An Application of the Dirichlet Integrals to the Summation of Series

Total Page:16

File Type:pdf, Size:1020Kb

An Application of the Dirichlet Integrals to the Summation of Series Some applications of the Dirichlet integrals to the summation of series and the evaluation of integrals involving the Riemann zeta function Donal F. Connon [email protected] 1 December 2012 Abstract Using the Dirichlet integrals, which are employed in the theory of Fourier series, this paper develops a useful method for the summation of series and the evaluation of integrals. CONTENTS Page 1. Introduction 1 2. Application of Dirichlet’s integrals 5 3. Some applications of the Poisson summation formula: 3.1 Digamma function 11 3.2 Log gamma function 13 3.3 Logarithm function 14 3.4 An integral due to Ramanujan 16 3.5 Stieltjes constants 36 3.6 Connections with other summation formulae 40 3.7 Hurwitz zeta function 48 4. Some integrals involving cot x 62 5. Open access to our own work 68 1. Introduction In two earlier papers [19] and [20] we considered the following identity (which is easily verified by multiplying the numerator and the denominator by the complex conjugate (1− e−ix ) ) 1 1 isin x 1 i (1.1) = − = + cot(x / 2) 1− eix 2 2 1− cos x 2 2 Using the basic geometric series identity 1 N y N +1 =∑ yn + 1− y n=0 1− y we obtain N 1 inx (1.1a) ix =∑e + RN () x 1− e n=0 1 iN()+ x iN(1)+ x 2 ei1 iN(1)+ x e where RxN () == e[]1 + i cot( x / 2) = 12− exix 2sin(/2) Separating the real and imaginary parts of (1.1a) produces the following two identities (the first of which is called Lagrange’s trigonometric identity and contains the Dirichlet kernel DxN () which is employed in the theory of Fourier series [55, p.49]) 1 N sin(N+ 1/ 2) x (1.2) = ∑cosnx − 2 n=0 2sin(x / 2) 1 N cos(N+ 1/ 2) x (1.3) cot(x / 2)=∑ sin nx + 2 n=0 2sin(x / 2) (and these equations may obviously be generalised by substituting αx in place of x . We first multiply by p()x and, in this section, p()x is assumed to be twice continuously differentiable on [,ab]. Integration then gives us 1 bbN p(xixdxpxedxR ) 1+= cot( / 2) ( ) inx + ∫∫()∑ N 2 aan=0 and we have bb1iN cos(+−+ 1/2) x sin( N 1/2) x R ==p()xR () xdx p ()xdx NN∫∫ aa2sin(x /2) 1 b ⎧ cos(x / 2) ⎫ =− ∫ p()cosxNx⎨ + sin Nx⎬dx 2sa ⎩⎭in(x /2) ixb ⎧ cos( / 2) ⎫ +−+∫ p()xNx⎨ sin cos Nx⎬dx 2sa ⎩⎭in(x /2) Therefore, provided sin(x / 2) has no zero in [ ab, ], a weak version of the Riemann- Lebesgue lemma (which was previously employed in Eq.(2.17) in [19] where we required p()x to be twice continuously differentiable) tells us that limRN = 0. N →∞ 2 This will also be the case if sin()a / 2= 0 , provided p( a )= 0. From the above we can therefore derive the following trigonometric integral identities: 1 bb∞ (1.4) ∫∫p()xdx= ∑ px ()cos nxdx 2 aan=0 1 bb∞ (1.4a) ∫∫p()cot(/2)x x dx= ∑ p ()sin x nx dx 2 aan=0 and more generally we have (by letting x → α x in (1.2) and (1.3) respectively) 1 bb∞ (1.5) ∫∫p()xdx= ∑ px ()cosα nxdx 2 aan=0 1 bb∞ (1.5a) ∫∫p(x )cot(αα x / 2) dx= ∑ p ( x )sin nx dx 2 aan=0 Equations (1.5) and (1.5a) are valid provided (i) sin(α x / 2)≠ 0 ∀∈ xab[ , ] or, alternatively, (ii) if sin(αa / 2)= 0 then pa()= 0 also. Similarly, using the identity 1 1 isin x 1 i (1.6) = − = − tan(x / 2) 1+ eix 2 2 1+ cos x 2 2 we may