Math 371 Lecture #21 §6.1: Ideals and Congruence, Part II §6.2: Quotients and Homomorphisms, Part I

Total Page:16

File Type:pdf, Size:1020Kb

Math 371 Lecture #21 §6.1: Ideals and Congruence, Part II §6.2: Quotients and Homomorphisms, Part I Math 371 Lecture #21 x6.1: Ideals and Congruence, Part II x6.2: Quotients and Homomorphisms, Part I Why ideals? For a commutative ring R with identity, we have for a; b 2 R, that a j b if and only if (b) ⊆ (a), and u 2 R is a unit if and only if (u) = R. Having the established the notion of an ideal, we can now extend the notion of congruence to any ring with respect to any ideal. Definition. Let I be an ideal in a ring R. For a; b 2 R, we say a is congruent to b modulo I, written a ≡ b (mod I), provided that a − b 2 I. Examples. (a) For any integer k ≥ 1, we see that congruence modulo k in Z is the same thing as congruence modulo the principal ideal (k) = kZ of Z. (b) For a field F and any nonconstant polynomial p(x) in F [x], we see that congruence modulo p(x) in F [x] is the same thing as congruence modulo the principal ideal (p(x)) of F [x]. We know that the congruence in each of these examples is an equivalence relation, but is this the case for an arbitrary ideal of an arbitrary ring? Theorem 6.4. For any ideal I of a ring R, the relation of congruence modulo I is an equivalence relation. Proof. Reflexive: Since I is closed under subtraction, we have for any a 2 R that a − a = 0R 2 I; hence a ≡ a (mod I). Symmetric: For a; b 2 R, suppose that a ≡ b (mod I). Then a − b 2 I, so there is i 2 I such that a − b = i. Since 0R and i both belong to I, then −i = 0R − i 2 I; hence b − a = −i 2 I; so b ≡ a (mod I). Transitive: For a; b; c 2 R, suppose that a ≡ b (mod I) and b ≡ c (mod I). Then there are i; j 2 I such that a − b = i and b − c = j. Hence a − c = a − b + b − c = i + j 2 I, so that a ≡ c (mod I). Congruence modulo an ideal I respects the addition and multiplication of the ring. Theorem 6.5. For an ideal I in a ring R, if a ≡ b (mod I) and c ≡ d (mod I), then (1) a + c ≡ b + d (mod I), and (2) ac ≡ bd (mod I). Proof. (1) We have i; j 2 I such that a − b = i and c − d = j. Then (a + c) − (b + d) = (a − b) + (c − d) = i + j 2 I, so that a + c ≡ b + d (mod I). For the product we have ac − bd = ac − bc + bc − bd = (a − b)c + b(c − d) = ic + bj: Since the ideal I absorbs on the left and on the right, we have that ic and bj both belong to I. Hence ac − bd = ic + bj 2 I. For an ideal I in a ring R, the congruence class of a 2 R modulo I is fb 2 R : b ≡ a (mod I)g = fb 2 R : b − a 2 Ig = fb 2 R : b − a = i for some i 2 Ig = fb 2 R : b = a + i with i 2 Ig = fa + i : i 2 Ig: We use the notation a + I to denote the congruence class of a modulo I. Definition. A (left) coset of an ideal I in a ring R is a subset of the form a + I. Note: if R is commutative, then the right coset I + a = fi + a : i 2 Ig is the same as the left coset a + I, in which case we call a + I a coset of I. Example. We can think of the + in the coset a + I as an addition: for integers a; n with n ≥ 1, we have [a] = a + (n) = a + nZ = fa; a ± n; a ± 2n; : : : g. Theorem 6.6. For an ideal I in a ring R, and a; c 2 R, we have a ≡ c (mod I) if and only if a + I = c + I. Proof. Suppose a ≡ c (mod I). If b 2 a + I, then b ≡ a (mod I), so by transitivity b ≡ c (mod I), hence b 2 c + I. If b 2 c + I, then b ≡ c (mod I), so by symmetry and transitivity b ≡ a (mod I), hence b 2 a + I. Thus a + I = c + I. Now suppose that a + I = c + I. Then by reflexivity, a 2 a + I = c + I, so that a ≡ c (mod I). Corollary 6.7. For an ideal I in a ring R, two cosets of I are either disjoint or identical. Notation. For an ideal I in a ring R, the set of all cosets of I is denoted by R=I. We attempt to turn R=I into a ring by defining an addition and multiplication of cosets: for a; c 2 R, (a + I) + (c + I) = (a + c) + I; (a + I)(c + I) = ac + I: Since we are defining these operations via representatives of the cosets, there is a question of whether they are well-defined. Theorem 6.8. For an ideal I in a ring R, if a + I = b + I and c + I = d + I, then (a + c) + I = (b + d) + I and ac + I = bd + I: Proof. Suppose a + I = b + I and c + I = d + I. By Theorem 6.6, we know that a ≡ b (mod I), and hence that a − b 2 I. Similarly, we have that c − d 2 I. Since the ideal I is closed under addition, we obtain (a+c)−(b+d) = (a−b)+(c−d) 2 I, and so a + c ≡ b + d (mod I). By Theorem 6.6, we conclude that (a + c) + I = (b + d) + I. For the product we have ac − bd = ac − bc + bc − bd = (a − b)c + b(c − d): Since the ideal I absorbs multiplication on the left and on the right, we have ac − bd 2 I. Hence by Theorem 6.6, we conclude that ac + I = bd + I. Theorem 6.9. For an ideal I in a ring R, we have (1) R=I is a ring, (2) R=I is commutative if R is commutative, and (3) R=I has an identity if R has an identity. Proof. (1) One verifies that the eight axioms of a ring hold for the set of (left) cosets R=I with the addition and multiplication of cosets as defined. (2) If R is commutative, then for a; c 2 R, we have (a + I)(c + I) = ac + I = ca + I = (c + I)(a + I): (3) If R has a multiplicative identity 1R, then the coset 1R + I satisfies, for any a 2 R, (a + I)(1R + I) = a1R + I = a + I; (1R + I)(a + I) = 1Ra + I = a + I: Thus 1R + I is the multiplicative identity of R=I. Definition. For a ring R and an ideal I in R, the ring R=I is called the quotient ring (or factor ring) of R by I. One may say that we factor R by I to obtain the quotient ring R=I. a b Example. Let R = : a; b; c; d 2 , and I the subset of R consisting of matri- c d Z ces with even entries. The subset I is nonempty (it contains the zero matrix). The nonempty subset I is an ideal because it is closed under subtraction and it absorbs multiplication by elements of R on the both the left and the right. What is R=I? We investigate a coset for a randomly chosen matrix in R: 7 8 1 0 6 8 1 0 + I = + + I = + I: 5 −3 1 1 4 −4 1 1 This demonstrates that representatives of the distinct cosets of I in R are given by matrices with 0 or 1 as entries. So the quotient ring R=I consists of 24 = 16 elements, the distinct cosets of I in R..
