Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Introduction to Homological Algebra, 85
Introduction to Homological Algebra, 85
Introduction to Homological Algebra, 85
Ebook681 pages7 hours

Introduction to Homological Algebra, 85

Rating: 4 out of 5 stars

4/5

()

Read preview

About this ebook

An Introduction to Homological Algebra discusses the origins of algebraic topology. It also presents the study of homological algebra as a two-stage affair. First, one must learn the language of Ext and Tor and what it describes. Second, one must be able to compute these things, and often, this involves yet another language: spectral sequences. Homological algebra is an accessible subject to those who wish to learn it, and this book is the author’s attempt to make it lovable. This book comprises 11 chapters, with an introductory chapter that focuses on line integrals and independence of path, categories and functors, tensor products, and singular homology. Succeeding chapters discuss Hom and ?; projectives, injectives, and flats; specific rings; extensions of groups; homology; Ext; Tor; son of specific rings; the return of cohomology of groups; and spectral sequences, such as bicomplexes, Kunneth Theorems, and Grothendieck Spectral Sequences. This book will be of interest to practitioners in the field of pure and applied mathematics.
LanguageEnglish
Release dateSep 7, 1979
ISBN9780080874012
Introduction to Homological Algebra, 85

Related to Introduction to Homological Algebra, 85

Related ebooks

Mathematics For You

View More

Related articles

Reviews for Introduction to Homological Algebra, 85

Rating: 4 out of 5 stars
4/5

1 rating0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Introduction to Homological Algebra, 85 - Joseph J. Rotman

    Preface

    In 1968, I taught a course in homological algebra at Urbana; those lectures were published in 1970 as my Notes on Homological Algebra. Since much of the basic material there has now drifted into earlier courses, I have rewritten the notes, incorporating new items to take advantage of the present times. As it is convenient that a book be essentially self-contained, the old material remains, though redone. Therefore, students at different levels should find something of interest here.

    Learning homological algebra is a two-stage affair. First, one must learn the language of Ext and Tor and what it describes. Second, one must be able to compute these things, and, often, this involves yet another language: spectral sequences. The following exercise appears on page 105 of Lang’s book, Algebra: Take any book on homological algebra and prove all the theorems without looking at the proofs given in that book. My guess is that Lang was trying to foster the healthy attitude that, in spite of its cumbersome, perhaps forbidding, appearance, homological algebra is actually accessible to those who wish to learn it. If nothing else, this book is my attempt to make the subject lovable.

    The origins of our subject are the origins of algebraic topology. At first glance, our opening sentence, We begin with a brief discussion of line integrals in the plane, must appear strange; upon reflection, how could we begin any other way? A simple definition may mystify a reader if it seems to have been offered up by some oracle. There are no oracles here. Indeed, one role of the introductory first chapter is to give the etymology, both linguistic and mathematical, of the word homology.

    With Chapter 1, the next three chapters form a short course in module theory. The main subject of homological algebra is the pair of bifunctors Hom and tensor. Thus, Chapter 2 studies their behavior on all modules, whereas Chapter 3 studies how they act on certain special modules: projectives, injectives, and flats. These special modules are used at once to characterize Hom and tensor (Watts’ theorems, which gave rise to adjoint functor theorems in category theory). We also introduce localization for later applications to polynomial rings and to regular local rings. To make all more concrete, to see why certain earlier hypotheses are needed, and to make contact with honest mathematics, Chapter 4 examines the interplay between the ring of scalars and the special modules. In particular, we prove the Quillen–Suslin theorem that projective R[x1, …, xn]–modules are free when R is a field.

    Chapter 5 is important etymology: group extensions are considered so that one can see the familiar formulas being born, crying for a projective resolution. At last, Chapter 6 defines homology and derived functors. The next two chapters treat Ext", the derived functors of Hom, and Torn, the derived functors of tensor. We show that Ext was named for its relation to extensions, and that Tor was named for its relation to torsion.

