Princeton University—Physics 205 Notes on Approximations

Total Page:16

File Type:pdf, Size:1020Kb

Princeton University—Physics 205 Notes on Approximations Princeton University—Physics 205 Notes on Approximations Introduction One aspect of the course that many students find particularly vexing is the need to make approximations. Unlike freshman physics, where problems are simple enough to be exactly soluble, many of the problems in Physics 205 can be solved only by making suitable approximations. This can make life seem much more complicated. Afterall, limiting oneself to what is exactly true has a certain simplicity. However, the use of approximations makes tractable a large number of difficult-to-solve or impossible-to-solve problems, so it’s worth some effort to learn how to do it. In the days before computers and pocket calculators, approximations were indispen- sible since even something as simple as evaluating a sine function could be a really big deal (they had to be looked up in tables). Although this is not such a big factor anymore, the ability to extract simple analytical solutions is still generally a big plus, even if those solutions are only approximately correct. Analytical solutions are often very useful for providing insight—e.g., how things change if one or another variable changes. Believe it or not, there actually is some rhyme and reason in the art of approximation making, which when mastered will increase your comfort level. In what follows I’ll try to present some general guidelines. When and Why to Make Approximations There is no set rule, but generally approximations are useful when they render a difficult or impossible integral or differential equation soluble. A classic example is the simple pendulum. Starting with the Lagrangian for this system 1 L = T − V = m2θ˙2 + mg cos θ 2 from which Lagrange’s equation yields d ∂L ∂L − = m2θ¨ + mg sin θ =0 dt ∂θ˙ ∂θ or g θ¨ + sin θ =0 1 This problem can actually be solved exactly in terms of “elliptic integrals” (not a pretty sight! See M&T Section 4.4), but we immediately gain insight by making the approxima- tion sin θ θ, θ 1 at which point we obtain g θ¨ + θ =0 which is just simple harmonic motion with ω = g/. One hallmark of a good approximation is whether or not it approaches the exact so- lution as its underlying assumption becomes more and more true. In the simple pendulum example, the exact result for the period is given by l k2 9 τ =2π 1+ + k4 + ... g 4 64 where k =sinθ0/2 is a measure of the pendulum’s amplitude. For an amplitude of θ0 = 45◦, k =0.3827 and the period of the pendulum is about 4% larger than what is predicted using the sin θ θ approximation. This sort of accuracy is already close enough for many purposes. If the amplitude is small, say 1◦, the exact and approximate results differ by less than 20 parts per million. At that level many other considerations come into play: air resistance, the mass of the pendulm string, etc. Exercise: Show that small θ equation of motion above can be obtained by expanding the Lagrangian in a Taylor series around θ = θ0 = 0. In other words, show that the small θ approximation corresponds to the “standard trick” of approximating a potential energy function by a parabola. Other Standard Approximations In the example and exercise above, trignometric functions (sine or cosine) were re- placed by their Taylor series expansions. Another common practice is to approximate the sum of two (or more) terms raised to some power using the binomial expansion. This works best√ when one of the two terms is much smaller than the other. For example, if we have z = x + y and x y,wecanwrite 1 √ y 2 √ 1 y z = x 1+ x 1+ x 2 x If x =10y, this approximation is accurate to almost one part per thousand. In general we have n(n − 1) n(n − 1)(n − 2) (1 + x)n 1+nx + x2 + x3 ... 2! 3! 2 If n is a positive integer, the series has a finite number of terms (n + 2). Note that this is just a special case of the Taylor series and is applicable for positive and negative n and for integer and non-integer n. This approximation is used frequently since is “linearizes” expressions, often making them much more amenable to algebraic solution. One also frequently encounters exponentials and logarithms, which can be expanded as x x2 x2 x3 ex =1+ + + ... and log(1 + x)=x − + ... 1! 2! 