ALGEBRAIC TORSION 1. Introduction 1.1. Definition of K 1(R)

Total Page:16

File Type:pdf, Size:1020Kb

ALGEBRAIC TORSION 1. Introduction 1.1. Definition of K 1(R) ALGEBRAIC TORSION 1. Introduction 1.1. Definition of K1(R). By a ring we always mean an associative ring R with 1 6= 0. For such a ring R define GL(R) to be the group of all N × N-matrices, that are the infinite identity matrix except for finitely many entries. More rigorously, GL(R) = lim GL(R; d), where we have the following maps in the directed system: ! GL(R; d) ! GL(R; d + 1) A 0 A 7! : 0 1 Given a 2 R and i 6= j 2 f1; : : : ; dg we denote by Eij the d × d-matrix which is zero except for the (i; j)-entry which equals 1. We refer to all matrices of form id +aEij with a 2 R as elementary matrices. We denote by E(R) the subgroup of GL(R) generated by all elementary matrices. In the following, given a group G and g; h 2 G we denote by [g; h] = ghg−1h−1 the commutator. Furthermore, we denote by n n o Y [G; G] := [gi; hi] g1; h1; : : : ; gn; hn 2 G i=1 the commutator subgroup of G. The following proposition gives a convenient descrip- tion of the commutator subgroup of GL(R). Proposition 1.1. E(R) = [GL(R); GL(R)]: Proof. For i 6= j 6= k one can easily verify that id +aEik = (id +aEij)(id +Ejk)(id −aEij)(id −Ejk) −1 −1 = (id +aEij)(id +Ejk)(id +aEij) (id +Ejk) : This shows that each elementary matrix in GL(d; A) with d ≥ 3 is indeed a commu- tator. Conversely, using the following three identities one can easily show that each com- mutator XYX−1Y −1 2 GL(d; R) is a product of elementary matrices within the Date: July 2, 2015. 2010 Mathematics Subject Classification. Primary 57M27; Secondary 57Q10. 1 2 ALGEBRAIC TORSION group GL(2d; R): XYX−1Y −1 0 X 0 Y 0 (YX)−1 0 (1) = ; 0 id 0 X−1 0 Y −1 0 YX X 0 id X id 0 id − id id 0 (2) = ; 0 X−1 0 id id −X−1 id 0 id id −X id d 2d id X Q Q (3) = (id +xijEij): 0 id i=1 j=d+1 The Whitehead group K1(R) is defined as the abelian group K1(R) := GL(R)=[GL(R); GL(R)] = GL(R)=E(R): For every d there exists a canonical homomorphism GL(R; d) ! K1(R), in particular there exists a homomorphism from the units of R into K1(R). By abuse of notation we denote the image of A 2 GL(R; d) in K1(R) by A as well. Lemma 1.2. For any two matrices A 2 GL(R; d1) and B 2 GL(R; d2) the product AB 2 K1(R) is given by A 0 AB = 2 K (R): 0 B 1 Proof. Without loss of generality we can assume that d1 = d2. Then AB 0 A 0 B 0 AB = = 2 K (R): 0 id 0 B 0 B−1 1 In the proof of Proposition 1.1 we had seen that the second matrix on the right hand side is trivial in K1(R). The proof of the following lemma is left as an exercise. Lemma 1.3. Let a; b be units in R. Then 0 a −a 0 and b 0 0 b represent the same element in K1(R). Corollary 1.4. If F is a commutative field, then the map ∗ K1(F) ! F := F n f0g [A] 7! det(A) is an isomorphism. In the following, given a commutative field F, we will use the isomorphism from ∗ Corollary 1.4 to identify K1(F) with F . ALGEBRAIC TORSION 3 Proof. It is straightforward to see that the map is a well-defined epimorphism. It remains to show that the map is injective. So let A be any matrix in GL(d; F). Using elementary linear algebra and Lemmas 1.2 and 1.3 one can show that A and the 1 × 1-matrix (det A) represent the same element in K1(F). It follows that the map under consideration is a monomorphism. In the following we will mostly work with the reduced Whitehead group K1(R) = K1(R)=subgroup of order 2 generated by (−1): On several occasions we will make use of the fact that a ring homomorphism ': R ! S induces group homomorphisms '∗ : K1(R) ! K1(S) and '∗ : K1(R) ! K1(S). Also, the maps GL(d; M(n × n; R)) ! GL(dn; R) A viewed as an d × d-matrix A with entries 7! nd × nd-matrix that are n × n-matrices over R with entries in R are homomorphisms that induce an isomorphism K1(M(n × n; R)) ! K1(R) that is sometimes referred to as Morita equivalence. 