How Do Elements Really Factor in Z[ √ −5]?

Total Page:16

File Type:pdf, Size:1020Kb

How Do Elements Really Factor in Z[ √ −5]? p How Do Elements Really Factor in Z[ −5]? Scott T. Chapman, Felix Gotti, and Marly Gotti p Abstract Most undergraduate level abstract algebra texts use Z[ −5] as an example of an integral domain which is not a unique factorization domain (or UFD) by ex- hibiting two distinct irreducible factorizations of a nonzero element. But such a brief example, which requires merely an understanding of basic norms, only scratches the surface of how elements actually factor in this ring of algebraicp integers. We offer here an interactive framework which shows that while Z[ −5] is not a UFD, it does satisfy a slightly weaker factorization condition, known as half-factoriality. The arguments involved revolve around the Fundamental Theorem of Ideal Theory in algebraic number fields. Dedicated to David F. Anderson on the occasion of his retirement. 1 Introduction Consider the integral domain p p Z[ −5] = fa + b −5 j a;b 2 Zg: Your undergraduate abstract algebra text probably used it as the base example of an integral domain that is not a unique factorizationp domain (or UFD). The Fundamen- tal Theorem of Arithmetic fails in Z[ −5] as this domain contains elements with Scott T. Chapman Sam Houston State University, Huntsville, TX 77341, e-mail: [email protected] Felix Gotti UC Berkeley, Berkeley, CA 94720 e-mail: [email protected] Marly Gotti University of Florida, Gainesville, FL 32611 e-mail: [email protected] 2010 AMS Mathematics Subject Classification: Primary 11R11, 13F15, 20M13. 1 2 Scott T. Chapman, Felix Gotti, and Marly Gotti multiple factorizations into irreducibles; for example, p p 6 = 2 · 3 = (1 − −5)(1 + −5) (1) p p even thoughp 2;3;1 − −5, and 1 + −5 are pairwisep non-associate irreducible el- ements in Z[ −5]. To argue this, the norm on Z[ −5], i.e., p N(a + b −5) = a2 + 5b2; (2) plays an important role, as it is a multiplicative function satisfying the following properties: • N(a) = 0 if and only if a = 0; p • N(ab) = N(a)N(b) for all a;b 2 [ −5]; Z p • a is a unit if and only if N(a) = 1 (i.e., ±1 are the only units of Z[ −5]); • if N(a) is prime, then a is irreducible. However, introductory abstract algebra books seldom dig deeper than what Equation (1) does. Thep goal of this paper is to use ideal theory to describe exactly how ele- mentsp in Z[ −5] factor into products of irreducibles. In doing so, we will show that Z[ −5] satisfies a nicep factorization property, which is known as half-factoriality. Thus, we say that Z[ −5] is a half-factorial domain (or HFD). Our journey will require nothing more than elementary algebra, but will give the reader a glimpse of how The Fundamentalp Theorem of Ideal Theory resolves the non-unique factoriza- tions of Z[ −5]. The notion that unique factorization in rings of integers could be recovered via ideals was important in the late 1800’s in attempts to prove Fermat’s Last Theorem (see [9, Chapter 11]). Our presentation is somewhat interactive, as many steps that follow from stan- dard techniques of basic algebra are left to the reader as exercises. The only back- ground we expect from the reader are introductory courses in linear algebra and abstract algebra. Assuming such prerequisites, we have tried to present here a self- contained and friendlyp approach to the phenomenon of non-uniqueness of factoriza- tions occurring in Z[ −5]. More advanced and general arguments (which apply to any ring of integers) can be found in [8] and [9]. 2 Integral Bases and Discriminants Although in this paper we are primarily concerned withp the phenomenon of non- unique factorizations in the particular ring of integers Z[ −5], it is more enlight- ening from an algebraic perspective to introduce our needed concepts for arbitrary commutative rings with identity, rings of integers, or quadratic rings of integers, depending on the most appropriate context for each concept being introduced. In what follows, we shall proceed in this manner while trying, by all means, to keep the exposition as elementary as possible. p How Do Elements Really Factor in Z[ −5]? 