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

\DeclareSymbolFont

AMSbUmsbmn

Unbounded negativity on rational
surfaces in positive characteristic

Raymond Cheng Department of Mathematics
Columbia University
2990 Broadway
New York, NY, 10027, United States of America
[email protected]
 and  Remy van Dobben de Bruyn Department of Mathematics
Princeton University
Fine Hall
Washington Road
Princeton, NJ 08544
United States of America
Institute for Advanced Study
1 Einstein Drive
Princeton, NJ 08540
United States of America
[email protected]
(Date: 2 March 2021)
Abstract.

We give explicit blowups of the projective plane in positive characteristic that contain smooth rational curves of arbitrarily negative self-intersection, showing that the Bounded Negativity Conjecture fails even for rational surfaces in positive characteristic.

Key words and phrases:
Bounded negativity conjecture, rational surfaces, positive characteristic, Fermat varieties, line configurations, Bogomolov–Miyaoka–Yau inequality.
2010 Mathematics Subject Classification:
14C17 (primary); 14C20, 14G17, 14J26, 14E05 (secondary)

Introduction

A smooth projective surface XX over an algebraically closed field is said to have Bounded Negativity if there exists a positive integer b(X)b(X) such that C2b(X)C^{2}\geq-b(X) for any reduced curve CXC\subset X. A folklore conjecture, going back to Enriques and discussed in [Har10, Conjecture I.2.1] and [BHK+13, Conjecture 1.1], is the

Bounded Negativity Conjecture. — Any smooth projective surface in characteristic 0 has Bounded Negativity.

The assumption on the characteristic cannot be dropped: if CC is a curve over 𝐅¯p\bar{\mathbf{F}}_{p}, then the graph ΓFeC×C\Gamma_{F^{e}}\subseteq C\times C of the pep^{e}-th power Frobenius endomorphism has self-intersection pe(22g)p^{e}(2-2g), which becomes arbitrarily negative as ee\to\infty when g2g\geq 2. Nonetheless, it is conceivable that certain geometric assumptions on the surface may still guarantee Bounded Negativity in positive characteristic. For instance, [BBC+12, discussion preceding Example 3.3.3] and [Har18, Conjecture 2.1.2] ask whether smooth rational surfaces over a field of positive characteristic have Bounded Negativity. We give a negative answer to this question:

Main Theorem. — Let kk be an algebraically closed field of characteristic p>0p>0, let mm be a positive integer invertible in kk, and let RmR_{m} be the blowup of 𝐏2\mathbf{P}^{2} along

Zm{[x0:x1:x2]|x0m=x1m=x2m}.Z_{m}\coloneqq\Set{[x_{0}:x_{1}:x_{2}]}{x_{0}^{m}=x_{1}^{m}=x_{2}^{m}}.

Let C1=V(x0+x1+x2)𝐏2C_{1}=V(x_{0}+x_{1}+x_{2})\subseteq\mathbf{P}^{2}, and for d1d\geq 1 invertible in kk, write Cd𝐏2C_{d}\subseteq\mathbf{P}^{2} for the image of

ϕd:C1\displaystyle\phi_{d}\colon C_{1} 𝐏2\displaystyle\to\mathbf{P}^{2}
[x0:x1:x2]\displaystyle[x_{0}:x_{1}:x_{2}] [x0d:x1d:x2d].\displaystyle\mapsto[x_{0}^{d}:x_{1}^{d}:x_{2}^{d}].

If dm=pe1dm=p^{e}-1 for some positive integer ee, then the strict transform C~dRm\widetilde{C}_{d}\subseteq R_{m} of CdC_{d} is a smooth rational curve with C~d2=d(3m)1\widetilde{C}_{d}^{2}=d(3-m)-1. Thus, if m>3m>3, the rational surface RmR_{m} does not have Bounded Negativity over kk.

Since 𝐏2\mathbf{P}^{2} has Bounded Negativity, this shows that [BDRH+15, Problem 1.2] has a negative answer in positive characteristic:

Corollary. — Bounded Negativity is not a birational property of smooth projective surfaces in positive characteristic.   \blacksquare\blacksquare

In fact, since every smooth projective surface XX admits a finite morphism X𝐏2X\to\mathbf{P}^{2}, pulling back the blowup Rm𝐏2R_{m}\to\mathbf{P}^{2} gives a blowup X~X\widetilde{X}\to X with a finite morphism X~Rm\widetilde{X}\to R_{m}. Pulling back the curves C~d\widetilde{C}_{d} to X~\widetilde{X} shows:

Corollary. — If XX is a smooth projective surface over an algebraically closed field kk of positive characteristic, then there exists a blowup X~X\widetilde{X}\to X such that X~\widetilde{X} does not have Bounded Negativity.   \blacksquare\blacksquare

In §​​ 1, we give a direct proof of the Main Theorem. In §​​ 2, we realise the plane curves CdC_{d} as norms of line configuration, thereby deriving equations for them. In §​​ 3, we view RmR_{m} as an isotrivial family of diagonal curves over C1C_{1} and relate the curves C~d\widetilde{C}_{d} on RmR_{m} to graphs of Frobenius morphisms on Fermat curves. Finally, we close in §​​ 4 with some questions and remarks towards characteristic zero.

Sections 2 and 3 each give alternative methods for computing the self-intersections of C~d\widetilde{C}_{d}. Given the simplicity of the formulas for C~d\widetilde{C}_{d} and the many connections to other well-studied examples, it is surprising that these curves have not been found before.

Notation

Throughout the paper, kk will be an algebraically closed field of arbitrary characteristic, and mm and dd will denote positive integers invertible in kk. We will use the notation of the Main Theorem throughout.

1. Proof of Main Theorem

In this section, fix mm and write RRmR\coloneqq R_{m} for the blowup of 𝐏2\mathbf{P}^{2} along ZZmZ\coloneqq Z_{m}. The generators s0=x1mx2ms_{0}=x_{1}^{m}-x_{2}^{m}, s1=x2mx0ms_{1}=x_{2}^{m}-x_{0}^{m}, and s2=x0mx1ms_{2}=x_{0}^{m}-x_{1}^{m} of the ideal of ZZ give a closed immersion R𝐏2×𝐏2R\hookrightarrow\mathbf{P}^{2}\times\mathbf{P}^{2}. Since s0+s1+s2=0s_{0}+s_{1}+s_{2}=0, one of the sis_{i} can be eliminated at the expense of breaking the symmetry in the computations below.

1.1. Lemma. — The embedding R𝐏2×𝐏2R\hookrightarrow\mathbf{P}^{2}\times\mathbf{P}^{2} realises RR as the complete intersection

{([x0:x1:x2],[y0:y1:y2])𝐏2×𝐏2|y0+y1+y2=0x0my0+x1my1+x2my2=0}\Set{\big{(}[x_{0}:x_{1}:x_{2}],[y_{0}:y_{1}:y_{2}]\big{)}\in\mathbf{P}^{2}\times\mathbf{P}^{2}}{\begin{array}[]{c}y_{0}+y_{1}+y_{2}=0\\ x_{0}^{m}y_{0}+x_{1}^{m}y_{1}+x_{2}^{m}y_{2}=0\end{array}}

of degrees (0,1)(0,1) and (m,1)(m,1) in 𝐏2×𝐏2\mathbf{P}^{2}\times\mathbf{P}^{2}. In particular, KR=𝒪R(m3,1)K_{R}=\mathcal{O}_{R}(m-3,-1).

Proof. The generators s0,s1,s2s_{0},s_{1},s_{2} of the ideal of ZZ identify RR as

{([x0:x1:x2],[y0:y1:y2])𝐏2×𝐏2|y0(x2mx0m)=y1(x1mx2m)y1(x0mx1m)=y2(x2mx0m)y2(x1mx2m)=y0(x0mx1m)}.\Set{\big{(}[x_{0}:x_{1}:x_{2}],[y_{0}:y_{1}:y_{2}]\big{)}\in\mathbf{P}^{2}\times\mathbf{P}^{2}}{\begin{array}[]{c}y_{0}(x_{2}^{m}-x_{0}^{m})=y_{1}(x_{1}^{m}-x_{2}^{m})\\ y_{1}(x_{0}^{m}-x_{1}^{m})=y_{2}(x_{2}^{m}-x_{0}^{m})\\ y_{2}(x_{1}^{m}-x_{2}^{m})=y_{0}(x_{0}^{m}-x_{1}^{m})\end{array}}.

