0% found this document useful (0 votes)
100 views15 pages

Linear Algebra

The document discusses vector spaces and linear algebra concepts. It defines a vector space as a set that is closed under finite addition and scalar multiplication. A vector space has properties like closure under addition and scalar multiplication, commutativity, distributivity, existence of additive inverses and a zero vector. Vectors can be represented by matrices. Linear dependence and independence of vectors is introduced. The maximum number of linearly independent vectors is the dimension of the vector space. Dual spaces and Dirac's bra-ket notation for representing vectors and dual vectors are also covered.
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)
100 views15 pages

Linear Algebra

The document discusses vector spaces and linear algebra concepts. It defines a vector space as a set that is closed under finite addition and scalar multiplication. A vector space has properties like closure under addition and scalar multiplication, commutativity, distributivity, existence of additive inverses and a zero vector. Vectors can be represented by matrices. Linear dependence and independence of vectors is introduced. The maximum number of linearly independent vectors is the dimension of the vector space. Dual spaces and Dirac's bra-ket notation for representing vectors and dual vectors are also covered.
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/ 15

Chapter 2

Linear Algebra

2.1 Vector Spaces

COP Y
A vector space over a field is a set that us closed under the finite addition and scalar multiplication, where
the scalar belongs to the field. In this context, we shall deal with vector spaces over the field of complex
numbers, denoted by C. A vector space has the following properties:

DR F T
1. Closure under vector addition:

A
(u + v) ∈ V ∀ (u,v) ∈ V

2. Closure under scalar multiplication:


cu ∈ V ∀ u ∈ V , c ∈ F

3. Vector addition and scalar multiplication are commutative


u + v = v + u ∀ u,v ∈ V
(cd)u = c(du) = d(cu) ∀ u ∈ V , (c,d) ∈ F

4. Vector addition distributive


c(u + v) = cu + cv ∀ (u,v) ∈ V , c ∈ F

5. Existance of an additive inverse


∀ u ∈ V, ∃ v ∈ V such that u + v = 0
The vector v is denoted as -u and called the inverse of u, under addition.

6. Existance of a scalar identity under multiplication


∀ u ∈ V, ∃ c ∈ F such that cu = u
The identity is denoted by 1.

7. Existance of a zero vector


∀ u ∈ V, ∃ v ∈ V such that u + v = u
v is then called the zero vector and denoted by 0.

In this context, we shall deal with vector spaces over the field of complex numbers, denoted by C.
A vector can be represented by a matrix. The entries of the matrix are called the components of the vector.

35
2.2. LINEAR DEPENDENCE AND INDEPENDENCE CHAPTER 2. LINEAR ALGEBRA

For example:
  
  
v1 u1 v1 + u1
 v2   u2   v2 + u2 
     
 v3   u3   v3 + u3 
     
v= 
.
 , u= 
 . 
 , v + u= 
 . 

.  .   . 
     
.  .   . 
vn un vn + un

Here n is called the dimension of the vector space. The vector space can also be infinite dimensinal.

2.2 Linear dependence and independence

A set of vectors −
→, −
v → − → −

1 v2 , v3 . . . . vn are said to be linearly dependent iff ∃ some constants ci ’s such
that:
n

ci →

vi = 0.

P Y
i=1

O
Exception: The trivial case where all ci ’s are = 0.

A F T C
Each vector depends on the rest of the vectors by a linear relationship. Hence the name linearly dependent.
The vectors are called linearly independent iff they do not satisfy the above conditions. For linearly
independent vectors, no vector is related to the rest by a linear relationship. The maximum number of

DR
linearly independent vectors in a vector space is called the dimension of the vector space, denoted by n.
Any vector →−
x in this vector space can be uniquely described as a linear combination of these n vectors of the
vector space. The proof for this is quite simple. A n-dimensional vector space will have a set of n linearly
independent vectors. If we add the vector → −
x to this set, then we will still have a n-dimensional space. So,

→ −
→ −
→ −
→ →

this means the set v1 , v2 , v3 . . . . vn , x is linearly dependent. So:
n

ci →

vi + cx →

x =0
i=1
n

so: ci →

vi = cx� →

x
i=1
�n
⇒ ci� →

vi = →

x
i=1

Hence, we have shown that the general vector → −x can be expressed uniquely as a linear combination of the
n-linearly independent vectors of the vector space. The set of n linearly independent vectors is called the
basis of the vector space. The set of all the vectors represented by this basis is called the spanning set of
the basis.

