8.2 The Jordan Canonical Form

In this section, we will show that any linear operator L on an n-dimensional vector space V can be represented by a block diagonal matrix whose diagonal blocks are simple Jordan matrices. We will apply this result to solving systems of linear differential equations of the form Y=AY, where A is defective.

Let us begin by considering the case where L has more than one distinct eigenvalue. We wish to show that if L has distinct eigenvalues λ1,,λk, then V can be decomposed into a direct sum of invariant subspaces S1,,Sk such that Lλi is nilpotent on Si for each i=1,,k. To do this, we must first prove the following lemma and theorem.

Lemma 8.2.1

If L is a linear operator mapping an n-dimensional vector space V into itself, then there exists a positive integer k0 such that ker(Lk0)=ker(Lk0+k) for all k>0.

Proof

If i<j, then clearly ker(Li) is a subspace of ker(Lj). We claim that if ker(Li)=ker(Li+1) for some i, then ker(Li)=ker(Li+k) for all k1. We will prove this by induction on k. In the case k=1, there is nothing to prove. Assume that for some k>1 the result holds all indices less than k. If vker(Li+k), then

0=Li+k(v)=Li+k1(L(v))

Thus, L(v)ker(Li+k1). By the induction hypothesis, ker(Li+k1)=ker(Li). Therefore, L(v)ker(Li) and hence vker(Li+1). Since ker(Li+1)=ker(Li), it follows that vker(Li) and hence ker(Li)=ker(Li+k). Thus, if ker(Li+1)=ker(Li) for some i, then

ker(Li)=ker(Li+1)=ker(Li+1)=

Since V is finite dimensional, the dimension of ker(Lk) cannot keep increasing as k increases. Thus for some k0, we must have dim(ker(Lk0))=dim(ker(Lk0+1)) and hence ker(Lk0) and ker(Lk0+1) must be equal. It follows that

ker(Lk0)=ker(Lk0+1)=ker(Lk0+2)=

Theorem 8.2.2

If L is a linear transformation on an n-dimensional vector space V, then there exist invariant subspaces X and Y such that V=XY, L is nilpotent on X, and L[Y] is invertible.

Proof

Choose k0 to be the smallest positive integer such that ker(Lk0)=ker(Lk0+1). It follows from Lemma 8.2.1 that ker(Lk0)=ker(Lk0+j) for all j1. Let X=ker(Lk0). Clearly, X is invariant under L for if xX, then L(x)ker(Lk01), which is a proper subspace of ker(Lk0). Let Y=R(Lk0). If wXY, then w=Lk0(v) for some v and hence

0=Lk0(w)=Lk0(Lk0(v))=L2k0(v)

Thus, vker(L2k0)=ker(Lk0) and hence

w=Lk0(v)=0

Therefore, XY={0}. We claim V=XY. Let {x1,,xr} be a basis for X and let {y1,,ynr} be a basis for Y. By Lemma 8.2.1, it suffices to show that x1,,xr,y1,,ynr are linearly independent and hence form a basis for V. If

α1x1++αrxr+β1y1++βnrynr=0
(1)

then applying Lk0 to both sides gives

β1Lk0(y1)++βnrLk0(ynr)=0

or

Lk0(β1y1++βnrynr)=0

Therefore, β1y1++βnrynrXY and hence

β1y1++βnrynr=0

Since the yi’s are linearly independent, it follows that

β1=β2==βnr=0

and hence (1) simplifies to

α1x1++αrxr=0

Since the xi’s are linearly independent, it follows that

α1=α2==αr=0

Thus, x1,,xr,y1,,ynr are linearly independent and therefore V=XY. L is invariant and nilpotent on X. We claim that L is invariant and invertible on Y. Let yY; then y=Lk0(v) for some vV. Thus,

L(y)=L(Lk0(v))=Lk0+1(v)=Lk0(L(v))

Therefore, L(y)Y and hence Y is invariant under L. To prove L[Y] is invertible, it suffices to show that

ker(L[Y])=Yker(L)={0}

This, however, follows immediately since ker(L)X and XY={0}.

