Grundlehren der mathematischen Wissenschaften 334 A Series of Comprehensive Studies in Mathematics

Series editors

M. Berger B. Eckmann P. de la Harpe F. Hirzebruch N. Hitchin L. Hörmander M.-A. Knus A. Kupiainen G. Lebeau M.Ratner D.Serre Ya.G.Sinai N.J.A. Sloane B. Totaro A. Vershik M. Waldschmidt

Editor-in-Chief A. Chenciner J. Coates S.R.S. Varadhan Edoardo Sernesi

Deformations of Algebraic Schemes

ABC Edoardo Sernesi Universitá "Roma Tre" Department of Mathematics Largo San Leonardo Murialdo 1 00146 Roma, Italy e-mail: [email protected]

Library of Congress Control Number: 2006924565

Mathematics Subject Classification (2000): 14D15, 14B12

ISSN 0072-7830 ISBN-10 3-540-30608-0 Springer Berlin Heidelberg New York ISBN-13 978-3-540-30608-5 Springer Berlin Heidelberg New York

This work is subject to copyright. All rights are reserved, whether the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation, broadcasting, reproduction on microfilm or in any other way, and storage in data banks. Duplication of this publication or parts thereof is permitted only under the provisions of the German Copyright Law of September 9, 1965, in its current version, and permission for use must always be obtained from Springer. Violations are liable for prosecution under the German Copyright Law. Springer is a part of Springer Science+Business Media springer.com c Springer-Verlag Berlin Heidelberg 2006 Printed in The Netherlands The use of general descriptive names, registered names, trademarks, etc. in this publication does not imply, even in the absence of a specific statement, that such names are exempt from the relevant protective laws and regulations and therefore free for general use.

Typesetting: by the author and SPI Publisher Services using a Springer LTA EX macro package Cover design: design & production GmbH, Heidelberg Printed on acid-free paper SPIN: 11371335 41/SPI 543210 Preface

In one sense, deformation theory is as old as itself: this is because all algebro-geometric objects can be “deformed” by suitably varying the coefficients of their defining equations, and this has of course always been known by the classical geometers. Nevertheless, a correct understanding of what “deforming” means leads into the technically most difficult parts of our discipline. It is fair to say that such technical obstacles have had a vast impact on the crisis of the classical language and on the development of the modern one, based on the theory of schemes and on cohomological methods. The modern point of view originates from the seminal work of Kodaira and Spencer on small deformations of complex analytic manifolds and from its forma- lization and translation into the language of schemes given by Grothendieck. I will not recount the history of the subject here since good surveys already exist (e.g. [27], [138], [145], [168]). Today, while this area is rapidly developing, a self-contained text covering the basic results of what we can call “classical deformation theory” seems to be missing. Moreover, a number of technicalities and “well-known” facts are scattered in a vast literature as folklore, sometimes with proofs available only in the complex analytic category. This book is an attempt to fill such a gap, at least par- tially. More precisely, it aims at giving an account with complete proofs of the results and techniques which are needed to understand the local deformation theory of alge- braic schemes over an algebraically closed field, thus providing the tools needed, for example, in the local study of Hilbert schemes and moduli problems. The existing monographs, like [14], [93], [105], [109], [124], [163], [175], [176], [184], all aim at goals different from the above. For these reasons my approach has been to work exclusively in the category of locally noetherian schemes over a fixed algebraically closed field k, to avoid switch- ing back and forth between the algebraic and the analytic category. I tried to make the text self-contained as much as possible, but without forgetting that all the technical ideas and prerequisites can be found in [3] and [2]: therefore the reader is advised to keep a copy of them to hand while reading this text. In any case a good fami- liarity with [84] and with a standard text in like [48] or [127] VI Preface will be generally sufficient; the classical [167] and [190] will be also useful. A good acquaintance with homological algebra is assumed throughout. One of the difficulties of writing about this subject is that it needs a great num- ber of technical results, which make it hard to maintain a proper balance between generality and understandability. In order to overcome this problem I tried to keep the technicalities to a minimum, and I introduced the main deformation problems in an elementary fashion in Chapter 1; they are then reconsidered as functors of Artin rings in Chapter 2, where the main results of the theory are proved. The first two chapters therefore give a self-contained treatment of formal deformation theory via the “classical” approach; cotangent complexes and functors are not introduced, nor the method of differential graded Lie algebras. Another chapter treats in more detail the most important deformation functors, with the single exception of vector bun- dles; this was motivated by reasons of space and because good monographs on the subject are already available (e.g. [90], [118], [59]). Although they are not the central issue of the book, I considered it necessary to include a chapter on Hilbert schemes and Quot schemes, since it would be impossible to give meaningful examples and applications without them, and because of the lack of an appropriate reference. De- formation theory is closely tied with classical algebraic geometry because some of the issues which had remained controversial and unclear in the old language have found a natural explanation using the methods discussed here. I have included a sec- tion on plane curves which gives a good illustration of this point. Unfortunately, important topics and results have been omitted because of lack of space, energy and competence. In particular, I did not include the construction of any global moduli spaces/stacks, which would have taken me too far from the main theme. The book is organized in the following way. Chapter 1 starts with a concise treat- ment of algebra extensions which are fundamental in deformation theory. It then discusses locally trivial infinitesimal deformations of algebraic schemes. Chapter 2 deals with “functors of Artin rings”, the abstract tool for the study of formal defor- mation theory. The main result of this theory is Schlessinger’s theorem. A section on obstruction theory, an elementary but crucial technical point, is included. We discuss the relation between formal and algebraic deformations and the algebraization prob- lem. This part is not entirely self-contained since Artin’s algebraization theorem is not proved in general and the approximation theorem is only stated. The last section explains the role of automorphisms and the related notion of “isotriviality”. Chap- ter 3 is an introduction to the most important deformation problems. By applying Schlessinger’s theorem to them, we derive the existence of formal (semi)universal deformations. Many examples are discussed in detail so that all the basic principles of deformation theory become visible. This chapter can be used as a reference for several standard facts of deformation theory, and it can be also helpful in supplement- ing the study of the more abstract Chapter 2. Chapter 4 is devoted to the construction and general properties of Hilbert schemes, Quot schemes and their variants, the “flag Hilbert schemes”. It ends with a section on plane curves, where the main properties of Severi varieties are discussed. My approach to the proof of existence of nodal curves with any number of nodes uses multiple point schemes and is apparently new. Preface VII

In the Appendices I have collected several topics which are well known and standard but I felt it would be convenient for the reader to have them available here. Acknowledgements. Firstly, I would like to express my deepest gratitude to D. Lieberman, who introduced me to the study of deformations of complex mani- folds a long time ago. More recently, R. Hartshorne, after a careful reading of a previous draft of this book, made a number of comments and suggestions which sig- nificantly contributed in improving it. I warmly thank him for his generous help. For many extremely useful remarks on another draft of the book I am also indebted to M. Brion. I gratefully acknowledge comments and suggestions from other colleagues and students, in particular: L. Badescu, I. Bauer, A. Bruno, M. Gonzalez, A. Lopez, M. Manetti, A. Molina Rojas, D. Tossici, A. Verra, A. Vistoli. I would also like to thank L. Caporaso, F. Catanese, C. Ciliberto, L. Ein, D. Laksov and H. Lange for encouragement and support which have been of great help. Contents

Terminology and notation ...... 1

Introduction ...... 3

1 Infinitesimal deformations ...... 9 1.1 Extensions...... 9 1.1.1 Generalities ...... 9 1.1.2 The module ExA(R, I ) ...... 12 1.1.3 Extensions of schemes ...... 15 1.2 Locallytrivialdeformations...... 20 1.2.1 Generalities on deformations ...... 20 1.2.2 Infinitesimal deformations of nonsingular affine schemes . . . 23 1.2.3 Extending automorphisms of deformations ...... 26 1.2.4 First-orderlocallytrivialdeformations ...... 29 1.2.5 Higher-order deformations – obstructions ...... 32

2 Formal deformation theory ...... 37 2.1 Obstructions...... 37 2.2 Functors of Artin rings ...... 44 2.3 The theorem of Schlessinger ...... 54 2.4 The local moduli functors ...... 64 2.4.1 Generalities ...... 64 2.4.2 Obstruction spaces ...... 69 2.4.3 Algebraic surfaces ...... 72 2.5 Formal versus algebraic deformations...... 75 2.6 Automorphisms and prorepresentability ...... 89 2.6.1 The automorphism functor ...... 89 2.6.2 Isotriviality ...... 96 X Contents

3 Examples of deformation functors ...... 103 3.1 Affine schemes ...... 103 3.1.1 First-orderdeformations...... 103 3.1.2 The second cotangent module and obstructions ...... 109 3.1.3 Comparison with deformations of the nonsingular locus . . . . 117 3.1.4 Quotient singularities ...... 120 3.2 Closed subschemes ...... 122 3.2.1 The local Hilbert functor ...... 122 3.2.2 Obstructions...... 129 3.2.3 Theforgetfulmorphism...... 132 3.2.4 The local relative Hilbert functor ...... 136 3.3 Invertiblesheaves...... 137 3.3.1 The local Picard functors ...... 137 3.3.2 Deformationsofsections,I...... 141 3.3.3 Deformations of pairs (X, L) ...... 145 3.3.4 Deformationsofsections,II...... 152 3.4 Morphisms...... 156 3.4.1 Deformations of a morphism leaving domain and target fixed 157 3.4.2 Deformations of a morphism leaving the target fixed ...... 161 3.4.3 Morphisms from a nonsingular curve with fixed target . . . . . 172 3.4.4 Deformations of a closed embedding ...... 176 3.4.5 Stability and costability ...... 180

4 Hilbert and Quot schemes ...... 187 4.1 Castelnuovo–Mumfordregularity...... 187 4.2 Flatnessintheprojectivecase...... 194 4.2.1 Flatness and Hilbert polynomials ...... 194 4.2.2 Stratifications...... 198 4.2.3 Flatteningstratifications...... 200 4.3 Hilbert schemes ...... 206 4.3.1 Generalities ...... 206 4.3.2 Linear systems ...... 207 4.3.3 Grassmannians ...... 209 4.3.4 Existence ...... 213 4.4 Quot schemes ...... 219 4.4.1 Existence ...... 219 4.4.2 Local properties ...... 223 4.5 Flag Hilbert schemes...... 227 4.5.1 Existence ...... 227 4.5.2 Local properties ...... 230 4.6 Examplesandapplications...... 235 4.6.1 Completeintersections...... 235 4.6.2 An obstructed nonsingular curve in IP 3 ...... 237 4.6.3 An obstructed (nonreduced) ...... 238 4.6.4 Relative grassmannians and projective bundles ...... 240 Contents XI

4.6.5 Hilbert schemes of points ...... 247 4.6.6 Schemes of morphisms ...... 249 4.6.7 Focal loci ...... 251 4.7 Planecurves...... 254 4.7.1 Equisingular infinitesimal deformations ...... 254 4.7.2 TheSeverivarieties...... 256 4.7.3 Nonemptiness of Severi varieties ...... 262

A Flatness ...... 269

B Differentials ...... 279

C Smoothness ...... 293

D Complete intersections ...... 305 D.1 Regular embeddings ...... 305 D.2 Relativecompleteintersectionmorphisms...... 307

E Functorial language ...... 313

References ...... 321

List of symbols ...... 329

Index ...... 333 Terminology and notation

All rings will be commutative with 1. A ring homomorphism A → B is called essentially of finite type (e.f.t.) if B is a localization of an A-algebra of finite type. We will also say that B is e.f.t. over A. We will always denote by k a fixed algebraically closed field. All schemes will be assumed to be defined over k, locally noetherian and separated, and all algebraic sheaves will be quasi-coherent unless otherwise specified. If X and Y are schemes we will write X × Y instead of X ×k Y .IfS is a scheme and s ∈ S we denote by k(s) = OS,s/mS,s the residue field of S at s. As is customary, various categories will be denoted by indicating their objects within parentheses when it will be clear what the morphisms in the category are. For instance (sets), (A-modules), etc. The class of objects of a category C will be denoted by ob(C). The dual of a category C will be denoted by C◦. Given categories C and D, a contravariant functor F from C to D will be always denoted as a covariant functor F : C◦ → D. We will consider the following categories of k-algebras:

A = the category of local artinian k-algebras with residue field k Aˆ = the category of complete local noetherian k-algebras with residue field k A∗ = the category of local noetherian k-algebras with residue field k (k-algebras) = the category of noetherian k-algebras

Morphisms are unitary k-homomorphisms, which are local in A, Aˆ and A∗.Fora given Λ in ob(A∗) we will consider the following:

AΛ = the category of local artinian Λ-algebras with residue field k ∗ AΛ = the category of local noetherian Λ-algebras with residue field k 2 Terminology and notation

They are subcategories of A and A∗ respectively. If Λ is in ob(Aˆ) then we will let ˆ AΛ = the category of complete local noetherian Λ-algebras with residue field k which is a subcategory of Aˆ. Moreover, we will set:

(schemes) = the category of schemes

(i.e. of locally noetherian separated k-schemes) and

(algschemes) = the category of algebraic schemes

For a given scheme Z we set

(schemes/Z) = the category of Z-schemes (algschemes/Z) = the category of algebraic Z-schemes hi (X, F) denotes dim[H i (X, F)] where F is a coherent on the complete scheme X. When no confusion is possible we will sometimes write H i (F) and hi (F) i i instead of H (X, F) and h (X, F) respectively. i Xi denotes the disjoint union of the schemes Xi . If E is a vector space or a locally free sheaf we will always denote its dual by E∨. If V is a k-vector space we will denote by IP (V ) the projective space Proj(Sym(V ∨)) (where Sym(−) is the symmetric algebra of −): thus the closed points of IP (V ) are the one-dimensional subspaces of V . Similarly, if E is a locally free sheaf on an algebraic scheme S, the projective bundle associated to E will be defined as

∨ IP (E) = Proj(Sym(E ))

Note that this definition is dual to the one given in [84], p. 162. For all definitions not explicitly given we will refer to [84]. Introduction

La methode´ gen´ erale´ consiste toujours a` faire des constructions formelles, ce qui consiste essentiellement a` faire de la geom´ etrie´ algebrique´ sur un anneau artinien,eta` en tirer des conclusions de nature “algebrique”´ en utilisant les trois theor´ emes` fondamentaux (Grothendieck [71], p. 11).

Deformation theory is a formalization of the Kodaira–Nirenberg–Spencer–Kuranishi (KNSK) approach to the study of small deformations of complex manifolds. Its main ideas are clearly outlined in the series of Bourbaki seminar expositions by Grothendieck which go under the name of “Fondements de la Geom´ etrie´ Algebrique”´ [2]; in particular they are explained in detail in [72] (see especially page 17), while the technical foundations are laid in [71]. The quotation at the top of this page gives a concise description of the method employed. The first step of this formalization consists in studying infinitesimal deforma- tions, and this is accomplished via the notion of “functor of Artin rings”; the study of such functors leads to the construction of “formal deformations”. This method enhances the analogies between the analytic and the algebraic cases, and at the same time hides some delicate phenomena typical of the algebraic geometrical world. These phenomena become visible when one tries to pass from formal to algebraic deformations. The techniques of deformation theory have a variety of applications which make them an extremely useful tool, especially in understanding the local structure of schemes defined by geometrical conditions or by functorial construc- tions. In this introduction we shall explain in outline the logical structure of deforma- tion theory; for this purpose we will start by outlining the KNSK theory of small deformations of compact complex manifolds. Given a compact complex manifold X,afamily of deformations of X is a com- mutative diagram of holomorphic maps between complex manifolds X ⊂ X ξ : ↓↓π t  −→o B with π proper and smooth (i.e. with everywhere surjective differential), B connected −1 and where  denotes the singleton space. We denote by Xt the fibre π (t), t ∈ B. It is a standard fact that, locally on B, X is differentiably a product so that π can be viewed locally as a family of complex structures on the differentiable manifold Xdiff. 4 Introduction

The family ξ is trivial at to if there is a neighbourhood U ⊂ B of to such that we have π −1(U) =∼ X × U analytically. ∂ ∈ Kodaira and Spencer started by defining, for every tangent vector ∂t Tto B, the π ∂ derivative of the family along ∂t as an element ∂X t ∈ H 1(X, T ) ∂t X thus giving a linear map κ : → 1( , ) Tto B H X TX called the Kodaira–Spencer map of the family π. They showed that if π is trivial at to κ( ∂ ) = ∂ ∈ then ∂t 0 for all ∂t Tto B. Then they investigated the problem of classifying all small deformations of X, by constructing a “complete family” of deformations of X. A family ξ as above is called complete if for every other family of deformations of X: X ⊂ Y η : ↓↓p m  −→o M there is an open neighbourhood V ⊂ M and a commutative diagram X

p−1(V ) → X ↓↓ V → B

−1 ∼ inducing an isomorphism p (V ) = V ×B X . The family is called universal if it is complete and moreover, the morphism V → B is unique locally around mo for each family η as above. Kodaira and Spencer proved that if κ is surjective then the family ξ is complete. The following existence result was then proved: 2 Theorem 0.0.1 (Kodaira–Nirenberg–Spencer [106]). If H (X, TX ) = 0 then there exists a complete family of deformations of X whose Kodaira–Spencer map is an 0 isomorphism. If, moreover, H (X, TX ) = 0 then such complete family is universal. Later Kuranishi [111] generalized this result by showing that a complete family of deformations of X such that κ is an isomorphism exists without assumptions on 2 H (X, TX ) provided the base B is allowed to be an analytic space. We want to rephrase everything algebraically as far as possible. Let’s fix an alge- braically closed field k and consider an algebraic k-scheme X.Alocal deformation, or a local family of deformations,ofX is a cartesian diagram X → X ξ : ↓↓π Spec(k) ⊂ S where π is a flat morphism, S = Spec(A) where A is a local k-algebra with residue field k, and X is identified with the fibre over the closed point. If X is nonsingular Introduction 5 and/or projective we will require π to be smooth and/or projective. We say that ξ is a deformation over Spec(A) or over A. If in particular A is an artinian local k-algebra then we speak of an infinitesimal deformation. The notion of local family has the fundamental property of being funtorial.Given two deformations of X: X → X X → Y ξ : ↓↓π and η : ↓↓ρ Spec(k) ⊂ Spec(A) Spec(k) ⊂ Spec(A) parametrized by the same Spec(A), an isomorphism ξ =∼ η is defined to be a mor- phism f : X → Y of schemes over Spec(A) inducing the identity on the closed fibre, i.e. such that the following diagram

X

X −→f Y

Spec(A) is commutative. Consider the category ∗ A = (noetherian local k-algebras with residue field k) and its full subcategory

A = (artinian local k-algebras with residue field k)

One defines a covariant functor ∗ DefX : A → (sets) by DefX (A) ={local deformations of X over Spec(A)}/(isomorphism) This is the functor of local deformations of X; its restriction to A is the functor of infinitesimal deformations of X. One may now ask whether DefX is representable, namely if there is a noetherian local k-algebra O and a local deformation X → X ◦ υ : ↓↓p Spec(k) ⊂ Spec(O) which is universal, i.e. such that any other local deformation ξ is obtained by pulling back υ under a unique Spec(A) → Spec(O). The approach of Grothendieck to this problem was to formalize the method of Kodaira and Spencer, which consists in a formal construction followed by a proof of convergence. In the search for the universal deformation υ the formal construction 6 Introduction corresponds to the construction of the sequence of its restrictions to the truncations n+1 Spec(O/mO ): → X ◦ X n un : ↓↓ n ≥ 0 n+1 Spec(k) → Spec(O/mO ) n+1 These are infinitesimal deformations of X because the rings O/mO are in A.The sequence uˆ ={un} can be considered as a formal approximation of υ. It is a special case of a formal deformation: more precisely, a formal deformation of X is given by a complete local k-algebra R with residue field k and by a sequence of infinitesimal deformations X → Xn ξn : ↓↓ n ≥ 0 ( ) → ( / n+1) Spec k Spec R m R ξ → ξ / n+1 → / n = Oˆ such that n n−1 under the truncation R m R R m R. In our case R . The goal of the formal step in deformation theory is the construction of uˆ for a given X, i.e. of a formal deformation having a suitable universal property which is inherited from the corresponding property of υ, and which we do not need to specify now. Observe that in trying to perform the formal step we will at best succeed in de- scribing Oˆ and not O. Since a formal deformation consists of infinitesimal deforma- tions, for the construction of uˆ we will only need to work with the covariant functor

DefX : A → (sets)

A covariant functor F : A → (sets) is called a functor of Artin rings.Toevery complete local k-algebra R we can associate a functor of Artin rings h R by

h R(A) = HomA(R, A)

A functor of this form is called prorepresentable. By categorical general nonsense one shows that a formal deformation ξˆ defines a morphism of functors (a natural transformation) h R → DefX and that this morphism is an isomorphism precisely when ξˆ is universal. Therefore we see that the search for uˆ is a problem of prorep- resentability of DefX . More generally, to every local deformation problem there cor- responds a functor of Artin rings F analogous to DefX ; the task of constructing a formal universal deformation for the given problem consists in showing that F is prorepresentable, producing the ring R prorepresenting F and the formal uni- versal deformation defining the isomorphism h R → F. This is the scheme of ap- proach to the formal part of every local deformation problem as it was outlined by Grothendieck. What one needs is to find criteria for the prorepresentability of a functor of Artin rings; we will also need to consider properties weaker than prorep- resentability (semiuniversality) satisfied by more general classes of functors coming from interesting deformation theoretic problems. Necessary and sufficient conditions of prorepresentability and of semiuniversality are given by Schlessinger’s theorem. After having solved the problem of existence of a formal universal (or semiuni- versal) deformation by means of necessary and sufficient conditions for its existence Introduction 7 one still has to decide whether O and υ exist and to find them. To pass from Oˆ to O is the analogue of the convergence step in the Kodaira–Spencer theory, and it is a very difficult problem, the algebraization problem. Under reasonably general as- sumptions one shows that there exists a deformation υ over an algebraic local ring (i.e. the henselization of a local k-algebra essentially of finite type) which does not quite represent the functor DefX but at least has a universal (or semiuniversal) associ- ated formal deformation. The further property of representing DefX is not in general satisfied by (O,υ), being related to the existence of nontrivial automorphisms of X. This part of the theory is largely due to the work of M. Artin, and based on the notions of effectivity of a formal deformation and of local finite presentation of a functor, al- ready introduced by Grothendieck. The main technical tool is Artin’s approximation theorem. 1 Infinitesimal deformations

The purpose of this chapter is to introduce the reader to deformation theory in an elementary and direct fashion. We will be especially interested in first-order defor- mations and obstructions and in giving them appropriate interpretation mostly by elementary Cechˇ cohomology computations. We will start by introducing some al- gebraic tools needed. For other notions used we refer the reader to the appendices.

1.1 Extensions

1.1.1 Generalities

Let A → R be a ring homomorphism. An A-extension of R (or of R by I )isan exact sequence ϕ (R ,ϕ): 0 → I → R −→ R → 0 where R is an A-algebra and ϕ is a homomorphism of A-algebras whose kernel I is an ideal of R satisfying I 2 = (0). This condition implies that I has a structure of R-module. (R ,ϕ)is also called an extension of A-algebras. If (R ,ϕ) and (R ,ψ) are A-extensions of R by I ,anA-homomorphism ξ : R → R is called an isomorphism of extensions if the following diagram commutes: 0 → I → R → R → 0 ↓ξ  0 → I → R → R → 0 Such a ξ is necessarily an isomorphism of A-algebras. More generally, given A- extensions (R ,ϕ)and (R ,ψ)of R, not necessarily having the same kernel, a homo- morphism of A-algebras r : R → R such that ψr = ϕ is called a homomorphism of extensions. The following lemma is immediate. Lemma 1.1.1. Let (R ,ϕ) be an extension as above. Given an A-algebra B and two A-homomorphisms f1, f2 : B → R such that ϕ f1 = ϕ f2 the induced map 10 1 Infinitesimal deformations f2 − f1 : B → I is an A-derivation. In particular, given two homomorphisms of extensions

r1, r2 : (R ,ϕ)→ (R ,ψ) the induced map r2 − r1 : R → ker(ψ) is an A-derivation.

The A-extension (R ,ϕ)is called trivial if it has a section, that is, if there exists a homomorphism of A-algebras σ : R → R such that ϕσ = 1R. We also say that (R ,ϕ)splits, and we call σ a splitting. Given an R-module I , a trivial A-extension of R by I can be constructed by considering the A-algebra R⊕˜ I whose underlying A-module is R ⊕ I and with mul- tiplication defined by: (r, i)(s, j) = (rs, rj + si) The first projection p : R⊕˜ I → R defines an A-extension of R by I which is trivial: a section q is given by q(r) = (r, 0). The sections of p can be identified with the A-derivations d : R → I . Indeed, if we have a section σ : R → R⊕˜ I with σ(r) = (r, d(r)) then for all r, r ∈ R:

σ(rr ) = (rr , d(rr )) = σ(r)σ (r ) = (r, d(r))(r , d(r )) = (rr , rd(r ) + r d(r)) and if a ∈ A then:

σ(ar) = (ar, d(ar)) = aσ(r) = a(r, d(r)) = (ar, ad(r)) hence d : R → I is an A-derivation. Conversely, every A-derivation d : R → I ˜ defines a section σd : R → R⊕I by σd (r) = (r, d(r)). Every trivial A-extension (R ,ϕ) of R by I is isomorphic to (R⊕˜ I, p).If σ : R → R is a section an isomorphism ξ : R⊕˜ I → R is given by:

ξ((r, i)) = σ(r) + i and its inverse is ξ −1(r ) = (ϕ(r ), r − σϕ(r )) An A-extension (P, f ) of R will be called versal if for every other A-extension (R ,ϕ) of R there is a homomorphism of extensions r : (P, f ) → (R ,ϕ).If R = P/I where P is a polynomial algebra over A then

0 → I/I 2 → P/I 2 → R → 0 is a versal A-extension of R. Therefore, since such a P always exists, we see that every A-algebra R has a versal extension. 1.1 Extensions 11

Examples 1.1.2. (i) Every A-extension of A is trivial because by definition it has a section. Therefore it is of the form A⊕˜ V for an A-module V . In particular, if t is an indeterminate the A-extension A[t]/(t2) of A is trivial, and is denoted by A[ ] (where = t mod (t2) satisfies 2 = 0). The corresponding exact sequence is:

0 → ( ) → A[ ]→A → 0

A[ ] is called the algebra of dual numbers over A. (ii) Assume that K is a field. If R is a local K -algebra with residue field K a K -extension of R by K is called a small extension of R.Let

f (R , f ) : 0 → (t) → R −→ R → 0 beasmallK -extension; in other words t ∈ m R is annihilated by m R so that (t) is a K -vector space of dimension one. (R , f ) is trivial if and only if the surjective linear map induced by f :

m m f : R → R 1 2 2 m R m R is not bijective. Indeed for the trivial K -extension

0 → (t) → R⊕˜ (t) → R → 0

2 we have t ∈ m ˜ \ m , hence the map f is not injective because f (t¯) = 0. R⊕(t) R⊕˜ (t) 1 1 ¯ ¯ Conversely, if f1 is not injective then f1(t) = 0; choose a vector subspace

⊂ / 2 / 2 = ⊕ (¯) ⊂ U m R m R such that m R m R U t and let V R be the subring generated by U. Then V is a subring mapped isomorphically onto R by f .Thein- verse of f|V is a section of f , therefore (R , f ) is trivial. For example, it follows from this criterion that the extension of K -algebras

(tn) K [t] K [t] 0 → → → → 0 (tn+1) (tn+1) (tn) n ≥ 2, is nontrivial. (iii) Let K be a field. The K -algebra

K [ , ]:=K [t, t ]/(t, t )2 is a K -extension of K [ ] by K in two different ways. The first

p 0 → ( ) → K [ , ] −→ K [ ]→0 12 1 Infinitesimal deformations is a trivial extension, isomorphic to p∗((K [ ], p )):

0 → ( ) → K [ ]×K K [ ]→K [ ]→0 ↓ ↓p p 0 → ( ) → K [ ] −→ K → 0 The isomorphism is given by

K [ , ]−→K [ ]×K K [ ] a + b + b −→ (a + b , a + b ) The second way is by “sum”: + 0 → ( − ) → K [ , ] −→ K [ ]→0

a + b + b → a + (b + b )

We leave it as an exercise to show that (K [ , ], +) is isomorphic to (K [ , ], p ).

1.1.2 The module ExA(R, I)

Let A → R be a ring homomorphism. In this subsection we will show how to give an R-module structure to the set of isomorphism classes of extensions of an A-algebra R by a module I , closely following the analogous theory of extensions in an abelian category as explained, for example, in Chapter III of [123]. Let (R ,ϕ) be an A-extension of R by I and f : S → R a homomorphism of A-algebras. We can define an A-extension f ∗(R ,ϕ)of S by I , called the pullback of (R ,ϕ)by f , in the following way: ∗ f (R ,ϕ): 0 → I → R ×R S → S → 0 ↓↓f (R ,ϕ): 0 → I → R → R → 0 where R ×R S denotes the fibred product defined in the usual way. Let λ : I → J be a homomorphism of R-modules. The pushout of (R ,ϕ)by λ is the A-extension λ∗(R ,ϕ)of R by J defined by the following commutative diagram: α ϕ 0 → I −→ R −→ R → 0 ↓ λ ↓ → → → → 0 J R I J R 0 where R ⊕˜ J R J = {(−α(i), λ(i)), i ∈ I } I Definition 1.1.3. For every A-algebra R and for every R-module I we define

ExA(R, I ) to be the set of isomorphism classes of A-extensions of R by I . If (R ,ϕ) is such an extension we will denote by [R ,ϕ]∈ExA(R, I ) its class. 1.1 Extensions 13

Using the operations of pullback and pushout it is possible to define an R-module structure on ExA(R, I ).

If r ∈ R and [R ,ϕ]∈ExA(R, I ) we define

r[R ,ϕ]=[r∗(R ,ϕ)] where r : I → I is the multiplication by r.

Given [R ,ϕ], [R ,ψ]∈ExA(R, I ), to define their sum we use the following diagram: 000 ↓↓ I ⊕ II= I ↓ ↓

0 → I → R ×R R → R → 0 ↓ ↓ 0 → I → R → R → 0 ↓↓ 000 which defines an A-extension:

ζ (R ×R R ,ζ): 0 → I ⊕ I → R ×R R −→ R → 0 We define

[R ,ϕ]+[R ,ψ]:=[δ∗(R ×R R ,ζ)] where δ : I ⊕ I → I is the “sum homomorphism”: δ(i ⊕ j) = i + j. Proposition 1.1.4. Let A → R be a ring homomorphism and I an R-module. With the operations defined above ExA(R, I ) is an R-module whose zero element is [R⊕˜ I, p]. This construction defines a covariant functor: (R-modules) −→ (R-modules) I −→ ExA(R, I ) ( f : I → J) −→ ( f∗ : ExA(R, I ) → ExA(R, J)) Proof. Straightforward.  It is likewise straightforward to check that if f : R → S is a homomorphism of A-algebras and I is an S-module, then the operation of pullback induces an applica- tion: ∗ f : ExA(S, I ) → ExA(R, I ) which is a homomorphism of R-modules. We have the following useful result. Proposition 1.1.5. Let A be a ring, f : S → R a homomorphism of A-algebras and let I be an R-module. Then there is an exact sequence of R-modules: ρ 0 → DerS(R, I ) → DerA(R, I ) → DerA(S, I ) ⊗S R −→ v f ∗ → ExS(R, I ) −→ ExA(R, I ) −→ ExA(S, I ) ⊗S R 14 1 Infinitesimal deformations

Proof. v is the obvious application sending an S-extension to itself considered as an A-extension. An A-extension ϕ 0 → I → R −→ R → 0 is also an S-extension if and only if there exists f : S → R such that the triangle R → R ↑ S commutes, and this is equivalent to saying that f ∗(R ,ϕ)is trivial. This proves the exactness in ExA(R, I ). The homomorphism ρ is defined by letting ρ(d) = (R⊕˜ I, p) where the structure of S-algebra on R⊕˜ I is given by the homomorphism s → ( f (s), d(s)) Clearly, vρ = 0. On the other hand, for ϕ (R ,ϕ): 0 → I → R −→ R → 0 ↑ S to define an element of ker(v) there must exist an isomorphism of A-algebras R → R⊕˜ I inducing the identity on I and on R. Hence the composition S → R → R⊕˜ I is of the form

s → ( f (s), d(s)) for some d ∈ DerA(S, I ): therefore the sequence is exact at ExS(R, I ). To prove the ˜ exactness at DerA(S, I ) note that ρ(d) = 0 if and only if p : R⊕I → R has a section as a homomorphism of S-algebras, if and only if there exists an A-derivation R → I whose restriction to S is d: this proves the assertion. The exactness at DerS(R, I ) and DerA(R, I ) is straightforward. 

Definition 1.1.6. The R-module ExA(R, R) is called the first cotangent module of R 1 = 1 over A and it is denoted by TR/A. In the case A k we will write TR instead of 1 TR/k. Proposition 1.1.7. Let A → B be an e.f.t. ring homomorphism and let B = P/J where P is a smooth A-algebra. Then for every B-module M we have an exact sequence: 2 DerA(P, M) → HomB(J/J , M) → ExA(B, M) → 0 (1.1)

If A → B is a smooth homomorphism then ExA(B, M) = 0 for every B-module M. Proof. We have a natural surjective homomorphism 2 HomB(J/J , M) → ExA(B, M) λ → λ∗(η) 1.1 Extensions 15 where η : 0 → J/J 2 → P/J 2 → B → 0

The surjectivity follows from the fact that η is versal. The extension λ∗(η) is trivial if and only if we have a commutative diagram 0 → J/J 2 → P/J 2 → B → 0 ↓ λ ↓ 0 → M → B⊕˜ M → B → 0 if and only if λ extends to an A-derivation D¯ : P/J 2 → M, equivalently to an A-derivation D : P → M. The last assertion is immediate (see Theorem C.9).  Corollary 1.1.8. If A → B is an e.f.t. ring homomorphism and M is a finitely gener- ( , ) 1 ated B-module then ExA B M is a finitely generated B-module. In particular TB/A is a finitely generated B-module and we have an exact sequence:

0 → HomB(ΩB/A, M) → HomB(ΩP/A ⊗P B, M) → (1.2) 2 → HomB(I/I , M) → ExA(B, M) → 0 if B = P/J for a smooth A-algebra P and an ideal J ⊂ P. Proof. It is a direct consequence of the exact sequence (1.1). 

1.1.3 Extensions of schemes

Let X → S be a morphism of schemes. An extension of X/S is a closed immersion

X ⊂ X , where X is an S-scheme, defined by a sheaf of ideals I ⊂ OX such 2 that I = 0. It follows that I is, in a natural way, a sheaf of OX -modules, which coincides with the conormal sheaf of X ⊂ X . To give an extension X ⊂ X of X/S is equivalent to giving an exact sequence on X: ϕ E : 0 → I → OX −→ OX → 0 2 where I is an OX -module, ϕ is a homomorphism of OS-algebras and I = 0in OX ; we call E an extension of X/SbyI or with kernel I. Two such extensions OX and OX are called isomorphic if there is an OS-homomorphism α : OX → OX inducing the identity on both I and OX . It follows that α must necessarily be an S-isomorphism. We denote by Ex(X/S, I) the set of isomorphism classes of extensions of X/S with kernel I. In the case where Spec(B) → Spec(A) is a morphism of affine schemes and I = M˜ we have an obvious identification:

ExA(B, M) = Ex(X/S, I)

If S = Spec(A) is affine we will sometimes write ExA(X, I) instead of Ex(X/Spec(A), I). Exactly as in the affine case one proves that Ex(X/S, I) is a Γ(X, OX )-module with identity element the class of the trivial extension: ˜ 0 → I → OX ⊕I → OX → 0 16 1 Infinitesimal deformations ˜ where OX ⊕I is defined as in the affine case (see Section 1.1). The correspondence

I → Ex(X/S, I) defines a covariant functor from OX -modules to Γ(X, OX )-modules. In deformation theory the case I = OX is the most important one, being related to first-order deformations. If, more generally, I is a locally free sheaf we get the notions of ribbon, carpet etc. (see [17]). Using the fact that the exact sequence (1.2) of page 15 localizes, it is immediate to check that the cotangent module localizes. More specifically, it is straightforward to show that given a morphism of finite type of schemes f : X → S one can define a 1 = ( ) quasi-coherent sheaf TX/S on X with the following properties. If U Spec A is an affine open subset of S and V = Spec(B) is an affine open subset of f −1(U), then Γ( , 1 ) = 1 V TX/S TB/A

1 1 It follows from the properties of the cotangent modules that TX/S is coherent. TX/S is / 1 = ( ) called the first cotangent sheaf of X S. We will write TX if S Spec k . For future reference it will be convenient to state the following:

1 Proposition 1.1.9. (i) If X is an algebraic scheme then TX is supported on the sin- gular locus of X. More generally, if X → S is a morphism of finite type of 1 algebraic schemes then TX/S is supported on the locus where X is not smooth over S. (ii) If we have a closed embedding X ⊂ Y with Y nonsingular, then we have an exact sequence of coherent sheaves on X: → → → → 1 → 0 TX TY |X NX/Y TX 0 (1.3) = [ → 1] so that, letting NX/Y ker NX/Y TX , we have the short exact sequence → → → → 0 TX TY |X NX/Y 0 (1.4)

NX/Y is called the equisingular normal sheaf of X in Y. (iii) For every scheme S and morphism of S-schemes f : X → Y we have an exact sequence of sheaves → → → ( ∗Ω1 , O ) → 1 → 1 → ∗ 1 0 TX/Y TX/S Hom f Y/S X TX/Y TX/S f TY/S (1.5) (iv) When S = Spec(k) and f is a closed embedding of algebraic schemes, with Y = = 1 nonsingular, we have TX/Y 0 and NX/Y TX/Y . Moreover, (1.3) is a special case of (1.5) in this case.

Proof. (i) Use Proposition 1.1.7. (ii) (1.3) globalizes the exact sequence (1.2). (iii) (1.5) globalizes the exact sequence of Proposition 1.1.5. (iv) Follows from (1.2) and 1.1.5.  1.1 Extensions 17

Note that the first half of the exact sequence (1.5) is the dual of the cotangent sequence of f . For more about (1.5) see also (3.43), page 162. The following is a basic result:

Theorem 1.1.10. Let X → S be a morphism of finite type of algebraic schemes and I a coherent locally free sheaf on X. Assume that X is reduced and S-smooth on a dense open subset. Then there is a canonical identification

Ex(X/S, I) = Ext1 (Ω1 , I) OX X/S which to the isomorphism class of an extension of X/S:

E : 0 → I → OX → OX → 0 associates the isomorphism class of the relative conormal sequence of X ⊂ X :

: → I →δ (Ω1 ) → Ω1 → cE 0 X /S |X X/S 0 (which is exact also on the left).

Proof. Suppose given an extension E. Since I is locally free and X is reduced in order to show that cE is exact on the left it suffices to prove that ker(δ) is torsion, equivalently that cE is exact at every general closed point x of any irreducible com- ponent of X. Since X is smooth over S at x it follows from 1.1.7 that there is an affine open neighbourhood U of x such that E|U is trivial. From Theorem B.3 we de- duce that the relative conormal sequence of E|U is split exact. Since it coincides with the restriction of cX to U we see that δ|U is injective; this shows that ker(δ) is tor- sion and cE is exact. Since isomorphic extensions have isomorphic relative cotangent sequences we have a well-defined map

1 1 c− : Ex(X/S, I) → Ext (Ω , I) OX X/S

Now let η : → I → A →p Ω1 → 0 X/S 0 define an element of Ext1 (Ω1 , I). Letting d : O → Ω1 be the canonical OX X/S X X/S derivation, consider the sheaf of OS algebras O = A ×Ω1 OX : over an open subset X/S U ⊂ X we have Γ(U, O) ={(a, f ) : p(a) = d( f )} and the multiplication rule is

(a, f )(a , f ) = ( fa + f a, ff )

Then we have an exact commutative diagram

0 → I → O → OX → 0 ↓d¯ ↓ d → I → A → Ω1 → 0 X/S 0 18 1 Infinitesimal deformations ¯ where one immediately checks that the projection d is an OS-derivation and therefore it must factor as 1 O → Ω ⊗O O → A O/OS X and we have an exact commutative diagram

0 → I → O → OX → 0 ↓ ↓d I → Ω1 ⊗ O → Ω1 → O/O O X X/S 0 (1.6) ↓S  → I → A → Ω1 → 0 X/S 0 1 ∼ which implies Ω ⊗O OX = A. Therefore, letting eη be the extension given by O/OS the first row of (1.6), we see that ceη = η. Similarly, one shows that ecE = E for any [E]∈Ex(X/S, I). Therefore c− and e− are inverse of each other and the conclusion follows. 

Corollary 1.1.11. Let X → S be a morphism of finite type of algebraic schemes, smooth on a dense open subset of X. Assume X reduced. Then there is a canonical isomorphism of coherent sheaves on X:

T 1 =∼ Ext1 (Ω1 , O ) X/S OX X/S X In particular, if X is a reduced algebraic scheme then

T 1 =∼ Ext1 (Ω1 , O ) X OX X X and if, moreover, X = Spec(B0) then

1 ∼ 1 T = Ext (Ω / , B ) B0 k B0 k 0 Proof. An immediate consequence of the above theorem.  A closer analysis of the proof of Theorem 1.1.10 shows that without assuming X reduced we only have an inclusion

Ext1 (Ω1 , O ) ⊂ T 1 OX X X X

NOTES 1. An alternative approach to the topics treated in this section can be obtained by means of the so-called “truncated cotangent complex”, first introduced in [76]. A more general version of the cotangent complex was introduced in [120] and later incorporated in general theories of Andre´ [4], Quillen [146] and Tate [181]. Here we will just recall the main facts about the truncated cotangent complex, without entering into any details, with the only purpose of showing the relation to the notions introduced in this section. For details we refer to [94]. Let A be a ring and R an A-algebra. To every A-extension

ϕ η : 0 → I → R −→ R → 0 1.1 Extensions 19 we associate a complex c•(η) of R-modules (also denoted c•(ϕ)) defined as follows:

c0(η) = ΩR /A ⊗R R c1(η) = I cn(η) = (0) n = 0, 1 d1 : c1(η) → c0(η) is the map x → d(x) ⊗ 1. In other words c•(η) consists of the first map in the conormal sequence of ϕ.Ifr : (R ,ϕ)→ (R ,ψ)is a homomorphism of A-extensions then r induces a homomorphism of complexes

c•(r) : c•(ϕ) → c•(ψ) in an obvious way. The following is easy to establish:

• Let r1, r2 : (R ,ϕ) → (R ,ψ) be two homomorphisms of A-extensions of R.Then c•(r1) and c•(r2) are homotopic. As an immediate consequence we have: • if (E, p) and (F, q) are two versal A-extensions of R then the complexes c•(p) and c•(q) are homotopically equivalent.

Definition 1.1.12. Let A be a ring, R an A-algebra and (E, p) a versal A-extension of R. The homotopy class of the complex c•(p) is called the (truncated) cotangent complex of R over A and denoted by Tˇ (R/A).

If R = P/I for a polynomial A-algebra P the A-extension

0 → I/I 2 → P/I 2 → R → 0 (1.7) is versal and therefore the complex

2 δ I/I −→ ΩP/A ⊗P R (1.8) where δ is the map appearing in the conormal sequence of A → P → R,representsTˇ (R/A). If R is e.f.t. and P is replaced by a smooth A-algebra then (1.7) is again a versal A-extension and (1.8) again represents Tˇ (R/A). From the fact that every A-algebra R can be obtained as the quotient of a polynomial A-algebra P it follows that the cotangent complex Tˇ (R/A) exists for every A-algebra R. The cotangent complex can be used to define “upper and lower cotangent functors”, as follows.

Definition 1.1.13. Let A → B be a ring homomorphism, M a B-module, and let d1 ˇ c• ={C1 −→ C0} represent T (B/A). Then for i = 0, 1 the lower cotangent module of B over A relative to M is: ˇ Ti (B/A, M) = Hi (c• ⊗B M) and the upper cotangent module of B over A relative to M is:

Tˇ i (B/A, M) = Hi (Hom(c•, M))

Because of the definition of cotangent complex it follows that the cotangent modules are independent on the choice of the complex c• representing Tˇ (B/A),butonlydependon 20 1 Infinitesimal deformations

A, B, M. Moreover, one immediately checks that the definition is functorial in M and there- fore we have covariant functors: ˇ Ti (B/A, −) : (B-modules) → (B-modules) i = 0, 1 and Tˇ i (B/A, −) : (B-modules) → (B-modules) i = 0, 1 One immediately sees that for i = 0 the cotangent functors are: ˇ T0(B/A, M) = ΩB/A ⊗B M and ˇ 0 T (B/A, M) = DerA(B, M) From the extension (1.7) we obtain the exact sequences:

ˇ 2 0 → T1(B/A, M) → I/I ⊗B M → ΩP/A ⊗P M → ΩB/A ⊗B M → 0 and 0 → HomB(ΩB/A, M) → HomB(ΩP/A ⊗P B, M) →

2 ˇ 1 HomB(I/I , M) → T (B/A, M) → 0 If B is e.f.t. then in (1.7) P can be chosen to be a smooth A-algebra; in this case it follows that ˇ ˇ i Ti (B/A, M) and T (B/A, M) are finitely generated B-modules if M is finitely generated. Moreover, recalling Corollary 1.1.8, we see that we have an identification

ˇ 1 T (B/A, M) = ExA(B, M)

2. The topics of this section originate from [76]. See also [1], Ch. 0IV, § 18. The proof of Theorem 1.1.10 has been taken from [17]; see also [66].

1.2 Locally trivial deformations

1.2.1 Generalities on deformations

Let X be an algebraic scheme. A cartesian diagram of morphisms of schemes

X → X η : ↓↓π s Spec(k) −→ S where π is flat and surjective, and S is connected, is called a family of deforma- tions,orsimplyadeformation,ofX parametrized by S, or over S; we call S and X respectively the parameter scheme and the total scheme of the deformation. If S is al- gebraic, for each k-rational point t ∈ S the scheme-theoretic fibre X (t) is also called a deformation of X. When S = Spec(A) with A in ob(A∗) and s ∈ S is the closed point we have a local family of deformations (shortly a local deformation)ofX over A. The deformation η will be also denoted by (S,η)or (A,η)when S = Spec(A). 1.2 Locally trivial deformations 21

The local deformation (A,η)is infinitesimal (resp. first-order)ifA ∈ ob(A) (resp. A = k[ ]). Given another deformation X → Y ξ : ↓↓ Spec(k) → S of X over S, an isomorphism of η with ξ is an S-isomorphism φ : X → Y inducing the identity on X, i.e. such that the following diagram is commutative: X

φ X −→ Y

S By a pointed scheme we will mean a pair (S, s) where S is a scheme and s ∈ S.IfK is a field we call (S, s) a K -pointed scheme if K =∼ k(s). Observe that for every X and for every k-pointed scheme (S, s) there exists at least one family of deformation of X over S, namely the product family: X → X × S ↓↓ s Spec(k) −→ S A deformation of X over S is called trivial if it is isomorphic to the product family. It will be also called a trivial family with fibre X. All fibres over k-rational points of a trivial deformation of X parametrized by an algebraic scheme are isomorphic to X. The converse is not true: there are deformations which are not trivial but have isomorphic fibres over all the k-rational points (see Example 1.2.2(ii) below). The scheme X is called rigid if every infinitesimal deformation of X over A is trivial for every A in ob(A). Given a deformation η of X over S as above and a morphism (S , s ) → (S, s) of k-pointed schemes there is induced a commutative diagram by base change

X → X ×S S ↓↓ Spec(k) → S which is clearly a deformation of X over S . This operation is functorial, in the sense that it commutes with composition of morphisms and the identity morphism does not change η. Moreover, it carries isomorphic deformations to isomorphic ones. An infinitesimal deformation η of X is called locally trivial if every point x ∈ X has an open neighbourhood Ux ⊂ X such that → X Ux |Ux ↓↓ Spec(k) → S is a trivial deformation of Ux . 22 1 Infinitesimal deformations

Remark 1.2.1. Let j X −→ X η : ↓↓π s Spec(k) −→ S be a family of deformations of an algebraic scheme X parametrized by an algebraic scheme S and let Z ⊂ X be a proper closed subset. Then

j X\Z −→ X \ j(Z) ↓↓π s Spec(k) −→ S is a family of deformations of X\Z having the same fibres as π over t ∈ S for t = s: thus such fibres are deformations both of X and of X\Z. This shows that the defi- nition of family of deformations given above is somewhat ambiguous unless we as- sume that π is projective or that the deformation is infinitesimal. In what follows we will restrict to the consideration of deformations of projective schemes and/or of in- finitesimal deformations when discussing the general theory, so that such ambiguity will be removed; only occasionally will we consider non-infinitesimal deformations of affine schemes.

Examples 1.2.2. (i) The quadric Q ⊂ A3 of equation xy − t = 0 defines, via the projection A3 → A1 (x, y, t) → t a flat family Q → A1 whose fibres are affine conics. This family is not trivial since the fibre Q(0) is singular, hence not isomorphic to the fibres Q(t), t = 0, which are nonsingular. (ii) Consider, for a given integer m ≥ 0, the rational ruled surface = (O ( ) ⊕ O ) Fm IP IP 1 m IP 1

1 The structural morphism π : Fm → IP defines a flat family whose fibres are all 1 ∼ isomorphic to IP ;butifm > 0 then π is not a trivial family because Fm = F0 = IP 1 × IP 1 (see Example B.11(iii)). (iii) Let 0 ≤ n < m be two distinct nonnegative integers having the same parity = 1 ( − ) 2 × 1 and let k 2 m n . Consider two copies of A IP given as ( [ , ,ξ ,ξ ]) =: ( [ , ,ξ ,ξ ]) =: Proj k t z 0 1 W and Proj k t z 0 1 W (here the rings are graded ξ ξ ξ = ξ /ξ ξ = ξ /ξ with respect to the variables i and i ). Letting 1 0 and 1 0 consider the open subsets

Spec(k[t, z,ξ]) ⊂ W, Spec(k[t, z ,ξ ]) ⊂ W and glue them together along the open subsets

− Spec(k[t, z, z 1,ξ]) ⊂ Spec(k[t, z,ξ]) 1.2 Locally trivial deformations 23 and − Spec(k[t, z , z 1,ξ ]) ⊂ Spec(k[t, z ,ξ ]) according to the following rules:

− z = z 1,ξ= zmξ + tzk (1.9)

This induces a gluing of W and W along ( [ , , −1,ξ ,ξ ]) ( [ , , −1,ξ ,ξ ]) Proj k t z z 0 1 and Proj k t z z 0 1 Call the resulting scheme W and f : W → A1 = Spec(k[t]) the morphism induced by the projections. Then f is a flat morphism because it is locally a projection; moreover, ∼ W(0) = Fm Let W◦ = f −1(A1\{0}) and f ◦ : W◦ → A1\{0} the restriction of f . In k[t, t−1, z,ξ] define kξ − ζ = z t tξ and in k[t, t−1, z ,ξ ] ξ ζ = tz m−kξ + t2 It is straightforward to verify that the gluing (1.9) induces the relation

ζ = znζ

This means that we have an isomorphism

◦ ∼ 1 W = Fn × (A \{0}) compatible with the projections to A1\{0}. Therefore the family f ◦ is trivial, in par- ticular all its fibres are isomorphic to Fn, but the family f is not trivial because ∼ W(0) = Fm. (iv) Let f : X → Y be a surjective morphism of algebraic schemes, with X integral and Y an irreducible and nonsingular curve. Then f is flat. This is a special case of Prop. III.9.7 of [84]. Therefore f defines a family of deformations of any of its closed fibres.

1.2.2 Infinitesimal deformations of nonsingular affine schemes

We will start by considering infinitesimal deformations of affine schemes. We need the following:

Lemma 1.2.3. Let Z0 be a closed subscheme of a scheme Z, defined by a sheaf of nilpotent ideals N ⊂ OZ .IfZ0 is affine then Z is affine as well. 24 1 Infinitesimal deformations

Proof. Let r ≥ 2 be the smallest integer such that N r = (0). Since we have a chain of inclusions

r−1 r−2 Z ⊃ V (N ) ⊃ V (N ) ⊃···⊃ V (N) = Z0 it suffices to prove the assertion in the case r = 2. In this case N is a coherent O Z0 -module, and therefore

1 1 H (Z, N) = H (Z0, N) = 0

Let R0 be the k-algebra such that Z0 = Spec(R0). We have the exact sequence: 0 0 0 → H (Z, N) → H (Z, OZ ) → R0 → 0

0 Put R = H (Z, OZ ) and let Z = Spec(R). We have a commutative diagram: θ Z −→ Z  Z0 −1 The sheaf homomorphism θ OZ → OZ is clearly injective and θ is a homeomor- −1 phism. It will therefore suffice to prove that θ OZ → OZ is surjective. Let z ∈ Z and f ∈ Γ(U, OZ ) for some affine open neighbourhood U of z.Let ϕ = ϕ ,ψ ∈ = 0 ψ ( ) = f0 f|U∩Z0 . It is possible to find 0 0 R0 such that f0 ψ , 0 z 0 ψ = \ ψ ∈ 0ψ = ψ and 0 0onZ0 U, because Z0 is affine. Let R be such that |Z0 0 (it exists by the surjectivity of R → R0). Then ψ(z) = 0 and ψ = 0onZ\U. There exists n  0 such that ψn f =: g ∈ R (it suffices to cover Z with affines). Then − = g ∈ θ 1O  f ψn Z . Let B0 be a k-algebra, and let X0 = Spec(B0). Consider an infinitesimal defor- mation of X0 parametrized by Spec(A), where A is in ob(A). By definition this is a cartesian diagram X0 → X ↓↓ Spec(k) → Spec(A) where X is a scheme flat over Spec(A). By Lemma 1.2.3 X is necessarily affine. Therefore, equivalently, we can talk about an infinitesimal deformation of B0 over A as a cartesian diagram of k-algebras:

B → B0 ↑↑ (1.10) A → k with A → B flat. Note that to give this diagram is the same as to give A → B flat and a k-isomorphism B ⊗A k → B0. We will sometimes abbreviate by calling A → B the deformation.

Given another deformation A → B of B0 over A, an isomorphism of deforma- tions of A → B to A → B is a homomorphism ϕ : B → B of A-algebras inducing a commutative diagram: 1.2 Locally trivial deformations 25

B0  ϕ B −→ B  A It follows from Lemma A.4 that such a ϕ is an isomorphism. An infinitesimal deformation of B0 over A is trivial if it is isomorphic to the product deformation B0 ⊗k A → B0 ↑↑ A → k

The k-algebra B0 is called rigid if Spec(B0) is rigid. Theorem 1.2.4. Every smooth k-algebra is rigid. In particular, every affine nonsin- gular algebraic variety is rigid.

Proof. Suppose k → B0 is smooth, and suppose given a first-order deformation of B0: B → B0 η0 : ↑ f ↑ k[ ]→ k Consider the commutative diagram:

B → B0 ↑ f ↑ k[ ]→B0[ ] where B0[ ]=B0⊗k[ ]. Since f is smooth (because flat with smooth fibre, see [84], ch. III, Th. 10.2) and the right vertical morphism is a k[ ]-extension, by Theorem C.9 there exists a k[ ]-homomorphism φ : B → B0[ ] making the diagram

B → B0 ↑ f ↑ k[ ]→B0[ ] commutative. Therefore φ is an isomorphism of deformations and η0 is trivial. Consider more generally a deformation of B0

B → B0 η : ↑ f ↑ A → k parametrized by A in ob(A). To show that η is trivial we proceed by induction on d = dimk(A). The case d = 2 has been already proved; assume d ≥ 3 and let

0 → (t) → A → A → 0 26 1 Infinitesimal deformations be a small extension. Consider the commutative diagram: ∼ B → B ⊗A A = B0 ⊗k A ↑ f ↑ A → B0 ⊗k A f is smooth, the upper right isomorphism is by the inductive hypothesis, and the right vertical homomorphism is an A-extension. By the smoothness of f and by Theorem C.9 we deduce the existence of an A-homomorphism B → B0 ⊗k A which is an isomorphism of deformations. 

2 Example 1.2.5. Let λ ∈ k and B0 = k[X, Y ]/(Y − X(X − 1)(X − λ)).Ifλ = 0, 1 then B0 is a smooth k-algebra, being the coordinate ring of a nonsingular plane cubic curve. By Theorem 1.2.4, B0 is rigid. On the other hand, the elementary theory of elliptic curves (see [84]) shows that the following flat family of affine curves Speck[X, Y ]/(Y 2 − X(X − 1)(X − (λ + t))) ↓ Spec(k[t]) is not trivial around the origin t = 0 so that it defines a nontrivial (non-infinitesimal) deformation of B0. This example shows that by studying infinitesimal deformations of affine schemes we are losing some information. In this specific case we will see that this information is recovered by considering the infinitesimal deformations of the projective closure of Spec(B0) (see Corollary 2.6.6, page 94).

1.2.3 Extending automorphisms of deformations

In deformation theory it is very important to have a good control of automorphisms of deformations and of their extendability properties. We will now begin to introduce such matters and to recall some terminology. In Section 2.6 we will consider these problems again in general, and we will relate them with the property of “prorepre- sentability”. Let’s start with a basic lemma.

Lemma 1.2.6. Let B0 be a k-algebra, and

e : 0 → (t) → A˜ → A → 0 a small extension in A. Then there is a canonical isomorphism of groups: ˜ automorphisms of the trivial deformation B0 ⊗k A → Derk(B0, B0) inducing the identity on B0 ⊗k A In particular the group on the left is abelian. ˜ ˜ Proof. Every automorphism θ : B0 ⊗k A → B0 ⊗k A belonging to the first group must be A˜-linear and induce the identity mod t. Therefore:

θ(x) = x + tdx 1.2 Locally trivial deformations 27 ˜ ˜ where d : B0 ⊗k A → B0 is a A-derivation (Lemma 1.1.1). But ˜ Der ˜ (B ⊗ A, B ) = Hom ˜ (Ω ˜ ˜ , B ) A 0 k 0 B0⊗k A B0⊗k A/A 0 = (Ω , ) = ( , ) HomB0 B0/k B0 Derk B0 B0 By sending θ → d we define the correspondence of the statement. Since θ is deter- mined by d the correspondence is one to one. Clearly the identity corresponds to the zero derivation. If we compose two automorphisms: ˜ θ ˜ σ ˜ B0 ⊗k A −→ B0 ⊗k A −→ B0 ⊗k A where θ(x) = x + tdx, σ(x) = x + tδx, we obtain:

σ(θ(x)) = θ(x) + tδ(θ(x)) = x + tdx + t(δx + tδ(dx)) = x + t(dx + δx) therefore the correspondence is a group isomorphism.  Recall the following well-known definition. Definition 1.2.7. Let G be a group acting on a set T and let

π : G × T → T be the map defining the action. T is called a homogeneous space under (the action of) G if π is transitive, i.e. if π(G ×{t}) = T for some (equivalently for any) t ∈ T (i.e. if there is only one orbit). The action is called free if for every point t ∈ T the stabilizer Gt ={g ∈ G : gt = t} is trivial, i.e. gt = t implies g = 1G for all t ∈ T . If the action is both transitive and free then T is called a principal homogeneous space (or a torsor) under (the action of) G. To an action π : G × T → T we can associate the map: p : G × T → T × T

(g, t) → (gt, t) The condition that the action is transitive (resp. free) is equivalent to p being sur- jective (resp. injective); therefore T is a torsor under G if and only if p is bijective. π = Note that pr1 p is determined by p. More generally, suppose that we have a map of sets f : T → T . Then the map p factors through T ×T T ⊂ T × T if and only if the action π is compatible with f , i.e. if f (t) = f (gt) for all t ∈ T , g ∈ G. As before, the map

p : G × T → T ×T T is surjective (resp. injective) if and only if the action of G is transitive (resp. free) on all the non-empty fibres of f . In particular, p is bijective if and only if all the non- empty fibres of f are torsors under G. In this case one also says, according to [3], 28 1 Infinitesimal deformations p. 114, that T over T is a formal principal homogeneous space (or a pseudo-torsor) under G. Now we come back to deformations and we prove a generalization of Lemma 1.2.6.

Lemma 1.2.8. Let B0 be a k-algebra,

e : 0 → (t) → A˜ → A → 0 A ˜ → ˜ → = ˜ ⊗ a small extension in , A B a deformation of B0 and A B B A˜ Athe induced deformation of B0 over A. Let σ : B → B be an automorphism of the deformation. Then: (i) If ( ˜ ) := τ : ˜ → ˜ τ ⊗ = σ =∅ Autσ B automorphisms B B such that A˜ A

then there is a free and transitive action ˜ ˜ Derk(B0, B0) × Autσ (B) → Autσ (B)

defined by (d,τ) → τ + td ˜ (ii) If B0 is a smooth k-algebra then Autσ (B) =∅for any σ .

Proof. (i) Recall that we have a chain of natural identifications

( ˜ , ) = (Ω , ) = (Ω ⊗ , ) DerA˜ B B0 HomB˜ B˜ /A˜ B0 HomB0 B˜ /A˜ A˜ k B0

= (Ω , ) = ( , ) HomB0 B0/k B0 Derk B0 B0 Therefore the action in the statement is well defined once we consider a ˜ ˜ d ∈ Derk(B0, B0) as an A-derivation of B into B0. Given any two elements ˜ τ,η ∈ Autσ (B) we have by definition:

qτ = σq = qη where q : B˜ → B is the projection; hence by Lemma 1.1.1, ˜ η − τ : B → tB0 = ker(q) is an A˜-derivation which is 0 if and only if η = τ. This implies that the action is free and transitive. ˜ ˜ (ii) Since B0 is smooth the deformation A → B is trivial (Theorem 1.2.4), so that ˜ ˜ ∼ ˜ ˜ ˜ we have an A-isomorphism B = B0 ⊗k A; moreover, B0 ⊗k A is a smooth A-algebra (Proposition C.2(iii)) hence B˜ is A˜-smooth. Let σ : B → B be any automorphism of the deformation and consider the diagram of A˜-algebras: 1.2 Locally trivial deformations 29 q σ B˜ → B → B ↑ q B˜ ˜ ˜ Since ker(q) = tB0 is a square-zero ideal, and since B is A-smooth, we deduce that there is σ˜ : B˜ → B˜ such that qσ˜ = σq. It is immediate to check that σ˜ is an ˜ isomorphism and therefore σ˜ ∈ Autσ (B).  In (ii) the condition that B0 is smooth cannot be removed. A simple example is given in 2.6.8(i). The extendability of automorphisms of deformations of not neces- sarily affine schemes will be studied in § 2.6.

1.2.4 First-order locally trivial deformations

We will now apply 1.2.6 to first-order deformations of any algebraic variety.

Proposition 1.2.9. Let X be an algebraic variety. There is a 1–1 correspondence: isomorphism classes of first order 1 κ : → H (X, TX ) locally trivial deformations of X

= (Ω1 , O ) = called the Kodaira–Spencer correspondence, where TX Hom X X Derk(OX , OX ), such that κ(ξ) = 0 if and only if ξ is the trivial deformation class. In particular if X is nonsingular then κ is a 1–1 correspondence isomorphism classes of 1 κ : → H (X, TX ) first-order deformations of X

Proof. Given a first-order locally trivial deformation

X → X ↓↓ Spec(k) → Spec(k[ ]) U ={ } X choose an affine open cover Ui i∈I of X such that |Ui is trivial for all i.For each index i we therefore have an isomorphism of deformations: θ : × ( [ ]) → X i Ui Spec k |Ui by 1.2.4. Then for each i, j ∈ I θ := θ −1θ : × ( [ ]) → × ( [ ]) ij i j Uij Spec k Uij Spec k is an automorphism of the trivial deformation Uij × Spec(k[ ]). By Lemma 1.2.6, θij corresponds to a dij ∈ Γ(Uij, TX ). Since on each Uijk we have

θ θ θ −1 = ij jk ik 1Uijk×Spec(k[ ]) (1.11) 30 1 Infinitesimal deformations it follows that dij + d jk − dik = 0 1 i.e. {dij} is a Cechˇ 1-cocycle and therefore defines an element of H (X, TX ).Itis easy to check that this element does not depend on the choice of the open cover U. If we have another deformation X → X ↓↓ Spec(k) → Spec(k[ ]) and Φ : X → X is an isomorphism of deformations then for each i ∈ I there is induced an automorphism:

− Φ| θ 1 θi Ui i α : U × Spec(k[ ]) −→ X| −→ X −→ U × Spec(k[ ]) i i Ui |Ui i ∈ Γ( , ) θ α = Φ θ and therefore a corresponding ai Ui TX .Wehave i i |Ui i and therefore

−1 −1 −1 −1 (θ α ) (θ α ) = θ Φ Φ| θ = θ θ i i j j i |Uij Uij j i j thus α−1θ α = θ i ij j ij Equivalently: + − = dij a j ai dij { } { } namely, dij and dij are cohomologous, and therefore define the same element of 1 H (X, TX ). 1 Conversely, given θ ∈ H (X, TX ) we can represent it by a Cechˇ 1-cocycle 1 {dij}∈Z (U, TX ) with respect to some affine open cover U. To each dij we can associate an automorphism θij of the trivial deformation Uij×Spec(k[ ]) by Lemma 1.2.6. They satisfy the identities (1.11). We can therefore use these automorphisms to patch the schemes Ui × Spec(k[ ]) by the well-known procedure (see [84], p. 69). We obtain a Spec(k[ ])-scheme X which, by construction, defines a locally trivial first-order deformation of X. The equivalence between κ(ξ) = 0 and the triviality of ξ is easily proved. The last assertion follows from the first one because all deforma- tions of a nonsingular variety are locally trivial by Theorem 1.2.4.  Definition 1.2.10. For every locally trivial first-order deformation ξ of a variety X 1 the cohomology class κ(ξ) ∈ H (X, TX ) is called the Kodaira–Spencer class of ξ. Let X → X ξ : ↓↓f (1.12) s Spec(k) −→ S be a family of deformations of a nonsingular variety X. By pulling back this family by morphisms Spec(k[ ]) → S with image s and applying the Kodaira–Spencer correspondence (Proposition 1.2.9) we define a linear map 1.2 Locally trivial deformations 31

1 κξ : TS,s → H (X, TX ) also denoted by κ f,s or κX /S,s, which is called the Kodaira–Spencer map of the family ξ.

1 Examples 1.2.11. (i) Let m ≥ 1 and let π : Fm → IP be the structural morphism of the rational ruled surface Fm (see B.11(iii)). Then π is not a trivial family but has 1 a trivial restriction around each closed point s ∈ IP , thus κπ,s = 0. (ii) Consider an unramified covering π : X → S of degree n ≥ 2 where X and S are projective nonsingular and irreducible algebraic curves. All fibres of π over the closed points consist of n distinct points, hence they are all isomorphic. More- over, each such fibre is rigid and unobstructed as an abstract variety. In particular the Kodaira–Spencer map is zero at each closed point s ∈ S. On the other hand, π −1(U) is irreducible for each open subset U ⊂ S and therefore the restriction −1 πU : π (U) → U is a nontrivial family; this follows also from the fact that π does not have rational sections. This example exhibits a phenomenon which is not detected by infinitesimal con- siderations and in some sense opposite to the one described in Example 1.2.5: we can have a flat projective family of deformations, all of whose geometric fibres are isomorphic, but which is nevertheless nontrivial over every Zariski open subset of the 1 base. Note that this is different from what happens with the projections Fm → IP , m ≥ 1 of Example (i), which are nontrivial but have trivial restriction to a Zariski open neighbourhood of every point of IP 1. See Subsection 2.6.2 for more about this. (iii) Let 0 ≤ n < m be integers having the same parity, and let k = 1 (m − n). Consider the smooth proper morphism f : W → A1 introduced 2 ∼ ∼ in Example 1.2.2(iii), whose fibres are W(0) = Fm, and W(t) = Fn for t = 0. Recall that the family f is given as the gluing of two copies of A2 × IP 1: = ( [ , ,ξ ,ξ ]), = ( [ , ,ξ ,ξ ]) W Proj k t z 0 1 W Proj k t z 0 1 ( [ , , −1,ξ ,ξ ]) ( [ , , −1,ξ ,ξ ]) along Proj k t z z 0 1 and Proj k t z z 0 1 according to the rules:

z = z−1,ξ = zmξ + tzk ξ = ξ /ξ ξ = ξ /ξ where 1 0 and 1 0 are nonhomogeneous coordinates on the corre- sponding copies of IP 1. κ κ ( d ) Let’s compute the local Kodaira–Spencer map f,0 of f at 0. The image f,0 dt 1( , ) is the element of H Fm TFm corresponding to the first-order deformation of Fm obtained by gluing := ( [ , ,ξ ,ξ ]) := ( [ , ,ξ ,ξ ]) W0 Proj k z 0 1 W0 Proj k z 0 1 ( [ , , −1,ξ ,ξ ]) ( [ , , −1,ξ ,ξ ]) along Proj k z z 0 1 and Proj k z z 0 1 according to the rules

z = z−1,ξ = zmξ + zk 32 1 Infinitesimal deformations

κ ( d ) 1(U, ) By definition we have that f,0 dt is the element of H TFm , where U ={ , } ∩ W0 W0 , defined by the 1-cocycle corresponding to the vector field on W0 W0 ∂ zk ∂ξ

According to Example B.11(iii) this element is nonzero; therefore κ f,0 is injective. Similarly, we can consider a smooth proper family F : Y → Am−1 defined as follows. Y is the gluing of

Y := Proj(k[t1,...,tm−1, z,ξ0,ξ1]) and := ( [ ,..., , ,ξ ,ξ ]) Y Proj k t1 tm−1 z 0 1 ( [ ,..., , , −1,ξ ,ξ ]) ( [ ,..., , , −1,ξ ,ξ ]) along Proj k t1 tm−1 z z 0 1 and Proj k t1 tm−1 z z 0 1 according to the rules:

m−1 −1 m j z = z ,ξ= z ξ + t j z j=1

The morphism F is defined by the projections onto Spec(k[t1,...,tm−1]); the fibre ∼ of F over 0 is Y(0) = Fm. The computation we just did immediately implies that the local Kodaira–Spencer map κ : m−1 → 1( , ) F,0 T0A H Fm TFm is an isomorphism.

1.2.5 Higher-order deformations – obstructions

Let X be a nonsingular algebraic variety. Consider a small extension

e : 0 → (t) → A˜ → A → 0 in A and let X → X ξ : ↓↓ Spec(k) → Spec(A) be an infinitesimal deformation of X.Alifting of ξ to A˜ consists in a deformation

X → X˜ ξ˜ : ↓↓ Spec(k) → Spec(A˜) 1.2 Locally trivial deformations 33 and an isomorphism of deformations X

φ X −→ X˜ × ( ) Spec(A˜) Spec A

Spec(A) If we want to study arbitrary infinitesimal deformations, and not only first-order ones, it is important to know whether, given ξ and e, a lifting of ξ to A˜ exists, and how many there are. Such information can then be used to build an inductive procedure for the description of infinitesimal deformations. The following proposition addresses this question.

Proposition 1.2.12. Given A in ob(A) and an infinitesimal deformation ξ of X over A: 2 (i) To every small extension e of A there is associated an element oξ (e) ∈ H (X, TX ) called the obstruction to lifting ξ to A,˜ which is 0 if and only if a lifting of ξ to A˜ exists. 1 (ii) If oξ (e) = 0 then there is a natural transitive action of H (X, TX ) on the set of isomorphism classes of liftings of ξ to A.˜ (iii) The correspondence e → oξ (e) defines a k-linear map

2 oξ : Exk(A, k) → H (X, TX )

Proof. Let U ={Ui }i∈I be an affine open cover of X. We have isomorphisms θ : × ( ) → X i Ui Spec A |Ui θ := θ −1θ and consequently, ij i j is an automorphism of the trivial deformation Uij × Spec(A). Moreover, θijθ jk = θik (1.13) ˜ ˜ on Uijk × Spec(A). To give a lifting ξ of ξ to A it is necessary and sufficient to give ˜ ˜ a collection of automorphisms {θij} of the trivial deformations Uij × Spec(A) such that ˜ ˜ ˜ (a) θijθ jk = θik ˜ (b) θij restricts to θij on Uij × Spec(A) In fact from such data we will be able to define X˜ by patching the local pieces ˜ ˜ Ui × Spec(A) along the open subsets Uij × Spec(A) in the usual way. To estab- ˜ ˜ lish the existence of the collection {θij} let’s choose arbitrarily automorphisms {θij} satisfying condition (b); they exist by Lemma 1.2.8(ii). Let θ˜ = θ˜ θ˜ θ˜−1 ijk ij jk ik ˜ This is an automorphism of the trivial deformation Uijk×Spec(A). Since by (1.13) it ˜ restricts to the identity on Uijk×Spec(A), by Lemma 1.2.6 we can identify each θijk 34 1 Infinitesimal deformations

˜ ˜ 2 with a dijk ∈ Γ(Uijk, TX ) and it is immediate to check that {dijk}∈Z (U, TX ).If ˜ we choose different automorphisms {Φij} of the trivial deformations Uij × Spec(A) satisfying the analogue of condition (b) then ˜ Φij = θij + tdij (1.14) for some dij ∈ Γ(Uij, TX ), by Lemma 1.2.8(i). For each i, j, k the automorphism Φ Φ Φ−1 ij jk ik corresponds to the derivation ˜ δijk = dijk + (dij + d jk − dik) ˜ and therefore we see that the 2-cocycles {dijk} and {δijk} are cohomologous. Their cohomology class 2 oξ (e) ∈ H (X, TX ) depends only on ξ and e and is 0 if and only if we can find a collection of automor- phisms {Φij} such that δijk = 0 for all i, j, k ∈ I . In such a case {Φij} defines a lifting ξ˜ of ξ. This proves (i). ˜ Assume that oξ (e) = 0, i.e. that the lifting ξ of ξ exists. Then we can choose ˜ the collection {θij} of automorphisms satisfying conditions (a) and (b) as above, in ˜ ¯ ˜ particular dijk = 0, all i, j, k. Any other choice of a lifting ξ of ξ to A corresponds to a choice of automorphisms {Φij} satisfying (1.14) and the analogue of condition (b). Therefore, for all i, j, k,wehave

0 = δijk = dij + d jk − dik

1 ¯ 1 so that {dij}∈Z (U, TX ) defines an element d ∈ H (X, TX ). As before, one checks that this element depends only on the isomorphism class of ξ¯; it follows in a straightforward way that the correspondence (ξ,˜ d¯) → ξ¯ defines a transitive action 1 ˜ of H (X, TX ) on the set of isomorphism classes of liftings of ξ to A. This proves (ii). (iii) is left to the reader. 

Definition 1.2.13. The deformation ξ is called unobstructed if oξ is the zero map; otherwise ξ is called obstructed.Xisunobstructed if every infinitesimal deformation of X is unobstructed; otherwise X is obstructed. Corollary 1.2.14. A nonsingular variety X is unobstructed if

2 H (X, TX ) = (0) The proof is obvious. Corollary 1.2.15. A nonsingular variety X is rigid if and only if

1 H (X, TX ) = (0) 1.2 Locally trivial deformations 35

Proof. The hypothesis implies, by Proposition 1.2.9, that all first-order deformations of X are trivial; moreover, by Proposition 1.2.12(ii), it implies that every infinitesimal deformation of X over any A in ob(A) has at most one lifting to any small extension of A. These two facts together easily give the conclusion. 

Examples 1.2.16. (i) If X is a projective nonsingular curve of genus g then from the Riemann–Roch theorem it follows that 0ifg = 0 1 h (X, TX ) = 1ifg = 1 3g − 3ifg ≥ 2

2 and h (X, TX ) = 0. In particular, projective nonsingular curves are unobstructed. (ii) If X is a projective, irreducible and nonsingular surface X then

2( , ) =∼ 0( ,Ω1 ⊗ )∨ H X TX H X X K X by Serre duality, and this rarely vanishes. For example, a nonsingular surface of de- 3 2 gree ≥ 5inIP satisfies H (X, TX ) = (0), but it is nevertheless unobstructed (see Example 3.2.11(i)); therefore the sufficient condition of Corollary 1.2.14 is not nec- 2 essary. In general a surface such that H (X, TX ) = (0) can be obstructed, but explicit examples are not elementary (see [96], [24], [87]). We will describe a class of such examples in Theorem 3.4.26, page 185. In § 2.4 we will show how to construct exam- ples of obstructed 3-folds (see remarks following Proposition 3.4.25, page 185). The first examples of obstructed compact complex manifolds where given in Kodaira– Spencer [107], § 16: they are of the form T × IP 1, where T is a two-dimensional complex torus. (iii) The projective space IP n is rigid for every n ≥ 1. In fact it follows immedi- ately from the Euler sequence:

n+1 0 → OIP n → OIP n (1) → TIP n → 0

1 n that H (IP , TIP n ) = 0. Similarly, one shows that finite products

IP n1 ×···×IP nk of projective spaces are rigid.

(iv) The ruled surfaces Fm are unobstructed because 2( , ) = h Fm TFm 0

(see (B.13), page 292). 2 Formal deformation theory

In this chapter we develop the theory of “functors of Artin rings”. The main result of this theory is a theorem of Schlessinger giving necessary and sufficient conditions for a functor of Artin rings to have a semiuniversal or a universal formal element. We then apply the functorial machinery to the construction of formal semiuniversal, or universal, deformations, which is the final goal of formal deformation theory, and we explain the relation between formal and algebraic deformations. In order to make the arguments easier to follow, these applications are given only in the case of the deformation functors DefX and DefX of an algebraic scheme. Only in Chapter 3 will we apply the general theory to other deformation functors.

2.1 Obstructions

In this section we investigate the notion of formal smoothness in the category A∗ us- ing the language of extensions. The results we prove are crucial for the understanding of obstructions in deformation theory. Our treatment is an expansion of [157]; for a more systematic treatment we refer to [54]. ∗ ∗ Let Λ ∈ ob(A ) and µ : Λ → R be in ob(AΛ).Therelative obstruction space of R/Λ is o(R/Λ) := ExΛ(R, k) If Λ = k then o(R/k) is called the (absolute) obstruction space of R and sim- ply denoted by o(R). We say that R is unobstructed (resp. obstructed) over Λ if o(R/Λ) = (0) (resp. if o(R/Λ) = (0)); R is said to be unobstructed (resp. ob- structed)ifo(R) = (0) (resp. if o(R) = (0)). Given a homomorphism f : R → S ∗ in AΛ we denote by o( f/Λ) : o(S/Λ) → o(R/Λ) the linear map induced by pullback:

∗ o( f/Λ)([η]) =[f η]∈ExΛ(R, k) 38 2 Formal deformation theory for all [η]∈ExΛ(S, k). Since this definition is functorial we have a contravariant functor: ∗ o(−/Λ) : AΛ → (vector spaces/k)

When Λ = k we write o( f ) instead of o( f/k).Ifµ is such that o(µ) is injective one sometimes says that R is less obstructed than Λ. By applying Proposition 1.1.5 we ∗ obtain an exact sequence for each f : R → S in AΛ:

o( f/Λ) 0 → tS/R → tS/Λ → tR/Λ → o(S/R) → o(S/Λ) −→ o(R/Λ) (2.1)

In the case Λ = k we obtain the exact sequence:

o( f ) 0 → tS/R → tS → tR → o(S/R) → o(S) −→ o(R) (2.2) which relates the absolute and the relative obstruction spaces. The next result gives a description of o(R/Λ) and an interpretation of formal ∗ smoothness of a Λ-algebra (R, m) in AΛ.

Proposition 2.1.1. Assume that Λ is in ob(Aˆ). ∗ ˆ (i) Let (R, m) be in ob(AΛ) and let χ : R → R be the natural homomorphism of R into its m-adic completion R.ˆ Then the induced map:

o(χ/Λ) : o(Rˆ/Λ) → o(R/Λ)

is an isomorphism. ∗ (ii) For every (R, m) in ob(AΛ) let d = dimk(tR/Λ) and let

ˆ R = Λ[[X1,...,Xd ]]/J

with J ⊂ (X)2, be a presentation of the m-adic completion R.ˆ Then there is a natural isomorphism: ∨ o(R/Λ) =∼ J/(X)J

In particular, R is unobstructed over Λ if and only if it is a formally smooth Λ-algebra.

Proof. (i) Let η : 0 → k → S → Rˆ → 0 be a small extension; denote by m the maximal ideal of S. Claim: S is complete. ¯ ˆ Let { fn}⊂S be a Cauchy sequence; then the image sequence { fn} in R is Cauchy, hence it converges to a limit which we may assume to be zero, after possibly 2.1 Obstructions 39

¯ e(n) subtracting a constant sequence from { fn}.Wehave fn ∈ˆm , with limn[e(n)]= e(n) ¯ ∞. For every n we may find gn ∈ m lying above fn. The sequence {gn} in S is Cauchy and converges to zero, and { fn − gn} is a Cauchy sequence in k. Since k is complete as an S-module, because it is annihilated by the maximal ideal, { fn − gn} converges to a limit f ∈ k. This is also the limit of { fn} because

fn − f = ( fn − gn − f ) + gn

Therefore S is complete. If χ ∗(η/Λ) is trivial the section induces a homomorphism g : R → S which factors through Rˆ because S is complete. Hence η is trivial. This proves that o(χ/Λ) is injective. Given a small Λ-extension of R:

(S,ϕ): 0 → k → S → R → 0 ˆ ˆ ˆ ˆ ˆ the map ϕˆ : S → R is surjective and ker(ϕ)ˆ = k = k. Therefore [S, ϕˆ]∈ExΛ(R, k) and o(χ/Λ)([Sˆ, ϕˆ]) =[S,ϕ]: this means that o(χ/Λ) is also surjective. (ii) R is a formally smooth Λ-algebra if and only if Rˆ is a power series ring over Λ, i.e. if and only if J = (0). Therefore the last assertion follows from the fact that J/(X)J = (0) if and only if J = (0), by Nakayama’s lemma. ˆ In order to prove the first assertion we may assume that R is in ob(AΛ), since o(R/Λ) = o(Rˆ/Λ) by the first part of the proposition. Hence R = Λ[[X]]/J with J ⊂ (X)2. The extension of R:

Φ : 0 → J/(X)J → Λ[[X]]/(X)J → R → 0 induces by pushouts a homomorphism: ∨ α : J/(X)J → ExΛ(R, k) = o(R)

d −→ [ d∗Φ]

Letting M be the maximal ideal of Λ[[X]]/(X)J we have J/(X)J ⊂ M2.Ifd ∈ ∨ (J/(X)J) is such that [d∗Φ]=0 then we have: Φ : 0 → J/(X)J → Λ[[X]]/(X)J → R → 0 ↓ d ↓ h  d∗Φ : 0 → k → A → R → 0 with d∗Φ trivial. From Example 1.1.2(ii) it follows that the generator of k in A is \ 2 ( /( ) ) ⊂ 2 = contained in m A m A. Since h J X J m A we deduce that d 0. It follows that α is injective. Conversely, given a Λ-extension (A,ϕ)of R by k it is possible to find a lifting: Λ[[X]] ↓˜ϕ ϕ : A → R 40 2 Formal deformation theory because A is complete (see C.3(ii)). From the fact that ker(ϕ) = k it follows that ker(ϕ)˜ ⊃ (X)J and therefore we have a commutative diagram:

Φ : 0 → J/(X)J → Λ[[X]]/(X)J → R → 0 ↓ d ↓˜ϕ  (A,ϕ): 0 → k → A → R → 0 in which d is the map induced by ϕ˜.Itfollowsthat(A,ϕ) = d∗Φ; hence α is surjective. 

Corollary 2.1.2. For every R in ob(A∗) the following are true:

(i) dimk[o(R)] < ∞

(ii) dimk(tR) ≥ dim(R) ≥ dimk(tR) − dimk[o(R)] (where dim(R) means Krull dimension of R). In (ii) the first equality holds if and only if R is formally smooth; the second equality holds if and only if Rˆ = 2 k[[X1,...,Xd ]]/J, with J ⊂ (X) and J generated by a regular sequence. ˆ ˆ Proof. We may assume that R is in ob(A); hence R = R = k[[X1,...,Xd ]]/J, with J ⊂ (X)2 and o(R) =∼ (J/(X)J)∨. Then (i) and (ii) follow from the fact that dimk[J/(X)J] is the number of elements of a minimal set of generators of J. 

Remarks 2.1.3. The only formally smooth k-algebra in A is k itself. By 2.1.1(ii) this means that o(k) = (0) and that o(A) = (0) for every A = k in A. The following are some special cases. • If A = k⊕˜ V , a trivial extension of k by a vector space V of dimension d, then ∼ 2 2 3 ∨ A = k[X1,...,Xd ]/(X) , and o(A) =[(X) /(X) ] . • If A = k[X]/(X)k then o(A) =[(X)k/(X)k+1]∨. In particular, if A = k[t]/(tn), n ≥ 2, then o(A) =[(tn)/(tn+1)]∨ is one-dimensional; from the proof of 2.1.1(ii) it follows immediately that o(A) is generated by the class of the ex- tension: + + 0 → (tn)/(tn 1) → k[t]/(tn 1) → k[t]/(tn) → 0

We will need the following:

Lemma 2.1.4. (i) Let µ : Λ → R be a homomorphism in A∗. Given a small ex- tension η : B → AinA and a homomorphism ϕ : R → A, the condition ϕ∗(η) ∈ ker(o(µ)) is equivalent to the existence of a commutative diagram: ϕ A ←− R ↑ η ↑ µ (2.3) B ←− Λ ϕ˜ Moreover, ϕ∗(η) = 0 if and only if there exists ϕ : R → B such that the resulting diagram 2.1 Obstructions 41 ϕ A ←− R

↑ η ϕ ↑ µ (2.4)

B ←− Λ ϕ˜ is commutative. ˆ ∗ (ii) For every Λ in ob(A) and µ : Λ → Rinob(AΛ) there exists A in ob(AΛ) and a homomorphism p : R → A such that o(p/Λ) : o(A/Λ) → o(R/Λ) is surjective.

Proof. (i) is left to the reader. n+1 (ii) We will show something more precise, namely that if pn : R → R/m is the natural map, then

n+1 o(pn/Λ) : o((R/m )/Λ) → o(R/Λ) is surjective for all n  0. ˆ Since o(pn/Λ) factors through o(R/Λ), which is isomorphic to o(R/Λ),wemay ˆ assume that R is in AΛ. Let’s write:

R = Λ[[X]]/J

2 n+2 where J = (g1,...,gs) ⊂ (X) .Letn  0 be such that g j ∈/ (X) for all j = 1,...,s. Then we have: Λ[[ ]] R = X mn+1 J,(X)n+1 and therefore ∨ (J,(X)n+1) o((R/mn+1)/Λ) = ((X)J,(X)n+2)

The map o(pn/Λ) is the transpose of J (J,(X)n+1) ı : → n (X)J ((X)J,(X)n+2) induced by the inclusion J ⊂ (J,(X)n+1). From the hypothesis on n it follows that n+2 if γ ∈ J ∩ ((X)J,(X) ) then γ ∈ (X)J; this means that ın is injective, i.e. that o(pn/Λ) is surjective.  The following theorem gives a characterization of formally smooth homomor- phisms in A∗. Theorem 2.1.5. Let µ : Λ → R be a homomorphism in A∗. The following condi- tions are equivalent: (i) For every commutative diagram (2.3) with η a small extension in A∗ there exists ϕ : R → B such that diagram (2.4) is commutative. 42 2 Formal deformation theory

(ii) µ is formally smooth. (iii) dµ : tR → tΛ is surjective and o(µ) is injective. (iv) o(R/Λ) = (0).

Proof. (i) ⇒ (ii) is trivial.

