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

Families of embeddings of the alternating group of rank 55 into the Cremona group

I. Krylov Igor Krylov
Korea Institute for Advanced Study, 85 Hoegiro, Dongdaemun-gu, Seoul 02455, Republic of Korea
[email protected]
Abstract.

I study embeddings of the alternating group of rank five into the Cremona group of rank three. I find all embeddings induced by 𝒜5\mathcal{A}_{5}\mathds{Q}-del Pezzo fibrations and I study their conjugacy. As an application, I show that there is a series of continuous families of pairwise non-conjugate embeddings of alternating group of rank five into Cr3()\operatorname{Cr}_{3}(\mathds{C}).

1. Introduction

The Cremona group Crn\operatorname{Cr}_{n} of rank nn is a group of birational transformations of n\mathds{P}^{n}. It is natural to study the group by studying its subgroups and finite subgroups in particular. The classification of finite subgroups in Cr2\operatorname{Cr}_{2} up to conjugacy is almost complete [7]. It is not feasible to achieve a classification for Cr3\operatorname{Cr}_{3}, nevertheless we can say something about its finite subgroups, for example we know that Crn\operatorname{Cr}_{n} is Jordan [16].

If we limit the group types, for example to pp-subgroups ([17]) or simple non-abelian, then the problem becomes more manageable. The motivating problem for me is the following.

Problem 1.1.

Classify the embeddings of finite simple non-abelian subgroups of Cr3\operatorname{Cr}_{3} up to conjugacy.

The isomorphism types of simple non-abelian groups were classified in [13, Theorem 1.3], the possibilities are: 𝒜5\mathcal{A}_{5}, 𝒜6\mathcal{A}_{6}, 𝒜7\mathcal{A}_{7}, PSL2(7)\operatorname{PSL}_{2}(7), SL2(8)\operatorname{SL}_{2}(8), and PSp4(3)\operatorname{PSp}_{4}(3). In this paper I study the families of subgroups of Cr3\operatorname{Cr}_{3}.

Question 1.2.

Let GG be a group. Is there a continuous family of embeddings GCrnG\hookrightarrow\operatorname{Cr}_{n} which are not pairwise conjugate to each other?

In general, one expects that the asnwer is positive for small groups and negative for big groups. For example there are huge families of pairwise non-conjugate embeddings of /2\mathds{Z}/2\mathds{Z} induced by Bertini and Geiser involutions [15]. Also it was recently shown in [1] that there is a continuous family of pairwise non-conjugate embeddings of 𝒮4\mathcal{S}_{4} into Cr3\operatorname{Cr}_{3}.

On the other hand, finite non-abelian groups are big, thus one expects that there should be no continuous families of pairwise non-conjugate embeddings of these groups. Indeed, there are only six simple non-abelian subgroups of Cr2\operatorname{Cr}_{2} up to conjugacy: three subgroups isomorphic to 𝒜5\mathcal{A}_{5}, two subgroups isomorphic to PSL2(7)\operatorname{PSL}_{2}(7), and one subgroup isomorphic to 𝒜6\mathcal{A}_{6}. In dimension three we know by [13, Theorem 1.5] that there are only finitely many embeddings for the three largest finite simple non-abelian subgroups. I study the embeddings of the smallest finite simple non-abelian subgroup: 𝒜5\mathcal{A}_{5}.

Theorem 1.3.

For each k2k\geqslant 2 there is a 2k32k-3-dimensional family of embeddings 𝒜5Cr3\mathcal{A}_{5}\hookrightarrow\operatorname{Cr}_{3} which are not pairwise conjugate to each other.

The remaining cases are 𝒜6\mathcal{A}_{6} and PSL2(7)\operatorname{PSL}_{2}(7). Some progress toward classification the embeddings of these groups has been made in [3] and [11], so far the results agree with the following expectation.

Conjecture 1.4.

The embeddings into Cr3\operatorname{Cr}_{3} of 𝒜6\mathcal{A}_{6} and PSL2(7)\operatorname{PSL}_{2}(7) up to conjugation form discrete families.

1.1. GG\mathds{Q}-Mori fiber spaces

The study of rational group actions on 3\mathds{P}^{3} may be replaced with the study of regular group actions on suitable rational varieties.

Definition 1.5.

I say that π:YZ\pi\colon Y\to Z (or simply Y/ZY/Z if π\pi is understood) is a GG\mathds{Q}-Mori fiber space if

  • dimZ<dimY\dim Z<\dim Y and π\pi has connected fibers,

  • GG-action on YY is faithful and ZZ admits a GG-action such that the map π\pi is GG-equivariant,

  • YY is terminal and GG\mathds{Q}-factorial, that is GG-invariant divisors on YY are \mathds{Q}-factorial,

  • KY-K_{Y} is π\pi-ample and the relative GG-invariant Picard rank is ρG(Y/Z)=1\rho^{G}(Y/Z)=1.

When YY is Gorenstein I omit \mathds{Q} and say that Y/ZY/Z is a GG-Mori fiber space. If dimYdimZ=2\dim Y-\dim Z=2, then I say that Y/ZY/Z is a GG\mathds{Q}-del Pezzo fibration.

By [14, Proposition 1.2] for a rational action of GG on a rationally connected variety XX there is a GG\mathds{Q}-Mori fiber space π:YZ\pi\colon Y\to Z such that XX is GG-equivariantly birational to YY, hence the GG-action on YY induces a conjugate embedding of GG into Bir(X)\operatorname{Bir}(X). The classification of subgroups of Crn\operatorname{Cr}_{n} up to conjugation is equivalent to the classification of the rational GG\mathds{Q}-Mori fiber spaces up to GG-equivariant birational equivalence.

Now I give examples of 𝒜5\mathcal{A}_{5}-del Pezzo fibrations.

Example 1.6.

The projective plane with 𝒜5\mathcal{A}_{5}-action induced by a non-trivial three-dimensional representation W3W_{3} is an 𝒜5\mathcal{A}_{5}-del Pezzo surface. A del Pezzo surface S5S_{5} of degree 55 satisfies 𝒜5AutS5\mathcal{A}_{5}\subset\operatorname{Aut}S_{5} and is 𝒜5\mathcal{A}_{5}-minimal, hence it is an 𝒜5\mathcal{A}_{5}-del Pezzo surface.

Let SS be a GG-del Pezzo surface, then S×1/1S\times\mathds{P}^{1}/\mathds{P}^{1} is a GG-del Pezzo fibration. It follows that 2×1/1\mathds{P}^{2}\times\mathds{P}^{1}/\mathds{P}^{1} and S5×1/1S_{5}\times\mathds{P}^{1}/\mathds{P}^{1} are 𝒜5\mathcal{A}_{5}-del Pezzo fibrations.

Example 1.7.

Let W3W_{3} and W3W_{3}^{\prime} be the irreducible 33-dimensional representations of 𝒜5\mathcal{A}_{5}. I consider 𝒜5\mathcal{A}_{5}-linearized vector bundles defined as

n=(W3×1)𝒪1(n)andn=(W3×1)𝒪1(n).\mathcal{E}_{n}=\big{(}W_{3}\times\mathds{P}^{1}\big{)}\oplus\mathcal{O}_{\mathds{P}^{1}}(-n)\quad\text{and}\quad\mathcal{E}_{n}^{\prime}=\big{(}W_{3}^{\prime}\times\mathds{P}^{1}\big{)}\oplus\mathcal{O}_{\mathds{P}^{1}}(-n).

Let Tn=(n)T_{n}=\mathds{P}(\mathcal{E}_{n}) and Tn=(n)T_{n}^{\prime}=\mathds{P}(\mathcal{E}_{n}^{\prime}) and denote by πT\pi_{T} and πT\pi_{T^{\prime}} the corresponding projections onto 1\mathds{P}^{1}.

I introduce coordinates on TnT_{n} (resp. TnT_{n}^{\prime}) as follows. Let u,vu,v be the coordinates on the base 1\mathds{P}^{1}, x,y,zx,y,z be the coordinates on W3W_{3} (resp. W3W_{3}^{\prime}), and let ww be the coordinate on the fiber of 𝒪1(n)1\mathcal{O}_{\mathds{P}^{1}}(-n)\to\mathds{P}^{1}. The degrees of the coordinates are given by

(uvxyzw00111111000n)\left(\begin{array}[]{ccccccc}u&v&x&y&z&w\\ 0&0&1&1&1&1\\ 1&1&0&0&0&-n\end{array}\right)

There is a unique 𝒜5\mathcal{A}_{5}-invariant conic Δ\Delta on (W3)\mathds{P}(W_{3}) (resp. (W3)\mathds{P}(W_{3}^{\prime})). Up to a change of coordinates on W3W_{3} (res. W3)W_{3}^{\prime}) I may assume that its equation is xzy2=0xz-y^{2}=0. Let XnX_{n} (resp. XnX_{n}^{\prime}) be the hypersurface in TnT_{n} (resp. TnT_{n}^{\prime}) given by the equation

a2n(u,v)w2=xzy2,a_{2n}(u,v)w^{2}=xz-y^{2},

where dega2n=2n\deg a_{2n}=2n. Observe that this equation is 𝒜5\mathcal{A}_{5}-invariant, thus XnX_{n} and XnX_{n}^{\prime} admit a faithful 𝒜5\mathcal{A}_{5}-action. The 𝒜5\mathcal{A}_{5}-varieties XnX_{n} and XnX_{n}^{\prime} are smooth if and only if a2na_{2n} has no multiple factors. If l(u,v)l(u,v) is a multiple factor of a2na_{2n}, then XnX_{n} and XnX_{n}^{\prime} have a cA1cA_{1}-singularity at x=y=z=l(u,v)=0x=y=z=l(u,v)=0. We have ρ(Xn)=ρ(Xn)=2\rho(X_{n})=\rho(X_{n}^{\prime})=2 if a2na_{2n} is not a square (Lemma 3.4). The restriction π\pi (resp. π\pi^{\prime}) of πT\pi_{T} (resp. πT\pi_{T}^{\prime}) to XnX_{n} (resp. XnX_{n}^{\prime}) induces the structure of 𝒜5\mathcal{A}_{5}-del Pezzo fibration on XnX_{n} (resp. XnX_{n}^{\prime}).

Note that the varieties X1X_{1} and X1X_{1}^{\prime} are unique while for n2n\geqslant 2 the families of varieties XnX_{n} and XnX_{n}^{\prime} are of dimension 2n32n-3.

I show that these are essentially the only 𝒜5\mathcal{A}_{5}\mathds{Q}-del Pezzo fibrations with the trivial action on the base.

Theorem 1.8.

Let π:V1\pi:V\to\mathds{P}^{1} be an 𝒜5\mathcal{A}_{5}\mathds{Q}-del Pezzo fibration and suppose 𝒜5\mathcal{A}_{5} acts trivially on the base. Then one of the following holds:

  1. (1)

    The general fiber of π\pi is 2\mathds{P}^{2} and VV is 𝒜5\mathcal{A}_{5}-equivariantly birational to 2×1\mathds{P}^{2}\times\mathds{P}^{1} with 𝒜5\mathcal{A}_{5} acting only on the first factor;

  2. (2)

    V𝒜5S5×1V\cong_{\mathcal{A}_{5}}S_{5}\times\mathds{P}^{1}, where S5S_{5} is the del Pezzo surface of degree 55 and 𝒜5\mathcal{A}_{5} acts only on the first factor;

  3. (3)

    The general fiber of π\pi is a quadric and VV is 𝒜5\mathcal{A}_{5}-equivariantly birational to a smooth XnX_{n}.

The varieties from the case (3) are inducing families of embeddings of 𝒜5\mathcal{A}_{5} into Cr3\operatorname{Cr}_{3}. In order to show that these embeddings are not pairwise conjugate I use the notion of GG-equivariant birational superrigidity.

Definition 1.9.

Let πX:XS\pi_{X}\colon X\to S and πY:YZ\pi_{Y}\colon Y\to Z be GG\mathds{Q}-Mori fiber spaces. A GG-equivariant birational map χ:XY\chi\colon X\dasharrow Y is called square if it fits into a commutative diagram

X\textstyle{X\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}χ\scriptstyle{\chi}πX\scriptstyle{\pi_{X}}Y\textstyle{\ignorespaces\ignorespaces\ignorespaces\ignorespaces Y}πY\scriptstyle{\pi_{Y}}S\textstyle{S\ignorespaces\ignorespaces\ignorespaces\ignorespaces}g\scriptstyle{g}Z,\textstyle{Z,}

where gg is birational and, in addition, the induced map on the generic fibers χη:XηYη\chi_{\eta}\colon X_{\eta}\dasharrow Y_{\eta} is isomorphism of GG-varieties. In this case I say that X/SX/S and Y/ZY/Z are GG-equivariantly square-birational.

