Math Note
Math Note
Notes
Charlie Cruz
Contents
1
Chapter 8 Operators on Complex Vector Spaces Page 93
2
𝜙
𝐴 𝐵
𝜓
𝜂
𝐶 𝐷
Information about the class: Math 354 Honors Linear Algebra.
Professor Varilly-Alvarado (Dr. V.) with his email being [email protected].
Office: HBH 222
Office Hours: Monday 3:30 – 5:00 pm, F 3-4PM (to be confirmed)
We will be using the book Linear Algebra Done Right by Sheldon Axler.
Notation Definition
ℕ The set of natural numbers = {1, 2, 3, 4, . . .}
ℤ The set of integers = {. . . , −2, −1, 0, 1, 2 . . .}
ℚ The set of rational numbers = { 𝑏𝑎 | 𝑎, 𝑏 ∈ ℤ, 𝑏 ≠ 0}
ℝ The set of real numbers
ℂ The set of complex numbers = {𝑎 + 𝑏𝑖 | 𝑎, 𝑏 ∈ ℝ, 𝑖 2 = −1}
∈ The symbol ∈ means “is an element of” or “belongs to”
∉ The symbol ∉ means “is not an element of” or “does not belong to”
⊂ The symbol ⊂ means “is a subset of”
⊆ The symbol ⊆ means “is a subset of or equal to”
∩ The symbol ∩ means “intersection of”
∪ The symbol ∪ means “union of”
\ The symbol \ means “set difference of”
∅ The symbol ∅ means “the empty set”
∀ The symbol ∀ means “for all”
∃ The symbol ∃ means “there exists”
| The symbol | in {𝑎 | 𝑎 ∈ ℝ} means “such that”
=⇒ The symbol =⇒ means “implies”
⇐⇒ The symbol ⇐⇒ means “if and only if”
𝑎® The symbol 𝑎® means “the vector 𝑎”
Example 0.0.1
For all 𝑛 ∈ ℕ, it is true that:
𝑛(𝑛 + 1)
1+2+3+...+𝑛 =
2
e.g.: 𝑛 = 𝑞 1= 2 ,
1·2
and so on for 𝑛 = 2, 3, . . ..
3
Note:-
Think about this as a domino effect.
𝑛(𝑛+1)
Example: Let 𝑃(𝑛) : 1 + 2 + . . . + 𝑛 = 2 .
𝑘(𝑘 + 1)
1 + 2 + . . . + 𝑘 + (𝑘 + 1) = + (𝑘 + 1)
2
𝑘(𝑘 + 1) + 2(𝑘 + 1)
=
2
(𝑘 + 1)(𝑘 + 2)
=
2
(𝑘 + 1)((𝑘 + 1) + 1)
=
2
𝑛(𝑛 + 1)
= where 𝑛 = 𝑘 + 1
2
Note:-
Baby Version Let 𝐴 ⊆ ℕ be a subset. Suppose that
(i) 1 ∈ 𝐴
(ii) If 𝑘 ∈ 𝐴, then 𝑘 + 1 ∈ 𝐴
Then 𝐴 = ℕ
Baby version =⇒ PMI (let 𝐴 {𝑛 ∈ ℕ : 𝑝(𝑛) is true})
1 1 1 𝑘
+ +...+ = (1)
1∗2 2∗3 𝑘(𝑘 + 1) 𝑘+1
We want to deduce that 𝑝(𝑘 + 1) is true i.e.
1 1 1 1 𝑘+1
+ +...+ + = (2)
1∗2 2∗3 𝑘(𝑘 + 1) (𝑘 + 1)(𝑘 + 2) 𝑘+2
To do this, add 1
(𝑘+1)(𝑘+2)
to both sides of (1).
4
1 1 1 1 𝑘 1
+ +...+ + = + .
1∗2 2∗3 𝑘(𝑘 + 1) (𝑘 + 1)(𝑘 + 2) 𝑘 + 1 (𝑘 + 1)(𝑘 + 2)
The RHS of this equation is:
1 1 1 𝑘
+ +...+ =
1∗2 2∗3 𝑘(𝑘 + 1) 𝑘+1
Add 1
(𝑘+1)(𝑘+2)
to both sides of the equation to obtain
1 1 1 1 𝑘 1
+ +...+ + = +
1∗2 2∗3 𝑘(𝑘 + 1) (𝑘 + 1)(𝑘 + 2) 𝑘 + 1 (𝑘 + 1)(𝑘 + 2)
𝑘(𝑘 + 2) + 1
=
(𝑘 + 1)(𝑘 + 2)
𝑘 2 + 2𝑘 + 1
=
(𝑘 + 1)(𝑘 + 2)
(𝑘 + 1)2
=
(𝑘 + 1)(𝑘 + 2)
𝑘+1
=
𝑘+2
Then, the claim is true for 𝑘 + 1, completing the inductive step.
We deduce that the claim is true for all 𝑛 ∈ ℕ by induction.
5
Chapter 1
Vector Spaces
1.1 Fields
Definition 1.1.1: Fields or ”sets of scalars”
Example 1.1.1
𝐹2 = {0, 1} is a field.
If we write the addition table:
+ 0 1
0 0 1
1 1 0
· 0 1
0 0 0
1 0 1
6
Example 1.1.2
Let’s define 𝔽 = {4, , ◦}
With the addition and multiplication tables as followed:
+ 4 ◦ · 4 ◦
4 4 ◦ 4 4 4 4
◦ 4 4 ◦
◦ ◦ 4 ◦ 4 ◦
𝔽 is a field.
With 𝑜 𝔽 = 4 and 1𝔽 = .
Now, ”2𝐹 ” = 1𝔽 + 1𝔽 = + = ◦.
Let’s also define, − = ◦.
Theorem 1.1.2
Let 𝔽 be a field and let 𝑎 ∈ 𝔽. Then 𝑎 · 𝑂 𝐹 = 0𝐹 .
Note:-
Don’t think of zero is nothing, think of its meaning and how important it is to a field
0𝐹 + 0𝐹 = 0𝐹 as 0𝐹 is an additive identity
𝑎 · (0𝐹 + 0𝐹 ) = 𝑎 · 0𝐹
𝑎 · 0𝐹 + 𝑎 · 0𝐹 = 𝑎 · 0𝐹 by distributivity
(𝑎 · 0𝐹 + 𝑎 · 0𝐹 ) + (−𝑎 · 0𝐹 ) = 𝑎 · 0𝐹 + (−𝑎 · 0𝐹 )
(𝑎 · 0𝐹 + 𝑎 · 0𝐹 ) + −𝑎 · 0𝐹 = 0𝐹 by additive inverse
𝑎 · 0𝐹 + (𝑎 · 0𝐹 + −𝑎 · 0𝐹 ) = 0𝐹 by associativity
𝑎 · 0𝐹 + 0𝐹 = 0𝐹 by additive inverse
𝑎 · 0𝐹 = 0𝐹 by additive identity
Theorem 1.1.3
Let 𝔽 be a field and let 𝑎 ∈ 𝔽. Then −(−𝑎) = 𝑎.
7
(−𝑎) + 𝑎 = 0𝐹
This says that 𝑎 is an additive inverse of −𝑎.
Additive inverses are unique, so 𝑎 must be the additive inverse of −(−𝑎).
Note:-
you can try this at home:
(−1𝐹 ) · (𝑎) = −𝑎
Where −1 is the additive inverse of 1, and −𝑎 is the additive inverse of 𝑎.
Hint: (−1) + 1 = 0 𝑓
ℂ = {(𝑎, 𝑏) : 𝑎, 𝑏 ∈ ℝ}
(i) Addition: (𝑎, 𝑏) + (𝑐, 𝑑) : = (𝑎 + 𝑐, 𝑏 + 𝑑), where (𝑎, 𝑏) = 𝑧1 and (𝑐, 𝑑) = 𝑧2 . As such, we use ℝ addition to
define ℂ addition.
(ii) Multiplication: (𝑎, 𝑏) · (𝑐, 𝑑) : = (𝑎𝑐 − 𝑏𝑑, 𝑎𝑑 + 𝑏𝑐), where (𝑎, 𝑏) = 𝑧1 and (𝑐, 𝑑) = 𝑧 2
Note:-
You might want to think of (𝑎, 𝑏) as 𝑎 + 𝑏𝑖, where 𝑖 2 = −1. As such:
8
Definition 1.1.2: Lists 𝑇 tuples
𝑥® + 0® = (𝑥 1 , . . . , 𝑥 𝑛 ) + (0𝐹𝐹 , . . . , 0𝐹𝐹 )
= (𝑥 1 + 0𝐹𝐹 , . . . , 𝑥 𝑛 + 0𝐹𝐹 )
= (𝑥 1 , . . . , 𝑥 𝑛 )
= 𝑥®
𝑥® + (− 𝑥®) = 0®
Note:-
Warning! In general, elements of 𝔽𝑛 cannot be multiplied with each other unless we define a multipli-
cation operation on 𝔽𝑛 .
9
1.2 Vectors Spaces
Definition 1.2.1: Vector Space in general
+ : 𝑉 × 𝑉 ↦→ 𝑉 , (𝑢, 𝑣) ↦→ 𝑢 + 𝑣
And scalar multiplication as
· : 𝔽 × 𝑉 ↦→ 𝑉 , (𝜆, 𝑣) ↦→ 𝜆 · 𝑣
These operations satisfy the following properties:
1 · 𝜆2 ) ·𝑉 𝑣 = 𝜆1 ·𝑉 (𝜆2 ·𝑉 𝑣)
(𝜆𝔽 𝔽 𝔽 𝔽
for all 𝜆1 , 𝜆2 ∈ 𝔽 and 𝑣 ∈ 𝑉
Note:-
If 𝔽 = ℝ, call 𝑉 a real vector space.
If 𝔽 = ℂ, call 𝑉 a complex vector space.
𝑥® + 𝑦® = (𝑥 1 , . . . , 𝑥 𝑛 ) + (𝑦1 , . . . , 𝑦𝑛 )
= (𝑥 1 + 𝑦1 , . . . , 𝑥 𝑛 + 𝑦𝑛 )
𝜆 · 𝑥® = 𝜆 · (𝑥1 , . . . , 𝑥 𝑛 )
= (𝜆 · 𝑥1 , . . . , 𝜆 · 𝑥 𝑛 )
10
Example 1.2.1
(iii) (a) (𝑉 , 𝔽, +, ·) = (ℂ2 , ℝ, +ℝ2 , ·ℝ2 ) is a complex vector space, but we are using real numbers to scale.
(b) Addition is the same as (ii), but 𝑧1 , 𝑧 2 ∈ ℂ2 i.e. 𝑎 + 𝑏𝑖
(c) Scalar multiplication: 𝜆 ∈ ℝ, 𝜆 ·ℝ (𝑧1 , 𝑧 2 ) = (𝜆 ·ℝ 𝑧 1 , 𝜆 ·ℝ 𝑧2 ).
2
Note:-
This is a bijection!
Example 1.2.2
Let 𝑆 = [0, 1] and 𝐹 = ℝ.
Now set 𝑉 = 𝔽𝑠 = ℝ[0,1] = {functions 𝑓 : [0, 1] → ℝ}.
Another Example.
Let 𝔽 = ℝ.
Now 𝑉 = {polynomials of degree ≤ 19 with coefficients in ℝ}.
Now, +𝑉 = usual addition of polynomials and ·𝑉 = usual scalar multiplication of polynomials.
For instance,
𝑥 19 + 𝑥 + 1 ∈ 𝑉
−𝑥 19 + 𝑥 17 − 𝑥 2 ∈ 𝑉
9 ∗ (𝑥 2 + 2) = 9𝑥 2 + 18 ∈ 𝑉
Note:-
Sometimes we denote that the degree of 0 (the zero polynomial) is −∞.
11
First properties of vector spaces: Let 𝑉 be a vector space over a field 𝔽.
(i) Additive identities are unique. Suppose 0® , 0®0 ∈ 𝑉 are additive identities. Then 0® = 0® + 0®0 = 0®0.
(ii) Additive inverses are unique:
Say 𝑤, 𝑤 0 are additive inverse of 𝑣 ∈ 𝑉.
𝑤 = 𝑤 + 0® = 𝑤 + (𝑣 + 𝑤 0) = (𝑤 + 𝑣) + 𝑤 0 = 0® + 𝑤 0 = 𝑤 0.
(iii) 0 · 𝑣 = 0® , ∀𝑣 ∈ 𝑉.
0𝐹 = 0𝐹 + 0𝐹
=⇒ 0𝐹 · 𝑣 = (0𝐹 + 0𝐹 ) · 𝑣
=⇒ 0𝐹 · 𝑣 = 0𝐹 · 𝑣 + 0𝐹 · 𝑣
0𝐹 · 𝑣 + (−0𝐹 · 𝑣) = (0𝐹 · 𝑣 + 0𝐹 · 𝑣) + (−0𝐹 · 𝑣)
0® = 0𝐹 · 𝑣 + (−0𝐹 · 𝑣 + 0𝐹 · 𝑣)
0® = 0𝐹 · 𝑣 + 0®
0® = 0𝐹 · 𝑣
1.3 Subspaces
Definition 1.3.1: Subspaces
Example 1.3.1
Let 𝑉 = ℝ3 and 𝔽 = ℝ.
𝑈 = {(𝑥1 , 𝑥 2 , 0) : 𝑥 1 , 𝑥 2 ∈ ℝ} * ℝ3 = 𝑉
1.34 in book / conditions for a subspace: To check that 𝑈 ⊆ 𝑉 is a subspace, it is enough to check:
(i) 0® ∈ 𝑈
Reason: These three conditions ensure that 𝑈 has an additive identity vector, and that addition and scalar
multipliation makes sense in 𝑈.
The remaining axioms for 𝑈 to be a vector space are inherited from 𝑉.
Example 1.3.2
Let’s check associativity of addition: Let 𝑢, 𝑣, 𝑤 ∈ 𝑈.
But we know that 𝑢, 𝑣, 𝑤 ∈ 𝑉 as 𝑈 ⊆ 𝑉, so 𝑢 + (𝑣 + 𝑤) = (𝑢 + 𝑣) + 𝑤(★) in 𝑉.
Since 𝑈 is closed under addition, 𝑢 + 𝑣 ∈ 𝑈.
Again, since 𝑢 + 𝑣 ∈ 𝑈, and 𝑤 ∈ 𝑈, we know that (𝑢 + 𝑣) + 𝑤 ∈ 𝑈.
Likewise, 𝑢 + (𝑣 + 𝑤) ∈ 𝑈. This means that (★) is also true in 𝑈.
Ditto for the other axioms. Thus, we would be proving the same thing twice.
12
Example 1.3.3 (Charlie add the graphs)
𝑉 = ℝ2
Example 1.3.4
Let 𝔽 = ℝ and 𝑉 = ℝ(0,3) = {functions 𝑓 : (0, 3) → ℝ}.
Let 𝑈 ⊆ 𝑉 be the subset {functions 𝑓 : (0, 3) → ℝ | 𝑓 is differentiable and 𝑓 0(2) = 0}.
Proof: Let’s check that 𝑈 ⊆ 𝑉 is a subspace:
Example 1.3.5
Our field will be 𝔽 = ℝ, vectors space will be 𝑉 = ℝ3 .
Let 𝑈1 = {(𝑥, 0, 0) : 𝑥 ∈ ℝ} , 𝑈2 = {(0, 𝑦, 0) : 𝑦 ∈ ℝ}.
Let 𝑈1 + 𝑈2 = {𝑉1 + 𝑉2 : 𝑉1 ∈ 𝑈1 , 𝑉2 ∈ 𝑈2 }.
This means that this is equal to {(𝑥, 0, 0) + (0, 𝑦, 0) : 𝑥, 𝑦 ∈ ℝ} = {(𝑥, 𝑦, 0) : 𝑥, 𝑦 ∈ ℝ}
Theorem 1.3.1
If 𝑈1 , . . . , 𝑈𝑚 are subspaces of 𝑉, then 𝑈1 + . . . + 𝑈𝑚 is the smallest subspace of 𝑉 containing 𝑈1 , . . . , 𝑈𝑚 .
Have to prove that:
13
(i) 𝑈1 + . . . + 𝑈𝑚 is a subspace of 𝑉 (not just a subset).
(ii) 𝑈1 ⊆ 𝑈1 + . . . , 𝑈𝑚 , 𝑈2 ⊆ 𝑈1 + . . . + 𝑈𝑚 , . . . , 𝑈𝑚 ⊆ 𝑈1 + . . . + 𝑈𝑚 .
(iii) 𝑈1 + . . . + 𝑈𝑚 is the smallest subspace of 𝑉 containing 𝑈1 , . . . , 𝑈𝑚 .
0® = 0®∈𝑢1 + . . . + 0®∈𝑈𝑚 ∈ 𝑈1 + . . . + 𝑈𝑚
Thus, we have shown that the additive identity is in 𝑈1 + . . . + 𝑈𝑚 .
Now, we want to show that this sum is closed under addition.
Let 𝑣® , 𝑤
® ∈ 𝑈1 + . . . + 𝑈𝑚 . We need to show that 𝑣® + 𝑤
® ∈ 𝑈1 + . . . + 𝑈 𝑚 .
Then, 𝑉 = 𝑉1 + . . . + 𝑉𝑚 , 𝑊 = 𝑤®1 + . . . + 𝑊𝑚
® ® ® ® ®
As such
𝑤
® + 𝑣® = 𝑤®1 + . . . + 𝑤®𝑚 + 𝑣®1 + . . . + 𝑣®𝑚
∈ 𝑈1 + . . . + 𝑈 𝑚
® = 𝜆 ∗ 𝑣®1 + . . . + 𝑣®𝑚
𝜆∗𝑉
= 𝜆 ∗ 𝑣®1 + . . . + 𝜆 ∗ 𝑣®𝑚
∈ 𝑈1 + . . . + 𝑈 𝑚
14
Definition 1.3.3: Direct sum
Example 1.3.6
Let 𝑈1 = {(𝑥, 𝑦, 0) : 𝑥, 𝑦 ∈ ℝ}, 𝑈2 = {(0, 0, 𝑧) : 𝑧 ∈ ℝ}.
0® = 0®∈𝑢1 + 0®∈𝑢2 + 0®∈𝑢3 = (0, −1, 0)∈𝑢1 + (0, 0, −1)∈𝑢2 + (0, 1, 1)∈𝑢3
Theorem 1.3.2
𝑈1 + . . . + 𝑈𝑚 is a direct sum if and only if
0® = 0®∈𝑢1 + . . . + 0®∈𝑢𝑚 , 𝑢𝑖 ∈ 𝑈 𝑖
Let 𝑣® ∈ 𝑈1 + . . . + 𝑈𝑚 be arbitrary, and suppose
15
Then,
Thus, we have shown that 𝑈1 + . . . + 𝑈𝑚 is a direct sum if and only if 0® = 0®∈𝑢1 + . . . + 0®∈𝑢𝑚 is the only way
to write 0® as a sum of vectors from 𝑈1 , . . . , 𝑈𝑚 .
