A plane autonomous system is a pair of simultaneous first-order differential equations,

Similar documents
Nonlinear Autonomous Systems of Differential

STABILITY. Phase portraits and local stability

+ i. cos(t) + 2 sin(t) + c 2.

Math 266: Phase Plane Portrait

Section 9.3 Phase Plane Portraits (for Planar Systems)

Nonlinear dynamics & chaos BECS

2.10 Saddles, Nodes, Foci and Centers

4 Second-Order Systems

ENGI Duffing s Equation Page 4.65

1. < 0: the eigenvalues are real and have opposite signs; the fixed point is a saddle point

EN Nonlinear Control and Planning in Robotics Lecture 3: Stability February 4, 2015

1 The pendulum equation

Math 216 Final Exam 24 April, 2017

Math 312 Lecture Notes Linear Two-dimensional Systems of Differential Equations

Calculus and Differential Equations II

Chapter 6 Nonlinear Systems and Phenomena. Friday, November 2, 12

154 Chapter 9 Hints, Answers, and Solutions The particular trajectories are highlighted in the phase portraits below.

Models Involving Interactions between Predator and Prey Populations

MAS212 Assignment #2: The damped driven pendulum

The dynamics of the Forced Damped Pendulum. John Hubbard Cornell University and Université de Provence

Consider a particle in 1D at position x(t), subject to a force F (x), so that mẍ = F (x). Define the kinetic energy to be.

ENGI 9420 Lecture Notes 4 - Stability Analysis Page Stability Analysis for Non-linear Ordinary Differential Equations

EE222 - Spring 16 - Lecture 2 Notes 1

Problem Set Number 5, j/2.036j MIT (Fall 2014)

Solutions to Dynamical Systems 2010 exam. Each question is worth 25 marks.

Copyright (c) 2006 Warren Weckesser

Department of Mathematics IIT Guwahati

Nonlinear Control Lecture 2:Phase Plane Analysis

Chapter 13 Lecture. Essential University Physics Richard Wolfson 2 nd Edition. Oscillatory Motion Pearson Education, Inc.

Math 331 Homework Assignment Chapter 7 Page 1 of 9

Problem set 6 Math 207A, Fall 2011 Solutions. 1. A two-dimensional gradient system has the form

B5.6 Nonlinear Systems

Phase portraits in two dimensions

Def. (a, b) is a critical point of the autonomous system. 1 Proper node (stable or unstable) 2 Improper node (stable or unstable)

Lecture 38. Almost Linear Systems

7 Two-dimensional bifurcations

Nonlinear Control Lecture 2:Phase Plane Analysis

Chaotic motion. Phys 420/580 Lecture 10

Introduction to the Phase Plane

Chapter 7. Nonlinear Systems. 7.1 Introduction

Chaotic motion. Phys 750 Lecture 9

Theoretical physics. Deterministic chaos in classical physics. Martin Scholtz

Autonomous Systems and Stability

Linear and Nonlinear Oscillators (Lecture 2)

Nonlinear Oscillators: Free Response

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

Problem Sheet 1.1 First order linear equations;

Complex Dynamic Systems: Qualitative vs Quantitative analysis

ENGI Linear Approximation (2) Page Linear Approximation to a System of Non-Linear ODEs (2)

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

8 Example 1: The van der Pol oscillator (Strogatz Chapter 7)

Understand the existence and uniqueness theorems and what they tell you about solutions to initial value problems.

Oscillatory Motion. Simple pendulum: linear Hooke s Law restoring force for small angular deviations. small angle approximation. Oscillatory solution

Oscillatory Motion. Simple pendulum: linear Hooke s Law restoring force for small angular deviations. Oscillatory solution

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

Autonomous systems. Ordinary differential equations which do not contain the independent variable explicitly are said to be autonomous.

LECTURE 8: DYNAMICAL SYSTEMS 7

L = 1 2 a(q) q2 V (q).

Transitioning to Chaos in a Simple Mechanical Oscillator

Chapter 4: First-order differential equations. Similarity and Transport Phenomena in Fluid Dynamics Christophe Ancey

7 Planar systems of linear ODE

Math 4B Notes. Written by Victoria Kala SH 6432u Office Hours: T 12:45 1:45pm Last updated 7/24/2016

Problem set 7 Math 207A, Fall 2011 Solutions

Classification of Phase Portraits at Equilibria for u (t) = f( u(t))

MCE693/793: Analysis and Control of Nonlinear Systems

Differential Equations 2280 Sample Midterm Exam 3 with Solutions Exam Date: 24 April 2015 at 12:50pm

CDS 101/110a: Lecture 2.1 Dynamic Behavior

One Dimensional Dynamical Systems

ANSWERS Final Exam Math 250b, Section 2 (Professor J. M. Cushing), 15 May 2008 PART 1

