95
FROBENIUS-STABLE LATTICES IN RIGID COHOMOLOGY OF CURVES vorgelegt von Diplom-Mathematiker und Diplom-Informatiker MORITZ MINZLAFF Karlsruhe Von der Fakult¨ at II – Mathematik und Naturwissenschaften der Technischen Universit¨ at Berlin zur Erlangung des akademischen Grades Doktor der Naturwissenschaften Dr. rer. nat. genehmigte Dissertation Promotionsausschuss: Vorsitzender: Prof. John Sullivan, Ph. D. Berichter: Prof. Dr. rer. nat. Florian Heß Berichter: Prof. Dr. rer. nat. Remke Kloosterman Tag der wissenschaftlichen Aussprache: 8. M¨ arz 2013 Berlin 2013 D83

Frobenius-stable lattices in rigid cohomology of curves€¦ · case the curve is a ne, the second cohomology group vanishes. To see this, use, e.g. the comparison theorems with crystalline

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

  • FROBENIUS-STABLE LATTICES INRIGID COHOMOLOGY OF CURVES

    vorgelegt vonDiplom-Mathematiker und Diplom-Informatiker

    MORITZ MINZLAFFKarlsruhe

    Von der Fakultät II – Mathematik und Naturwissenschaftender Technischen Universität Berlin

    zur Erlangung des akademischen GradesDoktor der Naturwissenschaften

    Dr. rer. nat.

    genehmigte Dissertation

    Promotionsausschuss:

    Vorsitzender: Prof. John Sullivan, Ph. D.Berichter: Prof. Dr. rer. nat. Florian HeßBerichter: Prof. Dr. rer. nat. Remke Kloosterman

    Tag der wissenschaftlichen Aussprache: 8. März 2013

    Berlin 2013

    D83

    MAILTO:[email protected]

  • Document History

    Version Remarks Date

    1.1 Accepted version; minor revisions 2013-04-06

    1.0 Submitted version; Section 2.2 and Section 2.3 finished,Proposition 3.13 fixed, minor revisions

    2012-12-18

    0.9 Complete and proofread draft; except for Section 2.2and Section 2.3

    2012-05-31

    Please send comments and corrections to me at [email protected].

    mailto:[email protected]

  • Ich bedanke mich bei Florian Heß für den spannenden Themenvorschlag und beiRemke Kloosterman für die

    ”vor Ort“-Betreuung seit Florian Hess nicht mehr in

    Berlin war. Desweiteren bedanke ich mich bei der Berlin Mathematical Schoolsowie meinen Eltern Angelika und Volker Minzlaff für die finanzielle Unterstützungwährend der Anfertigung dieser Doktorarbeit. Anja Bewersdorff, Tanja Fagel, Do-minique Schneider, Mariusz Szmerlo und Nadja Wisniewski vom BMS One-StopOffice möchte ich dafür danken, dass jegliche administrativen Gänge so angenehmwaren, wie man es sich nur wünschen kann. Viel Dank gilt auch John Sullivan fürdie kurzfristige Übernahme des Vorsitzes im Promotionsausschuss.

    Bei Claus Diem, Felix Fontein, Frank Herrlich und Gabriela Weitze-Schmithüsen,Kiran Kedlaya, Thorsten Lagemann, Alan Lauder und Jan Tuitman, George Walkerund Stefan Wewers bedanke ich mich für Einladungen zu Seminarvorträgen und dieZeit verschiedene Aspekte meiner Arbeit mit mir zu diskutieren. Ganz besondersdanke ich Nils Bruin dafür, daß er mir einen mehrmonatigen Besuch bei ihm an derSimon Fraser University in British Columbia ermöglicht hat.

    Für das Korrekturlesen und für Verbesserungsvorschläge bedanke ich mich bei FabianJanuszewski, Stefan Keil, Thorsten Lagemann, Fabian Müller, Christopher Özbek undOsmanbey Uzunkol. Christina Neuhaus schließlich gebührt ein ganz großer Dank fürdie moralische Unterstützung auf der Zielgeraden dieser Arbeit.

  • Attribution-NonCommercial-ShareAlike 3.0 Germany

    This work by Moritz Minzlaff is licensed under the Creative Commons Attribution-NonCommercial-ShareAlike 3.0 Germany License. To view a copy of this license,visit http://creativecommons.org/licenses/by-nc-sa/3.0/de/ or send a letterto Creative Commons, 171 Second Street, Suite 300, San Francisco, CA, 94105, USA.

    You are free:

    To Share To copy, distribute and transmit the work.

    To Remix To adapt the work.

    Under the following conditions:

    Attribution You must attribute the work in the manner specified by theauthor or licensor (but not in any way that suggests thatthey endorse you or your use of the work).

    Noncommercial You may not use this work for commercial purposes.

    Share Alike If you alter, transform, or build upon this work, you maydistribute the resulting work only under the same or similarlicense to this one.

    With the understanding that:

    Waiver Any of the above conditions can be waived if you get permission fromthe copyright holder.

    Other Rights In no way are any of the following rights affected by the license:Your fair dealing or fair use rights; the author’s moral rights;rights other persons may have either in the work itself or in howthe work is used, such as publicity or privacy rights.

    Notice For any reuse or distribution, you must make clear to others the li-cense terms of this work. The best way to do this is with a link tohttp://creativecommons.org/licenses/by-nc-sa/3.0/de/deed.en.

    xmlns:cc="http://creativecommons.org/ns#" property="cc:attributionName"http://creativecommons.org/licenses/by-nc-sa/3.0/de/http://creativecommons.org/licenses/by-nc-sa/3.0/de/deed.en

  • CONTENTS

    1. Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.1. State of the art. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21.2. This work. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41.3. Notations and conventions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

    2. Lifting birational equivalence classes of curves. . . . . . . . . . . . . . . . . . . . . . . . . 72.1. Teissier’s criterion for lifting classes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92.2. The equimultiple local Hilbert functor. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142.3. Computing equimultiple proper liftings. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

    3. An explicit description of Frobenius-stable lattices. . . . . . . . . . . . . . . . . . . . 313.1. De Rham cohomology of smooth proper pairs. . . . . . . . . . . . . . . . . . . . . . . . . . . . 333.2. Integral Monsky-Washnitzer cohomology. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45

    4. Computing the absolute Frobenius action in larger characteristic . . . 514.1. A basis for the cohomology. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 534.2. Approximating the absolute Frobenius on differentials. . . . . . . . . . . . . . . . . . . 554.3. Reducing differentials modulo exact differentials. . . . . . . . . . . . . . . . . . . . . . . . . 574.4. Correctness and complexity analysis. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

    5. Implementations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 695.1. Computing equimultiple proper liftings. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 695.2. Computing the absolute Frobenius action in larger characteristic. . . . . . . . . 70

    A. Curves over complete discrete valuation rings. . . . . . . . . . . . . . . . . . . . . . . . . 73

    Bibliography. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77

    Index. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85

  • CHAPTER 1

    INTRODUCTION

    It is probably fair to say that a major part of mathematics is ultimately motivated

    by the desire to solve equations or to study their sets of solutions. The present thesis

    is no exception. Its guiding problem is that of efficiently counting the number of

    solutions to bivariate polynomial equations

    f(x, y) = 0

    over finite fields. Formulated from the viewpoint of algebraic geometry, the problem

    is that of efficiently counting the number of rational points on plane curves. More

    generally, let X be any curve over a finite field. If N1, N2, N3, . . . are the numbers of

    rational points of X over the extensions of the finite field of degrees 1, 2, 3, . . ., then

    the zeta function of X is defined as

    ZX(t) = exp

    ( ∞∑r=1

    Nrrtr

    ).

    It is uniquely determined by N1, . . . , Nm whenever m is large enough. For example,

    if X is smooth, proper, geometrically integral, then one may take m equal to the

    genus [Sti09, Cor. V.1.17]. Conversely, the zeta function is a rational function in t

    and if αi are the reciprocal roots of the numerator and βj the reciprocal roots of the

    denominator, then one recovers the number of points as Nr =∑j β

    rj −∑i α

    ri . A more

    detailed account of these statements is given by Wan [Wan08]. The aim of this thesis

    is to improve the state of the art of computing zeta functions.

    The zeta function has a cohomological interpretation: Let X be a smooth, geomet-

    rically integral curve over the finite field k. Its rigid cohomlogy H•rig(X) consists of

    finite dimensional vector spaces over the fraction field K of the ring of Witt vec-

    tors W (k) [Ber97, Thm. 3.1]. The Frobenius morphism F on X induces linear maps

  • 2 CHAPTER 1. INTRODUCTION

    on those spaces and

    ZX(t) =det(id− Ft | H1rig(X)

    )det(

    id− Ft | H0rig(X))

    det(

    id− Ft | H2rig(X))

    whenever X is proper. If X is affine, then the relation is

    ZX(t) =det(id−#k · F−1t | H1rig(X)

    )det(

    id−#k · F−1t | H0rig(X))

    det(

    id−#k · F−1t | H2rig(X)) .

    The only interesting map in the above determinants is the one on the first cohomology

    group. Indeed, in both cases the zeroeth cohomology group is 1-dimensional and the

    Frobenius acts as the identity. If the curve is proper, then the second cohomology

    group is also 1-dimensional and the Frobenius action is multiplication by #k. In

    case the curve is affine, the second cohomology group vanishes. To see this, use, e.g.

    the comparison theorems with crystalline cohomology and with Monsky-Washnitzer

    cohomology [Ber97, Prop. 1.9 & 1.10] and the properties of the respective cohomolo-

    gies [Ber74, MW68, vdP86].

    1.1. State of the art

    The task of computing the Frobenius action on H1rig(X) has received a lot of attention

    over the past few years. First came Kedlaya’s algorithm for hyperelliptic curves in

    odd characteristic [Ked01] which has been generalised several times to superelliptic

    curves [GG01], Ca,b-curves [DV06a], and nondegenerate curves [CDV06]. Further-

    more there are algorithms for hyperelliptic curves in even characteristic [DV06b] and

    smooth plane curves (as special case of smooth hypersurfaces) [AKR11].

    More recent are algorithms that add a deformation step: The given curve is first

    embedded in a family of curves with a particular “nice” fibre. The Frobenius action

    on the cohomology of this fibre is then computed and finally transformed to give the

    action on the cohomology of the original curve. By now this has been applied to most

    classes of curves mentioned above [Ger07, Hub07, Hub08, CHV08, Tui11].

    Related to the approach with a deformation step is the approach that uses a fibration

    on X: Computing the Frobenius action for a given variety is reduced to the same prob-

    lem for a smooth hyperplane section. This method was introduced by Lauder [Lau06]

    and improved by Walker [Wal09]. In the case of curves, Walker focuses on smooth

    plane curves and on Ca,b-curves.

    Typically the runtimes of these algorithms are polynomial in the genus g of the curve

    and the degree n of k, but linear or worse in the characteristic p. In particular, it

  • 1.1. STATE OF THE ART 3

    is exponential in log(p). For example, the time complexities of the algorithms for

    hyperelliptic, superelliptic, and Ca,b-curves (without deformation or fibration) lie in

    Õ(p1n3g5).

    To contrast, if X is proper, then O(log(p)ng2) bits suffice to determine X [Hes,

    Thm. 56] and to write down the determinant of 1−Ft on H1rig(X) [Sti09, Thm. V.1.15].Using deformation can yield a lower exponent for n as demonstrated in case of hyper-

    elliptic curves. Regarding p, Harvey describes an algorithm for hyperelliptic curves

    in odd characteristic which improves the dependence on the characteristic to Õ(p0.5)

    at the cost of higher exponents for the other parameters [Har07].

