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

Similar documents
MATH 415, WEEK 11: Bifurcations in Multiple Dimensions, Hopf Bifurcation

MATH 353 LECTURE NOTES: WEEK 1 FIRST ORDER ODES

1 The pendulum equation

Nonlinear dynamics & chaos BECS

MATH 320, WEEK 2: Slope Fields, Uniqueness of Solutions, Initial Value Problems, Separable Equations

STABILITY. Phase portraits and local stability

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

MATH 116, LECTURE 13, 14 & 15: Derivatives

Why are Discrete Maps Sufficient?

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

MATH 319, WEEK 2: Initial Value Problems, Existence/Uniqueness, First-Order Linear DEs

Coordinate Curves for Trajectories

Math 266: Phase Plane Portrait

MATH 415, WEEKS 14 & 15: 1 Recurrence Relations / Difference Equations

Math 232, Final Test, 20 March 2007

3 Algebraic Methods. we can differentiate both sides implicitly to obtain a differential equation involving x and y:

2.10 Saddles, Nodes, Foci and Centers

LECTURE 8: DYNAMICAL SYSTEMS 7

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

Nonlinear Autonomous Systems of Differential

Nonlinear Oscillators: Free Response

Lecture 6, September 1, 2017

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

September Math Course: First Order Derivative

Math 312 Lecture Notes Linearization

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

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

Predator - Prey Model Trajectories and the nonlinear conservation law

4.9 Anti-derivatives. Definition. An anti-derivative of a function f is a function F such that F (x) = f (x) for all x.

The Liapunov Method for Determining Stability (DRAFT)

8.3 Partial Fraction Decomposition

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

Mass on a Horizontal Spring

AP Calculus (BC) Chapter 10 Test No Calculator Section. Name: Date: Period:

Updated 2013 (Mathematica Version) M1.1. Lab M1: The Simple Pendulum

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

Learning Objectives for Math 165

Problem set 7 Math 207A, Fall 2011 Solutions

Daba Meshesha Gusu and O.Chandra Sekhara Reddy 1

Compound Damped Pendulum: An Example

Math 234 Exam 3 Review Sheet

Computers, Lies and the Fishing Season

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

MATH 320, WEEK 4: Exact Differential Equations, Applications

Copyright (c) 2006 Warren Weckesser

CDS 101 Precourse Phase Plane Analysis and Stability

Question: Total. Points:

MCE693/793: Analysis and Control of Nonlinear Systems

Practice Problems for Final Exam

Lecture 10. (2) Functions of two variables. Partial derivatives. Dan Nichols February 27, 2018

Parametric Equations, Function Composition and the Chain Rule: A Worksheet

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

Half of Final Exam Name: Practice Problems October 28, 2014

Physics 351, Spring 2017, Homework #4. Due at start of class, Friday, February 10, 2017

Physics 141 Energy 1 Page 1. Energy 1

4 Partial Differentiation

Math 212-Lecture 8. The chain rule with one independent variable

Computer Problems for Methods of Solving Ordinary Differential Equations

MATH 521, WEEK 2: Rational and Real Numbers, Ordered Sets, Countable Sets

FIRST-ORDER SYSTEMS OF ORDINARY DIFFERENTIAL EQUATIONS III: Autonomous Planar Systems David Levermore Department of Mathematics University of Maryland

Slope Fields: Graphing Solutions Without the Solutions

Systems of Linear ODEs

Autonomous Systems and Stability

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

Complex Differentials and the Stokes, Goursat and Cauchy Theorems

Chapter 4. Oscillatory Motion. 4.1 The Important Stuff Simple Harmonic Motion

Infinite series, improper integrals, and Taylor series

Calculus Review Session. Brian Prest Duke University Nicholas School of the Environment August 18, 2017

Limit. Chapter Introduction

Substitutions and by Parts, Area Between Curves. Goals: The Method of Substitution Areas Integration by Parts

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

Basic Theory of Dynamical Systems

VANDERBILT UNIVERSITY. MATH 2300 MULTIVARIABLE CALCULUS Practice Test 1 Solutions

McGill University April Calculus 3. Tuesday April 29, 2014 Solutions

FINAL EXAM MATH303 Theory of Ordinary Differential Equations. Spring dx dt = x + 3y dy dt = x y.