    Chapters 9 and 10 apply Ext and Tor to rings and to groups. An introduction must show why its subject is valuable, and we have chosen these two areas as the most familiar ones in algebra. Dimension theory of rings organizes much of the material in Chapter 4 and also yields new results (we describe Serre’s characterization of regular local rings, and the theorem of Auslander–Buchsbaum–Nagata that these rings have unique factorization). Applications to groups include a discussion of the Schur multiplier, Maschke’s theorem, the Schur–Zassenhaus theorem, and a description of the theorem of Stallings–Swan characterizing free groups as those groups of cohomological dimension one.

    We have tried to keep the needs of algebraic topologists in mind, partly because one ought, and partly because the algebra is better understood when one sees its topological sources. A short discussion of Eilenberg–MacLane spaces does point out the intimate connection between topology and homological algebra; indeed, we even prove the universal coefficient theorem for group cohomology by looking at its counterpart for singular cohomology.

    The final chapter, spectral sequences, is almost a text by itself. This is not unexpected, for we have already mentioned that it is a second language. (Here I have not given etymology, although Lyndon [1948] is recommended for an algebraic viewpoint.) I think this exposition differs from others in its emphasis on illustrating how spectral sequences are used. Also, exact couples seem to eliminate most of the horrors of many indices. If I have succeeded, the reader will know when a spectral sequence is likely to arise, and he will know what to do with it if it does.

    And now the pleasant paragraph in which I can thank people from whom I have learned. I. Kaplansky, S. MacLane, J. Gray, and S. Gersten gave lectures that taught me much. L. McCulloh has made many valuable suggestions. E. Davis allowed me to use his account of Quillen’s solution of Serre’s problem. R. Treger told me of Vaserstein’s proof of Suslin’s solution of Serre’s problem, and was kind enough to translate it for me from the Russian. An incomplete list of others who helped: P. M. Cohn, A. Fauntleroy, E. Green, P. Griffith, the late A. Learner, A. Mann, D. Robinson, M. Schacher, and A. Zaks. Of course, I am indebted to the fine books on homological algebra, especially those of Cartan–Eilenberg [1956] and MacLane [1963].

    I am grateful to the Lady Davis Foundation; most of this book was written while I was a Lady Davis Fellow and Visiting Professor at the Technion, Israel Institute of Technology, Haifa, and at the Hebrew University of Jerusalem. My warm thanks to both of these institutions for their hospitality.

    Finally, I thank Mrs. Esther Tuval for a superb job of typing my manuscript.

    1

    Introduction

    Line Integrals and Independence of Path

    We begin with a brief discussion of line integrals in the plane. Assume X is an open disc in the plane with a finite number of points z1, …, zn deleted. Fix points a and b in X. Given a curve β in X from a to b, and given a pair of real-valued functions P and Q on X, is a second path in X from a to b,

    (i.e., go from a to b to a. Thus, we are led to consideration of closed curves in X. Now this last line integral is affected by whether any of the bad.points z1, …, zn lie inside γ (one of the functions P or Q may have a singularity at some zi). Consider

    Each γi is a (simple) closed curve in X with the point zi inside (and the other z’s outside); γ is a (simple) closed curve in X having all the γi inside. If γ is oriented counterclockwise and each of the γi is oriented clockwise, then Green’s theorem states that

    where R is the shaded 2-dimensional region in the picture. Of course, one is tempted to, and does, write the sum of line integrals more concisely as

    indeed, instead of mentioning the orientations explicitly, one often writes a formal linear combination of curves (negative coefficients being a convenient way to indicate how the orientations are aligned). Observe that for simple curves the notion of inside.always makes sense. However, we may allow curves to wind about some zi several times, assuming these curves are not so pathological as to fail to have an inside, and this introduces integer coefficients other than ± in the formal linear combinations. Let us now restrict attention to those pairs of functions P and Q satisfying ∂Q/∂x = ∂P/∂y. . An equivalence relation on formal linear combinations (or, oriented unions.) of closed curves in X equivalent if the values of

    agree for all pairs P and Q as restricted above. In particular, note that combinations α whose curves comprise the boundary of a 2-dimensional region in X are trivial (the integrals have value 0). The equivalence classes of formal linear combinations of closed curves in X are called homology classes, from the Greek homologos.= homo.+ logos.meaning agreeing. In short, integration is always independent of paths lying in the same homology class. Visibly, these homology classes reflect the complexity of the ambient space X.