1.2. Definition of algebraic torsion. Henceforth we will assume that all rings in these lecture notes have the following property: r s if r 6= s 2 N0, then R is not isomorphic to R as an R{module. It is clear that any field, Z, and more generally any domain have this property. The following lemma gives another important example. Lemma 1.5. If G is a group, then the group ring Z[G] has the above property. r ∼ s Proof. We view Z as a Z[G]-module via the trivial action of G on Z. If Z[G] = Z[G] , then we also have r ∼ r ∼ s ∼ s Z = Z ⊗Z[G] Z[G] = Z ⊗Z[G] Z[G] = Z ; hence r = s. In the following, given a ring R we mean by an R-module a finitely generated left R-module, unless we say explicitly otherwise. We view elements in Rn as row vectors, and an n × m-matrix over R gives rise to a homomorphism Rn ! Rm of left R-modules via right multiplication on row vectors. A based R-module is a free R-module together with an (unordered) basis. Given a free R-module V and two bases d = fd1; : : : ; dng and e = fe1; : : : ; eng we denote by [d=e] 2 K1(R) the element represented by the matrix of the basis change, P i.e. [d=e] = (aij) where di = j aijej. We say that two bases are equivalent if [d=e] is trivial in K1(R). 4 ALGEBRAIC TORSION p Given a short exact sequence 0 ! A −!i B −! C ! 0 of free R-modules and bases A = faig for A and C = fcjg for C we pick liftsc ~j to B and we denote by AC the equivalence classes of bases represented by the ai and thec ~j. It is straightforward to verify that AC does not depend on the choice of the lifts. If B is also equipped with a basis B, then we say that the bases A; B and C are compatible, if B and AC are equivalent. Let (C∗; C∗) be a finite chain complex of based R{modules. In the following we will associate to (C∗; C∗) its torsion τ(C∗; C∗) 2 K1(R) [ {∗}: If C∗ is not acyclic, then we define its torsion τ(C∗; C∗) to be ∗. Now suppose that C∗ is acyclic. In the following we make the technical assumption that each Bi := Im(Ci+1) ⊂ Ci is a free R-module. For each i we pick a basis Bi of Bi. Since C∗ is acyclic the sequences 0 ! Bi ! Ci ! Bi−1 ! 0 are exact. We define the torsion (sometimes also called Reidemeister-torsion) of the based acyclic complex (C∗; C∗) to be Y (−1)i+1 τ(C∗; C∗) = [BiBi−1=Ci] 2 K1(R): It is straightforward to see that τ(C∗; C∗) is independent of the choice of the Bi. If the R{modules Bi are not free, then one can show that they are stably free and a stable basis will then make the definition work again. We refer to [Mi66, p. 369] or [Tu01, p. 13] for the full details.1 For future reference we record the simplest example in the following lemma. @k Lemma 1.6. If 0 ! Ck −! Ck−1 ! 0 is an isomorphism between free based R- modules of rank n, then (−1)k−1 τ(C∗) = [Bk] 2 K1(R); where Bk denotes the n × n-matrix over R which represents @k with respect to the given bases. The statement of the lemma is an immediate consequence of the definitions. It says in particular that the torsion of a chain complex of based R-modules can be viewed as the generalization of the determinant of a homomorphism between two based R-modules. Example. Let A be an n × n-matrix over Z with det(A) 6= 0. Then n A n τ 0 ! Q −! Q ! 0 = det(A) 1The torsion introduced in [Tu01] is the multiplicative inverse of the torsion introduced in [Mi66]. We follow the convention of [Mi66]. ALGEBRAIC TORSION 5 which up to sign equals the number of elements in the finite abelian group Zn=AZn. This explains, perhaps, the name torsion for the invariant τ(C∗). The term `torsion' was introduced by Franz [Fra35]. In the following we extend the multiplication of K1(R) to τ(C∗; C∗) 2 K1(R) [ {∗} by defining the product of 0 with any other element to be 0. The following elementary lemma now shows how the torsion depends on the choice of the bases. 0 Lemma 1.7. Let C∗ be a chain complex of free R-modules and let C∗ and C∗ be two choices of bases.
