Euler's Factorial Series and Global Relations

Total Page:16

File Type:pdf, Size:1020Kb

Euler's Factorial Series and Global Relations Euler’s factorial series and global relations Tapani Matala-aho1 and Wadim Zudilin2,3 1Matematiikka, PL 3000, 90014 Oulun yliopisto, Finland 2Institute for Mathematics, Astrophysics and Particle Physics, Radboud Universiteit, PO Box 9010, 6500 GL Nijmegen, The Netherlands 3School of Mathematical and Physical Sciences, The University of Newcastle, Callaghan, NSW 2308, Australia 7 March 2017 To the memory of Marc Huttner (1947–2015) Abstract k Using Pad´eapproximations to the series E(z) = ∞ k! ( z) , we address k=0 − arithmetic and analytical questions related to its valuesP in both p-adic and Archimedean valuations. 1 Introduction In his 1760 paper on divergent series [5], L. Euler introduced and studied the formal (hypergeometric) series ∞ E(z) := k!( z)k (1) − Xk=0 arXiv:1703.02633v1 [math.NT] 7 Mar 2017 (see also [1, 2] and, in particular, [8, Section 2.5]), which specializes at z = 1 to Wallis’ series ∞ W := ( 1)kk! , (2) − Xk=0 which has teased people’s imagination since the times of Euler. Notice that the series (1) is a perfectly convergent p-adic series in the disc z p 1 for all primes p, so that one can discuss the arithmetic properties of its values,| for| example≤ , at z = 1 (the Wallis case) and at z = 1. The irrationality of the latter specialization, − ∞ K = Kp := k! , (3) Xk=0 1 for any prime p is a folklore conjecture [12, p. 17] (see [11], also for a link of the problem to a combinatorial conjecture of H. Wilf). Already the expectation K = 1 (so that | p|p Kp is a p-adic unit) for all primes p, which is an equivalent form of Kurepa’s conjecture [7] from 1971, remains open; see [4] for the latest achievements in this direction. In what follows, we refer to the series in (1) as Euler’s factorial series and label it Ep(z) when treat it as the function in the p-adic domain for a given prime p. Theorem 1. Given ξ Z 0 , let be a subset of prime numbers such that ∈ \{ } P n 2 2 lim sup c n! n! p =0, where c = c(ξ; ) := 4 ξ ξ p. (4) n | | P | | | | →∞ pY pY ∈P ∈P Then either there exists a prime p for which Ep(ξ) is irrational, or there are two distinct primes p, q such that E∈( Pξ) = E (ξ) (while E (ξ), E (ξ) Q). ∈ P p 6 q p q ∈ Because n! =1/n! when the product is taken over all primes, condition (4) is p | |p clearly satisfiedQ for any subset whose complement in the set of all primes is finite. P This also suggests more exotic choices of . Furthermore, the conclusion of the theorem is contrasted with Euler’s sum P ∞ k k!= 1, · − Xk=0 n 1 which is valid in any p-adic valuation (and follows from − k k! = n! 1)—an k=0 · − example of what is called a global relation. P The idea behind the proof of Theorem 1 is to construct approximations pn/qn to the number ωp = Ep(ξ) in question, which do not depend on p and approximate the number considerably well for each p-adic valuation: q ω p = r for n =0, 1, 2,... ; n p − n n,p r 0 as n and there are infinitely many indices n for which r = 0 for | n,p|p → → ∞ n,p 6 at least one prime p . Assume that ωp = a/b, the same rational number, for all p . Then q a p∈b P Z 0 , so that 0 < q a p b 1, for infinitely many ∈ P n − n ∈ \{ } | n − n |p ≤ indices n and all primes p, hence 1= q a p b q a p b q a p b q a p b | n − n | | n − n |p ≤| n − n | | n − n |p Yp pY ∈P ( a + b ) max q , p br ≤ | | | | {| n| | n|} | n,p|p pY ∈P ( a + b ) max q , p r ≤ | | | | {| n| | n|} | n,p|p pY ∈P for those n. This means that the condition lim sup max qn , pn rn,p p =0 (5) n {| | | |} | | →∞ pY ∈P contradicts the latter estimate, thus making