easily prove that 1 bb∞ (1.7) ∫∫p()xdx=−∑ px ()(1)cosn nxdx 2 aan=0 1 bb∞ (1.7a) ∫∫p()tan(/2)x x dx=−∑ p ()(1)sin xn+1 nx dx 2 aan=0 and more generally 1 bb∞ (1.8) ∫∫p()xdx=−∑ px ()(1)cosn α nxdx 2 aan=0 1 bb∞ (1.8a) ∫∫p()tan(xαα x /2) dx=−∑ p ()(1)sin x n+1 nx dx 2 aan=0 3 From (1.6) we note that the denominator is cos(x / 2) and hence (1.7) and (1.7a) are only valid provided either (i) cos(x / 2) has no zero in [ ab, ] or (ii) if cos(a / 2) = 0 , then p( a) must also be zero. Equations (1.8) and (1.8a) are valid provided (i) cos(α x / 2)≠ 0 ∀∈ xab[ , ] or, alternatively, (ii) if cos(αa / 2)= 0 then pa() = 0 also. The following simple trigonometric identities are easily proved 2 (1.9) cot(x / 2)+ tan( x / 2) = sin x (1.10) cot(x / 2)− tan( x / 2)= 2cot x Therefore, combining (1.5) and (1.7) produces the following identity bbpx() ∞∞ b (1.11) ∫∫dx=−−∑∑ p( x )sin nx dx ∫ p( x )( 1)n sin nx dx aasin x nn==00 a which simplifies to bbpx() ∞ (1.12) ∫∫dx=+2()sin(21)∑ p x n x dx aasin x n=0 Similarly, using (1.10) we obtain b ∞∞bb (1.13) ∫ p()cotxxdx=+∑∑∫∫p()sinxnxdx (1)()sin−n pxnxdx a nn==00aa which simplifies to bb∞ (1.14) ∫∫p()cotx x dx= 2∑ p ()sin2 x nx dx aan=1 It should be noted that in the above formulae we require either (i) both sin(x / 2) and cos(x / 2) have no zero in [ ab, ] or (ii) if either sin(a / 2) or cos(a / 2) is equal to zero then p( a) must also be zero. Condition (i) is equivalent to the requirement that sin x has no zero in [ ab, ]. Note that (1.14) is equivalent to (1.5a) with α = 2 . More generally we have 4 bbpx() ∞ (1.14a) ∫∫dx=+2()sin(21)∑ p x[] nα x dx aasinα x n=0 bb∞ (1.14b) ∫∫p()cotxαα x dx= 2∑ p ()sin2 x nx dx aan=1 Equations (1.14a) and (1.14b) are valid provided (i) sin(α x)≠ 0 ∀∈ xab[ , ] or, alternatively, (ii) if sin(αa )= 0 then pa( )= 0 also. We may also generalise Poisson’s integral [17, p.250] by applying the above analysis to 1 1 the quotient and useful results may also be obtained by using . 1± reix 1± ireix 2. Application of Dirichlet’s integrals In the previous section we required that p()x was twice continuously differentiable on the interval [,ab]; with the assistance of Dirichlet we now show that this condition may be significantly relaxed. Dirichlet [26] considered the following integrals in his classical treatment of Fourier series in 1829 (when he was then 23 years old) b sin(μt ) (2.1) limf (tdt ) = 0 μ→∞ ∫ a sin t b sin(μπt ) (2.2) limft ( ) dt= f(0+ ) μ→∞ ∫ 0 sint 2 where 0 < a < b < π . Dirichlet showed that these limits are valid if f ()x is continuous on [0,b ] and f ()x only has a finite number of maxima and minima on [0,b] . Jordan [37] subsequently showed in 1881 that the less restrictive condition that f ()x be of bounded variation on [0,]b is sufficient. Proofs may be found in [5, p.314] and [17, p.219]. We recall from [5, p.128] that if f ()x is monotonic on [ab,], then f ()x is of bounded variation on [ab,]. In addition, we also note that if f (x) is continuous on [ab,] and if f ′()x exists and is bounded on the open interval, then f ()x is of bounded variation on [,]ab. Even though Jordan wanted to relax the smoothness requirement for the pointwise convergence of Fourier series, it subsequently turned out that functions of bounded variation are in fact fairly smooth