Solutions to Exercises from Kenneth Brown's Cohomology of Groups

Total Page:16

File Type:pdf, Size:1020Kb

Solutions to Exercises from Kenneth Brown's Cohomology of Groups Solutions to Exercises from Kenneth Brown's Cohomology of Groups Christopher A. Gerig, Cornell University (College of Engineering) August 2008 - May 2009 I appreciate emails concerning any errors/corrections: [email protected]. Any errors would be due to solely myself, or at least the undergraduate-version of myself when I last looked over this. Remark made on 1/28/14. 1 major: Applied & Engineering Physics year: Sophomore Undergraduate Hey Ken, thanks for taking me under your wing. 2 Contents 0 Errata to Cohomology of Groups 4 1 Chapter I: Some Homological Algebra 5 2 Chapter II: The Homology of a Group 12 3 Chapter III: Homology and Cohomology with Coefficients 21 4 Chapter IV: Low-Dimensional Cohomology and Group Extensions 31 5 Chapter V: Products 41 6 Chapter VI: Cohomology Theory of Finite Groups 47 7 Chapter VII: Equivariant Homology and Spectral Sequences 55 8 Chapter VIII: Finiteness Conditions 58 9 Chapter IX: Euler Characteristics 59 10 Additional Exercises 61 11 References 76 3 0 Errata to Cohomology of Groups pg62, line 11 missing a paranthesis ) at the end. pg67, line 15 from bottom missing word, should say \as an abelian group". pg71, last line of Exercise 4 hint should be on a new line (for whole exercise). P −1 pg85, line 9 from bottom incorrect function, should be g2C=H g ⊗ gm. pg114, first line of Exercise 4 misspelled endomorphism with an extra r. pg115, line 5 from bottom missing word, should say \4.4 is a chain map". ∗ pg141, line 4 from bottom missing a hat b on the last H . pg149, line 13 the ideal I should be italicized. pg158, line 20 there should be a space between SL2(Fp) and (p odd). ∗ pg160, line 5 from bottom missing the coefficient in cohomology, Hb (H; M). pg165, line 26 the comma in the homology should be lowered, Hq(C∗;p). 4 1 Chapter I: Some Homological Algebra P P 2.1(a): The set A = fg − 1 j g 2 G; g 6= 1g is linearly independent because αg(g − 1) = 0 ) αgg = P P P P αg = αg1 + 0g, and since ZG is a free Z-module, αg has a unique expression, yielding P P αg = 0 8g 2 A. To show that I = span(A), let αgg 2 I and hence αg = 0. Thus we can write P P P P P αgg = αgg − 0 = αgg − αg = αg(g − 1). Since A is a linearly independent set which generates I, it is a basis for I. 2.1(b): Consider the left ideal A = (fs − 1 j s 2 Sg) over ZG. P P P P A ⊆ I since "(A) = " ( rgg)(s − 1) = "( rgg)"(s − 1) = y · 0 = 0. P s Pg Ps Pg If x 2 I then x = αigi − βjgj such that αi = βj. P P P P P P x = ( αigi − αi) − ( βjgj − βj) + ( αi − βj) P P Thus x = αi(gi − 1) − βj(gj − 1) and it suffices to show g − 1 2 A for any g 2 G so that x 2 A and ±1 ±1 I ⊆ A. Since G = hSi, we have a representation x = s1 ··· sn . By using ab − 1 = a(b − 1) + (a − 1) and c−1 − 1 = −c−1(c − 1), the result follows immediately. 2.1(c): Suppose S ⊂ G j I = (fs − 1 j s 2 Sg). Then every element of I is a sum of elements of 0 0 ±1 Pn−1 0 the form g − g j g = g s . For g 2 G, g − 1 2 I, so we have a finite sum g − 1 = i=1 (gi − gi). Since this is a sum of elements in G where G is written multiplicatively, 9 i0 j gi0 = g, say i0 = 1. 0 Pn−1 0 0 Pn−1 0 0 Thus g − 1 = g − g1 + i=2 (gi − gi) ) g1 − 1 = i=2 (gi − gi). Another iteration yields g2 = g1 and 0 ±1 ±1 0 ±1 ±1 0 0 0 0 g = g1 = g1s1 = g2s1 = g2s2 s1 . At the last iteration, gn−2 − 1 = gn−1 − gn−1 ) gn−1 = 1, gn−2 = 0 ±1 ±1 ±1 gn−1 = gn−1sn−1 = 1sn−1. Through this method we obtain a sequence g1; g2; :::; gn−1; gn j gi = gi+1si and g1 = g; gn = 1. ) g has a representation in terms of elements of S and so G = hSi since g was arbitrary. 