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

Factorization centers in dimension two and the Grothendieck ring of varieties

Hsueh-Yung Lin, Evgeny Shinder, Susanna Zimmermann Department of Mathematics, National Taiwan University, and National Center for Theoretical Sciences, Taipei, Taiwan. [email protected] School of Mathematics and Statistics, University of Sheffield, Hounsfield Road, S3 7RH, UK, and Hausdorff Center for Mathematics at the University of Bonn, Endenicher Allee 60, 53115. [email protected] Laboratoire angevin de recherche en mathématiques (LAREMA), CNRS, Université d’Angers, 49045 Angers cedex 1, France [email protected]
Abstract.

We initiate the study of factorization centers of birational maps, and complete it for surfaces over a perfect field in this article. We prove that for every birational automorphism ϕ:XX\phi:X\dashrightarrow X of a smooth projective surface XX over a perfect field kk, the blowup centers are isomorphic to the blowdown centers in every weak factorization of ϕ\phi. This implies that nontrivial L-equivalences of 0-dimensional varieties cannot be constructed based on birational automorphisms of a surface. It also implies that rationality centers are well-defined for every rational surface XX, namely there exists a 0-dimensional variety intrinsic to XX, which is blown up in any rationality construction of XX.

Key words and phrases:
Grothendieck ring, Birational automorphism, Algebraic Surface, Factorization center, Sarkisov link
2010 Mathematics Subject Classification:
14E07, 14E30, 14J26 14F20

1. Introduction

One source of motivation in birational geometry comes from studying groups of birational automorphisms of algebraic varieties, in particular the Cremona groups Crn(k)=Bir(kn){\operatorname{Cr}}_{n}(k)={\operatorname{Bir}}(\mathbb{P}^{n}_{k}). Each birational automorphism blows up some subschemes and contracts some exceptional divisors. The primary question we study in this paper is:

Question 1.1.

Let ϕ:XX\phi:X\dashrightarrow X be a birational automorphism of a smooth projective variety. Do centers blown up by ϕ\phi correspond, up to stable birational equivalence, to the exceptional divisors blown down by ϕ\phi?

We note that exceptional divisors are ruled over the corresponding blow up centers, so asking about stable birational equivalence classes is a natural way to compare exceptional divisors with the blow up centers. We give a complete answer to a stronger version of Question 1.1 for surfaces over an arbitrary perfect field, see Theorem 1.3 below. Our proof is an application of the two-dimensional Minimal Model Program [Isk79, Isk96] combined with the Grothendieck ring of varieties, étale cohomology groups and the so-called Gassmann equivalence of Galois sets.

There are three main reasons to study Question 1.1. First of all, it is the structure of birational automorphisms and Cremona groups, in particular their generation by involutions or regularizable elements. In the recent paper [LS22], we explain that the answer to Question 1.1 is negative in various contexts in dimension n3n\geq 3, and give applications to the structure of the higher Cremona groups. Secondly, Question 1.1 has a tight relationship to the structure of the Grothendieck ring of varieties.

Question 1.2.
  1. (a)

    (Larsen-Lunts [LL03], slightly reformulated) If classes of smooth projective varieties [X][X] and [Y][Y] coincide in the Grothendieck ring, how can we compare the geometry of XX and YY? For instance, are XX and YY birational?

  2. (b)

    [KS18] What is the geometric meaning of L-equivalence? For instance, if zero-dimensional schemes are L-equivalent, do they have to be isomorphic?

  3. (c)

    [(ht] Does the Grothendieck ring of varieties K0(Var/k)K_{0}({\mathrm{Var}}/k) have torsion elements?

In the direction of (a), the main result of [LL03], which also follows from [Bit04], is that if char(k)=0{\operatorname{char}}(k)=0, equality of classes [X][X], [Y][Y] of smooth projective connected varieties implies that XX and YY are stably birational. On the other hand, for non-projective smooth connected varieties, the second part of Question (a).(a) is known to have a negative answer (the first such example is [Bor18, proof of Theorem 2.13]). For (b), see [KS18] for conjectural relations to derived equivalence. In §2.2, we explain that L-equivalence of smooth zero-dimensional schemes implies Gassmann equivalence of the corresponding Galois sets, but this does not rule out the possibility of nontrivial L-equivalence between such schemes. Nothing is known about (c). As the Grothendieck ring is a colimit of the truncated groups K0(Varn/k)K_{0}({\mathrm{Var}}^{\leq n}/k) generated by varieties of dimension up to nn, we can ask each of the questions in these truncated groups. As a direct consequence of our positive answer to Question 1.1 in dimension two, we are able to answer Question 1.2 completely for K0(Var2/k)K_{0}({\mathrm{Var}}^{\leq 2}/k); see Corollary 3.11. Namely, (a) equality of classes of smooth projective varieties in K0(Var2/k)K_{0}({\mathrm{Var}}^{\leq 2}/k) implies birationality, (b) L-equivalence in K0(Var2/k)K_{0}({\mathrm{Var}}^{\leq 2}/k) is trivial and (c) K0(Var2/k)K_{0}({\mathrm{Var}}^{\leq 2}/k) is a free abelian group. We expect that studying Question 1.1 in dimensions n\leq n would lead to good control over Question 1.2 for K0(Varn/k)K_{0}({\mathrm{Var}}^{\leq n}/k).

Finally, answering Question 1.1 positively, or explaining all ways in which it can fail, allows to control all rationality constructions for every rational variety, see the discussion of rationality centers below and §5.2 for more details.

We now explain our answer to Question 1.1 in dimension 22 over perfect fields. Our main result can be stated in the following way.

Theorem 1.3 (see Theorem 3.5).

Let kk be a perfect field. Let X/kX/k be a smooth projective surface and let ϕBir(X)\phi\in{\operatorname{Bir}}(X) be a birational automorphism. For any factorization of ϕ\phi into a sequence of blow ups and blow downs at connected smooth zero-dimensional subschemes, let Z1,,ZrZ_{1},\dots,Z_{r} (resp. Z1,,ZrZ_{1}^{\prime},\dots,Z_{r^{\prime}}^{\prime}) be the centers which get blown up (resp. blown down). Then r=rr=r^{\prime} and there is a reordering under which ZiZiZ_{i}\simeq Z_{i}^{\prime} over kk for all i=1,,ri=1,\dots,r.

Note that Theorem 1.3 is easily seen to hold when kk is algebraically closed field, as the Galois actions are trivial and the number of points blown up is equal to the number of exceptional divisors contracted. Over an arbitrary perfect field, Theorem 1.3 is a non-trivial statement. In particular, by this we mean that it is not possible to recover centers of the blow ups simply from the Galois action on the cohomology of the surface; see Remark 2.11 for the technical formulation of this statement in terms of Chow motives, and Example 2.14 for an explicit construction. Neither there seems to exist a straightforward geometric argument: see Example 3.8 for an illustration of the birational geometry involved.

To put Theorem 1.3 into an appropriate context, we introduce a general invariant c(ϕ)c(\phi), keeping track of the factorization centers of the birational map ϕ\phi, which is a homomorphism from the groupoid of birational types of surfaces to a free abelian group generated by reduced kk-schemes of dimension 0; see Corollary 3.2 for an axiomatic definition of c(ϕc(\phi). We then show that cc is constant on each Bir(X,Y){\operatorname{Bir}}(X,Y) and as a consequence c(ϕ)=0c(\phi)=0 for every self-map, which implies Theorem 1.3.

To prove that cc is constant on Bir(X,Y){\operatorname{Bir}}(X,Y), we have to consider each birational type of surfaces that can occur, with geometrically rational (and especially, rational) surfaces being the most interesting ones. For geometrically rational surfaces, by the two-dimensional minimal model program we have to consider birational maps between del Pezzo surfaces and conic bundles. Our proof for uniqueness of factorization centers uses two ingredients: Sarkisov links and Gassmann equivalence. Sarkisov links are certain elementary birational transformations between del Pezzo surfaces and conic bundles which generate the groupoid of birational maps between geometrically rational surfaces. In dimension two, the existence of decomposition into links has been proved and all the links have been classified by Iskovskikh into a finite list [Isk96]. The largest variety of links occurs for what we call models of large degree: these include all minimal geometrically rational surfaces with KX25K_{X}^{2}\geq 5.

We derive Theorem 1.3 for geometrically rational surfaces from a uniform claim we make for all links between minimal geometrically rational surfaces (see Proposition 5.5), which we check for each link in Iskovskikh’s classification. The latter uniform claim is formulated in terms of the (virtual) Néron-Severi Galois set, which is closely related to the Néron-Severi lattice of the surface as a Galois module. These Néron-Severi sets are defined in terms of the Galois action on linear systems of rational curves of low degree (typically pencils of conics and nets of twisted cubics) on del Pezzo surfaces.

On top of the classification of links, to prove the result we use étale cohomology, permutation modules, and the so-called Gassmann triples from group theory: these are triples (G,H,H)(G,H,H^{\prime}) with HH and HH^{\prime} subgroups of finite group GG such that [G/H]\mathbb{C}[G/H] and [G/H]\mathbb{C}[G/H^{\prime}] are isomorphic GG-representations (this holds if HH, HH^{\prime} are conjugate, but the converse if false). Gassmann triples are used to produce arithmetically equivalent fields, see e.g. [Per77, BdS02], isospectral manifolds [Sun85], as well as curves with isomorphic Jacobians [Pra17]. In our dealing with Gassmann equivalence, we follow the approach of [Par13] which generalizes Gassmann triples from a pair of subgroups of GG to a pair of GG-sets with possibly non-transitive actions. Our Gassmann equivalent sets come from étale cohomology groups and in particular, from Néron-Severi groups of geometrically rational surfaces with their Galois group action. We show that Gassmann equivalence provides a cohomological expression of L-equivalence, see Lemma 2.10 and Remark 2.11. Using the fact that Gassmann triples of small order are trivial allows us to significantly limit the number of Sarkisov links we have to consider in the proof of the main result.

As a by-product of our results on Gassmann equivalence we obtain conditions which forbid non-trivial L-equivalence of zero-dimensional schemes to exist; for example L-equivalent zero-dimensional connected reduced schemes (resp. reduced schemes) of degree 6\leq 6 (resp. 5\leq 5) are isomorphic and L-equivalence always implies isomorphism for fields with procyclic Galois groups such as k=k=\mathbb{R}, k=𝔽qk=\mathbb{F}_{q}, see Example 2.12.

Finally for rational surfaces, our main result has the following consequence: if XX is a smooth projective rational surface, then there exists a zero-dimensional scheme depending only on XX, which will have to be blown up by any birational isomorphism ϕ:2X\phi:\mathbb{P}^{2}\dashrightarrow X, see Corollary 5.9; we think of these associated schemes as rationality centers of XX. This is in contrast with the higher-dimensional geometry, where the associated rationality centers are not well-defined, even up to stable birational equivalence. For instance, a K3 surface associated to a cubic fourfolds should not be unique up to isomorphism: it should be unique up to derived equivalence [Has16, Remark 27], or possibly up to L-equivalence, as hinted in [KS18, (2.6.1)].

This text is organized as follows. §2 is devoted to Gassmann equivalence of GG-sets. We relate it to L-equivalence and provide sufficient conditions for two Gassmann equivalent GG-sets to be isomorphic. In §3, we define the invariant c(ϕ)c(\phi) which captures the factorization center of a birational map ϕ\phi and formulate the main theorem (Theorem 3.5) of the paper, which we prove in §4 and §5. We also study rational curves on del Pezzo surfaces in §4, which will be used to define and study the virtual Néron-Severi sets in §5. At the end of the paper we explain the concept of rationality center for rational surfaces, and illustrate it in the case of del Pezzo surfaces.

Acknowledgements

H.Y.L. was supported by the World Premier International Research centre Initiative (WPI), MEXT, Japan, then by the Ministry of Education Yushan Young Scholar Fellowship (NTU-110VV006) and the National Science and Technology Council (110-2628-M-002-006-). E.S. is supported by EP/T019379/1 “Derived categories and algebraic K-theory of singularities”, and by the ERC Synergy grant “Modern Aspects of Geometry: Categories, Cycles and Cohomology of Hyperkähler Varieties”. S.Z. was supported by the ANR Project FIBALGA ANR-18-CE40-0003-01 and the Project Étoiles montantes of the Région Pays de la Loire. E.S. would like to thank Yujiro Kawamata, Keiji Oguiso, Atsushi Takahashi and Shinnosuke Okawa for supporting his visit to Japan where some of this work has originated. E.S. would also like to thank Ivan Cheltsov, Sergey Galkin, Alexander Kuznetsov, Yuri Prokhorov, Constantin Shramov for discussions and encouragement, and Jean-Louis Colliot-Thélène, Brendan Hassett and Claire Voisin for their comments on this work. H.Y.L. would like to thank the NCTS in Taipei for the hospitality and support during the preparation of this paper. S.Z. would like to thank Jean-Louis Colliot-Thélène for his comments on this work and Jean-Louis Colliot-Thélène, Alexander Merkurjev and Jean-Pierre Serre for indicating references to examples of integral Gassmann triples. The authors would like to thank Keiji Oguiso for the initial discussion he involved on this topic and for making this collaboration possible.

Conventions

We work over a perfect field kk unless otherwise specified. All schemes are of finite type over kk. By a surface we mean a connected smooth projective (but not necessarily geometrically irreducible) surface XX over kk.

2. Gassmann equivalence

We explain the basics about the Burnside ring of a profinite group GG, the Grothendieck ring of varieties and a homomorphism between them when GG is the absolute Galois group. Our presentation is similar to [R1̈1]111 Note however that [R1̈1] makes an erroneous statement on p. 943 that the homomorphism from the Burnside ring to the representation ring is injective on transitive GG-sets (cf. Example 2.7(2)).. This homomorphism allows us to relate L-equivalence of reduced zero-dimensional schemes to Gassmann equivalence of GG-sets, and thus to rule out the possibility of nontrivial L-equivalence in small degree.

2.1. Definition and basic properties

Let GG be a profinite group and let 𝐆𝐬𝐞𝐭\mathbf{Gset} be the semi-ring of isomorphism classes of finite GG-sets on which GG acts continuously for the profinite topology, where finite sets are considered with discrete topology. Continuity of the action is equivalent to the requirement that stabilizer of any point is open (and in particular, a finite index subgroup). Here in 𝐆𝐬𝐞𝐭\mathbf{Gset}, the addition (resp. multiplication) is defined by disjoint unions (resp. Cartesian products). We define the Burnside ring Burn(G){\operatorname{Burn}}(G) of GG to be the Grothendieck ring associated to 𝐆𝐬𝐞𝐭\mathbf{Gset}. When GG is a finite group, the definition of Burn(G){\operatorname{Burn}}(G) is the classical one. We sometimes refer to elements of Burn(G){\operatorname{Burn}}(G), that is, combinations of isomorphism classes of GG-sets with integer coefficients, as virtual GG-sets. We note that the number of elements (resp. the number of orbits) in a GG-set gives rise to a ring homomorphism (resp. a group homomorphism) Burn(G){\operatorname{Burn}}(G)\to\mathbb{Z}.

Let FF be a field of characteristic zero and let Rep(G,F){\operatorname{Rep}}(G,F) be the abelian monoidal category of finite dimensional GG-representations over FF. Let K0(Rep(G,F))K_{0}({\operatorname{Rep}}(G,F)) be the Grothendieck ring of Rep(G,F){\operatorname{Rep}}(G,F). There is a well-defined ring homomorphism

(2.1) μG:Burn(G)K0(Rep(G,F))\mu_{G}:{\operatorname{Burn}}(G)\to K_{0}({\operatorname{Rep}}(G,F))

which sends the class of a continuous finite GG-set AA to the class of the permutation representation F[A]F[A]. We are interested in the kernel of this homomorphism.

Lemma-Definition 2.1.

Let GG be a profinite group and let AA, BB be continuous finite GG-sets. Fix a field FF of characteristic zero. The following conditions are equivalent:

  1. (1)

    μG(A)=μG(B)\mu_{G}(A)=\mu_{G}(B) in K0(Rep(G,F))K_{0}({\operatorname{Rep}}(G,F)).