Recommended publications
  • Whitehead Torsion, Part II (Lecture 4)
    Whitehead Torsion, Part II (Lecture 4) September 9, 2014 In this lecture, we will continue our discussion of the Whitehead torsion of a homotopy equivalence f : X ! Y between finite CW complexes. In the previous lecture, we gave a definition in the special case where f is cellular. To remove this hypothesis, we need the following: Proposition 1. Let X and Y be connected finite CW complexes and suppose we are given cellular homotopy equivalences f; g : X ! Y . If f and g are homotopic, then τ(f) = τ(g) 2 Wh(π1X). Lemma 2. Suppose we are given quasi-isomorphisms f :(X∗; d) ! (Y∗; d) and g :(Y∗; d) ! (Z∗; d) between finite based complexes with χ(X∗; d) = χ(Y∗; d) = χ(Z∗; d). Then τ(g ◦ f) = τ(g)τ(f) in Ke1(R). Proof. We define a based chain complex (W∗; d) by the formula W∗ = X∗−1 ⊕ Y∗ ⊕ Y∗−1 ⊕ Z∗ d(x; y; y0; z) = (−dx; f(x) + dy + y0; −dy0; g(y0) + dz): Then (W∗; d) contains (C(f)∗; d) as a based subcomplex with quotient (C(g)∗; d), so an Exercise from the previous lecture gives τ(W∗; d) = τ(g)τ(f). We now choose a new basis for each W∗ by replacing each basis element of y 2 Y∗ by (0; y; 0; g(y)); this is an upper triangular change of coordinates and therefore does not 0 0 affect the torsion τ(W∗; d). Now the construction (y ; y) 7! (0; y; y ; g(y)) identifies C(idY )∗ with a based subcomplex of W∗ having quotient C(−g ◦ f)∗.
    [Show full text]
  • Math200b, Lecture 11
    Math200b, lecture 11 Golsefidy We have proved that a module is projective if and only if it is a direct summand of a free module; in particular, any free module is projective. For a given ring A, we would like to know to what extent the converse of this statement holds; and if it fails, we would like to somehow “measure" how much it does! In genral this is hard question; in your HW assignement you will show that for a local commutative ring A, any finitely generated projective module is free. By a result of Kaplansky the same statement holds for a module that is not necessarily finitely generated. Next we show that for a PID, any finitely generated projective module is free. Proposition 1 Let D be an integral domain. Then free D-module ) Projective ) torsion-free. 1 If D is a PID, for a finitely generated D-module all the above proper- ties are equivalent. Proof. We have already discussed that a free module is projec- tive. A projective module is a direct summand of a free module and a free module of an integral domain is torsion free. By the fundamental theorem of finitely generated modules over a PID, a torsion free finitely generated module over D is free; and claim follows. p Next we show A » −10¼ has a finitely generated projec- tive module that is not free. In fact any ideal of A is projective; and since it is not a PID, it has an ideal that is not free. Based on the mentioned result of Kaplansky, a projective module is locally free.
    [Show full text]
  • Braids in Pau -- an Introduction
    ANNALES MATHÉMATIQUES BLAISE PASCAL Enrique Artal Bartolo and Vincent Florens Braids in Pau – An Introduction Volume 18, no 1 (2011), p. 1-14. <http://ambp.cedram.org/item?id=AMBP_2011__18_1_1_0> © Annales mathématiques Blaise Pascal, 2011, tous droits réservés. L’accès aux articles de la revue « Annales mathématiques Blaise Pas- cal » (http://ambp.cedram.org/), implique l’accord avec les condi- tions générales d’utilisation (http://ambp.cedram.org/legal/). Toute utilisation commerciale ou impression systématique est constitutive d’une infraction pénale. Toute copie ou impression de ce fichier doit contenir la présente mention de copyright. Publication éditée par le laboratoire de mathématiques de l’université Blaise-Pascal, UMR 6620 du CNRS Clermont-Ferrand — France cedram Article mis en ligne dans le cadre du Centre de diffusion des revues académiques de mathématiques http://www.cedram.org/ Annales mathématiques Blaise Pascal 18, 1-14 (2011) Braids in Pau – An Introduction Enrique Artal Bartolo Vincent Florens Abstract In this work, we describe the historic links between the study of 3-dimensional manifolds (specially knot theory) and the study of the topology of complex plane curves with a particular attention to the role of braid groups and Alexander-like invariants (torsions, different instances of Alexander polynomials). We finish with detailed computations in an example. Tresses à Pau – une introduction Résumé Dans ce travail, nous décrivons les liaisons historiques entre l’étude de variétés de dimension 3 (notamment, la théorie de nœuds) et l’étude de la topologie des courbes planes complexes, dont l’accent est posé sur le rôle des groupes de tresses et des invariantes du type Alexander (torsions, différents incarnations des polynômes d’Alexander).
