10248
10248
10248
https://ebookfinal.com
https://ebookfinal.com/download/quadratic-
mappings-and-clifford-algebras-1st-edition-
jacques-helmstetter/
https://ebookfinal.com/download/clifford-algebras-applications-to-
mathematics-physics-and-engineering-1st-edition-carlos-a-berenstein/
ebookfinal.com
https://ebookfinal.com/download/handbook-of-conformal-mappings-and-
applications-1st-edition-prem-k-kythe/
ebookfinal.com
https://ebookfinal.com/download/hyperbolic-manifolds-and-holomorphic-
mappings-an-introduction-second-edition-shoshichi-kobayashi/
ebookfinal.com
Reflections on America by Jacques Maritain Jacques
Maritain
https://ebookfinal.com/download/reflections-on-america-by-jacques-
maritain-jacques-maritain/
ebookfinal.com
https://ebookfinal.com/download/re-mapping-archaeology-critical-
perspectives-alternative-mappings-mark-gillings/
ebookfinal.com
https://ebookfinal.com/download/monomial-algebras-second-edition-
rafael-villarreal/
ebookfinal.com
https://ebookfinal.com/download/neuroradiology-cases-1st-edition-
clifford-j-eskey/
ebookfinal.com
Quadratic mappings and Clifford algebras 1st Edition
Jacques Helmstetter Digital Instant Download
Author(s): Jacques Helmstetter, Artibano Micali
ISBN(s): 9783764386061, 3764386061
Edition: 1
File Details: PDF, 3.18 MB
Year: 2008
Language: english
Jacques Helmstetter
Artibano Micali
Quadratic Mappings
and Clifford Algebras
Birkhäuser
Basel · Boston · Berlin
Authors:
Bibliographic information published by Die Deutsche Bibliothek. Die Deutsche Bibliothek lists
this publication in the Deutsche Nationalbibliografie; detailed bibliographic data is available in
the Internet at http://dnb.ddb.de
This work is subject to copyright. All rights are reserved, whether the whole or part of the
material is concerned, specifically the rights of translation, reprinting, re-use of illustrations,
recitation, broadcasting, reproduction on microfilms or in other ways, and storage in data
banks. For any kind of use permission of the copyright owner must be obtained.
987654321 www.birkhauser.ch
Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ix
1 Algebraic Preliminaries
1.1 Some general notation and definitions . . . . . . . . . . . . . . . . 1
1.2 Universal objects in a category . . . . . . . . . . . . . . . . . . . . 2
1.3 Examples of universal objects . . . . . . . . . . . . . . . . . . . . . 4
1.4 Tensor algebras and symmetric algebras . . . . . . . . . . . . . . . 8
1.5 Functors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.6 Exact sequences . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.7 Projective modules and flat modules . . . . . . . . . . . . . . . . . 14
1.8 Finitely presented modules . . . . . . . . . . . . . . . . . . . . . . 15
1.9 Changes of basic rings . . . . . . . . . . . . . . . . . . . . . . . . . 17
1.10 Rings and modules of fractions . . . . . . . . . . . . . . . . . . . . 21
1.11 Localization and globalization . . . . . . . . . . . . . . . . . . . . . 25
1.12 Finitely generated modules . . . . . . . . . . . . . . . . . . . . . . 30
1.13 Some applications . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
2 Quadratic Mappings
2.1 Generalities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
2.2 Changes of basic ring and localizations . . . . . . . . . . . . . . . . 57
2.3 Nondegenerate quadratic mappings . . . . . . . . . . . . . . . . . . 59
2.4 Operations on quadratic mappings and symmetric
bilinear mappings . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
2.5 Hyperbolic and metabolic spaces . . . . . . . . . . . . . . . . . . . 67
2.6 Orthogonal decompositions of quadratic spaces . . . . . . . . . . . 72
2.7 Witt rings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
2.8 Examples of Witt rings . . . . . . . . . . . . . . . . . . . . . . . . 83
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
vi Contents
3 Clifford Algebras
3.1 Definitions and elementary properties . . . . . . . . . . . . . . . . 105
3.2 The parity grading of Clifford algebras . . . . . . . . . . . . . . . . 112
3.3 Clifford algebras of free modules of rank 2 . . . . . . . . . . . . . . 117
3.4 Graded quadratic extensions . . . . . . . . . . . . . . . . . . . . . . 122
3.5 Graded Azumaya algebras . . . . . . . . . . . . . . . . . . . . . . . 131
3.6 Traces and determinants . . . . . . . . . . . . . . . . . . . . . . . . 144
3.7 Clifford algebras of quadratic spaces . . . . . . . . . . . . . . . . . 149
3.8 Discriminant modules, quadratic extensions and
quaternion algebras . . . . . . . . . . . . . . . . . . . . . . . . . . . 154
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164
4 Comultiplications. Exponentials. Deformations
4.1 Coalgebras and comodules . . . . . . . . . . . . . . . . . . . . . . . 175
4.2 Algebras and coalgebras graded by parities . . . . . . . . . . . . . 181
4.3 Exterior algebras . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185
4.4 Interior products in Clifford algebras . . . . . . . . . . . . . . . . . 190
4.5 Exponentials in even exterior subalgebras . . . . . . . . . . . . . . 195
4.6 Systems of divided powers . . . . . . . . . . . . . . . . . . . . . . . 199
4.7 Deformations of Clifford algebras . . . . . . . . . . . . . . . . . . . 201
4.8 Applications of deformations . . . . . . . . . . . . . . . . . . . . . 210
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219
5 Orthogonal Groups and Lipschitz Groups
5.1 Twisted inner automorphisms and orthogonal groups . . . . . . . . 232
5.2 Filtrations of Clifford algebras . . . . . . . . . . . . . . . . . . . . . 243
5.3 Lipschitz monoids and derived groups . . . . . . . . . . . . . . . . 249
5.4 The invariance property . . . . . . . . . . . . . . . . . . . . . . . . 258
5.5 Associated Lie algebras . . . . . . . . . . . . . . . . . . . . . . . . 261
5.6 First results about orthogonal transformations . . . . . . . . . . . 267
5.7 Products of reflections when K is a local ring . . . . . . . . . . . . 272
5.8 Further results about orthogonal transformations . . . . . . . . . . 281
5.9 More information about exterior algebras . . . . . . . . . . . . . . 287
5.10 The Lipschitz monoid Lip(M ) when K is a field . . . . . . . . . . . 296
Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 303
6 Further Algebraic Developments
6.1 Modules over a noncommutative algebra . . . . . . . . . . . . . . . 323
6.2 Graded modules over a graded algebra . . . . . . . . . . . . . . . . 327
6.3 Graded semi-simple modules . . . . . . . . . . . . . . . . . . . . . 334
6.4 Graded Morita theory . . . . . . . . . . . . . . . . . . . . . . . . . 337
6.5 Graded separable algebras . . . . . . . . . . . . . . . . . . . . . . . 344
Contents vii
ercises of Chapter 6 propose a smart and effective path to the essential algebraic
properties that are needed in quantum mechanics.
