Riemann Zeta Function

Total Page:16

File Type:pdf, Size:1020Kb

Riemann Zeta Function MATH 324 Summer 2006 Elementary Number Theory Notes on the Riemann Zeta Function Department of Mathematical and Statistical Sciences University of Alberta Riemann Zeta Function In this note we give an introduction to the Riemann zeta function, which connects the ideas of real analysis with the arithmetic of the integers. Define a function of a real variable s as follows, 1 1 ζ(s) = : ns n=1 X This function is called the Riemann zeta function, after the German mathematician Bernhard Riemann who systematically studied the deeper properties of this function beginning in 1859: The infinite series for ζ(s) was actually first introduced by Euler nearly 100 years before Riemann's work. Note that from the integral test, the series for ζ(s) converges for s > 1; and therefore the function ζ(s) is defined for all real numbers s > 1: Also, since 1 n+1 dx ; ns ≥ xs Zn for all n 1; we have ≥ 1 1 1 dx 1 = ns ≥ xs s 1 n=1 1 X Z − for all s > 1: Now we let s 1+; and we see that ! lim ζ(s) = : s 1+ 1 ! Infinite Products In order to discuss the connection of ζ(s) with the primes, we need the concept of an infinite product, which is very similar to the notion of an infinite sum. Definition. Given a sequence of real numbers an n 1; let f g ≥ p1 = a1; p = a a ; 2 1 · 2 p = a a a ; 3 1 · 2 · 3 . n p = a a a = a : n 1 · 2 · · · n k kY=1 th The ordered pair of sequences an ; pn is called an infinite product. The real number pn is called the n ff g fth gg partial product and an is called the n factor of the product, and the following symbols are used to denote the infinite product defined above: 1 a a a or a : 1 · 2 · · · n · · · n n=1 Y Using the the analogy with infinite series, where we say that the series converges if and only if the sequence of partial sums converges, it is tempting to say that the infinite product converges if and only if the sequence of partial products converges. However, if we do this, then every product which has one factor equal to zero would converge, regardless of the behavior of the remaining factors. The definition below is more useful: n 1 th Definition. Given an infinite product an; let pn = ak be the n partial product. n=1 Y kY=1 (a) If infinitely many factors an are zero, then we say the product diverges to zero. (b) If no factor a is zero, then we say that the product converges if and only if there exists a p = 0; such n 6 1 that p p as n : In this case, p is called the value of the product, and we write p = a : n ! ! 1 n n=1 If p 0 as n ; then we say the product diverges to zero. Y n ! ! 1 1 (c) If there exists an integer N such that n > N implies that a = 0; then we say that a converges, n 6 n n=1 Y 1 provided that an converges as described in part (b). In this case, the value of the product is n=YN+1 1 a a a a : 1 · 2 · · · N · n n=YN+1 1 (d) an is said to be divergent if it does not converge as described in parts (b) and (c) above. n=1 Y Note: The value of a convergent infinite product can be zero, but this is the case if and only if a finite number of the factors are zero. Also, the convergence of an infinite product is not affected by inserting or removing a finite number of factors, zero or not. Now we give a criterion for the convergence of infinite products, completely analogous to the corresponding criterion for convergence of infinite series. Theorem (Cauchy Criteria for Convergence of an Infinite Product). 