    [Show full text]
  • Lecture 15. De Rham Cohomology
    Lecture 15. de Rham cohomology In this lecture we will show how differential forms can be used to define topo- logical invariants of manifolds. This is closely related to other constructions in algebraic topology such as simplicial homology and cohomology, singular homology and cohomology, and Cechˇ cohomology. 15.1 Cocycles and coboundaries Let us first note some applications of Stokes’ theorem: Let ω be a k-form on a differentiable manifold M.For any oriented k-dimensional compact sub- manifold Σ of M, this gives us a real number by integration: " ω : Σ → ω. Σ (Here we really mean the integral over Σ of the form obtained by pulling back ω under the inclusion map). Now suppose we have two such submanifolds, Σ0 and Σ1, which are (smoothly) homotopic. That is, we have a smooth map F : Σ × [0, 1] → M with F |Σ×{i} an immersion describing Σi for i =0, 1. Then d(F∗ω)isa (k + 1)-form on the (k + 1)-dimensional oriented manifold with boundary Σ × [0, 1], and Stokes’ theorem gives " " " d(F∗ω)= ω − ω. Σ×[0,1] Σ1 Σ1 In particular, if dω =0,then d(F∗ω)=F∗(dω)=0, and we deduce that ω = ω. Σ1 Σ0 This says that k-forms with exterior derivative zero give a well-defined functional on homotopy classes of compact oriented k-dimensional submani- folds of M. We know some examples of k-forms with exterior derivative zero, namely those of the form ω = dη for some (k − 1)-form η. But Stokes’ theorem then gives that Σ ω = Σ dη =0,sointhese cases the functional we defined on homotopy classes of submanifolds is trivial.
    [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]
  • Homological Mirror Symmetry for the Genus 2 Curve in an Abelian Variety and Its Generalized Strominger-Yau-Zaslow Mirror by Cath
    Homological mirror symmetry for the genus 2 curve in an abelian variety and its generalized Strominger-Yau-Zaslow mirror by Catherine Kendall Asaro Cannizzo A dissertation submitted in partial satisfaction of the requirements for the degree of Doctor of Philosophy in Mathematics in the Graduate Division of the University of California, Berkeley Committee in charge: Professor Denis Auroux, Chair Professor David Nadler Professor Marjorie Shapiro Spring 2019 Homological mirror symmetry for the genus 2 curve in an abelian variety and its generalized Strominger-Yau-Zaslow mirror Copyright 2019 by Catherine Kendall Asaro Cannizzo 1 Abstract Homological mirror symmetry for the genus 2 curve in an abelian variety and its generalized Strominger-Yau-Zaslow mirror by Catherine Kendall Asaro Cannizzo Doctor of Philosophy in Mathematics University of California, Berkeley Professor Denis Auroux, Chair Motivated by observations in physics, mirror symmetry is the concept that certain mani- folds come in pairs X and Y such that the complex geometry on X mirrors the symplectic geometry on Y . It allows one to deduce information about Y from known properties of X. Strominger-Yau-Zaslow (1996) described how such pairs arise geometrically as torus fibra- tions with the same base and related fibers, known as SYZ mirror symmetry. Kontsevich (1994) conjectured that a complex invariant on X (the bounded derived category of coherent sheaves) should be equivalent to a symplectic invariant of Y (the Fukaya category). This is known as homological mirror symmetry. In this project, we first use the construction of SYZ mirrors for hypersurfaces in abelian varieties following Abouzaid-Auroux-Katzarkov, in order to obtain X and Y as manifolds.
