Book
Book
Contributors
Mateo Carmona, Tim Holzschuh, Tim Hosgood, Ryan Keleti, Thiago Solovera & Nery, Wayne Yang
Contents
What this is 8
In defense of a translation 8
Notes from the translators 9
EGA IV sections 10
Mathematical warnings 10
Introduction 11
(From EGA I)
1. Rings of fractions 15
1.0 Rings and Algebras 15
1.1 Radical of an ideal. Nilradical and radical of a ring 16
1.2 Modules and rings of fractions 16
1.3 Functorial properties 17
1.4 Change of multiplicative subset 18
1.5 Change of ring 19
1.6 Identification of the module M f as an inductive limit 21
1.7 Support of a module 22
2. Irreducible spaces. Noetherian spaces 22
2.1 Irreducible spaces 22
2.2 Noetherian spaces 24
3. Supplement on sheaves 24
3.1 Sheaves with values in a category 24
3.2 Presheaves on an open basis 26
3.3 Gluing sheaves 27
3.4 Direct images of presheaves 28
3.5 Inverse images of presheaves 29
3.6 Simple and locally simple sheaves 31
3.7 Inverse images of presheaves of groups or rings 31
3.8 Sheaves on pseudo-discrete spaces 32
4. Ringed spaces 32
4.1 Ringed spaces, sheaves of A -modules, A -algebras 32
4.2 Direct image of an A -module 35
4.3 Inverse image of an A -module 36
4.4 Relation between direct and inverse images 37
5. Quasi-coherent and coherent sheaves 39
5.1 Quasi-coherent sheaves 39
5.2 Sheaves of finite type 39
5.3 Coherent sheaves 40
5.4 Locally free sheaves 42
5.5 Sheaves on a locally ringed space 45
6. Flatness 46
6.1 Flat modules 46
6.2 Change of ring 47
2
CONTENTS 3
Chapter II. Elementary global study of some classes of morphisms (EGA II) 200
Summary 200
1. Affine morphisms 200
1.1 S-preschemes and OS -algebras 200
1.2 Affine preschemes over a prescheme 201
1.3 Affine preschemes over S associated to an OS -algebra 202
1.4 Quasi-coherent sheaves over an affine prescheme over S 203
1.5 Change of base prescheme 205
1.6 Affine morphisms 206
1.7 Vector bundle associated to a sheaf of modules 207
2. Homogeneous prime spectra 210
2.1 Generalities on graded rings and modules 210
2.2 Rings of fractions of a graded ring 213
2.3 Homogeneous prime spectrum of a graded ring 215
2.4 The scheme structure on Proj(S) 217
2.5 The sheaf associated to a graded module 219
2.6 The graded S-module associated to a sheaf on Proj(S) 224
2.7 Finiteness conditions 225
2.8 Functorial behaviour 228
2.9 Closed subschemes of a scheme Proj(S) 232
3. Homogeneous spectrum of a sheaf of graded algebras 233
3.1 Homogeneous spectrum of a quasi-coherent graded OY -algebra 233
3.2 Sheaf on Proj(S ) associated to a graded S -module 237
3.3 Graded S -module associated to a sheaf on Proj(S ) 239
3.4 Finiteness conditions 241
3.5 Functorial behaviour 243
3.6 Closed subpreschemes of a prescheme Proj(S ) 245
3.7 Morphisms from a prescheme to a homogeneous spectrum 246
3.8 Criteria for immersion into a homogeneous spectrum 248
4. Projective bundles; ample sheaves 250
4.1 Definition of projective bundles 250
4.2 Morphisms from a prescheme to a projective bundle 251
4.3 The Segre morphism 254
4.4 Immersions into projective bundles; very ample sheaves 256
4.5 Ample sheaves 259
4.6 Relatively ample sheaves 263
5. Quasi-affine morphisms; quasi-projective morphisms; proper morphisms; projective
morphisms 267
5.1 Quasi-affine morphisms 267
CONTENTS 6
Chapter IV. Local study of schemes and their morphisms (EGA IV) 366
Summary 366
1. Relative finiteness conditions. Constructible sets in preschemes 367
1.1 Quasi-compact morphisms 368
2. Base change and flatness 368
3. Associated prime cycles and primary decomposition 368
4. Change of base field for algebraic preschemes 368
5. Dimension, depth, and regularity in locally Noetherian preschemes 368
6. Flat morphisms of locally Noetherian preschemes 368
7. Relations between a local Notherian ring and its completion. Excellent rings 368
8. Projective limits of preschemes 368
9. Constructible properties 368
10. Jacobson preschemes 368
11. Topological properties of finitely presented flat morphisms. Flatness criteria 368
12. Study of fibres of finitely presented flat morphisms 368
13. Equidimensional morphisms 368
14. Universally open morphisms 368
15. Study of fibres of a universally open morphism 368
16. Differential invariants. Differentially smooth morphisms 368
16.1 Normal invariants of an immersion 369
16.2 Functorial properties of the normal invariants of an immersion 371
16.3 Fundamental differential invariants of morphisms of preschemes 375
16.4 Functorial properties of differential invariants 377
16.5 Relative tangent sheaves and bundles; derivations. 385
16.6 Sheaf of p-differentials and exterior differentials. 390
16.7 The PX/Sn (F ). 392
16.8 Differential operators. 394
16.9 Regular and quasi-regular immersions 399
16.10 Differentially smooth morphisms. 403
16.11 Differential operators on a differentially smooth S-prescheme 404
16.12 Case of characteristic zero: Jacobian criterion for differentially smooth morphisms.406
17. Smooth morphisms, unramified (or net) morphisms, and étale morphisms. 406
17.1 Formally smooth morphisms, formally unramified morphisms, formally étale
morphisms. 406
17.2 General properties of differentials 409
17.3 Smooth morphisms, unramified morphisms, étale morphisms 410
17.4 Characterizations of unramified morphisms. 412
18. Supplement on étale morphisms. Henselian local rings and strictly local rings 412
19. Regular immersions and normal flatness 412
20. Meromorphic functions and pseudo-morphisms 412
20.0 Introduction 412
20.1 Meromorphic functions 412
20.2 Pseudo-morphisms and pseudo-functions 416
21. Divisors 416
Bibliography 417
What this is
I N DEFENSE OF A TRANSLATION
2
From Wikipedia :
In January 2010, Grothendieck wrote the letter “Déclaration d’intention de non-
publication” to Luc Illusie, claiming that all materials published in his absence
have been published without his permission. He asks that none of his work be
reproduced in whole or in part and that copies of this work be removed from
libraries.3 [...] This order may have been reversed later in 2010.4
It is a matter of often heated contention as to whether or not any translation of Grothendieck’s
work should take place, given his extremely explicit views on the matter. By no means do we mean
to argue that somehow Grothendieck’s wishes should be invalidated or ignored, nor do we wish
to somehow twist his earlier words around in order to justify what we have done: we fully accept
that he himself would probably have branded this project “an abomination”. With this in mind, it
remains to explain why we have gone ahead anyway.
First, and possibly foremost, it does not make sense (to us) for an individual to own the rights to
knowledge. Arguments can be made about how the EGA is the product of years and years of intense
work by Grothendieck (and Dieudonné, amongst others), and so this is something that he ‘owns’ and
has full control over. Indeed, it is true that there are almost innumerably many sentences in these
works that only Grothendieck himself could have engineered, but, in translation, we have never
improved anything, but only (regrettably, but almost certainly) worsened. The work in these pages
is that of Grothendieck; we have been not much more than typesetters and eager readers. However,
there is some important point to be made about the fact that Grothendieck collaborated and worked
with many other incredibly proficient mathematicians during the writing of this treatise; although it
is impossible to pinpoint which parts exactly others may have contributed (and by no means do we
wish to imply that any of this work is derivative or fraudulent in any way whatsoever—EGA was
written by Grothendieck) it seems fair that, in some amount, there are bits of the EGAs that ‘belong’ to
a broader collection of minds.
It is a very good idea here to repeat the oft-quoted aphorism: “the work here is not ours, but
any mistakes are”—it is very understandable for an author to not want their name on something
that they have not themselves written, or, at the very least, read. This may be, in part, a reason for
Grothendieck’s wishes, but that is pure speculation. Even so, we include this above disclaimer.
1https://github.com/ryankeleti/ega/issues
2https://en.wikipedia.org/wiki/Alexander_Grothendieck#Retirement_into_reclusion_and_death
3
Grothendieck’s letter. Secret Blogging Seminar. 9 February 2010. Retrieved 3 September 2019.
4
Réédition des SGA. Archived from the original on 29 June 2016. Retrieved 12 November 2013.
8
NOTES FROM THE TRANSLATORS 9
Secondly, then, we note that the French version of EGA is still entirely readily accessible.
Anybody reading these copies who is not a native French speaker, will probably be translating
at least some part of EGA into English in their head, or into their notebooks, as they read. This
document is just the product of a few people doing exactly that, but then passing on their efforts to
make things just that little bit easier for anyone else who follows.
Lastly, to quote another adage, “the guilty person is often the loudest”. If it seems like we are
over-eager to defend ourselves because we know that we are somehow in the wrong, it is because
we are, at least partially. Working on this translation has meant going against Grothendieck’s explicit
requests, and for that we are sorry. We only hope that the freedom of knowledge is an excusable
defense.
5
PDFs of which can be found online, hosted by the Grothendieck circle.
MATHEMATICAL WARNINGS 10
EGA IV SECTIONS
In EGA IV-1, the summary included a tentative list of section that EGA IV would contain. As
EGA IV was written, §5, §7, and §§11–21 would contain different sections than initially envisaged.
We include the original listing here:
M ATHEMATICAL WARNINGS
EGA uses prescheme for what is now usually called a scheme, and scheme for what is now usually
called a separated scheme — we have decided to translate “literally” or “historically”, and thus
continue to use the word scheme to mean separated scheme, and prescheme to mean scheme.
In some cases, we (the translators) have changed “→” to “7→” where appropriate.
Introduction
This memoir, and the many others will undoubtedly follow, are intended to form a treatise on I | 5
the foundations of algebraic geometry. They do not, in principle, presume any particular knowledge
of the subject, and it has even been recognised that such knowledge, despite its obvious advantages,
could sometimes (because of the much-too-narrow interpretation—through the birational point of
view—that it usually implies) be a hindrance to the one who wants to become familiar with the point
of view and techniques presented here. However, we assume that the reader has a good knowledge
of the following topics:
(a) Commutative algebra, as it is laid out, for example, in the volumes (in progress of being
written) of the Éléments of N. Bourbaki (and, pending the publication of these volumes, in
Samuel–Zariski [SZ60] and Samuel [Sam53b, Sam53a]).
(b) Homological algebra, for which we refer to Cartan–Eilenberg [CE56] (cited as (M)) and
Godement [God58] (cited as (G)), as well as the recent article by A. Grothendieck [Gro57]
(cited as (T)).
(c) Sheaf theory, where our main references will be (G) and (T); this theory provides the essential
language for interpreting, in “geometric” terms, the essential notions of commutative
algebra, and for “globalizing” them.
(d) Finally, it will be useful for the reader to have some familiarity with functorial language,
which will be constantly used in this treatise, and for which the reader may consult (M),
(G), and especially (T); the principles of this language and the main results of the general
theory of functors will be described in more detail in a book currently in preparation by
the authors of this treatise.
***
It is not the place, in this introduction, to give a more or less summary description from the
“schemes” point of view in algebraic geometry, nor the long list of reasons which made its adoption
necessary, and in particular the systematic acceptance of nilpotent elements in the local rings of
“manifolds” that we consider (which necessarily shifts the idea of rational maps into the background,
in favor of those of regular maps or “morphisms”). To be precise, this treatise aims to systematically
develop the language of schemes, and will demonstrate, we hope, its necessity. Although it would
be easy to do so, we will not try to give here an “intuitive” introduction to the notions developed in I | 6
Chapter I. For the reader who would like to have a glimpse of the preliminary study of the subject
matter of this treatise, we refer them to the conference by A. Grothendieck at the International
Congress of Mathematicians in Edinburgh in 1958 [Gro58], and the exposé [Gro] of the same author.
The work [Ser55a] (cited as (FAC)) of J.-P. Serre can also be considered as an intermediary exposition
between the classical point of view and the schemes point of view in algebraic geometry, and, as
such, its reading may be an excellent preparation for the reading of our Éléments.
***
We give below the general outline planned for this treatise, subject to later modifications,
especially concerning the later chapters.
11
INTRODUCTION 12
In principle, all chapters are considered open to changes, and supplementary sections could
always be added later; such sections would appear in separate fascicles in order to minimize the
inconveniences accompanying whatever mode of publication adopted. When the writing of such
a section is foreseen or in progress during the publication of a chapter, it will be mentioned in the
summary of the chapter in question, even if, owing to certain orders of urgency, its actual publication
clearly ought to have been later. For the convenience of the reader, we give in “Chapter 0” the
necessary tools in commutative algebra, homological algebra, and sheaf theory, that will be used
throughout this treatise, that are more or less well known but for which it was not possible to give
convenient references. It is recommended for the reader to not read Chapter 0 except whilst reading
the actual treatise, when the results to which we refer seem unfamiliar. Besides, we think that in this I|7
way, the reading of this treatise could be a good method for the beginner to familiarize themselves
with commutative algebra and homological algebra, whose study, when not accompanied with
tangible applications, is considered tedious, or even depressing, by many.
***
It is outside of our capabilities to give a historic overview, or even a summary thereof, of the
ideas and results described herein. The text will contain only those references considered particularly
useful for comprehension, and we indicate the origin of only the most important results. Formally, at
least, the subjects discussed in our work are reasonably new, which explains the scarcity of references
made to the fathers of algebraic geometry from the 19th to the beginning of the 20th century, whose
works we know only by hear-say. It is suitable, however, to say some words here about the works
which have most directly influenced the authors and contributed to the development of scheme-
theoretic point of view. We absolutely must mention the fundamental work (FAC) of J.-P. Serre first,
which has served as an introduction to algebraic geometry for more that one young student (the
author of this treatise being one), deterred by the dryness of the classic Foundations of A. Weil [Wei46].
It is there that it is shown, for the first time, that the “Zariski topology” of an “abstract” algebraic
variety is perfectly suited to applying certain techniques from algebraic topology, and notably to be
able to define a cohomology theory. Further, the definition of an algebraic variety given therein is
that which translates most naturally to the idea that we develop here6. Serre himself had incidentally
noted that the cohomology theory of affine algebraic varieties could be translated without difficulty
by replacing the affine algebras over a field by arbitrary commutative rings. Chapters I and II of
this treatise, and the first two paragraphs of Chapter III, can thus be considered, for the most part,
as easy translations, to this bigger framework, of the principal results of (FAC) and a later article
6Just as J.-P. Serre informed us, it is right to note that the idea of defining the structure of a manifold by the data of a
sheaf of rings is due to H. Cartan, who took this idea as the starting point of his theory of analytic spaces. Of course, just as in
algebraic geometry, it would be important in “analytic geometry” to allow the use of nilpotent elements in local rings of
analytic spaces. This extension of the definition of H. Cartan and J.-P. Serre has recently been broached by H. Grauert [Gra60],
and there is room to hope that a systematic report of analytic geometry in this setting will soon see the light of day. It is also
evident that the ideas and techniques developed in this treatise retain a sense of analytic geometry, even though one must
expect more considerable technical difficulties in this latter theory. We can foresee that algebraic geometry, by the simplicity
of its methods, will be able to serve as a sort of formal model for future developments in the theory of analytic spaces.
INTRODUCTION 13
of the same author [Ser57]. We have also vastly profited from the Séminaire de géométrie algébrique
de C. Chevalley [CC]; in particular, the systematic usage of “constructible sets” introduced by him
has turned out to be extremely useful in the theory of schemes (cf. Chapter IV). We have also
borrowed from him the study of morphisms from the point of view of dimension (Chapter IV), that
translates with negligible change to the framework of schemes. It also merits noting that the idea
of “schemes of local rings”, introduced by Chevalley, naturally lends itself to being extended to
algebraic geometry (not having, however, all the flexibility and generality that we intend to give it
here); for the connections between this idea and our theory, see Chapter I, §8. One such extension
has been developed by M. Nagata in a series of memoirs [Nag58a], which contain many special
results concerning algebraic geometry over Dedekind rings7.
***
It goes without saying that a book on algebraic geometry, and especially a book dealing with
the fundamentals, is of course influenced, if only by proxy, by mathematicians such as O. Zariski
and A. Weil. In particular, the Théorie des fonctions holomorphes by Zariski [Zar51], reasonably flexible
thanks to the cohomological methods and an existence theorem (Chapter III, §§4 and 5), is (along
with the method of descent described in Chapter VI) one of the principal tools used in this treatise,
and it seems to us one of the most powerful at our disposal in algebraic geometry.
The general technique in which it is employed can be sketched as follows (a typical example
of which will be given in Chapter XI, in the study of the fundamental group). We have a proper
morphism (Chapter II) f : X → Y of algebraic varieties (or, more generally, of schemes) that we wish
to study on the neighbourhood of a point y ∈ Y, with the aim of resolving a problem P relative to a
neighbourhood of y. We proceed step by step:
1st We can suppose that Y is affine, so that X becomes a scheme defined on the affine ring A
of Y, and we can even replace A by the local ring of y. This reduction is always easy in
practice (Chapter V) and brings us to the case where A is a local ring.
2nd We study the problem in question when A is a local Artinian ring. So that the problem P
still makes sense when A is not assumed to be integral, we sometimes have to reformulate
P, and it appears that we often obtain a better understanding of the problem in doing so, in
an “infinitesimal” way.
3rd The theory of formal schemes (Chapter III, §§3, 4, and 5) lets us pass from the case of an
Artinian ring to a complete local ring.
4th Finally, if A is an arbitrary local ring, considering “multiform (?) sections” of suitable
schemes over X, approximating a given “formal” section (Chapter IV), will let us pass,
by extension of scalars, to the completion of A, from a known result (about the scheme I | 9
induced by X by extension of scalars to the completion of A) to an analogous result for a
finite simple (e.g. unramified) extension of A.
This sketch shows the importance of the systematic study of schemes defined over an Artinian
ring A. The point of view of Serre in his formulation of the theory of local class fields, and the
recent works of Greenberg, seem to suggest that such a study could be undertaken by functorially
attaching, to some such scheme X, a scheme X ′ over the residue field k of A (assumed perfect) of
dimension equal (in nice cases) to n dim X, where n is the height of A.
As for the influence of A. Weil, it suffices to say that it is the need to develop the tools necessary
to formulate, with full generality, the definition of “Weil cohomology”, and to tackle the proof8 of all
the formal properties necessary to establish the famous conjectures in Diophantine geometry [Wei49],
that has been one of the principal motivations for the writing of this treatise, as well as the desire
to find the natural setting of the usual ideas and methods of algebraic geometry, and to give the
authors the chance to understand said ideas and methods.
***
7Among the works that come close to our point of view of algebraic geometry, we pick out the work of E. Kähler [Käh58]
and a recent note of Chow and Igusa [CI58], which go back over certain results of (FAC) in the context of Nagata–Chevalley
theory, as well as giving a Künneth formula.
8To avoid any misunderstanding, we point out that this task has barely been undertaken at the moment of writing this
introduction, and still hasn’t led to the proof of the Weil conjectures.
INTRODUCTION 14
Finally, we believe it useful to warn the reader that they, as did all the authors themselves, will
almost certainly have difficulty before becoming accustomed to the language of schemes, and to
convince themselves that the usual constructions that suggest geometric intuition can be translated,
in essentially only one sensible way, to this language. As in many parts of modern mathematics, the
first intuition seems further and further away, in appearance, from the correct language needed to
express the mathematics in question with complete precision and the desired level of generality. In
practice, the psychological difficulty comes from the need to replicate some familiar set-theoretic
constructions to a category that is already quite different from that of sets (the category of preschemes,
or the category of preschemes over a given prescheme): Cartesian products, group laws, ring laws,
module laws, fibre bundles, principal homogeneous fibre bundles, etc. It will most likely be difficult
for the mathematician, in the future, to shy away from this new effort of abstraction (maybe rather
negligible, on the whole, in comparison with that supplied by our fathers) to familiarize themselves
with the theory of sets.
***
The references are given following the numerical system; for example, in III, 4.9.3, the III
indicates the volume, the 4 the chapter, the 9 the section, and the 3 the paragraph.9 If we reference a
volume from within itself then we omit the volume number.
9[Trans] This is not a direct translation of the original, but instead uses the language more familiar to modern book (and LAT X
E
document) layouts.
CHAPTER 0
Preliminaries (EGA 0)
(1.0.7). A local ring is a ring A in which there exists a unique maximal ideal, which is thus the
complement of the invertible elements, and contains all the ideals ̸= A. If A and B are local rings,
and m and n their respective maximal ideals, then we say that a homomorphism ϕ : A → B is local if
ϕ(m) ⊂ n (or, equivalently, if ϕ−1 (n) = m). By passing to quotients, such a homomorphism then
defines a monomorphism from the residue field A/m to the residue field B/n. The composition of
any two local homomorphisms is a local homomorphism.
1.1. Radical of an ideal. Nilradical and radical of a ring
(1.1.1). Let a be an ideal of a ring A; the radical of a, denoted by r(a), is the set of x ∈ A such that
x n ∈ a for an integer n > 0; it is an ideal containing a. We have r(r(a)) = r(a); the relation a ⊂ b
implies r(a) ⊂ r(b); the radical of a finite intersection of ideals is the intersection of their radicals.
If ϕ is a homomorphism from another ring A′ to A, then we have r(ϕ−1 (a)) = ϕ−1 (r(a)) for any
ideal a ⊂ A. For an ideal to be the radical of an ideal, it is necessary and sufficient that it be an
intersection of prime ideals. The radical of an ideal a is the intersection of the minimal prime ideals
which contain a; if A is Noetherian, there are finitely many of these minimal prime ideals.
The radical of the ideal (0) is also called the nilradical of A; it is the set N of the nilpotent elements
of A. We say that the ring A is reduced if N = (0); for every ring A, the quotient A/N of A by its
nilradical is a reduced ring.
(1.1.2). Recall that the radical r( A) of a (not-necessarily-commutative) ring A is the intersection of the
maximal left ideals of A (and also the intersection of maximal right ideals). The radical of A/r( A) is
(0).
1.2. Modules and rings of fractions 0I | 13
(1.2.1). We say that a subset S of a ring A is multiplicative if 1 ∈ S and the product of two elements
of S is in S. The examples which will be the most important in what follows are: 1st, the set S f of
powers f n (n ⩾ 0) of an element f ∈ A; and 2nd, the complement A − p of a prime ideal p of A.
(1.2.2). Let S be a multiplicative subset of a ring A, and M an A-module; on the set M × S, the
relation between pairs (m1 , s1 ) and (m2 , s2 ):
“there exists an s ∈ S such that s(s1 m2 − s2 m1 ) = 0”
is an equivalence relation. We denote by S−1 M the quotient set of M × S by this relation, and by
m/s the canonical image of the pair (m, s) in S−1 M; we call iSM : m 7→ m/1 (also denoted iS ) the
canonical map from M to S−1 M. This map is, in general, neither injective nor surjective; its kernel is
the set of m ∈ M such that there exists an s ∈ S for which sm = 0.
On S−1 M we define an additive group law by setting
(m1 /s1 ) + (m2 /s2 ) = (s2 m1 + s1 m2 )/(s1 s2 )
(one can check that it is independent of the choice of representative of the elements of S−1 M, which
are equivalence classes). On S−1 A we further define a multiplicative law by setting ( a1 /s1 )( a2 /s2 ) =
( a1 a2 )/(s1 s2 ), and finally an exterior law on S−1 M, acted on by the set of elements of S−1 A, by setting
( a/s)(m/s′ ) = ( am)/(ss′ ). It can then be shown that S−1 A is endowed with a ring structure (called
the ring of fractions of A with denominators in S) and S−1 M with the structure of an S−1 A-module
(called the module of fractions of M with denominators in S); for all s ∈ S, s/1 is invertible in S−1 A, its
inverse being 1/s. The canonical map iSA (resp. iSM ) is a ring homomorphism (resp. a homomorphism
of A-modules, S−1 M being considered as an A-module by means of the homomorphism iSA : A →
S−1 A).
(1.2.3). If S f = { f n }n⩾0 for a f ∈ A, we write A f and M f instead of S− 1 −1
f A and S f M; when A f is
considered as algebra over A, we can write A f = A[1/ f ]. A f is isomorphic to the quotient algebra
A[ T ]/( f T − 1) A[ T ]. When f = 1, A f and M f are canonically identified with A and M; if f is
nilpotent, then A f and M f are 0.
When S = A − p, with p a prime ideal of A, we write Ap and Mp instead of S−1 A and S−1 M; Ap
is a local ring whose maximal ideal q is generated by iSA (p), and we have (iSA )−1 (q) = p; by passing
to quotients, iSA gives a monomorphism from the integral ring A/p to the field Ap /q, which can be
identified with the field of fractions of A/p.
1. RINGS OF FRACTIONS 17
(1.2.4). The ring of fractions S−1 A and the canonical homomorphism iSA are a solution to a universal
mapping problem: any homomorphism u from A to a ring B such that u(S) is composed of invertible
elements in B factors uniquely as
iS u∗
u:A−
→A
S−1 A −→ B
∗
where u is a ring homomorphism. Under the same hypotheses, let M be an A-module, N a B- 0I | 14
module, and v : M → N a homomorphism of A-modules (for the B-module structure on N defined
by u : A → B); then v factors uniquely as
iSM v∗
v : M −→ S−1 M −
→N
where v∗ is a homomorphism of S−1 A-modules (for the S−1 A-module structure on N defined by
u∗ ).
(1.2.5). We define a canonical isomorphism S−1 A ⊗ A M ≃ S−1 M of S−1 A-modules, sending the
element ( a/s) ⊗ m to the element ( am)/s, with the inverse isomorphism sending m/s to (1/s) ⊗ m.
(1.2.6). For every ideal a′ of S−1 A, a = (iSA )−1 (a′ ) is an ideal of A, and a′ is the ideal of S−1 A
generated by iSA (a), which can be identified with S−1 a (1.3.2). The map p′ 7→ (iSA )−1 (p′ ) is an
isomorphism, for the structure given by ordering, from the set of prime ideals of S−1 A to the set
of prime ideals p of A such that p ∩ S = ∅. In addition, the local rings Ap and (S−1 A)S−1 p are
canonically isomorphic (1.5.1).
(1.2.7). When A is an integral ring, for which K denotes its field of fractions, the canonical map
iSA : A → S−1 A is injective for any multiplicative subset S not containing 0, and S−1 A is then
canonically identified with a subring of K containing A. In particular, for every prime ideal p of A,
Ap is a local ring containing A, with maximal ideal pAp , and we have pAp ∩ A = p.
(1.2.8). If A is a reduced ring (1.1.1), so is S−1 A: indeed, if ( x/s)n = 0 for x ∈ A, s ∈ S, then this
means that there exists an s′ ∈ S such that s′ x n = 0, hence (s′ x )n = 0, which, by hypothesis, implies
s′ x = 0, so x/s = 0.
1.3. Functorial properties
(1.3.1). Let M and N be A-modules, and u an A-homomorphism M → N. If S is a multiplicative
subset of A, we define a S−1 A-homomorphism S−1 M → S−1 N, denoted by S−1 u, by setting
S−1 u(m/s) = u(m)/s; if S−1 M and S−1 N are canonically identified with S−1 A ⊗ A M and S−1 A ⊗ A
N (1.2.5), then S−1 u is considered as 1 ⊗ u. If P is a third A-module, and v an A-homomorphism
N → P, we have S−1 (v ◦ u) = (S−1 v) ◦ (S−1 u); in other words, S−1 M is a covariant functor in M,
from the category of A-modules to that of S−1 A-modules (A and S being fixed).
(1.3.2). The functor S−1 M is exact; in other words, if the sequence
u v
M−
→N−
→P
is exact, then so is the sequence
S −1 u S −1 v
S−1 M −−→ S−1 N −−→ S−1 P.
In particular, if u : M → N is injective (resp. surjective), the same is true for S−1 u; if N and P are 0I | 15
submodules of M, S−1 N and S−1 P are canonically identified with submodules of S−1 M, and we
have
S−1 ( N + P) = S−1 N + S−1 P and S−1 ( N ∩ P) = (S−1 N ) ∩ (S−1 P).
(1.3.3). Let ( Mα , ϕβα ) be an inductive system of A-modules; then (S−1 Mα , S−1 ϕβα ) is an inductive
system of S−1 A-modules. Expressing the S−1 Mα and S−1 ϕβα as tensor products ((1.2.5) and (1.3.1)),
it follows from the permutability of the tensor product and inductive limit operations that we have
a canonical isomorphism
S−1 lim Mα ≃ lim S−1 Mα
−→ −→
which is we can further express by saying that the functor S−1 M (in M) commutes with inductive
limits.
1. RINGS OF FRACTIONS 18
(1.3.4). Let M and N be A-modules; there is a canonical functorial (in M and N) isomorphism
( S −1 M ) ⊗ S −1 A ( S −1 N ) ≃ S −1 ( M ⊗ A N )
which sends (m/s) ⊗ (n/t) to (m ⊗ n)/st.
(1.3.5). We also have a functorial (in M and N) homomorphism
S−1 Hom A ( M, N ) −→ HomS−1 A (S−1 M, S−1 N )
which sends u/s to the homomorphism m/t 7→ u(m)/st. When M has a finite presentation, the
above homomorphism is an isomorphism: it is immediate when M is of the form Ar , and we pass to
the general case by starting with the exact sequence A p → Aq → M → 0 and using the exactness of
the functor S−1 M and the left-exactness of the functor Hom A ( M, N ) in M. Note that this is always
the case when A is Noetherian and the A-module M is of finite type.
1.4. Change of multiplicative subset
(1.4.1). Let S and T be multiplicative subsets of a ring A such that S ⊂ T; there exists a canonical
homomorphism ρ T,S A (or simply ρ
T,S ) from S−1 A to T −1 A, sending the element denoted a/s of S−1 A
( S −1 M ) ⊗ S −1 A ( S −1 N )
∼ / S −1 ( M ⊗ A N )
( T −1 M ) ⊗ T −1 A ( T −1 N )
∼ / T −1 ( M ⊗ A N )
T −1 Hom A ( M, N ) / Hom −1 ( T −1 M, T −1 N )
T A
are commutative.
(1.4.3). There is an important case, in which the homomorphism ρ T,S is bijective, when we then
know that every element of T is a divisor of an element of S; we then identify the modules S−1 M
and T −1 M via ρ T,S . We say that S is saturated if every divisor in A of an element of S is in S; by
replacing S with the set T of all the divisors of the elements of S (a set which is multiplicative and
saturated), we see that we can always, if we wish, consider only modules of fractions S−1 M, where
S is saturated.
(1.4.4). If S, T, and U are three multiplicative subsets of A such that S ⊂ T ⊂ U, then we have
ρU,S = ρU,T ◦ ρ T,S .
1. RINGS OF FRACTIONS 19
(1.4.5). Consider an increasing filtered family (Sα ) of multiplicative subsets of A (we write α ⩽ β for
S β ,Sα
Sα ⊂ S β ), and let S be the multiplicative subset α Sα ; let us put ρ βα = ρ A
S
for α ⩽ β; according to
′
(1.4.4), the homomorphisms ρ βα define a ring A as the inductive limit of the inductive system of rings
(Sα−1 A, ρ βα ). Let ρα be the canonical map Sα−1 A → A′ , and let ϕα = ρS,S
A ; as ϕα = ϕβ ◦ ρ βα for α ⩽ β
α
according to (1.4.4), we can uniquely define a homomorphism ϕ : A → S−1 A such that the diagram
′
Sα−1 A
ρ βα
S−
ρα 1 ϕα
β A (α ⩽ β)
}
ρβ ϕβ #
/ S −1 A
ϕ
A′
is commutative. In fact, ϕ is an isomorphism; it is indeed immediate by construction that ϕ is surjective.
On the other hand, if ρα ( a/sα ) ∈ A′ is such that ϕ(ρα ( a/sα )) = 0, then this means that a/sα = 0
in S−1 A, that is to say that there exists an s ∈ S such that sa = 0; but there is a β ⩾ α such that
s ∈ S β , and consequently, as ρα ( a/sα ) = ρ β (sa/ssα ) = 0, we find that ϕ is injective. The case for an
A-module M is treated likewise, and we have thus defined canonical isomorphisms
lim Sα−1 A ≃ (lim Sα )−1 A, lim Sα−1 M ≃ (lim Sα )−1 M,
−→ −→ −→ −→
the second being functorial in M.
0I | 17
(1.4.6). Let S1 and S2 be multiplicative subsets of A; then S1 S2 is also a multiplicative subset of A.
Let us denote by S2′ the canonical image of S2 in the ring S1−1 A, which is a multiplicative subset of
this ring. For every A-module M there is then a functorial isomorphism
−1
S2′ (S1−1 M) ≃ (S1 S2 )−1 M
which sends (m/s1 )/(s2 /1) to the element m/(s1 s2 ).
1.5. Change of ring
(1.5.1). Let A and A′ be rings, ϕ a homomorphism A′ → A, and S (resp. S′ ) a multiplicative subset
ϕ
of A (resp. A′ ), such that ϕ(S′ ) ⊂ S; the composition homomorphism A′ −
→ A → S−1 A factors as
′
−1 ϕS
A′ −→ S′ A′ −→ S−1 A,
′ ′
by (1.2.4); where ϕS ( a′ /s′ ) = ϕ( a′ )/ϕ(s′ ). If A = ϕ( A′ ) and S = ϕ(S′ ), then ϕS is surjective. If
′ ′
A′ = A and ϕ is the identity, then ϕS is exactly the homomorphism ρS,S A defined in (1.4.1).
(1.5.2). Under the hypotheses of (1.5.1), let M be an A-module. There exists a canonical functorial
morphism
−1
σ : S′ ( M[ϕ] ) −→ (S−1 M)[ϕS′ ]
−1 −1
of S′ A′ -modules, sending each element m/s′ of S′
( M[ϕ] ) to the element m/ϕ(s′ ) of (S−1 M)[ϕS′ ] ;
indeed, we immediately see that this definition does not depend on the representative m/s′ of the
element in question. When S = ϕ(S′ ), the homomorphism σ is bijective. When A′ = A and ϕ is the
′
identity, σ is none other than the homomorphism ρS,SM defined in (1.4.1).
When, in particular, we take M = A the homomorphism ϕ defines an A′ -algebra structure on A;
−1
S′ ( A[ϕ] ) is then endowed with a ring structure, with which it can be identified with (ϕ(S′ ))−1 A,
−1
and the homomorphism σ : S′ ( A[ϕ] ) → S−1 A is a homomorphism of S′ −1 A′ -algebras.
(1.5.3). Let M and N be A-modules; by composing the homomorphisms defined in (1.3.4) and (1.5.2),
we obtain a homomorphism
−1
(S−1 M ⊗S−1 A S−1 N )[ϕS′ ] ←− S′ (( M ⊗ N )[ϕ] )
1. RINGS OF FRACTIONS 20
(1.5.6). Under the hypotheses of (1.5.1), let T (resp. T ′ ) be another multiplicative subset of A (resp.
A′ ) such that S ⊂ T (resp. S′ ⊂ T ′ ) and ϕ( T ′ ) ⊂ T. Then the diagram
′
ϕS
S′
−1
A′ / S −1 A
′ ,S′
ρT ρ T,S
′
ϕT
T′
−1
A′ / T −1 A
is commutative. If M is an A-module, then the diagram
S′
−1
( M[ ϕ ] )
σ / ( S −1 M ) S ′
[ϕ ]
′ ,S′
ρT ρ T,S
′ −1
T ( M[ ϕ ] )
σ / ( T −1 M ) T ′
[ϕ ]
( S ′ −1 N ′ ) ⊗ S ′ −1 A ′ ( S −1 A ) [ ϕ S ′ ]
∼ / S −1 ( N ′ ⊗ A ′ A )
τ [ϕ]
ρ T,S
( T ′ −1 N ′ ) ⊗ T ′ −1 A ′ ( T −1 A ) [ ϕ T ′ ]
∼ / T −1 ( N ′ ⊗ A ′ A )
τ [ϕ]
′ ′ −1
is commutative, the left vertical arrow obtained by applying ρ TN ′,S to S′ N ′ and ρ T,S −1
A to S A.
0I | 19
(1.5.7). Let A′′ be a third ring, ϕ′ : A′′ → A′ a ring homomorphism, and S′′ a multiplicative subset of
A′′ such that ϕ′ (S′′ ) ⊂ S′ . Let ϕ′′ = ϕ ◦ ϕ′ ; then we have
S′′ ′ S′′
ϕ′′ = ϕS ◦ ϕ′ .
1. RINGS OF FRACTIONS 21
Let M be an A-module; evidently we have M[ϕ′′ ] = ( M[ϕ] )[ϕ′ ] ; if σ′ and σ′′ are the homomorphisms
defined by ϕ′ and ϕ′′ in the same way as how σ is defined in (1.5.2) by ϕ, then we have the transitivity
formula
σ′′ = σ ◦ σ′ .
Finally, let N ′′ be an A′′ -module; the A-module N ′′ ⊗ A′′ A[ϕ′′ ] is canonically identified with
( N ′′ ⊗ A′′ A′ [ϕ′ ] ) ⊗ A′ A[ϕ] , and likewise the S−1 A-module (S′′ −1 N ′′ ) ⊗S′′ −1 A′′ (S−1 A)[ϕ′′ S′′ ] is canon-
−1 −1
ically identified with ((S′′ N ′′ ) ⊗S′′ −1 A′′ (S′ A′ ) ′′ )⊗
S ′ −1 A ′
(S−1 A)[ϕS′ ] . With these identifica-
[ϕ′ S ]
tions, if τ ′ and τ ′′ are the isomorphisms defined by ϕ and ϕ′′ in
′ the same way as how τ is defined
in (1.5.4) by ϕ, then we have the transitivity formula
τ ′′ = τ ◦ (τ ′ ⊗ 1).
(1.5.8). Let A be a subring of a ring B; for every minimal prime ideal p of A, there exists a minimal
prime ideal q of B such that p = A ∩ q. Indeed, Ap is a subring of Bp (1.3.2) and has a single prime
ideal p′ (1.2.6); since Bp is not 0, it has at least one prime ideal q′ and we necessarily have q′ ∩ Ap = p′ ;
the prime ideal q1 of B, the inverse image of q′ , is thus such that q1 ∩ A = p, and a fortiori we have
q ∩ A = p for every minimal prime ideal q of B contained in q1 .
1.6. Identification of the module M f as an inductive limit
(1.6.1). Let M be an A-module and f an element of A. Consider a sequence ( Mn ) of A-modules, all
identical to M, and for each pair of integers m ⩽ n, let ϕnm be the homomorphism z 7→ f n−m z from
Mm to Mn ; it is immediate that (( Mn ), (ϕnm )) is an inductive system of A-modules; let N = lim Mn
−→
be the inductive limit of this system. We define a canonical functorial A-isomorphism from N to
M f . For this, let us note that, for all n, θn : z 7→ z/ f n is an A-homomorphism from M = Mn to M f ,
and it follows from the definitions that we have θn ◦ ϕnm = θm for m ⩽ n. As a result, there exists
an A-homomorphism θ : N → M f such that, if ϕn denotes the canonical homomorphism Mn → N,
then we have θn = θ ◦ ϕn for all n. Since, by hypothesis, every element of M f is of the form z/ f n for
at least one n, it is clear that θ is surjective. On the other hand, if θ (ϕn (z)) = 0, or, in other words,
if z/ f n = 0, then there exists an integer k > 0 such that f k z = 0, so ϕn+k,n (z) = 0, which gives
ϕn (z) = 0. We can therefore identify M f with lim Mn via θ.
−→
f f
(1.6.2). Now write M f ,n , ϕnm , and ϕn instead of Mn , ϕnm , and ϕn . Let g be another element of A.
Since f n divides f n gn , we have a functorial homomorphism
ρ f g, f : M f −→ M f g ((1.4.1) and (1.4.3));
if we identify M f and M f g with lim M f ,n and lim M f g,n respectively, then ρ f g, f identifies with the 0I | 20
−→ −→
inductive limit of the maps ρnfg, f : M f ,n → M f g,n , defined by ρnfg, f (z) = gn z. Indeed, this follows
immediately from the commutativity of the diagram
ρnfg, f
M f ,n / M f g,n
f fg
ϕn ϕn
ρ f g, f
Mf / M f g.
2. IRREDUCIBLE SPACES. NOETHERIAN SPACES 22
If an irreducible Kolmogoroff space admits a generic point, it admits exactly one, since a
nonempty open set contains any generic point.
Recall that a topological space X is said to be quasi-compact if, from any collection of open sets of
X, one can extract a finite cover of X (or, equivalently, if any decreasing filtered family of nonempty
closed sets has a nonempty intersection). If X is a quasi-compact space, then any nonempty closed
subset A of X contains a minimal nonempty closed set M, because the set of nonempty closed subsets
of A is inductive under the relation ⊃; if, in addition, X is a Kolmogoroff space, M is necessarily a
single point (or, as we say by abuse of language, is a closed point).
(2.1.4). In an irreducible space X, every nonempty open subspace U is irreducible, and if X admits a
generic point x, x is also a generic point of U.
To prove this, let (Uα ) be a cover (whose set of indices is nonempty) of a topological space
X, consisting of nonempty open sets; if X is irreducible, it is necessary and sufficient that Uα is
irreducible for all α, and that Uα ∩ Uβ ̸= ∅ for any α, β. The condition is clearly necessary; to see that
it is sufficient, it suffices to prove that if V is a nonempty open subset of X, then V ∩ Uα is nonempty
for all α, since then V ∩ Uα is dense in Uα for all α, and consequently V is dense in X. Now there is
at least one index γ such that V ∩ Uγ ̸= ∅, so V ∩ Uγ is dense in Uγ , and as for all α, Uα ∩ Vα ̸= ∅,
we also have V ∩ Uα ∩ Uγ ̸= ∅.
0I | 22
(2.1.5). Let X be an irreducible space, and f a continuous map from X into a topological space Y.
Then f ( X ) is irreducible, and if x is a generic point of X, then f ( x ) is a generic point of f ( X ) and
hence also of f ( X ). In particular, if, in addition, Y is irreducible and with a single generic point y,
then for f ( X ) to be everywhere dense, it is necessary and sufficient that f ( x ) = y.
(2.1.6). Any irreducible subspace of a topological space X is contained in a maximal irreducible
subspace, which is necessarily closed. Maximal irreducible subspaces of X are called the irreducible
components of X. If Z1 and Z2 are two irreducible components distinct from the space X, then Z1 ∩ Z2
is a closed rare set in each of the subspaces Z1 , Z2 ; in particular, if an irreducible component of
X admits a generic point (2.1.2), such a point cannot belong to any other irreducible component.
If X has only a finite number of irreducible components Zi (1 ⩽ i ⩽ n), and if, for each i, we put
S
Ui = j̸=i Zj , then the Ui are open, irreducible, disjoint, and their union is dense in X. Let U be an ErrII
open subset of a topological space X. If Z is an irreducible subset of X that intersects U, then Z ∩ U
is open and dense in Z, thus irreducible; conversely, for any irreducible closed subset Y of U, the
closure Y of Y in X is irreducible and Y ∩ U = Y. We conclude that there is a bijective correspondence
between the irreducible components of U and the irreducible components of X which intersect U.
(2.1.7). If a topological space X is a union of a finite number of irreducible closed subspaces Yi , then
the irreducible components of X are the maximal elements of the set of the Yi , because if Z is an
irreducible closed subset of X, then Z is the union of the Z ∩ Yi , from which one sees that Z must
be contained in one of the Yi . Let Y be a subspace of a topological space X, and suppose that Y has
only a finite number of irreducible components Yi , (1 ⩽ i ⩽ n); then the closures Yi in X are the
irreducible components of Y.
(2.1.8). Let Y be an irreducible space admitting a single generic point y. Let X be a topological space,
and f a continuous map from X to Y. Then, for any irreducible component Z of X intersecting f −1 (y),
f ( Z ) is dense in Y. The converse is not necessarily true; however, if Z has a generic point z, and if
f ( Z ) is dense in Y, then we must have f (z) = y (2.1.5); in addition, Z ∩ f −1 (y) is then the closure
of {z} in f −1 (y) and is therefore irreducible, and as an irreducible subset of f −1 (y) containing z is
necessarily contained in Z (2.1.6), z is a generic point of Z ∩ f −1 (y). As any irreducible component
of f −1 (y) is contained in an irreducible component of X, we see that, if any irreducible component
Z of X intersecting f −1 (y) admits a generic point, then there is a bijective correspondence between all
these components and all the irreducible components Z ∩ f −1 (y) of f −1 (y), the generic points of Z
being identical to those of Z ∩ f −1 (y).
3. SUPPLEMENT ON SHEAVES 24
Equivalently, we can say that, for each object T of C, that the family U 7→ Hom( T, F (U )) is a
sheaf of sets.
(3.1.3). Assume that C is the category defined by a “type of structure with morphisms” Σ, the objects
of C being the sets with structures of type Σ and morphisms those of Σ. Suppose that the category C
also satisfies the following condition:
(E) If ( A, (ρα )) is a solution of a universal mapping problem in the category C for families ( Aα ),
( Aαβ ), (ραβ ), then it is also a solution of the universal mapping problem for the same
families in the category of sets (that is, when we consider A, Aα , and Aαβ as sets, ρα and ραβ
as functions)3.
2This is a special case of the more general notion of a (non-filtered) projective limit (see (T, I, 1.8) and the book in
preparation announced in the introduction).
3It can be proved that it also means that the canonical functor C → Set commutes with projective limits (not necessarily
filtered).
3. SUPPLEMENT ON SHEAVES 25
Under these conditions, the condition (F) gives that, when considered as a presheaf of sets,
U 7→ F (U ) is a sheaf. In addition, for a map u : T → F (U ) to be a morphism of C, it is necessary
and sufficient, according to (F), that each map ρα ◦ u is a morphism T → F (Uα ), which means that
the structure of type Σ on F (U ) is the initial structure for the morphisms ρα . Conversely, suppose a
presheaf U 7→ F (U ) on X, with values in C, is a sheaf of sets and satisfies the previous condition; it is
then clear that it satisfies (F), so it is a sheaf with values in C.
(3.1.4). When Σ is a type of a group or ring structure, the fact that the presheaf U 7→ F (U ) with
values in C is a sheaf of sets implies ipso facto that it is a sheaf with values in C (in other words, a
sheaf of groups or rings within the meaning of (G))4. But it is not the same when, for example, C
is the category of topological rings (with morphisms as continuous homomorphisms): a sheaf with
values in C is a sheaf of rings U 7→ F (U ) such that for any open U and any covering of U by
open sets Uα ⊂ U, the topology of the ring F (U ) is to be the least fine making the homomorphisms
F (U ) → F (Uα ) continuous. We will say in this case that U 7→ F (U ), considered as a sheaf of
rings (without a topology), is underlying the sheaf of topological rings U 7→ F (U ). Morphisms
uV : F (V ) → G (V ) (V an arbitrary open subset of X) of sheaves of topological rings are therefore
homomorphisms of the underlying sheaves of rings, such that uV is continuous for all open V ⊂ X;
to distinguish them from any homomorphisms of the sheaves of the underlying rings, we will call
them continuous homomorphisms of sheaves of topological rings. We have similar definitions and
conventions for sheaves of topological spaces or topological groups.
0I | 25
(3.1.5). It is clear that for any category C, if there is a presheaf (respectively a sheaf) F on X with
values in C and U is an open set of X, the F (V ) for open V ⊂ U constitute a presheaf (or a sheaf)
with values in C, which we call the presheaf (or sheaf) induced by F on U and denote it by F |U.
For any morphism u : F → G of presheaves on X with values in C, we denote by u|U the
morphism F |U → G |U consisting of the uV for V ⊂ U.
(3.1.6). Suppose now that the category C admits inductive limits (T, 1.8); then, for any presheaf (and
in particular any sheaf) F on X with values in C and each x ∈ X, we can define the stalk Fx as the
object of C defined by the inductive limit of the F (U ) with respect to the filtered set (for ⊃) of the
open neighbourhoods U of x in X, and the morphisms ρU V : F (V ) → F (U ). If u : F → G is a
defined for any pair (U, V ) of elements of B such that U ⊂ V, with the conditions ρU U = identity and 0I | 26
W V W
ρU = ρU ◦ ρV if U, V, W in B are such that U ⊂ V ⊂ W. We can associate a presheaf with values in C:
U 7→ F (U ) in the ordinary sense, taking for all open U, F ′ (U ) = lim F (V ), where V runs through
←−
the ordered set (for ⊂, not filtered in general) of V ∈ B sets such that V ⊂ U, since the (V ) form a
projective system for the ρW ′
V (V ⊂ W ⊂ U, V ∈ B, W ∈ B). Indeed, if U, U are two open sets of
U′
X such that U ⊂ U ′ , we define ρ′ U as the projective limit (for V ⊂ U) of the canonical morphisms
F ′ (U ′ ) → F (V ), in other words the unique morphism F ′ (U ′ ) → F ′ (U ), which, when composed
with the canonical morphisms F ′ (U ) → F (V ), gives the canonical morphisms F ′ (U ′ ) → F (V );
U′
the verification of the transitivity of ρ′ U is then immediate. Moreover, if U ∈ B, the canonical
morphism F ′ (U ) → F (U ) is an isomorphism, allowing us to identify these two objects5.
(3.2.2). For the presheaf F ′ thus defined to be a sheaf, it is necessary and sufficient that the presheaf
F on B satisfies the condition:
(F0 ) For any covering (Uα ) of U ∈ B by sets Uα ∈ B contained in U, and for any object T ∈ C, the map
which sends f ∈ Hom( T, F (U )) to the family (ρUUα ◦ f ) ∈ Πα Hom( T, F (Uα )) is a bijection
Uβ
from Hom( T, F (U )) to the set of all ( f α ) such that ρU
V ◦ f α = ρV ◦ f β for any pair of indices
α
5If X is a Noetherian space, we can still define F ′ (U ) and show that it is a presheaf (in the ordinary sense) when one
supposes only that C admits projective limits for finite projective systems. Indeed, if U is any open set of X, there is a finite
covering (Vi ) of U consisting of sets of B; for every couple (i, j) of indices, let (Vijk ) be a finite covering of Vi ∩ Vj formed by
sets of B. Let I be the set of i and triples (i, j, k ), ordered only by the relations i > (i, j, k ), j > (i, j, k ); we then take F ′ (U ) to
be the projective limit of the system of F (Vi ) and F (Vijk ); it is easy to verify that this does not depend on the coverings (Vi )
and (Vijk ) and that U 7→ F ′ (U ) is a presheaf.
6It also means that the pair formed by F (U ) and the ρ = ρU is a solution to the universal problem defined in (3.1.1) by
α Uα
the data of Aα = F (Uα ), Aαβ = ΠF (V ) (for V ∈ B such that V ⊂ Uα ∩ Uβ ) and ραβ = (ρV ′′ ) : F (U ) → ΠF (V ) defined
α
′
by the condition that for V ∈ B, V ′ ∈ B, W ∈ B, V ∪ V ′ ⊂ Uα ∩ Uβ , W ⊂ V ∩ V ′ , ρW
V ◦ ρ′′ = ρV ◦ ρ′′ .
V W V′
3. SUPPLEMENT ON SHEAVES 27
U′
verification of the compatibility conditions with the ρ′ U follows from the functorial properties of
the projective limit.
(3.2.4). If the category C admits inductive limits, and if F is a presheaf on the basis B, with values
in C, for each x ∈ X the neighbourhoods of x belonging to B form a cofinal set (for ⊃) in the set of
neighbourhoods of x, therefore, if F ′ is the (ordinary) presheaf corresponding to F , the stalk Fx′ is
equal to limB F (V ) over the set of V ∈ B containing x. If u : F → G is morphism of presheaves
−→
on B with values in C, u′ : F ′ → G ′ the corresponding morphism of ordinary presheaves, u′x is
likewise the inductive limit of the morphisms uV : F (V ) → G (V ) for V ∈ B, x ∈ V.
(3.2.5). We return to the general conditions of (3.2.1). If F is an ordinary sheaf with values in C, F1
the sheaf on B obtained by the restriction of F to B, then the ordinary sheaf F1′ obtained from F1
by the procedure of (3.2.1) is canonically isomorphic to F , by virtue of the condition (F) and the
uniqueness properties of the projective limit. We identify the ordinary sheaf F with F1′ .
If G is a second (ordinary) sheaf on X with values in C, and u : F → G a morphism, the
preceding remark shows that the data of the uV : F (V ) → G (V ) for only the V ∈ B completely
determines u; conversely, it is sufficient, the uV being given for V ∈ B, to verify the commutative
diagram with the restriction morphisms ρW V for V ∈ B, W ∈ B, and V ⊂ W, for there to exist a
morphism u′ and a unique F in G such that uV ′ = u for each V ∈ B (3.2.3).
V
(3.2.6). Suppose that C admits projective limits. Then the category of sheaves on X with values in
C admits projective limits; if (Fλ ) is a projective system of sheaves on X with values in C, the
F (U ) = limλ Fλ (U ) indeed define a presheaf with values in C, and the verification of the axiom (F)
←−
follows from the transitivity of projective limits; the fact that F is then the projective limit of the Fλ
is immediate.
When C is the category of sets, for each projective system (Hλ ) such that Hλ is a subsheaf of Fλ 0I | 28
for each λ, limλ Hλ canonically identifies with a subsheaf of limλ Fλ . If C is the category of abelian
←− ←−
groups, the covariant functor limλ Fλ is additive and left exact.
←−
3.3. Gluing sheaves
(3.3.1). Suppose still that the category C admits (generalized) projective limits. Let X be a topological
space, U = (Uλ )λ∈ L an open cover of X, and for each λ ∈ L, let Fλ be a sheaf on Uλ , with values in
C; for each pair of indices (λ, µ), suppose that we are given an isomorphism θλµ : Fµ |(Uλ ∩ Uµ ) ≃
F |(Uλ ∩ Uµ ); in addition, suppose that for each triple (λ, µ, ν), if we denote by θλµ ′ , θ ′ , θ ′ the
µν λν
′ = θ ′ ◦ θ ′ (gluing condition for the θ ).
restrictions of θλµ , θµν , θλν to Uλ ∩ Uµ ∩ Uν , then we have θλν λµ µν λµ
Then there exists a sheaf F on X, with values in C, and for each λ an isomorphism ηλ : F |Uλ ≃ Fλ
such that, for each pair (λ, µ), if we denote by ηλ′ and ηµ′ the restrictions of ηλ and ηµ to Uλ ∩ Uµ ,
then we have θλµ = ηλ′ ◦ ηµ′−1 ; in addition, F and the ηλ are determined up to unique isomorphism
by these conditions. The uniqueness indeed follows immediately from (3.2.5). To establish the
existence of F , denote by B the open basis consisting of the open sets contained in at least one
Uλ , and for each U ∈ B, choose (by the Hilbert function τ) one of the Fλ (U ) for one of the λ such
that U ⊂ Uλ ; if we denote this object by F (U ), the ρU V for U ⊂ V, U ∈ B, V ∈ B are defined in an
evident way (by means of the θλµ ), and the transitivity conditions is a consequence of the gluing
condition; in addition, the verification of (F0 ) is immediate, so the presheaf on B thus clearly defines
a sheaf, and we deduce by the general procedure (3.2.1) an (ordinary) sheaf still denoted F and
which answers the question. We say that F is obtained by gluing the Fλ by means of the θλµ and we
usually identify the Fλ and F |Uλ by means of the ηλ .
It is clear that each sheaf F on X with values in C can be considered as being obtained by the
gluing of the sheaves Fλ = F |Uλ (where (Uλ ) is an arbitrary open cover of X), by means of the
isomorphisms θλµ reduced to the identity.
(3.3.2). With the same notation, let Gλ be a second sheaf on Uλ (for each λ ∈ L) with values in
C, and for each pair (λ, µ) let us be given an isomorphism ωλµ : Gµ |(Uλ ∩ Uµ ) ≃ Gλ |(Uλ ∩ Uµ ),
these isomorphisms satisfying the gluing condition. Finally, suppose that we are given for each λ a
3. SUPPLEMENT ON SHEAVES 28
uµ
(3.3.2.1) Fµ |(Uλ ∩ Uµ ) / Gµ |(Uλ ∩ Uµ )
uλ
Fλ |(Uλ ∩ Uµ ) / Gλ |(Uλ ∩ Uµ )
are commutative. Then, if G is obtained by gluing the Gλ by means of the ωλµ , there exists a unique
morphism u : F → G such that the diagrams 0I | 29
u|Uλ
F |Uλ / G |Uλ
uλ
Fλ / Gλ
are commutative; this follows immediately from (3.2.3). The correspondence between the family
(uλ ) and u is in a functorial bijection with the subset of Πλ Hom(Fλ , Gλ ) satisfying the conditions
(3.3.2.1) on Hom(F , G ).
(3.3.3). With the notation of (3.3.1), let V be an open set of X; it is immediate that the restrictions to
V ∩ Uλ ∩ Uµ of the θλµ satisfy the gluing condition for the induced sheaves Fλ |(V ∩ Uλ ) and that
the sheaves on V obtained by gluing the latter identifies canonically with F |V.
that the G (U ) and the ρU define a presheaf on Y with values in C, that we call the direct image of F by
V
ψ and we denote it by ψ∗ (F ). If F is a sheaf, we immediately verify the axiom (F) for the presheaf
G , so ψ∗ (F ) is a sheaf.
(3.4.3). Let Z be a third topological space, ψ′ : Y → Z a continuous map, and let ψ′′ = ψ′ ◦ ψ. It is
clear that we have ψ∗′′ (F ) = ψ∗′ (ψ∗ (F )) for each presheaf F on X with values in C; in addition, for
each morphism u : F → G of such presheaves, we have ψ∗′′ (u) = ψ∗′ (ψ∗ (u)). In other words, ψ∗′′ is
the composition of the functors ψ∗′ and ψ∗ , and this can be written as
In addition, for each open set U of Y, the image under the restriction ψ|ψ−1 (U ) of the induced
presheaf F |ψ−1 (U ) is none other than the induced presheaf ψ∗ (F )|U.
(3.4.4). Suppose that the category C admits inductive limits, and let F be a presheaf on X with values
in C; for all x ∈ X, the morphisms Γ(ψ−1 (U ), F ) → Fx (U an open neighbourhood of ψ( x ) in Y)
form an inductive limit, which gives by passing to the limit a morphism ψx : (ψ∗ (F ))ψ( x) → Fx of 0I | 30
the stalks; in general, these morphisms are neither injective or surjective. It is functorial; indeed, if
3. SUPPLEMENT ON SHEAVES 29
/ ( F1 ) x
ψx
(ψ∗ (F1 ))ψ(x)
(ψ∗ (u))ψ( x) ux
/ ( F2 ) x
ψx
(ψ∗ (F2 ))ψ(x)
(3.4.5). Under the hypotheses of (3.4.4), suppose in addition that ψ is a homeomorphism from X to
the subspace ψ( X ) of Y. Then, for each x ∈ X, ψx is an isomorphism. This applies in particular to the
canonical injection j of a subset X of Y into Y.
(3.4.6). Suppose that C be the category of groups, or of rings, etc. If F is a sheaf on X with values in
C, of support S, and if y ̸∈ ψ(S), then it follows from the definition of ψ∗ (F ) that (ψ∗ (F ))y = {0},
or in other words, that the support of ψ∗ (F ) is contained in ψ(S); but it is not necessarily contained
in ψ(S). Under the same hypotheses, if j is the canonical injection of a subset X of Y into Y, the sheaf
j∗ (F ) induces F on X; if moreover X is closed in Y, j∗ (F ) is the sheaf on Y which induces F on X
and 0 on Y − X (G, II, 2.9.2), but it is in general distinct from the latter when we suppose that X is
locally closed but not closed.
(3.5.1). Under the hypotheses of (3.4.1), if F (resp. G ) is a presheaf on X (resp. Y) with values in
C, then each morphism u : G → ψ∗ (F ) of presheaves on Y is called a ϕ-morphism from G to F ,
and we denote it also by G → F . We denote also by Homϕ (G , F ) the set of HomY (G , ψ∗ (F )) the
ψ-morphisms from G to F . For each pair (U, V ), where U is an open set of X, V an open set of
Y such that ψ(U ) ⊂ V, we have a morphism uU,V : G (V ) → F (U ) by composing the restriction
morphism F (ψ−1 (V )) → F (U ) and the morphism uV : G (V ) → ψ∗ (F )(V ) = F (ψ−1 (V )); it is
immediate that these morphisms render commutative the diagrams
uU,V
(3.5.1.1) G (V ) / F (U )
uU ′ ,V ′
G (V ′ ) / F (U ′ )
(3.5.3). Let ψ : X → Y be a continuous map, G a presheaf on Y with values in C. We call the inverse
image of G under ψ the pair (G ′ , ρ), where G ′ is a sheaf on X with values in C, and ρ : G → G ′ a
ψ-morphism (in other words a homomorphism G → ψ∗ (G ′ )) such that, for each sheaf F on X with
values in C, the map
(3.5.3.1) HomX (G ′ , F ) −→ Homψ (G , F ) −→ HomY (G , ψ∗ (F ))
sending v to ψ∗ (v) ◦ ρ, is a bijection; this map, being functorial in F , then defines an isomorphism of
functors in F . The pair (G ′ , ρ) is the solution of a universal problem, and we say it is determined up
to unique isomorphism when it exists. We then write G ′ = ψ∗ (G ), ρ = ρG , and by abuse of language,
we say that ψ∗ (G ) is the inverse image sheaf of G under ψ, and we agree that ψ∗ (G ) is considered as
equipped with a canonical ψ-morphism ρG : G → ψ∗ (G ), that is to say the canonical homomorphism of
presheaves on Y:
(3.5.3.2) ρG : G −→ ψ∗ (ψ∗ (G )).
For each homomorphism v : ψ∗ (G ) → F (where F is a sheaf on X with values in C), we put
v♭ = ψ∗ (v) ◦ ρG : G → ψ∗ (F ). By definition, each morphism of presheaves u : G → ψ∗ (F ) is of the
form v♭ for a unique v, which we will denote u♯ . In other words, each morphism u : G → ψ∗ (F ) of
presheaves factorizes in a unique way as
ρG ψ∗ (u♯ )
(3.5.3.3) u : G −→ ψ∗ (ψ∗ (G )) −−−→ ψ∗ (F ).
0I | 32
(3.5.4). Suppose now that the category C be such7 that each presheaf F on Y with values in C admits
an inverse image under ψ, and we denote it by ψ∗ (G ).
We will see that we can define ψ∗ (G ) as a covariant functor in G , from the category of presheaves
on Y with values in C, to that of sheaves on X with values in C, in such a way that the isomorphism
v 7→ v♭ is an isomorphism of bifunctors
(3.5.4.1) HomX (ψ∗ (G ), F ) ≃ HomY (G , ψ∗ (F ))
in G and F .
Indeed, for each morphism w : G1 → G2 of presheaves on Y with values in C, consider the
w ρG
composite morphism G1 − → G2 −−→ 2
ψ∗ (ψ∗ (G2 )); to it corresponds a morphism (ρG2 ◦ w)♯ : ψ∗ (G1 ) →
ψ∗ (G2 ), that we denote by ψ∗ (w). We therefore have, according to (3.5.3.3),
(3.5.4.2) ψ∗ (ψ∗ (w)) ◦ ρG1 = ρG2 ◦ w.
For each morphism u : G2 → ψ∗ (F ), where F is a sheaf on X with values in C, we have, according
to (3.5.3.3), (3.5.4.2), and the definition of u♭ , that
(u♯ ◦ ψ∗ (w))♭ = ψ∗ (u♯ ) ◦ ψ∗ (ψ∗ (w)) ◦ ρG1 = ψ∗ (u♯ ) ◦ ρG2 ◦ w = u ◦ w
where again
(3.5.4.3) ( u ◦ w ) ♯ = u ♯ ◦ ψ ∗ ( w ).
w′ ρG3
If we take in particular for u a morphism G2 −→ G3 −−→ ψ∗ (ψ∗ (G3 )), it becomes ψ∗ (w′ ◦ w) =
(ρG3 ◦ w′ ◦ w)♯ = (ρG3 ◦ w′ )♯ ◦ ψ∗ (w) = ψ∗ (w′ ) ◦ ψ∗ (w), hence our assertion.
Finally, for each sheaf F on X with values in C, let iF be the identity morphism of ψ∗ (F ) and
denote by
σF : ψ∗ (ψ∗ (F )) −→ F
the morphism (iF )♯ ; the formula (3.5.4.3) gives in particular the factorization
ψ∗ (u) F σ
(3.5.4.4) u♯ : ψ∗ (G ) −−−→ ψ∗ (ψ∗ (F )) −−
→F
for each morphism u : G → ψ∗ (F ). We say that the morphism σF is canonical.
7In the book mentioned in the introduction, we will give very general conditions on the category C ensuring the existence
of inverse images of presheaves with values in C.
3. SUPPLEMENT ON SHEAVES 31
(3.5.5). Let ψ′ : Y → Z be a continuous map, and suppose that each presheaf H on Z with values in
∗
C admits an inverse image ψ′ (H ) under ψ′ . Then (with the hypotheses of (3.5.4)) each presheaf H
on Z with values in C admits an inverse image under ψ′′ = ψ′ ◦ ψ and we have a canonical functorial
isomorphism
∗ ∗
(3.5.5.1) ψ′′ (H ) ≃ ψ∗ (ψ′ (H )).
This indeed follows immediately from the definitions, taking into account that ψ∗′′ = ψ∗′ ◦ ψ∗ . In 0I | 33
addition, if u : G → ψ∗ (F ) is a ψ-morphism, v : H → ψ∗′ (G ) a ψ′ -morphism, and w = ψ∗′ (u) ◦ v
their composition (3.5.2), then we have immediately that w♯ is the composite morphism
∗ ψ∗ (v♯ ) u♯
w♯ : ψ∗ (ψ′ (H )) −−−→ ψ∗ (G ) −
→ F.
(3.5.6). We take in particular for ψ the identity map 1X : X → X. Then if the inverse image under ψ
of a presheaf F on X with values in C exists, we say that this inverse image is the sheaf associated to
the presheaf F . Each morphism u : F → F ′ from F to a sheaf F ′ with values in C factorizes in a
ρF u♯
unique way as F −−→ 1∗X (F ) −
→ F ′.
3.6. Simple and locally simple sheaves
(3.6.1). We say that a presheaf F on X, with values in C, is constant if the canonical morphisms
F ( X ) → F (U ) are isomorphisms for each nonempty open U ⊂ X; we note that F is not necessarily
a sheaf. We say that a sheaf is simple if it is the associated sheaf (3.5.6) of a constant presheaf. We
say that a sheaf F is locally simple if each x ∈ X admits an open neighbourhood U such that F |U is
simple.
(3.6.2). Suppose that X is irreducible (2.1.1); then the following properties are equivalent:
(a) F is a constant presheaf on X;
(b) F is a simple sheaf on X;
(c) F is a locally simple sheaf on X.
Indeed, let F be a constant presheaf on X; if U, V are two nonempty open sets in X, then U ∩ V
is nonempty, so F ( X ) → F (U ) → F (U ∩ V ) and F ( X ) → F (U ) are isomorphisms, and similarly
both F (U ) → F (U ∩ V ) and F (V ) → F (U ∩ V ) are isomorphisms. We therefore conclude
immediately that the axiom (F) of (3.1.2) is clearly satisfied, F is isomorphic to its associated sheaf,
and as a result (a) implies (b).
Now let (Uα ) be an open cover of X by nonempty open sets and let F be a sheaf on X such that
F |Uα is simple for each α; as Uα is irreducible, F |Uα is a constant presheaf according to the above.
As Uα ∩ Uβ is not empty, F (Uα ) → F (Uα ∩ Uβ ) and F (Uβ ) → F (Uα ∩ Uβ ) are isomorphisms,
hence we have a canonical isomorphism θαβ : F (Uα ) → F (Uβ ) for each pair of indices. But then if
we apply the condition (F) for U = X, we see that for each index α0 , F (Uα0 ) and the θα0 α are solutions
to the universal problem, which (according to the uniqueness) implies that F ( X ) → F (Uα0 ) is an
isomorphism, and hence proves that (c) implies (a).
3.7. Inverse images of presheaves of groups or rings 0I | 34
(3.7.1). We will show that when we take C to be the category of sets, the inverse image under ψ for
each presheaf G with values in C always exists (the notation and hypotheses on X, Y, ψ being that of
(3.5.3)). Indeed, for each open U ⊂ X, define G ′ (U ) as follows: an element s′ of G ′ (U ) is a family
(s′x ) x∈U , where s′x ∈ Gψ(x) for each x ∈ U, and where, for each x ∈ U, the following condition is
satisfied: there exists an open neighbourhood V of ψ( x ) in Y, a neighbourhood W ⊂ ψ−1 (V ) ∩ U
of x, and an element s ∈ G (V ) such that s′z = sψ( x) for all z ∈ W. We verify immediately that
U 7→ G ′ (U ) clearly satisfies the axioms of a sheaf.
Now let F be a sheaf of sets on X, and let u : G → ψ∗ (F ), v : G ′ → F be morphisms. We define
u and v♭ in the following manner: if s′ is a section of G ′ over a neighbourhood U of x ∈ X and if V is
♯
an open neighbourhood of ψ( x ) and s ∈ G (V ) such that we have s′z = sψ( x) for z in a neighbourhood
of x contained in ψ−1 (V ) ∩ U, we take u♯x (s′x ) = uψ( x) (sψ( x) ). Similarly, if s ∈ G (V ) (V open in Y),
v♭ (s) is the section of F over ψ−1 (V ), the image under v of the section s′ of G ′ such that s′x = sψ( x)
4. RINGED SPACES 32
for all x ∈ ψ−1 (V ). In addition, the canonical homomorphism (3.5.3) ρ : G → ψ∗ (ψ∗ (G )) is defined
in the following manner: for each open V ⊂ Y and each section s ∈ Γ(V, G ), ρ(s) is the section
(sψ(x) ) x∈ψ−1 (V ) of ψ∗ (G ) over ψ−1 (V ). The verification of the relations (u♯ )♭ = u, (v♭ )♯ = v, and
v♭ = ψ∗ (v) ◦ ρ is immediate, and proves our assertion.
We check that, if w : G1 → G2 is a homomorphism of sheaves of sets on Y, ψ∗ (w) is expressed in
the following manner: if s′ = (s′x ) x∈U is a section of ψ∗ (G1 ) over an open set U of X, then (ψ∗ (w))(s′ )
is the family (wψ( x) (s′x )) x∈U . Finally, it is immediate that for each open set V of Y, the inverse image
of G |V under the restriction of ψ to ψ−1 (V ) is identical to the induced sheaf ψ∗ (G )|ψ−1 (V ).
When ψ is the identity 1X , we recover the definition of a sheaf of sets associated to a presheaf (G,
II, 1.2). The above considerations apply without change when C is the category of groups or of rings
(not necessarily commutative).
When X is any subset of a topological space Y, and j the canonical injection X → Y, for each
sheaf G on Y with values in a category C, we call the induced sheaf of X by G the inverse image j∗ (G )
(whenever it exists); for the sheaves of sets (or of groups, or of rings) we recover the usual definition
(G, II, 1.5).
(3.7.2). Keeping the notation and hypotheses of (3.5.3), suppose that G is a sheaf of groups (resp.
of rings) on Y. The definition of sections of ψ∗ (G ) (3.7.1) shows (taking into account (3.4.4)) that
the homomorphism of stalks ψx ◦ ρψ( x) : Gψ( x) → (ψ∗ (G )) x is a functorial isomorphism in G , that
identifies the two stalks; with this identification, u♯x is identical to the homomorphism defined in
(3.5.1), and in particular, we have Supp(ψ∗ (G )) = ψ−1 (Supp(G )).
An immediate consequence of this result is that the functor ψ∗ (G ) is exact in G on the abelian
category of sheaves of abelian groups.
3.8. Sheaves on pseudo-discrete spaces 0I | 35
(3.8.1). Let X be a topological space whose topology admits a basis B consisting of open quasi-
compact subsets. Let F be a sheaf of sets on X; if we equip each of the F (U ) with the discrete topology,
U 7→ F (U ) is a presheaf of topological spaces. We will see that there exists a sheaf of topological spaces
F ′ associated to F (3.5.6) such that Γ(U, F ′ ) is the discrete space F (U ) for each open quasi-compact
subsets U. It will suffice to show that the presheaf U 7→ F (U ) of discrete topological spaces on B
satisfy the condition (F0 ) of (3.2.2), and more generally that if U is an open quasi-compact subset and
if (Uα ) is a cover of U by sets of B, then the least fine topology T on Γ(U, F ) renders continuous
the maps Γ(U, F ) → Γ(Uα , F ) is the discrete topology. There exists a finite number of indices αi
such that U = i Uαi . Let s ∈ Γ(U, F ) and let si be its image in Γ(Uαi , F ); the intersection of the
S
inverse images of the sets {si } is by definition a neighbourhood of s for T ; but since F is a sheaf of
sets and the Uαi cover U, this intersection is reduced to s, hence our assertion.
We note that if U is an open non quasi-compact subset of X, the topological space Γ(U, F ′ ) still
has Γ(U, F ) as the underlying set, but the topology is not discrete in general: it is the least fine
rendering commutative the maps Γ(U, F ) → Γ(V, F ), for V ∈ B and V ⊂ U (the Γ(V, F ) being
discrete).
The above considerations apply without modification to sheaves of groups or of rings (not
necessarily commutative), and associate to them sheaves of topological groups or topological rings,
respectively. To summarize, we say that the sheaf F ′ is the pseudo-discrete sheaf of spaces (resp.
groups, rings) associated to a sheaf of sets (resp. groups, rings) F .
(3.8.2). Let F , G be two sheaves of sets (resp. groups, rings) on X, u : F → G a homomorphism.
Then u is thus a continuous homomorphism F ′ → G ′ , if we denote by F ′ and G ′ the pseudo-discrete
sheaves associated to F and G ; this follows in effect from (3.2.5).
(3.8.3). Let F be a sheaf of sets, H a subsheaf of F , F ′ and H ′ the pseudo-discrete sheaves
associated to F and H respectively. Then, for each open U ⊂ X, Γ(U, H ′ ) is closed in Γ(U, F ′ ):
indeed, it is the intersection of the inverse images of the Γ(V, H ) (for V ∈ B, V ⊂ U) under the
continuous maps Γ(U, F ) → Γ(V, F ), and Γ(V, H ) is closed in the discrete space Γ(V, F ).
(4.1.1). A ringed space (resp. topologically ringed space) is a pair ( X, A ) consisting of a topological
space X and a sheaf of (not necessarily commutative) rings (resp. of a sheaf of topological rings)
A on X; we say that X is the underlying topological space of the ringed space ( X, A ), and A the
structure sheaf. The latter is denoted OX , and its stalk at a point x ∈ X is denoted OX,x or simply Ox
when there is no chance of confusion.
We denote by 1 or e the unit section of OX over X (the unit element of Γ( X, OX )).
As in this treatise we will have to consider in particular sheaves of commutative rings, it will
be understood, when we speak of a ringed space ( X, A ) without specification, that A is a sheaf of
commutative rings.
The ringed spaces with not-necessarily-commutative structure sheaves (resp. the topologically
ringed spaces) form a category, where we define a morphism ( X, A ) → (Y, B ) as a couple (ψ, θ ) = Ψ
consisting of a continuous map ψ : X → Y and a ψ-morphism θ : B → A (3.5.1) of sheaves of ErrII
rings (resp. of sheaves of topological rings); the composition of a second morphism Ψ′ = (ψ′ , θ ′ ) :
(Y, B ) → ( Z, C ) and of Ψ, denoted Ψ′′ = Ψ′ ◦ Ψ, is the morphism (ψ′′ , θ ′′ ) where ψ′′ = ψ′ ◦ ψ, and
θ ′′ is the composition of θ and θ ′ (equal to ψ∗′ (θ ) ◦ θ ′ , cf. (3.5.2)). For ringed spaces, remember that
♯ ♯ ♯
we then have θ ′′ = θ ♯ ◦ ψ∗ (θ ′ ) (3.5.5); therefore if θ ′ and θ ♯ are injective (resp. surjective), then the
♯
same is true of θ ′′ , taking into account that ψx ◦ ρψ( x) is an isomorphism for all x ∈ X (3.7.2). We
verify immediately, thanks to the above, that when ψ is an injective continuous map and when θ ♯ is a
surjective homomophism of sheaves of rings, the morphism (ψ, θ ) is a monomorphism (T, 1.1) in the
category of ringed spaces.
By abuse of language, we will often replace ψ by Ψ in notation, for example in writing Ψ−1 (U )
in place of ψ−1 (U ) for a subset U of Y, when the is no risk of confusion.
(4.1.2). For each subset M of X, the pair ( M, A | M ) is evidently a ringed space, said to be induced on
M by the ringed space ( X, A ) (and is still called the restriction of ( X, A ) to M). If j is the canonical
injection M → X and ω is the identity map of A | M, ( j, ω ♭ ) is a monomorphism ( M, A | M ) →
( X, A ) of ringed spaces, called the canonical injection. The composition of a morphism Ψ : ( X, A ) →
(Y, B ) and this injection is called the restriction of Ψ to M.
(4.1.3). We will not revisit the definitions of A -modules or algebraic sheaves on a ringed space ( X, A )
(G, II, 2.2); when A is a sheaf of not necessarily commutative rings, by A -module we will always
mean “left A -module” unless expressly stated otherwise. The A -submodules of A will be called
sheaves of ideals (left, right, or two-sided) in A or A -ideals.
When A is a sheaf of commutative rings, and in the definition of A -modules, we replace
everywhere the module structure by that of an algebra, we obtain the definition of an A -algebra on X.
It is the same to say that an A -algebra (not necessarily commutative) is a A -module C , given with a
homomorphism of A -modules ϕ : C ⊗A C → C and a section e over X, such that: 1st the diagram
ϕ ⊗1
C ⊗A C ⊗A C / C ⊗A C
1⊗ ϕ ϕ
/C
ϕ
C ⊗A C
is commutative; 2nd for each open U ⊂ X and each section s ∈ Γ(U, C ), we have ϕ((e|U ) ⊗ s) =
ϕ(s ⊗ (e|U )) = s. We say that C is a commutative A -algebra if the diagram
C ⊗A C
σ / C ⊗A C
ϕ ϕ
# {
C
is commutative, σ denoting the canonical symmetry (twist) map of the tensor product C ⊗A C .
The homomorphisms of A -algebras are also defined as the homomorphisms of A -modules in
(G, II, 2.2), but naturally no longer form an abelian group.
If M is an A -submodule of an A -algebra C , the A -subalgebra of C generated by M is the sum of
the images of the homomorphisms n M → C (for each n ⩾ 0). This is also the sheaf associated
N
4. RINGED SPACES 34
We can first consider the topological space obtained by gluing (by means of the ϕλµ ) of the Xλ along 0I | 39
the Vλµ ; if we identify Xλ with the corresponding open subset Xλ′ in X, the hypotheses imply that
the three sets Vλµ ∩ Vλν , Vµν ∩ Vµλ , Vνλ ∩ Vνµ identify with Xλ′ ∩ Xµ′ ∩ Xν′ . We can also transport to
Xλ′ the ringed space structure of Xλ , and if Aλ′ are the transported sheaves of rings corresponding to
the Aλ , the Aλ′ satisfy the gluing condition (3.3.1) and therefore define a sheaf of rings A on X; we
say that ( X, A ) is the ringed space obtained by gluing the ( Xλ , Aλ ) along the Vλµ , by means of the
ϕλµ .
(4.2.2.2) Ψ ∗ (M ) ⊗B Ψ ∗ (N ) ⊗B Ψ ∗ (P ) / Ψ ∗ (M ⊗A N ) ⊗B Ψ ∗ (P )
Ψ ∗ (M ) ⊗B Ψ ∗ (N ⊗A P ) / Ψ ∗ (M ⊗A N ⊗A P )
is commutative.
(4.2.3). Let M , N be two A -modules. For each open U ⊂ Y, we have by definition that Γ(ψ−1 (U ), H omA (M , N )) =
HomA |V (M |V, N |V ), where we put V = ψ−1 (U ); the map u 7→ Ψ∗ (u) is a homomorphism
HomA |V (M |V, N |V ) −→ HomB |U (Ψ∗ (M )|U, Ψ∗ (N )|U )
on the Γ(U, B )-module structures; these homomorphisms are compatible with the restriction opera-
tions, hence they define a canonical functorial homomorphism of B-modules
(4.2.3.1) Ψ∗ (H omA (M , N )) −→ H omB (Ψ∗ (M ), Ψ∗ (N )).
4. RINGED SPACES 36
Ψ∗ Gi = Ψ ∗ ( Gi ) .
M M
(4.3.2.1)
i i
By passing to the inductive limit, we deduce, in light of the above, that the above equality is still
true for any direct sum.
(4.3.3). Let G1 , G2 be two B-modules; from the definition of the inverse images of sheaves of
abelian groups (3.7.1), we obtain immediately a canonical homomorphism ψ∗ (G1 ) ⊗ψ∗ (B ) ψ∗ (G2 ) →
ψ∗ (G1 ⊗B G2 ) of ψ∗ (B )-modules, and the stalk at a point of a tensor product of sheaves being
the tensor product of the stalks at this point (G, II, 2.8), we deduce from (3.7.2) that the above
homomorphism is in fact a isomorphism. By tensoring with A , we obtain a canonical functorial
isomorphism
(4.3.3.1) Ψ∗ (G1 ) ⊗A Ψ∗ (G2 ) ≃ Ψ∗ (G1 ⊗B G2 ).
(4.3.4). Let C be a B-algebra; the data of the algebra structure on C is the same as the data of a B-
homomorphism C ⊗B C → C satisfying the associativity and commutativity conditions (conditions
which are checked stalk-wise); the above isomorphism allows us to consider this homomorphism
4. RINGED SPACES 37
these homomorphisms will be called canonical. They are in general neither injective or surjective. We
have canonical factorizations analogous to (3.5.3.3) and (3.5.4.4).
We note that if s is a section of G over an open set V of Y, ρG (s) is the section s′ ⊗ 1 of Ψ∗ )(G )
over ψ−1 (V ), s′ being such that s′x = sψ( x) for all x ∈ ψ−1 (V ). We also note that if u : G → ψ∗ (F ) is
a homomorphism, it defines for all x ∈ X a homomorphism u x : Gψ( x) → Fx on the stalks, obtained
by composing (u♯ ) x : (Ψ∗ (G )) x → Fx and the canonical homomorphism s x 7→ s x ⊗ 1 from Gψ( x) to
(Ψ∗ (G )) x = Gψ(x) ⊗Bψ(x) Ax . The homomorphism u x is obtained also by passing to the inductive
u
→ Γ(ψ−1 (V ), F ) → Fx , where V varies over the
limit relative to the homomorphisms Γ(V, G ) −
neighbourhoods of ψ( x ).
(4.4.5). Let (Gλ )λ∈ L be an inductive system of B-modules, and, for each λ ∈ L, let uλ be a homomor-
phism Gλ → Ψ∗ (F ), form an inductive limit; we put G = lim Gλ and u = lim uλ ; then the (uλ )♯
−→ −→
form an inductive system of homomorphisms Ψ∗ (Gλ ) → F , and the inductive limit of this system
is none other than u♯ .
(4.4.6). Let M , N be two B-modules, V an open set of Y, U = ψ−1 (V ); the map v 7→ Ψ∗ (v) is a
homomorphism
for the Γ(V, B )-module structures (HomA |U (Ψ∗ (M )|U, Ψ∗ (N )|U ) is normaly equipped with the
a Γ(U, ψ∗ (B ))-module structure, and thanks to the canonical homomorphism (3.7.2) Γ(V, B ) → 0I | 44
Γ(U, ψ∗ (B )), it is also a Γ(V, B )-module). We see immediately that these homomorphisms are com-
patible with the restriction morphisms, and as a result define a canonical functorial homomorphism
G ⊗B G /G
u
Ψ ∗ (F ⊗A F ) / Ψ ∗ (F )
(4.4.8). Let ( Z, C ) be a third ringed space, Ψ′ = (ψ′ , θ ′ ) a morphism (Y, B ) → ( Z, C ), and Ψ′′ :
( X, A ) → ( Z, C ) the composite morphism Ψ′ ◦ Ψ. Let H be a C -module, u′ a homomorphism from
H to G ; the composition v′′ = v ◦ v′ is by definition the homomorphism from H to F defined by
v′ Ψ′∗ (v) ♯
→ Ψ′∗ (G ) −−−→ Ψ′∗ (Ψ∗ (F )); we check that v′′ is the homomorphism
H −
♯
∗ Ψ∗ (v′ ) v♯
Ψ∗ (Ψ′ (H )) −−−−→ Ψ∗ (G ) −
→ F.
5. QUASI-COHERENT AND COHERENT SHEAVES 39
(5.1.3). We say that an OX -module F is quasi-coherent if for each x ∈ X there is an open neigh-
bourhood U of x such that F |U is isomorphic to the cokernel of a homomorphism of the form
(I) ( J)
OX |U → OX |U, where I and J are sets of arbitrary indices. It is clear that OX is itself a quasi-
coherent OX -module, and that any direct sum of quasi-coherent OX -modules is again a quasi-
coherent OX -module. We say that an OX -algebra A is quasi-coherent if it is quasi-coherent as an
OX -module.
(5.1.4). Let f : X → Y be a morphism of ringed spaces. If G is a quasi-coherent OY -module, then
f ∗ (G ) is a quasi-coherent OX -module. Indeed, for each x ∈ X, there is an open neighbourhood V
(I) ( J)
of f ( x ) in Y such that G |V is the cokernel of a homomorphism OY |V → OY |V. If U = f −1 (V ),
and if f U is the restriction of f to U, then we have f ∗ (G )|U = f U∗ (G |V ); as f U∗ is right exact and
(I) ( J)
commutes with direct sums, f U∗ (G |V ) is the cokernel of a homomorphism OX |U → OX |U.
neighbourhood V ⊂ U of x such that the (si )y generate Fy for all y ∈ Y (FAC, I, 2, 12, prop. 1). In
particular, we conclude that the support of F is closed.
Similarly, if u : F → G is a homomorphism such that u x = 0, then there exists a neighbourhood 0I | 46
U of x such that uy = 0 for all y ∈ U.
(5.2.3). Suppose that X is quasi-compact, and let F and G be two OX -modules such that G is of finite
type, u : F → G a surjective homomorphism. In addition, suppose that F is the inductive limit of an
inductive system (Fλ ) of OX -modules. Then there exists an index µ such that the homomorphism
Fµ → G is surjective. Indeed, for each x ∈ X, there exists a finite system of sections si of G over an
open neighbourhood U ( x ) of x such that the (si )y generate Gy for all y ∈ U ( x ); there is then an open
neighbourhood V ( x ) ⊂ U ( x ) of x and n sections ti of F over V ( x ) such that si |V (s) = u(ti ) for all
i; we can also suppose that the ti are the canonical images of sections of a similar sheaf Fλ( x) over
V ( x ). We then cover X with a finite number of neighbourhoods V ( xk ), and let µ be the maximal
index of the λ( xk ); it is clear that this index gives the answer.
Suppose still that X is quasi-compact, and let F be an OX -module of finite type generated by
its sections over X (5.1.1); then F is generated by a finite subfamily of these sections: indeed, it
suffices to cover X by a finite number of open neighbourhoods Uk such that, for each k, there is a
finite number of sections sik of F over X whose restrictions to Uk generate F |Uk ; it is clear that the
sik then generate F .
(5.2.4). Let f : X → Y be a morphism of ringed spaces. If G is an OY -module of finite type, then
f ∗ (G ) is an OX -module of finite type. Indeed, for each x ∈ X, there is an open neighbourhood
p
V of f ( x ) in Y and a surjective homomorphism v : OY |V → G |V. If U = f −1 (V ) and if f U is the
restriction of f to U, then we have f (G )|U = f U (G |V ); since f U∗ is right exact (4.3.1) and commutes
∗ ∗ ErrII
p
with direct sums (4.3.2), f U∗ (v) is a surjective homomorphism OX |U → f ∗ (G )|U.
(5.2.5). We say that an OX -module F admits a finite presentation if for each x ∈ X there exists an
open neighbourhood U of x such that F |U is isomorphic to a cokernel of a (OX |U )-homomorphism
p q
OX |U → OX |U, p and q being two integers > 0. Such an OX -module is therefore of finite type and
quasi-coherent. If f : X → Y is a morphism of ringed spaces, and if G is an OY -module admitting a
finite presentation, then f ∗ (G ) admits a finite presentation, as shown in the argument of (5.1.4).
(5.2.6). Let F be an OX -module admitting a finite presentation (5.2.5); then, for each OX -module H ,
the canonical functorial homomorphism
(H omOX (F , H )) x −→ HomOx (Fx , Hx )
is bijective (T, 4.1.1).
(5.2.7). Let F and G be two OX -modules admitting a finite presentation. If for some x ∈ X, Fx and
Gx are isomorphic as Ox -modules, then there exists an open neighbourhood U of x such that F |U and
G |U are isomorphic. Indeed, if ϕ : Fx → Gx and ψ : Gx → Fx are an isomorphism and its inverse
isomorphism, then there exists, according to (5.2.6), an open neighbourhood V of x and a section u
(resp. v) of H omOX (F , G ) (resp. H omOX (G , F )) over V such that u x = ϕ (resp. v x = ψ). As (u ◦ v) x 0I | 47
and (v ◦ u) x are the identity automorphisms, there exists an open neighbourhood U ⊂ V of x such
that (u ◦ v)|U and (v ◦ u)|U are the identity automorphisms, hence the proposition.
5.3. Coherent sheaves
(5.3.1). We say that an OX -module F is coherent if it satisfies the two following conditions:
(a) F is of finite type.
(b) for each open U ⊂ X, integer n > 0, and homomorphism u : OXn |U → F |U, the kernel of
u is of finite type.
We note that these two conditions are of a local nature.
For most of the proofs of the properties of coherent sheaves in what follows, cf. (FAC, I, 2).
(5.3.2). Each coherent OX -module admits a finite presentation (5.2.5); the inverse is not necessarily
true, since OX itself is not necessarily a coherent OX -module.
Each OX -submodule of finite type of a coherent OX -module is coherent; each finite direct sum of
coherent OX -modules is a coherent OX -module.
5. QUASI-COHERENT AND COHERENT SHEAVES 41
(5.3.5). If F and G are two coherent OX -modules, then so are F ⊗OX G are H omOX (F , G ).
(5.3.6). Let F be a coherent OX -module, J a coherent sheaf of ideals of OX . Then the OX -module
J F is coherent, as the image of J ⊗OX F under the canonical homomorphism J ⊗OX F → F
((5.3.4) and (5.3.5)).
(5.3.7). We say that an OX -algebra A is coherent if it is coherent as an OX -module. In particular, OX
is a coherent sheaf of rings if and only if for each open U ⊂ X and each homomorphism of the form
p
u : OX |U → OX |U, the kernel of u is an (OX |U )-module of finite type.
If OX is a coherent sheaf of rings, then each OX -module F admitting a finite presentation (5.2.5)
is coherent, according to (5.3.4).
The annihilator of an OX -module F is the kernel J of the canonical homomorphism OX →
H omOX (F , F ) which sends each section s ∈ Γ(U, OX ) to the multiplication by s map in Hom(F |U, F |U );
if OX is coherent and if F is a coherent OX -module, then J is coherent ((5.3.4) and (5.3.5)) and for
each x ∈ X, Jx is the annihilator of Fx (5.2.6).
(5.3.8). Suppose that OX is coherent; let F be a coherent OX -module, x a point of X, M a submodule of 0I | 48
finite type of Fx ; then there exists an open neighbourhood U of x and a coherent (OX |U )-submodule
G of F |U such that Gx = M (T, 4.1, Lemma 1).
This result, along with the properties of the OX -submodules of a coherent OX -module, impose
the necessary conditions on the rings Ox such that OX is coherent. For example (5.3.4), the intersection
of two ideals of finite type of Ox must still be an ideal of finite type.
(5.3.9). Suppose that OX is coherent, and let M be an Ox -module admitting a finite presentation,
p q
therefore isomorphic to a cokernel of a homomorphism ϕ : Ox → Ox ; then there exists an open
neighbourhood U of X and a coherent (OX |U )-module F such that Fx is isomorphic to M. Indeed,
p q
according to (5.2.6), there exists a section u of H omOX (OX , OX ) over an open neighbourhood U of x
p q
such that u x = ϕ; the cokernel F of the homomorphism u : OX |U → OX |U gives the answer (5.3.4).
(5.3.10). Suppose that OX is coherent, and let J be a coherent sheaf of ideals of OX . For a (OX /J )-
module F to be coherent, it is necessary and sufficient for it to be coherent as an OX -module. In
particular, OX /J is a coherent sheaf of rings.
(5.3.11). Let f : X → Y be a morphism of ringed spaces, and suppose that OX is coherent; then, for
each coherent OY -module G , f ∗ (G ) is a coherent OX -module. Indeed, with the notation of (5.2.4),
q p
we can assume that G |V is the cokernel of a homomorphism v : OY |V → OY |V; as f U∗ is right
q p
exact, f ∗ (G )|U = f U∗ (G |V ) is the cokernel of the homomorphism f U∗ (v) : OX |U → OX |U, hence our
assertion.
(5.3.12). Let Y be a closed subset of X, j : Y → X the canonical injection, OY a sheaf of rings on Y,
and set OX = j∗ (OY ). For an OY -module G to be of finite type (resp. quasi-coherent, coherent), it is
necessary and sufficient for j∗ (G ) to be an OX -module of finite type (resp. quasi-coherent, coherent).
5. QUASI-COHERENT AND COHERENT SHEAVES 42
(5.4.1). Let X be a ringed space. We say that an OX -module F is locally free if for each x ∈ X there
exists an open neighbourhood U of x such that F |U is isomorphic to a (OX |U )-module of the form
(I)
OX |U, where I can depend on U. If for each U, I is finite, then we say that F is of finite rank; if
for each U, I has the same finite number of elements n, we say that F is of rank n. A locally free
OX -module of rank 1 is called invertible (cf. (5.4.3)). If F is a locally free OX -module of finite rank,
then for each x ∈ X, Fx is a free Ox -module of finite rank n( x ), and there exists a neighbourhood U
of x such that F |U, is of rank n( x ); if X is connected, then n( x ) is constant.
It is clear that each locally free sheaf is quasi-coherent, and if OX is a coherent sheaf of rings,
then each locally free OX -module of finite rank is coherent.
If L is locally free, then L ⊗OX F is an exact functor in F to the category of OX -modules.
We will mostly consider locally free OX -modules of finite rank, and when we speak of locally 0I | 49
free sheaves without specifying, it will be understood that they are of finite rank.
Suppose that OX is coherent, and let F be a coherent OX -module. Then, if at a point x ∈ X, Fx is
an Ox -module free of rank n, there exists a neighbourhood U of x such that F |U is locally free of rank
n; in fact, Fx is then isomorphic to Oxn , and the proposition follows from (5.2.7). ErrII
(5.4.3). If L is invertible, then so is its dual L ∨ = H omOX (L , OX ), since we can immediately reduce
(as the question is local) to the case L = OX . In addition, we have a canonical isomorphism
(5.4.3.1) H omOX (L , OX ) ⊗OX L ≃ OX
as, according to (5.3.2), it suffices to define a canonical isomorphism H omOX (L , L ) ≃ OX . For each
OX -module F , we have a canonical homomorphism OX ≃ H omOX (F , F ) (5.3.7). It remains to
prove that if F = L is invertible, then this homomorphism is bijective, and as the question is local,
it reduces to the case L = OX , which is immediate.
Due to the above, we put L −1 = H omOX (L , OX ), and we say that L −1 is the inverse of L . The
terminology “invertible sheaf” can be justified in the following way when X is a point and OX is a
local ring A with maximal ideal m; if M and M′ are two A-modules (M being of finite type) such
that M ⊗ A M′ is isomorphic to A, then as ( A/m) ⊗ A ( M ⊗ A M′ ) identifies with ( M/mM ) ⊗ A/m
( M′ /mM′ ), this latter tensor product of vector spaces over the field A/m is isomorphic to A/m,
which requires M/mM and M′ /mM′ to be of dimension 1. For each element z ∈ M not in mM,
we have M = Az + mM, which implies that M = Az according to Nakayama’s Lemma, M being
of finite type. Moreover, as the annihilator of z kills M ⊗ A M′ , which is isomorphic to A, this
annihilator is {0}, and as a result M is isomorphic to A. In the general case, this shows that L is an
OX -module of finite type, such that there exists an OX -module F for which L ⊗OX F is isomorphic
to OX , and if in addition the rings Ox are local rings, then Lx is an Ox -module isomorphic to Ox
for each x ∈ X. If OX and L are assumed to be coherent, then we conclude that L is invertible
according to (5.2.7).
(5.4.4). If L and L ′ are two invertible OX -modules, then so is L ⊗OX L ′ , since the question is local,
we can assume that L = OX , and the result is then trivial. For each integer n ⩾ 1, we denote by
L ⊗n the tensor product of n copies of the sheaf L ; we set by convention L ⊗0 = OX , and for n ⩾ 1, 0I | 50
L ⊗(−n) = (L −1 )⊗n . With these notation, there is then a canonical functorial isomorphism
for any rational integers m and n: indeed, by definition, we immediately reduce to the case where
m = −1, n = 1, and the isomorphism in question is then that defined in (5.4.3).
(5.4.5). Let f : Y → X be a morphism of ringed spaces. If L is a locally free (resp. invertible)
OX -module, then f ∗ (L ) is a locally free (resp. invertible) OY -module: this follows immediately
from that the inverse images of two locally isomorphic OX -modules are locally isomorphic, that
f ∗ commutes with finite direct sums, and that f ∗ (OX ) = OY (4.3.4). In addition, we know that we
have a canonical functorial homomorphism f ∗ (L ∨ ) → ( f ∗ (L ))∨ (4.4.6), and when L is locally
free, this homomorphism is bijective: indeed, we again reduce to the case where L = OX which is
trivial. We conclude that if L is invertible, then f ∗ (L ⊗n ) canonically identifies with ( f ∗ (L ))⊗n for
each rational integer n.
(5.4.6). Let L be an invertible OX -module; we denote by Γ• ( X, L ) or simply Γ• (L ) the abelian
group direct sum n∈Z Γ( X, L ⊗n ); we equip it with the structure of a graded ring, by corresponding
L
n∈Z
We equip this abelian group with the structure of a graded module over the graded ring Γ• (L ) in the
following way: to a pair (sn , um ), where sn ∈ Γ( X, L ⊗n ) and um ∈ Γ( X, F ⊗OX L ⊗m ), we associate
the section of F ⊗OX L ⊗(m+n) which canonically corresponds (5.4.4.1) to sn ⊗ um ; the verification
of the module axioms are immediate. For X and L fixed, Γ• (L , F ) is a covariant functor in F with
values in the category of graded Γ• (L )-modules; for X and F fixed, it is a covariant functor in L
with values in the category of abelian groups.
If f : Y → X is a morphism of ringed spaces, the canonical homomorphism (4.4.3.2) ρ :
L ⊗n → f ∗ ( f ∗ (L ⊗n )) defines a homomorphism of abelian groups Γ( X, L ⊗n ) → Γ(Y, f ∗ (L ⊗n )),
and as f ∗ (L ⊗n ) = ( f ∗ (L ))⊗n ), it follows from the definitions of the canonical homomorphisms
(4.4.3.2) and (5.4.4.1) that the above homomorphisms define a functorial homomorphism of graded rings
Γ• (L ) → Γ• ( f ∗ (L )). The same canonical homomorphism (4.4.3) similarly defines a homomor-
phism of abelian groups Γ( X, F ⊗OX L ⊗n ) → Γ(Y, f ∗ (F ⊗OX L ⊗n )), and as
f ∗ (F ⊗OX L ⊗n ) = f ∗ (F ) ⊗OY ( f ∗ (L ))⊗n (4.3.3.1),
these homomorphism (for n variable) define a di-homomorphism of graded modules Γ• (L , F ) → 0I | 51
Γ• ( f ∗ (L ), f ∗ (F )).
(5.4.7). One can show that there exists a set M (also denoted M( X )) of invertible OX -modules
such that each invertible OX -module is isomorphic to a unique element of M 8; we define on M a
composition law by sending two elements L and L ′ of M to the unique element of M isomorphic
to L ⊗OX L ′ . With this composition law, M is a group isomorphic to the cohomology group H1 ( X, OX∗ ),
where OX∗ is the subsheaf of OX such that Γ(U, OX∗ ) is the group of invertible elements of the ring
Γ(U, OX ) for each open U ⊂ X (OX∗ is therefore a sheaf of multiplicative abelian groups).
We will note that for all open U ⊂ X, the group of sections Γ(U, OX∗ ) canonically identifies with
the automorphism group of the (OX |U )-module OX |U, the identification sending a section ε of OX∗
over U to the automorphism u of OX |U such that u x (s x ) = ε x s x for all x ∈ X and all s x ∈ Ox . Then
let U = (Uλ ) be an open cover of X; the data, for each pair of indices (λ, µ), of an automorphism θλµ
of OX |(Uλ ∩ Uµ ) is the same as giving a 1-cochain of the cover U, with values in OX∗ , and say that the
θλµ satisfy the gluing condition (3.3.1), meaning that the corresponding cochain is a cocycle. Similarly,
the data, for each λ, of an automorphism ωλ of OX |Uλ is the same as the data of a 0-cochain of
the cover U, with values in OX∗ , and its coboundary corresponds to the family of automorphisms
(ωλ |Uλ ∩ Uµ ) ◦ (ωµ |Uλ ∩ Uµ )−1 . We can send each 1-cocycle of U with values in OX∗ to the element
of M isomorphic to an invertible OX -module obtained by gluing with respect to the family of
8See the book in preparation cited in the introduction.
5. QUASI-COHERENT AND COHERENT SHEAVES 44
automorphisms (θλµ ) corresponding to this cocycle, and to two cohomologous coycles correspond
two equal elements of M (3.3.2); in other words, we thus define a map ϕU : H1 (U, OX∗ ) → M. In
addition, if B is a second open cover of X, finer than U, then the diagram
H1 (U, OX∗ )
ϕU
$
:M
ϕB
H1 (B, OX∗ )
where the vertical arrow is the canonical homomorphism (G, II, 5.7), is commutative, as a result
of (3.3.3). By passing to the inductive limit, we therefore obtain a map H1 ( X, OX∗ ) → M, the Čech
cohomology group Ȟ1 ( X, OX∗ ) identifying as we know with the first cohomology group H1 ( X, OX∗ )
(G, II, 5.9, Cor. of Thm. 5.9.1). This map is surjective: indeed, by definition, for each invertible
OX -module L , there is an open cover (Uλ ) of X such that L is obtained by gluing the sheaves
OX |Uλ (3.3.1). It is also injective, since it suffices to prove for the maps H1 (U, OX ) → M, and this
follows from (3.3.2). It remains to show that the bijection thus defined is a group homomorphism. 0I | 52
Given two invertible OX -modules L and L ′ , there is an open cover (Uλ ) such that L |Uλ and
L ′ |Uλ are isomorphic to OX |Uλ for each λ; so there is for each index λ an element aλ (resp. a′λ )
of Γ(Uλ , L ) (resp. Γ(Uλ , L ′ )) such that the elements of Γ(Uλ , L ) (resp. Γ(Uλ , L ′ )) are the sλ · aλ
(resp. sλ · a′λ ), where sλ varies over Γ(Uλ , OX ). The corresponding cocycles (ε λµ ), (ε′λµ ) are such that
sλ · aλ = sµ · aµ (resp. sλ · a′λ = sµ · a′µ ) over Uλ ∩ Uµ is equivalent to sλ = ε λµ sµ (resp. sλ = ε′λµ sµ )
over Uλ ∩ Uµ . As the sections of L ⊗OX L ′ over Uλ are the finite sums of the sλ s′λ · ( aλ ⊗ a′λ ) where
sλ and s′λ vary over Γ(Uλ , OX ), it is clear that the cocycle (ε λµ , ε′λµ ) corresponds to L ⊗OX L ′ , which
9
finishes the proof.
(5.4.8). Let f = (ψ, ω ) be a morphism Y → X of ringed spaces. The functor f ∗ (L ) to the category of
free OX -modules defines a map (which we still denote f ∗ by abuse of language) from the set M( X )
to the set M(Y ). Second, we have a canonical homomorphism (T, 3.2.2)
(5.4.8.1) H1 ( X, OX∗ ) −→ H1 (Y, OY∗ ).
When we canonically identify (5.4.7) M( X ) and H1 ( X, OX∗ ) (resp. M(Y ) and H1 (Y, OY∗ )), the homo-
morphism (5.4.8.1) identifies with the map f ∗ . Indeed, if L comes from a cocycle (ε λµ ) corresponding
to an open cover (Uλ ) of X, then it suffices to show that f ∗ (L ) comes from a cocycle whose coho-
mology class is the image under (5.4.8.1) of (ε λµ ). If θλµ is the automorphism of OX |(Uλ ∩ Uµ ) which
corresponds to ε λµ , then it is clear that f ∗ (L ) is obtained by gluing the OY |ψ−1 (Uλ ) by means of the
automorphisms f ∗ (θλµ ), and it then suffices to check that these latter automorphisms corresponds
to the cocycle (ω ♯ (ε λµ )), which follows immediately from the definitions (we can identify ε λµ with
its canonical image under ρ (3.7.2), a section of ψ∗ (OX∗ ) over ψ−1 (Uλ ∩ Uµ )).
(5.4.9). Let E and F be two OX -modules, F assumed to be locally free, and let G be an OX -module
i p
extension of F by E , in other words there exists an exact sequence 0 → E − →G − → F → 0. Then, for
each x ∈ X, there exists an open neighbourhood U of x such that G |U is isomorphic to the direct sum
E |U ⊕ F |U. We can reduce to the case where F = OXn ; let ei (1 ⩽ i ⩽ n) be the canonical sections
(5.5.5) of OXn ; there then exists an open neighbourhood U of x and n sections si of G over U such that
p(si |U ) = ei |U for 1 ⩽ i ⩽ n. That being so, let f be the homomorphism F |U → G |U defined by
the sections si |U (5.1.1). It is immediate that for each open V ⊂ U, and each section s ∈ Γ(V, G ) we
have s − f ( p(s)) ∈ Γ(V, E ), hence our assertion.
(5.4.10). Let f : X → Y be a morphism of ringed spaces, F an OX -module, and L a locally free
OY -module of finite rank. Then there exists a canonical isomorphism
(5.4.10.1) f ∗ (F ) ⊗OY L ≃ f ∗ (F ⊗OX f ∗ (L )).
9For a general form of this result, see the book cited in the note on p. 51.
5. QUASI-COHERENT AND COHERENT SHEAVES 45
ρ the homomorphism (4.4.3.2) and α the homomorphism (4.2.2.1). To show that when L is locally
free, this homomorphism is bijective, it suffices, since the questions is local, to consider the case
where L = OXn ; in addition, f ∗ and f ∗ commute with finite direct sums, so we can assume n = 1,
and in this case the proposition follows immediately from the definitions and from the relation
f ∗ (OY ) = OX .
5.5. Sheaves on a locally ringed space
(5.5.1). We say that a ringed space ( X, OX ) is a locally ringed space if for each x ∈ X, Ox is a local
ring; these ringed spaces will be by far the most frequent ringed spaces that we will consider in
this work. We then denote by mx the maximal ideal of Ox , by k( x ) the residue field Ox /mx ; for each
OX -module F , each open set U of X, each point x ∈ U, and each section f ∈ Γ(U, F ), we denote by
f ( x ) the class of the germ f x ∈ Fx mod. mx Fx , and we say that this is the value of f at the point x.
The relation f ( x ) = 0 then means that f x ∈ mx Fx ; when this is so, we say (by abuse of language)
that f is zero at x. We will take care not to confuse this relation with f x = 0.
(5.5.2). Let X be a locally ringed space, L an invertible OX -module, and f a section of L over X.
There is then an equivalence between the three following properties for a point x ∈ X:
(a) f x is a generator of Lx ;
(b) f x ̸∈ mx Lx (in other words, f ( x ) ̸= 0);
(c) there exists a section g of L −1 over an open neighbourhood V of x such that the canonical image of
f ⊗ g in Γ(V, OX ) (5.4.3) is the unit section.
Indeed, since the questions is local, we can reduce to the case where L = OX ; the equivalence
of (a) and (b) are then evident, and it is clear that (c) implies (b). Conversely, if f x ̸∈ mx , then f x
is invertible in Ox , say f x gx = 1x . By definition of germs of sections, this means that there exists a
neighbourhood V of x and a section g of OX over V such that f g = 1 in V, hence (c).
It follows immediately from the condition (c) that the set X f of x satisfying the equivalent
conditions (a), (b), (c) is open in X; following the terminology introduced in (5.5.1), this is the set of
the x for which f does not vanish.
(5.5.3). Under the hypotheses of (5.5.2), let L ′ be a second invertible OX -module; then, if f ∈
Γ( X, L ), g ∈ Γ( X, L ′ ), we have
X f ∩ Xg = X f ⊗g .
We can in fact reduce immediately to the case where L = L ′ = OX (since the questions is
local); as f ⊗ g then canonically identifies with the product f g, the proposition is evident.
0I | 54
(5.5.4). Let F be a locally free OX of rank n; it is immediate that ∧ p F is a locally free OX -module
of rank (np) if p ⩽ n and 0 if p > n, since the question is local and we can reduce to the case where
F = OXn ; in addition, for each x ∈ X, (∧ p F ) x /mx (∧ p F ) x is a vector space of dimension (np) over
k( x ), which canonically identifies with ∧ p (Fx /mx Fx ). Let s1 , . . . , s p be the sections of F over an
open subset U of X, and let s = s1 ∧ · · · ∧ s p , which is a section of ∧ p F over U (4.1.5); we have
s( x ) = s1 ( x ) ∧ · · · ∧ s p ( x ), and as a result, we say that the s1 ( x ), . . . , s p ( x ) are linearly dependent
means s( x ) = 0. We conclude that the set of the x ∈ X such that s1 ( x ), . . . , s p ( x ) are linearly independent
is open in X: it suffices in fact, by reducing to the case where F = OXn , to apply (5.5.2) to the section
(np)
image of s under one of the projections of ∧ p F = OX to the (np) factors.
In particular, if s1 , . . . , sn are n sections of F over U such that s1 ( x ), . . . , sn ( x ) are linearly
independent for each point x ∈ U, then the homomorphism u : OXn |U → F |U defined by the si
(5.1.1) is an isomorphism: indeed, we can restrict to the case where F = OXn and where we canonically
identify ∧n F and OX ; s = s1 ∧ · · · ∧ sn is then an invertible section of OX over U, and we define an
inverse homomorphism for u by means of the Cramer formulas.
6. FLATNESS 46
(5.5.5). Let E and F be two locally free OX -modules (of finite rank), and let u : E → F be a
homomorphism. For there to exist a neighbourhood U of x ∈ X such that u|U is injective and
that F |U is the direct sum of the u(E )|U and of a locally free (OX |U )-submodule G , it is necessary and
sufficient that u x : Ex → Fx gives, by passing to quotients, an injective homomorphism of vector
spaces Ex /mx Ex → Fx /mx Fx . The condition is indeed necessary, since Fx is then the direct sum
of the free Ox -modules u x (Ex ) and Gx , so Fx /mx Fx is the direct sum of u x (Ex )/mx u x (Ex ) and of
Gx /mx Gx . The condition is sufficient, since we can reduce to the case where E = OXm ; let s1 , . . . , sm
be the images under u of the sections ei of OXm such that (ei )y is equal to the i-th element of the
canonical basis of Oym for each y ∈ Y (canonical sections of OXm ); by hypothesis, the s1 ( x ), . . . , sm ( x )
are linearly independent, so if F is of rank n, then there exist n − m sections sm+1 , . . . , sn of F over
a neighbourhood V of x such that the si ( x ) (1 ⩽ i ⩽ n) form a basis for Fx /mx Fx . There then exists
(5.5.4) a neighbourhood U ⊂ V of x such that the si (y) (1 ⩽ i ⩽ n) form a basis for Fy /my Fy for
each y ∈ V, and we conclude (5.5.4) that there is an isomorphism from F |U to OXn |U, sending the
si |U (1 ⩽ i ⩽ m) to the ei |U, which finishes the proof.
§6. F LATNESS
(6.0). The notion of flatness is due to J.-P. Serre [Ser56]; in the following, we omit the proofs of the
results which are presented in the Algèbre commutative of N. Bourbaki, to which we refer the reader.
We assume that all rings are commutative.10
If M, N are two A-modules, M′ (resp. N ′ ) a submodule of M (resp. N), we denote by Im( M′ ⊗ A 0I | 55
N ′ ) the submodule of M ⊗ A N, the image under the canonical map M′ ⊗ A N ′ → M ⊗ A N.
6.1. Flat modules
(6.1.1). Let M be an A-module. The following conditions are equivalent:
(a) The functor M ⊗ A N is exact in N on the category of A-modules;
(b) ToriA ( M, N ) = 0 for each i > 0 and for each A-module N;
(c) Tor1A ( M, N ) = 0 for each A-module N.
When M satisfies these conditions, we say that M is a flat A-module. It is clear that each free
A-module is flat.
For M to be a flat A-module, it suffices that for each ideal J of A, of finite type, the canonical map
M ⊗ A J → M ⊗ A A = M is injective.
L
(6.1.2). Each inductive limit of flat A-modules is a flat A-module. For a direct sum λ∈ L Mλ of
A-modules to be a flat A-modules, it is necessary and sufficient that each of the A-modules Mλ is
flat. In particular, every projective A-module is flat.
Let 0 → M′ → M → M′′ → 0 be an exact sequence of A-modules, such that M′′ is flat. Then,
for each A-module N, the sequence
0 −→ M′ ⊗ A N −→ M ⊗ A N −→ M′′ ⊗ A N −→ 0
is exact. In addition, for M to be flat, is it necessary and sufficient that M′ is (but it can be that M
and M′ are flat without M′′ = M/M′ being so).
(6.1.3). Let M be a flat A-module, N any A-module; for two submodules N ′ N ′′ of N, we then have
Im( M ⊗ A ( N ′ + N ′′ )) = Im( M ⊗ A N ′ ) + Im( M ⊗ A N ′′ ),
Im( M ⊗ A ( N ′ ∩ N ′′ )) = Im( M ⊗ A N ′ ) ∩ Im( M ⊗ A N ′′ )
(images taken in M ⊗ A N).
(6.1.4). Let M and N be two A-modules, M′ (resp. N ′ ) a submodule of M (resp. N), and suppose that
one of the modules M/M′ , N/N ′ is flat. Then we have Im( M′ ⊗ A N ′ ) = Im( M′ ⊗ A N ) ∩ ( M ⊗ A N ′ )
(images in M ⊗ A N). In particular, if J is an ideal of A and if M/M′ is flat, then we have JM′ =
M′ ∩ JM.
10See the exposé cited of N. Bourbaki for the generalization from most of the results to the noncommutative case.
6. FLATNESS 47
6.2. Change of ring When an additive group M is equipped with multiple modules structures
relative to the rings A, B, ..., we say that M is flat as an A-module, B-module, ..., we sometimes also
say that M is A-flat, B-flat, ....
(6.2.1). Let A and B be two rings, M an A-module, N an ( A, B)-bimodule. If M is flat and if N is
B-flat, then M ⊗ A N is B-flat. In particular, if M and N are two flat A-modules, then M ⊗ A N is a
flat A-module. If B is an A-algebra and if M is a flat A-module, then the B-module M( B) = M ⊗ A B 0I | 56
is flat. Finally, if B is an A-algebra which is flat as an A-module, and if N is a flat B-module, then N
is also A-flat.
(6.2.2). Let A be a ring, B an A-algebra which is flat as an A-module. Let M, N be two A-modules,
such that M admits a finite presentation; then the canonical homomorphism
(6.2.2.1) Hom A ( M, N ) ⊗ A B −→ HomB ( M ⊗ A B, N ⊗ A B)
(sending u ⊗ b to the homomorphism m ⊗ b′ 7→ u(m) ⊗ b′ b) is an isomorphism.
(6.2.3). Let ( Aλ , ϕµλ ) be a filtered inductive system of rings; let A = lim Aλ . On the other hand, for
−→
each λ, let Mλ be an Aλ -module, and for λ ⩽ µ let θµλ : Mλ → Mµ be a ϕµλ -homomorphism, such
that ( Mλ , θµλ ) is an inductive system; M = lim Mλ is then an A-module. This being so, if for each
−→
λ, Mλ is a flat Aλ -module, then M is a flat A-module. Indeed, let J be an ideal of finite type of A; by
definition of the inductive limit, there exists an index λ and an ideal Jλ of Aλ such that J = Jλ A. If
we put J′µ = Jλ Aµ for µ ⩾ λ, we also have J = lim J′µ (where µ varies over the indices ⩾ λ), hence
−→
(the functor lim being exact and commuting with tensor products)
−→
M ⊗ A J = lim( Mµ ⊗ Aµ J′µ ) = lim J′µ Mµ = JM.
−→ −→
6.3. Local nature of flatness
(6.3.1). If A is a ring, S a multiplicative subset of A, S−1 A is a flat A-module. Indeed, for each
A-module N, N ⊗ A S−1 A identifies with S−1 N (1.2.5) and we know (1.3.2) that S−1 N is an exact
functor in N.
If now M is a flat A-module, S−1 M = M ⊗ A S−1 A is a flat S−1 A-module (6.2.1), so it is also
A-flat according to the above and from (6.2.1). In particular, if P is an S−1 A-module, we can consider
it as an A-module isomorphic to S−1 P; for P to be A-flat, it is necessary and sufficient that it is
S−1 A-flat.
(6.3.2). Let A be a ring, B an A-algebra, and T a multiplicative subset of B. If P is a B-module which
is A-flat, T −1 P is A-flat. Indeed, for each A-module N, we have ( T −1 P) ⊗ A N = ( T −1 B ⊗ B P) ⊗ A N
= T −1 B ⊗ B ( P ⊗ A N ) = T −1 ( P ⊗ A N ); T −1 ( P ⊗ A N ) is an exact functor in N, being the composition
of the two exact functors P ⊗ A N (in N) and T −1 Q (in Q). If S is a multiplicative subset of A such
that its image in B is contained in T, then T −1 P is equal to S−1 ( T −1 P), so it is also S−1 A-flat according
to (6.3.1).
(6.3.3). Let ϕ : A → B be a ring homomorphism, M a B-module. The following properties are
equivalent:
(a) M is a flat A-module.
(b) For each maximal ideal n of B, Mn is a flat A-module.
(c) For each maximal ideal n of B, by setting m = ϕ−1 (n), Mn is a flat Am -module.
Indeed, as Mn = ( Mn )m , the equivalence of (b) and (c) follows from (6.3.1), and the fact
that (a) implies (b) is a particular case of (6.3.2). It remains to see that (b) implies (a), that is 0I | 57
to say, that for each injective homomorhism u : N ′ → N of A-modules, the homomorphism
v = 1 ⊗ u : M ⊗ A N ′ → M ⊗ A N is injective. We have that v is also a homomorphism of B-
modules, and we know that for it to be injective, it suffices that for each maximal ideal n of B,
vn : ( M ⊗ A N ′ )n → ( M ⊗ A N )n is injective. But as
( M ⊗ A N )n = Bn ⊗ B ( M ⊗ A N ) = Mn ⊗ A N,
vn is none other that the homomorphism 1 ⊗ u : Mn ⊗ A N ′ → Mn ⊗ A N, which is injective since
Mn is A-flat.
6. FLATNESS 48
In particular (by taking B = A), for an A-module M to be flat, it is necessary and sufficient that
Mm is Am -flat for each maximal ideal m of A.
(6.3.4). Let M be an A-module; if M is flat, and if f ∈ A does not divide 0 in A, f does not kill
any element ̸= 0 in M, since the homomorphism m 7→ f · m is expressed as 1 ⊗ u, where u is the
multiplication a 7→ f · a on A and M is identified with M ⊗ A A; if u is injective, it is the same for
1 ⊗ u since M is flat. In particular, if A is integral, M is torsion-free.
Conversely, suppose that A is integral, M is torsion-free, and suppose that for each maximal
ideal m of A, Am is a discrete valuation ring; then M is A-flat. Indeed, it suffices (6.3.3) to prove that
Mm is Am -flat, and we can therefore suppose that A is already a discrete valuation ring. But as M is
the inductive limit of its submodules of finite type, and these latter submodules are torsion-free, we
can in addition reduce to the case where M is of finite type (6.1.2). The proposition follows in this
case from that M is a free A-module.
In particular, if A is an integral ring, ϕ : A → B a ring homomorphism making B a flat A-module
and ̸= {0}, ϕ is necessarily injective. Conversely, if B is integral, A a subring of B, and if for each
maximal ideal m of A, Am is a discrete valuation ring, then B is A-flat.
6.4. Faithfully flat modules
(6.4.1). For an A-module M, the following four properties are equivalent:
(a) For a sequence N ′ → N → N ′′ of A-modules to be exact, it is necessary and sufficient that
the sequence M ⊗ A N ′ → M ⊗ A N → M ⊗ A N ′′ is exact;
(b) M is flat for each A-module N, the relation M ⊗ A N = 0 implies N = 0;
(c) M is flat for each homomorphism v : N → N ′ of A-modules, the relation 1 M ⊗ v = 0, 1 M
being the identity automorphism of M;
(d) M is flat for each maximal ideal m of A, mM ̸= M.
When M satisfies these conditions, we say that M is a faithfully flat A-module; M is then
necessarily a faithful module. In addition, if u : N → N ′ is a homomorphism of A-modules, then
for u to be injective (resp. surjective, bijective), it is necessary and sufficient that 1 ⊗ u : M ⊗ A N →
M ⊗ A N ′ is so.
0I | 58
(6.4.2). A free module ̸= {0} is faithfully flat; it is the same for the direct sum of a flat module and a
faithfully flat module. If S is a multiplicative subset of A, then S−1 A is a faithfully flat A-module if
S consists of invertible elements (so S−1 A = A).
(6.4.3). Let 0 → M′ → M → M′′ → 0 be an exact sequence of A-modules; if M′ and M′′ are flat, and
if one of the two is faithfully flat, then M is also faithfully flat.
(6.4.4). Let A and B be two rings, M an A-module, N an ( A, B)-bimodule. If M is faithfully flat and
if N is a faithfully flat B-module, then M ⊗ A N is a faithfully flat B-module. In particular, if M and
N are two faithfully flat A-modules, then so is M ⊗ A N. If B is an A-algebra and if M is a faithfully
flat A-module, the B-module M( B) is faithfully flat.
(6.7.2). With the notation of (6.7.1), if F is f -flat at x, for each open neighbourhood U of y = f ( x ),
the functor ( f ∗ (G ) ⊗OX F ) x in G is exact on the category of (OY |U )-modules; indeed, this stalk
canonically identifies with Gy ⊗Oy Fx , and our assertion follows from the definition. In particular, if
f is a flat morphism, the functor f ∗ is exact on the category of OY -modules.
(6.7.3). Conversely, suppose the sheaf of rings OY is coherent, and suppose that for each open neigh-
bourhood U of y, the functor ( f ∗ (G ) ⊗OX F ) x is exact in G on the category of coherent (OY |U )-
modules. Then F is f -flat at x. In fact, it suffices to prove that for each ideal of finite type J of Oy ,
7. ADIC RINGS 50
the canonical homomorphism J ⊗Oy Fx → Fx is injective (6.1.1). We know (5.3.8) that there then
exists an open neighbourhood U of y and a coherent sheaf of ideals J of OY |U such that Jy = J, 0I | 60
hence the conclusion.
(6.7.4). The results of (6.1) for flat modules are immediately translated into propositions for sheaves
with are f -flat at a point:
If 0 → F ′ → F → F ′′ → 0 is an exact sequence of OX -modules and if F ′′ is f -flat at a point
x ∈ X, then, for each open neighbourhood U of y = f ( x ) and each (OY |U )-module G , the sequence
0 −→ ( f ∗ (G ) ⊗OX F ′ ) x −→ ( f ∗ (G ) ⊗OX F ) x −→ ( f ∗ (G ) ⊗OX F ′′ ) x −→ 0
is exact. For F to be f -flat at x, it is necessary and sufficient that F ′ is. We have similar statements
for the corresponding notions of a f -flat OX -modules over y ∈ Y, or of a f -flat OX -module.
(6.7.5). Let f : X → Y, g : Y → Z be two morphisms of ringed spaces; let x ∈ X, y = f ( x ), and F be
an OX -module. If F is f -flat at the point x and if the morphism g is flat at the point y, then F is
( g ◦ f )-flat at x (6.2.1). In particular, if f and g are flat morphisms, then g ◦ f is flat.
(6.7.6). Let X, Y be two ringed spaces, f : X → Y a flat morphism. Then the canonical homomorphism
of bifunctors (4.4.6)
(6.7.6.1) f ∗ (H omOY (F , G )) −→ H omOX ( f ∗ (F ), f ∗ (G ))
is an isomorphism when F admits a finite presentation (5.2.5).
Indeed, since the questions is local, we can assume that there exists an exact sequence OYm →
OY → F → 0. The two sides of (6.7.6.1) are right exact functors in F according to the hypotheses
n
on f ; we then have reduced to proving the proposition in the case where F = OY , in which the
result is trivial.
(6.7.8). We say that a morphism f : X → Y of ringed spaces is faithfully flat if f is surjective and if, for
each x ∈ X, Ox is a faithfully flat O f ( x) -module. When X and Y are locally ringed spaces (5.5.1), it is
equivalent to say that the morphism f is surjective and flat (6.6.2). When f is faithfully flat, f ∗ is an
exact and faithful functor on the category of OY -modules (6.6.1, a), and for an OY -module G to be
Y-flat, it is necessary and sufficient that f ∗ (G ) is (6.6.3).
P ROOF. (i) follows immediately from the definitions. To prove (ii), it suffices to note that for
each neighbourhood V of 0 in A, there exists an n > 0 such that Jn ⊂ V; if x ∈ A is such that x m ∈ J,
we have x mq ∈ V for q ⩾ n, so x is topologically nilpotent. □
Proposition (7.1.4). — Let A be a preadmissible ring, J an ideal of definition for A.
(i) For an ideal J′ of A to be contained in an ideal of definition, it is necessary and sufficient that there
n
exists an integer n > 0 such that J′ ⊂ J.
(ii) For an x ∈ A to be contained in an ideal of definition, it is necessary and sufficient that it is
topologically nilpotent.
P ROOF.
n
(i) If J′ ⊂ J, then for each open neighbourhood V of 0 in A, there exists an m such that
mn
Jm ⊂ V, thus J′ ⊂ V.
(ii) The condition is evidently necessary; it is sufficient, since if it satisfied, then there exists an
n
n such that x n ∈ J, so J′ = J + Ax is an ideal of definition, because it is open, and J′ ⊂ J.
□
Corollary (7.1.5). — In a preadmissible ring A, an open prime ideal contains all the ideals of definition.
Corollary (7.1.6). — The notation and hypotheses being that of (7.1.4), the following properties of an ideal
J0 of A are equivalent:
(a) J0 is the largest ideal of definition of A;
(b) J0 is a maximal ideal of definition;
(c) J0 is an ideal of definition such that the ring A/J0 is reduced.
For there to exist an ideal J0 to have these properties, it is necessary and sufficient that the nilradical of A/J
to be nilpotent; J0 is then equal to the ideal T of topologically nilpotent elements of A.
P ROOF. It is clear that (a) implies (b), and (b) implies (c) according to (7.1.4, ii), and (7.1.3, ii); for
the same reason, (c) implies (a). The latter assertion follows from (7.1.4, i) and (7.1.3, ii). □
When T/J, the nilradical of A/J, is nilpotent, and we denote by Ared the (reduced) quotient
ring A/T.
0I | 62
Corollary (7.1.7). — A preadmissible Noetherian ring admits a largest ideal of definition.
Corollary (7.1.8). — If a preadmissible ring A is such that, for an ideal of definition J, the powers Jn (n > 0)
n
form a fundamental system of neighbourhoods of 0, it is the same for the powers J′ for each ideal of definition
J′ of A.
Definition (7.1.9). — We say that a preadmissible ring A is preadic if there exists an ideal of definition
J for A such that the Jn form a fundamental system of neighbourhoods of 0 in A (or equivalently,
such that the Jn are open). We call a ring adic if it is a separated and complete preadic ring.
If J is an ideal of definition for a preadic (resp. adic) ring A, we say that A is a J-preadic (resp.
J-adic) ring, and that its topology is the J-preadic (resp. J-adic) topology. More generally, if M is
an A-module, the topology on M having for a fundamental system of neighbourhoods of 0 the
submodules Jn M is called the J-preadic (resp. J-adic) topology. According to (7.1.8), these topologies
are independent of the choice of the ideal of definition J.
Proposition (7.1.10). — Let A be an admissible ring, J an ideal of definition for A. Then J is contained in
the radical of A.
This statement is equivalent to any of the following corollaries:
Corollary (7.1.11). — For each x ∈ J, 1 + x is invertible in A.
Corollary (7.1.12). — For f ∈ A to be invertible in A, it is necessary and sufficient that its canonical image
in A/J is invertible in A/J.
Corollary (7.1.13). — For each A-module M of finite type, the relation M = JM (equivalent to M ⊗ A
( A/J) = 0) implies that M = 0.
7. ADIC RINGS 52
(7.2.3). Let A be an admissible topological ring, J an ideal of A contained in an ideal of definition (in
other words (7.1.4) such that (Jn ) tends to 0); we can consider on A the ring topology having for a
fundamental system of neighbourhoods of 0 the powers Jn (n > 0); we call again this the J-preadic
topology. The hypothesis that A is admissible implies that n Jn = 0, therefore the J-preadic
S
topology on A is separated; let A b = lim A/Jn be the completion of A for this topology (where the
n ←−
A/J are equipped with the discrete topology), and denote by u the (not necessarily continuous) ring
homomorphism A → A, b the projective limit of the sequence of homomorphisms un : A → A/Jn .
On the other hand, the J-preadic topology on A is finer than the given topology T on A; as A is
separated and complete for T , we can extend by continuity the identity map of A (equipped with
the J-preadic topology) to A equipped with T ; this gives a continuous representation v : A b → A.
Corollary (7.2.5). — Under the hypotheses of (7.2.3), the following conditions are equivalent:
(a) the homomorphism u is continuous; 0I | 64
(b) the homomorphism v is bicontinous;
(c) A is a J-adic ring.
7. ADIC RINGS 53
Corollary (7.2.6). — Let A be an admissible ring A, J an ideal of definition for A. For A to be Noetherian,
it is necessary and sufficient for A/J to be Noetherian and for J/J2 to be an A/J-module of finite type.
These conditions are evidently necessary. Conversely, suppose the conditions are satisfied; as
according to (7.2.4) A is complete for the J-preadic topology, for it to be Noetherian, it is necessary
and sufficient that the associated graded ring grad( A) (for the filtration on the Jn ) is ([CC, p .18–07,
th. 4]). Now, let a1 , . . . , an be the elements of J whose classes mod. J2 are the generators of J/J2
as a A/J-module. It is immediate by induction that the classes mod. Jm+1 of the monomials of
total degree m in the ai (1 ⩽ i ⩽ n) form a system of generators for the A/J-module Jm /Jm+1 . We
conclude that grad( A) is a ring isomorphic to a quotient of ( A/J)[ T1 , . . . , Tn ] (Ti indeterminates),
which finishes the proof.
Proposition (7.2.7). — Let ( Ai , uij ) be a projective system (i ∈ N) of discrete rings, and for each integer i,
let Ji be the kernel in Ai of the homomorphism u0i : Ai → A0 . We suppose that:
(a) For i ⩽ j, uij is surjective and its kernel is Jij+1 (therefore Ai is isomorphic to A j /Jij+1 ).
(b) J1 /J21 (= J1 ) is a module of finite type over A0 = A1 /J1 .
Let A = limi Ai , and for each integer n ⩾ 0, let un be the canonical homomorphism A → An , and let
←−
J(n+1) ⊂ A be its kernel. Then we have these conditions:
(i) A is an adic ring, having J = J(1) for an ideal of definition.
(ii) We have J(n) = Jn for each n ⩾ 1.
(iii) J/J2 is isomorphic to J1 = J1 /J21 , and as a result is a module of finite type over A0 = A/J.
P ROOF. The hypothesis of surjectivity on the uij implies that un is surjective; in addition, the
j +1
hypothesis (a) implies that J j = 0, therefore A is an admissible ring (7.2.2); by definition, the
J( n ) form a fundamental system of neighbourhoods of 0 in A, so (ii) implies (i). In addition, we
have J = limi Ji and the maps J → Ji are surjective, so (ii) implies (iii), and it remains to prove
←−
(ii). By definition, J(n) consists of the elements ( xk )k⩾0 of A such that xk = 0 for k < n, therefore
J(n) J(m) ⊂ J(n+m) , in other words the J(n) form a filtration of A. On the other hand, J(n) /J(n+1) is
isomorphic to the projection from J(n) to An ; as J(n) = limi>n Jin , this projection is none other than
←−
Jnn , which is a module over A0 = An /Jn . Now let a j = ( a jk )k⩾0 be r elements of J = J(1) such that
a11 , . . . , ar1 form a system of generators for J1 over A0 ; we will see that the set Sn of monomials of
total degree n and the a j generate the ideal J(n) of A. As Jii+1 = 0, it is clear first of all that Sn ⊂ J(n) ;
since A is complete for the filtration (J(m) ), it suffices to prove that the set Sn of classes mod. J(n+1)
of elements of Sn generate the graded module grad(J(n) ) over the graded ring grad( A) for the above
filtration ([CC, p. 18–06, lemme]); according to the definition of the multiplication on grad( A), it
suffices to prove that for each m, Sm is a system of generators for the A0 -module J(m) /J(m+1) , or
that Jm m is generated by the monomials of degree m in the a jm (1 ⩽ j ⩽ r). For this, it remains to
show that Jm is generated (as an Am -module) by the monomials of degree ⩽ m with respect to to
a jm ; the proposition being evident by definition for m = 1, we argue by induction on m, and let J′m
be the Am -submodule of Jm generated by these monomials. The relation Jm−1 = Jm /Jm m and the
induction hypothesis prove that Jm = J′m + Jm , hence, since J m+1 = 0, we have Jm = J′ m , and
m m m m
finally Jm = J′m . □
Corollary (7.2.8). — Under the conditions of Proposition (7.2.7), for A to be Noetherian, it is necessary and
sufficient that A0 is.
Proposition (7.2.9). — Suppose the hypotheses of Proposition (7.2.7): for each integer i, let Mi be an
Ai -module, and for i ⩽ j, let vij : M j → Mi be a uij -homomorphism, such that ( Mi , vij ) is a projective
system. In addition, suppose that M0 is an A0 -module of finite type and that the vij are surjective with kernel
Jij+1 M j . Then M = lim Mi is an A-module of finite type, and the kernel of the surjective un -homomorphism
←−
vn : M → Mn is Jn+1 M (such that Mn identifies with M/Jn+1 M = M ⊗ A ( A/Jn+1 )).
7. ADIC RINGS 54
P ROOF. Let zh = (zhk )k⩾0 be a system of s elements of M such that the zh0 (1 ⩽ h ⩽ s) forms a
system of generators for M0 ; we will show that the zh generate the A-module M. The A-module M
is separated and complete for the filtration by the M(n) , where M(n) is the set of y = (yk )k⩾0 in M
such that yk = 0 for k < n; it is clear that we have J(n) M ⊂ M(n) and that M(n) /M(n+1) = Jnn Mn .
We therefore have reduced to showing that the classes of the zh modulo M(0) generate the graded
module grad( M) (by the above filtration) over the graded ring grad( A) [CC, p. 18–06, lemme];
for this, we observe that it suffices to prove that the zhn (1 ⩽ h ⩽ s) generate the An -module Mn .
We argue by induction on n, the proposition being evident by definition for n = 0; the relation
Mn−1 = Mn /Jnn Mn and the induction hypothesis show that if Mn′ is the submodule of Mn generated
by the zhn , we have that Mn = Mn′ + Jnn Mn , and as Jn is nilpotent, this implies that Mn = Mn′ .
Similarly, passing to the associated graded modules shows that the canonical map from J(n) to M(n)
is surjective (thus bijection), in other words that J(n) M = Jn M is the kernel of M → Mn−1 . □ ErrII
Corollary (7.2.10). — Let ( Ni , wij ) be a second projective system of Ai -modules satisfying the conditions of
Proposition (7.2.9), and let N = lim Ni . There is a bijective correspondence between the projective systems
←−
(hi ) of Ai -homomorphisms hi : Mi → Ni and the homomorphisms of A-modules h : M → N (which is
necessarily continuous for the J-adic topologies).
P ROOF. It is clear that if h : M → N is an A-homomorphism, then we have h(Jn M) ⊂ Jn N,
hence the continuity of h; by passing to quotients, there corresponds to h a projective system of
Ai -homomorphisms hi : Mi → Ni , whose projective limit is h, hence the corollary. □
Remark (7.2.11). — Let A be an adic ring with an ideal of definition J such that J/J2 is an A/J-
module of finite type; it is clear that the Ai = A/Ji+1 satisfy the conditions of Proposition (7.2.7); as 0I | 66
A is the projective limit of the Ai , we see that Proposition (7.2.7) gives the description of all the adic
rings of the type considered (and in particular of all the adic Noetherian rings).
Example (7.2.12). — Let B be a ring, J an ideal of B such that J/J2 is a module of finite type over
B/J (or over B, equivalently); set A = limn B/Jn+1 ; A is the separated completion of B equipped
←−
with the J-preadic topology. If An = B/Jn+1 , then it is immediate that the An satisfy the conditions
of Proposition (7.2.7); therefore A is an adic ring and if J is the closure in A of the canonical image of
J, then J is an ideal of definition for A, Jn is the closure of the canonical image of Jn , A/Jn identifies
with B/Jn and J/J2 is isomorphic to J/J2 as an A/J-module. Similarly, if N is such that N/JN is a
B-module of finite type, and if we set Mi = N/Ji+1 N, then M = lim Mi is an A-module of finite
←−
type, isomorphic to the separated completion of N for the J-preadic topology, and Jn M identifies
with the closure of the canonical image of Jn N, and M/Jn M identifies with N/Jn N.
7.3. Preadic Noetherian rings
(7.3.1). Let A be a ring, J an ideal of A, and M an A-module; we denote by A b = lim A/Jn (resp.
←−n
M n
b = lim M/J M) the separated completion of A (resp. M) for the J-preadic topology. Let
u
←− v
n
M′ − → M − → M′′ → 0 be an exact sequence of A-modules; as M/Jn M = M ⊗ A ( A/Jn ), the
sequence
un vn
M′ /Jn M′ −→ M/Jn M −→ M′′ /Jn M′′ −→ 0
is exact for each n. In addition, as v(Jn M ) = Jn v( M) = Jn M′′ , vb = lim vn is surjective (Bourbaki,
←−
Top. gén., Chap. IX, 2nd ed., p. 60, Cor. 2). On the other hand, if z = (zk ) is an element of the kernel
of vb, then for each integer k, there exists a z′k ∈ M′ /Jk M′ such that uk (z′k ) = zk ; we conclude that
there exists a z′ = (z′n ) ∈ M
c′ such that the first k components of ub(z′ ) coincide with the z; in other
words, the image of M c′ under ub is dense in the kernel of vb.
If we suppose that A is Noetherian, then so is A,b according to (7.2.12), J/J2 is then an A-module
of finite type. In addition, we have the following theorem.
Theorem (7.3.2). — (Krull’s Theorem). Let A be a Noetherian ring, J an ideal of A, M an A-module of
finite type, and M′ a submodule of M; then the induced topology on M′ by the J-preadic topology of M is
identical to the J-preadic topology of M′ .
This follows immediately from
7. ADIC RINGS 55
Lemma (7.3.2.1). — (Artin–Rees Lemma). Under the hypotheses of (7.3.2), there exists an integer p such
that, for n ⩾ p, we have
M ′ ∩ Jn M = Jn − p ( M ′ ∩ J p M ).
For the proof, see [CC, p. 2–04].
0I | 67
Corollary (7.3.3). — Under the hypotheses of (7.3.2), the canonical map M ⊗ A A b→M b is bijective, and the
functor M ⊗ A Ab is exact in M on the category of A-modules of finite type; as a result, the separated J-adic
completion A
b is a flat A-module (6.1.1).
P ROOF. We first note that M b is an exact functor in M on the category of A-modules of finite
′ u v
type. Indeed, let 0 → M − →M− → M′′ → 0 be an exact sequence; we have seen that vb : M b →d M′′ is
surjective (7.3.1); on the other hand, if i is the canonical homomorphism M → M, b it follows from
′
Krull’s Theorem (7.3.2) that the closure of i (u( M )) in M identifies with the separated completion of
b
M′ for the J-preadic topology; thus ub is injective, and according to (7.3.1), the image of ub is equal to
the kernel of vb.
This being so, the canonical map M ⊗ A A b→M b is obtained by passing to the projective limit of
the maps M ⊗ A A n n
b → M ⊗ A ( A/J ) = M/J M. It is clear that this map is bijective when M is of the
form A p . If M is an A-module of finite type, then we have an exact sequence A p → Aq → M → 0,
hence, by virtue of the right exactness of the functors M ⊗ A A b and M b (in M) on the category of
A-modules of finite type, we have the commutative diagram
Ap ⊗A A
b / Aq ⊗ A A
b / M ⊗A A
b /0
cp
A /A
cq /M
b / 0,
where the two rows are exact and the first two vertical arrows are isomorphisms; this immediately
finishes the proof. □
Corollary (7.3.4). — Let A be a Noetherian ring, J an ideal of A, M and N two A-modules of finite type;
we have the canonical functorial isomorphisms
( M ⊗ A N )∧ ≃ M A
b (Hom A ( M, N ))∧ ≃ Hom b ( M,
b ⊗ b N,
A
b Nb ).
Corollary (7.3.5). — Let A be a Noetherian ring, J an ideal of A. The following conditions are equivalent:
(a) J is contained in the radical of A.
(b) A
b is a faithfully flat A-module (6.4.1).
(c) Each A-module of finite type is separated for the J-preadic topology.
(d) Each submodule of an A-module of finite type is closed for the J-preadic topology.
P ROOF. As A b is a flat A-module, the conditions (b) and (c) are equivalent, since (b) is equivalent
to saying that if M is an A-module of finite type, then the canonical map M → M b = M ⊗A A b is
injective (6.6.1, c). It is immediate that (c) implies (d), since if N is a submodule of an A-module
M of finite type, then M/N is separated for the J-preadic topology, so N is closed in M. We show
that (d) implies (a): if m is a maximal ideal of A, then m is closed in A for the J-preadic topology,
so m = p⩾0 (m + J p ), and as m + J p is necessarily equal to A or to m, we have that m + J p = m
T
for large enough p, hence J p ⊂ m, and J ⊂ m when m is prime. Finally, (a) implies (b): indeed, let 0I | 68
P be the closure of {0} in an A-module M of finite type, for the J-preadic topology; according to
Krull’s Theorem (7.3.2), the topology induced on P by the J-preadic topology of M is the J-preadic
topology of P, so JP = P; as P is of finite type, it follows from Nakayama’s Lemma that P = 0 (J
being contained in the radical of A). □
We note that the conditions of Corollary (7.3.5) are satisfied when A is a local Noetherian ring and
J ̸= A is any ideal of A.
7. ADIC RINGS 56
Corollary (7.3.6). — If A is a J-preadic Noetherian ring, then each A-module of finite type is separated and
complete for the J-preadic topology.
b = A, this follows immediately from Corollary (7.3.3).
P ROOF. As we then have A □
We conclude that Proposition (7.2.9) gives the description of all the modules of finite type over
an adic Noetherian ring.
Corollary (7.3.7). — Under the hypotheses of (7.3.2), the kernel of the canonical map M → M
b = M ⊗A A
b
is the set of the x ∈ M killed by an element of 1 + J.
P ROOF. For each x ∈ M in this kernel, it is necessary and sufficient that the separated completion
of the submodule Ax is 0 (by Krull’s Theorem (7.3.2)), in other words, that x ∈ Jx. □
When A is Noetherian, the separated completion M b of M for the m-preadic topology is then an
A-module
b 2
of finite type; indeed, as m/m is then an A-module of finite type, this follows from Example
(7.2.12) and from the hypothesis on M/mM.
In particular, if we suppose that in addition A is complete and M is separated for the m-preadic
topology (in other words, n mn M = 0), then M is also an A-module of finite type: indeed, M
T
b is then
an A-module of finite type, and as M identifies with a submodule of M, M is also of finite type (and
b
is indeed identical to its completion according to Corollary (7.3.6)).
Proposition (7.4.2). — Let A, B be two local rings, m, n their maximal ideals, and suppose that B is
Noetherian. Let ϕ : A → B be a local homomorphism, M a B-module of finite type. If M is a quasi-finite
A-module, then the m-preadic and n-preadic topologies on M are identical, thus separated.
P ROOF. We note that by hypothesis M/mM is of finite length as an A-module, thus also a fortiori
as a B-module. We conclude that n is the unique prime ideal of B containing the annihilator of M/mM:
indeed, we immediately reduce (according to (1.7.4) and (1.7.2)) to the case where M/mM is simple,
thus necessarily isomorphic to B/n, and our assertion is evident in this case. On the other hand, as
M is a B-module of finite type, the prime ideals which contain the annihilator of M/mM are those
which contain mB + b, where we denote by b the annihilator of the B-module M (1.7.5). As B is
Noetherian, we conclude ([Sam53b, p. 127, Cor. 4]) that mB + b is an ideal of definition for B, in 0I | 69
other words there exists a k > 0 such that nk ⊂ mB + b ⊂ n; as a result, for each h > 0, we have
nhk ⊂ (mB + b)h M = mh M ⊂ nh M,
which proves that the m-preadic and n-preadic topologies on M are the same; the second is separated
according to Corollary (7.3.5). □
Corollary (7.4.3). — Under the hypotheses of Proposition (7.2.4), if in addition A is Noetherian and complete
for the m-preadic topology, then M is an A-module of finite type.
P ROOF. Indeed, M is then separated for the m-preadic topology, and our assertion follows from
the remark after Definition (7.4.1). □
(7.4.4). The most important case of Proposition (7.4.2) is when B is a quasi-finite A-module, i.e.,
B/mB is an algebra of finite rank over k = A/m; furthermore, this condition can be broken down into
the combination of the following:
(i) mB is an ideal of definition for B;
(ii) B/n is an extension of finite rank of the field A/m.
When this is so, every B-module of finite type is evidently a quasi-finite A-module.
Corollary (7.4.5). — Under the hypotheses of Proposition (7.4.2), if b is the annihilator of the B-module M,
then B/b is a quasi-finite A-module.
7. ADIC RINGS 57
Proposition (7.5.4). —
(i) If A is an admissible ring, then so is A′ = A{ T1 , . . . , Tr }.
7. ADIC RINGS 58
(ii) Let A be an adic ring, J an ideal of definition for A such that J/J2 is of finite type over A/J. If we 0I | 71
2
J′
set = JA′ ,then A′
is also a J′ -adic ring, and J′ /J′ is of finite type over A′ /J′ . If in addition
A is Noetherian, then so is A′ .
P ROOF.