1 The infinite product an converges if and only if given any > 0; there exists an integer n0 such that n=1 Y m ak 1 < − k=n+1 Y whenever m > n n0: ≥ 1 Proof. Assume first that the product an converges, we may also assume that no an is zero (discarding n=1 a finite number of terms if necessary). LetY pn = a1 a2 an and p = lim pn; n · · · · !1 so that p = 0; and there exists an M > 0 such that p > M for all n 1: Now, since p satisfies the 6 j nj ≥ f ng Cauchy criteria for sequences, given an > 0; there exists an integer n0 such that p p < M j m − nj m whenever m > n n0: Dividing by pn we get ak 1 < whenever m > n n0: ≥ j j − ≥ k=n+1 Y Conversly, suppose that given any > 0; there exists an integer n0 such that m ak 1 < ( ) − ∗ k=n+1 Y whenever m > n n0: ≥ First note that if m > n0 this implies that am = 0; since (assuming that 0 < < 1), taking n = m 1; we have 6 − a 1 a 1 < , j mj − ≤ j m − j that is, 0 < 1 < a < 1 + − j mj for m n : ≥ 0 1 Now take = 2 ; and let q = a a a m n0+1 n0+2 · · · m for m > n0; then 1 < q < 3 ; 2 j mj 2 for all m > n : Therefore, if q converges, it cannot converge to zero. 0 f mg 1 To see that the sequence qn does actually converge, let 0 < < 2 be arbitrary, then there exists an n0 such that f g q m 1 < 2 q − 3 n whenever m > n n0; that is, ≥ q q < 2 q < 3 2 = j m − nj 3 j nj 2 · 3 whenever m > n n : Thus, the sequence q is a Cauchy sequence of real numbers, and hence converges, ≥ 0 f ng 1 and so the infinite product an converges. n=1 Y 1 Note: Taking n = m 1 in ( ); we see that if the infinite product a converges, then − ∗ n n=1 Y lim an = 1: n !1 Because of this, the factors of a product are often written as 1 + an; so that convergence of a product 1 (1 + an) implies that lim an = 0: n n=1 !1 Y 1 Theorem. If a > 0 for all n 1; then the infinite product (1 + a ) converges if and only if the infinite n ≥ n n=1 Y 1 series an converges. n=1 P Proof. Let the partial sums and partial products be given by s = a + a + + a and p = (1 + a )(1 + a ) (1 + a ) n 1 2 · · · n n 1 2 · · · n for n 1: Since a > 0 for all n 1; then both of the sequences s and p are monotone increasing, ≥ n ≥ f ng f ng and to prove the theorem, we only have to show that sn is bounded above if and only if pn is bounded above. f g f g First we note that s < p for all n 1: Now using the fact that 1 + x ex for all real numbers x; we have n n ≥ ≤ p = (1 + a )(1 + a ) (1 + a ) ea1 ea2 ean = esn n 1 2 · · · n ≤ · · · for all n 1: Therefore the sequence pn is bounded above if and only if the sequence sn is bounded above. ≥ f g f g Also, note that the sequence of partial products pn cannot converge to zero, since pn 1 for n 1: Finally, note that f g ≥ ≥ p + if and only if s + : n ! 1 n ! 1 1 Definition. The infinite product (1 + an) is said to converge absolutely if and only if the infinite product n=1 Y 1 (1 + a ) converges. j nj n=1 Y 1 Theorem. If the infinite product (1 + an) converges absolutely, then it converges. n=1 Y Proof. Use the Cauchy criterion together with the inequality (1 + a )(1 + a ) (1 + a ) 1 (1 + a )(1 + a ) (1 + a ) 1: j n+1 n+2 · · · m − j ≤ j n+1j j n+2j · · · j mj − 1 Note: For positive terms a > 0 for all n 1; we know that (1 + a ) converges if and only if 1 a n ≥ n n n=1 n=1 Y P 1 1 converges, so that (1 + an) converges absolutely if and only if an converges absolutely. However, if n=1 n=1 the terms are not pYositive, we have the following examples to show Pthe results need not be valid. ( 1)n 1 Exercise 1. Let a = − for n 1; show that (1 + a ) diverges, but 1 a converges. n pn ≥ n n n=1 n=1 Y P Exercise 2. Let 1 1 1 a2n 1 = − and a2n = + − pn pn n 1 1 for n = 1; 2; : : : ; show that (1 + an) converges, but an diverges. n=1 n=1 Y P We do have a theorem analogous to the positive term result: 1 Theorem. If a 0 for all n 1; then the infinite product (1 a ) converges if and only if the infinite n ≥ ≥ − n n=1 Y 1 series an converges. n=1 P 1 Proof. Note that convergence of 1 a implies the absolute convergence of (1 a ); and hence the n − n n=1 n=1 Y 1 P convergence of (1 a ): − n n=1 Y 1 Conversely, suppose that 1 a diverges, if a does not converge to zero, then (1 a ) also diverges.
