You are on page 1of 53

Chapter 1

OPERATORS & PRINCIPLES of WAVE


MECHANICS

Schrodinger PauliHaisenberg
Ehrenfest
Brillouin
Debey de Broglie
Compton
Bragg Dirac Born Bohr
Curie Einstein
Lorentz
Planck

Solvay Conference, 24-29 October 1927, Belgium


Vectors & Operators
A vector A can be represented as a linear combination of a basis vectors i,j,k,
i.e., 
A  Axiˆ  Ay ˆj  Az kˆ

Also it is possible to write a vector in a basis using a column matrix, that is,

 Ax 
  
A   Ay 
A 
 z
A basis is a set of linearly independent vectors such that every vector in a
given vector space can be represented uniquely as a linear combination of
these basis vectors.
The dot product between two vectors in matrix form is written as

 Bx 
   
A  B  Ax Ay Az  B y   Ax Bx  Ay B y  Az Bz
B 
 z
And the basis vectors are written in matrix form as
1 0 0
     
ˆi   0  ĵ   1  k̂   0 
0 0 1
     
To generalize to the N-dimensional space, a vector a can be written as
N
a   ai i
i 1
Where ai are the components of the vector and i are the basis vectors such
that
1 i  j
Ψ i Ψ j   ij  
0 i  j
In the Dirac notations the column vector is written as
 a1 
 
 a2  With a
a  a    is called a ket vector.

 
 aN 
The adjoint of a is called the bra vector written as

a


 a  a1 a2  aN 
N
The dot product is written now as a  b  a b   aib j ij
i 1
Definition: An operator A in a vector space is a rule that maps every function
 into a new function ', i.e.,
Ψ   AΨ
Or an operator A in a vector space is a rule that maps every vectora into
a new vector, i.e.,
b  A a

If the operator A satisfies the relation


AΨ a  Ψ b   AΨ a  AΨ b

with  and  are arbitrary complex numbers. Then the operator A is called a
linear operator.
Exercise:
Verify which of the following operators are linear?

d
(i) A  log (ii) B  (iii ) C (iv) D  
dx
Matrix Representation of a Linear Operator:

Let A be a linear operator such that

A a  b

with a   ai i and b   bi i
i i

 b  A ai i   ai A i
i i

Multiply the last equation by the bra vector  j 

 j b   ai  j A i
i
but  j b   bi  j i   bi ij  b j  b j   ai Aij
i i i
with Aij   j A i are called the matrix element of the operator A.

In matrix form the above equation can be written as

 b1   A11  A1N  a1 
    
       
b   A  
 N   N 1  ANN  a N 

That is, a linear operator can be represented by a NN matrix in a N-


dimensional basis.

The linear operators obey the following rules

(i )  AB C  ABC 
(ii ) A B  C   AB  AC
(iii) AB  BA
The Commutator

The square bracket  A, B  AB  BA


Is called the commutator of the two operators A and B.

If A, B  0 Then the operators A and B are called commute.

Deffinition: The adjoint operator

If A a  b and a A†  b

Then the operator A† is called the adjoint operator. In another word we have

 a A

b   b A a

Definition: The Hermitian operator

If A  A† the operator A is called Hermitian operator.

That is, for Hermitian operator we have

 a A b   b A a

Deffnition: The Unitary operator

If A†  A1 the operator A is called Unitary operator.

Now for unitary operator we have AA†  A† A  AA1  A1 A  1

Let i be a set of orthonormal bases vectors, then If U is unitary operator

U i  i taking the adjoint 

 j  j U † multiplying the 1st equation by the 2nd one 


 j  i   j U †U i   j i   j i   ij

The vector i are another orthonormal basis, i.e.,

the Unitary operator can be regarded as defining a translation, an inversion, or


a rotation.

In another word, under a unitary operator the vector preserves its magnitude
and its relative orientation (all scalars are invariant under unitary operations).

To show how an operator transforms under unitary operator we let

U i  i and U i  i

The number  i F i Should be invarient under unitary transformation, i.e.,

i F i  i F  i

With F  is the new operator under unitary transformation.


Now i F  i   i † F  i  U i † F U i

or i F  i   i U † F U i

Now since  i F i  i F  i 

i F i  i U † F U i  F  U † F U
Multiplying the last equation from the left by U and from the right by U-1 we get

UFU †  UFU -1  F 
Exercise:

(i) Prove that the operator U is unitary if and if A is Hermitian, with U  e i A


(ii) Show that if an operator G is invarient under unitary transformation then

G,U   0
Definition: The Projection operator

With each basis vector i we introduce a projection operator Pi.

