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

Multiplicative bounds for measures of irrationality on complete intersections

Nathan Chen
The author’s research was partially supported by an NSF postdoctoral fellowship, DMS-2103099.

The purpose of this paper is to show that measures of irrationality on very general codimension two complete intersections and very general complete intersection surfaces are multiplicative in the degrees of the defining equations.

In recent years, there has been growing interest in studying measures of irrationality for projective varieties. As a higher dimensional generalization of gonality, these birational invariants quantify in various ways how far a given variety XX is from being rational. We will focus primarily on two of these measures, the degree of irrationality and the covering gonality, which are defined as follows:

irr(X)=min{δ>0| degree δ rational covering XdimX};\operatorname{irr}(X)=\min\left\{\delta>0\ \middle|\ \exists\text{ degree $\delta$ rational covering }X\dashrightarrow\mathbb{P}^{\dim X}\right\};
cov.gon(X)=min{c>0|Given a general point pX, irreduciblecurve CX through p with gonality c}.\operatorname{cov.gon}(X)=\min\left\{c>0\ \middle|\ \begin{subarray}{c}\text{Given a general point }p\in X,\ \exists\text{ irreducible}\\[2.84544pt] \text{curve }C\subseteq X\text{ through $p$ with gonality }c\end{subarray}\right\}.

In the case of hypersurfaces of large degree, the situation is now fairly well understood. Specifically, let Xn+1X\subset\mathbb{P}^{n+1}_{\mathbb{C}} be a smooth hypersurface of degree dd and dimension n2n\geq 2. If XX is very general of degree d2n+1d\geq 2n+1, then Bastianelli, De Poi, Ein, Lazarsfeld, and Ullery [2] have shown that irr(X)=d1\operatorname{irr}(X)=d-1. In the same paper, the authors proved that if XX is arbitrary and dn+2d\geq n+2, then cov.gon(X)dn\operatorname{cov.gon}(X)\geq d-n. The central theme of [2] is that the positivity properties of the canonical bundle yield lower bounds for measures of irrationality. Using different methods, Bastianelli, Ciliberto, Flamini, and Supino [1] later computed cov.gon(X)d2n\operatorname{cov.gon}(X)\approx d-2\sqrt{n} for very general hypersurfaces of degree d0d\gg 0.

A logical next step is to investigate the behavior of these invariants for complete intersection varieties in projective space. For complete intersections over \mathbb{C}, the same techniques in [2] yield lower bounds for the covering gonality which are additive in the degrees of the defining equations. Recently, Smith [16] has extended these results about the covering gonality of complete intersections to positive characteristic.

However, it has been conjectured [2, Problem 4.1] that measures of irrationality on complete intersections should be multiplicative in the degrees. As evidence, Lazarsfeld [13, Exercise 4.12] had established that the gonality of a smooth complete intersection curve Ce+1C\subset\mathbb{P}_{\mathbb{C}}^{e+1} of type (a1,a2,,ae)(a_{1},a_{2},\ldots,a_{e}) with 2a1ae2\leq a_{1}\leq\cdots\leq a_{e} is bounded from below by gon(C)(a11)a2ae\operatorname{gon}(C)\geq(a_{1}-1)a_{2}\cdots a_{e}. Further refinements due to Hotchkiss, Lau, and Ullery [10] show that when 4a1<a2ae4\leq a_{1}<a_{2}\leq\cdots\leq a_{e} holds, the gonality of the curve CC is realized by projection from a suitable linear subspace. In higher dimensions, Stapleton [18] used results about Seshadri constants on hypersurfaces which were due to Ito [11] to give bounds for the covering gonality of codimension two complete intersections that were stronger than additive. Later, Stapleton and Ullery [19] computed the degree of irrationality for codimension two complete intersections of type (2,d)(2,d) and (3,d)(3,d).

Our first result shows that the covering gonality of very general codimension two complete intersections is multiplicative:

Theorem A.

Let Xn+2X\subset\mathbb{P}^{n+2}_{\mathbb{C}} be a very general smooth complete intersection of type (a,b)(a,b) and dimension n2n\geq 2. If a,b18n/7a,b\geq 18n/7, then

cov.gon(X)23(n+1)2ab.\operatorname{cov.gon}(X)\geq\frac{2}{3(n+1)^{2}}\cdot ab.

Since in general irr(X)cov.gon(X)\operatorname{irr}(X)\geq\operatorname{cov.gon}(X), we obtain the same inequality for the degree of irrationality.

Our second theorem gives multiplicative bounds for complete intersection surfaces:

Theorem B.

Let Xe+2X\subset\mathbb{P}_{\mathbb{C}}^{e+2} be a very general smooth complete intersection surface of type (d1,,de)(d_{1},\ldots,d_{e}). There exist positive constants A=A(e)A=A(e) and B=B(e)B=B(e) such that if diAd_{i}\geq A for all 1ie1\leq i\leq e, then

cov.gon(X)Bd1de.\operatorname{cov.gon}(X)\geq B\cdot d_{1}\cdots d_{e}.

See the end of §1, Proposition 2.8, and Remark 4.4 for explicit constants.

In §1, we will present a reduction step which first appeared in the work of Stapleton [18, §5.2]. The theorems will reduce to showing that certain families of line bundles on the blow-ups of complete intersection varieties are big and nef, which can be thought of as multi-point Seshadri constants. See Theorems 1.4 and 1.6 for the explicit statements and details. In §2, we will collect several tools that will be used to control the numerical invariants of curves on complete intersections. In §3, we will prove Theorem 1.4. The proof of Theorem 1.6 will take up most of §4. Throughout the paper, we work over \mathbb{C}.

Acknowledgements.

I am grateful to Robert Lazarsfeld for many valuable discussions and for suggesting the approach to Proposition 2.4. I would also like to thank Olivier Martin, Mihnea Popa, and David Stapleton for giving valuable feedback about an earlier draft of the paper and thank Aaron Landesman for helpful discussions. A version of Theorem A was part of the author’s Ph.D. thesis at Stony Brook University.

1. Reduction step

In this section, we will show how Theorems A and B follow from the nefness of certain families of line bundles. Our starting point is the following result [2, Theorem 1.10], which says that positivity properties of the canonical bundle lead to lower bounds on the covering gonality (see also [18, Remark 5.14]):

Proposition 1.1.

Let XX be a smooth projective variety and suppose that there exists an integer rr such that the canonical bundle KXK_{X} separates rr points on an open set. Then

cov.gon(X)r+1.\operatorname{cov.gon}(X)\geq r+1.

We will apply this proposition as follows. Consider the inclusions XYn+eX\subset Y\subset\mathbb{P}^{n+e}, where

  • \diamond

    YY is a complete intersection of dimension n+1n+1 and type (a1,,ae1)(a_{1},\ldots,a_{e-1}), and

  • \diamond

    X|𝒪Y(ae)|X\in\absolutevalue{\mathcal{O}_{Y}(a_{e})} is a complete intersection of dimension nn and type (a1,,ae)(a_{1},\ldots,a_{e}).

We would like to show that KXK_{X} separates rr points on an open set, for suitable rr. By passing to a complete intersection YY of larger dimension, we can take advantage of adjunction:

Proposition 1.2.

Suppose that there exists a positive integer rr such that on the blow-up μ:Y~Y\mu\colon\widetilde{Y}\rightarrow Y along any distinct points p1,,prXp_{1},\ldots,p_{r}\in X with exceptional divisors E1,,ErE_{1},\ldots,E_{r}, the line bundle

L\colonequalsμ𝒪Y(X)i=1r(n+1)EiL\colonequals\mu^{\ast}\mathcal{O}_{Y}(X)-\sum_{i=1}^{r}(n+1)E_{i}

is nef and big. Then cov.gon(X)r+1\operatorname{cov.gon}(X)\geq r+1.

Proof of Proposition 1.2.

By Kawamata-Viehweg vanishing for big and nef line bundles,

H1(Y,(KY+𝒪Y(X)){p1,,pr})\displaystyle H^{1}\left(Y,(K_{Y}+\mathcal{O}_{Y}(X))\otimes\mathcal{I}_{\{p_{1},\ldots,p_{r}\}}\right) =H1(Y~,μ(KY+𝒪Y(X))i=1rEi)\displaystyle=H^{1}\left(\widetilde{Y},\mu^{\ast}(K_{Y}+\mathcal{O}_{Y}(X))-\sum_{i=1}^{r}E_{i}\right)
=H1(Y~,KY~+L)=0.\displaystyle=H^{1}(\widetilde{Y},K_{\widetilde{Y}}+L)=0.

Here we use the fact that KY~μKY+nEK_{\widetilde{Y}}\cong\mu^{\ast}K_{Y}+nE. The vanishing above gives a surjection

H0(Y,KY+𝒪Y(X))H0(Y,(KY+𝒪Y(X))𝒪{p1,,pr}).H^{0}(Y,K_{Y}+\mathcal{O}_{Y}(X))\twoheadrightarrow H^{0}\left(Y,(K_{Y}+\mathcal{O}_{Y}(X))\otimes\mathcal{O}_{\{p_{1},\ldots,p_{r}\}}\right).

In other words, sections of the adjoint bundle KY+𝒪Y(X)K_{Y}+\mathcal{O}_{Y}(X) separate any finite set of rr distinct points in XX. By the adjunction formula, KX(KY+𝒪Y(X))|XK_{X}\cong(K_{Y}+\mathcal{O}_{Y}(X))\big{|}_{X} and hence sections of KXK_{X} separate any finite set of rr distinct points in XX. Proposition 1.1 implies that cov.gon(X)r+1\operatorname{cov.gon}(X)\geq r+1. ∎

In practice, once we prove nefness of LL, it will follow numerically that LL is big [14, Theorem 2.2.16]. By Proposition 1.2, it suffices to prove nefness of families of line bundles in two different settings:

Set-up 1.3 (Codimension two complete intersections).

Let n2n\geq 2 be arbitrary and set e=2e=2. Then XYn+2X\subset Y\subset\mathbb{P}^{n+2}, where Y|𝒪n+2(a)|Y\in\absolutevalue{\mathcal{O}_{\mathbb{P}^{n+2}}(a)} is a very general hypersurface and X|𝒪Y(b)|X\in\absolutevalue{\mathcal{O}_{Y}(b)} is a very general complete intersection of dimension nn such that ba18n/7b\geq a\geq 18n/7. With this, we will show:

Theorem 1.4.

Consider Set-up 1.3 and fix an integer

r23(n+1)2ab.r\leq\frac{2}{3(n+1)^{2}}ab.

For any set of distinct points p1,,prYp_{1},\ldots,p_{r}\in Y, if we let μ:Y~Y\mu\colon\widetilde{Y}\rightarrow Y denote the blow-up of YY at these points with exceptional divisor EiE_{i} over pip_{i} and set H=μ𝒪Y(1)H=\mu^{\ast}\mathcal{O}_{Y}(1), then the following divisor on Y~\widetilde{Y} is nef:

bHi=1r(n+1)Ei.bH-\sum_{i=1}^{r}(n+1)E_{i}.