Recommended publications
  • Ch. 15 Power Series, Taylor Series
    Ch. 15 Power Series, Taylor Series 서울대학교 조선해양공학과 서유택 2017.12 ※ 본 강의 자료는 이규열, 장범선, 노명일 교수님께서 만드신 자료를 바탕으로 일부 편집한 것입니다. Seoul National 1 Univ. 15.1 Sequences (수열), Series (급수), Convergence Tests (수렴판정) Sequences: Obtained by assigning to each positive integer n a number zn z . Term: zn z1, z 2, or z 1, z 2 , or briefly zn N . Real sequence (실수열): Sequence whose terms are real Convergence . Convergent sequence (수렴수열): Sequence that has a limit c limznn c or simply z c n . For every ε > 0, we can find N such that Convergent complex sequence |zn c | for all n N → all terms zn with n > N lie in the open disk of radius ε and center c. Divergent sequence (발산수열): Sequence that does not converge. Seoul National 2 Univ. 15.1 Sequences, Series, Convergence Tests Convergence . Convergent sequence: Sequence that has a limit c Ex. 1 Convergent and Divergent Sequences iin 11 Sequence i , , , , is convergent with limit 0. n 2 3 4 limznn c or simply z c n Sequence i n i , 1, i, 1, is divergent. n Sequence {zn} with zn = (1 + i ) is divergent. Seoul National 3 Univ. 15.1 Sequences, Series, Convergence Tests Theorem 1 Sequences of the Real and the Imaginary Parts . A sequence z1, z2, z3, … of complex numbers zn = xn + iyn converges to c = a + ib . if and only if the sequence of the real parts x1, x2, … converges to a . and the sequence of the imaginary parts y1, y2, … converges to b. Ex.
    [Show full text]
  • Topic 7 Notes 7 Taylor and Laurent Series
    Topic 7 Notes Jeremy Orloff 7 Taylor and Laurent series 7.1 Introduction We originally defined an analytic function as one where the derivative, defined as a limit of ratios, existed. We went on to prove Cauchy's theorem and Cauchy's integral formula. These revealed some deep properties of analytic functions, e.g. the existence of derivatives of all orders. Our goal in this topic is to express analytic functions as infinite power series. This will lead us to Taylor series. When a complex function has an isolated singularity at a point we will replace Taylor series by Laurent series. Not surprisingly we will derive these series from Cauchy's integral formula. Although we come to power series representations after exploring other properties of analytic functions, they will be one of our main tools in understanding and computing with analytic functions. 7.2 Geometric series Having a detailed understanding of geometric series will enable us to use Cauchy's integral formula to understand power series representations of analytic functions. We start with the definition: Definition. A finite geometric series has one of the following (all equivalent) forms. 2 3 n Sn = a(1 + r + r + r + ::: + r ) = a + ar + ar2 + ar3 + ::: + arn n X = arj j=0 n X = a rj j=0 The number r is called the ratio of the geometric series because it is the ratio of consecutive terms of the series. Theorem. The sum of a finite geometric series is given by a(1 − rn+1) S = a(1 + r + r2 + r3 + ::: + rn) = : (1) n 1 − r Proof.
    [Show full text]
  • 11.3-11.4 Integral and Comparison Tests
    11.3-11.4 Integral and Comparison Tests The Integral Test: Suppose a function f(x) is continuous, positive, and decreasing on [1; 1). Let an 1 P R 1 be defined by an = f(n). Then, the series an and the improper integral 1 f(x) dx either BOTH n=1 CONVERGE OR BOTH DIVERGE. Notes: • For the integral test, when we say that f must be decreasing, it is actually enough that f is EVENTUALLY ALWAYS DECREASING. In other words, as long as f is always decreasing after a certain point, the \decreasing" requirement is satisfied. • If the improper integral converges to a value A, this does NOT mean the sum of the series is A. Why? The integral of a function will give us all the area under a continuous curve, while the series is a sum of distinct, separate terms. • The index and interval do not always need to start with 1. Examples: Determine whether the following series converge or diverge. 1 n2 • X n2 + 9 n=1 1 2 • X n2 + 9 n=3 1 1 n • X n2 + 1 n=1 1 ln n • X n n=2 Z 1 1 p-series: We saw in Section 8.9 that the integral p dx converges if p > 1 and diverges if p ≤ 1. So, by 1 x 1 1 the Integral Test, the p-series X converges if p > 1 and diverges if p ≤ 1. np n=1 Notes: 1 1 • When p = 1, the series X is called the harmonic series. n n=1 • Any constant multiple of a convergent p-series is also convergent.
