Chapter 2 Complex Analysis

Total Page:16

File Type:pdf, Size:1020Kb

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. The emphasis in that section was on the algebraic properties of complex numbers, and 73 although these properties are of course important here as well and will be used all the time, we are now also interested in more geometric properties of the complex numbers. The set C of complex numbers is naturally identi¯ed with the plane R2. This is often called the Argand plane. Given a complex number z = x+i y, its real and imag- 6 2 inary parts de¯ne an element (x; y) of R , as shown in z = x + iy y the ¯gure. In fact this identi¯cation is one of real vec- 7 tor spaces, in the sense that adding complex numbers and multiplying them with real scalars mimic the simi- - lar operations one can do in R2. Indeed, if ® 2 R is real, x then to ® z = (® x) + i (® y) there corresponds the pair (® x; ® y) = ® (x; y). Similarly, if z1 = x1 + i y1 and z2 = x2 + i y2 are com- plex numbers, then z1 + z2 = (x1 + x2) + i (y1 + y2), whose associated pair is (x1 + x2; y1 + y2) = (x1; y1) + (x2; y2). In fact, the identi¯cation is even one of euclidean spaces. Given a complexp number z = x + i y, its modulus jzj, de¯ned by jzj2 = zz¤, is given by x2 + y2 which is precisely the norm k(x; y)k of the pair (x; y). Similarly, if z1 = x1 + i y1 and z2 = x2 + i y2, ¤ then Re(z1z2) = x1x2 + y1y2 which is the dot product of the pairs (x1; y1) and (x2; y2). In particular, it follows from these remarks and the triangle inequality for the norm in R2, that complex numbers obey a version of the triangle inequality: jz1 + z2j · jz1j + jz2j : (2.1) Polar form and the argument function Points in the plane can also be represented using polar coordinates, and this representation in turn translates into a representation of the complex numbers. iθ Let (x; y) be a point in the plane. If we de¯ne r = z = re p 2 2 r 7 x + y and θ by θ = arctan(y=x), then we can write θ (x; y) = (r cos θ; r sin θ) = r (cos θ; sin θ). The complex number z = x + i y can then be written as z = r (cos θ + i sin θ). The real number r, as we have seen, is the modulus jzj of z, and the complex number cos θ + i sin θ has unit modulus. Comparing the Taylor series for the cosine and sine functions and the exponential functions we notice that cos θ+i sin θ = eiθ. The angle θ is called the argument of z and is written arg(z). Therefore we 74 have the following polar form for a complex number z: z = jzj ei arg(z) : (2.2) Being an angle, the argument of a complex number is only de¯ned up to the addition of integer multiples of 2¼. In other words, it is a multiple-valued function. This ambiguity can be resolved by de¯ning the principal value Arg of the arg function to take values in the interval (¡¼; ¼]; that is, for any complex number z, one has ¡¼ < Arg(z) · ¼ : (2.3) Notice, however, that Arg is not a continuous function: it has a discontinuity along the negative real axis. Approaching a point on the negative real axis from the upper half-plane, the principal value of its argument approaches ¼, whereas if we approach it from the lower half-plane, the principal value of its argument approaches ¡¼. Notice ¯nally that whereas the modulus is a multiplicative function: jzwj = jzjjwj, the argument is additive: arg(z1 z2) = arg(z1) + arg(z2), provided that we understand the equation to hold up to integer multiples of 2¼. Also notice that whereas the modulus is invariant under conjugation jz¤j = jzj, the argument changes sign arg(z¤) = ¡ arg(z), again up to integer multiples of 2¼. Some important subsets of the complex plane We end this section with a brief discussion of some very important subsets of the complex plane. Let z0 be any complex number, and consider all those complex numbers z which are a distance at most " away from z0. These points form a disk of radius " centred at z0. More precisely, let us de¯ne the open "-disk around z0 to be the subset D"(z0) of the complex plane de¯ned by D"(z0) = fz 2 C j jz ¡ z0j < "g : (2.4) Similarly one de¯nes the closed "-disk around z0 to be the subset ¹ D"(z0) = fz 2 C j jz ¡ z0j · "g ; (2.5) which consists of the open "-disk and the circle jz ¡ z0j = " which forms its boundary. More generally a subset U ½ C of the complex plane is said to be open if given any z 2 U, there exists some positive real number " > 0 (which can depend on z) such that the open "-disk around z also belongs to U. A set C is said to be closed if its complement Cc = fz 2 C j z 62 Cg|that is, all 75 those points not in C|is open. One