2 3 The list goes on and on. Indeed Taylor’s theorem tells us that essentially everything has its expansion—a good thing since all computers can do is multiply, divide, add, and subtract. Don’t be fooled by that “sin” key on your calculator! In the end what it needs to do is to evaluate the appropriate series expansion. (Question: Most such series are infinite. How can the calculator complete the calculation in a finite time?) Some jargon and some pitfalls One often hears scientists (and others) say things like “We estimate to first order that ...”. What is meant is that the estimate to be described is accurate to the extent that the “expansion parameter” is small compared to its square. For example, the first order estimate for sin θ is θ. Note that there is an implicit caveat in such statements, because θ might not be small, or, the series being expanded might not converge nicely. One way to check the validity of such estimates is to evaluate the “higher order” terms. In the case of thesinewehave x3 x5 sin x = x − + − ... 3! 5! Thus for x =0.1 radians, the first order estimate for sin x is 0.1, the second order correction is zero, and the third order correction is 1.67 × 10−4. Here is a case where the first order estimate is pretty good. (The fact that the next order correction is small is usually, although not always, a pretty good indicator that the first-order estimate is reliable.) If, however, x =1.0, the first order estimate for sin x (i.e., 1.0) is subject to a 16% higher order correction, since the exact value is sin(1.0) = .8414710 Note that including the next term in the expansion, gives a value (0.833333) that is accurate to within 1%. Once again we could have guessed this by noting that the next term in the series x5/5! 0.008333 is pretty small. One thing one needs to watch out for is the case where two leading order terms cancel. Consider, for example, the expression y = 1000(x − sin x)forx 1. The second term is in the parentheses is usually approximated by x, which is also quite obviously an excellent approximation to the first term. One might argue, therefore, that y 0. In some cases, 3 this might be an acceptable answer, but in other cases it might not. For example, if y represents the probability of a major nuclear disaster in a crowded urban area and x =0.1, the exact value of 16.6% is unacceptably high. To get a better answer one needs to “go out to higher order”—i.e., to write x3 x3 y 1000 (x − x + ) = 1000 =0.166 3! 3! Even though the higher order term is still off by a tad, the basic conclusion is not altered— i.e., the risk is unacceptably high. Whether it’s 16% or 16.5% is a fine point. It is also important to carry out calculations to a consistent order. An example of this occurs is the solar oblateness problem in Homework 3. In that problem, one needs to expand the quantity 1 1 = r r2 + a2 − 2ar cos φ where r a. If we define a 2 2a x ≡ − cos φ r r it is tempting (but wrong!) to write 2 1 1 − 1 1 x 1 a 2a = (1 + x) 2 1 − = 1 − + cos φ r r r 2 r r r the reason this won’t work is that the next term in the square root expansion (3x2/8) contains terms of order (a/r)2. As it turns out, the leading order term (∝ (a/r)cosφ) cancels out during the integration that follows, making the higher order corrections quite important. Fortunately, this sort of cancellation isn’t too common, but one does need to watch out. Determining the “Order” of a Term It is difficult to come up with a well defined recipe for doing this, but it may be helpful to consider an example in some detail. The example in question is the “Not-So- Simple Pendulum” (see Lecture I Notes, pg. 11). Lagrange’s equations yield two coupled differential equations, the first of which is mx¨ + mθ¨ cos θ − mθ˙2 sin θ + kx =0 In the limit where θ and x are small, this becomes mx¨ + mθ¨ + kx 0 4 The first term is first order (one power of x and/or θ) the second term is also first order, but we use cos θ 1. Note that this would not have been a good approximation in the potential energy term of the Lagrangian. The reason for this is that the constant part of cos θ 1 − θ2/2 (i.e., “1”) has no effect at all when added to the potential energy, whereas by the time we get to the differential equation cos θ appears as a multiplicative factor and can therefore be safely neglected provided θ is small enough. The third term in the differential equation is third order: two powers of θ appear in the θ˙2 factor, and a third appears in the sin θ θ factor.