Definition 1.10.

I say that πX:XS\pi_{X}\colon X\to S is GG-equivariantly superrigid if for any GG\mathds{Q}-Mori fiber space πY:YZ\pi_{Y}\colon Y\to Z any GG-equivariant birational map χ:XY\chi\colon X\to Y is GG-equivariantly square birational.

Theorem 1.11.

Suppose XnX_{n} (resp. XnX_{n}^{\prime}) is smooth and n2n\geqslant 2, then XnX_{n} (resp. XnX_{n}^{\prime}) is 𝒜5\mathcal{A}_{5}-equivariantly birationally superrigid.

Corollary 1.12.

Suppose XnX_{n}, XkX_{k}, XmX_{m}^{\prime}, and XlX_{l}^{\prime} are smooth and n,m,k,l2n,m,k,l\geqslant 2, then:

  1. (1)

    The varieties XnX_{n} and XmX_{m}^{\prime} are not 𝒜5\mathcal{A}_{5}-equivariantly birational to 2×1\mathds{P}^{2}\times\mathds{P}^{1} and S5×1S_{5}\times\mathds{P}^{1};

  2. (2)

    The variety XnX_{n} is not 𝒜5\mathcal{A}_{5}-equivariantly birational to XmX_{m}^{\prime};

  3. (3)

    The variety XnX_{n} is 𝒜5\mathcal{A}_{5}-equivariantly birational to XkX_{k} if and only if XnX_{n} is 𝒜5\mathcal{A}_{5}-equivariantly isomorphic to XkX_{k} and the same holds for XmX_{m}^{\prime} and XlX_{l}^{\prime};

  4. (4)

    For n2n\geqslant 2 the family of 𝒜5\mathcal{A}_{5}-del Pezzo fibrations XnX_{n} (or Xm)X_{m}^{\prime}) induces a 2n32n-3-dimensional family (resp. 2m32m-3-dimensional) of pairwise non-conjugate embeddings 𝒜5Cr3\mathcal{A}_{5}\hookrightarrow\operatorname{Cr}_{3}.

The assertion (4) is Theorem 1.3.

Acknowledgments.

The author would like to thank Andrey Trepalin and Constantin Shramov for pointing out how to work with del Pezzo surfaces of degree 55 over non-closed fields. The author was supported by KIAS individual grant MG069801 at Korea Institute for Advanced Study.

Notations and conventions

I work over \mathds{C} unless stated otherwise. All varieties are considered to be projective and normal. I denote the symmetric and alternating groups of rank nn by 𝒮n\mathcal{S}_{n} and 𝒜n\mathcal{A}_{n} respectively. I denote the del Pezzo surface of degree 55 by S5S_{5} and Clebsch cubic by S3S_{3}. I denote the linear equivalence of divisors by \sim, the numerical equivalence of cycles by \equiv, and 𝒜5\mathcal{A}_{5}-equivariant biregular equivalence by 𝒜5\cong_{\mathcal{A}_{5}}. Given a curve Δ\Delta I denote its orbit by Δ¯\overline{\Delta}. By a slight abuse of notations I denote the curve CΔ¯C\sum_{C\in\overline{\Delta}}C by Δ¯\overline{\Delta} as well.

2. The represenatives of the 𝒜5\mathcal{A}_{5}-equivariant birational classes of 𝒜5\mathcal{A}_{5}\mathds{Q}-del Pezzo fibrations with the trivial action on the base

2.1. Representations of 𝒜5\mathcal{A}_{5} and its central extension

Recall that 𝒜5\mathcal{A}_{5} has the following irreducible representations: I=I1I=I_{1}, W3W_{3}, W3W_{3}^{\prime}, W4W_{4}, and W5W_{5} (see [4, Sections 5.2 and 5.6], for details). The lower index is the dimension of the representation. The action of 𝒜5\mathcal{A}_{5} on 1\mathds{P}^{1} is induced by irreducible representations of the central extension 2.𝒜52.\mathcal{A}_{5}: U2U_{2} and U2U_{2}^{\prime}.

I recall some facts about the action of 𝒜5\mathcal{A}_{5} on smooth curves and surfaces

Lemma 2.1 ( [4, Lemmas 5.1.3, 5.1.4 and Section 5.2]).

Let SS be a smooth surface and CC be a smooth curve admitting a non-trivial action of 𝒜5\mathcal{A}_{5}.

  1. (1)

    Let Σ\Sigma be an 𝒜5\mathcal{A}_{5}-orbit on CC, then |Σ|=12\lvert\Sigma\rvert=12, 2020, or 6060.

  2. (2)

    Let Σ\Sigma be an 𝒜5\mathcal{A}_{5}-orbit on SS, then |Σ|5\lvert\Sigma\rvert\geqslant 5.

Proposition 2.2 ([4, Chapter 6]).

Let π:X1\pi:X\to\mathds{P}^{1} be an 𝒜5\mathcal{A}_{5}\mathds{Q}-del Pezzo fibration and let FF be a general fiber of π\pi. Then one of the following holds

  1. (1)

    F𝒜5(W3)F\cong_{\mathcal{A}_{5}}\mathds{P}(W_{3}) or F𝒜5(W3)F\cong_{\mathcal{A}_{5}}\mathds{P}(W_{3}^{\prime}),

  2. (2)

    F𝒜5S5F\cong_{\mathcal{A}_{5}}S_{5},

  3. (3)

    F𝒜5S3F\cong_{\mathcal{A}_{5}}S_{3},

  4. (4a)

    F𝒜5(U2)×(U2)F\cong_{\mathcal{A}_{5}}\mathds{P}(U_{2})\times\mathds{P}(U_{2}) or F𝒜5(U2)×(U2)F\cong_{\mathcal{A}_{5}}\mathds{P}(U_{2}^{\prime})\times\mathds{P}(U_{2}^{\prime}), in this case I say FF is a quadric with a diagonal 𝒜5\mathcal{A}_{5}-action,

  5. (4b)

    F𝒜5(U2)×(U2)F\cong_{\mathcal{A}_{5}}\mathds{P}(U_{2})\times\mathds{P}(U_{2}^{\prime}), this case I say that FF is a quadric with a twisted diagonal 𝒜5\mathcal{A}_{5}-action,

  6. (4c)

    F𝒜5(U2)×(II)F\cong_{\mathcal{A}_{5}}\mathds{P}(U_{2})\times\mathds{P}(I\oplus I) or F𝒜5(U2)×(II)F\cong_{\mathcal{A}_{5}}\mathds{P}(U_{2}^{\prime})\times\mathds{P}(I\oplus I), in this case I say that FF is a quadric with a one-factor 𝒜5\mathcal{A}_{5}-action.

Knowing the induced 𝒜5\mathcal{A}_{5}-action on 3\mathds{P}^{3} is very useful for studying cases (4a), (4b), and (4c).

Lemma 2.3.

Let FF be a quadric with an action of 𝒜5\mathcal{A}_{5}, let VV be the 44-dimensional representation dual to H0(F,12KF)H^{0}(F,-\frac{1}{2}K_{F}), and let g:F(V)g:F\hookrightarrow\mathds{P}(V) be the 𝒜5\mathcal{A}_{5}-equivariant embedding induced by |12KF|\lvert-\frac{1}{2}K_{F}\rvert. Then

  1. (1)

    the 𝒜5\mathcal{A}_{5}-action on FF is diagonal if and only if V𝒜5W3IV\cong_{\mathcal{A}_{5}}W_{3}\oplus I or V𝒜5W3IV\cong_{\mathcal{A}_{5}}W_{3}^{\prime}\oplus I;

  2. (2)

    the 𝒜5\mathcal{A}_{5}-action on FF is twisted diagonal if and only if V𝒜5W4V\cong_{\mathcal{A}_{5}}W_{4};

  3. (3)

    the 𝒜5\mathcal{A}_{5}-action on FF is a one-factor action if and only if (V)𝒜5(U2U2)\mathds{P}(V)\cong_{\mathcal{A}_{5}}\mathds{P}(U_{2}\oplus U_{2}), or (V)𝒜5(U2U2)\mathds{P}(V)\cong_{\mathcal{A}_{5}}\mathds{P}(U_{2}^{\prime}\oplus U_{2}^{\prime}).

Proof.

Assertion (1) is equivalent to [4, Lemma 6.3.3, (i)].

Clearly, (V)\mathds{P}(V) has no invariant lines and planes if and only if FF has no invariant divisors of degree (1,1)(1,1) or (0,1)(0,1). It follows that the action on FF is twisted diagonal if and only if V𝒜5W4V\cong_{\mathcal{A}_{5}}W_{4}.

Suppose F𝒜5(U2)×(II)F\cong_{\mathcal{A}_{5}}\mathds{P}(U_{2})\times\mathds{P}(I\oplus I). Let L1L_{1} and L2L_{2} be the curves of bi-degree (0,1)(0,1) on FF. Then L1𝒜5L2𝒜5(U2)L_{1}\cong_{\mathcal{A}_{5}}L_{2}\cong_{\mathcal{A}_{5}}\mathds{P}(U_{2}). Since g(L1)g(L_{1}) and g(L2)g(L_{2}) are 𝒜5\mathcal{A}_{5}-invariant skew lines we see that VU2U2V\cong U_{2}\oplus U_{2}. The converse follows from the assertions (1) and (2). ∎

2.2. The general fiber is 2\mathds{P}^{2}

Lemma 2.4.

Let π:X1\pi\colon X\to\mathds{P}^{1} be an 𝒜5\mathcal{A}_{5}\mathds{Q}-del Pezzo fibration of degree 99 and suppose 𝒜5\mathcal{A}_{5} acts trivially on the base. Then XX is 𝒜5\mathcal{A}_{5}-equivariantly birational to 2×1\mathds{P}^{2}\times\mathds{P}^{1} with the trivial action of 𝒜5\mathcal{A}_{5} on 1\mathds{P}^{1}.

Proof.

Let XηX_{\eta} be the generic fiber of π\pi. It is a form of 2\mathds{P}^{2} and has a point, thus it is 2\mathds{P}^{2}. The isomorphism Xη2X_{\eta}\cong\mathds{P}^{2} induces the 𝒜5\mathcal{A}_{5}-equivariant birational map to 2×1\mathds{P}^{2}\times\mathds{P}^{1}. ∎

Note that 2×1\mathds{P}^{2}\times\mathds{P}^{1} is not the only 𝒜5\mathcal{A}_{5}\mathds{Q}-del Pezzo fibration of degree 99 with a trivial action on the base.

Example 2.5.

Let Δ\Delta be the 𝒜5\mathcal{A}_{5}-invariant curve of degree 22 in the fiber F2×1F\subset\mathds{P}^{2}\times\mathds{P}^{1} over a point P1P\in\mathds{P}^{1}. Then I may blow up 2×1\mathds{P}^{2}\times\mathds{P}^{1} at Δ\Delta and then contract the proper transform of FF to acquire a variety V1V_{1} with a 12(1,1,1)\frac{1}{2}(1,1,1)-singularity. The composition 2×1V1\mathds{P}^{2}\times\mathds{P}^{1}\dasharrow V_{1} of these maps is an elementary 𝒜5\mathcal{A}_{5}-equivariant Sarkisov link of 𝒜5\mathcal{A}_{5}\mathds{Q}-del Pezzo fibrations. The new fiber over PP is isomorphic to (1,1,4)\mathds{P}(1,1,4).

Example 2.6.

Consider a toric variety RnR_{n} such that CoxRn=[u,v,x,y,z,w]\operatorname{Cox}R_{n}=\mathds{C}[u,v,x,y,z,w], where the irrelevant ideal I=u,vx,y,z,wI=\langle u,v\rangle\cap\langle x,y,z,w\rangle and the grading given by

(uvxyzw00111211000n)\left(\begin{array}[]{ccccccc}u&v&x&y&z&w\\ 0&0&1&1&1&2\\ 1&1&0&0&0&-n\end{array}\right)

I assume that 𝒜5\mathcal{A}_{5}-action on RnR_{n} comes from the identification of x,y,z3W3\mathds{C}^{3}_{x,y,z}\cong W_{3}. Suppose that the equation of the 𝒜5\mathcal{A}_{5}-invariant quadric on W3W_{3} is xzy2=0xz-y^{2}=0. Then consider an 𝒜5\mathcal{A}_{5}-invariant hypersurface VnV_{n} in RnR_{n} given by the equation

an(u,v)w=xzy2.a_{n}(u,v)w=xz-y^{2}.