(ii) ⇒ (iii) Let v ∈ tΛ be given as a k-algebra homomorphism Λ → k[ ]. The formal smoothness of µ implies the existence of a homomorphism w : R → k[ ] which makes the following diagram commutative: k ← R ↑ ↑ k[ ]←Λ and this means that dµ(w) = v. Therefore dµ is surjective. Consider a commutative diagram of k-algebra homomorphisms (2.3) with η a small extension in A. Then o(ϕ)([η]) ∈ ker(o(µ)). By the formal smoothness of µ there exists ϕ : R → B making (2.4) commutative: this implies that o(ϕ)([η]) = 0. Since, by Lemma 2.1.4, ϕ and η can be chosen so that o(ϕ)([η]) is an arbitrary element of ker(o(µ)), we deduce that o(µ) is injective. (iii) ⇔ (iv) follows from the exact sequence (2.2). (iii) ⇒ (i). Consider a diagram (2.3) with η a small extension in A∗. Then

o(ϕ)([η]) ∈ ker(o(µ))

By assumption o(ϕ)([η]) = 0, and therefore there exists ϕ¯ : R → B such that ηϕ¯ = ϕ.Itfollowsthatϕ˜ −¯ϕµ : Λ → ker(η) = k is a k-derivation. By assumption there exists a k-derivation v : R → k such that ϕ˜ −¯ϕµ = dµ(v) = vµ. Then ϕ :=ϕ ¯ + v : R → B is a k-homomorphism which obviously satisfies ηϕ = ϕ. Moreover, ϕ µ = (ϕ¯ + v)µ =¯ϕµ + vµ =˜ϕ and therefore ϕ makes (2.4) commutative.  In the special case Λ = k we obtain that a k-algebra R in ob(A∗) is unobstructed if and only if R is formally smooth, a result already proven in 2.1.1. More generally, the theorem says that µ is formally smooth if and only if dµ is surjective and R is less obstructed than Λ. The following corollary is immediate.

Corollary 2.1.6. (i) Let µ : Λ → R be a homomorphism in A∗ such that dµ is surjective and R is formally smooth. Then Λ and µ are formally smooth. (ii) A homomorphism µ : Λ → RinA∗ is formally etale if and only if dµ is an isomorphism and o(µ) is injective. This happens in particular if dµ is an isomorphism and R is formally smooth. (iii) If R ∈ ob(A∗) then the natural homomorphism R → Rˆ is formally etale. 2.1 Obstructions 43

In practice it is seldom possible to compute the obstruction map o(µ) explicitly for a given µ : Λ → R in A∗. But for the purpose of studying the formal smoothness of µ all that counts is to have information about ker[o(µ)]. This can be achieved somewhat indirectly by means of the following simple result, which turns out to be very effective in practice. Proposition 2.1.7. Let µ : Λ → R be a morphism in A∗. Assume that there exists a ∗ k-vector space υ(R/Λ) such that for every homomorphism ϕ : R → AinAΛ with Ainob(AΛ) there is a k-linear map

ϕυ : ExΛ(A, k) → υ(R/Λ) satisfying ker(ϕυ ) = ker[o(ϕ/Λ)] where ϕυ ExΛ(A, k) −→ υ(R/Λ)

o(ϕ/Λ)

ExΛ(R, k) = o(R/Λ) Then there is a natural k-linear inclusion o(R/Λ) ⊂ υ(R/Λ) Proof. Choosing ϕ such that o(ϕ/Λ) surjects onto o(R/Λ) we obtain an inclusion o(R/Λ) ⊂ υ(R/Λ) as asserted.  In practice this proposition will be applied as follows. Given ϕ,togiveanele- ment of ExΛ(A, k) is the same as to give a commutative diagram (2.3). Assume that to each such diagram one associates in a linear way an element of υ(R/Λ) which vanishes if and only if there is an extension ϕ : R → B making (2.4) commutative. Then the proposition applies. Taking Λ = k we get the following absolute version of the proposition, where for the last assertion we apply 2.1.2(ii). Corollary 2.1.8. Let R be in ob(A∗). Assume that there exists a k-vector space υ(R) such that for every morphism ϕ : R → AinA∗ with A in ob(A) there is a k-linear map ϕυ : o(A) → υ(R) satisfying ker(ϕυ ) = ker[o(ϕ)] Then there is a natural k-linear inclusion o(R) ⊂ υ(R) If υ(R) is finite dimensional then

dimk(tR) ≥ dim(R) ≥ dimk(tR) − dimk[υ(R)] A k-vector space υ(R) satisfying the conditions of 2.1.8 will sometimes be called an obstruction space for R. 44 2 Formal deformation theory 2.2 Functors of Artin rings

A functor of Artin rings is a covariant functor

F : AΛ −→ (sets) ˆ where Λ ∈ ob(A).LetA be in ob(AΛ). An element ξ ∈ F(A) will be called an infinitesimal deformation of ξ0 ∈ F(k) if ξ → ξ0 under the map F(A) → F(k);if A = k[ ] then ξ is called a first-order deformation of ξ0. ˆ Examples of functors of Artin rings are obtained by fixing an R in ob(AΛ) and letting: h /Λ(A) = ˆ (R, A) A (AΛ) R HomAΛ for every in ob Such a functor is clearly nothing but the restriction to AΛ of a representable functor ˆ ˆ on AΛ. A functor of Artin rings isomorphic to h R/Λ for some R in ob(AΛ) is called prorepresentable. In the case Λ = k we write h R instead of h R/k. Every representable functor h R/Λ, R in ob(AΛ), is a (trivial) example of prorep- resentable functor. Typically, a prorepresentable functor of Artin rings arises as follows. One con- siders a scheme M and the restriction

Φ : A −→ (sets)

Φ(A) = Hom(Spec(A), M) of the functor of points ◦ Hom(−, M) : (schemes) −→ (sets)

Φ is a functor of Artin rings; if ϕ : Spec(A) → M is an element of Φ(A) then ϕ is an infinitesimal deformation of the composition ϕ Spec(k) → Spec(A) −→ M where the first morphism corresponds to A → A/m A = k.Forafixedk-rational point m ∈ M, one may consider the subfunctor

F : A −→ (sets) of Φ defined as follows: morphisms Spec(A) → M F(A) = Hom(Spec(A), M) = (2.5) m whose image is {m} ˆ We have F = h R, where R = OM,m,soF is prorepresentable. We will introduce several functors of Artin rings which are defined by deformation problems. The hope is that they are of the form (2.5) for some k-pointed algebraic scheme (M, m) which has some geometrical meaning with respect to the deformation problem under con- sideration (a “moduli scheme”). As we will see, this sometimes happens, but most 2.2 Functors of Artin rings 45 of the time it doesn’t. We will thus have to content ourselves to prove that such func- tors have weaker properties. In this section and in the following one we will study prorepresentability and other properties of functors of Artin rings from an abstract point of view. Later on we will consider specific examples of such functors related to deformation problems. A prorepresentable functor F = h R/Λ satisfies the following conditions:

H0) F(k) consists of one element (the canonical quotient R → R/m R = k) Let A A (2.6) A be a diagram in AΛ and consider the natural map

α : F(A ×A A ) → F(A ) ×F(A) F(A ) induced by the commutative diagram:

F(A ×A A ) → F(A ) ↓↓ F(A ) → F(A)

Then

H) (left exactness) For every diagram (2.6) α is bijective (straightforward to check). H f ) F(k[ ]) has a structure of finite-dimensional k-vector space. In fact

F(k[ ]) = HomΛ−alg(R, k[ ]) = DerΛ(R, k) = tR/Λ is the relative tangent space of R over Λ (here the Λ-algebra structure of k[ ] is given by the composition Λ → k → k[ ], (see B.9(vi)).

A property weaker than H satisfied by a prorepresentable functor F is the fol- lowing:

H )αis bijective if A = k and A = k[ ]. It is interesting to observe that if F is prorepresentable the structure of k-vector space on F(k[ ]) can be reconstructed in a purely functorial way using only proper- ties H0 and H , and without using the prorepresentability explicitly. Indeed it is an easy exercise to check that the homomorphism:

+: k[ ]×k k[ ]−→ k[ ]

(a + b , a + b ) −→ a + (b + b ) 46 2 Formal deformation theory induces the sum operation on F(k[ ]) as the composition

F(+) F(k[ ]) × F(k[ ]) → F(k[ ]×k k[ ]) −→ F(k[ ]) where the first map is the inverse of α. Associativity is checked using H . The zero is the image of F(k) → F(k[ ]). The multiplication by a scalar c ∈ k is induced in F(k[ ]) by the morphism k[ ]−→ k[ ]

a + b −→ a + (cb) We can therefore state the following:

Lemma 2.2.1. If F is a functor of Artin rings having properties H0 and H then the set F(k[ ]) has a structure of k-vector space in a functorial way. This vector space is called the tangent space of the functor F, and is denoted by tF .IfF= h R/Λ then tF = tR/Λ. If f : F → G is a morphism of functors of Artin rings then the induced map tF → tG is called the differential of f and it is denoted by df. It is straightforward to check that if F and G satisfy H0 and H then df is k-linear. Every functor of Artin rings F can be extended to a functor ˆ ˆ F : AΛ −→ (sets) ˆ by letting, for every R in ob(AΛ): + Fˆ(R) = lim F(R/mn 1) ←− R and for every ϕ : R → S: Fˆ(ϕ) : Fˆ(R) → Fˆ(S) ( / n ) → ( / n ) ≥ be the map induced by the maps F R m R F S mS , n 1. An element uˆ ∈ Fˆ(R) is called a formal element of F. By definition uˆ can be { ∈ ( / n+1)} ≥ represented as a system of elements un F R m R n≥0 such that for every n 1 the map ( / n+1) → ( / n ) F R m R F R m R / n+1 → / n induced by the projection R m R R m R sends

un −→ un−1 (2.7) uˆ is also called a formal deformation of u0.If f : F → G is a morphism of functors of Artin rings then it can be extended in an obvious way to a morphism of functors fˆ : Fˆ → Gˆ . ˆ ˆ Lemma 2.2.2. Let R be in ob(AΛ). There is a 1–1 correspondence between F(R) and the set of morphisms of functors

h R/Λ −→ F (2.8) 2.2 Functors of Artin rings 47

Proof. To a formal element uˆ ∈ Fˆ(R) there is associated a morphism of functors ∈ ( / n+1) (2.8) in the following way. Each un F R m R defines a morphism of func- → tors h(R/mn+1)/Λ F. The compatibility conditions (2.7) imply that the following diagram commutes: → h(R/mn)/Λ h(R/mn+1)/Λ ↓

F for every n. Since for each A in ob(AΛ)

( ) → ( ) h(R/mn)/Λ A h(R/mn+1)/Λ A is a bijection for all n  0 we may define

h R/Λ(A) → F(A) as

lim [h( / n+1)/Λ(A) → F(A)] n→∞ R m ˆ Conversely, each morphism (2.8) defines a formal element uˆ ∈ F(R), where un ∈ ( / n+1) → / n+1 F R m R is the image of the canonical projection R R m R via the map

( / n+1) → ( / n+1) h R/Λ R m R F R m R  ˆ ˆ Definition 2.2.3. If R is in ob(AΛ) and uˆ ∈ F(R), we call (R, uˆ) a formal couple for F. The differential tR/Λ → tF of the morphism h R/Λ → F defined by uˆ is called the characteristic map of uˆ (or of the formal couple (R, uˆ)) and is denoted by du.ˆ If (R, uˆ) is such that the induced morphism (2.8) is an isomorphism, then F is prorep- resentable, and we also say that F is prorepresented by the formal couple (R, uˆ).In this case uˆ is called a universal formal element for F, and (R, uˆ) is a universal formal couple.

An ordinary couple (R, u) with R in AΛ defines a special case of formal couple, in which (Rn, un) = (Rn+1, un+1) for all n  0. A universal formal couple sel- dom exists; we will therefore need to introduce some weaker properties of a formal couple. They are based on the following definition.

Definition 2.2.4. Let f : F → G be a morphism of functors of Artin rings. f is called smooth if for every surjection µ : B → AinAΛ the natural map:

F(B) → F(A) ×G(A) G(B) (2.9) 48 2 Formal deformation theory induced by the diagram: F(B) → G(B) ↓↓ F(A) → G(A) is surjective. The functor F is called smooth if the morphism from F to the constant functor G(A) ={one element} all A ∈ ob(AΛ) is smooth; equivalently, if F(µ) : F(B) → F(A) is surjective for every surjection µ : B → AinAΛ.

Note that, since every surjection in AΛ factors as a finite sequence of small ex- tensions (i.e. surjections with one-dimensional kernel), for f (or F) to be smooth it is necessary and sufficient that the defining condition is satisfied for small extensions in AΛ. The next proposition states some properties of the notion of smoothness of a morphism of functors of Artin rings.* ˆ Proposition 2.2.5. (i) Let f : R → S be a homomorphism in AΛ. Then f is formally smooth if and only if the morphism of functors h f : hS/Λ → h R/Λ induced by f is smooth. (ii) If f : F → G is smooth then f is surjective, i.e. F(B) → G(B) is surjective for every B in ob(AΛ). In particular the differential d f : tF → tG is surjective. (iii) If f : F → G is smooth, then F is smooth if and only if G is smooth. (iv) If f : F → G is smooth then the induced morphism of functors fˆ : Fˆ → Gisˆ ˆ ˆ ˆ surjective, i.e. F(R) → G(R) is surjective for every R in ob(AΛ). (v) If F → G and G → H are smooth morphisms of functors then the composition F → H is smooth. (vi) If F → G and H → G are morphisms of functors and F → G is smooth, then F ×G H → H is smooth. (vii) If F and G are smooth functors then F × G is smooth.

Proof. (i) Let B → A be a surjection in AΛ and let A ← S ↑↑f B ← R be a commutative diagram of homomorphisms of Λ-algebras. The formal smooth- ness of f is equivalent to the existence, for each such diagram, of a morphism S → B such that the resulting diagram A ← S ↑ ↑f B ← R ∗ Here all functors are assumed to satisfy Ho. 2.2 Functors of Artin rings 49 is commutative. This is another way to express the condition that the map ( ) −→ ( ) × ( ) hS/Λ B hS/Λ A h R/Λ (A) h R/Λ B is surjective, i.e. that h f is smooth. (ii) follows immediately from the definition. (iii) Let µ : B → A be a surjection in AΛ, and consider the diagram:

F(B) → F(A) ×G(A) G(B) → G(B) ↓↓G(µ) F(A) −→ G(A) f (A) Suppose G is smooth and let ξ ∈ F(A). By the smoothness of G there exists η ∈ G(B) such that G(µ)(η) = f (A)(ξ). By the smoothness of f there is ζ ∈ F(B) which is mapped to (ξ, η) ∈ F(A) ×G(A) G(B).ItfollowsthatF(µ)(ζ ) = ξ and F is smooth. The converse is proved similarly using the surjectivity of f (A). ˆ (iv) Let vˆ ={vn}∈G(R). Since f is smooth the map ( / 2 ) → ( ) × ( / 2 ) = ( / 2 ) F R m R F k G(k) G R m R G R m R w ∈ ( / 2 ) ( / 2 )(w ) = v is surjective. Therefore there exists 1 F R m R such that f R m R 1 1. ≥ w ∈ ( / i+1) = ,..., Let’s assume that for some n 1 there exist i F R m R , i 1 n such that: ( / i+1)(w ) = v f R m R i i w −→ w ( / i+1) −→ ( / i ) and i i−1 under F R m R F R m R . The surjectivity of the map: ( / n+1) −→ ( / n ) × ( / n+1) F R m R F R m R G(R/mn) G R m R w ∈ ( / n+1) (w ,v ) implies that there exists n+1 F R m R whose image is n n+1 . By induc- tion we conclude that there exists wˆ ∈ Fˆ(R) whose image under fˆ is vˆ. The proofs of (v), (vi) and (vii) are straightforward.  We now introduce the notions of “versality” and “semiuniversality”, which are slightly weaker than universality.

Definition 2.2.6. Let F be a functor of Artin rings. A formal element uˆ ∈ Fˆ(R), ˆ for some R in ob(AΛ), is called versal if the morphism h R/Λ → F defined by uˆ is smooth; uˆ is called semiuniversal if it is versal and moreover, the differential tR/Λ → tF is bijective. We will correspondingly speak of a versal formal couple (respectively a semiuniver- sal formal couple).

It is clear from the definitions that:

uˆ universal ⇒ˆu semiuniversal ⇒ˆu versal 50 2 Formal deformation theory but none of the inverse implications is true (see Examples 2.6.8). We can also de- scribe these properties as follows. Assume given µ : (B,ξB) → (A,ξA), a surjection of couples for F (i.e. µ is surjective), and a homomorphism ϕ : R → A such that ϕ(uˆ) = ξA. Then uˆ is versal if for all such data there is a lifting ψ : R → B of ϕ (i.e. µψ = ϕ) such that ψ(uˆ) = ξB: B ξB ψ   R ↓ µ uˆ ↓ ϕ A ξA uˆ is universal if, moreover, the lifting ψ is unique. uˆ is semiuniversal if it is versal and, moreover, the lifting ψ is unique when µ : k[ ]→k. Let (R, uˆ) and (S, v)ˆ be two formal couples for F.Amorphism of formal couples

f : (R, uˆ) → (S, v)ˆ is a morphism f : R → S in Aˆ such that Fˆ( f )(uˆ) =ˆv. We will call f an isomor- phism of formal couples if in addition f : R → S is an isomorphism. It is obvious that with this definition the formal couples for F and their morphisms form a category containing the category IF of couples for F as a full subcategory.

Proposition 2.2.7. Let F be a functor of Artin rings. Then: (i) If (R, uˆ) and (S, v)ˆ are universal formal couples for F there exists a unique iso- morphism of formal couples (R, uˆ) =∼ (S, v)ˆ . (ii) If (R, uˆ) and (S, v)ˆ are semiuniversal formal couples for F there exists an iso- morphism of formal couples (R, uˆ) =∼ (S, v)ˆ , which is not necessarily unique ∼ but the induced isomorphism tS = tR is uniquely determined. (iii) If (R, uˆ) is a semiuniversal formal couple for F and (S, v)ˆ is versal, then there is an isomorphism, not necessarily unique:

ϕ : R[[X1,...,Xr ]] → S ˆ for some r ≥ 0, such that F(ϕj)(uˆ) =ˆv, where j : R ⊂ R[[X1,...,Xr ]] is the inclusion.

Proof. (i) By the universality of (R, uˆ) for every n ≥ 1 there exists a unique fn ∈ ( / n+1) → v ∈ ( / n+1) h R/Λ S mS such that fn n F S mS under the isomorphism associated to uˆ. In this way we obtain

f = lim{ f }:(R, uˆ) → (S, v)ˆ ← n which is uniquely determined. Analogously, we can construct a uniquely determined g : (S, v)ˆ → (R, uˆ). By universality the compositions 2.2 Functors of Artin rings 51

gf : (R, uˆ) → (R, uˆ) and fg : (S, v)ˆ → (S, v)ˆ are the identity. (ii) Using versality, and proceeding as above we can construct morphisms of formal couples f : (R, uˆ) → (S, v)ˆ and g : (S, v)ˆ → (R, uˆ)

We obtain a commutative diagram:

tR/Λ ↑↓ tS/Λ −→ tF where the vertical arrows are the differentials of f and g, and the other arrows are the characteristic maps duˆ and dvˆ. From this diagram we deduce that df = (duˆ)−1dvˆ and dg = (dv)ˆ −1duˆ are uniquely determined. Since

d(gf) = (dg)(df) = identity of tS/Λ and d( fg) = (df)(dg) = identity of tR/Λ it follows that f and g are bijections inverse of each other. (iii) By the versality of (R, uˆ) we can find a morphism of formal couples

f : (R, uˆ) → (S, v)ˆ

We obtain a commutative diagram:

tR/Λ ↑ tS/Λ −→ tF where duˆ : tR/Λ → tF is bijective because (R, uˆ) is semiuniversal, and dvˆ : tS/Λ → tF is surjective because (S, v)ˆ is versal. Hence df : tS/Λ → tR/Λ is surjective. This means that f induces an inclusion ∨ ⊂ ∨ tR/Λ tS/Λ

Let x1,...,xr ∈ S be elements which induce a basis of ∨ = /[ 2 + ( )] tS/R mS mS f m R 52 2 Formal deformation theory and define: ϕ : R[[X1,...,Xr ]] → S by ϕ(Xi ) = xi , i = 1,...,r.Let j : R ⊂ R[[X1,...,Xr ]] be the inclusion. Letting ˆ ˆ wˆ = F( j)(uˆ) ∈ F(R[[X1,...,Xr ]]) we have a commutative diagram of formal couples: ϕ (R[[X1,...,Xr ]], w)ˆ −→ (S, v)ˆ

↑ j  f

(R, uˆ) such that Fˆ(ϕj)(uˆ) =ˆv. Since ϕ : [[ ,..., ]]/ 2 → / 2 1 R X1 Xr M S mS is an isomorphism (here M ⊂ R[[X1,...,Xr ]] denotes the maximal ideal) ϕ is surjective. Let ψ : / 2 → [[ ,..., ]]/ 2 1 S mS R X1 Xr M be the inverse of ϕ1.Wehave

F(ψ1)(v1) = w1 hence, by the versality of (S, v)ˆ , it is possible to find a lifting of ψ1:

ψ : S → R[[X1,...,Xr ]] such that Fˆ(ψ)(v)ˆ =ˆw. Since by construction

d(ψϕ) = dψ dϕ = identity of tR[[X]] it follows that ψϕ is an automorphism of R[[X1,...,Xr ]]. In particular, we deduce that ϕ is injective.  The following is useful:

Proposition 2.2.8. If a functor of Artin rings F : A → (sets) satisfying H0 has a semiuniversal formal element and tF = (0) then F = hk, the constant functor.

Proof. The assumptions imply that there is a smooth morphism hk → F. Since a smooth morphism of functors satisfying H0 is surjective, the conclusion follows.  ∗∗∗∗∗∗

For the rest of this section we will only consider functors of Artin rings satisfying conditions H0 and H . 2.2 Functors of Artin rings 53

Definition 2.2.9. Let F be a functor of Artin rings. Suppose that υ(F) is a k-vector space such that for every A in ob(AΛ) and for every ξ ∈ F(A) there is a k-linear map ξυ : ExΛ(A, k) → υ(F) with the following property: ˜ ker(ξυ ) consists of the isomorphism classes of extensions (A,ϕ)such that

ξ ∈ Im[F(A˜) → F(A)]

Then υ(F) is called an obstruction space for the functor F. If F has (0) as an ob- struction space then it is called unobstructed.

If F is prorepresented by the formal couple (R, uˆ), then o(R/Λ) is an obstruction space for F, as it follows from the functorial characterization of o(R/Λ) given in § 2.1. Conversely, if υ(F) is an obstruction space for F, and F is prorepresented by the formal couple (R, uˆ), then it follows immediately from the definition that υ(F) is a relative obstruction space for R/Λ. One can show that under relatively mild conditions a functor F has an obstruction space. We will not discuss this matter here, and we refer the interested reader to [54]. The following are some basic properties of obstruction spaces.

Proposition 2.2.10. (i) F has (0) as an obstruction space if and only if it is smooth. (ii) Let f : F → G be a smooth morphism of functors of Artin rings. If υ(G) is an obstruction space for G then it is also an obstruction space for F.

Proof. (i)isobvious.

(ii) Let A be in ob(AΛ) and ξ ∈ F(A), and let

ξυ := f (A)(ξ)υ : ExΛ(A, k) → υ(G) the map defined by f (A)(ξ) ∈ G(A).IfA˜ → A = A˜/(t) defines an element of ˜ ker(ξυ ) then, since υ(G) is an obstruction space for G, there is η ∈ G(A) such that η → f (A)(ξ) under the map G(A˜) → G(A). From the smoothness of f it follows that the map ˜ ˜ F(A) → F(A) ×G(A) G(A) is surjective, hence there is ξ˜ ∈ F(A˜) which maps to the formal couple (ξ, η).It follows that ξ˜ → ξ under F(A˜) → F(A). Conversely, if A˜ is such that there exists ξ˜ ∈ F(A˜) such that

F(A˜) → F(A)

ξ˜ → ξ 54 2 Formal deformation theory then from the diagram: F(A˜) → F(A)

↓ f (A˜) ↓ f (A)

G(A˜) → G(A) we see that f (A˜)(ξ)˜ → f (A)(ξ) hence f (A)(ξ) ∈ Im[G(A˜) → G(A)]. Therefore ˜ the extension A → A defines an element of ker[ f (A)(ξ)υ ]. This proves that ξυ satisfies the conditions of Definition 2.2.9.  Corollary 2.2.11. Let F : A → (sets) be a functor of Artin rings, and suppose that (R, uˆ) is a versal formal couple for F. If F has a finite-dimensional obstruction space υ(F) then

dimk(tR) ≥ dim(R) ≥ dimk(tR) − dimk[υ(F)] Proof. From the definition of versal formal couple and from Proposition 2.2.10 we deduce that υ(F) is an obstruction space for h R, hence for R. The conclusion follows from 2.1.6.  Corollary 2.2.11 is very important in deformation theory since it gives basic geo- metrical information about a deformation problem. If F is induced by a k-pointed scheme (M, m), i.e. it is of the form (2.5), then 2.2.11 gives information about the dimension of M at m. We refer the reader to Proposition 2.3.6 for a result about obstruction spaces which is often used in applications.

NOTES 1. For an interesting discussion on the geometric motivations behind the consideration of functors of Artin rings we refer the reader to [72], part C.1. 2. Corollary 2.2.11 is due to Mumford (unpublished, see [189], p. 126). A proof is outlined in [157], p. 153. The proof given here is based on such outline, and has appeared already in [163], Corollary (B2.4). ˆ 3. Let R, P be in A, ϕ : P → R a homomorphism and F : h R → h P be the correspond- ing morphism between the prorepresentable functors

h R, h P : A → (sets)

Then F(A) : h R(A) → h P (A) is injective for all A in ob(A) if and only if ϕ is surjective, i.e. if and only if R = P/J for an ideal J ⊂ P.(Proof: left to the reader.)

2.3 The theorem of Schlessinger

In this section we will prove a well-known theorem of Schlessinger which gives necessary and sufficient conditions, easy to verify in practice, for the existence of a (semi)universal element for a functor of Artin rings. Before stating the theorem we want to make some introductory remarks, which can be useful in what follows. Let’s fix Λ ∈ ob(Aˆ) throughout this section. We start with a characterization of prorepresentable functors: 2.3 The theorem of Schlessinger 55

Proposition 2.3.1. Let F : AΛ → (sets) be a functor of Artin rings satisfying con- dition H0. Then F is prorepresentable if and only if it is left exact and has finite- dimensional tangent space, i.e. it has properties H and H f . Proof. The “only if” implication is obvious (see also § 2.2). So let’s assume that F satisfies H and H f . Then, by Proposition E.9, the category IF of couples for F is cofiltered and F = lim h X → (X,ξ)

Let (Ai ,ξi ) ∈ ob(IF ) and consider all the subrings of Ai images of morphisms ¯ ¯ (A j ,ξj ) → (Ai ,ξi ). By the descending chain condition there is (Ai , ξi ) such that ¯ Ai = Im[A j → Ai ] ¯ ¯ By construction the couples (Ai , ξi ) form a full subcategory of IF in which all maps are surjective. Moreover, the corresponding category of representable functors h A¯ /Λ ◦ i is clearly cofinal in (IF ) , and therefore

F = lim h ¯ → Ai /Λ

Therefore replacing IF by this subcategory we can assume that all homomorphisms are surjective, and therefore we have: = F h Ai /Λ

( [ ]) = ( [ ])  Moreover, since H f holds, F k h Ai k when i 0. We can discard all those Ai for which this is not true and we get again a full subcategory. Therefore

F = lim h /Λ → Ai

( [ ]) = ( [ ]) and F k h Ai /Λ k for all i. Then for each i we can find a surjection

Λ[[X1,...,Xr ]] → Ai with r = dim(F(k[ ])), and these surjections are compatible, i.e. the diagrams

Λ[[X1,...,Xr ]] → Ai ↓ A j

= Λ[[ ,..., ]]/( )2 = / 2 are commutative. Define B1 X1 Xr X Ai mi for all i. Then fix ν ≥ = / ν+1 Λ[[ ,..., ]]/( )ν+1 2, and set Ai,ν Ai mi .AllAi,ν’s are quotients of X1 Xr X and form a projective system:

···→ A j,ν → Ai,ν →··· 56 2 Formal deformation theory

Let Bν = lim A ,ν = A ,ν i  0 ← i i Then by construction ···Bν+1 → Bν →··· form a projective system and F = lim h /Λ → Bν Then ˆ B := lim Bν ← ˆ is in ob(AΛ) and prorepresents F.  Unfortunately this characterization of prorepresentable functors is not very useful in practice because given homomorphisms A A

A in AΛ we have

Spec(A ×A A ) = Spec(A ) Spec(A ) Spec(A) and this is not easy to visualize. That’s why left exactness of a functor of Artin rings F is hard to check. On the other hand, if at least one of the above homomorphisms is surjective then Spec(A ×A A ) is easier to describe. The theorem of Schlessinger re- duces the prorepresentability to the verification of the condition of left exactness only in cases when at least one of the above maps is surjective. An analogous condition is given for the existence of a semiuniversal element. The result is the following.

Theorem 2.3.2 (Schlessinger [155]). Let F : AΛ → (sets) be a functor of Artin rings satisfying condition H0. Let A → A and A → A be homomorphisms in AΛ and let

α : F(A ×A A ) −→ F(A ) ×F(A) F(A ) (2.10) be the natural map. Then: (i) F has a semiuniversal formal element if and only if it satisfies the following con- ditions: H¯ ) if A → A is a small extension then the map (2.10) is surjective.

H ) If A = k and A = k[ ] then the map (2.10) is bijective. H f ) dimk(tF )<∞ (ii) F has a universal element if and only if it also satisfies the following additional condition: H) the natural map

F(A ×A A ) −→ F(A ) ×F(A) F(A )

is bijective for every small extension A → AinAΛ. 2.3 The theorem of Schlessinger 57

Remark 2.3.3. It will be helpful to explain the meaning of the conditions of the theorem before giving the proof. Consider a small extension in AΛ:

µ 0 → (t) → A −→ A → 0 ∼ Assume that F = h R/Λ is prorepresentable. Then two Λ-homomorphisms f, g :

R → A have the same image in HomΛ−alg(R, A) if and only if there exists a Λ- derivation d : R → k (which is uniquely determined) such that

g(r) = f (r) + d(r)t equivalently if and only if g and f differ by an element of

DerΛ(R, k) = tR/Λ

Therefore the fibres of

F(µ) : HomΛ−alg(R, A ) → HomΛ−alg(R, A) that are nonempty are principal homogeneous spaces under the above action of tR/Λ = tF . Assume now that F is just a functor of Artin rings having properties H0 and H , so that it has tangent space tF . We can define an action of tF on F(A ) by means of the composition

− α 1 F(b) τ : tF × F(A ) −→ F(k[ ]×k A ) −→ F(A )

− where α 1 exists by property H and F(b) is induced by the morphism

b : k[ ]×k A −→ A

(x + y , a ) −→ a + yt The action τ maps the fibres of F(µ) into themselves. Indeed the isomorphism

γ : k[ ]×k A −→ A ×A A

(x + y , a ) −→ (a + yt, a ) induces a map

− α 1 F(γ ) β : tF × F(A ) −→ F(k[ ]×k A ) −→ F(A ×A A ) −→ F(A ) ×F(A) F(A ) which coincides with

tF × F(A ) −→ F(A ) ×F(A) F(A )

(v, ξ) −→ (τ(v, ξ), ξ) 58 2 Formal deformation theory

In case F is prorepresentable we have just given another description of the action of tF on the fibres of F(µ) introduced before. In general the map β is neither injective (i.e. in general the action τ is not free on the fibres of F(µ)) nor surjective (i.e. τ is not transitive on the fibres of F(µ)). This depends on the properties of the map

α : F(A ×A A ) −→ F(A ) ×F(A) F(A )

If F is left exact then α is bijective, hence β is bijective, and the action of tF on the fibres of F(µ) is free and transitive, as expected, since F is prorepresentable by Proposition 2.3.1. Conversely, what this analysis shows is that for τ to be free and transitive on the ¯ fibres of F(µ) we only need α bijective, i.e. the condition H, weaker than H. H only guarantees the transitivity of such action: the failure from prorepresentability is therefore related to the existence of fixed points of this action. In applications this is usually due to the existence of automorphisms of geometric objects associated to an element ξ ∈ F(A) which don’t lift to automorphisms of objects associated to an element ξ ∈ F(µ)−1(ξ) (see § 2.6). Since Theorem 2.3.2 guarantees that F has a semiuniversal element if it satisfies ¯ H0, H and H, we can summarize this discussion in the following statement.

Proposition 2.3.4. Let F be a functor of Artin rings satisfying conditions H0 and H and let

0 → (t) → A → A → 0 be a small extension in A. Then there is an action τ of tF on F(A ) such that:

(a) if F has a semiuniversal formal element then tF acts transitively on the non- empty fibres of F(A ) → F(A); (b) if F is prorepresentable then tF acts freely and transitively on the non-empty fibres of F(A ) → F(A).

The following definition introduces a convenient terminology.

Definition 2.3.5. Let f : F → G be a morphism of functors of Artin rings having a semiuniversal formal element and obstruction spaces υ(F) and υ(G) respectively. An obstruction map for f is a linear map

o( f ) : υ(F) → υ(G) such that for each A in ob(A) and for each ξ ∈ F(A) the diagram:

Exk(A, k) ξυ ηυ (2.11) o( f ) υ(F) −→ υ(G) is commutative, where η denotes f (A)(ξ). If f has an injective obstruction map we say that F is less obstructed than G. 2.3 The theorem of Schlessinger 59

With this terminology we can state a result often used in concrete situations:

Proposition 2.3.6. Let f : F → G be a morphism of functors of Artin rings having a semiuniversal formal element and obstruction spaces υ(F) and υ(G) respectively. Consider the following conditions: (a) d f is surjective (b) F is less obstructed than G. If they are both satisfied then f is smooth. If condition (b) is satisfied (but not neces- sarily (a)) and G is smooth then F is smooth.

Proof. Let ϕ : A˜ → A be a small extension and assume (a) and (b) hold. Consider the map ˜ ˜ F(A) → F(A) ×G(A) G(A) (2.12) ˜ and let (ξ, η)˜ ∈ F(A) ×G(A) G(A). Since η˜ → η = f (A)(ξ) we have ηυ (ϕ) = 0. By (b) and the commutativity of (2.11) we also have ξυ (ϕ) = 0 and therefore there exists ξ˜ ∈ F(A˜) such that ξ˜ → ξ.Letη˜ = f (A˜)(ξ)˜ ∈ G(A˜).Wehave

η˜ η˜

η

¯ Since by Theorem 2.3.2 the functor G satisfies condition H, there is w ∈ tG such that τ(w,η˜ ) =˜η. From the surjectivity of df it follows that there is v ∈ tF such that v → w.Nowτ(v,ξ)˜ ∈ F(A˜) satisfies

τ(v,ξ)˜ → (ξ, η)˜ because the action τ is functorial (as it easily follows from its definition). This shows that (2.12) is surjective and f is smooth. Now assume that only (b) holds and that G is smooth. We must show that

F(A˜) → F(A) is surjective or equivalently that ξυ (ϕ) = 0. Let ξ ∈ F(A) and let η := f (A)(ξ) ∈ G(A). Since G is smooth we have ηυ (ϕ) = 0. By (b) and the commutativity of (2.11) we also have ξυ (ϕ) = 0. 

Corollary 2.3.7. Let f : F → G be a morphism of functors of Artin rings having a semiuniversal formal element. Assume that F is smooth, G has an obstruction space υ(G) and d f is surjective. Then f and G are smooth.

Proof. Since F is smooth it has (0) as an obstruction space and the obvious map (0) → υ(G) is an obstruction map. Therefore f is smooth by Proposition 2.3.6 and we conclude by Proposition 2.2.5(iii).  60 2 Formal deformation theory

Remark 2.3.8. When F and G are prorepresentable the first part of Proposition 2.3.6 is a reformulation of Theorem 2.1.5(iii) and Corollary 2.3.7 is a reformulation of Corollary 2.1.6(i). If in Corollary 2.3.7 we replace the condition “df surjective” by “df bijective” we cannot conclude that f is an isomorphism unless both functors are prorepresentable: if they are then f is an isomorphism by Corollary 2.1.6(ii) because a formally etale homomorphism of k-algebras in Aˆ is an isomorphism.

Proof of Theorem 2.3.2 (i) Let’s assume that F has a semiuniversal formal element (R, uˆ). Consider a homomorphism f : A → A and a small extension π : A → A both in AΛ, and let

(ξ ,ξ ) ∈ F(A ) ×F(A) F(A ) with F( f )(ξ ) = F(π)(ξ ) =: ξ ∈ F(A) By the versality of (R, uˆ) the following maps are surjective:

HomΛ−alg(R, A ) −→ F(A )

HomΛ−alg(R, A ) −→ HomΛ−alg(R, A) ×F(A) F(A ) (2.13) therefore there are

g ∈ HomΛ−alg(R, A ), g ∈ HomΛ−alg(R, A ) such that Fˆ(g )(uˆ) = ξ and g → ( fg ,ξ ) under the map (2.13). This last condition gives:

πg = fg , Fˆ(g )(uˆ) = ξ ˆ and consequently, F(πg )(uˆ) = ξ. Using the morphism g × g : R → A ×A A we obtain an element ˆ ζ := F(g × g )(uˆ) ∈ F(A ×A A ) which, by construction, is mapped to (ξ ,ξ ) by (2.13). This proves that (R, uˆ) sat- isfies condition H¯ . If A = k[ ] and A = k the map (2.13) reduces to the bijection

HomΛ−alg(R, k[ ]) −→ tF (2.14)

In this case if ζ1,ζ2 ∈ F(A ×A k[ ]) are such that α(ζ1) = α(ζ2) = (ξ ,ξ ) choose g ∈ HomΛ−alg(R, A ) as before. By the versality the map

HomΛ−alg(R, A ×k k[ ]) −→ F(A ×k k[ ]) ×F(k[ ]) HomΛ−alg(R, k[ ]) 

F(A ×k k[ ]) 2.3 The theorem of Schlessinger 61

induced by the projection A ×k k[ ]→k[ ] is surjective. Hence we obtain two morphisms:

g × gi : R −→ A ×k k[ ] such that ˆ F(g × gi )(uˆ) = ζi ˆ i = 1, 2. But then F(gi )(uˆ) = ξ , i = 1, 2, hence, by the bijectivity of (2.14), g1 = g2, i.e. ζ1 = ζ2. This proves that F satisfies condition H . Condition H f is satisfied because the differential tR/Λ → tF is linear and is a bijection by definition of a semiuniversal formal couple. ¯ Conversely, let’s assume that F satisfies conditions H, H and H f . We will find a semiuniversal formal couple (R, uˆ) by constructing a projective system {Rn; pn+1 : Rn+1 → Rn}n≥0 of Λ-algebras in AΛ and a sequence {un ∈ F(Rn)}n≥0 such that F(pn)(un) = un−1, n ≥ 1. We take R0 = k and u0 ∈ F(k) its unique element. Let r = dimk(tF ), {e1,...,er } a basis of tF and letting Λ[[X]] = Λ[[X1,...,Xr ]] we set

2 R1 = Λ[[X]]/((X) + mΛΛ[[X]])

Since we have R1 = k[ ]×k ···×k k[ ] (r times) from H we deduce that F(R1) = tF ×···×tF . Let’s take u1 = (e1,...,er ) ∈ F(R1). Note that the map induced by u1:

r ∼ 2 ∨ k = ((X)/(X) ) = HomΛ−alg(R1, k[ ]) −→ tF is the isomorphism (λ1,...,λr ) −→ λ j e j j Let’s proceed by induction on n: assume that couples

(R0, u0), (R1, u1), . . . , (Rn−1, un−1) such that

Rh = Λ[[X]]/Jh, uh ∈ F(Rh), uh → uh−1 h = 1,...,n − 1 have been already constructed. In order to construct (Rn, un) we consider the family I of all ideals J ⊂ Λ[[X]] having the following properties:

(a) Jn−1 ⊃ J ⊃ (X)Jn−1; (b) there exists u ∈ F(Λ[[X]]/J) such that the map

F(Λ[[X]]/J) −→ F(Rn−1)

sends u → un−1. 62 2 Formal deformation theory

I =∅because Jn−1 ∈ I. Moreover, I has a minimal element Jn. This will be proved if we show that I is closed with respect to finite intersections. Let I, J ∈ I and K = I ∩ J. It is obvious that K satisfies condition (a). We may assume that I + J = Jn−1 (making J larger without changing K if necessary). This implies that the natural homomorphism Λ[[ ]]/ −→ Λ[[ ]]/ × Λ[[ ]]/ X K X I Rn−1 X J is an isomorphism. By H¯ the map α : (Λ[[ ]]/ ) −→ (Λ[[ ]]/ ) × (Λ[[ ]]/ ) F X K F X I F(Rn−1) F X J is surjective, therefore there exists u ∈ F(Λ[[X]]/K ) whose image in F(Rn−1) is un−1; this means that K satisfies condition (b) as well, hence it is in I. We take Rn = Λ[[X]]/Jn and un ∈ F(Rn) an element which is mapped to un−1. By induction we have constructed a formal couple (R, uˆ). We now show that it is a semiuniversal formal couple for F. ∼ As already remarked, uˆ induces an isomorphism of tangent spaces tR/Λ = tF . Therefore we only have to prove versality, namely that the map

uˆπ : HomΛ−alg(R, A ) −→ HomΛ−alg(R, A) ×F(A) F(A ) is surjective for every small extension π : A → A. ˆ Let ( f,ξ ) ∈ HomΛ−alg(R, A)×F(A) F(A ), i.e. F( f )(uˆ) = F(π)(ξ ).Wemust

find f ∈ HomΛ−alg(R, A ) such that uˆπ ( f ) = ( f,ξ ), i.e. such that

(1)πf = f ; (2) Fˆ( f )(uˆ) = ξ

Let’s consider the commutative diagram:

HomΛ−alg(R, k[ ]) × HomΛ−alg(R, A ) → tF × F(A ) ↓ β1 ↓ β2

HomΛ−alg(R, A ) ×Hom(R,A) HomΛ−alg(R, A ) → F(A ) ×F(A) F(A ) (2.15) where the horizontal arrows are induced by uˆ.Themapβ1 is a bijection because the action of HomΛ−alg(R, k[ ]) on the fibres of

HomΛ−alg(R, A ) → HomΛ−alg(R, A) ¯ is free and transitive (see Remark 2.3.3). From H it follows that β2 is surjective

(Remark 2.3.3 again). Therefore if f ∈ HomΛ−alg(R, A ) satisfies (1) then, letting η := F( f )(uˆ) ∈ F(A ) there exist v ∈ HomΛ−alg(R, k[ ]) = tF , f ∈ HomΛ−alg(R, A ) such that in diagram (2.15) we have: (v, f ) −→ (v, η ) ↓↓ ( f , f ) −→ (ξ ,η ) 2.3 The theorem of Schlessinger 63

This means that uˆπ ( f ) = ( f,ξ ). It follows that it suffices to find f satisfying condition (1). Let n > 0 be such that f factors as

fn−1 R → Rn−1 −→ A

Then f exists if and only if there exists ϕ which makes the following diagram com- mutative: ϕ Rn −→ A ↓ pn ↓ π Rn−1 −→ A equivalently, if and only if the extension

∗( ,π): → (π) → × −→π → fn A 0 ker Rn A A Rn 0 ↓ ↓ fn fn π 0 → ker(π) → A −→ A → 0 is trivial. Suppose not. This means that π induces an isomorphism of tangent spaces (see 1.1.2(ii)), i.e. that there exists an ideal I ⊂ Λ[[X]] such that

Rn ×A A = Λ[[X]]/I

By construction Jn−1 ⊃ I ⊃ (X)Jn−1 Moreover, since by H¯ the map

F(Rn ×A A ) −→ F(Rn) ×F(A) F(A )

is surjective, there exists u ∈ F(Rn ×A A ) inducing un ∈ F(Rn), hence inducing un−1 ∈ F(Rn−1).ItfollowsthatI satisfies conditions (a) and (b) and, by the mini- mality of Jn in I,itfollowsthatJn ⊂ I . But this is a contradiction because from the fact that π is a surjection with nontrivial kernel it follows that I is properly contained ∗( ,π) in Jn. This proves that fn A is trivial and concludes the proof of the fact that (R, uˆ) is semiuniversal and of part (i) of the theorem.

(ii) If F is prorepresentable then it trivially satisfies conditions H, H and H f , as already remarked. ¯ Conversely, suppose that F satisfies conditions H, H, H and H f .Wehavejust proved that F has a semiuniversal formal couple (R, uˆ). We will prove that this is a universal formal couple by showing that for every A in AΛ the map

uˆ(A) : HomΛ−alg(R, A) −→ F(A) induced by uˆ is bijective. 64 2 Formal deformation theory

This is clearly true if A = k. We will proceed by induction on dimk(A).Let

π : A → A be a small extension in AΛ. By the inductive hypothesis

HomΛ−alg(R, A) −→ F(A) is bijective and, by the versality, the map ∼ uˆπ : HomΛ−alg(R, A ) → HomΛ−alg(R, A) ×F(A) F(A ) = F(A ) is surjective. The map β2 in diagram (2.15) is bijective by condition H, and this implies that uˆπ is bijective. 

NOTES 1. Theorem 2.3.2 has been published in [155]. It had also appeared in [154]. See also [119]. 2. From Theorem 2.3.2 it follows that if F has a semiuniversal element then it has a tangent space which is of finite dimension, because F satisfies H0, H and H f . This property was not explicitly stated in the definition.

2.4 The local moduli functors

2.4.1 Generalities

If X is an algebraic scheme then for every A in ob(A) we let

DefX (A) ={deformations of X over A}/isomorphism

By the functoriality properties already observed in § 1.2 this defines a functor of Artin rings DefX : A → (sets) which is called the local moduli functor of X.IfX = Spec(B0) is affine, we will often write DefB0 instead of DefX . We can define the subfunctor : A → DefX (sets) by ( ) ={ }/ DefX A locally trivial deformations of X over A isomorphism called the locally trivial moduli functor of X.

Theorem 2.4.1. (i) For any algebraic scheme X the functors DefX and Def satisfy ¯ X conditions H0, H, H of Schlessinger’s theorem. Therefore, if DefX (k[ ]) (resp. ( [ ]) DefX k ) is finite dimensional, then DefX (resp. DefX ) has a semiuniversal element. 2.4 The local moduli functors 65

(ii) There is a canonical identification of k-vector spaces ( [ ]) = 1( , ) DefX k H X TX (2.16) In particular, if X is nonsingular then ( [ ]) = ( [ ]) = 1( , ) DefX k DefX k H X TX (iii) If X is an arbitrary algebraic scheme then we have a natural identification

DefX (k[ ]) = Exk(X, OX )

and an exact sequence:

ρ → 1( , ) −→τ ( [ ]) −→ 0( , 1) −→ 2( , ) 0 H X TX DefX k H X TX H X TX (2.17) In particular, Def (k[ ]) = T 1 B0 B0

if X = Spec(B0) is affine. (iv) If X is a reduced algebraic scheme then there is an isomorphism

Def (k[ ]) =∼ Ext1 (Ω1 , O ) X OX X X and the exact sequence (2.17) is isomorphic to the local-to-global exact se- quence for Exts: 0 → H 1(X, T ) → Ext1 (Ω1 , O ) X OX X X (2.18) 0 1 1 2 → H (X, Ext (Ω , OX )) → H (X, TX ) OX X

Proof. Obviously DefX and DefX satisfy condition H0. To verify the other condi- tions we assume first that X = Spec(B0) is affine. ¯ Let’s prove that DefB0 satisfies H.Let A A

A ¯ be homomorphisms in A, with A → A a small extension. Letting A = A ×A A we have a commutative diagram with exact rows:

0 → ( ) → A¯ → A → 0 ↓↓ (2.19) 0 → ( ) → A → A → 0 Consider an element of

Def (A ) × ( ) Def (A ) B0 DefB0 A B0 66 2 Formal deformation theory which is represented by a pair of deformations f : A → B and f : A → B of

B0 such that A → B ⊗A A and A → B ⊗A A are isomorphic deformations. As- ∼ ∼ sume that the isomorphism is given by A-isomorphisms B ⊗A A = B = B ⊗A A, where A → B is a deformation. In order to check H¯ it suffices to find a deformation f¯ : A¯ → B¯ inducing ( f , f ).Let ¯ B = B ×B B endowed with the obvious homomorphism f¯ : A¯ → B¯ . It is elementary to check that ¯ ⊗ =∼ ¯ ⊗ =∼ there are an A -isomorphism B A¯ A B and an A -isomorphism B A¯ A B . Therefore we only need to check that f¯ is flat. Tensoring diagram (2.19) with ⊗ ¯ A¯ B we obtain the following diagram with exact rows: ( ) ⊗ ¯ → ¯ → → A¯ B B B 0 ↓↓

0 → B0 → B → B → 0 where the second row is given by Lemma A.9. This diagram shows that ( ) ⊗ ¯ → ¯ A¯ B B is injective: the flatness of f¯ follows from Lemma A.9. = [ ] Let’s prove that DefB0 satisfies H . Assume in the above situation that A k and A = k, and let f˜ : A¯ → B˜ be a deformation such that α( f˜) = ( f , f ). Then the diagram ˜ → ˜ ⊗ =∼ B B A¯ A B ↓↓ ˜ ⊗ =∼ → B A¯ A B B0 commutes: the universal property of the fibre product implies that we have a homo- morphism γ : B˜ → B¯ of deformations, hence an isomorphism by Lemma A.4. This proves that the fibres of α contain only one element, i.e. α is bijective. Therefore

DefB0 satisfies condition H . Let’s prove (i) for X arbitrary. Consider a diagram in A: A A

A ¯ with A → A a small extension and let A = A ×A A . Consider an element ([X ], [X ]) ∈ ( ) × ( ) DefX A DefX (A) DefX A Therefore we have a diagram of deformations: X X  f  f ↓ X ↓ Spec(A ) ↓ Spec(A )  Spec(A) 2.4 The local moduli functors 67 where the morphisms f and f induce isomorphisms of deformations

∼ ∼ X ×Spec(A ) Spec(A) = X = X ×Spec(A ) Spec(A)

¯ ¯ Consider the sheaf of A-algebras OX ×O OX on X. Then X := (|X|, OX ×O ¯ X X OX ) is a scheme over Spec(A) (by the proof of the affine case). Reducing to the affine case one shows that X¯ is flat over Spec(A¯). Therefore X¯ is a deformation of X over Spec(A¯) inducing the pair ([X ], [X ]). This shows that the map

( ¯) → ( ) × ( ) DefX A DefX A DefX (A) DefX A

¯ is surjective, proving H for DefX . Moreover, if the deformations X and X are X¯ ¯ locally trivial then so is , and therefore H holds for the functor DefX as well. Now assume that A = k[ ] and A = k. Then the previous diagram becomes

X X  f  f X

In this case any X˜ → Spec(A¯) inducing the pair

([X ], [X ]) ∈ ( ) × ( ) DefX A DefX (A) DefX A is such that the isomorphisms

X˜ × ( ) =∼ X , X˜ × ( ) =∼ X Spec(A¯) Spec A Spec(A¯) Spec A

= X˜ × ( ) X˜ induce the identity on X Spec(A¯) Spec k . Therefore fits into a commutative diagram X˜  X X  f  f X By the universal property of the fibred sum of schemes we then get a morphism of deformations X¯ → X˜, which is necessarily an isomorphism. This proves property

H for DefX . The proof for DefX is similar. (ii) The identification (2.16) has been proved in 1.2.9. The verification that it is k-linear is elementary and will be left to the reader. (iii) If X ⊂ X ↓↓ Spec(k) → Spec(k[ ]) 68 2 Formal deformation theory is a first-order deformation of X then X ⊂ X is an extension of X by OX because by ∼ the flatness of X over Spec(k[ ]) we have OX = OX (Lemma A.9). Conversely, given such an extension X ⊂ X we have an exact sequence

j 0 → OX → OX → OX → 0

OX has a natural structure of k[ ]-algebra by sending, for any open U ⊂ X, → j(1). It follows from Lemma A.9 that X is flat over Spec(k[ ]). Let’s assume first that X = Spec(B0) is affine. Then

Ex (X, O ) = H 0(X, T 1 ) = T 1 k X B0 B0 and the exact sequence (2.17) reduces to the isomorphism

Def (k[ ]) =∼ T 1 B0 B0 which holds by what we have just remarked. Let’s assume that X is general. The map τ corresponds to the inclusion ( [ ]) ⊂ ( [ ])  DefX k DefX k in view of (2.16). The map associates to a first-order ξ 1 {ξ } deformation of X the section of TX defined by the restrictions |Ui for some affine open cover {Ui }. It is clear that Im(τ) = ker(). ρ ∈ 0( , 1) We now define .Leth H X TX be represented, in a suitable affine open cover U ={Ui = Spec(Bi )} of X, by a collection of k-extensions Ei of Bi by Bi . σ : E =∼ E Since h is a global section there exist isomorphisms ij j|Ui ∩U j i|Ui ∩U j . These isomorphisms patch together to give an extension E if and only if h ∈ Im() if and σ only if we can find new isomorphisms ij such that σ σ = σ ij jk ik (2.20) on Ui ∩ U j ∩ Uk. Such isomorphisms are of the form σ = σ θ ij ij ij θ E where ij is an automorphism of the extension j|Ui ∩U j . The collection of auto- 1 morphisms (θij) corresponds, via Lemma 1.2.6, to a 1-cochain (tij) ∈ C (U, TX ); ( ) (σ ) conversely, every 1-cochain tij defines a system of isomorphisms ij ; and the condition (2.20) is satisfied if and only if (tij) is a 1-cocycle. Therefore we define ρ(h) to be the class of the 2-cocycle (tij + t jk − tik). With this definition we clearly have ker(ρ) = Im(). We leave to the reader to verify that the definition of ρ does not depend on the choices made.

(iv) Since we have a natural identification DefX (k[ ]) = Exk(X, OX ) we con- clude by Theorem 1.1.10. The exact sequences (2.17) and (2.18) are isomorphic in view of the isomorphism T 1 =∼ Ext1 (Ω1 , O ) X OX X X given by Corollary 1.1.11, page 18.  2.4 The local moduli functors 69

Corollary 2.4.2. If X is a projective scheme or an affine scheme with at most isolated singularities then DefX has a semiuniversal formal element. 1( , ) 0( , 1) Proof. The condition implies that H X TX and H X TX are finite-dimensional vector spaces. Therefore the conclusion follows from Theorem 2.4.1 and Sch- lessinger’s theorem.  The stronger property of being prorepresentable is not satisfied in general by DefX . We will discuss this matter in § 2.6. The case X affine will be taken up again and analysed in detail in Section 3.1.

Definition 2.4.3. If (R, uˆ) is a semiuniversal couple for DefX then the Krull di- mension of R (i.e. the maximum of the dimensions of the irreducible components of Spec(R)) is called the number of moduli of X and it is denoted by µ(X). Remark 2.4.4. Let X be a reduced scheme and let ξ : X → Spec(k[ ]) beafirst- order deformation of X. Then the conormal sequence of X ⊂ X → O → Ω1 → Ω1 → 0 X X |X X 0 (2.21) is exact and defines the element of Ext1 (Ω1 , O ) which corresponds to ξ in the OX X X identification Def (k[ ]) = Ext1 (Ω1 , O ) of Theorem 2.4.1(iv) (see also Theo- X OX X X rem 1.1.10). Given an infinitesimal deformation ξ : X → Spec(A) of X we have a Kodaira– Spencer map κξ : t → Ext1 (Ω1 , O ) which associates to a tangent vector θ ∈ t A OX X X A the conormal sequence 2.21 of the pullback of ξ to Spec(k[ ]) defined by θ.

2 Example 2.4.5. Let X be a projective curve. Then H (X, TX ) = (0) and the exact 0( , 1) = ( [ ]) = ( ) sequence (2.17) shows that if h X TX 0 then DefX k 0 and X is not rigid. In particular, if X is reduced and has at least one nonrigid singular point then X is not rigid. By the way, it is not known whether rigid curve singularities exist at all.

2.4.2 Obstruction spaces

The elementary analysis of obstructions to lifting infinitesimal deformations carried out in Chapter 1 can be interpreted as the description of obstruction spaces for the corresponding deformation functors. More precisely, we have the following: 2 Proposition 2.4.6. Let X be a nonsingular algebraic variety. Then H (X, TX ) is an obstruction space for the functor DefX . If X is an arbitrary algebraic scheme then 2( , ) H X TX is an obstruction space for the functor DefX . Proof. The proposition is just a rephrasing of 1.2.12.  Corollary 2.4.7. Let X be a nonsingular projective algebraic variety. Then 1 1 2 h (X, TX ) ≥ µ(R) ≥ h (X, TX ) − h (X, TX ) The first equality holds if and only if X is unobstructed. Proof. It is an immediate consequence of 2.4.6 and of Corollary 2.2.11.  70 2 Formal deformation theory

If X is a singular algebraic scheme the previous results give no information about obstructions of the functor DefX . The following proposition addresses the case of a reduced local complete intersection (l.c.i.) scheme.

Proposition 2.4.8. Let X be a reduced l.c.i. algebraic scheme X, and assume char(k) = 0. Then Ext2 (Ω1 , O ) is an obstruction space for the functor Def . OX X X X Proof. Let A be in ob(A) and let X → X ξ : ↓↓f Spec(k) → Spec(A) be a deformation of X over A. We need to define a k-linear map

2 1 oξ : Ex (A, k) → Ext (Ω , O ) k OX X X having the properties of an obstruction map according to Definition 2.2.9. Consider an element of Exk(A, k) represented by an extension

e : 0 → (t) → A˜ → A → 0 and the conormal sequence of e:

→ ( ) → Ω ⊗ → Ω → 0 t A˜/k A˜ A A/k 0 (2.22) which is exact by Lemma B.10. Since f is flat, pulling back (2.22) to X we obtain the exact sequence

∗ ∗ 0 → O → f (Ω1 ) → f (Ω1 ) → 0 (2.23) X Spec(A˜)|Spec(A) Spec(A)

Since X is a reduced l.c.i. the morphism f satisfies the hypothesis of Theorem D.2.8. Therefore the relative cotangent sequence of f :

→ ∗(Ω1 ) → Ω1 → Ω1 → 0 f Spec(A) X X /Spec(A) 0 (2.24) is exact. Composing (2.23) and (2.24) we obtain the 2-term extension

∗ 1 1 1 0 → OX → f (Ω ) → Ω → Ω → 0 Spec(A˜)|Spec(A) X X /Spec(A) which defines an element

2 1 2 1 oξ (e) ∈ Ext (Ω , O ) = Ext (Ω , O ) OX X /Spec(A) X OX X X

This defines the map oξ . The linearity of oξ is a consequence of the linearity of ( , ) → 1 (Ω , ) the map Exk A k ExtA A/k k associating to an extension e the conormal sequence (2.22). 2.4 The local moduli functors 71

Assume that there is a lifting of ξ to A˜, i.e. that we have a diagram:

X ⊂ X˜ ↓ f ↓ f˜ Spec(A) ⊂ Spec(A˜)

Then we have a commutative and exact diagram as follows:

0 ↓ ∗ 1 ∗ 1 0 → OX → f (Ω ) → f (Ω ) → 0 Spec(A˜)|Spec(A) Spec(A) ↓ ↓ → O → Ω1 → Ω1 → 0 X X˜ |X X 0 ↓↓ 1 1 (Ω )|X = Ω X˜ /Spec(A˜) X /Spec(A) ↓↓ 00

In this diagram the first row is the pullback of the second row and this implies that oξ (e) = 0. Conversely, assume that oξ (e) = 0. Then we have a commutative and exact diagram as follows:

0 ↓ ∗ 1 ∗ 1 0 → OX → f (Ω ) → f (Ω ) → 0 Spec(A˜)|Spec(A) Spec(A) ↓ ↓ 1 0 → OX → E → ΩX → 0 ↓ Ω1 X /Spec(A) ↓ 0 for some coherent sheaf E on X . By the construction of Theorem 1.1.10 one finds a sheaf of A-algebras OX˜ and an extension of sheaves of A-algebras

0 → OX → OX˜ → OX → 0

1 E = Ω O ˜ such that X˜ |X . It remains to be shown that X can be given a structure of sheaf of flat A˜-algebras. The shortest way to see this is to use Corollary 3.1.13(ii), to be proved in the next chapter: since X is a l.c.i. it implies that there are no obstruc- ˜ tions to lifting X locally to A. This means that the restriction of OX˜ to every affine ˜ open subset U ⊂ X defines a lifting of X|U to A, and therefore it is a sheaf of flat A˜-algebras.  72 2 Formal deformation theory

1 1 Example 2.4.9. If X is a reduced l.c.i. curve then Ext (Ω , OX ) is a torsion sheaf OX X 2 1 and Ext (Ω , OX ) = 0; therefore OX X

i j 1 H (X, Ext (Ω , OX )) = (0) OX X for all i, j with i + j = 2. It follows that Ext2 (Ω1 , O ) = 0 and X is unobstructed OX X X by Proposition 2.4.8. For a discussion of obstructions in the affine case we refer the reader to Subsec- tion 3.1.2.

2.4.3 Algebraic surfaces

In this subsection we will assume char(k) = 0. We will denote by S a projective nonsingular connected algebraic surface. Let (R, uˆ) be a semiuniversal formal de- formation of S, and denote by

µ(S) := dim(R) the number of moduli of S. Proposition 2.4.10.

2 0 1 10χ(OS) − 2(K ) + h (S, TS) ≤ µ(S) ≤ h (S, TS) (2.25)

2 If h (S, TS) = 0 then both inequalities are equalities. Proof. A direct application of the Riemann–Roch formula gives

1 2 2 0 h (S, TS) − h (S, TS) = 10χ(OS) − 2(K ) + h (S, TS) By applying Corollary 2.4.7 we obtain the conclusion.  The first inequality was proved by Enriques (see [52]). Examples 2.4.11. (i) If S is a minimal ruled surface and S → C is the ruling over a projective nonsingular curve C, then letting S(x) =∼ IP 1 be the fibre of any x ∈ C we have 1 h (S(x), TS|S(x)) = 0 as an immediate consequence of the exact sequence

0 → TS(x) → TS|S(x) → NS(x)/S → 0  OS(x)

1 Therefore R p∗TS = 0 by Corollary 4.2.6, and the Leray spectral sequence implies 2 1 that h (S, TS) = 0. Therefore S is unobstructed and µ(S) = h (S, TS). This com- putation for the rational ruled surfaces Fm is also done in Example B.11(iii), page 288. 2.4 The local moduli functors 73

(ii) Assume that S is a K3-surface. Then 2( , ) = 0( ,Ω1) = 1( , O ) = h S TS h S S h S S 0 Therefore S is unobstructed. Moreover, 0( , ) = 2( ,Ω1) = 1( ,ω ) = h S TS h S S h S S 0 and DefS is prorepresentable (Corollary 2.6.4). Formula (2.25) gives in this case

1 µ(S) = h (S, TS) = 20

(iii) Let π : X → S be the blow-up of S at a point s and E = π −1(s) the exceptional curve. Then we have a standard exact sequence

dπ ∗ 0 → TX −→ π TS → OE (−E) → 0

(to see it one can use the exact sequence (3.50), page 172, and note that coker(dπ) = Nπ , the normal sheaf of π, which is defined on page 162). We thus see that

2 ∼ 2 H (X, TX ) = H (S, TS)

This implies, for example, that non-minimal rational or ruled surfaces and blow-ups of K3-surfaces are unobstructed. (iv) When the Kodaira dimension of S is ≥ 1 then in general

2 0 1 h (S, TS) = h (S,Ω ⊗ ωS) = 0 and in fact such surfaces can be obstructed. Examples are given in Theorem 3.4.26, 0 page 185. If we assume that h (S, TS) = 0 the estimate for µ(S) given by Proposi- tion 2.4.10 becomes

2 1 2 0 1 10χ(OS) − 2(K ) ≤ µ(S) ≤ h (S, TS) = 10χ(OS) − 2(K ) + h (S,Ω ⊗ ωS)

0 1 and to give an upper bound for µ(S) amounts to giving one for h (S,Ω ⊗ ωS) in 2 terms of pa, K , q. We refer the reader to [27] for a more detailed discussion of this point. For examples of obstructed surfaces see also [24], [87], [96], [28]. 0 (v) If h (S,ωS)>0 and q > 0 then certainly

0 1 h (S,Ω ⊗ ωS)>0

This is because there is a bilinear pairing 0( ,Ω1) × 0( ,ω ) → 0( ,Ω1 ⊗ ω ) H S S H S S H S S which is non-degenerate on each factor. 74 2 Formal deformation theory

For example, take S = C1 × C2 where C1 and C2 are projective nonsingular 0 connected curves of genera g1 ≥ 2 and g2 ≥ 2 respectively. Then h (S,ωS) = g1g2, q = g1 + g2 and − − T = p∗ω 1 ⊕ p∗ω 1 S 1 C1 2 C2 where pi : S → Ci is the i-th projection, i = 1, 2. Therefore

0 1 2 h (TS) = 0, h (TS) = 3(g1 + g2) − 6, h (TS) = g2(3g1 − 3) + g1(3g2 − 3)

On the other hand, S is unobstructed. In fact there is a natural morphism of functors

: × → f DefC1 DefC2 DefS which is clearly an isomorphism on tangent spaces: in fact

( × )( [ ]) := ( [ ]) × ( [ ]) = DefC1 DefC2 k DefC1 k DefC2 k

− − = H 1(C ,ω 1) ⊕ H 1(C ,ω 1) = H 1(S, T ) 1 C1 2 C2 S × Therefore, since DefC1 DefC2 is smooth because both factors are (Proposition 2.2.5(vii)), DefS is smooth by Corollary 2.3.7. Actually, f is an isomorphism be- cause both functors are prorepresentable (see Remark 2.3.8, page 60). This example can be generalized to any finite product of nonsingular projective connected curves of genera ≥ 2. If S is an abelian surface then

0 1 0 h (S,Ω ⊗ ωS) = 2 = h (S, TS)

Formula (2.25) gives 1 2 ≤ µ(S) ≤ h (S, TS) = 4 In fact the second equality holds because abelian surfaces are unobstructed despite 2 the fact that h (S, TS) = 0. This is a property common to all abelian varieties (see [137] and [135]). (vi) One should keep in mind that µ(S) is defined as the number of moduli of S in a formal sense. This is because the semiuniversal formal deformation (R, uˆ) can be non-algebraizable (see § 2.5 for the notion of algebraizability). For exam- ple µ(S) = 20 for a K3-surface, but every algebraic family of K3-surfaces with injective Kodaira–Spencer map at every point has dimension ≤ 19 (see § 3.3). Simi- larly an abelian surface has µ(S) = 4 but every algebraic family of abelian surfaces with injective Kodaira–Spencer map at every point has dimension ≤ 3 (see Example 3.3.13). In order to give an algebraic meaning to the number of moduli one should take the maximum dimension of a semiuniversal formal deformation of a pair (S, L) where L is an ample invertible sheaf on S. See § 3.3 and the appendix by Mumford to Chapter V of [189]. 2.5 Formal versus algebraic deformations 75 2.5 Formal versus algebraic deformations

We have already mentioned (see Examples 1.2.5 and 1.2.11) that infinitesimal de- formations do not explain faithfully some of the phenomena which can occur when one considers deformations parametrized by algebraic schemes or by the spectrum of an arbitrary noetherian, or even e.f.t., local ring. In this section we will make such statements precise. We start with a few definitions and some terminology. Let (S, s) be a pointed scheme. An etale neighbourhood of s in S is an etale morphism of pointed schemes f : (T, t) → (S, s) such that the following diagram commutes: T

t ↓f

s Spec(k(s)) −→ S The definition implies that k(s) =∼ k(t), i.e. f induces a trivial extension of the ˆ ∼ ˆ residue fields at s and t; therefore OS,s = OT,t by C.7. Affine neighbourhoods of s are particular etale neighbourhoods. Given two etale neighbourhoods (T, t) and (U, u) of s ∈ S a morphism (T, t) → (U, u) is given by a commutative diagram of pointed schemes: (T, t) → (U, u)

(S, s) Lemma 2.5.1. Let f : X → Y be an etale morphism and g : Y → X a section of f . Then g is etale.

Proof. Use C.2(v). 

Proposition 2.5.2. Let S be a scheme. The etale neighbourhoods of a given s ∈ S form a filtered system of pointed schemes.

Proof. Given two etale neighbourhoods (S , s ) and (S , s ), they are dominated by a third, namely:

S ×S S → S ↓↓ S → S

Now let f1, f2 : (S , s ) → (S , s ) be two morphisms between etale neighbour- hoods. Then there exists a third etale neighbourhood (S , s ) and a morphism (S , s ) → (S , s ) which equalizes them. In fact, consider the diagram:

pr S ×S S −→ S ↓ pr ↓ S → S 76 2 Formal deformation theory

We can shrink S and S so that S is affine and S is connected. Then the graphs

Γ1 and Γ2 of f1 and f2 are closed, because S is affine, and open, because they are images of sections of the etale morphism pr , which are etale. Therefore they are connected components of S ×S S .But(s , s ) ∈ Γ1 ∩Γ2 and therefore Γ1 = Γ2.

It follows that f1 = f2 on S = Γ1 = Γ2.  Definition 2.5.3. Given a scheme S and a point s ∈ S we define the local ring of S in s in the etale topology to be ˜ OS,s := lim OS ,s → (S ,s ) where the limit is taken for (S , s ) varying through all the etale neighbourhoods of s. ˜ ˜ The ring OS,s is also called the henselization of OS,s. (Note that OS,s is a local ring, because it is a limit of a filtering system of local rings and local homomorphisms.) A local ring A is called henselian if for the closed point s of S = Spec(A) one has ˜ ˜ A := OS,s = OS,s = A The henselization of an e.f.t. local k-algebra is called an algebraic local ring. Therefore the local ring in the etale topology of a point of an algebraic scheme is an algebraic local ring. For a given scheme S and point s ∈ S there is a canonical homomorphism ˜ OS,s → OS,s which is flat and induces an isomorphism of the completions ˆ ∼ ˜ OS,s = OS,s because every OS,s → OS ,s does. Moreover, ˜ ˆ OS,s ⊂ OS,s ⊂ OS,s ˜ because OS,s → OS,s is faithfully flat and OS,s is separated for the m-adic topology. ˜ ˆ ˆ In particular, we see that OS,s = OS,s if OS,s = OS,s, i.e. a local k-algebra in A (i.e. complete noetherian with residue field k) is henselian. Theorem 2.5.4 (Nagata). If A is a noetherian local ring then A˜ is noetherian. Proof. We have A ⊂ A˜ ⊂ Aˆ and Aˆ = A˜. Moreover,

A˜ = lim A → with A local algebras etale over A and inducing trivial residue field extension. To prove that A˜ is noetherian it suffices to prove that every ascending chain of finitely generated ideals of A˜ ⊆ ⊆···⊆ ⊆ a1 a2 an { ˆ} ˆ stabilizes. The chain an A stabilizes because A is noetherian. Therefore it suffices to prove that if a, b ⊂ A˜ are finitely generated ideals such that a Aˆ = bAˆ then 2.5 Formal versus algebraic deformations 77 a = b. Since a and b are finitely generated one can find A ⊃ A as above and finitely generated ideals a , b ⊂ A such that a = a A˜, b = b A˜. It follows that a Aˆ = b Aˆ. But since A is noetherian it follows that a = b and therefore a = b.  The following proposition gives a geometrical characterization of the henselization.

Proposition 2.5.5. Let A be a local ring, S = Spec(A),s∈ S the closed point. A is henselian if and only if every morphism f : Z → S such that there is a point z ∈ Z with f (z) = s, k(s) = k(z) and f etale at z, admits a section.

Proof. Assume the condition is satisfied. If A → A is an etale homomorphism inducing an isomorphism of the residue fields then the induced morphism f : Spec(A ) → S admits a section, which defines an isomorphism A =∼ A; therefore A is henselian. Conversely, assume A is henselian and let f : Z → S be a morphism ∼ satisfying the stated conditions. Then f induces an isomorphism A = OZ,z because A is henselian. The section is the composition ∼ S = Spec(A) = Spec(OZ,z) ⊂ Z  ∗∗∗∗∗∗

We will need the notion of “formal deformation of an algebraic scheme”, already introduced in § 2.2 for a general functor of Artin rings. Let X be an algebraic scheme and let A¯ be in ob(Aˆ). Then a formal deformation of X over A¯ is just a formal ¯ element ηˆ ∈ DefX (A) for DefX : by definition it is a sequence {ηn} of infinitesimal deformations of X fn X −→ Xn η : n ↓↓πn ¯ Spec(k) → Spec(An) ¯ ¯ n+1 where An = A/m ¯ , such that for all n ≥ 1 ηn induces ηn−1 by pullback under the A ¯ ¯ natural inclusion Spec(An−1) → Spec(An), i.e. we have:

fn−1 ¯ X −→ Xn ⊗ ¯ An−1 = Xn−1 η : An n−1 ↓↓πn−1 ¯ Spec(k) → Spec(An−1) ¯ ¯ We will denote such a formal deformation by (A, {ηn}) or by (A, η)ˆ . It can be also viewed as the morphism of formal schemes

π¯ : X → Specf(A¯) where X = (X, lim OX ), π¯ = lim π ← n ← n (see [84] and [16] for the definition and main properties of formal schemes). 78 2 Formal deformation theory

Note that π¯ is a flat morphism of formal schemes, namely for every x ∈ X O ¯ = O the local ring X ,x is flat over A Specf(A¯),π(¯ x). This is an almost immediate consequence of the general version of the local criterion of flatness ([3], ExposeIV,´ Corollaire 5.8). ¯ The formal deformation (A, {ηn}) will be called trivial (resp. locally trivial)if each ηn is trivial (resp. locally trivial). ¯ A formal deformation (A, {ηn}) is not to be confused with a deformation of X over Spec(A¯), just as a formal scheme over Specf(A¯) is not in general a formal com- pletion of a scheme over Spec(A¯) (see Definition 2.5.10 below and the discussion following it). Let X be a projective scheme and consider a flat family of deformations of X parametrized by an affine scheme S = Spec(B), with B in (k-algebras)

f X −→ X η : ↓↓π s Spec(k) −→ S namely a cartesian diagram with π projective and flat. The deformation η is called algebraic if B is a k-algebra of finite type. If B is in ob(A∗) then η is called a local deformation of X.IfB is in ob(A) we obtain an infinitesimal deformation of X;it is simultaneously local and algebraic. We will identify the deformation η with the couple (S,η)or (B,η)and we will also denote it by (S, s,η)or (B, s,η). Given such a deformation (S, s,η) let ηn be the infinitesimal deformation in- duced by pulling back η under the natural closed embedding

n+1 Spec(OS,s/m ) → S

n+1 ˆ n+1 ˆ We have OS,s/m = OS,s/mˆ and therefore it follows that (OS,s, {ηn}) is a formal deformation of X. It will be called the formal deformation defined by (or associated to) η. (S, s,η) is called formally trivial (resp. formally locally trivial) if the formal deformation defined by η is trivial (resp. locally trivial).

Lemma 2.5.6. Let (S, s,η)be a deformation of an algebraic scheme X, f : (S˜, s˜) → (S, s) an etale neighbourhood of s in S and (S˜, s˜, η)˜ the deformation of X obtained by pulling back η by f . Then the formal deformations of X defined by η and by η˜ are isomorphic.

Proof. We have a cartesian diagram of formal schemes:

X˜ → X ↓ π˜¯ ↓¯π (Oˆ ) → (Oˆ ) Specf S˜,s˜ Specf S,s where π˜¯ and π¯ are the formal deformations defined by η˜ and by η respectively. Since Oˆ =∼ Oˆ  f is etale it induces an isomorphism S,s S˜,s˜ and the conclusion follows. 2.5 Formal versus algebraic deformations 79

Definition 2.5.7. A deformation (S, s,η)of X is called formally universal (resp. for- ˆ mally semiuniversal, formally versal) if the formal deformation (OS,s, {ηn}n≥0) as- sociated to η is universal (resp. semiuniversal, versal). An algebraic formally versal deformation of X is also called with general moduli. A flat family π : X → S is called formally universal (resp. formally semiuniver- sal, formally versal, with general moduli) at a k-rational point s ∈ Sif X (s) ⊂ X η : ↓↓π s Spec(k) −→ S is a formally universal, (resp. formally semiuniversal, formally versal, with general moduli) deformation of X (s).

The expression “general moduli” goes back to the classical geometers. Infor- mally, it means that the family parametrizes all possible “sufficiently small” deforma- tions of X (s); when the family π parametrizes varieties for which there is a moduli scheme, or just a moduli stack (whatever this means for the reader since we have not introduced these notions), π with general moduli means that the functorial morphism from S to the moduli scheme or stack is open at s. An expression like “consider a variety X with general moduli” is used to mean: “choose X as a general fibre in a family with general moduli”. The early literature on deformation theory of complex analytic manifolds (in the approach of Kodaira and Spencer) considered only families parametrized by complex analytic manifolds. In that context the expression “effectively parametrized” was used to mean “with injective Kodaira–Spencer map”, and the word “complete” was used to mean the analogue of the notion of formal versality, in the category of germs of complex analytic manifolds.

Proposition 2.5.8. Let (S, s,η)be a deformation of X. Then: (i) If η is formally versal (resp. formally semiuniversal or formally universal) then the Kodaira–Spencer map

κπ,s : Ts S → DefX (k[ ])

is surjective (resp. an isomorphism). (ii) If S is nonsingular at s and the Kodaira–Spencer map κπ,s is surjective (resp. an isomorphism) then η is formally versal (resp. formally semiuniversal) and X is unobstructed, i.e. the functor DefX is smooth.

Proof. (i) is obvious in view of the definitions of versality and semiuniversality of a formal couple applied to h ˆ → DefX . OS,s (ii) follows from Proposition 2.2.5(iii) applied to f : h ˆ → DefX .  OS,s The proposition applies in particular to an algebraic deformation, giving a crite- rion for it to have general moduli. A classical result (see [107]) states the complete- ness of a complex analytic family of compact complex manifolds if the map κπ,s is 80 2 Formal deformation theory surjective. Part (ii) of Proposition 2.5.8 is the algebraic version of this result. It turns out to be very useful because it reduces the verification of formal versality and of unobstructedness to the computation of the Kodaira–Spencer map. ¯ Definition 2.5.9. A formal deformation (A, {ηn}) of X is called algebraizable if there ˆ ∼ ¯ exists an algebraic deformation (S, s,ξ)of X and an isomorphism OS,s = A sending ¯ ηn to ξn for all n (i.e. (A, {ηn}) is isomorphic to the formal deformation defined by ¯ ξ). The deformation (S, s,ξ)is called an algebraization of (A, {ηn}). It goes without saying that an algebraization of a formal versal (resp. semiuniver- sal, universal) deformation is formally versal (resp. formally semiuniversal, formally universal). The existence of algebraizations is a highly nontrivial problem. It can be considered as the counterpart of the convergence step in the construction of local families of deformations in the Kodaira–Spencer theory of deformations of com- pact complex manifolds. But the algebraic case presents some characteristic features which make the two theories radically different in methods and in results. In par- ticular, a projective algebraic variety X defined over the field of complex numbers need not have an algebraic formally versal deformation, even in the case when DefX is prorepresentable and unobstructed (see Example 2.5.12 below); but, according to the KNSK theory (see the Introduction), such a variety has a complete deformation in the complex analytic sense. ∗∗∗∗∗∗

For the rest of this section we will need to assume some familiarity with formal schemes: for some definitions and results not contained in [84] we will refer to [1]. The following definition introduces an important notion weaker than algebraizability.

Definition 2.5.10. Let X be an algebraic scheme and let Abein¯ ob(Aˆ).Aformal deformation (A¯, η)ˆ of X is called effective if there exists a deformation X → X η¯ : ↓↓π Spec(k) → Spec(A¯) of X over A¯ such that ηˆ is the formal deformation associated to η¯. Equivalently, letting π¯ : X → Specf(A¯) be the morphism of formal schemes associated to the formal deformation (A¯, η)ˆ , its effectivity means that there is a deformation η¯ such that X is the formal completion of X along X and π¯ is the morphism of formal schemes induced by π; in symbols:

X =∼ ( ¯) × X Specf A Spec(A¯)

It is obvious that the trivial formal deformation X → Specf(A¯), where

X = X × Specf(A¯) = (X, lim O ⊗ A¯/mn) ← X k is effective: it is the formal completion of the trivial deformation 2.5 Formal versus algebraic deformations 81

X × Spec(A¯) → Spec(A¯)

In particular, for any A¯ in ob(Aˆ) the formal projective space over A¯

Pr := ( r , O r ) ¯ IP lim IP ¯ A ← An is effective, being the formal completion of IP r × Spec(A¯) along the closed fibre IP r = IP r ×Spec(k). A consequence of Grothendieck’s theorem of formal functions is that a formal deformation of a proper algebraic scheme can be effective in at most one way; more precisely we have:

Theorem 2.5.11. Let X be a proper algebraic scheme. If X → X X → Y η¯ : ↓↓π ξ¯ : ↓↓q Spec(k) → Spec(A¯) Spec(k) → Spec(A¯) are two deformations of X with Ain¯ ob(Aˆ) such that the associated formal defor- mations (A¯, η)ˆ , (A¯, ξ)ˆ are isomorphic, then η¯ and ξ¯ are isomorphic deformations.

Proof. Let S = Spec(A¯) and Sˆ = Specf(A¯). For a given proper S-scheme f : Z → S we denote by Zˆ the formal completion of Z along Z(s) = f −1(s), where s ∈ S Fˆ O is the closed point, and by the Zˆ -sheaf obtained as the formal completion of a quasi-coherent OZ -sheaf F. It suffices to show that the correspondence ϕ →ˆϕ defines a one-to-one correspondence (X , Y) → (Xˆ, Yˆ) HomS HomSˆ because it will follow that Xˆ =∼ Yˆ, i.e. (A¯, η)ˆ =∼ (A¯, ξ)ˆ , implies X =∼ Y. The proof goes as follows. For any proper S-scheme f : Z → S and coherent sheaves F and G on Z the theorem of formal functions ([84], p. 277) implies that there is an isomorphism ∧ ∼ ˆ ˆ Hom(F, G) = lim HomO (F , G ) = Hom ˆ (F, G) ← Zn n n Z

Moreover, since Hom(F, G) is a finitely generated A¯-module, it coincides with its m A¯ -adic completion ([127], Theorem 8.7, p. 60); thus we have an isomorphism (F, G) =∼ (Fˆ , Gˆ) Hom HomZˆ

It maps a homomorphism u : F → G to its completion uˆ : Fˆ → Gˆ. It is easy to see that uˆ is injective (resp. surjective) if and only if u is injective (resp. surjective) ([1], ch. III, Cor. 5.1.3). In particular, we have a one-to-one correspondence between O O sheaves of ideals of Z and of Zˆ , equivalently between closed subschemes of Z and closed formal subschemes of Zˆ . Now take Z = X ×S Y and view it as an S-scheme by any of the projections. ˆ = Xˆ × Yˆ Then we have a natural identification Z Sˆ ([1], ch I, Prop. 10.9.7). We 82 2 Formal deformation theory deduce a one-to-one correspondence between closed subschemes of X ×S Y and Xˆ × Yˆ X → Y of Sˆ , in particular between graphs of S-morphisms and graphs of morphisms Xˆ → Yˆ of Sˆ-formal schemes.  In general, not every formal deformation of a given algebraic scheme need to be effective, just as deformations of algebraic varieties defined over the field of complex numbers need not be algebraic. Here is a typical example.

Example 2.5.12 (K3-surfaces). For this example we will need a result to be proved in Chapter 3. Let X be a projective K 3-surface, and assume for simplicity that Pic(X) = ZZ[H] for some divisor H.Let(A¯, η)ˆ be the formal universal deforma- tion of X and X the corresponding formal scheme over Sˆ = Specf(A¯). By Example 2.4.11(ii) we know that ¯ ∼ A = k[[X1,...,X20]] where X1,...,X20 are indeterminates. Claim: (A¯, η)ˆ is not effective. Otherwise there is a proper smooth morphism f : X → S = Spec(A¯) such that Xˆ =∼ X , where Xˆ is the formal completion of X along X = f −1(s), s ∈ S the closed point. Since X is of finite type over the integral scheme S, it has a nontrivial line bundle L which we may assume to correspond to a Cartier divisor whose support does not contain X and has a nonempty intersection with X; therefore ∼ L has a nontrivial restriction to the closed fibre X,sayL ⊗ OX = nH, n = 0. By [1], Ch. III, Cor. 5.1.6, L corresponds to a line bundle Lˆ on X which extends ∼ L ⊗ OX = nH.ButnH cannot be extended to all of X because it extends to a 19-dimensional space of first-order deformations of X (see page 150). Therefore we have a contradiction and the claim is proved.

The following is a basic result of effectivity of formal deformations.

Theorem 2.5.13 (Grothendieck [71]). Let X be a projective scheme. Then: (i) Let π¯ : X → Specf(A¯), Ain¯ ob(Aˆ), be a formal deformation of X. Assume that there is a closed embedding of formal schemes j : X ⊂ Pr such that π¯ = pj A¯ where p : Pr → Specf(A¯) is the projection. Then π¯ is effective. A¯ 2 (ii) Assume that h (X, OX ) = 0. Then every formal deformation of X is effective.

Proof. (i) The structure sheaf OX and the ideal sheaf IX are coherent sheaves of OPr -modules. Repeating verbatim the proof of the classical result of Serre (see [84], A¯ Theorem II.5.15, p. 121), one shows that there is an exact sequence

N g¯ OPr (−ni ) −→ OPr → OX → 0 A¯ A¯ i=1 for some positive integers n1,...,nN .Wehave 2.5 Formal versus algebraic deformations 83

N 0 r g¯ ∈ H (P ¯ , OPr (ni )) A A¯ i=1 and, by the theorem on formal functions ([84], ch. III, Theorem 11.1) we have 0 r 0 r ∧ H (P ¯ , OPr (ni )) = H (IP ¯ , OIP r (ni )) A A¯ A A¯ 0 r ¯ Since H (IP ¯ , OIP r (ni )) is an A-module of finite type we have A A¯ 0 r ∧ 0 r H (IP ¯ , OIP r (ni )) = H (IP ¯ , OIP r (ni )) A A¯ A A¯ ([127], Theorem 8.7, p. 60) and therefore 0 r 0 r H (P ¯ , OPr (ni )) = H (IP ¯ , OIP r (ni )) A A¯ A A¯ This implies that the homomorphism g¯ is induced by a homomorphism N g : OIP r (−ni ) → OIP r A¯ A¯ i=1 whose cokernel is the structure sheaf of a closed subscheme X ⊂ IP r having X A¯ as formal completion along X (o). The flatness of X over Spec(A¯) follows from Theorem A.5. ¯ (ii) Let (A, {ηn}) be a formal deformation of X, where:

fn X −→ Xn η : n ↓↓πn ¯ Spec(k) → Spec(An) For each n ≥ 1 we have an exact sequence: n m ρ → A¯ ⊗ O −→exp O∗ −→ O∗ → 0 + k X X X 0 mn 1 n n−1 A¯ where ρ is the natural restriction, exp(σ ) = 1 + σ and where we have identified mn mn A¯ ⊗ O = A¯ ⊗ O + k X + ¯ Xn mn 1 mn 1 An A¯ A¯ 2 using the flatness of πn. From the hypothesis h (X, OX ) = 0 we deduce that the homomorphism of Picard groups Pic(Xn) → Pic(Xn−1) is surjective. This implies that if we fix a very ample line bundle L on X such that H 1(X, L) = 0 then for each ≥ X = n 1 we can find a line bundle Ln on n such that Ln|Xn−1 Ln−1.ByNote3 N of § 4.2, Ln is very ample and defines an embedding Xn ⊂ IP ¯ where N + 1 = An h0(X, L). This means that the formal scheme X is a formal closed subscheme of P N . Now we conclude by part (i).  A¯ After these preliminaries we can state the following special case of Artin’s alge- braization theorem. 84 2 Formal deformation theory

Theorem 2.5.14 (Artin [12]). Let X be a projective scheme and let (A¯, η)ˆ be an effective formal versal deformation of X. Then (A¯, η)ˆ is algebraizable.