SYDE 112, LECTURE 7: Integration by Parts

2t t dt.. So the distance is (t2 +6) 3/2

Physics 132 3/31/17. March 31, 2017 Physics 132 Prof. E. F. Redish Theme Music: Benny Goodman. Swing, Swing, Swing. Cartoon: Bill Watterson

8.1 Bifurcations of Equilibria

Differential Equations Handout A

Calculus I Review Solutions

AIMS Exercise Set # 1

3. On the grid below, sketch and label graphs of the following functions: y = sin x, y = cos x, and y = sin(x π/2). π/2 π 3π/2 2π 5π/2

Stability of Nonlinear Systems An Introduction

for changing independent variables. Most simply for a function f(x) the Legendre transformation f(x) B(s) takes the form B(s) = xs f(x) with s = df

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

Integrals. D. DeTurck. January 1, University of Pennsylvania. D. DeTurck Math A: Integrals 1 / 61

Differential Equations

Complex Dynamic Systems: Qualitative vs Quantitative analysis

Math 2930 Worksheet Final Exam Review

Sin, Cos and All That

MATH 320, WEEK 6: Linear Systems, Gaussian Elimination, Coefficient Matrices

Antiderivatives and Indefinite Integrals

Exam 2 Study Guide: MATH 2080: Summer I 2016

Chapter 9 Global Nonlinear Techniques

Coordinate systems and vectors in three spatial dimensions

Higher Mathematics Course Notes

Maps and differential equations

2.2 Separable Equations

Lecture 7 - Separable Equations

Transcription:

MATH 415, WEEKS 7 & 8: Conservative and Hamiltonian Systems, Non-linear Pendulum Reconsider the following example from last week: dx dt = x y dy dt = x2 y. We were able to determine many qualitative features of this system based on the vector field diagram and linear stability analysis about the fixed points (0, 0) and (1, 1). We determined that (0, 0) behaved locally like a saddle point and were able to get a sense of the long-term behavior of solutions by considering how the nullclines partitioned the state space. There were still, however, several important questions which we were not able to answer. In particular, linearization about the second fixed point (1, 1) yielded the strictly complex eigenvalues λ 1,2 = ±i. Since the real part of the eigenvalues was zero, we were not able to apply the Hartman- Grobman Theorem to guarantee that the local picture (a center) extended to the nonlinear system. In fact, it is still possible for the point to locally behaved like a spiral source, a spiral sink, or a center. We would like to determine which is the case. To address this, we take a slight detour. We will develop a more comprehensive theory as this week progresses, but for now let s just jump to the punchline and consider the function H(x, y) = x3 3 (1) + xy y2 2. (2) An important first observation to make about (2) is that, just like potentials for one-dimensional systems, the function H above assigns to each point (x, y) in the state space a numerical value. For one-dimensional systems, this was not a tremendous simplification of the problem, since our state space was one-dimensional to begin with. For two-dimensional (and higher!) systems, however, this will turn out to be a significant advantage. We perform the same analysis we did with one-dimensional potentials. That is, we ask the question of how the value of H changes along solutions 1