There is a natural 𝒜5\mathcal{A}_{5}-equivariant projection πR:Rnu,v1\pi_{R}\colon R_{n}\to\mathds{P}^{1}_{u,v} which is a (1,1,1,2)\mathds{P}(1,1,1,2)-bundle. The restriction π=πR|Vn:Vnu,v1\pi=\pi_{R}|_{V_{n}}\colon V_{n}\to\mathds{P}^{1}_{u,v} is an 𝒜5\mathcal{A}_{5}\mathds{Q}-del Pezzo fibration of degree 99 with the trivial action on the base. The variety VnV_{n} has 22-Goreinstein singularities at x=y=z=an(u,v)=0x=y=z=a_{n}(u,v)=0. For every simple root of an(u,v)a_{n}(u,v) there is a 12(1,1,1)\frac{1}{2}(1,1,1)-singularity and for a root of multiplicity kk there is a cA1/μ2cA_{1}/\mu_{2}-singularity, where μ2\mu_{2} is the cyclic group of order 22.

Question 2.7.

Let π:X1\pi\colon X\to\mathds{P}^{1} be an 𝒜5\mathcal{A}_{5}\mathds{Q}-del Pezzo fibration and suppose that the general fiber of π\pi is 2\mathds{P}^{2}. Suppose 𝒜5\mathcal{A}_{5} acts trivially on the base. Is it true that XVnX\cong V_{n} for some an(u,v)a_{n}(u,v)?

2.3. The general fiber is S5S_{5}

Proposition 2.8.

Let π:X1\pi:X\to\mathds{P}^{1} be an A5A_{5}\mathds{Q}-del Pezzo fibration of degree 55. Then it is 𝒜5\mathcal{A}_{5}-equivariantly birational to S5×1S_{5}\times\mathds{P}^{1}.

Proof.

Consider the generic fiber XηX_{\eta} of π\pi. It is a del Pezzo surface of degree 55 over (t)\mathds{C}(t) which admits an action of 𝒜5\mathcal{A}_{5}. Recall, that a del Pezzo surface of degree 55 over an algebraically closed field is unique and it admits a faithful 𝒜5\mathcal{A}_{5}-action. Hence the 𝒜5\mathcal{A}_{5}-action on Pic(Xη¯)\operatorname{Pic}(\overline{X_{\eta}}) is faithful and induces an embedding 𝒜5W(A4)𝒮5\mathcal{A}_{5}\hookrightarrow\operatorname{W}(A_{4})\cong\mathcal{S}_{5}, where A4A_{4} is the root lattice.

Let Γ=Gal((t)¯/(t))\Gamma=\operatorname{Gal}(\overline{\mathds{C}(t)}/\mathds{C}(t)), then the action of Γ\Gamma and 𝒜5\mathcal{A}_{5} on X¯η\overline{X}_{\eta} commutes. It follows that Γ\Gamma commutes with W(A4)=𝒮5\operatorname{W}(A_{4})=\mathcal{S}_{5}, therefore all (1)(-1)-curves of X¯η\overline{X}_{\eta} are defined over (t)\mathds{C}(t). Thus XηX_{\eta} is acquired by a blow up of 2\mathds{P}^{2} at 44 points defined over (t)\mathds{C}(t). In particular a del Pezzo surface of degree 55 with an 𝒜5\mathcal{A}_{5}-action is unique.

On the other hand the generic fiber of the projection S5×11S_{5}\times\mathds{P}^{1}\to\mathds{P}^{1} is a del Pezzo surface of degree 55 with an action of 𝒜5\mathcal{A}_{5}, hence isomorphic to XηX_{\eta}. The isomorphism induces the 𝒜5\mathcal{A}_{5}-equivariant birational map XS5×1X\dasharrow S_{5}\times\mathds{P}^{1}.

Unlike the degree 99 case, one can show that there are no other 𝒜5\mathcal{A}_{5}\mathds{Q}-del Pezzo fibrations of degree 55.

Lemma 2.9.

The threefold S5×1S_{5}\times\mathds{P}^{1} is the unique 𝒜5\mathcal{A}_{5}\mathds{Q}-del Pezzo fibration of degree 55.

Proof.

It is well known that lct𝒜5S51\operatorname{lct}_{\mathcal{A}_{5}}S_{5}\geqslant 1, thus the statement follows from [2, Theorem 1.5] (see [11, Theorem 3.6] for GG-invariant version). ∎

I expect that S5×1S_{5}\times\mathds{P}^{1} is the unique 𝒜5\mathcal{A}_{5}\mathds{Q}-del Pezzo fibration in its birational class but it is 𝒜5\mathcal{A}_{5}-equivariantly birational to other 𝒜5\mathcal{A}_{5}-Mori fiber spaces. We can easily construct many 1\mathds{P}^{1}-bundles 𝒜5\mathcal{A}_{5}-equivariantly birational to S5×1S_{5}\times\mathds{P}^{1} by blowing up orbits of fibers of the projection onto S5S_{5}.

Question 2.10.

Does there exist an 𝒜5\mathcal{A}_{5}\mathds{Q}-Mori fiber space which is not 𝒜5\mathcal{A}_{5}-equivariantly square birational to S5×1/1S_{5}\times\mathds{P}^{1}/\mathds{P}^{1} or S5×1/S5S_{5}\times\mathds{P}^{1}/S_{5} but is 𝒜5\mathcal{A}_{5}-equivariantly birational to S5×1S_{5}\times\mathds{P}^{1}?

2.4. The general fiber is S3S_{3}

Lemma 2.11.

Let XηX_{\eta} be a cubic over (t)\mathds{C}(t) admitting a faithful 𝒜5\mathcal{A}_{5}-action, then ρ𝒜5(Xη)=2\rho^{\mathcal{A}_{5}}(X_{\eta})=2. In particular, the 𝒜5\mathcal{A}_{5}\mathds{Q}-del Pezzo fibrations of degree 33 do not exist.

Proof.

There is the unique 𝒜5\mathcal{A}_{5}-invariant cubic S3S_{3} in 3\mathds{P}^{3}: Clebsh cubic. It follows that there is a unique 𝒜5\mathcal{A}_{5}-invariant cubic XηX_{\eta} in (t)3\mathds{P}^{3}_{\mathds{C}(t)}. I may realize XηX_{\eta} as the generic fiber of the projection S3×11S_{3}\times\mathds{P}^{1}\to\mathds{P}^{1}. Thus ρ𝒜5(Xη)=2\rho^{\mathcal{A}_{5}}(X_{\eta})=2. ∎

2.5. The general fiber is a quadric

Recall that there are three types of 𝒜5\mathcal{A}_{5}-action on 1×1\mathds{P}^{1}\times\mathds{P}^{1} as defined in Proposition 2.2: diagonal, twisted diagonal, and one-factor action. The 𝒜5\mathcal{A}_{5}-quadric fibration exist only for the former type.

Lemma 2.12.

Let π:X1\pi:X\to\mathds{P}^{1} be an 𝒜5\mathcal{A}_{5}-equivariant map such that the 𝒜5\mathcal{A}_{5}-action on a general fiber FF of π\pi is either twisted diagonal or a one-factor action. Let XηX_{\eta} be the generic fiber of π\pi, then ρ𝒜5(Xη)=2\rho^{\mathcal{A}_{5}}(X_{\eta})=2.

Proof.

Suppose the 𝒜5\mathcal{A}_{5}-action on FF is twisted diagonal, then F(W4)F\subset\mathds{P}(W_{4}) by Lemma 2.3. Since Sym2(W4)W5W4I\operatorname{Sym}^{2}(W_{4})\cong W_{5}\oplus W_{4}\oplus I, there is the unique 𝒜5\mathcal{A}_{5}-invariant quadric SηS_{\eta} in (W4(t))\mathds{P}(W_{4}\otimes\mathds{C}(t)). On the other hand, SηS_{\eta} is also the generic fiber of the projection

(U2)×(U2)×(II)(II),\mathds{P}(U_{2})\times\mathds{P}(U_{2}^{\prime})\times\mathds{P}(I\oplus I)\to\mathds{P}(I\oplus I),

thus we see that ρ𝒜5(Sη)=2\rho^{\mathcal{A}_{5}}(S_{\eta})=2.

Suppose the 𝒜5\mathcal{A}_{5}-action on FF is a one-factor action corresponding to the representation U2U_{2} (or U2U_{2}^{\prime}) of 2.𝒜52.\mathcal{A}_{5}. Similarly to the twisted diagonal case I only need to show that there is a unique 𝒜5\mathcal{A}_{5}-invariant quadric in (U2U2)\mathds{P}(U_{2}\oplus U_{2}) (resp. (U2U2)\mathds{P}(U_{2}^{\prime}\oplus U_{2}^{\prime})). There are 𝒜5\mathcal{A}_{5}-invariant skew lines Z1Z_{1} and Z2Z_{2} and any 𝒜5\mathcal{A}_{5}-invariant quadric must contain both Z1Z_{1} and Z2Z_{2} since there are no orbits of length 2\leqslant 2 on 1\mathds{P}^{1}. On the other hand, the space of quadrics containing a pair of 𝒜5\mathcal{A}_{5}-invariant skew lines is isomorphic to (W)\mathds{P}(W) for some 44-dimensional representation WW of 𝒜5\mathcal{A}_{5} or 2.𝒜52.\mathcal{A}_{5}. The image on (V)\mathds{P}(V) of the quadric with a one-factor action corresponds to an 𝒜5\mathcal{A}_{5}-fixed point on (W)\mathds{P}(W). On the other hand, the family of reducible quadrics containing Z1Z_{1} and Z2Z_{2} is of dimension 33 and is 𝒜5\mathcal{A}_{5}-invariant. It follows that W𝒜5W3IW\cong_{\mathcal{A}_{5}}W_{3}\oplus I or W𝒜5W3IW\cong_{\mathcal{A}_{5}}W_{3}^{\prime}\oplus I. Thus the 𝒜5\mathcal{A}_{5}-invariant quadric on (U2U2)\mathds{P}(U_{2}\oplus U_{2}) (resp. (U2U2)\mathds{P}(U_{2}^{\prime}\oplus U_{2}^{\prime})) is unique. ∎

Recall the construction of the 𝒜5\mathcal{A}_{5}\mathds{Q}-quadric fibrations XnX_{n} and XnX_{n}^{\prime} from Example 1.7.

Proposition 2.13.

Let πV:V1\pi_{V}\colon V\to\mathds{P}^{1} be an 𝒜5\mathcal{A}_{5}\mathds{Q}-quadric fibration. Suppose the action on a general fiber FF of πV\pi_{V} is diagonal. Then VV is 𝒜5\mathcal{A}_{5}-equivariantly birational to a smooth XnX_{n} or XnX_{n}^{\prime} for some a2n(u,v)a_{2n}(u,v).

Proof.

The fiber FF embeds into (W3I)\mathds{P}(W_{3}\oplus I) or (W3I)\mathds{P}(W_{3}^{\prime}\oplus I) by Lemma 2.3. The two cases are analogous, suppose the former, then

Xη((W3I)(t)).X_{\eta}\hookrightarrow\mathds{P}\big{(}(W_{3}\oplus I)\otimes\mathds{C}(t)\big{)}.

Let x,y,zx,y,z be the coordinates on W3W_{3} and let ww be the coordinate on II. Then, up to a change of coordinates on W3W_{3}, every 𝒜5\mathcal{A}_{5}-invariant quadric in (W3I)\mathds{P}(W_{3}\oplus I) has the equation

λ(t)w2=μ(t)(xzy2).\lambda(t)w^{2}=\mu(t)(xz-y^{2}).

It follows that up to a change of coordinates in ((W3I)(t))\mathds{P}\big{(}(W_{3}\oplus I)\otimes\mathds{C}(t)\big{)} the equation of XηX_{\eta} is

b(t)w2=xzy2,b(t)w^{2}=xz-y^{2},

where b(t)b(t) has no multiple roots.

