Examples include: (a) the Lorenz system for climate and weather modeling (b) the Hodgkin-Huxley system for neuron modeling

Similar documents
The goal of this chapter is to study linear systems of ordinary differential equations: dt,..., dx ) T

Department of Mathematics IIT Guwahati

Linear Systems Notes for CDS-140a

7 Planar systems of linear ODE

In these chapter 2A notes write vectors in boldface to reduce the ambiguity of the notation.

MTH 464: Computational Linear Algebra

Linear Algebra Practice Problems

Linear ODEs. Existence of solutions to linear IVPs. Resolvent matrix. Autonomous linear systems

Math Ordinary Differential Equations

Stability of Dynamical systems

DIAGONALIZABLE LINEAR SYSTEMS AND STABILITY. 1. Algebraic facts. We first recall two descriptions of matrix multiplication.

MA5510 Ordinary Differential Equation, Fall, 2014

Chapter 7. Canonical Forms. 7.1 Eigenvalues and Eigenvectors

NORMS ON SPACE OF MATRICES

Lecture Notes for Math 524

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

Complex Dynamic Systems: Qualitative vs Quantitative analysis

Third In-Class Exam Solutions Math 246, Professor David Levermore Thursday, 3 December 2009 (1) [6] Given that 2 is an eigenvalue of the matrix

A = 3 1. We conclude that the algebraic multiplicity of the eigenvalues are both one, that is,

Section 9.3 Phase Plane Portraits (for Planar Systems)

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

MA 527 first midterm review problems Hopefully final version as of October 2nd

B5.6 Nonlinear Systems

Jordan normal form notes (version date: 11/21/07)

B. Differential Equations A differential equation is an equation of the form

2.10 Saddles, Nodes, Foci and Centers

Calculus and Differential Equations II

Differential Topology Final Exam With Solutions

21 Linear State-Space Representations

MCE693/793: Analysis and Control of Nonlinear Systems

Entrance Exam, Differential Equations April, (Solve exactly 6 out of the 8 problems) y + 2y + y cos(x 2 y) = 0, y(0) = 2, y (0) = 4.

8.1 Bifurcations of Equilibria

Problem set 7 Math 207A, Fall 2011 Solutions

Linear ODEs. Types of systems. Linear ODEs. Definition (Linear ODE) Linear ODEs. Existence of solutions to linear IVPs.

0.1 Rational Canonical Forms

Homogeneous Constant Matrix Systems, Part II

Econ 204 Differential Equations. 1 Existence and Uniqueness of Solutions

1. Let m 1 and n 1 be two natural numbers such that m > n. Which of the following is/are true?

4 Second-Order Systems

Differential Equations and Modeling

22.2. Applications of Eigenvalues and Eigenvectors. Introduction. Prerequisites. Learning Outcomes

NOTES ON LINEAR ODES

2. Review of Linear Algebra

av 1 x 2 + 4y 2 + xy + 4z 2 = 16.

Copyright (c) 2006 Warren Weckesser

Diagonalization. P. Danziger. u B = A 1. B u S.

Linear Algebra- Final Exam Review

Solution of Linear State-space Systems

Homogeneous Linear Systems of Differential Equations with Constant Coefficients

Chapter III. Stability of Linear Systems

JORDAN NORMAL FORM NOTES

Math 308 Final Exam Practice Problems

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

Control Systems. Dynamic response in the time domain. L. Lanari

Definition: An n x n matrix, "A", is said to be diagonalizable if there exists a nonsingular matrix "X" and a diagonal matrix "D" such that X 1 A X

1. The Transition Matrix (Hint: Recall that the solution to the linear equation ẋ = Ax + Bu is

= A(λ, t)x. In this chapter we will focus on the case that the matrix A does not depend on time (so that the ODE is autonomous):

Econ Lecture 14. Outline

Solutions to Math 53 Math 53 Practice Final

First Midterm Exam Name: Practice Problems September 19, x = ax + sin x.

SYSTEMTEORI - ÖVNING 1. In this exercise, we will learn how to solve the following linear differential equation:

Lecture Notes 6: Dynamic Equations Part C: Linear Difference Equation Systems

Dynamic interpretation of eigenvectors

Solutions of the Sample Problems for the Third In-Class Exam Math 246, Fall 2017, Professor David Levermore

Remark 1 By definition, an eigenvector must be a nonzero vector, but eigenvalue could be zero.

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

Remark By definition, an eigenvector must be a nonzero vector, but eigenvalue could be zero.