Recommended publications
  • A Category-Theoretic Approach to Representation and Analysis of Inconsistency in Graph-Based Viewpoints
    A Category-Theoretic Approach to Representation and Analysis of Inconsistency in Graph-Based Viewpoints by Mehrdad Sabetzadeh A thesis submitted in conformity with the requirements for the degree of Master of Science Graduate Department of Computer Science University of Toronto Copyright c 2003 by Mehrdad Sabetzadeh Abstract A Category-Theoretic Approach to Representation and Analysis of Inconsistency in Graph-Based Viewpoints Mehrdad Sabetzadeh Master of Science Graduate Department of Computer Science University of Toronto 2003 Eliciting the requirements for a proposed system typically involves different stakeholders with different expertise, responsibilities, and perspectives. This may result in inconsis- tencies between the descriptions provided by stakeholders. Viewpoints-based approaches have been proposed as a way to manage incomplete and inconsistent models gathered from multiple sources. In this thesis, we propose a category-theoretic framework for the analysis of fuzzy viewpoints. Informally, a fuzzy viewpoint is a graph in which the elements of a lattice are used to specify the amount of knowledge available about the details of nodes and edges. By defining an appropriate notion of morphism between fuzzy viewpoints, we construct categories of fuzzy viewpoints and prove that these categories are (finitely) cocomplete. We then show how colimits can be employed to merge the viewpoints and detect the inconsistencies that arise independent of any particular choice of viewpoint semantics. Taking advantage of the same category-theoretic techniques used in defining fuzzy viewpoints, we will also introduce a more general graph-based formalism that may find applications in other contexts. ii To my mother and father with love and gratitude. Acknowledgements First of all, I wish to thank my supervisor Steve Easterbrook for his guidance, support, and patience.
    [Show full text]
  • Modular Arithmetic
    CS 70 Discrete Mathematics and Probability Theory Fall 2009 Satish Rao, David Tse Note 5 Modular Arithmetic One way to think of modular arithmetic is that it limits numbers to a predefined range f0;1;:::;N ¡ 1g, and wraps around whenever you try to leave this range — like the hand of a clock (where N = 12) or the days of the week (where N = 7). Example: Calculating the day of the week. Suppose that you have mapped the sequence of days of the week (Sunday, Monday, Tuesday, Wednesday, Thursday, Friday, Saturday) to the sequence of numbers (0;1;2;3;4;5;6) so that Sunday is 0, Monday is 1, etc. Suppose that today is Thursday (=4), and you want to calculate what day of the week will be 10 days from now. Intuitively, the answer is the remainder of 4 + 10 = 14 when divided by 7, that is, 0 —Sunday. In fact, it makes little sense to add a number like 10 in this context, you should probably find its remainder modulo 7, namely 3, and then add this to 4, to find 7, which is 0. What if we want to continue this in 10 day jumps? After 5 such jumps, we would have day 4 + 3 ¢ 5 = 19; which gives 5 modulo 7 (Friday). This example shows that in certain circumstances it makes sense to do arithmetic within the confines of a particular number (7 in this example), that is, to do arithmetic by always finding the remainder of each number modulo 7, say, and repeating this for the results, and so on.
    [Show full text]
  • Modulo of a Negative Number1 1 Modular Arithmetic
    Algologic Technical Report # 01/2017 Modulo of a negative number1 S. Parthasarathy [email protected] Ref.: negamodulo.tex Ver. code: 20170109e Abstract This short article discusses an enigmatic question in elementary num- ber theory { that of finding the modulo of a negative number. Is this feasible to compute ? If so, how do we compute this value ? 1 Modular arithmetic In mathematics, modular arithmetic is a system of arithmetic for integers, where numbers "wrap around" upon reaching a certain value - the modulus (plural moduli). The modern approach to modular arithmetic was developed by Carl Friedrich Gauss in his book Disquisitiones Arithmeticae, published in 1801. Modular arithmetic is a fascinating aspect of mathematics. The \modulo" operation is often called as the fifth arithmetic operation, and comes after + − ∗ and = . It has use in many practical situations, particu- larly in number theory and cryptology [1] [2]. If x and n are integers, n < x and n =6 0 , the operation \ x mod n " is the remainder we get when n divides x. In other words, (x mod n) = r if there exists some integer q such that x = q ∗ n + r 1This is a LATEX document. You can get the LATEX source of this document from [email protected]. Please mention the Reference Code, and Version code, given at the top of this document 1 When r = 0 , n is a factor of x. When n is a factor of x, we say n divides x evenly, and denote it by n j x . The modulo operation can be combined with other arithmetic operators.