(i) If J is an ideal of A, J′ the ideal of A′ consisting of the ∑α cα T α such that cα ∈ J for all α,
then (J′ )n ⊂ (Jn )′ ; if J is an ideal of definition for A, then J′ is also an ideal of definition
for A′ .
(ii) Set Ai = A/Ji+1 , and for i ⩽ j, let uij be the canonical homomorphism A/J j+1 → A/Ji+1 ;
set Ai′ = Ai [ T1 , . . . , Tr ], and let uij′ be the homomorphism A′j → Ai′ (i ⩽ j) obtained by
applying uij to the coefficients of the polynomials in A′j . We will show that the projective
system ( Ai′ , uij′ ) satisfies the conditions of Proposition (7.2.7); as J′ is the kernel of A′ → A0′ ,
this proves the first assertion of (ii). It is clear that the uij′ are surjective; the kernel Ji′ of
u0i is the set of polynomials in Ai [ T1 , . . . , Tr ] whose coefficients are in J/Ji+1 ; in particular,
J1′ is the set of polynomials in A1 [ T1 , . . . , Tr ] whose coefficients are in J/J2 . As J/J2 is
2
of finite type over A1 = A/J2 , we see that J1′ /J1′ is a module of finite type over A1′
i +1
(or equivalently, over A0′ = A1′ /J1′ ). We will now show that the kernel of uij is J′j . It
i +1
is evident that J′j is contained in this kernel. On the other hand, let a1 , . . . , am be the
elements of J whose classes mod J2 generate J/J2 ; we verify immediately that the classes
mod J j+1 of monomials of degree ⩽ j in the ak (1 ⩽ k ⩽ m) generate J/J j+1 , and the classes
of monomials of degree > i and ⩽ j generate Ji+1 /J j+1 ; a monomial in the Tk having such
an element for a coefficient is thus a product of i + 1 elements of Ji′ , which establishes our
assertion. Finally, if A is Noetherian, then so is A′ /J′ = ( A/J)[ T1 , . . . , Tr ], hence A′ is
Notherian (7.2.8).
□
Proposition (7.5.5). — Let A be a Notherian J-adic ring, B an admissible topological ring, ϕ : A → B a
continuous homomorphism, making B and A-algebra. The following conditions are equivalent:
(a) B is Noetherian and JB-adic, and B/JB is an algebra of finite type over A/J.
(b) B is topologically A-isomorphic to lim Bn , where Bn = Bm /Jn+1 Bm for m ⩾ n, and B1 is an
←−
algebra of finite type over A1 = A/J2 .
(c) B is topologically A-isomorphic to a quotient of an algebra of the form A{ T1 , . . . , Tr } by an ideal
(necessarily closed according to Corollary (7.3.6) and Proposition (7.5.4, ii)).
P ROOF. As A is Noetherian, so is A′ = A{ T1 , . . . , Tr } (7.5.4), so (c) implies that B is Noetherian;
n
as = JA′ is an open neighbourhood of 0 in A′ such that the J′ form a fundamental system of
J′
n
neighbourhoods of 0, the images Jn B of the J′ form a fundamental system of neighbourhoods of 0
in B, and as B is separated and complete, B is a JB-adic ring. Finally, B/JB is an algebra (over ( A/J)
quotient of A′ /JA′ = ( A/J)[ T1 , . . . , Tr ], so it is of finite type, which proves that (c) implies (a).
If B is JB-adic and Noetherian, then B is isomorphic to lim Bn , where Bn = B/Jn+1 B (7.2.11),
←−
and JB/J2 B is a module of finite type over B/JB. Let ( a j )1⩽j⩽s be a system of generators for the
B/JB-module JB/J2 B, and let (ci )1⩽i⩽r be a system of elements of B/J2 B such that the classes mod 0I | 72
JB/J2 B form a system of generators for the A/J-algebra B/JB; we see immediately that the ci a j
form a system of generators for the A/J2 -algebra B/J2 B, hence (a) implies (b).
It remains to prove that (b) implies (c). The hypotheses imply that B1 is a Noetherian ring,
and as B1 = B2 /J2 B2 , we have J2 B1 = 0, hence JB1 = JB1 /J2 B1 is a B0 -module of finite type.
The conditions of Proposition (7.2.7) are thus satisfied by the projective system ( Bn ) and B is a
JB-adic ring. Let (ci )1⩽i⩽r be a finite system of elements of B whose classes mod JB generate
the A/J-algebra B/JB, and whose linear combinations with coefficients in J are such that their
classes mod J2 B generate the B0 -module JB/J2 B. There exists a continuous A-homomorphism u
from A′ = A{ T1 , . . . , Tr } to B which reduces to ϕ on A and is such that u( Ti ) = ci for 1 ⩽ i ⩽ r
(7.5.3); if we prove that u is surjective, then (c) will be established, since from u( A′ ) = B we deduce
that u(Jn A′ ) = Jn B, in other words that u is a strict morphism of topological rings and B is this
7. ADIC RINGS 59
P ROOF. If vλ is the canonical homomorphism S−1 A → Sλ−1 Aλ induced by uλ , then the kernel
of vλ is surjective, hence the proposition (7.2.1). □
(7.6.5). We say that A{S−1 } is the completed ring of fractions of A with denominators in S. With the
above notation, it is clear that the inverse image of S−1 Jλ in A contains Jλ , hence the canonical map
A → S−1 A is continuous, and if we compose it with the canonical map S−1 A → A{S−1 }, we obtain
a canonical continuous homomorphism A → A{S−1 }, the projective limit of the homomorphisms
A → Sλ−1 Aλ .
(7.6.6). The couple consisting of A{S−1 } and the canonical homomorphism A → A{S−1 } are
characterized by the following universal property: every continuous homomorphism u from A to
a linearly topologized ring B, separated and complete, such that u(S) consists of the invertible
u′
elements of B, uniquely factorizes as A → A{S−1 } −
→ B, where u′ is continuous. Indeed, u uniquely
v′
factorizes as A → S−1 A − → B; as for each open ideal K of B we have that u−1 (K) contains a
−1
Jλ , v′ (K) necessarily contains S−1 Jλ , so v′ is continuous; since B is separated and complete, v′
u′
uniquely factorizes as S−1 A → A{S−1 } −
→ B, where u′ is continuous; hence our assertion.
(7.6.7). Let B be a second linearly topologized ring, T a multiplicative subset of B, ϕ : A → B a contin-
uous homomorphism such that ϕ(S) ⊂ T. According to the above, the continuous homomorphism
ϕ ϕ′
A− → B → B{ T −1 } uniquely factorizes as A → A{S−1 } − → B{ T −1 }, where ϕ′ is continuous. In par-
ticular, if B = A and if ϕ is the identity, we see that for S ⊂ T we have a continuous homomorphism
ρ T,S : A{S−1 } → A{ T −1 } obtained by passing to the separated completion from S−1 A → T −1 A; if
U is a third multiplicative subset of A such that S ⊂ T ⊂ U, then we have ρU,S = ρU,T ◦ ρ T,S .
7. ADIC RINGS 60
(7.6.8). Let S1 , S2 be two multiplicative subsets of A, and let S2′ be the canonical image of S2 in
−1
A{S1−1 }; we then have a canonical topological isomorphism A{(S1 S2 )−1 } ≃ A{S1−1 }{S2′ }, as we
−1
see from the canonical isomorphism (S1 S2 )−1 A ≃ S2′′ (S1−1 A) (where S2′′ is the canonical image of
−1
S2 in S1 A), which is bicontinuous.
(7.6.9). Let a be an open ideal of A; we can assume that Jλ ⊂ a for all λ, and as a result S−1 Jλ ⊂ S−1 a
in the ring S−1 A, in other words, S−1 a is an open ideal of S−1 A; we denote by a{S−1 } its separated
completion, equal to lim(S−1 a/S−1 Jλ ), which is an open ideal of A{S−1 }, isomorphic to the closure
←−
of the canonical image of S−1 a. In addition, the discrete ring A{S−1 }/a{S−1 } is canonically isomorphic
to S−1 A/S−1 a = S−1 ( A/a). Conversely, if a′ is an open ideal of A{S−1 }, then a′ contains an ideal
of the form Jλ {S−1 } which is the inverse image of an ideal of S−1 A/S−1 Jλ , which is necessarily
(1.2.6) of the form S−1 a, where a ⊃ Jλ . We conclude that a′ = a{S−1 }. In particular (1.2.6):
Proposition (7.6.10). — The map p 7→ p{S−1 } is an increasing bijection from the set of open prime ideals
p of A such that p ∩ S = ∅ to the set of open prime ideals of A{S−1 }; in addition, the field of fractions of 0I | 74
A{S−1 }/p{S−1 } is canonically isomorphic to that of A/p.
Proposition (7.6.11). —
(i) If A is an admissible ring, then so is A′ = A{S−1 }, and for every ideal of definition J for A,
J′ = J{S−1 } is an ideal of definition for A′ .
(ii) Let A be an adic ring, J an ideal of definition for A such that J/J2 is of finite type over A/J; then
2
A′ is a J′ -adic ring and J′ /J′ is of finite type over A′ /J′ . If in addition A is Noetherian, then so
is A′ .
P ROOF.
(i) If J is an ideal of definition for A, then it is clear that S−1 J is an ideal of definition for
the topological ring S−1 A, since we have (S−1 J)n = S−1 Jn . Let A′′ be the separated ring
n n
associated to S−1 A, J′′ the image of S−1 J in A′′ ; the image of S−1 Jn is J′′ , so J′′ tends to
n n
0 in A′′ ; as J′ is the closure of J′′ in A′ , J′ is contained in the closure of J′′ , hence tends to
′
0 in A .
(ii) Set Ai = A/Ji+1 , and for i ⩽ j, let uij be the canonical homomorphism A/J j+1 → A/Ji+1 ;
let Si be the canonical image of S in Ai , and set Ai′ = Si−1 Ai ; finally, let uij′ : A′j → Ai′ be the
homomorphism canonically induced by uij . We show that the projective system ( Ai′ , uij′ )
satisfies the conditions of Proposition (7.2.7): it is clear that the uij′ are surjective; on the other
i +1
hand, the kernel of uij′ is S− 1 i +1
j (J /J j+1 ) (1.3.2), equal to J′j , where J′j = S− 1
j (J/J
j +1 );
2 2
finally, J1′ /J1′ = S1−1 (J/J2 ), and as J/J2 is of finite type over A/J2 , J1′ /J1′ is of finite
type over A1′ . Finally, if A is Noetherian, then so is A0′ = S0−1 ( A/J), which finishes the
proof (7.2.8).
□
Corollary (7.6.12). — Under the hypotheses of Proposition (7.6.11, ii), we have (J{S−1 })n = Jn {S−1 }.
P ROOF. This follows from Proposition (7.2.7) and the proof of Proposition (7.6.11). □
Proposition (7.6.13). — Let A be an adic Noetherian ring, S a multiplicative subset of A; then A{S−1 } is a
flat A-module.
P ROOF. If J is an ideal of definition for A, then A{S−1 } is the separated completion of the
Noetherian ring S−1 A equipped with the S−1J-preadic topology; as a result (7.3.3) A{S−1 } is a flat
S−1 A-module; as S−1 A is a flat A-module (6.3.1), the proposition follows from the transitivity of
flatness (6.2.1). □
Corollary (7.6.14). — Under the hypotheses of Proposition (7.6.13), let S′ ⊂ S be a second multiplicative
−1
subset of A; then A{S−1 } is a flat A{S′ }-module.
7. ADIC RINGS 61
−1
P ROOF. By (7.6.8), A{S−1 } canonically identifies with A{S′ }{S0−1 }, where S0 is the canonical
−1 −1
image of S in A{S′ }, and A{S′ } is Noetherian (7.6.11). □
(7.6.15). For each element f of a linearly topologized ring A, we denote by A{ f } the completed ring
of fractions A{S− 1 n
f }, where S f is the multiplicative set of the f (n ⩾ 0); for each open ideal a of
A, we write a{ f } for a{S− 1
f }. If g is a second element of A, then we have a canonical continuous
homomorphism A{ f } → A{ f g} (7.6.7). When f varies over a multiplicative subset S of A, the A{ f }
form a filtered inductive system with the above homomorphisms; we set A{S} = lim f ∈S A{ f } . For
−→
every f ∈ S, we have a homomorphism A{ f } → A{S−1 } (7.6.7), and these homomorphisms form 0I | 75
an inductive system; by passing to the inductive limit, they thus define a canonical homomorphism
A { S } → A { S −1 } .
Proposition (7.6.16). — If A is a Noetherian ring, then A{S−1 } is a flat module over A{S} .
P ROOF. By (7.6.14), A{S−1 } is flat for each of the rings A{ f } for f ∈ S, and the conclusion
follows from (6.2.3). □
Proposition (7.6.17). — Let p be an open prime ideal in an admissible ring A, and let S = A − p. Then
the rings A{S−1 } and A{S} are local rings, the canonical homomorphism A{S} → A{S−1 } is local, and the
residue fields of A{S} and A{S−1 } are canonically isomorphic to the field of fractions of A/p.
P ROOF. Let J ⊂ p be an ideal of definition for A; we have S−1 J ⊂ S−1 p = pAp , so Ap /S−1 J is
a local ring; we conclude from Corollary (7.1.12), (7.6.9), and Proposition (7.6.11, i) that A{S−1 } is
a local ring. Set m = lim f ∈S p{ f } , which is an ideal in A{S} ; we will see that each element in A{S}
−→
not in m is invertible. Indeed, such an element is the image in A{S} of an element z ∈ A{ f } not in
p{ f } , for an f ∈ S; its canonical image z0 in A{ f } /J{ f } = S− 1 −1
f ( A/J) therefore is not in S f (p/J)
k
(7.6.9), which means that z0 = x/ f , where x ̸∈ p and x, f are the classes of x, f mod J. As x ∈ S,
we have g = x f ∈ S, and in S− 1 k +1 /gk of x/ f k ∈ S−1 A admits
g A, the canonical image y0 = x f
an inverse x k−1 f 2k /gk . This implies a fortiori that the image of y0 in S− 1 −1
g A/S g J is invertible, so
((7.6.9) and Corollary (7.1.12)) the canonical image y of z in A{ g} is invertible; the image of z in
A{S} (equal to that of y) is as a result invertible. We thus see that A{S} a local ring with maximal
ideal m; in addition, the image of p{ f } in A{S−1 } is contained in the maximal ideal p{S−1 } of this
ring; a fortiori, the image of m in A{S−1 } is contained in p{S−1 }, so the canonical homomorphism
A{S} → A{S−1 } is local. Finally, as each element of A{S−1 }/p{S−1 } is the image of an element in
the ring S− 1 −1 −1
f A for a suitable f ∈ S, the homomorphism A{S} → A { S } /p{ S } is surjective, and
gives an isomorphism of the residue fields by passing to quotients. □
Corollary (7.6.18). — Under the hypotheses of Proposition (7.6.17), if we suppose also that A is an adic
Noetherian ring, then the local rings A{S−1 } and A{S} are Noetherian, and A{S−1 } is a faithfully flat
A{S} -module.
P ROOF. We know from before (7.6.11, ii) that A{S−1 } is Noetherian and A{S} -flat (7.6.16); as the
homomorphism A{S} → A{S−1 } is local, we conclude that A{S−1 } is a faithfully flat A{S} -module
(6.6.2), and as a result that A{S} is Noetherian (6.5.2). □
7. ADIC RINGS 62
Proposition (7.7.8). — Let A be a preadic ring, J an ideal of defintion for A, M an A-module of finite type,
equipped with the J-preadic topology. For every topological adic Noetherian A-algebra B, B ⊗ A M identifies
with the completed tensor product ( B ⊗ A M)∧ .
P ROOF. If K is an ideal of definition for B, there exists by hypothesis an integer m such that
Jm B ⊂ K, so Im( B ⊗ A Jnm M) = Im(Jnm B ⊗ A M ) ⊂ Im(Kn B ⊗ A M ) = Kn ( B ⊗ A M ); we conclude
that over B ⊗ A M, the tensor products of the topologies of B and M is the K-preadic topology. As
B ⊗ A M is a B-module of finite type, the proposition follows from Corollary (7.3.6). □
(7.8.1). Let A be a Noetherian J-adic ring, M and N two A-modules of finite type, equipped with
the J-preadic topology; we know (7.3.6) that they are separated and complete; in addition, every
A-homomorphism M → N is automatically continuous, and the A-module Hom A ( M, N ) is of
finite type. For every integer i ⩾ 0, set Ai = A/Ji+1 , Mi = M/Ji+1 M, Ni = N/Ji+1 N; for i ⩽ j,
every homomorphism u j : M j → Nj maps Ji+1 M j to Ji+1 Nj , thus giving by passage to quotients
a homomorphism ui : Mi → Ni , which defines a canonical homomorphism Hom A j ( M j , Nj ) →
Hom Ai ( Mi , Ni ); in addition, the Hom Ai ( Mi , Ni ) form a projective system for these homomorphisms,
and it follows from Corollary (7.2.10) that there is a canonical isomorphism ϕ : Hom A ( M, N ) →
limi Hom Ai ( Mi , Ni ). In addition:
←−
Proposition (7.8.2). — If M and N are modules of finite type over a J-adic Noetherian ring A, then the
submodules Hom A ( M, Ji+1 N ) form a fundamental system of neighbourhoods of 0 in Hom A ( M, N ) for the
J-adic topology, and the canonical isomorphism ϕ : Hom A ( M, N ) → limi Hom Ai ( Mi , Ni ) is a topological
←−
isomorphism.
P ROOF. We can consider M as the quotient of a free A-module L of finite type, and as a result
identify Hom A ( M, N ) as a submodule of Hom A ( L, N ); in this identification, Hom A ( M, Ji+1 N ) is the
intersection of Hom A ( M, N ) and Hom A ( L, Ji+1 N ); as the induced topology on Hom A ( M, N ) by the
J-adic topology of Hom A ( L, N ) is the J-adic (7.3.2), we have reduced to proving the first assertion
for M = L = Am ; but then Hom A ( L, N ) = N m , Hom A ( L, Ji+1 N ) = (Ji+1 N )m = Ji+1 N m and the 0I | 78
result is evident. To establish the second assertion, we note that the image of Hom A ( M, Ji+1 N ) in
Hom A j ( M j , Nj ) is zero for j ⩽ i, hence ϕ is continuous; conversely, the inverse image in Hom A ( M, N )
of 0 of Hom Ai ( Mi , Ni ) is Hom A ( M, Ji+1 N ), so ϕ is bicontinuous. □
If we only suppose that A is a Noetherian J-preadic ring, M and N two A-modules of finite
type, separated for the J-preadic topology, then the following proof shows that the first assertion of
Proposition (7.8.2) remains valid, and that ϕ is a topological isomorphism from Hom A ( M, N ) to a
submodule of limi Hom Ai ( Mi , Ni ).
←−
Proposition (7.8.3). — Under the hypotheses of Proposition (7.8.2), the set of injective (resp. surjective,
bijective) homomorphisms from M to N is an open subset of Hom A ( M, N ).
h w (Y ′ ) h w (Y )
h X ′ (u)
h X ′ (Y ′ ) / h X ′ (Y )
g (Y ′ ) g (Y )
F (u)
F (Y ′ ) / F (Y )
( g(Y ))((h X (v))(1X )) = ( g(Y ))(v) by definition of h X (v), in other words, it is equal to g(Y ), which
finishes the proof. □
(8.1.5). Recall that a subcategory C′ of a category C is defined by the condition that its objects are objects
of C, and that if X ′ , Y ′ are two objects of C′ , then the set HomC′ ( X ′ , Y ′ ) of morphisms X ′ → Y ′ in C′
is a subset of the set HomC ( X ′ , Y ′ ) of morphisms X ′ → Y ′ in C, the canonical map of “composition
of morphisms”
HomC′ ( X ′ , Y ′ ) × HomC′ (Y ′ , Z ′ ) −→ HomC′ ( X ′ , Z ′ )
being the restriction of the canonical map 0III | 7
′ ′ ′ ′ ′ ′
HomC ( X , Y ) × HomC (Y , Z ) −→ HomC ( X , Z ).
We say that C′ is a full subcategory of C if HomC′ ( X ′ , Y ′ ) = HomC ( X ′ , Y ′ ) for every pair of objects
inC′ . The subcategory C′′ of C consisting of the objects of C isomorphic to objects of C′ is then again a
full subcategory of C, equivalent (T, 1.2) to C′ as we verify easily.
A covariant functor F : C1 → C2 is called fully faithful if for every pair of objects X1 , Y1 of C1 , the
map u 7→ F (u) from Hom( X1 , Y1 ) to Hom( F ( X1 ), F (Y1 )) is bijective; this implies that the subcategory
F (C1 ) of C2 is full. In addition, if two objects X1 , X1′ have the same image X2 , then there exists a
unique isomorphism u : X1 → X1′ such that F (u) = 1X1 . For each object X2 of F (C1 ), let G ( X2 ) be
one of the objects X1 of C1 such that F ( X1 ) = X2 (G is defined by means of the axiom of choice);
for each morphism v : X2 → Y2 in F (C1 ), G (v) will be the unique morphism u : G ( X2 ) → G (Y2 )
such that F (u) = v; G is then a functor from F (C1 ) to C1 ; FG is the identity functor on F (C1 ), and the
above shows that there exists an isomorphism of functors ϕ : 1C1 → GF such that F, G, ϕ, and the
identity 1F(C1 ) → FG defines an equivalence between the category C1 and the full subcategory F (C1 )
of C2 (T, 1.2).
(8.1.6). We apply Proposition (8.1.4) to the case where F is h X ′ , X ′ being any object of C; the map
β : Hom( X, X ′ ) → Hom(h X , h X ′ ) is none other than the map w 7→ hw defined in (8.1.2); this map
being bijective, we see with the terminology of (8.1.5) that:
Proposition (8.1.7). — The canonical functor h : C → Hom(Cop , Set) is fully faithful.
(8.1.8). Let F be a contravariant functor from C to Set; we say that F is representable if there exists
an object X ∈ C such that F is isomorphic to h X ; it follows from Proposition (8.1.7) that the data of
an X ∈ C and an isomorphism of functors g : h X → F determines X up to unique isomorphism.
Proposition (8.1.7) then implies that h defines an equivalence between C and the full subcategory of
Hom(Cop , Set) consisting of the contravariant representable functors. It follows from Proposition (8.1.4)
that the data of a natural transformation g : h X → F is equivalent to that of an element ξ ∈ F ( X );
to say that g is an isomorphism is equivalent to the following condition on ξ: for every object Y of C
the map v 7→ ( F (v))(ξ ) from Hom(Y, X ) to F (Y ) is bijective. When ξ satisfies this condition, we say
that the pair ( X, ξ ) represents the representable functor F, By abuse of language, we also say that the
object X ∈ C represents F if there exists a ξ ∈ F ( X ) such that ( X, ξ ) represents F, in other words if
h X is isomorphic to F.
Let F, F ′ be two contravariant representable functors from C to Set, h X → F and h X ′ → F ′
two isomorphisms of functors. Then it follows from (8.1.6) that there is a canonical bijective
correspondence between Hom( X, X ′ ) and the set Hom( F, F ′ ) of natural transformations F → F ′ .
(8.1.9). Example I. Projective limits. The notion of a contravariant representable functor covers in
particular the “dual” notion of the usual notion of a “solution to a universal problem”. More 0III | 8
generally, we will see that the notion of the projective limit is a special case of the notion of a
representable functor. Recall that in a category C, we define a projective system by the data of a
preordered set I, a family ( Aα )α∈ I of objects of C, and for every pair of indices (α, β) such that α ⩽ β,
a morphism uαβ : A β → Aα . A projective limit of this system in C consists of an object B of C (denoted
lim Aα ), and for each α ∈ I, a morphism uα : B → Aα such that: 1st . uα = uαβ u β for α ⩽ β; 2nd . for
←−
every object X of C and every family (vα )α∈ I of morphisms vα : X → Aα such that vα = uαβ v β for
α ⩽ β, there exists a unique morphism v : X → B (denoted lim vα ) such that vα = uα v for all α ∈ I
←−
(T, 1.8). This can be interpreted in the following way: the uαβ canonically define maps
uαβ : Hom( X, A β ) −→ Hom( X, Aα )
8. REPRESENTABLE FUNCTORS 66
which define a projective system of sets (Hom( X, Aα ), uαβ ), and (vα ) is by definition an element of the
set lim Hom( X, Aα ); it is clear that X 7→ lim Hom( X, Aα ) is a contravariant functor from C to Set, and
←− ←−
the existence of the projective limit B is equivalent to saying that (vα ) 7→ lim vα is an isomorphism of
←−
functors in X
(8.1.9.1) lim Hom( X, Aα ) ≃ Hom( X, B),
←−
in other words, that the functor X 7→ lim Hom( X, Aα ) is representable.
←−
(8.1.10). Example II. Final objects. Let C be a category, { a} a singleton set. Consider the contravariant
functor F : C → Set which sends every object X of C to the set { a}, and every morphism X → X ′
in C to the unique map { a} → { a}. To say that this functor is representable means that there exists
an object e ∈ C such that for every Y ∈ C, Hom(Y, e) = he (Y ) is a singleton set; we say that e is an
final object of C, and it is clear that two final objects of C are isomorphic (which allows us to define, in
general with the axiom of choice, one final object of C which we then denote eC ). For example, in the
category Set, the final objects are the singleton sets; in the category of augmented algebras over a field
K (where the morphisms are the algebra homomorphisms compatible with the augmentation), K is
a final object; in the category of S-preschemes (I, 2.5.1), S is a final object.
(8.1.11). For two objects X and Y of a category C, set h′X (Y ) = Hom( X, Y ), and for every morphism
u : Y → Y ′ , let h X (u) be the map v 7→ vu from Hom( X, Y ) to Hom( X, Y ′ ); h′X is then a covariant
functor C → Set, so we deduce as in (8.1.2) the definition of a canonical covariant functor h′ : Cop →
Hom(C, Set); a covariant functor F from C to Set, in other words an object of Hom(C, Set), is then
representable if there exists an object X ∈ C (necessarily unique up to unique isomorphism) such that
F is isomorphic to h′X ; we leave it to the reader to develop the “dual” notions of the above, which
this time cover the notion of an inductive limit, and in particular the usual notion of a “solution to a
universal problem”.
8.2. Algebraic structures in categories
(8.2.1). Given two contravariant functors F and F ′ from C to Set, recall that for every object Y ∈ C, 0III | 9
we set ( F × F ′ )(Y ) = F (Y ) × F ′ (Y ), and for every morphism u : Y → Y ′ in C, we set ( F × F ′ )(u) =
F (u) × F ′ (u), which is the map (t, t′ ) 7→ ( F (u)(t), F ′ (u)(t′ )) from F (Y ′ ) × F ′ (Y ′ ) to F (Y ) × F ′ (Y );
F × F ′ : C → Set is thus a contravariant functor (which is none other than the product of the objects F
and F ′ in the category Hom(Cop , Set)). Given an object X ∈ C, we call an internal composition law on X
a natural transformation
(8.2.1.1) γX : h X × h X −→ h X .
In other words (T, 1.2), for every object Y ∈ C, γX (Y ) is a map h X (Y ) × h X (Y ) → h X (Y ) (thus by
definition an internal composition law on the set h X (Y )) with the condition that for every morphism
u : Y → Y ′ in C, the diagram
h X (u)×h X (u)
h X (Y ′ ) × h X (Y ′ ) / h X (Y ) × h X (Y )
γ X (Y ′ ) γ X (Y )
h X (u)
h X (Y ′ ) / h X (Y )
is commutative; this implies that for the composition laws γX (Y ) and γX (Y ′ ), h X (u) is a homomor-
phism from h X (Y ′ ) to h X (Y ).
In a similar way, given two objects Z and X of C, we call an external composition law on X, with Z
as its domain of operators a natural transformation
(8.2.1.2) ωX,Z : h Z × h X −→ h X .
We see as above that for every Y ∈ C, ωX,Z (Y ) is an external composition law on h X (Y ), with
h Z (Y ) as its domain of operators and such that for every morphism u : Y → Y ′ , h X (u) and h Z (u)
form a di-homomorphism from (h Z (Y ′ ), h X (Y ′ )) to (h Z (Y ), h X (Y )).
8. REPRESENTABLE FUNCTORS 67
(8.2.2). Let X ′ be a second object of C, and suppose we are given an internal composition law γX ′ on
X ′ ; we say that a morphism w : X → X ′ in C is a homomorphism for the composition laws if for every
Y ∈ C, hw (Y ) : h X (Y ) → h X ′ (Y ) is a homomorphism for the composition laws γX (Y ) and γX ′ (Y ). If X ′′
is a third object of C equipped with an internal composition law γX ′′ and w′ : X ′ → X ′′ is a morphism 0III | 10
in C which is a homomorphism for γX ′ and γX ′′ , then it is clear that the morphism w′ w : X → X ′′ is
a homomorphism for the composition laws γX and γX ′′ . An isomorphism w : X ≃ X ′ in C is called
an isomorphism for the composition laws γX and γX ′ if w is a homomorphism for these composition
laws, and if its inverse morphism w−1 is a homomorphism for the composition laws γX ′ and γX .
We define in a similar way the di-homomorphisms for pairs of objects of C equipped with external
composition laws.
(8.2.3). When an internal composition law γX on an object X ∈ C is such that γX (Y ) is a group law on
h X (Y ) for every Y ∈ C, we say that X, equipped with this law, is a C-group or a group object in C. We
similarly define C-rings, C-modules, etc.
(8.2.4). Suppose that the product X × X of an object X ∈ C by itself exists in C; by definition, we then
have h X × X = h X × h X up to canonical isomorphism, since it is a particular case of the projective
limit (8.1.9); an internal composition law on X can thus be considered as a functorial morphism
γX : h X × X → h X , and thus canonically determine (8.1.6) an element c X ∈ Hom( X × X, X ) such that
hcX = γX ; in this case, the data of an internal composition law on X is equivalent to the data of a
morphism X × X → X; when C is the category Set, we recover the classical notion of an internal
composition law on a set. We have an analogous result for an external composition law when the
product Z × X exists in C.
(8.2.5). With the above notation, suppose that in addition X × X × X exists in C; the characterization of
the product as an object representing a functor (8.1.9) implies the existence of canonical isomorphisms
( X × X ) × X ≃ X × X × X ≃ X × ( X × X );
if we canonically identity X × X × X with ( X × X ) × X, then the map γX (Y ) × 1hX (Y ) identifies
with hcX ×1X (Y ) for all Y ∈ C. As a result, it is equivalent to say that for every Y ∈ C, the internal law
γX (Y ) is associative, or that the diagram of maps
γX (Y )×1
h X (Y ) × h X (Y ) × h X (Y ) / h X (Y ) × h X (Y )
1 × γ X (Y ) γ X (Y )
γ X (Y )
h X (Y ) × h X (Y ) / h X (Y )
c X ×1 X
X×X×X / X×X
1X ×c X cX
cX
X×X /X
is commutative.
(8.2.6). Under the hypotheses of (8.2.5), if we want to express, for every Y ∈ C, the internal law
γX (Y ) as a group law, then it is first necessary that it is associative, and second that there exists
a map α X (Y ) : h X (Y ) → h X (Y ) having the properties of the inverse operation of a group; as for
every morphism u : Y → Y ′ in C, we have seen that h X (u) must be a group homomorphism
h X (Y ′ ) → h X (Y ), we first see that α X : h X → h X must be a natural transformation. On the other hand,
9. CONSTRUCTIBLE SETS 68
one can express the characteristic properties of the inverse s 7→ s−1 of a group G without involving
the identity element: it suffices to check that the two composite maps
(s, t) 7−→ (s, s−1 , t) 7−→ (s, s−1 t) 7−→ s(s−1 t),
(s, t) 7−→ (s, s−1 , t) 7−→ (s, ts−1 ) 7−→ (ts−1 )s
are equal to the second projection (s, t) 7→ t from G × G to G. By (8.1.3), we have α X = h aX , where
a X ∈ Hom( X, X ); the first condition above then expresses that the composite morphism
(1X ,a X )×1X X X X 1 ×c c
X × X −−−−−−→ X × X × X −−−−
→ X × X −→ X
is the second projection X × X → X in C, and the second condition is similar.
(8.2.7). Now suppose that there exists a final object e (8.1.10) in C. Let us always assume that
γX (Y ) is a group law on h X (Y ) for every Y ∈ C, and denote by ηX (Y ) the identity element of
γX (Y ). As, for every morphism u : Y → Y ′ in C, h X (u) is a group homomorphism, we have
ηX (Y ) = (h X (u))(ηX (Y ′ )); taking in particular Y ′ = e, in which case u is the unique element ε
of Hom(Y, e), we see that the element ηX (e) completely determines ηX (Y ) for every Y ∈ C. Set
eX = ηX ( X ), the identity element of the group h X ( X ) = Hom( X, X ); the commutativity of the
diagram
h X (ε)
h X (e) / h X (Y )
heX (e) h e X (Y )
h X (ε)
h X (e) / h X (Y )
(cf. (8.1.2)) shows that, on the set h X (Y ), the map heX (Y ) is none other than s 7→ ηX (Y ) sending 0III | 12
every element to the identity element. We then verify that the fact that ηX (Y ) is the identity element
of γX (Y ) for every Y ∈ C is equivalent to saying that the composite morphism
(1X ,1X ) X X X1 ×e c
X −−−−→ X × X −−−→ X × X −→ X,
and the analog in which we swap 1X and eX , are both equal to 1X .
(8.2.8). One could of course easily extend the examples of algebraic structures in categories. The
example of groups was treated with enough detail, but latter on we will usually leave it to the reader
to develop analogous notions for the examples of algebraic structures we will encounter.
P ROOF. It is clear that the condition is sufficient. To see that it is necessary, consider the set G of
finite unions of sets of the form U ∩ ∁V, where U and V are retrocompact open sets in X; it suffices
to see that every complement of a set in G is in G. Let Z = i∈ I (Ui ∩ ∁Vi ), where I is finite, Ui and
S
Vi retrocompact open sets in X; we have ∁Z = i∈ I (Vi ∪ ∁Ui ), so Z is a finite union of sets which
T
are intersections of a certain number of the Vi and of a certain number of the ∁Ui , thus of the form
V ∩ ∁U, where U is the union of a certain number of the Ui and V is the intersection of a certain 0III | 13
number of the Vi ; but we have noted above that finite unions and intersections of retrocompact open
sets in X are retrocompact open sets in X, hence the conclusion. □
Corollary (9.1.4). — Every constructible subset of X is retrocompact in X.
P ROOF. It suffices to show that if U and V are retrocompact open sets in X, then U ∩ ∁V is
retrocompact in X; if W is a quasi-compact open set in X, then W ∩ U ∩ ∁V is closed in the quasi-
compact space W ∩ U, hence it is quasi-compact. □
In particular:
Corollary (9.1.5). — For an open subset U of X to be constructible, it is necessary and sufficient for it to be
retrocompact in X. For a closed subset F of X to be constructible, it is necessary and sufficient for the open set
∁F to be retrocompact.
(9.1.6). An important case is when every quasi-compact open subset of X is retrocompact, in other
words, when the intersection of two quasi-compact open subsets of X is quasi-compact (cf. (I, 5.5.6)).
When X is also quasi-compact, this implies that the retrocompact open subsets of X are identical to
the quasi-compact open subsets of X, and the constructible subsets of X are finite unions of sets of
the form U ∩ ∁V, where U and V are quasi-compact open sets.
Corollary (9.1.7). — For a subset of a Noetherian space to be constructible, it is necessary and sufficient for
it to be a finite union of locally closed subsets of X.
Proposition (9.1.8). — Let X be a topological space, U an open subset of X.
(i) If T is a constructible subset of X, then T ∩ U is a constructible subset of U.
(ii) In addition, suppose that U is retrocompact in X. For a subset Z of U to be constructible in X, it is
necessary and sufficient for it to be constructible in U.
P ROOF.
(i) Using Proposition (9.1.3), we reduce to showing that if T is a retrocompact open set in
X, then T ∩ U is a retrocompact open set in U, in other words, for every quasi-compact
open W ⊂ U, T ∩ U ∩ W = T ∩ W is quasi-compact, which immediately follows from the
hypothesis.
(ii) The condition is necessary by (i), so it remains to show that it is sufficient. By Proposition
(9.1.3), it suffices to consider the case where Z is a retrocompact open set in U, because
it will then follow that U − Z is constructible in X, and if Z and Z ′ are two retrocompact
opens in U, then Z ∩ (U − Z ′ ) will be constructible in X. If W is a quasi-compact open
set in X and Z a retrocompact open set in U, then we have Z ∩ W = Z ∩ (W ∩ U ), and by
hypothesis W ∩ U is a quasi-compact open set in U; so W ∩ Z is quasi-compact, and as a
result Z is a retrocompact open set in X, and a fortiori constructible in X.
□
Corollary (9.1.9). — Let X be a topological space, (Ui )i∈ I a finite cover of X by retrocompact open sets in X.
For a subset Z of X to be constructible in X, it is necessary and sufficient for Z ∩ Ui to be constructible in Ui
for all i ∈ I.
(9.1.10). In particular, suppose that X is quasi-compact and every point of X admits a fundamental 0III | 14
system of retrocompact open neighbourhoods in X (and a fortiori quasi-compact); then the condition
for a subset Z of X to be constructible in X is of a local nature, in other words, it is necessary and
sufficient that for every x ∈ X, there exists an open neighbourhood V of x such that V ∩ Z is
constructible in V. Indeed, if this condition is satisfied, then there exists for every x ∈ X an open
neighbourhood V of x which is retrocompact in X and such that V ∩ Z is constructible in V, by the
9. CONSTRUCTIBLE SETS 70
hypotheses on X and by Proposition (9.1.8, i); it then suffices to cover X by a finite number of these
neighbourhoods and to apply Corollary (9.1.9).
Definition (9.1.11). — Let X be a topological space. We say that a subset T of X is locally constructible
in X if for every x ∈ X there exists an open neighbourhood V of x such that T ∩ V is constructible in
V.
It follows from Proposition (9.1.8, i) that if V is such that V ∩ T is constructible in V, then for
every open W ⊂ V, W ∩ T is constructible in W. If T is locally constructible in X, then for every
open set U in X, T ∩ U is locally constructible in U, as a result of the above remark. The same
remark shows that the set of locally constructible subsets of X is stable under finite unions and finite
intersections; on the other hand, it is clear that it is also stable under taking complements.
Proposition (9.1.12). — Let X be a topological space. Every constructible set in X is locally constructible in
X. The converse is true if X is quasi-compact and if its topology admits a basis formed by the retrocompact
sets in X.
P ROOF. The first assertion follows from Definition (9.1.11) and the second from (9.1.10). □
Corollary (9.1.13). — Let X be a topological space whose topology admits a basis formed by the retrocompact
sets in X. Then every locally constructible subset T of X is retrocompact in X.
P ROOF. Let U be a quasi-compact open set in X; T ∩ U is locally constructible in U, hence
constructible in U by Proposition (9.1.12), and as a result quasi-compact by Corollary (9.1.4). □
When X admits a generic point x (0, 2.1.2), the condition of Proposition (9.2.2) is equivalent to
the relation x ∈ E.
Proposition (9.2.3). — Let X be a Noetherian space. For a subset E of X to be constructible, it is necessary
and sufficient that, for every irreducible closed subset Y of X, E ∩ Y is rare in Y or contains a nonempty open
subset of Y.
P ROOF. The necessity of the condition follows from the fact that E ∩ Y must be a constructible
subset of Y and from Proposition (9.2.2), since a nondense subset of Y is necessarily rare in the
irreducible space Y (0, 2.1.1). To prove that the condition is sufficient, apply the principle of
Noetherian induction (0, 2.2.2) to the set F of closed subsets Y of X such that Y ∩ E is constructible
(with respect to Y or with respect to X, which are equivalent): we can thus assume that for every
closed subset Y ̸= X of X, E ∩ Y is constructible. First suppose that X is not irreducible, and let Xi
(1 ⩽ i ⩽ m) are its irreducible components, necessarily of finite number (0, 2.2.5); by hypothesis the
E ∩ Xi are constructible, hence their union E is as well. Suppose now that X is irreducible; then by
hypothesis, if E is rare, then E ̸= X and E = E ∩ E is constructible; if E contains a nonempty open
subset U of X, then it is the union of U and E ∩ ( X − U ); but X − U is a closed set distinct from X,
so E ∩ ( X − U ) is constructible; as a result, E is itself constructible, which finishes the proof. □
9. CONSTRUCTIBLE SETS 71
Corollary (9.2.4). — Let X be a Noetherian space, ( Eα ) an increasing filtered family of constructible subsets
of X, such that
(1st) X is the union of the family ( Eα ).
(2nd) Every irreducible closed subset of X is contained in the closure of one of the Eα .
Then there exists an index α such that X = Eα .
When every irreducible closed subset of X admits a generic point, the hypothesis (1st) can be omitted.
P ROOF. We apply the principle of Noetherian induction (0, 2.2.2) to the set M of closed subsets
of X contained in at least one of the Eα ; we can thus suppose that every closed subset Y ̸= X of X is
contained in one of the Eα . The proposition is evident if X is not irreducible, because each of the
irreducible components Xi of X (1 ⩽ i ⩽ m) is contained in an Eαi , and there exists an Eα containing
all of the Eαi . Now suppose that X is irreducible. By hypothesis, there exists a β such that X = Eβ ,
so (9.2.2) Eβ contains a nonempty open subset U of X. But then the closed set X − U is contained in
an Eγ , and it suffices to take an Eα containing Eβ and Eγ . When every irreducible closed subset Y of
X admits a generic point y, there exists α such that y ∈ Eα , so Y = {y} ⊂ Eα , and condition (2nd) is 0III | 16
therefore a consequence of (1st). □
Proposition (9.2.5). — Let X be a Noetherian space, x a point of X, and E a constructible subset of X. For
E to be a neighbourhood of x, it is necessary and sufficient that for every irreducible closed subset Y of X
containing x, E ∩ Y is dense in Y (if there exists a generic point y of Y, this also implies (9.2.2) that y ∈ E).
P ROOF. The condition is evidently necessary; we will prove that it is sufficient. Applying the
principle of Noetherian induction to the set M of closed subsets Y of X containing x and such that
E ∩ Y is a neighbourhood of x in Y, we can assume that every closed subset Y ̸= X of X containing
x belongs to M. If X is not irreducible, then each of the irreducible components Xi of X containing
x are distinct from X, hence E ∩ Xi is a neighbourhood of x with respect to Xi ; as a result, E is a
neighbourhood of x in the union of the irreducible components of X containing x, and as this union
is a neighbourhood of x in X, so is E. If X is irreducible, then E is dense in X by hypothesis, so it
contains a nonempty open subset U of X (9.2.2); the proposition is then evident if x ∈ U; otherwise,
x is by hypothesis inside E ∩ ( X − U ) with respect to X − U, so the closure of X − E in X does not
contain x, and the complement of this closure is a neighbourhood of x contained in E, which finishes
the proof. □
Corollary (9.2.6). — Let X be a Noetherian space, E a subset of X. For E to be an open set in X, it is
necessary and sufficient that for every irreducible closed subset Y of X intersecting E, E ∩ Y contains a
nonempty open subset of Y.
P ROOF. The condition is evidently necessary; conversely, if it is satisfied, then it implies that
E is constructible by Proposition (9.2.3). In addition, Proposition (9.2.5) shows that E is then a
neighbourhood of each of its points, hence the conclusion. □
9.3. Constructible functions
Definition (9.3.1). — Let h be a map from a topological space X to a set T. We say that h is
constructible if h−1 (t) is constructible for every t ∈ T, and empty except for finitely many values of
t; then for every subset S of T, h−1 (S) is constructible. We say that h is locally constructible if every
x ∈ X has an open neighbourhood V such that h|V is constructible.
Every constructible function is locally constructible; the converse is true when X is quasi-
compact and admits a basis formed by the retrocompact open sets in X (in particular, when X is
Noetherian).
Proposition (9.3.2). — Let h be a map from a Noetherian space X to a set T. For h to be constructible, it is
necessary and sufficient that for every irreducible closed subset Y of X, there exists a nonempty subset U of Y,
open with respect to Y, in which h is constant.
P ROOF. The condition is necessary: indeed, by hypothesis, h does not take finitely many values
ti on Y, and each of the sets h−1 (ti ) ∩ Y is constructible in Y (9.2.1); as they can not all be rare subsets
of the space Y, at least one of them contains a nonempty open set (9.2.3).
10. SUPPLEMENT ON FLAT MODULES 72
To see that the condition is sufficient, we apply the principle of Noetherian induction on the set 0III | 17
M of closed subsets Y of X such that the restriction h|Y is constructible; we can thus assume that for
every closed subset Y ̸= X of X, h|Y is constructible. If X is not irreducible, then the restriction of h
to each of the (finitely many) irreducible components Xi of X is constructible, and it then follows
immediately from Definition (9.3.1) that h is constructible. If X is irreducible, then there exists by
hypothesis a nonempty open subset U of X on which h is constant; on the other hand, the restriction
of h to X − U is constructible by hypothesis, and it follows immediately that h is constructible. □
Corollary (9.3.3). — Let X be a Noetherian space in which every irreducible closed subset admits a generic
point. If h is a map from X to a set T such that, for every t ∈ T, h−1 (t) is constructible, then h is constructible.
P ROOF. If Y is an irreducible closed subset of X and y its generic point, then Y ∩ h−1 (h(y)) is
constructible and contains y, hence (9.2.2) this set contains a nonempty open subset of Y, and it
suffices to apply Proposition (9.3.2). □
Proposition (9.3.4). — Let X be a Noetherian space in which every irreducible closed subset admits a generic
point, h a constructible map from X to an ordered set. For h to be upper semi-continuous on X, it is necessary
and sufficient that for every x ∈ X and every specialization (0, 2.1.2) x ′ of x, we have h( x ′ ) ⩽ h( x ).
P ROOF. The function h does not take a finite number of values; therefore, to say that it is upper
semi-continuous means that for every x ∈ X, the set E of the y ∈ X such that h(y) ⩽ h( x ) is a
neighbourhood of x. By hypothesis, E is a constructible subset of X; on the other hand,to say that an
irreducible closed subset Y of X contains x means that its generic point y is a specialization of x; the
conclusion then follows from Proposition (9.2.5). □
0III-1 | 18
(10.1.2). Suppose that one of the following hypotheses is satisfied:
(i) J is nilpotent;
(ii) A is Noetherian, J is contained in the radical of A, and M is of finite type.
Then the following properties are equivalent.
(a) M is a free A-module.
(b) M/JM = N ⊗ A ( A/J) is a free ( A/J)-module, and Tor1A ( M, A/J) = 0.
(c) M/JM is a free ( A/J)-module, and the canonical homomorphism (10.1.1.2) is injective
(and thus bijective).
(10.1.3). Suppose that A/J is a field (in other words, that J is maximal), and that one of the hypotheses,
(i) and (ii), of (10.1.2) is satisfied. Then the following properties are equivalent.
(a) M is a free A-module.
(b) M is a projective A-module.
(c) M is a flat A-module.
10. SUPPLEMENT ON FLAT MODULES 73
1M ⊗ j 1 M/M ⊗ j
i
M ⊗A N / ( M/Mi ) ⊗ A N
Proposition (10.3.1). — Let A be a local Noetherian ring, with maximal ideal J, and residue field k = A/J.
Let K be a field extension of k. Then there exists a local homomorphism from A to a local Noetherian ring B,
such that B/JB is k-isomorphic to K, and such that B is a flat A-module.
(10.3.1.2). Now suppose that K = k(t) = k[t], where t is algebraic over k; let f ∈ k[ T ] be the minimal
polynomial of t; there exists a monic polynomial F ∈ A[ T ] whose canonical image in k[ T ] is f . So let
A′ = A[ T ], and let J′ be the ideal JA′ + ( F ) in A′ . We will see that the quotient ring B = A′ /( F ) is
that for which we are searching. First of all, it is clear that B is a free A-module, and thus flat. The
ring A′ /J′ is isomorphic to
( A′ /JA′ )/ (JA′ + ( F ))/JA′ = k[ T ]/( f ) = K;
the image L of J′ in B is thus maximal, and we evidently have that L = JB. Finally, B is a semi-local
ring, because it is an A-module of finite type (Bourbaki, Alg. comm., chap. IV, § 2, no 5, cor. 3 of
prop. 9), and its maximal ideals are in bijective correspondence with those of B/JB ([SZ60, vol. I,
p. 259]); the previous arguments then prove that B is a local ring.
Lemma (10.3.1.3). — Let ( Aλ , f µλ ) be a filtered inductive system of local rings, such that the f µλ are local
homomorphisms; let mλ be the maximal ideal of Aλ , and let Kλ = Aλ /mλ . Then A′ = lim Aλ is a local ring,
−→
with maximal ideal m = lim mλ , and residue field K = lim Kλ . Further, if mµ = mλ Aµ with λ < µ, then
−→ −→
we have m′ = mλ A′ for all λ. If, further, for λ < µ, Aµ is a flat Aλ -module, and if all the Aλ are Noetherian,
then A′ is a flat Noetherian Aλ -modules for all λ.
P ROOF. Since, by hypothesis, ( f µ λ)(mλ ) ⊂ mµ for λ < µ, the mλ form an inductive system, and
its limit m′ is evidently an ideal of A′ . Further, if x ′ ̸∈ m′ , there exists a λ such that x ′ = f λ ( xλ ) for
some xλ ∈ Aλ (where f λ : Aλ → A′ denotes the canonical homomorphism); because x ′ ̸∈ m′ , we
necessarily have that xλ ̸∈ mλ , and so xλ admits an inverse yλ in Aλ , and y′ = f λ (yλ ) is the inverse
of x ′ in A′ , which proves that A′ is a local ring with maximal ideal m′ ; the claim about K follows
immediately from the fact that lim is an exact functor. The hypothesis that mµ = mλ Aµ implies that
−→
the canonical map mλ ⊗ Aλ Aµ → mµ is surjective; the equality m′ = mλ A′ then follows from, again,
the fact that the functor lim is exact and commutes with the tensor product.
−→
Now suppose that, for λ < µ, we have mµ = mλ Aµ , and that Aµ is a flat Aλ -module. Then A′ is
a flat Aλ -module for all λ, by (0I , 6.2.3); since A′ and Aλ are local rings, and since m′ = mλ A′ , A′ is
even a faithfully flat Aλ -module (0I , 6.6.2). Finally, suppose further that the Aλ are Noetherian; the
mλ -adic topologies are then separated (0I , 7.3.5); we now show that, from this, it follows that the
′
m′ -adic topology on A′ is separated. Indeed, if x ′ ∈ A′ belongs to all the m n (n ⩾ 0), then it is the
′
image of some xµ ∈ Aµ for a specific index µ, and since the inverse image in Aµ of m n = mnµ A′ is
mnµ (0I , 6.6.1), xµ belongs to all the mnµ , so xµ = 0, by hypothesis, and so x ′ = 0. Denote by A b′ the
completion of A′ for the m′ -adic topology; the above shows that we have A′ ⊂ A b′ . We will now
show that A ′
b is Noetherian and Aλ -flat for all λ; from this, it will follow that A b is A′ -flat (0I , 6.2.3), 0III-1 | 22
′
since mnλ /mnλ+1 is a Kλ -vector space of finite dimension, m′λn /m′λn+1 is a K-vector space of finite
dimension for all n ⩾ 0. It thus follows from (0I , 7.2.12) and (0I , 7.2.8) that A b′ is Noetherian. We
further know that the maximal ideal of A ′ ′ ′
b is m A , and that A ′ ′
b /m A n b is isomorphic to A′ /m′n ; since
′
′ ′ n n ′ ′ ′ n n
A /m = ( Aλ /mλ ) ⊗ Aλ A , we see that A /m is a flat ( Aλ /mλ )-module (0I , 6.2.1); criterion (10.2.2)
is thus applicable to the Noetherian Aλ -algebra A b′ , and shows that A b′ is Aλ -flat. □
(10.3.1.4). We now treat the general case. There exists an ordinal γ and, for every ordinal λ ⩽ γ, a
subfield k λ of K that contains k, such that (i) for all λ < γ, k λ+1 is an extension of k λ generated by
S
a single element; (ii) for every limit ordinal µ, k µ = λ<µ k λ ; and (iii) K = k γ . In fact, it suffices to
consider a bijection ξ 7→ tξ from the set of ordinals ξ ⩽ β (for some suitable β) to K, and to define
k λ by transfinite induction (for λ ⩽ β) as the union of the k µ for µ < λ if λ is a limit ordinal, and
as k ν (tξ ) if λ = ν + 1, where ξ is the smallest ordinal such that tξ ̸∈ k ν ; γ is then, by definition, the
smallest ordinal ⩽ β such that k γ = K.
With this in mind, we will define, by transfinite induction, a family of local Noetherian rings Aλ
for λ ⩽ γ, and local homomorphisms f µλ : Aλ → Aµ for λ ⩽ µ, satisfying the following conditions:
(i) ( Aλ , f µλ ) is an inductive system, and A0 = A;
(ii) for all λ, we have a k-isomorphism Aλ /JAλ ≃ k λ ;
(iii) for λ ⩽ µ, Aµ is a flat Aλ -module.
So suppose that the Aλ and the f µλ are defined for λ < µ < ξ, and suppose, first of all, that
ξ = ζ + 1, so that k ξ = k ζ (t). If t is transcendental over k ζ , we define Aζ , following the procedure of
(10.3.1.1), to be equal to ( Aζ [t])JAζ [t] ; the canonical map is f ζξ , and, for λ < ζ, we take f ξλ = f ξζ ◦ f ζλ ;
the verification of conditions (i) to (iii) is then immediate, given that what we have shown in (10.3.1.1).
So now suppose that t is algebraic, and let h be its minimal polynomial in k ζ [ T ], and H a monic
polynomial in Aζ [ T ] whose image in k ζ [ T ] is h; we then take Aξ to be equal to Aζ [ T ]( H ), with the
f ξλ being defined as before; the verification of conditions (i) to (iii) then follows from what we have
shown in (10.3.1.2).
Now suppose that ξ has no predecessor; we then take Aξ to be the inductive limit of the
inductive system of local rings ( Aλ , f µλ ) for λ < ξ; we define f ξλ as the canonical map for λ < ξ.
The fact that Aξ is local and Noetherian, that the f ξλ are local homomorphism, and that conditions (i)
to (iii) are satisfied for λ ⩽ ξ then follows from the induction hypothesis, and from Lemma (10.3.1.3). 0III-1 | 23
With this construction, it is clear that the ring B = Aγ satisfies the conditions of (10.3.1). □
We note that, by (10.2.1, c), we have a canonical isomorphism
∼
(10.3.1.5) gr( A) ⊗k K −
→ gr( B).
We can also replace B by its JB-adic completion B b without changing the conclusions of (10.3.1),
because B is a flat B-module (0I , 7.3.3), and thus a flat A-module (0I , 6.2.1).
b
We have also shown the following:
Corollary (10.3.2). — If K is an extension of finite degree, then we can assume that B is a finite A-algebra.
pq pq pq
of this quotient identified via αr with the subobjects Bk ( Er+1 ) and Zk ( Er+1 ) respectively.
It is clear that we then have, up to isomorphism,
pq pq pq
(11.1.1.1) Zk ( Er )/Bk ( Er ) = Ek for k ⩾ r + 1,
pq pq pq
and, if we set Br ( Er ) = 0 and Zr ( Er ) = Er , then we have the inclusion relations
(11.1.1.1.2)
pq pq pq pq pq pq pq
0 = Br ( Er ) ⊂ Br+1 ( Er ) ⊂ Br+2 ( Er ) ⊂ · · · ⊂ Zr+2 ( Er ) ⊂ Zr+1 ( Er ) ⊂ Zr ( Er ) = Er .
The other parts of the data of E are then:
pq pq pq pq pq
(d) Two subobjects B∞ ( E2 ) and Z∞ ( E2 ) of E2 such that we have B∞ ( E2 ) ⊂ Z∞ ( E2 ) and,
for every k ⩾ 2,
pq pq pq pq
Bk ( E2 ) ⊂ B∞ ( E2 ) and Z∞ ( E2 ) ⊂ Zk ( E2 ).
We set
pq pq pq
(11.1.1.1.3) E∞ = Z∞ ( E2 )/B∞ ( E2 ).
(e) A family ( En ) of objects of C, each equipped with a decreasing filtration ( F p ( En )) p∈Z . As 0III | 24
usual, we denote by gr( En ) the graded object associated to the filtered object En , the direct
sum of the gr p ( En ) = F p ( En )/F p+1 ( En ).
pq
(f) For every pair ( p, q) ∈ Z × Z, an isomorphism β pq : E∞ ≃ gr p ( E p+q ).
The family ( En ), without the filtrations, is called the abutment (or limit) of the spectral sequence
E.
Suppose that the category C admits infinite direct sums, or that for every r ⩾ 2 and every n ∈ Z,
pq
there are finitely many pairs ( p, q) such that p + q = n and Er ̸= 0 (it suffices for it to hold for r = 2).
(n) pq (n) (n) ( n +1)
Then the Er = ∑ p+q=n Er are defined, and we if denote by dr the morphism Er → Er
pq pq ( n +1) (n)
whose restriction to Er is dr for every pair ( p, q) such that p + q = n, then dr ◦ dr = 0, in
(n) (•)
other words, ( Er )n∈Z is a complex Er in C, with differentials of degree +1, and it follows from (c)
(•) (n)
that Hn ( En ) is isomorphic to Er+1 for every r ⩾ 2.
′ pq
(11.1.2). A morphism u : E → E′ from a spectral sequence E to a spectral sequence E′ = ( Er , E′n )
pq pq ′ pq
consists of systems of morphisms ur : Er → Er and un : En → E′n , the un compatible with the
filtrations on En and E′n , and the diagrams
pq
dr
Er
pq / Erp+r,q−r+1
pq p+r,q−r +1
ur ur
′ pq
dr
Er
′ pq / Er′′ p+r,q−r+1
pq pq pq pq
being commutative; in addition, by passing to quotients, ur gives a morphism ur : Zr+1 ( Zr )/Br+1 ( Er ) →
′ pq ′ pq ′ pq pq pq pq pq pq
Zr+1 ( Er )/Br+1 ( Er ) and we must have αr ◦ ur = ur+1 ◦ αr ; finally, we must have u2 ( B∞ ( E2 )) ⊂
′ pq pq pq ′ pq pq
B∞ ( E2 ) and u2 ( Z∞ ( E2 )) ⊂ Z∞ ( E2 ); by passing to quotients, u2 then gives a morphism
′ pq pq ′ pq
u∞ : E∞ → E∞ , and the diagram
′ pq
u∞
E∞
pq / E∞
′ pq
β pq β′ pq
gr p (u p+q )
gr p ( E p+q ) / gr ( E′ p+q )
p
must be commutative.
pq
The above definitions show, by induction on r, that if the u2 are isomorphisms, then so are the
pq pq pq ′ pq pq pq ′ pq
ur for r ⩾ 2; if in addition we know that u2 ( B∞ ( E2 )) = B∞ ( E2 ) and u2 ( Z∞ ( E2 )) = Z∞ ( E2 ) and
the un are isomorphisms, then we can conclude that u is an isomorphism.
11. SUPPLEMENT ON HOMOLOGICAL ALGEBRA 78
(11.1.3). Recall that if ( F p ( X )) p∈Z is a (decreasing) filtration of an object X ∈ C, then we say that this
filtration is separated if inf( F p ( X )) = 0, discrete if there exists a p such that F p ( X ) = 0, exhaustive (or
coseparated) if sup( F p ( X )) = X, codiscrete if there exists a p such that F p ( X ) = X.
pq pq
We say that a spectral sequence E = ( Er , En ) is weakly convergent if we have B∞ ( E2 ) =
pq pq pq pq pq
supk ( Bk ( E2 )) and Z∞ ( E2 ) = infk ( Zk ( E2 )) (in other words, the objects of B∞ ( E2 ) and Z∞ ( E2 )
are determined from the data of (a) and (c) of the spectral sequence E). We say that the spectral
sequence E is regular if it is weakly convergent and if, in addition:
pq
(1st) For every pair ( p, q), the decreasing sequence ( Zk ( E2 ))k⩾2 is stable; the hypothesis that E
pq pq
is weakly convergent then implies that Z∞ ( E2 ) = Zk ( E2 ) for k large enough (depending
on p and q).
(2nd) For every n, the filtration ( F p ( En )) p∈Z of En is discrete and exhaustive.
We say that the spectral sequence E is coregular if it is weakly convergent and if, in addition:
pq
(3rd) For every pair ( p, q), the increasing sequence ( Bk ( E2 ))k⩾2 is stable, which implies that
pq pq pq pq
B∞ ( E2 ) = Bk ( E2 ), and as a result, E∞ = inf Ek .
n
(4th) For every n, the filtration of E is codiscrete.
Finally, we say that E is biregular if it is both regular and coregular, in other words if we have the
following conditions:
pq pq
(a) For every pair ( p, q), the sequences ( Bk ( E2 ))k⩾2 and ( Zk ( E2 ))k⩾2 are stable and we have
pq pq pq pq pq
B∞ ( E2 ) = Bk ( E2 ) and Z∞ ( E2 ) = Zk ( E2 ) for k large enough (which implies that E∞ =
pq
Ek ).
(b) For every n, the filatration ( F p ( En )) p∈Z is discrete and codiscrete (which we also call finite).
The spectral sequences defined in (T, 2.4) are thus biregular spectral sequences.
(11.1.4). Suppose that in the category C, filtered inductive limits exist and the functor lim is exact
−→
(which is equivalent to saying that the axiom (AB 5) of (T, 1.5) is satisfied (cf. T, 1.8)). The condition
that the filtration ( F p ( X )) p∈Z of an object X ∈ C is exhaustive is then expressed as lim p→−∞ F p ( X ) =
−→
pq pq
X. If a spectral sequence E is weakly convergent, then we have B∞ ( E2 ) = limk→∞ Bk ( E2 ); if in
−→
addition u : E → E′ is a morphism from E to a weakly convergent spectral sequence E′ in C, then we
pq pq ′ pq
have u2 ( B∞ ( E2 )) = B∞ ( E2 ), by the exactness of lim. In addition:
−→
Proposition (11.1.5). — Let C be an abelian category in which filtered inductive limits are exact, E and E′
pq
two regular spectral sequences in C, u : E → E′ a morphism of spectral sequences. If the u2 are isomorphisms,
then so is u.
pq
P ROOF. We already know (11.1.2) that the ur are isomorphisms and that
pq pq ′ pq
u2 ( B∞ ( E2 )) = B∞ ( E2 );
pq pq ′ pq pq
the hypothesis that E and E′ are regular also implies that u2 ( Z∞ ( E2 )) = Z∞ ( E2 ), and as u2 0III | 26
′ pq
is an isomorphism, so is u∞ ; we thus conclude that gr p (u p+q ) is also an isomorphism. But as
the filtrations of the En and the E′n are discrete and exhaustive, this implies that the un are also
isomorphisms (Bourbaki, Alg. comm., chap. III, §2, no 8, th. 1). □
(11.1.6). It follows from (11.1.1.2) and the definition (11.1.1.3) that if, for a spectral sequence E, we
pq pq pq
have Er = 0, then we have Ek = 0 for k ⩾ r and E∞ = 0. We say that a spectral sequence
degenerates if there exists an integer r ⩾ 2 and, for every integer n ∈ Z, an integer q(n) such that
n−q,q n−q,q
Er = 0 for every q ̸= q(n). We first deduce from the previous remark that we also have Ek =0
pq
for k ⩾ r (including k = ∞) and q ̸= q(n). In addition, the definition of Er+1 shows that we have
n−q(n),q(n) n−q(n),q(n) n−q(n),q(n) n−q(n),q(n)
Er+1 = Er ; if E is weakly convergent, then we also have E∞ = Er ; in
n−q(n),q(n)
other words, for every n ∈ Z, gr p ( En )
= 0 for p ̸= q(n) and grq(n) = ( En ) Er
If in addition .
the filtration of En is discrete and exhaustive, then the spectral sequence E is regular, and we have
n−q(n),q(n)
En = Er up to unique isomorphism.
11. SUPPLEMENT ON HOMOLOGICAL ALGEBRA 79
(11.1.7). Suppose that filtered inductive limits exist and are exact in the category C, and let ( Eλ , uµλ )
be an inductive system (over a filtered set of indices) of spectral sequeneces in C. Then the inductive
limit of this inductive system exists in the additive category of spectral sequences of objects of C: to
pq pq pq pq pq
see this, it suffices to define Er , dr , αr , B∞ ( E2 ), Z∞ ( E2 ), En , F p ( En ), and β pq as the respective
pq pq pq pq pq pq
inductive limits of the Er,λ , dr,λ , αr,λ , B∞ ( E2,λ ), Z∞ ( E2,λ ), Eλn , F p ( Eλn ), and β λ ; the verification of the
conditions of (11.1.1) follows from the exactness of the functor lim on C.
−→
Remark (11.1.8). — Suppose that the category C is the category of A-modules over a Noetherian ring
pq
A (resp. a ring A). Then the definitions of (11.1.1) show that if, for a given r, the Er are A-modules
pq pq
of finite type (resp. of finite length), then so are each of the modules Es for s ⩾ r, hence so is E∞ . If in
addition the filtration of the abutment/limit ( En ) is discrete or codiscrete for all n, then we conclude
that each of the En is also an A-module of finite type (resp. of finite length).
(11.1.9). We will have to consider conditions which ensure that a spectral sequence E is biregular is a
“uniform” way in p + q = n. We will then use the following lemma:
pq
Lemma (11.1.10). — Let ( Er ) be a family of objects of C related by the data of (a), (b), and (c) of (11.1.1).
For a fixed integer n, the following properties are equivalent:
(a) There exists an integer r (n) such that for r ⩾ r (n), p + q = n or p + q = n − 1, the morphisms
pq
dr are zero.
pq pq
(b) There exists an integer r (n) such that for p + q = n or p + q = n + 1, we have Br ( E2 ) = Bs ( E2 )
for s ⩾ r ⩾ r (n).
pq pq
(c) There exists an integer r (n) such that for p + q = n or p + q = n − 1, we have Zr ( E2 ) = Zs ( E2 )
for s ⩾ r ⩾ r (n).
pq pq pq
(d) There exists an integer r (n) such that for p + q = n, we have Br ( E2 ) = Bs ( E2 ) and Zr ( E2 ) =
pq
Zs ( E2 ) for s ⩾ r ⩾ r (n).
pq
P ROOF. According to the conditions (a), (b), and (c) of (11.1.1), we have that saying Zr+1 ( E2 ) =
pq pq p+r,q−r +1 p+r,q−r +1
Zr ( E2 ) is equivalent to saying that dr = 0 and that saying Br ( E2 ) = Br+1 ( E2 ) is
pq
equivalent to saying that dr = 0; the lemma immediately follows from this remark. □
11.2. The spectral sequence of a filtered complex
(11.2.1). Given an abelian category C, we will agree to denote by notation such as K • the complexes
(K i )i∈Z of objects of C whose differential is of degree +1, and by the notation such as K• the
complexes (Ki )i∈Z of objects of C whose differential is of degree −1. To each complex K • = (K i )
whose differential d is of degree +1, we can associate a complex K•′ = (Ki′ ) by setting Ki′ = K −i , the
differential Ki′ → Ki′−1 being the operator d : K −i → K −i−1 ; and vice versa, which, depending on the
circumstances, will allow one to consider either one of the types of complexes and translate any
result from one type into results for the other. We similarly denote by notation such as K •• = (K ij )
(resp. K•• = (Kij )) the bicomplexes (or double complexes) of objects of C in which the two differentials
are of degree +1 (resp. −1); we can still pass from one type to the other by changing the signs of
the indices, and we have similar notation and remarks for any multicomplexes. The notation K •
and K• will also be used for Z-graded objects of C, which are not necessarily complexes (they can be
considered as such for the zero differentials); for example, we write H• (K • ) = (Hi (K • ))i∈Z for the
cohomology of a complex K • whose differential is of degree +1, and H• (K• ) = (Hi (K• ))i∈Z for the
homology of a complex K• whose differential is of degree −1; when we pass from K • to K•′ by the
method described above, we have Hi (K•′ ) = H−i (K • ).
Recall in this case that for a complex K • (resp. K• ), we will write in general Zi (K • ) = Ker(K i →
K ) (“object of cocycles”) and Bi (K • ) = Im(K i−1 → K i ) (“object of coboundaries”) (resp. Zi (K• ) =
i + 1
Ker(Ki → Ki−1 ) (“object of cycles”) and Bi (K• ) = Im(Ki+1 → Ki ) (“object of boundaries”)) so that
Hi (K • ) = Zi (K • )/Bi (K • ) (resp. Hi (K• ) = Zi (K• )/Bi (K• )).
If K • = (K i ) (resp. K• = (Ki )) is a complex in C and T : C → C′ a functor from C to an abelian
category C′ , then we denote by T (K • ) (resp. T (K• )) the complex ( T (K i )) (resp. ( T (Ki ))) in C′ .
We will not review the definition of the ∂-functors (T, 2.1), except to note that we also say ∂-functor
in place of ∂∗ -functor when the morphism ∂ decreases the degree of a unit, the context clarifying
this point if there is cause for confusion.
11. SUPPLEMENT ON HOMOLOGICAL ALGEBRA 80
Finally, we say that a graded object ( Ai )i∈Z of C is bounded below (resp. above) if there exists an i0
such that Ai = 0 for i < i0 (resp. i > i0 ).
(11.2.2). Let K • be a complex in C whose differential d is of degree +1, and suppose it is equipped
with a filtration F (K • ) = ( F p (K • )) p∈Z consisting of graded subobjects of K • , in other words, F p (K • ) = 0III | 28
(K i ∩ F p (K • ))i∈Z ; in addition, we assume that d( F p (K • )) ⊂ F p (K • ) for every p ∈ Z. Let us quickly
recall how one functorially defines a spectral sequence E(K • ) from K • (M, XV, 4 and G, I, 4.3). For
r ⩾ 2, the canonical morphism F p (K • )/F p+r (K • ) → F p (K • )/F p+1 (K • ) defines a morphism in
cohomology
H p+q ( F p (K • )/F p+r (K • )) −→ H p+q ( F p (K • )/F p+1 (K • )).
pq
We denote by Zr (K • ) the image of this morphism. Similarly, from the exact sequence
0 −→ F p (K • )/F p+1 (K • ) −→ F p−r+1 (K • )/F p+1 (K • ) −→ F p−r+1 (K • )/F p (K • ) −→ 0,
we deduce from the exact sequence in cohomology a morphism
H p+q−1 ( F p−r+1 (K • )/F p (K • )) −→ H p+q ( F p (K • )/F p+1 (K • )),
pq pq pq
and we denote by Br (K • ) the image of this morphism; we show that Br (K • ) ⊂ Zr (K • ) and we
pq pq pq pq pq
take Er (K • ) = Zr (K • )/Br (K • ); we will not specify the definition of the dr or the αr .
pq • pq •
We note here that all the Zr (K ) and Br (K ), for p and q fixed, are subobjects of the same
pq pq
object H p+q ( F p (K • )/F p+1 (K • )), which we denote by Z1 (K • ); we set B1 (K • ) = 0, so that the above
pq pq pq pq
definitions of Zr (K • ) and Br (K • ) also work for r = 1; we still set E1 (K • ) = Z1 (K • ). We define
pq pq
d1 and α1 such that the conditions of (11.1.1) are satisfied for r = 1. On the other hand, we define
pq
the subobjects Z∞ (K • ) as the image of the morphism
pq
H p+q ( F p (K • )) −→ H p+q ( F p (K • )/F p+1 (K • )) = E1 (K • ),
pq
and B∞ (K • ) as the image of the morphism
pq
H p+q−1 (K • /F p (K • )) −→ H p+q ( F p (K • )/F p+1 (K • )) = E1 (K • ),
pq pq
induced as above from the exact sequence in cohomology. We set Z∞ ( E2 (K • )) and B∞ ( E2 (K • )) to
pq pq pq
be the canonical images of E2 (K • ) in Z∞ (K • ) and B∞ (K • ).
Finally, we denote by F (H (K )) the image in H (K • ) of the morphism Hn ( F p (K • )) → Hn (K • )
p n • n
induced from the canonical injection F p (K • ) → K • ; by the exact sequence in cohomology, this is
also the kernel of the morphism Hn (K • ) → Hn (K • /F p (K • )). This defines a filtration on En (K • ) =
Hn (K • ); we will not give here the definition of the isomorphisms β pq .
(11.2.3). The functorial nature of E(K • ) is understood in the following way: given two filtered com-
plexes K • and K ′• in C and a morphism of complexes u : K • → K ′• that is compatible with the
pq
filtrations, we induce in an evident way the morphisms ur (for r ⩾ 1) and un , and we show that
pq pq
these morphisms are compatible with the dr , αr , and β pq in the sense of (11.1.2), and thus given a
well-defined morphism E(u) : E(K • ) → E(K ′• ) of spectral sequences. In addition, we show that if u
pq pq
and v are morphisms K • → K ′• of the above type, homotopic in degree ⩽ k, then ur = vr for r > k
n n
and u = v for all n (M, XV, 3.1).
(11.2.4). Suppose that filtered inductive limits in C are exact. Then if the filtration ( F p (K • )) of 0III | 29
K • is exhaustive, then so is the filatration ( F p (Hn (K • ))) for all n, since by hypothesis we have
K • = lim p→−∞ F p (K • ) and since the hypothesis on C implies that cohomology commutes with
−→
pq pq
inductive limits. In addition, for the same reason, we have B∞ ( E2 (K • )) = supk Bk ( E2 (K • )). We
say that the filtration ( F p (K • )) of K • is regular if for every n there exists an integer u(n) such that
Hn ( F p (K • )) = 0 for p > u(n). This is particularly the case when the filtration of K • is discrete. When
the filtration of K • is regular and exhaustive, and filtered inductive limits are exact in C, we have
(M, XV, 4) that the spectral sequence E(K • ) is regular.
11. SUPPLEMENT ON HOMOLOGICAL ALGEBRA 81
P ROOF. The proposition follows immediately from the definitions (11.1.3) and from (11.2.4), as
well as from the following observations relating to the filtration FI (and similar observations that we
can deduce for FII by exchange the roles of the two indices in K •• ):
1◦ If there exists i0 such that K ij = 0 for i > i0 , then the filtration FI (K •• ) is discrete.
2◦ If there exists i0 such that K ij = 0 for i < i0 , then the filtration FI (K •• ) is co-discrete. We can
then immediate deduce that it is the same for the corresponding filtration FI (Hn (K •• )) for
pq
all n. Furthermore, the definition of the Br corresponding to the filtration FI (K •• ) (11.2.2)
pq
shows that for any pair ( p, q), the sequence ( Br )r⩽2 is stable.
3◦ If there exists j0 such that K ij = 0 for j < j0 , then we have
∑
p +r
FI (K •• ) ∩ ( K ij ) = 0
i + j=n
pq pq
as soon as p + r + j0 > n, so Zr = Z∞ ( E2 ) for r > q − j0 + 1. On the other hand,
p
Hn ( FI (K •• )) = 0 for p > n − j0 + 1.
4◦ If there exists j0 such that K ij = 0 for j > j0 , then we have
p −r +1
FI (K •• ) ∩ ( ∑ K ij ) = ∑ K ij
i + j=n i + j=n
pq pq
as soon as p − r + 1 + j0 < n, so = Br B∞ ( E2 ) for r < j0 − q + 1. On the other hand,
p
Hn ( FI (K •• )) = Hn (K •• ) for p + j0 < n − 1.
□
(11.3.4). Suppose that the bicomplex K •• = (K ij ) is such that K ij = 0 for i < 0 or j < 0. We know 0III | 31
that we can define, for all p ∈ Z, a canonical “edge-homomorphism”.
′ p0
(11.3.4.1) E2 (K •• ) → H p (K •• )
p0
(M, XV, 6). Recall that this is due to, on the one hand, the spectral sequence ′ E(K •• ), since Zr =
p p +1
ZI ( ZII0 (K •• )) for 2 ⩽ r ⩽ +∞, and on the other hand, that H p ( FI (K •• )) = 0, so that the isomor-
p0 ∼ p p +1 p0 p
phism β p0 : ′ E∞ −
→ H p ( F1 )/H p ( F1 ) gives a homomorphism ′ E∞ → H p ( F1 (K •• )) → H p (K •• ).
p0 p0 p0
The equality of all the Zr allows us to define a canonical homomorphism ′ Er →′ Es for r ⩽ s,
p0 p0
and, in particular, a homomorphism ′ E2 →′ E∞ , from the composition of the edge-homomorphism
′ E p0 ( K •• ) → H p ( K •• ). Furthermore, we can immediately verify that, in the class mod B p0 of
2 2
an element z ∈ Z0II (K •• ) ⊂ K p0 such that d′ z = 0, the edge-homomorphism thus defined as-
p0 p0
sociates, to the class of mod B∞ in ′ E∞ , the cohomology class of z in H p (K •• ). We therefore
finally see that the edge-homomorphism (11.3.4.1) comes from, by passing to cohomology, the canonical
injection ZII0 (K •• ) → K •• (where K •• is considered as simple complex). We naturally interpret the
edge-homomorphism in the same way
′′ p0
(11.3.4.2) E2 (K •• ) → H p (K •• )
as coming from the canonical injection ZI0 (K •• ) → K •• .
(11.3.5). Now let K•• = (Kij ) be a bicomplex in C whose two differential operators are of degrees −1.
We then write Ki,• (resp. K•,j ) to mean the simple complex (Kij ) j∈Z (resp. (Kij )i∈Z ), HIIp (Ki,• )
(resp. HIp (K•,j )) to mean the p-th homology object, HIIp (K•• ) (resp. HIp (K•• )) to mean the complex
(HIIp (Ki,• ))i∈Z (resp. (HIp (K•,j )) j∈Z ), and HIq (HIIp (K•• )) (resp. HIIq (HIp (K•• ))) to mean the q-th homol-
ogy object; we use analogous notation for the objects of cycles and objects of boundaries; finally,
we will denote by Hn (K•• ) the n-th homology object (when it exists) of the simple complex (with
a differential operator of degree −1) defined by K•• . Let K ′•• = (K ij ) with K ′•• = (K−i,− j ) be the
bicomplex with differential operators of degrees +1 associated to K•• . By definition, the spectral
sequences of K•• are those of K ′•• , that we write as ′ E(K•• ) and ′′ E(K•• ), where, however, we change
the notation, by putting
′ r − p,−q − p,−q
E pq (K•• ) =′ Er (K ′•• ) and ′′ r
E pq (K•• ) =′′ Er (K ′•• ),
11. SUPPLEMENT ON HOMOLOGICAL ALGEBRA 83
ε
0 / L0 / L1 / L2 / ...
such that the sequence
→ L0 → L1 . . .
ε
0→A−
is exact. Similarly, a left resolution (or homological resolution) of A is a complex 0 ← L0 ← L1 ← L2 . . .
ε
of objects in C, whose differential operator is of degree −1, endowed with an augmentation L0 − →A
such that the sequence
ε
0←A←
− L0 ← L1 . . .
is exact.
When the right resolution ( Li )i ⩾ 0 of a object A is such that Li = 0 for i ⩾ n + 1, we say that
this resolution is of length ⩽ n. We define a left resolution of length ⩽ n similarly. A resolution which
is of of length ⩽ n for some integer n is said to be finite.
A resolution of A is called projective (resp. injective) if the objects of C, apart from A, of which it is
composed are projective (resp. injective). When C is the category of (left, say) modules over a ring, we
say that a resolution of A is flat (resp. free) when the modules, apart from A, of which it is composed
are flat (resp. free).
(11.4.2). Let K • = (K i )i∈Z be a complex of objects of C, whose differential operator is of degree +1.
We define the right Cartan–Eilenberg resolution of K • to be a pair consisting of a bicomplex L•• = ( Lij )
with differential operator of degree +1 and such that Lij = 0 for j ⩽ 0, and a morphism of simple
complexes ε : K • → L•,0 , such that the following condition are fulfilled: 0III | 33
(i) For each index i, the sequences
0 →Ki −
→ Li0 → Li1 → . . .
ε
0 → Bi ( K • ) −
→ BIi ( L•,0 ) → BIi ( L•,1 ) → . . .
ε
0 → Zi (K • ) −
→ ZIi ( L•,0 ) → ZIi ( L•,1 ) → . . .
ε
0 → Hi ( K • ) −
→ HIi ( L•,0 ) → HIi ( L•,1 ) → . . .
ε
11. SUPPLEMENT ON HOMOLOGICAL ALGEBRA 84
are exact, or, in other words, ( Li,• ), ( BIi ( L•• )), ( ZIi ( L•• )) and (HIi ( L•• )) are resolutions of
K i , Bi (K • ), Zi (K • ) and Hi (K • ) (respectively).
(ii) For each j, the simple complex L•,j is split, or, in other words, the exact sequences
(11.4.3). Now let T be an additive covariant functor from C to an abelian category C′ . Given a complex
K • of C and an injective Cartan–Eilenberg resolution of L•• of K • , suppose that the (simple) complex
defined by the bicomplex T ( L•• ) exists (11.3.1); then the two spectral sequences ′ E( T ( L•• )) and
′′ E ( T ( L•• )) of this bicomplex are called the hypercohomology spectral sequences of T with respect to
K • ; by (11.4.2) and (11.4.3), they effectively only depend on the complex K • effectively, not on the
choice of the injective Cartan–Eilenberg resolution L•• ; furthermore, they depend functorially on K • .
They have the same abutment H• ( T ( L•• )), also called the hypercohomology of T with respect to K • ,
and denoted by R• T (K • ). We can show that the E2 terms of the two spectral sequences above are
given by
′ pq
(11.4.3.1) E2 = H p (Rq T (K • ))
′′ pq
(11.4.3.2) E2 = R p T ( H q (K • ))
where R p T denotes, as usual, the p-th derived functor of T for p ∈ Z; Rq T (K • ) denotes the complexes 0III | 34
(Rq T (K i ))i∈Z . Unless explicitly otherwise mentioned, we will henceforth assume that every object
of C is a subobject of an injective object of C, so that injective Cartan–Eilenberg resolutions exist for
any complex of C. Since Lij = 0 for j < 0, the criteria of (11.3.3) show that the two hypercohomology
spectral sequences of T with respect to K • exist and are biregular in the each of the following two
cases:
(1) K • is bounded below;
(2) Every object of C admits an injective resolution of length at most n, for some integer n
independent of the object in question.
Indeed, we can suppose, in the first case, that (11.4.2) there exists i0 such that Lij = 0 for i < i0 ,
and, in the second case, that there exists j1 such that Lij = 0 for j > j1 ; in each of these two cases, it
is furthermore clear that, for any given n, there exist only a finite number of pairs (i, j) such that
Lij ̸= 0 and i + j = n, which proves our claims.
If we assume that filtrant inductive limits exist in C′ and are exact (which implies, in particular,
the existence of infinite direct sums in C′ ), then the complex defined by the bicomplex T ( L•• ) exists,
and criterion (11.3.3) shows that the sequence ′ E( T ( L•• )) is always regular.
If K • is a complex such that all the K i are zero except for a single K i0 , then Rn T (K • ) is isomorphic
to R −i0 T (K • ), as follows immediately from the definitions by taking a Cartan–Eilenberg resolution
n
If K • and K ′• are two complexes of C, and f and g are homotopic morphism from K • to K ′• , then
the morphisms R• T (K • ) → R• T (K ′• ) induced by f and g are identical, and the same is true for the
morphisms of the cohomology spectral sequences.
Proposition (11.4.5). — Suppose that filtrant inductive limits exist in C′ and are exact. If Rn T (K i ) = 0 for
all n > 0 and all i ∈ Z, then we have functorial isomorphisms
∼
(11.4.5.1) Ri T ( K • ) −
→ Hi ( T (K • ))
for all i ∈ Z.
P ROOF. Indeed, the only non-zero E2 terms of the first spectral sequence (11.4.3.1) are then
′ E p0
= H p ( T (K • )); in other words, this sequence is degenerate; since it is regular (11.4.4), the
2
conclusion follows from (11.1.6). □
(11.4.6). Now consider, for example, a covariant bifunctor ( M, N ) → T ( M, N ) from C × C to C′′ ,
where C, C′ , and C′′ are three abelian categories; we assume, for simplicity, that T is additive in
each of its arguments, and furthermore that every object of C and every object of C′ are subobjects
of an injective object, and that filtrant inductive limits exist in C and are exact. We then define
the hypercohomology of T with respect to the complexes K • and K ′• of C and C′ (respectively), with
differential operators of degree +1, by taking K • (resp. K ′• ) to be an injective Cartan–Eilenberg
resolution L•• (resp. L′•• ); then T ( L•• , L′•• ) is a quadricomplex of C′′ , which we consider as a
bicomplex of C′ , with the degree of T ( Lij , L′hk ) being the integers i + h and j + k. The hypercohomology
of T with respect to K • and K ′• is by definition the cohomology H• ( T ( L•• , L′•• )) of this bicomplex
(in other words, that of the associated simple complex) and is denoted by R• T (K • , K ′• ); it is the 0III | 35
abutment of two spectral sequences whose E2 terms are given by
′ pq
(11.4.6.1) E2 = H p (Rq T (K • , K ′• ))
′ ′′
(11.4.6.2) ′′ pq
E2 = ∑′′ R p T (Hq K • , Hq K ′• )
q′ +q =q
hand, are identical, and the same is true for the morphisms of the hypercohomology spectral
sequences.
We can generalize easily to any additive covariant multifunctor.
Proposition (11.4.7). — Suppose that for any injective object I of C (resp. I′ of C′ ), A′ 7→ T (I, A′ ) (resp.
A 7→ T (A, I′ )) is an exact functor. Then, with the notation of (11.4.6), we have a canonical isomorphism
∼ ∼
R• T (K • , K ′• ) −
→ H• ( T ( L•• , K ′• )) −
→ H• ( T (K • , L′•• ))
where the last two terms are the cohomology of simple complexes defined by the tricomplexes T ( L•• , K ′• ) and
T (K • , L•• ) (respectively).
P ROOF. Let us define, for example, the first of these isomorphisms. The quadricomplex
T ( L , L′•• )
•• can be considered as a bicomplex, where the degrees of T ( Lij , L′hk ) are the integers i + j
h,• h
and h + k. Since, for each h, L′ is a resolution of K ′ , we have, for this bicomplex, by virtue of
q
the hypotheses on T, HII ( T ( L•• , L′•• )) = 0 for q ̸= 0, and H0II ( T ( L•• , L′•• )) = T ( L•• , K ′• ); the first
spectral sequence of this bicomplex is thus degenerate; since L′hk = 0 for k < 0, this sequence is also
regular (11.3.3), and the conclusion then follows from (11.1.6). □
14. COMBINATORIAL DIMENSION OF A TOPOLOGICAL SPACE 86
We have similar results for a covariant multifunctor of any number n of arguments: in the
calculation of the hypercohomology, it is not necessary to replace all the complexes by a Cartan–
Eilenberg resolution, but only n − 1 of them, provided that, when we fix n − 1 arbitrary arguments
by taking them to be injective objects, the covariant functor in the remaining argument is exact.
11.5. Passage to the inductive limit in the hypercohomology 0III | 36
11.6. Hypercohomology of a functor with respect to a complex K•
11.7. Hypercohomology of a functor with respect to a bicomplex K••
11.8. Supplement on the cohomology of simplicial complexes
11.9. A lemma on the complexes of finite type
11.10. Euler-Poincare characteristic of a complex of finite length modules
Summary
Almost all of the preceding sections have been focused on the exposition of ideas of commutative
algebra that will be used throughout Chapter IV. Even though a large amount of these ideas already
appear in multiple works ([CC], [Sam53a], [SZ60], [Ser55b], [Nag62]), we thought that it would be
more practical for the reader to have a coherent, vaguely independent exposition. Together with
§§5, 6, and 7 of Chapter IV (where we use the language of schemes), these sections constitute, in
the middle of our treatise, a miniature special treatise, somewhat independent of Chapters I to III,
and one that aims to present, in a coherent manner, the properties of rings that “behave well” with
respect to operations such as completion, or integral closure, by systematically associating these
properties to more general ideas.11
11The majority of properties which we discuss were discovered by Chevalley, Zariski, Nagata, and Serre. The method
used here was first developed in the autumn of 1961, in a course taught at Harvard University by A. Grothendieck.
14. COMBINATORIAL DIMENSION OF A TOPOLOGICAL SPACE 87
because every chain of irreducible closed subsets of X is, by definition, contained in some irreducible
component of X, and, conversely, the irreducible components are closed in X, so every irreducible
closed subset of an Xα is a irreducible closed subset of X.
Definition (14.1.3). — We say that a topological space X is equidimensional if all its irreducible
components have the same dimension (which is thus equal to dim( X ), by (14.1.2.1)).
Proposition (14.1.4). —
(i) For every closed subset Y of a topological space X, we have dim(Y ) ⩽ dim( X ).
(ii) If a topological space X is a finite union of closed subsets Xi , then we have dim( X ) = supi dim( Xi ).
P ROOF. For every irreducible closed subset Z of Y, the closure Z of Z in X is irreducible (0I , 2.1.2),
and Z ∩ Y = Z, whence (i). Now, if X = in=1 Xi , where the Xi are closed, then every irreducible
S
closed subset of X is contained in one of the Xi (0I , 2.1.1), and so every chain of irreducible closed
subsets of X is contained in one of the Xi , whence (ii). □
From (14.1.4, i), we see that, for all x ∈ X, we can also write
(14.1.4.1) dimx ( X ) = lim dim(U ),
U
where the limit is taken over the downward-directed set of open neighbourhoods of x in X.
0IV-1 | 7
Corollary (14.1.5). — Let X be a topological space, x a point of X, U a neighbourhood of x, and Yi
(1 ⩽ i ⩽ n) closed subsets of U such that, for all i, x ∈ Yi , and such that U is the union of the Yi . Then we
have
(14.1.5.1) dimx ( X ) = sup(dimx (Yi )).
i
P ROOF. It follows from (14.1.4, ii) that we have dimx ( X ) = infV (supi (dim(Yi ∩ V ))), where
V ranges over the set of open neighbourhoods of x that are contained in U; similarly, we have
dimx (Yi ) = infV (dim(Yi ∩ V )) for all i. The corollary is thus evident if
sup(dimx (Yi )) = +∞;
i
if this were not the case, then there would be an open neighbourhood V0 ⊂ U of x such that
dim(Yi ∩ V ) = dimx (Yi ) for 1 ⩽ i ⩽ n and for all V ⊂ V0 , whence the conclusion. □
Proposition (14.1.6). — For every topological space X, we have dim( X ) = supx∈X dimx ( X ).
14. COMBINATORIAL DIMENSION OF A TOPOLOGICAL SPACE 88
P ROOF. It follows from Definition (14.1.2) and Proposition (14.1.4) that dimx ( X ) ⩽ dim( X ) for
all x ∈ X. Now, let Z0 ⊂ Z1 ⊂ . . . ⊂ Zn be a chain of irreducible closed subsets of X, and let x ∈ Z0 ;
for every open subset U ⊂ X that contains x, U ∩ Zi is irreducible (0I , 2.1.6) and closed in U, and
since we have U ∩ Zi = Zi in X, the U ∩ Zi are pairwise distinct; thus dim(U ) ⩾ n, which finishes
the proof. □
Corollary (14.1.7). — If ( Xα ) is an open, or closed and locally finite, cover of X, then dim( X ) =
supα (dim( Xα )).
P ROOF. If Xα is a neighbourhood of x ∈ X, then dimx ( X ) ⩽ dim( Xα ), whence the claim for
open covers. On the other hand, if the Xα are closed, and U is a neighbourhood of x ∈ X which
meets only finitely many of the Xα , then
dimx ( X ) ⩽ dim(U ) = sup(dim(U ∩ Xα )) ⩽ sup(dim( Xα ))
α α
by (14.1.4), whence the other claim. □
Corollary (14.1.8). — Let X be a Noetherian Kolmogoroff space (0I , 2.1.3), and F the set of closed points of
X. Then dim( X ) = supx∈ F dimx ( X ).
P ROOF. With the notation from the proof of (14.1.6), it suffices to note that there exists a closed
point in Z0 (0I , 2.1.3). □
Proposition (14.1.9). — Let X be a nonempty Noetherian Kolmogoroff space. To have dim( X ) = 0, it is
necessary and sufficient for X to be finite and discrete.
P ROOF. If a space X is separated (and a fortiori if X is a discrete space), then all the irreducible
closed subsets of X are single points, and so dim( X ) = 0. Conversely, suppose that X is a Noetherian
Kolmogoroff space such that dim( X ) = 0; since every irreducible component of X contains a closed
point (0I , 2.1.3), it must be exactly this single point. Since X has only a finite number of irreducible
components, it is thus finite and discrete. □
Corollary (14.1.10). — Let X be a Noetherian Kolmogoroff space. For a point x ∈ X to be isolated, it is
necessary and sufficient to have dimx ( X ) = 0.
P ROOF. The condition is clearly necessary (without any hypotheses on X). It is also sufficient,
because it implies that dim(U ) = 0 for any open neighbourhood U of x, and since U is a Noetherian 0IV-1 | 8
Kolmogoroff space, U is finite and discrete. □
Proposition (14.1.11). — The function x 7→ dimx ( X ) is upper semi-continuous on X.
P ROOF. It is clear that this function is upper semi-continuous at every point where its value is
+∞. So suppose that dimx ( X ) = n < +∞; then Equation (14.1.4.1) shows that there exists an open
neighbourhood U0 of x such that dim(U ) = n for every open neighbourhood U ⊂ U0 of x. So, for
all y ∈ U0 and every open neighbourhood V ⊂ U0 of y, we have dim(V ) ⩽ dim(U0 ) = n (14.1.4);
we thus deduce from (14.1.4.1) that dimy ( X ) ⩽ n. □
Remark (14.1.12). — If X and Y are topological spaces, and f : X → Y a continuous map, then it
can be the case that dim( f ( X )) > dim( X ); we obtain such an example by taking X to be a discrete
space with 2 points, a and b, and Y to be the set { a, b} endowed with the topology for which12
the closed sets are ∅, { a}, and { a, b}; if f : X → Y is the identity map, then dim(Y ) = 1 and
dim( X ) = 0. We note that Y is the spectrum of a discrete valuation ring A, of which a is the unique
closed point, and b the generic point; if K and k are the field of fractions and the residue field of A
(respectively), then X is the spectrum of the ring k × K, and f is the continuous map corresponding
to the homomorphism (ϕ, ψ) : A → k × K, where ϕ : A → k and ψ : A → K are the canonical
homomorphism (cf. (IV, 5.4.3)).
12[Trans.] This is now often referred to as the Sierpiński space, or the connected two-point set.
14. COMBINATORIAL DIMENSION OF A TOPOLOGICAL SPACE 89
0IV-1 | 9
(ii) For every nonempty closed subset Y of X, we have
(14.2.2.2) dim(Y ) + codim(Y, X ) ⩽ dim( X ).
(iii) If Y, Z, and T are closed subsets of X such that Y ⊂ Z ⊂ T, then
(14.2.2.3) codim(Y, Z ) + codim( Z, T ) ⩽ codim(Y, T ).
(iv) For a closed subset Y of X to be such that codim(Y, X ) = 0, it is necessary and sufficient for Y to
contain an irreducible component of X.
P ROOF. Claims (i) and (iv) are immediate consequences of Definition (14.2.1). To show (ii), we
can restrict to the case where Y is irreducible, and then the equation follows from Definitions (14.1.1)
and (14.2.1). Finally, to show (iii), we can, by Definition (14.2.1), first restrict to the case where Y is
irreducible; then codim(Y, Z ) = supα (codim(Y, Zα )) for the irreducible components Zα of Z that
contain Y; it is clear that codim(Y, T ) ⩾ codim(Y, Z ), so the inequality is true if codim(Y, Z ) = +∞;
but if this were not the case, then there would exist some α such that codim(Y, Z ) = codim(Y, Zα ),
and by (14.2.1), we can restrict to the case where Z itself is irreducible; but then the inequality in
(14.2.2.3) is an evident consequence of Definition (14.2.1). □
Proposition (14.2.3). — Let X be a topological space, and Y a closed subset of X. For every open subset U
of X, we have
(14.2.3.1) codim(Y ∩ U, U ) ⩾ codim(Y, X ).
Furthermore, for this inequality (14.2.3.1) to be an equality, it is necessary and sufficient to have
codim(Y, X ) = infα (codim(Yα , X )), where (Yα ) is the family of irreducible components of Y that meet U.
P ROOF. We know (0I , 2.1.6) that Z 7→ Z is a bijection from the set of irreducible closed subsets
of U to the set of irreducible closed subsets of X that meet U, and, in particular, induces a correspon-
dence between the irreducible components of Y ∩ U and the irreducible components of Y that meet
U; if Yα is one of the latter such components, then we have codim(Yα , X ) = codim(Yα ∩ U, U ), and
the proposition then follows from Definition (14.2.1). □
Definition (14.2.4). — Let X be a topological space, Y a closed subset of X, and x a point of X.
We define the codimension of Y in X at the point x, denoted by codimx (Y, X ), to be the number
supU (codim(Y ∩ U, U )), where U ranges over the set of open neighbourhoods of x in X.
14. COMBINATORIAL DIMENSION OF A TOPOLOGICAL SPACE 90
(a) Any two maximal chains of irreducible closed subsets of X have the same length.
(b) X is equidimensional, equicodimensional, and catenary.
(c) X is equidimensional, and, for any irreducible closed subsets Y and Z of X with Y ⊂ Z, we have
(14.3.3.1) dim( Z ) = dim(Y ) + codim(Y, Z ).
(d) X is equicodimensional, and, for any irreducible closed subsets Y and Z of X with Y ⊂ Z, we have
(14.3.3.2) codim(Y, Z ) = codim(Y, Z ) + codim( Z, X ).
P ROOF. The hypotheses on X imply that the first and last elements of a maximal chain of
irreducible closed subsets of X are necessarily a closed point and an irreducible component of X
(respectively) (0I , 2.1.3); further, every saturated chain with first element Y and last element Z (thus
Y ⊂ Z) is contained in a maximal chain whose elements differ from those of the given chain, or are
contained in Y, or contain Z. These remarks immediately establish the equivalence between (a) and
(b), and also show that if (a) is satisfied, then we have, for every irreducible closed subset Y to X,
(14.3.3.3) dim(Y ) + codim(Y, X ) = dim( X );
from (14.3.2.1), we immediately deduce (14.3.3.1) and (14.3.3.2) from (14.3.3.3). Conversely, (14.3.3.1)
implies (14.3.2.1), and so (14.3.3.1) implies the chain condition, by (14.3.2); further, by applying
(14.3.3.1) to the case where Y is a single closed point x of X, and Z is an irreducible component of
X, we get that codim({ x }, X ) = dim( Z ); we thus conclude that (c) implies (b). Similarly, (14.3.3.2)
implies (14.3.2.1), and thus the chain condition; further, with the same choice of Y and Z as above,
(14.3.3.2) again implies that codim({ x }, X ) = dim( Z ), and so (since every irreducible component of
X contains a closed point, by (0I , 2.1.3)), (d) implies (b). □
We say that a Noetherian Kolmogoroff space is biequidimensional if it is of finite dimension and if
it verifies any of the (equivalent) conditions of (14.3.3).
Corollary (14.3.4). — Let X be a biequidimensional Noetherian Kolmogoroff space; then, for every closed
point x of X, and every irreducible component Z of X, we have
(14.3.4.1) dim( X ) = dim( Z ) = codim({ x }, X ) = dimx ( X ).
0IV-1 | 12
P ROOF. The last equality follows from the fact that, if Y0 = { x } ⊂ Y1 ⊂ · · · ⊂ Ym is a maximal
chain of irreducible closed subsets of X, and U an open neighbourhood of x, then the U ∩ Yi are
pairwise disjoint irreducible closed subsets of U (because U ∩ Yi = Yi ), whence dim(U ) = dim( X ),
by (14.1.4). □
Corollary (14.3.5). — Let X be a Noetherian Kolmogoroff space; if X is biequidimensional, then so is every
union of irreducible components of X, and every irreducible closed subset of X. In addition, for every closed
subset Y of X, we have
(14.3.5.1) dim(Y ) + codim(Y, X ) = dim( X ).
P ROOF. Every chain of irreducible closed subsets of X is contained in an irreducible component
of X, and so the first claim follows immediately from (14.3.3). Further, if X ′ is an irreducible closed
subset of X, then X ′ trivially satisfies the conditions of (14.3.3, c), whence the second claim.
Finally, to show (14.3.5.1), note that we have seen, in the proof of (14.3.3), that this equation is
true whenever Y is irreducible; if Yi (1 ⩽ i ⩽ m) are the irreducible components of Y, then the Yi
for which dim(Yi ) is the largest are also those for which codim(Yi , X ) is the smallest; so (14.3.5.1)
follows from the definitions of dim(Y ) and codim(Y, X ). □
Remark (14.3.6). — The reader will note that the proof of (14.3.2) applies to any ordered set, and the
fact that we are working with the example of a set of irreducible closed subsets of a topological space
is not used at all in the proof. It is the same in the proof of (14.3.3), which holds, more generally, for
any ordered set E such that, for all x ∈ E, there exists some z ⩽ x which is minimal in E, and such
that the length of chains of elements of E is bounded.
CHAPTER I
S UMMARY
In §§1–8 we do little more than develop a language to be used in what follows. It should be I | 79
noted, however, that, in accordance with the general spirit of this treatise, §§7–8 will be used less
than the others, and in a less essential way; we speak of Chevalley schemes only for the purpose of
relating to the language of Chevalley [CC] and Nagata [Nag58a]. Then, in §9, we give definitions
and results concerning quasi-coherent sheaves, some of which are no longer simply a translation
of known notions of commutative algebra into a “geometric” language, but are instead already of
a global nature; they will be indispensable, in the following chapters, when it comes to the global
study of morphisms. Finally, in §10, we introduce a generalization of the notion of a scheme, which
will be used as an intermediary in Chapter III to conveniently formulate and prove the fundamental
results of the cohomological study of proper morphisms; moreover, it should be noted that the
notion of formal schemes seems indispensable in expressing certain facts about the “theory of
modules” (classification problems of algebraic varieties). The results of §10 will not be used before
§3 of Chapter III, and it is recommended to skip their reading until then.
(1.1.1). Notation. Let A be a (commutative) ring, and M an A-module. In this chapter and the I | 80
following, we will constantly use the following notation:
• Spec( A) = set of prime ideals of A, also called the prime spectrum of A; for x ∈ X = Spec( A),
it will often be convenient to write jx instead of x. For Spec( A) to be empty, it is necessary
and sufficient for the ring A to be 0.
• A x = Ajx = (local) ring of fractions S−1 A, where S = A − jx .
• mx = jx Ajx = maximal ideal of A x .
• k( x ) = A x /mx = residue field of A x , canonically isomorphic to the field of fractions of the
integral ring A/jx , with which we identify it.
• f ( x ) = class of f mod. jx in A/jx ⊂ k( x ), for f ∈ A and x ∈ X. We also say that f ( x ) is the
value of f at a point x ∈ Spec( A); the equations f ( x ) = 0 and f ∈ jx are equivalent.
• Mx = M ⊗ A A x = module of fractions with denominators in A − jx .
• r( E) = radical of the ideal of A generated by a subset E of A.
92
1. AFFINE SCHEMES 93
• V ( E) = set of x ∈ X such that E ⊂ jx (or the set of x ∈ X such that f ( x ) = 0 for all f ∈ E),
for E ⊂ A. So we have
\
(1.1.1.1) r( E ) = jx .
x ∈V ( E )
• V ( f ) = V ({ f }) for f ∈ A.
• D ( f ) = X − V ( f ) = set of x ∈ X where f ( x ) ̸= 0.
Proposition (1.1.2). — We have the following properties:
(i) V (0) = X, V (1) = ∅.
(ii) The relation E ⊂ E′ implies V ( E) ⊃ V ( E′ ).
(iii) For each family ( Eλ ) of subsets of A, V ( λ Eλ ) = V (∑λ Eλ ) = λ V ( Eλ ).
S T
(iv) V ( EE′ ) = V ( E) ∪ V ( E′ ).
(v) V ( E) = V (r( E)).
P ROOF. The properties (i), (ii), (iii) are trivial, and (v) follows from (ii) and from equation
(1.1.1.1). It is evident that V ( EE′ ) ⊃ V ( E) ∩ V ( E′ ); conversely, if x ̸∈ V ( E) and x ̸∈ V ( E′ ), then
there exists f ∈ E and f ′ ∈ E′ such that f ( x ) ̸= 0 and f ′ ( x ) ̸= 0 in k( x ), hence f ( x ) f ′ ( x ) ̸= 0, i.e.,
x ̸∈ V ( EE′ ), which proves (iv). □
Proposition (1.1.2) shows, among other things, that sets of the form V ( E) (where E varies over
the subsets of A) are the closed sets of a topology on X, which we will call the spectral topology 1; unless
expressly stated otherwise, we always assume that X = Spec( A) is equipped with the spectral
topology.
(1.1.3). For each subset Y of X, we denote by j(Y ) the set of f ∈ A such that f (y) = 0 for all y ∈ Y; I | 81
equivalently, j(Y ) is the intersection of the prime ideals jy for y ∈ Y. It is clear that the relation
Y ⊂ Y ′ implies that j(Y ) ⊃ j(Y ′ ) and that we have
[ \
(1.1.3.1) j Yλ = j(Yλ )
λ λ
for each family (Yλ ) of subsets of X. Finally we have
(1.1.3.2) j({ x }) = jx .
Proposition (1.1.4). —
(i) For each subset E of A, we have j(V ( E)) = r( E).
(ii) For each subset Y of X, V (j(Y )) = Y, the closure of Y in X.
P ROOF. (i) is an immediate consequence of the definitions and (1.1.1.1); on the other hand,
V (j(Y )) is closed and contains Y; conversely, if Y ⊂ V ( E), we have f (y) = 0 for f ∈ E and all y ∈ Y,
so E ⊂ j(Y ), V ( E) ⊃ V (j(Y )), which proves (ii). □
Corollary (1.1.5). — The closed subsets of X = Spec( A) and the ideals of A equal to their radicals (in
other words, those that are the intersection of prime ideals) correspond bijectively by the inclusion-reversing
maps Y 7→ j(Y ), a 7→ V (a); the union Y1 ∪ Y2 of two closed subsets corresponds to j(Y1 ) ∩ j(Y2 ), and the
intersection of any family (Yλ ) of closed subsets corresponds to the radical of the sum of the j(Yλ ).
Corollary (1.1.6). — If A is a Noetherian ring, X = Spec( A) is a Noetherian space.
Note that the converse of this corollary is false, as shown by any non-Noetherian integral ring
with a single prime ideal ̸= {0} (for example a nondiscrete valuation ring of rank 1).
As an example of ring A whose spectrum is not a Noetherian space, one can consider the
ring C (Y ) of continuous real functions on an infinite compact space Y; we know that, as a set, Y
corresponds to the set of maximal ideals of A, and it is easy to see that the topology induced on Y by
that of X = Spec( A) is the original topology of Y. Since Y is not a Noetherian space, the same is
true for X.
1The introduction of this topology in algebraic geometry is due to Zariski. So this topology is usually called the “Zariski
topology” on X.
1. AFFINE SCHEMES 94
Corollary (1.1.7). — For each x ∈ X, the closure of { x } is the set of y ∈ X such that jx ⊂ jy . For { x } to be
closed, it is necessary and sufficient that jx is maximal.
Corollary (1.1.8). — The space X = Spec( A) is a Kolmogoroff space.
P ROOF. If x and y are two distinct points of X, we have either jx ̸⊂ jy or jy ̸⊂ jx , so one of the
points x, y does not belong to the closure of the other. □
Proposition (1.1.10). — I | 82
(i) When f ranges over A, the sets D ( f ) forms a basis for the topology of X.
(ii) For every f ∈ A, D ( f ) is quasi-compact. In particular, X = D (1) is quasi-compact.
P ROOF.
(i) Let U be an open set in X; by definition, we have U = X − V ( E) where E is a subset of A,
T S
and V ( E) = f ∈E V ( f ), hence U = f ∈E D ( f ).
(ii) By (i), it suffices to prove that, if ( f λ )λ∈ L is a family of elements of A such that D ( f ) ⊂
λ∈ L D ( f λ ), then there exists a finite subset J of L such that D ( f ) ⊂
S S
λ∈ J D ( f λ ). Let
a be the ideal of A generated by the f λ ; we have, by hypothesis, that V ( f ) ⊃ V (a),
so r( f ) ⊂ r(a); since f ∈ r( f ), there exists an integer n ⩾ 0 such that f n ∈ a. But
then f n belongs to the ideal b generated by the finite subfamily ( f λ )λ∈ J , and we have
V ( f ) = V ( f n ) ⊃ V (b) = λ∈ J V ( f λ ), that is to say, D ( f ) ⊃ λ∈ J D ( f λ ).
T S
Proposition (1.1.11). — For each ideal a of A, Spec( A/a) is canonically identified with the closed subspace
V (a) of Spec( A).
P ROOF. We know there is a canonical bijective correspondence (respecting the inclusion order
structure) between ideals (resp. prime ideals) of A/a and ideals (resp. prime ideals) of A containing
a. □
Recall that the set N of nilpotent elements of A (the nilradical of A) is an ideal equal to r(0), the
intersection of all the prime ideals of A (0, 1.1.1).
Corollary (1.1.12). — The topological spaces Spec( A) and Spec( A/N) are canonically homeomorphic.
Proposition (1.1.13). — For X = Spec( A) to be irreducible (0, 2.1.1), it is necessary and sufficient that the
ring A/N is integral (or, equivalently, that the ideal N is prime).
P ROOF. By virtue of Corollary (1.1.12), we can restrict to the case where N = 0. If X is reducible,
then there exist two distinct closed subsets Y1 and Y2 of X such that X = Y1 ∪ Y2 , so j( X ) =
j(Y1 ) ∩ j(Y2 ) = 0, since the ideals j(Y1 ) and j(Y2 ) are distinct from (0) (1.1.5); so A is not integral.
Conversely, if there are elements f ̸= 0, g ̸= 0 of A such that f g = 0, we have V ( f ) ̸= X, V ( g) ̸= X
(since the intersection of all the prime ideals of A is (0)), and X = V ( f g) = V ( f ) ∪ V ( g). □
Corollary (1.1.14). —
(i) In the bijective correspondence between closed subsets of X = Spec( A) and ideals of A equal to
their radicals, the irreducible closed subsets of X correspond to the prime ideals of A. In particular,
the irreducible components of X correspond to the minimal prime ideals of A.
(ii) The map x 7→ { x } establishes a bijective correspondence between X and the set of closed irreducible
subsets of X (in other words, all closed irreducible subsets of X admit exactly one generic point).
1. AFFINE SCHEMES 95
P ROOF. (i) follows immediately from (1.1.13) and (1.1.11); and for proving (ii), we can, by
(1.1.11), restrict to the case where X is irreducible; then, according to Proposition (1.1.13), there exists
a smaller prime ideal N in A, which corresponds to the generic point of X; in addition, X admits at I | 83
most one generic point since it is a Kolmogoroff space ((1.1.8) and (0, 2.1.3)). □
Proposition (1.1.15). — If J is an ideal in A containing the radical N( A), the only neighbourhood of V (J)
in X = Spec( A) is the whole space X.
P ROOF. Each maximal ideal of A belongs, by definition, to V (J). As each ideal a ̸= A of A is
contained in a maximal ideal, we have V (a) ∩ V (J) ̸= 0, whence the proposition. □
1.2. Functorial properties of prime spectra of rings
(1.2.1). Let A, A′ be two rings, and
ϕ : A′ −→ A
a homomorphism of rings. For each prime ideal x = jx ∈ Spec( A) = X, the ring A′ /ϕ−1 (jx ) is
canonically isomorphic to a subring of A/jx , and so it is integral, or, in other words, ϕ−1 (jx ) is a
prime ideal of A′ ; we denote it by a ϕ( x ), and we have thus defined a map
a
ϕ : X = Spec( A) −→ X ′ = Spec( A′ )
(also denoted Spec(ϕ)), that we call the map associated to the homomorphism ϕ. We denote by ϕ x
the injective homomorphism from A′ /ϕ−1 (jx ) to A/jx induced by ϕ by passing to quotients, as well
as its canonical extension to a monomorphism of fields
ϕ x : k ( a ϕ( x )) −→ k ( x );
for each f ′ ∈ A′ , we therefore have, by definition, ErrII
x ′ a ′
(1.2.1.1) ϕ ( f ( ϕ( x ))) = (ϕ( f ))( x ) ( x ∈ X ).
Proposition (1.2.2). —
(i) For each subset E′ of A′ , we have
a −1
(1.2.2.1) ϕ (V ( E′ )) = V (ϕ( E′ )),
and in particular, for each f ′ ∈ A′ ,
a −1
(1.2.2.2) ϕ ( D ( f ′ )) = D (ϕ( f ′ )).
(ii) For each ideal a of A, we have
(1.2.2.3) a ϕ (V (a)) = V (ϕ−1 (a)).
P ROOF. The relation a ϕ( x ) ∈ V ( E′ ) is, by definition, equivalent to E′ ⊂ ϕ−1 (jx ), so ϕ( E′ ) ⊂ jx ,
and finally x ∈ V (ϕ( E′ )), hence (i). To prove (ii), we can suppose that a is equal to its radical, since
V (r(a)) = V (a) (1.1.2, (v)) and ϕ−1 (r(a)) = r(ϕ−1 (a)); if we set Y = V (a), and a′ = j( a ϕ(Y )), then
we have a (Y ) = V (a′ ) ((1.1.4, (ii))) the relation f ′ ∈ a′ is, by definition, equivalent to f ′ ( x ′ ) = 0 for
each x ∈ a ϕ(Y ), so, by Equation (1.2.1.1), it is also equivalent to ϕ( f ′ )( x ) = 0 for each x ∈ Y, or to
ϕ( f ′ ) ∈ j(Y ) = a, since a is equal to its radical; hence (ii). □
Corollary (1.2.3). — The map a ϕ is continuous.
We remark that, if A′′ is a third ring, and ϕ′ a homomorphism A′′ → A′ , then we have a (ϕ′ ◦ ϕ) =
a ϕ ◦ a ϕ′ ;
this result, with Corollary (1.2.3), says that Spec( A) is a contravariant functor in A, from the
category of rings to that of topological spaces.
Corollary (1.2.4). — Suppose that ϕ is such that every f ∈ A can be written as f = hϕ( f ′ ), where h is I | 84
invertible in A (which, in particular, is the case when ϕ is surjective). Then a ϕ is a homeomorphism from
X to a ϕ( X ).
P ROOF. We show that for each subset E ⊂ A, there exists a subset E′ of A′ such that V ( E) =
V (ϕ( E′ )); according to the (T0 ) axiom (1.1.8) and the formula (1.2.2.1), this implies first of all that a ϕ
is injective, and then, by (1.2.2.1), that a ϕ is a homeomorphism. But it suffices, for each f ∈ E, to take
f ′ ∈ A′ such that hϕ( f ′ ) = f with h invertible in A; the set E′ of these elements f ′ is exactly what
we are searching for. □
1. AFFINE SCHEMES 96
(1.2.5). In particular, when ϕ is the canonical homomorphism from A to a ring quotient A/a, we again
get (1.1.12), and a ϕ is the canonical injection of V (a), identified with Spec( A/a), into X = Spec( A).
Another particular case of (1.2.4):
Corollary (1.2.6). — If S is a multiplicative subset of A, the spectrum Spec(S−1 A) is canonically identified
(with its topology) with the subspace of X = Spec( A) consisting of the x such that jx ∩ S = ∅.
P ROOF. We know by (0, 1.2.6) that the prime ideals of S−1 A are the ideals S−1 jx such that
jx ∩ S = ∅, and that we have jx = (iSA )−1 (S−1 jx ). It then suffices to apply Corollary (1.2.4) to the
iSA . □
Corollary (1.2.7). — For a ϕ( X ) to be dense in X ′ , it is necessary and sufficient for each element of the kernel
Ker ϕ to be nilpotent.
P ROOF. Applying Equation (1.2.2.3) to the ideal a = (0), we have a^ ϕ( X ) = V (Ker ϕ), and for
′
V (Ker ϕ) = X to hold, it is necessary and sufficient for Ker ϕ to be contained in all the prime ideals ErrII
of A′ , or, equivalently, in the nilradical r′ of A′ . □ ErrII
(1.3.3). If D ( f ) = D ( g), then Lemma (1.3.2, b) shows that M f = Mg . More generally, if D ( f ) ⊃ D ( g),
then S′f ⊂ S′g , and we know (0, 1.4.1) that there exists a canonical functorial homomorphism
ρ g, f : M f −→ Mg ,
and if D ( f ) ⊃ D ( g) ⊃ D (h), we have (0, 1.4.4)
(1.3.3.1) ρh,g ◦ ρ g, f = ρh, f .
When f ranges over the elements of A − jx (for a given x in X = Spec( A)), the sets S′f constitute I | 85
an increasing filtered set of subsets of A − jx , since for elements f and g of A − jx , S′f and S′g are
contained in S′f g ; since the union of the S′f over f ∈ A − jx is A − jx , we conclude (0, 1.4.5) that
the A x -module Mx is canonically identified with the inductive limit lim M f , relative to the family of
−→
homomorphisms (ρ g, f ). We denote by
f
ρ x : M f −→ Mx
the canonical homomorphism for f ∈ A − jx (or, equivalently, x ∈ D ( f )).
Definition (1.3.4). — We define the structure sheaf of the prime spectrum X = Spec( A) (resp. the
sheaf associated to the A-module M), denoted by A e or OX (resp. M) e as the sheaf of rings (resp. the
A-module) associated to the presheaf D ( f ) 7→ A f (resp. D ( f ) 7→ M f ), defined on the basis B of X
e
consisting of the D ( f ) for f ∈ A ((1.1.10), (0, 3.2.1), and (0, 3.5.6)).
1. AFFINE SCHEMES 97
(resp. θ f : M f −→ Γ( D ( f ), M
e )),
the canonical map, so that, for all x ∈ D ( f ) and all ξ ∈ M f , we have
f
(1.3.4.1) (θ f (ξ )) x = ρ x (ξ ).
Proposition (1.3.5). — M
e is an exact functor, covariant in M, from the category of A-modules to the
category of A-modules.
e
ρ g, f ρ g, f
ug
Mg / Ng
I | 86
Proposition (1.3.6). — For each f ∈ A, the open subset D ( f ) ⊂ X is canonically identified with the prime
spectrum Spec( A f ), and the sheaf g
M f associated to the A f -module M f is canonically identified with the
e | D ( f ).
restriction M
P ROOF. The first assertion is a particular case of (1.2.6). In addition, if g ∈ A is such that D ( g) ⊂
D ( f ), then Mg is canonically identified with the module of fractions of M f whose denominators
are the powers of the canonical image of g in A f (0, 1.4.6). The canonical identification of g M f with
Me | D ( f ) then follows from the definitions. □
is bijective (in other words, the presheaf D ( f ) 7→ M f is a sheaf ). In particular, M can be identified with
Γ( X, M e ) via θ1 .
P ROOF. Consider the canonical homomorphism v 7→ Γ(v) from Hom Ae ( M, e N e ) to Hom e (Γ( M
Γ( A)
e ), Γ ( N
e ));
the latter module is canonically identified with Hom A ( M, N ), by Theorem (1.3.7). It remains to show
that u 7→ ue and v 7→ Γ(v) are inverses of each other; it is evident that Γ(ue) = u by definition of ue; on
the other hand, if we let u = Γ(v) for v ∈ Hom Ae ( M,
e Ne ), then the map w : Γ( D ( f ), M
e ) → Γ ( D ( f ), N
e)
canonically induced from v is such that the diagram
M
u /N
ρ f ,1 ρ f ,1
Mf
w / Nf
is commutative; so we necessarily have that w = u f for all f ∈ A (0, 1.2.4), which shows that
Γg(v) = v. □
Corollary (1.3.9). —
(i) Let u be a homomorphism from an A-module M to an A-module N; then the sheaves associated to
Ker u, Im u, and Coker u, are Ker ue, Im ue, and Coker ue (respectively). In particular, for ue to be
injective (resp. surjective, bijective), it is necessary and sufficient for u to be so too.
(ii) If M is an inductive limit (resp. direct sum) of a family of A-modules ( Mλ ), then M e is the inductive
limit (resp. direct sum) of the family ( g Mλ ), via a canonical isomorphism.
P ROOF.
(i) It suffices to apply the fact that M
e is an exact functor in M (1.3.5) to the two exact sequences
of A-modules
0 −→ Ker u −→ M −→ Im u −→ 0,
0 −→ Im u −→ N −→ Coker u −→ 0.
The second claim then follows from Theorem (1.3.7).
(ii) Let ( Mλ , gµλ ) be an inductive system of A-modules, with inductive limit M, and let gλ be
the canonical homomorphism Mλ → M. Since we have gf νµ ◦ g
g µλ = gf gλ = f
νλ and f gµ ◦ ggµλ
for λ ⩽ µ ⩽ ν, it follows that ( Mλ , gg
g µλ ) is an inductive system of sheaves on X, and if
we denote by hλ the canonical homomorphism g Mλ → lim g M , then there is a unique
−→ λ
homomorphism v : lim g Mλ → M e such that v ◦ hλ = fgλ . To see that v is bijective, it suffices
−→
to check, for each x ∈ X, that v x is a bijection from (lim g M ) to M e x = Mx , and
e x ; but M
−→ λ x
(lim g
M ) = lim( g M ) = lim( M ) = Mx (0, 1.3.3).
−→ λ x −→ λ x −→ λ x
1. AFFINE SCHEMES 99
Conversely, it follows from the definitions that ( f gλ ) x and (hλ ) are both equal to the canoni-
gλ ) x = v x ◦ (hλ ) x , v x is the identity. Finally, if M is the I | 88
cal map from ( Mλ ) x to Mx ; since ( f
direct sum of two A-modules N and P, it is immediate that M e =N e ⊕ P;e each direct sum
being the inductive limit of finite direct sums, the claims of (ii) are thus proved.
□
We note that Corollary (1.3.8) proves that the sheaves isomorphic to the associated sheaves of
A-modules form an abelian category (T, I, 1.4).
We also note that Corollary (1.3.9) implies that, if M is an A-module of finite type (that is to say,
there exists a surjective homomorphism An → M) then there exists a surjective homomorphism
fn → M,
A e or, in other words, the A-module
e Me is generated by a finite family of sections over X (0, 5.1.1),
and vice versa.
(1.3.10). If N is a submodule of an A-module M, the canonical injection j : N → M gives, by (1.3.9),
an injective homomorphism N e → M,e which allows us to canonically identify N
e with an A-submodule
e
of M; we will always assume that we have made this identification. If N and P are submodules of
e
M, then we have
(1.3.10.1) ( N + P)e = Ne + P,
e
(1.3.10.2) ( N ∩ P)e = N
e ∩ P,
e
since N + P and N ∩ P are the image of the canonical homomorphism N ⊕ P → M and the kernel of
the canonical homomorphism M → ( M/N ) ⊕ ( M/P) (respectively), and it suffices to apply (1.3.9).
e = P,
We conclude from (1.3.10.1) and (1.3.10.2) that, if N e then we have N = P.
Corollary (1.3.11). — On the category of sheaves isomorphic to the associated sheaves of A-modules, the
functor Γ is exact.
e − ue e ve e
P ROOF. Let M → N −
→ P be an exact sequence corresponding to two homomorphisms u :
M → N and v : N → P of A-modules. If Q = Im u and R = Ker v, we have Q e = Im ue = Ker ve = R
e
(Corollary (1.3.9)), hence Q = R. □
Corollary (1.3.12). — Let M and N be A-modules.
(i) The sheaf associated to M ⊗ A N is canonically identified with M
e ⊗ e N.
A
e
(ii) If, in addition, M admits a finite presentation, then the sheaf associated to Hom A ( M, N ) is
canonically identified with H om Ae ( M,
e Ne ).
P ROOF.
(i) The sheaf F = M
e ⊗e N
A
e is associated to the presheaf
U 7−→ F (U ) = Γ(U, M
e) ⊗ e) Γ (U, N ),
Γ(U, A
e
1⊗ ϕ ϕ
ϕ
" |
ϕ
/B
ϕ
B ⊗A B B
(σ being the canonical symmetry map) are commutative; and (b) ϕ(e ⊗ x ) = ϕ( x ⊗ e) = x. By
Corollary (1.3.12), the homomorphism ϕ e : Be⊗e B e → B e of A-modules
e satisfies the analogous
A
conditions, and so it defines an A-algebra
e structure on B.
e In a similar way, the data of a B-module N
is the same as the data of an A-module N and an A-homomorphism ψ : B ⊗ A N → N such that the
diagram
ϕ ⊗1
B ⊗A B ⊗A N / B ⊗A N
1⊗ ψ ψ
/N
ψ
B ⊗A N
is commutative and ψ(e ⊗ n) = n; the homomorphism ψ e: B e⊗e N
A
e → N e satisfies the analogous
condition, and so defines a B-module
e structure on N.
e
In a similar way, we see that if u : B → B′ (resp. v : N → N ′ ) is a homomorphism of A-algebras
(resp. of B-modules), then ue (resp. ve) is a homomorphism of A-algebras
e (resp. of B-modules),
e and
Ker ue is a B-ideal
e (resp. Ker ve, Coker ve, and Im ve are B-modules).
e If N is a B-module, then N e is a
B-module
e of finite type if and only if N is a B-module of finite type (0, 5.2.3).
If M, N are B-modules, then the B-module
e Me ⊗e N e is canonically identified with ( M ⊗ B N )e; I | 90
B
similarly H omBe ( M,e N e ) is canonically identified with (HomB ( M, N ))e whenever M admits a finite
presentation; the proofs are similar to those for Corollary (1.3.12).
If J is an ideal of B, and N is a B-module, then we have (JN )e = e J · N.
e
Finally, if B is an A-algebra graded by the A-submodules Bn (n ∈ Z), then the A-algebra e B,
e the
direct sum of the A-modules
e B
fn (1.3.9), is graded by these A-submodules,
e the axiom of graded
algebras saying that the image of the homomorphism Bm ⊗ Bn → B is contained in Bm+n . Similarly,
if M is a B-module graded by the submodules Mn , then M e is a B-module
e graded by the g Mn .
(1.3.14). If B is an A-algebra, and M a submodule of B, then the A-subalgebra
e of B
e generated by M
e
(0, 4.1.3) is the A-subalgebra C, where we denote by C the subalgebra of B generated by M. Indeed,
e e
C is the direct sum of the submodules of B which are the images of the homomorphisms n M → B
N
zero means that there exists some n ⩾ 0 such that f n z′ = 0 in N, or, equivalently, f n t = 0.
To finish the proof, that (b) implies (d1) and (d2), it suffices to establish the following lemma.
Lemma (1.4.1.1). — Suppose that V is the finite union of sets of the form D ( gi ), and that all of the sheaves
F | D ( gi ) and F |( D ( gi ) ∩ D ( g j )) = F | D ( gi g j ) satisfy (d1) and (d2); then F has the following two
properties:
(d’1) For each f ∈ A and for each section s ∈ Γ( D ( f ) ∩ V, F ), there exists an integer n ⩾ 0 such that
f n s extends to a section of F over V.
(d’2) For each f ∈ A and for each section t ∈ Γ(V, F ) such that the restriction of t to D ( f ) ∩ V is 0,
there exists an integer n ⩾ 0 such that f n t = 0.
We first prove (d′ 2): since D ( f ) ∩ D ( gi ) = D ( f gi ), there exists, for each i, an integer ni such that
the restriction of ( f gi )ni t to D ( gi ) is zero: since the image of gi in A gi is invertible, the restriction of
f ni t to D ( gi ) is also zero; taking n to be the largest of the ni , we have proved (d′ 2).
To show (d′ 1), we apply (d1) to the sheaf F | D ( gi ): there exists an integer ni ⩾ 0 and a section
si′ of F over D ( gi ) extending the restriction of ( f gi )ni s to D ( f gi ); since the image of gi in A gi is
n
invertible, there is a section si of F over D ( gi ) such that si′ = gi i si , and si extends the restriction
n
of f s to D ( f gi ); in addition we can suppose that all the ni are equal to a single integer n. By
i
exists a section s′ ∈ Γ(V, F ) extending the f m si ; as a result, this section extends f n+m s, hence we
have proved (d′ 1).
It remains to show that (d1) and (d2) imply (a). We first show that (d1) and (d2) imply that these
conditions are satisfied for each sheaf F | D ( g), where g ∈ A is such that D ( g) ⊂ V. It is evident for
(d1); on the other hand, if t ∈ Γ( D ( g), F ) is such that its restriction to D ( f ) ⊂ D ( g) is zero, there
exists, by (d1), an integer m ⩾ 0 such that gm t extends to a section s of F over V; applying (d2), we I | 92
see that there exists an integer n ⩾ 0 such that f n gm t = 0, and as the image of g in A g is invertible,
f n t = 0.
That being so, since V is quasi-compact, Lemma (1.4.1.1) proves that the conditions (d′ 1) and (d′ 2)
are satisfied. Consider then the A-module M = Γ(V, F ), and define a homomorphism of A-modules e
u : M e → j∗ (F ), where j is the canonical injection V → X. Since the D ( f ) form a basis for the
topology of X, it suffices, for each f ∈ A, to define a homomorphism u f : M f → Γ( D ( f ), j∗ (F )) =
Γ( D ( f ) ∩ V, F ), with the usual compatibility conditions (0, 3.2.5). Since the canonical image of
f in A f is invertible, the restriction homomorphism M = Γ(V, F ) → Γ( D ( f ) ∩ V, F ) factors as
uf
M → M f −→ Γ( D ( f ) ∩ V, F ) (0, 1.2.4), and the verification of these compatibility conditions for
D ( g) ⊂ D ( f ) is immediate. This being so, we show that the condition (d′ 1) (resp. (d′ 2)) implies that
each of the u f are surjective (resp. injective), which proves that u is bijective, and as a result that
F is the restriction to V of an A-module
e isomorphic to M. e If s ∈ Γ( D ( f ) ∩ V, F ), there exists, by
′
(d 1), an integer n ⩾ 0 such that f s extends to a section z ∈ M; we then have u f (z/ f n ) = s, so u f is
n
surjective. Similarly, if z ∈ M is such that u f (z/1) = 0, this means that the restriction to D ( f ) ∩ V
of the section z is zero; according to (d′ 2), there exists an integer n ⩾ 0 such that f n z = 0, hence
z/1 = 0 in M f , and so u f is injective. □
Corollary (1.4.2). — Each quasi-coherent sheaf over a quasi-compact open subset of X is induced by a
quasi-coherent sheaf on X.
Corollary (1.4.3). — Every quasi-coherent OX -algebra over X = Spec( A) is isomorphic to an OX -algebra
of the form B,
e where B is an algebra over A; every quasi-coherent B-module
e is isomorphic to a B-module
e of the
form N, where N is a B-module.
e
P ROOF. (a) trivially implies (b). To see that (b) implies (c), note that, since V is quasi-compact I | 93
(0, 2.2.3), we have previously seen that F is isomorphic to a sheaf N e |V, where N is an A-module
(1.4.1). We have N = lim Mλ , where Mλ run over the set of A-submodules of N of finite type, hence
−→
(1.3.9) F = N e |V = lim gM |V; but since F is of finite type, and V is quasi-compact, there exists an
−→ λ
index λ such that F = g Mλ |V (0, 5.2.3).
Finally, we show that (c) implies (a). It is clear that F is then of finite type ((1.3.6) and (1.3.9));
in addition, since the questions is local, we can restrict to the case where V = D ( f ), f ∈ A. Since
A f is Noetherian, we see that it suffices to prove that the kernel of a homomorphism A fn → M, e
1. AFFINE SCHEMES 103
where M is an A-module, is of finite type. But such a homomorphism is of the form ue, where u
is a homomorphism An → M (1.3.8), and if P = Ker u then we have Pe = Ker ue (1.3.9). Since A is
Noetherian, P is of finite type, which finishes the proof. □
Corollary (1.5.2). — Under the hypotheses of (1.5.1), the sheaf OX is a quasi-coherent sheaf of rings.
Corollary (1.5.3). — Under the hypotheses of (1.5.1), every coherent sheaf over an open subset of X is
induced by a coherent sheaf on X.
Corollary (1.5.4). — Under the hypotheses of (1.5.1), every quasi-coherent OX -module F is the inductive
limit of the coherent OX -submodules of F .
P ROOF. Indeed, F = M, e where M is an A-module, and M is the inductive limit of its submod-
ules of finite type; we conclude the proof by appealing to (1.3.9) and (1.5.1). □
1.6. Functorial properties of quasi-coherent sheaves over a prime spectrum
(1.6.1). Let A, A′ be rings,
ϕ : A′ −→ A
a homomorphism, and
ϕ : X = Spec( A) −→ X ′ = Spec( A′ )
a
Γ ( D ( g ′ ), f
A′ ) / Γ( D ( g′ ), a ϕ∗ ( A
e)
is commutative (0, 1.5.1); we have thus defined a homomorphism of OX ′ -algebras, as the D ( f ′ ) form
a basis for the topology of X ′ (0, 3.2.3). The pair Φ = ( a ϕ, ϕ
e) is thus a morphism of ringed spaces
Φ : ( X, OX ) −→ ( X ′ , OX ′ ),
(0, 4.1.1).
We also note that, if we put x ′ = a ϕ( x ), then the homomorphism ϕ ex♯ (0, 3.7.1) is exactly the
homomorphism
ϕx : A′x′ −→ A x
canonically induced by ϕ : A′ → A (0, 1.5.1). Indeed, each z′ ∈ A′x′ can be written as g′ / f ′ , where
f ′ , g′ are in A′ and f ′ ̸∈ jx′ ; D ( f ′ ) is then a neighbourhood of x ′ in X ′ , and the homomorphism
Γ ( D ( f ′ ), f
A′ ) → Γ( a ϕ−1 ( D ( f ′ )), A
e) induced by ϕ e is exactly ϕ f ′ ; by considering the section s′ ∈
Γ ( D ( f ′ ), f
A′ ) corresponding to g′ / f ′ ∈ A′ ′ , we obtain ϕ ex♯ (z′ ) = ϕ( g′ )/ϕ( f ′ ) in A x .
f
P ROOF. For purposes of abbreviation, we write M′ = M[ϕ] , and for each f ′ ∈ A′ , we put
f = ϕ( f ′ ). The modules of sections Γ( D ( f ′ ), M
f′ ) and Γ( D ( f ), M
e ) are identified, respectively, with
the modules M f ′ and M f (over A f ′ and A f , respectively); in addition, the A′f ′ -module ( M f )[ϕ ′ ] is
′ ′
f
This proof also shows that, for each A-algebra B, the canonical functorial isomorphism ( B[ϕ] )e ≃ I | 95
e) is an isomorphism of OX ′ -algebras; if M is a B-module, the canonical functorial isomorphism
Φ∗ ( B
( M[ϕ] )e ≃ Φ∗ ( M
e ) is an isomorphism of Φ∗ ( Be)-modules.
Corollary (1.6.4). — The direct image functor Φ∗ is exact on the category of quasi-coherent OX -modules.
f′ is an exact functor in M′
P ROOF. Indeed, it is clear that M[ϕ] is an exact functor in M and M
(1.3.5). □
Proposition (1.6.5). — Let N ′ be an A′ -module, and N the A-module N ′ ⊗ A′ A[ϕ] ; then there exists a
canonical functorial isomorphism from the OX -module Φ∗ ( N
f′ ) to N.
e
h x (s′x ⊗ t x ) = t x · s x (0, 4.4.3). But we can canonically identify N f with N ′f ′ ⊗ A′ ′ ( A f )[ϕ ′ ] (0, 1.5.4);
f f
In addition, let v : N1′ → N2′ be a homomorphism of A′ -modules; since vex′ = v x′ for each x ′ ∈ X ′ ,
it follows immediately from the above that Φ∗ (ve) is canonically identified with (v ⊗ 1)e, which
finishes the proof of (1.6.5). □
If B′ is an A′ -algebra, the canonical isomorphism from Φ∗ ( Be′ ) to ( B′ ⊗ A′ A[ϕ] )e is an isomor-
phism of OX -algebras; if, in addition, N ′ is a B′ -module, then the canonical isomorphism from
Φ∗ ( N
f′ ) to ( N ′ ⊗ A′ A )e is an isomorphism of Φ∗ ( Be′ )-modules.
[ϕ]
Corollary (1.6.6). — The sections of Φ∗ ( Nf′ ), the canonical images of the sections s′ , where s′ varies over
the A′ -module Γ( N
f′ ), generate the A-module Γ(Φ∗ ( N ′ )).
P ROOF. Indeed, these images are identified with the elements z′ ⊗ 1 of N, when we identify N ′
and N with Γ( N e ) (respectively) (1.3.7), and z′ varies over N ′ .
f′ ) and Γ( N □
(1.6.7). In the proof of (1.6.5), we had proved in passing that the canonical map (0, 4.4.3.2) ρ : N f′ →
∗ ′ ′
Φ∗ (Φ ( N )) is exactly the homomorphism j, where j : N → N ⊗ A′ A[ϕ] is the homomorphism I | 96
f ′ e
z′ 7→ z′ ⊗ 1. Similarly, the canonical map (0, 4.4.3.3) σ : Φ∗ (Φ∗ ( M e )) → M e is exactly pe, where
p : M[ϕ] ⊗ A′ A[ϕ] → M is the canonical homomorphism, which sends each tensor product z ⊗ a
(z ∈ M, a ∈ A) to a · z; this follows immediately from the definitions ((0, 3.7.1), (0, 4.4.3), and (1.3.7)).
1. AFFINE SCHEMES 105
We conclude ((0, 4.4.3) and (0, 3.5.4.4)) that if v : N ′ → M[ϕ] is an A′ -homomorphism, then
ve♯ = (v ⊗ 1)e.
(1.6.8). Let N1′ and N2′ be A′ -modules, and assume N1′ admits a finite presentation; it then follows from
(1.6.7) and (1.3.12, ii) that the canonical homomorphism (0, 4.4.6)
Φ∗ (H om f
A′ 1
f′ )) −→ H om (Φ∗ ( N
f′ , N
(N 2 A
e
f′ ), Φ∗ ( N
1
f′ ))
2
e, where γ denotes the canonical homomorphism of A-modules Hom A′ ( N1′ , N2′ ) ⊗ A′ A →
is exactly γ
Hom A ( N1 ⊗ A′ A, N2′ ⊗ A′ A).
′
(1.6.9). Let J′ be an ideal of A′ , and M an A-module; since, by definition, Je′ M e is the image of
∗
the canonical homomorphism Φ (J ) ⊗ Ae M → M, it follows from Proposition (1.6.5) and Corol-
e′ e e
e canonically identifies with (J′ M )e; in particular, Φ∗ (Je′ ) A
lary (1.3.12, i) that Je′ M e is identified with
(J A)e, and, taking the right exactness of the functor Φ into account, the A-algebra Φ∗ (( A′ /J′ )e)
′ ∗ e
is identified with ( A/J′ A)e.
(1.6.10). Let A′′ be a third ring, ϕ′ a homomorphism A′′ → A′ , and write ϕ′′ = ϕ ◦ ϕ′ . It follows
immediately from the definitions that a ϕ′′ = ( a ϕ′ ) ◦ ( a ϕ), and ϕ
f′′ = ϕ
e ◦ ϕe′ (0, 1.5.7). We conclude
′′ ′
that Φ = Φ ◦ Φ; in other words, (Spec( A), A) is a functor from the category of rings to that of
e
ringed spaces.
1.7. Characterization of morphisms of affine schemes
Definition (1.7.1). — We say that a ringed space ( X, OX ) is an affine scheme if it is isomorphic to
a ringed space of the form (Spec( A), A e), where A is a ring; we then say that Γ( X, OX ), which is
canonically identified with the ring A (1.3.7), is the ring of the affine scheme ( X, OX ), and we denote
it by A( X ) when there is no chance of confusion.
By abuse of language, when we speak of an affine scheme Spec( A); it will always be the ringed
space (Spec( A), A
e).
(1.7.2). Let A and B be rings, and ( X, OX ) and (Y, OY ) the affine schemes corresponding to the prime
spectra X = Spec( A), Y = Spec( B). We have seen (1.6.1) that each ring homomorphism ϕ : B → A
corresponds to a morphism Φ = ( a ϕ, ϕ e) = Spec(ϕ) : ( X, OX ) → (Y, OY ). We note that ϕ is entirely
determined by Φ, since we have, by definition, ϕ = Γ(ϕ e) : Γ( B
e) → Γ( a ϕ∗ ( A
e) = Γ( A
e).
Theorem (1.7.3). — 2 Let ( X, OX ) (Y, OY ) be affine schemes. For a morphism of ringed spaces (ψ, θ ) :
( X, OX ) → (Y, OY ) to be of the form ( a ϕ, ϕe), where ϕ is a homomorphism of rings A(Y ) → A( X ), it is
necessary and sufficient that, for each x ∈ X, θ x♯ is a local homomorphism: Oψ( x) → Ox .
ex♯ is the I | 97
P ROOF. Let A = A( X ), B = A(Y ). The condition is necessary, since we saw (1.6.1) that ϕ
homomorphism from Ba ϕ( x) to A x canonically induced by ϕ, and, by definition, of a ϕ( x ) = ϕ−1 (jx ),
this homomorphism is local.
We now prove that the condition is sufficient. By definition, θ is a homomorphism OY → ψ∗ (OX ),
and we canonically obtain a ring homomorphism
ϕ = Γ(θ ) : B = Γ(Y, OY ) −→ Γ(Y, ψ∗ (OX )) = Γ( X, OX ) = A.
The hypotheses on θ x♯ mean that this homomorphism induces, by passing to quotients, a
monomorphism θ x from the residue field k(ψ( x )) to the residue field k( x ), such that, for each section
f ∈ Γ(Y, OY ) = B, we have θ x ( f (ψ( x ))) = ϕ( f )( x ). The relation f (ψ( x )) = 0 is therefore equivalent
to ϕ( f )( x ) = 0, which means that jψ( x) = ja ϕ( x) , and we now write ψ( x ) = a ϕ( x ) for each x ∈ X, or
ψ = a ϕ. We also know that the diagram
/ Γ( X, OX ) = A
ϕ
B = Γ(Y, OY )
♯
Bψ( x)
θx
/ Ax
is commutative (0, 3.7.2), which means that θ x♯ is equal to the homomorphism ϕx : Bψ( x) → A x
canonically induced by ϕ (0, 1.5.1). As the data of the θ x♯ completely characterize θ ♯ , and as a result θ
(0, 3.7.1), we conclude that we have θ = ϕe, by the definition of ϕe (1.6.1). □
We say that a morphism (ψ, θ ) of ringed spaces satisfying the condition of (1.7.3) is a morphism
of affine schemes.
Corollary (1.7.4). — If ( X, OX ) and (Y, OY ) are affine schemes, there exists a canonical isomorphism from
the set of morphisms of affine schemes Hom(( X, OX ), (Y, OY )) to the set of ring homomorphisms from B to
A, where A = Γ(OX ) and B = Γ(OY ).
e) in A and Γ( X, OX ) in ( X, OX ) define an
Furthermore, we can say that the functors (Spec( A), A
equivalence between the category of commutative rings and the opposite category of affine schemes
(T, I, 1.2).
Corollary (1.7.5). — If ϕ : B → A is surjective, then the corresponding morphism ( a ϕ, ϕ
e) is a monomor-
phism of ringed spaces (cf. (4.1.7)).
P ROOF. Indeed, we know that a ϕ is injective (1.2.5), and, since ϕ is surjective, for each x ∈ X,
ϕx♯
: Ba ϕ( x) → A x , which is induced by ϕ by passing to rings of fractions, is also surjective (0, 1.5.1);
hence the conclusion (0, 4.1.1). □
1.8. Morphisms from locally ringed spaces to affine schemes Due to a remark by J. Tate, the II | 217
statements of Theorem (1.7.3) and Proposition (2.2.4) can be generalized as follows:3
Proposition (1.8.1). — Let (S, OS ) be an affine scheme, and ( X, OX ) a locally ringed space. Then there
is a canonical bijection from the set of ring homomorphisms Γ(S, OS ) → Γ( X, OX ) to the set of morphisms II | 218
of ringed spaces (ψ, θ ) : ( X, OX ) → (S, OS ) such that, for each x ∈ X, θ x♯ is a local homomorphism
Oψ ( x ) → O x .
P ROOF. We note first that if ( X, OX ) and (S, OS ) are any two ringed spaces, then a morphism
(ψ, θ ) from ( X, OX ) to (S, OS ) canonically defines a ring homomorphism Γ(θ ) : Γ(S, OS ) → Γ( X, OX ),
hence a first map
(1.8.1.1) ρ : Hom(( X, OX ), (S, OS )) −→ Hom(Γ(S, OS ), Γ( X, OX )).
Conversely, under the stated hypotheses, we set A = Γ(S, OS ), and consider a ring homomorphism
ϕ : A → Γ( X, OX ). For each x ∈ X, it is clear that the set of the f ∈ A such that ϕ( f )( x ) = 0 is a
prime ideal of A, since Ox /mx = k( x ) is a field; it is therefore an element of S = Spec( A), which we
denote by a ϕ( x ). In addition, for each f ∈ A, we have, by definition (0, 5.5.2), that a ϕ−1 ( D ( f )) = X f ,
which proves that a ϕ is a continuous map X → S. We then define a homomorphism
e : OS −→ a ϕ∗ (OX )
ϕ
of OS -modules; for each f ∈ A, we have Γ( D ( f ), OS ) = A f (1.3.6); for each s ∈ A, we associate to
s/ f ∈ A f the element (ϕ(s)| X f )(ϕ( f )| X f )−1 of Γ( X f , OX ) = Γ( D ( f ), a ϕ∗ (OX )), and we immedi-
ately see (by passing from D ( f ) to D ( f g)) that this is a well-defined homomorphism of OS -modules,
hence a morphism ( a ϕ, ϕ
e) of ringed spaces. In addition, with the same notation, and setting y = a ϕ( x )
for brevity, we immediately see (0, 3.7.1) that we have ϕ ex♯ (sy / f y ) = (ϕ(s) x )(ϕ( f ) x )−1 ; since the
relation sy ∈ my is, by definition, equivalent to ϕ(s) x ∈ mx , we see that ϕ ex♯ is a local homomorphism
Oy → Ox , and we have thus defined a second map σ : Hom(Γ(S, OS ), Γ( X, OX )) → L, where L is
the set of the morphisms (ψ, θ ) : ( X, OX ) → (S, OS ) such that θ x♯ is local for each x ∈ X. It remains
to prove that σ and ρ (restricted to L) are inverses of each other; the definition of ϕ e immediately
shows that Γ(ϕ e) = ϕ, and, as a result, that ρ ◦ σ is the identity. To see that σ ◦ ρ is the identity,
start with a morphism (ψ, θ ) ∈ L and let ϕ = Γ(θ ); the hypotheses on θ x♯ mean that this morphism
induces, by passing to quotients, a monomorphism θ x : k(ψ( x )) → k( x ) such that for each section
3[Trans.] The following section (I.1.8) was added in the errata of EGA II, hence the temporary change in page numbers, which refer
to EGA II.
