαH_{\alpha}\cap H_{\alpha^{\prime}}\neq\varnothing for each α\alpha and α\alpha^{\prime}, ff is constant on the union of all the HαH_{\alpha}’s, hence on VV, hence on UU. ∎

Lemma 3.6.

Suppose that XX is a projective surface such that H1(X,𝒪X)=0H^{1}(X,\mathcal{O}_{X})=0, DXsmD\subset X_{\mathrm{sm}} is an ample irreducible projective curve, and rr is a positive integer. Then any germ of meromorphic function along CC, with possibly a pole of order r\leq r along CC and no other poles, is induced by a rational function on XX with possibly a pole of order r\leq r along CC and no other poles.

Proof.

Let C𝒪X\mathcal{I}_{C}\subset\mathcal{O}_{X} be the ideal sheaf of CC; put rC=Cm\mathcal{I}_{rC}=\mathcal{I}_{C}^{m}, 𝒪rC=𝒪X/Cm\mathcal{O}_{rC}=\mathcal{O}_{X}/\mathcal{I}_{C}^{m}, and

𝒩X|rC=Hom¯𝒪rC(rC/rC2,𝒪rC)=Hom¯𝒪X(rC,𝒪rC).\mathcal{N}_{X|rC}=\underline{\mathrm{Hom}}_{\mathcal{O}_{rC}}(\mathcal{I}_{rC}/\mathcal{I}_{rC}^{2},\mathcal{O}_{rC})=\underline{\mathrm{Hom}}_{\mathcal{O}_{X}}(\mathcal{I}_{rC},\mathcal{O}_{rC}).

Identifying 𝒪X(rC)\mathcal{O}_{X}(rC) with the sheaf of meromorphic functions having at worst a pole of order r\leq r along CC, one has the exact sequence

(5) 0𝒪X𝒪X(rC)𝛼𝒩X|rC0,0\to\mathcal{O}_{X}\to\mathcal{O}_{X}(rC)\xrightarrow{\alpha}\mathcal{N}_{X|rC}\to 0,

in which the homomorphism α\alpha has the form

g(sgsmodrC),g\mapsto(s\mapsto gs\bmod\mathcal{I}_{rC}),

where gg is a meromorphic function on an open subset of VXV\subset X, with at worst a pole of order r\leq r along CC and ss is a section of rC\mathcal{I}_{rC} over VV. In particular, if UCU\supset C is a neighborhood, then any section of gH0(U,𝒪X(rC))g\in H^{0}(U,\mathcal{O}_{X}(rC)) induces a global section of 𝒩X|rC\mathcal{N}_{X|rC}. Since H1(X,𝒪X)=0H^{1}(X,\mathcal{O}_{X})=0, the homomorphism α\alpha from (5) induces a surjection on global sections, so there exists a section fH0(X,𝒪X(rC))f\in H^{0}(X,\mathcal{O}_{X}(rC)) such that ff and gg induce the same global section of 𝒩X|rC\mathcal{N}_{X|rC}. Hence, the meromorphic function (f|U)g(f|_{U})-g has no pole in UU, so by virtue of Lemma 3.5 this function is equal to a constant cc on a (possibly smaller) neighborhood of CC. Thus, the germ of fcf-c along CC equals that of gg. ∎

Proof of Proposition 3.4.

It follows from the hypothesis that (C1C1)=(C2C2)(C_{1}\cdot C_{1})=(C_{2}\cdot C_{2}). If we denote these intersection indices by mm, then Corollary 3.2 implies that dim|C1|=dim|C2|=m+1\dim|C_{1}|=\dim|C_{2}|=m+1. Let f0,,fm+1f_{0},\dots,f_{m+1} be a basis of H0(𝒪X1(C1))H^{0}(\mathcal{O}_{X_{1}}(C_{1})) (i.e., the basis of space of meromorphic functions on X1X_{1} with at worst a simple pole along C1C_{1}), and similarly let (g0,,gm+1)(g_{0},\dots,g_{m+1}) be a basis of H0(𝒪X2(C2))H^{0}(\mathcal{O}_{X_{2}}(C_{2})). Embed X1X_{1} (resp. X2X_{2}) into m+1\mathbb{P}^{m+1} with the linear system |C1||C_{1}| (resp. |C2||C_{2}|), that is, with the mappings

x(f0(x)::fn+1(x))andy(g0(y)::gn+1(y)).x\mapsto(f_{0}(x):\dots:f_{n+1}(x))\quad\text{and}\quad y\mapsto(g_{0}(y):\dots:g_{n+1}(y)).