We are now ready to prove the main result of this section.

Theorem 8.2.3

Let L be a linear operator mapping a finite dimensional vector space V into itself. If λ1,,λk are the distinct eigenvalues of L, then V can be decomposed into a direct sum

X1X2Xk

such that Lλi is nilpotent on Xi and the dimension of Xi equals the multiplicity of λi.

Proof

Let L1=Lλ1. By Theorem 8.2.2, there exist subspaces X1 and Y1 that are invariant under L1 such that V=X1Y1,L1 is nilpotent on X1, and L1[Y] is invertible. It follows that X1 and Y1 are also invariant under L. By Corollary 8.1.2, L[X1] can be represented by a block diagonal matrix A1, where diagonal blocks are simple Jordan matrices whose diagonal elements all equal λ1. Thus,

det(A1λI)=(λ1λ)m1

where m1 is the dimension of X1. Let B1 be a matrix representing L[Y1]. Since L1 is invertible on Y1, it follows that λ1 is not an eigenvalue of B1. Thus,

det(B1λI)=q(λ)

where q(λ1)0. It follows from Lemma 8.1.2 that the operator L on V can be represented by the matrix

A=[A1B1]

Thus, if each eigenvalue λi of L has multiplicity ri, then

(λ1λ)r1(λ2λ)r2(λkλ)rk=det(AλI)=det(AλI)det(B1λI)=(λ1λ)m1q(λ)

Therefore, r1=m1 and

q(λ)=(λ2λ)r2(λkλ)rk

If we consider the operator L2=Lλ2 on the vector space Y1, then we can decompose Y1 into a direct sum X2Y2 such that X2 and Y2 are invariant under L,L2 is nilpotent on X2, and L[Y2] is invertible. Indeed, we can continue this process of decomposing Yi into a direct sum Xi+1Yi+1 until we obtain a direct sum of the form

V=X1X2Xk1Yk1

The vector space Yk1 will be of dimension rk with a single eigenvalue λk. Thus, if we set Xk=Yk1, then Lλk will be nilpotent on Xk and we will have the desired decomposition of V.

It follows from Theorem 8.2.3 that each operator L mapping an n-dimensional vector space V into itself can be represented by a block diagonal matrix of the form

J=[A1A2Ak]

where each Ai is an ri×ri block diagonal matrix (ri=multiplicityofλi) whose blocks consist of simple Jordan matrices with λi’s along the main diagonal.

If A is an n×n matrix, then A represents the operator LA with respect to the standard basis on Rn, where LA is defined by

LA(x)=Axfor eachxRn

By the preceding remarks, LA can be represented by a matrix J of the form just described. It follows that A is similar to J. Thus, each n×n matrix A with distinct eigenvalues λ1,,λk is similar to a matrix J of the form

J=[A1A2Ak]
(2)

where Ai is an ri×ri matrix (ri=multiplicity ofλi) of the form

Ai=[J1(λi)J2(λi)Js(λi)]
(3)

with the J(λi)’s being simple Jordan matrices. The matrix J defined by (2) and (3) is called the Jordan canonical form of A. The Jordan canonical form of a matrix is unique except for a reordering of the simple Jordan blocks along the diagonal.

Example 1 1

Find the Jordan canonical form of the matrix

A=[3101131011410213101141012]

SOLUTION

The characteristic polynomial of A is

|AλI|=λ4(1λ)

The eigenspace corresponding to λ=1 is spanned by x1=(1,1,1,1,2)T and the eigen-space corresponding to λ=0 is spanned by x2=(1,1,0,1,1)T and x3=(0,0,1,0,0)T. Thus, the Jordan canonical form of A then will consist of three simple Jordan blocks. Except for a reordering of the blocks, there are only two possibilities:

designer icon

To determine which of these forms is correct, we compute (A0I)2=A2.

