You are on page 1of 23

Identical particles

Masatsugu Sei Suzuki


Department of Physics, SUNY at Binghamton
(Date: March 19, 2015)

In classical mechanics, identical particles (such as electrons) do not lose their "individuality", despite the
identity of their physical properties. In quantum mechanics, the situation is entirely different. Because of the
Heisenberg's principle of uncertainty, the concept of the path of an electron ceases to have any meaning. Thus,
there is in principle no possibility of separately following each of a number of similar particles and thereby
distinguishing them. In quantum mechanics, identical particles entirely lose their individuality. The principle of
the in-distinguishability of similar particles plays a fundamental part in the quantum theory of system composed
of identical particles. Here we start to consider a system of only two particles. Because of the identity of the
particles, the states of the system obtained from each other by interchanging the two particles must be completely
equivalent physically.

1 System of particles 1 and 2

k ' , k"

 a  k ' 1 k" 2
 b  k" 1 k ' 2

Even though the two particles are indistinguishable, mathematically  a and  b are distinct kets for
k '  k " . In fact we have  a  b  0 . Suppose we make a measurement on the two particle system.

k ' : state of one particle and k " : state of the other particle.

We do not know a priori whether the state ket is  a  k ' 1 k " 2


or  b  k " 1 k ' 2
or -for that matter- any
linear combination of the two: ca  a  cb  b .

2 Exchange degeneracy
A specification of the eigenvalue of a complete set of observables does not completely determine the state
ket.

Mathematics of permutation symmetry:

Pˆ12  a  Pˆ12 k ' 1 k " 2


 k" 1 k ' 2
 b .

Clearly

Pˆ21  Pˆ12 , and Pˆ12 2  1.

Under Pˆ12 , particle 1 having k ' becomes particle 1 having k " ; particle 2 having k " becomes particle 1
having k ' . In other words, it has the effect of interchanging 1 and 2.

((Note))
Identical particles 1 11/10/2019
Matrix element of Pˆ21  Pˆ12 in terms of  a  k ' 1 k " 2
and  b  k " 1 k ' 2

Pˆ12  a  Pˆ12 k ' 1 k " 2


 k" 1 k ' 2
 b ,

Pˆ12  b  Pˆ12 k " 1 k ' 2


 k ' 1 k" 2   a .

Matrix element of Pˆ21  Pˆ12

0 1
Pˆ12   
1 0

or

a b
a 0 1
b 1 0

Eigensystem[ P̂12 ]

 = 1 (symmetric):

P12  symm   symm

1 1
 symm  ( k ' 1 k" 2  k" 1 k ' 2 )  (1̂  Pˆ12 ) k ' 1 k " 2
2 2

where the symmetrizer is defined by

1
Sˆ  (1̂  Pˆ12 )
2

= -1 (anti-symmetric)

P12  anti    anti

1 1
 anti  ( k ' 1 k " 2  k" 1 k ' 2 )  (1̂  Pˆ12 ) k ' 1 k " 2
2 2

where the antisymmetrizer is defined by

1
Aˆ  (1̂  Pˆ12 )
2

Identical particles 2 11/10/2019


Our consideration can be extended to a system made up of many identical particles. A transposition is a
permutation which simply exchange the role of two of the particles, without touching others.

Pˆij k ' 1 k " 2 ..... k i k i 1 ..... k j .....  k ' 1 k " 2 ..... k j k i 1 ..... k i .....
i i 1 j i i 1 j

The transposition operators Pˆij are Hermitian ( Pˆij   Pˆij )

Pˆij  1
2

So that Pˆij is an unitary operator. The allowed eigenvalues of Pˆij are ±1. It is important to note, however, that in
general

[ Pˆij , Pˆkl ]  0 .

(i) The first example

Now we consider a permutation operator Pˆ123 for

Pˆ123 k ' 1 k " 2


k" ' 3  k ' 2
k" 3
k" ' 1  k" ' 1 k ' 2
k" 3

for the system of 3 identical particles ( k ' 1 k " 2


k''' 3 )

1 2 3
P123   
2 3 1

This means replacement of 12, 2 3, 3 1.