1. General Vector Spaces

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

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

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

Putzer s Algorithm. Norman Lebovitz. September 8, 2016

Math 302 Outcome Statements Winter 2013

Lecture 4. Chapter 4: Lyapunov Stability. Eugenio Schuster. Mechanical Engineering and Mechanics Lehigh University.

AN ELEMENTARY PROOF OF THE SPECTRAL RADIUS FORMULA FOR MATRICES

The following definition is fundamental.

We shall finally briefly discuss the generalization of the solution methods to a system of n first order differential equations.

27. Topological classification of complex linear foliations

6 Linear Equation. 6.1 Equation with constant coefficients

Online Exercises for Linear Algebra XM511

Problem List MATH 5173 Spring, 2014

MAT Linear Algebra Collection of sample exams

Quadratic forms. Here. Thus symmetric matrices are diagonalizable, and the diagonalization can be performed by means of an orthogonal matrix.

Systems of Linear Differential Equations Chapter 7

1.1 Limits and Continuity. Precise definition of a limit and limit laws. Squeeze Theorem. Intermediate Value Theorem. Extreme Value Theorem.

Archive of past papers, solutions and homeworks for. MATH 224, Linear Algebra 2, Spring 2013, Laurence Barker

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

Lecture 9. Systems of Two First Order Linear ODEs

LINEAR ALGEBRA QUESTION BANK

University of Sydney MATH3063 LECTURE NOTES

ECEN 605 LINEAR SYSTEMS. Lecture 7 Solution of State Equations 1/77

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

INNER PRODUCT SPACE. Definition 1

Systems of Linear ODEs

EIGENVALUES AND EIGENVECTORS 3

LINEAR ALGEBRA MICHAEL PENKAVA

DIAGONALIZATION. In order to see the implications of this definition, let us consider the following example Example 1. Consider the matrix

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

EK102 Linear Algebra PRACTICE PROBLEMS for Final Exam Spring 2016

AN ITERATION. In part as motivation, we consider an iteration method for solving a system of linear equations which has the form x Ax = b

Transcription:

1 Introduction Many natural processes can be viewed as dynamical systems, where the system is represented by a set of state variables and its evolution governed by a set of differential equations. Examples include: (a) the Lorenz system for climate and weather modeling (b) the Hodgkin-Huxley system for neuron modeling The purpose of this course is to develop necessary mathematical tools to facilitate the study of dynamical systems.

2 1 Linear Systems The goal of this chapter is to study linear systems of ordinary differential equations: ẋ = Ax, x(0) = x 0, (1) where x R n, A is an n n matrix and ẋ = dx ( dt = dx1 dt,..., dx ) T n. dt It will be shown that the unique solution of Eq. (1) is given by x(t) = e At x 0, where e At is an n n matrix defined by its Taylor series. A good portion of this chapter is concerned with the computation of e At in terms of the eigenvalues and eigenvectors of A.

3 1.1 Uncoupled Linear Systems Let s start the solution of the linear system (1) with the simplest case, where the system contains only one equation (i.e. n = 1): ẋ = ax, x(0) = c. The method of separation of variables immediately gives x(t) = c e at.

4 2 2 Uncoupled Systems: an Example To move one step forward, let s consider the following 2 2 uncoupled system: ẋ 1 = x 1, ẋ 2 = 2x 2, x 1 (0) = c 1, x 2 (0) = c 2. Its solution is easily found to be: x 1 (t) = c 1 e t, x 2 (t) = c 2 e 2t, or in matrix form: x(t) = [ e t 0 0 e 2t ] c =: e At c.

5 The General Case Clearly, the same procedure can be applied to solve uncoupled systems of any size. For example, the solution of the following 3 3 system: ẋ 1 = x 1, ẋ 2 = x 2, ẋ 3 = x 3, x 1 (0) = c 1, x 2 (0) = c 2, x 3 (0) = c 3. is given by x(t) = e t 0 0 0 e t 0 c =: eat c. 0 0 e t

6 Phase Plane Analysis Before wrapping up this section, let s introduce some notations that will be useful in the study of the linear system ẋ = Ax. Let s first consider the 2 2 system: [ 1 0 ẋ = Ax, A = 0 2 ]. Recall it has the solution: x(t) = [ e t 0 0 e 2t ] c. Clearly, the above formula describes the dynamics of each of the system components x 1 and x 2.