2.3 Dual Spaces

A linear function can be defined over the elements of the vector space. Let f be a function which is linear.
f :→−
x → f (→ −
x ), such that f (→

x ) ∈ C. The function maps an element in the vector space to an element in
the field over which the vector space is defined. By our convention we represent a vector →
−x by a coloumn

GO TO FIRST PAGE↑ 36
CHAPTER 2. LINEAR ALGEBRA 2.4. DIRAC’S BRA KET NOTATION

vector. So, we can see by inspection that f should be represented using a row vector. Then on multiplying
the row and the column vector, we get a single scalar. The function satisfies the requirement:

f (a→

x + b→ −
y ) = af (→

x ) + bf (→

x)
where: {a, b} ∈ F and {→ −
x,→−y} ∈ V

So, we can propose a representation for the function:


 
x1
 x2 
 
 x3 
  � �
if → −
x =  
 .  then we can say that f = f1 f2 f3 . . .fn
 . 
 
 . 
xn
Hence, on matrix multiplication of f with→

x , we get f1 x1 + f2 x2 + f3 x3 ....fn xn
which is a dimensionless number ∈ F.

The space of such functions is called the dual space to the corresponding vector space. If the vector space
is V, then the corresponding dual space is represented by V∗ .

2.4

OP Y
Dirac’s Bra Ket notation

C
T
The Dirac bra-ket notation is widely used in Quantum Mechanics. Here any vector → −

F
x (represented as a

A
column matrix) is represented as |x� and this vector is called a ket vector. The corresponding dual vector

DR
is called the bra vector, denoted by �x|. Hence the name bra ket notation. We would be following this
method from now on. The reasons will become evident as we proceed. So, we have the conventions:

�x| ↔ |x� (2.1)


�x|c ↔ c∗ |x� (where c is some complex number) (2.2)

2.5 Inner and outer products

Consider a vector |x� and its dual vector �x|. We saw that the former was a column matrix and the latter
the row matrix. �x| is called the bra dual of |x�. So, the product of a ket vector and its corresponding bra
vector is a dimensionless scalar quantity (as we saw this is multiplying a row matrix with a column matrix).
The product is represented as: �x|x�. This product is called inner product. The inner product of a bra
and a ket vector is in short defined as: �
�x|y� = yi∗ xi (2.3)
i

The inner product also satisfies some linearity properties:

�x|c1 y1 + c2 y2 � = c1 �x1 |y� + c2 �x2 |y� (2.4)


�c1 x1 + c2 x2 |y� = c∗1 �x1 |y� + c∗2 �x2 |y� (2.5)

The first can be derived by pulling the constant out of the ket vector. The second can be obtained by the
same technique, but using the rule given in (2.2). Another important property satisfied by the inner product
is:

�x|y� = (�y|x�) . (2.6)

GO TO FIRST PAGE↑ 37
2.6. ORTHONORMAL BASIS AND COMPLETENESS RELATIONSCHAPTER 2. LINEAR ALGEBRA

This can be proved as follows:



given from equation (2.3): �x|y� = yi∗ xi
i

Consider the RHS: ⇒ yi∗ xi
i
� ∗
⇒ (yi x∗i )
i
� �∗

⇒ yi x∗i
i
coming back to the initial definition of the inner product, we get back:

⇒ (�y|x�)

Therefore, hence we have proved that: �x|y� = (�y|x�)

The inner product of a vector and its bra dual gives the square of the norm of the vector. It is represented
as:

�x|x� = || |x� ||2 (2.7)

COP Y
The norm is a scalar dimensionless quantity. However there is also another possibility which we haven’t yet
explored. What about the product |x��x|. This is a product of a column matrix with a row matrix. This
is a matrix whose dimensions will be (n×n) where n is the dimension of the vector space containing |x�.

T
This product is called an outer product. A matrix can be thought as any transformation on a vector. In

F
quantum mechanics, these transformations or matrices are called operators.

DR
2.6
A
Orthonormal Basis and Completeness Relations

We saw that any set of n linearly independent vectors {|v1 �, |v2 �, |v3 �, ...|vn �} form a basis for a n dimensional
vector space. We impose the condition that:

�vi |vj � = δij for all {i , j} ∈ {1..n}

The basis that satisfies this condition is called the orthonormal basis. The orthonormal basis is conven-
tionally represented by {|e1 �, |e2 �, |e3 �, ...|en �}. The basis has some useful properties. Consider a vector |x�
in the vector space spanned by the basis {ei }. We can use this othonormality property to get the coefficients
multiplying the basis states:

|x� = ci |ei �
i

multiply both sides by �ej |: �ej |x� = ci �ej |ei �
i
since �ej |ei � = δji , it will be 1 only when j = i

Hence, when summed over we get: �ej |x� = cj (2.8)

Therefore, to get the coefficient multiplying the jth basis (orthonormal) state, in some basis expansion for a
vector, we must take an inner product of this basis state with the vector. Now, we can see some results with
the outer product of the basis states. The outer product leads to an operator. Now, let us take the basis

GO TO FIRST PAGE↑ 38
CHAPTER 2. LINEAR ALGEBRA 2.7. PROJECTION OPERATOR

expansion of the vector, and substitute the result obtained in (2.8):



|x� = ci |ei �
i

from (2.8), |x� = �ei |x�|ei �
i
The expression inside the summation is a scalar multiplication with a vector, which is commutative.