The construction and use of a phase diagram to investigate the properties of a dynamic system

Chapter 4. Systems of ODEs. Phase Plane. Qualitative Methods

Section 5.4 (Systems of Linear Differential Equation); 9.5 Eigenvalues and Eigenvectors, cont d

L = 1 2 a(q) q2 V (q).

arxiv: v1 [physics.class-ph] 5 Jan 2012

Linearization of Differential Equation Models

CDS 101/110a: Lecture 2.1 Dynamic Behavior

28. Pendulum phase portrait Draw the phase portrait for the pendulum (supported by an inextensible rod)

1 Lyapunov theory of stability

Physics 235 Chapter 4. Chapter 4 Non-Linear Oscillations and Chaos

Control Systems. Internal Stability - LTI systems. L. Lanari

Math 215/255: Elementary Differential Equations I Harish N Dixit, Department of Mathematics, UBC

Coupled Oscillators Monday, 29 October 2012 normal mode Figure 1:

Homogeneous Constant Matrix Systems, Part II

Math 216 Final Exam 24 April, 2017

Stability of Dynamical systems

THE SEPARATRIX FOR A SECOND ORDER ORDINARY DIFFERENTIAL EQUATION OR A 2 2 SYSTEM OF FIRST ORDER ODE WHICH ALLOWS A PHASE PLANE QUANTITATIVE ANALYSIS

Solutions for B8b (Nonlinear Systems) Fake Past Exam (TT 10)

NONLINEAR DYNAMICS AND CHAOS. Numerical integration. Stability analysis

MAT 22B - Lecture Notes

Chapter 14 (Oscillations) Key concept: Downloaded from

CHALMERS, GÖTEBORGS UNIVERSITET. EXAM for DYNAMICAL SYSTEMS. COURSE CODES: TIF 155, FIM770GU, PhD

In which we count degrees of freedom and find the normal modes of a mess o masses and springs, which is a lovely model of a solid.

Econ 204 Differential Equations. 1 Existence and Uniqueness of Solutions

Lotka Volterra Predator-Prey Model with a Predating Scavenger

Kinematics of fluid motion

Nonlinear differential equations - phase plane analysis

Chapter 2 PARAMETRIC OSCILLATOR

Dynamical Systems Solutions to Exercises

Stability of Nonlinear Systems An Introduction

2 Lyapunov Stability. x(0) x 0 < δ x(t) x 0 < ɛ

Transcription:

Chapter 11 Phase-Plane Techniques 11.1 Plane Autonomous Systems A plane autonomous system is a pair of simultaneous first-order differential equations, ẋ = f(x, y), ẏ = g(x, y). This system has an equilibrium point (or fixed point or critical point or singular point) (x 0, y 0 ) when f(x 0, y 0 ) = g(x 0, y 0 ) = 0. We can illustrate the behaviour of the system by drawing trajectories (i.e., solution curves) in the (x, y)-plane, known in this context as the phase plane. The trajectories in such a phase portrait are marked with arrows to show the direction of increasing time. Note that trajectories can never cross, because the solution starting from any point in the plane is uniquely determined: so there cannot be two such solution curves starting at any given point. The only exception is at an equilibrium point (because the solution starting at an equilibrium point is just that single point, so it is no contradiction for two curves to meet there). 11.2 Classification of Equilibria We can examine the stability of an equilibrium point by setting x = x 0 + ξ, y = y 0 + η and using Taylor Series in 2D for small ξ and η: ξ = f(x 0 + ξ, y 0 + η) 75

= f(x 0, y 0 ) + ξ f x (x 0, y 0 ) + η f y (x 0, y 0 ) = ξf x (x 0, y 0 ) + ηf y (x 0, y 0 ) and similarly η = ξg x (x 0, y 0 ) + ηg y (x 0, y 0 ), where we have ignored terms of second order and higher. In matrix notation, ( ) ξ =, = J (11.1) η where J = ( f x g x f y g y ) (x0,y 0 ). Let the eigenvalues of this stability matrix J be λ 1, λ 2 with corresponding eigenvectors e 1, e 2. The general solution of (11.1) is = Ae λ 1t e 1 + Be λ 2t e 2 (11.2) where A, B are arbitrary constants. The behaviour of the solution therefore depends on the eigenvalues. Two real, positive eigenvalues (say λ 1 > λ 2 > 0). The solution for moves outwards in both the e 1 and e 2 directions (though more quickly in the e 1 direction). Therefore increases exponentially with time; trajectories move away from the equilibrium point (x 0, y 0 ). This is an unstable node. Two real, negative eigenvalues (say λ 1 < λ 2 < 0). In this case, decreases exponentially and trajectories move towards the equilibrium point. This is a stable node. The phase portrait is identical to that of an unstable node with the arrows reversed. Two real eigenvalues of opposite sign (say λ 1 < 0, λ 2 > 0). Trajectories move inwards along e 1 but outwards along e 2. Unless the initial value of lies exactly parallel to e 1, the solution will eventually move away from the equilibrium point, so it is unstable. This is a saddle point. 76