the -global linear relation ωp = a/b impossible. P 2 The result in Theorem 1 can be put in a general context of global relations for Euler-type series; the corresponding settings can be found in the paper [3]. We do not pursue this route here as our principal motivation is a sufficiently elementary arithmetic treatment of an analytical function that have some historical value. The rational approximations to E(z) we construct in Section 2 are Pad´eapproximations; in spite of being known for centuries, implicitly from the continued fraction for E(z) given by Euler himself [5] and explicitly from the work of T. J. Stieltjes [13], these Pad´e approximations remain a useful source for arithmetic and analytical investigations. In Section 3 we revisit Euler’s summation of Wallis’ series (2) using the approximations; we also complement the derivation by providing ‘Archimedean analogue(s)’ for the divergent series K = E( 1) from (3) — the case when the classical strategy does not − work. Our way of constructing the Pad´eapproximations is inspired by a related Pad´e construction of M. Hata and M. Huttner in [6]. The construction was a particular favourite of Marc Huttner who, for the span of his mathematical life, remained a passionate advocate of interplay between Picard–Fuchs linear differential equations and Pad´eapproximations. We dedicate this work to his memory. 2 Hypergeometric series and Pad´eapproximations Euler’s factorial series (1) is the particular a = 1 instance of the hypergeometric series ∞ k 2F0(a, 1 z)= (a) z . (6) | k Xk=0 Here and in what follows, we use the Pochhammer notation (a)k which is defined inductively by (a)0 = 1and(a)k+1 =(a+k)(a)k for k Z 0. Our Pad´eapproximations ≥ below are given more generally for the function (6). ∈ Theorem 2. For n, λ Z 0, take ∈ ≥ n i n i n ( 1) z − Bn,λ(z)= − . i (a) + Xi=0 i λ Then degz Bn,λ = n and for a polynomial An,λ(z) of degree degz An,λ n + λ 1 we have ≤ − B (z) 2F0(a, 1 z) A (z)= L (z), (7) n,λ | − n,λ n,λ where ordz=0 Ln,λ(z)=2n + λ. Explicitly, ∞ n + k n + k + a + λ 1 L (z)=( 1)nn! z2n+λ k! − zk. (8) n,λ − k k Xk=0 Proof. Relation (7) means that there is a ‘gap’ of length n in the power series expansion B (z) 2F0(a, 1 z)= A (z)+ L (z). n,λ | n,λ n,λ 3 n h Write Bn,λ(z)= h=0 bht and consider the series expansion of the product P ∞ l B (z) 2F0(a, 1 z)= r z , n,λ | l Xl=0 where rl = h+k=l bh(a)k; in particular, P n n i n (a)i+λ+m i n rl = ( 1) = ( 1) (a + i + λ)m − i (a) + − i Xi=0 i λ Xi=0 with m = l n λ for l > n + λ 1. To verify the desired ‘gap’ condition, − − − rn+λ = rn+λ+1 = = rn+λ+n 1 =0, (9) ··· − we introduce the shift operators N = N and ∆ = ∆ = N id defined on functions a a − f(a) by Nf(a)= f(a +1) and ∆f(a)= f(a + 1) f(a). − It follows from n n ∆ (a + λ)m =( 1) ( m)n(a + λ + n)m n for n Z 0 − − − ∈ ≥ that n n r = ( 1)i N i(a + λ) = (id N)n(a + λ) l − i m − m Xi=0 n =( ∆) (a + λ)m =( m)n(a + λ + n)m n, − − − which, in turn, implies (9) because of the vanishing of ( m) for m =0, 1,...,n 1. − n − The explicit expression for rl just found also gives n n + k n + k + a + λ 1 r2 + + =( n k) (a + λ + n) =( 1) n! k! − , n λ k − − n k − k k hence the closed form (8). We also have n+λ 1 n+λ 1 n − − l l i n (a)i+l n An,λ(z)= rlz = z ( 1) − . − i (a)i+λ Xl=0 Xl=0 Xi=0 i n l ≥ − This concludes our proof of the theorem. From now on we choose a = 1, λ = 0, change z to z and renormalize the − corresponding Pad´eapproximations produced by Theorem 2 by multiplying them by 4 ( 1)nn!: − n n 2 n n n! n i n i Q (z) := ( 1) n! B 0( z)= z − = i! z , n − n, − i i! i Xi=0 Xi=0 n 1 n − n n l n i n n!