If γi\gamma_{i}, 0im+10\leq i\leq m+1, is the germ along C1C_{1} of the meromorphic function giφg_{i}\circ\varphi, then by virtue of Lemma 3.6, which we apply in the case r=1r=1, each γi\gamma_{i} is the germ along C1C_{1} of a meromorphic function hiH0(𝒪X1(C1))h_{i}\in H^{0}(\mathcal{O}_{X_{1}}(C_{1})). If hi=aijfjh_{i}=\sum a_{ij}f_{j}, then the matrix aij\|a_{ij}\| defines a linear automorphism Φ:m+1m+1\Phi\colon\mathbb{P}^{m+1}\to\mathbb{P}^{m+1} such that its restriction to a neighborhood of C1C_{1} coincides with Φ\Phi. Hence, Φ\Phi maps X1X_{1} isomorphically onto X2X_{2}. ∎

Corollary 3.7.

Suppose that FmF\subset\mathbb{P}^{m} (resp. FmF^{\prime}\subset\mathbb{P}^{m^{\prime}}) is a surface of minimal degree and CC (resp. CC^{\prime}) is its smooth hyperplane section. Then the germs of neighborhoods of CC in FF and of CC^{\prime} in FF^{\prime} are isomorphic if and only if there exists a linear isomorphism Φ:mm\Phi\colon\mathbb{P}^{m}\to\mathbb{P}^{m^{\prime}} such that Φ(F)=F\Phi(F)=F^{\prime} and Φ(C)=C\Phi(C)=C^{\prime}.

Proof.

Immediate from Proposition 3.4 if one takes into account that all surfaces of minimal degree have rational singularities. ∎

Remark 3.8.

We see that any algebraic neighborhood of C1C\cong\mathbb{P}^{1} has one more discrete invariant, besides the self-intersection d=(CC)d=(C\cdot C): if this neighborhood is isomorphic to the germ of neighborhood of the minimal surface Fn,rF_{n,r}, where d=n+2rd=n+2r, this is the integer n0n\geq 0 (and if the surface is not Fn,rF_{n,r} but the Veronese surface V5V\subset\mathbb{P}^{5}, we assign to our neighborhood the tag VV instead). It should be noted however that, as a rule, the pair (d,n)(d,n) does not determine the germ of neighborhood up to isomorphism. Indeed, the dimension of the group of automorphisms of the surface 𝔽n\mathbb{F}_{n} is n+5n+5 if n>0n>0 and 66 if n=0n=0. In most cases this is less that the dimension of the space of hyperplanes in n+2r+1\mathbb{P}^{n+2r+1}, in which Fn,rF_{n,r} is embedded. On the other hand, linear automorphisms of Fn,0n+1F_{n,0}\subset\mathbb{P}^{n+1} act transitively on the set of smooth hyperplane sections of Fn,0F_{n,0}, and ditto for V5V\subset\mathbb{P}^{5}. Thus, tags (n,0)(n,0) or VV do determine an algebraic germ of neighborhood of 1\mathbb{P}^{1} up to isomorphism.

4. Blowups and blowdowns

Suppose that a smooth projective curve CC lies on a smooth complex analytic surface SS and that pCp\in C. Let σ:S~S\sigma\colon\tilde{S}\to S be the blowup of SS at pp, and let C~S~\tilde{C}\subset\tilde{S} be the strict transform of CC. It is clear that the germ of neighborhood C~S~\tilde{C}\subset\tilde{S} depends only on the germ of neighborhood CSC\subset S and on the point pCp\in C.

Definition 4.1.

In the above setting, the germ of neighborhood (C~,U~)(\tilde{C},\tilde{U}) will be called the blowup of the germ (C,U)(C,U) at the point pp.

For future reference we state the following obvious properties of blowups of neighborhoods.

Proposition 4.2.

Suppose that the germ of neighborhoods (C~,U~)(\tilde{C},\tilde{U}) is a blowup of (C,U)(C,U). Then

  1. (1)

    C~\tilde{C} is isomorphic to CC;

  2. (2)

    (C~C~)=(CC)1(\tilde{C}\cdot\tilde{C})=(C\cdot C)-1;

  3. (3)

    if (C,U)(C,U) is algebraic then (C~,U~)(\tilde{C},\tilde{U}) is algebraic.

