0% found this document useful (0 votes)
9 views22 pages

Final Strecth

mm

Uploaded by

girnarevaishali8
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
9 views22 pages

Final Strecth

mm

Uploaded by

girnarevaishali8
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 22

MTH 311: Advanced Linear Algebra

Semester 1, 2024-2025

Prahlad Vaidyanathan (substituting for Rohit Holkar)


I. Canonical Forms
1. Review
Definition 1.1. A matrix A = (ai,j ) ∈ Mn (F ) is said to be upper triangular if

ai,j = 0

whenever i > j.

Proposition 1.2. Let T ∈ L(V ) and B = {v1 , v2 , . . . , vn } be an ordered basis of V .


Then, the matrix [T ]B is upper triangular if and only if

T (vk ) ∈ span{v1 , v2 , . . . , vk }

for all 1 ≤ k ≤ n.

Definition 1.3.

(i) A matrix A ∈ Mn (F ) is said to be block diagonal if it can be written in the form


 
A1 0 0 ... 0 0
 0 A2 0 ... 0 0 
 
A =  0 0 A3 ... 0 0
 

 .. .. 
 . . 
0 0 0 . . . 0 Ak

where each Ai is an square matrix.

(ii) An operator T ∈ L(V ) is said to be block diagonalizable if there is an ordered


basis B of V such that [T ]B is block diagonal.

Note: For convenience, we may sometimes write such a matrix as


 
Ar1 ×r1 0 0
A= 0 Br2 ×r2 0 
0 0 Cr3 ×r3

where A ∈ Mr1 (F ), B ∈ Mr2 (F ) and C ∈ Mr3 (F ).

2
Definition 1.4. Let V be a vector space and W1 , W2 , . . . , Wk be subspaces of V . We
say that V is a direct sum of {W1 , W2 , . . . , Wk } if every v ∈ V can be expressed uniquely
in the form
v = w1 + w2 + . . . + wk
where wi ∈ Wi for all 1 ≤ i ≤ k. If this happens, we write
k
M
V = W1 ⊕ W2 ⊕ . . . ⊕ Wk = Wi .
i=1

Proposition
Lk 1.5. Let V be a vector space and W1 , W2 , . . . , Wk be subspaces of V . Then,
V = i=1 Wi if and only if the following two conditions are met:

(i) V = W1 + W2 + . . . + Wk .

(ii) For each 1 ≤ i ≤ k,

Wi ∩ (W1 + W2 + . . . + Wi−1 + Wi+1 + . . . + Wk ) = {0}.

Corollary 1.6. For T ∈ L(V ), there is an ordered basis B of V such that A = [T ]B is


block diagonal if and only if there are T -invariant subspaces W1 , W2 , . . . , Wk such that
k
M
V = Wi .
i=1

Lk
Corollary 1.7. If V = i=1 Wi , then
k
X
dim(V ) = dim(Wi ).
i=1

Moreover, if Bi is a basis for Wi , then the {Bi : 1 ≤ i ≤ k} are mutually disjoint and

B = tki=1 Bi

is a basis for V .

2. Diagonalizable and Triangulable Operators


Definition 2.1.

(i) An operator T ∈ L(V ) is said to be diagonalizable if there is a basis B such that


[T ]B is diagonal.

Note: [T ]B is diagonal if and only if every vector in B is an eigenvector.

3
(ii) A matrix A ∈ Mn (F ) is said to be diagonalizable if it is similar to a diagonal
matrix.
Proposition 2.2. Let T ∈ L(V ) and λ1 , λ2 , . . . , λk be distinct eigenvalues of T . If
v1 , v2 , . . . , vk are any corresponding eigenvectors, then {v1 , v2 , . . . , vk } is linearly inde-
pendent.
Corollary 2.3. If V is an n-dimensional vector space, then any operator T ∈ L(V ) has
atmost n distinct eigenvalues.
Proof. Any linearly independent set can have atmost n elements.
Corollary 2.4. Let V be an n-dimensional vector space. If T ∈ L(V ) has n distinct
eigenvalues, then T is diagonalizable.
Proof. Write c1 , c2 , . . . , cn for the eigenvalues and v1 , v2 , . . . , vn for some (any) corre-
sponding eigenvectors. Then, B = {v1 , v2 , . . . , vn } is linearly independent. Since dim(V ) =
n, it must be a basis.
Example 2.5.
(i) Let V = R2 and T ∈ L(V ) be the linear transformation

T (x, y) = (−y, x).