Hence: |x� = |ei ��ei |x�
i
Now, consider the RHS of the form A|x�, where A is some operator:
� �

Hence: |x� = |ei ��ei | |x�
i

Now, looking at the form of this operator, we can easily guess that the operator is the identity operator.
Therefore: �
|ei ��ei | = I (2.9)
i

This relation given above is called the completeness relation.

2.7 Projection operator

COP Y
Consider the operator |ek ��ek |. This operator when acts on any vector, gives the coefficient of |ek �, in the

T
{ek } basis expansion of that vector. So, in other words, it takes any vector and projects the vector along

A F
the |ek � direction. The projection occurs due to the formation of the inner product of the bra �ek | with the

DR
arbitrary ket vector. Hence this operator |ek ��ek | is called the projection operator. It is denoted by Pk .
The projection operator satisfies some important properties:
n
(Pk ) = (Pk ) (2.10)
Pi Pj = 0 (2.11)

Pi = I (2.12)
i

The first (2.10) can be proved by considering Pk = |ek ��ek |. So,


n
(Pk ) = |ek ��ek |ek ��ek |ek ��ek | . . . |ek ��ek |.

Now, we can simplify this statement by taking inner product. We use the fact that |ek ��ek | = 1. Only the
terminal terms |ek � and �ek | will remain after all the others become 1 through formation of inner product.
Hence, the RHS will be Pk .
The second statement (2.11) can be proved by considering: Pi Pj = |ei ��ei |ej ��ej |. Now the inner product
�ei |ej � is = 0. The entire expression reduces to 0. Hence the statement is proved.
The third statement (2.12) is already proved in (2.9)when we were looking at the orthonormal basis and
completeness relation.

2.8 Gram-Schmidt Orthonormalization


Given any linearly independent basis that span some vector space, we can construct an orthonormal basis
for the same vector space. If we could not do that, then the original set of vectors would not be a basis since
we have a vector that cannot be uniquely described by them. Therefore, we proceed by trying to construct
an orthonormal set of basis from a given basis. So, our objective is that we must construct a set {Oi },

GO TO FIRST PAGE↑ 39
2.8. GRAM-SCHMIDT ORTHONORMALIZATION CHAPTER 2. LINEAR ALGEBRA

corresponding to the set {vi }. By defenition of orthonormal basis, we need that �Oi |Oi � = 1 ; that is, the
norm of the vector must be = 1. So, we try to construct that basis. So, we define

|vi �
|Oi � = (2.13)
|||vi �||

Our set of vector are all of unit norm. So, we now need to construct orthogonal vectors out of these. Our
idea briefly is the following: We can illustrate this process taking an example of vectors on a 2 dimensional
real vector space, which can be represented on a plane.

1. Without any loss of generality, take the first vector from the given basis (call it |x1 �) as the first
orthonormal vector |O1 �. Hence, we have:

COP Y
DR A F T
Figure 2.1: V1 and v2 are the given basis vectors. choose O1 = v1

2. Take the second vector (|x2 �) from the given basis, and take the projcection of |x2 � along this vector
|O1 �. (take the projection by multiplying |O1 ��O1 | with |x2 �.) Now, we have a vector that is on |O1 �.
Call this vector |PO1 �.

Figure 2.2: diagram for step 2. Project x2 along O1

3. Now, subtract PO1 from |v2 � (using the triangle law of vector addition) to get |e2 �.

GO TO FIRST PAGE↑ 40
CHAPTER 2. LINEAR ALGEBRA 2.9. LINEAR OPERATORS

Figure 2.3: diagram for step 3. get O2 = Po1 - O1. Dotted line indicates that we are subtracting vectors
using the triangle law of addition

4. Now, by construction, this vector |e2 � is orthogonal to |e1 �.

COP Y
DR A F T Figure 2.4: final diagram: O1 and O2 are formed.