1 2 3 1 2 32 1 3


P123         P P
2 3 1 2 1 32 3 1 12 13

Quantum mechanically this is not correct. The correct one is

Pˆ123  Pˆ13Pˆ12

(ii) The second example

1 2 3 1 2 31 3 2


P123         P P
2 3 1 1 3 22 3 1 23 12

or quantum mechanically

Pˆ123  Pˆ12 Pˆ23

(iii) The third example


Identical particles 3 11/10/2019
1 2 3 1 2 33 2 1
P132         P P
3 1 2 3 2 13 1 2 13 12

or quantum mechanically

Pˆ132  Pˆ12 Pˆ13

((Proof))

Pˆ12 Pˆ13 k ' 1 k " 2


k ' ' ' 3  Pˆ12 k ' ' ' 1 k " 2
k ' 3  k" 1 k ' ' ' 2
k' 3

and

Pˆ132 k ' 1 k " 2


k''' 3  k' 3
k" 1 k ' ' ' 2
 k" 1 k ' ' ' 2
k' 3

Therefore we have Pˆ132  Pˆ12 Pˆ13 .

Any permutation operators can be broken into a product of transposition operators.

Pˆ132  Pˆ23 Pˆ12  Pˆ13 Pˆ23  Pˆ12 Pˆ13  Pˆ23 Pˆ12 Pˆ232  ...

The decomposition is not unique. However, for a given permutation, it can be shown that the parity of the
number of transposition into which it can be broken down is always the same: it is called the parity of the
permutation.

Pˆ132  Pˆ23 Pˆ12  Pˆ13 Pˆ23  Pˆ12 Pˆ13  Pˆ23 Pˆ12 Pˆ23
2
even permutation

Pˆ123  Pˆ12 Pˆ23  Pˆ13Pˆ12 : even permutation

Pˆ23 odd permutation

Pˆ12 odd permutation

Pˆ31 odd permutation

((Note))

Pˆ132  Pˆ321  Pˆ213

3 Symmetrizer and antisymmetrizer


Now we consider the two Hermitian operators

1
Sˆ   Pˆ : symmetrizer
N! 

Identical particles 4 11/10/2019


1
Aˆ  
N! 
  Pˆ : antisymmetrizer

where   =1 if Pˆ is an even permutation and   = -1 if Pˆ is an odd permutation.


Sˆ  Sˆ and Aˆ   Aˆ .

((Theorem-1))
If Pˆ 0 is an arbitrary permutation operator, we have

Pˆ 0 Sˆ  SˆPˆ 0  Sˆ

Pˆ 0 Aˆ  Aˆ Pˆ 0   0 A (1)

((Proof))
This is due to the fact that

Pˆ 0 Pˆ  Pˆ

such that

     0

or

      2    
0 0

1 1
Pˆ 0 Sˆ   Pˆ 0 Pˆ   Pˆ  Sˆ
N!  N!  

1 1
Pˆ 0 Aˆ  
N! 
  Pˆ 0 Pˆ    0   Pˆ    0 Aˆ
N! 

From Eq.(1) we see the following theorem

((Theorem-2))

Sˆ 2  Sˆ

Aˆ 2  Aˆ

and

Aˆ Sˆ  SˆAˆ  0

Identical particles 5 11/10/2019


((Proof))

1 1
Sˆ 2  
N! 
Pˆ Sˆ   Sˆ  Sˆ
N! 

1 1
Aˆ 2     Pˆ Aˆ     Aˆ  Aˆ
2

N!  N! 

1 1
Aˆ Sˆ  Aˆ  
N! 
  Pˆ Sˆ  Sˆ     0
N! 

since half the  are equal to 1 and half the  equal to -1. Sˆ and Aˆ are therefore projectors. Their action on any
ket  of the state space yields a completely symmetric or completely antisymmetric ket.