The universal formal deformation (A¯, η)ˆ of a K3-surface is not effective (Exam- ple 2.5.12) and in fact it is not algebraizable: this is well known if k = C because K3-surfaces have arbitrarily small deformations which are non-algebraic complex manifolds. ∗∗∗∗∗∗

To give a complete proof of Theorem 2.5.14 is beyond our goals. Nevertheless, we want to introduce some relevant definitions and preliminary results which are needed for its general statement (see below) and proof. Consider a covariant functor

F : (k-algebras) → (sets)

Suppose given R in ob(A∗) and an element u¯ ∈ F(Rˆ). An important question is whether it is possible to approximate u¯ insomewaybyanelementu ∈ F(R). Indeed, every algebraization problem can be reduced to such a question for an appropriate functor F. In this context the following is a natural definition of approximation:

Definition 2.5.15. Suppose given R in ob(A∗), u¯ ∈ F(Rˆ) and u ∈ F(R).Ifc> 0 is an integer we say that u¯ and u are congruent modulo mc, in symbols

u¯ ≡ u (modulo mc) if u¯ and u induce the same element in F(Rˆ/mc ) = F(R/mc ). Rˆ R In order to make the problem tractable it turns out to be natural to impose the following finiteness condition on the functor F.

Definition 2.5.16. A functor

F : (k-algebras) → (sets) is said to be locally of finite presentation if for every filtering inductive system of k-algebras {Bi } the canonical map

lim F(B ) → F(lim B ) → i → i is bijective. A functor locally of finite presentation is sometimes also called limit preserving.

This is a natural finiteness property, first introduced in [1], ch. IV, Prop. 8.14.2, which is usually satisfied by the functors arising in algebraic geometry. An illustration of the meaning of Definition 2.5.16 is given by the following proposition and by its corollary. 2.5 Formal versus algebraic deformations 85

Proposition 2.5.17. Let R be a noetherian ring and B an R-algebra. Then B is an R- algebra of finite type if and only if for every filtering inductive system of R-algebras {Ci } the canonical map lim Hom − (B, C ) → Hom − (B, lim C ) → R alg i R alg → i is bijective.

Proof. Assume that the condition of the statement is satisfied, and write

B = lim B → i where the Bi ⊂ B are R-subalgebras of finite type. Then the hypothesis implies that the identity B → B factors as B → Bi → B for some i, and this implies B = Bi . Therefore B is of finite type. Conversely, assume that B is of finite type, and let {t1,...,tn} be a system of gen- erators of B as an R-algebra. Let {Ci } be a filtering inductive system of R-algebras, C = lim Ci and let lim Hom − (B, C ) → Hom − (B, C) (2.26) → R alg i R alg be the natural map. Consider two elements of lim HomR−alg(B, Ci ) , given by two (θ ) (θ ) compatible systems of homomorphisms i and i , defining the same homomor- phism θ : B → C. Then, letting ϕi : Ci → C and ϕ ji : Ci → C j be the canonical homomorphisms of the inductive system, for each s = 1,...,n there is an index is such that θ(t ) = ϕ (θ (t )) = ϕ (θ (t )) s is is s is is s

We may assume that all the is are equal, say to i0. In the same way we can find an index j0 ≥ i0 such that

ϕ (θ (t )) = ϕ (θ (t )) j0i0 i0 s j0i0 i0 s for all s = 1,...,n; this means that θ (t ) = θ (t ), i.e. θ = θ . Therefore the j0 s j0 s j0 j0 map (2.26) is injective. Now let θ : B → C be a homomorphism, and let cs = θ(ts), s = 1,...,n. Write B = R[T1,...,Tn]/J, and let {Pk}k=1,...,m be a system of generators of the ideal J. Then we have

Pk(c1,...,cn) = θ(Pk(t1,...,tn)) = 0, k = 1,...,m

,..., ∈ ϕ ( ) = Let i0 be an index such that there exist elements x1 xn Ci0 such that i0 xs cs, i = 1,...,n. One has ϕ ( ( ,..., )) = ( ,..., ) = = ,..., i0 Pk x1 xn Pk c1 cn 0 k 1 m 86 2 Formal deformation theory

Then there is an index j0 ≥ i0 such that = ϕ (( ( ,..., )) = (ϕ ( ), . . . , ϕ ( )) = ,..., 0 j0i0 Pk x1 xn Pk j0i0 x1 j0i0 xn k 1 m θ : → One deduces the existence of a k-algebra homomorphism j0 B C j0 such that θ ( ) = ϕ ( ) = ,..., j0 ts j0i0 xs , s 1 n. This implies the existence of a homomorphism θ = ϕ θ : → ≥ θ j jj0 j0 B C j for every j j0.Itfollowsthat is the inductive limit of the system (θ j ), and therefore the map (2.26) is also surjective.  The following is an immediate consequence. Corollary 2.5.18. Let B be a k-algebra and

F = Homk−alg(B, −) : (k-algebras) → (sets) Then F is locally of finite presentation if and only if B is a k-algebra of finite type. Therefore we see that the condition of being locally of finite presentation for a representable functor coincides with the condition of being represented by an algebra of finite type. For the statement of the algebraization theorem we need some terminology which generalizes notions we have already introduced. Consider a covariant functor

F : (k-algebras) → (sets)

Let u0 ∈ F(k). A couple (A, u) where A is in ob(A) and u ∈ F(A) is called an infinitesimal deformation of u0 if u → u0 under the map

F(A) → F(A/m A) = F(k) We will denote by : A → Fu0 (sets) the functor of infinitesimal deformations of u0, i.e.: ( ) ={ ∈ ( ) : → ( ) → ( )} Fu0 A u F A u u0 under F A F k This is a functor of Artin rings and, according to the definitions given in § 2.2, we have the notions of formal deformation of u0, and of universal (resp. semiuniversal, resp. versal) formal deformation of u0.

Definition 2.5.19. A local deformation of u0 ∈ F(k) is a couple (A, u) where u ∈ ∗ F(A), A is in A and u → u0 under the map

F(A) → F(A/m A) = F(k)

An algebraic deformation of u0 is a triple (B, x, u) where B is a k-algebra of finite type, x ∈ Spec(B) is a k-rational point and u ∈ F(B) is such that u → u0 under the map F(B) → F(k) induced by the composition

B → OSpec(B),x → k(x) = k 2.5 Formal versus algebraic deformations 87

To every local deformation (A, u) of u0 one associates the formal deformation ( ˆ, { }) ˆ ∈ ( / n+1) = A un , where A is the m A-adic completion of A, and un F A m A + + F(Aˆ/mn 1) is the image of u under the map F(A) → F(A/mn 1). We call Aˆ A ˆ (A, {un}) the formal deformation of u0 associated to (or defined by) (A, u). Simi- larly, the formal deformation of u0 defined by an algebraic deformation (B, x, u) of ˆ u0 is the formal deformation (OSpec(B),x , {un}) where u → un under the map

ˆ n+1 F(B) → F(OSpec(B),x /m ) induced by the natural homomorphism

n+1 ˆ n+1 B → OSpec(B),x /m = OSpec(B),x /m

A local deformation is called formally universal (resp. formally semiuniversal, resp. formally versal) if the associated formal deformation has the corresponding property. A similar definition is given for algebraic deformations. ¯ Definition 2.5.20. A formal deformation (A, {un}) of u0 ∈ F(k) is called algebraiz- able if it is associated to some algebraic deformation (B, x, u). In this case the de- ¯ formation (B, x, u) is called an algebraization of (A, {un}). ¯ ¯ A formal deformation (A, {un}) of u0 is called effective if there exists u¯ ∈ F(A) ¯ which induces {un}. In this case we will call (A, u¯) an effective formal deformation as well.

It is clear from the definition that an algebraic deformation (B, x, u) of u0 is an ¯ ¯ ˆ algebraization of the effective formal deformation (A, u¯) if and only if A = OX,x and u ≡¯u mod mn for all n ≥ 0. ¯ Note that if the formal deformation (A, {un}) of u0 is associated to a local de- formation (A, u) then it is effective: in fact it follows that {un} is also associated to the deformation u¯ ∈ F(Aˆ) which is the image of u under the map F(A) → F(Aˆ). Similarly, an algebraizable deformation is effective. The algebrization theorem states that the converse is true for versal deformations. Namely:

Theorem 2.5.21 (algebraization theorem [12]). Let

F : (k-algebras) → (sets) be a functor locally of finite presentation and let u0 ∈ F(k). Then every effective versal formal deformation of u0 is algebraizable.

Theorem 2.5.21 is a generalization of Theorem 2.5.14 because the deformation functor DefX of a projective scheme X can be extended to a functor locally of finite presentation defined on the category (k-algebras). 88 2 Formal deformation theory

The main technical ingredient in the proof of the algebraization theorem is the following:

Theorem 2.5.22 (approximation theorem [11]). Let

F : (k-algebras) → (sets) be a functor locally of finite presentation, A an e.f.t. local k-algebra and u¯ ∈ F(Aˆ). Then for every positive integer c there is an element u ∈ F(A˜) such that u¯ ≡ u modulo mc.

Outline of the proof of Theorem 2.5.21 (in a special case, when it follows almost directly from the approximation theorem.) ¯ Let (A, u¯) be an effective formal versal deformation of u0. We will assume that A¯ = Aˆ, where A is an e.f.t. local k-algebra in A∗. Suppose that we can find u˜ ∈ F(A˜) whose associated formal deformation is {¯un}. Then, since ˜ A = lim O , → S s where (S, s) runs through all the etale neighbourhoods of (Spec(A), {m A}), and since F is locally of finite presentation there is an etale neighbourhood (Spec(B), x) of ˜ (Spec(A), {m A}) and u ∈ F(B) such that u →˜u under F(B) → F(A).Again, because F is locally of finite presentation we may assume that Spec(B) is an affine algebraic scheme: therefore (B, x, u) is an algebraization of (A¯, u¯). Therefore we only need to find u˜ as above. By the approximation theorem there exists u˜ ∈ F(A˜) 2 ¯ ˜ 2 such that u¯ ≡˜u mod m . Therefore the homomorphism ψ1 : A → A/m sends u¯ →˜u1 = u1. By versality there is a compatible sequence of homomorphisms ¯ ˜ n+1 ψn : A → A/m lifting ψ1 and such that F(ψn)(u¯) =˜un for all n. We obtain ¯ ˜ ¯ an induced local homomorphism ψ : A → A = A such that F(ψ)(u¯) ≡˜un mod mn+1 for all n. It suffices to prove that ψ is an isomorphism. By construction ψ is the identity mod m2: the conclusion now follows from Lemma C.5. 

NOTES 1. The terminology introduced in Definition 2.5.10 is the most commonly used today but it differs from Grothendieck’s: in [71] he calls a formal deformation “algebraizable” if it is effective. The same terminology is used in [84], p. 195. 2. Other references for Theorem 2.5.13 are [1], ch. III, Th. 5.4.5, and [3], Exp. III, Prop. 7.2. 2.6 Automorphisms and prorepresentability 89 2.6 Automorphisms and prorepresentability

. . . chaque fois que . . . une variet´ e´ de modules . . . ne peut exister, malgre´ de bonnes hypotheses´ de platitude, proprete,´ et non singularite´ eventuellement, la raison en est seulement l’existence d’automorphismes de la structure . . . (Grothendieck [40], p. 94.)

2.6.1 The automorphism functor

The following theorem gives a criterion on an algebraic scheme X to decide whether

DefX , resp. DefX , has a universal formal element and not merely a semiuniversal one.

Theorem 2.6.1. Assume that X is an algebraic scheme such that DefX has a semiu- niversal element (e.g. X affine with isolated singularities or X projective). Then the following conditions are equivalent:

(i) DefX is prorepresentable (ii) for each small extension A → AinA, and for each deformation X of X over Spec(A ), every automorphism of the deformation

X ×Spec(A ) Spec(A) → Spec(A)

is induced by an automorphism of X .

A similar statement holds for the functor DefX .

Proof. (i) ⇒ (ii) Let X = X ×Spec(A ) Spec(A) and let f : X → X be the induced ¯ morphism; assume that θ is an automorphism of X . Letting A = A ×A A , one can construct two deformations Z and W of X over A¯ as we did in the proof of Theorem 2.4.1 as fibred sums fitting into the two diagrams: Z W   X X X X  f θ  f  f  f X X ¯ Since [Z], [W]∈DefX (A) have the same image ([X ], [X ]) under the map ( ¯) → ( ) × ( ) DefX A DefX A DefX (A) DefX A and since this map is bijective by (i), we have an isomorphism of deformations ∼ ρ : Z = W. The isomorphism ρ induces automorphisms θ1 and θ2 of X and an automorphism ψ of X such that:

θ1 f θ = f ψ, θ2 f = f ψ θ = θ −1θ This equality implies f 1 2 f : 90 2 Formal deformation theory

θ−1θ X −→1 2 X ↑ f ↑ f X −→θ X

θ −1θ θ and therefore 1 2 induces . (ii) ⇒ (i) Since DefX has a semiuniversal element, it suffices to show that it satisfies condition H of Theorem 2.3.2. Let A → A be a small extension in A: ¯ letting A = A ×A A we must show that the map α : ( ¯) → ( ) × ( ) DefX A DefX A DefX (A) DefX A is bijective. Given deformations X and X˜ of X over A inducing the deformation X over A, we have the “fibred sum” deformation X¯ over A¯, which fits into the diagram: X¯  X X˜  f  f˜ X and satisfies α([X¯]) = ([X ], [X ]). Suppose that Z is another deformation of X over A¯ such that α([Z]) = ([X ], [X˜ ]). We have isomorphisms of deformations induced by the two projections: X =∼ Z × ( ) =∼ X˜ Spec(A¯) Spec A There remains induced an automorphism θ of X as the composition:

X =∼ X × ( ) =∼ Z × ( ) =∼ X˜ × ( ) =∼ X Spec(A ) Spec A Spec(A¯) Spec A Spec(A ) Spec A and θ fits into the commutative diagram: Z  X X˜  f  f˜ X −→θ X

By (ii) we can lift θ to an automorphism σ : X =∼ X . Replacing the lower left map f by σ f we obtain the commutative diagram Z  X X˜  σ f  f˜ X

By the universal property of the fibred sum we obtain an isomorphism X¯ =∼ Z which is an isomorphism of deformations. Therefore [Z]=[X¯] and α is bijective. 2.6 Automorphisms and prorepresentability 91

 In the case of DefX the proof is similar. When X is a projective scheme, condition (ii) of Theorem 2.6.1 can be stated in a different way by means of the automorphism functor, which we now introduce. Assume that X is an algebraic scheme such that DefX has a semiuniversal couple (R, uˆ). Consider the functor of Artin rings

Autuˆ : AR → ( sets) Autuˆ (A) = the group of automorphisms of the deformation XA where XA is the deformation induced by uˆ under the morphism R → A. Then we have the following: 0 Proposition 2.6.2. If X is projective then Autuˆ has H (X, TX ) as tangent space and is prorepresented by a complete local R-algebra S. Moreover, the deformation func- tor DefX is prorepresentable if and only if S is a formally smooth R-algebra, i.e. if it is a power series ring over R.

Proof. Obviously, Autuˆ satisfies condition H0 because by definition the only auto- morphism of the deformation Xk = X is the identity. Now consider a diagram in AR: A A

A ¯ with A → A a small extension and let A = A ×A A . There is induced a diagram of deformations: X A¯ 

XA XA  XA and therefore a natural homomorphism: O → O × O X ¯ X OX X (2.27) A A A A

Since OX = OX ⊗ ¯ A it follows from Lemma A.4 that (2.27) is an isomorphism; A A¯ A in particular, we obtain an induced isomorphism ∗ ∗ ∗ O =∼ O × ∗ O X ¯ X OX X A A A A and therefore 0 ∗ ∼ 0 ∗ 0 ∗ H (O ) = H (O ) × 0 ∗ H (O ) (2.28) X ¯ X H (OX ) X A A A A

Now note that for every A in ob(AR) the elements of Autuˆ (A) are identified with those elements of H 0(O∗ ) which restrict to 1 ∈ O∗ . Hence we see that (2.28) XA X immediately implies the bijection

( ¯) =∼ ( ) × ( ) Autuˆ A Autuˆ A Autuˆ (A) Autuˆ A

Therefore the functor Autuˆ also satisfies conditions H and H . 92 2 Formal deformation theory

From Lemma 1.2.6 it follows that

∼ 0 Autuˆ (k[ ]) = H (X, TX ) (2.29) which has finite dimension since X is projective, and also H f is satisfied. This con- cludes the proof of the first part. Condition (ii) of Theorem 2.6.1 can be rephrased by saying that the functor Autuˆ is smooth over DefX .  0 H (X, TX ) is usually called the space of infinitesimal automorphisms of X.Asa corollary we obtain:

Corollary 2.6.3. If X is a projective scheme then the following conditions are equiv- alent: 0 (i) h (X, TX ) = 0, i.e. X has no infinitesimal automorphisms; ∼ (ii) Autuˆ = DefX ; (iii) every infinitesimal deformation of X has no nontrivial automorphisms.

Proof. It is an immediate consequence of the proposition. It can be also proved di- 0 rectly without using Proposition 2.6.2: just observe that if H (X, TX ) = 0 then using Lemma 1.2.6 one shows by induction that every infinitesimal deformation of X has no automorphisms.  An important application of the above proposition is the following result, which can be considered as the scheme-theoretic version of a classical theorem due to Kodaira–Nirenberg–Spencer [106]:

0 Corollary 2.6.4. If X is a projective scheme such that h (X, TX ) = 0 then DefX 2 is prorepresentable. If, moreover, X is nonsingular and h (X, TX ) = 0 then DefX is prorepresented by a formal power series ring.

0 Proof. From (2.29) it follows that S = R if h (X, TX ) = 0 and in particular S is a formally smooth R-algebra. Then the first part follows from Proposition 2.6.2. The last assertion is a consequence of Corollary 2.4.7.  0 The condition h (X, TX ) = 0 (no infinitesimal automorphisms) implies that the group Aut(X) is finite. We thus see that the existence of a nontrivial automorphism group does not prevent DefX from being prorepresentable provided there are no in- finitesimal automorphisms. On the other hand, the existence of any automorphisms is a source of difficulties when one considers local deformations (see Subsection 2.6.2). The prorepresentability criterion for DefX given by Proposition 2.6.2 is not easy 0 0 to apply when h (X, TX )>0. Note that the condition h (X, TX ) = 0 is not r necessary for the prorepresentability of DefX . An example is given by X = IP , r ≥ 1: in this case DefX is trivially prorepresentable because X is rigid, but 0 2 h (X, TX ) = (r + 1) − 1 > 0. For another example see the following proposi- tion. Corollary 2.6.4 can be generalized in a straightforward way to conclude that any functor of Artin rings F classifying isomorphism classes of deformations of 2.6 Automorphisms and prorepresentability 93 a scheme with some additional structure or of any other algebro-geometric object  (a morphism, etc.) is prorepresentable provided F has a semiuniversal element and  has no infinitesimal automorphisms. This remark will be applied in the proof of the following:

Proposition 2.6.5. Let X be a projective irreducible and nonsingular curve of genus 1. Then DefX is prorepresentable. Proof. Fix a closed point p ∈ X. For each A in ob(A) we define a deformation of the pointed curve (X, p) to be a pair (ξ, σ ) where X → X ξ : ↓↓π Spec(k) → Spec(A) is an infinitesimal deformation of X over A and σ : Spec(A) → X is a section of π such that Im(σ ) ={p}. We have an obvious definition of isomorphism of two deformations of (X, p) over A, and we define a functor of Artin rings

Def(X,p) → (sets) by Def(X,p)(A) ={isom. classes of deformations of (X, p) over A} We have a morphism of functors:

φ : Def(X,p) → DefX induced by the correspondence (ξ, σ ) → ξ which forgets the section σ. The proposition is a consequence of the following two facts: (a) φ is an isomorphism of functors. (b) Def(X,p) is prorepresentable. To prove (a) let A ∈ ob(A) and consider an infinitesimal deformation ξ of X over A. The point p defines a morphism Spec(k) → X making the following diagram commutative: X p ↓π Spec(k) → Spec(A) By the smoothness of X over Spec(A) there is an extension of p to a section σ : Spec(A) → X of π; this proves that Def(X,p)(A) → DefX (A) is surjective. Now let (ξ, σ ) and (η, τ) be two deformations with section of X over A, where X → Y η : ↓↓q Spec(k) → Spec(A) 94 2 Formal deformation theory and suppose that there is an isomorphism of deformations X

ψ X −→ Y

Spec(A) Then ψσ,τ : Spec(A) → Y are two sections of q. We now use the fact that Y has a structure of group scheme over Spec(A) with identity τ (in outline this can be seen as follows: X is a group scheme with identity p and the group structure is given by a multiplication morphism µ : X × X → X; the group operation on Y is defined by a morphism µA : Y ×A Y → Y which extends µ and which exists because we have a commutative diagram: µ X × X −→ X ⊂ Y ∩↓ Y ×A Y → Spec(A) and Y is smooth over Spec(A)). Replacing ψ by by ψ = (ψσ )−1ψ we obtain an isomorphism of deformations ψ such that ψ σ = τ and therefore ψ defines an isomorphism of (ξ, σ ) with (η, τ); this proves that Def(X,p)(A) → DefX (A) is injective as well, and (a) is proved. In particular, it follows that Def(X,p) has a semiuniversal element because DefX does. Now observe that the vector space of automorphisms of the trivial deformation 0 0 of (X, p) can be identified with the vector subspace of H (X, TX ) = H (X, OX ) consisting of the derivations D : OX → OX vanishing at p, and this is equal to 0 H (X, OX (−p)) = (0). Now the remark following Corollary 2.6.4 applies to con- clude that Def(X,p) is prorepresentable, i.e. (b) holds.  The following corollary computes in particular the number of moduli of projec- tive nonsingular curves. Corollary 2.6.6. If X is a projective nonsingular connected curve of genus g then DefX is prorepresentable. More precisely, DefX = h R where ⎧ ⎨ k if g = 0 = k[[X]] if g = R ⎩ 1 k[[X1,...,X3g−3]] if g ≥ 2 . Proof. X is unobstructed by Proposition 2.4.6 and 0ifg = 0 1 h (X, TX ) = 1ifg = 1 3g − 3ifg ≥ 2

So it remains to be shown that DefX is prorepresentable. In the case g = 0thisis 1 because IP is rigid; in the case g ≥ 2, since deg(TX ) = 2 − 2g < 0wehave 0 H (X, TX ) = 0 and therefore 2.6.4 applies. If g = 1 we use 2.6.5.  2.6 Automorphisms and prorepresentability 95

Recalling Theorems 2.5.13 and 2.5.14 we deduce the following:

Theorem 2.6.7. Let X be a projective nonsingular connected curve. Then X has an algebraic formally universal deformation.

Examples 2.6.8. (i) ([154]) Let C = Spec(B), where B = k[x, y]/(xy),bea reducible affine plane conic. Then DefC is not prorepresentable although C has a semiuniversal deformation by Corollary 2.4.2. In fact, consider the deformation of C over k[ ] given by xy + = 0 and its automorphism: x → x + x y → y

This automorphism does not extend to an automorphism of xy+t¯ = 0 over k[t]/(t3); if it did it would be of the form x → x + xt¯ + at¯2 y → y + bt¯2 for some a, b ∈ k[x, y]. But this implies that bx + ay =−1ink[x, y], which is impossible. From Theorem 2.6.1 we deduce that DefC is not prorepresentable. This holds more generally for the union of the coordinate axes in An, n ≥ 2 (see [154]). (ii) The condition of Corollary 2.6.4 is not satisfied by the rational ruled surfaces ≥ 1( ) = = 1( ) Fm, m 0 (see Example B.11(iii)). Since h TF0 0 h TF1 we find that F0 and F1 are rigid; in particular, DefF0 and DefF1 are prorepresentable. On the other ≥ 2( ) = hand, DefFm is unobstructed when m 2 (since h TFm 0) and has a semiuni- versal element but it is not prorepresentable. To see it we can argue as follows. We 2 1 can identify Fm with the hypersurface m of IP × IP of equation

xmv − ymu = 0 where (u,v,w; x, y) are bihomogeneous coordinates in IP 2×IP 1 (for the elementary proof of this fact see [5], p. 55). For simplicity let’s consider the case m = 2. The linear pencil V ⊂ IP 2 × IP 1 × A1 of equation: x2v − y2u − txyw = 0 1 ∼ defines a flat family V → A such that V(0) = 2 and V(t) = 0 for all t = 0. In fact, we have an isomorphism V\V(0) → V(1) × A1\{0} over A1\{0} sending

(x, y; u,v,w; t) → (x, y; u,v,tw; t); ∼ on the other hand, 0 = V(1) by the isomorphism

(x, y; u,v,w) → (x, y;−x2uw − xyw2, xyu2 + y2vw,x2u2 + 2xyuw + y2w2) 96 2 Formal deformation theory

(V is essentially the family considered in Example 1.2.11(iii) for m = 2). The pull- back 2 2 2 1 V = V (x v − y u − xyw) ⊂ IP × IP × Spec(k[ ]) has the automorphism defined by sending w → w + u and leaving all the other coordinates unchanged. We leave the reader to check that this automorphism does not extend to the pullback of V over Spec(k[t]/(t3)). From Theorem 2.6.1 we deduce that DefF2 is not prorepresentable.

2.6.2 Isotriviality

The notions of triviality and of formal triviality of an algebraic deformation are re- lated in a quite subtle way, as shown by Example 1.2.11(ii). This example is a special case of an important phenomenon, called isotriviality, considered for the first time in the literature in [165].

Definition 2.6.9. Let π : X → S be a flat family of schemes. Then: (i) π is called isotrivial if there is an etale cover (i.e. a finite surjective etale mor-

phism) f : S → S such that the family πS : S ×S X → S is trivial. If ∼ S ×S X = X × S we say that π is isotrivial with fibre X. (ii) If s ∈ Sisak-rational point then π is called locally isotrivial at s if there

is an etale neighbourhood f : (S , s ) → (S, s) such that the pullback πS :

S ×S X → S of π is trivial. π is called locally isotrivial if it is locally isotrivial at every k-rational point of S.

1 Every trivial family is isotrivial. A rational ruled surface π : Fm → IP with m ≥ 2 is locally isotrivial because locally around each point of IP 1 it is the product family with fibre IP 1; on the other hand, π is not isotrivial because it is not trivial and the identity IP 1 → IP 1 is the only connected etale cover of IP 1. In particular, the trivial family IP 1 × IP 1 → IP 1 cannot be obtained by pulling back π by an etale cover of IP 1. Therefore the notions of isotriviality and of local isotriviality are different. If π : X → S is isotrivial then all the fibres over the k-rational points are isomorphic. The next proposition considers the opposite implication.

Proposition 2.6.10. Let π : X → S be a flat family of algebraic schemes, and let s ∈ S be a closed point. If π is locally isotrivial at s then the formal deformation of X (s) associated to π is trivial. If the morphism π is projective then the converse is also true.

Proof. The first part follows immediately from Lemma 2.5.6. Conversely, let X = X (s) and A = OS,s, and assume that π is projective and that the formal deformation (Aˆ, η)ˆ of X associated to the family π is trivial. Let (Aˆ, η)¯ be the deformation of X over Spec(Aˆ) induced by π under the morphism Spec(Aˆ) → S. By Theorem 2.5.11 this deformation is uniquely determined by the formal deformation (Aˆ, η)ˆ and therefore it is trivial. 2.6 Automorphisms and prorepresentability 97

Extend DefX to a functor F locally of finite presentation defined on (k-algebras). We may assume that S = Spec(B) is affine. We have a commutative diagram of functors defined on the category (k-algebras):

h Aˆ

h B h B

F

The lower arrows are induced by the two algebraic deformations π : X → S and p : X × S → S of X (by Yoneda’s lemma), while the upper arrows are the same and correspond to the natural homomorphism B → Aˆ. There remains induced a morphism of functors → × h Aˆ h B F h B which corresponds to an effective formal element for the functor h B ×F h B. Since F is locally of finite presentation and B is of finite type the functor h B ×F h B is locally of finite presentation as well (a fibred product of functors locally of finite presentation is again locally of finite presentation: see [12], p. 33) so that we can apply the approximation theorem 2.5.22. Therefore we can find an etale neighbour- hood Spec(B ) → Spec(B) such that both deformations π and p pull back to the same deformation over Spec(B ); in particular, π pulls back to a trivial deformation over Spec(B ).  The inverse implication of Proposition 2.6.10 is false in general: families of non- singular affine schemes are formally trivial but not isotrivial in general, as shown by Example 1.2.5, page 26. The following is immediate.

Corollary 2.6.11. If π : X → S is a smooth projective family of algebraic schemes which is locally isotrivial at a k-rational point s ∈ S then the Kodaira–Spencer map

1 κπ,s : Ts S → H (X (s), TX (s)) is 0.

Example 2.6.12. The converse of Corollary 2.6.11 is false. For example, consider a smooth projective family π : X → Spec(k[t]) such that the induced deformations over k[t]/(tn) are nontrivial for n  0; in particular, π is not locally isotrivial at the point t = 0. Let q : Spec(k[u]) → Spec(k[t]) be the morphism defined by q : k[t]→k[u] sending t → u2. Then q has zero differential at 0 and therefore the pulled back family

X ×Spec(k[t]) Spec(k[u]) → Spec(k[u]) has a vanishing Kodaira–Spencer map at 0. But the restriction of this family to k[u]/(un) is nontrivial for n  0 and therefore the family is not locally isotrivial at s by Proposition 2.6.10. 98 2 Formal deformation theory

The existence of nontrivial isotrivial deformations of a scheme X is closely re- lated to the existence of nontrivial automorphisms of X. Before investigating this fact we give some examples.

Examples 2.6.13. (i) Let X be a quasi-projective scheme such that there is a finite nontrivial subgroup G ⊂ Aut(X).LetS be a quasi-projective scheme on which G acts freely and let S := S /G be the quotient. Then G acts on X × S componentwise and the action is easily seen to be free. The quotient X := (X × S )/G exists and is an algebraic scheme (see [166] or [3], exp. V). Since the projection X × S → S is G-equivariant it induces a morphism

π : X → S and we have a commutative diagram: X × S → X ↓↓π (2.30) S → S where the horizontal arrows are etale morphisms and all the fibres of π over the closed points are isomorphic to X. Moreover, (2.30) is a cartesian diagram (there is an S -morphism X × S → S ×S X which is easily seen to be an isomorphism) and π is flat (use A.5). Claim: The family π is not trivial. In fact, since (2.30) is cartesian, if π were trivial the action of G on X × S would be trivial on the first factor, a contradiction. It follows that π is an isotrivial nontrivial family. (ii) Let G be a nontrivial finite group scheme and Z a quasi-projective scheme on which G acts freely. Then the quotient scheme Z/G exists (see [166] or [3], exp. V). Let π : Z → Z/G be the canonical morphism. Then π is an etale cover; in particular, it is flat. Moreover, we have a commutative diagram

Z ×Z/G Z → Z ↓ π ↓ π Z → Z/G and since the action is free we have an isomorphism G × Z → Z ×Z/G Z induced by the map (g, z) → (z, gz). Therefore π is the trivial family and π is isotrivial. If Z is integral then π has no sections and it follows that π is not trivial. The morphism π is called a principal G-bundle. (iii) Perhaps the simplest examples of isotrivial nontrivial families are those de- scribed in Example 1.2.11(ii), page 31. We leave to the reader to verify that they are isotrivial.

We need the following well-known lemma ([165], n. 1.5). 2.6 Automorphisms and prorepresentability 99

Lemma 2.6.14. Let f : Z → Y be an etale cover of algebraic schemes. Then there is an etale cover ϕ : P → Z such that the composition

f ϕ : P → Z → Y is a principal G-bundle with respect to a finite group G. In particular, f is isotrivial and, if Z is integral and deg( f )>1, it is nontrivial.

Proof. Let n = deg( f ).Inthen-fold fibre product Z ×Y Z ×Y ···×Y Z consider the set P of points (z1,...,zn) such that zi = z j for all i = j. Then P is a union of connected components of Z ×Y Z ×Y ···×Y Z which is stable under the natural action of the symmetric group Sn. The natural morphism φ : P → Y is an etale ∼ cover of degree n! and therefore it induces an isomorphism P/Sn = Y . Therefore φ is a principal Sn-bundle. Moreover, the first projection ϕ : P → Z is etale of degree (n − 1)! and satisfies f ϕ = φ. In order to prove the last assertion recall that by Example 2.6.13(ii) φ is isotrivial. More precisely, we have a commutative diagram: ∼ P ×Y P = Sn × P → P ↓↓φ P → Y The left vertical morphism is the projection and, because of the factorization f ϕ = φ, it factors as Sn × P → P ×Y Z → P

It follows that P ×Y Z → P is trivial as well and this implies that f is isotrivial. If Z is integral and deg( f )>1 then f cannot be trivial because it has no sections. 

Theorem 2.6.15. The following conditions are equivalent on a quasi-projective scheme X: (a) There exists a nontrivial isotrivial algebraic family with fibre X. (b) Aut(X) contains a nontrivial finite subgroup.

Proof. (b) ⇒ (a) has been proved in Example 2.6.13(i). (a) ⇒ (b). Let π : X → S be a family as in (a). By hypothesis there is an etale cover φ : P → S such that πP : P ×S X → P is trivial with fibre X; let’s identify X × P = P ×S X and let ψ : X × P → X be the projection:

ψ X × P −→ X ↓↓π φ P −→ S

For each p ∈ P denote by ψp : X → X the morphism defined by

ψp(x) = ψ(x, p) 100 2 Formal deformation theory

By Lemma 2.6.14 we may assume that φ is a principal G-bundle with respect to some finite group G. We define an action of G on X × P by the following rule: ( , ) = (ψ−1ψ( , ), ) g x p gp x p gp (X × P)/G exists and, since the action is clearly free and transitive on the fibres of ψ, there is an induced morphism (X × P)/G → X which is an etale cover of degree one, i.e. it is an isomorphism. Fix p ∈ P and define

G × X → X

= ψ−1ψ( , ) by gx gp x p . Then one checks immediately that this is an action of G on X. Since π is nontrivial it is quite clear that this action cannot be trivial for all p ∈ P. Therefore for some p ∈ P the above action defines a homomorphism G → Aut(X) whose image is ={1}.  If a scheme X has an isotrivial local deformation η which is nontrivial then the local moduli functor ∗ DefX : A → (sets) considered in the Introduction cannot be representable, i.e. the local deformation υ considered there cannot exist. Assume by contradiction that υ exists; then, since η is nontrivial it must be pulled back from υ by a nonconstant morphism g : Spec(A) → Spec(O). On the other hand, since η is isotrivial its pullback to Spec(A˜) is trivial and is therefore obtained by pulling back υ in two different ways: by the constant morphism and by the composition

ϕ g Spec(A˜) −→ Spec(A) −→ O which is nonconstant because ϕ is faithfully flat hence surjective; this contradicts the universality of υ. These remarks explain why we cannot expect to be able to construct families representing functors defined on A∗ or on (k-algebras) or on (schemes), which clas- sify isomorphism classes of schemes having nontrivial isotrivial deformations; and the existence of such deformations is closely related to the existence of nontrivial automorphisms of such schemes, as we saw in Theorem 2.6.15. This discussion suggests that while the consideration of isomorphism classes of deformations is not a drawback when one is studying infinitesimal deformations, it becomes inadequate for the classification of algebraic deformations and for global moduli problems. In other words, because of the presence of nontrivial automor- phisms we cannot in general expect to find a scheme structure on the set M of isomorphism classes of objects we want to classify in such a way that it reflects faithfully the functorial properties of families. For example, in the case of projective nonsingular curves of genus 0 one should have M = Spec(k) and the only candidate for being the universal family is IP 1 → Spec(k) because there is only one isomor- 1 phism class of such curves; but the families Fm → IP (Example B.11(iii)) cannot be pulled back from it, thus a universal family cannot exist in this case. 2.6 Automorphisms and prorepresentability 101

That’s why it would be more natural, instead of taking isomorphism classes of deformations, to consider all families together and analyse them and their isomor- phisms. This will result in a more general structure, called a stack, which contains all the information about families and deformations of the objects of the set M we want to classify. We refer to [43], [115] and [185] for foundational material about stacks, and to [18], [46], [53], [65] for expository treatments. 3 Examples of deformation functors

This chapter is devoted to the study of the most important functors of Artin rings which arise when one considers deformations of various algebro-geometric objects. We will verify Schlessinger’s conditions and we will describe first-order deforma- tions, i.e. the tangent spaces of these functors, and obstruction spaces. We will focus on several examples and applications. It will emerge from the treatment that there is a pattern common to almost all examples: the tangent space and the obstruction space of a given functor will be respectively isomorphic to H i and to H i+1 of a certain sheaf which depends on the functor. It will be i = 0 if the deformation problem has no automorphisms, while i = 1 if there are automorphisms; in this case the H 0 will classify the infinitesimal automorphisms.

3.1 Affine schemes

In this section we study the deformation functor of an affine scheme. We already ¯ know that such a functor verifies Schlessinger’s conditions H0, H, H and we com- puted its tangent space (Proposition 2.4.1). In particular, we proved that it has a semi- universal element if the scheme has isolated singularities (Corollary 2.4.2). Here we will analyse this case in more detail. We will start by recalling the description of the tangent space.

3.1.1 First-order deformations

Let B0 be a k-algebra, and let X0 = Spec(B0). We continue the study of infinitesimal deformations of X0, equivalently of B0, started in Section 1.2 in the nonsingular case. Let’s recall the following fact (see 2.4.1(iii)).

Proposition 3.1.1. There is a natural isomorphism

Def (k[ ]) =∼ T 1 B0 B0 104 3 Examples of deformation functors where the class of trivial deformations corresponds to 0 ∈ T 1 .IfB is reduced then B0 0 ( [ ]) =∼ 1(Ω , ) DefB0 k Extk B0/k B0

Proof. A first-order deformation of B0 consists of a flat k[ ]-algebra B,plusa ∼ k-isomorphism B ⊗k[ ] k = B0. This set of data determines a k-extension:

j 0 → B0 −→ B → B0 → 0  (3.1) B obtained after tensoring the exact sequence

0 → ( ) → k[ ]→k → 0 by ⊗k[ ] B. Isomorphic deformations give rise to isomorphic extensions. Conversely, given a k-extension (3.1), B has a structure of k[ ]-algebra given by

→ j(1)

B is k[ ]-flat by Lemma A.9. If B0 is reduced then

1 1 T = Ext (Ω / , B ) B0 k B0 k 0 by Corollary 1.1.11.  The following is an immediate consequence of 3.1.1, 2.2.8 and Schlessinger’s theorem (2.3.2).

Corollary 3.1.2. If dim (T 1 )<∞ then Def has a semiuniversal formal element. k B0 B0 This happens in particular if X = Spec(B ) has isolated singularities. If T 1 = 0 0 0 B0 then B0 is rigid.

We will give some indications for the practical computation of T 1 when B is B0 0 e.f.t.. Let B0 = P/J, where P is a smooth k-algebra of the form

−1 P = ∆ k[X1,...,Xd ] for some multiplicative system ∆ ⊂ k[X1,...,Xd ], and J ⊂ P is an ideal. Consider the exact sequence

∨ δ 2 1 0 → Hom(Ω / , B ) → Hom(Ω / ⊗B , B ) −→ Hom(J/J , B ) −→ T → 0 B0 k 0 P k 0 0 0 B0 The module Hom(ΩP/k ⊗ B0, B0) = Derk(P, B0) 3.1 Affine schemes 105 consists of all derivations D of the form

d ∂p D(p) = b j ∂ X j j=1 for given b j ∈ B0, and

d ∂ ∨ ¯ f δ (D)( f ) = D( f ) = b j f ∈ J ∂ X j j=1

Assume that J = ( f1,..., fn) and let

ι j 0 → R −→ Pn −→ J → 0 be the corresponding presentation. We have the exact sequence:

∨ ∨ → ( / 2, ) −→j ( n, ) −→ι ( , ) 0 Hom J J B0 Hom B0 B0 Hom R B0

∨ ( / 2, ) ( n, ) where j identifies Hom J J B0 with the submodule of Hom B0 B0 consisting ( n, ) = n of those homomorphisms which are 0 on R. Identifying Hom B0 B0 B0 , thereby viewing its elements as column vectors, we see that the condition for ⎛ ⎞ q1 = ⎝ . ⎠ ∈ n q . B0 qn

2 t to be in Hom(J/J , B0) is that q · r = 0 for each r ∈ R (where we are viewing R as consisting of column vectors as well). j∨ associates to a homomorphism

ϕ : / 2 → J J  B0 ¯ ¯ b j f j → b j ϕ( f j ) the column vector ⎛ ¯ ⎞ ϕ( f1) ⎝ . ⎠ . ¯ ϕ( fn) (δ∨) ⊂ n Therefore Im B0 is generated by the column vectors corresponding to ∂ ∂ δ∨( ),..., δ∨( ), i.e. by the classes mod J of: ∂ X1 ∂ Xd ⎛ ⎞ ⎛ ⎞ ∂ f1 ∂ f1 ∂ X ∂ X ⎜ . 1 ⎟ ⎜ . d ⎟ ⎝ . ⎠ , ···, ⎝ . ⎠ ∂ fn ∂ fn ∂ X1 ∂ Xd 106 3 Examples of deformation functors

∨ ∨ If J = ( f1,..., fn) is generated by a regular sequence then ι = 0, equivalently j is an isomorphism, and it follows that n 1 ∼ P T = ⎛⎛ ⎞ ⎛ ⎞⎞ ⊗P B0 B0 ∂ f1 ∂ f1 ∂ ∂ ⎜⎜ X1 ⎟ ⎜ Xd ⎟⎟ ⎜⎜ . ⎟ ⎜ . ⎟⎟ ⎝⎝ . ⎠ , ···, ⎝ . ⎠⎠ ∂ fn ∂ fn ∂ X1 ∂ Xd

In particular, if B0 = P/( f ) then P 1 =∼ TB ∂ ∂ 0 f, f , ···, f ∂ X1 ∂ Xd It follows from this description that the hypersurface V ( f ) is rigid if and only if it is nonsingular. A similar remark holds for complete intersections. In particular, recalling Definition D.1, we can state the following, for future reference:

Proposition 3.1.3. An e.f.t. complete intersection ring B0 such that Spec(B0) is sin- gular is not rigid. Example 3.1.4. Let P be the local ring of a nonsingular algebraic surface X at a k-rational point p, m = (x, y) its maximal ideal, and B0 = P/( f ) the local ring of a curve C ⊂ X at p. Let’s compute T 1 in some cases. B0 ˆ ∼ 2 2 (a) Node (ordinary double point). By definition B0 = k[[X, Y ]]/(X + Y ). Then f = x2 + y2+ higher-order terms, and B T 1 =∼ 0 =∼ k B0 (x, y) if char(k) = 2. ˆ ∼ 2 3 2 3 (b) Ordinary cusp. In this case B0 = k[[X, Y ]]/(X + Y ). Then f = x + y + higher-order terms, and B T 1 =∼ 0 =∼ k2 B0 (y2, x) if char(k) = 2, 3. ˆ ∼ 2 (c) Tacnode. We have in this case B0 = k[[X, Y ]]/(Y (Y + X )) and B T 1 =∼ 0 =∼ k3 B0 (x2 + 2y, xy) if char(k) = 2. Conversely, we have the following result: Proposition 3.1.5. Assume char(k) = 0. Let P be the local ring of a nonsingular algebraic surface X at a k-rational point p, m = (x, y) its maximal ideal, and B = P/( f ) the local ring of a curve C ⊂ Xatp;lett= dim T 1 . Then 0 k B0 3.1 Affine schemes 107

(a) t = 0 if and only if B0 is regular (a DVR). (b) t = 1 if and only if B0 is the local ring of a node. (c) t = 2 if and only if B0 is the local ring of an ordinary cusp. (d) t = 3 if and only if B0 is the local ring of a tacnode.

Proof. The “if” implication follows from the above computations. We have

t = dimk P/( f, fx , fy)

∈ 3 ≥ ∈ 2 f m P immediately implies t 4; then f m P and, after suitable choice of 2 n generators of m P we may suppose f = y + x + higher-order terms, n ≥ 2or f = y(y + xn)+ higher-order terms, n ≥ 2. Now the conclusion follows easily. 

Example 3.1.6. (The affine cone over IP1 × IP 2). Let

P = k[X0, X1, X2, Y0, Y1, Y2]

J = (X1Y2 − X2Y1, X2Y0 − X0Y2, X0Y1 − X1Y0)

Then B0 = P/J is the coordinate ring of the affine cone over the Segre embedding IP 1 × IP 2 ⊂ IP 5. We have the following presentation:

A 0 → P2 −→ P3 → J → 0 where X0 Y0 A = X1 Y1 X2 Y2 2 A direct computation shows that Hom(J/J , B0) is generated by the following col- umn vectors: Y1 Y2 0 X1 X2 0 −Y0 0 Y2 −X0 0 X2 0 −Y0 −Y1 0 −X0 −X1 Since these vectors are, up to permutation, ∂ ∂ ∂ ∂ ∂ ∂ δ∨ ,δ∨ ,δ∨ ,δ∨ ,δ∨ ,δ∨ ∂ X0 ∂ X1 ∂ X2 ∂Y0 ∂Y1 ∂Y2 we see that T 1 = 0. This implies that B is rigid (see Corollary 3.1.2). B0 0 More generally, one can prove that the coordinate ring of the affine cone over the Segre embedding IPn × IP m ⊂ IP (n+1)(m+1)−1 is rigid whenever n + m ≥ 3. This has been computed for the first time in [67] in the case n = m ≥ 2; the general case is in [157] (see Corollary 3.1.20 below).

Example 3.1.7. Let P = k[X1, X2, X3](X), J = (X2 X3, X1 X3, X1 X2). Then B0 = P/J is the local ring at the origin of the union of the coordinate axes in A3.Wehave the presentation 108 3 Examples of deformation functors

A P3 −→ P3 → J → 0 where X1 X1 0 A = −X2 0 X2 0 −X3 −X3 and the columns of A generate R. 2 Hom(J/J , B0) is generated by the following column vectors mod J:

X2 X3 0000 00X1 X3 00 0000X1 X2 and Im(δ∨) is generated by the column vectors mod J:

0 X3 X2 X3 0 X1 X2 X1 0 It follows at once that

∨ T 1 = Hom(J/J 2, B )/Im(δ ) =∼ k3 B0 0

2 because there are three generators of Hom(J/J , B0) which are linearly independent ∨ 2 modulo the generators of Im(δ ), and all other elements of Hom(J/J , B0) are in Im(δ∨). In a similar vein one can consider, for any d ≥ 3

B0 = k[X1,...,Xd ](X)/J where ˆ J = (...,X1 X2 · Xi · Xd ,...)i=1,...,d d Then B0 is the local ring at the origin of the union of the coordinate axes in A .One computes easily, along the same lines of the case d = 3, that

T 1 =∼ kd(d−2) B0 Example 3.1.8. Let P be the local ring of a nonsingular algebraic variety V of di- mension n ≥ 2atak-rational point p and let B0 = P/( f ) be the local ring of a hypersurface X ⊂ V at p. We call p a node for X if

ˆ =∼ [[ ,..., ]]/( 2 +···+ 2) B0 k X1 Xn X1 Xn ,..., equivalently, if we can choose generators x1 xn of the maximal ideal m P so that = 2+ f xi higher-order terms. It can be proved immediately that if p is a node for X and char(k) = 2 then T 1 =∼ k. In particular, T 1 =∼ k. B0 k[ ] 3.1 Affine schemes 109

3.1.2 The second cotangent module and obstructions

We will now describe an obstruction space for the functor DefB0 . Assume that B0 = P/J for a smooth k-algebra P, and an ideal J ⊂ P. Consider a presentation: ι j η : 0 → R −→ F −→ J → 0 (3.2) where F is a finitely generated free P-module. Let λ : 2 F → F be defined by:

λ(x ∧ y) = ( jx)y − ( jy)x and Rtr = Im(λ). Obviously Rtr ⊂ R and Rtr ⊂ JF. n If J = ( f1,..., fn) then F = P and R is the module of relations among tr f1,..., fn. R is called the module of trivial (or Koszul) relations; it is generated by the relations of the form

(0,...,− f j ,..., fi , ..., 0)

ij It follows that tr R/R = H1(K•( f1,..., fn)) the first homology module of the Koszul complex associated to f1,..., fn.

tr Lemma 3.1.9. The P-module R/R is annihilated by J and therefore it is a B0- module in a natural way.

Proof. Let x ∈ R, a ∈ J.Lety ∈ F be such that j(y) = a. Then

ax = j(y)x = j(y)x − j(x)y = λ(y ∧ x) ∈ Rtr 

tr Since R ⊂ JFthe presentation (3.2) induces an exact sequence of B0-modules:

¯ tr ι¯ j 2 R/R −→ F ⊗P B0 −→ J/J → 0 (3.3)

Definition 3.1.10. The second cotangent module of B is the B -module T 2 defined 0 0 B0 by the exact sequence:

Hom (J/J 2, B ) → Hom (F ⊗ B , B ) → Hom (R/Rtr, B ) → T 2 → 0 B0 0 B0 P 0 0 B0 0 B0 induced by (3.3). Obviously T 2 is a B -module of finite type. B0 0

Lemma 3.1.11. For every e.f.t. k-algebra B the B -module T 2 is independent of 0 0 B0 the presentation (3.2). 110 3 Examples of deformation functors

Proof. Assume that F =∼ Pn and that j : Pn → J is defined by the system of generators f1,..., fn of J.Let

0 → S → Pm → J → 0 be another presentation of J, defined by the system of generators g1,...,gm.We may assume that m ≥ n and that gk = fk, k = 1,...,n.Let gs = bks fk s = n + 1,...,m k n m for some bks ∈ P. Denote by α : P → P the map

α(a1,...,an) = (a1,...,an, 0,...,0) and by β : Pm → Pn the map ⎛ ⎞ m m ⎝ ⎠ β(a1,...,am) = a1 + b1sas,...,an + bnsas s=n+1 s=n+1

Evidently, α(R) ⊂ S and α(Rtr) ⊂ Str. It is easy to verify that β(S) ⊂ R and β(Str) ⊂ Rtr.Itfollowsthatα and β induce homomorphisms  tr tr β : Hom(R/R , B0) → Hom(S/S , B0)  tr tr α : Hom(S/S , B0) → Hom(R/R , B0) whence homomorphisms ˜ tr n tr m β : Hom(R/R , B0)/Hom(P , B0) → Hom(S/S , B0)/Hom(P , B0) tr m tr n α˜ : Hom(S/S , B0)/Hom(P , B0) → Hom(R/R , B0)/Hom(P , B0) Since   tr α β = identity of Hom(R/R , B0) it follows that

˜ tr n α˜ β = identity of Hom(R/R , B0)/Hom(P , B0)

We now prove that

˜ tr m βα˜ = identity of Hom(S/S , B0)/Hom(P , B0) (3.4)

tr Let g ∈ Hom(S/S , B0) be induced by the homomorphism G : S → B0. Then, if tr (a1,...,am) ∈ S and (a¯1,...,a¯m) ∈ S/S is its class, we have:   (β α )(g)(a¯1,...,a¯m) = G(α(β(a1,...,am))) =   ( + m ,..., + m , ,..., ) = G a1 s=n+1 b1sas an s=n+1 bnsas 0 0   ( ,..., ) + ( m ,..., m , − ,...,− ) = (∗) G a1 am G s=n+1 b1sas s=n+1 bnsas an+1 am 3.1 Affine schemes 111

Now note that ⎛ ⎞ m m ⎝ ⎠ b1s ps,..., bns ps, −pn+1,...,−pm ∈ S s=n+1 s=n+1

m for every (p1,...,pm) ∈ P . Therefore letting ⎛ ⎞ m m ⎝ ⎠ τ(p1,...,pm) = G b1s ps,..., bns ps, −pn+1,...,−pm s=n+1 s=n+1

m we define a homomorphism τ : P → B0. It follows that

(∗) = g(a¯1,...,a¯m) + τ(a1,...,am)

Hence   m tr (β α )(g) − g ∈ Im[Hom(P , B0) → Hom(S/S , B0)] or equivalently (3.4) holds.  From the definition it easily follows that T 2 localizes. Namely, for every multi- B0 plicative subset ∆ ⊂ P we have: ∆−1 2 = 2 T T −1 B0 ∆ B0 It follows that for any scheme X we can define in an obvious way the second cotan- 2 gent sheaf which we will denote by TX . It satisfies

T 2 = T 2 X,x OX,x

Proposition 3.1.12. Assume that B = P/J for a smooth k-algebra P. Then T 2 is 0 B0 an obstruction space for the functor DefB0 . Proof. Let A be an object of A and let

B → B0 ξ : ↑↑ A → k be an infinitesimal deformation of B0 over A. We must associate to ξ a k-linear map

2 ξv : Ex (A, k) → T k B0 satisfying the conditions of Definition 2.2.9. Let B0 = P/J for a smooth k-algebra P and an ideal J = ( f1,..., fn) ⊂ P. We have an exact sequence:

f 0 → R → Pn −→ J → 0 112 3 Examples of deformation functors

Then, by the smoothness, in particular flatness, of P we have

B = (P ⊗k A)/(F1,...,Fn) where f j = Fj (mod m A), j = 1,...,n, by Theorem A.10. The flatness of B over = ( ,..., ) ∈ ,..., ∈ ⊗ A implies that for every r r1 rn Rthere exist R1 Rn P k A such = = ,..., = that r j R j (mod. m A), j 1 n, and j R j Fj 0, again by Theorem A.10. Let [γ ]∈Exk(A, k) be represented by an extension

φ γ : 0 → (t) → A˜ −→ A → 0 ˜ ˜ ˜ ˜ ˜ Choose F1,...,Fn, R1,...,Rn ∈ P ⊗k A liftings of F1,...,Fn, R1,...,Rn re- spectively; then ˜ ˜ ˜ ∼ R j Fj ∈ ker[P ⊗k A → P ⊗k A] = P j

˜ ,..., ˜ ˜ ,..., ˜ It is easy to check that a different choice of R1 Rn or of F1 Fn modifies ˜ ˜ ∈ j R j Fj by an element of J or by one of the form j q j r j , where q j P, respec- tively. Therefore sending ˜ ˜ r → R j Fj (3.5) j defines an element of [ ( n, ) → ( , )] coker Hom B0 B0 Hom R B0 = = ( ,..., ,...,− ,..., ) Moreover, since if r rij 0 f j fi 0 we can take ˜ ˜ ˜ ˜ (R1,...,Rn) = (0,...,Fj ,...,−Fi ,...,0)  ˜ ˜ = ( tr, ) and we get j R j Fj 0, it follows that (3.5) is zero on Hom R B0 . Therefore ˜ ˜ ˜ the n-tuple of liftings (F) = (F1,...,Fn) defines an element ξv (γ ) of

coker[Hom(Bn, B ) → Hom(R/Rtr, B )]=T 2 0 0 0 B0

Let’s prove that the map γ → ξv (γ ) is k-linear. Let [ζ ]∈Exk(A, k) be another element defined by the extension:

ζ : 0 → (t) → A → A → 0 ( ) = ( ,..., ) ∈ ⊗ and let F F1 Fn , Fj P k A be the corresponding lifting, which defines ξv (ζ ). Then ξv (γ ) + ξv (ζ ) is defined by → ˜ ˜ + r R j Fj R j Fj j j ,..., ∈ ⊗ ,..., where R1 Rn P k A are liftings of R1 Rn. Consider the diagram: 3.1 Affine schemes 113 ˜ 0 → k ⊕ k → A ×A A → A → 0 ↓ δ ↓ σ  γ + ζ : 0 → k → C → A → 0

Then ξv (γ + ζ)is defined by r → # j Φ j j where #1,...,#n,Φ1,...,Φn ∈ P ⊗k C are liftings of R1,...,Rn, F1,...,Fn. Since ˜ ∼ ˜ P ⊗k (A ×A A ) = (P ⊗k A) ×A (P ⊗k A ) ˜ letting ρ : P ⊗k (A ×A A ) → P ⊗k C be the homomorphism induced by σ ,we may assume that Φ = ρ( ˜ , ); # = ρ( ˜ , ) j Fj Fj j R j R j Then:    # Φ = ρ( ˜ , )ρ( ˜ , ) = ρ ( ˜ , )( ˜ , ) j j j j R j R j Fj Fj j R j R j Fj Fj     = ρ ˜ ˜ , = δ ˜ ˜ , j R j Fj j R j Fj j R j Fj j R j Fj   = ˜ ˜ + j R j Fj j R j Fj

This proves that ξv (γ +ζ) = ξv (γ )+ξv (ζ ). A similar argument shows that ξv (λγ ) = λξv (γ ), λ ∈ k. Now assume that [γ ]∈Exk(A, k) is such that there exists an infinitesimal de- formation ˜ B → B0 ξ˜ : ↑↑ A˜ → k such that ( ˜) → ( ) DefB0 A DefB0 A ξ˜ → ξ ˜ ∈ ⊗ ˜ ∈ It follows that there exist liftings Fj P k A of the Fj ’s such that every r R has ˜ ∈ ( ⊗ ˜)N ˜ ˜ = ξ (γ ) = a lifting R P k A such that j R j Fj 0. This means that v 0. ˜ ˜ ˜ Conversely, assume that ξv (γ ) = 0, and let F1,...,Fn ∈ P ⊗k A be arbitrary n liftings of F1,...,Fn. Then there exists (h1,...,hn) ∈ P such that for every choice ˜ ˜ n of a lifting R ∈ (P ⊗k A) of a relation r ∈ R we have: ˜ ˜ ˜ R j Fj =−t r j h j =− R j h j j j j ˜ ˜ ˜ This means that the ideal (F1 + th1,...,Fn + thn) ⊂ P ⊗k A defines a flat defor- ˜ mation of B0 over A lifting the deformation B = (P ⊗k A)/(F1,...,Fn). 114 3 Examples of deformation functors

Any other choice of a lifting of the deformation ξ over A˜ is of the form ˜ ˜ (F1 + t(h1 + k1), . . . , Fn + t(hn + kn))  = ( ,..., ) ∈ n = ∈ where k k1 kn B0 satisfy j r j k j 0 for every relation r R. There- 2 ∨ fore k ∈ Hom(J/J , B0). It is straightforward to verify that if k ∈ Im(δ ) then F˜ + th and F˜ + t(h + k) define isomorphic liftings of ξ over A˜. This means that we have an action of T 1 on the set of liftings of ξ over A˜. By construction it follows B0 that this action is transitive. 

Corollary 3.1.13. Let X0 = Spec(B0) be an affine algebraic scheme such that dim (T 1 )<∞, and let (R, {η }) be a semiuniversal formal deformation of X . k B0 n 0 Then (i) dim (T 1 ) ≥ dim(R) ≥ dim (T 1 ) − dim (T 2 ) k B0 k B0 k B0

The first equality holds if and only if X0 is unobstructed. (ii) If B0 is an e.f.t. local complete intersection k-algebra then it is unobstructed. In particular, hypersurface singularities are unobstructed.

Proof. (i) follows from Corollary 2.2.11. (ii) If J is generated by a regular sequence then R = Rtr and therefore T 2 = (0).  B0

Proposition 3.1.14. Assume that B0 = P/J for a smooth k-algebra P. Then:

(i) If Spec(B0) is reduced then

T 2 =∼ Ext1 (J/J 2, B ) B0 B0 0

(ii) If Spec(B0) is reduced and has depth at least 2 along the locus where it is not an l.c.i. (e.g. Spec(B0) is normal of dimension ≥ 2) there is an isomorphism

2 ∼ 2 T = Ext (Ω / , B ) B0 B0 k 0 Proof. We keep the notations introduced on page 109. (i) From the exact sequence (3.3) we deduce the following commutative diagram with exact rows and columns:

Hom(ker(ι),¯ B0) ↑ ¯∨ 2 j tr 2 Hom(J/J , B0) → Hom(F ⊗ B0, B0) −→ Hom(R/R , B0) → T → 0  ∪∪ 2 1 Hom(J/J , B0) → Hom(F ⊗ B0, B0) → Hom(Im(ι),¯ B0) → E → 0 where E1 = Ext1(J/J 2, B ) and T 2 = T 2 . Since the exact sequence 0 B0

ι j η : 0 → R −→ F −→ J → 0 3.1 Affine schemes 115 from which (3.3) is obtained localizes, we see that ker(ι)¯ is supported on the lo- cus where B0 is not an l.c.i.; in particular ker(ι)¯ is torsion. Therefore we have Hom(ker(ι),¯ B0) = (0) and the conclusion follows. (ii) Consider the conormal sequence

/ 2 −→δ Ω ⊗ → Ω → J J P/k B0 B0/k 0

Since Spec(B0) is reduced, ker(δ) is supported in the locus where Spec(B0) ⊂ Spec(P) is not a regular embedding (Proposition D.1.4), and this locus coincides with the locus where Spec(B0) is not an l.c.i. (Proposition D.2.5). From the assumption about the depth of Spec(B0) it follows that

1 Hom(ker(δ), B0) = Ext (ker(δ), B0) = (0)

i Using this fact and recalling that Ext (ΩP/k ⊗ B0, B0) = (0), i > 0, we obtain:

2(Ω , ) =∼ 1( (δ), ) =∼ 1( / 2, )  Ext B0/k B0 Ext Im B0 Ext J J B0

Example 3.1.15 (Schlessinger [154]). (An obstructed affine curve.) Let s be an in- 7 8 9 10 determinate and let B0 = k[s , s , s , s ]⊂k[s] be the coordinate ring of the affine rational curve C ⊂ A4 = Spec(k[x, y, z,w]) having parametric equations:

x = s7, y = s8, z = s9,w= s10

Write P = k[x, y, z,w] and

B0 = P/I for an ideal I ⊂ P. One can check, using, for example, a computer algebra package, that I is generated by the six 2 × 2 minors of the following matrix: xyzw2 yzw x3 i.e. by the following polynomials:

2 2 f1 = y − xz; f2 = xw − yz; f3 = z − yw;

4 2 3 2 3 3 f4 = x − w y; f5 = x y − zw ; f6 = w − x z

This ideal is prime of height 3. Consider a presentation:

0 → R → F → I → 0

∼ 6 where F = P with generators, say e1,...,e6, so that ei → fi . To describe R one can use the beginning of the free resolution of I given by the Eagon–Northcott 116 3 Examples of deformation functors complex (see [48]). One obtains a set of generators for R given by the rows of the following matrix: R1 : zyx000 2 R2 : w 00y −x 0 2 R3 : 00w 0 zy 2 R4 : 0 −w 0 z 0 x R5 : w zy000 3 R6 : x 00z −y 0 3 R7 : 00x 0 w z R : 0 −x3 0 w 0 y 8  ∈ Here each row gives the coefficients ai of the linear combination i ai ei R.We then have an exact sequence

0 → R¯ → F¯ → I/I 2 → 0 where F¯ = F/IF and R¯ = R/(IF ∩ R). Reducing mod I the above relations one gets the following set of generators of R¯ as elements of F¯: 9 8 7 r1 : s s s 000 20 8 7 r2 : s 00s −s 0 20 9 8 r3 : 00s 0 s s 20 9 7 r4 : 0 −s 0 s 0 s 10 9 8 (3.6) r5 : s s s 000 21 9 8 r6 : s 00s −s 0 21 10 9 r7 : 00s 0 s s 21 10 8 r8 : 0 −s 0 s 0 s

Since B0 is reduced we have

T 2 = Ext1 (I/I 2, B ) = Hom(R¯ , B )/Hom(F¯, B ) B0 B0 0 0 0

¯ 8 Representing an element h ∈ Hom(R, B0) as the 8-tuple (h(r1),...,h(r8)) ∈ B ¯ 0 we see that the submodule Hom(F, B0) is generated by the columns of (3.6). There- fore to prove that B0 is obstructed it will suffice to produce a first-order deforma- 3 tion ξ ∈ DefB (k[ ]) whose obstruction to lifting to k[t]/(t ) is represented by an ¯ 0 h : R → B0 not in the submodule generated by the columns of (3.6). We define ξ by the ideal ( f1 + ∆f1,..., f6 + ∆f6) ⊂ k[ , x, y, z,w] where 2 3 ∆f := (∆f1,...,∆f6) = (0, 0, 0, zw, w , −x ) this defines a deformation because R j · ∆f ∈ I for all j = 1,...,8. More precisely:

(R1 · ∆f,...,R8 · ∆f ) = (0, −w f2, − f5, f4 − wf3, 0,wf3, f6, − f5)

Therefore we see that the obstruction to lifting ξ to second order is defined by the ¯ homomorphism h : R → B0 represented by 3.1 Affine schemes 117

(0, 0, −t20, t19, 0, 0, −t21, −t20) ¯ Now we can immediately check that this vector is not in Hom(F, B0) and therefore the deformation ξ cannot be lifted: thus B0 is obstructed.

3.1.3 Comparison with deformations of the nonsingular locus

Under certain conditions it is possible to compare the deformations of an affine scheme with the deformations of the open subscheme of its nonsingular points. In this and the following subsections we will describe the analysis made in [156] and [157], with applications to the study of certain quotient singularities. We will need a preliminary lemma.

Lemma 3.1.16. Let X be an affine scheme, Z ⊂ X a closed subscheme and ∨ G a coherent sheaf on X. Let G = Hom(G, OX ). If depthZ (OX ) ≥ 2 then ∨ depthZ (G ) ≥ 2 and therefore

H 0(X, G∨) =∼ H 0(X\Z, G∨)

Proof. Consider a presentation

0 → R → F → G → 0 where F is a free OX -module. Then we obtain an exact sequence ∨ ∨ 0 → G → F → Q → 0 (3.7) ⊂ ∨ ∨ ( ∨) = (O ) ≥ where Q R . Since F is free we have depthZ F depthZ X 2 and 0( ∨) = = 1( ∨) therefore HZ F 0 HZ F ([75], Theorem 3.8, p. 44); it follows that 0( ∨) = 0( ∨) = 0( ) = HZ G 0. Similarly, one proves that HZ R 0, and therefore HZ Q 0. From the sequence of local cohomology associated to (3.7) we obtain H 1(G∨) = 0 ∨ Z and therefore depthZ (G ) ≥ 2 by [75], Theorem 3.8, p. 44. The last assertion fol- lows from the exact sequence → 0( , ∨) → 0( \ , ∨) → 1 ( ∨) 0 H X G H X Z G HX G (see [84], p. 212).  Consider an affine scheme X = Spec(B) where B = P/J for a smooth k-algebra P.LetZ = Sing(X) be the singular locus of X and U = X\Z. Let Y = Spec(P) and consider the exact sequence

→ → → → 1 → 0 TX TY |X NX TX 0 (3.8)

= 1 where NX NX/Y . Since TX is supported on Z, by restricting to U we get the exact sequence:

0 → TU → TY |U → NX|U → 0 (3.9) 118 3 Examples of deformation functors

Proposition 3.1.17. (i) If depthZ (OX ) ≥ 2 (e.g. X is normal of dimension ≥ 2)we have an exact sequence → 1 → 1( , ) → 1( , ) 0 TB H U TU H U TY |U

(ii) If depthZ (OX ) ≥ 3 then 1 =∼ 1( , ) TB H U TU Proof. (i) We have the local cohomology exact sequences (see [84], p. 212):

→ 0( , ) → 0( , ) → 1( ) 0 H X NX H U NX|U HZ NX

→ 0( , ) → 0( , ) → 1( ) 0 H X TY |X H U TY |U HZ TY |X

If depthZ (OX ) ≥ 2 then from Lemma 3.1.16 we deduce that depthZ (NX ) ≥ 2 and ( ) ≥ 1( ) = = 1( ) depthZ TY |X 2. Therefore we have HZ NX 0 HZ TY |X ([75], Theorem 3.8, p. 44) and

0 ∼ 0 0 ∼ 0 H (X, NX ) = H (U, NX|U ), H (X, TY |X ) = H (U, TY |U )

Comparing the exact cohomology sequences of (3.8) and (3.9) we get an exact and commutative diagram:

0( , ) → 0( , ) → 1 → H X TY |X H X NX TB 0 ∩ 0 0 1 1 H (U, TY |U ) → H (U, NX|U ) → H (U, TU ) → H (U, TY |U ) which proves (i). (O ) ≥ 1( ) = = 2( ) If depthZ X 3 then X is normal and HZ TY |X 0 HZ TY |X because TY |X is locally free; from the local cohomology exact sequence we get

1 ∼ 1 H (U, TY |U ) = H (X, TY |X ) = 0 because X is affine. Using (i) we deduce (ii).  The above proposition can be applied to prove the rigidity of a large class of cones over projective varieties. We will need the following well-known lemmas, which we include for the reader’s convenience.

Lemma 3.1.18. Let W ⊂ IP r be a projective nonsingular variety, CW the affine cone over W, v ∈ CW the vertex, U = CW\{v} and p : U → W the projection. If ∗ G is a coherent sheaf on CW such that G|U = p F for some coherent F = (0) on W, then the following conditions are equivalent:

(i) depthv (G) ≥ dforsomed≥ 2; 0 0 k (ii) H (CW, G) =⊕ν∈ZZ H (W, F(ν)) and H (W, F(ν)) = 0 for all 1 ≤ k ≤ d−2 and ν ∈ ZZ. 3.1 Affine schemes 119

Proof. We will use the equivalence

k depthv (G) ≥ d ⇔ Hv (G) = 0, k < d

([75], Theorem 3.8, p. 44). We have an exact local cohomology sequence:

0 0 0 1 0 → Hv (G) → H (CW, G) → H (U, G|U ) → Hv (G) → 0 and isomorphisms:

k−1 ∼ k H (U, G|U ) = Hv (G), k ≥ 2

∗ 0 Since G|U = p F with F = (0) we have depthv (G) ≥ 1, thus Hv (G) = 0. On the ∗ other hand, since p∗G|U = p∗ p F =⊕ν∈ZZ F(ν),wehave 0 0 H (U, G|U ) =⊕ν∈ZZ H (W, F(ν)). Now the conclusion follows. 

Lemma 3.1.19. Let 0 =[0,...,0, 1]∈IP r+1,V = IP r+1\{0} and let π : V → IP r be the projection. Then ∗ TV/IP r = π O(1)

Proof. It is an immediate consequence of the commutative exact diagram

00 ↑↑ ∗ 0 → TV/IP r → TV → π TIP r → 0 ↑ ↑ r+2 r+1 0 → OV (1) → OV (1) → OV (1) → 0 ↑↑ OO ↑↑ 00 where the vertical sequences are restrictions of Euler sequences. 

Corollary 3.1.20. Let W ⊂ IP r be a projective nonsingular variety of dimension ≥ 2. Assume that 0 r 0 (i) H (IP , O(ν)) → H (W, OW (ν)) surjective for all ν ∈ ZZ (W is projectively normal). 1 (ii) H (W, OW (ν)) = 0 for all ν ∈ ZZ. 1 (iii) H (W, TW (ν)) = 0 for all ν ∈ ZZ. Then the affine cone CW over W is rigid.

Proof. CW has dimension ≥ 3 and hypothesis (i) implies that it is normal ([84], p. 126). Hypotheses (i) and (ii) imply that depthv (OCW) ≥ 3, by Lemma 3.1.18. 120 3 Examples of deformation functors

1 Therefore by 3.1.17(ii) it suffices to show that H (U, TU ) = 0 where U = CW\{v}. Let p : U → W be the projection. We have 1 1 H (U, OU ) =⊕ν∈ZZ H (W, OW (ν)) = 0 (3.10) 1 ∗ 1 H (U, p TW ) =⊕ν∈ZZ H (W, TW (ν)) = 0 by conditions (ii) and (iii). The relative tangent sequence of p takes the following form: ∗ 0 → OU → TU → p TW → 0 (3.11) In fact it follows from Lemma 3.1.19 and from Proposition B.1(i) that we have = π ∗O ( ) = π −1( )\ = Tπ−1(W)/W W 1 and therefore, since U W W,wehaveTU/W ∗ p OW (1) and this is clearly equal to OU . The conclusion follows from (3.10) and from the cohomology sequence of (3.11).  Corollary 3.1.21. (i) The affine cone over IPn × IP m in its Segre embedding is rigid for every positive n, m such that n + m ≥ 3. (ii) The affine cone over any Veronese embedding of IPn,n≥ 2, is rigid. (iii) If W ⊂ IP r is a projective nonsingular variety of dimension ≥ 2, such that 1 1 h (W, OW ) = 0 = h (W, TW ) then the affine cone over the m-th Veronese embedding W (m) of W is rigid for every m  0. Proof. Using 3.1.20, (i) and (ii) are easy computations and (iii) follows from Serre’s vanishing theorem ([84], Th. III.5.2).  The affine cone over the quadric IP 1 × IP 1 ⊂ IP 3 is not rigid because it is a hypersurface (see Proposition 3.1.3, page 106); it does not satisfy both conditions (ii) r and (iii) of Corollary 3.1.20. The affine cone over a rational normal curve Γr ⊂ IP , r ≥ 2, is not rigid either (for r = 2 it is a singular quadric surface in A3;forr ≥ 3 see [131], [143], [144]). For a generalization of Corollary 3.1.20 and applications see [177].

3.1.4 Quotient singularities

The analysis of the previous subsection can be applied to the study of deformations of a class of affine singular schemes obtained as quotients of nonsingular ones by the action of a finite group. Let Y = Spec(P) be an affine nonsingular algebraic variety on which a finite group G acts. Let X = Y/G be the quotient variety, q : Y → X the projection. Assume that the action is free outside a G-invariant closed subscheme W ⊂ Y . Set V = Y \W.

Proposition 3.1.22. Assume that depthW (OY ) ≥ 2. Then

∼ G TX = (q∗TY )

∼ g −1 where G acts on TY = Derk(P, P) by D → D := gDg for all D ∈ Derk(P, P) and g ∈ G. 3.1 Affine schemes 121

Proof. Consider the exact sequence of coherent sheaves on Y :

→ → → ∗ → 1 0 TY/ X TY q TX TY/ X

Ω1 = Y/ X is supported on W since q is etale outside W; then we have TY/ X 0. Simi- 1 larly, TY/ X is supported on W so that from the above exact sequence restricted to V we deduce an isomorphism

0 ∼ 0 ∗ H (V, TY ) = H (V, q TX )

Then by Lemma 3.1.16 we deduce that

0 ∼ 0 ∗ H (Y, TY ) = H (Y, q TX )

Note that, letting A = PG the ring of invariant elements, we have X = Spec(A) and the above isomorphism is equivalent to an isomorphism ∼ Derk(P, P) = Derk(A, P)

Therefore it will suffice to show that

∼ G Derk(A, A) = Derk(A, P)

−1 So let D ∈ Derk(A, P) be such that D = gDg for all g ∈ G. Then for every a ∈ A we have D(a) = g(D(g−1a)) = g(D(a)) G so D(a) ∈ A and therefore Derk(A, P) ⊂ Derk(A, A). Conversely, if D ∈ Derk(A, A) then it defines a k-derivation of A in P which is clearly G-invariant G and we also have Derk(A, A) ⊂ Derk(A, P) .  Corollary 3.1.23. Let n be the order of G. Under the assumptions of 3.1.22, if char(k) does not divide n then TX is a direct summand of q∗TY . G Proof. Define a homomorphism q∗TY → TX = (q∗TY ) by 1 D → Dg n g∈G

This defines the splitting. 

Theorem 3.1.24. In the above situation, if the action is free outside a G-invariant closed subscheme W of codimension ≥ 3, and char(k) does not divide the order of G then X = Y/G is rigid.

Proof. Let Z = q(W) where q : Y → X is the projection, V = Y \W, = \ = / (O ) ≥ U X Z V G. We have depthZ X 2 because X is normal, being the 122 3 Examples of deformation functors quotient of a nonsingular variety by a finite group (for this elementary fact see e.g. [166], p. 58). Therefore 1 ⊂ 1( , ) =∼ 1( ,( )G ) TX H U TU H U q∗TY (3.12) where the inclusion follows from 3.1.17(i) and the isomorphism is Proposition 3.1.22. We also have an exact sequence 1( , ) → 1( , ) → 2 ( ) H Y TY H V TY HW TY  1 H (U, q∗TY ) where the left vector space is 0 because Y is affine and the right one is 0 because 1 of the depth assumption on W.ItfollowsthatH (U, q∗TY ) = 0 and therefore 1 G H (U,(q∗TY ) ) = 0 as well because it is a direct summand of it by Corollary 3.1.23. The conclusion now follows from (3.12).  In the theorem the hypothesis on the codimension of W cannot be removed. In fact all rational two-dimensional double points are quotient singularities and they are hypersurfaces, therefore they are not rigid (see [15]). Example 3.1.25 (Kollar [110]). Let n ≥ 3. Consider the weighted projective space

X = IP (1, 1, 1, a3,...,an) = Proj(k[X0,...,Xn]) where the Xi ’s are indeterminates with weights ai = 1ifi = 0, 1, 2 and ai ≥ 1if i = 3,...,n (see [45] for details on weighted projective spaces). Then X is locally the quotient of an affine space by a finite cyclic group and by the choice of the weights all its singularities have codimension ≥ 3. It follows from Theorem 3.1.24 1 = 1( , ) that TX 0. Therefore the first-order deformations of X are classified by H X TX by the exact sequence (2.17). But the exact sequence on X: 0 → OX → O(ai ) → TX → 0 i

1 implies H (X, TX ) = 0. Therefore X is rigid.

NOTES 1. The main references for this section are [154], [120], [14]. 2. ([174]) Let B be an e.f.t. local k-algebra, I ⊂ B an ideal generated by a regular sequence. Prove that if B/I is rigid then B is rigid as well.

3.2 Closed subschemes

3.2.1 The local Hilbert functor

Let X ⊂ Y be a closed embedding of algebraic schemes. A cartesian diagram of morphisms of schemes 3.2 Closed subschemes 123 X → X ⊂ Y × S η : ↓↓π Spec(k) → S where π is flat, and it is induced by the projection from Y × S, is called a (flat) family of deformations of X in Y parametrized by S, or over S. We call S and X respectively the parameter scheme and the total scheme of the family. When S = Spec(A) with A in ob(A∗) (resp. in ob(A), resp. A = k[ ]) we say that η is a local, resp. infinitesimal, resp. first-order family of deformations of X in Y . We will also say that η is a deformation of X in Y over A. The family is called trivial if X = X × S. X is rigid in Y if every infinitesimal deformation of X in Y is trivial. When considering a family η we speak generally of a family of closed subschemes of Y . Let X ⊂ Y be a closed embedding of algebraic schemes. For each A in ob(A) we let Y ( ) ={ } HX A deformations of X in Y over A We can immediately verify that this defines a functor of Artin rings

Y : A → HX (sets) called the local Hilbert functor of X in Y . Proposition 3.2.1. Given a closed embedding of algebraic schemes X ⊂ Y then: Y ¯ (i) The local Hilbert functor HX satisfies conditions H0,H , H, H of Theorem 2.3.2. (ii) There is a natural identification

Y ( [ ]) = 0( , ) HX k H X NX/Y

where NX/Y is the normal sheaf of X in Y . Y Proof. (i) Obviously HX satisfies condition H0.Let A A

A ¯ be homomorphisms in A, with A → A a small extension. Letting A = A ×A A we have a commutative diagram with exact rows:

0 → ( ) → A¯ → A → 0 ↓↓ 0 → ( ) → A → A → 0

Take an element of Y Y H (A ) × Y ( ) H (A ) X HX A X 124 3 Examples of deformation functors which is represented by a pair of deformations X ⊂ Y × Spec(A ) and X ⊂ Y × Spec(A ) such that

X ×Spec(A ) Spec(A) = X ×Spec(A ) Spec(A) ⊂ Y × Spec(A)

¯ O × O Consider the sheaf of A-algebras X OX X on X. Then X¯ := (| |, O × O ) ( ¯) ( ¯) X X OX X is a scheme over Spec A , flat over Spec A (see the proof of 2.4.1). Therefore X¯ is a deformation of X over Spec(A¯) inducing X and X . We have a commutative diagram: X → X ↓↓ X → Y × Spec(A¯) and the universal property of the fibred sum implies that there is a morphism

Φ : X¯ → Y × Spec(A¯)

Pulling back Φ over Spec(A ) (resp. Spec(A )) we obtain the closed embedding X ⊂ Y × Spec(A ) (resp. X ⊂ Y × Spec(A )). Since Spec(A ) ⊂ Spec(A¯) is a closed embedding defined by a square zero ideal, it follows that Φ is a closed embedding as well (details are left to the reader). Therefore Φ : X¯ ⊂ Y × Spec(A¯) Y ( ¯) (X , X ) defines an element of HX A which is mapped to by the map: Y Y Y α : H (A ×A A ) → H (A ) × Y ( ) H (A ) X X HX A X It follows that α is surjective. Now let X˜ ⊂ Y × Spec(A¯) be another element of Y ( ¯) (X , X ) HX A which is mapped to . Then by the universal property of the fibred sum there is a morphism τ : X¯ → X˜ which, as in the proof of 2.4.1, is easily seen to be an isomorphism. Moreover, the following diagram commutes: Y × Spec(A¯)  X¯ −→τ X˜

Since the diagonal arrows are closed embeddings it follows that X¯ = X˜ as closed subschemes of Y ×Spec(A¯). This proves that α is actually a bijection and (i) follows. (ii) Let I ⊂ OY be the ideal sheaf of X. A first-order deformation of X in Y , i.e. a flat family: X → X ⊂ Y × Spec(k[ ]) ↓↓ (3.13) Spec(k) → Spec(k[ ]) is defined by a sheaf OX of flat k[ ]-algebras, with an isomorphism ∼ OX ⊗k[ ] k = OX . The closed embedding X ⊂ Y × Spec(k[ ]) is determined by a sheaf of ideals I ⊂ OY [ ]:=OY ⊗k k[ ] such that OX = OY [ ]/I .The above data are obtained by gluing together their restrictions to an affine open cover. 3.2 Closed subschemes 125

On an affine open set U = Spec(P) ⊂ Y ,letX ∩ U = Spec(B) where B = P/J for an ideal J = ( f1,..., fN ) ⊂ P. Consider the exact sequence

v f 0 → R −→ P N −→ J → 0 where R is the module of relations among f1,..., fN . Taking HomP (−, B) we obtain the exact sequence:

∨ 2 N v 0 → HomB(J/J , B) → HomP (P , B) −→ HomP (R, B)

2 ∨ ∨ which identifies HomB(J/J , B) with ker(v ). An element of ker(v ) can be rep- = ( ,..., ) resented as an N-tuple h h1 h N of elements of P which, interpreted as an ( N , ) ( ,..., ) = element of HomP P B by scalar product (i.e. h p1 pN j h j p j mod J) must be zero on R. Hence h j r j ∈ J for every (r1,...,rN ) ∈ R j

This means that there exist ∆r1,...,∆rN ∈ P such that h j r j =− f j ∆r j j j or, equivalently, such that

( f + h) t (r + ∆r) = 0 in P ⊗k k[ ]. Therefore from Corollary A.11 it follows that f + h generates an ideal in P ⊗k k[ ] which defines a first-order deformation of Spec(B) in Spec(P) because every relation among f1,..., fN extends to a relation among f1 + h1,..., fN + h N . Using the same argument backwards one sees that every first-order de- 2 formation of Spec(B) in Spec(P) defines an element of HomB(J/J , B). It follows that at the global level we have a canonical 1–1 correspondence 0 between first-order deformations of X in Y and H (X, NX/Y ).  Corollary 3.2.2. Let X ⊂ Y be a closed embedding of algebraic schemes. If 0( , )<∞ Y h X NX/Y , for example if X is projective, then HX is prorepresentable. Proof. Follows from Proposition 3.2.1 and from Theorem 2.3.2.  If X ⊂ Y is a closed embedding of projective schemes then the prorepresentabi- Y Y lity of HX follows directly from the existence of the Hilbert scheme Hilb because Y Oˆ HX is prorepresented by the complete local ring HilbY ,[X] (see § 4.3). Let A¯ be in ob(Aˆ). A formal deformation of X in Y is a sequence ¯ Xn ⊂ Y × Spec(An) ξn : ↓ ¯ Spec(An) 126 3 Examples of deformation functors

¯ ¯ n+1 of infinitesimal deformations of X in Y over An = A/m ¯ such that ξn induces ¯ A ¯ ξn−1 by pullback under the natural inclusion Spec(An−1) → Spec(An) for all n ≥ 1. ¯ We can describe the formal deformation (A, {ξn}) of X in Y as a diagram of formal schemes X ⊂ Y × Specf(A¯) ↓ Specf(A¯) ¯ As in the case of the functor DefX , a formal deformation (A, {ξn}) of X in Y defines ξˆ ∈ Y ( ¯) ( ¯, ξ)ˆ Y an element HX A , i.e. a formal couple A for HX , and conversely, every such element is defined by a formal deformation of X in Y . Considering not necessarily infinitesimal deformations of a closed subscheme one has the notion of characteristic map, as follows. Definition 3.2.3. Let X ⊂ Y be a closed embedding of algebraic schemes, (S, s) a pointed scheme and X ⊂ X ⊂ Y × S ξ : ↓↓π s Spec(k) −→ S be a family of deformations of X in Y . By pulling back this family by morphisms Spec(k[ ]) → S with image s and applying Proposition 3.2.1 we obtain a linear map 0 χξ : Ts S → H (X, NX/Y ) called the characteristic map of the family ξ at s.

The characteristic map is the analogous for embedded deformations of the Kodaira–Spencer map for abstract ones.

Examples 3.2.4. (i) If X ⊂ IP r , r ≥ 1, is the complete intersection of r − n hypersurfaces f1,..., fr−n of degrees d1 ≤ d2 ≤ ... ≤ dr−n respectively, we have a presentation

2 f [⊕ j O(−d j )] −→ ⊕ j O(−d j ) → IX → 0 (3.14)

Taking Hom(−, OX ) we obtain an exact sequence:

2 fˇ 0 → NX →⊕j OX (d j ) −→ [⊕ j OX (d j )] where NX = NX/IP r . Since each nonzero entry of the matrix defining f is one of the ˇ f j ’s, the map f is zero. Therefore ∼ NX = ⊕ j OX (d j )

Equivalently, one can remark that (3.14) induces a surjective homomorphism 3.2 Closed subschemes 127

⊕ O (− ) → I /I2 j X d j X X of locally free sheaves of the same rank, which must therefore be an isomorphism. ∼ In particular, if X is a hypersurface of degree d we have NX = OX (d) and therefore d + r h0(X, N ) = − 1 X r confirming the fact that X can be deformed in IP r only inside the linear system of d+r − hypersurfaces of degree d, which is a projective space of dimension r 1. ∼ ⊕r−n If X is a linear subspace, i.e. d1 = ··· = dr−n = 1, then NX = OX (1) and therefore 0 h (X, NX ) = (r − n)(n + 1) as expected, since such linear subspaces X are parametrized by the grassmannian G(n + 1, r + 1), which is nonsingular of dimension (r − n)(n + 1). (ii) Let X be a Cartier divisor on a connected projective scheme Y . Consider the exact sequence 0 → OY → OY (X) → NX/Y → 0 and the cohomology sequence:

χ 0 0 0 1 0 → H (Y, OY ) → H (Y, OY (X)) −→ H (X, NX/Y ) → H (Y, OY ) (3.15)

Classically, the map χ was called the characteristic map of the linear system |X|. ∼ 0 0 By definition, Im(χ) = H (Y, OY (X))/H (Y, OY ) is naturally identified with the tangent space to the linear system |X|. We can verify that this is so by identifying Im(χ) as a subvector space of first-order deformations of X in Y , as follows. Assume that X is defined by a system of local equations { fi }, fi ∈ Γ(Ui , OY ) not 0-divisor, with respect to an affine cover {Ui } of Y .Wehave := −1 ∈ Γ( ∩ , O∗ ) , { } fij fi f j Ui U j Y for all i j, and fij is a Cech 1-cocycle which defines the line bundle OY (X). A first-order deformation of X in Y is a Cartier di- visor X ⊂ Y × Spec(k[ ]) which is determined by a system {Fi = fi + gi }, ∈ Γ( , O ) = + ∈ Γ( ∩ , O∗ ) gi Ui Y , such that there exist Fij fij gij Ui U j Y ×Spec(k[ ]) (hence gij ∈ Γ(Ui ∩ U j , OY )) satisfying Fi = FijFj for all i, j. Therefore on Ui ∩ U j we have: fi + gi = ( fij + gij)( f j + g j ) which is equivalent to the identity:

gi = fijg j + gij f j

This identity shows that the system {¯gi = gi mod IX } defines a section of NX/Y ,as expected. We also see that {¯gi }∈Im(χ) if and only if gi = fijg j , in which case Fij = fij, i.e. X deforms inside the linear system |X|, as asserted. 128 3 Examples of deformation functors

(iii) In IP 2 × IP 1 with bihomogeneous coordinates (u,v,w; x, y) consider the hypersurface m, m ≥ 0, defined by the equation:

xmv − ymu = 0

∼ 1 One can show that m = Fm, and that the structure morphism π : Fm → IP is 2 1 1 induced by the projection IP × IP → IP . Therefore the surface m is a realiza- 2 tion of the rational ruled surface Fm as the total scheme of a family of lines of IP 1 ∼ 1 1 parametrized by IP . When m > 0 this family is nontrivial because Fm = IP × IP (for details see [5]).