5. Similarly, we go on to construct all {ei }. In the ith step, we need to subtract all the projections of vi
from vi .
So, summing up, we got |O2 � from performing (|v2 � - |PO1 �). We get |PO1 � by performing (|O1 ��O1 |)|v2 �.
Hence, we can write:
�j−1
|vj � − |Oi ��Oi |vj �
|Oj � = �ij−1 where (1 ≥ j ≥ n) (2.14)
|||vj � − i |Oi ��Oi |vj �||

The denominator is to normalize (setting its norm to 1) the resultant vector. n is the dimensionality of the
vector space spanned by the given basis.

2.9 Linear Operators


A linear operator between two vector spaces V and U is any function A: V → U which is linear in its
inputs:
� �
� �
A ai |vi � = ai (A|vi �) (2.15)
i i

From the above definition we see that this function has a value in U for each element in V. If V and
U are n and m dimensional respectively, then the function will transform each dimension out of n to a

GO TO FIRST PAGE↑ 41
2.10. HERMITIAN MATRICES CHAPTER 2. LINEAR ALGEBRA

dimension out of m. In other words, if V is spanned by the basis {|v1 �, |v2 �, |v3 �, ..., |vn �} and U is spanned
by {|u1 �, |u2 �, |u3 �, ..., |um �}, then for every i in 1 to n, there exists complex numbers A1i to Ami such
that

A|vi � = Aij |ui �
i

If V and U are n and m dimensional respectively, then A is m × n dimensional.

2.10 Hermitian Matrices


A matrix is called self adjoint or Hermitian if the matrix is equal to its own conjugate transpose. That
is, the element

Aij = (Aji )∗ (2.16)

The congugate transpose of a matrix A is denoted as A† , where, the relation (2.16) holds. Hence, a matrix
A is called Hermitian iff A = A† . Since we have this property for the Hermitian matrix, we can see that the
diagonal entries of a Hermitian matrix have to be real. A very similar defenition holds also for Hermitian
operators as well. An operator is Hermitian if its matrix representation has a hermitian form. The conjugate

Y
transpose for an operator is defined as:

T COP
Therefore if an operator is Hermitian,
�u|A|v� = (�v|A† |u�)∗ (2.17)

A F �u|A|v� = (�v|A|u�) (2.18)


DR
is satisfied. The defenition of a conjugate transpose applies to vectors as well. We have:
� �
(|x�)† ≡ x∗1 x∗2 x∗3 . . . x∗n ≡ �x| (2.19)

Let us look at some derivations that we need to know before proceeding:

(A + B)† = A† + B† (2.20)

(cA) = c A
† ∗
(2.21)
(AB)† = B† A† (2.22)

Let us consider the matrix A to have the elements {Aij } and the matrix B to have the elements {Bij }. If
we prove that the above laws hold for an arbitrary element in A and B, then we have proved the laws in
general.
For the first case, (A + B) will have the elements (Aij + Bij ). Let this sum be = Cij . Now, C† will have
the elements (Cji )∗ . By laws of matrix addition,

(Cji )∗ = (Aji )∗ + (Bji )∗ .


Since, (Aji ) + (Bji )∗ = (Aji + Bji )∗

therefore, we can say
(Cji ) = (Aji + Bji ) . Since it is true for Cji it should also hold for Cij :
∗ ∗

Therefore (Cji )∗ = (Aji + Bji )∗

Now since the law holds for the individual elements, in should also hold for the matrix or the operator.
Therefore, we can say:

C† = (A + B)†

GO TO FIRST PAGE↑ 42
CHAPTER 2. LINEAR ALGEBRA 2.11. SPECTRAL THEOREM

2.11 Spectral Theorem

Theorem: For every normal operator M acting on a vector space V, there exits an orthonormal basis for V
in which the operator M has a diagonal representation.
P roof : We try to prove this theorem by induction on the dimension of the vector space V. We know that
for a single dimensional vector space, any operator is diagonal. So, let M be a normal operator on a vector
space V which is n dimensional. Now, let the eigenvalues and the eigenvectors of M be λ and |a� respectively.
Let P be a projector on to the λ eigen space of M. Let Q be the complement of the projector operator.
We have the relation (P + Q) = I and M|a� = λ|a�.
We now make a series of manipulations that shall give us the operator A, that acts on V, in terms of
some operator(s) that/those act on subspace(s) of V. This is because, we already assume that the spectral
theorem holds in the lower dimensional spaces (subspaces), i,e; Every normal operator in the subspace of V
has diagonal representation in some orthonormal basis for that subspace.