7 Phase Portrait In many cases, however, we are interested in the dynamics of the entire system x = (x 1, x 2 ) T besides that of the individual components x 1 or x 2. To gain a better understanding of the entire system, we eliminate the variable t from the solution representation so that a single formula involving only x 1 and x 2 results: x 2 = c2 1c 2 x 2 1 For any fixed c 1, c 2, the above equation defines a curve in the x 1 x 2 -plane the so-called phase plane. The set of all solution curves for all possible values of c 1 and c 2 constitutes a phase portrait of the linear system ẋ = Ax..

8 The Phase Portrait of the 2 2 System The phase portrait of the 2 2 system is shown below. Figure 1: The phase portrait of the 2 2 system.

9 Vector Field The direction of the motion of the solution x = (x 1, x 2 ) T along the solution curves can be read off from the explicit formula x 1 (t) = c 1 e t, x 2 (t) = c 2 e 2t. On the other hand, it can also be directly determined from the right-hand side [ ] 1 0 f(x) = Ax = 0 2 x of the system, which defines a vector field on the phase plane. The vector field must be tangent to the solution curves at every point x on the phase plane.

10 The Vector Field of the 2 2 System The vector field of the 2 2 system is shown below. Figure 2: The vector field of the 2 2 system.

11 The Phase Portrait of the 3 3 System Similar analysis can be carried out for more general linear systems. For example, the following figure shows the phase portrait of the 3 3 system ẋ = Ax where A = diag[1, 1, 1]: Figure 3: The phase portrait of the 3 3 system.

12 1.2 Diagonalization In the last section, we have seen how to solve uncoupled linear systems of the form ẋ = Ax = diag[λ 1,...,λ n ]x. The purpose of this and the following sections is to develop solution techniques for general, coupled linear system where the matrix A is not necessarily diagonal. The key is to reduce A to its diagonal form, or in more general situations, to its Jordan form.

13 Matrices with Real Distinct Eigenvalues Let s start with the simple case where A has real, distinct eigenvalues. The following theorem provides the basis for the solution of the linear system ẋ = Ax. Theorem 1 If the eigenvalues λ 1, λ 2,...,λ n of an n n matrix A are real and distinct, then any set of corresponding eigenvectors {v 1,v 2,...,v n } forms a basis for R n, the matrix P = [v 1,v 2,...,v n ] is invertible and P 1 AP = diag[λ 1, λ 2,...,λ n ]. The proof of the theorem can be found in any standard linear algebra text, for example Lowenthal [Lo].

14 Matrices with Real Distinct Eigenvalues (Cont d) Using the above theorem, we may solve the linear system ẋ = Ax by introducing the change of variable: y = P 1 x. It reduces the original system to an uncoupled linear system: ẏ = diag[λ 1,...,λ n ]y, and the solution of the original system can then be easily found: x(t) = PE(t)P 1 x(0), where E(t) is the diagonal matrix E(t) = diag [ e λ 1t,...,e λ nt ].

15 Example As an example, consider the linear system ẋ 1 = x 1 3x 2, ẋ 2 = 2x 2, x 1 (0) = c 1, x 2 (0) = c 2. Using the procedure described above, the solution is found to be x 1 (t) = c 1 e t + c 2 (e t e 2t ), x 2 (t) = c 2 e 2t.

16 Example (Cont d) The phase portrait of the above system is shown below. Figure 4: The phase portrait of the example.

17 Stable and Unstable Subspaces Note that the subspaces spanned by the eigenvectors v 1 and v 2 of the matrix A determine the stable and unstable subspaces of the linear system ẋ = Ax, according to the following definition. Definition 2 Suppose that the n n matrix A has k negative eigenvalues λ 1,...,λ k and n k positive eigenvalues λ k+1,...,λ n and that these eigenvalues are distinct. Let {v 1,...,v n } be a corresponding set of eigenvectors. Then the stable and unstable subspaces of the linear system, E s and E u, are the linear subspaces spanned by {v 1,...,v k } and {v k+1,...,v n } respectively; i.e., E s = span{v 1,...,v k }, E u = span{v k+1,...,v n }.

18 1.3 Exponentials of Operators In the last section, we have seen how to solve the linear system ẋ = Ax when A has real distinct eigenvalues, or more generally when A is diagonalizable. The purpose of this and the following sections is to study the general case where A is not necessarily diagonalizable. The key is to define the matrix exponential e At and verify the identity d dt eat = Ae At.

