Maps and differential equations

Similar documents
Introduction to bifurcations

Dynamical Systems: Lecture 1 Naima Hammoud

Mathematical Foundations of Neuroscience - Lecture 7. Bifurcations II.

Lecture 10: Powers of Matrices, Difference Equations

Why are Discrete Maps Sufficient?

THREE DIMENSIONAL SYSTEMS. Lecture 6: The Lorenz Equations

24. Nonlinear programming

QUADRATICS 3.2 Breaking Symmetry: Factoring

ACCESS TO SCIENCE, ENGINEERING AND AGRICULTURE: MATHEMATICS 1 MATH00030 SEMESTER / Lines and Their Equations

Lecture 15: Exploding and Vanishing Gradients

v n+1 = v T + (v 0 - v T )exp(-[n +1]/ N )

DIFFERENTIAL EQUATIONS

6.2 Brief review of fundamental concepts about chaotic systems

11 Chaos in Continuous Dynamical Systems.

Introduction to Nonlinear Dynamics and Chaos

More Details Fixed point of mapping is point that maps into itself, i.e., x n+1 = x n.

Chaos and Liapunov exponents

Handout 2: Invariant Sets and Stability

ARE202A, Fall 2005 CONTENTS. 1. Graphical Overview of Optimization Theory (cont) Separating Hyperplanes 1

EE222 - Spring 16 - Lecture 2 Notes 1

Main topics for the First Midterm Exam

Dynamical Systems and Chaos Part I: Theoretical Techniques. Lecture 4: Discrete systems + Chaos. Ilya Potapov Mathematics Department, TUT Room TD325

Lesson 21 Not So Dramatic Quadratics

TWO DIMENSIONAL FLOWS. Lecture 5: Limit Cycles and Bifurcations

PH36010: Numerical Methods - Evaluating the Lorenz Attractor using Runge-Kutta methods Abstract

Fundamentals of Dynamical Systems / Discrete-Time Models. Dr. Dylan McNamara people.uncw.edu/ mcnamarad

Lecture 1 Monday, January 14

8.3 Partial Fraction Decomposition

Unit Ten Summary Introduction to Dynamical Systems and Chaos

Quadratic Equations Part I

Math101, Sections 2 and 3, Spring 2008 Review Sheet for Exam #2:

Chapter 9: Roots and Irrational Numbers

Math 1270 Honors ODE I Fall, 2008 Class notes # 14. x 0 = F (x; y) y 0 = G (x; y) u 0 = au + bv = cu + dv

Applications of Hurst Coefficient Analysis to Chaotic Response of ODE Systems: Part 1a, The Original Lorenz System of 1963

The Derivative of a Function

Lecture 4: Training a Classifier

Lab 5: Nonlinear Systems

22 Approximations - the method of least squares (1)

The Liapunov Method for Determining Stability (DRAFT)

MATH 415, WEEKS 7 & 8: Conservative and Hamiltonian Systems, Non-linear Pendulum

Chapter 6 - Ordinary Differential Equations

APPM 2460 CHAOTIC DYNAMICS

2. FUNCTIONS AND ALGEBRA

MATH 415, WEEK 12 & 13: Higher-Dimensional Systems, Lorenz Equations, Chaotic Behavior

Introduction to Algorithms / Algorithms I Lecturer: Michael Dinitz Topic: Intro to Learning Theory Date: 12/8/16

Scenarios for the transition to chaos

PHY411 Lecture notes Part 5

Lecture 3. Dynamical Systems in Continuous Time

Stability Analysis for ODEs

Chaos. Lendert Gelens. KU Leuven - Vrije Universiteit Brussel Nonlinear dynamics course - VUB

Algebra. Here are a couple of warnings to my students who may be here to get a copy of what happened on a day that you missed.

then the substitution z = ax + by + c reduces this equation to the separable one.

Lecture 27: Theory of Computation. Marvin Zhang 08/08/2016

Descriptive Statistics (And a little bit on rounding and significant digits)

( )( b + c) = ab + ac, but it can also be ( )( a) = ba + ca. Let s use the distributive property on a couple of

Section 0.6: Factoring from Precalculus Prerequisites a.k.a. Chapter 0 by Carl Stitz, PhD, and Jeff Zeager, PhD, is available under a Creative

1 Review of the dot product

Math 308 Midterm Answers and Comments July 18, Part A. Short answer questions

Business Statistics. Lecture 9: Simple Regression

Section 20: Arrow Diagrams on the Integers

Basic Theory of Dynamical Systems

Calculus II. Calculus II tends to be a very difficult course for many students. There are many reasons for this.