    Fixing the characteristic, the most general class of curves for which a practical al-

    gorithm exists arguably is the class of nondegenerate curves. If the cardinality of k

    is large enough, then this class includes all Ca,b-curves and thus all superelliptic and

    hyperelliptic curves. Nonetheless, nondegenerancy is quite special: The moduli space

    of nondegenerate genus g curves has dimension 2g + 1 whereas the moduli space of

    all genus g curves has dimension 3g − 3 [CV09].(1)

    Further algorithms. — If the only goal is to compute the zeta function and not

    the Frobenius action on rigid cohomology, there are further methods available. Most

    notably are those using “Dwork cohomology”, étale (or `-adic) cohomlogy, or the

    canonical lift of the Jacobian. Currently the most general class of curves these meth-

    ods can practically handle are ordinary hyperelliptic curves in even and superelliptic

    curves in any characteristic.

    Lauder and Wan proposed an algorithm based on Dwork cohomology for arbitrary

    affine (not necessarily smooth) varieties. It runs in polynomial time when the char-

    acteristic is kept fixed [LW08], but is deemed not very practical [LW04, p. 332].

    Related algorithms for superelliptic curves [Lau03], Artin-Schreier curves [LW04],

    and smooth hypersurfaces with a deformation step [Lau04] have time complexity

    comparable to those using rigid cohomology. In particular, they share the bad be-

    haviour with respect to log(p).

    Schoof described an algorithm for elliptic curves using étale cohomology [Sch85].

    Apart from several optimisations, there is an adaption to genus 2 [GH00] and even

    a generalisation to arbitrary genus [Pil90], but the latter runs doubly-exponential in

    the genus and is considered impractical [The04, Sec. F.3.b]

    (1)The only exceptions are genus 7 where the dimension is 16 and genera less than 4 where 3g− 3 isless than 2g + 1 and all curves are nondegenerate.

  • 4 CHAPTER 1. INTRODUCTION

    Canonical lifts of Jacobians are used in Satoh’s and Mestre’s algorithms for ordinary

    elliptic curves in odd respectively even characteristic [Sat00, Mes00]. The most gen-

    eral variants of these algorithms are for ordinary hyperelliptic curves [CL07, LL06].

    More in-depth accounts of the various approaches to compute zeta functions can be

    found in a book by Cohen et al. [CFA05] and a survey by Chambert-Loir [CL08].

    1.2. This work

    The previous section indicates that when one wants to compute the Frobenius action

    on rigid cohomology with current methods, one has to restrict to special classes of

    curves such as the class of nondegenerate curves and to “small” characteristic. This

    thesis makes three main contributions to weaken these restrictions.

    Lifting birational equivalence classes of curves. — A central ingredient to com-

    puting zeta functions with rigid cohomology is a comparison theorem with de Rham

    cohomology: Let X be a smooth, proper, geometrically integral curve and X be a

    proper lifting. If Z is a reduced strict relative normal crossing divisor on X, then

    there is a natural isomorphism

    H•dR(XK \ ZK)∼−→ H•rig(X \ Z) (1.1)

    between the rigid cohomology of the affine curve X \ Z over k and the de Rhamcohomology of the affine curve XK \ ZK over K (see (3.2)). For the types of curvesmentioned in the previous section a proper lifting is easily found. Indeed, each of

    those curves is a smooth hypersurface in some ambient surface which is obviously the

    reduction of some proper, integral scheme over the ring of Witt vectors: For smooth

    plane curves one may take the projective plane as the ambient scheme, for hyper-

    elliptic and Ca,b-curves a weighted projective plane, and for nondegenerate curves a

    toric surface. One may identify a hypersurface with an element of the homogeneous

    coordinate ring. Under this identification, lifting the hypersurface amounts to picking

    a lift of the element in the homogeneous coordinate ring of the lifted surface. The

    first contribution of this thesis is a treatment of the case when such a nice embedding

    is not known. It suffices to lift the birational equivalence class of the curve and we

    present an algorithm that approximates a lifting of the class of any curve (Chapter 2).

    An explicit description of Frobenius-stable lattices. — Much like lifting a

    root modulo p of a polynomial over the integers in general only yields an element

    of the p-adic integers, it seems that one cannot expect a proper lifting of X to be

    defined over the integers or a finite extension of them. Rather, one has to consider

    the ring of Witt vectors W (k). A Frobenius action on lattices over W (k) is therefore

    of interest for two reasons: When one is only able to write down an approximation of

  • 1.3. NOTATIONS AND CONVENTIONS 5

    a lifting over Witt vectors WN (k) of finite length, one cannot write down the generic

    fibre of the exact lifting as needed in (1.1). In this case, a Frobenius-stable lattice

    over W (k) naturally yields a WN (k)-module on which to approximate the Frobenius

    action. Another reason is that even in the case when one does obtain an exact lifting,

    if the chosen basis of H1rig(X \ Z) does not span a Frobenius-stable lattice, then thematrix representing the Frobenius will not have coefficients in W (k). Since we can

    only compute with Witt vectors of finite length, this causes a loss of precision. The

    second contribution of this thesis is therefore an explicit description of a Frobenius-

    stable lattice in the vector space H1rig(X \Z) for any choice of X and Z (Chapter 3).

    Computing the Frobenius action in larger characteristic. — The third con-

    tribution of this thesis is an extension of Harvey’s algorithm from hyperelliptic curves

    in odd characteristic to more general “Kummer coverings” of the rational line. It

    computes the Frobenius action in Õ(p0.5)-time (Chapter 4). Like Harvey’s algorithm,

    the runtime is polynomial in the degree of k and the genus.

    Finally, the effectiveness of all algorithms in this thesis is demonstrated using imple-

    mentations in Sage [S+12] and Magma [BCP97] (Chapter 5).

    Original contributions of this thesis are usually enclosed in numbered environments

    such as propositions, lemmas, and others. Conversely, to our best knowledge such

    numbered environment contain original contribution. Exceptions to this are marked

    as such. The main theorems are numbered with capital letters.

    1.3. Notations and conventions

    We denote by N the monoid of finite cardinals 0, 1, 2, . . . and by Z the ring of integers.

    Throughout this thesis, k is a finite field of characteristic p. The ring of Witt vectors

    over k is denoted by W (k), its fraction field by K, its valuation by v, and the ring

    of Witt vectors of length N by WN (k) = W (k)/pNW (k). Liu’s book “Algebraic

    Geometry and Arithmetic Curves” [Liu06] will be our default reference for algebraic

    geometry.

    Objects over k. — Objects over k usually come with an overline · in their notation,e.g. we write X for a scheme over k. We also sometimes use this notation to denote

    the base change of an object over W (k) to k, e.g. given a W (k)-scheme X, we might

    write X for the special fibre Xk.

    Frobenius morphisms. — Let X be a scheme over k of finite type. The morphism

    Fp : X → X given by the identity on the topological space of X and p-th powering onOX is called the absolute Frobenius morphism on X. Unless k is the prime field Fp, Fp

  • 6 CHAPTER 1. INTRODUCTION

    is not a k-morphism. In general, let n be the degree of k. Then we define the (k-linear

    or relative) Frobenius morphism F as the n-th iterate of the absolute Frobenius, i.e.

    F = Fnp . This is the Frobenius that was already mentioned earlier in this introduction.

    Curves. — Let S be a locally Noetherian scheme. A curve over S is a flat, separated

    scheme of finite type over S whose fibres are geometrically reduced and equidimen-

    sional of dimension 1. Curves are proper unless otherwise mentioned. A curve over S

    is plane if it is a hypersurface in P2S . In general, we say that a curve X has a certain

    property whenever X has that property as a scheme.

    The arithmetic genus of a curve X over a field is pa(X)df= 1−χ(OX). If the normali-

    sation X̃ of X is smooth, then the genus g(X) of X is the geometric genus of X̃, i.e.

    g(X)df= h0(ΩX̃).

    Deformations. — Our main references for deformation theory and its terminology

    are Hartshorne’s “Deformation Theory” [Har10] and Sernesi’s “Deformations of Al-

    gebraic Schemes” [Ser06]. For example, the deformation of a scheme X over k to a

    local Noetherian ring A with residue field k is a scheme X that is flat and of finite

    type over A together with an isomorphism of its special fibre with X. We use the

    notion of a deformation and a lifting interchangeably, in particular when we think of

    lifting objects from characteristic p to characteristic 0.

    Abbreviations. — Given a sheaf F on a scheme X, we will usually write Hi(F )

    for Hi(X,F ). Similarly, if X is understood to be a scheme over another scheme S, we

    will write ΩX for ΩX/S and HidR(X) instead of H

    idR(X/S). Further abbreviations like

    these may be found throughout the thesis and hopefully will not lead to confusion.

    A bold letter usually denotes a tuple or sequence, e.g. xmight be a sequence x1, . . . , xrof variables or α might be a tuple (α1, . . . , αr) ∈ Nr of finite cardinals. In the lattercase, we also have |α| = α1 + . . .+ αr.

    Complexity. — In addition to the usual Oh notation to measure the asymptotic be-

    haviour of functions, e.g. n2 +n ∈ O(n2), we also use the soft Oh notation [vzGG03,Sec. 25.7] that neglects polylogarithmic factors, e.g. n(log n)2 ∈ Õ(n).

  • CHAPTER 2

    LIFTING BIRATIONAL EQUIVALENCE CLASSES OF

    CURVES

    This chapter addresses the issue of computing a proper lifting to W (k) of a smooth

    curve X over k up to birational equivalence. Our motivation to study this problem

    arose in the context of describing the p-adic cohomology of open affine subschemes

    of X (Chapter 3). Deformation theory tells us that such a proper lifting always

    exists [Har10, Thm. 22.1] and our main result will be to turn this existence statement

    into an effective algorithm.

    Finding a lifting is easy when the curve is embedded in projective space as com-

    plete intersection: Pick any homogeneous liftings of its defining equations [Har10,

    Thm. 9.4]. But while every smooth curve over k can be embedded already in P3` for

    some finite extension ` of k, in general it cannot be embedded as smooth complete

    intersection in P3` . It is conjectured that in general an embedding in P3` does not

    lift [EH99, Conj. C].

    One idea to circumvent this problem is to embed curves as complete intersections

    in more general ambient schemes. This is done by most algorithms that compute

    zeta functions with p-adic cohomology. However, this approach also seems to have

    limitations: For example, the step from hyperelliptic curves (hypersurfaces in weighted

    projective planes) to nondegenerate curves (hypersurfaces in toric surfaces) only gives

    a modest increase in the dimension of the moduli space. Beyond toric surfaces, there is

    (to us) no clear candidate for ambient schemes that are suitable from a computational

    point of view.

    Therefore, we take a different approach: We restrict to the simplest ambient scheme,

    the projective plane, but allow singularities in a controlled fashion. In particular, we

    allow nodes so that X is always birational to such a curve [Har97, Thm IV.3.10]

    over some finite extension of k. The idea is to lift the plane curve along with its

    singularities.

  • 8 CHAPTER 2. LIFTING BIRATIONAL EQUIVALENCE CLASSES OF CURVES

    For this chapter, we use a more general definition for k and K: Let Λ

    be a complete discrete valuation ring with maximal ideal m, fraction field

    K, and perfect residue field k. Set ΛNdf= Λ/mNΛ and denote by S the

    spectrum of Λ.