    [Show full text]
  • Some Notes About Simplicial Complexes and Homology II
    Some notes about simplicial complexes and homology II J´onathanHeras J. Heras Some notes about simplicial homology II 1/19 Table of Contents 1 Simplicial Complexes 2 Chain Complexes 3 Differential matrices 4 Computing homology groups from Smith Normal Form J. Heras Some notes about simplicial homology II 2/19 Simplicial Complexes Table of Contents 1 Simplicial Complexes 2 Chain Complexes 3 Differential matrices 4 Computing homology groups from Smith Normal Form J. Heras Some notes about simplicial homology II 3/19 Simplicial Complexes Simplicial Complexes Definition Let V be an ordered set, called the vertex set. A simplex over V is any finite subset of V . Definition Let α and β be simplices over V , we say α is a face of β if α is a subset of β. Definition An ordered (abstract) simplicial complex over V is a set of simplices K over V satisfying the property: 8α 2 K; if β ⊆ α ) β 2 K Let K be a simplicial complex. Then the set Sn(K) of n-simplices of K is the set made of the simplices of cardinality n + 1. J. Heras Some notes about simplicial homology II 4/19 Simplicial Complexes Simplicial Complexes 2 5 3 4 0 6 1 V = (0; 1; 2; 3; 4; 5; 6) K = f;; (0); (1); (2); (3); (4); (5); (6); (0; 1); (0; 2); (0; 3); (1; 2); (1; 3); (2; 3); (3; 4); (4; 5); (4; 6); (5; 6); (0; 1; 2); (4; 5; 6)g J. Heras Some notes about simplicial homology II 5/19 Chain Complexes Table of Contents 1 Simplicial Complexes 2 Chain Complexes 3 Differential matrices 4 Computing homology groups from Smith Normal Form J.
    [Show full text]
  • Infinite Groups with Large Balls of Torsion Elements and Small Entropy
    INFINITE GROUPS WITH LARGE BALLS OF TORSION ELEMENTS AND SMALL ENTROPY LAURENT BARTHOLDI, YVES DE CORNULIER Abstract. We exhibit infinite, solvable, abelian-by-finite groups, with a fixed number of generators, with arbitrarily large balls consisting of torsion elements. We also provide a sequence of 3-generator non-nilpotent-by-finite polycyclic groups of algebraic entropy tending to zero. All these examples are obtained by taking ap- propriate quotients of finitely presented groups mapping onto the first Grigorchuk group. The Burnside Problem asks whether a finitely generated group all of whose ele- ments have finite order must be finite. We are interested in the following related question: fix n sufficiently large; given a group Γ, with a finite symmetric generating subset S such that every element in the n-ball is torsion, is Γ finite? Since the Burn- side problem has a negative answer, a fortiori the answer to our question is negative in general. However, it is natural to ask for it in some classes of finitely generated groups for which the Burnside Problem has a positive answer, such as linear groups or solvable groups. This motivates the following proposition, which in particularly answers a question of Breuillard to the authors. Proposition 1. For every n, there exists a group G, generated by a 3-element subset S consisting of elements of order 2, in which the n-ball consists of torsion elements, and satisfying one of the additional assumptions: (1) G is solvable, virtually abelian, and infinite (more precisely, it has a free abelian normal subgroup of finite 2-power index); in particular it is linear.
    [Show full text]
  • The Cobordism Group of Homology Cylinders
    THE COBORDISM GROUP OF HOMOLOGY CYLINDERS JAE CHOON CHA, STEFAN FRIEDL, AND TAEHEE KIM Abstract. Garoufalidis and Levine introduced the homology cobordism group of homology cylinders over a surface. This group can be regarded as an enlargement of the mapping class group. Using torsion invariants, we show that the abelianization of this group is infinitely generated provided that the first Betti number of the surface is positive. In particular, this shows that the group is not perfect. This answers questions of Garoufalidis-Levine and Goda- Sakasai. Furthermore we show that the abelianization of the group has infinite rank for the case that the surface has more than one boundary component. These results hold for the homology cylinder analogue of the Torelli group as well. 1. Introduction Given g ≥ 0 and n ≥ 0, let Σg;n be a fixed oriented, connected and compact + surface of genus g with n boundary components. We denote by Hom (Σg;n;@Σg;n) the group of orientation preserving diffeomorphisms of Σg;n which restrict to the identity on the boundary. The mapping class group Mg;n is defined to be the + set of isotopy classes of elements in Hom (Σg;n;@Σg;n), where the isotopies are understood to restrict to the identity on the boundary as well. We refer to [FM09, Section 2.1] for details. It is well known that the mapping class group is perfect provided that g ≥ 3 [Po78] (e.g., see [FM09, Theorem 5.1]) and that mapping class groups are finitely presented [BH71, Mc75] (e.g., see [FM09, Section 5.2]).