2.1(d): If G is finitely generated, then by part(b) above, I is finitely generated. For the converse, suppose I = (a1; :::; an) is a left ideal over ZG. Noting from part(a) that I = hg − 1ig2G as a Z-module, P each ai can be represented as a finite sum ai = j zj(gj −1). Since each ai is generated by finitely many elements, and there are finitely many ai, I is finitely generated as a left ideal by elements s − 1 where s 2 G. Therefore, we apply part(c) to have G = hs1; : : : ; ski and thus G is a finitely generated group. 2.2: Let G = hti with jGj = n and let t be the image of T in R = Z[T ]=(T n − 1) =∼ ZG. The el- ement T − 1 is prime in Z[T ] because Z[T ]=(T − 1) =∼ Z which is an integral domain (An ideal P of a commutative ring R is prime iff the quotient ring R=P is an integral domain; Proposition 7.4.13[2]). By Proposition 8.3.10[2] (In an integral domain a prime element is always irreducible), T − 1 is irre- ducible in Z[T ]. We also could have obtained this result by applying Eisenstein's Criterion with the substitution T = x + 3 and using the prime 2. Since Z[T ] is a Unique Factorization Domain, the specific factorization T n − 1 = (T − 1)(T n−1 + T n−2 + ··· + T + 1) with the irreducible T − 1 factor is unique, Pn−1 i considering the latter factor i=0 T as the expansion of its irreducible factors [note: if n is prime then Pn−1 i i=0 T = Φn(T ), a cyclotomic polynomial which is irreducible in Z[T ] by Theorem 13.6.41[2]]. Thus Pn−1 i every f 2 R is annihilated by t − 1 iff it is divisible by N = i=0 t (and vice versa), and so the desired free resolution of M = Z = ZG=(t − 1) is: ···! R −!t−1 R −!N R −!t−1 R ! M ! 0 . F 3.1: The right cosets Hgi are H-orbits of G with the H-action as group multiplication. Since G = Hgi L ∼ L −1 where gi ranges over E, ZG = Z[Hgi] = Z[H=Hgi ]. G is a free H-set because hg = g ) hgg = −1 gg ) h = 1, i.e. the isotropy groups Hg are trivial. Therefore, ZG is a free ZH-module with basis E. P 3.2: Let hSi = H ⊆ G and consider Z[G=H]. Now x 2 Z[G=H] has the expression x = zi(giH), and there exists an element fixed by H, namely, x0 = g0H = H where g0 2 H. H is annihilated by I = Ker" = (fs − 1g) since (s − 1)H = sH − H = H − H = 0 8s 2 S, and so Ix0 = 0. We have g − 1 2 I since "(g − 1) = "(g) − "(1) = 1 − 1 = 0. Hence (g − 1)x0 = 0 ) gx0 = x0 8g 2 G ) Gx0 = x0. Finally, GH = H ) G ⊆ H. ) G = H = hSi. n 4.1: Orienting each n-cell e gives a basis for Cn(X). If X is an arbitrary G-complex, then with 5 P n n ηiei 2 Cn(X), g 2 G can reverse the orientation of e by inversion (fixing the cell). Thus G need not permute the basis, and hence Cn(X) is not necessarily a permutation module. 4.2: Since X is a free G-complex, it is necessarily a Hausdorff space with no fixed points under the G-action. First, assume G is finite and take the set of elements in Gx0, which are distinct points gix0 with 1x0 = x0 for an arbitrary point x0 2 X. Applying the Hausdorff condition, we have open sets Ugi containing gix0 where each such set is disjoint from U1 containing x0. Form the intersection T −1 W = ( i gi Ugi) \ U1 which contains x0. Since gkW ⊆ Ugk, we have gkW \ W = ? for all nonidentity gk 2 G, and so W is the desired open neighborhood of x0 [This result does not follow for arbitrary G since an infinite intersection of open sets need not be open]. Assume the result has been proved for G infinite. Let ' : X ! X=G be the quotient map, which sends the disjoint collection of giW 's to '(W ). Since −1 ` ' '(W ) = i giW , giW ! '(W ) is a bijective map (restriction of ') and thus it is a homeomorphism (' continuous and open). This covering space is regular because G acts transitively on '−1(Gx) by def- inition. Elements of G are obviously deck transformations since Ggx = Gx, hence G ⊆ Aut(X). Given Γ 2 Aut(X) with Γ(a) = b, those two points are mapped to the same orbit in X=G (since ' ◦ Γ = '), and so 9 g 2 G sending a to b. By the Lifting Lemma (uniqueness) we have Γ = g, hence Aut(X) ⊆ G ) G is the group of covering transformations. ∼ ∼ If X is contractible then G = π1(X=G)/π1(X) = π1(X=G)=0 = π1(X=G) and ' is the universal cover of X=G, so X=G is a K(G; 1) with universal cover X.
Recommended publications
  • Classification of Finite Abelian Groups
    Math 317 C1 John Sullivan Spring 2003 Classification of Finite Abelian Groups (Notes based on an article by Navarro in the Amer. Math. Monthly, February 2003.) The fundamental theorem of finite abelian groups expresses any such group as a product of cyclic groups: Theorem. Suppose G is a finite abelian group. Then G is (in a unique way) a direct product of cyclic groups of order pk with p prime. Our first step will be a special case of Cauchy’s Theorem, which we will prove later for arbitrary groups: whenever p |G| then G has an element of order p. Theorem (Cauchy). If G is a finite group, and p |G| is a prime, then G has an element of order p (or, equivalently, a subgroup of order p). ∼ Proof when G is abelian. First note that if |G| is prime, then G = Zp and we are done. In general, we work by induction. If G has no nontrivial proper subgroups, it must be a prime cyclic group, the case we’ve already handled. So we can suppose there is a nontrivial subgroup H smaller than G. Either p |H| or p |G/H|. In the first case, by induction, H has an element of order p which is also order p in G so we’re done. In the second case, if ∼ g + H has order p in G/H then |g + H| |g|, so hgi = Zkp for some k, and then kg ∈ G has order p. Note that we write our abelian groups additively. Definition. Given a prime p, a p-group is a group in which every element has order pk for some k.
    [Show full text]
  • Faithful Abelian Groups of Infinite Rank Ulrich Albrecht
    PROCEEDINGS OF THE AMERICAN MATHEMATICAL SOCIETY Volume 103, Number 1, May 1988 FAITHFUL ABELIAN GROUPS OF INFINITE RANK ULRICH ALBRECHT (Communicated by Bhama Srinivasan) ABSTRACT. Let B be a subgroup of an abelian group G such that G/B is isomorphic to a direct sum of copies of an abelian group A. For B to be a direct summand of G, it is necessary that G be generated by B and all homomorphic images of A in G. However, if the functor Hom(A, —) preserves direct sums of copies of A, then this condition is sufficient too if and only if M ®e(A) A is nonzero for all nonzero right ¿ï(A)-modules M. Several examples and related results are given. 1. Introduction. There are only very few criteria for the splitting of exact sequences of torsion-free abelian groups. The most widely used of these was given by Baer in 1937 [F, Proposition 86.5]: If G is a pure subgroup of a torsion-free abelian group G, such that G/C is homogeneous completely decomposable of type f, and all elements of G\C are of type r, then G is a direct summand of G. Because of its numerous applications, many attempts have been made to extend the last result to situations in which G/C is not completely decomposable. Arnold and Lady succeeded in 1975 in the case that G is torsion-free of finite rank. Before we can state their result, we introduce some additional notation: Suppose that A and G are abelian groups.