Examples 3.2.5. The following set of examples deals with properties of projective curves. Proofs are straightforward and are left to the reader. (i) If Y is a projective scheme and C ⊂ Y is a projective integral l.c.i. curve, then the normal sheaf NC/Y is torsion free. If C is nonsingular then NC/Y is locally free. (ii) Consider a nonsingular curve C ⊂ IP 3 and a (possibly singular) surface S ⊂ IP 3 of degree n containing C. Prove that there is an exact sequence of locally free sheaves on C: ψ → −1 ⊗ (− + ) → −→ O ( ) →[O / ]( ) → 0 a KC n 4 NC/IP 3 C n C a n 0 (3.16) where a ⊂ OC is the ideal sheaf generated by the restriction to C of the partial derivatives ∂ ∂ F ,..., F ∂ X0 ∂ X3 where F = 0 is an equation of S. (ψ) = O ⊗ ( → ) (Hint: Im C Im TIP 3|S NS/IP 3 ). In the case where S is nonsingular we obtain the sequence:

0 → KC (−n + 4) → NC → OC (n) → 0 (3.17)

Deduce from (3.16) that if a = OC (i.e. if C ∩ Sing(S) =∅) then C is not regularly embedded in S (yet the normal sheaf NC/S is locally free.) (iii) Consider a nonsingular curve C ⊂ IP 3 and a point p ∈ IP 3\C. Then there is an exact sequence → O ( ) → → ω ( ) → 0 C 1 NC/IP 3 C 3 0 which is obtained as a special case of (3.16) by taking as S the cone projecting C from p. Deduce that 1( , ) ≤ 1( , O ( )) h C NC/IP 3 h C C 1

(iv) Let C ⊂ IP r , r ≥ 4, be a nonsingular irreducible projective curve. Let r r−k p1,...,pk ∈ IP ,1≤ k ≤ r − 3, be general points and let π : C → IP be the projection with centre the linear span (p1,...,pk). Prove that there is an exact sequence 3.2 Closed subschemes 129

→ O ( )k → → π ∗ → 0 C 1 NC/IP r Nπ(C)/IP r−k 0 (3.18) which splits if and only if C is contained in a hyperplane. Deduce that

1 1 h (C, NC/IP r ) ≤ (r − 2)h (C, OC (1)) (3.19)

1 1 In particular, H (C, NC/IP r ) = (0) if h (C, OC (1)) = 0. (Solution: see [163], Prop. 11.2). Inequality (3.19) was already known classically as a bound on the dimension of the Hilbert scheme ([169], s. 8). For a sharpening of (3.19) see [47]. (v) Let C ⊂ IP r be a nonsingular irreducible projective curve and let L = OC (1). Show that if the Petri map µ0(L) (see Example 3.3.9) is injective then 1 H (C, NC/IP r ) = (0). (vi) Let C ⊂ IP r be a projective irreducible nonsingular curve of degree d and genus g. Prove that

0 1 χ(NC/IP r ) := h (C, NC/IP r ) − h (C, NC/IP r ) = (r + 1)d + (r − 3)(1 − g)

In particular, for a canonical curve of genus r + 1 ≥ 3 and degree 2r in IP r we have

0 χ(NC/IP r ) = h (C, NC/IP r ) = r(r + 5)

1 because h (C, NC/IP r ) = 0 since the Petri map µ0(ωC ) is injective (see (v) above).

3.2.2 Obstructions

Let X ⊂ Y be a closed embedding, A in ob(A) and let

X → X ⊂ Y × Spec(A) ξ : ↓↓f Spec(k) → Spec(A) be a deformation of X in Y over A.Let

e : 0 → k → A˜ → A → 0 be a small extension. A lifting of ξ to A˜ is a deformation of X in Y over A˜:

X → X˜ ⊂ Y × Spec(A˜) ξ˜ : ↓↓f Spec(k) → Spec(A˜) whose pullback to Spec(A) is ξ.

Proposition 3.2.6. Let X ⊂ Y be a regular closed embedding (Definition D.1.1) of 1 algebraic schemes with X projective. Then H (X, NX/Y ) is an obstruction space for Y the local Hilbert functor HX . 130 3 Examples of deformation functors

Proof. Given X → X ⊂ Y × Spec(A) ξ : ↓↓f Spec(k) → Spec(A) where A is in ob(A), an infinitesimal deformation of X in Y , we will show that there is a natural linear map

1 oξ/Y : Exk(A, k) → H (X, NX/Y ) such that, for every extension

e : 0 → k → A˜ → A → 0 ˜ we have oξ/Y (e) = 0 if and only if ξ has a lifting to A; this will prove the proposition. Since X is regularly embedded in Y we can find an affine open cover U ={Ui }i∈I of Y such that Xi := X ∩ Ui is a complete intersection in Ui for each i.Let Ui = Spec(Pi ), Xi = Spec(Pi /Ii ) where Ii = ( fi1,..., fiN) with { fi1,..., fiN} X = ( /I ) a regular sequence in Pi . We then have |Ui Spec PiA iA where

IiA = (Fi1,...,FiN) ⊂ PiA := Pi ⊗ A = α = ,..., ˜ ,..., ˜ ∈ and fiα Fiα mod m A, 1 N. Choose arbitrarily Fi1 FiN Pi A˜ ˜ ˜ ˜ such that Fiα = Fiα mod . By Example A.12 {Fi1,...,FiN} and {Fi1,...,FiN} I = are regular sequences in PiA and in Pi A˜ respectively; in particular, letting i A˜ ˜ ˜ (Fi1,...,FiN), X˜ := ( /I ) ⊂ × ( ˜) i Spec Pi A˜ i A˜ Ui Spec A X ⊂ × ( ) ˜ X ˜ is a lifting of |Ui Ui Spec A to A. In order to find a lifting of to A we must ˜ be able to choose the Fiα’s in such a way that X˜ = X˜ ⊂ × ( ˜) i|Uij j|Uij Uij Spec A (3.20) for each i, j ∈ I . Letting Uij = Spec(Pij) and viewing the fiα’s and f jα’s as elements of Pij via the natural maps

Pi Pj

Pij we have ˜ ˜ Fjα − Fiα =: hijα := ( ,..., ) ∈ Γ( , ) and hij hij1 hijN Uij NX/Y , because NX/Y is locally free of rank { }∈Z1(U, ) N and is trivial on each Ui . By construction, hij NX/Y . The condition ˜ = , (3.20) means that we can choose the Fiα’s so that hij 0 all i j. A different choice ˜ ˜ + := ( ,..., ) ∈ Γ( , ) of the Fiα’s is of the form Fiα hiα and hi hi1 hiN Ui NX/Y . Since we have 3.2 Closed subschemes 131 ˜ ˜ (Fjα + h jα) − (Fiα + hiα) = (hijα + h jα − hiα) (3.21) { } ( ) ∈ 1( , ) we see that hij defines an element oξ/Y e H X NX/Y which is zero if and ˜ ˜ ˜ only if the Xi ’s satisfy condition (3.20) and define a lifting X of X to A.Itiseasyto show that the map oξ is k-linear.  1 ˜ The element oξ/Y (e) ∈ H (X, NX/Y ) is called the obstruction to lifting ξ to A; we call ξ obstructed if oξ/Y (e) = 0forsomee ∈ Exk(A, k); otherwise it is unob- structed. X is said to be unobstructed in Y if all its infinitesimal deformations in Y are unobstructed; otherwise X is said to be obstructed in Y . Examples of obstructed closed subschemes are usually quite subtle, especially if one is interested in nonsin- gular obstructed subvarieties. In order to be able to describe them in a natural way it is necessary to know the existence of the Hilbert scheme of a projective scheme. We will give examples in § 4.6.

Corollary 3.2.7. Let j : X ⊂ Y be a regular closed embedding of algebraic schemes with X projective and let (R, {ξn}) be the formal universal deformation of X in Y . Then (i) 0 0 1 h (X, NX/Y ) ≥ dim(R) ≥ h (X, NX/Y ) − h (X, NX/Y ) The first equality holds if and only if X is unobstructed in Y . 0 (ii) X is rigid in Y if and only if H (X, NX/Y ) = 0. 1 (iii) If H (X, NX/Y ) = (0) then X is unobstructed in Y .

The proof is left to the reader. In the case of a closed embedding X ⊂ Y which Y is not regular, Proposition 3.2.6 says nothing about the obstruction space of HX .We refer the reader to § 4.4 and § 4.6 for some information about the general case.

Examples 3.2.8. (i) Let C be a projective nonsingular curve contained in a nonsingu- lar surface S, and assume that C is negatively embedded in S, i.e. deg(OC (C)) < 0. 0 Then H (C, OC (C)) = 0 and therefore C is rigid in S. This happens in particular when C =∼ IP 1 is an exceptional curve of the first kind. Another example is when C has genus g ≥ 2, S = C × C and C is identified to the diagonal ∆ ⊂ S. In this case NC/S = TC , which has degree 2 − 2g < 0. Note that 1 H (C, TC ) = (0) but C is unobstructed in S, being rigid in S. This example shows that the sufficient condition of Corollary 3.2.7 is not necessary. (ii) Hypersurfaces of IP r are unobstructed. In fact, if X ⊂ IP r has degree d then

1 1 h (X, NX/IP r ) = h (X, OX (d)) = 0

More generally, complete intersections in IP r are unobstructed (see Subsection 4.6.1). (iii) Let Q ⊂ IP 3 be a quadric cone with vertex v, and L ⊂ Q a line. Then we have an inclusion ⊂ = O ( ) ⊕ O ( ) NL/Q NL/IP 3 L 1 L 1 132 3 Examples of deformation functors whose cokernel is OL (2)(−v) (see Example 3.2.5(ii)). It follows that NL/Q = 1 OL (1); in particular, it is locally trivial and H (L, NL/Q) = 0, despite the fact that L ⊂ Q is not a regular embedding (Example 3.2.5(ii)) and L is obstructed in Q (see Note 1 of § 4.6). (iv) Let C ⊂ IP r be a nonsingular irreducible projective curve and let L = OC (1). If the Petri map µ0(L) (see Example 3.3.9) is injective then C is unobstructed r 1 in IP because h (C, NC/IP r ) = 0 (Example 3.2.5(v)). In particular canonical curves of genus g ≥ 3 are unobstructed in IP g−1 (Example 3.2.5(vi)).

3.2.3 The forgetful morphism

Let X ⊂ Y be a closed embedding of algebraic schemes. The forgetful morphism

Φ : Y → HX DefX is the morphism which associates to an infinitesimal deformation of X in Y :

X → X ⊂ Y × Spec(A) ξ : ↓↓ Spec(k) → Spec(A) the isomorphism class of the deformation of X:

X → X ↓↓ Spec(k) → Spec(A)

Proposition 3.2.9. Assume that X and Y are nonsingular and that X is projective. Consider the exact sequence

0 → TX → TY |X → NX/Y → 0 (3.22)

Then: 0 1 (i) dΦ = δ : H (X, NX/Y ) → H (X, TX ) is the coboundary map coming from (3.22). (ii) The coboundary map

1 2 δ1 : H (X, NX/Y ) → H (X, TX )

arising from the exact sequence (3.22) is an obstruction map for Φ (see Defini- tion 2.3.5). 1 (iii) If H (X, TY |X ) = 0 then Φ is smooth. (iv) If X is unobstructed in Y and δ is surjective then Φ is smooth. In particular, X is unobstructed as an abstract variety and has rk(δ) number of moduli. 3.2 Closed subschemes 133

Proof. (i) Let X ⊂ X ⊂ Y × Spec(k[ ]) ξ : ↓↓π s Spec(k) −→ Spec(k[ ]) be a first-order deformation of X in Y . We must show that δχ(ξ) = κ(ξ).Let 0 χ(ξ) = h ∈ H (X, NX/Y ). Consider an affine open cover U ={Ui = Spec(Pi )} of Y , and let Xi = X ∩ Ui = Spec(Pi /( fi1,..., fiN)).Wehave X := X = ( [ ]/( + ,..., + )) i |Ui Spec P fi1 hi1 fiN hiN

Then (h ,...,h ) =: h = h ∈ Γ(U , N / ). Since X is affine and non- i1 iN i |Ui i X Y i singular the abstract deformation Xi of Xi is trivial: thus there exist isomorphisms 1 θi : Xi × Spec(k[ ]) → Xi and κ(ξ) ∈ H (X, TX ) is defined by the 1-cocycle 1 {dij}∈Z (U, TX ) corresponding to the system of automorphisms θ = θ −1θ : × ( [ ]) → × ( [ ]) ij i j Xij Spec k Xij Spec k where Xij = X ∩ Uij. Let’s compute δ(h). The isomorphism θi is given by an isomorphism of k[ ]-algebras

ti : Pi [ ]/( fi1 + hi1,..., fiN + hiN) → Pi [ ]/( fi1,..., fiN) which, by the smoothness of Pi , is induced by a k[ ]-automorphism

Ti : Pi [ ]→Pi [ ] of the form Ti (p + q) = p + (q + di (p)) where di ∈ Derk(Pi , Pi ) = Γ(Ui , TY ) 0 is such that di (hiα) =−fiα.Wehave{di }∈C (U, TY ) and δ(h) is defined by {( − ) } ( − ) = δ( ) = κ(ξ) d j di |Xi . Since d j di |Xi dij we conclude that h . (ii) One must show that, given an infinitesimal deformation X ⊂ X ⊂ Y × Spec(A) ξ : ↓↓π s Spec(k) −→ Spec(A) where A is in A, we have a commutative diagram

Exk(A, k) oξ/Y oξ δ 1 1 2 H (X, NX/Y ) −→ H (X, TX ) The proof of this fact is similar to the proof of part (i) and will be omitted. (iii) The exact sequence (3.22) shows that the hypothesis implies 0 1 dΦ : H (X, NX/Y ) → H (X, TX ) surjective and

1 2 δ1 : H (X, NX/Y ) → H (X, TX ) injective Therefore the assertion follows from Proposition 2.3.6. (iv) is left to the reader.  134 3 Examples of deformation functors

Remark 3.2.10. Let X ⊂ Y be a regular embedding of algebraic schemes with X I ⊂ O Φ : Y → reduced and Y nonsingular, Y the ideal sheaf of X, and let HX DefX be the forgetful morphism. The differential

0 2 1 1 dΦ : H (X, N / ) = HomO (I/I , O ) → Ext (Ω , O ) X Y X X OX X X

2 is the k-linear map which associates to σ : I/I → OX the pushout σ∗(S) where : → I/I2 → Ω1 → Ω1 → S 0 Y |X X 0 is the conormal sequence of X ⊂ Y . This generalizes Proposition 3.2.9. The proof consists in considering, for a first-order deformation of X in Y

X ⊂ X ⊂ Y × Spec(k[ ]) the induced diagram of conormal sequences → I/I2 ⊕ O → Ω1 ⊕ O → Ω1 → 0 X Y |X X X 0 ↓↓ → O → Ω1 → Ω1 → 0 X X |X X 0 and in recognizing the second row as the pushout of the first.

Part (iii) of the proposition often gives a very effective way of proving that a given X ⊂ Y is unobstructed as an abstract variety. If X is a curve in IP r the vanishing of 1 H (X, TIP r |X ) is related to the Petri map (see Example 3.3.9). The following examples are further applications of this principle.

Examples 3.2.11. (i) ([107]) Let X ⊂ IP r , r ≥ 3, be a nonsingular hypersurface of 1 1 degree d ≥ 2. Then h (NX/IP r ) = h (OX (d)) = 0 and therefore X is unobstructed in IP r . On the other hand, from the exact sequence:

0 → TIP r (−d) → TIP r → TIP r |X → 0 and the Euler sequence:

r+1 0 → OIP r → OIP r (1) → TIP r → 0 we deduce that: 1 2 h (TIP r |X ) = h (TIP r (−d)) = 0ifr ≥ 4 while for r = 3 we have the exact sequence: ← 2( (− ))∨ ← 0(O ( − )) ← 0(O ( − ))4 0 H TIP 3 d H IP 3 d 4 H IP 3 d 5  1( )∨ H TIP 3|X Therefore we see that 1 1ifr = 3 and d = 4; h (T r | ) = IP X 0 otherwise 3.2 Closed subschemes 135

IP r → From 3.2.9 we therefore deduce that HX DefX is smooth and X is unobstructed as an abstract variety unless r = 3 and d = 4 (this is precisely the case when X is a K3 surface). An analogous result holds more generally for complete intersections ([161]). 2 Using (3.22) one computes easily that H (TX ) = 0ifX is a nonsingular surface of degree d ≥ 5inIP 3: therefore the unobstructedness of X could not have been deduced from 2.4.6 in this case. One can generalize to singular hypersurfaces as follows. Consider a reduced hypersurface X ⊂ IP r , r ≥ 3, of degree d ≥ 2. Then the conormal sequence is

→ O (− ) → Ω1 → Ω1 → 0 X d IP r |X X 0

1 so that H (X, NX/IP r ) = 0 and we have the exact sequence

Φ 0( , ) −→d 1 (Ω1 , O ) → 1(Ω1 , O ) H X NX/IP r ExtO X X Ext IP r |X X X 0 1 H (X, OX (d)) H (X, TIP r |X )

Ω1 where the equality on the right is because IP r |X is locally free. Therefore as before, Φ : Y → ( , ) = ( , ) we see that HX DefX is smooth and X is unobstructed if r d 3 4 . (ii) The previous example can be easily generalized to nonsingular hypersurfaces of IP n × IP m,1≤ n ≤ m, n + m ≥ 3. Let

q IP n × IP m −→ IP m ↓ p IP n be the projections. Consider a nonsingular hypersurface X ⊂ IP n × IP m of bidegree (a, b), i.e. defined by an equation σ = 0forsomeσ ∈ H 0(O(a, b)), where

O(a, b) := p∗O(a) ⊗ q∗O(b)

From the exact sequence

0 → O → O(a, b) → NX/IP n×IP m → 0 one deduces that 1 H (NX/IP n×IP m ) = (0) and therefore X is unobstructed in IP n × IP m. For any coherent sheaf F on IP n × IP m we use the notation F(α, β) = F ⊗ O(α, β) Using the fact that ∗ ∗ TIP n×IP m = p TIP n ⊕ q TIP m 136 3 Examples of deformation functors and the Leray spectral sequence with respect to any one of the projections, one easily computes that i h (TIP n×IP m (α, β)) = 0 when n + m ≥ 4, i = 1, 2 and (α, β) arbitrary. Moreover, when (n, m) = (1, 2) one finds: i ( ) = ≥ h TIP 1×IP 2 0alli 1 2( (− , − )) = ( , ) = ( , ) h TIP 1×IP 2 a b 0 unless a b 2 3 Putting all this information together and using the exact sequence

0 → TIP n×IP m (−a, −b) → TIP n×IP m → TIP n×IP m |X → 0 one deduces that 1 h (TIP n×IP m |X ) = 0 unless (n, m) = (1, 2) and (a, b) = (2, 3) (this is precisely the case when X is a K3 surface). Now as before, we conclude that the forgetful morphism IP n×IP m → HX DefX is smooth and X is unobstructed as an abstract variety.

3.2.4 The local relative Hilbert functor

Given a projective morphism p : X → S of schemes and a k-rational point s ∈ S, consider the fibre X (s) and a closed subscheme Z ⊂ X (s). For each A in ob(A) an infinitesimal deformation of Z in X relative to p parametrized by A is a commutative diagram: Z ⊂ XA → X ↓ ↓p s Spec(A) −→ S where the right square is cartesian, the left diagonal morphism is flat and its closed fibre is Z; this means in particular that the morphism s has image {s} and therefore that A is an OS,s-algebra. Then, letting Λ = OS,s, we can define the local relative Hilbert functor X /S : A → HZ Λ (sets) by X / infinitesimal deformations of Z in X H S(A) = Z relative to p parametrized by A We have the following generalization of Corollaries 3.2.2 and 2.4.7:

Theorem 3.2.12. Let p : X → S be a projective morphism of schemes, s ∈ Sa k-rational point, and Z ⊂ X (s) a closed subscheme of the fibre X (s). Denote OS,s by Λ. Then: X /S : A → (i) The local relative Hilbert functor HZ Λ (sets) is prorepresentable and 0 has tangent space H (Z, NZ/X (s)). 3.3 Invertible sheaves 137

1 (ii) If Z is regularly embedded in X (s) and p is flat then H (Z, NZ/X (s)) is an X /S obstruction space for HZ , and we have an exact sequence:

0 1 0 → H (Z, NZ/X (s)) → tR → Ts S → H (Z, NZ/X (s)) (3.23)

Λ X /S where R is the local -algebra prorepresenting HZ .

X /S Proof. (i) The proof of 3.2.1 can be followed almost verbatim, showing that HZ ¯ 0 satisfies conditions H0, H , H, H and that H (Z, NZ/X (s)) is its tangent space. (ii) The proof of 3.2.6 can be easily adapted to this case. The exact sequence (3.23) follows from the above and from (2.2).  Another generalization of the local Hilbert functors can be obtained with no extra effort. Consider a projective scheme X and a formal deformation of X

π¯ : X → Specf(R) where R is in ob(Aˆ) and π¯ is a flat projective morphism of formal schemes; let Z ⊂ X be a closed subscheme. For each A in ob(A) define an infinitesimal defor- mation of Z in X relative to π¯ parametrized by A as a commutative diagram:

Z ⊂ X A → X ↓ ↓¯π s Spec(A) −→ Specf(R) where the right square is cartesian, the left diagonal morphism is flat and its closed fibre is Z. Note that Spec(A) = Specf(A) and the morphism s is defined by a surjec- tive homomorphism R → A, so that X A is just an ordinary scheme projective and flat over Spec(A). We can define the local relative Hilbert functor

X /Specf(R) : A → HZ R (sets) as above. A result analogous to 3.2.12 can be proved in this case as well with a similar proof. Details of this straightforward generalization are left to the reader.

3.3 Invertible sheaves

3.3.1 The local Picard functors

The local Picard functors are related to the Picard schemes in the same way as the lo- cal Hilbert functors are to the Hilbert schemes, in the sense that if the Picard scheme of a given scheme exists then the local Picard functors describe its infinitesimal and local properties. In this section we will prove the basic properties of the local Picard functors. The Picard schemes will not be treated here. For a full treatment the reader is referred to Kleiman [100]. 138 3 Examples of deformation functors

Let X be a scheme. Recall that the Picard group of X is defined to be the group Pic(X) of isomorphism classes of invertible sheaves on X; we have the identifica- ( ) = 1( , O∗ ) [ ]∈ ( ) tion Pic X H X X (see [84]). We denote by L Pic X the class of an invertible sheaf L. For every A in ob(A) we will write for brevity X A instead of X × Spec(A). For every morphism A → B in A, and for every invertible sheaf L on X A we denote by L ⊗A B the invertible sheaf it induces on X B by pullback. Similarly, given λ ∈ Pic(X A) we denote by λ ⊗A B ∈ Pic(X B) its image under the pullback operation. This defines a homomorphism of group

∗ ∗ Pic(X ) = H 1(X , O ) → H 1(X , O ) = Pic(X ) A A X A B X B B which makes A → Pic(X A) a covariant functor on A with values in the category of abelian groups. We fix an element λ0 ∈ Pic(X) once and for all and, for each A in ob(A),welet ( ) := {λ ∈ ( ) : λ ⊗ = λ } Pλ0 A Pic X A A k 0 ( ) ( ) Then Pλ0 A is a subset of Pic X A whose definition is functorial in A so that we have a functor of Artin rings: : A → Pλ0 (sets) which will be called the local Picard functor of (X,λ0). Let L be an invertible sheaf on X such that λ0 =[L]. For a given A ∈ ob(A) ( ) the elements of Pλ0 A are the isomorphism classes of infinitesimal deformations of L over A:aninfinitesimal deformation of L over A is an invertible sheaf L on X × Spec(A) such that L = L|X×Spec(k). In the case A = k[ ] we speak of a first- order deformation of L. The main result about this functor is the following.

Theorem 3.3.1. Let X be an algebraic scheme, λ0 ∈ Pic(X). Assume that the fol- lowing conditions are satisfied: 0 ∼ (a) H (X, OX ) = k 1 (b) dimk H (X, OX )<∞ ( [ ]) = 1( , O ) 2( , O ) Then Pλ0 is prorepresentable and Pλ0 k H X X . Moreover, H X X is an obstruction space for Pλ0 . = Proof. Let’s check the conditions of Theorem 2.3.2. It is clear that P Pλ0 satisfies condition H0.Let A A

A be a diagram in A with A → A surjective. Consider an element

(λ ,λ ) ∈ P(A ) ×P(A) P(A ) 3.3 Invertible sheaves 139

and let λ = λ ⊗A A = λ ⊗A A.LetL , L , L be invertible sheaves on

X A , X A , X A respectively such that [L ]=λ , [L ]=λ , [L]=λ. Then we have homomorphisms L → L and L → L of sheaves on |X| inducing isomorphisms ∼ ∼ L ⊗A A = L, L ⊗A A = L.LetB = A ×A A . ∼ Claim: O = O ×O O . X B X A X A X A For every open set U ⊂|X| we have by definition:

[O ×O O ](U) = O (U) ×O ( ) O (U) X A X A X A X A X A U X A and by the universal property of the fibred sum we have a homomorphism

φ : O → O ×O O X B X A X A X A O → O O → O which is induced by the homomorphisms X B X A and X B X A coming from B → A and B → A respectively. Since A → A is surjective φ ⊗ : O ⊗ → O B A X B B A X A is an isomorphism. From Lemma A.4 it follows that φ is an isomorphism.

From the claim we deduce that N := L ×L L is an invertible sheaf on X B and ∼ ∼ the projections induce isomorphisms N ⊗B A = L , N ⊗B A = L . Therefore

P(B) *[N] → (λ ,λ ) ∈ P(A ) ×P(A) P(A )

This shows that the map

α : P(B) → P(A ) ×P(A) P(A ) is surjective. Assume that M is an invertible sheaf such that α([M]) = α([N]). This means that there are homomorphisms M → L and M → L inducing isomorphisms ∼ ∼ M ⊗B A = L , M ⊗B A = L . It follows that we have an automorphism θ : L → L given by the composition:

∼ ∼ ∼ ∼ L = L ⊗A A = M ⊗B A = L ⊗A A = L which makes the following diagram commutative:

M q q L L u u θ L −→ L 140 3 Examples of deformation functors

By hypothesis (a) the isomorphism θ is the multiplication by a unit a ∈ A. Since A → A is surjective we can lift a to a ∈ A and we can change q to a q thus assuming that u q = u q . It follows that we have a commutative diagram M

L L

L and this implies that M =∼ N. This shows that α is bijective. Therefore P satisfies also conditions H and H. We have an exact sequence

exp ∗ ∗ 0 → O −→ O → O → 0 X Xk[ ] X where exp( f ) = 1 + f .Itfollowsthat

1 ∗ 1 ∗ 1 P(k[ ]) = ker[H (X [ ], O ) → H (X, O )]=H (X, O ) k Xk[ ] X X

Finally, given A in ob(A) and [e]∈Exk(A, k) represented by an extension

0 → (t) → A˜ → A → 0 we have an exact sequence:

∗ ∗ 0 → tOX → O → O → 0 X A˜ X A which induces an exact sequence

1 ∗ 1 ∗ δ 2 H (X ˜ , O ) → H (X A, O ) −→ H (X, OX ) A X A˜ X A

˜ ˜ 2 Given λ ∈ P(A) it can be lifted to a λ ∈ P(A) if and only if δ(λ) = 0 ∈ H (X, OX ). 2 This shows that H (X, OX ) is an obstruction space for P. 

Corollary 3.3.2. Let X be a projective integral scheme. Then Pλ0 is prorepresentable for every λ0 ∈ Pic(X).

Proof. A projective integral scheme satisfies both conditions (a) and (b) of the theorem. 

Remark 3.3.3. Let X be an algebraic scheme and λ0 ∈ Pic(X). Then the tangent and obstruction spaces of the functor Pλ0 , as described by Theorem 3.3.1, depend only on X and not on λ0. This is because, given any λ0,µ0 ∈ Pic(X), there is a canonical =∼ isomorphism of functors Pλ0 Pµ0 . We leave to the reader the easy proof of this fact. 3.3 Invertible sheaves 141

3.3.2 Deformations of sections, I

Let X be a projective integral scheme, and let L be an invertible sheaf on X. One can define a homomorphism of sheaves:

0 ∨ m : OX → H (X, L) ⊗ L (3.24) as follows. For every open set U ⊂ X

m(U) : O(U) → H 0(X, L)∨ ⊗ Γ(U, L)

f →[s → fs|U ] for every s ∈ H 0(X, L). The induced maps on global sections are just given by cup product: i 0 i mi : H (X, OX ) → Hom(H (X, L), H (X, L)) a →[s → a ∪ s] If L is base point free and

0 ∨ ϕL : X → IP := IP (H (X, L) ) is the morphism defined by the sections of L, then it is easy to check that the homo- morphism (3.24) is the same as the one appearing in the pulled back Euler sequence of IP : → O → 0( , )∨ ⊗ → ϕ∗ → 0 X H X L L L TIP 0 We leave this to the reader. The linear maps between cohomology groups induced by these sheaf homomorphisms have deformation theoretic interpretations which we will now explain. Consider a deformation L of L over A ∈ ob(A): we have an induced restriction map 0 0 ρL : H (X A, L) → H (X, L)

We say that a section σ ∈ H 0(X, L) extends to L if σ ∈ Im(ρL).

Proposition 3.3.4. Let La be a first-order deformation of L, corresponding to an 1 0 element a ∈ H (X, OX ). A section s ∈ H (X, L) extends to La if and only if

a ∪ s = 0 ∈ H 1(X, L) equivalently, if and only if s ∈ ker[m1(a)] where

1 0 1 m1 : H (X, OX ) → Hom(H (X, L), H (X, L)) is the map induced by (3.24). 142 3 Examples of deformation functors

Proof. Let U ={Uα} be an affine open covering of X such that L is represented by { } ∈ Γ( , O∗ ) a system of transition functions fαβ , fαβ Uαβ X . Then the first-order de- formation La of L can be represented, in the same covering {Uα} of X × Spec(k[ ]), by transition functions: ˜ ∈ Γ( , ∗ ) fαβ Uαβ OX×Spec(k[ ]) such that ˜ ˜ ˜ fαβ fβγ = fαγ (3.25) and which restrict to the fαβ ’s modulo . O∗ = O∗ + O Since X×Spec(k[ ]) X X we can write ˜ fαβ = fαβ + gαβ (3.26) for suitable gαβ ∈ Γ(Uαβ , OX ). Identity (3.25) gives

gαβ gβγ gαγ + = fαβ fβγ fαγ

( gαβ ) ˇ ∈ 1( , O ) and the system fαβ is a Cech 1-cocycle which defines the element a H X X . Let’s assume that s ∈ H 0(X, L) is represented by the cocycle (sα), sα ∈ Γ(Uα, OX ), such that sα = fαβ sβ on Uαβ .Fors to extend to a section

0 s˜ ∈ H (X × Spec(k[ ]), La) it is necessary and sufficient that there exist tα ∈ Γ(Uα, OX ) such that ˜ sα + tα = fαβ (sβ + tβ ) ˜ on Uαβ . After replacing the fαβ ’s by the expressions (3.26) we obtain the identities

sα + tα = ( fαβ + gαβ )(sβ + tβ ) which are equivalent to: gαβ sβ = tα − fαβ tβ These identities can be also written as: gαβ sα = tα − fαβ tβ fαβ and they mean exactly that the 1-cocycle gαβ 1 sα ∈ Z (U, L) fαβ is a coboundary, i.e. that a ∪ s = 0.  3.3 Invertible sheaves 143

Corollary 3.3.5. In the above situation:

(i) If a ∈ ker(m1) then all the sections of L extend to La. 1 0 1 (ii) For every a ∈ H (X, OX ), at least max{0, h (X, L) − h (X, L)} linearly inde- pendent sections of L extend to La. Proof. Immediate.  Let’s fix a section s ∈ H 0(X, L) and let D = div(s) ⊂ X be the divisor of s. X Consider the local Hilbert functor HD . We have a morphism of functors: : X → aD HD P[L] associating to a deformation of D in X over A in ob(A), given by an effective Cartier D ⊂ [O (D)]∈ ( ) divisor X A, the element X A P[L] A . aD is called the Abel–Jacobi morphism of D ⊂ X. Consider the exact sequence s 0 → OX −→ L → L D → 0 (3.27) It induces linear maps: 0 1 δ0 : H (D, L D) → H (X, OX )

1 2 δ1 : H (D, L D) → H (X, OX )

Noting that L D = ND/ X we have:

Proposition 3.3.6. In the above situation δ0 is the differential of aD and δ1 is an obstruction map for aD. Proof. We keep the notations of the proof of Proposition 3.3.4 and therefore we assume D defined by local equations sα = 0 where (sα) is a 0-cocycle defining s with respect to the covering U; in particular, sα = fαβ sβ . A first-order deformation D ⊂ X × Spec(k[ ]) of D is defined by local equations

sα + tα = 0 which satisfy the cocycle conditions:

sα + tα = ( fαβ + gαβ )(sβ + tβ ) These conditions can be also written as: gαβ tα = fαβ tβ + sα (3.28) fαβ They mean that (t¯α = tα mod sα) ¯ 0 define a section t ∈ H (D, L D) which is the one corresponding to the first-order de- ¯ formation D of D. From the identities (3.28) we also see that δ0(t) is represented by ( gαβ ) (D) ∈ 1( , O ) the 1-cocycle fαβ . Since this cocycle represents aD H X X this proves δ (¯) = (¯) δ that 0 t daD t . The proof that 1 is an obstruction map is similar and will be left to the reader.  144 3 Examples of deformation functors

Corollary 3.3.7. Under the assumptions of Proposition 3.3.6: 1 (i) If H (X, L) = 0 then aD is smooth. (ii) If the natural map 1 1 H (X, L) → H (D, L D) X is zero then HD is less obstructed than P[L]. If, moreover, P[L] is smooth then X HD is smooth. 1 Proof. (i) If H (X, L) = 0 then (3.27) implies that δ0 is surjective and δ1 is injec- tive. Therefore aD is smooth by Proposition 2.3.6. The proof of (ii) is similar using Proposition 2.3.6 again.  Remarks 3.3.8. (i) Assume X to be projective and nonsingular. A Cartier divisor D on X is called semiregular if the natural map

1 1 H (X, OX (D)) → H (D, OD(D)) is zero, i.e. if condition (ii) of the corollary is satisfied. Part (ii) of the corollary can thus be rephrased by saying that D is unobstructed in X if it is semiregular and P[L] is smooth. The smoothness of P[L] is known to be true if char(k) = 0, by a general theorem of Cartier: we thus recover a celebrated theorem of Severi, Kodaira and Spencer claiming the unobstructedness of semiregular divisors on a projective nonsingular complex variety X. These matters are covered in full detail in [130]. The original sources are [172] and [108]. For further developments and applications of the notion of semiregularity see [21], [149] and [23]. (ii) In view of Proposition 3.3.6 the cohomology sequence of (3.27) can be inter- preted as a sequence of tangent spaces and differentials as follows:

δ H 0(X,L) 0 0 1 0 → (σ) → H (D, L D) −→ H (X, OX )

 

d → | |→ X ( [ ]) −→aD ( [ ]) 0 TD D HD k P[L] k

H 0(X,L) where we used the identification of (σ) with the tangent space to the linear system |D| at D proved in Example 3.2.4(ii). Example 3.3.9. Let X be a Gorenstein curve (e.g. a local complete intersection curve). Then the map m1 is dual to the map: 0 0 −1 0 µ0(L) : H (X, L) ⊗ H (X,ωX L ) → H (X,ωX ) (3.29) given by multiplication of global sections (ωX denotes the dualizing sheaf of X). 0 µ0(L) is called the Petri map of L. More generally, if V ⊂ H (X, L) is a vector space of sections of L, one can consider the multiplication map: 3.3 Invertible sheaves 145

0 −1 0 µ0(V ) : V ⊗ H (X,ωX L ) → H (X,ωX ) which is called thePetrimapofV. From Proposition 3.3.4 it follows that ⊥ 1 coker(µ0(V )) ⊂ H (X, OX ) is the space of first-order deformations of L to which all sections of V extend. In particular, if h0(X, L) = r + 1 and deg(L) = d then (µ ( ))⊥ r ( ) coker 0 L is the tangent space at L to the scheme Wd X of linear systems of degree d and dimension ≥ r on X (see [9] for details).

3.3.3 Deformations of pairs (X, L)

: O → Ω1 Let X be a nonsingular projective algebraic variety and let d X X be the canonical derivation. We can define a homomorphism of sheaves of abelian groups O∗ → Ω1 X X by the rule du u → u ⊂ ∈ Γ( , O∗ ) for all open sets U X and u U X . We have an induced group homomor- phism: : 1( , O∗ ) → 1( ,Ω1 ) c H X X H X X To simplify the notation we write c(L) instead of c([L]) for a given invertible sheaf L on X.Ifk = C then c(L) is, up to a multiplicative constant, the Chern class of L Ω1 (see [128], p. 127). Since X is locally free we have an identification

H 1(X,Ω1 ) = Ext1 (T , O ) X OX X X so that we can associate to c(L) an extension

0 → OX → EL → TX → 0 (3.30) defined up to isomorphism, called the Atiyah extension of L. The sheaf EL is locally free of rank dim(X) + 1 and P := E∨ ⊗ L L OX L is called the sheaf of (first-order) principal parts of L. Let U ={Uα} be an affine open covering of X such that L is represented by a { } ∈ Γ( , O∗ ) ( ) system of transition functions fαβ , fαβ Uαβ X . Then c L is represented ˇ by the Cech 1-cocycle dfαβ ∈ Z1(U,Ω1 ) X fαβ

The sheaf EL|Uα is isomorphic to OUα ⊕ TX|Uα . A section (aα, dα) of OUα ⊕ TX|Uα and a section (aβ , dβ ) of OUβ ⊕ TX|Uβ are identified on Uαβ if and only if dα = dβ dα( fαβ ) β − α = and a a fαβ . 146 3 Examples of deformation functors

Remark 3.3.10. We have c(L ⊗ M) = c(L) + c(M), in particular c(Ln) = nc(L) for any n ∈ ZZ. Therefore the Atiyah extension of Ln is a constant multiple of the ∼ extension (3.30). This means in particular that if n = 0 then ELn = EL . Consider, for example, X = IP := IP (V ) for some finite-dimensional k-vector space V . Then one can easily compute that the Euler sequence

0 → OIP → V ⊗ OIP (1) → TIP → 0 ∼ is the Atiyah extension of OIP (1). Therefore EL = EO(1) = V ⊗ OIP (1) for every nontrivial line bundle L on IP .

Let A be in ob(A).Aninfinitesimal deformation of the pair (X, L) over A con- sists of a pair (ξ, L) (also denoted by (X , L)), where X → X ξ : ↓↓ Spec(k) → Spec(A) is an infinitesimal deformation of X over A and L is an invertible sheaf on X such that L = L|X . One can also say that L is a deformation of L along ξ. In the case A = k[ ] we speak of a first order-deformation of (X, L). Two deformations (X , L) and (X , L ) of (X, L) over A will be called isomorphic if there is an isomorphism of deformations f : X → X and an isomorphism L → f ∗L . By letting Def(X,L)(A) = deformations of (X, L) over A /isomorphism we define a functor of Artin rings

Def(X,L) : A → (sets) called the functor of infinitesimal deformations of the pair (X, L). We will denote by [X , L]∈Def(X,L)(A) the isomorphism class of a deformation (X , L) of (X, L) over A ∈ ob(A).

Theorem 3.3.11. Let (X, L) be a pair consisting of a nonsingular projective alge- braic variety X and an invertible sheaf L on X. Then:

(i) The functor Def(X,L) has a semiuniversal formal element. (ii) There is a canonical isomorphism ( , ) 1st-order deformations of X L ∼ 1 Def( , )(k[ ]) = = H (X, E ) X L isomorphism L

2 and H (X, EL ) is an obstruction space for Def(X,L). 3.3 Invertible sheaves 147

(iii) Given a first-order deformation ξ of X, there is a first-order deformation of L along ξ if and only if κ(ξ) · c(L) = 0 where “·” denotes the composition:

1( , ) × 1( ,Ω1 ) −→∪ 2( , ⊗ Ω1 ) → 2( , O ) H X TX H X X H X TX X H X X of the cup product of cohomology classes ∪ with the map induced by the du- ⊗ Ω1 → O ality pairing TX X X (therefore the left-hand side is an element of 2 H (X, OX )).

Proof. Let’s check the conditions of Theorem 2.3.2. It is clear that Def(X,L) satisfies condition H0.Let A A

A be a diagram in A with A → A surjective. Consider an element ([X , L ], [X , L ]) ∈ ( ) × ( ) Def(X,L) A Def(X,L)(A) Def(X,L) A

Then there is an [X , L]∈Def(X,L)(A) and a diagram of deformations: X X  f  f ↓ X ↓ Spec(A ) ↓ Spec(A )  Spec(A) where the morphisms f and f induce isomorphisms of deformations: ∼ ∼ X ×Spec(A ) Spec(A) = X = X ×Spec(A ) Spec(A)

Moreover, we have isomorphisms

f ∗L =∼ L =∼ f ∗L

Let B = A ×A A . Then, as in the proof of Theorem 2.4.1, one sees that

X¯ := (| |, O × O ) X X OX X is a deformation of X over B inducing the pair ([X ], [X ]). Define ¯ L := L ×L L 148 3 Examples of deformation functors

Then L¯ is an invertible sheaf over X¯ which restricts on X to L and on X to L . ¯ ¯ Therefore the pair (X , L) defines an element of Def(X,L)(B) such that

[X¯, L¯] → ([X , L ], [X , L ]) (3.31) under the map ( ) → ( ) × ( ) Def(X,L) B Def(X,L) A Def(X,L)(A) Def(X,L) A ¯ This proves that Def(X,L) satisfies condition H of Theorem 2.3.2. In order to show that Def(X,L) satisfies condition H we must prove that the element [X¯, L¯] constructed above is the unique one satisfying (3.31) if we assume ¯ that A = k[ ]. From the proof of Theorem 2.4.1 we know that [X ]∈DefX (B) is ([X ], [X ]) ∈ ( ) × ( ) the unique element inducing DefX A DefX (A) DefX A .Using this fact the proof can now be completed along the lines of the proof of Theorem 3.3.1. We still need to verify that Def(X,L) satisfies condition H f . This will result as a consequence of part (ii) to be proved next, because X is projective. (ii) Let (ξ, L) be a first-order deformation of (X, L), where X → X ξ : ↓↓ Spec(k) → Spec(k[ ])

Let U ={Uα} be an affine open covering such that L is given by a system of tran- ∈ Z1(U, O∗ ) κ(ξ) ∈ 1( , ) ˇ sition functions fαβ X and H X TX is given by a Cech 1 1-cocycle (dαβ ) ∈ Z (U, TX ).Letθαβ = 1 + dαβ be the automorphism of Uαβ × Spec(k[ ]) corresponding to dαβ . The invertible sheaf L can be given by a system of transition functions (Fαβ ) ∈ 1 ∗ Z (U, OX ) which reduces to { fαβ } mod . Therefore it can be represented on Uαβ × Spec(k[ ]) as

Fαβ = fαβ + gαβ , gαβ ∈ Γ(Uαβ , OX ) and the cocycle condition translates into

Fαβ θαβ (Fβγ ) = Fαγ equivalently: Fαβ (Fβγ + dαβ Fβγ ) = Fαγ which means:

( fαβ + gαβ )[ fβγ + gβγ + dαβ ( fβγ + gβγ )]= fαγ + gαγ

After dividing by fαγ and equating the coefficients of we obtain:

gαβ gβγ gαγ dαβ fβγ + − + = 0 (3.32) fαβ fβγ fαγ fβγ 3.3 Invertible sheaves 149

gαβ {( , αβ )} Z1(U, E ) This identity means that the data fαβ d define an element of L , and that conversely such an element defines a first-order deformation of (X, L).This proves that ∼ 1 Def(X,L)(k[ ]) = H (X, EL ) modulo verifying that this correspondence is independent from the choice of the covering U and of the cocycles representing (X, L); we leave this to the reader. Consider a small extension

e : 0 → (t) → A˜ → A → 0 in A and let (ξ, L) be an infinitesimal deformation of (X, L) over A, where X → X ξ : ↓↓ Spec(k) → Spec(A)

Let U ={Uα} be an affine open cover of X;let

θα : Uα × Spec(A) → X|Uα

−1 be isomorphisms so that θαβ := θα θβ is an automorphism of the trivial deformation Uαβ × Spec(A). We may assume that L is given by a system of transition functions (Fαβ ), where each Fαβ is a nowhere zero function on Uαβ × Spec(A), such that

Fαβ θαβ (Fβγ ) = Fαγ

In order to see if a lifting (ξ,˜ L˜) of (ξ, L) to Spec(A˜) exists we choose arbitrarily a collection ˜ {θ˜αβ , Fαβ } where, for each α, β, γ : ˜ ˜ (a) θαβ is an automorphism of the product family Uαβ × Spec(A) which restricts to θαβ on Uαβ × Spec(A). ˜ ˜ (b) Fαβ is a nowhere zero function on Uαβ × Spec(A) which restricts to Fαβ on Uαβ × Spec(A).

Such collection exists by Lemma 1.2.8. Because of (a) we have

˜ ˜ ˜−1 θαβ θβγ θαγ = id + tdαβγ

2 where “id” here means the identity of Γ(Uαβγ , OX ), and (dαβγ ) ∈ Z (U, TX ) is a 2-cocycle which represents the obstruction to lifting ξ over Spec(A˜) (see the proof of Proposition 1.2.12). Because of (b), for each α, β, γ there is gαβγ ∈ Γ(Uαβγ , OX ) such that ˜ ˜ ˜ ˜ −1 Fαβ θαβ (Fβγ )Fαγ = 1 + tgαβγ 150 3 Examples of deformation functors

Therefore we have: ˜ ˜ ˜ ˜ ˜ −1 θαβ Fβγ θβγ (Fγδ)Fβδ · ˜ ˜ ˜ ˜ −1 −1 ˜ ˜ ˜ ˜ −1 ˜ ˜ ˜ ˜ −1 −1 [Fαγ θαγ (Fγδ)Fαδ ] [Fαβ θαβ (Fβδ)Fαδ ][Fαβ θαβ (Fβγ )Fαγ ]

= 1 + t(gβγδ − gαγδ + gαβδ − gαβγ ) On the other hand, the left side can be written as ( ) ˜ ˜ −1 dαβγ fγδ θ˜αβ θ˜βγ (Fγδ)[θ˜αγ (Fγδ)] = t fγδ Therefore we see that 2 (gαβγ , dαβγ ) ∈ Z (U, EL ) (3.33) ˜ ˜ We are free to modify our choice by replacing the θαβ ’s and the Fαβ ’s by ˜ ˜ Φαβ = θαβ + tdαβ , Gαβ = Fαβ + tgαβ (3.34) for some dαβ ∈ Γ(Uαβ , TX ) and gαβ ∈ Γ(Uαβ , OX ). One checks easily that the cocycle (3.33) is replaced by the cohomologous one:

(gαβγ + gαβ − gαγ + gβγ , dαβγ + dαβ − dαγ + dβγ ) (3.35) and it is clear that the lifting (ξ,˜ L˜) exists if and only if the data (3.34) can be deter- mined so that (3.35) is zero. Therefore associating to the extension e the cohomology 2 class o(ξ,L)(e) ∈ H (X, EL ) defined by the cocycle (3.33) we have defined an ob- struction map 2 Exk(A, k) → H (X, EL ) 2 which we leave to the reader to verify to be linear. This makes H (X, EL ) an ob- struction space for the functor Def(X,L).

(iii) Observing that dαβ fβγ 2 ∈ Z (U, OX ) fβγ represents κ(ξ) · c(L), the identity (3.32) expresses the condition that this 2-cocycle is a coboundary, and proves (iii).  2 If X is a connected nonsingular projective curve then H (X, OX ) = 0 so that every line bundle can be deformed along any first-order deformation of X. Moreover, 2 H (X, EL ) = 0 for any L, thus Def(X,L) is smooth and from the exact sequence (3.30) it follows that 1 h (X, EL ) = 4g − 3 if X has genus g ≥ 2. For higher-dimensional varieties the situation is more complicated in general. For example, if X is a K3-surface then the cup product 1( , ) × 1( ,Ω1 ) → 2( , O ) =∼ H X TX H X X H X X k 3.3 Invertible sheaves 151 coincides with Serre duality. Therefore

· ( ) 1 c L 2 H (X, TX ) −→ H (X, OX ) is surjective for every nontrivial line bundle L. This means that L deforms along 1 1 a 19-dimensional subspace of H (X, TX ), because h (X, TX ) = 20 (see Example 2.4.11(ii), page 73). We have a natural forgetful morphism

Φ : Def(X,L) → DefX defined in the obvious way. The differential and the obstruction map of this mor- phism are described as follows.

Proposition 3.3.12. In the situation of Theorem 3.3.11:

(i) The differential dΦ : Def(X,L)(k[ ]) → DefX (k[ ]) coincides with the linear map

1 1 H (X, EL ) → H (X, TX )

coming from the exact sequence (3.30). (ii) The map 2 2 H (X, EL ) → H (X, TX ) coming from the exact sequence (3.30) is an obstruction map for Φ. 2 (iii) If H (X, OX ) = 0 the morphism Φ is smooth.

Proof. (i) and (ii) are left to the reader. For (iii) use Proposition 2.3.6. 

Example 3.3.13. Assume k = C.LetA = V/Λ be an abelian variety of dimension g, represented as the quotient of a g-dimensional complex vector space V by a lattice Λ ⊂ V . Then, letting Ω = V ∨ we have = ⊗ O ,Ω1 = Ω ⊗ O TA V A A A Moreover, 1( , O ) = Ω¯ := ( , ) H A A HomC¯ V C is the space of C-antilinear forms on V and

2 2 ¯ H (A, OA) = Ω

(see [20], Theorem 1.4.1). Therefore

1 1 ¯ H (A, TA) = V ⊗ H (A, OA) = V ⊗ Ω 152 3 Examples of deformation functors

1 2 In particular, h (A, TA) = g .Wealsohave: 1( ,Ω1 ) = Ω ⊗ 1( , O ) = Ω ⊗ Ω¯ H A A H A A which can be identified with the space of hermitian forms on V .LetL be an ample invertible sheaf on X. Then

( ) ∈ 1( ,Ω1 ) = Ω ⊗ Ω¯ c L H A A is identified with a positive definite hermitian form hL : V × V → C ([20], § 4.1). It follows that the map · ( ) 1 c L 2 H (A, TA) −→ H (A, OA) (3.36) of Theorem 3.3.11 is just the composition of the map

V ⊗ Ω¯ → Ω¯ ⊗ Ω¯ (3.37) v ⊗  → hL (v, −) ⊗  with the canonical surjection

2 Ω¯ ⊗ Ω¯ → Ω¯

Since hL is positive definite the map (3.37) is an isomorphism and it follows that the map (3.36) is surjective. The conclusion is that L deforms along a subspace of 1 H (A, TA) of dimension g(g + 1) g = g2 − 2 2 For the related case of jacobians see Example 3.4.24(iii), page 182.

3.3.4 Deformations of sections, II

Consider again a projective nonsingular variety X and an invertible sheaf L on X. Assume as above that L is given, in an affine open cover U ={Uα}, by transition ( ) ∈ Z1(U, O∗ ) functions fαβ X . We can define a homomorphism of sheaves

0 ∨ M : EL → H (X, L) ⊗ L in the following way. Consider a section η ∈ Γ(U, EL ), where U ⊂ X is an open set; it is given by a system (aα, dα) where aα ∈ Γ(U ∩ Uα, OX ), dα ∈ Γ(U ∩ Uα, TX ), subject to the conditions that dβ = dα and aβ − aα = dα( fαβ )/fαβ on U ∩ Uα ∩ Uβ . Then, for every s = (sα) ∈ H 0(X, L) we let

M(η)(sα) = aαsα + dα(sα) 3.3 Invertible sheaves 153

On U ∩ Uα ∩ Uβ we find:

fαβ M(η)(sβ ) = fαβ (aβ sβ + dβ (sβ )) = aβ sα + fαβ dβ (sβ ) = sα(aα + dα( fαβ )/fαβ ) + fαβ dβ (sβ ) = sαaα + dα( fαβ )sβ + fαβ dβ (sβ ) = sαaα + dα( fαβ sβ ) = sαaα + dα(sα) = M(η)(sα)

Therefore the functions M(η)(sα) ∈ Γ(U ∩ Uα, OX ) patch together to define a section M(η)(s) ∈ Γ(U, L). This defines M. It is obvious from the definition that we have a commutative diagram: 0 ∨ m : OX → H (X, L) ⊗ L ↓ (3.38) 0 ∨ M : EL → H (X, L) ⊗ L namely M extends m (see (3.24), page 141). The linear map 1 0 1 M1 : H (X, EL ) → Hom(H (X, L), H (X, L)) 1 induced by M can be explicitly described as follows. Let η1 ∈ H (X, EL ) be repre- 1 sented by the Cechˇ cocycle (aαβ , dαβ ) ∈ Z (U, EL ). Then 0 1 M1(η1) : H (X, L) → H (X, L)

(sα) → (aαβ sα + dαβ sα)

The map M1 has the following deformation theoretic interpretation. Let A ∈ ob(A) and let (X , L) be an infinitesimal deformation of (X, L) over Spec(A). Then we say that a section s ∈ H 0(X, L) extends to L if

s ∈ Im[H 0(X , L) → H 0(X, L)]

Proposition 3.3.14. Let X be a projective nonsingular variety, L a line bundle on X, and (X , L) a first-order deformation of the pair (X, L) defined by a cohomology 1 0 class η1 ∈ H (X, EL ) according to Theorem 3.3.11(ii). Then a section s ∈ H (X, L) extends to L if and only if s ∈ ker(M1(η1)).

Proof. With respect to an affine cover U ={Uα} of X we assume L represented ( ) ∈ Z1(U, O∗ ) η ( , ) ∈ Z1(U, E ) by fαβ X and 1 by aαβ dαβ L . Then, according to the description given in the proof of Theorem 3.3.11, X is determined by gluing the products Uα × Spec(k[ ]) along the open subsets Uαβ × Spec(k[ ]) by means of the automorphisms θαβ = 1 + dαβ , and L is determined by transition func- tions of the form Fαβ = fαβ (1 + aαβ ) satisfying the patching identities on Uαβγ × Spec(k[ ]): Fαβ θαβ (Fβγ ) = Fαγ

The condition that s = (sα) ∈ H 0(X, L) extends to L is equivalent to the existence of functions tα ∈ Γ(Uα, OX ) such that

θαβ [Fαβ (sβ + tβ )]=sα + tα 154 3 Examples of deformation functors for all α, β. After replacing the expressions for Fαβ and θαβ we obtain the identities:

fαβ (1 + aαβ )(sβ + tβ ) + dαβ ( fαβ sβ ) = sα + tα

By equating the coefficients of on both sides we obtain:

aαβ sα + dαβ (sα) = tα − fαβ tβ and this means exactly that M1(η1)(s) = 0.  Corollary 3.3.15. In the above situation:

(i) If η1 ∈ ker(M1) then all the sections of L extend to the first-order deformation of (X, L) corresponding to η1. 1 0 1 (ii) For any η1 ∈ H (X, EL ) at least max{0, h (X, L)−h (X, L)} linearly indepen- dent sections of L extend to the first-order deformation of (X, L) corresponding to η1. Proof. Immediate.  This statement, as well as Proposition 3.3.14, show the deformation theoretic interest of the map M1. It can be useful to have a closer picture of the relation between the maps m and M. Assuming that L is base point free and denoting by 0 ∨ ϕL : X → IP := IP (H (X, L) ) the morphism defined by the sections of L,we have a commutative and exact diagram, which extends (3.38):

0 ↓ 0 TX ↓↓ → O −→m 0( , )∨ ⊗ → ϕ∗ → 0 X H X L L L TIP 0 ↓↓ (3.39) → E −→M 0( , )∨ ⊗ → → 0 L H X L L NϕL 0 ↓↓ TX 0 ↓ 0

The first column is the Atiyah extension. The sheaf NϕL is the normal sheaf of the morphism ϕL . It will be considered more systematically in § 3.4, page 162. In the ϕ = case where L is a closed embedding, NϕL NϕL (X)/IP and this diagram shows that

1 ∗ coker(m1) ⊂ H (X,ϕ TIP ) where we recall that

1 0 1 m1 : H (X, OX ) → Hom(H (X, L), H (X, L)) is the map induced by m (see Proposition 3.3.4, page 141). Therefore, according 2 to Proposition 3.2.9(iii), h (X, OX ) = 0 and the surjectivity of m1 are sufficient 3.3 Invertible sheaves 155 conditions for the smoothness of the forgetful morphism Φ : H IP → Def .In ϕL (X) X particular, if X is a nonsingular curve and m1 is surjective then Φ is smooth (see also Example 3.2.5(v)).

Example 3.3.16. Let X be a connected projective nonsingular curve of genus g. Then the map M1 is dual to : 0( , ) ⊗ 0( ,ω −1) → 0( ,ω ⊗ E∨) PL H X L H X X L H X X L More generally, for any vector subspace V ⊂ H 0(X, L) we can consider the map obtained by restricting PL : : ⊗ 0( ,ω −1) → 0( ,ω ⊗ E∨) PV V H X X L H X X L

⊥ 1 It follows directly from Proposition 3.3.14 that coker(PV ) ⊂ H (X, EL ) is the space of first-order deformations of (X, L) to which all sections of V extend. PL (resp. PV )iscalledthe extended Petri map of L (resp. of V ). After dualizing the cohomology diagram of (3.39) we deduce the following commutative diagram which relates the Petri map and the extended Petri map in the case where L is globally generated: 0 ↓ 0(ω2 ) 0 H X ↓↓ (3.40) → 1( )∨ → 0( ) ⊗ 0(ω −1) −→PL 0(ω ⊗ E∨) 0 H NϕL H L H X L H X L ↓ ↓ µ ( ) → 1(ϕ∗ )∨ → 0( ) ⊗ 0(ω −1) −→0 L 0(ω ) 0 H L TIP H L H X L H X If ϕL : X ⊂ IP is an embedding then this diagram contains some significant infor- IP mation on the relations between HX and DefX . From what has been remarked just before this example it follows that the injectivity of µ0(L), being equivalent to the surjectivity of its dual m1, is a sufficient condition for the smoothness of the forget- Φ : IP → ful morphism HX DefX . Diagram (3.40) also shows that we have an exact sequence µ ( ) → 1( )∨ → 1( )∨ −→1 L 0(ω2 ) 0 H NX/IP H TIP |X H X  ker(PL ) ker(µ0(L)) thus the injectivity of PL is implied by the injectivity of µ0(L) or more generally, by the injectivity of µ1(L). For more about this topic we refer the reader to [8].

NOTES

1. The coboundary maps δk in the cohomology sequence of the Atiyah extension (3.30) are induced by cup product with c(L), since the extension is defined by c(L).Inparticular, 156 3 Examples of deformation functors

3.3.11(iii) just says that L deforms along ξ if and only if 1 2 κ(ξ) ∈ ker[δ1 : H (X, TX ) → H (X, OX )], which is obvious in view of (i) and (ii). 2. The content of Theorem 3.3.11(iii) is outlined in the Appendix by Mumford to Chapter V of [189]. See also [89]. This result is related to the notion of deformation of a polarization (see [132]). 3. For a discussion of Example 3.3.13 for abelian varieties defined over a field of positive characteristic see Oort [137], remark on page 226. 4. The Petri map has been considered classically in [141] and it reappeared in the modern literature for the first time in [10]. It is of great importance in the study of the Brill–Noether loci on a curve and of their relation to moduli (see [9] for more details).

3.4 Morphisms

In this section we study deformations of a morphism between algebraic schemes. In the analytic case the corresponding theory has been developed by Horikawa in [85], [86], [88] and [89]. For a treatment in the analytic case we refer to [134]. Related work in the algebraic case is in [114] and [148]. Definition 3.4.1. Let f : X → Y be a morphism of algebraic schemes and let A be in ob(A) (resp. A = k[ ], resp. A in ob(A∗)). An infinitesimal (resp. first-order, resp. local) family of deformations of f parametrized by A or by S (more briefly a deformation of f over A or over S) is a cartesian diagram X → X

↓ f ↓ F

Y → Y (3.41)

↓↓ψ

s Spec(k) −→ S where S = Spec(A), and ψ and ψ F are flat (“cartesian diagram” in this case means that the horizontal morphisms induce an isomorphism of the left column with the pullback of the right column by s). If we replace S by a pointed scheme (S, s) we will call (3.41) a family of deformations of f . Essentially, a deformation of f consists of a morphism F between deforma- tions of X and of Y and an assigned identification with f of the restriction of F to the closed fibre. Note that if Y = Spec(k) then a family of deformations of f : X → Spec(k) is just a family of deformations of X in the sense of the defi- nition given at the beginning of § 1.2. The notion of trivial deformation of f can be given in a obvious way, as well as the notion of rigid morphism. 3.4 Morphisms 157

Given an infinitesimal deformation (3.41) and a small extension

e : 0 → (t) → A˜ → A → 0 a lifting of (3.41) to A˜ is a cartesian diagram

X → X → X˜

↓ f ↓ F ↓ F˜

Y → Y → Y˜

↓↓ψ ↓ ψ˜

s Spec(k) −→ Spec(A) → Spec(A˜)

(i.e. the horizontal morphisms induce an isomorphism of each column with the pull- back of the column on the right by the lowest horizontal morphisms) where F˜ and ψ˜ F˜ are flat. Of course such a lifting defines in particular a deformation of f over A˜. Definition 3.4.1 gives the most general notion of deformation of a morphism. It can be modified in several ways so to obtain more restricted notions each having independent interest and applications. We will study some of them starting from the simplest ones. In each situation we will consider a corresponding functor of Artin rings which classifies deformations modulo an appropriate equivalence relation.

3.4.1 Deformations of a morphism leaving domain and target fixed

If in Definition 3.4.1 we impose that both X and Y are the product families, i.e. if we consider a cartesian diagram of the form:

X → X × S

↓ f ↓ F

Y → Y × S

↓↓ψ

Spec(k) → S where S = Spec(A) and ψ is the projection, then we obtain the notion of deformation of f with fixed domain and target (wfdat). Deformations of f wfdat can be interpreted as deformations of the graph of f in X × Y so that the methods introduced in § 3.2 apply. Precisely, let’s define a functor of Artin rings by setting 158 3 Examples of deformation functors DefX/f/Y (A) = deformations of f over A wfdat for all A in ob(A). Then we have the following: Proposition 3.4.2. Let f : X → Y be a morphism of algebraic schemes, with X projective and reduced and Y nonsingular. Then: (i) There is a natural isomorphism of functors ∼ X×Y Def / / = H X f Y Γ f

where Γ f ⊂ X×Y is the graph of f . In particular, DefX/f/Y is prorepresentable. (ii) We have a natural isomorphism of vector spaces: 0 ∗ DefX/f/Y (k[ ]) → H (X, f TY ) 1 ∗ (iii) H (X, f TY ) is an obstruction space for the functor DefX/f/Y . Proof. (i) Let A be in ob(A) and let F : X × Spec(A) → Y × Spec(A) be a deformation of f . Its graph ΓF ⊂ X × Y × Spec(A) defines a deformation of ∼ Γ f ⊂ X × Y over A because ΓF = X × Spec(A) and the projection ΓF → Spec(A) equals the composition ∼ ΓF = X × Spec(A) → Spec(A) in particular, it is flat. Therefore we can define X×Y Def / / (A) → H (A) X f Y Γ f by F → (ΓF ⊂ X × Y × Spec(A)) This is an isomomorphism of functors. In fact, given a deformation

ΓA ⊂ X × Y × Spec(A) of ΓF in X × Y the projection X × Y × Spec(A) → X × Spec(A) induces a morphism ΓA → X × Spec(A) which is an isomorphism of deformations. Therefore the composition ∼ X × Spec(A) = ΓA ⊂ X × Y × Spec(A) → Y × Spec(A) can be identified with a deformation of f . This defines the inverse of the morphism of functors of the statement. (ii) We have natural isomorphisms of vector spaces:

∼ X×Y ∼ 0 Def / / (k[ ]) = H (k[ ]) = H (Γ, NΓ/ × ) X f Y Γ f X Y Since the projection p : X × Y → X is smooth and the composition pj : Γ → X is an isomorphism, from Proposition D.2.5 it follows that j is a regular embedding. Therefore applying Proposition D.1.4 we obtain the exact sequence: 3.4 Morphisms 159

→ I /I2 → ∗Ω1 → Ω1 → 0 Γ Γ j X×Y Γ 0 On the other hand, we have the exact sequence: → ∗Ω1 → ∗Ω1 → ( )∗Ω1 → 0 f Y j X×Y pj X 0 obtained by restricting to Γ the sequence → ∗Ω1 → Ω1 → ∗Ω1 → 0 q Y X×Y p X 0 : × → ( )∗Ω1 =∼ Ω1 (where q X Y Y is the second projection). Since pj X Γ , comparing ∗Ω1 =∼ I /I2 the two sequences we deduce that f Y Γ Γ . Therefore 0(Γ, ) = (I /I2 , O ) =∼ ( ∗Ω1 , O ) = 0( , ∗ ) H NΓ/X×Y Hom Γ Γ Γ Hom f Y X H X f TY and (ii) follows. 1 1 ∗ Similarly, H (Γ, NΓ/X×Y ) = H (X, f TY ) and (iii) follows as well.  The notions of obstructed (resp. unobstructed) deformation, and of obstructed (resp. unobstructed) morphism can be given in the usual way. One can give the notion of rigid morphism wfdat in an obvious way. We leave to the reader the task of proving 0 ∗ that, under the hypothesis of Proposition 3.4.2, H (X, f TY ) = 0 implies that f is rigid wfdat. Let f : X → Y be a morphism of algebraic schemes with X reduced and Y nonsingular; let (3.41) be a family of deformations of f wfdat parametrized by a pointed scheme (S, s). To every tangent vector t ∈ TS,s, viewed as a morphism Spec(k[ ]) → S with image {s} we can associate the pullback of the family F over Spec(k[ ]) by t and, using the correspondence of 3.4.2(i), we obtain an element of 0 ∗ H (X, f TY ) associated to s. This defines a natural linear map

0 ∗ TS,s → H (X, f TY ) which will be called the characteristic map of the family (3.41) wfdat. The next corollary follows immediately from the proposition. It can also be de- duced as a consequence of Lemma 1.2.6.

Corollary 3.4.3. If X is a nonsingular projective scheme, then the space of first- 0 order deformations of the identity X → XisH (X, TX ).

Examples 3.4.4. (i) Let f : X → Y be a nonconstant morphism of projective nonsingular connected curves, with g(Y ) ≥ 2. Then deg(TY )<0 and therefore 0 ∗ h (X, f TY ) = 0. Thus f is rigid as a morphism wfdat. (ii) A morphism f from a scheme X to a projective space is given by a linear system on X: deformations of f wfdat can thus be interpreted as “deformations of linear systems” in an appropriate way. We briefly discuss the case of curves. Let X be a projective irreducible and nonsingular curve, f : X → IP r a morphism and let ∗ 0 L = f OIP r (1),deg(L) = n. Then f is defined by a vector subspace V ⊂ H (X, L) 160 3 Examples of deformation functors of dimension r + 1 plus the choice of a basis of V . From the Euler sequence pulled back to X we have: ∗ χ(f TIP r ) = (r + 1)χ(L) − χ(OX ) = ρ(g, r, n) + r(r + 2) where ρ(g, r, n) := g − (r + 1)(g − n + r) is the Brill–Noether number and

0 r(r + 2) = h (TIP r ) = dim[PGL(r + 1)]

Assume that r + 1 = h0(L), i.e. that f is defined by the complete linear series |L|, and consider the exact sequence

1 1 r+1 1 ∗ H (OX ) → H (L) → H ( f TIP r ) → 0 obtained from the Euler sequence. It dualizes as:

µ ( ) 1 ∗ ∨ 0 0 −1 0 L 0 0 → H ( f TIP r ) → H (L) ⊗ H (ωX L ) −→ H (ωX ) where µ0(L) is the Petri map (see Example 3.3.9, page 144). Therefore we see that f is unobstructed if µ0(L) is injective. A necessary condition for this to be true is that

0 0 −1 0 (r + 1)(g − n + r) = dim[H (L) ⊗ H (ωX L )]≤h (ωX ) = g i.e. that ρ(g, r, n) ≥ 0. This necessary condition is not sufficient. The simplest ex- ample is given by a nonsingular complete intersection X = Q ∩ S ⊂ IP 3 of a quadric cone Q and of a cubic surface S. Then X is a canonical curve of genus 4 and the pro- 1 | | ω −1 =∼ jection from the vertex of the cone defines a complete g3 L such that X L L. 1 In this case ρ(4, 1, 3) = 0 but dim[ker(µ0(L))]=1. The morphism f : X → IP 0( ∗ ) = 1 is obstructed because h f TIP 1 4butL is the unique g3 on X. Therefore the unobstructed first-order deformations of f are only those in the three-dimensional space coming from the automorphisms of IP 1. (iii) If IP 1 =∼ E ⊂ S is a nonsingular projective rational curve negatively embed- ded in a projective nonsingular surface S with E2 =−n < 0, n ≥ 1, then we have 0 0 h (E, NE/S) = 0 and h (E, TS|E ) = 3. More precisely, the exact sequence

0 → TE → TS|E → NE/S → 0 splits because Ext1 (N / , T ) = H 1(E, O (n + 2)) = 0: therefore OE E S E E ∼ TS|E = OE (2) ⊕ OE (−n)

This means that, despite the fact that E is rigid in S, the morphism f : E → S has a three-dimensional family of deformations obtained by composing it with the automorphisms of E. 3.4 Morphisms 161

More generally, whenever we have an embedding f : IP 1 → Y with Y nonsin- gular algebraic variety, we have an inclusion

0( 1, ) ⊂ 0( 1, ∗ ) H IP TIP 1 H IP f TY

0 1 ∗ which implies h (IP , f TY ) ≥ 3. In general, for any nonconstant morphism 1 ∗ f : IP → Y the sheaf f TY is locally free and splits as a direct sum of dim(Y ) − 1 invertible sheaves, by the structure theorem (see [136] p. 22). The study of such mor- phisms is closely related to the notions of uniruledness and rational connectedness. We refer to Debarre [42] and to Kollar [109] for a detailed treatment of these matters. (iv) Similarly, if E ⊂ Y is an embedding of a projective nonsingular curve of genus 1 into a nonsingular algebraic variety Y , from the inclusion OE = TE ⊂ TY |E we deduce 0 0 H (E, OE ) ⊂ H (E, TY |E )

0 which implies h (E, TY |E ) ≥ 1.

3.4.2 Deformations of a morphism leaving the target fixed

In this subsection we will follow quite faithfully the treatment given in [86]. Given a morphism f : X → Y of algebraic schemes, a notion slightly more general than the previous one is that of a deformation of f with target Y , obtained by specializing Definition 3.4.1 to the case when Y is the product family, i.e. by considering a cartesian diagram of the form

X → X

↓ f ↓ F

Y → Y × S

↓↓ψ

Spec(k) → S where S = Spec(A) with A in ob(A) (resp. in ob(A∗)) and ψ is the projection. Such a deformation can be denoted concisely by the diagram

F X −→ Y × S

(3.42)

S 162 3 Examples of deformation functors

The deformation (3.42) will be called locally trivial if its domain X defines a locally trivial deformation of X. Given a deformation (3.42) and another deformation of f with target Y parametrized by S:

F X −→ Y × S

S an isomorphism between them is an isomorphism of deformations of X:

X

X −→Φ X

S which makes the following diagram commutative: X

↓ Φ Y × S  X

Definition 3.4.5. To a morphism f : X → Y of algebraic schemes there is associ- ated an exact sequence of coherent sheaves on X

→ −→J −→df ( ∗Ω1 , O ) −→P → 0 TX/Y TX Hom f Y X N f 0 (3.43) which defines the sheaf N f called the normal sheaf of f . The morphism f is called non-degenerate when TX/Y = 0. The sequence (3.43) is of course related to the exact sequence (1.5) on page 16. = 1 Comparing (3.43) with (1.5) we see that N f TX/Y if X is nonsingular. If f is smooth then N f = 0. The condition that f is non-degenerate is equivalent to f being unramified on a dense open subset of X. 1 We now introduce vector spaces DX/Y and DX/Y which will be used to describe infinitesimal deformations of the morphism f . Definition 3.4.6. Let f : X → Y be a morphism between algebraic schemes, with X projective. Let U ={Ui }i∈I be an affine open cover of X and define 3.4 Morphisms 163

0 ∗ 1 {(v, t) ∈ C (U, f TY ) × Z (U, TX ) : δv = df(t)} D / = X Y 0 {(df(w), δw) : w ∈ C (U, TX )} and {(ζ, s) ∈ C1(U, f ∗T ) × Z2(U, T ) : δζ = df(s)} D1 = Y X X/Y 1 {(df(u), δu) : u ∈ C (U, TX )} where δ is the coboundary map in Cechˇ cohomology.

Lemma 3.4.7. In the situation of Definition 3.4.6: 1 U (i) DX/Y and DX/Y don’t depend on the choice of the affine cover of X. (ii) We have the following exact sequences:

0 0 ∗ 1 1 ∗ (a) H (X, TX ) → H (X, f TY ) → DX/Y → H (X, TX ) → H (X, f TY )

1 0 2 (b) 0 → H (X, TX/Y ) → DX/Y → H (X, N f ) → H (X, TX/Y ) (3.44) ( ) 1( , ) → 1( , ∗ ) → 1 → 2( , ) → 2( , ∗ ) c H X TX H X f TY DX/Y H X TX H X f TY

( ) → 2( , ) → 1 → 1( , ) → 3( , ) d 0 H X TX/Y DX/Y H X N f H X TX/Y (iii) If f is non-degenerate then

∼ 0 DX/Y = H (X, N f )

1 =∼ 1( , ) DX/Y H X N f (iv) If f is smooth then ∼ 1 DX/Y = H (X, TX/Y )

1 =∼ 2( , ) DX/Y H X TX/Y Proof. It is clear that (ii) ⇒ (i). Moreover, since f non-degenerate implies TX/Y = 0, (iii) follows from the exact sequences (3.44)(b) and (3.44)(d). Similarly, f smooth implies N f = 0 and (iv) follows again from (3.44)(b) and (3.44)(d). Therefore it suffices to prove (ii). In the first sequence the homomorphism

1 DX/Y → H (X, TX ) is given by sending (v, t) → t; similarly, we have

0 ∗ H (X, f TY ) → DX/Y which sends u → (u, 0). The proof of exactness is left to the reader. In the second exact sequence the map

1 H (X, TX/Y ) → DX/Y 164 3 Examples of deformation functors is given by 1 Z (U, TX/Y ) * ψ → (0, Jψ) ∈ DX/Y The map 0 DX/Y → H (X, N f ) 0 sends (v, t) → Pv. Every element of H (X, N f ) is represented by some 0 ∗ 1 v ∈ C (U, f TY ) such that δv = df(t) for some t ∈ C (U, TX ).Wehave 2 df(δt) = δδt = 0sothatδt can be viewed as an element of Z (U, TX/Y ).This defines the map 0 2 H (X, N f ) → H (X, TX/Y ) The proof of exactness is left to the reader. The other two sequences are defined and their exactness is checked similarly.  Let f : X → Y be a morphism between algebraic schemes. Define functors of Artin rings: , : A → Def f/Y Def f/Y (sets) by isomorphism classes of Def / (A) = f Y deformations of f over A with fixed target isomorphism classes of locally trivial Def (A) = f/Y deformations of f over A with fixed target (A) = for all A in ob . Obviously, Def f/Y Def f/Y when X is nonsingular. We will consider the locally trivial case. The main general result about Def f/Y is the follow- ing:

Theorem 3.4.8. Let f : X → Y be a morphism of algebraic schemes with X pro- jective. Then Def f/Y has a formal semiuniversal deformation. Its tangent space is 1 DX/Y and DX/Y is an obstruction space (see Definition 3.4.6).

Proof. Let’s check the conditions of Theorem 2.3.2. Def f/Y trivially satisfies condi- tion H0. Consider a diagram in A: A A

A ¯ with A → A a small extension and let A = A ×A A .Let

( f , f ) ∈ Def / (A ) × ( ) Def / (A ) A A f Y Def f/Y A f Y where

f A : X → Y × Spec(A ), f A : X → Y × Spec(A ) ¯ X¯ ¯ Since DefX satisfies H there is a deformation of X over A such that X =∼ X¯ × ( ), X =∼ X¯ × ( ) Spec(A¯) Spec A Spec(A¯) Spec A 3.4 Morphisms 165 and we have a commutative diagram:

f X −→A Y × Spec(A ) ↓ X → X¯

↓ f A

Y × Spec(A ) → Y × Spec(A¯) By the universal property of the fibred sum we obtain a morphism : X¯ → × ( ¯) f A¯ Y Spec A which pulls back over Spec(A ) and Spec(A ) to f A and f A respectively. Therefore → ( , ) f A¯ f A f A under the map ¯ α : Def / (A) → Def / (A ) × ( ) Def / (A ) f Y f Y Def f/Y A f Y and this proves H¯ . = = [ ] : X¯ → × ( ¯) Let A k and A k and suppose that f A¯ Y Spec A and

˜ : X˜ → × ( ¯) ( ¯) ( , ) α f Y Spec A are elements of Def f/Y A mapped to f A f A by .Bythe universal property of X¯ we have a morphism X¯ → X˜ which, being an isomorphism mod , is an isomorphism. Moreover, we have a commutative diagram X¯ → X˜ ↓ Y × Spec(A¯) ˜ ( ¯) α and therefore f A¯ and f define the same element of Def f/Y A ; this implies that is bijective in this case and H holds.

In order to describe the tangent space to Def f , consider a first-order locally trivial deformation X → X

↓ f ↓ F

Y → Y × Spec(k[ ])

↓↓ψ

Spec(k) → Spec(k[ ]) Then the deformation X → X

↓↓ψ F (3.45)

Spec(k) → Spec(k[ ]) 166 3 Examples of deformation functors is locally trivial. Choose an affine open cover U ={Ui }i∈I of X and, for each index i,let θ : × ( [ ]) → X i Ui Spec k |Ui be an isomorphism of deformations. The composition

Fi := Fθi : Ui × Spec(k[ ]) → Y × Spec(k[ ]) := is a deformation wfdat of fi f|Ui and therefore it corresponds to an element v ∈ Γ( , ∗ ) = Γ( , ∗ ) i Ui fi TY Ui f TY by Proposition 3.4.2. Therefore we get an ele- 0 ∗ ment v ={vi }∈C (U, f TY ). Restricting to Uij we have:

θ −1θ = Fi|Uij i j Fj|Uij (3.46) θ −1θ ∈ Γ( , ) and, by Lemma 1.2.6, i j corresponds to a section tij Uij TX ; the collec- 1 tion {tij} is an element t ∈ Z (U, TX ) which defines the Kodaira–Spencer class of the deformation (3.45). The identity (3.46) means that v j − vi = d(tij) or, equiva- lently, that the pair (v, t) satisfies δv = df(t). The pair (v, t) is defined up to a choice of the trivializations θi or, equivalently, up to an element of the form (df(w), δw), 0 w ∈ C (U, TX ). Similarly, if we replace F by an isomorphic deformation of the form F = Fσ , where σ : X × Spec(k[ ]) → X × Spec(k[ ]) is an automorphism of the trivial deformation of X, we obtain the same element of DX/Y . Conversely, suppose given an element of DX/Y , represented by a pair 0 ∗ 1 (v, t) = ({vi }, {tij}) ∈ C (U, f TY ) × Z (U, TX ) such that δv = df(t). The class ¯ 1 t ∈ H (X, TX ) defines a first-order deformation of X

X → X

↓↓

Spec(k) → Spec(k[ ]) which is locally trivial. The 1-cocycle t ={tij} defines local trivializations θ : × ( [ ]) → X i Ui Spec k |Ui and each vi defines a deformation wfdat Fi of fi and by composition we get a mor- phism: θ−1 X −→i × ( [ ]) −→Fi × ( [ ]) |Ui Ui Spec k Y Spec k θ −1θ = By construction Fi|Uij i j Fj|Uij and therefore θ −1 = θ −1 Fi|Uij i Fj|Uij j 3.4 Morphisms 167

θ −1 This means that the morphisms Fi i patch together and define a morphism F : X → Y × Spec(k[ ]). This obviously gives rise to a first-order deformation of f . It is a straightforward task to verify that the correspondences F →¯v and v¯ → F ( [ ]) =∼ are the inverse of each other. Therefore Def f/Y k DX/Y and in particular H f holds. Now we want to prove the assertion about obstructions to lifting deformations. We have to show that for every locally trivial infinitesimal deformation

F X −→ Y × S

S of f with target Y over S = Spec(A) there is a natural map : ( , ) → 1 oF/Y Exk A k DX/Y such that, for a given extension

e : 0 → k → A˜ → A → 0 ˜ we have oF/Y (e) = 0 if and only if F has a lifting to A which is a locally trivial deformation of f with target Y . Choose U ={Ui }i∈I an affine open cover of X and trivializations θ : × ( ) → X i Ui Spec A |Ui 1 ∗ Since H (Ui , f TY ) = 0 for each i ∈ I the morphism

Fi := Fθi : Ui × Spec(A) → Y × Spec(A) has a lifting as a deformation of fi wfdat:

Fi Ui × Spec(A) −→ Y × Spec(A)

˜ ˜ Fi ˜ Ui × Spec(A) −→ Y × Spec(A)

If we restrict to Uij we have the identity: θ = Fi|Uij ij Fj|Uij where we have written θ := θ −1θ : × ( ) → × ( ) ij i j Uij Spec A Uij Spec A 168 3 Examples of deformation functors

Let ˜ ˜ ˜ θij : Uij × Spec(A) → Uij × Spec(A) θ × ( ) ˜ θ˜ ˜ be an automorphism which restricts to ij on Uij Spec A . Then Fi|Uij ij and Fj|Uij Γ( , ∗ ) are both liftings of Fj|Uij. Therefore, since Uij f TY acts faithfully and transi- tively on the set of such liftings (Proposition 3.4.2 (iii)), there is a ζ ∈ Γ( , ∗ ) ˜ ˜ θ˜ ζ ={ζ }∈C1(U, ∗ ) ij Uij f TY which carries Fj|Uij into Fi|Uij ij;set ij f TY . Now let θ˜ = θ˜ θ˜ θ˜−1 : × ( ˜) → × ( ˜) ijk ij jk ik Uijk Spec A Uijk Spec A ˜ Since θijk restricts to the identity on Uijk × Spec(A), by Lemma 1.2.6 it corresponds to an sijk ∈ Γ(Uijk, TX ); by construction s := {sijk} is a 2-cocycle, i.e. it is an 2 element of Z (U, TX ), and the pair (ζ, s) satisfies

δ(ζ) = df(s)

( ) = (ζ, ) ∈ 1 Define oF/Y e s DX/Y . This definition is well posed because a different ˜ 1 choice of the θij’s will replace (ζ, s) by (ζ +df(u), s +δu) for some u ∈ C (U, TX ). 1 If oF/Y (e) = 0 then (ζ, s) = (df(u), δu) for some u ∈ C (U, TX ). The condi- tion s = δu means that for the deformation X → X ξ : ↓↓ Spec(k) → Spec(A)

2 ˜ we have oξ (e) = s = 0 ∈ H (X, TX ) so that ξ has a lifting to Spec(A). Such a lifting X → X → X˜ ξ˜ : ↓↓↓ Spec(k) → Spec(A) → Spec(A˜) ˜ ˜ ˜ ˜ is defined by an appropriate choice of the automorphisms θij such that θijθ jk = θik. ˜ ˜ The condition ζ = df(u) means that the liftings Fi and Fj can be chosen so that

˜ θ˜ = ˜ Fi|Uij ij Fj|Uij and this means that they patch together to define a lifting F˜ of F. Conversely, if such a lifting exists then one shows in the same way that oF/Y (e) = 0.  ˜ oF/Y (e) is called the obstruction to lifting F to A as a deformation with tar- get Y . The notions of obstructed/unobstructed deformation, obstructed/unobstructed morphism with fixed target can be given as usual. One can give the notion of rigid morphism with fixed target in an obvious way. We have the following:

Corollary 3.4.9. Under the hypotheses of Theorem 3.4.8 we have: = 1 = (i) If DX/Y 0 then f is rigid as a morphism with fixed target. If DX/Y 0 then f is unobstructed as a morphism with fixed target. 3.4 Morphisms 169

(ii) If f is non-degenerate then in the statement of 3.4.8 we can replace DX/Y and 1 0( , ) 1( , ) DX/Y by H X N f and H X N f respectively. 1 (iii) If f is smooth then in the statement of 3.4.8 we can replace DX/Y and DX/Y by 1 2 H (X, TX/Y ) and H (X, TX/Y ) respectively. Proof. The proof is immediate in view of Lemma 3.4.7. 

Remarks 3.4.10. (i) If f is a closed embedding of projective nonsingular algebraic schemes then N f = NX/Y and Theorem 3.4.8 asserts that first-order deformations, resp. obstructions, of f with target Y coincide with first-order deformations, resp. obstructions, of X in Y . This is because, more generally, every infinitesimal defor- mation of f with target Y is an infinitesimal deformation of X in Y : a proof is given in Note 3. (ii) If f : X → Y is a morphism between algebraic varieties with X projective, and if (3.42) is a family of locally trivial deformations of f with target Y , then, using Theorem 3.4.8, we can define a linear map

TS,s → DX/Y which associates to a tangent vector t : Spec(k[ ]) → S at s the element of DX/Y corresponding to the first-order deformation obtained by pulling back F by t.This mapiscalledthecharacteristic map of the family F.

Given a morphism f : X → Y between algebraic varieties with X projective we have a natural morphism of functors Φ : → f Def f/Y DefX called the forgetful morphism which associates to a locally trivial deformation (3.42) over S = Spec(A), A in ob(A), the family of deformations of X obtained by forget- ting the morphism F. This morphism of functors generalizes the analogous forgetful morphism defined for the local Hilbert functor in Subsection 3.2.3. We have the fol- lowing generalization of Proposition 3.2.9.