Set a(u,v)a(u,v) to be the homogeneous polynomial of degree 2n2n without multiple factors satisfying a(t,1)=b(t)a(t,1)=b(t). Let XnX_{n} be the hypersurface in TnT_{n} given by

a(u,v)w2=xzy2.a(u,v)w^{2}=xz-y^{2}.

Then the general fiber of Xn1X_{n}\to\mathds{P}^{1} is 𝒜5\mathcal{A}_{5}-equivariantly isomorphic to XηX_{\eta}. This isomorphism of the general fibers induces the 𝒜5\mathcal{A}_{5}-equivariant birational map VXnV\dasharrow X_{n}. ∎

This completes the proof of Theorem 1.8.

3. Rigidity results

This section is devoted to proving Theorem 1.11. First, I present some elementary results on the geometry of varieties XnX_{n}. Next I recall Noether-Fano method and the notion of maximal centers. Then I show that the only possible maximal centers are the points on the 𝒜5\mathcal{A}_{5}-invariant curve of degree 22. At last, I use the technique of supermaximal singularities [18] to prove 𝒜5\mathcal{A}_{5}-equivariant birational superrigidity of XnX_{n} and XnX_{n}^{\prime} for n2n\geqslant 2.

3.1. The geometry of XnX_{n}

Note that TnTnT_{n}\cong T_{n}^{\prime} and for a fixed a2na_{2n} we have XnXnX_{n}\cong X_{n}^{\prime}. These isomorphisms are not 𝒜5\mathcal{A}_{5}-equivariant. From now on I work with TnT_{n} and XnX_{n}, the proofs for XnX_{n}^{\prime} are identical. In this section I do not assume that a2n(u,v)a_{2n}(u,v) has no multiple factors. In that case XnX_{n} is no longer smooth. Indeed, for each multiple linear factor l(u,v)l(u,v) of a2n(u,v)a_{2n}(u,v) the point (x=y=z=l(u,v)=0)(x=y=z=l(u,v)=0) is a cA1cA_{1}-singularity.

Denote by HTH_{T} the divisor class of (x=0)(x=0) and by FTF_{T} the divisor class of (u=0)(u=0) on TnT_{n}. Denote H=HT|XnH=H_{T}|_{X_{n}} and F=FT|XnF=F_{T}|_{X_{n}}. Let sTA3(Tn)s_{T}\in A^{3}(T_{n}) and sA2(Xn)s\in A^{2}(X_{n}) be the classes of the curve x=y=w=0x=y=w=0. Let fTA3(Tn)f_{T}\in A^{3}(T_{n}) and fA2(Xn)f\in A^{2}(X_{n}) be the classes of the curve x=y=u=0x=y=u=0. Using the fact that

(x=y=z=w=0)=and(x=y=z=u=0)=pt(x=y=z=w=0)=\varnothing\quad\text{and}\quad(x=y=z=u=0)=\operatorname{pt}

I compute the intersections on TnT_{n}.

Lemma 3.1.

The following holds for TnT_{n}:

  1. (1)

    The classes sTs_{T} and fTf_{T} generate the cone of effective curves on TnT_{n};

  2. (2)

    HT2FTfTH_{T}^{2}\cdot F_{T}\equiv f_{T};

  3. (3)

    HT3sT+nfTH_{T}^{3}\equiv s_{T}+nf_{T};

  4. (4)

    HT3FT=1H_{T}^{3}\cdot F_{T}=1;

  5. (5)

    HT4=nH_{T}^{4}=n.

Corollary 3.2.

The following holds for XnX_{n}:

  1. (1)

    HF2fH\cdot F\equiv 2f;

  2. (2)

    H22s+2nfH^{2}\equiv 2s+2nf;

  3. (3)

    H2F=2H^{2}\cdot F=2;

  4. (4)

    H3=2nH^{3}=2n;

  5. (5)

    KXn2H+(n2)FK_{X_{n}}\sim-2H+(n-2)F;

  6. (6)

    KXn28s+24f8nfK_{X_{n}}^{2}\equiv 8s+24f-8nf.

In order to better understand a singular XnX_{n}, it is useful to know how is it related to a smooth XmX_{m}.

Lemma 3.3.

Let variety XnX_{n} correspond to a2n(u,v)a_{2n}(u,v) and Xn+1X_{n+1} to u2a2n(u,v)u^{2}a_{2n}(u,v). Denote the fiber u=0u=0 by FF. Let σ:X~Xn\sigma\colon\widetilde{X}\to X_{n} be the blow up of XnX_{n} along the curve u=w=0u=w=0. Then there is a map ψ:X~Xn+1\psi\colon\widetilde{X}\to X_{n+1} contracting σ1F\sigma^{-1}F to a cA1cA_{1} singularity x=y=z=u=0x=y=z=u=0. The composition φu=ψσ1\varphi_{u}=\psi\circ\sigma^{-1} is an 𝒜5\mathcal{A}_{5}-equivariant elementary Sarkisov link.

X~\textstyle{\widetilde{X}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}σ\scriptstyle{\sigma}ψ\scriptstyle{\psi}Xn\textstyle{X_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}φu\scriptstyle{\varphi_{u}}Xn+1\textstyle{X_{n+1}}
Proof.

Elementary calculations. ∎

Thus we see that any singular XnX_{n} is 𝒜5\mathcal{A}_{5}-equivariantly square birational to a smooth XmX_{m} with m=nkm=n-k for some k>0k>0.

Lemma 3.4.

Suppose a2n(u,v)a_{2n}(u,v) is not a square, then

  1. (1)

    PicXnHF\operatorname{Pic}X_{n}\cong\mathds{Z}\cdot H\oplus\mathds{Z}\cdot F,

  2. (2)

    The classes ss and ff generate the cone of effective curves NE(Xn)\operatorname{NE}(X_{n}).

Proof.

The assetion (1) holds since PicXnPicXηPic1\operatorname{Pic}X_{n}\cong\operatorname{Pic}X_{\eta}\oplus\operatorname{Pic}\mathds{P}^{1} and PicXη\operatorname{Pic}X_{\eta} is generated by the hyperplane section if and only if a2n(u,v)a_{2n}(u,v) is not a complete square.

The assertion (2) follows from (1) and Lemma 3.1 (1) for smooth XnX_{n} by Poincare duality. Any XnX_{n} is related to some smooth XmX_{m} by a sequence of elementary Sarkisov links described in Lemma 3.3. The elementary links preserve the dimension of NE(Xn)\operatorname{NE}(X_{n}), hence the assertion holds by Lemma 3.1 (1). ∎

Lemma 3.5.

Suppose n2n\geqslant 2, then XnX_{n} satisfies the KK-condition, that is KXn-K_{X_{n}} is not in the interior of the cone of movable divisors.

Proof.

The linear system |H|\lvert H\rvert defines a divisorial contraction σ:XnYn\sigma\colon X_{n}\to Y_{n}, where YnY_{n} is a hypersurface in (1,1,n,n,n)\mathds{P}(1,1,n,n,n) given by the equation

a2n(u,v)=xzy2.a_{2n}(u,v)=xz-y^{2}.

Thus the cone of movable divisors of XnX_{n} is generated by HH and FF which implies the statement of the lemma by Corollary 3.2 (5). ∎

3.2. Noether-Fano method

For definitions of canonical singularities of pairs we refer the reader to [10, pages 16-17] and [5, Definition 2.1]. Let π:X1\pi:X\to\mathds{P}^{1} be an 𝒜5\mathcal{A}_{5}-del Pezzo fibration.

Suppose we are given a birational map χ:XY\chi\colon X\dasharrow Y to a Mori fiber space πY:YZ\pi_{Y}\colon Y\to Z. Let Y\mathcal{M}_{Y} be a very ample complete linear system on YY. I say =χ1Y\mathcal{M}=\chi^{-1}\mathcal{M}_{Y} is a mobile linear system associated to χ\chi. There are numbers λ+\lambda\in\mathds{Q}_{+} and ll\in\mathds{Q} such that λKX+lF\lambda\mathcal{M}\sim-K_{X}+lF. The Noether-Fano inequality is the essential result used to prove birational rigidity-type results.

Theorem 3.6 (Noether-Fano inequality, [6, Theorem 4.2]).

Suppose l>0l>0 and (X,λ)(X,\lambda\mathcal{M}) has canonical singularities, then χ\chi is isomorphism.

Let EE be a divisorial valuation of (X)\mathds{C}(X). If (E,X,λ)<0(E,X,\lambda\mathcal{M})<0, then I say that EE is a maximal singularity of the pair (X,λ)(X,\lambda\mathcal{M}). I call the center ZZ of EE on XX a maximal center of the pair (X,λ)(X,\lambda\mathcal{M}).

Now I examine, which subvarieties ZXZ\subset X can be maximal centers.

Set X=XnX=X_{n} for some a2n(u,v)a_{2n}(u,v) that is not a square and let χ\chi, \mathcal{M}, λ\lambda, ll be as above.

Let CC be an irreducible curve on XX, I say CC is horizontal if π(C)=1\pi(C)=\mathds{P}^{1} and vertical if π(C)\pi(C) is a point. I say that a curve CC is horizontal (resp. vertical) if every irreducible component of CC is horizontal (resp. vertical).

Let CXC\subset X be a horizontal or a vertical curve. I define its degree as follows

deg(C)={KXC/2,if C is verticalCF,if C is horizontal\deg(C)=\begin{cases}-K_{X}\cdot C/2,&\text{if }C\text{ is vertical}\\ C\cdot F,&\text{if }C\text{ is horizontal}\end{cases}
Lemma 3.7.

Suppose (X,λ)(X,\lambda\mathcal{M}) is not canonical at a curve Δ\Delta, then Δ\Delta is an 𝒜5\mathcal{A}_{5}-invariant vertical curve of degree 22.

Proof.

Let Δ¯\overline{\Delta} be the 𝒜5\mathcal{A}_{5}-orbit of Δ\Delta. The pair (X,λ)(X,\lambda\mathcal{M}) is not canonical at a curve Δ\Delta if and only if multΔλ>1\operatorname{mult}_{\Delta}\lambda\mathcal{M}>1 and hence if and only if multΔ¯λ>1\operatorname{mult}_{\overline{\Delta}}\lambda\mathcal{M}>1.

Suppose Δ\Delta is horizontal, let FF be a general fiber of π\pi and let D1,D2D_{1},D_{2}\in\mathcal{M} be general divisors. Then the set-theoretic intersection

Σ=FΔ¯FD1D2\Sigma=F\cap\overline{\Delta}\subset F\cap D_{1}\cap D_{2}

is a union of orbits on FF. Hence

8=FλD1λD2>λ2FΔ¯|Σ|12,8=F\cdot\lambda D_{1}\cdot\lambda D_{2}>\lambda^{2}F\cdot\overline{\Delta}\geqslant\big{|}\Sigma\big{|}\geqslant 12,

a contradiction.

Suppose Δ¯F\overline{\Delta}\subset F and let DD be general in \mathcal{M}. Let HF=H|FH_{F}=H|_{F}, then D|F2HFD|_{F}\sim 2H_{F} and ordΔ¯D|F>1\operatorname{ord}_{\overline{\Delta}}D|_{F}>1. It follows that degΔ¯4\deg\overline{\Delta}\leqslant 4 and [4, Lemma 6.4.4] implies that Δ¯=Δ\overline{\Delta}=\Delta is the 𝒜5\mathcal{A}_{5}-invariant curve of degree 22.

Suppose FF is singular, then it is a cone over 1\mathds{P}^{1} with the 𝒜5\mathcal{A}_{5}-action inherited from 1\mathds{P}^{1}. Let CC be the unique 𝒜5\mathcal{A}_{5}-invariant curve of degree 22 on FF. Suppose ΔC\Delta\neq C and denote ΔkHF\Delta\sim kH_{F}. Set Σ=ΔC\Sigma=\Delta\cap C, it is a union of 𝒜5\mathcal{A}_{5}-orbits on CC hence |Σ|12\lvert\Sigma\rvert\geqslant 12 by Lemma 2.1. It follows that ΔC12\Delta\cdot C\geqslant 12 and k6k\geqslant 6. On the other hand, D|F2HFD|_{F}\sim 2H_{F}, a contradiction. ∎

Lemma 3.8.

Let PP be an 𝒜5\mathcal{A}_{5}-invariant point and suppose XX is smooth at PP. Then pair (X,λ)(X,\lambda\mathcal{M}) is canonical at PP.

Proof.