Recommended publications
  • A Short Course on Approximation Theory
    A Short Course on Approximation Theory N. L. Carothers Department of Mathematics and Statistics Bowling Green State University ii Preface These are notes for a topics course offered at Bowling Green State University on a variety of occasions. The course is typically offered during a somewhat abbreviated six week summer session and, consequently, there is a bit less material here than might be associated with a full semester course offered during the academic year. On the other hand, I have tried to make the notes self-contained by adding a number of short appendices and these might well be used to augment the course. The course title, approximation theory, covers a great deal of mathematical territory. In the present context, the focus is primarily on the approximation of real-valued continuous functions by some simpler class of functions, such as algebraic or trigonometric polynomials. Such issues have attracted the attention of thousands of mathematicians for at least two centuries now. We will have occasion to discuss both venerable and contemporary results, whose origins range anywhere from the dawn of time to the day before yesterday. This easily explains my interest in the subject. For me, reading these notes is like leafing through the family photo album: There are old friends, fondly remembered, fresh new faces, not yet familiar, and enough easily recognizable faces to make me feel right at home. The problems we will encounter are easy to state and easy to understand, and yet their solutions should prove intriguing to virtually anyone interested in mathematics. The techniques involved in these solutions entail nearly every topic covered in the standard undergraduate curriculum.
    [Show full text]
  • Appendix a Short Course in Taylor Series
    Appendix A Short Course in Taylor Series The Taylor series is mainly used for approximating functions when one can identify a small parameter. Expansion techniques are useful for many applications in physics, sometimes in unexpected ways. A.1 Taylor Series Expansions and Approximations In mathematics, the Taylor series is a representation of a function as an infinite sum of terms calculated from the values of its derivatives at a single point. It is named after the English mathematician Brook Taylor. If the series is centered at zero, the series is also called a Maclaurin series, named after the Scottish mathematician Colin Maclaurin. It is common practice to use a finite number of terms of the series to approximate a function. The Taylor series may be regarded as the limit of the Taylor polynomials. A.2 Definition A Taylor series is a series expansion of a function about a point. A one-dimensional Taylor series is an expansion of a real function f(x) about a point x ¼ a is given by; f 00ðÞa f 3ðÞa fxðÞ¼faðÞþf 0ðÞa ðÞþx À a ðÞx À a 2 þ ðÞx À a 3 þÁÁÁ 2! 3! f ðÞn ðÞa þ ðÞx À a n þÁÁÁ ðA:1Þ n! © Springer International Publishing Switzerland 2016 415 B. Zohuri, Directed Energy Weapons, DOI 10.1007/978-3-319-31289-7 416 Appendix A: Short Course in Taylor Series If a ¼ 0, the expansion is known as a Maclaurin Series. Equation A.1 can be written in the more compact sigma notation as follows: X1 f ðÞn ðÞa ðÞx À a n ðA:2Þ n! n¼0 where n ! is mathematical notation for factorial n and f(n)(a) denotes the n th derivation of function f evaluated at the point a.
    [Show full text]
  • February 2009
    How Euler Did It by Ed Sandifer Estimating π February 2009 On Friday, June 7, 1779, Leonhard Euler sent a paper [E705] to the regular twice-weekly meeting of the St. Petersburg Academy. Euler, blind and disillusioned with the corruption of Domaschneff, the President of the Academy, seldom attended the meetings himself, so he sent one of his assistants, Nicolas Fuss, to read the paper to the ten members of the Academy who attended the meeting. The paper bore the cumbersome title "Investigatio quarundam serierum quae ad rationem peripheriae circuli ad diametrum vero proxime definiendam maxime sunt accommodatae" (Investigation of certain series which are designed to approximate the true ratio of the circumference of a circle to its diameter very closely." Up to this point, Euler had shown relatively little interest in the value of π, though he had standardized its notation, using the symbol π to denote the ratio of a circumference to a diameter consistently since 1736, and he found π in a great many places outside circles. In a paper he wrote in 1737, [E74] Euler surveyed the history of calculating the value of π. He mentioned Archimedes, Machin, de Lagny, Leibniz and Sharp. The main result in E74 was to discover a number of arctangent identities along the lines of ! 1 1 1 = 4 arctan " arctan + arctan 4 5 70 99 and to propose using the Taylor series expansion for the arctangent function, which converges fairly rapidly for small values, to approximate π. Euler also spent some time in that paper finding ways to approximate the logarithms of trigonometric functions, important at the time in navigation tables.