should keep in mind that generic subsets of the complex plane are neither closed nor open. By a neighbourhood of a point z0 in the complex plane, we will mean any open set containing z0. For example, any open "-disk around z0 is a neighbourhood of z0. Let us see that the open and closed "-disks are indeed open and closed, respectively. Let z 2 D"(z0). This means that jz ¡ z0j = ± < ". Consider the disk D"¡±(z). We claim that this disk is contained in D"(z0). Indeed, if jw ¡ zj < " ¡ ± then, jw ¡ z0j = j(w ¡ z) + (z ¡ z0)j (adding and subtracting z) · jw ¡ zj + jz ¡ z0j (by the triangle inequality (2.1)) < " ¡ ± + ± = ": Therefore the disk D"(z0) is indeed open. Consider now the subset D¹"(z0). Its complement is the subset of points z in the complex plane such that jz ¡ z0j > ". We will show that it is an open set. Let z be such that jz ¡ z0j = ´ > ". Then consider the open disk D´¡"(z), and let w be a point in it. Then jz ¡ z0j = j(z ¡ w) + (w ¡ z0)j (adding and subtracting w) · jz ¡ wj + jw ¡ z0j : (by the triangle inequality (2.1)) We can rewrite this as jw ¡ z0j ¸ jz ¡ z0j ¡ jz ¡ wj > ´ ¡ (´ ¡ ") (since jz ¡ wj = jw ¡ zj < ´ ¡ ") = ": Therefore the complement of D¹"(z0) is open, whence D¹"(z0) is closed. We should remark that the closed disk D¹"(z0) is not open, since any open disk around a point z at the boundary of D¹"(z0)|that is, for which jz ¡ z0j = "|contains points which are not included in D"(z0). Notice that it follows from this de¯nition that every open set is made out of the union of (a possibly uncountable number of) open disks. 2.1.2 Complex-valued functions In this section we will discuss complex-valued functions. We start with a rather trivial case of a complex-valued function. Suppose that f is a complex-valued function of a real variable. That means that if x is a real number, f(x) is a complex number, which can be decomposed into its real and imaginary parts: f(x) = u(x)+i v(x), where u and v are real-valued functions of a real variable; that is, the objects you are familiar with from calculus. We say that f is continuous at x0 if u and v are continuous at x0. Let us recall the de¯nition of continuity. Let f be a real-valued function of a real variable. We say that f is continuous at x0, if for every " > 0, there is a ± > 0 such that jf(x) ¡ f(x0)j < " whenever jx ¡ x0j < ±. A function is said to be continuous if it is continuous at all points where it is de¯ned. 76 Now consider a complex-valued function f of a complex variable z.
Recommended publications
  • Best Subordinant for Differential Superordinations of Harmonic Complex-Valued Functions
    mathematics Article Best Subordinant for Differential Superordinations of Harmonic Complex-Valued Functions Georgia Irina Oros Department of Mathematics and Computer Sciences, Faculty of Informatics and Sciences, University of Oradea, 410087 Oradea, Romania; [email protected] or [email protected] Received: 17 September 2020; Accepted: 11 November 2020; Published: 16 November 2020 Abstract: The theory of differential subordinations has been extended from the analytic functions to the harmonic complex-valued functions in 2015. In a recent paper published in 2019, the authors have considered the dual problem of the differential subordination for the harmonic complex-valued functions and have defined the differential superordination for harmonic complex-valued functions. Finding the best subordinant of a differential superordination is among the main purposes in this research subject. In this article, conditions for a harmonic complex-valued function p to be the best subordinant of a differential superordination for harmonic complex-valued functions are given. Examples are also provided to show how the theoretical findings can be used and also to prove the connection with the results obtained in 2015. Keywords: differential subordination; differential superordination; harmonic function; analytic function; subordinant; best subordinant MSC: 30C80; 30C45 1. Introduction and Preliminaries Since Miller and Mocanu [1] (see also [2]) introduced the theory of differential subordination, this theory has inspired many researchers to produce a number of analogous notions, which are extended even to non-analytic functions, such as strong differential subordination and superordination, differential subordination for non-analytic functions, fuzzy differential subordination and fuzzy differential superordination. The notion of differential subordination was adapted to fit the harmonic complex-valued functions in the paper published by S.