  2. (2)

    F[A]F[B]F[A]\simeq F[B] as F[G]F[G]-modules.

We say that two continuous finite GG-sets AA and BB are Gassmann equivalent if they satisfy one of the above equivalent conditions.

Proof.

It is clear that (2) implies (1). Conversely, if μG(A)=μG(B)\mu_{G}(A)=\mu_{G}(B), then since F[A]F[A] and F[B]F[B] are finite dimensional (as FF-vector spaces), both F[A]F[A] and F[B]F[B] admit composition series in Rep(G,F){\operatorname{Rep}}(G,F) of finite length [EH18, Lemma 3.9]. The Jordan-Hölder theorem [Wei13, Excercice II.6.3] shows that F[A]F[A] and F[B]F[B] have the same collection of isomorphism classes of (simple) Jordan-Hölder factors. As both GG-modules F[A]F[A] and F[B]F[B] factor through a finite quotient G/HG/H of GG (we can take HH to be the intersection of the kernels of the two actions, which is an open subgroup of GG, hence has finite index), Maschke’s Theorem implies that F[A]F[A] and F[B]F[B] are semisimple GG-modules. Hence F[A]F[B]F[A]\simeq F[B]. \Box

The following lemma reduces Gassmann equivalence to the case of finite group actions. If AA is a GG-set, we write GAG^{A} for the kernel of the action; thus GAG^{A} is a normal subgroup and G/GAG/G^{A} acts on AA faithfully.

Lemma 2.2.

Two continuous finite GG-sets AA and BB are Gassmann equivalent if and only if GA=GB=:HG^{A}=G^{B}=:H and AA and BB are Gassmann equivalent as G/HG/H-sets.

Proof.

This follows from Lemma 2.1, together with the observation that GA=kerρAG^{A}=\ker{\rho_{A}} (and similarly GB=ker(ρB)G^{B}=\ker(\rho_{B})) where ρA:GGL(V,F)\rho_{A}:G\to{\mathrm{GL}}(V,F) is the homomorphism defining the F[G]F[G]-module structure of F[A]F[A] and VV is the underlying FF-vector space of F[A]F[A]. \Box

Lemma 2.3 ([Par13, Proposition 1], cf [Pra17, Definition 1]).

Let GG be a finite group and let AA and BB be finite GG-sets. Fix a field FF of characteristic zero. The following conditions are equivalent:

  1. (1)

    For every gGg\in G, gg fixes the same number of elements in AA and in BB.

  2. (2)

    There exist subgroups H1,,HrH_{1},\dots,H_{r}, H1,,HrH_{1}^{\prime},\dots,H_{r}^{\prime} of GG such that

    Ai=1rG/Hi,Bi=1rG/HiA\simeq\bigsqcup_{i=1}^{r}G/H_{i},\;\;\;\;B\simeq\bigsqcup_{i=1}^{r}G/H_{i}^{\prime}

    and for each conjugacy class TGT\subset G these subgroups satisfy

    (2.2) i=1r|THi||Hi|=i=1r|THi||Hi|.\sum_{i=1}^{r}\frac{|T\cap H_{i}|}{|H_{i}|}=\sum_{i=1}^{r}\frac{|T\cap H_{i}^{\prime}|}{|H_{i}^{\prime}|}.
  3. (3)

    AA and BB are Gassmann equivalent.

In particular, the kernel of μ\mu is independent of the choice of FF.

Proof.

Equivalence of (1), (2) and (3) for F=F=\mathbb{C} is proved in [Par13, Proposition 1]. Finally, condition (3) is independent of the choice of FF because the functor

Rep(G,)Rep(G,F){\operatorname{Rep}}(G,\mathbb{Q})\to{\operatorname{Rep}}(G,F)

is injective on isomorphism classes [Ser77, §14.6]. \Box

Remark 2.4.

The number of GG-orbits in Gassmann equivalent continuous GG-sets is the same. This follows from Lemma 2.2 and Lemma 2.3(2), or directly as the number of orbits equals the multiplicity of the trivial representation in the permutation representation.

Proposition 2.5.

Let GG be a profinite group and let AA and BB be Gassmann equivalent continuous finite GG-sets. If one of the following conditions is satisfied:

  • G/GAG/G^{A} is a cyclic group

  • AA is transitive and the stabilizer of a point is normal in GG

  • AA is transitive and |A|6|A|\leq 6

  • |A|5|A|\leq 5

then AA and BB are isomorphic GG-sets.

If αBurn(G)\alpha\in{\operatorname{Burn}}(G), then we write |α||\alpha| for the smallest max(|A|,|B|)\max(|A|,|B|) over all possible representations α=[A][B]\alpha=[A]-[B] with continuous GG-sets AA and BB. We have the following immediate corollary of Proposition 2.5.

Corollary 2.6.

Let αKer(μG)\alpha\in{\operatorname{Ker}}(\mu_{G}). If |α|5|\alpha|\leq 5, then α=0\alpha=0.

Proof of Proposition 2.5.

By Lemma 2.2, we can assume that GG is a finite group and the GG-actions on both AA and BB are faithful. When GG is a finite cyclic group, the result is [Par13, Proposition 4.1]. If AA is transitive, then by Remark 2.4, BB is also transitive. Let HH (resp. HH^{\prime}) be the stabilizer of an element aAa\in A (resp. bBb\in B). Then G/HG/H and G/HG/H^{\prime} are Gassmann equivalent; in this case one refers to (G,H,H)(G,H,H^{\prime}) as a Gassmann triple. It is well-known and easy to see (e.g. using (2.2) with r=1r=1) that if HH is normal, then H=HH^{\prime}=H. It is a nontrivial computation that if [G:H]6[G:H]\leq 6 in Gassmann triple (G,H,H)(G,H,H^{\prime}), HH and HH^{\prime} will be conjugate so that AA and BB are isomorphic, see [Per77, Proof of Theorem 3] or [BdS02, p.3].

Finally, we need to consider the case when AA may not be transitive but t:=|A|5t:=|A|\leq 5. We write A={1,,t}A=\{1,\dots,t\} and let StS_{t} be the permutation group of tt elements. Since the GG-action on AA is assumed to be faithful, it realizes GG as a subgroup of StS_{t}.

Up to adding trivial GG-sets {}\{*\}, we can assume that t=5t=5. By Remark 2.4, AA and BB have the same number of GG-orbits, and we argue according to the number of orbits mm. When m=5m=5, then both AA and BB are trivial GG-sets. When m=4m=4, then both AA and BB are isomorphic to {1,2}{}{}{}\{1,2\}\sqcup\{*\}\sqcup\{*\}\sqcup\{*\} and G/2G\simeq\mathbb{Z}/2\mathbb{Z} acts on {1,2}\{1,2\} by involution. The case m=1m=1 is already covered by the third case of Proposition 2.5. So it remains to study the case where m=2m=2 or 33.

First we show that GG uniquely determines the length of the orbits of the GG-set AA. If m=3m=3, then AA is isomorphic to either A3{}{}A_{3}\sqcup\{*\}\sqcup\{*\} or A2A2{}A_{2}\sqcup A^{\prime}_{2}\sqcup\{*\} with |A3|=3|A_{3}|=3 and |A2|=|A2|=2|A_{2}|=|A^{\prime}_{2}|=2. So GS3G\leq S_{3} in the former case and G/2×/2G\leq\mathbb{Z}/2\mathbb{Z}\times\mathbb{Z}/2\mathbb{Z} in the latter case. In the former case, since the GG-action is transitive on A3A_{3}, GG contains /3S3\mathbb{Z}/3\mathbb{Z}\leq S_{3}, so we cannot embed GG into /2×/2\mathbb{Z}/2\mathbb{Z}\times\mathbb{Z}/2\mathbb{Z}. Therefore GG uniquely determines the length of the orbits. If m=2m=2, then AA is isomorphic to either A3A2A_{3}\sqcup A_{2} or A4{}A_{4}\sqcup\{*\} with |Ai|=i|A_{i}|=i. So GS3×/2G\leq S_{3}\times\mathbb{Z}/2\mathbb{Z} in the former case and GS4G\leq S_{4} in the latter case. In the former case, since the GG-action is transitive on both A3A_{3} and A2A_{2}, GG contains /3×/2S3×/2\mathbb{Z}/3\mathbb{Z}\times\mathbb{Z}/2\mathbb{Z}\leq S_{3}\times\mathbb{Z}/2\mathbb{Z}. Since S4S_{4} has no elements of order six, we cannot embed GG into S4S_{4}. Hence the lengths of the orbits are determined by GG as well.

We still assume that m=2m=2 or 33. If A=AA′′A=A^{\prime}\sqcup A^{\prime\prime} as GG-sets where AA^{\prime} is transitive and A′′A^{\prime\prime} is a disjoint union of trivial GG-sets, then we also have the same type of decomposition B=BB′′B=B^{\prime}\sqcup B^{\prime\prime} with |A|=|B||A^{\prime}|=|B^{\prime}|. It follows that μG(A)=μG(B)\mu_{G}(A^{\prime})=\mu_{G}(B^{\prime}), so ABA^{\prime}\simeq B^{\prime} since both GG-sets are transitive. This covers the cases where A=A3{}{}A=A_{3}\sqcup\{*\}\sqcup\{*\} and A=A4{}A=A_{4}\sqcup\{*\}. If A=A2A2{}A=A_{2}\sqcup A^{\prime}_{2}\sqcup\{*\}, then either G=/2×/2G=\mathbb{Z}/2\mathbb{Z}\times\mathbb{Z}/2\mathbb{Z} or GG is the diagonal /2/2×/2\mathbb{Z}/2\mathbb{Z}\leq\mathbb{Z}/2\mathbb{Z}\times\mathbb{Z}/2\mathbb{Z}. In either case we verify that GG uniquely determines the GG-set structure of AA. If A=A3A2A=A_{3}\sqcup A_{2}, then either G/3×/2G\leq\mathbb{Z}/3\mathbb{Z}\times\mathbb{Z}/2\mathbb{Z} or G=S3×/2G=S_{3}\times\mathbb{Z}/2\mathbb{Z}. Once again in either case, we verify that GG uniquely determines the GG-set structure of AA. \Box

The following example shows that the lower bounds on the order of GG-sets in Proposition 2.5 are optimal.

Example 2.7.

(1) [Par13, 1.1] Let G=/2×/2G=\mathbb{Z}/2\mathbb{Z}\times\mathbb{Z}/2\mathbb{Z} and let

H1=(1,0),H2=(0,1),H3=(1,1),H_{1}=\langle(1,0)\rangle,\;\;H_{2}=\langle(0,1)\rangle,\;\;H_{3}=\langle(1,1)\rangle,

and

H1={(0,0)},H2=H3=G.H_{1}^{\prime}=\{(0,0)\},\;\;H_{2}^{\prime}=H_{3}^{\prime}=G.

Then H1,H2,H3H_{1},H_{2},H_{3} and H1,H2,H3H_{1}^{\prime},H_{2}^{\prime},H_{3}^{\prime} satisfy conditions of Lemma 2.3(2) thus give rise to nonisomorphic nontransitive Gassmann equivalent sets of order 66 with orbit decompositions 2+2+22+2+2 and 4+1+14+1+1 respectively. Indeed for abelian groups the condition (2.2) rewrites as

Hig1|Hi|=Hig1|Hi| for all gG\sum_{H_{i}\ni g}\frac{1}{|H_{i}|}=\sum_{H^{\prime}_{i}\ni g}\frac{1}{|H^{\prime}_{i}|}\text{ for all $g\in G$}

which is immediately verified.

(2) [Per77, p. 358] Let G=PSL2(𝔽7)PSL3(𝔽2)G=\mathrm{PSL}_{2}(\mathbb{F}_{7})\simeq\mathrm{PSL}_{3}(\mathbb{F}_{2}) be the simple group of order 168168. Let AA be the set of 𝔽2\mathbb{F}_{2}-points of the projective plane over 𝔽2\mathbb{F}_{2}, and BB be the set of 𝔽2\mathbb{F}_{2}-lines on this plane, that is points of the dual projective plane. Then AA, BB are transitive GG-sets of order 77. Using simple linear algebra of the PSL3(𝔽2)\mathrm{PSL}_{3}(\mathbb{F}_{2})-action on 𝔽22\mathbb{P}^{2}_{\mathbb{F}_{2}}, one shows that AA and BB satisfy Lemma 2.3(1), so that AA and BB are Gassmann equivalent, and one can check that they are not isomorphic. See also [BdS02, Theorem 3], which shows that this pair is the only nontrivial Gassmann triple of faithful transitive GG-sets of order 77.

2.2. L-equivalence and Gassmann equivalence

Let K0(Var/k)K_{0}({\mathrm{Var}}/k) denote the Grothendieck ring of varieties with generators given by isomorphism classes [X][X] of schemes of finite type over kk and relations generated by cut and paste relations

(2.3) [X]=[Z]+[XZ][X]=[Z]+[X\setminus Z]

for every closed ZXZ\subset X. The ring structure on K0(Var/k)K_{0}({\mathrm{Var}}/k) is induced by products of schemes. We write 𝕃=[𝔸1]\mathbb{L}=[\mathbb{A}^{1}].

It is known that 𝕃\mathbb{L} is a zero-divisor [Bor18] and that the annihilator of 𝕃k\mathbb{L}^{k}, k1k\geq 1 encodes deep geometric information. Following [KS18] we call two smooth projective connected varieties XX, YY L-equivalent if for some k0k\geq 0,

𝕃k([X][Y])=0.\mathbb{L}^{k}\cdot([X]-[Y])=0.

We sometimes refer to the [X]=[Y][X]=[Y] case as trivial L-equivalence. It is currently unknown if zero-dimensional varieties can be nontrivially L-equivalent. The smallest-dimensional example of nontrivial L-equivalence is that of genus one curves over non-closed fields [SZ20]. See [KS18, SZ20] for some details about conjectural relationship between L-equivalence and derived equivalence, and the references therein for the currently known examples. Note that the classes in the Grothendieck ring are insensitive to nonreduced structure, hence when studying L-equivalence we can always assume schemes to be reduced.

Remark 2.8.

For fields of positive characteristic, there exist two alternative definitions of the Grothendieck ring of varieties, hence alternative definitions of L-equivalence.

(1) First of all, one can define the Grothendieck ring K0bl(Var/k)K_{0}^{\operatorname{bl}}({\mathrm{Var}}/k) generated by classes of smooth projective varieties with blow up relations as in [Bit04, Theorem 3.1 (bl)]. We have an obvious homomorphism

K0bl(Var/k)K0(Var/k)K_{0}^{\operatorname{bl}}({\mathrm{Var}}/k)\to K_{0}({\mathrm{Var}}/k)

which is known to be an isomorphism if kk is a field of characteristic zero [Bit04, Theorem 3.1].

(2) Furthermore, in positive characteristic one can define a modified Grothendieck ring [NS09] by imposing an additional relation of identifying varieties related by universal homeomorphisms (originating from totally inseparable coverings in positive characteristic). It is not known if this additional relation in fact gives rise to a non-isomorphic ring as all the standard invariants which are used to distinguish elements in the Grothendieck ring factor through the modified ring as well [NS09].