    Our discussion mentioned three sorts of finite unions of oriented curves in X: arbitrary such (usually called chains.); closed unions (usually called cycles.); boundaries of 2-dimensional regions in X. We have just witnessed the origin of homology as an abelian group of classes of formal linear combinations of closed curves in X in which boundaries of 2-dimensional regions are trivial. The step from considering closed curves in a planar region X to generalized chains, cycles, and boundaries in any euclidean space, indeed, in any topological space, does not seem an awesome leap today; nevertheless, this leap was the birth of algebraic topology. We shall return to the picture of Green’s theorem, for homological algebra was born when analogies between topological homology theory and algebra were recognized.

    Categories and Functors

    Let us introduce some terms to facilitate our discussion. Recall a set-theoretical distinction: our primitive vocabulary contains the words element.and class.; the word set.is reserved for those classes that are small enough to have a cardinal number. Thus, we may correctly say the class of all rationals.or the set of all rationals., but we may only say the class of all groups.(this latter class cannot be a set, for there exist groups of any nonzero cardinal).

    Definition

    A category consists of a class of objects, , pairwise disjoint sets of morphisms, (A, B), for every ordered pair of objects, and compositions (A, B(B, C(A, C), denoted (f,g) gf, satisfying the following axioms:

    (i) for each object A, there exists an identity morphism 1A (A, A) such that f1A = f for all f (A, B) and lAg = g for all g (C, A);

    (ii) associativity of composition holds whenever possible: if f (A, B), g (B, C) and h (C, D) then

    Several remarks are in order. First, the only requirement on Hom(A, B) is that it be a set; it is allowed to be empty. Second, we usually write f: A B instead of f (A, B) this explains our notation for composition. Finally, for each object A, the identity morphism 1A is unique: if 2A (A, A) is a second morphism satisfying (i), then 2A = 2A1A = lA. There is thus a one—one correspondence A 1A solely in terms of morphisms and compositions. This is sometimes convenient for category-theorists, but we shall never forget objects.

    Examples

    = Sets. Here objects are sets, morphisms are functions, and composition is the usual composition of functions.

    = Top. Here objects are topological spaces, morphisms are continuous functions, and composition is the usual composition.

    = R . Here R is a ring (associative with 1), objects are left R-modules, morphisms are R-maps (i.e., R-homomorphisms), and usual composition.

    Recall that a leftR-module is an additive abelian group M equipped with an action of R making it a "vector space over R" : there is a function R × M M, denoted (r, m) rm, satisfying:

    A function f: M N between left R-modules M and N is an R-map if

    We will often write RM

    4. If R = Z, the ring of integers, then R-module = Z-module = abelian group, while Z-map = homomorphism. We usually denote the category of abelian groups by Ab rather than Z .

    R. Here R is a ring, objects are right R-modules, morphisms are R-maps, and composition is the usual composition.

    Recall that an abelian group M is a rightR-module if there is a function R × M M, denoted (r, m) rm, satisfying all the axioms of a left R-module save (iii). The third axiom is replaced by

    If the value of R × M M on (r, m) is denoted mr, has the more appealing form

    We shall consistently use the latter notation (having r on the right) when discussing right R-modules. A function f: M N between right R-modules M and N is an R-map if

    If R is commutative, we may regard every right R-module M as a left R-module by defining rm = mr for all r R and m M.

    6. Let S be a semigroup with (two-sided) unit, i.e., a monoid. We may construe S has only one object, *), Hom(*, *) = S, and composition as the given multiplication in S.

    is a set, indeed, a rather small set. Second, morphisms are not functions between objects, for the object * is not comprised of elements.

    7. Let X be a quasi-ordered set (this means X has a reflexive and transitive binary relation ≤). To say it another way, X would be partially ordered if, in addition, ≤ were antisymmetric. We may construe X = X,

    whenever x y z.

    This last example has the feature that some Hom sets may be empty.

    The list of examples can be continued, and other categories will appear when necessary. There is a rough analogy between one’s first meeting with point-set topology and one’s first meeting with category theory. The definitions of topological space and of category are simple, abstract, and admit many examples, both natural and artificial. The raison d’etre of topological spaces is that continuous functions live on them, and it is continuity that is really interesting. Similarly, the raison d’etre of categories is that functors.live on them; let us define these at once. If one views a category as a generalized monoid (forget the objects, so one is dealing with morphisms that can sometimes be multiplied), then a functor is just a homomorphism.