Suppose (X,λ)(X,\lambda\mathcal{M}) is not canonical at PP and let EE_{\infty} be the divisorial valuation over XX such that a(E,X,λ)<0a(E_{\infty},X,\lambda\mathcal{M})<0.

First, observe that a fiber FF of π\pi containing PP is a quadratic cone and PP is its vertex. Let σ:X~X\sigma\colon\widetilde{X}\to X be the blow up at PP and let EE be the exceptional divisor of σ\sigma. Let LL be a general line through PP, then for general DD\in\mathcal{M}

multPλDLλD=2.\operatorname{mult}_{P}\lambda D\leqslant L\cdot\lambda D=2.

It follows that a(E,X,λ)0a(E,X,\lambda\mathcal{M})\geqslant 0, hence the center BB of EE_{\infty} on X~\widetilde{X} is a point or a curve on EE.

Note that the action of 𝒜5\mathcal{A}_{5} on EE is non-trivial. Indeed, the point PP up to a change of coordinates on u,v1\mathds{P}^{1}_{u,v} has the equations u=x=y=z=0u=x=y=z=0 and the local equation of XX near PP is u=0u=0, thus E𝒜5(W3)E\cong_{\mathcal{A}_{5}}\mathds{P}(W_{3}). Denote λ~E=(σ1)|E\widetilde{\lambda}\mathcal{M}_{E}=(\sigma^{-1}\mathcal{M})|_{E}, then ordBλ~E>1\operatorname{ord}_{B}\lambda\widetilde{\mathcal{M}}_{E}>1 and deg~E=multPλ2\deg\widetilde{\mathcal{M}}_{E}=\operatorname{mult}_{P}\lambda\mathcal{M}\leqslant 2. On the other hand, if BB is a curve, then degB¯2\deg\overline{B}\geqslant 2, a contradiction.

Suppose BB is a point and let B¯\overline{B} be the 𝒜5\mathcal{A}_{5}-orbit of BB. Then |B¯|6\big{|}\overline{B}\big{|}\geqslant 6 and there are 44 points P1,P2,P3,P4B¯E2P_{1},P_{2},P_{3},P_{4}\in\overline{B}\subset E\cong\mathds{P}^{2} in general position. I claim that

(1) i=14multPiC2degC\sum_{i=1}^{4}\operatorname{mult}_{P_{i}}C\leqslant 2\deg C

for any curve CEC\subset E. Indeed, denote by LijL_{ij} the line on EE passing through PiP_{i} and PjP_{j}. Decomposing C=C+αijLijC=C^{\prime}+\sum\alpha_{ij}L_{ij} and counting multiplicities I conclude the inequality (1). Thus

4<4multBλE=i=14multPiλE4,4<4\operatorname{mult}_{B}\lambda\mathcal{M}_{E}=\sum_{i=1}^{4}\operatorname{mult}_{P_{i}}\lambda\mathcal{M}_{E}\leqslant 4,

a contradiction. ∎

It is possible that the pair (X,λ)(X,\lambda\mathcal{M}) is not canonical at 𝒜5\mathcal{A}_{5}-invariant curves of degree 22. On the other hand, the elementary Sarkisov links originating from them are described in Lemma 3.3, these links are 𝒜5\mathcal{A}_{5}-equivariant square birational maps. Using these links I acquire a new pair which is canonical at all curves.

Let b(u,v)b(u,v) be a polynomial of degree kk. Then there is the associated map φb:XXb\varphi_{b}\colon X\dasharrow X_{b}, where XbX_{b} is a hypersurface in Tn+kT_{n+k} given by the equation a(u,v)(b(u,v))2w2=xzy2a(u,v)(b(u,v))^{2}w^{2}=xz-y^{2}. The map φb\varphi_{b} is the composition of elementary links described in Lemma 3.3. Denote b=φb1\mathcal{M}_{b}=\varphi_{b}^{-1}\mathcal{M}.

Proposition 3.9.

Suppose XX is smooth, then there is b(u,v)b(u,v) such that the pair (Xb,λb)(X_{b},\lambda\mathcal{M}_{b}) is canonical at curves and 𝒜5\mathcal{A}_{5}-invariant points.

Proof.

I prove the proposition by playing the two-ray game. Suppose (X,λ)(X,\lambda\mathcal{M}) is not canonical at a curve Δ1\Delta_{1}. By Lemma 3.7 the curve Δ1\Delta_{1} is of degree 22 and is 𝒜5\mathcal{A}_{5}-invariant, hence its equations are l1(u,v)=w=0l_{1}(u,v)=w=0 for some linear l1(u,v)l_{1}(u,v). The elementary Sarkisov link starting at Δ1\Delta_{1} is the map φl1\varphi_{l_{1}}. Let l1=φl11\mathcal{M}_{l_{1}}=\varphi_{l_{1}}^{-1}\mathcal{M}. If the pair (Xl1,l1)(X_{l_{1}},\mathcal{M}_{l_{1}}) is canonical at curves, then I am done. Otherwise there is a curve Δ2\Delta_{2} and an elementary Sarkisov link φl2:Xl1Xl1l2\varphi_{l_{2}}\colon X_{l_{1}}\dasharrow X_{l_{1}l_{2}}, and I repeat the process as many times as required. The process terminates by [6, Theorem 6.1] and I set b=l1l2lkb=l_{1}l_{2}\dots l_{k}. The pair (Xb,λb)(X_{b},\lambda\mathcal{M}_{b}) is canonical at curves by construction and Lemma 3.7.

The pair (Xb,λb)(X_{b},\lambda\mathcal{M}_{b}) is canonical at smooth 𝒜5\mathcal{A}_{5}-invariant points by Lemma 3.8. I will now show that the pair (Xb,λb)(X_{b},\lambda\mathcal{M}_{b}) is canonical at singular 𝒜5\mathcal{A}_{5}-invariant points as well.

Recall that by [9, Theorem 1.1] if the pair (Xb,λb)(X_{b},\lambda\mathcal{M}_{b}) is not canonical at a cA1cA_{1}-point PP with the local equation

xzy2uN=0,xz-y^{2}-u^{N}=0,

then either a(Es,t,Xb,λb)<0a(E_{s,t},X_{b},\lambda\mathcal{M}_{b})<0, where Es,tE_{s,t} is the exceptional divisor of a (s,t,2ts,1)(s,t,2t-s,1)-weighted blow up at PP for some coprime stN/2s\leqslant t\leqslant N/2 or N=3N=3 and a(Esp,Xb,λb)<0a(E_{sp},X_{b},\lambda\mathcal{M}_{b})<0, where EspE_{sp} is the exceptional divisor of a (1,3,5,2)(1,3,5,2)-weighted blow up if N=3N=3.

Note that a(E1,1,Xb,λb)>0a(E_{1,1},X_{b},\lambda\mathcal{M}_{b})>0 by construction. Thus, if N=2N=2, we are done. We proceed by induction. If N=3N=3, then a(Esp,Xb,λb)<0a(E_{sp},X_{b},\lambda\mathcal{M}_{b})<0 and the pair (Xb/u,λb/u)(X_{b/u},\lambda\mathcal{M}_{b/u}) is not canonical at the 𝒜5\mathcal{A}_{5}-invariant point in the fiber u=0u=0, which contradicts Lemma 3.8.

Similarly if N4N\geqslant 4, then a(Es,t,Xb,λb)0a(E_{s,t},X_{b},\lambda\mathcal{M}_{b})\geqslant 0 for s2s\geqslant 2. Indeed, otherwise the pair (Xb/u,λb/u)(X_{b/u},\lambda\mathcal{M}_{b/u}) is not canonical at the 𝒜5\mathcal{A}_{5}-invariant point in the fiber u=0u=0, which contradicts the assumption of induction. Thus I may assume that s=1s=1 and t>1t>1. Let σ:X~bX\sigma\colon\widetilde{X}_{b}\to X be the blow up at PP and let ~b=σ1b\widetilde{\mathcal{M}}_{b}=\sigma^{-1}\mathcal{M}_{b}. Then the pair (X~n,λ~b)(\widetilde{X}_{n},\lambda\widetilde{\mathcal{M}}_{b}) is not canonical at a line LL on the exceptional divisor EE of σ\sigma. Hence, it is not canonical at each line in the orbit of LL. It follows that deg~b|E>6\deg\widetilde{\mathcal{M}}_{b}|_{E}>6 since the length of the orbit of LL is at least 1212, which contradicts a(E,Xb,b)>0a(E,X_{b},\mathcal{M}_{b})>0. ∎

It remains to show that the points that are not fixed by the 𝒜5\mathcal{A}_{5}-action cannot be maximal centers.

3.2.1. Orbits of points as non-canonical centers

Let PP be a maximal center of the pair (X,λ)(X,\lambda\mathcal{M}). I have already shown, that PP is a point which is not fixed by 𝒜5\mathcal{A}_{5}-action. In this subsection I show that PP must lie on an 𝒜5\mathcal{A}_{5}-invariant curve of degree 22.

Denote by FF the fiber containing PP and denote by P¯\overline{P} the 𝒜5\mathcal{A}_{5}-orbit of PP, then (X,λ)(X,\lambda\mathcal{M}) is not canonical at any point PiP¯P_{i}\in\overline{P}. Let Δ\Delta be the 𝒜5\mathcal{A}_{5}-invariant curve of degree 22 in the fiber containing PP.

Lemma 3.10.

Suppose PΔP\not\in\Delta, then one of the following holds:

  1. (1)

    There is an irreducible curve Γ\Gamma of degree 44 and distinct points P1,,P8P¯ΓP_{1},\dots,P_{8}\in\overline{P}\cap\Gamma such that Γ\Gamma is smooth at P1,,P8P_{1},\dots,P_{8};

  2. (2)

    There are smooth disintct irreducible curves Γ1,Γ2\Gamma_{1},\Gamma_{2} of degree 22, distinct points P1,,P4P¯Γ1P_{1},\dots,P_{4}\in\overline{P}\cap\Gamma_{1}, and distinct points P5,,P8P¯Γ2P_{5},\dots,P_{8}\in\overline{P}\cap\Gamma_{2};

  3. (3)

    The fiber FF is smooth, there is a smooth irreducible curve Γ\Gamma of bi-degree (2,1)(2,1) and there are distinct points P1,,P7P¯ΓP_{1},\dots,P_{7}\in\overline{P}\cap\Gamma.

  4. (4)

    The fiber FF is singular, there are disinct smooth irreducible curves Γ1,Γ2\Gamma_{1},\Gamma_{2} of degree 33, there are distinct points P1,,P7Γ1P¯P_{1},\dots,P_{7}\in\Gamma_{1}\cap\overline{P}, and there are distinct points P1,,P7Γ2P¯P_{1}^{\prime},\dots,P_{7}^{\prime}\in\Gamma_{2}\cap\overline{P}.

Proof.

For any 88 points on FF there is a unique quadric section passing through them. Let CC be the quadric section through P1,,P8P_{1},\dots,P_{8}, note that degC=4\deg C=4.

First, suppose CC is reducible, then CC has at most 33 components. If CC has 33 components, then there is a conic CC containing at least 66 points among P1,,P8P_{1},\dots,P_{8}. There is g𝒜5g\in\mathcal{A}_{5} such that gCgCgC\neq gC, thus I set Γ1=C\Gamma_{1}=C and Γ2=gC\Gamma_{2}=gC.

If CC is non-reduced, then simiarly to the previous case I set Γ1=C/2\Gamma_{1}=C/2 and Γ2=gΓ1\Gamma_{2}=g\Gamma_{1} for some gg such that gΓ1Γ1g\Gamma_{1}\neq\Gamma_{1}.

Suppose C=C1+C2C=C_{1}+C_{2} where C1,C2C_{1},C_{2} are irreducible. If degC1=degC2=2\deg C_{1}=\deg C_{2}=2, then I set Γi=Ci\Gamma_{i}=C_{i}. If degC1=1\deg C_{1}=1 and degC2=3\deg C_{2}=3, then C1C_{1} contains at most one points among P1,P8P_{1},\dots P_{8}. Thus I may assume that P1,,P7C2P_{1},\dots,P_{7}\in C_{2}. If FF is smooth, then it is the situation (3). For singular FF set Γ1=C2\Gamma_{1}=C_{2} and Γ2=gΓ1\Gamma_{2}=g\Gamma_{1} for some gg such that gΓ1Γ1g\Gamma_{1}\neq\Gamma_{1}.