The relation s0+s1+s2=0s_{0}+s_{1}+s_{2}=0 shows that RR is contained in the locus y0+y1+y2=0y_{0}+y_{1}+y_{2}=0. The equation y0(x2mx0m)=y1(x1mx2m)y_{0}(x_{2}^{m}-x_{0}^{m})=y_{1}(x_{1}^{m}-x_{2}^{m}) can be rewritten as (y0+y1)x2m=x0my0+x1my1(y_{0}+y_{1})x_{2}^{m}=x_{0}^{m}y_{0}+x_{1}^{m}y_{1}, which is equivalent to x0my0+x1my1+x2my2=0x_{0}^{m}y_{0}+x_{1}^{m}y_{1}+x_{2}^{m}y_{2}=0 since y0+y1+y2=0y_{0}+y_{1}+y_{2}=0. The same holds for the other two equations by symmetry. The final statement then follows from the adjunction formula since K𝐏2×𝐏2=𝒪(3,3)K_{\mathbf{P}^{2}\times\mathbf{P}^{2}}=\mathcal{O}(-3,-3).   \blacksquare\blacksquare

Alternatively, one can observe that the complete intersection of Lemma 1 maps birationally onto its first factor, where the fibres are points when [x0m:x1m:x2m][1:1:1][x_{0}^{m}:x_{1}^{m}:x_{2}^{m}]\neq[1:1:1] and lines otherwise.

1.2. Lemma. — If chark=p>0\operatorname{char}k=p>0 and dm=pe1dm=p^{e}-1 for some positive integer ee, then the map ϕ~d:C1𝐏2×𝐏2\widetilde{\phi}_{d}\colon C_{1}\to\mathbf{P}^{2}\times\mathbf{P}^{2} given by

[x0:x1:x2]([x0d:x1d:x2d],[x0:x1:x2])[x_{0}:x_{1}:x_{2}]\mapsto\big{(}[x_{0}^{d}:x_{1}^{d}:x_{2}^{d}],[x_{0}:x_{1}:x_{2}]\big{)}

lands in RR. In particular, it is the unique map lifting ϕd:C1𝐏2\phi_{d}\colon C_{1}\to\mathbf{P}^{2}.

Proof. Since x0+x1+x2=0x_{0}+x_{1}+x_{2}=0, the image of ϕ~d\widetilde{\phi}_{d} is contained in the locus y0+y1+y2=0y_{0}+y_{1}+y_{2}=0. Since dm=pe1dm=p^{e}-1, the expression x0my0+x1my1+x2my2x_{0}^{m}y_{0}+x_{1}^{m}y_{1}+x_{2}^{m}y_{2} pulls back to x0pe+x1pe+x2pex_{0}^{p^{e}}+x_{1}^{p^{e}}+x_{2}^{p^{e}}, which vanishes because the pep^{e}-th power Frobenius is an endomorphism. Thus ϕ~d\widetilde{\phi}_{d} is a lift of ϕd\phi_{d} to RR, and it is the unique lift since the first projection pr1:R𝐏2\operatorname{pr}_{1}\colon R\to\mathbf{P}^{2} is birational.   \blacksquare\blacksquare

1.3. Corollary. — The map ϕ~d:C1R\widetilde{\phi}_{d}\colon C_{1}\to R is a closed immersion, whose image C~d\widetilde{C}_{d} is a smooth rational curve in RR with C~d2=d(3m)1\widetilde{C}_{d}^{2}=d(3-m)-1.

Proof. The first two statements follow from the coordinate expression in Lemma 1, since ϕ~d\widetilde{\phi}_{d} embeds C1C_{1} linearly into the second factor. The same expression shows that ϕ~d𝒪R(a,b)=𝒪C1(da+b)\widetilde{\phi}_{d}^{*}\mathcal{O}_{R}(a,b)=\mathcal{O}_{C_{1}}(da+b), so KRC~d=d(m3)1K_{R}\cdot\widetilde{C}_{d}=d(m-3)-1 by Lemma 1. Then the adjunction formula shows that

C~d2=2KRC~d=d(3m)1.\widetilde{C}_{d}^{2}=-2-K_{R}\cdot\widetilde{C}_{d}=d(3-m)-1.

This completes the proof of the Main Theorem.   \blacksquare\blacksquare

A consequence of Corollary 1 is that the singularities of CdC_{d} are contained in ZZ. However, the individual multiplicities are not so easy to determine. For example, in Lemma 3 we will compute the multiplicity of CdC_{d} at [1:1:1][1:1:1] in terms of point counts on Fermat curves.

2. Relation with line configurations

In this section, we observe that the dd-th power maps πd:𝐏2𝐏2\pi_{d}\colon\mathbf{P}^{2}\to\mathbf{P}^{2} are finite Galois morphisms such that πdCd\pi_{d}^{*}C_{d} is the union of the Galois translates of C1C_{1}. Thus the CdC_{d} are norms of line configurations, from which we derive in Corollary 2 a formal product formula for the equation of the plane curves CdC_{d}. In the second half of this section, we observe that in characteristic p>0p>0 and for qq a power of pp, the curve Cq1C_{q-1} comes from a subconfiguration of the set of all 𝐅q\mathbf{F}_{q}-rational lines. This allows us to show in Corollary 2 that an equation of Cq1C_{q-1} in this case is the complete homogeneous polynomial of degree q1q-1.

2.1. Power Maps. For any integer a1a\geq 1 invertible in kk, write πa\pi_{a} for the aa-th power map 𝐏2𝐏2\mathbf{P}^{2}\to\mathbf{P}^{2}. Since πaZm=Zam\pi_{a}^{*}Z_{m}=Z_{am}, the map πa\pi_{a} lifts to a finite morphism π~a:RamRm\widetilde{\pi}_{a}\colon R_{am}\to R_{m} given by

([x0:x1:x2],[y0:y1,y2])([x0a:x1a:x2a],[y0:y1,y2]).\big{(}[x_{0}:x_{1}:x_{2}],[y_{0}:y_{1},y_{2}]\big{)}\mapsto\big{(}[x_{0}^{a}:x_{1}^{a}:x_{2}^{a}],[y_{0}:y_{1},y_{2}]\big{)}.

Since aa is invertible in kk, both πa\pi_{a} and π~a\widetilde{\pi}_{a} are finite Galois with group G=𝝁a3/𝝁aG=\boldsymbol{\mu}_{a}^{3}/\boldsymbol{\mu}_{a}, where (ζ0,ζ1,ζ2)G(\zeta_{0},\zeta_{1},\zeta_{2})\in G acts on 𝐏2\mathbf{P}^{2} via

[x0:x1:x2][ζ0x0:ζ1x1:ζ2x2].[x_{0}:x_{1}:x_{2}]\mapsto[\zeta_{0}x_{0}:\zeta_{1}x_{1}:\zeta_{2}x_{2}].

This gives a tower of extensions

R4{R_{4}}R6{R_{6}}R9{R_{9}}  

\iddots\iddots

R2{R_{2}}R3{R_{3}}  

\iddots\iddots

R1{R_{1}}

indexed by the poset of positive integers invertible in kk under the divisibility relation.

2.2. Lemma. — If a,d1a,d\geq 1 are invertible in kk, then

  1. (i)

    The map ϕd:C1𝐏2\phi_{d}\colon C_{1}\to\mathbf{P}^{2} is unramified and birational onto its image;

  2. (ii)

    The inverse image πaCad\pi_{a}^{*}C_{ad} is totally split into the GG-translates of CdC_{d}.