For algebraic neighborhoods of 1\mathbb{P}^{1} the assertion (3) of Proposition 4.2 can be made more explicit.

Proposition 4.3.

If an algebraic germ (C,U)(C,U) is isomorphic to the germ of neighborhood of a hyperplane section of a surface of minimal degree FNF\subset\mathbb{P}^{N}, then the blowup of this germ at a point pCp\in C is isomorphic to the germ of neighborhood of a hyperplane section of the surface FN1F^{\prime}\subset\mathbb{P}^{N-1} that is obtained from FF by projection from the point pp.

Proof.

Immediate from Proposition 2.5. ∎

Remark 4.4.

Even though 1\mathbb{P}^{1} is homogeneous, which means that its points are indistinguishable, germs of blowups of a given neighborhood of 1\mathbb{P}^{1} at different points are not necessarily isomorphic. Indeed, suppose that C1C\cong\mathbb{P}^{1} is a smooth hyperplane section of the surface Fn,rF_{n,r}, where n>0n>0 and r>0r>0. If pCp\in C does not lie on the exceptional section EFn,rE\subset F_{n,r}, then Proposition 2.4 implies that the blowup at pp of the germ of neighborhood of CC is isomorphic to a neighborhood of a hyperplane section of Fn1,rF_{n-1,r}, and if pp does lie on EE, Proposition 2.4 implies that the blowup in question is isomorphic to a neighborhood of a hyperplane section of Fn+1,r1F_{n+1,r-1} (observe that CEC\neq E and CEC\cap E\neq\varnothing, so points of both kinds are present). If the blowups at such points were isomorphic as germs of neighborhoods, then, by virtue of Proposition 3.4, this isomorphism would be induced by a linear isomorphism between Fn1,rF_{n-1,r} and Fn+1,r1F_{n+1,r-1}, which does not exist.

Proposition 4.5.

Suppose that a curve D1D\cong\mathbb{P}^{1} is embedded in a surface UU. If, blowing up s>0s>0 points of the germ of neighborhood (D,U)(D,U), one obtains a germ of neighborhood (C,W)(C,W) that is isomorphic to the germ of neighborhood of a non-degenerate conic in 2\mathbb{P}^{2}, then the original germ (D,U)(D,U) is not algebraic.

Proof.

Without loss of generality we may and will assume that CC is a conic in 2\mathbb{P}^{2} and WCW\supset C is an open subset of 2\mathbb{P}^{2}. The Veronese mapping v:2Vv\colon\mathbb{P}^{2}\to V (see (3)) identifies 2\mathbb{P}^{2} with the Veronese surface V5V\subset\mathbb{P}^{5} and C2C\subset\mathbb{P}^{2} with a smooth hyperplane section of VV.

Arguing by contradiction, suppose that the germ (D,U)(D,U) is algebraic. Then Proposition 3.3 implies that this germ is isomorphic to the germ of neighborhood of a hyperplane section of a surface of minimal degree m=4+s>4m=4+s>4. Hence, this surface is of the form Fn,rF_{n,r}, where n+2r=mn+2r=m (see Proposition 2.3).

By construction, the germ of (C,W)(C,W) can be obtained from the germ (D,U)(D,U) by blowing up ss points on DD. Hence, by Proposition 4.3, the germ of (C,W)(C,W) is isomorphic to the germ of a hyperplane section of a surface that can be obtained from Fn,rF_{n,r} by ss consecutive projections. However, Proposition 2.4 shows that the resulting surface cannot be projectively isomorphic to V5V\subset\mathbb{P}^{5} (cf. Figure 1). On the other hand, Proposition 3.4 implies that if germs of hyperplane sections of two surfaces of minimal degree are isomorphic then the surfaces are projectively isomorphic. This contradiction shows that the germ (D,U)(D,U) is not algebraic. ∎

Now we can construct our first series of non-algebraic examples.

Lemma 4.6.

Suppose that C2C\subset\mathbb{P}^{2} is a non-degenerate conic and ss is a positive integer. Then there exist ss lines L1,,Ls2L_{1},\ldots,L_{s}\subset\mathbb{P}^{2} and neighborhoods WCW\supset C, WjLjW_{j}\supset L_{j}, 1js1\leq j\leq s, in 2\mathbb{P}^{2} having the following properties.