    [Show full text]
  • Integration in the Complex Plane (Zill & Wright Chapter
    Integration in the Complex Plane (Zill & Wright Chapter 18) 1016-420-02: Complex Variables∗ Winter 2012-2013 Contents 1 Contour Integrals 2 1.1 Definition and Properties . 2 1.2 Evaluation . 3 1.2.1 Example: R z¯ dz ............................. 3 C1 1.2.2 Example: R z¯ dz ............................. 4 C2 R 2 1.2.3 Example: C z dz ............................. 4 1.3 The ML Limit . 5 1.4 Circulation and Flux . 5 2 The Cauchy-Goursat Theorem 7 2.1 Integral Around a Closed Loop . 7 2.2 Independence of Path for Analytic Functions . 8 2.3 Deformation of Closed Contours . 9 2.4 The Antiderivative . 10 3 Cauchy's Integral Formulas 12 3.1 Cauchy's Integral Formula . 12 3.1.1 Example #1 . 13 3.1.2 Example #2 . 13 3.2 Cauchy's Integral Formula for Derivatives . 14 3.3 Consequences of Cauchy's Integral Formulas . 16 3.3.1 Cauchy's Inequality . 16 3.3.2 Liouville's Theorem . 16 ∗Copyright 2013, John T. Whelan, and all that 1 Tuesday 18 December 2012 1 Contour Integrals 1.1 Definition and Properties Recall the definition of the definite integral Z xF X f(x) dx = lim f(xk) ∆xk (1.1) ∆xk!0 xI k We'd like to define a similar concept, integrating a function f(z) from some point zI to another point zF . The problem is that, since zI and zF are points in the complex plane, there are different ways to get between them, and adding up the value of the function along one path will not give the same result as doing it along another path, even if they have the same endpoints.
    [Show full text]
  • Class 1/28 1 Zeros of an Analytic Function
    Math 752 Spring 2011 Class 1/28 1 Zeros of an analytic function Towards the fundamental theorem of algebra and its statement for analytic functions. Definition 1. Let f : G → C be analytic and f(a) = 0. a is said to have multiplicity m ≥ 1 if there exists an analytic function g : G → C with g(a) 6= 0 so that f(z) = (z − a)mg(z). Definition 2. If f is analytic in C it is called entire. An entire function has a power series expansion with infinite radius of convergence. Theorem 1 (Liouville’s Theorem). If f is a bounded entire function then f is constant. 0 Proof. Assume |f(z)| ≤ M for all z ∈ C. Use Cauchy’s estimate for f to obtain that |f 0(z)| ≤ M/R for every R > 0 and hence equal to 0. Theorem 2 (Fundamental theorem of algebra). For every non-constant polynomial there exists a ∈ C with p(a) = 0. Proof. Two facts: If p has degree ≥ 1 then lim p(z) = ∞ z→∞ where the limit is taken along any path to ∞ in C∞. (Sometimes also written as |z| → ∞.) If p has no zero, its reciprocal is therefore entire and bounded. Invoke Liouville’s theorem. Corollary 1. If p is a polynomial with zeros aj (multiplicity kj) then p(z) = k k km c(z − a1) 1 (z − a2) 2 ...(z − am) . Proof. Induction, and the fact that p(z)/(z − a) is a polynomial of degree n − 1 if p(a) = 0. 1 The zero function is the only analytic function that has a zero of infinite order.