Let kk be a field and let Gk=Gal(k¯/k)G_{k}={\operatorname{Gal}}(\bar{k}/k) where k¯\bar{k} denotes the separable closure of kk. For a variety X/kX/k (not necessarily smooth or projective) we consider its \ell-adic cohomology Hét,ci(Xk¯,)H^{i}_{{\text{\'{e}t}},c}(X_{\overline{k}},\mathbb{Q}_{\ell}) (with char(k)\ell\neq{\operatorname{char}}(k)) with compact supports as a GkG_{k}-module. These groups are finite-dimensional \mathbb{Q}_{\ell}-vector spaces  [FK88, Remark I.12.16], they vanish outside the range 0i2dim(X)0\leq i\leq 2\dim(X) and give rise to a group homomorphism

(2.4) μét:K0(Var/k)K0(Rep(Gk,))\mu_{{\text{\'{e}t}}}:K_{0}({\mathrm{Var}}/k)\to K_{0}({\operatorname{Rep}}(G_{k},\mathbb{Q}_{\ell}))

defined by

μét([X])=i=02dim(X)(1)i[Hét,ci(Xk¯,)].\mu_{{\text{\'{e}t}}}([X])=\sum_{i=0}^{2\dim(X)}(-1)^{i}[H^{i}_{{\text{\'{e}t}},c}(X_{\overline{k}},\mathbb{Q}_{\ell})].

We have

(2.5) μét(𝕃)=μét([1]1)=[(1)]\mu_{{\text{\'{e}t}}}(\mathbb{L})=\mu_{{\text{\'{e}t}}}([\mathbb{P}^{1}]-1)=[\mathbb{Q}_{\ell}(-1)]

and furthermore, the projective bundle formula for étale cohomology implies that μét\mu_{{\text{\'{e}t}}} is a [𝕃]\mathbb{Z}[\mathbb{L}]-module homomorphism where 𝕃\mathbb{L} acts by multiplication by (1)\mathbb{Q}_{\ell}(-1) on the Galois representations. (The map μét\mu_{\text{\'{e}t}} is even a ring homomorphism by the Künneth formula [Mil80, Corollary VI.8.23] but we do not need this fact.)

The étale realization (2.4) is useful for extracting information from a class in the Grothendieck of varieties. We record the following example to be used later.

Example 2.9.

Let XX be a geometrically rational smooth projective surface. Then since all cohomology classes on XX are algebraic, μét([X])=[]+[(2)]+[NS(Xk¯)(1)]\mu_{{\text{\'{e}t}}}([X])=[\mathbb{Q}_{\ell}]+[\mathbb{Q}_{\ell}(-2)]+[{\operatorname{NS}}(X_{\overline{k}})\otimes\mathbb{Q}_{\ell}(-1)], where \mathbb{Q}_{\ell} is considered as a trivial one-dimensional GkG_{k}-representation. In particular, if XX, XX^{\prime} are two such surfaces, and [X]=[X][X]=[X^{\prime}], then NS(Xk¯){\operatorname{NS}}(X_{\overline{k}}) and NS(Xk¯){\operatorname{NS}}(X^{\prime}_{\overline{k}}) have the same class in K0(Rep(Gk,))K_{0}({\operatorname{Rep}}(G_{k},\mathbb{Q}_{\ell})).

We explain how Gassmann equivalence relates to L-equivalence of reduced kk-schemes of dimension 0. Let 𝐄𝐭𝐒𝐜𝐡k\mathbf{EtSch}_{k} be the semi-ring of kk-schemes which are étale over Spec(k){\operatorname{Spec}}(k). As we assume kk to be perfect, 𝐄𝐭𝐒𝐜𝐡k\mathbf{EtSch}_{k} is also the semi-ring of reduced kk-schemes of dimension 0. Here in 𝐄𝐭𝐒𝐜𝐡k\mathbf{EtSch}_{k}, the addition (resp. multiplication) is defined by disjoint unions (resp. products over Spec(k){\operatorname{Spec}}(k)). For every Z𝐄𝐭𝐒𝐜𝐡kZ\in\mathbf{EtSch}_{k}, its base change Zk¯Z_{\bar{k}} to the separable closure k¯\bar{k} of kk is endowed with a GkG_{k}-action. By Galois descent, The map 𝐄𝐭𝐒𝐜𝐡k𝐆𝐬𝐞𝐭\mathbf{EtSch}_{k}\to\mathbf{Gset} sending Z𝐄𝐭𝐒𝐜𝐡kZ\in\mathbf{EtSch}_{k} to the underlying continuous GkG_{k}-set of Zk¯Z_{\bar{k}} is an isomorphism of semi-rings. As we assume kk to be perfect, this induces a ring isomorphism

(2.6) [Var0/k]Burn(Gk)\mathbb{Z}[{\mathrm{Var}}^{0}/k]\simeq{\operatorname{Burn}}(G_{k})

where Var0/k{\mathrm{Var}}^{0}/k denotes the set of (irreducible) kk-varieties of dimension 0.

We have a natural ring homomorphism

(2.7) [Var0/k]K0(Var/k)\mathbb{Z}[{\mathrm{Var}}^{0}/k]\to K_{0}({\mathrm{Var}}/k)

which sends ZZ to [Z][Z]. It follows from the blow up presentation of the Grothendieck ring [Bit04, Theorem 3.1] that over fields of characteristic zero (2.7) admits a splitting given by XSpec(H0(X,𝒪X))X\mapsto{\operatorname{Spec}}(H^{0}(X,\mathcal{O}_{X})) (XX smooth projective), hence for characteristic zero fields (2.7) is injective.

Throughout this text, the same notation Zk¯Z_{\bar{k}} (and also ZZ itself, when it does not lead to any confusion) denotes the underlying GkG_{k}-set of Zk¯Z_{\bar{k}} for every étale kk-scheme ZZ.

We work with Gassmann equivalence of such schemes considered as sets with Galois group action.

Lemma 2.10.

Let ZZ and ZZ^{\prime} be étale kk-schemes. If ZZ and ZZ^{\prime} are L-equivalent in the Grothendieck ring or in any of the modifications explained in Remark 2.8, then Zk¯Z_{\bar{k}} and Zk¯Z^{\prime}_{\bar{k}} are Gassmann equivalent.

Proof.

We use the étale realization (2.4), which is defined by pre-composition on K0bl(Var/k)K_{0}^{\operatorname{bl}}({\mathrm{Var}}/k), and also factors through the modified Grothendieck ring from Remark 2.8 (2) [NS09, Proposition 4.1.(3)].

For any étale kk-scheme ZZ, we have

(2.8) μét([Z])=[Hét0(Zk¯,)]=[[Zk¯]]=μGk([Zk¯]).\mu_{{\text{\'{e}t}}}([Z])=[H^{0}_{\text{\'{e}t}}(Z_{\bar{k}},\mathbb{Q}_{\ell})]=[\mathbb{Q}_{\ell}[Z_{\bar{k}}]]=\mu_{G_{k}}([Z_{\bar{k}}]).

For each j0j\geq 0 we consider the composition

Burn(Gk)[Var0/k]{{\operatorname{Burn}}(G_{k})\simeq\mathbb{Z}[{\mathrm{Var}}^{0}/k]}K0(Var/k){K_{0}({\mathrm{Var}}/k)}K0(Rep(Gk,)){K_{0}({\operatorname{Rep}}(G_{k},\mathbb{Q}_{\ell}))}K0(Rep(Gk,)){K_{0}({\operatorname{Rep}}(G_{k},\mathbb{Q}_{\ell}))}×𝕃j\scriptstyle{\times\mathbb{L}^{j}}μét\scriptstyle{\mu_{\text{\'{e}t}}}(j)\scriptstyle{\otimes\mathbb{Q}_{\ell}(j)}

which is equal to μGk\mu_{G_{k}} (2.1) by (2.5) and (2.8). Thus if ZZ, ZZ^{\prime} are L-equivalent, then Zk¯Z_{\bar{k}}, Zk¯Z^{\prime}_{\bar{k}} are Gassmann equivalent. \Box

Remark 2.11.

Properties of classes [X][X] of smooth projective varieties in the Grothendieck rings often go in parallel with properties of their Chow motives [And04]. In our situation, by [And04, Exemple 4.1.6.1], if kk has characteristic zero, then two étale kk-schemes ZZ, ZZ^{\prime} are Gassmann equivalent if and only if the Chow motives of ZZ, ZZ^{\prime} with rational coefficients are isomorphic. Thus in this setting Lemma 2.10 says that L-equivalence implies isomorphism of Chow motives. The same result is expected for all smooth projective varieties, and it follows from the conjectural uniqueness of direct sum decompositions for Chow motives, see e.g. [G0̈1, Conjecture 2.5, Conjecture 2.6].

Example 2.12.

Let kk be a field such that all continuous GkG_{k}-actions on finite sets factor through a finite cyclic quotient of GkG_{k}, e.g. k=k=\mathbb{R} or k=𝔽qk=\mathbb{F}_{q}. By Proposition 2.5 cyclic groups do not allow nontrivial Gassmann equivalence, hence by Lemma 2.10 L-equivalence of étale kk-schemes implies their isomorphism.

Corollary 2.13.

Let ZZ and ZZ^{\prime} be étale kk-schemes. If ZZ and ZZ^{\prime} are L-equivalent in the Grothendieck ring or any of the modifications explained in Remark 2.8, then ZZZ\simeq Z^{\prime} as soon as one of the following conditions is satisfied:

  • ZZ is kk-irreducible and Galois over Spec(k){\operatorname{Spec}}(k)

  • ZZ is kk-irreducible with degkZ6\deg_{k}Z\leq 6

  • ZZ satisfies degkZ5\deg_{k}Z\leq 5.

Proof.

By Lemma 2.10, Zk¯Z_{\bar{k}} and Zk¯Z^{\prime}_{\bar{k}} are Gassmann equivalent and by Proposition 2.5 they are isomorphic. \Box

Example 2.14.

Translating Example 2.7(1) into the language of étale schemes, we obtain the following. Let k=k=\mathbb{Q}, and define degree 66 schemes

Z\displaystyle Z =Spec((α))Spec((β))Spec((αβ))\displaystyle={\operatorname{Spec}}(\mathbb{Q}(\sqrt{\alpha}))\sqcup{\operatorname{Spec}}(\mathbb{Q}(\sqrt{\beta}))\sqcup{\operatorname{Spec}}(\mathbb{Q}(\sqrt{\alpha\beta}))
Z\displaystyle Z^{\prime} =Spec((α,β))Spec()Spec(),\displaystyle={\operatorname{Spec}}(\mathbb{Q}(\sqrt{\alpha},\sqrt{\beta}))\sqcup{\operatorname{Spec}}(\mathbb{Q})\sqcup{\operatorname{Spec}}(\mathbb{Q}),

where we choose any α,β\alpha,\beta\in\mathbb{Q}^{\star} which are nontrivial and distinct in /()2\mathbb{Q}^{\star}/(\mathbb{Q}^{\star})^{2}. Then ZZ, ZZ^{\prime} are Gassmann equivalent, and have isomorphic Chow motives (see Remark 2.11) but we do not know how to check if they are L-equivalent or not.

Embedding ZZ, ZZ^{\prime} into 2\mathbb{P}^{2} so that both images are in general position (for instance, if α+β1\alpha+\beta\neq 1, we can send ZZ onto [±α:1:0],[0:±β:1],[1:0:±αβ][\pm\sqrt{\alpha}:1:0],[0:\pm\sqrt{\beta}:1],[1:0:\pm\sqrt{\alpha\beta}] and ZZ^{\prime} onto [±α:±β:1],[1:0:1],[0:1:1][\pm\sqrt{\alpha}:\pm\sqrt{\beta}:1],[1:0:1],[0:1:1]), we obtain two del Pezzo surfaces of degree 33: X=BlZ(2)X={\operatorname{Bl}}_{Z}(\mathbb{P}^{2}), X=BlZ(2)X^{\prime}={\operatorname{Bl}}_{Z^{\prime}}(\mathbb{P}^{2}). These two surfaces are not isomorphic, as the first one has three kk-rational lines, and the second one has five.

However, we have an isomorphism

NS(X¯)NS(X¯){\operatorname{NS}}(X_{\bar{\mathbb{Q}}})\otimes\mathbb{Q}\simeq{\operatorname{NS}}(X^{\prime}_{\bar{\mathbb{Q}}})\otimes\mathbb{Q}

of permutation Galois representations, which shows that the associated Galois set is not uniquely defined. One can make a similar example integrally, using integral Gassmann triples as in [Sco93], [Pra17]. See Example 5.12 where we explain that using our techniques we can recover ZZ (resp. ZZ^{\prime}) from XX (resp. XX^{\prime}), providing another proof that they are not isomorphic.

Finally note that we do not know if [X]=[X][X]=[X^{\prime}] as by the blow up relation in the Grothendieck ring of varieties, [X]=[2]+𝕃[Z][X]=[\mathbb{P}^{2}]+\mathbb{L}[Z] and [X]=[2]+𝕃[Z][X^{\prime}]=[\mathbb{P}^{2}]+\mathbb{L}[Z^{\prime}] so [X]=[X][X]=[X^{\prime}] would imply the L-equivalence of ZZ and ZZ^{\prime}, which is unknown.

3. Factorization centers

In this section we introduce the invariant c(ϕ)c(\phi) keeping track of the factorization centers in any decomposition of a birational isomorphism of ϕ\phi between surfaces into a sequence of blow ups and blow downs, formulate the main theorem and interpret it in terms of the truncated Grothendieck ring of varieties.

3.1. Formulation of the main result

Fix a perfect field kk. Let ϕ:XY\phi:X\dashrightarrow Y be a birational isomorphism of smooth projective kk-surfaces. By the strong factorization theorem [Man86, Corollary of Lemma III.4.4] (or [Sta18, Lemma 54.17.2]) we have a decomposition

(3.1) X~\textstyle{\widetilde{X}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}α\scriptstyle{\alpha}β\scriptstyle{\beta}X\textstyle{X\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ϕ\scriptstyle{\phi}Y\textstyle{Y}

with both α\alpha and β\beta being compositions of blow ups with smooth centers Z1,,ZrZ_{1},\dots,Z_{r} and Z1,,ZsZ_{1}^{\prime},\dots,Z_{s}^{\prime} respectively, which are zero-dimensional smooth schemes. The factorization center of ϕ\phi is defined as

(3.2) c(ϕ)=i=1r[Zi]i=1s[Zi][Var0/k]c(\phi)=\sum_{i=1}^{r}[Z_{i}]-\sum_{i=1}^{s}[Z_{i}^{\prime}]\in\mathbb{Z}[{\mathrm{Var}}^{0}/k]

(see §2.2 for the definition of [Var0/k]\mathbb{Z}[{\mathrm{Var}}^{0}/k]). We remark that 𝕃c(ϕ)\mathbb{L}\cdot c(\phi), regarded as an element of K0(Var/k)K_{0}({\mathrm{Var}}/k), measures the difference between classes [Y][Y] and [X][X] in the Grothendieck ring of varieties, see Lemma 3.9. Given that 𝕃\mathbb{L} is a zero-divisor in the Grothendieck ring, c(ϕ)c(\phi) potentially contains more information than 𝕃c(ϕ)\mathbb{L}\cdot c(\phi).

We explain the well-definedness of cc and its basic properties. To do that it is most convenient to consider the groupoid Bir2/k{\operatorname{Bir}}_{2}/k of birational types of surfaces, whose objects are smooth projective surfaces and morphisms are birational isomorphisms.

Recall that if 𝒞\mathcal{C} is a groupoid, and GG a group, a homomorphism from 𝒞\mathcal{C} to GG is a functor from 𝒞\mathcal{C} to GG, where GG is considered as a groupoid with one object.

Lemma 3.1.

c(ϕ)c(\phi) does not depend on the choice of factorization of ϕ\phi and defines a homomorphism c:Bir2/k[Var0/k]c:{\operatorname{Bir}}_{2}/k\to\mathbb{Z}[{\mathrm{Var}}^{0}/k]. Explicitly, for any two birational isomorphisms of surfaces ϕ:XX\phi:X\dashrightarrow X^{\prime}, ψ:XX′′\psi:X^{\prime}\dashrightarrow X^{\prime\prime} we have

c(ψϕ)=c(ψ)+c(ϕ).c(\psi\circ\phi)=c(\psi)+c(\phi).

In particular, for any surface XX we have a homomorphism

c:Bir(X)[Var0/k].c:{\operatorname{Bir}}(X)\to\mathbb{Z}[{\mathrm{Var}}^{0}/k].
Proof.