Pick a point P9P¯P_{9}\in\overline{P} distinct from P1,,P8P_{1},\dots,P_{8}. Let CiC_{i} be the quadric section through points P1,,Pi1,Pi+1,,P9P_{1},\dots,P_{i-1},P_{i+1},\dots,P_{9}. I may assume that all CiC_{i} are irreducible, otherwise one of the previous cases applies. I may also assume that every CiC_{i} is singular at one of the points PjP_{j}, otherwise I am done. Note that CiC_{i} is singular at one point at most, hence if all CiC_{i} coincide, I am also done. Since C9Ci=8C_{9}\cdot C_{i}=8 the curve CiC_{i} must be singular at P9P_{9} for i9i\neq 9. But then C1C210C_{1}\cdot C_{2}\geqslant 10, a contradiction. ∎

Corollary 3.11.

Suppose (X,λ)(X,\lambda\mathcal{M}) is not caninical at a point PP, then PΔP\in\Delta.

Proof.

The proof is the case by case analysis for the curves Γ\Gamma, Γ1\Gamma_{1}, and Γ2\Gamma_{2} from Lemma 3.10. The case (1) is analogous to the case (3) and the case (4) is analogous the case (2).

Suppose there is a curve Γ\Gamma as in the case (3). Let DD be a general divisor in \mathcal{M} and denote DF=D|FD_{F}=D|_{F}. Then multPiλDF>1\operatorname{mult}_{P_{i}}\lambda D_{F}>1 for i=1,,7i=1,\dots,7. I decompose λDF=αΓ+C\lambda D_{F}=\alpha\Gamma+C, where ΓSuppC\Gamma\not\subset\operatorname{Supp}C and α1\alpha\leqslant 1 since bi-degree of λDF\lambda D_{F} equals (2,2)(2,2). Thus

7<i=17multPiλDF7α+CΓ=7α+66α=6+α7,7<\sum_{i=1}^{7}\operatorname{mult}_{P_{i}}\lambda D_{F}\leqslant 7\alpha+C\cdot\Gamma=7\alpha+6-6\alpha=6+\alpha\leqslant 7,

a contradiction.

Suppose there are curves Γ1,Γ2\Gamma_{1},\Gamma_{2} as in the case (2). Let DD be a general divisor in \mathcal{M} and denote DF=D|FD_{F}=D|_{F}. Then multPiλDF>1\operatorname{mult}_{P_{i}}\lambda D_{F}>1 for i=1,,8i=1,\dots,8. I decompose λDF=α1Γ1+α2Γ2+C\lambda D_{F}=\alpha_{1}\Gamma_{1}+\alpha_{2}\Gamma_{2}+C, where Γ1,Γ2SuppC\Gamma_{1},\Gamma_{2}\not\subset\operatorname{Supp}C and α1+α22\alpha_{1}+\alpha_{2}\leqslant 2 since degλDF=4\deg\lambda D_{F}=4.

At most two points among P1,P2,P3,P4P_{1},P_{2},P_{3},P_{4} coincide with points among P5,P6,P7,P8P_{5},P_{6},P_{7},P_{8} since Γ1Γ2=2\Gamma_{1}\cdot\Gamma_{2}=2. Thus after renumbering points PiP_{i} I may assume that

Γ1({P5,P6,P7,P8}{P1,P2,P3,P4})=and\displaystyle\Gamma_{1}\cap\Big{(}\{P_{5},P_{6},P_{7},P_{8}\}\setminus\{P_{1},P_{2},P_{3},P_{4}\}\Big{)}=\varnothing\quad\text{and}
Γ2({P1,P2,P3,P4}{P5,P6,P7,P8})=.\displaystyle\Gamma_{2}\cap\Big{(}\{P_{1},P_{2},P_{3},P_{4}\}\setminus\{P_{5},P_{6},P_{7},P_{8}\}\Big{)}=\varnothing.

Thus

8<i=14multPiλDF+i=58multPiλDF4α1+CΓ1+4α2+CΓ2=4+2α1+2α28,8<\sum_{i=1}^{4}\operatorname{mult}_{P_{i}}\lambda D_{F}+\sum_{i=5}^{8}\operatorname{mult}_{P_{i}}\lambda D_{F}\leqslant 4\alpha_{1}+C\cdot\Gamma_{1}+4\alpha_{2}+C\cdot\Gamma_{2}=4+2\alpha_{1}+2\alpha_{2}\leqslant 8,

a contradiction. ∎

3.3. Supermaximal singularities

To finish the proof of 𝒜5\mathcal{A}_{5}-equivarian birational superrigidity I use the technique of supermaximal singularities. It has been introduced in [18] for proving birational rigidity of del Pezzo fibrations of degrees 11, 22, and 33.

First, I require a stronger version of Noether-Fano inequality.

Proposition 3.12 ([12, Proposition 2.7]).

Let π:X1\pi:X\to\mathds{P}^{1} be a del Pezzo fibration. Suppose that we are given a non-square birational map f:XYf\colon X\dasharrow Y to a Mori fiber space πY:YZ\pi_{Y}\colon Y\to Z and let \mathcal{M} be a movable linear system associated to ff. Define numbers λ+\lambda\in\mathds{Q}_{+} and ll\in\mathds{Q} by the equivalence λ+KXlF\lambda\mathcal{M}+K_{X}\sim lF. Suppose in addition that the pair (X,λ)(X,\lambda\mathcal{M}) is canonical at curves on XX and l0l\geqslant 0. Then there exist points Q1,,QkQ_{1},\dots,Q_{k} of XX contained in distinct π\pi-fibers and positive rational numbers γ1,,γk\gamma_{1},\dots,\gamma_{k} with the following properties:

  • (X,λγjFj)(X,\lambda\mathcal{M}-\sum\gamma_{j}F_{j}) is not canonical at Q1,,QkQ_{1},\dots,Q_{k}, where FiF_{i} is the π\pi-fiber containing Γi\Gamma_{i}.

  • j=1kγj>l\sum_{j=1}^{k}\gamma_{j}>l.

Let D1,D2D_{1},D_{2} be general divisors in \mathcal{M} and denote their scheme theoretic intersection Z=D1D2Z=D_{1}\cap D_{2}. I may decompose ZZ into the horizontal and the vertical parts Z=Zh+ZvZ=Z^{h}+Z^{v}. I may further decompose

Zv=Ziv, where SuppZivFi.Z^{v}=\sum Z^{v}_{i},\text{ where }\operatorname{Supp}Z_{i}^{v}\subset F_{i}.
Lemma 3.13.

Suppose n2n\geqslant 2, then

λ2degZh\displaystyle\lambda^{2}\deg Z^{h} =8\displaystyle=8
λ2degZv\displaystyle\lambda^{2}\deg Z^{v} 8l+8\displaystyle\leqslant 8l+8
Proof.

I compute

λ2D1D2(KX+lF)2KX22lFKX.\lambda^{2}D_{1}D_{2}\equiv(-K_{X}+lF)^{2}\equiv K_{X}^{2}-2lF\cdot K_{X}.

By Corollary 3.2

2lFKX8lf-2lF\cdot K_{X}\equiv 8lf

and

KX28s+24f8nfK_{X}^{2}\equiv 8s+24f-8nf

These equivalences imply the statement of the lemma. ∎

Corollary 3.14.

There is a fiber FjF_{j} and a divisorial valuation EE_{\infty} of (X)\mathds{C}(X) such that degλ2Zjv8+8γi\deg\lambda^{2}Z^{v}_{j}\leqslant 8+8\gamma_{i}, a(E,X,λγjFj)<0a(E_{\infty},X,\lambda\mathcal{M}-\gamma_{j}F_{j})<0, and the center QjQ_{j} of EE_{\infty} is a point on FjF_{j}.

The valuation EE_{\infty} is called supermaximal singularity. By the previous section QjΔQ_{j}\in\Delta, where Δ\Delta is the 𝒜5\mathcal{A}_{5}-invariant curve of degree 22 on FjF_{j}. From now on I denote FjF_{j} by FF, γi\gamma_{i} by γ\gamma, QjQ_{j} by PP, and ZivZ^{v}_{i} by ZvZ_{v}.

3.4. Pukhlikov’s inequality

Consider the tower of blow ups realizing EE_{\infty}

(2) XNσNσ2X1σ1X0=X,X_{N}\xrightarrow{\sigma_{N}}\dots\xrightarrow{\sigma_{2}}X_{1}\xrightarrow{\sigma_{1}}X_{0}=X,

that is σi\sigma_{i} is the blow up of Xi1X_{i-1} at the center Bi1B_{i-1} of EE_{\infty} on Xi1X_{i-1}, EiE_{i} is the exceptional divisor of σi\sigma_{i}, and EN=EE_{N}=E_{\infty} as divisorial valuations of (X)\mathds{C}(X).

Let AA be an object on XiX_{i}, then I denote its proper transform on XjX_{j}, j>ij>i by A(j)A^{(j)}. Denote

K=min{iBi is not a point},\displaystyle K=\min\{i\mid B_{i}\text{ is not a point}\},
K=min{iBiF(i)}K,\displaystyle K^{\prime}=\min\{i\mid B_{i}\not\in F^{(i)}\}\cup{K},
K′′=min{iBiΔ(i)},\displaystyle K^{\prime\prime}=\min\{i\mid B_{i}\not\in\Delta^{(i)}\},

clearly KKK′′K\geqslant K^{\prime}\geqslant K^{\prime\prime}.

There is an oriented graph associated to (2)(\ref{tower}). It consists of NN vertices viv_{i}, and there is an edge vivjv_{i}\to v_{j} if i>ji>j and Bi1Ej(i1)B_{i-1}\subset E_{j}^{(i-1)}. Denote by pip_{i} the number of paths from vNv_{N} to viv_{i} and set pN=1p_{N}=1. Also set

Σ0=i=1Kpi,Σ0=i=1Kpi,Σ0′′=i=1K′′pi,Σ1=i=K+1Npi.\displaystyle\Sigma_{0}=\sum_{i=1}^{K}p_{i},\quad\Sigma_{0}^{\prime}=\sum_{i=1}^{K^{\prime}}p_{i},\quad\Sigma_{0}^{\prime\prime}=\sum_{i=1}^{K^{\prime\prime}}p_{i},\quad\Sigma_{1}=\sum_{i=K+1}^{N}p_{i}.

Denote νi=multBi1λ(i1)\nu_{i}=\operatorname{mult}_{B_{i-1}}\lambda\mathcal{M}^{(i-1)}, then non-canonity of the pair (X,λγF)(X,\lambda\mathcal{M}-\gamma F) is equivalent to

(3) i=1Npiνi>2Σ0+Σ1+γΣ0.\sum_{i=1}^{N}p_{i}\nu_{i}>2\Sigma_{0}+\Sigma_{1}+\gamma\Sigma_{0}^{\prime}.
Theorem 3.15 ([18, Proposition 4.2] Pukhlikov’s inequality).
λ2(i=1KpimultPi1Zh(i1)+i=1KpimultPi1Zv(i1))4Σ0+4γΣ0+(Σ1γΣ0)2Σ0+Σ1.\lambda^{2}\Big{(}\sum_{i=1}^{K}p_{i}\operatorname{mult}_{P_{i-1}}Z_{h}^{(i-1)}+\sum_{i=1}^{K^{\prime}}p_{i}\operatorname{mult}_{P_{i-1}}Z_{v}^{(i-1)}\Big{)}\geqslant 4\Sigma_{0}+4\gamma\Sigma_{0}^{\prime}+\frac{(\Sigma_{1}-\gamma\Sigma_{0}^{\prime})^{2}}{\Sigma_{0}+\Sigma_{1}}.

I apply this inequality for each valuation in the orbit of EE to show a contradiction.

Denote the valuations in the 𝒜5\mathcal{A}_{5}-orbit of EE_{\infty} by Ek,E_{k,\infty}, in particular E1,=E=ENE_{1,\infty}=E_{\infty}=E_{N} as valuations, and let θ\theta be the length of the orbit. Recall that a(Ek,,X,λγF)<0a(E_{k,_{\infty}},X,\lambda\mathcal{M}-\gamma F)<0 for k=1,,θk=1,\dots,\theta. Thus there is a tower (2) for each valuation k=1,,θk=1,\dots,\theta. The graphs for the towers are identical for each Ek,NE_{k,N}, thus the invariants KK, KK^{\prime}, K′′K^{\prime\prime}, NN, pip_{i}, Σ0\Sigma_{0}, Σ0\Sigma_{0}^{\prime}, Σ0′′\Sigma_{0}^{\prime\prime}, and Σ1\Sigma_{1} are the same for each Ek,NE_{k,N} as well.