anyway; Lebesgue ([50, p.356 ] and [7, p.192]) proved 5 that any monotonic function on a closed interval is differentiable almost everywhere, and that its derivative is integrable. With the substitution t= α x in (2.1) and (2.2) we see that αb sin(μα x ) (2.3) limfx (α ) dx= 0 μ→∞ ∫ αa sinα x or equivalently b′ sin(μα x ) (2.4) limgx ( ) dx= 0 μ→∞ ∫ a′ sinα x where 0 < a′ < b′ < π . We may write (1.2) as 1sN in(2Nx+1)α (2.5) =−∑cos 2αnx 22n=0 sinα x and, after multiplying this by a function f ()x of bounded variation on [ab,], we easily see that 1sbbN bin(2Nx+1)α (2.6) ∫∫f ()xdx=−∑ f ()cos2 xα nxdx ∫ f () x dx 2 aan=0 a2sinα x Letting N →∞ we obtain 1 bb∞ (2.7) ∫∫f ()xdx= ∑ f ()cos2 xα nxdx 2 aan=0 provided 0 < αa < αb < π . This is a generalised version of (1.4) above and, in particular, it should be noted that we no longer require f ()x to be twice continuously differentiable on [,ab]. With α = π we may write (2.7) as 1 bb∞ (2.8) ∫∫f ()xdx= ∑ f ()cos2 xπ nxdx 2 aan=0 where 0 < a < b < 1. We now consider the case where (2.2) applies. With the substitution t= α x in (2.2) we see that 6 αb sin(μαx ) π (2.9) limααfx ( ) dxf= (0+ ) μ→∞ ∫ 0 sinα x 2 or equivalently b′ sin(μαx ) π (2.10) limgx ( ) dx= g(0+ ) μ→∞ ∫ 0 sinααx 2 where 0 < b′ < π . Letting N →∞ in (2.6) and using (2.10) we then obtain 1 bb∞ π ∫∫f() x dx=−∑ f ()cos2 xα nx dx f (0)+ 2400n=0 α where 0 < αb < π .
Recommended publications
  • Arxiv:2007.02955V4 [Quant-Ph] 7 Jan 2021
    Harvesting correlations in Schwarzschild and collapsing shell spacetimes Erickson Tjoa,a,b,1 Robert B. Manna,c aDepartment of Physics and Astronomy, University of Waterloo, Waterloo, Ontario, N2L 3G1, Canada bInstitute for Quantum Computing, University of Waterloo, Waterloo, Ontario, N2L 3G1, Canada cPerimeter Institute for Theoretical Physics, Waterloo, Ontario, N2L 2Y5, Canada E-mail: [email protected], [email protected] Abstract: We study the harvesting of correlations by two Unruh-DeWitt static detectors from the vacuum state of a massless scalar field in a background Vaidya spacetime con- sisting of a collapsing null shell that forms a Schwarzschild black hole (hereafter Vaidya spacetime for brevity), and we compare the results with those associated with the three pre- ferred vacua (Boulware, Unruh, Hartle-Hawking-Israel vacua) of the eternal Schwarzschild black hole spacetime. To do this we make use of the explicit Wightman functions for a massless scalar field available in (1+1)-dimensional models of the collapsing spacetime and Schwarzschild spacetimes, and the detectors couple to the proper time derivative of the field. First we find that, with respect to the harvesting protocol, the Unruh vacuum agrees very well with the Vaidya vacuum near the horizon even for finite-time interactions. Sec- ond, all four vacua have different capacities for creating correlations between the detectors, with the Vaidya vacuum interpolating between the Unruh vacuum near the horizon and the Boulware vacuum far from the horizon. Third, we show that the black hole horizon inhibits any correlations, not just entanglement. Finally, we show that the efficiency of the harvesting protocol depend strongly on the signalling ability of the detectors, which is highly non-trivial in presence of curvature.