x(t) = (x(t), y(t)) of (1). To compute this, we carefully apply the chain rule to get dh(x(t)) dt = dh dx dx dt + dh dy dy dt = ( x 2 + y)(x y) + (x y)(x 2 y) = 0. This computation tells us that the value of H(x, y) does not change along solutions x(t) of (1). This is quite remarkable! For instance, supposing we start our trajectory at a state x(0) which has the value H(x(0)) = 5, we know that the trajectory may only ever pass through points also corresponding to the value of 5. We will never encounter a time where H(x(t)) = 4, or H(x(t)) = 6, or even H(x(t)) = 5.000001. To make the mathematical consequences of this more clear, we integrate with respect to t to obtain dh(x(t)) dt = (0) dt = H(x(t)) = C (3) dt where C is some constant. We should recognize (3) from multivariate calculus (e.g. Math 234) as it corresponds to the equation for a level curve of the multivariate function H. A collection of level curves for various values of C is called a contour plot. We have determined is that every solution x(t) of (2) remains on the level curve corresponding to the initial condition x(0) = x 0. For this example, those level curves are difficult to obtain analytically, but we can easily plot them with the aid of a computer (see Figure 1). We make the following notes: 1. We will say that any nontrivial function H(x, y) which has the property (3) along solutions of a system is a conservation function, and that the corresponding system as a conservative system. 2. We can determine that (1, 1) is a local maximum of H(x, y) by computing the Hessian at (1, 1): 2 H x 2 2 H x y 2 H x y 2 H y 2 x=1,y=1 = [ 2x 1 1 1 ] x=1,y=1 = [ 2 1 1 1 ]. 2

Figure 1: The level curves of the function H(x, y) = x 3 /3 + xy y 2 /2 (solid lines) trap the trajectories of the system (1). To test for a maximum or minimum we check [ 2 H 2 ( H 2 )] x 2 y 2 H = ( 2)( 1) (1) 2 = 1 > 0 x y x=1,y=1 and 2 H = 2 < 0. x2 It follows that(1, 1) is a local maximum of H(x, y) by the secondderivative test for multivariate functions. This is important since it guarantees that the system behaves like a center near (1, 1). 3. Although solutions of (1) are restricted to the level curves of H(x, y), we do not know which direction they go along these curves. For this information, we need to consider the original vector field. 4. If we are careful, we can divide the long-term behavior of solutions in the following way: (a) There is a closed loop which begins at (0, 0) (from the right), wraps around (1, 1), and then ends at (0, 0) (from the top). This type of unusual curve is called a homoclinic orbit. (b) Inside the homoclinic orbit, all solutions form closed curves. That is, every solution is periodic (in a nonlinear way). 3

(c) Outside the homoclinic orbit, almost all solutions drift off to infinity in the second quadrant. The lone exception is the stable manifold in quadrant three which divides those curves which divert to the left of (0, 0) and those which divert to the right. (This fragile curve actually converges to (0, 0).) This is a very interesting result, but we have only considered one example. We have no idea how robust/common the phenomenon of having a conserved quantity is or how we might systematically search for one. As we will see in the next few sections, this phenomenon is surprisingly robust, and there are several hints that a system can given us as to whether there is such a function and, if so, what it is. 1 State-Dependent Mechanical Systems We do not have to look very far in order to determine a classification of systems which necessarily conservative. Reconsider the model for a physical system obeying Newton s second law Force = mass acceleration. In general, the force function may depend upon x, x, and t, so that we have the system of equations mx = F (t, x, x ) (4) where m is the mass of the physical object being modeled. Systems of the form (4) do not always admit a conservation function; however, a very broad subclass of such systems does. Suppose we can write mx = F (x). (5) We will call such systems state-dependent mechanical systems, since the force function depends only upon the state x. A simple example of a state-dependent mechanical system which we have already considered is the pendulum/spring system with a restoring force kx but no friction. This gave us the second-order system mx = kx = mx + kx = 0. This system is linear and can be solved explicitly to get the periodic solution x(t) = C 1 sin( k/m t) + C 2 cos( k/m t). We will consider a more physically realistic pendulum model shortly. 4