3 An element a 2 C is said to be algebraic provided that it is a root of a nonzero polynomial with rational coefficients, while a is said to be an algebraic integer provided that it is a root of a monic polynomial with integer coefficients. It is not hard to argue that every subfield of C contains Q and is a Q-vector space. Definition 2.1. A subfield K of C is called an algebraic number field provided that it has finite dimension as a vector space over Q. The subset OK := fa 2 K j a is an algebraic integerg of K is called the ring of integers of K. The ring of integers of any algebraic number field is, indeed, a ring. The reader is invited to verify this observation. If a is a complex number, then Q(a) denotes the smallest subfield of C containing a. It is well known that a subfield K of C is an algebraic number field if and only if there exists an algebraic number a 2 C such that K = Q(a) (see, for example, [6, Theorem 2.17]). Among all algebraic number fields, we are primarily interested in those that are two-dimensional vector spaces over Q. Definition 2.2. An algebraic number field that is a two-dimensional vector space over Q is called a quadratic number field. If K is a quadratic number field, then OK is called a quadratic ring of integers. For a 2 C, let Z[a] denote the set of all polynomial expressions in a having integer coefficients.p Clearly, Z[a] is a subring of Q(a). It is also clear that, for d 2 Z, the field Q( d) has dimension at most two as a Q-vector space and, therefore, it is an algebraic number field. Moreover, if d 2= f0;1g and d is squarefree (i.e.,pd is not divisible by the square of any prime), then it immediately follows that Q( d) is a two-dimensional vector space over Q and, as a result, a quadratic number field. As we are mainly interested in the case when d = −5, we propose the following exercise. Exercise 2.3.pLet d 2 Z n f0;1g be a squarefree integer such that d ≡ 2;3p(mod 4). Prove that Z[ d] is the ring of integers of the quadratic number field Q( d). Remark. When d 2 nf0;1g is a squarefree integer satisfying d ≡ 1 (mod 4), it is Z p p 1+ d not hard to argue that the ring of integers of Q( d) is Z[ 2 ]. However, we will not be concerned with this case as our case of interest is d = −5. p For d asp specified in Exercise 2.3, the elementsp of Z[ d] can be written in the form a + b d for a;b 2 Z. The norm N on Z[ d] is defined by p N(a + b d) = a2 − db2 p (cf. Equation (2)). The norm N on Z[ d] also satisfies the four properties listed in the introduction. 4 Scott T. Chapman, Felix Gotti, and Marly Gotti Let us now take a look at the structure of an algebraic number field K with linear algebra in mind. For a 2 K consider the function ma : K ! K defined via multipli- cation by a, i.e., ma (x) = ax for all x 2 K. One can easily see that ma is a linear transformation of Q-vector spaces. Therefore, after fixing a basis for the Q-vector space K, we can represent ma by a matrix M. The trace of a, which is denoted by Tr(a), is defined to be the trace of the matrix M. It is worth noting that Tr(a) does not depend on the chosen basis for K. Also, notice that Tr(a) 2 Q. Furthermore, if a 2 OKp, then Tr(a) 2 Z (see [10, Lemma 4.1.1], or Exercise 2.5 for the case when K = Q( d)). Definition 2.4. Let K be an algebraic number field that has dimension n as a Q- vector space. The discriminant of a subset fw1;:::;wng of K, which is denoted by D[w1;:::;wn], is det T, where T is the n × n matrix Tr(wiw j) 1≤i; j≤n. With K as introduced above, if fw1;:::;wng is a subset of OpK, then it follows that D[w1;:::;wn] 2 Z (see Exercise 2.5 for the case when K = Q( d)). In addition, the discriminantp of any basis for the Q-vector space K is nonzero; we will prove this for K = Q( d) in Proposition 2.10. Exercise 2.5. Let d 2 Z n f0;1g be a squarefree integer such that d ≡ 2;3 (mod 4). p p 1. If a = a + a d 2 [ d], then Tr(a) = 2a . 1 2 Z p p 1 2. If, in addition, b = b1 + b2 d 2 Z[ d], then 2 a s(a) D[a;b] = det = 4d(a b − a b )2; b s(b) 1 2 2 1 p p where s(x + y d) = x − y d for all x;y 2 Z. Example 2.6. Let d 2= f0;1g be a squarefree integerp such that d ≡ 2;3 (modp 4).