Complex eigenvalues (a conjugate pair, say α ± iβ). We can rewrite the general solution (11.2) in the form = e αt{ A ( cos(βt) + i sin(βt) ) e 1 + B ( cos(βt) i sin(βt) ) } e 2 = e αt( a cos(βt) + b sin(βt) ) where a = Ae 1 + Be 2, b = i(ae 1 Be 2 ). (Of course, a and b must be real, because is real at all times, even though e 1, e 2, A and B are complex.) Therefore the solution spirals around the equilibrium point (starting with a displacement in the direction of a at t = 0, then moving round to b, then a, then b, before getting back to a and starting again). If α > 0 then increases with each loop, so this is an unstable focus; if α < 0 then it is a stable focus. If α = 0 (i.e., if the eigenvalues are purely imaginary) then the solution just goes round and round a closed loop (an ellipse with semi-axes a and b): this is a centre. The sense of the rotation (i.e., either clockwise or anticlockwise) is best discovered on a case-by-case basis (usually by considering the sign of either f or g close to the equilibrium point). Note that e 1 and e 2 are not necessarily orthogonal. Similarly, for complex eigenvalues α ± iβ, a and b are not necessarily orthogonal; so in this case the diagrams above might need to be skewed. In the case of a centre (α = 0) this would result in a sheared ellipse: but a sheared ellipse is still an ellipse (albeit one with different axes of symmetry). The diagrams are qualitatively correct. Using the Trace and Determinant In practice, it is not necessary to find the exact values of the eigenvalues and eigenvectors; the signs of the eigenvalues (or of their real parts) are all that is required to perform the categorisation. The characteristic equation for J is (f x λ)(g y λ) f y g x = 0 or, equivalently, λ 2 T λ + = 0 where T = f x + g y is the trace and = f x g y f y g x the determinant of J. We note that the eigenvalues are real iff T 2 4 0; and that the product of the eigenvalues is while their sum is T. This enables us to deduce the required signs. 77

Summary of Results T 2 4 Eigenvalues of J Classification ve [+ve] Real, opposite signs Saddle { +ve +ve Real, same signs Node: T < 0 T > 0 T < 0 +ve ve Complex conjugate pair Focus: T = 0 T > 0 stable unstable stable centre unstable In the above analysis we have ignored the following degenerate cases: One of the eigenvalues is zero, which occurs iff = 0. In this situation we would have to consider second-order terms in the Taylor series, which we could neglect in the analysis above. These terms could either stabilise or destabilise the equilibrium point. The two eigenvalues are equal (and real), which occurs iff T 2 4 = 0. In this situation there may be two non-parallel eigenvectors, just as normal; or there may be only one eigenvector, in which case the general solution is not of the form given above. In either case, the classification still holds: the equilibrium point is an unstable node if λ 1 = λ 2 > 0 and a stable node if λ 1 = λ 2 < 0. The phase diagram looks somewhat different, however: a star node (in the case of two non-parallel eigenvectors) or an inflected node (in the case of only one). 11.3 The Phase Plane for a Conservative System A second-order differential equation for a variable x(t) can always be converted to two first-order differential equations by defining y = ẋ. For example, a general force equation in one dimension, can be converted to mẍ = F (x, ẋ), ẋ = y, ẏ = 1 F (x, y) m which is of the form given in 11.1 for a plane autonomous system. 78

Now consider a conservative force field in one dimension described by a potential V (x) so that F = V (x). Then ẋ = y, ẏ = 1 m V (x). Equilibrium points occur where V (x 0 ) = 0 and y 0 = 0 (so they are all on the x-axis). At such a point, ( ) 0 1 J =, V (x 0 )/m 0 so T = 0 and = V (x 0 )/m. Hence, using the summary table in the previous section, all equilibrium points must be either saddles (if V (x 0 ) < 0) or centres (if V (x 0 ) > 0). This analysis agrees with the stability analysis of 4.2, with the following phase plane diagrams locally (i.e., close to the equilibrium point): We also have the energy equation 1 2 my2 + V (x) = E. This equation, for different values of the constant E, defines the trajectory curves in the (x, y)-plane. Trajectories for this system are therefore symmetric in y (which enables us to draw the diagrams above, knowing that the directions given by the eigenvectors must also be reflectionally symmetric in the x-axis). On a given trajectory, at any value of x there are just two values of y, one positive and one negative. If a trajectory is bounded then it must have y = 0 at each end, at which points it joins up. Trajectories for a conservative system must therefore be either closed (i.e., they come back to where they started) or unbounded. Example: the Duffing oscillator for a particle of unit mass, ẍ = x x 3. This corresponds to V (x) = 1 2 x2 + 1 4 x4. We will have a saddle at x = 0 and centres at x = ±1. The phase portrait is as follows: 79