1. AFFINE SCHEMES 107
/ Γ( X, OX )
ϕ
A
♯
Aψ( x )
θx
/ Ox
is commutative, and it is the same for the analogous diagram where θ x♯ is replaced by ϕ
ex♯ , hence
ex♯ = θ x♯ (0, 1.2.4), and, as a result, ϕ
ϕ e = θ. □
(1.8.2). When ( X, OX ) and (Y, OY ) are locally ringed spaces, we will consider the morphisms (ψ, θ ) :
( X, OX ) → (Y, OY ) such that, for each x ∈ X, θ x♯ is a local homomorphism Oψ(x) → Ox . Henceforth
when we speak of a morphism of locally ringed spaces, it will always be a morphism like the above; II | 219
with this definition of morphisms, it is clear that the locally ringed spaces form a category; for any
two objects X and Y of this category, Hom( X, Y ) thus denotes the set of morphisms of locally ringed
spaces from X to Y (the set denoted L in (1.8.1)); when we consider the set of morphisms of ringed
spaces from X to Y, we will denote it by Homrs ( X, Y ) to avoid any confusion. The map (1.8.1.1) is
then written as
(1.8.2.1) ρ : Homrs ( X, Y ) −→ Hom(Γ(Y, OY ), Γ( X, OX ))
and its restriction
(1.8.2.2) ρ′ : Hom( X, Y ) −→ Hom(Γ(Y, OY ), Γ( X, OX ))
is a functorial map in X and Y on the category of locally ringed spaces.
Corollary (1.8.3). — Let (Y, OY ) be a locally ringed space. For Y to be an affine scheme, it is necessary and
sufficient that, for each locally ringed space ( X, OX ), the map (1.8.2.2) be bijective.
P ROOF. Proposition (1.8.1) shows that the condition is necessary. Conversely, if we suppose that
the condition is satisfied, and if we put A = Γ(Y, OY ), then it follows from the hypotheses and from
(1.8.1) that the functors X 7→ Hom( X, Y ) and X 7→ Hom( X, Spec( A)), from the category of locally
ringed spaces to that of sets, are isomorphic. We know that this implies the existence of a canonical
isomorphism X → Spec( A) (cf. 0, 8). □
(1.8.4). Let S = Spec( A) be an affine scheme; denote by (S′ , A′ ) the ringed space whose underlying
space is a point and the structure sheaf A′ is the (necessarily simple) sheaf on S′ defined by the
ring A. Let π : S → S′ be the unique map from S to S′ ; on the other hand, we note that, for each
open subset U of S, we have a canonical map Γ(S′ , A′ ) = Γ(S, OS ) → Γ(U, OS ) which thus defines a
π-morphism ι : A′ → OS of sheaves of rings. We have thus canonically defined a morphism of ringed
spaces i = (π, ι) : (S, OS ) → (S′ , A′ ). For each A-module M, we denote by M′ the simple sheaf on S′
defined by M, which is evidently an A′ -module. It is clear that i∗ ( Me ) = M′ (1.3.7).
Lemma (1.8.5). — With the notation of (1.8.4), for each A-module M, the canonical functorial OS -
homomorphism (0, 4.4.3.3)
(1.8.5.1) i ∗ (i ∗ ( M
e )) −→ M
e
is an isomorphism.
P ROOF. Indeed, the two parts of (1.8.5.1) are right exact (the functor M 7→ i∗ ( M
e ) evidently
being exact) and commute with direct sums; by considering M as the cokernel of a homomorphism
A( I ) → A( J ) , we can reduce to proving the lemma for the case where M = A, and it is evident in
this case. □
Corollary (1.8.6). — Let ( X, OX ) be a ringed space, and u : X → S a morphism of ringed spaces. For each II | 220
A-module M, we have (with the notation of (1.8.4)) a canonical functorial isomorphism of OX -modules
(1.8.6.1) u∗ ( M
e ) ≃ u∗ (i∗ ( M′ )).
1. AFFINE SCHEMES 108
Corollary (1.8.7). — Under the hypotheses of (1.8.6), we have, for each A-module M and each OX -module
F , a canonical isomorphism, functorial in M and F ,
P ROOF. We have, according to (0, 4.4.3) and Lemma (1.8.5), a canonical isomorphism of bifunc-
tors
e u∗ (F )) ≃ Hom A′ ( M′ , i∗ (u∗ (F )))
HomOS ( M,
and it is clear that the right hand side is exactly Hom A ( M, Γ( X, F )). We note that the canonical
homomorphism (1.8.7.1) sends each OS -homomorphism h : M e → u∗ (F ) (in other words, each
u-morphism M e → F ) to the A-homomorphism Γ(h) : M → Γ(S, u∗ (F )) = Γ( X, F ). □
(1.8.8). With the notation of (1.8.4), it is clear (0, 4.1.1) that each morphism of ringed spaces (ψ, θ ) :
X → S′ is equivalent to the data of a ring homomorphism A → Γ( X, OX ). We can thus interpret
Proposition (1.8.1) as defining a canonical bijection Hom( X, S) ≃ Hom( X, S′ ) (where we understand
that the right-hand side is the collection of morphisms of ringed spaces, since in general A is not a
local ring). More generally, if X and Y are locally ringed spaces, and if (Y ′ , A′ ) is the ringed space
whose underlying space is a point and whose sheaf of rings A′ is the simple sheaf defined by the
ring Γ(Y, OY ), we can interpret (1.8.2.1) as a map
The result of Corollary (1.8.3) is interpreted by saying that affine schemes are characterized among
locally ringed spaces as those for which the restriction of ρ to Hom( X, Y ):
is bijective for every locally ringed space X. In the following chapter, we generalize this definition,
which allows us to associate to any ringed space Z (and not only to a ringed space whose underlying
space is a point) a locally ringed space which we will call Spec( Z ); this will be the starting point for
a “relative” theory of preschemes over any ringed space, extending the results of Chapter I.
(1.8.9). We can consider the pairs ( X, F ) consisting of a locally ringed space X and an OX -module
F as forming a category, a morphism in this category being a pair (u, h) consisting of a morphism of
locally ringed spaces u : X → Y and a u-morphism h : G → F of modules; these morphisms (for II | 221
( X, F ) and (Y, G ) fixed) form a set which we denote by Hom(( X, F ), (Y, G )); the map (u, h) 7→
(ρ′ (u), Γ(h)) is a canonical map
functorial in ( X, F ) and (Y, G ), the right-hand side being the set of di-homomorphisms correspond-
ing to the rings and modules considered (0, 1.0.2).
Corollary (1.8.10). — Let Y be a locally ringed space, and G an OY -module. For Y to be an affine scheme
and G to be a quasi-coherent OY -module, it is necessary and sufficient that, for each pair ( X, F ) consisting of
a locally ringed space X and an OX -module F , the canonical map (1.8.9.1) be bijective.
We leave the proof, which is modelled on that of (1.8.3), using (1.8.1) and (1.8.7), to the reader.
Remark (1.8.11). — The statements (1.7.3), (1.7.4), and (2.2.4) are particular cases of (1.8.1), as well
as the definition in (1.6.1); similarly, (2.2.5) follows from (1.8.7). Corollary (1.8.7) also implies (1.6.3)
(and, as a result, (1.6.4)) as a particular case, since if X is an affine scheme and Γ( X, F ) = N, then
the functors M 7→ HomOS ( M, e )) and M 7→ HomO ( M,
e u∗ ( N e ( N[ϕ] )e) (where ϕ : A → Γ( X, OX )
S
corresponds to u) are isomorphic, by Corollaries (1.8.7) and (1.3.8). Finally, (1.6.5) (and, as a result,
(1.6.6)) follow from (1.8.6), and the fact that, for each f ∈ A, the A f -modules N ′ ⊗ A′ A f and
( N ′ ⊗ A′ A) f (with the notation of (1.6.5)) are canonically isomorphic.
2. PRESCHEMES AND MORPHISMS OF PRESCHEMES 109
Definition (2.2.1). — Given two preschemes, ( X, OX ) and (Y, OY ), we define a morphism (of
preschemes) from ( X, OX ) to (Y, OY ) to be a morphism of ringed spaces (ψ, θ ) such that, for all
x ∈ X, θ x♯ is a local homomorphism Oψ( x) → Ox .
(2.2.2). The composition (ψ′′ , θ ′′ ) of two morphisms (ψ, θ ), (ψ′ , θ ′ ) of preschemes is also a morphism
♯ ♯
of preschemes, since it is given by the formula θ ′′ = θ ♯ ◦ ψ∗ (θ ′ ) (0, 3.5.5). From this we conclude
that preschemes form a category; using the usual notation, we will write Hom( X, Y ) to mean the set
of morphisms from a prescheme X to a prescheme Y.
Example (2.2.3). — If U is an open subset of X, then the canonical injection (0, 4.1.2) of the induced
prescheme (U, OX |U ) into ( X, OX ) is a morphism of preschemes; it is further a monomorphism of
ringed spaces (and a fortiori a monomorphism of preschemes), which follows rapidly from (0, 4.1.1).
Proposition (2.2.4). — 4 Let ( X, OX ) be a prescheme, and (S, OS ) an affine scheme associated to a ring A.
Then there exists a canonical bijective correspondence between morphisms of preschemes from ( X, OX ) to
(S, OS ) and ring homomorphisms from A to Γ( X, OX ).
P ROOF. First note that, if ( X, OX ) and (Y, OY ) are two arbitrary ringed spaces, a morphism
(ψ, θ ) from ( X, OX ) to (Y, OY ) canonically defines a ring homomorphism Γ(θ ) : Γ(Y, OY ) →
Γ(Y, ψ∗ (OX )) = Γ( X, OX ). In the case that we consider, everything boils down to showing that any
homomorphism ϕ : A → Γ( X, OX ) is of the form Γ(θ ) for exactly one θ. Now, by hypothesis, there
is a covering (Vα ) of X by affine open subsets; by composing ϕ with the restriction homomorphism
Γ( X, OX ) → Γ(Vα , OX |Vα ), we obtain a homomorphism ϕα : A → Γ(Vα , OX |Vα ) that corresponds to
a unique morphism (ψα , θα ) from the prescheme (Vα , OX |Vα ) to (S, OS ), by Theorem (1.7.3). Further-
more, for each pair of indices (α, β), each point of Vα ∩ Vβ admits an affine open neighbourhood
W contained inside Vα ∩ Vβ (2.1.3); it is clear that, by composing ϕα and ϕβ with the restriction
homomorphisms to W, we obtain the same homomorphism Γ(S, OS ) → Γ(W, OX |W ), so, with the
equation (θα♯ ) x = (ϕα ) x for all x ∈ Vα and all α (1.6.1), the restriction to W of the morphisms (ψα , θα )
and (ψβ , θ β ) coincide. From this we conclude that there is a morphism (ψ, θ ) : ( X, OX ) → (S, OS )
of ringed spaces, and only one such that its restriction to each Vα is (ψα , θα ), and it is clear that this
morphism is a morphism of preschemes and such that Γ(θ ) = ϕ.
Let u : A → Γ( X, OX ) be a ring homomorphism, and v = (ψ, θ ) the corresponding morphism
( X, OX ) → (S, OS ). For each f ∈ A, we have that
(2.2.4.1) ψ−1 ( D ( f )) = Xu( f )
with the notation of (0, 5.5.2) relative to the locally free sheaf OX . In fact, it suffices to verify this
formula when X itself is affine, and then this is nothing but (1.2.2.2). □
P ROOF. Reasoning as in Proposition (2.2.4), we reduce to the case where X is affine, and the
proposition then follows from Proposition (1.6.3) and from Corollary (1.3.8). □
(2.2.6). We say that a morphism of preschemes (ψ, θ ) : ( X, OX ) → (Y, OY ) is open (resp. closed) if, for
all open subsets U of X (resp. all closed subsets F of X), ψ(U ) is open (resp. ψ( F ) is closed) in Y. We
say that (ψ, θ ) is dominant if ψ( X ) is dense in Y, and surjective if ψ is surjective. We note that these
conditions rely only on the continuous map ψ.
B′ /B
Oy
We see (2.1.5) that there is a bijective correspondence between Spec(Oy ) and the set of closed irreducible
subsets of Y containing y.
Corollary (2.4.3). — For y ∈ Y to be the generic point of an irreducible component of Y, it is necessary and
sufficient for the only prime ideal of the local ring Oy to be its maximal ideal (in other words, for Oy to be
of dimension zero).
Proposition (2.4.4). — Let ( X, OX ) be a local scheme of some ring A, a its unique closed point, and
(Y, OY ) a prescheme. Every morphism u = (ψ, θ ) : ( X, OX ) → (Y, OY ) then factors uniquely as X →
Spec(Oψ(a) ) → Y, where the second arrow denotes the canonical morphism, and the first corresponds to a
local homomorphism Oψ(a) → A. This establishes a canonical bijective correspondence between the set of
morphisms ( X, OX ) → (Y, OY ) and the set of local homomorphisms Oy → A for (y ∈ Y).
Indeed, for all x ∈ X, we have that a ∈ { x }, so ψ( a) ∈ {ψ( x )}, which shows that ψ( X ) is
contained in every affine open subset that contains ψ( a). So it suffices to consider the case where
(Y, OY ) is an affine scheme of ring B, and then we have that u = ( a ϕ, ϕe), where ϕ ∈ Hom( B, A)
(1.7.3). Further, we have that ϕ−1 (ja ) = jψ(a) , and hence that the image under ϕ of any element of
B − jψ(a) is invertible in the local ring A; the factorization in the result follows from the universal
property of the ring of fractions (0, 1.2.4). Conversely, to each local homomorphism Oy → A there
is a unique corresponding morphism (ψ, θ ) : X → Spec(Oy ) such that ψ( a) = y (1.7.3), and, by
composing with the canonical morphism Spec(Oy ) → Y, we obtain a morphism X → Y, which
proves the proposition.
2. PRESCHEMES AND MORPHISMS OF PRESCHEMES 113
(2.4.5). The affine schemes whose ring is a field K have an underlying space that is just a point. If A
is a local ring with maximal ideal m, then each local homomorphism A → K has kernel equal to m,
and so factors as A → A/m → K, where the second arrow is a monomorphism. The morphisms
Spec(K ) → Spec( A) thus correspond bijectively to monomorphisms of fields A/m → K.
Let (Y, OY ) be a prescheme; for each y ∈ Y and each ideal ay of Oy , the canonical homomorphism
Oy → Oy /ay defines a morphism Spec(Oy /ay ) → Spec(Oy ); if we compose this with the canonical
morphism Spec(Oy ) → Y, then we obtain a morphism Spec(Oy /ay ) → Y, again said to be canonical.
For ay = my , this says that Oy /ay = k(y), and so Proposition (2.4.4) says that:
Corollary (2.4.6). — Let ( X, OX ) be a local scheme whose ring K is a field, ξ the unique point of I | 103
X, and (Y, OY ) a prescheme. Then each morphism u : ( X, OX ) → (Y, OY ) factors uniquely as X →
Spec(k(ψ(ξ ))) → Y, where the second arrow denotes the canonical morphism, and the first corresponds
to a monomorphism k(ψ(ξ )) → K. This establishes a canonical bijective correspondence between the set of
morphisms ( X, OX ) → (Y, OY ) and the set of monomorphisms k (y) → K (for y ∈ Y).
Corollary (2.4.7). — For all y ∈ Y, every canonical morphism Spec(Oy /ay ) → Y is a monomorphism of
ringed spaces.
P ROOF. We have already seen this when ay = 0 (2.4.2), and it suffices to apply Corollary (1.7.5).
□
Remark. — 2.4.8 Let X be a local scheme, and a its unique closed point. Since every affine open
subset containing a is necessarily equal to the whole of X, every invertible OX -module (0, 5.4.1) is
necessarily isomorphic to OX (or, as we say, again, trivial). This property does not hold in general for
an arbitrary affine scheme Spec( A); we will see in Chapter V that if A is a normal ring then this is
true when A is a unique factorisation domain.
2.5. Preschemes over a prescheme
Definition (2.5.1). — Given a prescheme S, we say that the data of a prescheme X and a morphism
of preschemes ϕ : X → S defines a prescheme X over the prescheme S, or an S-prescheme; we say that
S is the base prescheme of the S-prescheme X. The morphism ϕ is called the structure morphism of
the S-prescheme X. When S is an affine scheme of ring A, we also say that X endowed with ϕ is a
prescheme over the ring A (or an A-prescheme).
It follows from (2.2.4) that the data of a prescheme over a ring A is equivalent to the data of
a prescheme ( X, OX ) whose structure sheaf OX is a sheaf of A-algebras. An arbitrary prescheme can
always be considered as a Z-prescheme in a unique way.
If ϕ : X → S is the structure morphism of an S-prescheme X, we say that a point x ∈ X is over a
point s ∈ S if ϕ( x ) = s. We say that X dominates S if ϕ is a dominant morphism (2.2.6).
(2.5.2). Let X and Y be S-preschemes; we say that a morphism of preschemes u : X → Y is a morphism
of preschemes over S (or an S-morphism) if the diagram
X
u /Y
S
(where the diagonal arrows are the structure morphisms) is commutative: this ensures that, for all
s ∈ S and x ∈ X over s, u( x ) also lies over s.
It follows immediately from this definition that the composition of any two S-morphisms is an
S-morphism; S-preschemes thus form a category.
We denote by HomS ( X, Y ) the set of S-morphisms from an S-prescheme X to an S-prescheme Y;
the identity morphism of an S-prescheme X is denoted by 1X .
When S is an affine scheme of ring A, we will also say A-morphism instead of S-morphism.
(2.5.3). If X is an S-prescheme, and v : X ′ → X a morphism of preschemes, then the composition I | 104
X ′ → X → S endows X ′ with the structure of an S-prescheme; in particular, every prescheme
induced by an open set U of X can be considered as an S-prescheme by the canonical injection.
3. PRODUCTS OF PRESCHEMES 114
is a bijection from the set Hom( X, Y ) to the product set ∏α Hom( Xα , Y ). In particular, if the Xα are
S-preschemes, with structure morphisms ψα , then X is an S-prescheme by the unique morphism
ψ : X → S such that ψ ◦ ϕα = ψα for each α. The sum of two preschemes X and Y is denoted by
X ⊔ Y. It is immediate that, if X = Spec( A) and Y = Spec( B), then X ⊔ Y is canonically identified
with Spec( A × B).
the restrictions of f αλ and f βµ to Tαλ ∩ Tβµ coincide, which will finish the proof of Lemma (3.2.6.2).
The images of Tαλ ∩ Tβµ under f αλ and f βµ are contained in Wαλ ∩ Wβµ by definition. Since
Lemma (3.2.6.3). — Let (Uα ) be an open cover of X, (Vλ ) an open cover of Y, and suppose that, for each I | 107
pair (α, λ), there exists a product of Uα and Vλ ; then there exists a product of X and Y.
Applying Lemma (3.2.6.1) to the open sets Uα ∩ Uβ and Vλ ∩ Vµ , we see that there exists a
product of S-preschemes induced, respectively, by X and Y on these open sets; in addition, the
uniqueness of the product shows that, if we set i = (α, λ) and j = ( β, µ), then there is a canonical
isomorphism hij (resp. h ji ) from this product to an S-prescheme Wij (resp. Wji ) induced by Uα ×S Vλ
(resp. Uβ ×S Vµ ) on an open set; then f ij = hij ◦ h− 1
ji is an isomorphism from Wji to Wij . In addition,
for a third pair k = (γ, ν), we have f ik = f ij ◦ f jk on Wki ∩ Wkj , by applying Lemma (3.2.6.1) to
the open sets Uα ∩ Uβ ∩ Uγ and Vλ ∩ Vµ ∩ Vν in Uβ and Vµ , respectively. It follows that we have
a prescheme Z, an open cover ( Zi ) of the underlying space of Z, and, for each i, an isomorphism
gi from the induced prescheme Zi to the prescheme Uα ×S Vλ , so that, for each pair (i, j), we have
f ij = gi ◦ g−
j
1
(2.3.1); in addition, we have gi ( Zi ∩ Zj ) = Wij . If pi , qi , and θi are the projections
and the structure morphism of the S-prescheme Uα ×S Vλ (respectively), we immediately see that
pi ◦ gi = p j ◦ g j on Zi ∩ Zj , and similarly for the two other morphisms. We can thus define the
morphisms of preschemes p : Z → X (resp. q : Z → Y, θ : Z → S) by the condition that p (resp.
q, θ) coincide with pi ◦ gi (resp. qi ◦ gi , θi ◦ gi ) on each of the Zi ; Z, equipped with θ, is then an
S-prescheme. We now show that Zi′ = p−1 (Uα ) ∩ q−1 (Vλ ) is equal to Zi . For each index j = ( β, µ),
we have Zj ∩ Zi′ = g− 1 −1 −1
j ( p j (Uα ) ∩ q j (Vλ )). We have, by Lemma (3.2.6.1),
p− 1 −1 −1 −1
j (Uα ) ∩ q j (Vλ ) = p j (Uα ∩ Uβ ) ∩ q j (Vλ ∩ Vµ );
U ×S′ V is canonically identified with the prescheme induced on Z by p−1 (U ) ∩ q−1 (V ) (considered as an S′ -
prescheme). In addition, if f : T → X and g : T → Y are S-morphisms such that f ( T ) ⊂ U and g( T ) ⊂ V,
then the S′ -morphism ( f , g)S′ can be identified with the restriction of ( f , g)S to p−1 (U ) ∩ q−1 (V ).
(3.2.8). Let ( Xα ) and (Yλ ) be families of S-preschemes, and X (resp. Y) the sum of the family ( Xα )
(resp. (Yλ )) (3.1). Then X ×S Y can be identified with the sum of the family ( Xα ×S Yλ ); this follows
immediately from Lemma (3.2.6.3).
(3.2.9). 6 It follows from (1.8.1) that we can state (3.2.2) in the following manner: Z = Spec( B ⊗ A C ) II | 221
is not only a product of X = Spec( B) and Y = Spec(C ) in the category of S-preschemes, but also in
the category of locally ringed spaces over S (with a definition of S-morphisms modelled on that of
(2.5.2)). The proof of (3.2.6) also proves that, for any two S-preschemes X and Y, the prescheme
X ×S Y is not only the product of X and Y in the category of S-preschemes, but also in the category
of locally ringed spaces over the prescheme S.
(3.3.1). The reader will notice that all the properties stated in this section, except (3.3.13) and (3.3.15),
are true without modification in any category, whenever the products involved in the statements
exist (since it is clear that the notions of an S-object and of an S-morphism can be defined exactly as
in (2.5) for any object S of the category).
f ×1 f ′ ×1
X×Y / X′ × Y / X ′′ × Y
f f′
X / X′ / X ′′
is commutative.
Proposition (3.3.3). — For each S-prescheme X, the first (resp. second) projection from X ×S S (resp.
S ×S X) is a functorial isomorphism from X ×S S (resp. S ×S X) to X, whose inverse isomorphism is (1X , ϕ)S
(resp. (ϕ, 1X )S ), where we denote by ϕ the structure morphism X → S; we can therefore write, up to a
canonical isomorphism,
X ×S S = S ×S X = X.
(3.3.5). We can define, in a manner similar to (3.2), the product of a finite number n of S-preschemes, I | 109
and the existence of these products follows from (3.2.6) by induction on n, and by noting that
( X1 ×S X2 ×S · · · ×S Xn−1 ) ×S Xn satisfies the definition of a product. The uniqueness of the
product implies, as in any category, its commutativity and associativity properties. If, for example,
p1 , p2 , and p3 denote the projections from X1 ×S X2 ×S X3 , and if we identify this prescheme with
( X1 ×S X2 ) ×S X3 , then the projection to X1 ×S X2 is identified with ( p1 , p2 )S .
6[Trans.] (3.2.9) is from the errata of EGA II, on page 221, whence the change in page numbering.
3. PRODUCTS OF PRESCHEMES 118
(3.3.8). The prescheme X(S′ ) can be considered as a solution to a universal mapping problem: each
S′ -prescheme T is also an S-prescheme via ϕ; each S-morphism g : T → X is then uniquely written
as g = p ◦ f , where f is an S′ -morphism T → X(S′ ) , as follows from the definition of the product
applied to the S-morphisms f and ψ : T → S′ (the structure morphism of T).
Proposition (3.3.9). — (“Transitivity of base change”). Let S′′ be a prescheme, and ϕ′ : S′′ → S a
morphism. For each S-prescheme X, there exists a canonical functorial isomorphism from the S′′ -prescheme
( X(ϕ) )(ϕ′ ) to the S′′ -prescheme X(ϕ◦ϕ′ ) .
P ROOF. Let T be a S′′ -prescheme, ψ its structure morphism, and g an S-morphism from T to
X (T being considered as an S-prescheme with structure morphism ϕ ◦ ϕ′ ◦ ψ). Since T is also an
S′ -prescheme with structure morphism ϕ′ ◦ ψ, we can write g = p ◦ g′ , where g′ is an S′ -morphism
T → X(ϕ) , and then g′ = p′ ◦ g′′ , where g′′ is an S′′ -morphism T → ( X(ϕ) )(ϕ′ ) :
p p′
Xo X( ϕ ) o ( X( ϕ ) ) ( ϕ ′ )
π π′ π ′′
ϕ′
So So
ϕ
S′′ .
So the result follows by the uniqueness of the solution to a universal mapping problem. □ I | 110
This result can be written as the equality (up to a canonical isomorphism) ( X(S′ ) )(S′′ ) = X(S′′ ) (if
there is no chance of confusion), or also as
(3.3.9.1) ( X ×S S′ ) ×S′ S′′ = X ×S S′′ ;
the functorial nature of the isomorphism defined in (3.3.9) can similarly be expressed by the transi-
tivity formula for base change morphisms
(3.3.9.2) ( f (S′ ) )(S′′ ) = f (S′′ )
for each S-morphism f : X → Y.
Corollary (3.3.10). — If X and Y are S-preschemes, then there exists a canonical functorial isomorphism
from the S′ -prescheme X(S′ ) ×S′ Y(S′ ) to the S′ -prescheme ( X ×S Y )(S′ ) .
The functorial nature of the isomorphism defined in Corollary (3.3.10) can be expressed by the
formula
(3.3.10.1) (u(S′ ) , v(S′ ) )S′ = ((u, v)S )(S′ )
for each pair of S-morphisms u : T → X, v : T → Y.
In other words, the base change functor X(S′ ) commutes with products; it also commutes with
sums (3.2.8).
Corollary (3.3.11). — Let Y be an S-prescheme, and f : X → Y a morphism which makes X a Y-prescheme
(and, as a result, also an S-prescheme). The prescheme X(S′ ) is then identified with the product X ×Y Y(S′ ) ,
the projection X ×Y Y(S′ ) → Y(S′ ) being identified with f (S′ ) .
P ROOF. Let ψ : Y → S be the structure morphism of Y; we have the commutative diagram
f (S′ )
S′ o Y(S′ ) o X( S ′ )
f
So Yo
ψ
X.
We have that Y(S′ ) is identified with S(′ ψ) , and X(S′ ) with S(′ ψ◦ f ) ; taking (3.3.9) and (3.3.4) into account,
we thus deduce the corollary. □
(3.3.12). Let f : X → X ′ and g : Y → Y ′ be S-morphisms which are monomorphisms of preschemes
(T, I, 1.1); then f ×S g is a monomorphism. Indeed, if p and q are the projections of X ×S Y, p′ and
q′ the projections of X ′ ×S Y ′ , and u and v both S-morphisms T → X ×S Y, then the equation
( f ×S g) ◦ u = ( f ×S g) ◦ v implies that p′ ◦ ( f ×S g) ◦ u = p′ ◦ ( f ×S g) ◦ v, or, in other words,
that f ◦ p ◦ u = f ◦ p ◦ v, and since f is a monomorphism, p ◦ u = p ◦ v; using the fact that g is a
monomorphism, we similarly obtain q ◦ u = q ◦ v, hence u = v.
It follows that, for each base change S′ → S, I | 111
f (S′ ) : X(S′ ) −→ Y(S′ )
is a monomorphism.
(3.3.13). Let S and S′ be affine schemes of rings A and A′ (respectively); a morphism S′ → S
then corresponds to a ring homomorphism A → A′ . If X is an S-prescheme, we denote by X( A′ ) or
X ⊗ A A′ the S′ -prescheme X(S′ ) ; when X is also affine of ring B, X( A′ ) is affine of ring B( A′ ) = B ⊗ A A′
obtained by extension of scalars from the A-algebra B to A′ .
(3.3.14). With the notation of (3.3.6), for each S-morphism f : S′ → X, we have that f ′ = ( f , 1S′ )S is
an S′ -morphism S′ → X ′ = X(S′ ) such that p ◦ f ′ = f , π ′ ◦ f ′ = 1S′ , or, in other words, an S′ -section
of of X ′ ; conversely, if f ′ is such an S′ -section, then f = p ◦ f ′ is an S-morphism S′ → X. We thus
define a canonical bijective correspondence
HomS (S′ , X ) ≃ HomS′ (S′ , X ′ ).
We say that f ′ is the graph morphism of f , and we denote it by Γ f .
(3.3.15). Given a prescheme X, which we can always consider as a Z-prescheme, it follows, in
particular, from (3.3.14) that the X-sections of X ⊗Z Z[ T ] (where T is an indeterminate) correspond
bijectively with morphisms Z[ T ] → X. We will show that these X-sections also correspond bijectively
with sections of the structure sheaf OX over X. Indeed, let (Uα ) be a cover of X by affine open subsets;
let u : X → X ⊗Z Z[ T ] be an X-morphism, and let uα be its restriction to Uα ; if Aα is the ring of
the affine scheme Uα , then Uα ⊗Z Z[ T ] is an affine scheme of ring Aα [ T ] (3.2.2), and uα canonically
corresponds to an Aα -homomorphism Aα [ T ] → Aα (1.7.3). Now, since such a homomorphism is
completely determined by the data of the image of T in Aα , let sα ∈ Aα = Γ(Uα , OX ), and if we
suppose that the restrictions of uα and u β to an affine open subset V ⊂ Uα ∩ Uβ coincide, then we
see immediately that sα and s β coincide on V; thus the family (sα ) consists of the restrictions to Uα of
a section s of OX over X; conversely, it is clear that such a section defines a family (uα ) of morphisms
which are the restrictions to Uα of an X-morphism X → X ⊗Z Z[ T ]. This result is generalized in
(II, 1.7.12).
3. PRODUCTS OF PRESCHEMES 120
s′ ϕ′
ψ′
Y ( T )S / Z ( T )S
Proposition (3.4.7). — Let Xi (1 ⩽ i ⩽ n) be S-preschemes, and, for each index i, let xi be a point of Xi .
For there to exist a point y of Y = X1 ×S X2 × . . . ×S Xn whose image is xi under the ith projection for each
1 ⩽ i ⩽ n, it is necessary and sufficient that the xi all lie above the same point s of S.
P ROOF. The condition is evidently necessary; Lemma (3.4.6) proves that it is sufficient. □
In other words, if we denote by ( X ) the underlying set of X, we see that we have a canonical
surjective function ( X ×S Y ) → ( X ) ×(S) (Y ); we must point out that this function is not injective in
general; in other words, there can exist multiple distinct points z in X ×S Y that have the same projections
x ∈ X and y ∈ Y; we have already seen this when S, X, and Y are prime spectra of fields k, K, and
K ′ (respectively), since the tensor product K ⊗k K ′ has, in general, multiple distinct prime ideals
(cf. (3.4.9)).
Corollary (3.4.8). — Let f : X → Y be an S-morphism, and f (S′ ) : X(S′ ) → Y(S′ ) the S′ -morphism induced
by f by an extension S′ → S of the base prescheme. Let p (resp. q) be the projection X(S′ ) → X (resp.
Y(S′ ) → Y); for every subset M of X, we have
I | 114
P ROOF. Indeed (3.3.11), X(S′ ) can be identified with the product X ×Y Y(S′ ) thanks to the com-
mutative diagram
p
Xo X( S ′ )
f f (S′ )
q
Yo Y(S′ )
By (3.4.7), the equation q(y′ ) = f ( x ) for x ∈ M and y′ ∈ Y(S) is equivalent to the existence of some
x ′ ∈ X(S′ ) such that p( x ′ ) = x and f (S′ ) ( x ′ ) = y′ , whence the corollary. □
Proposition (3.5.2). —
(i) If f : X → X ′ and g : Y → Y ′ are surjective S-morphisms, then f ×S g is surjective.
(ii) If f : X → Y is a surjective S-morphism, then f (S′ ) is surjective for every extension S′ of the base
prescheme.
P ROOF. The composition of any two surjections being a surjection, it suffices to prove (ii); but
this proposition follows immediately from (3.4.8) applied to M = X. □
Proposition (3.5.3). — For a morphism f : X → Y to be surjective, it is necessary and sufficient that,
for every field K and every morphism Spec(K ) → Y, there exist an extension K ′ of K and a morphism
Spec(K ′ ) → X that make the following diagram commute:
Xo Spec(K ′ )
f
Yo Spec(K ).
P ROOF. The condition is sufficient because, for all y ∈ Y, it suffices to apply it to a morphism
Spec(K ) → Y corresponding to a monomorphism k(y) → K, with K being an extension of k(y)
(2.4.6). Conversely, suppose that f is surjective, and let y ∈ Y be the image of the unique point
of Spec(K ); there exists some x ∈ X such that f ( x ) = y; we will consider the corresponding
monomorphism k(y) → k( x ) (2.2.1); it then suffices to take K ′ to be the extension of k(y) such that
there exist k(y)-monomorphisms from k ( x ) and K to K ′ (Bourbaki, Alg., chap. V, §4, prop. 2); the
morphism Spec(K ′ ) → X corresponding to k ( x ) → K ′ is exactly that for which we are searching. □
With the language introduced in (3.4.5), we can say that every geometric point of Y with values in K
comes from a geometric point of X with values in an extension of K.
3. PRODUCTS OF PRESCHEMES 123
ϕ ϕ′
α′ / Y (K ′ )
X (K ′ )
commutes, where ϕ and ϕ′ come from the morphism Spec(K ′ ) → Spec(K ), and α and α′ corre-
sponding to f . However, ϕ is injective, and so too is α′ , by hypothesis; hence α is necessarily
injective.
Proposition (3.5.6). — Let f : X → Y and g : Y → Z be two morphisms of preschemes.
(i) If f and g are radicial, then so is g ◦ f .
(ii) Conversely, if g ◦ f is radicial, then so is f .
P ROOF. Taking into account Definition (3.5.4), the proposition reduces to the corresponding
claims for the maps X (K ) → Y (K ) → Z (K ), and these claims are evident. □
Proposition (3.5.7). —
(i) If the S-morphisms f : X → X ′ and g : X → X ′ are radicial, then so is f ×S g.
(ii) If the S-morphism f : X → Y is radicial, then so is f (S′ ) : X(S′ ) → Y(S′ ) for every extension
S′ → S of the base prescheme.
P ROOF. Given (3.5.1), it suffices to prove (i). We have seen (3.4.2.1) that
( X ×S Y )(K ) = X (K ) ×S(K) Y (K ),
( X ′ ×S Y ′ )(K ) = X ′ (K ) ×S(K) Y ′ (K ),
with the map ( X ×S Y )(K ) → ( X ′ ×S Y ′ )(K ) corresponding to f ×S g thus being identified with
(u, v) → ( f ◦ u, g ◦ v), and the proposition then follows. □
Proposition (3.5.8). — For a morphism f = (ψ, θ ) : X → Y to be radicial, it is necessary and sufficient
for ψ to be injective and for the monomorphism θ x : k(ψ( x )) → k( x ) to make k( x ) a radicial extension of
k(ψ( x )) for every x ∈ X.
P ROOF. We suppose that f is radicial and first show that the equation ψ( x1 ) = ψ( x2 ) = y
necessarily implies that x1 = x2 . Indeed, there exists a field K, and an extension of k(y), along
with k (y)-monomorphisms k( x1 ) → K and k( x2 ) → K (Bourbaki, Alg., chap. V, §4, prop. 2); the
corresponding morphisms u1 : Spec(K ) → X and u2 : Spec(K ) → X are then such that f ◦ u1 =
f ◦ u2 , and so u1 = u2 by hypothesis, and this implies, in particular, that x1 = x2 . We now consider
k( x ) as the extension of k (ψ( x )) by means of θ x : if k ( x ) is not a radicial algebraically-closed extension,
then there exist two distinct k (ψ( x ))-monomorphisms from k( x ) to an algebraically-closed extension
K of k(ψ( x )), and the two corresponding morphisms Spec(K ) → X would contradict the hypothesis.
Conversely, taking (2.4.6) into account, it is immediate that the conditions stated are sufficient for f
to be radicial. □
Corollary (3.5.9). — If A is a ring, and S is a multiplicative set of A, then the canonical morphism
Spec(S−1 A) → Spec( A) is radicial.
P ROOF. Indeed, this morphism is a monomorphism (1.6.2). □
Corollary (3.5.10). — Let f : X → Y be a radicial morphism, g : Y ′ → Y a morphism, and X ′ =
X(Y ′ ) = X ×Y Y ′ . Then the radicial morphism f (Y ′ ) (3.5.7, ii) is a bijection from the underlying space of X to
g−1 ( f ( X )); further, for every field K, the set X ′ (K ) can be identified with the subset of Y ′ (K ) given by the
inverse image of the map Y ′ (K ) → Y (K ) (corresponding to g) from the subset X (K ) of Y (K ).
3. PRODUCTS OF PRESCHEMES 124
P ROOF. The first claim follows from (3.5.8) and (3.4.8); the second, from the commutativity of
the following diagram: I | 117
X ′ (K ) / Y ′ (K )
X (K ) / Y (K )
□
Remark (3.5.11). — We say that a morphism f = (ψ, θ ) of preschemes is injective if the map ψ
is injective. For a morphism f = (ψ, θ ) : X → Y to be radicial, it is necessary and sufficient that,
for every morphism Y ′ → Y, the morphism f (Y ′ ) : X(Y ′ ) → Y ′ be injective (which justifies the
terminology of a universally injective morphism). In fact, the condition is necessary by (3.5.7, ii) and
(3.5.8). Conversely, the condition implies that ψ is injective; if, for some x ∈ X, the monomorphism
θ x : k (ψ( x )) → k( x ) were not radicial, then there would be an extension K of k (ψ( x )), and two
distinct morphisms Spec(K ) → X corresponding to the same morphism Spec(K ) → Y (3.5.8).
But then, setting Y ′ = Spec(K ), there would be two distinct Y ′ -sections of X(Y ′ ) (3.3.14), which
contradicts the hypothesis that f (Y ′ ) is injective.
3.6. Fibres
Proposition (3.6.1). — Let f : X → Y be a morphism, y a point of Y, and ay an ideal of definition for Oy
for the my -preadic topology. Then the projection p : X ×Y Spec(Oy /ay ) → X is a homeomorphism from
the underlying space of the prescheme X ×Y Spec(Oy /ay ) to the fibre f −1 (y) equipped with the topology
induced from that of the underlying space of X.
P ROOF. Since Spec(Oy /ay ) → Y is radicial ((3.5.4) and (2.4.7)), since Spec(Oy /ay ) is a single
point, and since the ideal my /ay is nilpotent by hypothesis (1.1.12), we already know ((3.5.10) and
(3.3.4)) that p identifies, as sets, the underlying space of X ×Y Spec(Oy /ay ) with f −1 (y); everything
reduces to proving that p is a homeomorphism. By (3.2.7), the question is local on X and Y, and so
we can suppose that X = Spec( B) and Y = Spec( A), with B being an A-algebra. The morphism
p then corresponds to the homomorphism 1 ⊗ ϕ : B → B ⊗ A A′ , where A′ = Ay /ay and ϕ is the
canonical map from A to A′ . Then every element of B ⊗ A A′ can be written as
!
∑ bi ⊗ ϕ(ai )/ϕ(s) = ∑(ai bi ⊗ 1) (1 ⊗ ϕ(s))−1 ,
i i
(3.6.2). Throughout the rest of this treatise, whenever we consider a fibre f −1 (y) of a morphism
as having the structure of a k(y)-prescheme, it will always be the prescheme obtained by transporting
the structure of X ×Y Spec(k(y)) by the projection to X. We will also write this (latter) product as
X ×Y k(y), or X ⊗OY k (y); more generally, if B is an Oy -algebra, we will denote by X ×Y B or
X ⊗OY B the product X ×Y Spec( B).
With the preceding convention, it follows from (3.5.10) that the points of X with values in an
extension K of k (y) are identified with the points of f −1 (y) with values in K.
(3.6.3). Let f : X → Y and g : Y → Z be two morphisms, and h = g ◦ f their composition; for all
z ∈ Z, the fibre h−1 (z) is a prescheme isomorphic to
X × Z Spec(k (z)) = ( X ×Y Y ) × Z Spec(k(z)) = X ×Y g−1 (z).
Inparticular, if U is an open subset of X, then the prescheme induced on U ∩ f −1 (y) by the prescheme I | 118
f −1 (y) is isomorphic to f U−1 (y) ( f U being the restriction of f to U),
Proposition (3.6.4). — (Transitivity of fibres) Let f : X → Y and g : Y ′ → Y be morphisms; let
X ′ = X ×Y Y ′ = X(Y ′ ) and f ′ = f (Y ′ ) : X ′ → Y ′ . For every y′ ∈ Y ′ , if we let y = g(y′ ), then the
prescheme f ′−1 (y′ ) is isomorphic to f −1 (y) ⊗k(y) k(y′ ).
3. PRODUCTS OF PRESCHEMES 125
P ROOF. Indeed, it suffices to remark that the two preschemes ( X ⊗Y k(y)) ⊗k(y) k(y′ ) and
( X ×Y Y ′ ) ⊗Y′ k(y′ ) are both canonically isomorphic to X ×Y Spec(k(y′ )) by (3.3.9.1). □
In particular, if V is an open neighbourhood of y in Y, and we denote by f V the restriction of
f to the induced prescheme on f −1 (V ), then the preschemes f −1 (y) and f V−1 (y) are canonically
identified.
Proposition (3.6.5). — Let f : X → Y be a morphism, y a point of Y, Z the local prescheme Spec(Oy ),
and p = (ψ, θ ) the projection X ×Y Z → X; then p is a homeomorphism from the underlying space of
X ×Y Z to the subspace f −1 ( Z ) of X (when the underlying space of Z is identified with a subspace of
Y, cf. (2.4.2)), and, for all t ∈ X ×Y Z, letting z = ψ(t), we have that θt♯ is an isomorphism from Ox to Ot .
P ROOF. Since Z (identified as a subspace of Y) is contained inside every affine open containing
y (2.4.2), we can, as in (3.6.1), reduce to the case where X = Spec( A) and Y = Spec( B) are affine
schemes, with A being a B-algebra. Then X ×Y Z is the prime spectrum of A ⊗ B By , and this ring
is canonically identified with S−1 A, where S is the image of B − jy in A (0, 1.5.2); since p then
corresponds to the canonical homomorphism A → S−1 A, the proposition follows from (1.6.2). □
3.7. Application: reduction of a prescheme mod. J This section, which makes use of notions and
results from Chapter I and Chapter II, will not be used in what follows in this treatise, and is only intended for
readers familiar with classical algebraic geometry.
(3.7.1). Let A be a ring, X an A-prescheme, and J an ideal of A; then X0 = X ⊗ A ( A/J) is an
( A/J)-prescheme, which we sometimes say is induced from X by reduction mod. J.
. This terminology is used foremost when A is a local ring and J its maximal ideal, in such a way that
X0 is a prescheme over the residue field k = A/J of A.
When A is also integral, with field of fractions K, we can consider the K-prescheme X ′ = X ⊗ A K.
By an abuse of language which we will not use, it has been said, up until now, that X0 is induced by
X ′ by reduction mod. J. In the case where this language was used, A was a local ring of dimension 1
(most often a discrete valuation ring) and it was implied (be it more or less explicitly) that the given
K-prescheme X ′ was a closed subprescheme of a K-prescheme P′ (in fact, a projective space of the
form (?) PrK , cf. (II, 4.1.1)), itself of the form P′ = P ⊗ A K, where P is a given A-prescheme (in fact,
the A-scheme PrA , with the notation of (II, 4.1.1)). In our language, the definition of X0 in terms of
X ′ is formulated as follows:
We consider the affine scheme Y = Spec( A), formed of two points, the unique closed point y = J I | 119
and the generic point (0), the singleton set U of the generic point being thus an open U = Spec(K )
in Y. If X is an A-prescheme (or, in other words, a Y-prescheme), then X ⊗ A K = X ′ is exactly the
prescheme induced by X on ψ−1 (U ), denoting by ψ the structure morphism X → Y. In particular,
if ϕ is the structure morphism P → Y, then a closed subprescheme X ′ of P′ = ϕ−1 (U ) is a (locally
closed) subprescheme of P. If P is Noetherian (for example, if A is Noetherian and P is of finite type
over A), then there exists a smaller closed subprescheme X = X ′ of G that through which X ′ factors
(9.5.10), and X ′ is the prescheme induced by X on the open ϕ−1 (U ) ∩ X, and so is isomorphic to
X ⊗ A K (9.5.10). The immersion of X ′ into P′ = P ⊗ A K thus lets us canonically consider X ′ as being of the
form X ′ = X ⊗ A K, where X is an A-prescheme. We can then consider the reduced mod. J prescheme
X0 = X ⊗ A k, which is exactly the fibre ψ−1 (y) of the closed point y. Up until now, lacking the
adequate terminology, we had avoided explicitly introducing the A-prescheme X. One ought to,
however, note that all the claims normally made about the “reduced mod. J” prescheme X0 should
be seen as consequences of more complicated claims about X itself, and cannot be satisfactorily
formulated or understood except by interpreting them as such. It seems also that any hypotheses
made on X0 always reduce to hypotheses on X itself (independent of the prior data of an immersion
of X ′ in PrK ), which lets us give more intrinsic statements.
(3.7.3). Lastly, we draw attention to a very particular fact, which has undoubtedly contributed to
slowing the conceptual clarification of the situation envisaged here: if A is a discrete valuation ring,
and if X is proper over A (which is indeed the case if X is a closed subprescheme of some PrA , cf.
(II, 5.5.4)), then the points of X with values in A and the points of X ′ with values in k are in bijective
correspondence (II, 7.3.8). This is why we often believe that facts about X ′ have been proved, when
4. SUBPRESCHEMES AND IMMERSION MORPHISMS 126
in reality we have proved facts about X, and these remain valuable (in this form) whenever we no
longer suppose that the base local ring is of dimension 1.
We say that (Y, OY ) is the subprescheme of ( X, OX ) defined by the sheaf of ideals I ; this is a particular
case of the more general notion of a subprescheme:
Definition (4.1.3). — We say that a ringed space (Y, OY ) is a subprescheme of a prescheme ( X, OX )
if:
1st. Y is a locally closed subspace of X;
2nd. if U denotes the largest open subset of X containing Y such that Y is closed in U (equiv-
alently, the complement in X of the boundary of Y with respect to Y), then (Y, OY ) is a
subprescheme of (U, OX |U ) defined by a quasi-coherent sheaf of ideals of OX |U.
We say that the subprescheme (Y, OY ) of ( X, OX ) is closed if Y is closed in X (in which case U = X).
It follows immediately from this definition and Proposition (4.1.2) that the closed subpreschemes
of X are in canonical bijective correspondence with the quasi-coherent sheaves of ideals J of OX , since
if two such sheaves J and J ′ have the same (closed) support Y, and if the restrictions of OX /J
and OX /J ′ to Y are identical, then we have J ′ = J .
(4.1.4). Let (Y, OY ) be a subprescheme of X, U the largest open subset of X such that Y is closed
(and thus contained) in U, and V an open subset of X contained in U; then V ∩ Y is closed in
V. In addition, if Y is defined by the quasi-coherent sheaf of ideals J of OX |U, then J |V is a
quasi-coherent sheaf of ideals of OX |V, and it is immediate that the prescheme induced by Y on
Y ∩ V is the closed subprescheme of V defined by the sheaf of ideals J |V. Conversely:
Proposition (4.1.5). — Let (Y, OY ) be a ringed space such that Y is a subspace of X, and there exists a
cover (Vα ) of Y by open subsets of X such that, for each α, Y ∩ Vα is closed in Vα , and the ringed space
(Y ∩ Vα , OY |(Y ∩ Vα )) is a closed subprescheme of the prescheme induced on Vα by X. Then (Y, OY ) is a
subprescheme of X.
P ROOF. The hypotheses imply that Y is locally closed in X and that the largest open U in which
Y is closed (and thus contained) contains all the Vα ; we can thus reduce to the case where U = X and
Y is closed in X. We then define a quasi-coherent sheaf of ideals J of OX by taking J |Vα to be the
sheaf of ideals of OX |Vα which define the closed subprescheme (Y ∩ Vα , OY |(Y ∩ Vα )), and, for each
open subset W of X not intersecting Y, J |W = OX |W. We see immediately, by Definition (4.1.3) and
(4.1.4), that there exists a unique sheaf of ideals J satisfying these conditions, and that it defines the
closed subprescheme (Y, OY ). □
4. SUBPRESCHEMES AND IMMERSION MORPHISMS 127
the first arrow being the canonical homomorphism. Let ψ′ be the continuous map Z → Y coinciding
∗
with ψ; it is clear that we have ψ′ (OY ) = ψ∗ (OX /J ); on the other hand, ω is evidently a local
homomorphism, so g = (ψ′ , ω ♭ ) is a morphism of preschemes Z → Y (2.2.1), which, according to I | 122
the above, is such that f = j ◦ g, hence the proposition. □
Corollary (4.1.10). — For an injection morphism Z → X to be factor through the injection morphism
Y → X, it is necessary and sufficient for Z to be a subprescheme of Y.
We then write Z ⩽ Y, and this condition is evidently an ordering on the set of subpreschemes of
X.
8[Trans.] There doesn’t seem to be an English equivalent of this, except for ‘bounded above’, which doesn’t make much sense in this
context. We would normally just say that ‘ f factors through j’, but to avoid having to entirely restructure the often-lengthy sentences in
the original, we sometimes (but as little as we can) use ‘majoré’.
4. SUBPRESCHEMES AND IMMERSION MORPHISMS 128
The subprescheme Z and the isomorphism g are then determined in a unique way, since if Z ′
is a second subprescheme of X, j′ the injection Z ′ → X, and g′ an isomorphism Y → Z ′ such that
−1
j ◦ g = j′ ◦ g′ , then we have j′ = j ◦ g ◦ g′ , hence Z ′ ⩽ Z (4.1.10), and we can similarly show that
Z ⩽ Z ′ , hence Z ′ = Z, and, since j is a monomorphism of preschemes, g′ = g.
We say that f = j ◦ g is the canonical factorization of the immersion f , and the subprescheme Z
and the isomorphism g are those associated to f .
It is clear that an immersion is a monomorphism of preschemes (4.1.7) and a fortiori a radicial
morphism (3.5.4).
Proposition (4.2.2). —
(a) For a morphism f = (ψ, θ ) : Y → X to be an open immersion, it is necessary and sufficient for ψ to
be a homeomorphism from Y to some open subset of X, and, for all y ∈ Y, that the homomorphism
θy♯ : Oψ(y) → Oy be bijective.
(b) For a morphism f = (ψ, θ ) : Y → X to be an immersion (resp. a closed immersion), it is necessary
and sufficient for ψ to be a homeomorphism from Y to some locally closed (resp. closed) subset of X,
and, for all y ∈ Y, that the homomorphism θy♯ : Oψ(y) → Oy be surjective.
P ROOF.
(a) The conditions are evidently necessary. Conversely, if they are satisfied, then it is clear that
θ ♯ is an isomorphism from OY to ψ∗ (OX ), and ψ∗ (OX ) is the sheaf induced by “transport
of structure” via ψ−1 from OX |ψ(Y ); hence the conclusion.
(b) The conditions are evidently necessary—we prove that they are sufficient. Consider first
the particular case where we suppose that X is an affine scheme, and that Z = ψ(Y ) is
closed in X. We then know (0, 3.4.6) that ψ∗ (OY ) has support equal to Z, and that, denoting
its restriction to Z by OZ′ , the ringed space ( Z, OZ′ ) is induced from (Y, OY ) by transport
of structure via the homeomorphism ψ considered as a map from Y to Z. Let us now
show that f ∗ (OY ) = ψ∗ (OY ) is a quasi-coherent OX -module. Indeed, for all x ̸∈ Z, ψ∗ (OY )
restricted to a suitable neighbourhood of x is zero. On the contrary, if z ∈ Z, then we have
x = ψ(y) for some well-defined y ∈ Y; let V be an affine open neighbourhood of y in Y;
ψ(V ) is then open in Z, and so equal to the intersection of Z with an open subset U of X,
and the restriction of U to ψ∗ (OY ) is identical to the restriction of U to the direct image
(ψV )∗ (OY |V ), where ψV is the restriction of ψ to V. The restriction of the morphism (ψ, θ ) I | 123
to (V, OY |V ) is a morphism from this aforementioned prescheme to ( X, OX ), and, as a
e), where ϕ is the homomorphism from the ring A = Γ( X, OX ) to
result, is of the form ( a ϕ, ϕ
the ring Γ(V, OY ) (1.7.3); we conclude that (ψV )∗ (OY |V ) is a quasi-coherent OX -module
(1.6.3), which proves our assertion, due to the local nature of quasi-coherent sheaves. In
addition, the hypothesis that ψ is a homeomorphism implies (0, 3.4.5) that, for all y ∈ Y, ψy
is an isomorphism (ψ∗ (OY ))ψ(y) → Oy ; since the diagram
θψ(y)
Oψ ( y ) / (ψ∗ (OY ))
ψ(y)
ψy ◦ρψ(y) ψy
♯
θy
(ψ∗ (OX ))y / Oy
is commutative, and the vertical arrows are the isomorphisms (0, 3.7.2), the hypothesis
that θy♯ is surjective implies that θψ(y) is surjective as well. Since the support of ψ∗ (OY )
4. SUBPRESCHEMES AND IMMERSION MORPHISMS 129
Corollary (4.2.4). —
(a) Let f be a morphism Y → X, and (Vλ ) a cover of f (Y ) by open subsets of X. For f to be an
immersion (resp. an open immersion), it is necessary and sufficient for its restriction to each of the I | 124
induced preschemes f −1 (Vλ ) to be an immersion (resp. an open immersion) into Vλ .
(b) Let f be a morphism Y → X, and (Vλ ) an open cover of X. For f to be a closed immersion, it is
necessary and sufficient for its restriction to each of the induced preschemes f −1 (Vλ ) to be a closed
immersion into Vλ .
P ROOF. Let f = (ψ, θ ); in the case (a), θy♯ is surjective (resp. bijective) for all y ∈ Y, and in the
case (b), θy♯ is surjective for all y ∈ Y; it thus suffices to check, in case (a), that ψ is a homeomorphism
from Y to a locally closed (resp. open) subset of X, and, in case (b), that ψ is a homeomorphism from
Y to a closed subset of X. Now ψ is evidently injective, and sends each neighbourhood of y in Y to a
neighbourhood of ψ(y) is ψ(Y ) for all y ∈ Y, by virtue of the hypothesis; in case (a), ψ(Y ) ∩ Vλ is
locally closed (resp. open) in Vλ , so ψ(Y ) is locally closed (resp. open) in the union of the Vλ , and a
fortiori in X; in case (b), ψ(Y ) ∩ Vλ is closed in Vλ , so ψ(Y ) is closed in X since X = λ Vλ .
S
□
Proposition (4.2.5). — The composition of any two immersions (resp. of two open immersions, of two closed
immersions) is an immersion (resp. an open immersion, a closed immersion).
P ROOF. This follows easily from (4.1.6). □
4. SUBPRESCHEMES AND IMMERSION MORPHISMS 130
P ROOF. This follows immediately from Proposition (4.4.1) and Corollary (4.1.10). □
Corollary (4.4.4). — Let f : X → Y be a morphism, and Y′ and Y ′′ subpreschemes of Y; then we have
f −1 (inf(Y ′ , Y ′′ )) = inf( f −1 (Y ′ ), f −1 (Y ′′ )).
P ROOF. This follows from the existence of the canonical isomorphism between ( X ×Y Y ′ ) × X
( X ×Y Y ′′ ) and X ×Y (Y ′ ×Y Y ′′ ) (3.3.9.1). □
Proposition (4.4.5). — Let f : X → Y be a morphism, and Y ′ a closed subprescheme of Y defined by a
quasi-coherent sheaf of ideals K of OY (4.1.3); the closed subprescheme f −1 (Y ′ ) of X is then defined by the
quasi-coherent sheaf of ideals f ∗ (K )OX of OX .
P ROOF. The statement is evidently local on X and Y; it thus suffices to note that if A is a
B-algebra and K an ideal of B, then we have A ⊗ B ( B/K) = A/KA, and to then apply (1.6.9). □ ErrII
Proposition (4.5.5). —
(i) The composition of any two local immersions (resp. of two local isomorphisms) is a local immersion
(resp. a local isomorphism).
(ii) Let f : X → X ′ and g : Y → Y ′ be two S-morphisms. If f and g are local immersions (resp. local
isomorphisms), then so too is f ×S g.
(iii) If an S-morphism f is a local immersion (resp. a local isomorphism), then so too is f (S′ ) for every
extension S′ → S of the base prescheme.
P ROOF. According to (3.5.1), it suffices to prove (i) and (ii).
(i) follows immediately from the transitivity of closed (resp. open) immersions (4.2.5) and from
the fact that if f is a homeomorphism from X to a closed subset of Y, then for every open U ⊂ X,
f (U ) is open in f ( X ), so there exists an open subset V of Y such that f (U ) = V ∩ f ( X ), and, as a
result, f (U ) is closed in V.
To prove (ii), let p and q be the projections from X × X Y, and p′ and q′ the projections from
X ×S Y ′ . There exists, by hypothesis, open neighbourhoods U, U ′ , V, and V ′ of x = p(z), x ′ = p′ (z′ ),
′
y = q(z), and y′ = q′ (z′ ) (respectively), such that the restrictions of f and g to U and V (respectively)
are closed (resp. open) immersions into U ′ and V ′ (respectively). Since the underlying spaces of U ×S
−1
V and U ′ ×S V ′ can be identified with the open neighbourhoods p−1 (U ) ∩ q−1 (V ) and p′ (U ′ ) ∩
− 1
q′ (V ′ ) of z and z′ (respectively) (3.2.7), the proposition follows from Proposition (4.3.1). □
We say that the quasi-coherent OX -module N thus defined is the nilradical of the OX -algebra B;
in particular, we denote by NX the nilradical of OX .
Corollary (5.1.2). — Let X be a prescheme; the closed subprescheme of X defined by the sheaf of ideals NX
is the only reduced subprescheme (0, 4.1.4) of X that has X as its underlying space; it is also the smallest
subprescheme of X that has X as its underlying space.
P ROOF. Since the structure sheaf of the closed subprescheme of Y defined by NX is OX /NX , it
is immediate that Y is reduced and has X as its underlying space, because Nx ̸= Ox for any x ∈ X.
To show the other claims, note that a subprescheme Z of X that has X as its underlying space is
defined by a sheaf of ideals I (4.1.3) such that Ix ̸= Ox for any x ∈ X. We can restrict to the case
where X is affine, say X = Spec( A) and I = e I, where I is an ideal of A; then, for every x ∈ X, we
have Ix ⊂ jx , and so I is contained in every prime ideal of A, and so also in their intersection N, the
nilradical of A. This proves that Y is the small subprescheme of X that has X as its underlying space
(4.1.9); furthermore, if Z is distinct from Y, we necessarily have Ix ̸= Nx for at least one x ∈ X, and
so (5.1.1) Z is not reduced. □
Proposition (5.1.4). — For the prime spectrum of a ring A to be a reduced (resp. integral) prescheme (2.1.7),
it is necessary and sufficient for A to be a reduced (resp. integral) ring.
P ROOF. Indeed, it follows immediately from (5.1.1) that the condition N = (0) is necessary
and sufficient for X = Spec( A) to be reduced; the claim corresponding to integral rings is then a
consequence of (1.1.13). □
Since every ring of fractions ̸= {0} of an integral ring is integral, it follows from (5.1.4) that,
for every locally integral prescheme X, Ox is an integral ring for every x ∈ X. The converse is true
whenever the underlying space of X is locally Noetherian: indeed, X is then reduced, and if U is an
affine open subset of X, which is a Noetherian space, then U has only a finite number of irreducible
components, and so its ring A has only a finite number of minimal prime ideals (1.1.14). If two of
the irreducible components of U had a common point x, then Ox would have at least two distinct
minimal prime ideals, and would thus not be integral; the components of U are thus open subsets
that are pairwise disjoint, and each of them is thus integral.
f red
Xred / Yred
f
X /Y
is commutative, where the vertical arrows are the injection morphisms; in other words, Xred → X is
a functorial morphism. We note in particular that, if X is reduced, then every morphism f : X → Y
f red
factors as X −−→ Yred → Y; in other words, f factors through the injection morphism Yred → Y.
5. REDUCED PRESCHEMES; THE SEPARATION CONDITION 134
P ROOF. The proposition is trivial if f is surjective; if f is radicial, then the proposition follows
from the fact that, for every x ∈ X, the field k( x ) is the same for the preschemes X and Xred (3.5.8).
Finally, if f = (ψ, θ ) is an immersion, a closed immersion, or a local immersion (resp. an open
immersion, or a local isomorphism), then the proposition follows from the fact that, if θ x♯ is surjective
(resp. bijective), then so too is the homomorphism obtained by passing to the quotients by the
nilradicals Oψ( x) and Ox ((5.1.2) and (4.2.2)) (cf. (5.5.12)). □
Proposition (5.1.7). — If X and Y are S-preschemes, then the preschemes Xred ×Sred Yred and Xred ×S Yred
are identical, and canonically identified with a subprescheme of X ×S Y that has the same underlying subspace
as the two aforementioned products.
P ROOF. The canonical identification of Xred ×S Yred with a subprescheme of X ×S Y that has the
same underlying space follows from (4.3.1). Furthermore, if ϕ and ψ are the structure morphisms
Xred → S and Yred → S (respectively), then they factor through Sred (5.1.5), and since Sred → S is a
monomorphism, the first claim of the proposition follows from (3.2.4). □
Corollary (5.1.8). — The preschemes ( X ×S Y )red and ( Xred ×Sred Yred )red are canonically identified with
one another.
We note that, even if X and Y are reduced preschemes, X ×S Y might not be reduced, because
the tensor product of two reduced algebras can have nilpotent elements.
Proposition (5.1.9). — Let X be a prescheme, and I a quasi-coherent sheaf of ideals of OX such that I | 130
I n = 0 for some integer n > 0. Let X0 be the closed subprescheme ( X, OX /I ) of X; for X to be an affine
scheme, it is necessary and sufficient for X0 to be an affine scheme.
A = Γ( X, OX )
A0 = Γ( X0 , OX0 ) = Γ( X, OX /I ).
(5.1.9.1) 0 −→ Γ( X, I ) −→ Γ( X, OX ) −→ Γ( X, OX /I ) −→ 0
is an exact sequence. We now prove, assuming that this is true, the proposition. Note that K =
Γ( X, I ) is an ideal whose square is zero in A, and thus a module over A0 = A/K. By hypothesis,
we have X0 = Spec( A), and, since the underlying spaces of X0 and X are identical, K = Γ( X0 , I );
Additionally, since I 2 = 0, I is a quasi-coherent (OX /I )-module, so we have I ∼ = Ke and
′
Kx = Ix for all x ∈ X0 (1.4.1). With this in mind, let X = Spec( A), and consider the morphism
f = (ψ, θ ) : X → X ′ of preschemes that corresponds to the identity map A → Γ( X, OX ) (2.2.4). For
every affine open subset V of X, the diagram
A / Γ(V, OX |V )
A0 = A/K / Γ(V, OX |V )
0
5. REDUCED PRESCHEMES; THE SEPARATION CONDITION 135
P ROOF. There is indeed an open subset U of Y such that X = U ∩ X; since, by (5.2.2), X is I | 132
a reduced subprescheme of Z, the subprescheme X is induced by Z on the open subspace X by
uniqueness (5.2.1). □
Corollary (5.2.4). — Let f : X → Y be a morphism, and X ′ (resp. Y ′ ) a closed subprescheme of X (resp. Y)
defined by a quasi-coherent sheaf of ideals J (resp. K ) of OX (resp. OY ). Suppose that X ′ is reduced, and
that f ( X ′ ) ⊂ Y ′ . Then f ∗ (K )OX ⊂ J .
P ROOF. Since, by (5.2.2), the restriction of f to X ′ factors as X ′ → Y ′ → Y, it suffices to apply
(4.4.6). □
5.3. Diagonal; graph of a morphism
(5.3.1). Let X be an S-prescheme; we define the diagonal morphism of X in X ×S X, denoted by ∆ X |S ,
or ∆ X , or even ∆ if no confusion may arise, to be the S-morphism (1X , 1X )S , or, in other words, the
unique S-morphism ∆ X such that
(5.3.1.1) p1 ◦ ∆ X = p2 ◦ ∆ X = 1 X ,
where p1 and p2 are the projections of X ×S X (Definition (3.2.1)). If f : T → X and g : T → Y are
S-morphisms, we immediately have that
(5.3.1.2) ( f , g )S = ( f ×S g ) ◦ ∆ T |S .
The reader will note that the preceding definition and the results stated in (5.3.1) to (5.3.8) are
true in any category, provided that the products used within exist in the category.
Proposition (5.3.2). — Let X and Y be S-preschemes; if we make the canonical identification between
( X × Y ) × ( X × Y ) and ( X × X ) × (Y × Y ), then the morphism ∆ X ×Y is identified with ∆ X × ∆Y .
P ROOF. Indeed, if p1 : X × X → X and q1 : Y × Y → Y are the projections onto the first
component, then the projection onto the first component ( X × Y ) × ( X × Y ) → X × Y is identified
with p1 × q1 , and we have
( p 1 × q 1 ) ◦ ( ∆ X × ∆ Y ) = ( p 1 ◦ ∆ X ) × ( q 1 ◦ ∆ Y ) = 1 X ×Y
and we can argue similarly for the projections onto the second component. □
.
Corollary (5.3.4). — For every extension S′ → S of the base prescheme, ∆ XS′ is canonically identified with
( ∆ X )(S′ ) .
P ROOF. It suffices to remark that ( X ×S X )(S′ ) is canonically identified with X(S′ ) ×S′ X(S′ )
(3.3.10). □
Proposition (5.3.5). — Let X and Y be S-preschemes, and ϕ : S → T a morphism of preschemes, which lets
us consider every S-prescheme as a T-prescheme. Let f : X → S and g : Y → S be the structure morphisms,
p and q the projections of X ×S Y, and π = f ◦ p = g ◦ q the structure morphism X ×S Y → S. Then the
diagram
( p,q) T
(5.3.5.1) X ×S Y / X ×T Y
π f ×T g
∆S| T
S / S ×T S
commutes, and identifies X ×S Y with the product of the (S × T S)-preschemes S and X × T Y, and the I | 133
projections with π and ( p, q) T .
P ROOF. By (3.4.3), we are led to proving the corresponding proposition in the category of sets,
replacing X, Y, and S by X ( Z ) T , Y ( Z ) T , and S( Z ) T (respectively), with Z being an arbitrary T-
prescheme. But, in the category of sets, the proof is immediate and left to the reader. □
Corollary (5.3.6). — The morphism ( p, q) T can be identified (letting P = S × T S) with 1X ×T Y × P ∆S .
5. REDUCED PRESCHEMES; THE SEPARATION CONDITION 137
f f ×S 1Y
∆Y
Y / Y ×S Y
f f ×S f
∆Y
Y / Y ×S Y
fi
Y /X
commute. The diagram
( g1 ,g2 )S
Spec(k (y)) / Spec(k( x )) ×S Spec(k ( x ))
( f 1 , f 2 )S
Y / X ×S X
thus also commutes. But it follows from the equality g1 = g2 that the image under ( g1 , g2 )S of the
unique point of Spec(k (y)) belongs to the diagonal of Spec(k( x )) ×S Spec(k( x )); the conclusion then
follows from (5.3.15). □
5. REDUCED PRESCHEMES; THE SEPARATION CONDITION 139
Proposition (5.5.1). —
(i)Every monomorphism of preschemes (and, in particular, every immersion) is a separated morphism.
(ii)The composition of any two separated morphisms is separated.
(iii)If f : X → X ′ and g : Y → Y ′ are separated S-morphisms, then f ×S g is separated.
(iv) If f : X → Y is a separated S-morphism, then the S′ -morphism f (S′ ) is separated for every extension
S′ → S of the base prescheme.
(v) If the composition g ◦ f is separated, then f is separated.
(vi) For a morphism f to be separated, it is necessary and sufficient for f red (5.1.5) to be separated.
P ROOF. Note that (i) is an immediate consequence of (5.3.8). If f : X → Y and g : Y → Z are
morphisms, then the diagram
∆X|Z
(5.5.1.1) X / X ×Z X
9
∆ X |Y # j
X ×Y X
commutes (where j denotes the canonical immersion (5.3.10)), as can be immediately verified. If f
and g are separated, then ∆ X |Y is a closed immersion by definition, and j is a closed immersion by
(5.4.2), whence ∆ X | Z is a closed immersion by (4.2.4), which proves (ii). Given (i) and (ii), (iii) and (iv) I | 137
are equivalent (3.5.1), so it suffices to prove (iv). But X(S′ ) ×Y(S′ ) X(S′ ) is canonically identified with
( X ×Y X ) ×Y Y(S′ ) by (3.3.11) and (3.3.9.1), and we immediately see that the diagonal morphism
∆ X(S′ ) can then be identified with ∆ X ×Y 1Y(S′ ) ; the proposition then follows from (4.3.1).
Γf p2
To prove (v), consider, as in (5.3.13), the factorisation X −→ X × Z Y −→ Y of f , noting that
p2 = ( g ◦ f ) × Z 1Y ; the hypothesis (that g ◦ f is separated) implies that g2 is separated, by (iii) and
(i), and, since Γ f is an immersion, Γ f is separated, by (i), whence f is separated, by (ii). Finally, to
prove (vi), recall that the preschemes Xred ×Yred Xred and Xred ×Y Xred are canonically identified
with one another (5.1.7); if we denote by j the injection Xred → X, then the diagram
∆X
Xred / Xred ×Y Xred
red
j j ×Y j
∆X
X / X ×Y X
commutes (5.3.15), and the proposition follows from the fact that the vertical arrows are homeomor-
phisms of the underlying spaces (4.3.1). □
Corollary (5.5.3). — If X and Y are S-preschemes such that Y is separated over S, then X ×S Y is separated
over X.
P ROOF. This is a particular case of (5.5.1, iv). □
Proposition (5.5.4). — Let X be a prescheme, and assume that its underlying space is a finite union of closed
subsets Xk (1 ⩽ k ⩽ n); for each k, consider the reduced subprescheme of X that has Xk as its underlying
space (5.2.1), and denote this again by Xk . Let f : X → Y be a morphism, and, for each k, let Yk be a closed
subset of Y such that f ( Xk ) ⊂ Yk ; we again denote by Yk the reduced subprescheme of Y that has Yk as its
fk
underlying space, so that the restriction Xk → Y of f to Xk factors as Xk −→ Yk → Y (5.2.2). For f to be
separated, it is necessary and sufficient for all the f k to be separated.
5. REDUCED PRESCHEMES; THE SEPARATION CONDITION 141
P ROOF. The necessity follows from (5.5.1, i, ii, and v). Conversely, if the condition of the
statement is satisfied, then each of the restrictions Xk → Y of f is separated (5.5.1, (i) and (ii)); if p1
and p2 are the projections of X ×Y X, then the subspace ∆ Xk ( Xk ) can be identified with the subspace
∆ X ( X ) ∩ p1−1 ( Xk ) of the underlying space of X ×Y X (5.3.16); these subspaces are closed in X ×Y X,
and thus so too is their union ∆ X ( X ). □
Suppose, in particular, that the Xk are the irreducible components of X; then we can suppose that
the Yk are the irreducible components of Y (0, 2.1.5); Proposition (5.5.4) then, in this case, leads to the
idea of separation in the case of integral preschemes (2.1.7).
I | 138
Proposition (5.5.5). — Let (Yλ ) be an open cover of a prescheme Y; for a morphism f : X → Y to be
separated, it is necessary and sufficient for each of its restrictions f −1 (Yλ ) → Yλ to be separated.
P ROOF. If we set Xλ = f −1 (Yλ ), everything reduces, by taking (4.2.4, b) and the identification
of the products Xλ ×Y Xλ and Xλ ×Yλ Xλ into account, to proving that the Xλ ×Y Xλ form a cover
of X ×Y X. But if we set Yλµ = Yλ ∩ Yµ and Xλµ = Xλ ∩ Xµ = f −1 (Yλµ ), then Xλ ×Y Xµ can be
identified with the product Xλµ ×Yλµ Xλµ (3.2.6.4), and so also with Xλµ ×Y Xλµ (3.2.5), and finally
with an open subset of Xλ ×Y Xλ , which proves our claim (3.2.7). □
Proposition (5.5.4) allows us, by taking an affine open cover of Y, to restrict our study of
separated morphisms to just those that take values in affine schemes.
Proposition (5.5.6). — Let Y be an affine scheme, X a prescheme, and (Uα ) a cover of X by affine open
subsets. For a morphism f : X → Y to be separated, it is necessary and sufficient for Uα ∩ Uβ to be, for every
pair of indices (α, β), an affine open subset, and for the ring Γ(Uα ∩ Uβ , OX ) to be generated by the union of
the canonical images of the rings Γ(Uα , OX ) and Γ(Uβ , OX ).
P ROOF. The Uα ×Y Uβ form an open cover of X ×Y X (3.2.7); denoting the projections of X ×Y X
by p and q, we have
∆− 1 −1 −1 −1
X (Uα ×Y Uβ ) = ∆ X ( p (Uα ) ∩ q (Uβ ))
= ∆− 1 −1 −1 −1
X ( p (Uα )) ∩ ∆ X ( q (Uβ )) = Uα ∩ Uβ ;
Corollary (5.5.7). — An affine scheme is separated (and is thus a scheme, which justifies the terminology
of (5.4.1)).
Corollary (5.5.8). — Let Y be an affine scheme; for f : X → Y to be a separated morphism, it is necessary
and sufficient for X to be separated (in other words, for X to be a scheme).
P ROOF. Indeed, we see that the criteria of (5.5.6) do not depend on f . □
Corollary (5.5.9). — For a morphism f : X → Y to be separated, it is necessary and sufficient for the
induced prescheme f −1 (U ) to be separated, for every open subset of U on which Y induces a separated
prescheme, and it is sufficient for it to be the case for every affine open subset U ⊂ Y.
P ROOF. The necessity of the condition follows from (5.5.4) and (5.5.1, ii); the sufficiency follows
from (5.5.4) and (5.5.8), taking into account the existence of affine open covers of Y. □
Examples (5.5.11). — The prescheme from Example (2.3.2) (“the projective line over a field K”) is
separated, because, for the cover ( X1 , X2 ) of X by affine open subsets, X1 ∩ X2 = U12 is affine, and
Γ(U12 , OX ), the ring of rational fractions of the form f (s)/sm with f ∈ K [s], is generated by K [s] and
1/s, so the conditions of (5.5.6) are satisfied.
With the same choice of X1 , X2 , U12 , and U21 as in Example (2.3.2), now take u12 to be the
isomorphism which sends f (s) to f (t); we now obtain, by gluing, a non-separated integral prescheme
X, because the first condition of (5.5.6) is satisfied, but not the second. It is immediate here that
Γ( X, OX ) → Γ( X1 , OX ) = K [s] is an isomorphism; the inverse isomorphism defines a morphism
f : X → Spec(K [s]) that is surjective, and for every y ∈ Spec(K [s]) such that jy ̸= (0), f −1 (y)
consists of a single point, but for jy = (0), f −1 (y) consists of two distinct points (we say that X is the
“affine line over K with the point 0 doubled”).
We can also give examples where neither of the two conditions of (5.5.6) are satisfied. First note
that, in the prime spectrum Y of the ring A = K [s, t] of polynomials in two indeterminates over a
field K, the open subset U given by the union of D (s) and D (t) is not an affine open subset. Indeed, if
z is a section of OY over U, there exist two integers m, n ⩾ 0 such that sm z and tn z are the restrictions
of polynomials in s and t to U (1.4.1), which is clearly possible only if the section z extends to a
section over the whole of Y, identified with a polynomial in s and t. If U were an affine open subset,
then the injection morphism U → Y would be an isomorphism (1.7.3), which is a contradiction,
since U ̸= Y.
With the above in mind, take two affine schemes Y1 and Y2 , prime spectra of the rings A1 =
K [s1 , t2 ] and A2 = K [s2 , t2 ] (respectively); take U12 = D (s1 ) ∪ D (t1 ) and U21 = D (s2 ) ∪ D (t2 ), and
take u12 to be the restriction of an isomorphism Y2 → Y1 to U21 corresponding to the isomorphism
of rings that sends f (s1 , t1 ) to f (s2 , t2 ); we then have an example where the conditions of (5.5.6) are
not satisfied (the integral prescheme thus obtained is called “the affine plane over K with the point 0
doubled”).
Remark (5.5.12). — Given some property P of morphisms of preschemes, consider the following
propositions.
If we suppose that (i) and (ii) are both true, then (iii) and (iv) are equivalent, and (v) and (vi) are
consequences of (i), (ii), and (iii).
Γf
The first claim has already been shown (3.5.1). Consider the factorisation (5.3.13) X −→ X × Z
p2
Y −→ Y of f ; the equation p2 = ( g ◦ f ) × Z 1Y shows that, if g ◦ f has property P, then so too does
p2 , by (iii); if g is separated, then Γ f is a closed immersion (5.4.3), and so also has property P, by (i);
finally, by (ii), f has property P.
6. FINITENESS CONDITIONS 143
f
X / Y,
where the vertical arrows are the closed immersions (5.1.5), and thus have property P, by (i). The
f red
hypothesis that f has property P implies, by (ii), that Xred −−→ Yred → Y has property P; finally,
since a closed immersion is separated (5.5.1, i), f red has property P, by (v).
Note that, if we consider the propositions
(i’) Every immersion has property P;
(v’) If g ◦ f has property P, then so too does f ;
then the above arguments show that (v’) is a consequence of (i’), (ii), and (iii).
(5.5.13). Note that (v) and (vi) are again consequences of (i), (iii), and
(ii’) If j : X → Y is a closed immersion, and g : Y → Z is a morphism that has property P, then g ◦ j
has property P.
Similarly, (v’) is a consequence of (i’), (iii), and
(ii”) If j : X → Y is an immersion, and g : Y → Z is a morphism that has property P, then g ◦ j has
property P.
This follows immediately from the arguments of (5.5.12).
the inverse image of ean under the canonical injection Vi → X (0, 5.1.4); e an |Vi is thus of the form e ani ,
where ani is an ideal of Ai (1.3.7). Since Ai is Noetherian, the sequence (ani ) is stable for all i, whence
the proposition. □
We note that the above argument proves also that if X is a Noetherian prescheme, then every
increasing sequence of coherent sheaves of ideals of OX is stable (?).
Proposition (6.1.4). — Every subprescheme of a Noetherian (resp. locally Noetherian) prescheme is
Noetherian (resp. locally Noetherian).
P ROOF. If suffices to give a proof for a Noetherian prescheme X; further, by definition (6.1.1),
we can also restrict to the case where X is an affine scheme. Since every subprescheme of X is a
closed subprescheme of a prescheme induced on an open subset (4.1.3), we can restrict to the case of
a subprescheme Y, either closed or induced on an open subset of X. The proof in the case where Y is
closed is immediate, since if A is the ring of X, we know that Y is an affine scheme given by the ring
A/J, where J is an ideal of A (4.2.3); since A is Noetherian (6.1.3), so too is A/J.
Now suppose that Y is open in X; the underlying space of Y is Noetherian (6.1.2), hence quasi-
compact, and thus a finite union of open subsets D ( f i ) ( f i ∈ A); everything reduces to showing the
proposition in the case where Y = D ( f ) with f ∈ A. But then Y is an affine scheme whose ring is
isomorphic to A f (1.3.6); since A is Noetherian (6.1.3), so too is A f . □
(6.1.5). We note that the product of two Noetherian S-preschemes is not necessarily Noetherian, even
if the preschemes are affine, since the tensor product of two Noetherian algebras in not necessarily a
Noetherian ring (cf. (6.3.8)).
Proposition (6.1.6). — If X is a Noetherian prescheme, the nilradical NX of OX is nilpotent.
P ROOF. We can in fact cover X with a finite number of affine open subsets Ui , and it suffices
to prove that there exists whole numbers ni such that (NX |Ui )ni = 0; if n is the largest of the ni ,
then we will have NXn = 0. We can thus restrict to the case where X = Spec( A) is affine, with A
a Noetherian ring; by (5.1.1) and (1.3.13), it suffices to observe that the nilradical of A is nilpotent
([Sam53b, p. 127, cor. 4]). □
Corollary (6.1.7). — Let X be a Noetherian prescheme; for X to be an affine scheme, it is necessary and
sufficient that Xred be affine.
P ROOF. This follows from (6.1.6) and (5.1.10). □
Lemma (6.1.8). — Let X be a topological space, x a point of X, and U an open neighbourhood of x having
only a finite number of irreducible components. Then there exists a neighbourhood V of x such that every
open neighbourhood of x contained in V is connected.
P ROOF. Let Ui (1 ⩽ i ⩽ m) be the irreducible components of U not containing x; the complement
(in U) of the union of the Ui is an open neighbourhood V of X inside U, and thus so too in X; it is
also, incidentally, the complement (in X) of the union of the irreducible components of X that do
not contain x (0, 2.1.6). So let W be an open neighbourhood of X contained in V. The irreducible
components of W are the intersections of W with the irreducible components of U (0, 2.1.6), so these
components contain x; since they are connected, so too is W. □
Corollary (6.1.9). — A locally Noetherian topological space is locally connected (which implies, amongst
other things, that its connected components are open).
Proposition (6.1.10). — Let X be a locally Noetherian topological space. The following conditions are
equivalent.
(a) The irreducible components of X are open.
(b) The irreducible components of X are exactly its connected components.
(c) The connected components of X are irreducible.
(d) Two distinct irreducible components of X have an empty intersection.
Finally, if X is a prescheme, then these conditions are also equivalent to
(e) For every x ∈ X, Spec(Ox ) is irreducible (or, in other words, the nilradical of Ox is prime).
6. FINITENESS CONDITIONS 145
P ROOF. It is immediate that (a) implies (b), because an irreducible space is connected, and (a)
implies that the irreducible components of X are the sets that are both open and closed. It is trivial
that (b) implies (c); conversely, a closed set F containing a connected component C of X, with C I | 143
distinct from F, cannot be irreducible, because not being connected means that F is the union of
two disjoint nonempty sets that are both open and closed in F, and thus closed in X; as a result, (c)
implies (b). We immediately conclude from this that (c) implies (d), since two distinct connected
components have no points in common.
We have not yet used the fact that X is locally Noetherian. Suppose now that this is indeed the
case, and we will show that (d) implies (a): by (0, 2.1.6), we can restrict ourselves to the case where
the space X is Noetherian, and so has only a finite number of irreducible components. Since they are
closed and pairwise disjoint, they are open.
Finally, the equivalence between (d) and (e) holds true even without the assumption that the
underlying space of the prescheme X is locally Noetherian. We can in fact restrict ourselves to
the case where X = Spec( A) is affine, by (0, 2.1.6); to say that x is contained in only one single
irreducible component of X is to say that jx contains only one single minimal ideal of A (1.1.14),
which is equivalent to saying that jx Ox contains only one single minimal ideal of Ox , whence the
conclusion. □
Corollary (6.1.11). — Let X be a locally Noetherian space. For X to be irreducible, it is necessary and
sufficient that X be connected and nonempty, and that any two distinct irreducible components of X have
an empty intersection. If X is a prescheme, this latter condition is equivalent to asking that Spec(Ox ) be
irreducible for all x ∈ X.
P ROOF. The second claim has already been shown in (6.1.10); the only thing thus remaining to
show is that the conditions in the first claim are sufficient. But by (6.1.10), these conditions imply
that the irreducible components of X are exactly its connected components, and since X is connected
and nonempty, it is irreducible. □
Corollary (6.1.12). — Let X be a locally Noetherian prescheme. For X to be integral, it is necessary and
sufficient that X be connected and that Ox be integral for all x ∈ X.
Proposition (6.1.13). — Let X be a locally Noetherian prescheme, and let x ∈ X be a point such that
the nilradical Nx of Ox is prime (resp. such that Ox is reduced, resp. integral); then there exists an open
neighbourhood U of x that is irreducible (resp. reduced, resp. integral).
P ROOF. It suffices to consider two cases: where Nx is prime, and where Nx = 0; the third
hypotheses is a combination of the first two. If Nx is prime, then x belongs to only one single
irreducible component Y of X (6.1.10); the union of the irreducible components of X that do not
contain x is closed (the set of these components being locally finite), and the complement U of this
union is thus open and contained in Y, and thus irreducible (0, 2.1.6) If Nx = 0, we also have Ny = 0
for any y in a neighbourhood of x, because N is quasi-coherent (5.1.1), and thus coherent, since X is
locally Noetherian, and the conclusion then follows from (0, 5.2.2). □
6.2. Artinian preschemes
Definition (6.2.1). — We say that a prescheme is Artinian if it is affine, and given by an Artinian
ring.
Proposition (6.2.2). — Given a prescheme X, the following conditions are equivalent: I | 144
(a) X is an Artinian scheme;
(b) X is Noetherian and its underlying space is discrete;
(c) X is Noetherian and the points of its underlying space are closed (the T1 condition).
When any of the above hold, the underlying space of X is finite, and the ring A of X is the direct sum of local
(Artinian) rings of points of X.
P ROOF. We know that (a) implies the last claim ([SZ60, p. 205, th. 3]), so every prime ideal of A
is thus maximal and is the inverse image of a maximal ideal of one of the local components of A,
and so the space X is finite and discrete; (a) thus implies (b), and (b) clearly implies (c). To see that
(c) implies (a), we first show that X is then finite; we can indeed restrict to the case where X is affine,
6. FINITENESS CONDITIONS 146
and we know that a Noetherian ring whose prime ideals are all maximal is Artinian ([SZ60, p. 203]),
whence our claim. The underlying space X is then discrete, the topological sum of a finite number
of points xi , and the local rings Oxi = Ai are Artinian; it is clear that X is isomorphic to the prime
spectrum affine scheme of the ring A (the direct sum of the Ai ) (1.7.3). □
6.3. Morphisms of finite type
Definition (6.3.1). — We say that a morphism f : X → Y is of finite type if Y is the union of a family
(Vα ) of affine open subsets having the following property:
(P) f −1 (Vα ) is a finite union of affine open subsets Uαi that are such that each ring A(Uαi ) is an
algebra of finite type over A(Vα ).
We then say that X is a prescheme of finite type over Y, or a Y-prescheme of finite type.
Proposition (6.3.2). — If f : X → Y is a morphism of finite type, then every affine open subset W of Y
satisfies property (P) of (6.3.1).
We first show
Lemma (6.3.2.1). — If T ⊂ Y is an affine open subset, satisfying property (P), then, for every g ∈ A( T ),
D ( g) also satisfies property (P).
P ROOF. By hypothesis, f −1 ( T ) is a finite union of affine open subsets Zj , that are such that
A( Zj ) is an algebra of finite type over A( T ); let ϕj : A( T ) → A( Zj ) be the homomorphism of rings
corresponding to the restriction of f to Zj (2.2.4), and set g j = ϕj ( g); we then have f −1 ( D ( g)) ∩ Zj =
D ( g j ) (1.2.2.2). But A( D ( g j )) = A( Zj ) g j = A( Zj )[1/g j ] is of finite type over A( Zj ), and a fortiori
over A( T ) by the hypothesis, and so also over A( D ( g)) = A( T )[1/g], which proves the lemma. □
P ROOF. With the above lemma, since W is quasi-compact (1.1.10), there exists a finite covering
of W by sets of the form D ( gi ) ⊂ W, where each gi belongs to a ring A(Vα(i) ). Each D ( gi ), being ErrII
quasi-compact, is a finite union of sets D (hik ), where hik ∈ A(W ); if ϕi : A(W ) → A( D ( gi )) is the
canonical map, then we have D (hik ) = D (ϕi (hik )) by (1.2.2.2). By (6.3.2.1), each of the f −1 ( D (hik ))
admits a finite covering by affine open subsets Uijk , that are such that the A(Uijk ) are algebras of
finite type over A( D (hik )) = A(W )[1/hik ], whence the proposition. □
We can thus say that the notion of a prescheme of finite type over Y is local on Y. I | 145
Proposition (6.3.3). — Let X and Y be affine schemes; for X to be of finite type over Y, it is necessary and
sufficient that A( X ) be an algebra of finite type over A(Y ).
P ROOF. Since the condition clearly suffices, we show that it is necessary. Set A = A(Y ) and
B = A( X ); by (6.3.2), there exists a finite affine open cover (Vi ) of X such that each of the rings
A(Vi ) is an A-algebra of finite type. Further, since the Vi are quasi-compact, we can cover each
of them with a finite number of open subsets of the form D ( gij ) ⊂ Vi , where gij ∈ B; if ϕi is a
homomorphism B → A(Vi ) that corresponds to the canonical injection Vi → X, then we have
Bgij = ( A(Vi ))ϕi ( gij ) = A(Vi )[1/ϕi ( gij )], so Bgij is an A-algebra of finite type. We can thus restrict
to the case where Vi = D ( gi ) with gi ∈ B. By hypothesis, there exists a finite subset Fi of B and an
n
integer ni ⩾ 0 such that Bgi is the algebra generated over A by the elements bi /gi i , where the bi run
over all of Fi . Since there are only finitely many of the gi , we can assume that all the ni are equal to
the same integer n. Further, since the D ( gi ) form a cover of X, the ideal generated in B by the gi is
equal to B, or, in other words, there exist hi ∈ B such that ∑i hi gi = 1. So let F be the finite subset of
B given by the union of the Fi , the set of the gi , and the set of the hi ; we will show that the subring
B′ = A[ F ] of B is equal to B. By hypothesis, for every b ∈ B and every i, the canonical image of b in
m
Bgi is of the form bi′ /gi i , where bi′ ∈ B′ ; by multiplying the bi′ by suitable powers of the gi , we can
again assume that all the mi are equal to the same integer m. By the definition of the ring of fractions,
there is thus an integer N (dependant on b) such that N ⩾ m and giN b ∈ B′ for all i; but, in the ring
B′ , the giN generate the ideal B′ , because the gi do (and the hi belong to B′ ); there are thus ci ∈ B′
such that ∑i ci giN = 1, whence b = ∑i ci giN b ∈ B′ , Q.E.D. □
Proposition (6.3.4). —
6. FINITENESS CONDITIONS 147
Corollary (6.3.8). — Let X be a prescheme of finite type over S. For every base extension S′ → S with S′
Noetherian (resp. locally Noetherian), X(S′ ) is Noetherian (resp. locally Noetherian).
P ROOF. This follows from (6.3.7), since X(S′ ) is of finite type over S′ by (6.3.4, iv). □
6. FINITENESS CONDITIONS 148
We can also says that, for a product X ×S Y of S-preschemes, if one of the factors X or Y is of I | 147
finite type over S and the other is Noetherian (resp. locally Noetherian), then X ×S Y is Noetherian (resp.
locally Noetherian).
Corollary (6.3.9). — Let X be a prescheme of finite type over a locally Noetherian prescheme S. Then every
S-morphism f : X → Y is of finite type.
P ROOF. In fact, we can assume that S is Noetherian; if ϕ : X → S and ψ : Y → S are the
structure morphisms, then we have ϕ = ψ ◦ f , and X is Noetherian by (6.3.7); f is thus of finite type
by (6.3.6). □
Xc
f Spec(Ω)
{
Y
(cf. (3.5.3)). Conversely, suppose that f is surjective, and let g : {ξ } = Spec(Ω) → Y be a morphism,
where Ω is an algebraically closed field. If we consider the diagram
Xo X( Ω )
f f (Ω)
Yo Spec(Ω),
then it suffices to show that there exists a rational point over Ω in X(Ω) ((3.3.14), (3.4.3), and (3.4.4)).
Since f is surjective, X(Ω) is nonempty (3.5.10), and since f is of finite type, so too is f (Ω) (6.4.3, iv);
thus X(Ω) contains a nonempty affine open subset Z such that A( Z ) is an non-null algebra of finite
type over Ω. By Hilbert’s Nullstellensatz [Zar47], there exists an Ω-homomorphism A( Z ) → Ω, and
thus a section of X(Ω) over Spec(Ω), which proves the proposition. □
Corollary (6.4.3). — Let K be an algebraically-closed field, and X an algebraic K-prescheme; the closed
points of X are then the rational points over K (3.4.4) and can be canonically identified with the points of X
with values in K.
Proposition (6.4.4). — Let X be an algebraic prescheme over a field K. The following properties are
equivalent.
(a) X is Artinian.
(b) The underlying space of X is discrete.
(c) The underlying space of X has only a finite number of closed points.
(c’) The underlying space of X is finite.
(d) The points of X are closed.
(e) X is isomorphic to Spec( A), where A is a K-algebra of finite rank.
P ROOF. Since X is Noetherian, it follows from (6.2.2) that the conditions (a), (b), and (d) are
equivalent, and imply (c) and (c′ ); it is also clear that (e) implies (a). It remains to see that (c) implies
(d) and (e); we can restrict to the case where X is affine. Then A( X ) is a K-algebra of finite type
(6.3.3), and thus a Jacobson ring ([CC, p. 3-11 and 3-12]), in which there are, by hypothesis, only a
finite number of maximal ideals. Since a finite intersection of prime ideals can only be a prime ideal
if it is equal to one of the prime ideals being intersected, every prime ideal of A( X ) is thus maximal,
whence (d). Further, we then know (6.2.2) that A( X ) is an Artinian K-algebra of finite type, and so
necessarily of finite rank [Zar47]. □
(6.4.5). When the conditions of (6.4.4) are satisfied, we say that X is a scheme finite over K (cf.
(II, 6.1.1)), or a finite K-scheme, of rank [ A : K ], which we also denote by rgK ( X ); if X and Y are finite
K-schemes, we have
(6.4.5.1) rgK ( X ⊔ Y ) = rgK ( X ) + rgK (Y ),
Remark (6.4.10). — We will see in chapter IV that the conclusion of (6.4.9) still holds without the
hypothesis on the transcendence degree of Ω over K.
Proposition (6.4.11). — If f : X → Y is a morphism of finite type, then, for every y ∈ Y, the fibre f −1 (y)
is an algebraic prescheme over the residue field k(y), and for every x ∈ f −1 (y), k( x ) is an extension of k( x )
of finite type.
P ROOF. Since f −1 (y) = X ⊗Y k(y) (6.3.6), the proposition follows from (6.3.4, iv) and (6.3.3). □
(6.4.13). Proposition (6.4.11) shows that the morphisms of finite type that correspond, intuitively, to
the “algebraic families of algebraic varieties”, with the points of Y playing the role of “parameters”, I | 150
which gives these morphisms a “geometric” meaning. The morphisms which are not of finite type
will show up in the following mostly in questions of “changing the base prescheme”, by localisation
or completion, for example.
6.5. Local determination of a morphism
Proposition (6.5.1). — Let X and Y be S-preschemes, with Y of finite type over S; let x ∈ X and y ∈ Y lie
over the same point s ∈ S.
(i) If two S-morphisms f = (ψ, θ ) and f ′ = (ψ′ , θ ′ ) from X to Y are such that ψ( x ) = ψ′ ( x ) = y,
♯
and the (local) Os -homomorphisms θ x♯ and θ ′ x from Oy to Ox are identical, then f and f ′ agree on
an open neighbourhood of x.
(ii) Suppose further that S is locally Noetherian. For every local Os -homomorphism ϕ : Oy → Ox , there
exists an open neighbourhood U of x in X, and an S-morphism f = (ψ, θ ) from U to Y such that
ψ( x ) = y and θ x♯ = ϕ.
P ROOF.
(i) Since the question is local on S, X, and Y, we can assume that S, X, and Y are affine,
given by rings A, B, and C (respectively), and with f and f ′ of the form ( a ϕ, ϕ e) and
( a ϕ′ , ϕe′ ) (respectively), where ϕ and ϕ′ are A-homomorphisms from C to B such that
ϕ−1 (jx ) = ϕ′−1 (jx ) = jy , and the homomorphisms ϕx and ϕx′ from Cy to Bx , induced by
ϕ and ϕ′ , are identical; we can further suppose that C is an A-algebra of finite type. Let ci
(1 ⩽ i ⩽ n) be the generators of the A-algebra C, and set bi = ϕ(ci ) and bi′ = ϕ′ (ci ); by
hypothesis, we have bi /1 = bi′ /1 in the ring of fractions Bx (1 ⩽ i ⩽ n). This implies that
there exist elements si ∈ B − jx such that si (bi − bi′ ) = 0 for 1 ⩽ i ⩽ n, and we can clearly
assume that all the si are equal to a single element g ∈ B − jx . From this, we conclude
that we have bi /1 = bi′ /1 for 1 ⩽ i ⩽ n in the ring of fractions Bg ; if i g is the canonical
homomorphism B → Bg , we then have i g ◦ ϕ = i g ◦ ϕ′ ; so the restrictions of f and f ′ to
D ( g) are identical.
(ii) We can restrict to the situation as in (i), and further assume that the ring A is Noetherian.
Let ci (1 ⩽ i ⩽ n) be the generators of the A-algebra C, and let α : A[ X1 , . . . , Xn ] → C be
6. FINITENESS CONDITIONS 151
(6.5.1.1) C
γ
/ Bhg
iy ix
/ Bx
ϕ
Cy
P ROOF.
(ii) By (6.5.1), there exists an open neighbourhood V of Y and a morphism g : V → X such that
g ◦ f (resp. f ◦ g) is defined and agrees with the identity on a neighbourhood of x (resp. y),
whence we can easily see that f is a local isomorphism.
(i) Since the question is local on X and Y, we can assume that X and Y are affine, given
by rings A and B (respectively); we have f = ( a ϕ, ϕ e), where ϕ is a homomorphism of
rings B → A that makes A a B-algebra of finite type; we have ϕ−1 (jx ) = jy , and the
homomorphism ϕx : By → A x induced by ϕ is surjective. Let (ti ) (1 ⩽ i ⩽ n) be a system
6. FINITENESS CONDITIONS 152
of generators of the B-algebra A; the hypothesis on ϕx implies that there exist bi ∈ B and
some c ∈ B − jx such that, in the ring of fractions A x , we have ti /1 = ϕ(bi )/ϕ(c) for
1 ⩽ i ⩽ n. Then (1.3.3) there exists some a ∈ A − jx such that, if we let g = aϕ(c), we also
have ti /1 = aϕ(bi )/g in the ring of fractions A g . With this, there exists, by hypothesis, a
polynomial Q( X1 , . . . , Xn ), with coefficients in the ring ϕ( B), such that a = Q(t1 , . . . , tn );
let Q( X1 /T, . . . , Xn /T ) = P( X1 , . . . , Xn , T )/T m , where P is homogeneous of degree m. In I | 152
the ring A g , we have
a/1 = am P(ϕ(b1 ), . . . , ϕ(bn ), ϕ(c))/gm = am ϕ(d)/gm
where d ∈ B. Since, in A g , g/1 = ( a/1)(ϕ(c)/1) is invertible by definition, so too are a/1
and ϕ(c)/1, and we can thus write a/1 = (ϕ(d)/1)(ϕ(c)/1)−m . From this we conclude that
ϕ(d)/1 is also invertible in A g . So let h = cd; since ϕ(h)/1 is invertible in A g , the composite
ϕ γ
homomorphism B − → A → A g factors as B → Bh − → A g (0, 1.2.4). We will show that γ
is surjective; it suffices to show that the image of Bh in A g contains the ti /1 and ( g/1)−1 .
But we have ( g/1)−1 = (ϕ(c)/1)m−1 (ϕ(d)/1)−1 = γ(cm /h), and a/1 = γ(dm+1 /hm ), so
( aϕ(bi ))/1 = γ(bi dm+1 /hm ), and since ti /1 = ( aϕ(bi )/1)( g/1)−1 , our claim is proved.
The choice of h implies that ψ( D ( g)) ⊂ D (h), and we also know that the restriction of f to
D ( g) is equal to ( a γ, γ
e); since γ is surjective, this restriction is a closed immersion of D ( g)
into D (h) (4.2.3).
□
Corollary (6.5.5). — Let f = (ψ, θ ) : X → Y be a morphism of finite type. Assume that X is irreducible,
and denote by x its generic point, and let y = ψ( x ).
(i) For f to be a local immersion at any point of X, it is necessary and sufficient that θ x♯ : Oy → Ox be
surjective.
(ii) Assume further that Y is irreducible and locally Noetherian. For f to be a local isomorphism at any
point of X, it is necessary and sufficient that y be the generic point of Y (or, equivalently (0, 2.1.4),
that f be a dominant morphism) and that θ x♯ be an isomorphism (in other words, that f be
birational (2.2.9)).
P ROOF. It is clear that (i) follows from (6.5.4, i), taking into account the fact that every nonempty
open subset of X contains x; similarly, (ii) follows from (6.5.4, ii). □
6.6. Quasi-compact morphisms and morphisms locally of finite type
Definition (6.6.1). — We say that a morphism f : X → Y is quasi-compact if the inverse image of any
quasi-compact open subset of Y under f is quasi-compact.
Let B be a base for the topology of Y consisting of quasi-compact open subsets (for example,
affine open subsets); for f to be quasi-compact, it is necessary and sufficient that the inverse image
of every set of B under f be quasi-compact (or, equivalently, a finite union of affine open subsets),
because every quasi-compact open subset of Y is a finite union of sets of B. For example, if X is
quasi-compact and Y affine, then every morphism f : X → Y is quasi-compact: indeed, X is a finite
union of affine open subsets Ui , and for every affine open subset V of Y, Ui ∩ f −1 (V ) is affine (5.5.10),
and so quasi-compact.
If f : X → Y is a quasi-compact morphism, it is clear that, for every open subset V of Y, the
restriction of f to f −1 (V ) is a quasi-compact morphism f −1 (V ) → V. Conversely, if (Uα ) is an open
cover of Y, and f : X → Y a morphism such that the restrictions f −1 (Uα ) → Uα are quasi-compact,
then f is quasi-compact.
Definition (6.6.2). — We say that a morphism f : X → Y is locally of finite type if, for every x ∈ X,
there exists an open neighbourhood U of x and an open neighbourhood V ⊃ f (U ) of y such that the
restriction of f to U is a morphism of finite type from U to V. We then also say that X is a prescheme I | 153
locally of finite type over Y, or a Y-prescheme locally of finite type.
It follows immediately from (6.3.2) that, if f is locally of finite type, then, for every open subset
W of Y, the restriction of f to f −1 (W ) is a morphism f −1 (W ) → W that is locally of finite type.
6. FINITENESS CONDITIONS 153
If Y is locally Noetherian and X locally of finite type over Y, then X is locally Noetherian thanks
to (6.3.7).
Proposition (6.6.3). — For a morphism f : X → Y to be of finite type, it is necessary and sufficient that it
be quasi-compact and locally of finite type.
P ROOF. The necessity of the conditions is immediate, given (6.3.1) and the remark following
(6.6.1). Conversely, suppose that the conditions are satisfied, and let U be an affine open subset of Y,
given by some ring A; for all x ∈ f −1 (U ), there is, by hypothesis, a neighbourhood V ( x ) ⊂ f −1 (U )
of x, and a neighbourhood W ( x ) ⊂ U of y = f ( x ) containing f (V ( x )), and such that the restriction
of f to V ( x ) is a morphism V ( x ) → W ( x ) of finite type. Replacing W ( x ) with a neighbourhood
W1 ( x ) ⊂ W ( x ) of x of the form D ( g) (with g ∈ A), and V ( x ) with V ( x ) ∩ f −1 (W1 ( x )), we can
assume that W ( x ) is of the form D ( g), and thus of finite type over U (because its ring can be written
as A[1/g]); so V ( x ) is of finite type over U. Further, f −1 (U ) is quasi-compact by hypothesis, and so
the finite union of open subsets V ( xi ), which finishes the proof. □
Proposition (6.6.4). —
(i) An immersion X → Y is quasi-compact if it is closed, or if the underlying space of Y is locally
Noetherian, or if the underlying space of X is Noetherian.
(ii) The composition of any two quasi-compact morphisms is quasi-compact.
(iii) If f : X → Y is a quasi-compact S-morphism, then so too is f (S′ ) : X(S′ ) → Y(S′ ) for any extension
g : S → S′ of the base prescheme.
(iv) If f : X → X ′ and g : Y → Y ′ are two quasi-compact S-morphisms, then f ×S g is quasi-compact.
(v) If the composition of any two morphisms f : X → Y and g : Y → Z is quasi-compact, and if either
g is separated or the underlying space of X is locally Noetherian, then f is quasi-compact.
(vi) For a morphism f to be quasi-compact, it is necessary and sufficient that f red be quasi-compact.
P ROOF. We note that (vi) is evident because the property of being quasi-compact, for a mor-
phism, depends only on the corresponding continuous map of underlying spaces. We will similarly
prove the part of (v) corresponding to the case where the underlying space of X is locally Noetherian.
Set h = g ◦ f , and let U be a quasi-compact open subset of Y; g(U ) is quasi-compact (but not
necessarily open) in Z, and so contained in a finite union of quasi-compact open subsets Vj (2.1.3),
and f −1 (U ) is thus contained in the union of the h−1 (Vj ), which are quasi-compact subspaces of X,
and thus Noetherian subspaces. We thus conclude (0, 2.2.3) that f −1 (U ) is a Noetherian space, and a
fortiori quasi-compact.
To prove the other claims, it suffices to prove (i), (ii), and (iii) (5.5.12). But (ii) is evident, and (i)
follows from (6.3.5) whenever the space Y is locally Noetherian or the space X is Noetherian, and
is evident for a closed immersion. To show (iii), we can restrict to the case where S = Y (3.3.11); I | 154
let f ′ = f (S′ ) , and let U ′ be a quasi-compact open subset of S′ . For every s′ ∈ U ′ , let T be an affine
open neighbourhood of g(s′ ) in S, and let W be an affine open neighbourhood of s′ contained in
U ′ ∩ g−1 ( T ); it will suffice to show that f ′−1 (W ) is quasi-compact; in other words, we can restrict to
showing that, when S and S′ are affine, the underlying space of X ×S S′ is quasi-compact. But since
X is then, by hypothesis, a finite union of affine open subsets Vj , X ×S S′ is a union of the underlying
spaces of the affine schemes Vj ×S S′ ((3.2.2) and (3.2.7)), which proves the proposition. □
We note also that, if X = X ′ ⊔ X ′′ is the sum of two preschemes, a morphism f : X → Y is
quasi-compact if and only if its restrictions to both X ′ and X ′′ are quasi-compact.
Proposition (6.6.5). — Let f : X → Y be a quasi-compact morphism. For f to be dominant, it is necessary
and sufficient that, for every generic point y of an irreducible component of Y, f −1 (y) contain the generic
point of an irreducible component of X.
P ROOF. It is immediate that the condition is sufficient (even without assuming that f is quasi-
compact). To see that it is necessary, consider an affine open neighbourhood U of y; f −1 (U ) is
quasi-compact, and so a finite union of affine open subsets Vi , and the hypothesis that f be dominant
implies that y belongs to the closure in U of one of the f (Vi ). We can clearly assume X and Y to be
reduced; since the closure in X of an irreducible component of Vi is an irreducible component on X
7. RATIONAL MAPS 154
(0, 2.1.6), we can replace X by Vi , and Y by the closed reduced subprescheme of U that has f (Vi ) ∩ U
as its underlying space (5.2.1), and we are thus led to proving the proposition when X = Spec( A)
and Y = Spec( B) are affine and reduced. Since f is dominant, B is a subring of A (1.2.7), and the
proposition then follows from the fact that every minimal prime ideal of B is the intersection of B
with a minimal prime ideal of A (0, 1.5.8). □
Proposition (6.6.6). —
(i) Every local immersion is locally of finite type.
(ii) If two morphisms f : X → Y and g : Y → Z are locally of finite type, then so too is g ◦ f .
(iii) If f : X → Y is an S-morphism locally of finite type, then f (S′ ) : X(S′ ) → Y(S′ ) is locally of finite
type for any extension S′ → S of the base prescheme.
(iv) If f : X → X ′ and g : Y → Y ′ are S-morphisms locally of finite type, then f ×S g is locally of finite
type.
(v) If the composition g ◦ f of two morphisms is locally of finite type, then f is locally of finite type.
(vi) If a morphism f is locally of finite type, then so too is f red .
P ROOF. By (5.5.12), it suffices to prove (i), (ii), and (iii). If j : X → Y is a local immersion then,
for every x ∈ X, there is an open neighbourhood V of j( x ) in Y and an open neighbourhood U of x
in X such that the restriction of j to U is a closed immersion U → V (4.5.1), and so this restriction is
of finite type. To prove (ii), consider a point x ∈ X; by hypothesis, there is an open neighbourhood
W of g( f ( x )) and an open neighbourhood V of f ( x ) such that g(V ) ⊂ W and such that V is of
of finite type over W; furthermore, f −1 (V ) is locally of finite type over V (6.6.2), so there is an
open neighbourhood U of x that is contained in f −1 (V ) and of is finite type over V; thus we have I | 155
g( f (U )) ⊂ W, and that U is of finite type over W (6.3.4, ii). Finally, to prove (iii), we can restrict to
the case where Y = S (3.3.11); for every x ′ ∈ X ′ = X(S′ ) , let x be the image of x ′ in X, s the image of
x in S, T an open neighbourhood of s, T ′ the inverse image of T in S′ , and U an open neighbourhood
of x that is of finite type over T and whose image is contained in T; then U ×S T ′ = U × T T ′ is an
open neighbourhood of x ′ (3.2.7) that is of finite type over T ′ (6.3.4, iv). □
Corollary (6.6.7). — Let X and Y be S-preschemes that are locally of finite type over S. If S is locally
Noetherian, then X ×S Y is locally Noetherian.
P ROOF. Indeed, X being locally of finite type over S means that it is locally Noetherian, and
that X ×S Y is locally of finite type over X, and so X ×S Y is also locally Noetherian. □
Remark (6.6.8). — Proposition (6.3.10) and its proof extend immediately to the case where we
suppose only that the morphism f is locally of finite type. Similarly, propositions (6.4.2) and (6.4.9)
hold true when we suppose only that the preschemes X and Y in the claim are locally of finite type
over the field K.
Let f be a rational map from X to Y, and U an open subset of X; if f 1 and f 2 are two morphisms
belonging to the class of f , defined (respectively) on dense open subsets V and W of X, then the
restrictions f 1 |(U ∩ V ) and f 2 |(U ∩ W ) agree on U ∩ V ∩ W, which is dense in U; the class f of
morphisms thus defines a rational map from U to Y, called the restriction of f to U, and denoted by
f |U.
If, to every S-morphism f : X → Y, we take the corresponding rational S-map to which f
belongs, we obtain a canonical map from HomS ( X, Y ) to the set of rational S-maps from X to Y.
We denote by Γrat ( X/Y ) the set of rational Y-sections on X, and we thus have a canonical map
Γ( X/Y ) → Γrat ( X/Y ). It is also clear that, if X and Y are S-preschemes, then the set of rational
S-maps from X to Y is canonically identified with Γrat (( X ×S Y )/X ) (3.3.14).
(7.1.3). It also follows from (7.1.2) and (3.3.14) that the rational functions on X are canonically
identified with equivalence classes of sections of the structure sheaf OX over dense open subsets of X,
where two such sections are equivalent if the agree on some dense open subset of X contained inside
the intersection of the subsets on which they are defined. In particular, it follows that the rational
functions on X form a ring R( X ).
(7.1.4). When X is an irreducible prescheme, every nonempty open subset of X is dense in X; so we
can say that the nonempty open subsets of X are the open neighbourhoods of the generic point x of X. To
say that two morphisms from nonempty open subsets of X to Y are equivalent thus means, in this
case, that they have the same germ at the point x. In other words, the rational maps (resp. rational
S-maps) X → Y are identified with the germs of morphisms (resp. S-morphisms) from nonempty open
subsets of X to Y at the generic point x of X. In particular:
Proposition (7.1.5). — If X is an irreducible prescheme, then the ring R( X ) of rational maps on X is
canonically identified with the local ring Ox of the generic point x of X. It is a local ring of dimension 0, and
thus a local Artinian ring when X is Noetherian; it is a field when X is integral, and, when X is further an
affine scheme, it is identified with the field of fractions of A( X ).
P ROOF. Given the above, and the identification of rational functions with sections of OX over a
dense open subset of X, the first claim is nothing but the definition of the fibre of a sheaf above a
point. For the other claims, we can reduce to the case where X is affine, given by some ring A; then
jx is the nilradical of A, and Ox is thus of dimension 0; if A is integral, then jx = (0), and Ox is thus
the field of fractions of A. Finally, if A is Noetherian, we know ([Sam53b, p. 127, cor. 4]) that jx is
nilpotent, and Ox = A x Artinian. □
If X is integral, then the ring Oz is integral for all z ∈ X; every affine open subset U containing
z also contains x, and R(U ), being equal to the field of fractions of A(U ), is identified with R( X );
we thus conclude that R( X ) can also be identified with the field of fractions of Oz : the canonical
identification of Oz to a subring of R( X ) consists of associating, to every germ of a section s ∈ Oz ,
the unique rational function on X, class of a section of OX , (necessarily defined on a dense open
subset of X) having s as its germ at the point z.
(7.1.6). Now suppose that X has a finite number of irreducible components Xi (1 ⩽ i ⩽ n) (which
will be the case whenever the underlying space of X is Noetherian); let Xi′ be the open subset of X
given by the complement of the X j ∩ Xi for j ̸= i inside Xi ; Xi′ is irreducible, its generic point xi is the
generic point of Xi , and the Xi′ are pairwise disjoint, with their union being dense in X (0, 2.1.6). For
every dense open subset U of X, Ui = U ∩ Xi′ is a nonempty dense open subset of Xi′ , with the Ui
being pairwise disjoint, and so U ′ = i Ui′ is dense in X. Giving a morphism from U ′ to Y consists
S
Corollary (7.1.9). — Let A be a Noetherian ring, and X = Spec( A). If Q is the complement of the union of
the minimal prime ideals of A, then the ring of rational functions on X can be canonically identified with the
ring of fractions Q−1 A.
This will follow from the following lemma:
Lemma (7.1.9.1). — For an element f ∈ A to be such that D ( f ) is dense in X, it is necessary and sufficient
that f ∈ Q; every dense open subset of X contains an open subset of the form D ( f ), where f ∈ Q.
P ROOF. To show (7.1.9.1), we again denote by Xi (1 ⩽ i ⩽ n) the irreducible components of X;
if D ( f ) is dense in X then D ( f ) ∩ Xi ̸= ∅ for 1 ⩽ i ⩽ n, and vice-versa; but this means that f ̸∈ pi
for 1 ⩽ i ⩽ n, where we set pi = j( Xi ), and since the pi are the minimal prime ideals of A (1.1.14),
the conditions f ̸∈ pi (1 ⩽ i ⩽ n) are equivalent to f ∈ Q, whence the first claim of the lemma. For
the other claim, if U is a dense open subset of X, the complement of U is a set of the form V (a),
where a is an ideal which is not contained in any of the pi ; it is thus not contained in their union
([Nor53, p. 13]), and there thus exists some f ∈ a belonging to Q; whence D ( f ) ⊂ U, which finishes
the proof. □
(7.1.10). Suppose again that X is irreducible, with generic point x. Since every nonempty open subset
U of X contains x, and thus also contains every z ∈ X such that x ∈ {z}, every morphism U → Y
can be composed with the canonical morphism Spec(Ox ) → X (2.4.1); and any two morphisms into
Y from two nonempty open subsets of X which agree on a nonempty open subset of X give, by
composition, the same morphism Spec(Ox ) → Y. In other words, to every rational map from X to Y
there is a corresponding well-defined morphism Spec(Ox ) → Y.
Proposition (7.1.11). — Let X and Y be two S-preschemes; suppose that X is irreducible with generic point
x, and that Y is of finite type over S. Any two rational S-maps from X to Y that correspond to the same
S-morphism Spec(Ox ) → Y are then identical. If we further suppose S to be locally Noetherian, then every
S-morphism from Spec(Ox ) to Y corresponds to exactly one rational S-map from X to Y.
P ROOF. Taking into account that every nonempty subset of X is dense in X, this follows from
(6.5.1). □
Corollary (7.1.12). — Suppose that S is locally Noetherian, and that the other hypotheses of (7.1.11) are
satisfied. The rational S-maps from X to Y can then be identified with points of the S-prescheme Y, with
values in the S-prescheme Spec(Ox ).
P ROOF. This is nothing but (7.1.11), with the terminology introduced in (3.4.1). □
Corollary (7.1.13). — Suppose that the conditions of (7.1.12) are satisfied. Let s be the image of x in S.
The data of a rational S-map from X to Y is equivalent to the data of a point y of Y over s along with a local
Os -homomorphism Oy → Ox = R( X ).
P ROOF. This follows from (7.1.11) and (2.4.4). □
In particular:
Corollary (7.1.14). — Under the conditions of (7.1.12), rational S-maps from X to Y depend only (for
any given Y) on the S-prescheme Spec(Ox ), and, in particular, remain the same whenever X is replaced by
Spec(Oz ), for any z ∈ X.
P ROOF. Since z ∈ { x }, x is the generic point of Z = Spec(Oz ), and OX,x = OZ,z . □
When X is integral, R( X ) = Ox = k( x ) is a field (7.1.5); the preceding corollaries then specialize
to the following:
Corollary (7.1.15). — Suppose that the conditions of (7.1.12) are satisfied, and further that X is integral.
Let s be the image of x in S. Then rational S-maps from X to Y can be identified with the geometric points of
Y ⊗S k(s) with values in the extension R( X ) of k (s), or, in other words, every such map is equivalent to the
data of a point y ∈ Y above s along with a k(s)-monomorphism from k (y) to k ( x ) = R( X ).
P ROOF. The points of Y above s are identified with the points of Y ⊗S k(s) (3.6.3), and the local
Os -homomorphisms Oy → R( X ) with the k (s)-monomorphisms k(y) → R( X ). □
7. RATIONAL MAPS 157
More precisely:
Corollary (7.1.16). — Let k be a field, and X and Y two algebraic preschemes over k (6.4.1); suppose further
that X is integral. Then the rational k-maps from X to Y can be identified with the geometric points of Y with
values in the extension R( X ) of k (3.4.4).
7.2. Domain of definition of a rational map
(7.2.1). Let X and Y be preschemes, and f rational map from X to Y. We say that f is defined at a point
x ∈ X if there exists a dense open subset U of X that contains x, and a morphism U → Y belonging
to the equivalence class of f . The set of points x ∈ X where f is defined is called the domain of
definition of f ; it is clear that it is an open dense subset of X.
Proposition (7.2.2). — Let X and Y be S-preschemes such that X is reduced and Y is separated over S. Let I | 159
f be a rational S-map from X to Y, with domain of definition U0 . Then there exists exactly one S-morphism
U0 → Y belonging to the class of f .
Since, for every morphism U → Y belonging to the class of f , we necessarily have U ⊂ U0 , it is
clear that the proposition will be a consequence of the following:
Lemma (7.2.2.1). — Under the hypotheses of (7.2.2), let U1 and U2 be two dense open subsets of X, and
f i : Ui → Y (i = 1, 2) two S-morphisms such that there exists an open subset V ⊂ U1 ∩ U2 , dense in X, and
on which f 1 and f 2 agree. Then f 1 and f 2 agree on U1 ∩ U2 .
P ROOF. We can clearly restrict to the case where X = U1 = U2 . Since X (and thus V) is reduced,
X is the smallest closed subprescheme of X containing V (5.2.2). Let g = ( f 1 , f 2 )S : X → Y ×S Y;
since, by hypothesis, the diagonal T = ∆Y (Y ) is a closed subprescheme of Y ×S Y, Z = g−1 ( T ) is
a closed subprescheme of X (4.4.1). If h : V → Y is the common restriction of f 1 and f 2 to V, then
the restriction of g to V is g′ = (h, h)S , which factors as g′ = ∆Y ◦ h; since ∆Y−1 ( T ) = Y, we have that
g′−1 ( T ) = V, and so Z is a closed subprescheme of X inducing V, thus containing V, which implies
that Z = X. From the equation g−1 ( T ) = X, we deduce (4.4.1) that g factors as ∆Y ◦ f , where f is a
morphism X → Y, which implies, by the definition of the diagonal morphism, that f 1 = f 2 = f . □
It is clear that the morphism U0 → Y defined in (7.2.2) is the unique morphism of the class f
that cannot be extended to a morphism from an open subset of X that strictly contains U0 . Under the
hypotheses of (7.2.2), we can thus identify the rational maps from X to Y with the non-extendible (to
strictly larger open subsets) morphisms from dense open subsets of X to Y. With this identification,
Proposition (7.2.2) implies:
Corollary (7.2.3). — With the hypotheses from (7.2.2) on X and Y, let U be a dense open subset of X. Then
there exists a canonical bijective correspondence between S-morphisms from U to Y and rational S-maps from
X to Y that are defined at all points of U.
P ROOF. By (7.2.2), for every S-morphism f from U to Y, there exists exactly one rational S-map
f from X to Y which extends f . □
Corollary (7.2.5). — Let X and Y be two S-preschemes; suppose that X is reduced, and that X and Y are
separated over S. Let p : Y → X be an S-morphism (making Y an X-prescheme), U a dense open subset of X,
and f a U-section of Y; then the rational map f from X to Y extending f is a rational X-section of Y.
P ROOF. We have to show that p ◦ f is the identity on the domain of definition of f ; since X is I | 160
separated over S, this again follows from (7.2.2.1). □
7. RATIONAL MAPS 158
Corollary (7.2.6). — Let X be a reduced prescheme, and U a dense open subset of X. Then there is a
canonical bijective correspondence between sections of OX over U and rational functions on X defined at every
point of U.
P ROOF. Taking (7.2.3), (7.1.2), and (7.1.3) into account, it suffices to note that the X-prescheme
X ⊗Z Z[ T ] is separated over X (5.5.1, iv). □
Corollary (7.2.7). — Let Y be a reduced prescheme, f : X → Y a separated morphism, U a dense open
subset of Y, g : U → f −1 (U ) a U-section of f −1 (U ), and Z the reduced subprescheme of X that has g(U )
as its underlying space (5.2.1). For g to be the restriction of a Y-section of X (in other words (7.2.5), for the
rational map from Y to X extending g to be defined everywhere), it is necessary and sufficient for the
restriction of f to Z to be an isomorphism from Z to Y.
P ROOF. The restriction of f to f −1 (U ) is a separated morphism (5.5.1, i), so g is a closed
immersion (5.4.6), and so g(U ) = Z ∩ f −1 (U ), and the subprescheme induced by Z on the open
subset g(U ) of Z is identical to the closed subprescheme of f −1 (U ) associated to g (5.2.1). It is then
clear that the stated condition is sufficient, because, if satisfied, and if f Z : Z → Y is the restriction of
f to Z, and g : Y → Z is the inverse isomorphism, then g extends g. Conversely, if g is the restriction
to U of a Y-section h of X, then h is a closed immersion (5.4.6), and so h(Y ) is closed, and, since it is
contained in Z, is equal to Z, and it follows from (5.2.1) that h is necessarily an isomorphism from Y
to the closed subprescheme Z of X. □
(7.2.8). Let X and Y be two S-preschemes, with X reduced, and Y separated over S. Let f be a rational
S-map from X to Y, and let x be a point of X; we can compose f with the canonical S-morphism
Spec(Ox ) → X (2.4.1) provided that the intersection of Spec(Ox ) with the domain of definition of f
is dense in Spec(Ox ) (identified with the set of z ∈ X such that x ∈ {z} (2.4.2)). This will happen in
the follow cases:
1st. X is irreducible (and thus integral), because then the generic point ξ of X is the generic point
of Spec(Ox ); since the domain of definition U of f contains ξ, U ∩ Spec(Ox ) contains ξ, and
so is dense in Spec(Ox ).
2nd. X is locally Noetherian; our claim then follows from:
Lemma (7.2.8.1). — Let X be a prescheme whose underlying space is locally Noetherian, and x a point of X.
The irreducible components of Spec(Ox ) are the intersections of Spec(Ox ) with the irreducible components of
X containing x. For an open subset U ⊂ X to be such that U ∩ Spec(Ox ) is dense Spec(Ox ), it is necessary
and sufficient for it to have a nonempty intersection with the irreducible components of X that contain x
(which will be the case whenever U is dense in X).
P ROOF. It suffices to show just the first claim, since the second then follows. Since Spec(Ox ) is
contained in every affine open subset U that contains x, and since the irreducible components of U
that contain x are the intersections of U with the irreducible components of X containing x (0, 2.1.6),
we can suppose that X is affine, given by some ring A. Since the prime ideals of A x correspond
bijectively to the prime ideals of A that are contained in jx (2.1.6), the minimal prime ideals of A x I | 161
correspond to the minimal prime ideals of A that are contained in jx , hence the lemma. □
With this in mind, suppose that we are in one of the two cases mentioned in (7.2.8). If U is the
domain of definition of the rational S-map f , then we denote by f ′ the rational map from Spec(Ox )
to Y which agrees (taking (2.4.2) into account) with f on U ∩ Spec(Ox ); we say that this rational map
is induced by f .
Proposition (7.2.9). — Let S be a locally Noetherian prescheme, X a reduced S-prescheme, and Y an
S-scheme of finite type. Suppose further that X is either irreducible or locally Noetherian. Then let f be a
rational S-map from X to Y, and x a point of X. For f to be defined at a point x, it is necessary and sufficient
for the rational map f ′ from Spec(Ox ) to Y, induced by f (7.2.8), to be a morphism.
P ROOF. The condition clearly being necessary (since Spec(Ox ) is contained in every open subset
containing x), we show that it is sufficient. By (6.5.1), there exists an open neighbourhood U of x
in X, and an S-morphism g from U to Y that induces f ′ on Spec(Ox ). If X is irreducible, then U is
dense in X, and, by (7.2.3), we can suppose that g is a rational S-map. Further, the generic point of
7. RATIONAL MAPS 159
X belongs to both Spec(Ox ) and the domain of definition of f , and so s and g agree at this point,
and thus on a nonempty open subset of X (6.5.1). But since f and g are rational S-maps, they are
identical (7.2.3), and so f is defined at x.
If we now suppose that X is locally Noetherian, then we can suppose that U is Noetherian; then
there are only a finite number of irreducible components Xi of X that contain x (7.2.8.1), and we
can suppose that they are the only ones that have a nonempty intersection with U, by replacing, if
needed, U with a smaller open subset (since there are only a finite number of irreducible components
of X that have a nonempty intersection with U, because U is Noetherian). We then have, as above,
that f and g agree on a nonempty open subset of each of the Xi . Taking into account the fact that
each of the Xi is contained in U, we consider the morphism f 1 , defined on a dense open subset of
U ∪ ( X − U ), equal to g on U, and to f on the intersection of X − U with the domain of definition
of f . Since U ∪ ( X − U ) is dense in X, f 1 and f agree on a dense open subset of X, and since f is a
rational map, f is an extension of f 1 (7.2.3), and is thus defined at the point x. □
7.3. Sheaf of rational functions
(7.3.1). Let X be a prescheme. For every open subset U ⊂ X, we denote by R(U ) the ring of rational
functions on U (7.1.3); this is a Γ(U, OX )-algebra. Further, if V ⊂ U is a second open subset of X,
then every section of OX over a dense (in X) open subset of V gives, by restriction to V, a section
over a dense (in X) open subset of V, and if two sections agree on a dense (in X) open subset of
U, then their restrictions to V agree on a dense (in X) open subset of V. We can thus define a
di-homomorphism of algebras ρV,U : R(U ) → R(V ), and it is clear that, if U ⊃ V ⊃ W are open
subsets of X, then we have ρW,U = ρW,V ◦ ρV,U ; the R(U ) thus define a presheaf of algebras on X.
Definition (7.3.2). — We define the sheaf of rational functions on a prescheme X, denoted by R ( X ), I | 162
to be the OX -algebra associated to the presheaf defined by the R(U ).
For every prescheme X and open subset U ⊂ X, it is clear that the induced sheaf R ( X )|U is
exactly R (U ).
Proposition (7.3.3). — Let X be a prescheme such that the family ( Xλ ) of its irreducible components is
locally finite (which is the case whenever the underlying space of X is locally Noetherian). Then the OX -module
R ( X ) is quasi-coherent, and for every open subset U of X that has a nonempty intersection with only finitely
many of the components Xλ , R(U ) is equal to Γ(U, R ( X )), and can be canonically identified with the direct
sum of the local rings of the generic points of the Xλ such that U ∩ Xλ ̸= ∅.
P ROOF. We can evidently restrict to the case where X has only a finite number of irreducible
components Xi , with generic points xi (1 ⩽ i ⩽ n). The fact that R(U ) can be canonically identified
with the direct sum of the Oxi = R( Xi ) such that U ∩ Xi ̸= ∅ then follows from (7.1.7). We will show
that the presheaf U → R(U ) satisfies the sheaf axioms, which will prove that R(U ) = Γ(U, R ( X )).
Indeed, it satisfies (F1) by what has already been discussed. To see that it satisfies (F2), consider
a cover of an open subset U of X by open subsets Vα ⊂ U; if the sα ∈ R(Vα ) are such that the
restrictions of sα and s β to Vα ∩ Vβ agree for every pair of indices, then we can conclude that, for
every index i such that U ∩ Xi ̸= ∅, the components in R( Xi ) of all the sα such that Vα ∩ Xi ̸= ∅ are
all the same; denoting this component by ti , it is clear that the element of R(U ) that has the ti as its
components has sα as its restriction to each Vα . Finally, to see that R ( X ) is quasi-coherent, we can
restrict to the case where X = Spec( A) is affine; by taking U to be an affine open subset of the form
D ( f ), where f ∈ A, it follows from the above and from Definition (1.3.4) that we have R ( X ) = M, e
where M is the direct sum of the A-modules A xi . □
Corollary (7.3.4). — Let X be a reduced prescheme that has only a finite number of irreducible components,
and let Xi (1 ⩽ i ⩽ n) be the closed reduced preschemes of X that have the irreducible components of X as
their underlying spaces (5.2.1). If hi is the canonical injection Xi → X, then R ( X ) is the direct sum of the
OX -algebras (hi )∗ (R ( Xi )).
Corollary (7.3.5). — If X is irreducible, then every quasi-coherent R ( X )-module F is a simple sheaf.
P ROOF. It suffices to show that every x ∈ X admits a neighbourhood U such that F |U is a
simple sheaf (0, 3.6.2); in other words, we are led to considering the case where X is affine; we can
7. RATIONAL MAPS 160
Proposition (7.3.7). — Suppose that the prescheme X is locally integral or locally Noetherian. Then R ( X ) I | 163
is a quasi-coherent OX -algebra; if, further, X is reduced (which will be the case whenever X is locally integral),
then the canonical homomorphism OX → R ( X ) is injective.
P ROOF. Since the questions is local, the first claim follows from (7.3.3); the second follows from
(7.2.3). □
(7.3.8). 10 Let X and Y be two integral preschemes, which implies that R ( X ) (resp. R (Y )) is a quasi-
coherent OX -module (resp. OY -module) (7.3.3). Let f : X → Y be a dominant morphism; then there
exists a canonical homomorphism of OX -modules
(7.3.8.1) τ : f ∗ (R (Y )) −→ R ( X ).
P ROOF. Suppose first that X = Spec( A) and Y = Spec( B) are affine, given by integral rings
A and B, with f thus corresponding to an injective homomorphism B → A which extends to a
monomorphism L → K from the field of fractions L of B to the field of fractions K of A. The
homomorphism (7.3.8.1) then corresponds to the canonical homomorphism L ⊗ B A → K (1.6.5). In
the general case, for each pair of nonempty affine open sets U ⊂ X and V ⊂ Y such that f (U ) ⊂ V,
we define, as above, a homomorphism τU,V and we immediately have that, if U ′ ⊂ U, V ′ ⊂ V,
f (U ′ ) ⊂ V ′ , then τU,V extends τU ′ ,V ′ , and hence our assertion. If x and y are the generic points of X
and Y respectively, then we have f ( x ) = y,
( f ∗ (R (Y ))) x = Oy ⊗Oy Ox = Ox
(0, 4.3.1) and τx is thus an isomorphism. □
10[Trans.] This paragraph was changed entirely in the Errata of EGA II.
8. CHEVALLEY SCHEMES 161
Proposition (7.4.5). — Let X and Y be integral preschemes, and f : X → Y a dominant morphism. For
every torsion-free quasi-coherent OX -module F , f ∗ (F ) is a torsion-free OY -module.
P ROOF. Since f ∗ is left exact (0, 4.2.1), it suffices, by (7.4.2), to prove the proposition in the case I | 164
where F = (R ( X ))( I ) . But every nonempty open subset U of Y contains the generic point of Y, so
f −1 (U ) contains the generic point of X (0, 2.1.5), so we have that Γ(U, f ∗ (F )) = Γ( f −1 (U ), F ) =
( R( X ))( I ) ; in other words, f ∗ (F ) is the simple sheaf with fibre ( R( X ))( I ) , considered as a R (Y )-
module, and it is clearly torsion free. □
Proposition (7.4.6). — Let X be an integral prescheme, and x its generic point. For every quasi-coherent
OX -module F of finite type, the following conditions are equivalent: (a) F is a torsion sheaf; (b) Fx = 0; (c)
Supp(F ) ̸= X.
P ROOF. By (7.3.5) and (7.4.1), the equations Fx = 0 and F ⊗OX R ( X ) = 0 are equivalent, so
(a) and (b) are equivalent; then Supp(F ) is closed in X (0, 5.2.2), and since every nonempty open
subset of X contains x, (b) and (c) are equivalent. □
(7.4.7). We generalise (by an abuse of language) the definitions of (7.4.1) to the case where X is a
reduced prescheme having only a finite number of irreducible components; it then follows from (7.3.4)
that the equivalence between a) and c) in (7.4.6) still holds true for such a prescheme.
(8.1.2). Now consider a field R. For all subrings A of R, we denote by L( A) the set of local rings Ap ,
where p ranges over the prime spectrum of A; such local rings are identified with the subrings of R
containing A. Since p = (pAp ) ∩ A, the map p 7→ Ap from Spec( A) to L( A) is bijective.
Lemma (8.1.3). — Let R be a field, and A a subring of R. For a local subring M of R to dominate a ring
Ap ∈ L( A), it is necessary and sufficient that A ⊂ M; the local ring Ap dominated by M is then unique,
and corresponds to p = m( M ) ∩ A.
P ROOF. If M dominates Ap , then m( M) ∩ Ap = pAp , by (8.1.1), whence the uniqueness of p;
on the other hand, if A ⊂ M, then mM ∩ A = p is prime in A, and since A − p ⊂ M, we have that
Ap ⊂ M and pAp ⊂ m( M ), so M dominates Ap . □
Lemma (8.1.4). — Let R be a field, M and N local subrings of R, and P the subring of R generated by I | 165
M ∪ N. Then the following conditions are equivalent.
(i) There exists a prime ideal p of P such that m( M ) = p ∩ M and m( N ) = p ∩ N.
(ii) The ideal a generated in P by m( M ) ∪ m( N ) is distinct from P.
(iii) There exists a local subring Q of R simultaneously dominating both M and N.
8. CHEVALLEY SCHEMES 162
P ROOF. It is clear that (i) implies (ii); conversely, if a ̸= P, then a is contained in a maximal
ideal n of P, and since 1 ̸∈ n, n ∩ M contains m( M) and is distinct from M, so n ∩ M = m( M),
and similarly n ∩ N = m( N ). It is clear that, if Q dominates both M and N, then P ⊂ Q and
m( M) = m( Q) ∩ M = (m( Q) ∩ P) ∩ M, and m( N ) = (m( Q) ∩ P) ∩ N, so (iii) implies (i); the
converse is evident when we take Q = Pp . □
When the conditions of (8.1.4) are satisfied, we say, with C. Chevalley, that the local rings M and
N are allied.
Proposition (8.1.5). — Let A and B be subrings of a field R, and C the subring of R generated by A ∪ B.
Then the following conditions are equivalent.
(i) For every local ring Q containing A and B, we have that Ap = Bq , where p = m( Q) ∩ A and
q = m( Q) ∩ B.
(ii) For all prime ideals r of C, we have that Ap = Bq , where p = r ∩ A and q = r ∩ B.
(iii) If M ∈ L( A) and N ∈ L( B) are allied, then they are identical.
(iv) L( A) ∩ L( B) = L(C ).
P ROOF. Lemmas (8.1.3) and (8.1.4) prove that (i) and (iii) are equivalent; it is clear that (i) implies
(ii) by taking Q = Cr ; conversely, (ii) implies (i), because if Q contains A ∪ B then it contains C,
and if r = m( Q) ∩ C, then p = r ∩ A and q = r ∩ B, by (8.1.3). It is immediate that (iv) implies (i),
because if Q contains A ∪ B then it dominates a local ring Cr ∈ L(C ) by (8.1.3); by hypothesis we
have that Cr ∈ L( A) ∩ L( B), and (8.1.1) and (8.1.3) prove that Cr = Ap = Bq . We prove finally that
(iii) implies (iv). Let Q ∈ L(C ); Q dominates some M ∈ L( A) and some N ∈ L( B) (8.1.3), so M
and N, being allied, are identical by hypothesis. As we then have that C ⊂ M, we know that M
dominates some Q′ ∈ L(C ) (8.1.3), so Q dominates Q′ , whence necessarily (8.1.3) Q = Q′ = M,
so Q ∈ L( A) ∩ L( B). Conversely, if Q ∈ L( A) ∩ L( B), then C ⊂ Q, so (8.1.3) Q dominates some
Q′′ ∈ L(C ) ⊂ L( A) ∩ L( B); Q and Q′′ , being allied, are identical, so Q′′ = Q ∈ L(C ), which
completes the proof. □
8.2. Local rings of an integral scheme
(8.2.1). Let X be an integral prescheme, and R its field of rational functions, identical to the local
ring of the generic point a of X; for all x ∈ X, we know that Ox can be canonically identified with a
subring of R (7.1.5), and for every rational function f ∈ R, the domain of definition δ( f ) of f is the
open set of x ∈ X such that f ∈ Ox . It thus follows, from (7.2.6), that, for every open U ⊂ X, we
have
Γ(U, OX ) = Ox .
\
(8.2.1.1)
x ∈U
Proposition (8.2.2). — Let X be an integral prescheme, and R its field of rational fractions. For X to be a I | 166
scheme, it is necessary and sufficient for the relation “Ox and Oy are allied” (8.1.4), for points x and y of X,
to imply that x = y.
P ROOF. We suppose that this condition is satisfied, and aim to show that X is separated. Let
U and V be two distinct affine open subsets of X, given by rings A and B (respectively), identified
with subrings of R; U (resp. V) is thus identified (8.1.2) with L( A) (resp. L( B)), and the hypotheses
tell us (8.1.5) that C is the subring of R generated by A ∪ B, and W = U ∩ V is identified with
L( A) ∩ L( B) = L(C ). Furthermore, we know ([CC], p. 5-03, 4 bis) that every subring E of R is equal
to the intersection of the local rings belonging to L( E); C is thus identified with the intersection of the
rings Oz for z ∈ W, or, equivalently (8.2.1.1), with Γ(W, OX ). So consider the subprescheme induced
by X on W; to the identity morphism ϕ : C → Γ(W, OX ) there corresponds (2.2.4) a morphism
Φ = (ψ, θ ) : W → Spec(C ); we will see that Φ is an isomorphism of preschemes, whence W is an
affine open subset. The identification of W with L(C ) = Spec(C ) shows that ψ is bijective. On the
other hand, for all x ∈ W, θ x♯ is the injection Cr → Ox , where r = mx ∩ C, and, by definition, Cr
is identified with Ox , so θ x♯ is bijective. It thus remains to show that ψ is a homeomorphism, or, in
other words, that for every closed subset F ⊂ W, ψ( F ) is closed in Spec(C ). But F is the intersection
of W with a closed subspace of U of the form V (a), where a is an ideal of A; we will show that
ψ( F ) = V (aC ), which proves our claim. In fact, the prime ideals of C containing aC are the prime
8. CHEVALLEY SCHEMES 163
ideals of C containing a, and so are the ideals of the form ψ( x ) = mx ∩ C, where a ⊂ mx and x ∈ W;
since a ⊂ mx is equivalent to x ∈ V (a) = W ∩ F for x ∈ U, we do indeed have that ψ( F ) = V (aC ).
It follows that X is separated, because U ∩ V is affine and its ring C is generated by the union
A ∪ B of the rings of U and V (5.5.6).
Conversely, suppose that X is separated, and let x and y be points of X such that Ox and Oy
are allied. Let U (resp. V) be an affine open subset containing x (resp. y), of ring A (resp. B); we
then know that U ∩ V is affine and that its ring C is generated by A ∪ B (5.5.6). If p = mx ∩ A and
q = my ∩ B, then Ap = Ox and Bq = Oy , and since Ap and Bq are allied, there exists a prime ideal
r of C such that p = r ∩ A and q = r ∩ B (8.1.4). But then there exists a point z ∈ U ∩ V such that
r = mz ∩ C, since U ∩ V is affine, and so evidently x = z and y = z, whence x = y. □
Corollary (8.2.3). — Let X be an integral scheme, and x and y points of X. In order for x ∈ {y}, it is
necessary and sufficient for Ox ⊂ Oy , or, equivalently, for every rational function defined at x to also be
defined at y.
P ROOF. The condition is evidently necessary because the domain of definition δ( f ) of a rational
function f ∈ R is open; we now show that it is sufficient. If Ox ⊂ Oy , then there exists a prime ideal
p of Ox such that Oy dominates (Ox )p (8.1.3); but (2.4.2) there exists some z ∈ X such that x ∈ {z}
and Oz = (Ox )p ; since Oz and Oy are allied, we have that z = y by (8.2.2), whence the corollary. □
Corollary (8.2.4). — If X is an integral scheme then the map x → Ox is injective; equivalently, if x and y
are two distinct points of X, then there exists a rational function defined at one of these points but not the
other.
P ROOF. This follows from (8.2.3) and the axiom (T0 ) (2.1.4). □ I | 167
Corollary (8.2.5). — Let X be an integral scheme whose underlying space is Noetherian; letting f range
over the field R of rational functions on X, the sets δ( f ) generate the topology of X.
In fact, every closed subset of X is thus a finite union of irreducible closed subsets, or, in other
words, of the form {y} (2.1.5). But, if x ̸∈ {y}, then there exists a rational function f defined at x but
not at y (8.2.3), or, equivalently, we have that x ∈ δ( f ) and that δ( f ) is not contained in {y}. The
complement of {y} is thus a union of sets of the form δ( f ), and, by virtue of the first remark, every
open subset of X is the union of finite intersections of open sets of the form δ( f ).
(8.2.6). Corollary (8.2.5) shows that the topology of X is entirely characterised by the data of the local
rings (Ox ) x∈X that have R as their field of fractions. It is equivalent to say that the closed subsets of
X are defined in the following manner: given a finite subset { x1 , . . . , xn } of X, consider the set of
y ∈ X such that Oy ⊂ Oxi for at least one index i, and these sets (over all choices of { x1 , . . . , xn }) are
the closed subsets of X. Further, once the topology on X is known, the structure sheaf OX is also
determined by the family of the Ox , since Γ(U, OX ) = x∈U Ox , by (8.2.1.1). The family (OX ) x∈X
T
thus completely determines the prescheme X when X is an integral scheme whose underlying space
is Noetherian.
Proposition (8.2.7). — Let X and Y be integral schemes, f : X → Y a dominant morphism (2.2.6), and K
(resp.L) the field of rational functions on X (resp.Y). Then L can be identified with a subfield of K, and, for all
x ∈ X, O f ( x) is the unique local ring of Y dominated by Ox .
P ROOF. If f = (ψ, θ ) and a is the generic point of X, then ψ( a) is the generic point of Y (0, 2.1.5);
θ a♯ is then a monomorphism of fields, from L = Oψ(a) to K = Oa . Since every nonempty affine
open subset U of Y contains ψ( a), it follows from (2.2.4) that the homomorphism Γ(U, OY ) →
Γ(ψ−1 (U ), OX ) corresponding to f is the restriction of θ a♯ to Γ(U, OY ). So, for every x ∈ X, θ x♯
is the restriction to Oψ(a) of θ a♯ , and is thus a monomorphism. We also know that θ x♯ is a local
homomorphism, so, if we identify L with a subfield of K by θ a♯ , Oψ( x) is dominated by Ox (8.1.1); it is
also the only local ring of Y dominated by Ox , since two local rings of Y that are allied are identical
(8.2.2). □
Proposition (8.2.8). — Let X be an irreducible prescheme, f : X → Y a local immersion (resp. local
isomorphism), and suppose further that f is separated. Then f is an immersion (resp. an open immersion).
9. SUPPLEMENT ON QUASI-COHERENT SHEAVES 164
P ROOF. Let f = (ψ, θ ); it suffices, in both cases, to prove that ψ is a homeomorphism from X to
ψ( X ) (4.5.3). Replacing f by f red ((5.1.6) and (5.5.1, vi)), we can assume that X and Y are reduced. If
Y ′ is the closed reduced subprescheme of Y that has ψ( X ) as its underlying space, then f factors as
f′ j
X− → Y′ − → Y, where j is the canonical injection (5.2.2). It follows from (5.5.1, v) that f ′ is again a
separated morphism; further, f ′ is again a local immersion (resp. a local isomorphism), because, I | 168
since the condition is local on X and Y, we can restrict to the case where f is a closed immersion
(resp. open immersion), and our claim then follows immediately from (4.2.2).
We can thus suppose that f is a dominant morphism, which leads to the fact that Y is, itself,
irreducible (0, 2.1.5), and so X and Y are both integral. Further, since the condition is local on Y, we
can suppose that Y is an affine scheme; since f is separated, X is a scheme (5.5.1, ii), and we are
finally at the hypotheses of Proposition (8.2.7). Then, for all x ∈ X, θ x♯ is injective; but the hypothesis
that f is a local immersion implies that θ x♯ is surjective (4.2.2), so θ x♯ is bijective, or, equivalently (with
the identification of Proposition (8.2.7)) we have that Oψ( x) = Ox . This implies, by Corollary (8.2.4),
that ψ is an injective map, which already proves the proposition when f is a local isomorphism (4.5.3).
When we suppose that f is only a local immersion, for all x ∈ X there exists an open neighbourhood
U of x in X and an open neighbourhood V of ψ( x ) in Y such that the restriction of ψ to U is a
homeomorphism from U to a closed subset of V. But U is dense in X, so ψ(U ) is dense in Y and a
fortiori in V, which proves that ψ(U ) = V; since ψ is injective, ψ−1 (V ) = U and this proves that ψ is
a homeomorphism from X to ψ( X ). □
(8.3.1). Let X be a Noetherian integral scheme, and R its field of rational functions; we denote by
X ′ the set of local subrings Ox ⊂ R, where x ranges over all points of X. The set X ′ satisfies the
following three conditions.
(Sch. 1) For all M ∈ X ′ , R is the field of fractions of M.
(Sch. 2) There exists a finite set of Noetherian subrings Ai of R such that X ′ = i L( Ai ), and, for all
S
We have seen in (8.2.1) that (Sch. 1) is satisfied, and (Sch. 3) follows from (8.2.2). To show (Sch. 2),
it suffices to cover X by a finite number of affine open subsets Ui whose rings are Noetherian, and to
take Ai = Γ(Ui , OX ); the hypothesis that X is a scheme implies that Ui ∩ Uj is affine, and also that
Γ(Ui ∩ Uj , OX ) = Aij (5.5.6); further, since the space Ui is Noetherian, the immersion Ui ∩ Uj → Ui
is of finite type (6.3.5), so Aij is an Ai -algebra of finite type (6.3.3).