    Definition

    be categories. A functorF: is a function satisfying:

    (i) If A , then FA

    (ii) If f: A B , then Ff: FA FB

    , then

    (iv) for every A , we have F(lA) = lFA.

    Examples

    8. The identity functorF: defined by FA = A and Ff = f.

    9. The Hom functors Sets. Fix an object A and define FC = Hom(A, Cby g fg. One usually denotes Ff by f*.

    , then the Hom functor F = HomR(A,) actually takes values in Ab: HomR(A, C) = {all R-maps A → C} is an abelian group if we define g + h by a g(a) + h(a); moreover, one easily checks that every f* is a homomorphism of abelian groups.

    11. A statement similar to that in R.

    12. If D , de fine the constant functor for every C , for every morphism f .

    Our description of the Hom functors is incomplete—what if we fix the second variable?

    Definition

    be categories. A contravariant functorF: is a function satisfying:

    (i) If A , then FA

    (ii)′. if f: A B , then Ff: FB FA (thus F reverses arrows);

    , then

    (iv) for every A .

    Examples

    13. Fix an object B and define F: Sets by FA = Hom(A, Bby g gf. One usually denotes Ff by f*.

    , then F = HomR(, B) actually takes its values in Ab. A R.

    A functor is often called a covariant functor to emphasize it preserves the direction of arrows.

    Exercises

    1.1. A morphism f: A B is called an equivalence such that gf = 1A and fg = 1B. What are the equivalences in the categories described above?

    1.2. If F: is a functor (either variance) and f , then Ff .

    Definition

    is pre-additive (A, B) is an (additive) abelian group and the distributivity laws hold, when defined.

    . As every nonempty set admits some abelian group structure, it would be foolish not to demand distributivity in the definition of pre-additive.

    Of course, R R are pre-additive categories, then we say a functor (contra or co) F: is additive if F(f + g)=Ff+Fg for every pair of morphisms f and g (C, C′(FC, FC′) given by f Ff is a homomorphism. Note that the Hom functors on module categories are additive.

    Exercises

    1.3. In a pre-additive category, zero morphisms 0: A B are defined (namely, the zero element of Hom(A, B)). If F: is an additive functor (either variance) between pre-additive categories, then F are module categories, then F(0) = 0, where 0 is the zero module.

    1.4. If R is a commutative ring, then HomR(M, N) is an R-module if we define rf by m r · fm. Moreover, either Hom functor takes values in R .

    (A special case of Exercise 1.4 is R a field (so that R-modules are vector spaces and R-maps are linear transformations). The contravariant functor HomR(, R) is the dual space functor that assigns to each space its space of linear functionals and to each linear transformation its transpose.)

    1.5. If R is a ring, its center is the subring

    For every ring R, prove that HomR(M, N) is a Z(R)-module and that the Hom functors take values in Z(R) .

    1.6. Let A be a left R-module and let r Z(R). Prove that the function μr: A → A defined by a ra is an R-map. Give an example to show this may be false if r ∉ Z(R). (Hint: Take A = RR.)

    1.7. Let R and S : R S a ring map. Every left S-module M may be construed as a left R-module if one defines r · m = ϕ(r)m (similarly for right modules).

    Tensor Products

    There is a second family of functors, tensor products, that are as fundamental in homological algebra as the Hom functors. Before describing them, we give a construction for abelian groups; it will be generalized to modules in Chapter 3.

    Definition

    Let G be an (additive) abelian group with subset X; we say G is a free (abelian) group with basisX if each g ∈ G has a unique expression of the form

    where mx Z and almost allmx = 0 (i.e., all but a finite number of mx = 0).

    The analogy between free groups and vector spaces is a good one; the next result shows that the notions of basis in each are kindred.

    Theorem 1.1

    Let G be free with basis X; let H be any abelian group and and let f: X H for all x X.

    Proof

    Gis unique and at most finitely many mx is unique follows from the fact that X generates G.

    from f is called extending by linearity.