Granting Theorem 1.4 for now, we will first show how it implies Theorem A.

Proof of Theorem A.

In the setting of Set-up 1.3, fix n2n\geq 2, choose b,a18n/7b,a\geq 18n/7, and set r=23(n+1)2abr=\left\lfloor\frac{2}{3(n+1)^{2}}\cdot ab\right\rfloor. Theorem 1.4 shows that for any tuple of rr distinct points p1,,prXp_{1},\ldots,p_{r}\in X, the divisor

L\colonequalsbHi=1r(n+1)EiL\colonequals bH-\sum_{i=1}^{r}(n+1)E_{i}

on the blow-up μ:Y~Y\mu\colon\widetilde{Y}\rightarrow Y is nef. It is straightforward to check that (Ln+1)>0(L^{n+1})>0 on Y~\widetilde{Y}:

(bHi=1r(n+1)Ei)n+1=abn+1r(n+1)n+1abn+123(n+1)n1ab>0\left(bH-\sum_{i=1}^{r}(n+1)E_{i}\right)^{n+1}=ab^{n+1}-r\cdot(n+1)^{n+1}\geq ab^{n+1}-\frac{2}{3}(n+1)^{n-1}ab>0

holds as long as ban+1b\geq a\geq n+1, so LL is also big. By Proposition 1.2,

cov.gon(X)r+123(n+1)2ab.\operatorname{cov.gon}(X)\geq r+1\geq\frac{2}{3(n+1)^{2}}\cdot ab.\qed
Set-up 1.5 (Complete intersection surfaces).

Set n=2n=2 and let e2e\geq 2 be arbitrary. We will first consider a very general smooth complete intersection surface

XYe+2of type(a1,,ae),X\subset Y\subset\mathbb{P}^{e+2}\qquad\text{of type}\qquad(a_{1},\ldots,a_{e}),

where the aia_{i} are integers of a special form. More precisely, assume that 3ea1ae3e\leq a_{1}\leq\cdots\leq a_{e} and

ai\colonequals(e+1)!qifor i=1,,e1,a_{i}\colonequals(e+1)!\cdot q_{i}\qquad\text{for }i=1,\ldots,e-1,

where the qiq_{i} are positive integers which are pairwise coprime. Let YY be the smooth complete intersection threefold of type (a1,,ae1)(a_{1},\ldots,a_{e-1}) containing XX. We will prove:

Theorem 1.6.

Consider Set-up 1.5 and fix a positive integer

r234(3e+2)((e+1)!)ea1ae.r\leq\frac{2}{3^{4}(3e+2)((e+1)!)^{e}}\cdot a_{1}\cdots a_{e}.

For any collection p1,,prYp_{1},\ldots,p_{r}\in Y of rr points, if we write Y~\colonequalsBl{p1,,pr}Y𝜇Y\widetilde{Y}\colonequals\operatorname{Bl}_{\{p_{1},\ldots,p_{r}\}}Y\xrightarrow{\mu}Y for the blow-up with exceptional divisor EiE_{i} over pip_{i}, then the divisor

μ𝒪Y(ae)i=1r3Eiis nef on Y~.\mu^{\ast}\mathcal{O}_{Y}(a_{e})-\sum_{i=1}^{r}3E_{i}\qquad\text{is nef on }\widetilde{Y}.

Granting Theorem 1.6 for now, we will use it to prove Theorem B.

Proof of Theorem B.

In the setting of Set-up 1.5, fix e3e\geq 3, choose d(e+1)!4(e1)log4(e1)d\geq(e+1)!\cdot 4(e-1)\log 4(e-1), and set

r=234(3e+2)((e+1)!)ea1ae.r=\left\lfloor\frac{2}{3^{4}(3e+2)((e+1)!)^{e}}\cdot a_{1}\cdots a_{e}\right\rfloor.

Theorem 1.6 shows that for any tuple of rr distinct points p1,,prXp_{1},\ldots,p_{r}\in X, the divisor

L\colonequalsaeHi=1r(n+1)EiL\colonequals a_{e}H-\sum_{i=1}^{r}(n+1)E_{i}

on the blow-up μ:Y~Y\mu\colon\widetilde{Y}\rightarrow Y is nef. It is straightforward to check that (L3)>0(L^{3})>0 on Y~\widetilde{Y}, so LL is also big. By Proposition 1.2,

cov.gon(X)r+1234(3e+2)((e+1)!)ea1ae.\operatorname{cov.gon}(X)\geq r+1\geq\frac{2}{3^{4}(3e+2)((e+1)!)^{e}}\cdot a_{1}\cdots a_{e}.

So far, this only gives lower bounds on the covering gonality of complete intersection surfaces of special degrees. Now consider a very general complete intersection surface Xe+2X^{\prime}\subset\mathbb{P}^{e+2} of type (d1,,de)(d_{1},\ldots,d_{e}) where d1ded_{1}\leq\cdots\leq d_{e}. If d1d_{1} is sufficiently large, then we may choose a1,,aea_{1},\ldots,a_{e} such that

(a1,,ae1,ae)\colonequals((e+1)!q1,,(e+1)!qe1,de),(a_{1},\ldots,a_{e-1},a_{e})\colonequals((e+1)!q_{1},\ldots,(e+1)!q_{e-1},d_{e}),

the inegers qiq_{i} are distinct primes, and di/2<aidid_{i}/2<a_{i}\leq d_{i} for 1ie11\leq i\leq e-1.111In Remark 4.4, we will explain how to choose these aia_{i}. Next, degenerate XX^{\prime} to a union of varieties, with one component consisting of a very general complete intersection Xe+2X\subset\mathbb{P}^{e+2} of type (a1,,ae)(a_{1},\ldots,a_{e}). The family can be chosen so that the total space is irreducible. In order to prove that

cov.gon(X)cov.gon(X),\operatorname{cov.gon}(X^{\prime})\geq\operatorname{cov.gon}(X),

we need a strengthened version of [8, Proposition 2.2] involving families where the central fiber is possibly reducible:

Proposition 1.7.

Let f:𝒳Tf\colon\mathcal{X}\rightarrow T be a flat family of irreducible projective varieties over an irreducible one-dimensional base. Assume that the total space 𝒳\mathcal{X} is irreducible and suppose that for all 0tT0\not=t\in T, the fiber 𝒳t\mathcal{X}_{t} has covering gonality d\leq d. Then every component 𝒳𝒳0red\mathcal{X}^{\prime}\subset\mathcal{X}_{0}^{\text{red}} of the reduced special fiber has covering gonality d\leq d.

Proof of Proposition 1.7.

Adaptating the proof of [8, Proposition 2.2], the key point is that irreducibility of the total space of the family means that if the base of the covering family coming from the compactified Kontsevich moduli space of stable maps is irreducible (this may be assumed), then the covering family automatically covers every component of the central fiber. The rest of argument follows through. ∎

By the proposition above, we have

cov.gon(X)cov.gon(X)234(3e+2)((e+1)!)ed12de12de,\operatorname{cov.gon}(X^{\prime})\geq\operatorname{cov.gon}(X)\geq\frac{2}{3^{4}(3e+2)((e+1)!)^{e}}\cdot\frac{d_{1}}{2}\cdots\frac{d_{e-1}}{2}\cdot d_{e},

which simplifies to give the desired bound with a constant of

B(e)=234(3e+2)((e+1)!)e2e1B(e)=\frac{2}{3^{4}(3e+2)((e+1)!)^{e}2^{e-1}}

This completes the proof of Theorem B. ∎

Remark 1.8.

Note that Theorems 1.4 and 1.6 are false without the very general assumption. For instance, one may take a complete intersection YY which contains a line and then choose points which all lie on the line.

We will now give a conceptual outline of the proofs of Theorems 1.4 and 1.6. Proceeding by contradiction, the failure of the line bundle LL on Y~\widetilde{Y} to be nef means that there exists a curve C~Y~\widetilde{C}\subset\widetilde{Y} which intersects negatively against LL. By projecting to the complete intersection YY, this roughly says that the image curve C\colonequalsμ(C~)C\colonequals\mu(\widetilde{C}) passes through the points pip_{i} with large multiplicities. We then relate this to the geometry of curves on very general complete intersections to reach a contradiction. For Theorem 1.4, we will need lower bounds on the geometric genus of CC, which follow from work of Ein [6] and Voisin [20]. The argument for Theorem 1.6 will require (i) a more precise estimate of the arithmetic genus of CC and its relationship to the multiplicities mim_{i}, and (ii) lower bounds on the degree of CC. The first ingredient (i) will incorporate some ideas originally due to Castelnuovo [4] about estimating the genus of a space curve, while (ii) will involve certain degeneration arguments of Kollár [12].

2. Numerical invariants of curves on complete intersections

In this section, we will collect some results about the geometry of curves in complete intersections, which will be used in the proofs of Theorems 1.4 and 1.6. We begin by giving lower bounds for the geometric genus of curves on generic complete intersections, which arise from calculations of Ein [6] and Voisin [20] (for comparison, see [2, proof of Proposition 3.8]):

Proposition 2.1.

Let Xn+eX\subset\mathbb{P}^{n+e} be a very general complete intersection of dimension n2n\geq 2 and type (d1,,de)(d_{1},\ldots,d_{e}). For any integral curve CXC\subset X, we have

pg(C)1+12(i=1edi2ne)degn+e(C).p_{g}(C)\geq 1+\frac{1}{2}(\sum_{i=1}^{e}d_{i}-2n-e)\cdot\deg_{\mathbb{P}^{n+e}}(C).
Proof.