  1. (1)

    each LjL_{j} intersects the conic CC at precisely two points, pjp_{j} and qjq_{j};

  2. (2)

    all the points p1,,psp_{1},\ldots,p_{s} are distinct, all the points q1,,qsq_{1},\ldots,q_{s} are distinct, and piqjp_{i}\neq q_{j} for any i,ji,j.

  3. (3)

    WjW{p1,,ps,q1,,qs}={pj,qj}W_{j}\cap W\cap\{p_{1},\ldots,p_{s},q_{1},\ldots,q_{s}\}=\{p_{j},q_{j}\} for each jj.

  4. (4)

    The open subset WjWW_{j}\cap W has precisely two connected components PjpjP_{j}\ni p_{j} and QjqjQ_{j}\ni q_{j}.

  5. (5)

    PiPj={P_{i}}\cap{P_{j}}=\varnothing whenever iji\neq j, QiQj={Q_{i}}\cap{Q_{j}}=\varnothing whenever iji\neq j, and PiQj={P_{i}}\cap{Q_{j}}=\varnothing for any i,ji,j.

Proof.

Only the assertions (3) and (4) deserve a sketch of proof. To justify them choose a Hermitian metric on 2\mathbb{P}^{2} and let WW be a small enough tubular neighborhood of CC and W1,,WsW_{1},\ldots,W_{s} be small enough tubular neighborhoods of L1,,LsL_{1},\ldots,L_{s}. ∎

The proof of the following lemma is left to the reader.

Lemma 4.7.

Suppose that X1X_{1} and X2X_{2} are Hausdorff topological spaces, O1X1O_{1}\subset X_{1} and O2X2O_{2}\subset X_{2} are open subsets, and XX is the topological space obtained by gluing X1X_{1} and X2X_{2} via a homeomorphism Φ:O1O2\Phi\colon O_{1}\to O_{2}.

If for any x1bd(O1)X1x_{1}\in\operatorname{bd}(O_{1})\subset X_{1}, x2bd(O2)X2x_{2}\in\operatorname{bd}(O_{2})\subset X_{2}, where bd\operatorname{bd} means “boundary”, there exist open neighborhoods V1x1V_{1}\ni x_{1} in X1X_{1} and V2x2V_{2}\ni x_{2} in X2X_{2} such that

Φ(V1O1)V2=,\Phi(V_{1}\cap O_{1})\cap V_{2}=\varnothing,

then XX is Hausdorff.∎

Construction 4.8.

Suppose that CC, L1,,LsL_{1},\ldots,L_{s}, W1,,WsW_{1},\ldots,W_{s}, and WW are as in Lemma 4.6.

Put

U0=WW1WsU_{0}=W\sqcup W_{1}\dots\sqcup W_{s}

(disjoint sum), and let π0:U02\pi_{0}\colon U_{0}\to\mathbb{P}^{2} be the natural projection. Define the equivalence relation \sim on U0U_{0} as follows: if x,yU0x,y\in U_{0} and xyx\neq y, then xyx\sim y if and only if π0(x)=π0(y)P1Ps\pi_{0}(x)=\pi_{0}(y)\in P_{1}\cup\dots\cup P_{s}.

Let U1U_{1} be the quotient of U0U_{0} by the equivalence relation \sim, and let π1:U12\pi_{1}\colon U_{1}\to\mathbb{P}^{2} be the natural projection.

Denote the images of CWC\subset W and LjWjL_{j}\subset W_{j} in U1U_{1} by C1C_{1} and LjL_{j}^{\prime}.

Lemma 4.9.

In the above setting, U1U_{1} is a Hausdorff and connected complex surface and π1:U12\pi_{1}\colon U_{1}\to\mathbb{P}^{2} is a local holomorphic isomorphism.

The curves C1C_{1} and LjL_{j}^{\prime} are isomorphic to 1\mathbb{P}^{1}, (LjLj)=(LjLj)=1(L^{\prime}_{j}\cdot L^{\prime}_{j})=(L_{j}\cdot L_{j})=1, (LjC1)=1(L^{\prime}_{j}\cdot C_{1})=1 for each jj, and (C1C1)=(CC)=4(C_{1}\cdot C_{1})=(C\cdot C)=4.

Proof.