(8.3.2). The structures whose axioms are (Sch. 1), (Sch. 2), and (Sch. 3) generalise “schemes”, in the
sense of C. Chevalley, who additionally supposes that R is an extension of finite type of a field K, and
that the Ai are K-algebras of finite type (which renders a part of (Sch. 2) useless) [CC]. Conversely, if
we have such a structure on a set X ′ , then we can associate to it an integral scheme X by using the
remarks from (8.2.6): the underlying space of X is equal to X ′ endowed with the topology defined in
(8.2.6), and with the sheaf OX such that Γ(U, OX ) = x∈U Ox for all open U ⊂ X, with the evident
T
definition of restriction homomorphisms. We leave to the reader the task of verifying that we thus
obtain an integral scheme, whose local rings are the elements of X ′ ; we will not use this result in
what follows.
Proposition (9.1.1). — Let X be a prescheme (resp. a locally Noetherian prescheme). Let F and G be I | 169
quasi-coherent (resp. coherent) OX -modules; then F ⊗OX G is quasi-coherent (resp. coherent); it is further
of finite type if both F and G are of finite type. If F admits a finite presentation and if G is quasi-coherent
(resp. coherent), then H om(F , G ) is quasi-coherent (resp. coherent).
9. SUPPLEMENT ON QUASI-COHERENT SHEAVES 165
P ROOF. Being a local proposition, we can suppose that X is affine (resp. Noetherian affine);
further, if F is coherent, then we can assume that it is the cokernel of a homomorphism OXm → OXn .
The claims pertaining to quasi-coherent sheaves then follow from Corollaries (1.3.12) and (1.3.9);
the claims pertaining to coherent sheaves follow from Theorem (1.5.1) and from the fact that if M
and N are modules of finite type over a Noetherian ring A then M ⊗ A N and Hom A ( M, N ) are both
A-modules of finite type. □
P ROOF. This follows immediately from the commutativity of the diagram I | 171
f (S′ )
X( S ′ ) /Y ′
(S )
f
X / Y.
□
Corollary (9.1.11). — Let X and Y be S-preschemes, and F (resp. G ) a quasi-coherent OX -module (resp.
quasi-coherent OY -module). Then the inverse image of the sheaf (F(S′ ) ) ⊗(S′ ) (G(S′ ) ) by the canonical
isomorphism ( X ×S Y )(S′ ) ≃ ( X(S′ ) ) ×S′ (Y(S′ ) ) (3.3.10) is equal to (F ⊗S G )(S′ ) .
P ROOF. If p and q are the projections of X ×S Y, then the isomorphism in question is nothing
but ( p(S′ ) , q(S′ ) )S′ ; the corollary then follows from Propositions (9.1.4) and (9.1.10). □
Proposition (9.1.12). — With the notation from Definition (9.1.2), let z be a point of X ×S Y, and
let x = p(z), and y = q(z); the stalk (F ⊗S G )z is isomorphic to (Fx ⊗Ox Oz ) ⊗Oz (Gy ⊗Oy Oz ) =
F x ⊗ O x Oz ⊗ Oy ⊗ Gy .
P ROOF. Since we can reduce to the affine case, the proposition follows from Equation (1.6.5.1).
□
Corollary (9.1.13). — If F and G are of finite type, then
Supp(F ⊗S G ) = p−1 (Supp(F )) ∩ q−1 (Supp(G )).
P ROOF. Since p∗ (F ) and q∗ (G ) are both of finite type over OX ×S Y , we reduce, by Proposi-
tion (9.1.12) and by (0, 1.7.5), to the case where G = OY , that is, it remains to prove the following
equation:
(9.1.13.1) Supp( p∗ (F )) = p−1 (Supp(F )).
The same reasoning as in (0, 1.7.5) leads us to prove that, for all z ∈ X ×S Y, we have Oz /mx Oz ̸=
0 (with x = p(z)), which follows from the fact that the homomorphism Ox → Oz is local, by
hypothesis. □
9. SUPPLEMENT ON QUASI-COHERENT SHEAVES 167
We leave it to the reader to extend the results in this section to the more general case of an
arbitrary (but finite) number of factors, instead of just two.
9.2. Direct image of a quasi-coherent sheaf
Proposition (9.2.1). — Let f : X → Y be a morphism of preschemes. We suppose that there exists a cover
(Yα ) of Y by affine opens having the following property: every f −1 (Yα ) admits a finite cover ( Xαi ) by affine
opens that are contained in f −1 (Yα ) and that are such that every intersection Xαi ∩ Xαj is itself a finite union
of affine opens. With these hypotheses, for every quasi-coherent OX -module F , f ∗ (F ) is a quasi-coherent
OY -module.
P ROOF. Since this is a local condition on Y, we can assume that Y is equal to one of the Yα , and
thus omit the indices α.
(a) First, suppose that the Xi ∩ X j are themselves affine opens. We set Fi = F | Xi and Fij =
F |( Xi ∩ X j ), and let Fi′ and Fij′ be the images of Fi and Fij (respectively) by the restriction
of f to Xi and to Xi ∩ X j (respectively); we know that the Fi′ and Fij′ are quasi-coherent
(1.6.3). Set G = i Fi′ and H = i,j Fij′ ; G and H are quasi-coherent OY -modules; we
L L
that satisfies the usual compatibility conditions when we let W vary. If, for every section
si ∈ Γ( f −1 (W ) ∩ Xi , F ), we denote by si| j its restriction to f −1 (W ) ∩ Xi ∩ X j , then we set
uW ( s i ) = ( s i | j − s j | i )
and the compatibility conditions are clearly satisfied. To prove that the kernel R of u
is f ∗ (F ), we define a homomorphism from f ∗ (F ) to R by sending each section s ∈
Γ( f −1 (W ), F ) to the family (si ), where si is the restriction of s to f −1 (W ) ∩ Xi ; axioms (F1)
and (F2) of sheaves (G, II, 1.1) tell us that this homomorphism is bijective, which finishes
the proof in this case.
(b) In the general case, the same reasoning applies once we have established that the Fij are
quasi-coherent. But, by hypothesis, Xi ∩ X j is a finite union of affine opens Xijk ; and since
the Xijk are affine opens in a scheme, the intersection of any two of them is again an affine
open (5.5.6). We are thus led to the first case, and so we have proved Proposition (9.2.1).
□
Corollary (9.2.2). — The conclusion of Proposition (9.2.1) holds true in each of the following cases:
(a) f is separated and quasi-compact;
(b) f is separated and of finite type;
(c) f is quasi-compact, and the underlying space of X is locally Noetherian.
P ROOF. In case (a), the Xαi ∩ Xαj are affine (5.5.6). Case (b) is a particular example of case (a)
(6.6.3). Finally, in case (c), we can reduce to the case where Y is affine and the underlying space of X
is Noetherian; then X admits a finite cover of affine opens ( Xi ), and the Xi ∩ X j , being quasi-compact,
are finite unions of affine opens (2.1.3). □
9. SUPPLEMENT ON QUASI-COHERENT SHEAVES 168
Theorem (9.3.1). — Let X be a prescheme whose underlying space is Noetherian, or a scheme whose
underlying space is quasi-compact. Let L be an invertible OX -module (0, 5.4.1), f a section of L over X, X f
the open set of x ∈ X such that f ( x ) ̸= 0 (0, 5.5.1), and F a quasi-coherent OX -module.
(i) If s ∈ Γ( X, F ) is such that s| X f = 0, then there exists an integer n > 0 such that s ⊗ f ⊗n = 0.
(ii) For every section s ∈ Γ( X f , F ), there exists an integer n > 0 such that s ⊗ f ⊗n extends to a
section of F ⊗ L ⊗n over X.
P ROOF.
(i) Since the underlying space of X is quasi-compact, and thus the union of finitely-many
affine opens Ui with L |Ui isomorphic to OX |Ui , we can reduce to the case where X is affine
and L = OX . In this case, f can be identified with an element of A( X ), and we have that
X f = D ( f ); s can be identified with an element of an A( X )-module M, and s| X f to the
corresponding element of M f , and the result is then trivial, recalling the definition of a
module of fractions. I | 173
(ii) Again, X is a finite union of affine opens Ui (1 ⩽ i ⩽ r) such that L |Ui ∼ = OX |Ui , and, for
every i, (s ⊗ f ⊗n )|(Ui ∩ X f ) can be identified (by the aforementioned isomorphism) with
( f |(Ui ∩ X f ))n (s|(Ui ∩ X f )). We then know (1.4.1) that there exists an integer n > 0 such
that, for all i, (s ⊗ f ⊗n )|(Ui ∩ X f ) extends to a section si of F ⊗ L ⊗n over Ui . Let si| j be
the restriction of si to Ui ∩ Uj ; by definition we have that si| j − s j|i = 0 on X f ∩ Ui ∩ Uj . But,
if X is a Noetherian space, then Ui ∩ Uj is quasi-compact; if X is a scheme, then Ui ∩ Uj
is an affine open (5.5.6), and so again quasi-compact. By virtue of (i), there thus exists an
integer m (independent of i and j) such that (si| j − s j|i ) ⊗ f ⊗m = 0. It immediately follows
that there exists a section s′ of F ⊗ L ⊗(n+m) over X that restricts to si ⊗ f ⊗m over each Ui ,
and restricts to s ⊗ f ⊗(n+m) over X f .
□
The following corollaries give an interpretation of Theorem (9.3.1) in a more algebraic language:
Corollary (9.3.2). — With the hypotheses of (9.3.1), consider the graded ring A• = Γ• (L ) and the graded
A• -module M∗ = Γ• (L , F ) (0, 5.4.6). If f ∈ An , where n ∈ Z, then there is a canonical isomorphism
Γ( X f , F ) ≃ (( M∗ ) f )0 (the subgroup of the module of fractions ( M∗ ) f consisting of elements of
degree 0).
Corollary (9.3.3). — Suppose that the hypotheses of (9.3.1) are satisfied, and suppose further that L = OX .
Then, setting A = Γ( X, OX ) and M = Γ( X, F ), the A f -module Γ( X f , F ) is canonically isomorphic to
Mf .
P ROOF. Since X is a union of finitely-many affine opens whose rings are Noetherian, we can
suppose that X is affine, given by some Noetherian ring A; then F = M, e where M = Γ( X, F ) is
an A-module of finite type, and J = J, where J = Γ( X, J ) is an ideal of A ((1.4.1) and (1.5.1)).
e
Since A is Noetherian, J admits a finite system of generators f i (1 ⩽ i ⩽ m). By hypothesis, every
section of F over X is zero on each of the D ( f i ); if s j (1 ⩽ j ⩽ q) are sections of F that generate M,
then there exists an integer h, independent of i and j, such that f ih s j = 0 (1.4.1), whence f ih s = 0 for
all s ∈ M. We thus conclude that, if n = mh, then Jn M = 0, and so the corresponding OX -module
J n F = J] n M (1.3.13) is zero. □
Corollary (9.3.5). — With the hypotheses of (9.3.4), there exists a closed subprescheme Y of X, whose
underlying space is the support of OX /J , such that, if j : Y → X is the canonical injection, then
F = j∗ ( j∗ (F )).
9. SUPPLEMENT ON QUASI-COHERENT SHEAVES 169
P ROOF. First, note that the supports of OX /J and OX /J n are the same, since, if Jx = Ox ,
then Jxn = Ox , and we also have that Jxn ⊂ Jx for all x ∈ X. We can, thanks to (9.3.4), thus
suppose that J F = 0; we can then take Y to be the closed subprescheme of X defined by J , and
since F is then an (OX /J )-module, the conclusion follows immediately. □
9.4. Extension of quasi-coherent sheaves
(9.4.1). Let X be a topological space, F a sheaf of sets (resp. of groups, of rings) on X, U an open I | 174
subset of X, ψ : U → X the canonical injection, and G a subsheaf of F |U = ψ∗ (F ). Since ψ∗ is
left exact, ψ∗ (G ) is a subsheaf of ψ∗ (ψ∗ (F )); we denote by ρ the canonical homomorphism F →
ψ∗ (ψ∗ (F )) (0, 3.5.3), and we denote by G the subsheaf ρ−1 (ψ∗ (G )) of F . It follows immediately
from the definitions that, for every open subset V of X, Γ(V, G ) consists of sections s ∈ Γ(V, F )
whose restriction to V ∩ U is a section of G over V ∩ U. We thus have that G |U = ψ∗ (G ) = G , and
that G is the largest subsheaf of F that restricts to G over U; we say that G is the canonical extension
of the subsheaf G of F |U to a subsheaf of F .
Proposition (9.4.2). — Let X be a prescheme, and U an open subset of X such that the canonical injection
j : U → X is a quasi-compact morphism (which will be the case for all U if the underlying space of X is
locally Noetherian (6.6.4, i)). Then:
(i) for every quasi-coherent (OX |U )-module G , j∗ (G ) is a quasi-coherent OX -module, and j∗ (G )|U =
j∗ ( j∗ (G )) = G ;
(ii) for every quasi-coherent OX -module F and every quasi-coherent (OX |U )-submodule G , the canon-
ical extension G of G (9.4.1) is a quasi-coherent OX -submodule of F .
P ROOF. If j = (ψ, θ ) (ψ being the injection U → X of underlying spaces), then, by definition, we
have that j∗ (G ) = ψ∗ (G ) for every (OX |U )-module G , and, further, that j∗ (H ) = ψ∗ (H ) = H |U
for every OX -module H , by definition of the prescheme induced over an open subset. So (i) is thus
a particular case of ((9.2.2, a)); for the same reason, j∗ ( j∗ (F )) is quasi-coherent, and since G is the
inverse image of j∗ (G ) by the homomorphism ρ : F → j∗ ( j∗ (F )), (ii) follows from (4.1.1). □
Note that the hypothesis that the morphism j : U → X is quasi-compact holds whenever the
open subset U is quasi-compact and X is a scheme: indeed, U is then a union of finitely-many affine
opens Ui , and, for every affine open V of X, V ∩ Ui is an affine open (5.5.6), and thus quasi-compact.
Corollary (9.4.3). — Let X be a prescheme, and U a quasi-compact open subset of X such that the injection
morphism j : U → X is quasi-compact. Suppose as well that every quasi-coherent OX -module is the inductive
limit of its quasi-coherent OX -submodules of finite type (which will be the case if X is an affine scheme).
Then let F be a quasi-coherent OX -module, and G a quasi-coherent (OX |U )-submodule of F |U of finite
type. Then there exists a quasi-coherent OX -submodule G ′ of F of finite type such that G ′ |U = G .
P ROOF. We have G = G |U, and G is quasi-coherent, from (9.4.2), so the inductive limit of its
quasi-coherent OX -submodules Hλ of finite type. It follows that G is the inductive limit of the Hλ |U,
and thus equal to one of the Hλ |U, since it is of finite type (0, 5.2.3). □
Remark (9.4.4). — Suppose that for every affine open U ⊂ X, the injection morphism U → X
is quasi-compact. Then, if the conclusion of (9.4.3) holds for every affine open U and for every
quasi-coherent (OX |U )-submodule G of F |U of finite type, it follows that F is the inductive limit I | 175
of its quasi-coherent OX -submodules of finite type. Indeed, for every affine open U ⊂ X, we
have that F |U = M, e where M is an A(U )-module, and since the latter is the inductive limit of its
quasi-coherent submodules of finite type, F |U is the inductive limit of its (OX |U )-submodules of
finite type (1.3.9). But, by hypothesis, each of these submodules is induced on U by a quasi-coherent
OX -submodule Gλ,U of F of finite type. The finite sums of the Gλ,U are again quasi-coherent OX -
modules of finite type, because this is a local property, and the case where X is affine was covered in
(1.3.10); it is clear then that F is the inductive limit of these finite sums, whence our claim.
Corollary (9.4.5). — Under the hypotheses of Corollary (9.4.3), for every quasi-coherent (OX |U )-module G
of finite type, there exists a quasi-coherent OX -module G ′ of finite type such that G ′ |U = G .
P ROOF. Since F = j∗ (G ) is quasi-coherent (9.4.2) and F |U = G , it suffices to apply Corol-
lary (9.4.3) to F . □
9. SUPPLEMENT ON QUASI-COHERENT SHEAVES 170
Lemma (9.4.6). — Let X be a prescheme, L a well-ordered set, (Vλ )λ∈ L a cover of X by affine opens,
and U an open subset of X; for all λ ∈ L, we set Wλ = µ<λ Vµ . Suppose that: (1) for every λ ∈ L,
S
Vλ ∩ Wλ is quasi-compact; and (2) the immersion morphism U → X is quasi-compact. Then, for every
quasi-coherent OX -module F and every quasi-coherent (OX |U )-submodule G of F |U of finite type, there
exists a quasi-coherent OX -submodule G ′ of F of finite type such that G ′ |U = G .
P ROOF. Let Uλ = U ∪ Wλ ; we will define a family (Gλ′ ) by induction, where Gλ′ is a quasi-
coherent (OX |Uλ )-submodule of F |Uλ of finite type, such that Gλ′ |Uµ = Gµ′ for µ < λ and Gλ′ |U = G .
The unique OX -submodule G ′ of F such that G ′ |Uλ = G ′ for all λ ∈ L (0, 3.3.1) will then give us
what we want. So suppose that the Gµ′ are defined and have the preceding properties for µ < λ;
if λ does not have a predecessor then we take Gλ′ to be the unique (OX |Uλ )-submodule of F |Uλ
such that Gλ′ |Uµ = Gµ′ for all µ < λ, which is allowed since the Uµ with µ < λ then form a cover
of Uλ . If, conversely, λ = µ + 1, then Uλ = Uµ ∪ Vµ , and it suffices to define a quasi-coherent
(OX |Vµ )-submodule Gµ′′ of F |Vµ of finite type such that
Gµ′′ |(Uµ ∩ Vµ ) = Gµ′ |(Uµ ∩ Vµ );
and then to take Gλ′ to be the (OX |Uλ )-submodule of F |Uλ such that Gλ′ |Uµ = Gµ′ and Gλ′ |Vµ = Gµ′′
(0, 3.3.1). But, since Vµ is affine, the existence of Gµ′′ is guaranteed by (9.4.3) as soon as we show
that Uµ ∩ Vµ is quasi-compact; but Uµ ∩ Vµ is the union of U ∩ Vµ and Wµ ∩ Vµ , which are both
quasi-compact by virtue of the hypotheses. □
Theorem (9.4.7). — Let X be a prescheme, and U an open subset of X. Suppose that one of the following
conditions is verified:
(a) the underlying space of X is locally Noetherian;
(b) X is a quasi-compact scheme and U is a quasi-compact open.
Then, for every quasi-coherent OX -module F and every quasi-coherent (OX |U )-submodule G of F |U of
finite type, there exists a quasi-coherent OX -submodule G ′ of F of finite type such that G ′ |U = G .
P ROOF. Let (Vλ )λ∈ L be a cover of X by affine opens, with L assumed to be finite in case (b); I | 176
since L is equipped with the structure of a well-ordered set, it suffices to check that the conditions of
(9.4.6) are satisfied. It is clear in the case of (a), since the spaces Vλ are Noetherian. For case (b), the
Vλ ∩ λµ are affine (5.5.6), and thus quasi-compact, and, since L is finite, Vλ ∩ Wλ is quasi-compact.
Whence the theorem. □
Corollary (9.4.8). — Under the hypotheses of (9.4.7), for every quasi-coherent (OX |U )-module G of finite
type, there exists a quasi-coherent OX -module G ′ of finite type such that G ′ |U = G .
P ROOF. It suffices to apply (9.4.7) to F = j∗ (G ), which is quasi-coherent (9.4.2) and such that
F |U = G . □
Corollary (9.4.9). — Let X be a prescheme whose underlying space is locally Noetherian, or a quasi-compact
scheme. Then every quasi-coherent OX -module is the inductive limit of its quasi-coherent OX -submodules of
finite type.
P ROOF. This follows from Theorem (9.4.7) and Remark (9.4.4). □
Corollary (9.4.10). — Under the hypotheses of (9.4.9), if a quasi-coherent OX -module F is such that every
quasi-coherent OX -submodule of finite type of F is generated by its sections over X, then F is generated by
its sections over X.
P ROOF. Let U be an affine open neighbourhood of a point x ∈ X, and let s be a section of F over
U; the OX -submodule G of F |U generated by s is quasi-coherent and of finite type, so there exists a
quasi-coherent OX -submodule G ′ of F of finite type such that G ′ |U = G (9.4.7). By hypothesis, there
thus exists a finite number of sections ti of G ′ over X and of sections ai of OX over a neighbourhood
V ⊂ U of x such that s|V = ∑i ai (ti |V ), which proves the corollary. □
9. SUPPLEMENT ON QUASI-COHERENT SHEAVES 171
Definition (9.5.3). — Whenever there exists a smaller subprescheme Y ′ of Y such that f factors
through the canonical injection j : Y ′ → Y, we say that Y ′ is the closed image prescheme of X under
the morphism f .
Proposition (9.5.4). — If f ∗ (OX ) is a quasi-coherent OY -module, then the underlying space of the closed
image of X under f is the closure f ( X ) in Y.
P ROOF. As the support of f ∗ (OX ) is contained in f ( X ), we have (with the notation of (9.5.2))
Jy = Oy for y ̸∈ f ( X ), thus the support of OY /J is contained in f ( X ). In addition, this support is
closed and contains f ( X ): indeed, if y ∈ f ( X ), the unit element of the ring (ψ∗ (OX ))y is not zero,
being the germ at y of the section
1 ∈ Γ( X, OX ) = Γ(Y, ψ∗ (OX ));
since it is the image under θ of the unit element of Oy , the latter does not belong to Jy , hence
Oy /Jy ̸= 0; this finishes the proof. □
Corollary (9.5.6). — Let f : X → Y be an S-morphism such that Y is the closed image of X under f . Let Z
be an S-scheme; if two S-morphisms g1 , g2 from Y to Z are such that g1 ◦ f = g2 ◦ f , then g1 = g2 .
9. SUPPLEMENT ON QUASI-COHERENT SHEAVES 172
Proposition (9.5.1) are satisfied, then it follows from (9.5.4) that f ( X ) is dense in Y ′ ; if X is irreducible,
then so is Y ′ (0, 2.1.5). The claim about integral preschemes follows from the conjunction of the two
others. □
Proposition (9.5.10). — Let Y be a subprescheme of a prescheme X, such that the canonical injection
i : Y → X is a quasi-compact morphism. Then there exists a smaller closed subprescheme Y of X containing
Y; its underlying space is the closure of that of Y; the latter is open in its closure, and the prescheme Y is
induced on this open by Y.
P ROOF. It suffices to apply Proposition (9.5.1) to the injection j, which is separated (5.5.1) and
quasi-compact by hypothesis; (9.5.1) thus proves the existence of Y, and (9.5.4) shows that its
underlying space is the closure of Y in X; since Y is locally closed in X, it is open in Y, and the last
claim comes from (9.5.8) applied to an open subset V of X such that Y is closed in V. □
With the above notation, if the injection V → X is quasi-compact, and if J is the quasi-coherent
sheaf of ideals of OX |V defining the closed subprescheme Y of V, it follows from Proposition (9.5.1)
that the quasi-coherent sheaf of ideals of OX defining Y is the canonical extension (9.4.1) J of J ,
because it is clearly the largest quasi-coherent subsheaf of ideals of OX inducing J on V.
Corollary (9.5.11). — Under the hypotheses of Proposition (9.5.10), every section of OY over an open V of I | 179
Y that is zero on V ∩ Y is zero.
P ROOF. By Proposition (9.5.8), we can reduce to the case where V = Y. If we take into account
that the sections of OY over Y canonically correspond to the Y-sections of Y ⊗ Z Z [ T ] (3.3.15) and
that the latter is separated over Y, then the corollary appears as a specific case of (9.5.6). □
9. SUPPLEMENT ON QUASI-COHERENT SHEAVES 173
Proposition (9.6.1). — Let X be a prescheme, and B a quasi-coherent OX -algebra (0, 5.1.3). For a B-module
F to be quasi-coherent (on the ringed space ( X, B )) it is necessary and sufficient that F be a quasi-coherent
OX -module.
P ROOF. Since the question is local, we can assume X to be affine, given by some ring A, and
thus B = B,e where B is an A-algebra (1.4.3). If F is quasi-coherent on the ringed space ( X, B )
then we can also assume that F is the cokernel of a B-homomorphism B ( I ) → B ( J ) ; since this
homomorphism is also an OX -homomorphism of OX -modules, and B ( I ) and B ( J ) are quasi-coherent
OX -modules (1.3.9, ii), F is also a quasi-coherent OX -module (1.3.9, i).
Conversely, if F is a quasi-coherent OX -module, then F = M, e where M is a B-module (1.4.3);
M is isomorphic to the cokernel of a B-homomorphism B → B ) , so F is a B-module isomorphic
( I ) ( J
(9.6.2). Given a prescheme X, we say that a quasi-coherent OX -algebra B is of finite type if, for
all x ∈ X, there exists an open affine neighbourhood U of x such that Γ(U, B ) = B is an algebra
of finite type over Γ(U, OX ) = A. We then have that B |U = B e and, for all f ∈ A, the induced
(OX | D ( f ))-algebra B | D ( f ) is of finite type, because it is isomorphic to ( B f )∼ , and B f = B ⊗ A A f is
clearly an algebra of finite type over A f . Since the D ( f ) form a basis for the topology of U, we thus
conclude that if B is a quasi-coherent OX -algebra of finite type then, for every open V of X, B |V is
a quasi-coherent (OX |V )-algebra of finite type.
Proposition (9.6.3). — Let X be a locally Noetherian prescheme. Then every quasi-coherent OX -algebra B
of finite type is a coherent sheaf of rings (0, 5.3.7).
P ROOF. We can once again restrict to the case where X is an affine scheme given by a Noetherian
ring A, and where B = B, e with B being an A-algebra of finite type; B is then a Noetherian ring.
With this, it remains to prove that the kernel N of a B-homomorphism B m → B is a B-module of I | 180
finite type; but it is isomorphic (as a B-module) to N,
e where N is the kernel of the corresponding
homomorphism of B-modules Bm → B (1.3.13). Since B is Noetherian, the submodule N of Bm
is a B-module of finite type, so there exists a homomorphism B p → Bm with image N; since the
sequence B p → Bm → B is exact, so too is the corresponding sequence B p → B m → B (1.3.5), and
since N is the image of B p → B m (1.3.9, i), this proves the proposition. □
Corollary (9.6.4). — Under the hypotheses of (9.6.3), for a B-module F to be coherent, it is necessary and
sufficient that it be a quasi-coherent OX -module and a B-module of finite type. If this is the case, and if G is a
B-submodule or a quotient module of F , then in order for G to be a coherent B-module, it is necessary and
sufficient that it be a quasi-coherent OX -module.
P ROOF. Taking (9.6.1) into account, the conditions on F are clearly necessary; we will show
that they are sufficient. We can restrict to the case where X is affine, given by some Noetherian
ring A, where B = B, e with B an A-algebra of finite type, where F = M, e with M a B-module, and
where there exists a surjective B-homomorphism B m → F → 0. We then have the corresponding
exact sequence Bm → M → 0, so M is a B-module of finite type; further, the kernel P of the
homomorphism Bm → M is then a B-module of finite type, since B is Noetherian. We thus conclude
(1.3.13) that F is the cokernel of a B-homomorphism B m → B n , and is thus coherent, since B is a
coherent sheaf of rings (0, 5.3.4). The same reasoning shows that a quasi-coherent B-submodule
(resp. a quotient B-module) of F is of finite type, from whence the second part of the corollary. □
10. FORMAL SCHEMES 174
Proposition (9.6.6). — Let X be a prescheme whose underlying space is locally Noetherian, or a quasi-
compact scheme. Then every quasi-coherent OX -algebra B is the inductive limit of its quasi-coherent
OX -subalgebras of finite type.
P ROOF. In fact, it follows from (9.4.9) that B is the inductive limit (as an OX -module) of its
quasi-coherent OX -submodules of finite type; the latter generating quasi-coherent OX -subalgebras of
B of finite type (1.3.14), and so B is a fortiori their inductive limit. □
Proposition (10.1.4). — Let A be an admissible ring, X = Spf( A), and, for every f ∈ A, let D( f ) =
D ( f ) ∩ X; then the topologically ringed space (D( f ), OX |D( f )) is isomorphic to the formal affine spectrum
Spf( A{ f } ) (0, 7.6.15).
P ROOF. The condition is necessary: let p = jx ∈ Spf( A), and let q = ϕ−1 (jx ); if g ̸∈ q, then
we have ϕ( g) ̸∈ p, and it is immediate that the homomorphism B{ g} → A{ϕ( g)} induced by ϕ
(0, 7.6.7) sends q{ g} to a subset of p{ϕ( g)} ; by passing to the inductive limit, we see (taking (10.1.5)
and (0, 7.6.17) into account) that ϕex♯ is a local homomorphism.
Conversely, let (ψ, θ ) be a morphism satisfying the condition of the proposition; by (10.1.3), θ
defines a continuous ring homomorphism
ϕ = Γ(θ ) : B = Γ(Y, OY ) −→ Γ(X, OX ) = A.
By virtue of the hypothesis on θ, for the section ϕ( g) of OX over X to be an invertible germ at the
point x, it is necessary and sufficient that g be an invertible germ at the point ψ( x ). But, by (0, 7.6.17),
the sections of OX (resp. OY ) over X (resp. Y) that have a non-invertible germ at the point x (resp.
ψ( x )) are exactly the elements of jx (resp. jψ( x) ); the above remark thus shows that a ϕ = ψ. Finally, I | 183
for all g ∈ B the diagram
/ Γ(X, OX ) = A
ϕ
B = Γ(Y, OY )
Γ ( θD( g ) )
B{ g} = Γ(D( g), OY ) / Γ(D(ϕ( g)), OX ) = A
{ϕ( g)}
10. FORMAL SCHEMES 176
is commutative; by the universal property of completed rings of fractions (0, 7.6.6), θD( g) is equal to
eD( g) for all g ∈ B, and so (0, 3.2.5) we have θ = ϕ
ϕ e. □
We say that a morphism (ψ, θ ) of topologically ringed spaces satisfying the condition of Propo-
sition (10.2.2) is a morphism of formal affine schemes. We can say that the functors Spf( A) in A and
Γ(X, OX ) in X define an equivalence between the category of admissible rings and the opposite
category of formal affine schemes (T, I, 1.2).
(10.2.3). As a particular case of (10.2.2), note that, for f ∈ A, the canonical injection of the formal
affine scheme induced by X on D( f ) corresponds to the continuous canonical homomorphism
A → A{ f } . Under the hypotheses of Proposition (10.2.2), let h be an element of B, and g an element
of A that is a multiple of ϕ(h); we then have ψ(D( g)) ⊂ D(h); the restriction of u to D( g), considered
as a morphism from D( g) to D(h), is the unique morphism v making the diagram
D( g )
v / D( h )
X
u /Y
commutate.
This morphism corresponds to the unique continuous homomorphism ϕ′ : B{h} → A{ g} (0, 7.6.7)
making the diagram
Ao
ϕ
B
ϕ′
A{ g} o B{h}
commutate.
10.3. Ideals of definition for a formal affine scheme
(10.3.1). Let A be an admissible ring, J an open ideal of A, and X the formal affine scheme Spf( A).
Let (Jλ ) be the set of those ideals of definition for A that are contained in J; then e J/eJλ is a sheaf of
ideals of A/e e ∆
Jλ . Denote by J the projective limit of the sheaves on X induced by e J/eJλ , which is
identified with a sheaf of ideals of OX (0, 3.2.6). For every f ∈ A, Γ(D( f ), J∆ ) is the projective limit
of the S− 1 −1
f J/S f Jλ , or, in other words, can be identified with the open ideal J{ f } of the ring A{ f }
(0, 7.6.9), and, in particular, Γ(X, J∆ ) = J; we conclude (the D( f ) forming a basis for the topology of
X) that
(10.3.1.1) J∆ |D ( f ) = (J{ f } ) ∆ .
(10.3.2). With the notation of (10.3.1), for all f ∈ A, the canonical map from A{ f } = Γ(D( f ), OX ) I | 184
to Γ(D( f ), ( A/
e eJ)|X) = S−1 A/S−1 J is surjective and has Γ(D( f ), J∆ ) = J{ f } as its kernel (0, 7.6.9);
f f
these maps thus define a surjective continuous homomorphism, said to be canonical, from the sheaf of
topological rings OX to the sheaf of discrete rings ( A/
e eJ)|X, whose kernel is J∆ ; this homomorphism
is none other than ϕ e (10.2.1), where ϕ is the continuous homomorphism A → A/J; the morphism
( ϕ, ϕ) : Spec( A/J) → X of formal affine schemes (where a ϕ is the identity homeomorphism from
a e
X to itself) is also said to be canonical. We thus have, according to the above, a canonical isomorphism
(10.3.2.1) OX /J∆ ≃ ( A/
e eJ)|X.
It is clear (since Γ(X, J∆ ) = J) that the map J → J∆ is strictly increasing; according to the above,
∆
for J ⊂ J′ , the sheaf J′ /J∆ is canonically isomorphic to Je′ /e
J = (J′ /J)e.
(10.3.3). The hypotheses and notation being the same as in (10.3.1), we say that a sheaf of ideals J
of OX is a sheaf of ideals of definition for X (or an ideal sheaf of definition for X) if, for all x ∈ X, there
exists an open neighbourhood of x of the form D( f ), where f ∈ A, such that J |D( f ) is of the form
H∆ , where H is an ideal of definition for A{ f } .
10. FORMAL SCHEMES 177
Proposition (10.3.4). — For all f ∈ A, each sheaf of ideals of definition for X induces a sheaf of ideals of
definition for D( f ).
P ROOF. This follows from (10.3.1.1). □
Proposition (10.3.5). — If A is an admissible ring, then every sheaf of ideals of definition for X = Spf( A)
is of the form J∆ , where J is a uniquely determined ideal of definition for A.
P ROOF. Let J be a sheaf of ideals of definition of X; by hypothesis, and since X is quasi-compact,
there are finitely-many elements f i ∈ A such that the D( f i ) cover X and such that J |D( f i ) = Hi∆ ,
where Hi is an ideal of definition for A{ f i } . For each i, there exists an open ideal Ki of A such that
(Ki ){ f i } = Hi (0, 7.6.9); let K be an ideal of definition for A containing all the Ki . The canonical image
of J /K∆ in the structure sheaf ( A/K)e of Spec( A/K) (10.3.2) is thus such that its restriction to
D( f i ) is equal to its restriction to (Ki /K)e; we conclude that this canonical image is a quasi-coherent
sheaf on Spec( A/K), and so is of the form (J/K)e, where J is an ideal of definition for A containing
K (1.4.1), whence J = J∆ (10.3.2); in addition, since for each i there exists an integer ni such that
Hi i ⊂ K{ f i } , we will have, by setting n to be the largest of the ni , that (J /K∆ )n = 0, and, as a result
n
(10.3.2), that ((J/K)e)n = 0, whence finally that (J/K)n = 0 (1.3.13), which proves that J is an ideal
of definition for A (0, 7.1.4). □
Proposition (10.3.6). — Let A be an adic ring, and J an ideal of definition for A such that J/J2 is an
( A/J)-module of finite type. For any integer n > 0, we then have (J∆ )n = (Jn )∆ .
P ROOF. For all f ∈ A, we have (since Jn is an open ideal)
by (10.3.1.1) and (0, 7.6.12). The result then follows from the fact that (J∆ )n is associated to the I | 185
presheaf U 7→ (Γ(U, J∆ ))n (0, 4.1.6), since the D( f ) form a basis for the topology of X. □
(10.3.7). We say that a family (Jλ ) of sheaves of ideals of definition for X is a fundamental system of
sheaves of ideals of definition if each sheaf of ideals of definition for X contains one of the Jλ ; since
Jλ = J∆ λ , it is equivalent to say that the Jλ form a fundamental system of neighbourhoods of 0 in A. Let
( f α ) be a family of elements of A such that the D( f α ) cover X. If (Jλ ) is a filtered decreasing family
of sheaves of ideals of OX such that, for each α, the family (Jλ |D( f α )) is a fundamental system of
sheaves of ideals of definition for D( f α ), then (Jλ ) is a fundamental system of sheaves of ideals of
definition for X. Indeed, for each sheaf of ideals of definition J for X, there is a finite cover of X
by D( f i ) such that, for each i, Jλi |D( f i ) is a sheaf of ideals of definition for D( f i ) that is contained
in J |D( f i ). If µ is an index such that Jµ ⊂ Jλi for all i, then it follows from (10.3.3) that Jµ is a
sheaf of ideals of definition for X, evidently contained in J , whence our claim.
10.4. Formal preschemes and morphisms of formal preschemes
(10.4.1). Given a topologically ringed space X, we say that an open U ⊂ X is a formal affine open
(resp. a formal adic affine open, resp. a formal Noetherian affine open) if the topologically ringed space
induced on U by X is a formal affine scheme (resp. a scheme whose ring is adic, resp. adic and
Noetherian).
Definition (10.4.2). — A formal prescheme is a topologically ringed space X which admits a formal
affine open neighbourhood for each point. We say that the formal prescheme X is adic (resp. locally
Noetherian) if each point of X admits a formal adic (resp. Noetherian) open neighbourhood. We say
that X is Noetherian if it is locally Noetherian and its underlying space is quasi-compact (and hence
Noetherian).
Proposition (10.4.3). — If X is a formal prescheme (resp. a locally Noetherian formal prescheme), then the
formal affine (resp. Noetherian affine) open sets form a basis for the topology of X.
P ROOF. This follows from Definition (10.4.2) and Proposition (10.1.4) by taking into account
that, if A is an adic Noetherian ring, then so too is A{ f } for all f ∈ A (0, 7.6.11). □
10. FORMAL SCHEMES 178
Corollary (10.4.4). — If X is a formal prescheme (resp. a locally Noetherian formal prescheme, resp. a
Noetherian formal prescheme), then the topologically ringed space induced on each open set of X is also a
formal prescheme (resp. a locally Noetherian formal prescheme, resp. a Noetherian formal prescheme).
Definition (10.4.5). — Given two formal preschemes X and Y, a morphism (of formal preschemes)
from X to Y is a morphism (ψ, θ ) of topologically ringed spaces such that, for all x ∈ X, θ x♯ is a local
homomorphism Oψ( x) → Ox .
It is immediate that the composition of any two morphisms of formal preschemes is again a
morphism of formal preschemes; the formal preschemes thus form a category, and we denote by
Hom(X, Y) the set of morphisms from a formal prescheme X to a formal prescheme Y.
If U is an open subset of X, then the canonical injection into X of the formal prescheme induced I | 186
on U by X is a morphism of formal preschemes (and in fact a monomorphism of topologically ringed
spaces (0, 4.1.1)).
Proposition (10.4.6). — Let X be a formal prescheme, and S = Spf( A) a formal affine scheme. There exists
a canonical bijective equivalence between the morphisms from the formal prescheme X to the formal prescheme
S and the continuous homomorphisms from the ring A to the topological ring Γ(X, OX ).
P ROOF. The proof is similar to that of (2.2.4), by replacing “homomorphism” with “continuous
homomorphism”, “affine open” with “formal affine open”, and by using Proposition (10.2.2) instead
of Theorem (1.7.3); we leave the details to the reader. □
(10.4.7). Given a formal prescheme S, we say that the data of a formal prescheme X and a morphism
ϕ : X → S defines a formal prescheme X over S or an formal S-prescheme, ϕ being called the structure
morphism of the S-prescheme X. If S = Spf( A), where A is an admissible ring, then we also say that
the formal S-prescheme X is a formal A-prescheme or a formal prescheme over A. An arbitrary formal
prescheme can be considered as a formal prescheme over Z (equipped with the discrete topology).
If X and Y are formal S-preschemes, then we say that a morphism u : X → Y is a S-morphism if
the diagram
X
u /Y
S
(where the downwards arrows are the structure morphisms) is commutative. With this definition,
the formal S-preschemes (for some fixed S) form a category. We denote by HomS (X, Y) the set of
S-morphisms from a formal S-prescheme X to a formal S-prescheme Y. When S = Spf( A), we
sometimes say A-morphism instead of S-morphism.
(10.4.8). Since every affine scheme can be considered as a formal affine scheme (10.1.2), every (usual)
prescheme can be considered as a formal prescheme. In addition, it follows from Definition (10.4.5)
that, for the usual preschemes, the morphisms (resp. S-morphisms) of formal preschemes coincide
with the morphisms (resp. S-morphisms) defined in §2.
10.5. Sheaves of ideals of definition for formal preschemes
(10.5.1). Let X be a formal prescheme; we say that an OX -ideal J is a sheaf of ideals of definition (or
an ideal sheaf of definition) for X if every x ∈ X has a formal affine open neighbourhood U such that
J |U is a sheaf of ideals of definition for the formal affine scheme induced on U by X (10.3.3); by
(10.3.1.1) and Proposition (10.4.3), for each open V ⊂ X, J |V is then a sheaf of ideals of definition
for the formal prescheme induced on V by X.
We say that a family (Jλ ) of sheaves of ideals of definition for X is a fundamental system of I | 187
sheaves of ideals of definition if there exists a cover (Uα ) of X by formal affine open sets such that, for
each α, the family of the Jλ |Uα is a fundamental system of sheaves of ideals of definition (10.3.6)
for the formal affine scheme induced on Uα by X. It follows from the last remark of (10.3.7) that,
when X is a formal affine scheme, this definition coincides with the definition given in (10.3.7). For
an open subset V of X, the restrictions Jλ |V then form a fundamental system of sheaves of ideals of
definition for the formal prescheme induced on V, according to (10.3.1.1). If X is a locally Noetherian
10. FORMAL SCHEMES 179
formal prescheme, and J is a sheaf of ideals of definition for X, then it follows from Proposition
(10.3.6) that the powers J n form a fundamental system of sheaves of ideals of definition for X.
(10.5.2). Let X be a formal prescheme, and J a sheaf of ideals of definition for X. Then the ringed
space (X, OX /J ) is a (usual) prescheme, which is affine (resp. locally Noetherian, resp. Noetherian)
when X is a formal affine scheme (resp. a locally Noetherian formal scheme, resp. a Noetherian
formal scheme); we can reduce to the affine case, and then the proposition has already been proved
in (10.3.2). In addition, if θ : OX → OX /J is the canonical homomorphism, then u = (1X , θ ) is a
morphism (said to be canonical) of formal preschemes (X, OX /J ) → (X, OX ), because, again, this
was proved in the affine case (10.3.2), to which it is immediately reduced.
Proposition (10.5.3). — Let X be a formal prescheme, and (Jλ ) a fundamental system of sheaves of ideals
of definition for X. Then the sheaf of topological rings OX is the projective limit of the pseudo-discrete sheaves
of rings (0, 3.8.1) OX /Jλ .
P ROOF. Since the topology of X admits a basis of formal quasi-compact affine open sets (10.4.3),
we reduce to the affine case, where the proposition is a consequence of Proposition (10.3.5), (10.3.2),
and Definition (10.1.1). □
It is not true that any formal prescheme admits a sheaf of ideals of definition. However:
Proposition (10.5.4). — Let X be a locally Noetherian formal prescheme. There exists a largest sheaf of ideals
of definition T for X; this is the unique sheaf of ideals of definition J such that the prescheme (X, OX /J )
is reduced. If J is a sheaf of ideals of definition for X, then T is the inverse image under OX → OX /J of
the nilradical of OX /J .
P ROOF. Suppose first that X = Spf( A), where A is an adic Noetherian ring. The existence and
the properties of T follow immediately from Propositions (10.3.5) and (5.1.1), taking into account ErrII
the existence and the properties of the largest ideal of definition for A ((0, 7.1.6) and (0, 7.1.7)).
To prove the existence and the properties of T in the general case, it suffices to show that, if U ⊃
V are two Noetherian formal affine open subsets of X, then the largest sheaf of ideals of definition
TU for U induces the largest sheaf of ideals of definition TV for V; but as (V, (OX |V )/(TU |V )) is
reduced, this follows from the above. □
f′
(X, OX /J ) / (Y, OY /K )
P ROOF.
10. FORMAL SCHEMES 180
(i) If f = (ψ, θ ), then the hypotheses imply that the image under θ ♯ : ψ∗ (OY ) → OX of the
sheaf of ideals ψ∗ (K ) of ψ∗ (OY ) is contained in J (0, 4.3.5). By passing to quotients, we
thus obtain from θ ♯ a homomorphism of sheaves of rings
ω : ψ∗ (OY /K ) = ψ∗ (OY )/ψ∗ (K ) −→ OX /J ;
furthermore, since, for all x ∈ X, θ x♯ is a local homomorphism, so too is ωx . The morphism
of ringed spaces (ψ, ω ♭ ) is thus (2.2.1) the unique morphism f ′ of ringed spaces whose
existence was claimed.
(ii) The canonical functorial correspondence between morphisms of formal affine schemes
and continuous homomorphisms of rings (10.2.2) shows that, in the case considered, the
relation f ∗ (K )OX ⊂ J implies that we have f ′ = ( a ϕ′ , ϕe′ ), where ϕ′ : B/K → A/J is the
unique homomorphism making the diagram
/A
ϕ
(10.5.6.2) B
ϕ′
B/K / A/J
commutate. The existence of ϕ′ thus implies that ϕ(K) ⊂ J. Conversely, if this condition is
satisfied, then, denoting by ϕ′ the unique homomorphism making the diagram (10.5.6.2)
commutate and setting f ′ = ( a ϕ′ , ϕe′ ), it is clear that the diagram (10.5.6.1) is commu-
∗
tative; considering the homomorphisms a ϕ∗ (OY ) → OX and a ϕ′ (OY /K ) → OX /J
corresponding to f and f ′ respectively then leads to the fact that this implies the relation
f ∗ (K )OX ⊂ J .
□
It is clear that the correspondence f 7→ f ′ defined above is functorial.
10.6. Formal preschemes as inductive limits of preschemes
(10.6.1). Let X be a formal prescheme, and (Jλ ) a fundamental system of sheaves of ideals of
definition for X; for each λ, let f λ be the canonical morphism (X, OX /Jλ ) → X (10.5.2); for
Jµ ⊂ Jλ , the canonical morphism OX /Jµ → OX /Jλ defines a canonical morphism f µλ : I | 189
(X, OX /Jλ ) → (X, OX /Jµ ) of (usual) preschemes such that f λ = f µ ◦ f µλ . The preschemes
Xλ = (X, OX /Jλ ) and the morphisms f µλ thus form (by (10.4.8)) an inductive system in the category
of formal preschemes.
Proposition (10.6.2). — With the notation of (10.6.1), the formal prescheme X and the morphisms f λ form
an inductive limit (T, I, 1.8) of the system ( Xλ , f µλ ) in the category of formal preschemes.
P ROOF. Let Y be a formal prescheme, and, for each index λ, let
gλ = (ψλ , θλ ) : Xλ −→ Y
be a morphism such that gλ = gµ ◦ f µλ for Jµ ⊂ Jλ . This latter condition and the definition of the
Xλ imply first of all that the ψλ are identical to a single continuous map ψ : X → Y of the underlying
spaces; in addition, the homomorphisms θλ♯ : ψ∗ (OY ) → OXi = OX /Jλ form a projective system
of homomorphisms of sheaves of rings. By passing to the projective limit, there is an induced
homomorphism ω : ψ∗ (OY ) → lim OX /Jλ = OX , and it is clear that the morphism g = (ψ, ω ♭ ) of
←−
ringed spaces is the unique morphism making the diagrams
gλ
(10.6.2.1) Xλ /Y
?
fλ g
X
commutative. It remains to prove that g is a morphism of formal preschemes; the question is local
on X and Y, so we can assume that X = Spf( A) and Y = Spf( B), with A and B admissible rings,
and with Jλ = J∆ λ , where (Jλ ) is a fundamental system of ideals of definition for A (10.3.5);
10. FORMAL SCHEMES 181
since A = lim A/Jλ , the existence of a morphism g of formal affine schemes making the diagrams
←−
(10.6.2.1) commutate then follows from the bijective correspondence (10.2.2) between morphisms
of formal affine schemes and continuous ring homomorphisms, and from the definition of the
projective limit. But the uniqueness of g as a morphism of ringed spaces shows that it coincides with
the morphism in the beginning of the proof. □
The following proposition establishes, under certain additional conditions, the existence of
the inductive limit of a given inductive system of (usual) preschemes in the category of formal
preschemes:
Proposition (10.6.3). — Let X be a topological space, and (Oi , u ji ) a projective system of sheaves of rings on
X, with N for its set of indices. Let Ji be the kernel of u0i : Oi → O0 . Suppose that:
(a) the ringed space (X, Oi ) is a prescheme Xi ;
(b) for all x ∈ X and all i, there exists an open neighbourhood Ui of x in X such that the restriction ErrII
Ji |Ui is nilpotent; and
(c) the homomorphisms u ji are surjective.
Let OX be the sheaf of topological rings given by the projective limit of the pseudo-discrete sheaves of I | 190
rings Oi , and let ui : OX → Oi be the canonical homomorphism. Then the topologically ringed space (X, OX )
is a formal prescheme; the homomorphisms ui are surjective; their kernels J (i) form a fundamental system of
sheaves of ideals of definition for X, and J (0) is the projective limit of the sheaves of ideals Ji .
P ROOF. We first note that, on each stalk, u ji is a surjective homomorphism and a fortiori a local
homomorphism; thus vij = (1X , u ji ) is a morphism of preschemes X j → Xi (i ⩾ j) (2.2.1). Suppose
first that each Xi is an affine scheme of ring Ai . There exists a ring homomorphism ϕji : Ai → A j
such that u ji = ϕ fji (1.7.3); as a result (1.6.3), the sheaf O j is a quasi-coherent Oi -module over Xi (for
the external law defined by u ji ), associated to A j considered as an Ai -module by means of ϕji . For
all f ∈ Ai , let f ′ = ϕji ( f ); by hypothesis, the open sets D ( f ) and D ( f ′ ) are identical in X, and the
homomorphism from Γ( D ( f ), Oi ) = ( Ai ) f to Γ( D ( f ), O j ) = ( A j ) f ′ corresponding to u ji is exactly
(ϕji ) f (1.6.1). But when we consider A j as an Ai -module, ( A j ) f ′ is the ( Ai ) f -module ( A j ) f , so we
also have u ji = ϕ fji , where ϕji is now considered as a homomorphism of Ai -modules. Then, since u ji is
surjective, we conclude that ϕji is also surjective (1.3.9), and if J ji is the kernel of ϕji , then the kernel
of u ji is a quasi-coherent Oi -module equal to J fji . In particular, we have Ji = Jei , where Ji is the
kernel of ϕ0i : Ai → A0 . Hypothesis (b) implies that Ji is nilpotent: indeed, since X is quasi-compact,
we can cover X by a finite number of open sets Uk such that (Ji |Uk )nk = 0, and, by setting n to be
the largest of the nk , we have Jin = 0. We thus conclude that Ji is nilpotent (1.3.13). Then the ring
A = lim Ai is admissible (0, 7.2.2), the canonical homomorphism ϕi : A → Ai is surjective, and its
←−
kernel J(i) is equal to the projective limit of the Jik for k ⩾ i; the J(i) form a fundamental system of
neighbourhoods of 0 in A. The claims of Proposition (10.6.3) follow in this case from (10.1.1) and
(10.3.2), with (X, OX ) being Spf( A).
Again, in this particular case, we note that, if f = ( f i ) is an element of the projective limit
A = lim Ai , then all the open sets D ( f i ) (which are affine open sets in Xi ) can be identified with the
←−
open subset D( f ) of X, and the prescheme induced on D( f ) by Xi thus being identified with the
affine scheme Spec(( Ai ) f i ).
In the general case, we remark first that, for every quasi-compact open subset U of X, each of
the Ji |U is nilpotent, as shown by the above reasoning. We will show that, for every x ∈ X, there
exists an open neighbourhood U of x in X which is an affine open set for all the Xi . Indeed, we take
U to be an affine open set for X0 , and observe that OX0 = OXi /Ji . Since, by the above, Ji |U is
nilpotent, U is also an affine open set for each Xi , by Proposition (5.1.9). This being so, for each U
satisfying the preceding conditions, the study of the affine case as above shows that (U, OX |U ) is a
formal prescheme whose J (i) |U form a fundamental system of sheaves of ideals of definition, and
J (0) |U is the projective limit of the Ji |U; whence the conclusion. □
j +1
Corollary (10.6.4). — Suppose that, for i ⩾ j, the kernel of u ji is Ji and that J1 /J12 is of finite I | 191
type over O0 = O1 /J1 . Then X is an adic formal prescheme, and if J (n) is the kernel of OX → On ,
10. FORMAL SCHEMES 182
then J (n) = J n+1 and J /J 2 is isomorphic to J1 . If, in addition, X0 is locally Noetherian (resp.
Noetherian), then X is locally Noetherian (resp. Noetherian).
P ROOF. Since the underlying spaces of X and X0 are the same, the question is local, and we
can suppose that all the Xi are affine; taking into account the fact that Jij = J fji (with the notation
of Proposition (10.3.6)), we can immediately reduce the problem to the corresponding claims of
Proposition (0, 7.2.7) and Corollary (0, 7.2.8), by noting that J1 /J21 is then an A0 -module of finite
type (1.3.9). □
In particular, every locally Noetherian formal prescheme X is the inductive limit of a sequence
( Xn ) of locally Noetherian (usual) preschemes satisfying the conditions of Proposition (10.3.6) and
Corollary (10.6.4): it suffices to consider a sheaf of ideals of definition J for X (10.5.4) and to take
Xn = (X, OX /J n+1 ) ((10.5.1) and Proposition (10.6.2)).
Corollary (10.6.5). — Let A be an admissible ring. For the formal affine scheme X = Spf( A) to be
Noetherian, it is necessary and sufficient for A to be adic and Noetherian.
P ROOF. The condition is evidently sufficient. Conversely, suppose that X is Noetherian, and
let J be an ideal of definition for A, and J = J∆ the corresponding sheaf of ideals of definition
for X. The (usual) preschemes Xn = (X, OX /J n+1 ) are then affine and Noetherian, so the rings
An = A/Jn+1 are Noetherian (6.1.3), whence we conclude that J/J2 is an A/J-module of finite
type. Since the J n form a fundamental system of sheaves of ideals of definition for X (10.5.1), we
have OX = lim OX /J n (10.5.3); we thus conclude (10.1.3) that A is topologically isomorphic to
←−
lim A/Jn , which is adic and Noetherian (0, 7.2.8). □
←−
Remark (10.6.6). — With the notation of Proposition (10.6.3), let Fi be an Oi -module, and suppose
we are given, for i ⩾ i, a vij -morphism θ ji : Fi → F j such that θkj ◦ θ ji = θki for k ⩽ j ⩽ i. Since the
underlying continuous map of vij is the identity, θ ji is a homomorphism of sheaves of abelian groups
to the space X; in addition, if F is the projective limit of the projective system (Fi ) of sheaves of
abelian groups, then the fact that the θ ji are vij -morphisms lets us define an OX -module structure
on F by passing to the projective limit; when equipped with this structure, we say that F is the
projective limit (with respect to the θ ji ) of the system of Oi -modules (Fi ). In the particular case where
vij∗ (Fi ) = F j and θ ji is the identity, we say that F is the projective limit of a system (Fi ) such that
vij∗ (Fi ) = F j for j ⩽ i (without mentioning the θ ji ).
(10.6.7). Let X and Y be formal preschemes, J (resp. K ) a sheaf of ideals of definition for X (resp.
Y), and f : X → Y a morphism such that f ∗ (K )OX ⊂ J . We then have, for every integer n > 0,
that f ∗ (K n )OX = ( f ∗ (K )OX )n ⊂ J n ; f thus induces (10.5.6) a morphism of (usual) preschemes
f n : Xn → Yn by setting Xn = (X, OX /J n+1 ) and Yn = (Y , OY /K n+1 ), and it immediately
follows from the definitions that the diagrams
fm
(10.6.7.1) Xm / Ym
fn
Xn / Yn
f
X /Y
commutate.
10. FORMAL SCHEMES 183
Proposition (10.6.9). — Let X and Y be locally Noetherian formal preschemes, and J (resp. K ) be a
sheaf of ideals of definition for X (resp. Y); the map f 7→ ( f n ) defined in (10.6.7) is a bijection from the set of
morphisms f : X → Y such that f ∗ (K )OX ⊂ J to the set of sequences ( f n ) of morphisms that make the
diagrams (10.6.7.1) commutate.
P ROOF. If f is the inductive limit of this sequence, then it remains to show that f ∗ (K )OX ⊂ J .
The statement, being local on X and Y, can be reduced to the case where X = Spf( A) and Y = Spf( B)
are affine, with A and B adic Noetherian rings, and with J = J∆ and K = K∆ , where J (resp.
K) is an ideal of definition for A (resp. B). We then have that Xn = Spec( An ) and Yn = Spec( Bn ),
with An = A/Jn+1 and Bn = B/Kn+1 , by Proposition (10.3.6) and (10.3.2); f n = ( a ϕn , ϕ fn ), where
the homomorphisms ϕn : Bn → An forms a projective system, thus f = ( a ϕ, ϕ e), and so f = ( a ϕ, ϕ
e),
where ϕ = lim ϕn . The commutativity of the diagram (10.6.7.1) for m = 0 then gives the condition
←−
ϕn (K/Kn+1 ) ⊂ J/Jn+1 for all n, so, by passing to the projective limit, we have ϕ(K) ⊂ J, which
implies that f ∗ (K )OX ⊂ J (10.5.6, ii). □
Corollary (10.6.10). — Let X and Y be locally Noetherian formal preschemes, and T the largest sheaf of
ideals of definition for X (10.5.4).
(i) For every sheaf of ideals of definition K for Y and every morphism f : X → Y, we have
f ∗ (K )OX ⊂ T .
(ii) There is a canonical bijective correspondence between Hom(X, Y) and the set of sequences ( f n )
of morphisms making the diagrams (10.6.7.1) commute, where Xn = (X, OX /T n+1 ) and Yn =
(Y, OY /K n+1 ).
P ROOF. (ii) follows immediately from (i) and Proposition (10.6.9). To prove (i), we can reduce
to the case where X = Spf( A) and Y = Spf( B), with A and B Noetherian, and with T = T∆
and K = K∆ , where T is the largest ideal of definition for A and K is an ideal of definition for
B. Let f = ( a ϕ, ϕ
e), where ϕ : B → A is a continuous homomorphism; since the elements of K are
topologically nilpotent (0, 7.1.4, ii), so too are those of ϕ(K), and so ϕ(K) ⊂ T, since T is the set of
topologically nilpotent elements of A (0, 7.1.6); hence, by Proposition (10.5.6, ii), we are done. □
Corollary (10.6.11). — Let S, X, Y be locally Noetherian formal preschemes, and f : X → S and
g : Y → S the morphisms that make X and Y formal S-preschemes. Let J (resp. K , L ) be a sheaf
of ideals of definition for S (resp. X, Y), and suppose that f ∗ (J )OX ⊂ K and g∗ (J )OY = L ; set
Sn = (S, OS /J n+1 ), Xn = (X, OX /K n+1 ), and Yn = (Y, OY /L n+1 ). Then there exists a canonical
bijective correspondence between HomS (X, Y) and the set of sequences (un ) of Sn -morphisms un : Xn → Yn I | 193
making the diagrams (10.6.7.1) commute.
P ROOF. For each S-morphism u : X → Y, we have by definition that f = g ◦ u, whence
u∗ (L )OX = u∗ ( g∗ (J OY )OX = f ∗ (J )OX ⊂ K ,
and the corollary then follows from Proposition (10.6.9). □
We note that, for m ⩽ n, the data of a morphism f n : Xn → Yn determines a unique morphism
f m : Xm → Ym making the diagram (10.6.7.1) commutate, since we immediately see that we can
reduce to the affine case; we have thus defined a map ϕmn : HomSn ( Xn , Yn ) → HomSm ( Xm , Ym ),
and the HomSn ( Xn , Yn ) form, with the ϕmn , a projective system of sets; Corollary (10.6.11) then says
that there is a canonical bijection
HomS (X, Y) ≃ lim HomSn ( Xn , Yn ).
← −
n
10. FORMAL SCHEMES 184
(10.8.3). Now let F be a coherent OX -module; for all J ∈ Φ, we have that F ⊗OX (OX /J ) is a
coherent OX -module (9.1.1) with support contained in X ′ , and we will usually identify it with its
restriction to X ′ . When J varies over Φ, these sheaves form a projective system of sheaves of abelian
groups.
Definition (10.8.4). — Given a closed subset X ′ of a locally Noetherian prescheme X and a coherent
OX -module F , we define the completion of F along X ′ , denoted by F/X ′ (or Fc when there is little
′
chance of confusion), to be the restriction to X of the sheaf limΦ (F ⊗OX (OX /J )); we say that its I | 195
←−
sections over X ′ are the formal sections of F along X ′ .
It is immediate that, for every open U ⊂ X, we have (F |U )/(U ∩X ′ ) = (F/X ′ )|(U ∩ X ′ ).
By passing to the projective limit, it is clear that (OX )/X ′ is a sheaf of rings, and that F/X ′ can
be considered as an (OX )/X ′ -module. In addition, since there exists a basis for the topology of X ′
consisting of quasi-compact open sets, we can consider (OX )/X ′ (resp. F/X ′ ) as a sheaf of topological
rings (resp. of topological groups), the projective limit of the pseudo-discrete sheaves of rings (resp.
groups) OX /F (resp. F ⊗OX (OX /F ) = F /J F ), and, by passing to the projective limit, F/X ′
then becomes a topological (OX )/X ′ -module ((0, 3.8.1) and (0, 3.8.2)); recall that, for every quasi-compact
open U ⊂ X, Γ(U ∩ X ′ , (OX )/X ′ ) (resp. Γ(U ∩ X ′ , F/X ′ )) is then the projective limit of the discrete
rings (resp. groups) Γ(U, OX /J ) (resp. Γ(U, F /J F )).
Now, if u : F → G is a homomorphism of OX -modules, then there are canonically induced
homomorphisms uJ : F ⊗OX (OX /J ) → G ⊗OX (OX /J ) for all J ∈ Φ, and these homomor-
phisms form a projective system. By passing to the projective limit and restricting to X ′ , these give
a continuous (OX )/X ′ -homomorphism F/X ′ → G/X ′ , denoted u/X ′ or ub, and called the completion
of the homomorphism u along X ′ . It is clear that, if v : G → H is a second homomorphism of
OX -modules, then we have (v ◦ u)/X ′ = (v/X ′ ) ◦ (u/X ′ ), hence F/X ′ is a covariant additive functor in
F from the category of coherent OX -modules to the category of topological (OX )/X ′ -modules.
Proposition (10.8.5). — The support of (OX )/X ′ is X ′ ; the topologically ringed space ( X ′ , (OX )/X ′ ) is
a locally Noetherian formal prescheme, and, if J ∈ Φ, then J/X ′ is a sheaf of ideals of definition for this
formal prescheme. If X = Spec( A) is an affine scheme with Noetherian ring A, J = e J for some ideal J
of A, and X ′ = V (J), then ( X ′ , (OX )/X ′ ) is canonically identified with Spf( A
b), where A
b is the separated
completion of A with respect to the J-preadic topology.
P ROOF. We can evidently reduce to proving the latter claim. We know (0, 7.3.3) that the
separated completion bJ of J with respect to the J-preadic topology can be identified with the ideal
J A of A, where A is the Noetherian b
b b b J-adic ring such that A/ Jn = A/Jn (0, 7.2.6). This latter
b b
equation shows that the open prime ideals of A are the ideals b
b p = pb J, where p is a prime ideal of A
containing J, and that we have bp ∩ A = p, and hence Spf( A b) = X ′ . Since OX /J n = ( A/Jn )e, the
proposition follows immediately from the definitions. □
We say that the formal prescheme defined above is the completion of X along X ′ , and we denote it
b when there is little chance of confusion. When we take X ′ = X, we can set J = 0,
by X/X ′ or X
and we thus have X/X = X.
10. FORMAL SCHEMES 186
Corollary (10.8.6). — The (usual) prescheme X b red is the unique reduced subprescheme of X having X ′ as its
underlying space (5.2.1). For Xb to be Noetherian, it is necessary and sufficient for X
b red to be Noetherian, and
it suffices that X be Noetherian.
P ROOF. Since Xb red is determined locally (10.5.4), we can assume that X is an affine scheme
of some Noetherian ring A; with the notation of Proposition (10.8.5), the ideal T of topologically
nilpotent elements of A b is the inverse image under the canonical map A b → A/
b b J = A/J of the
nilradical of A/J (0, 7.1.3), so A/T is isomorphic to the quotient of A/J by its nilradical. The first
b
claim then follows from Propositions (10.5.4) and (5.1.1). If X b red is Noetherian, then so too is its
′ ′
underlying space X , and so the Xn = Spec(OX /J ) are Noetherian (6.1.2), and thus so too is X
n b
(10.6.4); the converse is immediate, by Proposition (6.1.2). □
(ψ, θ ) : X/X ′ −→ X
of ringed spaces.
By taking tensor products, for every coherent OX -module F , the canonical homomorphisms
OX → OX /J give homomorphisms F → F ⊗OX (OX /J ) of OX -modules which form a projective
system, and thus give, by passing to the projective limit, a canonical functorial homomorphism
γ : F → ψ∗ (F/X ′ ) of OX -modules.
Proposition (10.8.8). —
(i) The functor F/X ′ (in F ) is exact.
(ii) The functorial homomorphism γ♯ : i∗ (F ) → F/X ′ of (OX )/X ′ -modules is an isomorphism.
P ROOF.
(i) It suffices to prove that, if 0 → F ′ → F → F ′′ → 0 is an exact sequence of coherent
OX -modules, and if U is an affine open subset of X with Noetherian ring A, then the
sequence
0 −→ Γ(U ∩ X ′ , F/X
′ ′ ′ ′′
′ ) −→ Γ (U ∩ X , F/X ′ ) −→ Γ (U ∩ X , F/X ′ ) −→ 0
the separated completion of M with respect to the J-preadic topology, and similarly
Γ(U ∩ X ′ , F/X
′ c′ ′ ′′
′ ) = M , Γ (U ∩ X , F/X ′ ) = M ;
d′′
our claim then follows, since A is Noetherian, and since the functor M
b in M is exact on the
category of A-modules of finite type (0, 7.3.3). I | 197
10. FORMAL SCHEMES 187
(ii) The question is local, so we can assume that we have an exact sequence OXm → OXn → F →
0 (0, 5.3.2); since γ♯ is functorial, and the functors i∗ (F ) and F/X ′ are right exact (by (i)
and (0, 4.3.1)), we have the commutative diagram
(10.8.8.1) i∗ (OXm ) / i ∗ (O n ) / i ∗ (F ) /0
X
γ♯ γ♯ γ♯
(OXm )/X ′ / (O n )/X ′ / F/X ′ /0
X
whose rows are exact. Furthermore, the functors i∗ (F ) and F/X ′ commute with finite
direct sums ((0, 3.2.6) and (0, 4.3.2)), and we thus reduce to proving our claim for F = OX .
We have i∗ (OX ) = (OX )/X ′ = OXb (0, 4.3.4), and that γ♯ is a homomorphism of OXb -modules;
so it suffices to check that γ♯ sends the unit section of OXb over an open subset of X ′ to itself,
which is immediate, and shows, in this case, that γ♯ is the identity.
□
Corollary (10.8.10). — If F and G are coherent OX -modules, then there exist canonical functorial (in F
and G ) isomorphisms
(10.8.10.1) (F/X ′ ) ⊗(OX )/X′ (G/X ′ ) ≃ (F ⊗OX G )/X ′ ,
P ROOF. This follows from the canonical identification of i∗ (F ) with F/X ′ ; the existence of the
first isomorphism is then a result which holds for all morphisms of ringed spaces (0, 4.3.3.1), and the
second is a result which holds for all flat morphisms (0, 6.7.6), by Corollary (10.8.9). □
iX iY
f
X /Y
commutes, with the vertical arrows being the canonical morphisms (10.8.7).
(10.9.3). Let Z be a third prescheme, g : Y → Z a morphism, and Z ′ a closed subset of Z such that
g(Y ′ ) ⊂ Z ′ . If gb denotes the completion of the morphism g along Y ′ and Z ′ , then it immediately
follows from (10.9.1) that we have ( g ◦ f )∧ = gb ◦ fb.
Proposition (10.9.4). — Let X and Y be locally Noetherian S-preschemes, with Y of finite type over S. Let f
and g be S-morphisms from X to Y such that f ( X ′ ) ⊂ Y ′ and g( X ′ ) ⊂ Y ′ . For fb = gb to hold, it is necessary
and sufficient for f and g to coincide on a neighbourhood of X ′ .
P ROOF. The condition is evidently sufficient (even without the finiteness hypothesis on Y).
To see that it is necessary, we remark first that the hypothesis fb = gb implies that f ( x ) = g( x )
for all x ∈ X ′ . Also, since the questions is local, we can assume that X and Y are affine open
neighbourhoods of x and y = f ( x ) = g( x ) respectively (with Noetherian rings), that S is affine,
10. FORMAL SCHEMES 189
and that Γ(Y, OY ) is a Γ(S, OS )-algebra of finite type (6.3.3). Then f and g correspond to Γ(S, OS )-
homomorphisms ρ and σ (respectively) from Γ(Y, OY ) to Γ( X, OX ) (1.7.3), and, by hypothesis, the
extensions by continuity of these homomorphisms to the separated completion of Γ(Y, OY ) are the
same. We conclude from Proposition (10.8.11) that, for every section s ∈ Γ(Y, OY ), the sections
ρ(s) and σ (s) coincide on a neighbourhood (depending on s) of X ′ ; since Γ(Y, OY ) is an algebra of
finite type over Γ(S, OS ), we have that there exists a neighbourhood V of X ′ such that ρ(s) and σ (s)
coincide on V for every section s ∈ Γ(Y, OY ). If h ∈ Γ( X, OX ) is such that D (h) is a neighbourhood
of x contained in V, then we conclude from the above and from Theorem (1.4.1, d) that f and g
coincide on D (h). □
Proposition (10.9.5). — Under the hypotheses of (10.9.1), for every coherent OY -module G , there exists a
canonical functorial isomorphism of (OX )/X ′ -modules
P ROOF. If we canonically identify ( f ∗ (G ))/X ′ with i∗X ( f ∗ (G )), and fb∗ (G/Y ′ ) with fb∗ (iY∗ (G ))
(10.8.8), then the proposition follows immediately from the commutativity of the diagram in (10.9.2).
□
P ROOF. It is immediate that the question is local for S, X, and Y, and we thus reduce to the case
where S = Spec( A), X = Spec( B), Y = Spec(C ), S′ = V (J), X ′ = V (K), and Y ′ = V (L), with J, K,
and L ideals such that ϕ(J) ⊂ K and ψ(J) ⊂ L, where we denote by ϕ and ψ the homomorphisms
A → B and A → C which correspond to g and h (respectively). We know that Z = Spec( B ⊗ A C ) and
that Z ′ = V (M), where M is the ideal Im(K ⊗ A C ) + Im( B ⊗ A L). The conclusion follows (10.7.2)
from the fact that the completed tensor product ( B b⊗bC b)∧ (where A,b B,
b and C
b are, respectively,
A
the separated completions of A, B, and C with respect to the J-, K-, and L-preadic topologies) is
the separated completion of the tensor product B ⊗ A C with respect to the M-preadic topology
(0, 7.7.2). □
Corollary (10.9.8). — Let X and Y be locally Noetherian S-preschemes such that X ×S Y is locally
Noetherian; let S′ be a closed subset of S, and X ′ (resp. Y ′ ) a closed subset of X (resp. Y) whose image in S is
contained in S′ . For every S-morphism f : X → Y such that f ( X ′ ) ⊂ Y ′ , the graph morphism Γ fb can be
identified with the extension (Γ f )∧ of the graph morphism of f .
10. FORMAL SCHEMES 190
Remark (10.9.10). — If S is the sum X1 ⊔ X2 (3.1), X ′ the union X1′ ∪ X2′ , where Xi′ is a closed subset I | 201
of Xi (i = 1, 2), then we have X/X ′ = X1 /X ′ ⊔ X2 /X ′ .
1 2
Proposition (10.10.2). —
(i) M∆ is an exact functor in M, and there exists a canonical functorial isomorphism of A-modules
Γ(X, M∆ ) ≃ M.
(ii) If M and N are A-modules of finite type, then there exist canonical functorial isomorphisms
(10.10.2.1) ( M ⊗ A N )∆ ≃ M∆ ⊗OX N ∆ ,
P ROOF. The exactness of M∆ follows from the exactness of the functors M e (1.3.5) and F/X ′
∆
(10.8.8). By definition, Γ( X, M ) is the separated completion of the A-module Γ( X, M e ) = M with
respect to the J-preadic topology; but, since A is complete and M is of finite type, we know (0, 7.3.6)
that M is separated and complete, which proves (i). The isomorphism (10.10.2.1) (resp. (10.10.2.2))
comes from the composition of the isomorphisms (1.3.12, i) and (10.8.10.1) (resp. (1.3.12, ii) and
(10.8.10.2)). Finally, since Hom A ( M, N ) is an A-module of finite type, we can apply (i), which
identifies Γ(X, (Hom A ( M, N ))∆ ) with Hom A ( M, N ), and we can use (10.10.2.2), which proves that
the homomorphism (10.10.2.3) is an isomorphism. □
We deduce from Proposition (10.10.2) a series of results analogous to those of Theorem (1.3.7)
and Corollary (1.3.12), whose formulation we leave to the reader.
We note that the exactness property of M∆ , applied to the exact sequence 0 → J → A → A/J → I | 202
0, shows that the sheaf of ideals of OX denoted here by J∆ coincides with the one denoted also by
J∆ in (10.3.1), by (10.3.2).
Proposition (10.10.3). — Under the hypotheses of (10.10.1), OX is a coherent sheaf of rings.
10. FORMAL SCHEMES 191
P ROOF. If f ∈ A, then we know that A{ f } is an adic Noetherian ring (0, 7.6.11), and, since the
question is local, we reduce (10.1.4) to proving that the kernel of the homomorphism v : OX n →O
X
∆
is an OX -module of finite type. We then have v = u , where u is an A-homomorphism An →
A (10.10.2); since A is Noetherian, the kernel of u is of finite type, or, equivalently, we have a
w w u
homomorphism Am − → An such that the sequence Am − → An − → A is exact. We conclude (10.10.2)
w∆ v
that the sequence OX
m −→ O n −
X → OX is exact, which proves that the kernel of v is of finite type. □
(10.10.4). With the above notation, set An = A/Jn+1 , and let Sn be the affine scheme Spec( An ) =
(X, OX /J n+1 ), with J = J∆ the sheaf of ideals of definition for OX corresponding to the ideal J.
Let umn be the morphism of preschemes Xm → Xn corresponding to the canonical homomorphism
An → Am for m ⩽ n; the formal scheme X is the inductive limit of the Xn with respect to the umn
(10.6.3).
Proposition (10.10.5). — Under the hypothesis of (10.10.1), let F be an OX -module. The following
conditions are equivalent:
(a) F is a coherent OX -module;
(b) F is isomorphic to the projective limit (10.6.6) of a sequence (Fn ) of coherent OXn -modules such
that u∗mn (Fn ) = Fm ; and
(c) there exists an A-module M of finite type (determined up to canonical isomorphism by Proposition
(10.10.2, i)) such that F is isomorphic to M∆ .
P ROOF. We first show that (b) implies (c). We have Fn = g Mn , where Mn is an An -module
of finite type, and the hypotheses imply that Mm = Mn ⊗ An Am for m ⩽ n (1.6.5); the Mn thus
form a projective system for the canonical di-homomorphisms Mn → Mm (m ⩽ n), and it follows
immediately from the definition of the An that this projective system satisfies the conditions of
(0, 7.2.9); as a result, its projective limit M is an A-module of finite type such that Mn = M ⊗ A An for
all n. We deduce that Fn is induced over Xn by M e ⊗O (O X /e
X
Jn+1 ), and so F = M∆ by Definition
(10.8.4).
Conversely, (c) implies (b); indeed, if un is the immersion morphism Xn → X, then u∗n ( M e) =
( M ⊗ A An )e is induced over Xn by M e ⊗O (O X /e J n + 1 ∆ ∗
), and M = lim un ( M e ) by Definition (10.8.4);
X ←−
∗
since um = un ◦ umn for m ⩽ n, the Fn = un ( M ) satisfy the conditions of (b), whence our claim.
e
We now show that (c) implies (a): indeed, we have, by definition, that OX = A∆ ; since M is
the cokernel of a homomorphism Am → An , it follows from Proposition (10.10.2) that M∆ is the
cokernel of a homomorphism OX m → O n , and, since the sheaf of rings O is coherent (10.10.3), so
X X
too is M∆ (0, 5.3.4).
Finally, (a) implies (b). Considered as an OX -module, we have that OXn = OX /J n+1 = A∆ n ; I | 203
but Fn = F ⊗OX OXn is a coherent OX -module (0, 5.3.5), and, since it is also an OXn -module, and
J n+1 is coherent, we conclude that Fn is a coherent OXn -module (0, 5.3.10), and it is immediate
that u∗mn (Fn ) = Fm for m ⩽ n (recalling that the continuous map Xm → Xn of the underlying
spaces is the identity on X). The sheaf G = lim Fn is thus a coherent OX -module, since we have
←−
seen that (b) implies (a). The canonical homomorphisms F → Fn form a projective system, which,
by passing to the limit, gives a canonical homomorphism w : F → G , and it remains only to
prove that w is bijective. The question is now local, so we can reduce to the case where F is the
p q
cokernel of a homomorphism OX → OX ; since this homomorphism is of the form v∆ , where v is
a homomorphism Am → An (10.10.2), F is isomorphic to M∆ , where M = Coker v (10.10.2). We
then have, by Proposition (10.10.2), that Fn = M∆ ⊗OX A∆ ∆
n = ( M ⊗ A An ) , and, since the J-adic
∆
topology on M ⊗ A An is discrete, we have ( M ⊗ A An ) = ( M ⊗ A An )e (as an OXn -module); we
have seen above that M∆ = lim Fn , and w is thus the identity in this case. □
←−
Corollary (10.10.6). — If F satisfies condition (b) of Proposition (10.10.5), then the projective system (Fn )
is isomorphic to the system of the F ⊗OX OXn .
(10.10.7). Now let A and B be adic Noetherian rings, and ϕ : B → A a continuous homomorphism;
we denote by J (resp. K) an ideal of definition for A (resp. B) such that ϕ(K) ⊂ J, and we set
X = Spec( A), Y = Spec( B), X = Spf( A), and Y = Spf( B). Let f : X → Y be the morphism of
10. FORMAL SCHEMES 192
Proposition (10.10.8). — For every B-module N of finite type, there exists a canonical functorial isomor-
phism of OX -modules
fb∗ ( N ∆ ) ≃ ( N ⊗ B A)∆ .
(1.6.5); the proposition then follows from the commutativity of the diagram in (10.9.2). □
P ROOF. Let j be the canonical injection b → B, to which corresponds the canonical injection j∆ :
b∆ → OY of sheaves of OY -modules; by definition, fb∗ (b∆ )OX is the image of the homomorphism
fb∗ ( j∆ ) : fb∗ (b∆ ) → OX = fb∗ (OY ); but this homomorphism can be identified with ( j ⊗ 1)∆ :
(b ⊗ B A)∆ → OX = ( B ⊗ B A)∆ by Proposition (10.10.8). Since the image of j ⊗ 1 is the ideal bA of
A, the image of ( j ⊗ 1)∆ is thus (bA)∆ , by Proposition (10.10.2), whence the conclusion. □
Proposition (10.11.1). — If X is a locally Noetherian formal prescheme, then the sheaf of rings OX is I | 204
coherent, and every sheaf of ideals of definition for X is coherent.
P ROOF. The question is local, so we can reduce to the case of a Noetherian affine formal scheme,
and the proposition then follows from Propositions (10.10.3) and (10.10.5). □
(10.11.2). Let X be a locally Noetherian formal prescheme, J a sheaf of ideals of definition for X,
and Xn the locally Noetherian (usual) prescheme (X, OX /J n+1 ), so that X is the inductive limit
of the sequence ( Xn ) with respect to the canonical morphisms umn : Xm → Xn (10.6.3). With this
notation:
P ROOF. The question is local, so we can reduce to the case where X is a Noetherian affine formal
scheme, and the theorem then is a consequence of Proposition (10.10.5) and Corollary (10.10.6). □
We can thus say that the data of a coherent OX -module is equivalent to the data of a projective system
(Fn ) of coherent OXn -modules such that umn (Fn ) = Fm for m ⩽ n.
Corollary (10.11.4). — If F and G are coherent OX -modules, then we can (with the notation of Theorem
(10.11.3)) define a canonical functorial isomorphism
(10.11.4.1) HomOX (F , G ) ≃ lim HomOXn (Fn , Gn ).
← −
n
P ROOF. The projective limit on the right-hand side is understood to be taken with respect to
the maps θn 7→ u∗mn (θn ) (m ⩽ n) from HomOXn (Fn , Gn ) to HomOXm (Fm , Fm ). The homomorphism
(10.11.4.1) sends an element θ ∈ HomOX (F , G ) to the sequence (u∗n (θ )); we see that we can define an
inverse homomorphism of the above by sending a projective system (θn ) ∈ limn HomOXn (Fn , Gn )
←−
to its projective limit in HomOX (F , G ), taking into account Theorem (10.11.3). □
P ROOF. The question is local, so we reduce to the case where X = Spf( A) with A an adic
Noetherian ring, F = M∆ , G = N ∆ , and θ = u∆ , where M and N are A-modules of finite type,
and u is a homomorphism M → N; we then have that θ0 = ue0 , where u0 is the homomorphism
u ⊗ 1 : M ⊗ A A/J → N ⊗ A A/J; the conclusion follows from the fact θ and u (resp. θ0 and u0 )
are simultaneously surjective ((1.3.9) and (10.10.2)) and that u and u0 are simultaneously surjective
(0, 7.1.14). □
(10.11.6). Theorem (10.11.3) shows that we can consider every coherent OX -module F as a topological
OX -module, considering it as a projective limit of pseudo-discrete sheaves of groups Fn (0, 3.8.1).
It then follows from Corollary (10.11.4) that every homomorphism u : F → G of coherent OX -
modules is automatically continuous (0, 3.8.2). Furthermore, if H is a coherent OX -submodule of a I | 205
coherent OX -module F , then, for every open U ⊂ X, Γ(U, H ) is a closed subgroup of the topological
group Γ(U, F ), since the functor Γ is left exact, and Γ(U, H ) is the kernel of the homomorphism
Γ(U, F ) → Γ(U, F /H ), which is continuous by the above, since F /G is coherent (0, 5.3.4); our
claim follows from the fact that Γ(U, F /H ) is a separated topological group.
Proposition (10.11.7). — Let F and G be coherent OX -modules. We can define (with the notation of
Theorem (10.11.3)) canonical functorial isomorphisms of topological OX -modules (10.11.6)
(10.11.7.1) F ⊗OX G ≃ lim(Fn ⊗OXn Gn ),
← −
n
P ROOF. The existence of the isomorphism (10.11.7.1) follows from the formula
Fn ⊗OXn Gn = (F ⊗OX OXn ) ⊗OXn (G ⊗OX OXn ) = (F ⊗OX G ) ⊗OX OXn
and from Theorem (10.11.3). The isomorphism (10.11.7.2), where both sides are considered as
sheaves of modules without topology, follows from the definition of the sections of H omOX (F , G )
and H omOXn (Fn , Gn ), and from the existence of the isomorphism (10.11.4.1), mapping a prescheme
induced on an arbitrary Noetherian formal affine open set to X. It remains to prove that the
isomorphism (10.11.7.2) is bicontinuous over a quasi-compact set, and we can thus reduce to the
case where X = Spf( A) with A an adic Noetherian ring, and hence (10.10.5) to the case where
F = M∆ and G = N ∆ , with M and N both A-modules of finite type; taking (10.10.2.1), (10.10.2.3),
and Corollary (1.3.12, ii) into account, we reduce to showing that the canonical isomorphism
Hom A ( M, N ) ≃ limn Hom An ( Mn , Nn ) (with Mn = M ⊗ A An and Nn = N ⊗ A An ) is continuous,
←−
which has already been proved in (0, 7.8.2). □
(10.11.8). Since HomOX (F , G ) is the group of sections of the sheaf of topological groups H omOX (F , G ),
it is equipped with a group topology. If X is Noetherian, then it follows from (10.11.7.2) that the
subgroups HomOX (F , J n G ) (for arbitrary n) form a fundamental system of neighbourhoods of 0
in this group.
Proposition (10.11.9). — Let X be a Noetherian formal prescheme, and F and G coherent OX -modules. In
the topological group HomOX (F , G ), the surjective (resp. injective, bijective) homomorphisms form an open
set.
P ROOF. By Corollary (10.11.5), the set of surjective homomorphisms in HomOX (F , G ) is the
inverse image under the continuous map HomOX (F , G ) → HomOX (F0 , G0 ) of a subset of the
0
discrete group HomOX (F0 , G0 ), whence the first claim. To show the second, we cover X by a finite
0
number of Noetherian formal affine subsets Ui . For θ ∈ HomOX (F , G ) to be injective, it is necessary
and sufficient for all of the images under the (continuous) restriction maps HomOX (F , G ) →
HomOX |Ui ) (F |Ui , G |Ui ) to be injective; we can thus reduce to the affine case, and then this has
already been proved in (0, 7.8.3). □
10. FORMAL SCHEMES 194
since A1 is Noetherian, J1 is of finite type over A1 , and so K1 = K1 /K21 is of finite type over B1 , and
a fortiori of finite type over B0 = B1 /K1 ; the fact that X is Noetherian then follows from (10.6.4);
if B = lim Bn , then we have X = Spf( B), and, if K is the kernel of B → B0 , then Bn = B/Kn+1 . If
←−
ρn : A/Jn+1 → B/Kn+1 is the homomorphism corresponding to f n , then we have that
K/Kn+1 = ( B/Kn+1 )ρn (J/Jn+1 )
since the homomorphism ρ : A → B corresponding to f is equal to lim ρn , and that the ideal JB of B
←−
is dense in K, and, since every ideal of B is closed (0, 7.3.5), we also have that K = JB. If K = K∆ ,
the equality f ∗ (J )OX = K then follows from (10.10.9), and finishes the proof. □
(10.12.3.2). The above equivalence gives, for adic S-preschemes X and Y, a canonical bijection
HomS (X, Y) ≃ lim HomSn ( Xn , Yn )
← −
n
where the projective limit is relative to the maps un → (un )(Sm ) for m ⩽ n.
We note that, if the conditions of Proposition (10.13.1) are satisfied, then property (a) holds true
for any ideal of definition K of Y (by (c)), and so, in property (b), all the f n are morphisms of finite
type.
Corollary (10.13.2). — If the conditions of (10.13.1) are satisfied, then every Noetherian formal affine open
subset V of Y has property (Q), and, if Y is Noetherian, then so too is X.
P ROOF. This follows immediately from (10.13.1) and (6.3.2). □
Definition (10.13.3). — When the equivalent properties (a), (b), and (c) of (10.13.1) are satisfied, we
say that the morphism f is of finite type, or that X is a formal Y-prescheme of finite type, or a formal
prescheme of finite type over Y.
10. FORMAL SCHEMES 196
Corollary (10.13.4). — Let X = Spf( A) and Y = Spf( B) be Noetherian formal affine schemes; for X to
be of finite type over Y, it is necessary and sufficient for the Noetherian adic ring A to be isomorphic to the
quotient of a formal series algebra, restricted to B, by some closed ideal.
P ROOF. With the notation of (10.13.1), if X is of finite type over Y, then A/KA is a ( B/K)-
algebra of finite type by (6.3.3), and KA is an ideal of definition of A (10.10.9). We are then done, by
(0, 7.5.5). □
Proposition (10.13.5). —
(i) The composition of any two morphisms (of formal preschemes) of finite type is again of finite type. I | 209
(ii) Let X, S, and S′ be locally Noetherian (resp. Noetherian) formal preschemes, and f : X → S and
X → S′ morphisms. If f is of finite type, then X ×S S′ is locally Noetherian (resp. Noetherian)
and of finite type over S′ .
(iii) Let S be a locally Noetherian formal prescheme, and X′ and Y′ formal S-preschemes such that
X′ ×S Y′ is locally Noetherian. If X and Y are locally Noetherian formal S-preschemes, and
f : X → X′ and g : Y → Y′ are S-morphisms of finite type, then X ×S Y is locally Noetherian,
and f ×S g is a S-morphism of finite type.
P ROOF. By the formal argument of (3.5.1), (iii) follows from (i) and (ii), so it suffices to prove (i)
and (ii).
Let X, Y, and Z be locally Noetherian formal preschemes, and f : X → Y and g : Y → Z
morphisms of finite type. If L is an ideal of definition of Z, then K = g∗ (L )OY is an ideal of
definition of Y, and J = f ∗ ( g∗ (L ))OX is an ideal of definition for X. Let X0 = (X, OX /J ),
Y0 = (Y, OY /K ), and Z0 = (Z, OZ /L ), and let f 0 : X0 → Y0 and g0 : Y0 → Z0 be the morphisms
corresponding to f and g (respectively). Since, by hypothesis, f 0 and g0 are of finite type, so too is
g0 ◦ f 0 (6.3.4), which corresponds to g ◦ f ; thus g ◦ f is of finite type, by (10.13.1).
Under the conditions of (ii), S (resp. X, S′ ) is the inductive limit of a sequence (Sn ) (resp. ( Xn ),
′
(Sn )) of locally Noetherian preschemes, and we can assume (10.13.1) that Xm = Xn ×Sn Sm for m ⩽ n.
The formal prescheme X ×S S′ is then the inductive limit of the preschemes Xn ×Sn Sn′ (10.7.4), and
we have that
′ ′
Xm × Sm Sm = ( Xn × Sn Sm ) × Sm Sm = ( Xn ×Sn Sn′ ) ×Sn′ Sm
′
.
Furthermore, X0 ×S0 S0′ is locally Noetherian, since X0 is of finite type over S0 (6.3.8). We thus
conclude (10.12.3.1), first of all, that X ×S S′ is locally Noetherian; then, since X0 ×S0 S0′ is of finite
type over S0′ (6.3.8), it follows from (10.12.3.1) and (10.13.1) that X ×S S′ is of finite type over S′ ,
which proves (ii) (the claim about Noetherian preschemes being an immediate consequence of
(6.3.8)). □
Corollary (10.13.6). — Under the hypotheses of (10.9.9), if f is a morphism of finite type, then so too is its
extension fb to the completions.
Proposition (10.14.1). — Let X be a locally Noetherian formal prescheme, and A a coherent sheaf of ideals
of OX . If Y if the (closed) support of OX /A , then the topologically ringed space (Y, (OX /A )|Y) is a locally
Noetherian formal prescheme that is Noetherian if X is.
P ROOF. Note that OX /A is coherent by (10.10.3) and (0, 5.3.4), so its support Y is closed (0, 5.2.2).
Let J be an ideal of definition of X, and let Xn = (X/OX /J n+1 ); the sheaf of rings OX /A is
the projective limit of the sheaves OX /(A + J n+1 ) = (OX /A ) ⊗OX (OX /J n+1 ) (10.11.3), all of
which have support Y. The sheaf (A + J n+1 )/J n+1 is a coherent OX -module, since J n+1 is
coherent, and so (A + J n+1 )/J n+1 is also a coherent (OX /J n+1 )-module (0, 5.3.10); if Yn is the
closed subprescheme of Xn defined by this sheaf of ideals, it is immediate that (Y, (OX /A )|Y) is I | 210
the formal prescheme given by the inductive limit of the Yn , and, since the conditions of (10.6.4) are
satisfied, this proves that this formal prescheme is locally Noetherian, and further Noetherian if X is
(since then Y0 is, by (6.1.4)). □
10. FORMAL SCHEMES 197
Proposition (10.14.5). —
(i) If f : Z → Y and g : Y → X are closed immersions of locally Noetherian formal preschemes, then
g ◦ f is a closed immersion. I | 211
(ii) Let X, Y, and S be locally Noetherian formal preschemes, f : X → S a closed immersion, and
g : Y → S a morphism. Then the morphism X ×S Y → Y is a closed immersion.
(iii) Let S be a locally Noetherian formal prescheme, and X′ and Y′ locally Noetherian formal S-
preschemes such that X′ ×S Y′ is locally Noetherian. If X and Y are locally Noetherian S-
preschemes, and f : X → X′ and g : Y → Y′ are S-morphisms that are closed immersions, then
f ×S g is a closed immersion.
P ROOF. By (3.5.1), it again suffices to prove (i) and (ii).
To prove (i), we can assume that Y (resp. Z) is a closed subprescheme of X (resp. Y) defined by
a coherent sheaf J (resp. K ) of ideals of OX (resp. OY ); if ψ is the injection Y → X of underlying
spaces, then ψ∗ (K ) is a coherent sheaf of ideals of ψ∗ (OY ) = OX /J (0, 5.3.12), and thus also
a coherent OX -module (0, 5.3.10); the kernel K1 of OX → (OX /J )/ψ∗ (K ) is thus a coherent
sheaf of ideals of OX (0, 5.3.4), and OX /K1 is isomorphic to ψ∗ (OY /K ), which proves that Z is an
isomorphism to a closed subprescheme of X.
10. FORMAL SCHEMES 198
To prove (ii), we can immediately restrict to the case where S = Spf( A), X = Spf( B), and
Y = Spf(C ), with A a Noetherian J-adic ring, B = A/a (where a is an ideal of A), and C a
Noetherian topological adic A-algebra. Everything then reduces to proving that the homomorphism
C → C⊗ b A ( A/a) is surjective: but A/a is an A-module of finite type, and its topology is the J-adic
topology; it then follows from (0, 7.7.8) that C ⊗
b A ( A/a) can be identified with C ⊗ A ( A/a) = C/aC,
whence our claim. □
Corollary (10.14.6). — Under the hypotheses of (10.14.5, ii), let p : X ×S Y → X and q : X ×S Y → Y
be the projections, so that the diagram
p
Xo X ×S Y
f q
g
So Y
commutes. For every coherent OX module F , we then have a canonical isomorphism of OY -modules
(10.14.6.1) u : g ∗ f ∗ (F ) ≃ q ∗ p ∗ (F ).
P ROOF. We know that defining a homomorphism g∗ f ∗ (F ) → q∗ p∗ (F ) is equivalent to defining
a homomorphism f ∗ (F ) → g∗ q∗ p∗ (F ) = f ∗ p∗ p∗ (F ) (0, 4.4.3): we take u = f ∗ (ρ), where ρ is the
canonical homomorphism F → p∗ p∗ (F ) (0, 4.4.3). To see that u is an isomorphism, we can
immediately restrict to the case where S, X, and Y are formal spectra of Noetherian adic rings
A, B, and C (respectively), satisfying the conditions in (10.14.5, ii) above; we then have F = M∆ ,
where M is an ( A/a)-module of finite type (10.10.5), and the two sides of (10.14.6.1) can then be
identified, respectively, by (10.10.8), with (C ⊗ A M )∆ and ((C/aC ) ⊗ A/a M)∆ , whence the corollary,
since (C/aC ) ⊗ A/a M = (C ⊗ A ( A/a)) ⊗ A/a M is canonically identified with C ⊗ A M. □
I | 212
Corollary (10.14.7). — Let X be a locally Noetherian usual prescheme, Y a closed subprescheme of X, j
the canonical injection Y → X, X ′ a closed subset of X, and Y ′ = Y ∩ X ′ ; then bj : Y/Y ′ → X/X ′ is a closed
immersion, and, for every coherent OY -module F , we have
bj∗ (F/Y ′ ) = ( j∗ (F ))/X ′ .
P ROOF. Since Y ′ = j−1 ( X ′ ), it suffices to use (10.9.9) and to apply (10.14.5) and (10.14.6). □
10.15. Separated formal preschemes
Definition (10.15.1). — Let S be a formal prescheme, X a formal S-prescheme, and f : X → S the
structure morphism. We define the diagonal morphism ∆X|S : X → X ×S X (also denoted by ∆X )
to be the morphism (1X , 1X )S . We say that X is separated over S, or is a formal S-scheme, or that
f is a separated morphism, if the image of the underlying space of X under ∆X is a closed subset
of the underlying space of X ×S X. We say that a formal prescheme X is separated, or is a formal
scheme, if it is separated over Z.
Proposition (10.15.2). — Suppose that the formal preschemes S and X are inductive limits of sequences
(Sn ) and ( Xn ) (respectively) of usual preschemes, and that the morphism f : X → S is the inductive limit of
a sequence of morphisms f n : Xn → Sn . For f to be separated, it is necessary and sufficient for the morphism
f 0 : X0 → S0 to be separated.
P ROOF. Indeed, ∆X|S is then the inductive limit of the sequence of morphisms ∆ Xn |Sn (10.7.4),
and the image of the underlying space of X (resp. of X ×S X under ∆X|S ) is identical to the image
of the underlying space of X0 (resp. of X0 ×S0 X0 ) under ∆ X0 |S0 ; whence the conclusion. □
Proposition (10.15.3). — Suppose that all the formal preschemes (resp. morphisms of formal preschemes) in
what follows are inductive limits of sequences of usual preschemes (resp. of morphisms of usual preschemes).
(i) The composition of any two separated morphisms is separated.
(ii) If f : X → X′ and g : Y → Y′ are separated S-morphisms, then f ×S g is separated.
(iii) If f : X → Y is a separated S-morphism, then the S′ -morphism f (S′ ) is separated for every
extension S′ → S of the base formal prescheme.
10. FORMAL SCHEMES 199
S UMMARY
The various classes of morphisms studied in this chapter are used extensively in cohomological II | 5
methods; further study using these methods will be done in Chapter III, where we make particular
use of §§2, 4, and 5 of Chapter II. On a first reading, §8 can be omitted: it supplements the formalism
developed in §§1 and 3, reducing to easy applications of this formalism, and we will use it less
consistently than the other results of this chapter.
(1.1.2.1) X
h /Y
g
f
S.
We have by definition h = (ψ, θ ), where θ : OY → h∗ (OX ) = ψ∗ (OX ) is a homomorphism
of sheaves of rings; we thus obtain (0, 4.2.2) a homomorphism of OS -algebras g∗ (θ ) : g∗ (OY ) →
g∗ (h∗ (OX )) = f ∗ (OX ), in other words, a homomorphism of OS -algebras A (Y ) → A ( X ), which
we denote by A (h). If h′ : Y → Z is a second S-morphism, then it is immediate that A (h′ ◦ h) =
A (h) ◦ A (h′ ). We have thus defined a contravariant functor A ( X ) from the category of S-preschemes
to the category of OS -algebras.
Now let F be an OX -module, G an OY -module, and u : G → F an h-morphism, that is (0, 4.4.1)
a homomorphism of OY -modules G → h∗ (F ). Then g∗ (u) : g∗ (G ) → g∗ (h∗ (F )) = f ∗ (F ) is a
200
1. AFFINE MORPHISMS 201
/ Γ(Y, OY )
ϕ
B
/ Γ( g−1 (U ), OY )
ϕU
Bµ
1. AFFINE MORPHISMS 202
(1.3.6). Let X be a prescheme affine over S. To define a prescheme Y affine over X, it is equivalent,
by Corollary (1.3.5), to give a prescheme Y affine over S, and an S-morphism g : Y → X; in other
words (Proposition (1.3.1) and (1.2.7)), it is equivalent to give a quasi-coherent OS -algebra B and
a homomorphism A ( X ) → B of OS -algebras (which can be considered as defining on B an
A ( X )-algebra structure). If f : X → S is the structure morphism, then we have B = f ∗ ( g∗ (OY )).
Corollary (1.3.7). — Let X be a prescheme affine over S; for X to be of finite type over S, it is necessary and
sufficient for the quasi-coherent OS -algebra A ( X ) to be of finite type (I, 9.6.2).
P ROOF. By definition (I, 9.6.2), we can reduce to the case where S is affine; then X is an affine
scheme (1.3.4), and if S = Spec( A), X = Spec( B), then A ( X ) is the OS -algebra B;
e as Γ(U, B
e) = B,
the corollary follows from (I, 9.6.2) and (I, 6.3.3). □
Corollary (1.3.8). — Let X be a prescheme affine over S; for X to be reduced, it is necessary and sufficient
for the quasi-coherent OX -algebra A ( X ) to be reduced (0, 4.1.4).
P ROOF. The question is local on S; by Corollary (1.3.2), the corollary follows from (I, 5.1.1) and
(I, 5.1.4). □
1.4. Quasi-coherent sheaves over an affine prescheme over S
Proposition (1.4.1). — Let X be a prescheme affine over S, Y an S-prescheme, and F (resp. G ) a quasi-
coherent OX -module (resp. an OY -module). Then the map (h, u) 7→ (A (h), A (u)) from the set of morphism
(Y, G ) → ( X, F ) to the set of di-homomorphisms (A ( X ), A (F )) → (A (Y ), A (G )) ((1.1.2) and (1.1.3))
is bijective.
P ROOF. The proof follows exactly as that of Proposition (1.2.7) by using (I, 2.2.5) and (I, 2.2.4),
and the details are left to the reader. □
Corollary (1.4.2). — If, in addition to the hypotheses of Proposition (1.4.1), we suppose that Y is affine over S,
then for (h, u) to be an isomorphism, it is necessary and sufficient for (A (h), A (u)) to be a di-isomorphism.
Proposition (1.4.3). — For every pair (B, M ) consisting of a quasi-coherent OS -algebra B and a quasi- II | 10
coherent B-module M (considered as an OS -module or as a B-module, which are equivalent (I, 9.6.1)),
there exists a pair ( X, F ) consisting of a prescheme X affine over S and of a quasi-coherent OX -module F ,
such that A ( X ) = B and A (F ) = M ; in addition, this couple is determined up to unique isomorphism.
P ROOF. The uniqueness follows from Proposition (1.4.1) and Corollary (1.4.2); the existence is
proved as in Proposition (1.3.1), and we leave the details to the reader. □
(I, 9.1.3), and the corollary follows from the canonical identification of the OS -modules ( M ⊗ B M′ )e
and M e ⊗e M f′ (where M, M′ , and B are considered as C-modules) ((I, 1.3.12) and (I, 1.6.3)). □
B
If we apply Corollary (1.4.7) in particular to the case where X = Y and F ′ = OX ′ , then we see
that the A ′ -module A ( f ′∗ (F )) identifies with A (F ) ⊗B A ′ .
(1.4.8). In particular, when X = X ′ = Y (X being affine over S), we see that if F and G are two
quasi-coherent OX -modules, then we have
(1.4.8.1) A (F ⊗O X G ) = A (F ) ⊗A ( X ) A (G )
up to canonical functorial isomorphism. If in addition F admits a finite presentation, then it follows
from (I, 1.6.3) and (I, 1.3.12) that
(1.4.8.2) A (H omX (F , G )) = H omA (X ) (A (F ), A (G ))
up to canonical isomorphism.
Remark (1.4.9). — If X and X ′ are two preschemes affine over S, then the sum X ⊔ X ′ is also affine
over S, as the sum of two affine schemes is an affine scheme.
Proposition (1.4.10). — Let S be a prescheme, B a quasi-coherent OS -algebra, and X = Spec(B ). For a
quasi-coherent sheaf of ideals J of B, Jfis quasi-coherent sheaf of ideals of OX , and the closed subprescheme
Y of X defined by J fis canonically isomorphic to Spec(B/J ).
P ROOF. It follows immediately from (I, 4.2.3) that Y is affine over S; by Proposition (1.3.1), we
reduce to the case where S is affine, and the proposition then follows immediately from (I, 4.1.2). □
We can also express the result of Proposition (1.4.10) by saying that if h : B → B ′ is a surjective
homomorphism of quasi-coherent OS -algebras, A (h) : Spec(B ′ ) → Spec(B ) is a closed immersion.
Proposition (1.4.11). — Let S be a prescheme, B a quasi-coherent OS -algebra, and X = Spec(B ). For II | 12
every quasi-coherent sheaf of ideals K of OS , we have (denoting by f the structure morphism X → S)
f ∗ (K )OX = (K B )e up to canonical isomorphism.
1. AFFINE MORPHISMS 205
P ROOF. Since the questions is local on S, we can reduce to the case where S = Spec( A) is affine,
and in this case the proposition is none other than (I, 1.6.9). □
Proposition (1.5.1). — Let X be a prescheme affine over S. For every extension g : S′ → S of the base
prescheme, X ′ = X(S′ ) = X ×S S′ is affine over S′ .
P ROOF. If f ′ is the projection X ′ → S′ , then it suffices to prove that f ′−1 (U ′ ) is an affine open
set for every affine open subset U ′ of S′ such that g(U ′ ) is contained in an affine open subset U
of S (1.2.1); we can thus reduce to the case where S and S′ are affine, and it suffices to prove that
X ′ is then an affine scheme (1.3.4). But then (1.3.4) X is an affine scheme, and if A, A′ , and B are
the rings of S, S′ , and X respectively, then we know that X ′ is the affine scheme with ring A′ ⊗ A B
(I, 3.2.2). □
Corollary (1.5.2). — Under the hypotheses of Proposition (1.5.1), let f : X → S be the structure morphism,
f ′ : X ′ → S′ and g′ : X ′ → X the projections, such that the diagram
g′
Xo X′
f f′
g
So S′
is commutative. For every quasi-coherent OX -module F , there exists a canonical isomorphism of OS′ -modules
and to set u = v♯ (0, 4.4.3). We take v = f ∗ (ρ), where ρ is the canonical homomorphism F →
g∗′ ( g′∗ (F )) (0, 4.4.3). To prove that u is an isomorphism, we can reduce to the case where S and S′ ,
hence X and X ′ , are affine; with the notation of Proposition (1.5.1), we then have F = M, e where
∗ ′ ′∗
M is a B-module. We then note immediately that g ( f ∗ (F )) and f ∗ ( g (F )) are both equal to the
OS′ -module associated to the A′ -module A′ ⊗ A M (where M is considered as an A-module), and
that u is the homomorphism associated to the identity ((I, 1.6.3), (I, 1.6.5), (I, 1.6.7)). □
Remark (1.5.3). — We do not have that Corollary (1.5.2) remains true when X is not assumed affine
over S, even when S′ = Spec(k(s)) (s ∈ S) and S′ → S is the canonical morphism (I, 2.4.5)—in
which case X ′ is none other than the fibre f −1 (s) (I, 3.6.2). In other words, when X is not affine over
S, the operation “direct image of quasi-coherent sheaves” does not commute with the operation II | 13
of “passing to fibres”. However, we will see in Chapter III (III, 4.2.4) a result in this sense, of an
“asymptotic” nature, valid for coherent sheaves on X when f is proper (5.4) and S is Noetherian.
Corollary (1.5.4). — For every prescheme X affine over S and every s ∈ S, the fibre f −1 (s) (where f denoted
the structure morphism X → S) is an affine scheme.
P ROOF. It suffices to apply Proposition (1.5.1) with S′ = Spec(k (s)) and to use Corollary (1.3.4).
□
Corollary (1.5.5). — Let X be an S-prescheme, S′ a prescheme affine over S; then X ′ = X(S′ ) is a prescheme
affine over X. In addition, if f : X → S is the structure morphism, then there is a canonical isomorphism of OX -
algebras A ( X ′ ) ≃ f ∗ (A (S′ )), and for every quasi-coherent A (S′ )-module M , a canonical di-isomorphism
f ∗ (M ) ≃ A ( f ′∗ (Mf)), denoting by f ′ = f (S′ ) the structure morphism X ′ → S′ .
(1.5.6.1) X′
v /X
q
S′ /S
is commutative (the vertical arrows being the structure morphisms). Indeed, the data of u is
equivalent to that of a homomorphism of quasi-coherent OS′ -algebras u♯ : q∗ (B ) → B ′ (0, 4.4.3);
this thus canonically defines an S′ -morphism
w : Spec(B ′ ) −→ Spec(q∗ (B ))
such that A (w) = u♯ (1.2.7). On the other hand, it follows from (1.5.2) that Spec(q∗ (B )) canonically
w p1
identifies with X ×S S′ ; the morphism v is the composition X ′ − → X ×S S′ −→ X of w with the
first projection, and the commutativity of (1.5.6.1) follows from the definitions. Let U (resp. U ′ )
be an affine open of S (resp. S′ ) such that q(U ′ ) ⊂ U, A = Γ(U, OS ), A′ = Γ(U ′ , OS′ ) their rings,
B = Γ(U, B ), B′ = Γ(U ′ , B ′ ); the restriction of u to a (q|U ′ )-morphism: B |U → B ′ |U ′ corresponds
to a di-homomorphism of algebras B → B′ ; if V, V ′ are the inverse images of U, U ′ in X, X ′
respectively, under the structure morphisms, then the morphism V ′ → V, the restriction of v,
corresponds (I, 1.7.3) to the above di-homomorphism.
(1.5.7). Under the same hypotheses as in (1.5.6), let M be a quasi-coherent B-module; there is then a
canonical isomorphism of OX ′ -modules
(1.5.7.1) v ∗ (Mf) ≃ (q∗ (M ) ⊗q∗ (B ) B ′ )e.
Indeed, the canonical isomorphism (1.5.2.1) gives a canonical isomorphism from p1∗ (M f) to the II | 14
sheaf on Spec(q∗ (B )) associated to the q∗ (B )-module q∗ (M ), and it then suffices to apply (1.4.7).
1.6. Affine morphisms
(1.6.1). We say that a morphism f : X → Y of preschemes is affine if it defines X as a prescheme affine
over Y. The properties of preschemes affine over another translates as follows in this language:
Proposition (1.6.2). —
(i) A closed immersion is affine.
(ii) The composition of two affine morphisms is affine.
(iii) If f : X → Y is an affine S-morphism, then f (S′ ) : X(S′ ) → Y(Y ′ ) is affine for every base change
S′ → S.
(iv) If f : X → Y and f ′ : X ′ → Y ′ are two affine S-morphisms, then
f ×S f ′ : X ×S X ′ −→ Y ×S Y ′
is affine.
(v) If f : X → Y and g : Y → Z are two morphisms such that g ◦ f is affine and g is separated, then f
is affine.
(vi) If f is affine, then so if f red .
P ROOF. By (I, 5.5.12), it suffices to prove (i), (ii), and (iii). But (i) is none other than Exam-
ple (1.2.2), and (ii) is none other than Corollary (1.3.5); finally, (iii) follows from Proposition (1.5.1),
since X(S′ ) identifies with the product X ×Y Y(S′ ) (I, 3.3.11). □
Corollary (1.6.3). — If X is an affine scheme and Y is a scheme, then every morphism f : X → Y is affine.
Proposition (1.6.4). — Let Y be a locally Noetherian prescheme, f : X → Y a morphism of finite type. For
f to be affine, it is necessary and sufficient for f red to be.
1. AFFINE MORPHISMS 207
P ROOF. By (1.6.2, vi), we see only need to prove the sufficiency of the condition. It suffices to
prove that if Y is affine and Noetherian, then X is affine; but Yred is then affine, so the same is true
for Xred by hypothesis. Now X is Noetherian, so the conclusion follows from (I, 6.1.7). □
1.7. Vector bundle associated to a sheaf of modules
(1.7.1). Let A be a ring, E an A-module. Recall that we call the symmetric algebra on E and denote by
S( E) (or S A ( E)) the quotient algebra of the tensor algebra T( E) by the two-sided ideal generated
by the elements x ⊗ y − y ⊗ x, where x and y vary over E. The algebra S( E) is characterized by the
following universal property: if σ is the canonical map E → S( E) (obtained by composing E → T( E)
with the canonical map T( E) → S( E)), then every A-linear map E → B, where B is a commutative
σ g
A-algebra, factors uniquely as E −
→ S( E) −
→ B, where g is an A-homomorphism of algebras. We
immediately deduce from this characterization that for two A-modules E and F, we have
S( E ⊕ F ) = S( E) ⊗ S( F )
up to canonical isomorphism; in addition, S( E) is a covariant functor in E, from the category of II | 15
A-modules to that of commutative A-algebras; finally, the above characterization also shows that if
E = lim Eλ , then we have S( E) = lim S( Eλ ) up to canonical isomorphism. By abuse of language, a
−→ −→
product σ ( x1 )σ ( x2 ) · · · σ ( xn ), where xi ∈ E, is often denoted by x1 x2 · · · xn if no confusion follows.
The algebra S( E) is graded, Sn ( E) being the set of linear combinations of n elements of E (n ⩾ 0); the
algebra S( A) is canonically isomorphic to the polynomial algebra A[ T ] is an indeterminate, and the
algebra S( An ) with the polynomial algebra in n indeterminates A[ T1 , . . . , Tn ].
(1.7.2). Let ϕ be a ring homomorphism A → B. If F is a B-module, then the canonical map F → S( F )
gives a canonical map F[ϕ] → S( F )[ϕ] , which thus factors as F[ϕ] → S( F[ϕ] ) → S( F )[ϕ] ; the canonical
homomorphism S( F[ϕ] ) → S( F )[ϕ] is surjective, but not necessarily bijective. If E is an A-module,
then every di-homomorphism E → F (that is to say, every A-homomorphism E → F[ϕ] ) thus
canonically gives an A-homomorphism of algebras S( E) → S( F[ϕ] ) → S( F )[ϕ] , that is to say a
di-homomorphism of algebras S( E) → S( F ).
With the same notations, for every A-module E, S( E ⊗ A B) canonically identifies with the
algebra S( E) ⊗ A B; this follows immediately from the universal property of S( E) (1.7.1).
(1.7.3). Let R be a multiplicative subset of the ring A; apply (1.7.2) to the ring B = R−1 A, and
remembering that R−1 E = E ⊗ A R−1 A, we see that we have S( R−1 E) = R−1 S( E) up to canonical
isomorphism. In addition, if R′ ⊃ R is a second multiplicative subset of A, then the diagram
R −1 E / R ′ −1 E
S ( R −1 E ) / S ( R ′ −1 E )
is commutative.
(1.7.4). Now let (S, A ) be a ringed space, and let E be a A -module over S. If to any open U ⊂ S we
associate the Γ(U, A )-module S(Γ(U, E )), then we define (see the functorial nature of S( E) (1.7.2))
a presheaf of algebras; we say that the associated sheaf, which we denote by S(E ) or SA (E ) is the
symmetric A -algebra on the A -module E . It follows immediately from (1.7.1) that S(E ) is a solution
to a universal problem: every homomorphism of A -modules E → B, where B is an A -algebra,
factors uniquely as E → S(E ) → B, the second arrow being a homomorphism of A -algebras.
There is thus a bijective correspondence between homomorphisms E → B of A -modules and
homomorphisms S(E ) → B of A -algebras. In particular, every homomorphism u : E → F of
A -modules defines a homomorphism S(u) : S(E ) → S(F ) of A -algebras, and S(E ) is thus a
covariant functor in E .
By (1.7.2) and the commutativity of S with inductive limits, we have (S(E )) x = S(Ex ) for every II | 16
point x ∈ S. If E , F are two A -modules, then S(E ⊕ F ) canonically identifies with S(E ) ⊗A S(F ),
as we see for the corresponding presheaves.
We also note that S(E ) is a graded A -algebra, the infinite direct sum of the Sn (E ), where the
A -module Sn (E ) is the sheaf associated to the presheaf U 7→ Sn (Γ(U, E )). If we take in particular
1. AFFINE MORPHISMS 208
Definition (1.7.8). — Let E be a quasi-coherent OS -module. We call the vector bundle over S defined
by E and denote by V(E ) the spectrum (1.3.1) of the quasi-coherent OS -algebra S(E ).
By (1.2.7), for every S-prescheme X, there is a canonical bijective correspondence between the
S-morphisms X → V(E ) and the homomorphisms of OS -algebras S(E ) → A ( X ), thus also between
these S-morphisms and the homomorphisms of OS -modules E → A ( X ) = f ∗ (OX ) (where f is the
structure morphism X → S). In particular:
(1.7.9). Take for X a subprescheme induced by S on an open U ⊂ S. Then the S-morphisms U → V(E )
are none other than the U-sections (I, 2.5.5) of the U-prescheme induced by V(E ) on the open p−1 (U )
(where p is the structure morphism V(E ) → S). From what we have just seen, these U-sections
bijectively correspond to homomorphisms of OS -modules E → j∗ (OS |U ) (where j is the canonical
injection U → S), or equivalently (0, 4.4.3) with the (OS |U )-homomorphisms j∗ (E ) = E |U → OS |U. II | 17
In addition, it is immediate that the restriction to an open U ′ ⊂ U of an S-morphism U → V(E )
corresponds to the restriction to U ′ of the corresponding homomorphism E |U → OS |U. We conclude
that the sheaf of germs of S-sections of V(E ) is canonically identified with the dual E ∨ of E .
In particular, if we set X = U = S, then the zero homomorphism E → OS corresponds to a
canonical S-section of V(E ), called the zero S-section (cf. (8.3.3)).
(1.7.10). Now take X to be the spectrum {ξ } of a field K; the structure morphism f : X → S then
corresponds to a monomorphism k(s) → K, where s = f (ξ ) (I, 2.4.6); the S-morphisms {ξ } → V(E )
are none other than the geometric points of V(E ) with values in the extension K of k (s) (I, 3.4.5), points
which are localized at the points of p−1 (s). The set of these points, which we can call the rational
geometric fibre over K of V(E ) over the point s, is identified by (1.7.8) with the set of homomorphisms of
OS -modules E → f ∗ (OX ), or, equivalently (0, 4.4.3) with the set of homomorphisms of OX -modules
f ∗ (E ) → OX = K. But we have by definition (0, 4.3.1) f ∗ (E ) = Es ⊗Os K = E s ⊗k(s) K, setting
E s = Es /ms Es ; the geometric fibre of V(E ) rational over K over s thus identifies with the dual of the
K-vector space E s ⊗k(s) K; if E s or K is of finite dimension over k (s), then this dual also identifies with
(E s )∨ ⊗k(s) K, denoting by (E s )∨ the dual of the k(s)-vector space E s .
Proposition (1.7.11). —
1. AFFINE MORPHISMS 209
(i) V(E ) is a contravariant functor in E from the category of quasi-coherent OS -modules to the category
of affine S-schemes.
(ii) If E is an OS -module of finite type, then V(E ) is of finite type over S.
(iii) If E and F are two quasi-coherent OS -modules, then V(E ⊕ F ) canonically identifies with
V (E ) × S V (F ).
(iv) Let g : S′ → S be a morphism; for every quasi-coherent OS -module E , V( g∗ (E )) canonically
identifies with V(E )(S′ ) = V(E ) ×S S′ .
(v) A surjective homomorphism E → F of quasi-coherent OS -modules corresponds to a closed immer-
sion V(F ) → V(E ).
P ROOF. (i) is an immediate consequence of (1.2.7), taking into account that every homomor-
phism of OS -modules E → F canonically defines a homomorphism of OS -algebras S(E ) → S(F ).
(ii) follows immediately from the definition (I, 6.3.1) and the fact that if E is an A-module of finite
type, then S( E) is an A-algebra of finite type. To prove (iii), it suffices to start with the canonical
isomorphism S(E ⊕ F ) ≃ S(E ) ⊗OS S(F ) (1.7.4) and to apply (1.4.6). Similarly, to prove (iv), it
suffices to start with the canonical isomorphism S( g∗ (E )) ≃ g∗ (S(E )) (1.7.5) and to apply (1.5.2).
Finally, to establish (v), it suffices to remark that if the homomorphism E → F is surjective, then
so is the corresponding homomorphism S(E ) → S(F ) of OS -algebras, and the conclusion follows
from (1.4.10). □
(1.7.12). Take in particular E = OS ; the prescheme V(OS ) is the affine S-scheme, spectrum of the
OS -algebra S(OS ) which identifies with the OS -algebra OS [ T ] = OS ⊗Z Z[ T ] (T indeterminate); this II | 18
is evident when S = Spec(Z), by virtue of (1.7.6), and we pass from there to the general case by
considering the structure morphism S → Spec(Z) and using (1.7.11, iv). Because of this result, we
set V(OS ) = S[ T ], and we thus have the formula
(1.7.12.1) S [ T ] = S ⊗ Z Z [ T ].
The identification of the sheaf of germs of S-sections of S[ T ] with OS , already seen in (I, 3.3.15),
here in a more general context, as a special case of (1.7.9).
(1.7.13). For every S-prescheme X, we have seen (1.7.8) that HomS ( X, S[ T ]) canonically identifies
with HomOS (OS , A ( X )), which is canonically isomorphic to Γ(S, A ( X )), and as a result is equipped
with the structure of a ring; in addition, to every S-morphism h : X → Y there corresponds a
morphism Γ(A (h)) : Γ(S, A (Y )) → Γ(S, A ( X )) for the ring structures (1.1.2). When we equip
HomS ( X, S[ T ]) with a ring structure as defined, then we can see that Hom( X, S[ T ]) can be consid-
ered as a contravariant functor in X, from the category of S-preschemes to that of rings. On the other
hand, HomS ( X, V(E )) is likewise identified with HomOS (E , A ( X )) (where A ( X ) is considered
as an OS -module); as a result, we can canonically equip it with a module structure over the ring
HomS ( X, S[ T ]), and we see as above that the pair
(HomS ( X, S[ T ]), Hom( X, V(E )))
is a contravariant functor in X, with values in the category whose elements are the pairs ( A, M)
consisting of a ring A and an A-module M, the morphisms being di-homomorphisms.
We will interpret these facts by saying that S[ T ] is an S-scheme of rings and that V(E ) is an
S-scheme of modules on the S-scheme of rings S[ T ] (cf. Chapter 0, §8).
(1.7.14). We will see that the structure of an S-scheme of modules defined on the S-scheme V(E )
allows us to reconstruct the OS -module E up to unique isomorphism: for this, we will show
that E is canonically isomorphic to an OS -submodule of S(E ) = A (V(E )), defined by means
of this structure. Indeed (1.7.4) the set HomOS (S(E ), A ( X )) of homomorphisms of OS -algebras is
canonically identified with HomOS (E , A ( X )), the set of homomorphisms of OS -modules: if h and
h′ are two elements of this latter set, si (1 ⩽ i ⩽ n) sections of E over an open U ⊂ S, t a section of
A ( X ) over U, then we have by definition
n
(h + h′ )(s1 s2 · · · sn ) = ∏(h(si ) + h′ (si ))
i =1
2. HOMOGENEOUS PRIME SPECTRA 210
and
n
(t · h)(s1 s2 · · · sn ) = tn ∏ h(si ).
i =1
This being so, if z is a section of S(E ) over U, then h 7→ h(z) is a map from HomS ( X, V(E )) =
HomOS (S(E ), A ( X )) to Γ(U, A ( X )). We will show that E is identified with a submodule of S(E ) II | 19
such that, for every open U ⊂ S, every section z of this OS -submodule of U, and every S-prescheme X, the map
h 7→ h(z) from HomOS (S(E )|U, A ( X )|U ) to Γ(U, A ( X )) is a homomorphism of Γ(U, A ( X ))-modules.
It is immediate that E has this property; to show the converse, we can reduce to proving that
when S = Spec( A), E = M, e a section z of S(E ) over S that (for U = S) has the property stated above
is necessarily a section of E ; we then have z = ∑∞ n=0 zn , where zn ∈ Sn ( M ), and it is a question of
proving that zn = 0 for n ̸= 1. Set B = S( M ) and take for X the prescheme Spec( B[ T ]), where T
is an indeterminate. The set HomOS (S(E ), A ( X )) identifies with the set of ring homomorphisms
h : B → B[ T ] (I, 1.3.13), and from what we saw above, we have ( T · h)(z) = ∑∞ n
n=0 T h ( zn ): the
∞ n ∞
hypothesis on z implies that we have ∑n=0 T h(zn ) = T · ∑n=0 h(zn ) for every homomorphism h.
In in particular we take for h the canonical injection, then ∑∞ n ∞
n=0 T zn = T · ∑n=0 zn , which implies
the conclusion zn = 0 for n ̸= 1.
Proposition (1.7.15). — Let Y be a prescheme whose underlying space is Noetherian, or a quasi-compact
scheme. Every affine Y-scheme X of finite type over Y is Y-isomorphic to a closed Y-subscheme of a Y-scheme
of the form V(E ), where E is a quasi-coherent OY -module of finite type.
P ROOF. The quasi-coherent OY -algebra A ( X ) is of finite type (1.3.7). The hypotheses imply
that A ( X ) is generated by a quasi-coherent OY -submodule of finite type E (I, 9.6.5); by definition,
this implies that the canonical homomorphism S(E ) → A ( X ) canonically extending the injection
E → A ( X ) is surjective; the conclusion then follows from (1.4.10). □
and N are respectively direct sums of the Mm and the Nn , we know that we can canonically identify
M ⊗Z N with the direct sum of all the Mm ⊗Z Nn ). This being so, we have M ⊗S N = ( M ⊗Z N )/P,
where P is the Z-submodule of M ⊗Z N generated by the elements ( xs) ⊗ y − x ⊗ (sy) for x ∈ M,
2. HOMOGENEOUS PRIME SPECTRA 211
(v) There exists an integer h > 0 such that Smh = (Sh )m for all m > 0.
n for all m ⩾ m .
(vi) For every integer n > 0, there exists an integer m0 such that Sm ⊂ S+ 0
are only finitely many systems (αi ) satisfying this equation, since the hi are > 0, hence the first
h/h
assertion of (i); the second is evident. On the other hand, let h be the l.c.m. of the hi and set gi = f i i
α1
(1 ⩽ i ⩽ r) such that all the gi are of degree h; let zµ be the elements of M of the form f 1 · · · f rαr x j
with 0 ⩽ αi < h/hi for 1 ⩽ i ⩽ r; there are finitely many of these elements, so let n1 be the largest
of their degrees. It is clear that for n ⩾ n1 , every element of Mn+h is a linear combination of the zµ
whose coefficients are monomials of degree > 0 with respect to the gi , so we have Mn+h = Sh Mn ,
which establishes (ii). In a similar way, we see (for all d > 0) that an element of M(d,k) is a linear
combinations, with coefficients in S0 , of elements of the form gd f 1 1 · · · f rαr x j with 0 ⩽ αi < d, g being
α
a homogeneous element of S; hence (iii); (iv) then follows from (iii) and from Lemma (2.1.3), by
taking M = S+ , since (S+ )(d) = (S(d) )+ . The assertion of (v) is deduced from (ii) by taking M = S.
Finally, for a given n, there are finitely many systems (αi ) such that αi ⩾ 0 and ∑i αi < n, so if m0
is the largest value of the sum ∑i αi hi of these systems, then we have Sm ⊂ S+ n for m > m , which
0
proves (vi). □
(2.1.8). Let p be a graded prime ideal of the graded ring S; p is thus a direct sum of the subgroups
pn = p ∩ Sn . Suppose that p does not contain S+ . Then if f ∈ S+ is not in p, the relation f n x ∈ p is
equivalent to x ∈ p; in particular, if f ∈ Sd (d > 0), for all x ∈ Sm−nd , then the relation f n x ∈ pm is
equivalent to x ∈ pm−nd .
Proposition (2.1.9). — Let n0 be an integer > 0; for all n ⩾ n0 , let pn be a subgroup of Sn . For there to
exist a graded prime ideal p of S not containing S+ and such that p ∩ Sn = pn for all n ⩾ n0 , it is necessary
and sufficient for the following conditions to be satisfied:
(1st) Sm pn ⊂ pm+n for all m ⩾ 0 and all n ⩾ n0 .
(2nd) For m ⩾ n0 , n ⩾ n0 , f ∈ Sm , g ∈ Sn , the relation f g ∈ pm+n implies f ∈ pm or g ∈ pn .
(3rd) pn ̸= Sn for at least one n ⩾ n0 .
In addition, the graded prime ideal p is then unique.
P ROOF. It is evident that the conditions (1st) and (2nd) are necessary. In addition, if p ̸⊃ S+ ,
then there exists at least one k > 0 such that p ∩ Sk ̸= Sk ; if f ∈ Sk is not in p, the relation p ∩ Sn = Sn
implies p ∩ Sn−mk = Sn−mk according to (2.1.8); therefore, if p ∩ Sn = Sn for a certain value of n, we
would have p ⊃ S+ contrary to the hypothesis, which proves that (3rd) is necessary. Conversely,
suppose that the conditions (1st), (2nd), and (3rd) are satisfied. Note that if for an integer d ⩾ n0 ,
f ∈ Sd is not in pd , then, if p exists, pm , for m < n0 , is necessarily equal to the set of the x ∈ Sm such
that f r x ∈ pm+rd , except for a finite number of values of r. This already proves that if p exists, then it
is unique. It remains to show that if we define the pm for m < n0 by the previous condition, then
p = ∑∞ n=0 pn is a prime ideal. First, note that by virtue of (2nd), for m ⩾ n0 , pm is also defined as
the set of the x ∈ Sm such that f r x ∈ pm+rd except for a finite number of values of r. This being II | 23
so, if g ∈ Sm , x ∈ pn , then we have f r gx ∈ pm+n+rd except for a finite number of values of r, so
gx ∈ pm+n , which proves that p is an ideal of S. To establish that this ideal is prime, in other words
that the ring S/p, graded by the subgroups Sn /pn , is an integral domain, it suffices (by considering
the components of higher degree of two elements of S/p) to prove that if x ∈ Sm and y ∈ Sn are such
that x ̸∈ pm and y ̸∈ pn , then xy ̸∈ pm+n . If not, for r large enough, we would have f 2r xy ∈ pm+n+2rd ;
but we have f r y ̸∈ pn+rd for all r > 0; it then follows from (2nd) that, except for a finite number of
values of r, we have f r x ∈ pm+rd , and we conclude that x ∈ pm contrary to the hypothesis. □
2. HOMOGENEOUS PRIME SPECTRA 213
(2.1.11). We note that if, in the graded ring S, an element x is a zero-divisor, then so is its component
of highest degree. We say that a ring S is essentially integral if the ring S+ (without the unit element)
does not contain a zero-divisor and is ̸= 0; it suffices that a homogeneous element ̸= 0 in S+ is not
a zero-divisor in this ring. It is clear that if p is a graded prime ideal of S+ , then S/p is essentially
integral.
Let S be an essentially integral graded ring, and let x0 ∈ S0 : if there then exists a homogeneous
element f ̸= 0 of S+ such that x0 f = 0, then we have x0 S+ = 0, since we have ( x0 g) f = ( x0 f ) g = 0
for all g ∈ S+ , and the hypothesis thus implies x0 g = 0. For S to be integral, it is necessary and
sufficient for S0 to be integral and the annihilator of S+ in S0 to be 0.
(2.2.1). Let S be a graded ring, in positive degrees, f a homogeneous element of S, of degree d > 0;
then the ring of fractions S′ = S f is graded, taking for Sn′ the set of the x/ f k , where x ∈ Sn+kd with
k ⩾ 0 (we observe here that n can take arbitrary negative values); we denote the subring S0′ = (S f )0
of S′ consisting of elements of degree 0 by the notation S( f ) .
If f ∈ Sd , then the monomials ( f /1)h in S f (h a positive or negative integer) form a free system
over the ring S( f ) , and the set of their linear combinations is none other than the ring (S(d) ) f , which II | 24
is thus isomorphic to S( f ) [ T, T −1 ] = S( f ) ⊗Z Z[ T, T −1 ] (where T is an indeterminate). Indeed, if we
have a relation ∑bh=− a zh ( f /1)h = 0 with zh = xh / f m , where the xh are in Smd , then this relation is
equivalent by definition to the existence of a k > − a such that ∑bh=− a f h+k xh = 0, and as the degrees
of the terms of this sum are distinct, we have f h+k xh = 0 for all h, hence zh = 0 for all h.
If M is a graded S-module, then M′ = M f is a graded S f -module, Mn′ being the set of the z/ f k
with z ∈ Mn+kd (k ⩾ 0); we denote by M( f ) the set of the homomogenous elements of degree 0 of
M′ ; it is immediate that M( f ) is an S( f ) -module and that we have ( M(d) ) f = M( f ) ⊗S( f ) (S(d) ) f .
Lemma (2.2.2). — Let d and e be integers > 0, f ∈ Sd , g ∈ Se . There exists a canonical ring isomorphism
S( f g ) ≃ ( S( f ) ) g d / f e ;
if we canonically identify these two rings, then there exists a canonical module isomorphism
M( f g ) ≃ ( M( f ) ) g d / f e .
P ROOF. Indeed, f g divides f e gd , and this latter element divides ( f g)de , so the graded rings S f g
and S f e gd are canonically identified; on the other hand, S f e gd also identifies with (S f e ) gd /1 (0, 1.4.6),
and as f e /1 is invertible in S f e , S f e gd also identifies with (S f e ) gd / f e . The element gd / f e is of degree 0
in S f e ; we immediately conclude that the subring of (S f e ) gd / f e consisting of elements of degree 0 is
(S( f e ) ) gd / f e , and as we evidently have S( f e ) = S( f ) , this proves the first part of the proposition; the
second is established in a similar way. □
2. HOMOGENEOUS PRIME SPECTRA 214
(2.2.3). Under the hypotheses of (2.2.2), it is clear that the canonical homomorphism S f → S f g
(0, 1.4.1), which sends x/ f k to gk x/( f g)k , is of degree 0, thus gives by restriction a canonical homo-
morphism S( f ) → S( f g) , such that the diagram
S( f )
} $
S( f g )
∼ / (S ) d e
(f) g /f
P ROOF. By virtue of Lemma (2.2.2), it suffices to see that 1/( gd / f e ) = f d+e /( f g)d belongs to
the canonical image of S( g) in S( f g) , which is evident by definition. □
Proposition (2.2.5). — Let d be an integer > 0 and let f ∈ Sd . Then there exists a canonical ring
isomorphisms S( f ) ≃ S(d) /( f − 1)S(d) ; if we identify these two rings by this isomorphism, then there exists
a canonical module isomorphism M( f ) ≃ M(d) /( f − 1) M(d) .
P ROOF. The first of these isomorphisms is defined by sending x/ f n , where x ∈ Snd , to the
element x, the class of x mod. ( f − 1)S(d) ; this map is well-defined, because we have the congruence
f h x ≡ x (mod. ( f − 1)S(d) ) for all x ∈ S(d) , so if f h x = 0 for an h > 0, then we have x = 0. On the II | 25
other hand, if x ∈ Snd is such that x = ( f − 1)y with y = yhd + y(h+1)d + · · · + ykd with y jd ∈ S jd and
yhd ̸= 0, then we necessarily have h = n and x = −yhd , as well as the relations y( j+1)d = f y jd for
h ⩽ j ⩽ k − 1, f ykd = 0, which ultimately gives f k−n x = 0; we send every class x mod. ( f − 1)S(d)
of an element x ∈ Snd to the element x/ f n of S( f ) , since the preceding remark shows that this map
is well-defined. It is immediate that these two maps thus defined are ring homomorphisms, each
the reciprocal of the other. We proceed exactly the same way for M. □
Corollary (2.2.6). — If S is Noetherian, then so is S( f ) for f homogeneous of degree > 0.
P ROOF. This follows immediately from Corollary (2.1.7) and Proposition (2.2.5). □
(2.2.7). Let T be a multiplicative subset of S+ consisting of homogeneous elements; T0 = T ∪ {1}
is then a multiplicative subset of S; as the elements of T0 are homogeneous, the ring T0−1 S is still
graded in the evident way; we denote by S(T ) the subring of T0−1 S consisting of elements of order
0, that is to say, the elements of the form x/h, where h ∈ T and x is homogeneous of degree equal
to that of h. We know (0, 1.4.5) that T0−1 S is canonically identified with the inductive limit of the
rings S f , where f varies over T (with respect to the canonical homomorphisms S f → S f g ); as this
identification respects the degrees, it identifies S(T ) with the inductive limit of the S( f ) for f ∈ T. For
every graded S-module M, we similarly define the module M(T ) (over the ring S(T ) ) consisting of
elements of degree 0 of T0−1 M, and we see that this module is the inductive limit of the M( f ) for
f ∈ T.
If p is a graded prime ideal of S+ , then we denote by S(p) and M(p) the ring S(T ) and the module
M(T ) respectively, where T is the set of homogeneous elements of S+ which do not belong to p.
2. HOMOGENEOUS PRIME SPECTRA 215
(2.3.2.3) V+ ( EE′ ) = V+ ( E) ∪ V+ ( E′ ).
We do not change V+ ( E) by replacing E with the graded ideal generated by E; in addition, if J
is a graded ideal of S, then we have
V+ (J) = V+ q⩾n (J ∩ Sq )
S
(2.3.2.4)
for all n > 0: indeed, if p ∈ Proj(S) contains the homogeneous elements of J of degree ⩾ n, then as II | 26
by hypothesis there exists a homogeneous element f ∈ Sd not contained in p, for every m ⩾ 0 and
every x ∈ Sm ∩ J, we have f r x ∈ J ∩ Sm+rd for all but finitely many values of r, so f r x ∈ p ∩ Sm+rd ,
which implies that x ∈ p ∩ Sm (2.1.9).
Finally, we have, for every graded ideal J of S,
(2.3.2.5) V+ (J) = V+ (r+ (J)).
(2.3.3). By definition, the V+ ( E) are the closed subsets of X = Proj(S) for the topology induced by
the spectral topology of Spec(S), which we also call the spectral topology on X. For all f ∈ S, we set
(2.3.3.1) D+ ( f ) = D ( f ) ∩ Proj(S) = Proj(S) − V+ ( f )
and so, for any two elements f and g of S (I, 1.1.9.1),
(2.3.3.2) D+ ( f g ) = D+ ( f ) ∩ D+ ( g ) .
Proposition (2.3.4). — The D+ ( f ), as f runs over the set of homogeneous elements of S+ , form a base for
the topology of X = Proj(S).
P ROOF. It follows from (2.3.2.2) and (2.3.2.4) that every closed subset of X is the intersection of
sets of the form V+ ( f ), where f is homogeneous of degree > 0. □
(2.3.5). Let f be a homogeneous element of S+ , of degree d > 0; for every graded prime ideal p of S
that does not contain f , we know that the set of the x/ f n , where x ∈ p and n ⩾ 0, is a prime ideal of
the ring of fractions S f (0, 1.2.6); its intersection with S( f ) is thus a prime ideal of S( f ) , which we
denote by ψ f (p): it is the set of the x/ f n for n ⩾ 0 and x ∈ p ∩ Snd . We have thus defined a map
ψ f : D+ ( f ) −→ Spec(S( f ) );
furthermore, if g ∈ Se is another homogeneous element of S+ , then we have a commutative diagram
ψf
(2.3.5.1) D+ ( f ) / Spec(S )
O O
(f)
D+ ( f g ) / Spec(S
ψf g ( f g) )
where the vertical arrow on the left is the inclusion, and the vertical arrow on the right is the map
aω
f g, f induced by the canonical homomorphism ω = ω f g, f : S( f ) → S( f g) (I, 1.2.1). Indeed, if
x/ f n ∈ ω −1 (ψ f g (p)), with f g ̸∈ p, then, by definition, gn x/( f g)n ∈ ψ f g (p), so gn x ∈ p, and so
x ∈ p; the converse is evident.
Proposition (2.3.6). — The map ψ f is a homeomorphism from D+ ( f ) to Spec(S( f ) ).
2. HOMOGENEOUS PRIME SPECTRA 216
P ROOF. Firstly, ψ f is continuous; this is since, if h ∈ Snd is such that h/ f n ∈ ψ f (p), then, by
definition, h ∈ p, and conversely, and so ψ− 1 n
f ( D ( h/ f )) = D+ ( h f ), and our claim then follows from
(2.3.3.2). Furthermore, the D+ (h f ), where h runs over the sets Snd , form a topology of D+ ( f ), by
(2.3.4) and (2.3.3.2); the above thus proves, taking into account the (T0 ) axiom, which holds in D+ ( f ) II | 27
and in Spec(S( f ) ), that ψ f is injective and that the inverse map ψ f ( D+ ( f )) → D+ ( f ) is continuous.
Finally, to see that ψ f is surjective, we note that, if q0 is a prime ideal of S( f ) , and if, for all n > 0,
we denote by pn the set of x ∈ Sn such that x d / f n ∈ q0 , then the pn satisfy the conditions of (2.1.9):
if x, y ∈ Sn are such that x d / f n , yd / f n ∈ q0 , then ( x + y)2d / f 2n ∈ q0 , whence ( x + y)d / f n ∈ q0 ,
since q0 is prime; this proves that the pn are subgroups of the Sn , and the verification of the other
conditions of (2.1.9) is immediate, taking into account the fact that q0 is prime. If p is the graded
prime ideal of S thus defined, then indeed ψ f (p) = q0 , since, if x ∈ Snd , then having x/ f n ∈ q0 and
x d / f nd ∈ q0 is equivalent to q0 being prime. □
Corollary (2.3.7). — To have D+ ( f ) = ∅, it is necessary and sufficient for f to be nilpotent.
P ROOF. To have Spec(S( f ) ) = ∅, it is necessary and sufficient to have S( f ) = 0, or indeed to
have 1 = 0 in S f , which means, by definition, that f is nilpotent. □
Corollary (2.3.8). — Let E be a subset of S+ . Then the following conditions are equivalent:
(a) V+ ( E) = X = Proj(S).
(b) Every element of E is nilpotent.
(c) The homogeneous components of every element of E are nilpotent.
P ROOF. It is clear that (c) implies (b), and that (b) implies (a). If J is the graded ideal of S
generated by E, then condition (a) is equivalent to requiring that V+ (J) = X; a fortiori, (a) implies
that every homogeneous element f ∈ J is such that V+ ( f ) = X, and so f is nilpotent by (2.3.7). □
Corollary (2.3.9). — If J is a graded ideal of S+ , then r+ (J) is the intersection of the graded prime ideals of
S+ that contain J.
P ROOF. By considering the graded ring S/J, we can reduce to the case where J = 0. We need
to prove that, if f ∈ S+ is not nilpotent, then there exists a graded prime ideal of S that does not
contain f ; but at least one of the homogeneous components of f is not nilpotent, and we can thus
suppose f to be homogeneous; the claim then follows from (2.3.7). □
(2.3.10). For every subset Y of X = Proj(S), let j+ (Y ) be the set of f ∈ S+ such that Y ⊂ V+ ( f ); this
is equivalent to saying that j+ (Y ) = j(Y ) ∩ S+ ; then j+ (Y ) is an ideal of S+ that is equal to its radical
in S+ .
Proposition (2.3.11). — (i) For every subset E of S+ , j+ (V+ ( E)) is the radical in S+ of the graded
ideal of S+ generated by E.
(ii) For every subset Y of X, V+ (j+ (Y )) = Y, where Y is the closure of Y in X.
P ROOF.
(i) If J is the graded ideal of S+ generated by E, then V+ ( E) = V+ (J), and the claim then
follows from (2.3.9).
(ii) Since V+ (J) = f ∈J V+ ( f ), having Y ⊂ V+ (J) implies that Y ⊂ V+ ( f ) for every f ∈ J,
T
and thus j+ (Y ) ⊃ J, whence V+ (j+ (Y )) ⊂ V+ (J), which proves (ii) by the definition of the
closed subsets.
□
Corollary (2.3.12). — The closed subsets Y of X = Proj(S) are in bijective correspondence with the
graded ideals of S+ that are equal to their radical in S+ , via the inclusion-reversing maps Y 7→ j+ (Y )
and J 7→ V+ (J); the union Y1 ∪ Y2 of two closed subsets of X corresponds to j+ (Y1 ) ∩ j+ (Y2 ), and the II | 28
intersection of an arbitrary family (Yλ ) of closed subsets corresponds to the radical in S+ of the sum of the
j+ (Yλ ).
Corollary (2.3.13). — Let J be a graded ideal of S+ ; to have V+ (J) = ∅, it is necessary and sufficient for
every element f of S+ to have a power f n in J.
2. HOMOGENEOUS PRIME SPECTRA 217
This above corollary can also be expressed in one of the following equivalent forms:
Corollary (2.3.14). — Let ( f α ) be a family of homogeneous elements of S+ . For the D+ ( f α ) to form a cover
of X = Proj(S), it is necessary and sufficient for every element of S+ to have a power in the ideal generated
by the f α .
Corollary (2.3.15). — Let ( f α ) be a family of homogeneous elements of S+ , and f an element of S+ . Then
the following are equivalent:
(a) D+ ( f ) ⊂ α D+ ( f α );
S
(b) V+ ( f ) ⊃ α V+ ( f α );
T
P ROOF.
2. HOMOGENEOUS PRIME SPECTRA 218
(i) Let S be the graded ring S/N+ , and denote by x 7→ x the canonical homomorphism S → S
of degree 0. For all f ∈ Sd (d > 0), the canonical homomorphism S f → S (0, 1.5.1) is
surjective and of degree 0, and thus gives, by restriction, a surjective homomorphism
S( f ) → S( f ) ; if we suppose that f ̸∈ N+ , then we immediately see that S( f ) is reduced,
and that the kernel of the above homomorphism is the nilradical of S( f ) , or, in other
words, that S( f ) = (S( f ) )red . So to this homomorphism corresponds a closed immersion
D+ ( f ) → D+ ( f ) that identifies D+ ( f ) with ( D+ ( f ))red (I, 5.1.2), and which is, in particular,
a homeomorphism of the underlying spaces of these two affine schemes. Furthermore, if
g ̸∈ N+ is another homogeneous element of S+ , then the diagram
S( f ) /S
(f)
S( f g ) /S
( f g)
A
is commutative, for homogeneous f , g in S+ . □
Proposition (2.4.7). — Let S be a positively graded ring.
(i) For every integer d > 0, there exists a canonical isomorphism from the scheme Proj(S) to the scheme
Proj(S(d) ).
(ii) Let S′ be the graded ring such that S0 = Z and Sn′ = Sn (considered as a Z-module) for n > 0.
Then there exists a canonical isomorphism from the scheme Proj(S) to the scheme Proj(S′ ).
P ROOF.
(i) We first show that the map p 7→ p ∩ S(d) is a bijection from the set Proj(S) to the set
Proj(S(d) ). Indeed, suppose that we have a graded prime ideal p′ ∈ Proj(S(d) ), and let
pnd = p′ ∩ Snd (n ⩾ 0). For all n > 0 that are not multiples of d, define pn as the set of
2. HOMOGENEOUS PRIME SPECTRA 219
x ∈ Sn such that x d ∈ pnd ; if x, y ∈ pn , then ( x + y)2d ∈ p2nd , and so ( x + y)d ∈ pnd , since p′
is prime; it is immediate that the pn thus defined, for n ⩾ 0, satisfy the conditions of (2.1.9),
and so there exists a unique prime ideal p ∈ Proj(S) such that p ∩ S(d) = p′ . Since, for
every homogeneous f in S+ , we have that V+ ( f ) = V+ ( f d ) (2.3.2.3), we see that the above
bijection is a homeomorphism of topological spaces. Finally, with the same notation, S( f )
and S( f d ) are canonically identified (2.2.2), and so Proj(S) and Proj(S(d) ) are canonically
identified as schemes.
(ii) If, to each p ∈ Proj(S), we associate the unique prime ideal p′ ∈ Proj(S′ ) such that p′ ∩
Sn = p ∩ Sn for every n > 0 (2.1.9), then it is clear that we have defined a canonical
∼
homeomorphism Proj(S) − → Proj(S′ ) of the underlying spaces, since V+ ( f ) is the same set
for S and S when f is a homogeneous element of S+ . Since, further, S( f ) = S(′ f ) , Proj(S)
′
Corollary (2.4.8). — If S is a graded A-algebra, and S′A the graded A-algebra such that (S′A )0 = A and
(S′A )n = Sn for n > 0, then there exists a canonical isomorphism from Proj(S) to Proj(S′A ).