Alternative proof that 𝑈1 + 𝑈2 (from our example) is direct using criterion from our prvious theorem.
Theorem 1.3.3
If 𝑈1 , 𝑈2 are subspaces of of a vector space 𝑉, then
n o
(𝑈1 + 𝑈2 is direct) ⇐⇒ (𝑈1 ∩ 𝑈2 = 0® )
n o
Proof of =⇒ : Suppose 𝑈1 + 𝑈2 is direct, then we want to show that 𝑈1 ∩ 𝑈2 = 0® .
In other words we need to prove subset inclusion in both directions.
⊆ : We have {0} ⊆ 𝑈1 ∩ 𝑈2 since 0® ∈ 𝑈1 and 0® ∈ 𝑈2 .
=⇒ 𝑣® ∈ 𝑈1 and 𝑣® ∈ 𝑈2
=⇒ −®𝑣 ∈ 𝑈1 and − 𝑣® ∈ 𝑈2 as they are closed under scalar multiplication
𝑣 ) ∈ 𝑈1 + 𝑈2 by our previous theorem
=⇒ 0® = 𝑣® + (−®
=⇒ 𝑣® = 0® , −®
𝑣 = 0®
n o
Thus, 𝑈1 ∩ 𝑈2 = 0® .
n o
Proof of ⇐= : Suppose 𝑈1 ∩ 𝑈2 = 0® , then we want to show that 𝑈1 + 𝑈2 is direct.
Suppose 0® = 𝑢®1 + 𝑢®2 for some 𝑢®1 ∈ 𝑈1 , 𝑢®2 ∈ 𝑈2 .
We want to show that 𝑢®1 = 𝑢®2 = 0®.
16
0 = 𝑢®1 + 𝑢®2 =⇒ 𝑢®1 = −𝑢®2 =⇒ 𝑢®1 ∈ 𝑈1 and 𝑢®1 ∈ 𝑈2
n o
so 𝑢®1 ∈ 𝑈1 ∩ 𝑈2 = 0®
By our previous theorem, 𝑈1 + 𝑈2 is direct because we can only write 0® in one way as a sum of vectors
from 𝑈1 and 𝑈2 .
n o
Thus, we have shown that 𝑈1 + 𝑈2 is direct if and only if 𝑈1 ∩ 𝑈2 = 0® .
17
Chapter 2
Example 2.1.1
Let 𝑉 = ℝ, and our field being 𝔽 = ℝ.
So 17, −4, 2 is a linear combination of (2, 1, −3) and (1, −2, 4).
Definition 2.1.2
The span of 𝑣®1 , 𝑣®2 , . . . , 𝑣®𝑚 ∈ 𝑉 is the set of all linear combinations of 𝑣®1 , 𝑣®2 , . . . , 𝑣®𝑚 .
Note:-
n o
We have a few convention: span () ≔ 0®𝑣 .
Proposition 2.1.1
The span (𝑣 1 , . . . , 𝑣 𝑚 ) is the smallest subspace of 𝑉 that contains 𝑣1 , . . . , 𝑣 𝑚 .
Proof: We have to show three things in 1.34.
(a) We know that 0®𝑣 = 0𝔽 · 𝑣®1 + 0𝔽 · 𝑣®2 + · · · + 0𝔽 · 𝑣®𝑚 ∈ span 𝑣®1 , . . . , 𝑣®𝑚 . Thus, we are done
18
(c) Now closed under scalar multiplication:
Example 2.1.2
Let 𝑉 = ℝ3 , and the field 𝔽 = ℝ.
Then span ((1, 0, 0), (0, 1, 0), (0, 0, 1)) = ℝ3 .
Definition 2.1.3
Example 2.1.3
𝑃𝑚 (𝐹) = {Polys of degree ≤ 𝑚 with coefficients in 𝐹}
And we claim that this is spanned by 1, 𝑥, 𝑥 2 , . . . , 𝑥 𝑚 .
Because any 𝑝(𝑥) ∈ 𝑃𝑚 (𝐹) has the form 𝑎 𝑚 · 𝑥 𝑚 + . . . + 𝑎 1 · 𝑥 + 𝑎 0 for some 𝑎 0 , . . . , 𝑎 𝑚 ∈ 𝐹.
Proposition 2.1.2
𝑃(𝐹) = {Polys with coefficients in 𝐹} is not finite-dimensional.
19
Then, there exists a finite list 𝑝1 (𝑥), . . . , 𝑝 𝑚 (𝑥) that spans 𝑃(𝐹).
In other words, span (𝑝1 (𝑥), . . . , 𝑝 𝑚 (𝑥)) = 𝑃(𝐹).
Let 𝑛 = 𝑚𝑎𝑥 (deg(𝑝1 (𝑥)), . . . , deg(𝑝 𝑚 (𝑥))).
Then, 𝑑𝑒 𝑔(𝑎1 · 𝑝1 (𝑥) + . . . + 𝑎 𝑚 · 𝑝 𝑚 (𝑥)) ≤ 𝑛 for all 𝑎 1 , . . . , 𝑎 𝑚 ∈ 𝐹.
So the degree of every element of span (𝑝1 (𝑥), . . . , 𝑝 𝑚 (𝑥)) is at most 𝑛.
Hence, 1𝔽 · 𝑋 𝑛+1 ∉ span (𝑝1 (𝑥), . . . , 𝑝 𝑚 (𝑥)).
This means that span (𝑝1 (𝑥), . . . , 𝑝 𝑚 (𝑥)) ( 𝑃(𝐹).
This is absurd!
So our assumption that 𝑃(𝐹) is finite-dimensional is false.
Definition 2.1.4
Linear (In)depence.
Let (𝑉 , 𝔽, +, ·) be a vector space.
A list 𝑣®1 , . . . , 𝑣®𝑚 ∈ 𝑉 is linearly independent if the only way to write
Example 2.1.4
We want to show that (1, 0, 0), (0, 1, 0)0, (0, 0, 1) are linearly independent in ℝ3 = 𝑉.
because if
Example 2.1.5
We want to show that (1, 0, 0), (0, 1, 0),
√ (0, 0, 1) spans
√ ℝ3 . √ √
This implies that the list (2, −1, 𝜋), ( 3, −7, 𝑒), ( 19, −1, 7), (0, −5, 2 + 3) is not linearly independent.
Since the length of the first list is 3, and the length of the second list is 4.
Thus, this list cannot be linearly independent.
Prep work: Say 𝑣®1 , . . . , 𝑣®𝑛 ∈ 𝑉 is linearly dependent. Then there is a 𝑗 ∈ {1, . . . , 𝑚}
such that
20
(i) 𝑣 𝑗 ∈ span 𝑣®1 , . . . , 𝑣 𝑗 ®− 1
(ii) span 𝑣®1 , . . . , 𝑣®𝑚 = span 𝑣1 , . . . , 𝑣ˆ𝑗 , . . . , 𝑣 𝑚 , where 𝑣ˆ𝑗 means that we remove 𝑣 𝑗 from the list.
𝑎1 𝑎 𝑗−1
1
=⇒ 𝑣 𝑗 = − 𝑎1 𝑣1 + . . . + 𝑎 𝑗−1 𝑣 𝑗−1 = − 𝑣1 + . . . + − 𝑣 𝑗−1
𝑎𝑗 𝑎𝑗 𝑎𝑗
=⇒ 𝑣 𝑗 ∈ span 𝑣1 , . . . , 𝑣 𝑗−1
Note:-
We have to do the one above as well.
Now, we want to show the other direction as well.
𝑎1 𝑎 𝑗−1
=⇒ 𝑣 = 𝑏 1 𝑣1 + . . . + 𝑏 𝑗 − 𝑣1 + . . . + − 𝑣 𝑗−1 + 𝑏 𝑗+1 𝑣 𝑗+1 + . . . + 𝑏 𝑚 𝑣 𝑚 where 𝑣 𝑗 (from (i))
𝑎𝑗 𝑎𝑗
=⇒ 𝑣 ∈ span 𝑣1 , . . . , 𝑣ˆ𝑗 , . . . , 𝑣 𝑚
And
span 𝑣®1 , 𝑢®1 , . . . , 𝑢®𝑛 = span 𝑣®1 , 𝑢®1 , . . . , 𝑢®ˆ𝑗1 , . . . , 𝑢®𝑛
21
Note:-
NB means nota bene, which means note well.
Notice that 𝑣 1 is not plucked out from our listn when
o we apply LDL.
If it were, then LDL would say 𝑣1 ∈ span () = 0®𝑣 .
This implies that 𝑣1 = 0®𝑣 ,
But 𝑣®1 , . . . , 𝑣®𝑚 is linearly independent.
As 0®𝑣 = 1𝔽 · 𝑣®1 + 0𝔽 · 𝑣®2 + . . . + 0𝔽 · 𝑣®𝑚 is the only way to write 0®𝑣 as a linear combination of 𝑣®1 , . . . , 𝑣®𝑚 .
Thus, 𝑣1 ≠ 0®𝑣 .
Step 2: 𝑣2 ∈ span 𝑢®1 , . . . , 𝑢®𝑛 = span 𝑣®1 , 𝑢®1 , . . . , 𝑢®𝑛 = span 𝑣®1 , 𝑢®1 , . . . , 𝑢®ˆ𝑗1 , . . . , 𝑢®𝑛
After 𝑚 steps: Our list is 𝑣®1 , . . . , 𝑣®𝑚 , some 𝑢’s implies that 𝑚 ≤ 𝑛
Thus, we have shown that 𝑚 ≤ 𝑛.
22
2.2 Basis
Definition 2.2.1
Example 2.2.1
𝑣1 = (1, 0, . . . , 0)
𝑣2 = (0, 1, . . . , 0)
..
.
𝑣 𝑛 = (0, 0, . . . , 1)
e.g., 𝑉 = ℝ3 = span ((1, 0, 0), (0, 1, 0), (0, 0, 1)). This list is linearly independent.
(ii) 𝑉 = ℝ2 The list (1, 2), (2, 3) is a basis.
Linearly Independence: If
(iii) 𝑉 = 𝑃𝑚 (ℝ)
Thus, the list 1, 𝑥, 𝑥 2 , . . . , 𝑥 𝑚 is a basis for 𝑉
Proposition 2.2.1
𝑣®1 , . . . , 𝑣®𝑛 ∈ 𝑉 is a bais for 𝑉 if and only if every 𝑣® ∈ 𝑉 can be written uniquely as a linear combination
of 𝑣®1 , . . . , 𝑣®𝑛 .
Proof of =⇒ : Say 𝑣®1 , . . . , 𝑣®𝑛 ∈ 𝑉 is a basis.
Let 𝑣® ∈ 𝑉. Since 𝑉 = span 𝑣®1 , . . . , 𝑣®𝑛 ,
we know that 𝑣® = 𝑎 1 𝑣®1 + . . . + 𝑎 𝑛 𝑣®𝑛 for some 𝑎 1 , . . . , 𝑎 𝑛 ∈ 𝔽.
Since 𝑣®1 , . . . , 𝑣®𝑛 are linearly independent, we know this representation is unique.
Proof of ⇐= : Suppose that every 𝑣® ∈ 𝑉 can be written uniquely as 𝑣® = 𝑎1 𝑣®1 + . . . + 𝑎 𝑛 𝑣®𝑛 for some
𝑎1 , . . . , 𝑎 𝑛 ∈ 𝔽.
Then 𝑣® ∈ span 𝑣®1 , . . . , 𝑣®𝑛 , so 𝑉 ⊆ span 𝑣®1 , . . . , 𝑣®𝑛 .
By the definition of span, we know that span 𝑣®1 , . . . , 𝑣®𝑛 ⊆ 𝑉.
Thus, 𝑉 = span 𝑣®1 , . . . , 𝑣®𝑛 .
Next, let 𝑣® = 0®𝑣 .
We know that 0®𝑣 = 𝑎1 𝑣®1 + . . . + 𝑎 𝑛 𝑣®𝑛 for unique 𝑎 1 , . . . , 𝑎 𝑛 ∈ 𝔽.
On the other hand (OTOH): taking 𝑎1 = . . . = 𝑎 𝑛 = 0 works!
Therefore, the only way to write 0®𝑣 is a linearly combination of 𝑣®1 , . . . , 𝑣®𝑛
23
is to take 𝑎1 = . . . = 𝑎 𝑛 = 0𝔽 .
The definition implies that 𝑣®1 , . . . , 𝑣®𝑛 is linearly independent.
Thus, we have shown that 𝑣®1 , . . . , 𝑣®𝑛 ∈ 𝑉 is a basis for 𝑉 if and only if every 𝑣® ∈ 𝑉 can be written uniquely
as a linear combination of 𝑣®1 , . . . , 𝑣®𝑛 .
Theorem 2.2.1
Let (𝑉 , 𝔽, +, ·) be a finite-dimensional vector space (fdvs).
Then every spanning list for 𝑉 can be trimmed to a basis.
𝐵 = 𝑣®1 , . . . , 𝑣®𝑛 ;
1 /* Note that 𝐵 has no order. */
2 for 𝑗 = 1, . . . , 𝑛 do
if 𝑣 𝑗 ∈ span 𝑣®1 , . . . , 𝑣 𝑗−1
® ∩ 𝐵 then
3
4 Delete 𝑣 𝑗 from 𝐵;
5 end
When the loop is finished, the set 𝐵 gives rise to a basis (any order).
Example 2.2.2
𝑉 = ℝ3 .
Let 𝑣1 =(1, 0, 0), 𝑣2 = (1, 1, 1), 𝑣3 = (0, 1, 1), and 𝑣4 = (0, 0, 1).
Let 𝐵 = 𝑣®1 , 𝑣®2 , 𝑣®3 , 𝑣®4
n o
Step 1: Is 𝑣1 ∈ span (∅ ∩ 𝐵) = span () = 0®𝑣 ?
NO. Leave 𝐵 alone.
𝑣3 = −𝑣1 + 𝑣2
New 𝐵 = {𝑣1 , 𝑣 2 , 𝑣 4 }
Step 4: Is 𝑣4 ∈ span ({𝑣 1 , 𝑣 2 , 𝑣 3 } ∩ 𝐵) = span (𝑣1 , 𝑣 2 )?
Does 𝑣 4 = 𝑎 1 · 𝑣1 + 𝑎2 · 𝑣2 ?
No!
Leave 𝐵 alone.
Thus, 𝐵 = {𝑣1 , 𝑣 2 , 𝑣 4 } is a basis for 𝑉 through trimming.
Corollary 2.2.1
Any linearly independence list 𝑣®1 , . . . , 𝑣®𝑚 on 𝑉 can be extended to a basis.
Proof: Let 𝑢®1 , . . . , 𝑢®𝑛 be any basis for 𝑉 .
Trim the enlarged list 𝑣®1 , . . . , 𝑣®𝑚 , 𝑢®1 , . . . , 𝑢®𝑛 .
24
No 𝑣®𝑖 is deleted during trimming (LDL).
As such, 𝑉 = 𝑈 + 𝑊
(ii) Let 𝑣 ∈ 𝑈 ∩ 𝑊.
Since 𝑢’s and 𝑤’s are linearly independent in 𝑉, this forces 𝑎’s and 𝑏’s to be all 0𝔽
This implies that 𝑣 = 0®𝑣
n o
Thus, 𝑈 ∩ 𝑊 ⊆ 0®𝑣 .
n o
Thus, 𝑈 ∩ 𝑊 = 0®𝑣 and 𝑈 + 𝑊 = 𝑉.
Therefore, 𝑉 = 𝑈 ⊕ 𝑊.
n o
proof of claim: If 𝑈 = 0®𝑣 then we are done!
This is because 𝑈 = span ()
Otherwise, there
is a 𝑣®1 ≠ 0𝑣 in 𝑈 .
®
If 𝑈 = span 𝑣®1 , then we are done.
This is because 𝑈 is finite-dimensional.
Otherwise, there is a 𝑣®2 ∈ 𝑈 such that 𝑣®2 ∉ span 𝑣®1 .
25
This implies that (𝑣1 , 𝑣 2 ) is a linearly independent list in 𝑈 .
Which means that the list is also linearly independent in 𝑉 .
If 𝑈 = span 𝑣®1 , 𝑣®2 , then we are done.
Otherwise there is a 𝑣®3 ∈ 𝑈 such that 𝑣®3 ∉ span 𝑣®1 , 𝑣®2 .
This implies that (𝑣1 , 𝑣 2 , 𝑣 3 ) is a linearly independent list in 𝑈 .
Which means that the list is also linearly independent in 𝑉 .
This process terminates:
𝑉 is finite dimensional, which implies 𝑉 = span 𝑥®1 , . . . , 𝑥®𝑝
At step 𝑚 we produce a linearly independent list 𝑣®1 , . . . , 𝑣®𝑚 of 𝑉 .
The key result we have proved in class: 𝑚 ≤ 𝑝.
2.3 Dimension
Theorem 2.3.1
Any two bases of a finite-dimensional vector space 𝑉 have the same length.
Proof: Say 𝑣®1 , . . . , 𝑣®𝑚 and 𝑢®1 , . . . , 𝑢®𝑛 are bases for 𝑉 .
Let 𝑣®1 , . . . , 𝑣®𝑚 are linearly independent in 𝑉 .
Let 𝑢®1 , . . . , 𝑢®𝑛 span 𝑉 .
By the key result 𝑚 ≤ 𝑛. Reverse roles to get 𝑛 ≤ 𝑚.
Thus, 𝑚 = 𝑛.
The length of any basis for 𝑉 is called the dimension of 𝑉.
Example 2.3.1
26
Note:-
Do this as an exercise
.
Example 2.3.2
Take 𝑉 = {𝑝(𝑥) ∈ 𝑃3 (ℝ) : 𝑝‘(5) = 0} ⊆ 𝑃3 (ℝ)
We know that 𝑃3 (ℝ) is 4-dimensional, with a basis 1, 𝑥, 𝑥 2 , 𝑥 3 .
27
Definition 2.3.1: Dimension of a sum
Claim: Let 𝑢®1 , . . . , 𝑢®𝑛 , 𝑣®1 , . . . , 𝑣®𝑚 , 𝑤®1 , . . . , 𝑤®𝑝 is a basis for 𝑈1 + 𝑈2 .
Assume claim true for now.
28
Chapter 3
Linear Transformations
0®𝑤 = 𝑇(0®𝑣 )
Example 3.1.1
We will be showing a lot of examples today!
(i) Zero map:
0: 𝑉 → 𝑊
𝑣 ↦→ 0𝑤
id𝑉 : 𝑉 → 𝑉
𝑣 ↦→ 𝑣
Note:-
Notation ℒ(𝑉 , 𝑊) = {𝑇 : 𝑉 → 𝑊 | 𝑇 is linear}.