Proposition 3.4.11. Let f : X → Y be a morphism between algebraic varieties with X projective, and let Φ : → f Def f/Y DefX be the forgetful morphism. Then (i) 1 dΦ f : DX/Y → H (X, TX ) is the map occurring in the exact sequence (3.44)(a). (ii) The map 1 → 2( , ) DX/Y H X TX

occurring in the exact sequence (3.44)(c) is an obstruction map for Φ f . 170 3 Examples of deformation functors

(iii) Assume that 1 ∗ H (X, f TY ) = 0

Then Φ f is smooth. Proof. The proofs of (i) and (ii) are straightforward. (iii) The hypothesis and the exact sequences (3.44)(a) and (3.44)(c) imply that 1 DX/Y → H (X, TX ) is surjective and

1 → 2( , ) DX/Y H X TX is injective Now, using parts (i) and (ii), the conclusion follows from Proposition 2.3.6.  Corollary 3.4.12. (i) Let f : X → Y be a non-degenerate morphism of algebraic schemes with X projective. Then 0 1 dΦ f : H (X, N f ) → H (X, TX ) and 1 2 o(Φ f ) : H (X, N f ) → H (X, TX ) are the coboundary maps coming from the exact sequence ∗ 0 → TX → f TY → N f → 0 (ii) Let f : X → Y be a smooth morphism of projective nonsingular algebraic schemes. Then 1 1 dΦ f : H (X, TX/Y ) → H (X, TX ) and 2 2 o(Φ f ) : H (X, TX/Y ) → H (X, TX )

are the maps induced by the natural inclusion of sheaves TX/Y ⊂ TX . Proof. The corollary is just a special case of the proposition. 

Examples 3.4.13. (i) Let m ≥ 0 be an integer, Fm = IP (Fm) where F is the locally free rank two sheaf on IP 1: F = O ( ) ⊕ O m IP 1 m IP 1 1 and let π : Fm → IP be the projection. Then we have 1( ,π∗ ) = H Fm TIP 1 0 by an easy calculation using the Leray spectral sequence. Therefore, since π is smooth, we can apply Proposition 3.4.11 to conclude that Φπ is smooth. Moreover, since Fm is unobstructed as an abstract variety because 2( , ) = h Fm TFm 0 (see (B.13)), it follows from Proposition 2.2.5(iii) that π is unobstructed. 3.4 Morphisms 171

We can actually be more precise because we have an exact sequence of locally free sheaves on IP 1:

∨ → O 1 → F ⊗ F → π∗ 1 → 0 IP m m TFm /IP 0 which can be deduced easily from the exact sequence (4.28). Since

F ⊗ F ∨ ∼ O⊕2 ⊕ O( ) ⊕ O(− ) m m = m m

1 ( , 1 ) = − using the Leray spectral sequence we deduce that h Fm TFm /IP m 1for m ≥ 1. Therefore, recalling (B.12), we see that the map

1 1 ( , 1 ) → ( , ) H Fm TFm /IP H Fm TFm is not only surjective but it is actually an isomorphism. (ii) Let f : X → Y be a smooth family of projective curves of genus ≥ 2 with X a projective nonsingular surface and Y a projective nonsingular connected curve. 1 Assume that f is non-isotrivial (see Definition 2.6.9). Then H (X, TX/Y ) = 0 and therefore, by (i) and (iii) of Corollary 3.4.9, f is rigid as a morphism with fixed target. This theorem is due to Parshin [139] in char 0. For an exposition we refer the reader to Szpiro [178], where the theorem, and its generalization due to Arakelov [6], is proved without the restriction char(k) = 0. (iii) Let f : X → Y be an etale morphism of projective nonsingular schemes of ∗ dimension n. Then f is non-degenerate and df : TX → f TY is an isomorphism; therefore N f = 0 and f is rigid as a morphism with fixed target. If we only assume f to be non-degenerate, but not necessarily etale, then df degenerates on a divisor R ⊂ X (which is the divisor Dn−1(df) of Example 4.2.8), called the ramification divisor of f . N f is supported on R, and in general f is not rigid as a morphism with fixed target. For the case when X and Y are curves see Subsection 3.4.3. (iv) Let Y be a projective nonsingular variety, γ ⊂ Y a nonsingular closed subva- riety of pure codimension r ≥ 2 and π : X → Y the blow-up of Y with centre γ .Let −1 ∼ E = π (γ ) ⊂ X be the exceptional divisor. Then E = IP (Nγ/Y ) is a projective bundle over γ :letq : E → γ be the structure morphism. Then NE/ X = OE (E) and it is well known that the restriction of NE/ X to each fibre IP of q is OIP (−1). Therefore by the Leray spectral sequence of q we immediately deduce that

i h (E, NE/ X ) = 0 (3.47) for all i.WehaveTX/Y = 0 because TX/Y is a subsheaf of the locally free TX and is i supported on E; therefore π is non-degenerate. Since π∗OX = OY and R π∗OX = 0 for i ≥ 1, from the Leray spectral sequence we deduce that

i ∗ i ∗ i H (X,π TY ) = H (Y,(π∗π OX ) ⊗ TY ) = H (Y, TY ), i ≥ 0 (3.48) 172 3 Examples of deformation functors

We have an exact and commutative diagram of locally free sheaves on E: 00 ↓↓ ∗ 0 → TE/γ → TE → q Tγ → 0 ↓ ↓ ∗ 0 → TE/γ → TX|E → q TY → Nπ → 0 (3.49) ↓↓ ∗ 0 → NE/ X → q Nγ/Y → Nπ → 0 ↓↓ 00 In particular, we see that we have an exact sequence of locally free sheaves on E:

∗ 0 → NE/ X → q Nγ/Y → Nπ → 0 (3.50)

The verification of these facts is straightforward and it is left to the reader. 2 2 2 For example, let π : X = Bl[1,0,0] IP → IP be the blow-up of IP with centre the point [1, 0, 0]. From the exact sequence (3.50) we deduce that Nπ = OE (1). Therefore 0 i h (X, Nπ ) = 2, h (X, Nπ ) = 0, i ≥ 1 In particular, π is unobstructed. Moreover, 0( , ) = 0( 2, ⊗ I ) = h X TX h IP TIP 2 [1,0,0] 6 as can be easily checked using the Euler sequence. Therefore from the exact sequence 1 (3.43) we see that h (X, TX ) = 0, i.e. X is rigid.

3.4.3 Morphisms from a nonsingular curve with fixed target

Theorem 3.4.8 applies in particular to a morphism

ϕ : C → Y where C and Y are projective and nonsingular, C is a curve, and ϕ is not constant on each component of C. Consider the exact sequence

dϕ ∗ 0 → TC −→ ϕ TY → Nϕ → 0

The vanishing divisor (see Example 4.2.8 page 200)

Z := D0(dϕ) of dϕ is called the ramification divisor of ϕ;theindex of ramification of ϕ at p ∈ C is the coefficient of p in Z. ϕ is unramified if and only if Z = 0. The homomorphism dϕ extends to a homomorphism 3.4 Morphisms 173

∗ TC (Z) → ϕ TY ¯ whose cokernel we denote by Nϕ; it is locally free. We have ¯ Nϕ = Nϕ/Hϕ where Hϕ ⊂ Nϕ is the torsion subsheaf; it is supported on Z. The following com- mutative and exact diagram summarizes the situation: 0 ↓ Hϕ ↓ dϕ ∗ 0 → TC −→ ϕ TY → Nϕ → 0 (3.51) ↓↓ ∗ ¯ 0 → TC (Z) → ϕ TY → Nϕ → 0 ↓ 0 We obtain: ∗ χ(Nϕ) = χ(ϕ TY ) + 3g − 3 (3.52) Example 3.4.14. Assume that C is connected of genus g, that Y is a projective con- nected nonsingular curve of genus γ , and that ϕ has degree d; then ¯ Nϕ = 0, Nϕ = Hϕ = OZ ∗ where O(Z) = ϕ (TY ) ⊗ KC , so that 0 χ(Nϕ) = h (Nϕ) = deg(Z) = 2[g − 1 + (1 − γ)d] and ϕ is unobstructed because h1(Nϕ) = 0. This corresponds to the fact that the deformations of ϕ leaving Y fixed are obtained by varying the branch points of ϕ. Note that ϕ is rigid as a morphism with fixed target if g ≥ 2 and Z = 0, i.e. if it is unramified.

∗∗∗∗∗∗ Assume now that ϕ : C → S is a nonconstant morphism from an irreducible projective nonsingular curve C of genus g to a projective nonsingular surface S and that ϕ is birational onto its image; let Γ = ϕ(C) ⊂ S. Then we have a commutative and exact diagram: 0 ↓ dϕ ∗ 0 → TC −→ ϕ TS → Nϕ → 0 ↓↓j ∗ ∗ ∗ 0 → ϕ TΓ → ϕ TS → ϕ NΓ/S 174 3 Examples of deformation functors ∗ ¯ Since ϕ NΓ/S is invertible the homomorphism j factors through Nϕ and the above diagram gives rise to the following:

0 ↓ ∗ ¯ 0 → TC (Z) −→ ϕ TS → Nϕ → 0 ↓↓ → ϕ∗ → ϕ∗ → ϕ∗ → 0 TΓ TS NΓ/S 0 ϕ = [ → 1 ] where Z is the ramification divisor of and NΓ/S ker NΓ/S TΓ is the equi- singular normal sheaf (see also § 4.7). This diagram implies the following isomor- phisms:

( ) =∼ ϕ∗ , ¯ =∼ ϕ∗ , H =∼ [ → ϕ∗( )]=: TC Z TΓ Nϕ NΓ/S ϕ coker TC TΓ Nϕ¯ where we have denoted by ϕ¯ : C → Γ the morphism induced by ϕ. In particular, ϕ∗ ϕ∗ TΓ and NΓ/S are invertible and

ϕ [ ( )] =∼ ⊗ ϕ O ,ϕ¯ =∼ ⊗ ϕ O ∗ TC Z TΓ ∗ C ∗ Nϕ NΓ/S ∗ C On Γ we have a natural exact sequence:

0 → OΓ → ϕ∗OC → t → 0 where t is a torsion sheaf supported on the singular locus of Γ . Since NΓ/S is inver- tible the homomorphism NΓ/S → NΓ/S ⊗ ϕ∗OC is injective and it follows that we have an exact sequence → → ϕ ¯ → ⊗ → 0 NΓ/S ∗ Nϕ NΓ/S t 0  (3.53) ⊗ ϕ O NΓ/S ∗ C This sequence implies in particular: 0( ) ≤ 0( ¯ ) ≤ 0( ) h NΓ/S h Nϕ h Nϕ

1( ) ≥ 1( ¯ ) = 1( ) h NΓ/S h Nϕ h Nϕ Lemma 3.4.15. If the singularities of Γ are nodes and ordinary cusps then ⊗ = NΓ/S t 0; equivalently, =∼ ϕ ¯ NΓ/S ∗ Nϕ In particular, if Γ has only nodes as singularities then

=∼ ϕ NΓ/S ∗ Nϕ 3.4 Morphisms 175

Proof. The exact sequence (3.53) can be embedded in the following exact and com- mutative diagram: 00 ↓↓ → → ϕ ¯ → ⊗ → 0 NΓ/S ∗ Nϕ NΓ/S t 0 ↓↓a ↓ d 0 → NΓ/S → NΓ/S ⊗ ϕ∗OC → NΓ/S ⊗ t → 0 ↓↓↓c 1 b 1 1 0 → TΓ −→ TΓ ⊗ ϕ∗OC → TΓ ⊗ t → 0 ↓↓ ↓ 00 0 The arrow a is injective because it is a nonzero homomorphism of torsion free rank one sheaves. Because of the assumptions made on the singularities, at each singular ∼ point p ∈ Γ we have tp = ϕ∗OC,p/OΓ,p = k. Therefore the arrow c is an isomor- ∼ ∼ 1 phism because NΓ/S ⊗ t = t = TΓ ⊗ t. Thus d = 0. The arrow b is injective because at each singular point p ∈ Γ we have 2 1 ∗ 1 ∼ k if p is a node (TΓ ⊗ ϕ∗O ) = ϕ∗ϕ (TΓ, ) = C p p k3 if p is a cusp (proved by easy local computation) while k if p is a node T 1 =∼ Γ,p k2 if p is a cusp (recall Example 3.1.4). The conclusion now follows from the “Snake Lemma”. If Γ ¯ = =∼ ϕ  has only nodes then Nϕ Nϕ and we deduce that NΓ/S ∗ Nϕ. ⊗ = Γ It is possible to show that conversely, if NΓ/S t 0 then the singularities of are nodes and ordinary cusps (see [68]). If S = IP 2 then, letting L = ϕ∗O(1), d = deg(L), from the Euler sequence restricted to C: → O → ⊕3 → ϕ∗ → 0 C L TIP 2 0 χ(ϕ∗ ) = + − we deduce TIP 2 3d 2 2g and from (3.52) χ(Nϕ) = 3d + g − 1 (3.54)

By Corollary 3.4.9 the unobstructedness of ϕ is related to the vanishing of H 1(C, Nϕ). From (3.51) we see that ( ) = (ϕ∗ ) − ( ) = + − deg Nϕ c1 TIP 2 deg TC 3d 2g 2 and that 1 1 ¯ h (Nϕ) = h (Nϕ) But ¯ deg(Nϕ) = c1(Nϕ) − deg(Z) ¯ and therefore Nϕ is a nonspecial line bundle whenever deg(Z)<3d. We can there- fore state the following result: 176 3 Examples of deformation functors

Proposition 3.4.16. Let ϕ : C → IP 2 be a morphism from an irreducible projective nonsingular curve C of genus g, birational onto its image. Let d = deg(ϕ∗O(1)), and let Z be the ramification divisor of ϕ. Then 0 h (C, Nϕ) ≥ 3d + g − 1 If deg(Z)<3d then ϕ is unobstructed and the above inequality is an equality. In particular, if ϕ(C) is a plane curve having nodes and cusps as its only singu- larities and the number κ of cusps satisfies κ<3d then 0 1 h (C, Nϕ) = 3d + g − 1, h (C, Nϕ) = 0

3.4.4 Deformations of a closed embedding

The deformation theory of morphisms is more subtle if we want to allow both the domain and the target to deform nontrivially. In this subsection we will address this case, considering only the simplest situation of a closed embedding. Let j : X ⊂ Y be a closed embedding of algebraic schemes. If

X −→J Y

S where S = Spec(A), A in ob(A), is an infinitesimal deformation of j then J is a closed embedding (see Note 3 at the end of of this section). It is obvious that an infinitesimal deformation of j with fixed target is nothing but a deformation of X in Y . Given another infinitesimal deformation of j:

X −→J Y

S over the same S = Spec(A), an isomorphism between them is a pair of isomorphisms of deformations: α : X → X ,β: Y → Y which make the diagram X −→J Y

↓ α ↓ β

X −→J Y commutative. 3.4 Morphisms 177

We define isomorphism classes of Def (A) = j deformations of j over A isomorphism classes of locally Def (A) = j trivial deformations of j over A for each A in ob(A). These are the functor of infinitesimal deformations of j and of locally trivial infinitesimal deformations of j respectively. The locally trivial infinitesimal deformations of a closed embedding are studied by means of a sheaf which we now introduce. Let’s now assume that Y is nonsingular and let IX ⊂ OY be the ideal sheaf of X. Let TY (X)⊂TY be the inverse image of TX ⊂ TY |X under the natural restriction homomorphism TY → TY |X . Then TY (X) is called the sheaf of germs of tangent vectors to Y which are tangent to X. We clearly have an inclusion IX TY ⊂ TY (X) such that

TX = TY (X)/IX TY and an exact sequence → ( )→ → → 0 TY X TY NX/Y 0 (3.55) ⊂ where NX/Y NX/Y is the equisingular normal sheaf of X in Y (introduced in Proposition 1.1.9, page 16). From the definition it follows that, for every open set U ⊂ Y , Γ(U, TY (X)) consists of those k-derivations D ∈ Γ(U, TY ) such that D(g) ∈ Γ(U, IX ) for every g ∈ Γ(U, IX ). We also have the following exact com- mutative diagram:

00 ↓↓ IX TY = IX TY

↓↓

→ ( )→ → → 0 TY X TY NX/Y 0 (3.56)

↓↓

→ → → → 0 TX TY |X NX/Y 0 ↓↓ 00

Of course, NX/Y is replaced by NX/Y in the case where X is nonsingular. We will describe locally trivial infinitesimal deformations of a closed embedding by means of the sheaf TY (X). 178 3 Examples of deformation functors

Proposition 3.4.17. Let j : X ⊂ Y be a closed embedding of projective algebraic schemes with Y nonsingular. Then Def j has a formal semiuniversal deformation. Its 1 2 tangent space is H (Y, TY (X)) and H (Y, TY (X)) is an obstruction space. ¯ Proof. The proof that Def j satisfies Schlessinger’s conditions H0, H and H is sim- ilar to the proof given in Theorem 3.4.8 and will be left to the reader. Since Y is projective the existence of a semiuniversal formal deformation will follow if we will 1( , ( )) prove the assertion about the tangent space of Def j , because H Y TY X is finite dimensional. Let U ={Ui }i∈I be an affine open cover of Y and

V ={Vi = X ∩ Ui }i∈I the induced affine open cover of X. Every locally trivial first-order deformation of j is obtained by gluing the trivial deformations

Vi ⊂ Vi × Spec(k[ ])

Ui ⊂ Ui × Spec(k[ ]) along Vij × Spec(k[ ]) and Uij × Spec(k[ ]). It is therefore necessary to describe the automorphisms of the trivial deformations

Vij ⊂ Vij × Spec(k[ ])

Uij ⊂ Uij × Spec(k[ ])

Every such automorphism Aij consists of a pair (θij,Θij) where

θij : Vij × Spec(k[ ]) → Vij × Spec(k[ ]) and Θij : Uij × Spec(k[ ]) → Uij × Spec(k[ ]) are automorphisms of deformations such that the following diagram commutes:

θij Vij × Spec(k[ ]) −→ Vij × Spec(k[ ])

Θij Uij × Spec(k[ ]) −→ Uij × Spec(k[ ]) θ = Θ Θ θ equivalently, such that ij ij|Vij. According to Lemma 1.2.6, ij and ij cor- respond to sections Dij ∈ Γ(Uij, TY ) and dij ∈ Γ(Vij, TX ) respectively such that Dij → dij when restricted to X. It follows that Dij ∈ Γ(Uij, TY (X)) and that to give Aij is the same as to give Dij. This said, the proof of the statement about tan- gent and obstruction spaces of Def j now proceeds in a straightforward way along the lines of the analogous proofs of 1.2.9 and of 1.2.12. We omit the details.  Also in this case we have the notions of obstructed (resp. unobstructed) deforma- tion, obstructed (resp. unobstructed) embedding, and of rigid embedding. It follows 3.4 Morphisms 179 from Proposition 3.4.17 that a closed embedding j : X ⊂ Y of projective nonsingu- lar varieties is rigid if and only if 1 H (Y, TY (X)) = 0

Let X −→J Y

S be a locally trivial deformation of j : X ⊂ Y parametrized by a pointed scheme (S, s). Then we can define a characteristic map 1 χJ : TS,s → H (Y, TY (X)) by associating to a tangent vector t : Spec(k[ ]) → S at s the element of 1 H (Y, TY (X)) corresponding to the first-order deformation of f obtained by pulling back J by t. Remark 3.4.18. Let j : X ⊂ Y be a closed embedding of projective schemes with Y nonsingular. If 0 H (Y, TY (X)) = 0 ( , ˆ) then Def j is prorepresentable. In fact, letting R u be the formal semiuniversal deformation of Def j , one can generalize Theorem 2.6.1 and its corollaries by intro- ducing an automorphism functor ˆ Autuˆ : AR → (sets) in an obvious way and proving that it is prorepresentable with tangent space 0 H (Y, TY (X)). This is called the space of infinitesimal automorphisms of j.The details of this straightforward generalization are left to the reader. Example 3.4.19. Let Y be a projective scheme, p ∈ Y a closed point and j :{p}⊂Y . Then TY ((p)) = IpTY I ⊂ O where p Y is the ideal sheaf of p. In this case Def j is the functor of locally trivial deformations of the pointed scheme (Y, p). If, for example, Y is a projective nonsingular connected curve of genus g then TY ((p)) = TY (−p) and we get: 0ifg = 0 1 h (Y, TY (−p)) = 1ifg = 1 3g − 2ifg ≥ 2 2 while h (Y, TY (−p)) = 0. Of course, we can generalize by considering a set of m distinct closed points { ,..., } :{ ,..., }→ p1 pm of Y , and the inclusion j p1 pm Y . Then Def j is the functor of locally trivial deformations of the m-pointed scheme (Y ; p1,...,pm). 180 3 Examples of deformation functors

3.4.5 Stability and costability

Whenever we have a locally trivial infinitesimal deformation X → Y

Spec(A) of a closed embedding j : X ⊂ Y of projective schemes we also have a deformation of X and a deformation of Y (both locally trivial):

X → X Y → Y ξ : ↓↓η : ↓↓ Spec(k) → Spec(A) Spec(k) → Spec(A)

This means that we have two forgetful morphisms of functors:

Φ −→Y Def j DefY ↓ ΦX

DefX The differentials and obstruction maps of these morphisms are described as follows.

Proposition 3.4.20. If j : X → Y is a closed embedding of projective schemes with Y nonsingular then (i) 1 1 dΦY : H (Y, TY (X)) → H (Y, TY ) and 2 2 o(ΦY ) : H (Y, TY (X)) → H (Y, TY )

are the maps induced in cohomology by the inclusion TY (X)⊂TY . (ii) 1 1 dΦX : H (Y, TY (X)) → H (X, TX ) and 2 2 o(ΦX ) : H (Y, TY (X)) → H (X, TX )

are the maps induced in cohomology by the restriction TY (X)→TX .

Proof. It is a straightforward consequence of the above analysis. Details are left to the reader. 

Remark 3.4.21. Let j : X ⊂ Y be a closed embedding of projective nonsingular schemes. Then there is a natural morphism of functors Y → HX Def j whose differential is easily seen to be the coboundary map 3.4 Morphisms 181

0 1 δ : H (X, NX/Y ) → H (Y, TY (X)) determined by the exact sequence (3.55). It follows from the proposition that

ker(dΦY ) = Im(δ) as expected, because deformations of X in Y are precisely those deformations of the embedding X ⊂ Y which induce the trivial deformation of Y . The proposition implies also that

β 1 1 ker(dΦX ) = Im[H (Y, IX TY ) −→ H (Y, TY (X))] where β is the map induced by the inclusion IX TY ⊂ TY (X). This kernel consists of the first-order deformations of j which induce the trivial deformation of X.If,in 0 particular, H (X, TX ) = 0 then

1 ker(dΦX ) = H (Y, IX TY )

Definition 3.4.22. If j : X ⊂ Y is a closed embedding of projective schemes then X is called stable in Y if the morphism of functors Φ : → Y Def j DefY is smooth. X is called costable in Y if the morphism of functors Φ : → X Def j DefX is smooth.

The notion of stability was introduced and studied in [104] for a compact com- plex submanifold of a complex manifold. As stated in [104], stability means that “no local deformation of Y makes X disappear”. Our definition of stability implies that every infinitesimal locally trivial deformation of Y is induced by a locally trivial de- formation of j. Costability implies that every infinitesimal locally trivial deformation of X is induced by a locally trivial deformation of j. The notion of costability has been introduced in [88].

Proposition 3.4.23. Let j : X ⊂ Y be a closed embedding of projective schemes, Y nonsingular. 1( , ) = ( ) (i) If H X NX/Y 0 then X is stable in Y . 2 (ii) If H (Y, IX TY ) = 0 then X is costable in Y . (iii) If X is nonsingular and is both stable and costable in Y then X is obstructed if and only if Y is obstructed (as abstract varieties).

Proof. (i) From the exact sequence (3.55) it follows that dΦY is surjective and that

2 2 H (Y, TY (X)) → H (Y, TY ) 182 3 Examples of deformation functors

is injective; but by Proposition 3.4.20(i) this last condition means that Def f is less obstructed than DefY and the conclusion follows from Proposition 2.3.6. (ii) The proof is similar using the exact cohomology sequence of the first column of diagram (3.56), Proposition 3.4.20 and Proposition 2.3.6. (iii) Since the morphisms of functors

ΦX ΦY DefX ←− Def j −→ DefY are both smooth we deduce that any one of the functors DefX , Def j , DefY is smooth if and only if the others are. 

Examples 3.4.24. (i) (Kodaira [104], Th. 5) Let Y be a projective nonsingular vari- ety, γ ⊂ Y a nonsingular closed subvariety and π : X → Y the blow-up of Y with centre γ .LetE = π −1(γ ) ⊂ X be the exceptional divisor; then

i h (E, NE/ X ) = 0 for all i (see (3.47)). From Proposition 3.4.23 we obtain that E is a stable subvariety of X. This is remarkable because γ has not been required to be stable in Y . (ii) Let X be a projective nonsingular algebraic surface and Z ⊂ X an irreducible nonsingular rational curve with self-intersection ν = Z 2. Then Z is stable in X if 1 ν ≥−1 because H (Z, NZ/ X ) = 0 in this case. On the other hand, if ν ≤−2 then in general Z is not stable in X. An example is provided by the negative section E 2 in the rational ruled surface Fm,form ≥ 2. In fact, E =−m and we have seen 1 in Example 1.2.2(ii) that there is a family f : W → A of deformations of Fm for which [E] does not extend to the other fibres W(t), t = 0, since they are isomorphic to Fn for some 0 ≤ n < m. This implies that E is not stable. (iii) Assume k = C.LetC be a projective irreducible nonsingular curve of genus g ≥ 3, and let α : C → JC be the Abel–Jacobi embedding of C into its jacobian variety. Then: 1 ∼ 1 • H (JC, OJC) = H (C, OC ) ([20], Lemma 11.3.1) 1 • TJC = H (C, OC ) ⊗ OJC.

Thus the restriction TJC → TJC|C induces isomorphisms

0 ∼ 0 ∼ 1 H (JC, TJC) = H (C, TJC|C ) = H (C, OC ) (3.57) 1 ∼ 1 ∼ 1 1 H (JC, TJC) = H (C, TJC|C ) = H (C, OC ) ⊗ H (C, OC )

In particular, in view of the second column of diagram (3.56), we have

0 1 H (JC, IC TJC) = 0 = H (JC, IC TJC) and therefore, taking cohomology of (3.56), we obtain the commutative and exact diagram: 3.4 Morphisms 183

Φ 0 1 d JC 1 1 H (NC/JC) → H (JC, TJC(C)) −→ H (JC, TJC) → H (NC/JC) → 0 ∩  (3.58) σ 0 1 1 1 H (NC/JC) → H (C, TC ) −→ H (C, TJC|C ) → H (NC/JC) → 0 which implies that 1 ∼ 1 H (JC, TJC(C)) = H (C, TC ) (3.59) and 2 2 H (JC, TJC(C)) ⊂ H (JC, TJC) (3.60) Since (3.60) is the obstruction map of

ΦJC : Defα → DefJC from Proposition 2.3.6 we deduce that Defα is smooth, being less obstructed than DefJC which is smooth (see Example 2.4.11(v), page 74). On the other hand, since (3.59) is the differential of the forgetful morphism

ΦC : Defα → DefC and Defα is smooth, we deduce by Corollary 2.3.7 that ΦC is smooth, i.e. C is costable in JC. 0 Note that H (JC, TJC(C)) = 0 by the first column of diagram (3.56): therefore α has no infinitesimal automorphisms and Defα is prorepresentable (Remark 3.4.18). It follows that ΦC is actually an isomorphism of functors (Remark 2.3.8). We can identify Defα with DefC and the differential 1 1 dΦJC : H (JC, TJC(C)) → H (JC, TJC) with the map σ in diagram (3.58). Therefore ΦJC is a closed embedding if and only if σ is injective. In view of the isomorphisms (3.57) σ is Serre-dual to the natural multiplication map: ∨ 0 0 0 σ : H (C, KC ) ⊗ H (C, KC ) → H (C, 2KC ) This map is surjective if and only if C is non-hyperelliptic: therefore in this case DefC is a closed smooth subfunctor of DefJC. In the natural decomposition 2 1 1 2 1 1 H (C, OC ) ⊗ H (C, OC ) = S H (OC ) ⊕ H (OC ) we have Im(σ ) ⊂ S2 H 1(O ) C 2 1 ∨ 2 1 because H (OC ) ⊂ ker(σ ). Therefore Im(dΦJC) is contained in S H (OC ) which is the space of first-order deformations of JC preserving the principal po- larization (compare with Example 3.3.13, page 151, and observe that in this case ¯ 1 1 Ω = H (OC ) = V by the Hodge decomposition of H (C, C)). The validity of the condition “σ injective” is called the infinitesimal Torelli theo- rem: thus it holds if and only if C is non-hyperelliptic. It is not difficult to show that α(C) is unobstructed in JC if and only if C is non-hyperelliptic (see [70], [113]). 184 3 Examples of deformation functors

The following result, due to Kodaira [104], gives the possibility of relating a local Hilbert functor with the deformation functor of an abstract variety. Proposition 3.4.25. Let Y be a projective nonsingular variety, γ ⊂ Y a closed non- singular subvariety of pure codimension r ≥ 2, and let π : X → Y be the blow-up of Y with centre γ . Then: (i) There is a natural isomorphism of functors

Y B : Hγ → Defπ/Y

In particular, Defπ/Y is prorepresentable. 1 (ii) Assume that H (Y, TY ) = 0, i.e. that Y is rigid. Then the forgetful morphism Φπ : Defπ/Y → DefX is smooth. Therefore we have a smooth morphism of Y functors BΦπ : Hγ → DefX . In particular, if γ is obstructed in Y then X is obstructed as an abstract variety. Proof. (i) We define Y B : Hγ → Defπ/Y by associating to a family of deformations

γA ⊂ Y × Spec(A) ↓ Spec(A) of γ in Y over A the blow-up π : := ( × ( )) → × ( ) A X A BlγA Y Spec A Y Spec A × ( ) γ O of Y Spec A along A. Note that, since γA is A-flat and we have the exact sequence on Y × Spec(A): → I → O ⊗ → O → 0 γA Y A γA 0 + Iγ A Ik /Ik 1 the sheaf A is -flat as well (Proposition A.2(VI)); moreover, γA γA is locally free over OY ⊗ A for all k ≥ 1 because γA is regularly embedded in Y × Spec(A) Ik A k ≥ by Lemma D.2.3. From this it is easy to deduce that γA is -flat for all 1 and therefore X = Ik A Proj γA k is A-flat by Proposition A.2(V). We leave to the reader to check the functoriality of B. The differential of B is the composition

0 ∼ 0 ∗ 0 dB : H (γ, Nγ/Y ) = H (E, q Nγ/Y ) → H (X, Nπ ) where the first map is the obvious isomorphism and the second one comes from the exact sequence (3.50); in a similar way, one describes the obstruction map of B as the one induced by the composition 3.4 Morphisms 185

1 ∼ 1 ∗ 1 H (γ, Nγ/Y ) = H (E, q Nγ/Y ) → H (X, Nπ ) deduced from the exact sequence (3.50). These facts can be easily verified by chasing i diagram (3.49). Since H (E, NE/ X ) = 0 for all i, we see that these maps are both bijective, and the conclusion follows. 1 ∗ (ii) From (3.48) it follows that H (X,π TY ) = 0. The exact sequence ∗ 0 → TX → π TY → Nπ → 0 and Proposition 3.4.20(i) imply that dΦπ is surjective and o(Φπ ) is injective. The conclusion is a consequence of Proposition 2.3.6. The last assertion is an obvious Y consequence of the fact that the composition Φπ B : Hγ → DefX is smooth.  By applying Proposition 3.4.25 to any obstructed nonsingular curve γ ⊂ IP 3 (e.g. the curve of degree 14 and genus 24 described in § 4.6) we obtain an exam- ple of obstructed projective variety of dimension 3. As an application we obtain the following result, which gives examples of obstructed surfaces.

Theorem 3.4.26 (Horikawa [88]). Let γ ⊂ IP 3 be an obstructed nonsingular curve, and X the blow-up of IP3 with centre γ .IfS ⊂ X is a sufficiently ample nonsingular surface then S is obstructed as an abstract variety.

Proof. By Proposition 3.4.23(iii) it is sufficient to show that S is both stable and 2 costable in X.Wehaveh (X, OX ) = 0 and, by the ampleness of S, 1 h (X, OX (S)) = 0 by Serre’s vanishing theorem. From the exact sequence

0 → OX → OX (S) → NS/ X → 0

1 we deduce that h (S, NS/ X ) = 0 and therefore S is stable in X. On the other hand, we have 2 H (X, TX (−S)) = 0 by Serre’s vanishing theorem again. Therefore S is costable in X as well.  We can immediately verify, using the adjunction formula, that the surfaces S constructed in the theorem are regular and with ample canonical class. For related examples see [29].

NOTES 1. The analysis of morphisms from a nonsingular curve is taken from [8]; see also [7]. For Lemma 3.4.15 see [180].

2. For the study of the functors Def f/Y and Def f under more general assumptions than those of Theorem 3.4.8 a more careful analysis of first-order deformations and obstructions is needed. We refer the reader to [148] and [150] for more information about this. The deforma- tion theory of closed embeddings is studied in [134] in the analytic category.

3. Let Φ X −→ Y

S 186 3 Examples of deformation functors be a commutative diagram of morphisms of algebraic schemes, with X and Y S-flat and Φ projective. Assume that Φo : X (o) → Y(o) is a closed embedding, for some k-rational point o ∈ S. Then there is an open neighbourhood U ⊂ S of o such that the restriction Φ(U) : X (U) → Y(U) is a closed embedding.

Proof. Let K = coker[OY → Φ∗(OX )].SinceΦ is projective, Φ∗(OX ) is a coherent sheaf and so is K. Moreover, K(o) = (0) because Φo is a closed embedding. It follows that there is an open subset U ⊂ S containing o such that K|Y(U) = (0).LetZ = Spec(Φ∗(OX )), h : Z → Y the induced S-morphism and Φ X −→ Y g  h Z the Stein factorization of Φ. Then it follows that h(U) : Z(U) → Y(U) is a closed em- bedding. Moreover, since g has connected fibres and is bijective over Z(o), it follows that, modulo shrinking U if necessary, g(U) : X (U) → Z(U) is an isomorphism. The conclusion follows.  4 The Hilbert schemes and the Quot schemes

Even though this book is centred around the theme of infinitesimal and local defor- mations, in this chapter we turn our attention to global deformations. We will intro- duce the Hilbert schemes and other related objects, which are important examples of parameter schemes for global families of deformations of algebro-geometric objects. They are used to describe and classify “extrinsic” deformations, i.e. deformations of objects within a given ambient space (e.g. closed subschemes of a given scheme). Their study is preliminary to the construction of “moduli schemes”. Moreover, they provide some of the most typical examples of constructions in algebraic geometry by the functorial approach. We will study some of their properties and consider a few applications of the local theory developed so far.

4.1 Castelnuovo–Mumford regularity

In this section we introduce the notion of m-regularity, also called Castelnuovo– Mumford regularity, and we prove its main properties. They will be needed for the construction of the Hilbert schemes and of the Quot schemes. Let m ∈ ZZ. A coherent sheaf F on IP r is m-regular if

H i (F(m − i)) = (0) for all i ≥ 1. Because of Serre’s vanishing theorem, every coherent sheaf F on IP r is m- regular for some m ∈ ZZ. The definition of m-regularity makes sense for a coherent sheaf on any projective scheme X endowed with a very ample line bundle O(1). For simplicity we will consider the case X = IP r only, leaving to the reader the obvious modifications of the statements and of the proofs in the general case. 188 4 Hilbert and Quot schemes

Proposition 4.1.1. If F is m-regular then: (i) The natural map

0 0 0 H (F(k)) ⊗k H (O(1)) → H (F(k + 1))

is surjective for all k ≥ m. (ii) H i (F(k)) = (0) for all i ≥ 1 and k ≥ m − i; in particular, F is n-regular for all n ≥ m. (iii) F(m), and therefore also F(k) for all k ≥ m, is generated by its global sections.

Proof. We prove (i) and (ii) by induction on r.Ifr = 0 there is nothing to prove. Assume r ≥ 1 and let H be a hyperplane not containing any point of Ass(F);it exists because Ass(F) is a finite set. Tensoring by F(k) the exact sequence:

0 → O(−H) → O → OH → 0 we get an exact sequence:

0 → F(k − 1) → F(k) → FH (k) → 0 where FH = F ⊗ OH . For each i > 0 we obtain an exact sequence

i i i+1 H (F(m − i)) → H (FH (m − i)) → H (F(m − i − 1)) which implies that FH is m-regular on H. It follows by induction that (i) and (ii) are true for FH . Let’s consider the exact sequence

i+1 i+1 i+1 H (F(m − i − 1)) → H (F(m − i)) → H (FH (m − i))

If i ≥ 0 the two extremes are zero (the right one by (ii) for FH , the left one by the m regularity of F), therefore F is (m + 1)-regular. By iteration this proves (ii). To prove (i) we consider the commutative diagram:

0 0 0 0 H (F(k)) ⊗k H (O(1)) −→ H (FH (k)) ⊗k H (OH (1)) u ↓ w ↓ t H 0(F(k)) → H 0(F(k + 1)) −→ H 0(F (k + 1)) v H The map u is surjective for k ≥ m because H 1(F(k − 1)) = (0); moreover, t is surjective for k ≥ m by (i) for FH . Therefore vw is surjective. It follows that H 0(F(k + 1)) is generated by Im(w) and by H 0(F(k)) for all k ≥ m.But H 0(F(k)) ⊂ Im(w) because the inclusion H 0(F(k)) ⊂ H 0(F(k + 1)) is multipli- cation by H. Therefore w is surjective. Let’s prove (iii). Let h  0 be such that F(m + h) is generated by its global sections. Then the composition

0 0 0 H (F(m)) ⊗k H (O(h)) ⊗k O → H (F(m + h)) ⊗k O → F(m + h) 4.1 Castelnuovo–Mumford regularity 189 is surjective because from (i) it follows that the first map is; we deduce that the composition

0 0 0 H (F(m)) ⊗k H (O(h)) ⊗k O(−h) → H (F(m)) ⊗k O → F(m) is also surjective, hence the second map is surjective too.  Note that if F is m-regular then the graded k[X1,...,Xr ]-module 0 Γ∗(F) := H (F(k)) k∈ZZ can be generated by elements of degree ≤ m. In fact, this is equivalent to the surjec- tivity of the multiplication maps

0 0 0 H (F(m)) ⊗k H (O(h)) → H (F(m + h)) for h ≥ 1, and follows from part (i) of the proposition. In particular, if an ideal sheaf I ⊂ OIP r is m-regular then the homogeneous ideal

I = Γ∗(I) ⊂ k[X0,...,Xr ] is generated by elements of degree ≤ m. Note also that in proving 4.1.1 we have proved the following:

Proposition 4.1.2. If F is m-regular and

0 → F(−1) → F → G → 0 is an exact sequence, then G is m-regular.

Conversely, we have the following:

Proposition 4.1.3. Let

0 → F(−1) → F → G → 0 be an exact sequence of coherent sheaves on IPr , and assume that G is m-regular. Then: (i) H i (F(k)) = 0 for i ≥ 2 and k ≥ m − i

(ii) h1(F(k − 1)) ≥ h1(F(k)) for k ≥ m − 1

(iii) H 1(F(k)) = 0 for k ≥ (m − 1) + h1(F(m − 1))

In particular, F is m + h1(F(m − 1))-regular. 190 4 Hilbert and Quot schemes

Proof. (i) In the exact sequence − H i 1(G(k)) → H i (F(k − 1)) → H i (F(k)) → H i (G(k)) the first and the last group are zero for i ≥ 2 and k ≥ m − (i − 1). Therefore

H i (F(m − i)) =∼ H i (F(m − i + 1)) =∼ H i (F(m − i + 2)) =∼ ···

From Serre’s vanishing theorem we get H i (F(m − i + h)) = 0 for all h  0 and (i) follows. (ii) For k ≥ m − 1 we have the exact sequence v 0 → H 0(F(k−1)) → H 0(F(k)) −→k H 0(G(k)) → H 1(F(k−1)) → H 1(F(k)) → 0 which implies (ii). (iii) Assume vk surjective, and consider the commutative diagram: v ⊗id H 0(F(k)) ⊗ H 0(O(1)) −→k H 0(G(k)) ⊗ H 0(O(1)) ↓↓wk v + H 0(F(k + 1)) −→k 1 H 0(G(k + 1))

Since wk is surjective for k ≥ m, we have that vk+1 is surjective too. Therefore H 1(F(k − 1)) =∼ H 1(F(k)) =∼ H 1(F(k + 1)) =∼ ···=∼ 0

1 1 If vk is not surjective then h (F(k − 1)) > h (F(k)). Therefore the function k → h1(F(k)) is strictly decreasing for k ≥ m − 1, and this implies (iii).  The following is a useful characterization of m-regularity.

Theorem 4.1.4. A coherent sheaf F on IPr is m-regular if and only if it has a reso- lution of the form:

···→O(−m − 2)b2 → O(−m − 1)b1 → O(−m)b0 → F → 0 (4.1) for some nonnegative b0, b1, b2,.... Proof. Assume that F has a resolution (4.1) and let b0 R1 = ker O(−m) → F b j− b j− R j = ker O(−m − j + 1) 1 → O(−m − j + 2) 2 j = 2,...,r

br+1 Rr+1 = O(−m − r) Using the short exact sequences:

b0 0 → R1(m − i) → O(−i) → F(m − i) → 0

b j−1 0 → R j (m − i) → O(−i − j + 1) → R j−1(m − i) → 0

br+1 br 0 → O(−i − r − 1) → O(−i − r) → Rr (m − i) → 0 4.1 Castelnuovo–Mumford regularity 191 we see that for all 1 ≤ i ≤ r we have: i ∼ i+1 ∼ H (F(m − i)) = H (R1(m − i)) = ···

∼ r ∼ r+1 ···= H (Rr−i (m − i)) = H (Rr−i+1(m − i)) = (0) and F is m-regular. Assume conversely that F is m-regular. By 4.1.1(iii) we have an exact sequence:

b0 0 → R1 → O(−m) → F → 0 0 with b0 = h (F(m)), which defines R1. If R1 = 0 we are done; if R1 = 0 from the sequences:

b0 0 → R1(m − i + 1) → O(−i + 1) → F(m − i + 1) → 0 we deduce that

i ∼ i−1 H (R1(m − i + 1)) = H (F(m − i + 1)) 1 ≤ i ≤ r hence R1 is (m + 1)-regular. Applying the same argument to R1 we find an exact sequence: b1 b0 0 → R2 → O(−m − 1) → O(−m) → F → 0 with R2 (m + 2)-regular. This process can be repeated and gives a resolution as required.  We will now turn to the problem of finding numerical criteria of m-regularity for a coherent sheaf F on IP r . Consider a sequence σ1,...,σN of N sections of OIP r (1).WewillcallitF- regular if the sequences of sheaf homomorphisms induced by multiplication by σ1,...,σN : σ1 0 → F(−1) −→ F → F1 → 0

σ2 0 → F1(−1) −→ F1 → F2 → 0 etc., are exact. By choosing σi+1 not containing any point of Ass(Fi ) one shows that F- regular sequences of any length exist. Therefore any general N-tuple (σ1,...,σN ) ∈ 0 N H (OIP r (1)) is an F-sequence.

r Definition 4.1.5. Let F be a coherent sheaf on IP , and (b) = (b0, b1,...,bN ) a sequence of nonnegative integers such that N ≥ dim[Supp(F)].We will call F a (b)-sheaf if there exists an F-regular sequence σ1,...,σN of sec- 0 tions of OIP r (1) such that h (Fi (−1)) ≤ bi ,i = 0,...,N where F0 = F, and Fi = F/(σ1,...,σi )F(−1), i ≥ 1.

Note that from the definition it follows immediately that if F is a (b)-sheaf then F1 is a (b1,...,bN )-sheaf. 192 4 Hilbert and Quot schemes

It is likewise clear that for every coherent sheaf F on IP r there is a sequence (b) such that F is a (b)-sheaf. Moreover, a subsheaf of a (b)-sheaf is easily seen to be a (b)-sheaf. For example, every ideal sheaf I ⊂ OIP r is a (0)-sheaf, because OIP r is clearly a (0)-sheaf. Lemma 4.1.6. Let 0 → F(−1) → F → G → 0 be an exact sequence of coherent sheaves on IPr .If r k + i χ(F(k)) = a i i i=0 then − r 1 k + i χ(G(k)) = a + i 1 i i=0 The proof is left to the reader.

Proposition 4.1.7. Let F be a (b)-sheaf, let s = dim[Supp(F)] and s k + i χ(F(k)) = a i i i=0 Then:  ≥− 0(F( )) ≤ s k+i (i) For each k 1 we have h k i=0 bi i . (ii) as ≤ bs and F is also a (b0,...,bs−1, as)-sheaf. 0 0 Proof. (i) By induction on s.Ifs = 0 then a0 = h (F) = h (F(−1)) ≤ b0 and the conclusion is obvious. Assume s ≥ 1. We have an exact sequence

0 → F(−1) → F → F1 → 0 with F1 a (b1,...,bN )-sheaf and dim[Supp(F1)]=s − 1. Then:

0 0 0 h (F(k)) − h (F(k − 1)) ≤ h (F1(k)) and s−1 + 0 k i h (F (k)) ≤ b + 1 i 1 i i=0 0 by the inductive hypothesis. Since h (F(−1)) ≤ b0 by induction on k ≥−1weget the conclusion. (ii) By Lemma 4.1.6 and by induction on s we get as ≤ bs and F1 is a (b1,...,bs−1, as)-sheaf. The conclusion follows.  4.1 Castelnuovo–Mumford regularity 193

Definition 4.1.8. The following polynomials, defined by induction for each integer r ≥−1:

P−1 = 0  ( ,..., )− + ( ,..., ) = ( ,..., ) + r Pr−1 X1 Xr 1 i Pr X0 Xr Pr−1 X1 Xr i=0 Xi i are called (b)-polynomials. One immediately sees that

Pr (X0,...,Xt , 0,...,0) = Pt (X0,...,Xt ) (4.2) for each t < r. The following theorem gives a numerical criterion of m-regularity. r Theorem 4.1.9. Let F be a (b)-sheaf on IP , with (b) = (b0, b1,...,bN ), and let r k + i χ(F(k)) = a i i i=0 be its Hilbert polynomial. Let (c0,...,cr ) be a sequence of integers such that ci ≥ bi − ai ,fori = 0,...,r, and m = Pr (c0,...,cr ). Then m ≥ 0 and F is m-regular. In particular, F is Ps−1(c0,...,cs−1)-regular if s = dim[Supp(F)]. Proof. By induction on r.Ifr = 0 then m = 0 and F is n-regular for every n ∈ ZZ, so the theorem is true in this case. Assume r ≥ 1. We have an exact sequence:

0 → F(−1) → F → F1 → 0 r−1 with F1 a (b1,...,bN )-sheaf supported on IP . From Lemma 4.1.6 and from the inductive hypothesis we deduce that n ≥ 0 and F1 is n-regular, where n = 1 Pr−1(c1,...,cr ). From 4.1.3 we deduce that F is [n + h (F(n − 1)]-regular and hi (F(n − 1)) = 0fori ≥ 2. Therefore: r n − 1 + i h1(F(n − 1)) = h0(F(n − 1)) − χ(F(n − 1)) ≤ (b − a ) i i i i=0  F + r n−1+i by 4.1.7(i). It follows that is n i=0 ci i -regular, by 4.1.1(ii). This proves the first assertion. The last assertion follows from 4.1.7(ii) and from (4.2).  Note that the integer m in the statement of the theorem depends on the coeffi- cients of the Hilbert polynomial of F as well as on the integers bi . In the special case when F is a sheaf of ideals we can determine an m for which F is m-regular which depends only on the Hilbert polynomial of F, as stated in the next corollary.

Corollary 4.1.10. For each r ≥ 0 there exists a polynomial Fr (X0,...,Xr ) such that every sheaf of ideals I ⊂ OPr having the Hilbert polynomial r k + r χ(I(k)) = a i i i=0 is m-regular, where m = Fr (a0,...,ar ), and m ≥ 0. 194 4 Hilbert and Quot schemes

Proof. It suffices to observe that I is a (0)-sheaf. Therefore the corollary follows from Theorem 4.1.9, taking Fr (X0,...,Xr ) = Pr (−X0,...,−Xr ). 

NOTES 1. Corollary 4.1.10 is in general false for coherent sheaves which are not sheaves of ideals. An example from [130] is F = O ( ) ⊕ O (− ) IP 1 k IP 1 k In fact χ(F) = 2 is independent of k but the least m such F is m-regular is |k|. 2. If I is the sheaf of ideals of the closed subscheme X ⊂ IP r and I is m-regular with m ≥ 0, then OX is (m − 1)-regular. Conversely, if OX is (m − 1)-regular and the restriction map 0 r 0 H (IP , OIP r (m − 1)) → H (X, OX (m − 1)) is surjective, then I is m-regular. This follows from the exact sequences

0 → I(k) → OIP r (k) → OX (k) → 0 k ≥ m − 1. 3. The notion of m-regularity is related to that of bounded collection of sheaves, important in moduli theory. A collection of coherent sheaves {Fj } j∈J on a projective scheme X is said to be bounded if there is an algebraic scheme S and a coherent sheaf F on X × S such that for each j ∈ J there is a closed point s ∈ S such that Fj is isomorphic to the sheaf F(s) = F|X×{s}.One also says that the collection {Fj } j∈J is bounded by the sheaf F on X × S. For details we refer to [98]. 4. The notion of Castelnuovo–Mumford regularity has been introduced in [130]. Casteln- uovo studied the properties of m-regularity of projective curves in [26], where his upper bound for the genus of projective curves was proved. The treatment of (b)-sheaves has been taken from [98].

4.2 Flatness in the projective case

This section is devoted to some properties of flat families of projective schemes which will be needed in this chapter. In particular, we will prove a powerful technical result due to Grothendieck and Mumford, the existence of flattening stratifications, which is a key ingredient in the construction of the Hilbert schemes, of the Quot schemes, and of related schemes like the Severi varieties. The treatment of stratifica- tions closely follows Lecture 8 of [130].

4.2.1 Flatness and Hilbert polynomials

The following result gives the name to the “Hilbert scheme”. Proposition 4.2.1. (i) Let S be a scheme, F a coherent sheaf on IPr × S and p : IP r × S → S the projection. Then F is flat over S if and only if p∗F(h) is locally free on S for all h  0. 4.2 Flatness in the projective case 195

(ii) Assume that S is connected. For each s ∈ Slet i i r Ps(t) = χ(F(s)(t)) = (−1) h (IP (s), F(s)(t)) i

be the Hilbert polynomial of F(s).IfF is flat over S then Ps(t) is independent of s ∈ S. Conversely, if S is integral and Ps(t) is independent of s for all s ∈ S, then F is flat over S. If S is integral and algebraic and Ps(t) is independent of s for all closed s ∈ S, then F is flat over S.

For the proof of this proposition we refer the reader to [84], Theorem III.9.9.

Corollary 4.2.2. If X ⊂ IP r × S ↓ S is a flat family of closed subschemes of IPr with S connected, then all fibres X (s) have the same Hilbert polynomial; in particular, they have the same degree.

Proof. It follows from 4.2.1 applied to F = OX .  1 Examples 4.2.3. (i) Let Ui ={(z0, z1) ∈ IP : zi = 0}, U = U0 U1 and f : U → IP 1 the natural morphism. Then f is flat surjective and quasi-finite. The fibres of f are 0-dimensional, hence projective, but their degree is not constant. This is not a contradiction with Corollary 4.2.2 because the morphism f is not projective, since U is an affine variety. 3 (ii) In IP with homogeneous coordinates X = (X0, X1, X2, X3) consider the curve

Cu = Proj(k[X]/(X2, X3)) ∪ Proj(k[X]/(X1, X3 − uX0))

1 for every u ∈ A .Ifu = 0 then Cu consists of two disjoint lines, while

C0 = Proj(k[X]/(X1 X2, X3)) is a reducible conic in the plane X3 = 0. The Hilbert polynomials are

Pu(t) = 2t + 2 u = 0 P0(t) = 2t + 1

From Corollary 4.2.2 it follows that {Cu} cannot be the set of fibres of a flat family of closed subschemes of IP 3. We may try to construct a morphism whose fibres are the Cu’s by considering the closed subscheme X ⊂ IP 3 × A1 defined by the ideal

J = (X2, X3) ∩ (X1, X3 − uX0) = (X1 X2, X1 X3, X2(X3 − uX0), X3(X3 − uX0)) 196 4 Hilbert and Quot schemes

1 of k[u, X0,...,X3]. From [84], Prop. III.9.7, it follows that X is flat over A .We have: X (u) = Cu, u = 0

X ( ) = ( [ ]/( , , , 2)) 0 Proj k X X1 X2 X1 X3 X2 X3 X3 and X (0) = C0: indeed X ( ) = ∪ ( [ ]/( , , 2)) 0 C0 Proj k X X1 X2 X3 is a nonreduced scheme obtained from C0 by adjoining an embedded point in (1, 0, 0, 0). In particular, we see that X (0) and C0 have the same support. Prop. III.9.8 of [84] implies that X (0) is uniquely determined by the other fibres, i.e. by X ∩[IP 3 × (A1\{0})].

Fix a scheme S and a coherent sheaf F on IP r × S. Consider a morphism g : T → S and the diagram

h IP r × T −→ IP r × S ↓ q ↓ p g T −→ S where h = id × g. For every open set U ⊂ S we have homomorphisms

j r j r −1 ∗ 0 −1 j ∗ H (IP × U, F) → H (IP × g (U), h F) → H (g (U), R q∗(h F)) and therefore a homomorphism

j j ∗ R p∗F → g∗[R q∗(h F)] which corresponds to a homomorphism

∗ j j ∗ g (R p∗F) → R q∗(h F)

In the case j = 0 we have the following asymptotic result which will be applied later in this section:

Proposition 4.2.4. For all m  0 the homomorphism

∗ ∗ g (p∗F(m)) → q∗(h F(m))

∗ is an isomorphism and, if T is noetherian, R j q∗(h F(m)) = 0 all j ≥ 1.

Proof. We have ∗ ∗ ∗ h F = Γ∗(h F) := ⊕mq∗(h F(m))

Since F = Γ∗(F)we also have ∗ ∗ ∗ h F = h Γ∗(F) = ⊕m g (p∗F)(m) 4.2 Flatness in the projective case 197 and therefore for all m  0

∗ ∼ ∗ g (p∗F(m)) = q∗(h F(m))

For the last assertion cover T by finitely many affine open sets and apply Theorem III.5.2 of [84].  The homomorphism of Proposition 4.2.4 is particularly important when g : Spec(k(s)) → S is the inclusion in S of a point s ∈ S; it is denoted by

j j j r t (s) : R p∗(F)s ⊗ k(s) → H (IP (s), F(s))

The study of these homomorphisms is carried out in [1], Ch. III2 (see also Chapter III, Section 12, of [84]). Their main properties are summarized in the following theorem and in its corollary.

Theorem 4.2.5. Let S be a scheme, F a coherent sheaf on IPr × S, flat over S, s ∈ S and j ≥ 0 an integer. Then: (i) If t j (s) is surjective then it is an isomorphism. + + (ii) If t j 1(s) is an isomorphism then R j 1 p∗(F) is free at s if and only if t j (s) is an isomorphism. j j (iii) If R p∗(F) is free at s for all j ≥ j0 + 1 then t (s) is an isomorphism for all j ≥ j0. Proof. (i) and (ii) are Theorem III.12.11 of [84]. (iii) follows from (i) and (ii) by descending induction on j0.  Corollary 4.2.6. Let X → S be a projective morphism, and let F be a coherent sheaf on X , flat over S. Then: j+1 j+1 (i) If H (X (s), F(s)) = 0 for some s ∈ S and j ≥ 0 then R p∗(F)s = (0), and j j j t (s) : R p∗(F)s ⊗ k(s) → H (X (s), F(s)) is an isomorphism. (ii) Let j0 be an integer such that

H j (X (s), F(s)) = 0

j for all j ≥ j0 + 1 and s ∈ S (e.g. j0 = maxs∈S{dim[Supp(F(s))]}). Then t 0 (s) is an isomorphism for all s ∈ S. (iii) Let j0 ≥ 0 be an integer. Then there is a nonempty open set U ⊂ S such that t j0 (s) is an isomorphism for all s ∈ U.

Proof. (i) follows immediately from 4.2.5. (ii) is a special case of (i). (iii) It is the open set U = U , where U ={s ∈ S : R j p∗(F) is free} j≥ j0 j j s (apply 4.2.5(iii)).  198 4 Hilbert and Quot schemes

4.2.2 Stratifications

Let S be a scheme. A stratification of S consists of a set of finitely many locally closed subschemes {S1,...,Sn} of S, called strata, pairwise disjoint and such that

S = S1 ∪ ...∪ Sn i.e. such that we have a surjective morphism Si → S i

Let F be a coherent sheaf on S and for each s ∈ S let

e(s) := dimk(s)[Fs ⊗ k(s)]

Fix a point s ∈ S,lete = e(s) and let a1,...,ae ∈ Fs be such that their images in Fs ⊗ k(s) form a basis. From Nakayama’s lemma it follows that the homomor- : Oe → F ,..., phism fs S,s s defined by a1 ae is surjective; therefore there is an open neighbourhood U of s to which f extends defining a surjective homomorphism : Oe → F ( ) f U |U . With a similar argument applied to ker fs we may find an affine open neighbourhood U(s) of s contained in U and an exact sequence

Od −→g Oe −→f F → U(s) U(s) |U(s) 0 (4.3)

It follows that: (i) e(s ) ≤ e(s) for all s ∈ U(s): therefore s → e(s) is an upper semicontinuous function from S to ZZ. 0 (ii) Let (gij) be the e × d matrix with entries in H (U(s), OS) which defines g. 0 The ideal generated by the gij’s in H (U(s), OS) defines a closed subscheme Zs of U(s) with support equal to Ye ∩ U(s), where for each e ≥ 0wehaveset Ye ={s ∈ S : e(s) = e}. In particular, Ye is a locally closed subset of S. Moreover: (iii) If q : T → U(s) is a morphism, q∗(F) is locally free of rank e if and only if q factors through the subscheme Zs. ∗ Proof. q factors through Zs if and only if all the functions q (gij) are zero on T . Since the sequence ∗( ) ∗( ) Od q−→g Oe q−→f ∗(F) → T T q 0 is exact on T , this is equivalent to q∗( f ) being an isomorphism and this condition implies that q∗(F) is locally free of rank e. Conversely, if q∗(F) is locally free of rank e,letG = ker[q∗( f )]. At every point t ∈ T we have an exact sequence: ∗ 0 → G ⊗ k(t) → k(t)e → q (F) ⊗ k(t) → 0 4.2 Flatness in the projective case 199

Since q∗(F) ⊗ k(t) is a vector space of dimension e we have G ⊗ k(t) = (0). By Nakayama’s lemma, G = (0) in a neighbourhood of t and therefore G = (0) everywhere.

(iv) Since property (iii) characterizes the scheme Zs and does not depend on the

presentation (4.3), for any s, s ∈ S the schemes Zs and Zs coincide on

U(s) ∩ U(s ); therefore the collection of schemes {Zs : s ∈ S} defines a lo- cally closed subscheme Ze of S supported on Ye. Evidently, {Ze : e ≥ 0} is a stratification of S.

(v) Because of (i), for each e the closure of Ze is contained in e ≥e Ze . In partic- ular, if E is the highest integer such that Z E =∅, then Z E is closed. (vi) By the right exactness of tensor product, the construction of the schemes Ze commutes with base change (the proof is similar to that of (iii)). In other words, if q : T → S is a morphism then

˜ −1 Ze = f (Ze) ˜ ∗ for all e, where Ze ⊂ T is the locally closed stratum associated to the sheaf q F. We have proved the following:

Theorem 4.2.7. Let S be a scheme and F a coherent sheaf on S. There is a unique stratification {Ze}e≥0 of S such that if q : T → S is a morphism the sheaf ∗ q (F)is locally free if and only if q factors through the disjoint union of the Ze’s: → → . T e Ze S Moreover, the strata Z0, Z1,... are indexed so that for each e = 0, 1,... the restriction of F to Ze is locally free of rank e. ¯ ⊂ For a given e, Ze e ≥e Ze . In particular, if E is the highest integer such that Z E =∅, then Z E is closed. The stratification {Ze}e≥0 commutes with base change. Theorem 4.2.7 describes a natural way to construct stratifications on a scheme. {Ze}e≥0 is called the stratification defined by the sheaf F. An alternative approach to the construction of this stratification is by the “Fitting ideals” of the sheaf F.Letk ≥ 0; the k-th Fitting ideal of F is the ideal sheaf F ⊂ O ( − ) ( ) Fittk S locally defined by the minors of order e k of the matrix gij in the presentation (4.3). The proof of Theorem 4.2.7 essentially shows that the Fitting ideals are independent of the choice of the presentations (4.3). The closed subscheme F (F) of S defined by Fittk−1 is denoted by Nk . It follows directly from the definition that Supp(Nk(F)) ={s ∈ S : dimk(s)(Fs ⊗ k(s)) ≥ k} and that Nk(F) commutes with base change. Therefore the stratification defined in Theorem 4.2.7 can be also described as follows:

Ze = Ne(F)\Ne+1(F) (4.4)

For details about the properties of the Fitting ideals see [48]. 200 4 Hilbert and Quot schemes

Example 4.2.8. Let ϕ : A → B be a homomorphism of locally free sheaves on the scheme S, of ranks a and b respectively. Applying Theorem 4.2.7 to coker(ϕ) we obtain a stratification of S with the property that Zb−e is supported on the locus

{s ∈ S : rk[ϕ(s) : A(s) → B(s)]=e}

The scheme Zb−e of this stratification will be denoted by De(ϕ). Note that in partic- ular, the subscheme D0(ϕ), called the vanishing scheme of ϕ,isclosedinS because of (v) above. It has the property that a morphism f : T → S satisfies f ∗(ϕ) = 0if and only if f factors through D0(ϕ). The ideal sheaf of D0(ϕ) is locally generated by the entries of a matrix repre- senting ϕ. More intrinsically, it can be obtained as follows. Since ϕ ∈ Hom(A, B),it induces by adjunction a homomorphism:

ϕ∨ Hom(B, A) −→ OS whose image is just the ideal sheaf of D0(ϕ). Example 4.2.9. Let f : X → S be a finite morphism of algebraic schemes. Then f∗OX is a coherent sheaf on S. The scheme Nk( f∗OX ) ⊂ S is supported on the set of points of S having ≥ k preimages (counting multiplicities). It is usually denoted by Nk( f ) and it is called the k-th multiple point scheme of f . The corresponding stratification is the multiple point stratification of S relative to f . There is a vast literature on this stratification. For more about it we refer the reader to [61] and to the literature quoted there.

4.2.3 Flattening stratifications

Definition 4.2.10. Let S be a scheme and F a coherent sheaf on IPr × S. A flattening stratification for F is a stratification {S1,...,Sn} of S such that for every morphism g : T → S the sheaf ∗ Fg := (1 × g) (F) r on IP × T is flat over T if and only if g factors through Si . Note that if such a stratification exists it is clearly unique. In the special case r = 0 we obtain again the notion of stratification defined by the sheaf F. The following is a basic technical result. Theorem 4.2.11. For every coherent sheaf F on IPr × S the flattening stratification exists.

Proof. The theorem has already been proved in the case r = 0 (Theorem 4.2.7). Therefore we may assume r ≥ 1. We will proceed in several steps. Step 1: There are finitely many locally closed subsets Y 1,...,Y k of S such that for each i = 1,...,k if we consider on Y i the reduced scheme structure then F ⊗ O i Y i ×IP r is flat over Y . 4.2 Flatness in the projective case 201

It follows immediately from a repeated use of the fact that there is a nonempty ⊂ F r open subset U S such that |IP ×Ured is flat over Ured (see Note 7). Step 2: Only finitely many polynomials P1,...,Ph occur as Hilbert polynomials of the sheaves F(s), s ∈ S. In fact from Corollary 4.2.2 it follows that, at most, as many Hilbert polynomials occur as the number of connected components of the sets Y 1,...,Y k. Step 3: There is an integer N such that for every m ≥ 0 and for every s ∈ S we have: H j (IP r (s), F(s)(N + m)) = (0)

for j ≥ 1 and the natural map:

0 r [p∗F(N + m)]s ⊗ k(s) → H (IP (s), F(s)(N + m))

is an isomorphism, where p : IP r × S → S is the projection. For each i = 1,...,k consider the diagram hi : IP r × Y i → IP r × S ↓ pi ↓ p (4.5) Y i → S

j i∗ and let ni  0 be so that R pi∗[h F(ni + m)]=(0) for all m ≥ 0 and all j ≥ 1 (apply Proposition 4.2.4). Letting

N  max{n1,...,nk} we may apply Proposition 4.2.4 to the diagrams (4.5) and to the sheaf F and we obtain isomorphisms [ F( + )]⊗O ∼ [ i∗F( + )] p∗ N m Y i = pi∗ h N m for all s ∈ Y i and for all i = 1,...,k. In particular, we have isomorphisms

∼ i∗ [p∗F(N + m)]⊗k(s) = pi∗[h F(N + m)]s ⊗ k(s) (4.6) for all s ∈ Y i and for all i = 1,...,k. We may also apply Corollary 4.2.6 to the i∗ sheaves h F and to the projections pi for j0 = 0 to deduce that

H j (IP r (s), F(s)(N + m)) = (0) for all s ∈ S, j ≥ 1 and m ≥ 0, and that

i∗ ∼ 0 r pi∗[h F(N + m)]s ⊗ k(s) = H (IP (s), F(s)(N + m)) (4.7) for all s ∈ Y i and for all i = 1,...,k and all m ≥ 0. 202 4 Hilbert and Quot schemes

Comparing (4.6) and (4.7) we obtain the conclusion.

Step 4: Let N be as in Step 3, and let g : T → S be a morphism. Then Fg is flat ∗ over T if and only if g [p∗F(N + m)] is locally free for all m ≥ 0.

r Suppose that Fg is flat over T and let q : IP × T → T be the projection. Since

j r j r H (IP (t), Fg(t)(N + m)) = H (IP (g(t)), F(g(t))(N + m)) = (0) for all t ∈ T , m ≥ 0 and j ≥ 1, from Corollary 4.2.6(ii) we deduce that

0 r q∗Fg(N + m)t ⊗ k(t) → H (IP (g(t)), F(g(t))(N + m)) (4.8) is an isomorphism for all t ∈ T . Theorem 4.2.5(ii) applied for j =−1 implies that q∗Fg(N + m) is locally free for all m ≥ 0. For all t ∈ T the natural homomorphism ∗ ϕ : g [p∗F(N + m)]→q∗Fg(N + m) induces an isomorphism: ∗ ∼ g [p∗F(N + m)]t ⊗ k(t) = q∗Fg(N + m)t ⊗ k(t) because both sides are isomorphic to H 0(IP r (g(t)), F(g(t))(N + m)) (the first be- cause of Step 3, the second because of (4.8)). From the fact that q∗Fg(N + m) is locally free and from Nakayama’s lemma it follows that ϕ is an isomorphism. There- ∗ fore g [p∗F(N + m)] is locally free for every m ≥ 0. ∗ Conversely, suppose that g [p∗F(N + m)] is locally free for all m ≥ 0. Since for all m  0 the natural map ϕ is an isomorphism (Prop. 4.2.4) it follows that q∗Fg(N + m) is locally free for all m  0: Proposition 4.2.1 implies that Fg is flat.

Step 5: For every m ≥ 0 apply Theorem 4.2.7 to the sheaf p∗F(N + m) and let Ym, j be the component of the corresponding stratification of S where p∗F(N + m) becomes locally free of rank j. Then for each j = 1,...,h we have the following equality of subsets of S: ( ) = ( ) Supp Ym,Pi (N+m) Supp Ym,Pi (N+m) m≥0 m=0,...,r

The inclusion ⊂ is obvious. For s ∈ S let Ps(t) be the Hilbert polynomial of F( ) ∈∩ ( ) s . Then s m≥0Supp Ym,Pi (N+m) if and only if

0 r i Ps(N +m) = h (IP (s), F(s)(N +m)) = dim[p∗F(N +m)s ⊗k(s)]=P (N +m)

i for all m ≥ 0, and this happens if and only if Ps(t) = P (t) as polynomials. On the ∈∩ ( ) ( + ) = i ( + ) other hand, s m=0,...,r Supp Ym,Pi (N+m) if and only if Ps N m P N m i for m = 0,...,r. Since both Ps(t) and P (t) have degree ≤ r,itfollowsthatPs(t) = i ( ) ∈∩ ( ) P t and therefore s m≥0Supp Ym,Pi (N+m) . 4.2 Flatness in the projective case 203

Step 6: Fix i between 1 and h. For each integer c ≥ 0 the finite intersection Ym,Pi (N+m) m=0,...,c

is a well-defined locally closed subscheme of S. Because of Step 5 the sub- schemes , = , + ,... Ym,Pi (N+m) c r r 1 m=0,...,c for a descending chain with fixed support; in particular, they form a descend- ing chain of closed subschemes of a fixed open set V ⊂ S, and therefore they stabilize. In other words the intersection i = ( ) Z Supp Ym,Pi (N+m) m≥0

is a well-defined locally closed subscheme of S.ByStep5wehave:

i Supp(Z ) ={s ∈ S : Ps(t) = Pi (t)}

Step 7: The subschemes Z 1,...,Z h form a stratification of S. It follows immedi- ately from Step 4 that this is the flattening stratification for F. This concludes the proof of Theorem 4.2.11. 

NOTES 1. From the proof of Theorem 4.2.11 it follows that the strata Z1,...,Zh of the flattening stratification for F are indexed by the Hilbert polynomials of the sheaves F(s), s ∈ S. 0 2. (The see-saw theorem) Let X be a projective scheme such that H (X, OX ) = k,San algebraic scheme and L an invertible sheaf on X × S. Then: (i) The locus ∼ S0 ={s ∈ S : L|X×{s} = OX } is closed in S. (ii) Letting p0 : X × S0 → S0 be the projection, there is an invertible sheaf M on S0 such that L =∼ ∗ |X×S0 p0 M

Proof. A line bundle L on X is trivial if and only if it satisfies h0(X, L)>0and − − h0(X, L 1)>0. In fact, nonzero sections σ ∈ H0(X, L) and τ ∈ H0(X, L 1) correspond to homomorphisms σ τ ∨ OX −→ L −→ OX 0 whose composition is multiplication by a constant (by the assumption H (X, OX ) = k), so that they are isomorphisms. 204 4 Hilbert and Quot schemes

+ − Therefore S0 = S ∩ S where

+ 0 S ={s ∈ S : h (X ×{s}, L|X×{s})>0}

− ={ ∈ : 0( ×{ }, L−1 )> } S s S h X s |X×{s} 0 + − It follows from the semicontinuity theorem that both S and S are closed in S,and(i) follows. 0 (ii) Since h (X ×{s}, L|X×{s}) = 1 is constant in s ∈ S0, from Corollary 12.9 of [84] we ∼ 0 deduce that p0∗L⊗k(s) = H (X ×{s}, L|X×{s}) for all s ∈ S0 and in particular, p0∗L =: M is an invertible sheaf. We have a homomorphism of invertible sheaves on X × S0: ∗ = ∗ L → L p0 M p0 p0∗ |X×S0 which is surjective because L|X×{s} is globally generated for all s ∈ S0. Then it is an isomor- phism.  3. Let X → S be a flat projective morphism of algebraic schemes, and L an invertible sheaf on X . Assume that for some k-rational point o ∈ SthesheafL(o) is very ample on X (o) and satisfies H1(X (o), L(o)) = 0. Then there is an open neighbourhood V ⊂ Sofo such that LV := L|X (V ) is very ample relative to V . In particular, L(s) is very ample on X (s) for every s ∈ V.

Proof. By Corollary 4.2.6, there is an open neighbourhood U ⊂ S of o such that 1 (R f∗L)|U = (0) and

0 0 t (u) : ( f∗L)u ⊗ k(u) → H (X (u), L(u)) is an isomorphism for all u ∈ U. We may even assume that f∗L is locally free of rank h0(X (o), L(o)) on U. From the surjectivity of the map t0(o) and from the fact that L(o) is globally generated we deduce that the canonical homomorphism:

∗ f ( f∗L) → L is surjective on X (o).Since f is projective it follows that there is an open W ⊂ U containing o such that ∗ [ f ( f∗L)]|X (W) → LW (4.9) ∗ is surjective and moreover, [ f ( f∗L)]|X (W) is locally free. The homomorphism (4.9) defines a W-morphism ∗ X (W) → IP ( f ( f∗L))|X (W) (4.10) W whose restriction to X (o) is the embedding defined by the global sections of L(o).From Note 3 of § 3.4 above it follows that there is an open subset V ⊂ W containing o such that the restriction of (4.10) to X (V ) is an embedding. This implies the conclusion (see [116], Theorem 1.2.17 for the proof of a more general statement without flatness assumption). 4. Let E be a locally free sheaf over IP1 × S, with S an algebraic integral scheme. Let ∈ E( ) =∼ ⊕ O( i ) o Sbeak-rational point, and o i n0 the fibre over o. Then: 4.2 Flatness in the projective case 205

(i) There is an open set U ⊂ S such that for each s ∈ U we have

E( ) ∼ ⊕ O( i ) s = i ns

with { i }≤ { i } { i }≥ { i } maxi ns maxi no and mini ns mini no E( ) i = j , E( ) ∼ E( ) ∈ Moreover, if o is balanced (i.e. no no for all i j) then s = o for all s U. (ii) For each s ∈ S we have i = i ns n0 i i

Proof. ([25]) (i) By the structure theorem for locally free sheaves on IP 1 (see [136]) we know ∈ E( ) ∼ ⊕ O( i ) i that for each s S we have an isomorphism s = i ns for some integers ns.Let = { i } E¯ := E ⊗ ∗O(− − ) : 1 × → 1 M0 maxi n0 and consider the sheaf p M0 1 ,wherep IP S IP is the projection. Since h0(E¯(0)) = 0, from the semicontinuity theorem it follows that there is an open neighbourhood U of 0 such that h0(E¯(s)) = 0foralls ∈ U; but this means that { i }≤ ∈ maxi ns M0 for all s U, which is the first statement of the proposition. The statement about the minimum is proved similarly after replacing E by its dual. The last assertion is obvious. (ii) Applying (i) to det(E) we find that every point t ∈ S has an open neighbourhood Ut i = i ∈ i where i ns i nt for all s Ut .SinceS is connected we deduce that i ns is constant. 5. Let X → S be a flat projective morphism with S an algebraic scheme, and let o ∈ S be a k-rational point. Prove that:

(i) If X (o) is connected and X (s) is disconnected for all s = o in an open neighbourhood of o then X (o) is nonreduced. In particular: (ii) If X (o) is connected and reduced then X (s) is connected for all s in an open neighbour- hood of o. (iii) If X (o) is disconnected then X (s) is disconnected for all s in an open neighbourhood of o.

6. Let f : X → Y be a proper morphism of algebraic schemes with finite fibres. Let g : Y → Y be an arbitrary morphism, X = X ×Y Y ,f : X → Y and g : X → Xthe projections. Then for every quasi-coherent OX -module F we have a canonical isomorphism

∗ ∼ ∗ g ( f∗F) = f∗(g F)

Proof. Since it is proper and quasi-finite, f is finite, in particular it is affine. The conclusion follows from [1] Ch. II, 1.5.2.  7. (Generic flatness) Let f : X → S be a morphism of finite type with S integral, and let F be a coherent sheaf on X . There is a dense open subset U ⊂ S such that the restriction of F to f −1(U) is flat over U.

Proof. See [1], ch. IV, Th. 6.9.1 or [48], Theorem 14.4, p. 308.  − Note that if f is not dominant then U = S\ f (X ) and f 1(U) =∅. 206 4 Hilbert and Quot schemes 4.3 Hilbert schemes

4.3.1 Generalities

Consider a projective scheme Y , and a closed embedding Y ⊂ IP r . Let’s fix a nu- merical polynomial of degree ≤ r, i.e. a polynomial P(t) ∈ Q[t] of the form: r t + r P(t) = a i i i=0 with ai ∈ ZZ for all i. For every scheme S we let: flat families X ⊂ Y × S of closed subschemes Y ( ) = HilbP(t) S of Y parametrized by S with fibres having Hilbert polynomial P(t) Since flatness is preserved under base change, this defines a contravariant functor Y : ◦ → HilbP(t) (schemes) (sets) called the Hilbert functor of Y relative to P(t). = r r In the case Y IP we will denote the Hilbert functor with the symbol HilbP(t). Y If the functor HilbP(t) is representable, the scheme representing it will be called ( ) Y r the Hilbert scheme of Y relative to P t , and will be denoted by HilbP(t) (or HilbP(t) = r ( ) = Y in case Y IP ). If P t n a constant polynomial then HilbP(t) is usually denoted by Y [n]. Y If the Hilbert scheme HilbP(t) exists then there is a universal element, i.e. there is a flat family of closed subschemes of Y having Hilbert polynomial equal to P(t): W ⊂ × Y Y HilbP(t) (4.11)

Y parametrized by HilbP(t) and possessing the following: Universal property: for each scheme S and for each flat family X ⊂ Y × S of closed subschemes of Y having Hilbert polynomial P(t) there is a unique morphism → Y S HilbP(t), called the classifying morphism, such that

X = S × Y W ⊂ Y × S HilbP(t) ( Y , W) The family (4.11) is called the universal family, and the pair HilbP(t) rep- Y resents the functor HilbP(t). W Y ( Y ) The family is the universal element of HilbP(t) HilbP(t) , namely the element corresponding to the identity under the identification ( Y , Y ) = Y ( Y ) Hom HilbP(t) HilbP(t) HilbP(t) HilbP(t) 4.3 Hilbert schemes 207

Example 4.3.1. Consider the constant polynomial P(t) = 1. Then we have a canon- ical identification Y [1] = Y and the universal family is the diagonal ∆ ⊂ Y × Y . To prove it consider an element of Y [1](S) for some scheme S: Γ ⊂ S × Y ↓ f S

Then f is an isomorphism: in fact it is a one-to-one morphism and OS → f∗OΓ is an isomorphism since f∗OΓ is an OS-algebra which is locally free of rank one −1 over OS. We therefore have the well-defined morphism g = qf : S → Y where q : S × Y → Y is the projection. The morphism

(gf, q) : Γ → Y × Y factors through ∆ and induces a commutative diagram Γ → ∆ ↓↓ S −→ Y g such that Γ =∼ g∗∆. Therefore the family Γ is induced by ∆ via the morphism g.

Before proving the existence of the Hilbert schemes in general we will consider two important special cases.

4.3.2 Linear systems

If X ⊂ IP r is a hypersurface of degree d it has Hilbert polynomial t + r t + r − d d h(t) = − = tr−1 +··· r r (r − 1)! Conversely, if a closed subscheme Y of IP r has the Hilbert polynomial h(t) then it is a hypersurface of degree d. In fact, since h(t) has degree r − 1, Y has dimension r − 1, so Y = Y1 ∪ Z, with Y1 a hypersurface and dim(Z)

I , I ⊂ O r where Y Y1 IP are the ideal sheaves of Y and Y1. We deduce that

h(t) = h1(t) + k(t) where h1(t) is the Hilbert polynomial of Y1 and k(t) is the Hilbert polynomial of I /I ( ( )) < − Y1 Y . Since this sheaf is supported on Z,wehavedeg k t r 1; therefore we see that d h (t) = tr−1 +··· 1 (r − 1)! 208 4 Hilbert and Quot schemes so Y1 is a hypersurface of degree d, and therefore h1(t) = h(t). It follows that ( ) = I = I = k t 0, i.e. Y1 Y ; equivalently, Y Y1. r Therefore Hilbh(t), if it exists, parametrizes a universal family of hypersurfaces of degree d in IP r . To prove its existence let V := H 0(IP r , O(d)) and in IP (V ) take homogeneous coordinates

(. . . , ci(0),...,i(r),...)i(0)+···+i(r)=d

The hypersurface H ⊂ IP r × IP (V ) defined by the equation i(0) ··· i(r) = ci(0),...,i(r) X0 Xr 0 projects onto IP (V ) with fibres hypersurfaces of degree d. It follows from Proposi- tion 4.2.1 that H is flat over IP (V ). Let’s denote by p : IP r × IP (V ) → IP (V ) the projection, and let IH ⊂ OIP r ×IP (V ) be the ideal sheaf of H. For all x ∈ IP (V ) we have 0 r 0 r 1 = h (IP (x), IH(x)(d)) = h (IP (x), IH(d)(x)) and i r i r 0 = h (IP (x), IH(x)(d)) = h (IP (x), IH(d)(x))

i 0 = h (H(x), OH(x)(d)) for all i ≥ 1. Applying 4.2.5 and 4.2.2 we deduce that:

(a) R1 p∗IH(d) = 0; (b) p∗IH(d) is an invertible subsheaf of p∗OIP r ×IP (V )(d) = V ⊗k OIP (V ); (c) p∗OIP r ×IP (V )(d)/p∗IH(d) = p∗OH(d) is locally free. It follows that p∗IH(d) = OIP (V )(−1) the tautological invertible sheaf on IP (V ), and the natural map

∗ p p∗IH(d) → IH(d) is an isomorphism. Therefore

∗ IH =[p OIP (V )(−1)](−d)

Let’s prove that H ⊂ IP r × IP (V ) is a universal family. Suppose that X ⊂ IP r × S ↓ f S is a flat family of closed subschemes of IP r with Hilbert polynomial h(t), i.e. hy- persurfaces of degree d, and let IX ⊂ OIP r ×S be the ideal sheaf of X . Arguing as above we deduce that f∗IX (d) is an invertible subsheaf of V ⊗k OS with locally free cokernel f∗OX (d), and that 4.3 Hilbert schemes 209

∗ IX =[f f∗IX (d)](−d)

We have an induced morphism g : S → IP (V ) such that

∗ g [OIP (V )(−1)]= f∗IX (d)

r The subscheme S ×IP (V ) H ⊂ IP × S is defined by the ideal sheaf ∗ ∗ (1 × g) IH = (1 × g) [OIP (V )(−1)(−d)]

∗ ∗ ∗ = f [g OIP (V )(−1)](−d) =[f f∗IX (d)](−d) = IX

Hence S ×IP (V ) H = X . The proof of the uniqueness of g having this property is left to the reader. Therefore we see that H ⊂ IP r × IP (V ) is a universal family, and r = ( ) Hilbh(t) IP V .

4.3.3 Grassmannians

The classical grassmannians are special cases of Hilbert schemes, since they parame- trize linear spaces, which are the closed subschemes with Hilbert polynomials of the t+n−1 − form n−1 , n 1 being their dimension. Let’s fix a k-vector space V of dimension N, and let 1 ≤ n ≤ N. Letting

∨ GV,n(S) ={loc. free rk n quotients of the free sheaf V ⊗k OS on S} we define a contravariant functor:

o GV,n : (schemes) → (sets) called the Grassmann functor; we will denote it simply by G when no confusion is possible.

Theorem 4.3.2. The Grassmann functor G is represented by a scheme Gn(V ) to- gether with a locally free quotient of rank n ∨ ⊗ O → V k Gn(V ) Q called the universal quotient bundle.

Proof. Given a scheme S and an open cover {Ui } of S, to give a locally free rank n ∨ quotient of V ⊗k OS is equivalent to giving one such quotient over each open set Ui so that they patch together on the intersections Ui ∩ U j . Therefore G is a sheaf. ∨ Let’s fix a basis {ek} of V and choose a set J of n distinct indices in {1,...,N}. We have an induced decomposition V ∨ = E ⊕ E , with E (resp. E ) a vector subspace of rank n (resp. N − n). We can define a subfunctor GJ of G letting: ∨ locally free rk n quotients V ⊗k OS → F GJ (S) = inducing E ⊗k OS → F surjective 210 4 Hilbert and Quot schemes

Let S be any scheme and f : Hom(−, S) → G a morphism of functors correspond- ing to a locally free rank n quotient ∨ V ⊗k OS → F

The fibred product SJ := Hom(−, S) ×G GJ is clearly represented by the open subscheme of S supported on the points where the map E ⊗k OS → F is surjective; this proves that GJ is an open subfunctor of G. Since clearly the SJ ’s cover S,we also see that the family of subfunctors {GJ } is an open covering of G. To prove that GJ is representable note that if ∨ q : V ⊗k OS → F is an element of G(S) then the induced map

η : E ⊗k OS → F is surjective if and only if it is an isomorphism; in this case the composition

−1 ∨ η ◦ q : V ⊗k OS → E ⊗k OS restricts to the identity on E ⊗k OS, hence it is determined by the composition ∨ E ⊗k OS → V ⊗k OS → E ⊗k OS It follows that we can identify

GJ (S) = Hom(E ⊗k OS, E ⊗k OS) = Hom(E , E ) ⊗k OS n(N−n) This proves that GJ is isomorphic to Hom(−, A ), hence it is representable. Now the theorem follows from Proposition E.10.  Gn(V ) is called the grassmannian of n-dimensional subspaces of V ;itisalso called the grassmannian of (n−1)-dimensional projective subspaces of IP(V ). When N N V = k the grassmannian Gn(k ) is denoted by G(n, N). When n = 1 the functor GV,1 is represented by ∨ G1(V ) = IP (V ) = Proj(Sym(V )), the (N − 1)-dimensional projective space associated to V . In this case Q = OIP (V )(1). From Theorem 4.3.2 it follows that for all schemes S the morphisms f : S → Gn(V ) are in 1–1 correspondence with the locally free rank n quotients ∨ V ⊗k OS → F ∗ via f ↔ f Q. This is the universal property of Gn(V ). The universal quotient bundle defines an exact sequence of locally free sheaves on Gn(V ): → → ∨ ⊗ O → → 0 K V k Gn(V ) Q 0 called the tautological exact sequence; K is called the universal subbundle. 4.3 Hilbert schemes 211

Let S be a scheme. Associating to every locally free quotient of rank n ∨ V ⊗k OS → F the quotient n ∨ n (∧ V ) ⊗k OS →∧ F we define a morphism of functors GV,n → G∧n V,1, which is induced by a morphism

n π : Gn(V ) → IP (∧ V )

π is called the Plucker¨ morphism.

Proposition 4.3.3. The Plucker¨ morphism is a closed embedding. In particular, Gn(V ) is a projective variety. Proof. As in the proof of 4.3.2, we fix a basis of V ∨ and we choose a set J of n distinct indices in {1,...,N}. We obtain a decomposition V ∨ = E ⊕ E with dim(E ) = n,dim(E ) = N − n, and an induced one: ∧n ∨ =⊕n (∧n−i ) ⊗ ∧i =∧n ⊕ V i=0 E k E E F =⊕n (∧n−i ) ⊗ ∧i where F i=1 E k E . For every scheme S let locally free rk 1 quotients ∧n V ∨ → L IP (S) = J s.t. the induced ∧n E → L is surjective

We obtain a subfunctor IP J of G∧n V,1. As in the proof of 4.3.2, we see that the IP J ’s form an open cover of G∧n V,1 by functors representable by affine spaces. ∨ Note that for every locally free rank n quotient V ⊗k OS → F the induced homomorphism:

E ⊗k OS → F n n −1 is surjective if and only if ∧ E →∧ F is. Therefore GJ = π (IP J ) and it suffices to prove that π : GJ → IP J is a closed embedding.