In other words, if B denotes the standard ordered basis for V , then


 
0 −1
A := [T ]B =
1 0

We claim that T is not diagonalizable. Suppose it was, then there would be a


basis B 0 = {v, w} consisting of eigenvectors of T . Now, the eigenvalues of T are
the roots of the polynomial
 
−x −1
fT (x) = fA (x) = det(A − xI) = det = x2 + 1.
1 −x

This polynomial does not have any roots in R, so T cannot be diagonalizable.

(ii) If V := C2 , however and T ∈ L(V ) is given exactly as above, then T is diagonal-


izable because it has two distinct eigenvalues i and −i.

(iii) Let V = R2 and T ∈ L(V ) be the linear transformation

T (x, y) = (y, 0).

In other words, if B denotes the standard ordered basis for V , then


 
0 1
A := [T ]B =
0 0

4
We claim that T is not diagonalizable. Suppose it was, then there would be a
basis B 0 = {v, w} consisting of eigenvectors of T . Now, the eigenvalues of T are
the roots of the polynomial
 
−x 1
fT (x) = fA (x) = det(A − xI) = det = x2 .
0 −x

Hence, the only eigenvalue of T is zero. In particular, it must happen that T (v) =
0 = T (w). In that case,  
0 0
B := [T ]B0 =
0 0
In turn, this would imply that A is similar to the zero matrix. But since A 6= 0,
this is impossible. Therefore, T is not diagonalizable.

Remark 2.6. By Corollary 2.4, if an operator has n distinct eigenvalues, then it is


diagonalizable. Indeed, the matrix is of the form
 
λ1 0 0 ... 0
0 λ2 0 . . . 0 
A =  ..
 
.. .. .. .. 
. . . . .
0 0 0 . . . λn

where the λi are all distinct. However, this is not necessary because it can happen that
the eigenvalues of a diagonal matrix are repeated. In that case, we would have a matrix
of the form  
λ1 I1 0 0 ... 0
 0 λ2 I2 0 . . . 0 
A =  ..
 
.. .. .. .. 
 . . . . . 
0 0 0 . . . λk Ik
where Ij are identity matrices of various sizes.

(i) In this case, consider the characteristic polynomial of A.


 
(x − λ1 )I1 0 0 ... 0
 0 (x − λ2 )I2 0 . . . 0 
det(xI − A) = det 
 
.. .. .. .. .. 
 . . . . . 
0 0 0 . . . (x − λk )Ik
= (x − λ1 )d1 (x − λ2 )d2 . . . (x − λk )dk

So,
cT (x) = cA (x) = (x − λ1 )d1 (x − λ2 )d2 . . . (x − λk )dk

5
(ii) Moreover, consider the matrix
 
0d1 0 0 ... 0
 0 (λ1 − λ2 )I2 0 ... 0 
λ1 I − A =  ..
 
.. .. .. .. 
 . . . . . 
0 0 0 . . . (λ1 − λk )Ik

Clearly,
B1 = {e1 , e2 , . . . , ed1 } ⊂ ker(λ1 I − A)
Moreover, if X = (x1 , x2 , . . . , xn ) ∈ ker(λ1 I − A), then

(λ1 − λ2 )xd1 +1 = 0 ⇒ xd1 +1 = 0.

Similarly, xj = 0 for all j > d1 . Hence, X = (x1 , x2 , . . . , xd1 , 0, 0, . . . , 0). In other


words,
ker(λ1 I − A) = span{B1 }.
In particular, dim(λ1 I − A) = d1 . Similarly, we see that

dim(ker(λi I − A)) = di

for all 1 ≤ i ≤ k.

(End of Day 1)

Definition 2.7. Let T ∈ L(V ) and λ ∈ F be an eigenvalue of T .

(i) The algebraic multiplicity of λ is the number of times (x − λ) occurs as a factor


of cT (x), the characteristic polynomial of T .

(ii) The geometric multiplicity of λ is the dimension of ker(λI − T ).

We denote these by a(T, λ) and g(T, λ) respectively. We may define these terms for a
matrix A ∈ Mn (F ) as well, in an analogous fashion.