P ROOF. In fact, these two schemes are both canonically isomorphic to Proj(S′ ), using the
notation of (2.4.7, (ii)). □
(2.5.1). Let M be a graded S-module. The, for every homogeneous f in S+ , M( f ) is an S( f ) -module, and
thus has a corresponding quasi-coherent associated sheaf ( M( f ) )e on the affine scheme Spec(S( f ) ),
identified with D+ ( f ) (I, 1.3.4).
II | 31
Proposition (2.5.2). — There exists on X = Proj(S) exactly one quasi-coherent OX -module M e such that, for
f homogeneous in S+ , we have Γ( D+ ( f ), M e ) = M( f ) , with the restriction homomorphism Γ( D+ ( f ), M
e) →
Γ ( D+ ( f g ) , M
e ), for f and g homogeneous in S+ , being the canonical homomorphism M( f ) → M( f g) (2.2.3).
P ROOF. Suppose that f ∈ Sd and g ∈ Se . Since D+ ( f g) can be identified with the prime spec-
trum of (S( f ) ) gd / f e by (2.2.2), the restriction to D+ ( f g) of the sheaf ( M( f ) )e on D+ ( f ) is canonically
identified with the sheaf associated to the module ( M( f ) ) gd / f e (I, 1.3.6), and thus also with ( M( f g) )e
(2.2.2); we thus conclude that there exists a canonical isomorphism
∼
θ g, f : ( M( f ) )e| D+ ( f g) −
→ ( M( g) )e| D+ ( f g)
such that, if h is a third homogeneous element of S+ , then θ f ,h = θ f ,g ◦ θ g,h in D+ ( f gh). Consequently
(0, 3.3.1) there exists a quasi-coherent OX -module F on X, and, for every homogeneous f in S+ , an
isomorphism η f from F | D+ ( f ) to ( M f )e such that θ g, f = ηg ◦ η − 1
f . If we then consider the sheaf
G associated to the presheaf (on the base for the topology of X given by the D+ ( f )) defined by
D+ ( f ) 7→ M( f ) , with the canonical homomorphisms M( f ) → M( f g) as restriction homomorphisms,
then the above proves that F and G are isomorphic (taking (I, 1.3.7) into account); the sheaf G
is denoted by M, e and indeed satisfies the conditions of the statement. We have, in particular,
S = OX .
e □
Recall that the graded S-modules form a category when we restrict from arbitrary homomor-
phisms of graded modules to homomorphisms of degree 0. With this convention:
Proposition (2.5.4). — The functor M 7→ M e is an exact additive covariant functor from the category of
graded S-modules to the category of quasi-coherent OX -modules, and it commutes with inductive limits and
direct sums.
2. HOMOGENEOUS PRIME SPECTRA 220
P ROOF. Indeed, since these properties are local, it suffices to show that they are satisfied for
the sheaves of the form M e | D+ ( f ) = ( M( f ) )e; but the functors M 7→ M f , N 7→ N0 (to the category
of graded S f -modules), and P 7→ Pe (to the category of S( f ) -modules) all have the three properties
of exactness and of commutativity with inductive limits and direct sums ((I, 1.3.5) and (I, 1.3.9));
whence the proposition. □
e →N
We denote by ue the homomorphism M e corresponding to a homomorphism u : M → N of
degree 0. We immediately deduce from (2.5.4) that the results of (I, 1.3.9) and (I, 1.3.10) still hold
for graded S-modules and homomorphisms of degree 0 (with the sense given here to M), e with the
proofs being purely formal.
e p = lim Γ( D+ ( f ), M
P ROOF. By definition, M e ), where f runs over the set of homogeneous
−→
elements of S+ such that f ̸∈ p; since Γ( D+ ( f ), M
e ) = M( f ) , the proposition follows from the
definition of M(p) (2.2.7) □
In particular, the local ring (OX )p is exactly the ring S(p) , the set of fractions x/ f with f homoge- II | 32
neous in S+ and not belonging to p, and with x homogeneous of the same degree as f .
Even more particularly, if S is essential integral, so that Proj(S) = X is integral (2.4.4), and if
ξ = (0) is the generic point of X, then the field of rational functions R( X ) = Oξ is field consisting of
elements f g−1 with f and g homogeneous of the same degree in S+ , and with g ̸= 0.
Proposition (2.5.6). — If, for all z ∈ M and all homogeneous f in S+ , there exists a power of f that
e = 0. This sufficient condition is also necessary if the S0 -algebra S is generated by the
annihilates z, then M
set S1 of homogeneous elements of degree 1.
Proposition (2.5.7). — Let d > 0 be an integer, and let f ∈ Sd . Then, for all n ∈ Z, the (OX | D+ ( f ))-
module (S(nd))e| D+ ( f ) is canonically isomorphic to OX | D+ ( f ).
P ROOF. Indeed, multiplication by the invertible element ( f /1)n of S f gives a bijection from
S( f ) = (S f )0 to (S f )nd = (S f (nd))0 = (S(nd) f )0 = S(nd)( f ) ; in other words, the S( f ) -modules S( f )
and S(nd)( f ) are canonically isomorphic, whence the proposition. □
(2.5.11). Let M and N be graded S-modules. For all f ∈ Sd (d > 0), we define a canonical functorial
homomorphism of S( f ) -modules by
(2.5.11.1) λ f : M( f ) ⊗S( f ) N( f ) −→ ( M ⊗S N )( f )
It immediately follows from this definition that, if g ∈ Se (e > 0), then the diagram
λf
M( f ) ⊗S( f ) N( f ) / ( M ⊗S N )
(f)
M( f g) ⊗S( f g) N( f g) / ( M ⊗S N )
λfg ( f g)
(where the vertical arrow on the right is the canonical homomorphism, and the one on the left
comes from the canonical homomorphisms) commutes. Thus λ induces a canonical functorial
homomorphism of OX -modules
(2.5.11.2) e ⊗O N
λ:M e −→ ( M ⊗S N )e.
X
(2.5.11.3) J ⊗O X K
e e λ / (e
J ⊗S K
e )e
# z
OX
then commutes. Indeed, we can reduce to verifying this on each open subset D+ ( f ) (for f ho-
mogeneous in S+ ), and this follows immediately from the definition (2.5.11.1) of λ f and from
(I, 1.3.13).
Finally, note that, if M, N, and P are graded S-modules, then the diagram
λ ⊗1
(2.5.11.4) e ⊗O N
M e ⊗O Pe / ( M ⊗S N )e ⊗O Pe
X X X
1⊗ λ λ
e ⊗O ( N ⊗S P)e
M / ( M ⊗S N ⊗S P)e
X λ
commutes. It again suffices to verify this on each open subset D+ ( f ), and this follows immediately II | 34
from the definitions and from (I, 1.3.13).
diagram:
µf
(HomS ( M, N ))( f ) / HomS ( M , N )
(f) (f) (f)
(HomS ( M, N ))( f g) / HomS ( M
µfg ( f g) ( f g) , N( f g) )
(where the vertical arrow on the left is the canonical homomorphism, and the one on the right comes
from the canonical homomorphisms). We thus again conclude (taking (I, 1.3.8) into account) that the
µ f define a functorial canonical homomorphism of OX -modules
(2.5.12.2) µ : (HomS ( M, N ))e −→ H omOX ( M,
e Ne)
Proposition (2.5.13). — Suppose that the ideal S+ is generated by S1 . Then the homomorphism λ (2.5.11.2)
is an isomorphism; so too is the homomorphism µ (2.5.12.2) if the graded S-module M admits a finite
presentation (2.1.1)
P ROOF. Since X is the union of the D+ ( f ) for f ∈ S1 (2.3.14), we are led to proving that λ f and
µ f are isomorphisms, under the given hypotheses, whenever f is homogeneous and of degree 1. But
we can then define a Z-bilinear map Mm × Nn → M( f ) ⊗S( f ) N( f ) by sending ( x, y) to the element
( x/ f m ) ⊗ (y/ f n ) (if m < 0, we write x/ f m to mean f −m x/1); these maps define a Z-linear map
M ⊗Z N → M( f ) ⊗S( f ) N( f ) , and, if s ∈ Sq , this map sends (sx ) ⊗ y to (s/ f q )(( x/ f m ) ⊗ (y/ f n )) (for
x ∈ Mm and y ∈ Nn ). We thus obtain a di-homomorphism of modules γ f : M ⊗S N → M( f ) ⊗S( f )
N( f ) , with respect to the canonical homomorphism S → S( f ) (sending s ∈ Sq to s/ f q ). Suppose
furthermore that, for an element ∑i ( xi ⊗ yi ) of M ⊗S N (with xi and yi homogeneous of degree mi
and ni , respectively), we have that f r ∑i ( xi ⊗ yi ) = 0, or, in other words, that ∑i ( f r xi ⊗ yi ) = 0. We
thus deduce, by (0, 1.3.4), that ∑i ( f r xi / f mi +r ) ⊗ (yi / f ni ) = 0, i.e. γ f (∑i ( xi ⊗ yi )) = 0. Then γ f
γ′f
factors as M ⊗S N → ( M ⊗S N ) f −→ M( f ) ⊗S( f ) N( f ) ; if λ′f is the restriction of γ′ to ( M ⊗S N )( f ) , II | 35
then we can immediately show that λ f and λ′f are inverse S( f ) -homomorphisms, whence the first
part of the proposition.
To prove the second part, suppose that M is the cokernel of a homomorphism P → Q of graded
S-modules, with P and Q being direct sums of a finite number of modules of the form S(n); using
the left-exactness of HomS ( L, N ) in L, and the exactness of M( f ) in M, we can immediately reduce
to proving that µ f is an isomorphism whenever M = S(n). But, for any homogeneous z in N, let uz
be the homomorphism from S(n) to N such that uz (1) = z; we immediately see that η : z → uz is an
isomorphism of degree 0 from N (−n) to HomS (S(n), N ). There is a corresponding isomorphism
η f : ( N (−n))( f ) −→ (HomS (S(n), N ))( f ) .
Now let η ′f be the isomorphism N( f ) → HomS( f ) (S(n)( f ) , N( f ) ) that, to any z′ ∈ N( f ) , associates
the homomorphism vz′ that is such that vz′ (s/ f k ) = sz′ / f n+k (for s ∈ Sn+k = (S(n))k ). We easily
note that the composed map
−1
ηf µf η ′f
( N (−n))( f ) −→ (HomS (S(n), N ))( f ) −→ HomS( f ) (S(n)( f ) , N( f ) ) −−→ N( f )
is the isomorphism z/ f h 7→ z/ f h−n from ( N (−n))( f ) to N( f ) , and thus µ f is an isomorphism. □
If the ideal S+ is generated by S1 , then we deduce from (2.5.13) that, for every graded ideal J of
S, and for every graded S-module M, we have
(2.5.13.1) J·M
e e = (J · M )e
up to canonical isomorphism; this follows from the commutativity of the diagram
J ⊗O X M
e e λ / (J ⊗S M)e
# z
M
e
2. HOMOGENEOUS PRIME SPECTRA 223
2.6. The graded S-module associated to a sheaf on Proj(S) We suppose all throughout this section
that the ideal S+ of S is generated by the set S1 of homogeneous elements of degree 1.
(2.6.1). The OX -module OX (1) is invertible (2.5.9); we thus define, for every OX -module F (0, 5.4.6),
Γ • (F ) = Γ • (O X (1), F ) = Γ( X, F (n))
M
(2.6.1.1)
n∈Z
taking (2.5.14.2) into account. Recall (0, 5.4.6) that Γ• (OX ) is endowed with the structure of a graded
ring, and Γ• (F ) with the structure of a graded Γ• (OX )-module.
Since OX (n) is locally free, Γ• (F ) is a left exact additive covariant functor in F ; in particular, if
J is a sheaf of ideals of OX , then Γ• (J ) is canonically identified with a graded idea of Γ• (OX ).
(2.6.2). Let M be a graded S-module; for every f ∈ Sd (d > 0), x 7→ x/1 is a homomorphism of
abelian groups M0 → M( f ) , and, since M( f ) is canonically identified with Γ( D+ ( f ), M e ), we thus II | 37
f
obtain a homomorphism of abelian groups α0 : M0 → Γ( D+ ( f ), M e ). It is clear that, for every g ∈ Se
(e > 0), the diagram
Γ ( D+ ( f ) , M
e)
f 8
α0
M0
fg
α0 &
Γ ( D+ ( f g ) , M
e)
f g
commutes; this implies that, for all x ∈ M0 , the sections α0 ( x ) and α0 ( x ) of M agree on D+ ( f ) ∩
D+ ( g), and thus there exists a unique section α0 ( x ) ∈ Γ( X, M
e ) whose restriction to each D+ ( f ) is
f
α0 ( x ). We have thus defined (without using the hypothesis that S be generated by S1 ) a homomor-
phism of abelian groups
(2.6.2.1) α0 : M0 −→ Γ( X, M
e ).
Applying this result to the graded S-module M (n) (for each n ∈ Z), we obtain, for each n ∈ Z, a
homomorphism of abelian groups
(2.6.2.2) αn : Mn = ( M(n))0 −→ Γ( X, M
e (n))
(taking (2.5.15)); whence we obtain a functorial homomorphism (of degree 0) of graded abelian
groups
(2.6.2.3) α : M −→ Γ• ( M
e)
Proposition (2.6.3). — For every f ∈ Sd (d > 0), D+ ( f ) is identical to the set of p ∈ X on which the
section αd ( f ) of OX (d) does not vanish (0, 5.5.2).
P ROOF. Since X = g∈S1 D+ ( g) by hypothesis, it suffices to show that, for all g ∈ S1 , the set
S
(2.6.4). Now let F be an OX -modules, and set M = Γ• (F ); by the existence of the homomorphism of
graded rings α : S → Γ• (OX ), we can consider M as a graded S-module. For every f ∈ Sd (d > 0), it
follows from (2.6.3) that the restriction to D+ ( f ) of the section αd ( f ) of OX (d) is invertible; thus so too
is the restriction to D+ ( f ) of the section αd ( f n ) of OX (nd), for all n > 0. So let z ∈ Mnd = Γ( X, F (nd)
(n > 0); if there exists an integer k ⩾ 0 such that the restriction to D+ ( f ) of f k z, i.e. the section II | 38
(z| D+ ( f ))(αd ( f k )| D+ ( f )) of F ((n + k)d), is zero, then, by the above remark, we also have that
z| D+ ( f ) = 0. This shows that we have defined an S( f ) -homomorphism β f : M( f ) → Γ( D+ ( f ), F )
by sending the element z/ f n to the section (z| D+ ( f ))(αd ( f n )| D+ ( f ))−1 of F over D+ ( f ). We can
further immediately show that the diagram
βf
(2.6.4.1) M( f ) / Γ ( D+ ( f ) , F )
M( f g ) / Γ ( D+ ( f g ) , F )
βfg
Γ• ( β)
Γ • (F ) −
→ Γ• ((Γ• (F ))e) −−−→ Γ• (F )
α
(2.6.5.2)
are the identity isomorphisms.
P ROOF. The proof for (2.6.5.1) is local: in an open subset D+ ( f ), it follows immediately from
the definitions, along with the fact that β, applied to quasi-coherent sheaves, is determined by its
action on the sections over D+ ( f ) (I, 1.3.8). The proof for (2.6.5.2) is done for each degree separately:
if we set M = Γ• (F ), then Mn = Γ( X, F (n)), and (Γ• ( M e ))n = Γ( X, M e (n)) = Γ( X, ( M(n))e).
f
But if f ∈ S1 and z ∈ Mn , then αn (z) is the element z/1 of ( M (n))( f ) , equal to ( f /1)n (z/ f n ); it
corresponds, via β f , to the section
n −1
z| D+ ( f ) (α1 ( f ))n | D+ ( f )
α 1 ( f ) | D+ ( f )
over D+ ( f ), i.e. the restriction of z to D+ ( f ), which finishes the proof for (2.6.5.2). □
P ROOF.
(i) If S is Noetherian, then the ideal S+ admits a finite system of homogeneous generators f i
(1 ⩽ i ⩽ p), thus (2.3.14) the underlying space X is the union of the D+ ( f i ) = Spec(S( f i ) ),
and everything then reduces to showing that each of the S( f i ) is Noetherian, which follows
from (2.2.6).
2. HOMOGENEOUS PRIME SPECTRA 226
(ii) The hypothesis implies that S0 is an A-algebra of finite type, and that S is an S0 -algebra
of finite type, and so S+ is an ideal of finite type (2.1.4). We are thus reduced, as in (i), to
showing that S( f ) is an A-algebra of finite type for all f ∈ Sd . By (2.2.5), it suffices to show
that S(d) is an A-algebra of finite type, which follows from (2.1.6).
□
(2.7.2). In what follows, we consider the following finiteness conditions for a graded S-module M:
L
(TF) There exists an integer n such that the submodule k⩾n Mk is an S-module of finite type.
(TN) There exists an integer n such that Mk = 0 for k ⩾ n.
If M satisfies (TN), then M( f ) = 0 for all homogeneous f in S+ , and thus M
e = 0.
Let M and N be graded S-modules; we say that a homomorphism u : M → N of degree 0 is
(TN)-injective (resp. (TN)-surjective, (TN)-bijective) if there exists an integer n such that uk : Mk → Nk
is injective (resp. surjective, bijective) for k ⩾ n. Saying that u is (TN)-injective (resp. (TN)-surjective)
thus reduces to saying that Ker u (resp. Coker u) satisfies (TN). By (2.5.4), if u is (TN)-injective (resp.
(TN)-surjective, (TN)-bijective), then ue is injective (resp. surjective, bijective); if u is (TN)-bijective,
then we also say that u is a (TN)-isomorphism.
Proposition (2.7.3). — Let S be a graded ring such that the ideal S+ is of finite type, and let M be a graded
S-module.
(i) If M satisfies condition (TF) then the OX -module Me is of finite type.
(ii) Suppose that M satisfies (TF); for M = 0, it is necessary and sufficient for M to satisfy (TN).
e
P ROOF. We have just seen that condition (TN) implies that M e = 0. If M satisfies (TF), then
the graded submodule M = k⩾n Mk , which is, by hypothesis, of finite type, is such that M/M′
′ L
satisfies (TN); thus ( M/M′ )e, and the exactness of the functor M e (2.5.4) implies that M e =Mf′ ; to
prove that M is of finite type, we can thus reduce to the case where M is of finite type . Since the
e
question is local, it suffices to prove that M( f ) is an S( f ) -module of finite type (I, 1.3.9); but M(d) is
an S(d) -module of finite type (2.1.6, iii), and our claim then follows from (2.2.5).
Now suppose that M satisfies (TF) and that M e = 0; then, with the same notation as above, we
′ ′
have that M = 0, and condition (TN) for M is equivalent to condition (TN) for M, so to prove that
f
Me = 0 implies that M satisfies (TN), we can again restrict to the case where M is generated by a
finite number of homogeneous elements xi (1 ⩽ i ⩽ p); let ( f j )1⩽j⩽q be a system of homogeneous
generators of the ideal S+ . By hypothesis, M( f j ) = 0 for all j, and so there exists an integer n such
that f jn xi = 0 for any i and j. Let n j be the degree of f j , and let m be the largest value of ∑ j r j n j for the
system of finitely many integers (r j ) such that ∑ j r j ⩽ nq; it is then clear that, if k > m, then Sk xi = 0 II | 40
for all i; if h is the largest of the degrees of the xi , then we conclude that Mk = 0 for k > h + m, which
finishes the proof. □
Corollary (2.7.4). — Let S be a graded ring such that the ideal S+ is of finite type; for X = Proj(S) = ∅, it
is necessary and sufficient for there to exist n such that Sk = 0 for k ⩾ n.
Theorem (2.7.5). — Suppose that the ideal S+ is generated by a finite number of homogeneous elements
of degree 1; let X = Proj(S). Then, for every quasi-coherent OX -module F , the canonical homomorphism
β : (Γ• (F ))e → F (2.6.4) is an isomorphism.
P ROOF. If S+ is generated by a finite number of elements f i ∈ S1 , then X is the union of the
subspaces Spec(S( f i ) ) (2.3.6), which are quasi-compact, and so X is quasi-compact; furthermore, X
is a scheme (2.4.2); by (I, 9.3.2), (2.5.14.2), and (2.6.3), we have, for all f ∈ Sd (d > 0), a canonical
∼
isomorphism (Γ• (F ))(αd ( f )) − → Γ( D+ ( f ), F ); also, by definition, (Γ• (F ))(αd ( f )) (where Γ• (F ) is
considered as a Γ• (OX )-module) is exactly (Γ• (F ))( f ) (where Γ• (F ) is considered as an S-module);
if we refer to the definition (I, 9.3.1) of the above canonical isomorphism, then we see that it agrees
with the homomorphism β f , whence the theorem. □
2. HOMOGENEOUS PRIME SPECTRA 227
Remark (2.7.6). — If we suppose that the graded ring S is Noetherian, then the condition of (2.7.5) is
satisfied ipso facto as soon as we suppose that the ideal S+ is generated by the set S1 of homogeneous
elements of degree 1.
Corollary (2.7.7). — Under the hypotheses of (2.7.5), every quasi-coherent OX -module F is isomorphic to
an OX -module of the form M,
e where M is a graded S-module.
Corollary (2.7.8). — Under the hypotheses of (2.7.5), every quasi-coherent OX -module F of finite type is
isomorphic to an OX -module of the form N,
e where N is a graded S-module of finite type.
P ROOF. We can suppose that F = M, e where M is a graded S-module (2.7.7). Let ( f λ )λ∈ L be
a system of homogeneous generators of M; for every finite subset H of L, let M H be the graded
submodule of M generated by the f λ such that λ ∈ H; it is clear that M is the inductive limit of its
submodules M H , and so F is the inductive limit of its sub-OX -modules M g H (2.5.4). But, since F is
of finite type, and since the underlying space of X is quasi-compact, it follows from (0, 5.2.3) that
F =M g H for some finite subset H of L. □
Corollary (2.7.9). — Under the hypotheses of (2.7.5), let F be a quasi-coherent OX -module of finite type.
Then there exists an integer n0 such that, for all n ⩾ n0 , F (n) is isomorphic to a quotient of an OX -module
of the form OXk (for some k > 0 depending on n), and is thus generated by a finite number of its sections over
X (0, 5.1.1).
P ROOF. By (2.7.8), we can suppose that F = M,e where M is a quotient of a finite direct sum of
S-modules of the form S(mi ); by (2.5.4), we can thus restrict to the case where M = S(m), and so
F (n) = (S(m + n))e = OX (m + n) (2.5.15). It thus suffices to prove
| 41
Lemma (2.7.9.1). — Under the hypotheses of (2.7.5), for all n ⩾ 0, there exists an integer k (depending on
n) and a surjective homomorphism OXk → OX (n).
It suffices (2.7.2) to show that, for suitable k, there is a (TN)-surjective homomorphism u of
degree 0 from the graded product S-module Sk to S(n). But (S(n))0 = Sn , and, by hypothesis,
Sh = S1h for all h > 0, and so SSn = Sn + Sn+1 + . . .. Since Sn is an S0 -module of finite type ((2.1.5)
and (2.1.6, i)), consider a system of generators ( ai )1⩽i⩽k of this module; consider the homomorphism
u that sends ai to the i-th element of the canonical basis of Sk (1 ⩽ i ⩽ k); since Coker u can then be
identified with (S(n))−n + . . . + (S(n))−1 , u is indeed the desired homomorphism. □
Corollary (2.7.10). — Under the hypotheses of (2.7.5), let F be a quasi-coherent OX -module of finite type.
Then there exists an integer n0 such that, for all n ⩾ n0 , F is isomorphic to a quotient of an OX -module of
the form (OX (−n))k (for some k depending on n).
Proposition (2.7.11). — Suppose that the hypotheses of (2.7.5) are satisfied, and let M be a graded S-module.
Then:
(i) The canonical homomorphism e α:M e → (Γ• ( M
e ))e is an isomorphism.
(ii) Let G be a quasi-coherent sub-OX -module of M,
e and let N be the graded sub-S-module of M given
by the inverse image of Γ• (G ) under α. Then Ne = G (where N e is identified, by (2.5.4), with a
sub-OX -module of M).
e
P ROOF. Since β : (Γ• ( Me ))e → M e is an isomorphism (by (2.7.5)), e α is the inverse isomorphism
(by (2.6.5.1)), whence (i). Let P be the graded submodule α( M) of Γ• ( M ); since M
e e is an exact functor
(2.5.4), the image of Me under e e and so, by (i), Pe = (Γ• ( M
α is equal to P, e ))e. Set Q = Γ• (G ) ∩ P; by the
above, and by (2.5.4), we have that Q e = (Γ• (G ))e, and so the restriction of β to Q e is an isomorphism
from this OX -module to G by (2.7.5). But, by the definition of N, and by (2.5.4), the restriction of the
isomorphism e α to N
e is an isomorphism from N e to Q,
e whence the conclusion, by (2.6.5.1). □
2. HOMOGENEOUS PRIME SPECTRA 228
(2.8.1). Let S and S′ be positively graded rings, and ϕ : S′ → S a homomorphism of graded rings.
We denote by G (ϕ) the open subset of X = Proj(S) given by the complement of V+ (ϕ(S+ ′ )), or,
′ ′
equivalently, the union of the D+ (ϕ( f )) where f runs over the set of homogeneous elements of
S+ ′ . The restriction to G ( ϕ ) of the continuous map a ϕ from Spec( S ) to Spec( S′ ) (I, 1.2.1) is thus a
continuous map from G (ϕ) to Proj(S′ ), which we again denote, with an abuse of language, by a ϕ. If
f ′ ∈ S+
′ is homogeneous, then
a −1
(2.8.1.1) ϕ ( D+ ( f ′ )) = D+ (ϕ( f ′ ))
taking into account the fact that a ϕ sends G (ϕ) to Proj(S′ ), as well as (I, 1.2.2.2). The homomorphism
ϕ also canonically defines (with the same notation) a homomorphism of graded rings S′f ′ → S f ,
whence, by restriction to the degree 0 elements, a homomorphism S(′ f ′ ) → S( f ) , which we denote II | 42
by ϕ( f ) ; there is a corresponding (I, 1.6.1) morphism of affine schemes ( a ϕ( f ) , ϕ
e( f ) ) : Spec(S( f ) ) →
Spec(S(′ f ′ ) ). If we canonically identify Spec(S( f ) ) with the scheme induced by Proj(S) on D+ ( f )
(2.3.6), then we have defined a morphism Φ f : D+ ( f ) → D+ ( f ′ ), and a ϕ( f ) is identified with the
restriction of a ϕ to D+ ( f ). It is also immediate that, if g′ is another homogeneous element of S+
′ , and
′
g = ϕ( g ), then the diagram
Φf
D+ ( f ) / D+ ( f ′ )
O O
D+ ( f g ) / D+ ( f ′ g ′ )
Φfg
ω f ′ g′ , f ′ ω f g, f
S(′ f ′ g′ ) /S
ϕ( f g) ( f g)
commutes. Taking the definition of G (ϕ), along with (2.3.3.2), we thus see that:
Proposition (2.8.2). — Given a homomorphism of graded rings ϕ : S′ → S, there exists exactly one
e) from the induced prescheme G (ϕ) to Proj(S′ ) (said to be associated to ϕ, and denoted by
morphism ( a ϕ, ϕ
Proj(ϕ)) such that, for every homogeneous element f ′ ∈ S+′ , the restriction of this morphism to D ( ϕ ( f ′ ))
+
agrees with the morphism associated to the homomorphism S(′ f ′ ) → S(ϕ( f ′ )) corresponding to ϕ.
ϕ( f )
(2.8.2.1) S(′ f ′ ) /S
(f)
∼ ∼
S′
(d)
/ ( f ′ − 1) S ′
(d) / S ( d ) / ( f − 1) S ( d )
P ROOF. Claim (i) is an immediate consequence of (2.8.2) and (2.8.1.1). Claim (ii) follows since, if
Ker(ϕ) is nilpotent, then, for every homogeneous f ′ in S+ ′ , we immediately see that Ker( ϕ ) (with
f
′
f = ϕ( f )) is nilpotent, and thus so too is Ker(ϕ( f ) ); we then apply (2.8.2) and (I, 1.2.7) □
2. HOMOGENEOUS PRIME SPECTRA 229
We note that there are, in general, morphisms from Proj(S) to Proj(S′ ) that are not affine, and II | 43
that thus do not come from homomorphisms of graded rings S′ → S; an example is the structure
morphism Proj(S) → Spec( A) when A is a field (Spec( A) thus being identified with Proj( A[ T ]),
cf. (3.1.7)); indeed, this follows from (I, 2.3.2).
(2.8.4). Let S′′ be another positively graded ring, and ϕ′ : S′′ → S′ a homomorphism of graded
rings, and set ϕ′′ = ϕ ◦ ϕ′ . By (2.8.1.1) and the formula a ϕ′′ = ( a ϕ′ ) ◦ ( a ϕ), we immediately see
that G (ϕ′′ ) ⊂ G (ϕ), and that, if Φ, Φ′ , and Φ′′ are the morphisms associated to ϕ, ϕ′ , and ϕ′′
(respectively), then Φ′′ = Φ′ ◦ (Φ| G (ϕ′′ )).
(2.8.5). Suppose that S (resp. S′ ) is a graded A-algebra (resp. a graded A′ -algebra), and let ψ : A′ → A
be a ring homomorphism such that the diagram
/A
ψ
(2.8.5.1) A′
S′ /S
ϕ
commutes. We can then consider G (ϕ) (resp. Proj(S′ )) as a scheme over Spec( A) (resp. Spec( A′ )); if
Φ (resp. Ψ) is the morphism associated to ϕ (resp. ψ), then the diagram
Φ / Proj(S′ )
G (ϕ)
Spec( A) / Spec( A′ )
Ψ
/A
ψ
A′
S(′ f ′ ) /S
ϕ( f ) (f)
commutes.
(2.8.6). Now let M be a graded S-module, and consider the S′ -module M[ϕ] , which is evidently
graded. Let f ′ be homogeneous in S+ ′ , and let f = ϕ ( f ′ ); we know (0, 1.5.2) that there is a
∼
canonical isomorphism ( M[ϕ] ) f ′ −
→ ( M f )[ϕ f ] , and it is immediate that this isomorphism preserves
∼
degree, whence a canonical isomorphism ( M[ϕ] )( f ′ ) −
→ ( M( f ) )[ϕ( f ) ] . To this isomorphism, there
∼
canonically corresponds an isomorphism of sheaves ( M )e| D+ ( f ′ ) − → (Φ )∗ (e( M)| D+ ( f )) ((2.5.2)
[ϕ] f
and (I, 1.6.3)). Furthermore, if g′ is another homogeneous element of S+
′ , and g = ϕ ( g′ ), then the | 43
diagram
∼ /
(M ) ′ (M )
[ϕ] ( f ) ( f ) [ϕ( f ) ]
( M[ ϕ ] ) ( f ′ g ′ )
∼ / (M
( f g) )[ϕ( f g) ]
Proposition (2.8.7). — There exists a canonical functorial isomorphism from the OX ′ -module ( M[ϕ] )e to
the OX ′ -module Φ∗ ( M
e | G (ϕ)).
We thus immediately deduce a canonical functorial map from the set of ϕ-morphisms M′ → M
from a graded S′ -module to the graded S-module M, to the set of Φ-morphisms M f′ → M e | G ( ϕ ).
′′ ′′
With the notation of (2.8.4), if M is a graded S -module, then, to the composition of a ϕ-morphism
M′ → M and a ϕ′ -morphism M′′ → M′ , canonically corresponds the composition of M f′ G (ϕ′ ) →
M ′′
e | G (ϕ ) and g′′
M →M ′ ′
f | G ( ϕ ).
Proposition (2.8.8). — Under the hypotheses of (2.8.1), let M′ be a graded S′ -module. Then there exists
a canonical functorial homomorphism ν from the (OX | G (ϕ))-module Φ∗ ( M f′ ) to the (OX | G (ϕ))-module
′ ′ ′
( M ⊗S′ S)e| G (ϕ). If the ideal S+ is generated by S1 , then ν is an isomorphism.
P ROOF. Indeed, for f ′ ∈ Sd′ (d > 0), we define a canonical functorial homomorphism of S( f ) -
modules (where f = ϕ( f ′ ))
(2.8.8.1) ν f : M(′ f ′ ) ⊗S′ S( f ) −→ ( M′ ⊗S′ S)( f )
( f ′)
by composing the homomorphism M(′ f ′ ) ⊗S′ S( f ) → M′f ′ ⊗S′ ′ S f and the canonical isomorphism
( f ′) f
∼
M′f ′ ⊗S′ ′ S f −
→ ( M′ ⊗S′ S) f (0, 1.5.4), and noting that the latter preserves degrees. We can immedi-
f
ately verify the compatibility of ν f with the restriction operators from D+ ( f ) to D+ ( f g) (for any
g ′ ∈ S+
′ and g = ϕ ( g′ )), whence the definition of the homomorphism
ν : Φ∗ ( M
f′ ) −→ ( M′ ⊗S′ S)e| G (ϕ)
taking (I, 1.6.5) into account. To prove the second claim, it suffices to show that ν f is an isomorphism
for all f ′ ∈ S1 , since G (ϕ) is then a union of the D+ (ϕ( f ′ )). We first define a Z-bilinear Mm
′ ×S →
n
′ ′ ′ ′ m n
M( f ′ ) ⊗S′ ′ S( f ) by sending ( x , s) to the element ( x / f ) ⊗ (s/ f ) (with the convention that
(f )
m −m
x ′ / f ′ is f ′ x ′ /1 when m < 0). We claim that, in the proof of (2.5.13), this map gives rise to a II | 45
di-homomorphism of modules
η f : M′ ⊗S′ S −→ M(′ f ′ ) ⊗S( f ′ ) S( f ) .
r
Furthermore, if, for r > 0, we have f r ∑i ( xi′ ⊗ si ) = 0, then this can also be written as ∑i ( f ′ xi′ ⊗ si ) =
r m +r
0, whence, by (0, 1.5.4), ∑i ( f ′ xi / f ′ i ) ⊗ (si / f ni ) = 0, i.e. η f (∑i xi ⊗ yi )0 =, which proves that
η ′f
η f factors as M′ ⊗S′ S → ( M′ ⊗S′ S) f −→ M(′ f ′ ) ⊗S′ S( f ) ; we finally can prove that η ′f and ν f are
( f ′)
inverse isomorphisms to one another.
In particular, it follows from (2.1.2.1) that we have a canonical homomorphism
∼
(2.8.8.2) Φ∗ (OX ′ (n)) −
→ OX (n)| G (ϕ)
for all n ∈ Z. □
(2.8.9). Let A and A′ be rings, and ψ : A′ → A a ring homomorphism, defining a morphism
Ψ : Spec( A) → Spec( A′ ). Let S′ be a positively graded A′ -algebra, and set S = S′ ⊗ A′ A, which
is evidently an A-algebra graded by the Sn′ ⊗ A′ A; the map ϕ : s′ → s′ ⊗ 1 is then a graded ring
homomorphism that makes the diagram (2.8.5.1) commute. Since S+ is here the A-module generated
by ϕ(S+ ′ ), we have G ( ϕ ) = Proj( S ) = X; whence, setting X ′ = Proj( S′ ), we have the commutative
diagram
Φ / X′
(2.8.9.1) X
p
Y / Y′
Ψ
Now let M′ be a graded S′ -module, and set M = M′ ⊗ A′ A = M′ ⊗S′ S. Under these conditions:
Proposition (2.8.10). — The diagram (2.8.9.1) identifies the scheme X with the product X ′ ×Y ′ Y; further-
more, the canonical homomorphism ν : Φ∗ ( M
f′ ) → Me (2.8.8) is an isomorphism.
2. HOMOGENEOUS PRIME SPECTRA 231
P ROOF. The first claim will be proven if we show that, for every homogenous f ′ in S+ ′ , setting
(I, 3.2.6.2); in other words, it suffices (I, 3.2.2) to prove that S( f ) is canonically identified with
∼ ∼
→ S′f ′ ⊗ A′ A, which is immediate by the existence of the canonical isomorphism S f −
Sf − → S′f ′ ⊗ A′ A
′
that preserves degrees (0, 1.5.4). The second claim then follows from the fact that M( f ′ ) ⊗S′ ′ S( f )
(f )
can be identified, by the above, with M(′ f ′ ) ⊗ A′ A, and this can be identified with M( f ) , since M f is
canonically identified with M′f ′ ⊗ A′ A by an isomorphism that preserves degrees. □
Corollary (2.8.11). — For every integer n ∈ Z, Me (n) can be identified with Φ∗ ( M f′ (n) ⊗Y ′ OY ;
f′ (n)) = M
∗
in particular, OX (n) = Φ (OX ′ (n)) = OX ′ (n) ⊗Y ′ OY .
P ROOF. This follows from (2.8.10) and (2.5.15). □
II | 46
(2.8.12). Under the hypotheses of (2.8.9), for f ′ ∈ Sd′ (d > 0) and f = ϕ( f ′ ), the diagram
(2.8.12.1) M(′ f ′ )
∼ / M ′ ( d ) / ( f ′ − 1) M ′ ( d )
M( f )
∼ / M ( d ) / ( f − 1) M ( d )
M0 /M = Γ ( D+ ( f ) , M
e)
(f)
thus commutes; we immediately conclude from this remark, and from the definition of the homo-
morphism α (2.6.2), that the diagram
α M′
(2.8.13.1) M′ / Γ• ( M
f′ )
M / Γ• ( Me)
αM
(Γ• (F ))e /F
βF
commutes (the vertical arrow on the right being the canonical Φ-morphism F ′ → Φ∗ (F ′ ) = F ).
2. HOMOGENEOUS PRIME SPECTRA 232
II | 47
(2.8.14). Still keeping the notation and hypotheses of (2.8.9), let N ′ be another graded S′ -module,
and let N = N ′ ⊗ A′ A. It is immediate that the canonical di-homomorphisms M′ → M and N ′ → N
give a di-homomorphism M′ ⊗S′ N ′ → M ⊗S N (with respect to the canonical ring homomorphism
S′ → S), and thus also an S-homomorphism ( M′ ⊗S′ N ′ ) ⊗ A′ A → M ⊗S N of degree 0, to which
corresponds (taking (2.8.10) into account) a homomorphism of OX -modules
(2.8.14.2) Φ∗ ( M
f′ ⊗O N f′ ) ∼ /M
e ⊗O Ne = Φ∗ ( M
f′ ) ⊗O Φ∗ ( N
f′ )
X′ X X
Φ∗ (λ) λ
Φ∗ (( M′ ⊗S′ N ′ )e) / ( M ⊗S N )e
commutes, with the first line being the canonical isomorphism (0, 4.3.3). If the ideal S+′ is generated
′
by S1 , then it is clear that S+ is generated by S1 , and the two vertical arrows of (2.8.14.2) are then
isomorphisms (2.5.13); it is thus also the case for (2.8.14.1).
We similarly have a canonical di-homomorphism HomS′ ( M′ , N ′ ) → HomS ( M, N ) by sending a
homomorphism u′ of degree k to the homomorphism u′ ⊗ 1, which is also of degree k; from this, we
again deduce a S-homomorphism of degree 0
(HomS′ ( M′ , N ′ )) ⊗ A′ A −→ HomS ( M, N )
whence a homomorphism of OX -modules
Φ∗ (µ) µ
Φ∗ (H omOX′ ( M
f′ , N
f′ )) / H omO ( M,
e Ne)
X
commutes (the second horizontal line being the canonical homomorphism (0, 4.4.6)).
II | 48
(2.8.15). With the notation and hypotheses of (2.8.1), it follows from (2.4.7) that we do not change the
morphism Φ, up to isomorphism, when we replace S0 and S0′ by Z, and ϕ0 by the identity map, and
(d)
thus when we replace S and S′ by S(d) and S′ (respectively) (d > 0), and ϕ by its restriction ϕ(d) to
S(d) .
(ii) Suppose further that the ideal S+ is generated by a finite number of homogeneous elements of
degree 1. Let X ′ be a closed subscheme of X defined by a quasi-coherent sheaf of ideals J of OX . Let
J be the graded ideal of S given by the inverse image of Γ• (J ) under the canonical homomorphism
α : S → Γ• (OX ) (2.6.2), and set S′ = S/J. Then X ′ is the subscheme associated to the closed
immersion Proj(S′ ) → X corresponding to the canonical homomorphism of graded rings S → S′ .
P ROOF.
(i) We can suppose that ϕ is surjective (2.9.1). Since, by hypothesis, ϕ(S+ ) generates S+ ′ , we
′
have G (ϕ) = Proj(S ). Now, the second claim can be checked locally on X; so let f be a
homogeneous element of S+ , and set f ′ = ϕ( f ). Since ϕ is a surjective homomorphism of
graded rings, we immediately see that ϕ( f ′ ) : S( f ) → S(′ f ′ ) is surjective, and that its kernel
is J( f ) , which proves (i) (I, 4.2.3).
(ii) By (i), we are led to proving that the homomorphism ej : e J → OX induced by the canonical
injection j : J → S is an isomorphism from J to J , which follows from (2.7.11).
e
□
We note that J is the largest of the graded ideals J′ of S such that ej(Je′ ) = J , since we can
immediately show, using the definitions (2.6.2), that this equation implies that α(J′ ) ⊂ Γ• (J ).
Corollary (2.9.3). — Suppose that the hypotheses of (2.9.2, (i)) are satisfied, and further that the ideal S+ is
generated by S1 ; then Φ∗ ((S(n))e) is canonically isomorphic to (S′ (n))e for all n ∈ Z, and so Φ∗ (F (n))
is canonically isomorphic to Φ∗ (F )(n) for every OX -module F .
P ROOF. This is a particular case of (2.8.8), taking (2.5.10.2) into account. □
Corollary (2.9.4). — Suppose that the hypotheses of (2.9.2, (ii)) are satisfied. For the closed sub-prescheme
X ′ of X to be integral, it is necessary and sufficient for the graded ideal J to be prime in S.
II | 49
P ROOF. Since X ′ is isomorphic to Proj(S/J), the condition is sufficient by (2.4.4). To see that it
is necessary, consider the exact sequence 0 → J → OX → OX /J , which gives the exact sequence
0 −→ Γ• (J ) −→ Γ• (OX ) −→ Γ• (OX /J ).
It suffices to prove that, if f ∈ Sm and g ∈ Sn are such that the image in Γ• (OX /J ) of αn+m ( f g)
is zero, then the image of either αm ( f ) or αn ( g) is zero. But, by definition, these images are sections
of invertible (OX /J )-modules L = (OX /J )(m) and L ′ = (OX /J )(n) over the integral scheme
X ′ ; the hypothesis implies that the product of these two sections is zero in L ⊗ L ′ ((2.9.3) and
(2.5.14.1)), and so one of them is zero by (I, 7.4.4). □
Corollary (2.9.5). — Let A be a ring, M an A-module, S a graded A-algebra generated by the set S1 of
homogeneous elements of degree 1, u : M → S1 a surjective homomorphism of A-modules, and u : S( M) → S
the homomorphism (of A-algebras) from the symmetric algebra S( M) of M to S that extends u. Then the
morphism corresponding to u is a closed immersion of Proj(S) into Proj(S( M )).
P ROOF. Indeed, u is surjective by hypothesis, and so it suffices to apply (2.9.2) □
the graded S -module such that (M (n))k = Mn+k for all k ∈ Z; if S and M are quasi-coherent,
then M (n) is a quasi-coherent graded S -module (I, 9.6.1).
We say that M is a graded S -module of finite type (resp. admitting a finite presentation) if, for
all y ∈ Y, there exists an open neighbourhood U of y, along with integers ni (resp. integers mi
and n j ) such that there is a surjective degree 0 homomorphism ri=1 (S (ni )|U ) → M |U (resp.
L
such that M |U is isomorphic to the cokernel of a degree 0 homomorphism ri=1 (S (mi )|U ) →
L
Ls
j1 (S ( n J )|U )).
Let U be an affine open of Y, with ring A = Γ(U, OY ); by hypothesis, the graded (OY |U )-algebra
S |U is isomorphic to S, e where S = Γ(U, S ) is a graded A-algebra (I, 1.4.3); set XU = Proj(Γ(U, S )). II | 50
Let U ′ ⊂ U be another affine open of Y, with ring A′ , and let j : U ′ → U be the canonical injection,
which corresponds to the restriction homomorphism A → A′ ; we have that S |U ′ = j∗ (S |U ),
and so S′ = Γ(U ′ , S ) is canonically identified with XU ×U U ′ , and thus also with f U−1 (U ′ ), where
we denote by f U the structure morphism XU → U (I, 4.4.1). We denote by σU ′ ,U the canonical
∼
isomorphism f U−1 (U ′ ) −→ XU ′ thus defined, and by ρU ′ ,U the open immersion XU ′ → XU obtained
by composing σU−′1,U with the canonical injection f U−1 (U ′ ) → XU . It is immediate that, if U ′′ ⊂ U ′ is
another affine open of Y, then ρU ′′ ,U = ρU ′′ ,U ′ ◦ ρU ′ ,U .
Proposition (3.1.2). — Let Y be a prescheme. For every quasi-coherent positively graded OY -algebra, there
exists exactly one (up to Y-isomorphism) prescheme X over Y satisfying the following property: if f : X → Y
is the structure morphism, then, for every affine open U of Y, there exists an isomorphism ηU from the
induced prescheme f −1 (U ) to XU = Proj(Γ(U, S )) such that, if V is another affine open of Y that is
contained in U, then the diagram
/ XV
ηV
(3.1.2.1) f −1 ( V )
ρV,U
f −1 (U ) / XU
ηV
commutes.
P ROOF. Given affine opens U and V of Y, let XU,V be the prescheme induced on f U−1 (U ∩ V ) by
∼
XU ; we are going to define a Y-isomorphism θU,V : XV,U − → XU,V . For this, we consider an affine
open W ⊂ U ∩ V: by composing the isomorphisms
−1
σW,U σW,V
f U−1 (W ) −−→ XW −−→ f V−1 (W ),
we obtain an isomorphism τW , and we immediately see that, if W ′ ⊂ W is an affine open, then τW ′
is the restriction of τW to f U−1 (W ′ ); the τW are thus indeed the restrictions of a Y-isomorphism θV,U .
Further, if U, V, and W are affine open subsets of Y, and θU,V ′ ′
, θV,W ′
, and θU,W the restrictions of θU,V ,
θV,W , and θU,W (respectively) to the inverse images of U ∩ V ∩ W in XV , XW , and XW (respectively),
then it follows from the above definitions that we have θU,V ′ ′
◦ θV,W ′
= θU,W . The existence of
some X satisfying the properties in the statement thus follows from (I, 2.3.1); its uniqueness up to
Y-isomorphism is trivial, taking (3.1.2.1) into account. □
(3.1.3). We say that the prescheme X defined in (3.1.2) is the homogeneous spectrum of the quasi-
coherent graded OY -algebra S , and we denote it by Proj(S ). It is immediate that Proj(S ) is
separated over Y ((2.4.2) and (I, 5.5.5)); if S is an OY -algebra of finite type (I, 9.6.2), then Proj(S ) is of
finite type over Y ((2.7.1, (ii)) and (I, 6.3.1)).
If f is the structure morphism X → Y, then it is immediate that, for every prescheme induced by
Y on an open subset U of Y, f −1 (U ) can be identified with the homogeneous spectrum Proj(S |U ).
Proposition (3.1.4). — Let f ∈ Γ(Y, Sd ) for d > 0. Then there exists an open subset X f of the
underlying space of X = Proj(S ) that satisfies the following property: for every affine open U of Y, we have
X f ∩ ϕ−1 (U ) = D+ ( f |U ) in ϕ−1 (U ) identified with XU = Proj(Γ(U, S )), where ϕ denotes the structure
morphism X → Y. Furthermore, the prescheme induced on X f is affine over Y, and is canonically isomorphic II | 51
to Spec(S (d) /( f − 1)S [d] ) (1.3.1).
3. HOMOGENEOUS SPECTRUM OF A SHEAF OF GRADED ALGEBRAS 235
P ROOF. We have f |U ∈ Γ(U, Sd ) = (Γ(U, S ))d . If U and U ′ are affine opens of Y such that
U′ ⊂ U, then f |U ′ is the image of f |U under the restriction homomorphism
Γ(U, S ) −→ Γ(U ′ , S )
and so D+ ( f |U ′ ) is equal (with the notation of (3.1.1)) to the prescheme induced on the inverse
−1
image ρU ′ ,U ( D+ ( f |U )) in XU ′ (2.8.1); whence the first claim. Furthermore, the prescheme induced
on D+ ( f |U ) by XU is canonically identified with Spec((Γ(U, S )) f |U ), with these identifications
being compatible with the restriction homomorphisms (2.8.1); the second claim then follows from
(2.2.5) and from the commutativity of the diagram (2.8.2.1). □
We also say that X f (as an open subset of the underlying space X) is the set of x ∈ X where f
does not vanish.
Corollary (3.1.5). — If f ∈ Γ(Y, Sd ) and g ∈ Γ(Y, Se ), then
(3.1.5.1) X f g = X f ∩ Xg .
P ROOF. It suffices to consider the intersection of the two sets with a set ϕ−1 (U ), where U is an
affine open in Y, and to then apply formula (2.3.3.2). □
Corollary (3.1.6). — Let ( f α ) be a family of sections of S over Y such that f α ∈ Γ(Y, Sdα ); if the sheaf
of ideals of S generated by this family (0, 5.1.1) contains all the Sn starting from a certain rank, then the
underlying space X is the union of the X f α .
P ROOF. For every affine open U of Y, ϕ−1 (U ) is the union of the X f α ∩ ϕ−1 (U ) (2.3.14). □
Corollary (3.1.7). — Let A be a quasi-coherent OY -algebra; set
S = A [ T ] = A ⊗Z Z[ T ]
where T is an indeterminate (and Z and Z[ T ] are considered as simple sheaves over Y). Then X = Proj(S )
is canonically identified with Spec(A ). In particular, Proj(OY [ T ]) is identified with Y.
P ROOF. By applying (3.1.6) to the unique section f ∈ Γ(Y, S ) that is equal to T at each point
of Y< we see that X f = X. Further, here we have d = 1, and S (1) /( f − 1)S (1) = S /( f − 1)S is
canonically isomorphic to A , whence the corollary (1.2.2). □
Let g ∈ Γ(Y, OY ); if we take S = OY [ T ], then g ∈ Γ(Y, S0 ); let
h = gT ∈ Γ(Y, S1 ).
If X = Proj(S ), then the canonical identification defined in (3.1.7) identifies Xh with the open subset
Yg of Y (in the sense of (0, 5.5.2)): indeed, we can restrict to the case where Y = Spec( A) is affine,
and everything then reduces (taking (2.2.5) into account) to the fact that the ring of fractions A g is
canonically identified with A[ T ]/( gT − 1) A[ T ] (0, 1.2.3).
Proposition (3.1.8). — Let S be a quasi-coherent positively-graded OY -algebra. Then
(i) For all d > 0, there exists a canonical Y-isomorphism from Proj(S ) to Proj(S (d) ). II | 52
(ii) Let S ′ be the graded OY -algebra given by the direct sum of OY with the Sn (for n ⩾ 0); then
Proj(S ′ ) and Proj(S ) are canonically Y-isomorphic.
(iii) Let L be an invertible OY -module (0, 5.4.1), and let S(L ) be the graded OY -algebra given by
the direct sum of the Sd ⊗ L ⊗d (for d ⩾ 0); then Proj(S ) and Proj(S(L ) ) are canonically
Y-isomorphic.
P ROOF. In each of the three cases, it suffices to define the isomorphism locally on Y, since the
verification of compatibility with the restriction operations from one open subset to a smaller one is
trivial. We can thus suppose that Y is affine, and then (i) follows from (2.4.7, (i)), and (ii) follows from
(2.4.8). As for (iii), if we further suppose that L is isomorphic to OY (which we are allowed to do,
since the question is local on Y), then the isomorphism between Proj(S ) and Proj(S(L ) ) is evident;
to define a canonical isomorphism, let Y = Spec( A) and S = S, e where S is a graded A-algebra, and
let c be a generator of the free A-module L such that L = L; then, for all n > 0, xn 7→ xn ⊗ c⊗n is
e
an A-isomorphism from Sn to Sn ⊗ L⊗n , and these A-isomorphisms define an A-isomorphism of
3. HOMOGENEOUS SPECTRUM OF A SHEAF OF GRADED ALGEBRAS 236
x ∈ Snd , we have that ( x ⊗ cnd )/( f ⊗ cd )n = ( x ⊗ (εc)nd )/( f ⊗ (εc)d )n for every invertible element
ε ∈ A, which shows that the isomorphism S( f ) → (S( L) )( f ⊗cd ) induced from ϕc is independent of the
generator c of L considered, and thus finishes the proof. □
(3.1.9). Recall ((0, 4.1.3) and (I, 1.3.14)) that, for the quasi-coherent graded OY -algebra S to be
generated by the OY -module S1 , it is necessary and sufficient for there to exist an affine open cover
(Uα ) of Y such that the graded algebra Γ(Uα , S ) over Γ(Uα , S0 ) is generated by the set Γ(Uα , S1 )
of its homogeneous elements of degree 1. For every open V of Y, S |V is then generated by the
(OY |V )-module S1 |V.
Proposition (3.1.10). — Suppose that there exists a finite affine open cover (Ui ) of Y such that each graded
algebra Γ(Ui , S ) is of finite type over Γ(Ui OY ). Then there exists d > 0 such that S (d) is generated by Sd ,
with Sd an OY -module of finite type.
P ROOF. Indeed, it follows from (2.1.6, (v)) that, for each i, there exists an integer mi such that
Γ(Ui , Snmi ) = (Γ(Ui , Smi ))n for all n > 0; it suffices to take d to be a common multiple of all the mi ,
taking (2.1.6, (i)) into account. □
Corollary (3.1.11). — Under the hypotheses of (3.1.10), Proj(S ) is Y-isomorphic to a homogeneous
spectrum Proj(S ′ ), where S ′ is a graded OY -algebra generated by S1′ , with S1′ an OY -module of finite type.
P ROOF. It suffices to take S ′ = S (d) , where d satisfies the property of (3.1.10), and to then
apply (3.1.8, (i)) □
(3.1.12). If S is a quasi-coherent positively-graded OY -algebra, we know (I, 5.1.1) that its nilradical
N is a quasi-coherent OY -module; we say that N+ = N ∩ S+ is the nilradical of S+ ; this is a quasi-
coherent graded S0 -module, since we can immediately reduce to the case where Y is affine, and the
proposition then follows from (2.1.10). For all y ∈ Y, (N+ )y is then the nilradical of (S+ )y = (Sy )+
(I, 5.1.1). We say that the graded OY -algebra S is essentially reduced if N+ = 0, which is equivalent
to saying that Sy is an essentially reduced graded Oy -algebra for all y ∈ Y. For every graded II | 53
OY -algebra S , S /N+ is essentially reduced.
We say that S is integral if, for all y ∈ Y, Sy is an integral ring and, furthermore, (Sy )+ =
(S+ )y ̸= 0.
Proposition (3.1.13). — Let S be a positively-graded OY -algebra. If X = Proj(S ), then the Y-scheme
Xred is canonically isomorphic to Proj(S /N+ ); in particular, if S is essentially reduced, then X is reduced.
P ROOF. The fact that X ′ = Proj(S /N+ ) is reduced follows immediately from (2.4.4, (i)), since
−1
the property is local; further, for every affine open U ⊂ Y, ϕ′ (U ) is equal to (ϕ−1 (U ))red (where we
′ ′
denote by ϕ and ϕ the structure morphisms X → Y and X → Y, respectively); we immediately see
−1
that the canonical U-morphisms ϕ′ (U ) → ϕ−1 (U ) are compatible with the restriction operations,
and thus define a closed immersion X ′ → X that is a homeomorphism of the underlying spaces;
whence the conclusion (I, 5.1.2). □
Proposition (3.1.14). — Let Y be an integral prescheme, and S a quasi-coherent graded OY -algebra such
that S0 = OY .
(i) If S is integral (3.1.12), then X = Proj(S ) is integral, and the structure morphism ϕ : X → Y is
dominant.
(ii) Suppose furthermore that S is essentially reduced. Then, conversely, if X is integral and ϕ is
dominant, then S is integral.
P ROOF.
(i) If (Uα ) is a base of Y consisting of non-empty affine opens, then it suffices to prove the
proposition in the case where Y is replaced by one of the Uα , and S by S |Uα : indeed, one
one hand it will follow that the underlying space ϕ−1 (Uα ) are irreducible opens (and thus
non-empty) of X such that ϕ−1 (Uα ) ∩ ϕ−1 (Uβ ) ̸= ∅ for any pair of indices (since Uα ∩ Uβ
contains some Uγ ), and so X is irreducible (0, 2.1.4); on the other hand, X will be reduced,
since this is a local property, and so X will indeed be integral, with ϕ( X ) dense in Y.
3. HOMOGENEOUS SPECTRUM OF A SHEAF OF GRADED ALGEBRAS 237
So suppose that Y = Spec( A), where A is integral (I, 5.1.4), and that S = S, e where S
is a graded A-algebra; the hypothesis is that, for every y ∈ Y, Sy = Sy is an integral graded
e
ring such that (Sy )+ ̸= 0. It suffices to prove that S is an integral ring, since then we will
have that S+ ̸= 0, and we can then apply (2.4.4, (ii)). But let f , g ̸= 0 be elements of S, and
suppose that f g = 0; for all y ∈ Y we then have that ( f /1)( g/1) = 0 in Sy , and so f /1 = 0
or g/1 = 0 by hypothesis. Suppose, for example, that f /1 = 0 in Sy ; this implies that there
exists a ∈ A such that a ̸∈ jy and a f = 0; we then have, for all z ∈ Y, that ( a/1)( f /1) = 0
in the integral ring Sz , and since a/1 ̸= 0 (since A is integral), f /1 = 0, which implies that
f = 0.
(ii) Since the question is local on Y, we can again suppose that Y = Spec( A), with A integral,
and that S = S. e By hypothesis, for all y ∈ Y, (Sy )+ does not contain any non-zero nilpotent
elements, and the same is true of (S0 )y = Ay by hypothesis; so Sy is a reduced ring for all
y ∈ Y, and we thus conclude first of all that S itself is reduced (I, 5.1.1). The hypothesis that
X is integral implies that S is essentially integral (2.4.4, (ii)), and everything then reduces to
showing that the annihilator J of S+ in A = S0 is just 0 (2.1.11). If this were not the case,
we would have that (Sh )+ = 0 for some h ̸= 0 in J, and thus (3.1.1) that ϕ−1 ( D (h)) = ∅,
and ϕ( X ) would not be dense in Y, contradicting the hypothesis (since D (h) ̸= ∅, since h
is not nilpotent). II | 54
□
We say that Mfis the OX -module associated to the quasi-coherent graded S -module M .
Proposition (3.2.3). — Let M be a quasi-coherent graded S -module, and let f ∈ Γ(Y, Sd ) (for d > 0). If
ξ f is the canonical isomorphism from X f to the Y-prescheme Z f = Spec(S (d) /( f − 1)S (d) ) (3.1.4), then
(ξ f )∗ (Mf| X f ) is the OZ -module (M (l ) /( f − 1)M (d) ) (1.4.3).
f
P ROOF. Since the question is local on Y, we can immediately reduce to (2.2.5), taking into
account the commutativity of the diagram in (2.8.12.1). □
Proposition (3.2.4). — The OX -module M fis an exact additive covariant functor in M , from the category
of quasi-coherent graded S -modules to the category of quasi-coherent OX -modules, that commutes with
inductive limits and direct sums.
P ROOF. Since the question is local on Y, we can reduce to (I, 1.3.11), (I, 1.3.9), and (2.5.4). □
(where the right-hand side is defined as in (0, 4.3.5)). It suffices to verify this formula in the case
where Y = Spec( A) is affine, S = S, e with S a graded A-algebra, M = M, e with M a graded
S-module, and I = I, with I an ideal of A. For every homogeneous element f of S+ , the restriction II | 55
to D+ ( f ) = Spec(S( f ) ) of the left-hand side of (3.2.4.1) can be associated with (IM )( f ) = I · M( f ) ,
and the same is true of the restriction of the right-hand side, given (I, 1.3.13) and (i, 1.6.9).
Proposition (3.2.5). — Let f ∈ Γ(Y, Sd ) (for d > 0). On the open subset X f , the (OX | X f )-module
(S (nd))eX f is canonically isomorphic to OX | X f for all n ∈ Z. In particular, if the OY -algebra S is
generated by S1 (3.1.9), then the OX -modules (S (n))e are invertible for all n ∈ Z.
P ROOF. Indeed, for every affine open U of Y, we defined in (2.5.7) a canonical isomorphism
from (S (nd))e|( X f ∩ ϕ−1 (U )) to OX |( X f ∩ ϕ−1 (U )), taking (3.1.4) into account (where ϕ is the
structure morphism X → Y); it is immediate that these isomorphisms are compatible with the
restriction from U to an affine open U ′ ⊂ U, whence the first claim. To prove the second, it suffices to
note that, if S is generated by S1 , then there is a cover (Uα ) of Y by affine opens such that Γ(Uα , S )
is generated by Γ(Uα , S )1 = Γ(Uα , S1 ); we can then apply the result of (2.5.9), since the property of
being invertible is local. □
We again set, for all n ∈ Z,
(3.2.5.1) OX (n) = (S (n))e
and, for all OX -modules F ,
(3.2.5.2) F ( n ) = F ⊗O X O X ( n ).
It follows immediately from these definitions that, for every open subset U of Y, we have
((S |U )(n))e = OX (n)| f −1 (U )
where f is the structure morphism X → Y.
Proposition (3.2.6). — Let M and N be quasi-coherent graded S -modules. Then there exists a canonical
functorial (in M and N ) homomorphism
(3.2.6.1) λ:Mf⊗O Nf−→ (M ⊗S N )e
X
Remarks (3.2.9). —
3. HOMOGENEOUS SPECTRUM OF A SHEAF OF GRADED ALGEBRAS 239
(1) If S = A [ T ], with A a quasi-coherent OY -algebra (3.1.7), then we immediately see that all
the invertible O-modules O (n) are canonically isomorphic to OX .
Furthermore, let N be a quasi-coherent A -module, and set M = N ⊗A A [ T ]. It then
follows from (3.2.3) and (3.1.7) that, under the canonical identification of X = Proj(A [ T ])
with X ′ = Spec(A ), the OX -module M fis identified with the OX ′ -module Nfassociated to
N (in the sense of (1.4.3)).
(2) Let S be an arbitrary graded OY -algebra, and S ′ the graded OY -algebra such that S0′ = OY
and Sn′ = Sn for all n > 0; then the canonical isomorphism from X = Proj(S ) to
X ′ = Proj(S ′ ) (3.1.8, (ii)) identifies OX (n) with OX ′ (n) for all n ∈ Z. This follows from the
same proposition for the affine case (2.5.16) and from the fact that the identifications, for the
affine opens of Y, commute with the restriction operations. Similarly, let X (d) = Proj(S (d) );
then the canonical isomorphism from X to X (d) (3.1.8, (i)) identifies OX (nd) with OX (d) (n)
for all n ∈ Z.
Proposition (3.2.10). — Let L be an invertible OY -module, and g the canonical isomorphism from
X(L ) = Proj(S(L ) ) to X = Proj(S ) (3.1.8, (iii)). Then, for any n ∈ Z, g∗ (OX(L ) (n)) is canonically
isomorphic to OX (n) ⊗Y L ⊗n .
P ROOF. Suppose first of all that Y is affine, of ring A, and that L = e L, where L is a free
monogenous A-module. With the notation from the proof of (3.1.8, (iii)), we define, for f ∈ Sd , an
isomorphism from (S(n))( f ) ⊗ A L⊗n to (S( L) (n))( f ⊗cd ) by sending ( x/ f k ) ⊗ cn , where x ∈ Skd+n ,
to the element ( x ⊗ cn+kd )/( f ⊗ cd )k ; it is immediate that this isomorphism is independent of the
chosen generator c of L; further, the isomorphisms thus defined for each f ∈ S+ are compatible
with the restriction operators D+ ( f ) → D+ ( f g). Finally, in the general case, we easily see, from the
definitions (3.1.1), that the isomorphisms thus defined for each affine open U of Y are compatible
with passing from U to an affine open U ′ ⊂ U. □
3.3. Graded S -module associated to a sheaf on Proj(S ) Throughout this entire section we suppose
that the graded OY -algebra S is generated by S1 (3.1.9). Recall that, by (3.1.8, (i)), this restrictive
assumption is not essential, thanks to finiteness conditions (3.1.10).
(3.3.1). Let p be the structure morphism X = Proj(S ) → Y. For every OX -module F , set
Γ ∗ (F ) = p∗ (F (n))
M
(3.3.1)
n∈Z
and, in particular, II | 57
Γ ∗ (O X ) = p∗ (OX (n)).
M
(3.3.1.2)
n∈Z
We know (0, 4.2.2) that there exists a canonical homomorphism
p∗ (F ) ⊗OY p∗ (G ) −→ p∗ (F ⊗OX G )
for OX -modules F and G ; we thus deduce from (3.2.7.1) that Γ∗ (OX ) is endowed with the structure
of a graded OY -algebra, and (3.2.5.2) defines the structure of a graded Γ∗ (OX )-module on Γ∗ (F ).
By (3.2.5), and by left exactness of the functor p∗ (0, 4.2.1), Γ∗ (F ) is an additive and left exact
covariant functor in F from the category of OX -modules to the category of graded OY -modules
(where the morphisms are the homomorphisms of degree 0). In particular, if J is a sheaf of ideals
of OX , then Γ∗ (J ) can be identified with a graded sheaf of ideals in Γ∗ (OX ).
(3.3.2). Let M be a quasi-coherent graded S -module. For every affine open U of Y, we defined in
(2.6.2) a homomorphism of abelian groups
α0,U : Γ(U, M0 ) −→ Γ( p−1 (U ), Mf).
It is immediate that these homomorphisms commute with the restriction operations (2.8.13.1) and
thus define (without using the hypothesis that S is generated by S1 ) a homomorphism of sheaves
of abelian groups
(3.3.2.1) α0 : M0 −→ p∗ (Mf).
3. HOMOGENEOUS SPECTRUM OF A SHEAF OF GRADED ALGEBRAS 240
Applying this result to each of the Mn = (M (n))0 , and taking (3.2.8.1) into account, we can
define a homomorphism of sheaves of abelian groups
(3.3.2.2) αn : Mn −→ p∗ (Mf(n))
for all n ∈ Z, whence a functorial homomorphism (of degree 0) of graded sheaves of abelian groups
(3.3.2.3) α : M −→ Γ∗ (Mf)
(also denoted αM ).
By taking M = S in particular, we see that α : S → Γ∗ (OX ) is a homomorphism of graded
OY -algebra, and that (3.3.2.3) is a di-homomorphism of graded modules, with respect to this
homomorphism of graded algebras.
We again note that to each of the αn there corresponds (0, 4.4.3) a canonical homomorphism of
OX -modules
(3.3.2.4) α♯n : p∗ (Mn ) −→ Mf(n).
We can easily verify that this homomorphism is exactly the one which corresponds functorially
(3.2.4) to the canonical homomorphism (of degree 0) of graded OY -modules
(3.3.2.5) Mn ⊗OY S −→ M (n)
where the grading of the right-hand side comes naturally from that of S . We can restrict to the case II | 57
where Y = Spec( A) is affine, M = M, e and S = S, e with the graded A-algebra S being generated
by S1 , so that, as f runs over S1 , the D+ ( f ) form a cover of X. By the definitions (2.6.2), we see
then see, taking (I, 1.6.7) into account, that the restriction to D+ ( f ) of the homomorphism (3.3.2.4)
corresponds (I, 1.3.8) to the homomorphism of S( f ) -modules Mn ⊗ A S( f ) → (S(n))( f ) that sends
x ⊗ 1 (where x ∈ Mn ) to x/1; this proves the claim.
Proposition (3.3.3). — For every section f ∈ Γ(Y, Sd ) (where d > 0), X f is identical to the set of points of
X where αd ( f ) (considered as a section of OX (d)) does not vanish (0, 5.5.2).
P ROOF. (Note that αd ( f ) is a section of p∗ (OX (d)) over Y, but by definition such a section is
also a section of O (d) over X (0, 4.2.1)). The definition of X f (3.1.4) lets us reduce to the case where
Y is affine, which has already been dealt with in (2.6.3). □
(3.3.4). From now on, we suppose, in addition to the hypothesis at the start of this section, that,
for every quasi-coherent OX -module F , the p∗ (F (n)) are quasi-coherent on Y, so that Γ∗ (F ) =
n∈Z p∗ (F ( n )) is also a quasi-coherent OY -module ((I, 1.4.1) and (I, 1.3.9)); this will always be
L
the case if X is of finite type over Y (I, 9.2.2). We thus conclude that (Γ∗ (F ))e is defined, and is a
quasi-coherent OX -module. For every affine open U of Y< we have ((I, 1.3.9) and (I, 2.5.4))
M
Γ(U, p∗ (F (n))) e = Γ(U, p∗ (F (n))) e
M
n∈Z n∈Z
M
= Γ( p−1 (U ), F (n)) e
n∈Z
M
= Γ( p−1 (U ), F (n)) e
n∈Z
= (Γ∗ (F | p−1 (U )))e
and so (2.6.4) we have a canonical homomorphism
βU : Γ(U, ) p∗ (F (n)) e −→ F | p−1 (U ).
M
n∈Z
Furthermore, the commutativity of (2.8.13.2) shows that these homomorphism commute with
the restriction operations on Y; we thus obtain a canonical functorial homomorphism
(3.3.4.1) β : (Γ∗ (F ))e −→ F
(also denoted β F ) for quasi-coherent OX -modules.
3. HOMOGENEOUS SPECTRUM OF A SHEAF OF GRADED ALGEBRAS 241
Γ∗ ( β)
Γ ∗ (F ) −
→ Γ∗ ((Γ∗ (F ))e) −−−→ Γ∗ (F )
α
(3.3.5.2)
are the identity isomorphisms.
P ROOF. The question is local on Y, so we can reduce to (2.6.5). □
P ROOF. Since the question is local on Y, we can suppose that Y is affine of ring A; then S = S, e
where S = Γ(Y, S ), and by hypothesis S is an A-algebra generated by S1 = Γ(Y, S1 ), where we can
further suppose that S1 is an A-module of finite type ((I, 1.3.9) and (I, 1.3.12)). Then S is a graded
A-algebra of finite type, and we can reduce to (2.7.1, (ii)). □
P ROOF. Since the questions are local on Y, we can reduce to the case where Y is affine of ring A,
S = S,e where S is a graded A-algebra such that the ideal S+ is of finite type, and M = M, e where
M is a graded S-module; the proposition then follows from (2.7.3). □
Corollary (3.4.5). — Under the hypotheses of (3.4.4), every quasi-coherent OX -module F is isomorphic to
an OX -module of the form M f, where M is a quasi-coherent graded S -module. If, further, F is of finite type,
and if we assume that Y is a quasi-compact scheme, or that the underlying space of Y is Noetherian, then we
can take M to be of finite type.
II | 60
P ROOF. The first claim follows immediately from (3.4.4) by taking M = Γ∗ (F ). To prove the
second, it suffices to show that M is the inductive limit of its graded sub-S -modules of finite type
Nλ : indeed, it will follow from this that Mfis the inductive limit of the Nf
λ (3.2.4), and so F is the
3. HOMOGENEOUS SPECTRUM OF A SHEAF OF GRADED ALGEBRAS 242
/ SU
ϕU
(3.5.1.1) ′
SU
′
SV / SV
ϕV
Proposition (3.5.2). —
(i) If M is a quasi-coherent graded S -module, then there exists a canonical functorial isomorphism
from the OX ′ -module (M[ϕ] )e to the OX ′ -module Φ∗ (M f| G (ϕ)).
(ii) If M ′ is a quasi-coherent graded S ′ -module, then there exists a canonical functorial isomorphism ν
from the (OX | G (ϕ))-module Φ∗ (M g′ ) to the (OX | G (ϕ))-module (M ′ ⊗S ′ S )e| G (ϕ). If S ′ is
′
generated by S1 , then ν is an isomorphism.
P ROOF. The homomorphisms in question are indeed already defined if Y is affine ((2.8.7) and
(2.8.8)), and in the general case it suffices to check that they are compatible with the restriction
of an affine open of Y to a smaller open, which follows immediately from the commutativity of
(3.5.1.1). □
In particular, for all n ∈ Z, we have a canonical homomorphism
(3.5.2.1) Φ∗ (OX ′ (n)) −→ OX (n)| G (ϕ).
Proposition (3.5.3). — Let Y and Y ′ be preschemes, ψ : Y ′ → Y a morphism, and S a quasi-coherent
graded OY -algebra; set S ′ = ψ∗ (S ). Then the Y ′ -scheme X ′ = Proj(S ′ ) is canonically identified with
Proj(S ) ×Y Y ′ . Furthermore, if M is a quasi-coherent graded S -module, then the OX ′ -module (ϕ∗ (M ))e
can be identified with Mf⊗Y OY ′ .
P ROOF. Note first of all that ψ∗ (S ) and ψ∗ (M ) are quasi-coherent OY ′ -modules, as are their
homogenous components (0, 5.1.4). Let U be an affine open of Y, U ′ ⊂ ψ−1 (U ) an affine open of Y ′ ,
and A and A′ the rings of U and U ′ , respectively; then S |U = S, e where S is a graded A-algebra,
′ ′ ′
and S |U can be identified with (S ⊗ A A )e (I, 1.6.5); the first claim then follows from (2.8.10) and
(I, 3.2.6.2), since we immediately see that the projection Proj(S ′ |U ′ ) → Proj(S |U ) defined by the
above identification is compatible with the restriction operations on U and U ′ , and thus indeed
defines a morphism Proj(S ′ ) → Proj(S ). Now let
q : Proj(S ) −→ Y
q′ : Proj(S ′ ) −→ Y ′
be the structure morphisms; q′−1 (U ′ ) can then be identified with q−1 (U ) ×U U ′ , and the two sheaves
(ψ∗ (M ))e|q′−1 (U ′ ) and (Mf⊗Y OY′ )|q′−1 (U ′ ) are then both canonically identified with ( M ⊗ A A′ )e,
3. HOMOGENEOUS SPECTRUM OF A SHEAF OF GRADED ALGEBRAS 244
where we set M = Γ(U, M ), by (2.8.10) and (I, 1.6.5); whence the second claim, since we can again
immediately see the compatibility of the above identifications with the restriction operations. □
Corollary (3.5.4). — With the notation of (3.5.3), OX ′ (n) is canonically identified with OX (n) ⊗Y OY ′ for
all n ∈ Z (where X = Proj(S )).
P ROOF. Indeed, with the notation of (3.5.3), it is clear that ψ∗ (S (n)) = S ′ (n) for all n ∈ Z. □
(3.5.5). Keeping the above notation, denote by Ψ the canonical projection X ′ → X, and set M ′ =
ψ∗ (M ); we further suppose that S is generated by S1 , and that X is of finite type over Y; it then II | 63
follows that S ′ is generated by S1′ (as can be seen by reducing to the case where Y and Y ′ are affine),
and that X ′ is of finite type over Y ′ (I, 6.3.4). Let F be an OX -module, and set F ′ = Ψ∗ (F ); it then
follows from (3.5.4) and (0, 4.3.3) that F ′ (n) = Ψ∗ (F (n)) for all n ∈ Z. We further define a canonical
Ψ-homomorphism θn : q∗ (F (n)) → q′∗ (F ′ (n)) in the following way: given the commutativity of
the diagram
Ψ
Xo X′
q q′
Yo Y′
ψ
it is enough to define a homomorphism q∗ (F (n)) → ψ∗ (q′∗ (Ψ∗ (F (n)))) = q∗ (Ψ∗ (Ψ∗ (F (n)))),
and it suffices to take the homomorphism θn = q∗ (ρn ), where ρn is the canonical homomorphism
F (n) → Ψ∗ (Ψ∗ (F (n))) (0, 4.4.3). It is immediate that, for every affine open U of Y and every
affine open U ′ of Y ′ such that U ′ ⊂ ψ−1 (U ), the homomorphism θn gives, on sections, the canonical
homomorphism (0, 3.7.2) Γ(q−1 (U ), F (n)) → Γ(q′−1 (U ′ ), F ′ (n)). The commutativity of (2.8.13.2)
then shows that, if F is quasi-coherent, the diagram
/ F′
ρ
FO O
βF βF ′
commutes (the top horizontal arrow being the canonical Ψ-morphism F → Ψ∗ (F )).
Similarly, the commutativity of (2.8.13.1) shows that the diagram
Γ ∗ (M f) θ / Γ ∗ (M
g′ )
O O
αM αM ′
M / M′
ρ
commutes (the bottom horizontal arrow being the canonical ψ-morphism M → ψ∗ (M )).
(3.5.6). Now consider preschemes Y and Y ′ , a morphism g : Y ′ → Y, a quasi-coherent graded
OY -algebra (resp. quasi-coherent graded OY ′ -algebra) S (resp. S ′ ), and a g-morphism of graded
algebras u : S → S ′ , i.e. a OY -homomorphism of graded algebras S → g∗ (S ′ ); we already know
that this is equivalent to giving a OY ′ -homomorphism of graded algebras u♯ : g∗ (S ) → S ′ . We
thus canonically obtain from u♯ a Y ′ -morphism W = Proj(u♯ ) : G (u♯ ) → Proj( g∗ (S )), where G (u♯ )
is an open of X ′ = Proj(S ′ ) (3.5.1). We also know that X ′′ = Proj( g∗ (S )) is canonically identified
with X ×Y Y ′ , by taking X = Proj(S ) (3.5.3); composing the first projection p : X ×Y Y ′ → X with II | 64
Proj(u♯ ), we thus obtain a morphism v : G (u♯ ) → X, which we denote by Proj(u), and which is such
that the diagram
G (u♯ )
v /X
Y′ /Y
g
3. HOMOGENEOUS SPECTRUM OF A SHEAF OF GRADED ALGEBRAS 245
commutes.
Furthermore, for every quasi-coherent graded OY -module M , we have a canonical v-morphism
(3.5.6.1) ν:Mf−→ ( g∗ (M ) ⊗ g∗ (S ) S ′ )e| G (u♯ ).
P ROOF.
(i) We can assume that ϕ is surjective (3.6.1). Then, for every affine open U of Y, Γ(U, S ) →
Γ(U, S ′ ) is surjective (I, 1.3.9), so (3.5.1) G (ϕ) = X. We can then immediately reduce to
proving the proposition in the case where Y is affine, and this follows from (2.9.2, (i)).
(ii) We can reduce to proving that the homomorphism J f → OX induced by the canonical
injection J → S is an isomorphism from J to I ; since the question is local on Y, we
f
can take Y to be affine of ring A, which implies that S = S, e where S is a graded A-algebra
generated by S1 , with S1 of finite type over A. It then suffices to apply (2.9.2, (ii)).
□
Corollary (3.6.3). — Under the conditions of (3.6.2, (i)), suppose further that S is generated by S1 . Then
Φ∗ (OX (n)) is canonically identified with OX ′ (n) for all n ∈ Z.
P ROOF. We have defined such a canonical isomorphism when Y is affine (2.9.3); in the general
case, it suffices to show that the isomorphisms thus defined for each affine open U of Y are compatible
with the passage from U to an affine open U ′ ⊂ U, which is immediate. □
Corollary (3.6.4). — Let Y be a prescheme, S a quasi-coherent graded OY -algebra generated by S1 , M
a quasi-coherent OY -module, u a surjective OY -homomorphism M → S1 , and u : SOY (M ) → S the
homomorphism of graded OY -algebras that extends u (1.7.4). Then the morphism corresponding to u is a
closed immersion of Proj(S ) into Proj(SOY (M )).
P ROOF. Indeed, u is surjective by hypothesis, and we apply (3.6.1, (i)). □
3. HOMOGENEOUS SPECTRUM OF A SHEAF OF GRADED ALGEBRAS 246
n⩾0
which is equivalent to having a q-morphism of graded algebras
ψ♭ : S −→ q∗ (S ′ ).
We know that Proj(S ′ ) is canonically identified with X ((3.1.7) and (3.1.8, (iii))); we thus
canonically obtain from ψ an open subset G (ψ) of X, and a Y-morphism
(3.7.1.1) rL ,ψ : G (ψ) −→ Proj(S ) = P
which we call the morphism associated to L and ψ; recall (3.5.6) that this morphism is obtained by II | 66
composing the Y-morphism
τ = Proj(ψ) : G (ψ) −→ Proj(q∗ (S ))
with the first projection π : Proj(q∗ (S )) = P ×Y X → P.
(3.7.2). We will explicitly describe r = rL ,ψ in the case where Y = Spec( A) is affine, so that
S = S, e where S is a positively-graded A-algebra. Suppose first of all that X = Spec( B) is also
affine, and that we have L = e L, where L is a free B-module of rank 1. Then q∗ (S ) = (S ⊗ A B)e
(I, 1.6.5); if c is a generator of L, then ψn : q∗ (Sn ) → L ⊗n corresponds to a homomorphism
wn : s ⊗ b 7→ bvn (s)c⊗n from Sn ⊗ A B to L⊗n , where vn : Sn → B is a homomorphism of A-modules,
constituting a homomorphism of algebras S → B. Let f ∈ Sd (where d > 0), and set g = vd ( f ); we
have π −1 ( D+ ( f )) = D+ ( f ⊗ 1) by (2.8.10) and the identification of D+ ( f ) with Spec(S( f ) ) (2.3.6);
we also know, from (2.8.1.1) (taking into account the canonical identification of X with Proj(S ′ )),
that
τ −1 ( D+ ( f ⊗ 1)) = D ( g)
whence
(3.7.2.1) r −1 ( D+ ( f )) = D ( g).
Furthermore, the morphism τ = Proj(ψ), restricted to D ( g), corresponds to the homomorphism
that sends (s ⊗ 1)/( f ⊗ 1)n (where s ∈ Snd ) to vnd (s)/gn (2.8.1), and the projection π, restricted
to D+ ( f ⊗ 1), corresponds t the homomorphism s/ f n 7→ (s ⊗ 1)/( f ⊗ 1)n ; we thus conclude that
r, restricted to D ( g), corresponds to the homomorphism of A-algebras ω : S( f ) → Bg such that
ω (s/ f n ) = vnd (s)/gn (where s ∈ Snd and n > 0). Passing to the case where X is arbitrary (but Y
still affine), we thus obtain, taking (2.8.1) into account, the following:
Proposition (3.7.3). — If Y = Spec( A) is affine and S = S,
e where S is a graded A-algebra, then, for all
f ∈ Sd = Γ(Y, Sd ), we have
−1
(3.7.3.1) rL ,ψ ( D+ ( f )) = Xψ♭ ( f )
Corollary (3.7.4). — Under the hypotheses of (3.7.3), for rL ,ψ to be everywhere defined, it is necessary and
sufficient that, for every x ∈ X, there exist an integer n > 0 and a section s of Sn over Y such that, if we set
t(= ψ♭ (s) ∈ Γ( X, L ⊗n ), then t( x ) ̸= 0.
Corollary (3.7.5). — Under the hypotheses of (3.7.3), for rL ,ψ to be dominant, it is necessary and sufficient
that, for every integer n > 0, every section s ∈ Sn such that ψ♭ (s) ∈ Γ( X, L ⊗n ) is locally nilpotent is itself
nilpotent.
−1
P ROOF. We have to show that rL ,ψ ( D+ ( s )) is not empty if D+ ( s ) is empty, and the corollary
thus follows from (3.7.3.1) and (2.3.7). □
rL ,ψ′ is everywhere defined, then so too is rL ,ψ ; if u is (TN)-surjective, then if rL ,ψ′ is dominant, so too is
rL ,ψ ; conversely, if u is (TN)-injective, then if rL ,ψ is dominant, so too is rL ,ψ′ .
P ROOF. We know that G (ψ′ ) ⊂ G (ψ) (2.8.4), whence the first claim; if u is (TN)-surjective,
then Proj(u) : Proj(S) → Proj(S′ ) is everywhere defined and a closed immersion; since rL ,ψ′ is
the composition of Proj(u) and the restriction to G (ψ′ ) of rL ,ψ , we thus conclude that, if rL ,ψ′ is
dominant, then so too is rL ,ψ . Finally, if u is (TN)-injective, then we know that Proj(u) is a dominant
morphism from G (u) to Proj(S′ ) (2.8.3); since G (ψ′ ) is the inverse image of G (u) under rL ,ψ , we see
that, if rL ,ψ is dominant, then so too is rL ,ψ′ . □
everywhere defined, it is necessary and sufficient that there exist some λ such that rL ,ψλ be everywhere
defined; rL ,ψµ is then everywhere defined for µ ⩾ λ.
Corollary (3.7.8). — Under the hypotheses of (3.7.7), if the rL ,ψλ are dominant, then so too is rL ,ψ ; the
converse is true if the ϕλ are all injective.
II | 68
P ROOF. The second claim is a particular case of (3.7.6); to show that rL ,ψ is dominant, we can
restrict to the case where Y is affine; if s ∈ S is such that ψ♭ (s) is locally nilpotent, since we can write
s = ϕλ (sλ ) for at least one λ, we conclude from the hypotheses and from (3.7.5) that sλ is nilpotent,
and thus so too is s, and the criteria of (3.7.5) then apply. □
Remarks (3.7.9). —
3. HOMOGENEOUS SPECTRUM OF A SHEAF OF GRADED ALGEBRAS 248
(i) With the notation of (3.7.1), and taking (3.2.10) into account, we have, for all n ∈ Z, a
canonical homomorphism
∗ ⊗n
(3.7.9.1) θ : rL ,ψ (O P ( n )) −→ L
defined in a general way in (3.5.6.2). We immediately see that, under the conditions of
(3.7.3), the restriction of this homomorphism to Xψ♭ ( f ) sends s/ f k (where s ∈ Sn+kd ) to the
element (ψ♭ (s)| Xψ♭ ( f ) )(ψ♭ ( f )| Xψ♭ ( f ) )−k .
(ii) Let F be a quasi-coherent OX -module, and suppose that q is quasi-compact and sepa-
rated, so that, for all n ⩾ 0, q∗ (F ⊗ L ⊗n ) is a quasi-coherent OY -module (I, 9.2.2). Let
M ′ = n⩾0 F ⊗ L ⊗n , which is a quasi-coherent graded S ′ -module, and consider its
L
P ROOF. For r to be an immersion (resp. a closed immersion), it is necessary and sufficient that
there exist a cover of r ( G (ψ)) (resp. of P) by the D+ (sα ), such that, if we set Vα = r −1 ( D+ (sα )),
then the restriction of r to Uα is a closed immersion of Vα into D+ (sα ) (I, 4.2.4). Condition (i) says
that r is everywhere defined and also that the D+ (sα ) cover r ( X ), by (3.7.3.1); since D+ (sα ) is affine,
conditions (ii) and (iii) say that the restriction of r to X f α is a closed immersion into D+ (sα ), by
(I, 4.2.3); finally, since conditions (iii) and (iv) say that ψ(♭s ) is bijective (using the notation of (3.7.3.2)),
α
conditions (ii), (iii), and (iv) say that the restriction of r to X f α is an isomorphism to D+ (sα ) for all α,
and so conditions (i), (ii), (iii), and (iv) together say that r is an open immersion. □
Corollary (3.8.3). — Under the hypotheses of (3.7.6), if rL ,ψ′ is everywhere defined and an immersion,
then so too is rL ,ψ . If we further suppose that u is (TN)-surjective, then if rL ,ψ′ is an open (resp. closed)
immersion, so too is rL ,ψ .
P ROOF. By (3.8.2), there exists a family s′α ∈ Sn′ α such that, if we set f α = ψ′♭ (s′α ), conditions (i),
(ii), and (iii) are satisfied. But if we set sα = u(s′α ), then we also have that f α = ψ♭ (sα ), and if t =
(ψ′♭ (s′ )| X f α )( f α | X f α )−m , then also t = (ψ♭ (s)| X f α )( f α | X f α )−m by setting s = u(s′ ), whence the first
claim. The second claim follows immediately from the fact that Proj(u) is a closed immersion. □
Proposition (3.8.4). — Suppose that the hypotheses of (3.7.7) are satisfied, and further that q : X → Y is
a morphism of finite type. Then, for rL ,ψ to be everywhere defined and an immersion, it is necessary and
sufficient that there exist some λ such that rL ,ψλ is everywhere defined and an immersion; in this case, rL ,ψµ
is also everywhere defined and an immersion for all µ ⩾ λ.
II | 70
P ROOF. Taking (3.8.3) into account, it suffices to show that, if rL ,ψ is everywhere defined and
an immersion, then so too is rL ,ψλ for at least one λ. By the same argument as in (3.7.7), using the
fact that Y is quasi-compact, (3.8.2) shows the existence of a finite family (si ∈ Sni ) of elements of S
satisfying conditions (i), (ii), and (iii). The morphism X f i → Y (where f i = ψ♭ (si )) is of finite type:
it is a morphism of affine schemes, and is thus quasi-compact (I, 6.6.1), and also locally of finite
type, since q is of finite type (I, 6.3.2), and the conclusion then follows from (I, 6.6.3). The ring Bi of
X f i is thus an A-algebra of finite type (I, 6.6.3); let (tij ) be a family of generators of this algebra. By
hypothesis, there exist elements sij′ ∈ Smij ni such that
Also by hypothesis, there exists some λ and elements siλ ∈ Snλi and sijλ ′ ∈ Sλ
mij ni whose images under
′
ϕλ are si and sij , respectively; it is clear that rL ,ψλ satisfies conditions (i), (ii), and (iii) of (3.8.2). □
defined and an immersion, it is necessary and sufficient that there exist an integer n > 0 and a quasi-coherent
sub-OY -module of finite type E of Sn such that
a) the homomorphism ψn ◦ q∗ ( jn ) : q∗ (E ) → L ⊗n (where jn : E → S is the canonical injection) is
surjective; and
b) if we denote by S ′ the (graded) sub-OY -algebra of S generated by E , and by ψ′ the homomorphism
ψ ◦ q∗ ( j′ ) (where j′ is the injection of S ′ into S ), rL ,ψ′ is everywhere defined and an immersion.
If this is the case, then every quasi-coherent sub-OY -module E ′ of Sn that contains E also possesses the
same property, as does the image of ⊗k E in Skn for all k > 0.
P ROOF. The fact that the condition is sufficient, and the last two claims, are particular cases of
(3.8.3), taking into account the canonical isomorphism between Proj(S ) and Proj(S (k) ) (3.1.8).
Let (Ui ) be a finite cover of Y by affine opens, and set Ai = A(Ui ). Since q−1 (Ui ) is compact,
the hypothesis that rL ,ψ be everywhere defined and an immersion implies, along with (3.8.2), the
(i )
existence of a finite family (sij ∈ Snij ) of elements of S(i) = Γ(Ui , S ) satisfying conditions (i), (ii), and
4. PROJECTIVE BUNDLES; AMPLE SHEAVES 250
(iii). We see, as in the proof of (3.8.4), that the morphism X f ij → Ui (where f ij = ψ♭ (sij )) is of finite
type, and that the ring Bij of X f ij is thus an Ai -algebra of finite type, having a system of generators of
(i )
the form (ψ♭ (tijk ) X f ij )( f ij | X f ij )−mijk , where tijk ∈ Smijk nij . Let n be a common multiple of the mijk nij ;
ρ
replacing (for each (i, j, k)) sij by some power sij such that ρmijk nij = n, and multiplying tijk by
ρ−mijk (i )
sij , we can assume that, for each i, the sij and tijk belong to Sn and that mijk = 1. Let Ei be
the sub-Ai -module of S(i) generated by these elements (for fixed i). Then there exists a coherent
sub-OY -module Ei of Sn of finite type such that Ei |Ui = ( Ei )e (I, 9.4.7). It is clear that the sub- II | 71
OY -module E of Sn given by the sum of the Ei is the desired object (since each section f ij is such
that, for all x ∈ X f ij , there exists an affine neighbourhood V ⊂ X f ij of x such that f |V is a basis for
Γ(V, L ⊗n )). □
P ROOF. Note first of all that, if A is a ring, E an A-module, and L a free monogenous A-module,
then we can canonically define a homomorphism of A-modules
Sn ( E ⊗ L) −→ Sn ( E) ⊗ L⊗n
4. PROJECTIVE BUNDLES; AMPLE SHEAVES 251
P ROOF. We have seen, in (3.3.2), that α1♯ corresponds functorially to the canonical homomor-
phism E ⊗OY SOY (E ) → (SOY (E ))(1); since, by definition, E generates SOY (E ), this homomorphism
is surjective, whence the conclusion, by (3.2.4) □
4.2. Morphisms from a prescheme to a projective bundle
(4.2.1). Keeping the notation of (4.1.5), let X be a Y-prescheme, q : X → Y the structure morphism,
and let r : X → P be a Y-morphism such that the following diagram commutes:
Po
r
X
p
q
Y
II | 73
r∗
Since the functor is right exact (0, 4.3.1), we obtain, from the surjective homomorphism in
(4.1.5.1), a surjective homomorphism
r ∗ (α1♯ ) : r ∗ ( p∗ (E )) −→ r ∗ (OP (1)).
But r ∗ ( p∗ (E )) = q∗ (E ), and r ∗ (OP (1)) is locally isomorphic to r ∗ (OP ) = OX , or, in other words,
the latter is an invertible sheaf Lr on OX , and so we have defined, given r, a canonical surjective
OX -homomorphism
(4.2.1.1) ϕr : q∗ (E ) −→ Lr .
When Y = Spec( A) is affine, and E = E, e we can further clarify this homomorphism in the
following way: given f ∈ E, it follows from (2.6.3) that
(4.2.1.2) r −1 ( D+ ( f )) = Xϕ♭ ( f ) .
r
Now let V be an affine open subset of X that is contained inside r −1 ( D+ ( f )), and let B be
its ring, which is an A-algebra; let S = S A ( E); the restriction of r to V corresponds to an A-
homomorphism ω : S f → B, and we have that q∗ (E )|V = ( E ⊗ A B)e and Lr |V = Ler , whence Lr =
(S(1))( f ) ⊗S( f ) B[ω ] (I, 1.6.5). The restriction of ϕr to q∗ (E )|V corresponds to the B-homomorphism
u : E ⊗ A B → Lr , which sends x ⊗ 1 to ( x/1) ⊗ f = ( f /1) ⊗ ω ( x/ f ). The canonical extension of ϕr
to a homomorphism of OX -algebras
ψr : q∗ (S(E )) = S(q∗ (E )) −→ S(Lr ) = Lr⊗n
M
n⩾0
n⩾0
In particular: II | 75
(1) The corollary (4.2.6) applies whenever Y is a local scheme (I, 2.4.8). Let Y be an arbitrary
prescheme, y a point of Y, and Y ′ = Spec(k(y)); then the fibre p−1 (y) of P(E ) can, by
(4.1.3.1), be identified with P(E y ), where E y = Ey ⊗Oy k(y) = Ey /my Ey is considered as a
vector space over k(y). More generally, if K is an extension of k(y), then p−1 (y) ⊗k(y) K can
be identified with P(E y ⊗k(y) K ). The corollary (4.2.6) then shows that the set of geometric
points of P(E ) with values in the extension K of k(y) (I, 3.4.5), which we can also call the
rational geometric fibre over K of P(E ) over the point y, can be identified with the projective
space associated to the dual of the K-vector space E y ⊗k(y) K.
(2) Suppose that Y is affine of ring A, and, further, that every invertible OY -module is trivial;
further, take E = OYn ; then, in (4.2.6), V can be identified with An (I, 1.3.8), and V ∗ with the
sets of systems ( f i )1⩽i⩽n of elements of A that generate the ideal A; any two such systems
define the same Y-section of PYn−1 = PnA−1 , or, in other words, the same point of PnA−1 with
values in A, if and only if one of them can be obtained from the other by multiplication by
an invertible element of A.
These properties justify the terminology “projective bundle” for P(E ). We note that the defi-
nitions that we will similarly obtain for “projective space” is in fact dual to the classical definition;
this is imposed upon us by the necessity of being able to define P(E ) for arbitrary quasi-coherent
OY -modules E , and not just locally free ones.
Remark (4.2.7). — We will see, in Chapter V, that, if Y is connected and locally Noetherian, and if E
is locally free, then, letting P = P(E ), every invertible OP -module is isomorphic to an OP -module of
the form L ′ ⊗OY OP (m), with L ′ some invertible OY -module, well defined up to isomorphism, and
m some well defined integer. In other words, H1 ( P, OP∗ ) is isomorphic to Z × H1 (Y, OY∗ ) (0, 5.4.7).
We will also see ((III, 2.1.14), taking (0, 5.4.10) into account) that p∗ (L ⊗m ) = 0 if m < 0, and
p∗ (L ⊗m ) is isomorphic to L ′ ⊗OY (SOY (E ))m if m ⩾ 0. If F is a quasi-coherent OY -module, then
every Y-morphism P(E ) → P(F ) is determined by the data of an invertible OY -module, an integer
m ⩾ 0, and an OY -homomorphism ψ : F → L ′ ⊗OY (SOY (E ))m such that the corresponding
homomorphism ψ♯ of OP(F ) -modules is surjective. We will also see that, if the Y-morphism in
question is an isomorphism, then m = 1 and F is isomorphic to E ⊗OY L ′ (the converse of (4.1.4)).
This will allow us to determine the sheaf of germs of automorphisms of P(E ) as the quotient of the
sheaf of groups Aut(E ) (which is locally isomorphic to GL(n, OY ) is E is of rank n) by OY∗ .
(4.2.8). Keeping the notation of (4.2.1), let u : X ′ → X be a morphism; if the Y-morphism r : X → P
corresponds to the homomorphism ϕ : q∗ (E ) → L , then the Y-morphism r ◦ u corresponds to
u∗ (ϕ) : u∗ (q∗ (E )) → u∗ (L ), as follows immediately from the definitions.
(4.2.9). Let E and F be quasi-coherent OY -modules, v : E → F a surjective homomorphism,
and j = P(v) the corresponding closed immersion P(F ) → P(E ) (4.1.2). If the Y-morphism
r : X → P(F ) corresponds to the homomorphism ϕ : q∗ (F ) → L , then the Y-morphism j ◦ r II | 76
q∗ (v) ϕ
corresponds to q∗ (E ) −−−→ q∗ (F ) −
→ L ; this again follows from the definition given in (4.2.1).
(4.2.10). Let ψ : Y ′ → Y be a morphism, and let E ′ = ψ∗ (E ). If the Y-morphism r : X → P
corresponds to the homomorphism ϕ : q∗ (E ) → L , then the Y ′ -morphism
r(Y ′ ) : X(Y ′ ) −→ P′ = P(E ′ )
corresponds to ϕ(Y ′ ) : q∗(Y ′ ) (E ′ ) = q∗ (E ) ⊗Y OY ′ → L ⊗Y OY ′ . Indeed, by (4.1.3.1), we have the
commutative diagram
p (Y ′ ) r ′
Y′ o P′ = P(Y ′ ) o
(Y )
X (Y ′ )
u v
p
Yo Po
r
X
From (4.1.3.1), we have
(r(Y′ ) )∗ (OP′ (1)) = (r(Y′ ) )∗ (u∗ (OP (1))) = v∗ (r ∗ (OP (1))) = v∗ (L ) = L ⊗Y OY′ ;
4. PROJECTIVE BUNDLES; AMPLE SHEAVES 254
we also know that u∗ (α1♯ ) is exactly the canonical homomorphism α1♯ : ( p(Y ′ ) )∗ (E ′ ) → OP′ (1); we
can see this by explicitly calculating the canonical homomorphisms α1♯ to P and P′ as in (4.1.6).
Whence our claim.
4.3. The Segre morphism
(4.3.1). Let Y be a prescheme, and E and F quasi-coherent OY -modules; let P1 = P(E ) and P2 =
P(F ), and denote the structure morphisms by p1 : P1 → Y and p2 : P2 → Y. Let Q = P1 ×Y P2 ,
and let q1 : Q → P1 and q2 : Q → P2 be the canonical projections; then the OQ -module L =
OP1 (1) ⊗Y OP2 (1) = q1∗ (OP1 (1)) ⊗OQ q2∗ (OP2 (1)) is invertible, since it is the tensor product of of two
invertible OQ -modules (0, 5.4.4). Also, if r = p1 ◦ q1 = p2 ◦ q2 is the structure morphism Q → Y,
then r ∗ (E ⊗OY F ) = q1∗ ( p1∗ (E )) ⊗OQ q2∗ ( p2∗ (F )) (0, 4.3.3); the canonical surjective homomorphisms
(4.1.5.1) p1∗ (E ) → OP1 (1) and p2∗ (F ) → OP2 (1) thus give, by taking the tensor product, a canonical
homomorphism
(4.3.1.1) s : r ∗ (E ⊗OY F ) −→ L
which is evidently surjective; from this we obtain (4.2.2) a canonical morphism, called the Segre
morphism:
(4.3.1.2) ς : P(E ) ×Y P(F ) −→ P(E ⊗OY F ).
We can study the morphism ς more explicitly in the case where Y = Spec( A) is affine, and
E =Ee and F = F, e where E and F are A-modules, whence E ⊗O F = ( E ⊗ A F )∼ (I, 1.3.12); let
Y
R = S A ( E), S = S A ( F ), and T = S A ( E ⊗ A F ); let f ∈ E and g ∈ F, and consider the affine open
D+ ( f ) ×Y D+ ( g) = Spec( B)
of Q, where B = R( f ) ⊗ A S( g) ; the restriction of L to this affine open is e
L, where II | 77
L = ( R(1))( f ) ⊗ A (S(1))( g)
and the element c = ( f /1) ⊗ ( g/1) is a generator of L considered as a free B-module (2.5.7). The
homomorphism (4.3.1.1) corresponds to the homomorphism
( x ⊗ y) ⊗ b 7−→ b(( x/1) ⊗ (y/1))
from ( E ⊗ A F ) ⊗ A B to L. With the notation of (3.7.2), we thus have that v1 ( x ⊗ y) = ( x/ f ) ⊗
(y/g); the restriction of ς to D+ ( f ) ×Y D+ ( g) is a morphism from this affine scheme to D+ ( f ⊗ g),
corresponding to the ring homomorphism ω : T( f ⊗ g) → R( f ) ⊗ A S( g) defined by
(4.3.1.3) ω (( x ⊗ y)/( f ⊗ g)) = ( x/ f ) ⊗ (y/g)
for x ∈ E and y ∈ F.
(4.3.2). It follows from (4.2.3) that we have a canonical isomorphism
∼
(4.3.2.1) τ : ς∗ (OP (1)) −
→ OP1 (1) ⊗Y OP2 (1)
where we let P = P(E ⊗OY F ). We will show that, for x ∈ Γ(Y, E ) and y ∈ Γ(Y, F ), we have
(4.3.2.2) τ (α1 ( x ⊗ y)) = α1 ( x ) ⊗ α1 (y).
Indeed, we can restrict to the case where Y is affine, and we then have, with the notation of
f ⊗g f
(4.3.1) and (2.6.2), that α1 ( x ⊗ y) = ( x ⊗ y)/1 in ( T (1))( f ⊗ g) , that α1 ( x ) = x/1 in ( R(1))( f ) , and
g
that α1 (y) = y/1 in (S(1))( g) . The definition of τ given in (4.2.3) and the calculation of v1 done in
(4.3.1) then immediately prove the claim (4.3.2.2). From this we obtain the equation
(4.3.2.3) ς−1 ( Px⊗y ) = ( P1 ) x ×Y ( P2 )y
with the notation of (3.1.4). Indeed, taking (3.3.3) into account, the equation (4.3.2.2) (by restricting to
the affine case, with the help of (I, 3.2.7) and (I, 3.2.3)) leaves us only to prove the following lemma:
Lemma (4.3.2.4). — Let B and B′ be A-algebras, and let Y = Spec( A), Z = Spec( B), and Z ′ = Spec( B′ );
then D (t ⊗ t′ ) = D (t) ×Y D (t′ ) for any t ∈ B, t′ ∈ B′ .
4. PROJECTIVE BUNDLES; AMPLE SHEAVES 255
ς ς
P (E ′ ⊗ F ) / P (E ⊗ F )
commutes, where j denotes the canonical closed immersion P(E ′ ) → P(E ). Let P1′ = P(E ′ ) and
keep the notation from (4.3.1); then j × 1 is a closed immersion (I, 4.3.1) and, up to isomorphism,
( j × 1)∗ (OP1 (1) ⊗ OP2 (1)) = j∗ (OP1 (1)) ⊗ OP2 (1) = OP′ (1) ⊗ OP2 (1)
1
by (4.1.2.1) and (I, 9.1.5); our claim then follows from (4.2.8) and (4.2.9).
(4.3.5). With the notation of (4.3.1), let ψ : Y ′ → Y be a morphism, and let E ′ = ψ∗ (E ) and
F ′ = ψ∗ (F ); then the Segre morphism P(E ′ ) × P(F ′ ) → P(E ′ ⊗ F ′ ) can be identified with ς (Y ′ ) .
Indeed, keeping the notation of (4.3.1), let P1′ = P(E ′ ) and P2 = P(F ′ ); we know (4.1.3.1) that Pi′ can
be identified with ( Pi )(Y ′ ) (i = 1, 2), and so the structure morphism P1′ × P2′ → Y ′ can be identified
with r(Y ′ ) . Also E ′ ⊗ F ′ can be identified with ψ∗ (E ⊗ F ), and so P(E ′ ⊗ F ′ ) can be identified
with (P(E ⊗ F ))(Y ′ ) (4.1.3.1). Finally, OP′ (1) ⊗Y OP′ (1) = L ′ can be identified with L ⊗Y OY ′ , by
1 2
(4.1.3.1) and (I, 9.1.11). The canonical homomorphism (r(Y ) )∗ (E ′ ⊗ F ′ ) → L ′ can then be identified
with s(Y ′ ) , and our claim follows from (4.2.10).
Remark (4.3.6). — The prescheme given by the sum of P(E ) and P(F ) is even canonically iso-
morphic to a closed subprescheme of P(E ⊕ F ). Indeed, the surjective homomorphisms E ⊕ F → E
and E ⊕ F → F correspond to closed immersions P(E ) → P(E ⊕ F ) and P(F ) → P(E ⊕ F );
everything then reduces to showing that the underlying spaces of the closed subpreschemes of
P(E ⊕ F ) obtained in this way have empty intersection. Since the question is local on Y, we can
adopt the notation of (4.3.1); but Sn ( E) and Sn ( F ) can be identified with submodules of Sn ( E ⊕ F )
with intersection consisting only of 0; if p is a graded prime ideal of S( E) such that p ∩ Sn ( E) ̸= Sn ( E)
for any n ⩾ 0, then there exists a corresponding graded prime ideal of S( E ⊕ F ) whose intersection
with Sn ( E) is p ∩ Sn ( E), but who also contains S+ ( F ), as we immediately see; thus no point in
Proj(S( E)) can have the same image in Proj(S( E ⊕ F )) as any point in Proj(S( F )).
4. PROJECTIVE BUNDLES; AMPLE SHEAVES 256
P ROOF.
(i) The fact that rL ,ϕ is everywhere defined and is an immersion is equivalent, by (3.8.5), to
the existence of some n ⩾ 0 and E such that, if S ′ is the subalgebra of S generated by E ,
the homomorphism q∗ (E ) → L ⊗n is surjective and the morphism rL ,ψ′ : X → Proj(S ′ )
is everywhere defined and is an immersion. We already have a canonical surjective
homomorphism S(E ) → S ′ to which there exists a corresponding closed immersion
Proj(S ′ ) → P(E ) (3.6.2); whence the conclusion.
(ii) Since F is the inductive limit of its quasi-coherent submodules of finite type Eλ (I, 9.4.9),
S(F ) is the inductive limit of the S(Eλ ); the conclusion then follows from (3.8.4), by
observing that we can take all the ni in the proof of (3.8.4) to be equal to 1: indeed,
supposing that Y is affine, if r = rL ,ϕ is an immersion, then r ( X ) is a quasi-compact
subspace of P(F ) that we can cover by finitely many open subsets of P(F ) of the form
D+ ( f ), with f ∈ F, such that D+ ( f ) ∩ r ( X ) is closed.
□
Definition (4.4.2). — Let Y be a prescheme, and q : X → Y a morphism. We say that an invertible
OX -module L is very ample for q, or relative to q (or very ample for (or relative to) Y, or simply very
ample, if q is clear from the context) if there exists a quasi-coherent OY -module E and a Y-immersion
i from X to P = P(E ) such that L is isomorphic to i∗ (OP (1)).
It is equivalent (4.2.3) to say that there exists a quasi-coherent OY -module E and a surjective
homomorphism ϕ : q∗ (E ) → L such that rL ,ϕ : X → P(E ) is an immersion.
We note that the existence of a very ample (for Y) OX -module implies that q is separated ((3.1.3)
and (I, 5.5.1, (i) and (ii))).
Corollary (4.4.3). — Suppose that there exists a quasi-coherent graded OY -algebra S , generated by S1 ,
and a Y-immersion i : X → P = Proj(S ) such that L is isomorphic to i∗ (OP (1)); then L is very ample
relative to q.
P ROOF. If F = S1 , then the canonical homomorphism S(F ) → S is surjective, and so, by
compositing with the corresponding closed immersion Proj(S ) → P(F ) (3.6.2) and the immersion
i, we obtain an immersion j : X → P(F ) = P′ such that L is isomorphic to j∗ (OP′ (1)) (3.6.3). □
Proposition (4.4.4). — Let q : X → Y be a quasi-compact morphism, and L an invertible OX -module.
Then the following properties are equivalent:
(a) L is very ample relative to q.
(b) q∗ (L ) is quasi-coherent, the canonical homomorphism σ : q∗ (q∗ (L )) → L is surjective, and the
morphism rL ,σ : X → P(q∗ (L )) is an immersion.
P ROOF. Since q is quasi-compact, we know that q∗ (L ) is quasi-coherent if q is separated
(I, 9.2.2).
II | 80
4. PROJECTIVE BUNDLES; AMPLE SHEAVES 257
n⩾0
and so (3.8.3) the hypothesis that rL ,ϕ is an immersion implies that so too is j = rL ,σ ; furthermore
(4.2.4), L is isomorphic to j∗ (OP′ (1)), where P′ = P(q∗ (L )). We thus see that (a) and (b) are
equivalent. □
Corollary (4.4.5). — Suppose that q is quasi-compact. For L to be very ample relative to Y, it is necessary
and sufficient for there to exist an open cover (Uα ) of Y such that L |q−1 (Uα ) is very ample relative to Uα for
every α.
P ROOF. Indeed, condition (b) of (4.4.4) is local on Y. □
Proposition (4.4.6). — Let Y be a quasi-compact scheme, or a prescheme whose underlying space is
Noetherian, q : X → Y a morphism of finite type, and L an invertible OX -module. Then conditions (a) and
(b) of (4.4.4) are equivalent to the following:
(a’) There exists a quasi-coherent OY -module E of finite type and a surjective homomorphism ϕ :
q∗ (E ) → L such that rL ,ϕ is an immersion.
(b’) There exists a coherent sub-OY -module E of q∗ (L ) of finite type that has the properties stated in
condition (a’).
P ROOF. It is clear that (a’) or (b’) imply (a); also (a) implies (a’), by (4.4.1), and similarly (b)
implies (b’). □
Corollary (4.4.7). — Suppose that Y is a quasi-compact scheme, or a Noetherian prescheme. If L is
very ample for q, then there exists a quasi-coherent graded OY -algebra S such that S1 is of finite type and
generates S , and also a dominant open Y-immersion i : X → P = Proj(S ) such that L is isomorphic to
i∗ (OP (1)).
P ROOF. Indeed, condition (b) of (4.4.6) is satisfied; the structure morphism p : P(E ) = P′ → Y
is then separated and of finite type (3.1.3), and so P′ is a quasi-compact scheme (resp. a Noetherian
prescheme) if Y is a quasi-compact scheme (resp. a Noetherian prescheme). Let Z be the closure
(I, 9.5.11) of the subprescheme X ′ of P′ associated to the immersion j = rL ,ϕ from X into P′ ; it is
clear that j factors as a dominant open immersion i : X → Z followed by the canonical injection
Z → P′ . But Z can be identified with a prescheme Proj(S ), where S is a graded OY -algebra equal
to the quotient of S(E ) by a quasi-coherent graded sheaf of ideals (3.6.2), and it is clear that S1 is
of finite type and generates S ; furthermore, OZ (1) is the inverse image of OP′ (1) by the canonical
injection (3.6.3), and so L = i∗ (OZ (1)). □
Proposition (4.4.8). — Let q : X → Y be a morphism, L a very ample (relative to q) OX -module, and L ′ an
invertible OX -module, such that there exists a quasi-coherent OY -module E ′ and a surjective homomorphism
q∗ (E ′ ) → L ′ . Then L ⊗OX L ′ is very ample relative to q.
P ROOF. The hypothesis implies the existence of a Y-morphism r ′ : X → P′ = P(E ′ ) such that
L′ = r ′∗ (OP′ (1)) (4.2.1). There is, by hypothesis, a quasi-coherent OY -module E and a Y-immersion II | 81
r : X → P = P(E ) such that L = r ∗ (OP (1)). Let Q = P(E ⊗ E ′ ), and consider the Segre morphism
ς : P ×Y P′ → Q (4.3.1). Since r is an immersion, so too is (r, r ′ )Y : X → P ×Y P′ (I, 5.3.14); but
(r,r ′ ) ς
since ς is an immersion (4.3.3), so too is r ′′ : X −−→ P ×Y P′ − → Q. But also (4.3.2.1) ς(OQ (1)) is
isomorphic to OP (1) ⊗Y OP′ (1), and so (I, 9.1.4) r ′′∗ (OQ (1)) is isomorphic to L ⊗ L ′ , which proves
the proposition. □
Corollary (4.4.9). — Let q : X → Y be a morphism.
(1) Let L be an invertible OX -module, and K an invertible OY -module. For L to be very ample
relative to q, it is necessary and sufficient for L ⊗ q∗ (K ) to be so.
(2) If L and L ′ are very ample (relative to q) OX -modules, then so too is L ⊗ L ′ ; in particular, L ⊗n
is very ample relative to q for all n > 0.
4. PROJECTIVE BUNDLES; AMPLE SHEAVES 258
P ROOF. Claim (ii) is an immediate consequence of (4.4.8), as well as the necessity of condition (i);
conversely, if L ⊗ q∗ (K ) is very ample, then so too is (L ⊗ q∗ (K )) ⊗ q∗ (K −1 ), by the above, and
the latter OX -module is isomorphic to L ((0, 5.4.3) and (0, 5.4.5)). □
Proposition (4.4.10). —
(i) For every prescheme Y, every invertible OY -module L is very ample relative to the identity
morphism 1Y .
(i bis) Let f : X → Y be a morphism, and j : X ′ → X an immersion. If L is a very ample (relative to f )
OX -module, then j∗ (L ) is very ample relative to f ◦ j.
(ii) Let Z be a quasi-compact prescheme, f : X → Y a morphism of finite type, g : Y → Z a quasi-
compact morphism, L a very ample (relative to f ) OX -module, and K a very ample (relative to g)
OY -module. Then there exists some integer n0 > 0 such that L ⊗ f ∗ (K ⊗n ) is very ample relative
to g ◦ f for all n ⩾ n0 .
(iii) Let f : X → Y and g : Y ′ → Y be morphisms, and let X ′ = X(Y ′ ) . If L is a very ample (relative
to f ) OX -module, then L ′ = L ⊗Y OY ′ is a very ample (relative to f (Y ′ ) ) OX -module.
(iv) Let f i : Xi → Yi (i = 1, 2) be S-morphism. If Li is a very ample (relative to f i ) OXi -module
(i = 1, 2), then L1 ⊗S L2 is very ample relative to f 1 ×S f 2 .
(v) Let f : X → Y and g : Y → Z be morphisms. If an OX -module L is very ample relative to g ◦ f ,
then it is also very ample relative to f .
(vi) Let f : X → Y be a morphism, and j the canonical injection Xred → X. If an OX -module L is
very ample relative to f , then j∗ (L ) is very ample relative to f red .
P ROOF. Property (i bis) follows immediately from the definition (4.4.2), and it is immediate that
(vi) follows formally from (i bis) and (v), by an argument copied from the proof of (I, 5.5.12), which we
Γf p2
leave to the reader. To prove (v), we consider, as in (I, 5.5.12), the factorisation X −→ X × Z Y −→ Y,
where p2 = ( g ◦ f ) × 1Y . It follows from the hypothesis and from (i) and (iv) that L ⊗OZ OY is very
ample for p2 ; but also L = Γ∗f (L ⊗OZ OY ) (I, 9.1.4), and Γ f is an immersion (I, 5.3.11); we can thus
apply (i bis).
To prove (i), we apply the definition (4.4.2) with E = L , and note that then P(E ) can be II | 82
identified with Y (4.1.4).
Now we prove (iii). There exists a quasi-coherent OY -module E and a Y-immersion i : X →
P(E ) = P such that L = i∗ (OP (1)); if we let E ′ = g∗ (E ), then E ′ is a quasi-coherent OY ′ -module,
and we have that P′ = P(E ′ ) = P(Y ′ ) (4.1.3.1), that i(Y ′ ) is an immersion from X(Y ′ ) into P′ (I, 4.3.2),
and that L ′ is isomorphic to (i(Y ′ ) )∗ (OP′ (1)) (4.2.10).
To prove (iv), note that there is, by hypothesis, a Yi -immersion ri : Xi → Pi = P(Ei ), where Ei is
a quasi-coherent OYi -module, and Li = ri∗ (OPi (1)) (i = 1, 2); r1 ×S r2 is an S-immersion of X1 ×S X2
into P1 ×S P2 (I, 4.3.1), and the inverse image of OP1 (1) ⊗S OP2 (1) under this immersion is L1 ⊗S L2 .
Now let T = Y1 ×S Y2 , and let p1 and p2 be the projections from T to Y1 and Y2 , respectively. If we
let Pi′ = P( pi∗ (Ei )) (i = 1, 2), then Pi′ = Pi ×Yi T, by (4.1.3.1), and so
P1′ × T P2′ = ( P1 ×Y1 T ) × T ( P2 ×Y2 T ) = P1 ×Y1 ( T ×Y2 P2 ) = P1 ×Y1 (Y1 ×S P2 ) = P1 ×S P2
up to canonical isomorphism. Similarly, OP′ (1) = OPi (1) ⊗Yi OT (4.1.3.2), and an analogous cal-
i
culation (based in particular on (I, 9.1.9.1) and (I, 9.1.2)) shows that, in the above identification,
OP′ (1) ⊗ T OP′ (1) can be identified with OP1 ⊗S OP2 (1). We can thus consider r1 ×S r2 as a T-
1 2
immersion from X1 ×S X2 into P1′ × T P2′ , with the inverse image of OP′ (1) ⊗ T OP′ (1) under this
1 2
immersion being L1 ⊗S L2 . We then finish the argument as in (4.4.8) by using the Segre morphism.
It remains only to prove (ii). We can first of all restrict to the case where Z is an affine scheme,
since, in general, there exists a finite cover (Ui ) of Z by affine opens; if the proposition were proven
for K | g−1 (Ui ), L | f −1 ( g−1 (U )), and an integer ni , then it would suffice to take n0 to be the largest
of the ni to prove the proposition for K and L (4.4.5). The hypothesis implies that f and g are
separated morphisms, and so X and Y are quasi-compact schemes.
There is an immersion r : X → P = P(E ), where E is a quasi-coherent OY -module of finite type,
and L = r ∗ (OP (1)), by (4.4.6). We will see that there exists a very ample (relative to the composed
h g ⊗(−m)
morphism P − → Z) OP -module M such that OP (1) is isomorphic to M ⊗Y K
→Y− for some
4. PROJECTIVE BUNDLES; AMPLE SHEAVES 259
integer m. For n ⩾ m + 1, OP (1) ⊗Y K ⊗n will then be very ample for Z, by hypothesis and by
(iv) applied to the morphisms h : P → Y and 1Y ; since r is an immersion and L ⊗ f ∗ (K ⊗n ) =
r ∗ (OP (1) ⊗Y K ⊗n ), the conclusion will then follow from (i bis). To prove our claim concerning
OP (1), we will use the following lemma:
Lemma (4.4.10.1). — Let Z be a quasi-compact scheme, or a prescheme whose underlying space is Noetherian,
and let g : Y → Z be a quasi-compact morphism, K a very ample (with respect to g) invertible OY -module,
and E a quasi-coherent OY -module of finite type. Then there exists an integer m0 such that, for all m ⩾ m0 ,
E is isomorphic to a quotient of an OY -module of the form g∗ (F ) ⊗ K ⊗(−m) , where F is a quasi-coherent
OZ -module of finite type (depending on m).
This lemma will be proven in (4.5.10.1); the reader can verify that (4.4.10) is not used anywhere
in (4.5).
Assuming this lemma, there exists a closed immersion j1 from P to II | 83
∗ ⊗(−m)
P1 = P( g (F ) ⊗ K )
such that OP (1) is isomorphic to j1∗ (OP1 (1)) (4.1.2). Now, there exists an isomorphism from P1 to
P2 = P( g∗ (F )), sending OP2 (1) ⊗Y K ⊗(−m) to OP1 (1) (4.1.4); we thus have a closed immersion
j2 : P → P2 such that OP (1) is isomorphic to j2∗ (OP2 (1)) ⊗Y K ⊗(−m) . Finally, P2 can be identified
with P3 × Z Y, where P3 = P(F ), and OP2 (1) with OP3 (1) ⊗ Z OY (4.1.3). By definition, OP3 (1) is very
ample for Z; since so too is K , we conclude, from (iv), that OP2 (1) ⊗Y K is very ample for Z; so too
is M = j2∗ (OP2 (1) ⊗Y K ) by (i bis), and OP (1) is isomorphic to M ⊗Y K ⊗(−m−1) , which finishes
the proof. □
Proposition (4.4.11). — Let f : X → Y and f ′ : X ′ → Y be morphisms, X ′′ the sum prescheme X ⊔ X ′ ,
and f ′′ the morphism X ′′ → Y that agrees with f (resp. f ′ ) on X (resp. X ′ ). Let L (resp. L ′ ) be an
invertible OX -module (resp. invertible OX ′ -module), and let L ′′ be the invertible OX ′′ -module that agrees
with L (resp. L ′ ) on X (resp. X ′ ). For L ′′ to be very ample relative to f ′′ , it is necessary and sufficient for
L to be very ample relative to f and for L ′ to be very ample relative to f ′ .
P ROOF. We can immediately restrict to the case where Y is affine. If L ′′ is very ample then
so too are L and L ′ , by (4.4.10, (i bis)). Conversely, if L and L ′ are very ample, then it follows
immediately from the definition (4.4.2) and from (4.3.6) that L ′′ is very ample. □
4.5. Ample sheaves
(4.5.1). Given a prescheme X and an invertible OX -module L , we define, for every OX -module F
(when there will be no confusion possible over L ) F (n) = F ⊗OX L ⊗n (n ∈ Z); we also define
S = n⩾0 Γ( X, L ⊗n ) (a graded subring of the ring Γ• (L ) defined in (0, 5.4.6)). If we consider
L
and endomorphisms of the graded ring S (I, 2.2.5); the homomorphism ε : p∗ (Se) → n⩾0 L ⊗n that
L
equivalent:
(a) When f runs over the set of homogeneous elements of S+ , the X f form a base for the topology of X.
(a’) When f runs over the set of homogeneous elements of S+ , the X f that are affine form a cover of X.
(b) The canonical morphism G (ε) → Proj(S) (4.5.1) is everywhere defined and is a dominant open
immersion. II | 84
(b’) The canonical morphism G (ε) → Proj(S) is everywhere defined and is a homeomorphism from the
underlying space of X to a subspace of Proj(S).
(c) For every quasi-coherent OX -module F , if we denote by Fn the sub-OX -module of F (n) generated
by the sections of F (n) over X, then F is the sum of the sub-OX -modules Fn (−n) over the
integers n > 0.
(c’) Property (c) holds for every quasi-coherent sheaf of ideals of OX .
4. PROJECTIVE BUNDLES; AMPLE SHEAVES 260
Furthermore, if ( f α ) is a family of homogeneous elements of S+ such that the X f α are affine, then
the restriction to α X f α of the canonical morphism X → Proj(S) is an isomorphism from α X f α to
S S
S
α (Proj( S )) f α .
P ROOF. It is clear that (b) implies (b’), and (b’) implies (a) by (3.7.3.1) (taking into account the
fact that ε♭ is the identity). Condition (a) implies (a’), since every x ∈ X has an affine neighbourhood
U such that L |U is isomorphic to OX |U; if f ∈ Γ( X, L ⊗n ) is such that x ∈ X f ⊂ U, then X f is
also the set of x ′ ∈ U such that ( f |U )( x ′ ) ̸= 0, and it is thus an affine open subset (I, 1.3.6). To
prove that (a’) implies (b), it suffices to prove the last claim of the theorem, and to further prove
S
that, if X = α X f α , then condition (iv) of (3.8.2) is satisfied. This latter point follows immediately
from (I, 9.3.1, (i)). As for the last claim of (4.5.2), since X f α is the inverse image of (Proj(S)) f α under
G (ε) → Proj(S), it suffices to apply (I, 9.3.2). Thus (a), (a’), (b), and (b’) are all equivalent.
To show that (a’) implies (c), note that, if X f is affine (with f ∈ Sk ), then F | X f is generated
by its sections over X f (I, 1.3.9); on the other hand (I, 9.3.1, (ii)), such a section s is of the form
(t| X f ) ⊗ ( f | X f )−m , where t ∈ Γ( X, F (km)); by definition, t is also a section of Fkm , so s is indeed
a section of Fkm (−km) over X f , which proves (c). It is clear that (c) implies (c’), so it remains
only to show that (c’) implies (a). But let U be an open neighbourhood of x ∈ X, and let J be
a quasi-coherent sheaf of ideals of OX defining a closed subprescheme of X that has X − U as its
underlying space (I, 5.2.1). Hypothesis (c’) implies that there exists an integer n > 0 and a section f
of J (n) over X such that f ( x ) ̸= 0. But we clearly have f ∈ Sn , and x ∈ X f ⊂ U, which proves
(a). □
When X is a prescheme whose underlying space is Noetherian, the equivalent conditions of
(4.5.2) imply that X is a scheme, since it is isomorphic to a subprescheme of the scheme S = Proj( A),
by (4.5.2, (b)).
Definition (4.5.3). — We say that an invertible OX -module L is ample if X is a quasi-compact
scheme and if the equivalent conditions of (4.5.2) are satisfied.
It evidently follows from criterion (a) of (4.5.2) that, if L is an ample OX -module, then, for every
open subset U of X, L |U is an ample (OX |U )-module.
It follows from the proof of (4.5.2) that the affine X f form a base for the topology of X. Further-
more:
Corollary (4.5.4). — Let L be an ample OX -module. For every finite subspace Z of X and every neighbour-
hood U of Z, there exists an integer n and some f ∈ Γ( X, L ⊗n ) such that X f is an affine neighbourhood of
Z contained in U.
P ROOF. By (4.5.2, (b)), it suffices to prove that, for every finite subset Z ′ of Proj(S) and ev- II | 85
ery open neighbourhood U of Z ′ , there exists a homogeneous element f ∈ S+ such that Z ⊂
(Proj(S)) f ⊂ U (2.4.1). But, by definition, the closed set Y, complement of U in Proj(S), is of the
form V+ (I), where I is a graded ideal of S that does not contain S+ (2.3.2); also, the points of Z ′ are,
by definition, graded ideals pi of S+ that do not contain J (2.3.1). There thus exists an element f ∈ I
that does not belong to any of the pi (Bourbaki, Alg. comm., chap. II, §1, no. 1, prop. 2), and, since the
pi are graded, the argument made loc. cit. shows that we can even take f to be homogeneous; this
element then satisfies the claim. □
Proposition (4.5.5). — Suppose that X is a quasi-compact scheme or a prescheme whose underlying space is
Noetherian. Then conditions (a) to (c’) of (4.5.2) are equivalent to the following:
(d) For every quasi-coherent OX -module F of finite type, there exists an integer n0 such that, for all
n ⩾ n0 , F (n) is generated by its sections over X.
(d’) For every quasi-coherent OX -module F of finite type, there exist integers n > 0 and k > 0 such
that F is isomorphic to a quotient of the OX -module L ⊗(−n) ⊗ OXk .
(d”) Property (d’) holds for every quasi-coherent sheaf of ideals of OX of finite type.
P ROOF. Since X is quasi-compact, if a quasi-coherent OX -module F of finite type is such that
F (n) (which is of finite type) is generated by its sections over X, then F (n) is generated by a
finite number of these sections (0, 5.2.3), and so (d) implies (d’), and it is clear that (d’) implies (d”).
4. PROJECTIVE BUNDLES; AMPLE SHEAVES 261
Since every quasi-coherent OX -module G is the inductive limits of its sub-OX -modules of finite type
(I, 9.4.9), to satisfy condition (c’) of (4.5.2), it suffices to do so for a quasi-coherent sheaf of ideals
of OX that is of finite type, and (d”) thus implies (c’). It remains only to show that, if L is ample,
then property (d) is satisfied. Consider a finite cover of X by X f i ( f i ∈ Sni ), that we can assume to
be affine; by replacing the f i with suitable powers (which does not alter the X f i ), we can assume
that all the ni are equal to one single integer m. The sheaf F | X f i , being of finite type, by hypothesis,
is generated by a finite number of its sections hij over X f i (I, 1.3.13); so there exists an integer k0
⊗k0
such that the section hij ⊗ f i extends to a section of F (k0 m) over X for every pair (i, j) (I, 9.3.1).
⊗k
fi 0
A fortiori, the hij ⊗ extend to sections of F (km) over X for every k ⩾ k0 , and, for these values of
k, F (km) is thus generated by its sections over X. For every p such that 0 < p < m, F ( p) is also of
finite type, and so there exists an integer k p such that F ( p)(km) = F ( p + km) is generated by its
sections over X for all k ⩾ k p . Taking n0 to be the largest of the k p m, we thus conclude that F (n)
is generated by its sections over X for all n ⩾ n0 , since such an n is of the form n = km + p, with
k ⩾ k p and 0 ⩽ p < m. □
Remark (4.5.9). — Let P = H1 ( X, OX× ) be the group of classes of invertible OX -modules (0, 5.4.7),
and let P+ be the subset of P consisting of classes of ample sheaves. Suppose that P+ is non-empty.
Then it follows from (4.5.7) and (4.5.8) that
P+ + P+ ⊂ P+ and P+ − P+ = P
or, in other words, P+ ∪ {0} is the set of positive elements in P for a preorder structure on P that is
compatible with its group structure, and is even archimedean, by (4.5.8). This is why we sometimes
say “positive sheaf” instead of ample sheaf, and “negative sheaf” for the inverse of an ample sheaf
(but we will not use this terminology).
Proposition (4.5.10). — Let Y be an affine scheme, q : X → Y a quasi-compact separated morphism, and
L an invertible OX -module.
(i) If L is very ample for q, then L is ample.
(ii) Suppose further that the morphism q is of finite type. Then, for L to be ample, it is necessary and
sufficient for it to posses one of the following properties:
(e) There exists n0 > 0 such that, for every integer n ⩾ n0 , L ⊗n is very ample for q.
(e’) There exists n > 0 such that L ⊗n is very ample for q.
P ROOF. The first claim follows from the definition (4.4.2) of a very ample OX -module: if A is
the ring of Y, then there exists an A-module E and a surjective homomorphism
ψ : q∗ ((S( E))e ) −→ L ⊗n
M
n⩾0
4. PROJECTIVE BUNDLES; AMPLE SHEAVES 262
Remark (4.5.12). — We do not know if the hypothesis that an OX -module L is such that L ⊗n is
very ample (relative to q) implies the same conclusion for L ⊗(n+1) .
Proposition (4.5.13). — Let X be a quasi-compact prescheme, Z a closed prescheme of X defined by a
nilpotent quasi-coherent sheaf J of ideals of OX , and j the canonical injection Z → X. For an invertible
OX -module L to be ample, it is necessary and sufficient for L ′ = j∗ (L ) to be an ample OZ -module.
P ROOF. The condition is necessary. Indeed, for every section f of L ⊗n over X, let f ′ be its
canonical image f ⊗ 1, which is a section of L ′⊗n = L ⊗n ⊗OX (OX /J ) over the space Z (which is
identical to X); it is clear that X f = Z f ′ , and so the criterion (a) of (4.5.2) shows that L ′ is ample.
To see that the condition is sufficient, note first of all that we can restrict to the case where J 2 = 0,
by considering the (finite) sequence of preschemes Xk = ( X, OX /J k+1 ) with each prescheme being
a closed subprescheme of the next, defined by a square-zero sheaf of ideals. But X is a scheme, since
Xred is a scheme by hypothesis ((4.5.3) and (I, 5.5.1)). Criterion (a) of (4.5.2) shows that it suffices to
prove
Lemma (4.5.13.1). — Under the hypotheses of (4.5.13), suppose further that J is square-zero; with L
being an invertible OX -module, let g be a section of L ′⊗n over Z such that Zg is affine. Then there exists an
integer m > 0 such that g⊗m is the canonical image of a section f of L ⊗nm over X.
P ROOF. We have the exact sequence of OX -modules
0 −→ J (n) −→ OX (n) = L ⊗n −→ OZ (n) = L ′⊗n −→ 0
since F (n) is an exact functor in F ; from this, we have the exact sequence of cohomology II | 88
0 −→ Γ( X, J (n)) −→ Γ( X, L ⊗n ) −→ Γ( X, L ′⊗n ) −
→ H1 ( X, J (n))
∂
Note that, since J 2 = 0, J can be considered as a quasi-coherent OZ -module, and, for all
k, L ′⊗k ⊗OZ J (n) = J (n + k ); for every section s ∈ Γ( X, L ′⊗k ), tensor multiplication with s is
4. PROJECTIVE BUNDLES; AMPLE SHEAVES 263
(a) L is f -ample.
4. PROJECTIVE BUNDLES; AMPLE SHEAVES 264
homomorphism
∗
(4.6.3.2) rL ,σ (M ) −→ F
f
P ROOF. We note that (a) implies that f is separated, and thus that S is quasi-coherent (I, 9.2.2, (a)).
Since rL ,σ being an everywhere defined immersion is of a local (on Y) character, to prove that (a)
implies (b) we can suppose that Y is affine and L ample; the claim then follows from (4.5.2, (b)). It II | 90
is clear that (b) implies (b’); finally, to prove that (b’) implies (a), it suffices to consider an affine open
cover (Uα ) of Y and to apply the criterion (4.5.2, (b’)) to each sheaf L | f −1 (Uα ).
For the final two claims, we use the fact that σ♭ is here the identity, and the clarification of
the homomorphisms (3.7.9.1) and (3.7.9.2); from this, it immediately follows that (4.6.3.1) is an
isomorphism. As for (4.6.3.2), we can restrict to the case where Y is affine, and thus L ample; it is
clear that the homomorphism (4.6.3.2) is injective, and criterion (4.5.2, (c)) shows that it is surjective,
whence the conclusion. □
Corollary (4.6.4). — Let (Uα ) be an open cover of Y. For L to be ample relative to Y, it is necessary and
sufficient for L | f −1 (Uα ) to be ample relative to Uα for all α.
Corollary (4.6.5). — Let K be an invertible OY -module. For L to be Y-ample, it is necessary and sufficient
for L ⊗ f ∗ (K ) to be Y-ample.
P ROOF. This is an evident consequence of (4.6.4), by taking the Uα to be such that K |Uα is
isomorphic to OY |Uα for all α. □
Corollary (4.6.6). — Suppose that Y is affine; for L to be Y-ample, it is necessary and sufficient for L to be
ample.
P ROOF. This is an immediate consequence of the definition (4.6.1) and the criteria (4.6.3, (b))
and (4.5.2, (b)), since here Proj(S ) = Proj(Γ(Y, S )) by definition. □
Corollary (4.6.7). — Let f : X → Y be a quasi-compact morphism. Suppose that there exists a quasi-
coherent OY -module E , and a Y-morphism g : X → P = P(E ) that is a homeomorphism from the underlying
space of X to a subspace of P; then L = g∗ (OP (1)) is Y-ample.
P ROOF. We can assume Y to be affine; the corollary then follows from the criterion (4.5.2, (a)),
from equation (3.7.3.1), and from (4.2.3). □
P ROOF. Since X is quasi-compact, so too is f ( X ), and so there exists a finite cover (Ui ) of f ( X )
consisting of affine open subsets Ui of Y. To prove condition (c) when L is f -ample, we can replace
Y by the Ui , and X by the f −1 (Ui ), since, if we obtain, for each i, an integer ni such that (c) holds
true (for Ui , f −1 (Ui ), and L | f −1 (Ui )) for all n ⩾ ni , then it suffices to take n0 to be the largest of
the ni in order to obtain (c) for Y, X, and L . But if Y is affine, condition (c) follows from (4.5.5, (d)),
taking (4.6.6) into account. It is trivial that (c) implies (c’). Finally, to prove that (c’) implies that L is
f -ample, we can again restrict to the case where Y is affine: in fact, every quasi-coherent sheaf Ji of
ideals of OX | f −1 (Ui ) of finite type is the restriction of a coherent sheaf of ideals of OX of finite type
(I, 9.4.7), and hypothesis (c’) implies that Ji ⊗ (L ⊗n | f −1 (Ui )) is generated by its sections (taking II | 91
(I, 9.2.2) and (3.4.7) into account); it thus suffices to apply criterion (4.5.5, (d”)). □
Proposition (4.6.9). — Let f : X → Y be a quasi-compact morphism, and L an invertible OX -module.
(i) Let n > 0 be an integer. For L to be f -ample, it is necessary and sufficient for L ⊗n to be f -ample.
(ii) Let L ′ be an invertible OX -module, and suppose that there exists an integer n > 0 such that the
canonical homomorphism σ : f ∗ ( f ∗ (L ′⊗n )) → L ′⊗n is surjective. Then, if L if f -ample, so too
is L ⊗ L ′ .
P ROOF. We can in fact immediately restrict to the case where Y is affine, and the proposition is
then an immediate consequence of (4.5.6). □
Corollary (4.6.10). — The tensor product of two f -ample OX -modules is f -ample.
Proposition (4.6.11). — Let Y be a quasi-compact prescheme, f : X → Y a morphism of finite type, and
L an invertible OX -module. For L to be f -ample, it is necessary and sufficient for it to posses one of the
following equivalent properties:
(d) There exists some n0 > 0 such that, for every integer n ⩾ n0 , L ⊗n is very ample relative to f .
(d’) There exists some n > 0 such that L ⊗n is very ample relative to f .
P ROOF. If L is ample relative to f , then there exists a finite cover (Ui ) of Y by affine open
subsets such that the L | f −1 (Ui ) are ample. We thus conclude (4.5.10) that there exists an integer n0
such that L ⊗n | f −1 (Ui ) is very ample relative to f −1 (Ui ) → Ui for all n ⩾ n0 and every i, and so
L ⊗n is very ample relative to f (4.5.5). Conversely, (d’) already implies that L ⊗n is f -ample (4.6.2),
and thus so too is L (4.6.9, (i)). □
Corollary (4.6.12). — Let Y be a quasi-compact prescheme, f : X → Y a morphism of finite type, and L
and L ′ invertible OX -modules. If L is f ample, then there exists some n0 such that L ⊗n ⊗ L ′ is very
ample relative to f for all n ⩾ n0 .
P ROOF. We argue as in (4.6.11), by using a finite affine open cover of Y and (4.5.11). □
Proposition (4.6.13). — (i) For every prescheme Y, every invertible OY -module L is ample relative
to the identity morphism 1Y .
(i bis) Let f : X → Y be a quasi-compact morphism, and j : X ′ → X a quasi-compact morphism that is a
homeomorphism from the underlying space of X ′ to a subspace of X. If L is an OX -module that is
ample relative to f , then j∗ (L ) is ample relative to f ◦ j.
(ii) Let Z be a quasi-compact prescheme, f : X → Y and g : Y → Z quasi-compact morphisms, L an
OX -module that is ample relative to f , and K an OY -module that is ample relative to g. Then there
exists an integer n0 > 0 such that L ⊗ f ∗ (K ⊗n ) is ample relative to g ◦ f for all n ⩾ n0 .
(iii) Let f : X → Y be a quasi-compact morphism, and g : Y ′ → Y a morphism, and let X ′ = X(Y ′ ) . If
L is an OX -module that is ample relative to f , then L ′ = L ⊗Y OY ′ is an OX ′ -module that is
ample relative to f (Y ′ ) .
(iv) Let f i : Xi → Yi (i = 1, 2) be quasi-compact S-morphisms. If Li is an OXi -modules that is ample
relative to f i (i = 1, 2), then L1 ⊗S L2 is ample relative to f 1 ×S f 2 .
(v) Let f : X → Y and g : Y → Z be morphisms such that g ◦ f is quasi-compact. If an OX -module II | 92
L is ample relative to g ◦ f , and if g is separated or the underlying space of X is locally Noetherian,
then L is ample relative to f .
(vi) Let f : X → Y be a quasi-compact morphism, and j the canonical injection Xred → X. If L is an
OX -module that is ample relative to f , then j∗ (L ) is ample relative to f red .
4. PROJECTIVE BUNDLES; AMPLE SHEAVES 266
P ROOF. Note first of all that (v) and (vi) follow from (i), (i bis), and (iv) by the same argument
as in (4.4.10), by using (4.6.4) instead of (4.4.5); we leave the details of the argument to the reader.
Claim (i) is trivially a consequence of (4.4.10, (i)) and (4.6.2). To prove (i bis), (iii), and (iv), we will
use the following lemma:
Lemma (4.6.13.1). — (i) Let u : Z → S be a morphism, L an invertible OS -module, and s a section
of L over S, and let s′ be the canonically corresponding section u∗ (L ) = L ′ over Z. Then
Zs′ = u−1 (Ss ).
(ii) Let Z and Z ′ be S-preschemes, p and p′ the projections of T = Z ×S Z ′ , L (resp. L ′ ) an invertible
OZ -module (resp. invertible OZ′ -module), and let t (resp. t′ ) be a section of L (resp. L ′ ) over Z
(resp. Z ′ ), and s (resp. s′ ) the corresponding section of p∗ (L ) (resp. p′∗ (L ′ )) over Z ×S Z ′ . Then
Ts⊗s′ = Zt ×S Zt′′ .
P ROOF. It follows from the definitions that we can restrict to the case where all the preschemes
in question are affine. Furthermore, in (i), we can suppose that L = OB ; claim (i) then follows
immediately from (I, 1.2.2.2). Similarly, in (ii), we can restrict to the case where L = OZ and
L ′ = OZ′ , and then the claim reduces to Lemma (4.3.2.4). □
We now prove (i bis). We can assume that Y is affine (4.6.4), and thus that L is ample (4.6.6); if s
runs over the set of the union of the Γ( X, L ⊗n ) (n > 0), then the Xs form a base for the topology of
X (4.5.2, (a)), and so, by hypothesis, the j−1 ( Xs ) form a base for the topology of X ′ ; we thus conclude,
by Lemma (4.6.13.1, (i)) and (4.5.2, (a)), that j∗ (L ) is ample.
Next we prove (iii). We can again suppose that Y and Y ′ are affine (4.6.4), whence it follows that
the projection h : X ′ → X is affine (1.5.5). Since L is ample (4.6.6), if s runs over the set of sections
of the L ⊗n (n > 0) over X such that Xs is affine, then the Xs cover X (4.5.2, (a’)), and so the h−1 ( Xs )
are affine (1.2.5) and cover X ′ ; it thus follows again from Lemma (4.6.13.1, (i)) and from (4.5.2, (a’))
that L ′ is ample, since the morphism f (Y ′ ) is quasi-compact (I, 6.6.4, (iii)).
To prove (iv), note first of all that f 1 ×S f 2 is quasi-compact (I, 6.6.4, (iv)). We can further suppose
that S, Y1 , and Y2 are affine ((4.6.4) and (I, 3.2.7)), and thus that Li is ample (i = 1, 2) (4.6.6). The
⊗n
open subsets ( X1 )s1 ×S ( X2 )s2 form a cover of X1 ×S X2 , where si runs over the sections of Li i
such that ( Xi )si is affine (4.5.2, (a’)). Then, replacing s1 and s2 with suitable powers, which does not
change the ( Xi )si , we can assume that n1 = n2 . We thus deduce, from (4.6.13.1, (ii)) and (4.5.2, (a’)),
that L1 ⊗S L2 is ample, whence the claim, since Y1 ×S Y2 is affine (4.6.6).
It remains only to prove (ii). By the same argument as in (4.4.10), but here using (4.6.4), we
can restrict to the case where Z is affine. Since K is then ample, and Y quasi-compact, there exists
a finite number of sections si ∈ Γ(Y, K ⊗ki ) such that the Ysi are affine and cover Y (4.5.2, (a’)); II | 93
replacing the si with suitable powers, we can further suppose that all the k i are equal to one single
integer k. Let si′ be the sections of f ∗ (K ⊗k ) over X that canonically correspond to the si , so that the
Xs′ = f −1 (Ysi ) (4.6.13.1, (i)) cover X. Since L | Xs′ is ample ((4.6.4) and (4.6.6)), there exists, for each
i i
i, a finite number of sections tij ∈ Γ( X, L ⊗nij ) such that the Xij are affine, contained in the Xs′ , and
i
cover Xs′ (4.5.2, (a’)); we can also suppose that all the nij are equal to one single integer n. With this
i
in mind, X is separated and quasi-compact, and so there exists an integer m > 0, and, for every (i, j),
a section
uij ∈ Γ( X, L ⊗n ⊗ X f ∗ (K ⊗mk ))
such that tij ⊗ si′⊗m is the restriction to Xs′ of uij (I, 9.3.1); furthermore, Xuij = Xtij , and so the Xuij
i
are affine and cover X. We can also suppose that m is of the form nr; if we set n0 = rk, then we
see (4.5.2, (a’)) that L ⊗OX f ∗ (K ⊗n0 ) is ample. Furthermore, there exists h0 > 0 such that K ⊗h is
generated by its sections over Y for all h ⩾ h0 (4.5.5); a fortiori, f ∗ (K ⊗h ) is generated by its sections
over X for all h ⩾ h0 , by definition of the inverse images ((0, 3.7.1) and (4.4.1)). We thus conclude
that L ⊗ f ∗ (K ⊗n0 +h ) is ample for all h ⩾ h0 (4.5.6), which finishes the proof. □
Remark (4.6.14). — Under the conditions of (ii), we refrain from believing that L ⊗ f ∗ (K ) is ample
for g ◦ f ; in fact, since L ⊗ f ∗ (K −1 ) is also ample for f (4.6.5), we would conclude that L is ample
for g ◦ f ; taking, in particular, g to be the identity morphism, every invertible OX -module would be
ample for f , which is not the case in general (see (5.1.6), (5.3.4, (i)), and (5.3.1)).
5. QUASI-AFFINE MORPHISMS; QUASI-PROJECTIVE MORPHISMS; PROPER MORPHISMS; PROJECTIVE MORPHISMS 267
(b’) The canonical morphism u : X → Spec( A) is a homeomorphism from X to some subspace of the
underlying space of Spec( A).
(c) The OX -module OX is very ample relative to u (4.4.2).
(c’) The OX -module OX is ample (4.5.1).
(d) When f ranges over A, the X f form a basis for the topology of X.