A2=[1000110001100011000120002]

Next we consider the systems

A2x=xi

for i=2,3. Since these systems turn out to be inconsistent, the Jordan canonical form of A cannot have any 3×3 simple Jordan blocks and, consequently, it must be of the form

designer icon

To find X, we must solve

Ax=xi

for i=2,3. The system Ax=x2 has infinitely many solutions. We need choose only one of these, say, x4=(1,3,0,0,1)T. Similarly, Ax=x3 has infinitely many solutions, one of which is x5=(1,0,0,2,1)T. Let

X=[x1x2x3x4x5]=[1101111030101001100221011]

The reader may verify that X1AX=J.

One of the main applications of the Jordan canonical form is in solving systems of linear differential equations that have defective coefficient matrices. Given such a system

Y(t)=AY(t)

we can simplify it by using the Jordan canonical form of A. Indeed, if A=XJX1, then

Y=(XJX1)Y

Thus, if we set Z=X1Y, then Y=XZ and the system simplifies to

XZ=XJZ

Multiplying by X1, we get

Z=JZ
(4)

Because of the structure of J, this new system is much easier to solve. Indeed, solving (4) will only involve solving a number of smaller systems, each of the form

z1=λz1+z2z2=λz2+z3zk1=λzk1+zkzk=λzk

These equations can be solved one at a time starting with the last. The solution to the last equation is clearly

zk=ceλt

The solution to any equation of the form

z(t)λz(t)=u(t)

is given by

z(t)=eλteλtu(t)dt

Thus, we can solve

zk1λzk1=zk

for zk1 and then solve

zk2λzk2=zk1

for zk2, etc.

Example 2

Solve the initial value problem

[y1y2y3y4]=[1001011001121021][y1y2y3y4]y1(0)=y2(0)=y3(0)=y4(0)=2

SOLUTION

The coefficient matrix A has two distinct eigenvalues λ1=0 and λ2=2, each of multiplicity 2. The corresponding eigenspaces are both dimension 1. Using the methods of this section, A can be factored into a product XJX1, where

J=[0100000000210002]

The choice of X is not unique. The reader may verify that the one we have calculated:

X=[1111111110101010]

does the job. If we now change variable and set Z=X1Y, then we can rewrite the system in the form

Z=JZ

The block structure of J allows us to break up the system into two simpler systems:

z1=z2z2=0andz3=2z3+z4z4=2z4

The first system is not difficult to solve.

z1=c1t+c2z2=c1(c1andc2are  constants)

To solve the second system, we first solve

z4=2z4

getting

z4=c3e2t

Thus,

z32z3=c3e2t

and hence

z3=e2te2t(c3e2t)dt=e2t(c3t+c4)

Finally, we have

Y=XZ=[(c1t+c2)c1(c3t+c4)e2t+c3e2t(c1t+c2)+c1(c3t+c4)e2tc3e2t(c1t+c2)+(c3t+c4)e2t(c1t+c2)+(c3t+c4)e2t]

If we set t=0 and use the initial conditions to solve for the ci’s, we get

c1=1,c2=c3=c4=1

Thus, the solution to the initial value problem is

y1=tte2ty2=t+te2ty3=1+t+(1+t)e2ty4=1t+(1+t)e2t

The Jordan canonical form not only provides a nice representation of an operator, but it also allows us to solve systems of the form Y=AY even when the coefficient matrix is defective. From a theoretical point of view, its importance cannot be questioned. As far as practical applications go, however, it is generally not very useful.

If n5, it is usually necessary to calculate the eigenvalues of A by some numerical method. The calculated λi’s are only approximations to the actual eigenvalues. Thus, we could have calculated values λ1 and λ2, which are unequal while actually λ1=λ2. So in practice, it may be difficult to determine the correct multiplicity of the eigenvalues. Furthermore, in order to solve Y=AY, we need to find the similarity matrix X such that A=XJX1. However, when A has multiple eigenvalues, the matrix X may be very sensitive to perturbations and, in practice, one is not guaranteed that the entries of the computed similarity matrix will have any digits of accuracy whatsoever. A recommended alternative is to compute the matrix exponential eA and use it to solve the system Y=AY.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset
18.190.159.10