    [Show full text]
  • Lecture 2 Complex Analysis
    Complex Analysis Lecture 2 Complex Analysis A Complex Numbers and Complex Variables In this chapter we give a short discussion of complex numbers and the theory of a function of a complex variable. Before we get to complex numbers, let us first say a few words about real numbers. All real numbers have meanings in the real world. Ever since the beginning of civilization, people have found great use of real positive integers such as 2 and 30, which came up in, as an example, the sentence “my neighbor has two pigs, and I have thirty chickens”. The concept of negative real integers, say ?5, is a bit more difficult, but found its use when a person owed another person five ounces of silver. It was also natural to extend the concept of integers to numbers which are not integers. For example, when six persons share equally a melon, the number describing the fraction of melon each of them has is not an integer but the rational number 1/6. The need for other real numbers was found as mathematicians pondered the length of the circumference of a perfectly circular hole with unit radius. This length is a real number which can be expressed neither as an integer nor as a ratio of two integers. This real number is denoted as Z, and is called an irrational number. Another example of irrational number is e. Each of the real numbers, be it integral, rational or irrational, can be geometrically represented by a point on an infinite line, and vice versa.
    [Show full text]
  • Dynamics of Contour Integration
    Vision Research 41 (2001) 1023–1037 www.elsevier.com/locate/visres Dynamics of contour integration Robert F. Hess *, William H.A. Beaudot, Kathy T. Mullen Department of Ophthalmology, McGill Vision Research, McGill Uni6ersity, 687 Pine A6enue West (H4-14), Montre´al, Que´bec, Canada H3A 1A1 Received 3 May 2000; received in revised form 6 October 2000 Abstract To determine the dynamics of contour integration the temporal properties of the individual contour elements were varied as well as those of the contour they form. A temporal version of a contour integration paradigm (Field, D. J., Hayes, A., & Hess, R. F. (1993) Vision Research, 33, 173–193) was used to assess these two temporal dynamics as a function of the contrast of individual elements and the curvature of the contour. The results show that the dynamics of contour integration are good when the contrast of the individual elements is modulated in time (10–30 Hz), but are poor when contour linking per se is temporally modulated (1–12 Hz). The dynamics of contour linking is not dependent on the absolute contrast of the linking elements, so long as they are visible, but does vary with the curvature of the contour. For straight contours, temporal resolution is around 6–12 Hz but falls to around 1–2 Hz for curved contours. © 2001 Elsevier Science Ltd. All rights reserved. Keywords: Contour integration; Temporal sensitivity; Dynamics; Curvature; Contrast 1. Introduction columns are interconnected across disparate regions of cortex through long-range connections. There is now There is a growing awareness of the importance of neurophysiological evidence that these network interac- network interactions within the visual cortex as op- tions are context-dependent (e.g.
    [Show full text]
  • Lightning Overview of Contour Integration Wednesday, ÕÉ October Óþõõ Physics ÕÕÕ
    Lightning Overview of Contour Integration Wednesday, ÕÉ October óþÕÕ Physics ÕÕÕ e prettiest math I know, and extremely useful. Õ. Overview is handout summarizes key results in the theory of functions of a complex variable, leading to the inverse Laplace transform. I make no pretense at rigor, but I’ll do my best to ensure that I don’t say anything false. e main points are: Õ. e Cauchy-Riemann conditions, which must be satised by a dierentiable function f z of a complex variable z. ó. f z dz þ if f z is continuous and dierentiable everywhere inside the closed path.( ) ì. If∮ f (z) blows= up inside( ) a closed contour in the complex plane, then the value of f z dz is óπi times the sum of the residues of all the poles inside the contour. ( ) ∞ ikx ¦. e Fourier representation of the delta function is −∞ e dk óπδ x . ¢. ∮ e( inverse) Laplace transform equation. ∫ = ( ) ó. Cauchy-Riemann For the derivative of a function of a complex variable to exist, the value of the derivative must be independent of the manner in which it is taken. If z x iy is a complex variable, which we may wish to represent on the x y plane, and f z u x, y iυ x, y , then the d f = + value of should be the same at a point z whether we approach the point along a line parallel dz ( ) = ( ) + ( ) to the x axis or a line parallel to the y axis. erefore, d f ∂u ∂υ i (along x) dz ∂x ∂y d f = ∂u+ ∂υ ∂υ ∂u i i (along y) dz ∂ iy ∂ iy ∂y ∂y = + = − Comparing real and imaginary parts,( we) obtain( the) Cauchy-Riemann equations: ∂u ∂υ ∂υ ∂u and (Õ) ∂x ∂y ∂x ∂y = = − ì.