Lemma 2.8. If  
Ar×r Cr×(n−r)
X=
0 B(n−r)×(n−r)
is a block upper triangular matrix, then

det(X) = det(A) det(B).

Proof.

6
(i) Suppose first that A = Ir×r . Then, by expanding along the first column

det(X) = det(X 0 )

where
0
 
0 I(r−1)×(r−1) C(r−1)×(n−r)
X =
0 B(n−r)×(n−r)
By induction on r, it follows that det(X) = det(B).
(ii) Now suppose B = In−r×n−r , then the same argument shows that det(X) = det(A).
 
Ar×r Cr×(n−r)
(iii) For the general case, suppose X = . We consider two cases:
0 B(n−r)×(n−r)
(a) If det(A) = 0, then A is not invertible. So the columns of A are linearly
dependent. In that case, the columns of X are also linearly dependent, so X
is not invertible. Hence, det(X) = 0 as well.
(b) If det(A) 6= 0, then A is invertible. In that case,

Ir×r A−1 C
  
A 0
X=
0 In−r×n−r 0 B

Since det is multiplicative, we use part (i) and (ii) to obtain

Ir×r A−1 C
   
A 0
det(X) = det det = det(A) det(B).
0 In−r×n−r 0 B

Lemma 2.9. For any T ∈ L(V ) and eigenvalue λ ∈ F , g(T, λ) ≤ a(T, λ).
Proof. Suppose r = g(T, λ) = dim(ker(λI − T )), then there is a basis B of ker(λI − T )
with |B| = r. This is a linearly independent set in V , so it may be extended to form a
basis B 0 of V . Consider
A = [T ]B0
and observe that A has the form
 
λIr×r Cr×(n−r)
A=
0(n−r)×r Dn−r×(n−r)
Hence,  
(xI − λI)r×r Cr×n−r
(xI − A) =
0 (xI − D)
To the characteristic polynomial of A (by Lemma 2.8) is

cT (x) = cA (x) = (x − λ)r det(xI − D).

Hence, r ≤ a(T, λ) as required.

7
Theorem 2.10. An operator T ∈ L(V ) is diagonalizable if and only if the following two
conditions hold:

(i) The characteristic polynomial of T is of the form

cT (x) = (x − λ1 )d1 (x − λ2 )d2 . . . (x − λk )dk

(i.e. The characteristic polynomial splits).

(ii) For each 1 ≤ i ≤ k, g(T, λi ) = a(T, λi ).

Proof.

(i) If T is diagonalizable, then the two conditions hold by Remark 2.6.

(ii) Conversely, suppose these two conditions hold. Write Wi = ker(T −λi I). We claim
that
M k
V = Wi .
i=1

To do this, we verify the two conditions of Proposition 1.5.


(a) We first verify the second condition: Fix 1 ≤ i ≤ k, and suppose

v ∈ Wi ∩ (W1 + W2 + . . . + Wi−1 + Wi+1 + . . . + Wk ).

Then write v = w1 + w2 + . . . + wi−1 + wi+1 + . . . + wk with wj ∈ Wj . More-


over, we may assume that all these terms are non-zero (otherwise, drop them).
However, each such vector is an eigenvector associated to a different eigen-
value, so the set {v, w1 , w2 , . . . , wi−1 , wi+1 , . . . , wk } is linearly independent by
Proposition 2.2. This is impossible unless v = wj = 0. Hence,

Wi ∩ (W1 + W2 + . . . + Wi−1 + Wi+1 + . . . + Wk ) = {0}.

(b) Now consider W = W1 + W2 + . . . + Wk . This is a subspace of V . Moreover,


by part (a),
Mk
W = Wi .
i=1

By Corollary 1.7,
k
X k
X
dim(W ) = dim(Wi ) = di = deg(cT (x)) = n = dim(V ).
i=1 i=1

Hence, W = V . Therefore,
k
M
V = Wi .
i=1

8
Now choose a basis Bi for each Wi , and set
k
G
B= Bi .
i=1

This is a basis of V and [T ]B is diagonal.