29
(iii) Differentiation map: 𝐷 ∈ ℒ(𝑃(ℝ), 𝑃(ℝ))
𝐷 : 𝑃(ℝ) → 𝑃(ℝ)
𝑝(𝑥) ↦→ 𝑝0(𝑥)
Let’s check!
Linear:
𝐼 : 𝑃(ℝ) → 𝑃(ℝ)
∫ 1
𝑝(𝑥) ↦→ 𝑝(𝑥)𝑑𝑥
0
Let’s check!
Linear:
∫1 ∫1 ∫1
(a) 𝐼(𝑝(𝑥) +𝑃(ℝ) 𝑞(𝑥)) = 0
(𝑝(𝑥) + 𝑞(𝑥))𝑑𝑥 = 0
𝑝(𝑥)𝑑𝑥 + 0
𝑞(𝑥)𝑑𝑥 = 𝐼(𝑝(𝑥)) +ℝ 𝐼(𝑞(𝑥))
∫1 ∫1
(b) 𝐼(𝜆 ·𝑃(ℝ) 𝑝(𝑥)) = 0
(𝜆𝑝(𝑥))𝑑𝑥 = 𝜆 0
𝑝(𝑥)𝑑𝑥 = 𝜆𝐼(𝑝(𝑥))
𝑆 : 𝔽∞ → 𝔽∞
(𝑥1 , 𝑥 2 , . . .) ↦→ (0, 𝑥 1 , 𝑥 2 , . . .)
(vi) 𝑇 : ℝ3 → ℝ2
𝑇 : ℝ3 → ℝ2
(𝑥1 , 𝑥 2 , 𝑥 3 ) ↦→ (5𝑥 + 7𝑦 − 𝑧, 2𝑥 − 𝑦)
ℒ(𝑉 , 𝑊) = {𝑇 : 𝑉 → 𝑊 | 𝑇 is linear}
The set ℒ(𝑉 , 𝑊) can be given the structure of a vector space over 𝔽.
(𝑇 + 𝑆) : 𝑉 → 𝑊
𝑣 ↦→ 𝑇(𝑣) +𝑊 𝑆(𝑣)
(𝜆𝑇) : 𝑉 → 𝑊
𝑣 ↦→ 𝜆 ·𝑊 𝑇(𝑣)
𝑇 𝑆
z}|{ z}|{
𝑈 → 𝑉 → 𝑊
(a) Associativity: 𝑈 → 𝑉 → 𝑊 → 𝑋
z}|{ z}|{ z}|{
𝑇3 · (𝑇2 · 𝑇1 ) = (𝑇3 · 𝑇2 ) · 𝑇1
(b) Identities: 𝑇 : 𝑉 → 𝑊
id𝑉 : 𝑉 → 𝑉
𝑣 ↦→ 𝑣
id𝑊 : 𝑊 → 𝑊
𝑤 ↦→ 𝑤
𝑇 · (𝑆1 + 𝑆2 ) = 𝑇 · 𝑆1 + 𝑇 · 𝑆2
Important: Say 𝑉 is a finite dimensional vector space over 𝔽1 , and 𝑣®1 , . . . , 𝑣®𝑛 is a basis for 𝑉 .
Then a linear map 𝑇 : 𝑉 → 𝑊 is determined by the values 𝑇(𝑣®1 ), . . . , 𝑇(𝑣®𝑛 ).
Reason: Let 𝑣® ∈ 𝑉 = span 𝑣®1 , . . . , 𝑣®𝑛 .
This implies that 𝑣® = 𝜆1 𝑣®1 + · · · + 𝜆𝑛 𝑣®𝑛 for some and unique 𝜆1 , . . . , 𝜆𝑛 ∈ 𝔽.
Then:
𝑇(®
𝑣 ) = 𝑇(𝜆1 𝑣®1 + · · · + 𝜆𝑛 𝑣®𝑛 )
= 𝑇(𝜆1 𝑣®1 ) + · · · + 𝑇(𝜆𝑛 𝑣®𝑛 )
= 𝜆1 𝑇(𝑣®1 ) + · · · + 𝜆𝑛 𝑇(𝑣®𝑛 )
𝑇(®
𝑣 ) = 𝑇(𝜆1 𝑣®1 + · · · + 𝜆𝑛 𝑣®𝑛 ) = 𝜆1 𝑇(𝑣®1 ) + · · · + 𝜆𝑛 𝑇(𝑣®𝑛 ) = 𝜆1 𝑤®1 + · · · + 𝜆𝑛 𝑤®𝑛
31
3.2 Null spaces and Ranges
Definition 3.2.1: Kernels or null spaces
Note:-
The image on our canvas page is this definition.
Note:-
We know that 𝑇(0®𝑉 ) = 0®𝑊 , so 0®𝑉 ∈ ker 𝑇.
Example 3.2.1
(a) ker(0) = 𝑉
0: 𝑉 → 𝑊
𝑣 ↦→ 0𝑊
n o
(b) ker(id𝑉 ) = 0®𝑉
id𝑉 : 𝑉 → 𝑉
𝑣 ↦→ 𝑣
(c)
𝐷 : 𝑃(ℝ) → 𝑃(ℝ)
𝑝(𝑥) ↦→ 𝑝0(𝑥)
Then:
(d) Shift
𝑆 : 𝔽∞ → 𝔽∞
(𝑥 1 , 𝑥 2 , . . .) ↦→ (𝑥 2 , 𝑥 3 , . . .)
Then:
n o
ker 𝑆 = (𝑥 1 , 𝑥 2 , . . .) ∈ 𝔽∞ | (𝑥2 , 𝑥 3 , . . .) = 0®𝔽∞
= {(𝑥 1 , 0, 0, . . .) ∈ 𝔽∞ | 𝑥 1 ∈ 𝔽}
32
Proposition 3.2.1
In general, ker 𝑇 is a subspace of 𝑉.
Proof: Let 𝑇 : 𝑉 → 𝑊 be a linear map.
Now we want to check 1.34:
𝑇(®
𝑢 +𝑉 𝑣® ) = 𝑇(®
𝑢 ) +𝑊 𝑇(®
𝑣)
= 0®𝑊 +𝑊 0®𝑊
= 0®𝑊
Thus, 𝑢® +𝑉 𝑣® ∈ ker 𝑇.
𝑇(𝜆 ·𝑉 𝑢® ) = 𝜆 ·𝑊 𝑇(®
𝑢)
= 𝜆 ·𝑊 0®𝑊
= 0®𝑊
Thus, 𝜆 ·𝑉 𝑢® ∈ ker 𝑇.
𝑇(®
𝑢) = 𝑇(®
𝑣) =⇒ 𝑢=𝑣
|{z} |{z} |{z}
equal outputs must come from equal inputs
𝑢≠𝑣 =⇒ 𝑇(®
𝑢 ) ≠ 𝑇(®
𝑣)
|{z} | {z }
unequal inputs unequal outputs
Proposition 3.2.2
Let 𝑇 : 𝑉 → 𝑊 be a linear map. n o
Then T is injective if and only if ker 𝑇 = 0®𝑉 .
33
Let 𝑣® ∈ ker 𝑇.
Then 𝑇(® 𝑣 ) = 𝑇(0®𝑉 + 0®𝑉 ) = 𝑇(0®𝑉 ) + 𝑇(0®𝑉 ) = 0®𝑊 + 0®𝑊 = 0®𝑊 .
n o
Proof of ⇐= : We are given that ker 𝑇 = 0®𝑉 .
We want to show that 𝑇 is injective.
Suppose 𝑇(® 𝑢 ) = 𝑇(®𝑣 ).
Then 𝑇(® 𝑢 ) − 𝑇(®𝑣 ) = 0®𝑊 .
By linearity, 𝑇(® 𝑢 − 𝑣® ) = 0®𝑊 .
Thus, 𝑢® − 𝑣® ∈ ker 𝑇.
This means that 𝑢® − 𝑣® = 0®𝑉 .
Therefore, 𝑢® − 𝑣® = 0®𝑉 =⇒ 𝑢® = 𝑣® .
Let 𝑇 ∈ ℒ(𝑉 , 𝑊). Then the image of 𝑇 is 𝐼𝑚(𝑇) = {𝑤 ∈ 𝑊 | 𝑤 = 𝑇(𝑣) for some 𝑣 ∈ 𝑉 }.
Also denoted as Range (𝑇).
It is a subspace of 𝑊 (Axle 3.19)
Example 3.2.2
n o
(i) 𝐼𝑚(0) = 0®𝑊
0: 𝑉 → 𝑊
𝑣 ↦→ 0®𝑊
(ii) 𝐼𝑚(id𝑉 ) = 𝑉
id𝑉 : 𝑉 → 𝑉
𝑣 ↦→ 𝑣
𝐷 : 𝑃(ℝ) → 𝑃(ℝ)
𝑝(𝑥) ↦→ 𝑝0(𝑥)
𝐷 : 𝑃5 (ℝ) → 𝑃5 (ℝ)
Note: 𝑥 5 ∉ 𝐼𝑚(𝐷5 )
34
Definition 3.2.4: Surjective
A map 𝑇 : 𝑉 → 𝑊 is surjective if
for any 𝑤 ∈ 𝑊 there is a 𝑣 ∈ 𝑉 such that 𝑇(𝑣) = 𝑤.
i.e., 𝑇 is surjective if (and only if) 𝐼𝑚(𝑇) = 𝑊.
Proof: Let 𝑉 be a finite dimensional vector space over 𝔽 and ker 𝑇 ⊆ 𝑉 be a subspace.
This means that ker 𝑇 is finite dimensional.
Let 𝑢®1 , . . . , 𝑢®𝑚 be a basis for ker 𝑇.
Which means that 𝑢®1 , . . . , 𝑢®𝑛 is linearly independent in ker 𝑇.
Therefore, it also linearly independent in 𝑉.
We can extend this list to a full basis 𝑢®1 , . . . , 𝑢®𝑛 , 𝑣®1 , . . . , 𝑣®𝑚 for 𝑉.
Then dim 𝑉 = 𝑛 + 𝑚, and dim ker 𝑇 = 𝑛
Claim: 𝑇(𝑣®1 ), . . . , 𝑇(𝑣®𝑚 ) is a basis for 𝐼𝑚(𝑇).
Thus, if the claim is true, then 𝐼𝑚(𝑇) is finite dimensional and dim 𝐼𝑚(𝑇) = 𝑚.
Thus, dim 𝑉 = dim ker 𝑇 + dim 𝐼𝑚(𝑇).
𝑇(𝑣) = 𝑇(𝑎1 𝑢®1 + · · · + 𝑎 𝑛 𝑢®𝑛 + 𝑏 1 𝑣®1 + · · · + 𝑏 𝑚 𝑣®𝑚 ) = 𝑇(𝑎1 𝑣®1 ) + · · · + 𝑇(𝑎 𝑛 𝑣®𝑛 ) + 𝑇(𝑏1 𝑣®1 ) + · · · + 𝑇(𝑏 𝑚 𝑣®𝑚 )
= 𝑎 1 𝑇(𝑣®1 ) + · · · + 𝑎 𝑛 𝑇(𝑣®𝑛 ) + 𝑏 1 𝑇(𝑣®1 ) + · · · + 𝑏 𝑚 𝑇(𝑣®𝑚 )
we know that 𝑇(𝑢®1 ) = · · · = 𝑇(𝑢®𝑛 ) = 0®𝑊
= 𝑏 1 𝑇(𝑣®1 ) + · · · + 𝑏 𝑚 𝑇(𝑣®𝑚 )
∈ span 𝑇(𝑣®1 ), . . . , 𝑇(𝑣®𝑚 )
35
Since 𝑢®1 , . . . , 𝑢®𝑛 , 𝑣®1 , . . . , 𝑣®𝑚 are linearly independent in 𝑉,
it follows that 𝑑1 = · · · = 𝑑𝑛 = 𝑐 1 = · · · = 𝑐 𝑚 = 0.
Thus, 𝑇(𝑣®1 ), . . . , 𝑇(𝑣®𝑚 ) are linearly independent in 𝐼𝑚(𝑇).
As we have shown that 𝑇(𝑣®1 ), . . . , 𝑇(𝑣®𝑚 ) are linearly independent in 𝐼𝑚(𝑇) and span 𝐼𝑚(𝑇), we have proven
the claim.
𝑇 : ℝ𝑛 → ℝ𝑚
𝑥1 𝑎1,1 𝑥1 + · · · + 𝑎1,𝑛 𝑥 𝑛
.. ..
. ↦→
.
𝑥 𝑛 𝑎 𝑚,1 𝑥1 + · · · + 𝑎 𝑚,𝑛 𝑥 𝑛
We can check 𝑇 is linear!
Thus, 𝑥1 = · · · = 𝑥 𝑛 = 0 is 0®ℝ ∈ ker 𝑇.
Rank-nullity: By the theorem, we know that dim ℝ𝑛 = dim ker 𝑇 + dim 𝐼𝑚(𝑇).
| {z } | {z }
𝑛 ≤𝑚
Thus, 𝑛 ≤ dim ker 𝑇 + 𝑚.
As such, dim ker 𝑇 ≥ 𝑛 − 𝑚
Suppose that 𝑛 − 𝑚 > 0 (more variables than equations).
Therefore, dim ker 𝑇 > 0.
Meaning that there are non-zero solutions to the system of equations.
Note:-
0
n o .
Is ker 𝑇 = 0®ℝ𝑛 = .. ?
0
Or is there something else?
Theorem 3.2.2
Let 𝑉 , 𝑊 be a finite dimensional vector space over 𝔽 and dimn𝑉 >o dim 𝑊.
Then any linear map 𝑇 : 𝑉 → 𝑊 is not injective, i.e., ker 𝑇 ≠ 0®𝑉 .
36
Proof:
Note:-
Going back to systems of linear equations:
37
3.3 Matrix of a linear map
Definition 3.3.1: Matrix of a linear map
𝑇(𝑣®1 ) ∈ 𝑊 = span 𝑤®1 , . . . , 𝑤®𝑚 =⇒ 𝑇(𝑣®1 ) = 𝑎1,1 𝑤®1 + · · · + 𝑎 𝑚,1 𝑤®𝑚 , 𝑎 𝑖,1 ∈ 𝔽
..
.
𝑇(𝑣®𝑛 ) ∈ 𝑊 = span 𝑤®1 , . . . , 𝑤®𝑚 =⇒ 𝑇(𝑣®𝑛 ) = 𝑎1,𝑛 𝑤®1 + · · · + 𝑎 𝑚,𝑛 𝑤®𝑚 , 𝑎 𝑖,𝑛 ∈ 𝔽
Example 3.3.1
Let 𝑇 : ℝ2 → ℝ3 be a linear map.
With (𝑥, 𝑦) ↦→ (𝑥 + 3𝑦, 2𝑥 + 5𝑦, 7𝑥 + 9𝑦).
Choose standard bases:
38
1 3
Thus, ℳ(𝑇) = 2 5 .
7 9
Example 3.3.2
Differentiation:
Check bases:
1, 𝑥, 𝑥 2 , 𝑥 3 for 𝑃3 (ℝ)
| {z }
𝑉1 ,𝑉2 ,𝑉3 ,𝑉4
1, 𝑥, 𝑥 2
for 𝑃2 (ℝ)
| {z }
𝑊1 ,𝑊2 ,𝑊3
Then:
𝐷(𝑣1 ) = 𝐷(1) = 0 = 0 ·ℝ 1 + 0 ·ℝ 𝑥 + 0 ·ℝ 𝑥 2
𝐷(𝑣2 ) = 𝐷(𝑥) = 1 = 1 ·ℝ 1 + 0 ·ℝ 𝑥 + 0 ·ℝ 𝑥 2
𝐷(𝑣3 ) = 𝐷(𝑥 2 ) = 2𝑥 = 0 ·ℝ 1 + 2 ·ℝ 𝑥 + 0 ·ℝ 𝑥 2
𝐷(𝑣4 ) = 𝐷(𝑥 3 ) = 3𝑥 2 = 0 ·ℝ 1 + 0 ·ℝ 𝑥 + 3 ·ℝ 𝑥 2
Thus,
0 1 0 0
ℳ(𝐷) = 0 0 2 0
0 0 0 3
Thus,
𝑎1,1 ··· 𝑎 1,𝑛
ℳ(𝑇, {𝑣’s} , {𝑤’s}) = ... .. ..
.
.
𝑎 𝑚,1 𝑎 𝑚,𝑛
···
And,
39
𝑏1,1 ··· 𝑏1,𝑛
ℳ(𝑆, {𝑣’s} , {𝑤’s}) = ... .. ..
.
.
𝑏 𝑚,1 𝑏 𝑚,𝑛
···
Therefore:
Thus,
𝑎1,1 + 𝑏1,1 ··· 𝑎1,𝑛 + 𝑏 1,𝑛
.. .. ..
ℳ(𝑆 + 𝑇, {𝑣’s} , {𝑤’s}) = .
. .
𝑎 𝑚,1 + 𝑏 𝑚,1 𝑎 𝑚,𝑛 + 𝑏 𝑚,𝑛
···
So we define addition of matrices so that:
Scalar Multiplication: Let 𝑇 ∈ ℒ(𝑉 , 𝑊) with bases 𝑣®1 , . . . , 𝑣®𝑛 for 𝑉 and 𝑤®1 , . . . , 𝑤®𝑚 for 𝑊.
Remember that 𝑀(𝑇) = 𝑀(𝑇, (𝑣®1 , . . . , 𝑣®𝑛 ), (𝑤®1 , . . . , 𝑤®𝑚 )).
This looks like:
𝑎1,1 ··· 𝑎1,𝑛
.. .. ..
. . .
𝑎 𝑚,1 𝑎 𝑚,𝑛
···
i.e,. 𝑇(𝑣®𝑘 ) = 𝑎1,𝑘 𝑤®1 + · · · + 𝑎 𝑚,𝑘 𝑤®𝑚 .
Then, for 𝜆 ∈ 𝔽, we define 𝜆𝑇 ∈ ℒ(𝑉 , 𝑊) by:
𝜆 · 𝑀(𝑇) ≔ 𝑀(𝜆𝑇)
We compute:
(𝜆 · 𝑇)(𝑣®𝑘 ) ≔ 𝜆 · 𝑇(𝑣®𝑘 )
= 𝜆 · (𝑎1,𝑘 𝑤®1 + · · · + 𝑎 𝑚,𝑘 𝑤®𝑚 )
𝜆𝑎1,1 ··· 𝜆𝑎 1,𝑛
=⇒ 𝑀(𝜆 · 𝑇) = ... .. ..
.
.
𝜆𝑎 𝑚,1 𝜆𝑎 𝑚,𝑛
···
In other words:
𝑎1,1 ··· 𝑎1,𝑛 𝜆𝜆𝑎1,1 ··· 𝜆𝑎1,𝑛
𝜆 · ... .. =
.. .. .. ..
. .
. . .