If we put, in Lemma 4.7, X1=WX_{1}=W, X2=W1WsX_{2}=W_{1}\sqcup\dots\sqcup W_{s}, O1=P1PsWO_{1}=P_{1}\cup\ldots\cup P_{s}\subset W, O2=P1PsW1WsO_{2}=P_{1}\sqcup\dots\sqcup P_{s}\subset W_{1}\sqcup\dots\sqcup W_{s}, Φ=Id\Phi=\mathrm{Id}, then the hypothesis of this lemma is satisfied if, putting P=P1PsP=P_{1}\cup\dots\cup P_{s} and Q=Q1QsQ=Q_{1}\cup\dots\cup Q_{s}, one has

(6) ((P¯W)P)(P¯(W1Ws)P)=.((\bar{P}\cap W)\setminus P)\cap(\bar{P}\cap(W_{1}\cup\dots\cup W_{s})\setminus P)=\varnothing.

The left-hand side of (6) is equal to

(P¯W(W1Ws))P=(P¯P)(PQ)P¯Q,(\bar{P}\cap W\cap(W_{1}\cup\dots\cup W_{s}))\setminus P=(\bar{P}\setminus P)\cap(P\cup Q)\subset\bar{P}\cap Q,

which is empty by Lemma 4.6(5). Hence, U1U_{1} is Hausdorff.

The rest is obvious. ∎

Construction 4.10.

In the above setting, for each jj, 1js1\leq j\leq s, choose two distinct points uj,vjLju_{j},v_{j}\in L^{\prime}_{j}, different from the intersection point of LjL^{\prime}_{j} and C1C_{1}. Let U2U_{2} be the blowup of the surface U1U_{1} at the points u1,,us,v1,,vsu_{1},\ldots,u_{s},v_{1},\ldots,v_{s}. If σ:U2U1\sigma\colon U_{2}\to U_{1} is the natural morphism, put C2=σ1(C)U2C_{2}=\sigma^{-1}(C^{\prime})\subset U_{2}; it is clear that σ\sigma is an isomorphism on a neighborhood of C2C_{2}; in particular, C2C11C_{2}\cong C_{1}\cong\mathbb{P}^{1} and (C2C2)=4(C_{2}\cdot C_{2})=4.

For each jj, let L~jU2\tilde{L}_{j}\subset U_{2} be the strict transform of LjL^{\prime}_{j} with respect to the blowup σ:U2U1\sigma\colon U_{2}\to U_{1}. By construction, L~j1\tilde{L}_{j}\cong\mathbb{P}^{1}, (L~jL~j)=1(\tilde{L}_{j}\cdot\tilde{L}_{j})=-1 for each jj, and the curves L~j\tilde{L}_{j} are pairwise disjoint. Hence, one can blow down the curves L~1,,L~s\tilde{L}_{1},\ldots,\tilde{L}_{s} to obtain a smooth complex surface U3U_{3} and a curve C3U3C_{3}\subset U_{3}, which is the image of C2C_{2}; one has C31C_{3}\cong\mathbb{P}^{1} and (C3C3)=(C2C2)+s=4+s>4(C_{3}\cdot C_{3})=(C_{2}\cdot C_{2})+s=4+s>4.

Proposition 4.11.

If (C3,U3)(C_{3},U_{3}) is the germ of neighborhood from Construction 4.10, then this germ is not algebraic and tr.deg(U3,C3)=2\operatorname{tr.deg}_{\mathbb{C}}\mathcal{M}(U_{3},C_{3})=2.

Proof.

It follows from the construction that the blowup of the germ of neighborhood of C3C_{3} in U3U_{3} at the ss points to which L~1,,L~s\tilde{L}_{1},\ldots,\tilde{L}_{s} were blown down, is isomorphic to the germ of neighborhood of C2C_{2} in U2U_{2}, which is isomorphic to the of neighborhood of the conic CC in 2\mathbb{P}^{2}. Now Proposition 4.5 implies that the germ (C3,U3)(C_{3},U_{3}) is not algebraic.

Since π1:U12\pi_{1}\colon U_{1}\to\mathbb{P}^{2} is a local isomorphism, the filed of meromorphic functions on 2\mathbb{P}^{2}, which is isomorphic to the field (X,Y)\mathbb{C}(X,Y) of rational functions in two variables, can be embedded in the field of meromorphic functions on U1U_{1}. Since the surface U3U_{3} is obtained from U1U_{1} by a sequence of blowups and blowdowns, the fields of meromorphic functions on U1U_{1} and U3U_{3} are isomorphic. Hence, (X,Y)\mathbb{C}(X,Y) can be embedded in the field of meromorphic functions on U3U_{3}, which can be embedded in (U3,C3)\mathcal{M}(U_{3},C_{3}). Thus, tr.deg(U3,C3)2\operatorname{tr.deg}_{\mathbb{C}}\mathcal{M}(U_{3},C_{3})\geq 2, whence tr.deg(U3,C3)=2\operatorname{tr.deg}_{\mathbb{C}}\mathcal{M}(U_{3},C_{3})=2. This completes the proof. ∎