    This chapter is organised as follows: In Section 2.1, we formally define liftings of

    birational equivalence classes of curves. We will see that a proper lifting X of a

    representative may fail to lift the class. The solution is an adaption of Teissier’s

    δ-invariant criterion to our situation: A proper lifting X lifts the class if and only

    if the δ-invariants of its fibres coincide (Proposition 2.5). For plane curves whose

    singularities are all ordinary, e.g. nodes, the δ-invariant is completely determined by

    the multiplicities of the singularities and the degrees of their underlying points. An

    “equimultiple proper lifting” of a plane ordinary curve therefore solves the problem of

    lifting its birational equivalence class (Proposition 2.11).

    The definition of an“equimultiple”local Hilbert functor (Definition 2.15) in Section 2.2

    frames the problem of computing such liftings in the context of formal deformation

    theory. Finally in Section 2.3, we explain how the infinitesimal lifting property of

    smooth schemes can be used to compute equimultiple proper liftings of plane curves

    over k to ΛN (Algorithm 1). We use the formal deformation theory introduced in the

    previous section to show the algorithm is successful whenever the plane curve has only

    nodes and not too many ordinary singularities of higher multiplicity (Corollary 2.27).

    Most results of this chapter are classical knowledge when one takes Λ as the ring of

    formal power series over the complex numbers. We are not aware, however, of any

    references for the general case.

    Further notation and terminology. — Let Set be the category of sets. The

    category of Artinian local Λ-algebras with residue field k is denoted by Ar. We write

    mA for the the maximal ideal of a local ring A. The category of complete Noetherian

    local Λ-algebras A such that A/miA is in Ar, all i, will be denoted by Âr. A surjection

    A′ � A in Ar is a small extension (of A by I) if the kernel I is annihilated by mA′ .The kernel has thus the structure of a k-vector space. We write Ex(A, I) for the set

    of small extensions of A by I. One can give Ex(A,−) the structure of a covariantfunctor. In case of a surjection I � J of finitely generated k-vector spaces with kernelC, this functor maps the extension A′ � A to A′/C � A.

  • 2.1. TEISSIER’S CRITERION FOR LIFTING CLASSES 9

    2.1. Teissier’s criterion for lifting classes

    The (birational equivalence) class of a curve over a field contains a unique-up-to-

    isomorphism normal curve.(1) Consequently, we make

    Definition 2.1. — A lifting of a class [X] of a curve X over k to Λ is a class of

    curves over Λ that contains a proper lifting of the normal curve of [X].

    While a proper lifting Y of a normal curve over k is a normal curve over Λ, there may

    be several nonisomorphic normal curves in its class. Nonetheless, the generic fibres of

    all normal curves in [Y ] are isomorphic and thus define the same class [YK ] over K.

    The existence of a unique representative in [Y ] is not relevant to our goals; we refer

    the interested reader to Liu’s book [Liu06, Chp. 9 & 10].

    Example 2.2. — Let X be a curve over k. By definition, a proper lifting of the nor-

    mal curve in [X] also lifts the class. However, given any representative of [X], a proper

    lifting does not need to lift the class. Take, for example, the geometrically integral

    genus 1 curve X defined by y2 = x3(x−1)(x−2), where the characteristic of k does notdivide 6. Assume that K is perfect. The equation y2 = x(x−m)(x+m)(x−1)(x−2),m ∈ m nonzero, defines a proper lifting X to Λ. Let Y be a proper lifting of the nor-mal curve in [X]. Then Y is smooth since it is a proper lifting of a normal, hence

    smooth, curve over k. So the genera of Yk and YK must coincide and be equal to 1.

    But the genus of XK is 2, so Y and X cannot be birational.

    The general case. — Let X be a curve over Λ. We want to know when [X] lifts

    [Xk]. The above example suggests that the genera of the fibres of X should be equal.

    As it turns out, that is the exact criterion except that when Xk is not irreducible we

    replace the genera by the arithmetic genera of the normalisations of the fibres. In

    this subsection, we follow a paper by Diaz and Harris which treats curves over the

    complex numbers [DH88, Sec. 2].

    Lemma 2.3. — Let X be a curve over Λ. Its normalisation π : Y → X is a finitemorphism of curves over S.

    Proof. — Since X is of finite type over Λ, X is excellent [Gro65, Sch. 7.8.3]. So

    its normalisation π is finite [Gro65, Prop. 7.8.6]. Moreover, the structure map

    Y → X → S is flat: A reduced scheme over Λ is flat if and only if the generic pointof each irreducible component maps to the generic point of S [Gro66, Prop 14.5.6].

    Since X is flat over Λ and since the normalisation induces a bijection between generic

    points of Y and X, Y is flat over Λ.

    (1)In contrast to the definition in Liu’s book [Liu06, Def. 4.1.1], we do not assume that a normal

    scheme is irreducible.

  • 10 CHAPTER 2. LIFTING BIRATIONAL EQUIVALENCE CLASSES OF CURVES

    Hence, Y is a proper lifting of Yk and it remains to show that Yk is a curve. Since

    Y → X is the normalisation, the normalisation of the special fibre Xk must fac-tor through Yk → Xk and Yk must be equidimensional of dimension 1. To showthat Yk is reduced (this implies Yk geometrically reduced as k is perfect), consider a

    point P ∈ Yk. Its local ring is reduced if and only if the zero ideal is radical. Let0 = ∩p∈Ass(OYk,P )q(p) be a reduced primary decomposition, i.e. q(p) is p-primary.Since Y is normal and of dimension 2, it is Cohen-Macaulay [Liu06, Cor. 8.2.22] and

    so is Yk. Hence Ass(OYk,P ) is the set of minimal primes [Liu06, Prop. 8.2.15]. Let Z

    be the irreducible component of Yk passing through P and corresponding to p. Let

    ξ be its generic point. The image π(ξ) is a generic point of Xk. Since Xk must be

    regular along an open subset of π(Z), Xk must be regular at π(ξ). Hence, X must be

    regular at π(ξ). Thus, OY,ξ and OX,π(ξ) are isomorphic since π is the normalisation.

    In particular, OYk,ξ is isomorphic to OXk,π(ξ) and thus reduced. So q(p) equals p. This

    argument applies to all minimal primes p of OYk,P and so its zero ideal is radical.

    Let X be a curve over Λ and s ∈ S. Let π(s) : X̃s → Xs be the normalisation of thefibre over s. Denote by SXs the quotient sheaf of π(s)∗OX̃s by OXs , i.e. there is the

    short exact sequence

    0→ OXs → π(s)∗OX̃s → SXs → 0. (2.1)

    One defines the δ-invariant of Xs as

    δ(Xs)df= h0(SXs).

    The δ-invariant of a closed point P of Xs is δP (Xs)df= lengthOXs,P (SXs,P )

    = [k(P ) : k(s)]−1 dimk(k) SXs,P . Since SXs is a skyscraper sheaf, there is an equality

    δ(Xs) =∑P∈X0s

    [k(P ) : k(s)]δP (Xs).

    The sequence (2.1) gives an identity χ(π(s)∗OX̃s) = χ(OXs) + χ(SXs) [Liu06,

    Lem. 7.3.16]. Since χ(π(s)∗OX̃s) and χ(OX̃s) are equal, one arrives at

    pa(Xs) = pa(X̃s) + δ(Xs). (2.2)

    Lemma 2.4. — Let X be a curve over Λ. Let X̃k → Xk and X̃K → XK be thenormalisations of its fibres. Then pa(X̃k) is at most pa(X̃K).

    Proof. — Let Y → X be the normalisation of X. Then X̃K equals YK and the nor-malisation X̃k → Xk factors through Yk → Xk. Since Y is flat over Λ by Lemma 2.3,the arithmetic genera of its fibres coincide. Hence, the genus formula (2.2) gives

    pa(X̃K) = pa(YK) = pa(Yk) = pa(X̃k) + δ(Yk). This proves the claim.

    Proposition 2.5. — Let X be a curve over Λ. Let π : Y → X, π(k) : X̃k → Xk, andπ(K) : X̃K → XK be the normalisations. The following are equivalent:

  • 2.1. TEISSIER’S CRITERION FOR LIFTING CLASSES 11

    (1) The δ-invariants δ(Xk) and δ(XK) coincide,

    (2) The arithmetic genera pa(X̃k) and pa(X̃K) coincide,

    (3) the fibres of π are π(k) and π(K), and

    (4) the class [X] lifts the class [Xk].

    Proof. — The genus formula (2.2) implies that (1) and (2) are equivalent. Now

    fix a very ample sheaf OX(1) on X and let s ∈ S. As in the lead-up tothe genus formula, tensoring (2.1) with OXs(n) = OX(n)|Xs yields equalitiesχ(π(s)∗OX̃s(n)) = χ(OXs(n)) + χ(SXs(n)) for all n ∈ N. In terms of Hilbertpolynomials this means P (π(s)∗OX̃s) = P (OXs) + P (SXs). Note that SXs is a

    skyscraper sheaf, so χ(SXs(n)) equals δ(SXs) for all n ∈ N. Since X is flat over Λ,P (OXs) is independent of s. Therefore (1) holds if and only if the Hilbert polynomial

    P (π(s)∗OX̃s) is independent of s. In this formulation, the equivalence of (1) and (3)

    was proven by Chiang-Hsieh and Lipman [CHL06, Cor. 3.4.2]. (In the context of

    complex analytic spaces the result goes back to Teissier [Tei80].)

    To finish, we will show that (3) implies (4) implies (2). Assume (3) holds. By

    Lemma 2.3, Y is a proper lifting of Yk = X̃k. Hence, [X] lifts [Xk] and (4) holds.

    Assuming (4), one can now argue as in Example 2.2: If [X] contains a proper lifting Z

    of X̃k, then Z is normal. In particular, Zs is isomorphic to X̃s for both points s of S.

    Since Z is flat over Λ, the arithmetic genera of its fibres agree, i.e. (2) is satisfied.

    Definition 2.6. — A curve over Λ is equinormalisable if it satisfies any of the equiv-

    alent properties in Proposition 2.5.

    Example 2.7. — In the later sections of this chapter, we will use the δ-invariant to

    lift classes of curves. Hurwitz’s theorem gives another approach to finding a lifting:

    Let Y be a curve over Λ with normal, geometrically integral fibres. Moreover, let

    φ : Y → P1S be a finite morphism and s be a point of S. Hurwitz relates the arithmeticgenus of Ys to the degree a(s) of φs and the degree of the ramification divisor (or

    different) Rφs of φs:

    2pa(Ys) = 2− 2a(s) + degRφs .Now consider the geometrically integral curve X ⊂ P2k defined by ya = h(x) where thecharacteristic of k does not divide a and h(x) is a polynomial over k. For simplicity,

    assume that a and deg h are coprime. Let Y be the normalisation of X. The function x

    defines a finite morphism φ : Y → P1k of degree a. Let h =∏i h

    rii be the factorisation

    into irreducibles and si be the greatest common divisor of a and ri, each i. The degree

    of Rφ is∑i(a− si) deg hi [Sti09, Sec. 3.7]. So for each hi, choose a degree-preserving

    lifting hi ∈ Λ[x]. Each hi is necessarily irreducible. The equation ya =∏i h

    rii defines

    a proper lifting X of X to Λ. Let Y be its normalisation. The rational function x

    defines a finite, flat morphism φ : Y → P1Λ of degree a whose special fibre is φ. Forsimplicity, assume that K is perfect. The analysis of the ramification divisor of the

  • 12 CHAPTER 2. LIFTING BIRATIONAL EQUIVALENCE CLASSES OF CURVES

    generic fibre φK is completely analogous to above, i.e. its degree is∑i(a− si) deg hi.