Consider the diagram (3.1). Let E1,,EmXE_{1},\dots,E_{m}\subset X (resp. E1,,EnYE_{1}^{\prime},\dots,E_{n}^{\prime}\subset Y) be the irreducible components of the exceptional divisor of ϕ\phi (resp. ϕ1\phi^{-1}). Let D1,,DtX~D_{1},\dots,D_{t}\subset\widetilde{X} be the irreducible divisors which are contracted by both α\alpha and β\beta. This way the centers Zi,i=1,,rZ_{i},\;i=1,\dots,r (resp. Zi,i=1,,rZ_{i}^{\prime},\;i=1,\dots,r^{\prime}) are in one-to-one correspondence with the collection of divisors {Ei}i=1,,m{Dj}j=1,,t\{E_{i}\}_{i=1,\dots,m}\cup\{D_{j}\}_{j=1,\dots,t} (resp. {Ei}i=1,,n{Dj}j=1,,t\{E^{\prime}_{i}\}_{i=1,\dots,n}\cup\{D_{j}\}_{j=1,\dots,t}).

Note that each prime divisor DD contracted by α\alpha or β\beta is birational to 1×Z\mathbb{P}^{1}\times Z, where ZZ is the center of the corresponding blow up, which can be recovered from DD as ZSpec(H0(D,𝒪D))Z\simeq{\operatorname{Spec}}(H^{0}(D,\mathcal{O}_{D})). Cancelling out the centers corresponding to D1,,DtD_{1},\dots,D_{t} we obtain

c(ϕ)=i=1r[Zi]i=1s[Zi]=i=1m[Spec(H0(Ei,𝒪))]i=1n[Spec(H0(Ei,𝒪))]\displaystyle c(\phi)=\sum_{i=1}^{r}[Z_{i}]-\sum_{i=1}^{s}[Z_{i}^{\prime}]=\sum_{i=1}^{m}[{\operatorname{Spec}}(H^{0}(E_{i},\mathcal{O}))]-\sum_{i=1}^{n}[{\operatorname{Spec}}(H^{0}(E_{i}^{\prime},\mathcal{O}))]

thus the expression (3.2) only depends on the exceptional divisors of ϕ\phi and ϕ1\phi^{-1}, hence c(ϕ)c(\phi) is independent of the choice of strong factorization (3.1).

To show that c(ψϕ)=c(ψ)+c(ϕ)c(\psi\circ\phi)=c(\psi)+c(\phi) consider the diagram

X3{X_{3}}X1{X_{1}}X2{X_{2}}X{X}X{X^{\prime}}X′′{X^{\prime\prime}}μ\scriptstyle{\mu}μ\scriptstyle{\mu^{\prime}}τ\scriptstyle{\tau}τ\scriptstyle{\tau^{\prime}}ν\scriptstyle{\nu}ν\scriptstyle{\nu^{\prime}}ϕ\scriptstyle{\phi}ψ\scriptstyle{\psi}

where τ\tau, τ\tau^{\prime} (resp. ν\nu, ν\nu^{\prime}) provide a strong factorization for ϕ\phi (resp. ψ\psi), and μ\mu, μ\mu^{\prime} provide a strong factorization for ν1τ\nu^{-1}\tau^{\prime}. It is clear that cc is additive on regular birational isomorphisms, hence

c(ψϕ)=c(νμ)c(τμ)=c(νμ)c(νμ)+c(τμ)c(τμ)=c(ν)c(ν)+c(τ)c(τ)=c(ψ)+c(ϕ).\begin{split}c(\psi\circ\phi)&=c(\nu^{\prime}\circ\mu^{\prime})-c(\tau\circ\mu)\\ &=c(\nu^{\prime}\circ\mu^{\prime})-c(\nu\circ\mu^{\prime})+c(\tau^{\prime}\circ\mu)-c(\tau\circ\mu)\\ &=c(\nu^{\prime})-c(\nu)+c(\tau^{\prime})-c(\tau)=c(\psi)+c(\phi).\end{split}

\Box

Corollary 3.2.

There is a unique assignment ϕc(ϕ)[Var0/k]\phi\mapsto c(\phi)\in\mathbb{Z}[{\mathrm{Var}}^{0}/k] defined for all birational isomorphisms ϕ\phi between smooth projective kk-surfaces, satisfying the following axioms

  1. (i)

    For any (biregular) isomorphism ϕ\phi, c(ϕ)=0c(\phi)=0

  2. (ii)

    If X~\widetilde{X} is a blow up of XX with smooth connected center ZZ and ϕ:X~X\phi:\widetilde{X}\to X the contraction map, then

    c(ϕ)=[Z],c(ϕ1)=[Z]c(\phi)=-[Z],\;c(\phi^{-1})=[Z]
  3. (iii)

    For composable birational isomorphisms

    c(ψϕ)=c(ψ)+c(ϕ).c(\psi\circ\phi)=c(\psi)+c(\phi).
Proof.

Uniqueness follows from the strong factorization (3.1), and existence is the content of Lemma 3.1. \Box

Remark 3.3.

The invariant c(ϕ)c(\phi) is generalized to higher-dimensional varieties in [LS22], to the equivariant setting in [KT22] and to varieties with a logarithmic form in [CLKT23]. In all these contexts properties of c(ϕ)c(\phi) are analogous to those stated in Corollary 3.2.

Proposition 3.4.
  1. (i)

    For any field extension L/kL/k we have a commutative diagram

    Bir(X,Y)\textstyle{{\operatorname{Bir}}(X,Y)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}c\scriptstyle{c}[Var0/k]\textstyle{\mathbb{Z}[{\mathrm{Var}}^{0}/k]\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\simeq}(2.6)\scriptstyle{\eqref{eq:burn-galois}}Burn(Gk)\textstyle{{\operatorname{Burn}}(G_{k})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Bir(XL,YL)\textstyle{{\operatorname{Bir}}(X_{L},Y_{L})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}c\scriptstyle{c}[Var0/L]\textstyle{\mathbb{Z}[{\mathrm{Var}}^{0}/L]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\simeq}(2.6)\scriptstyle{\eqref{eq:burn-galois}}Burn(GL)\textstyle{{\operatorname{Burn}}(G_{L})}

    where the left vertical arrow is an extension of scalars for birational map, the middle vertical arrow is defined by extension of scalars, that is it maps an étale kk-scheme ZZ to the sum of the connected components of Z×kLZ\times_{k}L, and the right vertical arrow is the restriction of the group action defined through the map Gal(L¯/L)Gal(k¯/k){\operatorname{Gal}}(\bar{L}/L)\to{\operatorname{Gal}}(\bar{k}/k) induced by any choice of embedding k¯L¯\bar{k}\hookrightarrow\bar{L} compatible with kLk\hookrightarrow L.

    In particular, applying the diagram to an automorphism σAut(L/k)\sigma\in{\operatorname{Aut}}(L/k) the left square gives c(σ(ϕ))=σ(c(ϕ))c(\sigma(\phi))=\sigma(c(\phi)), that is cc commutes with the group action by the automorphisms of the field.

  2. (ii)

    For any finite field extension L/kL/k and surfaces XX, YY over LL, let X|kX|_{k}, Y|kY|_{k} denote the underlying kk-surfaces.222Here by X|kX|_{k} we mean the kk-surface given by the composition XSpec(L)Spec(k)X\to{\operatorname{Spec}}(L)\to{\operatorname{Spec}}(k) (this is not the Weil restriction of scalars). If L/kL/k is nontrivial, the surface X|kX|_{k} is not geometrically connected: X|k×kLX|_{k}\times_{k}L is isomorphic to a disjoint union of [L:k][L:k] copies of XX. We have the following commutative diagram:

    Bir(X,Y)\textstyle{{\operatorname{Bir}}(X,Y)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}c\scriptstyle{c}[Var0/L]\textstyle{\mathbb{Z}[{\mathrm{Var}}^{0}/L]\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\simeq}(2.6)\scriptstyle{\eqref{eq:burn-galois}}Burn(GL)\textstyle{{\operatorname{Burn}}(G_{L})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Bir(X|k,Y|k)\textstyle{{\operatorname{Bir}}(X|_{k},Y|_{k})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}c\scriptstyle{c}[Var0/k]\textstyle{\mathbb{Z}[{\mathrm{Var}}^{0}/k]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\simeq}(2.6)\scriptstyle{\eqref{eq:burn-galois}}Burn(Gk)\textstyle{{\operatorname{Burn}}(G_{k})}

    where the middle vertical map is restriction of scalars.

Proof.

By Lemma 3.1, in both (i), (ii) it is sufficient to check a single blow up where the statements are clear. \Box

The main result of the paper is the following.

Theorem 3.5.

For any two smooth projective kk-surfaces XX, YY and any two birational isomorphisms ϕ,ψ:XY\phi,\psi:X\dashrightarrow Y, we have c(ϕ)=c(ψ)c(\phi)=c(\psi). In particular, c:Bir(X)[Var0/k]c:{\operatorname{Bir}}(X)\to\mathbb{Z}[{\mathrm{Var}}^{0}/k] is a zero map.

We prove Theorem 3.5 at the end of §5. The result is straightforward when kk is algebraically closed, because in this case the invariant takes values in [Spec(k)]\mathbb{Z}\cdot[{\operatorname{Spec}}(k)]\simeq\mathbb{Z}, and measures the difference of the ranks of two Néron-Severi groups (cf Proposition 4.4). However, Theorem 3.5 becomes a nontrivial statement when kk is an arbitrary perfect field, and its proof depends on the two-dimensional Minimal Model Program. This result is also specific for surfaces and fails to be true in higher dimension, even over algebraically closed fields [LS22].

Example 3.6.

Let k=k=\mathbb{R}. In this case Theorem 3.5 says that birational automorphisms of surfaces blow up the same number of rational points as the number of rational divisors they blow down, and that they blow up the same number of pairs of complex conjugate points as the number of pairs of complex conjugate divisors they blow down. This can be proved directly by considering the Galois action on the Néron-Severi group (see also Example 2.12).

Remark 3.7.

A different proof of Theorem 3.5 in the case when the surfaces XX, YY are rational can be deduced from a more recent result by Lamy and Schneider [LS21] who analyze relations between Sarkisov links to prove that Cr2(k){\operatorname{Cr}}_{2}(k) is generated by involutions. This implies that every homomorphism from Cr2(k){\operatorname{Cr}}_{2}(k) to a free abelian group is trivial (for k=k=\mathbb{R} one can also deduce this from [Zim18, Theorem 1.1]). On the other hand this is not true for other types of del Pezzo surfaces: Blanc, Schneider and Yasinsky have constructed nontrivial homomorphisms Bir(X){\operatorname{Bir}}(X)\to\mathbb{Z} for Severi-Brauer surfaces XX [BSY22].

Example 3.8.

Consider the following composition ϕ\phi of type IIII links (see Definition 4.2 for links) between del Pezzo surfaces dPddP_{d} (where d=K2d=K^{2} stands for the degree)

dP7\textstyle{dP_{7}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Bl1\scriptstyle{{\operatorname{Bl}}_{1}}Bl2\scriptstyle{{\operatorname{Bl}}_{2}}dP3\textstyle{dP_{3}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Bl2\scriptstyle{{\operatorname{Bl}}_{2}}Bl5\scriptstyle{{\operatorname{Bl}}_{5}}dP4\textstyle{dP_{4}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Bl5\scriptstyle{{\operatorname{Bl}}_{5}}Bl1\scriptstyle{{\operatorname{Bl}}_{1}}ϕ:\textstyle{\phi:}2\textstyle{\mathbb{P}^{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}dP8\textstyle{dP_{8}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}dP5\textstyle{dP_{5}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}2\textstyle{\mathbb{P}^{2}}

where each map Bli{\operatorname{Bl}}_{i} is the blow up (resp. blow down) along a Galois orbit ZiZ_{i} (resp. ZiZ^{\prime}_{i}) of degree ii. See [Isk96] for the general results on links, or see the explicit constructions explained below for the existence of the three links above. We have

c(ϕ)=[Z2][k]+[Z5][Z2]+[k][Z5]=[Z2][Z2]+[Z5][Z5],c(\phi)=[Z_{2}]-[k]+[Z_{5}]-[Z_{2}^{\prime}]+[k]-[Z_{5}^{\prime}]=[Z_{2}]-[Z_{2}^{\prime}]+[Z_{5}]-[Z_{5}^{\prime}],

so that c(ϕ)=0c(\phi)=0 is equivalent to Z2Z2 and Z5Z5Z_{2}\simeq Z_{2}^{\prime}\text{ and }Z_{5}\simeq Z_{5}^{\prime}.

One way to understand this is to read the diagram as a composition of two different well-known rationality constructions for del Pezzo surace dP5dP_{5}: one blows up a point PP and contracts a Galois orbit of five (1)(-1)-curves which are proper preimages of conics through PP (connecting dP5dP_{5} with 2\mathbb{P}^{2} moving right) or one blows up a Galois orbit of two points Q1Q_{1}, Q2Q_{2} and contracts five (1)(-1)-curves obtained as proper preimages of cubics passing through Q1Q_{1}, Q2Q_{2} onto a rational quadric dP8dP_{8} and then transforms it to 2\mathbb{P}^{2} (connecting dP5dP_{5} with 2\mathbb{P}^{2} moving left). This way dP5dP_{5} has two potentially different rationality centers that can be blown up by 2dP5\mathbb{P}^{2}\dashrightarrow dP_{5}, namely Z5Z_{5} and Z5Z_{5}^{\prime} and the result is that these centers are in fact isomorphic (see Corollary 5.9 and Example 5.10).

3.2. Interpretation in terms of the Grothendieck ring of varieties

From the perspective of the Grothendieck ring of varieties, we have the following interpretation of c(ϕ)c(\phi).

Lemma 3.9.

For any birational isomorphism ϕ:XY\phi:X\dashrightarrow Y between smooth projective surfaces we have the following identity in K0(Var/k)K_{0}({\mathrm{Var}}/k)

(3.3) [Y]=[X]+𝕃c¯(ϕ),[Y]=[X]+\mathbb{L}\cdot\overline{c}(\phi),

where c¯\overline{c} is the composition of cc with the natural homomorphism (2.7) [Var0/k]K0(Var/k)\mathbb{Z}[{\mathrm{Var}}^{0}/k]\to K_{0}({\mathrm{Var}}/k).

Proof.

Using Lemma 3.1 we see that (3.3) is preserved under compositions of birational isomorphisms. Since birational isomorphisms are decomposed into blow ups and blow downs along kk-étale subschemes, it suffices to check Lemma 3.9 for a single blow up, where the result is clear by Corollary 3.2(2) and the blow up formula in the Grothendieck ring. \Box

Remark 3.10.

We see that 𝕃c¯(ϕ)K0(Var/k)\mathbb{L}\cdot\overline{c}(\phi)\in K_{0}({\mathrm{Var}}/k) only depends on [X][X] and [Y][Y], however it is known that 𝕃\mathbb{L} is a zero-divisor [Bor18] so a priori we cannot divide 𝕃c¯(ϕ)\mathbb{L}\cdot\overline{c}(\phi) by 𝕃\mathbb{L} and deduce that c¯(ϕ)\overline{c}(\phi) (or c(ϕ)c(\phi)) only depends on XX and YY but not on ϕ\phi.

We can informally reformulate Theorem 3.5 by the statement that 0-dimensional kk-varieties cannot be L-equivalent via surfaces. Let us explain this. For each n0n\geq 0 consider K0(Varn/k)K_{0}({\mathrm{Var}}^{\leq n}/k), the abelian group generated by isomorphism classes of varieties of dimension n\leq n, modulo cut and paste relations (2.3). For each nmn\leq m there is a (generally non-injective) group homomorphism

K0(Varn/k)K0(Varm/k)K_{0}({\mathrm{Var}}^{\leq n}/k)\to K_{0}({\mathrm{Var}}^{\leq m}/k)

and K0(Var/k)K_{0}({\mathrm{Var}}/k), as an abelian group, is the colimit of this system. For every nn there is a surjective homomorphism

K0(Varn/k)[Birn/k],K_{0}({\mathrm{Var}}^{\leq n}/k)\to\mathbb{Z}[{\operatorname{Bir}}_{n}/k],

where Birn/k{\operatorname{Bir}}_{n}/k is the set of kk-birational classes of dimension nn. The kernel of this homomorphism is spanned by varieties of dimension (n1)\leq(n-1), but it may not be isomorphic to K0(Varn1/k)K_{0}({\mathrm{Var}}^{\leq n-1}/k); see [Zak17] for interpretation of this in terms of an algebraic K{\mathrm{K}}-theory spectral sequence.