    [Show full text]
  • Arxiv:1511.02907V1 [Math.AT] 9 Nov 2015 Oal Cci Ope Fpoetvs Hs R Xc Complexes Exact Algebr Are Homological These Gorenstein Projectives
    THE PROJECTIVE STABLE CATEGORY OF A COHERENT SCHEME SERGIO ESTRADA AND JAMES GILLESPIE Abstract. We define the projective stable category of a coherent scheme. It is the homotopy category of an abelian model structure on the category of unbounded chain complexes of quasi-coherent sheaves. We study the cofibrant objects of this model structure, which are certain complexes of flat quasi- coherent sheaves satisfying a special acyclicity condition. 1. introduction Let R be a ring and R-Mod the category of left R-modules. The projective stable module category of R was introduced in [BGH13]. The construction provides a triangulated category Sprj and a product-preserving functor γ : R-Mod −→Sprj taking short exact sequences to exact triangles, and kills all injective and projective modules (but typically will kill more than just these modules). The motivation for this paper is to extend this construction to schemes, that is, to replace R with a scheme X and introduce the projective stable (quasi-coherent sheaf) category of a scheme X. Although we don’t yet understand the situation in full generality, this paper does make significant progress towards this goal. Let us back up and explain the projective stable module category of a ring R. The idea is based on a familiar concept in Gorenstein homological algebra, that of a totally acyclic complex of projectives. These are exact complexes P of projective R-modules such that HomR(P,Q) is also exact for all projective R-modules Q. Such complexes historically arose in group cohomology theory, since the Tate cohomology groups are defined using totally acyclic complexes of projectives.
    [Show full text]
  • Homotopy Equivalences and Free Modules Steven
    Topology, Vol. 21, No. 1, pp. 91-99, 1982 0040-93831821010091-09502.00/0 Printed in Great Britain. Pergamon Press Ltd. HOMOTOPY EQUIVALENCES AND FREE MODULES STEVEN PLOTNICKt:~ (Received 15 December 1980) INTRODUCTION IN THIS PAPER, we are concerned with the problem of determining the group of self-homotopy equivalences of some familiar spaces. We have in mind the following spaces: Let ~- be a finite group acting freely and cellulary on a CW-complex with the homotopy type of S 2m+1. Removing a point from S2m+1/Tr yields a 2m-complex X. We will (theoretically) describe the self-homotopy equivalences of X. (I use the word theoretically since the results is given in terms of units in the integral group ring Z Tr and certain projective modules over ZTr, and these are far from being completely understood.) Once this has been done, it will be fairly straightforward to describe the self-homotopy equivalences of $2~+1/7r and of the 2m-complex which results from removing more than one point from s2m+l/Tr (i.e. X v v $2~). As one might expect, the more points one removes, the easier it is to realize automorphisms of ~r by homotopy equivalences. It seems quite nice that this can be phrased as a question in Ko(Z~). The results can be summarised as follows. First, some standard facts, definitions, and notation: If ~r has order n and acts freely on S :m+~, then H:m+1(~r; Z)m Zn. Thus, a E Aut ~r acts on H2m+l(~r) via multiplication by an integer r~, where (r~, n) = 1.