Now consider the following argument. In analogy with potentials for one-dimensional systems, we rewrite (5) in the form mx = dv dx (6) where V (x) is an unknown function. That is, we take F (x) = V (x). We now apply a trick which will seem very unusual, but will produce the desired result almost directly. We rearrange (5) and multiply the whole expression by x. This gives mx x + x dv dx = 0. This trick should not feel intuitive, but the key feature is that we can actually integrate this expression. Using the chain rule in reverse, we realize that we have d [ m dt 2 (x ) 2 + V (x)] = 0. In other words, if we define E(x, x ) = m 2 (x ) 2 +V (x), we have that the value of E(x, x ) does not change along trajectories of (5). So E(x, x ) is exactly the conservation function we have been looking for! The function E(x, x ) should be familiar to those well-versed in physics. The term m 2 (x ) 2 corresponds to the kinetic energy of the system while the term V (x) corresponds to the potential energy of the system. This result tells us that, for state-dependent mechanical systems, any increase in the kinetic, or potential, energy must be offset by a corresponding decrease in the other. In other words, the total energy E is conserved. Example 1: Determine the total energy function E for the undamped pendulum/spring model mx = kx. (7) Solution: The argument above holds for all systems of the form (5), so we do not need to go through it all over again. It suffices to determine the potential energy function V (x). We have that V (x) = kx = V (x) = k x2 2. It follows that the total energy function is given by E(x, x ) = m 2 (x ) 2 + k 2 x2. 5

Figure 2: The level curves of the function E(x, x ) = m 2 (x ) 2 + k 2 x2 trap the trajectories of the system (7). The system above is pictured with m = k = 1. If we look at the level sets of this function in the (x, x )-plane (which is the same as making the substitutions x 1 = x, x 2 = x ) we can see that trajectories must move along ovals (see Figure 2). This should not come as a surprise. This system is linear, so we could transform it into a system and see that the corresponding coefficient matrix has strictly complex eigenvalues. This guarantees that the system is a center. If were are unconvinced by argument, we can check explicitly that d dt E(x, x ) = d dt [ m 2 (x ) 2 + k 2 x2 = mx x + kx x = x [mx + kx] = 0 by the original differential equation. It follows that, as expected, the total energy of the system is conserved. Example 2: Determine the total energy function E for the statedependent mechanical system ] x = x x 3. (8) Solution: As before, because the system is state-dependent only, it suffices to find the potential energy V (x). We have V (x) = F (x) dx = (x 3 x) dx = x4 4 x2 2. 6

It follows that the required energy function is E(x, x ) = (x ) 2 2 + x4 4 x2 2. The level curves of E, which correspond to solutions of the original system, are shown in the (x, x )-plane in Figure 3. Figure 3: The level curves of the function E(x, x ) = (x ) 2 x2 2 2 + x4 4 trap the trajectories of the system (8). The fixed points at ( 1, 0) and (1, 0) are nonlinear centers while the fixed point at (0, 0) is a nonlinear saddle. Notice the homoclinic orbits which extend from (0, 0), wrap around (1, 0) or ( 1, 0), and then return to (0, 0). 2 Nonlinear Pendulum We noted at the beginning of the course that the model mx = kx (9) for an undamped pendulum was at best an approximation. It captures the idea that if the pendulum is positioned to the right of its center, the force will be to the left, and that if it is positioned to the left, the force will be to the right. It does not, however, capture the dynamical behavior far from the center. For instance, this model does not permit the pendulum to swing over the top as we would expect of a realistic physical model. In order to accommodate this more exotic behavior associated with pendulums, we need to consider the nonlinear model mθ = k sin(θ) (10) 7

where θ R is the angle the pendulum makes with the vertical axis (that is, θ = 0 corresponds to a pendulum hanging straight down). This relationship can be derived by considering the relevant forces. For our purposes, we will just be interested in the observation that, when θ 0, this system behaves like the linear system because sin(θ) θ for small θ. For large θ, however, this model has behavior not characterized by the linear model (9). The nonlinear system (10) is certainly more challenging to analyze than the linear system (9). The methods of Math 319 and 320 do not apply to (10) in fact, there is no way to determine the explicit solution. Instead we will rely on the qualitative methods we have been developing. First of all, we rewrite this in the first-order system form dx 1 dt = x 2 dx 2 dt = k m sin(x 1). (11) where x 1 (t) = θ(t) and x 2 (t) = θ (t). We now consider the vector field and linear stability analysis. We start by identifying the nullclines and fixed points of the system. We have x 1 = 0 = x 2 = 0 and x 2 = 0 = k m sin(x 1) = 0 = x 1 = nπ, n Z. It follows that we have the fixed points ( x 1, x 2 ) = (nπ, 0), n Z. Accounting for vector flow in the bounded regions, we arrive at the general picture contained in Figure 4. We now consider linear stability analysis. We have so that and [ Df(x 1, x 2 ) = Df(2nπ, 0) = 0 1 k m cos(x 1) 0 [ 0 1 k m 0 [ ] 0 1 Df((2n + 1)π, 0) = ] k m 0 where, for both cases, we take n Z. The eigenvalues of Df(2nπ, 0) are λ = ±i which does not yield any information as a result of linearization. The 8 ]