5. Ramified coverings

In this section we construct another series of examples of non-algebraic neighborhoods (C,U)(C,U), where C1C\cong\mathbb{P}^{1} and tr.deg(U,C)=2\operatorname{tr.deg}_{\mathbb{C}}\mathcal{M}(U,C)=2. In these examples, the self-intersection (CC)(C\cdot C) may be an arbitrary positive integer. We begin with two simple lemmas.

Lemma 5.1.

Suppose that XX is a smooth complex surface, CXC\subset X is a projective curve, C1C\cong\mathbb{P}^{1}, and UXU\subset X is a tubular neighborhood of CC. Then π1(UC)/m\pi_{1}(U\setminus C)\cong\mathbb{Z}/m\mathbb{Z}, where m=(CC)m=(C\cdot C).

Proof.

Immediate from the homotopy exact sequence of the fiber bundle UCCU\setminus C\to C. ∎

Lemma 5.2.

Suppose that f:STf\colon S\to T is a dominant morphism of smooth projective algebraic surfaces and that BTB\subset T is the branch divisor of ff (see the definition in Section 1.1). If TBT\setminus B is simply connected, then degf=1\deg f=1.

Proof.

The critical locus RTR\subset T of ff is of the form BEB\cup E, where BB is the branch divisor and EE is a finite set. The mapping

f|Sf1(R):Sf1(R)TRf|_{S\setminus f^{-1}(R)}\colon S\setminus f^{-1}(R)\to T\setminus R

is a topological covering. Since the subset ETBE\subset T\setminus B is finite and TBT\setminus B is smooth, fundamental groups of TRT\setminus R and TBT\setminus B are isomorphic, so TRT\setminus R is also simply connected, whence the result. ∎

Construction 5.3.

Fix an integer n>0n>0. We are going to construct a certain neighborhood of 1\mathbb{P}^{1} with self-intersection nn.

To that end, suppose that X2n+1X\subset\mathbb{P}^{2n+1} is a non-degenerate surface of degree 2n2n which is not the cone F2n,0F_{2n,0} and, if n=2n=2, not the Veronese surface VV. Let CXC\subset X be a smooth hyperplane section, and let UCU\supset C, UXU\subset X be a tubular neighborhood of CC. By virtue of Lemma 5.1 one has π1(UC)=/2n\pi_{1}(U\setminus C)=\mathbb{Z}/2n\mathbb{Z}. Hence, there exists a two-sheeted ramified covering π:VU\pi\colon V\to U that is ramified along CC with index 22. If C=π1(C)C^{\prime}=\pi^{-1}(C), then C1C^{\prime}\cong\mathbb{P}^{1} and (CC)=n(C^{\prime}\cdot C^{\prime})=n.

Proposition 5.4.

If (C,V)(C^{\prime},V) is the neighborhood from Construction 5.3, then tr.deg(V,C)=2\operatorname{tr.deg}_{\mathbb{C}}\mathcal{M}(V,C^{\prime})=2 and the neighborhood (C,V)(C^{\prime},V) is not algebraic.

Proof.

The neighborhood (C,U)(C,U) is algebraic, so tr.deg(U,C)=2\operatorname{tr.deg}_{\mathbb{C}}\mathcal{M}(U,C)=2, and the morphism π:VU\pi\colon V\to U induces an embedding of (U,C)\mathcal{M}(U,C) in (V,C)\mathcal{M}(V,C^{\prime}), hence tr.deg(V,C)2\operatorname{tr.deg}_{\mathbb{C}}\mathcal{M}(V,C^{\prime})\geq 2, hence this transcendence degree equals 22.

To prove the non-algebraicity of (C,V)(C^{\prime},V), assume the converse. Then, by virtue of Proposition 3.3, the germ of the neighborhood (C,V)(C^{\prime},V) is isomorphic to the germ of neighborhood of a smooth hyperplane section of a non-degenerate surface Xn+1X^{\prime}\subset\mathbb{P}^{n+1}, degX=n\deg X^{\prime}=n; we will identify CC^{\prime} with this hyperplane section and VV with a neighborhood of CC^{\prime} in XX^{\prime}.