Figure 1: Doing work on a block by pushing it across the floor.

Chapter 1, Section 1.2, Example 9 (page 13) and Exercise 29 (page 15). Use the Uniqueness Tool. Select the option ẋ = x

Computational Physics (6810): Session 8

Midterm 1 Review. Distance = (x 1 x 0 ) 2 + (y 1 y 0 ) 2.

Simplest Chaotic Flows with Involutional Symmetries

2 Discrete Dynamical Systems (DDS)

Relevant sections from AMATH 351 Course Notes (Wainwright): 1.3 Relevant sections from AMATH 351 Course Notes (Poulin and Ingalls): 1.1.

Lecture 22. r i+1 = b Ax i+1 = b A(x i + α i r i ) =(b Ax i ) α i Ar i = r i α i Ar i

Final Review Sheet. B = (1, 1 + 3x, 1 + x 2 ) then 2 + 3x + 6x 2

Chaotic motion. Phys 750 Lecture 9

Lecture 6, September 1, 2017

MATH 215/255 Solutions to Additional Practice Problems April dy dt

Physics 202 Laboratory 3. Root-Finding 1. Laboratory 3. Physics 202 Laboratory

MATH 308 COURSE SUMMARY

LECTURE 8: DYNAMICAL SYSTEMS 7

An Undergraduate s Guide to the Hartman-Grobman and Poincaré-Bendixon Theorems

Invariant manifolds. Marc R. Roussel. October 4, 2005

CHAOS -SOME BASIC CONCEPTS

Topic 5 Notes Jeremy Orloff. 5 Homogeneous, linear, constant coefficient differential equations

Solving simple equations, a review/primer

CS 61A/CS Mehrdad Niknami. University of California, Berkeley. Mehrdad Niknami (UC Berkeley) CS 61A/CS / 25

What is proof? Lesson 1

6x 2 8x + 5 ) = 12x 8

Logistic Model with Harvesting

One dimensional Maps

Hypothesis testing I. - In particular, we are talking about statistical hypotheses. [get everyone s finger length!] n =

Introduction to Classical Chaos

2.3 Damping, phases and all that

... it may happen that small differences in the initial conditions produce very great ones in the final phenomena. Henri Poincaré

2 Systems of Linear Equations

Solving Equations by Adding and Subtracting

Class 4: More Pendulum results

TAYLOR POLYNOMIALS DARYL DEFORD

Algebra & Trig Review

CS 542G: Robustifying Newton, Constraints, Nonlinear Least Squares

6.080 / Great Ideas in Theoretical Computer Science Spring 2008

Conceptual Explanations: Simultaneous Equations Distance, rate, and time

Qualitative Analysis of Tumor-Immune ODE System

Transcription:

Maps and differential equations Marc R. Roussel November 8, 2005 Maps are algebraic rules for computing the next state of dynamical systems in discrete time. Differential equations and maps have a number of important connections which we explore in this lecture. Many of the topics mentioned in these notes are also connected to each other. You may be able to see some of the connections immediately, but in any case you should certainly think about them. 1 Numerical methods as maps In an earlier lecture, we noted that Euler s method for integrating differential equations, z n+1 = z n + hf(z n ), is a map: Given the state in step n, the state in step n + 1 can be computed from the differential equation (f(z n )) and the size of the time step h. Many numerical integration methods can be understood as maps, albeit usually more complicated ones than this. In the limit of small step size, sensible numerical integration methods converge to a solution of the corresponding differential equation. However, at larger step sizes, there are no guarantees. Example 1.1 Consider the logistic 1 differential equation dx dt Euler s method for this differential equation is = x(1 x). x n+1 = x n + hx n (1 x n ) = x n (1 + h hx n ). (1) This is called the logistic map. (This map is actually better known in a slightly different form. Let y n = hx n /(1 + h) and λ = 1 + h. Then, y n+1 = λy n (1 y n ).) The logistic differential equation has very simple behavior: x = 0 is an unstable equilibrium point, while x = 1 is a stable equilibrium. The logistic map on the other hand, can display an astonishing variety of attractors, depending on the parameters. Figure 1 gives some examples. For smallish values of h (up to about h = 2), the iterates of the logistic map 1 converge 1 This differential equation has applications in some problems in logistics, which is the systematic study of problems associated with the movement of goods or people to meet needs in distant locations. 1