Consider the spaces Vdi\colonequalsH0(n,𝒪n+k(di))V^{d_{i}}\colonequals H^{0}(\mathbb{P}^{n},\mathcal{O}_{\mathbb{P}^{n+k}}(d_{i})) for di2d_{i}\geq 2 and let V=iVdiV=\prod_{i}V^{d_{i}}. Consider the universal complete intersection 𝒳V×n+k\mathcal{X}\subseteq V\times\mathbb{P}^{n+k} of type (d1,,de)(d_{1},\ldots,d_{e}) with the two projections pr1:𝒳Vpr_{1}\colon\mathcal{X}\longrightarrow V and pr2:𝒳n+epr_{2}\colon\mathcal{X}\longrightarrow\mathbb{P}^{n+e}. Let v=dimVv=\dim V and suppose that a very general complete intersection of type (d1,,de)(d_{1},\ldots,d_{e}) in n+e\mathbb{P}^{n+e} contains an irreducible curve of geometric genus gg. By standard arguments, there is a diagram

𝒞{\mathcal{C}}𝒳{\mathcal{X}}T{T}V{V}π\scriptstyle{\pi}f\scriptstyle{f}pr1\scriptstyle{pr_{1}}ρ\scriptstyle{\rho}

where π:𝒞T\pi\colon\mathcal{C}\rightarrow T is a family of curves of geometric genus gg whose general member 𝒞t=π1(t)\mathcal{C}_{t}=\pi^{-1}(t) is smooth, ρ\rho is étale, and ft:𝒞tXρ(t)f_{t}\colon\mathcal{C}_{t}\rightarrow X_{\rho(t)} is birational onto its image. In this setting, Ein and Voisin show that if tTt\in T is a general point, then

Ω𝒞v+1((pr2f)𝒪n+e(2(n+e)i=1edie))|𝒞t\Omega_{\mathcal{C}}^{v+1}\otimes\left((pr_{2}\circ f)^{\ast}\mathcal{O}_{\mathbb{P}^{n+e}}(2(n+e)-\sum_{i=1}^{e}d_{i}-e)\middle)\right|_{\mathcal{C}_{t}}

is generically generated by its global sections. This implies that the canonical bundle of the general curve 𝒞t\mathcal{C}_{t} is of the form

K𝒞t(i=1edi2ne)H𝒞t+(Effective),K_{\mathcal{C}_{t}}\cong(\sum_{i=1}^{e}d_{i}-2n-e)H_{\mathcal{C}_{t}}+(\text{Effective}),

where H𝒞tH_{\mathcal{C}_{t}} is the pull-back of the hyperplane bundle from n+e\mathbb{P}^{n+e}. Comparing degrees on both sides, we arrive at the desired result. ∎

We will also need the following result:

Lemma 2.2.

Let CNC\subset\mathbb{P}^{N} (for N3N\geq 3) be a reduced and irreducible curve of degree kk with a finite collection of points pip_{i} (i=1,,i=1,\ldots,\ell) which have multiplicity mim_{i}. After a generic projection of φ:CC2\varphi\colon C\rightarrow C^{\prime}\subset\mathbb{P}^{2}, the multiplicities of the image points φ(pi)\varphi(p_{i}) in CC^{\prime} remain the same. This leads to the estimate:

i=1mi(mi1)2pa(C)pg(C)=(k1)(k2)2pg(C).\sum_{i=1}^{\ell}\frac{m_{i}(m_{i}-1)}{2}\leq p_{a}(C^{\prime})-p_{g}(C^{\prime})=\frac{(k-1)(k-2)}{2}-p_{g}(C).
Remark 2.3.

Given a smooth variety XX and a curve CXC\subset X, the multiplicity of CC at a point pp is equal to the intersection of the strict transform C~\widetilde{C} against the exceptional divisor EpE_{p} of the blow-up μ:X~X\mu\colon\widetilde{X}\rightarrow X at pp (see [7, pg. 79]).

In order to prove Theorem 1.6, we will need a finer analysis of the contribution of the multiplicity of a singular point to the arithmetic genus of a curve. This is captured in:

Proposition 2.4.

Let VV be a smooth variety of dimension nn. Let CVC\subset V be an irreducible and reduced curve with a singular point pp of multiplicity m\colonequalsmultpCm\colonequals\operatorname{mult}_{p}C. Then the discrepancy between the arithmetic genus and the geometric genus of CC is bounded from below by

pa(C)pg(C)(n1)nn1nmnn1nm.p_{a}(C)-p_{g}(C)\geq\frac{(n-1)^{\frac{n}{n-1}}}{n}\cdot m^{\frac{n}{n-1}}-nm.
Proof.

Consider the blow-up V~\colonequalsBlpV𝜋V\widetilde{V}\colonequals\operatorname{Bl}_{p}V\xrightarrow{\pi}V, with exceptional divisor EE. Let C~V~\widetilde{C}\subset\widetilde{V} be the strict transform of CC. We may pushforward the ideal sheaf sequence for C~V~\widetilde{C}\subset\widetilde{V} along π\pi to obtain

0C𝒪Vπ𝒪C~τ\colonequalsR1πC~0.0\rightarrow\mathcal{I}_{C}\rightarrow\mathcal{O}_{V}\rightarrow\pi_{\ast}\mathcal{O}_{\widetilde{C}}\rightarrow\tau\colonequals R^{1}\pi_{\ast}\mathcal{I}_{\widetilde{C}}\rightarrow 0.

To compute the length of the sheaf τ\tau, we will use the theorem on formal functions. Write \mathcal{E}_{\ell} for the \ell-th infinitesimal neighborhood of EE. Then

(R1μC~)=limH1(C~𝒪),\left(R^{1}\mu_{\ast}\mathcal{I}_{\widetilde{C}}\right)^{\wedge}=\varprojlim_{\ell}H^{1}\left(\mathcal{I}_{\widetilde{C}}\otimes\mathcal{O}_{\mathcal{E}_{\ell}}\right),

so

length(τ)=limh1(C~𝒪).\text{length}(\tau)=\lim_{\ell\rightarrow\infty}h^{1}\left(\mathcal{I}_{\widetilde{C}}\otimes\mathcal{O}_{\mathcal{E}_{\ell}}\right).

The spaces on the right can be studied using the sequences:

0C~|E(1)C~𝒪C~𝒪10.0\rightarrow\mathcal{I}_{\widetilde{C}}\big{|}_{E}(\ell-1)\rightarrow\mathcal{I}_{\widetilde{C}}\otimes\mathcal{O}_{\mathcal{E}_{\ell}}\rightarrow\mathcal{I}_{\widetilde{C}}\otimes\mathcal{O}_{\mathcal{E}_{\ell-1}}\rightarrow 0.

Define

k\displaystyle k_{\ell} =dimker(H0(C~𝒪1)H1(C~|E(1)))\displaystyle=\dim\ker\left(H^{0}\left(\mathcal{I}_{\widetilde{C}}\otimes\mathcal{O}_{\mathcal{E}_{\ell-1}}\right)\rightarrow H^{1}\left(\mathcal{I}_{\widetilde{C}}\big{|}_{E}(\ell-1)\right)\right)
c\displaystyle c_{\ell} =dimcoker(H0(C~𝒪1)H1(C~|E(1))).\displaystyle=\dim\operatorname{coker}\left(H^{0}(\mathcal{I}_{\widetilde{C}}\otimes\mathcal{O}_{\mathcal{E}_{\ell-1}})\rightarrow H^{1}\left(\mathcal{I}_{\widetilde{C}}\big{|}_{E}(\ell-1)\right)\right).

Since all higher cohomology HiH^{i} terms are zero for i2i\geq 2, it follows that

(1) h0(C~𝒪)=h0(C~|E(1))+k,h1(C~𝒪)=c+h1(C~𝒪1).\displaystyle\begin{split}h^{0}\left(\mathcal{I}_{\widetilde{C}}\otimes\mathcal{O}_{\mathcal{E}_{\ell}}\right)&=h^{0}\left(\mathcal{I}_{\widetilde{C}}\big{|}_{E}(\ell-1)\right)+k_{\ell},\\ h^{1}\left(\mathcal{I}_{\widetilde{C}}\otimes\mathcal{O}_{\mathcal{E}_{\ell}}\right)&=c_{\ell}+h^{1}\left(\mathcal{I}_{\widetilde{C}}\otimes\mathcal{O}_{\mathcal{E}_{\ell-1}}\right).\end{split}

In addition,

(2) ck=h1(C~|E(1))h0(C~𝒪1).\displaystyle c_{\ell}-k_{\ell}=h^{1}\left(\mathcal{I}_{\widetilde{C}}\big{|}_{E}(\ell-1)\right)-h^{0}\left(\mathcal{I}_{\widetilde{C}}\otimes\mathcal{O}_{\mathcal{E}_{\ell-1}}\right).

Now we will need the following:

Claim.

With the set-up above:

  1. (i)

    The dimensions h0(C~𝒪1)h^{0}\left(\mathcal{I}_{\widetilde{C}}\otimes\mathcal{O}_{\ell-1}\right) are non-decreasing in \ell and stabilize for 0\ell\gg 0.

  2. (ii)

    For any >0\ell>0,

    h1(C~𝒪)=i=02χ(C~|E(i))+h1(C~|E(1))+k.h^{1}\left(\mathcal{I}_{\widetilde{C}}\otimes\mathcal{O}_{\mathcal{E}_{\ell}}\right)=-\sum_{i=0}^{\ell-2}\chi(\mathcal{I}_{\widetilde{C}}\big{|}_{E}(i))+h^{1}(\mathcal{I}_{\widetilde{C}}\big{|}_{E}(\ell-1))+k_{\ell}.

Granting the Claim for now, we will use it to complete the proof of Proposition 2.4. Observe that

χ(C~|E(i))=χ(𝒪E(i))χ(𝒪C~E(i))=h0(𝒪E(i))multpC=(i+n1n1)m.\chi\left(\mathcal{I}_{\widetilde{C}}\big{|}_{E}(i)\right)=\chi\left(\mathcal{O}_{E}(i)\right)-\chi\left(\mathcal{O}_{\widetilde{C}\cap E}(i)\right)=h^{0}\left(\mathcal{O}_{E}(i)\right)-\operatorname{mult}_{p}C=\binom{i+n-1}{n-1}-m.

Negate both sides to get

(3) χ(C~|E(i))=m(i+n1n1),-\chi\left(\mathcal{I}_{\widetilde{C}}\big{|}_{E}(i)\right)=m-\binom{i+n-1}{n-1},

and let 0\ell_{0} be the largest integer such that

m(0+n1n1)0.m-\binom{\ell_{0}+n-1}{n-1}\geq 0.

Then by definition

(4) m(0+nn1)<0(0+n)n1n1>m0>((n1)m)1n1n.m-\binom{\ell_{0}+n}{n-1}<0\implies\frac{(\ell_{0}+n)^{n-1}}{n-1}>m\implies\ell_{0}>\left((n-1)m\right)^{\frac{1}{n-1}}-n.

We will now use (3) to add up the contributions of each term:

i=02χ(C~|E(i))=i=00[m(i+n1n1)]=m0(0+nn)m(n1n01).-\sum_{i=0}^{\ell-2}\chi\left(\mathcal{I}_{\widetilde{C}}\big{|}_{E}(i)\right)=\sum_{i=0}^{\ell_{0}}\left[m-\binom{i+n-1}{n-1}\right]=m\cdot\ell_{0}-\binom{\ell_{0}+n}{n}\geq m\cdot\left(\frac{n-1}{n}\cdot\ell_{0}-1\right).

By the Claim and (4), it follows that

length(τ)>m((n1)nn1nm1n1n)=(n1)nn1nmnn1nm.\text{length}(\tau)>m\cdot\left(\frac{(n-1)^{\frac{n}{n-1}}}{n}\cdot m^{\frac{1}{n-1}}-n\right)=\frac{(n-1)^{\frac{n}{n-1}}}{n}\cdot m^{\frac{n}{n-1}}-nm.