Recommended publications
  • Formal Power Series - Wikipedia, the Free Encyclopedia
    Formal power series - Wikipedia, the free encyclopedia http://en.wikipedia.org/wiki/Formal_power_series Formal power series From Wikipedia, the free encyclopedia In mathematics, formal power series are a generalization of polynomials as formal objects, where the number of terms is allowed to be infinite; this implies giving up the possibility to substitute arbitrary values for indeterminates. This perspective contrasts with that of power series, whose variables designate numerical values, and which series therefore only have a definite value if convergence can be established. Formal power series are often used merely to represent the whole collection of their coefficients. In combinatorics, they provide representations of numerical sequences and of multisets, and for instance allow giving concise expressions for recursively defined sequences regardless of whether the recursion can be explicitly solved; this is known as the method of generating functions. Contents 1 Introduction 2 The ring of formal power series 2.1 Definition of the formal power series ring 2.1.1 Ring structure 2.1.2 Topological structure 2.1.3 Alternative topologies 2.2 Universal property 3 Operations on formal power series 3.1 Multiplying series 3.2 Power series raised to powers 3.3 Inverting series 3.4 Dividing series 3.5 Extracting coefficients 3.6 Composition of series 3.6.1 Example 3.7 Composition inverse 3.8 Formal differentiation of series 4 Properties 4.1 Algebraic properties of the formal power series ring 4.2 Topological properties of the formal power series
    [Show full text]
  • MAXIMAL and NON-MAXIMAL ORDERS 1. Introduction Let K Be A
    MAXIMAL AND NON-MAXIMAL ORDERS LUKE WOLCOTT Abstract. In this paper we compare and contrast various prop- erties of maximal and non-maximal orders in the ring of integers of a number field. 1. Introduction Let K be a number field, and let OK be its ring of integers. An order in K is a subring R ⊆ OK such that OK /R is finite as a quotient of abelian groups. From this definition, it’s clear that there is a unique maximal order, namely OK . There are many properties that all orders in K share, but there are also differences. The main difference between a non-maximal order and the maximal order is that OK is integrally closed, but every non-maximal order is not. The result is that many of the nice properties of Dedekind domains do not hold in arbitrary orders. First we describe properties common to both maximal and non-maximal orders. In doing so, some useful results and constructions given in class for OK are generalized to arbitrary orders. Then we de- scribe a few important differences between maximal and non-maximal orders, and give proofs or counterexamples. Several proofs covered in lecture or the text will not be reproduced here. Throughout, all rings are commutative with unity. 2. Common Properties Proposition 2.1. The ring of integers OK of a number field K is a Noetherian ring, a finitely generated Z-algebra, and a free abelian group of rank n = [K : Q]. Proof: In class we showed that OK is a free abelian group of rank n = [K : Q], and is a ring.
    [Show full text]
  • Super-Multiplicativity of Ideal Norms in Number Fields
    Super-multiplicativity of ideal norms in number fields Stefano Marseglia Abstract In this article we study inequalities of ideal norms. We prove that in a subring R of a number field every ideal can be generated by at most 3 elements if and only if the ideal norm satisfies N(IJ) ≥ N(I)N(J) for every pair of non-zero ideals I and J of every ring extension of R contained in the normalization R˜. 1 Introduction When we are studying a number ring R, that is a subring of a number field K, it can be useful to understand the size of its ideals compared to the whole ring. The main tool for this purpose is the norm map which associates to every non-zero ideal I of R its index as an abelian subgroup N(I) = [R : I]. If R is the maximal order, or ring of integers, of K then this map is multiplicative, that is, for every pair of non-zero ideals I,J ⊆ R we have N(I)N(J) = N(IJ). If the number ring is not the maximal order this equality may not hold for some pair of non-zero ideals. For example, arXiv:1810.02238v2 [math.NT] 1 Apr 2020 if we consider the quadratic order Z[2i] and the ideal I = (2, 2i), then we have that N(I)=2 and N(I2)=8, so we have the inequality N(I2) > N(I)2. Observe that if every maximal ideal p of a number ring R satisfies N(p2) ≤ N(p)2, then we can conclude that R is the maximal order of K (see Corollary 2.8).