    [Show full text]
  • Math 259: Introduction to Analytic Number Theory the Contour Integral Formula for Ψ(X)
    Math 259: Introduction to Analytic Number Theory The contour integral formula for (x) We now have several examples of Dirichlet series, that is, series of the form1 1 F (s) = a n s (1) X n − n=1 from which we want to extract information about the growth of n<x an as x . The key to this is a contour integral. We regard F (s) asP a function of!1 a complex variable s = σ + it. For real y > 0 we have seen already that s σ 2 y− = y− . Thus if the sum (1) converges absolutely for some real σ0, then itj convergesj uniformly and absolutely to an analytic function on the half-plane Re(s) σ0; and if the sum converges absolutely for all real s > σ0, then it ≥ converges absolutely to an analytic function on the half-plane Re(s) > σ0. Now for y > 0 and c > 0 we have 1; if y > 1; 1 c+i ds 8 1 ys = 1 (2) Z <> 2 ; if y = 1; 2πi c i s − 1 0; if y < 1; :> in the following sense: the contour of integration is the vertical line Re(s) = c, and since the integral is then not absolutely convergent it is regarded as a principal value: c+i c+iT 1 Z f(s) ds := lim Z f(s) ds: c i T c iT − 1 !1 − Thus interpreted, (2) is an easy exercise in contour integration for y = 1, and an elementary manipulation of log s for y = 1. So we expect that if (1)6 converges absolutely in Re(s) > σ0 then c+i 1 1 s ds an = Z x F (s) (3) X 2πi c i s n<x − 1 for any c > σ0, using the principal value of the integral and adding ax=2 to the sum if x happens to be an integer.
    [Show full text]
  • Chapter 2 Complex Analysis
    Chapter 2 Complex Analysis In this part of the course we will study some basic complex analysis. This is an extremely useful and beautiful part of mathematics and forms the basis of many techniques employed in many branches of mathematics and physics. We will extend the notions of derivatives and integrals, familiar from calculus, to the case of complex functions of a complex variable. In so doing we will come across analytic functions, which form the centerpiece of this part of the course. In fact, to a large extent complex analysis is the study of analytic functions. After a brief review of complex numbers as points in the complex plane, we will ¯rst discuss analyticity and give plenty of examples of analytic functions. We will then discuss complex integration, culminating with the generalised Cauchy Integral Formula, and some of its applications. We then go on to discuss the power series representations of analytic functions and the residue calculus, which will allow us to compute many real integrals and in¯nite sums very easily via complex integration. 2.1 Analytic functions In this section we will study complex functions of a complex variable. We will see that di®erentiability of such a function is a non-trivial property, giving rise to the concept of an analytic function. We will then study many examples of analytic functions. In fact, the construction of analytic functions will form a basic leitmotif for this part of the course. 2.1.1 The complex plane We already discussed complex numbers briefly in Section 1.3.5.
    [Show full text]
  • 9.1 Contour Integral
    THE CHINESE UNIVERSITY OF HONG KONG DEPARTMENT OF MATHEMATICS MATH2230A (First term, 2015{2016) Complex Variables and Applications Notes 9 Contour Integration 9.1 Contour Integral Previously, we have learned how to find the harmonic conjugate of a given harmonic function u(x; y). In other words, we would like to find v(x; y) such that u+iv defines an analytic function. The process is to use the Cauchy-Riemann Equations successively. For example, we start with vy = ux = this is known as u is given; and obtain Z v = ux dy + '(x) : 0 Then, we use vx = −uy to get information about ' (x). Next, integrating wrt x will get the result. Observing from both analytically and geometrically, we see that the process corresponds to an integration along the vertical and then horizontal line in the picture below. Sometimes, in order to avoid holes in the domain Ω where u is defined, we may need to take more vertical and horizontal lines (dotted in picture). Nevertheless, one may expect the general theory should be related to integration along a curve. 9.1.1 Contours and Integrals Let us start with the building block of a contour. A smooth path γ in Ω ⊂ C is a continuous function γ :[a; b] ! Ω from an interval with γ(t) = x(t) + iy(t) such that x0(t) and y0(t) exist and are continuous for all t 2 (a; b). The function γ is called a parametrization of the path. It determines a \direction" for the path. We denote γ0(t) = x0(t) + iy0(t).