x 2 '=0 x 2 '=0 x 2 '=0 x 1 '=0 Figure 4: Phase portrait in the (x 1, x 2 )-plane (i.e. (θ, θ )-plane) of the system (11) with k/m = 1. The nullclines x 1 = 0 (red) and x 2 = 0 (blue) are overlain. eigenvalue/eigenvectors pairs for Df((2k+1)π, 0), however, are λ 1 = k/m, v 1 = (1, k/m), and λ = k/m, v 2 = (1, k/m). We therefore have saddles at each point ((2n+1)π, 0) with the unstable manifold corresponding to a positively sloped curve through the point, and a stable manifold corresponding to a negatively sloped curve. This is consistent with our picture in Figure 4. A number of things are now clear. Solutions with high values of x 2 (i.e. high velocity) travel to the right, while solutions with low values (i.e. high negative velocity) travel to the left. These cases correspond to pendulums which spin over the top in one direction or the other. For solutions with small x 2 (i.e. low velocity), trajectories swirl from the left to the right of a given fixed point. This corresponds to pendulums wobbling about their equilibrium position. There is still, however, something mysterious about the fixed point ( x 1, x 2 ) = ((2n + 1)π, 0). Because linear stability analysis failed, we have not determined whether these fixed points correspond to spiral sources, spiral sinks, or a nonlinear center. All three are consistent with the rough picture contained in Figure 4. To answer this question, we consider the previous discussion of energy functions and conserved quantities. While the extension to the nonlinear system (10) has changed many properties of the system, and destroyed our ability to solve it explicitly, it is still a state-dependent mechanical system since the force function depends only upon our state variable θ. All of our previous analysis applies! We need only find the potential function V (x), 9

which is given by V (x) = k sin(θ) dθ = k cos(θ). We therefore have that the value of the energy function E(θ, θ ) = m 2 (θ ) 2 k cos(θ) (12) is conserved along solutions of (10). We can check this explicitly, if we are unconvinced. We have de dt = mθ θ + k sin(θ)θ = θ [ mθ + k sin(θ) ] = 0. The level curves of (12) is once again not something we plot by hand, but our computers will happily do it for us (see Figure 5). We can now see that the previously unclassified fixed points are nonlinear centers. For low amplitudes, trajectories just oscillate around these fixed points, which corresponds to the pendulum swinging back and forth like a grandfather clock. Figure 5: Level curves of the energy function E(x, x ) = m 2 (x ) 2 cos(x) overlain on the vector field plot of (11) reverted back to the (x, x )-plane. 3 Hamiltonian Systems As nice as state-dependent mechanical systems are, they do not capture every system for which there is a conservation relation. For example, we determined earlier that the function H(x, y) = x3 3 10 + xy y2 2 was conserved

along solutions of the system dx dt = x y dy dt = x2 y. This system is not a state-dependent mechanical system, so how was the conserved quantity determined? There is another very general class of systems which guarantees the existence of a conserved quantity (which can be thought of as an energy, or just as a general function, depending on your preferred point of view). They are called Hamiltonian systems. As with conservative systems, the basis for Hamiltonian systems comes from conserved quantities in physics and, in particular, Hamiltonian mechanics. We will not be interested in delving too deeply into those waters. We have the following. Definition 3.1. Consider a two-dimensional autonomous system of differential equations. We will say the system is Hamiltonian if there exists a function H(x, y) such that dx dt = H y dy dt = H x The function H(x, y) is called a Hamiltonian function. (13) Theorem 3.1. The Hamiltonian function is conserved along solutions x(t) of a Hamiltonian system. Proof. This can be computed directly by the chain rule. We have dh(x(t)) dt = H dx x = H x dt + H y ( H y dy dt ) + H y ( H ) = 0. x What is even better is that, for two-dimensional systems, it turns out to be very easy to check whether a system is Hamiltonian or not. We have the following result. 11