Pˆ 0 Sˆ   Sˆ 

Pˆ 0 Aˆ     0 Aˆ 

 S  Sˆ 

 A  Aˆ 

((Example))

For N = 3,

1
Sˆ  [1̂  Pˆ12  Pˆ23  Pˆ31  Pˆ123  Pˆ132 ]
6

and

1
Aˆ  [1̂  Pˆ12  Pˆ23  Pˆ31  Pˆ123  Pˆ132 ]
6

where

Pˆ123  Pˆ12 Pˆ23 , Pˆ132  Pˆ12 Pˆ13

1
Sˆ  Aˆ  (1̂  Pˆ123  Pˆ132 )  1̂
3

4 Symmetrization postulate
The system containing N identical particles are either totally symmetrical under the interchange of any pair
(boson), or totally antisymmetrical under the interchange of any pair (fermion).

Identical particles 6 11/10/2019


Pˆij  N , B   N , B ,

Pˆij  N ,F    N ,F ,

where  N , B is the eigenket of N identical boson systems and  N , F is the eigenket of N identical fermion
systems.

((Note)) It is an empirical fact that a mixed symmetry does not occur.

Even more remarkable is that there is a connection between the spin of a particle and the statistics obeyed by it:

Half-integer spin particles are fermion, while integer-spin particles are bosons.

5 Pauli exclusion principle


Wolfgang Ernst Pauli (April 25, 1900 – December 15, 1958) was an Austrian theoretical physicist and one
of the pioneers of quantum physics. In 1945, after being nominated by Albert Einstein, he received the Nobel
Prize in Physics for his "decisive contribution through his discovery of a new law of Nature, the exclusion
principle or Pauli principle," involving spin theory, underpinning the structure of matter and the whole of
chemistry.

http://en.wikipedia.org/wiki/Wolfgang_Pauli

Electron is a fermion. No two electrons can occupy the same state. We discuss the framatic difference
between fermions and bosons. Let us consider two particles. Each of which can occupy only two states k ' and
k" .
For a system of two fermions, we have no choice

1 1 k' 1 k' 2
[ k ' 1 k" 2  k ' 2
k" 1 ] 
2 2 k' ' 1 k" 2

For bosons, there are three states.

Identical particles 7 11/10/2019


k' 1 k' 2 ,

k" 1 k" 2 ,

1
( k ' 1 k" 2  k" 1 k ' 2 ) .
2

In contrast, for “classical particles” satisfying Maxwell-Boltzmann (M-B) statitics with no restriction on
symmetry, we have altogether four independent states.

k ' 1 k" 2 , k " 1 k ' 2 , k ' 1 k ' 2


and k " 1 k " 2

We see that in the fermion case, it is impossible for both particles to occupy the same state.

6 Transformation of observables by permutation


For simplicity, we consider a specific case where the two particle state ket is completely specified by the
eigenvalues of a single observable Aˆ for each of the particle.

Aˆ1 a' 1 a" 2  a ' a ' 1 a" 2

and

Aˆ 2 a ' 1 a" 2  a" a' 1 a" 2

Since

1 1
Pˆ12 Aˆ1 Pˆ12 Pˆ12 a ' 1 a" 2  Pˆ12 Aˆ1 Pˆ12 a" 1 a' 2

or

1
Pˆ12 Aˆ1 Pˆ12 Pˆ12 a' 1 a" 2  Pˆ12 Aˆ1 a ' 1 a" 2

 a' Pˆ12 a' 1 a" 2

 a' a" 1 a' 2


 Aˆ 2 a" 1 a ' 2
 Aˆ 2 Pˆ12 a ' 1 a" 1

we obtain

Pˆ12 Aˆ1Pˆ12 1  Aˆ2 .

Similarly, we have

Pˆ21 Aˆ2 Pˆ211  Aˆ1 .

It follows that Pˆ12 must change the particle label of observables.

Identical particles 8 11/10/2019


There are also observables, such as Aˆ1  Bˆ2 , Aˆ1 Bˆ 2 , which involve both indices simultaneously.

Pˆ12 ( Aˆ1  Bˆ 2 ) Pˆ121  Aˆ2  Bˆ1

Pˆ12 Aˆ1Bˆ2 Pˆ12 1  Pˆ12 Aˆ1Pˆ12 1 Pˆ12 Bˆ 2 Pˆ12 1  Aˆ2 Bˆ1

These results can be generalized to all observables which can be expressed in terms of observables which can be
expressed in terms of observables of the type of Aˆ1 and Bˆ2 , to be denoted by Oˆ (1,2).