    [Show full text]
  • 1.1 Constructing the Real Numbers
    18.095 Lecture Series in Mathematics IAP 2015 Lecture #1 01/05/2015 What is number theory? The study of numbers of course! But what is a number? • N = f0; 1; 2; 3;:::g can be defined in several ways: { Finite ordinals (0 := fg, n + 1 := n [ fng). { Finite cardinals (isomorphism classes of finite sets). { Strings over a unary alphabet (\", \1", \11", \111", . ). N is a commutative semiring: addition and multiplication satisfy the usual commu- tative/associative/distributive properties with identities 0 and 1 (and 0 annihilates). Totally ordered (as ordinals/cardinals), making it a (positive) ordered semiring. • Z = {±n : n 2 Ng.A commutative ring (commutative semiring, additive inverses). Contains Z>0 = N − f0g closed under +; × with Z = −Z>0 t f0g t Z>0. This makes Z an ordered ring (in fact, an ordered domain). • Q = fa=b : a; 2 Z; b 6= 0g= ∼, where a=b ∼ c=d if ad = bc. A field (commutative ring, multiplicative inverses, 0 6= 1) containing Z = fn=1g. Contains Q>0 = fa=b : a; b 2 Z>0g closed under +; × with Q = −Q>0 t f0g t Q>0. This makes Q an ordered field. • R is the completion of Q, making it a complete ordered field. Each of the algebraic structures N; Z; Q; R is canonical in the following sense: every non-trivial algebraic structure of the same type (ordered semiring, ordered ring, ordered field, complete ordered field) contains a copy of N; Z; Q; R inside it. 1.1 Constructing the real numbers What do we mean by the completion of Q? There are two possibilities: 1.
    [Show full text]
  • Be a Metric Space
    2 The University of Sydney show that Z is closed in R. The complement of Z in R is the union of all the Pure Mathematics 3901 open intervals (n, n + 1), where n runs through all of Z, and this is open since every union of open sets is open. So Z is closed. Metric Spaces 2000 Alternatively, let (an) be a Cauchy sequence in Z. Choose an integer N such that d(xn, xm) < 1 for all n ≥ N. Put x = xN . Then for all n ≥ N we have Tutorial 5 |xn − x| = d(xn, xN ) < 1. But xn, x ∈ Z, and since two distinct integers always differ by at least 1 it follows that xn = x. This holds for all n > N. 1. Let X = (X, d) be a metric space. Let (xn) and (yn) be two sequences in X So xn → x as n → ∞ (since for all ε > 0 we have 0 = d(xn, x) < ε for all such that (yn) is a Cauchy sequence and d(xn, yn) → 0 as n → ∞. Prove that n > N). (i)(xn) is a Cauchy sequence in X, and 4. (i) Show that if D is a metric on the set X and f: Y → X is an injective (ii)(xn) converges to a limit x if and only if (yn) also converges to x. function then the formula d(a, b) = D(f(a), f(b)) defines a metric d on Y , and use this to show that d(m, n) = |m−1 − n−1| defines a metric Solution.
    [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]
  • Absolute Convergence in Ordered Fields
    Absolute Convergence in Ordered Fields Kristine Hampton Abstract This paper reviews Absolute Convergence in Ordered Fields by Clark and Diepeveen [1]. Contents 1 Introduction 2 2 Definition of Terms 2 2.1 Field . 2 2.2 Ordered Field . 3 2.3 The Symbol .......................... 3 2.4 Sequential Completeness and Cauchy Sequences . 3 2.5 New Sequences . 4 3 Summary of Results 4 4 Proof of Main Theorem 5 4.1 Main Theorem . 5 4.2 Proof . 6 5 Further Applications 8 5.1 Infinite Series in Ordered Fields . 8 5.2 Implications for Convergence Tests . 9 5.3 Uniform Convergence in Ordered Fields . 9 5.4 Integral Convergence in Ordered Fields . 9 1 1 Introduction In Absolute Convergence in Ordered Fields [1], the authors attempt to dis- tinguish between convergence and absolute convergence in ordered fields. In particular, Archimedean and non-Archimedean fields (to be defined later) are examined. In each field, the possibilities for absolute convergence with and without convergence are considered. Ultimately, the paper [1] attempts to offer conditions on various fields that guarantee if a series is convergent or not in that field if it is absolutely convergent. The results end up exposing a reliance upon sequential completeness in a field for any statements on the behavior of convergence in relation to absolute convergence to be made, and vice versa. The paper makes a noted attempt to be readable for a variety of mathematic levels, explaining new topics and ideas that might be confusing along the way. To understand the paper, only a basic understanding of series and convergence in R is required, although having a basic understanding of ordered fields would be ideal.