    By Hurwitz, the arithmetic genera of Y and YK agree and since YK is a normalisation

    of XK , [X] lifts [X].

    Remark 2.8. — Except potentially for “special” cases, all geometrically integral

    curves X in Prk have an equinormalisable proper lifting in PrΛ: The normalisation

    π : Y → X corresponds to a closed immersion φ : Y → Prk and thus to a base-pointfree linear system on Y . Set L = φ∗OPrk(1). The linear system is given by a subspace

    V of H0(L ). Choose a proper lifting Y of Y to Λ and an invertible sheaf L on Y that

    restricts to L . If L is acyclic, then so is L and the natural map H0(L )→ H0(L ) issurjective. The inverse image V of V under this map defines a base-point free linear

    system. Indeed, let P ∈ Y . Then LP ⊗ k equals L P . Since L is generated by Vat P , L is generated by V at P by Nakayama. Now the set Z of points at which

    L is not generated by V is closed. As Y is proper over Λ, Z must be empty (oth-

    erwise it would intersect Y ). Hence, V and L define a morphism φ : Y → PrΛ. Byconstruction φ lifts φ and induces a finite morphism π : Y → X = φ(Y ) that lifts π.If ξ denotes the generic point of Y , then π(ξ) is the generic point of X and OY,ξ is a

    finite extension of OX,π(ξ). In particular, the genus of XK is at most YK . This gives

    g(XK) ≤ g(YK) = g(Yk) = g(Xk), so by Lemma 2.4, the genera of the fibres of Xcoincide, i.e. X is equinormalisable.

    Example 2.9. — Let X ⊂ P2k be a geometrically integral curve of degree d andgenus g. With the notation of the above remark, the sheaf L is acyclic whenever d is

    at least 2g−1. The genus formula (2.2) for X is (d−1)(d−2)/2 = g+δ(X). So d is atleast 2g−1 if and only if δ(X) is at least 2g2−6g+3. In other words, any sufficientlysingular geometrically integral plane curve over k admits an equinormalisable proper

    lifting in P2Λ.

    The case of plane ordinary curves. — The δ-invariant of a plane curve whose

    singularities are all ordinary and whose underlying points are rational is completely de-

    termined by the multiplicities of its singularities. Similar to Wahl’s [Wah74b, Sec. 1]

    and Markwig’s [Mar07, Paper V] definitions over the complex numbers, we will now

    define equimultiple proper liftings of a plane curve X over k. When all singularities

    of X are ordinary and have residue field k, such liftings are equinormalisable.

    Let F be a field and U ⊆ P2F be open. Let X ⊆ U be a curve and P be an F -rationalpoint of P2F with maximal ideal mP . The multiplicity of X at P is the largest integer

    m such that mmP contains the ideal OU (−X)P . The point P lies on X if and onlyif the multiplicity is at least 1. If the multiplicity of X at P is exactly 1, then the

    local ring OX,P is regular. If it is greater than 1, then P is called a multiple point or

    singularity of X. To generalise these notions to curves over an algebra A of Âr, note

    that the F -rational point P corresponds to an F -section of U . So now let U ⊆ P2Λ be

  • 2.1. TEISSIER’S CRITERION FOR LIFTING CLASSES 13

    open and σ be an A-section of U . Then σ : SpecA → U is a closed immersion withideal sheaf Iσ. We define the multiplicity mσ(X) of a curve X ⊆ UA along σ as

    mσ(X)df= max{n ∈ N | OP2A(−X) ⊆ I

    nσ }.

    From now on we assume that the underlying point of a singularity is

    rational.

    Definition 2.10. — Let A be an object of Âr, U ⊆ P2Λ be open, and σ ∈ U(A). Aplane curve X over A is equimultiple along σ (of multiplicity m) if mσk(Xk) = mσ(X)

    (= m). The curve X is called equimultiple if there are sections σ1, . . . , σe ∈ U(A) suchthat: (1) X is equimultiple along σi for each i = 1, . . . , e, and (2) σ1, . . . , σe induce

    the k-sections that correspond to the multiple points of Xk.

    Let x0, x1, x2 be coordinates of P2Λ over Λ. The image of an A-section σ of P

    is contained in D(xi) for some i = 0, 1, 2, say for i = 0. Then its ideal sheaf Iσis trivial outside D(x0) and on D(x0) it is given by an ideal of A[x1, x2] of the form

    (x1−λ1, x2−λ2) with λ1, λ2 ∈ A. Now let X be a plane curve over A and F (x0, x1, x2)be a defining homogeneous polynomial. Then X has multiplicity m along σ if and

    only if the homogeneous parts of degrees 0, 1, . . . ,m−1 of F (1, x1 +λ1, x2 +λ2) vanishand the the degree m part does not vanish. Now let A be equal to k. Assume that

    σ is the section corresponding to a singularity P of X. We say that P is an ordinary

    singularity if the degree m part of F (1, x1 +λ1, x2 +λ2) factors into m distinct linear

    forms over k.(2) If in addition m equals 2, then P is a node of X. We call X an

    ordinary curve when all its singularities are ordinary and a nodal curve when all its

    singularities are nodes. The δ-invariant of a singularity P of multiplicity mP satisfies

    δP (X) ≥ mP (mP − 1)/2. (2.3)

    The inequality is an equality if and only if P is ordinary [Hir57, Thm. 1].

    Proposition 2.11. — Let X be a plane curve over Λ with ordinary special fibre. If

    X is equimultiple, then it is equinormalisable.

    Proof. — Lemma 2.4 and the genus formula (2.2) imply that δ(XK) is at most δ(Xk).

    Let us use the notation from Definition 2.10 and let mi be the multiplicity of X along

    σi. Each section σi gives rise to a section σi,K ∈ P2K(K) such that XK has multiplicitymi along σi,K . In other words, each σi corresponds to a singularity of multiplicity

    mi on XK . Moreover, pairwise distinct σi give pairwise distinct singularities. From

    the formula (2.3) for the δ-invariant of a singularity, we deduce that δ(XK) is at least

    (2)The definition of ordinary singularities given here is more restrictive than the general one. In

    particular, ordinarity can be defined when the multiple points are not rational. For our purposes,

    the definition given here suffices.

  • 14 CHAPTER 2. LIFTING BIRATIONAL EQUIVALENCE CLASSES OF CURVES

    δ(Xk). Hence, the δ-invariants of the fibres ofX must agree andX is equinormalisable.

    Example 2.12. — The converse to the proposition does not hold: A trivial example

    is the union X of the three lines defined by x, y, and x + y + mz in the projective

    plane over Λ, m ∈ m nonzero. Its generic fibre has three singularities, each a node, at[0 : 0 : 1], [m : 0 : 1], and [0 : m : 1]. Therefore, δ(XK) is 3. However, its special fibre

    has a unique singularity, an ordinary multiple point of multiplicity 3, at [0 : 0 : 1]. So

    δ(Xk) also equals 3. Hence, X is equinormalisable but not equimultiple.

    2.2. The equimultiple local Hilbert functor

    For this section, we fix a plane curve X of degree d over k and a finite

    set Σ ⊆ P2k(k) of sections. Let m = {mσ}σ be the multiplicities of Xalong the σ ∈ Σ.

    Due to Proposition 2.5 and Proposition 2.11, we are interested in equimultiple proper

    liftings of X to Λ. The goal of this section is to introduce and understand the following

    natural diagram of functors. The first half of this section introduces the necessary

    terminology and the functors on the top square. The bottom functors are the topic

    of the second half.

    HX� // P

    d(d+3)/2Λ

    H(X,Σ)� //

    99 99

    Pd(d+3)/2Λ ×

    ∏σ∈Σ P

    99 99

    EX?�

    OO

    E(X,Σ)� //

    ?�

    OO

    88 88

    E(d,m)?�

    OO

    (2.4)

    The three functors on the right-hand side are (the functors of sections of) projective

    schemes over S. The left-hand square is made up of prorepresentable functors of

    Artin rings. They are the “local counterparts” at X and Σ to the functors on the

    right. The functors on the top square are smooth. If A is an object of Ar and Σ is

    the set of sections corresponding to the multiple points of X, then EX(A) is the set

    of equimultiple proper liftings of X to A.

  • 2.2. THE EQUIMULTIPLE LOCAL HILBERT FUNCTOR 15

    Formal deformation theory. — A functor of Artin rings is a covariant functor

    F : Ar→ Set, #F (k) = 1.

    For example, if R is an object of Âr, then HomΛ(R,−) is a functor of Artin rings.A functor of Artin rings that is isomorphic to some HomΛ(R,−) is said to be prorep-resentable or prorepresented by R. The functor F of Artin rings is called smooth if

    F (A′) → F (A) is surjective whenever A′ → A is. If F is prorepresented by R, thenF is smooth if and only if R is a power series ring over Λ [Sch68, Prop. 2.5]. Any

    functor F of Artin rings can be extended to a functor F̂ on Âr by

    F̂ (A)df= lim←−F (A/m

    NA ).

    Let k[ε] be the ring of dual number of k, i.e. ε2 equals 0. The tangent space of F is

    defined as

    t(F )df= F (k[ε]).

    Given surjections A′ � A and A′′ � A in Ar, there is the natural map

    F (A′ ×A A′′)→ F (A′)×F (A) F (A′′). (2.5)

    Consider the following “Schlessinger’s conditions”:

    (H1) (2.5) is surjective whenever A′′ � A is small,(H2) (2.5) is bijective whenever A = k and A′′ = k[ε],

    (H3) t(F ) is a k-vector space of finite dimension, and

    (H4) (2.5) is bijective whenever A′′ � A is small.

    If F satisfies (H2), then the tangent space t(F ) has a natural structure as k-vector

    space [Sch68, Lem. 2.10]. Moreover, let A′ � A be a small extension with kernel I.There is a natural action

    F (A′)× (t(F )⊗ I)→ F (A′)×F (A) F (A′)

    of t(F ) ⊗ I on the nonempty fibres of F (A′) → F (A). If F satisfies (H1) and (H2),then the action is transitive. If F satisfies (H2) and (H4), then the action is transitive

    and free [Sch68, (2.15)]. Finally, a functor of Artin rings is prorepresentable if and

    only if it satisfies (H2), (H3), and (H4) [Sch68, Thm. 2.11].

    Example 2.13. — Let Z be a closed subscheme of Prk. The local Hilbert functor

    HZ : Ar→ Set sends an algebra A to the set

    HZ(A)df={Z ⊆ PrA | Z is a deformation of Z to A

    }.

    The functor HZ is prorepresentable [Har10, Thm. 17.1]. Let NZ/Prkbe the normal

    sheaf of the closed immersion Z ↪→ Prk. The tangent space of the local Hilbert functoris t(HZ)

    ∼= H0(NZ/Prk) [Har10, Thm. 2.4]. A particular case is that of a k-rationalpoint P of Prk. Its local Hilbert functor is smooth and can be identified with a

    subfunctor of PrΛ.

  • 16 CHAPTER 2. LIFTING BIRATIONAL EQUIVALENCE CLASSES OF CURVES

    When σ is a k-section of PrΛ and P is the corresponding k-rational point,

    we write Hσ for HP . Similarly, we may identify a plane curve X of degree

    d over an object A of Ar with an A-section of Pd(d+3)/2Λ . In particular,

    the local Hilbert functor HX of a plane curve over k is a subfunctor of

    some Pd(d+3)/2Λ .