    Theorem 1.2

    Given a set X, there exists a free abelian group G having X as a basis.

    Proof

    be a family of copies of the integers Z. its elements are vectors(mx)—this notation indicates that the xth coordinate is mx Z. The cartesian product is an abelian group under coordinatewise addition:

    Define X′ = {all vectors having 1 as xth coordinate (fixed x X) and all other coordinates 0}. Clearly X′ is in one–one correspondence with X. Define G′ generated by X′. It is immediate that each g G′ is a linear combination of elements of X′; . Thus, G′ is a free abelian group with basis X′.

    If one actually wants X (i.e., replace X′ by X) and define the obvious addition on the new set G; the new group G is free abelian with X as basis.

    Here is a very formal but useful notion.

    Definition

    A diagram

    commutes if βα = α′β′; a diagram

    commutes if α = γβ a larger diagram comprised of squares and triangles n commutes if each of its squares and triangles commutes.

    (Of course, commutativity of a triangle is the special case of a commutative square one of whose sides is an identity map.)

    Definition

    Let R be an associative ring with 1. If A is a right R-module, B a left R-module, and G is an additive abelian group, then an R-biadditive function¹ is a function

    such that for all a, a′ A, b, b′ B, and r ∈ R,

    We define a tensor product of A and B that converts R-biadditive functions A × B G .

    Definition

    A tensor product of A R and B R and an R-biadditive function h which solves the following universal mapping problem :

    for every abelian group G and every R-biadditive f, there exists a unique homomorphism f′ making the diagram commute.

    Theorem 1.3

    Any two tensor products of A and B are isomorphic.

    Proof

    Suppose there is a second group X and an R-biadditive k : A × B X that also solves the universal mapping problem. There is a diagram

    where k′ and h′ are homomorphisms such that

    There is a second diagram

    for the dashed arrow, we may choose either the identity or h′k′ (for h′k′h = h′k = h); uniqueness yields 1 = h′k′. A similar argument shows that k′h′ = 1x, so that k′ .

    The proof of Theorem 1.3 should serve as a model for a more general fact: any two solutions of a universal mapping problem, if, indeed, there are any, are isomorphic. Having proved uniqueness of tensor product, we now prove its existence.

    Theorem 1.4

    The tensor product of a right R-module A and a left R-module B exists.

    Proof

    Let F be the free abelian group with basis A × B, i.e., F is the group of Z-linear combinations of all ordered pairs (a, b). Define S as the subgroup of F generated by all elements of the following three forms:

    . If we denote the coset (a, b) + S by a b, defined by (a, b) a b is R-biadditive.

    Assume G is an abelian group, and we have the diagram

    where f is R-biadditive. Since F is free on A × B: F G ((a, b)) = f(a, b). Moreover, f being R-biadditive implies S induces a homomorphism f′: F/S G by

    , and this says f′h = f. We leave the proof of uniqueness of f′ is generated by all a b).

    It is important to realize that a typical element of A RB and may not have an expression of the form a bgiven by specifying its value on each a bis not well defined unless it preserves all the relations. The safest scheme is to rely on the universal mapping problem, for it is easier to define an R-biadditive function on the cartesian product A × B.

    Theorem 1.5

    Let f: A A′ be an R-map of right R-modules and let g: B B′ be an R-map of left R-modules

    Proof

    is obviously R-biadditive; now use universality.

    Definition

    is denoted

    Theorem 1.6

    Assume are R-maps of right Rare R-maps of left R-modules Then.

    Proof

    Use Theorem 1.5.

    Corollary 1.7

    If defined by

    if f: B B′ is an R-map between left R-modules B and B′,

    Similarly, for fixed , there is an additive functor with and (where g: A A′ is an R-map).

    Proof

    That F preserves composition follows immediately from Theorem 1.6, setting g = g′ = 1Asends

    .

    We shall usually write A R instead of F RB instead of G. When no confusion can arise, we suppress the subscript R.

    is only an abelian group; however, there is a common circumstance in which it is a module.

    Definition

    Let R and S be rings. An abelian group B is an (R–S) bimodule, denoted RBS, if B is a left R-module, a right S-module, and the two actions are related by an associative law.

    Enjoying the preview?
    Page 1 of 1