Let f0,,f2nf_{0},\ldots,f_{2n} be a basis of the space H0(X,𝒪X(C))H^{0}(X,\mathcal{O}_{X}(C)) (i.e., of the space of meromorphic functions on XX with at worst a simple pole along CC). For each jj, the function (fj|U)π(f_{j}|_{U})\circ\pi is a meromorphic function on VV with at worst a pole of order 22 along CC^{\prime}. Using Lemma 3.6 (in which one puts r=2r=2), one sees that there exist meromorphic functions g0,,g2nH0(X,𝒪X(2C))g_{0},\ldots,g_{2n}\in H^{0}(X^{\prime},\mathcal{O}_{X^{\prime}}(2C^{\prime})) such that, for each jj, the germ of gjg_{j} along CC is the same as that of (fj|U)π(f_{j}|_{U})\circ\pi.

Choose a basis h0,,hNh_{0},\ldots,h_{N} is of H0(X,𝒪X(2C))H^{0}(X^{\prime},\mathcal{O}_{X^{\prime}}(2C^{\prime})), and let X1NX_{1}\subset\mathbb{P}^{N} be the image of XX^{\prime} under the embedding

x(h0(x)::hN(x))x\mapsto(h_{0}(x):\dots:h_{N}(x))

(this is the embedding defined by the complete linear system |2C|=|𝒪X(2)||2C|=|\mathcal{O}_{X^{\prime}}(2)|). If gj=ajkhkg_{j}=\sum a_{jk}h_{k} for each jj, 0j2n0\leq j\leq 2n, then the matrix ajk\|a_{jk}\| defines a rational mapping p:X1Xp\colon X_{1}\dasharrow X, induced by a linear projection p¯:N2n\bar{p}\colon\mathbb{P}^{N}\dasharrow\mathbb{P}^{2n}. Hence,

(7) degXdegX1degp,\deg X\leq\frac{\deg X_{1}}{\deg p},

and the equality is attained if and only if the rational mapping pp is regular.

On the other hand, degX1=4degX=4n\deg X_{1}=4\deg X^{\prime}=4n, degX=2n\deg X=2n, and degp2\deg p\geq 2 since the restriction of pp to VXX1V\subset X^{\prime}\cong X_{1} coincides with our ramified covering π:VU\pi\colon V\to U. Now it follows from (7) that the projection pp is regular (i.e., its center does not intersect X1X_{1}) and degp=2\deg p=2.

If Xn+1X^{\prime}\subset\mathbb{P}^{n+1} is not a cone (i.e., if XFn,0X^{\prime}\neq F_{n,0}), put X2=X1X_{2}=X_{1} and q=pq=p, and if XX^{\prime} is the cone Fn,0F_{n,0}, put X2=𝔽nX_{2}=\mathbb{F}_{n} and q=pσq=p\circ\sigma, where σ:𝔽nFn,0=X\sigma\colon\mathbb{F}_{n}\to F_{n,0}=X^{\prime} is the standard resolution. So, we have a holomorphic mapping q:X2Xq\colon X_{2}\to X, degq=2\deg q=2. One has either X2𝔽nX_{2}\cong\mathbb{F}_{n}, or X22X_{2}\cong\mathbb{P}^{2} (the latter case is possible only if n=4n=4 and X5X^{\prime}\subset\mathbb{P}^{5} is the Veronese surface).

Since qq agrees with π\pi on a neighborhood of the curve CX2C^{\prime}\subset X_{2}, the curve CXC\subset X is contained in the branch divisor of qq. Let us show that the branch divisor of q:X2Xq\colon X_{2}\to X coincides with CC.

To that end, denote this branch divisor by BXB\subset X. Let DXD\subset X be a general hyperplane section, and put D2=q1(D)X2D_{2}=q^{-1}(D)\subset X_{2}. For a general DD, one has D1D\cong\mathbb{P}^{1}, D2D_{2} is a smooth and connected projective curve, and the morphism q|D2:D21Dq|_{D_{2}}\colon D_{2}\to\mathbb{P}^{1}\cong D is ramified over degB\deg B points.

If n=4n=4 and XX^{\prime} is the Veronese surface V5V\subset\mathbb{P}^{5}, then (X2,𝒪X2(D))(2,𝒪2(4))(X_{2},\mathcal{O}_{X_{2}}(D))\cong(\mathbb{P}^{2},\mathcal{O}_{\mathbb{P}^{2}}(4)), so the curve D2D_{2} is isomorphic to a smooth plane quartic. Such a curve does not admit a mapping to 1\mathbb{P}^{1} of degree 22, so this case is impossible.