    [Show full text]
  • Functor Homology and Operadic Homology
    FUNCTOR HOMOLOGY AND OPERADIC HOMOLOGY BENOIT FRESSE Abstract. The purpose of these notes is to define an equivalence between the natural homology theories associated to operads and the homology of functors over certain categories of operators (PROPs) related to operads. Introduction The aim of these notes is to prove that the natural homology theory associated to an operad is equivalent the homology of a category of functors over a certain category of operators associated to our operad. Recall that an operad P in a symmetric monoidal category C basically consists of a sequence of objects P(n) 2 C, n 2 N, of which elements p 2 P(n) (whenever the notion of an element makes sense) intuitively represent operations on n inputs and with 1 output: p A⊗n = A ⊗ · · · ⊗ A −! A; | {z } n for any n 2 N. In short, an operad is defined axiomatically as such a sequence of objects P = fP(n); n 2 Ng equipped with an action of the symmetric group Σn on the term P(n), for each n 2 N, together with composition products which are shaped on composition schemes associate with such operations. The notion of an operad is mostly used to define a category of algebras, which basically consists of objects A 2 C on which the operations of our operad p 2 P(n) act. We use the term of a P-algebra, and the notation P C, to refer to this category of algebras associated to any given operad P. Recall simply that the usual category of associative algebras in a category of modules over a ring k, the category of (associative and) commutative algebras, and the category of Lie algebras, are associated to operads, which we respectively denote by P = As; Com; Lie.
    [Show full text]
  • Algebraic Number Theory
    Algebraic Number Theory Sergey Shpectorov January{March, 2010 This course in on algebraic number theory. This means studying problems from number theory with methods from abstract algebra. For a long time the main motivation behind the development of algebraic number theory was the Fermat Last Theorem. Proven in 1995 by Wiles with the help of Taylor, this theorem states that there are no positive integers x, y and z satisfying the equation xn + yn = zn; where n ≥ 3 is an integer. The proof of this statement for the particular case n = 4 goes back to Fibonacci, who lived four hundred years before Fermat. Modulo Fibonacci's result, Fermat Last Theorem needs to be proven only for the cases where n = p is an odd prime. By the end of the course we will hopefully see, as an application of our theory, how to prove the Fermat Last Theorem for the so-called regular primes. The idea of this belongs to Kummer, although we will, of course, use more modern notation and methods. Another accepted definition of algebraic number theory is that it studies the so-called number fields, which are the finite extensions of the field of ra- tional numbers Q. We mention right away, however, that most of this theory applies also in the second important case, known as the case of function fields. For example, finite extensions of the field of complex rational functions C(x) are function fields. We will stress the similarities and differences between the two types of fields, as appropriate. Finite extensions of Q are algebraic, and this ties algebraic number the- ory with Galois theory, which is an important prerequisite for us.
    [Show full text]
  • Projective Dimensions and Almost Split Sequences
    View metadata, citation and similar papers at core.ac.uk brought to you by CORE provided by Elsevier - Publisher Connector Journal of Algebra 271 (2004) 652–672 www.elsevier.com/locate/jalgebra Projective dimensions and almost split sequences Dag Madsen 1 Department of Mathematical Sciences, NTNU, NO-7491 Trondheim, Norway Received 11 September 2002 Communicated by Kent R. Fuller Abstract Let Λ be an Artin algebra and let 0 → A → B → C → 0 be an almost split sequence. In this paper we discuss under which conditions the inequality pd B max{pd A,pd C} is strict. 2004 Elsevier Inc. All rights reserved. Keywords: Almost split sequences; Homological dimensions Let R be a ring. If we have an exact sequence ε :0→ A → B → C → 0ofR-modules, then pd B max{pdA,pd C}. In some cases, for instance if the sequence is split exact, equality holds, but in general the inequality may be strict. In this paper we will discuss a problem first considered by Auslander (see [5]): Let Λ be an Artin algebra (for example a finite dimensional algebra over a field) and let ε :0→ A → B → C → 0 be an almost split sequence. To which extent does the equality pdB = max{pdA,pd C} hold? Given an exact sequence 0 → A → B → C → 0, we investigate in Section 1 what can be said in complete generality about when pd B<max{pdA,pd C}.Inthissectionwedo not make any assumptions on the rings or the exact sequences involved. In all sections except Section 1 the rings we consider are Artin algebras.