As for the Claim, part (i) follows from the equation h1(C~𝒪)=c+h1(C~𝒪1)h^{1}(\mathcal{I}_{\widetilde{C}}\otimes\mathcal{O}_{\mathcal{E}_{\ell}})=c_{\ell}+h^{1}(\mathcal{I}_{\widetilde{C}}\otimes\mathcal{O}_{\mathcal{E}_{\ell-1}}) and the fact that

H1(C~|E(1))=0for 0.H^{1}\left(\mathcal{I}_{\widetilde{C}}\big{|}_{E}(\ell-1)\right)=0\qquad\text{for }\ell\gg 0.

For part (ii), we will argue by induction. The base case =1\ell=1 follows from relations (1) and (2). Now assume that the equation holds for some positive integer =j1\ell=j-1. By these same relations,

h1(C~𝒪j)\displaystyle h^{1}\left(\mathcal{I}_{\widetilde{C}}\otimes\mathcal{O}_{\mathcal{E}_{j}}\right) =h1(C~𝒪j1)+cj\displaystyle=h^{1}\left(\mathcal{I}_{\widetilde{C}}\otimes\mathcal{O}_{\mathcal{E}_{j-1}}\right)+c_{j}
=h1(C~𝒪j1)+h1(C~|E(j1))h0(C~𝒪j1)+kj\displaystyle=h^{1}\left(\mathcal{I}_{\widetilde{C}}\otimes\mathcal{O}_{\mathcal{E}_{j-1}}\right)+h^{1}\left(\mathcal{I}_{\widetilde{C}}\big{|}_{E}(j-1)\right)-h^{0}\left(\mathcal{I}_{\widetilde{C}}\otimes\mathcal{O}_{\mathcal{E}_{j-1}}\right)+k_{j}
=h1(C~𝒪j1)+h1(C~|E(j1))(h0(C~|E(j2))+kj1)+kj\displaystyle=h^{1}\left(\mathcal{I}_{\widetilde{C}}\otimes\mathcal{O}_{\mathcal{E}_{j-1}}\right)+h^{1}\left(\mathcal{I}_{\widetilde{C}}\big{|}_{E}(j-1)\right)-\left(h^{0}\left(\mathcal{I}_{\widetilde{C}}\big{|}_{E}(j-2)\right)+k_{j-1}\right)+k_{j}

We may rewrite the inductive hypothesis as

h1(C~𝒪j1)h0(C~|E(j2))=i=0j2χ(C~|E(i))+kj1.h^{1}\left(\mathcal{I}_{\widetilde{C}}\otimes\mathcal{O}_{\mathcal{E}_{j-1}}\right)-h^{0}\left(\mathcal{I}_{\widetilde{C}}\big{|}_{E}(j-2)\right)=-\sum_{i=0}^{j-2}\chi\left(\mathcal{I}_{\widetilde{C}}\big{|}_{E}(i)\right)+k_{j-1}.

Therefore,

h1(C~𝒪j)=i=0j2χ(C~|E(i))+h1(C~|E(j1))+kj,h^{1}\left(\mathcal{I}_{\widetilde{C}}\otimes\mathcal{O}_{\mathcal{E}_{j}}\right)=-\sum_{i=0}^{j-2}\chi\left(\mathcal{I}_{\widetilde{C}}\big{|}_{E}(i)\right)+h^{1}\left(\mathcal{I}_{\widetilde{C}}\big{|}_{E}(j-1)\right)+k_{j},

which is what we want. ∎

For dimension counts, the following expression will be useful:

Lemma 2.5.

Let Yn+fY\subset\mathbb{P}^{n+f} be a complete intersection of dimension n2n\geq 2 and type (a1,,af)(a_{1},\ldots,a_{f}). Then

h0(Y,𝒪())=j1=0a11jf=0af1h0(n,𝒪(j1jf)).h^{0}(Y,\mathcal{O}(\ell))=\sum_{j_{1}=0}^{a_{1}-1}\cdots\sum_{j_{f}=0}^{a_{f}-1}h^{0}\left(\mathbb{P}^{n},\mathcal{O}(\ell-j_{1}-\cdots-j_{f})\right).
Proof.

Consider the Koszul resolution for a complete intersection Yn+fY\subset\mathbb{P}^{n+f} of dimension n2n\geq 2 and type (a1,,af)(a_{1},\ldots,a_{f}). Using the vanishing of higher cohomology of line bundles on n+f\mathbb{P}^{n+f}, we have

h0(Y,𝒪())\displaystyle h^{0}(Y,\mathcal{O}(\ell)) =h0(n+f,𝒪())j=1fh0(n+f,𝒪(aj))\displaystyle=h^{0}(\mathbb{P}^{n+f},\mathcal{O}(\ell))-\sum_{j=1}^{f}h^{0}(\mathbb{P}^{n+f},\mathcal{O}(\ell-a_{j}))
+1u<vn+1h0(n+f,𝒪(auav))\displaystyle\quad+\sum_{1\leq u<v\leq n+1}h^{0}(\mathbb{P}^{n+f},\mathcal{O}(\ell-a_{u}-a_{v}))-\cdots

We may rearrange the terms in pairs as follows:

[h0(n+f,𝒪())h0(n+f,𝒪(a1))]\displaystyle[h^{0}(\mathbb{P}^{n+f},\mathcal{O}(\ell))-h^{0}(\mathbb{P}^{n+f},\mathcal{O}(\ell-a_{1}))]
j=2f[h0(n+f,𝒪(aj))h0(n+f,𝒪(aja1))]+\displaystyle\qquad-\sum_{j=2}^{f}[h^{0}(\mathbb{P}^{n+f},\mathcal{O}(\ell-a_{j}))-h^{0}(\mathbb{P}^{n+f},\mathcal{O}(\ell-a_{j}-a_{1}))]+\cdots
=[(+n+fn+f)(a1+n+fn+f)]\displaystyle=\left[\binom{\ell+n+f}{n+f}-\binom{\ell-a_{1}+n+f}{n+f}\right]
j=2f[(aj+n+fn+f)(aja1+n+fn+f)]+,\displaystyle\qquad-\sum_{j=2}^{f}\left[\binom{\ell-a_{j}+n+f}{n+f}-\binom{\ell-a_{j}-a_{1}+n+f}{n+f}\right]+\cdots,

where the convention we adopt is that the binomial coefficients are zero if the upper index is smaller than the lower index. The expression in each bracket can be replaced by a sum of binomial coefficients. For instance,

(+n+fn+f)(a1+n+fn+f)\displaystyle\binom{\ell+n+f}{n+f}-\binom{\ell-a_{1}+n+f}{n+f} =j=1a1(j+a1+n+f1n+f1)\displaystyle=\sum_{j=1}^{a_{1}}\binom{j+\ell-a_{1}+n+f-1}{n+f-1}
=j=0a11(j+n+f1n+f1).\displaystyle=\sum_{j=0}^{a_{1}-1}\binom{\ell-j+n+f-1}{n+f-1}.

Similarly, the next term can be rewritten as

(aj+n+fn+f)(aja1+n+fn+f)=j=0a11(jaj+n+f1n+f1),\binom{\ell-a_{j}+n+f}{n+f}-\binom{\ell-a_{j}-a_{1}+n+f}{n+f}=\sum_{j=0}^{a_{1}-1}\binom{\ell-j-a_{j}+n+f-1}{n+f-1},

and so on. Adding up the contribution of each term, it follows that

h0(Y,𝒪())=j=0a11h0(W,𝒪(j)),h^{0}(Y,\mathcal{O}(\ell))=\sum_{j=0}^{a_{1}-1}h^{0}(W,\mathcal{O}(\ell-j)),

where Wn+f1W\subset\mathbb{P}^{n+f-1} is a complete intersection of dimension nn and type (a2,,af)(a_{2},\ldots,a_{f}) (compare with the Koszul resolution for WW). By repeating this process, we obtain

h0(Y,𝒪())=j1=0a11jf=0af1h0(n,𝒪(j1jf)).h^{0}(Y,\mathcal{O}(\ell))=\sum_{j_{1}=0}^{a_{1}-1}\cdots\sum_{j_{f}=0}^{a_{f}-1}h^{0}(\mathbb{P}^{n},\mathcal{O}(\ell-j_{1}-\cdots-j_{f})).\qed

Another ingredient that goes into the proof of Theorem 1.6 is an estimate of the minimal degree of curves contained in a generic complete intersection of special type (see Proposition 2.8). The following results are a straightforward generalization of those in [12]. One may also compare with [15, §2]:

Lemma 2.6.

Let d,kd,k be integers. Assume that there is a smooth projective variety YY of dimension NN with a line bundle LL such that LN=dL^{N}=d and kBLk\mid B\cdot L holds for every curves BYB\subset Y. If XN+1X\subset\mathbb{P}^{N+1} is a very general hypersurface of degree dd and CXC\subset X is any curve, then k|N!degCk\big{|}N!\cdot\deg C.

Proof.

We embed YY by LL to get YNY\subset\mathbb{P}^{N} and project generically to get a finite morphism

φ:YY~n+1.\varphi\colon Y\rightarrow\widetilde{Y}\subset\mathbb{P}^{n+1}.

We know that φ\varphi is an isomorphism on an open set, ramified 2:12:1 over a divisor in Y~\widetilde{Y}, and so on up to n:1n:1 over a curve. So for any irreducible curve CY~C\subset\widetilde{Y}, the projection formula gives

φ1(C)L=deg(φ1(C)C)degn+1C\varphi^{-1}(C)\cdot L=\deg(\varphi^{-1}(C)\rightarrow C)\cdot\deg_{\mathbb{P}^{n+1}}C

where deg(φ1(C)C)\deg(\varphi^{-1}(C)\rightarrow C) is some positive integer less than or equal to nn. By our hypothesis, it follows that

kφ1(C)Ln!degC.k\mid\varphi^{-1}(C)\cdot L\mid n!\cdot\deg C.

If Xn+1X\subset\mathbb{P}^{n+1} is a very general hypersurface of degree dd and CXXC_{X}\subset X is a curve, then CXC_{X} can be specialized to some curve CY~Y~C_{\widetilde{Y}}\subset\widetilde{Y}. Therefore, kn!degCXk\mid n!\cdot\deg C_{X}. ∎

Example 2.7 (Van Geemen).

Let (A,L)(A,L) be a very general Abelian variety of dimension m3m\geq 3 with a polarization of type (1,,1,δ)(1,\ldots,1,\delta). Then NS(A)=[L]\operatorname{NS}(A)=\mathbb{Z}[L] and Lm=m!δL^{m}=m!\cdot\delta. In H1(A,)H^{1}(A,\mathbb{Z}), choose a basis {dxi}\{dx_{i}\} such that

c1(L)=dx1dxm+1++dxm1dx2m1+δdxmdx2m.c_{1}(L)=dx_{1}\wedge dx_{m+1}+\cdots+dx_{m-1}\wedge dx_{2m-1}+\delta\cdot dx_{m}\wedge dx_{2m}.