Thus, X=Fk,lX^{\prime}=F_{k,l}, where k+2l=nk+2l=n. In the notation of Section 2, the divisor D2X2D_{2}\subset X_{2} is equivalent to 2(e+(k+l)f)2(e+(k+l)f); since the canonical class KX2K_{X_{2}} of the surface X2𝔽kX_{2}\cong\mathbb{F}_{k} is equivalent to 2e(k+2)f-2e-(k+2)f, one has, denoting the genus of D2D_{2} by gg,

2g2=(D2D2+KX2)=(2e+2(k+l)f(k+2l2)f)=2(n2),2g-2=(D_{2}\cdot D_{2}+K_{X_{2}})=(2e+2(k+l)f\cdot(k+2l-2)f)=2(n-2),

whence g=n1g=n-1. Applying Riemann–Hurwitz formula to the degree 22 morphism q|D2:D2Dq|_{D_{2}}\colon D_{2}\to D, one sees that the number of its branch points equals 2n2n. So, degB=2n\deg B=2n. Since BCB\supset C and degC=2n\deg C=2n, one has B=CB=C.

Observe now that XCX\setminus C is simply connected since XCX\setminus C, as a topological space, is a fiber bundle over 1\mathbb{P}^{1} with the fiber \mathbb{C}, and the complement X2q1(C)X_{2}\setminus q^{-1}(C) is connected. Applying Lemma 5.2 to the mapping q:X2Xq\colon X_{2}\to X, one sees that degq=1\deg q=1. We arrived at a contradiction. ∎

References

  • [1] Aldo Andreotti, Théorèmes de dépendance algébrique sur les espaces complexes pseudo-concaves, Bull. Soc. Math. France 91 (1963), 1–38. MR 152674
  • [2] Wolf P. Barth, Klaus Hulek, Chris A. M. Peters, and Antonius Van de Ven, Compact complex surfaces, second ed., Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge. A Series of Modern Surveys in Mathematics, vol. 4, Springer-Verlag, Berlin, 2004. MR 2030225
  • [3] David Eisenbud and Joe Harris, On varieties of minimal degree (a centennial account), Algebraic geometry, Bowdoin, 1985 (Brunswick, Maine, 1985), Proc. Sympos. Pure Math., vol. 46, Amer. Math. Soc., Providence, RI, 1987, pp. 3–13. MR 927946
  • [4] Maycol Falla Luza and Frank Loray, Neighborhoods of rational curves without functions, Math. Ann. 382 (2022), no. 3-4, 1047–1058. MR 4403217
  • [5] by same author, Projective structures, neighborhoods of rational curves and Painlevé equations, Mosc. Math. J. 23 (2023), no. 1, 59–95.
  • [6] Hans Grauert, Über Modifikationen und exzeptionelle analytische Mengen, Math. Ann. 146 (1962), 331–368. MR 137127
  • [7] Robin Hartshorne, Ample subvarieties of algebraic varieties, Lecture Notes in Mathematics, Vol. 156, Springer-Verlag, Berlin-New York, 1970, Notes written in collaboration with C. Musili. MR 0282977
  • [8] J. C. Hurtubise and N. Kamran, Projective connections, double fibrations, and formal neighbourhoods of lines, Math. Ann. 292 (1992), no. 3, 383–409. MR 1152943
  • [9] M. B. Mishustin, Neighborhoods of the Riemann sphere in complex surfaces, Funktsional. Anal. i Prilozhen. 27 (1993), no. 3, 29–41, 95 (Russian), English translation: Funct. Anal. Appl., 27 (1993), No 3, 176–185. MR 1250979
  • [10] S. Yu. Orevkov, An example in connection with the Jacobian conjecture, Mat. Zametki 47 (1990), no. 1, 127–136, 173 (Russian), English translation: Mathematical Notes, 1990, Volume 47, Issue 1, Pages 82–88; erratum: Mathematical Notes, 1991, Volume 49, Issue 6, Page 659. MR 1048269
  • [11] V. I. Savelyev, Zero-type imbedding of a sphere into complex surfaces, Vestnik Moskov. Univ. Ser. I Mat. Mekh. (1982), no. 4, 28–32, 85 (Russian), English translation: Mosc. Univ.Math. Bull. 37, 34–39 (1982). MR 671883