Pˆ12 Oˆ (1,2)Pˆ12 1  Oˆ (2,1)

where Oˆ (2,1) is the observable obtained from Oˆ (1,2) by exchanging indices 1 and 2 throughout.

Oˆ s (1,2) is said to be symmetric if

Oˆ s (1,2)  Oˆ s (2,1)

or

[Oˆs (1,2), Pˆ12 ]  0

Symbolic observables commute with the permutation operator.

In general. the observables Oˆ s (1,2,3,..., N) which are completely symmetric under exchange of indices 1, 2, ..., N
commute with all the transposition operators, and with all the permutation operators

7 Example
Let us now consider a Hamiltonian of a system of two identical particles.

1 ˆ 2 1 ˆ 2
Hˆ  p1  p  Vpair ( ˆr1  rˆ2 )  Vext (rˆ1 )  Vext (rˆ2 )
2m 2m 2

Clearly we have

Pˆ12 Hˆ Pˆ12 1  Hˆ

or

[ Pˆ12, , Hˆ ]  0

Pˆ12 is a constant of the motion. Since Pˆ12 2  1, the eigenvalue of Pˆ12 allowed are ±1.

Hˆ   E 

P̂12    

Identical particles 9 11/10/2019


Pˆ12   Pˆ12   2   
2

or

= ±1.

It therefore follows that if the two-particle state ket is symmteric (antisymmettric) to start with, it remains so at
all times.

(i) N = 2 case
We can define the symmetrizer and antisymmetrtizer as follows.

1 1
Sˆ  (1  Pˆ12 ) Aˆ  (1  Pˆ12 )
2 2

Sˆ  Aˆ  1̂

1 1 1 1
Sˆ 2  (1̂  Pˆ12 ) (1̂  Pˆ12 )  (1̂  2 Pˆ12  1̂)  (1̂  Pˆ12 )  Sˆ
2 2 4 2

1 1 1 1
Aˆ 2  (1̂  Pˆ12 ) (1̂  Pˆ12 )  (1̂  2 Pˆ12  1̂)  (1̂  Pˆ12 )  Aˆ
2 2 4 2

Then we have

1
S  ( k ' 1 k" 2  k" 1 k ' 2 )
2

and

1
A  ( k ' 1 k " 2  k" 1 k ' 2 )
2

where

1 1
Sˆ k ' 1 k " 2  (1  Pˆ12 ) k ' 1 k " 2  ( k ' 1 k " 2  k " 1 k ' 2 )
2 2

1 1
Aˆ k ' 1 k " 2  (1  Pˆ12 ) k ' 1 k " 2  ( k ' 1 k " 2  k " 1 k ' 2 )
2 2

(ii) N = 3 Cases

1
Sˆ  (1  Pˆ12  Pˆ23  Pˆ31  Pˆ123  Pˆ132 )
6

Identical particles 10 11/10/2019


1
Aˆ  (1  Pˆ12  Pˆ23  Pˆ31  Pˆ123  Pˆ132 )
6

1
Sˆ  Aˆ  (1 Pˆ123  Pˆ132 )  1
3

k' 1 k" 1 k''' 1


1
Aˆ k ' 1 k " 2
k''' 3  k' 2 k" 2
k''' 2
3!
k' 3 k" 3
k'' ' 3

Slater determinant
Aˆ k ' 1 k " 2 k ' ' ' 3 is zero if two of individual states coincide. We obtain Pauli’s exclusion principle.

8 Generalized method (Tomonaga)


We now consider a system consisting of many spins.

Sˆ  Sˆ1  Sˆ 2  Sˆ 3  ...  Sˆ N

Sˆ  Sˆ  ( Sˆ1  Sˆ 2  Sˆ 3  ...  Sˆ N )  ( Sˆ1  Sˆ 2  Sˆ 3  ...  Sˆ N )

or

N
ℏ2 N 2 1 2
S   S n  2  ( Sn  Sn ' )   σˆ n  ℏ  (σˆ n  σˆ n ' )
ˆ ˆ ˆ ˆ
2 2

n 1 n n ' 4 n 1 2 n  n'

or

Sˆ 2 3 N 1
2
 1̂   (σˆ n  σˆ n ' )
ℏ 4 2 nn'