Example 2.11. Let T ∈ L(R3 ) be the linear operator represented in the standard basis
by the matrix  
5 −6 −6
A = −1 4 2
3 −6 −4

(i) We first compute the characteristic polynomial of A: cA (x) = det(xI − A)


 
(x − 5) 6 6
det  1 x − 4 −2 
−3 6 x+4

Subtracting column 3 from column 2 gives us a new matrix with the same deter-
minant. Hence,
   
x−5 0 6 x−5 0 6
f = det  1 x − 2 −2  = (x − 2) det  1 1 −2 
−3 2 − x x + 4 −3 −1 x + 4

Now adding row 2 to row 3 does not change the determinant, so


 
x−5 0 6
cA (x) = (x − 2) det  1 1 −2 
−2 0 x+2

Expanding along column 2 now gives


 
x−5 6
cA (x) = (x − 2) det = (x − 2)(x2 − 3x + 2) = (x − 2)2 (x − 1)
−2 x + 2

(ii) Hence, the characteristic values of T are 1 and 2. Moreover,

a(A, 2) = 2 and a(A, 1) = 1.

(iii) We wish to determine the dimensions of the characteristic spaces, W1 and W2 .

9
(a) Consider the case c = 1, and the matrix
 
4 −6 −6
(A − I) = −1 3 2
3 −6 −5
Row reducing this matrix gives
   
4 −6 −6 4 −6 −6
0 3 1 
7→ 0 32 1 
=: B
2 2 2
−3 −1
0 2 2
0 0 0
Hence, rank(A − I) = 2 so by the Rank-Nullity theorem,
g(A, 1) = 1.

(b) Consider the case c = 2: We know that


rank(A − 2I) ≥ 1
since (A − 2I) is non-zero. Furthermore, we know that
2 + rank(A − 2I) = rank(A − I) + rank(A − 2I) ≤ dim(R3 ) = 3
So it follows that rank(A − 2I) = 1. As before, this implies that
g(A, 2) = 2.

Hence, we conclude by Theorem 2.10 that T is diagonalizable. Indeed, A is similar


to the matrix  
1 0 0
D = 0 2 0
0 0 2
(iv) We now determine a basis consisting of characteristic vectors:
(a) Consider the case c = 1: Using the matrix B above, we solve the system of
linear equations BX = 0. This gives a solution
α1 = (3, −1, 3)

(b) Consider the case c = 2, and the matrix


 
3 −6 −6
(A − 2I) = −1 2 2
3 −6 −6
Row reducing gives  
3 −6 −6
C = 0 0 0
0 0 0
So solving the system CX = 0 gives two solutions
α2 = (2, 1, 0) and α3 = (2, 0, 1)

10
(v) Thus, we get an ordered basis

B = {(3, −1, 3), (2, 1, 0), (2, 0, 1)}

consisting of characteristic vectors of T . Furthermore,


 
1 0 0
[T ]B = 0 2 0 := D
0 0 2

(vi) Furthermore, if P is the matrix


 
3 2 2
P = −1 1 0
3 0 1

Then
P −1 AP = D

(End of Day 2)

Example 2.12. Let V = R3 and T ∈ L(V ) be the map T (v) = A(v) where
 
3 1 −1
A = 2 2 −1
2 2 0

(i) Now, cT (x) = cA (x) and


 
3−x 1 −1
cA (x) = − det(A − xI) = − det  2 2 − x −1 
2 2 −x
= (x − 3)[(2 − x)(−x) + 2] + 2[(−x) + 2] − 2[−1 + (2 − x)]
= x3 − 5x2 + 8x − 4
= (x − 1)(x − 2)2 .

So a(A, 1) = 1 and a(A, 2) = 2. Therefore, λ = 1 and λ = 2 are the eigenvalues of


T.

(ii) Since g(A, 1) ≤ 1 and g(A, 1) 6= 0, it follows that

g(A, 1) = 1.

11
(iii) To find eigenvectors associated to c = 2: We need to solve the equation (A −
2I)(X) = 0. Observe that
 
1 1 −1
(A − 2I) = 2
 0 −1
2 2 −2

Writing X = (x, y, z), we wish to solve the system

x+y−z =0
2x − z = 0
2x + 2y − 2z = 0

This yields that X is a scalar multiple of β = (1, 1, 2). Hence,

g(A, 2) = 1.

In particular, g(A, 2) = 1 < a(A, 2). Hence, A is not diagonalizable.


Definition 2.13.
(i) An operator T ∈ L(V ) is said to be triangulable if there is a basis B such that
[T ]B is upper triangular.