The effect of Pi on an arbitrary vector a is to produce a new vector in

the direction of i and its magnitude is i a , i.e.,

Pi a  i a i  i i a 

Pi  i i Dyad product

Now Pi  j  i i  j   ij i 

The basis vectors are eigenvectors of the projection operator with eigen values
of ij .

Now we say that a basis is complete if any arbitrary vector can be represented
completely in that basis, i.e., if i is complete 
N N N
a   i a i   i i a   Pi a 
i 1 i 1 i 1
N N
 Pi   i i  1 Completeness relation
i 1 i 1

Theorem: The eigenvalues of a Hermitian operator are all real and their
eigenvectors corresponding to different eigenvalues are orthogonal.
Proof: Let A be a Hermitian operator and

A a  ai a & b A†  b bi

Multiplying the 1st by b  b A a  ai b a (1)

Multiplying the 2nd by a  b A† a  bi b a (2)

Now subtracting the last two equations we get



b A a  b A† a  ai  bi b a  (3)

Since A is Hermitian, the L.H.S of Eq.(3) is zero, i.e.,

ai  bi  b a 0 (4)

Setting a=b and noting that a a  0 

ai  ai   0  ai  ai

If the two eigenvalues are different, i.e., ai  bi 

b a  0

i.e., the two eigenvectors are orthogonal.


The Concept of Degeneracy
If two or more of the eigen values of an operator are equal, the operator is
called degenerate, otherwise it is non degenerate. In another ward

If more than one linearly independent eigen functions ( 1  2 )


corresponding to the same eigen value, we say that we have a degeneracy.

Theorem: If two operators have a common set of eigen vectors, then they
commute.
Proof Let A a , b  ai a , b
and B a , b  bi a , b

Operate on the 1st equation by B  BA a , b  ai B a , b


Operate on the 2nd equation by A  AB a , b  bi A a , b
Now subtracting the last two-equations we get

BA  AB a , b  aibi  bi ai  a , b  0  BA  AB  B, A  0


Theorem: If two non degenerate operators commute, then they have a
common set of eigen vectors.

Proof Let A a  a a and B b  b b

Operate on the 1st equation from the left by B 

BA a  aB a
but since A and B are commute  AB a   aB a  

B a  Is an eigen vector of A with the eigenvalue a.

But the operator A is nondegenrate. 

B a  Must be at worst proportional to a 

B a   b a With b here is the proportionality constant

But the operator B is also nondegenrate.  a  b  a , b


Theorems:

1- The sum of two Hermitian operators is Hermitian.

2- The identity operator, which takes every function into itself, is Hermitian.

3- If F is non-Hermitian, then the operators F  F † and


Hermitians

i F  F†  are

4- If F and G are two arbitrary operators, the adjoint of their product is given by

FG   G † F †
Proof of (4) It is known that F      F†  
FG       FG †  (1)

Now, let G    (2)  Eq. (1) becomes

F       FG †  
 F †    FG †  (3)

Now let F †    (4)  Eq. (3) becomes

    FG †  (5)

Taking the adjoint of Eq. (2), we get

 G †   (6)
Now substituting eqs. (4&6) into eq.(5) we get

 G † F †    FG †  


FG   G † F †
Corollary: The product of two Hermitian operators is Hermitian if and only if
they commute.

Commutator Algebra: It is easy to prove the following identities.

(1) A, A  0 (2) A, B  B, A  0


(3) A, B  C   A, B  A, C  (4) A  B, C   A, C   B, C 

(5) A, BC  BA, C   A, BC (6) AB, C   AB, C   A, C B
(7 )  A, B, C   C ,  A, B   B, C , A  0
A 1 1
(8) A
e Be  B   A, B    A,  A, B    A,  A,  A, B   
2! 3!
A B A B  12  A, B 
(9) e e e with A and B commute with their commutator

Identity (9) is called Baker-Hausdorf identity (Glauber’s theorem).


Proof of (8)
Let f    eA BeA
Making Taylor expansion of the function defined above 

df 2 d 2 f
f    f 0    
d  0 2! d2
 0

But
df
d
  
 A eA BeA  eA BeA A  Af    f   A   A, f   

d2 f  df 
and   A,    A,  A, f  
d2  d 
1
Knowing that f 0  B  f    B   A, B    A,  A, B   
2!
Letting =1 we get

A 1 1
A
e Be  B   A, B    A,  A, B   A,  A,  A, B  
2! 3!
Proof of (9)

Let f    eAeBe  A B  

A B    A B 
 AeAeBe  A B   eA BeBe  A B   e e  A  B e
df (1)
d
 A  B    A B 
 AeAeBe  A B   e e Ae
df
(2)
d
Using identity (8) with the fact that A&B commute with [A,B] 

eA BeA  B  A, B  

eA B  B  A, B eA  BeA  A, B eA (3)