Here we define an operator

2
Oˆ   Pˆnn'
N ( N  1) n  n '

Ô is Hermitian and [ Pˆ , Oˆ ]  0 . We assume that

1
Pˆnn '  (1̂  σˆ n  σˆ n' ) (Dirac exchange interaction)
2

2 1 2 1 N ( N  1) 1
Oˆ  
N ( N  1) n  n ' 2
(1̂  σˆ n  σˆ n ' )  [
N ( N  1) 2 2
1̂   σˆ n  σˆ n' ]
2 n n '

or
Identical particles 11 11/10/2019
1 2
Oˆ  [1̂ 
2
 σˆ n  σˆ n ' ]
N ( N  1) n  n '

Using the relation,

Sˆ 2 3 N 1
2
 1̂   (σˆ n  σˆ n ' ) ,
ℏ 4 2 nn'

we get

1 2 2 3N 1 N 4 4 1 ˆ2
Oˆ  [1̂  ( 2 Sˆ 2  )]  [ 1̂  S ]
2 N ( N  1) ℏ 2 2 ( N  1) N ( N  1) ℏ 2

[Sˆ 2 , Oˆ ]  0 . When the eigenvalue of Sˆ 2 is given by ℏ 2 S ( S  1) , the eigenvalue of Ô is equal to

1 N  4 4 S ( S  1)
 [  ]
2 ( N  1) N ( N  1)

The eigenvalue of Ô , (  ) , specifies the symmetry.

(i) For N = 2,

D1/2 x D1/2 = D1 + D0

1
  [2  2 S (S  1)]  1  S ( S  1)
2

When S = 1,  = 1 (symmetric).
When S = 0,  = -1 (anti-symmetric).

(ii) For N = 3,

D1/2 x D1/2 x D1/2 = D3/2 + 2 D1/2

1 1 2
  [  S ( S  1)]
2 2 3

When S = 3/2,  = 1 (symmetric).


When S = 1/2,  = 0.

(iii) For N = 4,

D1/2 x D1/2 x D1/2 x D1/2 = D2 + 3D1 + 2D0

Identical particles 12 11/10/2019


S ( S  1)

6

For S = 2,  = 1 (symmetric).
For S = 1,  = 1/3.
For S = 0,  = 0.

9. Summary: fermion and boson


The Hamiltonian H of the system (in the absence of a magnetic field) does not contain the spin operators, and
hence, when it is applied to the wave function, it has no effect on the spin variables. The wave function of the
system of particles can be written in the form of product,

   space  spin ,

where  space depends only on the coordinate of the particles and  spin only on their spins.    space  spin
should be anti-symmetric since electrons are fermions.
The system containing N identical particles are either totally symmetrical under the interchange of any pair
(boson), or totally antisymmetrical under the interchange of any pair (fermion).

Pˆij  N , B   N , B ,

Pˆij  N , F    N , F ,

where  N , B is the eigenket of N identical boson systems and  N , F is the eigenket of N identical fermion
systms.

((Note)) It is an empirical fact that a mixed symmetry does not occur.

Even more remarkable is that there is a connection between the spin of a particle and the statistics obeyed by it:

Half-integer spin particles are fermion, obeying the Fermi-Dirac statistic, while integer-spin particles are
bosons, obeying the Bose-Einstein statistics.

10. Ground state for systems with many electrons


We construct the form of wavefunctions for the system with two, three, four, and five electrons (identical
particles, fermion) using the Slater determinant. We assume that each electron has spin states

   ,    .

We also assume the orbital state: a  n  1 b  n  2 , c  n  3 .., where n is the principal quantum
number. It is reasonable to consider that the ground state can be expressed by electrons occupied by

a  , a 

Identical particles 13 11/10/2019


taking into account of the Pauli’s exclusion principle. For the system with three electrons, the ground state can
be expressed by

a  , a  , b 

For the system with the four electrons, the ground state can be expressed by

a  , a  , b  , b 

For the system with the five electrons, the ground state can be expressed by

a  , a  , b  , b  ,` c 

(a) Two particles state


For the two states,

a  a  , and a  a 

we have the Slater determinant

1 a11 a11 1
   a1a2 (1 2  1 2 )
2 a2 2 a2  2 2

where index 1, 2 denotes the particle 1 and particle 2. Clearly, the state vector is antisymmetric under the
exchange of two particles.