Write ωi\colonequalsdxidxm+i\omega_{i}\colonequals dx_{i}\wedge dx_{m+i} for i=1,,mi=1,\ldots,m. Then

c1(L)(m1)(m1)!=ω1ω2ωm1+δi=1m1ω1ωi^ωm\frac{c_{1}(L)^{\wedge(m-1)}}{(m-1)!}=\omega_{1}\wedge\omega_{2}\wedge\cdots\wedge\omega_{m-1}+\delta\cdot\sum_{i=1}^{m-1}\omega_{1}\wedge\cdots\wedge\widehat{\omega_{i}}\wedge\cdots\wedge\omega_{m}

represents an indivisible integral class in H2m2(A,)H^{2m-2}(A,\mathbb{Z}). By the Hard Lefschetz theorem, the cohomology class of every curve in AA is a rational multiple of c1(L)(m1)c_{1}(L)^{\wedge(m-1)} and hence an integral multiple of c1(L)(m1)/(m1)!c_{1}(L)^{\wedge(m-1)}/(m-1)!, so mδCLm\delta\mid C\cdot L for every curve CAC\subset A.

Proposition 2.8.

Fix integers q1,,qf>2n+f1q_{1},\ldots,q_{f}>2^{n+f-1} such that (qi,qj)=1(q_{i},q_{j})=1 for all 1i<jf1\leq i<j\leq f. Let Xn+fX\subset\mathbb{P}^{n+f} be a very general complete intersection of dimension nn and type

((n+f1)!q1,,(n+f1)!qf).((n+f-1)!\cdot q_{1},\ldots,(n+f-1)!\cdot q_{f}).

If CXC\subset X is a curve, then q1qf(n+f2)!degn+fCq_{1}\cdots q_{f}\mid(n+f-2)!\cdot\deg_{\mathbb{P}^{n+f}}C.

Proof.

Fix ii and let qi>2n+f1q_{i}>2^{n+f-1} be an integer as above. If (A,L)(A,L) is a very general abelian variety of dimension n+f1n+f-1 with a polarization of type (1,,1,qi)(1,\ldots,1,q_{i}), then Debarre-Hulek-Spandaw [5] have shown that LL is very ample. Example 2.7 demonstrates that (n+f1)qiCL(n+f-1)\cdot q_{i}\mid C\cdot L for every curve CAC\subset A. By applying the pair (A,L)(A,L) to Lemma 2.6 and using the fact that Ln+f1=(n+f1)!qiL^{n+f-1}=(n+f-1)!\cdot q_{i}, we see that if X(n+f1)!qin+fX_{(n+f-1)!\cdot q_{i}}\subset\mathbb{P}^{n+f} is a generic hypersurface of degree (n+f1)!qi(n+f-1)!\cdot q_{i} and CX(n+f1)!qiC\subset X_{(n+f-1)!\cdot q_{i}} is any curve, then

(n+f1)qi(n+f1)!degCqi(n+f2)!degC.(n+f-1)q_{i}\mid(n+f-1)!\cdot\deg C\implies q_{i}\mid(n+f-2)!\cdot\deg C.

Now we can vary this argument for all i=1,,fi=1,\ldots,f; in other words, we will apply it to the ff hypersurfaces that intersect to give XX. From the hypothesis that (qi,qj)=1(q_{i},q_{j})=1 for all 1i<jf1\leq i<j\leq f, it follows that for any curve CXC\subset X,

q1qf(n+f2)!degC.q_{1}\cdots q_{f}\mid(n+f-2)!\cdot\deg C.\qed

Proposition 2.8 suggests that there should be uniform lower bounds for the degrees of curves in very general complete intersections of any type (d1,,dk)(d_{1},\ldots,d_{k}), i.e. the following should hold:

Conjecture 2.9.

Given a very general complete intersection variety Xn+fX\subset\mathbb{P}^{n+f} of dimension nn and type (d1,d2,,df)(d_{1},d_{2},\ldots,d_{f}), any curve in XX has degree Dd1df\geq D\cdot d_{1}\cdots d_{f}, where DD is some positive constant depending only on nn and ff.

Note that a positive answer to the question above would streamline the proof of Theorem B by eliminating the need for Proposition 1.7 and Remark 4.4.

3. Proof of Theorem 1.4

Recalling Set-up 1.3, we will induct on the number of points rr. For the base case, the statement is trivial for r2r\leq 2 as soon as b2b\geq 2 since 𝒪Y(1)\mathcal{O}_{Y}(1) is very ample. By induction, we may assume that the theorem holds for r=sr=s where

2s23(n+1)2ab1.2\leq s\leq\frac{2}{3(n+1)^{2}}ab-1.

We want to prove that the theorem holds for r=s+1r=s+1.

Suppose for the sake of contradiction that the theorem fails when r=s+1r=s+1. Then p1,,ps+1Y\exists p_{1},\ldots,p_{s+1}\in Y such that the corresponding divisor

L:=bHi=1s+1(n+1)EiL:=bH-\sum_{i=1}^{s+1}(n+1)E_{i}

on the blow-up Y~𝜇Y\widetilde{Y}\xrightarrow{\mu}Y is not nef. Here, H=μ𝒪Y(1)H=\mu^{\ast}\mathcal{O}_{Y}(1) and EiE_{i} is the exceptional divisor over pip_{i}.

By definition, this means that there is an integral curve C~Y~\widetilde{C}\subset\widetilde{Y} such that

LC~<0.L\cdot\widetilde{C}<0.

We claim that C~\widetilde{C} cannot be contained in some exceptional divisor EjE_{j}, because otherwise C~Ej<0\widetilde{C}\cdot E_{j}<0 would imply C~L>0\widetilde{C}\cdot L>0, which is a contradiction. By the Lefschetz hyperplane theorem and Poincaré duality, C~\widetilde{C} is numerically a \mathbb{Q}-linear combination of terms involving HnH^{n} and EinE_{i}^{n} (note that the mixed terms involving HEiH\cdot E_{i} must vanish because we have blown up a collection of points). One can check that Hn+1=aH^{n+1}=a and (Ei)n+1=1(-E_{i})^{n+1}=-1. Furthermore, the intersection numbers C~H1\widetilde{C}\cdot H\geq 1 and C~Ei0\widetilde{C}\cdot E_{i}\geq 0 must be integers. It follows that the numerical class of C~\widetilde{C} is given by

(5) C~numkaHn+(1)ni=1s+1miEin\widetilde{C}\equiv_{\text{num}}\frac{k}{a}H^{n}+(-1)^{n}\sum_{i=1}^{s+1}m_{i}E_{i}^{n}

where k1k\geq 1 is the degree of the image curve C\colonequalsμ(C~)n+2C\colonequals\mu(\widetilde{C})\subset\mathbb{P}^{n+2} and mi0m_{i}\geq 0 are the multiplicities of CC at pip_{i}. Note that

LC~<0i=1s+1mi>1n+1bk.L\cdot\widetilde{C}<0\implies\sum_{i=1}^{s+1}m_{i}>\frac{1}{n+1}bk.

Since the quantity mi2\sum m_{i}^{2} is minimized when all of the mim_{i} are the same, it follows that

(6) i=1s+1mi>1n+1kb=:γi=1s+1mi2>(γs+1)2(s+1)=k2b2(n+1)2(s+1).\sum_{i=1}^{s+1}m_{i}>\frac{1}{n+1}kb=:\gamma\implies\sum_{i=1}^{s+1}m_{i}^{2}>\left(\frac{\gamma}{s+1}\right)^{2}\cdot(s+1)=\frac{k^{2}b^{2}}{(n+1)^{2}(s+1)}.

On the other hand, our induction hypothesis implies that the divisor LI:=bHYiI(n+1)EiL_{I}:=bH_{Y}-\sum_{i\in I}(n+1)E_{i} is nef for any subset I{1,2,,s+1}I\subset\{1,2,\ldots,s+1\} with #I=s\#I=s. Averaging over all II shows that

Ls+1:=s+1sbHYi=1s+1(n+1)EiL_{s+1}:=\frac{s+1}{s}bH_{Y}-\sum_{i=1}^{s+1}(n+1)E_{i}

is nef. This implies that Ls+1C~0L_{s+1}\cdot\widetilde{C}\geq 0, and hence

(7) i=1s+1mi(s+1)s(n+1)bk.\sum_{i=1}^{s+1}m_{i}\leq\frac{(s+1)}{s(n+1)}bk.

By Lemma 2.2 and Proposition 2.1 applied to CYn+2C\subset Y\subset\mathbb{P}^{n+2}, we have

i=1rmi(mi1)2\displaystyle\sum_{i=1}^{r}\frac{m_{i}(m_{i}-1)}{2} 12(k1)(k2)pg(C)\displaystyle\leq\frac{1}{2}(k-1)(k-2)-p_{g}(C)
12(k1)(k2)12(a2n3)k1\displaystyle\leq\frac{1}{2}(k-1)(k-2)-\frac{1}{2}(a-2n-3)k-1
(8) =12(k2+(2na)k)\displaystyle=\frac{1}{2}(k^{2}+(2n-a)k)

Next, we can combine this with the inequalities in (6) and (7):

k2b2(n+1)2(s+1)<i=1rmi2=i=1rmi(mi1)+i=1rmik2+(2na)k+(s+1)s(n+1)bk.\frac{k^{2}b^{2}}{(n+1)^{2}(s+1)}<\sum_{i=1}^{r}m_{i}^{2}=\sum_{i=1}^{r}m_{i}(m_{i}-1)+\sum_{i=1}^{r}m_{i}\leq k^{2}+(2n-a)k+\frac{(s+1)}{s(n+1)}bk.

After simplifying, we get

(9) [b2(n+1)2(s+1)1]k<2na+(s+1)s(n+1)b.\left[\frac{b^{2}}{(n+1)^{2}\cdot(s+1)}-1\right]\cdot k<2n-a+\frac{(s+1)}{s(n+1)}b.

Recall from our induction set-up that

3s+123(n+1)2abandn2(s+1)s(n+1)13.3\leq s+1\leq\frac{2}{3(n+1)^{2}}ab\quad\text{and}\quad n\geq 2\implies\frac{(s+1)}{s(n+1)}\leq\frac{1}{3}.

In addition, bab\geq a so term on the left hand side of (9) is positive. Solving for kk yields:

(10) k<(2na+b/3)2a3b2a19(2a+(18n7a)2a3b2a)29a.k<\frac{(2n-a+b/3)\cdot 2a}{3b-2a}\leq\frac{1}{9}\left(2a+\frac{(18n-7a)\cdot 2a}{3b-2a}\right)\leq\frac{2}{9}a.