    [Show full text]
  • Contour Integration with Vectors
    Integrating Functions Over a Line (or Contour) Example: When dry, a climbing rope weighs 0.175 #/foot. The weight density for a wet rope hanging vertically is: wh()=+0.175 1 0.7 e−0.005h # per foot 0.30 0.28 0.26 Find the weight of a wet 2000’ 0.24 rope hanging vertically 0.22 0.20 Note: a dry rope 2000’ long would weigh 0.18 2000 feet x 0.175 #/foot = 350 # 0.16 Weight Density (pounds/foot)Weight Density 0 500 1000 1500 2000 Height Above Ground (feet) The weight of a wet rope is: h 2000 Total weight =∫∫whdh( )=+ 0.175() 1 0.7 e−.005h dh 00 2000 =−0.175he 140−.005h = 0.175() 2000 −+= 0.0064 140 374.5# ()0 Contour Integration with Vectors Example: if the force acting on a particle is Fa=−ˆy Newtons, how much work is required to move the particle from (0,0) to (1,1) along the path yx= ? F y W = work = force times distance, where (1,1) force is the force tangent to the path. Path, or Contour x (0,0) In this example, the angle between the path and the force is 45° over the entire path: FFTAN =°cos 45 The length of the path is 2 ⇒=°=work 2 cos 45 1 Nm To handle more general cases, where the vector or path varies, we will need to integrate. To do this, define a vector differential element, dl 1 Differential Length Vector dl is an infintesimally-short vector tangent to a curve or li ne In general, for Cartesian coordinates dl dl ddxadyadzal =++ˆˆˆx yz This can be put in terms of dx, dy, or dz only to facilitate integration.
    [Show full text]
  • CONTOUR INTEGRALS LECTURE Recall the Residue Theorem
    MATH 311: COMPLEX ANALYSIS | CONTOUR INTEGRALS LECTURE Recall the Residue Theorem: Let γ be a simple closed loop, traversed counter- clockwise. Let f be a function that is analytic on γ and meromorphic inside γ. Then Z X f(z) dz = 2πi Resc(f): γ c inside γ This writeup shows how the Residue Theorem can be applied to integrals that arise with no reference to complex analysis. 1. Computing Residues Proposition 1.1. Let f have a simple pole at c. Then Resc(f) = lim(z − c)f(z): z!c Proposition 1.2. Let f have a pole of order n ≥ 1 at c. Define a modified function g(z) = (z − c)nf(z). Then 1 (n−1) Resc(f) = lim g (z): (n − 1)! z!c Proof. It suffices to prove the second proposition, since it subsumes the first. Recall that the residue is the the coefficient a−1 of 1=(z − c) in the Laurent series. The proof is merely a matter of inspection: a a f(z) = −n + ··· + −1 + a + ··· ; (z − c)n z − c 0 and so the modified function g is n−1 n g(z) = a−n + ··· + a−1(z − c) + a0(z − c) + ··· ; whose (n − 1)st derivative is (n−1) g (z) = (n − 1)! a−1 + n(n − 1) ··· 2 a0(z − c) + ··· : Thus (n−1) (n−1) lim g (z) = g (c) = (n − 1)! a−1: z!c This is the desired result. In applying the propositions, we do not go through these calculations again every time. L'Hospital's Rule will let us take the limits without computing the Laurent series.