For n=0n=0, K0(Var0/k)K_{0}({\mathrm{Var}}^{\leq 0}/k) is canonically isomorphic to [Var0/k]\mathbb{Z}[{\mathrm{Var}}^{0}/k], and K0(Var1/k)K_{0}({\mathrm{Var}}^{\leq 1}/k) fits into a split exact sequence

(3.4) 0[Var0/k]K0(Var1/k)[Bir1/k]0.0\to\mathbb{Z}[{\mathrm{Var}}^{0}/k]\to K_{0}({\mathrm{Var}}^{\leq 1}/k)\to\mathbb{Z}[{\operatorname{Bir}}_{1}/k]\to 0.

Our main results can be reformulated as results about K0(Var2/k)K_{0}({\mathrm{Var}}^{\leq 2}/k):

Corollary 3.11 (of Theorem 3.5).

(i) We have a short exact sequence

(3.5) 0[Var0/k][Bir1/k]K0(Var2/k)[Bir2/k]0,0\to\mathbb{Z}[{\mathrm{Var}}^{0}/k]\oplus\mathbb{Z}[{\operatorname{Bir}}_{1}/k]\to K_{0}({\mathrm{Var}}^{\leq 2}/k)\to\mathbb{Z}[{\operatorname{Bir}}_{2}/k]\to 0,

where the last map sends a combination of varieties to the birational classes of its 22-dimensional components, and the first map sends a kk-variety ZZ of dimension 0 to [Z][Z] and a birational class of a curve to the class of its unique smooth projective model.

(ii) We have a (noncanonical) splitting

K0(Var2/k)[Var0/k][Bir1/k][Bir2/k],K_{0}({\mathrm{Var}}^{\leq 2}/k)\simeq\mathbb{Z}[{\mathrm{Var}}^{0}/k]\oplus\mathbb{Z}[{\operatorname{Bir}}_{1}/k]\oplus\mathbb{Z}[{\operatorname{Bir}}_{2}/k],

and in particular K0(Var2/k)K_{0}({\mathrm{Var}}^{\leq 2}/k) is a (torsion-)free abelian group.

(iii) If ZZ and ZZ^{\prime} are étale kk-schemes, and

𝕃i([Z][Z])=0\mathbb{L}^{i}\cdot([Z]-[Z^{\prime}])=0

in K0(Var2/k)K_{0}({\mathrm{Var}}^{\leq 2}/k) with i{0,1,2}i\in\{0,1,2\} (where 𝕃i([Z][Z])\mathbb{L}^{i}\cdot([Z]-[Z^{\prime}]) is represented by [𝔸i×kZ][𝔸i×kZ][\mathbb{A}^{i}\times_{k}Z]-[\mathbb{A}^{i}\times_{k}Z^{\prime}]), then ZZ and ZZ^{\prime} are isomorphic.

Proof.

(i) Because resolution of singularities and weak factorization (even strong factorization) are known for surfaces over arbitrary perfect fields, the proof of [Bit04] goes through to show that we have an isomorphism

K0bl(Var2/k)K0(Var2/k),K_{0}^{\operatorname{bl}}({\mathrm{Var}}^{\leq 2}/k)\to K_{0}({\mathrm{Var}}^{\leq 2}/k),

where the first group is defined by smooth projective varieties of dimension 2\leq 2 and Bittner’s blow up relations. These relations can be equivalently presented as

(3.6) [Y]=[X]+𝕃c¯(ϕ)[Y]=[X]+\mathbb{L}\cdot\bar{c}(\phi)

for all birational isomorphisms ϕ:XY\phi:X\dashrightarrow Y between smooth projective surfaces. Indeed, (3.6) include the blow up relations if ϕ:YX\phi:Y\to X is a smooth blow up as a particular case and conversely, (3.6) follow as soon as we impose the blow up relations as in the proof of Lemma 3.9. Since curves admit unique smooth projective models, we have the canonical splitting of (3.4) giving

K0(Var1/k)=K0bl(Var1/k)[Var0/k][Bir1/k].K_{0}({\mathrm{Var}}^{\leq 1}/k)=K_{0}^{\operatorname{bl}}({\mathrm{Var}}^{\leq 1}/k)\simeq\mathbb{Z}[{\mathrm{Var}}^{0}/k]\oplus\mathbb{Z}[{\operatorname{Bir}}_{1}/k].

Furthermore we have an obvious short exact sequence

K0bl(Var1/k)K0bl(Var2/k)[Bir2/k]0K_{0}^{\operatorname{bl}}({\mathrm{Var}}^{\leq 1}/k)\to K_{0}^{\operatorname{bl}}({\mathrm{Var}}^{\leq 2}/k)\to\mathbb{Z}[{\operatorname{Bir}}_{2}/k]\to 0

and to prove (3.5) it suffices to show that the first map in the sequence is split-injective.

The splitting is based on Theorem 3.5 and is not canonical. First of all, we choose a smooth projective representative for each birational class of surfaces. If LL is a two-dimensional function field, we write MLM_{L} for the chosen model. We define the splitting

ϵ:K0bl(Var2/k)K0bl(Var1/k)\epsilon:K_{0}^{\operatorname{bl}}({\mathrm{Var}}^{\leq 2}/k)\to K_{0}^{\operatorname{bl}}({\mathrm{Var}}^{\leq 1}/k)

by identity on classes of smooth projective curves and zero-dimensional schemes and if XX is a smooth projective surface we define

ϵ([X])=𝕃c¯(Mk(X)𝜓X)\epsilon([X])=\mathbb{L}\cdot\overline{c}(M_{k(X)}\overset{\psi}{\dashrightarrow}X)

for any choice of a birational isomorphism ψ\psi between XX and its model Mk(X)M_{k(X)}. The fact that this is independent of ψ\psi is the content of Theorem 3.5, and the fact that ϵ\epsilon is well-defined, that is preserves the relations (3.6), follows immediately from the property that cc is additive on compositions (Lemma 3.1) applied to a composition of birational isomorphisms M𝜓XϕYM\overset{\psi}{\dashrightarrow}X\overset{\phi}{\dashrightarrow}Y (where M:=Mk(X)=Mk(Y)M:=M_{k(X)}=M_{k(Y)}).

(ii) follows from (3.5); explicit splittings are constructed in the proof of (i).

(iii) The case i=2i=2 is clear as if 𝔸i×kZ\mathbb{A}^{i}\times_{k}Z and 𝔸i×kZ\mathbb{A}^{i}\times_{k}Z^{\prime} are kk-birational, then ZZ and ZZ^{\prime} are isomorphic. For i=1i=1, the element 𝕃([Z][Z])\mathbb{L}\cdot([Z]-[Z^{\prime}]) is the image of ([Z]+[Z],[1×kZ][1×kZ])(-[Z]+[Z^{\prime}],[\mathbb{P}^{1}\times_{k}Z]-[\mathbb{P}^{1}\times_{k}Z^{\prime}]) under the first map in (3.5), and since this map is injective, [Z]=[Z][Z]=[Z^{\prime}], that is ZZ and ZZ^{\prime} are isomorphic. Finally the case i=0i=0 follows again from (3.5). \Box

4. Birational geometry of surfaces

In this section we recall the Minimal Model Program and Sarkisov link decomposition for surfaces, which is used in our proof of Theorem 3.5. In Proposition 4.4 we prove Theorem 3.5 for birational types with particularly simple links. In §4.2 we investigate linear systems of rational curves on del Pezzo surfaces, which will be needed to finish the proof of Theorem 3.5 in §5.

4.1. Birational classification of surfaces and links

Let X/kX/k be a geometrically irreducible surface. Recall that XX is called minimal if it does not have a Galois orbit of disjoint (1)(-1)-curves, or equivalently, every regular birational map XXX\to X^{\prime} from XX to a smooth surface is an isomorphism. We say that XX is rational if it is birational to 2\mathbb{P}^{2} over kk and XX is geometrically rational if it is birational to 2\mathbb{P}^{2} over k¯\overline{k}. A del Pezzo surface is a smooth projective geometrically irreducible surface with ample anticanonical class.

A conic bundle is a fibered surface π:XC\pi:X\to C, with CC a smooth projective curve, such that the generic fiber of π\pi is a smooth rational curve and XX has Picard rank two. For a conic bundle π:XC\pi:X\to C, πωX/C\pi_{*}\omega_{X/C}^{\vee} is locally free of rank 33 and we have an embedding X(πωX/C)X\hookrightarrow\mathbb{P}(\pi_{*}\omega_{X/C}^{\vee}) over CC realizing each fiber of π:XC\pi:X\to C as a plane conic (see [Has09, Corollary 3.7]). The number mm of geometric singular fibers of a conic bundle equals

(4.1) m=8KX2m=8-K_{X}^{2}

see e.g. [Isk79, Theorem 3], in particular KX28K_{X}^{2}\leq 8.

We write NS(X){\operatorname{NS}}(X) for the Néron-Severi group of divisors modulo algebraic equivalence; by the theorem of Néron-Severi it is a finitely generated abelian group. Each contraction of a Galois orbit of (1)(-1)-curves decreases the rank of the Néron-Severi group, hence a sequence of contractions always terminates to produce a minimal surface birationally equivalent to the given one.

We have the following classification result going back to Enriques, Manin and Iskovskikh, see [Kol96, Theorem III.2.2].

Theorem 4.1 (Minimal Model Program in dimension two).

Any geometrically irreducible minimal surface X/kX/k is isomorphic to exactly one of the following:

  • Non-geometrically rational case

    1. (1)

      Surface with nef KXK_{X}

    2. (2)

      Conic bundle π:XC\pi:X\to C with g(C)>0g(C)>0

  • Geometrically rational case

    1. (3)

      Conic bundle XCX\to C with g(C)=0g(C)=0

    2. (4)

      Minimal del Pezzo surface of Picard rank one

In cases (2) and (3), NS(X){\operatorname{NS}}(X) is of rank two.

Note that if kk is algebraically closed, any conic bundle in (2), (3) has no singular fibers (as components of singular fibers are (1)(-1)-curves), thus (2) consists of ruled surfaces and (3) consists of Hirzebruch surfaces 𝔽n\mathbb{F}_{n}.

Definition 4.2 (Sarkisov links, see 2.2 in [Isk96]).

Let XBX\to B and XBX^{\prime}\to B^{\prime} be Mori fiber spaces with dimX=dimX=2\dim X=\dim X^{\prime}=2. Here each Mori fiber space is either a del Pezzo surface of Picard rank one or a conic bundle of Picard rank two over a smooth curve. A Sarkisov link between XX and XX^{\prime} is a birational map ν:XX\nu:X\dashrightarrow X^{\prime} satisfying one of the following descriptions.

  • (Type I) ν1\nu^{-1} is the blow up of XX at a smooth closed point with B=Spec(k)B={\operatorname{Spec}}(k) and XBX^{\prime}\to B^{\prime} a conic bundle.

  • (Type II) We have a commutative diagram

    (4.2) X{X}Y{Y}X{X^{\prime}}B{B}B{B^{\prime}}α\scriptstyle{\alpha}β\scriptstyle{\beta}\scriptstyle{\sim}

    and ν=βα1\nu=\beta\circ\alpha^{-1} where both α\alpha and β\beta are blow ups at a smooth closed point. In this case,

    • (Type IIC) either both XBX\to B and XBX^{\prime}\to B^{\prime} are conic bundles;

    • (Type IID) or B=B=Spec(k)B=B^{\prime}={\operatorname{Spec}}(k).

  • (Type III) This is the inverse of link (I).

  • (Type IV) XBX\rightarrow B, XBX^{\prime}\to B^{\prime} are conic bundles and ν\nu is an isomorphism not respecting the conic bundle structures.

We rely on decomposing birational isomorphisms of surfaces into Sarkisov links [Isk96]. Some of the links are easier to deal with.

Lemma 4.3.

If ν:XX\nu:X\dashrightarrow X^{\prime} is a Sarkisov link of type IIC, then KX2=KX2K_{X}^{2}=K_{X^{\prime}}^{2}, rkNS(X)=rkNS(X){\operatorname{rk}}\;{\operatorname{NS}}(X)={\operatorname{rk}}\;{\operatorname{NS}}(X^{\prime}) and c(ν)=0c(\nu)=0.

Proof.

Let ϕ:XB\phi:X\to B be the conic bundle as in (4.2). By [Isk96, Theorem 2.6], ν\nu is an elementary transformation of XBX\to B. More precisely, α\alpha in (4.2) is the blow up at a smooth closed point pXp\in X lying in a smooth fiber CC of ϕ\phi and β\beta in (4.2) is the contraction of the proper transformation C~\tilde{C} of CC under α\alpha. Since C~C1×Spec(k){p}\tilde{C}\simeq C\simeq\mathbb{P}^{1}\times_{{\operatorname{Spec}}(k)}\{p\}, the blow up center of β\beta is isomorphic to pp. Hence rkNS(X)=rkNS(X){\operatorname{rk}}\;{\operatorname{NS}}(X)={\operatorname{rk}}\;{\operatorname{NS}}(X^{\prime}) and c(ν)=0c(\nu)=0. The equality KX2=KX2K_{X}^{2}=K_{X^{\prime}}^{2} follows from the fact that XBX\to B and XBX^{\prime}\to B^{\prime} have the same number of geometric singular fibers using (4.1). \Box

Recall that the degree of a geometrically rational surface XX is defined to be KX2K_{X}^{2}. The next result is a step towards Theorem 3.5.

Proposition 4.4.

Let XX, XX^{\prime} be a pair of birational minimal geometrically irreducible surfaces. Assume that XX is either (1) geometrically irrational or (2) geometrically rational and of degree 4\leq 4, then the same holds for XX^{\prime} and for any birational isomorphism ϕ:XX\phi:X\dashrightarrow X^{\prime} we have

(4.3) c(ϕ)=(rkNS(X)rkNS(X))[Spec(k)].c(\phi)=({\operatorname{rk}}\;{\operatorname{NS}}(X^{\prime})-{\operatorname{rk}}\;{\operatorname{NS}}(X))\cdot[{\operatorname{Spec}}(k)].

In particular c(ϕ)=0c(\phi)=0 for birational automorphisms of such surfaces.

Proof.

(1) It is clear that XX^{\prime} is geometrically irrational and moreover, XX and XX^{\prime} have the same type in Theorem 4.1. If XX, XX^{\prime} have nef canonical class, then every birational isomorphism is a biregular isomorphism by [IS96, Corollary 1 in II.7.3] so that XXX\simeq X^{\prime} and c(ϕ)=0c(\phi)=0 as required.

If XX, XX^{\prime} are conic bundles over a curve of positive genus, then by [Sch22, Corollary 3.2], birational maps between them can be decomposed into Sarkisov links of type IIC, in which case rkNS(X)=rkNS(X){\operatorname{rk}}\;{\operatorname{NS}}(X)={\operatorname{rk}}\;{\operatorname{NS}}(X^{\prime}) and c(ϕ)=0c(\phi)=0 by Lemma 4.3.

(2) It follows from the classification of links in [Isk96, Theorem 2.6] that elementary links will only connect XX to minimal surfaces of degree 4\leq 4 or conic bundles of degree 33 and Picard rank 22, and that birational isomorphisms between such surfaces will be decomposed into

  • Biregular isomorphisms

  • Type IIC links

  • Bertini and Geiser involutions

  • Blow ups of a rational point (between degree 44 and degree 33 surfaces).