1.4 1.3 h=1.9 h=2.3 h=2.5 1.2 1.1 1 x 0.9 0.8 0.7 0.6 0.5 5000 5005 5010 5015 5020 5025 5030 5035 5040 5045 5050 n 1.4 h=2.7 1.2 1 x 0.8 0.6 0.4 5000 5050 5100 5150 5200 n Figure 1: Some time series obtained from the application of the Euler method to the logistic differential equation for various values of h. The lines connecting the points are only there to guide the eye since values of x are only generated at integer values of n. 2

onto the stable equilibrium point of the logistic differential equation. However, for h > 2, we see periodic solutions, and eventually solutions which look chaotic. Thus, numerical methods can produce results which aren t even qualitatively correct. It is therefore a good idea to verify integration results carefully, varying the step size and, if possible, changing methods. Maps don t just turn up in the integration of differential equations. Consider for instance the problem of solving the algebraic equation f (x) = 0 for x. One common way to do this is to use Newton s method. Expand f (x) in a Taylor series near x n : f (x) = f (x n ) + (x x n ) f (x n ) +... Assuming that the first two terms are a reasonable approximation to the function and setting this expression to zero, we get x n+1 x n f (x n) f (x n ). We label the solution of the truncated equation x n+1 in recognition of the fact that we won t get the exact answer in one step, so we ll have to iterate this procedure many times to get the correct answer, if this is going to work at all. When it doesn t work, Newton s method often generates interesting dynamics. In fact, a lot of the early work on fractals came from a study of Newton s method in the complex plane. 2 Solution maps of differential equations It sometimes occurs that we have a differential equation for a system, but are only interested in the behavior at fixed time intervals. For instance, the logistic differential equation is sometimes used to model population growth, but we might only have census data at intervals of five or ten years. It then makes little sense to look at the whole continuous solution. Suppose for instance that we want solutions of the logistic differential equation (not solutions of some numerical approximation like the Euler method iterates) at fixed intervals T. The solution of this differential equation is > dsolve({diff(x(t),t)=x(t)*(1-x(t)),x(0)=x0},x(t)); or x(t) = 1 1 e( t) ( 1 + x0) x0 x(t) = x 0 x 0 e t (x 0 1). (2) e t = x 0(x(t) 1) x(t)(x 0 1), (3) x 0 and x(t + T ) = x 0 e (t+t) (x 0 1). (4) 3

If we substitute for the exponential in equation 4 using equation 3 we get, after a little rearranging, x(t + T ) = x(t) x(t) e T x 0 (x(t) 1). (5) This last equation is a solution map. It lets us calculate x(t + T ) knowing only x(t) and some parameters. Unlike equation 1 which gives approximations to the solution of the logistic differential equation at fixed intervals h, equation 5 is exact. We were able to obtain this equation because we were able to solve the differential equation. In general of course, we can t do that, but we can still obtain numerical representations of the solution map by sampling the numerical solution (obtained with a good numerical method, of course) at fixed time intervals and plotting x(t + T ) vs x(t). This is sometimes called a Ruelle plot. 3 Poincaré maps of systems with periodic nonautonomous terms The solution maps and Ruelle plots described above are clearly related to the Poincaré sections described in the last lecture for systems with periodic driving terms. To go from a Poincaré section to a Poincaré map, we need only mention that, in a deterministic dynamical system, there always has to be some rule (even if we don t know what it is) that can be applied to calculate the next intersection of a trajectory with the surface of section given the last intersection. For this reason, Poincaré maps are often called next-return maps. Formally, let {z 1,z 2,z 3,...} be the set of intersections of a trajectory with the surface of section. The Poincaré map is a relation P such that P (z n ) = z n+1. In some very simple cases, Poincaré maps can be computed analytically, but this isn t usually the case. Instead, we calculate the Poincaré map numerically. In principle, if we have a d-dimensional state space (and so a d + 1-dimensional phase space in a nonautonomous system), then the Poincaré map is a map from R d into R d, which is hard to picture for any d > 1. However, it is often sufficient to use just one variable, i.e. to plot z i (t + T ) vs z i (t). Example 3.1 We return to the driven Brusselator which we studied recently: ẋ = a 0 + αsin(2πt/t) bx + x 2 y x, ẏ = bx x 2 y. Suppose that we want to look at the Poincaré map using xpp. We use the xpp input file which recast the equations as follows: ẋ = a 0 + αsin(2πθ/t) bx + x 2 y x, ẏ = bx x 2 y, θ = 1. We set up xpp with a periodic phase space in θ of period T. We also set up a Poincaré section at some reasonable value of θ [0,T) and then run a trajectory with a long transient. 4