    [Show full text]
  • Countable Torsion-Free Abelian Groups
    COUNTABLE TORSION-FREE ABELIAN GROUPS By M. O'N. CAMPBELL [Received 27 January 1959.—Read 19 February 1959] Introduction IN this paper we present a new classification of the countable torsion-free abelian groups. Since any abelian group 0 may be regarded as an extension of a periodic group (namely the torsion part of G) by a torsion-free group, and since there exists the well-known complete classification of the count- able periodic abelian groups due to Ulm, it is clear that the classification problem studied here is of great interest. By a group we shall always mean an abelian group, and the additive notation will be employed throughout. It is well known that all the maximal linearly independent sets of elements of a given group are car- dinally equivalent; the corresponding cardinal number is the rank of the group. Derry (2) and Mal'cev (4) have studied the torsion-free groups of finite rank only, and have given classifications of these groups in terms of certain equivalence classes of systems of matrices with £>-adic elements. Szekeres (5) attempted to abolish the finiteness restriction on rank; he extracted certain invariants, but did not derive a complete classification. By a basis of a torsion-free group 0 we mean any maximal linearly independent subset of G. A subgroup generated by a basis of G will be called a basal subgroup or described as basal in G. Thus U is a basal sub- group of G if and only if (i) U is free abelian, and (ii) the factor group G/U is periodic.
    [Show full text]
  • Torsion Theories Over Commutative Rings
    View metadata, citation and similar papers at core.ac.uk brought to you by CORE provided by Elsevier - Publisher Connector JOURNAL OF ALGEBRA 101, 136150 (1986) Torsion Theories over Commutative Rings WILLY BRANDALAND EROL BARBUT Department of Mathematics and Applied Statistics, University of Idaho, Moscow, Idaho 83843 Communicated by I. N. Herstein Received October 20, 1982 The definition of an h-local integral domain is generalized to commutative rings. This new definition is in terms of Gabriel topologies; i.e., in terms of hereditary tor- sion theories over commutative rings. Such commutative rings are characterized by the decomposition of torsion modules relative to a given torsion theory. Min-local commutative rings constitute a special case. If R is a min-local commutative ring, then an injective cogenerator of a nonminimal Gabriel topology of R is the direct product of the injective cogenerators of all the locahzations of the given Gabriel topology. The ring of quotients of a min-local commutative ring with respect to a nonminimal Gabriel topology can be canonically embedded into the product of rings of quotients of localizations. All the Gabriel topologies of commutative valuation rings and their rings of quotients are described. If R is a min-local Priifer commutative ring, then the ring of quotients of R with respect to any nonminimal Gabriel topology of R can be canonically embedded into a product of rings of quotients of locahzations, each of which is a valuation ring or a topological com- pletion of a valuation ring. ‘c 1986 Academic Press, Inc R will always denote a commutative ring with identity, and all rings con- sidered, except some endomorphism rings, will be commutative rings.
    [Show full text]
  • Homology Groups of Homeomorphic Topological Spaces
    An Introduction to Homology Prerna Nadathur August 16, 2007 Abstract This paper explores the basic ideas of simplicial structures that lead to simplicial homology theory, and introduces singular homology in order to demonstrate the equivalence of homology groups of homeomorphic topological spaces. It concludes with a proof of the equivalence of simplicial and singular homology groups. Contents 1 Simplices and Simplicial Complexes 1 2 Homology Groups 2 3 Singular Homology 8 4 Chain Complexes, Exact Sequences, and Relative Homology Groups 9 ∆ 5 The Equivalence of H n and Hn 13 1 Simplices and Simplicial Complexes Definition 1.1. The n-simplex, ∆n, is the simplest geometric figure determined by a collection of n n + 1 points in Euclidean space R . Geometrically, it can be thought of as the complete graph on (n + 1) vertices, which is solid in n dimensions. Figure 1: Some simplices Extrapolating from Figure 1, we see that the 3-simplex is a tetrahedron. Note: The n-simplex is topologically equivalent to Dn, the n-ball. Definition 1.2. An n-face of a simplex is a subset of the set of vertices of the simplex with order n + 1. The faces of an n-simplex with dimension less than n are called its proper faces. 1 Two simplices are said to be properly situated if their intersection is either empty or a face of both simplices (i.e., a simplex itself). By \gluing" (identifying) simplices along entire faces, we get what are known as simplicial complexes. More formally: Definition 1.3. A simplicial complex K is a finite set of simplices satisfying the following condi- tions: 1 For all simplices A 2 K with α a face of A, we have α 2 K.
    [Show full text]