Similarly by replacing A by B one can show that

eB A  AeB  B, AeB (4)

Now and using Equation (4), Eq. (2) becomes


df
d

 AeAeB e   A B   eA AeB  A, B eB e   A B  

df
df
 A, B eAeB e   A B    A, B  f   
d d
df
   A, B d (5)
f
Integrating Equation (5) from 0 to 1 we get
21 f 1 1
1 1 
df
 f    d  A, B  ln f   0 
1
 A, B  ln  2  A, B 
2 f (0)
0 0 0

Noting that f 0  1 & f 1  e Ae B e  A B  

A B  A B 
1  A, B 
A B A B  12 A, B 
e e e  e2  e e e
Postulates of Quantum Mechanics

The postulates provide the connections between the physical (real) world and
the mathematics of quantum mechanics

Postulate (I): Every physical quantity can be represented by a Hermitian


operator with a complete set of eigenvectors. Such operators
are called observables.


x  x & p 
i

Postulate (II): The quantum values allowed to any observable are determined
by the eigenvalues of the corresponding operator.
Postulate (III): The state of any physical system is characterized by a state
vector of unit length in a complex space or by a normalized
state function (r,t) which is single valued, continuous and
differentiable.

The state function (r,t) contains all the information we can know about the
system.

r ,t 
2 is the probability density for finding the particle at position r.

Then the probability of finding the particle in some finite region of space is then
proportional to the integral of r ,t  over this region.
2

f
(r )    r , t  2 d 3r
i
Postulate (IV): If a system is characterized by a state vector  and if

A i  ai i

then the probability of observing the system with the value ai is given by

Pai   i 
2
1

Now  Pai    i     i i  
2

i i i
N
From the completeness relation we have  i i  1
i 1
  Pai      1
i

The expectation value of an observable is defined as

A   A

A    A  dx
Postulate (V) The time development of a state vector  (r , t is determined by
the equation

H  r , t   i  r , t 
t

or H r , t   i  r , t 
t
Where H=T+V is the Hamiltonian operator representing the total energy of the
system.
To solve the above equation we rewrite it as
d r , t  H
 dt 
 r , t  i 
t
d r , t  H t
 r , t   i  dt (with H is assumed to be t-independent) 
t
o t
o

 r , t   exp  H t  to   r ,0  U t  to  r ,0


 i
  
U t  to   exp H t  to 
 i
with
  
is called the time evolution operator.

The Schwarz Inequality: for any two functions f and g the following inequality
holds

 f d  g d   f
2 2 
gd
2

Or for any two vectors a and b

2
a a b b  a b
The Heisenberg Uncertainty Relation:

If A, B and C are Hermitian operators and satisfying the relation

 A, B   iC 

AB   C 1
2
where A  A2  A
2

Proof
Let a   A  A   and b  B  B   

a a    A  A  A  A     A   2 A  A   A 
2 2

2 2
 A2  2 A  A  A2  A
2

 a a  A2 (1)
Similarly

b b  B 2 (2)
Substituting Equations (1) and (2) in the Schwarz Inequality we have

A2 B 2    A  A B  B   2


(3)

It is known that If F is non-Hermitian, then


F  F † and i F  F †  are Hermitian.

 F can be written as a linear combination of two Hermitian operators, i.e.,


F  F† F  F†
F i (4)
2 2i
Letting F   A  A B  B  Using Eq.(4) we can write

A   A A B  B   B  B  A  A 
A B  B  
2
 A  A B  B   B  B  A  A 
i
2i
Denoting the 1st term by G and the 2nd term by H we have
A  A B  B   B  B  A  A 
H i
2i
AB  A B  A B  A B  BA  B A  B A  B A
i
2i
AB  BA  A, B  
AB  BA C
i
i i i
2i 2i 2i 2

A  C 
 A B  B   G  i  (5)
2
Substituting equation (5) into equation (3) 
2 2
i
A2 B 2   G  i C   A2 B 2  G  C
2 2
But since G and C are Hermitian  G and C are reals 
2
i 2 2
G  C  G  14 C 
2
A2 B 2  G 2  14 C 2

A2 B 2  14 C 2  AB  12 C

This means that only commuting observables can in principle be measured and
specified with perfect precision simultaneously. If A and B are two Hermitian
operators that do not commute, the physical quantities A and B cannot both be
sharply defined simultaneously.