Since all mim_{i} are nonnegative integers (in fact, the induction hypothesis tells us that mi1m_{i}\geq 1 for all ii), the inequality in (3) also gives

12(k1)(k2)12k(a2n3)1i=1rmi(mi1)20ka2n,\frac{1}{2}(k-1)(k-2)-\frac{1}{2}k(a-2n-3)-1\geq\sum_{i=1}^{r}\frac{m_{i}(m_{i}-1)}{2}\geq 0\implies k\geq a-2n,

which contradicts (10) as soon as a18n/7a\geq 18n/7. This completes the proof of Theorem 1.4.

4. Proof of Theorem 1.6

Recalling Set-up 1.5, we will induct on the number of points rr. Let α\colonequalsa1ae1\alpha\colonequals a_{1}\cdots a_{e-1} be the degree of the threefold Ye+2Y\subset\mathbb{P}^{e+2}. For the base case r2r\leq 2, the statement is trivial as soon as d2d\geq 2 since HYH_{Y} is very ample. By induction, we may assume that the theorem holds for r=sr=s where

2s234(3e+2)((e+1)!)eαae1.2\leq s\leq\frac{2}{3^{4}(3e+2)((e+1)!)^{e}}\cdot\alpha a_{e}-1.

We want to prove that the theorem holds for r=s+1r=s+1.

Suppose for the sake of contradiction that the theorem fails when r=s+1r=s+1. Then p1,,ps+1Y\exists p_{1},\ldots,p_{s+1}\in Y such that the corresponding divisor

L:=aeHi=1s+13EiL:=a_{e}H-\sum_{i=1}^{s+1}3E_{i}

on the blow-up Y~\widetilde{Y} is not nef.

By definition, this means that there is an integral curve C~Y~\widetilde{C}\subset\widetilde{Y} such that

LC~<0.L\cdot\widetilde{C}<0.

As before, note that C~\widetilde{C} cannot be contained in some exceptional divisor EjE_{j}. By the Lefschetz hyperplane theorem and Poincaré duality, C~\widetilde{C} is numerically a \mathbb{Q}-linear combination of terms involving H2H^{2} and Ei2E_{i}^{2} (note that the mixed terms involving HEiH\cdot E_{i} must vanish because we have blown up a collection of points). One can check that H3=αH^{3}=\alpha and (Ei)3=1(-E_{i})^{3}=-1. Furthermore, the intersection numbers C~H1\widetilde{C}\cdot H\geq 1 and C~Ei0\widetilde{C}\cdot E_{i}\geq 0 must be integers. It follows that the numerical class of C~\widetilde{C} is given by

(11) C~numkαH2+i=1s+1miEi2\widetilde{C}\equiv_{\text{num}}\frac{k}{\alpha}H^{2}+\sum_{i=1}^{s+1}m_{i}E_{i}^{2}

for some integers k1k\geq 1 and mi0m_{i}\geq 0.

Remark 4.1.

Note that kk is the degree of the image curve C\colonequalsμ(C~)Ye+2C\colonequals\mu(\widetilde{C})\subset Y\subset\mathbb{P}^{e+2} and the integers mim_{i} are the multiplicities of CC at pip_{i}. By the induction hypothesis, we know that mi1m_{i}\geq 1 for all ii.

Using the description (11) for C~\widetilde{C}, the condition LC~<0L\cdot\widetilde{C}<0 reduces to

(12) i=1s+1mi>13aek.\sum_{i=1}^{s+1}m_{i}>\frac{1}{3}a_{e}k.

For a fixed i=1s+1mi\sum_{i=1}^{s+1}m_{i}, the quantity mi3/2\sum m_{i}^{3/2} is minimized when all of the mim_{i} are the same, so setting N\colonequalsaek/3N\colonequals a_{e}k/3 for the right hand side yields

(13) i=1s+1mi3/2>(Ns+1)3/2(s+1)=133/2(s+1)1/2k3/2ae3/2.\sum_{i=1}^{s+1}m_{i}^{3/2}>\left(\frac{N}{s+1}\right)^{3/2}\cdot(s+1)=\frac{1}{3^{3/2}\cdot(s+1)^{1/2}}k^{3/2}a_{e}^{3/2}.

The rest of the proof will be devoted to bounding the expression i=1s+1mi3/2\sum_{i=1}^{s+1}m_{i}^{3/2} from above. We will first use a dimension count to approximate CC as a complete intersection curve. The key point is that if we can find an effective divisor V|𝒪Y()|V\in\absolutevalue{\mathcal{O}_{Y}(\ell)} of degree 13ae\ell\leq\frac{1}{3}a_{e} passing through all of the points pip_{i}, then

LC~<0V~C~<0,L\cdot\widetilde{C}<0\implies\widetilde{V}\cdot\widetilde{C}<0,

where V~\widetilde{V} is the strict transform of VV. It must then follow that C~V~\widetilde{C}\subset\widetilde{V}. This is the basic idea behind:

Lemma 4.2.

With Set-up 1.5 in mind and integers aeae1a1a_{e}\geq a_{e-1}\geq\cdots\geq a_{1}, there are surfaces Vj|𝒪Y(bj)|V_{j}\in\absolutevalue{\mathcal{O}_{Y}(b_{j})} with bj13aeb_{j}\leq\frac{1}{3}a_{e} for j=1,2j=1,2 such that their intersection V1V2V_{1}\cap V_{2} is a curve (not necessarily irreducible) which contains CC as an irreducible component.

Proof.

Recall that

s+1234(3e+2)((e+1)!)eαae.s+1\leq\frac{2}{3^{4}(3e+2)((e+1)!)^{e}}\cdot\alpha a_{e}.

Choose b1b_{1} to be the minimal degree such that there exists a hypersurface V1|𝒪Y(b1)|V_{1}\in\absolutevalue{\mathcal{O}_{Y}(b_{1})} passing through all of the points pip_{i} (1is+11\leq i\leq s+1). We claim that b1ae/3b_{1}\leq\lfloor a_{e}/3\rfloor. For =ae/3\ell=\lfloor a_{e}/3\rfloor, we can apply Lemma 2.5 with f=e1f=e-1, n=3n=3, and Ye+2Y\subset\mathbb{P}^{e+2} a threefold of type (a1,,ae1)(a_{1},\ldots,a_{e-1}) to see that

h0(Y,𝒪())\displaystyle h^{0}(Y,\mathcal{O}(\ell)) =j1=0a1je1=0ae1h0(3,𝒪(j1je1))\displaystyle=\sum_{j_{1}=0}^{a_{1}}\cdots\sum_{j_{e-1}=0}^{a_{e-1}}h^{0}(\mathbb{P}^{3},\mathcal{O}(\ell-j_{1}-\cdots-j_{e-1}))
j1=0a1/(3e)je1=0ae1/(3e)h0(3,𝒪(j1je1))(by truncating the sum)\displaystyle\geq\sum_{j_{1}=0}^{\lfloor a_{1}/(3e)\rfloor}\cdots\sum_{j_{e-1}=0}^{\lfloor a_{e-1}/(3e)\rfloor}h^{0}(\mathbb{P}^{3},\mathcal{O}(\ell-j_{1}-\cdots-j_{e-1}))\qquad(\text{by truncating the sum})
>j1=0a1/(3e)je1=0ae1/(3e)13!(a13eae13e+1)3\displaystyle>\sum_{j_{1}=0}^{\lfloor a_{1}/(3e)\rfloor}\cdots\sum_{j_{e-1}=0}^{\lfloor a_{e-1}/(3e)\rfloor}\frac{1}{3!}\left(\ell-\left\lfloor\frac{a_{1}}{3e}\right\rfloor-\cdots-\left\lfloor\frac{a_{e-1}}{3e}\right\rfloor+1\right)^{3}
16a13eae13e(13ae(e1)ae3e+1)3(setting =ae/3)\displaystyle\geq\frac{1}{6}\cdot\frac{a_{1}}{3e}\cdots\frac{a_{e-1}}{3e}\cdot\left(\left\lfloor\frac{1}{3}a_{e}\right\rfloor-(e-1)\cdot\left\lfloor\frac{a_{e}}{3e}\right\rfloor+1\right)^{3}\qquad(\text{setting }\ell=\lfloor a_{e}/3\rfloor)
16(3e)e+2αae3\displaystyle\geq\frac{1}{6\cdot(3e)^{e+2}}\alpha a_{e}^{3}
>s+1(for e2).\displaystyle>s+1\qquad(\text{for }e\geq 2).

Hence, there exists a hypersurface V1|𝒪Y(ae/3)|V_{1}\in\absolutevalue{\mathcal{O}_{Y}(\lfloor a_{e}/3\rfloor)} of degree b1ae/3b_{1}\leq\lfloor a_{e}/3\rfloor which passes through all of the pip_{i}. We claim that CV1C\subset V_{1}. To see this, observe that the class of the strict transform V~1Y~\widetilde{V}_{1}\subset\widetilde{Y} is given by

V1~linb1Hi=1s+1ciEi\widetilde{V_{1}}\equiv_{\text{lin}}b_{1}H-\sum_{i=1}^{s+1}c_{i}E_{i}

for some integers ci1c_{i}\geq 1 with 1is+11\leq i\leq s+1. By comparing with (12), the condition b1ae/3b_{1}\leq a_{e}/3 implies that V1~C~<0\widetilde{V_{1}}\cdot\widetilde{C}<0 and hence C~V1~\widetilde{C}\subset\widetilde{V_{1}}. Therefore, CV1C\subset V_{1}. Since CC is irreducible and passes through all of the pip_{i} by Remark 4.1, we may assume that V1V_{1} is irreducible.