Proof. The Jacobian (dx0d1,dx1d1,dx2d1)(d\cdot x_{0}^{d-1},d\cdot x_{1}^{d-1},d\cdot x_{2}^{d-1}) of ϕd\phi_{d} only vanishes when x0=x1=x2=0x_{0}=x_{1}=x_{2}=0, showing that ϕd\phi_{d} is unramified. Then the map C1CdνC_{1}\to C_{d}^{\nu} to the normalisation of CdC_{d} is unramified, hence an isomorphism since it is an étale map of smooth projective rational curves, proving (i).

Since πaϕd=ϕad\pi_{a}\circ\phi_{d}=\phi_{ad}, part (i) shows that πa\pi_{a} maps CdC_{d} birationally onto its image. This shows that the decomposition group of CdC_{d} is trivial, so no two GG-translates ζCd\zeta C_{d} of CdC_{d} coincide and CadC_{ad} is totally split under πa\pi_{a}.   \blacksquare\blacksquare

2.3. Corollary. — If dd is invertible in kk, then the homogeneous ideal of Cd𝐏2C_{d}\subseteq\mathbf{P}^{2} is generated by

fdNπd,𝒪𝐏2/𝒪𝐏2(x01/d+x11/d+x21/d)=ζ,ζ𝝁d(x01/d+ζx11/d+ζx21/d).f_{d}\coloneqq N_{\pi_{d,*}\mathcal{O}_{\mathbf{P}^{2}}/\mathcal{O}_{\mathbf{P}^{2}}}\big{(}x_{0}^{1/d}+x_{1}^{1/d}+x_{2}^{1/d}\big{)}=\prod_{\zeta,\zeta^{\prime}\in\boldsymbol{\mu}_{d}}\big{(}x_{0}^{1/d}+\zeta x_{1}^{1/d}+\zeta^{\prime}x_{2}^{1/d}\big{)}.

Proof. By Lemma 2 (ii), the inverse image πd1(Cd)\pi_{d}^{-1}(C_{d}) is the union of lines ζ𝝁d3/𝝁dζC1\bigcup_{\zeta\in\boldsymbol{\mu}_{d}^{3}/\boldsymbol{\mu}_{d}}\zeta C_{1}. The result follows since C1C_{1} is cut out by x0+x1+x2=0x_{0}+x_{1}+x_{2}=0.   \blacksquare\blacksquare

2.4. In general, the fdf_{d} are complicated symmetric polynomials. However, in Corollary 2 we will show that the coefficients of fq1f_{q-1} are congruent to 11 modulo pp if qq is a power of a prime pp. For example, for q=p=3q=p=3, we get

N(x012+x112+x212)\displaystyle N\big{(}x_{0}^{\tfrac{1}{2}}+x_{1}^{\tfrac{1}{2}}+x_{2}^{\tfrac{1}{2}}\big{)} =(x012+x112+x212)(x012+x112x212)(x012x112+x212)(x012x112x212)\displaystyle=\Big{(}x^{\tfrac{1}{2}}_{0}+x^{\tfrac{1}{2}}_{1}+x^{\tfrac{1}{2}}_{2}\Big{)}\Big{(}x_{0}^{\tfrac{1}{2}}+x_{1}^{\tfrac{1}{2}}-x_{2}^{\tfrac{1}{2}}\Big{)}\Big{(}x_{0}^{\tfrac{1}{2}}-x_{1}^{\tfrac{1}{2}}+x_{2}^{\tfrac{1}{2}}\Big{)}\Big{(}x_{0}^{\tfrac{1}{2}}-x_{1}^{\tfrac{1}{2}}-x_{2}^{\tfrac{1}{2}}\Big{)}
=x02+x12+x222x0x12x1x22x2x0.\displaystyle=x_{0}^{2}+x_{1}^{2}+x_{2}^{2}-2x_{0}x_{1}-2x_{1}x_{2}-2x_{2}x_{0}.
x02+x12+x22+x0x1+x1x2+x2x0(mod3).\displaystyle\equiv x_{0}^{2}+x_{1}^{2}+x_{2}^{2}+x_{0}x_{1}+x_{1}x_{2}+x_{2}x_{0}\pmod{3}.

In the remainder of this section, assume chark=p>0\operatorname{char}k=p>0 and let qq be a power of pp.

2.5. Finite Field Line Configurations. The configuration of 𝐅q\mathbf{F}_{q}-rational lines in 𝐏2\mathbf{P}^{2} is the union of the lines La={a0x0+a1x1+a2x2=0}L_{a}=\{a_{0}x_{0}+a_{1}x_{1}+a_{2}x_{2}=0\} indexed by a=[a0:a1:a2]𝐏ˇ2(𝐅q)a=[a_{0}:a_{1}:a_{2}]\in\check{\mathbf{P}}^{2}(\mathbf{F}_{q}). Their union is the divisor in 𝐏2\mathbf{P}^{2} cut out by the polynomial

det(x0x1x2x0qx1qx2qx0q2x1q2x2q2)=x0qx1q2x2x0q2x1qx2+x0x1qx2q2x0x1q2x2q+x0q2x1x2qx0qx1x2q2,\det\left(\begin{smallmatrix}x_{0}&x_{1}&x_{2}\\ x_{0}^{q}&x_{1}^{q}&x_{2}^{q}\\ x_{0}^{q^{2}}&x_{1}^{q^{2}}&x_{2}^{q^{2}}\end{smallmatrix}\right)=x_{0}^{q}x_{1}^{q^{2}}x_{2}-x_{0}^{q^{2}}x_{1}^{q}x_{2}+x_{0}x_{1}^{q}x_{2}^{q^{2}}-x_{0}x_{1}^{q^{2}}x_{2}^{q}+x_{0}^{q^{2}}x_{1}x_{2}^{q}-x_{0}^{q}x_{1}x_{2}^{q^{2}},

since the three columns become linearly dependent when x0x_{0}, x1x_{1}, and x2x_{2} satisfy a linear relation over 𝐅q\mathbf{F}_{q}, and the degree equals q2+q+1=|𝐏ˇ2(𝐅q)|q^{2}+q+1=|\check{\mathbf{P}}^{2}(\mathbf{F}_{q})|. Now Lemma 2(ii) shows that πq1Cq1\pi_{q-1}^{*}C_{q-1} consists of the q22q+1q^{2}-2q+1 lines LaL_{a} with all coordinates of a=[a0:a1:a2]a=[a_{0}:a_{1}:a_{2}] nonzero. We can thus derive an equation for Cq1C_{q-1} by extracting factors cutting out the lines LaL_{a} in which aa has a vanishing coordinate. A neat description of the result comes from the following polynomial identity, also observed in [RVVZ01, p. 90]:

2.6. Lemma. — For any nonnegative integer nn, define the polynomials

gnn0+n1+n2=nx0n0x1n1x2n2andhnx0x1nx0nx1+x1x2nx1nx2+x2x0nx2nx0g_{n}\coloneqq\sum_{n_{0}+n_{1}+n_{2}=n}x_{0}^{n_{0}}x_{1}^{n_{1}}x_{2}^{n_{2}}\quad\text{and}\quad h_{n}\coloneqq x_{0}x_{1}^{n}-x_{0}^{n}x_{1}+x_{1}x_{2}^{n}-x_{1}^{n}x_{2}+x_{2}x_{0}^{n}-x_{2}^{n}x_{0}

in 𝐙[x0,x1,x2]\mathbf{Z}[x_{0},x_{1},x_{2}]. Then h2=(x2x1)(x0x2)(x1x0)h_{2}=(x_{2}-x_{1})(x_{0}-x_{2})(x_{1}-x_{0}) and hn=h2gn2h_{n}=h_{2}g_{n-2} for n3n\geq 3.

Proof. Let G(t)n0gntnG(t)\coloneqq\sum_{n\geq 0}g_{n}t^{n} and H(t)n0hntnH(t)\coloneqq\sum_{n\geq 0}h_{n}t^{n} be the generating functions of the gng_{n} and hnh_{n}, respectively. A standard computation gives

G(t)=1(1x0t)(1x1t)(1x2t).G(t)=\frac{1}{(1-x_{0}t)(1-x_{1}t)(1-x_{2}t)}.