    [Show full text]
  • Approximation Atkinson Chapter 4, Dahlquist & Bjork Section 4.5
    Approximation Atkinson Chapter 4, Dahlquist & Bjork Section 4.5, Trefethen's book Topics marked with ∗ are not on the exam 1 In approximation theory we want to find a function p(x) that is `close' to another function f(x). We can define closeness using any metric or norm, e.g. Z 2 2 kf(x) − p(x)k2 = (f(x) − p(x)) dx or kf(x) − p(x)k1 = sup jf(x) − p(x)j or Z kf(x) − p(x)k1 = jf(x) − p(x)jdx: In order for these norms to make sense we need to restrict the functions f and p to suitable function spaces. The polynomial approximation problem takes the form: Find a polynomial of degree at most n that minimizes the norm of the error. Naturally we will consider (i) whether a solution exists and is unique, (ii) whether the approximation converges as n ! 1. In our section on approximation (loosely following Atkinson, Chapter 4), we will first focus on approximation in the infinity norm, then in the 2 norm and related norms. 2 Existence for optimal polynomial approximation. Theorem (no reference): For every n ≥ 0 and f 2 C([a; b]) there is a polynomial of degree ≤ n that minimizes kf(x) − p(x)k where k · k is some norm on C([a; b]). Proof: To show that a minimum/minimizer exists, we want to find some compact subset of the set of polynomials of degree ≤ n (which is a finite-dimensional space) and show that the inf over this subset is less than the inf over everything else.
    [Show full text]
  • Student Understanding of Taylor Series Expansions in Statistical Mechanics
    PHYSICAL REVIEW SPECIAL TOPICS - PHYSICS EDUCATION RESEARCH 9, 020110 (2013) Student understanding of Taylor series expansions in statistical mechanics Trevor I. Smith,1 John R. Thompson,2,3 and Donald B. Mountcastle2 1Department of Physics and Astronomy, Dickinson College, Carlisle, Pennsylvania 17013, USA 2Department of Physics and Astronomy, University of Maine, Orono, Maine 04469, USA 3Maine Center for Research in STEM Education, University of Maine, Orono, Maine 04469, USA (Received 2 March 2012; revised manuscript received 23 May 2013; published 8 August 2013) One goal of physics instruction is to have students learn to make physical meaning of specific mathematical expressions, concepts, and procedures in different physical settings. As part of research investigating student learning in statistical physics, we are developing curriculum materials that guide students through a derivation of the Boltzmann factor using a Taylor series expansion of entropy. Using results from written surveys, classroom observations, and both individual think-aloud and teaching interviews, we present evidence that many students can recognize and interpret series expansions, but they often lack fluency in creating and using a Taylor series appropriately, despite previous exposures in both calculus and physics courses. DOI: 10.1103/PhysRevSTPER.9.020110 PACS numbers: 01.40.Fk, 05.20.Ày, 02.30.Lt, 02.30.Mv students’ use of Taylor series expansion to derive and I. INTRODUCTION understand the Boltzmann factor as a component of proba- Over the past several decades, much work has been done bility in the canonical ensemble. While this is a narrow investigating ways in which students understand and learn topic, we feel that it serves as an excellent example of physics (see Refs.
    [Show full text]
  • Nonlinear Optics
    Nonlinear Optics: A very brief introduction to the phenomena Paul Tandy ADVANCED OPTICS (PHYS 545), Spring 2010 Dr. Mendes INTRO So far we have considered only the case where the polarization density is proportionate to the electric field. The electric field itself is generally supplied by a laser source. PE= ε0 χ Where P is the polarization density, E is the electric field, χ is the susceptibility, and ε 0 is the permittivity of free space. We can imagine that the electric field somehow induces a polarization in the material. As the field increases the polarization also in turn increases. But we can also imagine that this process does not continue indefinitely. As the field grows to be very large the field itself starts to become on the order of the electric fields between the atoms in the material (medium). As this occurs the material itself becomes slightly altered and the behavior is no longer linear in nature but instead starts to become weakly nonlinear. We can represent this as a Taylor series expansion model in the electric field. ∞ n 23 PAEAEAEAE==++∑ n 12 3+... n=1 This model has become so popular that the coefficient scaling has become standardized using the letters d and χ (3) as parameters corresponding the quadratic and cubic terms. Higher order terms would play a role too if the fields were extremely high. 2(3)3 PEdEE=++()εχ0 (2) ( 4χ ) + ... To get an idea of how large a typical field would have to be before we see a nonlinear effect, let us scale the above equation by the first order term.