𝑎 𝑚,1 𝑎 𝑚,𝑛 𝜆𝑎 𝑚,1 𝜆𝑎 𝑚,𝑛
··· ···
40
Composition of maps:
𝑆 𝑇
𝑈→
− 𝑉→
− 𝑍
Now pick bases: 𝑢®1 , . . . , 𝑢®𝑝 for 𝑈, 𝑣®1 , . . . , 𝑣®𝑛 for 𝑉, and 𝑤®1 , . . . , 𝑤®𝑚 for 𝑊.
Let 𝑗 = 1, . . . , 𝑝 and 𝑘 = 1, . . . , 𝑛.
Then:
(𝑇 ◦ 𝑆)(𝑢®𝑗 ) = 𝑇(𝑆(𝑢®𝑗 ))
= 𝑇(𝑏 1,𝑗 𝑣®1 + · · · + 𝑏 𝑛,𝑗 𝑣®𝑛 )
= 𝑏 1,𝑗 𝑇(𝑣®1 ) + · · · + 𝑏 𝑛,𝑗 𝑇(𝑣®𝑛 ) by linearity
= 𝑏 1,𝑗 (𝑎 1,1 𝑤®1 + · · · + 𝑎 𝑚,1 𝑤®𝑚 ) + · · · + 𝑏 𝑛,𝑗 (𝑎1,𝑛 𝑤®1 + · · · + 𝑎 𝑚,𝑛 𝑤®𝑚 )
= (𝑎 1,1 · 𝑏 1,𝑗 + · · · + 𝑎 1,𝑛 · 𝑏 𝑛,𝑗 )𝑤®1 + · · · + (𝑎 𝑚,1 · 𝑏 1,𝑗 + · · · + 𝑎 𝑚,𝑛 · 𝑏 𝑛,𝑗 )𝑤®𝑚
All told:
𝑛 𝑛 𝑛
! ! !
Õ Õ Õ
(𝑇 ◦ 𝑆)(𝑢®𝑗 ) = 𝑎 1,𝑘 · 𝑏 𝑘,𝑗 𝑤®1 + 𝑎 2,𝑘 · 𝑏 𝑘,𝑗 𝑤®2 + · · · + 𝑎 𝑚,𝑘 · 𝑏 𝑘,𝑗 𝑤®𝑚
𝑘=1 𝑘=1 𝑘=1
41
Theorem 3.3.1 Matrix multiplication is associative
𝑎 · · · 𝑎1,𝑛
© 1,1
Let 𝐴 = ... ..
.
.. ª®, where 𝑎 ∈ ℝ.
. ® 𝑖,𝑗
𝑇𝐴 : ℝ𝑛 → ℝ𝑚
𝑒 𝑘 = (0, . . . , 1, . . . , 0) ↦→ (𝑎1,𝑘 , 𝑎 2,𝑘 , . . . , 𝑎 𝑚,𝑘 )
| {z }
𝑘 th place
𝐴 · (𝐵 · 𝐶) = (𝐴 · 𝐵) · 𝐶
Proof: Let
𝑇𝐴 : ℝ𝑚 → ℝ𝑛
𝐴 = 𝑀(𝑇𝐴 )
𝑇𝐵 : ℝ𝑛 → ℝ𝑝
𝐵 = 𝑀(𝑇𝐵 )
𝑇𝐶 : ℝ𝑝 → ℝ𝑟
𝐶 = 𝑀(𝑇𝐶 )
Then:
42
3.4 Invertible Linear Maps
Definition 3.4.1: Invertible
𝑆1 = 𝑆1 ◦ 𝐼𝑑𝑊 = 𝑆1 ◦ (𝑇 ◦ 𝑆2 )
|{z}
since 𝑆2 is an inverse
= (𝑆1 ◦ 𝑇) ◦ 𝑆2 = 𝐼𝑑𝑉 ◦ 𝑆2
|{z}
since 𝑆1 is an inverse
= 𝑆2
Theorem 3.4.1
Let 𝑇 ∈ ℒ(𝑉 , 𝑊) is invertible if and only if 𝑇 is bijective (injective and surjective).
𝑢 = 𝑇 −1 (𝑇(𝑢)) = 𝑇 −1 (𝑇(𝑣)) = 𝑣
So 𝑊 ⊆ 𝐼𝑚(𝑇).
Thus, 𝑇 is bijective.
𝑆: 𝑊 → 𝑉
𝑤 the unique 𝑣 ∈ 𝑉 such that 𝑇(𝑣) = 𝑤
43
(ii) 𝑆 ◦ 𝑇 = 𝐼𝑑𝑣 . We want 𝑆(𝑇(𝑣)) = 𝑣 for all 𝑣 ∈ 𝑉.
(a) Additivity:
One on hand we have:
44
Definition 3.4.2
Proposition 3.4.1
Say 𝑉 , 𝑊 are finite dimensional vector spaces over 𝔽, and 𝑉 𝑊.
Then dim 𝑉 = dim 𝑊.
Proof: If 𝑉 𝑊, then there is an invertible linear map 𝑇 : 𝑉 → 𝑊.
By the rank-nullity theorem, we know that:
Converse is also true (Axler 3.5): If 𝑉 , 𝑊 are finite dimensional vector spaces over 𝔽 and dim 𝑉 = dim 𝑊,
then 𝑉 𝑊.
𝑤
® = 𝑎1 𝑤®1 + · · · + 𝑎 𝑛 𝑤®𝑛 , 𝑎 𝑖 ∈ 𝔽
= 𝑎1 𝑇(𝑣®1 ) + · · · + 𝑎 𝑛 𝑇(𝑣®𝑛 )
= 𝑇(𝑎1 𝑣®1 ) + · · · + 𝑇(𝑎 𝑛 𝑣®𝑛 )
= 𝑇(𝑎1 𝑣®1 + · · · + 𝑎 𝑛 𝑣®𝑛 )
Thus, 𝑇 is injective.
Thus, we have shown that 𝑇 is bijective, and thus 𝑇 is an isomorphism.
45
Example 3.4.1
We know that 𝑃3 (ℂ) and ℂ4 are isomorphic.
Proof gives us:
𝑇 : 1 ↦→ (1, 0, 0, 0)
𝑥 ↦→ (0, 1, 0, 0)
𝑥 2 ↦→ (0, 0, 1, 0)
𝑥 3 ↦→ (0, 0, 0, 1)
Example 3.4.2
Let 𝑉 , 𝑊 be finite dimensional vector spaces over 𝔽.
Choose bases 𝑣®1 , . . . , 𝑣®𝑛 for 𝑉 and 𝑤®1 , . . . , 𝑤®𝑚 for 𝑊.
Let’s define:
𝑀 : ℒ(𝑉 , 𝑊) → 𝐹 𝑚,𝑛
𝑇 ↦→ 𝑀(𝑇, (𝑣®1 , . . . , 𝑣®𝑛 ), (𝑤®1 , . . . , 𝑤®𝑚 ))
Example 3.4.3
Here are some examples:
(i)
𝑇 : 𝑃(ℝ) → 𝑃(ℝ)
𝑝(𝑥) ↦→ 𝑥 2 𝑝(𝑥)
46
(ii)
𝑆 : ℂ∞ → ℂ∞
(𝑥1 , 𝑥 2 , 𝑥 3 , . . .) ↦→ (𝑥2 , 𝑥 3 , . . .)
Theorem 3.4.2
Let 𝑉 be a finite dimensional vector space over 𝔽.
Let 𝑇 ∈ ℒ(𝑉).
Then the following are equivalent:
(i) 𝑇 is injective
(ii) 𝑇 is surjective
(iii) 𝑇 is invertible
Corollary 3.4.1
If 𝑇 : ℝ𝑛 → ℝ𝑛 is linear, then:
47
Question 1
Show that, given 𝑞(𝑥) ∈ 𝑃(ℝ), there exists another polynomial 𝑝(𝑥) such that:
00
𝑞(𝑥) = 𝑥 2 + 2𝑥 + 3 · 𝑝(𝑥)
𝑇 : 𝑃𝑚 (ℝ) → 𝑃𝑚 (ℝ)
00
𝑝(𝑥) ↦→ 𝑥 2 + 2𝑥 + 3 · 𝑝(𝑥)
Thus, the only way for this to be true is if ker 𝑇 = 0𝑃𝑚 (ℝ) .
This implies that 𝑇 is injective.
Then, by the previous theorem, we know that if 𝑇 is injective, then 𝑇 is surjective.
Thus, given 𝑞(𝑥) ∈ 𝑃(ℝ), there exists another polynomial 𝑝(𝑥) such that 𝑇(𝑝(𝑥)) = 𝑞(𝑥).
00
Therefore, 𝑥 2 + 2𝑥 + 3 · 𝑝(𝑥) = 𝑞(𝑥).
Linear Maps as Matrix multiplication: Let 𝑉 be a finite dimensional vector space over 𝔽.
Let 𝑣®1 , . . . , 𝑣®𝑛 be a basis for 𝑉.
Now for any 𝑣 ∈ 𝑉, we can write for some scalars 𝑐 1 , . . . , 𝑐 𝑛 ∈ 𝔽:
𝑣 = 𝑐1 𝑣®1 + · · · + 𝑐 𝑛 𝑣®𝑛
Let’s define:
𝑐1
𝑀(𝑣) ≔ ...
𝑐 𝑛
Example 3.4.4
Let 𝑉 = 𝑃3 (ℝ) with basis 1, 𝑥, 𝑥 2 , 𝑥 3 .
Then,
𝑣 = 2 − 7𝑥 + 5𝑥 3 = 2 · 1 − 7 · 𝑥 + 0 · 𝑥 2 + 5 · 𝑥 3
Or in other words:
2
−7
𝑀(𝑣) =
0
5
48
Note: 𝑀(𝑣0 + 𝑤 0 ) = 𝑀(𝑣 0 ) + 𝑀(𝑤0 ) and 𝑀(𝜆𝑣) = 𝜆𝑀(𝑣).
Say that 𝑇 ∈ ℒ(𝑉 , 𝑊).
Let 𝑤®1 , . . . , 𝑤®𝑚 be a basis for 𝑊.
Then, for any 𝑣 ∈ 𝑉, we can write:
𝑇𝐴 : 𝔽𝑛 ↦→ 𝔽𝑚 linear map
𝑥® ↦→ 𝐴 𝑥® = 𝑏®
49
𝑏
© .1 ª
Question is: 𝐼s .. ®® ∈ image(𝑇𝐴 )?
«𝑏 𝑚 ¬
Row operations are used on the augmented matrix:
𝑎1,1 𝑎 1,2 ... 𝑎1,𝑛 𝑏1
[𝐴 | 𝐵] = ... .. .. .. ..
. ∈ 𝔽𝑚,𝑛+1
. . .
𝑎 𝑚,1 𝑎 𝑚,2 ... 𝑎 𝑚,𝑛 𝑏𝑚
to simplify the original systems of equations.
Need elementary matrices to express row operations: 𝐸 ∈ 𝔽𝑚,𝑚
Thus, we get three types:
𝑎 𝑖,𝑖 ↦→ 0 𝑎 𝑖,𝑗 ↦→ 1
𝑎 𝑗,𝑗 ↦→ 0 𝑎 𝑗,𝑖 ↦→ 1
Then:
1
..
.
0 1
𝐸 = ..
.
1 0
..
.
1
1
..
.
𝐸= 𝑐
..
.
1
Example 3.4.5
50
(i)
1 7 0 1 2 3 29 37 45
0 1 0 4 5 6 = 4 5 6
0 0 1 7 8 9 7 8 9
| {z }
𝐸(𝑖)
(ii)
0 0 1 1 2 3 7 8 9
0 1 0 4 5 6 = 4 5 6
1 0 0 7 8 9 1 2 3
| {z }
𝐸(𝑖𝑖)
(iii)
1 0 0 1 2 3 1 2 3
𝐸 = 0
1 0 4 5 6 = 4 5 6
0 0 3 7 8 9 21 24 27
| {z }
𝐸(𝑖𝑖𝑖)
Lenma 3.4.1
Elementary matrices are invertible:
if 𝐸 is an elementary matrix, then there exists a matrix 𝐸−1 such that 𝐸 · 𝐸−1 = 𝐸−1 · 𝐸 = 𝐼.
1-1
Upshot: Elementary row operations ⇐=⇒ Elementary matrices.
Example 3.4.6
51
1 1 2 1 5 1 1 2 1 5
−𝑅1 +𝑅2 ↦→𝑅2
𝐴 = 1 1 2 6 10 −−−−−−−−−−→
0 0 0 5 5
1 2 5 2 7 1 2 5 2 7
1 1 2 1 5
−𝑅1 +𝑅3 ↦→𝑅3
−−−−−−−−−−→ 0 0 0 5 5
0 1 3 1 2
1 1 2 1 5
𝑅2 ↔𝑅 3
−−−−−→ 0 1 3 1 2
0 0 0 5 5
1 1 2 1 5
5 𝑅 3 ↦→𝑅 3
1
−−−−−−−→ 0 1 3 1 2
0 0 0 1 1
1 0 −1 0 3
−𝑅2 +𝑅1 ↦→𝑅1
−−−−−−−−−−→ 0 1 3 1 2
0 0 0 1 1
1 0 −1 0 3
−𝑅3 +𝑅2 ↦→𝑅2
−−−−−−−−−−→ 0 1 3 0 1 ≔ 𝐴0
0 0 0 1 1
In other words:
1 −1 0
0 1 0 0
𝐴 =
· 0 1 0
0 1 −1 0 0 1
0 0 1
| {z }
−𝑅2 +𝑅1 ↦→𝑅1
Note:-
I didn’t finish the above but therey are equal
𝐴 · 𝑥® = 𝐵𝔽𝑚
|{z} |{z} |{z}
𝔽𝑚,𝑚 𝔽𝑛
𝑀0 = 𝐸 𝑘 · . . . · 𝐸1 ·𝑀
| {z }
elementary matrices (𝑚×𝑚)
𝑎0 𝑎1,2
0
... 𝑎1,𝑛
0
𝑏 10
1,1
= [𝐴0 | 𝐵0] = ... .. .. .. ..
.
. . .
𝑎 𝑎 0𝑚,2 ... 𝑎 0𝑚,𝑛 𝑏 0𝑚
0
𝑚,1
Important: ★ 𝑥® ∈ 𝔽𝑛 | 𝐴 · 𝑥® = 𝐵 = 𝑥® ∈ 𝔽 | 𝐴 · 𝑥® = 𝐵0
𝑛 0
Meaning, the solutions to our original system of equations are the same as the solutions to our modified
system of equations.
Proof: Let 𝑃 = 𝐸 𝑘 · . . . · 𝐸1 is invertible.
Where, 𝑃 −1 = 𝐸1−1 · . . . · 𝐸−1
𝑘
And 𝐼 = 𝑃 −1 · 𝑃 = 𝐸1−1 · . . . · 𝐸−1
𝑘
· 𝐸 𝑘 · . . . · 𝐸1
Say 𝑥® ∈ LHS of ★:
52
Where 𝑀0 = 𝑃 · 𝑀 = 𝑃 ∗ 𝐴 | 𝑃 ∗ 𝐵
|{z} |{z}
𝐴0 𝐵0
𝐴 · 𝑥® = 𝐵
𝑃 · 𝐴 · 𝑥® = 𝑃 · 𝐵
𝐴0 · 𝑥® = 𝐵0
=⇒ 𝑥® ∈ RHS.
(Reduced) Row-Echelon form: Notation: 𝑀 ∈ 𝔽𝑚,𝑛 , write 𝑀 𝑖 for the 𝑖th row of 𝑀.
Definition 3.4.4
(ii) If 𝑀 𝑖 ≠ (0, . . . , 0), then the left most nonzero entry is a 1 (pivot).
(iii) If 𝑀 𝑖+1 ≠ (0, . . . , 0) as well, then the pivot in 𝑀 𝑖+1 is to the right of the pivot in 𝑀 𝑖 .
(iv) The entries above and below a pivot are 0.
Example 3.4.7
Think 𝔽 = ℚ, ℝ, or ℂ.
1 0 −1 0 3
0 1 3 0 1
0 0 0 1 1
0 0 0 0 0
Theorem 3.4.3
Let 𝑀 ∈ 𝔽𝑚,𝑛 . There is a sequence of elementary row operations, 𝐸 𝑘 , . . . , 𝐸1 ,
such that 𝑀 0 = 𝐸 𝑘 · . . . · 𝐸1 · 𝑀 is in row-echelon form.
𝑀 0 is unique.
𝐴0 · 𝑥® =𝐵0
𝑥1 + 6𝑥2 + 𝑥4 =0
𝑥 3 + 2𝑥4 =0
0 =1
53
If instead we had:
1 6 0 1 1
0 0 0
𝑀 = [𝐴 | 𝐵 ] = 0 0 1 2 3
0 0 0 0 0
Thus would imply that:
𝑥1 + 6𝑥 2 + 𝑥4 =1
𝑥3 + 2𝑥4 =3
0 =0
𝑥1 = 1 − 6𝑎 − 𝑏
𝑥3 = 3 − 2𝑏
=⇒ 𝑥® = (𝑥1 , 𝑥 2 , 𝑥 3 , 𝑥 4 ) = (1 − 6𝑎 − 𝑏, 𝑎, 3 − 2𝑏, 𝑏)
In general:
Let 𝑀 0 = [𝐴0 | 𝐵0] be in row-echelon form.
Lenma 3.4.2
Let 𝑥®𝑠 be a solution to 𝑇( 𝑥®) = 𝑏.
®
Where 𝑇 is a linear map that maps 𝑥® ∈ ℝ𝑛 to 𝑏® ∈ ℝ𝑚 by a matrix 𝐴: 𝑇( 𝑥®) = 𝐴 · 𝑥®.
Then, if there are other solutions, 𝑥®★, to 𝑇( 𝑥®) = 𝑏,
®
Then there exists an 𝑥®𝑘 ∈ ker 𝑇 such that every other solution is given by:
𝑇𝐴 : ℝ𝑛 → ℝ𝑚
𝑥® ↦→ 𝐴 · 𝑥®
54
Remember:
We can write:
𝐴 = 𝑀(𝑇𝐴 , (𝑒1 , . . . , 𝑒 𝑛 ), ( 𝑓1 , . . . , 𝑓𝑚 ))
n o
ker 𝑇𝐴 = 𝑥® ∈ ℝ𝑛 | 𝐴 · 𝑥® = 0®ℝ𝑚
n o
= 𝑥® ∈ ℝ𝑛 | 𝐴0 · 𝑥® = 0®ℝ𝑚 where 𝐴0 is in row-echelon form
= ker 𝑇𝐴0
Example 3.4.8
Let:
1 6 0 1
𝐴0 = 0
0 1 2 ∈ ℝ3,4
0 0 0 0
This implies:
𝐴0 · 𝑥® = 0®ℝ3
𝑥1 + 6𝑥 2 + 𝑥4 = 0
𝑥3 + 2𝑥4 = 0
0=0
𝑥1 = −6𝑎 − 𝑏
𝑥3 = −2𝑏
Solutions:
𝑎 = 1, 𝑏 = 0 =⇒ (−6, 1, 0, 0)
𝑎 = 0, 𝑏 = 1 =⇒ (−1, 0, −2, 1)
Thus:
55
Images: Given 𝑇𝐴 : ℝ𝑛 → ℝ𝑚 .