(ii) A matrix A in Mn (F ) is said to be triangulable if it is similar to an upper triangular


matrix.
Remark 2.14. Suppose T ∈ L(V ) is triangulable, then there is a basis B such that
 
a1,1 ∗ ∗ ... ∗ ∗
 0 a2,2 ∗ ... ∗ ∗ 
A = [T ]B =  
 0 0 a3,3 ... ∗ ∗ 
0 0 0 . . . 0 an,n

is upper triangular. Therefore,

fT (x) = fA (x) = det(xI − A) = (x − a1,1 )(x − a2,2 ) . . . (x − an,n ).

Therefore, the entries {a1,1 , . . . , an,n } are the eigenvalues of T . Combining like terms,
we may write
fT (x) = (x − λ1 )d1 (x − λ2 )d2 . . . (x − λk )dk .
Theorem 2.15. An operator T ∈ L(V ) in triangulable if and only if the characteristic
polynomial of T is of the form

fT (x) = (x − λ1 )d1 (x − λ2 )d2 . . . (x − λk )dk .

(i.e. the characteristic polynomial splits).

12
Proof. If T is triangulable, then fT (x) splits by Remark 2.14.

Conversely, suppose the characteristic polynomial splits as above. We induct on n :=


dim(V ).

(i) If n = 1, there is nothing to prove.

(ii) If n > 1, assume that the result is true for any vector space over F with dimension
≤ n − 1. Now suppose fT (x) splits, then T has an eigenvalue λ ∈ F . Let x1 ∈ V
be an eigenvector corresponding to λ and let

W := hx1 i

Then, W is a T -invariant subspace, so T induces an operator

T̂ : V /W → V /W given by x + W 7→ T (x) + W.

Moreover, cT̂ (x) divides cT (x), so cT̂ (x) also splits over F . Note that

dim(V /W ) = dim(V ) − dim(W ) ≤ n − 1.

So by the induction hypothesis, V /W has a basis B 0 = {x2 +W, x3 +W, . . . , xn +W }


such that [T̂ ]B0 is upper triangular. In other words, for each 2 ≤ k ≤ n,

T̂ (xk + W ) ∈ span{x2 + W, x3 + W, . . . , xk + W }.

Fix 2 ≤ k ≤ n and scalars {αk,2 , αk,3 . . . , αk,k } so that

T̂ (xk + W ) = αk,2 (x2 + W ) + αk,3 (x3 + W ) + . . . + αk,k (xk + W ).

Since T̂ (xk + W ) = T (xk ) + W , it follows that


k
X
T (xk ) − αk,i xi ∈ W = hx1 i
i=2

So there exists αk,1 ∈ F such that


k
X
T (xk ) = αk,i xi ∈ span{x1 , x2 , . . . , xk }
i=1

This is true for each 2 ≤ k ≤ n, and T (x1 ) ∈ span{x1 }, so T is triangulable.

Corollary 2.16. If V is a vector space over C, then every T ∈ L(V ) is triangulable.

(End of Day 3)

13
3. The Jordan Canonical Form
Definition 3.1. For t ∈ N and λ ∈ F , define Jt (λ) ∈ Mt (F ) to be the matrix
 
λ 1 0 ... 0 0
0 λ 1 . . . 0 0
 
 .. .. .. .. .. .. 
Jt (λ) =  . . . . . . 
 
0 0 0 . . . λ 1
0 0 0 ... 0 λ

Such a matrix is called Jordan matrix (or block) of size t.


The main theorem we wish to discuss is the following.
Theorem 3.2 (Jordan Canonical Form). Let V be a complex vector space and T ∈ L(V ).
Let {λ1 , λ2 , . . . , λk } be the set of distinct eigenvalues of T . Then, there is a basis B of
V such that  
A1 0 0 . . . 0
 0 A2 0 . . . 0 
[T ]B =  ..
 
.. .. .. .. 
 . . . . . 
0 0 0 . . . Ak
is a block diagonal matrix, and each Ai is a block diagonal matrix of the form
 
Jti1 (λi ) 0 0 ... 0
 0 Jti2 (λi ) 0 . . . 0 
Ai =  ..
 