(i + l n)! P (z) := ( 1) n! A 0( z)=( 1) ( z) ( 1) − − n − n, − − − − i i! Xl=0 i=Xn l − n 1 l 2 − n =( 1)n ( z)l ( 1)ii!(l i)! − − − − i Xl=0 Xi=0 and 2 ∞ n 2 2n k n + k k R (z) := ( 1) n! L 0( z)= n! z ( 1) k! z . n − n, − − k Xk=0 In this notation, the Pad´eapproximation formula (7) may be rewritten as Q (z)E(z) P (z)= R (z) (10) n − n n for n Z 0. Observe that ∈ > 2 2n Q (z)P +1(z) Q +1(z)P (z)= n! z . (11) n n − n n Indeed, the standard Pad´ewalkabout proves the identity Q (z)P +1(z) Q +1(z)P (z)= Q +1(z)R (z) Q (z)R +1(z), n n − n n n n − n n in which the degree of the left-hand side is at most 2n while the order of the right-hand side is at least 2n.
Recommended publications
  • 1 Approximating Integrals Using Taylor Polynomials 1 1.1 Definitions
    Seunghee Ye Ma 8: Week 7 Nov 10 Week 7 Summary This week, we will learn how we can approximate integrals using Taylor series and numerical methods. Topics Page 1 Approximating Integrals using Taylor Polynomials 1 1.1 Definitions . .1 1.2 Examples . .2 1.3 Approximating Integrals . .3 2 Numerical Integration 5 1 Approximating Integrals using Taylor Polynomials 1.1 Definitions When we first defined the derivative, recall that it was supposed to be the \instantaneous rate of change" of a function f(x) at a given point c. In other words, f 0 gives us a linear approximation of f(x) near c: for small values of " 2 R, we have f(c + ") ≈ f(c) + "f 0(c) But if f(x) has higher order derivatives, why stop with a linear approximation? Taylor series take this idea of linear approximation and extends it to higher order derivatives, giving us a better approximation of f(x) near c. Definition (Taylor Polynomial and Taylor Series) Let f(x) be a Cn function i.e. f is n-times continuously differentiable. Then, the n-th order Taylor polynomial of f(x) about c is: n X f (k)(c) T (f)(x) = (x − c)k n k! k=0 The n-th order remainder of f(x) is: Rn(f)(x) = f(x) − Tn(f)(x) If f(x) is C1, then the Taylor series of f(x) about c is: 1 X f (k)(c) T (f)(x) = (x − c)k 1 k! k=0 Note that the first order Taylor polynomial of f(x) is precisely the linear approximation we wrote down in the beginning.
    [Show full text]
  • A Quotient Rule Integration by Parts Formula Jennifer Switkes ([email protected]), California State Polytechnic Univer- Sity, Pomona, CA 91768
    A Quotient Rule Integration by Parts Formula Jennifer Switkes ([email protected]), California State Polytechnic Univer- sity, Pomona, CA 91768 In a recent calculus course, I introduced the technique of Integration by Parts as an integration rule corresponding to the Product Rule for differentiation. I showed my students the standard derivation of the Integration by Parts formula as presented in [1]: By the Product Rule, if f (x) and g(x) are differentiable functions, then d f (x)g(x) = f (x)g(x) + g(x) f (x). dx Integrating on both sides of this equation, f (x)g(x) + g(x) f (x) dx = f (x)g(x), which may be rearranged to obtain f (x)g(x) dx = f (x)g(x) − g(x) f (x) dx. Letting U = f (x) and V = g(x) and observing that dU = f (x) dx and dV = g(x) dx, we obtain the familiar Integration by Parts formula UdV= UV − VdU. (1) My student Victor asked if we could do a similar thing with the Quotient Rule. While the other students thought this was a crazy idea, I was intrigued. Below, I derive a Quotient Rule Integration by Parts formula, apply the resulting integration formula to an example, and discuss reasons why this formula does not appear in calculus texts. By the Quotient Rule, if f (x) and g(x) are differentiable functions, then ( ) ( ) ( ) − ( ) ( ) d f x = g x f x f x g x . dx g(x) [g(x)]2 Integrating both sides of this equation, we get f (x) g(x) f (x) − f (x)g(x) = dx.