This comment about spinor spaces is just one example of the constant em-
phasis put on parity gradings (from Chapter 3 to the end), in full agreement with
C.T.C. Wall and H. Bass. In many cases the reversion of two odd factors must
be compensated by a multiplication by −1, and here this rule is systematically
enforced in all contexts in which it is relevant; indeed only a systematic treatment
of parity gradings can avoid repeated hesitations about such multiplications by
−1. For instance if f and g are
2 linear forms on M , their exterior product can be
defined as the linear form on (M ) that takes this value on the exterior product
of two elements x and y of M :
the sign − before f (x)g(y) comes from the reversion of the odd factors g and x;
but in f (y)g(x) the odd factor y has jumped over two odd factors x and g, whence
the sign +.
Weyl algebras
Weyl algebras represent for alternate bilinear forms the same structure as Clifford
algebras for quadratic forms, and in some publications they are even called “sym-
plectic Clifford algebras”. In [Dixmier 1968] you can find a concise exposition of
what was known about them before cliffordian mathematicians became interested
in them. Revoy was probably the first to propose a cliffordian treatment of Weyl
algebras; see [Nouazé, Revoy 1972] and [Revoy 1978]. Later and independently,
Crumeyrolle in France and the Japanese team Sato–Miwa–Jimbo produced some
publications developing the cliffordian treatment of Weyl algebras, although they
ignored (at least in their first publications) that these algebras had been already
studied, and had received H. Weyl’s name. Revoy’s isolated work was hardly no-
ticed, the cliffordian ideas of the Japanese team (which the renewal of Lipschitz’s
ideas mentioned above) were inserted in a very long and difficult work devoted to
differential operators, which discouraged many people, and Crumeyrolle’s state-
ments bumped up against severe and serious objections. That is why the cliffordian
treatment of Weyl algebras has not yet won complete acknowledgement.
Introduction xiii
Acknowledgements
A. Micali is pleased to recall that he has worked for a long time with Orlando E.
Villamayor (1923–1998), who efficiently contributed with some other authors to
much progress that is now common knowledge and presented as such in this book.
J. Helmstetter declares that he is indebted to Chevalley for his interest in Clifford
algebras. Both authors have also learned much from H. Bass’s publications. For
some difficult topics we have also consulted [Knus 1991].
During the writing of our text, we took advantage of the services of the Math-
ematical Department (Fourier Institute) of the University of Grenoble; our text
was prepared in this Institute and its Library was very often visited. Therefore we
are grateful to the Director and to the Librarian for their help. Several colleagues
in this Institute suggested some good ideas, or tried to answer embarrassing ques-
tions; they also deserve our gratitude.
The authors also thank the Director and the Librarian of the Mathemati-
cal Department of the University of Montpellier, who always gave us a helpful
welcome. Our colleague Philippe Revoy followed our work and gave much good
advice, for which we are indebted to him.
The authors are also grateful to Max-Albert Knus (Zürich, ETH-Zentrum)
for his critical reading of a large part of the book, and for his judicious suggestions.
Finally we thank all the persons that have worked on our text to improve its
literary quality and its presentation; despite our efforts, we yet needed their help.
Without the benevolence of the editorial board of Birkhäuser Verlag, our project
would not have been completed.
Grenoble, September 2007.
Algebraic Preliminaries
This preliminary chapter is devoted to the following three subjects, which to-
gether allows us to review a great part of the prerequisites, and to add some more
specialized knowledge:
(a) a very simple presentation of the notion of universal property, with many
examples; Sections 1.2 to 1.4 are devoted to this subject.
(b) additional information about categories of modules; Sections 1.5 to 1.9 con-
tain reminders about exact sequences, usual functors, projective modules and
changes of basic rings; but in 1.8 finitely presented modules are treated with
more detail.
(c) a self-contained presentation of rings and modules of fractions, localization
and globalization; although this material is already well treated in the existing
literature, a concise exposition of the exact part that is here actually useful
should prevent beginners from wandering in too-specialized topics.
Complete knowledge of all this chapter is not indispensable, because precise refer-
ences will always be given when the most difficult or specialized results are needed
in the following chapters.
required that the equality h(gf ) = (hg)f is true whenever it is meaningful (in other
words, whenever the targets of f and g are respectively the sources of g and h). In
the definition of a category is also mentioned the existence of an identity morphism
idN for each object N , with the requirement that the equalities idN f = f and
g idN = g hold whenever they are meaningful. A morphism f : M → N is called
an isomorphism if there exists g : N → M such that gf = idM and f g = idN ;
this morphism g is unique and is called the reciprocal isomorphism.
An object U in a category C is called an initial universal object (resp. a
final universal object) if for every object M in C the set HomC (U, M ) (resp.
HomC (M, U )) contains exactly one element. This definition implies that idU is
the only element in HomC (U, U ). Here almost all universal objects under consid-
eration will be initial ones. We must keep in mind the following evident theorem.
(1.2.1) Theorem. If a category contains two initial universal objects (resp. two final
universal objects) U and V , the only morphism from U to V is an isomorphism.
Proof. There is one morphism f from U to V and also one morphism g from V to
U ; since gf (resp. f g) is a morphism from U (resp. V ) to itself, it must be equal
to idU (resp. idV ); therefore f and g are reciprocal isomorphisms.
The most evident category is the category of all sets in which any mapping
is a morphism; the empty set is the only initial universal object, and any set
containing exactly one element is a final universal object.
In the category Mod(K) of all modules over K, the morphisms are the K-
linear mappings; the notation Hom(M, N ) (or HomK (M, N ) when more precision
is necessary) is the usual abbreviation for HomMod(K) (M, N ). In this category any
module reduced to zero is both an initial universal object and a final one.
Now let us consider the category Alg(K) of all algebras over the ring K, in
which the morphisms are the K-linear ring morphisms. The algebras reduced to
zero are the only ones in which the unit element is equal to 0, and they are final
universal objects. Since the morphisms must respect the unit elements (see 1.1),
the ring K itself is an initial universal object. The category Alg(Z) is the category
of all rings; for any ring A, the only morphism Z → A determines the characteristic
of A : it is the integer n ≥ 0 such that nZ is the kernel of this morphism.
Before more interesting examples are given in the next two sections, it must
be recalled that sometimes the objects of a category are morphisms in another
category. In particular, with each category C is associated a category Chom in
which the objects are all the morphisms of C; if u : M1 → M2 and v : N1 → N2
are two objects of Chom , a morphism from u to v is a couple (f1 , f2 ) of morphisms
f1 : M1 → N1 and f2 : M2 → N2 such that vf1 = f2 u. The composition of two
morphisms (f1 , f2 ) and (g1 , g2 ) in Chom is given by the formula (g1 , g2 ) ◦ (f1 , f2 ) =
(g1 f1 , g2 f2 ) whenever it is meaningful. If U is an initial universal object in C, then
idU is an initial universal object in Chom .