.. .. .. .. 
 . . . . . 
0 0 0 . . . Jtisi (λi )

Moreover, for each 1 ≤ i ≤ k.

ti1 + ti2 + . . . + tisi = dim(Vλi ) = a(T, λi ) and


si = dim(ker(T − λi I)) = g(T, λi ).

Such a matrix is called a Jordan matrix, or the Jordan form of T .


Corollary 3.3. Let V be a complex vector space and T ∈ L(V ). Let {λ1 , λ2 , . . . , λk }
be the set of eigenvalues of T . If a(T, λi ) = g(T, λi ) for all 1 ≤ i ≤ k, then T is
diagonalizable.
Proof. Consider the Jordan Canonical form of A as above. Then,

ti1 + ti2 + . . . + tisi = dim(Vλi ) = a(T, λi ) and


si = dim(ker(T − λi I)) = g(T, λi ).

So if g(T, λi ) = a(T, λi ), it follows that ti1 = ti2 = . . . = tisi = 1. Hence, each Ai is a


diagonal matrix, so A is a diagonal matrix as well.

14
Example 3.4. The following matrices are in Jordan form:

•  
1 0 0 0
0 2 0 0
 
0 0 3 0
0 0 0 4
Here, fA (x) = (x − 1)(x − 2)(x − 3)(x − 4) and a(A, λ) = g(A, λ) = 1 for all
λ ∈ {1, 2, 3, 4}.

•  
2 1
0 2
Here, fA (x) = (x − 2)2 , so a(A, 2) = 2. Moreover,
 
0 1
(A − 2I) = 6= 0
0 0

to g(A, 2) 6= 2. Since 1 ≤ g(A, 2) ≤ 2, it follows that g(A, 2) = 1.

•  
2 1 0
0 2 0
0 0 3
Here, fA (x) = (x − 2)2 (x − 3), so a(A, 2) = 2, a(A, 3) = 1. Moreover, g(A, 2) = 1
and g(A, 3) = 1.

•  
0 1 0 0
0 0 0 0
 
0 0 3 0
0 0 0 3
Here, fA (x) = x2 (x − 3)2 so a(A, 0) = 2, a(A, 3) = 2. Moreover, g(A, 0) = 1 and
g(A, 3) = 2.

Example 3.5. Find the Jordan form of the matrix


 
2 2 3
A= 1 3 3
−1 −2 −2

(i) Find the characteristic polynomial:


 
x − 2 −2 −3
fA (x) = det  −1 x − 3 −3  = (x − 1)3 .
1 2 x+2

15
(ii) Find eigenvalues and algebraic multiplicity: λ = 1 with a(A, 1) = 3.

(iii) Guess the possible Jordan forms: Hence, the Jordan form of A is of the form
 
Jt1 (1) 0 ... 0
B= 0 Jt2 (1) ... 0 
0 0 . . . Jtk (1)

such that t1 + t2 + . . . + tk = 3 and k = g(A, 1). The only possibilities are


     
1 1 0 1 1 0 1 0 0
B1 := 0 1 1 or B2 := 0 1 0 or B3 := 0 1 0
0 0 1 0 0 1 0 0 1

(iv) Determine the eigenspaces: Consider


 
1 2 3
(A − I) =  1 2 3
−1 −2 −3

Row reducing gives us a row equivalent matrix


 
1 2 3
C = 0 0 0
0 0 0

Solving B(X) = 0 gives us two linearly independent solutions (−2, 1, 0) and


(−3, 0, 1). Hence,
g(A, 1) = 2.
Since g(A, 1) < a(A, 1), A is not diagonalizable.

(v) Determine the Jordan form: Since g(A, 1) = 2, the Jordan form of A has two
blocks. Hence, A is similar to
 
1 1 0
B2 = 0 1 0
0 0 1

Example 3.6. Find the Jordan canonical form for


 
2 −1 0 1
0 3 −1 0
A= 0 1

1 0
0 −1 0 3

16
(i) Find the characteristic polynomial:
 
x−2 1 0 −1
 0 x−3 1 0 
fA (x) = det  
 0 −1 x − 1 0 
0 1 0 x−3
 
x−3 1 0
= (x − 2) det  −1 x − 1 0 
1 0 x−3
 
x−3 1
= (x − 2)(x − 3) det
−1 x − 1
= (x − 2)(x − 3)[(x − 3)(x − 1) + 1]
= (x − 2)(x − 3)[x3 − 4x + 4]
= (x − 2)3 (x − 3)