19 Matrices as Linear Operators We shall define e At through the Taylor series e At = k=0 1 k! Ak t k, but first we need to make sure that the series converges in appropriate norms. To introduce a norm (i.e. a measure ) for an n n matrix A, we view it as a linear operator T that maps an element in R n (i.e. an n-vector) to another element in R n : T : R n R n, T(x) = Ax. It can be shown that the converse is also true, i.e. any linear operator that maps R m to R n can be identified with an n m matrix. So matrices are indeed synonyms for linear operators.

20 Operator Norm For a linear operator T : R n R n, we define the operator norm: T = sup x =0 T(x), x where x denotes the Euclidean norm of x R n : x = x 2 1 + + x2 n. It can be readily verified that the operator norm has the following equivalent definitions: T = sup T(x) or T = sup T(x). x 1 x =1 Remark. The induced norm of the matrix representation A of the operator T is called the 2-norm of A.

21 Example To illustrate the concept of operator norm, let s compute A for the 2 2 matrix [ ] a b A =. b a For any x R 2, it is easy to see that Ax 2 = (a 2 + b 2 ) x 2. Hence the induced norm (or 2-norm) of A is given by A = a 2 + b 2. In general, it can be shown that the 2-norm of an m n matrix A is the largest singular value of A, i.e., A = λ max (A H A).

22 Properties of the Operator Norm The operator norm has all of the usual properties of a norm, namely for any linear operators S, T : R n R n, (a) T 0 and T = 0 iff T = 0 (positive definiteness) (b) at = a T for any a R (positive homogeneity) (c) S + T S + T (triangle inequality or subadditivity) It can be shown that the space L(R n ) of linear operators T : R n R n equipped with the norm is a complete normed space, or in other words, a Banach space. The convergence of a sequence of operators T k L(R n ) can then be defined in terms of the norm.

23 Convergence in Operator Norm Definition 3 A sequence of linear operators T k L(R n ) is said to converge to a linear operator T L(R n ) as k, i.e., lim T k = T, k if for any ǫ > 0, there exists an N such that T T k < ǫ for all k N. Now we can show that the infinite Taylor series e Tt = k=0 1 k! T k t k, converges in the operator norm.

24 The Operator Exponential e Tt Theorem 4 Given T L(R n ) and t 0 > 0, the series e Tt := k=0 1 k! T k t k is absolutely and uniformly convergent for all t t 0. Moreover, e Tt e Tt. To prove this theorem, we need the following lemma. Lemma 5 For S, T L(R n ) and x R n, (a) T(x) T x (b) TS T S (c) T k T k for k = 0, 1, 2,...

25 The Matrix Exponential e At By identifying an n n matrix A with a linear operator T L(R n ) via the relation T(x) = Ax, we may define the matrix exponential e At as follows. Definition 6 Let A be an n n matrix. Then for t R, e At is the n n matrix defined by the Taylor series e At = k=0 1 k! Ak t k. As will be shown later, the matrix exponential e At can be computed in terms of the eigenvalues and eigenvectors of A.

26 Properties of the Matrix Exponential We next establish some basic properties of the operator exponential e T in order to facilitate the computation of the corresponding matrix exponential e A. Proposition 7 If P and T are linear operators on R n and S = PTP 1, then e S = Pe T P 1. Corollary 8 If P 1 AP = diag[λ j ], then e At = P diag[e λ jt ]P 1. Proposition 9 If S and T are linear operators on R n which commute, i.e., ST = TS, then e S+T = e S e T = e T e S. Corollary 10 If T is a linear operator on R n, the inverse of the linear operator e T is given by (e T ) 1 = e T.

27 Properties of the Matrix Exponential (Cont d) Corollary 11 (Complex Conjugate Eigenvalues) If [ ] a b A =, b a then e A = e a [ cosb sinb sinb cosb ]. Corollary 12 (Nontrivial Jordan Block) If [ ] a b A =, 0 a then e A = e a [ 1 b 0 1 ].

28 Matrix Exponential for 2 2 Matrices It will be shown in Section 1.8 that, for any 2 2 matrix A, there is an invertible 2 2 matrix P (whose columns consist of generalized eigenvectors of A) such that the matrix B = P 1 AP has one of the following forms: [ ] [ ] [ ] λ 0 λ 1 a b B =, B =, or B =. 0 µ 0 λ b a It then follows that e At = Pe Bt P 1 where [ ] [ e λt e Bt 0 1 t =, e Bt = e λt 0 e µt 0 1 [ ] cosbt sinbt or e Bt = e at. sinbt cosbt ],

29 1.4 The Fundamental Theorem for Linear Systems In the last section, we have seen how to define the matrix exponential e At through the Taylor series e At = k=0 1 k! Ak t k. The purpose of this section is to verify the identity d dt eat = Ae At and solve the linear system ẋ = Ax in terms of the matrix exponential e At : x(t) = e At x(0).