(b) Three particles state


For the three states,

a  a  , a  a  , and b  b 

we have the Slater determinant

a11 a11 b11


1
  a2 2 a2  2 b2 2
6
a3 3 a3 3 b3 3
1
 [a2 a3 ( 2  3   2 3 )b11  a1a3 (1 3  13 )b2 2  a1a2 (1 2  1 2 )b3 3 ]
6

This state vector is antisymmetric under the exchange between any two particles (such that 1  2 ).
Identical particles 14 11/10/2019
(c) Four particle state
For the four states,

a  a  , a  a  , b  b  , and b  b 

we have the Slater determinant

a11 a11 b11 b11


1 a2 2 a2  2 b2 2 b2  2
 
24 a3 3 a3  3 b3 3 b3  3
a4 4 a4  4 b4 4 b4 4
 b1b2 (1 2  1 2 )a3a4 ( 3  4   3 4 )
 b2b3 ( 2  3   2 3 )a1a4 (1 4  1 4 )
 b2b4 ( 2  4   2 4 )a1a4 (1 4  1 4 )
 b1b3 (1 3  1 3 )a2 a4 ( 2  4   2 4 )
 b1b4 (1 4  1 4 )a2 a3 ( 2  3   2 3 )
 b3b4 ( 3 4   3 4 )a1a2 (1 2  1 2 )

This state vector is antisymmetric under the exchange between any two particles (such that 1  2 ), but
symmetric under the two types exchange (such that 1  2 and 3  4 )

(d) Five particle state


For the five states

a  a  , a  a  , b  b  , b  b  , c  c 

we have the Slater determinant

a11 a11 b11 b11 c11


a2 2 a2  2 b2 2 b2  2 c2 2
1
  a3 3 a3 3 b3 3 b3 3 c3 3
120
a4 4 a4  4 b4 4 b4  4 c4 4
a5 5 a5 5 b5 5 b5  5 c5 5
1
 [a1a2 (1 2  1 2 )b3b4 ( 3  4   4 3 )c5 5  ...]
120

((Mathematica))

Identical particles 15 11/10/2019


a1 1 a1 1
Clear "Global` " ; A1 ;
a2 2 a2 2
Det A1 FullSimplify
a1 a2 2 1 1 2

a1 1 a1 1 b1 1
B1 a2 2 a2 2 b2 2 ;
a3 3 a3 3 b3 3
Det B1 FullSimplify
a1 a3 b2 a2 b3 2 3 1
a2 a3 b1 a1 b3 1 3 2 a3 a2 b1 a1 b2 1 2 3

a1 1 a1 1 b1 1 b1 1
a2 2 a2 2 b2 2 b2 2
C1 ;
a3 3 a3 3 b3 3 b3 3
a4 4 a4 4 b4 4 b4 4
Det C1 FullSimplify
b2 a3 a4 b1 2 1 1 2 4 3 3 4
a1 a4 b3 3 2 2 3 4 1 1 4
a3 b4 3 1 1 3 4 2 2 4
a2 a4 b1 b3 3 1 1 3 4 2 2 4
b4 a3 b1 3 2 2 3 4 1 1 4
a1 b3 2 1 1 2 4 3 3 4

________________________________________________________________________
REFERENCES
S. Tomonaga Angular momentum and spin (Misuzo Syobo, Tokyo, 1989) [in Japanese].
J.J. Sakurai Modern Quantum Mechanics, Revised Edition (Addison-Wesley, Reading Massachusetts, 1994).
L.D. Landau and E.M. Lifshitz, Quantum Mechanics (Pergamon Press, Oxford, 1977).
John S. Townsend , A Modern Approach to Quantum Mechanics, second edition (University Science Books,
2012).
David H. McIntyre, Quantum Mechanics A Paradigms Approach (Pearson Education, Inc., 2012).