From now on suppose that σi\sigma_{i} is the blow up of Xi1X_{i-1} at the centers of Ek,NE_{k,N}. Denote the center of Ek,NE_{k,N} on XiX_{i} by Bk,iB_{k,i} and the exceptional divisor of σi\sigma_{i} over Bk,i1B_{k,i-1} by Ek,iE_{k,i}. Note that \mathcal{M} is 𝒜5\mathcal{A}_{5}-invariant, therefore multBk,i(i)=multBj,i(i)=νi+1\operatorname{mult}_{B_{k,i}}\mathcal{M}^{(i)}=\operatorname{mult}_{B_{j,i}}\mathcal{M}^{(i)}=\nu_{i+1} for any 1j,kθ1\leqslant j,k\leqslant\theta, 0iN10\leqslant i\leqslant N-1. Applying the Pukhlikov’s inequality to Ek,NE_{k,N} and taking a sum I get the following inequality.

Corollary 3.16.
(4) λ2k=1θ(i=1KpimultBk,i1Zh(i1)+i=1KpimultBk,i1Zv(i1))θ(4Σ0+4γΣ0+(Σ1γΣ0)2Σ0+Σ1).\lambda^{2}\sum_{k=1}^{\theta}\Big{(}\sum_{i=1}^{K}p_{i}\operatorname{mult}_{B_{k,i-1}}Z_{h}^{(i-1)}+\sum_{i=1}^{K^{\prime}}p_{i}\operatorname{mult}_{B_{k,i-1}}Z_{v}^{(i-1)}\Big{)}\geqslant\theta\big{(}4\Sigma_{0}+4\gamma\Sigma_{0}^{\prime}+\frac{(\Sigma_{1}-\gamma\Sigma_{0}^{\prime})^{2}}{\Sigma_{0}+\Sigma_{1}}\big{)}.

To find a bound for the left-hand side, I decompose λ2Zv=ξΔ+C\lambda^{2}Z_{v}=\xi\Delta+C, where ΔSuppC\Delta\not\in\operatorname{Supp}C.

Lemma 3.17.

The following inequalities hold

  1. (1)
    λ2k=1θi=1KpimultBk,i1Zh(i1)8Σ0,\lambda^{2}\sum_{k=1}^{\theta}\sum_{i=1}^{K}p_{i}\operatorname{mult}_{B_{k,i-1}}Z_{h}^{(i-1)}\leqslant 8\Sigma_{0},
  2. (2)
    k=1θmultBk,0C8+8γ2ξ,\sum_{k=1}^{\theta}\operatorname{mult}_{B_{k,0}}C\leqslant 8+8\gamma-2\xi,

    in particular, ξ4+4γ\xi\leqslant 4+4\gamma.

Proof.

The inequality

λ2k=1θmultBk,0Zhλ2ZhF=8.\lambda^{2}\sum_{k=1}^{\theta}\operatorname{mult}_{B_{k,0}}Z_{h}\leqslant\lambda^{2}Z_{h}\cdot F=8.

implies (1).

Assertion (2) follows from

k=1θmultBk,0CCΔ=degC8+8γ2ξ.\sum_{k=1}^{\theta}\operatorname{mult}_{B_{k,0}}C\leqslant C\cdot\Delta=\deg C\leqslant 8+8\gamma-2\xi.

Proposition 3.18.

The following inequality holds:

8Σ0>θ(Σ1γΣ0)2Σ0+Σ1.8\Sigma_{0}>\theta\frac{(\Sigma_{1}-\gamma\Sigma_{0}^{\prime})^{2}}{\Sigma_{0}+\Sigma_{1}}.
Proof.

Combining Lemma 3.17 with (4) I get

(5) 8Σ0+θξΣ0′′+(8+8γ2ξ)Σ0>θ(4Σ0+4γΣ0+(Σ1γΣ0)2Σ0+Σ1).8\Sigma_{0}+\theta\xi\Sigma_{0}^{\prime\prime}+(8+8\gamma-2\xi)\Sigma_{0}^{\prime}>\theta\Big{(}4\Sigma_{0}+4\gamma\Sigma_{0}^{\prime}+\frac{(\Sigma_{1}-\gamma\Sigma_{0}^{\prime})^{2}}{\Sigma_{0}+\Sigma_{1}}\Big{)}.

Observe that

θΣ0′′>2Σ0.\theta\Sigma_{0}^{\prime\prime}>2\Sigma_{0}^{\prime}.

Indeed, otherwise (5) becomes

8Σ0+(8+8γ)Σ0>4θΣ0+4θγΣ048Σ0+48γΣ0,8\Sigma_{0}+(8+8\gamma)\Sigma_{0}^{\prime}>4\theta\Sigma_{0}+4\theta\gamma\Sigma_{0}^{\prime}\geqslant 48\Sigma_{0}+48\gamma\Sigma_{0}^{\prime},

a contradiction.

I apply the bound ξ4+4γ\xi\leqslant 4+4\gamma to (5) to get the inequality

8Σ0+(4+4γ)θΣ0′′>4θΣ0+4θγΣ0+θ(Σ1γΣ0)2Σ0+Σ1.8\Sigma_{0}+(4+4\gamma)\theta\Sigma_{0}^{\prime\prime}>4\theta\Sigma_{0}+4\theta\gamma\Sigma_{0}^{\prime}+\theta\frac{(\Sigma_{1}-\gamma\Sigma_{0}^{\prime})^{2}}{\Sigma_{0}+\Sigma_{1}}.

Since Σ0Σ0′′\Sigma_{0}^{\prime}\geqslant\Sigma_{0}^{\prime\prime} I can further simplify

8Σ0+4θΣ0′′>4θΣ0+θ(Σ1γΣ0)2Σ0+Σ1,8\Sigma_{0}+4\theta\Sigma_{0}^{\prime\prime}>4\theta\Sigma_{0}+\theta\frac{(\Sigma_{1}-\gamma\Sigma_{0}^{\prime})^{2}}{\Sigma_{0}+\Sigma_{1}},

which implies the statement of the proposition. ∎

I use a different technique to get a bound contradicting this one.

3.5. The technique of restricted multiplicities

The technique I am using is inspired by [8, Proof of Theorem 3.1].

Consider the linear system 𝒮|γH+F|\mathcal{S}\subset\lvert\gamma H+F\rvert of divisors containing Δ\Delta. Let SS be general in 𝒮\mathcal{S} and let DD be general in \mathcal{M}. Denote DS=D|SD_{S}=D|_{S} and multΓD=α\operatorname{mult}_{\Gamma}D=\alpha, then ordΓDS=α\operatorname{ord}_{\Gamma}D_{S}=\alpha. Indeed, any S𝒮S\in\mathcal{S} has the equation a(u,v)w+up(x,y,z,w)=a(u,v)w+up(x,y,z,w)= for some linear pp and aa. Thus ordΓS1S2=1\operatorname{ord}_{\Gamma}S_{1}\cap S_{2}=1 for general S1,S2𝒮S_{1},S_{2}\in\mathcal{S}. It follows that DS=αΓ+DSD_{S}=\alpha\Gamma+D_{S}^{\prime}, where ΓSuppDS\Gamma\not\subset\operatorname{Supp}D_{S}^{\prime}.

Lemma 3.19.

Let BB be a point or a smooth curve on a smooth threefold XX. Let SS be a surface on XX, suppose B~=BS\widetilde{B}=B\cap S is a point and suppose SS is smooth at B~\widetilde{B}. Let σ:YX\sigma\colon Y\to X be the blow up at BB, let EE be its exceptional divisor, let SY=σ1SS_{Y}=\sigma^{-1}S and let e=ESYe=E\cap S_{Y}. Let DD be an effective divisor on XX and let DY=σ1DD_{Y}=\sigma^{-1}D, then

multB~D|S=multBD+ordeDY|SY.\operatorname{mult}_{\widetilde{B}}D|_{S}=\operatorname{mult}_{B}D+\operatorname{ord}_{e}D_{Y}|_{S_{Y}}.
Proof.

Elementary calculations in local coordinates. ∎

Note that multBk,i1DS(i1)=multBj,i1DS(i1)\operatorname{mult}_{B_{k,i-1}}D_{S}^{(i-1)}=\operatorname{mult}_{B_{j,i-1}}D_{S}^{(i-1)} for general DD and SS and 1j,kθ1\leqslant j,k\leqslant\theta. Denote ν~i=λmultBk,i1DS(i1)\widetilde{\nu}_{i}=\lambda\operatorname{mult}_{B_{k,i-1}}D_{S}^{(i-1)}, then one can bound νi\nu_{i} using ν~i\widetilde{\nu}_{i}.

Lemma 3.20.

For any iK′′i\leqslant K^{\prime\prime} we have

ν1++νiν~1++ν~i.\nu_{1}+\dots+\nu_{i}\leqslant\widetilde{\nu}_{1}+\dots+\widetilde{\nu}_{i}.
Proof.

Let ek,i=S(i)Ek,ie_{k,i}=S^{(i)}\cap E_{k,i} and denote mi=ordek,iλD(i)|S(i)m_{i}=\operatorname{ord}_{e_{k,i}}\lambda D^{(i)}|_{S^{(i)}}, then

λD(i)|S(i)=λDS(i)+m1ek,1(i)++mi1ek,i1(i)+miek,i.\lambda D^{(i)}|_{S^{(i)}}=\lambda D_{S}^{(i)}+m_{1}e_{k,1}^{(i)}+\dots+m_{i-1}e_{k,i-1}^{(i)}+m_{i}e_{k,i}.

Observe that Bk,iEk,i1(i)B_{k,i}\not\in E_{k,i-1}^{(i)} for i<K′′i<K^{\prime\prime} since Δ\Delta is smooth. In particular Bk,iek,j(i)B_{k,i}\not\in e_{k,j}^{(i)} for j<ij<i, thus

multBiλD(i)|S(i)=ν~i+1+mifori1.\operatorname{mult}_{B_{i}}\lambda D^{(i)}|_{S^{(i)}}=\widetilde{\nu}_{i+1}+m_{i}\quad\text{for}~{}i\geqslant 1.

On the other hand by Lemma 3.19

multBiλD(i)|S(i)=νi+1+mi+1.\operatorname{mult}_{B_{i}}\lambda D^{(i)}|_{S^{(i)}}=\nu_{i+1}+m_{i+1}.

Thus I get the equalities

ν~1=ν1+m1\displaystyle\widetilde{\nu}_{1}=\nu_{1}+m_{1}
ν~2+m1=ν2+m2\displaystyle\widetilde{\nu}_{2}+m_{1}=\nu_{2}+m_{2}
\displaystyle\quad\quad\quad\dots\dots
ν~K′′+mK′′1=νK′′+mK′′,\displaystyle\widetilde{\nu}_{K^{\prime\prime}}+m_{K^{\prime\prime}-1}=\nu_{K^{\prime\prime}}+m_{K^{\prime\prime}},

which together imply the statement of the lemma. ∎

Proposition 3.21.

There is a bound

ν1++νK′′K′′+4θ,\nu_{1}+\dots+\nu_{K^{\prime\prime}}\leqslant K^{\prime\prime}+\frac{4}{\theta},

in particular

(6) p1ν1++pNνN(Σ0+Σ1)(1+4θK′′),p_{1}\nu_{1}+\dots+p_{N}\nu_{N}\leqslant(\Sigma_{0}+\Sigma_{1})(1+\frac{4}{\theta K^{\prime\prime}}),
Proof.

The curve Γ\Gamma is smooth at PiP_{i}, thus

pj=p1andν~j=λα+λmultB~i,j1(DS)(j1)forjK′′\displaystyle p_{j}=p_{1}\quad\text{and}\quad\widetilde{\nu}_{j}=\lambda\alpha+\lambda\operatorname{mult}_{\widetilde{B}_{i,j-1}}(D_{S}^{\prime})^{(j-1)}\quad\text{for}~{}j\leqslant K^{\prime\prime}

On the other hand SF=ΓS\cap F=\Gamma, therefore

λi=12γj=1LmultB~i,j(DS)(j)λDSF=λDFS=4.\lambda\sum_{i=1}^{2\gamma}\sum_{j=1}^{L}\operatorname{mult}_{\widetilde{B}_{i,j}}(D_{S}^{\prime})^{(j)}\leqslant\lambda D_{S}\cdot F=\lambda D\cdot F\cdot S=4.