Sometimes it is convenient to associate with a given category C a dual category
C ∗ containing the same objects and the same morphisms; but the target (resp. the
4 Chapter 1. Algebraic Preliminaries
Quotient modules
Let M be a module over K, and N a submodule of M ; when the quotient module
M/N and the quotient morphism ϕ : M → M/N are involved in an argument,
the following proposition is often referred to.
(1.3.1) Proposition. For every linear mapping f : M → P such that f (N ) = 0,
there exists a unique linear mapping f : M/N → P such that f = f ϕ. Moreover
if M is a K-algebra and N an ideal of M , then M/N is also a K-algebra and ϕ
is an algebra morphism.
This property of M/N and ϕ is said to be a universal property because it
means that ϕ is an initial universal object in the category of all linear mappings f
from M into any module P such that f (N ) = 0. If f : M → P and g : M → Q are
objects in this category, a morphism from f to g is a linear mapping u : P → Q
such that g = uf . When M is a K-algebra and N an ideal, we may require f and u
to be algebra morphisms. The proof of (1.3.1) is based on the fact that the image
and kernel of ϕ are exactly M/N and N ; therefore these properties characterize
M/N and ϕ up to isomorphy.
The null morphism M → 0 with a trivial target is obviously a final uni-
versal object, but quite uninteresting. In the following examples such trivial final
universal objects will not even be mentioned.
The objects of the category under consideration are the bilinear mappings f
defined on M × N ; if f : M × N → P and g : M × N → Q are two objects, a
morphism from f to g is a linear mapping u : P → Q such that g = uf .
The precise definition of M ⊗ N is very seldom needed, since it is character-
ized by (1.3.2) up to isomorphy. For instance (1.3.2) is sufficient to explain that
two linear mappings v : M → M and w : N → N determine a linear mapping
v ⊗ w from M ⊗ M into N ⊗ N ; indeed v ⊗ w is the only linear mapping such
that (v ⊗ w)(x ⊗ y) = v(x) ⊗ w(y) for all (x, y) ∈ M × N .
Of course it is possible to state an analogous universal property for a ten-
sor product of several modules. The so-called commutativity property of tensor
products refers to the evident isomorphisms M ⊗ N ←→ N ⊗ M . There is also an
associativity property:
(M ⊗K N ) ⊗L P ∼
= M ⊗K (N ⊗L P );
For a family of two modules M and N the direct product and the direct sum
coincide; but the notation M ⊕ N is generally preferred when it is the source of
a linear mapping, whereas M × N is preferred when it is the source of a bilinear
mapping or the target of a mapping of any kind.
Obviously the categories C and C have been derived from the category C =
Mod(K) and the family (Mj ) in such a way that no special properties of Mod(K)
have been needed; consequently it is possible to repeat the same construction of
C and C by starting from any category C and any family (Mj ) of objects of C.
If C contains a final universal object, its source is called the direct product of the
family (Mj ) in C; and if C contains an initial universal object, its target is called
its direct sum in C.
When the objects of C are sets provided with some common structure, and
when its morphisms are the mappings respecting this structure, it often happens
that every cartesian product of objects is still an object, that the canonical pro-
jections from this cartesian product onto its components are morphisms in C, and
that finally this cartesian product is the direct product in C. This is true for the
category of all sets, the category Alg(K) and its subcategory Com(K).
The concepts of direct product and direct sum may be understood as dual
to each other if we remember the dual category C ∗ defined in 1.2. Indeed if we
replace C with C ∗ in the construction of C (without changing the family (Mj )),
we get the dual category of C ; and conversely. Therefore a direct sum in C is a
direct product in C ∗ , and conversely.
Despite this duality, the existence of direct sums is often a more difficult prob-
lem than the existence of direct products when the objects of C are sets provided
with some structure as above. Besides the previous example with C = Mod(K),
we will also look for direct sums of two objects in two other categories. First, in the
category of all sets, the direct sum of two sets M and N is the so-called disjoint
union. When M and N are actually disjoint, it is merely M ∪ N ; if not, it may be
the union of two disjoints sets M and N respectively isomorphic to M and N .
Secondly the direct sum of two commutative algebras A and B in Com(K)
is their tensor product A ⊗ B ; indeed (1.3.3) implies that there is a canonical
bijection
HomCom(K) (A ⊗ B, P ) −→ HomCom(K) (A, P ) × HomCom(K) (B, P ).
8 Chapter 1. Algebraic Preliminaries
T0 (M ) = K , T1 (M ) = M , T2 (M ) = M ⊗M , T3 (M ) = M ⊗M ⊗M, . . . ;
because of the so-called associativity of tensor products, for all (i, j) ∈ N2 there
is a canonical isomorphism from Ti (M ) ⊗ Tj (M ) onto Ti+j (M ), whence a mul-
tiplication mapping Ti (M ) × Tj (M ) −→ Ti+j (M ); this multiplication can be
extended by linearity to the direct sum T(M ) of all the tensor powers Tk (M ), and
thus T(M ) is an associative algebra (even a graded algebra). The next proposition
means that the natural injection ϕ : M → T(M ) is the desired universal object.
(1.4.1) Proposition. Any linear mapping f from M into an algebra P extends
in a unique way to an algebra morphism f : T(M ) → P . For every sequence
(x1 , x2 , . . . , xk ) of elements of M (of any length k ≥ 1),
f (x1 ∨ x2 ∨ · · · ∨ xk ) = f (x1 , x2 , . . . , xk )
for all x1 , x2 , . . . , xk ∈ M .
Proof. We suppose k ≥ 2, otherwise the symmetry hypothesis is empty. First f
induces a linear mapping f : Tk (M ) → P . As above, we treat Sk (M ) as the
quotient Tk (M )/Rk , where Rk is spanned by the products
1.5 Functors
Let C and D be two categories. A covariant functor F from C to D associates
with each object M of C an object F (M ) of D, and with each morphism f of C a
morphism F (f ) of D in such a way that the following three conditions are always
fulfilled:
if f ∈ HomC (M, N ) , then F (f ) ∈ HomD (F (M ), F (N )) ;
F (gf ) = F (g)F (f ) whenever gf exists;
if f is an identity morphism, then so is F (f ).
A contravariant functor F from C to D is a covariant functor from the dual category
of C (see 1.2) to the category D; this means that:
if f ∈ HomC (N, M ) , then F (f ) ∈ HomD (F (M ), F (N )) ;
F (f g) = F (g)F (f ) whenever f g exists;
if f is an identity morphism, then so is F (f ).
Many universal properties give rise to a functor. For instance there is a func-
tor T associated with tensor algebras; it is a covariant functor from Mod(K) to
Alg(K); the tensor algebra T(M ) of a module M has been defined in 1.4; if f
is a linear mapping from M into N , then T(f ) is the unique algebra morphism
T(M ) → T(N ) extending the linear mapping M → T(N ) defined by x −→ f (x).