(ii) Find the eigenvalues and algebraic multiplicities:


• λ = 2 and a(A, 2) = 3.
• λ = 3 and a(A, 3) = 1.
(iii) Guess the possible Jordan forms: It must be one of the following:
     
3 0 0 0 3 0 0 0 3 0 0 0
0 2 0 0 0 2 1 0 0 2 1 0
B1 = 
0 0 2 0 , B2 = 0 0
   or B3 =  
2 0 0 0 2 1
0 0 0 2 0 0 0 2 0 0 0 2

(iv) Determine the eigenspaces:


(a) λ = 2: Consider  
0 −1 0 1
0 1 −1 0
(A − 2I) = 
0 1 −1 0

0 −1 0 1
Row reducing, gives  
0 −1 0 1
0 1 −1 0
C=
0 0

0 0
0 0 0 0
Hence, the solution space to (A − 2I)(X) = 0 is the same as the solution
space to C(X) = 0, which is spanned by {(1, 0, 0, 0), (0, 1, 1, 1)}. Hence,

g(A, 2) = 2 < a(A, 2).

Thus, A is not diagonalizable.

17
(b) λ = 3: Since a(A, 3) = 1 it follows that g(A, 3) = 1 as well.
(v) Determine the Jordan form for A: Since g(A, 2) = 2, the Jordan block associated
to λ = 2 has two sub-blocks. Therefore, A is similar to
 
3 0 0 0
0 2 1 0
B2 = 
0 0 2 0

0 0 0 2

(End of Day 4)

4. Generalized Eigenspaces
Definition 4.1. Let T ∈ L(V ) and λ ∈ F be an eigenvalue of T .
(i) The generalized eigenspace associated to λ is the set

[
Vλ := {v ∈ V : (T − λI)j (v) = 0 for some j ≥ 1} = ker(T − λI)j
j=1

(ii) A vector v ∈ Vλ is called a generalized eigenvector of T associated to λ.


For convenience, we write Vλj := ker(T − λI)j .
Example 4.2.
(i) Let A ∈ M2 (R) be the matrix  
1 1
A=
0 1
Then λ = 1 is an eigenvalue. Solving the equation
  
0 1 x
(A − I)X = 0 ⇒ = 0 ⇒ y = 0.
0 0 y
Hence, ker(A − I) = span{e1 }. However,
  
2 0 1 0 1
(A − I) = = 0.
0 0 0 0
Hence, ker(A − I)2 = R2 . Therefore, all vectors in R2 are generalized eigenvectors
of A.
(ii) Let A ∈ M3 (R) be the matrix
 
3 1 0
A = 0 3 0
0 0 2

18
(a) Then, λ = 3 is an eigenvalue and

ker(A − 3I) = span{e1 }.

Moreover, as above  
0 0 0
(A − 3I)2 = 0 0 0
0 0 4
More generally,  
0 0 0
(A − 3I)j = 0 0 0 
0 0 2j
for all j ≥ 2. Hence,

ker(A − 3I)j = ker(A − 3I)2 = span{e1 , e2 }

for all j ≥ 2.
(b) Finally, λ = 2 is an eigenvalue and
 
1 1 0
A − 2I = 0 1 0
0 0 0

Therefore,
ker(A − 2I) = span{e3 }.
Now observe that
    
1 1 0 1 1 0 1 2 0
(A − 2I)2 = 0 1 0 0 1 0 = 0 1 0
0 0 0 0 0 0 0 0 0

More generally,  
1 j 0
(A − 2I)j = 0 1 0 .
0 0 0
Therefore,
ker(A − 2I)j = ker(A − 2I) = span{e3 }
for all j ≥ 1. Hence, all generalized eigenvectors are eigenvectors.

Lemma 4.3. Fix T ∈ L(V ) and λ ∈ F be an eigenvalue of T .

(i) For each j ∈ N, Vλj ⊂ Vλj+1 .

(ii) Vλ is a T -invariant subspace of V .

19
(iii) There exists r ∈ N such that Vλ = ker(T − λI)r . In other words,

Vλr = Vλr+1 = Vλr+2 = . . . .

(iv) Let Wλ := Range(T − λI)r . Then, V is a direct sum

V = Vλ ⊕ Wλ

and both subspaces are T -invariant.