    [Show full text]
  • The Development and Understanding of the Concept of Quotient Group
    HISTORIA MATHEMATICA 20 (1993), 68-88 The Development and Understanding of the Concept of Quotient Group JULIA NICHOLSON Merton College, Oxford University, Oxford OXI 4JD, United Kingdom This paper describes the way in which the concept of quotient group was discovered and developed during the 19th century, and examines possible reasons for this development. The contributions of seven mathematicians in particular are discussed: Galois, Betti, Jordan, Dedekind, Dyck, Frobenius, and H61der. The important link between the development of this concept and the abstraction of group theory is considered. © 1993 AcademicPress. Inc. Cet article decrit la d6couverte et le d6veloppement du concept du groupe quotient au cours du 19i~me si~cle, et examine les raisons possibles de ce d6veloppement. Les contributions de sept math6maticiens seront discut6es en particulier: Galois, Betti, Jordan, Dedekind, Dyck, Frobenius, et H61der. Le lien important entre le d6veloppement de ce concept et l'abstraction de la th6orie des groupes sera consider6e. © 1993 AcademicPress, Inc. Diese Arbeit stellt die Entdeckung beziehungsweise die Entwicklung des Begriffs der Faktorgruppe wfihrend des 19ten Jahrhunderts dar und untersucht die m/)glichen Griinde for diese Entwicklung. Die Beitrfi.ge von sieben Mathematikern werden vor allem diskutiert: Galois, Betti, Jordan, Dedekind, Dyck, Frobenius und H61der. Die wichtige Verbindung zwischen der Entwicklung dieses Begriffs und der Abstraktion der Gruppentheorie wird betrachtet. © 1993 AcademicPress. Inc. AMS subject classifications: 01A55, 20-03 KEY WORDS: quotient group, group theory, Galois theory I. INTRODUCTION Nowadays one can hardly conceive any more of a group theory without factor groups... B. L. van der Waerden, from his obituary of Otto H61der [1939] Although the concept of quotient group is now considered to be fundamental to the study of groups, it is a concept which was unknown to early group theorists.
    [Show full text]
  • Grade 7/8 Math Circles Modular Arithmetic the Modulus Operator
    Faculty of Mathematics Centre for Education in Waterloo, Ontario N2L 3G1 Mathematics and Computing Grade 7/8 Math Circles April 3, 2014 Modular Arithmetic The Modulus Operator The modulo operator has symbol \mod", is written as A mod N, and is read \A modulo N" or "A mod N". • A is our - what is being divided (just like in division) • N is our (the plural is ) When we divide two numbers A and N, we get a quotient and a remainder. • When we ask for A ÷ N, we look for the quotient of the division. • When we ask for A mod N, we are looking for the remainder of the division. • The remainder A mod N is always an integer between 0 and N − 1 (inclusive) Examples 1. 0 mod 4 = 0 since 0 ÷ 4 = 0 remainder 0 2. 1 mod 4 = 1 since 1 ÷ 4 = 0 remainder 1 3. 2 mod 4 = 2 since 2 ÷ 4 = 0 remainder 2 4. 3 mod 4 = 3 since 3 ÷ 4 = 0 remainder 3 5. 4 mod 4 = since 4 ÷ 4 = 1 remainder 6. 5 mod 4 = since 5 ÷ 4 = 1 remainder 7. 6 mod 4 = since 6 ÷ 4 = 1 remainder 8. 15 mod 4 = since 15 ÷ 4 = 3 remainder 9.* −15 mod 4 = since −15 ÷ 4 = −4 remainder Using a Calculator For really large numbers and/or moduli, sometimes we can use calculators to quickly find A mod N. This method only works if A ≥ 0. Example: Find 373 mod 6 1. Divide 373 by 6 ! 2. Round the number you got above down to the nearest integer ! 3.