It is easy to verify that a functor has been defined in this way. In the same way we
can define a functor S from Mod(K) to Com(K) by means of symmetric algebras.
The covariant functors between two categories are themselves the objects of
a category. When F and G are covariant functors from C to D, a morphism Φ from
F to G associates with each object M of C a morphism Φ(M ) from F (M ) to G(M )
in D, in such a way that G(f ) ◦ Φ(M ) = Φ(N ) ◦ F (f ) for every f ∈ HomC (M, N ).
This morphism Φ : F −→ G is an isomorphism if and only if all morphisms Φ(M )
are isomorphisms.
Let F be a covariant functor from C to D, and M and N two objects of C;
we assume that C contains morphisms ϕ1 : M → P and ϕ2 : N → P for which P
is the direct sum of M and N in this category, and that D contains morphisms
ψ1 : F (M ) → Q and ψ2 : F (N ) → Q for which Q is the direct sum of F (M ) and
F (N ) in this category. Now F (ϕ1 ) and F (ϕ2 ) have the same sources as ψ1 and
ψ2 , but their common target is F (P ); since (ψ1 , ψ2 ) is universal, all this results in
a morphism Q → F (P ). In other words, the existence of direct sums in C and in
D implies the existence of canonical morphisms
F (M ) ⊕D F (N ) −→ F (M ⊕C N ).
Some functors F have the nice property that these canonical morphisms are iso-
morphisms. For instance the functor S defined by means of symmetric algebras
has this property, as stated in the next theorem; remember that the direct sum of
two objects in Com(K) is their tensor product (see 1.3).
1.5. Functors 11
because of the universal property of tensor products, we derive from it the recip-
rocal isomorphism.
12 Chapter 1. Algebraic Preliminaries
M −→ M −→ M ;
u v
0 −→ M −→M −→M −→ 0
u v
(1.6.1)
0 −→ M −→M −→M −→ 0,
u v
u v
0 ←− M ←−M ←−M ←− 0.
Let F be an additive functor from the category Mod(K) into itself; the
additiveness of F means that for every couple of modules (M, N ) the mapping
u −→ F (u) is a group morphism from Hom(M, N ) into Hom(F (M ), F (N )) if F
is covariant, into Hom(F (N ), F (M )) if F is contravariant. With this hypothesis,
any equality vu = 0 implies F (v)F (u) = 0 or F (u)F (v) = 0 ; but this is not
sufficient to conclude that every exact sequence is transformed by F into an exact
sequence; F is called an exact functor if it transforms every exact sequence into
an exact sequence. It is not difficult to prove that F is exact if and only if it
transforms every exact sequence like (1.6.1) into an exact sequence.
Every splitting exact sequence is transformed into a splitting exact sequence,
because four mappings u, v, u , v satisfying the five equalities (1.6.2) are trans-
formed by F into four mappings satisfying the analogous five equalities that prove
the exactness and the splitting of the transformed sequence. Therefore if a functor
is not exact, its lack of exactness can be observed only on exact sequences that do
not split.
Unfortunately the most usual additive functors Hom and ⊗ are not exact for
all rings K; the former is only left exact (for both variables), and the latter is right
exact. This means that for all modules P and all exact sequences
0 −→ M −→ M −→ M ,
N −→ N −→ N −→ 0 ,
From the right exactness of the functor ⊗ the following statement can be
immediately derived: the exactness of the two sequences
M −→ M −→ M −→ 0 and N −→ N −→ N −→ 0
(M ⊗ N ) ⊕ (M ⊗ N ) −→ M ⊗ N −→ M ⊗ N −→ 0; (1.6.3)
(P ⊗ N ) + (P ⊗ N ) = P ⊗ (N + N ),
(P ⊗ N ) ∩ (P ⊗ N ) = P ⊗ (N ∩ N );
indeed the former equality is obvious, and the latter one follows from it because
of the exact sequence
0 −→ N ∩ N −→ N ⊕ N −→ N + N −→ 0
in which the second arrow is c −→ (c, −c) and the third one is (a, b) −→ a + b.
Obviously every inclusion N ⊂ N in M gives a similar inclusion in P ⊗ M . When
P is faithfully flat, the inclusion N ⊂ N is equivalent to P ⊗ N ⊂ P ⊗ N . Indeed
these inclusions are respectively equivalent to the surjectiveness of the mappings
N → N + N and P ⊗ N → P ⊗ (N + N ).
The finitely generated projective modules afford the most convenient frame to
generalize the classical properties of vector spaces of finite dimension. For instance
we may associate with each module M the dual module M ∗ = Hom(M, K), and
there is a canonical mapping from M into the bidual module (M ∗ )∗ ; if P is finitely
generated and projective, the mapping P → (P ∗ )∗ is an isomorphism; indeed this
is obviously true when P is a free module with a finite basis; when P is not free,
there exists a module P such that P ⊕ P is free with a finite basis; whence an
isomorphism P ⊕ P → ((P ⊕ P )∗ )∗ ; but (P ⊕ P )∗ is canonically isomorphic to
P ∗ ⊕ P ∗ , and finally we get an isomorphism
P ⊕ P −→ (P ∗ )∗ ⊕ (P ∗ )∗ ;
since this isomorphism maps each of the two components on the left side into
the corresponding component on the right side, it gives two isomorphisms, among
which is the announced isomorphism.
in order to get the exact sequence that symbolizes the finite presentation of M :
M1 −→ M0 −→ M −→ 0.
M1 −→ M0 −→ M −→ 0 and N1 −→ N0 −→ N −→ 0,
(M1 ⊗ N0 ) ⊕ (M0 ⊗ N1 ) −→ M0 ⊗ N0 −→ M ⊗ N −→ 0.
there exists P such that P ⊕ P is free with finite bases; thus Hom(P ⊕ P , M0 ) is
also free with finite bases, and consequently Hom(P, M0 ) is finitely generated and
projective; and the same for Hom(P, M1 ).
TK (M ) −→ TL (M ) and SK (M ) −→ SL (M );
each degree k ≥ 2 the mappings TkK (M ) → TkL (M ) and SkK (M ) → SkL (M ) are
surjective, and even bijective when f is surjective.
Let us now suppose that M is merely a K-module; we may derive from it
two L-modules, called the extensions of M , which are L ⊗K M and HomK (L, M );
an element µ of L multiplies an element λ ⊗ x of L ⊗K M or an element ξ of
HomK (L, M ) in this way:
defined by x −→ 1L ⊗ x and ξ −→ ξ(1L ) ; the former is not always injective,
and the latter is not always surjective. When L = K, both are bijective and
usually K ⊗K M and HomK (K, M ) are identified with M . But in general these
two extensions are different, and isomorphisms can be found between them only
under restrictive hypotheses. The next lemma gives more details in a particular
case.
(1.9.1) Lemma. Let us suppose that f : K → L is surjective; let a be its kernel.