(d’) When f varies over A, the X f that are affine form a cover of X. II | 95
(e) Every quasi-coherent OX -module is generated by its sections over X.
(e’) Every quasi-coherent sheaf of ideals of OX of finite type is generated by its sections over X.
P ROOF. It is clear that (b) implies (a), and (a) implies (c) by (4.4.4, b) applied to the identity
morphism (taking into account the remark preceding this proposition); Furthermore, (c) implies (c′ )
(4.5.10, i), and (c′ ), (b), and (b′ ) are all equivalent by (4.5.2, b) and (4.5.2, b′ ). Finally, (c′ ) is the same
as each of (d), (d′ ), (e), and (e′ ) by (4.5.2, a), (4.5.2, a′ ), (4.5.2, c), and (4.5.5, d′′ ). □
We further observe that, with the previous notation, the X f that are affine form a basis for the
topology of X, and that the canonical morphism u is dominant (4.5.2).
Corollary (5.1.3). — Let X be a quasi-compact prescheme. If there exists a morphism v : X → Y from X
to some affine scheme Y (which would be a homeomorphism from X to some open subspace of Y), then X is
quasi-affine.
P ROOF. There exists a family ( gα ) of sections of OY over Y such that the D ( gα ) form a basis
for the topology of v( X ); if v = (ψ, θ ) and we set f α = θ ( gα ), then we have X f α = ψ−1 ( D ( gα ))
(I, 2.2.4.1), so the X f α form a basis for the topology of X, and the criterion (5.1.2, d) is satisfied. □
Corollary (5.1.4). — If X is a quasi-affine scheme, then every invertible OX -module is very ample (relative
to the canonical morphism), and a fortiori ample.
P ROOF. Such a module L is generated by its sections over X (5.1.2, e), so L ⊗ OX = L is very
ample (4.4.8). □
Corollary (5.1.5). — Let X be a quasi-compact prescheme. If there exists an invertible OX -module L such
that L and L −1 are ample, then X is a quasi-affine scheme.
P ROOF. Indeed, OX = L ⊗ L −1 is then ample (4.5.7). □
Proposition (5.1.6). — Let f : X → Y be a quasi-compact morphism. Then the following conditions are
equivalent.
(a) The morphism f is quasi-affine.
(b) The OY -algebra f ∗ (OX ) = A ( X ) is quasi-coherent, and the canonical morphism X → Spec(A ( X ))
corresponding to the identity morphism A ( X ) → A ( X ) (1.2.7) is an open immersion.
(b’) The OY -algebra A ( X ) is quasi-coherent, and the canonical morphism X → Spec(A ( X )) is a
homeomorphism from X to some subspace of Spec(A ( X )).
(c) The OX -module OX is very ample for f .
(c’) The OX -module OX is ample for f .
(d) The morphism f is separated, and, for every quasi-coherent OX -module F , the canonical homomor-
phism σ : f ∗ ( f ∗ (F )) → F (0, 4.4.3) is surjective.
Furthermore, whenever f is quasi-affine, every invertible OX -module L is very ample relative to f .
P ROOF. The equivalence between (a) and (c′ ) follows from the local (on Y) character of the
f -ampleness (4.6.4), Definition (5.1.1), and the criterion (5.1.2, c′ ). The other properties are local on
Y and thus follow immediately from (5.1.2) and (5.1.4), taking into account the fact that f ∗ (F ) is II | 96
quasi-coherent whenever f is separated (I, 9.2.2, a). □
Corollary (5.1.7). — Let f : X → Y be a quasi-affine morphism. For every open subset U of Y, the
restriction f −1 (U ) → U of f is quasi-affine.
Corollary (5.1.8). — Let Y be an affine scheme, and f : X → Y a quasi-compact morphism. For f to be
quasi-affine, it is necessary and sufficient for X to be a quasi-affine scheme.
P ROOF. This is an immediate consequence of (5.1.6) and (4.6.6). □
5. QUASI-AFFINE MORPHISMS; QUASI-PROJECTIVE MORPHISMS; PROPER MORPHISMS; PROJECTIVE MORPHISMS 269
Corollary (5.1.9). — Let Y be a quasi-compact scheme or a prescheme whose underlying space is Noetherian,
and f : X → Y a morphism of finite type. If f is quasi-affine, then there exists a quasi-coherent OY -subalgebra
B of A ( X ) = f ∗ (OX ) of finite type (I, 9.6.2) such that the morphism X → Spec(B ) corresponding to the
canonical injection B → A ( X ) is an immersion. Further, every quasi-coherent OY -subalgebra B ′ of finite
type over A ( X ) containing B has the same property.
P ROOF. Indeed, A ( X ) is the inductive limit of its quasi-coherent OY -subalgebras of finite type
(I, 9.6.5); the result is then a particular case of (3.8.4), taking into account the identification of
Spec(A ( X )) with Proj(A ( X )[ T ]) (3.1.7). □
Proposition (5.1.10). —
(i) A quasi-compact morphism X → Y that is a homeomorphism from the underlying space of X to
some subspace of the underlying space of Y (so, in particular, any closed immersion) is quasi-affine.
(ii) The composition of any two quasi-affine morphisms is quasi-affine.
(iii) If f : X → Y is a quasi-affine S-morphism, then f (S′ ) : X(S′ ) → Y(S′ ) is a quasi-affine morphism
for any extension S′ → S of the base prescheme.
(iv) If f : X → Y and g : X ′ → Y ′ are quasi-affine S-morphisms, then f ×S g is quasi-affine.
(v) If f : X → Y and g : Y → Z are morphisms such that g ◦ f is quasi-affine, and if g is separated or
the underlying space of X is locally Noetherian, then f is quasi-affine.
(vi) If f is a quasi-affine morphism, then so is f red .
P ROOF. Taking into account the criterion (5.1.6, c′ ), all of (i), (iii), (iv), (v), and (vi) follow
immediately from (4.6.13, i bis), (4.6.13, iii), (4.6.13, iv), (4.6.13, v), and (4.6.13, vi) (respectively). To
prove (ii), we can restrict to the case where Z is affine, and then the claim follows directly from
applying (4.6.13, ii) to L = OX and K = OY . □
Remark (5.1.11). — Let f : X → Y and g : Y → Z be morphisms such that X × Z Y is locally
Noetherian. Then the graph immersion Γ f : X → X × Z Y is quasi-affine, since it is quasi-compact
(I, 6.3.5), and since (I, 5.5.12) shows that, in (v), the conclusion still holds true if we remove the
hypothesis that g is separated.
Proposition (5.1.12). — Let f : X → Y be a quasi-compact morphism, and g : X ′ → X a quasi-affine
morphism. If L is an ample (for f ) OX -module, then g∗ (L ) is an ample (for f ◦ g) OX ′ -module.
P ROOF. Since OX ′ is very ample for g, and the question is local on Y (4.6.4), it follows from
(4.6.13, ii) that there exists (for Y affine) an integer n such that
g∗ (L ⊗n ) = ( g∗ (L ))⊗n
is ample for f ◦ g, and so g∗ (L ) is ample for f ◦ g (4.6.9) □
5.2. Serre’s criterion
Theorem (5.2.1). — (Serre’s criterion). Let X be a quasi-compact scheme or a prescheme whose underlying
space is Noetherian. The following conditions are equivalent.
(a) X is an affine scheme.
(b) There exists a family of elements f α ∈ A = Γ( X, OX ) such that the X f α are affine, and such that
the ideal generated by the f α in A is equal to A itself.
(c) The functor Γ( X, F ) is exact in F on the category of quasi-coherent OX -modules, or, in other
words, if
(*) 0 −→ F ′ −→ F −→ F ′′ −→ 0
is an exact sequence of quasi-coherent OX -modules, then the sequence
0 −→ Γ( X, F ′ ) −→ Γ( X, F ) −→ Γ( X, F ′′ ) −→ 0
is also exact.
(c’) Condition (c) holds for every exact sequence (∗) of quasi-coherent OX -modules such that F is
isomorphic to an OX -submodule of OXn for some finite n.
(d) H1 ( X, F ) = 0 for every quasi-coherent OX -module F .
(d′ ) H1 ( X, J ) = 0 for every quasi-coherent sheaf of ideals J of OX .
5. QUASI-AFFINE MORPHISMS; QUASI-PROJECTIVE MORPHISMS; PROPER MORPHISMS; PROJECTIVE MORPHISMS 270
P ROOF. It is evident that (a) implies (b); furthermore, (b) implies that the X f α cover X, because,
by hypothesis, the section 1 is a linear combination of the f α , and the D ( f α ) thus cover Spec( A). The
final claim of (4.5.2) thus implies that X → Spec( A) is an isomorphism.
We know that (a) implies (c) (I, 1.3.11), and (c) trivially implies (c′ ). We now prove that (c′ )
implies (b). First of all, (c′ ) implies that, for every closed point x ∈ X and every open neighbourhood
U of x, there exists some f ∈ A such that x ∈ X f ⊂ X − U. Let J (resp. J ′ ) be the quasi-
coherent sheaf of ideals of OX defining the closed reduced subprescheme of X that has X − U
(resp. ( X − U ) ∪ { x }) as its underlying space (I, 5.2.1); it is clear that we have J ′ ⊂ J , and
that J ′′ = J /J ′ is a quasi-coherent OX module that has support equal to { x }, and such that
Jx′′ = k( x ). Hypothesis (c′ ) applied to the exact sequence 0 → J ′ → J → J ′′ → 0 shows that
Γ( X, J ) → Γ( X, J ′′ ) is surjective. The section of J ′′ whose germ at x is 1x is thus the image of
some section f ∈ Γ( X, J ) ⊂ Γ( X, OX ), and we have, by definition, that f ( x ) = 1x and f (y) = 0
in X − U, which establishes our claim. Now, if U is affine, then so is X f (I, 1.3.6), so the union of
the X f that are affine ( f ∈ A) is an open set Z that contains all the closed points of X; since X is a
quasi-compact Kolmogoroff space, we necessarily have Z = X (0, 2.1.3). Because X is quasi-compact,
there are a finite number of elements f i ∈ A (1 ⩽ i ⩽ n) such that the X f i are affine and cover X.
So consider the homomorphism OXn → OX defined by the sections f i (0, 5.1.1); since, for all x ∈ X,
at least one of the ( f i ) x is invertible, this homomorphism is surjective, and we thus have an exact
sequence 0 → R → OXn → OX → 0, where R is a quasi-coherent OX -submodule of OX . It then
follows from (c′ ) that the corresponding homomorphism Γ( X, OXn ) → Γ( X, OX ) is surjective, which II | 98
proves (b).
Finally, (a) implies (d) (I, 5.1.9.2), and (d) trivially implies (d′ ). It remains to show that (d′ )
implies (c′ ). But if F ′ is a quasi-coherent OX -submodule of OXn , then the filtration 0 ⊂ OX ⊂
OX2 ⊂ . . . ⊂ OXn defines a filtration of F ′ given by the Fk′ = F ∩ OXk (0 ⩽ k ⩽ n), which are quasi-
coherent OX -modules (I, 4.1.1), and Fk′+1 /Fk′ is isomorphic to a quasi-coherent OX -submodule
of OXk+1 /OXk = OX , which is to say, a quasi-coherent sheaf of ideals of OX . Hypothesis (d′ ) thus
implies that H1 ( X, Fk′+1 /Fk′ ) = 0; the exact cohomology sequence H1 ( X, Fk′ ) → H1 ( X, Fk′+1 ) →
H1 ( X, Fk′+1 /Fk′ ) = 0 then lets us prove by induction on k that H 1 ( X, Fk′ ) = 0 for all k. □
Corollary (5.2.2). — Let f : X → Y be a separated quasi-compact morphism. The following conditions are
equivalent.
(a) The morphism f is an affine morphism.
(b) The functor f ∗ is exact on the category of quasi-coherent OX -modules.
(c) For every quasi-coherent OX -module F , we have R1 f ∗ (F ) = 0.
(c’) for every quasi-coherent sheaf of ideals J of OX , we have R1 f ∗ (J ) = 0.
P ROOF. All these conditions are local on Y, by definition of the functor R1 f ∗ (T, 3.7.3), and so
we can assume that Y is affine. If f is affine, then X is affine, and property (b) is nothing more than
(I, 1.6.4). Conversely, we now show that (b) implies (a): for every quasi-coherent OX -module F ,
we have that f ∗ (F ) is a quasi-coherent OY -module (I, 9.2.2, a). By hypothesis, the functor f ∗ (F )
is exact in F , and the functor Γ(Y, G ) is exact in G (in the category of quasi-coherent OY -modules)
because Y is affine (I, 1.3.11); so Γ(Y, f ∗ (F )) = Γ( X, F ) is exact in F , which proves our claim, by
(5.2.1, c).
If f is affine, then f −1 (U ) is affine for every affine open subset U of Y (1.3.2), and so H1 ( f −1 (U ), F ) =
0 (5.2.1, d), which, by definition, implies that R1 f ∗ (F ) = 0. Finally, suppose that condition (c′ ) is
satisfied; the exact sequence of terms of low degree in the Leray spectral sequence (G, II, 4.17.1 and
I, 4.5.1) give, in particular, the exact sequence
Since Y is affine, and f ∗ (J ) quasi-coherent (I, 9.2.2, a), we have that H1 (Y, f ∗ (J )) = 0 (5.2.1);
hypothesis (c′ ) thus implies that H1 ( X, J ) = 0, and we conclude, by (5.2.1), that X is an affine
scheme. □
Corollary (5.2.3). — If f : X → Y is an affine morphism, then, for every quasi-coherent OX -module F , the
canonical homomorphism H1 (Y, f ∗ (F )) → H1 ( X, F ) is bijective.
II | 99
P ROOF. We have the exact sequence
0 −→ H1 (Y, f ∗ (F )) −→ H1 ( X, F ) −→ H0 (Y, R1 f ∗ (F ))
of terms of low degree in the Leray spectral sequence, and the conclusion follows from (5.2.2). □
Remark (5.2.4). — In Chapter III, §1, we prove that, if X is affine, then we have Hi ( X, F ) = 0 for all
i > 0 and all quasi-coherent OX -modules F .
5.3. Quasi-projective morphisms
Definition (5.3.1). — We say that a morphism f : X → Y is quasi-projective, or that X (considered as
a Y-prescheme via f ) is quasi-projective over Y, or that X is a quasi-projective Y-scheme, if f is of finite
type and there exists an invertible f -ample OX -module.
We note that this notion is not local on Y: the counterexamples of Nagata [Nag58b] and Hironaka
show that, even if X and Y are non-singular algebraic schemes over an algebraically closed field,
every point of Y can have an affine neighbourhood U such that f −1 (U ) is quasi-projective over U,
without f being quasi-projective.
We note that a quasi-projective morphism is necessarily separated (4.6.1). When Y is quasi-
compact, it is equivalent to say either that f is quasi-projective, or that f is of finite type and there
exists a very ample (relative to f ) OX -module ((4.6.2) and (4.6.11)). Further:
Proposition (5.3.2). — Let Y be a quasi-compact scheme or a prescheme whose underlying space is
Noetherian, and let X be a Y-prescheme. The following conditions are equivalent.
(a) X is a quasi-projective Y-scheme.
(b) X is of finite type over Y, and there exists some quasi-coherent OY -module E of finite type such that
X is Y-isomorphic to a subprescheme of P(E ).
(c) X is of finite type over Y, and there exists some quasi-coherent graded OY -algebra S such that S1
is of finite type and generates S , and such that X is Y-isomorphic to a induced subprescheme on
some everywhere-dense open subset of Proj(S ).
P ROOF. This follows immediately from the previous remark and from (4.4.3), (4.4.6), and
(4.4.7). □
We note that, whenever Y is a Noetherian prescheme, we can, in conditions (b) and (c) of (5.3.2),
remove the hypothesis that X is of finite type over Y, since this automatically satisfied (I, 6.3.5).
Corollary (5.3.3). — Let Y be a quasi-compact scheme such that there exists an ample OY -module L (4.5.3).
For a Y-scheme X to be quasi-projective, it is necessary and sufficient for it to be of finite type over Y and also
isomorphic to a Y-subscheme of a projective bundle of the form PYr .
P ROOF. If E is a quasi-coherent OY -module of finite type, then E is isomorphic to a quotient
of an OY -module L ⊗(−n) ⊗OY OYk (4.5.5), and so P(E ) is isomorphic to a closed subscheme of PYk−1
((4.1.2) and (4.1.4)). □
Proposition (5.3.4). —
(i) A quasi-affine morphism of finite type (and, in particular, a quasi-compact immersion, or an affine
morphism of finite type) is quasi-projective.
(ii) If f : X → Y and g : Y → Z are quasi-projective, and if Z is quasi-compact, then g ◦ f is
quasi-projective. II | 100
(iii) If f : X → Y is a quasi-projective S-morphism, then f (S′ ) : X(S′ ) → Y(S′ ) is quasi-projective for
every extension S′ → S of the base prescheme.
5. QUASI-AFFINE MORPHISMS; QUASI-PROJECTIVE MORPHISMS; PROPER MORPHISMS; PROJECTIVE MORPHISMS 272
Proposition (5.4.2). —
(i) A closed immersion is a proper morphism.
(ii) The composition of two proper morphisms is proper.
(iii) If X and Y are S-preschemes, and f : X → Y a proper S-morphism, then f (S′ ) : X(S′ ) → Y(S′ ) is
proper for every extension S′ → S of the base prescheme.
(iv) If f : X → Y and g : X ′ → Y ′ are proper S-morphisms, then f ×S g : X ×S Y → X ′ ×S Y ′ is a
proper S-morphism.
II | 101
P ROOF. It suffices to prove (i), (ii), and (iii) (I, 3.5.1). In each of the three cases, verifying
condition (a) of (5.4.1) follows from previous results ((I, 5.5.1) and (6.4.3)); it remains to verify
condition (b). It is immediate in case (i), because if X → Y is a closed immersion, then so is
X ×Y Y ′ → Y ×Y Y ′ = Y ′ ((I, 4.3.2) and (3.3.3)). To prove (ii), consider two proper morphisms
X → Y and Y → Z, and a morphism Z ′ → Z. We can write X × Z Z ′ = X ×Y (Y × Z Z ′ ) (I, 3.3.9.1),
and so the projection X × Z Z ′ → Z ′ factors as X ×Y (Y × Z Z ′ ) → Y × Z Z ′ → Z ′ . Taking the initial
remark into account, (ii) follows from the fact that the composition of two closed morphisms is
closed. Finally, for every morphism S′ → S, we can identify X(S′ ) with X ×Y Y(S′ ) (I, 3.3.11); for
every morphism Z → Y(S′ ) , we can write
X(S′ ) ×Y(S′ ) Z = ( X ×Y Y(S′ ) ) ×Y(S′ ) Z = X ×Y Z;
since by hypothesis X ×Y Z → Z is closed, this proves (iii). □
Corollary (5.4.3). — Let f : X → Y and g : Y → Z be morphisms such that g ◦ f is proper.
5. QUASI-AFFINE MORPHISMS; QUASI-PROJECTIVE MORPHISMS; PROPER MORPHISMS; PROJECTIVE MORPHISMS 273
p′
p
% ′
Z
(where p and p′ are the projections) commutes (I, 3.2.1); furthermore, f × 1Z′ is surjective because f
is surjective (I, 3.5.2), and p is a closed morphism by hypothesis. Every closed subset F of Y × Z Z ′ is
thus the image under f × 1Z′ of some closed subset E of X × Z Z ′ , so p′ ( F ) = p( E) is closed in Z ′ by
hypothesis, whence the corollary. □
Corollary (5.4.4). — If X is a proper prescheme over Y, and S a quasi-coherent OY -algebra, then every
Y-morphism f : X → Proj(S ) is proper (and a fortiori closed).
P ROOF. The structure morphism p : Proj(S ) → Y is separated, and p ◦ f is proper by hypothe-
sis. □
Corollary (5.4.5). — Let f : X → Y be a separated morphism of finite type. Let ( Xi )1⩽i⩽n (resp. (Yi )1⩽i⩽n )
be a finite family of closed subpreschemes of X (resp. Y), and ji (resp. hi ) the canonical injection Xi → X
(resp. Yi → Y). Suppose that the underlying space of X is the union of the Xi , and that, for all i, there is a
morphism f i : Xi → Yi , such that the diagram
fi
Xi / Yi
ji hi
f
X /Y
commutes. Then, for f to be proper, it is necessary and sufficient for all of the f i to be proper.
II | 102
Corollary (5.4.6). — Let f : X → Y be a separated morphism of finite type; for f to be proper, it is necessary
and sufficient for f red : Xred → Yred to be proper.
P ROOF. This is a particular case of (5.4.5), with n = 1, X1 = Xred , and Y1 = Yred (I, 5.1.5). □
(5.4.7). If X and Y are Noetherian preschemes, and f : X → Y a separated morphism of finite type,
then we can, to show that f is proper, restrict to the the case of dominant morphisms and integral
preschemes. Indeed, let Xi (1 ⩽ i ⩽ n) be the (finitely many) irreducible components of X, and
consider, for each i, the unique closed reduced subprescheme of X that has Xi as its underlying
space, which we again denote by Xi (I, 5.2.1). Let Yi be the unique closed reduced subprescheme
of Y that has f ( Xi ) as its underlying space. If gi (resp. hi ) is the injection morphism Xi → X (resp.
Yi → Y), then we conclude that f ◦ gi = hi ◦ f i , where f i is a dominant morphism Xi → Yi (I, 5.2.2);
we are then under the right conditions to apply (5.4.5), and for f to be proper, it is necessary and
sufficient for all the f i to be proper.
5. QUASI-AFFINE MORPHISMS; QUASI-PROJECTIVE MORPHISMS; PROPER MORPHISMS; PROJECTIVE MORPHISMS 274
Corollary (5.4.8). — Let X and Y be separated S-preschemes of finite type over S, and f : X → Y an
S-morphism. For f to be proper, it is necessary and sufficient that, for every S-prescheme S′ , the morphism
f ×S 1S′ : X ×S S′ → Y ×S S′ be closed.
P ROOF. First note that, if g : X → S and h : Y → S are the structure morphisms, then we have,
by definition, g = h ◦ f , and so f is separated and of finite type ((I, 5.5.1) and (6.3.4)). If f is proper,
then so is f ×S 1S′ (5.4.2); a fortiori, f ×S 1S′ is closed. Conversely, suppose that the conditions of the
statement are satisfied, and let Y ′ be a Y-prescheme; Y ′ can also be considered as an S-prescheme,
and since Y → S is separated, X ×Y Y ′ can be identified with a closed subprescheme of X ×S Y ′
(I, 5.4.2). In the commutative diagram
f ×1Y ′
X ×Y Y ′ / Y ×Y Y ′ = Y ′
f ×1S ′
X ×S Y ′ / Y ×S Y ′ ,
the vertical arrows are closed immersions; it thus immediately follows that if f × 1S′ is a closed
morphism, then so is f × 1Y ′ □
Remark (5.4.9). — We say that a morphism f : X → Y is universally closed if it satisfies condition (b)
of Definition (5.4.1). The reader will observe that, in (5.4.2) to (5.4.8), we can replace every occurrence II | 103
of “proper” with “universally closed” without changing the validity of the results (and in the
hypotheses of (5.4.3), (5.4.5), (5.4.6), and (5.4.8), we can omit the finiteness conditions).
(5.4.10). Let f : X → Y be a morphism of finite type. We say that a closed subset Z of X is proper on Y
(or Y-proper, or proper for f ) if the restriction of f to a closed subprescheme of X, with underlying
space Z (I, 5.2.1), is proper. Since this restriction is then separated, it follows from (5.4.6) and
(I, 5.5.1, vi) that the preceding property does not depend on the closed subprescheme of X that has Z
as its underlying space. If g : X ′ → X is a proper morphism, then g−1 ( Z ) is a proper subset of X ′ : if T
is a subprescheme of X that has Z as its underlying space, it suffices to note that the restriction of
g to the closed subprescheme g−1 ( T ) of X ′ is a proper morphism g−1 ( T ) → T, by (5.4.2, iii), and
to then apply (5.4.2, ii). Further, if X ′′ is a Y-scheme of finite type, and u : X → X ′′ a Y-morphism,
then u( Z ) is a proper subset of X ′′ ; indeed, let us take T to be the closed reduced subprescheme
of X having Z as its underlying space; then the restriction of f to T is proper, and thus so is the
restriction of u to T (5.4.3, i), thus u( Z ) is closed in X ′′ ; let T ′′ be a closed subprescheme of X ′′
v j
having u( Z ) as its underlying space (I, 5.2.1), such that u| T factors as T − → T ′′ −
→ X ′′ , where j is
the canonical injection (I, 5.2.2), and v is thus proper and surjective (5.4.5); if g is the restriction to
T ′′ of the structure morphism X ′′ → Y, then g is separated and of finite type, and we have that
f | T = g ◦ v; it thus follows from (5.4.3, ii) that g is proper, whence our assertion.
It follows, in particular, from these remarks that, if Z is a Y-proper subset of X, then
(1) for every closed subprescheme X ′ of X, Z ∩ X ′ is a Y-proper subset of X ′ ; and
(2) if X is a subprescheme of a Y-scheme of finite type X ′′ , then Z is also a Y-proper subset of
X ′′ (and so, in particular, is closed in X ′′ ).
5. QUASI-AFFINE MORPHISMS; QUASI-PROJECTIVE MORPHISMS; PROPER MORPHISMS; PROJECTIVE MORPHISMS 275
Theorem (5.5.3). —
(i) Every projective morphism is quasi-projective and proper.
(ii) Conversely, let Y be a quasi-compact scheme or a prescheme whose underlying space is Noetherian;
then every morphism f : X → Y that is quasi-projective and proper is projective.
P ROOF.
(i) It is clear that if f : X → Y is projective, then it is of finite type and quasi-projective (thus,
in particular, separated); furthermore, it follows immediately from (5.5.1, b) and (3.5.3)
that if f is projective, then so is f ×Y 1Y ′ : X ×Y Y ′ → Y ′ for every morphism Y ′ → Y. To
show that f is universally closed, it is thus enough to show that a projective morphism f
is closed. Since the question is local on Y, we can suppose that Y = Spec( A), thus (5.5.1)
X = Proj(S), where S is a graded A-algebra generated by a finite number of elements of S1 .
For all y ∈ Y, the fibre f −1 (y) can be identified with Proj(S) ×Y Spec(k(y)) (I, 3.6.1), and
so also with Proj(S ⊗ A k(y)) (2.8.10); so f −1 (y) is empty if and only if S ⊗ A k(y) satisfies
condition (TN) (2.7.4), or, in other words, if Sn ⊗ A k(y) = 0 for sufficiently large n. But since
(Sn )y is an Oy -module of finite type, the preceding condition implies that (Sn )y = 0 for
sufficiently large N, by Nakayama’s lemma. If an is the annihilator in A of the A-module
Sn , then the preceding condition also implies that an ⊂ jn for sufficiently large n (0, 1.7.4).
But since Sn S1 = Sn+1 , by hypothesis, we have that an ⊂ an+1 , and if a is the union of the
an , then we see that f ( X ) = V (a), which proves that f ( X ) is closed in Y. If now X ′ is an
arbitrary closed subset of X, then there exists a closed subprescheme of X that has X ′ as
f
its underlying space (I, 5.2.1), and it is clear (5.5.1, a) that the morphism X ′ → X −
→ Y is
projective, and so f ( X ′ ) is closed in Y.
(ii) The hypothesis on Y and the fact that f is quasi-projective implies the existence of a quasi-
coherent OY -module E of finite type, as well as a Y-immersion j : X → P(E ) (5.3.2). But
since f is proper, j is closed, by (5.4.4), and so f is projective.
□
Remark (5.5.4). —
(i) Let f : X → Y be a morphism such that f is proper, such that there exists a very ample
(relative to f ) OX -module L , and such that the quasi-coherent OY -module E = f ∗ (L ) is
of finite type. Then f is a projective morphism: indeed (4.4.4), there is then a Y-immersion
r : X → P(E ), and, since f is proper, r is a closed immersion (5.4.4). We will see in
Chapter III, §3, that when Y is locally Noetherian, the third condition above (E being of finite
5. QUASI-AFFINE MORPHISMS; QUASI-PROJECTIVE MORPHISMS; PROPER MORPHISMS; PROJECTIVE MORPHISMS 276
type) is a consequence of the first two, and so the first two conditions characterise, in this
case, the projective morphisms, and if Y is quasi-compact, then we can replace the second
condition (the existence of a very ample (relative to f ) OX -module L ) by the hypothesis
that there exists an ample (relative to f ) OX -module (4.6.11).
(ii) Let Y be a quasi-compact scheme such that there exists an ample OY -module. For a Y-
scheme X to be projective, it is necessary and sufficient for it to be Y-isomorphic to a closed
Y-subscheme of a projective bundle of the form PYr . The condition is clearly sufficient.
Conversely, if X is projective over Y, then it is quasi-projective, and so there exists a II | 105
Y-immersion j of X into some PYr (5.3.3) that is closed, by (5.4.4) and (5.5.3).
(iii) The argument of (5.5.3) shows that, for every prescheme Y and every integer r ⩾ 0, the
structure morphism PYr → Y is surjective, because if we set S = SOY (OYr+1 ), then we
evidently have Sy = Sk(y) (k (y)r+1 ) (1.7.3), and so (Sn )y ̸= 0 for any y ∈ Y or any n ⩾ 0.
(iv) It follows from the examples of Nagata [Nag58b] that there exist proper morphisms that
are not quasi-projective.
Proposition (5.5.5). —
(i) A closed immersion is a projective morphism.
(ii) If f : X → Y and g : Y → Z are projective morphisms, and if Z is a quasi-compact scheme or a
prescheme whose underlying space is Noetherian, then g ◦ f is projective.
(iii) If f : X → Y is a projective S-morphism, then f (S′ ) : X(S′ ) → Y(S′ ) is projective for every extension
S′ → S of the base prescheme.
(iv) If f : X → Y and g : X ′ → Y ′ are projective S-morphisms, then so is f ×S g.
(v) If g ◦ f is a projective morphism, and if g is separated, then f is projective.
(vi) If f is projective, then so is f red .
P ROOF. (i) follows immediately from (3.1.7). We have to show (iii) and (iv) separately, because
of the restriction introduced on Z in (ii) (cf. (I, 3.5.1)). To show (iii), we restrict to the case where
S = Y (I, 3.3.11), and the claim then immediately follows from (5.5.1, b) and (3.5.3). To show (iv),
we are immediately led to the case where X = P(E ) and X = P(E ′ ), where E (resp. E ′ ) is a quasi-
coherent OY -module (resp. quasi-coherent OY ′ -module) of finite type. Let p and p′ be the canonical
projections of T = Y ×S Y ′ to Y and Y ′ (respectively); by (4.1.3.1), we have P( p∗ (E )) = P(E ) ×Y T
and P( p′∗ (E ′ )) = P(E ′ ) ×Y ′ T; whence
P( p∗ (E )) × T P( p′∗ (E ′ )) = (P(E ) ×Y T ) × T ( T ×Y ′ P(E ′ ))
= P(E ) ×Y ( T ×Y′ P(E ′ )) = P(E ) ×S P(E ′ )
by replacing T with Y ×S Y ′ , and using (I, 3.3.9.1). But p∗ (E ) and p′∗ (E ′ ) are of finite type over
T (0, 5.2.4), and thus so is p∗ (E ) ⊗OT p′∗ (E ′ ); since P( p∗ (E )) × T P( p′∗ (E ′ )) can be identified with
a closed subprescheme of p∗ (E ) ⊗OT p′∗ (E ′ ) (4.3.3), this proves (iv). To show (v) and (vi), we
can apply (I, 5.5.13), because every closed subprescheme of a projective Y-scheme is a projective
Y-scheme, by (5.5.1, a).
It remains to prove (ii); by the hypothesis on Z, this follows from (5.5.3), (5.3.4, ii), and (5.4.2, ii).
□
Proposition (5.5.6). — If X and X ′ are projective Y-schemes, then X ⊔ X ′ is a projective Y-scheme.
P ROOF. This is an evident consequence of (5.5.2) and (4.3.6). □
Proposition (5.5.7). — Let X be a projective Y-scheme, and L a Y-ample OX -module; then, for every
section f of L over X, X f is affine over Y.
II | 106
P ROOF. Since the question is local on Y, we can assume that Y = Spec( A); furthermore,
X f ⊗n = X f , so by replacing L with some suitable L ⊗n , we can assume that L is very ample relative
to the structure morphism q : X → Y (4.6.11). The canonical homomorphism σ : q∗ (q∗ (L )) → L is
thus surjective, and the corresponding morphism
r = rL ,σ : X −→ P = P(q∗ (L ))
5. QUASI-AFFINE MORPHISMS; QUASI-PROJECTIVE MORPHISMS; PROPER MORPHISMS; PROJECTIVE MORPHISMS 277
is an immersion such that L = r ∗ (OP (1)) (4.4.4); furthermore, since X is proper over Y, the
immersion r is closed (5.4.4). But by definition, f ∈ Γ(Y, q∗ (L )), and σ♭ is the identity of q∗ (L ); it
then follows from Equation (3.7.3.1) that we have X f = r −1 ( D+ ( f )); so X f is a closed subprescheme
of the affine scheme D+ ( f ), and is thus also an affine scheme. □
In the particular case where Y = X, we obtain (taking (4.6.13, i) into account) the following
corollary, whose direct proof is immediate anyway:
Corollary (5.5.8). — Let X be a prescheme, and L an invertible OX -module. For every section f of L over
X, X f is affine over X (and thus also an affine scheme whenever X is an affine scheme).
5.6. Chow’s lemma
Theorem (5.6.1). — (Chow’s lemma). Let S be a prescheme, and X an S-scheme of finite type. Suppose
that the following conditions are satisfied:
(a) S is Noetherian;
(b) S is a quasi-compact scheme, and X has a finite number of irreducible components.
Under these hypotheses,
(i) there exists a quasi-projective S-scheme X ′ , and an S-morphism f : X ′ → X that is bothprojective
and surjective;
(ii) we can take X ′ and f to be such that there exists an open subset U ⊂ X for which U ′ = f −1 (U ) is
dense in X ′ , and for which the restriction of f to U ′ is an isomorphism U ′ ≃ U; and
(iii) if X is reduced (resp. irreducible, integral), then we can assume that X ′ is reduced (resp. irreducible,
integral).
P ROOF. The proof proceeds in multiple steps.
(A) We can first restrict to the case where X is irreducible. Indeed, in hypothesis (a), X is
Noetherian, and so, in the two hypotheses, the irreducible components Xi of X are finite in
number. If the theorem is shown to be true for each of the closed reduced preschemes of
X having the Xi as their underlying spaces, and if Xi′ and f i : Xi′ → Xi are the prescheme
and the morphism corresponding to Xi (respectively), then the prescheme X ′ given by the
sum of the Xi′ , and the morphism f : X ′ → X whose restriction to each Xi′ is ji ◦ f i (where ji
is the canonical injection Xi → X) satisfy the conclusion of the theorem. It is immediate
that X ′ is reduced if all
of the Xi′are; furthermore, we can satisfy (ii) by taking U to be the
union of the sets Ui ∩ ∁ X j . Finally, since the Xi′ are quasi-projective over S, so is X ′
S
j ̸ =i
(5.3.6); similarly, the morphisms Xi′ → X are projective by (5.5.5, i) and (5.5.5, ii), and so f II | 107
is projective (5.5.6), and is clearly surjective, by definition.
(B) Now suppose that X is irreducible. Since the structure morphism r : X → S is of finite type,
there exists a finite cover (Si ) of S by affine open subsets, and for each i there is a finite
cover ( Tij ) of r −1 (Si ) by affine open subsets, and the morphisms Tij → Si are of finite type,
and so quasi-projective (5.3.4, i); since in both hypotheses (a) and (b) the immersion Si → S
is quasi-compact, it is also quasi-projective (5.3.4, i), and so the restriction of r to Tij is a
quasi-projective morphism (5.3.4, ii). Denote the Tij by Uk (1 ⩽ k ⩽ n). There exists, for
each index k, an open immersion ϕk : Uk → Pk , where Pk is projective over S ((5.3.2) and
T
(5.5.2)). Let U = k Uk ; since X is irreducible, and the Uk nonempty, U is nonempty, and
thus dense in X; the restrictions of the ϕk to U define a morphism
ϕ : U −→ P = P1 ×S P2 ×S · · · ×S Pn
such that the diagrams
/P
ϕ
(5.6.1.1) U
jk pk
/ Pk
ϕk
Uk
commute, where jk is the canonical injection U → Uk , and pk the canonical projection
P → Pk . If j is the canonical injection U → X, then the morphism ψ = ( j, ϕ)S : U → X ×S P
5. QUASI-AFFINE MORPHISMS; QUASI-PROJECTIVE MORPHISMS; PROPER MORPHISMS; PROJECTIVE MORPHISMS 278
is an immersion (I, 5.3.14). In hypothesis (a), X ×S P is locally Noetherian ((3.4.1), (I, 6.3.7),
and (I, 6.3.8)); in hypothesis (b), X ×S P is a quasi-compact scheme ((I, 5.5.1) and (I, 6.6.4));
in both cases, the closure X ′ in X ×S P of the subprescheme Z associated to ψ (and so with
underlying space ψ(U )) exists, and ψ factors as
ψ′ h
(5.6.1.2) → X′ −
ψ:U− → X ×S P
where ψ′ is an open immersion and h a closed immersion (I, 9.5.10). Let q1 : X ×S P → X and
q2 : X ×S P → P be the canonical projections; we set
h q1
(5.6.1.3) f : X′ −
→ X ×S P −
→ X,
h q2
(5.6.1.4) g : X′ −
→ X ×S P −
→ P.
We will see that X ′ and f satisfy the conclusion of the theorem.
(C) First we show that f is projective and surjective, and that the restriction of f to U ′ = f −1 (U )
is an isomorphism from U ′ to U. Since the Pk are projective over S, so is P (5.5.5, iv), and so
X ×S P is projective over X (5.5.5, iii), and thus so is X ′ , which is a closed subprescheme
of X ×S P. Furthermore, we have f ◦ ψ′ = q1 ◦ (h ◦ ψ′ ) = q1 ◦ ψ = j, so f ( X ′ ) contains the
open everywhere-dense subset U of X; but f is a closed morphism (5.5.3), so f ( X ′ ) = X.
Now note that q1−1 (U ) = U ×S P is induced on an open subset of X ×S P, and, by definition,
the prescheme U ′ = h−1 (U ×S P) is induced by X ′ on the open subset U ′ ; it is thus the
closure relative to U ×S P of the prescheme Z (I, 9.5.8). But the immersion ψ factors as II | 108
Γϕ j ×1
U −→ U ×S P −−→ X ×S P, and since P is separated over S, the graph morphism Γϕ is a
closed immersion (I, 5.4.3), and so Z is a closed subprescheme of U ×S P, whence U ′ = Z.
Since ψ is an immersion, the restriction of f to U ′ is an isomorphism onto U, and the inverse
of ψ′ ; finally, by the definition of X ′ , U ′ is dense in X ′ .
(D) We now show that g is an immersion, which will imply that X ′ is quasi-projective over S,
because P is projective over S. Set
Vk = ϕk (Uk ) (open subset of Pk )
Wk = p− 1
k (Vk ) (open subset of P)
Uk′ = f −1 (Uk ) (open subset of X ′ )
Uk′′ = g−1 (Wk ) (open subset of X ′ ).
It is clear that the Uk′ form an open cover of X ′ ; we will first see that the Uk′′ also form
an open cover of X ′ , by showing that Uk′ ⊂ Uk′′ . For this, it will suffice to show that the
diagram
g|Uk′
(5.6.1.5) Uk′ /P
f |Uk′ pk
/ Pk
ϕk
Uk
commutes. But the prescheme Uk′ = h−1 (Uk ×S P) is induced by X ′ on the open subset Uk′ ,
and is thus the closure of Z = U ′ ⊂ Uk′ relative to Uk′ (I, 9.5.8). To show the commutativity
of (5.6.1.5), it thus suffices (since Pk is an S-scheme) to show that composing the diagram
with the canonical injection U ′ → Uk′ (or, equivalently, thanks to the isomorphism from U ′
to U, with ψ) gives us a commutative diagram (I, 9.5.6). But, by definition, the diagram
thus obtains is exactly (5.6.1.1), whence our claim.
The Wk thus form an open cover of g( X ′ ); to show that g is an immersion, it suffices
to show that each of the restrictions g|Uk′′ is an immersion into Wk (I, 4.2.4). For this,
pk
k ϕ −1
consider the morphism uk : Wk − → Vk −− → Uk → X; since X is separated over S, the
graph morphism Γuk : Wk → X ×S Wk is a closed immersion (I, 5.4.3), and so the graph
Tk = Γuk (Wk ) is a closed subprescheme of X ×S W; if we show that U ′ → X ×S Wk factors
5. QUASI-AFFINE MORPHISMS; QUASI-PROJECTIVE MORPHISMS; PROPER MORPHISMS; PROJECTIVE MORPHISMS 279
through this subprescheme, then the map from the subprescheme induced by X ′ on the
open subset Xk′′ of X ′ to X ×S Wk will also factor through this graph, by (I, 9.5.8). Since
the restriction of q2 to Tk is an isomorphism onto Wk , the restriction of g to Xk′′ will be
an immersion into Wk , and our claim will be proven. Let vk be the canonical injection
U ′ → X ×S Wk ; we have to show that there exists a morphism wk : U ′ → Wk such that
vk = Γuk ◦ wk . By the definition of the product, it suffices to prove that q1 ◦ vk = uk ◦ q2 ◦ vk
(I, 3.2.1), or, by composing on the right with the isomorphism ψ′ : U → U ′ , that q1 ◦ ψ = II | 109
uk ◦ q2 ◦ ψ. But since q1 ◦ ψ = j and q2 ◦ ψ = ϕ, our claim follows from the commutativity
of (5.6.1.1), taking into account the definition of uk .
(E) It is clear that since U, and thus U ′ , is irreducible, so is the X ′ from the preceding construc-
tion, and the morphism f is thus birational (I, 2.2.9). If in addition X is reduced, then so is
U ′ , and hence X ′ is also reduced (I, 9.5.9). This finishes the proof.
□
Corollary (5.6.2). — Suppose that one of the hypotheses, (a) and (b), of (5.6.1) is satisfied. For X to be
proper over S, it is necessary and sufficient for there to exist a projective scheme X ′ over S, and a surjective
S−morphism f : X ′ → X (which is thus projective, by (5.5.5, v)). Whenever this is the case, we can further
choose f to be such that there exists a dense open subset U of X for which the restriction of f to f −1 (U ) is
an isomorphism f −1 (U ) ≃ U, and for which f −1 (U ) is dense in X ′ . If in addition X is irreducible (resp.
reduced), then we can assume that X ′ is also irreducible (resp. reduced); when X and X ′ are irreducible, f is a
birational morphism.
P ROOF. The condition is sufficient, by (5.5.3) and (5.4.3, ii). It is necessary because, with the
notation of (5.6.1), if X is proper over S, then X ′ is proper over S, because it is projective over
X, and thus proper over X (5.5.3), and our claim follows from (5.4.2, ii); furthermore, since X ′ is
quasi-projective over S, it is projective over S, by (5.5.3). □
Corollary (5.6.3). — Let S be a locally Noetherian prescheme, and X an S-scheme of finite type over S,
with structure morphism f 0 : X → S. For X to be proper over S, it is necessary and sufficient that, for
every morphism of finite type S′ → S, ( f 0 )(S′ ) : X(S′ ) → S′ be a closed morphism. It even suffices for this
condition to be verified only for every S-prescheme of the form S′ = S ⊗Z Z[ T1 , . . . , Tn ] (where the Ti are
indeterminates).
P ROOF. The condition being clearly necessary, we now show that it is sufficient. Since the
question is local on S and S′ (5.4.1), we can suppose that S and S′ are affine and Noetherian. By
Chow’s lemma, there exists a projective S-scheme P, an immersion j : X ′ → P, and a surjective
projective morphism f : X ′ → X, such that the diagram
f
Xo X′
f0 j
So
r
P
commutes. Since P is of finite type over S, the first hypothesis implies that the projection q2 :
X ×S P → P is a closed morphism. But the immersion j is the composition of q2 and the morphism
f × 1 from X ′ ×S P to X ×S P; but f , being projective, is proper (5.5.3), and so f × 1 is closed. We
thus conclude that j is a closed immersion, and thus proper (5.4.2, i). Furthermore, the structure
morphism r : P → S is projective, and thus proper (5.5.3), so f 0 ◦ f = r ◦ j is proper (5.4.2, ii); finally,
since f is surjective, f 0 is proper, by (5.4.3).
To prove the proposition using only the second, weaker hypothesis (where S′ is of the form
S ⊗Z Z[ T1 , . . . , Tn ]), it suffices to show that it implies the first. But, if S′ is affine and of finite type over
S = Spec( A), then we have S′ = Spec( A[c1 , . . . , cn ]) (I, 6.3.3), and S′ is thus isomorphic to a closed II | 110
subprescheme of S′′ = Spec( A[ T1 , . . . , Tn ]) (where the Ti are indeterminates). In the commutative
6. INTEGRAL MORPHISMS AND FINITE MORPHISMS 280
diagram
1X × j
X ×S S′ / X ×S S′′
( f 0 )(S′ ) ( f 0 )(S′′ )
j
S′ / S′′
both j and 1X × j are closed immersions (I, 4.3.1), and ( f 0 )(S′ ) is closed by hypothesis; thus ( f 0 )(S′′ )
is also closed. □
If S is affine of ring A, then we also say “integral (resp. finite) over A” instead of “integral (resp.
finite) over S”.
(6.1.2). It is clear that, if X is integral over S, then it is affine over S. For an affine prescheme X over S
to be integral (resp. finite) over S it is necessary and sufficient that the associated quasi-coherent
OS -algebra A ( X ) be such that there exist a cover (Sα ) of S by affine opens having the property
that, for all α, Γ(Sα , A ( X )) is an integral (resp. integral and of finite type) algebra over Γ(Sα , OS ). A
quasi-coherent OS -algebra with this property is said to be integral (resp. finite) over OS . Giving an
integral (resp. finite) prescheme over S is thus (1.3.1) the same as giving a quasi-coherent OS -algebra
that is integral (resp. finite) over OS . Note that a quasi-coherent OS -algebra B is finite if and only if
it is an OS -module of finite type (I, 1.3.9); it is equivalent to say that B is an integral OS -algebra of finite
type, since an algebra that is integral and of finite type over a ring A is an A-module of finite type.
Proposition (6.1.3). — Let S be a locally Noetherian prescheme. For an affine prescheme X over S to be
finite over S, it is necessary and sufficient that the OS -algebra A ( X ) be coherent.
P ROOF. Taking the preceding remark into account, this reduces to noting that, if S is locally
Noetherian, then the quasi-coherent OS -modules of finite type are exactly the coherent OS -modules
(I, 1.5.1). □
Proposition (6.1.4). — Let X be an integral (resp. finite) prescheme over S, with structure morphism
f : X → S. Then, for every affine open U ⊂ S of ring A, f −1 (U ) is an affine scheme whose ring B is an
integral (resp. finite) algebra over A.
II | 111
Lemma (6.1.4.1). — Let A be a ring, M an A-module, and ( gi )1⩽i⩽m a finite system of elements of A such
that the D ( gi ) (for 1 ⩽ i ⩽ m) cover Spec( A). If, for all i, Mgi is an A gi -module of finite type, then M is an
A-module of finite type.
P ROOF. We can assume that Mgi admits a finite system of generators (mij /gin ) with mij ∈ M,
with n the same for all indices i. We will show that the mij for a system of generators of M. By
hypothesis, for each i, there exist aij ∈ A and some integer p (independent of i) such that, in Mgi ,
p
m/1 = (∑i aij mij )/gi ; this implies that there exists an integer r ⩾ p such that, for all i, we have
gir m ∈ M′ . But, since the D ( gir ) = D ( gi ) cover Spec( A), the ideal of A generated by the gir is equal
to A, or, in other words, there exist elements ai ∈ A such that ∑i ai gir ; then m = (∑i ai gir )m ∈ M′ ,
whence the lemma. □
6. INTEGRAL MORPHISMS AND FINITE MORPHISMS 281
Proposition (6.1.5). —
(i) A closed immersion is finite (and a fortiori integral).
(ii) The composition of two finite (resp. integral) morphisms is finite (resp. integral).
(iii) If f : X → Y is a finite (resp. integral) S-morphism, then f (S′ ) : X(S′ ) → Y(S′ ) is finite (resp.
integral) for any base extension S′ → S.
(iv) If f : X → Y and g : X ′ → Y ′ are finite (resp. integral) S-morphisms, then f ×S g : X ×S Y →
X ′ ×S Y ′ is finite (resp. integral).
(v) If f : X → Y and g : Y → Z are morphisms such that g ◦ f is finite (resp. integral), if g is
separated, then f is finite (resp. integral).
(vi) If f : X → Y is a finite (resp. integral) morphism, then f red is finite (resp. integral).
II | 112
P ROOF. By (I, 5.5.12), it suffices to prove (i), (ii), and (iii). To prove that a closed immersion
X → S is finite, we can restrict to the case where S = Spec( A), and everything then follows from
noting that a quotient ring A/J is a monogeneous A-module. To prove that the composition of two
finite (resp. integral) morphism X → Y, Y → Z is finite (resp. integral), we can again assume that Z
(and thus X and Y (1.3.4)) is affine, and then the claim is equivalent to saying that, if B is a finite
(resp. integral) A-algebra and C a finite (resp. integral) B-algebra, then C is a finite (resp. integral)
A-algebra, which is immediate. Finally, to prove (iii), we can restrict to the case where S = Y, since
X(S′ ) can be identified with X ×Y Y(S′ ) (I, 3.3.11); we can further suppose that S = Spec( A) and
S′ = Spec( A′ ); then X is affine of ring B (1.3.4), and X(S′ ) affine of ring A′ ⊗ A B, and it suffices to note
that, if B is a finite (resp. integral) A-algebra, then A′ ⊗ A B is a finite (resp. integral) A′ -algebra. □
We also note that, if X and Y are S-preschemes that are finite (resp. integral) over S, then their
sum X ⊔ Y is a finite (resp. integral) prescheme over S, since this reduces to showing that, if B and C
are finite (resp. integral) A-algebras over A, then so too is B × C.
Corollary (6.1.6). — If X is an integral (resp. finite) prescheme over S, then, for every open U ⊂ S, f −1 (U )
is integral (resp. finite) over U.
P ROOF. This is a particular case of (6.1.5, (iii)). □
Corollary (6.1.7). — Let f : X → Y be a finite morphism. Then, for all y ∈ Y, the fibre f −1 (y) is a finite
algebraic scheme over k (y), and a fortiori its underlying space is discrete and finite.
P ROOF. Indeed, as a k(y)-prescheme, f −1 (y) can be identified with X ×Y Spec(k(y)) (I, 3.6.1),
which is finite over Spec(k(y)) (6.1.5, (iii)); it is thus an affine scheme whose ring is an algebra of
finite rank over k(y) (6.1.4). The corollary then follows from (I, 6.4.4). □
6. INTEGRAL MORPHISMS AND FINITE MORPHISMS 282
neighbourhood Vj of z j such that the restrictions of the ti to Vj defined an isomorphism OYr |Vj → E |Vj
(0, 5.5.4); the neighbourhood U given by the union of the Vj is the desired neighbourhood. □
Proposition (6.1.13). — Let g : X ′ → X be an integral morphism of preschemes, Y a normal locally integral
prescheme, and f a rational map from Y to X ′ such that g ◦ f is an everywhere defined rational map (I, 7.2.1).
Then f is everywhere defined.
P ROOF. If f 1 and f 2 are morphisms (from dense open subsets of Y to X ′ ) in the class f , then it
is clear that g ◦ f 1 and g ◦ f 2 are equivalent morphisms, which justifies the notation g ◦ f for their
equivalence class. We recall also that, if we further suppose Y to be locally Noetherian, then the
hypothesis that Y is normal already implies that Y is locally integral (I, 6.1.13).
To prove the proposition, note first of all that the question is local on Y, and so we can suppose
that there exists in the class g ◦ f a morphism h : Y → X. Consider the inverse image Y ′ = X(′ h) = X(′ Y ) ,
and note that the morphism g′ = g(Y ) : Y ′ → Y is integral (6.1.5, (iii)). Given the correspondence
between rational maps from Y to X ′ and rational Y-sections of Y ′ (I, 7.1.2), we see that it suffices to
prove the specific case of (6.1.13) where X = Y; in other words, the following: □
Corollary (6.1.14). — Let X be a normal locally integral prescheme, g : X ′ → X an integral morphism,
and f a rational X-section of X ′ . Then f is everywhere defined.
P ROOF. Since the question is local on X, we can assume that X is integral, and then f is
identified with a morphism from an open U of X to X ′ (I, 7.2.2) that is a U-section of g−1 (U ). Since g
is separated, f is a closed immersion from U into g−1 (U ) (I, 5.4.6); let Z be the closed subprescheme
of g−1 (U ) associated to f (I, 4.2.1), which is isomorphic to U, and thus integral; let X1 be the reduced
subprescheme of X ′ whose underlying space is the closure Z of Z in X ′ (I, 5.2.1); then Z is an induced
subprescheme on an open of X1 (I, 5.2.3), and, since it is irreducible, so too is X1 , which is thus
integral. The morphism f can then be considered as a rational X-section of X1 ; since the restriction
of g to X1 is an integral morphism (6.1.5, (i) and (ii)), we can finally reduce to proving (6.1.14) in the
specific case where X ′ = X1 ; in other words, the following: □
Corollary (6.1.15). — Let X be a normal integral prescheme, X ′ an integral prescheme, and g : X ′ → X an
integral morphism. If there exists a rational X-section f of X ′ , then g is an isomorphism.
P ROOF. Since the question is local on X, we can assume that X is affine of integral ring A, and
then X ′ is affine of ring A′ with A′ integral over A (6.1.4) and integral; furthermore, the argument
of (6.1.14) shows that there exists a dense open of X that is isomorphic to a dense open of X ′ , and
so A and A′ have the same field of fractions. Also, by (I, 8.2.1.1), and the hypothesis that the Ox
are integrally closed, the ring A is integrally closed, and so A′ = A, which finishes the proof of
(6.1.13) □
6.2. Quasi-finite morphisms
Proposition (6.2.1). — Let f : X → Y be morphism locally of finite type, and x a point of X. Then the
following conditions are equivalent: II | 115
a) The point x is isolated in its fibre f −1 ( f ( x )).
b) The ring Ox is a quasi-finite O f ( x) -module (0, 7.4.1).
P ROOF. Since the question is clearly local on X and on Y, we can assume that X = Spec( A) and
Y = Spec( B) are affine, with A a B-algebra of finite type (I, 6.3.3). Furthermore, we can replace X
by X ×Y Spec(O f ( x) ) without changing either the fibre f −1 ( f ( x )) or the local ring Ox (I, 3.6.5); we
can thus assume that B is a local ring (equal to O f ( x) ); if n is the maximal ideal of B, then f −1 ( f ( x ))
is an affine scheme of ring A/nA, of finite type over k( f ( x )) = B/n (I, 6.4.11). With this, if (a) is
satisfied then we can further suppose that f −1 ( f ( x )) consists of the single point x; thus A/nA is of
finite rank over B/n (I, 6.4.4), or, in other words, A is a quasi-finite B-module. Conversely, if A is a
quasi-finite B-module, then f −1 ( f ( x )) is an Artinian affine scheme, and thus discrete (I, 6.4.4); so x
is isolated in its fibre, and this shows that (b) implies (a). □
Corollary (6.2.2). — Let f : X → Y be a morphism of finite type. Then the following conditions are
equivalent:
6. INTEGRAL MORPHISMS AND FINITE MORPHISMS 284
a) Every point x ∈ X is isolated in its fibre f −1 ( f ( x )) (or, in other words, the subspace f −1 ( f ( x )) is
discrete).
b) For all x ∈ X, the prescheme f −1 ( f ( x )) is a finite k ( f ( x ))-prescheme.
c) For all x ∈ X, the ring Ox is a quasi-finite O f ( x) -module.
P ROOF. The equivalence of (a) and (c) follows from (6.2.1). Since f −1 ( f ( x )) is an algebraic
k( f ( x ))-prescheme (I, 6.4.11), the equivalence of (a) and (b) follows from (I, 6.4.4). □
Definition (6.2.3). — If f : X → Y is a morphism of finite type satisfying the equivalent conditions
of (6.2.2), we say that f is quasi-finite, or that X is quasi-finite over Y.
It is clear that every finite morphism is quasi-finite (6.1.8).
Proposition (6.2.4). —
(i) An immersion X → Y that is closed, or such that X is Noetherian, is a quasi-finite morphism.
(ii) If f : X → Y and g : Y → Z are quasi-finite morphisms, then g ◦ f is quasi-finite.
(iii) If X and Y are S-preschemes, and f : X → Y a quasi-finite S-morphism, then f (S′ ) : X(S′ ) → Y(S′ )
is quasi-finite for any base extension g : S′ → S.
(iv) If f : X → Y and g : X ′ → Y ′ are quasi-finite S-morphisms, then
f ×S g : X ×S Y −→ X ′ ×S Y ′
is quasi-finite.
(v) Let f : X → Y and g : Y → Z be morphisms such that g ◦ f is quasi-finite; if, further, g is
separated, or X is Noetherian, or X × Z Y is locally Noetherian, then f is quasi-finite.
(vi) If f is quasi-finite, then f red is quasi-finite.
P ROOF. If f : X → Y is an immersion, then every fibre consists of a single point, and claim (i)
then follows from ((I, 6.3.4, (i)) and (I, 6.3.5)). To prove (ii), note first of all that h = g ◦ f is of finite
type (I, 6.3.4, (ii)); furthermore, if z = h( x ) and y = f ( x ), then y is isolated in g−1 (z), and so there
exists an open neighbourhood V of y in Y that does not meet any point of g−1 (z) apart from y; thus
f −1 (V ) is an open neighbourhood of x that does not meet any of the f −1 (y′ ), where y′ ̸= y is in
g−1 (z); since x is isolated in f −1 (y), it is isolated in h−1 (z) = f −1 ( g−1 (z)). To prove (iii), we can II | 116
reduce to the case where Y = S (I, 3.3.11); we again note first of all that f ′ = f (S′ ) is of finite type
−1
(I, 6.3.4, (iii)); also, if x ′ ∈ X ′ = X(S′ ) , and if we set y′ = f ′ ( x ′ ) and y = g(y′ ), then f ′ (y′ ) can be
identified with f −1 (y) ⊗k(y) k(y′ ) (I, 3.6.5); since f −1 (y) is of finite rank over k(y) by hypothesis,
−1
f ′ (y′ ) is of finite rank over k (y′ ), and thus discrete. Claims (iv), (v), and (vi) follows from (i),
(ii), and (iii) by the general method (I, 5.5.12), except for when the hypotheses in (v) are not “g is
separated”; in these cases, we remark first of all that, if x is isolated in f −1 ( g−1 ( g( f ( x )))), then it is
a fortiori isolated in f −1 ( f ( x )); the fact that f is of finite type then follows from (I, 6.3.6). □
Proposition (6.2.5). — Let A be a complete local Noetherian ring, X an A-scheme locally of finite type,
and x a point of X over the closed point y of Y = Spec( A). Suppose that x is isolated in its fibre f −1 (y)
(where f is the structure morphism X → Y). Then Ox is an A-module of finite type, and X is Y-isomorphic
to the sum (I, 3.1) of X ′ = Spec(Ox ) (which is a finite Y-scheme) and an A-scheme X ′′ .
P ROOF. It follows from (6.2.1) that Ox is a quasi-finite A-module. Since Ox is Noetherian
(I, 6.3.7), and the homomorphism A → Ox is local, the hypothesis that A is complete implies that Ox
is an A-module of finite type (0, 7.4.3). Let X ′ = Spec(Ox ) be the local scheme of X at the point x
g f
(I, 2.4.1), and let g : X ′ → X be the canonical morphism. Since the composite X ′ − →X− → Y is finite
′
(6.1.1), and since f is separated, g is finite (6.1.5, (v)), and so g( X ) is closed in X (6.1.10); also, since
g is of finite type, g is a local isomorphism at the closed point x ′ of X ′ by the definition of g and
(I, 6.5.4); but since X ′ is the only open neighbourhood of x ′ , this implies that g is an open immersion,
and so g( X ′ ) is also open in X, which finishes the proof. □
Corollary (6.2.6). — Let A be a complete local Noetherian ring, Y = Spec( A), and f : X → Y a separated
quasi-finite morphism. Then X is Y-isomorphic to a sum X ′ ⊔ X ′′ , where X ′ is a finite Y-scheme, and X ′′ is a
quasi-finite Y-scheme such that, if y is the closed point of Y, then X ′′ ∩ f −1 (y) = ∅.
6. INTEGRAL MORPHISMS AND FINITE MORPHISMS 285
P ROOF. Indeed, the fibre f −1 (y) is finite and discrete by hypothesis, and the corollary then
follows, by induction on the number of points in this fibre, from (6.2.5). □
Remark (6.2.7). — In Chapter V, we will see that, if Y is locally Noetherian, then a separated
quasi-finite morphism X → Y is necessarily quasi-affine.
6.3. Integral closure of a prescheme
Proposition (6.3.1). — Let ( X, A ) be a ringed space, B a (commutative) A -algebra, and f a section of B
over X. Then the following properties are equivalent:
a) The sub-A -module of B generated by the f n for n ⩾ (0, 5.1.1) is of finite type.
b) There exists a sub-A -algebra C of B that is an A -module of finite type and such that f ∈ Γ( X, C ).
c) For all x ∈ X, f x is integral over the fibre Ax .
II | 117
P ROOF. Since the sub-A -module of B generated by the f n is an A -algebra, it is clear that (a)
implies (b). On the other hand, (b) implies that, for all x ∈ X, the Ax -module Cx is of finite type,
which implies that every element of the algebra Cx , and, in particular, f x , is integral over Ax . Finally,
if we have, for some x ∈ X, a relation of the form
f xn + ( a1 ) x f xn−1 + . . . + ( an ) x = 0
where the ai (for 1 ⩽ i ⩽ n) are sections of A over an open neighbourhood U of x, then the section
f n |U + a1 · f n−1 |U + . . . + an is zero over a neighbourhood V ⊂ U of x, from which it immediately
follows that the f k |V (for k ⩾ 0) are linear combinations of the f j |V (for 0 ⩽ j ⩽ n − 1) with
coefficients in Γ(V, A ); we thus conclude that (c) implies (a). □
If the equivalent conditions of (6.3.1) are satisfied then we say that the section f is integral over
A.
Corollary (6.3.2). — Under the hypotheses of (6.3.1), there exists a (unique) sub-A -module A ′ of B such
that, for all x ∈ X, Ax′ is the set of germs f x ∈ Bx that are integral over Ax . For every open U ⊂ X, the
sections of A ′ over U are the sections f ∈ Γ(U, B ) that are integral over A |U.
P ROOF. The existence of A ′ is immediate, by taking Γ(U, A ′ ) to be the set of f ∈ Γ(U, B ) such
that f x is integral over Ax for all x ∈ U. The second claim follows immediately from (6.3.1). □
v = u♯ ◦ g∗ ( j) : j∗ (B ′ ) −→ C ′
sends g∗ (B ′ ) to A ′ . But an element of g∗ (B ′ ) x = Bψ′ ( x) ⊗Bψ(x) Ax is integral over Ax by the
definition of B ′ , and thus so too is its images under v x , which proves the claim.
Proposition (6.3.4). — Let X be a prescheme, and A a quasi-coherent OX -algebra. Then the integral closure
OX′ of OX in A is a quasi-coherent OX -algebra, and, for every affine open U of X, Γ(U, OX′ ) is the integral
closure of Γ(U, OX ) in Γ(U, A ).
6. INTEGRAL MORPHISMS AND FINITE MORPHISMS 286
P ROOF. We can restrict to the case where X = Spec( B) is affine, and A = A, e where A is a
′
B-algebra; let B be the integral closure of B in A. Everything then reduces to proving that, for all II | 118
x ∈ X, an element of A x which is integral over Bx necessarily belongs to Bx′ , which follows from
the fact that, for a commutative ring C, the operations of taking the integral closure in a C-algebra
and passing to a ring of fractions (with respect to a multiplicative subset of C) commute [SZ60, t. I,
pp. 261 and 257]. □
The X-scheme X ′ = Spec(OX′ ) is then called the integral closure of X with respect to A (or with
respect to Spec(A )); it is clear that X ′ is integral over X (6.1.2).
We immediately deduce from (6.1.4) that, if f : X ′ → X is the structure morphism, then, for
every open U of X, f −1 (U ) is the integral closure of the prescheme induced by X on U, with respect to
A |U.
(6.3.5). Let X and Y be preschemes, f : X → Y a morphism, A (resp. B) a quasi-coherent OX -algebra
(resp. quasi-coherent OY -algebra), and let u : B → A be an f -morphism. We have seen (6.3.3) that
we obtain an f -morphism u′ : OY′ → OX′ , where OX′ (resp. OY′ ) is the integral closure of OX (resp.
OY ) in A (resp. B). Then, if X ′ (resp. Y ′ ) is the integral closure of X (resp. Y) with respect to A
(resp. B), we canonically obtain from u a morphism f ′ = Spec(u′ ) : X ′ → Y ′ (1.5.6) such that the
diagram
f′
(6.3.5.1) X′ / Y′
X /Y
f
commutes.
(6.3.6). Suppose that X has only a finite number of irreducible components Xi (for 1 ⩽ i ⩽ r), with
generic points ξ i , and consider, in particular, the integral closure of X with respect to some quasi-
coherent R ( X )-algebra A (quasi-coherent as either an OX -algebra or as an R ( X )-algebra, since the
two are equivalent). We know (I, 7.3.5) that A is a direct sum of r quasi-coherent OX -algebras Ai ,
where the support of Ai is contained inside Xi , and the sheaf induced by Ai on Xi is a simple sheaf
whose fibre Ai is an algebra over Oξ i . It is then clear (6.3.4) that the integral closure OX′ of OX in A
(i )
is the direct sum of the integral closures OX of OX in each of the A, and thus the integral closure
(i )
X ′ = Spec(OX′ ) of X with respect to A is the X-scheme given by the sum of the Spec(OX ) = Xi′ (for
1 ⩽ i ⩽ r).
Suppose further that the OX -algebra A is reduced, or, equivalently, that each of the algebras
Ai are reduced, and can thus be considered as an algebra over the field k (ξ i ) (equal to the field of
rational functions of the reduced subprescheme of X that has Xi as its underlying space); then (1.3.8)
each of the Xi′ is a reduced X-scheme, and X ′ is also the integral closure of Xred . Suppose further
that each of the algebras Ai is a direct sum of a finite number of fields Kij (for 1 ⩽ j ⩽ si ); if Kij is
(i )
the sub-algebra of Ai corresponding to Kij , then it is clear that OX is the direct sum of the integral
(ij) (ij)
closures OX of OX in each of the Kij . Then Xi′ is the sum of the X-schemes Xij′ = Spec(OX ) (for
1 ⩽ j ⩽ si ). Furthermore, under these hypotheses, and with this notation, we have:
II | 119
Proposition (6.3.7). — Each of the Xij′ is an integral normal X-prescheme, and its field of rational functions
R( Xij′ ) is canonically identified with the algebraic closure Kij′ of k (ξ i ) in Kij .
P ROOF. By the above, we can suppose that X is integral, and so r = 1 and s1 = 1, so that the
unique algebra A1 is a field K; let ξ be the generic point of X, and let f : X → X ′ be the structure
morphism. For every non-empty affine open U of X, f −1 (U ) can be identified with the spectrum of
the integral closure BU ′ in the field K of the integral ring B = Γ (Y, O ) (6.3.4); since the ring B′
U X U
is integral and integrally closed, so too are the local rings of the points of its spectrum, and thus
f −1 (U ) is by definition an integral and normal scheme ((0, 4.1.4) and (I, 5.1.4)). Furthermore, since
6. INTEGRAL MORPHISMS AND FINITE MORPHISMS 287
(0) is the only prime ideal of BU ′ over the prime ideal (0) of B [SZ60, t. 1, p. 259], f −1 ( ξ ) consists of
U
a single point ξ , and k(ξ ) is the field of fractions K ′ of BU
′ ′ ′ , which is exactly the algebraic closure of
k(ξ ) in K. Finally, X ′ is irreducible, since, if U runs over the set of non-empty affine opens of X, then
the f −1 (U ) give an open cover of X ′ by irreducible opens; furthermore, the intersection f −1 (U ∩ V )
of any two of these opens contains ξ ′ , and is thus non-empty, and we conclude by (0, 2.1.4). □
Corollary (6.3.8). — Let X be a reduced prescheme that has only a finite number of irreducible components
Xi (where 1 ⩽ i ⩽ r), and let ξ i be the generic point of Xi . Then the integral closure X ′ of X with respect
to R ( X ) is the sum of the r X-schemes Xi′ , which are integral and normal. If f : X ′ → X is the structure
morphism, then f −1 (ξ i ) consists only of the generic point ξ i′ of Xi′ , and we have k (ξ i′ ) = k(ξ i ), or, in other
words, that f is birational.
In this case, we say that X ′ is the normalisation of the reduced prescheme X; we note that f ,
since it is birational and integral, is surjective (6.1.10). Then in order to have X ′ = X, it is necessary
and sufficient that X be normal. If X is an integral prescheme, then it follows from (6.3.8) that its
normalisation X ′ is integral.
(6.3.9). Let X and Y be integral preschemes, f : X → Y a dominant morphism, and L = R( X )
(resp. K = R(Y )) the field of rational functions of X (resp. Y); there is a canonical injection K → L
corresponding to f , and if we identify K (resp. L) with the simple sheaf R (Y ) (resp. R ( X )), this
injection is an f -morphism. Let K1 (resp. L1 ) be an extension of K (resp. L), and suppose we have a
morphism K1 → L1 such that the diagram
K1 / L1
O O
K /L
commutes; if K1 (resp. L1 ) is considered as a simple sheaf on Y (resp. X), and thus as an R (Y )-algebra
(resp. R ( X )-algebra), this implies that K1 → L1 is an f -morphism. With this, if X ′ (resp. Y ′ ) is the
integral closure of X (resp. Y) with respect to L1 (resp. K1 ), then X ′ (resp. Y ′ ) is a normal integral
prescheme (6.3.6) whose field of rational functions is canonically identified with the algebraic closure
L′ (resp. K ′ ) of L (resp. K) in L1 (resp K1 ), and there exists a (necessarily dominant) canonical
morphism f ′ : X ′ → Y ′ that makes the diagram (6.3.5.1) commute. II | 120
The most important case is when we take L1 = L, with K1 then being an extension of K contained
in L, and when we suppose X to be integral and normal, so that X ′ = X. The above then shows that,
since X is normal, if Y ′ is the integral closure of Y with respect to a field K1 ⊂ L = R( X ), then every
dominant morphism f : X → Y factors as
f′
→ Y ′ −→ Y
f :X−
where f ′ is dominant; also, since the monomorphism K1 → L is given, f ′ is necessarily unique,
as we see by reducing to the case where X and Y are affine. We thus see that, given Y, L, and a
K-monomorphism K1 → L, the integral closure Y ′ of Y with respect to K1 is the solution of a universal
problem.
Remark (6.3.10). — Consider the hypothesis of (6.3.6), and suppose further that each of the algebras
Ai is of finite rank over k(ξ i ) (which implies that Ai is a direct sum of a finite number of fields); we
can then show, in certain cases, that the structure morphism X ′ → X is not just integral, but even
finite. Restrict to the case where X is reduced; since the question is local on X, we can further suppose
that X is affine of ring C, and that C has only a finite number of minimal ideals pi (for 1 ⩽ i ⩽ r)
such the that Ci = C/pi are integral; then X ′ is finite over X if the integral closure of each Ci in
finite-dgree extension of its field of fractions is a C-module of finite type (6.3.4). We know that this
condition is always satisfied if C is an algebra of finite type over a field [SZ60, t. I, p. 267, th. 9], or over
Z [Nag58a, I,p. 93, th. 3], or over a complete Noetherian local ring [Nag55, p. 298]. We thus conclude
that X ′ → X will be a finite morphism whenever X is a scheme of finite type over a field, or over Z,
or over a complete Noetherian local ring.
6. INTEGRAL MORPHISMS AND FINITE MORPHISMS 288
Corollary (6.4.4). — Under the hypotheses of (6.4.3), if we further suppose A to be normal, then the σi (u)
(and, in particular, Tr u and det u) belong to A.
Proposition (6.4.5). — Let A be an integral ring, E an A-module of finite type of rank n, and u an
endomorphism of E such that the σi (u) belong to A. For u to be an automorphism of E, it is necessary that
det u be invertible in A; this condition is sufficient if E is torsion free.
P ROOF. The condition is sufficient when E is torsion free, since the hypothesis and equa-
tions (6.4.1.4) and (6.4.1.5) (which hold in E, and not just in E ⊗ A K, since E is torsion free) prove
that (det u)−1 Q(u) is the inverse of u.
The condition is necessary, since, if u is invertible, then it follows from (6.4.3) that det(u−1 )
belongs to the integral closure A′ of A in its field of fractions K, and is evidently the inverse of det u
in A′ . Our claim then follows from: □
Lemma (6.4.5.1). — Let A be a subring of a ring A′ such that A′ is integral over A. If an element x ∈ A is
invertible in A′ , then it is also invertible in A.
P ROOF. In the contrary case, x would belong to a maximal ideal m of A, and it follows from the
first theorem of Cohen–Seidenberg [SZ60, t. I, p. 257, th. 3] that there would be a maximal ideal m′
of A′ such that m = m′ ∩ A; we would then have that x ∈ m′ , which is a contradiction. □
Corollary (6.4.6). — Let A be an integral and integrally closed ring, E a torsion-free A-module of finite
type, and u an endomorphism of E> For u to be an automorphism of E, it is necessary and sufficient that
det u be invertible in A.
P ROOF. This follows from (6.4.4) and (6.4.5). □
Remark (6.4.7). — We will later need a generalisation of the above results. Consider a reduced
Noetherian ring A; let pα (for 1 ⩽ α ⩽ r) be its minimal ideals, Kα the field of fractions of the integral
ring A/pα , and K the total ring of fractions of A, which is the direct sum of the fields Kα . Let E
be an A-module of finite type, and suppose that E ⊗ A K is a free K-module of rank n (which here
is merely a consequence of the other hypotheses); equivalently, we can ask that all the Kα -vector
spaces E ⊗ A Kα = Eα have the same dimension n; if then u is an endomorphism of E, we again set
P(u, T ) = P(u ⊗ 1, T ) and σj (u) = σj (u ⊗ 1), and, in particular, det u = det(u ⊗ 1); the σj (u) are
thus elements of K. It is immediate that E ⊗ A K is the direct sum of the Eα , and that the latter are
stable under u ⊗ 1; the restriction of u ⊗ 1 to Eα is exactly the extension uα of u to this Kα -vector
space; we thus conclude that σj (u) is the element of K who components in the Kα are the σj (uα ).
Since the integral closure of A in K is the direct sum of the integral closures of A in the Kα , the σj (u)
are integral over A.
Lemma (6.4.7.1). — The sub-A-algebra of K generated by all the elements σj (u) (for 1 ⩽ j ⩽ n), where u
runs over Hom A ( E, E), is an A-module of finite type.
P ROOF. It suffices to prove that the sub-A[ T ]-algebra of K [ T ] generated by the P(u, T ) is an
A[ T ]-module of finite type, since if the Fi ( T ) (for 1 ⩽ i ⩽ m) form a system of generators for this
A[ T ]-module, then the coefficients of the Fi ( T ) are integral over A, and thus generate an A-algebra
that is an A-module of finite type [SZ60, t. I, p. 255, th. 1]. We can thus replace A by A[ T ] (which II | 123
is Noetherian), and E by E ⊗ A A[ T ] = E′ , which is such that E′ ⊗ A[T ] K [ T ] = E ⊗ A K [ T ] is a free
K [ T ]-module of rank n. Using the initial notation, we thus see that it suffices to prove that the
A-module generated by the elements det u, where u runs over Hom A ( E, E), is of finite type; a fortiori
(since every submodule of an A-module of finite type is of finite type) it suffices to prove that, as v
runs over the set of endomorphisms of ∧n E, the A-module generated by the det v is of finite type; in
other words, we can again reduce to the case where n = 1. But then the proposition follows from the
fact that Hom A ( E, E) is an A-module of finite type, and that v 7→ det v is a homomorphism from
this A-module into K. □
Let F be the kernel of the canonical homomorphism E → E ⊗ A K, and let E0 = E/F; we
see, as above, that E ⊗ A K can be identified with E0 ⊗ A K, that u( F ) ⊂ F, and that, if u0 is the
endomorphism of E0 induced from u by passing to the quotient, then u ⊗ 1 can be identified with
u0 ⊗ 1, and σj (u) = σj (u0 ) for all j. If we have F = 0, then equations (6.4.1.3) and (6.4.1.5) still
6. INTEGRAL MORPHISMS AND FINITE MORPHISMS 290
make sense and hold true whenever E is identified with a submodule of E ⊗ A K, and the u j with
homomorphisms E → E ⊗ A K; then Proposition (6.4.5) extends also to this case, with the same
proof.
(6.4.8). Let ( X, A ) be a ringed space, E a locally free A -module (of finite rank), and u an endomor-
phism of E . By hypothesis, there exists a base B for the topology of X such that, for all V ∈ B, E |V
is isomorphic to some A n |V (where n may vary with V). Let u be an endomorphism of E ; for all
V ∈ B, uV is thus an endomorphism of the Γ(V, A )-module Γ(V, E ), which is free by hypothesis;
the determinant det uV is thus defined and belongs to Γ(V, A ). Furthermore, if e1 , . . . , en form a
basis of Γ(V, E ), then their restrictions to every open W ⊂ V form a basis of Γ(W, E ) over Γ(W, A ),
and so det uW is the restriction of det uV to W. There thus exists exactly one section of A over X,
which we denote by det u, and call the determinant of u, such that the restriction of det u to each
V ∈ B is det uV . It is clear that, for all x ∈ X, we have (det u) x = det u x ; for endomorphisms u and
v of E , we have
(6.4.8.1) det(u ◦ v) = (det u)(det v)
as well as
(6.4.8.2) det(1E ) = 1A
and, if the rank of E is constant (which will be the case (0, 5.4.1) if X is connected) and equal to n,
(6.4.8.3) det(s · u) = sn det u
for all s ∈ Γ( X, A ) (note that det(0) = 0A if n ⩾ 1, but det(0) = 1A for n = 0). Furthermore, for u
to be an automorphism of E , it is necessary and sufficient that u be invertible in Γ( X, A ).
If the rank of E is constant, then we can similarly define the elementary symmetric functions
σi (u), which are elements of Γ( X, A ); we again have the relations in (6.4.1.3) and (6.4.1.5).
We have thus defined a homomorphism det : HomA (E , E ) → Γ( X, A ) of multiplicative II | 124
monoids; if we note that HomA (E , E ) = Γ( X, H omA (E , E )) by definition, then we see that we can
replace X in this definition by an arbitrary open U of X, which immediately defines a homomorphism
det : H omA (E , E ) → A of sheaves of multiplicative monoids. If E has constant rank, then
we similarly define homomorphisms σi : H omA (E , E ) → A of sheaves of sets; for i = 1, the
homomorphism σ1 = Tr is a homomorphism of A -modules.
Let (Y, B ) be another ringed space, and let f : ( X, A ) → (Y, B ) be a morphism of ringed
spaces; if F is a locally free B-module, then f ∗ (F ) is a locally free A -module (which is of the
same rank as F if the latter is of constant rank) (0, 5.4.5). For every endomorphism v of F , f ∗ (v) is
an endomorphism of f ∗ (F ), and it follows immediately from the definitions that det f ∗ (v) is the
section of A = f ∗ (B ) over X that canonically corresponds to det v ∈ Γ(Y, B ). We can further say
that the homomorphism f ∗ (det) : f ∗ (H omB (F , F )) → f ∗ (B ) = A is the composition
γ♯ det
f ∗ (H omB (F , F )) −→ H omA ( f ∗ (F ), f ∗ (F )) −→ A
(0, 4.4.6). We have analogous results for the σi .
(6.4.9). Now suppose that X is a locally integral prescheme, so that the sheaf R ( X ) of rational functions
on X is a locally simple sheaf of fields (I, 7.4.3), and quasi-coherent as an OX -module. If E is a quasi-
coherent OX -module of finite type, then E ′ = E ⊗OX R ( X ) is a locally free R ( X )-module (I, 7.3.6); for
every endomorphism u of E , u ⊗ 1R (X ) is then an endomorphism of E ′ , and det(u ⊗ 1) is a section
of R ( X ) over X, which we also call the determinant of u, and also denote by det u. It follows from
(6.4.3) that det u is a section of the integral closure of OX in R ( X ) (6.3.2); if, further, X is normal, then
det u is a section of OX over X, and if we further suppose that E is torsion free, then for u to be an
automorphism of E it is necessary and sufficient that det u be invertible, by (6.4.6). Equations (6.4.8.1)
to (6.4.8.3) still hold true; from the homomorphism u 7→ det u, applied to the modules of sections
of H omOX (E , E ), we obtain a homomorphism of sheaves det : H omOX (E , E ) → R ( X ), which takes
values in OX if X is normal. We have analogous definitions and results for the other elementary
symmetric functions σj (u), if E ′ is of constant rank; if, further, X is normal, then the σj (u) are sections
of OX over X.
6. INTEGRAL MORPHISMS AND FINITE MORPHISMS 291
Finally, let X and Y be integral preschemes, and f : X → Y a dominant morphism. We know that
there exists a canonical homomorphism f ∗ (R (Y )) → R ( X ) (I, 7.3.8), whence we obtain, for every
quasi-coherent OY -module F of finite type, a canonical homomorphism θ : f ∗ (F ⊗OY R (Y )) →
f ∗ (F ) ⊗OX R ( X ). If v is an endomorphism of F , then f ∗ (v ⊗ 1R (Y ) ) is an endomorphism of II | 125
f ∗ (F ⊗OY R (Y )), and we have a commutative diagram
f ∗ ( v ⊗1)
f ∗ (F ⊗OY R (Y )) / f ∗ (F ⊗ OY R (Y ))
θ θ
f ∗ (F ) ⊗O X R ( X ) / f ∗ (F ) ⊗O R ( X )
f ∗ (v)⊗1 X
We thus easily conclude that det f ∗ (v) is the canonical image, under the homomorphism
f ∗ (R (Y ))→ R ( X ), of the section det v of R (Y ); indeed, we can immediately reduce to the case
where X = Spec( A) and Y = Spec( B) are affine, with A and B integral rings with fields of fractions
K and L (resp.), with the homomorphism B → A an injection and thus extending to a monomor-
phism L → K; if F = M, e where M is a B-module of finite type, then the rank of M ⊗ B L over L is
equal to that of ( M ⊗ B A) ⊗ A K over K, and det((u ⊗ 1) ⊗ 1) is the image in K of det(u ⊗ 1) for any
endomorphism u of M, whence the conclusion.
(6.4.10). Finally, suppose that X is a locally Noetherian reduced prescheme, so that the sheaf R ( X )
of rational functions on X is again a quasi-coherent OX -module (I, 7.3.4); let E be a coherent OX -
module such that E ′ = E ⊗OX R ( X ) is locally free (of finite rank). By (6.4.7), if E ′ is of constant
rank, then we can, for any endomorphism u of E , define the σj (u), which are sections of R ( X )
over X. If we do not suppose E ′ to be of constant rank, we can still define the homomorphism
det : H omOX (E , E ) → R ( X ).
6.5. Norm of an invertible sheaf
(6.5.1). Let ( X, A ) be a ringed space and B a (commutative) A -algebra. The A -module B is
canonically identified with a sub-A -module of H omA (B, B ), and a section f of B over an open U
of X is identified with multiplication by this section. If ( X, A ) and B satisfy one of the conditions
given in (6.4.8), (6.4.9), or (6.4.10), then we can define det( f ) (and, in certain cases, the σj ( f )) as
sections of A or R ( X ) over U, that we call the norm of f (resp. the elementary symmetric functions of
f ), and denote by NB /A ( f ). We will suppose that one of the following conditions is satisfied:
(I) B is a locally free A -module (of finite rank).
(II) ( X, A ) is a locally Noetherian reduced prescheme, B is a coherent A -module such that B ⊗A
R ( X ) is a locally free R ( X )-module, and, for every section f ∈ Γ(U, B ) over an open U of X, the
norm NB /A ( f ) is a section of A over U.
Note that the latter condition is automatically satisfied whenever the locally Noetherian prescheme II | 126
X is normal (6.4.9).
The hypothesis that B ⊗A R ( X ) is locally free can also be expressed in the following way:
denote by Xα the closed reduced subpreschemes of X whose underlying spaces are the irreducible
components of X (I, 5.2.1), which are thus locally Noetherian integral preschemes. Every x ∈ X
belongs to a finite number of the subspaces Xα ; on the other hand, B ⊗A R ( Xα ) is a locally free
R ( Xα )-module of constant rank k α (I, 7.3.6); to say that B ⊗A R ( X ) is a locally free R ( X )-module
implies that, for all x ∈ X, the ranks k α that correspond to the indices such that x ∈ Xα are equal. This
is a local statement, and we can reduce to the case X = Spec(C ), where C is a reduced Noetherian
ring, and B = D, e where D is a C-algebra that is a C-module of finite type; if pi (for 1 ⩽ i ⩽ m) are
the minimal prime ideals of C, then the total ring of fractions L of C is the direct sum of the fields
of fractions Ki of the integral rings C/pi , and D ⊗C L is the direct sum of the D ⊗C Ki , whence the
conclusion.
It is clear that, under hypothesis (I) or (II), we have thus defined a homomorphism of sheaves of
multiplicative monoids NB /A : B → A , also denoted by N if no confusion may arise, and we call
this homomorphism the norm. For sections f and g of B over the same open U, we thus have
(6.5.1.1) NB /A ( f g) = NB/A ( f )B/A ( g)
6. INTEGRAL MORPHISMS AND FINITE MORPHISMS 292
to suppose either that g∗ (OX ′ ) is a locally free OX -module (of finite type) or that ( X, OX ) and g∗ (OX ′ )
satisfy hypothesis (II). For every invertible OX ′ -module L ′ , we thus set
(6.5.5.1) NX ′ /X (L ′ ) = Ng∗ (O ( g∗ (L ′ ))
X ′ ) /O X
and we call this the norm (with respect to g) of L ′ . Similarly, if h′ : L1′ → L2′ is a homomorphism of
invertible OX ′ -modules, then we set
(6.5.5.2) NX ′ /X (h′ ) = Ng∗ (O ( g∗ (h′ )) : NX ′ /X (L1′ ) −→ NX ′ /X (L2′ ).
X ′ ) /O X
commutes.
Suppose that either g∗ (OX ′ ) is a locally free OX -module or that ( X, OX ) and g∗ (OX ′ ) satisfy (II). Suppose
further that Y is a locally Noetherian reduced prescheme, and that the restriction of f to any irreducible
component of Y is a dominant morphism to an irreducible component of X. Then, for every invertible
OX ′ -module L ′ , we have
∗
NY ′ /Y ( f ′ (L ′ )) = f ∗ ( NX ′ /X (L ′ ))
up to canonical isomorphism.
∗
P ROOF. Note that we have f ∗ ( g∗ (L ′ )) = g∗′ ( f ′ (L ′ )), by (1.5.2), and, in particular, g∗′ (OY ′ ) =
f ∗ (g ′
∗ (OX ′ )); if g∗ (OX ′ ) is locally free, then so too is g∗ (OY ′ ). The conclusion then follows from the
definitions and (6.5.4). □
Remark (6.5.9). — We later generalise the notion of norm developed above, by placing it relation to
the notion of direct image of a divisor.
6. INTEGRAL MORPHISMS AND FINITE MORPHISMS 295
P ROOF. We can suppose Y to be affine (4.6.4), and then, by (4.6.6), the statement is equivalent
to: □
P ROOF. In the second hypothesis, we can further suppose that the canonical homomorphism
g∗ (OX ′ ) → g∗ (OX ′ ) ⊗OX R ( X ) is injective. Indeed, if not, let T be the kernel of this homomorphism, II | 131
which is a coherent ideal of B = g∗ (OX ′ ) (I, 6.1.1), and set X ′′ = Spec(B/T ); we thus have a
commutative diagram
X ′′
h / X′
g′ g
~
X
where h is a closed immersion (1.4.10). Furthermore, we know that the support of T is a closed set
(0, 5.2.2) that is rare in X (I, 7.4.6), whence we conclude that, for the generic point x of an irreducible
component of X, there is an affine open neighbourhood U of x such that B |U = (B/T )|U. Since g
is, by hypothesis, surjective, we thus conclude that x ∈ g′ ( X ′′ ); g′ is thus dominant, and, since it is a
finite morphism, it is surjective (6.1.10); by definition,
g∗′ (OX ′′ ) ⊗OX R ( X ) = (B/T ) ⊗OX R ( X ) = g∗ (OX ′ ) ⊗OX R ( X )
thus ( X, OX ) and g∗′ (OX ′′ ) satisfy (II bis), and furthermore g∗′ (OX ′′ ) → g∗′ (OX ′′ ) ⊗OX R ( X ) is injective.
Finally, h∗ (L ′ ) = L ′′ is an ample OX ′′ -module (4.6.13, (i bis)), and we have that NX ′′ /X (L ′′ ) =
NX ′ /X (L ′ ). Indeed, to define these two invertible OX -modules, we can use the same affine open
cover (Uλ ) of X such that the restrictions of g∗ (L ′ ) and g∗′ (L ′′ ) to Uλ are isomorphic to B |Uλ
and (B/T )|Uλ (respectively). We immediately see that, for every isomorphism ηλ : g∗ (L ′ )|Uλ →
B |Uλ , there is a canonically corresponding isomorphism
ηλ′ : g∗′ (L ′′ )|Uλ −→ (B/T )|Uλ
′ ) are the 1-cocycles corresponding to the systems of isomorphisms ( η )
so that, if (ωλµ ) and (ωλµ λ
and (ηλ′ ) (6.5.2), ωλµ
′ is the canonical image in Γ (U ∩ U , B/T ) of ω
λ µ λµ ∈ Γ ( Uλ ∩ Uµ , B ) . By the
definition of T , we thus conclude that
′
NB /A (ωλµ ) = N(B /T )/A (ωλµ )
(where A = OX ), whence the claimed equality.
So suppose that the homomorphism g∗ (OX ′ ) → g∗ (OX ′ ) ⊗OX R ( X ) is injective when we are in
hypothesis (II bis). It suffices to prove that, as f runs over the sections of L ⊗n (for n > 0) over X, the
X f form a base for the topology of X (4.5.2). But let x ∈ X, and let U be an arbitrary neighbourhood
of x; since g−1 ( x ) is finite (6.1.7) and L ′ is ample, there exists an integer n > 0 and a section f ′ of
⊗n
L ′ over X ′ such that X ′f ′ is a neighbourhood of g−1 ( x ) contained inside g−1 (U ) (4.5.4). Since
⊗n
L ⊗n = NX ′ /X (L ′ )
it suffices to take f = NX ′ /X ( f ′ ); indeed, then X \ X f = g( X ′ \ X ′f ′ ) (6.5.7), and so x ∈ X f ⊂ U. □
Corollary (6.6.3). — Under the hypotheses of (6.6.1), for an invertible OX -module L to be ample for f , it is
necessary and sufficient that L ′ = g∗ (L ) be ample for f ◦ g.
6. INTEGRAL MORPHISMS AND FINITE MORPHISMS 296
P ROOF. The condition is necessary, since g is affine (5.1.12). To prove that the condition is
sufficient, we can suppose that Y is affine (4.6.4), and so X and X ′ are quasi-compact and L ′
is ample (4.6.6), and we need to show that L is ample. But the set of points x ∈ X that admit II | 132
a neighbourhood where g∗ (OX ′ ) (resp. g∗ (OX ′ ) ⊗OX R ( X )) has a given rank n in the first (resp.
second) hypothesis is simultaneously open and closed in X, and so X is the prescheme given by the
sum of a finite number of these opens, and we can thus suppose that it is equal to one of them (4.6.17).
But then NX ′ /X (L ′ ) = L ⊗n , and so L ⊗n is ample by (6.6.2), and thus so too is L (4.5.6). □
Corollary (6.6.4). — Suppose that the hypotheses of (6.6.1) are satisfied, and suppose further that f : X → Y
is of finite type. Then, for f to be quasi-projective, it is necessary and sufficient that f ◦ g be quasi-projective.
If we further suppose that Y is a quasi-compact scheme, or a prescheme whose underlying space is Noetherian,
then for f to be projective, it is necessary and sufficient that f ◦ g be projective.
P ROOF. The hypothesis implies that f ◦ g is of finite type. Taking into account the definition
of quasi-projective morphisms (5.3.1), the first claim follows from (6.6.1) and (6.6.3). Taking into
account this result, along with (5.5.3, (ii)), it remains to show that, if f is quasi-projective, then for f
to be proper, it is necessary and sufficient that f ◦ g be proper. But f is then separated (5.3.1) and of
finite type; since g is surjective, our claim follows from (5.4.2, (ii)) and (5.4.3, (ii)). □
In particular:
Corollary (6.6.5). — Let X be a prescheme of finite type over a field K, and K ′ a finite-degree extension of K.
For X to be projective (resp. quasi-projective) over K, it is necessary and sufficient that X ′ = X ⊗K K ′ be
projective (resp. quasi-projective) over K ′ .
P ROOF. The condition is necessary ((5.3.4, (iii)) and (5.5.5, (iii))). Conversely, suppose that it
is satisfied, and let g : X ′ → X be the canonical projection. It is clear that g is a finite morphism
(6.1.5, (iii)) and surjective (I, 3.5.2, (ii)). Furthermore, g∗ (OX ′ ) is a locally free OX -module, since it is
isomorphic to OX ⊗K K ′ (1.5.2). It then follows from the hypothesis and from (6.1.11) and (5.5.5, (ii))
that X ′ is projective (resp. quasi-projective) over K; we then deduce from (6.6.4) that X is projective
(resp. quasi-projective) over K. □
In Chapter V, we will show that the statement of (6.6.5) remains true when K ′ is an arbitrary
extension of K.
The rest of this section is dedicated to the proof of the criterion in (6.6.11), which is a rather
technical refinement of (6.6.1); it can be omitted on a first reading.
Lemma (6.6.6). — Let X be a reduced Noetherian prescheme, and E a coherent OX -module such that
E ⊗OX R ( X ) is a locally free R ( X )-module of constant rank n. Then there exists a reduced Noetherian
prescheme Z and a birational finite morphism h : Z → X that has the following property: the morphisms
of sheaves of sets σi : H omOX (E , E ) → R ( X ) (for 1 ⩽ i ⩽ n) (cf. (6.4.10)) send H omOX (E , E ) to the
coherent OX -algebra h∗ (OZ ).
P ROOF. Consider an affine open U of X, of ring A(U ) = A; let E = Γ(U, E ), and let CU be the
subalgebra of R(U ) generated by the σi (u) where u runs over Hom A ( E, E); we have seen (6.4.7.1)
that this A-algebra is of finite rank. Furthermore, it is clear that forming the algebras CU commutes
with the restriction operations of an affine open U to an affine open U ′ ⊂ U. We have thus defined
a finite sub-OX -algebra C of R ( X ) such that Γ(U, C ) = CU for every affine open U of X. We take
Z = Spec(C ), and take h to be the structure morphism, which is thus finite (6.1.2); since C is reduced,
Z is a reduced Noetherian prescheme (1.3.8). Finally, the total ring of fractions of CU is R(U ), by
definition, and since CU is contained inside the integral closure of A(U ) in R(U ), there is a bijective
correspondence between minimal prime ideals of A(U ) and minimal prime ideals of CU [SZ60, t. I,
p. 259], which proves that h is birational and finishes the proof. □
Corollary (6.6.7). — Under the hypotheses of (6.6.6), let W be an open of X such that, for all x ∈ W, either
X is normal at the point x, or Ex is a free Ox -module. Then we can suppose h to be defined such that the
restriction of h to h−1 (W ) is an isomorphism from h−1 (W ) to W.
6. INTEGRAL MORPHISMS AND FINITE MORPHISMS 297
P ROOF. Either hypothesis implies that, if U ⊂ W is an affine open, then, with the notation of
(6.6.6), (σi (u)) x ∈ A x for all x ∈ U (6.4.3), and so σi (u) ∈ A, and the conclusion follows from the
definition of h given in (6.6.6). □
(6.6.8). Let X be a reduced Noetherian prescheme, and g : X ′ → X a finite surjective morphism, so
that B = g∗ (OX ′ ) is a coherent OX -algebra; suppose further that B ⊗OX R ( X ) is a locally free
R ( X )-module of constant rank n. We can then apply Lemma (6.6.6), taking E = B, whence,
with the notation of (6.6.6), we obtain a homomorphism of sheaves of multiplicative monoids
σn : H omOX (B, B ) → h∗ (OZ ), and by composing this homomorphism with the canonical homo-
morphism B → H omOX (B, B ) (6.5.1), we thus obtain a homomorphism of sheaves of multiplicative
monoids:
(6.6.8.1) N ′ : B = g∗ (OX ′ ) −→ h∗ (OZ ) = C .
With this in mind, for every invertible OX ′ -module L ′ , g∗ (L ′ ) is an invertible B-module
(6.1.12), and the method of (6.5.2) allows us to functorially associate to L ′ an invertible C -module,
which we denote by N ′ ( g∗ (L ′ )).
Lemma (6.6.9). — Let X be a reduced Noetherian prescheme, and g : X ′ → X a finite surjective morphism
such that g∗ (OX ′ ) ⊗OX R ( X ) is a locally free R ( X )-module of constant rank. Then there exists a reduced
Noetherian prescheme Z and a finite birational morphism h : Z → X that has the following property: for
every ample OX ′ -module L ′ , the invertible OZ -module M such that h∗ (M ) = N ′ ( g∗ (L ′ )) (using the
notation of (6.6.8)) is ample.
P ROOF. Suppose first of all that the homomorphism B → B ⊗OX R ( X ) is injective. Define Z
and h as in (6.6.6) (with E = g∗ (OX ′ )). Let z ∈ Z; we have to show that there exists an integer m > 0
and a section t of M ⊗m over Z such that Zt is an affine open that contains z (4.5.2). Let x = h(z),
and let U be an affine open of X that contains x; then h−1 (U ) is an affine open neighbourhood of
z, and it suffices to find t such that z ∈ Zt ⊂ h−1 (U ), since Zt will then necessarily be affine (5.5.8).
⊗n
There exists, by hypothesis, an integer n > 0 and a section s′ of L ′ over X ′ such that
(6.6.9.1) g−1 ( x ) ⊂ Xs′ ′ ⊂ g−1 (U )
by (4.5.4). By definition, s′ is also a section of g∗ (L ′ ) over X, and it corresponds, as in (6.5.2), to
a section s = N ′ (s′ ) of N ′ ( g∗ (L ′ )) over X. We will show that, if t is the section s considered as a II | 134
section of M over Z, then t is the desired section. Set
(6.6.9.2) V = X \ g( X ′ \ Xs′ ′ )
which is an open of X that contains x and is contained in U, by (6.6.9.1) and (6.1.10). We will show
that
(6.6.9.3) h−1 (V ) ⊂ Zt ⊂ h−1 (U )
which will finish the proof. It is equivalent to say that the set T of y ∈ X such that sy is invertible
contains V and is contained in U. For this, consider first of all an affine open W contained in V;
then g−1 (W ) is an affine open in X ′ , and by (6.6.9.2) s′y′ is invertible for all y′ ∈ g−1 (W ); by the
hypotheses on X and B, we can apply the results of (6.4.7), and we see that, if y = g(y′ ), then sy is
invertible; in other words, V ⊂ T. On the other hand, it also follows from (6.4.7) that, conversely, if
sy is invertible, then so too is s′y′ , which implies that y′ ∈ g−1 (U ) by (6.6.9.1), and so y ∈ U, whence
T ⊂ U in this case.
We pass from this to the general case by the same argument as in (6.6.2), replacing X ′ by X ′′
such that g∗′ (OX ′′ ) → g∗′ (OX ′′ ) ⊗OX R ( X ) is injective, and L ′ by an ample OX ′′ -module L ′′ such
that N ′ ( g∗ (L ′ )) = N ′ ( g∗′ (L ′′ )). Lemma (6.6.9) is then proven in all cases (with a suitable choice of
h). □
Corollary (6.6.10). — Suppose that the hypotheses of (6.6.9) are satisfied; for every invertible OX -module L
such that g∗ (L ) is ample, h∗ (L ) is ample.
P ROOF. If we set L ′ = g∗ (L ), then g∗ (L ′ ) = L ⊗OX B (0, 5.4.10), so
N ′ ( g∗ (L ′ )) = (L ⊗OX C )⊗n
6. INTEGRAL MORPHISMS AND FINITE MORPHISMS 298
(by the same argument as for (6.5.2.4)). We thus conclude that M = (h∗ (L ))⊗n , and since M is
ample, so too is h∗ (L ) (4.5.6). □
Proposition (6.6.11). — Let Y be an affine scheme, X an reduced Noetherian prescheme, f : X → Y a
quasi-compact morphism, and g : X ′ → X a finite surjective morphism. Let W be an open subset of X such
that, for all x ∈ W, either X is normal at the point x or there exists an open neighbourhood T ⊂ W of x such
that ( g∗ (OX ′ ))| T is a locally free (OX | T )-module. Then there exists a reduced Y-prescheme Z and a finite
birational Y-morphism h : Z → X such that the restriction of h to h−1 (W ) is an isomorphism from h−1 (W )
to W and has the following property: for every invertible OX -module L such that g∗ (L ) is ample for f ◦ g,
h∗ (L ) is ample for f ◦ h.
P ROOF. Since Y is affine, g∗ (L ) is ample, and the problem thus reduces to proving that, for a
suitable choice of h, h∗ (L ) is ample (4.6.6). We will show that we can replace g by a finite surjective
∗
morphism g′ : X ′′ → X such that g′ (L ) is ample and g∗′ (OX ′′ ) ⊗OX R ( X ) is a locally free R ( X )-
module of constant rank; we will then have reduced to the conditions of (6.6.10), and the proposition
will thus be proven.
For this, let B = g∗ (OX ′ ); denote by Xi (for 1 ⩽ i ⩽ n) the closed reduced subpreschemes of
X that have the irreducible components of X as their underlying space (I, 5.2.1); they are integral
by hypothesis. Let Xi′ be the closed subprescheme g−1 ( Xi ) of X ′ , and gi : Xi′ → Xi the morphism II | 135
g restricted to Xi′ , which is finite (6.1.5, (iii)) and surjective; let k i be the rank of the OXi -algebra
Bi = ( gi )∗ (OX ′ ). Since Bi ⊗OX R ( Xi ) is a constant presheaf (I, 7.3.5), the rank k i is also the rank of
i i
the (OX |U )-algebra B |U for every open U of X that does not meet the only irreducible component
Xi . If T is an open subset of X such that B | T is isomorphic to OXm | T, then it follows from the above
remark that the numbers k i are equal to m for all the indices i such that T ∩ Xi ̸= ∅. So let U be
the open subset of X given by the points that admit a neighbourhood on which B is a locally free
OX -module, and let Uj (for 1 ⩽ j ⩽ s) be its connected components, which are open in X and finitely
many (since U is Noetherian); denote by Vj the closed subprescheme of X ′ given by the closure of
the subprescheme induced on the open g−1 (Uj ) (I, 9.5.11). By the above, for all indices i such that
Xi ∩ Uj ̸= ∅, the ranks k i are all equal to a single integer m j ; note also that a single Xi cannot meet
two Uj with different indices j. Let iλ be the indices i such that Xi ∩ U = ∅. Consider the product k
of all the k i , set ni = k/k i , and let X ′′ be the prescheme defined as follows. For each j (for 1 ⩽ j ⩽ s),
consider k/m j preschemes isomorphic to Vj , and for each λ, k/k iλ preschemes isomorphic to Xi′ ;
λ
then X ′′ is the sum of all these preschemes. We define a morphism g′′ : X ′′ → X ′ that reduces
to the canonical injection on each of the summands of X ′′ ; it is clear that g′′ is a finite dominant
morphism, and thus surjective (since a finite morphism is closed (6.1.10)); set g′ = g ◦ g′′ , which is
∗ ∗ ∗
a finite surjective morphism X ′′ → X; we have g′ (L ) = g′′ ( g∗ (L )), and so g′ (L ) is an ample
OX ′′ -module (5.1.12). It is then clear that, for this new prescheme X ′′ , the ranks defined as the k i for
X ′ are all equal to k; taking (I, 7.3.3) into account, we immediately conclude that, for every affine
open T of X, ( g∗′ (OX ′′ ) ⊗OX R ( X ))| T is an (R ( X )| T )-module isomorphic to (R ( X )| T )k . □
Corollary (6.6.12). — If, in the statement of (6.6.11), we have W = X, then for an invertible OX -module L
to be ample for f , it is necessary and sufficient that g∗ (L ) be ample for f ◦ g.
Remark (6.6.13). — In Chapter III, we will see that, if Y is Noetherian, and f of finite type, and if
the restriction of f to the closed reduced subprescheme of X that has X \ W as its underlying space
is proper, then the conclusion of (6.6.12) still holds true. But we will give, in Chapter V, examples of
algebraic schemes X over a field K (with the structure morphism X → Spec(K ) not proper) whose
normalisation X ′ is quasi-affine, but which are not quasi-affine (in that OX is not ample, even though
OX ′ is (5.1.12) and the morphism X ′ → X is finite and surjective (6.3.10)). We will see in the next
section that this circumstance cannot happen when we replace “quasi-affine” by “affine”.
6. INTEGRAL MORPHISMS AND FINITE MORPHISMS 299
6.7. Chevalley’s theorem We are going to prove (with the help of Serre’s criterion (5.2.1)) the
following theorem, which was proven by C. Chevalley by other methods, in the case of algebraic
schemes.
II | 136
Theorem (6.7.1). — Let X be an affine scheme, Y a Noetherian prescheme, and f : X → Y a finite surjective
morphism. Then Y is an affine scheme.
P ROOF. It is clear that f red : Xred → Yred is finite (6.1.5, (vi)); since Xred is an affine scheme,
and since saying that Y is affine is equivalent to saying that Yred is affine (since Y is Noetherian
(I, 6.1.7)), we see that we can suppose Y to be reduced. For every closed subset Y ′ of Y, there exists
exactly one reduced subprescheme of Y that has Y ′ as its underlying space (I, 5.1.2); its inverse image
f −1 (Y ′ ), which is canonically isomorphic to X ×Y Y ′ (I, 4.4.1), is affine as a closed subprescheme
of X, and the restriction of f to f −1 (Y ′ ), which can be identified with f ×Y 1Y ′ , is a finite surjective
morphism (6.1.5, (iii)). By the principal of Noetherian induction (0, 2.2.2), we can thus (taking
(I, 6.1.7) into account) reduce to proving the theorem under the hypothesis that for every closed
subset Y ′ ̸= Y, every closed subprescheme of Y that has Y ′ as its underlying space is affine. We
thus conclude that, for every coherent OY -module F whose (closed) support Z is distinct from Y, we have
H1 (Y, F ) = 0. Indeed, there exists a closed subprescheme Y ′ of Y that has Z as its underlying space
and is such that, if j : Y → Y is the canonical injection, then F = j∗ ( j∗ (F )) (I, 9.3.5); then (5.2.3)
H1 (Y, F ) = H1 (Y ′ , j∗ (F )) = 0, by (I, 5.1.9.2).
Suppose first of all that Y is not irreducible, and let Y ′ be an irreducible component of Y,
and Y ′′ = Y \ Y ′ ; we again denote by Y ′ the closed reduced subprescheme of Y that has Y ′ as its
underlying space, and by j the canonical injection Y ′ → Y. Let F be a coherent OY -module, and
consider the canonical homomorphism
ρ : F −→ F ′ = j∗ ( j∗ (F ))
(0, 4.4.3); F ′ is a coherent OY -module by (0, 5.3.10) and (0, 5.3.12), since j∗ (OY ′ ) = OY /J , where
we denote by J the sheaf of ideals of OY that defines the subprescheme Y ′ . Then G = Ker ρ and
K = Im ρ are also coherent OY -modules (0, 5.3.4); but, by definition, the fibre Fy′ of F ′ at the
generic point y of Y ′ is equal to Fy , since y is interior to Y ′ and thus Jy = 0, since Y is reduced. We
thus conclude that y is not contained in the (closed) support of G ; also, the support of F ′ (and a
fortiori that of K ) is contained in Y ′ ; in other words, the supports of G and K are distinct from Y.
We thus deduce that H1 (Y, G ) = H1 (Y, K ) = 0, and the exact sequence of cohomology applied to
the exact sequence 0 → G → F → K → 0 implies that H1 (Y, F ) = 0. We thus conclude by Serre’s
criterion (5.2.1).
So suppose that Y is irreducible, and thus integral. We can also suppose that X is integral: if we
denote by Xi the closed reduced subpreschemes of X that have the irreducible components of X
as their underlying space (I, 5.2.1), and by gi the restriction of g to Xi , then at least one of the gi is
dominant, and since it is a finite morphism (6.1.5), it is surjective (6.1.10); since Xi is also an affine
scheme, we see that we can replace X by Xi in the statement.
Lemma (6.7.1.1). — Let X and Y be integral Noetherian preschemes, x (resp. y) the generic point of X (resp.
Y), and f : X → Y a finite surjective morphism. Let L be an invertible OX -module such that there exists an
affine open neighbourhood U of y and a section g ∈ Γ( X, L ) such that x ∈ Xg ⊂ f −1 (U ). Then there exist II | 137
integers m, n > 0, a homomorphism u : OYm → f ∗ (L ⊗n ), and an open neighbourhood V of y such that the
restriction u|V is an isomorphism from OYm |V to f ∗ (L ⊗n )|V.
P ROOF. Let C be the (integral) ring of U, and k = Oy its field of fractions: since f is finite,
U ′ = f −1 (U ) is affine (1.3.2); let D be its (integral) ring with field of fractions K = Ox ; by hypothesis,
D is a C-module of finite type (6.1.4), and so K is an extension of finite rank of k. The fibre
f −1 (y) = X ×Y Spec(k(y)) = X ×Y Spec(Oy ) can be identified with Spec(K ) (I, 3.6.5); let si (for
1 ⩽ i ⩽ m) be elements of D that form a basis of K over k. There exists n > 0 such that the
sections (si | Xg ) g⊗n of L ⊗n over Xg extend to sections bi (for 1 ⩽ i ⩽ m) of L ⊗n over X (I, 9.3.1).
The bi are also, by definition, sections of f ∗ (L ⊗n ) over Y, and thus define a homomorphism
u : OYm → f ∗ (L ⊗n ) (0, 5.1.1); we will show that u is the desired homomorphism. We have that
7. VALUATIVE CRITERIA 300
L ⊗n |U ′ = M,
e where M is a D-module of finite type, so if ϕ is the injection C → D corresponding to
the morphism f −1 (U ) → U given by the restriction of f , then M[ϕ] is a C-module of finite type; then
f ∗ (L ⊗n )|U = ( M[ϕ] )e
(I, 1.6.3) is coherent, and since U is an arbitrary affine open of Y, f ∗ (L ⊗n ) is coherent; furthermore,
u|U = θ, e where θ is a C-homomorphism C m → M[ϕ] , and uy = θy is the homomorphism θ ⊗ 1 :
K = C m ⊗ K → M[ϕ] ⊗ K; but the latter is, by definition, an isomorphism, since the (bi ) x form a
m
With this, the hypotheses of Lemma (6.7.1.1) are satisfied in the case that we are considering,
by taking L = OX , since X is affine (I, 1.1.10); set A = OY and B = f ∗ (OX ). By Serre’s criterion
(5.2.1.1), it suffices to prove that, for every coherent OY -module F , we have that H1 (Y, F ) = 0; it
even suffices to prove this in the case where F ⊂ OY , which implies that F is torsion free, since Y is
integral; in fact, we will show that H1 (Y, OY ) = 0 for every coherent torsion-free OY -module F . But
the homomorphism u : A m → B defines a homomorphism
v : G = H omA (B, F ) −→ H omA (A m , F ) = F m .
We will first show that v is injective: by hypothesis, T = Coker u has a support that does not
meet V, and is thus a torsion OY -module (I, 7.4.6); the exact sequence
A m −→ B −→ T −→ 0
gives, by left exactness of the functor H omA , the exact sequence
v
→ F m.
0 −→ H omA (T , F ) −→ G −
But since F is torsion free, we have that H omA (T , F ) = 0 (0, 5.2.6), whence our claim. We thus
have the exact sequence
0 −→ G −→ F m −→ Coker v −→ 0
where G and Coker v are coherent OY -modules ((0, 5.3.4) and (0, 5.3.5)); by the exact sequence II | 138
of cohomology, it would suffice to show that H1 (Y, G ) = H1 (Y, Coker v) = 0 in order to deduce
that H1 (Y, F m ) = (H1 (Y, F ))m = 0, and thus that H1 (Y, F ) = 0. But the restriction v|V is an
isomorphism, and so the support of Coker v is distinct from Y, whence H1 (Y, Coker v) = 0 by the
hypothesis at the start. Now, G is a coherent B-module (I, 9.6.4); since X is affine over Y, there exists
a quasi-coherent OX -module K such that G is isomorphic to f ∗ (K ) (1.4.3); since X is affine, we
have that H1 ( X, K ) = 0 (I, 5.1.9.2), and so H1 (Y, G ) = 0 by (5.2.3), which finishes the proof of
Theorem (6.7.1). □
Corollary (6.7.2). — Let X be a Noetherian prescheme, ( Xi )1⩽i⩽n a finite cover of the space X consisting of
closed subsets. For X to be affine, it is necessary and sufficient that, for each i, there exist a closed subprescheme
of X that has Xi as its underlying space and is affine.
P ROOF. Indeed, if this is the case, then let X ′ be the scheme given by the sum of the Xi ; it is clear
that X ′ is affine, and we define a surjective morphism f : X ′ → X by taking the restriction of f to Xi
to be the canonical injection. Everything reduces to showing that f is finite, by (6.7.1), and this we
have already seen in (6.1.5). □
Corollary (6.7.3). — For a Noetherian prescheme X to be affine, it is necessary and sufficient that the closed
reduced subpreschemes whose underlying spaces are the irreducible components of X be affine.
concern ourselves with in all that follows, and we introduce arbitrary valuation rings, in the general
case, only to discuss the links with the classical study of such objects.
7.1. Reminder on valuation rings
(7.1.1). Amongst the many diverse equivalent properties that characterise valuation rings, we will
use the following: a ring A is said to be a valuation ring if it is an integral ring which is not a field,
and A is maximal in the set of local rings strictly contained in the field of fractions K of A under the
domination relation (I, 8.1.1). Recall that a valuation ring is integrally closed. If A is a valuation ring,
then so too is Ap for any prime ideal p ̸= 0 of A.
(7.1.2). Let K be a field, and A a local subring of K that is not a field; then there exists a valuation II | 139
ring that both dominates A and has K as its field of fractions ([CC, p. 1-07, lemma 2]).
Now let B be a valuation ring, k its residue field, K its field of fractions, and L an extension of k.
Then there exists a complete valuation ring C that dominates B and whose residue field is L. Indeed,
L is the algebraic extension of a pure transcendental extension L′ = k( Tµ )µ∈ M ; we know that we can
extend the valuation of B corresponding to B to a valuation of K ′ = K ( Tµ )µ∈ M in such a way that
L′ is the residue field of this valuation ([Jaf60, p. 98]); replacing B by the completion of the ring of
this extended valuation, we see that that we can restrict to the case where B is complete and L is an
algebraic closure of k. If K is an algebraic closure of K, we can then extend the valuation that defines
B to K, and the corresponding residue field is an algebraic closure of k, as we can see by lifting to K
the coefficients of a unitary polynomial of k[ T ]. We are thus finally led to the case where L = k, and
it then suffices to take C to be the completion of B in order to satisfy our claim.
(7.1.3). Let K be a field, and A a subring of K; the integral closure A′ of A in K is the intersection
of the valuation rings that contain A and have K as their field of fractions ([Sam53b, p. 51, th. 2]).
Proposition (7.1.2) can then be expressed geometrically in an equivalent form:
Proposition (7.1.4). — Let Y be a prescheme, p : X → Y a morphism, x a point of X, y = p( x ), and y′ ̸= y
a specialisation (0, 2.1.2) of y. Then there exists a local scheme Y ′ which is the spectrum of some valuation
ring, and a separated morphism f : Y ′ → Y such that, denoting the unique closed point of Y ′ by a and the
generic point of Y ′ by b, we have f ( a) = y′ and f (b) = y. We can furthermore suppose that one of the two
additional following properties are satisfied:
(i) Y ′ is the spectrum of a complete valuation ring whose residue field is algebraically closed, and there
exists a k (y)-homomorphism k ( x ) → k(b).
∼
(ii) There exists a k (y)-isomorphism k ( x ) −
→ k ( b ).
P ROOF. Let Y1 be the closed reduced subprescheme of Y that has {y} as its underlying space
(I, 5.2.1), and let X1 be the closed subprescheme given by the inverse image p−1 (Y1 ); since y′ ∈ {y}
by hypothesis, and since k( x ) is the same in X and in X1 , we can assume that Y is integral, with
generic point y; Oy′ is then an integral local ring that is not a field, and whose field of fractions is
Oy = k(y), and k( x ) is then an extension of k(y). To satisfy the conditions f ( a) = y′ and f (b) = y as
well as the additional condition (i) (resp. (ii)), we take Y ′ = Spec( A′ ), where A′ is a valuation ring
that dominates Oy′ (resp. a valuation ring that dominates Oy′ and whose field of fractions is k ( x ));
the existence such an of A′ is guaranteed by (7.1.2). □
(7.1.5). Recall that a local ring A is said to be of dimension 1 if there exists a prime ideal distinct from
the maximal ideal m, and if every prime ideal of A distinct from m is a minimal prime ideal; when
A is integral, it is equivalent to ask that m and (0) be the only prime ideals, with m ̸= (0); in other
words, Y = Spec( A) consists of two points a and b: a is the unique closed point, we have ja = m, and II | 140
k( a) = k is the residue field k = A/m; b is the generic point of Y, jb = (0), with the set {b} being the
unique open subset of Y distinct from both ∅ and Y (an open subset which is thus everywhere dense),
and k(b) = K is the field of fractions of A.
(7.1.6). For a local ring A, Noetherian and of dimension 1, we know ([CC, pp. 2-08 and 17-01]) that
the following conditions are equivalent:
(a) A is normal;
(b) A is regular;
7. VALUATIVE CRITERIA 302
(where π, ϕ, and γ are the homomorphisms corresponding to p, f , and g, respectively) the morphism γ is a
bijection.
P ROOF. As in (7.1.4), we can restrict to the case where Y is integral with generic point y (taking
(I, 6.4.3, iv) into account), and, since the question is local on X and Y, we can assume that p is
of finite type; we are then in the situation of (7.1.4), with the additional property that k ( x ) is an
extension of finite type of k(y) (I, 6.4.11) and that Oy′ is Noetherian; this lets us apply (7.1.7) and take
Y ′ = Spec( A′ ), where A′ is a discrete valuation ring that dominates Oy′ and whose field of fractions
7. VALUATIVE CRITERIA 303
is k( x ). We have thus defined a commutative diagram (7.1.9.1) where γ is a bijection, with π and
ϕ corresponding to the morphisms p and f . Furthermore, since X and Y are locally Noetherian
(I, 6.6.2) and since Y ′ is integral, there exists exactly one rational Y-map g from Y ′ to X to which
corresponds the isomorphism γ (I, 7.1.15), which finishes the proof. □
P ROOF. Condition (b) implies that f ({ x }) = {y}, and is thus a consequence of (a). To show that
(b) implies (a), consider a closed subset X ′ of the underlying space X; let Y ′ = f ( X ′ ), and show that
Y ′ = f ( X ′ ) as follows. Consider the closed reduced subpreschemes of X and Y whose underlying
spaces are X ′ and Y ′ (respectively) (I, 5.2.1); there then exists a morphism f ′ : X ′ → Y ′ such that the
diagram
f′
X′ / Y′
X /Y
f
commutes (I, 5.2.2), and, since f is quasi-compact, so too is f ′ . We are thus led to proving that, if f is
a quasi-compact and dominant morphism, then condition (b) implies that f ( X ) = Y. But let y′ be a II | 142
point of Y, and let y be the generic point of an irreducible component of Y that contains y′ ; by (b), it
suffices to show that f −1 (y) is not empty. But we know that this property is a consequence of the
fact that f is quasi-compact and dominant (I, 6.6.5). □
Corollary (7.2.2). — Let f : X → Y be a quasi-compact immersion. For the underlying space X to be closed
in Y, it is necessary and sufficient for it to contain every specialisation (in Y) of all of its points.
P ROOF. The equivalence of (b) and (c) follows from the bijective correspondence between Y-
morphisms from Y ′ to X and Y ′ -sections of X ′ (I, 3.3.14). If X is separated over Y, condition (b) is
satisfied, by (I, 7.2.2.1), since Y ′ is integral. It remains to show that (b) implies that the diagonal
morphism ∆ : X → X ×Y X is closed, and it is equivalent to show that it satisfies the criteria of
(7.2.2). But let z be a point of the diagonal ∆( X ), and z′ ̸= z a specialisation of z in X ×Y X. There
then exists (7.1.4) a valuation ring A and a morphism f from Y ′ = Spec( A) to X ×Y X such that f
sends the closed point a of Y ′ to z′ , and the generic point b of Y ′ to z; this morphism makes Y ′ an
( X ×Y X )-prescheme, and a fortiori a Y-prescheme. If we compose the two projections of X ×Y X
with f , then we obtain two Y-morphisms, g1 and g2 , from Y ′ to X, which, by hypothesis, agree on
the point b; they are thus equal to one single morphism g, which implies (I, 5.3.1) that f factors as
f = ∆ ◦ g, and thus z′ ∈ ∆( X ). If we suppose that Y is locally Noetherian and f is locally of finite
type, then X ×Y X is locally Noetherian (I, 6.6.7); we can thus follow the same argument as before
by supposing that A is a discrete valuation ring, by (7.1.9). □
7. VALUATIVE CRITERIA 304
Remark (7.2.4). — (i) The hypothesis that the morphism ∆ is quasi-compact is always satis-
fied whenever Y is locally Noetherian and f is locally of finite type, because X ×Y X is then
locally Noetherian (I, 6.6.4, i). In the general case, this also implies that, for every cover
(Uα ) of X by affine opens, the sets Uα ∩ Uβ are quasi-compact.
(ii) For f to be separated, it is sufficient for condition (b) or (c) to be satisfied for some valuation
ring A that is complete and whose residue field is algebraically closed; this follows from the
proofs of (7.2.3) and (7.2.4).
II | 143
7.3. Valuative criterion for properness
Proposition (7.3.1). — Let A be a valuation ring, Y = Spec( A), b the generic point of Y, X an integral
scheme, and f : X → Y a closed morphism such that f −1 (b) consists of a single point x and such that the
corresponding homomorphism k (b) → k( x ) is bijective. Then f is an isomorphism.
P ROOF. Since f if closed and dominant, we have that f ( X ) = Y; it suffices (I, 4.2.2) to prove that,
for all y′ ̸= b in Y, there exists exactly one point x ′ such that f ( x ′ ) = y′ , and that the corresponding
homomorphism Oy′ → Ox′ is bijective, since then f will be a homeomorphism. But if f ( x ′ ) = y′
then Ox′ is a local ring contained in K = k( x ) = k(b) and dominating Oy′ ; the latter is the local ring
Ay′ , and is thus a valuation ring (7.1.1) that has K as its field of fractions. Also, Ox′ ̸= K, since x ′ is
not the generic point of X (0, 2.1.3); we thus conclude that Ox′ = Oy′ . Since X is an integral scheme,
the fact that Ox′ = Ox′′ implies that x ′ = x ′′ (I, 8.2.2), which finishes the proof. □
(7.3.2). Let A be a valuation ring, K its field of fractions, Y = Spec( A), and b the generic point of
Y, such that Ob = k(b) is equal to K; let f : X → Y be a morphism. We know (I, 7.1.4) that the
rational Y-sections of X are in bijective correspondence with the germs of Y-sections (defined in a
neighbourhood of b) at the point b, whence we have a canonical map
(7.3.2.1) Γrat ( X/Y ) −→ Γ( f −1 (b)/ Spec(K ))
with the elements of Γ( f −1 (b)/ Spec(K )) being identified, by definition (I, 3.4.5), with the points of
f −1 (b) = X ⊗ A K that are rational over K. When f is separated, it follows from (I, 5.4.7) that the map
(7.3.2.1) is injective, since Y is an integral scheme.
Composing (7.3.2.1) with the canonical map Γ( X/Y ) → Γrat ( X/Y ) (I, 7.1.2), we obtain a canoni-
cal map
(7.3.2.2) Γ( X/Y ) −→ Γ( f −1 (b)/ Spec(K )).
When f is separated, this map is again injective (I, 5.4.7).
Proposition (7.3.3). — Let A be a valuation ring with field of fractions K, Y = Spec( A), b the generic point
of Y, and f : X → Y a separated and closed morphism. Then the canonical map (7.3.2.2) is bijective (which
is equivalent to saying that it is surjective, and implies that the rational Y-sections of X are everywhere
defined).
P ROOF. So let x be a point of f −1 (b) that is rational over K. Since f is separated, so too is the
morphism f −1 (b) → Spec(K ) corresponding to f (I, 5.5.1, iv), and, since every section of f −1 (b) is
a closed immersion (I, 5.4.6), { x } is closed in f −1 (b). Consider the closed reduced subprescheme
X ′ of X that has the closure { x } of { x } in X as its underlying space. It is clear that the restriction of
f to X ′ satisfies the hypotheses of (7.3.1), and is thus an isomorphism from X ′ to Y, whose inverse
isomorphism is the desired Y-section of X. □
(7.3.4). To state the two following results, we use a terminology that will be justified and discussed in
chapter IV: if F is a subset of a prescheme Y, we define the codimension of F in Y, denoted codimY F, II | 144
to be the lower bound of the integers dim(Oz ) over all z in F.
Corollary (7.3.5). — Let Y be a locally Noetherian reduced prescheme, and N the set of points y ∈ Y where
Y is not regular (0, 4.1.4); suppose that codimY N ⩾ 2. Let f : X → Y be a morphism of finite type, both
separated and closed, and let g be a rational Y-section of X; if Y ′ is the set of points of Y where g is not
defined (a set which is closed (I, 7.2.1)), then codimY Y ′ ⩾ 2.
7. VALUATIVE CRITERIA 305
P ROOF. It suffices to prove that g is defined at every point z ∈ Y such that dim Oz ⩽ 1. If
dim Oz = 0, then z is the generic point of an irreducible component of Y (I, 1.1.14), and so belongs
to every everywhere-dense open subset of Y, and, in particular, to the domain of definition of g.
So suppose that dim Oz = 1; by hypothesis, Oz is then a regular Noetherian local ring, and thus
(7.1.6) a discrete valuation ring. Let Z = Spec(Oz ); since U = Y − Y ′ is everywhere dense, U ∩ Z is
nonempty (I, 2.4.2); let g′ be the rational map from Z to X induced by g (I, 7.2.8); it suffices to show
that g′ is a morphism (I, 7.2.9). But g′ can be considered as a rational Z-section of the Z-prescheme
f −1 ( Z ) = X ×Y Z; it is clear that the morphism f −1 ( Z ) → Z corresponding to f is closed, and it
follows from (I, 5.5.1, i) that it is separated; we thus conclude from (7.3.3) that g′ is everywhere
defined; since Z is reduced, and X is separated over Y, g′ is a morphism (I, 7.2.2). □
Corollary (7.3.6). — Let S be a locally Noetherian prescheme, and X and Y both S-preschemes; suppose that
Y is reduced, and further that the set N of points y ∈ Y where Y is not regular is such that codimY N ⩾ 2;
suppose finally that the structure morphism X → S is proper. Let f be a rational S-map from Y to X, and let
Y ′ be the points of Y where f is not defined; then codimY Y ′ ⩾ 2.
P ROOF. We know (I, 7.1.2) that we can identify the rational S-maps from Y to X with the rational
Y-sections of X ×S Y; since the structure morphism X ×S Y → Y is closed (5.4.1), we can apply
(7.3.5), whence the corollary. □
Remark (7.3.7). — The hypotheses on Y in (7.3.5) and (7.3.6) will be satisfied in particular when Y is
normal (0, 4.1.4), by (7.1.6).
We can characterise the universally closed morphisms (resp. proper morphisms) by a converse
of (7.3.3):
Theorem (7.3.8). — Let Y be a prescheme (resp. a locally Noetherian prescheme), and f : X → Y a
quasi-compact separated morphism (resp. a morphism of finite type). The following conditions are equivalent:
(a) f is universally closed (resp. proper).
(b) For every Y-scheme of the form Y ′ = Spec( A), where A is a valuation ring (resp. a discrete
valuation ring) with field of fractions K, the canonical map
HomY (Y ′ , X ) −→ HomY (Spec(K ), X )
corresponding to the canonical injection A → K is surjective (resp. bijective). II | 145
(c) For every Y-scheme of the form Y ′ = Spec( A), where A is a valuation ring (resp. a discrete
valuation ring), the canonical map (7.3.2.2) with respect to the Y ′ -prescheme X(Y ′ ) is surjective
(resp. bijective).
P ROOF. The equivalence of (b) and (c) follows immediately from (I, 3.3.14); (a) implies (b), since
(a) implies, in either case, that f ( Y ′ ) is separated (I, 5.5.1, iv) and closed, and it suffices to apply
(7.3.3). It remains to prove that (b) implies (a). We first consider the case where Y is arbitrary,
and f is separated and quasi-compact. If condition (b) is satisfied by f , then it is also satisfied by
f (Y ′′ ) : X(Y ′′ ) → Y ′′ , where Y ′′ is an arbitrary Y-prescheme, thanks to the equivalence between (b)
and (c), and the fact that X(Y ′′ ) ×Y ′′ Y ′ = X ×Y Y ′ for every morphism Y ′ → Y ′′ (I, 3.3.9.1); since,
further, f (Y ′′ ) is separated and quasi-compact whenever f is ((I, 5.5.1, iv) and (I, 6.6.4, iii)), we are led
to proving that (b) implies that f is closed. For this, it suffices to verify condition (b) of (7.2.1). So let
x ∈ X, and y′ be a specialisation of y = f ( x ), distinct from y; by (7.1.4), there exists a scheme Y ′ , the
spectrum of some valuation ring, and a separated morphism g : Y ′ → Y such that, letting a denote
the closed point and b the generic point of Y ′ , we have that g( a) = y′ , g(b) = y, and that there
exists a k(y)-homomorphism k( x ) → k(b). The latter corresponds canonically to a Y-morphism
Spec(k(b)) → X (I, 2.4.6), and it thus follows from (b) that there exists a Y-morphism h : Y ′ → X to
which the previous morphism corresponds. We then have that h(b) = x; if we set h( a) = x ′ , then x ′
is a specialisation of x, and we have that f ( x ′ ) = f (h( a)) = g( a) = y′ .
If now Y is locally Noetherian and f of finite type, then hypothesis (b) implies, first of all, that
f is separated, by (7.2.3), with the diagonal morphism X → X ×Y X being quasi-compact (7.2.4).
Further, to show that f is proper, it suffices to show that f (Y ′′ ) : X(Y ′′ ) → Y ′′ is closed for every
Y-prescheme Y ′′ of finite type, taking (5.6.3) into account. Since Y ′′ is then locally Noetherian, we can
7. VALUATIVE CRITERIA 306
follow the same reasoning as in the first case by taking Y ′ to be the spectrum of a discrete valuation
ring, and applying (7.1.9) instead of (7.1.4). □
Remarks (7.3.9). — (i) Whenever Y is an arbitrary prescheme and f a separated morphism,
for f to be universally closed, it suffices that condition (b) or (c) be satisfied for the complete
valuation rings A whose residue field is algebraically closed; this follows from the above
proof and from (7.1.4).
(ii) From criterion (c) of (7.3.8) we obtain a new proof of the fact that a projective morphism
X → Y is closed (5.5.3), and it is closer to the classical approach. We can indeed assume
that Y is affine, and thus that X can be identified with a closed subprescheme of a projective
bundle PYn (5.3.3); to prove that X → Y is closed, it suffices to verify that the structure
morphism PYn → Y is closed, and criteria (c) of (7.3.8), combined with (4.1.3.1), tells us that
we can reduce to proving the following fact: if Y is the spectrum of a valuation ring A, with
field of fractions K, then every point of PYn with values in K comes from (by restriction to the generic
point of Y) a point of PYn with values in A. But every invertible OY -module is trivial (I, 2.4.8);
so it follows from (4.2.6) that a point of PYn with values in K can be identified with a class
of elements (ζc0 , ζc1 , . . . , ζcn ) of K, where ζ ̸= 0 and the c are elements of K that are not
all zero. However, by multiplying the ci by an element of A of suitable valuation, we can II | 146
suppose that the ci all belong to A, and that at least one of them is invertible. But then
(4.2.6) the system (c0 , . . . , cn ) also defines a point of PYn with values in A, which proves our
claim.
(iii) Criteria (7.2.3) and (7.3.8) are particularly simple when we consider the data of a Y-
prescheme X as being equivalent to the data of the functor
X (Y ′ ) = HomY (Y ′ , X )
for Y-preschemes Y ′ ; these criteria allow us, for example, to prove that, under certain
conditions, the “Picard schemes” are proper.
Corollary (7.3.10). — Let Y be an integral scheme (resp. a locally Noetherian integral scheme), X an integral
scheme, and f : X → Y a dominant morphism.
(i) If f is quasi-compact and universally closed, then every valuation ring whose field of fractions is the
field R( X ) of rational functions on X, and which is dominated by a local ring Y, also dominates by
a local ring of X.
(ii) Conversely, suppose that f is of finite type, and that the property described in (i) is verified by every
valuation ring (resp. every discrete valuation ring) that has R( X ) as its field of fractions. Then f is
proper.
P ROOF. Note first of all that the hypotheses imply, in any case, that f is separated (I, 5.5.9).
(i) Let K = R(Y ), L = R( X ), y a point of Y, and A a valuation ring that dominates Oy
and has L as its field of fractions; the injection Oy → A then defines a morphism h from
Y ′ = Spec( A) to Y (I, 2.4.4) such that h( a) = y, where we write a to denote the closed
point of Y ′ ; furthermore, if η is the generic point of Y, which is also the generic point of
Spec(Oy ), then we have h(b) = η, writing b to denote the generic point of Y ′ (since K ⊂ L
by hypothesis). If ξ is the generic point of X, then k (ξ ) = k (b) = L by hypothesis, whence
we have a Y-morphism g : Spec( L) → X such that g(b) = ξ; by (7.3.8), g comes from a
Y-morphism g′ : Y ′ → X. If x = g′ ( a), it is clear that A dominates Ox .
(ii) Since the questions is local on Y, we can always suppose that Y is affine (resp. affine and
Noetherian). Since f is of finite type, we can apply, in either case, Chow’s lemma (5.6.1).
There is thus a projective morphism p : P → Y, an immersion morphism j : X ′ → P, and
a projective morphism g : X ′ → X that is both surjective and birational, with X integral,
such that the diagram
j
Po X′
p g
Yo X
f
7. VALUATIVE CRITERIA 307
7.4. Algebraic curves and function fields of dimension 1 The aim of this section is to show
how to formulate the classical notion of algebraic curves (as introduced, by example, in the book of
C. Chevalley [Che51]) in the language of schemes. All throughout this section, we write k to mean a
field, all the schemes in question are k-schemes of finite type, and all the morphisms are k-morphisms.
Proposition (7.4.1). — Let X be a prescheme of finite type over k (and thus Noetherian); let xi (1 ⩽ i ⩽ n)
be the generic points of the irreducible components Xi of X, and let Ki = k( xi ) (1 ⩽ i ⩽ n). Then the
following conditions are equivalent:
(a) Each of the Ki is an extension of k with transcendence degree equal to 1.
(b) For every closed point x of X, the local ring Ox is of dimension 1 (7.1.5).
(c) The closed irreducible subsets of X that are distinct from the Xi are exactly the closed points of X.
P ROOF. Since X is quasi-compact, every closed irreducible subset F of X contains a closed point
(0, 2.1.3). By (I, 2.4.2), there is a bijective correspondence between the prime ideals of Ox and the
closed irreducible subsets of X that contain x (I, 1.1.14); the equivalence between (b) and (c) follows
immediately from this. Now, if pα (1 ⩽ α ⩽ r) are the minimal prime ideals of the local Noetherian
ring Ox , then the local rings Ox /pα are integral, and have the Ki such that x ∈ Xi as their fields
of fractions. Furthermore, we know ([CC, p. 4-06, th. 2]) that the dimension of a local integral
k-algebra of finite type is equal to the transcendence degree over k of its field of fractions. Finally,
the dimension of Ox is bounded above by the dimensions of the Ox /pα ; but condition (a) implies
that these dimensions are equal to 1, and so (a) implies (b); conversely, if Ox is of dimension 1, then
none of the pα can be equal to the maximal ideal of Ox , otherwise Ox would be of dimension 0; thus
each of the Ox /pα are of dimension 1, which shows that (b) implies (a). □
We note that, under the conditions of (7.4.1), the set X is either empty or infinite, as an immediate
result of (I, 6.4.4).
Definition (7.4.2). — We define an algebraic curve over k to be a non-empty algebraic scheme over k
that satisfies the conditions of (7.4.1).
In the language of dimensions, which will be introduced in Chapter IV, this can be expressed by
saying that an algebraic curve over k is a non-empty algebraic k-scheme whose irreducible components
are all of dimension 1.
We note that, if X is an algebraic curve over k, then the closed reduced subpreschemes Xi
(1 ⩽ i ⩽ n) of X that have the irreducible components of X as their underlying space are also
algebraic curves over k.
Corollary (7.4.3). — Let X be an irreducible algebraic curve. The only non-closed point of X is its generic
point. The closed subsets of X that are distinct from X are the finite sets of closed points; these are also the
only subsets of X that are not everywhere dense.
P ROOF. If a point x ∈ X is not closed, then its closure in X is an irreducible closed subset of X,
and thus necessarily the whole of X, by (7.4.1), and thus x is the generic point of X. A closed subset
F of X that is distinct from X cannot contain the generic point of X, and so all its points are closed II | 149
(in X, and a fortiori in F); by considering the closed reduced subpreschemes of X that have F as their
underlying space (I, 5.2.1), it thus follows from (I, 6.2.2) that F is finite and discrete. The closure in X
of any infinite subset of X is thus necessarily equal to X itself. □
If X is an arbitrary algebraic curve, by applying (7.4.3) to the irreducible components of X, we
see that the only non-closed points of X are the generic points of these components.
Corollary (7.4.4). — Let X and Y be irreducible algebraic curves over k, and f : X → Y a k-morphism. For
f to be dominant, it is necessary and sufficient for f −1 (y) to be finite for all y ∈ Y.
P ROOF. Indeed, if f is not dominant, then f ( X ) is necessarily a finite subset of Y, by (7.4.3), and
so it is not possible for f −1 (y) to be finite for every point of Y, since otherwise X would be finite,
which is a contradiction (7.4.1). Conversely, if f is dominant, then for any y ∈ Y distinct from the
generic point η of Y, we have that f −1 (y) is closed in X, since {y} is closed in Y (7.4.3); also, by
hypothesis, f −1 (y) does not contain the generic point ξ of X, and is thus finite, by (7.4.3). Finally, to
see that, when f is dominant, f −1 (η ) is finite, we note that the fibre f −1 (η ) is an irreducible scheme
7. VALUATIVE CRITERIA 309
of finite type over k(η ), and with generic point ξ ((I, 6.3.9) and (I, 6.4.11)). Since k(ξ ) and k (η ) are
extensions of finite type of k, both of transcendence degree 1, we have that k(ξ ) is necessarily an
extension of finite degree of k(η ), and so ξ is closed in f −1 (η ) (I, 6.4.2), and f −1 (η ) thus consists of a
single point ξ. □
Corollary (7.4.5). — Let X be an algebraic curve over k. For X to be regular, it is necessary and sufficient
for X to be normal, or for the local rings of its closed points to be discrete valuation rings.
Corollary (7.4.6). — Let X be a reduced algebraic curve, and A a reduced coherent R ( X )-algebra; then the
integral closure X ′ of X with respect to A (6.3.4) is a normal algebraic curve, and the canonical morphism
X ′ → X is finite.
P ROOF. The fact that X ′ → X is finite follows from (6.3.10); X ′ is thus an algebraic k-scheme;
furthermore, if xi (1 ⩽ i ⩽ n) are the generic points of the irreducible components of X, and x ′j
(1 ⩽ j ⩽ m) the generic points of the irreducible components of X ′ , then each of the k( x ′j ) is a finite
algebraic extension of one of the k ( xi ) (6.3.6), and thus of transcendence degree 1 over k. So X ′
is indeed an algebraic curve over k, and, furthermore, we know that X ′ is a finite sum of normal
integral schemes ((6.3.6) and (6.3.7)). □
Corollary (7.4.8). — For a reduced algebraic curve X over k to be complete, it is necessary and sufficient for
its normalisation X ′ to be complete.
II | 150
P ROOF. The canonical morphism f : X ′ → X is finite (7.4.6), and thus proper (6.1.11) and
surjective (6.3.8); if g : X → Spec(k) is the structure morphism, then g and ◦ f are both proper, by
(5.4.2, (ii)) and (5.4.3, (ii)), since g is separated by hypothesis. □
Proposition (7.4.9). — Let X be a normal algebraic curve over k, and Y a proper algebraic k-scheme over
k. Then every rational k-map from X to Y is everywhere defined, or, in other words, is a morphism.
P ROOF. It follows from (7.3.7) that, at the points x ∈ X where such a map is not defined, the
dimension of Ox must be ⩾ 2, and so the set of such points is empty; the final claim follows from
(I, 7.2.3). □
P ROOF. Since X is a finite sum of normal integral algebraic curves (6.3.8), we can restrict to the
case where X is integral (5.3.6). Since X is quasi-compact, it is covered by a finite number of affine
open subsets Ui (1 ⩽ i ⩽ n), and, since each of these Ui is of finite type over k, for each i there exists
n
some integer ni along with a k-immersion f i : Ui → Pk i ((5.3.3) and (5.3.4, (i))). Since Ui is dense in
n
X, it follows from (7.4.9) that f i can be extended to a k-morphism gi : X → Pk i , whence we obtain
ni
a k-morphism g = ( g1 , . . . , gn )k from X to the product P of the Pk over k. Furthermore, for each i,
since the restriction of gi to Ui is an immersion, so too is the restriction of g to Ui (I, 5.3.14). Since the
Ui cover X, and since g is separated (I, 5.5.1, (v)), g is an immersion from X into P (I, 8.2.8). Since the
Segre morphism (4.3.3) gives an immersion of P into PkN , this proves that X is quasi-projective. □
Corollary (7.4.11). — Any normal algebraic curve X is isomorphic to the scheme induced by some
complete normal algebraic curve X
b on some everywhere dense open subset, and this X
b is unique up to unique
isomorphism.
7. VALUATIVE CRITERIA 310
P ROOF. If X1 and X2 are complete normal curves, then it follows from (7.4.9) that every isomor-
phism from any dense open U1 in X1 to any dense open U2 in X2 can be uniquely extended to an
isomorphism from X1 to X2 ; whence the uniqueness claim. To prove the existence of X, b it suffices to
n
note that we can consider X as a subscheme of a projective bundle Pk (7.4.10). Let X be the closure of
X in Pnk (I, 9.5.11); since X is induced by X on a dense open subset of X (I, 9.5.10), the generic points
xi of the irreducible components of X are also the generic points of the irreducible components of
X, and the k( xi ) are the same for both of these schemes, and so (7.4.1) X is an algebraic curve over
k that is reduced (I, 9.5.9) and projective over k (5.5.1), whence complete (5.5.3). So we take for X b
the normalisation of X, which is again complete (7.4.8); furthermore, if h : X b → X is the canonical
morphism, then the restriction of h to h−1 ( X ) is an isomorphism to X, since X is normal (6.3.4), and
since h−1 ( X ) contains the generic points of the irreducible components of X b (6.3.8), it is dense in X,
b
which finishes the proof. □
II | 151
Remark (7.4.12). — We will show, in Chapter V, that the conclusion of (7.4.10) still holds true
without the assumption that the curve is normal (or even reduced); we will also show that, for an
algebraic curve (reduced or not) to be affine, it is necessary and sufficient for its (reduced) irreducible
components to not be complete.
Corollary (7.4.13). — Let X be a normal irreducible curve over the field K = R( X ), and Y a complete
integral curve over the field L = R(Y ). Then there is a canonical bijective correspondence between dominant
k-morphisms X → Y and k-monomorphisms L → K.
P ROOF. By (7.4.9), rational k-map from X to Y can be identified with k-morphisms u : X → Y.
Since the dominant morphisms u : X → Y are characterised by being those such that u( x ) = y
(writing x and y to denote the generic points of X and Y, respectively), the corollary follows from
these remarks and from (I, 7.1.13). □
(7.4.14). We can refine the result of (7.4.13) in the case where Y is the projective line P1k = Proj(k[ T0 , T ]),
where T0 and T are indeterminates. Then Y is an integral scheme (2.4.4), and the scheme induced on
the open subset D+ ( T0 ) of Y is isomorphic to Spec(k[ T ]) (2.3.6), and so the generic point of Y is the
ideal (0) of k [ T ], and the field of rational functions of Y is k( T ), which proves that Y is a complete
algebraic curve over k. Furthermore, the only graded prime ideal of S = k[ T0 , T ] that contains T0
and is distinct from S+ is the principal ideal ( T0 ), and so the complement of D+ ( T0 ) in Y = P1k
consists of one closed point, called the “point at infinity”, which we denote by ∞ (for a general study
of the links between vector bundles and projective bundles, see (8.4)). With these notations:
Corollary (7.4.15). — Let X be a normal irreducible curve over the field K = R( X ). Then there exists
a canonical bijective correspondence between the set K and the set of morphisms u from X to P1k that are
distinct from the constant morphism with value ∞. For such a rational map to be dominant, it is necessary
and sufficient for the corresponding element of K to be transcendent over k.
P ROOF. This claim follows immediately from (7.4.9) and the following:
Lemma. — Let X be an integral prescheme over k, and let K = R( X ) be its field of rational functions. Then
there exists a canonical bijective correspondence between the set K and the set of rational maps u from X to
P1k that are distinct from the constant morphism with value ∞. For such a rational map to be dominant, it is
necessary and sufficient for the corresponding element of K to be transcendent over k.
First of all, rational maps from X to P1k correspond bijectively to points of P1k with values in the
extension K of k (I, 7.1.12). If such a point is located (I, 3.4.5) at the generic point of P1k , then the
corresponding rational map is clearly dominant. In the converse case, since every point of P1k that
is distinct from the generic point is closed (7.4.3), the image of the domain of definition U of u by
the unique morphism U → P1k of the class u (I, 7.2.2) consists of one closed point y of P1k , and this
morphism (which is not necessarily everywhere defined on X) is thus not dominant; as an abuse of
language, we thus say that the rational map u is “constant, of value y”. It remains to place in bijective
correspondence the points of P1k with value in K that are located (I, 3.4.5) not at ∞, and the elements
8. BLOWUP SCHEMES; BASED CONES; PROJECTIVE CLOSURE 311
of K, and then to verify that the location of such a point is the generic point of P1k if and only if it
corresponds to an element that is transcendental over k. But this is immediate (4.2.6, example 1). □ II | 152
Corollary (7.4.16). — Let X and Y be algebraic curves over k that are normal, complete, and irreducible;
let K = R( X ) and L = R(Y ) be their fields. Then there exists a canonical bijective correspondence between
∼ ∼
the set of k-isomorphisms X −
→ Y and the set of k-isomorphisms L − → K.