    [Show full text]
  • RING THEORY 1. Ring Theory a Ring Is a Set a with Two Binary Operations
    CHAPTER IV RING THEORY 1. Ring Theory A ring is a set A with two binary operations satisfying the rules given below. Usually one binary operation is denoted `+' and called \addition," and the other is denoted by juxtaposition and is called \multiplication." The rules required of these operations are: 1) A is an abelian group under the operation + (identity denoted 0 and inverse of x denoted x); 2) A is a monoid under the operation of multiplication (i.e., multiplication is associative and there− is a two-sided identity usually denoted 1); 3) the distributive laws (x + y)z = xy + xz x(y + z)=xy + xz hold for all x, y,andz A. Sometimes one does∈ not require that a ring have a multiplicative identity. The word ring may also be used for a system satisfying just conditions (1) and (3) (i.e., where the associative law for multiplication may fail and for which there is no multiplicative identity.) Lie rings are examples of non-associative rings without identities. Almost all interesting associative rings do have identities. If 1 = 0, then the ring consists of one element 0; otherwise 1 = 0. In many theorems, it is necessary to specify that rings under consideration are not trivial, i.e. that 1 6= 0, but often that hypothesis will not be stated explicitly. 6 If the multiplicative operation is commutative, we call the ring commutative. Commutative Algebra is the study of commutative rings and related structures. It is closely related to algebraic number theory and algebraic geometry. If A is a ring, an element x A is called a unit if it has a two-sided inverse y, i.e.
    [Show full text]
  • Lecture 7: Unit Group Structure
    MATH 361: NUMBER THEORY | SEVENTH LECTURE 1. The Unit Group of Z=nZ Consider a nonunit positive integer, Y n = pep > 1: The Sun Ze Theorem gives a ring isomorphism, Y e Z=nZ =∼ Z=p p Z: The right side is the cartesian product of the rings Z=pep Z, meaning that addition and multiplication are carried out componentwise. It follows that the corresponding unit group is × Y e × (Z=nZ) =∼ (Z=p p Z) : Thus to study the unit group (Z=nZ)× it suffices to consider (Z=peZ)× where p is prime and e > 0. Recall that in general, × j(Z=nZ) j = φ(n); so that for prime powers, e × e e−1 j(Z=p Z) j = φ(p ) = p (p − 1); and especially for primes, × j(Z=pZ) j = p − 1: Here are some examples of unit groups modulo prime powers, most but not quite all cyclic. × 0 (Z=2Z) = (f1g; ·) = (f2 g; ·) =∼ (f0g; +) = Z=Z; × 0 1 (Z=3Z) = (f1; 2g; ·) = (f2 ; 2 g; ·) =∼ (f0; 1g; +) = Z=2Z; × 0 1 (Z=4Z) = (f1; 3g; ·) = (f3 ; 3 g; ·) =∼ (f0; 1g; +) = Z=2Z; × 0 1 2 3 (Z=5Z) = (f1; 2; 3; 4g; ·) = (f2 ; 2 ; 2 ; 2 g; ·) =∼ (f0; 1; 2; 3g; +) = Z=4Z; × 0 1 2 3 4 5 (Z=7Z) = (f1; 2; 3; 4; 5; 6g; ·) = (f3 ; 3 ; 3 ; 3 ; 3 ; 3 g; ·) =∼ (f0; 1; 2; 3; 4; 5g; +) = Z=6Z; × 0 0 1 0 0 1 1 1 (Z=8Z) = (f1; 3; 5; 7g; ·) = (f3 5 ; 3 5 ; 3 5 ; 3 5 g; ·) =∼ (f0; 1g × f0; 1g; +) = Z=2Z × Z=2Z; × 0 1 2 3 4 5 (Z=9Z) = (f1; 2; 4; 5; 7; 8g; ·) = (f2 ; 2 ; 2 ; 2 ; 2 ; 2 g; ·) =∼ (f0; 1; 2; 3; 4; 5g; +) = Z=6Z: 1 2 MATH 361: NUMBER THEORY | SEVENTH LECTURE 2.