As a direct application of Heisenberg uncertainty relation we have

x, px   i  x px  12 

The Heisenberg uncertainty relation is thus seen to be a direct consequence


of the noncommutivity of the position and momentum operators.
The Virial Theorem: The expectation value of the operator A is given by

A  r , t  A r , t  

d   r , t   A   r , t  
A   A r , t    r , t   r , t   r , t  A 
dt  t  t  t 

Now, from postulate (v) we have

 1
 r , t   H  r , t  And its complex conjugate
t i
 1
 r , t    H  r , t  
t i
d 1 A
A    r , t  HA  r , t    r , t   r , t 
dt i t
1
  r , t  AH  r , t 
i
d 1 A
A    r , t  H , A  r , t    r , t   r , t  
dt i t

d 1 A
A   A, H  
dt i t
If A commute with H and doesn’t depend on t explicitly 

d
A 0 
dt
A is constant of motion or conserved.

Example: Note that the Hamiltonian is given by

p2
H  V (r )   p, H    p,V    ,V 
2m i
p commute with H if V is constant  p is conserved if V is constant (F=0).
d 1
Now let
A r p  r  p  r  p, H 
dt i
 p x2  i 2
but xp x , H    xp x ,  V   p x  x p x , V 
 2  

Similarly for the other two components we can write

 i 2
but yp y , H  p y  y p y ,V

  and zp z , H  
i 2

p z  z p z , V 

p2 p2
  r  V  V  
d
r p    r  ,V   
dt  
p2
  r  V     V    V   
d
 r p 
dt 
p2
   r  V 

For stationary state we have

d
r p 0 
dt

2 T  r  V The Virial theorem

As an application for the one-dimensional harmonic oscillator we have

V  x   12 kx 2  r  V  x V  kx 2 
x
2 T  kx 2  2 V  T  V

E
but H  T  V  2 V  2 T  T  V  12 H  n
2
The Equation of Motion:

From postulate (v) we have

 1
 r , t   H  r , t  Or in wave function formulation
t i
 1 2 2
 r , t   H r , t  with H  T  V     V r  
t i 2
Letting r , t   r  f t  

2 d
 f t  2 r   V r  f t  r   i r  f t 
2 dt
2 1 i d
Deviding by (r)f(t)     r   V r  
2
f t  
2  r  f t  dt

Since each side depends on different variable  each side must equal to
some constant, i.e.,
i d
f t   constant  E 
f t  dt
E
i t
f t   e   e it
d E
f t   f t  
dt i

2 2
and    r   V r   E  r  Time-independent Schrödinger
2 equation
The Continuity Equation

From postulate (v) we have


 1
 r , t   H  r , t  Or in wave function formulation
t i

 1 2 2
 r , t   H r , t  with H  T  V     V r  
t i 2

2 2 
   r , t  V r  r , t  i  r , t 
      (1)
2 t
And its complex comjugate
2 2  
   r , t   V r  r , t   i  r , t  (2)
2 t

Multiplying the 1st equation by   and the 2nd one by  and then subtract 


2i
 
 
 2   2      
t t



2i
 
      
t
  
Letting J  

2i

    be the probability current density

and noting that     2   



 J  0 Continuity Equation
t
The time rate of the probability density in a volume is equal to the negative of
the net flow of particles through that volume.

Now we have  d      J d By Gauss's theorem we have
t


 t d    J  da
S
Since the first integral is over all the space and noting that

  0 at  
the second integral is zero 
 
 t d  t  d  0  the probability density is continuous

The wave equation guarantees the conservation of normalization: If  was


normalized at t = 0, it will remain normalized at all times.

Let us derive the continuity equation for charge and current. It is known that the
charge inside a volume is given by

dS
q    r d  I

the current I is

dq  r 
I  d
dt t
Also for the current I flowing out of a surface we have

  dq
I   J  dS  
dt
The minus sign is because when the current flows out from the volume the
charge decreases within that volume. Equating the last two equations we get

  
 t d    J  dS
  
Using the Divergence theorem we get  t d     Jd

  
   J
t
 
It is easy to show that  and J are related as J  v
Expectation Values of Dynamical Variables:

It is known that x     x  d   x  2 d   x  d 

d 
x   x d 
dt t Using the continuity equation

d
x   x    J  d       xJ  d    x   J d
dt
d
 x       xJ  d   J x d
dt
By Gauss's theorem again we have

    xJ  d   xJ  da  0 as R   
S
d d
x   J x d or r   J d
dt dt
Substituting for J we get

d
dt
r   J d 

2i
 
   d 
Knowing that

 
        

d
dt
r 

2i
 
 
   