Then L ⊗K M is canonically isomorphic to the quotient M/aM , and HomK (L, M )
to the submodule of all elements x ∈ M such that ax = 0.
Proof. The right exactness of the functor ⊗ gives the exact sequence
a ⊗K M −→ K ⊗K M −→ L ⊗K M −→ 0;
TL (L ⊗K M ) ∼
= L ⊗K TK (M ), (1.9.2)
∼ L ⊗K SK (M ).
SL (L ⊗K M ) = (1.9.3)
Most of the here mentioned isomorphisms are easy consequences of the definitions
and universal properties of the objects under consideration, and the following
explanations about (1.9.2) should be a sufficient model for all the others. Observe
that the L-algebras TL (L ⊗K M ) and L ⊗K TK (M ) are generated by the elements
which in each algebra are written 1L ⊗ x (with x ∈ M ). The mapping which
maps every λ ⊗ x in L ⊗K M to the corresponding λ ⊗ x in L ⊗K TK (M ), is
L-linear, and therefore extends to a morphism of L-algebras TL (L ⊗K M ) →
L ⊗K TK (M ). Conversely the mapping x −→ 1L ⊗ x extends to a morphism
1.9. Changes of basic rings 19
λ µ tλ + sµ λµ λµ
+ = and = ;
s t st s t st
it is easy to prove that this addition and this multiplication are well defined on the
set of equivalence classes and satisfy the required properties for S −1 K to be a ring;
the zero and unit elements are respectively the fractions 0/1 and 1/1. Moreover
the mapping f : K −→ S −1 K which maps every λ to the fraction λ/1 is a ring
morphism, and it is easy to prove that it is an initial universal object in C, as
stated in the following theorem.
(1.10.1) Theorem. For every ring morphism f : K → L such that f (s) is invertible
for all s ∈ S, there exists a unique ring morphism f : S −1 K → L such that
f (λ) = f (λ/1) for all λ ∈ K.
It is clear that f (λ/s) = f (λ)f (s)−1 for all fractions λ/s. Moreover an element
λ ∈ K belongs to the kernel of the canonical morphism f : K → S −1 K if and only
if there exists t ∈ S such that tλ = 0.
If S is a multiplicative subset containing S, the universal property of S −1 K
implies that the canonical mapping K → S −1 K can be factorized through S −1 K,
and thus we get a ring morphism S −1 K → S −1 K.
When K is an integral domain (a ring without divisors of zero, and not
reduced to 0), the set of all nonzero elements of K is a multiplicative subset; the
corresponding ring of fractions is a field, which is called the field of fractions of
K; all rings of fractions of K and K itself can be identified with subrings of this
field.
Now let M be any K-module, and D the category in which the objects are
the K-linear mappings g : M → P from M into any (S −1 K)-module P ; a mor-
phism from g to g : M → P is a (S −1 K)-linear mapping u : P → P such that
22 Chapter 1. Algebraic Preliminaries
x y tx + sy λ y λy
+ = and = .
s t st s t st
Two kinds of multiplicative subsets will be used later. First from any element
s ∈ K we can derive the multiplicative subset S = {1, s, s2 , s3 , . . . } of all powers
of s; then the ring S −1 K is usually denoted by Ks ; it is reduced to 0 if and only
if s is nilpotent. Similarly Ms is the module of fractions of M with denominator a
power of s.
An ideal p of K is called a prime ideal when the following four equivalent
statements are true:
(a) p = K, and whenever p contains a product xy of elements of K, it contains
x or y;
(b) p = K, and whenever p contains a product ab of ideals of K, it contains
a or b;
(c) the quotient K/p is an integral domain (without divisors of zero and not
reduced to 0);
(d) the complementary subset S = K \ p is a multiplicative subset.
The corresponding ring S −1 K and modules S −1 M are denoted by Kp and Mp ,
and are called the localizations of K and M at the prime ideal p.
1.10. Rings and modules of fractions 23
Every ring K (not reduced to 0) contains prime ideals; indeed Zorn’s Lemma
implies the existence of maximal ideals, and maximal ideals are prime, because
the following two assertions are equivalent:
(a) the ideal m is maximal (it is contained in no ideal other than m and K, and
m = K);
(b) the quotient K/m is a field.
Zorn’s Lemma even implies that every ideal other than K is contained in a maximal
ideal.
Let us also recall that for any ring K the following two statements are equiv-
alent:
(a) K contains exactly one maximal ideal;
(b) K is not reduced to 0, and a sum of noninvertible elements is never invertible.
When these statements are true, K is called a local ring. In a local ring the unique
maximal ideal m is the subset of all noninvertible elements. The quotient K/m is
called the residue field of the local ring K.
When p is a prime ideal of K, the elements of Kp which are not invertible,
are the elements of pKp (this notation is meaningful since Kp is a K-module); this
proves that the localized ring Kp is a local ring.
Here is a first application of these notions.
(1.10.2) Theorem. The radical of K, which is the subset Rad(K) of all its nilpotent
elements, is also the intersection of all its prime ideals.
Proof. It is clear that every prime ideal contains all nilpotent elements. Conversely
we prove that when s is not nilpotent, there exists a prime ideal that does not
contain it. Indeed let us consider the ring Ks of fractions with denominator in
the set of powers of s, and the canonical morphism f : K → Ks . Since s is not
nilpotent, Ks is not reduced to 0. If m is a maximal ideal of Ks , f −1 (m) is a prime
ideal p of K because K/p is isomorphic to a subring of the field Ks /m. Since f (s)
is invertible and cannot belong to m, we are sure that s ∈ / p.
Every multiplicative subset S affords a functor from Mod(K) toward
Mod(S −1 K); with every K-module M we associate the (S −1 K)-module S −1 M ,
and with every K-linear mapping f : M −→ N we associate the (S −1 K)-linear
mapping S −1 f : S −1 M → S −1 N defined in this way: because of the universal
property of S −1 M , the mapping M −→ N → S −1 N can be factorized in a unique
way through S −1 M .
(1.10.3) Theorem. The functor M −→ S −1 M is exact. Moreover it is equivalent
to the functor M −→ S −1 K ⊗ M corresponding to the extension K → S −1 K of
the basic ring; in other words, for each K-module M there is a canonical (S −1 K)-
linear isomorphism S −1 M → S −1 K ⊗ M , and for each K-linear mapping f :
M −→ N the canonical isomorphisms S −1 M → S −1 K ⊗M and S −1 N → S −1 K ⊗
N intertwine S −1 f and S −1 K ⊗ f .
24 Chapter 1. Algebraic Preliminaries
S −1 (M ⊗K M ) ∼
= S −1 M ⊗S −1 K S −1 M .
S −1 K ⊗K (M ⊗K M ) ∼
= (S −1 K ⊗K M ) ⊗S −1 K (S −1 K ⊗K M ).