To set up the Ruelle plot, we start by doing something that doesn t on the face of it make much sense: We click on Viewaxes 2D and then tell xpp to plot y on both axes. Having done this, click on numerics ruelle plot. Set the x-axis shift to 1, and leave the other two shifts at zero. This plots y n 1 on the x axis, and y n on the y axis. Finally, hit Window/zoom Fit to see the plot. Fig. 2 shows two examples of Poincaré maps for the driven Brusselator. Panel (a) shows a case where we previously found quasiperiodic behavior. You will note that the Poincaré map is, like the section itself, the image of a closed curve. On the other hand, we get a different kind of figure in panel (b), at parameter values which we previously identified as being chaotic. The curve isn t continuous, and it certainly isn t closed. Note that the shape of this curve is roughly parabolic, just like the logistic map (equation 1). This isn t a coincidence, as we will see later. The fact that the chaotic map of Fig. 2(b) is single-valued means that we can actually use it to predict the next value of y given the last value of y at the section. For instance, if y n 1 = 2.2, y n = 3.15 (roughly). This means that we can compute the sequence of y s at the Poincaré section from this map. There s just one problem: If you try to do this, you will find that small errors in y n 1 for values near 3.2 produce large errors in y n. In other words, the Poincaré map displays sensitive dependence on initial conditions and is chaotic. Finally note that the reason we picked y rather than x to construct the Poincaré map is that the plot of x n vs x n 1 isn t as obviously informative. Again, there s no guarantee that we will get a sensible map from one variable. It was just luck that y gave us a nice, single-valued map in the chaotic regime. 4 Poincaré sections and maps in autonomous systems The concept of a Poincaré section or map is also useful in autonomous systems with some type of cyclic attractor. If one of the variables, say x, passes repeatedly through some particular level, we can define a Poincaré section by collecting values of the other variables every time x passes through the selected level in a particular direction, say when x is increasing. Figure 3 illustrates this procedure. As an illustration, we return to the autocatalator, last discussed in our lectures on bifurcation theory. The equations of motion were ȧ = µ(κ + c) ab 2 a, ḃ = 1 σ ( ab 2 + a b ), ċ = 1 (b c). δ Suppose that we re interested in studying the chaotic regime in this system. Figure 4 shows a projection of the attractor in the chaotic regime for this system. Note that there is hole in the middle of the attractor. We imagine a half-plane of constant b starting inside this hole which intersects the attractor to the right (going toward larger values of a). We can for instance take our section at 5

3.3 3.2 (a) 3.1 3 2.9 y n 2.8 2.7 2.6 2.5 2.4 2.4 2.5 2.6 2.7 2.8 2.9 3 3.1 3.2 3.3 y n-1 3.4 3.2 (b) 3 2.8 y n 2.6 2.4 2.2 2 2 2.2 2.4 2.6 2.8 3 3.2 3.4 y n-1 Figure 2: Poincaré maps for the driven Brusselator at parameter values a 0 = 0.4, b = 1.2, T = 7.8, and (a) α = 0.008 or (b) α = 0.05. 6

y x Figure 3: Computation of the Poincare section of a system with a cyclic attractor. We pick some particular level of one of the variables, in this case x, and save the values of the other variables (e.g. y in the illustration) when x passes through the selected level in a particular direction. In this case, we decided to collect values of the other variables when x increases while passing through the section, shown as a dashed line. log(b) 5 4 3 2 1 µ=0.154 0-1 -8-7 -6-5 -4-3 log(a) -2-1 0 Figure 4: Strange attractor of the autocatalator at µ = 0.154, κ = 65, δ = 0.02 and σ = 0.005. 7