P ROOF. This is an evident consequence of (7.4.13). □
(7.4.17). This corollary (7.4.16) shows that an algebraic curve over k that is normal, complete, and
irreducible, is determined by its field of rational functions K up to unique isomorphism; by definition, K is
an extension of finite type of k, of transcendence degree 1 (we classically call this a field of algebraic
functions of one variable). Furthermore:
Proposition (7.4.18). — For every extension K of k of finite type and of transcendence degree 1, there exists
an algebraic curve X (determined up to unique isomorphism) that is normal, complete, and irreducible, and
such that R( X ) = K. The set of local rings of X can be identified (I, 8.2.1) with the set consisting of the
elements of K and the elements of the valuation rings that contain k and have K as their field of fractions.
P ROOF. Indeed, K is an extension of finite degree of a pure transcendental extension k ( T ) of k,
which can be identified, as we have seen, with the field of rational functions of the projective line
Y = P1k . Let X be the integral closure of Y with respect to K (6.3.4); then X is a normal algebraic curve
over the field K (6.3.7), and it is complete, since the morphism X → Y is finite (7.4.6). The local rings
Ox of X are either the field K, when x is the generic point, or discrete valuation rings that contain k
and have K as their field of fractions, when x is distinct from the generic point (7.4.5). Conversely,
let A be such a ring; since the morphism X → Spec(k ) is proper, the fact that A dominates k implies
that A also dominates a local ring Ox of X (7.3.10); since the latter is a valuation ring that has K as a
field of fractions, it is necessarily equal to A. □
Remarks (7.4.19). — It follows from (7.4.16) and (7.4.18) that the data of an algebraic curve over k that
is normal, complete, and irreducible, is essentially equivalent to the data of an extension K of k that is of finite
type and of transcendence degree 1. We note that, if k′ is an extension of the base field k, then X ⊗k k′
will again be a complete algebraic curve over k′ (5.4.2, (iii)), but, in general, it will be neither reduced
nor irreducible. It will, however, be both reduced and irreducible if K is a separable extension of k,
and k is algebraically closed in K (this can be expressed, in classical terminology, which we will not
use, by saying that K is a “regular extension” of k). But even in this case, it is possible for X ⊗k k′ to
not be normal. The reader will find details on these questions in Chapter IV.
It follows from (8.1.1.1) and (8.1.1.2) that S is a quasi-coherent graded OY -algebra, and thus
defines a Y-scheme X = Proj(S ). If J is an invertible sheaf of ideals of OY , then In ⊗OY J ⊗n is
canonically identified with In J n . If we then replace the In by the In J n , and, in doing so, replace
S by a quasi-coherent OY -algebra S(J ) , then X(J ) = Proj(S(J ) ) is canonically isomorphic to X
(3.1.8).
8. BLOWUP SCHEMES; BASED CONES; PROJECTIVE CLOSURE 312
(8.1.2). Suppose that Y is locally integral, so that the sheaf R (Y ) of rational functions is a quasi-
coherent OY -algebra (I, 7.3.7). We say that an OY -submodule I of R (Y ) is a fractional ideal of R (Y )
if it is of finite type (0, 5.2.1). Suppose we have, for all n ⩾ 0, a quasi-coherent fractional ideal
In of R (Y ), such that I0 = OY , and such that condition (8.1.1.2) (but not necessarily the second
condition (8.1.1.1)) is satisfied; we can then again define a quasi-coherent graded OY -algebra by
Equation (8.1.1.4), and the corresponding Y-scheme X = Proj(S ); we will again have a canonical
isomorphism from X to XJ for every invertible fractional ideal J of R (Y ).
Definition (8.1.3). — Let Y be a prescheme (resp. a locally integral prescheme), and I a quasi-
coherent ideal of OY (resp. a quasi-coherent fractional ideal of R (Y )). We say that the Y-scheme
X = Proj( n⩾0 I n ) is obtained by blowing up the ideal I , or is the blow-up prescheme of Y
L
then X is projective over Y (5.5.2). Without any hypotheses on I , the OX -module OX (1) is invertible
(3.2.5) and very ample, by (4.4.3) applied to the structure morphism X → Y.
We note that, if j : X → Y is the structure morphism, then the restriction of f to f −1 (Y − Y ′ )
is an isomorphism to Y − Y ′ whenever I is an ideal of OY and Y ′ is the closed subprescheme that it
defines: indeed, since the questions is local on Y, it suffices to assume that I = OY , and our claim
then follows from (3.1.7).
If we replace I by I d (d > 0), then the blow-up Y-scheme X is replaced by a canonically
isomorphic Y-scheme X ′ (8.1.1); similarly, for every invertible ideal (resp. invertible fractional ideal)
J , the blow-up prescheme X(J ) relative to the ideal I J is canonically isomorphic to X (8.1.1).
In particular, whenever I is an invertible ideal (resp. invertible fractional ideal), the Y-scheme
obtained by blowing up I is isomorphic to Y (3.1.7).
Proposition (8.1.3). — Let Y be an integral prescheme.
(i) For every sequence (In ) of quasi-coherent fractional ideals of R (Y ) that satisfies (8.1.1.2) and II | 154
such that I0 = OY , the Y-scheme X = Proj( n⩾0 I n ) is integral, and the structure morphism
L
f : X → Y is dominant.
(ii) Let I be a quasi-coherent fractional ideal of R (Y ), and let X be the Y-scheme given by the blow
up of Y relative to I . If I ̸= 0, then the structure morphism f : X → Y is then birational and
surjective.
P ROOF.
(i) This follows from the fact that S = n⩾0 In is an integral OY -algebra ((3.1.12) and (3.1.14)),
L
fractional ideal of R (Y ) that is not zero, Iy ̸= 0 (I, 7.3.6), whence Iy = k (y); then Proj(S ′ )
can be identified with Spec(k(y)) (3.1.7), whence the conclusion.
□
We show a converse of (8.1.4) in (III, 2.3.8).
(8.1.5). We return to the setting and notation of (8.1.1). By definition, the injection homomorphisms
In+1 → In (8.1.1.1) define, for every k ∈ Z, an injective homomorphism of degree zero of graded
S -modules
(8.1.5.1) uk : S+ (k + 1) −→ S (k );
since S+ (k + 1) and S (k + 1) are canonically (TN)-isomorphic, they give a canonical correspon-
dence between uk and an injective homomorphism of OX -modules (3.4.2):
(8.1.5.2) uek : OX (k + 1) −→ OX (k).
8. BLOWUP SCHEMES; BASED CONES; PROJECTIVE CLOSURE 313
S ( h ) ⊗S S ( k ) ⊗S S ( l ) / S ( h + k ) ⊗S S ( l )
S ( h ) ⊗S S ( k + l ) / S (h + k + l )
commutes, it follows from the functoriality of the λ (3.2.6) that the homomorphisms (8.1.5.3) define
the structure of a quasi-coherent graded OX -algebra on
SX = O X ( n ).
M
(8.1.5.4)
n∈Z
Furthermore, the diagram
S ( h ) ⊗S S ( k + 1) / S ( h + k + 1)
1⊗ u k uk+h
S ( h ) ⊗S S ( k ) / S (h + k)
commutes; the functoriality of the λ then implies that we have a commutative diagram
(8.1.5.5) O X ( h ) ⊗O X O X ( k + 1)
λ / O X ( h + k + 1)
1⊗uek uek+h
O X ( h ) ⊗O X O X ( k )
λ / OX ( h + k )
where the horizontal arrows are the canonical homomorphisms. We can thus say that the uek define
an injective homomorphism (of degree zero) of graded SX -modules
(8.1.5.6) ue : SX (1) −→ SX .
(8.1.6). Keeping the notation from (8.1.5), we now note that, for n ⩾ 0, the composite homomorphism
ven = uen−1 ◦ uen−2 ◦ . . . ◦ ue0 is an injective homomorphism OX (n) → OX ; we denote by In,X its image,
which is thus a quasi-coherent ideal of OX , isomorphic to OX (n). Furthermore, the diagram
O X ( m ) ⊗O X O X ( n )
λ / OX (m + n)
the Y-scheme given by blowing up I . We then have, for all n > 0, a canonical isomorphism
∼
(8.1.7.1) → I n OX = In,X
OX (n) −
(cf. (0, 4.3.5)), and thus that I n OX is a very-ample invertible OX -module if n > 0.
P ROOF. The last claim is immediate, since OX (1) is invertible (3.2.5) and very ample for Y by
definition ((4.4.3) and (4.4.9)). Also by definition, the image of vn is exactly I n S , and (8.1.7.1) then
follows from the exactness of the functor M f(3.2.4) and from Equation (3.2.4.1). □
8. BLOWUP SCHEMES; BASED CONES; PROJECTIVE CLOSURE 314
Corollary (8.1.8). — Under the hypotheses of (8.1.7), if f : X → Y is the structure morphism, and Y ′ the
closed subprescheme of Y defined by I , then the closed subprescheme X ′ = f −1 (Y ′ ) of X is defined by I OX
(which is canonically isomorphic to OX (1)), from which we obtain a canonical short exact sequence
(8.1.8.1) 0 −→ OX (1) −→ OX −→ OX ′ −→ 0.
P ROOF. This follows from (8.1.7.1) and from (I, 4.4.5). □
(8.1.9). Under the hypotheses of (8.1.7), we can be more precise about the structure of the In,X . Note
that the homomorphism
ue−1 : OX −→ OX (−1)
canonically corresponds to a section s of OX (−1) over X, which we call the canonical section (relative
to I ) (0, 5.1.1). In the diagram in (8.1.5.5), the horizontal arrows are isomorphisms (3.2.7); by
replacing h with k, and k with −1 in this diagram, we obtain that uek = 1k ⊗ ue−1 (where 1k denotes
the identity on OX (h)), or, equivalently, that the homomorphism uek is given exactly by tensoring with
the canonical section s (for all k ∈ Z). The homomorphism ue (8.1.5.6) can then be understood in the
same way.
Thus, for all n ⩾ 0, the homomorphism ven : OX (n) → OX is given exactly by tensoring with s⊗n ;
we thus deduce:
Corollary (8.1.10). — With the notation of (8.1.8), the underlying space of X ′ is the set of x ∈ X such that
s( x ) = 0, where s denotes the canonical section of OX (−1).
P ROOF. Indeed, if c x is a generator of the fibre (OX (1)) x at a point x, then s x ⊗ c x is canonically
identified with a generator of the fibre of I1,X at the point x, and is thus invertible if and only if
s x ̸∈ mx (Ox (−1)) x , or, equivalently, if and only if s( x ) ̸= 0. □
is an A-isomorphism from In+1 = (S(1))n to aIn+1 = aISn ⊂ In = Sn , and thus defines a (TN)- II | 157
isomorphism of degree zero of graded S-modules S+ (1) → aIS. On the other hand, x 7→ a−1 x is
∼
an isomorphism of degree zero of graded S-modules aIS − → IS. We thus obtain, by composition
∼
(3.2.4), an isomorphism of OX -modules OX (1) − → I OX , and, since S is generated by S1 = I, OX (1)
is invertible (3.2.5) and very ample ((4.4.3) and (4.4.9)), whence our claim. □
which are both graded subrings of S, in only positive and negative degrees (respectively). If f is a
homogeneous elements of degree d (positive or negative) of S, then the ring of fraction S f = S′ is
again endowed with the structure of a graded ring, by taking Sn′ (n ∈ Z) to be the set of the x/ f k for
′⩾ ′
x ∈ Sn+kd (k ⩾ 0); we define S( f ) = S0′ , and will write S⩾ ⩽
f and S f for S and S ⩽ (respectively). If
d > 0, then
(8.2.1.2) ( S⩾ ) f = S f
since, if x ∈ Sn+kd with n + kd < 0, then we can write x/ f k = x f h / f h+k , and we also have that
n + (h + k)d > 0 for h sufficiently large and > 0. We thus conclude, by definition, that
(8.2.1.3) ( S⩾ ) ( f ) = ( S⩾
f )0 = S( f ) .
8. BLOWUP SCHEMES; BASED CONES; PROJECTIVE CLOSURE 315
M⩾ = M⩽ =
M M
(8.2.1.4) Mn , Mn
n⩾0 n⩽0
which are (respectively) a graded S⩾ -module and a graded S⩽ -module, and their intersection is the
S0 module M0 . If f ∈ Sd , then we define M f to be the graded S f -module whose elements of degree n
are the z/ f k for z ∈ Mn+kd (k ⩾ 0); we denote by M( f ) the set of elements of degree zero of M f , and
this is an S( f ) -module, and we will write M⩾ ⩽ ⩾ ⩽
f and M f to mean ( M f ) and ( M f ) (respectively). If
d > 0, then we see, as above, that
(8.2.1.5) ( M⩾ ) f = M f
and
(8.2.1.6) ( M⩾ )( f ) = ( M⩾
f ) 0 = M( f ) .
(8.2.2). Let z be an indeterminate, we we will call the homogenisation variable. If S is a graded ring (in
positive or negative degrees), then the polynomial algebra1
(8.2.2.1) Sb = S[z]
is a graded S-algebra, where we define the degree of f zn (n ⩾ 0), with f homogeneous, as II | 158
P ROOF. The first of the isomorphisms in (8.2.3.1) was defined in (2.2.5), and the second is trivial;
∼
the isomorphism Sb(z) − → S thus defined thus gives a correspondence between xzn /zn+k (where
deg( x ) = k for k ⩾ −n) and the element x. The homomorphism (8.2.3.2) gives a correspondence
between xzn / f k (where deg( x ) = kd − n) and the element x/ f k of degree −n in S⩽ f , and it is again
clear that this does indeed give an isomorphism. □
(8.2.4.1) b = M ⊗S Sb = M ⊗S S[z]
M
is the direct sum of the S-modules M ⊗ Szn , and thus of the abelian groups Mk ⊗ Szn (k ∈ Z, n ⩾ 0);
we define on M b the structure of a graded S-module
b by setting
for all homogeneous x in M. We leave it to the reader to prove the analogue of (8.2.3):
1This should not be confused with the use of the notation Sb to denote the completed separation of a ring.
8. BLOWUP SCHEMES; BASED CONES; PROJECTIVE CLOSURE 316
(8.2.6). Let S be a positively-graded ring, and consider the decreasing sequence of graded ideals of S
M
(8.2.6.1) S[ n ] = Sm (n ⩾ 0)
m⩾n
(so, in particular, we have S[0] = S and S[1] = S+ ). Since it is evident that S[m] S[n] ⊂ S[m+n] , we can
define a graded ring S♮ by setting
M ♮
(8.2.6.2) S♮ = Sn with Sn♮ = S[n] .
n⩾0
S0♮ is then the ring S considered as a non-graded ring, and S♮ is thus an S0♮ -algebra. For every
homogeneous element f ∈ Sd (d > 0), we denote by f ♮ the element f considered as belonging to
S[d] = Sd♮ . With this notation:
II | 159
Lemma (8.2.7). — Let S be a positively-graded ring, and f a homogeneous element of Sd (d > 0). There are
canonical ring isomorphisms
∼ M
(8.2.7.1) Sf −
→ S(n)( f )
n∈Z
∼
(8.2.7.2) ( S⩾
f ) f /1 −
→ Sf
∼
(8.2.7.3) S(♮ f ♮ ) −
→ S⩾
f
and, since S[m] M[n] ⊂ M[m+n] (m ⩾ 0), we can define a graded S♮ -module M♮ by setting
M♮ = Mn♮ = M[n] .
M
(8.2.8.2) with
n∈Z
∼
(8.2.9.2) ( M⩾
f ) f /1 −
→ Mf
∼
(8.2.9.3) M(♮ f ♮ ) −
→ M⩾
f
P ROOF.
(i) If S♮ is an S0♮ -algebra of finite type, then S+ = S1♮ is a module of finite type over S = S0♮ , by
(2.1.6, i), and so S is an S0 -algebra of finite type (2.1.4); if S♮ is a Noetherian ring, then so
too is S0♮ = S (2.1.5). Conversely, if S is an S0 -algebra of finite type, then we know (2.1.6, ii) II | 160
that there exist h > 0 and m0 > 0 such that Sn+h = Sh Sn for n ⩾ m0 ; we can clearly assume
that m0 ⩾ h. Furthermore, the Sm are S0 -modules of finite type (2.1.6, i). So, if n ⩾ m0 + h,
then Sn♮ = Sh Sn♮ −h = Sh♮ Sn♮ −h ; and if m < m0 + h then, letting E = Sm0 + . . . + Sm0 +h−1 , we
have that
♮
Sm = Sm + . . . + Sm0 +h−1 + Sh E + S2h E + . . . .
For 1 ⩽ m ⩽ m0 , let Gm be the union of the finite systems of generators of the S0 -modules Si
for m ⩽ i ⩽ m0 + h − 1, considered as a subset of S[m] . For m0 + 1 ⩽ m ⩽ m0 + h − 1, let Gm
be the union of the finite system of generators of the S0 -modules Si for m ⩽ i ⩽ m0 + h − 1
♮
and of Sh E, considered as a subset of S[m] . It is clear that Sm = S0♮ Gm for 1 ⩽ m ⩽ m0 + h − 1,
and thus the union G of the Gm for 1 ⩽ m ⩽ m0 + h − 1 is a system of generators of the
S0♮ -algebra S♮ . We thus conclude that, if S = S0♮ is a Noetherian ring, then so too is S♮ .
(ii) It is clear that, if Sn+1 = S1 Sn for n ⩾ n0 , then Sn♮ +1 = S1 Sn♮ , and a fortiori Sn♮ +1 = S1♮ Sn♮ for
n ⩾ n0 . Conversely, this last equality can be written as
Sn+1 + Sn+2 + . . . = (S1 + S2 + . . .)(Sn + Sn+1 + . . .)
and comparing terms of degree n + 1 (in S) on both sides gives that Sn+1 = S1 Sn .
(iii) If Sn = S1n for n ⩾ n0 , then Sn♮ = S1n + S1n+1 + . . .; since S1♮ contains S1 + S12 + . . ., we
have that Sn♮ ⊂ S1♮n , and thus Sn♮ = S1♮n for n ⩾ n0 . Conversely, the only terms of S1♮n =
(S1 + S2 + . . .)n that are of degree n in S are those of S1n ; the equality Sn♮ = S1♮n thus implies
that Sn = S1n .
(iv) It suffices to show that, if an element g ∈ Sk+h is considered as an element of Sk♮ (k > 0,
♮
h ⩾ 0), then there exists an integer n > 0 such that gn is a linear combination (in Skn )
of the f α♮ with coefficients in S♮ . By hypothesis, there exists an integer m0 such that, for
m ⩾ m0 , we have, in S, that gm = ∑α cαm f α , where the indices α here are independent of m;
furthermore, we can clearly assume that the cαm are homogeneous, with
deg(cαm ) = m(k + h) − deg f α
in S. So take m0 sufficiently large enough to ensure that km0 > deg f α for all the f α that
appear in gm0 ; for all α, let c′αm be the element cαm considered as having degree km − deg f α
in S♮ ; we then have, in S♮ , that gm = ∑α c′αm f α♮ , which finishes the proof.
□
Since Sn is a quotient S0 -module of S[n] /S+ S[n] , there is a canonical homomorphism of graded
S0 -algebras
(8.2.11.2) S♮ ⊗S S0 −→ S
which is clearly surjective, and thus corresponds (2.9.2) to a canonical closed immersion
(8.2.11.3) Proj(S) −→ Proj(S♮ ⊗S S0 ).
II | 161
8. BLOWUP SCHEMES; BASED CONES; PROJECTIVE CLOSURE 318
Proposition (8.2.12). — The canonical morphism (8.2.11.3) is bijective. For the homomorphism (8.2.11.2) to
be (TN)-bijective, it is necessary and sufficient for there to exist some n0 such that Sn+1 = S1 Sn for n ⩾ n0 .
If this latter condition is satisfied, then (8.2.11.3) is an isomorphism; the converse is true whenever S is
Noetherian.
P ROOF. To prove the first claim, it suffices (2.8.3) to show that the kernel I of the homomorphism
(8.2.11.2) consists of nilpotent elements. But if f ∈ S[n] is an element whose class modulo S+ S[n]
belongs to this kernel, then this implies that f ∈ S[n+1] ; then f n+1 , considered as an element of
S[n(n+1)] , is also an element of S+ S[n(n+1)] , since it can be written as f · f n ; so the class of f n+1 modulo
S+ S[n(n+1)] is zero, which proves our claim. Since the hypothesis that Sn+1 = S1 Sn for n ⩾ n0 is
equivalent to Sn♮ +1 = S1♮ Sn♮ for n ⩾ n0 (8.2.10, ii), this hypothesis is equivalent, by definition, to the
fact that (8.2.11.2) is (TN)-injective, and thus (TN)-bijective, and so (8.2.11.3) is an isomorphism, by
(2.9.1). Conversely, if (8.2.11.3) is an isomorphism, then the sheaf e I on Proj(S♮ ⊗S S0 ) is zero (2.9.2, i);
♮ ♮
since S ⊗S S0 is Noetherian, as a quotient of S (8.2.10, i), we conclude from (2.7.3) that I satisfies
condition (TN), and so Sn♮ +1 = S1♮ Sn♮ for n ⩾ n0 , and this finishes the proof, by (8.2.10, ii). □
(8.2.13). Consider now the canonical injections (S+ )n → S[n] , which define an injective homomor-
phism of degree zero of graded rings
(S+ )n −→ S♮ .
M
(8.2.13.1)
n⩾0
Proj( T ) → Proj( n⩾0 T1n ), then there exists some n0 such that Tn = T1n for n ⩾ n0 .
L
Let gi (1 ⩽ i ⩽ r) be generators of the T0 -module T1 . The hypothesis implies first of all that
the D+ ( gi ) cover Proj( T ) (2.8.1). Let (h j )1⩽j⩽s be a system of homogeneous elements of T+ , with
deg(h j ) = n j , that form, with the gi , a system of generators of the ideal T+ , or, equivalently (2.1.3), a
nj
system of generators of T as a T0 -algebra; if we set T ′ = n⩾0 T1n , then the element h j /gi of the
L
ring T( gi ) must, by hypothesis, belong to the subring T(′g ) , and so there exists some integer k such
i
k+n j
that T1k h j ⊂ T1 for all j. We thus conclude, by induction on r, that T1k hrj ⊂ T ′ for all r ⩾ 1, and,
by definition of the h j , we thus have that T1k T ⊂ T ′ . Also, there exists, for all j, an integer m j such
mj mj mj k
that h j belongs to the ideal of T generated by the gi (2.3.14), so h j ∈ T1 T, and h j ∈ T1k T ⊂ T ′ . II | 162
There is thus an integer m0 ⩾ k such that hm mn
j ∈ T1 for m ⩾ m0 . So, if q is the largest of the integers
n j , then n0 = qsm0 + k is the required number. Indeed, an element of Sn , for n ⩾ n0 , is the sum of
monomials belonging to T1α u, where u is a product of powers of the h j ; if α ⩾ k, then it follows from
β
the above that T1α u ⊂ T1n ; in the other case, one of the exponents of the h j is ⩾ m0 , so u ∈ T1 v, where
β ⩾ k and v is again a product of powers of the h j ; we can then reduce to the previous case, and so
we conclude that T1α u ⊂ T1n in all cases. □
Remark (8.2.15). — The condition Sn = S1n for n ⩾ n0 clearly implies that Sn+1 = S1 Sn for n ⩾ n0 ,
but the converse is not necessarily true, even if we assume that S is Noetherian. For example, let K
be a field, A = K [x], and B = K [y]/y2 K [y], where x and y are indeterminates, with x taken to have
degree 1 and y to have degree 2, and let S = A ⊗K B, so that S is a graded algebra over K that has
8. BLOWUP SCHEMES; BASED CONES; PROJECTIVE CLOSURE 319
a basis given by the elements 1, xn (n ⩾ 1), and xn y (n ⩾ 0). It is immediate that Sn+1 = S1 Sn for
n ⩾ 2, but S1n = Kxn while Sn = Kxn + Kxn y for n ⩾ 2.
(where, in the definition of C, we consider S as a non-graded OY -algebra), and we say that C (resp.
b is the affine cone (resp. projective cone) defined by S ; we will sometimes say “cone” instead of
C)
“affine cone”. By an abuse of language, we also say that C (resp. C) b is the affine cone based at X (?)
2
(resp. the projective cone based at X (?)) , with the implicit understanding that the prescheme X is
given in the form Proj(S ); finally, we say that C b is the projective closure of C (with the data of S
being implicit in the structure of C).
Proposition (8.3.2). — There exist canonical Y-morphisms
ε i
(8.3.2.1) Y−
→C−
→ Cb
j
(8.3.2.2) X−
→ Cb
such that ε and j are closed immersions, and i is an affine morphism, which is a dominant open immersion, for
which
(8.3.2.3) b − j ( X );
i (C ) = C
furthermore, C b containing i (C ).
b is the smallest closed subprescheme of C
(8.3.2.4) bz = Spec(S
C c/(z − 1)S
c)
∼
(3.1.4), where z is canonically identified with a section of S over Y. The isomorphism i : C −
→ Cbz
then corresponds to the canonical isomorphism (8.2.3.1)
∼
S
c/(z − 1)S
c−→ S.
The morphism ε corresponds to the augmentation homomorphism S → S0 = OY , which
has kernel S+ (1.2.7), and, since the latter is surjective, ε is a closed immersion (1.4.10). Finally,
j corresponds (3.5.1) to the surjective homomorphism of degree zero S c → S , which restricts to
the identity on S and is zero on zS c, which is its kernel; j is everywhere defined, and is a closed
immersion, by (3.6.2).
To prove the other claims of (8.3.2), we can clearly restrict to the case where Y = Spec( A) is
affine, and S = S, e with S a graded A-algebra, whence S c= (Sb)e; the homogeneous elements f of
S+ can then be identified with sections of S cover Y, and the open subset of C, b denoted D+ ( f ) in
(2.3.3), can then be written as C f (3.1.4); similarly, the open subset of C denoted D ( f ) in (I, 1.1.1) can
b
be written as C f (0, 5.5.2). With this in mind, it follows from (2.3.14) and from the definition of Sb
2[Trans.] A more literal translation of the French (cône projetant (affine/projectif)) would be the projecting (affine/projective)
cone, but it seems that this terminology already exists to mean something else.
8. BLOWUP SCHEMES; BASED CONES; PROJECTIVE CLOSURE 320
bz = i (C ) and C
that, in this case, the open subsets C b f (with f homogeneous in S+ ) form a cover of C.
b
Furthermore, with this notation,
(8.3.2.5) i −1 ( C
bf ) = C f ;
indeed, C b f ∩ i (C ) = C
bf ∩ Cbz = C b f z = Spec(Sb( f z) ). But, if d = deg( f ), then Sb( f z) is canonically
isomorphic to (Sb(z) ) f /zd (2.2.2), and it follows from the definition of the isomorphism in (8.2.3.1)
that the image of (Sb(z) ) d under the corresponding isomorphism of rings of fractions is exactly S f .
f /z
Since C f = Spec(S f ), this proves (8.3.2.5) and shows, at the same time, that the morphism i is affine;
furthermore, the restriction of i to C f , considered as a morphism to Cb f , corresponds (I, 1.7.3) to the
canonical homomorphism Sb( f ) → Sb( f z) , and, by the above and (8.2.3.2), we can claim the following
result:
Now note that (for Y affine) the complement of C bz in Cb = Proj(Sb) is, by definition, the set of II | 164
graded prime ideals of Sb containing z, which is exactly j( X ), by definition of j, which proves (8.3.2.3).
Finally, to prove the last claim of (8.3.2), we can assume that Y is affine. With the above notation,
b z is not a zero divisor; since i (C ) = C,
note that, in the ring S, b it suffices to prove the following
lemma:
By (I, 4.1.9), the question is local on Z; for every homogeneous element h ∈ Te (e > 0), it thus
suffices to prove that Zh is the smallest closed subprescheme of Zh that contains Zgh ; it follows
from the definitions and from (I, 4.3.2) that this condition is equivalent to asking for the canonical
homomorphism T(h) → T( gh) to be injective. But this homomorphism can be identified with the
canonical homomorphism T(h) → ( T(h) ) ge /hd (2.2.3). But since ge is not a zero divisor in T, ge /hd is
not a zero divisor in Th (nor a fortiori in T(h) ), since the fact that ( ge /hd )(t/hm ) = 0 (for t ∈ T and
m > 0) implies the existence of some n > 0 such that hn ge t = 0, whence hn t = 0, and thus t/hm = 0
in Th . This thus finishes the proof (0, 1.2.2). □
(8.3.3). We will often identify the affine cone C with the subprescheme induced by the projective
cone C b on the open subset i (C ) by means of the open immersion i. The closed subprescheme of
C associated to the closed immersion ε is called the vertex prescheme (?) of C; we also say that
ε, which is a Y-section of C, is the vertex section (?), or the null section, or C; we can identify Y
with the vertex prescheme (?) of C by means of ε. Also, i ◦ ε is a Y-section of C, b and thus also a
closed immersion (I, 5.4.6), corresponding to the canonical surjective homomorphism of degree zero
Sc= S [z] → OY [z] (3.1.7), whose kernel is S+ [z] = S+ S c; the subprescheme of C b associated to
this closed immersion is also called the vertex prescheme (?) of C, and i ◦ ε the vertex section (?) of C;
b b it
can be identified with Y by means of i ◦ ε. Finally, the closed subprescheme of C associated to j is
b
called the part at infinity of C,
b and can be identified with X by means of j.
(8.3.4.1) E = C − ε (Y ) , b − i (ε(Y ))
b=C
E
are called (by an abuse of language) the pointed affine cone and the pointed projective cone (respectively)
defined by S ; we note that, despite this nomenclature, E is not necessarily affine over Y, nor E b
projective over Y (8.4.3). When we identify C with i (C ), we thus have the underlying spaces
(8.3.4.2) C∪E
b = C,
b C∩E
b=E
8. BLOWUP SCHEMES; BASED CONES; PROJECTIVE CLOSURE 321
so that C
b can be considered as being obtained by gluing the open subpreschemes C and E;
b further-
more, by (8.3.2.3),
(8.3.4.3) b − j ( X ).
E=E
If Y = Spec( A) is affine, then, with the notation of (8.3.2),
[ [
(8.3.4.4) E= Cf , b=
E C
bf , Cf = C ∩ C
bf
where f runs over the set of homogeneous elements of S+ (or only a subset M of this set, with M
generating an ideal of S+ whose radical in S+ is S+ itself, or, equivalently, such that the X f for f ∈ M
cover X (2.3.14)). The gluing of C and C b f along C f is thus determined by the injection morphisms
C f → C and C f → C b f , which, as we have seen (8.3.2.6), correspond (respectively) to the canonical
homomorphisms S → S f and S⩽
f → Sf .
Proposition (8.3.5). — With the notation of (8.3.1) and (8.3.4), the morphism associated (3.5.1) to the
canonical injection ϕ : S → S
c= S [z] is a surjective affine morphism (called the canonical retraction)
(8.3.5.1) b −→ X
p:E
such that
(8.3.5.2) p ◦ j = 1X .
P ROOF. To prove the proposition, we can restrict to the case where Y is affine. Taking into
b the fact that the domain of definition G (ϕ) of p is equal to E
account the expression in (8.3.4.4) for E, b
will follow from the first of the following claims:
(8.3.5.3). If Y = Spec( A) is affine, and S = S,
e then, for all homogeneous f ∈ S+ ,
(8.3.5.4) p −1 ( X f ) = C
bf
Indeed, to prove this equation, it suffices to consider some homogeneous f in S+ such that
p ∈ D ( f ), and to note that q( f ) is the inverse image of p f under the injection S( f ) → S f .
Corollary (8.3.8). — If S is generated by S1 , then the morphisms p and π are of finite type; for all x ∈ X,
the fibre p−1 ( x ) is isomorphic to Spec(k( x )[ T ]), and the fibre π −1 isomorphic to Spec(k( x )[ T, T −1 ])
P ROOF. This follows immediately from (8.3.5) and (8.3.6) by noting that, whenever Y is affine
and S is generated by S1 , the X f , for f ∈ S1 , form a cover of X (2.3.14). □
Remark (8.3.9). — The pointed affine cone corresponding to the graded OY -algebra OY [ T ] (where
T is an indeterminate) can be identified with Gm = Spec(OY [ T, T −1 ]), since it is exactly CT , as we
have seen in (8.3.2) (see (8.4.4) for a more general result). This prescheme is canonical endowed with
the structure of a “Y-scheme in commutative groups”. This idea will be explained in detail later on, but,
for now, can be quickly summarised as follows. A Y-scheme in groups is a Y-scheme G endowed
with two Y-morphisms, p : G ×Y G → G and s : G → G, that satisfy conditions formally analogous
to the axioms of the composition law and the symmetry law of a group: the diagram
p ×1
G×G×G / G×G
1× p p
G×G /G
p
should commute (“associativity”), and there should be a condition which corresponds to the fact
that, for groups, the maps
( x, y) 7−→ ( x, x −1 , y) 7−→ ( x, x −1 y) 7−→ x ( x −1 y)
and
( x, y) 7−→ ( x, x −1 , y) 7−→ ( x, yx −1 ) 7−→ (yx −1 ) x
should both reduce to ( x, y) 7→ y; the sequence of morphisms corresponding, for example, to the
first composite map is
(1,s)×1 1× p p
G × G −−−−→ G × G × G −−→ G × G −
→G
and the reader should write down the second sequence.
It is immediate (I, 3.4.3) that the data of a structure of a Y-scheme in groups on a Y-scheme II | 167
G is equivalent to the data, for every Y-prescheme Z, of a group structure on the set HomY ( Z, G ),
where these structures should be such that, for every Y-morphism Z → Z ′ , the corresponding
map HomY ( Z ′ , G ) → HomY ( Z, G ) is a group homomorphism. In the particular case of Gm that we
consider here, HomY ( Z, G ) can be identified with the set of Z-sections of Z ×Y Gm (I, 3.3.14), and
thus with the set of Z-sections of Spec(OZ [ T, T −1 ]); finally, the same reasoning as in (I, 3.3.15) shows
that this set is canonically identified with the set of invertible elements of the ring Γ( Z, OZ ), and the
group structure on this set is the structure coming from the multiplication in the ring Γ( Z, OZ ). The
reader can verify that the morphisms p and s from above are obtained in the following way: they
correspond, by (1.2.7) and (1.4.6), to the homomorphisms of OY -algebras
′
π :OY [ T, T −1 ] −→ OY [ T, T −1 , T ′ , T −1 ]
σ :OY [ T, T −1 ] −→ OY [ T, T −1 ]
which associates, to each section sn ∈ Γ(U, Sn ) (where U is an open subset of Y), the section
sn T n ∈ Γ(U, S ⊗OY OY [ T, T −1 ]).
Conversely, suppose that we are given a quasi-coherent, a priori non-graded, OY -algebra, and,
on C = Spec(S ), a structure of a “Y-scheme in sets endowed with a group of operators” that has the Y-
scheme in groups Gm as its domain of operators; then we canonically obtain a grading of OY -algebras
on S . Indeed, the data of a Y-morphism Gm ×Y C → C is equivalent to that of a homomorphism of
OY -algebras ψ : S → S [ T, T −1 ], which can be written as ψ = ∑n∈Z ψn T n , where the ψn : S → S
are homomorphisms of OY -modules (with ψn (s) = 0 except for finitely many n for every section
s ∈ Γ(U, S ), for any open subset U of Y). We can then prove that the axioms of sets endowed with
a group of operators imply that the ψn (S ) = Sn define a grading (in positive or negative degree) of
OY -algebras on S , with the ψn being the corresponding projectors. We also have the notation of a
structure of an “affine cone” on every affine Y-scheme, defined in a “geometric” way without any
reference to any prior grading. We will not further develop this point of view here, and we leave the II | 168
work of precisely formulating the definitions and results corresponding to the information given
above to the reader.
8.4. Projective closure of a vector bundle
(8.4.1). Let Y be a prescheme, and E a quasi-coherent OY -module. If we take S to be the graded
OY -algebra SOY (E ), then Definition (8.3.1.1) shows that S
ccan be identified with SO (E ⊕ OY ). With
Y
the affine cone Spec(S ) defined by S being, by definition, V(E ), and Proj(S ) being, by definition,
P(E ), we see that:
Proposition (8.4.2). — The projective closure of a vector bundle V(E ) on Y is canonically isomorphic to
P(E ⊕ OY ), and the part at infinity of the latter is canonically isomorphic to P(E ).
Remark (8.4.3). — Take, for example, E = OYr with r ⩾ 2; then the pointed cones E and E b defined
by S are nether affine nor projective on Y if Y ̸= ∅. The second claim is immediate, because
Cb = P(O r+1 ) is projective on Y, and the underlying spaces of E and E b are non-closed open subsets
Y
b and so the canonical immersions E → C
of C, b→C
b and E b are not projective (5.5.3), and we conclude
by appealing to (5.5.5, v). Now, supposing, for example, that Y = Spec( A) is affine, and r = 2, then
C = Spec( A[ T1 , T2 ]), and E is then the prescheme induced by C on the open subset D ( T1 ) ∪ D ( T2 );
but we have already seen that the latter is not affine (I, 5.5.11); a fortiori E
b cannot be affine, since E is
the open subset where the section z over E b does not vanish (8.3.2).
However:
Proposition (8.4.4). — If L is an invertible OY -module, then there are canonical isomorphisms for both the
b corresponding to C = V(L ):
pointed cones E and E
!
∼
L ⊗n
M
(8.4.4.1) Spec −
→E
n∈Z
∼ b
(8.4.4.2) V (L −1 ) −
→ E.
Furthermore, there exists a canonical isomorphism from the projective closure of V(L ) to the projective
closure of V(L −1 ) that sends the null section (resp. the part at infinity) of the former to the part at infinity
(resp. the null section) of the second.
P ROOF. We have here that S = ⊗n ;
n⩾0 L
L
the canonical injection
S −→ L ⊗n
M
n∈Z
defines a canonical dominant morphism
! !
⊗n ⊗n
L −→ V(L ) = Spec L
M M
(8.4.4.3) Spec
n∈Z n⩾0
and it suffices to prove that this morphism is an isomorphism from the scheme Spec( n∈Z L ⊗n ) to
L
E. Since the questions is local on Y, we can assume that Y = Spec( A) is affine and that L = OY , II | 169
8. BLOWUP SCHEMES; BASED CONES; PROJECTIVE CLOSURE 324
of A[ T ], and thus (8.4.4.3) identifies n∈Z L ⊗n (?) with the prescheme induced by C = V(L ) on
L
the open subset D ( T ); the complement V ( T ) of this open subset in C is the underlying space of the
closed subprescheme of C defined by the ideal TA[ T ], which is exactly the null section of C, and so
E = D ( T ).
The isomorphism in (8.4.4.2) will be a consequence of the last claim, since V(L −1 ) is the
complement of the part at infinity of its projective closure, and E b is the complement of the null
section of the projective closure C = V(L ). But these projective closures are P(L −1 ⊕ OY ) and
P(L ⊕ OY ) (respectively); but we can write L ⊕ OY = L ⊗ (L −1 ⊕ OY ). The existence of the
desired canonical isomorphism then follows from (4.1.4), and everything reduces to showing that
this isomorphism swaps the null sections and the parts at infinity. For this, we can reduce to the case
where Y = Spec( A) is affine, L = Ac, and L−1 = Ac′ , with the canonical isomorphism L ⊗ L−1 → A
sending c ⊗ c′ to the element 1 of A. Then S( L ⊕ A) is the tensor product of A[z] with n⩾0 Ac⊗n ,
L
′
and S( L−1 ⊕ A) is the tensor product of A[z] with n⩾0 Ac ⊗n , and the isomorphism defined in
L
′
(4.1.4) sends zh ⊗ c ⊗(n−h) to the element zn−h ⊗ c⊗h . But, in P(L −1 ⊕ OY ), the part at infinity is
the set of points where the section z vanishes, and the null section is the section of points where
the section c′ vanishes; since we have analogous definitions for P(L ⊕ OY ), the conclusion follows
immediately from the above explanation. □
8.5. Functorial behaviour
(8.5.1). Let Y and Y ′ be prescheme, q : Y ′ → Y a morphism, and S (resp. S ′ ) a quasi-coherent
positively-graded OY -algebra (resp. quasi-coherent positively-graded OY ′ -algebra). Consider a q-
morphism of graded algebras
(8.5.1.1) ϕ : S −→ S ′ .
We know (1.5.6) that this corresponds, canonically, to a morphism
Φ = Spec(ϕ) : Spec(S ′ ) −→ Spec(S )
such that the diagram
Φ /C
(8.5.1.2) C′
Y′ /Y
q
commutes, where we write C = Spec(S ) and C ′ = Spec(S ′ ). Suppose, further, that S0 = OY and
S0′ = OY ′ ; let ε : Y → C and ε : Y ′ → C ′ be the canonical immersions (8.3.2); we then have a
commutative diagram
q
(8.5.1.3) Y′ /Y
ε′ ε
C′ /C
Φ
OY / OY ′
where the vertical arrows are the augmentation homomorphisms, and so the commutativity follows
from the hypothesis that ϕ is assumed to be a homomorphism of graded algebras.
Proposition (8.5.2). — If E (resp. E′ ) is the pointed affine cone defined by S (resp. S ′ ), then Φ−1 ( E) ⊂ E′ ;
if, further, Proj(ϕ) : G (ϕ) → Proj(S ) is everywhere defined (or, equivalently, if G (ϕ) = Proj(S ′ )), then
Φ−1 ( E) = E′ , and conversely.
8. BLOWUP SCHEMES; BASED CONES; PROJECTIVE CLOSURE 325
P ROOF. The first claim follows from the commutativity of (8.5.1.3). To prove the second, we can
restrict to the case where Y = Spec( A) and Y ′ = Spec( A′ ) are affine, and S = Se and S ′ = Se′ . For
every homogeneous f in S+ , writing f ′ = ϕ( f ), we have that Φ−1 (C f ) = C ′f ′ (I, 2.2.4.1); saying that
G (ϕ) = Proj(S′ ) implies that the radical (in S+ ′ ) of the ideal generated by the f ′ = ϕ ( f ) is S′ itself
+
((2.8.1) and (2.3.14)), and this is equivalent to saying that the C ′f ′ cover E′ (8.3.4.4). □
Y′ /Y
q
commutes (3.5.6). It follows immediately from the definitions that, if we write i : C → C b and
′ ′ b′ ′ ′
i : C → C to mean the canonical open immersions (8.3.2), then i (C ) ⊂ G (ϕ), and the diagram
b
Φ /C
(8.5.3.2) C′
i i′
G (ϕ b) /C
b
Φ
b
commutes. Finally, if we let X = Proj(S ) and X ′ = Proj(S ′ ), and if j : X → C b and j′ : X ′ → Cb′ are
the canonical closed immersions (8.3.2), then it follows from the definition of these immersions that
j′ ( G (ϕ)) ⊂ G (ϕ
b), and that the diagram II | 171
Proj(ϕ)
(8.5.3.3) G (ϕ) /X
j′ j
G (ϕ b) /C
b
Φ
b
commutes.
Proposition (8.5.4). — If E b (resp. Eb′ ) is the pointed projective cone defined by S (resp. by S ′ ),
then Φ −
b (E1 b → X and p′ : Eb′ → X ′ are the canonical retractions, then
b) ⊂ Eb′ ; furthermore, if p : E
p′ (Φ
b −1 ( E
b)) ⊂ G (ϕ
b), and the diagram
b −1 ( E Φ /E
Φ
b
(8.5.4.1) b) b
p′ p
G (ϕ) /X
Proj(ϕ)
P ROOF. The first claim follows from the commutativity of Diagrams (8.5.1.3) and (8.5.3.2), and
the two following claims from the definition of the canonical retractions (8.3.5) and the definition
b. To see that Φ
of ϕ b is everywhere defined whenever Proj(ϕ) is, we can restrict to the case where
Y = Spec( A) and Y ′ = Spec( A′ ) are affine, and where S = Se and S ′ = Se′ ; the hypothesis is that,
′ of the ideal generated in
when f runs over the set of homogeneous elements of S+ , the radical in S+
S+ by the ϕ( f ) is S+ itself; we thus immediately conclude that the radical in (S′ [z])+ of the ideal
′ ′
8. BLOWUP SCHEMES; BASED CONES; PROJECTIVE CLOSURE 326
generated by z and the ϕ( f ) is (S′ [z])+ itself, whence our claim; this also shows that Eb′ is the union
b −1 ( E
of the Cb′ ϕ( f ) , and hence equal to Φ b). □
Corollary (8.5.5). — Whenever Proj(ϕ) is everywhere defined, the inverse image under Φ b of the underlying
′
space of the part at infinity (resp. of the vertex prescheme) of C is the underlying space of the part at infinity
b
(resp. of the vertex prescheme) of C.b
P ROOF. This follows immediately from (8.5.4) and (8.5.2), taking into account the equalities
(8.3.4.1) and (8.3.4.2). □
8.6. A canonical isomorphism for pointed cones
(8.6.1). Let Y be a prescheme, S a quasi-coherent positively-graded OY -algebra such that S0 = OY ,
and let X be the Y-scheme Proj(S ). We are going to apply the results of (8.5) to the case where
Y ′ = X, and q : X → Y is the structure morphism; let
SX = OX (n)
M
(8.6.1.1)
n∈Z
which is a quasi-coherent graded OX -algebra, with multiplication defined by means of the canonical II | 172
homomorphisms (3.2.6.1)
OX (m) ⊗OX OX (n) −→ OX (m + n)
whose associativity is ensured by the commutative diagram in (2.5.11.4). Let S ′ be the quasi-
coherent positively-graded OX -subalgebra SX⩾ = n⩾0 OX (n) of SX .
L
(8.6.2.1) Γ( X f , SX⩾ ) = S⩾
f .
So u−1 ( X f ) = Proj(S⩾ ⩾
f ). But if f ∈ Sd (d > 0), then Proj( S f ) is canonically isomorphic
to Proj((S⩾ (d) ⩾ (d)
f ) ) (2.4.7), and we also know that ( S f ) = ( S ( d ) )⩾
f can be identified with S( f ) [ T ]
(2.2.1) by the map T 7→ f /1; we thus conclude (3.1.7) that the structure morphism u−1 ( X f ) →
X f is an isomorphism, whence the first claim. To prove the second, note that the restriction
u−1 ( X f ) ∩ G (α) → X = Proj(S) of Proj(α) corresponds to the canonical map x 7→ x/1 from
S to S⩾ ′
f (2.6.2); we thus deduce, first of all, that G ( α ) = X , and then, taking into account the
fact that u−1 ( X f ) = (u−1 ( X f )) f /1 , that it follows from (2.8.1.1) that the image of u−1 ( X f ) under
Proj(α) is contained in X f , and the restriction of Proj(α) to u−1 ( X f ), considered as a morphism to
X f = Spec(S( f ) ), is indeed identical to that of u. Finally, applying (8.3.5.4) to p X instead of p, we see
⩾ ⩽
that p− 1 −1
X ( u ( X f )) = Spec(( S f ) f /1 ), and this open subset is, by (8.5.4.1), the inverse image under
α) of p−1 ( X f ) = Spec(S⩽
Proj(b f ) (8.3.5.3). Taking (8.2.3.2) into account, the restriction of Proj( α ) to
b
8. BLOWUP SCHEMES; BASED CONES; PROJECTIVE CLOSURE 327
⩽
p− 1 −1
X ( u ( X f )) corresponds to the isomorphism inverse to (8.2.7.2), restricted to S f , whence the third
claim; the last claim is evident by definition.
We note also that it follows from the commutative diagram in (8.5.3.2) that the restriction to CX of II | 173
Proj(bα) is exactly the morphism Spec(α). □
Corollary (8.6.4). — Assume that OX (1) is an invertible OX -module, and that SX is isomorphic to
⊗n (which will be the case, in particular, whenever S is generated by S ((3.2.5) and (3.2.7))).
n∈Z (OX (1))
L
1
Then the pointed projective cone E b can be identified with the rank-1 vector bundle V(OX (−1)) on X, and
the pointed affine cone E with the subprescheme of this vector bundle induced on the complement of the null
b → X is identified with the structure morphism of
section. With this identification, the canonical retraction E
the X-scheme V(OX (−1)). Finally, there exists a canonical Y-morphism V(OX (1)) → C, whose restriction
to the complement of the null section of V(OX (1)) is an isomorphism from this complement to the pointed
affine cone E.
of Proj(b
α), and the claims of the corollary are then particular cases of (8.6.2). □
We note that the inverse image under the morphism V(OX (1)) → C of the underlying space
of the vertex prescheme of C is the underlying space of the null section of V(OX (1)) (8.5.5); but, in
general, the corresponding subpreschemes of C and of V(OX (1)) are not isomorphic. This problem
will be studied below.
q r
Y
i◦ε /C
b
by (8.5.1.3) and (8.5.3.2); furthermore, the restriction of r to the complement C bX − i X (ε X ( X )) of the II | 174
null section is an isomorphism to the complement C b − i (ε(Y )) of the null section, by (8.6.2). If we
suppose, to simplify things, that Y is affine, that S is of finite type and generated by S1 , and that X
is projective over Y and C bX projective over X (5.5.1), then CbX is projective over Y (5.5.5, ii), and a
fortiori over C
b (5.5.5, v). We then have a projective Y-morphism r : C bX → Cb (whose restriction to CX
is a projective Y-morphism CX → C) that contracts X to Y (?) and that induces an isomorphism when
we restrict to the complements of X and Y. We thus have a connection between CX and C, analogous
to that which exists between a blow-up prescheme and the original prescheme (8.1.3). We will
effectively show that CX can be identified with the homogeneous spectrum of a graded OC -algebra.
8. BLOWUP SCHEMES; BASED CONES; PROJECTIVE CLOSURE 328
(8.7.2). Keeping the notation of (8.6.1), consider, for all n ⩾ 0, the quasi-coherent ideal
S[ n ] = Sm
M
(8.7.2.1)
m⩾n
of the graded OY -algebra S . It is clear that
(8.7.2.2) S[0] = S , S[ n ] ⊂ S[ m ] for m ⩽ n
(8.7.2.3) Sn S[ m ] ⊂ S[ m + n ] .
Consider the OC -module associated to S[n] , which is a quasi-coherent ideal of OC = S
f(1.4.4)
(8.7.2.6) In I[ m ] ⊂ I[ m + n ] .
We are thus in the setting of (8.1.1), which leads us to introduce the quasi-coherent graded
OC -algebra
!
S♮ = In = S[ n ]
M M
(8.7.2.7) e.
n⩾0 n⩾0
But with the notation of (8.6.2), if v : CX → X is the structure morphism, then it follows from II | 175
♮
(8.6.2.1) that v−1 ( X f ) = Spec(S⩾ ♮
f ). We also have that Spec( S( f ♮ ) ) = D+ ( f ), which means that
(8.7.3.2) defines an isomorphism v−1 ( X f ) → D+ ( f ♮ ). Furthermore, if g ∈ Se (e > 0), then the
diagram
∼ /
v −1 ( X f g ) D+ ( f ♮ g ♮ )
v −1 ( X f )
∼ / D+ ( f ♮ )
(8.7.3.3) CU ′ / Proj(S′ ♮ )
CU / Proj(S♮ )
′
commutes. But S is canonically identified with S ⊗ A A′ , and so S ♮ is canonically identified with
S♮ ⊗S S′ = S♮ ⊗ A A′ ;
8. BLOWUP SCHEMES; BASED CONES; PROJECTIVE CLOSURE 329
′
thus Proj(S ♮ ) = Proj(S♮ ) ×U U ′ (2.8.10); similarly, if X = Proj(S) and X ′ = Proj(S′ ), then X ′ =
X ×U U ′ and SX ′ = SX ⊗OU U ′ (3.5.4), or, equivalently, SX ′ = j∗ (SX ), where j is the projection
X ′ → X. We then (1.5.2) have that CU ′ = CU × X X ′ = CU ×U U ′ , and the commutativity of (8.7.3.3)
is then immediate. □
Remark (8.7.4). — (i) The end of the proof of (8.7.3) can be immediately generalised in the
following way. Let g : Y ′ → Y be a morphism, S ′ = g∗ (S ), and X ′ = Proj(S ′ ); then we
have a commutative diagram
(8.7.4.1) CX ′ / Proj(S ′ ♮ )
CX / Proj(S ♮ )
(8.7.4.2) CX ′′
∼ / Proj(S ′′♮ )
w Proj(ψ)
CX
∼ / Proj(S ♮ )
Corollary (8.7.5). — Suppose that there exists some n0 > 0 such that
(8.7.5.1) Sn+1 = S1 Sn for n ⩾ n0 .
Then the vertex subprescheme (?) of CX (isomorphic to X) is the inverse image under the canonical
morphism r : CX → C of the vertex subprescheme of C (isomorphic to Y). Conversely, if this property is true,
and if we further assume that Y is Noetherian and that S is of finite type, then there exists some n0 > 0 such
that (8.7.5.1) holds true.
P ROOF. Since the first claim is local on Y, we can assume that Y = Spec( A) is affine, so that
S = S, e with S a positively-graded A-algebra. The claim then follows from (8.2.12), since Proj(S♮ ⊗S
S0 ) = CX ×C ε(Y ) (by the identification in (8.7.3.1)), or, in other words, since this prescheme is
the inverse image of ε(Y ) in CX (I, 4.4.1). The converse also follows from (8.2.12) whenever Y is
8. BLOWUP SCHEMES; BASED CONES; PROJECTIVE CLOSURE 330
Noetherian affine and S is of finite type. If Y is Noetherian (but not necessarily affine) and S is of II | 177
finite type, then there exists a finite cover of Y by Noetherian affine open subsets Ui , and we then
deduce from the above that, for all i, there exists an integer ni such that Sn+1 |Ui = (S1 |Ui )(Sn |Ui )
for n ⩾ ni ; the largest of the ni then ensures that (8.7.5.1) holds true. □
(8.7.6). Now consider the C-prescheme Z given by blowing up the vertex subprescheme ε(Y ) in the affine
cone C; by Definition (8.1.3), it is exactly the prescheme Proj( n⩾0 S+n ); the canonical injection
L
S+n −→ S ♮
M
(8.7.6.1) ι:
n⩾0
(8.7.6.2) G (ι) −→ Z
where G (ι) is an open subset of CX (3.5.1); note that it could be the case that G (ι) ̸= CX , as shown
by the example where Y = Spec(K ), with K a field, and S = S, e with S = K [y], where y is an
indeterminate of degree 2; if Rn denotes the set (S+ ) , considered as a subset of S[n] = Sn♮ , then S+
n ♮
is
♮
not the radical in S+ of the ideal generated by the union of the Rn (cf. (2.3.14)).
Corollary (8.7.7). — Assume that there exists some n0 > 0 such that
P ROOF. The first claim is local on Y, and thus follows from (8.2.14); the converse follows
similarly, arguing as in (8.7.5). □
Remark (8.7.8). — Since condition (8.7.7.1) implies (8.7.5.1), we see that, whenever it holds true, not
only can CX be identified with the prescheme given by blowing up the vertex (identified with Y)
of the affine cone C, but also the vertex (identified with X) of CX can be identified with the closed
subprescheme given by the inverse image of the vertex Y of C. Furthermore, hypothesis (8.7.7.1)
implies that, on X = Proj(S ), the OX -modules OX (n) are invertible ((3.2.5) and (3.2.9)), and that
OX (n) = L ⊗n with L = OX (1) ((3.2.7) and (3.2.9)); by Definition (8.6.1.1), CX is thus the vector
bundle V(L ) on X, and its vertex is the null section of this vector bundle.
S = OY ⊕ f ∗ (L ⊗ n )
M
(8.8.1.1)
n⩾1
which is quasi-coherent (I, 9.2.2, a). There is a canonical homomorphisms of graded OX -algebras II | 178
τ : f ∗ (S ) −→ L ⊗n
M
(8.8.1.2)
n⩾0
(8.8.1.3) r = rL ,τ : X −→ P = Proj(S )
Proposition (8.8.2). — Let C = Spec(S ) be the affine cone defined by S ; if L is f -ample, then there
exists a canonical Y-morphism
(8.8.2.1) g : V = V(L ) −→ C
such that the diagram
j
(8.8.2.2) X / V (L ) π /X
f g f
/C /Y
ε ψ
Y
commutes, where ψ and π are the structure morphisms, and j and ε the canonical immersions sending X and
Y (respectively) to the null section of V(L ) and the vertex prescheme of C (respectively). Furthermore, the
restriction of g to V(L ) − j( X ) is an open immersion
(8.8.2.3) V(L ) − j( X ) −→ E = C − ε(Y )
into the pointed affine cone E corresponding to S .
P ROOF. With the notation of (8.8.1), let SP⩾ = n⩾0 OP (n) and C p = Spec(SP⩾ ). We know
L
(8.6.2) that there is a canonical morphism h = Spec(α) : C p → C such that the diagram
(8.8.2.4) CP /P
h p
/Y
ψ
C
commutes; furthermore, if ε P : P → CP is the canonical immersion, then the diagram
p
(8.8.2.5) P / CP
εP h
Y
ε /C
commutes (8.7.1.1), and, finally, the restriction of H to the pointed affine cone EP is an isomorphism
∼
EP −
→ E (8.6.2). It follows from (8.8.1.4) that
r ∗ (SP⩾ ) = SOX (L )
and so we have a canonical P-morphism q : V(L ) → CP , with the commutative diagram II | 179
(8.8.2.6) V (L )
π /X
q r
CP /P
identifying V(L ) with the product CP × P X (1.5.2); since r is an open immersion, so too is q (I, 4.3.2).
Furthermore, the restriction of q to V(L ) − j( X ) sends this prescheme to E p , by (8.5.2), and the
diagram
j
(8.8.2.7) X / V (L )
r q
P
εP
/ CP
is commutative (since it is a particular case of (8.5.1.3)). The claims of (8.8.2) immediately follow
from these facts, by taking g to be the composite morphism h ◦ q. □
Remark (8.8.3). — Assume further that Y is a Noetherian prescheme, and that f is a proper morphism.
Since r is then proper(5.4.4), and thus closed, and since it is also a dominant open immersion, r is
∼
necessarily an isomorphism X − → P. Furthermore, we will see, in Chapter III (III, 2.3.5.1), that S is
then necessarily an OY -algebra of finite type. It then follows that S ♮ is an S0♮ -algebra of finite type
8. BLOWUP SCHEMES; BASED CONES; PROJECTIVE CLOSURE 332
((8.2.10, i) and (8.7.2.7)); since CP is C-isomorphic to Proj(S ♮ ) (8.7.3), we see that the morphism
h : CP → C is projective; since the morphism r is an isomorphism, so too is q : V(L ) → CP , and we
thus conclude that the morphism g : V(L ) → C is projective. Furthermore, since the restriction of
h to EP is an isomorphism to E, and since q is an isomorphism, the restriction (8.8.2.3) of g is an
∼
isomorphism V(L ) − j( X ) − → E.
If we further assume that L is very ample for f , then, as we will also see in Chapter III (III, 2.3.5.1),
there exists some integer n0 > 0 such that Sn = S1n for n ⩾ n0 . We then conclude, by (8.7.7), that
V(L ) can be identified with the prescheme Z given by blowing up the vertex prescheme (identified
with Y) in the affine cone C, and that the null section of V(L ) (identified with Y) is the inverse image of
the vertex subprescheme Y of C.
Some of the above results can in fact be proven even without the Noetherian hypothesis:
Corollary (8.8.4). — Let Y be a prescheme (resp. a quasi-compact scheme), f : X → Y a proper morphism,
and L an invertible OX -module that is ample relative to f . Then the morphism in (8.8.2.1) is proper (resp.
projective), and its restriction (8.8.2.3) is an isomorphism.
P ROOF. To prove that g is proper, we can restrict to the case where Y is affine, and it then suffices
to consider the case where Y is a quasi-compact scheme. The same arguments as in (8.8.3) first of all
∼
show that r is an isomorphism X − → P; then q is also an isomorphism, and, since the restriction of h to
∼
EP is an isomorphism EP − → E, we have already seen that (8.8.2.3) is an isomorphism. It remains
only to prove that g is projective.
Since f is of finite type, by hypothesis, we can apply (3.8.5) to the homomorphism τ from II | 180
(8.8.1.2): there is an integer d > 0 and a quasi-coherent OY -submodule E of finite type of Sd such
that, if S ′ is the OY -subalgebra of S generated by E , and τ ′ = τ ◦ q∗ (ϕ) (where ϕ is the canonical
injection S ′ → S ), then r ′ = rL ,τ ′ is an immersion
X −→ P′ = Proj(S ′ ).
Furthermore, since ϕ is injective, r ′ is also a dominant immersion (3.7.6); the same argument as for
r Φ
r then shows that r ′ is a surjective closed immersion; since r ′ factors as X −→ Proj(S ) − → Proj(S ′ ),
where Φ = Proj(ϕ), we thus conclude that Φ is also a surjective closed immersion. But this implies that
Φ is an isomorphism; we can restrict to the case where Y = Spec( A) is affine, and S = Se and S ′ = Se′ ,
with S a graded A-algebra and S′ a graded subalgebra of S. For every homogeneous element t ∈ S′ ,
we have that S(′ t) is a subring of S(t) ; if we return to the definition of Proj(ϕ) (2.8.1), we see that
it suffices to prove that, if B′ is a subring of a ring B, and if the morphism Spec( B) → Spec( B′ )
corresponding to the canonical injection B′ → B is a closed immersion, then this morphism is
necessarily an isomorphism; but this follows from (I, 4.2.3). Furthermore, Φ∗ (OP′ (n)) = OP (n)
′ ′
((3.5.2, ii) and (3.5.4)), and so r ∗ (OP′ (n)) is isomorphic to L ⊗n (4.6.3). Let S ′′ = S (d) , so that
(3.1.8, i) X is canonically identified with P′′ = Proj(S ′′ ), and L ′′ = L ⊗d with OP′′ (1) (3.2.9, ii).
Now, if C ′′ = Spec(S ′′ ), then SP⩾′′ = n⩾0 OP′′ (n) can be identified with n⩾0 L ′′⊗n , and
L L
thus CP′′ = Spec(SP⩾′′ ) with V(L ′′ ); we also know (8.7.3) that CP′′ is C ′′ -isomorphic to Proj(S ′′♮ );
by the definition of S ′′ , we know that S ′′♮ is generated by S1′′♮ , and that S1′′♮ is of finite type over
S0′′♮ = S ′′ ((8.2.10, i and iii)), and so Proj(S ′′♮ ) is projective over C ′′ (5.5.1). Consider the diagram
g
(8.8.4.1) V (L ) / Spec(S ) = C
u v
g′′
V(L ′′ ) / Spec(S ′′ ) = C ′′
n⩾0 n⩾0
(3.3.2.3) (see (8.8.5) below), and v and u to the inclusion morphisms S ′′ → S and n⩾0 L ⊗nd →
L
⊗n (respectively); it is immediate (3.3.2) that this diagram is commutative. We have just seen
n⩾0 L
L
′′
that g is a projective morphism; we also know that u is a finite morphism. Since the question is local
8. BLOWUP SCHEMES; BASED CONES; PROJECTIVE CLOSURE 333
on X, we can assume that X is affine of ring A, and that L = OX ; everything then reduces to noting
that the ring A[ T ] is a module of finite type over its subring A[ T d ] (with T an indeterminate). Since
Y is a quasi-compact scheme, and since C ′′ is affine over Y, we know that C ′′ is also a quasi-compact
scheme, and so g′′ ◦ u is a projective morphism (5.5.5, ii); by commutativity of (8.8.4.1), v ◦ g is also II | 181
projective, and, since v is affine, thus separated, we finally conclude that g is projective (5.5.5, v). □
(8.8.5). Consider again the situation in (8.8.1). We will see that the morphism g : V(L ) → C can be
also be defined in a way that works for any invertible (but not necessarily ample) OX -module L .
For this, consider the f -morphism
τ ♭ : S −→ L ⊗n
M
(8.8.5.1)
n⩾0
corresponding to the morphism τ from (8.8.1.2). This induces (1.5.6) a morphism g′ : V → C such
that, if π : V → X and ψ : C → Y are the structure morphisms, the diagrams
j
Xo /V
π
(8.8.5.2) V X
f g′ f g′
Yo /C
ψ ε
C Y
commute ((8.5.1.2) and (8.5.1.3)). We will show that (if we assume that L is f -ample) the morphisms
g and g′ are identical.
Since the questions is local on Y, we can assume that Y = Spec( A) is affine, and (by (8.8.1.3))
identify X with an open subset of P = Proj(S), where S = A ⊕ n⩾0 Γ( X, L ⊗n ); we then deduce,
L
by (8.8.1.4), that Γ( X, OP (n)) = Γ( X, L ⊗n ) for all n ∈ Z. Taking into account the definition of
h = Spec(α), where α is the canonical p-morphism Se → SP⩾ (8.6.1.2), we have to show that the
restriction to X of α♯ : p∗ (Se) → SP⩾ is identical to τ. Taking (0, 4.4.3) into account, it suffices to
show that, if we compose the canonical homomorphism αn : Sn → Γ( P, OP (n)) with the restriction
homomorphism Γ( P, OP (n)) → Γ( X, OP (n)) = Γ( X, L ⊗n ), then we obtain the identity, for all n > 0;
but this follows immediately from the definition of the algebra S and of αn (2.6.2).
Proposition (8.8.6). — Assume (with the notation of (8.8.5)) that, if we write f = ( f 0 , λ), then the
homomorphism λ : OY → j∗ (OX ) is bijective; then:
(i) if we write g = ( g0 , µ), then µ : OC → g∗ (OV ) is an isomorphism; and
(ii) if X is integral (resp. locally integral and normal), then C is integral (resp. normal).
P ROOF. Indeed, the f -morphism τ ♭ is then an isomorphism
τ ♭ : S = ψ∗ (OC ) −→ f ∗ (π∗ (OV )) = ψ∗ ( g∗ (OV ))
and the Y-morphism g can be considered as that for which the homomorphism A ( g) (1.1.2) is equal
to τ ♭ . To see that µ is an isomorphism of OC -modules, it suffices (1.4.2) to see that A (µ) : ψ∗ (OC ) →
ψ∗ ( g∗ (OV )) is an isomorphism. But, by Definition (1.1.2), we have that A (µ) = A ( g), whence the
conclusion of (i).
e with S = Ln⩾0 Γ( X, L ⊗n ); II | 182
To prove (ii), we can restrict to the case where Y is affine, and so S = S,
the hypothesis that X is integral implies that the ring S is integral (I, 7.4.4), and thus so too is C
(I, 5.1.4). To show that C is normal, we will use the following lemma:
Lemma (8.8.6.1). — Let Z be a normal integral prescheme. Then the ring Γ( Z, OZ ) is integral and integrally
closed.
P ROOF. It follows from (I, 8.2.1.1) that Γ( Z, OZ ) is the intersection, in the field of rational
functions R( Z ), of the integrally closed rings Oz over all z ∈ Z. □
With this in mind, we first show that V is locally integral and normal; for this, we can restrict to
the case where X = Spec( A) is affine, with ring A integral and integrally closed (6.3.8), and where
L = OX . Since then V = Spec( A[ T ]), and A[ T ] is integral and integrally closed [Jaf60, p. 99], this
proves our claim. For every affine open subset U of C, g−1 (U ) is quasi-compact, since the morphism
g is quasi-compact; since V is locally integral, the connected components of g−1 (U ) are open integral
8. BLOWUP SCHEMES; BASED CONES; PROJECTIVE CLOSURE 334
preschemes in g−1 (U ), and thus finite in number, and, since V is normal, these preschemes are
also normal (6.3.8). Then Γ(U, OC ), which is equal to Γ( g−1 (U ), OV ), by (i), is the direct sum (?) of
finitely-many integral and integrally closed rings (8.8.6.1), which proves that C is normal (6.3.4). □
8.9. Grauert’s ampleness criterion: statement We intend to show that the properties proven in
(8.8.2) characterise f -ample OX -modules, and, more precisely, to prove the following criterion:
Theorem (8.9.1). — (Grauert’s criterion). Let Y be a prescheme, p : X → Y a separated and quasi-compact
morphism, and L an invertible OX -module. For L to be ample relative to p, it is necessary and sufficient for
there to exist a Y-prescheme C, a Y-section ε : Y → C of C, and a Y-morphism q : V(L ) → C, satisfying
the following properties:
(i) the diagram
j
(8.9.1.1) X / V (L )
p q
Y
ε /C
commutes, where j is the null section of the vector bundle V(L ); and
(ii) the restriction of q to V(L ) − j( X ) is a quasi-compact open immersion
V(L ) − j( X ) −→ X
whose image does not intersect ε(Y ).
Note that, if C is separated over Y, we can, in condition (ii), remove the hypothesis that the open
immersion is quasi-compact; to see that this property (of quasi-compactness) is in fact a consequence
of the other conditions, we can restrict to the case where Y is affine, and the claim then follows from
(I, 5.5.1)i and (I, 5.5.10). We can also remove the same hypothesis if we assume that X is Noetherian, II | 183
since then V is also Noetherian, and the claim follows from (I, 6.3.5).
Corollary (8.9.2). — If the morphism p : X → Y is proper, then we can, in the statement of Theorem (8.9.1),
assume that q is proper, and replace “open immersion” by “isomorphism”.
In a more suggestive manner, we can say (whenever p : X → Y is proper) that L is ample relative
to p if and only if we can “contract” the null section of the vector bundle V(L ) to the base prescheme Y.
An important particular case is that where Y is the spectrum of a field, and where the operation of
“contraction” consists of contract the null section V(L ) to a single point.
(8.9.3). The necessity of the conditions in Theorem (8.9.1) and Corollary (8.9.2) follow immediately
from (8.8.2) and (8.8.4).
To show that the conditions of (8.9.1) suffices, consider a slightly more general situation. For
this, let (with the notation of (8.8.2))
S′ = L ⊗n
M
n⩾0
and
V = V(L ) = Spec(S ′ ).
The closed subprescheme j( X ), null section of V(L ), is defined by the quasi-coherent sheaf
of ideals J = (S+′ )e of OV (1.4.10). This OV -module is invertible, since this property is local on
X, and this reduces to remarking that the ideal TA[ T ] in a ring of polynomials A[ T ] is a free cyclic
A[ T ]-module. Furthermore, it is immediate (again, because the question is local on X) that
L = j ∗ (J )
and
j∗ (L ) = J /J 2 .
Now, if
π : V(L ) −→ X
is the structure morphism, then π∗ (J ) = S+′ and π∗ (J /J 2 ) = L ; there are thus canonical ho-
momorphisms L → π∗ (J ) → L , the first being the canonical injection L → S+′ , and the second
8. BLOWUP SCHEMES; BASED CONES; PROJECTIVE CLOSURE 335
the canonical projection from S+′ to S1′ = L , and their composition being the identity. We can also
canonically embed π∗ (J ) = S+′ = n⩾1 L ⊗n into the product ∏n⩾1 L ⊗n = limn π∗ (J /J n+1 )
L
←−
(since π∗ (J /J n+1 ) = L ⊕ L ⊗2 ⊕ . . . ⊕ L ⊗n ), and we thus have canonical homomorphisms
(8.9.3.1) L −→ lim π∗ (J /J n+1 ) −→ L
←−
whose composition is the identity.
With this in mind, the generalisation of (8.9.1) that we are going to prove is the following:
Proposition (8.9.4). — Let Y be a prescheme, V a Y-prescheme, and X a closed subprescheme of V defined
by an ideal J of OV , which is an invertible OV -module; if j : X → V is the canonical injection, then let II | 184
L = j∗ (J ) = J ⊗OV OX , so that j∗ (L ) = J /J 2 . Assume that the structure morphism p : X → Y
is separated and quasi-compact, and that the following conditions are satisfied:
(i) there exists a Y-morphism π : V → X of finite type such that π ◦ j = 1X , and so π∗ (J /J 2 ) =
L;
(ii) there exists a homomorphism of OX -modules ϕ : L → lim π∗ (J /J n+1 ) such that the composi-
←−
tion
ϕ
L −→ lim π∗ (J /J n+1 ) − → π∗ (J /J 2 ) = L
α
←−
(where α is the canonical homomorphism) is the identity;
(iii) there exists a Y-prescheme C, a Y-section ε of C, and a Y-morphism q : V → C such that the
diagram
j
(8.9.4.1) X /V
p q
Y /C
ε
commutes; and
(iv) the restriction of q to W = V − j( X ) is a quasi-compact open immersion into C, whose image does
not intersect ε(Y ).
Then L is ample relative to p.
8.10. Grauert’s ampleness criterion: proof
Lemma (8.10.1). — Let π : V → X be a morphism, j : X → V an X-section of V that is also a closed
immersion, and J a quasi-coherent ideal of OV that defines the closed subprescheme of V associated to j.
Then the following all hold true.
(i) For all n ⩾ 0, π∗ (OV /J n+1 ) and π∗ (J /J n+1 ) are quasi-coherent OX -modules, and π∗ (OV /J ) =
OX and π∗ (J /J 2 ) = j∗ (J ).
(ii) If X = {ξ } = Spec(k), where k is a field, then lim π∗ (OV /J n+1 ) is isomorphic to the separated
←−
completion of the local ring O j(ξ ) for the m j(ξ ) -preadic topology.
(iii) Assume that J is an invertible OV -module (which implies that
L = j∗ (J ) = π∗ (J /J 2 )
is an invertible OX -module), and that there exists a homomorphism ϕ : L → lim π∗ (J /J n+1 )
←−
ϕ
such that the composition L − → lim π∗ (J /J n+1 ) − → π∗ (J /J 2 ) (where α is the canonical
α
←−
homomorphism) is the identity. If we write S = n⩾0 L ⊗n , then ϕ canonically induces an
L
isomorphism of OX -algebras from the completion S cof S relative to its canonical filtration (the
completion being isomorphic to the product ∏n⩾0 L ⊗n ) to lim π∗ (OV /J n+1 ).
←−
P ROOF. Note first of all that the support of the OV -module OV /J n+1 is j( X ), and the support of
J /J n+1 is contained in j( X ). In the case of (ii), j( X ) is a closed point j(ξ ) of V, and, by definition, II | 185
π∗ (OV /J n+1 ) is the fibre of OV /J n+1 at the point j(ξ ), or, equivalently, setting C = O j(ξ ) , and
denoting by m the maximal ideal of C, the C-module C/mn+1 ; claim (ii) is then evident.
To prove (i), note that the question is local on X; we can thus restrict to the case where X
is affine. Let U be an affine open subset of V; then j( X ) ∩ U is an affine open subset of j( X ),
8. BLOWUP SCHEMES; BASED CONES; PROJECTIVE CLOSURE 336
Lemma (8.10.2). — Under the hypotheses of Lemma (8.10.1), let g : X ′ → X be a morphism, write II | 186
V ′ = V × X X ′ , and let π ′ : V ′ → X ′ and g : V ′ → V be the canonical projections, so that we have the
commutative diagram
g′
Vo V′
π π′
Xo g
X′
′
Then j′ = j × 1X ′ is an X ′ -section of V ′ that is also a closed immersion, and J ′ = g ∗ (J )OV ′ is
the quasi-coherent ideal of OV ′ that defines the closed subprescheme of V ′ associated to j′ . Furthermore,
π∗′ (OV ′ /J ′ n + 1) = g∗ (π∗ (OV /J n+1 )). Finally, J ′ is an OV ′ -module that is canonically isomorphic
′
to g ∗ (J ), and is, in particular, invertible if J is an invertible OV -module.
P ROOF. The fact that j′ is a closed immersion follows from (I, 4.3.1), and it is an X ′ -section of
V′ by functoriality of extension of the base prescheme. Furthermore, if Z (resp. Z ′ ) is the closed
′
subprescheme of V (resp. V ′ ) associated to j (resp. j′ ), then Z ′ = g −1 ( Z ) (I, 4.3.1), and the second
claim then follows from (I, 4.4.5). To prove the other claims, we see, as in (8.10.1), that we can restrict
to the case where X, V, and X ′ (and thus also V ′ ) are affine; we keep the notation from the proof
of (8.10.1), and let X ′ = Spec( A′ ). Then V ′ = Spec( B′ ), where B′ = B ⊗ A A′ , and J ′ = J f′′ , where
′ ′ ′ ′ n +1 n + 1 ′
J = Im(J ⊗ A A ). Then B /J = ( B/J ) ⊗ A A ; furthermore, since J is a direct factor (as an
A-module) of B, J ⊗ A A′ is a direct factor (as an A′ -module) of B′ , and is thus canonically identified
with J′ . □
8. BLOWUP SCHEMES; BASED CONES; PROJECTIVE CLOSURE 337
Corollary (8.10.3). — Assume that the hypotheses of Lemma (8.10.1) are satisfied, and assume further that
π is of finite type, and that J is an invertible OV -module. Then, for all x ∈ X, the local ring at the point
j( x ) of the fibre π −1 ( x ) is a regular (thus integral) ring of dimension 1, whose completion is isomorphic to the
formal series ring k( x )[[ T ]] (where T is an indeterminate); furthermore, there exists exactly one irreducible
component of π −1 ( x ) that contains j( x ).
P ROOF. Since π −1 ( x ) = V × X Spec(k( x )), we are led, by (8.10.2), to the case where X is the
spectrum of a field K. Since π is of finite type (I, 6.4.3, iv), O j( x) is a Noetherian local ring, and
thus separated for the m j( x) -preadic topology (0, 7.3.5); it follows from (8.10.1, ii and iii) that the
completion of this ring is isomorphic to K [[ T ]], and so O j( x) is regular and of dimension 1 ([CC,
p. 17-01, th. 1]); finally, since O j( x) is integral, j( x ) belongs to exactly one of the (finitely many)
irreducible components of V (I, 5.1.4). □
Corollary (8.10.4). — Suppose that the hypotheses of Lemma (8.10.1) are satisfied, and assume further
that J is an invertible OV -module. Let W = V − j( X ); for every quasi-coherent ideal K of OX , let
KV = π ∗ (K )OV and KW = KV |W. Then KV is the largest quasi-coherent ideal of OV whose restriction
to W is KW .
P ROOF. Indeed, we see as in (8.10.1) that the question is local on X and V; we can thus reuse
the notation from the proof of (8.10.1), with J = Bt, where t is not a zero divisor in B. Furthermore,
we have W = Spec( Bt ) and K = K, e where K is an ideal of A; whence π ∗ (K )OV = (K.B)e (I, 1.6.9),
KW = (K.Bt )e, and the largest ideal of B whose canonical image in Bt is K.Bt is the inverse image of II | 187
K.Bt , that is, the set of s ∈ B such that, for some integer n > 0, we have tn s ∈ K.B. We have to show
that this last relation implies that s ∈ K.B, or again that the canonical image of t is not a zero divisor
in B/KB = ( A/K) ⊗ A B, which follows from (8.10.2) applied to X ′ = Spec( A/K). □
Corollary (8.10.5). — Suppose that the hypotheses of (8.10.3) are satisfied; let W = V − j( X ), x be a point
of X, K a quasi-coherent ideal of OX , and z the generic point of the irreducible component of π −1 ( x ) that
contains j( x ) (8.10.3).
(i) Let g be a section of OV over V such that g|W is a section of KW over W (using the notation from
(8.10.4)). Then g is a section of KV ; if further g(z) ̸= 0, and if, for every integer m > 0, we denote
by gm x the germ at the point x of the canonical image g of g in Γ ( X, π (O /J m+1 )), then there
m ∗ V
exists an integer m > 0 such that the image of gm x in
P ROOF.
(i) Since the ideal of OW generated by g|W is contained in KW by hypothesis, the ideal of
OV generated by g is contained in KV by (8.10.4), or, in other words, g is a section of KV .
To prove the second claim of (i), we can again assume that X and V are affine, and reuse
the notation from (8.10.1); the fibre π −1 ( x ) is then affine of ring B′ = B ⊗ A k( x ), and there
exists in B′ an element t′ which is not a zero divisor and is such that B′ = k( x ) ⊕ B′ t′ . Since
j( x ) is a specialisation of z and since g(z) ̸= 0, we necessarily have that g( j) x ̸= 0. But O j( x)
is a separated local ring (8.10.3), and thus embeds into its completion, and the image of g
′
in this completion is thus not null. But this completion is isomorphic to limn ( B′ /B′ t n+1 )
←−′
(8.10.3); if g′ = g ⊗ 1 ∈ B′ , there then exists an integer m such that g′ ̸∈ B′ t m+1 , or, again,
the image gm ′ of g′ in B′ /B′ t′ m+1 is not null. But since g′ is exactly the image of g x , our
m m
claim is proved.
(ii) By (8.10.1, iii), π∗ (OV /J m+1 ) is isomorphic to the direct sum of the L ⊗k for 0 ⩽ k ⩽
m; we denote by f k the section of L ⊗k over X that is the component of the element of
Lm ⊗k ) which corresponds to g by this isomorphism. Choosing m as in (i), there
k =0 Γ ( X, L m
is thus an index k such that f k ( x ) ̸= 0, by (i). To see that f k is a section of K L ⊗k , it suffices
8. BLOWUP SCHEMES; BASED CONES; PROJECTIVE CLOSURE 338
to consider, as above, the case where X and V are affine, and this follows immediately from
the fact that g ∈ K.B (with the notation from (8.10.4)). The final claim follows from the fact
that the hypothesis g ∈ Γ(V, J ) implies that f 0 = 0.
□
(8.10.6). Proof of (8.9.4). The question is local on Y (4.6.4); since ε is a Y-section, we can thus replace C
by an affine open neighbourhood U of a point of ε(Y ) such that ε(Y ) ∩ U is closed in U. In other
words, we can assume that C is affine, and that Y is a closed subprescheme of C (and thus also
affine) defined by a quasi-coherent sheaf I of ideals of OC . Since p is separated and quasi-compact, II | 188
X is thus a quasi-compact scheme, and we are reduced to proving that L is ample (4.6.4). By
criterion (4.5.2, a)), we must thus prove the following: for every quasi-coherent ideal K of OX and
every point x ∈ X not belonging to the support of OX /K , there exists an integer n > 0 and a section
f of K ⊗ L ⊗n over X such that f ( x ) ̸= 0.
For this, set
KV = π ∗ ( K ) OV
KW = KV |W
where W = V − j( X ); since the restriction of q to W is a quasi-compact immersion to C, it follows
from (I, 9.4.2) that KW is the restriction to W of a quasi-coherent ideal KV′ of OV of the form
KV′ = q∗ (KC )OV
where KC is a quasi-coherent ideal of OC . Furthermore, since, by hypotheses, q−1 (Y ) ⊂ j( X ), and
since Y is defined by the ideal I , the restriction to W of q∗ (I )OV is identical to that of OV , and so
KW is also the restriction to W of q∗ (I KC )OV , and we can thus suppose that KC ⊂ I , whence
(8.10.6.1) KV′ ⊂ q∗ (I )OV ⊂ J
taking into account (I, 4.4.6) and the commutativity of (8.9.4.1). Furthermore, we deduce from
(8.10.4) that
(8.10.6.2) KV′ ⊂ KV .
With this in mind, it follows from (8.10.3) that j( x ) belongs to exactly one irreducible component
of π −1 ( x ); let z be the generic point of this component, and let z′ = q(z). By (8.10.5), the proof
will be finished (taking (8.10.6.1) and (8.10.6.2) into account) if we show the existence of a section
g of KV′ over V such that g(z) ̸= 0. But, by hypothesis, K has a restriction equal to that of OX in
an open neighbourhood of x; also, it follows from (8.10.3) that z ̸= j( x ), and so z ∈ W, and thus
(KW )z = OV,z , whence, by definition, (KC )z′ = OC,z . Since C is affine, there is thus a section g′ of
KC over C such that g′ (z′ ) ̸= 0, and by taking g to be the section of KV′ corresponding canonically
to g′ , we indeed have g(z) ̸= 0, which finishes the proof.
Remark (8.10.7). — We ignore the question of whether or not condition (ii) in (8.9.4) is superfluous
or not. In any case, the conclusion does not hold if we do not assume the existence of a Y-morphism
π : V → X such that π ◦ j = 1X ; we briefly point out how we can indeed construct a counterexample,
whose details will not be developed until later on. We take Y = Spec(k), where k is a field,
and C = Spec( A), where A = k[ T1 , T2 ], and the Y-section ε corresponding to the augmentation
homomorphism A → k. We denote by C ′ the scheme induced by C by blowing up the closed point
a = ε(Y ) of C; if D is the inverse image of a in C ′ , we consider in D a closed point b, and we denote II | 189
by V the scheme induced by C ′ by blowing up b; X is the closed subprescheme of V given by the
inverse image of a by the structure morphism q : V → C. We now show that X is the union of two
irreducible components, X1 and X2 , where X1 is the inverse image of b in V. It is immediate that
the ideal J of OV that defines X is again invertible, we we can show that j∗ (J ) = L (where j is
the canonical injection X → V) is not ample, by considering the “degree” of the inverse image of
L in X1 , which would be > 0 if L were ample, but we can show (by an elementary intersection
calculation) that it is in fact equal to 0.
8. BLOWUP SCHEMES; BASED CONES; PROJECTIVE CLOSURE 339
that the diagram in (8.9.1.1) commutes. Suppose further that, if p = ( p0 , θ ), then θ : OY → p∗ (OX ) is an
isomorphism. Let S ′ = n⩾0 p∗ (L ⊗n ) and C ′ = Spec(S ′ ), and let q′ : V(L ) → C ′ be the canonical
L
Y-morphism (8.8.5). Then there exists exactly one Y-morphism u : C ′ → C such that q = u ◦ q′ .
P ROOF. The hypothesis on θ implies, in particular, that p is surjective; since, by (8.8.4), the
restriction of q′ to V(L ) − j( X ) is an isomorphism to C ′ − ε′ (Y ) (where ε is the vertex section of C ′ ), it
follows from (8.8.4) that q′ is proper and surjective; furthermore, by (8.8.6), if we let q′ = (q0′ , τ ), then
τ : OC′ → q′∗ (OV ) is an isomorphism. We are thus in a situation where we can apply (8.11.1), and
′
we will have proven the lemma if we show that q is constant on every fibre q −1 (z′ ), where z′ ∈ C ′ .
But this condition is trivially satisfied for z′ ̸∈ ε′ (Y ). If z′ ∈ ε′ (Y ), then there exists exactly one
y ∈ Y such that z′ = ε′ (y), and, by commutativity of (8.8.5.2) and the fact that q′ sends V(L ) − j( X )
′
to C ′ − ε′ (Y ), q −1 (z′ ) = j( p−1 (y)); the commutativity of the diagram in (8.9.1.1) then proves our
claim. □
Corollary (8.11.6). — Under the hypotheses of (8.11.5), suppose further that q is proper, and that the
restriction of q to V(L ) − j( X ) is an isomorphism to C − ε(Y ). Then the morphism u is universally closed,
surjective, and radicial, and its restriction to C ′ − ε′ (Y ) is an isomorphism to C − ε(Y ).
P ROOF. Since q′ is an isomorphism from V(L ) − j( X ) to C ′ − ε′ (Y ) (8.8.4), the last claim follows
immediately from the fact that q = u ◦ q′ . Furthermore, the commutativity of the diagrams in (8.8.5.2) II | 191
and (8.9.1.1) shows that the restriction of u to the closed subprescheme ε′ (Y ) of C ′ is an isomorphism
to the closed subprescheme ε(Y ) of C, from which we immediately deduce that, for all z′ ∈ ε′ (Y ),
if z = u(z′ ), then u defines an isomorphism from k(z) to k (z′ ). These remarks prove that u is
bijective and radicial; furthermore, if ψ : C → Y and ψ′ : C ′ → Y are the structure morphisms, then
ψ′ = ψ ◦ u, and, since ψ′ is separated (1.2.4), so too is u (I, 5.5.1, v). We have already seen, in the
proof of (8.11.5), that q′ is surjective; since q = u ◦ q′ is proper, we finally conclude, from (5.4.3) and
(5.4.9), that u is universally closed. □
Proposition (8.11.7). — Let Y be a prescheme, X an integral prescheme, p : X → Y a proper morphism, L
a p-ample invertible OX -module, C a normal Y-prescheme, ε : Y → C a Y-section, and q : V = V(L ) → C
a Y morphism, all such that the diagram in (8.9.1.1) commutes. Suppose further that, if p = ( p0 , θ ),
then θ : OY → p∗ (OX ) is an isomorphism. Let S ′ = n⩾0 p∗ (L ⊗n ) and C ′ = Spec(S ′ ), and let
L
q′ : V(L ) → C ′ be the canonical Y-morphism (8.8.5). Then the unique Y-morphism u : C ′ → C such that
q = u ◦ q′ is an isomorphism.
P ROOF. It follows from (8.8.6) that C ′ is integral; since u is a homeomorphism of the underlying
subspaces C ′ → C (u being bijective and closed, by (8.11.6)), C is irreducible, thus integral, and, since
the restriction of u to a non-empty open subset of C ′ is an isomorphism to an open subset of C, u is
birational. Since C is assumed to be normal, it suffices to apply (8.11.2) to obtain the conclusion. □
Remark (8.11.8). — (i) The hypothesis that C is normal implies that X is also normal. Indeed,
C ′ = Spec(S ′ ) is then normal, being isomorphic to C, and integral, by (8.8.6); we thus
conclude that Proj(S ′ ) is normal. Indeed, the question is local on Y; if Y is affine, with
S ′ = Se′ , then the ring S′ = Γ(C ′ , S ′ ) is integral and integrally closed (8.8.6.1), and so, for
every homogeneous element f ∈ S+ ′ , the graded ring S′ is integral and integrally closed
f
[SZ60, t. I, p. 257 and 261], and thus so too is the ring S(′ f ) of its degree-zero terms, because
the intersection of S′f with the field of fractions of S(′ f ) is equal to S(′ f ) ; this proves our
claim (6.3.4). Finally, since X is isomorphic to an open subprescheme of Proj(S ′ ) (8.8.1),
X is indeed normal. We can thus express (8.11.7) in the following form: If X is integral
and normal, and p = ( p0 , θ ) : X → Y is a proper morphism such that θ : OY → p∗ (OX ) is an
isomorphism, then, for every p-ample OX -module L , there exists exactly one way of contracting the
null section of V = V(L ) to obtain a normal Y-scheme C and a proper Y-morphism q : V → C.
(ii) When p is proper, the hypothesis p∗ (OX ) = OY can be considered as an auxiliary hypoth-
esis, not really restricting the generality of the result. Indeed, if it is not satisfied, then
it suffices to replace Y with the Y-scheme Y ′ = Spec( p∗ (OX )), and to consider X as a
Y ′ -scheme. We will return to this general method in chapter III, § 4.
8. BLOWUP SCHEMES; BASED CONES; PROJECTIVE CLOSURE 341
the quasi-coherent graded OX -algebra SX being defined by (8.6.1.1), Proj(M ) is equipped with a
structure of a (quasi-coherent) graded SX -module, by means of the canonical homomorphisms (3.2.6.1)
(8.12.1.2) OX (m) ⊗OX Proj0 (M (n)) −→ Proj0 (S (m) ⊗S M (n)) −→ Proj0 (M (m + n)),
the verification of the axioms of sheaves of modules being done using the commutative diagram in
(2.5.11.4).
If Y = Spec( A) is affine, S = S,
e and M = M,e where S is a graded A-algebra and M is a graded
S-module, then, for every homogeneous element f ∈ S+ , we have
(8.12.1.3) Γ( X f , Proj( M
e )) = M f
(8.12.1.4) Mc= M ⊗S S
c
(S
cbeing defined by (8.3.1.1)); this induces a quasi-coherent graded O b-module Proj (M
C 0
c), which
we will also denote by
(8.12.1.5) M □ = Proj0 (Mc).
It is clear (3.2.4) that M □ is an additive functor which is exact in M , commuting with direct
sums and with inductive limits.
Proposition (8.12.2). — With the notation of (8.3.2), we have canonical functorial isomorphisms
∼ ∼
(8.12.2.1) i ∗ (M □ ) −
→ Mf, j ∗ (M □ ) −
→ Proj0 (M ).
Indeed, i∗ (M □ ) is canonically identified with (Mc/(z − 1)Mc)e on Spec(S c/(z − 1)S c) by (3.2.3); the first
of the canonical isomorphisms (8.12.2.1) is then immediately induced (1.4.1) by the canonical isomorphism
∼
M c/(z − 1)M c−→ M . The canonical immersion j : X → C corresponds to the canonical homomorphism
Sc→ S with kernel zS c(8.3.2); the second homomorphism (8.12.2.1) is the particular case of the canonical
homomorphism (3.5.2, ii), since here we have M c⊗ cS = M ; to verify that this is an isomorphism, we can
S
restrict to the case where Y = Spec( A) is affine, S = S, e and M = M; e by appealing to (2.8.8), the proof
that, for all homogeneous f in S+ , the preceding homomorphism, restricted to X f , restricts to an isomorphism,
is then immediate.
By an abuse of language, we again say, thanks to the existence of the first isomorphism (8.12.2.1), II | 193
that M □ is the projective closure of the OX -module Mf(it being implicit that the data of the OC -module
Mfincludes the grading of the S -module M ).
Indeed, this is a particular case of the homomorphism ν♯ defined more generally in (3.5.6). If
Y = Spec( A) is affine, S = S,e and M = M, e then, by appealing to (2.8.8), the restriction of (8.12.3.1)
− 1
to p ( X f ) = C f (for some homogeneous f in S+ ) corresponds to the canonical homomorphism
b
(8.12.3.2) M( f ) ⊗ S ( f ) S ⩽ ⩽
f −→ M f
(8.12.4). Let us place ourselves in the settings of (8.5.1), and assume its hypotheses and keep its
notation. It follows from (1.5.6) that, for every quasi-coherent graded S -module S , we have, on
one hand, a canonical isomorphism
∼
(8.12.4.1) Φ ∗ (M → (q∗ (M ) ⊗q∗ (S ) S ′ )e
f) −
of OC′ -modules; on the other hand, (3.5.6) implies the existence of a canonical Proj(ϕ)-morphism
(8.12.4.2) Proj0 M −→ (Proj0 (q∗ (M )) ⊗q∗ (S ) S ′ )| G (ϕ)
(8.12.5). Consider now the setting of (8.6.1), with the same notation; we thus take Y ′ = X, the
morphism q : X → Y being the structure morphism, and ϕ the canonical q-morphism (8.6.1.2). We
then have a canonical isomorphism
∼
(8.12.5.1) q∗ (M ) ⊗q∗ (S ) SX⩾ −
→ MX⩾
by setting MX⩾ = n⩾0 Proj0 (M ( n )). We can indeed restrict to the case where Y = Spec( A )
L
is affine, S = S,e and M = M, e and define the isomorphism (8.12.5.1) on each of the affine
open subsets X f (where f is homogeneous in S+ ), by verifying the compatibility with taking
a homogeneous multiple of f . But the restriction to X f of the left-hand side of (8.12.5.1) is
Me ′ = (( M ⊗ A S( f ) ) ⊗S⊗ S S⩾ )e by (8.6.2.1); since we have a canonical isomorphism from
A (f) f
M ⊗ A S( f ) to M ⊗S (S ⊗ A S( f ) ), we have an induced isomorphism from M e ′ to ( M ⊗S S⩾ )e, and
f
the latter is canonically isomorphic, by (8.2.9.1), to the restriction to X f of the right-hand side of
(8.12.5.1), and satisfies the required compatibility conditions. II | 194
Replacing M by M c, S by S c, and SX by (S ⩾ )∧ in the previous argument, we similarly have
X
a canonical isomorphism
c) ⊗ ∗ c (S ⩾ )∧ − ∼
(8.12.5.2) q ∗ (M q (S ) X → (MX⩾ )∧ .
(8.12.5.4) M □ −→ (MX⩾ )□ .
Recall now (8.6.2) that the restrictions of r to the pointed cones E
bX and EX are isomorphisms to E
b
and E (respectively). Furthermore:
is an isomorphism; but, by (8.2.3.2) and (8.2.5.2), the left-hand side is canonically identified with
M⩽ ⩾ ⩽ ⩽ ⩾ ⩽
f ⊗S⩽ ( S f ) f /1 , and thus with M f , by (8.2.7.2), and the right-hand side with ( M f ) f /1 , and thus
f
also with M⩽ f , by (8.2.9.2), whence the conclusion concerning (8.12.6.1); (8.12.6.2) then follows from
(8.12.6.1) and (8.12.2.1). □
Corollary (8.12.7). — With the identifications of (8.6.3), the restriction of (MX⩾ )□ to E
bX can be identified
⩽ ⩾ □
with (MX )e, and the restriction of (MX ) to Ex with MX . f
P ROOF. We can restrict to the affine case, and this follows from the identification of ( M⩾ ⩽
f ) f /1
with M⩽ ⩾
f , and of ( M f ) f /1 with M f (8.2.9.2). □
Proposition (8.12.8). — Under the hypotheses of (8.6.4), the canonical homomorphism (8.12.3.1) is an
isomorphism.
P ROOF. Taking into account the fact that Proj(SX⩾ ) → X is an isomorphism (8.6.2), and the
isomorphisms (8.12.5.4) and (8.12.6.1), we are led to proving the corresponding proposition for the II | 195
canonical homomorphism p∗X (Proj0 (MX⩾ )) → (MX⩾ )□ | EX , or, in other words, we can restrict to
the case where S1 is an invertible OY -module, and where S is generated by S1 . With the notation
of (8.12.3), we then have, for some f ∈ S1 , that S⩽
f = S( f ) [1/ f ], and the canonical homomorphism
M( f ) ⊗ S ( f ) S ⩽ ⩽ ⩽
f → M f is an isomorphism, by the definition of M f . □
m⩾n
to the section ∑i si ∈ Γ(U, M ). Denote by Nn′ the inverse image of N by this homomorphism, which
is a quasi-coherent sub-S -module of i⩽n Mi . Now consider the homomorphism i⩽n Mi →
L L
M c = M [z] which sends (si ) to the section ∑i⩽n si zn−i ∈ Γ(U, M cn ), and let Nn be the image of
8. BLOWUP SCHEMES; BASED CONES; PROJECTIVE CLOSURE 344
Proposition (8.13.3). — The OCb-module Proj0 (N ) is the canonical extension (Nf)− of Nfto C.
b
P ROOF. The question is local on Y and C b by the definition of the canonical extension (I, 9.4.1).
We can thus already suppose that Y = Spec( A) is affine, with S = S, e M = M, e and N = N, e
where N is a non-necessarily-graded sub-S-module of M. Furthermore (8.3.2.6), C is a union of
b
affine opens C bz = C and C b f = Spec(S⩽ ) (with f homogeneous in S+ ). It thus suffices to show
f
that: (1) the restriction of Proj0 (N ) to C is Nf; (2) the restriction of Proj0 (N ) to each Cb f is the
canonical extension of the restriction of N to C ∩ C f = Spec(S f ) (8.3.2.6). For the first point, note
b
that Proj0 (N )|C can be identified with ( N (z) )e (8.3.2.4); but N (z) is canonically identified (2.2.5)
b (z − 1) M,
with the image of N in M/ b and by the canonical isomorphism of the latter with M (8.2.5),
this image can be identified with N, by the definition of N given in (8.13.2).
To prove the second point, note that the injection i : C ∩ C bf → C b corresponds to the canonical
⩽ ⩽
injection S → S f (8.3.2.6); we also have that Γ(C
f
b f , M □ ) = M , that Γ(C b f , i∗ (Nf)) = N, and, by
f
(8.12.2.1), that Γ(C b f , i∗ (i∗ (M □ ))) = M f . Taking (I, 9.4.2) into account, we are thus led to showing
b ( f ) = M⩽ is canonically identified with the inverse image of N f under the canonical
that N ( f ) ⊂ M f
injection M⩽
f → M f . Indeed, let d = deg( f ) > 0, and suppose that an element (∑k⩽md xk )/ f m of M f
(with xk ∈ Mk ) is of the form y/ f m with y ∈ N. By multiplying y and the xk by one single suitable
f h , we can already assume that ∑k⩽md xk = y. But in the identification of (8.2.5.2), (∑k⩽md xk )/ f m
corresponds to ∑k⩽md xk zmd−k / f m , and this is indeed an element of N ( f ) , since ∑k⩽md xk ∈ N; the
converse is evident. □
Remark (8.13.4). — (i) The most important case of application of (8.13.3) is that where M =
S , with Nf then being an arbitrary quasi-coherent sheaf of ideals J of OC (1.4.3), cor-
responding bijectively to a closed subprescheme Z of C. Then the canonical extension J
of J is the quasi-coherent sheaf of ideals of OCb that defines the closure Z of Z in C b
(I, 9.5.10); Proposition (8.13.3) gives a canonical way of defining Z by using a graded ideal
in S c= S [z].
(ii) Suppose, to simplify things, that Y is affine, and adopt the notation from the proof of
(8.13.3). For every non-zero x ∈ N, let d( x ) be the largest degree of the homogeneous
components xi of x in M; by definition, N is the submodule of M b consisting of 0 and
k
elements of the form h( x, k ) = z ∑i⩽d( x) xi z d ( x )− i (for integral k ⩾ 0); it is thus generated,
as a module over Sb = S[z], by the elements of the form
h( x, 0) = ∑ xi zd( x)−i .
i⩽d( x )
We say that h( x, 0) is induced from x by homogenisation via the “homogenising variable” z. II | 197
But since h( x, 0) does not depend additively on x (nor a fortiori S-linearly), we will refrain
from believing (even when M = S) that the h( x, 0) form a system of generators of the graded
S-module
b N when we let x run over a system of generators of the S-module N. This is,
however, the case (considered only in elementary algebraic geometry) when N is a free
cyclic S-module, since, if t is a basis of N, then h(t, 0) generates the S-module
b N.
8. BLOWUP SCHEMES; BASED CONES; PROJECTIVE CLOSURE 345
P ROOF. Since q(U ) is quasi-compact, it is covered by a finite number of affine opens Vi , and so
every x ∈ U is contained in some affine open of the form D+ ( f ), where f is a homogeneous element
of degree > 0 of one of the rings Γ(Vi , S ). Since U is quasi-compact, we can cover it by a finite
number of such opens D+ ( f j ); let d be a common multiple of the degrees of the f j . This d satisfies
the desired property, by (2.5.17). □
(8.14.5). With the hypotheses and notation of (8.14.1), we defined, in (3.3.2), canonical homomor-
phisms of OY -modules
(8.14.5.1) αn : Mn −→ q∗ (Proj0 (M (n))) ( n ∈ Z ).
Generalising the notation of (3.3.1), we set, for every graded SX -module F ,
Γ ∗ (F ) = q ∗ (Fn ).
M
(8.14.5.2)
n∈Z
In particular, Γ(SX ) = n∈Z q∗ (OX (n)) is the graded OY -algebra denoted by Γ∗ (OX ) in (3.3.1.2);
L
it is clear that Γ(F ) is a graded Γ∗ (SX )-algebra (0, 4.2.2). When we take M = S in the homomor- II | 199
phisms (8.14.5.1), we obtain the homomorphism of graded OY -algebras
(8.14.5.3) α : S −→ Γ(SX )
previously defined in (3.3.2), and which makes Γ∗ (F ) a graded S -module; the homomorphisms
(8.14.5.1) then define a homomorphism (of degree 0) of graded S -modules
(8.14.5.4) α : M −→ Γ∗ (Proj(M )).
(8.14.6). In general, for a quasi-coherent graded SX -module F , it is not certain that the graded S -
module Γ∗ (F ) will necessarily be quasi-coherent. Consider an open X ′ of X such that the restriction
q′ : X ′ → Y of q to X ′ is a quasi-compact morphism. Since q′ is further separated, q′∗ (F ′ ) is then a
quasi-coherent OY -module for every quasi-coherent OX ′ module F ′ (I, 9.2.2, b). We set
SX ′ = SX | X ′ = OX (n)| X ′
M
(8.14.6.1)
n∈Z
The previous remark then shows that, if F′ is a quasi-coherent SX ′ -module, then Γ′∗ (F ′ ) is a
graded quasi-coherent S -module (I, 9.6.1).
We note also that the canonical injection j : X ′ → X is quasi-compact, because q′ = q ◦ j is quasi-
compact and q is separated (I, 6.6.4, v). Then F = j∗ (F ′ ) is a quasi-coherent graded SX -module
for every quasi-coherent graded SX ′ -module F ’, and it follows from the previous definitions that
(8.14.6.3) Γ′∗ (F ′ ) = Γ∗ (F ).
With the same hypotheses on X ′ , for every quasi-coherent graded S -module M , we set
(8.14.6.4) Proj′ (M ) = Proj(M )| X ′
which is a quasi-coherent graded SX ′ -module. The canonical homomorphism
Proj(M ) −→ j∗ (Proj′ (M ))
(0, 4.4.3) thus gives a canonical homomorphism Γ∗ (Proj(M )) → Γ′∗ (Proj′ (M )) of graded S -
modules, and, by composition with (8.14.5.4), we obtain a functorial canonical homomorphism (of
degree 0) of quasi-coherent graded S -modules
(8.14.6.5) α′ : M −→ Γ′∗ (Proj′ (M )).
(8.14.7). Keeping the hypotheses on X ′ from (8.14.6), let F ′ be a quasi-coherent graded SX ′ -module
such that Proj′ (Γ′∗ (F ′ )) is also a graded quasi-coherent SX ′ -module. We will define a functorial II | 200
canonical homomorphism (of degree 0) of graded SX ′ -modules
(8.14.7.1) β′ : Proj′ (Γ′∗ (F ′ )) −→ F ′ .
8. BLOWUP SCHEMES; BASED CONES; PROJECTIVE CLOSURE 347
Suppose first of all that Y = Spec( A) is affine, and that S = S, e where S is a positively-graded
′ ′
A-algebra; then Γ∗ (F ) = M, where M =
e L ′ ′
n ∈ ZΓ( X , Fn ) is a graded S-module. Let f ∈ Sd be
such that D+ ( f ) ⊂ X ′ ; by definition (2.6.2), αd ( f ) restricted to D+ ( f ) is the section of OX (d) over
D+ ( f ) corresponding to the element f /1 of (S(d))( f ) , and is thus invertible; thus so too is αd ( f n )
for every n > 0. From this, we immediately conclude that we have defined an S f -homomorphism
(of degree 0) of graded modules β f : M f → Γ( D+ ( f ), F ′ ) by sending each element z/ f n ∈ M f
(where z ∈ M) to the section (z| D+ ( f ))(αd ( f n )| D+ ( f ))−1 of F ′ over D+ ( f ). Furthermore, we have
a commutative diagram corresponding to (2.6.4.1), whence the definition of β′ in this case. To pass
to the general case, we must consider an A-algebra A′ , the graded A′ -algebra S′ = S ⊗ A A′ , and use
the commutative diagram analogous to (2.8.13.2); we leave the details to the reader.
Proposition (8.14.8). — If X ′ is an open of X = Proj(S ) such that q′ : X ′ → Y is quasi-compact, then the
homomorphism β′ defined in (8.14.7) is bijective.
P ROOF. We can clearly restrict to the case where Y is affine, and everything then reduces to
proving (with the notation of (8.14.7)) that the homomorphism β f : M f → Γ( D+ ( f ), F ′ ) is an
isomorphism. But replacing f by one of its powers changes neither D+ ( f ) nor β f ; since X ′ is
quasi-compact by hypothesis, we can always assume, by (8.14.4), that the sheaf OX (d) is invertible.
Since X ′ is a scheme (because q′ is separated), the proposition is then exactly (I, 9.3.1). □
Corollary (8.14.9). — Under the hypotheses of (8.14.8), every quasi-coherent graded SX ′ -module is
isomorphic to a graded SX ′ -module of the form Proj′ (M ), where M is a quasi-coherent graded S -module.
Further, if F ′ is of finite type, and if we assume that Y is a quasi-compact scheme, or a prescheme whose
underlying space is Noetherian, then we can assume that M is of finite type.
P ROOF. The proof starting from (8.14.8) follows exactly the same route as the proof of (3.4.5)
starting from (3.4.4), and we leave the details to the reader. □
Proposition (8.14.10). — Under the hypotheses of (8.14.7), let M be a quasi-coherent graded S -module,
and F ′ a quasi-coherent graded SX ′ -module; the composite homomorphisms
Proj′ (α′ ) β′
(8.14.10.1) Proj′ (M ) −−−−−→ Proj′ (Γ′∗ (Proj′ (M ))) −
→ Proj′ (M )
α′ Γ′∗ ( β′ )
(8.14.10.2) → Γ′∗ (Proj′ (Γ′∗ (F ′ ))) −−−→ Γ′∗ (F ′ )
Γ′∗ (F ′ ) −
are the identity isomorphisms.
P ROOF. The question is local on Y, and the proof follows as in (2.6.5); we leave the details to the
reader. □
Remark (8.14.11). — In chapter III (III, 2.3.1), we will see that, when Y is locally Noetherian, and
S is a quasi-coherent graded OY -algebra of finite type (in which case we can take X ′ = X), then II | 201
the homomorphism α (8.14.5.4) is (TN)-bijective for every quasi-coherent graded S -module M
satisfying condition (TF).
Remark (8.14.12). — The situation described in (8.14.4) is a particular case of the following. Let
X be a ringed space, and S a (positively- and negatively-) graded OX -algebra; suppose that there
exists an integer d > 0 such that Sd and S−d are invertible, with the canonical homomorphism
(8.14.12.1) Sd ⊗OX S−d −→ OX
being an isomorphism (such that S−d is identified with Sd−1 ). We then say that the graded OX -algebra
S is periodic, of period d. This nomenclature stems from the following property: under the preceding
hypotheses, for every graded S -module F , the canonical homomorphism
(8.14.12.1) Sd ⊗ Fn −→ Fn+d
is an isomorphism for all n ∈ Z. Indeed, the question is local on X, and we can assume that Sd
has an invertible section s over X, with its inverse s′ being a section of S−d . The homomorphism
Fn+d → Sd ⊗ Fn , which sends each section z ∈ Γ(U, Fn+d ) to the section (s|U ) ⊗ (s′ |U )z of
8. BLOWUP SCHEMES; BASED CONES; PROJECTIVE CLOSURE 348
Sd ⊗ Fn over U, is then the inverse of (8.14.12.2), whence our claim. This induces, for all k ∈ Z, a
canonical isomorphism
∼
(Sd ) ⊗ k ⊗ Fn −
→ Fn+kd .
Then the data of a graded S -module F is equivalent to the data of S0 -modules Fi (0 ⩽ i ⩽ d − 1) and
canonical homomorphisms
Si ⊗ F j −→ Fi+ j for 0 ⩽ i, j ⩽ d − 1
(setting Fi+ j = Sd ⊗S0 Fi+ j−d whenever i + j ⩾ d). Of course, for theses homomorphisms to give a
well-defined S -module structure on the direct sum of the (Sd )⊗k ⊗ Fi (k ∈ Z, 0 ⩽ i ⩽ d − 1), they
should satisfy some associativity conditions that we will not explain.
In the case where d = 1 (which is the one considered in (3.3)), we can thus say that the category
of graded S -modules (resp. quasi-coherent S -modules if X is a prescheme and S is quasi-coherent)
is equivalent to the category of arbitrary S0 -modules (resp. quasi-coherent S0 -modules); it is in this
way that we can think of the results of this paragraph as generalising those of §3. Furthermore, we
see that, under suitable finiteness conditions, the results of this paragraph (along with (8.14.11))
reduces, in some sense, the study of quasi-coherent graded algebras on a prescheme, and graded
modules “modulo (TN)” on such algebras, to the study of the particular case where the algebras in
question are periodic (and where condition (TN) for M (3.4.2) thus implies that M = 0).
Remark (8.14.13). — Under the hypotheses of (8.14.1), let d be an integer > 0; we have defined a
canonical Y-isomorphism h from X to X (d) = Proj(S (d) ) (3.1.8). For every quasi-coherent graded II | 202
S -module M and every integer k such that 0 ⩽ k ⩽ d − 1, we also have (with the notation of (3.1.1))
a canonical h-isomorphism
∼
(8.14.13.1) (Proj(M ))(d,k) ←
− Proj(M (d,k) ).
Suppose, first of all, that Y = Spec( A) is affine, S = S,
e and M = M, e where S is a positively-
graded A-algebra, and M a graded S-module. We know, for every f ∈ Se (e > 0), that h sends D+ ( f )
to D+ ( f d ), and corresponds to the canonical isomorphism S( f d ) → S( f ) (2.2.2). The restriction of
(8.14.13.1) to D+ ( f d ) then corresponds to the canonical di-isomorphism M f d → M f restricted to
the elements of M f d whose degree is congruent to k (modulo d). We leave to the reader the task
of showing that these isomorphisms are compatible with passing from f to some homogeneous
multiple f g, and then that there is an analogous compatibility with passing from S to a graded
A′ -algebra S′ = S ⊗ A A′ , where A′ is some A-algebra. In particular, this gives us an h-isomorphism
∼
(8.14.13.2) (S ( d ) ) X (d ) −
→ (S X ) ( d )
that respects the multiplicative structures of both the source and the target, and that, thanks to
(8.14.13.1), becomes an h-di-isomorphism from a graded (S (d) ) X (d) -module to a graded (SX )(d) -
module. Similarly, we have an h-isomorphism
∼
(8.14.13.3) Proj0 (S (d,k) (n)) −
→ OX (nd + k),
which completes the result of (3.2.9, ii).
The isomorphism in (8.14.13.1) immediately induces an isomorphism of graded S (d) -modules
(d) ∼
(8.14.13.4) Γ∗ (Proj(M (d,k) )) −
→ Γ∗ ((Proj(M ))(d,k) )
(d)
where Γ∗ corresponds to the structure morphism q(d) : X (d) → Y; it can be immediately verified
that the canonical homomorphism α (8.14.5.4), and the analogous homomorphism α(d) for X (d) ,
make the following diagram commute:
(8.14.13.5) M (d,k)
α(d) α
w '
(d)
Γ∗ (Proj(M (d,k) ))
∼ / Γ∗ ((Proj(M ))(d,k) )
where we proceed by supposing that Y is affine and then calculating the restrictions of the images
under α(d) and α of some single element of M(d,k) to the open subsets D+ ( f d ) and D+ ( f ) (using the
same notation as above).
8. BLOWUP SCHEMES; BASED CONES; PROJECTIVE CLOSURE 349
S UMMARY
§1. Cohomology of affine schemes.
§2. Cohomological study of projective morphisms.
§3. Finiteness theorem for proper morphisms.
§4. The fundamental theorem of proper morphisms. Applications.
§5. An existence theorem for coherent algebraic sheaves.
§6. Local and global Tor functors; Künneth formula.
§7. Base change for homological functors of sheaves of modules.
§8. The duality theorem for projective bundles
§9. Relative cohomology and local cohomology; local duality
§10. Relations between projective cohomology and local cohomology. Formal completion technique along a divisor
§11. Global and local Picard groups1
This chapter gives the fundamental theorems concerning the cohomology of coherent algebraic
sheaves, with the exception of theorems explaining the theory of residues (duality theorems), which
will be the subject of a later chapter. Amongst all those included here, there are essentially six
fundamental theorems, and each one is the subject of one of the first six chapters. These results will
prove to be essential tools in all that follows, even in questions which are not truly cohomological
in their nature; the reader will see the first such examples starting from §4. §7 gives some more
technical results, but ones which are constantly used in applications. Finally, in §§8–11, we will
develop certain results, related to the duality of coherent sheaves, that are particularly important for
applications, and which can be explained even before the introduction of the full general theory of
residues.
The content of §§1 and 2 is due to J.-P. Serre, and the reader will observe that we have had
only to follow (FAC). §8 and 9 are equally inspired by (FAC) (the changes necessitated by the
different contexts, however, being less evident). Finally, as we said in the Introduction, §4 should be
considered as the formalisation, in modern language, of the fundamental “invariance theorem” of
Zariski’s “theory of holomorphic functions”.
We draw attention to the fact that the results of no 3.4 (and the preliminary propositions of
(0, 13.4 to 13.7)) will not be used in what follows Chapter III, and can thus be skipped in a first
reading.
350
1. COHOMOLOGY OF AFFINE SCHEMES 351
antiderivation of degree −1 of ∧( Ar ), and if (ei )1⩽i⩽r is the canonical basis of Ar , then we have
if (ei ) = f i ; the verification of the condition if ◦ if = 0 is immediate.
An equivalent definition is the following: for each i, we consider a chain complex K• ( f i ) defined
as follows: K0 ( f i ) = K1 ( f i ) = A, Kn ( f i ) = 0 for n ̸= 0, 1: the boundary map is defined by the
condition that d1 : A → A is multiplication by f i . We then take K• (f) to be the tensor product
K• ( f 1 ) ⊗ K• ( f 2 ) ⊗ · · · ⊗ K• ( f r ) (G, I, 2.7) with its total degree; the verification of the isomorphism
from this complex to the complex defined above is immediate.
(1.1.2). For every A-module M, we define the chain complex
(1.1.2.1) K• (f, M ) = K• (f) ⊗ A M
and the cochain complex (G, I, 2.2)
(1.1.2.2) K • (f, M) = Hom A (K• (f, M).
If g is a k-cochain of this latter complex, and if we set
g ( i 1 , . . . , i k ) = g ( e i1 ∧ · · · ∧ e i k ),
then g identifies with an alternating map from [1, r ]k to M, and it follows from the above definitions
that we have
k +1
(1.1.2.3) d k g ( i 1 , i 2 , . . . , i k +1 ) = ∑ (−1)h−1 fih g(i1 , . . . , ibh , . . . , ik+1 ).
h =1
(1.1.3). From the above complexes, we deduce as usual the homology and cohomology A-modules (G, I, III | 83
2.2)
(1.1.3.1) H• (f, M) = H• (K• (f, M)),
Lemma (1.1.4.1). — Let K• be a chain complex of free A-modules, zero except in dimensions 0 and 1. For
every chain complex L• of A-modules, we have an exact sequence
0 −→ H0 (K• ⊗ H p ( L• )) −→ H p (K• ⊗ L• ) −→ H1 (K• ⊗ H p−1 ( L• )) −→ 0
for every index p.
This is a particular case of an exact sequence of low-order terms of the Künneth spectral sequence III | 84
(M, XVII, 5.2 (a) and G, I, 5.5.2); it can be proved directly as follows. Consider K0 and K1 as chain
complexes (zero in dimensions ̸= 0 and ̸= 1 respectively); we then have an exact sequence of
complexes
0 −→ K0 ⊗ L• −→ K• ⊗ L• −→ K1 ⊗ L• −→ 0,
to which we can apply the exact sequence in homology
∂ ∂
· · · −→ H p+1 (K1 ⊗ L• ) −
→ H p (K0 ⊗ L• ) −→ H p (K• ⊗ L• ) −→ H p (K1 ⊗ L• ) −
→ H p−1 (K0 ⊗ L• ) −→ · · · .
But it is evident that H p (K0 ⊗ L• ) = K0 ⊗ H p ( L• ) and H p (K1 ⊗ L• ) = K1 ⊗ H p−1 ( L• ) for all p; in
addition, we verify immediately that the operator ∂ : K1 ⊗ H p ( L• ) → K0 ⊗ H p ( L• ) is none other than
d1 ⊗ 1; the lemma thus follows from the above exact sequence and the definition of H0 (K• ⊗ H p ( L• ))
and H1 (K• ⊗ H p−1 ( L• )).
The lemma having been established, the end of the proof of Proposition (1.1.4) is immediate: the
induction hypothesis of Lemma (1.1.4.1) gives H p (K• ⊗ L• ) = 0 for p ⩾ 2; in addition if we show
that H1 (K• , H0 ( L• )) = 0, then we also deduce from Lemma (1.1.4.1) that H1 (K• ⊗ L• ) = 0; but by
definition, H1 (K• , H0 ( L• )) is none other than the kernel of the scaling z 7→ f r · z on Mr−1 , and as by
hypothesis this kernel is zero, this finishes the proof.
(1.1.5). Let g = ( gi )1⩽i⩽r be a second sequence of r elements of A, and set fg = ( f i gi )1⩽i⩽r . We can
define a canonical homomorphism of complexes
(1.1.5.1) ϕg : K• (fg) −→ K• (f)
as the canonical extension to the exterior algebra ∧( Ar ) of the A-linear map ( x1 , . . . , xr ) 7→ ( g1 x1 , . . . , gr xr )
from Ar to itself. To see that we have a homomorphism of complexes, it suffices to note, in general,
that if u : E → F is an A-linear map, and if x ∈ F ∨ and y = t u(x) ∈ E∨ , then we have the formula
(1.1.5.2) (∧u) ◦ iy = ix ◦ (∧u);
indeed, the two elements are antiderivations of ∧ F, and it suffices to check that they coincide on F,
which follows immediately from the definitions.
When we identify K• (f) with the tensor product of the K• ( f i ) (1.1.1), ϕg is the tensor product of
the ϕgi , where ϕgi is the identity in degree 0 and multiplication by gi in degree 1.
(1.1.6). In particular, for every pair of integers m and n such that 0 ⩽ n ⩽ m, we have homomor-
phisms of complexes
(1.1.6.1) ϕfm−n : K• (fm ) −→ K• (fn )
and as a result, homomorphisms
(1.1.6.2) ϕfm−n : K • (fn , M) −→ K • (fm , M ),
Remark (1.1.10). — With the same notations as above, set X = Spec( A) and Y the closed sub-
prescheme of X defined by the ideal (f). We will prove in §9 that H• ((f), M ) is isomorphic to the
cohomology HY• ( X, M
e ) corresponding to the antifilter Φ of closed subsets of Y (T, 3.2). We will also
show that Proposition (1.2.3) applied to X and to F = M e is a particular case of an exact sequence in
cohomology
p p +1
· · · −→ HY ( X, F ) −→ H p ( X, F ) −→ H p ( X − Y, F ) −→ HY ( X, F ) −→ · · · .
1.2. Čech cohomology of an open cover
Notation (1.2.1). — In this section, we denote:
(1) X a prescheme;
(2) F a quasi-coherent OX -module; III | 86
(3) A = Γ( X, OX ), M = Γ( X, F );
(4) f = ( f i )1⩽i⩽r a finite system of elements of A;
(5) Ui = X f i , the open set (0, 5.5.2) of the x ∈ X such that f i ( x ) ̸= 0;
U = ri=1 Ui ;
S
(6)
(7) U the cover (Ui )1⩽i⩽r of U.
(1.2.2). Suppose that X is either a prescheme whose underlying space is Noetherian or a scheme whose
underlying space is quasi-compact. We then know (I, 9.3.3) that we have Γ(Ui , F ) = M f i . We set
p
\
Ui0 i1 ···i p = Uik = X f i f ··· f i p
0 i1
k =0
(n)
We have (0, 1.6.1) that M f i identifies with the inductive limit limn Mi i ···i p , where the
f ··· f i p
0 i1 −→ 0 1
(n) (m) (n)
inductive system is formed by the Mi i ···i p = M, the homomorphisms ϕnm : Mi i ···i p → Mi i ···i p
0 1 0 1 0 1
p
being multiplication by ( f i0 f i1 · · · f i p )n−m for m ⩽ n. We denote by Cn ( M ) the set of alternating maps
from [1, r ] p+1 to M (for all n); these A-modules also form an inductive system with respect to the
ϕnm . If C p (U, F ) is the group of alternating Čech p-cochains relative to the cover U, with coefficients
in F (G, II, 5.1), then it follows from the above that we can write
p
(1.2.2.2) C p (U, F ) = lim Cn ( M ).
−→n
1. COHOMOLOGY OF AFFINE SCHEMES 354
p
With the notations of (1.1.2), Cn ( M ) identifies with K p+1 (fn , M), and the map ϕnm identifies with
the map ϕfn−m defined in (1.1.6). We thus have, for every p ⩾ 0, a canonical functorial isomorphism
(1.2.2.3) C p (U, F ) ≃ C p+1 ((f), M ).
In addition, the formula (1.1.2.3) and the definition of the cohomology of a cover (G, II, 5.1)
shows that the isomorphisms (1.2.2.3) are compatible with the coboundary maps.
Proposition (1.2.3). — If X is a prescheme whose underlying space in Noetherian or a scheme whose
underlying space is quasi-compact, then there exists a canonical functorial isomorphism in F
(1.2.3.1) H p (U, F ) ≃ H p+1 ((f), M ) for p ⩾ 1.
In addition, we have a functorial exact sequence in F
(1.2.3.2) 0 −→ H0 ((f), M ) −→ M −→ H0 (U, F ) −→ H1 ((f), M ) −→ 0.
III | 87
P ROOF. The isomorphisms (1.2.3.1) are immediate consequences of what we saw in (1.2.2). On
the other hand, we have C0 (U, F ) = C1 ((f), M ); as a result, H0 (U, F ) identifies with the subgroup
of 1-cocycles of C1 ((f), M); as M = C0 ((f), M), the exact sequence (1.2.3.2) is none other than the
one given by the definition of the cohomology groups H0 ((f), M ) and H1 ((f), M ). □
Corollary (1.2.4). — Suppose that the X f i are quasi-compact and that there exists gi ∈ Γ(U, F ) such that
∑i gi ( f i |U ) = 1|U. Then for every quasi-coherent (OX |U )-module G , we have H p (U, G ) = 0 for p > 0; if
in addition U = X, then the canonical homomorphism (1.2.3.2) M → H0 (U, F ) is bijective.
P ROOF. As by hypothesis the Ui = X f i are quasi-compact, so is U, and we can reduce to the case
where U = X; the hypothesis then implies that H p ((f), M) = 0 for all p ⩾ 0 (1.1.9). The corollary
then follows immediately from (1.2.3.1) and (1.2.3.2). □
We note that since H0 (U, F ) = H0 (U, F ) (G, II, 5.2.2), we have again proved (I, 1.3.7) as a
special case.
Remark (1.2.5). — Suppose that X is an affine scheme; then the Ui = X f i = D ( f i ) are affine open
sets, as well as the Ui0 i1 ···i p (but U is not necessarily affine). In this case, the functors Γ( X, F ) and
Γ(Ui0 i1 ···i p , F ) are exact in F (I, 1.3.11). If we have an exact sequence 0 → F ′ → F → F ′′ → 0 of
quasi-coherent OX -modules, then the sequence of complexes
0 −→ C • (U, F ′ ) −→ C • (U, F ) −→ C • (U, F ′′ ) −→ 0
is exact, and thus gives an exact sequence in cohomology
0 / C • ((f), M′ ) / C • ((f), M) / C • ((f), M′′ ) /0
(1.2.5.1) H p (U, F ′′ )
∂ / H p+1 (U, F ′ )
H p+1 ((f), M′′ )
∂ / H p+2 ((f), M′ )
1. COHOMOLOGY OF AFFINE SCHEMES 355
(1.4.4.1) Hq (U, F ′′ )
∂ / Hq+1 (U, F ′ )
Hq+1 ((f), M′′ )
∂ / Hq+2 ((f), M′ )
P ROOF.
(i) If X f is affine, then we have Hi ( X f , F ) = 0 for all i > 0 (1.3.1), so the assertion follows
directly from Corollary (1.4.6).
(ii) By virtue of Serre’s criterion (II, 5.2.1), it suffices to prove that for every quasi-coherent
sheaf of ideals K of OX | X f , we have H1 ( X f , K ) = 0. As X f is a quasi-compact open set in
a quasi-compact scheme X, there exists a quasi-coherent sheaf of ideals J of OX such that
K = J | X f (I, 9.4.2). According to Corollary (1.4.6), we have H1 ( X f , K ) = (H1 ( X, J )) f ,
and the hypothesis implies that the right hand side is zero, hence the assertion.
□
Remark (1.4.8). — We note that Corollary (1.4.7, i) gives a simpler proof of the relation (II, 4.5.13.2).
1. COHOMOLOGY OF AFFINE SCHEMES 357
Lemma (1.4.9). — Let X be a quasi-compact scheme, U = (Ui )1⩽i⩽n a finite cover of X by affine open sets, III | 91
and F a quasi-coherent OX -module. The complex of sheaves C • (U, F ) defined by the cover U (G, II, 5.2) is
then a quasi-coherent OX -module.
P ROOF. It follows from the definitions (G, II, 5.2) that C p (U, F ) is the direct sum of the direct
image sheaves of the F |Ui0 ···i p under the canonical injection Ui0 ···i p → X. The hypothesis that X
is a scheme implies that these injections are affine morphisms (I, 5.5.6), hence the C p (U, F ) are
quasi-coherent (II, 1.2.6). □
Proposition (1.4.10). — Let u : X → Y be a separated and quasi-compact morphism. For every quasi-
coherent OX -module F , the Rq u∗ (F ) are quasi-coherent OY -modules.
P ROOF. The question is local on Y, so we can suppose that Y is affine. Then X is a finite union
of affine open sets Ui (1 ⩽ i ⩽ n); let U be the cover (Ui ). In addition, as Y is a scheme, it follows
from (I, 5.5.10) that for every affine open V ⊂ Y, the canonical injection u−1 (V ) → X is an affine
morphism; we conclude (Proposition (1.4.1) and (G, II, 5.2)) that we have a canonical isomorphism
(1.4.10.1) H• (u−1 (V ), F ) ≃ H• (Γ(V, K • )),
where we set K • = u∗ (C • (U, F )). According to Lemma (1.4.1) and (I, 9.2.2), K • is a quasi-coherent
OY -module; moreover, it constitutes a complex of sheaves since so is C • (U, F ). It then follows from
the definition of the cohomology H • (K • ) (G, II, 4.1) that the latter consists of quasi-coherent
OY -modules (I, 4.1.1). As (for V affine in Y) the functor Γ(V, G ) is exact in G on the category of
quasi-coherent OY -modules, we have (G, II, 4.1)
(1.4.10.2) H• (Γ(V, K • )) = Γ(V, H • (K • )).
Finally, we note that it follows from the definition of the canonical homomorphism
H• (U, F ) −→ H• ( X, F ),
given in (G, II, 5.2), that if V ′ ⊂ V is a second affine open subset of Y, then the diagram
H• ( u −1 ( V ), F )
∼ / H• (Γ(V, K • ))
H• ( u −1 ( V ′ ), F )
∼ / H• (Γ(V ′ , K • ))
is commutative. We thus conclude from the above that the isomorphisms (1.4.10.1) define an
isomorphism of OY -modules
(1.4.10.3) R• u ∗ (F ) ≃ H • (K • ),
and as a result, R• u∗ (F ) is quasi-coherent. □ III | 92
Corollary (1.4.13). — Suppose that the hypotheses of Proposition (1.4.10) are satisfied, and in addition
suppose that Y = Spec( A) is affine. Then for every quasi-coherent OX -module F and every f ∈ A, we have
Γ(Y f , Rq u∗ (F )) = (Γ(Y, Rq u∗ (F )) f
up to canonical isomorphism.
P ROOF. This follows from the fact that Rq u∗ (F ) is a quasi-coherent OY -module (I, 1.3.7). □
Proposition (1.4.14). — Let f : X → Y be a separated and quasi-compact morphism, g : Y → Z an
affine morphism. For every quasi-coherent OX -module F , the canonical homomorphism R p ( g ◦ f )∗ (F ) →
g∗ (R p f ∗ (F )) (0, 12.2.5.2) is bijective for all p.
P ROOF. For every affine open subset W of Z, g−1 (W ) is an affine open subset of Y. The
homomorphism of presheaves defining the canonical homomorphism
R p ( g ◦ f )∗ (F ) −→ g∗ (R p f ∗ (F ))
(0, 12.2.5) is thus bijective by Corollary (1.4.11). □
Proposition (1.4.15). — Let u : X → Y be a separated morphism of finite type, v : Y ′ → Y a flat morphism
of preschemes (0, 6.7.1); let u′ = u(Y ′ ) , such that we have the commutative diagram
v′
(1.4.15.1) Xo X ′ = X (Y ′ )
u u′
Yo
v
Y′ .
Then for every quasi-coherent OX -module F , Rq u′∗ (F ′ ) is canonically isomorphic to Rq u∗ (F ) ⊗OY OY ′ =
∗
v∗ (Rq u∗ (F )) for all q ⩾ 0, where F ′ = v′ (F ) = F ⊗OY OY ′ .
III | 93
∗
P ROOF. The canonical homomorphism ρ : F → v′∗ (v′ (F )) (0, 4.4.3.2) defines by functoriality
a homomorphism
(1.4.15.2) Rq u∗ (F ) −→ Rq u∗ (v′∗ (F ′ )).
On the other hand, we have, by setting w = u ◦ v′ = v ◦ u′ , the canonical homomorphisms
(0 ,12.2.5.1 and 12.2.5.2)
(1.4.15.3) Rq u∗ (v′∗ (F ′ )) −→ Rq w∗ (F ′ ) −→ v∗ (Rq u′∗ (F ′ )).
Composing (1.4.15.3) and (1.4.15.2), we have a homomorphism
ψ : Rq u∗ (F ) −→ v∗ (Rq u′∗ (F ′ )),
and finally we obtain a canonical homomorphism (whose definition does not make any assumptions
on v)
(1.4.15.4) ψ♯ : v∗ (Rq u∗ (F )) −→ Rq u′∗ (F ′ ),
and it is necessary to prove that it is an isomorphism when v is flat. It is clear that the question is
local on Y and Y ′ , and we can therefore suppose that Y = Spec( A) and Y ′ = Spec( B); we will also
use the following lemma:
Lemma (1.4.15.5). — Let ϕ : A → B be a ring homomorphism, Y = Spec( A), X = Spec( B), f : X → Y
the morphism corresponding to ϕ, and M a B-module. For the OX -module M e to be f -flat (0, 6.7.1), it is
necessary and sufficient for M to be a flat A-module. In particular, for the morphism f to be flat, it is necessary
and sufficient for B to be a flat A-module.
This follows from the definition (0, 6.7.1) and from (0, 6.3.3), taking into account (I, 1.3.4).
This being so, it follows from (1.4.11.1) and the definitions of the homomorphisms (1.4.15.3)
(cf. (0, 12.2.5)) that ψ then corresponds to the composite morphism
ρq ∗ θq ∗ ∗ σq ∗
→ Hq ( X, v′∗ (v′ (F ))) −
Hq ( X, F ) − → Hq ( X ′ , v′ (v′∗ (v′ (F )))) −
→ Hq ( X ′ , v′ (F )),
1. COHOMOLOGY OF AFFINE SCHEMES 359
where ρq and σq are the homomorphisms in cohomology corresponding to the canonical morphisms
∗
ρ and σ : v′ (v′∗ (G ′ )) → G ′ , and θq is the ϕ-morphism (0, 12.1.3.1) relative to the OX -module
′ ′ ∗
v∗ (v (F )). But by the functoriality of θq , we have the commutative diagram
ρq
Hq ( X, F ) / Hq ( X, v′∗ (v′ ∗ (F )))
θq θq
∗
v′ (ρq )
∗
Hq ( X ′ , v′ (F )) / Hq ( X ′ , v′ ∗ (v′∗ (v′ ∗ (F )))),
∗
and as by definition (0, 4.4.3) v′ (ρ) is the inverse of σ, we see that the composite morphism III | 94
considered above is finally none other than θq ; as a result, ψ♯ is the associated B-homomorphism
Hq ( X, F ) ⊗ A B → Hq ( X ′ , F ′ ). As u is of finite type, X is a finite union of affine open sets Ui
(1 ⩽ i ⩽ r); let U be the cover (Ui ). As v is an affine morphism, so is v′ (II, 1.6.2, iii), and as a result
−1
the Ui′ = v′ (Ui ) form an affine open cover U′ of X ′ . We then know (0, 12.1.4.2) that the diagram
θq
Hq (U, F ) / Hq (U′ , F ′ )
θq
Hq ( X, F ) / Hq ( X ′ , F ′ )
is commutative, and the vertical arrows are isomorphisms since X and X ′ are schemes (I, 1.4.1). As
a result, it suffices to prove that the canonical ϕ-morphism θq : Hq (U, F ) → Hq (U′ , F ′ ) is such that
the associated B-homomorphism
Hq (U, F ) ⊗ A B −→ Hq (U′ , F ′ )
Tp
is an isomorphism. For every sequence s = (ik )0⩽k⩽p of p + 1 indices of [1, r ], set Us = k =0 Uik ,
Tp −1
Us′ = k=0 Ui′ = v′ (Us ), Ms = Γ(Us , F ), and ′Ms′′
= Γ(Us , F ). The canonical map Ms ⊗ A B →
k
Ms′ is an isomorphism (I, 1.6.5), hence the canonical map C p (U, F ) ⊗ A B → C p (U′ , F ′ ) is an
isomorphism, by which d ⊗ 1 identifies with the coboundary map C p (U′ , F ′ ) → C p+1 (U′ , F ′ ). As B
is a flat A-module, it follows from the definition of the cohomology modules that the canonical map
Hq (U, F ) ⊗ A B → Hq (U′ , F ′ ) is an isomorphism (0, 6.1.1). This result will later be generalized in
§6. □
Corollary (1.4.16). — Let A be a ring, X an A-scheme of finite type, and B an A-algebra which is faithfully
flat over A. For X to be affine, it is necessary and sufficient for X ⊗ A B to be.
P ROOF. The condition is evidently necessary (I, 3.2.2); we show that it is sufficient. As X is
separated over A and the morphism Spec( B) → Spec( A) is flat, it follows from Proposition (1.4.1)
that we have
(1.4.16.1) Hi ( X ⊗ A B, F ⊗ A B) = Hi ( X, F ) ⊗ A B
for every i ⩾ 0 and every quasi-coherent OX -module F . If X ⊗ A B is affine, the left hand side of
(1.4.16.1) is zero for i = 1, hence so is H1 ( X, F ) since B is a faithfully flat A-module. As X is a
quasi-compact scheme, we finish the proof by Serre’s criterion (II, 5.2.1). □
u v
Proposition (1.4.17). — Let X be a prescheme, 0 → F −
→G −
→ H → 0 an exact sequence of OX -modules.
If F and H are quasi-coherent, then so is G .
III | 95
P ROOF. The question is local on X, so we can suppose that X = Spec( A) is affine, and it
then suffices to prove that G satisfies the conditions (d1) and (d2) of (I, 1.4.1) (with V = X). The
verification of (d2) is immediate, because if t ∈ Γ( X, G ) is zero when restricted to D ( f ), then so is
its image v(t) ∈ Γ( X, H ); therefore there exists an m > 0 such that f m v(t) = v( f m t) = 0 (I, 1.4.1),
and as Γ is left exact, f m t = u(s), where s ∈ Γ( X, F ); as u is injective, the restriction of s to D ( f )
3. FINITENESS THEOREM FOR PROPER MORPHISMS 360
is zero, hence (I, 1.4.1) there exists an integer n > 0 such that f n s = 0; we finally deduce that
f m+n t = u( f n s) = 0.
We now check (d1); let t′ ∈ Γ( D ( f ), G ); as H is quasi-coherent, there exists an integer m such
that f m v(t′ ) = v( f m t′ ) extends to a section z ∈ Γ( X, H ) (I, 1.4.1). But in virtue of Theorem (1.3.1)
(or (I, 5.1.9.2)) applied to the quasi-coherent OX -module F , the sequence Γ( X, G ) → Γ( X, H ) → 0
is exact, so there exists t ∈ Γ( X, G ) such that z = v(t); we thus see that v( f m t′ − t′′ ) = 0, denoting
by t′′ the restriction of t to D ( f ); thus we have f m t′ − t′′ = u(s′ ), where s′ ∈ Γ( D ( f ), F ). But as F
is quasi-coherent, there exists an integer n > 0 such that f n s′ extends to a section s ∈ Γ( X, F ); as
f m+n t′ − f n t′′ = u( f n s′ ), we see that f m+n t′ is the restriction to D ( f ) of a section f n t + u( f n s) ∈
Γ( X, G ), which finishes the proof. □
0 −→ Ker u −→ F −→ Im u −→ 0,
and the hypothesis that C′ is exact.
(b) Y is irreducible, and as a result, the subprescheme Y of X is integral. If y is its generic point,
then we have (OY )y = k (y), and as j∗ (F ) is a coherent OY -module, Fy = ( j∗ (F ))y is a
k (y)-vector space of finite dimension m. By hypothesis, there is a coherent OX -module
G ∈ C′ (necessarily of support Y) such that Gy is a k(y)-vector space of dimension 1. As
a result, there is a k(y)-isomorphism (Gy )m ≃ Fy , which is also an OY -isomorphism,
and as G m and F are coherent, there exists an open neighbourhood W of y in X and an
isomorphism G m |W ≃ F |W (0, 5.2.7). Let H be the graph of this isomorphism, which is
a coherent (OX |W )-submodule of (G m ⊕ F )|W, canonically isomorphic to G m |W and to
F |W; there thus exists a coherent OX -submodule H0 of G m ⊕ F , inducing H on W and 0
on X − Y, since G m and F have Y for their support (I, 9.4.7). The restrictions v : H0 → G m
and w : H0 → F of the canonical projections of G m ⊕ F are then homomorphisms of
coherent OX -modules, which, on W and on X − Y, reduce to isomorphisms; in other words,
the kernels and cokernels of v and w have their support in the closed set Y − (Y ∩ W ),
distinct from Y. They are in C′ ; on the other hand, we have G m ∈ C′ since G ∈ C′ and since
C′ is exact. We conclude successively, by the exactness of C′ , that H0 ∈ C′ , then F ∈ C′ .
Q.E.D.
□
Corollary (3.1.3). — Suppose that the exact subset C′ of C has in addition the property that any coherent
direct factor of a coherent OX -module M ∈ C′ is also in C′ . In this case, the conclusion of Theorem (3.1.2)
is still valid when the condition “Gy is a k(y)-vector space of dimension 1” is replaced by Gy ̸= 0 (this is
equivalent to Supp(G ) = Y).
P ROOF. The reasoning of Theorem (3.1.2) must be modified only in the case (b); now Gy is a
k(y)-vector space of dimension q > 0, and as a result, we have an OY -isomorphism (Gy )m ≃ (Fy )q ;
the end of the reasoning in Theorem (3.1.2) then proves that F q ∈ C′ , and the additional hypothesis
on C′ implies that F ∈ C′ . □
Rq−1 f ∗ (F ′′ ) −
→ Rq f ∗ (F ′ ) −→ Rq f ∗ (F ) −→ Rq f ∗ (F ′′ ) −
→ Rq +1 f ∗ (F ′ ),
∂ ∂
in which by hypothesis the outer four terms are coherent; it is the same for the middle term Rq f ∗ (F )
by ((0, 5.3.4) and (0, 5.3.3)). We show in the same way that when F and F ′ (resp. F and F ′′ )
are in C′ , then so is F ′′ (resp. F ′ ). In addition, every coherent direct factor F ′ of an OX -module
F ∈ C′ belongs to C′ : indeed, Rq f ∗ (F ′ ) is then a direct factor of Rq f ∗ (F ) (G, II, 4.4.4), therefore it
is of finite type, and as it is quasi-coherent (1.4.10), it is coherent, as Y is Noetherian. By virtue of
Corollary (3.1.3), we reduce to proving that when X is irreducible with generic point x, there exists one
coherent OX -module F belonging to C′ , such that Fx ̸= 0: indeed, if this point is established, then it
can be applied to any irreducible closed subprescheme Y of X, since if j : Y → X is the canonical
injection, then f ◦ j is proper (II, 5.4.2), and if G is a coherent OY -module with support Y, then j∗ (G )
is a coherent OX -module such that Rq ( f ◦ j)∗ (G ) = Rq f ∗ ( j∗ (G )) (G, II, 4.9.1), therefore we can apply
Corollary (3.1.3).
By virtue of Chow’s lemma (II, 5.6.2), there exists an irreducible prescheme X ′ an a projective
and surjective morphism g : X ′ → X such that f ◦ g : X ′ → Y is projective. There exists an ample
OX -module L for g (II, 5.3.1); we apply the fundamental theorem of projective morphisms (2.2.1) to
3. FINITENESS THEOREM FOR PROPER MORPHISMS 362
g : X ′ → X and with L : there thus exists an integer n such that F = g∗ (OX ′ (n)) is a coherent OX -
module and Rq g∗ (OX ′ (n)) = 0 for all q > 0; in addition, as g∗ ( g∗ (OX ′ (n))) → OX ′ (n) is surjective
for n large enough (2.2.1), we see that we can suppose, at the generic point x of X, that we have
Fx ̸= 0 (II, 3.4.7). On the other hand, as f ◦ g is projective as Y is Noetherian, the Rq ( f ◦ g)∗ (OX ′ (n))
are coherent (2.2.1). This being so, R• ( f ◦ g)∗ (OX ′ (n)) is the abutment of a Leray spectral sequence,
pq
whose E2 -term is given by E2 = R p f ∗ (Rq g∗ (OX ′ (n))); the above shows that this spectral sequence
p0
degenerates, and we then know (0, 11.1.6) that E2 = R p f ∗ (F ) is isomorphic to R p ( f ◦ g)∗ (OX ′ (n)),
which finishes the proof. □
Corollary (3.2.2). — Let Y be a locally Noetherian prescheme. For every proper morphism f : X → Y, the
direct image under f of any coherent OX -module is a coherent OY -module.
Corollary (3.2.3). — Let A be a Noetherian ring, X a proper scheme over A; for every coherent OX -module
F , the H p ( X, F ) are A-modules of finite type, and there exists an integer r > 0 such that for every coherent
OX -module F and all p > r, H p ( X, F ) = 0.
P ROOF. The second assertion has already been proved (1.4.12); the first follows from the finite-
ness theorem (3.2.1), taking into account Corollary (1.4.11). □
In particular, if X is a proper algebraic scheme over a field k, then, for every coherent OX -module
F , the H p ( X, F ) are finite-dimensional k-vector spaces.
Corollary (3.2.4). — Let Y be a locally Noetherian prescheme, f : X → Y a morphism of finite type. For
every coherent OX -module F whose support in proper over Y (II, 5.4.10), the OY -modules Rq f ∗ (F ) are
coherent.
III | 118
P ROOF. Since the questions is local on Y, we can suppose Y Noetherian, and it is the same for
X (I, 6.3.7). By hypothesis, every closed subprescheme Z of X whose underlying space is Supp(F )
is proper over Y, in other words, if j : Z → X is the canonical injection, then f ◦ j : Z → Y is proper.
We can suppose that Z is such that F = j∗ (G ), where G = j∗ (F ) is a coherent OZ -module (I, 9.3.5);
as we have Rq f ∗ (F ) = Rq ( f ◦ j)∗ (G ) by Corollary (1.3.4), the conclusion follows immediately from
Theorem (3.2.1). □
the R p f ∗ (M ) = k∈Z R p f ∗ (Mk ) are graded S -modules of finite type for all p. Suppose in addition that
L
the S are generated by S1 ; then, for every p ∈ Z, there exists an integer k p such that for all k ⩾ k p and all
r > 0, we have
(3.3.1.1) R p f ∗ ( M k + r ) = Sr R p f ∗ ( M k ) .
P ROOF. The first assertion is identical to the statement of Theorem (2.4.1, i), where we have
simply replaced “projective morphism” by “proper morphism”. In the proof of Theorem (2.4.1, i),
the hypothesis on f was only used to show (with the notation of this proof) that R p f ∗′ (M f) is a
′
coherent OY ′ -module. With the hypothesis of Proposition (3.3.1), f is proper (II, 5.4.2, iii), so we can
resume without change in the proof of Theorem (2.4.1, i), thanks to the finiteness theorem (3.2.1).