Recall that \mathcal{M} is canonical at Γ\Gamma by Proposition 3.9, that is λα1\lambda\alpha\leqslant 1. Putting the bounds together I get the first inequality. It follows that νi1+4θK′′\nu_{i}\leqslant 1+\frac{4}{\theta K^{\prime\prime}} for i>K′′i>K^{\prime\prime} which implies the second inequality in the lemma. ∎

Corollary 3.22.

Let PP be a point which is not 𝒜5\mathcal{A}_{5}-fixed. Then there is no supermaximal singularity at PP.

Proof.

I combine (6) with (3) to get a lower bound on Σ1\Sigma_{1}:

γΣ0+2Σ0+Σ1<(Σ0+Σ1)(1+4θK′′),\gamma\Sigma_{0}^{\prime}+2\Sigma_{0}+\Sigma_{1}<(\Sigma_{0}+\Sigma_{1})(1+\frac{4}{\theta K^{\prime\prime}}),

equivalently

θK′′γΣ0+(θ4)K′′Σ0<4Σ1.\theta K^{\prime\prime}\gamma\Sigma_{0}+(\theta-4)K^{\prime\prime}\Sigma_{0}<4\Sigma_{1}.

In particular, since K′′1K^{\prime\prime}\geqslant 1 and θ12\theta\geqslant 12 I get

Σ1>2Σ0+3γΣ0.\Sigma_{1}>2\Sigma_{0}+3\gamma\Sigma_{0}^{\prime}.

Applying this bound to Proposition 3.18 I get

8Σ04θ(Σ0Σ′′)>124Σ02+8γΣ0Σ03Σ0+3γΣ0>8Σ0,8\Sigma_{0}-4\theta(\Sigma_{0}-\Sigma^{\prime\prime})>12\frac{4\Sigma_{0}^{2}+8\gamma\Sigma_{0}\Sigma_{0}^{\prime}}{3\Sigma_{0}+3\gamma\Sigma_{0}^{\prime}}>8\Sigma_{0},

a contradiction. ∎

At last I am ready to prove Theorem 1.11.

Proof of Theorem 1.11.

Let χ:XnY\chi\colon X_{n}\dasharrow Y be an 𝒜5\mathcal{A}_{5}-equivariant birational map to a 𝒜5\mathcal{A}_{5}\mathds{Q}-Mori fiber space Y/ZY/Z and let \mathcal{M} be the mobile 𝒜5\mathcal{A}_{5}-invariant linear system associated χ\chi. Define the numbers λ\lambda and ll by the equivalence λ+KXnlF\lambda\mathcal{M}+K_{X_{n}}\sim lF. By Lemma 3.9 there is another birational model πb:Xb1\pi_{b}\colon X_{b}\to\mathds{P}^{1} and an 𝒜5\mathcal{A}_{5}-equivariant square birational map φb:XnXb\varphi_{b}\colon X_{n}\dasharrow X_{b} such that the corresponding pair (Xb,λb)(X_{b},\lambda\mathcal{M}_{b}) is canonical at curves and 𝒜5\mathcal{A}_{5}-fixed points. By Corollary 3.22 the pair (Xb,λb)(X_{b},\lambda\mathcal{M}_{b}) is also canonical at points not fixed by 𝒜5\mathcal{A}_{5}-action.

By Proposition 3.12 and Corollary 3.14 if the map χφb1:XbY\chi\circ\varphi_{b}^{-1}\colon X_{b}\to Y is not 𝒜5\mathcal{A}_{5}-equivariantly square, then there is a supermaximal singularity EE_{\infty}. But by Corollary 3.22 there are no supermaximal singularities. It follows that χφb1\chi\circ\varphi_{b}^{-1} is 𝒜5\mathcal{A}_{5}-equivariantly square and hence χ\chi is 𝒜5\mathcal{A}_{5}-equivariantly square. ∎

4. The 𝒜5\mathcal{A}_{5}-equivariant birational geometry of X1X_{1}

In this section I describe some 𝒜5\mathcal{A}_{5}-Mori fiber spaces 𝒜5\mathcal{A}_{5}-equivariantly birational to the variety X1X_{1}.

Let Q(W3II)Q\subset\mathds{P}(W_{3}\oplus I\oplus I) be a smooth 𝒜5\mathcal{A}_{5}-invariant quadric. Let x,y,zx,y,z be the coordinates on W3W_{3} and u,vu,v be the coordinates on III\oplus I. Let Γ\Gamma be the 𝒜5\mathcal{A}_{5}-invariant point curve of degree 22, then it has equations (xzy2=u=v=0)(xz-y^{2}=u=v=0) after a change of coordinates on W3W_{3}. Thus the blow up of QQ at Γ\Gamma is σΓ:X1Q\sigma_{\Gamma}\colon X_{1}\to Q. On the other hand QQ is a quadric with an 𝒜5\mathcal{A}_{5}-invariant point, hence it is 𝒜5\mathcal{A}_{5}-equivariantly birational to (W3I)\mathds{P}(W_{3}\oplus I).

Consider the blow up at the 𝒜5\mathcal{A}_{5}-invariant point σY:Y1(W3I)\sigma_{Y}\colon Y_{1}\to\mathds{P}(W_{3}\oplus I), where

Y1=(𝒪(W3)𝒪(W3)(1)).Y_{1}=\mathds{P}(\mathcal{O}_{\mathds{P}(W_{3})}\oplus\mathcal{O}_{\mathds{P}(W_{3})}(-1)).

The 1\mathds{P}^{1}-bundle τ:Y1(W3)\tau\colon Y_{1}\to\mathds{P}(W_{3}) has many 𝒜5\mathcal{A}_{5}-equivariantly square birational to it models. For example, an 𝒜5\mathcal{A}_{5}-invariant conic on the exceptional divisor of σY\sigma_{Y} induces the 𝒜5\mathcal{A}_{5}-equivariant elementary Sarkisov link Y1Y3Y_{1}\dasharrow Y_{3}, where

Y3=(𝒪(W3)𝒪(W3)(3)).Y_{3}=\mathds{P}(\mathcal{O}_{\mathds{P}(W_{3})}\oplus\mathcal{O}_{\mathds{P}(W_{3})}(-3)).

Similarly, Y1Y_{1} is 𝒜5\mathcal{A}_{5}-equivariantly birational to Y2k+1Y_{2k+1} for any kk. Alternatively, we can take a fiber ff of τ\tau and the blow up σ:Y~Y1\sigma\colon\widetilde{Y}\to Y_{1} at the orbit of ff, it is easy to see that Y~\widetilde{Y} admits an 𝒜5\mathcal{A}_{5}-equivariant 1\mathds{P}^{1}-bundle.

As we can see, X1X_{1} has a rich 𝒜5\mathcal{A}_{5}-equivariant birational geometry with the following 𝒜5\mathcal{A}_{5}-Mori fiber spaces structures:

  1. (1)

    𝒜5\mathcal{A}_{5}-Fano variety Q(W3II)Q\in\mathds{P}(W_{3}\oplus I\oplus I),

  2. (2)

    𝒜5\mathcal{A}_{5}-Fano variety (W3I)\mathds{P}(W_{3}\oplus I),

  3. (3)

    𝒜5\mathcal{A}_{5}-del Pezzo fibration π:X11\pi\colon X_{1}\to\mathds{P}^{1},

  4. (4)

    𝒜5\mathcal{A}_{5}-conic bundle τ:Y1(W3)\tau\colon Y_{1}\to\mathds{P}(W_{3}).

I believe that these are the only 𝒜5\mathcal{A}_{5}-Mori fiber space structures.

Suppose χ:QY\chi\colon Q\to Y is a birational 𝒜5\mathcal{A}_{5}-equivariant map to a 𝒜5\mathcal{A}_{5}\mathds{Q}-Mori fiber space πY:YZ\pi_{Y}\colon Y\to Z. Let Q\mathcal{M}_{Q} be the associated mobile linear system. If (Q,λQQ)(Q,\lambda_{Q}\mathcal{M}_{Q}) is not canonical at the 𝒜5\mathcal{A}_{5}-invariant conic, then the map χ\chi factors through X1X_{1}. Let \mathcal{M} be the corresponding mobile linear system on X1X_{1}. Elementary calculations show that λKX1+lF\lambda\mathcal{M}\sim-K_{X_{1}}+lF, where l>0l>0. I believe that the results of Section 3 can be refined to show that Y/ZY/Z is 𝒜5\mathcal{A}_{5}-equivariantly square birational to X1/1X_{1}/\mathds{P}^{1}.

Question 4.1.

Does there exist an 𝒜5\mathcal{A}_{5}\mathds{Q}-Mori fiber space 𝒜5\mathcal{A}_{5}-equivariantly birational to QQ which is not 𝒜5\mathcal{A}_{5}-equivariantly square birational to X1/1X_{1}/\mathds{P}^{1}, Y1/(W3)Y_{1}/\mathds{P}(W_{3}), QQ, or (W3I)\mathds{P}(W_{3}\oplus I)?

References

  • [1] Hamid Ahmadinezhad, Ivan Cheltsov, Jihun Park, and Constantin Shramov, Double veronese cones with 28 nodes, 2019.
  • [2] I. Cheltsov, On singular cubic surfaces, Asian J. Math. 13 (2009), no. 2, 191–214. MR 2559108
  • [3] Ivan Cheltsov and Constantin Shramov, Five embeddings of one simple group, Trans. Amer. Math. Soc. 366 (2014), no. 3, 1289–1331. MR 3145732
  • [4] by same author, Cremona groups and the icosahedron, Monographs and Research Notes in Mathematics, CRC Press, Boca Raton, FL, 2016. MR 3444095
  • [5] A. Corti, Singularities of linear systems and 33-fold birational geometry, Explicit birational geometry of 3-folds, London Math. Soc. Lecture Note Ser., vol. 281, Cambridge Univ. Press, Cambridge, 2000, pp. 259–312.
  • [6] Alessio Corti, Factoring birational maps of threefolds after Sarkisov, J. Algebraic Geom. 4 (1995), no. 2, 223–254. MR 1311348
  • [7] Igor V. Dolgachev and Vasily A. Iskovskikh, Finite subgroups of the plane Cremona group, Algebra, arithmetic, and geometry: in honor of Yu. I. Manin. Vol. I, Progr. Math., vol. 269, Birkhäuser Boston, Boston, MA, 2009, pp. 443–548. MR 2641179
  • [8] M. M. Grinenko, On the double cone over the Veronese surface, Izv. Ross. Akad. Nauk Ser. Mat. 67 (2003), no. 3, 5–22. MR 1992191
  • [9] Masayuki Kawakita, Divisorial contractions in dimension three which contract divisors to compound A1A_{1} points, Compositio Math. 133 (2002), no. 1, 95–116. MR 1918291
  • [10] János Kollár, Singularities of pairs, Algebraic geometry—Santa Cruz 1995, Proc. Sympos. Pure Math., vol. 62, Amer. Math. Soc., Providence, RI, 1997, pp. 221–287. MR 1492525
  • [11] Igor Krylov, Birational geometry of del Pezzo fibrations with terminal quotient singularities, J. Lond. Math. Soc. (2) 97 (2018), no. 2, 222–246. MR 3789845
  • [12] Takuzo Okada, On birational rigidity of singular del pezzo fibrations of degree 1, (2018).
  • [13] Yuri Prokhorov, Simple finite subgroups of the Cremona group of rank 3, J. Algebraic Geom. 21 (2012), no. 3, 563–600. MR 2914804
  • [14] by same author, GG-Fano threefolds, I, Adv. Geom. 13 (2013), no. 3, 389–418. MR 3100917
  • [15] by same author, On stable conjugacy of finite subgroups of the plane cremona group, ii, The Michigan Mathematical Journal 64 (2015), no. 2, 293–318.
  • [16] Yuri Prokhorov and Constantin Shramov, Jordan property for Cremona groups, Amer. J. Math. 138 (2016), no. 2, 403–418. MR 3483470
  • [17] by same author, pp-subgroups in the space Cremona group, Math. Nachr. 291 (2018), no. 8-9, 1374–1389. MR 3817323
  • [18] A. Pukhlikov, Birational automorphisms of three-dimensional algebraic varieties with a pencil of del Pezzo surfaces, Izv. Ross. Akad. Nauk Ser. Mat. 62 (1998), no. 1, 123–164.