Proof. Since L is a direct sum of flat modules, it is flat. Let us suppose that
L ⊗ M = 0 ; therefore for each x ∈ M there exists a positive integer k such
thatski x = 0 for i = 1, 2, . . . , n. If a = K, there exist λ1 , λ2 ,. . . ,λ
nn such that
n
1 = i=1 λi si ; let us set m = n(k − 1) + 1 ; from the equality 1 = n i=1 λ
( i si )
m
k
we can deduce the existence of µ1 , µ2 , . . . , µn such that 1 = i=1 µi si ; this
shows that the equalities ski x = 0 imply x = 0 ; therefore M = 0. Conversely if
a = K, the equality (K/a) ⊗ L = 0 (a consequence of (1.9.1)) shows that L is not
faithfully flat.
S −1 (N + N ) = S −1 N + S −1 N and S −1 (N ∩ N ) = S −1 N ∩ S −1 N .
1.11. Localization and globalization 25
Each element g/s of S −1 Hom(M, N ) is mapped to x/t −→ g(x)/st. This proposi-
tion is an immediate corollary of (1.9.9) because the extension K → S −1 K is flat,
as stated in (1.10.4).
(1.10.9) Lemma. The mapping q −→ f −1 (q) is a bijection from the set of prime
ideals q of S −1 K onto the set of prime ideals p of K such that p ∩ S is empty; the
converse bijection is p −→ S −1 p.
V(aj ) = V( aj );
j∈J j∈J
All these statements are evident except perhaps the last one, a consequence of the
inclusions
V(a) ∪ V(b) ⊂ V(a ∩ b) ⊂ V(ab) ⊂ V(a) ∪ V(b).
Lemma (1.11.1) proves that there is a topology on Spec(K) for which the closed
subsets are the subsets V(a); it is called the Zariski topology of Spec(K). For every
p ∈ Spec(A) the topological closure of {p} is V(p); thus the point {p} is closed if
and only if p is a maximal ideal; this topology is almost never a Hausdorff topology.
As explained in 1.10, with each p ∈ Spec(K) is associated a localized ring Kp
with maximal ideal pKp , and a residue field Kp /pKp. The kernel of the ring mor-
phism K → Kp → Kp /pKp is exactly p. Therefore there is an injective morphism
from the integral domain K/p into the residue field, which extends to a morphism
from the field of fractions of K/p into the residue field. This morphism is surjec-
tive, consequently bijective (since every field morphism is injective). Therefore the
residue field Kp /pKp can be identified with the field of fractions of K/p. At every
maximal ideal m the residue field is K/m.
Every module M gives a localized module Mp at the point p, and a vector
space Fp ⊗K M over the residue field Fp = Kp /pKp . The associativity of tensor
products allows us to write
Fp ⊗K M = Fp ⊗Kp Kp ⊗K M = Fp ⊗Kp Mp ;
and then (1.9.1) allows us to identify Fp ⊗ M with Mp /pMp . For a maximal ideal
m we can write Fm = K/m and Fm ⊗ M = M/mM .
Let f : K → L be any ring morphism. When q is a prime ideal of L, then
f −1 (q) is a prime ideal of K, because the quotient K/f −1 (q) is isomorphic to a
subring of L/q. This defines a mapping Spec(f ) from Spec(L) into Spec(K), and
it is easy to prove that it is continuous. Thus we have constructed a contravariant
functor from the category Com(Z) of all commutative rings to the category of
topological spaces.
(1.11.2) Example. Let F be an algebraically closed field, and K = F [x1 , x2 , . . . , xn ]
a ring of polynomials over F ; K can be identified with the ring of polynomial
functions on F n and its field of fractions F (x1 , x2 , . . . , xn ) with the field of rational
functions on F n ; every ring of fractions of K is a subring of this field. At every
point a = (a1 , a2 , . . . , an ) ∈ F n there is a valuation morphism K → F defined
by f −→ f (a); if ma is its kernel, the quotient K/ma is isomorphic to F and
consequently ma is a maximal ideal; a theorem of Hilbert (Nullstellensatz) proves
that the image of the mapping a −→ ma is the set of all maximal ideals. We
identify the point a ∈ F n with the point ma ∈ Spec(K). The localized ring Ka
at the maximal ideal ma is the subring of all rational functions g that are defined
at the point a, and the image of g in the residue field (which is isomorphic to
K/ma = F ) can be identified with the value g(a) of g at the point a. Now let p
be any prime ideal of K, and V (p) the subset of all a ∈ F n such that f (a) = 0
for all f ∈ p; a subset like V (p) is called an irreducible algebraic submanifold of
1.11. Localization and globalization 27
F n , and its points are the maximal ideals belonging to V(p). The localized ring
at p is the subring of all rational functions that are defined at least at one point
of V (p) (consequently at almost all points of V (p)). The ideal p is the kernel of
the ring morphism that maps every polynomial function f ∈ K to its restriction
to the subset V (p), and thus K/p is identified with a ring of functions on V (p),
the so-called regular functions. The residue field at p is the field of fractions of
K/p and its elements are called the rational functions on V (p). If g belongs to the
localized ring Kp , its image in the residue field is its restriction to V (p).
Let s be a nonzero element in the previous ring K = F [x1 , . . . , xn ]. It is (in an
essentially unique way) a product of prime elements (each one generates a prime
ideal in K); with each prime divisor of s is associated an irreducible algebraic
hypersurface in F n , and V (Ks) (which is the subset of all a ∈ F n such that
s(a) = 0) is the union of all these hypersurfaces. Let Us be the complementary
subset of V (Ks) in F n . The elements of Ks are the rational functions on F n
that are defined at all points of Us ; their restrictions to Us are called the regular
functions on Us . According to Lemma (1.10.9) the mapping Spec(Ks ) → Spec(K)
is a bijection from Spec(Ks ) onto the subset Us of all prime ideals of K that do
not contain s; this subset Us is open because the complementary subset is V(Ks).
The elements of Us correspond to the irreducible algebraic submanifolds V (p)
contained in Us .
Let us examine what happens when Spec(K) is not a connected topological
space.
(1.11.3) Theorem. When K is a direct sum n of ideals K1 , K2 , . . . , Kn , each
Ki is generated by an idempotent ei and i=1 ei = 1 ; moreover the mapping
Spec(Ki ) → Spec(K) associated with the projection K → Ki induces a bijection
from Spec(Ki ) onto an open subset Ui of Spec(K), and Spec(K) is the disjoint
union of the open subsets U1 , U2 , . . . ,Un . Conversely when Spec(K) is a dis-
joint union of open subsets U1 , U2 , . . . , Un , then K is a direct sum of ideals K1 ,
K2 , . . . , Kn in such a way that each open subset Ui is the image of the mapping
Spec(Ki ) → Spec(K).
n
Proof. Let us suppose that K = i=1 Ki and let ei be the component of 1 in Ki
for i = 1, 2, . . . , n. The equality λ = ni=1 λei holds for every λ ∈ K, and proves
that λei is the projection of λ in Ki ; in particular e2i = ei and Ki = Kei . A prime
ideal cannot contain all the n idempotents ei because their sum is 1; if it does
not contain ei , it must contain all ej such that j = i, because ei ej = 0. Therefore
each prime ideal contains all the n idempotents ei except one, and Spec(K) is the
disjoint union of the n subsets Ui defined in this way: Ui is the set of all prime
ideals containing ej whenever j = i. Since Ui is equal both to V(K(1 − ei )) and
to the subset complementary to V(Kei ), it is open and closed. The continuous
mapping Spec(Ki ) → Spec(K) maps each prime ideal pi of Ki to the prime ideal
of K which is the direct sum of pi and all Kj with j = i; thus we get a bijection
Spec(Ki ) → Ui .