Compute a basis 𝑣®1 , . . . , 𝑣®𝑟 for the kernel.
Let 𝑖1 , . . . , 𝑖 𝑛−𝑟 be the indices of the pivot columns of 𝐴0
Example 3.4.9
1 1 2 1
𝐴 = 1 1 2 6
1 2 5 2
This implies: 𝑇𝐴 : ℝ4 → ℝ3 .
Row-echelon form of 𝐴 is:
1 0 −1 0
0
𝐴 = 0 1 3 0 , 𝑖 1 = 1, 𝑖 2 = 2, 𝑖 3 = 4
0 0 0 1
𝑇𝐴 : 𝔽𝑛 → 𝔽𝑛
𝑥® ↦→ 𝐴 · 𝑥®
Theorem 3.4.4
Let 𝐴 ∈ 𝔽𝑛,𝑛 . The following are equivalent (TFAE):
56
Proof of 1 =⇒ 2: Let 𝐼𝑛 = 𝐴0 = 𝐸 𝑘 · . . . · 𝐸1 · 𝐴.
Since elementary matrices are invertible: (𝐸 𝑘 · . . . · 𝐸1 )−1 = 𝐸1−1 · . . . · 𝐸−1
𝑘
.
Then, 𝐴 = 𝐸1−1 · . . . · 𝐸−1
𝑘
.
But 𝐸−1
𝑖
is elementary for 1 ≤ 𝑖 ≤ 𝑘 are also elementary matrices.
Thus, 𝐴 is a product of elementary matrices.
𝐸1 · . . . · 𝐸 𝑘 = 𝐴
𝐸−1 −1
𝑘 · . . . · 𝐸1 = 𝐵 = 𝐴
−1
Thus:
𝐴 · 𝐵 = 𝐸1 · . . . · 𝐸 𝑘 · 𝐸−1 −1
𝑘 · . . . · 𝐸1 = 𝐼 𝑛
Thus, 𝐴 is invertible.
Proof of 3 =⇒ 1 : Assume 𝐴 is invertible.
Let 𝐴0 = 𝐸 𝑘 · . . . · 𝐸1 · 𝐴 be the row-echelon form of 𝐴.
Either 𝐴0 = 𝐼𝑛 or the bottom row of 𝐴0 is (0, . . . , 0).
If the bottom row of 𝐴0 has all zeros, then:
Note:-
Notice that we started to talk about determinants after section 3.C.
I’ve moved this to chapter 10 to correspond with the textbook.
Click here to go to the determinants section: Determinants
𝑉1 × . . . × 𝑉𝑚 = {(𝑣1 , . . . , 𝑣 𝑚 ) | 𝑣 𝑖 ∈ 𝑉𝑖 for 1 ≤ 𝑖 ≤ 𝑚}
I.e., think of this in terms of a cartesian product.
Example 3.5.1
Elements of ℝ2 × ℝ3 look like:
57
Example 3.5.2
Vectors in 𝑃2 (ℝ) × ℝ looks like:
Definition 3.5.2
(𝑣1 , . . . , 𝑣 𝑚 ) + (𝑤 1 , . . . , 𝑤 𝑚 ) = (𝑣1 + 𝑤 1 , . . . , 𝑣 𝑚 + 𝑤 𝑚 )
𝜆 · (𝑣1 , . . . , 𝑣 𝑚 ) = (𝜆 · 𝑣1 , . . . , 𝜆 · 𝑣 𝑚 )
Proposition 3.5.1
If 𝑉1 , . . . , 𝑉𝑚 are finite dimensional over 𝔽, then so is 𝑉1 × . . . × 𝑉𝑚 .
In fact, the dimension of 𝑉1 × . . . × 𝑉𝑚 is:
{(𝑣 1,1 , 0, . . . , 0), . . . , (𝑣 1,𝑚 , 0, . . . , 0), (0, 𝑣 2,1 , 0, . . . , 0), . . . , (0, 𝑣 𝑚,𝑚 )}
Example 3.5.3
𝑃2 (ℝ) × ℝ2 :
(1, (0, 0)), (𝑥, (0, 0)), (𝑥 2 , (0, 0)), (0, (1, 0)), (0, (0, 1))
Γ : 𝑈1 × . . . × 𝑈 𝑚 → 𝑈1 + . . . + 𝑈 𝑚
So, Γ(𝑢1 , . . . , 𝑢𝑚 ) = 𝑢1 + . . . + 𝑢𝑚 .
Is Γ a linear map?
Proof of linear map: Vector addition:
Thus, it is homogeneous.
Therefore, Γ is a linear map as desired.
Moreover:
(i) Γ is surjective:
if 𝑢1 + . . . + 𝑢𝑚 ∈ 𝑈1 + . . . + 𝑈𝑚 , then Γ((𝑢1 , . . . , 𝑢𝑚 )) = 𝑢1 + . . . + 𝑢𝑚 .
Example 3.5.4
𝑉 = ℝ2 , 𝑈 = {(𝑥, 2𝑥) | 𝑥 ∈ ℝ}.
Then 𝑣 + 𝑈 is the set of all lines parallel to 𝑈.
Let 𝑣1 = (3, 1) and 𝑣 2 = (4, 3).
59
𝑣 + 𝑈 = {(3, 1) + (𝑥, 2𝑥) | 𝑥 ∈ ℝ}
= {(3 + 𝑥, 1 + 2𝑥) | 𝑥 ∈ ℝ}
= {(4 + 𝑥, 3 + 2𝑥) | 𝑥 ∈ ℝ}
Lenma 3.5.1
(i) 𝑣1 + 𝑈 = 𝑣2 + 𝑈
(ii) 𝑣2 − 𝑣1 ∈ 𝑈
Proof of 𝑖𝑖 =⇒ 𝑖 : Let 𝑣 ∈ 𝑣1 + 𝑈.
So 𝑣 = 𝑣1 + 𝑢 for some 𝑢 ∈ 𝑈.
𝑣 = 𝑣1 + 𝑢 = 𝑣2 − 𝑣2 − 𝑣1 + 𝑢
= 𝑣 2 + (𝑣 1 − 𝑣 2 ) + 𝑢 ∈ 𝑣 2 + 𝑈
Similarly, 𝑣2 + 𝑈 ⊆ 𝑣1 + 𝑈.
(𝑣 + 𝑈) +𝑉\𝑈 (𝑤 + 𝑈) = (𝑣 +𝑉 𝑤) + 𝑈
Scaler multiplication
60
𝜆 · (𝑣 + 𝑈) ≔ 𝜆 · 𝑣 + 𝑈
Have to check that, e.g, addition is well defined:
Say 𝑣1 + 𝑈 = 𝑣2 + 𝑈 and 𝑤 1 + 𝑈 = 𝑤 2 + 𝑈.
Then we need to show that:
?
z}|{
(𝑣1 + 𝑈) + (𝑤 1 + 𝑈) = (𝑣 2 + 𝑈) + (𝑤 2 + 𝑈)
?
z}|{
(𝑣 1 + 𝑤 1 ) + 𝑈 = (𝑣 2 + 𝑤 2 ) + 𝑈
⇐⇒ (𝑣1 + 𝑤 1 ) − (𝑣2 + 𝑤 2 ) ∈ 𝑈
⇐⇒ (𝑣1 − 𝑣2 ) + (𝑤 1 − 𝑤 2 ) ∈ 𝑈
| {z } | {z }
∈𝑈 ∈𝑈
=⇒ 𝑣1 − 𝑣2 ∈ 𝑈
=⇒ 𝜆 · (𝑣 1 − 𝑣 2 ) ∈ 𝑈
=⇒ 𝜆 · 𝑣 1 − 𝜆 · 𝑣2 ∈ 𝑈
=⇒ 𝜆 · 𝑣 1 + 𝑈 = 𝜆 · 𝑣2 + 𝑈
=⇒ 𝜆 · (𝑣 1 + 𝑈) = 𝜆 · (𝑣 2 + 𝑈)
Let’s give an example:
Example 3.5.6
Let:
𝜋 :𝑉 \𝑈
𝑣 ↦→ 𝑣 + 𝑈
n n oo
ker 𝜋 = 𝑣 ∈ 𝑉 : 𝜋(𝑣) = 0𝑉\𝑈
®
n o
= 𝑣 ∈ 𝑉 : 𝜋(𝑣) = 0®𝑣 + 𝑈
= {𝑣 ∈ 𝑉 : 𝑣 + 𝑈 = 𝑈 }
= {𝑣 ∈ 𝑉 : 𝑣 ∈ 𝑈 }
=𝑉 \𝑈
61
Theorem 3.5.1 1st isomorphism theorem
Let 𝑇 ∈ ℒ(𝑉 , 𝑊) with 𝐼𝑚(𝑇) ⊆ 𝑊, ker 𝑇 ⊆ 𝑉.
Which means that 𝑉 → 𝑉 \ ker 𝑇.
Let’s define the following:
𝑇
e : 𝑉 \ ker 𝑇 → 𝐼𝑚(𝑇)
𝑣 + ker 𝑇 ↦→ 𝑇(𝑣)
(i) 𝑇
e is well defined.
If 𝑣 1 + ker 𝑇 = 𝑣2 + ker 𝑇, then we want to show that 𝑇(𝑣
e 1 + ker 𝑇) = 𝑇(𝑣
e 2 + ker 𝑇).
We have:
𝑣1 − 𝑣2 ∈ ker 𝑇
𝑇(𝑣 1 − 𝑣 2 ) = 0®𝑊
𝑇(𝑣1 ) − 𝑇(𝑣2 )
(ii) 𝑇
e is linear.
e.g., 𝑇((𝑣
e + ker 𝑇) + (𝑤 + ker 𝑇)) = 𝑇((𝑣
e + 𝑤) + ker 𝑇).
This is equal to 𝑇(𝑣 + 𝑤) = 𝑇(𝑣) + 𝑇(𝑤)
Meaning that 𝑇(𝑣
e + ker 𝑇) + 𝑇(𝑤
e + ker 𝑇).
We leave homogeneity as an exercise.
(iii) 𝑇
e is injective!
Say 𝑇(𝑣
e + ker 𝑇) = 0®𝑊
This means that 𝑇(𝑣) = 0®𝑊 .
Which implies that 𝑣 ∈ ker 𝑇.
Hence, 0® + ker 𝑇 = 𝑣 + ker 𝑇.
This is 0𝑉\ker
® 𝑇
(iv) 𝑇
e is surjective!
Let 𝑤 ∈ 𝐼𝑚(𝑇).
Then 𝑤 = 𝑇(𝑣) for some 𝑣 ∈ 𝑉.
Which means that 𝑤 = 𝑇(𝑣
e + ker 𝑇).
Thus, 𝑇
e ∈ ℒ(𝑉 \ ker 𝑇, 𝐼𝑚(𝑇)) is an isomorphism of 𝔽 vector spaces.
i.e.,
𝑉 \ ker 𝑇 𝐼𝑚(𝑇)
62
Chapter 4
Polynomials
Definition 4.0.1
Definition 4.0.2
Let 𝑧 ∈ ℂ, then
The complex conjugate of 𝑧 is 𝑧p= <(𝑧) − =(𝑧)𝑖.
The absolute value of 𝑧 is |𝑧| = <(𝑧)2 + =(𝑧)2 .
𝑧 = 𝑎 + 𝑏𝑖
𝑤 = 𝑐 + 𝑑𝑖
𝑧 = 𝑎 − 𝑏𝑖
𝑤 = 𝑐 − 𝑑𝑖
(i) Sum of 𝑧 and 𝑧: 𝑧 + 𝑧 = 2<(𝑧)
Proof:
𝑧 + 𝑧 = (𝑎 + 𝑏𝑖) + (𝑎 − 𝑏𝑖)
= 2𝑎
= 2<(𝑧)
Proof:
𝑧 − 𝑧 = (𝑎 + 𝑏𝑖) − (𝑎 − 𝑏𝑖)
= 2𝑏𝑖
= 2=(𝑧)𝑖
63
(iii) Product of 𝑧 and 𝑧: 𝑧𝑧 = |𝑧| 2
Proof:
𝑧𝑧 = (𝑎 + 𝑏𝑖)(𝑎 − 𝑏𝑖)
= 𝑎 2 − 𝑎𝑏𝑖 + 𝑎𝑏𝑖 − 𝑏 2 𝑖 2
= 𝑎2 + 𝑏2
= |𝑧| 2
Proof:
𝑧 + 𝑤 = (𝑎 − 𝑏𝑖) + (𝑐 − 𝑑𝑖)
= (𝑎 + 𝑐) − (𝑏 + 𝑑)𝑖
=𝑤+𝑧
Proof:
𝑤 · 𝑧 = (𝑐 − 𝑑𝑖)(𝑎 − 𝑏𝑖)
= 𝑎𝑐 − 𝑎𝑑𝑖 − 𝑏𝑐𝑖 − 𝑏𝑑𝑖 2
= 𝑎𝑐 − 𝑎𝑑𝑖 − 𝑏𝑐𝑖 + 𝑏𝑑
= (𝑎𝑐 + 𝑏𝑑) − (𝑎𝑑 + 𝑏𝑐)𝑖
= 𝑤𝑧
Proof:
𝑧 = 𝑎 − 𝑏𝑖
= 𝑎 + 𝑏𝑖
=𝑧
Proof:
|𝑧| 2 = 𝑧𝑧
= (𝑎 + 𝑏𝑖)(𝑎 − 𝑏𝑖)
= 𝑎2 + 𝑏2
|𝑧| 2 ¾ 𝑎 2
|𝑧| 2 ¾ 𝑏 2
|𝑧| ¾ 𝑎
|𝑧| ¾ 𝑏
64
(viii) Absolute value of the complex conjugate: |𝑧| = |𝑧|
Proof:
|𝑧| = |𝑎 − 𝑏𝑖|
√
= 𝑎2 + 𝑏2
= |𝑧|
Proof:
|𝑤𝑧| 2 = (𝑤𝑧)(𝑤𝑧)
p
|𝑤𝑧| = (𝑤𝑧)(𝑤𝑧)
p
= (𝑤𝑤)(𝑧𝑧)
√ √
= 𝑤𝑤 𝑧𝑧
= |𝑤||𝑧|
Proof:
|𝑤 + 𝑧| 2 = (𝑤 + 𝑧)(𝑤 + 𝑧)
= 𝑤𝑤 + 𝑤𝑧 + 𝑧𝑤 + 𝑧𝑧
= |𝑤| 2 + 𝑤𝑧 + 𝑤𝑧 + |𝑧| 2
= |𝑤| 2 + |𝑧| 2 + 2<(𝑤𝑧)
¶ |𝑤| 2 + |𝑧| 2 + 2 |𝑤𝑧|
¶ |𝑤| 2 + |𝑧| 2 + 2 |𝑤| |𝑧|
= (|𝑤| + |𝑧|)2
65
Definition 4.0.3
Theorem 4.0.1
Let 𝑎 0 , . . . , 𝑎 𝑚 ∈ 𝔽. If:
𝑎0 + 𝑎1 𝑥 + . . . + 𝑎 𝑚 𝑥 𝑚 = 0
For every 𝑥 ∈ 𝔽, then 𝑎0 = . . . = 𝑎 𝑚 = 0.
Proof: Assume the contrapositive. Let our polynomial be given by
𝑝(𝑥) = 𝑎0 + 𝑎1 𝑥 + · · · + 𝑎 𝑚 𝑥 𝑚
If this polynomial is not the zero function, then there exists some coefficient 𝑎 𝑘 ≠ 0.
Without loss of generality, let’s assume that 𝑎 𝑚 is that coefficient.
We want to show that there exists some value 𝑥 = 𝑧 for which the polynomial does not evaluate to zero.
Specifically, we’ll show that the term 𝑎 𝑚 𝑧 𝑚 will dominate all other terms for a sufficiently large 𝑧,
such that the polynomial cannot evaluate to zero.
To do this, let’s choose 𝑧 such that
Í𝑚−1
𝑗=0 |𝑎 𝑗 |
𝑧>
|𝑎 𝑚 |
Given this choice of 𝑧, the magnitude of the term 𝑎 𝑚 𝑧 𝑚 will exceed the combined magnitudes of all the
other terms:
66
Question 2
(a) Show that if 𝑝 has degree 𝑛 > 0, then there is some monomial 𝑞 such that 𝑝 − (𝑥 − 𝑐)𝑞 is a polynomial
of degree less than 𝑛. (A monomial is a polynomial that has only one non-zero term.)
(b) Suppose that 𝑝 is a polynomial with a root at 𝑥 = i.e., 𝑝(𝑐) = 0. Show that (𝑥 − 𝑐) is a factor of 𝑝
(that is, there is some polynomial 𝑟 such that 𝑝 = (𝑥 − 𝑐)𝑟 ).
Proof of 𝑎: Given polynomial 𝑝 with degree 𝑛 > 0, and the form 𝑝(𝑥) = 𝑎0 + 𝑎1 𝑥 + 𝑎2 𝑥 2 + . . . + 𝑎 𝑛 𝑥 𝑛 , for 𝑎 𝑖 ∈ ℝ.
Let’s fix 𝑐, now, we want to show that there is some monomial 𝑞
such that 𝑝 − (𝑥 − 𝑐)𝑞 is a polynomial of degree less than 𝑛.
Let’s proceed by induction on 𝑛 ∈ ℕ,
Base Case: Let 𝑛 = 1, which means that 𝑝(𝑥) has a degree of 1.
𝑝(𝑥) = 𝑎0 + 𝑎1 𝑥
Clearly, we can pick 𝑞 = 𝑎1 (as it is a monomial).
Moreover, if we solve for 𝑝 − (𝑥 − 𝑐)𝑞, we get
𝑝 − (𝑥 − 𝑐)𝑞 = (𝑎 0 + 𝑎1 𝑥) − (𝑥 − 𝑐)(𝑎1 )
= 𝑎0 + 𝑎1 𝑥 − 𝑎1 𝑥 + 𝑎1 𝑐
= 𝑎0 + 𝑎1 𝑐
Notice, that 𝑎0 + 𝑎1 𝑐 is a constant polynomial, meaning that its degree is 0, which is less than 1.
Hence, the base case holds.
Inductive Step: Assume the statement holds for all polynomials 𝑝 with degree less than 𝑛.
Thus, for all 𝑘 < 𝑛, 𝑘 ∈ ℕ, we have a monomial 𝑞 s.t. 𝑝 − (𝑥 − 𝑐)𝑞 is a polynomial of degree less than 𝑘.
Now, we want to show that the statement holds for 𝑛.