On the other hand, writing hn=(x2x1)x0n+(x0x2)x1n+(x1x0)x2nh_{n}=(x_{2}-x_{1})x_{0}^{n}+(x_{0}-x_{2})x_{1}^{n}+(x_{1}-x_{0})x_{2}^{n} gives

H(t)=x2x11x0t+x0x21x1t+x1x01x2t=(x2x1)x2x1+(x0x2)x0x2+(x1x0)x0x1(1x0t)(1x1t)(1x2t)t2.H(t)=\frac{x_{2}-x_{1}}{1-x_{0}t}+\frac{x_{0}-x_{2}}{1-x_{1}t}+\frac{x_{1}-x_{0}}{1-x_{2}t}=\frac{(x_{2}-x_{1})x_{2}x_{1}+(x_{0}-x_{2})x_{0}x_{2}+(x_{1}-x_{0})x_{0}x_{1}}{(1-x_{0}t)(1-x_{1}t)(1-x_{2}t)}t^{2}.

The result follows by recognising the numerator as h2h_{2}.   \blacksquare\blacksquare

2.7. Corollary. — Suppose chark=p>0\operatorname{char}k=p>0 and qq is a power of pp. Then gq1g_{q-1} generates the homogeneous ideal of Cq1𝐏2C_{q-1}\subseteq\mathbf{P}^{2}. In particular, fq1gq1(modp)f_{q-1}\equiv g_{q-1}\pmod{p}.

Proof. Since C1=L[1:1:1]C_{1}=L_{[1:1:1]} is among the 𝐅q\mathbf{F}_{q}-rational lines of 2 and is not a coordinate axis, the points [x0:x1:x2][x_{0}:x_{1}:x_{2}] of C1=V(x0+x1+x2)C_{1}=V(x_{0}+x_{1}+x_{2}) satisfy the equation there divided by x0x1x2x_{0}x_{1}x_{2}:

x0q1x1q21x0q21x1q1+x1q1x2q21x1q21x2q1+x2q1x0q21x2q21x0q1=0.x_{0}^{q-1}x_{1}^{q^{2}-1}-x_{0}^{q^{2}-1}x_{1}^{q-1}+x_{1}^{q-1}x_{2}^{q^{2}-1}-x_{1}^{q^{2}-1}x_{2}^{q-1}+x_{2}^{q-1}x_{0}^{q^{2}-1}-x_{2}^{q^{2}-1}x_{0}^{q-1}=0.

Since C1Cq1C_{1}\to C_{q-1} is [x0:x1:x2][x0q1:x1q1:x2q1][x_{0}:x_{1}:x_{2}]\mapsto[x_{0}^{q-1}:x_{1}^{q-1}:x_{2}^{q-1}], any [x0:x1:x2]Cq1[x_{0}:x_{1}:x_{2}]\in C_{q-1} satisfies

x0x1q+1x0q+1x1+x1x2q+1x1q+1x2+x2x0q+1x2q+1x0=0.x_{0}x_{1}^{q+1}-x_{0}^{q+1}x_{1}+x_{1}x_{2}^{q+1}-x_{1}^{q+1}x_{2}+x_{2}x_{0}^{q+1}-x_{2}^{q+1}x_{0}=0.

So hq+1h_{q+1} vanishes on Cq1C_{q-1}, which by Lemma 2 equals (x2x1)(x0x2)(x1x0)gq1(x_{2}-x_{1})(x_{0}-x_{2})(x_{1}-x_{0})g_{q-1}. The result follows since Cq1C_{q-1} is not contained in any of the lines {x2=x1}\{x_{2}=x_{1}\}, {x0=x2}\{x_{0}=x_{2}\}, or {x1=x0}\{x_{1}=x_{0}\}, and deggq1=q1=degCq1\deg g_{q-1}=q-1=\deg C_{q-1}.   \blacksquare\blacksquare

2.8. Negative Curves via Equations. If m>3m>3 and qq is a power of pp congruent to 11 modulo mm, then the curves C~dRm\widetilde{C}_{d}\subseteq R_{m} with dm=qe1dm=q^{e}-1 of the Main Theorem can therefore be obtained by starting with the very explicit equations

Cqe1=V(n0+n1+n2=qe1x0n0x1n1x2n2)𝐏2,C_{q^{e}-1}=V\left(\sum_{n_{0}+n_{1}+n_{2}=q^{e}-1}x_{0}^{n_{0}}x_{1}^{n_{1}}x_{2}^{n_{2}}\right)\subseteq\mathbf{P}^{2},

blowing up at [1:1:1][1:1:1], pulling back along π~m:RmR1\widetilde{\pi}_{m}\colon R_{m}\to R_{1}, and taking one of the m2m^{2} isomorphic components ζC~d\zeta\widetilde{C}_{d} for ζ𝝁m3/𝝁m\zeta\in\boldsymbol{\mu}_{m}^{3}/\boldsymbol{\mu}_{m}. From this point of view, the self-intersection may be computed as

C~d2=C~dπ~m(C~qe1)ζ1C~d(ζC~d)=(2dm1)3(m1)d=d(3m)1,\widetilde{C}_{d}^{2}=\widetilde{C}_{d}\cdot\widetilde{\pi}_{m}^{*}(\widetilde{C}_{q^{e}-1})-\sum_{\zeta\neq 1}\widetilde{C}_{d}\cdot(\zeta\widetilde{C}_{d})=(2dm-1)-3(m-1)d=d(3-m)-1,

since the intersection number between C~d\widetilde{C}_{d} and a Galois translate by ζ=(ζ0,ζ1,ζ2)G{1}\zeta=(\zeta_{0},\zeta_{1},\zeta_{2})\in G\setminus\{1\} is

C~d(ζC~d)={d,ζ0=ζ1 or ζ1=ζ2 or ζ2=ζ0,0,otherwise.\widetilde{C}_{d}\cdot(\zeta\widetilde{C}_{d})=\begin{cases}d,&\zeta_{0}=\zeta_{1}\text{ or }\zeta_{1}=\zeta_{2}\text{ or }\zeta_{2}=\zeta_{0},\\ 0,&\text{otherwise}.\end{cases}

Indeed, ζC~d\zeta\widetilde{C}_{d} is the image of the morphism ζϕ~d\zeta\circ\widetilde{\phi}_{d} given by

[x0:x1:x2]([ζ0x0d:ζ1x1d:ζ2x2d],[x0:x1:x2]).[x_{0}:x_{1}:x_{2}]\mapsto\big{(}[\zeta_{0}x_{0}^{d}:\zeta_{1}x_{1}^{d}:\zeta_{2}x_{2}^{d}],[x_{0}:x_{1}:x_{2}]\big{)}.

Thus, C~d\widetilde{C}_{d} and ζC~d\zeta\widetilde{C}_{d} only intersect when ζϕ~d([x0:x1:x2])=ϕ~d([x0:x1:x2])\zeta\widetilde{\phi}_{d}([x_{0}:x_{1}:x_{2}])=\widetilde{\phi}_{d}([x_{0}:x_{1}:x_{2}]). At most one of the xix_{i} can vanish since x0+x1+x2=0x_{0}+x_{1}+x_{2}=0, so there are no intersections when ζiζj\zeta_{i}\neq\zeta_{j} for iji\neq j, and a single intersection with multiplicity dd at V(xk)V(x_{k}) when ζi=ζj\zeta_{i}=\zeta_{j} and {i,j,k}={0,1,2}\{i,j,k\}=\{0,1,2\}.

3. Relation with Fermat varieties and Frobenius morphisms

By Lemma 1, the second projection pr2:RmV(y0+y1+y2)\operatorname{pr}_{2}\colon R_{m}\to V(y_{0}+y_{1}+y_{2}) realises RmR_{m} as the family of diagonal degree mm curves over C1𝐏1C_{1}\cong\mathbf{P}^{1} given by

x0my0+x1my1+x2my2=0.x_{0}^{m}y_{0}+x_{1}^{m}y_{1}+x_{2}^{m}y_{2}=0.