Theorem 3.2. A two-dimensional system is Hamiltonian if and only if f 1 x + f 2 y = 0. Proof. We will start by rewriting the general nonlinear system in the form (13). That is, we consider dx dt = f 1(x, y) = H y dy dt = f 2(x, y) = H x. This is a notational association, since we do not know that such a function H(x, y) exists. We now consider equality of mixed-order partial derivatives which must hold if a (sufficiently smooth) function H(x, y) exists: 2 H x y = 2 H y x. This can be rearranged to give ( ) H = ( ) H x y y x x f 1(x, y) = y f 2(x, y). After rearranging we have the desired condition, and we are done. We can now verify that our previous example is a Hamiltonian system, and also determine the Hamiltonian function H(x, y). We have f 1 (x, y) = x y and f 2 (x, y) = x 2 y so that f 1 x + f 2 y = x (x y) + y (x2 y) = 1 + ( 1) = 0. It follows by Theorem 3.2 that the system is Hamiltonian, but it remains to determine the Hamiltonian function H(x, y). To accomplish this, we need to solve the system H y = f 1(x, y) = x y H x = f 2(x, y) = x 2 y. There are a number of ways to accomplish this. The most straight-forward is generally to integrate one expression and then taking the corresponding 12

partial derivative required to match it with the second. The trick will be to be careful with the variable dependencies. We compute H(x, y) = H(x, y) dy = y (x y) dy = xy y2 2 + g(x) (14) where g(x) is a general unknown function which depends on x alone. Taking the partial derivative of (14) gives H x = y + g (x). Comparing this expression with the earlier expression gives g (x) = x 2 = g(x) = x3 3. Substituting this into the earlier form for H(x, y) gives H(x, y) = x3 3 + xy y2 2 as expected, and we are done. Example: Show that the following system is Hamiltonian and determine the Hamiltonian function: dx dt = ln( y ) dy dt = x2 1. (15) Solution: We apply the test of Theorem 3.2. We have df 1 dx + df 2 dy = d dx [ln( y )] + d dy [x2 1] = (0) + (0) = 0. Although this should seem trivial, it is still sufficient to guarantee that the system is Hamiltonian. To determine the Hamiltonian function H(x, y), we consider the system of equations H y = ln( y ) H x = x2 1. (16) 13

To illustrate the flexibility allowed at this step, we integrate the second equation to get H(x, y) = (1 x 2 ) dx = x x3 3 + g(y). To match this with the remaining equation in (16) we differentiate with respect to y. We have H y = g (y) = ln( y ) = g(y) = y ln( y ) y. It follows that the Hamiltonian function is H(x, y) = x x3 3 + y ln( y ) y. (17) As nice as having this conserved quantity is, it is really only the beginning of the story. We would like to use this information to say something about the dynamical behavior, but the function H(x, y) is too complicated to imagine graphing by hand. Instead, we go to our favorite computer graphic software package, and obtain the contour plot contained in Figure 6. We can now see very clearly that the fixed points (1, 1) and ( 1, 1) are saddles and the fixed points (1, 1) and ( 1, 1) are nonlinear centers. Each center has a clearly demarcated region of influence, outside of which solutions seem to slingshot around the fixed points. If you look closely, you can see one particularly interesting solution which is whipped around both centers before eventually being thrown off in the first quadrant (although we might wonder what really happens when it passes through y = 0 since this is not strictly allowed by (15). Figure 6: Level curves of the Hamiltonian function H(x, y) (17) overlain on the vector field plot of (15). 14