    [Show full text]
  • The Ratio Test, Integral Test, and Absolute Convergence
    MATH 1D, WEEK 3 { THE RATIO TEST, INTEGRAL TEST, AND ABSOLUTE CONVERGENCE INSTRUCTOR: PADRAIC BARTLETT Abstract. These are the lecture notes from week 3 of Ma1d, the Caltech mathematics course on sequences and series. 1. Homework 1 data • HW average: 91%. • Comments: none in particular { people seemed to be pretty capable with this material. 2. The Ratio Test So: thus far, we've developed a few tools for determining the convergence or divergence of series. Specifically, the main tools we developed last week were the two comparison tests, which gave us a number of tools for determining whether a series converged by comparing it to other, known series. However, sometimes this P 1 P n isn't enough; simply comparing things to n and r usually can only get us so far. Hence, the creation of the following test: Theorem 2.1. If fang is a sequence of positive numbers such that a lim n+1 = r; n!1 an for some real number r, then P1 • n=1 an converges if r < 1, while P1 • n=1 an diverges if r > 1. If r = 1, this test is inconclusive and tells us nothing. Proof. The basic idea motivating this theorem is the following: if the ratios an+1 an eventually approach some value r, then this series eventually \looks like" the geo- metric series P rn; consequently, it should converge whenever r < 1 and diverge whenever r > 1. an+1 To make the above concrete: suppose first that r > 1. Then, because limn!1 = an r;, we know that there is some N 2 N such that 8n ≥ N, a n+1 > 1 an )an+1 > an: But this means that the an's are eventually an increasing sequence starting at some value aN > 0; thus, we have that limn!1 an 6= 0 and thus that the sum P1 n=1 an cannot converge.
    [Show full text]
  • Infinitesimals Via Cauchy Sequences: Refining the Classical Equivalence
    INFINITESIMALS VIA CAUCHY SEQUENCES: REFINING THE CLASSICAL EQUIVALENCE EMANUELE BOTTAZZI AND MIKHAIL G. KATZ Abstract. A refinement of the classic equivalence relation among Cauchy sequences yields a useful infinitesimal-enriched number sys- tem. Such an approach can be seen as formalizing Cauchy’s senti- ment that a null sequence “becomes” an infinitesimal. We signal a little-noticed construction of a system with infinitesimals in a 1910 publication by Giuseppe Peano, reversing his earlier endorsement of Cantor’s belittling of infinitesimals. June 2, 2021 1. Historical background Robinson developed his framework for analysis with infinitesimals in his 1966 book [1]. There have been various attempts either (1) to obtain clarity and uniformity in developing analysis with in- finitesimals by working in a version of set theory with a concept of infinitesimal built in syntactically (see e.g., [2], [3]), or (2) to define nonstandard universes in a simplified way by providing mathematical structures that would not have all the features of a Hewitt–Luxemburg-style ultrapower [4], [5], but nevertheless would suffice for basic application and possibly could be helpful in teaching. The second approach has been carried out, for example, by James Henle arXiv:2106.00229v1 [math.LO] 1 Jun 2021 in his non-nonstandard Analysis [6] (based on Schmieden–Laugwitz [7]; see also [8]), as well as by Terry Tao in his “cheap nonstandard analysis” [9]. These frameworks are based upon the identification of real sequences whenever they are eventually equal.1 A precursor of this approach is found in the work of Giuseppe Peano. In his 1910 paper [10] (see also [11]) he introduced the notion of the end (fine; pl.