Let’s consider a polynomial 𝑝 with degree 𝑛, then we can write:
𝑝(𝑥) = 𝑎 𝑛 𝑥 𝑛 + 𝑝 𝑛−1 (𝑥), where 𝑝 𝑛−1 (𝑥) is a polynomial of degree less than 𝑛
By our inductive hypothesis, we know that there is some monomial 𝑞 𝑛−1 (𝑥)
such that 𝑝 𝑛−1 − (𝑥 − 𝑐)𝑞 𝑛−1 is a polynomial of degree less than 𝑛 − 1.
Combining this information, let’s pick 𝑞(𝑥) = 𝑎 𝑛 𝑥 𝑛−1 . Clearly, 𝑞(𝑥) is a monomial.
Thus, we have:
𝑠(𝑥) = (𝑥 − 𝑐)𝑡(𝑥)
Substituting this into our original equation, we get:
68
Chapter 5
Definition 5.1.1
𝑢 ∈ 𝑈 =⇒ 𝑇(𝑢) ∈ 𝑈
in other words, if 𝐼𝑚(𝑇 |𝑈 ) ⊆ 𝑈,
or 𝑇 |𝑈 : 𝑈 → 𝑈, i.e., 𝑇 | 𝑢 ∈ ℒ(𝑈) where 𝑇 : 𝑉 → 𝑉.
Example 5.1.1
What does a 1 dimensional invariant subspace under 𝑇 look like?
𝑈 = span (𝑣). Then 𝑇(𝑣) ∈ 𝑈, so 𝑇(𝑣) = 𝜆𝑣 for some 𝜆 ∈ 𝔽.
Conversely, if 𝑣 ≠ 0®𝑣 and 𝑇(𝑣) = 𝜆𝑣 for some 𝜆 ∈ 𝔽,
then 𝑈 = span (𝑣) is 1-dimensional invariant subspace under 𝑇.
We call 𝜆 an eigenvalue of 𝑇.
If 𝑣 ≠ 0®𝑣 , then 𝑣 is an eigenvector for the eigenvalue 𝜆.
Proposition 5.1.1
Suppose that 𝑉 is a finite-dimensional vector space over 𝔽 and 𝑇 ∈ ℒ(𝑉).
Then the following are equivalent:
(a) 𝜆 ∈ 𝔽 is an eigenvalue of 𝑇.
(b) 𝑇 − 𝜆𝐼𝑑 is not injective.
(c) 𝑇 − 𝜆𝐼𝑑 is not surjective.
(d) 𝑇 − 𝜆𝐼𝑑 is not invertible.
69
(𝑇 − 𝜆𝐼𝑑)(𝑣) = 𝑇(𝑣) − 𝜆𝐼𝑑(𝑣)
= 𝑇(𝑣) − 𝜆𝑣
Claim 5.1.1
Eigenvectors corresponding to distinct eigenvalues are linearly independent.
Let 𝑇 ∈ ℒ(𝑉), and let 𝜆1 , . . . , 𝜆𝑚 be distinct eigenvalues of 𝑇,
with eigenvectors 𝑣 1 , . . . , 𝑣 𝑚 respectively.
Then 𝑣1 , . . . , 𝑣 𝑚 are linearly independent.
Proof: Suppose for contradiction that 𝑣1 , . . . , 𝑣 𝑚 are linearly dependent.
Then by the linear dependence lemma, there exists a (smallest) 𝑘 ∈ {1, . . . , 𝑚} such that
𝑎1 (𝜆 𝑘 − 𝜆1 ) = . . . = 𝑎 𝑘−1 (𝜆 𝑘 − 𝜆 𝑘−1 ) = 0
Since we are given that 𝜆1 , . . . , 𝜆𝑚 are distinct, we have that 𝜆 𝑘 − 𝜆 𝑖 ≠ 0 for all 𝑖 ∈ {1, . . . , 𝑘 − 1}.
This means that 𝑎1 = . . . = 𝑎 𝑘−1 = 0𝔽 .
Thus, 𝑣 𝑘 = 0®𝑣 thus 𝑣 𝑘 is not an eigenvector.
Which is a contradiction!
70
The first part shows that there exists 𝑣 ≠ 0®𝑣 , 𝑇(𝑣) = 𝜆 · 𝑣
Theoremn 5.1.1
o
Let 𝑣 ≠ 0®𝑣 be a finite-dimensional vector space over ℂ.
Let 𝑇 ∈ ℒ(𝑉).
Then 𝑇 has at least one eigenvalue.
n o
Proof: Let 𝑛 = dim 𝑉. Note 𝑛 ¾ 1, since 𝑣 ≠ 0®𝑣 .
Then det(𝐴 − 𝑥𝐼𝑛 ) is a polynomial of degree 𝑛 with complex coefficients.
By the fundamental theorem of algebra (proved in Math 427),
a non-constant polynomial with complex coefficients has a root in ℂ.
Thus, there exists 𝜆 ∈ ℂ such that det(𝐴 − 𝜆𝐼𝑛 ) = 0.
Example 5.1.2
Let
1 4 5
𝐴 = 0 2 6®
© ª
«0 0 3¬
Let:
1 4 5 𝑥 0 0
det(𝐴 − 𝑥 · 𝐼3 ) = det 0 2 6® − 0 𝑥 0 ®®
©© ª © ªª
« «0 0 3¬ « 0 0 𝑥 ¬¬
Thus, we get the following:
1−𝑥 4 5
det 0 2−𝑥 6 ®® = (1 − 𝑥)(2 − 𝑥)(3 − 𝑥)
©© ªª
«« 0 0 3 − 𝑥 ¬¬
Roots of characteristic polynomial are 𝑥 = 1, 2, or3.
Change of basis: Does the characteristic polynomial depend on 𝐴, or does it depend only on 𝑇?
𝑇 : 𝔽𝑛 → 𝔽𝑛
We can have:
𝐴 = ℳ(𝑇, (𝑒1 , . . . , 𝑒 𝑛 ))
𝐴0 = ℳ(𝑇, ( 𝑓1 , . . . , 𝑓𝑛 )) 𝑓 𝑗 = 𝑎 1,𝑗 𝑒1 + . . . + 𝑎 𝑛,𝑗 𝑒 𝑛
𝑎 ... 𝑎1,𝑛
© 1,1
𝑃 = ... ..
.
.. ª®
. ®
« 𝑎 𝑛,1 ... 𝑎 𝑛,𝑛 ¬
Where we get a new basis in terms of old basis.
To get from 𝐴 to 𝐴0:
𝐴0 = 𝑃 −1 · 𝐴 · 𝑃
|{z} |{z} |{z}
converts e’s to f’s apply T WRT e’s converts f’s to e’s
71
What does this mean for our characteristic polynomial?: We have that:
1 2 3
0 4 5®
© ª
« 0 0 6¬
Proposition 5.1.2
Let 𝑇 ∈ ℒ(𝑉). 𝑉 is a finite-dimensional vector space over 𝔽.
Let 𝑣®1 , . . . , 𝑣®𝑛 be a basis for 𝑉.
The following are equivalent:
1. ℳ(𝑇, (𝑣1 , . . . , 𝑣 𝑛 )) is upper triangular.
2. 𝑇(𝑣 𝑗 ) ∈ span 𝑣1 , . . . , 𝑣 𝑗 for all 𝑗 ∈ {1, . . . , 𝑛}.
..
.
𝑇(𝑣 𝑗 ) ∈ span 𝑣1 , . . . , 𝑣 𝑗
Since 𝑇 is linear.
Then this implies that 𝑇(𝑎1 𝑣1 + . . . + 𝑎 𝑗 𝑣 𝑗 ) = 𝑎1 𝑇(𝑣1 ) + . . . + 𝑎 𝑗 𝑇(𝑣 𝑗 ) ∈ span 𝑣1 , . . . , 𝑣 𝑗 .
Hence, 𝑇(span 𝑣1 , . . . , 𝑣 𝑗 ) ⊆ span 𝑣1 , . . . , 𝑣 𝑗 .
72
Theorem 5.1.2
Let 𝑉 be a finite-dimensional vector space over ℂ and 𝑇 ∈ ℒ(𝑉).
Then there exists a basis 𝑣®1 , . . . , 𝑣®𝑛 of 𝑉 such that ℳ(𝑇, (𝑣1 , . . . , 𝑣 𝑛 )) is upper triangular (UT).
For this, we need the above proposition.
Proof: Let’s proceed on induction on 𝑛 = dim 𝑉.
Base Claim: Let 𝑛 = 1, then clearly every 1 × 1 matrix is upper triangular.
Inductive Step: Assume that the statement holds for all 𝑆 ∈ ℒ(𝑊) with dim 𝑊 < dim 𝑉.
Let 𝜆 ∈ ℂ be an eigenvalue for 𝑇. This means it exists such that 𝑣 ≠ 0®𝑣 .
Now consider (𝑇 − 𝜆 · 𝐼𝑑 : 𝑉 → 𝑉).
Set 𝑊 = 𝐼𝑚(𝑇 − 𝜆 · 𝐼𝑑) ⊆ 𝑉.
𝑇(𝑤) = 𝑇(𝑤) − 𝜆 · 𝑤 + 𝜆 · 𝑤
= (𝑇 − 𝜆 · 𝐼𝑑)(𝑤) + 𝜆 · 𝑤
| {z } |{z}
∈𝑊 by definition ∈𝑊
𝑇 |𝑊 : 𝑊 → 𝑊
This implies that 𝑇 |𝑊 ∈ ℒ(𝑊).
By our inductive hypothesis, there exists a basis 𝑤®1 , . . . , 𝑤®𝑚 of 𝑊 such that
ℳ(𝑇 |𝑊 , (𝑤 1 , . . . , 𝑤 𝑚 )) is upper triangular.
By our proposition, 𝑇(𝑤 𝑗 ) ∈ span 𝑤1 , . . . , 𝑤 𝑗 for some 𝑗.
Extend to a basis for 𝑉: 𝑤®1 , . . . , 𝑤®𝑚 , 𝑣®1 , . . . , 𝑣 𝑛−𝑚
® .
Then, for 𝑘 = 1, . . . , 𝑛 − 𝑚, we have:
𝑇(𝑣 𝑘 ) = 𝑇(𝑣 𝑘 ) − 𝜆 · 𝑣 𝑘 + 𝜆 · 𝑣 𝑘
= (𝑇 − 𝜆 · 𝐼𝑑)(𝑣 𝑘 ) +𝜆 · 𝑣 𝑘
| {z }
∈𝑊=span 𝑤®1 ,..., 𝑤®𝑚
∈ span (𝑤 1 , . . . , 𝑤 𝑚 , 𝑣 𝑘 ) ⊆ span (𝑤 1 , . . . , 𝑤 𝑚 , 𝑣 1 , . . . , 𝑣 𝑘 )
73
𝜆 ★
© 1 ..
ℳ(𝑇, (𝑣 1 , . . . , 𝑣 𝑛 )) =
ª
. ®
®
«0 𝜆𝑛 ¬
Now, we proofed with the biconditional.
𝑎 1,2 𝑎1,2
1 1
𝑣2 = 𝑇(𝑣2 ) − 𝑣1 = 𝑇 𝑣2 − 𝑣1
𝜆2 𝜆2 𝜆2 𝜆2 |{z}
| {z } ∈𝐼𝑚(𝑇)
∈𝐼𝑚(𝑇)
| {z }
∈𝐼𝑚(𝑇) as it is a subspace
Theorem 5.1.3
If ℳ(𝑇) is UT then the eigenvalues of 𝑇 are the diagonal entries of ℳ(𝑇).
Proof: Say:
𝜆 ★
© 1 ..
ℳ(𝑇, (𝑣 1 , . . . , 𝑣 𝑛 )) =
ª
. ®
®
«0 𝜆𝑛 ¬
Let 𝜆 ∈ 𝔽, then ℳ(𝑇 − 𝜆 · 𝐼𝑑) is UT.
74
𝜆1 − 𝜆 ★
ℳ(𝑇 − 𝜆 · 𝐼𝑑, (𝑣1 , . . . , 𝑣 𝑛 )) = ..
© ª
. ®
®
« 0 𝜆 𝑛 − 𝜆¬
Hence:
5.2 Eigenspaces
Definition 5.2.1: Eigenspaces
(𝑇 − 𝜆 · 𝐼𝑑)(𝑣) = 0
𝑇(𝑣) − (𝜆 · 𝐼𝑑)(𝑣) = 0
𝑇(𝑣) = 𝜆 · 𝑣
Proposition 5.2.1
Let 𝑉 be a finite-dimensional vector space over 𝔽 and 𝑇 ∈ ℒ(𝑉).
Let 𝜆1 , . . . , 𝜆𝑚 be distinct eigenvalues.
Then 𝐸(𝜆1 , 𝑇), . . . , 𝐸(𝜆𝑚 , 𝑇) ⊆ 𝑉 is a direct sum.
Moreover:
75
Definition 5.2.2
𝜆 0
© 1 ..
𝐴 =
ª
. ®
®
«0 𝜆𝑛 ¬
Then, 𝑇 ∈ ℒ(𝑉) is diagonalizable if there is a basis 𝑣1 , . . . , 𝑣 𝑛 of 𝑉 such that:
ℳ(𝑇, (𝑣1 , . . . , 𝑣 𝑛 )) is diagonal.
Example 5.2.1
𝑇 : ℝ3 → ℝ3 with ℳ(𝑇, (𝑒1 , 𝑒2 , 𝑒3 ) is:
5 0 0
0 8 0®
© ª
« 0 0 8¬
Then 𝑇(𝑥, 𝑦, 𝑧) = (5𝑥, 8𝑦, 8𝑧).
With 𝑇(𝑒1 ) = 5𝑒1 , 𝑇(𝑒2 ) = 8𝑒2 , 𝑇(𝑒3 ) = 8𝑒3 .
Then 𝐸(5, 𝑇) = span (𝑒1 ), 𝐸(8, 𝑇) = span (𝑒2 , 𝑒3 ), and 𝐸(0, 𝑇) = ℝ3 .
Thus:
Example 5.2.2
Let 𝑇 : ℝ2 → ℝ3 with (𝑥, 𝑦) ↦→ (41𝑥 + 7𝑦, −20𝑥 + 74𝑦
Use standard basis: 𝑒1 = (1, 0), 𝑒2 = (0, 1)
Thus:
41 7
ℳ(𝑇, (𝑒1 , 𝑒2 )) =
−20 74
Now try 𝑣1 = (1, 4), 𝑣 2 = (7, 5).
We claim that 𝑣1 , 𝑣 2 is a basis for ℝ2 .
Thus, we get:
69 0
ℳ(𝑇, (𝑣 1 , 𝑣 2 )) =
0 46
So 𝑇 is diagonalizable.
76
Hence:
Theorem 5.2.1
Let 𝑉 be a finite-dimensional vector space over 𝔽 and 𝑇 ∈ ℒ(𝑉).
Let 𝜆1 , . . . , 𝜆𝑚 be a complete list of the distinct eigenvalues of 𝑇.
The following are equivalent:
1. 𝑇 is diagonalizable.
2. 𝑉 has a basis consisting of eigenvectors of 𝑇.
1 ⇐⇒ 2
2 =⇒ 3
3 =⇒ 4
4 =⇒ 2
Let’s start:
Proof of 1 ⇐⇒ 2 : This is trivial.
If ℳ(𝑇, (𝑣1 , . . . , 𝑣 𝑛 )) is diagonal, then
77
Hence, 𝑢𝑖 Í= 0®𝑣 for 𝑖 = 1, . . . , 𝑚.
But 𝑢𝑖 = 𝑘∈𝐾 𝑖 𝑎 𝑘 𝑣 𝑘 .
Since these 𝑣 𝑘 ’s are LI (they are all in 𝐸(𝜆 𝑖 , 𝑇)).
Which implies that 𝑎 𝑘 = 0 for 𝑘 ∈ 𝐾 𝑖 and 𝑖 = 1, . . . , 𝑚.
Thus, 𝑣1 , . . . , 𝑣 𝑛 is linearly independent.
Now, our condition says that dim 𝑉 = dim 𝐸(𝜆1 , 𝑇) + . . . + dim 𝐸(𝜆𝑚 , 𝑇).
Let’s denote this as 𝑛, which is the dimension of 𝑉.
Hence, it’s a linearly independent list with an appropriate dimension, which implies that it is a basis for
𝑉.
Example 5.2.3
Let 𝑇 : ℂ2 → ℂ2 with 𝑇(𝑤, 𝑧) = (𝑧, 0).
0 1
Standard basis 𝑒1 , 𝑒2 , then ℳ(𝑇, (𝑒1 , 𝑒2 )) = .
0 0
𝐸(0, 𝑇) = 𝑣 ∈ ℂ2 | 𝑇(𝑣) = 0
Thus, you will never be able to find a basis of eigenvectors for 𝑇 that makes ℳ(𝑇) diagonal.
Since 2 = dim ℂ2 ≠ dim 𝐸(0, 𝑇) = 1, we conclude that 𝑇 is not diagonalizable.
78
Chapter 6
h, i : 𝑉 ×𝑉 → 𝔽
(𝑣, 𝑤) ↦→ h𝑣, 𝑤i
Such that:
2. h𝑣, 𝑣i = 0 ⇐⇒ 𝑣 = 0®𝑣
3. h𝑢 + 𝑤, 𝑣i = h𝑢, 𝑣i + h𝑤, 𝑣i for all 𝑢, 𝑤, 𝑣 ∈ 𝑉
4. h𝜆 ·𝑣 𝑣, 𝑤i = 𝜆 ·𝔽 h𝑣, 𝑤i for all 𝜆 ∈ 𝔽 and 𝑣, 𝑤 ∈ 𝑉
Example 6.1.1
h(𝑥1 , . . . , 𝑥 𝑛 ), (𝑦1 , . . . , 𝑦𝑛 )i ≔ 𝑥 1 𝑦1 + . . . + 𝑥 𝑛 𝑦𝑛
And:
h(𝑧1 , . . . , 𝑧 𝑛 ), (𝑤 1 , . . . , 𝑤 𝑛 )i ≔ 𝑧 1 𝑤 1 + . . . + 𝑧 𝑛 𝑤 𝑛
And:
79
h(𝑧1 , . . . , 𝑧 𝑛 ), (𝑧1 , . . . , 𝑧 𝑛 )i = 𝑧1 𝑧1 + . . . + 𝑧 𝑛 𝑧 𝑛 = |𝑧1 | 2 + . . . + |𝑧 𝑛 | 2 ≥ 0
80
Definition 6.1.2: Inner product space
𝑇𝑢 : 𝑉 → 𝔽
𝑣 ↦→ h𝑣, 𝑢i
Homogeneity:
𝑇𝑢 (𝜆 ·𝑣 𝑣) = h𝜆 ·𝑣 𝑣, 𝑢i = 𝜆 ·𝔽 h𝑣, 𝑢i = 𝜆 ·𝔽 𝑇𝑢 (𝑣)
D E D E
(ii) 0®𝑣 , 𝑣 = 0𝔽 : 0®𝑣 , 𝑣 = 𝑇𝑣 (0®𝑣 ) = 0𝔽
h𝑣, 𝑢 + 𝑤i = h𝑢 + 𝑤, 𝑣i
= h𝑢, 𝑣i + h𝑤, 𝑣i
= h𝑢, 𝑣i + h𝑤, 𝑣i
= h𝑣, 𝑢i + h𝑣, 𝑤i
= h𝑣, 𝑢i + h𝑣, 𝑤i
D E D E D E
(iv) 𝑣, 0®𝑣 = 0𝔽 : 𝑣, 0®𝑣 = 0®𝑣 , 𝑣 = 0𝔽 = 0𝔽
h𝑣, 𝜆 · 𝑤i = h𝜆 · 𝑤, 𝑣i
= 𝜆 · h𝑤, 𝑣i
= 𝜆 · h𝑤, 𝑣i
= 𝜆 · h𝑣, 𝑤i
Definition 6.1.3
81
Example 6.1.2
1. 𝑉 = ℝ𝑛 , 𝔽 = ℝ, h , i = dot product.
q
k(𝑥1 , . . . , 𝑥 𝑛 )k = 𝑥12 + . . . + 𝑥 𝑛2
∫1
2. 𝑉 = { 𝑓 : [−1, 1] → ℝ | 𝑓 is continuous} , 𝔽 = ℝ, h 𝑓 (𝑥), 𝑔(𝑥)i = −1
𝑓 (𝑥)𝑔(𝑥)𝑑𝑥
s
∫ 1
k𝑓k = 𝑓 (𝑥)2 𝑑𝑥
−1
Meaning k 𝑓 − 𝑔𝑛 k → 0 as 𝑛 → ∞.