M = IAI
using P + Q = I:
M = (P + Q)M(P + Q) ⇒ PMP + PMQ + QMP + QMQ

Now from these terms, we need to filter out those on the RHS which evaluate to 0. Taking QMP: P projects
a vector on to the λ eigen space. In other words, P acts on a vector to give the coefficient of the λ eigen state

Y
(when the vector is expressed as a linear combination of the λ eigenvectors). M Then acts on these eigen

P
vectors (produced by P), giving the same eigen vector as the result. Q now acts on a vector and projects

O
it to the space which is an orthogonal complement of the λ eigen space. Since after the action of P and M,

C
the result is an eigen vector of the lambda eigen space, the action of Q will now try to express this eigen

T
vector as a linear combination of the eigen vectors which do not belong to the λ eigen space. But this is not

F
possible since we have assumed that the eigen vectors are orthonormal (linearly independent). So, the action

DR A
of Q, shall produce 0. Therefore, QMP = 0.
We can now look at the term PMQ. Let us first consider that M is a normal matrix.

MM† = M† M
MM† |a� = M† M|a�
⇒ M(M† |a�) = λ(M† |a�)

We now see that M† |a� is also an eigenvector of M, with eigenvalue λ. Therefore, from the same above
argument, M† acting on an eigenvector of the λ eigen space, gives another eigen vector in the λ eigen space.
Hence, PA† Q = 0. Now, taking the adjoint of this equation gives: (PM† Q)† = 0 ⇒ QMP = 0 (as P and
Q are Hermitian operators).
Therefore, we have:

M = PMP + QMQ (2.23)

The operators on the RHS act on a subspace of V and we had already assumed that the spectral theorem
holds for the subspace. So, if these operators are normal, then we can say that they are diagonal in some
basis.
We can easily show that PMP is normal. Since, P projects any vector onto the λ eigen space, and then
the vector is left unchanged by action of PM (except for a scalar multiplication of λ by M), we can say
that:

(PMP)† = λP (2.24)

Now since P is hermitian, it is obviously normal. Therefore we conclude that PMP is also normal.

GO TO FIRST PAGE↑ 43
2.12. OPERATOR FUNCTIONS CHAPTER 2. LINEAR ALGEBRA

Similarly, we have that QMQ is also normal. (Note: Q2 = Q and Q is hermitian.) We also have:

QM = QMI ⇒ QM(P + Q)
∴ QM = QMP + QMQ ⇒ QMQ (2.25)
† †
Similarly: QM = QM Q (2.26)

So, using the above equations, we can prove that QMQ is normal.

(QMQ)† = QM† Q
∴ (QMQ)† (QMQ) = QM† QQMQ
since Q2 = Q: ⇒ QM† QMQ
since QM† Q = QM† : ⇒ QM† MQ
since MM† = M† M: ⇒ QMM† Q
since QM = QMQ: ⇒ QMQM† Q
since Q = Q2 : ⇒ QMQQM† Q
Therefore, we have: ⇒ (QMQ)(QMQ)†

Therefore, QMQ is normal. Now since PMP and QMQ are normal operators on subspaces P, and Q, they

Y
are diagonal in some orthonormal basis for those respective spaces.

P
Now since PMP and QMQ are diagonal, their sum is also diagonal in the combined basis. Therefore,

O
M = PMP + QMQ is diagonal in some basis for V. Hence, proved.

2.12

F T C
Operator functions

A
DR
It is possible to extend the notion of functions, defined over complex numbers, to functions that are defined
over operators. It is necessary that these operators are normal. A function on a normal matrix or an normal
operator
� is defined in the following way: Let A be some normal operator which has a spectral decomposition
as |a��a|, then:
a

f (A) = f (a)|a��a| (2.27)
a

So, in the above equation, we try to represent the operator in a diagonal form and then apply the function
to each diagonal entry. We can try to prove the above equation for special functions. Take for example An
for some positive n.

An |a��a| = λa An−1 |a��a| ⇒ λ2a An−2 |a��a| ⇒ . . . · · · ⇒ λna |a��a| (2.28)


� �
� � �
From completeness theorem, we have: |a��a| = I. ∴ An = An |a��a| ⇒ An |a��a|
a a a

From equation (eq. 2.28), we see: A |a��a| =
n
λna |a��a| ∴A =
n
λna |a��a| (2.29)
a

Hence, we prove equation (eq. 2.27) for the special case of the function being a power operation.

1. Why must we, in general, consider only normal operators or, operators having a spectral decomposition
?
2. How do √we prove equation (eq. 2.27) for the case of square-root or a logarithm operation, that is
f (A) = A and f (A) = log A ?