If chark=p>0\operatorname{char}k=p>0 and mm is invertible in kk, then the curves C~dRm\widetilde{C}_{d}\subseteq R_{m} for dm=pe1dm=p^{e}-1 are given by sections ϕ~d:C1Rm\widetilde{\phi}_{d}\colon C_{1}\to R_{m} of pr2\operatorname{pr}_{2}. In this section, we pull back the family RmC1R_{m}\to C_{1} and the sections ϕ~d\widetilde{\phi}_{d} along finite covers of C1C_{1}. Pulling back along covers by Fermat curves allows us to relate the C~d\widetilde{C}_{d} in Corollary 3 with graphs of Frobenius on products of Fermat curves. Pulling back along the Frobenius morphism of C1C_{1} allows us to realise the C~d\widetilde{C}_{d} in Corollary 3 as pullbacks of a constant section C~0\widetilde{C}_{0} under powers of a horizontal Frobenius morphism of RmR_{m} over C1C_{1}.

3.1. Intermediate Surfaces. For positive integers mm and nn invertible in kk and r𝐍r\in\mathbf{N}, denote by Rm,n,rR_{m,n,r} the normal surface

Rm,n,r={([x0:x1:x2],[y0:y1:y2])𝐏2×𝐏2|y0n+y1n+y2n=0x0my0r+x1my1r+x2my2r=0}.R_{m,n,r}=\Set{\big{(}[x_{0}:x_{1}:x_{2}],[y_{0}:y_{1}:y_{2}]\big{)}\in\mathbf{P}^{2}\times\mathbf{P}^{2}}{\begin{array}[]{c}y_{0}^{n}+y_{1}^{n}+y_{2}^{n}=0\\ x_{0}^{m}y_{0}^{r}+x_{1}^{m}y_{1}^{r}+x_{2}^{m}y_{2}^{r}=0\end{array}}.

It is smooth if and only if m=1m=1 or r{0,1}r\in\{0,1\}; in all other cases, the singular locus V(x0y0,x1y1,x2y2)V(x_{0}y_{0},x_{1}y_{1},x_{2}y_{2}) consists of the 3n3n points

{([1:0:0],[0:s:t]),([0:1:0],[s:0:t]),([0:0:1],[s:t:0])|sn+tn=0}.\left\{\big{(}[1:0:0],[0:s:t]\big{)},\big{(}[0:1:0],[s:0:t]\big{)},\big{(}[0:0:1],[s:t:0]\big{)}\ \Big{|}\ s^{n}+t^{n}=0\right\}.

Note that Rm,1,1R_{m,1,1} is none other than the surface RmR_{m} of Lemma 1. If XnX_{n} denotes the Fermat curve V(y0n+y1n+y2n)𝐏2V(y_{0}^{n}+y_{1}^{n}+y_{2}^{n})\subseteq\mathbf{P}^{2} of degree nn, then Rm,n,0R_{m,n,0} coincides with Xm×XnX_{m}\times X_{n}. The surfaces Rm,n,rR_{m,n,r} for r>0r>0 come with a projection

pr2:Rm,n,rXn\operatorname{pr}_{2}\colon R_{m,n,r}\to X_{n}

that is smooth away from the 3n3n fibres above V(y0y1y2)XnV(y_{0}y_{1}y_{2})\subseteq X_{n}, and whose singular fibres consist of mm lines meeting at a point.

3.2. Generalized Power Maps. For positive integers aa and bb invertible in kk, define the finite morphism

πa,b:Ram,bn,br\displaystyle\pi_{a,b}\colon R_{am,bn,br} Rm,n,r\displaystyle\to R_{m,n,r}
([x0:x1:x2],[y0:y1:y2])\displaystyle\big{(}[x_{0}:x_{1}:x_{2}],[y_{0}:y_{1}:y_{2}]\big{)} ([x0a:x1a:x2a],[y0b:y1b:y2b]).\displaystyle\mapsto\big{(}[x_{0}^{a}:x_{1}^{a}:x_{2}^{a}],[y_{0}^{b}:y_{1}^{b}:y_{2}^{b}]\big{)}.

For b=1b=1 and n=r=1n=r=1, it coincides with the morphism π~a\widetilde{\pi}_{a} from 2. When a=1a=1, these fit into pullback squares

Rm,bn,br{R_{m,bn,br}}Rm,n,r{R_{m,n,r}}Xbn{X_{bn}}Xn .{X_{n}\makebox[0.0pt][l]{\,.}}π1,b\scriptstyle{\pi_{1,b}}pr2\scriptstyle{\operatorname{pr}_{2}}pr2\scriptstyle{\operatorname{pr}_{2}}

If Fe:XnXnF^{e}\colon X_{n}\to X_{n} is the pep^{e}-th power Frobenius morphism of XnX_{n}, there are pullback squares

Rm,n,per{R_{m,n,p^{e}r}}Rm,n,r{R_{m,n,r}}Xn{X_{n}}Xn ,{X_{n}\makebox[0.0pt][l]{\,,}}pr2\scriptstyle{\operatorname{pr}_{2}}pr2\scriptstyle{\operatorname{pr}_{2}}Fe\scriptstyle{F^{e}}

so Rm,n,perR_{m,n,p^{e}r} is the Frobenius twist Rm,n,r(e)R_{m,n,r}^{(e)} of Rm,n,rR_{m,n,r} over XnX_{n}. We denote the top map by π(e)\pi^{(e)}.

3.3. Lemma. — Let mm and nn be positive integers invertible in kk, let a,r𝐙a,r\in\mathbf{Z}, and assume that rr and r+amr+am are nonnegative. Then the map

ψa:𝐏2×𝐏2\displaystyle\psi_{a}\colon\mathbf{P}^{2}\times\mathbf{P}^{2} 𝐏2×𝐏2\displaystyle\stackrel{{\scriptstyle\sim}}{{\dashrightarrow}}\mathbf{P}^{2}\times\mathbf{P}^{2}
([x0:x1:x2],[y0:y1:y2])\displaystyle\big{(}[x_{0}:x_{1}:x_{2}],[y_{0}:y_{1}:y_{2}]\big{)} ([x0y0a:x1y1a:x2y2a],[y0:y1:y2])\displaystyle\longmapsto\big{(}[x_{0}y_{0}^{a}:x_{1}y_{1}^{a}:x_{2}y_{2}^{a}],[y_{0}:y_{1}:y_{2}]\big{)}

maps Rm,n,r+amR_{m,n,r+am} birationally onto Rm,n,rR_{m,n,r}.

Proof. Note that ψa\psi_{a} is a birational map with rational inverse ψa\psi_{-a}. The result follows since ψa\psi_{a} takes Rm,n,r+amR_{m,n,r+am} into Rm,n,rR_{m,n,r} and ψa\psi_{-a} does the opposite, and neither surface is contained in the locus where ψa\psi_{a} or ψa\psi_{-a} is undefined.   \blacksquare\blacksquare

This allows us to relate Rm,m,mR_{m,m,m} and Xm×XmX_{m}\times X_{m}:

3.4. Corollary. — The surfaces Xm×XmRm,m,0X_{m}\times X_{m}\cong R_{m,m,0} and Rm,m,mR_{m,m,m} are birational via

ψ:Xm×Xm\displaystyle\psi\colon X_{m}\times X_{m} Rm,m,m\displaystyle\stackrel{{\scriptstyle\sim}}{{\dashrightarrow}}R_{m,m,m}
([x0:x1:x2],[y0:y1:y2])\displaystyle\big{(}[x_{0}:x_{1}:x_{2}],[y_{0}:y_{1}:y_{2}]\big{)} ([x0y0:x1y1:x2y2],[y0:y1:y2]).\displaystyle\longmapsto\Big{(}\big{[}\tfrac{x_{0}}{y_{0}}:\tfrac{x_{1}}{y_{1}}:\tfrac{x_{2}}{y_{2}}\big{]},[y_{0}:y_{1}:y_{2}]\Big{)}.