We may identify GJ with (the affine space associated to) Homk(E , E ) and IP J n with Homk(F, ∧ E ). Considering that

n−i n ∼ i Homk(∧ E , ∧ E ) = ∧ E canonically via the perfect pairing:

∧i E ×∧n−i E →∧n E we have: ( , ∧n ) =⊕n ((∧n−i ) ⊗ ∧i , ∧n ) Homk F E i=1Homk E k E E =⊕n (∧i , (∧n−i , ∧n )) i=1Homk E Homk E E =⊕n (∧i , ∧i ) i=1Homk E E 212 4 Hilbert and Quot schemes and the map n π : Homk(E , E ) → Homk(F, ∧ E ) is λ → (λ, ∧2λ,...,∧nλ)

This is the graph of a morphism of affine schemes, hence it is a closed embedding.  For some 1 ≤ n ≤ r,letG = G(n + 1, r + 1) be the grassmannian of n- dimensional projective subspaces of IP r . Consider the projective bundle over G:

∨ I := IP (Q ) = Proj(Sym(Q))

r+1 → where Q is the universal quotient bundle on G. Because of the surjection OG Q we have a closed embedding:

I ⊂ IP r × G ↓ p G

∨ = I ( ) ⊂ O ( ) = Or+1 v ∈ note that Q p∗ I 1 p∗ IP r ×G 1 G . For every closed point G the fibre I(v) is the projective subspace IP (v) ⊂ IP r ; for this reasonI is called t+n the incidence relation. Since all fibres of p have Hilbert polynomial n ,from Proposition 4.2.1 it follows that p is a flat family. Suppose now that

Λ ⊂ IP r × S ↓ q S t+n is another flat family whose fibres have Hilbert polynomial n . We have an inclu- sion of sheaves on S

I ( ) ⊂ O ( ) = Or+1 q∗ Λ 1 q∗ IP r ×S 1 S which has locally free cokernel q∗OΛ(1). By the universal property of G the above inclusion induces a unique morphism

g : S → G

∗ ∨ such that g (Q ) = q∗IΛ(1). Since Λ = IP (q∗IΛ(1)) it follows that

Λ = S ×G I namely, the family q is obtained by base change from the incidence relation via the morphism g. This proves that ( + , + ) = r G n 1 r 1 Hilb(t+n) n 4.3 Hilbert schemes 213

4.3.4 Existence

Theorem 4.3.4. For every projective scheme Y ⊂ IP r and every numerical polyno- ( ) Y mial P t , the Hilbert scheme HilbP(t) exists and is a projective scheme. Proof. We will first prove the theorem in the case Y = IP r . From Corollary 4.1.10 r it follows that there is an integer m0 such that for every closed subscheme X ⊂ IP with Hilbert polynomial P(t) the sheaf of ideals IX is m0-regular. It suffices to take

m0 = Fr (−a0,...,−ar−1, 1 − ar )

It follows that for every k ≥ m0

i r h (IP , IX (k)) = 0 (4.12)

≥ for i 1 and k + r h0(IP r , I (k)) = − P(k) X r depends only on P(k). Moreover, by Note 2 of Section 4.1 we have

i h (X, OX (k)) = 0 (4.13) all k ≥ m0 and all i ≥ 1. Let m + r N = 0 − P(m ) r 0

0 r V = H (IP , OIP r (m0))

Consider the grassmannian G = G N (V ) of N-dimensional vector subspaces of V . ∨ Let V ⊗k OG → Q be the universal quotient bundle, which is a locally free sheaf of rank N on G, and p : IP r × G → G the projection. We may identify V ⊗k OG = p∗[OIP r ×G (m0)]. The image of the composition ∗ ∨ p Q (−m0) −→ V ⊗k OIP r ×G (−m0) −→ OIP r ×G  ∗ p p∗[OIP r ×G (m0)]⊗OIP r ×G (−m0) is a sheaf of ideals: we will denote it by J. Let Z ⊂ IP r ×G be the closed subscheme defined by J and denote by q : Z → G the restriction of p to Z. Consider the flattening stratification G1 G2 ...⊂ G 214 4 Hilbert and Quot schemes for OZ and let H be the stratum relative to the polynomial P(t). We will prove that = r H HilbP(t) and that the universal family is the pullback of q to H:

W := H ×G Z → Z ↓ π ↓ q H → G By the choice of H, W defines a flat family of closed subschemes of IP r with Hilbert polynomial equal to P(t). Let’s prove that W has the universal property. Consider a flat family of closed subschemes of IP r with Hilbert polynomial P(t): X ⊂ IP r × S ↓ f S From (4.12) and (4.13) and from Theorem 4.2.5 it follows that

1 1 R f∗IX (m0) = (0) = R f∗OX (m0) In particular, we have an exact sequence on S:

0 → f∗IX (m0) → f∗OIP r ×S(m0) → f∗OX (m0) → 0  V ⊗k OS

If we apply Theorem 4.2.5 for j =−1 we deduce that f∗IX (m0) and f∗OX (m0) are locally free and f∗IX (m0) has rank N. From the universal property of G it follows that there exists a unique morphism g : S → G such that ∗ ∨ f∗IX (m0) = g Q

∗ Claim: For all m  m0 we have f∗OX (m) = g p∗OZ (m).

Proof of the Claim: For all m  m0 we have exact sequences:

0 → p∗J(m) → p∗OIP r ×G (m) → q∗OZ (m) → 0 on G and 0 → f∗IX (m) → f∗OIP r ×S(m) → f∗OX (m) → 0 ∗ on S; since g p∗OIP r ×G (m) = f∗OIP r ×S(m) it suffices to show that: ∼ ∗ f∗IX (m) = g p∗J(m) for all m  m0. For all such m we have the equality on G: ∨ p∗J(m) = Im[Q ⊗ p∗O(m − m0) → p∗OIP r ×G (m)]

∗ ∨ r induced by the surjections p Q (m − m0) → J(m) of sheaves on IP × G. Hence for all m  m0 we have: 4.3 Hilbert schemes 215

∗ ∗ ∨ g p∗J(m) = g Im[Q ⊗ p∗OIP r ×G (m − m0) → p∗OIP r ×G (m)] ∗ ∨ = Im[g Q ⊗ f∗OIP r ×S(m − m0) → f∗OIP r ×S(m)] = Im[ f∗IX (m0) ⊗ f∗OIP r ×S(m − m0) → f∗OIP r ×S(m)]= f∗IX (m) and this proves the Claim. From the Claim it follows that:

(i) g factors through H.

Indeed, from 4.2.4 it follows that for all m  m0: ∗ ∗ g q∗OZ (m) = f∗(1 × g) OZ (m)

∗ Since g q∗OZ (m) = f∗OX (m) is locally free of rank P(m) for all such m Proposi- ∗ tion 4.2.1 implies that (1×g) OZ is flat over S with Hilbert polynomial P(t). Hence g factors by the definition of H.

(ii) X = S ×H W. Indeed: ∗ X = Proj[ f∗OX (m)]=Proj[ g q∗OZ (m)]   m0 m 0 ∗ = Proj[ g π∗OW (m)]=S ×H Proj[ π∗OW (m)]=S ×H W m0 m0 = r π Properties (i) and (ii) imply that H HilbP(t) and that is the universal family. r By construction, HilbP(t) is a quasi-projective scheme. To prove that it is projec- tive it suffices to show that it is proper over k. We will use the valuative criterion of properness. Let A be a discrete valuation k-algebra with quotient field L and residue field K , and let ϕ : ( ) → r Spec L HilbP(t) be any morphism. We must show that ϕ extends to a morphism ϕ˜ : ( ) → r Spec A HilbP(t) Pulling back the universal family by ϕ we obtain a flat family X ⊂ IP r × Spec(L) of closed subschemes of IP r with Hilbert polynomial P(t). Since Spec(A) is non- singular of dimension one and

Spec(L) = Spec(A)\{closed point}

Proposition III.9.8 of [84] implies the existence of a flat family

X ⊂ IP r × Spec(A) 216 4 Hilbert and Quot schemes

X r which extends . By the universal property of HilbP(t) this family corresponds to a ϕ˜ : ( ) → r ϕ morphism Spec A HilbP(t) which extends . This concludes the proof of the theorem in the case Y = IP r . Let’s now assume that Y is an arbitrary closed subscheme of IP r .Itwillsuffice Y r to show that the functor HilbP(t) is represented by a closed subscheme of HilbP(t). Applying Corollary 4.1.10 twice we can find an integer µ such that IY ⊂ OIP r is µ-regular and such that for every closed subscheme X ⊂ IP r with Hilbert polyno- mial P(t) the ideal sheaf IX ⊂ OIP r is µ-regular. Let

0 r 0 r V = H (IP , OIP r (µ)), U = H (IP , IY (µ))

It follows from 4.2.5 and 4.2.6 that π∗IW (µ) is a locally free subsheaf of V ⊗k OHilb with locally free cokernel. r On HilbP(t) consider the composition

# : U ⊗k OHilb → V ⊗k OHilb → V ⊗k OHilb/π∗IW (µ) ⊂ r # = Let Z HilbP(t) be the closed subscheme defined by the condition 0, or equivalently, by the condition

U ⊗k OZ ⊂ π∗IW (µ) ⊗ OZ (4.14) : → r Letting j Z HilbP(t) be the inclusion, one easily sees that condition (4.14) implies that ∗ IY ×Z ⊂ (1 × j) IW ⊂ OIP r ×Z hence that r Z ×Hilb W ⊂ Y × Z ⊂ IP × Z (4.15) = Y It is straightforward to check that Z HilbP(t) and that (4.15) is the universal family. This concludes the proof of Theorem 4.3.4.  For any projective scheme Y ⊂ IP r it is often convenient to consider the functor:

HilbY : (schemes) → (sets) defined as: Y ( ) = Y ( ) Hilb S HilbP(t) S P(t) This functor is represented by the disjoint union Y = Y Hilb HilbP(t) P(t) which is a scheme locally of finite type (but not of finite type because it has infinitely many connected components unless dim(Y ) = 0).Itisthe Hilbert scheme of Y .One convenient feature of HilbY is that it is independent on the projective embedding of Y Y , even though the indexing of its components HilbP(t) by Hilbert polynomials does 4.3 Hilbert schemes 217 depend on the embedding. For this reason, when considering HilbY we will not need to specify a projective embedding of Y . Let’s fix a projective scheme Y , and in the Hilbert scheme HilbY let’s consider a k-rational point [X] which parametrizes a closed subscheme X ⊂ Y . Denote by I ⊂ O Y Y the ideal sheaf of X in Y . The local Hilbert functor HX is a subfunctor of the restriction to A of the Hilbert functor; since HilbY represents the Hilbert functor we have, with the notation introduced in § 2.2: Y ( ) = ( ( ), Y ) HX A Hom Spec A Hilb [X] (A) Y Oˆ for every A in ob . In particular, HX is prorepresented by the local ring Hilb,[X]. We can therefore apply the results proved in § 2.4 to obtain information about the local properties of HilbY at [X]. In particular, we have the following:

Theorem 4.3.5. (i) There is a canonical isomorphism of k-vector spaces:

Y ∼ 0 T[X]Hilb = H (X, NX/Y ) = (I/I2, O ) where NX/Y HomOX X is the normal sheaf of X in Y . ⊂ O (ii) If X Y is a regular embedding then the obstruction space of HilbY ,[X] is a 1 subspace of H (X, NX/Y ).

The simplest illustration of Theorem 4.3.5 is for Y [1] = Y . In this case 4.3.5(i) ( / 2 , ) simply says that Hom m p m p k is the Zariski tangent space of Y at a k-rational ∈ (O ) = (O ) point p Y . The obstruction space is o Y [1],[p] o Y,p . Of course if p is a 1 singular point then it is not regularly embedded in Y , and H (p, N p/Y ) = 0 is not an obstruction space for the local Hilbert functor. Consider a flat family of closed subschemes of Y : X ⊂ Y × S ↓ f S

It induces a functorial morphism χ : S → HilbY (the classifying morphism of the family) whose differential at a k-rational point s ∈ S is a linear map

0 dχs : Ts S → H (X (s), NX (s)/Y ) (4.16) called the characteristic map of the family f . Obviously, the surjectivity of dχs is a necessary condition for the smoothness of χ at s. We have the following more precise result.

Proposition 4.3.6. Let Y be a projective scheme, X ⊂ Y × S ↓ f S 218 4 Hilbert and Quot schemes a flat family of closed subschemes of Y with S algebraic, and

χ : S → HilbY the classifying map of the family. Then: (i) If s is a nonsingular point of S and if the characteristic map

0 dχs : Ts S → H (X (s), NX (s)/Y ) is surjective then χ is smooth at s and X (s) is unobstructed in Y . 1 (ii) If, moreover, h (X (s), TY |X (s)) = 0 and X (s) is nonsingular then f has general moduli at s and X (s) is unobstructed as an abstract variety.

Proof. (i) The smoothness of χ is a consequence of Theorem 2.1.5 and of the non- singularity of S at s. The unobstructedness of X (s) in Y , i.e. the nonsingularity of HilbY at χ(s), follows from the smoothness of χ and the nonsingularity of S at s. 1 (ii) The condition h (X (s), TY |X (s)) = 0 implies that the forgetful morphism Y → HX (s) DefX (s) is smooth (Proposition 3.2.9) and therefore the Kodaira–Spencer map of f at s is surjective, being the composition of two surjective maps. We obtain the conclusion from Proposition 2.5.8.  The following criterion follows at once from Proposition 3.2.9:

Proposition 4.3.7. Let X ⊂ Y be a closed embedding of projective nonsingular 1 schemes such that h (X, TY |X ) = 0. Then the universal family X ⊂ Y × HilbY ↓ HilbY has general moduli at the point [X].

NOTES 1. The construction of the grassmannian given here is taken from [97]. r ( ) 2. It is a classical result of Hartshorne that HilbP(t) is connected for all r and P t (see [83] and [140]). For general Y this is no longer true: for example, if Y ⊂ IP 3 is a nonsingular Y quadric then Hilbt+1 has two connected components. 3. Let X ⊂ Y be a closed embedding of projective schemes. It can be easily verified that for any closed subscheme Z ⊂ X, the induced injective linear map 0 0 H (Z, NZ/ X ) → H (Z, NZ/Y ) coincides with the differential at [Z] of the closed embedding

HilbX ⊂ HilbY

4. If Z ⊂ Y is a closed embedding of projective schemes and P(t) a numerical polyno- mial, then one can define a functor 4.4 Quot schemes 219

(− )Y : ( )◦ → ( ) Hilb Z P(t) schemes sets by flat families X ⊂ Y × S with Hilbert polynomial Hilb(−Z)Y (S) = P(t) P(t) and such that X (s) ∩ Z =∅, ∀s ∈ S Y (− )Y Then Hilb Z P(t) is representable by a closed subscheme

(− )Y ⊂ Y Hilb Z P(t) HilbP(t) Let Y \Z := Y \ (− )Y ⊂ Y HilbP(t) HilbP(t) Hilb Z P(t) HilbP(t) be the corresponding open subscheme and define Y \Z := Y \Z Hilb HilbP(t) P(t)

Then this is a scheme locally of finite type which depends only on W := Y \Z and not on ⊂ ( ) = W the embedding W Y . In the case P t n, a constant polynomial, we denote HilbP(t) by [ ] W n .IfW is affine then [ ] HilbW = W n n The proofs of these facts are left to the reader. See also [81].

4.4 Quot schemes

4.4.1 Existence

We will now introduce an important class of schemes, the so-called Quot schemes, which generalize the Hilbert schemes. As special cases we will obtain the relative Hilbert schemes.

Let p : X → S be a projective morphism of algebraic schemes, and let OX (1) be a line bundle on X very ample with respect to p. Fix a coherent sheaf H on X and a numerical polynomial P(t) ∈ Q[t]. We define a functor

X/S : ( / )◦ → ( ) QuotH,P(t) schemes S sets called the Quot functor of X/S relative to H and P(t), in the following way: H → X/S ( → ) = coherent quotients Z F, flat over Z, with QuotH,P(t) Z S Hilbert polyn. P(t) on the fibres of X Z → Z where X Z = Z ×S X and HZ denotes the pullback of H on X Z , as usual. When = ( ) X X/Spec(k) S Spec k we write QuotH,P(t) instead of QuotH,P(t) . This definition generalizes the Hilbert functors which are obtained in the case S = Spec(k) and H = OX . 220 4 Hilbert and Quot schemes

= X/S Theorem 4.4.1. The functor Quot QuotH,P(t) is represented by a projective S- scheme X/S → QuotH,P(t) S Proof. We first consider the case S = Spec(k) and X = IP r . From Theorem 4.1.9 it follows that there is an integer m such that for each scheme Z and for each (ϕ : HZ → F) ∈ Quot(Z), letting N = ker(ϕ), all the sheaves N(z), H(z) = r H, F(z), z ∈ Z,arem-regular. Therefore, letting pZ : IP × Z → Z be the projec- tion, we obtain an exact sequence of locally free sheaves on Z:

0 r 0 → pZ∗ N(m) → H (IP , H(m)) ⊗k OZ → pZ∗ F(m) → 0

Moreover, for each m ≥ m there is an exact sequence

0 r 0 r H (IP , O(m − m)) ⊗k pZ∗ N(m) → H (IP , H(m )) ⊗k OZ → pZ∗ F(m ) → 0 where the first map is given by multiplication of sections. This shows that pZ∗ F(m) O [ ,..., ] ( ) uniquely determines the sheaf of graded Z X0 Xr -modules k≥m pZ∗ F k , which in turn determines F. Therefore, letting

H 0(IP r , H(m)) = V we have an injective morphism of functors: → Quot GVˇ ,P(m) given by: ( ) → ( ) Quot Z GVˇ ,P(m) Z (HZ → F) →[V ⊗k OZ → pZ∗ F(m)] ˇ On G = G P(m)(V ) consider the tautological exact sequence

0 → K → V ⊗k OG → Q → 0

r r r Let, moreover, p2 : IP × G → G and p1 : IP × G → IP be the projections. On G we have Γ∗(H) ⊗k OG , which is a sheaf of graded OG [X0,...,Xr ]-modules, ∗(H) O [ ,..., ] F and determines p1 . Consider the subsheaf K G X0 Xr and the sheaf r on G × IP corresponding to the quotient Γ∗(H) ⊗k OG /K OG [X0,...,Xr ], and let G P ⊂ G be the stratum corresponding to P of the flattening stratification of F. Then we claim that a morphism of schemes f : Z → G defines an element of Quot(Z) if and only if f factors through G P , and therefore Quot is represented by G P .The proof of this fact is similar to the one given for the proof of Theorem 4.3.4 and will be left to the reader. Since G P is quasi-projective, to prove that it is projective amounts to proving that it is proper over k, and this can be done using the valuative criterion of properness. Let A be a discrete valuation k-algebra with quotient field L and residue field K , and let ϕ : Spec(L) → G P 4.4 Quot schemes 221 be any morphism. We must show that ϕ extends to a morphism

ϕ˜ : Spec(A) → G P

The datum of ϕ corresponds to an element (ϕL : HL → FL ) of Quot(Spec(L)).The r existence of ϕ˜ will be proved if there is a quotient ϕA : HA → FA on IP × Spec(A) r which is flat over Spec(A) and which restricts to FL over IP × Spec(L).Leti : r r IP ×Spec(L) → IP ×Spec(A) be the inclusion, and take FA = i∗(FL ). Obviously, 1 FA restricts to FL . Moreover, if K L = ker(ϕL ),wehaveR i∗(K L ) = 0 and therefore a surjection HA = i∗(HL ) → FA. We need the following: Lemma 4.4.2. Let X be a scheme, U an open subset of X and i : U → Xthe inclusion. Then for every coherent sheaf F on U we have

Ass(i∗(F)) = Ass(F)

Proof. Since i∗(F)|U = F we have Ass(i∗(F)) ∩ U = Ass(F). Therefore we only need to prove that Ass(i∗(F)) ⊂ U. We may assume that X = Spec(A) and U = Spec(B) are affine. The inclu- sion i corresponds to an injective homomorphism A → B and F = M∼ for an f.g. B-module M.Letx ∈ Ass(i∗(F)) and assume that x ∈ X\U. Then the ideal px ⊂ A annihilates an element mx ∈ i∗(F)x which corresponds to a section m ∈ Γ(V, i∗(F)) for some open neighbourhood V of x. Up to shrinking X we may assume V = X, so that m ∈ Γ(X, i∗(F)) = Γ(U, F) = M is annihilated by the ideal px B.Butpx B = B because x ∈/ U and therefore m = 0: this is a contradiction. The lemma is proved.

From the lemma it follows that Ass(FA) = Ass(FL ): therefore, using the fact that FL is flat over Spec(L) and [84], Prop. III.9.7, we deduce that FA is flat over Spec(A). This concludes the proof of the theorem in the case S = Spec(k) and X = IP r . Assume now that S and X are arbitrary. Consider the closed embedding r j : X → IP × S determined by OX (1). Replacing H by j∗H we can assume that X = IP r × S.Leth, h  0 be such that we have an exact sequence:

M M OIP r ×S(−h ) → OIP r ×S(−h) → H → 0 for some M, M . Then for each S-scheme Z → S and for each (H → ) ∈ IP r ×S/S( → ) Z F QuotH,P(t) Z S we obtain that the composition

M OIP r ×Z (−h) → HZ → F → 0 r × / QuotIP S S (Z → S) is a surjection, i.e. is an element of O(−h)M ,P(t) . This proves that the r × / r × / QuotIP S S QuotIP S S functor H,P(t) is a subfunctor of the functor O(−h)M ,P(t), and this functor r IP × S is evidently represented by QuotO(−h)M ,P(t) . 222 4 Hilbert and Quot schemes

Conversely, a quotient

(O (− )M → ) ∈ IP r ×S/S ( → ) IP r ×Z h F Quot M Z S OX (−h) ,P(t)

IP r ×S/S( → ) is in QuotH,P(t) Z S if and only if the composition

M M OIP r ×Z (−h ) → OIP r ×Z (−h) → F is zero. This means that the condition for an S-morphism

r × / r / Z → IP S S = IP k × S QuotO(−h)M ,P(t) QuotO(−h)M ,P(t)

IP r ×S/S( → ) to define an element of QuotH,P(t) Z S is that it factors through the closed subscheme defined by the entries of the matrix of the homomorphism:

M M OIP r ×Quot(−h ) → OIP r ×Quot(−h)

IP r ×S/S and this is a closed condition. This proves that QuotH,P(t) is represented by a r / IP k × S  closed subscheme of QuotO(−h)M ,P(t) . X/S X/S From the fact that QuotH,P(t) represents the functor QuotH,P(t) it follows that there is a universal quotient

(H → F) ∈ X/S ( X/S ) Quot QuotH,P(t) QuotH,P(t) corresponding to the identity morphism under the identification

Hom(Quot, Quot) = Quot(Quot)

X/S X/S In the case H = OX the scheme Quot is denoted by Hilb and called the OX ,P(t) P(t) relative Hilbert scheme of X/S with respect to the polynomial P(t). It will be sometimes convenient to consider the functor

X/S ◦ QuotH : (schemes/S) → (sets) defined as: X/S( → ) = X/S ( → ) QuotH Z S QuotH,P(t) Z S P(t) This functor is represented by the disjoint union X/S = X/S QuotH QuotH,P(t) P(t) 4.4 Quot schemes 223 which is a scheme locally of finite type, called the Quot scheme of X over S relative to H; it carries a universal quotient HQuot → F. Similarly, we will consider the relative Hilbert scheme of X over S: X/S = X/S Hilb HilbP(t) P(t)

The construction of the Quot scheme commutes with base change; this is a result which follows quite directly from the definition, but it is worth pointing it out:

Proposition 4.4.3 (base change property). Given a projective morphism X → S, a coherent sheaf H on X, and a morphism T → S, there is a natural identification:

XT /T X/S Quot = T ×S Quot HT H Proof. Consider the product diagram X/S X/S T × QuotH → QuotH ↓↓ T → S

X/S X/S The universal quotient H → F on QuotH pulls back to a quotient X /T X/S H T → FT on T × QuotH . We have immediately that the T -scheme / / T × QuotX S endowed with this quotient represents the functor QuotXT T .  H HT

4.4.2 Local properties

Proposition 4.4.4. Let X → S be a projective morphism of algebraic schemes, H X/S a coherent sheaf on X, flat over S, and π : Q = QuotH → S the associated Quot scheme over S. Let s ∈ Sbeak-rational point and q ∈ π −1(s) = Q(s) corresponding to a coherent quotient f : H → F with kernel K. Let

fs : H(s) → F(s) be the restriction of f to the fibre X(s), whose kernel is K(s) = K ⊗ OX(s) (by the flatness of F). Then there is an exact sequence

dπq 0 → Hom(K(s), F(s)) → t Q −→ t S → Ext1 (K(s), F(s)) q s OX (s) and an inclusion:

ker[o(π )]⊂Ext1 (K(s), F(s)) q OX (s) where o(πq ) : o(OQ,q ) → o(OS,s) is the obstruction map of the local homomor- phism πq : OS,s → OQ,q . In particular, π is smooth at q if Ext1 (K(s), F(s)) = 0. OX (s) 224 4 Hilbert and Quot schemes

Proof. A vector in ker(dπq ) corresponds to a commutative diagram:

Spec(k[ ]) → Q ↓↓π Spec(k) → S such that the upper horizontal arrow has image {q}. The above diagram corresponds to an exact and commutative diagram of sheaves on X(s):

00 ↓↓ K(s) −→ K(s) → 0 ↓↓ 0 → H(s) → H(s)[ ]−→H(s) → 0 ↓↓ ↓ 0 → F(s) → F(s) −→ F(s) → 0 ↓↓ ↓ 00 0 where the middle row is exact by the flatness of H. Replacing the middle row by its pushout under H(s) → F(s) we see that this diagram is equivalent to the following one:

00 ↓↓ K(s) = K(s) → 0 ↓↓ 0 → F(s) → P −→ H(s) → 0 ↓ ↓ 0 → F(s) → F(s) −→ F(s) → 0 ↓↓ ↓ 00 0 and therefore we deduce that ker(dπq ) = Hom(K(s), F(s)). Now consider A in A and a commutative diagram

ϕ A ←− OQ,q

↑ η ↑ πq (4.17) B ←− OS,s ϕ˜ where η is a small extension in A. This diagram corresponds to an exact diagram of sheaves on X: 4.4 Quot schemes 225 0 ↓ KA ↓ γ : 0 → H(s) → H ⊗k B −→ H ⊗k A → 0 (4.18) ↓ FA ↓ 0 where the row is exact by the flatness of H over S. By pushing out by the quotient fs : H(s) → F(s) and then pulling back by α : KA → H ⊗k A we obtain an element

∗ 1 1 [α f ∗(γ )]∈Ext (K , F(s)) = Ext (K(s), F(s)) s OX ⊗A A OX (s) By construction this element vanishes if and only if the previous diagram can be embedded in a commutative diagram with exact rows and columns:

00 0 ↓↓ ↓ 0 → K(s) → KB → KA → 0 ↓↓ ↓ 0 → H(s) → H ⊗k B −→ H ⊗k A → 0 ↓↓ ↓ 0 → F(s) → FB → FA ↓↓ ↓ 00 0

X/S The middle column of this diagram is an element of QuotH (Spec(B)), which cor- responds to a homomorphism ϕ : OQ,q → B making the diagram ϕ A ←− OQ,q

↑ η ϕ ↑ πq

B ←− OS,s ϕ˜ commutative. Therefore we have associated an element of Ext1 (K(s), F(s)) to OX (s) each diagram (4.18). It is straightforward to check that this correspondence is linear. Taking η : k[ ]→k we get an inclusion

coker(dπ ) ⊂ Ext1 (K(s), F(s)) q OX (s) Taking any small extension η in A we can apply Proposition 2.1.7 to yield the con- clusion.  226 4 Hilbert and Quot schemes

Corollary 4.4.5. Under the assumptions of 4.4.4, if

Ext1 (K(s), F(s)) = 0 OX (s) then π : Q → S is smooth at q of relative dimension dim Hom(K(s), F(s)) .

When S = Spec(k) we obtain the following “absolute” version of Proposition 4.4.4.

Corollary 4.4.6. If X is a projective scheme, H a coherent sheaf on X and X f : H → F a coherent quotient of H with ker( f ) = K then, letting Q = QuotH, we have:

T[ f ] Q = Hom(K, F)

1 and the obstruction space of OQ,[ f ] is a subspace of Ext (K, F). In particular, if Ext1(K, F) = 0 then Q is nonsingular of dimension dim Hom(K, F) at [ f ].

A special case of Proposition 4.4.4 is the following:

Proposition 4.4.7. Let p : X → S be a projective flat morphism of algebraic X / schemes, and π : Hilb S → S the relative Hilbert scheme. For a closed point s ∈ SletX= X (s) be the fibre over s and let Z ⊂ X be a closed subscheme with ideal sheaf I ⊂ OX . Then there is an exact sequence:

π 0 X /S d [Z] 1 0 → H (Z, N / ) → T[ ]Hilb −→ T S → Ext (I, O ) Z X Z s OX Z

If, moreover, Z ⊂ X is a regular embedding then the above exact sequence becomes:

π 0 X /S d [Z] 1 0 → H (Z, NZ/ X ) → T[Z]Hilb −→ Ts S → H (Z, NZ/ X ) (4.19)

If Ext1 (I, O ) = (0) (resp. H 1(Z, N / ) = (0) in the case where Z ⊂ Xisa OX Z Z X 0 regular embedding) then π is smooth at [Z] of relative dimension h (Z, NZ/ X ).

NOTES 1. One should compare the statement of 4.4.7 with 3.2.12, since the local relative Hilbert ˆ functor of Z in X relative to X → S is prorepresented by the local ring O X/S . Hilb[Z] 2. Our presentation of the Quot schemes is an adaptation of the one given in [90]. For a description of the sheaf of differentials of the Quot schemes see [117]. 3. When p : X → S is a projective flat morphism of integral schemes the relative Hilbert X / scheme Hilb S has a component isomorphic to S, parametrizing the fibres of p, and a com- [ ] ponent isomorphic to X , identified with X 1 . 4.5 Flag Hilbert schemes 227 4.5 Flag Hilbert schemes

4.5.1 Existence

Fix an integer r ≥ 1 and an m-tuple of numerical polynomials

P(t) = (P1(t),...,Pm(t)), m ≥ 1

For every scheme S we let:

r X1 ⊂···⊂Xm ⊂ IP × S r ( ) = (X ,...,X ) : FHP(t) S 1 m S-flat closed subschemes with Hilbert polynomials P(t)

This clearly defines a contravariant functor: r : ◦ → FHP(t) (schemes) (sets) called the flag Hilbert functor of IPr relative to P(t). When m = 1 the flag Hilbert functors are just ordinary Hilbert functors. ≥ r Theorem 4.5.1. For every r 1 and P(t) as above, the flag Hilbert functor F HP(t) r r is represented by a projective scheme FHP(t), called the flag Hilbert scheme of IP relative to P(t), and by a universal family: W ⊂···⊂W ⊂ r × r 1 m IP FHP(t) ↓ r FHP(t) Proof. We will only prove the theorem in the case m = 2, leaving to the reader the task of extending the proof to the general case. By applying Corollary 4.1.10 twice we can find an integer µ such that simulta- r neously for i = 1, 2 we have that for every closed subscheme Xi ⊂ IP with Hilbert ( ) I µ X X polynomial Pi t the sheaf of ideals Xi is -regular. For every such 1 and 2 we thus have: µ + 0 r r h (IP , IX (µ)) = − P (t) =: N i = 1, 2 i r i i

= 0( r , O (µ)) = r Let V H IP IP r . Consider the Hilbert scheme Hi HilbP (t) with uni- r 1 versal family Vi ⊂ IP × Hi , i = 1, 2. On the product H1 × H2 consider the pullback of the two universal families with respect to the projections: V × ( × ) ⊂ r × × , = , i Hi H1 H2 IP H1 H2 i 1 2

r and denote by q : IP × H1 × H2 → H1 × H2 the projection. Because of the choice of µ and by Theorem 4.2.5 we have that q∗Ii (µ) is a locally free subsheaf 228 4 Hilbert and Quot schemes

⊗ O = , of V k H1×H2 of rank Ni , with locally free cokernel, i 1 2. Consider the composition ϕ : I (µ) ⊂ ⊗ O → ⊗ O / I (µ) q∗ 2 V k H1×H2 V k H1×H2 q∗ 1 and let F ⊂ H1 × H2 be the vanishing scheme of ϕ (Example 4.2.8). Note that we have q∗I2(µ) ⊗ OF ⊂ q∗I1(µ) ⊗ OF ⊂ V ⊗ OF (4.20) We now pull back to F the two universal families, i = 1, 2:

r r IP × F IP × Hi ∪∪ W := V × → V i i Hi F i

↓ k ↓

F → Hi I (µ) ⊗ O = I (µ) = , Claim: q∗ i F k∗ Wi , i 1 2. We have natural homomorphisms: β : I (µ) ⊗ O → I (µ) i q∗ i F k∗ Wi Because of the µ-regularity and of Theorem 4.2.5 for every x ∈ F we have isomor- phisms I (µ) ⊗ ( ) =∼ 0( r , I (µ)) =∼ I (µ) ⊗ ( ) q∗ i x k x H IP Wi (x) k∗ Wi x k x I (µ) β Since the sheaf k∗ Wi is locally free, from Nakayama’s lemma it follows that 1 and β2 are isomorphisms, and this proves the claim. From the claim and from (4.20) we deduce that I (µ) ⊂ I (µ) k∗ W2 k∗ W1 (4.21) The µ-regularity and Proposition 4.1.1(iii) imply that the natural homomorphisms ∗ I (µ) → I (µ), = , k k∗ Wi Wi i 1 2 ⊂ I W ⊂ W are surjective; because of (4.21) we deduce that IW2 W1 , hence 1 2. (W , W ) ∈ r ( ) Therefore 1 2 FHP(t) F . ( ,(W , W )) r Claim: The pair F 1 2 represents the functor FHP(t). (X , X ) ∈ r ( ) X ⊂ X ⊂ Let S be a scheme and let 1 2 FHP(t) S . By definition 1 2 r IP × S are flat over S with Hilbert polynomials P1(t) and P2(t) respectively. Let f : IP r × S → S be the projection. We have induced classifying morphisms

g1 : S → H1, g2 : S → H2 4.5 Flag Hilbert schemes 229 which together define a morphism

g : S → H1 × H2

Arguing as before, we see that ∗ I (µ) =∼ I (µ), = , g q∗ i f∗ Xi i 1 2 (4.22) X ⊂ X I (µ) ⊂ I (µ) The fact that 1 2 implies that X2 X1 , hence that I (µ) ⊂ I (µ) f∗ X2 f∗ X1 This, together with (4.22), in turn implies that g factors through F. Since clearly we have X1 = S ×F W1, X2 = S ×F W2 = r r the claim follows, hence F FHP(t). Moreover, FHP(t) is projective because it is a closed subscheme of H1 × H2.  r From the definition it follows that the closed points of FHP(t) are in 1–1 corre- r spondence with the m-tuples (X1,...,Xm) of closed subschemes of IP such that Xi has Hilbert polynomial Pi (t) and

X1 ⊂ X2 ⊂···⊂ Xm

Such an m-tuple is called a flag of closed subschemes of IP r . We will denote by [ ,..., ] r X1 Xm the point of FHP(t) parametrizing such a flag. From the proof of The- r orem 4.5.1 it follows that FHP(t) is a closed subscheme of m Hilbr Pi (t) i=1 We will denote the projections by

pr : FHr → Hilbr , i = 1,...,m i P(t) Pi (t) For every subset I ⊂{1,...,m} with cardinality µ we can consider the µ-tuple of r polynomials P (t) = (P ,...,P µ ) and the flag Hilbert scheme FH .Wehave I i1 i PI(t) natural projection morphisms

pr : FHr → FHr I P(t) PI(t) of which the pri ’s are special cases. The flag Hilbert schemes are generalizations of the flag varieties (fibres´ en drapeaux in [77], ch. 1, § 9.9), which parametrize flags of linear subspaces of IP r . If Z ⊂ IP r is a closed subscheme having Hilbert polynomial Q(t), and P(t) is an m-tuple of numerical polynomials as above, one can define the flag Hilbert scheme of Z relative to P(t) Z : ◦ → FHP(t) (schemes) (sets) 230 4 Hilbert and Quot schemes

r by an obvious modification of the definition of FHP(t). It is straightforward to prove that FHZ is represented by a closed subscheme FHZ of the scheme FHr , where P(t) P(t) P˜ (t) ˜ P(t) = (P1(t),...,Pm(t), Q(t))

Precisely, letting pr : FHr → Hilbr Q P˜ (t) Q(t) Z be the projection, one has an identification of FHP(t) with the scheme-theoretic fibre −1([ ]) [ ]∈ r prQ Z of prQ over the point Z HilbQ(t). It is convenient to consider the disjoint union Z = Z FH FHP(t) P(t) which is a scheme locally of finite type, and call it the flag Hilbert scheme of Z. Another variation on the same theme is the following. Given closed subschemes X ⊂ Z ⊂ IP r , and an m-tuple of numerical polynomials P(t), one can consider flags of closed subschemes of Z containing X, namely m-tuples (Y1,...,Ym) of closed subschemes of Z having Hilbert polynomials P(t) and such that

X ⊂ Y1 ⊂ Y2 ···⊂Ym ⊂ Z

Again there is an obvious generalization of the definition of the corresponding flag Hilbert functor and of the proof of its representability by a projective scheme. These generalized flag Hilbert schemes can be fruitfully used in concrete geometrical sit- uations, like, for example, in the study of families of pointed closed subschemes, or of reducible closed subschemes, of a given projective scheme.

4.5.2 Local properties

Consider a projective scheme Z and let (X1,...,Xm) be a flag of closed subschemes of Z;letIi ⊂ OZ be the ideal sheaf of Xi . We have inclusions

Im ⊂ Im−1 ⊂···⊂I1 and surjections O → O →···→O → Xm Xm−1 X1 0

Definition 4.5.2. The normal sheaf of (X1,...,Xm) in Z is the sheaf of germs of commutative diagrams of homomorphisms of OZ -modules of the following form:

Im ⊂ Im−1 ⊂···⊂ I1 ↓ σm ↓ σm−1 ↓ σ1 O → O →···→ O Xm Xm−1 X1

It is denoted by N(X1,...,Xm )/Z . 4.5 Flag Hilbert schemes 231

Note that we have an obvious homomorphism of “projection”:

N( ,..., )/ → N( ,..., )/ X1 Xm Z Xi1 Xiµ Z for each choice of a subset I ={i1,...,iµ}⊂{1,...,m}. In particular, we have homomorphisms:

→ , = ,..., N(X1,...,Xm )/Z NXi /Z i 1 m

Proposition 4.5.3. Let Z be a projective scheme and let (X1,...,Xm) be a flag of closed subschemes of Z. Then: (i) There is a natural identification:

Z = 0( , ) T[X1,...,Xm ]FH H Z N(X1,...,Xm )/Z (4.23)

(ii) If Xi ⊂ Xi+1 is a regular embedding for all i = 1,...,m−1 and Xm is regularly embedded in Z then the obstruction space of the local ring O Z is FH ,[X1,...,Xm ] 1( , ) contained in H Z N(X1,...,Xm )/Z . Proof. For simplicity we give the proof in the case m = 2, i.e. in the case of a flag (X, Y ) of closed subschemes of Z. The general case can be treated similarly. (i) Let IX ⊂ OZ and IY ⊂ OZ be the ideal sheaves of X and of Y . We can represent a first-order deformation Y ⊂ Z × Spec(k[ ]) of Y in Z by an ideal sheaf IY ⊂ OZ [ ] fitting into the commutative and exact diagram:

00 ↓↓ IY → IY → 0 ↓↓ 0 → OZ → OZ [ ]→OZ → 0 ↓↓↓ 0 → OY → OY → OY → 0

The flatness of Y over Spec(k[ ]) follows from the exactness of the last row (Lemma A.9). This diagram is equivalent to the following one, deduced after pushing out the second row by the homomorphism OZ → OY :

00 ↓↓ IY = IY → 0 ↓↓ 0 → OY → OY ⊕ OZ → OZ → 0 ↓↓ 0 → OY → OY → OY → 0 232 4 Hilbert and Quot schemes

Therefore we have a convenient representation of the given first-order deformation as the middle vertical exact sequence of the previous diagram, which we rewrite:

0 → IY → OY ⊕ OZ → OY → 0

Composing the first inclusion with the first projection OY ⊕ OZ → OY we ob- σ ∈ 0( , ) = (I , O ) Y tain the section Y H Y NY/Z HomOZ Y Y corresponding to .Let’s assume now that we have an element (X , Y) ∈ FHZ (k[ ]). By putting together the above constructions for X and for Y we obtain the following commutative diagram:

0 → IY → OY ⊕ OZ → OY → 0 ∩↓ ↓ 0 → IX → OX ⊕ OZ → OX → 0 where the condition X ⊂ Y corresponds to the condition OY → OX , and this is in turn equivalent to the condition that composing the inclusions on the left with the first projections in the middle terms we obtain a commutative diagram:

IY ⊂ IX ↓ σY ↓ σX OY → OX

This is the global section of N(X,Y )/Z corresponding to (X , Y) in (4.23). Conversely, given a global section

IY ⊂ IX 0 ↓ σ ↓ τ ∈ H (Z, N(X,Y )/Z ) OY → OX one finds a first-order deformation of (X, Y ) by repeating backwards the above con- struction. (ii) Let A be in ob(A) and let

ξ : X ⊂ Y ⊂ Z × Spec(A) be an infinitesimal deformation of the flag (X, Y ).Let

η : 0 → k → A˜ → A → 0 [ ]∈ ( , ) = ⊂ define an element e Exk A k , where m A˜ 0. Since the embeddings X Y and Y ⊂ Z are regular we can find an affine open cover U ={Ui = Spec(Pi )}i∈I of Z such that Yi := Y ∩ Ui is a complete intersection in Ui and Xi := X ∩ Ui is a complete intersection in Yi . Therefore, letting Xi = X ∩ Ui , Yi = Y ∩ Ui , there are liftings ˜ ˜ Yi ⊂ Ui × Spec(A) ∪∪ Yi ⊂ Ui × Spec(A) 4.5 Flag Hilbert schemes 233 and ˜ ˜ Xi ⊂ Yi ∪∪ Xi ⊂ Yi which together give local liftings of the flags (Xi , Yi ):

˜ ˜ ˜ Xi ⊂ Yi ⊂ Ui × Spec(A) ∪∪ ∪ Xi ⊂ Yi ⊂ Ui × Spec(A)

In order to find a lifting (X˜, Y˜) of (X , Y) we must be able to choose the local liftings ˜ ˜ (Xi , Yi ) so that on every Uij := Ui ∩ U j we have

(X˜ , Y˜ ) = (X˜ , Y˜ ) i|Uij i|Uij j|Uij j|Uij

At the level of ideals we have: ⊂ ⊂ IYi IXi Pi IY ⊂ IX ⊂ PiA := Pi ⊗k A i i ˜ I ˜ ⊂ I ˜ ⊂ P ˜ := Pi ⊗k A Yi Xi i A We have a commutative and exact diagram:

I ˜ → IX → 0 Xi i ↓↓ → → → → 0 Pi Pi A˜ PiA 0 ↓↓↓ 0 → Pi /IX → P ˜ /I ˜ → PiA/IX → 0 i i A Xi i → / which, after pushing out the middle row by Pi Pi IXi , gives the following one:

= IXi IXi ↓↓ → / → → → 0 Pi IXi QiA PiA 0 ↓↓ 0 → Pi /IX → P ˜ /I ˜ → PiA/IX → 0 i i A Xi i where = ( / ) QiA Pi IXi PiA Pi ˜ therefore the datum of Xi corresponds to the middle vertical sequence of this dia- gram:

0 → IX → QiA → P ˜ /I ˜ → 0 (4.24) i i A Xi 234 4 Hilbert and Quot schemes ˜ ˜ Repeating the analogous construction for Yi we obtain that Yi is determined by an exact sequence: 0 → IY → QiA → P ˜ /I ˜ → 0 (4.25) i i A Yi ˜ ˜ Since Xi ⊂ Yi the exact sequences (4.24) and (4.25) fit together:

0 → IY → QiA → P ˜ /I ˜ → 0 i i A Yi ∩ ↓ 0 → IX → QiA → P ˜ /I ˜ → 0 i i A Xi ˜ A different choice of the lifting Xi corresponds to a different homomorphism → IXi QiA, and they differ by an element σ ∈ ( , / ) = ( , / ) ˜ HomP ˜ IX Pi IX HomP IX Pi IX Xi i A i i i i i ˜ Similarly, a different choice of the lifting Yi corresponds to an element

σ ˜ ∈ HomP (IY , Pi /IY ) Yi i i i

The condition that the pair (σ ˜ ,σ˜ ) defines another lifting of the flag (Xi , Yi ) is Xi Yi that the following diagram ⊂ IYi IXi ↓ σ ˜ ↓ σ ˜ Yi Xi / → / Pi IYi Pi IXi commutes. This is precisely the condition that (σ ˜ ,σ˜ ) ∈ Γ(Ui , N(X,Y )/Z ). There- Xi Yi ˜ fore the set of liftings of (Xi , Yi ) over Spec(A) is in 1–1 correspondence with Γ(Ui , N(X,Y )/Z ). It follows from this analysis that for all i, j ∈ I we have a (σ ,τ ) ∈ Γ( , ) (X˜ , Y˜ ) section ij ij Uij N(X,Y )/Z such that j|Uij j|Uij is obtained from (X˜ , Y˜ ) (σ ,τ ) i|Uij i|Uij after modifying it by the section ij ij . The collection of these sections is a 1-cocycle

1 {(σij,τij)}∈Z (U, N(X,Y )/Z ) which defines a cohomology class o(X ,Y)([e]). It is straightforward to verify that this class is independent of the choices made and that o(X ,Y)([e]) = 0 if and only if a lifting (X˜, Y˜) exists.  If a flag (X1,...,Xm) of closed subschemes of a projective scheme Z satisfies the conditions of Proposition 4.5.3(ii), i.e; if Xi ⊂ Xi+1 is a regular embedding for all i = 1,...,m − 1 and Xm is regularly embedded in Z, we say that the flag (X1,...,Xm) is regularly embedded in Z. Remarks 4.5.4. (i) One can adapt the above proof to the case m = 1 to obtain an- other, more intrinsic, proof of Propositions 3.2.1(ii) and 3.2.6. (ii) In the case m = 2 considered in the proof of Proposition 4.5.3, to the flag X ⊂ Y of closed subschemes of Z there is associated a diagram of normal sheaves: 4.6 Examples and applications 235

NY/Z ↓ → → → ⊗ O 0 NX/Y NX/Z NY/Z OZ X and we can immediately check that there is a natural identification

0( , ) = 0( , ) × 0( , ) H Z N(X,Y )/Z H X NX/Z 0( , / ⊗O O ) H Y NY/Z H X NY Z Z X

(iii) It is an immediate consequence of the definition that, given a flag (X1,...,Xm) of closed subschemes of a projective scheme Z, the normal sheaf N(X1,...,Xm )/Z is an O Xm -module. This implies that in the case when we have a regularly embedded flag consisting of 0-dimensional subschemes of Z,wehave

1( , ) = 1( , ) = H Z N(X1,...,Xm )/Z H Xm N(X1,...,Xm )/Z 0

Z and from Proposition 4.5.3(ii) we deduce that FH is nonsingular at [X1,...,Xm]. For example, if we consider a projective nonsingular curve Z then it follows that FHZ , which parametrizes flags of effective divisors (D ,...,D ) of degrees (n1,...,nm ) 1 m n1 < n2 < ···< nm, is nonsingular. (iv) It is possibile to give a notion of flag Quot scheme generalizing the flag Hilbert schemes. This seems not to have been considered in the literature yet.

NOTES 1. The flag Hilbert schemes have been considered in [102]. The proof of Theorem 4.5.1 given here has appeared in [163]. More recent references about flag Hilbert schemes are [30] and [62] (where they are called “nested Hilbert schemes”) and [151].

4.6 Examples and applications

4.6.1 Complete intersections

We have already discussed some properties of the local Hilbert functor of a complete intersection X ⊂ IP r , which of course correspond to local properties of the Hilbert r 1 scheme Hilb at [X]. It is easy to check that, despite the fact that H (X, NX ) = (0) in general, every complete intersection X is unobstructed in IP r . We may assume dim(X)>0. Let’s suppose that X ⊂ IP r , r ≥ 2, is the complete intersection of r − n hypersurfaces f1,..., fr−n of degrees d1 ≤ d2 ≤ ... ≤ dr−n respectively, n < r. (1) (m) 0 r Consider a basis Φ ,...,Φ of ⊕ j H (IP , O(d j )) where

Φ(h) = (φ(h),...,φ(h) ) 1 r−n 236 4 Hilbert and Quot schemes

= ,..., φ(h) ∈ [ ,..., ] ,..., h 1 m, and the j k X0 Xr . Consider indeterminates u1 um and the (r − n)-tuple m + Φ(h) = ( + φ(1) +···+ φ(m),..., + φ(1) +···+ φ(m) ) f uh f1 u1 1 um 1 fr−n u1 r−n um r−n h=1 (4.26) [ , ]= [ ,..., , ,..., ] of elements of the polynomial ring k u x k u1 um X0 Xr . ( + Φ(h)) Let K• f h uh be the Koszul complex relative to (4.26) and (h) m+r+1 ∆ := Supp[H1(K•(f + uhΦ ))]⊂A = Spec(k[u, x]) h m+r+1 m m Denoting by p : A → A the projection, U := A \p(∆) is the set of points ∈ m ( + Φ(h)) u A such that K• f h th is exact; U is an open set containing the origin. In IP r × Am consider the closed subscheme X = [ , ]/( + φ(1)+···+ φ(m),..., + φ(1) +···+ φ(m) ) Proj k u x f1 u1 1 um 1 fr−n u1 r−n um r−n m the projection π : X → A and its restriction πU : XU → U, where −1 XU := π (U). All the fibres of πU are complete intersections of multidegree (d1,...,dr−n) and X (0) = X. The Hilbert polynomial of a complete intersection depends only on its multidegree because it can be computed using the Koszul com- plex: it follows that all the fibres of πU have the same Hilbert polynomial P(t) and r therefore πU is a flat family of deformations of X in IP . In an obvious way the tan- 0 r gent space of U at 0 can be identified with ⊕ j H (IP , O(d j )), and the characteristic map with the restriction 0 r 0 ϕ :⊕j H (IP , O(d j )) →⊕j H (X, OX (d j )) Since dim(X)>0themapϕ is surjective, as one easily verifies using the Koszul complex; since, moreover, U is nonsingular at 0 from Proposition 4.3.6 it follows r [ ] that HilbP(t) is smooth at X and the classifying map is smooth. From this it also r follows that complete intersections are parametrized by an open subset of HilbP(t). It is interesting to observe that the closure of this open set may contain points parametrizing nonsingular subschemes of IP r which are not complete intersections. An example of such a subscheme is given by a trigonal canonical curve C ⊂ IP 4:the quadrics containing C intersect in a rational cubic surface S, so it is not a complete intersection since it has degree 8; but [C] is in the closure of the family of complete intersections of three quadrics. It is apparently unknown whether a similar phenom- enon may occur in IP 3, namely whether there are nonsingular curves in IP 3 which are flat limits of complete intersections without being complete intersections. See [51] for more about this. The Kodaira–Spencer map of the families πU has been studied in [161] in the case of complete intersections of dimension ≥ 2: πU has general moduli except for surfaces of multidegrees (4), (2, 3), (2, 2, 2) (respectively in IP 3,inIP 4 and in IP 5), i.e. for complete intersection K3-surfaces. The special case of hypersurfaces had already been considered in [107] (see Example 3.2.11(i), page 134). 4.6 Examples and applications 237

4.6.2 An obstructed nonsingular curve in IP 3

3 We will prove that the Hilbert scheme HilbIP has an everywhere nonreduced com- ponent whose general point parametrizes a nonsingular curve of degree 14 and genus 24. It will follow that every curve parametrized by a general point of is obstructed in IP 3. This example is due to Mumford ([129]). A general element of is constructed as follows. Let F ⊂ IP 3 be a nonsingular cubic surface, E, H ⊂ F respectively a line and a plane section in F.LetC ⊂ F be a general member of the linear system |4H +2E|. Using Bertini’s theorem one easily checks that C is irreducible and nonsingular; its degree and genus are (C · H) = 14 1 ( − · ) + = and 2 C H C 1 24. From the exact sequence:

0 → KC (H) → NC → OC (3H) → 0 we see that

1 1 0 0 h (C, NC ) = h (C, OC (3H)) = h (C, KC (−3H)) = h (C, OC (2E)) = 1 where the last equality follows easily from the exact sequence

0 → OF (−4H) → OF (2E) → OC (2E) → 0

0 1 0 0 and from h (OF (−4H)) = 0 = h (OF (−4H)) and h (OF (2E)) = h (OF (E)) = 1. Moreover, the linear system |C|=|4H + 2E| has dimension

0 dim(|C|) = 1 + dim(|C|C ) = h (C, KC (H)) = 37 and therefore, since every curve C is contained in a unique cubic surface (because 9 < 14), the dimension of the family W of all curves C we are considering is 0 1 19 + 37 = 56 = 4 · 14 but they satisfy h (C, NC ) = 56 + h (C, NC ) = 57. 3 We will prove that W is an open set of a component of HilbIP and this will imply that is everywhere nonreduced. Our assertion will be proved if we show that our curves C are not contained in a family whose general member is a curve D not con- tained in a cubic surface. But on every such D the line bundle OD(4) is nonspecial 0 and therefore, by Riemann-Roch, h (D, OD(4)) = 33, hence D is contained in a pencil of quartic surfaces. Let G1, G2 be two linearly independent quartics contain- ing D: they are both irreducible because otherwise D would be either contained in a plane or in a quadric, which is not the case because there are no nonsingular curves of degree 14 and genus 24 on such surfaces. We have G1 ∩ G2 = D ∪ q where q is a conic; since q has at most double points D has at most triple points and therefore G1 and G2 cannot be simultaneously singular at any point of D, thus the general quartic surface G containing D is nonsingular along D. By applying Riemann-Roch on G we obtain dim(|D|G) = 24. Therefore, since G is not a general quartic surface (because D is not a complete intersection), we see that the family of pairs (D, G) has dimension ≤ 33 + 24 = 57 so that the family Z of curves D has dimension ≤ 56. This shows that the family W, which has dimension 56, cannot be in the closure of Z and this proves the assertion. 238 4 Hilbert and Quot schemes

It is instructive to observe that we can write the linear system |C| on a nonsingular cubic surface F as |4H + 2E|=|6H − 2(H − E)| and this means that we can find a sextic surface F6 such that F ∩ F6 = C ∪ q1 ∪ q2 where q1 and q2 are disjoint conics; if [C]∈ is general then one can show that q1, q2 and F6 can be chosen to be nonsingular. 3 There is another component R of HilbIP whose general point parametrizes a nonsingular curve C of degree 14 and genus 24 such that

C ∪ E ∪ Γ = F3 ∩ F6 where E is a line and Γ is a rational normal cubic which are disjoint. We have in this case |C |=|6H − E − Γ | and

1 1 0 h (C , NC ) = h (C , OC (3H)) = h (C , KC (−3H))

= 0( , O ( − − Γ))= 0( , O ( − − Γ))= h C C 2H E h F3 F3 2H E 0 Thus C is unobstructed. We refer the reader to [41] for another point of view about this example. This is the first published example of an obstructed space curve. Many others have since appeared in the literature (see [78], [162], [79], [50], [103], [187], [22], [125], [80]). A final word on the search for pathologies of this kind is contained in [183], where 3 it is shown that virtually every singularity can appear as a point of HilbIP parame- trizing a nonsingular curve. For examples of obstructed nonsingular curves in higher dimensional projective spaces see [35], [55].

4.6.3 An obstructed (nonreduced) scheme

In IP 3 consider the scheme X = Proj k[X0,...,X3]/J where = ( , , , 2) J X1 X2 X1 X3 X2 X3 X3 X is supported on the reducible conic defined by the equations

X1 X2 = 0, X3 = 0 has an embedded point at (1, 0, 0, 0) and has Hilbert polynomial 2(t + 1) (see Ex- ample 4.2.3(ii)). As in 4.2.3(ii) we consider the flat family parametrized by A1: 3 1 X = Proj k[u, X]/(X1 X2, X1 X3, X2(X3 − uX0), X3(X3 − uX0)) ⊂ IP × A where k[u, X]=k[u, X0,...,X3].WehaveX = X (0).Ifu = 0 then X (u) is a pair of disjoint lines. Let : 1 → 3 g A Hilb2(t+1) 4.6 Examples and applications 239 be the classifying map. If u = 0wehave

1 0 h (X (u), NX (u)) = 0; h (X (u), NX (u)) = 8

Therefore g(u) is a smooth point and the tangent space has dimension 8. In order to show that X is obstructed it suffices to show that

0 h (X, NX )>8 (4.27)

( ) ( ) 3 because g 0 and g u belong to the same irreducible component of Hilb2(t+1). Consider the surjection

: O (− )⊕4 → I → f IP 3 2 X 0 determined by the four equations of degree 2 which define X. Elementary compu- tations, based on the fact that the generators of the ideal are monomials, lead to the following resolution of OX which extends f:

→ O (− ) −→B O (− )⊕4 −→A O (− )⊕4 −→f O → O → 0 IP 3 4 IP 3 3 IP 3 2 IP 3 X 0 (4.28)

A and B being given by the following matrices: ⎛ ⎞ ⎛ ⎞ X3 X3 X3 00 ⎜ −X ⎟ ⎜ −X 0 X 0 ⎟ B = ⎝ 3 ⎠ A = ⎝ 2 3 ⎠ X2 0 −X1 0 X3 −X1 00−X1 −X2

By taking Hom(−, OX ) we obtain the following exact sequence:

t ⊕4 A ⊕4 0 → NX → OX (2) −→ OX (3) from which we deduce that

t 0 0 ⊕4 A 0 ⊕4 H (X, NX ) = ker[H (OX (2)) −→ H (X, OX (3)) ] (4.29)

Using resolution (4.28) it is easy to show that the restriction maps

0 3 0 ϕn : H (IP , O(n)) → H (X, OX (n))

0 0 are surjective if n ≥ 2. This allows us to identify H (X, OX (2)) and H (X, OX (3)) with the homogeneous parts of degree 2 and 3 respectively of k[X0,...,X3]/J. 0 Hence using (4.29) we can represent H (X, NX ) by 4-tuples of polynomials. Pre- 0 cisely, H (X, NX ) is, modulo J, the vector space of 4-tuples

q = (q1, q2, q3, q4) 240 4 Hilbert and Quot schemes

t 4 of homogeneous polynomials of degree 2 such that A q ∈ (J3) . It is easy to find all of them because J is generated by monomials. Computing, one finds that a basis 0 of H (X, NX ) is defined by the following column vectors: 2 2 X1 X1 X0 X2 X2 X0 X3 X0 00 000 0 0 2 0000 0X1 X1 X0 X3 X0 00 0 0 2 0000 000 0X2 X2 X0 X3 X0 0 0000 000 000 0X3 X0 0 In particular, we see that h (X, NX ) = 12, and this proves (4.27). A little extra work shows that [X]=g(0) belongs to two irreducible components 3 ( ) = of Hilb2(t+1). We already know one of them of dimension 8: it contains g u , u 0, and a general point of it parametrizes a pair of disjoint lines. The other component has dimension 11 and a general point of it parametrizes the disjoint union Y = Q ∪{p} of a conic Q and a point p. Note that

0 0 0 h (Y, NY ) = h (Q, NQ) + h (p, N p) = 8 + 3 = 11

1 and h (Y, NY ) = 0. Hence Y is a smooth point of a component of dimension 11 3 of Hilb2(t+1). Therefore it suffices to produce a flat family parametrized by an irre- ducible curve, e.g. A1, Y ⊂ IP 3 × A1 such that Y(0) = X, Y(1) = Y .Hereitis: Y = Proj k[v, X0, X1, X2, X3]/I where = ( , + v , + v , 2 − v2 2) I X1 X2 X1 X3 X1 X0 X2 X3 X2 X0 X3 X0 Clearly Y(0) = X; since

I = (X1, X2, X3 − v X0) ∩ (X3 + v X0, X1 X2) it follows that for all v = 0, Y(v) is the disjoint union of a conic and a point. The flatness of Y follows from [84], Prop. III.9.8. This example shows that in general the Hilbert schemes are reducible and not 3 equidimensional. For a description of Hilb3t+1, which presents several analogies 3 with the one given here of Hilb2(t+1), we refer to [142].

4.6.4 Relative grassmannians and projective bundles

Consider a coherent sheaf E on an algebraic scheme S, and let P(t) = n, where n is S/S a positive integer, be a constant polynomial. Then QuotE,n is a projective S-scheme which will be denoted by Quotn(E) in what follows. We will denote by

ρ : Quotn(E) → S 4.6 Examples and applications 241 the structural projective morphism and by ρ∗E → Q the universal quotient bundle; Q is locally free of rank n.Ifn = 1 then Q is an O ( ) O( ) invertible sheaf: it will be denoted by Quot1(E) 1 ,orsimplyby 1 if no confusion arises. The pair (Quotn(E)/S, Q) represents the functor ◦ Quotn(E) : (schemes/S) → (sets) defined by: ∗ Quotn(E)( f : T → S) ={locally free rk n quotients f E → F on T }

On Quotn(E) we have a tautological exact sequence ∗ 0 → K → ρ (E) → Q → 0 If E is locally free then K is locally free as well and it is called the universal subbun- dle. If E is locally free we define ∨ Gn(E) := Quotn(E ) and call it the grassmannian bundle of subbundles of rank n of E;forn = 1wehave

G1(E) = IP (E) the projective bundle associated to E, according to our definition (which differs from the one adopted in [84], p. 162). The tautological exact sequence on IP (E) is: ∗ ∨ 0 → K → ρ (E ) → OIP (E)(1) → 0 In particular, for a finite-dimensional k-vector space V we have

IP (V ⊗k OS) = IP (V ) × S and more generally, Gn(V ⊗k OS) = Gn(V ) × S

Therefore, if E is locally free on S then Quotn(E) is locally the product of S by a grassmannian; in particular, the projection ρ : Quotn(E) → S is a smooth morphism. Proposition 4.6.1. Let E be a locally free sheaf on the algebraic scheme S, and let ∗ 0 → K → ρ (E) → Q → 0 (4.30) be the tautological exact sequence on Quotn(E) for some 1 ≤ n ≤ rk(E). Then there is a natural isomorphism Ω1 =∼ Hom(Q, K) Quotn(E)/S and therefore =∼ (K, Q) TQuotn(E)/S Hom 242 4 Hilbert and Quot schemes

Proof. Letting B = Quotn(E) consider the product B ×S B with projections : × → = , E E × pri B S B B, i 1 2, and let B×S B be the pullback of on B S B. I ⊂ O ∆ ⊂ × Denote by ∆ B×S B the ideal sheaf of the diagonal B S B. The tautolog- ical exact sequence (4.30) pulls back to two exact sequences:

∗ ∨ ∗ 0 → pr K → E → pr Q → 0 i B×S B i on B ×S B whose restrictions to ∆ coincide, and ∆ is characterized by this prop- erty. This can be also expressed by saying that ∆ is the vanishing scheme of the composition ∗ ∨ ∗ pr K → E → pr Q 1 B×S B 2 Therefore we have a surjective homomorphism: ( ∗Q, ∗K) → I Hom pr2 pr1 ∆ (see 4.2.8) which, restricted to ∆, gives a surjective homomorphism: (Q, K) → I /I2 = Ω1 Hom ∆ ∆ B/S which has to be an isomorphism since both sheaves are locally free and have the same rank. 

Proposition 4.6.2. Let α β 0 → E −→ F −→ G → 0 be an exact sequence of locally free sheaves on the algebraic scheme S, and n ≥ 1 an integer. Then there is a closed regular embedding

Quotn(G) ⊂ Quotn(F) and a natural identification: = ρ∗E∨ ⊗ Q ⊗ O NQuotn(G)/Quotn(F) Quotn(G) ∗ where ρ : Quotn(F) → S is the structure morphism and ρ F → Q is the universal quotient. In particular, if n = 1 we have = ρ∗E∨ ⊗ O( ) ⊗ O NQuot1(G)/Quot1(F) 1 Quot1(G) Proof. Let f : T → S be a morphism. For every locally free rank n quotient

∗ ( f G → H) ∈ Quotn(G)(T ) there is associated, by composition with the surjective homomorphism f ∗(β) : f ∗F → f ∗G, an element

∗ ( f F → H) ∈ Quotn(F)(T ) 4.6 Examples and applications 243

Therefore Quotn(G) is a subfunctor of Quotn(F). Consider the diagram of homo- morphisms on Quotn(F): ρ∗(E)

↓ ρ∗(α)

γ ρ∗(F) −→ Q → 0

↓ ρ∗(β)

ρ∗(G) ↓ 0 Given a morphism f : T → S, an element of

Quotn(F)(T ) = HomS(T, Quotn(F)) belongs to Quotn(G)(T ) if and only if it factors through the closed subscheme ∗ D0(γρ (α)) of Quotn(F). This proves that Quotn(G) is a closed subfunctor of Quotn(F), and therefore the embedding Quotn(G) ⊂ Quotn(F) is closed. More ∗ precisely, this analysis shows that Quotn(G) = D0(γρ (α)). Since Quotn(G) has codimension rk(E)n in Quotn(F) it follows that it is regularly embedded. According to Example 4.2.8 we have a surjective homomorphism: Hom(Q,ρ∗(E)) → I I ⊂ O (G) (G) where Quotn(F) is the ideal sheaf of Quotn . By restricting to Quotn we obtain a surjective homomorphism: (Q,ρ∗E) ⊗ O → I/I2 → Hom Quotn(G) 0 which is an isomorphism because both are locally free and of the same rank.  Corollary 4.6.3. Let 0 → E → F → G → 0 (4.31) be an exact sequence of locally free sheaves on the algebraic scheme S. Then there is a closed immersion IP (E) ⊂ IP (F) and a natural identification: ∗ NIP (E)/IP (F) = ρ G ⊗ OIP (E)(1) In particular, to every exact sequence (4.31) with E an invertible sheaf there corre- sponds a section σ : S → IP (E) ⊂ IP (F) of the projective bundle IP(F) → S whose normal bundle is G ⊗ E∨. 244 4 Hilbert and Quot schemes

Proof. Only the last assertion requires a proof. It follows by observing that ∨ OIP (F)(1) ⊗ OIP (E) = OIP (E)(1) = E and therefore the formula for the normal bundle is a direct consequence of Proposi- tion 4.6.2.  Remark 4.6.4. Let E be a locally free sheaf on an algebraic scheme S, and let ∨ ∗ ∨ 0 → Q → ρ (E) → K → 0 (4.32) ∨ be the dual of the tautological exact sequence (4.30) on Gn(E) = Quotn(E ). Ten- soring with Q we obtain the exact sequence: 0 → Q∨ ⊗ Q → ρ∗(E) ⊗ Q → K∨ ⊗ Q → 0  (4.33)

TGn(E)/S

In the case n = 1 and S = Spec(k) we have E = V a vector space and G1(V ) = IP (V ) = IP ; the dual of the tautological sequence is

0 → OIP (−1) → V ⊗ OIP → TIP (−1) → 0 (4.34) and the sequence (4.33) is the Euler sequence

0 → OIP → V ⊗ OIP (1) → TIP → 0 Therefore (4.33) is a generalization of the Euler sequence. Dualizing (4.34) we get an inclusion = (Ω1 ) ⊂ × ∨ I IP IP IP IP where I := {(x, H) : x ∈ H} is the incidence relation (see Note 2 of Appendix B). From Corollary 4.6.3 we obtain ∗ ∗ ∨ = O ( ) ⊗ O ∨ ( ) ⊗ O NI/IP ×IP p1 IP 1 p2 IP 1 I where p p IP ←−1 IP × IP ∨ −→2 IP ∨ are the projections. Example 4.6.5. Let X be a projective nonsingular variety and L an invertible sheaf on X. Consider the Atiyah extension

0 → OX → EL → TX → 0 r (see page 145). Then ρ : IP (EL ) → X is a IP -bundle, r = dim(X), and IP (OX ) ⊂ IP (EL ) is a section of ρ with normal bundle TX . In the case X = IP := IP (V ), where V is a finite-dimensional vector space, and L = O(1) the Atiyah extension coincides with the Euler sequence (Remark 3.3.10) so that we have IP (EL ) = IP × IP and

IP = IP (OIP ) ⊂ IP (EL ) = IP × IP is the diagonal embedding. 4.6 Examples and applications 245

Example 4.6.6. Let V be a finite-dimensional k-vector space, X ⊂ IP := IP (V ) a projective irreducible nonsingular variety and I ⊂ OIP its ideal sheaf. Then we have inclusions of locally free sheaves on X: ∨ = I/I2 ⊂ Ω1 ⊂ ∨ ⊗ O (− ) NX/IP IP |X V X 1 which induce closed embeddings of projective bundles: ( ∨ ) ⊂ (Ω1 ) ⊂ × ∨ IP NX/IP IP IP |X X IP  {(x, H) : x ∈ H} Recalling the exact sequence → ∨ → Ω1 → Ω1 → 0 NX/IP IP |X X 0 (4.35) we see that we have an identification: ( ∨ ) ={( , ) : ⊂ }⊂ × ∨ IP NX/IP x H Tx X H X IP (4.36) Letting ρ : ( ∨ ) → IP NX/IP X be the structural morphism, n = dim(X), and r = dim(IP ), from the description (4.36) we deduce that for every x ∈ X, the fibre of ρ is identified with the (r −n−1)- dimensional linear system of hyperplanes which are tangent to X at x. Let’s consider the special case when X is a rational normal curve in IP = IP r , r r ≥ 2. Denote by λ the invertible sheaf of degree 1 on X, so that OX (1) = λ .We have a morphism : ( ∨ ) →|λr−2| =∼ r−2 q IP NX/IP IP which associates to (x, H) the divisor q(H) of degree r − 2 such that H · X = 2x + q(H) ( ∨ ) =∼ × r−2 The existence of q implies that IP NX/IP X IP and therefore ∨ = λs ⊕···⊕λs NX/IP r−1 for some s ∈ ZZ. From the exact sequence (4.35) one immediately computes that (r − 1)s = (r + 2)(1 − r) i.e. s =−(r + 2). The conclusion is the following formula for the normal bundle of a rational normal curve X ⊂ IP r : ∼ λr+2 ⊕···⊕λr+2 NX/IP r = (4.37) r−1 For another approach to the same computation see [131]. The normal bundles of rational curves of any degree in IP 3 (resp. in IP r ) have been computed in [63] and [49] (resp. in [153]). 246 4 Hilbert and Quot schemes

Example 4.6.7 (Zappa [188]). Let C be a projective nonsingular connected curve of genus 1 and let 0 → OC → E → OC → 0 (4.38) be an exact sequence of locally free sheaves corresponding to a nonzero element 1 1 of ExtO (OC , OC ) = H (C, OC ); in particular, (4.38) does not split. Consider the C ruled surface ρ : IP (E) → C and let C := IP (OC ) ⊂ IP (E) be the section of ρ corresponding to the subsheaf OC ⊂ E in (4.38). By Corollary 4.6.3 we have 0 N := NC /IP (E) = OC and therefore h (C , N) = 1. If C ⊂ IP (E) is a curve E) such that [C ] and [C ] belong to the same component of HilbIP ( then, since N is the trivial sheaf, either C = C or C ∩ C =∅. But both C and C are sections of ρ and the second possibility implies that the exact sequence (4.38) splits. This E) is a contradiction, and we conclude that [C ] is isolated in HilbIP ( , with a one- dimensional Zariski tangent space. Therefore C is obstructed in IP (E).

∗∗∗∗∗∗

Let H be a coherent sheaf on the projective scheme X, and let P(t) = n be a constant polynomial, where n is a positive integer. Then we have two different Quot schemes associated to these data. X/ X H → F The first one is QuotH,n , whose k-rational points are quotients which are locally free of rank n: it has just been considered. X The other one is QuotH,n.Ak-rational point of this scheme is nothing but a quotient H → F such that F is a torsion sheaf with finite support and h0(F) = n. When n = 1 then F =∼ k(x) for some closed point x ∈ X: therefore we have a natural morphism : X → q QuotH,1 X

(H → F) → Supp(F) X H → F X and QuotH,1 is a scheme over X.Let be a k-rational point of QuotH,1; then ker[H → F] is called an elementary transform of H. The process of passing from H to ker[H → F] is called an elementary transformation centred at x. This construc- tion is classical when X is a projective nonsingular curve and H is locally free. For X generalizations of it see [126]. QuotH,1 is the scheme of elementary transformations of H. (H) X Proposition 4.6.8. Assume that Supp is connected. Then QuotH,1 is connected. Proof. The natural morphism : X → q QuotH,1 X

(H → F) → Supp(F) has image Supp(H).EveryH → F factors as 4.6 Examples and applications 247

H → H ⊗ Ox ↓ ∼ F = Ox ↓ 0

−1 0 ∨ and therefore the fibre q (x) is identified with IP (H (H ⊗ Ox )) which is con- nected. The conclusion follows. 

4.6.5 Hilbert schemes of points

The Hilbert schemes parametrizing 0-dimensional subschemes of a given scheme are inexpectedly complicated and have a variety of applications. Since it is beyond our scope to give an exhaustive overview of the subject, we will only explain some of the basic facts. The reader is referred to [91], [92], [133] for more details. See also [81] for another approach to Hilbert schemes of points. Consider a projective scheme Y and, for a positive integer n, the Hilbert scheme Y [n]. We have already seen in § 4.3 that Y [1] =∼ Y : we will therefore assume n ≥ 2. Let Z ⊂ Y be a closed subscheme of length n. Then 0( , ) = (I , O ) H Z NZ Hom Z y Z y y∈Supp(Z) and 1 H (Z, NZ ) = (0)

Moreover, from the spectral sequence of Exts (see [64], p. 265) we see that 1(I , O ) = 0( , 1(I , O )) = 1(I , O ) Ext Z Z H Y Ext Z Z Ext Z y Z y y∈Supp(Z)

It follows that the local properties of Y [n] at [Z] are determined by the indepen- dent contributions from each of the points of Supp([Z]). We immediately have the following properties: (a) If Z is reduced and supported at n distinct points of Y then [Z] is a nonsingular point of Y [n] if and only if it is supported at nonsingular points of Y . (b) If Y is reduced then the set of [Z]’s with Z supported at n nonsingular points of Y is an open set of dimension n dim(Y ) contained in the nonsingular locus of Y [n].

Another important property is:

(c) [58] If Y is connected then Y [n] is also connected. 248 4 Hilbert and Quot schemes

≥ I ⊂ O Proof. Let n 1 and let Y ×Y [n] be the ideal sheaf of the universal family in Y × Y [n]. Then we have a diagram of morphisms:

Y ×Y [n] QuotI,1

p q

Y [n+1] Y × Y [n] where q is the natural morphism, which is surjective because Supp(I) = Y × Y [n]. The morphism p is defined as follows. Let (y, [Z]) ∈ Y × Y [n] be a k-rational point and let γ : I → k(y, [Z]) be Y ×Y [n] (γ ) ⊂ O [ ] a quotient, which is a k-rational point of QuotI,1 . Then ker Y ×Y n is an ideal sheaf such that ker(γ )OY ×[Z] has colength 1 in I ⊗ OY ×[Z]. Therefore ker(γ )OY ×[Z] defines a subscheme W ⊂ Y of length n + 1 containing Z and x; we define p(γ ) =[W]. The morphism p is clearly surjective. Since Y × Y [n] is connected by induction, we conclude that Y [n+1] is connected by Proposition 4.6.8.  In general, Y [n] is singular and reducible even if Y is nonsingular connected. Notable exceptions are the cases dim(Y ) = 1, 2. If C is a projective nonsingular curve and n ≥ 1 an integer, every closed sub- scheme D ⊂ C of length n is a Cartier divisor, therefore regularly embedded in C. It follows that C[n] is nonsingular of dimension

0 0 h (D, ND) = h (D, OD) = n 1( , ) = ( ) O because H D ND 0 is an obstruction space for C[n],[D] (Theorem 4.3.5(ii)). Actually, C[n] is naturally isomorphic to C(n),then-th symmetric product of C,the isomorphism being given by the cycle map

C[n] → C(n) whichmapsaclosedD ⊂ C to the associated Weil divisor. The case of surfaces is more subtle.

Theorem 4.6.9 (Fogarty [58]). If Y is a projective nonsingular connected surface then Y [n] is nonsingular connected of dimension 2n.

Proof. Let [Z]∈Y [n]. We then have:

i Ext (IZ , OZ ) = (0) i ≥ 3

Moreover, from the exact sequence:

0 → IZ → OY → OZ → 0 we obtain the sequence: 4.6 Examples and applications 249

0 → Hom(OZ , OZ ) → Hom(OY , OZ ) → Hom(IZ , OZ ) 1 1 1 → Ext (OZ , OZ ) → Ext (OY , OZ ) → Ext (IZ , OZ ) 2 2 2 → Ext (OZ , OZ ) → Ext (OY , OZ ) → Ext (IZ , OZ ) → 0

i i Since Ext (OY , OZ ) = H (Y, OZ ) = (0) for i ≥ 1 we see that

2 Ext (IZ , OZ ) = (0) and 1 ∼ 2 ∨ Ext (IZ , OZ ) = Ext (OZ , OZ ) = Hom(OZ , OZ ⊗ ωY )

∨ 0 ∨ = Hom(OZ , OZ ) = H (Z, OZ ) Therefore 2 i i 0 0 0 (−1) dim[Ext (IZ , OZ )]=h (Z, NZ ) − h (Z, OZ ) = h (Z, NZ ) − n i=0

0 Since the left-hand side is independent of Z, it follows that h (Z, NZ ) is also inde- pendent of Z.ButY [n] is connected and has an open set which is nonsingular and of dimension 2n: the conclusion follows.  To see that Y [n] is singular if dim(Y ) = 3 consider IP 3 with homogeneous coor- dinates X0,...,X3 and the subscheme = ( 2, 2, 2, , , ) Z V X1 X2 X3 X1 X2 X1 X3 X2 X3 Then [Z]∈(IP 3)[4]. A computation similar to that of the example of Subsection 4.6.3 shows that the Zariski tangent space of (IP 3)[4] at [Z] has dimension 18. But (IP 3)[4] is connected and has a component which is nonsingular of dimension 12 at its general point: it follows that (IP 3)[4] is singular. For more examples and for a useful detailed introduction to punctual Hilbert schemes we refer the reader to [91].

4.6.6 Schemes of morphisms

Let X and Y be schemes, with X projective and Y quasi-projective. For every scheme S let: F(S) = HomS(X × S, Y × S) This defines a contravariant functor: ◦ F : (schemes) → (sets) called the functor of morphisms from X to Y . ∼ For every Φ ∈ F(S) let ΓΦ ⊂ X × Y × S be its graph. Then ΓΦ = X × S is flat over S and therefore defines a flat family of closed subschemes of X × Y parametrized by S. This means that F is a subfunctor of HilbX×Y . If G ⊂ X × Y × S is a flat family of closed subschemes of X × Y , proper over S, then the projection π : G → X × S is a family of morphisms into X and the locus 250 4 Hilbert and Quot schemes of points s ∈ S such that π(s) is an isomorphism is open (Note 2 of § 4.2). This means that F is an open subfunctor of HilbX×Y , represented by an open subscheme of HilbX×Y , which we denote by Hom(X, Y ). It is called the scheme of morphisms from X to Y . The scheme Hom(X, Y ) contains an open and closed subscheme isomorphic to Y and consisting of the constant morphisms. In particular, if the only morphisms f : X → Y are the constant ones then Hom(X, Y ) =∼ Y Let X and Y be as above, and consider the contravariant functor ◦ G : (schemes) → (sets) defined as follows: G(S) ={S-isomorphisms X × S → Y × S} Clearly, G is a subfunctor of F. It is easy to prove that G is represented by an open subscheme Isom(X, Y ) of Hom(X, Y ), called the scheme of isomorphisms from X to Y . When X = Y it is denoted by Aut(X) and called the scheme of automorphisms of X. It is a group scheme. The following result follows immediately from Proposition 3.4.2 and Corollary 3.4.3: Proposition 4.6.10. Let f : X → Y be a morphism of algebraic schemes, with X reduced and projective and Y nonsingular and quasiprojective. Then ∼ 0 ∗ T[ f ](Hom(X, Y )) = H (X, f TY ) 1 ∗ and the obstruction space of Hom(X, Y ) at [ f ] is contained in H (X, f TY ). 0 If X is nonsingular then the tangent space to Aut(X) at 1X is H (X, TX ). Let j : X ⊂ Y be a closed embedding of projective nonsingular schemes. Then j induces an inclusion J : Aut(X) ⊂ Hom(X, Y ) such that J(1X ) = j and which is induced by the closed embedding

1X × j : X × X ⊂ X × Y.

It follows that J is a closed embedding. Its differential at 1X is the injective linear map 0 0 H (TX ) → H (TY |X ) coming from the natural inclusion TX ⊂ TY |X . In fact, from the diagram of inclu- sions: X × X ⊂ X × Y ∪∪ ∼ ∆ = Γ j we deduce the commutative diagram: ⊂ N∆/X×X NΓ j / X×Y  TX ⊂ TY |X and we conclude according to § 4.3, Note 3. 4.6 Examples and applications 251

Example 4.6.11. Consider X = IP 1, Y = IP r and j : IP 1 → IP r the r-th Veronese embedding. Locally around [ j] we have a well-defined morphism

r M : Hom(IP 1, IP r ) → HilbIP sending [ j] →[j(IP 1)] with fibre M−1([ j(IP 1)]) an open neighbourhood of the identity in Aut(IP 1). Consider the following diagram consisting of two exact se- quences: 0 ↑ → → ∗ → → 0 TIP 1 j TIP r N j 0 ↑ O ( )r+1 IP 1 r ↑ O IP 1 ↑ 0 From the vertical sequence (the Euler sequence restricted to IP 1) we get

1 ∗ 0 ∗ h ( j TIP r ) = 0, h ( j TIP r ) = r(r + 2) 1( ) = 1( ) = Since h TIP 1 0 from the other sequence we obtain h N j 0 and the exact sequence

→ 0( ) → 0( ∗ ) −→q 0( ) → 0 H TIP 1 H j TIP r H N j 0   1 1 r IP r T Aut(IP ) T[ ]Hom(IP , IP ) T[ ( 1)]Hilb 1IP 1 j j IP 1 r Since the map q can be identified with dM[ j] we see that M and Hom(IP , IP ) are r smooth at [ j] and HilbIP is smooth at [ j(IP 1)]; moreover, ( ( 1, r ) = ( + ) = ( IP r ) + dim[ j] Hom IP IP r r 2 dim j(IP 1) Hilb 3

For more on the schemes Hom(IP 1, X) and applications to uniruledness see [42].

4.6.7 Focal loci

We assume char(k) = 0. Consider a flat family of closed subschemes of a projective scheme Y : q  ⊂ Y × B −→1 Y ↓ q2 (4.39) B parametrized by a scheme B.Let

f :  ⊂ Y × B → Y 252 4 Hilbert and Quot schemes be the composition of the inclusion (4.39) with the projection q1. Denote by

N := N/Y ×B the normal sheaf of  in Y × B, and let

( ∗ ) q2 TB | be the restriction to  of the tangent sheaf along the fibres of q1.Theglobal charac- teristic map of the family (4.39) is the homomorphism:

χ : ( ∗ ) → N q2 TB | defined by the following exact and commutative diagram: 0 ↓ χ ( ∗ ) −→ N q2 TB | ↓ (4.40) 0 → T → TY ×B| → N → 0 ↓ df ↓ ∗ ∗ f TY = f TY For each b ∈ B the homomorphism χ induces a homomorphism

χb : TB,b ⊗ O(b) → N(b)/Y called the local characteristic map of the family (4.39) at the point b. Let ϕ : B → HilbY be the classifying morphism induced by the flat family (4.39). Then if (b) is con- nected, the linear map:

0 0 H (χb) : TB,b → H (N(b)/Y ) is dϕb, the characteristic map of (4.39) at the point b (see page 217). Assume that Y, B and the family (4.39) are smooth. In this case all the sheaves in (4.40) are locally free. From a diagram-chasing it follows that

ker(χ) = ker(df) and therefore:

Proposition 4.6.12.

dim[ f ()]=dim() − rk[ker(χ)] 4.6 Examples and applications 253

Let’s denote by V (χ) the closed subscheme of  defined by the condition: (χ) < { [( ∗ ) ], (N )}= { ( ), ()} rk min rk q2 TB | rk min dim B codimY ×B We will call the points of V (χ) first-order foci of the family (4.39). V (χ) is the scheme of first-order foci, and the fibre of V (χ) over a point b ∈ B:

V (χ)b = V (χb) ⊂ (b) is the scheme of first-order foci at b. If χ has maximal rank, i.e. if χ is injective or has torsion cokernel, then V (χ) is a proper closed subscheme of .Ifχ does not have maximal rank then V (χ) = . The definition of first-order foci at a point b depends only on the geometry of the family (4.39) in a neighbourhood of b. A focus y ∈ V (χ)b is a point where there is an intersection between the fibre (b) and the infinitesimally near ones. One defines higher-order foci inductively: second-order foci are the first-order foci of the family of first-order foci, and so on. Focal loci have been studied classically in the case of families of linear spaces (see e.g. [158]). Recently, they have been applied to the geometry of the theta divisor of an algebraic curve in [37] and [38]. Related work is [31], [33], [39].

NOTES 1. Let Q ⊂ IP 3 be a quadric cone with vertex v,andL ⊂ Q a line. Then = O ( ) ⊂ = O ( ) ⊕ O ( ) NL/Q L 1 NL/IP 3 L 1 L 1 0 1 (see Example 3.2.8(iii), page 131); in particular, H (L, NL/Q) = 2andH (L, NL/Q) = 0. On the other hand, the Hilbert scheme HilbQ is one-dimensional at the point [L] since L moves in a one-dimensional family. It follows that L is obstructed in Q (see [44] for general- izations of this example). 2. Another example of reducible Hilbert scheme is the following, which appears in [170].

Let Y = C × C where C and C are projective nonsingular connected curves of genera g and g respectively and let p Y −→ C ↓ p C be the projections; assume g, g ≥ 2. Consider an effective divisor D = x1+···+xg of degree = +···+ g on C, and an effective divisor D x1 xg of degree g on C , both consisting of distinct points, and let − − Γ = 1( ) + 1( ) = +···+ + +···+ p D p D Cx Cx Cx Cx 1 g 1 g −1 −1 where C = p (x) and Cx = p (x ). Γ is a reduced divisor, has gg nodes and no other x singularity. If either D or D is nonspecial the curve Γ belongs to an irreducible component Y H1 of Hilb of dimension g + g generically consisting of curves of the same form, obtained by moving D and D . When both D and D are special divisors the curve Γ belongs to a linear system of dimension ≥ 3 whose general member is a nonsingular curve and therefore Y belongs to another irreducible component H2 of Hilb which has dimension g + g − 1. The intersection H1 ∩ H2 is irreducible of dimension g + g − 2. 254 4 Hilbert and Quot schemes 4.7 Plane curves

An important refinement of the Hilbert functors derives from the consideration of flat families of closed subschemes of a projective scheme Y having prescribed singulari- ties, i.e. of families all of whose members have the same type of singularity in some specified sense. This leads to the notion of equisingularity and to a related vast area of research. In this section we will only concentrate on the specific case of families of plane curves with assigned number of nodes and cusps: we will show how to con- struct universal families of such curves, whose parameter schemes are called Severi varieties for historical reasons. This is a subject with a long history and a wealth of important results, both classical and modern. Here we will limit ourselves to prov- ing a few basic results and to indicating some of their generalizations and the main references in the literature. We will assume char(k) = 0 in this section.