    [Show full text]
  • Fundamentals of Duct Acoustics
    Fundamentals of Duct Acoustics Sjoerd W. Rienstra Technische Universiteit Eindhoven 16 November 2015 Contents 1 Introduction 3 2 General Formulation 4 3 The Equations 8 3.1 AHierarchyofEquations ........................... 8 3.2 BoundaryConditions. Impedance.. 13 3.3 Non-dimensionalisation . 15 4 Uniform Medium, No Mean Flow 16 4.1 Hard-walled Cylindrical Ducts . 16 4.2 RectangularDucts ............................... 21 4.3 SoftWallModes ................................ 21 4.4 AttenuationofSound.............................. 24 5 Uniform Medium with Mean Flow 26 5.1 Hard-walled Cylindrical Ducts . 26 5.2 SoftWallandUniformMeanFlow . 29 6 Source Expansion 32 6.1 ModalAmplitudes ............................... 32 6.2 RotatingFan .................................. 32 6.3 Tyler and Sofrin Rule for Rotor-Stator Interaction . ..... 33 6.4 PointSourceinaLinedFlowDuct . 35 6.5 PointSourceinaDuctWall .......................... 38 7 Reflection and Transmission 40 7.1 A Discontinuity in Diameter . 40 7.2 OpenEndReflection .............................. 43 VKI - 1 - CONTENTS CONTENTS A Appendix 49 A.1 BesselFunctions ................................ 49 A.2 AnImportantComplexSquareRoot . 51 A.3 Myers’EnergyCorollary ............................ 52 VKI - 2 - 1. INTRODUCTION CONTENTS 1 Introduction In a duct of constant cross section, with a medium and boundary conditions independent of the axial position, the wave equation for time-harmonic perturbations may be solved by means of a series expansion in a particular family of self-similar solutions, called modes. They are related to the eigensolutions of a two-dimensional operator, that results from the wave equation, on a cross section of the duct. For the common situation of a uniform medium without flow, this operator is the well-known Laplace operator 2. For a non- uniform medium, and in particular with mean flow, the details become mo∇re complicated, but the concept of duct modes remains by and large the same1.
    [Show full text]
  • Fourier Series
    Chapter 10 Fourier Series 10.1 Periodic Functions and Orthogonality Relations The differential equation ′′ y + 2y = F cos !t models a mass-spring system with natural frequency with a pure cosine forcing function of frequency !. If 2 = !2 a particular solution is easily found by undetermined coefficients (or by∕ using Laplace transforms) to be F y = cos !t. p 2 !2 − If the forcing function is a linear combination of simple cosine functions, so that the differential equation is N ′′ 2 y + y = Fn cos !nt n=1 X 2 2 where = !n for any n, then, by linearity, a particular solution is obtained as a sum ∕ N F y (t)= n cos ! t. p 2 !2 n n=1 n X − This simple procedure can be extended to any function that can be repre- sented as a sum of cosine (and sine) functions, even if that summation is not a finite sum. It turns out that the functions that can be represented as sums in this form are very general, and include most of the periodic functions that are usually encountered in applications. 723 724 10 Fourier Series Periodic Functions A function f is said to be periodic with period p> 0 if f(t + p)= f(t) for all t in the domain of f. This means that the graph of f repeats in successive intervals of length p, as can be seen in the graph in Figure 10.1. y p 2p 3p 4p 5p Fig. 10.1 An example of a periodic function with period p.