    [Show full text]
  • Operations on Power Series Related to Taylor Series
    Operations on Power Series Related to Taylor Series In this problem, we perform elementary operations on Taylor series – term by term differen­ tiation and integration – to obtain new examples of power series for which we know their sum. Suppose that a function f has a power series representation of the form: 1 2 X n f(x) = a0 + a1(x − c) + a2(x − c) + · · · = an(x − c) n=0 convergent on the interval (c − R; c + R) for some R. The results we use in this example are: • (Differentiation) Given f as above, f 0(x) has a power series expansion obtained by by differ­ entiating each term in the expansion of f(x): 1 0 X n−1 f (x) = a1 + a2(x − c) + 2a3(x − c) + · · · = nan(x − c) n=1 • (Integration) Given f as above, R f(x) dx has a power series expansion obtained by by inte­ grating each term in the expansion of f(x): 1 Z a1 a2 X an f(x) dx = C + a (x − c) + (x − c)2 + (x − c)3 + · · · = C + (x − c)n+1 0 2 3 n + 1 n=0 for some constant C depending on the choice of antiderivative of f. Questions: 1. Find a power series representation for the function f(x) = arctan(5x): (Note: arctan x is the inverse function to tan x.) 2. Use power series to approximate Z 1 2 sin(x ) dx 0 (Note: sin(x2) is a function whose antiderivative is not an elementary function.) Solution: 1 For question (1), we know that arctan x has a simple derivative: , which then has a power 1 + x2 1 2 series representation similar to that of , where we subsitute −x for x.
    [Show full text]
  • Derivative of Power Series and Complex Exponential
    LECTURE 4: DERIVATIVE OF POWER SERIES AND COMPLEX EXPONENTIAL The reason of dealing with power series is that they provide examples of analytic functions. P1 n Theorem 1. If n=0 anz has radius of convergence R > 0; then the function P1 n F (z) = n=0 anz is di®erentiable on S = fz 2 C : jzj < Rg; and the derivative is P1 n¡1 f(z) = n=0 nanz : Proof. (¤) We will show that j F (z+h)¡F (z) ¡ f(z)j ! 0 as h ! 0 (in C), whenever h ¡ ¢ n Pn n k n¡k jzj < R: Using the binomial theorem (z + h) = k=0 k h z we get F (z + h) ¡ F (z) X1 (z + h)n ¡ zn ¡ hnzn¡1 ¡ f(z) = a h n h n=0 µ ¶ X1 a Xn n = n ( hkzn¡k) h k n=0 k=2 µ ¶ X1 Xn n = a h( hk¡2zn¡k) n k n=0 k=2 µ ¶ X1 Xn¡2 n = a h( hjzn¡2¡j) (by putting j = k ¡ 2): n j + 2 n=0 j=0 ¡ n ¢ ¡n¡2¢ By using the easily veri¯able fact that j+2 · n(n ¡ 1) j ; we obtain µ ¶ F (z + h) ¡ F (z) X1 Xn¡2 n ¡ 2 j ¡ f(z)j · jhj n(n ¡ 1)ja j( jhjjjzjn¡2¡j) h n j n=0 j=0 X1 n¡2 = jhj n(n ¡ 1)janj(jzj + jhj) : n=0 P1 n¡2 We already know that the series n=0 n(n ¡ 1)janjjzj converges for jzj < R: Now, for jzj < R and h ! 0 we have jzj + jhj < R eventually.