(v) If S := T |Wλ ∈ L(Wλ ), then λ is not an eigenvalue of S.

Proof.

(i) If v ∈ Vλj then (T − λI)j (v) = 0. Hence, (T − λI)j+1 (v) = 0 as well. Hence,
v ∈ Vλj+1 .

(ii) Since λ is an eigenvalue, Vλ 6= {0}. If α, β ∈ Vλ and c ∈ F , then there exists


i, j ∈ N such that v ∈ Vλi and w ∈ Vλj . Assume i ≤ j, then by part (i), v ∈ Vλj .
However, Vλj is a subspace, so (cv + w) ∈ Vλj ⊂ Vλ . Therefore, Vλ is a subspace.

Moreover, if v ∈ Vλ , then (T − λI)j (v) = 0 for some j ∈ N. Hence,

(T − λI)j T (v) = T (T − λI)j (v) = T (0) = 0.

Therefore, T (v) ∈ Vλ as well so Vλ is T -invariant.

(iii) Choose a basis B = {v1 , v2 , . . . , vk } of Vλ . For each 1 ≤ i ≤ k, there exists si ∈ N


such that vi ∈ Vλsi . Let r = max{s1 , s2 , . . . , sk }. Then by part (i),

vi ∈ Vλr

for all 1 ≤ i ≤ k. Hence, Vλ ⊂ Vλr ⊂ Vλ .

(iv) If R := (T − λI)r , then we wish to show that

V = ker(R) ⊕ Range(R).

We verify the conditions of Proposition 1.5.


(a) We first show that ker(R)∩Range(R) = {0}: Suppose v ∈ ker(R)∩Range(R),
then there exists w ∈ V such that

v = R(w) = (T − λI)r (w).

Since v ∈ ker(R), we have (T − λI)2r (w) = 0, so w ∈ Vλ2r . By part (ii),


Vλ2r = Vλr , so
v = (T − λI)r (w) = 0.

20
(b) Note that ker(R) and Range(R) are both subspaces of V . Hence, W :=
ker(R) + Range(R) is a subspace of V . By the Rank-Nullity theorem,
dim(W ) = dim(V ).
Hence, W = V .
(v) Let S := T |Wλ : Wλ → Wλ . Suppose v ∈ Wλ such that S(v) = λv. Then,
(T − λI)(v) = 0.
Hence, v ∈ ker(T − λI) ⊂ Vλ . But by part (iv),
Vλ ∩ Wλ = {0}.
Hence, v = 0, so this is impossible.

Theorem 4.4. Let V be a complex vector space and T ∈ L(V ). Let {λ1 , λ2 , . . . , λk } be
the distinct eigenvalues of T . Then,
k
M
V = Vλi
i=1

Proof. We induct on k = the number of distinct eigenvalues of T (note that k ≤ dim(V )


by Corollary 2.3). If k = 1, then T has only one eigenvalue λ. By Lemma 4.3,
V = Vλ ⊕ Wλ
where Wλ = Range(T − λI)r . Moreover, T |Wλ has no eigenvalues (since T has only one
eigenvalue). Therefore, it must happen (by the Fundamental Theorem of Algebra) that
Wλ = {0}. Hence,
V = Vλ .
Now suppose k ≥ 2 and assume that the result is true for any operator R on a vector
space W with ≤ k − 1 eigenvalues. By the Fundamental Theorem of Algebra, T has an
eigenvalue λ1 , so by Lemma 4.3,
V = Vλ1 ⊕ Wλ1 .
Now consider S := T |Wλ1 ∈ L(Wλ1 ). Note that any eigenvalue of S is also an eigenvalue
of T and λ1 is not an eigenvalue of S (by Lemma 4.3). Hence, S has ≤ k − 1 distinct
eigenvalues. By induction, we may express
Wλ1 = Vλ2 ⊕ . . . ⊕ Vλk .
Therefore,
k
M
V = Vλi .
i=1

21
Remark 4.5. In F the above theorem, each Vλi is T -invariant. Therefore, if Bi is a basis
for Vλi , then B = ki=1 Bi is a basis for V and

[T ]B

is block diagonal. It turns out that we can choose Bi in such a way that [T ]B is in Jordan
form.

(End of Day 5)

22

You might also like