4.7.1 Equisingular infinitesimal deformations

Let Y be a projective nonsingular variety and X ⊂ Y a closed subscheme with ideal sheaf IX ⊂ OY . Recall ((1.3)) that on X we have an exact sequence of coherent sheaves: → → → → 1 → 0 TX TY |X NX/Y TX 0 Recall that the sheaf := [ → 1] NX/Y ker NX/Y TX is called the equisingular normal sheaf of X in Y (see Proposition 1.1.9, page 16). = Clearly, NX/Y NX/Y if X is nonsingular. By definition, sections of the equisin- gular normal sheaf parametrize first-order deformations of X in Y which are locally trivial, because they induce trivial deformations around every point of X. We recall that an alternative description of the equisingular normal sheaf can be given by means of the sheaf of germs of tangent vectors to Y which are tangent to X

TY (X)⊂TY introduced in § 3.4. In fact, there is an exact sequence (see (3.55)) → ( )→ → → 0 TY X TY NX/Y 0 (4.41)

From the definition it follows that, for every open set U ⊂ Y , Γ(U, TY (X)) con- sists of those k-derivations D ∈ Γ(U, TY ) such that D(g) ∈ Γ(U, I) for every g ∈ Γ(U, I).

∼ Examples 4.7.1. (i) Assume that X is a hypersurface in Y ; then NX/Y = OX (X). ∼ Locally on an affine open set U ⊂ Y we have (NX/Y )|U∩X = OU∩X . Assume that X can be represented by an equation f (x1,...,xn) = 0 in local coordinates on U; then 1 ( ) ⊂ O from the definition of TX it follows that NX/Y |U∩X U∩X is the image of the ideal sheaf (∂ f/∂x1,...,∂f/∂xn) ⊂ OU . We deduce that an equisingular first-order 4.7 Plane curves 255

¯ deformation of X in Y corresponding to a local section g of NX/Y can be written locally as f (x) + g(x) = 0 where ∂ f ∂ f g(x) = a1(x) +···+an(x) ∂x1 ∂xn restricts to g¯. Therefore if Y = IP n and X is a hypersurface of degree d we have an exact sequence → O → I( ) → → 0 IP n d NX/IP n 0 where I ⊂ OIP n is the ideal sheaf locally generated by the partial derivatives of a local equation of X. In the special case of a curve X in a surface Y assume that p ∈ X is a singular point and let f (x, y) = 0 be a local equation of X around p. If p is a node then (∂ f/∂x,∂f/∂y) = m p is just the maximal ideal of p;ifp is an ordinary cusp with principal tangent say y = 0 then (∂ f/∂x,∂f/∂y) = (x, y2) is an ideal of colength 2. (ii) Let X ⊂ IP 2 be a (possibly reducible) plane curve of degree d of equation F(X0, X1, X2) = 0, having δ nodes p1,...,pδ and no other singularity. This case is important because every nonsingular projective curve is birationally equivalent to 2 a nodal plane curve. Denote by ∆ ={p1,...,pδ}⊂IP the 0-dimensional reduced scheme of the nodes of X and by ν : C → X the normalization map. The above analysis shows that sections of H 0(I∆(d)), i.e. curves of degree d which are adjoint to X, cut on X sections of N . This means that X/IP 2 ∗ ∗ ν (N ) = ν [O (d) ⊗ I∆] X/IP 2 X = ν∗O ( )(− − −···− − ) = ω ⊗ ν∗O( ) X d p1 p1 pδ pδ C 3 ν−1( ) ={ , } = ,...,δ where pi pi pi , i 1 , and therefore we have 0( ,ν∗( )) = + − , 1( ,ν∗( )) = h C NX/IP 2 3d g 1 h C NX/IP 2 0 where g is the geometric genus of X. Moreover, since ∗ ∗ ν∗[ν (N )]=ν∗[ν (O (d) ⊗ I∆] X/IP 2 X

= O (d) ⊗ ν∗O (−p − p −···− p − p ) = O (d) ⊗ I∆ = N X C 1 1 δ δ X X/IP 2 we have + h0(X, N ) = h0(C,ν∗(N )) = 3d + g − 1 = d 2 − δ − 1 X/IP 2 X/IP 2 2 (4.42) h1(X, N ) = h1(C,ν∗(N )) = 0 X/IP 2 X/IP 2 Finally, since + 0 d 2 h (I∆(d)) ≥ − δ = 3d + g 2 256 4 Hilbert and Quot schemes

and H 0(I∆(d))/(F) ⊂ H 0(N ), comparing with (4.42) we see that ∆ imposes X/IP 2 independent conditions to curves of degree d and that

0 0 ( ) = (I∆( ))/( ) H NX/IP 2 H d F or, equivalently, the restriction map

0 0 (I∆( )) → ( ) H d H NX/IP 2

is surjective. Note that N is a noninvertible subsheaf of N 2 . X/IP 2 X/IP (iii) Another interesting case is obtained by taking an irreducible curve X ⊂ IP 2 of degree d having δ nodes p1,...,pδ and κ ordinary cusps q1,...,qκ as its only singularities. This case is important because branch curves of generic projection on IP 2 of projective nonsingular surfaces are curves of this type. −1 Let ν : C → X be the normalization map. Letting q¯ j = ν (q j ), j = 1,...,κ, we have in this case, according to the above description

∗ ν (N ) = O (d)(−p − p −···− p − p − 3q¯ −···−3q¯κ ) X/P2 C 1 1 δ δ 1 ∗ = ωC ⊗ ν OX (3)(−¯q1 −···−q ¯κ )

∗ As before, one shows that ν∗[ν (N )]=N and therefore X/P2 X/P2 + 0 0 ∗ d 2 h (X, N ) = h (C,ωC ⊗ ν OX (3)(−¯q −···−q ¯κ )) ≥ − δ − 2κ − 1 X/IP 2 1 2 (4.43) and in general we may have strict inequality and h1(X, N ) = 0 because the X/IP 2 ∗ invertible sheaf ωC ⊗ ν OX (3)(−¯q1 −···−q ¯κ ) can be special. But if κ<3d then it is certainly nonspecial and therefore in such a case we have

+ h0(X, N ) = d 2 − δ − 2κ − 1 = 3d + g − 1 − κ X/IP 2 2

h1(X, N ) = 0 X/IP 2

4.7.2 The Severi varieties

Given an integer d > 0, consider the complete linear system |O(d)| of plane curves of degree d. It is a flat family of closed subschemes of IP 2 parametrized by the 0 2 projective space d = IP [H (IP , O(d))]: 2 H ⊂ IP × d |O(d)|: ↓ d 4.7 Plane curves 257

The linear system |O(d)| has a universal property with respect to families of plane curves of degree d because the pair ( d , |O(d)|) represents the Hilbert functor: ◦ Λd : (algschemes) → (sets) given by: flat families C ⊂ IP 2 × S of plane Λ (S) = d curves of degree d parametrized by S (see § 4.3). In this subsection we want to consider the problem of constructing a universal family of reduced curves in IP 2 having degree d, an assigned number δ of nodes and κ of ordinary cusps and no other singularity. If such a universal family Vδ,κ exists it is parametrized by a scheme which we denote by d . These schemes have been studied classically: the foundations of their theory are given in [171] and they are therefore called Severi schemes or Severi varieties. Vδ,κ If the Severi scheme d exists then, by the universal property, there is a functo- rially defined morphism Vδ,κ → d d (4.44) We start from the definition of the functor we want to represent.

Definition 4.7.2. Let d,δ,κ as above. Then δ,κ : ( )◦ → ( ) Vd algschemes sets is defined as follows. For each algebraic scheme S flat families C ⊂ IP 2 × S of plane curves of deg. d formally δ,κ( ) = ∈ Vd S locally trivial at each k-rational s S whose geometric fibres are curves with δ nodes and κ cusps as singularities

δ,κ (see § 2.5 for the definition of formal local triviality). Obviously, Vd is a subfunctor Λ δ,κ of d . The main result about Vd is the following: ,δ,κ δ,κ Theorem 4.7.3. For each d as above, the functor Vd is represented by an Vδ,κ algebraic scheme d which is a (possibly empty) locally closed subscheme of d . κ = Vδ Vδ,0 In the case 0 we write d instead of d . The first published proof of this result is in [186]. We will not reproduce it in full generality here, but we will only consider the case κ = 0, i.e. the case of nodal curves. This assumption allows a technically simpler argument without changing the structure of the original proof. We need some lemmas.

2 2 Lemma 4.7.4. Let p ∈ IP , O the local ring of IP at p, B0 = O/( f0) the local ring of a plane curve having a node at p. Assume that for some A in ob(A) we have → 1 a deformation A BofB0 over A such that TB/A is A-flat. Then B is trivial and 1 =∼ TB/A A. 258 4 Hilbert and Quot schemes

Proof. By induction on dimk(A). The case A = k is trivial because T 1 = O/( f , f , f ) = O/( f , f ) =∼ k B0 0 0X 0Y 0X 0Y (see 3.1.4). In the general case consider a small extension

0 → ( ) → A → A → 0 and the induced deformation A → B .WehaveB = (A ⊗k O)/( f ) for some f 1 = /( , ) which reduces to f0 modulo m A, and TB/A B fX fY . Therefore

B = (A ⊗k O)/( f ) where f is obtained from f by reducing the coefficients to A , and 1 = /( , ) = /( , ) ⊗ = 1 ⊗ TB /A B fX fY B fX fY A A TB/A A A

1 It follows that TB /A is A -flat and, by induction, we have

B = (A ⊗k O)/( f0) and 1 = ⊗ [O/( , , )]= TB /A A k f0 f0X f0Y A Thus f = f0 + g where g ∈ k.Wehave: 1 = ( ⊗ O)/( + , , ) = /( ) TB/A A k f0 g f0X f0Y A g where the last equality follows from the fact that f0 ∈ ( f0X , f0Y ). Since A/( g) is A-flat if and only if g = 0 it follows that B = (A ⊗k O)/( f0) = A ⊗k (O/( f0)) is 1 =  the trivial deformation and TB/A A. Lemma 4.7.5. Let f : X → S be a flat morphism of algebraic schemes which factors as j X −→ Y ↓q S where j is a regular embedding of codimension 1 and q is smooth. Then for every morphism of algebraic schemes ϕ : S → S we have

1 ∼ ∗ 1 T = Φ T S ×S X/S X/S Φ : × → 1 where S S X X is the projection (i.e. TX/S commutes with base change). Proof. Since the question is local we may reduce to a diagram of k-algebras of the form: B = P/( f ) → B = PA /( f ) ↑↑ A → A where P is a smooth A-algebra, f ∈ P is a regular element, and f is the image of f in PA = P ⊗A A . Then we have 4.7 Plane curves 259 δ 2 ∼ ( f )/( f ) = B −→ ΩP/A ⊗P B f¯ → df ⊗ 1 and δ 2 ∼ ( f )/( f ) = B −→ ΩP /A ⊗P B f¯ → df ⊗ 1 But since

ΩP /A ⊗P B = (ΩP/A ⊗P B) ⊗B B we have δ = δ ⊗B B and 1 ⊗ = (δ∨) ⊗ = (δ ∨) = 1  TB/A B B coker B B coker TB /A Lemma 4.7.6. Let S be an algebraic scheme and C ⊂ IP 2 × S a flat family of plane curves of degree d. Let s ∈ Sbeak-rational point such that the fibre C(s) is a curve 1 ∈ C( ) having at most nodes as singularities. Then TC/S is S-flat at a point p s if and only if the family is formally locally trivial at s around p.

Proof. If p is a nonsingular point of C(s) then C → S is smooth at p, hence 1 T (C/S, OC)p = 0 and the assertion is obvious. α Let’s assume that p is a node of C(s).LetA = OS,s, Aα := A/m , Sα = Spec(Aα) and Cα → Sα the induced infinitesimal deformation of C1 = C(s); = O = O let B C,p and Bα Cα,p. By Lemma 4.7.5 we have:

1 1 ∼ 1 1 = = (C/ , O ) ⊗ α = ⊗ α TCα/Sα,p TBα/Aα T S C p OC,p B TB/A B B (4.45)

1(C/ , O ) = 1 1 α Assume that T S C p TB/A is A-flat. Then TBα/Aα is A -flat by (4.45) and therefore Bα is the trivial deformation of B0 = B/m A B = OC(s),p, by Lemma 4.7.4. Since this is true for every α we conclude that the family C → S is formally locally trivial at p. ∼ Conversely, assume that C → S is formally locally trivial at p. Then Bα = ⊗ α α 1 α B0 k A for all and TBα/Aα is A -flat by Lemma 4.7.4. But by (4.45) we have

1 = 1 ⊗ = 1 ⊗ TBα/Aα TB/A B Bα TB/A A Aα

1 = 1(C/ , O ) and from the local criterion of flatness we deduce that TB/A T S C p is A-flat.  2 Proof of Theorem 4.7.3. Consider the universal family H ⊂ IP × d of plane curves 1 of degree d and let {Wi } be the flattening stratification of T .LetW = Wi be H/ d a stratum containing a k-rational point s ∈ d parametrizing a reduced curve H(s) having δ nodes and no other singularity, and let H ⊂ IP 2 × W be the induced family of degree d curves. By Lemma 4.7.5 we have

1 ∼ 1 T = T ⊗ OH H /W H/ d 260 4 Hilbert and Quot schemes

1 H ⊂ 2 × and therefore, by construction, TH /W is flat over W. Moreover, since IP W 1 O is a regular embedding of codimension 1, TH /W is of the form V for some closed subscheme V ⊂ IP 2 × W. By applying Lemma 4.7.5 again we deduce

O = 1 ⊗ =∼ 1 V (s) TH /W k TH(s) which is a reduced scheme of length δ supported at Sing(H(s)). This implies that V → W is etale at the δ points of V (s). Therefore there is an open neighbourhood U of s ∈ W such that V (U) → U is etale of degree δ.Ifu ∈ U is a k-rational H( ) 1 =∼ ( ) H( ) δ point then u is a curve such that TH(u) V u , hence u has singular points 1 ∼ p1,...,pδ and no other singularity, such that TO = k. From Proposition 3.1.5 it p j follows that H(u) is a δ-nodal curve. Therefore by applying Lemma 4.7.6 we see H ( ) → δ,0( ) that the family U U is an element of Vd U . Putting together all these open sets we obtain a locally closed subset Ui ⊂ Wi H ⊂ 2 × δ,0( ) such that the induced family i IP Ui defines an element of Vd Ui .Nowlet Vδ = d Ui i

H ⊂ 2 × Vδ C ⊂ 2 × δ,0( ) and let IP d be the induced family. If IP S is an element of Vd S for an algebraic scheme S then by the universal property of |O(d)| we obtain a unique morphism S → d inducing the given family by pullback. By Lemma 4.7.6 and the Vδ defining property of the flattening stratification this morphism factors through d . (Vδ, H) δ,0  Thus d represents the functor Vd . We now consider the local properties of the Severi varieties.

Proposition 4.7.7. Let C ⊂ IP 2 be a reduced curve having degree d, δ nodes and κ [ ]∈V = Vδ,κ ordinary cusps and no other singularity. Let C d be the point parame- trizing C. Then there is a natural identification:

0 [ ]V = ( , ) T C H C NC/IP 2

and H 1(C, N ) is an obstruction space for OV,[ ]. C/IP 2 C

0 Proof. T[C]V is the subspace of T[C] d = H (C, OC (d)) corresponding to lo- cally trivial first-order deformations, and these are the elements of H 0(C, N ) C/IP 2 by the very definition of N . From the proof of Proposition 3.2.6 it is ob- C/IP 2 vious that obstructions to deforming locally trivial deformations lie in the space H 1(C, N ).  C/IP 2 Vδ,κ [ ] According to the classical terminology, we call d regular at a point C if δ,κ H 1(C, N ) = 0; otherwise V is called superabundant at [C]. An irreducible C/IP 2 d Vδ,κ component W of d is called regular (resp. superabundant) if it is regular (resp. 4.7 Plane curves 261

Vδ,κ superabundant) on a nonempty open subset. d is called regular if all its compo- nents are regular; otherwise it is called superabundant. From 4.7.7 and from Example Vδ,κ 4.7.1(iii) it follows that if a component W of d is regular then it is generically non- singular of dimension 3d + g − 1 − κ, where g is the geometric genus of C, i.e. of pure codimension δ + 2κ in d . κ< Vδ,κ Vδ Corollary 4.7.8. If 3d then d is regular at every point. In particular, d is regular at every point, thus it is nonsingular of pure dimension d + 2 3d + g − 1 = − 1 − δ 2 if it is nonempty.

Proof. It follows from Proposition 4.7.7 and from Example 4.7.1(iii).  Vδ,κ =∅ Remark 4.7.9. Note that the corollary does not claim that d . The nonempti- ness of the Severi varieties will be discussed in the next subsection. Corollary 4.7.8 follows also from Proposition 3.4.16 recalling that, by Lemma 3.4.15, we have

i ∼ i ( , ) = ( ˜ , ϕ), = , H C NC/IP 2 H C N i 0 1 where ϕ : C˜ → C is the normalization map. The description of the deformations of C given by Proposition 3.4.16 can be considered as the “parametric” counterpart of the “cartesian” point of view of Corollary 4.7.8. Vδ,κ The Severi varieties d may have a complicated structure. If there are too many δ,κ cusps then in general a [C]∈V satisfies H 1(C, N ) = (0) (see Example d C/IP 2 Vδ,κ [ ] 4.7.10 below) and in fact, d can be singular at such a C . To decide whether this effectively happens has been a long-standing classical problem (see [189], ch. VIII, where this topic is discussed). The first example of a singular point of a Severi variety was given by Wahl [186]: it is a plane irreducible curve of degree 104 with 3636 nodes and 900 cusps. For other examples see [122], [80] and [183]. κ> Vδ,κ Example 4.7.10. If 3d then d can be superabundant. The following classical example is due to B. Segre (see [159] and [189], p. 220). Consider plane curves of the following type:

3 2 C :[f2m(x, y)] +[f3m(x, y)] = 0 where f2m(x, y) and f3m(x, y) are general polynomials of the indicated degrees, and m > 2. Then d = deg(C) = 6m, δ = 0 and κ = 6m2 because the only singularities of C are the points of intersection of the curves f2m = 0 and f3m = 0 and they are easily seen to be cusps. C is irreducible of geometric genus 6m − 1 g = − 6m2 2 262 4 Hilbert and Quot schemes

The dimension of the family of curves C is 2m + 2 3m + 2 1 R := + − 1 = (13m + 2)(m + 1) 2 2 2 which is larger than 6m(6m + 3) r = − 2κ = 6m2 + 9m 2 − = m−1 V0,6m2 [ ] In fact, R r 2 . Therefore 6m is superabundant at all points C . Let’s compute h1(C, N ). By the analysis of Example 4.7.1(iii) we know C/IP 2 that h1(C, N ) equals the index of speciality ι of the linear system cut on the C/IP 2 normalization C˜ of C by the curves of degree 6m passing through the cusps and tangent there to the cuspidal tangents. It is an easy computation (see [189] p. 220 for details) that ι = R − r The conclusion is that each [C] is a nonsingular point of a superbundant component V0,6m2 of 6m of dimension R. For a modern treatment of this example see [180].

4.7.3 Nonemptiness of Severi varieties

Even though we have proved precise results about the structure of the Severi vari- eties, it is not clear that nodal curves of given degree and number of nodes exist at all: the task of writing down explicitly the equation of such a curve is too concrete and precise to be within the reach of known techniques. Nevertheless Severi himself outlined a method to prove the existence of nodal curves. His approach is based on δ V the notion of “analytic branch” and consists in analysing the local structure of d Vδ+1 along d . While his proof makes perfectly good sense over the field of complex numbers, it is not straightforward to translate it into an algebraic proof. In this sub- Vδ section we will show that the Severi varieties d of nodal curves are nonempty in the expected range of δ using a different method which is entirely algebraic and el- ementary. It is based on the theory of multiple point schemes, which we will now recall. Consider a finite unramified morphism f : X → Y of algebraic schemes and let Nk( f ) ⊂ Y be the k-th multiple point scheme of f (for the definition see Example 4.2.9, page 200). Define

−1 Mk( f ) = f (Nk( f )) ⊂ X

(scheme theoretic inverse image). Note that, in particular, N1( f ) is supported on the image of f and M1( f ) = X.Let X ×Y X = ∆ X2 4.7 Plane curves 263

(where ∆ is the diagonal and the union is disjoint because f is unramified) and let f1 : X2 → X be the morphism induced by the first projection. Then f1 is called the first iteration morphism of f .

Lemma 4.7.11. Let f : X → Y be a finite unramified morphism of algebraic schemes. Then:

(i) The first iteration morphism f1 is finite and unramified. (ii) Nk−1( f1) = Mk( f ) for all k ≥ 2.

Proof. (i) f1 is finite and unramified because both properties are invariant under base change and composition with a closed embedding. (ii) By the base change property of the Fitting ideals we have

−1 Mk( f ) = f (Nk( f )) = Nk(π1) = Nk−1( f1) where π1 : X ×Y X → X is the first projection.  Lemma 4.7.12. Let f : X → Y be a finite unramified morphism of purely dimen- sional algebraic schemes, such that dim(X) = dim(Y ) − 1 and Y is nonsingular. ˜ Then every irreducible component X2 of X2 satisfies ˜ dim(X2) ≥ dim(Y ) − 2

Proof. We have −1 X ×Y X = ( f × f ) (∆Y ) where ∆Y ⊂ Y ×Y is the diagonal. Since Y is nonsingular, ∆Y is regularly embedded of codimension dim(Y ) in Y × Y . It follows that every component of X ×Y X has codimension ≤ dim(Y ) in X × X (Lemma D.1.2). We deduce that every component ˜ X2 of X2 satisfies ˜ dim(X2) ≥ 2dim(X) − dim(Y ) = dim(Y ) − 2 

If the morphism f satisfies some further assumptions then one can describe its behaviour quite precisely.

Definition 4.7.13. Let f : X → Y be a finite unramified morphism of algebraic schemes. Then f is called self-transverse of codimension 1 if X and Y are nonsin- gular and purely dimensional, dim(X) = dim(Y ) − 1, and for any closed point −1 y ∈ Y and for any r distinct points x1,...,xr ∈ f (y) the tangent spaces ,..., Tx1 X Txr X, viewed as subspaces of TyY , are in general position (i.e. their in- tersection has codimension r).

Lemma 4.7.14. If f : X → Y is a self-transverse codimension 1 morphism then f1 : X2 → X is a self-transverse codimension 1 morphism. 264 4 Hilbert and Quot schemes

Proof. Let (x1, x2) ∈ X2 and let y = f1(x1, x2) = f (x1) = f (x2). Then = × T(x1,x2) X2 Tx1 X Ty Y Tx2 X ( ) = ( )− Since f is self-transverse we have dim T(x1,x2) X2 dim Y 2. On the other hand, dim(X2) ≥ dim(Y ) − 2, by Lemma 4.7.12. It follows that X2 is nonsingular of pure dimension dim(Y ) − 2. ∈ ( , ) ∈ −1( ) For any x X and x x f1 x the differential

df1(x,x ) : T(x,x ) X2 → Tx X ( , ), . . . , ( , ) ∈ −1( ) is an injection of codimension 1. For any x x2 x xs f1 x we have ( ) ··· ( ) = ··· df1(x,x2) T(x,x2) X2 df1(x,xs ) T(x,xs ) X2 Tx X Tx2 X Txs X viewed as subspaces of TyY . The self-transversality of f1 now follows from the analogous property of f .  With this terminology we can state the following useful result.

Proposition 4.7.15. Assume that f : X → Y is a self-transverse codimension 1 morphism and that Nr ( f ) =∅for some r ≥ 2. Then Nr ( f ) has pure codimension r in Y and Ns( f ) =∅for all 1 ≤ s ≤ r −1. In particular, Ns( f ) has pure codimension s for all 1 ≤ s ≤ r − 1.

Proof. By induction on r.Ifr = 2 then M2( f ) = N1( f1), by Lemma 4.7.11(ii), and has pure dimension equal to dim(X2) = dim(Y ) − 2 by Lemma 4.7.14: therefore N2( f ) is of pure codimension 2. Moreover, N1( f ) = f (X) =∅and has pure dimension dim(X) = dim(Y ) − 1. Now assume r ≥ 3. Nr ( f ) has pure codimension r in Y if and only if Mr ( f ) has pure codimension r − 1inX. By Lemma 4.7.11(ii) we have

Mr ( f ) = Nr−1( f1) and by the inductive hypothesis Nr−1( f1) has pure codimension r − 1 because f1 is self-transverse of codimension 1. Again by the inductive hypothesis

Ms( f ) = Ns−1( f1) =∅ for all 1 ≤ s − 1 ≤ r − 2. This implies that Ns( f ) =∅for all 2 ≤ s ≤ r − 1. The case s = 1 is a consequence of the first part of the proof.  Next we will see how Proposition 4.7.15 can be applied to the study of the Severi varieties of nodal curves. ∗∗∗∗∗∗

2 Consider the universal family H ⊂ IP × d of plane curves of degree d ≥ 2 1 2 and the cotangent sheaf T . Since H ⊂ IP × d is a regular embedding H/ d 1 of codimension 1, T is a quotient of O 2× . Therefore we can identify H/ d IP d 4.7 Plane curves 265

1 T = OZ where Z ⊂ H is a closed subscheme; evidently, the set of closed H/ d points of Z is

2 {(p, s) : p is a singular point of H(s)}⊂IP × d Z can be described more precisely as follows. We let n n n ( , , ) = , , 0 1 2 = F X0 X1 X2 An0 n1 n2 X0 X1 X2 0 n0+n1+n2=d

2 be the equation of the universal curve H inside IP × d . Then Z is defined by the three equations: ∂ F ∂ F ∂ F = 0, = 0, = 0 ∂ X0 ∂ X1 ∂ X2 Denote by π π 2 1 2 IP ←− Z −→ d the projections. Lemma 4.7.16. (i) Z is irreducible, nonsingular, rational of codimension 3 in IP2 × d . (ii) π2 maps Z birationally onto its image W ⊂ d , which is an irreducible rational divisor parametrizing all singular curves of degree d.

Proof. π1 is surjective with fibres linear systems of dimension dim( d ) − 3 because ∈ 2 π −1( ) (− ) for each p IP the fibre 1 p is the linear system d 2p of all curves of degree d which are singular at p. This proves (i). Moreover, by an easy application of Bertini’s theorem one gets that a general element of d (−2p) is a curve having a node at p as its only singular point. This takes care of (ii).  Z contains an open subset Z1 whose set of closed points is {(p, s) : H(s) has only nodes as singularities and p is a node of H(s)}

From the proof of Lemma 4.7.16 it follows immediately that Z1 ⊂ Z is a nonempty (dense) open subset. Let W1 := π2(Z1)

W1 is a dense open subset of W. It parametrizes all singular curves of degree d having only nodes as singularities; a general point of W1 parametrizes a curve of degree d having one node and no other singularities. Consider the closed subset Bd := π2(Z\Z1) ⊂ d , which parametrizes all singular non-nodal curves of degree d, and the morphism

π : Z1 → d \Bd obtained by restricting π2.

Proposition 4.7.17. π is birational onto its image W1, finite, unramified and self- transverse of codimension 1. 266 4 Hilbert and Quot schemes

−1 Proof. If s ∈ W1 = π(Z1) is a closed point then π (s) is the scheme of nodes of H(s) which is finite and reduced. Therefore π is unramified. Moreover, π is the restriction over an open subset of d of a projective morphism. Therefore π is finite. The birationality onto its image follows from that of π2 (Lemma 4.7.16). In order to prove that π is self-transverse of codimension 1, let X ⊂ IP 2 be a curve of degree d having δ nodes {p1,...,pδ} and no other singularity. Then (pi , [X]) ∈ Z1, i = 1,...,δ. The local analysis of Example 4.7.1(i) and the fact that π is unramified show that we have [ π ]= 0( , (− )) Im d (pi ,[X]) H X NX/IP 2 pi where 1 N 2 (−p ) = ker[N 2 → T ] X/IP i X/IP X,pi Therefore δ 0 δ [ π( ,[ ])]= ( , ) = [ ]V Im d pi X H X NX/IP 2 T X d i=1 Vδ δ Since T[X] d has codimension by Corollary 4.7.8, this equality means that the [ π ] = 0( , ) subspaces Im d (pi ,[X]) are hyperplanes of T[X] d H X NX/IP 2 in general position.  We can now prove the main result of this subsection.

Theorem 4.7.18 (Severi [173]). Let X ⊂ IP 2 be a curve of degree d having δ nodes [ ]∈Vδ and no other singularity, i.e. such that X d . Then δ− ∅ = Vδ ⊂ V 1 ⊂···⊂V1 ⊂ V0 = d d d d d

Vs Vs−1 = ,...,δ and d has pure codimension 1 in d , for every s 1 . Proof. From the previous analysis it follows that for each s Vs = (π)\ (π) d Ns Ns+1 Vδ =∅ The existence of X implies that d . The theorem is now an immediate conse- quence of Propositions 4.7.17 and 4.7.15. 

Corollary 4.7.19. (i) For every d ≥ 2 and d 0 ≤ δ ≤ 2 Vδ the Severi variety d is nonempty. (ii) For every d ≥ 2 and d − 1 0 ≤ δ ≤ 2 Vδ the Severi variety d contains irreducible curves. 4.7 Plane curves 267

Proof. (i) A curve X consisting of d distinct lines no three of which pass through δ V o δ = d Vδ the same point defines a point of d with o 2 and the non-emptiness of d for δ ≤ δo follows from the theorem. 2 d (ii) A general projection in IP of a rational normal curve Γd ⊂ IP is an irre- d−1 ducible curve Y of degree d having 2 nodes. The theorem implies that Y is in Vδ δ ≤ d−1 the closure of d for every 2 . The following lemma guarantees that for each δ Vδ  such we can find irreducible curves in d . Lemma 4.7.20. Let C ⊂ IP 2 × S be a flat family of plane curves of degree d ≥ 2, with S an algebraic scheme, and let o ∈ Sbeak-rational point. Assume that the fibre C(o) is reduced and irreducible. Then C(s) is reduced and irreducible for all s ∈ S in an open neighbourhood of o.

i Proof. Denote by f : C → S the projection. Since H (C(s), OC(s)(m)) = 0 for all m ≥ d − 2, i ≥ 1, and for all s ∈ S, by Theorem 4.2.5 0 ( ) : O ( ) ⊗ ( ) → 0(C( ), O ( )) tm s f∗ C m s k s H s C(s) m is bijective and f∗OC(m)s is free for all m ≥ d − 2 and s ∈ S. The multiplication map

0 0 0 µs : H (OC(s)(d − 1)) ⊗ H (OC(s)(d − 1)) → H (OC(s)(2d − 2)) is surjective for all s ∈ S. Then the multiplication map

µ : f∗OC(d − 1) ⊗ f∗OC(d − 1) → f∗OC(2d − 2) is surjective and ker(µ) is locally free. Let R ⊂ f∗OC(d − 1) ⊗ f∗OC(d − 1) be the cone of reducible tensors. For 0 each s ∈ S it restricts to the cone R(s) of reducible tensors in H (OC(s)(d − 1)) ⊗ 0(O ( − )) 0 ( ) = − , − H C(s) d 1 , by the bijectivity of the maps tm s for m d 1 2d 2. The condition that C(o) is reduced and irreducible is equivalent to R(o) ∩ ker(µo) ={0}. Therefore there is an open neighbourhood of U of o ∈ S such that R(s) ∩ ker(µs) = {0}, i.e. such that C(s) is reduced and irreducible, for all s ∈ U.  Vδ Remarks 4.7.21. (i) It is easy to see that d is reducible in general. For example, V3 V9 4 has two irreducible components and 6 has five irreducible components. An im- portant classical problem, known as the “Severi problem”, has been to decide about Vδ the irreducibility of the open set of d parametrizing irreducible nodal curves. This problem has been solved affirmatively in [82] and, independently, in [147] (see also [182]). A report of Harris’ proof is given in [121]. It is known that the open set of Vδ,κ κ> d parametrizing irreducible curves is reducible in general if 0. For examples see [160] (such examples are also reported in [189]). (ii) Theorem 4.7.18 and some generalizations of it have been reconsidered in [179], but the proof given there is based on infinitesimal considerations which seem to need a further insight. For an interesting discussion see [60]. The analogous prob- Vδ,κ lem for the varieties d is still open, i.e. we don’t know a characterization of the 268 4 Hilbert and Quot schemes

,δ,κ Vδ,κ =∅ values of d for which d . For partial results see [95]; for a classical dis- cussion see [159]. The results on multiple point schemes used here are special cases of a general theory for which we refer the reader to [99], [101] and to the references quoted there. (iii) The proof of Theorem 4.7.3 can be easily modified to prove the existence of universal families of curves with nodes and cusps (generalized Severi varieties) on a Y projective nonsingular surface Y . In such a proof one replaces d by Hilb and uses the existence and the universal property of HilbY . Vδ,κ Such generalized Severi varieties behave in a way relatively similar to the d ’s as long as Y has Kodaira dimension ≤ 0 (see [179], [112]). On surfaces of general type the situation changes radically. On such a surface Y the generalized Severi va- rieties can be superabundant even when κ = 0 and it is not known in which range of δ they are not empty. A systematic study of them has started relatively recently. We refer the reader to [34], [69], [32], [56], [57] for details. A Flatness

The algebraic notion of flatness, introduced for the first time in [164], is the basic technical tool for the study of families of algebraic varieties and schemes. In this appendix we will overview the main algebraic results needed. For the properties of flat morphisms between schemes we refer to [84]. See also § 4.2. A module M over a ring A is A-flat (or flat over A,orsimplyflat) if the func- tor N → M ⊗A N from the category of A-modules into itself is exact. Since this functor is always right exact, the flatness means that it takes monomorphisms into monomorphisms. An A-algebra B is flat over A if B is flat as an A-module. The A-module M is said to be faithfully flat if for every sequence of A-modules N → N → N the sequence

M ⊗A N → M ⊗A N → M ⊗A N is exact if and only if the original sequence is exact. Obviously, if M is faithfully flat then it is flat. In a similar way we give the notion of faithfully flat A-algebra. It is straightforward to check that if A → B is a local homomorphism of local rings, then a B-module of finite type is faithfully A-flat if and only if it is A-flat and nonzero. Recall that the flatness of an A-module M is equivalent to any of the following conditions: A( , ) = ( ) > (1) Tori M N 0 for all i 0 and for every A-module N. A( , ) = ( ) (2) Tor1 M N 0 for every A-module N. A( , ) = ( ) (3) Tor1 M N 0 for every finitely generated A-module N. A( , / ) = ( ) ⊂ (4) Tor1 M A I 0 for every ideal I A. (5) I ⊗A M → M is injective for every ideal I ⊂ A. (6) I ⊗A M → IM is an isomorphism for every ideal I ⊂ A.

Example A.1. Let k be a ring, u,v indeterminates and f : k[u, uv]→k[u,v] the inclusion. Then k[u, uv] u k[u, uv] = k[u] −→ k[u]= (uv) (uv) 270 A Flatness is injective. Tensoring by ⊗k[u,uv]k[u,v] we obtain:

[ ,v] [ ,v] k u −→u k u (uv) (uv) which is not injective. Therefore f is not flat.

We list without proof a few basic properties of flat modules:

Proposition A.2. (I) M is A-flat if and only if Mp is A p-flat for every prime ideal p. (II) Every projective module is flat. (III) Assume M is finitely generated. Then M is flat if and only if it is projective; if A is local then M is flat if and only if it is free. (IV) If S ⊂ A is a multiplicative subset then AS is A-flat. (V) A direct sum M =⊕i∈I Mi is flat if and only if all Mi ’s are flat. (VI) Let

0 → M → M → M → 0 be an exact sequence of A-modules with M flat. Then M is flat if and only if M is flat. (VII) Base change: if M is A-flat and f : A → B is a ring homomorphism, then M ⊗A B is B-flat. (VIII) Transitivity: if B is a flat A-algebra and N is a flat B-module, then N is A-flat. (IX) If A is a noetherian ring and I is an ideal, the I -adic completion Aˆ is a flat A-algebra. If I is contained in the Jacobson radical of A then Aˆ is a faithfully flat A-algebra. (X) If B is an A-algebra and if there exists a B-module M which is faithfully flat, then the morphism Spec(B) → Spec(A) is surjective. (XI) If X1,...,Xr are indeterminates, then A[X1,...,Xr ] and A[[X1,...,Xr ]] are A-flat.

The following result is frequently used:

Proposition A.3. If A is an artinian local ring with residue field k the following are equivalent for an A-module M: (i) M is free (ii) M is flat A( , ) = ( ) (iii) Tor1 M k 0 Proof. (i) ⇒ (ii) ⇒ (iii) are clear. (iii) ⇒ (ii).LetN be a finitely generated A-module and let

N = N0 ⊃···⊃ Nn = (0) be a composition series for N such that ∼ Ni /Ni+1 = k A Flatness 271 for i = 0,...,n − 1. Using the Tor exact sequences from hypothesis (iii) we deduce that Tor1(M, N) = (0) and the flatness of M follows from (3). Let’s now prove (ii) ⇒ (i).Let{e j } j∈J be a system of elements of M which induces a basis of M ⊗A k over k. The system {e j } defines a homomorphism J J f : A → M which induces an isomorphism k → M ⊗A k. From the follow- ing lemma we find that f is an isomorphism, and therefore M is free. 

Lemma A.4. Let R be a ring, I an ideal and f : F → G a homomorphism of R-modules with G flat. Assume that one of the following conditions is satisfied: (a) I is nilpotent. (b) R is noetherian, I is contained in the Jacobson radical of R and F and G are finitely generated. If the induced homomorphism F/IF → G/I G is an isomorphism, then f is an isomorphism.

Proof. Let K = coker( f ). Tensoring the exact sequence

F → G → K → 0 with R/I we get K/IK = 0: from Nakayama’s lemma (which holds in either of hypotheses (a) and (b)) it follows that K = 0, and therefore F is surjective. Letting H = ker( f ) we deduce an exact sequence

0 → H/IH → F/IF → G/IG → 0 using the flatness of G. By Nakayama again we deduce H = 0 and the conclusion follows.  The following is a basic criterion of flatness.

Theorem A.5 (Local criterion of flatness). Suppose that ϕ : A → B is a local ho- momorphism of local noetherian rings, and let k = A/m A be the residue field of A. If M is a finitely generated B-module, then the following conditions are equivalent: (i) M is A-flat. A( , ) = (ii) Tor1 M k 0. ⊗ ( / n ) / n ≥ (iii) M A A m A is flat over A m A for every integer n 1. ⊗ ( / n ) / n ≥ (iv) M A A m A is free over A m A for every integer n 1. Proof. (i) ⇒ (ii) is obvious. (ii) ⇒ (i) see [48], Th. 6.8, p. 167. (i) ⇒ (iii) is obvious. (iii) ⇒ (i) It suffices to show that for every inclusion N → N of A-modules of

finite type we have an inclusion M ⊗A N → M ⊗A N. For this purpose it suffices to show that the kernel of this last map is contained in := [ ⊗ → ⊗ ( / ∩ n )] Kn ker M A N M A N N m A N 272 A Flatness = ( ) for all n, because n Kn 0 . We have a commutative diagram with exact rows: → → ⊗ → ⊗ ( / ∩ n ) → 0 Kn M A N M A N N m A N 0 ↓↓ ⊗ → ⊗ ( / n ) → M A N M A N m A N 0 The last vertical arrow coincides with the map obtained from the injection / ∩ n → / n N N m A N N m A N / n / n ⊗ ( / n ) after tensoring over A m A with the A m A-flat module M A A m A , and therefore it is injective. The conclusion follows from the above diagram. ( ) ⇔ ( v) / n  iii i follows from Proposition A.3 because A m A is artinian. For a more general version of the local criterion we refer to [3], exp. IV, Theor´ eme` 5.6. Note that A.3 is a special case of A.5.

Corollary A.6. Suppose that ϕ : A → B is a local homomorphism of local noetherian rings, let k = A/m A be the residue field of A, M, N two finitely generated B-modules, and suppose that N is A-flat. Let u : M → Nbea B-homomorphism. Then the following are equivalent: (i) u is injective and coker(u) is A-flat. (ii) u ⊗ 1 : M ⊗ k → N ⊗ k is injective.

Proof. (i) ⇒ (ii). Let G = coker(u). Tensoring by k the exact sequence

u 0 → M −→ N → G → 0 by k we obtain the exact sequence:

A( , ) → ⊗ −→u⊗1 ⊗ → ⊗ → Tor1 G k M A k N A k G A k 0 A( , ) = ⊗ Since G is A-flat we have Tor1 G k 0, and it follows that u 1 is injective. (ii) ⇒ (i). Factor u ⊗ 1as

α β M ⊗A k −→ Im(u) ⊗A k −→ N ⊗A k

Then α is an isomorphism and β is injective. Tensoring by k the exact sequence

0 → Im(u) → N → G → 0(A.1) we obtain the exact sequence:

β A( , ) → A( , ) → ( ) ⊗ −→ ⊗ → ⊗ → Tor1 N k Tor1 G k Im u A k N A k G A k 0 A( , ) = β Since N is A-flat we have Tor1 N k 0; from the injectivity of we deduce A( , ) = Tor1 G k 0 and from A.5 it follows that G is A-flat. Applying (VI) to the exact sequence (A.1) we deduce that Im(u) is A-flat as well. Consider the exact sequence: A Flatness 273

0 → ker(u) → M → Im(u) → 0 and tensor by k. We obtain the exact sequence:

α 0 → ker(u) ⊗A k → M ⊗A k −→ Im(u) ⊗A k → 0

Since α is an isomorphism we deduce that ker(u)⊗A k = 0, and therefore ker(u) = 0 by Nakayama’s lemma.  A related result is the following:

Lemma A.7. Let B be a local ring with residue field K , and let d : G → Fbea homomorphism of finitely generated B-modules, with F free. Then d is split injective if and only if d ⊗B K : G ⊗B K → F ⊗B K is injective. In such a case G is also free.

Proof. d is split injective if and only if coker(d) is free and d is injective. If this last condition is satisfied then clearly d ⊗B K is injective. Conversely, assume that d ⊗B K is injective, and factor d as

G → Im(d) → F

We see that G ⊗B K → Im(d) ⊗B K is bijective Im(d) ⊗B K → F ⊗B K is injective From the exact sequence

0 → Im(d) → F → coker(d) → 0 we get

0 → Tor1(coker(d), K ) → Im(d) ⊗B K → F ⊗B K so Tor1(coker(d), K ) = (0) and this implies that coker(d) is free. From the above exact sequence we deduce that Im(d) is free as well, so that

0 → ker(d) → G → Im(d) → 0 ∼ is split exact. Recalling that G ⊗B K = Im(d) ⊗B K we deduce that ker(d) ⊗B K = (0), hence ker(d) = (0) by Nakayama.  For the reader’s convenience we include the proof of the following well-known lemma:

Lemma A.8. Let (B, m) be a noetherian local , with residue field K and quotient field L. If M is a finitely generated B-module and if

dimK (M ⊗B K ) = dimL (M ⊗B L) = r then M is free of rank r. 274 A Flatness

Proof. Let m1,...,mr ∈ M be such that their images in M ⊗B K = M/mM form a basis. Then they define a homomorphism ϕ : Br → M and we have an exact sequence: ϕ 0 → N → Br −→ M → Q → 0 where N and Q are kernel and cokernel of ϕ. Since tensoring with K we get

ϕ¯ K r −→ M/mM → Q/mQ → 0 and ϕ¯ is surjective, we get Q/mQ = (0) and from Nakayama’s lemma it follows that Q = (0): hence ϕ is surjective. Now we tensor the above exact sequence with L, which is flat over B (by (IV)), and we obtain the exact sequence:

ϕ˜ r 0 → N ⊗B L → L −→ M ⊗B L → 0

∼ r Since M ⊗B L = L and ϕ˜ is surjective, it follows that N ⊗B L = ker(ϕ)˜ = (0). Therefore N is a torsion module. But N ⊂ Br and therefore N = (0).  We have the following useful criterion:

Lemma A.9. Let A → A be a small extension in A, and let g : A → Rbea homomorphism of k-algebras. Let R0 = R ⊗A k. Then g is flat if and only if ∼ ker(R → R ⊗A A ) = R0 and the homomorphism g : A → R ⊗A A induced by g is flat. ∼ Proof. Assume that g is flat. Then since R ⊗A ( ) = R ⊗A k = R0 and A( , ) = Tor1 R A 0, from the exact sequence → A( , ) → ⊗ ( ) → → ⊗ → 0 Tor1 R A R A R R A A 0(A.2) we deduce that the first condition is satisfied. The flatness of g is obvious. Assume conversely that the conditions of the statement are satisfied. Then the A( , ) = = sequence (A.2) implies that Tor1 R A 0. If A k the conclusion follows from A.3. If not, from the exact sequence

0 → m A → A → k → 0 one gets the exact sequence:

∂ A( , ) → A( , ) −→ ⊗ → → ⊗ → Tor1 R A Tor1 R k R A m A R R A k 0 

0 R ⊗A m A R ⊗A k ∂ = A( , ) = From the flatness of R over A we deduce that 0, hence Tor1 R k 0, and we conclude by A.3.  ∗∗∗∗∗∗ A Flatness 275

Flatness in terms of generators and relations Let P be a noetherian k-algebra, J ⊂ P an ideal. Let A be in ob(A), ∼ PA = P ⊗k A, and J ⊂ PA an ideal such that (PA/J) ⊗A k = P/J. We want to find the conditions J has to satisfy so that PA/J is A-flat. We have the following:

Theorem A.10. Let n N %0 : P → P → P → P/J → 0 be a presentation of P/J as a P-module. Then the following conditions are equiva- lent for an ideal J ⊂ PA: ∼ (i) PA/J is A-flat and (PA/J) ⊗A k = P/J. (ii) There is an exact sequence % : n → N → → / → PA PA PA PA J 0

such that %0 = % ⊗A k (= %/m A%). (iii) There is a complex ϕ % : n −→ N → → / → PA PA PA PA J 0 N % = % ⊗ which is exact except possibly at PA , such that 0 A k. Proof. (ii) ⇒ (i).Wehave: A( / , ) = (% ⊗ ) = (% ) = ( ) Tor1 PA J k H1 k H1 0 0

From A.3 it follows that PA/J is A-flat. Moreover, (ii) implies that ∼ (PA/J) ⊗A k = P/J. ( ) ⇒ ( ) : N → i ii . Choose a PA-homomorphism p PA J which makes the following diagram commute: : N → p PA J ↓↓ N p0 : P → J where p0 is the surjective homomorphism defined by the presentation %0.Fromthe / A( / , ) = ( ) flatness of PA J it follows that Tor1 PA J k 0 ; hence the exact sequence → A( / , ) → ⊗ → ⊗ → ( / ) ⊗ → 0 Tor1 PA J k J k PA k PA J A k 0  PP/J implies that J ⊗ k = J.Itfollowsthatp ⊗A k = p0 and therefore

coker(p) ⊗A k = coker(p0) = (0) so that coker(p) = (0) by Nakayama’s lemma. Hence p is surjective. 276 A Flatness

Now consider the exact sequence → ( ) → N → → 0 ker p PA J 0 and the associated Tor sequence: A( , ) → ( )/ ( ) → N → → Tor1 J k ker p m Aker p P J 0(A.3)

From the flatness of PA/J and from the exact sequence

0 → J → PA → PA/J → 0 A( , ) = A( / , ) = ( ) we have Tor1 J k Tor2 PA J k 0 . Therefore from (A.3) we see that ∼ ker(p)/m Aker(p) = ker(p0) : n → ( ) Arguing as before we can find a surjective homomorphism q PA ker p which makes the following diagram commutative:

n −→q ( ) PA ker p ↓↓ n P → ker(p0)

(ii) ⇒ (iii) is obvious. ( ) ⇒ ( ) % N iii i If is not exact at PA then we can add finitely many generators N → of the kernel of PA PA to obtain an exact sequence

ϕ % : n −→ N → → / → PA PA PA PA J 0

Then % ⊗A k has the form:

ϕ ⊗k Pn −→ P N → P → P/J → 0

Since

Im(ϕ ⊗ k) ⊂ Im(ϕ ⊗ k) ⊂ ker[P N → P] N we see that Im(ϕ ⊗ k) = ker[P → P] and therefore % ⊗A k is exact. Now (i) follows from A.3. 

Corollary A.11. Assume that J = ( f1,..., fN ) ⊂ P and that

J = (F1,...,FN ) ⊂ PA with f j = Fj (mod m A PA), j = 1,...,N. Then every relation among f1,..., fN lifts to a relation among F1,...,FN if and only if PA/J is A-flat and ∼ (PA/J) ⊗A k = P/J. A Flatness 277

Proof. The condition that the Fj ’s reduce to the f j ’s modulo m A PA implies that the exact sequence N −→F → / → PA PA PA J 0 reduces to f P N −→ P → P/J → 0(A.4) when tensored by ⊗Ak. Complete (A.4) to a presentation %0 of P/J. The condition that every relation among f1,..., fN lifts to a relation among F1,...,FN is a re- statement of condition (iii) of A.10. Therefore the conclusion follows from Theorem A.10. 

Example A.12. Let A be in ob(A). Suppose that f1,..., fN ∈ P form a regular sequence, and let F1,...,FN ∈ PA be any liftings of f1,..., fN , i.e. such that f j = Fj (mod mPA), j = 1,...,N. Then J = (F1,...,FN ) ⊂ PA defines a flat family of deformations of X = Spec(P/J), where J = ( f1,..., fN ). In fact, every relation among f1,..., fN is a linear combination of the trivial ones rij = (0,..., f j ,...,− fi ,...,0) 1 ≤ i < j ≤ N and these can be lifted to the corresponding trivial relations

Rij = (0,...,Fj ,...,−Fi ,...,0) among F1,...,FN . Applying Corollary A.11 it is easy to show that F1,...,FN form a regular sequence.

NOTES 1. In the proof of Theorem A.10 the condition that A is artinian has only been used in the proof of (i) ⇒ (ii) in order to apply Nakayama’s lemma. In particular, the implications ∗ (ii) ⇒ (i), (iii) ⇒ (i) and (ii) ⇒ (iii) hold for any A ∈ ob(A ). Using the local criterion of flatness it is easy to verify that the implication (i) ⇒ (ii) (and therefore the equivalence of the three conditions) holds as well if A is in ob(Aˆ). B Differentials

Let A → B be a ring homomorphism. As usual, we will denote by ΩB/A the module of differentials of B over A, and by dB/A : B → ΩB/A the canonical A-derivation. Recall that 2 ΩB/A := I/I µ where I = ker(B ⊗A B −→ B) is the natural map, and for each b ∈ B

dB/A(b) = b ⊗ 1 − 1 ⊗ b is called the differential of b. We have a natural isomorphism of B-modules ∼ DerA(B, M) = HomB(ΩB/A, M)

Note that the exact sequence

µ 2 0 → ΩB/A → (B ⊗A B)/I −→ B → 0(B.1) where µ is induced by µ,isanA-extension of B. The ring

2 PB/A := (B ⊗A B)/I is called the algebra of principal parts of B over A.TheA-extension (B.1) is trivial because we have splittings:

λ1,λ2 : B → PB/A defined by λ1(b) = b ⊗ 1, λ2(b) = 1 ⊗ b; note that dB/A = λ1 − λ2. We will consider PB/A as a B-algebra via λ1. The following are some fundamental properties of the modules of differentials: Proposition B.1. (i) If B ↑ A −→ A 280 B Differentials

are ring homomorphisms, then:

Ω ⊗ =∼ Ω B/A A A B⊗A A /A (ii) If A → B is a ring homomorphism and ∆ ⊂ B is a multiplicative system, then: Ω ∼ ∆−1Ω ∆−1 B/A = B/A (iii) Let K → L be a finitely generated extension of fields. Then

dimL (ΩL/K ) ≥ trdeg(L/K )

and equality holds if and only if L is separably generated over K . In particular, ΩL/K = (0) if and only if K ⊂ L is a finite algebraic separable extension.

Proof. See [48].  We now introduce two standard exact sequences.

Theorem B.2 (Relative cotangent sequence). Given ring homomorphisms

f g A −→ B −→ C there is an exact sequence of C-modules:

α β ΩB/A ⊗B C −→ ΩC/A −→ ΩC/B → 0(B.2) where the maps are given by:

α(dB/A(b) ⊗ c) = cdC/A(g(b)); β(dC/A(r)) = dC/B(r) b ∈ B, c ∈ C

Proof. See [48], prop. 16.2.  When B → C is surjective we have ΩC/B = (0) and the next theorem describes ker(α).

Theorem B.3 (Conormal sequence). Let

f g A −→ B −→ C be ring homomorphisms with g surjective, and let J = ker(g), so that C = B/J. Then: (i) We have an exact sequence

2 δ α J/J −→ ΩB/A ⊗B C −→ ΩC/A → 0(B.3)

where δ is the C-linear map defined by δ(x¯) = dB/A(x) ⊗ 1. B Differentials 281

(ii) There is an isomorphism Ω ⊗ ∼ Ω ⊗ (B/J 2)/A (B/J 2) C = B/A B C In other words the conormal sequence (B.3) depends only on the first infinitesi- mal neighbourhood of Spec(C) in Spec(B). (iii) The map δ is a split injection if and only if there is a map of A-algebras C → B/J 2 splitting the projection B/J 2 → C. Proof. (i) see e.g. [48], prop. 16.3. (ii) Comparing the exact sequence (B.3) with the analogous sequence associated to A → B/J 2 → C we get a commutative diagram: 2 J/J → ΩB/A ⊗B C → ΩC/A → 0 ↓ / 2 → Ω ⊗ → Ω → J J (B/J 2)/A (B/J 2) C C/A 0 and the vertical arrow, which is induced by B → B/J 2, must be an isomorphism. (iii) By (ii) we may assume that J 2 = 0, i.e. that 0 → J → B → C → 0 is an A-extension. Assume that δ : J → ΩB/A ⊗B C is a split injection, and let σ : ΩB/A ⊗B C → J be a splitting. Then the composition

d¯ σ B −→ ΩB/A ⊗B C −→ J is an A-derivation. It follows that 1 − σd¯ : B → B is an A-homomorphism such that (1 − σd¯)(J) = 0 and therefore it induces an A-homomorphism C → B which splits g. Conversely, assume that g : B → C has a section τ : C → B. Then we have a derivation D : B → J ⊕ ΩC/A given by D(b) = (b − (τg)(b), dC/A(g(b))). One easily checks that D induces an ∼ isomorphism ΩB/A ⊗B C = J ⊕ ΩC/A, thus proving the assertion.  As an application we have the following: Proposition B.4. Let K be a field and (B, m) a local K -algebra with residue field B/m = K . Then the map

2 δ : m/m → ΩB/K ⊗B K in the exact sequence (B.3) relative to K → B → K is injective if and only if K ⊂ K is a separable field extension. In particular, if B/m = K then

2 δ : m/m → ΩB/K ⊗B K is an isomorphism. Therefore

dim(B) ≤ dimK (ΩB/K ⊗B K ) 282 B Differentials

Proof. See [48], cor. 16.13. The last assertion follows from the conormal sequence relative to K → B → K .  The following theorem describes the module of differentials for regular local rings.

Theorem B.5. Assume that K is a field and B is a local noetherian K -algebra with residue field B/m = K.IfΩB/K is a free B-module of rank equal to dim(B) then B is a regular local ring. If K is perfect (e.g. algebraically closed) and B is e.f.t. over K then the converse is also true.

Proof. Assume first that ΩB/K is free of rank equal to dim(B). Then

2 dimK (m/m ) = dim(B) by B.4, so B is a regular local ring. Assume conversely that K is perfect and that B is a regular local ring, e.f.t. over K . Then we have

2 dimK (ΩB/K ⊗B K ) = dimK (m/m ) = dim(B)

Let L be the quotient field of B. Then, by B.1(3), we have

ΩB/K ⊗B L = ΩL/K and dimL (ΩL/K ) = trdeg(L/K ) = dim(B) because L is separably algebraic over K , since K is perfect. Therefore we have

dimK (ΩB/K ⊗B K ) = dim(B) = dimL ΩB/K ⊗B L

Since B is e.f.t. over K , ΩB/K is a finitely generated B-module, and from Lemma A.8 it follows that it is free of rank equal to dim(B).  In particular, we have the following:

Corollary B.6. Let k be an algebraically closed field, and let B be an integral k- algebra of finite type. Then B is a regular ring if and only if ΩB/k is a projective B-module of rank equal to dim(B).

Proof. Both conditions are satisfied if and only if they are satisfied after localizing at the maximal ideals of B. For every maximal ideal m ⊂ B the local ring Bm is a k-algebra e.f.t. with residue field k.ByB.5,Bm is a regular local ring if and only if Ω = (Ω ) ( )  Bm /k B/k m is free of rank equal to dim B . The conclusion follows.

Proposition B.7. If the ring homomorphism A → B is e.f.t. then ΩB/A is a B- module of finite type. −1 If, in particular, B = S A[X1,...,Xn] for some multiplicative system S, then ΩB/A is a free B-module of rank n with basis {dB/A(X1),...,dB/A(Xn)}. B Differentials 283

Proof. The last assertion is elementary (see [48]). To prove the first, let −1 B = (S P)/J, where P = A[X1,...,Xn] and S ⊂ P is a multiplicative Ω Ω ⊗ system. Then B/A is a quotient of S−1 P/A S−1 P B, by the conormal sequence. 

Remark B.8. If A and B are only assumed to be noetherian then ΩB/A is not neces- sarily a B-module of finite type even if A is a field. An example is given by ΩQ[[X]]/Q (see [1] ch. 0IV, n. 20.7.16). −1 Examples B.9. (i) Assume that B = S A[X1,...,Xn] for some multiplicative system S. Then DerA(B, B) = HomB(ΩB/A, B) is a free module of rank n with basis ∂ ∂ ,..., ∂ X1 ∂ Xn which is the dual of the basis

{dB/A(X1),...,dB/A(Xn)}

∂ of Ω / , and where : B → B is the partial A-derivation with respect to X . B A ∂ X j j Let Y1,...,Yn ∈ B be such that the jacobian determinant ∂Y det i ∂ X j is a unit in B. Then {dB/A(Y1),...,dB/A(Yn)} is another basis of ΩB/A and we have:

∂ X j ∂ X j dB/A(X j ) = dB/A(Y1) +···+ dB/A(Yn) ∂Y1 ∂Yn Dually: ∂ ∂Y ∂ ∂Y ∂ = 1 +···+ n (B.4) ∂ X j ∂ X j ∂Y1 ∂ X j ∂Yn The proof of these statements is straightforward. (ii) Let k be a field and let B = k[X, Y ]/(XY), where X, Y are indeterminates. ∼ Then, since Ωk[X,Y ]/K ⊗ B = BdX ⊕ BdY, using the conormal sequence we deduce that ∼ BdX ⊕ BdY Ω / = B k (YdX ⊕ XdY) It follows that the element YdX =−XdY is killed by the maximal ideal (X, Y ) and therefore it generates a torsion submodule ∼ T := (YdX) = k ⊂ ΩB/k 284 B Differentials

The quotient is

Ω / BdX ⊕ BdY B k = =∼ k[X]dX ⊕ k[Y ]dY =∼ (X, Y ) ⊂ B T (YdX, XdY) where the last isomorphism is given by

f (X)dX ⊕ g(Y )dY → f (X)X + g(Y )Y

Therefore we have an exact sequence:

0 → T → ΩB/k → B → k → 0

(iii) Let k be a field and let B = k[t, X, Y ]/( f ) where t, X, Y are indeterminates and f = XY + t. Then arguing as before we see that

∼ BdX ⊕ BdY Ω / [ ] = B k t (YdX ⊕ XdY)

The element YdX =−XdY is not killed by any b ∈ B; therefore ΩB/k[t] is torsion free of rank one. The homomorphism

ΩB/k[t] → B sending f (t, X)dX⊕ g(t, Y )dY → f (t, X)X + g(t, Y )Y is bijective onto the max- imal ideal (t, X, Y ) so that we have an exact sequence:

0 → ΩB/k[t] → B → k → 0

(iv) Let k be a field and let k[ ]:=k[t]/(t2), where we have denoted by the class of t mod (t2). Then the conormal sequence of k → k[t]→k[ ] is

2 4 δ (t )/(t ) −→ Ωk[t]/k ⊗k[t] k[ ]→Ωk[ ]/k → 0 and the middle term is isomorphic to k[ ].Themapδ acts as

t¯2 → 2 t¯3 → 0 In particular, we see that δ is not injective. Therefore kd if char(k) = 2; Ω [ ]/ = k k k[ ]d if char(k) = 2 and d : k[ ]→Ωk[ ]/k acts as d(α + β) = βd . (v) An obvious generalization of the above computation shows that if A = k[t]/(tn), n ≥ 2 and char(k) = 0 or char(k)>n then B Differentials 285

¯n−1 ΩA/k = A/(t )

∈ (A∗) ∨ := / 2 := ( / 2 )∨ (vi) If B ob then tB m B m B and tB m B m B are the (Zariski) / 2 =∼ Ω ⊗ cotangent space, respectively tangent space of B.Wehavem B m B B/k B k by Prop. B.4, and therefore ( , ) = (Ω , ) = (Ω ⊗ , ) = ( / 2 )∨ Derk B k HomB B/k k Homk B/k B k k m B m B Moreover, there is a natural identification

Derk(B, k) = Homk−alg(B, k[ ]) which we leave to the reader to verify. If µ : Λ → B is a homomorphism in A∗, the induced homomorphism µ∨ : / 2 → / 2 d mΛ mΛ m B m B is the codifferential of µ, while its transpose

dµ : tB → tΛ is the differential of µ. We define the relative cotangent space of B over Λ to be ∨ := ( µ∨) = /( 2 + ) tB/Λ coker d m B m B mΛ B and the relative tangent space of B over Λ as its dual: = ( µ) = /( 2 + ) ∨ tB/Λ ker d m B m B mΛ B From the exact sequence

ΩΛ/k ⊗Λ B → ΩB/k → ΩB/Λ → 0 ∨ = Ω ⊗ tensored by k we deduce an identification tB/Λ B/Λ B k and therefore

tB/Λ = HomB(ΩB/Λ, k) = DerΛ(B, k) = HomΛ−alg(B, k[ ]) where the Λ-algebra structure on k[ ] is defined by the composition Λ → k → k[ ] (the last equality is straightforward to verify). The following lemma describes a situation where the conormal sequence is exact. Lemma B.10. Assume char(k) = 0. Let

e : 0 → (t) → R → R → 0 be a small extension in A. Then the conormal sequence

δ η : 0 → (t) −→ ΩR /k ⊗R R → ΩR/k → 0 is exact also on the left. 286 B Differentials

Proof. Assume first that e is trivial, so that R = R⊕˜ k. Then the codifferential

ΩR /k ⊗R k → ΩR/k ⊗R k  / 2 / 2 m R m R m R m R has a nontrivial kernel (Example 1.1.2) so that a fortiori

ΩR /k ⊗R R → ΩR/k has a nontrivial kernel. Assume now that e is not trivial. Then, letting m = dim(tR ) = dim(tR), we can write R = P/J , R = P/J 2 where P = k[X1,...,Xm], a polynomial algebra, and J , J ⊂ (X) ⊂ P ideals such that J ⊂ J and J/J =∼ (t).LetT ∈ J be such that

t = T + J

Since e is small we have (X)J ⊂ J and therefore T ∈/ (X)J. Claim: We can choose T so that ∂T ∈/ J for some i ∂ Xi ∂ ∂ If T ∈ J then X T ∈ J so that we can replace T by ∂ Xi i ∂ Xi ∂T T1 := T − Xi ∂ Xi If ∂2T ∂T −X = 1 ∈/ J i ∂ 2 ∂ Xi Xi we are done, otherwise we replace T1 by

X ∂T ∂T X 2 ∂2T T := T − i 1 = T − X + i 2 1 ∂ i ∂ ∂ 2 2 Xi Xi 2 Xi and we apply the same argument. After ν steps of this process we obtain

ν s ∂ ν− X ∂s := − Xi T 1 = + (− )s i T Tν Tν−1 T 1 s ν ∂ Xi s! ∂ X s≥1 i

∂ ∂ Since Tν = 0forν  0 we see that either Tν ∈/ J for some ν and we replace T ∂ Xi ∂ Xi by the first Tν with this property, or we can replace T by a Tν which is constant with respect to Xi . Repeating this process for every index i we will end up by replacing B Differentials 287

T by a T¯ having the required property or otherwise constant with respect to every variable, which is clearly a contradiction. The claim is proved. From the claim we deduce that ∂T dT = dX ∈/ JΩ / (B.5) ∂ X i P k i i where d = dP/k : P → ΩP/k is the universal derivation. But we have:

ΩR /k = ΩP/k/ J ΩP/k + (dg )g ∈J so that

ΩR /k ⊗R R = ΩP/k/ JΩP/k + (dg )g ∈J But since clearly dt = dg for all g ∈ J , (B.5) implies that

dt = 0 ∈ ΩR /k ⊗R R  ∗∗∗∗∗∗ : → Ω1 If f X Y is a morphism of schemes, we denote by X/Y the sheaf of relative differentials,ortherelative cotangent sheaf,onX. It satisfies

Ω1 = Ω X/Y,x OX,x /OY, f (x) for all x ∈ X.If f : Spec(B) → Spec(A) is a morphism of affine schemes then Ω1 = (Ω )∼ Spec(B)/Spec(A) B/A We denote by := (Ω1 , O ) TX/Y Hom X/Y X the sheaf of relative derivations,ortherelative tangent sheaf of f . Ω1 Ω1 We will write X and TX instead of X/Spec(k) and TX/Spec(k) respectively; they are the cotangent sheaf and the tangent sheaf of X, respectively (cotangent and tan- gent bundles if locally free). If X is algebraic and x ∈ X is closed then, by B.4:

m , Ω1 ⊗ k(x) = X x X,x 2 m X,x is the cotangent space of X at x, and

∨ m X,x ∼ T X := T , ⊗ k(x) = = (O , , k) x X x 2 Derk X x m X,x is the Zariski tangent space of X at x. 288 B Differentials

Let S be a scheme and g X −→ Y a morphism of S-schemes. The induced homomorphism of sheaves on X: ∗Ω1 → Ω1 g Y/S X/S is called the relative codifferential of g. The dual homomorphism: → ( ∗Ω1 , O ) TX/S Hom g Y/S X

= ( ) ∗Ω1 → Ω1 is the relative differential of g. When S Spec k we have g Y X , which is the codifferential of g, while its dual : → ( ∗Ω1 , O ) dg TX Hom g Y X Ω1 is the differential of g. Note that if Y/S is locally free then ( ∗Ω1 , O ) = ∗ (Ω1 , O ) = ∗ Hom g Y/S X g Hom Y/S Y g TY/S but in general the first and the second sheaves are different. The relative cotangent sequence is ∗Ω1 → Ω1 → Ω1 → g Y/S X/S X/Y 0(B.6) Conditions for the injectivity of the first map in this sequence are given in Theorem C.15, page 302 and Theorem D.2.8, page 310. If X ⊂ Y is an embedding of schemes and I = IX/Y ⊂ OY is the ideal sheaf of 2 X in Y , then I/I is a sheaf of OX -modules in a natural way, called the conormal sheaf of X in Y . Its dual := (I/I2, O ) = (I, O ) NX/Y HomOX X HomOY X

2 is called the normal sheaf of X in Y . NX/Y (resp. I/I ) is called the normal bundle (resp. the conormal bundle)ofX in Y if it is locally free. Given a closed embedding of S-schemes i : X ⊂ Y , we have an exact sequence of sheaves on X: I/I2 → ∗Ω1 → Ω1 → i Y/S X/S 0(B.7) where I ⊂ OY is the ideal sheaf of X in Y . (B.7) is called the relative conormal sequence. When S = Spec(k) we obtain the conormal sequence I/I2 → ∗Ω1 → Ω1 → i Y X 0 Conditions for the injectivity of the first map in these sequences are given in Propo- sition D.1.4, page 306, and Theorem D.2.7, page 310.

Examples B.11. In the following examples we will describe the global vector fields on the given schemes by exhibiting their restrictions to an affine open set. All will be done by explicit computation. B Differentials 289

0( ) 1 = ∪ (i) H TIP 1 can be described explicitly as follows. Consider IP U0 U1 −1 where U0 = Spec(k[ξ]) and U1 = Spec(k[η]) with η = ξ on U0 ∩ U1.Wehave ∂ ∂ξ ∂ ∂ ∂ = =−1 =−ξ 2 ∂η ∂η ∂ξ η2 ∂ξ ∂ξ

∩ θ ∈ 0( ) on U0 U1.Let H TIP 1 ; then ∂ θ| = g(ξ) g(ξ) ∈ k[ξ] U0 ∂ξ and ∂ θ| = h(η) , h(η) ∈ k[η] U1 ∂η

On U0 ∩ U1 we have ∂ ∂ ∂ g(ξ) = h(η) =−h(ξ −1)ξ 2 ∂ξ ∂η ∂ξ

−1 2 2 and therefore g(ξ) =−h(ξ )ξ . It follows that g(ξ) = a0 + a1ξ + a2ξ and (η) =−( η2 + η + ) , , ∈ 0( ) ∼ 3 h a0 a1 a2 , with a0 a1 a2 k. In particular, H TIP 1 = k . i ( ) = ≥ ≥ θ ∈ 1( ) Moreover, H TIP 1 0ifi 1. For i 2 it is obvious. Let H TIP 1 ˇ θ ∈ Γ( ∩ , ) be represented by a Cech 1-cocycle defined by 01 U0 U1 TIP 1 . It can be written as n i θ01 = ai ξ i=−m   θ = −1 η−i θ =− n ξ i Letting 1 i=−m ai and 0 i=0 ai we obtain:

θ01 = θ1 − θ0 so (θ01) is a coboundary. 0( ) 1 × 1 = ∪ (ii) We want to describe H TA1×IP 1 .LetA IP V0 V1 where 1 V0 = A × U0 = Spec[z,ξ]) 1 V1 = A × U1 = Spec[z,η]) −1 −1 and η = ξ on V0 ∩ V1 = Spec(k[z,ξ,ξ ]).Wehave ∂ ∂ξ ∂ ∂ ∂ = =−1 =−ξ 2 ∂η ∂η ∂ξ η2 ∂ξ ∂ξ

∩ θ ∈ 0( ) on V0 V1.Let H TA1×IP 1 ; then ∂ ∂ θ| = g(z,ξ) + h(z,ξ) g(z,ξ),h(z,ξ)∈ k[z,ξ] V0 ∂z ∂ξ ∂ ∂ θ| = γ(z,η) + χ(z,η) γ(z,η),χ(z,η)∈ k[z,η] V1 ∂z ∂η 290 B Differentials

On V0 ∩ V1 we have: g(z,ξ)= γ(z,ξ−1) and therefore g(z,ξ)= g(z) is constant with respect to ξ. Moreover, ∂ ∂ ∂ h(z,ξ) = χ(z,η) =−χ(z,ξ−1)ξ 2 ∂ξ ∂η ∂ξ and therefore h(z,ξ)=−χ(z,ξ−1)ξ 2 It follows that h(z,ξ) = a(z) + b(z)ξ + c(z)ξ 2, with a(z), b(z), c(z) ∈ k[z].In θ ∈ 0( ) conclusion, every H TA1×IP 1 restricts to V0 as a vector field of the form ∂ ∂ 2 θ| = g(z) + (a(z) + b(z)ξ + c(z)ξ ) (B.8) V0 ∂z ∂ξ with g(z), a(z), b(z), c(z) ∈ k[z], and conversely, every such vector field is the restriction of a global section of TA1×IP 1 . As in example (i) we also deduce that i ( ) = ≥ H TA1×IP 1 0ifi 1. 0( ) In a similar way, one describes H T(A1\{0})×IP 1 by showing that the image of the restriction 0 0 H (T 1 1 ) → H (T 1 ) (A \{0})×IP (A \{0})×U0 consists of the vector fields of the form (B.8) with g(z), a(z), b(z), c(z) ∈ k[z, z−1]. (iii) We now consider, for a given integer m ≥ 0, the rational ruled surface = (O ( ) ⊕ O ) Fm IP IP 1 m IP 1

1 Let π : Fm → IP be the projection. Then Fm can be represented as

−1 −1 1 1 Fm = π (U) ∪ π (U ) = (U × IP ) ∪ (U × IP ) where U = Spec(k[z]), U = Spec(k[z ]) and z = z−1 on U ∩ U . We consider the affine open sets 1 V0 = Spec(k[z,ξ]) ⊂ U × IP = ( [ ,ξ ]) ⊂ × 1 V0 Spec k z U IP ∩ = ( [ , −1,ξ]) = ( [ , −1,ξ ]) where on V0 V0 Spec k z z Spec k z z we have:

z = z−1,ξ = zmξ

Therefore we have: ∂ =−2 ∂ + ξ ∂ ∂z z ∂z mz ∂ξ (B.9) ∂ −m ∂ ∂ξ = z ∂ξ 0 We will describe a typical element θ ∈ H (TF ) by describing its restriction to the m open sets V0 and V0. We have, by example (ii) above: B Differentials 291 ∂ ∂ 2 θ| = g(z) + (a(z) + b(z)ξ + c(z)ξ ) V0 ∂z ∂ξ with g(z), a(z), b(z), c(z) ∈ k[z] and similarly

∂ ∂ θ = ρ( ) + (α( ) + β( )ξ + γ( )ξ 2) |V z z z z 0 ∂z ∂ξ ρ( ), α( ), β( ), γ ( ) ∈ [ ] ∩ with z z z z k z . Imposing their equality on V0 V0 and using (B.9) we obtain the following conditions: g(z) =−ρ(z−1)z2 a(z) = α(z−1)z−m (B.10) b(z) = β(z−1) + ρ(z−1)mz c(z) = γ(z−1)zm We distinguish the cases m = 0 and m > 0. If m = 0 (B.10) give:

2 g(z) = g0 + g1z + g2z a(z) = a g , g , g , a, b, c ∈ k b(z) = b 0 1 2 c(z) = c In the case m > 0wehave: 2 g(z) = g0 + g1z + g2z a(z) = 0 g0, g1, g2, b, c0,...,cm ∈ k b(z) = b − mz(g1 + g2z) m c(z) = c0 + c1z +···+cm z 0( ) → 0( ) Since the restriction H TFm H TV0 is injective and we have described its image, we can conclude: 0( ) =∼ 6 H TF0 k 0( ) =∼ m+5 H TFm k ∼ 1 1 In particular, Fm and Fn are not isomorphic if m = n (note that F0 = IP × IP is ∼ 2 not isomorphic to F1 = Bl(1,0,0) IP ). Since, by the calculations of the previous example (ii) i ( ) = i ( ) = i ( ) = , ≥ h TU×IP 1 h TU ×IP 1 h T(U∩U )×IP 1 0 i 1 (B.11) we deduce that:

1( ) = 0( )/ 0( ) + 0( ) H TFm H T(U∩U )×IP 1 H TU×IP 1 H TU ×IP 1

≥ 1( ) An easy computation based on (B.10) shows that, for m 1, H TFm consists of 0( ) + 0( ) the classes, modulo H TU×IP 1 H TU ×IP 1 , of the vector fields ∂ m−1 (b z +···+b − z ) 1 m 1 ∂ξ 292 B Differentials

In particular, 1( ) =∼ m−1 H TFm k (B.12) It also follows from (B.11) that 2( ) = ( ) H TFm 0 (B.13)

NOTES 1. Let X → Y be a morphism of algebraic schemes. Prove that there is an exact sequence → Ω1 → P1 → O → 0 X/Y X/Y X 0 (B.14) P1 which globalizes (B.1). X/Y is called the sheaf of principal parts of X over Y , denoted by P1 = ( ) X if Y Spec k . Let X = IP (V ) for a finite-dimensional k-vector space V . Then the exact sequence (B.14) is the dual of the Euler sequence; in particular,

P1 =∼ O (− ) ⊗ ∨ IP (V ) IP (V ) 1 V Therefore (B.14) is a generalization of the Euler sequence to any X → Y . 2. Consider IP = IP (V ) for a finite-dimensional k-vector space V and the incidence relation: ∨ I ={(x, H) : x ∈ H}⊂IP × IP (B.15) Consider the twisted and dualized Euler sequence: → Ω1 ( ) → O ⊗ ∨ → O ( ) → 0 IP (V ) 1 IP (V ) V IP (V ) 1 0 = (Ω1 ( )) × ∨ = (O ⊗ ∨) From its definition it follows that I IP IP (V ) 1 and IP IP IP IP (V ) V and the inclusion in (B.15) is induced by the first homomorphism in the above sequence. C Smoothness

The notion of “formal smoothness”, introduced in [1], Ch. IV § 17, is of crucial importance in deformation theory, and therefore plays a special role in this book. In this appendix we introduce this concept from scratch, and we show how it is related to the notion of “smooth morphism” as introduced in [3] Expose´ II, and [84]. We will not give a systematic treatment of the properties of smooth morphisms in algebraic geometry: the reader is referred to the above quoted references for them. For more details on the approach taken in this section the reader can also consult [13] and [94].

Definition C.1. A ring homomorphism f : R → B is called formally smooth, and B is called a formally smooth R-algebra, if for every exact sequence:

η 0 → I → A −→ A → 0(C.1) where A and A are local artinian R-algebras, each R-algebra homomorphism B → A has a lifting B → A; equivalently, if the map:

HomR−alg(B, A) → HomR−alg(B, A ) (C.2) is surjective. f is called smooth if it is formally smooth and essentially of finite type (shortly e.f.t.). If the map (C.2) is bijective (instead of only being surjective) for all exact sequences (C.1), then f is formally etale;fisetale if it is formally etale and e.f.t..

Recall that e.f.t. means that B is a localization of an R-algebra of finite type (see e.g. [127]). It is easy to prove by induction that it suffices to check the above conditions only for the exact sequences (C.1) such that I 2 = (0), i.e. for extensions of local artinian R-algebras.

Proposition C.2. (i) If B is a ring and ∆ ⊂ B is a multiplicative system, then B → ∆−1 B is formally etale. In particular, B is a formally etale B-algebra. (ii) The composition of formally smooth (resp. formally etale) homomorphisms is formally smooth (resp. formally etale). 294 C Smoothness

(iii) If f : R → B is formally smooth (resp. formally etale) and C is an R-algebra, then C → C ⊗R B is formally smooth (resp. formally etale). (iv) A finitely generated field extension K ⊂ L is smooth if and only if L is separable over K . f g (v) Let R −→ B −→ C be ring homomorphisms, and assume that f is formally etale. Then g f is formally smooth (resp. formally etale) if and only if g is for- mally smooth (resp. formally etale).

Proof. (i) Given an exact sequence (C.1) and a commutative diagram

B → ∆−1 B ↓ ϕ ↓ ϕ A −→ A p we must find ϕ˜ : ∆−1 B → A which makes it commutative. For every s ∈ ∆ choose −1 as ∈ A such that ϕ(s) = p(as). Since

−1 p(ϕ (s)as) = ϕ(s)ϕ(s) = 1A we have

ϕ (s)as = 1A + is is ∈ I for every s ∈ ∆. Therefore

ϕ (s)as(1A − is) = 1A

Hence ϕ (s) ∈ A is invertible. Now define ϕ(˜ r/s) = ϕ (r)ϕ (s)−1. Noting that ϕ˜ is uniquely determined by ϕ we get the assertion. (ii) and (iii) are straightforward. (iv) Assume first that K ⊂ L is separable. By (ii) it suffices to consider the cases L = K (X) and L = K [X]/( f (X)) where f is irreducible and f (x) = 0. The first case is left to the reader (see remark C.3(i)). In the second case consider an extension A¯ = A/I of local artinian K -algebras, where I ⊂ A is an ideal with I 2 = (0).Letϕ : K [X]/( f (X)) → A¯ be a homomor- phism, sending X¯ →¯α. Choose arbitrarily α ∈ A such that α¯ = α mod I .Itwill suffice to find e ∈ I such that f (α + e) = 0

We have f (α + e) = f (α) + f (α)e. Since f (α) is a unit mod I it is also a unit in A, and therefore we can take e =−f (α)/f (α). Assume conversely that K ⊂ L is smooth. Then L = F[X]/J where F is a purely transcendental extension of K and J is a principal ideal. We have an exact sequence of finite-dimensional L-vector spaces:

2 J/J → ΩF[X]/K ⊗ L → ΩL/K → 0 C Smoothness 295 where J/J 2 is one-dimensional. By the first part of the proof F is smooth over K and by B.3(ii) the left map is injective because, by the smoothness of L over K ,the surjection F[X]/J 2 → L splits. It follows that (Ω ) = (Ω ⊗ ) − = ( [ ]) − dim L/K dim F[X]/K L 1 trdegK F X 1 = ( ) = ( ) trdegK F trdegK L From B.1(iii) it follows that K ⊂ L is separable. (v) “if” follows immediately from (ii); “only if” is left to the reader. 

Remarks C.3. (i) Any polynomial algebra R[X1, X2 ...] is trivially a formally smooth R-algebra. From C.2(i) it follows that a localization of a polynomial R- algebra is also a formally smooth R-algebra. −1 More precisely, a localization P = S R[X1, X2,...] of a polynomial algebra over a ring R satisfies the following condition, stronger than formal smoothness: For every extension of R-algebras:

0 → I → A → A → 0

where A and A are R-algebras and I 2 = 0themap

HomR−alg(P, A) → HomR−alg(P, A )

is surjective. Every R-algebra B is a quotient of a formally smooth R-algebra, because it is a quotient of a polynomial R-algebra. From C.2(i) it follows that every e.f.t. R-algebra is a quotient of a smooth R-algebra. This is trivial for polynomial rings, and in the general case it can be proved adapting the proof of C.2(i) in an obvious way. ˆ (ii) if R is in ob(A) then every formal power series ring R[[X1, X2,...]] is a formally smooth R-algebra, because local artinian R-algebras are complete. More precisely, a formal power series ring R[[X1, X2,...]] satisfies the following condi- tion, stronger than formal smoothness over R: For every extension:

0 → I → A → A → 0 of complete local R-algebras the map

HomR−alg(P, A) → HomR−alg(P, A )

is surjective. The proof is straightforward and is left to the reader.

The following result characterizes an important class of formally smooth algebras. 296 C Smoothness

Theorem C.4. Let k be a field and let (B, m) be a noetherian local k-algebra with residue field K . Suppose that K is finitely generated and separable over k. Then the following are equivalent: (i) B is regular. ˆ ∼ (ii) B = K [[X1,...,Xd ]], where d = dim(B). (iii) B is a formally smooth k-algebra.

Proof. (i) ⇔ (ii) is standard (see [48], prop. 10.16 and exercise 19.1). (ii) ⇒ (iii). It follows directly from the definition that B is formally smooth over k if and only if Bˆ is. Since Bˆ is formally smooth over K (remark C.3(ii)), and since K is smooth over k by C.2(iv), the conclusion follows by transitivity. 2 (iii) ⇒ (i). Let {x1,...,xd } be a system of generators of m. Then, since B/m is complete and K is separable over k, B/m2 contains a coefficient field ([48], Theorem 7.8). Therefore there exists an isomorphism

2 ∼ 2 v1 : B/m = K [X1,...,Xd ]/M M = (X1,...,Xd ) v 2 1 2 Let v : B → B/m −→ K [X1,...,Xd ]/M . By the formal smoothness of B and by induction we can find a lifting of v:

n+1 vn : B → K [X1,...,Xd ]/M for every n ≥ 2. Consider the elements

n+1 vn(x1),...,vn(xd ) ∈ M/M

Their classes generate M/M2, hence they generate M/Mn+1, by Nakayama. Then we have:

n+1 n+1 K [X1,...,Xd ]/M = vn(B) + (M/M )  = v ( ) + v ( ) v ( ) + ( / n+1) = v ( ) + ( / n+1)2 =··· n B i n xi n B M M n B M M

n+1 n+1 = vn(B) + (M/M ) = vn(B)

n+1 hence vn is surjective. Since m ⊂ ker(vn) we have: d + n (B/mn+1) ≥ (K [X ,...,X ]/Mn+1) = 1 d d and this implies that dim(B) ≥ d. Since m is generated by d elements it follows that B is regular.  For the reader’s convenience we include the proof of the following well-known lemma:

Lemma C.5. (i) A surjective endomorphism f : A → A of a noetherian ring is an automorphism. C Smoothness 297

(ii) Let A be a complete noetherian local ring and ψ : A → A an endomorphism ψ : / 2 → / 2 ψ inducing an isomorphism 1 A m A A m A. Then is an automorphism. Proof. (i) We have an ascending chain of ideals

ker( f ) ⊆ ker( f 2) ⊆ ker( f 3) ⊆···

Since A is noetherian we have ker( f n) = ker( f n+1) = ker( f n+2) = ···for some n, and it suffices to prove that ker( f n) = (0). After replacing f by f n we may assume ker( f ) = ker( f 2).Leta ∈ ker( f ); by assumption there exists b ∈ A such that a = f (b). Then 0 = f (a) = f 2(b) and therefore b ∈ ker( f 2) = ker( f ), i.e. a = f (b) = 0. (ii) Let gr(A) = A/m⊕m/m2⊕···be the associated graded ring. Since gr(A) is generated by m/m2 over A/m the endomorphism gr(ψ) : gr(A) → gr(A) induced by ψ is surjective. It follows that ψ is also surjective. In fact, given a ∈ A the surjectivity of gr(ψ) implies that there are a1, a2, a3,...,b1, b2, b3,... ∈ A such i−1 i that ai ∈ m , bi ∈ m , and

a = f (a1) + b1, b1 = f (a2) + b2, b2 = f (a3) + b3,...

We obtain a convergent power series a¯ = a1 + a2 + a3 +···such that

n+1 a − ψ(a1 + a2 +···+an) = bn ∈ m

On the limit we therefore get a = ψ(a¯). The conclusion is now a consequence of (i). 

Proposition C.6. Let f : R → B be a local homomorphism of noetherian local rings containing a field k isomorphic to their residue fields. Then the following con- ditions are equivalent: (i) f is formally smooth. (ii) Bˆ is isomorphic to a formal power series ring over R.ˆ (iii) The homomorphism fˆ : Rˆ → Bˆ induced by f is formally smooth.

Proof. (i) ⇒ (ii). Let m ⊂ B and n ⊂ R be the maximal ideals. Choose elements ˆ ˆ 2 ˆ ˆ x1,...,xd ∈ B inducing a k-basis of B/(mˆ + f (nˆ)), and let F = R[[X1,...,Xd ]], where X1,...,Xd are indeterminates. Denote by M ⊂ F the maximal ideal. The homomorphism u : F → Bˆ Xi → xi induces an isomorphism

2 ˆ 2 ˆ u1 : F/(M +ˆnF) → B/(mˆ + f (nˆ))

By the formal smoothness of f the composition

− u 1 ˆ ˆ 2 ˆ 1 2 v1 : B → B → B/(mˆ + f (nˆ)) −→ F/(M +ˆnF) 298 C Smoothness can be lifted to an R-homomorphism

k vk : B → F/M ˆ for each k ≥ 2. Therefore the sequence {vk} defines an R-homomorphism

v : Bˆ → F such that vu : F → F and uv : Bˆ → Bˆ induce isomorphisms 2 2 ˆ 2 ˆ 2 (vu)1 : F/M → F/M and (uv)1 : B/mˆ → B/mˆ respectively. From Lemma C.5 it follows that u and v are isomorphisms. (ii) ⇒ (iii) is obvious. (iii) ⇒ (i) is left to the reader. 

Corollary C.7. Let f : R → B be a local homomorphism of noetherian local rings containing a field k isomorphic to their residue fields. Then the following conditions are equivalent: (i) f is formally etale. (ii) The homomorphism fˆ : Rˆ → Bˆ induced by f is an isomorphism.

Proof. Left to the reader.

Corollary C.8. Let R be in ob(A∗). The inclusion f : R → Rˆ is formally etale.

The proof is obvious. We now restrict our attention to smooth homomorphisms, i.e. we add the condition that the homomorphism is e.f.t.. In this case the module of differentials comes into play; moreover, the defining condition of Definition C.1 can be replaced by the more general condition (i) in the following statement.