    [Show full text]
  • Global Subanalytic Cmc Surfaces
    GLOBALLY SUBANALYTIC CMC SURFACES IN R3 WITH SINGULARITIES JOSE´ EDSON SAMPAIO Abstract. In this paper we present a classification of a class of globally sub- 3 analytic CMC surfaces in R that generalizes the recent classification made by Barbosa and do Carmo in 2016. We show that a globally subanalytic CMC 3 surface in R with isolated singularities and a suitable condition of local con- nectedness is a plane or a finite union of round spheres and right circular cylinders touching at the singularities. As a consequence, we obtain that a 3 globally subanalytic CMC surface in R that is a topological manifold does not have isolated singularities. It is also proved that a connected closed glob- 3 ally subanalytic CMC surface in R with isolated singularities which is locally Lipschitz normally embedded needs to be a plane or a round sphere or a right circular cylinder. A result in the case of non-isolated singularities is also pre- sented. It is also presented some results on regularity of semialgebraic sets and, in particular, it is proved a real version of Mumford's Theorem on regu- larity of normal complex analytic surfaces and a result about C1 regularity of minimal varieties. 1. Introduction The question of describing minimal surfaces or, more generally, surfaces of con- stant mean curvature (CMC surfaces) is known in Analysis and Differential Geom- etry since the classical papers of Bernstein [4], Bombieri, De Giorgi and Giusti [9], Hopf [26] and Alexandrov [1]. Recently, in the paper [2], Barbosa and do Carmo showed that the connected algebraic smooth CMC surfaces in R3 are only the planes, round spheres and right circular cylinders.
    [Show full text]
  • A Property of the Derivative of an Entire Function
    A property of the derivative of an entire function Walter Bergweiler∗ and Alexandre Eremenko† July 21, 2011 Abstract We prove that the derivative of a non-linear entire function is un- bounded on the preimage of an unbounded set. MSC 2010: 30D30. Keywords: entire function, normal family. 1 Introduction and results The main result of this paper is the following theorem conjectured by Allen Weitsman (private communication): Theorem 1. Let f be a non-linear entire function and M an unbounded set in C. Then f ′(f −1(M)) is unbounded. We note that there exist entire functions f such that f ′(f −1(M)) is bounded for every bounded set M, for example, f(z)= ez or f(z) = cos z. Theorem 1 is a consequence of the following stronger result: Theorem 2. Let f be a transcendental entire function and ε > 0. Then there exists R> 0 such that for every w C satisfying w >R there exists ∈ | | z C with f(z)= w and f ′(z) w 1−ε. ∈ | | ≥ | | ∗Supported by the Deutsche Forschungsgemeinschaft, Be 1508/7-1, and the ESF Net- working Programme HCAA. †Supported by NSF grant DMS-1067886. 1 The example f(z)= √z sin √z shows that that the exponent 1 ε in the − last inequality cannot be replaced by 1. The function f(z) = cos √z has the property that for every w C we have f ′(z) 0 as z , z f −1(w). ∈ → → ∞ ∈ We note that the Wiman–Valiron theory [20, 12, 4] says that there exists a set F [1, ) of finite logarithmic measure such that if ⊂ ∞ zr = r / F and f(zr) = max f(z) , | | ∈ | | |z|=r | | then ν(r,f) z ′ ν(r, f) f(z) f(zr) and f (z) f(z) ∼ zr ∼ r −1/2−δ for z zr rν(r, f) as r .
    [Show full text]
  • Informal Lecture Notes for Complex Analysis
    Informal lecture notes for complex analysis Robert Neel I'll assume you're familiar with the review of complex numbers and their algebra as contained in Appendix G of Stewart's book, so we'll pick up where that leaves off. 1 Elementary complex functions In one-variable real calculus, we have a collection of basic functions, like poly- nomials, rational functions, the exponential and log functions, and the trig functions, which we understand well and which serve as the building blocks for more general functions. The same is true in one complex variable; in fact, the real functions we just listed can be extended to complex functions. 1.1 Polynomials and rational functions We start with polynomials and rational functions. We know how to multiply and add complex numbers, and thus we understand polynomial functions. To be specific, a degree n polynomial, for some non-negative integer n, is a function of the form n n−1 f(z) = cnz + cn−1z + ··· + c1z + c0; 3 where the ci are complex numbers with cn 6= 0. For example, f(z) = 2z + (1 − i)z + 2i is a degree three (complex) polynomial. Polynomials are clearly defined on all of C. A rational function is the quotient of two polynomials, and it is defined everywhere where the denominator is non-zero. z2+1 Example: The function f(z) = z2−1 is a rational function. The denomina- tor will be zero precisely when z2 = 1. We know that every non-zero complex number has n distinct nth roots, and thus there will be two points at which the denominator is zero.