30 The Fundamental Theorem We first compute the derivative of e At in the following lemma. Lemma 13 Let A be a square matrix, then d dt eat = Ae At. Now we are ready to formulate the main theorem. Theorem 14 Let A be an n n matrix. Then for a given x 0 R n, the initial value problem has a unique solution given by ẋ = Ax, x(0) = x 0, x(t) = e At x 0.

31 Example Solve the initial value problem [ ] 2 1 ẋ = x, x(0) = 1 2 [ 1 0 ], and sketch the solution curves in the phase plane R 2. By applying the fundamental theorem and the results from the last section, the solution is easily found to be [ ] cost x(t) = e 2t. sint

32 Example (Cont d) The phase portrait of the above system is shown below. Figure 5: The phase portrait of the example.

33 1.5 Linear Systems in R 2 Now we have learned how to solve the linear system ẋ = Ax in terms of the matrix exponential e At : x(t) = e At x(0), and the only task that remains is to compute e At for arbitrary square matrices A. Before moving on, we shall take a short discursion and discuss in this section the various phase portraits that are possible for linear systems in R 2, i.e. systems with 2 2 matrices A.

34 Reduction to Jordan Form Based on the results summarized at the end of Section 1.3, for any 2 2 matrix A, there is an invertible 2 2 matrix P such that the matrix B = P 1 AP has one of the following forms: [ ] [ ] [ ] λ 0 λ 1 a b B =, B =, or B =. 0 µ 0 λ b a It suffices to describe the phase portraits for the linear system ẏ = By since the phase portrait for ẋ = Ax can be obtained from that for ẋ = Bx under the linear transformation x = Py.

35 The Solution It follows from the fundamental theorem in Section 1.4 and the form of the matrix e Bt computed in Section 1.3 that the solution of ẏ = By with y(0) = y 0 is given by [ ] [ ] e λt 0 1 t y(t) = y 0, y(t) = e λt y 0, 0 e µt 0 1 [ cosbt sinbt or y(t) = e at sinbt cosbt We now list the various phase portraits that result from these solutions. ] y 0.

36 Case I: Saddle [ ] λ 0 Consider the case B = with λ < 0 < µ. The system 0 µ ẏ = By has a saddle at the origin in this case. Figure 6: A saddle at the origin.

37 Case II: Stable/Unstable Node [ ] λ 0 Consider the case B = with λ µ < 0 or B = 0 µ [ ] λ 1 with λ < 0. The system ẏ = By has a stable node 0 λ at the origin in these cases. If λ µ > 0 or λ > 0, the node is called an unstable node. Figure 7: A stable node at the origin.

38 Case III: Stable/Unstable Focus [ ] a b Consider the case B = with a < 0. The system b a ẏ = By has a stable focus at the origin in this case. If a > 0, the focus is called an unstable focus. Figure 8: A stable focus at the origin.

39 Case IV: Center [ 0 b Consider the case B = b 0 center at the origin in this case. ]. The system ẏ = By has a Figure 9: A center at the origin. Remark. If one (or both) of the eigenvalues of B is zero, i.e., if detb = 0, the origin is called a degenerate equilibrium point of the system ẏ = By.

40 Example As an example, consider the linear system [ ] 0 4 ẋ = Ax = x, x(0) = c. 1 0 It can be reduced to the Jordan form [ 0 2 ẏ = By = 2 0 ] y, and its solution can be found to be [ cos2t x(t) = 2 sin2t sin2t cos2t 1 2 ] c.

41 Example (Cont d) The phase portrait of the above system consists of concentric ellipses as shown in the following figure. Figure 10: The phase portrait of the example.

42 Classification of Linear Systems in R 2 Now we may classify the linear systems ẋ = Ax in R 2. Definition 15 The linear system ẋ = Ax is said to have a saddle, a node, a focus or a center at the origin if the matrix A is similar to one of the matrices B in Cases I, II, III or IV respectively. Remark. We say the matrix A is similar to the matrix B if there is a nonsingular matrix P such that P 1 AP = B. Furthermore, the direction of rotation of trajectories in the phase portraits for the systems ẋ = Ax and ẏ = By will be the same if detp > 0 (i.e., if P is orientation preserving) and it will be opposite if detp < 0 (i.e., if P is orientation reversing).