By a similar argument, choose b2b_{2} to be the smallest positive integer such that there exists a section V2|𝒪V1(b2)|V_{2}\in\absolutevalue{\mathcal{O}_{V_{1}}(b_{2})} containing pip_{i} (1is+11\leq i\leq s+1). Since V1V_{1} is a complete intersection, a similar dimension count using Lemma 2.5 for V1e+2V_{1}\subset\mathbb{P}^{e+2} leads to:

h0(V1,𝒪())\displaystyle h^{0}(V_{1},\mathcal{O}(\ell)) =j0=0b11j1=0a11je1=0ae11h0(2,𝒪(j0j1je1))\displaystyle=\sum_{j_{0}=0}^{b_{1}-1}\sum_{j_{1}=0}^{a_{1}-1}\cdots\sum_{j_{e-1}=0}^{a_{e-1}-1}h^{0}(\mathbb{P}^{2},\mathcal{O}(\ell-j_{0}-j_{1}-\cdots-j_{e-1}))
j1=0a1/(3e)je1=0ae1/(3e)h0(2,𝒪(j1je1))(taking the first term j0=0 and truncating)\displaystyle\geq\sum_{j_{1}=0}^{\lfloor a_{1}/(3e)\rfloor}\cdots\sum_{j_{e-1}=0}^{\lfloor a_{e-1}/(3e)\rfloor}h^{0}(\mathbb{P}^{2},\mathcal{O}(\ell-j_{1}-\cdots-j_{e-1}))\quad\begin{subarray}{c}(\text{taking the first term }\\ j_{0}=0\text{ and truncating})\end{subarray}
>j1=0a1/(3e)je1=0ae1/(3e)12(a13eae13e+1)2(since =ae/3)\displaystyle>\sum_{j_{1}=0}^{\lfloor a_{1}/(3e)\rfloor}\cdots\sum_{j_{e-1}=0}^{\lfloor a_{e-1}/(3e)\rfloor}\frac{1}{2}\left(\ell-\left\lfloor\frac{a_{1}}{3e}\right\rfloor-\cdots-\left\lfloor\frac{a_{e-1}}{3e}\right\rfloor+1\right)^{2}\quad(\text{since }\ell=\lfloor a_{e}/3\rfloor)
12a13eae13e(13ae(e1)ae3e+1)2\displaystyle\geq\frac{1}{2}\cdot\frac{a_{1}}{3e}\cdots\frac{a_{e-1}}{3e}\cdot\left(\lfloor\frac{1}{3}a_{e}\rfloor-(e-1)\cdot\lfloor\frac{a_{e}}{3e}\rfloor+1\right)^{2}
12(3e)e+1αae2\displaystyle\geq\frac{1}{2\cdot(3e)^{e+1}}\alpha a_{e}^{2}
>s+1.\displaystyle>s+1.

Thus, there exists a hypersurface V2|𝒪V1(b2)|V_{2}\in\absolutevalue{\mathcal{O}_{V_{1}}(b_{2})} of degree b2ae/3b_{2}\leq\lfloor a_{e}/3\rfloor passing through all of the pip_{i}. The restriction map H0(Y,𝒪(b2))H0(V1,𝒪(b2))H^{0}(Y,\mathcal{O}(b_{2}))\rightarrow H^{0}(V_{1},\mathcal{O}(b_{2})) is surjective (see [3, Lemma VIII.9]), so without loss of generality we may take V2V_{2} to be a surface in |𝒪Y(b2)|\absolutevalue{\mathcal{O}_{Y}(b_{2})} such that the intersection V1V2V_{1}\cap V_{2} is a curve (not necessarily irreducible). By the same reasoning as above, the condition b2ae/3b_{2}\leq a_{e}/3 implies that V2V_{2} contains CC. This proves that CV1V2C\subset V_{1}\cap V_{2}. ∎

We can now give a bound for the arithmetic genus of CYC\subset Y.

Lemma 4.3.

With the same assumptions as in the previous lemma, the arithmetic genus of CC is bounded from above by

pa(C)(a1++ae1+2ae/3)degCp_{a}(C)\leq(a_{1}+\cdots+a_{e-1}+2a_{e}/3)\cdot\deg C
Proof.

The approach used here follows from ideas of Castelnuovo [4] about bounding the genus of space curves as well as subsequent work of Harris [9]. In summary, let α\alpha_{\ell} be the dimension of the image of the restriction map

H0(Y,𝒪())ρH0(C,𝒪()).H^{0}(Y,\mathcal{O}(\ell))\xrightarrow{\rho_{\ell}}H^{0}(C,\mathcal{O}(\ell)).

We would like to bound the differences αα1\alpha_{\ell}-\alpha_{\ell-1} from below in order to get a lower bound for h0(C,𝒪())h^{0}(C,\mathcal{O}(\ell)). For 0\ell\gg 0 we can then apply Riemann-Roch to obtain a bound on pa(C)p_{a}(C).

Let HH be a generic hyperplane section in e+2\mathbb{P}^{e+2} and consider the intersections

Γ\colonequalsHC,Y2\colonequalsYH.\Gamma\colonequals H\cap C,\qquad Y_{2}\colonequals Y\cap H.

There is another restriction map

H0(Y,Γ())σH0(C,Γ()|C)H0(C,𝒪(1)).H^{0}(Y,\mathcal{I}_{\Gamma}(\ell))\xrightarrow{\sigma_{\ell}}H^{0}(C,\mathcal{I}_{\Gamma}(\ell)\big{|}_{C})\cong H^{0}(C,\mathcal{O}(\ell-1)).

Since H0(Y,𝒪(1))H^{0}(Y,\mathcal{O}(\ell-1)) injects into H0(Y,Γ())H^{0}(Y,\mathcal{I}_{\Gamma}(\ell)) via the multiplication map by the defining equation of HH and the kernels of ρ\rho_{\ell} and σ\sigma_{\ell} are isomorphic, i.e. equal to H0(Y,C())H^{0}(Y,\mathcal{I}_{C}(\ell)), we see that

β\colonequalsh0(Y,𝒪)h0(Y,Γ())αα1.\beta_{\ell}\colonequals h^{0}(Y,\mathcal{O}_{\ell})-h^{0}(Y,\mathcal{I}_{\Gamma}(\ell))\leq\alpha_{\ell}-\alpha_{\ell-1}.

The natural restriction maps H0(Y,𝒪())H0(Y2,𝒪())H^{0}(Y,\mathcal{O}(\ell))\rightarrow H^{0}(Y_{2},\mathcal{O}(\ell)) and H0(Y,Γ())H0(Y2,Γ())H^{0}(Y,\mathcal{I}_{\Gamma}(\ell))\rightarrow H^{0}(Y_{2},\mathcal{I}_{\Gamma}(\ell)) are both surjective, so

β=h0(Y2,𝒪())h0(Y2,Γ()).\beta_{\ell}=h^{0}(Y_{2},\mathcal{O}(\ell))-h^{0}(Y_{2},\mathcal{I}_{\Gamma}(\ell)).

In order to understand the β\beta_{\ell}, set γ0\colonequalsβ0\gamma_{0}\colonequals\beta_{0} and define the differences

γ\colonequalsββ1.\gamma_{\ell}\colonequals\beta_{\ell}-\beta_{\ell-1}.

Furthermore, take a generic hyperplane section Y1\colonequalsHY2Y_{1}\colonequals H^{\prime}\cap Y_{2} (note that dimYi=i\dim Y_{i}=i). Assuming HH^{\prime} is disjoint from Γ\Gamma, we can write down a long sequence on cohomology:

(14) 0H0(Y2,Γ(1))H0(Y2,Γ())()H0(Y1,𝒪())H1(Y2,Γ(1)).0\rightarrow H^{0}(Y_{2},\mathcal{I}_{\Gamma}(\ell-1))\rightarrow H^{0}(Y_{2},\mathcal{I}_{\Gamma}(\ell))\xrightarrow{(\ast)}H^{0}(Y_{1},\mathcal{O}(\ell))\rightarrow H^{1}(Y_{2},\mathcal{I}_{\Gamma}(\ell-1)).

If we let ee_{\ell} be the dimension of the image of ()(\ast), then one can show that

(15) γ=h0(Y1,𝒪())e.\gamma_{\ell}=h^{0}(Y_{1},\mathcal{O}(\ell))-e_{\ell}.

By Serre vanishing, γ=0\gamma_{\ell}=0 for 0\ell\gg 0. For our purposes, we will need an effective bound for when the numbers γ\gamma_{\ell} become zero. This is where Lemma 4.2 will be utilized. Recall that CV1V2C\subset V_{1}\cap V_{2} where Vj|𝒪Y(bj)|V_{j}\in\absolutevalue{\mathcal{O}_{Y}(b_{j})} are surfaces in YY and V1V2V_{1}\cap V_{2} is a (possibly reducible) curve. So

ΓHV1V2\Gamma\subset H\cap V_{1}\cap V_{2}

where the right hand side is also finite. From the Koszul resolution for HV1V2Y2H\cap V_{1}\cap V_{2}\subset Y_{2}, Serre duality, and [3, Lemma VIII.9], it follows that

H1(Y2,IHV1V2())=0for>a1++ae1e2+b1+b2.H^{1}(Y_{2},I_{H\cap V_{1}\cap V_{2}}(\ell))=0\quad\text{for}\quad\ell>a_{1}+\cdots+a_{e-1}-e-2+b_{1}+b_{2}.

The exact sequence 0IHV1V2IΓ𝒪HV1V2Γ00\rightarrow I_{H\cap V_{1}\cap V_{2}}\rightarrow I_{\Gamma}\rightarrow\mathcal{O}_{H\cap V_{1}\cap V_{2}\setminus\Gamma}\rightarrow 0 on Y2Y_{2} implies that

H1(Y2,Γ())=0if>a1++ae1e2+b1+b2.H^{1}(Y_{2},\mathcal{I}_{\Gamma}(\ell))=0\quad\text{if}\quad\ell>a_{1}+\cdots+a_{e-1}-e-2+b_{1}+b_{2}.

Since b1,b2ae/3b_{1},b_{2}\leq a_{e}/3, by (14) and (15) we see that γ=0\gamma_{\ell}=0 for

>a1++ae1+2ae/3e1.\ell>a_{1}+\cdots+a_{e-1}+2a_{e}/3-e-1.

Note that βi=v=0iγv\beta_{i}=\sum_{v=0}^{i}\gamma_{v} so

i=0βi=i=0v=0iγv=i=0(i+1)γi.\sum_{i=0}^{\ell}\beta_{i}=\sum_{i=0}^{\ell}\sum_{v=0}^{i}\gamma_{v}=\sum_{i=0}^{\ell}(\ell-i+1)\gamma_{i}.

From the ideal sheaf sequence for ΓY\Gamma\subset Y and the fact that H1(Y,Γ())=0H^{1}(Y,\mathcal{I}_{\Gamma}(\ell))=0 for 0\ell\gg 0, it follows that

i=0γi=β=d,0.\sum_{i=0}^{\ell}\gamma_{i}=\beta_{\ell}=d,\qquad\ell\gg 0.

We may substitute this into the expression above to get

i=0βi=di=0(i1)γi,0.\sum_{i=0}^{\ell}\beta_{i}=\ell d-\sum_{i=0}^{\ell}(i-1)\gamma_{i},\qquad\ell\gg 0.

On the other hand, for 0\ell\gg 0 we have αi=0βi\alpha_{\ell}\geq\sum_{i=0}^{\ell}\beta_{i}. From the definition of α\alpha_{\ell} and Riemann-Roch, it follows that for 0\ell\gg 0:

pa(C)=dα+1i=0(i1)γi+1.p_{a}(C)=d\ell-\alpha_{\ell}+1\leq\sum_{i=0}^{\ell}(i-1)\gamma_{i}+1.