    [Show full text]
  • Lecture 5: Complex Logarithm and Trigonometric Functions
    LECTURE 5: COMPLEX LOGARITHM AND TRIGONOMETRIC FUNCTIONS Let C∗ = C \{0}. Recall that exp : C → C∗ is surjective (onto), that is, given w ∈ C∗ with w = ρ(cos φ + i sin φ), ρ = |w|, φ = Arg w we have ez = w where z = ln ρ + iφ (ln stands for the real log) Since exponential is not injective (one one) it does not make sense to talk about the inverse of this function. However, we also know that exp : H → C∗ is bijective. So, what is the inverse of this function? Well, that is the logarithm. We start with a general definition Definition 1. For z ∈ C∗ we define log z = ln |z| + i argz. Here ln |z| stands for the real logarithm of |z|. Since argz = Argz + 2kπ, k ∈ Z it follows that log z is not well defined as a function (it is multivalued), which is something we find difficult to handle. It is time for another definition. Definition 2. For z ∈ C∗ the principal value of the logarithm is defined as Log z = ln |z| + i Argz. Thus the connection between the two definitions is Log z + 2kπ = log z for some k ∈ Z. Also note that Log : C∗ → H is well defined (now it is single valued). Remark: We have the following observations to make, (1) If z 6= 0 then eLog z = eln |z|+i Argz = z (What about Log (ez)?). (2) Suppose x is a positive real number then Log x = ln x + i Argx = ln x (for positive real numbers we do not get anything new).
    [Show full text]
  • MATH 305 Complex Analysis, Spring 2016 Using Residues to Evaluate Improper Integrals Worksheet for Sections 78 and 79
    MATH 305 Complex Analysis, Spring 2016 Using Residues to Evaluate Improper Integrals Worksheet for Sections 78 and 79 One of the interesting applications of Cauchy's Residue Theorem is to find exact values of real improper integrals. The idea is to integrate a complex rational function around a closed contour C that can be arbitrarily large. As the size of the contour becomes infinite, the piece in the complex plane (typically an arc of a circle) contributes 0 to the integral, while the part remaining covers the entire real axis (e.g., an improper integral from −∞ to 1). An Example Let us use residues to derive the formula p Z 1 x2 2 π 4 dx = : (1) 0 x + 1 4 Note the somewhat surprising appearance of π for the value of this integral. z2 First, let f(z) = and let C = L + C be the contour that consists of the line segment L z4 + 1 R R R on the real axis from −R to R, followed by the semi-circle CR of radius R traversed CCW (see figure below). Note that C is a positively oriented, simple, closed contour. We will assume that R > 1. Next, notice that f(z) has two singular points (simple poles) inside C. Call them z0 and z1, as shown in the figure. By Cauchy's Residue Theorem. we have I f(z) dz = 2πi Res f(z) + Res f(z) C z=z0 z=z1 On the other hand, we can parametrize the line segment LR by z = x; −R ≤ x ≤ R, so that I Z R x2 Z z2 f(z) dz = 4 dx + 4 dz; C −R x + 1 CR z + 1 since C = LR + CR.
    [Show full text]
  • Harmonic Functions
    Lecture 1 Harmonic Functions 1.1 The Definition Definition 1.1. Let Ω denote an open set in R3. A real valued function u(x, y, z) on Ω with continuous second partials is said to be harmonic if and only if the Laplacian ∆u = 0 identically on Ω. Note that the Laplacian ∆u is defined by ∂2u ∂2u ∂2u ∆u = + + . ∂x2 ∂y2 ∂z2 We can make a similar definition for an open set Ω in R2.Inthatcase, u is harmonic if and only if ∂2u ∂2u ∆u = + =0 ∂x2 ∂y2 on Ω. Some basic examples of harmonic functions are 2 2 2 3 u = x + y 2z , Ω=R , − 1 3 u = , Ω=R (0, 0, 0), r − where r = x2 + y2 + z2. Moreover, by a theorem on complex variables, the real part of an analytic function on an open set Ω in 2 is always harmonic. p R Thus a function such as u = rn cos nθ is a harmonic function on R2 since u is the real part of zn. 1 2 1.2 The Maximum Principle The basic result about harmonic functions is called the maximum principle. What the maximum principle says is this: if u is a harmonic function on Ω, and B is a closed and bounded region contained in Ω, then the max (and min) of u on B is always assumed on the boundary of B. Recall that since u is necessarily continuous on Ω, an absolute max and min on B are assumed. The max and min can also be assumed inside B, but a harmonic function cannot have any local extrema inside B.