    [Show full text]
  • BLECHER • All Numbers Below Are Integers, Usually Positive. • Read Up
    MATH 4389{ALGEBRA `CHEATSHEET'{BLECHER • All numbers below are integers, usually positive. • Read up on e.g. wikipedia on the Euclidean algorithm. • Diophantine equation ax + by = c has solutions iff gcd(a; b) doesn't divide c. One may find the solutions using the Euclidean algorithm (google this). • Euclid's lemma: prime pjab implies pja or pjb. • gcd(m; n) · lcm(m; n) = mn. • a ≡ b (mod m) means mj(a − b). An equivalence relation (reflexive, symmetric, transitive) on the integers. • ax ≡ b (mod m) has a solution iff gcd(m; a)jb. • Division Algorithm: there exist unique integers q and r with a = bq + r and 0 ≤ r < b; so a ≡ r (mod b). • Fermat's theorem: If p is prime then ap ≡ a (mod p) Also, ap−1 ≡ 1 (mod p) if p does not divide a. • For definitions for groups, rings, fields, subgroups, order of element, etc see Algebra fact sheet" below. • Zm = f0; 1; ··· ; m−1g with addition and multiplication mod m, is a group and commutative ring (= Z =m Z) • Order of n in Zm (i.e. smallest k s.t. kn ≡ 0 (mod m)) is m=gcd(m; n). • hai denotes the subgroup generated by a (the smallest subgroup containing a). For example h2i = f0; 2; 2+2 = 4; 2 + 2 + 2 = 6g in Z8. (In ring theory it may be the subring generated by a). • A cyclic group is a group that can be generated by a single element (the group generator). They are abelian. • A finite group is cyclic iff G = fg; g + g; ··· ; g + g + ··· + g = 0g; so G = hgi.
    [Show full text]
  • Arxiv:2010.00369V6 [Math.KT] 29 Jun 2021 H Uco Faeinzto Se[8 H II, Derived Simplicial Ch
    ON HOMOLOGY OF LIE ALGEBRAS OVER COMMUTATIVE RINGS SERGEI O. IVANOV, FEDOR PAVUTNITSKIY, VLADISLAV ROMANOVSKII, AND ANATOLII ZAIKOVSKII Abstract. We study five different types of the homology of a Lie algebra over a commutative ring which are naturally isomorphic over fields. We show that they are not isomorphic over commutative rings, even over Z, and study connections between them. In particular, we show that they are naturally isomorphic in the case of a Lie algebra which is flat as a module. As an auxiliary result we prove that the Koszul complex of a module M over a principal ideal domain that connects the exterior and the symmetric powers n n−1 n−1 n 0 → Λ M → M ⊗ Λ M → ⋅⋅⋅ → S M ⊗ M → S M → 0 is purely acyclic. Introduction There are several equivalent purely algebraic definitions of group homology of a group G. Namely one can define it using Tor functor over the group ring; or the relative Tor functor; or in the simplicial manner, as a simplicial derived functor of the functor of abelianization (see [28, Ch. II, §5]). For Lie algebras over a field we can also define homology in several equivalent ways, including the homology of the Chevalley–Eilenberg complex. Moreover, for a Lie algebra g over a commutative ring k, which is free as a k-module we also have several equivalent definitions (see [7, Ch. XIII]). However, in general, even in the case k = Z, these definitions are not equivalent. For example, if we consider g = Z 2 as an abelian Lie algebra over Z, the / Ug Z Z higher homology of the Chevalley–Eilenberg complex vanishes but Tor2n+1 , = Z 2 for any n ≥ 0.