Random documents with unrelated
content Scribd suggests to you:
And at these encouraging words Piglet felt quite happy again, and
decided not to be a Sailor after all. So when Christopher Robin had
helped them out of the Gravel Pit, they all went off together hand-in-
hand.
And two days later Rabbit happened to meet Eeyore in the Forest.
"Hallo, Eeyore," he said, "what are you looking for?"
"Small, of course," said Eeyore. "Haven't you any brain?"
"Oh, but didn't I tell you?" said Rabbit. "Small was found two days
ago."
There was a moment's silence.
"Ha-ha," said Eeyore bitterly. "Merriment and what-not. Don't
apologize. It's just what would happen."
CHAPTER IV
IN WHICH It Is Shown That Tiggers Don't Climb Trees
One day when Pooh was thinking, he thought he would go and see
Eeyore, because he hadn't seen him since yesterday. And as he
walked through the heather, singing to himself, he suddenly
remembered that he hadn't seen Owl since the day before yesterday,
so he thought that he would just look in at the Hundred Acre Wood on
the way and see if Owl was at home.
Well, he went on singing, until he came to the part of the stream
where the stepping-stones were, and when he was in the middle of
the third stone he began to wonder how Kanga and Roo and Tigger
were getting on, because they all lived together in a different part of
the Forest. And he thought, "I haven't seen Roo for a long time, and if
I don't see him today it will be a still longer time." So he sat down on
the stone in the middle of the stream, and sang another verse of his
song, while he wondered what to do.
The other verse of the song was like this:
So when he had sung this, he got up off his stone, walked back
across the stream, and set off for Rabbit's house.
But he hadn't got far before he began to say to himself:
"Yes, but suppose Rabbit is out?"
"Or suppose I get stuck in his front door again, coming out, as I did
once when his front door wasn't big enough?"
"Because I know I'm not getting fatter, but his front door may be
getting thinner."
"So wouldn't it be better if——"
And all the time he was saying things like this he was going more and
more westerly, without thinking ... until suddenly he found himself at
his own front door again.
And it was eleven o'clock.
Which was Time-for-a-little-something....
Half an hour later he was doing what he had always really meant to
do, he was stumping off to Piglet's house. And as he walked, he
wiped his mouth with the back of his paw, and sang rather a fluffy
song through the fur. It went like this:
Written down, like this, it doesn't seem a very good song, but coming
through pale fawn fluff at about half-past eleven on a very sunny
morning, it seemed to Pooh to be one of the best songs he had ever
sung. So he went on singing it.
Piglet was busy digging a small hole in the ground outside his house.
"Hallo, Piglet," said Pooh.
"Hallo, Pooh," said Piglet, giving a jump of surprise. "I knew it was
you."
"So did I," said Pooh. "What are you doing?"
"I'm planting a haycorn, Pooh, so that it can grow up into an oak-tree,
and have lots of haycorns just outside the front door instead of having
to walk miles and miles, do you see, Pooh?"
"Supposing it doesn't?" said Pooh.
"It will, because Christopher Robin says it will, so that's why I'm
planting it."
"Well," said Pooh, "if I plant a honeycomb outside my house, then it
will grow up into a beehive."
Piglet wasn't quite sure about this.
"Or a piece of a honeycomb," said Pooh, "so as not to waste too
much. Only then I might only get a piece of a beehive, and it might be
the wrong piece, where the bees were buzzing and not hunnying.
Bother."
Piglet agreed that that would be rather bothering.
"Besides, Pooh, it's a very difficult thing, planting unless you know
how to do it," he said; and he put the acorn in the hole he had made,
and covered it up with earth, and jumped on it.
And as they went, Tigger told Roo (who wanted to know) all about the
things that Tiggers could do.
"Can they fly?" asked Roo.
"Yes," said Tigger, "they're very good flyers, Tiggers are. Stornry good
flyers."
"Oo!" said Roo. "Can they fly as well as Owl?"
"Yes," said Tigger. "Only they don't want to."
"Why don't they want to?"
"Well, they just don't like it, somehow."
Roo couldn't understand this, because he thought it would be lovely
to be able to fly, but Tigger said it was difficult to explain to anybody
who wasn't a Tigger himself.
"Well," said Roo, "can they jump as far as Kangas?"
"Yes," said Tigger. "When they want to."
"I love jumping," said Roo. "Let's see who can jump farthest, you or
me."
"I can," said Tigger. "But we mustn't stop now, or we shall be late."
"Late for what?"
"For whatever we want to be in time for," said Tigger, hurrying on.
In a little while they came to the Six Pine Trees.
"I can swim," said Roo. "I fell into the river, and I swimmed. Can
Tiggers swim?"
"Of course they can. Tiggers can do everything."
"Can they climb trees better than Pooh?" asked Roo, stopping under
the tallest Pine Tree, and looking up at it.
"Climbing trees is what they do best," said Tigger. "Much better than
Poohs."
"Could they climb this one?"
"They're always climbing trees like that," said Tigger. "Up and down
all day."
"Oo, Tigger, are they really?"
"I'll show you," said Tigger bravely, "and you can sit on my back and
watch me." For of all the things which he had said Tiggers could do,
the only one he felt really certain about suddenly was climbing trees.
"Oo, Tigger, oo, Tigger, oo, Tigger!" squeaked Roo excitedly.
So he sat on Tigger's back and up they went.
And for the first ten feet Tigger said happily to himself, "Up we go!"
And for the next ten feet he said:
"I always said Tiggers could climb trees."
And for the next ten feet he said:
"Not that it's easy, mind you."
And for the next ten feet he said:
"Of course, there's the coming-down too. Backwards."
And then he said:
"Which will be difficult ...
"Unless one fell ...
"when it would be ...
"EASY."
And at the word "easy" the branch he was standing on broke
suddenly, and he just managed to clutch at the one above him as he
felt himself going ... and then slowly he got his chin over it ... and then
one back paw ... and then the other ... until at last he was sitting on it,
breathing very quickly, and wishing that he had gone in for swimming
instead.
Roo climbed off, and sat down next to him.
"Oo, Tigger," he said excitedly, "are we at the top?"
"No," said Tigger.
"Are we going to the top?"
"No," said Tigger.
"Oh!" said Roo rather sadly. And then he went on hopefully: "That
was a lovely bit just now, when you pretended we were going to fall-
bump-to-the-bottom, and we didn't. Will you do that bit again?"