    Remark 2.14. — Let {Fi}i∈I be a finite family of functors of Artin rings. Thentheir product F =

    ∏i∈I Fi in the category of functors Ar→ Set is a functor of Artin

    rings. Let (H*) be any of Schlessinger’s conditions or smoothness. If each Fi satisfies

    (H*), then so does F .

    By the above remark and the previous example, the functor

    H(X,Σ)df= HX ×

    ∏σ∈Σ

    is prorepresentable and smooth.

    Equimultiple Hilbert functors. — We are now able to consider equimultiple

    proper liftings of plane curves in the framework just introduced. Following Wahl

    again [Wah74b, Sec. 1], we make

    Definition 2.15. — The equimultiple local Hilbert functor of (X,Σ) sends an alge-

    bra A of Ar to

    E(X,Σ)(A)df={

    (X,Σ) ∈ H(X,Σ)(A) | For all σ ∈ Σ, X is equimultiple along σ}.

    Remark 2.16. — Let EX be the image of E(X,Σ) in HX . Assume that the sections

    in Σ correspond to the multiple points of X. Then for any object A of Âr, ÊX(A) is

    the set of equimultiple proper liftings of X to A.

    The equimultiple local Hilbert functor is prorepresentable and – at least in some cases

    – smooth. To see this, we introduce a scheme that represents a “global equimultiple

    Hilbert functor”. Let σ be a lifting of σ ∈ Σ to an object A of Âr. Let (x0, x1, x2)be coordinates of P2Λ over Λ. If the image of σ is contained in D(xi), then so is

    the image of σ. Hence, the ideal sheaf of σ is trivial outside D(xi) and on D(xi) it

    corresponds to an ideal generated by xj − tj , j ∈ {0, 1, 2} \ {i}, for some tj ∈ A. Asnoted following Definition 2.10, the multiplicity of a plane curve V (F ) over A is given

    by the least degree among the nonvanishing homogeneous parts of F (1, x1+t1, x2+t2)

    (or F (x0 + t0, 1, x2 + t2) or F (x0 + t0, x1 + t1, 1) depending on i). This motivates the

    following notation and definition:

    Let x = (x0, x1, x2) be fixed coordinates for P2Λ. For each section σ ∈ Σ,

    pick a fixed xσ ∈ x such that the image of σ is contained in D(xσ). Letx̂σ the tuple obtained from x by removing xσ.

  • 2.2. THE EQUIMULTIPLE LOCAL HILBERT FUNCTOR 17

    Definition 2.17. — Pick coordinates λ = (λα)α∈N3,|α|=d and tσ = (tσ,0, tσ,1, tσ,2),

    σ ∈ Σ, for the multiprojective space Pd(d+3)/2Λ ×∏σ∈Σ P

    2Λ. Set F =

    ∑α λαx

    α.

    For each σ ∈ Σ, let fσ ∈ Λ[x̂σ][λ, tσ] be the polynomial F (1, x1 + tσ,1, x2 + tσ,2) orF (x0 + tσ,0, 1, x2 + tσ,2) or F (x0 + tσ,0, x1 + tσ,1, 1) (choose the polynomial that sets

    xσ to 1). For each σ ∈ Σ, β ∈ N2, and |β| ≤ mσ − 1, let fσ,β ∈ Λ[λ, tσ] be thecoefficient of x̂βσ in fσ. We set

    E(d,m)df= V ({fσ,β}σ,β) ⊆ Pd(d+3)/2Λ ×

    ∏σ∈Σ

    P2Λ.

    Even though the notation suggests that the scheme E(d,m) depends only on d and m,

    in general it also depends on how the k-rational points specified by Σ are spread out in

    the projective plane: Unless k is infinite, there may not be coordinates x for P2Λ such

    that a single D(xi) contains all points specified by Σ. For an example over the field of

    2 elements, consider the union of the lines V (x0), V (x1), V (x2) and V (x0 + x1 + x2)

    and let Σ be the set of sections corresponding to the multiple points.

    Remark 2.18. — Let A be an object of Ar. By the discussion preceding Defini-

    tion 2.17,

    E(X,Σ)(A) = H(X,Σ)(A) ∩ E(d,m)(A),

    where the intersection is taken in(Pd(d+3)/2Λ ×

    ∏σ∈Σ P

    )(A).

    We now turn to prorepresentability and smoothness of E(X,Σ) and EX .

    Proposition 2.19. — The functor E(X,Σ) is prorepresented by the completion Ô of

    the local ring O = OE(d,m),(X,Σ).

    Proof. — We need to show that the functors HomΛ(Ô,−) and E(X,Σ) are isomorphic.So let A be an object of Ar and let N be its length. Write RN = R/m

    NR for any

    Λ-algebra R. On the one hand, there is a natural bijection between HomΛ(Ô, A) and

    HomΛN (ÔN , A). Moreover, ÔN and ON are naturally isomorphic.

    On the other hand, by definition an element of E(X,Σ)(A) may be interpreted as an

    A-section of E(d,m) that lifts the unique element of E(X,Σ)(k). In other words, the

    elements of E(X,Σ)(A) correspond bijectively to local morphisms O � ON → A of Λ-algebras. These in turn correspond bijectively to the elements of HomΛN (ON , A).

    Lemma 2.20. — If Σ contains at most one section, then E(X,Σ) is smooth.

    Proof. — If Σ is empty, then EX,Σ equals HX and we already mentioned that this

    functor is smooth in Example 2.13. So let Σ = {σ}. Let A′ � A be a surjectionin Ar and let (X,Σ) ∈ E(X,Σ)(A). We need to show that there exists an element

  • 18 CHAPTER 2. LIFTING BIRATIONAL EQUIVALENCE CLASSES OF CURVES

    (X ′,Σ′) ∈ E(X,Σ)(A′) that maps to (X,Σ). We may choose coordinates (x, y, z) forP2Λ such that the ideal sheaf of σ is the ideal generated by x and y in A[x, y]. Let∑di=0 Fi(x, y)z

    i be a defining polynomial of X and Fi(x, y) be homogeneous part of

    degree d − i. Let m be the unique member of m, i.e. the multiplicity of X along σ.Since (X,Σ) is a section of E(d,m), the polynomials F0, F1, . . . , Fm−1 vanish. For

    i = m, . . . , d, pick a homogeneous lifting F ′i of Fi to A′. Let X ′ be the plane curve

    over A′ defined by∑di=m F

    ′izi and σ′ be the A′-section of P2Λ that corresponds to

    the ideal generated by x and y in A′[x, y]. Then (X ′,Σ′) is the desired element of

    E(X,{σ})(A′).

    In general, the question of smoothness seems more complicated. Since E(X,Σ) maps

    onto EX by definition, the latter functor is smooth whenever the first is. The converse

    holds when all points corresponding to the sections in Σ are ordinary singularities

    on X. In fact, we have

    Lemma 2.21. — Assume that the point corresponding to the section σ ∈ Σ is anordinary singularity on X. Then the natural map E(X,Σ) → E(X,Σ\{σ}) is injective.In particular, when all sections in Σ correspond to ordinary singularities on X, then

    the natural surjection E(X,Σ) � EX is an isomorphism.

    For Λ the power series ring over the complex numbers and X an algebroid curve, the

    lemma was proven by Wahl [Wah74b, Prop. 1.9]. We follow his proof.

    Proof. — Since E(X,Σ) and E(X,Σ\{σ}) are both prorepresentable by the previous

    proposition, it suffices to show that the map on tangent spaces t(E(X,Σ))→ t(E(X,Σ\{σ}))is injective [Wah74a, Prop. 1.1.4]. (Wahl’s proof is given in a more restrictive setting

    than needed here. However, his proof can be used verbatim also in our context.)

    Let (X,Σ) and (X,Σ′) be elements of t(E(X,Σ)) that map to the same element of

    t(E(X,Σ\{σ})), e.g. Σ = Θ ∪ {σ} and Σ′ = Θ ∪ {σ′} and σ, σ′ are liftings of σ. Let mbe the multiplicity of X along σ. We may choose coordinates (x, y, z) of P2Λ such that

    σ and σ′ correspond to the ideals (x, y) and (x− εa, y − εb) of k[ε][x, y] respectively.We may also assume that k is algebraically closed. Then in addition we can choose

    the coordinates so that X is defined by a polynomial F (x, y, z) such that the degree

    m part of f(x, y) = F (x, y, 1) is equal to yh(x, y).

    We may assume that a and b are elements of k and thus must show that both

    equal 0. So let F + εG be a defining polynomial of X in k[ε][x, y, z] and write

    g(x, y) = G(x, y, 1). Since (X,Σ′) is an element of t(E(X,Σ)), we calculate

    f(x+εa, y+εb)+εg(x+εa, y+εb) = f(x, y)+εa∂f

    ∂x(x, y)+εb

    ∂f

    ∂y(x, y)+εg(x, y) ∈ (x, y)m.

    (2.6)

  • 2.2. THE EQUIMULTIPLE LOCAL HILBERT FUNCTOR 19

    But we also know that (X,Σ) is an element of t(E(X,Σ)), so f(x, y) + εg(x, y) also

    belongs to (x, y)m. Hence,

    a∂f

    ∂x(x, y) + b

    ∂f

    ∂y(x, y) ∈ (x, y)m.

    For the degree m part of f(x, y) written as yh(x, y), this means

    ay∂h

    ∂x(x, y) + bh(x, y) + by

    ∂h

    ∂y(x, y) = 0.

    Therefore, y divides bh(x, y). Since σ corresponds to an ordinary singularity, the

    degree m part of f splits into pairwise distinct linear factors. Hence, y does not

    divide h(x, y), but b and so b is equal to 0. This in turn implies that a(∂h/∂x)(x, y)

    vanishes. So either a equals 0 as desired or h is a p-th power. (Use that y does

    not divide h(x, y).) The latter cannot happen since σ corresponds to an ordinary

    singularity.

    Remark 2.22. — With the notions from the proof, we can identify the tangent space

    of EX when Σ contains a unique section: Indeed, let I denote the ideal sheaf in OP2kdefined by the ideal (f, ∂f/∂x, ∂f/∂y)+(x, y)m ⊆ k[x, y]. Recall that there is a shortexact sequence

    0→ OP2k → I (X)→ IX(X)→ 0, (2.7)

    where I (X) is the tensor product I ⊗ OP2k(X). Equation (2.6) implies

    t(EX)∼= H0(IX(X)). (2.8)

    In general, define an ideal sheaf for each section in Σ as above and let I be their

    sum. Then an isomorphism as in (2.8) still exists. Note also that this identification is

    independent of the types of singularities that the elements of Σ correspond to.

    Definition 2.23. — The ideal sheaf I defined in the remark above is called the

    equimultiple ideal sheaf for X and Σ. If Σ is the set of all sections of X along which

    X has multiplicity greater than 1, then we also say that I is the equimultiple ideal

    sheaf of X.

    Proposition 2.24. — Let I be the equimultiple ideal sheaf. The functor EX is

    smooth if H1(IX(X)) vanishes.

    For curves over complex surfaces (in fact, characteristic 0 seems to suffice) Greuel

    and Lossen show that EX is smooth already when Im(H1(I (X))→ H1(IX(X))

    )vanishes [GL01, Prop. 1.3]. In case that no sections are considered, i.e. Σ is the

    empty set, their statement is the “Severi-Kodeira-Spencer” theorem. A proof is in

    Mumford’s “Lectures on Curves on an Algebraic Surface” [Mum66, Lec. 23]. We

    follow this proof.