    [Show full text]
  • Notes on Euler's Work on Divergent Factorial Series and Their Associated
    Indian J. Pure Appl. Math., 41(1): 39-66, February 2010 °c Indian National Science Academy NOTES ON EULER’S WORK ON DIVERGENT FACTORIAL SERIES AND THEIR ASSOCIATED CONTINUED FRACTIONS Trond Digernes¤ and V. S. Varadarajan¤¤ ¤University of Trondheim, Trondheim, Norway e-mail: [email protected] ¤¤University of California, Los Angeles, CA, USA e-mail: [email protected] Abstract Factorial series which diverge everywhere were first considered by Euler from the point of view of summing divergent series. He discovered a way to sum such series and was led to certain integrals and continued fractions. His method of summation was essentialy what we call Borel summation now. In this paper, we discuss these aspects of Euler’s work from the modern perspective. Key words Divergent series, factorial series, continued fractions, hypergeometric continued fractions, Sturmian sequences. 1. Introductory Remarks Euler was the first mathematician to develop a systematic theory of divergent se- ries. In his great 1760 paper De seriebus divergentibus [1, 2] and in his letters to Bernoulli he championed the view, which was truly revolutionary for his epoch, that one should be able to assign a numerical value to any divergent series, thus allowing the possibility of working systematically with them (see [3]). He antic- ipated by over a century the methods of summation of divergent series which are known today as the summation methods of Cesaro, Holder,¨ Abel, Euler, Borel, and so on. Eventually his views would find their proper place in the modern theory of divergent series [4]. But from the beginning Euler realized that almost none of his methods could be applied to the series X1 1 ¡ 1!x + 2!x2 ¡ 3!x3 + ::: = (¡1)nn!xn (1) n=0 40 TROND DIGERNES AND V.
    [Show full text]
  • Calculus Terminology
    AP Calculus BC Calculus Terminology Absolute Convergence Asymptote Continued Sum Absolute Maximum Average Rate of Change Continuous Function Absolute Minimum Average Value of a Function Continuously Differentiable Function Absolutely Convergent Axis of Rotation Converge Acceleration Boundary Value Problem Converge Absolutely Alternating Series Bounded Function Converge Conditionally Alternating Series Remainder Bounded Sequence Convergence Tests Alternating Series Test Bounds of Integration Convergent Sequence Analytic Methods Calculus Convergent Series Annulus Cartesian Form Critical Number Antiderivative of a Function Cavalieri’s Principle Critical Point Approximation by Differentials Center of Mass Formula Critical Value Arc Length of a Curve Centroid Curly d Area below a Curve Chain Rule Curve Area between Curves Comparison Test Curve Sketching Area of an Ellipse Concave Cusp Area of a Parabolic Segment Concave Down Cylindrical Shell Method Area under a Curve Concave Up Decreasing Function Area Using Parametric Equations Conditional Convergence Definite Integral Area Using Polar Coordinates Constant Term Definite Integral Rules Degenerate Divergent Series Function Operations Del Operator e Fundamental Theorem of Calculus Deleted Neighborhood Ellipsoid GLB Derivative End Behavior Global Maximum Derivative of a Power Series Essential Discontinuity Global Minimum Derivative Rules Explicit Differentiation Golden Spiral Difference Quotient Explicit Function Graphic Methods Differentiable Exponential Decay Greatest Lower Bound Differential
    [Show full text]
  • Euler and His Work on Infinite Series
    BULLETIN (New Series) OF THE AMERICAN MATHEMATICAL SOCIETY Volume 44, Number 4, October 2007, Pages 515–539 S 0273-0979(07)01175-5 Article electronically published on June 26, 2007 EULER AND HIS WORK ON INFINITE SERIES V. S. VARADARAJAN For the 300th anniversary of Leonhard Euler’s birth Table of contents 1. Introduction 2. Zeta values 3. Divergent series 4. Summation formula 5. Concluding remarks 1. Introduction Leonhard Euler is one of the greatest and most astounding icons in the history of science. His work, dating back to the early eighteenth century, is still with us, very much alive and generating intense interest. Like Shakespeare and Mozart, he has remained fresh and captivating because of his personality as well as his ideas and achievements in mathematics. The reasons for this phenomenon lie in his universality, his uniqueness, and the immense output he left behind in papers, correspondence, diaries, and other memorabilia. Opera Omnia [E], his collected works and correspondence, is still in the process of completion, close to eighty volumes and 31,000+ pages and counting. A volume of brief summaries of his letters runs to several hundred pages. It is hard to comprehend the prodigious energy and creativity of this man who fueled such a monumental output. Even more remarkable, and in stark contrast to men like Newton and Gauss, is the sunny and equable temperament that informed all of his work, his correspondence, and his interactions with other people, both common and scientific. It was often said of him that he did mathematics as other people breathed, effortlessly and continuously.