The composition ρ:Xm×XmRm\rho\colon X_{m}\times X_{m}\dashrightarrow R_{m} of ψ\psi with π1,m\pi_{1,m} is given by

([x0:x1:x2],[y0:y1:y2])([x0y0:x1y1:x2y2],[y0m:y1m:y2m]).\big{(}[x_{0}:x_{1}:x_{2}],[y_{0}:y_{1}:y_{2}]\big{)}\longmapsto\Big{(}\big{[}\tfrac{x_{0}}{y_{0}}:\tfrac{x_{1}}{y_{1}}:\tfrac{x_{2}}{y_{2}}\big{]},[y_{0}^{m}:y_{1}^{m}:y_{2}^{m}]\Big{)}.

If chark=p>0\operatorname{char}k=p>0, mm is invertible in kk, and dm=pe1dm=p^{e}-1 for some positive integer ee, then the strict transform of π1,mC~d\pi_{1,m}^{*}\widetilde{C}_{d} under ψ\psi is the transpose ΓFe\Gamma_{F^{e}}^{\top} of the graph of the pep^{e}-power Frobenius.

Proof. The first statement follows by applying Lemma 3 to m=n=rm=n=r and a=1a=-1, and the second is immediate from the definitions. For the final statement, recall that ΓFe\Gamma_{F^{e}}^{\top} is given by the section s:XmXm×Xms\colon X_{m}\to X_{m}\times X_{m} of pr2\operatorname{pr}_{2} given by

[y0:y1:y2]([y0pe:y1pe:y2pe],[y0:y1:y2]).[y_{0}:y_{1}:y_{2}]\mapsto\Big{(}\big{[}y_{0}^{p^{e}}:y_{1}^{p^{e}}:y_{2}^{p^{e}}\big{]},[y_{0}:y_{1}:y_{2}]\Big{)}.

By the first pullback square of 3, the curve π1,mC~d\pi_{1,m}^{*}\widetilde{C}_{d} is the image of the section XmRm,m,mX_{m}\to R_{m,m,m} given by

[y0:y1:y2]([y0dm:y1dm:y2dm],[y0:y1:y2]),[y_{0}:y_{1}:y_{2}]\mapsto\big{(}[y_{0}^{dm}:y_{1}^{dm}:y_{2}^{dm}],[y_{0}:y_{1}:y_{2}]\big{)},

which agrees with ψs\psi\circ s.   \blacksquare\blacksquare

3.5. The curves ΓFeXm×Xm\Gamma_{F^{e}}\subseteq X_{m}\times X_{m} are the standard example of curves with unbounded negative self-intersection: the condition m>3m>3 of the Main Theorem is exactly the condition g(Xm)>1g(X_{m})>1 that makes ΓFe2=pe(22g)\Gamma_{F^{e}}^{2}=p^{e}(2-2g) negative. In fact, since ΓFe\Gamma_{F^{e}}^{\top} passes through 3m3m of the 3m23m^{2} points of indeterminacy of ψ\psi, resolving the map shows that m2C~d2=ΓFe23mm^{2}\,\widetilde{C}_{d}^{2}=\Gamma_{F^{e}}^{2}-3m.

On the other hand, RmC1R_{m}\to C_{1} is an isotrivial family of diagonal degree mm curves that becomes rationally trivialised over the mm-th power cover XmC1X_{m}\to C_{1}. Thus, we can also look directly at the pullback π(e):Rm(e)Rm\pi^{(e)}\colon R_{m}^{(e)}\to R_{m} of the Frobenius Fe:C1C1F^{e}\colon C_{1}\to C_{1}. Note that Rm(e)=Rm,1,peR_{m}^{(e)}=R_{m,1,p^{e}} by 3, so we get:

3.6. Corollary. — If pe=dm+1p^{e}=dm+1, then Rm(e)R_{m}^{(e)} is birational to RmR_{m} via

ψ:Rm\displaystyle\psi\colon R_{m} Rm(e)\displaystyle\stackrel{{\scriptstyle\sim}}{{\dashrightarrow}}R_{m}^{(e)}
([x0:x1:x2],[y0:y1:y2])\displaystyle\big{(}[x_{0}:x_{1}:x_{2}],[y_{0}:y_{1}:y_{2}]\big{)} ([x0y0d:x1y1d:x2y2d],[y0:y1:y2]).\displaystyle\longmapsto\Big{(}\big{[}\tfrac{x_{0}}{y_{0}^{d}}:\tfrac{x_{1}}{y_{1}^{d}}:\tfrac{x_{2}}{y_{2}^{d}}\big{]},[y_{0}:y_{1}:y_{2}]\Big{)}.

If ϕ~0:C1Rm\widetilde{\phi}_{0}\colon C_{1}\to R_{m} denotes the constant section [y0:y1:y2]([1:1:1],[y0:y1:y2])[y_{0}:y_{1}:y_{2}]\mapsto\big{(}[1:1:1],[y_{0}:y_{1}:y_{2}]\big{)} and C~0Rm\widetilde{C}_{0}\subseteq R_{m} denotes its image, then C~d\widetilde{C}_{d} is the strict transform of π(e),C~0\pi^{(e),*}\widetilde{C}_{0} under ψ\psi.

Proof. The first statement follows from Lemma 3 applied to n=1n=1, r=per=p^{e}, and a=da=-d. For the second, by the second pullback square of 3, the curve π(e),C~0\pi^{(e),*}\widetilde{C}_{0} is the image of the constant section C1Rm(e)C_{1}\to R_{m}^{(e)} given by

[y0:y1:y2]([1:1:1],[y0:y1:y2]),[y_{0}:y_{1}:y_{2}]\mapsto\big{(}[1:1:1],[y_{0}:y_{1}:y_{2}]\big{)},

which agrees with ψϕ~d\psi\circ\widetilde{\phi}_{d}.   \blacksquare\blacksquare

3.7. Instead of the transpose ΓFe\Gamma_{F^{e}}^{\top} of the graph of Fe:XmXmF^{e}\colon X_{m}\to X_{m}, one can also look at the negative curves ΓFeXm×Xm\Gamma_{F^{e}}\subseteq X_{m}\times X_{m}, which are given by pulling back the diagonal along the relative Frobenius of pr2:Xm×XmXm\operatorname{pr}_{2}\colon X_{m}\times X_{m}\to X_{m}. Their images under the rational map ρ\rho of Corollary 3 are given by the parametrised rational curves

C1\displaystyle C_{1} Rm\displaystyle\to R_{m}
[y0:y1:y2]\displaystyle[y_{0}:y_{1}:y_{2}] ([y1dy2d:y0dy2d:y0dy1d],[y0pe:y1pe:y2pe]),\displaystyle\mapsto\Big{(}\big{[}y_{1}^{d}y_{2}^{d}:y_{0}^{d}y_{2}^{d}:y_{0}^{d}y_{1}^{d}\big{]},\big{[}y_{0}^{p^{e}}:y_{1}^{p^{e}}:y_{2}^{p^{e}}\big{]}\Big{)},

where dm=pe1dm=p^{e}-1 as usual. These are obtained from the curve C~0\widetilde{C}_{0} of Corollary 3 by pulling back the strict transform of C~0\widetilde{C}_{0} under ψ1:Rm(e)Rm\psi^{-1}\colon R_{m}^{(e)}\dashrightarrow R_{m} along the relative Frobenius FRm/C1eF_{R_{m}/C_{1}}^{e}. Note that the images of these curves in 𝐏2\mathbf{P}^{2} differ from the curves CdC_{d} by the Cremona transformation

[x0:x1:x2][x01:x11:x21].[x_{0}:x_{1}:x_{2}]\mapsto[x_{0}^{-1}:x_{1}^{-1}:x_{2}^{-1}].

Finally, we relate the multiplicity of CdC_{d} at [1:1:1][1:1:1] to point counts on the Fermat curve XmX_{m} if dm=pe1dm=p^{e}-1 for some positive integer ee.

3.8. Lemma. — If dm=pe1dm=p^{e}-1, then a point xC1x\in C_{1} maps to [1:1:1][1:1:1] in CdC_{d} if and only if there exists yXm(𝐅pe)y\in X_{m}(\mathbf{F}_{p^{e}}) with nonzero coordinates mapping to xx under the mm-th power map XmC1X_{m}\to C_{1}. In particular,

mult[1:1:1]Cd=|Xm(𝐅pe)|3mm2.\operatorname{mult}_{[1:1:1]}C_{d}=\frac{|X_{m}(\mathbf{F}_{p^{e}})|-3m}{m^{2}}.