Example: a simple pendulum of length l (as in 2.5) has θ = g l sin θ. Let y = θ, so that θ = y, ẏ = g l sin θ. The stable equilibria are at θ = 2nπ and the unstable ones at θ = (2n + 1)π (n Z), by considering ( ) 0 1 J =. (g/l) cos θ 0 The phase diagram has this form: Around stable equilibria, the pendulum oscillates back and forth (e.g., curve A). This motion is known as libration. If the pendulum has sufficiently large energy then it can instead undergo rotation (e.g., curve C) where it always has the same sign of θ. The curves which join the saddle points (e.g., B) are known as separatrices because they separate the phase plane into regions containing these two different kinds of motion. They correspond physically to the (highly unlikely) motion where the pendulum starts vertically upwards then executes precisely one revolution, ending vertically upwards again. 80

11.4 Damped Systems Consider a simple pendulum with damping: θ = g l sin θ k θ (11.3) where k is a small positive constant. Letting y = θ we have θ = y, ẏ = g l sin θ ky. The equilibrium points are still where sin θ = 0; we have ( ) 0 1 J =. (g/l) cos θ k At θ = 2nπ, T = k and = g/l, corresponding to a stable focus (from the table in 11.2, assuming that k is small enough that k 2 < 4g/l). At θ = (2n + 1)π, T = k and = g/l so we have a saddle. Because this system is not conservative (the force depends on θ as well as θ, prohibiting the existence of a potential V (θ)), the solution curves are not symmetric in y and are not closed. The phase portrait is as follows: Note that if we define energy in the same way as for an undamped pendulum, E = 1 2 ml2 θ2 mgl cos θ, then de dt = ml2 θ θ + mgl θ sin θ = mkl θ 2 (from (11.3)) 0, 81

so the energy is decreasing. (Without the damping term we would have had Ė = 0.) This allows us to deduce the portrait above from the undamped portrait of 11.3, because the solution continuously moves to curves of lower energy. A more interesting example is the van der Pol oscillator ẍ = x (x 2 1)ẋ. For large x we have damping (because x 2 1 > 0); but for small x we have negative damping (x 2 1 < 0). The only equilibrium is at the origin and is an unstable focus. The phase portrait looks like this: So all trajectories (whether starting from small or large x, except for the one at the origin itself) tend towards a limit cycle. Therefore, after an initial transient the system always settles down into this finite amplitude oscillation. If we had studied only the linear stability near the origin we would have concluded that disturbances grow exponentially; the fact that this growth is in fact limited is useful in practical situations. 11.5 The Forced Pendulum Finally, add time-periodic forcing to the pendulum: θ = g l sin θ k θ + F cos Ωt where Ω is the forcing frequency and F its amplitude. This is no longer an autonomous system (unlike in 11.1), because t appears in the governing equation in addition to θ and θ. The phase plane becomes three-dimensional and consequently much more elaborate: for instance, trajectories can now twist around each other, which is prevented in two dimensions by the rule that trajectories cannot cross except at equilibrium points. When F is small, we obtain a resonant response as described in 2.2. But for larger forcing amplitudes the system exhibits much more complicated behaviour because the nonlinearity of the sin θ term interacts with the forcing. Eventually chaos results. 82

For small values of F, a graph of oscillation amplitude A versus forcing frequency Ω has a simple resonant peak. As F is increased, the oscillations of the pendulum become larger; the restoring force (g/l) sin θ is smaller than the linear approximation (g/l)θ that applies for small F, so the period of the oscillations increases. This means that the forcing frequency Ω required to produce resonance decreases for larger A, so the peak bends backwards. When F becomes larger still, the oscillation amplitude can take more than one value for a range of Ω. Suppose that we slowly increase Ω from small values: then the amplitude will move along the lower curve shown until it has to jump to the upper curve. But when Ω is slowly decreased again the amplitude moves back in a different way, sticking to the upper curve until it has to fall down to the lower one. This phenomenon, whereby the solution of a system depends not just on its parameters but also on the history of those parameters, is known as hysteresis. Many nonlinear systems also exhibit chaos, as does the forced pendulum for very large F. In a chaotic system, trajectories never settle into a limit cycle or other repeating pattern. Furthermore, the solutions are very sensitive to the initial conditions, so that if we start with two sets of conditions very close to each other (say 10 10 apart) at t = 0, within a relatively short time the solutions are a long way apart (several orders of magnitude greater, e.g., 1). This makes the solution effectively unpredictable, even on a very accurate computer. 83