14 12 10 c 8 6 4 2 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55 0.6 0.65 a Figure 5: Poincaré section for b increasing through the value 12 in the autocatalator at the parameters of Fig. 4. logb = 2.5, which corresponds to b = e 2.5 12. We choose to cut to the right because there is a larger dispersion of values of a in this direction than in the other. This usually gives better Poincaré maps, although as in most other things in nonlinear dynamics, it s often necessary to experiment to get the right cut. We now need to know whether b increases or decreases in this part of the attractor. If a is large then, from the differential equation, ḃ > 0. Having gathered all this intelligence, we re ready to go to xpp. We run the system for a little while, discarding a transient to make sure that we are on the attractor. We then set up a Poincaré section at b = 12, setting the Direction in the appropriate dialog box to 1, which indicates an increasing value of b. If we run the model again, we get the image shown in Fig. 5. We can compute the Poincaré map using xpp s Ruelle plot function. I decided to plot a n vs a n 1, but c would have done just as nicely. The result appears in Fig. 6. The map isn t single-valued between a n 1 = 0.3 and 0.4 (roughly), but it is pretty close to just being a simple curve. In fact, the map is pretty close to being parabolic from a n 1 = 0.15 to 0.3 or so. This again suggests that we could learn something about this system from studying analytic maps which have similar shapes. Finally note that Poincaré sections don t have to be orthogonal to a coordinate axis and that it is sometimes convenient to use a surface of section which has some other orientation in phase space. 5 Next-amplitude maps In some cases, we get a nicer looking map is we use a slightly different construction. Suppose that instead of collecting data on a surface of section, we collect data when a variable x reaches a 8

0.65 0.6 0.55 0.5 0.45 a n 0.4 0.35 0.3 0.25 0.2 0.15 0.1 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55 0.6 0.65 a n-1 Figure 6: Poincaré map corresponding to the Poincaré section in Fig. 5. maximum (or a minimum, for that matter). If we then plot the next maximum vs the last, we have next-amplitude map. The next-amplitude map has some practical applications, but is perhaps most famous because it was one of the devices used by Edward Lorenz in his famous study of the attractor which bears his name. We retrace Lorenz s steps here. The Lorenz equations 2 are dx dt dy dt dz dt = σ(y x), = x(r z) y, = xy bz. These equations are a radical simplication of equations describing convection in a fluid heated from below. Integrating these equations at appropriate parameter values gives the famous Lorenz butterfly attractor shown in Fig. 7(a). Since the chaotic solution shown in Fig. 7 was generated by deterministic differential equations, Lorenz thought there should be some regular trends within the data. For some reason, his attention fixed itself on the variable z. He intuited by looking at data such as those shown in Fig. 7(b) that the maximum in z might be a good predictor of what would happen on the next loop around the attractor. Based on this intuition, he constructed the first next-amplitude map, plotting one 2 Lorenz himself modestly calls this the three-variable model. 9

45 45 40 40 35 35 30 30 z 25 z 25 20 20 15 15 10 10 (a) 5-20 -15-10 -5 0 5 10 15 20 x (b) 5 450 455 460 465 470 475 480 485 490 495 500 t Figure 7: Lorenz attractor (a) and time series of z vs t for b = 2.67, σ = 10 and r = 28. maximum in z against the next. We can repeat this calculation in xpp by setting up a Poincaré section, but this time choosing the Max/min option. We set the Variable to be z. The Section value isn t used in searching for maxima, while a Direction of 1 indicates that we want maxima rather than minima ( 1) or both (0). If we then set up a Ruelle plot in the usual way and plot z max,n vs z max,n 1, we get the next-amplitude map shown in Fig. 8. In this case, the points don t exactly lie on a curve. They never do, but in this case it s obvious that the curves which make up our map have some thickness. However, we can again conclude that to a good approximation, the solutions of these differential equations define a map, this time roughly tent-shaped. 6 Concluding comments Whether we are interested in the dynamics of numerics, as the study of the dynamics induced by numerical methods is called, or in learning something about the chaotic solutions of certain differential equations, students of differential equations frequently end up studying maps. Maps also turn up directly as models of certain natural phenomena, particularly in population biology. In many ways, this is a fortunate thing because maps are easier to analyze than differential equations, as we will see in our next lecture. 10

48 46 44 42 z max,n 40 38 36 34 32 30 30 32 34 36 38 40 42 44 46 48 z max,n-1 Figure 8: Lorenz next-amplitude map for the attractor shown in Fig. 7. 11