So we would like to produce a function γimax\gamma_{i}^{\text{max}} which maximizes the sum above, subject to the constraint i=1γi=d\sum_{i=1}^{\ell}\gamma_{i}=d for 0\ell\gg 0.

Using the fact that γ=0\gamma_{\ell}=0 for >a1++ae1+2ae/3e1\ell>a_{1}+\cdots+a_{e-1}+2a_{e}/3-e-1, we can write down a simple example of a function γimax\gamma_{i}^{\text{max}} that produces a rough upper bound for the arithmetic genus:

γmax={degC if =a1++ae1+2ae/3e1,0 otherwise.\gamma_{\ell}^{\text{max}}=\begin{cases}\deg C&\text{ if }\ell=a_{1}+\cdots+a_{e-1}+\lfloor 2a_{e}/3\rfloor-e-1,\\ 0&\text{ otherwise}.\end{cases}

We then obtain

pa(C)\displaystyle p_{a}(C) 1+i=0(i1)γimax\displaystyle\leq 1+\sum_{i=0}^{\ell}(i-1)\gamma_{i}^{\text{max}}
=1+(a1++ae1+2ae/3e2)degC\displaystyle=1+(a_{1}+\cdots+a_{e-1}+\lfloor 2a_{e}/3\rfloor-e-2)\deg C
(a1++ae1+2ae/3)degC\displaystyle\leq(a_{1}+\cdots+a_{e-1}+2a_{e}/3)\deg C\qed

Now we will use the induction hypothesis, which says that the divisor

LI:=aeHiI3EiL_{I}:=a_{e}H-\sum_{i\in I}3E_{i}

is nef for any subset I{1,2,,s+1}I\subset\{1,2,\ldots,s+1\} with #I=s\#I=s. Averaging over all II shows that

Ls+1:=s+1saeHYi=1s+13EiL_{s+1}:=\frac{s+1}{s}a_{e}H_{Y}-\sum_{i=1}^{s+1}3E_{i}

is nef. This implies that Ls+1C~0L_{s+1}\cdot\widetilde{C}\geq 0, and hence

(16) i=1s+1mis+13saek.\sum_{i=1}^{s+1}m_{i}\leq\frac{s+1}{3s}a_{e}k.

We can now give an upper bound for

23/23i=1s+1mi3/2\frac{2^{3/2}}{3}\sum_{i=1}^{s+1}m_{i}^{3/2}

by integrating the various estimates that have been established so far. Applying Proposition 2.4 to piCYp_{i}\in C\subset Y (1is+11\leq i\leq s+1) with n=3n=3 gives:

pa(C)pg(C)i=1s+123/23mi3/22mi.p_{a}(C)-p_{g}(C)\geq\sum_{i=1}^{s+1}\frac{2^{3/2}}{3}\cdot m_{i}^{3/2}-2m_{i}.

This may be rewritten as

i=1s+123/23mi3/2\displaystyle\sum_{i=1}^{s+1}\frac{2^{3/2}}{3}\cdot m_{i}^{3/2} pa(C)pg(C)+2i=1s+1mi\displaystyle\leq p_{a}(C)-p_{g}(C)+2\sum_{i=1}^{s+1}m_{i}
pa(C)+aek\displaystyle\leq p_{a}(C)+a_{e}k
(a1++ae1+53ae)k,\displaystyle\leq(a_{1}+\cdots+a_{e-1}+\frac{5}{3}a_{e})k,

where the second step follows from s2s\geq 2 and (16). Next, combine this with the lower bound coming from (13):

23/235/2(s+1)1/2k3/2ae3/2<i=1s+123/23mi3/2(a1++ae1+53ae)k.\frac{2^{3/2}}{3^{5/2}\cdot(s+1)^{1/2}}k^{3/2}a_{e}^{3/2}<\sum_{i=1}^{s+1}\frac{2^{3/2}}{3}\cdot m_{i}^{3/2}\leq(a_{1}+\cdots+a_{e-1}+\frac{5}{3}a_{e})k.

We can square both sides, solve for kk, and simplify using the inequalities a1aea_{1}\leq\cdots\leq a_{e}:

k\displaystyle k <3523(a1++ae1+5ae/3)2ae3(s+1)\displaystyle<\frac{3^{5}}{2^{3}}\cdot\frac{(a_{1}+\cdots+a_{e-1}+5a_{e}/3)^{2}}{a_{e}^{3}}(s+1)
352322[(e+2/3)ae]2ae3(s+1)\displaystyle\leq\frac{3^{5}}{2^{3}}\cdot\frac{2^{2}[(e+2/3)a_{e}]^{2}}{a_{e}^{3}}(s+1)
=33(3e+2)22ae(s+1)\displaystyle=\frac{3^{3}(3e+2)^{2}}{2a_{e}}(s+1)
(3e+2)3((e+1)!)eα<1e!((e+1)!)e1α.\displaystyle\leq\frac{(3e+2)}{3\cdot((e+1)!)^{e}}\cdot\alpha<\frac{1}{e!((e+1)!)^{e-1}}\alpha.

On the other hand, recall from Set-up 1.5 that the integers aia_{i} are of the form ai=(e+1)!qia_{i}=(e+1)!\cdot q_{i} for some integers qiq_{i} which are pairwise coprime (i=1,,e1i=1,\ldots,e-1). By applying Proposition 2.8 to CYC\subset Y with n=3n=3 and f=e1f=e-1, it follows that

k1e!q1qe1=1e!((e+1)!)e1α,k\geq\frac{1}{e!}\cdot q_{1}\cdots q_{e-1}=\frac{1}{e!((e+1)!)^{e-1}}\alpha,

which is a contradiction. This completes the proof of Theorem 1.6.

Remark 4.4.

Lastly, we would like to explain how to choose the numbers aia_{i} with the desired properties in the proof of Theorem B. One method is to invoke a theorem of Sondow [17], which can be viewed as an extension of Bertrand’s postulate. Let π(x)\pi(x) denote the number of primes not exceeding xx, and recall:

Definition 4.5.

For n1n\geq 1, the nnth Ramanujan prime is the smallest positive integer RnR_{n} such that if xRnx\geq R_{n}, then π(x)π(12x)n\pi(x)-\pi(\frac{1}{2}x)\geq n.

By [17, Theorem 2], the nnth Ramanujan prime satisfies the inequalities

2nlog2n<Rn<4nlog4n,(n1).2n\log 2n<R_{n}<4n\log 4n,\qquad(n\geq 1).

For our purposes, the upper bound will suffice. Set n=e1n=e-1 and let d1de1d_{1}\leq\cdots\leq d_{e-1} be integers such that

d1(e+1)!4(e1)log4(e1).d_{1}\geq(e+1)!\cdot 4(e-1)\log 4(e-1).

Then this lower bound holds for all did_{i} so Sondow’s theorem tells us that there are always e1e-1 primes between di/(2(e+1)!)d_{i}/(2\cdot(e+1)!) and di/(e+1)!d_{i}/(e+1)! for i=1,,e1i=1,\ldots,e-1. Working in reverse order i=e1,e2,,1i=e-1,e-2,\ldots,1, it follows that we can always choose distinct primes qiq_{i} such that

12di(e+1)!<qidi(e+1)!.\frac{1}{2}\cdot\frac{d_{i}}{(e+1)!}<q_{i}\leq\frac{d_{i}}{(e+1)!}.

Finally, we may define ai\colonequals(e+1)!qia_{i}\colonequals(e+1)!\cdot q_{i}.

References

  • [1] F. Bastianelli, C. Ciliberto, F. Flamini, and P. Supino, A note on gonality of curves on general hypersurfaces, Bollettino dell’Unione Matematica Italiana, 11 (2018), pp. 31–38.
  • [2] F. Bastianelli, P. De Poi, L. Ein, R. Lazarsfeld, and B. Ullery, Measures of irrationality for hypersurfaces of large degree, Compositio Mathematica, 153 (2017), pp. 2368–2393.
  • [3] A. Beauville, Complex algebraic surfaces, vol. 34 of London Mathematical Society Student Texts, Cambridge University Press, Cambridge, second ed., 1996.
  • [4] G. Castelnuovo, Sui multipli di una serie lineare di gruppi di punti appartenente ad una curva algebrica, Rendiconti del circolo Matematico di Palermo, 7 (1893), pp. 89–110.
  • [5] O. Debarre, K. Hulek, and J. Spandaw, Very ample linear systems on abelian varieties, Mathematische Annalen, 300 (1994), pp. 181–202.
  • [6] L. Ein, Subvarieties of generic complete intersections, Inventiones Mathematicae, 94 (1988), pp. 163–169.
  • [7] W. Fulton, Intersection Theory, vol. 2 of Ergebnisse der Math. und ihrer Grenzgebiete (3), Springer-Verlag, Berlin, second ed., 1998.
  • [8] F. Gounelas and A. Kouvidakis, Measures of irrationality of the Fano surface of a cubic threefold, Transactions of the American Mathematical Society, 371 (2019), pp. 7111–7133.
  • [9] J. Harris, The genus of space curves, Mathematische Annalen, 249 (1980), pp. 191–204.
  • [10] J. Hotchkiss, C. C. Lau, and B. Ullery, The gonality of complete intersection curves, Journal of Algebra, 560 (2020), pp. 579–608.
  • [11] A. Ito, Seshadri constants via toric degenerations, Journal für die reine und angewandte Mathematik, 2014 (2014), pp. 151–174.
  • [12] J. Kollár, Trento examples, Classification of irregular varieties (Trento, 1990), 1515 (1992), pp. 136–139.
  • [13] R. Lazarsfeld, Lectures on linear series, with the assistance of Guillermo Fernández del Busto, IAS/Park City Math. Ser, 3 (1997), pp. 161–219.
  • [14]  , Positivity in algebraic geometry I: Classical setting: line bundles and linear series, vol. 48, Springer, 2004.
  • [15] M. Paulsen, On the degree of algebraic cycles on hypersurfaces, arXiv:2109.06303, (2021).
  • [16] G. Smith, Covering gonalities of complete intersections in positive characteristic, arXiv:2005.08878, (2020).
  • [17] J. Sondow, Ramanujan primes and Bertrand’s postulate, The American Mathematical Monthly, 116 (2009), pp. 630–635.
  • [18] D. Stapleton, The degree of irrationality of very general hypersurfaces in some homogeneous spaces, PhD thesis, PhD thesis, Stony Brook University, 2017.
  • [19] D. Stapleton and B. Ullery, The degree of irrationality of hypersurfaces in various Fano varieties, Manuscripta Mathematica, 161 (2020), pp. 377–408.
  • [20] C. Voisin, On a conjecture of Clemens on rational curves on hypersurfaces, Journal of Differential Geometry, 44 (1996), pp. 200–213.

Department of Mathematics, Harvard University, Cambridge, Massachusetts 02138

E-mail address: [email protected]