    [Show full text]
  • Practice Problems for the Final
    Practice problems for the final 1. Show the following using just the definitions (and no theorems). (a) The function ( x2; x ≥ 0 f(x) = 0; x < 0; is differentiable on R, while f 0(x) is not differentiable at 0. Solution: For x > 0 and x < 0, the function is clearly differentiable. We check at x = 0. The difference quotient is f(x) − f(0) f(x) '(x) = = : x x So x2 '(0+) = lim '(x) = lim = 0; x!0+ x!0+ x 0 '(0−) = lim '(x) = lim = 0: x!0− x!0− x 0 Since '(0+) = '(0−), limx!0 '(x) = 0 and hence f is differentiable at 0 with f (0) = 0. So together we get ( 2x; x > 0 f 0(x) = 0; x ≤ 0: f 0(x) is again clearly differentiable for x > 0 and x < 0. To test differentiability at x = 0, we consider the difference quotient f 0(x) − f 0(0) f 0(x) (x) = = : x x Then it is easy to see that (0+) = 2 while (0−) = 0. Since (0+) 6= (0−), limx!0 (x) does not exist, and f 0 is not differentiable at x = 0. (b) The function f(x) = x3 + x is continuous but not uniformly continuous on R. Solution: We show that f is continuous at some x = a. That is given " > 0 we have to find a corresponding δ. jf(x) − f(a)j = jx3 + x − a3 − aj = j(x − a)(x2 + xa + a2) + (x − a)j = jx − ajjx2 + xa + a2 + 1j: Now suppose we pick δ < 1, then jx − aj < δ =) jxj < jaj + 1, and so jx2 + xa + a2 + 1j ≤ jxj2 + jxjjaj + a2 + 1 ≤ 3(jaj + 1)2: To see the final inequality, note that jxjjaj < (jaj + 1)jaj < (jaj + 1)2, and a2 + 1 < (jaj + 1)2.
    [Show full text]
  • Infinitesimals
    Infinitesimals: History & Application Joel A. Tropp Plan II Honors Program, WCH 4.104, The University of Texas at Austin, Austin, TX 78712 Abstract. An infinitesimal is a number whose magnitude ex- ceeds zero but somehow fails to exceed any finite, positive num- ber. Although logically problematic, infinitesimals are extremely appealing for investigating continuous phenomena. They were used extensively by mathematicians until the late 19th century, at which point they were purged because they lacked a rigorous founda- tion. In 1960, the logician Abraham Robinson revived them by constructing a number system, the hyperreals, which contains in- finitesimals and infinitely large quantities. This thesis introduces Nonstandard Analysis (NSA), the set of techniques which Robinson invented. It contains a rigorous de- velopment of the hyperreals and shows how they can be used to prove the fundamental theorems of real analysis in a direct, natural way. (Incredibly, a great deal of the presentation echoes the work of Leibniz, which was performed in the 17th century.) NSA has also extended mathematics in directions which exceed the scope of this thesis. These investigations may eventually result in fruitful discoveries. Contents Introduction: Why Infinitesimals? vi Chapter 1. Historical Background 1 1.1. Overview 1 1.2. Origins 1 1.3. Continuity 3 1.4. Eudoxus and Archimedes 5 1.5. Apply when Necessary 7 1.6. Banished 10 1.7. Regained 12 1.8. The Future 13 Chapter 2. Rigorous Infinitesimals 15 2.1. Developing Nonstandard Analysis 15 2.2. Direct Ultrapower Construction of ∗R 17 2.3. Principles of NSA 28 2.4. Working with Hyperreals 32 Chapter 3.
    [Show full text]
  • Uniform Convergence
    2018 Spring MATH2060A Mathematical Analysis II 1 Notes 3. UNIFORM CONVERGENCE Uniform convergence is the main theme of this chapter. In Section 1 pointwise and uniform convergence of sequences of functions are discussed and examples are given. In Section 2 the three theorems on exchange of pointwise limits, inte- gration and differentiation which are corner stones for all later development are proven. They are reformulated in the context of infinite series of functions in Section 3. The last two important sections demonstrate the power of uniform convergence. In Sections 4 and 5 we introduce the exponential function, sine and cosine functions based on differential equations. Although various definitions of these elementary functions were given in more elementary courses, here the def- initions are the most rigorous one and all old ones should be abandoned. Once these functions are defined, other elementary functions such as the logarithmic function, power functions, and other trigonometric functions can be defined ac- cordingly. A notable point is at the end of the section, a rigorous definition of the number π is given and showed to be consistent with its geometric meaning. 3.1 Uniform Convergence of Functions Let E be a (non-empty) subset of R and consider a sequence of real-valued func- tions ffng; n ≥ 1 and f defined on E. We call ffng pointwisely converges to f on E if for every x 2 E, the sequence ffn(x)g of real numbers converges to the number f(x). The function f is called the pointwise limit of the sequence. According to the limit of sequence, pointwise convergence means, for each x 2 E, given " > 0, there is some n0(x) such that jfn(x) − f(x)j < " ; 8n ≥ n0(x) : We use the notation n0(x) to emphasis the dependence of n0(x) on " and x.
    [Show full text]