    [Show full text]
  • A Summary of Contour Integration Vs 0.2, 25/1/08 © Niels Walet, University of Manchester
    A Summary of Contour Integration vs 0.2, 25/1/08 © Niels Walet, University of Manchester In this document I will summarise some very simple results of contour integration 1. The Basics The key word linked to contour integration is "analyticity" or the absence thereof: A function is called analytic in a region R in the complex plane iff all the derivatives of the function (1st, 2nd, ....) exist for every point inside R. This means that the Taylor series H Ln f HzL = Ú¥ f HnLHcL z-c (1) n=0 n! exists for every point c inside R. In most cases we are actually interested in functions that are not analytic; if this only happens at isolated points (i.e., we don't consider a "line of singularities", usually called a "cut" or "branch-cut") we can still expand the function in a Laurent series H Ln f HzL = Ú¥ f HnLHcL z-c (2) n=-¥ n! How we obtain the coefficients f HnLHcL is closely linked to the problem of contour integration. 2. Contour Integration Let us look at the effects of integrating the powers of z along a line in the complex plane (note that we implicitely assume that the answer is independent of the position of the line, and only depends on beginning and end!) z 1 n à z âz, n Î Z. (3) z0 We know how to integrate powers, so apart from the case n = -1, we get z z1 1 1 zn âz = B zn+1F , n ¹ -1, à + z0 n 1 z0 (4) z1 à z-1 âz = @log zDz1 , n = -1.
    [Show full text]
  • Math 3964 - Complex Analysis
    MATH 3964 - COMPLEX ANALYSIS ANDREW TULLOCH AND GILES GARDAM Contents 1. Contour Integration and Cauchy’s Theorem 2 1.1. Analytic functions 2 1.2. Contour integration 3 1.3. Cauchy’s theorem and extensions 3 1.4. Cauchy’s integral formula 3 1.5. The Cauchy-Taylor theorem and analytic continuation 4 1.6. Laurent’s theorem and the residue theorem 5 1.7. The Gamma function Γ(z) 6 1.8. The residue theorem 7 2. Analytic Theory of Differential Equations 8 2.1. Existence and uniqueness 8 2.2. Singular point analysis of nonlinear differential equations 8 2.3. Painlevé transcendents 9 2.4. Fuchsian theory 10 2.5. Hypergeometric functions 11 2.6. Gauss’ formula at z = 1 12 2.7. Kummer’s formula at z = −1 12 1 2.8. Evaluation at z = 2 12 2.9. Connection formulae 12 2.10. Elliptic functions 12 2.11. Elliptic integrals 13 2.12. Inversion of elliptic integrals 14 2.13. Doubly periodic meromorphic functions 14 2.14. Jacobi elliptic functions 15 2.15. Addition theorems 15 2.16. Liouville theory 15 1 MATH 3964 - COMPLEX ANALYSIS 2 2.17. Weierstrass’ Theorem 15 MATH 3964 - COMPLEX ANALYSIS 3 1. Contour Integration and Cauchy’s Theorem 1.1. Analytic functions. Definition 1.1. A function f(z) is differentiable at z0 if the limit 0 f(ζ) − f(z0) f (z0) = lim ! ζ z0 ζ − z0 exists independently of the path of approach. Definition 1.2. A function f(z) is analytic on a region D if it is differentiable everywhere on D.
    [Show full text]
  • INTEGRATION: the FEYNMAN WAY 1. Introduction Many Up-And
    INTEGRATION: THE FEYNMAN WAY ANONYMOUS Abstract. In this paper we will learn a common technique not often de­ scribed in collegiate calculus courses. After reviewing the necessary theory, we will proceed to work through some typical examples. Throughout this pro­ cess, we will see trivial integrals that can be evaluated using basic techniques of integration (such as integration by parts), however we will also encounter inte­ grals that would otherwise require more advanced techniques such as contour integration. 1. Introduction Many up-and-coming mathematicians, before every reaching the university level, heard about a certain method for evaluating definite integrals from the following passage in [1]: One thing I never did learn was contour integration. I had learned to do integrals by various methods show in a book that my high school physics teacher Mr. Bader had given me. The book also showed how to differentiate parameters under the integral sign - It’s a certain operation. It turns out that’s not taught very much in the universities; they don’t emphasize it. But I caught on how to use that method, and I used that one damn tool again and again. So because I was self-taught using that book, I had peculiar methods of doing integrals. The result was that, when guys at MIT or Princeton had trouble doing a certain integral, it was because they couldn’t do it with the standard methods they had learned in school. If it was contour integration, they would have found it; if it was a simple series expansion, they would have found it.
    [Show full text]