For each of these types the claim of Proposition 4.4 is true, namely in the first three cases rkNS(X)=rkNS(X){\operatorname{rk}}\;{\operatorname{NS}}(X)={\operatorname{rk}}\;{\operatorname{NS}}(X^{\prime}) and c(ϕ)=0c(\phi)=0 (using Lemma 4.3 for the second case), while in the last case the ranks differ by one and the result is true by definition of cc. Equality (4.3) follows by additivity of cc under compositions (Lemma 3.1). \Box

Remark 4.5.

From the proof of Proposition 4.4, we note that in (4.3), we have actually c(ϕ)=0c(\phi)=0 or c(ϕ)=±[Spec(k)]c(\phi)=\pm[{\operatorname{Spec}}(k)]. Moreover, c(ϕ)=±[Spec(k)]c(\phi)=\pm[{\operatorname{Spec}}(k)] only happens when ϕ\phi is a map between a del Pezzo surface of degree 44 and a conic bundle of degree 33.

4.2. Rational curves on del Pezzo surfaces

Linear systems of rational curves on del Pezzo surfaces are closely related to factorization centers: for instance, to create a rational two-dimensional quadric XX, one needs to blow up 2\mathbb{P}^{2} in Z2Z_{2}, where Z2Z_{2} is degree two étale kk-scheme, and to contract a line through the center. This way the original scheme Z2Z_{2} can be recovered as the scheme parametrizing rulings on XX (cf Definition 5.2).

Definition 4.6.

Let j1j\geq 1. We call a complete linear system |L||L| of curves on a del Pezzo surface XX a linear system of degree jj rational curves if a general member C|L|C\in|L| is a smooth rational curve and (KX)L=j(-K_{X})\cdot L=j.

For each j1j\geq 1 we consider j(X)\mathcal{H}^{j}(X), the Hilbert scheme of curves on XX with general members of each component being smooth rational curves of degree jj. By an easy computation (see Lemma 4.7(i)) H1(X,L)=0H^{1}(X,L)=0, hence j(X)\mathcal{H}^{j}(X) is smooth [Kol96, Theorem 2.8]. In fact, over the algebraic closure k¯\overline{k}, each j(Xk¯)\mathcal{H}^{j}(X_{\overline{k}}) is a disjoint union of projective spaces parametrizing effective divisors in the corresponding linear systems. We refer to points of 1(X)\mathcal{H}^{1}(X) as lines on XX, points of 2(X)\mathcal{H}^{2}(X) as conics and so on. When XX is a twisted form of 1×1\mathbb{P}^{1}\times\mathbb{P}^{1} (a minimal del Pezzo surface of degree 88), then all curves have even degree, and families of conics on XX are also called rulings.

We note that by adjunction formula, a linear system LL of rational curves of degree jj satisfies

(4.4) L2=j2,(KX)L=j.L^{2}=j-2,\;\;(-K_{X})\cdot L=j.

However the latter numerical conditions are not sufficient to deduce that a linear system consists of rational curves. Indeed the linear system 3H+E3H+E on BlP(2){\operatorname{Bl}}_{P}(\mathbb{P}^{2}), where HH is the class of a line on 2\mathbb{P}^{2}, and EE is the exceptional divisor, satisfies (4.4) with j=10j=10 but has a fixed component and contains no rational curves.

We investigate how to find all linear systems of rational curves of a given degree following [Man86]. The first step is to solve (4.4). Let XX be a del Pezzo surface of degree d=KX2d=K_{X}^{2}. Assume Xk¯X_{\overline{k}} is not isomorphic to 1×1\mathbb{P}^{1}\times\mathbb{P}^{1} so that Xk¯X_{\overline{k}} is a blow up of 2\mathbb{P}^{2} in r:=9dr:=9-d points by [Man86, Theorem IV.2.5][Isk79, §3]. Write D=aHi=1rbiEiD=aH-\sum_{i=1}^{r}b_{i}E_{i}, where HH is the pullback of the hyperplane class and EiE_{i} are the irreducible components of the exceptional divisor. Since KX=3Hi=1rEi-K_{X}=3H-\sum_{i=1}^{r}E_{i}, (4.4) translates into:

(4.5) {i=1rbi=3aji=0rbi2=a2j+2\left\{\begin{array}[]{l}\sum_{i=1}^{r}b_{i}=3a-j\\ \sum_{i=0}^{r}b_{i}^{2}=a^{2}-j+2\\ \end{array}\right.

The case j=1j=1 is that of lines on del Pezzo surfaces [Man86].

Lemma 4.7.

Let XX be a del Pezzo surface of degree dd over an algebraically closed field kk.

  1. (i)

    A linear system of rational curves of degree jj satisfies h0(X,L)=jh^{0}(X,L)=j, and higher cohomology of LL vanish.

  2. (ii)

    Any linear system satisfying (4.4) with j1j\geq 1 is non-empty.

  3. (iii)

    A linear system satisfying (4.4) with 1jd11\leq j\leq d-1 is a linear system of rational curves.

  4. (iv)

    Let 1jd11\leq j\leq d-1. The assignment |D||KXD||D|\mapsto|-K_{X}-D| establishes a bijection between linear systems of rational curves of degrees jj and djd-j.

Proof.

(i) Let C|L|C\in|L| be a smooth rational curve. Taking cohomology for the short exact sequence

0𝒪X𝒪X(C)𝒪C(C)00\to\mathcal{O}_{X}\to\mathcal{O}_{X}(C)\to\mathcal{O}_{C}(C)\to 0

we get H1(X,𝒪X(C))=H2(X,𝒪X(C))=0H^{1}(X,\mathcal{O}_{X}(C))=H^{2}(X,\mathcal{O}_{X}(C))=0, and hence h0(X,𝒪X(C))=jh^{0}(X,\mathcal{O}_{X}(C))=j by Riemann-Roch.

(ii) We have h2(X,L)=h0(X,LωX)=0h^{2}(X,L)=h^{0}(X,L^{\vee}\otimes\omega_{X})=0 (because LωXL^{\vee}\otimes\omega_{X} has negative anticanonical degree) and the Riemann-Roch theorem implies

h0(X,L)j.h^{0}(X,L)\geq j.

(iii) If X1×1X\simeq\mathbb{P}^{1}\times\mathbb{P}^{1}, then it is easy to see that solutions of (4.4) are precisely divisors (a,1)(a,1), a0a\geq 0 and (1,b)(1,b), b0b\geq 0, and these are linear systems of rational curves (no upper bound on the degree needed). Now assume that XX is a blow up of 2\mathbb{P}^{2} in 0r80\leq r\leq 8 points. Since DD satisfies (4.4), it follows that KXD-K_{X}-D satisfies (4.4) with degree djd-j. By (ii) both |D||D| and |KXD||-K_{X}-D| are nonempty. Therefore if D=aHi=1rbiEiD=aH-\sum_{i=1}^{r}b_{i}E_{i}, then 0a30\leq a\leq 3. Solving equations (4.5) with bib_{i}\in\mathbb{Z}, under the assumption 1jd11\leq j\leq d-1 gives rise to the following divisors, which we list up to reordering of the exceptional divisors:

  • a=0a=0: E1E_{1}

  • a=1a=1: Hi=1tEiH-\sum_{i=1}^{t}E_{i}

  • a=2a=2: 2Hi=1rtEi2H-\sum_{i=1}^{r-t}E_{i}

  • a=3a=3: 3H2E1i=2rEi3H-2E_{1}-\sum_{i=2}^{r}E_{i}

Indeed, the cases a=0,1a=0,1 are straightforward and the case a=2,3a=2,3 follow via the DKXDD\mapsto-K_{X}-D substitution. The a=1,2a=1,2 cases are only possible when r7r\leq 7 (d2d\geq 2). We must have 0t20\leq t\leq 2, and for r=6r=6 (resp. r=7r=7), only allow t=1,2t=1,2 (resp. t=2t=2). Under these conditions each of the divisors in the list is linearly equivalent to a smooth rational curve.

(iv) Follows from (iii). \Box

Proposition 4.8.

Let X/kX/k be a del Pezzo surface of degree dd.

  1. (i)

    For each j1j\geq 1, j(X)\mathcal{H}^{j}(X) is either empty or a smooth Severi-Brauer fibration of relative dimension j1j-1 over a smooth zero-dimensional scheme j(X)\mathcal{M}^{j}(X). For each 1jd11\leq j\leq d-1 we have a natural isomorphism j(X)dj(X)\mathcal{M}^{j}(X)\simeq\mathcal{M}^{d-j}(X).

  2. (ii)

    Assume j(X)\mathcal{H}^{j}(X) is nonempty and let ZXZ\subset X be a zero-dimensional subscheme of degree j1j-1. Let j(X,Z)\mathcal{H}^{j}(X,Z) be the closed subvariety of j(X)\mathcal{H}^{j}(X) parametrizing curves containing ZZ. If j(X,Z)\mathcal{H}^{j}(X,Z) is nonempty and zero-dimensional, then it is isomorphic to j(X)\mathcal{M}^{j}(X).

  3. (iii)

    Suppose that Xk¯X_{\overline{k}} is obtained by blowing up r5r\leq 5 points on 2\mathbb{P}^{2} with exceptional divisors E1,,ErE_{1},\ldots,E_{r}. Let HNS(Xk¯)H\in{\operatorname{NS}}(X_{\overline{k}}) be the pullback of the hyperplane class. Then the classes of conics (resp. cubics) on Xk¯X_{\overline{k}} are HEiH-E_{i} and 2Hi=14Ei2H-\sum_{i=1}^{4}E_{i} (resp. HH and 2Hi=13Ei2H-\sum_{i=1}^{3}E_{i}).

  4. (iv)

    Consider the following cases.

    1. (dP8dP_{8})

      Let XX be a minimal del Pezzo surface of degree 88. Then 4(X)Spec(k)\mathcal{M}^{4}(X)\simeq{\operatorname{Spec}}(k) and 2(X)6(X)Z2\mathcal{M}^{2}(X)\simeq\mathcal{M}^{6}(X)\simeq Z_{2}, where Z2Z_{2} is an étale kk-scheme of degree 2.

    2. (dP6dP_{6})

      Let XX be a del Pezzo surface of degree 66. Then

      2(X)4(X)Z3,3(X)Z2\mathcal{M}^{2}(X)\simeq\mathcal{M}^{4}(X)\simeq Z_{3},\;\mathcal{M}^{3}(X)\simeq Z_{2}

      for degree two (resp. degree three) étale kk-schemes Z2,Z3Z_{2},Z_{3}.

    3. (dP5dP_{5})

      Let XX be a del Pezzo surface of degree 55. Then 2(X)3(X)Z5\mathcal{M}^{2}(X)\simeq\mathcal{M}^{3}(X)\simeq Z_{5} for a degree five étale kk-scheme Z5Z_{5}.

Proof.

(i) Since H1(X,𝒪X)=0H^{1}(X,\mathcal{O}_{X})=0 [Kol96, III.3.2.1], deforming an effective divisor DD as a subscheme is equivalent to deforming it in its linear system, so j(X)k¯\mathcal{H}^{j}(X)_{\overline{k}} is either empty or a disjoint union of complete linear systems (see [Kol96, I.1.14.2]). Since the Hilbert polynomial of a rational curve CC on a del Pezzo surface is determined by its degree C(KX)C\cdot(-K_{X}), and since the Hilbert scheme is projective [Kol96, Theorem I.1.4], in particular of finite type, it follows that j(X)k¯\mathcal{H}^{j}(X)_{\overline{k}} is a finite disjoint union of projective spaces. (Finiteness also follows from the fact there are finitely many solutions for (4.4)\eqref{eq:num-rat} with fixed jj.) By Lemma 4.7(i), these projective spaces have dimension j1j-1.

Thus we have shown that j(X)\mathcal{H}^{j}(X) is a smooth scheme of finite type over kk isomorphic over k¯\overline{k} to a finite disjoint union of projective spaces. Let j(X)=Spec(H0(j(X),𝒪))\mathcal{M}^{j}(X)={\operatorname{Spec}}(H^{0}(\mathcal{H}^{j}(X),\mathcal{O})); it is a smooth zero-dimensional scheme, and j(X)j(X)\mathcal{H}^{j}(X)\to\mathcal{M}^{j}(X) is a Severi-Brauer fibration. Finally, the isomorphism j(X)dj(X)\mathcal{M}^{j}(X)\simeq\mathcal{M}^{d-j}(X) is given by |C||KXC||C|\mapsto|-K_{X}-C|, using Lemma 4.7 (iv).

(ii) We claim that the projection j(X,Z)j(X)\mathcal{H}^{j}(X,Z)\to\mathcal{M}^{j}(X) is an isomorphism. By Galois descent it is sufficient to verify this over the algebraic closure. Each fiber of j(X)k¯j(X)k¯\mathcal{H}^{j}(X)_{\bar{k}}\to\mathcal{M}^{j}(X)_{\bar{k}} is equal to |L||L| for some line bundle LL. It follows from the assumption that the linear system defined by H0(Xk¯,LIZ)H0(Xk¯,L)H^{0}(X_{\bar{k}},L\otimes I_{Z})\subset H^{0}(X_{\bar{k}},L) is a point in |L||L| where IZI_{Z} be the ideal sheaf of Zk¯Xk¯Z_{\bar{k}}\subset X_{\bar{k}}. Hence j(X,Z)k¯j(X)k¯\mathcal{H}^{j}(X,Z)_{\bar{k}}\to\mathcal{M}^{j}(X)_{\bar{k}} is an isomorphism.

(iii) is proven by solving (4.5).

(iv) Galois descent identifies smooth zero-dimensional scheme with a set with Galois action. Thus it suffices to verify the numbers of conic bundles and nets of cubics over the algebraic closure in each case. For (dP8dP_{8}), the linear system of quartic rational curves on 1×1\mathbb{P}^{1}\times\mathbb{P}^{1} is defined by 𝒪(1,1)\mathcal{O}(1,1), and the two pencils of conics 𝒪(1,0)\mathcal{O}(1,0) and 𝒪(0,1)\mathcal{O}(0,1). The cases (dP6dP_{6}) and (dP5dP_{5}) follow from (iii). \Box

5. Models of large degree

In this section we deal with surfaces admitting the most interesting birational automorphisms. These are geometrically rational surfaces with minimal models of degree 5\geq 5. At the end of this section we prove Theorem 3.5 and discuss rationality centers for rational surfaces.

5.1. Virtual Néron-Severi sets

We introduce a type of surfaces X/kX/k which we call models of large degree:

  1. (dP9dP_{9})

    2\mathbb{P}^{2} or a Severi-Brauer surface

  2. (dP8dP_{8})

    minimal del Pezzo surface of degree 88; in this case Xk¯1×1X_{\overline{k}}\simeq\mathbb{P}^{1}\times\mathbb{P}^{1}

  3. (C8C_{8})

    smooth conic bundle over a conic; in this case Xk¯𝔽nX_{\overline{k}}\simeq\mathbb{F}_{n}, and we assume n1n\geq 1 to avoid the overlap with (dP8dP_{8})

  4. (dP6dP_{6})

    del Pezzo surface of degree 66

  5. (dP5dP_{5})

    del Pezzo surface of degree 55

Note that we do not make any assumption on the Picard rank of the surfaces in the list above.

Proposition 5.1.

All minimal geometrically rational surfaces XX with KX25K_{X}^{2}\geq 5 and all conic bundles over a curve of genus 0 with KX25K_{X}^{2}\geq 5 are among models of large degree.

Proof.

By the work of Iskovskikh [Isk79] explained in Theorem 4.1 (3), (4), XX is a del Pezzo surface of Picard rank one or a conic bundle of Picard rank two. Minimal del Pezzo surfaces of degree KX25K_{X}^{2}\geq 5 are all among (dP9dP_{9}), (dP8dP_{8}), (dP6dP_{6}), (dP5dP_{5}) (del Pezzo surface of degree 77 is not minimal, see e.g. [Isk79, Corollary on p.37]).