    [Show full text]
  • 20-Finding Taylor Coefficients
    Taylor Coefficients Learning goal: Let’s generalize the process of finding the coefficients of a Taylor polynomial. Let’s find a general formula for the coefficients of a Taylor polynomial. (Students complete worksheet series03 (Taylor coefficients).) 2 What we have discovered is that if we have a polynomial C0 + C1x + C2x + ! that we are trying to match values and derivatives to a function, then when we differentiate it a bunch of times, we 0 get Cnn!x + Cn+1(n+1)!x + !. Plugging in x = 0 and everything except Cnn! goes away. So to (n) match the nth derivative, we must have Cn = f (0)/n!. If we want to match the values and derivatives at some point other than zero, we will use the 2 polynomial C0 + C1(x – a) + C2(x – a) + !. Then when we differentiate, we get Cnn! + Cn+1(n+1)!(x – a) + !. Now, plugging in x = a makes everything go away, and we get (n) Cn = f (a)/n!. So we can generate a polynomial of any degree (well, as long as the function has enough derivatives!) that will match the function and its derivatives to that degree at a particular point a. We can also extend to create a power series called the Taylor series (or Maclaurin series if a is specifically 0). We have natural questions: • For what x does the series converge? • What does the series converge to? Is it, hopefully, f(x)? • If the series converges to the function, we can use parts of the series—Taylor polynomials—to approximate the function, at least nearby a.
    [Show full text]
  • On the Euler Integral for the Positive and Negative Factorial
    On the Euler Integral for the positive and negative Factorial Tai-Choon Yoon ∗ and Yina Yoon (Dated: Dec. 13th., 2020) Abstract We reviewed the Euler integral for the factorial, Gauss’ Pi function, Legendre’s gamma function and beta function, and found that gamma function is defective in Γ(0) and Γ( x) − because they are undefined or indefinable. And we came to a conclusion that the definition of a negative factorial, that covers the domain of the negative space, is needed to the Euler integral for the factorial, as well as the Euler Y function and the Euler Z function, that supersede Legendre’s gamma function and beta function. (Subject Class: 05A10, 11S80) A. The positive factorial and the Euler Y function Leonhard Euler (1707–1783) developed a transcendental progression in 1730[1] 1, which is read xedx (1 x)n. (1) Z − f From this, Euler transformed the above by changing e to g for generalization into f x g dx (1 x)n. (2) Z − Whence, Euler set f = 1 and g = 0, and got an integral for the factorial (!) 2, dx ( lx )n, (3) Z − where l represents logarithm . This is called the Euler integral of the second kind 3, and the equation (1) is called the Euler integral of the first kind. 4, 5 Rewriting the formula (3) as follows with limitation of domain for a positive half space, 1 1 n ln dx, n 0. (4) Z0 x ≥ ∗ Electronic address: [email protected] 1 “On Transcendental progressions that is, those whose general terms cannot be given algebraically” by Leonhard Euler p.3 2 ibid.