Proof. The first statement follows since XmC1X_{m}\to C_{1} is surjective and a point y=[y0:y1:y2]y=[y_{0}:y_{1}:y_{2}] on XmX_{m} with nonzero coordinates maps to [1:1:1][1:1:1] under the (pe1)(p^{e}-1)-st power map XmCdX_{m}\to C_{d} if and only if yXm(𝐅pe)y\in X_{m}(\mathbf{F}_{p^{e}}).

For the second statement, note that mult[1:1:1]Cd\operatorname{mult}_{[1:1:1]}C_{d} equals the number of preimages of [1:1:1][1:1:1] in C1C_{1}, since ϕ~d:C1C~d\widetilde{\phi}_{d}\colon C_{1}\to\widetilde{C}_{d} is an isomorphism by Corollary 1. The result now follows since XmC1X_{m}\to C_{1} is finite étale of degree m2m^{2} away from the coordinate axes, so each point xC1V(x0x1x2)x\in C_{1}\setminus V(x_{0}x_{1}x_{2}) has exactly m2m^{2} preimages in XmX_{m}.   \blacksquare\blacksquare

For example, if pν1(modm)p^{\nu}\equiv-1\pmod{m} for some positive integer ν\nu, then

|Xm(𝐅pe)|=1(m1)(m2)2pe/2+pe|X_{m}(\mathbf{F}_{p^{e}})|=1-\frac{(m-1)(m-2)}{2}p^{e/2}+p^{e}

whenever pe1(modm)p^{e}\equiv 1\pmod{m} [SK79, Lem. 3.3].

4. Remarks towards characteristic 0

4.1. Although the Bounded Negativity Conjecture is currently still open in characteristic 0, the Weak Bounded Negativity Conjecture is known [Hao19]: for any smooth projective complex surface XX and any g𝐍g\in\mathbf{N}, there exists a constant b(X,g)b(X,g) such that C2b(X,g)C^{2}\geq-b(X,g) for every reduced curve C=iCiC=\sum_{i}C_{i} whose components CiC_{i} have geometric genus at most gg.

Our examples in the Main Theorem certainly violate this, and, as we now verify, arise from the failure of the logarithmic Bogomolov–Miyaoka–Yau inequality for the pair (Rm,C~d)(R_{m},\widetilde{C}_{d}) when dd is large with respect to mm. In the next three paragraphs, assume chark=p>0\operatorname{char}k=p>0 and dm=pe1dm=p^{e}-1. To ease notation, write (R,C~)(R,\widetilde{C}) for (Rm,C~d)(R_{m},\widetilde{C}_{d}). We will use logarithmic sheaves of differentials; see for example [EV92, §2].

4.2. Lemma. — The Chern numbers of the pair (R,C~)(R,\widetilde{C}) are

c12(R,C~)\displaystyle c_{1}^{2}(R,\widetilde{C}) c12(ΩR1(logC~))=d(m3)m2+6,\displaystyle\coloneqq c_{1}^{2}\big{(}\Omega^{1}_{R}(\log\widetilde{C})\big{)}=d(m-3)-m^{2}+6,
c2(R,C~)\displaystyle c_{2}(R,\widetilde{C}) c2(ΩR1(logC~))=m2+1.\displaystyle\coloneqq c_{2}\big{(}\Omega^{1}_{R}(\log\widetilde{C})\big{)}=m^{2}+1.

In particular, the Chern slopes c12(R,C~)/c2(R,C~)c_{1}^{2}(R,\widetilde{C})/c_{2}(R,\widetilde{C}) are unbounded for fixed mm and growing dd.

Proof. The logarithmic sheaf of differentials fit into a short exact sequence

0ΩR1ΩR1(logC~)𝒪C~0,0\to\Omega^{1}_{R}\to\Omega^{1}_{R}(\log\widetilde{C})\to\mathcal{O}_{\widetilde{C}}\to 0,

so c12(R,C~)=(KR+C~)2c_{1}^{2}(R,\widetilde{C})=(K_{R}+\widetilde{C})^{2} and c2(R,C~)=c2(ΩR1)+C~(KR+C~)c_{2}(R,\widetilde{C})=c_{2}(\Omega^{1}_{R})+\widetilde{C}(K_{R}+\widetilde{C}). Since RR is the blowup of 𝐏2\mathbf{P}^{2} in m2m^{2} points, we get KR2=9m2K_{R}^{2}=9-m^{2} and c2(ΩR1)=3+m2c_{2}(\Omega^{1}_{R})=3+m^{2}, so the result follows from the computations of the intersection numbers in Corollary 1.   \blacksquare\blacksquare

4.3. Lemma. — If m>3m>3 and dd is such that

χ(2(KR+C~))=d(m3)m2+5>0\chi\big{(}2(K_{R}+\widetilde{C})\big{)}=d(m-3)-m^{2}+5>0

then H0(R,2(KR+C~))0\mathrm{H}^{0}(R,2(K_{R}+\widetilde{C}))\neq 0. In particular, KR+C~K_{R}+\widetilde{C} is pseudoeffective.

Proof. The Euler characteristic statement follows from Riemann–Roch, so it remains to show that H0(R,2(KR+C~))0\mathrm{H}^{0}\big{(}R,2(K_{R}+\widetilde{C})\big{)}\neq 0 once χ(2(KR+C~))>0\chi\big{(}2(K_{R}+\widetilde{C})\big{)}>0. But H2(R,2(KR+C~))=H0(R,KR2C~)\mathrm{H}^{2}\big{(}R,2(K_{R}+\widetilde{C})\big{)}=\mathrm{H}^{0}(R,-K_{R}-2\widetilde{C})^{\vee}, and the latter vanishes since C~\widetilde{C} is effective and KR=𝒪R(3m,1)-K_{R}=\mathcal{O}_{R}(3-m,1) by Lemma 1.   \blacksquare\blacksquare

For dd large with respect to mm, this shows that (R,C~)(R,\widetilde{C}) falls into the final case considered in [Hao19, §1.2, Case 2], and that the failure of Weak Bounded Negativity stems from the failure of the logarithmic Bogomolov–Miyaoka–Yau inequality:

4.4. Corollary. — If m>3m>3 and d>5m22m3d>\frac{5m^{2}-2}{m-3}, then KR+C~K_{R}+\widetilde{C} is pseudoeffective and

c12(R,C~)/c2(R,C~)>4.c_{1}^{2}(R,\widetilde{C})/c_{2}(R,\widetilde{C})>4.

Moreover, the pair (R,C~)(R,\widetilde{C}) does not lift to the second Witt vectors W2(k)W_{2}(k).

Proof. The first part follows from Lemma 4 and Lemma 4. The final statement follows from [Lan16, Proposition 4.3], since (R,C~)(R,\widetilde{C}) violates the logarithmic Bogomolov–Miyaoka–Yau inequality.   \blacksquare\blacksquare

4.5. On the other hand, the surface RmR_{m} itself does lift to characteristic 0. This gives new examples of surfaces XSpec𝐙X\to\operatorname{Spec}\mathbf{Z} such that almost all special fibres X𝐅¯pX_{\bar{\mathbf{F}}_{p}} (namely those with p\nmidmp\nmid m) violate bounded negativity. The same property holds for the square C×CC\times C of a curve CSpec𝐙C\to\operatorname{Spec}\mathbf{Z} of genus 2\geq 2, which is the classical counterexample to bounded negativity in positive characteristic. However, the rational surface X=RmX=R_{m} has the additional property that the specialisation maps NS(X𝐐¯)NS(X𝐅¯p)\operatorname{NS}(X_{\bar{\mathbf{Q}}})\to\operatorname{NS}(X_{\bar{\mathbf{F}}_{p}}) are isomorphisms for every prime p\nmidmp\nmid m.

4.6. Question. — Is it possible to determine the effective cone of RmR_{m} for some m4m\geq 4? How does it depend on the characteristic of kk?