    [Show full text]
  • Function Approximation Through an Efficient Neural Networks Method
    Paper ID #25637 Function Approximation through an Efficient Neural Networks Method Dr. Chaomin Luo, Mississippi State University Dr. Chaomin Luo received his Ph.D. in Department of Electrical and Computer Engineering at Univer- sity of Waterloo, in 2008, his M.Sc. in Engineering Systems and Computing at University of Guelph, Canada, and his B.Eng. in Electrical Engineering from Southeast University, Nanjing, China. He is cur- rently Associate Professor in the Department of Electrical and Computer Engineering, at the Mississippi State University (MSU). He was panelist in the Department of Defense, USA, 2015-2016, 2016-2017 NDSEG Fellowship program and panelist in 2017 NSF GRFP Panelist program. He was the General Co-Chair of 2015 IEEE International Workshop on Computational Intelligence in Smart Technologies, and Journal Special Issues Chair, IEEE 2016 International Conference on Smart Technologies, Cleveland, OH. Currently, he is Associate Editor of International Journal of Robotics and Automation, and Interna- tional Journal of Swarm Intelligence Research. He was the Publicity Chair in 2011 IEEE International Conference on Automation and Logistics. He was on the Conference Committee in 2012 International Conference on Information and Automation and International Symposium on Biomedical Engineering and Publicity Chair in 2012 IEEE International Conference on Automation and Logistics. He was a Chair of IEEE SEM - Computational Intelligence Chapter; a Vice Chair of IEEE SEM- Robotics and Automa- tion and Chair of Education Committee of IEEE SEM. He has extensively published in reputed journal and conference proceedings, such as IEEE Transactions on Neural Networks, IEEE Transactions on SMC, IEEE-ICRA, and IEEE-IROS, etc. His research interests include engineering education, computational intelligence, intelligent systems and control, robotics and autonomous systems, and applied artificial in- telligence and machine learning for autonomous systems.
    [Show full text]
  • Graduate Macro Theory II: Notes on Log-Linearization
    Graduate Macro Theory II: Notes on Log-Linearization Eric Sims University of Notre Dame Spring 2017 The solutions to many discrete time dynamic economic problems take the form of a system of non-linear difference equations. There generally exists no closed-form solution for such problems. As such, we must result to numerical and/or approximation techniques. One particularly easy and very common approximation technique is that of log linearization. We first take natural logs of the system of non-linear difference equations. We then linearize the logged difference equations about a particular point (usually a steady state), and simplify until we have a system of linear difference equations where the variables of interest are percentage deviations about a point (again, usually a steady state). Linearization is nice because we know how to work with linear difference equations. Putting things in percentage terms (that's the \log" part) is nice because it provides natural interpretations of the units (i.e. everything is in percentage terms). First consider some arbitrary univariate function, f(x). Taylor's theorem tells us that this can be expressed as a power series about a particular point x∗, where x∗ belongs to the set of possible x values: f 0(x∗) f 00 (x∗) f (3)(x∗) f(x) = f(x∗) + (x − x∗) + (x − x∗)2 + (x − x∗)3 + ::: 1! 2! 3! Here f 0(x∗) is the first derivative of f with respect to x evaluated at the point x∗, f 00 (x∗) is the second derivative evaluated at the same point, f (3) is the third derivative, and so on.
    [Show full text]
  • A Rational Approximation to the Logarithm
    A Rational Approximation to the Logarithm By R. P. Kelisky and T. J. Rivlin We shall use the r-method of Lanczos to obtain rational approximations to the logarithm which converge in the entire complex plane slit along the nonpositive real axis, uniformly in certain closed regions in the slit plane. We also provide error esti- mates valid in these regions. 1. Suppose z is a complex number, and s(l, z) denotes the closed line segment joining 1 and z. Let y be a positive real number, y ¿¿ 1. As a polynomial approxima- tion, of degree n, to log x on s(l, y) we propose p* G Pn, Pn being the set of poly- nomials with real coefficients of degree at most n, which minimizes ||xp'(x) — X|J===== max \xp'(x) — 1\ *e«(l.v) among all p G Pn which satisfy p(l) = 0. (This procedure is suggested by the observation that w = log x satisfies xto'(x) — 1=0, to(l) = 0.) We determine p* as follows. Suppose p(x) = a0 + aix + ■• • + a„xn, p(l) = 0 and (1) xp'(x) - 1 = x(x) = b0 + bix +-h bnxn . Then (2) 6o = -1 , (3) a,- = bj/j, j = 1, ■•-,«, and n (4) a0= - E Oj • .-i Thus, minimizing \\xp' — 1|| subject to p(l) = 0 is equivalent to minimizing ||«-|| among all i£P„ which satisfy —ir(0) = 1. This, in turn, is equivalent to maximiz- ing |7r(0) I among all v G Pn which satisfy ||ir|| = 1, in the sense that if vo maximizes |7r(0)[ subject to ||x|| = 1, then** = —iro/-7ro(0) minimizes ||x|| subject to —ir(0) = 1.