    [Show full text]
  • Computing Mod Without Mod
    Computing Mod Without Mod Mark A. Will and Ryan K. L. Ko Cyber Security Lab The University of Waikato, New Zealand fwillm,[email protected] Abstract. Encryption algorithms are designed to be difficult to break without knowledge of the secrets or keys. To achieve this, the algorithms require the keys to be large, with some algorithms having a recommend size of 2048-bits or more. However most modern processors only support computation on 64-bits at a time. Therefore standard operations with large numbers are more complicated to implement. One operation that is particularly challenging to implement efficiently is modular reduction. In this paper we propose a highly-efficient algorithm for solving large modulo operations; it has several advantages over current approaches as it supports the use of a variable sized lookup table, has good spatial and temporal locality allowing data to be streamed, and only requires basic processor instructions. Our proposed algorithm is theoretically compared to widely used modular algorithms, before practically compared against the state-of-the-art GNU Multiple Precision (GMP) large number li- brary. 1 Introduction Modular reduction, also known as the modulo or mod operation, is a value within Y, such that it is the remainder after Euclidean division of X by Y. This operation is heavily used in encryption algorithms [6][8][14], since it can \hide" values within large prime numbers, often called keys. Encryption algorithms also make use of the modulo operation because it involves more computation when compared to other operations like add. Meaning that computing the modulo operation is not as simple as just adding bits.
    [Show full text]
  • Version 0.2.1 Copyright C 2003-2009 Richard P
    Modern Computer Arithmetic Richard P. Brent and Paul Zimmermann Version 0.2.1 Copyright c 2003-2009 Richard P. Brent and Paul Zimmermann This electronic version is distributed under the terms and conditions of the Creative Commons license “Attribution-Noncommercial-No Derivative Works 3.0”. You are free to copy, distribute and transmit this book under the following conditions: Attribution. You must attribute the work in the manner specified • by the author or licensor (but not in any way that suggests that they endorse you or your use of the work). Noncommercial. You may not use this work for commercial purposes. • No Derivative Works. You may not alter, transform, or build upon • this work. For any reuse or distribution, you must make clear to others the license terms of this work. The best way to do this is with a link to the web page below. Any of the above conditions can be waived if you get permission from the copyright holder. Nothing in this license impairs or restricts the author’s moral rights. For more information about the license, visit http://creativecommons.org/licenses/by-nc-nd/3.0/ Preface This is a book about algorithms for performing arithmetic, and their imple- mentation on modern computers. We are concerned with software more than hardware — we do not cover computer architecture or the design of computer hardware since good books are already available on these topics. Instead we focus on algorithms for efficiently performing arithmetic operations such as addition, multiplication and division, and their connections to topics such as modular arithmetic, greatest common divisors, the Fast Fourier Transform (FFT), and the computation of special functions.
    [Show full text]
  • Modular Arithmetics Before C.F. Gauss. Maarten Bullynck
    Modular Arithmetics before C.F. Gauss. Maarten Bullynck To cite this version: Maarten Bullynck. Modular Arithmetics before C.F. Gauss.: Systematizations and discussions on remainder problems in 18th-century Germany.. Historia Mathematica, Elsevier, 2009, 36 (1), pp.48- 72. halshs-00663292 HAL Id: halshs-00663292 https://halshs.archives-ouvertes.fr/halshs-00663292 Submitted on 26 Jan 2012 HAL is a multi-disciplinary open access L’archive ouverte pluridisciplinaire HAL, est archive for the deposit and dissemination of sci- destinée au dépôt et à la diffusion de documents entific research documents, whether they are pub- scientifiques de niveau recherche, publiés ou non, lished or not. The documents may come from émanant des établissements d’enseignement et de teaching and research institutions in France or recherche français ou étrangers, des laboratoires abroad, or from public or private research centers. publics ou privés. Modular Arithmetic before C.F. Gauss. Systematisations and discussions on remainder problems in 18th century Germany Maarten Bullynck IZWT, Bergische Universit¨at Wuppertal, Gaußstraße 20, 42119 Wuppertal, Germany Forschungsstipendiat der Alexander-von-Humboldt-Stiftung Email: [email protected] Abstract Remainder problems have a long tradition and were widely disseminated in books on calculation, algebra and recreational mathematics from the 13th century until the 18th century. Many singular solution methods for particular cases were known, but Bachet de M´eziriac was the first to see how these methods connected with the Euclidean algorithm and with Diophantine analysis (1624). His general solution method contributed to the theory of equations in France, but went largely unno- ticed elsewhere. Later Euler independently rediscovered similar methods, while von Clausberg generalised and systematised methods which used the greatest common divisor procedure.