    [Show full text]
  • Orders on Computable Torsion-Free Abelian Groups
    Orders on Computable Torsion-Free Abelian Groups Asher M. Kach (Joint Work with Karen Lange and Reed Solomon) University of Chicago 12th Asian Logic Conference Victoria University of Wellington December 2011 Asher M. Kach (U of C) Orders on Computable TFAGs ALC 2011 1 / 24 Outline 1 Classical Algebra Background 2 Computing a Basis 3 Computing an Order With A Basis Without A Basis 4 Open Questions Asher M. Kach (U of C) Orders on Computable TFAGs ALC 2011 2 / 24 Torsion-Free Abelian Groups Remark Disclaimer: Hereout, the word group will always refer to a countable torsion-free abelian group. The words computable group will always refer to a (fixed) computable presentation. Definition A group G = (G : +; 0) is torsion-free if non-zero multiples of non-zero elements are non-zero, i.e., if (8x 2 G)(8n 2 !)[x 6= 0 ^ n 6= 0 =) nx 6= 0] : Asher M. Kach (U of C) Orders on Computable TFAGs ALC 2011 3 / 24 Rank Theorem A countable abelian group is torsion-free if and only if it is a subgroup ! of Q . Definition The rank of a countable torsion-free abelian group G is the least κ cardinal κ such that G is a subgroup of Q . Asher M. Kach (U of C) Orders on Computable TFAGs ALC 2011 4 / 24 Example The subgroup H of Q ⊕ Q (viewed as having generators b1 and b2) b1+b2 generated by b1, b2, and 2 b1+b2 So elements of H look like β1b1 + β2b2 + α 2 for β1; β2; α 2 Z.
    [Show full text]
  • Cohomology Theory of Lie Groups and Lie Algebras
    COHOMOLOGY THEORY OF LIE GROUPS AND LIE ALGEBRAS BY CLAUDE CHEVALLEY AND SAMUEL EILENBERG Introduction The present paper lays no claim to deep originality. Its main purpose is to give a systematic treatment of the methods by which topological questions concerning compact Lie groups may be reduced to algebraic questions con- cerning Lie algebras^). This reduction proceeds in three steps: (1) replacing questions on homology groups by questions on differential forms. This is accomplished by de Rham's theorems(2) (which, incidentally, seem to have been conjectured by Cartan for this very purpose); (2) replacing the con- sideration of arbitrary differential forms by that of invariant differential forms: this is accomplished by using invariant integration on the group manifold; (3) replacing the consideration of invariant differential forms by that of alternating multilinear forms on the Lie algebra of the group. We study here the question not only of the topological nature of the whole group, but also of the manifolds on which the group operates. Chapter I is concerned essentially with step 2 of the list above (step 1 depending here, as in the case of the whole group, on de Rham's theorems). Besides consider- ing invariant forms, we also introduce "equivariant" forms, defined in terms of a suitable linear representation of the group; Theorem 2.2 states that, when this representation does not contain the trivial representation, equi- variant forms are of no use for topology; however, it states this negative result in the form of a positive property of equivariant forms which is of interest by itself, since it is the key to Levi's theorem (cf.
    [Show full text]
  • The Exchange Property and Direct Sums of Indecomposable Injective Modules
    Pacific Journal of Mathematics THE EXCHANGE PROPERTY AND DIRECT SUMS OF INDECOMPOSABLE INJECTIVE MODULES KUNIO YAMAGATA Vol. 55, No. 1 September 1974 PACIFIC JOURNAL OF MATHEMATICS Vol. 55, No. 1, 1974 THE EXCHANGE PROPERTY AND DIRECT SUMS OF INDECOMPOSABLE INJECTIVE MODULES KUNIO YAMAGATA This paper contains two main results. The first gives a necessary and sufficient condition for a direct sum of inde- composable injective modules to have the exchange property. It is seen that the class of these modules satisfying the con- dition is a new one of modules having the exchange property. The second gives a necessary and sufficient condition on a ring for all direct sums of indecomposable injective modules to have the exchange property. Throughout this paper R will be an associative ring with identity and all modules will be right i?-modules. A module M has the exchange property [5] if for any module A and any two direct sum decompositions iel f with M ~ M, there exist submodules A\ £ At such that The module M has the finite exchange property if this holds whenever the index set I is finite. As examples of modules which have the exchange property, we know quasi-injective modules and modules whose endomorphism rings are local (see [16], [7], [15] and for the other ones [5]). It is well known that a finite direct sum M = φj=1 Mt has the exchange property if and only if each of the modules Λft has the same property ([5, Lemma 3.10]). In general, however, an infinite direct sum M = ®i&IMi has not the exchange property even if each of Λf/s has the same property.
    [Show full text]