    [Show full text]
  • 6. Localization
    52 Andreas Gathmann 6. Localization Localization is a very powerful technique in commutative algebra that often allows to reduce ques- tions on rings and modules to a union of smaller “local” problems. It can easily be motivated both from an algebraic and a geometric point of view, so let us start by explaining the idea behind it in these two settings. Remark 6.1 (Motivation for localization). (a) Algebraic motivation: Let R be a ring which is not a field, i. e. in which not all non-zero elements are units. The algebraic idea of localization is then to make more (or even all) non-zero elements invertible by introducing fractions, in the same way as one passes from the integers Z to the rational numbers Q. Let us have a more precise look at this particular example: in order to construct the rational numbers from the integers we start with R = Z, and let S = Znf0g be the subset of the elements of R that we would like to become invertible. On the set R×S we then consider the equivalence relation (a;s) ∼ (a0;s0) , as0 − a0s = 0 a and denote the equivalence class of a pair (a;s) by s . The set of these “fractions” is then obviously Q, and we can define addition and multiplication on it in the expected way by a a0 as0+a0s a a0 aa0 s + s0 := ss0 and s · s0 := ss0 . (b) Geometric motivation: Now let R = A(X) be the ring of polynomial functions on a variety X. In the same way as in (a) we can ask if it makes sense to consider fractions of such polynomials, i.
    [Show full text]
  • Formal Power Series Rings, Inverse Limits, and I-Adic Completions of Rings
    Formal power series rings, inverse limits, and I-adic completions of rings Formal semigroup rings and formal power series rings We next want to explore the notion of a (formal) power series ring in finitely many variables over a ring R, and show that it is Noetherian when R is. But we begin with a definition in much greater generality. Let S be a commutative semigroup (which will have identity 1S = 1) written multi- plicatively. The semigroup ring of S with coefficients in R may be thought of as the free R-module with basis S, with multiplication defined by the rule h k X X 0 0 X X 0 ( risi)( rjsj) = ( rirj)s: i=1 j=1 s2S 0 sisj =s We next want to construct a much larger ring in which infinite sums of multiples of elements of S are allowed. In order to insure that multiplication is well-defined, from now on we assume that S has the following additional property: (#) For all s 2 S, f(s1; s2) 2 S × S : s1s2 = sg is finite. Thus, each element of S has only finitely many factorizations as a product of two k1 kn elements. For example, we may take S to be the set of all monomials fx1 ··· xn : n (k1; : : : ; kn) 2 N g in n variables. For this chocie of S, the usual semigroup ring R[S] may be identified with the polynomial ring R[x1; : : : ; xn] in n indeterminates over R. We next construct a formal semigroup ring denoted R[[S]]: we may think of this ring formally as consisting of all functions from S to R, but we shall indicate elements of the P ring notationally as (possibly infinite) formal sums s2S rss, where the function corre- sponding to this formal sum maps s to rs for all s 2 S.
    [Show full text]
  • 9. Ideals and the Zariski Topology Definition 9.1. Let X ⊂ a N Be A
    9. Ideals and the Zariski Topology Definition 9.1. Let X ⊂ An be a subset. The ideal of X, denoted I(X), is simply the set of all polynomials which vanish on X. Let S ⊂ K[x1; x2; : : : ; xn]. Then the vanishing locus of S, denoted V (S), is n f p 2 A j f(x) = 0; 8f 2 S g: Lemma 9.2. Let X, Y ⊂ An and I, J ⊂ K[x] be any subsets. (1) X ⊂ V (I(X)). (2) I ⊂ I(V (I)). (3) If X ⊂ Y then I(Y ) ⊂ I(X). (4) If I ⊂ J then V (J) ⊂ V (I). (5) If X is a closed subset then V (I(X)) = X. Proof. Easy exercise. Note that a similar version of (5), with ideals replacing closed subsets, does not hold. For example take the ideal I ⊂ K[x], given as hx2i. Then V (I) = f0g, and the ideal of functions vanishing at the origin is hxi. It is natural then to ask what is the relation between I and I(V (I)). Clearly if f n 2 I then f 2 I(V (I)). Definitionp 9.3. Let I be an ideal in a ring R. The radical of I, denoted I, is f r 2 R j rn 2 I, some n g: It is not hard to check that the radical is an ideal. Theorem 9.4 (Hilbert's Nullstellensatz). Let K be an algebraically closed field and let pI be an ideal. Then I(V (I)) = I. p Proof. One inclusion is clear, I(V (I)) ⊃ I.