APPENDIX: r
((Townsend textbook))

Identical particles 16 11/10/2019


1. Formula-1
For a symmetric spatial state under the exchange of two particles, we have

 S   dr ' dr" r ' , r"; S r ' , r"; S  S


  dr ' dr "Sˆ r ' , r " r ' , r " Sˆ  S
  dr ' dr" r ' , r " r ' , r "  S

and for an anti-symmetric spatial state under the exchange of two particles, we have

 A   dr ' dr " r ' , r"; A r ' , r "; A  A


  dr ' dr "Aˆ r ' , r " r ' , r " Aˆ  A
  dr ' dr " r ' , r " r ' , r "  A

where

r ' , r"  r ' 1 r" 2

1 1
Sˆ  (1̂  Pˆ12 ) , Aˆ  (1̂  Pˆ12 )
2 2

((Proof))

 S   dr ' dr" r ' , r"; S r ' , r"; S  S


  dr ' dr "Sˆ r ' , r " r ' , r " Sˆ  S
1
  dr ' dr " (1̂  Pˆ12 ) r ' , r " r ' , r "  S
2
1
  dr ' dr "(1̂  Pˆ12 ) r ' , r " r ' , r "  S
2
1
  dr ' dr "( r ' , r "  r " , r ' ) r ' , r "  S
2
1
  dr ' dr "( r ' , r " r ' , r "  S  r ", r ' r ' , r " Pˆ12  S
2
1
  dr ' dr "( r ' , r " r ' , r "  S  r ", r ' r ", r '  S )
2
  dr ' dr "( r ' , r " r ' , r "  S

where

Sˆ 2   Sˆ  S  Sˆ    S

Sˆ  S   S ,
Identical particles 17 11/10/2019
r ' , r"  S  r ' , r" Pˆ12  S  r", r '  S .

Pˆ12  S   S

Similarly,

 A   dr ' dr"Aˆ r ' , r" ; A r ' , r" ; A  A


  dr ' dr "Aˆ r ' , r " r ' , r " Aˆ  A
1
2 
 dr ' dr "(1̂  Pˆ12 ) r ' , r " r ' , r "  A

1
2 
 dr ' dr "( r ' , r "  r " , r ' r ' , r "  A

1
2 
 dr ' dr "( r ' , r " r ' , r "  A  r " , r ' r ' , r "  A )

1
2 
 dr ' dr "( r ' , r " r ' , r "  A  r " , r ' r ' , r " Pˆ12  A

1
2 
 dr ' dr "( r ' , r " r ' , r "  A  r " , r ' r " , r '  A )

  dr '  dr " r ' , r " r ' , r "  A

where

Aˆ 2   Aˆ  A  Aˆ    A

r ' , r"  A   r ' , r" Pˆ12  A   r" , r '  A

Pˆ12  A    A

2. Formula-2: symmetric case

1
k ' , k "; S  (1̂  Pˆ12 ) k ' 1 k " 2  Sˆ k ' , k "
2

1
x ' , x"; S  (1  Pˆ12 ) x ' 1 x" 2  Sˆ x' , x"
2

Identical particles 18 11/10/2019


x' , x"; S k ' , k " ; S  x ' , x" Sˆ  k ' , k " ; S
 x ' , x" Sˆ k ' , k " ; S
 x ' , x" k ' , k "; S
1 1
 [ x' k ' x" k "  x' k " x" k ' ]
2 2

or

2 x' , x"; S k ' , k " ; S   S ( x ' , x" )


1
 [ x' k ' x" k "  x ' k " x" k ' ]
2

3. Formula-3: antisymmetric case

1
k ' , k "; A  (1̂  Pˆ12 )[ k ' 1 k " 2  Aˆ k ' , k "
2

x' , x"; A k ' , k "; A  x' , x" Aˆ  k ' , k "; A


 x' , x" Aˆ k ' , k "; A
 x' , x" Aˆ 2 k ' , k
 x' , x" Aˆ k ' , k
 x' , x" k ' , k "; A
1
 [ x ' k ' x" k "  x ' k " x" k ' ]
2
1 1
 [ x' k ' x" k "  x' k " x" k ' ]
2 2

or

2 x' , x"; A k ' , k "; A   A ( x ' , x" )


1 x' k ' x' k "

2 x" k ' x" k "