Assume that XX is a conic bundle over a curve of genus 0; using (4.1) we have 5KX285\leq K_{X}^{2}\leq 8. By [Isk79, Theorem 5], XX is model of large degree. \Box

Definition 5.2.

For each model of large degree we define its virtual Néron-Severi Galois set AXBurn(Gk)[Var0/k]A_{X}\in{\operatorname{Burn}}(G_{k})\simeq\mathbb{Z}[{\mathrm{Var}}^{0}/k] as follows. In each case, ZiZ_{i} refers to one of the finite étale kk-schemes of degree ii introduced in Proposition 4.8(iv).

  • (dP9dP_{9})

    AX=[Spec(k)]A_{X}=[{\operatorname{Spec}}(k)]

  • (dP8dP_{8})

    AX=[Z2]A_{X}=[Z_{2}], where Z2Z_{2} parametrizes the rulings (that is, the conic bundle structures) on Xk¯X_{\overline{k}}

  • (C8C_{8})

    AX=2[Spec(k)]A_{X}=2[{\operatorname{Spec}}(k)]

  • (dP6dP_{6})

    AX=[Z2]+[Z3][Spec(k)]A_{X}=[Z_{2}]+[Z_{3}]-[{\operatorname{Spec}}(k)], where Z3Z_{3} parametrizes three pencils of conics on Xk¯X_{\overline{k}} and Z2Z_{2} parametrizes two families of cubics

  • (dP5dP_{5})

    AX=[Z5]A_{X}=[Z_{5}], where Z5Z_{5} parametrizes five pencils of conics on Xk¯X_{\overline{k}}

Note that in all cases except for (dP6dP_{6}), the virtual Néron-Severi set is realized as a set. The following Lemma explains the name Néron-Severi set. See (2.1) for the definition of μGk\mu_{G_{k}}.

Lemma 5.3.

If AXBurn(Gk)A_{X}\in{\operatorname{Burn}}(G_{k}) is the virtual Néron-Severi set of XX, and FF is any field of characteristic zero, then μGk(AX)=[NS(Xk¯)F]K0(Rep(Gk,F))\mu_{G_{k}}(A_{X})=[{\operatorname{NS}}(X_{\overline{k}})\otimes F]\in K_{0}({\operatorname{Rep}}(G_{k},F)).

Proof.
  • (dP9dP_{9})

    AXA_{X} is a set consisting of one element and NS(Xk¯){\operatorname{NS}}(X_{\overline{k}}) is the one-dimensional trivial GkG_{k}-representation.

  • (dP8dP_{8})

    AXA_{X} is the set of rulings on XX, and NS(Xk¯){\operatorname{NS}}(X_{\overline{k}}) is freely generated by these rulings.

  • (C8C_{8})

    AXA_{X} is a set with two elements and trivial action. Since KXK_{X} and the class of a fiber form a basis of NS(X)F{\operatorname{NS}}(X)\otimes F and rkNS(Xk¯)=rkNS(𝔽n)=2{\operatorname{rk}}\,{\operatorname{NS}}(X_{\overline{k}})={\operatorname{rk}}\,{\operatorname{NS}}(\mathbb{F}_{n})=2, necessarily NS(Xk¯){\operatorname{NS}}(X_{\overline{k}}) is of rank two with trivial Galois action.

  • (dP6dP_{6})

    By Proposition 4.8 (iii), NS(Xk¯)F{\operatorname{NS}}(X_{\overline{k}})\otimes F is generated by three classes of conics and two classes of cubics, modulo a one-dimensional space with a trivial GkG_{k}-action; this implies the result.

  • (dP5dP_{5})

    By Proposition 4.8 (iii), NS(Xk¯)F{\operatorname{NS}}(X_{\overline{k}})\otimes F is freely generated by classes of conics AXA_{X}.

\Box

It follows from Lemma 5.3 that the rank of NS(X){\operatorname{NS}}(X) equals the number of orbits of the virtual Néron-Severi set AXA_{X}. In particular, XX has Picard rank one if and only AXA_{X} has only one (virtual) orbit. Since μGk\mu_{G_{k}} is not injective, we can not simply define the virtual Galois set of a surface using μGk(AX)=[NS(Xk¯)F]\mu_{G_{k}}(A_{X})=[{\operatorname{NS}}(X_{\overline{k}})\otimes F], see Example 2.14 for an illustration; this is why we define them case by case in Definition 5.2.

Remark 5.4.

The Galois sets forming AXA_{X} naturally appear in the Chow motive of XX [Gil15, (9) and the following Remark] and as semiorthogonal components in the derived category of the corresponding surface [BSS11], [AB18, Propositions 9.8, 10.1]. Furthermore, singular versions of those kk-algebras show up in the study of the derived categories of the corresponding singular del Pezzo surfaces [Kuz19], [KKS22], [Xie21].

Proposition 5.5.

Let XX be a model of large degree, and let ϕ:XX\phi:X\dashrightarrow X^{\prime} be a birational isomorphism to another minimal surface; then XX^{\prime} is also a model of large degree, and

(5.1) c(ϕ)=AXAX.c(\phi)=A_{X^{\prime}}-A_{X}.

First we prove that if we apply μGk\mu_{G_{k}} to both sides of (5.1), we have equality.

Lemma 5.6.

Let XX, XX^{\prime} be models of large degree. Let ϕ:XYX\phi:X\leftarrow Y\to X^{\prime} be a composition of a blow up and a blow down (the centers of the blow ups can be disconnected, or empty). We have

μGk(AX)μGk(AX)=μGk(c(ϕ)).\mu_{G_{k}}(A_{X}^{\prime})-\mu_{G_{k}}(A_{X})=\mu_{G_{k}}(c(\phi)).
Proof.

Let ZZ, ZZ^{\prime} be the centers of the two blow ups. By the blow up formula, we have [X]+𝕃[Z]=[Y]=[X]+𝕃[Z][X]+\mathbb{L}\cdot{[Z]}=[Y]=[X^{\prime}]+\mathbb{L}\cdot{[Z^{\prime}]} in the Grothendieck ring. Applying the étale realization (2.4) to this equality (cf. Example 2.9), and using Lemma 5.3 with F=F=\mathbb{Q}_{\ell} we get

(5.2) μGk(AX)+[[Zk¯]]=[NS(Yk¯)]=μGk(AX)+[[Zk¯]]K0(Rep(Gk,)).\mu_{G_{k}}(A_{X})+[\mathbb{Q}_{\ell}[Z_{\bar{k}}]]=[{\operatorname{NS}}(Y_{\overline{k}})\otimes\mathbb{Q}_{\ell}]=\mu_{G_{k}}(A_{X^{\prime}})+[\mathbb{Q}_{\ell}[Z^{\prime}_{\bar{k}}]]\in K_{0}({\operatorname{Rep}}(G_{k},\mathbb{Q}_{\ell})).

Hence μGk(AX)μGk(AX)=μGk(c(ϕ))\mu_{G_{k}}(A_{X}^{\prime})-\mu_{G_{k}}(A_{X})=\mu_{G_{k}}(c(\phi)). \Box

Before proving Proposition 5.5, we establish some particular cases, which rely on Gassmann equivalence being trivial for Galois sets of order 5\leq 5.

Lemma 5.7.

Let XX, XX^{\prime} be models of large degree. Let ϕ:XYX\phi:X\leftarrow Y\to X^{\prime} be a composition of a blow up and a blow down. Assume that either (1) the two blow ups have isomorphic centers or (2) KY25K_{Y}^{2}\geq 5. Then (5.1) holds for ϕ\phi.

Proof.

Let ZZ, ZZ^{\prime} be the centers of the two blow ups. We need to show that

α:=(AX+[Z])(AX+[Z])=0.\alpha\mathrel{:=}(A_{X}+[Z])-(A_{X^{\prime}}+[Z^{\prime}])=0.

By Lemma 5.6 and Corollary 2.6, it is sufficient to show that |α|5|\alpha|\leq 5, that is to represent α\alpha as a difference of two GkG_{k}-sets, each having order 5\leq 5. Recall that the virtual GkG_{k}-set AXA_{X} is represented by a GkG_{k}-set of order 5\leq 5, except possibly when KX2=6K_{X}^{2}=6 in which case AX+[Spec(k)]A_{X}+[{\operatorname{Spec}}(k)] is a GkG_{k}-set, of order 55.

(1) Since [Z]=[Z][Z]=[Z^{\prime}], we have α=AXAX\alpha=A_{X}-A_{X^{\prime}}. If KX2=KX26K_{X}^{2}=K_{X^{\prime}}^{2}\neq 6, then |α|5|\alpha|\leq 5, as |AX|=|AX|5|A_{X}|=|A_{X^{\prime}}|\leq 5. If KX2=KX2=6K_{X}^{2}=K_{X^{\prime}}^{2}=6, then α=(AX+[Spec(k)])(AX+[Spec(k)])\alpha=(A_{X}+[{\operatorname{Spec}}(k)])-(A_{X^{\prime}}+[{\operatorname{Spec}}(k)]) also shows that |α|5|\alpha|\leq 5.

(2) Assume both KX2K_{X}^{2}, KX2K_{X^{\prime}}^{2} are not equal to 66. In this case by (5.2) both AX+[Z]A_{X}+[Z] and AX+[Z]A_{X^{\prime}}+[Z^{\prime}] have order equal to rkNS(Xk¯)=10KY25{\operatorname{rk}}\;{\operatorname{NS}}(X_{\overline{k}})=10-K_{Y}^{2}\leq 5, and the original representation of α\alpha shows that |α|5|\alpha|\leq 5. The cases when KX2=6K_{X}^{2}=6 or KX2=6K_{X^{\prime}}^{2}=6 are analogous and are left to the reader. \Box

Proof of Proposition 5.5.

By [Isk96, Theorem 2.5], any birational isomorphism between minimal geometrically rational surfaces is a composition of Sarkisov links explained in Definition 4.2. Since cc is a homomorphism and sends isomorphisms to zero, it suffices to prove Proposition 5.5 for every link of type I, II, or III.

For type I links we write aba\leftarrow b for a link ϕ:XX\phi:X\leftarrow X^{\prime} with KX2=aK_{X}^{2}=a, KX2=bK_{X^{\prime}}^{2}=b. We have the following possibilities according to [Isk96, Theorem 2.6(i)]: 989\leftarrow 8, 959\leftarrow 5, 868\leftarrow 6. Here XBX^{\prime}\to B^{\prime} is a conic bundle of degree 5\geq 5, hence XX^{\prime} is a model of large degree by Proposition 5.1, and (5.1) follows from Lemma 5.7(2) (with one of the centers empty). Exactly the same argument proves the claim for links of type III.

For type IIC links, the result is true by Lemma 4.3 and Proposition 5.1. For a type IID link XYXX\leftarrow Y\to X^{\prime}, by the first statement of Proposition 4.4(2), KX25K_{X}^{2}\geq 5 if and only if KX25K^{2}_{X^{\prime}}\geq 5. Hence XX is a model of large degree if and only if XX^{\prime} is.

It remains to show (5.1) for each link of type IID. We write adba\leftarrow d\to b for a type IID link XYXX\leftarrow Y\to X^{\prime} between surfaces of degree aa, dd, bb. Since cc takes values in a torsion-free abelian group, it vanishes on involutions and in particular the Bertini and Geiser involutions; these are links with d=1d=1 and d=2d=2 respectively in the list of links in [Isk96, Theorem 2.6(ii)]. On the other hand, links with d5d\geq 5 are covered by Lemma 5.7(2).

Thus we only have to consider links with d=3d=3 or d=4d=4.

Claim 5.8.

Let XϕYϕXX\xleftarrow{\phi}Y\xrightarrow{\phi^{\prime}}X^{\prime} be a link of type IID such that KY23K_{Y}^{2}\geq 3. Let ZXZ\subset X be the blowup center of ϕ\phi and DXD\subset X the divisor contracted by ϕϕ1\phi^{\prime}\circ\phi^{-1}. Then each irreducible component of Dk¯D_{\bar{k}} is a smooth rational curve of degree δ\delta (according to Definition 4.6) containing exactly δ1\delta-1 points of Zk¯Z_{\bar{k}}. For the type IID links with KY2{3,4}K_{Y}^{2}\in\{3,4\} listed below, δ\delta has the following description:

  • 945,9399\leftarrow 4\to 5,9\leftarrow 3\rightarrow 9: δ=6\delta=6 (which are conics in 2\mathbb{P}^{2} in the classical sense).

  • 8488\leftarrow 4\rightarrow 8: δ=4\delta=4.

  • 8358\leftarrow 3\to 5: δ=6\delta=6.

  • 6466\leftarrow 4\to 6: δ=3\delta=3.

  • 6366\leftarrow 3\to 6: δ=4\delta=4.

  • 5495\leftarrow 4\to 9: δ=2\delta=2.

  • 5385\leftarrow 3\to 8: δ=3\delta=3.

Proof.

Since KY23K_{Y}^{2}\geq 3, Yk¯Y_{\bar{k}} is obtained by blowing up r6r\leq 6 points on 2\mathbb{P}^{2} with exceptional divisors E1,,ErE_{1},\ldots,E_{r}. Let HNS(Yk¯)H\in{\operatorname{NS}}(Y_{\bar{k}}) be the pullback of the hyperplane class of 2\mathbb{P}^{2}. By solving (4.5), the (1)(-1)-classes on Yk¯Y_{\bar{k}} are one of the following:

  • EiE_{i}, with i{1,,r}i\in\{1,\ldots,r\};

  • HEiEjH-E_{i}-E_{j}, with i,j{1,,r}i,j\in\{1,\ldots,r\} and iji\neq j;

  • (only when r=6r=6) KYH+Ei-K_{Y}-H+E_{i}, with i{1,,6}i\in\{1,\ldots,6\}.

From the above description, any pair of (1)(-1)-curves EE and EE^{\prime} on Yk¯Y_{\bar{k}} satisfies EE1E\cdot E^{\prime}\leq 1. Since Xk¯X_{\bar{k}} is a simultaneous contraction of disjoint (1)(-1)-curves on Yk¯Y_{\bar{k}}, it follows that each irreducible component CC of Dk¯D_{\bar{k}} is smooth. It also follows that if CC has degree δ\delta, then CC contains δ1\delta-1 points of Zk¯Z_{\bar{k}}.

The value of δ\delta follows from the matrix description in [Isk96, Theorem 2.6] of the action of each link in the classification on the Picard-Manin space. For instance for 6366\leftarrow 3\to 6, the description in [Isk96, Theorem 2.6(ii), KX2=6K_{X}^{2}=6, (c)] implies that D=2KXD=-2K_{X}, which shows that δ=(2KX2)/3=4\delta=(-2K_{X}^{2})/3=4. The same argument works for other links. \Box

We first consider symmetric links where the centers of the blow up and the blow down are isomorphic, so that they are covered by Lemma 5.7(1):

  • 9399\leftarrow 3\rightarrow 9: the first map blows up a Galois orbit of six points, and the second one contracts the Galois orbit of the proper preimages of six conics (in the classical sense: they have degree 6 according to Definition 4.6) passing through five of the points by Claim 5.8; these two Galois orbits are isomorphic.

  • 8488\leftarrow 4\rightarrow 8: we blow up a Galois orbit of four general points on XX and contract the Galois orbit of the proper preimages of quartic curves passing through three of the four points by Claim 5.8; these two Galois orbits are isomorphic.

Finally we need to deal with the following links (we list them up to inverses):

  • 8358\leftarrow 3\rightarrow 5: here XX is a quadric with AX=[Z2]A_{X}=[Z_{2}] and XX^{\prime} is a degree five del Pezzo surface with AX=[Z5]A_{X^{\prime}}=[Z_{5}^{\prime}]. We need to show that the first map has center Z5Z_{5}^{\prime} and the second map has center Z2Z_{2}. By Claim 5.8, the center of the second map parametrizes smooth rational curves of degree 66 on XX passing through an orbit of five points. By Proposition 4.8(ii) and (iv)(dP8)(dP_{8}), the center of the second map is a subscheme isomorphic to Z2Z_{2}. The center of the first map parametrizes cubics passing through an orbit of two points on XX^{\prime}, and thus by Proposition 4.8(ii) and (iv)(dP5)(dP_{5}) is a scheme isomorphic to Z5Z_{5}^{\prime}.

