Final Strecth
Final Strecth
Semester 1, 2024-2025
ai,j = 0
whenever i > j.
T (vk ) ∈ span{v1 , v2 , . . . , vk }
for all 1 ≤ k ≤ n.
Definition 1.3.
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 .
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 .
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
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.
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.
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
dim(ker(λi I − A)) = di
for all 1 ≤ i ≤ k.
(End of Day 1)
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
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
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
7
Theorem 2.10. An operator T ∈ L(V ) is diagonalizable if and only if the following two
conditions hold:
Proof.
(ii) Conversely, suppose these two conditions hold. Write Wi = ker(T −λi I). We claim
that
M k
V = 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
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
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
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.
10
(v) Thus, we get an ordered basis
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
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
x+y−z =0
2x − z = 0
2x + 2y − 2z = 0
g(A, 2) = 1.
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
12
Proof. If T is triangulable, then fT (x) splits by Remark 2.14.
(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
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
T̂ (xk + W ) ∈ span{x2 + W, x3 + W, . . . , xk + W }.
(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
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
•
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.
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)
(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
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)
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,
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
18
(a) Then, λ = 3 is an eigenvalue and
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,
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.
19
(iii) There exists r ∈ N such that Vλ = ker(T − λI)r . In other words,
V = Vλ ⊕ Wλ
Proof.
(i) If v ∈ Vλj then (T − λI)j (v) = 0. Hence, (T − λI)j+1 (v) = 0 as well. Hence,
v ∈ Vλj+1 .
vi ∈ Vλr
V = ker(R) ⊕ Range(R).
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
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