Theorem C.9. Let f : R → B be an e.f.t. ring homomorphism. Then the following conditions are equivalent: (i) For every extension of R-algebras:

0 → I → A → A → 0(C.3)

the map

HomR−alg(B, A) → HomR−alg(B, A ) is surjective. −1 (ii) If B = P/J, where P = S R[X1,...,Xd ],S⊂ R[X1,...,Xd ] is a multi- plicative system and J ⊂ P is an ideal, the conormal sequence

2 δ 0 → J/J −→ ΩP/R ⊗P B → ΩB/R → 0

2 is split exact. In particular, J/J and ΩB/R are finitely generated projective B-modules. C Smoothness 299

(iii) B is a smooth R-algebra. (iv) (Jacobian criterion of smoothness) If P and J are as in (ii) the map

δ⊗ ( ) ( / 2) ⊗ ( ) −→B K p Ω ⊗ ( ) ( ) = / J J B K p P/R P K p where K p Bp m Bp

is injective for every prime ideal p ⊂ B.

Proof. (i) ⇒ (ii). The hypothesis implies that the extension:

0 → J/J 2 → P/J 2 → B → 0 splits. Therefore the conormal sequence is split exact by B.3(iii) and it follows that 2 J/J and ΩB/R are finitely generated projective because the module ΩP/R ⊗P B is free of finite rank. (ii) ⇒ (i). Consider an exact sequence (C.3) and a homomorphism of R-algebras f : B → A . By Remark C.3(i) there exists an R-homomorphism g : P → A making the following diagram commute: P → B ↓ g ↓ f A → A

Since g(J) ⊂ I , we see that g factors through P/J 2, so that we have a commutative diagram: P/J 2 → B ↓¯g ↓ f A → A The hypothesis implies, via B.3(iii), that there exists h : B → P/J 2 a splitting of P/J 2 → B. The composition f =¯gh : B → A gives a lifting of f . (i) ⇒ (iii) is obvious. (iii) ⇒ (iv). We may assume B and P local with residue field K . To prove that δ ⊗B K is injective, it suffices to show that for every K -vector space V the map induced by δ:

2 HomK (ΩP/R ⊗P K, V ) → HomK ((J/J ) ⊗B K, V )  2 DerR(P, V ) HomB(J/J , V ) is surjective. Consider a homomorphism g : J/J 2 → V , and the associated pushout diagram (see § 1.1 for the definition):

Λ : 0 → J/J 2 → P/J 2 → B → 0 ↓ g ↓ g∗(Λ) : 0 → V → Q → B → 0 300 C Smoothness

We can write m Q = V ⊕ m , where m ⊂ Q is an ideal, because V is annihilated by m Q. Therefore the previous diagram can be embedded in the following:

P ↓ Λ : 0 → J/J 2 → P/J 2 → B → 0 ↓ g ↓ g∗(Λ) : 0 → V → Q → B → 0 ↓↓¯v η : 0 → V → Q/m → K → 0 where η is an extension of local artinian R-algebras. From the smoothness of B we deduce the existence of v : B → Q/m lifting the projection v¯ : B → K . Denoting by r : P → B the natural map, and by w : P → P/J 2 → Q → Q/m the composition, consider the homomorphism:

d = w − vr : P → V

It is easy to show that this is an R-derivation, which induces g.

(iv) ⇒ (ii). From Nakayama’s lemma it follows that ker(δ) ⊗ Bp = (0) for all prime ideals p ⊂ B and therefore ker(δ) = (0). Moreover, since ΩP/R ⊗B Bp is free Bp (Ω ⊗ , ( )) = and finitely generated it follows that Tor1 B/R B Bp K p 0: it follows that ΩB/R ⊗B Bp is flat, and therefore, being finitely generated, it is free. Thus ΩB/R is projective, δ has a splitting and J/J 2 is also projective and finitely generated.  The following result follows easily from what we have seen so far.

Theorem C.10. Let B be an integral k-algebra of finite type and of dimension d. Then the following are equivalent:

(i) Bp is smooth over k for each prime ideal p ∈ Spec(B). (ii) B is a regular ring. (iii) ΩB/k is projective of rank d. (iv) B is smooth over k.

Proof. (ii) ⇔ (iii) is Corollary B.6. (i) ⇔ (ii) follows from Theorem C.4. (iv) ⇒ (i). for each p ∈ Spec(B), Bp is smooth over B by Proposition C.2(i); from Proposition C.2(ii) it follows that Bp is smooth over k. (i) ⇒ (iv). (i) implies that condition (iv) of Theorem C.9 is satisfied for all p ∈ Spec(B), so that B is smooth by Theorem C.9.  From now on we will freely replace the defining property for smooth homo- morphisms given in Definition C.1 by condition (i) of Theorem C.9. Here is a first example.

Proposition C.11. Let R be a ring, P an R-algebra and B = P/J for an ideal J ⊂ P. If B is a smooth R-algebra the conormal sequence C Smoothness 301

2 0 → J/J → ΩP/R ⊗P B → ΩB/R → 0 is split exact and ΩB/R is projective and finitely generated. If, moreover, P is a smooth R-algebra then J/J 2 is finitely generated and projective as well.

Proof. Since B is smooth the R-algebra extension

0 → J/J 2 → P/J 2 → B → 0 splits. Therefore the conormal sequence splits by Theorem B.3(iii) and ΩB/R is fi- nitely generated and projective by Theorem C.9. If P is smooth then ΩP/R is finitely generated and projective as well and so is J/J 2. 

Corollary C.12. Let P be an e.f.t. k-algebra and B = P/J for an ideal J ⊂ P. Assume that B is reduced. Then in the conormal sequence

2 δ J/J −→ ΩP/k ⊗P B → ΩB/k → 0(C.4) ker(δ) is a torsion B-module whose support is contained in the singular locus of Spec(B). If J/J 2 is torsion free then δ is injective.

Proof. Since B is reduced there is a dense open subset U ⊂ Spec(B) such that Bp is a regular local ring for all p ∈ U. From Theorem C.4 it follows that Bp is a smooth k-algebra for all such p and, by Propositions C.11 and B.1(ii), the conormal sequence (C.4) localized at p is split exact. It follows that ker(δ)p = (0) for all p ∈ U and the conclusion follows. The last assertion is an obvious consequence of the first part.  The next result explains the relation between smoothness and the relative cotan- gent sequence.

f g Theorem C.13. Let K −→ R −→ B be ring homomorphisms, with g smooth. Then the relative cotangent sequence:

α 0 → ΩR/K ⊗R B −→ ΩB/K → ΩB/R → 0 is split exact.

Proof. By Theorem B.2 it suffices to prove that α is a split injection; this is equivalent to showing that, for any B-module M, the induced map:

α∨ HomB(ΩB/K , M) −→ HomB(ΩR/K ⊗R B, M)  DerK (B, M) DerK (R, M)

D → D g 302 C Smoothness is split surjective. Let D : R → M be a K -derivation and consider the commutative diagram: 1 B −→B B ↑ g ↑ γ R −→ B⊕˜ M where γ(r) = (g(r), D(r)), r ∈ R. By the smoothness of g we can find a homomor- phism of R-algebras ψ : B → B⊕˜ M making the diagram 1 B −→B B ↑ g ψ ↑ R −→ B⊕˜ M γ commutative. The homomorphism ψ is necessarily of the form:

ψ(b) = (b, D (b)) and D : B → M is a K -derivation such that D = D g. This proves the surjectivity ∨ of α .NowtakeM = ΩR/K ⊗R B and D = dR/K ⊗ g : R → ΩR/K ⊗R B and let

α : ΩB/K → ΩR/K ⊗R B be the B-linear map corresponding to D : B → ΩR/K ⊗R B. Then α α = 1M and this proves that α is split injective. 

f g Corollary C.14. Let K −→ R −→ B be ring homomorphisms, with g etale. Then ∼ ΩR/K ⊗R B = ΩB/K and ΩB/R = (0) Proof. By the relative cotangent sequence the two assertions are equivalent. We will prove the first. Keeping the notations of the proof of C.13, the hypothesis that g is etale implies that the derivation D is unique and consequently α is an isomorphism.  ∗∗∗∗∗∗

A morphism ϕ : X → Y of algebraic schemes is smooth at a point x ∈ X if OX,x is a smooth OY,ϕ(x)-algebra; ϕ is smooth if it is smooth at every point. The definition of etale morphism is given similarly. This definition is equivalent to the definition of smooth (resp. etale) morphism as given in [3] and in [84]. The equivalence can be seen by means of the jacobian criterion of smoothness, proved in Theorem C.9, and using [3], Expose´ II, Corollaire 5.9. By translating into geometrical language the algebraic results proved above we deduce in particular the following. Theorem C.15. Let S be an algebraic scheme, and ϕ : X → Y a morphism of algebraic S-schemes. Then: C Smoothness 303

(i) If ϕ is smooth at x ∈ X then the relative cotangent sequence → ϕ∗Ω1 → Ω1 → Ω1 → 0 Y/S X/S X/Y 0

Ω1 is split exact at x and X/Y is locally free at x. The rank of the free module Ω1 ϕ X/Y,x is called the relative dimension of at x. (ii) ϕ is etale at x ∈ X if and only if it is smooth of relative dimension zero at x. Ω1 = ϕ In particular, X/Y,x 0 (i.e. is unramified at x) and therefore we have an isomorphism ϕ∗Ω1 =∼ Ω1 Y/S,x X/S,x

(iii) If X is smooth over S at x and ϕ is a closed embedding with ideal sheaf I ⊂ OY then the relative conormal sequence → I/I2 → ϕ∗Ω1 → Ω1 → O Y/S X/S 0

Ω1 ϕ( ) is exact at x and X/S is free at x; if, moreover, Y is also smooth over S at x then I/I2 is free at x as well.

The exactness of the relative conormal sequence in part (iii) holds under more general assumptions as well (see Theorem D.2.7). D Complete intersections

D.1 Regular embeddings

Definition D.1.1. An embedding of schemes j : X ⊂ Yisaregular embedding of codimension n at the point x ∈ Xifj(x) has an affine open neighbourhood Spec(R) in Y such that the ideal of j(X) ∩ Spec(R) in R can be generated by a regular sequence of length n. If this happens at every point of X we say that j is a regular embedding of codimension n.

An open embedding is a regular embedding of codimension 0. If X and Y are both nonsingular then X ⊂ Y is a regular embedding. The set of points of X where an embedding j : X ⊂ Y is regular is open. 2 If X ⊂ Y is a regular embedding of codimension n then I/I and NX/Y are both locally free of rank n ([48], Exercise 17.12, p. 440). It follows from standard facts in commutative algebra (see [48], Exercise 17.16, p. 441) that Ik/Ik+1 is locally free as well for every k ≥ 2. AringB is called a complete intersection if Spec(B) can be regularly embedded in Spec(R) where R is a regular ring. A scheme X is a local complete intersection (l.c.i.) if every local ring OX,x is a complete intersection ring. A nonsingular scheme X, i.e. a scheme all of whose local rings are regular, is an example of an l.c.i. scheme. If X ⊂ Y is a regular embedding and Y is an l.c.i. scheme, then X is an l.c.i. scheme.

Lemma D.1.2. Let f : X → Y be a morphism of schemes and let Z ⊂ Ybea regular embedding of codimension n. Then the induced embedding j : X ×Y Z ⊂ X has codimension ≤ n at every point and if equality holds at a point x ∈ X ×Y Z then j is regular at x.

−1 Proof. If IZ ⊂ OY is the ideal sheaf of Z in Y then the ideal sheaf f IZ of X ×Y Z ⊂ X is locally generated at a point x by the n images of the local generators of IZ, f (x). The conclusion follows easily from this fact.  306 D Complete intersections

If we have a flag of embeddings of schemes X ⊂ Y ⊂ Z and IY ⊂ IX ⊂ OZ are the ideal sheaves of X and Y , we have the exact sequence

0 → IY → IX → IX/Y → 0(D.1) I ⊂ O ⊗ O where X/Y Y is the ideal sheaf of X in Y . After tensoring by OZ X we obtain an exact sequence of coherent OX -modules:

I α I IX/Y Y ⊗ O −→ X → → 0(D.2) I2 X I2 I2 Y X X/Y Its dual is the sequence:

0 → NX/Y → NX/Z → NY/Z ⊗ OX (D.3)

Lemma D.1.3. (i) If f : X ⊂ Y and g : Y ⊂ Z are regular embeddings of codi- mensions m and n respectively, then g f : X → Z is a regular embedding of codimension m + n. (ii) If the embeddings f and g are both regular then we have exact sequences of locally free sheaves on X:

I α I IX/Y 0 → Y ⊗ O −→ X → → 0(D.4) I2 X I2 I2 Y X X/Y

0 → NX/Y → NX/Z → NY/Z ⊗ OX → 0(D.5)

Proof. (i) Left to the reader. (ii) All sheaves in (D.4) are locally free because they are conormal bundles of regu- (α) (I /I2 ) ⊗ lar embeddings. Since Im is a torsion free sheaf of the same rank of Y Y OX , it follows that α must be injective. The sequence (D.5) is exact because 1 2 Ext (IX/Y /I , OX ) = 0.  OX X/Y Proposition D.1.4. Let j : X ⊂ Y be an embedding of algebraic schemes, with X reduced and Y nonsingular. Consider the conormal sequence

I/I2 −→δ Ω1 → Ω1 → Y |X X 0(D.6)

(where I ⊂ OY is the ideal sheaf of X). Then: (i) The homomorphism δ is injective on the open set where j is a regular embedding. (ii) If X and Y are nonsingular then the dual sequence

0 → TX → TY |X → NX/Y → 0(D.7)

is exact. D.2 Relative complete intersection morphisms 307

Proof. (i) It suffices to show that δ is injective under the assumption that j is a regular embedding. In this case the conormal sheaf I/I2 is locally free of rank equal to the codimension of X. The sequence (D.6) is exact at every nonsingular point x ∈ X by Theorem C.15(iii). Since X is reduced, this happens on a dense open subset so that ker(δ) is supported on a nowhere dense subset. But X has no embedded points because it is regularly embedded in Y : it follows that ker(δ) = 0. Ω1 (ii) Under the stated hypothesis, j is a regular embedding and X is locally free, 1(Ω1 , O ) =  so we have Ext X X 0 and the exactness of (D.7) follows. Remark D.1.5. If we don’t assume X reduced, part (i) of the proposition is false in general. An example is given by the closed regular embedding of codimension 1:

Spec(k[ ]) ⊂ Spec(k[t]) = A1

(see Example B.9(iv)).

A morphism f : X → Y of schemes will be called a cover (or a covering)ifitis finite and surjective. Recall that a morphism of schemes f : X → Y is called unramified at a point ∈ Ω1 = ∈ x X if X/Y,x 0; f is unramified if it is unramified at every x X.After ∆ ⊂ × Ω1 identifying X with the diagonal X Y X, we see that X/Y gets identified with the conormal sheaf of this embedding. It follows that f is unramified at x if and only if ∆ ⊂ X ×Y X is an open embedding at x, and that the locus of x ∈ X such that f is unramified at x is open. Moreover, f is unramified if and only if ∆ is both open and closed in X ×Y X.

D.2 Relative complete intersection morphisms

We now introduce a natural class of morphisms which generalize smooth morphisms and behave well with respect to differentials and base change.

Definition D.2.1. A flat morphism of finite type f : X → S is called a relative complete intersection (r.c.i.) morphism at the point x ∈ X if there is an open neigh- bourhood U of x such that the restriction of f to U can be obtained as a composition

j g U −→ V −→ S where j is a regular embedding and g is smooth. If f is an r.c.i. morphism at every point we call it an r.c.i. morphism, and we call X a complete intersection over S.

This definition is equivalent to Def. 19.3.6 of Ch. IV of [1]; the equivalence is proved in [19], Prop. 1.4. Note that in case S = Spec(k) the morphism f is an r.c.i. if and only if X is an l.c.i. of finite type. 308 D Complete intersections

If X → S is a flat morphism of finite type of nonsingular varieties then f is an r.c.i. because it factors as X → X × S → S where the first morphism is the graph of f . Before discussing the main properties of this notion we need two lemmas.

Lemma D.2.2. Let A → B be a ring homomorphism, M a B-module and f1,..., fn an M-regular sequence of elements of B. Assume that for each i = 1,...,nthe /( i−1 ) → module M j=1 f j M is A-flat. Then, for every ring homomorphism A A , = ⊗ = ⊗ = ⊗ ≤ ≤ letting B B A A ,M M A A , and fi fi 1 (1 i n), the sequence ,..., /( i−1 ) f1 fn of elements of B is M -regular and the modules M j=1 f j M are A -flat.

Proof. Consider the exact sequence:

f1 0 → M −→ M → M/f1 M → 0

Since M/f1 M is A-flat, the sequence:

f1⊗1 0 → M ⊗A A −→ M ⊗A A → (M/f1 M) ⊗A A → 0  is exact, and therefore f is not a zero-divisor for M .LetM = M/( i f M),  1 i j=1 j = /( i ) = ⊗ = / Mi M j=1 f j M ; then we have Mi Mi A A , Mi+1 Mi fi+1 Mi ,

M + = M /f + M . Replacing M and f1 by Mi and fi+1 in the above argument, i 1 i i 1 i one deduces that fi+1 is not a zero-divisor for Mi , thereby proving the first assertion by induction. The last assertion follows from A.2(VII). 

Lemma D.2.3. Let A → B be a local homomorphism of noetherian local rings, M a B-module of finite type, flat over A, and f1,..., fn ∈ m B.For1 ≤ i ≤ nletgi be the image of fi in B ⊗A k, where k = A/m A is the residue field of A. Then the following conditions are equivalent:  ,..., = /( i ) (i) f1 fn is an M-regular sequence, and Mi M j=1 f j M is A-flat for all 1 ≤ i ≤ n. (ii) g1,...,gn is an (M ⊗A k)-regular sequence. Proof. (i) ⇒ (ii) follows from D.2.2 applied to A = k. (ii) ⇒ (i) Applying Corollary A.6, from the injectivity of g1 : M ⊗A k → M ⊗A k we deduce that f1 : M → M is injective and that M1 = M/f1 M is A-flat. Proceeding by induction on i, assume Mi flat over A. Since gi+1 : Mi ⊗A k → Mi ⊗A k is injective from A.6, again we deduce that fi+1 : Mi → Mi is injective and that Mi+1 is A-flat.  In the next proposition some general properties of r.c.i. morphisms are proved.

Proposition D.2.4. (i) An open embedding is an r.c.i. morphism. A smooth morphism of finite type is an r.c.i. morphism. D.2 Relative complete intersection morphisms 309

(ii) If f : X → S is an r.c.i. morphism and h : S → S is a morphism, then

the morphism f : X ×S S → S induced by f after base change is an r.c.i. morphism.

Proof. (i) is an immediate consequence of the definition and (ii) follows easily from Lemma D.2.2.  From D.2.4(ii) it follows in particular that if f : X → S is an r.c.i. morphism then Xs is an l.c.i. for every k-rational point s ∈ S. So for example, a non-l.c.i. algebraic scheme cannot be the fibre of a flat morphism of algebraic nonsingular varieties. The next result gives a useful characterization of r.c.i. morphisms.

Proposition D.2.5. Let j X −→ Y f g (D.8) S be a commutative diagram of morphisms of algebraic schemes, where f is flat, g is smooth and j is an embedding. Then the following conditions are equivalent for a k-rational point x ∈ X: (i) f is an r.c.i. morphism at x. (ii) Letting s = f (x), the fibre Xs is an l.c.i. at x. (iii) j is a regular embedding at x.

Proof. (i) ⇒ (ii) follows from D.2.4(ii) and (iii) ⇒ (i)isobvious. (ii) ⇒ (iii) From (ii) it follows that the embedding js : Xs ⊂ Ys is regular at x. Let I ⊂ OY be the ideal sheaf of X. Tensoring the exact sequence

0 → I → OY → OX → 0

−⊗ by OS k we obtain the sequence

→ I ⊗ → O → O → 0 OS k Ys Xs 0

I ⊗ ( ) which is exact because f is flat. Therefore OS k is the ideal sheaf of j Xs in Ys. Consider a sequence f1,..., fn of sections of I in an open neighbourhood of j(x) I /( I + I2 ) O /( O + I ) which induce a basis of j(x) ms j(x) j(x) as a Y, j(x) ms Y, j(x) j(x) - module. Then the images f1 ⊗ 1 = g1,..., fn ⊗ 1 = gn are generating sections I ⊗ ( ) of OS k in an open neighbourhood of j x in Ys which form a regular sequence in j(x). From Nakayama’s lemma it follows that f1,..., fn generate I in an open neighbourhood of j(x) in Y . From Lemma D.2.3 it follows that f1,..., fn form a regular sequence in j(x) and therefore (iii) holds. 

Corollary D.2.6. Under the hypothesis of Proposition D.2.5, the locus of points x ∈ X such that f is an r.c.i. at x is open. If f is proper then the locus of points s ∈ S such that Xs is an l.c.i. is open. 310 D Complete intersections

Proof. The last assertion follows from the first because a proper map is closed. The first assertion can be proved using characterization D.2.5(iii) of r.c.i. morphism and the fact that the locus where an embedding is regular is open.  We conclude this section with two results about the relative conormal sequence and cotangent sequence for r.c.i. morphisms. Theorem D.2.7. Let j X −→ Y

f ↓ g

S be a commutative diagram of morphisms of algebraic schemes, with f an r.c.i., j an immersion and g smooth. Let J ⊂ OY be the ideal sheaf of j(X). If f is smooth on a dense open subset intersecting every fibre then the relative conormal sequence

→ J /J 2 −→δ ∗Ω1 → Ω1 → 0 j Y/S X/S 0 is exact and J /J 2 is locally free. Proof. From the equivalence (i) ⇔ (iii) in Proposition D.2.5 it follows that j is a regular embedding and therefore J /J 2 is locally free. Moreover, the support of ker(δ) does not contain any generic point of X nor any fibre of f because it is con- tained in the locus where f is not smooth. Since f is generically smooth and j is a regular embedding, X has no embedded components except possibly for some union of fibres. It follows that δ is injective.  Theorem D.2.8. Let f : X → S be an r.c.i morphism of algebraic schemes, and assume f smooth on a dense open subset intersecting every fibre. Then the relative cotangent sequence → ∗Ω1 → Ω1 → Ω1 → 0 f S X X/S 0(D.9) is exact. Proof. We only have to prove the injectivity of the left homomorphism and the ques- tion is local on X. Since all schemes are algebraic, locally on X we can construct the following commutative diagram:

j i X −→ V −→ U

f ↓ ψ ↓ ϕ

h S −→ W where W, U,ψ,ϕ are smooth, i, h are closed embeddings and j is a regular closed embedding. From the smooth morphism ϕ we deduce the exact sequence of locally free sheaves on U: D.2 Relative complete intersection morphisms 311

→ ϕ∗Ω1 → Ω1 → Ω1 → 0 W U U/W 0 which restricts on X to the exact sequence: → ( )∗Ω1 → ( )∗Ω1 → ∗Ω1 → 0 hf W ij U j V/S 0

Let J ⊂ OV and I ⊂ OU be the ideal sheaves of the embeddings j and ij respec- tively, and H ⊂ OW the ideal sheaf of the embedding h. Then we have an exact and commutative diagram of coherent sheaves on X:

0 ↓ f ∗(H/H2) → I/I2 → J /J 2 → 0 ↓↓↓δ j → ( )∗Ω1 → ( )∗Ω1 → ∗Ω1 → 0 hf W ij U j V/S 0 ↓↓↓ ∨ ∗Ω1 −→df Ω1 → Ω1 → f S X X/S 0 ↓↓↓ 000 where the second and the third columns are the relative conormal sequences of ij and of j respectively; the first column is the pullback to X of the conormal sequence ∗ 2 of h; the first row is exact because ψ (H/H ) is the conormal sheaf of i;themapδ j is injective by Theorem D.2.7 and by the assumptions made on X and f . A diagram chasing shows that the codifferential df∨ is injective and proves the theorem. 

NOTES 1. An algebraic scheme can have different embeddings in IP r , i.e. by means of noniso- morphic invertible sheaves, but with same normal sheaf. An example is given by a projective nonsingular curve C of genus 1, and by the embeddings in IP 3 given by two nonisomorphic 2 = 2 invertible sheaves L1and L2 of degree 4 such that L1 L2.ThenC is embedded as a non- singular complete intersection of two quadrics by both sheaves, and the normal bundles are 2 ⊕ 2 = 2 ⊕ 2 L1 L1 L2 L2. 2. Let S be a scheme, and X, Y smooth over S. Prove that every closed S-embedding X ⊂ Y is regular. In particular, every section of a smooth morphism f : Y → S is a regular embedding of codimension equal to the relative dimension of f . 3. Let f : X → S be a morphism of finite type and s ∈ S a k-rational point. Let ms ⊂ OS,s be the maximal ideal and I = IX (s) the ideal sheaf of the fibre X (s) of f over s. Prove that we have a surjective homomorphism

ms 2 ⊗ OX ( ) → I/I 2 k s ms and an injection: NX (s)/X ⊂ TS,s ⊗k OX (s) If f is flat then they are isomorphisms; in particular, if f is flat then NX (s)/X is free. E Functorial language

Let C be a category. A covariant (resp. contravariant) functor F from C to (sets) is said to be representable if there is an object X in C such that F is isomorphic to the functor Y → Hom(X, Y ) (E.1)

(resp. Y → Hom(Y, X)). We will denote by h X a functor of the form (E.1). The representable functors are a full subcategory, isomorphic to C◦ (resp. to C in the contravariant case), of the category Funct(C,(sets)) of covariant functors (resp. Funct(C◦,(sets)) of contravariant functors) from C to (sets). To fix ideas let’s consider covariant functors. In order to investigate conditions for the representability of a given functor F it is convenient to study functorial mor- phisms h X → F. Such morphisms turn out to be easy to describe, thanks to the following elementary lemma:

Lemma E.1 (Yoneda). Let F : C → (sets) be a covariant functor. For each object XinC there is a canonical bijection:

Hom(h X , F) ↔ F(X) Φ → Φ(X)(1X )

Let’s mention, in passing, that functorial morphisms F → h X are much harder to control. They are related to the notion of “coarse moduli scheme”. We may consider couples of the form (X,ξ), where X is an object of C and ξ ∈ F(X). Yoneda’s lemma implies that to give such a couple is equivalent to giving a morphism of functors h X → F; if this morphism is an isomorphism then (X,ξ) is called a universal couple, and ξ a universal element,forF. The existence of a universal couple is equivalent to the representability of F. The couples for F are the objects of a category in which a morphism (X,ξ) → (Y,η)between two couples is by definition a morphism f : X → Y in C 314 E Functorial language such that F( f )(ξ) = η. We denote this category by IF . A morphism f : (X,ξ) → (Y,η)in IF corresponds to a commutative diagram of morphisms of functors: ξ h X −→ F ↑ f  η hY We have an obvious “forgetful functor”

IF → C The fibres of this functor are precisely the sets F(X), which are embedded as sub- categories of IF by ξ → (X,ξ). (Recall that, given a functor G : C → D, the fibre G−1(D) of G over an object D of D is a subcategory of C, consisting of all objects C such that G(C) = D and of all morphisms f such that G( f ) = 1D. A set can be viewed as a category whose objects are its elements and the only morphisms are the identity morphisms.)

Lemma E.2. The functor F is representable if and only if the category IF has an initial object (X,ξ). If this is the case, (X,ξ)is a universal couple for F. The proof is immediate. Note that, since an initial object is unique up to isomor- phism, it follows that a representable functor has a unique universal couple, up to isomorphism. ∗∗∗∗∗∗ Let I and D be two categories. Given an object A of D, the constant functor cA : I → D is defined as cA(i) = A for each object i of I and cA( f ) = 1A for each morphism f in I . Note that cA is both covariant and contravariant. Every morphism α : A → B in D induces an obvious morphism of functors cα : cA → cB. Consider a covariant functor Φ : I → D.Aninductive limit of Φ is an object A of D and a functorial morphism λ : Φ → cA such that for every other morphism µ : Φ → cB there is a morphism α : A → B such that µ = cαλ. λ Φ −→ cA µ ↓ cα cB From the definition it follows that an inductive limit of Φ, if it exists, is unique up to unique isomorphism, and is denoted by

lim Φ → In practice an inductive limit is an object A of D such that there is a morphism Φ(i) → A for each i ∈ Ob(I ) with the condition that the diagram Φ(i) → A ↓ Φ( f )  Φ(j) E Functorial language 315 is commutative for each morphism f : i → j in I ; moreover, these data must satisfy a universal property. Dually, one has the notion of projective limit of a covariant functor Φ : I → D:it is an object A of D and a morphism π : cA → Φ such that for every other morphism ρ : cB → Φ there is a morphism β : B → A such that ρ = πcβ . The projective limit of Φ, if it exists, is denoted by

lim Φ ←

The above notions can be defined without changes replacing the covariant functor Φ by a contravariant one. We will write Φi for Φ(i), for each object i of I , and sometimes

lim Φ (resp. lim Φ ) instead of lim Φ(resp. lim Φ) → i ← i → ←

Example E.3. Let J be a partially ordered set. We define a category Ord(J) as follows. The objects of Ord(J) are the elements of J; for any i, j ∈ J the set HomOrd(J)(i, j) consists of one element if i ≤ j and is ∅ otherwise. A covariant (resp. contravariant) functor Φ : Ord(J) → D is called an inductive system (resp. a projective system)inD indexed by J; in the case D =(sets), we obtain the usual notions of inductive (projective) system and of inductive (projective) limit. If I is a set and Φ : I → D is a functor, where D is a category with arbitrary coproducts, then lim Φ = Φ → i i Similarly, if D has products then lim Φ = Φ ← i i Proposition E.4. The inductive limit and projective limit exist for every functor Φ : I → (sets) from any category I .

Proof. We take lim Φ = Φ /R → i i where R is the equivalence relation generated by pairs (x, y), x ∈ Φi and y ∈ Φ j , such that there exists ϕ : i → j with Φ(x) = y. Similarly for the projective limit. 

Example E.5. Let F : C → (sets) be a covariant functor, and let IF be the category of couples for F. Then we have a contravariant functor

Φ : IF → Funct(C,(sets)) which sends a couple (X,ξ) to the functor h X : C → (sets), and a morphism f : (X,ξ) → (Y,η) to the functorial morphism h f : hY → h X induced by f . 316 E Functorial language

By construction there is a morphism Φ → cF . This morphism makes F the induc- tive limit of the functor Φ (the proof is an easy exercise). We will write:

F = lim h X → (X,ξ)

Definition E.6. A category I is filtered if: (a) for every pair of objects i, j in I there exists an object k in I and morphisms:

i ↓ j → k

→ → → (b) each pair of morphisms i → j has a coequalizer i → j k. The category I is cofiltered if the dual category I ◦ is filtered.

Assume from now on that C is a category with products and fibred products.

Definition E.7. A covariant functor F : C → (sets) is called left exact if F(B×C) = F(B) × F(C) and F(B ×A C) = F(B) ×F(A) F(C) for each diagram C ↓ B → A in C (i.e. F commutes with finite products and finite fibred products).

Every representable functor is left exact by definition of product and fibred prod- uct.

Lemma E.8. Let I be a filtered category and Φ : I → Funct(C,(sets)) a covariant functor. Then, for each diagram in C:

C ↓ B → A there is a bijection: ∼ lim Φ (B) × Φ ( ) lim Φ (C) = lim[Φ (B) ×Φ ( ) φ (C)] → i lim i A → i → i i A i

The proof of this lemma is straightforward and we omit it. The following result is a useful characterization of left-exact functors.

Proposition E.9. A covariant functor F : C → (sets) is left exact if and only if the category IF is cofiltered. E Functorial language 317

Proof. Assume that IF is cofiltered. Applying Lemma E.8 to the functor Φ of Ex- ample E.5, we see that the inductive limit F = lim(X,ξ) h X is left exact because each functor h X is left exact. Conversely, assume that F is left exact. Let (X,ξ),(Y,η) ∈ Ob(IF );wemust find (Z,ζ) → (X,ξ) ↓ (Y,η) → ( ,ζ) = ( × ,(ξ,η)) ( ,ξ) ( ,η) Take Z X Y . Now consider X → Y coming from φ,ψ : X → Y .Wehave F(φ)(ξ) = F(ψ)(ξ) = η Consider the diagram: Γφ X −→ X × Y ↑↑Γψ K → X where Γφ = (1X ,φ)and Γψ = (1X ,ψ)and K = X ×X×Y X. Since F is left exact

F(K ) = F(X) ×F(X×Y ) F(X) and there is χ ∈ F(K ) corresponding to (ξ, ξ): (Γ ) ξ F −→φ (ξ, η)

↑↑F(Γψ )

χ −→ ξ

Then (K,χ)is the equalizer of φ and ψ. Therefore IF is cofiltered.  Let I be a category. A full subcategory J of I is cofinal if for each i ∈ Ob(I ) there is a morphism f : i → j for some j ∈ Ob(J). It follows immediately from the definitions that if Φ : I → D is a covariant functor and ΦJ : J → D is its restriction, then lim Φ = lim Φ → → J ∗∗∗∗∗∗ Let Z be a scheme. In this subsection we will consider contravariant functors defined on (schemes/Z). All we will say holds, with obvious modifications, for functors defined on (algschemes/Z), the full subcategory of algebraic Z-schemes. A contravariant functor ◦ F : (schemes/Z) → (sets) defines on every Z-scheme S a presheaf of sets:

U → F(U) 318 E Functorial language for all open sets U ⊂ S. For this reason a functor as above is also called a presheaf. F is called a sheaf (more precisely a sheaf in the Zariski topology) if it defines a sheaf on every scheme; namely, if for all Z-schemes S and for all open coverings {Ui } of S the following is an exact sequence of sets: → ( ) → ( ) ( ∩ ) F S F Ui → F Ui U j i i, j The most important sheaves are the representable functors, i.e. functors isomorphic to one of the form: S → HomZ (S, X) for some Z-scheme X. Such a functor is called the functor of points of X/Z. It is very important to have conditions, easy to verify in practice, for a contravari- ant functor F : (schemes/Z)◦ → (sets) to be representable. Certainly a necessary condition is that F is a sheaf. Another necessary condition is the following. Recall that a subfunctor G of F is said to be an open (resp. closed) subfunctor if for every scheme S and for every morphism of functors

Hom(−, S) → F the fibred product Hom(−, S) ×F G, which is a subfunctor of Hom(−, S), is repre- sented by an open (resp. closed) subscheme of S. A family of open subfunctors {Gi } of F is a covering of F if for every Z-scheme S and for every morphism of functors Hom(−, S) → F the family {Hom(−, S) ×F Gi } of subschemes of S is an open covering of S. An obvious example is obtained by considering an open (resp. closed) subscheme X of a Z-scheme X: correspondingly, we obtain an open (resp. closed) subfunctor

Hom(−, X ) of Hom(−, X). An open cover {Xi } of X defines a cover of Hom(−, X) by open subfunctors. Therefore a second obvious necessary condition for a functor F to be repre- sentable is that it can be covered by representable open subfunctors. We will now show that these two necessary conditions are also sufficient. Proposition E.10. Let ◦ F : (schemes/Z) → (sets) be a contravariant functor. Suppose that: (a) F is a sheaf; (b) F admits a covering by representable open subfunctors Fi . Then F is representable.

Proof. Letting Fij = Fi ×F Fj , by (b) the projections Fij → Fi correspond to open embedding of schemes Xij → Xi . Therefore the Fi ’s patch together to form a representable functor Hom(−, X), where X is the scheme obtained by patching the Xi ’s together along the Xij’s. By (a), F and Hom(−, X) are isomorphic.  The following is an easy but important remark. E Functorial language 319

Lemma E.11. If F is a sheaf then F is determined by its restriction to the category of affine schemes.

Proof. In fact, if S is any Z-scheme we can consider an affine open cover {Ui }.For any i, j we take an affine open cover {Vi, j,α} of Ui ∩ U j ; composing the map → ( ) ( ∩ ) F Ui → F Ui U j i, j with the inclusions F(Ui ∩ U j ) → α F(Vi, j,α) we obtain the exact sequence: → ( ) → ( ) ( ) F S F Ui → F Vi, j,α i i, j,α which shows that F(S) is determined by its values on affine schemes.  This lemma implies that the functor of points of a scheme X

◦ F = HomZ (−, X) : (schemes) → (sets) is determined by its restriction to the category of affine schemes, or equivalently, by its covariant version: F : (k-algebras) → (sets) Since the category of schemes is isomorphic to the category of functors of points, this means that we can define schemes as certain types of functors on (k-algebras). Thanks to Proposition E.10, we can say that these functors are precisely the sheaves admitting an open cover by affine schemes, i.e. by representable functors. This point of view is very fruitful because it gives the possibility of generalizing the notion of scheme by considering more general functors. The notion of algebraic space is such a generalization (see Artin [13]). References

1. Elements´ de Geom´ etrie´ Algebrique´ , par A. Grothendieck, redig´ es´ avec la collaboration de J. Dieudonne,´ Publ. Math. IHES, n. 4, 8, 11, 17, 20, 24, 32. 2. Fondements de la Geom´ etrie´ Algebrique´ , by A. Grothendieck, exposes´ at Seminaire´ Bourbaki 1957–1962, Secretariat Math. Paris 1962 (some of them are listed separately below). 3. Revetements Etales et Groupe Fondamentale,Seminaire´ de Geom´ etrie´ Algebrique´ du Bois Marie 1960/61, Springer Verlag Lecture Notes in Mathematics vol. 224. 4. AndreM.:´ Homologie des algebres´ commutatives, Springer Grundlehren b. 206 (1974). 5. Andreotti A.: On the complex structures on a class of simply-connected manifolds, Al- gebraic Geometry and Topology – A symposium in honor of S. Lefschetz, Princeton U.P. (1957). 6. Arakelov S.: Families of algebraic curves with fixed degeneracies, Math. USSR Izv. 5 (1971), 1277–1302. 7. Arbarello E., Cornalba M.: Su una proprieta` notevole dei morfismi di una curva a moduli generali in uno spazio proiettivo. Rend. Sem. Mat. Univers. Politecn. Torino 38 (1980), 87–99. 8. Arbarello E., Cornalba M.: Su una congettura di Petri, Comm. Math. Helvetici 56 (1981), 1–38. 9. Arbarello E., Cornalba M., Griffiths Ph., Harris J.: Geometry of Algebraic Curves, vol. I, Springer Grundlehren b. 267 (1985). 10. Arbarello E., Sernesi E.: Petri’s approach to the study of the ideal associated to a special divisor, Inventiones Math. 49 (1978), 99–119. 11. Artin M.: Algebraic approximation of structures over complete local rings, Publ. Math. IHES 36 (1969), 23–58. 12. Artin M.: Algebraization of formal moduli I, in Global Analysis – Papers in honor of K. Kodaira, Princeton Math. Series n. 29 (1969), 21–71. 13. Artin M.: Theoremes de representabilite pour les espaces algebriques, Les Presses de L’universite´ de Montreal (1972). 14. Artin M.: Deformations of Singularities, TATA Lecture Notes vol. 54 (1976). 15. Badescu L.: Algebraic Surfaces, Springer Universitext (2001). 16. Badescu L.: Projective Geometry and Formal Geometry, Birkhauser (2004). 17. Bayer D., Eisenbud D.: Ribbons and their canonical embeddings, Trans. Amer. Math. Soc. 347 (1995), 719–756. 322 References

18. Behrend K.: Localization and Gromow-Witten invariants, in Quantum Cohomology,ed- itors P. de Bartolomeis, B. Dubrovin, C. Reina, Springer Lecture Notes in Mathematics vol. 1776 (1998). 19. Berthelot P.: Le theoreme de Riemann-Roch, SGA 6, exp. VIII. Springer Lecture Notes in Mathematics vol. 225 (1971). 20. Birkenhake C., Lange H.: Complex Abelian Varieties, 2nd ed., Springer Grundlehren b. 302 (2004). 21. Bloch S.: Semiregularity and De Rham cohomology, Inventiones Math. 17 (1972), 51–66. 22. Bolondi G., Kleppe J.O., Miro-Roig` R.M.: Maximal rank curves and singular points of the Hilbert scheme, Compositio Math. 77 (1991), 269–291. 23. Buchweitz R.O., Flenner H.: A semiregularity map for modules and applications to de- formations, Compositio Math. 137 (2003), 135–210. 24. Burns D., Wahl J.: Local contributions to global deformations of surfaces, Inventiones Math. 26 (1974), 67–88. 25. Caporaso L., Sernesi E.: Characterizing curves by their odd theta-characteristics, J. fur die reine u. angew. Math. 562 (2003), 101–135. 26. Castelnuovo G.: Sui multipli di una serie lineare di gruppi di punti appartenente ad una curva algebrica, Rend. Circolo Mat. Palermo 7 (1893), 89–110. 27. Catanese F.: Moduli of algebraic surfaces, in Theory of Moduli – Montecatini Terme, 1985, E. Sernesi Ed., Springer Lecture Notes in Mathematics vol. 1337 (1988), 1–83. 28. Catanese F.: Everywhere non-reduced moduli spaces, Inventiones Math. 98 (1989), 293–310. 29. Chang M.C.: Some obstructed manifolds with very ample canonical class, J. Algebraic Geom. 1 (1992), 1–4. 30. Cheah J.: Cellular decompositions for nested Hilbert schemes of points, Pacific J. Math. 183 (1998), 39–90. 31. Chiantini L., Ciliberto C.: A few remarks on the lifting problem, Asterisque 218 (1993), 95–109. 32. Chiantini L., Ciliberto C.: On the Severi varieties of surfaces in IP 3, J. Algebraic Geom. 8 (1999), 67–83. 33. Chiantini L., Lopez A.F.: Focal loci of families and the genus of curves on surfaces, Proc. Amer. Math. Soc. 127 (1999), 3451–3459. 34. Chiantini L., Sernesi E.: Nodal curves on surfaces of general type, Math. Annalen 307 (1997), 41–56. 35. Ciliberto C.: On the Hilbert scheme of curves of maximal genus in a projective space, Math. Z. 194 (1987), 351–363. 36. Ciliberto C., Sernesi E.: Families of varieties and the Hilbert scheme, in Lectures on Riemann Surfaces (Trieste 1987), 428–499, World Sci. Publishing (1989). 37. Ciliberto C., Sernesi E.: Singularities of the theta divisor and congruences of planes, J. Algebraic Geom. 1 (1992), 231–250. 38. Ciliberto C., Sernesi E.: Singularities of the theta divisor and families of secant spaces to a canonical curve, J. Algebra 171 (1995), 867–893. 39. Ciliberto C., Sernesi E.: On the geometry of canonical curves of odd genus, Comm. in Algebra 28 (2000), 5993–6001. 40. Colmez P., Serre J.P.: Correspondance Grothendieck-Serre, Documents Mathematiques´ 2 (2001), Societ´ eMath´ ematique´ de France. 41. Curtin D.J.: Obstructions to deforming a space curve, Trans. A.M.S. 267 (1981), 83–94. 42. Debarre O.: Higher-Dimensional Algebraic Geometry, Springer Universitext (2001). References 323

43. Deligne P., Mumford D.: The irreducibility of the space of curves of given genus, Publ. Math. IHES 36 (1969), 75–110. 44. Di Gennaro V.: Alcune osservazioni sulle singolarita` dello schema di Hilbert che para- metrizza le varieta` lineari contenute in una varieta` proiettiva, Ricerche di Matematica 39 (1990), 259–291. 45. Dolgachev I.: Weighted projective varieties, in Group Actions and Vector Fields (Vancouver B.C. 1981), Springer Lecture Notes in Mathematics vol. 956 (1982), 34–71. 46. Edidin D.: Notes on the construction of the moduli space of curves, in Recent Progress in Intersection Theory (Bologna, 1997), G. Ellingsrud, W. Fulton, A. Vistoli Eds., Birkhauser (2000), 85–113. 47. Ein L.: Hilbert schemes of smooth space curves, Ann. Sci. E.N.S. 19 (1986), 469–478. 48. Eisenbud D.: Commutative Algebra with a View toward Algebraic Geometry, Springer Graduate Texts in Mathematics vol. 150 (1995). 49. Eisenbud D., Van de Ven A.: On the normal bundles of smooth rational space curves, Math. Annalen 256 (1981) 453–463. 50. Ellia Ph., Fiorentini M.: Defaut de postulation et singularites´ du schema de Hilbert, Annali Univ. Ferrara 30 (1984), 185–198. 51. Ellia Ph., Hartshorne R.: Smooth specializations of space curves: questions and exam- ples, in: Commutative Algebra and Algebraic Geometry, ed. by F. van Oystaeyen, Marcel Dekker Lecture Notes in Pure and Applied Math. n. 206 (1999), 53–79. 52. Enriques F.: Le superficie algebriche, Zanichelli, Bologna (1949). 53. Fantechi B.: Stacks for everybody, in European Congress of Mathematics, Vol. I (Barcelona, 2000), Birkhauser (2001), 349–359. 54. Fantechi B., Manetti M.: Obstruction calculus for functors of Artin rings, I, J. of Algebra 202 (1998), 541–576. 55. Fantechi B., Pardini R.: On the Hilbert scheme of curves in higher-dimensional projec- tive space, Manuscripta Math. 90 (1996), 1–15. 56. Flamini F.: Some results of regularity for Severi varieties of projective surfaces, Comm. Algebra 29 (2001), 2297–2311. 57. Flamini F.: Moduli of nodal curves on smooth surfaces of general type, J. Algebraic Geom. 11 (2002), 725–760. 58. Fogarty J.: Algebraic families on an algebraic surface, Amer. J. Math. 90 (1968), 511–521. 59. Friedman R.: Algebraic Surfaces and Holomorphic Vector Bundles, Springer Universitext (1998). 60. Fulton W.: On nodal curves, in Algebraic Geometry – Open Problems (Ravello, 1982), 146–155, Springer Lecture Notes in Mathematics vol. 997 (1983). 61. Fulton W.: Intersection Theory, Springer Ergebnisse (3) b. 2 (1984). 62. Geertsen J.A., Hirschowitz A.: Saturation theory and very ample Hilbert functions, Math. Z. 245 (2003), 155–181. 63. Ghione F., Sacchiero G.: Normal bundles of rational curves in IP 3, Manuscripta Math. 33 (1980), 111–128. 64. Godement R.: Topologie algebrique´ et theorie´ des fasceaux, Hermann, Paris (1964). 65. Gomez´ T.L.: Algebraic stacks, Proc. Indian Acad. Sci. Math. Sci. 111 (2001), 1–31. 66. Gonzalez M.: Alisamiento de cintas sobre curvas, thesis, Univ. Complutense Madrid (2004). 67. Grauert H., Kerner H.: Deformationen von Singularitaten¨ komplexer Raume,¨ Math. Annalen 153 (1964), 236–260. 324 References

68. Greuel G.M.: A remark on the paper of A. Tannenbaum, Compositio Math. 51 (1984), 185–187. 69. Greuel G.M., Lossen C., Shustin E.: New asymptotics in the geometry of equisingular families of curves, Intern. Math. Res. Notices 13 (1997), 595–611. 70. Griffiths Ph.: Some remarks and examples on continuous systems and moduli, J. Math. and Mech. 16 (1967), 789–802. 71. Grothendieck A.: Geom´ etrie´ formelle et geom´ etrie´ algebrique,´ Seminaire Bourbaki, exp. 182 (1959). 72. Grothendieck A.: Le theoreme d’existence en theorie formelle des modules, Seminaire Bourbaki, exp. 195 (1960). 73. Grothendieck A.: Les schemas de Hilbert. Seminaire Bourbaki, exp. 221 (1960). 74. Grothendieck A.: Familles d’espaces complexes et fondements de la geometrie analy- tique, Seminaire H. Cartan 1960/61, exp. I-X, Secret. Math., Paris 1962. 75. Grothendieck A.: Local Cohomology, Springer Lecture Notes in Mathematics vol. 41 (1967). 76. Grothendieck A.: Categories cofibrees additives et complexe cotangent relatif, Springer Lecture Notes in Mathematics vol. 79 (1968). 77. Grothendieck A., Dieudonne´ J.A.: Elements´ de Geom´ etrie´ Algebrique´ I, Springer Grundlehren b. 166 (1971). 78. Gruson L., Peskine Ch.: Genre des courbes de l’espace projectif, in Algebraic Geom- etry – Proceedings, Tromsø, Norway 1977, L.D. Olson Ed., Springer Lecture Notes in Mathematics vol. 687 (1978), 31–59. 79. Gruson L., Peskine Ch.: Genre des courbes de l’espace projectif (II), Ann. Sci. Ec. Norm. Sup. 15 (1982), 401–418. 80. Guffroy S.: Sur l’incompletude´ de la serie´ lineare´ caracteristique´ d’une famille de courbes planes a` noeuds et a` cusps, Nagoya Math. J. 171 (2003), 51–83. 81. Gustavsen T.S., Laksov D., Skjelnes R.M.: An elementary, explicit, proof of the exis- tence of the Hilbert schemes of points, AG/0506161. 82. Harris J.: On the Severi problem, Inventiones Math. 84 (1986), 445–461. 83. Hartshorne R.: Connectedness of the Hilbert scheme, Publ. Math IHES 29 (1966), 261–304. 84. Hartshorne R.: Algebraic Geometry, Springer Verlag Graduate Texts in Mathematics vol. 52 (1977). 85. Horikawa E.: On deformations of holomorphic maps I, J. Math. Soc. Japan 25 (1973), 372–396. 86. Horikawa E.: On deformations of holomorphic maps II, J. Math. Soc. Japan 26 (1974), 647–667. 87. Horikawa E.: On deformations of quintic surfaces, Inventiones Math. 31 (1975), 43–85. 88. Horikawa E.: On deformations of holomorphic maps III, Math. Annalen 222 (1976), 275–282. 89. Horikawa E.: On deformations of rational maps, J. Fac. Sci. Univ. Tokyo 23 (1976), 581–600. 90. Huybrechts D., Lehn M.: The Geometry of Moduli Spaces of Sheaves, Aspects of Math- ematics vol. E31 (1997), Vieweg. 91. Iarrobino A.: Hilbert scheme of points: overview of last ten years, in Algebraic Geometry – Bowdoin 1985, Proc. of Symposia in Pure Math. 46, part 2, 297–320 (1987). 92. Iarrobino A., Kanev V.: Power sums, Gorenstein Algebras, and Determinantal Loci, Springer Lecture Notes in Mathematics vol. 1721 (1999). 93. Illusie L.: Complexe cotangent et deformations I, II. Springer Lecture Notes in Mathe- matics vol. 239, 283 (1971). References 325

94. Iversen B.: Generic Local Structure in Commutative Algebra, Springer Lecture Notes in Mathematics vol. 310 (1973). 95. Kang P.: A note on the variety of plane curves with nodes and cusps, Proceedings AMS 106 (1989), 309–312. 96. Kas A.: On obstructions to deformations of complex analytic surfaces, Proc. Nat. Acad. Sci. USA 58 (1967), 402–404. 97. Kleiman S.: Geometry on grassmannians and applications to splitting bundles and smoothing cycles, Publ. Math. IHES 36 (1969), 281–298. 98. Kleiman S.: Les theor´ emes` de finitude pour le foncteur de Picard. SGA 6, exp. XIII. Springer Lecture Notes in Mathematics vol. 225 (1971). 99. Kleiman S.: Multiple point formulas I: iteration, Acta Mathematica 147 (1981), 13–49. 100. Kleiman S.: The Picard scheme, AG/0504020. 101. Kleiman S., Lipman J., Ulrich B.: The multiple point schemes of a finite curvilinear map of codimension one, Ark. Mat. 34 (1996), 285–326. 102. Kleppe J.O.: The Hilbert-flag scheme, its properties and its connection with the Hilbert scheme. Applications to curves in 3-space. Preprint, Inst. of Math. Univ. Oslo (1981). 103. Kleppe J.O.: Non-reduced components of the Hilbert scheme of smooth space curves. Springer Lecture Notes in Mathematics vol. 1266 (1987), 181–207. 104. Kodaira K.: On stability of compact submanifolds of complex manifolds, Amer.J.Math. 85 (1963), 79–94. 105. Kodaira K.: Complex Manifolds and Deformations of Complex Structures, Springer Grundlehren b. 283 (1986). 106. Kodaira K., Nirenberg L., Spencer D.C.: On the existence of deformations of complex analytic structures, Annals of Math. 68 (1958), 450–459. 107. Kodaira K., Spencer D.C.: On deformations of complex analytic structures I, II. Ann. of Math. 67 (1958), 328–466. 108. Kodaira K., Spencer D.C.: A theorem of completeness of characteristic systems of com- plete continuous systems, Amer. J. Math. 81 (1959), 477–500. 109. Kollar J.: Rational Curves on Algebraic Varieties, Springer Ergebnisse b. 32 (1996). 110. Kollar J.: Non quasi-projective moduli spaces, AG/0501294. 111. Kuranishi M.: On the locally complete families of complex analytic structures, Ann. of Math. 75 (1962), 536–577. 112. Lange H., Sernesi E.: Severi varieties and branch curves of abelian surfaces of type (1, 3), International J. of Math. 13 (2002), 227–244. 113. Lange H., Sernesi E.: On the Hilbert scheme of a Prym variety, Annali di Matematica 183 (2004), 375–386. 114. Laudal O.A.: Formal Moduli of Algebraic Structures, Springer Lecture Notes in Mathe- matics vol. 754 (1979). 115. Laumon G., Moret-Bailly L.: Champs Algebriques´ , Springer Ergebnisse b. 39 (2000). 116. Lazarsfeld R.: Positivity in Algebraic Geometry, vol. I, II, Springer Ergebnisse b. 48, 49 (2004). 117. Lehn M.: On the cotangent sheaf of Quot schemes, Intern.J.ofMath.9 (1998), 513–522. 118. Le Potier J.: Lectures on Vector Bundles, Cambridge U.P. (1997). 119. Levelt A.H.M.: Functeurs exacts a` gauche, Inventiones Math. 8 (1969), 114–140. 120. Lichtenbaum S., Schlessinger M.: The cotangent complex of a morphism, Transactions AMS 128 (1967), 41–70. 121. Loeser F.: Deformations´ de courbes planes (d’apres´ Severi et Harris), Seminaire´ Bourbaki 1986/87, Asterisque´ 152–153 (1987), 187–205. 122. Luengo I.: On the existence of complete families of projective plane curves which are obstructed, J. London Math. Soc. 36 (1987), 33–43. 326 References

123. MacLane S.: Homology, Springer Grundlehren b. 114 (1967). 124. Manetti M.: Lectures on Deformations of Complex Manifolds, Rendiconti di Matematica s. VII n. 24 (2004), 1–183. 125. Martin-Deschamps M., Perrin D.: Le schema´ de Hilbert des courbes gauches localement Cohen-Macaulay n’est (presque) jamais reduit,´ Ann. Scient. Ec. Norm. Sup. 29 (1996), 757–785. 126. Maruyama M.: On a generalization of elementary transformations of algebraic vector bundles, Rend. Sem. Mat. Univ. Politec. Torino, special issue (1986), 1–13. 127. Matsumura H.: Commutative Ring Theory, Cambridge University Press (1986). 128. Morrow J., Kodaira K.: Complex Manifolds, Holt, Rinehart & Winston, New York (1971). 129. Mumford D.: Further pathologies in algebraic geometry, Amer.J.ofMath.84 (1962), 642–648. 130. Mumford D.: Lectures on Curves on an Algebraic Surface, Annals of Mathematics Stud- ies vol. 59, Princeton U.P. (1966). 131. Mumford D.: A remark on the paper of M. Schlessinger, Rice Univ. Studies 59 (1973), 113–117. 132. Mumford D., Fogarty J.: Geometric Invariant Theory, second enlarged edition, Springer Ergebnisse b. 34 (1982). 133. Nakajima H.: Lectures on Hilbert Schemes of Points on Surfaces, AMS Univ. Lecture Series, vol. 18 (1999). 134. Namba M.: Families of Meromorphic Functions on Compact Riemann Surfaces, Springer Lecture Notes in Mathematics vol. 767 (1979). 135. Norman P., Oort F.: Moduli of abelian varieties, Ann. of Math. (2) 112 (1980), 413-439. 136. Okonek C., Schneider M., Spindler H.: Vector Bundles on Complex Projective Space, Progress in Math. vol. 3, Birkhauser (1980). 137. Oort F.: Finite group schemes, local moduli for abelian varieties, and lifting problems, Compositio Math. 23 (1971), 265-296. 138. Palamodov V.P.: Deformations of complex spaces, in Several Complex Variables IV: Algebraic Aspects of Complex Analysis, Encyclopaedia of Mathematical Sciences, vol. 10, 105–194. Springer (1990). 139. Parshin A.N.: Algebraic curves over function fields I, Math. SSSR Izv., 2 (1968), 1145–1170. 140. Peeva I., Stillman M.: Connectedness of the Hilbert schemes, J. Algebraic Geom. 14 (2005), 193–211. 141. Petri K.: Uber¨ Spezialkurven I, Math. Ann. 93 (1925), 182–209. 142. Piene R., Schlessinger M.: On the Hilbert scheme compactification of the space of twisted cubics, Amer. J. Math. 107 (1985), 761–774. 143. Pinkham H.: Deformations of cones with negative grading, J. of Algebra 30 (1974), 92–102. 144. Pinkham H.: Deformations of algebraic varieties with Gm-action, Asterisque 20 (1974). 145. Popp H.: Moduli Theory and Classification Theory of Algebraic Varieties, Springer Lecture Notes in Mathematics vol. 620 (1977). 146. Quillen D.: On the (co)-homology of commutative rings. Proc. Symp. in Pure Math. 17 (1970), 65–87. 147. Ran Z.: On nodal plane curves, Inventiones Math. 86 (1986), 529–534. 148. Ran Z.: Deformations of maps, in Algebraic Curves and Projective Geometry – Proceed- ings, Trento 1988, E. Ballico and C. Ciliberto Eds., Springer Lecture Notes in Mathe- matics vol. 1389 (1989), 246–253. References 327

149. Ran Z.: Hodge theory and the Hilbert scheme, J. Diff. Geometry 37 (1993), 191–198. 150. Ran Z.: Hodge theory and deformations of maps, Compositio Math. 97 (1995), 309–328. 151. Ran Z.: A note on Hilbert schemes of nodal curves, AG/0410037. 152. Reid M.: Chapters on algebraic surfaces, in Complex Algebraic Geometry (Park City, UT, 1993), 3–159, Amer. Math. Soc., Providence, RI (1997). 153. Sacchiero G.: Fibrati normali di curve razionali dello spazio proiettivo, Annali Univ. Ferrara (VII) 26 (1980) 33–40. 154. Schlessinger M.: Infinitesimal deformations of singularities, PhD thesis, Harvard University (1964). 155. Schlessinger M.: Functors of Artin rings, Transactions AMS 130 (1968), 208–222. 156. Schlessinger M.: Rigidity of quotient singularities, Inventiones Math. 14 (1971), 17–26. 157. Schlessinger M.: On rigid singularities, Rice University Studies vol. 59 (1973), No. 1, 147–162. 158. Segre C.: Sui fuochi di secondo ordine dei sistemi infiniti di piani, e sulle curve iperspaziali con una doppia infinita` di piani plurisecanti, Atti R. Accad. Lincei (5) 30 (1921), 67–71. 159. Segre B.: Esistenza e dimensione dei sistemi continui di curve piane algebriche con dati caratteri, Rendiconti Accad. Naz. Lincei (6) 10 (1929), 31–38. 160. Segre B.: Esistenza e dimensione dei sistemi continui di curve piane algebriche con dati numeri plueckeriani, Rendiconti Accad. Naz. Lincei (6) 10 (1929), 557–560. 161. Sernesi E.: Small deformations of global complete intersections, Bollettino UMI (4) 12 (1975), 138–146. 162. Sernesi E.: Un esempio di curva ostruita in IP 3, Seminario di variabili complesse, Bologna (1981). 163. Sernesi E.: Topics on Families of Projective Schemes, Queen’s Papers in Pure and Applied Mathematics, vol. 73, Kingston, Ontario, Canada (1986). 164. Serre J.P.: Geom´ etrie´ algebrique´ et geom´ etrie´ analytique, Ann. Inst. Fourier, Grenoble 6 (1955-56), 1–42. 165. Serre J.P.: Espaces fibres´ algebriques,´ Seminaire Chevalley exp. 1 (1958). 166. Serre J.P.: Groupes algebriques´ et corps de classes, Hermann, Paris (1959). 167. Serre J.P.: Local Algebra, Springer Monographs in Mathematics (2000). 168. Seshadri C.S.: Theory of moduli. In Algebraic Geometry, Proc. Symposia Pure Math., vol. 29 (1975), pp. 263–304. Amer. Math. Soc., Providence, R. I. 169. Severi F.: Sulla classificazione delle curve algebriche e sul teorema di esistenza di Riemann, Rend. Acc. Naz. Lincei (5) 241 (1915), 877–888 and 1011–1020. 170. Severi F.: Nuovi contributi alla teoria dei sistemi continui di curve appartenenti ad una superficie algebrica, Rend. Acc. Naz. Lincei (5) 25 (1916), 459–471. 171. Severi F.: Vorlesungen uber¨ Algebraische Geometrie, Teubner, Leipzig (1921). 172. Severi F.: Sul teorema fondamentale dei sistemi continui di curve sopra una superficie algebrica, Annali di matematica (4) 23 (1944), 149–181. 173. Severi F.: Geometria dei sistemi algebrici sopra una superficie e sopra una varieta` algebrica II, Cremonese, Roma, 1958. 174. Sheets R.: A criterion for the rigidity of fibered products of rigid algebras, J. of Algebra 46 (1977), 37–53. 175. Stevens J.: Deformations of Singularities, Springer Lecture Notes in Mathematics vol. 1811 (2003). 176. Sundararaman D.: Moduli, Deformations and Classification of Compact Complex Manifolds, Pitman Research Notes in Math. vol. 45 (1980). 177. Svanes T.: Some criteria for rigidity of noetherian rings, Math. Z. 144(1975), 135–145. 328 References

178. Szpiro L.: Sur le theor´ eme` de rigidite´ de Parshin et Arakelov, Asterisque 64 (1978), 169–202. 179. Tannenbaum A.: Families of algebraic curves with nodes, Compositio Math., 41 (1980), 107–126. 180. Tannenbaum A.: On the classical characteristic linear series of plane curves with nodes and cuspidal points: two examples of Beniamino Segre, Compositio Math. 51 (1984), 169–183. 181. Tate J.: Homology of noetherian rings and local rings, Illinois J. Math. 1 (1957), 14–25. 182. Treger R.: The local Severi problem, Bull. AMS 19 (1988), 325–327. 183. Vakil R.: Murphy’s law in algebraic geometry: badly-behaved deformation spaces. AG/0411469. 184. Viehweg E.: Quasi-Projective Moduli for Polarized Manifolds, Springer Ergebnisse (3) b. 30 (1995). 185. Vistoli A.: Intersection theory on algebraic stacks and on their moduli spaces, Inven- tiones Math. 97 (1989), 613–670. 186. Wahl J.: Deformations of plane curves with nodes and cusps, Amer. J. of Math. 96 (1974), 529–577. 187. Walter C.: Some examples of obstructed curves in IP 3, London Math. Soc. LNS 179 (1992), 325–340. 188. Zappa G.: Sull’esistenza, sopra le superficie algebriche, di sistemi continui completi infiniti, la cui curva generica e` a serie caratteristica incompleta, Pont. Acad. Sci. Acta 9 (1945), 91–93. 189. Zariski O.: Algebraic Surfaces, second supplemented edition, Springer Ergebnisse b. 61 (1971). 190. Zariski O., Samuel P.: Commutative Algebra, Van Nostrand, Princeton (1960). List of Symbols

k,1 κX /S,s,31 k(s),1 oξ (e),33 ob(C),1 o(R/Λ),37 C◦,1 o(R),37 A,1 o( f/Λ),37 Aˆ,1 o( f ),38 ∗ A ,1 tF ,46 AΛ,1 df,46 ∗ ˆ AΛ,1 F,46 ˆ AΛ,2 (R, uˆ),47 ∨ E ,2 DefX ,64

IP (V ),2 DefX ,64 IP (E),2 µ(X),69 ˜ (R ,ϕ),9 OS,s,76 ˜ R⊕˜ I ,10 A,76 A[ ],11 DefX ,77 ( ¯, {η }) ExA(R, I ),12 A n ,77 1 ( ¯, η)ˆ T / ,14 A ,77 R A X T 1,14 ,77 R ( , ,η) Ex(X/S, I),15 S s ,78 ( , ,η) O ⊕˜ I,16 B s ,78 X r 1 P ¯ ,81 TX/S,16 A Aut ˆ ,91 T 1,16 u X Def( , ),93 N ,16 X p X/Y T 2 , 109 F ,22 B0 m 2 κ,29 TX , 111 Y κ(ξ),30 HX , 123 κξ ,31 oξ/Y , 130 κ X /S f,s,31 HZ , 136 330 List of Symbols

X /Specf(R) Y HZ , 137 Hilb , 216 Y X A, 138 Hilb , 216 ⊗ (− )Y L A B, 138 Hilb Z P(t), 219 Pλ , 138 (− )Y 0 Hilb Z P(t), 219 m1, 141 Y \Z HilbP(t), 219 aD, 143 HilbY \Z , 219 µ0(L), 144 X/S QuotH, ( ), 219 µ0(V ), 145 P t X c(L), 145 QuotH,P(t), 219 X/S EL , 145 QuotH,P(t), 220 PL , 145 X/S Hilb ( ), 222 Def( , ), 146 P t X L X/S M1, 153 QuotH , 222 X/S PL , 155 QuotH , 222 / PV , 155 HilbX S, 223 wfdat, 157 P(t), 227 Def / / , 158 r X f Y FHP(t), 227 ρ(g, r, n), 160 r FHP(t), 227 N f , 162 [X1,...,Xm], 229 DX/Y , 162 pri , 229 1 ( ) DX/Y , 162 PI t , 229 Def f/Y , 164 prI, 229 Z Def f/Y , 164 FHP(t), 229 oF/Y , 167 Z FHP(t), 230 Φ f , 169 ˜ ¯ P(t), 230 Nϕ, 173 FHZ , 230 Hϕ, 173 N( ,..., )/ , 230 Def , 177 X1 Xm Z j Quot (E), 240 Def , 177 n j Quotn(E), 241 TY (X), 177 Gn(E), 241 FittF k , 199 Hom(X, Y ), 250 (F) Nk , 199 Isom(X, Y ), 250 (ϕ) De , 200 Aut(X), 250 Nk( f ), 200 Y d , 256 HilbP(t), 206 |O(d)|, 256 Hilbr , 206 P(t) Λd , 257 HilbY , 206 Vδ,κ P(t) d , 257 r δ,κ Hilb ( ), 206 P t Vd , 257 [n] Y , 206 Mk( f ), 262 GV,n, 209 ΩB/A, 279 G, 209 dB/A, 279 Gn(V ), 209 PB/A, 279 G(n, N), 210 k[ ], 284 I, 212 tB/Λ, 285 List of Symbols 331

Ω1 P1 X/Y , 287 X , 292 TX/Y , 287 h X , 313 Ω1 X , 287 Funct(C,(sets)), 313 TX , 287 (X,ξ), 313 Tx X, 287 IF , 314 dg, 288 lim Φ, 314 NX/Y , 288 → P1 lim Φ, 315 X/Y , 292 ← Index

Abel–Jacobi carpet, 16 embedding, 182 Cartier, 144 morphism, 143 Castelnuovo, 194 abelian variety, 151 Castelnuovo–Mumford regularity, 194 action, 57 Castelnuovo–Mumford regularity algebra numerical criterion, 193 etale, 293 Cauchy sequence, 39 formally etale, 293 characteristic map, 47, 126, 159, 169, 179, formally smooth, 293 217, 218, 252 less obstructed, 38 global/local, 252 obstructed/unobstructed, 37 of a linear system, 127 of dual numbers, 11 Chern class, 145 of principal parts, 279 codifferential, 288 rigid, 25, 26, 104, 122 relative, 288 smooth, 293 cofinal subcategory, 317 algebraic local ring, 76 complete intersection, 126, 135, 160, 235, algebraic space, 319 305, 307 algebraization, 80 K3 surface, 236 algebraization theorem, 83, 87 nonrigid, 106 ample complex torus, 35 canonical class, 185 cone, 131 surface, 185 over IP 1 × IP 2, 107 approximation theorem, 88 over IP n × IP m, 107, 120 Arakelov, 171 over Veronese embedding of IP n, 120 Artin, 84, 319 rigid, 107, 119, 120 Artin’s algebraization theorem, 83 congruence, 84 Atiyah extension, 145, 154, 244 conic, 95, 195, 240 automorphism functor, 90 conormal bundle, 288 Bertini’s theorem, 237 sequence, 134, 280, 288, 306 blow-up, 73, 171, 182, 184, 185 sheaf, 288 bounded collection of sheaves, 194 coordinate axes, 95, 107 Brill–Noether number, 160 costable subscheme, 181 334 Index cotangent formally versal/universal/semiuniversal, module 79, 87 first, 14 local, 20, 78, 86, 92, 123, 156 second, 109 locally trivial, 21 upper/lower, 19 of a morphism, 156 sequence (relative), 280, 288, 303, 310 of a morphism wfdat, 157 sheaf of a morphism with fixed target, 161 first, 16 of a polarization, 156 second, 111 of a scheme, 20 sheaf/bundle, 287 of a subscheme, 123 space, 285 product, 21 couple, 313 trivial, 21, 123, 156 formal, 47, 126 with general moduli, 79 universal, 313 depth, 117 cover/covering, 307 diagonal, 131, 207, 263, 307 covering of a functor, 318 differential curve, 155 of aD, 143 affine obstructed, 115 of a morphism of functors, 46 canonical, 129, 132 of a morphism of schemes, 288 canonical of genus 4, 160 of the forgetful morphism, 132, 134, 151, cubic, 26 169, 170, 180 elliptic, 26 divisor Gorenstein, 144 Cartier, 127 negatively embedded, 131 ramification, 171, 172 non-hyperelliptic, 183 semiregular, 144 nonsingular, 94, 150 duality pairing, 147 not rigid, 69 dualizing sheaf, 144 obstructed, 185 obstructed in IP 3, 237 Eagon–Northcott complex, 116 of genus ≥ 3, 182 element of genus 0, 100 universal, 206 of genus 1, 93 on a surface in IP 3, 128 elementary transformation, 246 pointed, 93 embedded point, 196 rational, 160, 182 embedding rational normal, 120 regular, 305 unobstructed, 35 rigid, 178 cusp, 106, 174 Segre, 107, 120 cycle map, 248 Veronese, 120, 251 Enriques, 72 Debarre, 161 equisingularity, 254 deformation etale algebraic, 78, 86 cover, 96 complete, 79 homomorphism, 293 effectively parametrized, 79 morphism, 171, 302 first-order/infinitesimal, see first- neighbourhood, 75 order/infinitesimal deformation topology, 76 formal, 46, 77, 78, 80, 84, 86 Euler sequence, 35, 119, 134, 146, 160, 172, formally locally trivial, 78 175, 244, 251, 292 formally trivial, 78 generalization of, 244 Index 335 extended Petri map, 155 forgetful morphism, 132, 134, 151, 155, 169, extension of algebras, 9 180, 183, 184 pullback of, 12 formal pushout of, 12 couple, 126 small, 11 scheme, 126 splits/splitting, 10 formal couple, 47 trivial, 10 universal, 47 versal, 10 versal/semiuniversal, 49 extension of schemes, 15 formal deformation, 46, 77, 86 trivial, 15 algebraizable, 80, 84, 88 effective, 80, 84, 87 faithfully flat module, 269 locally trivial, 78 family of a subscheme, 125 isotrivial, 96 semiuniversal, 74, 86, 114, 164, 178 locally isotrivial, 96 trivial, 78 non-isotrivial, 171 universal, 86, 95 of closed subschemes, 123 versal, 84, 86 product, 21 formal element, 46 universal, 206, 227 semiuniversal, 52, 56, 58, 59, 69, 104, 146 with general moduli, 79, 218 universal, 47 family of deformations, see deformation versal/semiuniversal, 49 filtered category, 316 formal projective space, 81 first iteration morphism, see morphism, first formal scheme, 77 iteration formally etale homomorphism, 293 first-order foci, 253 formally smooth homomorphism, 293 first-order/infinitesimal deformation functor of ξ0 ∈ F(k),44 automorphism, 90 of a morphism, 156 criterion of representability, 318 of a pair (X, L), 146 flag Hilbert, 227 of a scheme, 21 Grassmann, 209 of a subscheme, 123 left exact, 45, 316 of an algebra, 24 less obstructed, 58, 144, 183 of an invertible sheaf, 138 local Hilbert, 123 Fitting ideal, 199, 263 local moduli, 64 flag local Picard, 138 Hilbert functor, 227 local relative Hilbert, 136 Hilbert scheme, 227, 229 locally of finite presentation, 84, 97 of closed subschemes, 229 of deformations of a pair (X, L), 146 Quot scheme, 235 of Artin rings, 44 regularly embedded, 234 of deformations of a closed embedding, variety, 229 177 flat module, 269 of infinitesimal deformations of u0,86 flatness criterion of morphisms, 249 by Hilbert polynomial, 194 of points, 318 in terms of generators and relations, 275 prorepresentable, 44, 47, 60, 125, 136, local, 271 138, 158, 179, 184 foci, 253 Quot, 219 Fogarty’s theorem, 248 representable, 44, 313, 318 336 Index functor (Continued.) infinitesimal automorphisms, 92 smooth, 47, 150 infinitesimal deformation, see first- unobstructed, 53 order/infinitesimal deformation of u0 ∈ F(k),86 general moduli, 79, 218 infinitesimal Torelli theorem, 183 generic flatness, 205 isomorphism graph, 82, 158, 212, 249, 308 of deformations, 21 Grassmann functor, 209 of extensions, 9 grassmannian, 127, 210 grassmannian bundle, 241 jacobian criterion of smoothness, 299 Grothendieck, 82, 88, 194 jacobian variety, 182

Harris, 267 Kleiman, 137 Hartshorne, 218 Kodaira, 182, 184 henselian local ring, 76 dimension, 73, 268 henselization, 76 Kodaira–Nirenberg–Spencer, 92 hermitian form, 152 Kodaira–Spencer, 80 Hilbert functor, 206 class, 30, 166 local, 123 correspondence, 29 local relative, 136, 137 map, 31, 69, 74, 79, 97, 126, 236 Hilbert scheme, 206, 216 map (vanishing), 97 bound on the dimension, 129 Kollar, 122, 161 connectedness, 218 Koszul existence theorem, 213 complex, 109, 236 flag, 229 relations, 109 nested, 235 Krull dimension, 40, 69 nonreduced, 237 reducible, 253 relative, 222, 223 lifting, 129 Hodge decomposition, 183 limit homogeneous space, 27 inductive, 314 formal principal, 28 projective, 315 principal, 27, 57 line homomorphism in a quadric cone, 131 essentially of finite type, 1, 293 linear system, 127, 144, 207 etale, 293 deformation of, 159 formally etale, 293 local formally smooth, 293 complete intersection, 305 obstructed/unobstructed, 37 criterion of flatness, 271, 277 of extensions, 9 deformation, 92 smooth, 293 deformation of a scheme, 20 Horikawa, 156, 185 Hilbert functor, 123 hypersurface Picard functor, 138 of IP n × IP m, 135 relative Hilbert functor, 136, 137 of IP r , 134, 207 ring unobstructed, 114, 131 algebraic, 76 henselian, 76 incidence relation, 212, 244, 292 in the etale topology, 76 index of ramification, 172 local-to-global exact sequence, 65 Index 337 module of differentials, 279 surface, 35 moduli scheme, 44, 79, 187 obstructed/unobstructed deformation coarse, 313 of a morphism, 159 moduli stack, 79 of a morphism with fixed target, 168 morphism of a scheme, 34 Abel–Jacobi, 143 of a subscheme, 131 classifying, 206 of an embedding, 178 etale, 171, 302 obstruction, 33, 131, 168 first iteration, 263 map, 58, 223 forgetful, see forgetful morphism for aD, 143 non-degenerate, 162, 171 for the forgetful morphism, 132, 151, obstructed/unobstructed, 159 169, 170, 180 of nonsingular curves, 159 space Plucker,¨ 211 for DefB , 111 0 quasi-finite, 195 for Def j , 178 relative complete intersection, 307 for DefX ,69 rigid, 156, 159, 168, 171 for Def f/Y , 164 self-transverse codimension 1, 263 for DefX ,70 smooth, 302 for Def(X,L), 146 unramified, 162, 303, 307 for DefX/f/Y , 158 multiple point for Hom(X, Y ), 250 scheme, 200, 262 Y for HX , 129 stratification, 200 X /S Mumford, 54, 74, 156, 194 for HZ , 137 for Pλ , 138 example of, 237 0 for a functor, 53, 58 Nagata, 76 for an algebra, 43 node, 106, 108, 174 for Quot, 226 O non-degenerate morphism, 162 of HilbY ,[X], 217 of O Z , 231 normal bundle, 243, 288 FH ,[X1,...,Xm ] of a rational normal curve, 245 relative/absolute, 37 normal sheaf, 123, 128, 288 Oort, 156 equisingular, 16, 174, 177, 254 of (X1,...,Xm), 230 parameter scheme, 20, 123 of a morphism, 154, 162 Parshin, 171 number of moduli, 69, 72, 74, 94, 132 Petri map, 129, 132, 134, 144, 160 extended, 155 obstructed Picard functor (local), 138 nonreduced scheme, 238 Picard group, 138 affine curve, 115 Plucker¨ morphism, 211 curve, 185 pointed scheme, 21, 156, 159, 179 curve in IP 3, 237 polarization, 156 surface, 185 presheaf, 318 variety of dimension 3, 185 principal G-bundle, 98 obstructed/unobstructed principal parts embedding, 178 algebra of, 279 morphism, 159 sheaf of, 145, 292 scheme, 34 product of curves, 74 subscheme, 131 projection of a curve, 128 338 Index projective bundle, 2, 241 scheme prorepresentability flag Hilbert, 227 of DefX , 89, 92, 94 formal, 77, 126

of DefX ,89 Hilbert, 206 Y multiple point, 200, 262 of HX , 125 obstructed/unobstructed, 34 of Pλ0 , 140 pseudo-torsor, 28 of elementary transformations, 246 pullback of an extension of algebras, 12 of isomorphisms/automorphisms, 250 pushout, 134 of morphisms, 250 of an extension of algebras, 12 pointed, 21, 156, 159, 179 Quot, 223 quadric, 22, 120 rigid, 21, 121 cone, 131, 160 Severi, 257 Quot unobstructed, 79 functor, 219 vanishing, 200 scheme, 223 Schlessinger, 115 flag, 235 theorem of, 56, 64, 104 see-saw theorem, 203 ramification divisor, 171, 172 Segre rational connectedness, 161 embedding, 107, 120 regular example of, 261 embedding, 305 semiregular, 144 sequence, 40 Serre, 82 Severi variety, 260 duality, 151 relative vanishing theorem, 185, 187, 190 codifferential, 288 Severi, 262 complete intersection morphism, 307 problem, 267 conormal sequence, 288, 303, 310 scheme/variety, 257 cotangent sequence, 280, 288, 303, 310 variety cotangent sheaf, 287 regular/superabundant, 260 derivations, 287 Severi–Kodaira–Spencer, 144 differential, 288 sheaf, 209, 318 differentials (sheaf of), 287 (b), 191 dimension of a smooth morphism, 303 m-regular, 187 Hilbert scheme, 222 Castelnuovo–Mumford regular, 187 tangent sheaf, 287 conormal, 288 ribbon, 16 dualizing, 144 rigid normal, 128, 288 algebra, 25, 26, 104, 122 of germs of vectors tangent along a cone, 107, 119, 120 subvariety, 177 curve singularity, 69 of principal parts, 145, 292 embedding, 178 smooth morphism, 156, 159, 168, 171 functor, 47, 150 nonsingular variety, 34 homomorphism, 293 product of projective spaces, 35 morphism, 302 projective space, 35 morphism of functors, 47 scheme, 21, 121 smoothness of forgetful morphism, 155 subscheme, 123, 131, 184 stable subscheme, 181 weighted projective space, 122 stack, 101 Index 339

stratification, 198 of Pλ0 , 138 defined by a sheaf, 199 of FHZ, 231 flattening, 200 of a functor, 46 multiple point, 200 of Quot, 226 subfunctor, 211, 243 tautological open/closed, 318 exact sequence, 210, 241, 244 subscheme invertible sheaf, 208 costable, 181 torsor, 27 obstructed/unobstructed, 131 total scheme, 20, 123 rigid, 123, 131, 184 truncated cotangent complex, 19 stable, 181 unobstructed, 131 uniruledness, 161, 251 superabundant Severi variety, 260 universal surface, 106, 182 family, deformation, element, see family, abelian, 74 deformation, element K3, 73, 74, 82, 135, 136, 150, 236 property, 206, 210 minimal ruled, 72 quotient, 222 obstructed, 185 quotient bundle, 209, 241 obstructed/unobstructed, 35 subbundle, 210, 241 of general type, 268 unobstructed rational, 73 canonical curve, 132 rational ruled, 22, 31, 35, 72, 95, 96, 128, curve, 35 290 functor, 53 ruled, 73 hypersurface, 114, 131 symmetric product, 248 local complete intersection, 114 system morphism, 170 inductive/projective, 315 nonsingular projective variety, 69 Szpiro, 171 rational ruled surface, 35, 170 scheme, 79 tacnode, 106 subscheme, 131 tangent unramified covering of curves, 31 space, 285 unramified morphism, 162, 303, 307 tangent space upper semicontinuous function, 198 of Def j , 178

of DefX ,64 vanishing scheme, 200 of DefX ,64 Veronese embedding, 120, 251 X /S of H , 136 Z of Def f/Y , 164 Wahl, 261 of Def(X,L), 146 weighted projective space, 122 of DefX/f/Y , 158 of HilbY , 217 Yoneda’s lemma, 97, 313 of Hom(X, Y ), 250 Y of HX , 123 Zappa’s example, 246 Grundlehren der mathematischen Wissenschaften A Series of Comprehensive Studies in Mathematics A Selection

246. Naimark/Stern: Theory of Group Representations 247. Suzuki: Group Theory I 248. Suzuki: Group Theory II 249. Chung: Lectures from Markov Processes to Brownian Motion 250. Arnold: Geometrical Methods in the Theory of Ordinary Differential Equations 251. Chow/Hale: Methods of Bifurcation Theory 252. Aubin: Nonlinear Analysis on Manifolds. Monge-Ampère Equations 253. Dwork: Lectures on ρ-adic Differential Equations 254. Freitag: Siegelsche Modulfunktionen 255. Lang: Complex Multiplication 256. Hörmander: The Analysis of Linear Partial Differential Operators I 257. Hörmander: The Analysis of Linear Partial Differential Operators II 258. Smoller: Shock Waves and Reaction-Diffusion Equations 259. Duren: Univalent Functions 260. Freidlin/Wentzell: Random Perturbations of Dynamical Systems 261. Bosch/Güntzer/Remmert: Non Archimedian Analysis – A System Approach to Rigid Analytic Geometry 262. Doob: Classical Potential Theory and Its Probabilistic Counterpart 263. Krasnosel’skiˇı/Zabreˇıko: Geometrical Methods of Nonlinear Analysis 264. Aubin/Cellina: Differential Inclusions 265. Grauert/Remmert: Coherent Analytic Sheaves 266. de Rham: Differentiable Manifolds 267. Arbarello/Cornalba/Griffiths/Harris: Geometry of Algebraic Curves, Vol. I 268. Arbarello/Cornalba/Griffiths/Harris: Geometry of Algebraic Curves, Vol. II 269. Schapira: Microdifferential Systems in the Complex Domain 270. Scharlau: Quadratic and Hermitian Forms 271. Ellis: Entropy, Large Deviations, and Statistical Mechanics 272. Elliott: Arithmetic Functions and Integer Products 273. Nikol’skiˇı: Treatise on the shift Operator 274. Hörmander: The Analysis of Linear Partial Differential Operators III 275. Hörmander: The Analysis of Linear Partial Differential Operators IV 276. Liggett: Interacting Particle Systems 277. Fulton/Lang: Riemann-Roch Algebra 278. Barr/Wells: Toposes, Triples and Theories 279. Bishop/Bridges: Constructive Analysis 280. Neukirch: Class Field Theory 281. Chandrasekharan: Elliptic Functions 282. Lelong/Gruman: Entire Functions of Several Complex Variables 283. Kodaira: Complex Manifolds and Deformation of Complex Structures 284. Finn: Equilibrium Capillary Surfaces 285. Burago/Zalgaller: Geometric Inequalities 286. Andrianaov: Quadratic Forms and Hecke Operators 287. Maskit: Kleinian Groups 288. Jacod/Shiryaev: Limit Theorems for Stochastic Processes 289. Manin: Gauge Field Theory and Complex Geometry 290. Conway/Sloane: Sphere Packings, Lattices and Groups 291. Hahn/O’Meara: The Classical Groups and K-Theory 292. Kashiwara/Schapira: Sheaves on Manifolds 293. Revuz/Yor: Continuous Martingales and Brownian Motion 294. Knus: Quadratic and Hermitian Forms over Rings 295. Dierkes/Hildebrandt/Küster/Wohlrab: Minimal Surfaces I 296. Dierkes/Hildebrandt/Küster/Wohlrab: Minimal Surfaces II 297. Pastur/Figotin: Spectra of Random and Almost-Periodic Operators 298. Berline/Getzler/Vergne: Heat Kernels and Dirac Operators 299. Pommerenke: Boundary Behaviour of Conformal Maps 300. Orlik/Terao: Arrangements of Hyperplanes 301. Loday: Cyclic Homology 302. Lange/Birkenhake: Complex Abelian Varieties 303. DeVore/Lorentz: Constructive Approximation 304. Lorentz/v. Golitschek/Makovoz: Construcitve Approximation. Advanced Problems 305. Hiriart-Urruty/Lemaréchal: Convex Analysis and Minimization Algorithms I. Fundamentals 306. Hiriart-Urruty/Lemaréchal: Convex Analysis and Minimization Algorithms II. Advanced Theory and Bundle Methods 307. Schwarz: Quantum Field Theory and Topology 308. Schwarz: Topology for Physicists 309. Adem/Milgram: Cohomology of Finite Groups 310. Giaquinta/Hildebrandt: Calculus of Variations I: The Lagrangian Formalism 311. Giaquinta/Hildebrandt: Calculus of Variations II: The Hamiltonian Formalism 312. Chung/Zhao: From Brownian Motion to Schrödinger’s Equation 313. Malliavin: Stochastic Analysis 314. Adams/Hedberg: Function spaces and Potential Theory 315. Bürgisser/Clausen/Shokrollahi: Algebraic Complexity Theory 316. Saff/Totik: Logarithmic Potentials with External Fields 317. Rockafellar/Wets: Variational Analysis 318. Kobayashi: Hyperbolic Complex Spaces 319. Bridson/Haefliger: Metric Spaces of Non-Positive Curvature 320. Kipnis/Landim: Scaling Limits of Interacting Particle Systems 321. Grimmett: Percolation 322. Neukirch: Algebraic Number Theory 323. Neukirch/Schmidt/Wingberg: Cohomology of Number Fields 324. Liggett: Stochastic Interacting Systems: Contact, Voter and Exclusion Processes 325. Dafermos: Hyperbolic Conservation Laws in Continuum Physics 326. Waldschmidt: Diophantine Approximation on Linear Algebraic Groups 327. Martinet: Perfect Lattices in Euclidean Spaces 328. Van der Put/Singer: Galois Theory of Linear Differential Equations 329. Korevaar: Tauberian Theory. A Century of Developments 330. Mordukhovich: Variational Analysis and Generalized Differentiation I: Basic Theory 331. Mordukhovich: Variational Analysis and Generalized Differentiation II: Applications 332. Kashiwara/Schapira: Categories and Sheaves. An Introduction to Ind-Objects and Derived Categories 333. Grimmett: The Random-Cluster Model 334. Sernesi: Deformations of Algebraic Schemes 335. Bushnell/Henniart: The Local Langlands Conjecture for GL(2)