For norms:
p
k𝑣 k = 0𝔽 ⇐⇒ h𝑣, 𝑣i = 0𝔽
⇐⇒ h𝑣, 𝑣i = 0𝔽
⇐⇒ 𝑣 = 0®𝑣
Example 6.1.3
Let 𝑉 = ℝ2 :
h(𝑥 1 , 𝑦1 ), (𝑥2 , 𝑦2 )i = 0 ⇐⇒ 𝑥 1 𝑥2 + 𝑦1 𝑦2 = 0
𝑥2 𝑦1
⇐⇒ =−
𝑦2 𝑥1
⇐⇒ (𝑥 1 , 𝑦1 ) and (𝑥2 , 𝑦2 ) are perpendicular
k𝑢 + 𝑣k 2 = k𝑢k 2 + k𝑣k 2
Proof:
k𝑢 + 𝑣 k 2 = h𝑢 + 𝑣, 𝑢 + 𝑣i
= h𝑢, 𝑢 + 𝑣i + h𝑣, 𝑢 + 𝑣i
= h𝑢, 𝑢i + h𝑢, 𝑣i + h𝑣, 𝑢i + h𝑣, 𝑣i
= h𝑢, 𝑢i + 0𝔽 + 0𝔽 + h𝑣, 𝑣i
= k𝑢 k 2 + k𝑣 k 2
Loose End:
k𝜆𝑣 k = |𝜆| · k𝑣k
As both are non-negative.
82
Proof:
p
k𝜆𝑣k = h𝜆𝑣, 𝜆𝑣i
q
= 𝜆 · 𝜆 · h𝑣, 𝑣i
q
= |𝜆| 2 · h𝑣, 𝑣i
p
= |𝜆| · h𝑣, 𝑣i
= |𝜆| · k𝑣 k
𝑣
Find 𝑐 ∈ 𝔽, 𝑤 ∈ 𝑉 such that we complete the triangle, i.e. 𝑢 = 𝑐 · 𝑣 + 𝑤.
𝑢
𝑤
𝑣 𝑐−𝑣
Want:
h𝑤, 𝑣i ⇐⇒ h𝑢 − 𝑐 · 𝑣, 𝑣i = 0
⇐⇒ h𝑢, 𝑣i − 𝑐 · h𝑣, 𝑣i = 0
⇐⇒ h𝑢, 𝑣i − 𝑐 · k𝑣 k 2 = 0
h𝑢, 𝑣i
=⇒ 𝑐 =
k𝑣 k 2
Note:-
h𝑢, 𝑣i
𝑤 = 𝑢 − 𝑐𝑣 = 𝑢 − ·𝑣
k𝑣 k 2
Where 𝑣 and 𝑤 are orthogonal.
83
Pick 𝛼 = h𝑢,𝑣i2 , so that 𝑤 + 𝛼𝑣 = 𝑢
k𝑣 k
This implies that:
2
h𝑢, 𝑣i
k𝑢 k 2 = k𝑤k 2 + · k𝑣k 2
k𝑣k 2
Where the right term of the sum is:
|h𝑢, 𝑣i| 2
· k𝑣 k 2
k𝑣k 4
Which is:
|h𝑢, 𝑣i| 2
k𝑣 k 2 ¾ =⇒ k𝑢k · k𝑣k ¾ |h𝑢, 𝑣i|
k𝑣k 2
Notice that equality holds if and only if:
k𝑤 k = 0 ⇐⇒ 𝑤 = 0®𝑣
h𝑢, 𝑣i
⇐⇒ 𝑢 = ·𝑣 =0
k𝑣k 2
h𝑢, 𝑣i
⇐⇒ 𝑢 = · 𝑣 ⇐⇒ 𝑢, 𝑣 are linearly dependent
k𝑣k 2
Example 6.1.4
𝑥® = (𝑥 1 , . . . , 𝑥 𝑛 )
𝑦® = (𝑦1 , . . . , 𝑦𝑛 )
2 2 2
𝑥®, 𝑦® ¶ 𝑥® · 𝑦®
(𝑥1 𝑦1 + . . . + 𝑥 𝑛 𝑦𝑛 )2 ¶ (𝑥12 + . . . + 𝑥 𝑛2 ) · (𝑦12 + . . . + 𝑦𝑛2 )
∫1
2. Let 𝑉 = { 𝑓 : [−1, 1] → ℝ | 𝑓 is continuous} , 𝔽 = ℝ, h 𝑓 , 𝑔i = −1
𝑓 (𝑥)𝑔(𝑥)𝑑𝑥.
By C.S., we know that:
|h 𝑓 , 𝑔i| 2 ¶ k 𝑓 k 2 · k 𝑔 k 2
84
Thus:
∫ 1 2 ∫ 1 ∫ 1
𝑓 (𝑥)𝑔(𝑥)𝑑𝑥 ¶ 𝑓 (𝑥) 𝑑𝑥 ·
2
𝑔(𝑥) 𝑑𝑥
2
−1 −1 −1
𝑢+𝑣
The triangle inequality states that:
k𝑢 + 𝑣k ¶ k𝑢 k + k𝑣 k
Proof:
k𝑢 + 𝑣 k 2 = h𝑢 + 𝑣, 𝑢 + 𝑣i
= h𝑢, 𝑢i + h𝑢, 𝑣i + h𝑣, 𝑢i + h𝑣, 𝑣i
= k𝑢 k 2 + h𝑢, 𝑣i + h𝑢, 𝑣i + k𝑣 k 2
= k𝑢 k 2 + 2 · Re h𝑢, 𝑣i + k𝑣k 2 as (𝑎 + 𝑏𝑖) + (𝑎 − 𝑏𝑖) = 2𝑎
≤ k𝑢k 2 + 2 · |h𝑢, 𝑣i| + k𝑣k 2 by ★
≤ k𝑢k 2 + 2 · k𝑢 k · k𝑣k + k𝑣k 2 by C.S.
= (k𝑢 k + k𝑣 k)2
Thus, we get:
k𝑢 + 𝑣 k 2 ¶ (k𝑢 k + k𝑣 k)2
k𝑢 + 𝑣k ¶ k𝑢 k + k𝑣 k
So why is ★ true?
√
h𝑢, 𝑣i = 𝑎 + 𝑏𝑖 =⇒ |h𝑢, 𝑣i| = 𝑎2 + 𝑏2
Re h𝑢, 𝑣i = 𝑎
√
But why is 𝑎 ¶ 𝑎2 + 𝑏2?
True since:
𝑎 ≤ |𝑎|
√
= 𝑎2
√
≤ 𝑎2 + 𝑏2
85
Example 6.1.5
Let 𝑉 = ℝ𝑛 , 𝑥® = (𝑥1 , . . . , 𝑥 𝑛 ), 𝑦® = (𝑦1 , . . . , 𝑦𝑛 ).
We have:
2 2 2
𝑥® + 𝑦® = 𝑥® + 2 · 𝑥®, 𝑦® + 𝑦®
Thus:
v v 2
𝑛
t 𝑛 t 𝑛
Õ © Õ 2 Õ ª
(𝑥 𝑖 + 𝑦 𝑖 ) ¶
2
𝑥𝑖 + 𝑦 𝑖2 ®
𝑖=1 « 𝑖=1 𝑖=1 ¬v
v
𝑛
t 𝑛 t 𝑛 𝑛
Õ Õ Õ Õ
= 𝑥 2𝑖 +2· 𝑥 2𝑖 · 𝑦 𝑖2 + 𝑦 𝑖2
𝑖=1 𝑖=1 𝑖=1 𝑖=1
1 1 1 1 1
√ , √ , √ , − √ , √ , 0 ∈ ℝ3
3 3 3 2 2
k𝑎 1 𝑒1 + . . . + 𝑎 𝑛 𝑒 𝑛 k 2 = h𝑎1 𝑒1 + . . . + 𝑎 𝑛 𝑒 𝑛 , 𝑎 1 𝑒1 + . . . + 𝑎 𝑛 𝑒 𝑛 i
= h𝑎1 𝑒1 , 𝑎 1 𝑒1 i + . . . + h𝑎 𝑛 𝑒 𝑛 , 𝑎 𝑛 𝑒 𝑛 i
= 𝑎1 · 𝑎1 h𝑒1 , 𝑒1 i + . . . + 𝑎 𝑛 · 𝑎 𝑛 h𝑒 𝑛 , 𝑒 𝑛 i
= |𝑎1 | 2 + . . . + |𝑎 𝑛 | 2
86
Definition 6.2.2
Claim 6.2.1
Suppose {𝑉1 , . . . , 𝑉𝑛 } is orthonormal, then {𝑉1 , . . . , 𝑉𝑛 } is linearly independent.
𝑐1 𝑣1 + . . . 𝑐 𝑛 𝑣 𝑛 = 0
Then it follows:
h𝑐1 𝑣1 + . . . 𝑐 𝑛 𝑣 𝑛 , 𝑐 1 𝑣1 + . . . 𝑐 𝑛 𝑣 𝑛 i = k𝑐 1 𝑣1 + . . . 𝑐 𝑛 𝑣 𝑛 k 2 = 0
Which means:
|𝑐 1 | 2 + . . . + |𝑐 𝑛 | 2 = 0 =⇒ |𝑐1 | 2 = . . . = |𝑐 𝑛 | 2 = 0
Thus,
𝑐 1 = . . . = 𝑐 𝑛 = 0𝔽
/* We want h𝑒1 , 𝑒2 i = 0𝔽 */
𝑣®1
1 𝑒®1 = 𝑣®1
;
𝑣®2 − 𝑣®2 , 𝑒®1 · 𝑒®1
2 𝑒®2 = 𝑣®2 − 𝑣®2 , 𝑒®1 · 𝑒®1
;
𝑣®𝑗 − 𝑣®𝑗 , 𝑒®1 · 𝑒®1 −...− 𝑣®𝑗 , 𝑒 𝑗−1
® · 𝑒 𝑗−1
®
3 𝑒®𝑗 = 𝑣®𝑗 − 𝑣®𝑗 , 𝑒®1 · 𝑒®1 −...− 𝑣®𝑗 , 𝑒 𝑗−1
® · 𝑒 𝑗−1
®
;
Example 6.2.1
∫1
Let 𝑉 = 𝒫2 (ℝ), h𝑝, 𝑞i = −1 𝑝(𝑥)𝑞(𝑥)𝑑𝑥.
Start with 𝑣®1 = 1, 𝑣®2 = 𝑥, 𝑣®3 = 𝑥 2 .
1.
𝑣®1 1 √
∫
𝑒®1 = = 12 𝑑𝑥 = 2
𝑣®1 −1
1
=√
2
87
2.
∫ 1
1 1
𝑣2 − h𝑣2 , 𝑒1 i · 𝑒1 = 𝑥 − 𝑥 · √ 𝑑𝑥 · √
−1 2 2
notice that the integral is 0 as 𝑥 is odd
=𝑥
Remember that:
∫ 1
2 2
k𝑥 k = h𝑥, 𝑥i = 𝑥 2 𝑑𝑥 =
−1 3
r
2 𝑥
=⇒ k𝑥k = =⇒ 𝑒®2 = q
3 2
3
3.
1 1
𝑥 𝑥
∫ ∫
1 1
𝑣3 − h𝑣3 , 𝑒1 i · 𝑒1 − h𝑣 3 , 𝑒2 i · 𝑒2 = 𝑥 2 − 𝑥 2 · √ 𝑑𝑥 · √ − 𝑥 2 · q 𝑑𝑥 · q
−1 2 2 −1 2 2
3 3
Hence,
s
𝑥2 − 1
∫ 1
r
1 1 8
𝑥2 − = (𝑥 2 − )2 𝑑𝑥 = =⇒ 𝑒®3 = q 3
3 −1 3 45 8
45
In particular:
𝑢 = 𝑢0
⇐⇒ 𝑢 − 𝑢 0 = 0
⇐⇒ ∀𝑣 ∈ 𝑉 , h𝑣, 𝑢 − 𝑢 0i = 0
88
Definition 6.2.3
𝜙
A linear functional on 𝑉 is a linear map 𝑉 −
→ 𝔽.
i.e., 𝜙 ∈ ℒ(𝑉 , 𝔽)
Which is 𝑢!
Thus,
Uniqueness:
𝜙(𝑣) = h𝑣, 𝑢i = h𝑣, 𝑢 0i for all 𝑣 ∈ 𝑉
Show 𝑢 = 𝑢 0 ⇐⇒ show h𝑣, 𝑢 − 𝑢 0i = 0 for all 𝑣 ∈ 𝑉.
Thus, 𝑢 = 𝑢 0.
Note:-
Because of uniqueness the 𝑢 in the proof cannot / doesn’t depend on the choice of basis.
Example 6.2.2
∫1
Let 𝒫2 (ℝ) with h𝑝, 𝑞i = −1 𝑝(𝑥)𝑞(𝑥)𝑑𝑥.
This has an orthonormal basis:
r r r
1 3 45 2 1
𝑒1 = , 𝑒2 = 𝑥, 𝑒3 = (𝑥 − )
2 2 8 3
89
Let 𝜙 ∈ ℒ(𝒫2 (ℝ), ℝ) be defined by:
∫ 1
𝜙(𝑝) = 𝑝(𝑥) cos(𝜋𝑥)𝑑𝑥 ∈ ℒ(𝒫2 (ℝ), ℝ)
−1
Note:-
We have:
𝜙(𝑝) = h𝑝, 𝑢i
𝜙𝑤 (𝑣) = h𝑣, 𝑢𝑤 i𝑉
Now, notice:
𝑇 ∗ : 𝑊 → 𝑉 , 𝑤 ↦→ 𝑢𝑤 ≔ 𝑇 ∗ (𝑤)
Definition 6.2.4
The adjoint of a linear map 𝑇 : 𝑉 → 𝑊 between inner product spaces is the linear map 𝑇 ∗ : 𝑊 → 𝑉
characterized by:
Example 6.2.3
Let ℝ3 , ℝ2 with the standard inner product i.e., dot product.
90
h𝑇(𝑥1 , 𝑥 2 , 𝑥 3 ), (𝑦1 , 𝑦2 )i ℝ2 = h(𝑥1 + 𝑥2 , 2𝑥2 + 𝑥 3 ), (𝑦1 , 𝑦2 )i ℝ2
= h(𝑥1 , 𝑥 2 , 𝑥 3 ), 𝑇 ∗ (𝑦1 , 𝑦2 )i ℝ3
= (𝑥1 + 𝑥2 )𝑦1 + (2𝑥2 + 𝑥 3 )𝑦2
= 𝑥1 𝑦1 + 𝑥 2 𝑦1 + 2𝑥 2 𝑦2 + 𝑥3 𝑦2 = h(𝑥1 , 𝑥 2 , 𝑥 3 ), (𝑦1 , 𝑦1 + 2𝑦2 , 𝑦2 )i ℝ3
Note:-
Is𝑇∗ is linear?
Adjoins are linear:
If 𝑇 : 𝑉 → 𝑊 is linear, then 𝑇 ∗ : 𝑊 → 𝑉 is linear.
91
Chapter 7
92
Chapter 8
93
Chapter 9
94
Chapter 10
det : 𝔽𝑛,𝑛 =⇒ 𝔽
𝑎 𝑏
(b) If 𝑛 = 2, then det = 𝑎𝑑 − 𝑏𝑐.
𝑐 𝑑
Example 10.1.1
Let
1 2 3
𝐴 = 4 5 6®
© ª
«7 8 9¬
.
Then:
2 3
𝐴2,1 =
8 9
𝑛
Õ
det 𝐴 = (−1)𝑖+1 𝑎 𝑖,1 · det 𝐴 𝑖,1
𝑖=1
95
Example 10.1.2
Let:
1 0 3 𝑎 𝑎 12 𝑎 13
ª © 11
𝐴 = 2 1 2® = 𝑎21 𝑎 22 𝑎 23 ®
© ª
«0 5 1¬ « 𝑎31 𝑎 32 𝑎 33 ¬
Thus,
det 𝐴 = 𝑎1,1 · det 𝐴1,1 − 𝑎2,1 · det 𝐴2,1 + 𝑎3,1 · det 𝐴3,1
1 2 0 3 0 3
= 1 · det − 2 · det + 0 · det
5 1 5 1 1 2
= 1 · (−9) − 2 · (−15) + 0 · (−3)
= 21
1. 𝛿(𝐼𝑛 ) = 1
2. 𝛿 is row-linear.
1 2 3 1 2 3 1 2 3
𝛿 4𝜆 + 2𝜇 5𝜆 + 5𝜇 6𝜆 + 8𝜇®® = 𝜆 · 𝛿 4 5 6® + 𝜇 · 𝛿 2 5 8®
©© ªª © ª © ª
«« 7 8 9 ¬¬ «7 8 9¬ «7 8 9¬
Assume the previous theorem is true for now.
What is the value of 𝛿 on elementary matrices?
𝛿(𝐴)
if 𝐸 is type (i)
if 𝐸 is type (ii)
𝛿(𝐸 · 𝐴) = −𝛿(𝐴)
𝑐 · 𝛿(𝐴) if 𝐸 is type (iii)
𝑆 is determined on elementary matrices.