    [Show full text]
  • Approximation of a Function by a Polynomial Function
    1 Approximation of a function by a polynomial function 1. step: Assuming that f(x) is differentiable at x = a, from the picture we see: 푓(푥) = 푓(푎) + ∆푓 ≈ 푓(푎) + 푓′(푎)∆푥 Linear approximation of f(x) at point x around x = a 푓(푥) = 푓(푎) + 푓′(푎)(푥 − 푎) 푅푒푚푎푛푑푒푟 푹 = 푓(푥) − 푓(푎) − 푓′(푎)(푥 − 푎) will determine magnitude of error Let’s try to get better approximation of f(x). Let’s assume that f(x) has all derivatives at x = a. Let’s assume there is a possible power series expansion of f(x) around a. ∞ 2 3 푛 푓(푥) = 푐0 + 푐1(푥 − 푎) + 푐2(푥 − 푎) + 푐3(푥 − 푎) + ⋯ = ∑ 푐푛(푥 − 푎) ∀ 푥 푎푟표푢푛푑 푎 푛=0 n 푓(푛)(푥) 푓(푛)(푎) 2 3 0 푓 (푥) = 푐0 + 푐1(푥 − 푎) + 푐2(푥 − 푎) + 푐3(푥 − 푎) + ⋯ 푓(푎) = 푐0 2 1 푓′(푥) = 푐1 + 2푐2(푥 − 푎) + 3푐3(푥 − 푎) + ⋯ 푓′(푎) = 푐1 2 2 푓′′(푥) = 2푐2 + 3 ∙ 2푐3(푥 − 푎) + 4 ∙ 3(푥 − 푎) ∙∙∙ 푓′′(푎) = 2푐2 2 3 푓′′′(푥) = 3 ∙ 2푐3 + 4 ∙ 3 ∙ 2(푥 − 푎) + 5 ∙ 4 ∙ 3(푥 − 푎) ∙∙∙ 푓′′′(푎) = 3 ∙ 2푐3 ⋮ (푛) (푛) n 푓 (푥) = 푛(푛 − 1) ∙∙∙ 3 ∙ 2 ∙ 1푐푛 + (푛 + 1)푛(푛 − 1) ⋯ 2푐푛+1(푥 − 푎) 푓 (푎) = 푛! 푐푛 ⋮ Taylor’s theorem If f (x) has derivatives of all orders in an open interval I containing a, then for each x in I, ∞ 푓(푛)(푎) 푓(푛)(푎) 푓(푛+1)(푎) 푓(푥) = ∑ (푥 − 푎)푛 = 푓(푎) + 푓′(푎)(푥 − 푎) + ⋯ + (푥 − 푎)푛 + (푥 − 푎)푛+1 + ⋯ 푛! 푛! (푛 + 1)! 푛=0 푛 ∞ 푓(푘)(푎) 푓(푘)(푎) = ∑ (푥 − 푎)푘 + ∑ (푥 − 푎)푘 푘! 푘! 푘=0 푘=푛+1 convention for the sake of algebra: 0! = 1 = 푇푛 + 푅푛 푛 푓(푘)(푎) 푇 = ∑ (푥 − 푎)푘 푝표푙푦푛표푚푎푙 표푓 푛푡ℎ 표푟푑푒푟 푛 푘! 푘=0 푓(푛+1)(푧) 푛+1 푅푛(푥) = (푥 − 푎) 푠 푟푒푚푎푛푑푒푟 푅푛 (푛 + 1)! Taylor’s theorem says that there exists some value 푧 between 푎 and 푥 for which: ∞ 푓(푘)(푎) 푓(푛+1)(푧) ∑ (푥 − 푎)푘 can be replaced by 푅 (푥) = (푥 − 푎)푛+1 푘! 푛 (푛 + 1)! 푘=푛+1 2 Approximating function by a polynomial function So we are good to go.
    [Show full text]