GO TO FIRST PAGE↑ 44
CHAPTER 2. LINEAR ALGEBRA 2.13. SIMULTANEOUS DIAGONALIZABLE THEOREM

2.12.1 Trace

The trace of a matrix is also a function on the matrix. The trace of a matrix A is defined as the sum of all
the diagonal elements of A. (A need not be a diagonal matrix ). The trace can be defined using the matrix
notation as well as the outer product form:

tr(A) = Aii (2.30)
i

tr(A) = �i|A|i� (2.31)
i

2.13 Simultaneous Diagonalizable Theorem


Theorem: Any two hermitian matrices A and B commute if and only if there exists an orthonormal basis
in which the matrix representations of A and B are diagonal.
We say that A and B are simultaneously diagonalizable if there exists some basis where the matrix represen-
tations of A and B are diagonal.
Proof : Let A and B be two operators that have a common orthonormal basis in which they are diagonal.
This means they have a common eigen basis. Let the eigen basis be denoted by the set of eigenstates labelled
|a, b�. So, we have:

COP Y AB|a, b� = bA|a, b� = ab|a, b�


BA|a, b� = aB|a, b� = ba|a, b�
(2.32)
(2.33)

T
On subtracting the above two equations: (eq. 2.32) - (eq. 2.33), we get

DR A F AB - BA|a, b� = (ab − ba)|a, b�


AB - BA = 0 ⇒ [A, B] = 0

We have shown that A and B commute if and only if they have a common eigen basis. This proves the
simultaneous diagonalizable theorem.
(2.34)

Proof of the converse: Let |a, j� be the eigenstates of the operator A with the eigenvalue a. Here the index j
denotes the degeneracy. Let the vector space containing all eigenstates with eigenvalue a form the vector space
Va . Let the projection operator onto the Va eigenspace be called Pa . Now let us assume that [A, B] = 0.
Therefore, we have:

AB|a, j� = BA|a, j� = aB|a, j� (2.35)

Therefore, we have that B|a, j� is also an element of the eigenspace Va . Let us define an operator

Ba = Pa BPa (2.36)

We can now try to see how Ba acts on an arbitrary vector. From definition (def. 2.36), we can see that Pa
will cut off all the components of a vector which does not belong to the Va eigen space. The vector, produced
by action of Pa , shall belong to the Va eigenspace. The action of B on any vector inside the Va eigenspace
will leave the vector unchanged (except for multiplying it by a scalar). The action of Pa again on the vector
produced by the action of B, will leave the vector unchanged. Now we can see how B†a acts on any arbitrary
vector. We have B†a = PB† Pa . When an arbitrary vector is acted upon by Ba , the projection operator
Pa shall produce a vector that entirely lies in the Va eigen space. The action of B†a will now produce some
arbitrary vector (since the vector produced by Pa may not be an eigenstate of B†a . The projection operator
Pa again acts on this arbitrary vector (produced by B†a ), taking it again to a vector that entirely lies in the
Va eigen space.
Therefore, summing up, we can say that the action of Ba and B†a on any arbitrary vector, produce the same
vector in the Va eigenspace. Therefore, we say that restricting Ba to subspace Va , Ba and B†a are equivalent

GO TO FIRST PAGE↑ 45
2.14. POLAR DECOMPOSITION CHAPTER 2. LINEAR ALGEBRA

operators acting on Va . In other words, the restriction of Ba to space Va is hermitian on Va .


Since, Ba is hermitian operator on Va , it must have a spectral decomposition in terms of an orthonormal set
of eigenvectors in Va . These eigenvectors are both eigenvectors A (since they belong to the Va eigenspace)
and Ba (since they are part of the spectral decomposition of Ba ). Let us call these |a, b, k� where the indices
a, b denote the eigenstates of A and B operators respectively, and k denotes the degeneracy.
We have the eigenvalue equation Ba |a, b, k� = b|a, b, k�. Since, |a, b, k� is an element of the Va eigenspace, we
have:

Pa |a, b, k� = |a, b, k� (2.37)

From equation (eq. 2.35), we also have that B|a, b, k� is an element of space Va . So, similarly we can
say:

Pa B|a, b, k� = B|a, b, k�

We can now modify this (above) equation by replacing |a, b, k� on the LHS with Pa |a, b, k� (refer equation
(eq 2.37)):

B|a, b, k� = Pa BPa |a, b, k� (2.38)

If we compare the RHS of the above equation with equation (eq 2.36), we see that the RHS of the above
equation is the same as Ba |a, b, k�. Since |a, b, k� is an eigenstate of Ba with eigenvalue b, we can rewrite

Y
equation (eq. 2.38) as:

P
B|a, b, k� = b|a, b, k� (2.39)

F T CO
Therefore, in the above equation, we see that |a, b, k� is also an eigenstate of B with eigenvalue b. Hence, we
see that the set of vectors |a, b, k� form a common eigen basis for A and B. Hence, we have proved that if
[A, B] = 0, then there exists an orthonormal basis where A and B are diagonal.