43 A Simple Criterion For deta 0, the following simple criterion can be used to determine the type of the linear system ẋ = Ax without reducing A to its Jordan form. Theorem 16 Let δ = deta and τ = tracea and consider the linear system ẋ = Ax. (a) If δ < 0 then the system has a saddle at the origin. (b) If δ > 0 and τ 2 4δ 0 then the system has a node at the origin; it is stable if τ < 0 and unstable if τ > 0. (c) If δ > 0, τ 2 4δ < 0, and τ 0 then the system has a focus at the origin; it is stable if τ < 0 and unstable if τ > 0. (d) If δ > 0 and τ = 0 then the system has a center at the origin.

44 The Bifurcation Diagram The above results can be summarized in a bifurcation diagram as shown in the following figure. Note that a stable node or focus of the linear system ẋ = Ax is called a sink and an unstable node or focus is called a source. Figure 11: A bifurcation diagram for the linear system ẋ = Ax.

45 1.6 Complex Eigenvalues Having studied the classification of the linear system ẋ = Ax in R 2, we return in this section to the computation of the matrix exponential e At for arbitrary square matrices A. We have learned how to compute e At for matrices A that (1) have real eigenvalues and (2) have a complete set of eigenvectors, in which case A is diagonalizable (in R). We still need to handle the case where A (1) has complex eigenvalues and/or (2) has an incomplete set of eigenvectors, in which case A is not diagonalizable (in R). In this section, we shall first handle the case where A has complex eigenvalues. This has been done for 2 2 systems but we need to generalize the results to arbitrary dimensions.

46 Matrices with Distinct Complex Eigenvalues Let s start with the special case where A has distinct, complex (only) eigenvalues. The following theorem provides the basis for the solution of the linear system ẋ = Ax (recall that complex eigenvalues occur in complex conjugate pairs). Theorem 17 If the 2n 2n real matrix A has 2n distinct complex eigenvalues λ j = a j +ib j and λ j = a j ib j and corresponding complex eigenvectors w j = u j + iv j, and w j = u j + iv j, j = 1,...,n, then {u 1,v 1,...,u n,v n } is a basis for R 2n, the matrix P = [v 1 u 1 v 2 u 2 v n u n ] is invertible and P 1 AP = diag [ aj b j b j a j ], a real 2n 2n matrix with 2 2 blocks along the diagonal.

47 Matrices with Distinct Complex Eigenvalues (Cont d) Using the above theorem, we may solve the linear system ẋ = Ax by introducing the change of variable y = P 1 x, as we have done before. Corollary 18 Under the hypotheses of the above theorem, the solution of the initial value problem ẋ = Ax, x(0) = x 0 is given by x(t) = P diag e a jt [ cosbj t sinb j t sinb j t cos b j t ] P 1 x 0.

48 Example Solve the initial value problem 1 1 0 0 1 1 0 0 ẋ = x, x(0) = x 0 0 3 2 0. 0 0 1 1 By applying the above theorem, the solution is found to be e t cost e t sint 0 0 e t sint e t cost 0 0 x(t) = x 0 0 e 2t (cos t + sint) 2e 2t sint 0. 0 0 e 2t sint e 2t (cos t sint)

49 Matrices with Distinct Mixed Eigenvalues In case A has both real and complex eigenvalues and they are distinct, we have the following result. Theorem 19 If A has distinct real eigenvalues λ j and corresponding eigenvectors v j, j = 1,...,k and distinct complex eigenvalues λ j = a j + ib j and λ j = a j ib j and corresponding eigenvectors w j = u j +iv j and w j = u j iv j, j = k +1,...,n, then the matrix P = [v 1 v k v k+1 u k+1 v n u n ] is invertible and P 1 AP = diag[λ 1,...,λ k, B k+1,...,b n ], B j = [ aj b j b j a j ].

50 Example Solve the initial value problem 3 0 0 ẋ = 0 3 2 x, x(0) = x 0. 0 1 1 By applying the above theorem, the solution is found to be e 3t 0 0 x(t) = 0 e 2t (cos t + sint) 2e 2t sint x 0. 0 e 2t sint e 2t (cost sint)

51 Example (Cont d) The phase portrait of the above system is shown below. Figure 12: The phase portrait of the example.