For example, the curves in 𝐏2\mathbf{P}^{2} cut out by the polynomials gm1g_{m-1} of Lemma 2 are smooth of genus (m2)(m3)2\tfrac{(m-2)(m-3)}{2} in characteristic 0 [RVVZ01, Thm. 1], and the equation gm1h2=hm+1g_{m-1}h_{2}=h_{m+1} shows that V(gm1)V(x0x1)V(x1x2)V(x2x0)V(g_{m-1})\cup V(x_{0}-x_{1})\cup V(x_{1}-x_{2})\cup V(x_{2}-x_{0}) contains

ZmV(x0x1m+1x0m+1x1,x1x2m+1x1m+1x2,x2x0m+1x2m+1x0)=Zm{[s:t:0],[s:0:t],[0:s:t]|sm=tm}.Z_{m}^{\prime}\coloneqq V\left(\begin{array}[]{c}x_{0}x_{1}^{m+1}-x_{0}^{m+1}x_{1},\\ x_{1}x_{2}^{m+1}-x_{1}^{m+1}x_{2},\\ x_{2}x_{0}^{m+1}-x_{2}^{m+1}x_{0}\end{array}\right)=Z_{m}\cup\left\{[s:t:0],[s:0:t],[0:s:t]\ \big{|}\ s^{m}=t^{m}\right\}.

Since V(gm1)V(g_{m-1}) has self-intersection (m1)2(m-1)^{2} and passes through the m23m+2m^{2}-3m+2 points of ZmZ_{m} whose coordinates are pairwise distinct, its strict transform on RmR_{m} has self-intersection m1m-1. On the further blowup RmR^{\prime}_{m} of 𝐏2\mathbf{P}^{2} in ZmZ^{\prime}_{m}, the strict transform has self-intersection 2m+2-2m+2, but unlike the situation described in 2, there does not appear to be an obvious way to produce infinitely many negative curves on a single rational surface this way.

When m=pem=p^{e} for some prime pp, the specialisation to characteristic pp collapses ZmZ_{m} onto the point [1:1:1][1:1:1], and the smooth curve V(gm1)V(g_{m-1}) becomes a rational curve that is highly singular at [1:1:1][1:1:1]. Even though these curves are not negative yet (see 2), taking different values of ee does give infinitely many curves on the same rational surface.

4.7. As far as we are aware, all known counterexamples to bounded negativity on a smooth projective surface XX over an algebraically closed field kk of characteristic p>0p>0 consist of a family CiC_{i} of curves on XX for which there exist constants a,ba,b such that Ci2=api+bC_{i}^{2}=ap^{i}+b for all i𝐍i\in\mathbf{N}.

4.8. Question. — If XX is a surface over an algebraically closed field kk of characteristic p>0p>0, is there a finite set {(ai,bi)𝐐2|iI}\{(a_{i},b_{i})\in\mathbf{Q}^{2}\ |\ i\in I\} such that all integral curves CXC\subseteq X with C2<0C^{2}<0 satisfy

C2=aipe+biC^{2}=a_{i}p^{e}+b_{i}

for some positive integer ee and some iIi\in I? If not, is there some other way in which the self-intersections of negative curves on XX are “not too scattered”?

We can also consider the following uniform version:

4.9. Question. — If XSX\to S is a smooth projective surface over a finitely generated integral base scheme SS, does there exist a finite set {(ai,bi)𝐐2|iI}\{(a_{i},b_{i})\in\mathbf{Q}^{2}\ |\ i\in I\} such that every geometrically integral curve CXsC\subseteq X_{s} of negative self-intersection in a fibre XsX_{s} with charκ(s)>0\operatorname{char}\kappa(s)>0 satisfies

C2=aipe+biC^{2}=a_{i}p^{e}+b_{i}

for some positive integer ee and some iIi\in I, where p=charκ(s)p=\operatorname{char}\kappa(s)?

For example, for the surfaces RmSpec𝐙[1/m]R_{m}\to\operatorname{Spec}\mathbf{Z}[1/m] and the curves C~d\widetilde{C}_{d} of the Main Theorem, we may take a=3mma=\tfrac{3-m}{m} and b=3mb=\tfrac{-3}{m}, which do not depend on the characteristic of κ(s)\kappa(s).

4.10. Despite the failure of bounded negativity in positive characteristic, a positive answer to Question 4 still implies bounded negativity in characteristic 0 via reduction modulo primes. Indeed, the minimum bmin=min{bi|iI}b_{\text{min}}=\min\{b_{i}\ |\ i\in I\} is a lower bound for the self-intersection C2C^{2} of a geometrically integral curve CC in the generic fibre, since the specialisations CsC_{s} of CC satisfy Cs2=C2C_{s}^{2}=C^{2} for all sSs\in S and remain geometrically integral for ss in a dense open set USU\subseteq S, and

pP{aipe+bi|iI,e𝐙>0}[bmin,)\bigcap_{p\in P}\big{\{}a_{i}p^{e}+b_{i}\ \big{|}\ i\in I,e\in\mathbf{Z}_{>0}\big{\}}\subseteq[b_{\text{min}},\infty)

for any infinite set of primes PP. Thus, Question 4 is a natural analogue of the Bounded Negativity Conjecture in positive characteristic.

Acknowledgements

We thank Johan de Jong, Joaquín Moraga, Takumi Murayama, Will Sawin, and John Sheridan for helpful discussions. RvDdB was partly supported by the Oswald Veblen Fund at the Institute for Advanced Study.

References

  • [BBC+12] T. Bauer, C. Bocci, S. Cooper, S. Di Rocco, M. Dumnicki, B. Harbourne, K. Jabbusch, A. L. Knutsen, A. Küronya, R. Miranda, J. Roé, H. Schenck, T. Szemberg, and Z. Teitler, Recent developments and open problems in linear series. Contributions to algebraic geometry, EMS Ser. Congr. Rep., p. 93–140. Eur. Math. Soc., Zürich, 2012. doi:10.4171/114-1/4.
  • [BDRH+15] T. Bauer, S. Di Rocco, B. Harbourne, J. Huizenga, A. Lundman, P. Pokora, and T. Szemberg, Bounded negativity and arrangements of lines. Int. Math. Res. Not. IMRN 2015.19, p. 9456–9471 (2015). doi:10.1093/imrn/rnu236.
  • [BHK+13] T. Bauer, B. Harbourne, A. L. Knutsen, A. Küronya, S. Müller-Stach, X. Roulleau, and T. Szemberg, Negative curves on algebraic surfaces. Duke Math. J. 162.10, p. 1877–1894 (2013). doi:10.1215/00127094-2335368.
  • [EV92] H. Esnault and E. Viehweg, Lectures on vanishing theorems. DMV Seminar 20. Birkhäuser Verlag, Basel, 1992. doi:10.1007/978-3-0348-8600-0.
  • [Hao19] F. Hao, Weak bounded negativity conjecture. Proc. Amer. Math. Soc. 147.8, p. 3233–3238 (2019). doi:10.1090/proc/14376.
  • [Har10] B. Harbourne, Global aspects of the geometry of surfaces. Ann. Univ. Paedagog. Crac. Stud. Math. 9, p. 5–41 (2010).
  • [Har18] B. Harbourne, Asymptotics of linear systems, with connections to line arrangements. Phenomenological approach to algebraic geometry. Banach Center Publ. 116, p. 87–135. Polish Acad. Sci. Inst. Math., Warsaw, 2018.
  • [Lan16] A. Langer, The Bogomolov-Miyaoka-Yau inequality for logarithmic surfaces in positive characteristic. Duke Math. J. 165.14, p. 2737–2769 (2016). doi:10.1215/00127094-3627203.
  • [RVVZ01] F. Rodríguez Villegas, J. F. Voloch, and D. Zagier, Constructions of plane curves with many points. Acta Arith. 99.1, p. 85–96 (2001). doi:10.4064/aa99-1-8.
  • [SK79] T. Shioda and T. Katsura, On Fermat varieties. Tōhoku Math. J. (2) 31.1, p. 97–115 (1979). doi:10.2748/tmj/1178229881.