    [Show full text]
  • Commutative Algebra
    Commutative Algebra Andrew Kobin Spring 2016 / 2019 Contents Contents Contents 1 Preliminaries 1 1.1 Radicals . .1 1.2 Nakayama's Lemma and Consequences . .4 1.3 Localization . .5 1.4 Transcendence Degree . 10 2 Integral Dependence 14 2.1 Integral Extensions of Rings . 14 2.2 Integrality and Field Extensions . 18 2.3 Integrality, Ideals and Localization . 21 2.4 Normalization . 28 2.5 Valuation Rings . 32 2.6 Dimension and Transcendence Degree . 33 3 Noetherian and Artinian Rings 37 3.1 Ascending and Descending Chains . 37 3.2 Composition Series . 40 3.3 Noetherian Rings . 42 3.4 Primary Decomposition . 46 3.5 Artinian Rings . 53 3.6 Associated Primes . 56 4 Discrete Valuations and Dedekind Domains 60 4.1 Discrete Valuation Rings . 60 4.2 Dedekind Domains . 64 4.3 Fractional and Invertible Ideals . 65 4.4 The Class Group . 70 4.5 Dedekind Domains in Extensions . 72 5 Completion and Filtration 76 5.1 Topological Abelian Groups and Completion . 76 5.2 Inverse Limits . 78 5.3 Topological Rings and Module Filtrations . 82 5.4 Graded Rings and Modules . 84 6 Dimension Theory 89 6.1 Hilbert Functions . 89 6.2 Local Noetherian Rings . 94 6.3 Complete Local Rings . 98 7 Singularities 106 7.1 Derived Functors . 106 7.2 Regular Sequences and the Koszul Complex . 109 7.3 Projective Dimension . 114 i Contents Contents 7.4 Depth and Cohen-Macauley Rings . 118 7.5 Gorenstein Rings . 127 8 Algebraic Geometry 133 8.1 Affine Algebraic Varieties . 133 8.2 Morphisms of Affine Varieties . 142 8.3 Sheaves of Functions .
    [Show full text]
  • Are Maximal Ideals? Solution. No Principal
    Math 403A Assignment 6. Due March 1, 2013. 1.(8.1) Which principal ideals in Z[x] are maximal ideals? Solution. No principal ideal (f(x)) is maximal. If f(x) is an integer n = 1, then (n, x) is a bigger ideal that is not the whole ring. If f(x) has positive degree, then ± take any prime number p that does not divide the leading coefficient of f(x). (p, f(x)) is a bigger ideal that is not the whole ring, since Z[x]/(p, f(x)) = Z/pZ[x]/(f(x)) is not the zero ring. 2.(8.2) Determine the maximal ideals in the following rings. (a). R R. × Solution. Since (a, b) is a unit in this ring if a and b are nonzero, a maximal ideal will contain no pair (a, 0), (0, b) with a and b nonzero. Hence, the maximal ideals are R 0 and 0 R. ×{ } { } × (b). R[x]/(x2). Solution. Every ideal in R[x] is principal, i.e., generated by one element f(x). The ideals in R[x] that contain (x2) are (x2) and (x), since if x2 R[x]f(x), then x2 = f(x)g(x), which 2 ∈ 2 2 means that f(x) divides x . Up to scalar multiples, the only divisors of x are x and x . (c). R[x]/(x2 3x + 2). − Solution. The nonunit divisors of x2 3x +2=(x 1)(x 2) are, up to scalar multiples, x2 3x + 2 and x 1 and x 2. The− maximal ideals− are− (x 1) and (x 2), since the quotient− of R[x] by− either of those− ideals is the field R.
    [Show full text]
  • 6 Ideal Norms and the Dedekind-Kummer Theorem
    18.785 Number theory I Fall 2017 Lecture #6 09/25/2017 6 Ideal norms and the Dedekind-Kummer theorem In order to better understand how ideals split in Dedekind extensions we want to extend our definition of the norm map to ideals. Recall that for a ring extension B=A in which B is a free A-module of finite rank, we defined the norm map NB=A : B ! A as ×b NB=A(b) := det(B −! B); the determinant of the multiplication-by-b map with respect to an A-basis for B. If B is a free A-module we could define the norm of a B-ideal to be the A-ideal generated by the norms of its elements, but in the case we are most interested in (our \AKLB" setup) B is typically not a free A-module (even though it is finitely generated as an A-module). To get around this limitation, we introduce the notion of the module index, which we will use to define the norm of an ideal. In the special case where B is a free A-module, the norm of a B-ideal will be equal to the A-ideal generated by the norms of elements. 6.1 The module index Our strategy is to define the norm of a B-ideal as the intersection of the norms of its localizations at maximal ideals of A (note that B is an A-module, so we can view any ideal of B as an A-module). Recall that by Proposition 2.6 any A-module M in a K-vector space is equal to the intersection of its localizations at primes of A; this applies, in particular, to ideals (and fractional ideals) of A and B.