DR
2.14
A
Polar Decomposition
Theorem: For every matrix A, there exists an unitary matrix U and positive (semidefinite) matrices J and
K, such that:

A = UJ = KU (2.40)

The above theorem can be said as: for every positive operators J and K, there exists an unitary matrix U
such that: KU = UJ .

Proof: Consider the operator J = A† A. By construction, the operator J is Hermitian. Therefore, there
exists
� a spectral decomposition for A (involving its eigen values and its eigen states). Therefore, let J =
λi |i��i| , where λi (≥ 0) and |i� are the eigen values and the corresponding eigen states of J.
i
Let us define

|ψi � = A|i� (2.41)

From this, we can see that �ψ|ψ� = λ2i . Now, consider all the non-zero eigen values, that is all λi �= 0.
Let
|ψi �
|ei � = (2.42)
λi
�ψi |ψi �
Therefore, we have �ei |ej � = = δij (since all λi ’s are real). We are still considering the λi �= 0
λ i λj
case. Let us now construct an orthonormal basis using the gram-schmidt orthonormalization technique, by

GO TO FIRST PAGE↑ 46
CHAPTER 2. LINEAR ALGEBRA 2.15. SINGULAR VALUE DECOMPOSITION

starting off with the


�first vector as |ei �. We can now construct an orthonormal basis {|ei �}. Let us now define
an operator U = |ei ��i|. When λi �= 0, we have UJ|i� = U(λi |i�) since |i� is an eigenstate of J. Now,
i
U(λi |i�) = λi U|i� . On using the definition of the operator U, we have: U|i� = |ei �. Therefore, UJ|i� = λi |ei �
. From definition, (def 2.42), λi |ei � = |ψi �, where again from definition (2.41), we have |ψi � = A|i�. So, we
can summarize to say that
UJ|i� = A|i� (2.43)
When λi = 0, we have as before, UJ|i� = U(0), and even |ψi � = 0 from relation (def. 2.42). So, we again
have UJ|i� = A|i�. Therefore, we see that operators UJ and A agree with each other on all values in the |i�
basis. Hence we can say that there exists a basis where A = UJ.
A = UJ (2.44)
This gives the right polar decomposition of A. Also, we can prove that the matrix J is unique for every A.
This can be done by multiplying the equation (eq. 2.44) with its adjoint. So, we have:
A† A = JU† UJ
Since J is hermitian: A† A = J2 ⇒ J

∴ J = A† A (2.45)
We can also get an expression for the operator U from equation (eq. 2.44), by post-multiplying both sides

Y
by J−1 . We have:

T COP U = AJ−1 (2.46)


The above definition is however possible only if J is invertible or non-singular. Also whether J is invertible
or not depends on the existence of inverse for A. Therefore, we see from the equation (eq. 2.46) that if A is

F
invertible, then the U is also uniquely defined for every A.

DR A
We can also do the left polar decomposition, starting from equation (eq. 2.44). On post-multiplying the RHS
with U† U (since U is unitary, this will preserve the equality) we get:
A = UJU† U

Now, let UJU† = K. So, we can rewrite the above equation as:
A = KU (2.47)
This now gives us the left polar decomposition of A. We can also show that the matrix K is uniquely defined
for every A. For this, we multiply the equation (eq. 2.47) with its adjoint. We get
AA† = KUU† K
Since, U is unitary, we have

K= AA† (2.48)
Therefore, we see that K is uniquely defined. Similarly, we can show that if A is invertible, then the matrix
U is uniquely defined.
Therefore, combining both the left the and right parts, we have proved the polar decomposition.

2.15 Singular value decomposition


Theorem: For every matrix A, there exist unitary matrices U and V and a diagonal matrix D with positive
entries, such that:
A = UDV (2.49)
The diagonal entries of the matrix D are known as the singular values of A.

GO TO FIRST PAGE↑ 47
2.15. SINGULAR VALUE DECOMPOSITION CHAPTER 2. LINEAR ALGEBRA

2.15.1 Proving the theorem for the special case of square matrices:

Before proving, we can go through some simple assumptions:

1. If A and B are two unitary matrices, then their product: AB is also an unitary matrix.
Proof: To prove (AB)† AB = I , given A† A = AA† = B† B = BB† = I:

(AB)† (AB) = B† A† AB
Since A and B are unitary: ⇒ B† IB ⇒ IB† B ⇒ I
∴ (AB)† (AB) = I

Proof: From the polar decomposition of A, we know that there exist� an unitary matrix S and a positive
operator J such that A = SJ . Since J is hermitian (it is defined as A† A (Refer equation (eq. 2.45)) ,
we know that there exits a spectral decomposition for J. So we have some unitary matrix T and a diagonal
matrix D (with non-negative entries) such that J = TDT† . Now, we can rewrite A as:

A = SJ = STDT†

OP Y
Now, since S and T are unitary, from assumption (1), we can define another unitary operator U = ST.