"NO," said Tigger.
Roo was silent for a little while, and then he said, "Shall we eat our
sandwiches, Tigger?" And Tigger said, "Yes, where are they?" And
Roo said, "At the bottom of the tree." And Tigger said, "I don't think
we'd better eat them just yet." So they didn't.
By and by Pooh and Piglet came along. Pooh was telling Piglet in a
singing voice that it didn't seem to matter, if he didn't get any fatter,
and he didn't think he was getting any fatter, what he did; and Piglet
was wondering how long it would be before his haycorn came up.
"Look, Pooh!" said Piglet suddenly. "There's something in one of the
Pine Trees."
"So there is!" said Pooh, looking up wonderingly. "There's an Animal."
Piglet took Pooh's arm, in case Pooh was frightened.
"Is it One of the Fiercer Animals?" he said, looking the other way.
Pooh nodded.
"It's a Jagular," he said.
"What do Jagulars do?" asked Piglet, hoping that they wouldn't.
"They hide in the branches of trees, and drop on you as you go
underneath," said Pooh. "Christopher Robin told me."
"Perhaps we better hadn't go underneath, Pooh. In case he dropped
and hurt himself."
"They don't hurt themselves," said Pooh. "They're such very good
droppers."
Piglet still felt that to be underneath a Very Good Dropper would be a
Mistake, and he was just going to hurry back for something which he
had forgotten when the Jagular called out to them.
"Help! Help!" it called.
"That's what Jagulars always do," said Pooh, much interested. "They
call 'Help! Help!' and then when you look up, they drop on you."
"I'm looking down," cried Piglet loudly, so as the Jagular shouldn't do
the wrong thing by accident.
Something very excited next to the Jagular heard him, and squeaked:
"Pooh and Piglet! Pooh and Piglet!"
All of a sudden Piglet felt that it was a much nicer day than he had
thought it was. All warm and sunny——
"Pooh!" he cried. "I believe it's Tigger and Roo!"
"So it is," said Pooh. "I thought it was a Jagular and another Jagular."
"Hallo, Roo!" called Piglet. "What are you doing?"
"We can't get down, we can't get down!" cried Roo. "Isn't it fun? Pooh,
isn't it fun, Tigger and I are living in a tree, like Owl, and we're going
to stay here for ever and ever. I can see Piglet's house. Piglet, I can
see your house from here. Aren't we high? Is Owl's house as high up
as this?"
CHAPTER V
IN WHICH Rabbit Has a Busy Day, and We Learn What Christopher
Robin Does in the Mornings
It was going to be one of Rabbit's busy days. As soon as he woke up
he felt important, as if everything depended upon him. It was just the
day for Organizing Something, or for Writing a Notice Signed Rabbit,
or for Seeing What Everybody Else Thought About It. It was a perfect
morning for hurrying round to Pooh, and saying, "Very well, then, I'll
tell Piglet," and then going to Piglet, and saying, "Pooh thinks—but
perhaps I'd better see Owl first." It was a Captainish sort of day, when
everybody said, "Yes, Rabbit" and "No, Rabbit," and waited until he
had told them.
He came out of his house and sniffed the warm spring morning as he
wondered what he would do. Kanga's house was nearest, and at
Kanga's house was Roo, who said "Yes, Rabbit" and "No, Rabbit"
almost better than anybody else in the Forest; but there was another
animal there nowadays, the strange and Bouncy Tigger; and he was
the sort of Tigger who was always in front when you were showing
him the way anywhere, and was generally out of sight when at last
you came to the place and said proudly "Here we are!"
"No, not Kanga's," said Rabbit thoughtfully to himself, as he curled his
whiskers in the sun; and, to make quite sure that he wasn't going
there, he turned to the left and trotted off in the other direction, which
was the way to Christopher Robin's house.
"After all," said Rabbit to himself, "Christopher Robin depends on Me.
He's fond of Pooh and Piglet and Eeyore, and so am I, but they
haven't any Brain. Not to notice. And he respects Owl, because you
can't help respecting anybody who can spell TUESDAY, even if he
doesn't spell it right; but spelling isn't everything. There are days
when spelling Tuesday simply doesn't count. And Kanga is too busy
looking after Roo, and Roo is too young and Tigger is too bouncy to
be any help, so there's really nobody but Me, when you come to look
at it. I'll go and see if there's anything he wants doing, and then I'll do
it for him. It's just the day for doing things."
He trotted along happily, and by-and-by he crossed the stream and
came to the place where his friends-and-relations lived. There
seemed to be even more of them about than usual this morning, and
having nodded to a hedgehog or two, with whom he was too busy to
shake hands, and having said, "Good morning, good morning,"
importantly to some of the others, and "Ah, there you are," kindly, to
the smaller ones, he waved a paw at them over his shoulder, and was
gone; leaving such an air of excitement and I-don't-know-what behind
him, that several members of the Beetle family, including Henry Rush,
made their way at once to the Hundred Acre Wood and began
climbing trees, in the hope of getting to the top before it happened,
whatever it was, so that they might see it properly.
Rabbit hurried on by the edge of the Hundred Acre Wood, feeling
more important every minute, and soon he came to the tree where
Christopher Robin lived. He knocked at the door, and he called out
once or twice, and then he walked back a little way and put his paw
up to keep the sun out, and called to the top of the tree, and then he
turned all round and shouted "Hallo!" and "I say!" "It's Rabbit!"—but
nothing happened. Then he stopped and listened, and everything
stopped and listened with him, and the Forest was very lone and still
and peaceful in the sunshine, until suddenly a hundred miles above
him a lark began to sing.
"Bother!" said Rabbit. "He's gone out."
He went back to the green front door, just to make sure, and he was
turning away, feeling that his morning had got all spoilt, when he saw
a piece of paper on the ground. And there was a pin in it, as if it had
fallen off the door.
"Ha!" said Rabbit, feeling quite happy again. "Another notice!"
This is what it said:
GON OUT
BACKSON
BISY
BACKSON.
C. R.
"Ha!" said Rabbit again. "I must tell the others." And he hurried off
importantly.
The nearest house was Owl's, and to Owl's House in the Hundred
Acre Wood he made his way. He came to Owl's door, and he knocked
and he rang, and he rang and he knocked, and at last Owl's head
Welcome to our website – the ideal destination for book lovers and
knowledge seekers. With a mission to inspire endlessly, we offer a
vast collection of books, ranging from classic literary works to
specialized publications, self-development books, and children's
literature. Each book is a new journey of discovery, expanding
knowledge and enriching the soul of the reade
Our website is not just a platform for buying books, but a bridge
connecting readers to the timeless values of culture and wisdom. With
an elegant, user-friendly interface and an intelligent search system,
we are committed to providing a quick and convenient shopping
experience. Additionally, our special promotions and home delivery
services ensure that you save time and fully enjoy the joy of reading.
ebookfinal.com