    [Show full text]
  • DISCRIMINANTS in TOWERS Let a Be a Dedekind Domain with Fraction
    DISCRIMINANTS IN TOWERS JOSEPH RABINOFF Let A be a Dedekind domain with fraction field F, let K=F be a finite separable ex- tension field, and let B be the integral closure of A in K. In this note, we will define the discriminant ideal B=A and the relative ideal norm NB=A(b). The goal is to prove the formula D [L:K] C=A = NB=A C=B B=A , D D ·D where C is the integral closure of B in a finite separable extension field L=K. See Theo- rem 6.1. The main tool we will use is localizations, and in some sense the main purpose of this note is to demonstrate the utility of localizations in algebraic number theory via the discriminants in towers formula. Our treatment is self-contained in that it only uses results from Samuel’s Algebraic Theory of Numbers, cited as [Samuel]. Remark. All finite extensions of a perfect field are separable, so one can replace “Let K=F be a separable extension” by “suppose F is perfect” here and throughout. Note that Samuel generally assumes the base has characteristic zero when it suffices to assume that an extension is separable. We will use the more general fact, while quoting [Samuel] for the proof. 1. Notation and review. Here we fix some notations and recall some facts proved in [Samuel]. Let K=F be a finite field extension of degree n, and let x1,..., xn K. We define 2 n D x1,..., xn det TrK=F xi x j .
    [Show full text]
  • Super-Multiplicativity of Ideal Norms in Number Fields
    Universiteit Leiden Mathematisch Instituut Master Thesis Super-multiplicativity of ideal norms in number fields Academic year 2012-2013 Candidate: Advisor: Stefano Marseglia Prof. Bart de Smit Contents 1 Preliminaries 1 2 Quadratic and quartic case 10 3 Main theorem: first implication 14 4 Main theorem: second implication 19 Introduction When we are studying a number ring R, that is a subring of a number field K, it can be useful to understand \how big" its ideals are compared to the whole ring. The main tool for this purpose is the norm map: N : I(R) −! Z>0 I 7−! #R=I where I(R) is the set of non-zero ideals of R. It is well known that this map is multiplicative if R is the maximal order, or ring of integers of the number field. This means that for every pair of ideals I;J ⊆ R we have: N(I)N(J) = N(IJ): For an arbitrary number ring in general this equality fails. For example, if we consider the quadratic order Z[2i] and the ideal I = (2; 2i), then we have that N(I) = 2 and N(I2) = 8, so we have the inequality N(I2) > N(I)2. In the first chapter we will recall some theorems and useful techniques of commutative algebra and algebraic number theory that will help us to understand the behaviour of the ideal norm. In chapter 2 we will see that the inequality of the previous example is not a coincidence. More precisely we will prove that in any quadratic order, for every pair of ideals I;J we have that N(IJ) ≥ N(I)N(J).
    [Show full text]
  • NOTES in COMMUTATIVE ALGEBRA: PART 1 1. Results/Definitions Of
    NOTES IN COMMUTATIVE ALGEBRA: PART 1 KELLER VANDEBOGERT 1. Results/Definitions of Ring Theory It is in this section that a collection of standard results and definitions in commutative ring theory will be presented. For the rest of this paper, any ring R will be assumed commutative with identity. We shall also use "=" and "∼=" (isomorphism) interchangeably, where the context should make the meaning clear. 1.1. The Basics. Definition 1.1. A maximal ideal is any proper ideal that is not con- tained in any strictly larger proper ideal. The set of maximal ideals of a ring R is denoted m-Spec(R). Definition 1.2. A prime ideal p is such that for any a, b 2 R, ab 2 p implies that a or b 2 p. The set of prime ideals of R is denoted Spec(R). p Definition 1.3. The radical of an ideal I, denoted I, is the set of a 2 R such that an 2 I for some positive integer n. Definition 1.4. A primary ideal p is an ideal such that if ab 2 p and a2 = p, then bn 2 p for some positive integer n. In particular, any maximal ideal is prime, and the radical of a pri- mary ideal is prime. Date: September 3, 2017. 1 2 KELLER VANDEBOGERT Definition 1.5. The notation (R; m; k) shall denote the local ring R which has unique maximal ideal m and residue field k := R=m. Example 1.6. Consider the set of smooth functions on a manifold M.
    [Show full text]