 
  
d  

d
dt

r   d 
i

2i
   
d  
zero

i

   d 

2i

   
d   1   
r    d      d 
dt i  i 
d
p  r
dt
 
d d        d
Now p  i  d  i  
dt dt  t   t 
 
Using Schrödinger equation and its complex conjugate we get

d
dt
p 
2
2
  
2 
     2
   d   V
 
  
V  d 
Letting   v &   u the 1st integral of the last equation becomes

2
     
2
2   2 
           d     2
v u   v 2
u  d
2 2
Recalling the Green's theorem

 uv  vu   ds   u 2


v  v 2
u d
S V

 the 1st integral becomes


2

2 S
   
   2 
  ds  0  

d
dt
  
p   V    V  d  d  V    V      V  d
dt

d
p    V d   V  F
dt

From the above we conclude that: Expectation values of dynamical quantities


obey the law Classical Mechanics.
Stationary State Solution:

If at t=to=0 the system is characterized by a state function  r ,0  , then its


future evolution is obtained using the operation
iHt
 r , t   U t ,0 r ,0 with U t ,0   e 

Suppose that the initial state   r ,0  is an eigenstate of the Hamiltonian


denoted by n r , such that
 r ,0   n r 
From the above equations we can write
iHt

 r , t   e  
n r 
But H n r    n n r  

iHt i t
  n
e   n r   e   n r  

 i n t n
 r , t   e  n r  with n 

Now the expectation value of an observable A at t=0 is

A t 0    r ,0Ar ,0dr   n r A n r dr

At later time t we have

A t    r , t Ar , t dr  eint e int  n r A n r dr  A t  0

 In a stationary state we have A t  A t 0


We conclude that the eigensates of the Hamiltonian are called stationary states
given as

 r , t   e  i n t  n r 
stationary state

Note that the probability density of a stationary state is given by

(r )   r , t  2  eint e int  n r  2   n r  2 

The probability density for a stationary state is constant of time.


Theorem: The separation constant E is real.


Proof: From the continuity equation we have  J  0
t


i
E  
E 
t 
But    r , t   r 
2 2
e  

i
 

t

i


E  E e  
 E  E t
 r 
2 i

  E  E 

Substituting back in the continuity equation we have

 

E  E   J
i
Integrating over all the space we get

 E  E d  i    Jd
Using Gausses Theorem we get

 E  E d  i  J  dS
The right hand side of the last equation vanishes as R   

 

 E  E  d  0  E  E  0  E is real

Note that    r , t  2
  r , t  r, t 
E
i t
Substituting for  r , t   e   r     r  r   r  2

The probability density is constant of motion. For this reason the wavefunction
is said to represent stationary state.
Theorem: The expectation values of dynamical quantities which don't depend
on time explicitly are constants.

Proof: The expectation value of a an operator Q (not depend on time explicitly)


corresponding to a dynamical variable is given by

Q    r , t Q r , t  d   Q r  d
2

zero zero
d d    
Q   Q  r  d    Q   r  d   Q  r  d
2 2 2
dt dt  t  t

d
 Q 0  Q is constant.
dt
  r , t  
Now H    r , t  H  r , t  d    r , t  i  d
 t 
 r , t  f t  i
But   r    E r , t  
t t 
H  E   r , t  r , t  d  E  r , t  d  E
2

Also we have from continuity equation



 J  0
t

But   r   0   J  0
2

t
For stationary state  J  0
Symmetry Properties of Schrödinger Equation:

1- Space reflection (parity)

Let r  r  r   r , t      r , t 

  r  
Now        2  2 
r r  r r 
2 2
H   T   V      V  r   If V r   V  r  
2
H  H
For even potential energy, the Hamiltonian is invariant under reflection.
Parity operator: The parity operator is defined as

U p f r   f  r 

Let U p g r   g  r   g r  With  is the eigenvalue of the parity operator.

Operate again with Up 


Let U pU p g r   U p g  r   U p g r 

but U p g  r   g r  and U p g r   g r  

g r    2 g r    2 1    1 
The eigenvalue of the parity operator is 1

For the parity operator we can write

U p r U p †  r & U p p U p †   p 

U p r  rU p & U p p   pU p
2- Time Reversal: Recalling the time-dependent Schrödinger Eq.

The time reversal operator is defined as

U t f t    f t   f  t 

Recalling the time-dependent Schrodinger equation

2 2 
  r , t   V r  r , t   i  r , t  (1)
2 t

Taking the complex conjugate and replacing t  t   t


2 2  
   r ,t   V r  r ,t   i  r ,t  (2)
2 t
Eq.(1) & Eq.(2) have the same form   r , t  &   r ,t  are solutions of Eq.(1)

For the time reversal operator we can write

r †  r & p †   p

You might also like