  • 20 CHAPTER 2. LIFTING BIRATIONAL EQUIVALENCE CLASSES OF CURVES

    Proof. — To begin, let us fix some notation: Let {Ui}i∈I be a covering of P2k by openaffines such that each Ui contains the image of at most one section σ ∈ Σ. Let Σibe the subset of Σ that consists of this unique section or is empty when Ui does not

    contain the image of any section. Set Uij = Ui ∩ Uj and Uij` = Ui ∩ Uj ∩ U` fori, j, ` ∈ I. Note that P2A has the same topological space as P2k for any object A of Ar.

    Now let A′ � A be a surjection in Ar. We need to show that any element of EX(A)lifts to an element of EX(A

    ′). It suffices to do this when A′ � A is a small extensionwhose kernel is generated by a single element η as k-vector space. (Since any surjection

    in Ar is the composite of finitely many such extensions.) So let X ∈ EX(A) be given.For each i ∈ I, let E(i)

    Xbe the image of E(X,Σi) in HX . By Lemma 2.20 and the

    fact that E(X,Σi) → E(i)

    Xis surjective, E

    (i)

    Xis smooth. Moreover, there are natural

    injections EX ↪→ E(i)

    X. Hence, for each i ∈ I, there exists a lifting X ′i ∈ E

    (i)

    X(A′) of X.

    There exists a lifting X ′ ∈ E(X,Σ)(A′) of X if and only if there exist liftingsX ′i ∈ E

    (i)

    X(A′) of X that glue along the Uij . Assume that X = V (F ) for some

    homogeneous polynomial F (x, y, z) and for each i ∈ I, let X ′i = V (F ′i ), i ∈ I, be anylifting of X in E

    (i)

    X(A′). We may assume that each F ′i is a lifting of F to A

    ′. Write

    f ′i for the element induced in OP2A′

    (Ui) by F′i and fi for the image of f

    ′i in OP2A(Ui).

    Then

    f ′i = u′ijf′j + ηh

    ′ij (2.9)

    over Uij for some unit u′ij ∈ OP2

    A′(Uij)

    ∗ and h′ij ∈ OP2A′

    (Uij). We must show that

    the f ′i and u′ij can be chosen so that the h

    ′ij vanish. First, we calculate

    η(h′ij + u′ijh′j`) = f

    ′i − u′ijf ′j + u′ij(f ′j − u′j`f ′`)

    = f ′i − u′iju′j`f ′`= ηh′i` + (u

    ′i` − u′iju′j`)f ′`.

    Given an element a′ or a of OP2A′

    (Uij), let a denote its image in OP2k(Uij). From the

    above we get

    hij + uijhj` = hi` +

    (u′i` − u′iju′j`

    η

    )f `.

    Moreover, f i = uijf j . Hence,

    hij

    f i+hj`

    f j=hj`

    f i+

    (1− u′iju′j`u′i`

    −1

    η

    ).

    Since X ′i and X′j are equimultiple along the unique sections (if any) of Ui and Uj , the

    sections hij lie in I (Uij). Therefore, {hij/f i}i,j represents an element of H1(IX(X))which, by assumption, is equal to 0.

  • 2.3. COMPUTING EQUIMULTIPLE PROPER LIFTINGS 21

    Hence, there exist elements gi ∈ I (Ui) such that hij/f i equals gj/f j − gi/f i assection of IX(X)(Uij). Pick any liftings g

    ′i ∈ OP2

    A′(Ui) of the gi. Then f

    ′i + ηg

    ′i is an

    equimultiple(!) lifting of fi and

    f ′i + ηg′i = (u

    ′ij + η(hij + g

    ′i − g′ju′ij)/f ′j)(f ′j + ηg′j).

    Since uij is a unit, so is u′ij + η(hij + g

    ′i − g′ju′ij)/f ′j and the f ′i + ηg′i glue along the

    Uij as desired.

    2.3. Computing equimultiple proper liftings

    We are now ready to describe an algorithm that computes equimultiple proper liftings

    of a given plane curve X over k to WN (k). It is an application of formal smoothness

    of E(d,m) in neighbourhoods of smooth points.

    Since E(X,Σ) is prorepresented by the completion of OE(d,m),(X,Σ), the functor is

    smooth if and only if the completion is a power series ring over Λ. This in turn happens

    if and only if E(d,m) is smooth at (X,Z) [Bou06, Prop VIII.5.1 & Thm. VIII.5.2].

    Now let A be an algebra of Ar. We may identify an element of E(X,Σ)(A) with an

    A-section of E(d,m). If E(d,m) is smooth at (X,Z), then it is formally smooth in

    a neighbourhood of (X,Z) [Gro67, Thm. 17.5.1]. Hence, for any small extension

    A′ � A in Ar, there exists an A′-section of E(d,m) that lifts the A-section [Gro67,Def. 17.1.1]. This is also sometimes called the infinitesimal lifting property.

    Proposition 2.25 (Effective infinitesimal lifting property)

    Let V be a scheme of finite type over Λ, σ ∈ V (k), and Imσ be a smooth point of V .Given an integer N ≥ 1 and a presentation ΛN [x]/(f) of an open affine neighbourhoodU of Imσ, Algorithm 1 computes a section σ ∈ U(ΛN ) whose special fibre is σ.

    Proof. — The case N equal to 1 is trivial. Assume that N is at least 2. Let

    x = (x1, . . . , xr), f = (f1, . . . , fs), t be a generator of the maximal ideal of Λ, and

    n be an integer between 2 and N . A section in U(Λn) corresponds to an r-tuple of

    elements of Λn, i.e. an element of the free module Λrn. We need to show that if a

    represents a lifting of σ to U(Λn−1) at the beginning of the loop in Algorithm 1, then

    a represents a lifting to U(Λn) at the end of the loop.

    To find a lifting of a to Λrn that represents a section in U(Λn) the algorithm proceeds

    as follows: First, it picks any lifting a ∈ Λrn, i.e. a section in Ar(Λn). This sectionrestricts to an element of U(ΛN ) if and only if o = f(a) vanishes. Now any other

    lifting of a to Λrn is of the form a′ = a+ tN−1∆, ∆ ∈ ΛrN . This element a′ defines a

    section in U(Λn) if and only if f(a′) = f(a+ tN−1∆) = 0. This happens if and only

  • 22 CHAPTER 2. LIFTING BIRATIONAL EQUIVALENCE CLASSES OF CURVES

    Algorithm 1: Applies the infinitesimal lifting property to lift a k-section σ of

    an affine scheme U ⊆ ArΛN to a ΛN -section if Imσ is a smooth point.

    J ← Jacobian(U)1a← represent in kr(σ)2for n = 2 to N do3

    // Choose lifting to Ar(Λn)

    a← lift(Λn,a)4// Compute obstruction for lifting to lie in U(Λn)

    o← f(a)5// Modify lifting so that obstruction vanishes

    ∆← solve(J(a),−o/tn−1)6a← a+ tn−1 · lift(Λn,∆)7

    // Now a represents a lifting of σ to U(ΛN )

    return a8

    if

    tN−1J(a) ·∆ = −o,where J denotes the Jacobian matrix of U . Regard this as an equation in ∆. Both

    sides are multiples of tN−1, so the solutions to the equation are in one-to-one corre-

    spondence to the elements tN−1∆ where ∆ is any lifting to Λrn of a solution ∆ to the

    equation

    J(a) ·∆ = −o/tN−1. (2.10)Since U is formally smooth in a neighbourhood of Imσ, a lifting of a to Λrn that

    defines a section in U(Λn) has to exist, i.e. (2.10) has a solution ∆. The algorithm

    can now pick any lifting ∆ of ∆ and replace the initial choice of a by a + tN−1∆.

    By the above, this a represents the desired lifting of σ.

    Remark 2.26. — Li describes this algorithm in his thesis“Anwendung deformations-

    theoretischer Methoden zur Liftung des Frobeniusmorphismus” [Li08] when U is an

    affine plane curve. In this case, J(a) is a matrix in k1×2. Li argues that a solution

    to (2.10) always exists when U is smooth at Imσ since smoothness is equivalent to

    J(a) being of full rank. In the higher dimensional case of course, J(a) does not need

    to be of full rank if U is smooth at Imσ. A priori it is not clear that (2.10) has a

    solution and this is why we argue with formal smoothness.

    Theorem A. — Let X be an ordinary, plane curve over k with irreducible compo-

    nents Xi of degrees di, i = 1, . . . , s, Σ ⊆ P2k(k) be the set of sections that correspondto the multiple points of X, and I be the equimultiple ideal sheaf of X and Σ. As-

    sume that the curves Xi are geometrically irreducible. The space H1(IX(X)) vanishes

  • 2.3. COMPUTING EQUIMULTIPLE PROPER LIFTINGS 23

    whenever

    3di >∑

    mσ(X)>2

    mσ(Xi), for all i.

    Before we spent the remainder of this chapter on the proof of this theorem, let us first

    discuss the result. There is

    Corollary 2.27. — In the situation of Theorem A, E(d,m) is smooth at (X,Σ) and

    Algorithm 1 (applied to a suitable affine patch of E(d,m)) computes a lifting of (X,Σ)

    to ΛN .

    Proof. — Since H1(IX(X)) vanishes, EX is smooth (Proposition 2.24). As X is or-

    dinary, E(X,Σ) is smooth as well (Lemma 2.21). So by the arguments at the beginning

    of this section, the scheme Ed,m is smooth at (X,Σ). So Proposition 2.25 shows that

    the algorithm computes a lifting of (X,Σ) to ΛN .

    Remark 2.28. — Beerenwinkel also considered the problem of effectively lifting

    curves. In his Diplomarbeit [Bee99], he derives the same lifting algorithm as ours

    but with two restrictions: All multiple points of the given plane curve must be nodes

    and only at most b(d + 1)/2c nodes are allowed (where d is as usual the degree ofthe curve). This is due to the fact that Beerenwinkel relies on a Hensel-lifting lemma

    instead of formal deformation theory. Therefore, he needs that the Jacobian of Ed,mat (X,Σ) has full rank. However, when there are more than b(d + 1)/2c nodes, therank might no longer be full. (As example consider three nodes on a degree 4 curve

    that lie on a line).

    The first restriction – that all multiple points are nodes – is a mild one from a the-

    oretical point of view. However, in the context of lifting a smooth curve in practice,

    it is of course of interest to allow more general (ideally, any) plane images. A more

    serious restriction is the one on the number of nodes. If there are at most b(d+ 1)/2csuch points, then the genus formula (2.2) and formula (2.3) force

    d2 − 4d+ 3 ≤ 2g. (2.11)

    Hence, the Brill-Noether number ρ(g, 2, d) = g − 3(g − d + 2) is negative wheneverthe genus is at least 7. In particular, in general it is not clear whether smooth

    geometrically integral curves of large genus can be mapped birationally to a plane

    curve so that (2.11) is satisfied. Hence, it is not clear whether Beerenwinkel’s result

    is sufficient to lift arbitrary curves.