    [Show full text]
  • The Set of Continuous Functions with Everywhere Convergent Fourier Series
    TRANSACTIONS OF THE AMERICAN MATHEMATICAL SOCIETY Volume 302, Number 1, July 1987 THE SET OF CONTINUOUS FUNCTIONS WITH EVERYWHERE CONVERGENT FOURIER SERIES M. AJTAI AND A. S. KECHRIS ABSTRACT. This paper deals with the descriptive set theoretic properties of the class EC of continuous functions with everywhere convergent Fourier se- ries. It is shown that this set is a complete coanalytic set in C(T). A natural coanalytic rank function on EC is studied that assigns to each / € EC a count- able ordinal number, which measures the "complexity" of the convergence of the Fourier series of /. It is shown that there exist functions in EC (in fact even differentiable ones) which have arbitrarily large countable rank, so that this provides a proper hierarchy on EC with uii distinct levels. Let G(T) be the Banach space of continuous 27r-periodic functions on the reals with the sup norm. We study in this paper descriptive set theoretic aspects of the subset EC of G(T) consisting of the continuous functions with everywhere convergent Fourier series. It is easy to verify that EC is a coanalytic set. We show in §3 that EC is actually a complete coanalytic set and therefore in particular is not Borel. This answers a question in [Ku] and provides another example of a coanalytic non-Borel set. We also discuss in §2 (see in addition the Appendix) the relations of this kind of result to the studies in the literature concerning the classification of the set of points of divergence of the Fourier series of a continuous function.
    [Show full text]
  • Calculus of Variations Raju K George, IIST
    Calculus of Variations Raju K George, IIST Lecture-1 In Calculus of Variations, we will study maximum and minimum of a certain class of functions. We first recall some maxima/minima results from the classical calculus. Maxima and Minima Let X and Y be two arbitrary sets and f : X Y be a well-defined function having domain → X and range Y . The function values f(x) become comparable if Y is IR or a subset of IR. Thus, optimization problem is valid for real valued functions. Let f : X IR be a real valued → function having X as its domain. Now x X is said to be maximum point for the function f if 0 ∈ f(x ) f(x) x X. The value f(x ) is called the maximum value of f. Similarly, x X is 0 ≥ ∀ ∈ 0 0 ∈ said to be a minimum point for the function f if f(x ) f(x) x X and in this case f(x ) is 0 ≤ ∀ ∈ 0 the minimum value of f. Sufficient condition for having maximum and minimum: Theorem (Weierstrass Theorem) Let S IR and f : S IR be a well defined function. Then f will have a maximum/minimum ⊆ → under the following sufficient conditions. 1. f : S IR is a continuous function. → 2. S IR is a bound and closed (compact) subset of IR. ⊂ Note that the above conditions are just sufficient conditions but not necessary. Example 1: Let f : [ 1, 1] IR defined by − → 1 x = 0 f(x)= − x x = 0 | | 6 1 −1 +1 −1 Obviously f(x) is not continuous at x = 0.
    [Show full text]
  • Taylor's Series of Sin X
    Taylor's Series of sin x In order to use Taylor's formula to find the power series expansion of sin x we have to compute the derivatives of sin(x): sin0(x) = cos(x) sin00(x) = − sin(x) sin000(x) = − cos(x) sin(4)(x) = sin(x): Since sin(4)(x) = sin(x), this pattern will repeat. Next we need to evaluate the function and its derivatives at 0: sin(0) = 0 sin0(0) = 1 sin00(0) = 0 sin000(0) = −1 sin(4)(0) = 0: Again, the pattern repeats. Taylor's formula now tells us that: −1 sin(x) = 0 + 1x + 0x 2 + x 3 + 0x4 + · · · 3! x3 x5 x7 = x − + − + · · · 3! 5! 7! Notice that the signs alternate and the denominators get very big; factorials grow very fast. The radius of convergence R is infinity; let's see why. The terms in this sum look like: x2n+1 x x x x = · · · · · : (2n + 1)! 1 2 3 (2n + 1) Suppose x is some fixed number. Then as n goes to infinity, the terms on the right in the product above will be very, very small numbers and there will be more and more of them as n increases. In other words, the terms in the series will get smaller as n gets bigger; that's an indication that x may be inside the radius of convergence. But this would be true for any fixed value of x, so the radius of convergence is infinity. Why do we care what the power series expansion of sin(x) is? If we use enough terms of the series we can get a good estimate of the value of sin(x) for any value of x.
    [Show full text]