C
Also since T is unitary (which implies that T† too is unitary), we can define another new unitary operator

T
V = T† . Putting all the definitions together, we have:

DR A F A = STDT† ⇒ UDV

The above equation (eq. 2.50) now, proves the singular value decomposition theorem.
(2.50)

2.15.2 Proving the theorem for the general case:

Proof: Before getting into this proof, we first need to make some assumptions.

1. For every operator A, the operator A† A is Hermitian.

2. The eigenvalues of a Hermitian operator are real positive numbers.

3. The eigenvectors, corresponding to different eigenvalues, of a Hermitian operator are orthogonal.

Let us now construct a hermitian operator A† A. Consider the eigenvalue equation for this operator:

A† A|λi � = λi |λi � (2.51)

Also, we claim (from assumption no.2) that the set of eigenvalues {λi } are all real positive numbers.
That is if the set of eigenvalues {λi } are all > 0 for 0 < i ≤ r, then the set {λi } for (r+1) ≤ i ≤ n are all zero.
Therefore, the set of eigenvectors {|λ1 �, |λ2 �, |λ3 �, . . . |λr �} is orthogonal to the set of eigenvectors {|λr+1 �, |λr+2 �, |λr+3 �, . . .
as they correspond to different eigenvalues (see assumption 3).

GO TO FIRST PAGE↑ 48
CHAPTER 2. LINEAR ALGEBRA 2.15. SINGULAR VALUE DECOMPOSITION

Let us now construct a few operators:


� �
V = |λ1 � |λ2 � |λ3 � . . . |λr � |λr+1 � . . . |λn � (2.52)
 � 
λ1 �
 
 λ2 
 
 . 
 
 . � 
D=  (2.53)
 λr 
 
 . 
 
 . � 
λn
� �
U = |µ1 � |µ2 � |µ3 � . . . |µr � |µr+1 � . . . |µm � (2.54)
1
where |µi � = √ A|λi � (1 < i < r) (2.55)
λi
The set {|µ1 �, |µ2 � . . . |µr �} is orthogonal to {|µr+1 � . . . |µm �} (2.56)

Now, having these definitions, we can perform the product UDV :
� � � � � � � �
UD = λ1 |µ1 � λ2 |µ2 � λ3 |µ3 � . . . λr |µr � λr+1 |µr+1 � . . . λm |µm � (2.57)

From equation (eq. 2.55), we see that the entries (|µ1 �, |µ2 �, |µ3 � . . . |µr �) of the matrix UD can be simplified.

Y
The entries after |µr � can be left unchanged.

P
� �

O
� �
∴ UD = A|λ1 � A|λ2 � A|λ3 � . . . A|λr � λr+1 |µr+1 � . . . λm |µm � (2.58)

C
 

T
�λ1 |

F
 �λ2 | 
 

A
 . 
 

DR
� �  . 
� �  
λm |µm �  �λr | 


⇒ UDV = A|λ1 � A|λ2 � . . . A|λr � λr+1 |µr+1 � . . .  (2.59)
 �λr+1 | 
 
 . 
 
 . 
�λn |
� �
∴ UDV = A|λ1 ��λ1 | + A|λ2 ��λ2 | + · · · + A|λr ��λr | + λr+1 |µr+1 ��λr+1 | + · · · + λm |µm ��λm | (2.60)

Since, we assumed earlier that eigenvalues {λi } are all zero for i > r, can ignore the terms in equation (eq.
2.60) that succeed A|λr ��λr | .
r


UDV = A |λi ��λi | (2.61)
i=1

Since, we know that all the eigenvalues after λr are 0, we have A|λi � = λi |λi � = 0 for r < i < n. If A|λi � = 0,
then A|λi ��λi | = �λi |0 ⇒ 0 . So, we can write the above equation (eq. 2.61) as:
r
� n

UDV† = A |λi ��λi | + A|λi ��λi | (2.62)
i=1 i=r+1
n

⇒ UDV† = A |λi ��λi | (2.63)
i=1

The above is the completeness relation. So, we can write:


UDV† = AI
∴ UDV† = A (2.64)
Hence, we prove the singular value decomposition statement.

GO TO FIRST PAGE↑ 49

You might also like