+1
if 𝐸 is type (i)
if 𝐸 is type (ii)
𝛿(𝐸) = −1
𝑐 if 𝐸 is type (iii)
Proof: Take 𝐴 = 𝐼𝑛 in theorem 2.
96
Proof to det 2: For 𝐸 of type (𝑖𝑖𝑖) this is jut row-linearity of 𝛿.
Let 𝐴 𝑖 be the 𝑖th row of 𝐴.
1 −−
1 𝐴 −− −− 1 𝐴 −− 1
ª −− 𝐴1 −− ª® .. ..
.. ..
©© © ª © ª © ª
. ® −− 𝐴2 . . .
® © ® ® ®
−−®®®
ª ® ® ®
𝛿 𝑐 ®· .. ®® = 𝛿 −− 𝑐𝐴 𝑖 −−® = 𝑐·𝛿 −− 𝐴 𝑖 −−® = 𝑐·𝛿(𝐴) 𝑐
® ® ® ®
.
®
.. .. .. ..
® ®® ® ® ®
. . . .
®®
® −− 𝐴𝑛
® ® ® ®
−−¬® ® ® ®
1¬ −− 𝐴 −−¬ −− 𝐴 −−¬ 1¬
«
«« ¬ « 𝑛 « 𝑛 «
Since we did not require 𝑐 ≠ 0, then this is true for all 𝑐 ∈ 𝔽.
Thus, 𝛿(𝐸 · 𝐴) = 𝑐 · 𝛿(𝐴) for all 𝑐 ∈ 𝔽.
If a row contains only zeros, then 𝛿(𝐴) = 0.
For types (𝑖) and (𝑖𝑖), we do the special case when 𝐸 acts on consecutive rows.
(Type: i):
−− 𝐴1 −− −− 𝐴1 −−
1
ª −− 𝐴2 −−ª® 𝐴2
© ©−− −−ª®
.. .. ..
©
. 𝑎 𝑖,𝑗
®
® .
®
® .
®
®
𝐸 · 𝐴 = .. ® · −− 𝐴 𝑖 −−®® = −− 𝑎 𝑖,𝑗 𝐴 𝑖 + 𝐴 𝑗
® ® ®
. −−®®
® −− 𝐴 𝑖+1 𝐴 𝑖+1
®
.. −−®® −− −−®®
. ®
.. ..
. .
® ® ®
1¬ ® ®
«
« −− 𝐴 𝑛 −−¬ «−− 𝐴𝑛 −−¬
Special case, 𝑗 = 𝑖 + 1
−− 𝐴1 −− −− 𝐴1 −− −− 𝐴1 −−
©−− 𝐴2 −−ª® ©−− 𝐴2 −−ª® ©−− 𝐴2 −−ª®
.. .. ..
® ® ®
. ® . ® . ®
𝛿(𝐸 · 𝐴) = 𝛿 −− 𝐴 𝑖+1 −−®® = 𝑎 · 𝛿 −− 𝐴 𝑖 −−®® + 𝛿 −− 𝐴 𝑖
® ® ®
−−®®
−− 𝑎𝐴 𝑖 + 𝐴 𝑖+1 −−®® −− 𝐴 𝑖 −−®® −− 𝐴 𝑖+1 −−®®
.. .. ..
. . .
® ® ®
® ® ®
« −− 𝐴 𝑛 −−¬ « −− 𝐴 𝑛 −−¬ « −− 𝐴 𝑛 −−¬
But, the first matrix’s determinant is 0 since it has two identical rows.
Thus, 𝛿(𝐸 · 𝐴) = 𝑎 · 0 + 𝛿(𝐴) = 𝛿(𝐴).
(Type: ii): Swap rows.
Again, this assumes theorem 1.
Let’s assume that we swap row 𝑖 with row 𝑖 + 1
−− 𝐴1 −−
© .. ª
. ®
®
−− 𝐴 𝑖 −−®
𝛿(𝐴) = 𝛿
−− 𝐴 𝑖+1 −−®
®
®
..
.
®
®
«−− 𝐴𝑛 −−¬
by part one:
97
−− 𝐴1 −−
© .. ª
. ®
®
−− 𝐴 𝑖 − 𝐴 𝑖+1 −−®
= 𝛿
𝐴
®
−− 𝑖+1 −−®
..
®
.
®
®
« −− 𝐴 𝑛 −−¬
by part one again:
−− 𝐴1 −−
© .. ª
. ®
®
−− 𝐴 𝑖 − 𝐴 𝑖+1 −−®®
= 𝛿
−− 𝐴 𝑖+1 + (𝐴 𝑖 − 𝐴 𝑖+1 ) −−®
®
..
.
®
®
« −− 𝐴 𝑛 −− ¬
−− 𝐴1 −−
© .. ª
. ®
®
−− 𝐴 𝑖 − 𝐴 𝑖+1 −−®
= 𝛿
𝐴𝑖
®
−− −−®®
..
.
®
®
«−− 𝐴𝑛 −−¬
by row linearity:
−− 𝐴1 −− −− 𝐴1 −−
© .. ª © .. ª
. ®
®
. ®
®
−− 𝐴 𝑖 −−®® −− 𝐴 𝑖+1 −−®
= 𝛿 −𝛿
−− 𝐴 𝑖 −− 𝐴 𝑖 −−®
®
−−®® ®
.. ..
. .
® ®
® ®
«−− 𝐴𝑛 −−¬ «−− 𝐴𝑛 −−¬
but for the first matrix is zero since it has two identical rows:
−− 𝐴1 −−
© .. ª
. ®
®
−− 𝐴 𝑖+1 −−®
= −𝛿
−− 𝐴 𝑖 −−®
®
®
..
.
®
®
« −− 𝐴 𝑛 −−¬
= −𝛿(𝐸 · 𝐴)
𝛿(𝐸 · 𝐴) = −𝛿(𝐴)
𝑗 − 1 + (𝑗 − 1) − 𝑖 = 2(𝑗 − 𝑖) − 1
Which is odd!
This means that 𝛿(𝐸 · 𝐴) = (−1)2(𝑗−𝑖)−1 · 𝛿(𝐴) = −𝛿(𝐴).
Note:-
We can also do this for part 1.
Do this an exercise.
As such, we have proven theorem 2.
Corollary 10.1.2
𝛿(𝑆 · 𝐵) = 𝛿(𝐴) · 𝛿(𝐵) for any 𝐴, 𝐵 ∈ 𝔽𝑛,𝑛 .
We know that 𝛿(𝐸) · 𝛿(𝐴) = 𝛿(𝐸 · 𝐴) if 𝐸 is an elementary matrix.
Let 𝐴0 = 𝐸 𝑘 · · · 𝐸1 · 𝐴 be a reduced row echelon form of 𝐴.
Then either:
(i) 𝐴0 = 𝐼𝑛 or
(ii) the last row of 𝐴0 is all zeros. (could be more than the last row)
Then:
(i) If 𝐴0 = 𝐼𝑛 ,
𝐴0 = 𝐼𝑛 =⇒ 𝐴 = 𝐸1−1 · · · 𝐸−1
𝑘 · 𝐼𝑛
=⇒ 𝛿(𝐴) = 𝛿(𝐸1−1 ) · · · 𝛿(𝐸−1
𝑘 )
Proof of det 1: Proof of uniqueness: Suppose there are functions 𝛿 : 𝔽𝑛,𝑛 → 𝔽 and 𝛿0 : 𝔽𝑛,𝑛 → 𝔽.
Each satisfying the three desired properties.
Let 𝐴 ∈ 𝔽𝑛,𝑛 such that 𝐴0 = 𝐸 𝑘 · · · 𝐸1 · 𝐴 is a reduced row echelon form of 𝐴.
Either we get 𝐴0 = 𝐼𝑛 or its last row is all zeros.
In either case, 𝛿(𝐴0) = 𝛿0(𝐴0) = 1.
Or 𝛿(𝐴0) = 𝛿0(𝐴0) = 0.
That means that 𝛿(𝐸 𝑘 · · · 𝐸1 · 𝐴) = 𝛿0(𝐸 𝑘 · · · 𝐸1 · 𝐴) in either case.
Thus,
𝛿(𝐸 𝑘 ) · · · 𝛿(𝐸1 ) · 𝛿(𝐴) = 𝛿0(𝐸 𝑘 ) · · · 𝛿0(𝐸1 ) · 𝛿0(𝐴)
99
But by theorem 2, we get 𝛿(𝐸 𝑖 ) = 𝛿0(𝐸 𝑖 ).
Which means that 𝛿(𝐴) = 𝛿0(𝐴) for all 𝐴 ∈ 𝔽𝑛,𝑛 .
Proof of existence: We’ll show det : 𝔽𝑛,𝑛 → 𝔽 satisfies the three properties to be 𝛿.
Let’s proceed by induction on 𝑛 ∈ ℕ
Base Case: Let 𝑛 be 1.
Then det : 𝔽1,1 → 𝔽 is defined by det(𝑎) = 𝑎.
Thus, det(𝐼1 ) = 1.
Now, for row linear:
1
..
© ª
.
®
®
𝛿(𝐼𝑛 ) = 𝛿 1
®
®
.. ®
. ®
®
« 1¬
= 1 · 𝛿(𝐼𝑛−1 )
=1·1 by inductive hypothesis
=1
Claim 10.1.1
𝑑 𝑖,1 · det(𝐷𝑖,1 ) = 𝜆 · 𝑎 𝑖,1 · det(𝐴 𝑖,1 ) + 𝜇 · 𝑏 𝑖,1 · det(𝐵 𝑖,1 ) is true for all 𝑖 ∈ {1, . . . , 𝑛}.
If claim is true then we can:
Case (i) 𝑖 = 𝑘, then The minors 𝐴 𝑘,1 , 𝐵 𝑘,1 and 𝐷 𝑘,1 are equal.
I.e., the 𝑘th row of 𝐴, 𝐵, 𝐷 is deleted.
Which means:
Claim is true ⇐⇒ 𝑑 𝑖,1 = 𝜆 · 𝑎 𝑖,1 + 𝜇 · 𝑏 𝑖,1 .
This is true by our construction of 𝐷.
Case (ii) 𝑖 ≠ 𝑘, then
𝐴0 , 𝐵0 , 𝐷 0 rows with 𝑛 − 1 entries after deleting the 𝑘th row.
Then 𝐷 0𝑘 = 𝜆 · 𝐴0𝑘 + 𝜇 · 𝐵0𝑘 .
100
All other rows of 𝐴0 , 𝐵0 , 𝐷 0 are equal.
Thus, by inductive hypothesis:
(iii) Moved a bit ahead in these notes, you can see the final part of the proof.
Note:-
On Mondays’ class we showed that:
det(𝐴) = (−1) 𝑘+1 · 𝑎 𝑘,1 · det 𝐴 𝑘,1 + (−1) 𝑘+2 · 𝑎 𝑘+1,1 · det 𝐴 𝑘+1,1
Since 𝐴 𝑘 = 𝐴 𝑘+1 have 𝑎 𝑘,1 = 𝑎 𝑘+1,1 and 𝐴 𝑘,1 = 𝐴 𝑘+1,1
This implies that:
det 𝐴 = (−1) 𝑘+1 [𝑎 𝑘,1 · det 𝐴 𝑘,1 + (−1) · 𝑎 𝑘,1 · det 𝐴 𝑘,1 ] = 0
Thus, det 𝐴 = 0.
Therefore, by the principle of mathematical induction, det 𝐴 = 0 for all 𝐴 with two identical rows.
Corollary 10.1.3
These are given free by the theorem of det 1:
𝑆 1 2 3 4 5
𝜎(𝑆) 3 5 1 4 2
Then:
∼
n o
𝑆𝑛 ≔ permutations 𝜎 : {1, . . . , 𝑛} →
− {1, . . . , 𝑛}
|𝑆𝑛 | = 𝑛!
Can compare permutations:
∼
𝜏 : {1, . . . , 𝑛} →
− {1, . . . , 𝑛}
∼
𝜎 : {1, . . . , 𝑛} →
− {1, . . . , 𝑛}
Example:
𝑆 1 2 3 4
𝜎(𝑆) 4 1 3 2
Thus:
𝜎 = (142)(3)
= (142)
102
𝜎 = (134)(25)
𝜏 = (1452)
𝜏 ◦ 𝜎 = [(1452)] ◦ [(134)(25)]
| {z } | {z }
then this first do this
= (135)(2)(4)
= (135)
(𝜏 ◦ 𝜎)(1) = 𝜏(𝜎(1)) = 𝜏(3) = 5
Question 3
Problem 5. The trace of a square matrix 𝐴 is the sum of its diagonal entries:
Show that
(a) tr(𝐴 + 𝐵) = tr(𝐴) + tr(𝐵);
(b) tr(𝐴𝐵) = tr(𝐵𝐴);
(c) if 𝐵 is invertible, then tr(𝐴) = tr 𝐵𝐴𝐵−1 .
𝑎 +𝑏 𝑎12 + 𝑏 12 ··· 𝑎 1𝑛 + 𝑏 1𝑛
© 11 11
𝑎21 + 𝑏21 𝑎22 + 𝑏 22 ··· 𝑎 2𝑛 + 𝑏 2𝑛 ®
ª
𝐶 = .. .. .. .. ®
. . . .
®
®
« 𝑛1 𝑏 𝑛1
𝑎 + 𝑎 𝑛2 + 𝑏 𝑛2 ··· 𝑎 𝑛𝑛 + 𝑏 𝑛𝑛 ¬
Let’s take the trace of 𝐶:
𝑛
Õ
𝑡𝑟(𝐶) = 𝑐 𝑖𝑖
𝑖=1
= (𝑎 1,1 + 𝑏 11 ) + . . . + (𝑎 𝑛,𝑛 + 𝑏 𝑛,𝑛 )
= 𝑎11 + . . . + 𝑎 𝑛,𝑛 + 𝑏 11 + . . . + 𝑏 𝑛,𝑛
𝑛
Õ 𝑛
Õ
𝑡𝑟(𝐴) + 𝑡𝑟(𝑏) = 𝑎 𝑖𝑖 + 𝑏 𝑖𝑖
𝑖=1 𝑖=1
= 𝑎 11 + . . . + 𝑎 𝑛,𝑛 + 𝑏 11 + . . . + 𝑏 𝑛,𝑛
As both sides are equal, we have shown that 𝑡𝑟(𝐴 + 𝐵) = 𝑡𝑟(𝐴) + 𝑡𝑟(𝐵).
Proof of 𝑏 : Let 𝐴, 𝐵 be two matrices size 𝑛 × 𝑛 with entries in 𝔽.
If they are not of the same size, then we cannot multiply them.
So, let’s assume they are both matrices of size 𝑛 × 𝑛.
103
Note:-
Don’t worry, I have a program that generates these matrices for me.
Í𝑛 Í𝑛 Í𝑛
𝑎 𝑏 𝑎 𝑏 ··· Í𝑛𝑘=1 𝑎1𝑘 𝑏 𝑘𝑛 ª
©Í𝑛𝑘=1 1𝑘 𝑘1 Í𝑛𝑘=1 1𝑘 𝑘2
𝑘=1 𝑎2𝑘 𝑏 𝑘1 𝑘=1 𝑎 2𝑘 𝑏 𝑘2 ··· 𝑘=1 𝑎 2𝑘 𝑏 𝑘𝑛 ®
𝐶 = .. .. .. .. ®
. .
Í𝑛 . .
®
Í ®
𝑛 Í𝑛
𝑎 𝑏
« 𝑘=1 𝑛𝑘 𝑘1 𝑘=1 𝑎 𝑛𝑘 𝑏 𝑘2 ··· 𝑘=1 𝑎 𝑛 𝑘 𝑏 𝑘𝑛 ¬
Õ 𝑛
𝐶 𝑖,𝑗 = 𝑎 𝑖𝑘 𝑏 𝑘 𝑗
𝑘=1
Í𝑛 Í𝑛 Í𝑛
𝑏 𝑎 𝑏 𝑎 ··· Í𝑛𝑘=1 𝑏1𝑘 𝑎 𝑘𝑛 ª
©Í𝑛𝑘=1 1𝑘 𝑘1 Í𝑛𝑘=1 1𝑘 𝑘2
𝑘=1 𝑏2𝑘 𝑎 𝑘1 𝑘=1 𝑏 2𝑘 𝑎 𝑘2 ··· 𝑘=1 𝑏 2𝑘 𝑎 𝑘𝑛 ®
𝐷 = .. .. .. .. ®
Í . . . .
®
®
𝑛 𝑛 Í𝑛
𝑏 𝑎 𝑘=1 𝑛 𝑘 𝑎 𝑘2
𝑏 𝑏 𝑎
Í
« 𝑘=1 𝑛𝑘 𝑘1 ··· 𝑘=1 𝑛𝑘 𝑘𝑛 ¬
Õ 𝑛
𝐷𝑖,𝑗 = 𝑏 𝑖𝑘 𝑎 𝑘 𝑗
𝑘=1
𝑛
Õ
𝑡𝑟(𝐶) = 𝑐 𝑖𝑖
𝑖=1
Õ𝑛 Õ 𝑛
= 𝑎 𝑖𝑘 𝑏 𝑘𝑖
𝑖=1 𝑘=1
Õ𝑛
= (𝑎 𝑖1 𝑏1𝑖 + . . . + 𝑎 𝑖𝑛 𝑏 𝑛𝑖 )
𝑖=1
𝑛
(𝑏1𝑖 𝑎 𝑖1 + 𝑏2𝑖 𝑎 𝑖2 + . . . + 𝑏 𝑛𝑖 𝑎 𝑖𝑛 ) by commutativity of · in 𝔽
Õ
=
𝑖=1
= (𝑏11 𝑎11 + . . . + 𝑏 𝑛1 𝑎1𝑛 ) + . . . + (𝑏1𝑛 𝑎 𝑛1 + . . . + 𝑏 𝑛𝑛 𝑎 𝑛𝑛 )
= (𝑏11 𝑎11 + 𝑏 12 𝑎21 + . . . + 𝑏1𝑛 𝑎 𝑛1 ) + . . . + (𝑏 𝑛1 𝑎1𝑛 + . . . + 𝑏 𝑛𝑛 𝑎 𝑛𝑛 )
𝑛
Õ
= (𝑏 𝑖1 𝑎1𝑖 + 𝑏 𝑖2 𝑎2𝑖 + . . . + 𝑏 𝑖𝑛 𝑎 𝑛𝑖 )
𝑖=1
Õ𝑛 Õ 𝑛
= 𝑏 𝑖𝑘 𝑎 𝑘𝑖
𝑖=1 𝑘=1
Õ𝑛
= 𝑑 𝑖𝑖
𝑖=1
= 𝑡𝑟(𝐷)
104
Proof of 𝑐: This follows directly from part 𝑏.
Let 𝐴, 𝐵 be two matrices size 𝑛 × 𝑛 with entries in 𝔽.
Remember that we prove in part 𝑏, that 𝑡𝑟(𝐴𝐵) = 𝑡𝑟(𝐵𝐴). Thus:
105