    [Show full text]
  • Lecture Notes #5: Modular Arithmetic
    EECS 70 Discrete Mathematics and Probability Theory Spring 2014 Anant Sahai Note 5 Modular Arithmetic In several settings, such as error-correcting codes and cryptography, we sometimes wish to work over a smaller range of numbers. Modular arithmetic is useful in these settings, since it limits numbers to a prede- fined range f0;1;:::;N − 1g, and wraps around whenever you try to leave this range — like the hand of a clock (where N = 12) or the days of the week (where N = 7). Example: Calculating the time When you calculate the time, you automatically use modular arithmetic. For example, if you are asked what time it will be 13 hours from 1 pm, you say 2 am rather than 14. Let’s assume our clock displays 12 as 0. This is limiting numbers to a predefined range, f0;1;2;:::;11g. Whenever you add two numbers in this setting, you divide by 12 and provide the remainder as the answer. If we wanted to know what the time would be 24 hours from 2 pm, the answer is easy. It would be 2 pm. This is true not just for 24 hours, but for any multiple of 12 hours. What about 25 hours from 2 pm? Since the time 24 hours from 2 pm is still 2 pm, 25 hours later it would be 3 pm. Another way to say this is that we add 1 hour, which is the remainder when we divide 25 by 12. This example shows that under certain circumstances it makes sense to do arithmetic within the confines of a particular number (12 in this example).
    [Show full text]
  • CHAPTER 2 RING FUNDAMENTALS 2.1 Basic Definitions and Properties
    page 1 of Chapter 2 CHAPTER 2 RING FUNDAMENTALS 2.1 Basic Definitions and Properties 2.1.1 Definitions and Comments A ring R is an abelian group with a multiplication operation (a, b) → ab that is associative and satisfies the distributive laws: a(b+c)=ab+ac and (a + b)c = ab + ac for all a, b, c ∈ R. We will always assume that R has at least two elements,including a multiplicative identity 1 R satisfying a1R =1Ra = a for all a in R. The multiplicative identity is often written simply as 1,and the additive identity as 0. If a, b,and c are arbitrary elements of R,the following properties are derived quickly from the definition of a ring; we sketch the technique in each case. (1) a0=0a =0 [a0+a0=a(0+0)=a0; 0a +0a =(0+0)a =0a] (2) (−a)b = a(−b)=−(ab)[0=0b =(a+(−a))b = ab+(−a)b,so ( −a)b = −(ab); similarly, 0=a0=a(b +(−b)) = ab + a(−b), so a(−b)=−(ab)] (3) (−1)(−1) = 1 [take a =1,b= −1 in (2)] (4) (−a)(−b)=ab [replace b by −b in (2)] (5) a(b − c)=ab − ac [a(b +(−c)) = ab + a(−c)=ab +(−(ac)) = ab − ac] (6) (a − b)c = ac − bc [(a +(−b))c = ac +(−b)c)=ac − (bc)=ac − bc] (7) 1 = 0 [If 1 = 0 then for all a we have a = a1=a0 = 0,so R = {0},contradicting the assumption that R has at least two elements] (8) The multiplicative identity is unique [If 1 is another multiplicative identity then 1=11 =1] 2.1.2 Definitions and Comments If a and b are nonzero but ab = 0,we say that a and b are zero divisors;ifa ∈ R and for some b ∈ R we have ab = ba = 1,we say that a is a unit or that a is invertible.
    [Show full text]