    Let us now turn to the proof of Theorem A. We will use the notion of isomorphism

    defect : Let F and G be coherent sheaves on a curve Y over k and assume that both

    have the same rank on each irreducible component of Y . In the following, a ·̃ denotesthe reduction of a sheaf modulo torsion. Let Z be a closed subset of Y and denote

  • 24 CHAPTER 2. LIFTING BIRATIONAL EQUIVALENCE CLASSES OF CURVES

    by ·Z the restriction of a sheaf to Z. The local isomorphism defect of F in G at a(closed) point P of Z is

    isodZ,P (F ,G )df= min

    (lengthOZ,P

    (Coker(F̃Z,P −→ G̃Z,P )

    )),

    where the minimum is over all injective local homomorphisms induced by local homo-

    morphisms FP → GP . The local isomorphism defect is nonzero at only finitely manypoints of Z and one defines

    isodZ(F ,G )df=∑P∈Z

    [k(P ) : k]isodZ,P (F ,G )

    as the isomorphism defect of F in G along Z. When Z is the whole curve, then we

    drop the Z in the indices. A key lemma for the proof of Theorem A is

    Lemma 2.29. — Let Y be a curve on a smooth surface(3) over k, Y 1, . . . , Y s its

    irreducible components, ωY its canonical sheaf, and F be a coherent OY -module of

    rank 1 on each irreducible component of Y . Then H1(F ) vanishes if for all i = 1, . . . , s

    χ(F̃Y i) > χ(ωY ,Y i)− isodY i(F , ωY ).

    Proof. — Greuel and Karras’ proof over the complex numbers [GK89, Prop. 5.2]

    works over any field (the fact that k is perfect is not needed): Assume that H1(F )

    does not vanish. Since Y is a curve on a smooth surface, it is a local complete

    intersection and the canonical sheaf of Y is (isomorphic to) its dualising sheaf [Liu06,

    Thm. 6.4.32], so there exists a nonzero morphism ϕ : F → ωY [Liu06, Rem. 6.4.20].Since ωY is torsion-free, the image of ϕ must have rank 1 on at least one irreducible

    component Y i of Y . Let ϕ̃i : F̃Y i → ω̃Y ,Y i be the map induced by ϕ. Note thatω̃Y ,Y i equals ωY ,Y i as the latter sheaf is already torsion-free. By the above, ϕ̃i is

    injective and thus

    χ(F̃Y i) = χ(ωY ,Y i)− χ(Coker ϕ̃i) ≤ χ(ωY ,Y i)− isodY i(F , ωY ).

    We introduce one final notion that we will use in the following proof: Given a plane

    curve Y over k with equimultiple ideal sheaf I and a closed point P on Y , set

    τ emP

    (Y )df= lengthOY ,P (OY ,P

    /IY ,P )

    and

    τ em(Y )df= h0(OY ,P

    /IY ,P ) =

    ∑P

    [k(P ) : k]τ emP

    (Y ).

    (3)A surface over k is a geometrically reduced, separated k-scheme of finite type all of whose irre-

    ducible components are of dimension 2.

  • 2.3. COMPUTING EQUIMULTIPLE PROPER LIFTINGS 25

    Proof of Theorem A. — We follow the proof of Greuel and Lossen who consider the

    case that Λ is the ring of formal power series over the complex numbers [GL96,

    Cor. 5.1]. To begin, note that the sheaf IX(X) vanishes outside X, so we may regard

    it as OX -module when computing its cohomology [Har97, Lem. III.2.10]. Our goal

    is to apply Lemma 2.29. To this end, note that there is an exact sequence

    0→ IX(X)→ OX(X)→ Q → 0

    which we may take to define Q. By construction, the stalk of Q at a closed point P

    of X is

    QP∼= OX,P

    /IX,P ,

    Hence, Q is torsion and vanishes outside the singular locus of X. Let us indicate with

    an index i the restriction of a sheaf on X to Xi. Then Qi is also torsion and (with

    the ·̃-notation introduced before Lemma 2.29) there is an exact sequence

    0→ ĨX(X)i → OX(X)i → Qi → 0 (2.12)

    for all i = 1, . . . , s. Since OX(X) is an invertible sheaf, so is OX(X)i and we deduce

    from the exact sequences that IX(X) is a coherent OX -module of rank 1 on each

    irreducible component of X. Let ωX be the canonical sheaf of X. We apply

    Lemma 2.29 and find that H1(IX(X)) vanishes if

    χ(ĨX(X)i) > χ(ωX,i)− isodXi(IX(X), ωX) (2.13)

    for all i = 1, . . . , s. Let us first consider χ(ωX,i). Let ωP2k be the canonical sheaf

    of P2k. The adjunction formula tells us that ωX,i equals (OP2k(X)⊗ ωP2k)|Xi [Liu06,Thm. 9.1.37]. This in turn equals (OP2k(Xi)⊗ωP2k)|Xi⊗OP2k(X

    ′)|Xi withX = X′+Xi.

    By adjunction again, (OP2k(Xi) ⊗ ωP2k)|Xi is isomorphic to the canonical sheaf ωXiof Xi. For the Euler-Poincaré characteristics this means

    χ(ωX,i) = χ(ωXi) + χ(OP2k(X′)|Xi)− χ(OXi).

    Now, χ(OP2k(X′)|Xi) − χ(OXi) is just the intersection number X

    ′ · Xi of X′

    and

    Xi [Liu06, Thm. 9.1.12]. Furthermore, χ(ωXi) equals −χ(OXi) by duality [Liu06,Rem. 6.4.21] and therefore is equal to pa(Xi)− 1. Putting everything together, (2.13)is equivalent to

    χ(ĨX(X)i) > pa(Xi)− 1 +X′ ·Xi − isodXi(IX(X), ωX). (2.14)

    Let us now turn to χ(ĨX(X)i). By(2.12) and since the support of Qi is 0-dimensional,

    χ(ĨX(X)i) = χ(OX(X)i)− χ(Qi)

    = χ(OX(X)i)− h0(Qi).

    Now, χ(OX(X)i) equals 1−pa(Xi) +X ·Xi [Liu06, Thm. 9.1.12]. Plugging this into(2.14) gives

    X ·Xi − 2pa(Xi) + 2 > X′ ·Xi + h0(Qi)− isodXi(IX(X), ωX)

  • 26 CHAPTER 2. LIFTING BIRATIONAL EQUIVALENCE CLASSES OF CURVES

    and by adjunction

    X ·Xi − (W +Xi) ·Xi > X′ ·Xi + h0(Qi)− isodXi(IX(X),OX),

    where W is a canonical divisor of P2k. This simplifies to

    −W ·Xi > h0(Qi)− isodXi(IX(X), ωX). (2.15)

    Let H be a line on P2k. Then one may take W as −3H. So Bézout tells us that theabove inequality is equivalent to

    3di > h0(Qi)− isodXi(IX(X), ωX).

    We are left to show that∑mP (X)>2

    mP (Xi) is at least h0(Qi)− isodXi(IX(X), ωX).

    By Lemma 2.30, the latter equals∑mP (X)>0

    τ emP

    (Xi) + iP (Xi, X′)− isodXi,P (IX(X), ωX).

    By Lemma 2.33 and Corollary 2.31, this simplifies to∑mP (X)>0

    τ emP

    (Xi)− δP (Xi) + length(

    (IX,P/

    cond(IX,P ))⊗ OXi,P),

    where cond(·) denotes the conductor (see below). Clearly, the individual summandsare each 0 whenever P is a regular point of X, i.e. whenever mP (X) equals 1. So,

    together with Lemma 2.34, the above simplifies once more to∑mP (X)>1

    min{0,mP (Xi)− 2}+ length(

    (IX,P/

    cond(IX,P ))⊗ OXi,P).

    The theorem is proven with Lemma 2.35.

    Lemmas for the Proof. — This subsection collects several lemmas that were used

    in the proof of Theorem A. To this end, we assume that k is algebraically closed (this

    is not strictly needed for every lemma). Let Y be a plane, ordinary curve over k, P

    be a closed point of Y , m be the multiplicity of Y at P , I the equimultiple ideal

    sheaf of Y , and ωY be the canonical sheaf of Y . In the application to the proof, Y

    will play the role of X or Xi. Let x and y be local coordinates P2k at P . We may

    and do assume that Y is defined at P by f ∈ k[x, y] and that x does not divide thedegree m part of f .

    Lemma 2.30. — In the situation of the proof Theorem A, let P be a closed point

    of X. Then

    length(Qi,P ) = τemP

    (Xi) + iP (Xi, X′).

    Proof. — If P is not a multiple point of X, then the three cardinals in each of the

    two equations are each equal to 0. So let P be a multiple point of X. By our

    earlier assumption in Section 2.1, the degree of P is 1, so we find defining elements

  • 2.3. COMPUTING EQUIMULTIPLE PROPER LIFTINGS 27

    f i, f′ ∈ k[x, y] of Xi and X

    ′at P respectively. Let mi,m

    ′ be the multiplicities of Xiand X

    ′at P respectively. Then(

    OX,P/IX,P

    )⊗ OXi,P ∼= OXi,P

    /f′ ·(∂f i∂x

    ,∂f i∂y

    )+ (x, y)mi+m

    ∼= ÔXi,P/f′ ·(∂f i∂x

    ,∂f i∂y

    )+ (x, y)mi+m

    ′,

    where ÔXi,P is the completion of OXi,P . Hence, the lemma is a direct consequence

    of the snake lemma applied to the following commutative diagram of ÔXi,P -modules

    with exact rows:(∂fi∂x ,

    ∂fi∂y

    )+ (x, y)mi

    ·f ′//

    ��

    f′ ·(∂fi∂x ,

    ∂fi∂y

    )+ (x, y)mi+m

    ′//

    ��

    0 //

    ��

    0

    0 // ÔXi,P·f ′

    // ÔXi,P// ÔXi,P

    /f′ÔXi,P

    // 0

    The only question that remains is why the first row is exact. To answer it, we need to

    show that (x, y)mi+m′

    is in the image of the first map (the rest is immediate). To this

    end, it suffices to show that the ideal in k[[x, y]] generated by the degree mi part of

    f i and the degree m part of f′

    contains (x, y)mi+m′. Now recall that P is an ordinary

    multiple point of X. Hence, the degree mi part of f i and the degree m′ part of f

    ′are

    coprime. Some linear algebra in the coefficients of those polynomials shows that the

    first row is exact.

    Let R be reduced Noetherian ring and R̃ be its normalisation. The conductor

    of an ideal I of R is the ideal cond(I)df= {h ∈ I | hR̃ ⊆ I}. Conductors im-

    plicitely appeared already earlier in this chapter: The δ-invariant δP (Y ) equals

    length(OY ,P/

    cond(OY ,P )) [Ful02, Cor. 2.2]. If P is an ordinary singularity, then

    cond(OY ,P ) equals (x, y)m−1 [Ful02, Prop. 2.1]. So the same reasoning as in the

    proof of the above lemma gives

    Corollary 2.31. — In the situation of the proof Theorem A, let P be a closed point

    of X. Then

    length(

    (OX,P/

    cond(OX,P ))⊗ OXi,P)

    = δP (Xi) + iP (Xi, X′).

    The map that sends an element h of cond(I) to the multipliation-by-h map gives a

    natural isomorphism between cond(I) and HomR(R̃, I). In particular, any morphism

    I → J of ideals of R restricts to a morphism cond(I)→ cond(J).

    Lemma 2.32. — Assume that P is an ordinary singularity of Y and let ÕY ,P be the

    normalisation of OY ,P . If P is a node, then the conductor of IY ,P is IY ,P = (x, y).

  • 28 CHAPTER 2. LIFTING BIRATIONAL EQUIVALENCE CLASSES OF CURVES

    Otherwise m is at least 3 and the conductor is xmÕY ,P = (x, y)m. Moreover, the iso-

    morphism ϕ : xmÕY ,P∼−→ xm−1ÕY ,P = (x, y)m−1 extends to an injective morphism

    IY ,P ↪→ OY ,P .