  • 9459\leftarrow 4\rightarrow 5: it suffices to show that if Z5Z_{5} is the center of the first map and AX=[Z5]A_{X^{\prime}}=[Z_{5}^{\prime}], then Z5Z5Z_{5}\simeq Z_{5}^{\prime}. One can verify this directly, as in the previous link, or apply Corollary 2.13 as in the proof of Lemma 5.7(2).

  • 6366\leftarrow 3\rightarrow 6: we have AX=[Z3]+[Z2]1A_{X}=[Z_{3}]+[Z_{2}]-1 and AX=[Z3]+[Z2]1A_{X^{\prime}}=[Z_{3}^{\prime}]+[Z_{2}^{\prime}]-1. We claim that c(ϕ)=[Z3][Z3]c(\phi)=[Z_{3}^{\prime}]-[Z_{3}] and Z2Z2Z_{2}\simeq Z_{2}^{\prime}. Indeed, by Claim 5.8 the second arrow contracts the proper preimages of smooth rational curves of degree 44 passing through the Galois orbit of the points blown up by the second arrow, and the latter scheme of conics is isomorphic to Z3Z_{3} by Proposition 4.8(ii) and (iv)(dP6)(dP_{6}). The same argument with roles of XX, XX^{\prime} reversed implies that the first arrow blows up Z3Z_{3}^{\prime}. Thus c(ϕ)=[Z3][Z3]c(\phi)=[Z_{3}^{\prime}]-[Z_{3}]. By Lemma 5.6, we have μGk([Z2])=μGk([Z2])\mu_{G_{k}}([Z_{2}])=\mu_{G_{k}}([Z^{\prime}_{2}]), hence Z2Z2Z_{2}\simeq Z_{2}^{\prime} by Proposition 2.5.

  • 6466\leftarrow 4\rightarrow 6: we have Z3Z3Z_{3}\simeq Z_{3}^{\prime}, and c(ϕ)=[Z2][Z2]c(\phi)=[Z_{2}^{\prime}]-[Z_{2}] similarly to the previous case, using Proposition 4.8(ii), (iv)(dP6)(dP_{6}), and Proposition 2.5 again.

\Box

Proof of Theorem 3.5.

We first assume that XX and YY are geometrically irreducible. Composing with contractions to minimal models, and using the additivity of cc under composition, we may assume that XX and YY are minimal, hence belong to one of the classes from Theorem 4.1.

In the nongeometrically rational case and geometrically rational case with KX24K_{X}^{2}\leq 4 the result follows from Proposition 4.4. Finally, in the geometrically rational case with KX25K_{X}^{2}\geq 5 the result is Proposition 5.5.

In general, that is if the surfaces XX and YY are not geometrically irreducible, write LXL_{X} and LYL_{Y} for the fields of regular functions of XX and YY; these are finite field extensions of kk. Then ϕ\phi induces a kk-isomorphism σ:LXLY\sigma:L_{X}\to L_{Y}, which allows us to consider both XX and YY as smooth projective geometrically irreducible surfaces over L:=LXL:=L_{X} and ϕ\phi becomes a LL-birational isomorphism. This way XX and YY are restrictions of scalars of geometrically irreducible surfaces over LL (this can be thought of as the Stein factorizations for XX and YY over Spec(k){\operatorname{Spec}}(k)), and the result follows from the geometrically irreducible case considered above and Proposition 3.4(ii). \Box

5.2. Rationality centers

The following corollary tells us that the rationality center of a rational surface XX is well-defined, that is for any sequence of blow ups and blow downs connecting 2\mathbb{P}^{2} to XX, the virtual Galois set of blow up centers minus the blow down centers is independent of the choice of the birational isomorphism between them.

In the higher-dimensional case such rationality centers are not well-defined (however, see Definition 2.3 in [GS14] for a similar class in the localized Grothendieck ring).

Corollary 5.9 (of Theorem 3.5).

There exists a unique map

{Isomorphism classes of rational smooth projective surfaces}[Var0/k]\{\text{Isomorphism classes of rational smooth projective surfaces}\}\overset{\mathcal{M}}{\longrightarrow}\mathbb{Z}[{\mathrm{Var}}^{0}/k]

with the following properties:

  • (1)

    We have (X)=AX\mathcal{M}(X)=A_{X} as in Definition 5.2 for models of large degree

  • (2)

    For any birational isomorphism ϕ:2X\phi:\mathbb{P}^{2}\dashrightarrow X we have c(ϕ)=[(X)]1c(\phi)=[\mathcal{M}(X)]-1

Proof.

For any rational surface we define (X)\mathcal{M}(X) as c(ϕ)+1c(\phi)+1; by Theorem 3.5 this is independent of the choice of ϕ:2X\phi:\mathbb{P}^{2}\dashrightarrow X. As \mathcal{M} is required to satisfy [(X)]=c(ϕ)+1[\mathcal{M}(X)]=c(\phi)+1, this shows the uniqueness of \mathcal{M}. By Proposition 5.5 this is consistent with Definition 5.2. \Box

Example 5.10.

If XX is a del Pezzo surface of degree 55, by Definition 5.2, AX=[Z5]A_{X}=[Z_{5}]. Assume that XX has Picard rank one which by Lemma 5.3 is equivalent to Z5Z_{5} being irreducible. Then for any rationality construction of XX (see Example 3.8), that is a sequence of blow ups and blow downs starting with 2\mathbb{P}^{2} and ending with XX, one of the blow ups will have Z5Z_{5} as its center.

Example 5.11.

Consider the rationality question for a del Pezzo surface XX of degree d4d\leq 4. If XX is rational, then it is a result of Iskovskikh deduced from his classification of links [Isk96, Theorem 2.6] that such a surface XX can not be minimal, thus it admits a Galois orbit of disjoint (1)(-1)-curves which we can contract via some morphism ϕ:XX\phi:X\to X^{\prime} to obtain a minimal rational del Pezzo surface of degree 55, 66, 88 or 99. In each case the rationality center AXA_{X} is given in the table. As usual we write ZjZ_{j} or ZjZ_{j}^{\prime} for étale schemes of degree jj.

deg(X)\deg(X^{\prime}) c(ϕ1)c(\phi^{-1}) AXA_{X^{\prime}} AXA_{X}
55 [Z5d][Z_{5-d}] [Z5][Z_{5}^{\prime}] [Z5d]+[Z5][Z_{5-d}]+[Z_{5}^{\prime}]
66 [Z6d][Z_{6-d}] [Z2]+[Z3]1[Z_{2}^{\prime}]+[Z_{3}^{\prime}]-1 [Z6d]+[Z3]+[Z2]1[Z_{6-d}]+[Z_{3}^{\prime}]+[Z_{2}^{\prime}]-1
88 [Z8d][Z_{8-d}] [Z2][Z_{2}^{\prime}] [Z8d]+[Z2][Z_{8-d}]+[Z_{2}^{\prime}]
99 [Z9d][Z_{9-d}] 11 [Z9d]+1[Z_{9-d}]+1
Example 5.12.

Consider the two rational cubic surfaces XX, XX^{\prime} introduced in Example 2.14; they have isomorphic rational permutation Néron-Severi representations, and the associated Galois sets can not be read off from them.

However, the construction of X\mathcal{M}_{X} does determine these Galois sets from XX itself, as by Corollary 5.9 we have [Z]=X1[Z]=\mathcal{M}_{X}-1 and [Z]=X1[Z^{\prime}]=\mathcal{M}_{X^{\prime}}-1.

References

  • [AB18] Asher Auel and Marcello Bernardara. Semiorthogonal decompositions and birational geometry of del Pezzo surfaces over arbitrary fields. Proc. Lond. Math. Soc. (3), 117(1):1–64, 2018.
  • [And04] Yves André. Une introduction aux motifs (motifs purs, motifs mixtes, périodes), volume 17 of Panoramas et Synthèses [Panoramas and Syntheses]. Société Mathématique de France, Paris, 2004.
  • [BdS02] Wieb Bosma and Bart de Smit. On arithmetically equivalent number fields of small degree. In Algorithmic number theory (Sydney, 2002), volume 2369 of Lecture Notes in Comput. Sci., pages 67–79. Springer, Berlin, 2002.
  • [Bit04] Franziska Bittner. The universal Euler characteristic for varieties of characteristic zero. Compos. Math., 140(4):1011–1032, 2004.
  • [Bor18] Lev A. Borisov. The class of the affine line is a zero divisor in the Grothendieck ring. J. Algebraic Geom., 27(2):203–209, 2018.
  • [BSS11] M. Blunk, S. J. Sierra, and S. Paul Smith. A derived equivalence for a degree 6 del Pezzo surface over an arbitrary field. J. K-Theory, 8(3):481–492, 2011.
  • [BSY22] Jérémy Blanc, Julia Schneider, and Egor Yasinsky. Birational maps of severi-brauer surfaces, with applications to cremona groups of higher rank, 2022.
  • [CLKT23] Antoine Chambert-Loir, Maxim Kontsevich, and Yuri Tschinkel. Burnside rings and volume forms with logarithmic poles, 2023.
  • [EH18] Karin Erdmann and Thorsten Holm. Algebras and representation theory. Springer Undergraduate Mathematics Series. Springer, Cham, 2018.
  • [FK88] Eberhard Freitag and Reinhardt Kiehl. Étale cohomology and the Weil conjecture, volume 13 of Ergebnisse der Mathematik und ihrer Grenzgebiete (3) [Results in Mathematics and Related Areas (3)]. Springer-Verlag, Berlin, 1988. Translated from the German by Betty S. Waterhouse and William C. Waterhouse, With an historical introduction by J. A. Dieudonné.
  • [G0̈1] Lothar Göttsche. On the motive of the Hilbert scheme of points on a surface. Math. Res. Lett., 8(5-6):613–627, 2001.
  • [Gil15] Stefan Gille. Permutation modules and Chow motives of geometrically rational surfaces. J. Algebra, 440:443–463, 2015.
  • [GS14] S. Galkin and E. Shinder. The Fano variety of lines and rationality problem for a cubic hypersurface. arXiv:math/arXiv:1405.5154, 2014.
  • [Has09] Brendan Hassett. Rational surfaces over nonclosed fields. In Arithmetic geometry, volume 8 of Clay Math. Proc., pages 155–209. Amer. Math. Soc., Providence, RI, 2009.
  • [Has16] Brendan Hassett. Cubic fourfolds, K3 surfaces, and rationality questions. In Rationality problems in algebraic geometry, volume 2172 of Lecture Notes in Math., pages 29–66. Springer, Cham, 2016.
  • [(ht] Dominik (https://mathoverflow.net/users/37059/dominik). Does the Grothendieck ring of varieties contain torsion? MathOverflow. URL:https://mathoverflow.net/q/225599 (version: 2015-12-08).
  • [IS96] V. A. Iskovskikh and I. R. Shafarevich. Algebraic surfaces [ MR1060325 (91f:14029)]. In Algebraic geometry, II, volume 35 of Encyclopaedia Math. Sci., pages 127–262. Springer, Berlin, 1996.
  • [Isk79] Vasily A. Iskovskikh. Minimal model of rational surfaces over arbitrary fields. Izv. Akad. Nauk SSSR Ser. Mat., Translation in Math. USSR Izvestija, 43(1):19–43, 1979.
  • [Isk96] V. A. Iskovskikh. Factorization of birational mappings of rational surfaces from the point of view of Mori theory. Uspekhi Mat. Nauk, 51(4(310)):3–72, 1996.
  • [KKS22] Joseph Karmazyn, Alexander Kuznetsov, and Evgeny Shinder. Derived categories of singular surfaces. J. Eur. Math. Soc. (JEMS), 24(2):461–526, 2022.
  • [Kol96] János Kollár. Rational curves on algebraic varieties, volume 32 of Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge. A Series of Modern Surveys in Mathematics [Results in Mathematics and Related Areas. 3rd Series. A Series of Modern Surveys in Mathematics]. Springer-Verlag, Berlin, 1996.
  • [KS18] Alexander Kuznetsov and Evgeny Shinder. Grothendieck ring of varieties, D- and L-equivalence, and families of quadrics. Selecta Math. (N.S.), 24(4):3475–3500, 2018.
  • [KT22] Andrew Kresch and Yuri Tschinkel. Burnside groups and orbifold invariants of birational maps, 2022.
  • [Kuz19] Alexander Kuznetsov. Derived Categories of Families of Sextic del Pezzo Surfaces. International Mathematics Research Notices, 06 2019. rnz081.
  • [LL03] Michael Larsen and Valery A. Lunts. Motivic measures and stable birational geometry. Mosc. Math. J., 3(1):85–95, 259, 2003.
  • [LS21] Stéphane Lamy and Julia Schneider. Generating the plane cremona groups by involutions, 2021.
  • [LS22] Hsueh-Yung Lin and Evgeny Shinder. Motivic invariants of birational maps, 2022.
  • [Man86] Yu. I. Manin. Cubic forms, volume 4 of North-Holland Mathematical Library. North-Holland Publishing Co., Amsterdam, second edition, 1986. Algebra, geometry, arithmetic, Translated from the Russian by M. Hazewinkel.
  • [Mil80] James S. Milne. Étale cohomology, volume 33 of Princeton Mathematical Series. Princeton University Press, Princeton, N.J., 1980.
  • [NS09] Johannes Nicaise and Julien Sebag. A note on motivic integration in mixed characteristic, 2009.
  • [Par13] Ori Parzanchevski. On GG-sets and isospectrality. Ann. Inst. Fourier (Grenoble), 63(6):2307–2329, 2013.
  • [Per77] Robert Perlis. On the equation ζK(s)=ζK(s)\zeta_{K}(s)=\zeta_{K^{\prime}}(s). J. Number Theory, 9(3):342–360, 1977.
  • [Pra17] Dipendra Prasad. A refined notion of arithmetically equivalent number fields, and curves with isomorphic Jacobians. Adv. Math., 312:198–208, 2017.
  • [R1̈1] Karl Rökaeus. The class of a torus in the Grothendieck ring of varieties. Amer. J. Math., 133(4):939–967, 2011.
  • [Sch22] Julia Schneider. Relations in the Cremona group over a perfect field. Ann. Inst. Fourier (Grenoble), 72(1):1–42, 2022.
  • [Sco93] Leonard L. Scott. Integral equivalence of permutation representations. In Group theory (Granville, OH, 1992), pages 262–274. World Sci. Publ., River Edge, NJ, 1993.
  • [Ser77] Jean-Pierre Serre. Linear representations of finite groups. Springer-Verlag, New York-Heidelberg, 1977. Translated from the second French edition by Leonard L. Scott, Graduate Texts in Mathematics, Vol. 42.
  • [Sta18] The Stacks Project Authors. Stacks Project. http://stacks.math.columbia.edu, 2018.
  • [Sun85] Toshikazu Sunada. Riemannian coverings and isospectral manifolds. Ann. of Math. (2), 121(1):169–186, 1985.
  • [SZ20] Evgeny Shinder and Ziyu Zhang. L-equivalence for degree five elliptic curves, elliptic fibrations and K3 surfaces. Bull. Lond. Math. Soc., 52(2):395–409, 2020.
  • [Wei13] Charles A. Weibel. The KK-book, volume 145 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, 2013. An introduction to algebraic KK-theory.
  • [Xie21] Fei Xie. Derived categories of quintic del Pezzo fibrations. Selecta Math. (N.S.), 27(1):Paper No. 4, 32, 2021.
  • [Zak17] Inna Zakharevich. The annihilator of the Lefschetz motive. Duke Math. J., 166(11):1989–2022, 2017.
  • [Zim18] Susanna Zimmermann. The Abelianization of the real Cremona group. Duke Math. J., 167(2):211–267, 2018.