    [Show full text]
  • Niobrara County School District #1 Curriculum Guide
    NIOBRARA COUNTY SCHOOL DISTRICT #1 CURRICULUM GUIDE th SUBJECT: Math Algebra II TIMELINE: 4 quarter Domain: Student Friendly Level of Resource Academic Standard: Learning Objective Thinking Correlation/Exemplar Vocabulary [Assessment] [Mathematical Practices] Domain: Perform operations on matrices and use matrices in applications. Standards: N-VM.6.Use matrices to I can represent and Application Pearson Alg. II Textbook: Matrix represent and manipulated data, e.g., manipulate data to represent --Lesson 12-2 p. 772 (Matrices) to represent payoffs or incidence data. M --Concept Byte 12-2 p. relationships in a network. 780 Data --Lesson 12-5 p. 801 [Assessment]: --Concept Byte 12-2 p. 780 [Mathematical Practices]: Niobrara County School District #1, Fall 2012 Page 1 NIOBRARA COUNTY SCHOOL DISTRICT #1 CURRICULUM GUIDE th SUBJECT: Math Algebra II TIMELINE: 4 quarter Domain: Student Friendly Level of Resource Academic Standard: Learning Objective Thinking Correlation/Exemplar Vocabulary [Assessment] [Mathematical Practices] Domain: Perform operations on matrices and use matrices in applications. Standards: N-VM.7. Multiply matrices I can multiply matrices by Application Pearson Alg. II Textbook: Scalars by scalars to produce new matrices, scalars to produce new --Lesson 12-2 p. 772 e.g., as when all of the payoffs in a matrices. M --Lesson 12-5 p. 801 game are doubled. [Assessment]: [Mathematical Practices]: Domain: Perform operations on matrices and use matrices in applications. Standards:N-VM.8.Add, subtract and I can perform operations on Knowledge Pearson Alg. II Common Matrix multiply matrices of appropriate matrices and reiterate the Core Textbook: Dimensions dimensions. limitations on matrix --Lesson 12-1p. 764 [Assessment]: dimensions.
    [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]
  • 1 the Complex Plane
    Math 135A, Winter 2012 Complex numbers 1 The complex numbers C are important in just about every branch of mathematics. These notes present some basic facts about them. 1 The Complex Plane A complex number z is given by a pair of real numbers x and y and is written in the form z = x+iy, where i satisfies i2 = −1. The complex numbers may be represented as points in the plane, with the real number 1 represented by the point (1; 0), and the complex number i represented by the point (0; 1). The x-axis is called the \real axis," and the y-axis is called the \imaginary axis." For example, the complex numbers 1, i, 3 + 4i and 3 − 4i are illustrated in Fig 1a. 3 + 4i 6 + 4i 2 + 3i imag i 4 + i 1 real 3 − 4i Fig 1a Fig 1b Complex numbers are added in a natural way: If z1 = x1 + iy1 and z2 = x2 + iy2, then z1 + z2 = (x1 + x2) + i(y1 + y2) (1) It's just vector addition. Fig 1b illustrates the addition (4 + i) + (2 + 3i) = (6 + 4i). Multiplication is given by z1z2 = (x1x2 − y1y2) + i(x1y2 + x2y1) Note that the product behaves exactly like the product of any two algebraic expressions, keeping in mind that i2 = −1. Thus, (2 + i)(−2 + 4i) = 2(−2) + 8i − 2i + 4i2 = −8 + 6i We call x the real part of z and y the imaginary part, and we write x = Re z, y = Im z.(Remember: Im z is a real number.) The term \imaginary" is a historical holdover; it took mathematicians some time to accept the fact that i (for \imaginary," naturally) was a perfectly good mathematical object.
    [Show full text]