    Proof. — First, let m equal 2, i.e. P is a node. Then IY ,P equals (x, y) and Fulton’s

    proof that (x, y) is the conductor of OY ,P also shows that (x, y) is the conductor of

    IY ,P . Now let m be at least 3. A single blow-up resolves the singularity. Thus, ÕY ,Pis generated as OY ,P -module by t

    j , j = 0, . . . ,m− 1, with xt = y [Ful02, Prop. 2.1].Consequently, xmÕY ,P = (x, y)

    m is a subset of the conductor of IY ,P . Moreover

    xm−1ÕY ,P equals the conductor (x, y)m−1 of OY ,P . It is then not difficult to see that

    ϕ extends as claimed and is injective. So by the remarks above the lemma, the inverse

    image of xm−1ÕY ,P under ϕ contains the conductor of IY ,P . Since the inverse image

    is exactly xmÕY ,P , the claim is proven.

    Lemma 2.33. — In the situation of the proof Theorem A, let P be a closed point

    of X. There is the equality

    isodXi,P (IX(X), ωX) = length(

    (OX,P/

    cond(OX,P ))⊗ OXi,P)

    − length(

    (IX,P/

    cond(IX,P ))⊗ OXi,P),

    Proof. — If P is not a multiple point of X, then all three cardinals in the equation are

    equal to 0. So let P be a multiple point of X. Since ωX and OX(X) are both invertible

    sheaves, isodXi,P (IX(X), ωX) equals isodXi,P (IX(X),OX(X)). The latter, in turn,

    equals isodXi,P (IX ,OX). So let ϕ : IX,P → OX,P be a morphism such that theinduced morphism ϕ̃i : ˜IX,P ⊗ OXi,P → ˜OX,P ⊗ OXi,P = OXi,P is injective.

    Note that IX,P ⊗ OXi,P is the OXi,P -ideal generated by ∂f/∂x, ∂f/∂y and themonomials of degree at least m. It is therefore torsion-free, so the restriction ϕi of ϕ

    to Xi and ϕ̃i agree. Hence, there is the following exact commutative diagram.

    0

    ��

    cond(IX,P )⊗ OXi,P //

    ��

    cond(OX,P )⊗ OXi,P //

    ��

    Cokerϕi|cond(...) //

    ��

    0

    0 // IX,P ⊗ OXi,Pϕi

    // OX,P ⊗ OXi,P // Cokerϕi // 0

  • 2.3. COMPUTING EQUIMULTIPLE PROPER LIFTINGS 29

    By the snake lemma,

    length (Cokerϕi) = length(

    (OX,P/

    cond(OX,P ))⊗ OXi,P)

    − length(

    (IX,P/

    cond(IX,P ))⊗ OXi,P)

    + length(Cokerϕi|cond(...)

    )If we let ϕ be the map from Lemma 2.32 (take ϕ as the natural inclusion if P is a

    node), then ϕi is injective and the claim is proven.

    Lemma 2.34. — Let P be a multiple point of Y . Then τ emP

    (Y ) = m(m+ 1)/2− 2.

    Proof. — Recall that τ emP

    (Y ) is length(OY ,P /IY ,P ) and that the length equals the

    dimension over k since P is rational. Moreover, IY ,P equals (∂f/∂x, ∂f/∂y)+(x, y)m.

    Since P is an ordinary singularity, the degree m − 1 parts of ∂f/∂x and ∂f/∂y arelinearly independent over k (even in positive characteristic) and the claim follows.

    Lemma 2.35. — In the situation of the proof Theorem A, let P be a multiple point

    of X. Then

    length(

    (IX,P /cond(IX,P ))⊗ OXi,P)≤ min{mP (Xi), 2}.

    Moreover, if mP (X) is 2, then the length is 0.

    Proof. — If mP (X) equals 2, then the claim immediately follows from Lemma 2.32.

    So let mP (X) be larger than 2. If mP (Xi) is 0, then OXi,P vanishes and so does the

    length of (IX,P /cond(IX,P ))⊗ OXi,P .

    So letmP (Xi) be at least 1. Since P is rational, the length of (IX,P /cond(IX,P ))⊗OXi,Pequals its dimension over k. By Lemma 2.32, the k-space is generated by the images

    of ∂f/∂x and ∂f/∂y.

    As remarked earlier, IX,P ⊗ OXi,P is the OXi,P -ideal generated by ∂f/∂x and∂f/∂y) + (x, y)mP (X). So if mP (Xi) equals 1, then ∂f/∂x and ∂f/∂y are con-

    stant multiples of f′

    (f′

    as before a defining element of X′

    at P ). In particular,

    IX,P ⊗ OXi,P /cond(IX,P )⊗ OXi,P is generated by the image of f′.

  • CHAPTER 3

    AN EXPLICIT DESCRIPTION OF FROBENIUS-STABLE

    LATTICES

    The zeta function of a smooth geometrically integral curve X over k is uniquely

    determined by the Frobenius action on the first rigid cohomology group. It is therefore

    of interest to have an explicit description of this group. Let X be a proper lifting of

    X to W (k) and Z a nontrivial reduced strict relative normal crossing divisor on X.

    Then there is the following description for large enough `:

    H0(ΩXK ((`+ 1)ZK))/dH0(OXK (`ZK))

    ∼−→ H1dR(XK \ ZK)∼−→ H1rig(X \ Z). (3.1)

    Indeed, this is a consequence of (1.1) and the Riemann-Roch theorem. The remarks

    from Section 1.2 lead to the following natural and relevant question:(1)

    Does there exist and can one describe a sequence of isomorphic,

    Frobenius-stable lattices such that tensoring with K naturally yields(3.1)?

    Two approaches to answer this question come to mind. In this chapter we look at

    both of them and show that one succeeds while the other fails (and to what extent the

    other fails). The successful approach, considered in Section 3.1, is via crystalline and

    de Rham cohomology of the pair (X,Z). Crystalline cohomology is a contravariant

    functor from pairs (X,Z) to W (k)-modules Hicr(X,Z) [Shi02, Chp. 2 & 3]. Thanks to

    natural isomorphisms by Abott et al. [AKR11, Cor. 2.2.6], Kato [Kat89, Thm. 6.4],

    and Shiho [Shi02, Cor. 2.4.13 & Thm. 3.1.1], there is the following commutative

    diagram

    H1dR(X,Z)⊗K∼ // H1dR(XK \ ZK)

    ∼ // H1rig(X \ Z)∼ // H1cr(X,Z)⊗K

    H1dR(X,Z)∼ //

    OO

    H1cr(X,Z)

    OO(3.2)

    (1)A lattice in a vector space V of finite dimension n over K is a free sub-W (k)-module of V of

    rank n [Bou89a, Sec. VII.4].

  • 32 CHAPTER 3. AN EXPLICIT DESCRIPTION OF FROBENIUS-STABLE LATTICES

    in which (by naturality) the top, right-hand map is equivariant with respect to the

    Frobenius maps on crystalline and rigid cohomology. The main result of this section

    is as follows: Let ϕ be the map sending a differential ω ∈ H0(ΩX((` + 1)Z)) to theformal sum of its Laurent expansions at the closed points in the support of Z, cut off

    above t−2dt (where t is a uniformising parameter) and where the coefficient of tidt is

    taken modulo i+ 1. Then there is a natural isomorphism

    Kerϕ/dH0(OX(`Z))

    ∼−→ H1dR(X,Z)/

    (tor).(2) (3.3)

    The statement is made precise below (Theorem B).

    The second approach to give an answer is via Monsky-Washnitzer and de Rham coho-

    mology of X \Z. The question whether the image of H1dR(X \Z) in H1dR(XK \ZK) isa Frobenius-stable lattice was already asked by van den Bogaart in his thesis [vdB08,

    p. 82]. In Section 3.2, we provide this module with a Frobenius action using an integral

    definition of Monsky-Washnitzer cohomology (Definition 3.10 and Proposition 3.12).

    We will then express the failure of the image of H1dR(X \ Z) to be a lattice in termsof the genus and the Hasse-Witt invariant of X (Proposition 3.13).

    We end this overview of the chapter with the following remark: Let Ω(rf)XK

    and Ω(rf)

    Xbe

    the spaces of residue-free differentials on XK respectively X. Let g be the genus of

    X. The quotient of Ω(rf)XK

    by exact differentials has dimension 2g, while the quotient

    of Ω(rf)

    Xby exact differentials is infinite-dimensional. There are two ways to get a

    finite-dimensional quotient space in characteristic p: One can restrict to differentials

    of the second kind [Ros53], i.e. differentials whose Laurent tails are integrable (in

    characteristic 0 this notion coincides with being residue-free), or one can take the

    quotient by the larger space of pseudo-exact differentials [Lam58]; a definition is

    given below. The first approach loosely corresponds to de Rham cohomology of (X,Z)

    and the second approach to de Rham cohomology of X \ Z.

    Further notation and terminology. — The diagram (3.2) holds for any smooth

    proper pair (X,Z) overW (k) and in any degree. Here, a smooth proper pair over W (k)

    is a pair (X,Z) consisting of a smooth proper W (k)-scheme X of relative dimension

    n and a reduced strict relative normal crossing divisor Z on X. The strict relative

    normal crossing condition means that X can be covered by open affine subschemes U

    such that each U is étale over AnW (k) and Z|U is the fibre product of U and the unionof some (or all, or none) coordinate hyperplanes. In other words, there exists an étale

    map

    W (k)[x1, . . . , xn]étale−→ OX(U)

    xj 7→ tj

    (2)One can show that H1dR(X,Z) is torsion-free (Lemma 3.7).

  • 3.1. DE RHAM COHOMOLOGY OF SMOOTH PROPER PAIRS 33

    such that

    OX(−Z)(U) = t1 · · · trOX(U),

    for some 0 ≤ r ≤ n depending on U . We always assume that r in the above formula isminimal in the sense that no tj , j = 1, . . . , r, can be left out. For simplicity, we call U

    together with the t1, . . . , tn a pleasant open affine of X and a pleasant neighbourhood

    of any point P of U .

    For the remainder of this chapter, let (X,Z) be a smooth proper pair over

    W (k) where X is of relative dimension n and Z nontrivial.

    3.1. De Rham cohomology of smooth proper pairs

    This section is divided into three subsections. In the first we define differentials with

    logarithmic poles and the de Rham cohomology of (X,Z). The middle subsection

    contains Theorem B and its proof. The third part specialises to the case of curves.

    Differentials with logarithmic poles. — To define the de Rham cohomology of

    (X,Z) one introduces the sheaf Ω(X,Z) of differentials on X with logarithmic poles

    along Z: Let U , t1, . . . , tn, be a pleasant open affine. Since U is étale over AnW (k), the

    differentials dt1, . . . , dtn form an OX(U)-basis of ΩX(U). Set

    d̃tjdf=

    {dtj/tj if j ≤ rdtj otherwise.

    The sections of Ω(X,Z) over U are defined as

    Ω(X,Z)(U)df=

    n⊕j=1

    OX(U)d̃tj (3.4)

    ⊆ ΩX(Z)(U).

    This definition glues and gives rise to a subsheaf Ω(X,Z) of ΩX(Z). We let

    ∂̃j : OX(U)→ OX(U) (3.5)

    be the map sending an element f to the coefficient of d̃tj in df ∈ Ω(X,Z)(U).

    As usual, we set Ωi(X,Z)df=∧i