As for the second assertion, it suffices to remark that there is a finite affine open cover (Ui ) of Y
such that the restrictions to the Ui of the two sides of (3.3.1.1) are equal for all k ⩾ k p,i (II, 2.1.6, ii); it
suffices to take for k p the largest of the k p,i . □
Corollary (3.3.2). — Let A be a Noetherian ring, m an ideal of A, X a proper A-scheme, and F a coherent
OX -module. Then, for all p ⩾ 0, the direct sum k⩾0 H p ( X, mk F ) is a module of finite type over the ring
L
S = k⩾0 mk ; in particular, there exists an integer k p ⩾ 0 such that for all k ⩾ k p and all r > 0, we have
L
(3.3.2.1) H p ( X, mk+r F ) = mr H p ( X, mk F ).
3. FINITENESS THEOREM FOR PROPER MORPHISMS 363
considering, for every a ∈ mr , the map H p ( X, mk F ) → H p ( X, mk+r F ), which comes from the
passage to cohomology of the multiplication map mr F → mk+r F defined by a (2.4.1).
3.4. Finiteness theorem: the case of formal schemes The results of this section (except the III | 119
definition (3.4.1)) will not be used in the rest of this chapter.
(3.4.1). Let X and S be two locally Noetherian formal preschemes (I, 10.4.2), f : X → S a morphism
of formal preschemes. We say that f is a proper morphism if it satisfies the following conditions:
1st. f is a morphism of finite type (I, 10.13.3).
2nd. If K is a sheaf of ideals of definition for S and if we set J = f ∗ (K )OX , X0 = (X, OX /J ),
S0 = (S, OS /K ), then the morphism f 0 : X0 → S0 induced by f (I, 10.5.6) is proper.
It is immediate that this definition does not depend on the sheaf of ideals of definition K for S
considered; indeed, if K ′ is a second sheaf of ideals of definition such that K ′ ⊂ K , and if we
set J ′ = f ∗ (K ′ )OX , X0′ = (X, OX /J ′ ), S0′ = (S , OS /K ′ ), then the morphism f 0′ : X0′ → S0′
induced by f is such that the diagram
f0
X0 / S0
i j
f 0′
X0′ / S′
0
is commutative, i and j being surjective immersions; it is equivalent to say that f 0 or f 0′ is proper, by
virtue of (II, 5.4.5).
We note that, for all n ⩾ 0, if we set Xn = (X, OX /J n+1 ), Sn = (S, OS /K n+1 ), then the
morphism f n : Xn → Sn induced by f (I, 10.5.6) is proper for all n whenever it is for n = 0 (II, 5.4.6).
If g : Y → Z is a proper morphism of locally Noetherian usual preschemes, Z ′ a closed subset
of Z, Y ′ a closed subset of Y such that g(Y ′ ) ⊂ Z ′ , then the extension gb : Y/Y ′ → Z/Z′ of g to the
completions (I, 10.9.1) is a proper morphism of formal preschemes, as it follows from the definition
and from (II, 5.4.5).
Let X and S be two locally Noetherian formal preschemes, f : X → S a morphism of finite
type (I, 10.13.3); the notation being the same as above, we say that a subset Z of the underlying space
of X is proper over S (or proper for f ) if, considered as a subset of X0 , Z is proper over S0 (II, 5.4.10).
All the properties of proper subsets of usual preschemes stated in (II, 5.4.10) are still true for the
proper subsets of formal preschemes, as it follows immediately from the definitions.
Theorem (3.4.2). — Let X and Y be locally Noetherian formal preschemes, f : X → Y a proper morphism.
For every coherent OX -module F , the OY -modules Rq f ∗ (F ) are coherent for all q ⩾ 0.
Let J be a sheaf of ideals of definition for Y, K = f ∗ (J )OX , and consider the OX -modules
(3.4.2.1) Fk = F ⊗OY (OY /J k+1 ) = F /K k +1
F ( k ⩾ 0)
which evidently form a projective system of topological OX -modules, such that F = limk Fk (I, 10.11.3). III | 120
←−
On the other hand, it follows from Theorem (3.4.2) that each of the Rq f ∗ (F ), being coherent, is
naturally equipped with a topological OY -module structure (I, 10.11.6), and so are the Rq f ∗ (Fk ). The
canonical homomorphisms F → Fk = F /K k+1 F canonically correspond to homomorphisms
Rq f ∗ (F ) −→ Rq f ∗ (Fk ),
which are necessarily continuous for the topological OY -module structures above (I, 10.11.6), and
form a projective system, giving the limit a canonical functorial homomorphism
(3.4.2.2) Rq f ∗ (F ) −→ lim Rq f ∗ (Fk ),
←−
k
which will be a continuous homomorphism of topological OY -modules. We will prove along with
Theorem (3.4.2) the
3. FINITENESS THEOREM FOR PROPER MORPHISMS 364
We know that J∆ is a sheaf of ideals of definition for Y (I, 10.3.1); let K = f ∗ (J∆ )OX , X0 =
(X, OX /K ), Y0 = (Y, OY /J∆ ) = Spec( A0 ), with A0 = A/J. It is clear that the Mk = Jk F /Jk+1 F
are coherent OX0 -modules (I, 10.11.3). Consider on the other hand the quasi-coherent graded
OX0 -algebra
S = OX0 ⊗ A0 S = gr(OX ) = K k /K k +1
M
(3.4.4.4) .
k⩾0
The hypothesis that F is an OX -module of finite type implies first that M is a graded S -module III | 121
of finite type. Indeed, the question is local on X, and we can thus suppose that X = Spf( B), where
B is an adic Noetherian ring, and F = N ∆ , where N is a B-module of finite type (I, 10.10.5);
we have in addition X0 = Spec( B0 ), where B0 = B/JB, and the quasi-coherent OX0 -modules
S and M are respectively equal to Se′ and M f′ , where S′ = Lk⩾0 ((Jk /Jk+1 ) ⊗ A B0 ) and M′ =
0
k k +1 ) ⊗ ′ ′
A0 N0 ), with N0 = N/JN; we then evidently have M = S ⊗ B0 N0 , and as N0 is
L
k⩾0 ((J /J
a B0 -module of finite type, M′ is a S′ -module of finite type, hence our assertion (I, 1.3.13).
As the morphism f 0 : X0 → Y0 is proper by hypothesis, we can apply Corollary (3.3.2) to S ,
M , and the morphism f 0 : taking into account Corollary (1.4.11), we conclude that for all n ⩾ 0,
k⩾0 H ( X0 , Mk ) is a graded S-module of finite type. This proves that the condition (Fn ) of (0, 13.7.7)
L n
is satisfied for all n ⩾ 0, when we consider the strictly projective system (F /Jk F )k⩾0 of sheaves of
abelian groups on X0 , each equipped with its natural “filtered A-module” structure. We can thus
apply (0, 13.7.7), which proves that:
1st. The projective system (Hn (X, Fk ))k⩾0 satisfies the (ML)-condition.
2nd. If H′n = limk Hn (X, Fk ), then H′n is an A-module of finite type.
←−
3rd. The filtration defined on H′n by the kernels of the canonical homomorphisms H′n →
Hn (X, Fk ) is J-good.
Note that on the other hand, if we set Xk = (X, OX /K k+1 ), then Fk is a coherent OXk -
module (I, 10.11.3), and if U is an affine open set in X0 , then U is also an affine open set in each
of the Xk (I, 5.1.9), so Hn (U, Fk ) = 0 for all n > 0 and all k (1.3.1) and H0 (U, Fk ) → H0 (U, Fh ) is
surjective for h ⩽ k (I, 1.3.9). We are thus in the conditions of (0, 13.3.2) and applying (0, 13.3.1)
proves that H′n canonically identifies with Hn (X, limk Fk ) = Hn (X, F ); this finishes the proof of
←−
Corollary (3.4.4).
(3.4.5). We return to the proof of (3.4.2) and (3.4.3). We first prove the propositions for the case
Y = Spf( A) envisaged in (3.4.4); for this, for all g ∈ A, apply (3.4.4) to the Noetherian affine formal
7. BASE CHANGE FOR HOMOLOGICAL FUNCTORS OF SHEAVES OF MODULES 365
scheme induced on the open set Yg = D( g) of Y, which is equal to Spf( A{ g} ), and to the formal
prescheme induced by X on f −1 (Yg ); note that Yg is also an affine open set in the prescheme
Yk = (Y, OY /(J∆ )k+1 ), and as Fk is a coherent OXk -module, we have
Hn ( f −1 (Yg ), Fk ) = Γ(Yg , Rn f ∗ (Fk ))
for all k ⩾ 0 by virtue of Corollary (1.4.11). The canonical homomorphism
Hn ( f −1 (Yg ), F ) −→ lim Γ(Yg , Rn f ∗ (Fk ))
←−
k
is an isomorphism; but we have (0, 3.2.6)
lim Γ(Yg , Rn f ∗ (Fk )) = Γ(Yg , lim Rn f ∗ (Fk )),
←− ←−
k k
n −1 (Y ), F ) on the
and as the sheaf Rn f ∗ (F ) is the sheaf associated to the presheaf Y g 7 → H ( f g III | 122
Yg (0, 3.2.1), we have shown that the homomorphism (3.4.2.2) is bijective. Let us now prove that
Rn f ∗ (F ) is a coherent OY -module, and more precisely that we have
∆
(3.4.5.1) Rn f ∗ (F ) = Hn (X, F ) .
With the above notation, we have, since Fk is a coherent OXk -module (1.4.13),
Γ(Yg , Rn f ∗ (Fk )) = (Γ(Y, Rn f ∗ (Fk ))) g = (Hn (X, Fk )) g .
Now the Hn (X, Fk ) form a projective system satisfying (ML), and their projective limit Hn (X, F )
is an A-module of finite type. We conclude (0, 13.7.8) that we have
lim (Hn (X, Fk )) g = Hn (X, F ) ⊗ A A{ g} = Γ(Yg , (Hn (X, F ))∆ ),
←−
k
taking into account (I, 10.10.8) applied to A and A{ g} ; this proves (3.4.5.1) since Γ(Yg , Rn f ∗ (F )) =
limk Γ(Yg , Rn f ∗ (Fk )).
←−
As (3.4.2.2) is then an isomorphism of coherent OY -modules, it is necessarily a topological
isomorphism (I, 10.11.6). Finally, it follows from the relations Rn f ∗ (Fk ) = (Hn (X, Fk ))∆ that the
projective system (Rn f ∗ (Fk ))k⩾0 satisfies (ML) (I, 10.10.2).
Once (3.4.2) and (3.4.3) are proved in the case where the formal prescheme Y is affine Noetherian,
it is immediate to pass to the general case for (3.4.2) and the first assertion of (3.4.3), which are local
on Y. As for the second assertion of (3.4.3), it suffices, Y being Noetherian, to cover it by a finite
number of Noetherian affine open sets Ui and to note that the restrictions of the projective system
(Rq f ∗ (Fk )) to each of the Ui satisfies (ML).
Along the way, we have in addition proved:
Corollary (3.4.6). — Under the hypotheses of Corollary (3.4.4), the canonical homomorphism
(3.4.6.1) Hq (X, F ) −→ Γ(Y, Rq f ∗ (F ))
is bijective.
S UMMARY
IV-1 | 222
§1. Relative finiteness conditions. Constructible sets in preschemes.
§2. Base change and flatness.
§3. Associated prime cycles and primary decomposition.
§4. Change of base field for algebraic preschemes.
§5. Dimension, depth, and regularity for locally Noetherian preschemes.
§6. Flat morphisms of locally Noetherian preschemes.
§7. Relations between a local Noetherian ring and its completion. Excellent rings.
§8. Projective limits of preschemes.
§9. Constructible properties.
§10. Jacobson preschemes.
§11.1 Topological properties of finitely presented flat morphisms. Flatness criteria.
§12. Study of fibres of finitely presented flat morphisms.
§13. Equidimensional morphisms.
§14. Universally open morphisms.
§15. Study of fibres of a universally open morphism.
§16. Differential invariants. Differentially smooth morphisms.
§17. Smooth morphisms, unramified (or net) morphisms, and étale morphisms.
§18. Supplement on étale morphisms. Henselian local rings and strictly local rings.
§19. Regular immersions and normal flatness.
§20. Meromorphic functions and pseudo-morphisms
§21. Divisors.
The subjects discussed in the chapter call for the following remarks. IV-1 | 223
(a) The common property of all the subjects discussed is that they all related to local properties
of preschemes or morphisms, i.e. considered at a point, or the points of a fibre, or on a
(non-specified) neighbourhood of a point or of a fibre. These properties are generally of a
topological, differential, or dimensional nature (i.e. bringing the ideas of dimension and depth
into play), and are linked to the properties of the local rings at the points considered. One
type of problem is the relating, for a given morphisms f : X → Y and point x ∈ X, of the
properties of X at x with those of Y at y = f ( x ) and those of the fibre Xy = f −1 (y) at x.
Another is the determining of the topological nature (for example, the constructibility, or
the fact of being open or closed) of the set of points x ∈ X at which X has a certain property,
or for which the fibre X f ( x) passing through x has a certain property at x. Similarly, we
are interested in the topological nature of the set of points y ∈ Y such that X has a certain
property at all the points of the fibre Xy , or those such that this fibre itself has a certain
property.
1The order and content of §§11–21 are given only as an indication of what the titles will be, and will possibly be modified
before their publication. [Trans.] This was indeed the case: many of §§11–21 ended up having entirely different titles or content. See
here.
366
1. RELATIVE FINITENESS CONDITIONS. CONSTRUCTIBLE SETS IN PRESCHEMES 367
(b) The most important idea for the following chapters is that of flat morphisms of finite presenta-
tion, as well as the particular cases of smooth morphisms and étale morphisms. Their detailed
study (as well as that of connected questions) really starts in §11.
(c) Sections §§1–10 can be considered as being preliminary in nature, and as developing three
types of techniques, used, not only in the other sections of the chapter, but also, of course,
in the follow chapters:
(c1) Sections §§1–4 are envisaged as treating the diverse aspects of the idea of change of base,
above all in relation with the conditions of finiteness or flatness; we there initiate the
technique of descent, with its most elementary aspects (the questions of “effectiveness”
linked to this technique will be studied in Chapter V).
(c2) Sections §§5–7 are focused on what we may call Noetherian techniques, since the
preschemes considered are always locally Noetherian, whereas, on the contrary, there
is generally no finiteness condition imposed on the morphisms; this is essentially due
to the fact that the ideas of dimension and depth are hardly manageable except in
the case of Noetherian local rings. Recall that §7 constitutes a “delicate (?)” theory of
Noetherian local rings, not much used in what follows in the chapter.
(c3) Sections §§8–10 describe, amongst other things, the means of eliminating the Noetherian
hypotheses on the preschemes considered, by substituting such hypotheses for suitable
ones of finiteness (“finite presentation”) on the morphisms considered: the advantage
of this substitution is that the latter such hypotheses (those of finiteness on the mor-
phisms) are stable under base change, which is not the case for the Noetherian hypotheses
on the preschemes. The technique permitting this substitution relies, in some part, on
the use of the idea of the projective limit of preschemes, thanks to which we can reduce
a question to the same question with Noetherian hypotheses; on the other hand, it relies
on the systematic use of constructible sets, which have the double interest of being
preserved under taking inverse images (of arbitrary morphisms) and by direct images IV-1 | 224
(of morphisms of finite presentation), and having manageable topological properties
in locally Noetherian preschemes. The same techniques often even allow to restrict to
the case of more specific Noetherian rings, for example the Z-algebras of finite type, and
it is here that the properties of “excellent” rings (studied in §7) intervene in a decisive
manner. Independently of the question of elimination of Noetherian hypotheses, the
techniques of §§8–10, elementary in nature, find constant use in nearly all applications.
Definition (16.1.2). — The OY -augmented sheaf of rings ψ∗ (OX )/I fn+1 is called the n’th normal
invariant of f ; the ringed space (Y, ψ∗ (OX )/I fn+1 ) is called the n’th infinitesimal neighbourhood of Y
(n)
along f and is denoted by Y f or simply Y (n) . The sheaf of graded rings associated to the sheaf of
filtered rings ψ∗ (OX )
Gr• ( f ) = (I fn /I fn+1 )
M
(16.1.2.1)
n⩾0
is called the sheaf of graded rings associated to f . The sheaf Gr1 ( f ) = I f /I f2 is called the conormal
sheaf of f (that will be denoted by NY/X when there is no risk of confusion).
It is clear that the OY (n) = ψ∗ (OX )/I fn+1 (that we also denote O (n) ) form a projective system of IV-4 | 6
Yf
sheaves of rings on Y, the transition homomorphism ϕnm : OY (m) → OY (n) for n ⩽ m identifies OY (n)
with the quotient of OY (m) by the power (I f /I fn+1 )m of the augmentation ideal of OY (n) , kernel of
ϕ0n : OY (n) → OY . The Y (n) therefore form a inductive system of ringed spaces, all having underlying
space Y, and we have canonical morphisms of ringed spaces hn : Y (n) → X equal to (ψ, θn ), where
θn# is the canonical morphism ψ∗ (OX ) → ψ∗ (OX )/I fn+1 . It is clear that the sheaf Gr• ( f ) is a sheaf of
graded algebras over the sheaf of rings OY = Gr0 ( f ) and the Grk ( f ) of OY -modules.
As with every sheaf of filtered rings, we have a canonical surjective homomorphism of graded
OY -algebras
•
(16.1.2.2) SO Y
(Gr1 ( f )) −→ Gr• ( f )
which coincide in degrees 0 and 1 with the identities.
Examples (16.1.3). —
(i) Suppose that X is a locally ringed space, Y is reduced to a single point y (endowed with
a ring Oy ) and that, if x = ψ(y), θ # : Ox → Oy is a surjective homomorphism of rings
having as kernel the maximal ideal mx of Ox . So the OY (n) are identified with the rings
Ox /mnx +1 and Gr• ( f ) with the graded ring associated with the local ring Ox endowed with
the mx -preadic filtration.
(ii) Suppose that Y is a closed subset of an open subspace U of X and that the OY is induced
on Y by a quotient sheaf OU /I , where I is an ideal of OU such that Ix = Ox for every
x ̸∈ Y; if X is a locally ringed space we also suppose that Ix ̸= Ox for y ∈ Y so that (Y, OY )
is a locally ringed space.
Let ψ0 : Y → U be the canonical injection and denote by θ0 : OU → (ψ0 )∗ (OY ) the ho-
momorphism such that θ0# is the canonical homomorphism ψ0∗ (OU ) = OU |Y → (OU /I )|Y,
so that j0 = (ψ0 , θ0 ) : Y → U is a morphism of ringed spaces (and of locally ringed spaces
if X is a locally ringed space); if i : U → X is the canonical injection (morphism of ringed
spaces), j = i ◦ j0 is the morphism (ψ, θ ) of Y to X where ψ : Y → X is the canonical
injection and θ : OX → ψ∗ (OY ) is the homomorphism such that θ # = θ0# . Since θ # is
surjective we can apply the previous definitions; OY (n) is equal to ψ0∗ (OU /I n+1 ), and we
have (ψ0 )∗ (OY (n) ) = OU /I n+1 , and Grn ( j) = Grn ( j0 ) = ψ0∗ (I n /I n+1 ) = j0∗ (I n /I n+1 ).
(16.1.4). The example (16.1.3, (ii)) shows that in general the OY (n) are not canonically endowed with a
structure of an OY -module, or a fortiori with a structure of an OY -algebra. The data of such structure
is equivalent to the data of a homomorphism of sheaves of rings λn : OY → OY (n) , right inverse to
16. DIFFERENTIAL INVARIANTS. DIFFERENTIALLY SMOOTH MORPHISMS 370
the augmentation morphism ϕ0n ; it is also equivalent to the data of a morphism of ringed spaces
(1Y , λn ) : Y (n) → Y left inverse to the canonical morphism (1Y , ϕ0n ) : Y → Y (n) .
Proposition (16.1.5). — Let f = (ψ, θ ) : Y → X be an immersion of preschemes. Then:
(i) Gr• ( f ) is a quasi-coherent graded OY -algebra. IV-4 | 7
(ii) The Y (n) are preschemes, canonically isomorphic to subpreschemes of X.
(iii) Every homomorphism of sheaves of rings λn : OY → OY (n) , right inverse to the augmentation
homomorphism ϕ0n , makes the OY (n) and OY (k) for k ⩽ n quasi-coherent OY -algebras; the OY -
module structures induced from the above structures on the Grk ( f ) for k ⩽ n coincide with the ones
defined in (16.1.2).
P ROOF. (i) Since the question is local on X and Y, we can reduce to the case where Y is a closed
subpreschemes of X defined by an quasi-coherent ideal I of OX ; since OY is the restriction to
Y of OX /I the assertion (i) is evident, and Y (n) is the closed subprescheme of X defined by the
quasi-coherent ideal I n+1 of OX . Finally, to prove (iii) we notice that the data of λn makes the
ideal I /I n of the augmentation ϕ0n and their quotients I /I k+1 (1 ⩽ k ⩽ n) OY -modules, and
it suffices to prove by induction on k that the I /I k+1 are quasi-coherent OY -modules and the
structure of quotient OY -module induced on I k /I k+1 is the same as defined on (16.1.2). The
second assertion is immediate, I k /I k+1 being killed by I /I n+1 ; the first result, by induction on
k, is trivial for k = 1 and for I /I k+1 being an extension of I /I k by I k /I k+1 (1.4.17). □
Corollary (16.1.6). — Under the general hypotheses of (16.1.5), if the immersion f is locally of finite
presentation then the Grn ( f ) are quasi-coherent OY -modules of finite type.
P ROOF. Indeed, with the notation from the proof of (16.1.5), I is an ideal of finite type of OX
(1.4.7), therefore the I n /I n+1 are OY -modules of finite type, hence the conclusion. □
Corollary (16.1.7). — Under the general hypotheses of (16.1.5), let g : X → Y be a morphism of preschemes,
hn g
left inverse to f . Therefore, for every n, the composite morphism (1, λn ) : Y (n) −→ X −→ Y defines a
homomorphism of sheaves of rings λn : OY → OY (n) right inverse to the augmentation ϕ0n , making OY (n) a
quasi-coherent OY -algebra; via these homomorphisms, the transition homomorphism ϕnm : OY (m) → OY (n)
(n ⩽ m) are homomorphisms of OY -algebras. Also, if g is locally of finite type, then the OY (n) are quasi-coherent
OY -modules of finite type.
P ROOF. The first assertion is an immediate result from the definitions and (16.1.5). On the
other hand, if g is locally of finite type, then f is locally of finite presentation (1.4.3, (v)); the Grn ( f )
being then quasi-coherent OY -modules of finite type by (16.1.6), the same goes for the OY -modules
I /I n+1 , being extensions of a finite number of the Grk ( f ) (III, 1.4.17). □
Proposition (16.1.8). — Let X be a locally Noetherian prescheme, j : Y → X an immersion; Then the Y (n)
are locally Noetherian preschemes, the Grn ( j) are coherent OY -modules and the Gr• ( j) is a coherent sheaf of
rings over the space Y.
P ROOF. Everything is local on X and Y, so we reduce to the case where X is affine and j is a
closed immersion and therefore all the assertions are evident except for the last, which follows from
the fact that if A is a Noetherian ring and I is an ideal of A, then grI• ( A) is a Noetherian ring, taking
into account the exactness of the functor ψ∗ and (0, 5.3.7). □
Proposition (16.1.9). — Let X be a prescheme, j : Y → X an immersion locally of finite presentation, y a IV-4 | 8
point of Y. The following conditions are equivalent:
(a) There exists an open neighbourhood U of y in Y such that j|U is a homeomorphism of U onto an
open set of X.
(b) There is an integer n > 0 such that the canonical homomorphism
(ϕn−1,n )y : OY(n) ,y −→ OY(n−1) ,y
is bijective.
(c) There is an integer n > 0 such that (Grn ( j))y = 0.
In addition, if the integer n satisfies (b) or (c), then there is a neighbourhood V of y in Y such
that Grm ( j)|V = 0 for m ⩾ n and that ϕnm |V : OY (m) |V → OY (n) |V is bijective for m ⩾ n.
16. DIFFERENTIAL INVARIANTS. DIFFERENTIALLY SMOOTH MORPHISMS 371
P ROOF. Since the questions is local on Y, we can restrict ourselves to the case where j is a closed
immersion, Y being defined by a quasi-coherent ideal of finite type I of OX . The equivalence of (b) and
(c), for a given n, is immediate; also, since I n /I n+1 is an OX -module of finite type, there is an open
neighbourhood U of y in X such that I n |U = I n+1 |U (0, 5.2.2), so we also have I n |U = I m |U for
m ⩾ n proving the last assertions. To prove that (a) implies (b), we can restrict ourselves to the cases
where the underlying space of Y is equal to the underlying space of X and where I is generated by
a finite number of sections over X: since I is contained in the nilradical N of OX (I, 5.1.2), it is now
nilpotent which proves b). Finally, to prove that (b) implies (a), we can restrict ourselves to the case
where I n = I m ; therefore, for every y ∈ Y, since Iy ⊂ my , maximal ideal of OX,y , we must have
Iyn = 0 because of Nakayama’s lemma, since Iy is an ideal of finite type. The set of x ∈ X such that
Ixn = 0 is an open U of X contained in Y (0, 5.2.2); since on the other hand Ix ̸= 0 for x ∈ / Y, we
must have U = Y. □
Corollary (16.1.10). — For a restriction of the immersion j to an open neighbourhood of y in Y to be an open
immersion (in other words, for j to be a local isomorphism on the point y), it is necessary and sufficient that
(Gr1 ( j))y = (NY/X )y = 0.
P ROOF. The condition is clearly necessary, and the previous reasoning applied to n = 1 proves
that it is sufficient. □
Remark (16.1.11). —
(i) Under the conditions of the definition (16.1.1), the projective limit of the projective system
(OY(n) , ϕnm ) of sheaves of rings over Y is called the normal invariant of infinite order of f ,
and sometimes denoted by OY (∞) . When X is a locally noetherian prescheme, j : Y → X
a closed immersion, Y then is a closed subprescheme of X defined by a coherent ideal I
and OY (∞) is exactly the formal completion of OX along Y (I, 10.8.4), and Y (∞) = (Y, OY (∞) ) is
the formal prescheme that is the completion of X along Y (I, 10.8.5). In all cases, we could
say that Y (∞) is the formal neighbourhood of Y in X (via the morphism f ). In the particular
case we have just considered, it is the formal prescheme that is the inductive limit of the
infinitesimal neighbourhoods of order n.
(ii) Note that for a morphism of preschemes f = (ψ, θ ) : Y → X, it can happen that the
homomorphism θ # : ψ∗ (OX ) → OY is surjective without f being a local immersion and IV-4 | 9
without f being injective. We have an example by taking Y to be a sum of preschemes
Yλ all isomorphic to Spec(Ox ), where x ∈ X, ad taking f to be the morphism equal to the
canonical morphism in each of the Yλ .
16.2. Functorial properties of the normal invariants of an immersion
(16.2.1). Let f = (ψ, θ ) : Y → X and f ′ = (ψ′ , θ ′ ) : Y ′ → X ′ by two morphisms of ringed spaces such
that θ # and θ ′# are surjective; consider a commutative diagram of morphisms of ringed spaces
f
(16.2.1.1) YO /X
O
u v
Y′ / X′
f′
Let u = (ρ, λ), v = (σ, µ). We have ρ∗ (ψ∗ (OX )) = ψ′∗ (σ∗ (OX )) and as a result a commutative
diagram of homomorphisms of sheaves of rings over Y ′
ψ′∗ (µ# )
ρ∗ (ψ∗ (OX )) = ψ′∗ (σ∗ (OX )) / ψ′∗ (OX ′ )
ρ∗ (θ # ) θ ′#
ρ∗ (OY ) / OY ′
λ#
from which we conclude, if I and I ′ are the kernels of θ # and θ ′# , that we have ψ′∗ (µ# )(ρ∗ (I )) ⊂
I ′ , having in mind the exactness of the functor ρ∗ . We deduce that, for every integer n, ψ′∗ (µ# )(ρ∗ (I n )) ⊂
16. DIFFERENTIAL INVARIANTS. DIFFERENTIALLY SMOOTH MORPHISMS 372
n
I ′ , which shows that ψ′∗ (µ# ) defines, passing to the quotients, a homomorphism of sheaves of
rings
n +1
(16.2.1.2) νn : ρ∗ (ψ∗ (OX )/I n+1 ) −→ ψ′∗ (OX ′ )/I ′
and therefore a morphism of ringed spaces wn = (ρ, νn ) : Y ′(n) → Y (n) (which, for n = 0, is none
other than u). It follows immediately from this definition that the diagrams
hmn hm
Y (O n) / Y (m) /X
O O
wn wm v (n ⩽ m)
Y ′(n) / Y ′(m) / X′
h′mn h′m
(where the horizontal arrows are the canonical morphisms (16.1.2)) are commutative.
By passage to the quotients via the morphisms (16.2.1.2), and taking into account the exactness IV-4 | 10
of the functor ρ∗ , we obtain a di-homomorphism of graded algebras (relative to the morphism
λ# : ρ∗ (OY ) → OY ′ )
(16.2.1.3) gr(u) : ρ∗ (Gr• ( f )) −→ Gr• ( f ′ )
(or, if you like, a ρ-morphism (0, 3.5.1) Gr• ( f ) → Gr• ( f ′ )), and in particular a di-homomorphism of
conormal sheaves
gr1 (u) : ρ∗ (Gr1 ( f )) −→ Gr1 ( f ′ ).
It is also immediate that these homomorpisms give rise to a commutative diagram
(16.2.1.4) ρ ∗ ( SO
• (Gr ( f )))
1
/ ρ∗ (Gr• ( f ))
Y
YO ′ / X′
f′ O
u′ v′
Y ′′ / X ′′
f ′′
where f ′′ = (ψ′′ , θ ′′ ) is such that θ ′′# is surjective, and if wn′ and wn′′ are defined from u′ , v′ for one
and u′′ = u ◦ u′ , v′′ = v ◦ v′ for the other, we have wn′′ = wn ◦ wn′ , which follows immediately
from the definitions and from (0, 3.5.5); we have also gr(u′′ ) = gr(u′ ) ◦ ρ′∗ (gr(u)) if u′ = (ρ′ , λ′ ).
Therefore we can say that Y (n) and Gr• ( f ) depend functorially on f .
Proposition (16.2.2). — With the notation and hypotheses of (16.2.1), suppose also that f , f ′ , u, and v are
morphisms of preschemes. We have:
(i) The morphisms wn : Y ′(n) → Y (n) are morphisms of preschemes.
(ii) If Y ′ = Y × X X ′ , u and f ′ the canonical projections, and if f is an immersion or if v is flat, we have
Y ′(n) = Y (n) × X X ′ .
(iii) If Y ′ = Y × X X ′ and if v is flat (resp. if f is an immersion), the homomorphism
Gr(u) = gr(u) ⊗ I : Gr• ( f ) ⊗OY OY ′ −→ Gr• ( f ′ )
is bijective (resp. surjective).
16. DIFFERENTIAL INVARIANTS. DIFFERENTIALLY SMOOTH MORPHISMS 373
P ROOF.
(i) The hypotheses immediately imply that, for every y′ ∈ Y ′ , ρ∗y′ (θψ# ′ (y′ ) ) is a local homomor-
phism (I, 1.6.2), so wn is a morphism of preschemes (I, 2.2.1). IV-4 | 11
(ii) and (iii) If f is an immersion, we can restrict ourselves to the case where f is a closed immersion, Y
being defined by a quasi-coherent ideal I of OX and Y (n) by the ideal I n+1 ; the assertions
follows from (I, 4.4.5).
Second, suppose that v is flat; we can restrict ourselves to the case where X = Spec( A),
Y = Spec( B), X ′ = Spec( A′ ) are affines, A′ being a flat A-module; so Y ′ = Spec( B′ ) where
B′ = B ⊗ A A′ ; in addition, if I is the kernel of the homomorphism A → B, the kernel I′ of
n +1
A′ → B′ is identified with I ⊗ A A′ by flatness, and I ′n /I ′ is equal to
ψ′∗ (σ∗ ((In /In+1 )e) ⊗σ∗ (OX ) OX ′ ) =
ψ′∗ (σ∗ ((In /In+1 ))e)⊗ψ′∗ (σ∗ (OX )) ψ′∗ (OX ′ ) = ρ∗ (I n /I n+1 ) ⊗ρ∗ (ψ∗ (OX )) ψ′∗ (OX ′ )
and in particular for n = 0, we have
OY ′ = ρ∗ (OY ) ⊗ρ∗ (ψ∗ (OX )) ψ′∗ (OX ′ )
n +1
from which we have canonical isomorphism of I ′n /I ′ with
ρ∗ (I n /I n+1 ) ⊗ρ∗ (OY ) OY ′ = (I n /I n+1 ) ⊗OY OY ′
which proves (iii). Let now Cn = Γ(Y, OY (n) ), Cn′ = Γ(Y ′ , OY ′(n) ). As Y (n) and Y ′(n) are
affine schemes (16.1.5), the kernel Kn (resp. K′n ) of the homomorphism Cn → Cn−1 (resp.
n +1
Cn′ → Cn′ −1 ) is Γ(Y, I n /I n+1 ) (resp. Γ(Y, I ′n /I ′ )); therefore we can deduce from
′ ′
the above results that Kn = Kn ⊗ A A . Now, we have a commutative diagram
0 / Kn ⊗ A A ′ / Cn ⊗ A A′ / Cn−1 ⊗ A A′ /0
r sn s n −1
0 / K′n / Cn′ / C′ /0
n −1
where the vertical arrow of the left is bijective and the two lines are exact (A′ being a flat
A-module). We deduce by induction that sn is bijective for every n, because it is true by
hypothesis for n = 0, and is deduced by application of the five lemma for all n. That proves
the second assertion of (ii).
□
Corollary (16.2.3). — Let g : X → Y, u : Y ′ → Y be two morphisms of preschemes, X ′ = X ×Y Y ′ ,
g′ : X ′ → Y ′ and v : X ′ → X by the canonical projections. Let f : Y → X by a Y-section of X (and therefore
an immersion), f ′ = f (Y ′ ) : Y ′ → X ′ the Y ′ -section of X ′ deduced from f by the base change u. We have:
(n) (n)
(i) The morphism wn : Y f′ ′ → Yf corresponding to f , f ′ , u, v (16.2.1) and the canonical morphism
(n) (n) (n)
h′n : Y f′ ′ → X ′ identifies Yf′ ′ with the product Yf × X X ′ .
(ii) If we endow O (n) (resp. O ′ (n) ) with the structure of an OY -algebra defined by g (resp. with the
Yf Y′
f
structure of an OY ′ -algebra defined by g′ ) (16.1.5, (iii)), then the homomorphism of OY ′ -algebras
(16.2.3.1) ρ ∗ (O (n) ) ⊗OY OY′ −→ OY′ (n)
Yf f′
induced by the homomorphism νn (16.2.1.2) is bijective. Also, the homomorphism of OY ′ -modules IV-4 | 12
′
(16.2.3.2) Gr1 (u) : Gr1 ( f ) ⊗OY OY ′ −→ Gr1 ( f )
is bijective.
P ROOF.
(i) Let us first note that f ′ : Y ′ → X ′ and u : Y ′ → Y identifies Y ′ with the product Y × X X ′
(via the structure morphisms f : Y → X and v : X ′ → X) (14.5.12.1). The conclusion of (i)
now follows from (16.2.2, (ii)), the morphism g being an immersion.
16. DIFFERENTIAL INVARIANTS. DIFFERENTIALLY SMOOTH MORPHISMS 374
hn h′n
Xo v
X′
g g′
Yo u
Y′
(n) (n)
identifies Y f′ ′ with the product Y f × X X ′ , so (I, 3.3.9) it identifies (via the morphisms
(n) (n) (n) (n)
g′ ◦ h′n and wn ) Y f′ ′ to the product Y f ×Y Y ′ . Since Y f (resp. Y f′ ′ ) is the affine
prescheme over Y (resp. over Y ′ ) associated with the OY -algebran O (n) (resp. to the
Yf
OY ′ -algebran O ), the fact that the canonical homomorphism (16.2.3.1) is bijective
Y ′ ′ (n)
f
follows from (II, 1.5.2). Finally, the canonical homomorphism (16.2.3.1) is compatible
with the augmentations O (n) → OY and O ′ (n) → OY ′ ; since O (n) is a direct sum (as an
Yf Y′ Yf
f
OY -module) of OY and the augmentation ideal I /I n+1 ,
we can therefore see that the
canonical homomorphism (16.2.3.1), restricted to I /I n+1 ⊗OY OY ′ , is a bijection of the
latter onto I ′ /I ′n+1 . For n = 1 this shows that Gr1 (u) is bijective.
□
We note that, under the hypotheses of (16.2.3), the homomorphisms Grn (u) are surjective in view
of the above, but are not bijective in general for n ⩾ 2. However:
Corollary (16.2.4). — Under the hypotheses of (16.2.3), suppose that u : Y ′ → Y is a flat morphism (resp.
that the Grn ( f ) are flat OY -modules for n ⩽ m). Then the homomorphism
Grn (u) : Grn ( f ) ⊗OY OY ′ −→ Grn ( f ′ )
is bijective for all n (resp. for n ⩽ m).
P ROOF. If u is flat, then we deduce by base change that the same is true for v : X ′ → X, and
we already know in this case that Gr(u) is bijective (16.2.2, (iii)). If the Grn ( f ) are flat for n ⩽ m,
then we first see by induction on n that the same holds for I /I n+1 for n ⩽ m, because of the exact
sequences
0 / I n /I n+1 / I /I n+1 / I /I n /0
(0, 6.1.2); in addition, we have the commutative diagram IV-4 | 13
0 / (I n /I n+1 ) ⊗O OY ′ / (I /I n+1 ) ⊗O OY ′ / (I /I n ) ⊗O OY ′ /0
Y Y Y
0 / I ′ n /I ′ n+1 / I ′ /I ′ n+1 / I ′ /I ′ n /0
in which the lines are exact (the first by flatness (0, 6.1.2)) and the two last vertical arrows are bijective
by virtue of (16.2.2, (ii)); hence the conclusion. □
Remarks (16.2.5). —
(i) The reasoning of (16.2.2, (i)) still applies to (16.2.1.1) when these are morphisms of locally
ringed spaces (I, 1.8.2).
(ii) In (16.2.2, (ii)), the conclusion is no longer necessarily valid if we only suppose that v and f
are morphisms of preschemes ( f satisfying the condition of (16.1.1)). For example (with
the notation of the proof of (16.2.2, (ii))), it can happen that I = 0 but the kernel I′ of
A′ → B′ = B ⊗ A A′ is not zero and that B′ ̸= 0, in which case we have Y (n) = Y for all n,
but Y ′ ̸= Y ′ . We have an example of this by taking A = Z, B = Q, A′ = ∏∞
(n) h
h=1 ( Z/m Z )
where m > 1.
16. DIFFERENTIAL INVARIANTS. DIFFERENTIALLY SMOOTH MORPHISMS 375
(16.2.6). Consider the particular case of the diagram (16.2.1.1) where X ′ = X, v is the identity, X
a prescheme, Y a subprescheme of X, Y ′ a subprescheme of Y, f , u, and f ′ = f ◦ u the canon-
ical injections; the di-homomorphism (16.2.1.3) gives us, by tensoring with OY ′ over ρ∗ (OY ), a
homomorphism of graded OY ′ -algebras
(16.2.6.1) u∗ (Gr• ( f )) −→ Gr• ( f ′ ).
On the other hand, we identify OY to ψ∗ (OX )/I f and OY ′ to ρ∗ (OY )/Iu ; since ρ∗ is an exact functor,
we have ρ∗ (OY ) = ρ∗ (ψ∗ (OX ))/ρ∗ (I f ) = ψ′∗ (OX )/ρ∗ (I f ), and since OY ′ is moreover identified
with ψ′∗ OX /I f ′ , we see that Iu = I f ′ /ρ∗ (I f ). We deduce that for every integer n there is a
canonical homomorphism I fn′ /I fn′ +1 → Iun /Iun+1 , from which we have a canonical morphism of
graded OY ′ -algebras
(16.2.6.2) Gr• ( f ′ ) −→ Gr• (u).
(16.2.7.1) j∗ (NY/X ) / NY ′ /X / NY ′ /Y /0
where the arrows are the degree 1 components of the canonical homomorphisms (16.2.6.1) and (16.2.6.2).
P ROOF. The problem being local, we can restrict to the case where X = Spec( A), Y = Spec( A/I)
and Y ′ = Spec( A/K), I and K being ideals of A such that I ⊂ K; everything reduces to seeing
that the sequence of canonical morphisms I/KI → K/K2 → (K/I)/(K/I)2 → 0 is exact, which is IV-4 | 14
immediate given that the image of I/KI in K/K2 is (I + K2 )/K2 and that (K/I)/(K/I)2 is identified
with K/(I + K2 ). □
It is easy to give examples where the sequence (16.2.7.1) extended on the left by 0 is not exact;
with the above notation, it suffices to take A = k[ T ], I = AT 2 , K = AT, because then (I + K2 )/K2 = 0
and I/KI ̸= 0. See however (16.9.13) and (19.1.5) for some cases where the extended sequence is
indeed exact.
and call the sheaf of principal parts of order n of the S-prescheme X, the OX -augmented sheaf of rings,
n-th normal invariant of ∆ f (16.1.2). We will also write P ∞ ∞
f = PX/S = lim P n , Grn (P f ) =
←−n X/S
Grn (PX/S ) = Grn (∆ f ) (16.1.2); the OX -module Gr1 (∆ f ), augmentation sheaf of ideals of PX/S 1 , is
denoted by Ω1f or Ω1X/S , and is called the OX -module of 1-differentials of f , or of X with respect to S,
or of the S-prescheme X.
X).
(16.3.5). In what follows, the two OX -algebra structures defined on the PX/S n ∞ will play
and on PX/S
∞
very different roles: we will now agree, unless said otherwise, that when PX/S or PX/S is considered as an
n
(resp. principal part of infinite order) of t. We set dt = d1 t − t, and we say that dt is the differential of t
(an element of Γ(U, Ω1X/S ), also denoted d X/S (t)).
n⩾0
in particular Ω1f
= Ω1X/S
is the quasi-coherent OX -module corresponding to the B-module of
1-differentials of B over A, Ω1B/A (0, 20.4.3). The projection morphisms p1 : X ×S X → X, p2 :
X ×S X → X corresponding to the two homomorphisms of rings j1 : B → B ⊗ A B, j2 : B → B ⊗ A B
n
such that j1 (b) = b ⊗ 1, j2 (b) = 1 ⊗ b, so that (by the convention of (16.3.5)), PB/A is always
n ; the ring j1
considered as a B-algebra via the composite homomorphism B −
→ B ⊗ A B → PB/A
j2 n
homomorphism B − → B ⊗ A B → PB/A is denoted by dnB/A and corresponds to dnX/S acting on
Γ( X, OX ); for every t ∈ B, dt is equal to d B/A t, defined in (0, 20.4.6).
n
If πn : B ⊗ A B → PB/A is the canonical homomorphism, so we have, in light of the preceding
definitions,
(16.3.7.1) πn (b ⊗ b′ ) = b · πn (1 ⊗ b′ ) = b · dnB/A (b′ ) for b ∈ B, b′ ∈ B.
2[Trans.] This is, locally we have (0, 20.1.1).
16. DIFFERENTIAL INVARIANTS. DIFFERENTIALLY SMOOTH MORPHISMS 377
P ROOF. We immediately reduce to the case where X = Spec( B) and S = Spec( A) are affine
and the proposition follows from (16.3.7.1) since πn is surjective. We note that in general dnX/S is not
surjective (even for n = 1). □
Proposition (16.3.9). — Suppose that f : X → S is a morphism locally of finite type. Then the P nf and the
Grn (P f ) are quasi-coherent OX -modules of finite type.
P ROOF. This follows from (16.1.6) and from the fact that ∆ f is locally of finite presentation
(I, 4.3.1). □
Xo
u
(16.4.1.1) X′
f f′
So w
S′
Xo
u
X′
∆f ∆f′
X ×S X o v
X ′ ×S′ X ′
(16.4.3). We verify immediately (for example, by restricting ourselves to the affine case with help of
(16.3.7)) that with the notation of (16.4.1), the diagram
λ# / OX ′
(16.4.3.1) ρ ∗ (O X )
ρ∗ (PX/S
n ) / Pn′ ′
νn X /S
where the vertical arrows are the ones defining the algebra structure chosen in (16.3.5) (that is to say,
the ones coming from the first projections) is commutative; the same goes for the diagram
λ# / OX ′
(16.4.3.2) ρ ∗ (O X )
ρ∗ (dnX/S ) dnX ′ /S′
ρ∗ (PX/S
n ) / Pn′ ′
νn X /S
the vertical arrows defining here the algebra structure from the second projection; besides, if σ and IV-4 | 18
σ′ are the canonical symmetries corresponding to f and f ′ (16.3.4), we have
νn ◦ ρ∗ (σ ) = σ′ ◦ νn
which switches one diagram with the other. We deduce from (16.4.3.1) a canonical homomorphism
of augmented OX ′ -algebras
(16.4.3.3) Pn (u) : u∗ (PX/S
n n
) = PX/S ⊗OX OX ′ −→ PXn ′ /S′
and it follows from (16.4.3.2) that the diagram
(16.4.3.4) OX ′
id / OX ′
(16.4.3.7) OX ′
id / OX ′
d X/S ⊗1 d X ′ /S′
Ω1X/S ⊗OX OX ′ / Ω1 ′ ′
X /S
16. DIFFERENTIAL INVARIANTS. DIFFERENTIALLY SMOOTH MORPHISMS 379
(16.4.4). When S = Spec( A), S′ = Spec( A′ ), X = Spec( B), X ′ = Spec( B′ ) are affine, so that we have
a commutative diagram of ring homomorphisms
BO / B′
O
A / A′
the image of IB/A in B′ ⊗ A′ B′ is contained in IB′ /A′ , and the homomorphism νn corresponds to the
homomorphism of rings PB/A n → PBn′ /A′ induced from the homomorphism B ⊗ A B → B′ ⊗ A′ B′ by
passing to quotients. The homomorphism (16.4.3.6) corresponds to the homomorphism defined in
(0, 20.5.4.1), and the commutative diagram (16.4.3.7) to the diagram (0, 20.5.4.2).
Proposition (16.4.5). — Suppose that X ′ = X ×S S′ , f ′ and u the canonical projections. Then the canonical IV-4 | 19
homomorphisms Pn (u) (16.4.3.3) and Gr1 (u) (16.4.3.6) are bijective.
P ROOF. We have X ′ ×S′ X ′ = ( X ×S X ) ×S S′ , and it suffices to apply (16.2.3, (ii)) replacing g by
the first p1 : X ×S X → X and f by the diagonal ∆ f . □
We note that under the hypotheses of (16.4.5) the homomorphism Gr• (u) (16.4.3.5) is surjective,
but not bijective in general. However (16.2.4):
Corollary (16.4.6). — Under the hypotheses of (16.4.5), suppose in addition that w : S → S′ is flat (resp.
that Grn (PX/S
n ) are flat O -modules for n ⩽ m); then the homomorphism
X
(16.4.7). Let S be a prescheme, E a quasi-coherent OS -Module, and set X = V(E ) (II, 1.7.8), the
vector bundle associated to E , equal to Spec(SOS (E )). Let f : X → S be the structure morphism.
For every open U of S and every section t ∈ Γ(U, E ), t is identified with a section of SOS (E ) over U;
let t′ be its image in Γ( f −1 (U ), OX ) = Γ(U, f ∗ (OX )) = Γ(U, SOS (E )), and set
of OX -algebras
(16.4.7.2) f ∗ (SOS (E )) −→ PX/S
n
and in view of the above remark, if K is the ideal kernel of augmentation SOS (E ) → OS , the
image of K n+1 by (16.4.7.2) is zero, so that by factoring by K n+1 , we finally have a canonical
homomorphism
(16.4.7.3) δn : f ∗ (SOS (E )/K n +1 n
) −→ PX/S .
Proposition (16.4.8). — Under the conditions of (16.4.7), the homomorphisms δn are bijective and form a
projective system of isomorphisms; we deduce an isomorphism of graded OS -algebras
(16.4.8.1) f ∗ ( SO
•
S
(E )) −→ Gr• (PX/S ).
P ROOF. The fact that homomorphisms (16.4.7.3) form a projective system follows immediately
from their definition. To prove they are isomorphisms, it suffices to prove that (16.4.8.1) is an IV-4 | 20
isomorphism, since both filtrations involved in (16.4.7.3) are finite (Bourbaki, Alg. comm., chap. III,
§2, no 8, cor. 3 of th. 1). To do this, consider the split exact sequence of OS -modules
(16.4.8.2) 0 /E u / E ⊕E v /E /0
16. DIFFERENTIAL INVARIANTS. DIFFERENTIALLY SMOOTH MORPHISMS 380
where, for every pair of sections s, t of E over an open U of S, we take u(s) = (−s, s) and v(s, t) =
s + t. We have
X ×S X = Spec(SOS (E ) ⊗OS SOS (E )) = Spec(SOS (E ⊕ E ))
((II, 1.4.6) and (II, 1.7.11)), and the diagonal morphism X → X ×S X corresponds (II, 1.2.7) to the
homomorphism of OX -algebras S(v) : SOS (E ⊕ E ) → SOS (E ) (II, 1.7.4), such that if I is the kernel
of this homomorphism, then we have
n
PX/S = f ∗ (SOS (E ⊕ E )/I n+1 ).
The proposition now will be a consequence of the following lemma: □
u v
Lemma (16.4.8.3). — Let Y be a ringed space, 0 → F ′ − → F ′′ → 0 an exact sequence of OY
→F −
modules such that each point y ∈ Y has an open neighbourhood V such that the sequence 0 → F ′ |V →
F |V → F ′′ |V → 0 is split. Let I be the kernel ideal of S(v):
SOY (F ) −→ SOY (F ′′ ),
• ( S (F )) be the graded O -algebra associated to the O -algebra S (F ) endowed with the
and let grI OY Y Y OY
I -preadic filtration. Then the homomorphism of graded OY -algebras
•
(16.4.8.4) SO Y
(F ′ ) ⊗OY SO
•
Y
(F ′′ ) −→ grI
•
(SOY (F ))
(where the first member is the graded tensor product of symmetric OY -algebras endowed with the canonical
gradation (II, 1.7.4) and (II, 2.1.2)), induced by the canonical injection
F ′ −→ I = gr1I (SOY (F )),
is bijective.
P ROOF. The injection F ′ → I indeed canonically gives a homomorphism of graded OY -
algebras SO • (F ′ ) → gr• ( S• (F )), and since the second member is by definition a graded
Y I OY
SO• (F ′′ )-algebra, we induce the canonical homomorphism (16.4.8.4) by tensoring the above
Y
with SO • (F ′′ ). To prove the lemma we can, being a local problem, restrict to the case where
Y
F = F ′ ⊕ F ′′ , u and v the canonical homomorphisms. Then the graded algebra SO • (F ) is canoni-
Y
• ′ • ′′
cally identified with the graded tensor product SO (F ) ⊗OY SO (F ) (II, 1.7.4), and it is immediate
Y Y
that I is therefore the ideal I ′ ⊗OY SO
• (F ′′ ), where I ′ is the augmentation ideal of S• (F ′ ), that
OY
Y
m (F ′ ) for m ⩾ 1. We conclude that I n = I ′n ⊗ • ′′
is to say the (direct) sum of the SO
Y
OY SO (F ), Y
where this time I ′n is the direct sum of the SO
m (F ′ ) for m ⩾ n; we have therefore I n /I n+1 =
Y
n (F ) ⊗ • ′′
SO
Y
OY SO (F ), which proves that (16.4.8.4) is bijective.
Y
□
Having proved the lemma, it remains to see that the homomorphism (16.4.8.1) is the image by IV-4 | 21
f ∗ of the homomorphism (16.4.8.4) corresponding to the exact sequence (16.4.8.2); we can easily
see that it follows from the definition of u (16.4.8.2) and of δ (16.4.7.1), given the definition of the
OX -algebra structures of PX/S
n and of the dnX/S (16.3.5) and (16.4.3.6).
In particular:
Corollary (16.4.9). — Under the conditions of (16.4.7), we have a canonical isomorphism
(16.4.9.1) gr1 (δ) : f ∗ (E ) ≃ Ω1X/S .
Corollary (16.4.10). — If S = Spec( A), E = OSm , so that
X = Spec( A[ T1 , . . . , Tm ]),
then PX/S n is canonically identified with the OX -algebra corresponding to the quotient A[ T1 , . . . , Tm ]-algebra
A[ T1 , . . . , Tm , U1 , . . . , Um ]/Kn+1 , where the Ui (1 ⩽ i ⩽ m) are m new indeterminates and K is the ideal
generated by U1 , . . . , Um .
We thus recover in particular the structure of Ω1X/S in this case (0, 20.5.13).
In addition, note that the dnX/S then corresponds to a polynomial F ( T1 , . . . , Tm ), the class modulo
K n + 1 of F ( T1 + U1 , . . . , Tm + Um ), which follows from the definition (16.4.7.1).
16. DIFFERENTIAL INVARIANTS. DIFFERENTIALLY SMOOTH MORPHISMS 381
(via the OS -algebra structure on O (n) defined by f (16.1.7)), making the diagram
Sg
(16.4.11.2) OS = g ∗ (O X )
λn
/O (n)
: Sg
g∗ (dnX/S ) ϖn
&
g∗ (PX/S
n )
g ∆f g ∆f
X / X ×S X X / X ×S X
( g◦ f ,1X )S (1X ,g◦ f )S
identifies S with the product of the ( X ×S X )-preschemes X and X by the morphisms ∆ f and IV-4 | 22
( g ◦ f , 1X )S (resp. (1X , g ◦ f )S ). On the other hand, the diagrams
( g◦ f ,1X )S (1X ,g◦ f )S
(16.4.11.4) X /X X /X
f p1 f p2
S /X S /X
g g
identify X to the product of X-preschemes S and X ×S X via the morphisms g and p1 (resp. p2 )
(particular case of the associativity formula (I, 3.3.9.1)). We can say that ∆ f , considered as an X-
section of X ×S X (relative to p1 or p2 ) plays the role of a universal section for the S-sections of X: each
of these sections g in fact are deduced by base change ( g ◦ f , 1X )S : X → X ×S X. The definition of the
homomorphism ω̄n and the fact that it is bijective follows from the remarks of (16.2.3, (ii)) applied to
the first diagram (16.4.11.4). The commutativity of the first diagram (16.4.11.4) follows also from
(16.2.3, (ii)) this time applied to the second diagram (16.11.4). To explain ϖn , we can restrict ourselves
to the case where g is a closed immersion: Indeed, for every s ∈ S, there is an open neighbourhood
W of s in S such that g(W ) is closed in an open set U of X, and it is clear that g|W is a W-section of
the morphism U ∩ f −1 (W ). We can then suppose that S is a closed subprescheme of X defined by a
quasi-coherent ideal K . Then the preceding definitions show that if W is an open of S, t is a section
of OX over f −1 (W ), ϖn (dn t|W ) is equal to the canonical image of t in Γ(W, (OX /K n+1 )|W ). The
uniqueness of ϖn then follows since the image of OX under dnX/S generates the OX -module PX/S n
(16.3.8). □
Corollary (16.4.12). — Let k be a field, X a k-prescheme, x a point of X rational over k. Then (PX/S
n ) ⊗
x Ox
k ( x ) is canonically isomorphic (as an augmented k ( x )-algebra) to Ox /mx .
n + 1
maximal ideal of OXs ,x ; more precisely, this isomorphism sends (dn t) x ⊗ 1 (where t is a section of OX over
n +1
an open neighbourhood of x in X) to the class of t x ⊗ 1 modulo m′x .
16. DIFFERENTIAL INVARIANTS. DIFFERENTIALLY SMOOTH MORPHISMS 382
The preceding corollaries justify the terminology “sheaf of principal parts of order n”.
P ROOF. It suffices to remark that S−1 (( B ⊗ A B)/In+1 ) = S−1 ( B ⊗ A B)/(S−1 I)n+1 by flatness,
and that S−1 ( B ⊗ A B) = (S−1 B) ⊗ A (S−1 B) (I, 1.3.4). □
Corollary (16.4.15). — The notation being that of (16.4.14), let R be a multiplicative subset of A such that
ρ( R) ⊂ S. Then we have canonical isomorphisms
that is to say that we reduce to the case there ρ( R) is consists of invertible elements of B. But then the
isomorphism (16.4.15.2) is simply induced by the canonical isomorphism B ⊗ A B → B ⊗ R−1 A B by
passing to quotients (I, 1.5.3). □
We deduce from these isomorphisms of the associated graded rings, and in particular a canonical
isomorphism
Corollary (16.4.17). — Let k be a field, K the field of rational functions k( T1 , . . . , Tr ). Then, for every
n which sends U to dn T − T .1 is
integer n, the homomorphism of K [U1 , . . . , Ur ] (Ui indeterminates) into PK/k i i i
surjective and defines an isomorphism from the quotient K [U1 , . . . , Un ]/mn+1 (where m is the ideal generated
n .
by the Ui ) to PK/k
P ROOF. This follows from (16.4.8), (16.4.10) and (16.4.14), where we take A = k, B = k [ T1 , . . . , Tr ]
and S = B − {0}. □
We thus recover the fact that the dTi form a basis of the K-vector space Ω1K/k (0, 20.5.10).
Then gX/Y/Z is surjective, and its kernel is the sheaf of ideals generated by the image under f X/Y/Z
of the augmentation ideal of f ∗ (PX/Z
n ).
16. DIFFERENTIAL INVARIANTS. DIFFERENTIALLY SMOOTH MORPHISMS 383
P ROOF. First note that gX/Y/Z corresponds to the case in (16.4.3.3) where X ′ = X, S′ = Y and IV-4 | 24
S = Z, u = 1X , and f X/Y/Z to the case where we replace X ′ , X, S, S′ by X, Y, Z, Z respectively and u,
f by f , g respectively.
We have a commutative diagram (I, 3.5)
∆f j
(16.4.18.3) X / X ×Y X / X ×Z X
p f ×z f
#
f
Y / Y ×Z Y
∆g
OY -module in general).
P ROOF. We know (I, 5.3.8) that the diagonal ∆ j : X → X ×Y X is an isomorphism, from
which the first assertion follows. If ϖ1 and ϖ2 are the two canonical homomorphisms of algebras
OY → PY/Z n corresponding respectively to the two canonical projections p1 , p2 of Y × Z Y → Y,
recall that by definition ((16.3.5) and (16.3.6)) ϖ1 is the structure homomorphism of the OY -algebra
PY/Z
n n
and ϖ2 = dY/Z . The OX -algebra j∗ (PY/Z
n ) is therefore identified with P n /ϖ (K )P n
Y/Z 1 Y/Z
and its quotient by the ideal generated by j∗ (dY/Zn (K )) to PY/Z
n / ( ϖ (K ) + ϖ (K ))P n . Now
1 2 Y/Z
note that we have a commutative diagram IV-4 | 25
j
Yo X
∆f ∆ f ◦j
Y ×Z Y o X ×Z X
j× Z j
identifying X with the product of the (Y × Z Y )-preschemes Y and X × Z X (I, 5.3.7). Since j × Z j is
an immersion, we therefore deduce from this remark and from (16.2.2) that if ∆Y/Z
n and ∆nX/Z denote
the infinitesimal neighbourhoods of order n of Y and X by the canonical immersions ∆ f and ∆ f ◦ j
respectively, then we have a diagram
∆Y/Z
n o ∆nX/Z
Y ×Z Y o X ×Z X
j× Z j
16. DIFFERENTIAL INVARIANTS. DIFFERENTIALLY SMOOTH MORPHISMS 384
making ∆nX/Z the product of the (Y × Z Y )-preschemes ∆Y/Z n and X × Z X. We can also say that
PX/Z is identified with the sheaf of rings PY/Z ⊗OY× Y OX ×Z X . But we see immediately that
n n
Z
(for example, by restricting to the affine case) that OX ×Z X = OY ×Z Y /( p1∗ (K ) + p2∗ (K ))OY ×Z Y .
Therefore PX/Z
n is identified with the quotient of PY/Z
n by the ideal generated by the image in
∗ ∗
PY/Z of p1 (K ) + p2 (K ). But by definition this ideal is generated by ϖ1 (K ) + ϖ2 (K ).
n □
Corollary (16.4.21). — Let f : Y → Z be a morphism, j : X → Y an immersion. We have an exact sequence
of quasi-coherent OX -modules
(16.4.21.1) NX/Y / j ∗ ( Ω1 ) / Ω1 / 0.
Y/Z X/Z
is bijective.
P ROOF. The commutative diagram IV-4 | 26
q
Yo X ×S Y o
id
X ×S Y
g h p
So So X
id f
(resp. p∗ (PX/S
n )). We conclude the proposition by considering the case n = 1. □
We immediately generalize (16.4.23) to the case of a product of any finite number of S-preschemes.
Remarks (16.4.24). —
(i) We will see (17.2.3) that when the morphism f : X → Y in (16.4.18) is smooth, the homo-
morphism f X/Y/Z in (16.4.19.1) is locally left invertible and in particular injective. Similarly,
when the morphism f ◦ j : X → Z of (16.4.20) is smooth, the homomorphism on the left
in (16.4.21.1) is locally left invertible and a fortiori injective (17.2.5). In Chapter V, we will
also give a variant, in the case of modules over a prescheme, of the “imperfection modules”
studied in (0, 20.6), and the exact sequences where they occur.
16. DIFFERENTIAL INVARIANTS. DIFFERENTIALLY SMOOTH MORPHISMS 385
(ii) Let X be a topological space, A a sheaf of rings over X and B a A -algebra over X. Then it
is clear that
U 7−→ PΓn(U,B )/Γ(U,A ) (U open in X)
is a presheaf of augmented Γ(U, B )-algebras, and therefore the associated sheaf PB n
/A is
an augmented B-algebra. In the particular case where X is a prescheme, f = (ψ, θ ) : X → S
a morphism of preschemes, it follows easily from (16.4.16) and from the exactness of
the functor lim that PX/Sn is canonically isomorphic to PO n . It follows that the
X /ψ (OS )
∗
−→
formalism developed in the present paragraph could be considered as a particular case IV-4 | 27
of a differential formalism for ringed spaces endowed with a sheaf of algebras over the
structure sheaf. However, we did not start with this point of view, which is less intuitive
and less convenient for applications. It also seems that, for various kinds of “varieties”,
the “global” constructions of the P n analogous to those we have used here are also better
suited for applications.
16.5. Relative tangent sheaves and bundles; derivations.
(16.5.1). Let f = (ψ, θ ) : X → S be a morphism of ringed spaces. For every OX -module F , we say
S-derivation (or ( X/S)-derivation, or f -derivation) of OX to F for every homomorphism of sheaves of
additive groups D : OX → F satisfying the following conditions:
(a) for every open V of X, and all pair of sections (t1 , t2 ) of OX over V, we have
(16.5.1.1) D ( t1 t2 ) = t1 D ( t2 ) + D ( t1 ) t2 ;
(b) for every open V of X, every section t of OX over V, and every section s of OS over an open
U of S such that V ⊂ f −1 (U ), we have
(16.5.1.2) D ((s|V )t) = (s|V ) D (t).
It is clear that this amounts to saying that, for all x ∈ X, the homomorphism of additive groups
Dx : Ox → Fx is an O f ( x) -derivation.
Another interpretation consists of considering the OX -algebra DOX (F ) as equal to OX ⊕ F ,
the algebra structure being defined by the condition that for every open V of X, the product of
two sections of OX (resp. of a section of OX and a section of F ) over V is defined by the ring
structure of Γ(V, OX ) (resp. the Γ(V, OX )-module structure on Γ(V, F )), and the product of two
sections of F over V is chosen to be zero; then F is an ideal of DOX (F ), the kernel of the canonical
augmentation DOX (F ) → OX , and to say that D is an S-derivation of OX to F means that 1OX + D
is an OS -homomorphism of algebras from OX to DOX (F ), which, composed with the augmentation,
gives 1OX .
The S-derivations of OX to F clearly form a Γ( X, OX )-module DerOS (OX , F ).
When F = OX , an S-derivation of OX to itself is simply called an S-derivation of OX .
Proposition (16.5.2). — Let A be a ring, B an A-algebra, L a B-module; let S = Spec( A), X = Spec( B),
F =e L. Then the map D 7→ Γ( D ) which sends every S-derivation D of OX to F to the map Γ( D ) : t 7→ D (t)
of B to L, is an isomorphism of B-modules from DerS (OX , F ) to Der A ( B, L) (cf. (0, 20.1.2)).
P ROOF. This follows immediately from the given interpretation of S-derivations in terms of IV-4 | 28
homomorphisms of algebras, analogous to the interpretation given in (0, 20.1.6), and from the
canonical correspondence between homomorphisms of OX -algebras and homomorphisms of B-
algebras ((I, 1.3.13) and (I, 1.3.8)). □
Proposition (16.5.3). — Let f = (ψ, θ ) : X → S be a morphism of preschemes.
(i) The differential d X/S : OX → Ω1X/S (16.3.6) is an S-derivation.
(ii) For every OX -module F , the map u 7→ u ◦ d X/S is an isomorphism of Γ( X, OX )-modules
(16.5.3.1) HomOX (Ω1X/S , F ) ≃ DerS (OX , F ).
P ROOF. The assertion (i) has already been written (16.3.6). On the other hand, it is immediate
(in light of (0, 20.4.8)) that u 7→ u ◦ d X/S is injective, considering the restrictions to a fibre Ox of the
two sides and using (16.4.16.2). To see that the homomorphism (16.5.3.1) is surjective, consider an
S-derivation D : OX → F ; for every affine open V = Spec( B) of X, such that f (V ) is contained in
16. DIFFERENTIAL INVARIANTS. DIFFERENTIALLY SMOOTH MORPHISMS 386
(16.5.9). Let D1 , D2 be two S-derivations of OX . For every open U of X, if D1U , D2U are the corre-
sponding derivations of the ring Γ(U, OX ), the bracket
[ D1U , D2U ] = D1U ◦ D2U − D2U ◦ D1U
16. DIFFERENTIAL INVARIANTS. DIFFERENTIALLY SMOOTH MORPHISMS 387
from which we deduce, by (16.5.10.1), a canonical homomorphism (Bourbaki, Alg., chap. II, 3rd ed., IV-4 | 30
§5, no 3)
(16.5.10.2) GX/S ⊗OS OS′ −→ GX ′ /S′ ,
which is neither injective nor surjective in general. However:
Proposition (16.5.11). — (i) If g : S′ → S is a flat morphism and if f is locally of finite type (resp.
locally of finite presentation), then the homomorphism (16.5.10.2) is injective (resp. bijective).
(ii) If Ω1X/S is a locally free OX -module of finite type, then the homomorphism (16.5.10.2) is bijective.
P ROOF. The assertion (ii) follows from Bourbaki, Alg., chap. II, 3rd ed., §5, no 3, prop. 7. The
assertion (i) follows similarly from Bourbaki, Alg. Comm., chap. I, §2, no 10, prop. 11 and from the
fact that if f is locally of finite type (resp. locally of finite presentation), then Ω1X/S is an OX -module
of finite type (resp. of finite presentation) ((16.3.9) (16.4.22)). □
(16.5.12). Since Ω1X/S is a quasi-coherent OX -module, we can consider the vector bundle over X
defined by Ω1X/S (II, 1.7.8)
V( f ∗ (ΩY/S
1 1
)) = V(ΩY/S ) ×Y X (II, 1.7.11) ,
gives us an X-morphism TX/S ( f ) : TX/S → TY/S ×Y X. If g : Y → Z is a second S-morphism, we
have TX/S ( g ◦ f ) = ( TY/S ( g) × 1X ) ◦ TX/S ( f ) (0, 20.5.4.1).
It follows from (16.5.10.1) and from (II, 1.7.11) that for every base change g : S′ → S we have a
canonical isomorphism
(16.5.12.2) TX ′ /S′ ≃ TX/S ×S S′ = TX/S × X X ′ .
(16.5.13). For every point x ∈ X, we define the tangent space of X at the point x (relative to S) to be the
set of points in the fibre TX/S × X Spec(k( x )) that are rational over k ( x ), that is, the set
(16.5.13.1) TX/S ( x ) = Homk( x) (Ω1X/S ⊗Ox k( x ), k( x )),
which is the dual of the k( x )-vector space Ω1Ox /Os /mx · Ω1Ox /Os . When Ω1X/S is an OX -module of finite
type, then TX/S ( x ) is a vector space of finite rank over k ( x ), and for every base change g : S → S′ , IV-4 | 31
and every point x ′ ∈ X ′ = X ×S S′ over x, we have a canonical isomorphism
(16.5.13.2) TX ′ /S′ ( x ′ ) ≃ TX/S ⊗k( x) k ( x ′ ).
16. DIFFERENTIAL INVARIANTS. DIFFERENTIALLY SMOOTH MORPHISMS 388
If x is rational over k(s), where s = f ( x ) (so that k(s) → k( x ) is an isomorphism), it follows from
(16.4.13) that we have a canonical isomorphism
′
(16.5.13.3) TX/S ( x ) = TXs /k(s) ( x ) = Homk(s) (m′x /mx2 , k( x )),
where m′x is the maximal ideal of OXs ,x = OX,x /ms OX,x . In the case where S is the spectrum of a
field k, we recover the definition of the Zariski tangent space of a point x ∈ X rational over k, as the
dual of mx /m2x .
Let Y be a second S-prescheme and let g : Y → X be an S-morphism; then we have a canonical
homomorphism of OY -modules (16.4.19)
(16.5.13.4) gY/X/S : g∗ (Ω1X/S ) −→ ΩY/S
1
.
Now note that if y ∈ Y and x = g(y), then we have
g∗ (Ω1X/S ) ⊗OY k (y) = (Ω1X/S ⊗OX k( x )) ⊗k( x) k(y)
and we suggest looking for an S-morphism u : Y → X such that u0 = u ◦ j (in other words, if it is IV-4 | 32
possible to complete the diagram above by the dotted arrow u, keeping it commutative).
For that, consider an affine open U = Spec(C ) of Y; its inverse image j−1 (U ) is the affine open
U0 = Spec(C/L), where L = Γ(U, J ), a zero-square ideal in C; suppose that U is small enough so
that u0 (U0 ) is contained in an affine open V = Spec( B) of X and that g(U ) = f (u0 (U0 )) is contained
in an affine open W = Spec( A) of S, so that B and C are A-algebras and u0 |U0 corresponds to
an A-homomorphism ψ from B to C/L; Let P(U0 ) be the set of restrictions u|U of the sought
homomorphisms, which corresponds canonically to A-homomorphisms of algebras ϕ : B → C
ϕ
such that the composite B − → C → C/L is equal to ψ. So we know (0, 20.1.1) that the set of such
homomorphisms is either empty or of the form ϕ1 + Der A ( B, L); when P(U0 ) is not empty, the
additive group Der A ( B, L) acts by addition on P(U0 ), which is therefore an affine space for the
additive group Der A ( B, L) (or even a principal homogeneous space (or torsor) under Der A ( B, L)).
Now notice that, since L is equipped with a B-module structure via ψ, we have an isomor-
phism v 7→ v ◦ d B/A of HomB (Ω1B/A , L) onto Der A ( B, L) (0, 20.4.8). Besides, as L is square-zero,
therefore a (C/L)-module, every B-homomorphism v : Ω1B/A → L can be considered as a (C/L)-
homomorphism Ω1B/A ⊗ B (C/L) → L. As I is square-zero, it can be considered as a quasi-coherent
16. DIFFERENTIAL INVARIANTS. DIFFERENTIALLY SMOOTH MORPHISMS 389
Γ(U0′ , G ) × P(U0′ ) / P (U ′ )
h′ 0
is commutative (the vertical arrows being the restrictions). In light of the above remark, we reduce
to proving the commutativity of the above diagram when h is defined as such from affine opens V,
W and h′ from affine opens V ′ ⊂ V and W ′ ⊂ W. But because of the preceding description of h, this IV-4 | 33
follows from the commutativity of the diagram (0, 20.5.4.2).
The mapping Γ(U0 , G ) × P(U0 ) → P(U0 ) therefore define a homomorphism of sheaf of sets m :
G × P × P such that, for all open sets U0 for which Γ(U0 , P ) ̸= ∅, mU0 : Γ(U0 , G ) × Γ(U0 , P ) →
Γ(U0 , P ) is an external law defining in Γ(U0 , P ) a torsor structure for the group Γ(U0 , G ).
(16.5.15). In general, when we are given a sheaf of sets P over a topological space Z, a sheaf of
groups G (not necessarily commutative), and a homomorphism of sheaves of sets m : G × P → P
such that, for every open U ⊂ Z such that Γ(U, P ) ̸= ∅, mU : Γ(U, G ) × Γ(U, P ) → Γ(U, P )
makes Γ(U, P ) a torsor under the group Γ(U, G ), then we say that P is a pseudo-torsor (or formally
principal homogeneous sheaf ) under the sheaf of groups G . We say that P is a torsor (or principal
homogeneous sheaf ) under G 3 if in addition Γ(U, P ) ̸= ∅ for every open U ̸= ∅ in a suitable basis
for the topology of Z.
For the general theory of torsors, we refer to [eAG64]; we will limit ourselves to recalling the
canonical correspondence between isomorphism classes of torsors (for a given G ) and elements
from the cohomology set H1 ( Z, G ). Consider a torsor P under G and an open cover (Uλ ) of
Z such that Γ(Uλ , P ) ̸= ∅ for every λ; denote by pλ an element of Γ(Uλ , P ). For every pair
of indices λ, µ such that Uλ ∩ Uµ ̸= ∅, there then exists a unique element γλµ of Γ(Uλ ∩ Uµ , G )
such that γλµ · ( pµ |Uλ ∩ Uµ ) = pλ |Uλ ∩ Uµ ; in addition, if λ, µ, ν are three indices such that
Uλ ∩ Uµ ∩ Uν ̸= ∅, then the restrictions γλµ ′ , γ′ , γ′ of γ , γ , γ
µν λν λµ µν λν to Uλ ∩ Uµ ∩ Uν satisfy
′ ′ ′
the condition γλν = γλµ γµν ; in other words, (λ, µ) 7→ γλµ is a 1-cocycle of the cover (Uλ ) with
values in G . If, for every λ, p′λ is a second element of Γ(Uλ , P ), then there exists a unique element
β λ ∈ Γ(Uλ , G ) such that p′λ = β λ · pλ , and the 1-cocycle (γλµ
′ ) corresponding to the family ( p′ ) is
λ
given by γλµ′ = β γ β−1 , that is, it is cohomologous to γ . Conversely, the data of a 1-cocycle ( γ )
λ λµ µ λµ λµ
defines, for every pair (λ, µ), an automorphism θλµ of the sheaf of sets G |Uλ ∩ Uµ , namely the right
translation by γλµ , and the fact that it is a cocycle shows that we can glue the sheaves of sets G |Uλ
via the automorphisms θλµ (0, 3.3.1); we thus obtain a torsor under G , denoted P, and if we take for
pλ the unit section over Uλ , then the corresponding 1-cocycle is none other than the given 1-cocycle
(γλµ ); in addition, if we replace (γλµ ) by a 1-cocycle γλµ′ = β γ β−1 cohomologous to it, then we
λ λµ µ
check immediately that the torsor obtained is isomorphic to P.
In particular, if (γλµ ) is a 1-coboundary, in other words of the form γλµ = β λ β− 1
µ , then the torsor
P obtained is isomorphic to G (considered as a torsor under itself by left translations); we say in this
case that P is trivial, and the converse is evident.
In particular, it follows from (III, 1.3.1) that we have:
Proposition (16.5.16). — Let Z be an affine scheme, G a quasi-coherent OZ -module; then every torsor over
G is trivial.
3[Trans.] This is nowadays more commonly called a G -torsor rather then a torsor under G .
16. DIFFERENTIAL INVARIANTS. DIFFERENTIALLY SMOOTH MORPHISMS 390
In particular:
Corollary (16.5.18). — With the notation of (16.5.16), suppose that Y is affine and Ω1X/S is of finite
presentation; if there is a open cover (Uα ) of Y, and, for every index α, an S-morphism vα : Uα → X such
that, if Uα0 = j−1 (Uα ), we have vα ◦ ( j|Uα0 ) = u0 |Uα0 , then there is an S-morphism u : Y → X such that
u ◦ j = u0 .
P ROOF. Indeed, G is a quasi-coherent OY0 -module (I, 1.3.12); by (16.5.16) and the fact that Y0 is
then affine, the sheaf P, which is by hypothesis a torsor over G , and not only a pseudo-torsor, is
trivial; but if w is an isomorphism from G to P (as it is a torsor over G ), the image under w of the
zero section of G is the S-morphism we want. □
Theorem (16.6.2). — There is one and only one endomorphism d of the sheaf of additive groups Ω•X/S with
the following properties:
(i) d ◦ d = 0.
(ii) For every open set U of X and every section f ∈ Γ(U, OX ) we have d f = d X/S f . IV-4 | 35
′ p
(iii) For every open set U of X, every pair of integers p, q and every pair of sections ω p ∈ Γ(U, Ω X/S ),
q
ωq′′ ∈ Γ(U, Ω X/S ), we have
(16.6.2.1) d(ω ′p ∧ ωq′′ ) = (dω ′p ) ∧ ωq′′ + (−1) p ω ′p ∧ dωq′′ .
Also, d is an endomorphism of graded ψ∗ (OX )-modules of degree +1.
P ROOF. Suppose that we have proved the existence of an endomorphism d. For every affine
p
open U of X, every section of Ω X/S over U is (because of (ii)) a linear combination if a finite number
of elements of the form g(d f 1 ∧ d f 2 ∧ · · · ∧ d f p ), where g and the f i are sections of OX over U
(0, 20.4.7). The conditions (i) and (iii) then show, by induction on p, that we necessarily have
(16.6.2.2) d( g(d f 1 ∧ d f 2 ∧ · · · ∧ d f p )) = dg ∧ d f 1 ∧ d f 2 ∧ · · · ∧ d f p .
This therefore proves the uniqueness of d and the last claim of the theorem. By virtue of this
uniqueness property, to show the existence of d, we can restrict ourselves to the case where S =
Spec( A) and X = Spec( B) are affines. Now (Bourbaki, Alg., chap. III, 3rd ed., §10) to define an
16. DIFFERENTIAL INVARIANTS. DIFFERENTIALLY SMOOTH MORPHISMS 391
V
A-derivation D of degree +1 of an exterior algebra ( M ) (where M is a B-module and an A-
algebra), such derivation taking its values in a graded anti-commutative A-algebra C = ∞
L
n=0 Cn ,
whose elements of degree 1 are square-zero, it suffices to give arbitrarily an A-derivation D0 of B in
C1 and an A-homomorphism D1 of M in C2 ; then it exists one and only one A-anti-derivation D of
Λ( M) in C coinciding with D0 in B and D1 in M.
In the present case, D0 is necessarily equal to d B/A by (ii); we reduce to seeing, having (16.6.2.2)
in mind, that there is an A-homomorphism u of Ω1B/A in Ω2B/A such that
(16.6.2.3) u( g.d f ) = dg ∧ d f
for whichever f , g in A; it suffices to show that there is an A-homomorphism v : B ⊗ A Ω1B/A → Ω2B/A
such that
(16.6.2.4) v( g.ω ) = dg ∧ ω
for g ∈ B and ω ∈ Ω1B/A .Finally, since Ω1B/A
= I/I2 (where I = IB/A is the kernel of the canonical
homomorphism B × A B → B) and that Ω1B/A
is generated by elements of the form g.d f , it is enough
to define an A-homomorphism w : B ⊗ A ( B ⊗ A B) → Ω2B/A such that
(16.6.2.5) w( g′ ⊗ g ⊗ f ) = dg′ ∧ ( g.d f )
and such that w is zero on the image of B ⊗ AI2 . Or, since the second member of (16.6.2.5) is
A-trilinear in g′ , g and f , the existence of w verifying (16.6.2.5) is immediate. Since, on the other
hand, I is generated by elements of the form 1 ⊗ x − x ⊗ 1 (x ∈ B), we reduce to checking that when
z = (1 ⊗ x − x ⊗ 1)(1 ⊗ y − y ⊗ 1) we have w( g′ ⊗ z) = 0. Or, since z = 1 ⊗ ( xy) + ( xy) ⊗ 1 − x ⊗
y − y ⊗ x, the formula (16.6.2.4) shows that it is enough to see that we have d( xy) − x.dy − y.dx = 0, IV-4 | 36
which is to say that d is a derivation.
It remains to be shown that d verifies the condition (i). Now, the square of an anti-derivation is a
derivation (Bourbaki, loc. cit.), and since Ω•B/A is generated by Ω1B/A as a B-algebra, it is enough
to verify that d(dz) = 0 for z ∈ B and x ∈ Ω1B/A ; in the first case, this follows from the formula
(16.6.2.3) when g = 1; for the second, we can restrict ourselves to the case where z = g.d f with f , g
in B, and then we have, because of (16.6.2.1) and (16.6.2.3),
d(d( g.d f )) = d(dg ∧ d f ) = (d(dg)) ∧ (d f ) − (dg) ∧ (d(d f )) = 0.
□
Definition (16.6.3). — The anti-derivation d defined in (16.6.2) (also denoted by d X/S ) is called the
exterior differential on X (relative to S).
Proposition (16.6.4). — For every base change g : S′ → S, if we put X ′ = X ×S S′ , the canonical morphism
(16.6.4.1) Ω•X/S ⊗S S′ −→ Ω•X ′ /S′
deduced from the isomorphism (16.5.10.1) is bijective. Also, if s is a section of Ω•X/S over an open set U of
X, s ⊗ 1 its inverse image, section of Ω•X ′ /S′ over the inverse image U ′ of U in X ′ , we have d X ′ /S′ (s ⊗ 1) =
d X/S (s) ⊗ 1.
P ROOF. The first claim is immediate, the formation of the exterior algebra of a module commutes
with extending the scalar ring. To prove the second, we can, because of (16.6.2.2), restrict ourselves
to the case where s ∈ Γ(U, OX ), and in this case the claim has already been proven (16.4.3.7). □
(16.6.5). Suppose that Ω1X/S is an locally free OX -module of rank n in a point x, so that we have n
sections si ∈ Γ(U, OX ) such that the dsi form a basis for the Γ(U, OX )-module Γ(U, Ω1X/S ) (16.5.8).
Then, for every integer p ⩾ 1, the p-differentials dsi1 ∧ dsi2 ∧ · · · ∧ dsi p (for i1 ⩽ i2 ⩽ . . . ⩽ i p elements
p
of [1, n]) form a basis of (np) elements of Γ(U, Ω X/S ) over Γ(U, OX ). Also the formula (16.6.2.2) shows
that for every section g ∈ Γ(U, OX ), we have
∂g
(16.6.5.1) d( g.dsi1 ∧ dsi2 ∧ · · · ∧ dsi p ) = ∑(−1)r ∂sk dsi1 ∧ · · · ∧ dsir ∧ dsk ∧ dsir+1 ∧ · · · ∧ dsi p
k
where, in the second member, k varies in the set of the n − p indexes different from the ih , ir being
the biggest index < k.
16. DIFFERENTIAL INVARIANTS. DIFFERENTIALLY SMOOTH MORPHISMS 392
We note that the relation d(dg) = 0 for every section g ∈ Γ(U, OX ) expresses itself in the form
Di ( D j g ) = D j ( Di g ) for i ̸= j;
in other words, the derivations Di defined in (16.5.7) commute with each other.
16.7. The PX/S
n (F ).
(n)
(16.7.1). Let f : X → S be a morphism of preschemes, F an OX -module. We denote by X∆ the n’th
f
(n)
infinitesimal neighbourhood of X via the diagonal morphism ∆ f : X → X ×S X, by hn : X∆ → X ×S X IV-4 | 37
f
the canonical morphism (16.1.2), and consider the two composite morphisms
(n) (n) h
n p1 (n) n (n) h p2
p1 : X∆ −→ X ×S X −→ X, p2 : X∆ −→ X ×S X −→ X
f f
(n)
so that, by definition, p1 corresponds to the homomorphism of sheaves of rings OX → PX/S
n which
(n)
we have chosen to define the OX -algebra structure on PX/S
n (16.3.5), and p2 to the homomorphism
(n)
of sheaves of rings dnX/S : OX → PX/S
n (16.3.6). Since X∆ and X have the same underlying subspace,
f
we can write
(n) (n)
(16.7.1.1) n
PX/S = ( p1 )∗ (( p2 )∗ (OX )).
More generally, we define
(n) (n)
(16.7.1.2) n
PX/S (F ) = ( p1 )∗ (( p2 )∗ (F )).
so that PX/S
n = PX/S
n (O ); by definition, P n (F ) is an O -module.
X X/S X
(16.7.2). If we come back to the definition of the inverse image of modules on ringed spaces (0, 4.3.1)
(n)
and having in mind that X∆ and X have the same underlying space, we see that we can write the
f
definition (16.7.1.2) in the form
n n
(16.7.2.1) PX/S (F ) = PX/S ⊗O X F ,
but where you have to be careful that, in the interpretation of the symbol ⊗, PX/S
n is endowed with
the structure of OX -module defined by the homomorphism of sheaves of rings d X/S : OX → PX/S
n n . It
follows immediately from such formula (or directly from (16.7.1.2)) that PX/S (F ) is canonically
n
Proposition (16.7.3). — (i) The functor F 7→ PX/S n (F ) from the category of O -modules to the
X
category of PX/S -modules is right exact, and commutes with arbitrary inductive limits; it is exact
n
□
(16.7.4). The two structures of OX -module on PX/S n define in PX/S
n (F ) two structures of O -
X
modules, which happen to be permutable, and therefore an OX -bimodule structure. It is convenient
to denote the structure coming from the structure homomorphism OX → PX/S n (chosen in (16.3.5))
on the left and the one coming from the homomorphism d X/S : OX → PX/S on the right. On other
n n
words, for every open U of X, and every triplet a ∈ Γ(U, OX ), b ∈ Γ(U, PX/S
n ), t ∈ Γ (U, F ), we
have by definition
(16.7.4.1) a(b ⊗ t) = ( ab) ⊗ t, (b ⊗ t) a = (b.dn a) ⊗ t = b ⊗ ( at) = (dn a).(b ⊗ t).
The OX -module structure coming from the definition (16.7.1.2) is therefore, under these conventions,
the left OX -module structure. If F is a quasi-coherent OX -module, then the same is true for PX/S
n (F ) IV-4 | 38
for any one of its OX -module structures. If also F is of finite type (resp. of finite presentation) and
f : X → S is locally of finite type (resp. locally of finite presentation), PX/S
n (F ) is (for any one of its
16. DIFFERENTIAL INVARIANTS. DIFFERENTIALLY SMOOTH MORPHISMS 393
OX -module structures) of finite type (resp. of finite presentation), which is a consequence of (16.3.9)
and (16.4.22).
(16.7.5). The definition (16.7.2.1) entails the existence of a homomorphism of sheaves of commutative
groups
(16.7.5.1) dnX/S,F : F −→ PX/S
n
(F ) (also denoted dnX/S )
such that, in the notations of (16.7.4), we have
(16.7.5.2) dnX/S,F (t) = 1 ⊗ t
and consequently, because of (16.7.4.1)
(16.7.5.3) dnX/S,F ( at) = (1 ⊗ t) a = (dnX/S,F (t)).a
phism d X/S,F .
P ROOF. This is an immediate consequence of (16.7.5.3) and of the particular case F = OX
(16.3.8). □
(16.7.7). The canonical homomorphisms of sheaves of rings
m n
ϕnm : PX/S −→ PX/S
for n ⩽ m (16.1.2) define, because of (16.7.2.1), canonical homomorphisms
m n
PX/S (F ) −→ PX/S (F ) (n ⩽ m)
which are homomorphisms of OX -bimodules in light of (16.1.6) and (7.4.1); also we have commuta-
tive diagrams
PX/S
m (F ) / P n (F )
d : X/S
dnX/S,F dnX/S,F
F
We have therefore a projective system of OX -bimodules (PX/S
n (F )), and we define
∞ n
(16.7.7.1) PX/S (F ) = lim PX/S (F ).
←−
Also, this shows that the homomorphisms (16.7.5.1) form a projective system of homomorphisms,
and therefore define a canonical homomorphism
(16.7.7.2) d∞ ∞
X/S,F : F → PX/S (F ).
IV-4 | 39
(16.7.8). Let F , G be two OX -modules; it follows immediately from the definition (16.7.2.1) that we
have a canonical isomorphism of PX/Sn -modules
n n n
(16.7.8.1) PX/S (F ⊗OX G ) ≃ PX/S (F ) ⊗P n PX/S (G )
X/S
(Bourbaki, Alg., chap. II, 3rd ed., §5, n.1, prop. 3).
We conclude in particular (or we see directly from the definition (16.7.2.1)) that if F has an OX -
algebra structure (not necessarily associative), PX/S n (F ) has a canonical O -algebra structure; the
X
latter is associative (resp. commutative, res. unital, resp. a Lie algebra) if F is so. Also the canonical
homomorphisms PX/S m (F ) → P n (F ) for n ⩽ m (16.7.7) are then algebra di-homomorphisms;
X/S
similarly, (16.7.5.1) is then an OX -algebra homomorphisms when PX/S n (F ) is equipped with the
n n n
(16.7.8.2) PX/S (H omOX (F , G )) −→ H omP n (PX/S (F ), PX/S (G ))
X/S
16. DIFFERENTIAL INVARIANTS. DIFFERENTIALLY SMOOTH MORPHISMS 394
(Bourbaki, Alg., chap. II, 3rd ed., §5, n.3), which is bijective when PX/S
n is locally free of finite type (loc.
cit., prop. 7).
(16.7.9). Suppose we are in the situation described in (16.4.1); then from the canonical homomorphism
Pn (u) (16.4.3.3) we deduce immediately a canonical homomorphism of OX ′ -bimodules
(16.7.9.1) u∗ (PX/S
n
(F )) −→ PXn ′ /S′ (u∗ (F )).
We leave it to the reader to extend the properties seen in (16.4) in the case F = OX .
Remark (16.7.10). — The definition of PX/S
n (F ) in the form (16.7.1.2) still makes sense when F is
(n)
a sheaf of sets (the inverse image of a sheaf of sets by p2 being defined in (0, 3.7.1)); a variant of
this definition allows us to define the “jet schemes” (relatively to S) for any prescheme X.
4
16.8. Differential operators.
Definition (16.8.1). — Let f = (ψ, θ ) : X → S be a morphism of preschemes, F , G two OX -modules,
n an integer ⩾ 0. We say that a morphism D : F → G of sheaves of additive groups is a differential
operator of order ⩽ n (relative to S) if there is a homomorphism of OX -modules u : PX/S
n (F ) → G
(where PX/Sn (F ) is equipped with the structure of left O -modules (16.7.4)) such that we have
X
n
D = u ◦ d X/S,F .
4For a more general formalism, see the exposé VII of [eAG64] (due to P. Gabriel).
16. DIFFERENTIAL INVARIANTS. DIFFERENTIALLY SMOOTH MORPHISMS 395
(16.8.3). It is clear that the set of differential operators of order ⩽ n from F to G forms an ad-
ditive group, denoted by DiffnX/S (F , G ); when F = G = OX , we also write DiffnX/S instead of
DiffnX/S (OX , OX ).
We have seen (16.8.1), that given two open sets U ⊃ V of X, we have a restriction homomorphism
DiffnX/S (F |U, G |U ) −→ DiffnX/S (F |V, G |V )
n
from which we deduce that U 7→ DiffU/S (F |U, G |U ) is a presheaf of additive groups; in fact, it is IV-4 | 41
actually a sheaf, since for an open set U varying in X, the homomorphisms u 7→ u ◦ dU/S, n
F |U are
isomorphisms of sheaves of additive groups
n n
(16.8.3.1) HomOU (PU/S (F |U ), G |U ) ≃ DiffU/S (F |U, G |U ),
because of the fact that the image of F by dnX/S,F generates PX/S
n (F ) (16.7.6). We denote this sheaf
left OX -module structure on D iff nX/S (F , G ) explained as follows: for every open set U of X, every
section a ∈ Γ(U, OX ) and every differential operator D : F |U → G |U, aD is the differential operator
which, for every section t ∈ Γ(U, F ), makes correspond the section
(16.8.5.1) ( aD )(t) = a( D (t))
of Γ(U, G ). Similarly, the right OX -module structure on D iff nX/S (F , G ) is made explicit as fol-
lows: under the same notations as above, Da is the operator which, to every t ∈ Γ(U, F ), makes
correspond the section
(16.8.5.2) ( Da)(t) = D ( at).
Proposition (16.8.6). — If f : X → S is a morphism locally of finite presentation, F a quasi-coherent OX -
module of finite presentation and G a quasi-coherent OX -module, then D iff nX/S (F , G ) is a quasi-coherent
OX -module for any of the structures defined in (16.8.5).
P ROOF. The proposition follows from the fact that, under these hypothesis, PX/S
n is a quasi-
coherent OX -module of finite presentation (16.7.4) and of (I, 1.3.12) □
(16.8.7). The set of differential operators (of unspecified order (16.8.1)) is denoted by DiffX/S (F , G );
we also see as in (16.8.3) that U 7→ DiffU/S (F |U, G |U ) is a sheaf of additive groups, which we
will denote by D iff X/S (F , G ). It is immediate that D iff X/S (F , G ) is the reunion of the increasing
filtered family of its subsheaves D iff nX/S (F , G ); if X is quasi-compact, DiffX/S (F , G ) is similarly IV-4 | 42
the union of its subgroups DiffnX/S (F , G ) (16.8.1). The OX -bimodule structure on the D iff nX/S (F , G )
induce therefore an OX -bimodule structure on D iff X/S (F , G ), further explained in (16.8.5.1) and
(16.8.5.2).
Note that, for n ⩽ m, we have a commutative diagram
H omOX (PX/S
m (F ), G ) ∼ / D iff m (F , G )
X/S
where the horizontal arrows are the isomorphisms (16.8.4.1) and the horizontal arrow on the left
comes from the canonical morphism PX/Sm (F ) → P n (F ) (16.7.7). For every open set U of X, we
X/S
16. DIFFERENTIAL INVARIANTS. DIFFERENTIALLY SMOOTH MORPHISMS 396
then endow Γ(U, PX/S ∞ (F )) = lim Γ (U, P n (F )) of the projective limit topology of the discrete
←− X/S
n (F )), which defines on Γ (U, P ∞ (F )) a topological Γ (U, O )-bimodule
topologies on Γ(U, PX/S X/S X
∞ (F ) shows itself as a sheaf valued in the category of topological commutative
structure, so that PX/S
groups (0, 3.2.6). So [God58, II, 1.11], the limit of the inductive system of sheaves of commutative
groups (H omOX (PX/S n (F ), G )) is precisely the sheaf of continuous germs of homomorphisms from
∞ (F ) to G (the latter equipped with the discrete topology): the continuous homomorphisms
PX/S
Γ(U, PX/S∞ (F )) into the discrete group G indeed correspond bijectively to the inductive systems of
group homomorphisms Γ(U, PX/S n (F )) → Γ (U, G ). We can furthermore express (16.8.4) by saying
′
∑ (−1)Card( H ) ( ∏ ai ) Dan+1 (( ∏ ai )t) = 0.
H ⊂ In+1 i∈ H ′ / H′
i∈
But if we replace on this relation Dan+1 by the definition (16.8.8.1), we check immediately that we
have, up to sign, the first member of (16.8.8.2); from which we conclude. □
factors as
′
+n
dnX/S,F
F / P n + n ′ ⊗O F w / P n ′ ⊗O G
X/S X X/S X
dnX/S δ
PX/S
n / P n ′ (P n )
′ X/S X/S
dnX/S,P n
X/S
We will then have, indeed, a commutative diagram deduced from (16.8.9.3) by tensorization
with F
′
+n
dnX/S,F
F / P n + n ′ (F )
X/S
dnX/S,F δ ⊗1
PX/S
n (F ) / P n′ (P n ((F )))
′ X/S X/S
dnX/S,P n (F )
X/S
and on the other hand, we verify immediately the from definition (16.7.5) that the diagram
PX/S
n (F ) u /G
′ ′
dnX/S,P n dnX/S,G
(F )
X/S
n′ (P n ((F )))
PX/S / P n ′ (G )
X/S 1⊗ u X/S
□
P ROOF (16.8.9.3). It remains to prove the lemma (16.8.9.3). Considering (16.7.6), which proves
the uniqueness of δ, we are brought back to the case where S = Spec( A) and X = Spec( B) are
affines; letting I = IB/A , it suffices to define a canonical homomorphism of B-modules
′ ′
ϕ : ( B ⊗ A B)/In+n +1 −→ (( B ⊗ A B)/In +1 ) ⊗ B (( B ⊗ A B)/In+1 )
the B-module structure of the two members coming from the first B factor; recall that on tensor
′
product of the second member, ( B ⊗ A B)/In +1 must be considered as a right B-module by its IV-4 | 45
second B factor , and ( B ⊗ A B)/In+1 as a left B-module by its first B factor (16.7.2). It is the same to
define a homomorphism of B-modules
′
ϕ0 : B ⊗ A B −→ (( B ⊗ A B)/In +1 ) ⊗ B (( B ⊗ A B)/In+1 )
16. DIFFERENTIAL INVARIANTS. DIFFERENTIALLY SMOOTH MORPHISMS 398
′
and prove it is zero on In+n +1 . Now, we immediately define a homomorphism by the condition
that
ϕ0 (b ⊗ b′ ) = πn′ (b ⊗ 1) ⊗ πn (1 ⊗ b′ ) for b, b′ in B
under the notations of (16.3.7). Also, it is immediate that ϕ0 is a homomorphism of rings. Now, we
can write
ϕ0 (b ⊗ 1 − 1 ⊗ b) = πn′ (b ⊗ 1 − 1 ⊗ b) ⊗ πn (1 ⊗ 1) + πn′ (1 ⊗ b) ⊗ πn (1 ⊗ 1) − πn′ (1 ⊗ 1) ⊗ πn (1 ⊗ b)
and we have
πn′ (1 ⊗ b) ⊗ πn (1 ⊗ 1) = πn′ (1 ⊗ 1)b ⊗ πn (1 ⊗ 1) = πn′ (1 ⊗ 1) ⊗ bπn (1 ⊗ 1) = πn′ (1 ⊗ 1) ⊗ πn (b ⊗ 1)
from which, finally
(16.8.9.4) ϕ0 (b ⊗ 1 − 1 ⊗ b) = πn′ (b ⊗ 1 − 1 ⊗ b) ⊗ πn (1 ⊗ 1) + πn′ (1 ⊗ 1) ⊗ πn (b ⊗ 1 − 1 ⊗ b).
A product of n + n′ + 1 of terms of the form (16.8.9.4) is therefore necessarily zero, because the same
is true for the product of n + 1 terms of the form πn (b ⊗ 1 − 1 ⊗ b) and of n′ + 1 terms of the form
πn′ (b ⊗ 1 − 1 ⊗ b). The conclusion therefore results from (0, 20.4.4). □
Corollary (16.8.10). — The sheaf D iff X/S (OX , OX ) (also denoted D iff X/S ) is canonically endowed with
the structure of sheaf of rings, and the D iff nX/S form an increasing filtration compatible with such structure.
In particular, D iff 0X/S is a sheaf of subrings of D iff X/S , which is canonically identified with OX
(16.8.1). The formulas (16.8.5.1) and (16.8.5.2) show that the structure of OX -bimodule of D iff X/S
comes from the multiplication on the left and on the right by sections of OX considered as a sheaf of
subrings of D iff X/S .
Remarks (16.8.11). — (i) Suppose that F = ⊕λ∈ L Fλ ; then it is clear (16.7.2.1) that PX/S
n (F ) =
⊕λ∈ L PX/S
n (F ); since the functor U 7 → Γ (U, F ) commutes with the formation of arbitrary
λ
direct sums, dnX/S,F is the homomorphism whose restriction to each Fλ is dnX/S,F : Fλ →
λ
PX/S
n (F ); then we conclude immediately that we have
λ
DiffnX/S (F , G ) = ∏ DiffnX/S (F , Gµ ),
µ∈ M
(ii) So far, we have hardly encountered differential operators F → G where F and G are not
locally free of finite rank, in which case the structure is reduced locally, because of (i), to
the case of the sheaf D iff X/S ; the latter will be studied later (16.11) in a particular case.
16. DIFFERENTIAL INVARIANTS. DIFFERENTIALLY SMOOTH MORPHISMS 399
Corollary (16.9.5). — Let X be a locally ringed space, I a quasi-regular ideal of OX , ( f i )1⩽i⩽n a sequence
of sections of I over X, x a point of Supp(OX /I ). The following conditions are equivalent:
(a) There is a neighbourhood of x in X such that f i |U form a quasi-regular sequence of elements
Γ(U, OX ) generating I |U.
(b) The ( f i ) x form a system of generators of Ix whose size is as small as possible.
(b’) The ( f i ) x form a minimal set of generators of Ix .
(c) If f¯i is the canonical image of f i in Γ( X, I /I 2 ), the ( f¯i ) x form a basis for the (Ox /Ix )-module
Ix /Ix2 .
P ROOF. By hypothesis, Ox is a local ring, Ix an ideal of finite type of Ox contained in the
maximal ideal of Ox ; the equivalence of (b), (b’) and (c) results from Nakayama’s lemma (Bourbaki,
Alg. comm., chap. II, §3, no 2, prop. 5). It is clear that (a) implies (c) because of (16.9.3); on the other
hand, from (0, 5.2.2) it follows that if condition (c) is verified (and therefore so is (b)), there is a
neighbourhood U of x in X such that (I /I 2 )|U has constant rank equal to n, and that the f i |U
generate I |U; it suffices now to apply the last assertion of (16.9.3) to U. □
Remarks (16.9.6). — (i) Under the general hypothesis of (16.9.5), for the sequence ( f i ) to
generate I , it is not enough that the ( f¯i )y form a basis of the (Oy /Iy )-module (Iy /Iy2 )
for all y ∈ X. We have an example by taking X = Spec( A), where A is a Dedekind ring, IV-4 | 48
and I = e I, where I is a non-principal ideal of A; then indeed Iy /Iy2 = 0 in every point
y different from x ∈ X corresponding to I, and Ix /Ix2 has rank 1 over the field Ox /Ix ;
also, I is evidently a regular ideal.
(ii) In (16.9.5), one cannot replace “quasi-regular” by “regular”, even when X is a prescheme (cf.
(16.9.12)). Indeed, denote by B the ring of germs of infinitely differentiable functions on the
point 0 on R; it has a maximal ideal m generated by the germ of t of the identity mapping
to 0, and the intersection n of the mk for k > 0 is not reduced to 0. Now let A be the quotient
ring B[ T ]/nTB[ T ], and let f 1 , f 2 be the canonical images in A of the elements t and T of
B[ T ]. The sequence ( f 1 , f 2 ) is regular in A: indeed, f 1 is not a zero divisor in A, because
the relation tP[ T ] ∈ nTB[ T ], for a polynomial P ∈ B[ T ], implies that the products of t by
the coefficients of P belongs to the ideal n, and it results immediately that the coefficients
are the same in n, so P[ T ] ∈ nTB[ T ]. Since B/tB is isomorphic to R, A/ f 1 A is isomorphic
to the ring of polynomials R[ T ], therefore integral, and the image of f 2 in A/ f 1 A, being
equal to T, is not a zero divisor, so that our claim is true. However, f 2 is a zero divisor in A,
since for every non-zero element x ∈ n, the image of x in A is ̸= 0, but the image of xT is
zero. We conclude that the sequence ( f 2 , f 1 ) is not regular in A; on the other hand, the ideal
I = f 1 A + f 2 A is distinct from A, so the conditions (b), (b’) and (c) of (16.9.5) do not imply
the condition (a) when we replace “quasi-regular” by “regular”.
(16.9.7). If X = Spec( A) is an affine scheme, we’ll say that the ideal I of A is regular (resp. quasi-
regular) if the ideal J = e
I of OX is regular (resp. quasi-regular); we note that this notion is local and
does not imply the existence of a system of generators of I forming in A a regular (resp. quasi-regular)
sequence as the example (16.9.5) shows; however this is true if A is local (16.9.5).
The proposition (16.9.4) can be translated in terms of quasi-regular immersions in the following
manner:
Proposition (16.9.8). — Let j : Y → X be a morphism of preschemes; for j to be a quasi-regular immersion,
it is necessary and sufficient that j satisfies the following conditions:
(i) j is an immersion locally of finite presentation.
(ii) The conormal sheaf Gr1 ( j) = NY/X (16.1.2) is a locally free OY -module.
(iii) The canonical homomorphism
•
SO Y
(Gr1 ( j)) −→ Gr• ( j)
(16.1.2.2) is bijective.
P ROOF. The problem being local on Y, we can restrict ourselves to the case where j is the
canonical injection of a closed subprescheme Y of X, so the translation of (16.9.4) into (16.9.8) results
16. DIFFERENTIAL INVARIANTS. DIFFERENTIALLY SMOOTH MORPHISMS 401
from the description of Gr1 ( j) and Gr• ( j) in terms of the ideal I of OX defining the subprescheme Y
(16.1.3, (ii)). □
Corollary (16.9.11). — Let X be a locally Noetherian prescheme; then every quasi-regular ideal of OX is
regular.
Remarks (16.9.12). — (i) We note that a regular immersion is not in general a flat morphism,
and therefore a fortiori neither are quasi-regular morphisms in the sense of (6.8.1).
(ii) Let A be a local Noetherian ring; it follows immediately from (16.9.4) and from (0, 17.1.1)
that for A to be regular, it is necessary and sufficient that its maximal ideal m is quasi-regular
(or regular, which amounts to the same thing given that A is Noetherian). For an affine
Noetherian scheme X to be regular, it is necessary and sufficient that for every closed point
x ∈ X, the canonical injection Spec(k( x )) → X to be a regular immersion.
Proposition (16.9.13). — Let X be a locally Noetherian prescheme, Y a subprescheme of X, Y ′ a subprescheme
of Y, such that the canonical injection j : Y ′ → Y is regular. Then the sequence of OY ′ -modules
(16.9.13.1) 0 / j∗ (NY/X ) / NY ′ /X / NY ′ /Y /0
is exact; furthermore, for every x ∈ X, there is an open neighbourhood U of x such that the restrictions to U IV-4 | 50
of the homomorphisms of (16.9.13.1) form a split exact sequence.
Let us first prove the following lemma:
16. DIFFERENTIAL INVARIANTS. DIFFERENTIALLY SMOOTH MORPHISMS 402
is exact (see the demonstration of (16.2.7)), and the modules making up this sequence are precisely
the stalks in x of the sheaves of (16.9.13.1). The second claim follows from the fact that NX/Y is a
locally free OY ′ -module (16.9.8) and Bourbaki, Alg., chap. II, 3rd ed., §1, no 11, prop. 21. □
16. DIFFERENTIAL INVARIANTS. DIFFERENTIALLY SMOOTH MORPHISMS 403
We note that the integer n of the proposition (16.10.3) is the rank of Ω1X/S in the point x. We
shall see (17.13.5) that when f is differentially smooth and locally of finite type, n is equal to the
dimension of the fiber f −1 ( f ( x )).
Proposition (16.10.4). — Let f : X → S, g : S′ → S be two morphisms, and take X ′ = X ×S S′ ,
f ′ = f (S′ ) : X ′ → S′ .
(i) If f is differentially smooth, the same is true for f ′ .
(ii) Conversely, if g is faithfully flat and quasi-compact, and if f ′ is differentially smooth and Ω1X ′ /S′ is
a finite type OX ′ -module, f is differentially smooth and Ω1X/S is a finite type OX -module.
P ROOF. Indeed, if f is differentially smooth, the Grn (PX/S
n ) are flat O -modules; therefore by
X
(16.4.6), the homomorphism Grn (PX/S n )⊗
OX OX ′ → Grn (PX ′ /S′ ) are bijective for every n, because
n
of the commutative of the diagram (16.2.1.3), it follows from the definition (16.10.1) that f ′ is
16. DIFFERENTIAL INVARIANTS. DIFFERENTIALLY SMOOTH MORPHISMS 404
differentially smooth. On the other hand, if g is faithfully flat and quasi-compact, it follows also
from (16.4.6) that Grn (PX/Sn )⊗
OX OX ′ → Grn (PX ′ /S′ ) is bijective for every n. Suppose also that
n
f is differentially smooth and Ω X ′ /S′ is of finite rank. Since the canonical projection X ′ → X is a
′ 1
faithfully flat and quasi-compact morphism, it results first from (2.5.2) that Ω1X/S is an OX -module
locally free of finite rank, then from (2.2.7) that the canonical homomorphism (16.3.1.1) is bijective,
and therefore f is differentially smooth. □
Proposition (16.10.5). — For a morphism locally of finite type f : X → S to be differentially smooth, it is
necessary and sufficient that the diagonal immersion ∆ f : X → X ×S X to be quasi-regular.
IV-4 | 53
P ROOF. Being a local problem, we can reduce to the case where S and X are affines, and therefore
the diagonal subprescheme of X ×S X is closed. The hypothesis that f is locally of finite type implies
that ∆ f is locally of finite presentation (I, 4.3.1), therefore the diagonal prescheme of X ×S X is
defined by an ideal I of finite type, and Ω1X/S = I /I 2 is an OX -module of finite type. The
proposition is now immediate from the comparison of the conditions of (16.10.1) and (16.9.4). □
Remark (16.10.6). — Let f : X → S be a morphism such that the OX -module Ω1X/S is locally free
of finite rank. It results from (0, 20.4.7) that every x ∈ X has an open neighbourhood such that
there is a finite family (zλ )λ∈ L of sections of OX over U for which (dzλ )λ∈ L forms a basis of the
Γ(U, OX )-module Γ(U, Ω1X/S ).
(16.11.1.4) zp = ∏ ( zλ ) pλ , ζp = ∏ ( ζ λ ) pλ .
λ λ
We therefore have, with this notation
p
(16.11.1.5) m m
d (z ) = (d (z)) = (ζ + z)p =
p p
∑ q zp−q ζ q
q⩽p
p p−q m q
(16.11.1.6) ζ p = (dm z − z)p = ∑ (−1)|p−q| q
z d ( z ).
q⩽p
Since the dzλ generate Ω1X/S , and are the images of δzλ , and as the canonical morphism (16.3.1.1)
is surjective, we conclude that for finite m, the δzλ generate the OU -algebra PU/S n (Bourbaki, Alg.
comm., chap. III, §2, n 8, cor. 2 du th. 1). Therefore the ζ (for |p| ⩽ m) generate the OU -module
o p
n . A differential operator D ∈ Diffm
PU/S U/S is consequently entirely determined by the values of
p
⟨ζ , D ⟩for |p| ⩽ m, or, which amounts to the same by (16.11.1.5) and (16.11.1.6), by the values of the IV-4 | 54
⟨dm (z), D ⟩ = D (zp ) for |p| ⩽ m; more precisely, it follows from (16.11.5) that we have
p
(16.11.1.7) D ( z ) = ⟨ d ( z ), D ⟩ = ∑
p m p
⟨ζ q , D ⟩zp−q .
q⩽p
q
16. DIFFERENTIAL INVARIANTS. DIFFERENTIALLY SMOOTH MORPHISMS 405
Theorem (16.11.2). — Let f : X → S be a morphism, U an open set of X, (zλ )λ∈ L a family of sections
of OX over U such that the family (dzλ )λ∈ L generates Ω1X/S |U = ΩU/S 1 . The following conditions are
equivalent:
(a) f |U is differentially smooth and (dzλ ) is a basis of the OU -module ΩU/S 1 .
(b) There is a family ( Dp )p∈N( L) of differential operators of OU to itself verifying the conditions
q q−p
(16.11.2.1) Dp ( z q ) = z , (p, q in N( L) ).
p
Also, when these conditions are verified, the family ( Dp ) is uniquely determined by the conditions
(16.11.2.1) and satisfies the relations
(p + q)!
(16.11.2.2) Dq ◦ Dp = Dp ◦ Dq = Dp + q (p, q in N( L) ).
p!q!
Finally, if L is finite, for every integer m, the Dp such that |p| ⩽ m form a basis of the OU -module
D iff U/S
m , in other words, every differential operator of order ⩽ m on U can be written uniquely as
D= ∑ a p Dp
|p|⩽m
where the ap are sections of OX over U.
P ROOF. Note first that because of (16.11.1.6) and (16.11.1.5), we verify immediately that the
conditions (16.11.2.1) are equivalent to
(16.11.2.3) ⟨ζ p , Dq ⟩ = δpq (Kronecker’s symbol).
The existence of the family ( Dp ) verifying these conditions implies first (by taking |p| = 1) that
the dzλ are linearly independent, and therefore form a basis of the OU -module ΩU/S 1 . Then, for
p
every integer m ⩾ 1, we deduce similarly from (16.11.2.3) that the ζ such that |p| ⩽ m are linearly
independent; it follows that the canonical homomorphism (16.3.1.1) is injective, and therefore
bijective, which proves that (b) implies (a). The converse follows immediately from the definition
(16.10.1), the fact that ζ p form a basis of PU/S
n for |p| ⩽ m implies the existence and uniqueness of a
family of homomorphisms uq,m : PU/S → OU (|q| ⩽ m) such that ⟨ζ p , uq,m ⟩ = δp, q for |p| ⩽ m,
n
|q| ⩽ m. For a given value of q, the differential operators corresponding to uq,m for m ⩾ |q| are
identified with the same operator Dq . This proves that (a) implies (b), and also that the family ( Dq )
is uniquely determined, and that, if L is finite, for |p| ⩽ m, the Dp form a basis of the dual D iff U/S
m
of PU/S . Finally, the relations (16.11.2.2) follows immediately from the expression of the values of
n
the three operators considered on the zr , and of the fact that the ζ r for |r| ⩽ m generate PU/S
n . □
IV-4 | 55
Remarks (16.11.3). — (i) The fact that, because of (16.11.2.2), the Dp are pairwise permutable
naturally do not imply that the OU -algebra D iff U/S is commutative, the Dp do not com-
mute with the sections of OU unless n = 0.
(ii) The indices p such that |p| = 1 are the ϵ λ = (ϵλµ )µ∈ L where ϵλµ = 0 if µ ̸= λ and ϵλλ = 1;
when L is finite, the operators Dϵ λ are exactly the S-derivations Di introduced in (16.5.7).
We note that in general (contrary to what happens in classical analysis), it is not true that a
differential operator of any order can be written as a linear combination of powers of Di (cf.
(16.12)).
(iii) For every integer r ⩾ 1, we can define the notion of differentially smooth up to order r by
replacing in (16.10.1) the condition (ii) by the condition that the homomorphisms
m n
SO X
(Ω1X/S ) −→ Grm (PX/S )
are bijective for every m ⩽ r. The argument of (16.11.2) proves also that if, in the condition (a),
we replace “differentially smooth” by “differentially smooth up to order r”, this condition
is equivalent to (b) by restricting ourselves to p ∈ N( L) , q ∈ N( L) such that |p| ⩽ r, |q| ⩽ r.
17. SMOOTH MORPHISMS, UNRAMIFIED (OR NET) MORPHISMS, AND ÉTALE MORPHISMS. 406
16.12. Case of characteristic zero: Jacobian criterion for differentially smooth morphisms.
(16.12.1). We say that a prescheme X is of characteristic p (p equal to zero or a prime number) if,
for every affine open set U of X, the ring Γ(U, OX ) is of characteristic p (0, 21.1.1). It follows from
(0, 21.1.3) that for X to be of characteristic 0, it is necessary and sufficient that for every closed
point x of X, the residue field k ( x ) is of characteristic 0, or even that X can be given a structure of
Q-prescheme (necessarily unique).
P ROOF. The problem being local on X, we can suppose there is a family (zλ ) of sections of OX
over X such that the (dzλ ) is a basis for the OX -module Ω1X/S . Applying the criterion (16.11.2), it is
enough for the operators
Dp = (p!)−1 ∏ Dλλ
p
(where the Dλ are the coordinate forms corresponding to the basis (dzλ )) to verify the relations
(16.11.2.1), which is a consequence of the fact that the Dλ are derivations. □
(16.12.3). The theorem above is not true if we discard the hypothesis that X is of characteristic 0. For
example, if S = Spec(k), where k is a field of characteristic p > 0, X = Spec(K ) where K = k(α)
where α ∈ / k, α p ∈ k, we verify immediately that Ω1X/S has rank 1, and that the morphism X → S IV-4 | 56
has rank 1, and that the morphism X → S is differentially smooth up to order p − 1 (16.11.3, (iii)),
but not of order p. However, the proof of (16.12.2) proves that if Ω1X/S is locally free, and if n!1OX is
inversible in Γ( X, OX ), then X is differentially smooth over S up to order n.
17.1. Formally smooth morphisms, formally unramified morphisms, formally étale mor-
phisms.
induced by the canonical map Y0′ → Y ′ , is surjective (resp. injective, resp. bijective).
One also says that X is formally smooth (resp. formally unramified, resp. formally étale) over Y.
It is clear that for f to be formally étale, it is necessary and sufficient for f to be formally smooth
and formally unramified.
Remark (17.1.2). —
(i) Suppose that Y = Spec( A) and X = Spec( B) are affine, so that f comes from a homomor-
phism of rings ϕ : A → B. According to (0, 19.3.1) and (0, 19.10.1), saying that f is formally
smooth (resp. formally unramified, resp. formally étale) means that, via ϕ, B is a formally
smooth (resp. formally unramified, resp. formally étale) A-algebra, for the discrete topologies
on A and B.
17. SMOOTH MORPHISMS, UNRAMIFIED (OR NET) MORPHISMS, AND ÉTALE MORPHISMS. 407
(ii) To verify that f is formally smooth (resp. formally unramified, resp. formally étale), we
can, in Definition (17.1.1), restrict to the case where J 2 = 0. To see this, if f satisfies the
corresponding condition of Definition (17.1.1) in the particular case J 2 = 0, and if we have
J n = 0, then we consider the closed subscheme Yj′ of Y ′ defined by the sheaf of ideals
J j+1 for 0 ⩽ j ⩽ n − 1, so that Yj′ is a closed subscheme of Yj′+1 defined by a square-zero
sheaf of ideals; the hypotheses imply that each of the maps
HomY (Yj′+1 , X ) −→ HomY (Yj′ , X ) (0 ⩽ j ⩽ n − 1)
is surjective (resp. injective, resp. bijective); by composition, we conclude that the same IV | 57
holds for (17.1.1.1).
(iii) Note that the properties of the morphism f defined in (17.1.1) are properties of the repre-
sentable functor (0III , 8.1.8)
Y ′ 7−→ HomY (Y ′ , X )
from the category of Y-preschemes to the category of sets; they keep a meaning for any
contravariant functor with the same domain and codomain, representable or not.
(iv) Assume that the morphism f is formally unramified (resp. formally étale); consider an
arbitrary Y-prescheme Z and a closed subprescheme Z0 of Z defined by a locally nilpotent
sheaf of ideals J of OZ . Then the map
(17.1.2.1) HomY ( Z, X ) −→ HomY ( Z0 , X )
induced by the canonical injection Z0 → Z, is still injective (resp. bijective). To see this, let
(Uα ) be an affine open covering of Z such that the sheaves of ideals J |Uα are nilpotent,
and for each α, let Uα0 be the inverse image of Uα in Z0 , which is the closed subprescheme
of Uα defined by J |Uα . Let f 0 : Z0 → X by a Y-morphism; by hypothesis, for each α, there
is at most one (resp. one and only one) Y-morphism f α : Uα → X whose restriction to Z0
coincides with f 0 |Uα . We immediately conclude that if f α and f β are defined, then, for each
affine open V ⊂ Uα ∩ Uβ , we have f α |V = f β |V, as the restrictions of these morphisms to
the inverse image V0 of V in Z0 coincide. There is therefore at most one (resp. one and only
one) Y-morphism f : Z → X whose restriction to Z0 coincides with f 0 .
Proposition (17.1.3). —
(i) A monomorphism of preschemes is formally unramified; an open immersion is formally étale.
(ii) The composition of two formally smooth (resp. formally unramified, resp. formally étale) morphisms
is formally smooth (resp. formally unramified, resp. formally étale).
(iii) If f : X → Y is a formally smooth (resp. formally unramified, resp. formally étale) S-morphism,
then so is f (S′ ) : X(S′ ) → Y(S′ ) for any base extension S′ → S.
(iv) If f : X → X ′ and g : Y → Y ′ are two formally smooth (resp. formally unramified, resp. formally
étale) S-morphisms, then so is f ×S g : X ×S Y → X ′ ×S Y ′ .
(v) Let f : X → Y and g : Y → Z be two morphisms; if g ◦ f is formally unramified, then so is f .
(vi) If f : X → Y is a formally unramified morphism, then so is f red : Xred → Yred .
P ROOF. According to (I, 5.5.12), it suffices to prove (i), (ii), and (iii). The assertions in (i) are
both trivial. To prove (ii), consider two morphisms f : X → Y, g : Y → Z, an affine scheme Z ′ , a
closed subscheme Z0′ of Z defined by a nilpotent ideal and a morphism Z ′ → Z. Suppose that f and
g formally smooth, and consider a Z-morphism u0 : Z0′ → X; the hypothesis on g implies that there IV | 58
exists a Z-morphism v : Z ′ → Y such that f ◦ u0 = v ◦ j (where j : Z0′ → Z is the canonical injection);
the hypothesis on f then implies that there exists a morphism u : Z ′ → X such that f ◦ u = v and
u ◦ j = u0 , therefore ( g ◦ f ) ◦ u is equal to the given morphism Z ′ → Z and u ◦ j = u0 , which proves
that g ◦ f is formally smooth; we argue the same way when we suppose that f and g are formally
unramified.
Finally, to prove (iii), let X ′ = XS′ , Y ′ = YS′ , f ′ = f S′ ; consider an affine scheme Y ′′ , a closed
subscheme Y0′′ defined by a nilpotent sheaf of ideals, and a morphism g : Y ′′ → Y ′ making
Y ′′ a Y ′ -prescheme; we then know by (I, 3.3.8) that HomY ′ (Y ′′ , X ′ ) is canonically identified with
17. SMOOTH MORPHISMS, UNRAMIFIED (OR NET) MORPHISMS, AND ÉTALE MORPHISMS. 408
HomY (Y ′′ , X ), and HomY ′ (Y0′′ , X ′ ) with HomY (Y0′′ , X ), and the conclusion follows immediately
from Definition (17.1.1). □
We note that a closed immersion is not necessarily formally smooth.
Proposition (17.1.4). — Let f : X → Y and g : Y → Z be two morphisms, and suppose that g is formally
unramified. Then, if g ◦ f is formally smooth (resp. formally étale), so is f .
P ROOF. Let Y ′ be an affine scheme, Y0′ a closed subscheme of Y ′ defined by a nilpotent sheaf of
ideals, h : Y ′ → Y a morphism, j : Y0′ → Y ′ the canonical injection, u0 : Y0′ → Y a Y-morphism, such
that f ◦ u0 = h ◦ j. Suppose that g ◦ f is formally smooth; then there exists a morphism u : Y ′ → X
such that u ◦ j = u0 and ( g ◦ f ) ◦ u = g ◦ h. But these two relations imply that f ◦ u and h are
Z-morphisms from Y ′ to Y such that ( f ◦ u) ◦ j = h ◦ j; by virtue of the hypothesis that g is formally
unramified, we get that f ◦ u = h, in other words that u is a Y-morphism; thus f is formally smooth.
Taking into account (17.1.3, (v)), this proves the proposition. □
Corollary (17.1.5). — Suppose that g is formally étale; then, for g ◦ f to be formally smooth (resp. formally
unramified, resp. formally étale), it is necessary and sufficient that f is.
P ROOF. This follows from (17.1.4) and (17.1.3, (ii) and (iv)). □
Proposition (17.1.6). — Let f : X → Y be a morphism of preschemes.
(i) Let (Uα ) be an open covering of X and, for each α, let iα : Uα → X be the canonical injection. For f
to be formally smooth (resp. formally unramified, resp. formally étale), it is necessary and sufficient
that each f ◦ iα is.
(ii) Let (Vλ ) be an open covering of Y. For f to be formally smooth (resp. formally unramified, resp.
formally étale), it is necessary and sufficient that each of the restrictions f −1 (Vλ ) → Vλ of f is.
P ROOF. First note that (ii) is a consequence of (i): if jλ : Vλ → Y and iλ : f −1 (Vλ ) → X are the
canonical injections, then the restriction f λ : f −1 (Vλ ) → Vλ of f is such that jλ ◦ f λ = f ◦ iλ ; if f
is formally smooth (resp. formally unramified), then so is f ◦ iλ since iλ is formally étale (17.1.3);
but since jλ is formally étale, this means that f λ is formally smooth (resp. formally unramified), by
virtue of (17.1.5). Conversely, if all the f λ are formally smooth (resp. formally unramified), the same
applies to jλ ◦ f λ (17.1.3), so also to f in virtue of (i).
If we take into account that the iα are formally étale, everything comes down to proving that if IV | 59
the f ◦ iα are formally smooth (resp. formally unramified), then the same applies to f .
Therefore let Y ′ be an affine scheme, Y0′ a closed subscheme of Y ′ defined by a nilpotent ideal
J , which we may assume to satisfy J 2 = 0 (17.1.2, (ii)), and finally let g : Y ′ → Y be a morphism.
Suppose we are given a Y-morphism u0 : Y0′ → X; denote by Wα (resp. Wα0 ) the prescheme induced
by Y ′ (resp. Y0′ ) on the open subset u0−1 (Uα ) (we recall that Y ′ and Y0′ share the same underlying
topological space). Let us first suppose that the f ◦ iα are formally unramified, and show that, if u′ and u′′
are two Y-morphisms from Y ′ to X whose restrictions to Y0′ coincide, then we have u′ = u′′ . Indeed,
taking into account (17.1.2, (iv)), the hypothesis that the f ◦ iα are formally unramified implies that
for all α, we have u′ |Wα = u′′ |Wα , since the restrictions of both Y-morphisms to Wα0 coincide. Hence
the conclusion follows.
Now suppose that the f ◦ iα are formally smooth and prove the existence of a Y-morphism
u : Y ′ → X whose restriction to Y0′ is u0 . Now, since Y ′ is an affine scheme, we can apply (16.5.17),
the hypotheses of which are satisfied, and the conclusion of which precisely proves the existence of
u. □
We can therefore say that the notions introduced in (17.1.1) are local on X and Y, which always
allows, in virtue of (17.1.2, (i)), to be reduced to the study of formally smooth (resp. formally
unramified, resp. formally étale) algebras.
17. SMOOTH MORPHISMS, UNRAMIFIED (OR NET) MORPHISMS, AND ÉTALE MORPHISMS. 409
P ROOF. Taking into account (17.1.6), we reduce to the case where Y = Spec( A) and X = Spec( B)
are affine, and the conclusion then follows from (0, 20.7.4) and the interpretation of Ω1X/Y in this
case (16.3.7). □
Corollary (17.2.2). — Let f : X → Y and g : Y → Z be two morphisms. For f being formally unramified,
it is necessary and sufficient that the canonical morphism (16.4.19)
f ∗ (ΩY/Z
1
) −→ Ω1X/Z
is surjective.
P ROOF. This is an immediate consequence of (17.2.1) and the exact sequence (16.4.19.1). □
Proposition (17.2.3). — Let f : X → Y be a formally smooth morphism.
(i) The OX -module Ω1X/Y is locally projective (16.10.1). If f is locally of finite type, then Ω1X/Y is
locally free and of finite type.
(ii) For all morphisms g : Y → Z, the sequence (16.4.19) of OX -modules
(17.2.3.1) 0 −→ f ∗ (ΩY/Z
1
) −→ Ω1X/Z −→ Ω1X/Y −→ 0
is exact; moreover, for each x ∈ X, there exists an open neighbourhood U of x such that the
restrictions to U of the homomorphisms in (17.2.3.1) form a split exact sequence.
IV | 60
P ROOF.
(i) We know (16.3.9) that if f is locally of finite type, then Ω1f is an OX -module of finite type.
To prove that, in all cases, it is locally projective, we can reduce, by virtue of (17.1.6), to
the case where Y = Spec( A) and X = Spec( B) are affine, and the result follows from the
hypothesis on f and from (0, 20.4.9) and (0, 19.2.1).
(ii) Again, we can restrict to the case where X, Y, and Z are affine (17.1.6), and the conclusion in
this case follows from the interpretation of the sheaves of modules in the sequence (17.2.3.1)
and from (0, 20.5.7).
□
Corollary (17.2.4). — If f : X → Y is formally étale, then, for all morphisms g : Y → Z, the canonical
homomorphism of OX -modules
f ∗ (ΩY/Z
1
) −→ Ω1X/Z
is bijective.
P ROOF. This follows from the exactness of the sequence (17.2.3.1) and from the fact that we then
have Ω1X/Y = 0 (17.2.1). □
Proposition (17.2.5). — Let f : X → Y be a morphism, X ′ a subprescheme of X such that the composite
j f
morphism X ′ −→ X − → Y (where j is the canonical injection) is formally smooth. Then the sequence of
OX -modules (16.4.21)
(17.2.5.1) 0 −→ NX ′ /X −→ Ω1X/Y ⊗OX OX ′ −→ Ω1X ′ /Y −→ 0
is exact; moreover, for each x ∈ X, there exists an open neighbourhood U of x such that the restrictions to U
of the homomorphisms in (17.2.5.1) form a split exact sequence.
P ROOF. By virtue of (17.1.6), we reduce to the case where Y = Spec( A) and X = Spec( B) are
affine, and X ′ = Spec( B/J), where J is an ideal of B. The conormal sheaf NX ′ /X then corresponds
to the B-module J/J2 (16.1.3), and the conclusion follows from (0, 20.5.14). □
Proposition (17.2.6). — Let X and Y be two preschemes, f : X → Y a morphism locally of finite type. The
following conditions are equivalent:
17. SMOOTH MORPHISMS, UNRAMIFIED (OR NET) MORPHISMS, AND ÉTALE MORPHISMS. 410
(a) f is a monomorphism.
(b) f is radicial and formally unramified.
(c) For each y ∈ Y, the fibre f −1 (y) is empty or k(y)-isomorphic to Spec(k (y)) (in other words, it is
reduced to a single point z such that k (y) → Oz /my Oz is an isomorphism).
P ROOF. The fact that (a) implies (c) follows from (8.11.5.1). It is clear that (c) implies that f is
radicial; let us prove that it also follows from (c) that Ω1X/Y = 0, which will prove that (c) implies
(b) (17.2.1). Note that the OX -module Ω1X/Y is quasi-coherent of finite type (16.3.9). It follows from
(I, 9.1.13.1) that, for (Ω1X/Y ) x = 0, it is necessary and sufficient that if we set Y1 = Spec(k(y)),
X1 = f −1 (y) = X ×Y Y1 , then we have (Ω1X /Y ) x = 0; but as the morphism f 1 : X1 → Y1
1 1
induced by f is formally unramified by virtue of the hypothesis (c) (17.1.3), the conclusion follows
from (17.2.1). Finally, let us prove that (b) implies (a); for this, consider the diagonal morphism
g = ∆ f : X → X ×Y X; since f is radicial, g is surjective (1.8.7.1); on the other hand, Ω1X/Y is
by definition the conormal sheaf Gr1 ( g) of the immersion g (16.3.1), and to say that f is formally
unramified therefore means that Gr1 ( g) = 0 (17.2.1). In addition, g is locally of finite presentation IV | 61
(1.4.3.1); therefore the hypothesis Gr1 ( g) = 0 implies that g is an open immersion (16.1.10); being
surjective, this immersion is an isomorphism, hence f is a monomorphism (I, 5.3.8). □
17.3. Smooth morphisms, unramified morphisms, étale morphisms
Definition (17.3.1). — We say that a morphism f : X → Y is smooth (resp. unramified, or net 5 resp.
étale) if it is locally of finite presentation and formally smooth (resp. formally unramified, resp.
formally étale).
We then also say that X is smooth (resp. unramified, resp. étale) over Y.
We will see later (17.5.2) that this definition of a smooth morphism coincides with the definition
already given in (6.8.1); until then, we will exclusively use definition (17.3.1).
It is clear that saying that f is étale means that it is both smooth and unramified.
Remark (17.3.2). —
(i) Note that definition (17.3.1) can be phrased using only the functor
Y ′ 7−→ HomY (Y ′ , X )
considered in (17.1.2, (iii)) because to say that f is locally of finite presentation is equivalent
to saying that the preceding functor commutes with projective limits of affine schemes (8.14.2).
(ii) Let A be a ring and B an A-algebra. We say that B is a smooth (resp. unramified, resp. étale)
A-algebra if the corresponding morphism Spec( B) → Spec( A) is smooth (resp. unramified,
resp. étale). It is equivalent to say that B is an A-algebra of finite presentation (1.4.6) that
is furthermore formally smooth (resp. formally unramified, resp. formally étale) for the
discrete topologies.
(iii) It follows from (17.1.6) and the definition of a morphism locally of finite presentation (1.4.2)
that the notion of a smooth (resp. unramified, resp. étale) morphism is local on X and on Y.
Proposition (17.3.3). —
(i) An open immersion is étale. For an immersion to be unramified, it is necessary and sufficient to it be
locally of finite presentation.
(ii) The composition of two smooth (resp. unramified, resp. étale) morphisms is smooth (resp. unramified,
resp. étale).
(iii) If f : X → Y is a smooth (resp. unramified, resp. étale) S-morphism, then so is f (S′ ) : X(S′ ) → Y(S′ )
for any base extension S′ → S.
(iv) If f : X → X ′ and g : Y → Y ′ are smooth (resp. unramified, resp. étale) S-morphisms, then so is
f ×S g : X ×S Y → X ′ ×S Y ′ . IV | 62
5The words “net” and “formally net” seem more preferable to the terminology used in “unramified” (resp. formally
unramified”) and will be used almost exclusively in Chapter V. In this chapter, we have kept the old terminology so as not to
conflict with 0, 19.10.
17. SMOOTH MORPHISMS, UNRAMIFIED (OR NET) MORPHISMS, AND ÉTALE MORPHISMS. 411
dom(u). When F is strictly torsion-free, such that F is identified by (20.1.5.1) with a subsheaf of
MX (F ), then saying u is defined on U means that u|V is a section of F over U.
(20.1.8). In accordance with the general notation of (0I , 5.4.7), we denote by MX∗ the sheaf of
multiplicative groups such that Γ(U, MX∗ ) is (for every open U of X) the group of invertible ele-
ments of Γ(U, MX ). This sheaf is none other than the sheaf S (MX ) defined in (20.1.3): indeed, if
s ∈ Γ(U, S (MX )), then for every x ∈ U, there exists an open neighbourhood V ⊂ U of x such that
s|V is a regular element in the total ring of fractions of Γ(V, OX ), and we know that such an element is
necessarily invertible in this ring of fractions. We say that the sections of MX∗ over X are the regular
meromorphic functions (note that we are deviating here from the terminology followed by certain
authors, who call “regular” meromorphic functions those which are sections of OX , identified with a
subsheaf of MX ).
Let L be an invertible OX -module (0I , 5.4.1); then it is clear that MX (L ) = L ⊗OX MX is an IV-4 | 229
invertible MX -module. Let U be an open such that L |U is isomorphic to OU ; as every automorphism
of MU is multiplication by an invertible element of Γ(U, MX ) (0I , 5.4.7), it is equivalent to say that
a section s ∈ Γ(U, MX (L )) has an invertible image in Γ(U, MX ) under an isomorphism or by any
isomorphism on Γ(U, MX ); we say in this case that s is a regular meromorphic section of L over U; a
section s of L over X is called a regular meromorphic section of L if, for every open U such that L |U is
isomorphic to OU , s|U is a regular meromorphic section of L over U. We denote by (MX (L ))∗ the
subsheaf of MX (L ) such that for every open U, Γ(U, (MX (L ))∗ ) is the set of regular meromorphic
sections of L over U. Let s be a meromorphic section of L over X (i.e. a section of MX (L )); it
defines a homomorphism hs : MX → MX (L ) which sends every section t of MX over an open U to
(s|U )t. It follows immediately from the above that for s to be regular, it is necessary and sufficient
for hs to be injective, and in fact hs is then a bijective homomorphism from MX to MX (L ), and
its restriction to MX∗ is a bijection to (MX (L ))∗ . We conclude that the homothety t 7→ ts is an
isomorphism from M ( X ) to M ( X, L ).
(20.1.9). Let s be a regular meromorphic section of an invertible OX -module L over X; then for every
OX -module F , s similarly defines a homomorphism hs ⊗ 1F : MX (F ) → MX (F ⊗OX L ), which
is again bijective.
(20.1.10). Let s be a meromorphic section of an invertible OX -module L over X; for s to be regular,
it is necessary and sufficient for there to exists a meromorphic section s′ of L −1 over X such that
the canonical image of s ⊗ s′ in MX (0I , 5.4.3) is the unit section, and this section s′ is then unique:
indeed, the necessity of the local existence of such a section is evident, and its local uniqueness
implies its global (and unique) existence; moreover, the existence of s′ is trivially sufficient for s to
be regular. We will take s′ = s−1 .
Finally, if L ′ is a second invertible OX -module, s (resp. s′ ) a regular meromorphic section of L
(resp. L ′ ) over X, then s ⊗ s′ is evidently a regular meromorphic section of L ⊗ L ′ over X.
(20.1.11). If f : X ′ → X is a morphism of ringed spaces, then there is in general no natural map
sending a meromorphic function on X to a meromorphic function on X’. For example, if X is the
spectrum of a local integral domain A, X ′ its residue field k, then there is no natural homomorphism
from the field of fractions K of A to k, and we can only send an element of K to an element of k if it is
already in A.
In general, if f = (ψ, θ ), then for every open U of X, denote by S f (U ) the set of regular sections
s ∈ Γ(U, OX ) such that the image of s under
Γ(θ # ) : Γ(U, OX ) −→ Γ( f −1 (U ), OX ′ )
is a regular section. It is immediate that U 7→ S f (U ) is a subsheaf of the sheaf of sets S (OX ), which
we denote by S f . We set M f = OX [S f−1 ]; this is a subsheaf of rings of MX , and we canonically IV-4 | 230
obtain from θ# ψ ∗ (O
: X ) → OX ′ a homomorphism of sheaves of rings : θ ′# ψ ∗ (M
f ) → MX ′
extending θ (Bourbaki, Alg. comm., chap. II, §2, n 1, prop. 2); hence, recalling that f ∗ (M f ) =
# o
(20.1.11.1) f ∗ (M f ) −→ MX ′ .
20. MEROMORPHIC FUNCTIONS AND PSEUDO-MORPHISMS 415
For every meromorphic function ϕ on X that is a section of M f , Γ(θ ′# )(ϕ) is a meromorphic function
on X ′ , called the inverse image of ϕ under f , and denoted by ϕ ◦ f is there is no cause for confusion.
Similarly, if F is an OX -module, then we set M f (F ) = F ⊗OX M f , and we immediately obtain
from θ ′# a canonical homomorphism (which is also written as u 7→ u ◦ f )
Γ( X, M f (F )) −→ Γ( X ′ , MX ′ ( f ∗ (F ))).
In addition, if u ∈ Γ( X, M f (F )) is defined (20.1.7) at a point x, then u coincides, on a neighbourhood
U of x, with a section of the form ∑i hi ⊗ (ti /si ), where the hi belong to Γ(U, F ), the ti to Γ(U, OX ),
and the si to Γ(U, S f ). As by hypothesis the images of the si in Γ( f −1 (U ), OX ′ ) are regular, we see
that u ◦ f is defined at every point of f −1 (U ); in other words, we have
(20.1.11.2) f −1 (dom(u)) ⊂ dom(u ◦ f ).
We will see later (20.6.5, (i)) examples (with F = OX ) where the two sides of (20.1.11.2) can be
different.
Consider in particular the case where M f = MX ; then, if L is an invertible OX -module, the
image in MX ′ ( f ∗ (L )), under Γ(θ ′# ), of a regular meromorphic section of L over X (20.1.8) is a
regular meromorphic section of f ∗ (L ) over X ′ , as it follows immediately from the definition of
its sections, and from the fact that a homomorphism of rings sends an invertible element to an
invertible element.
Let f ′ : X ′′ → X ′ be a second morphism of ringed spaces, and suppose that M f = MX and
M f ′ = MX ′ ; then, if we set f ′′ = f ◦ f ′ , we also have M f ′′ = MX , and we immediately see that for
every meromorphic section u of F over X, we have u ◦ f ′′ = (u ◦ f ) ◦ f ′ .
Proposition (20.1.12). — If the morphism f : X ′ → X is flat (0I , 6.7.1), then we have M f = MX , and the
homomorphism ϕ 7→ ϕ ◦ f is defined on all of M ( X ). In addition, if f is a (flat) morphism of locally ringed
spaces, then we have dom(ϕ ◦ f ) = f −1 (dom(ϕ)); if in addition f is surjective (thus faithfully flat), then
the homomorphism ϕ 7→ ϕ ◦ f is injective.
P ROOF. The first assertion follows from the fact that, if B is an A-algebra which is a flat A-
module, then every element of A not a divisor of 0 in A is not a divisor of 0 in B (0I , 6.3.4). To prove
the other assertions, note that, for every x ′ ∈ X ′ , if x = f ( x ′ ), then OX ′ ,x′ is a flat OX,x -module, and
as the homomorpism OX,x → OX ′ ,x′ is local by hypothesis, it is injective ((0I , 6.5.1) and (0I , 6.6.2)); if
we set A = OX,x , B = OX ′ ,x′ , such that A identifies with a subring of B, then ( f ∗ (MX )) x′ is equal to
S−1 A ⊗ A B = S−1 B, where S is the set of regular elements of A, (MX ′ ) x′ is equal to T −1 B, where T is
the set of regular elements of B, and as we have seen that S ⊂ T, the homomorphism S−1 B → T −1 B IV-4 | 231
is injective; in other words, this proves that the homomorphism (20.1.11.1) f ∗ (MX ) → MX ′ is
injective (hence the last assertion of the statement). The quotient f ∗ (MX )/OX ′ identifies with an
OX ′ -submodule of MX ′ /OX ′ , and ( f ∗ (MX )/OX ′ ) x′ identifies with (MX /OX ) x ⊗OX,x OX ′ ,x′ . Then
suppose that x ̸∈ dom(ϕ); the image of ϕx in (MX /OX ) x is therefore ̸= 0; by faithful flatness, we
deduce that that it is the same for the image of (ϕ ◦ f ) x′ , so x ′ ̸∈ dom(ϕ ◦ f ), which finishes the
proof. □
Remark (20.1.13). — Let X be a reduced complex analytic space; then the notion of a meromorphic
function on X defined above coincides with the usual notion. Consider on the other hand a
prescheme Y, locally of finite type over the field C; we then know that we can associate to Y an
analytic space Y an having the same underlying topological space, and the canonical morphism
f : Y an → Y is flat [?]; by virtue of (20.1.12), the canonical homomorphism u 7→ u ◦ f from M (Y ) to
M(Y an ) is therefore always defined and is injective; but it is not surjective in general. For example,
when Y = V0r (ErrIII , 14) is the affine space of dimension r over C, M(Y ) canonically identifies with
the field R(Y ) of rational functions on Y (20.2.13, (i)), while M (Y an ) is the field of usual meromorphic
functions on Cr . Because of this fact, it is often preferable, in algebraic geometry, to abstain from the
terminology introduced in this section, and to use the equivalent terminology of “pseudo-function”
which will be defined below.
21. DIVISORS 416
§21. D IVISORS
Bibliography
[Car] H. Cartan, Séminaire de l’école normale supérieure, 13th year (1960–61), exposé no 11.
[CC] H. Cartan and C. Chevalley, Séminaire de l’école normale supérieure, 8th year (1955–56), géométrie algébrique.
[CE56] H. Cartan and S. Eilenberg, Homological Algebra, Princeton Math. Series (Princeton University Press), 1956.
[Che51] C. Chevalley, Introduction to the theory of algebraic functions of one variable, Math. Surveys, vol. 6, Amer. Math. Soc.,
New York, 1951.
[CI58] W. L. Chow and J. Igusa, Cohomology theory of varieties over rings, Proc. Nat. Acad. Sci. U.S.A. XLIV (1958), 1244–1248.
[eAG64] M. Demazure et A. Grothendieck, Schémas en groupes, 1963–1964.
[God58] R. Godement, Théorie des faisceaux, Actual. Scient. et Ind., no. 1252, Paris (Hermann), 1958.
[Gra60] H. Grauert, Ein Theorem der analytischen Garbentheorie und die Moldulräume komplexer Strukturen, Publ. Math. Inst.
Hautes Études Scient., no. 5, 1960.
[Gro] A. Grothendieck, Géométrie formelle et géométrie algébrique, Séminaire Bourbaki, 11th year (1958–59), exposé 182.
[Gro57] , Sur quelques points d’algèbre homologique, Tôhoku Math. Journ. IX (1957), 119–221.
[Gro58] , Cohomology theory of abstract algebraic varieties, Proc. Intern. Congress of Math., Edinburgh (1958), 103–118.
[Jaf60] P. Jaffard, Les systèmes d’idéaux, Paris (Dunod), 1960.
[Käh58] E. Kähler, Geometria Arithmetica, Ann. di Mat. (4) XLV (1958), 1–368.
[Nag55] M. Nagata, On the derived normal rings of noetherian integral domains, Mem. Coll. Sci. Kyoto A, XXIX (1955), 293–303.
[Nag58a] , A general theory of algebraic geometry over Dedekind domains, Amer. Math. Journ. I: LXXVIII, II: LXXX (1956,
1958), 78–116, 382–420.
[Nag58b] , Existence theorems for non projective complete algebraic varieties, Ill. J. Math II (1958), 490–498.
[Nag62] , Local rings, Interscience Tracts, vol. 13, Interscience, New York, 1962.
[Nor53] D. G. Northcott, Ideal theory, Cambridge Univ. Press, 1953.
[Sam53a] P. Samuel, Algèbre locale, Mém. Sci. Math., no. 123, Paris, 1953.
[Sam53b] , Commutative algebra (notes by D. Herzig), 1953.
[Ser56] J.-P. Serre, Géométrie algébrique and géométrie analytique, Ann. Inst. Fourier VI (1955–56), 1–42.
[Ser55a] , Faisceaux algébriques cohérents, Ann. of Math. LXI (1955), 197–278.
[Ser55b] , Sur la dimension homologique des anneaux et des modules Noethériens, Proc. Intern. Symp. on Alg. Number
theory, Tokyo–Nikko (1955), 176–189.
[Ser57] , Sur la cohomologie des variétés algébriques, Journ. of Math. (9) XXXVI (1957), 1–16.
[SZ60] P. Samuel and O. Zariski, Commutative algebra, 2 vol., New York (Van Nostrand), 1958–60.
[Wei46] A. Weil, Foundations of algebraic geometry, Amer. Math. Soc. Coll. Publ., no. 29, 1946.
[Wei49] , Numbers of solutions of equations in finite fields, Bull. Amer. Math. Soc. LV (1949), 497–508.
[Zar47] O. Zariski, A new proof of Hilbert’s Nullstellensatz, Bull. Amer. Math. Soc. LIII (1947), 362–368.
[Zar51] , Theory and applications of holomorphic functions on algebraic varieties over arbitrary ground fields, Mem. Amer.
Math. Soc., no. 5, 1951.
417