4. Inner product for the symmetric and antisymmetric cases


We consider the inner product for the symmetric case

Identical particles 19 11/10/2019


x' , x"; S y ' , y" ; S  x ' x" Sˆ  y ' , y" ; S
 x ' x" Sˆ y ' , y" ; S
 x ' x" Sˆ 2 y ' , y"
 x ' x" Sˆ y ' , y"
 x ' x" y ' , y"; S
1
 [ x ' y ' x" y"  x ' y" x" y ' ]
2
1
 [ ( x' y ' ) ( x" y" )   ( x' y" ) ( x" y ' )]
2

where

Sˆ   Sˆ (Hermitian operator)

Sˆ 2  Sˆ

y ' , y"; S  Sˆ y ' , y"

We also consider the inner product for the antisymmetric case,

x' , x"; A y ' , y"; A  x ' x" Aˆ  y ' , y"; A


 x ' x" Aˆ y ' , y"; A
 x ' x" Aˆ 2 y ' , y"
 x ' x" Aˆ y ' , y"
 x ' x" y ' , y"; A
1
 [ x' y ' x" y"  x' y" x" y ' ]
2
1
 [ ( x' y ' ) ( x" y" )   ( x ' y" ) ( x" y ' )]
2

where

Aˆ   Aˆ (Hermitian operator)

Aˆ 2  Aˆ

y ' , y"; A  Aˆ y ' , y"

5. Closure relation

(i) Symmetric case

Identical particles 20 11/10/2019


 drdr ' r ', r"; S r ' , r"; S  1̂ (completeness)

((Proof))

 dr ' dr" s' , s"; S r ' , r"; S r ' , r "; S  S


1
  dr ' dr" [ ( s' r ' ) ( s" r" )   ( s' r" ) ( s" r ' )] r ' , r" ; S  S
2
1
 [ s' , s" ; S  S  s" , s ' ; S  S
2
 s' , s"; S  S

since

1
s ' , s"; S r ' , r "; S  [ ( s ' r ' ) ( s" r " )   ( s ' r " ) (r " r ' )]
2

s' , s"; S  S  s' , s" Sˆ 2   s ' , s" Sˆ   s' , s"  S

s", s' ; S  S  s", s'  S  s' , s" Pˆ12  S  s' , s"  S

or

s ' , s"; S  S  s", s '; S  S

(ii) Antisymmetric case

 drdr ' r ' , r"; A r ' , r" ; A  1̂ (completeness)

((Proof))

 dr ' dr" s' , s"; A r ' , r"; A r ' , r "; A  S


1
  dr ' dr" [ ( s' r ' ) ( s" r " )   ( s' r" ) (r" r ' )] r ' , r"; A  A
2
1
 [ s' , s" ; A  A  s" , s' ; A  A ]
2
 s' , s" ; A  S

since

1
s ' , s"; A r ' , r "; A  [ ( s ' r ' ) ( s" r " )   ( s ' r " ) (r " r ' )]
2

Identical particles 21 11/10/2019


s' , s"; A  A  s' , s" Aˆ 2   s' , s" Aˆ   s' , s"  A

s", s'; A  A  s" , s'  A  s' , s" Pˆ12  A   s' , s"  A

or

s", s '; A  A   s ' , s"; A  A

6. Wave function representation (symmetric case)


We have the relation for the symmetric case, such that

1
x' , x"; S  S  [ x ' , x"  S  x" , x'  S
2

and

x' , x"; S  S  x' , x" Sˆ 2   x' , x" Sˆ   x' , x"  S

Noting that

x' , x"; S  S  x' , x"  S  x" , x' Pˆ12  S  x", x '  S

Pˆ12  S   S

we get

x' , x"; S  S  x' , x"  S  x" , x '  S

7. Wave function representation (anti-symmetric case)


We have the relation for the anti-symmetric case, such that

1
x' , x"; A  A  [ x' , x"  A  x" , x '  A
2

and

x' , x"; A  A  x' , x" Aˆ 2   x ' , x" Aˆ   x ' , x"  A

Noting tht

x' , x"; A  A  x ' , x"  A  x" , x' Pˆ12  A   x" , x'  A

Pˆ12  A    A

Identical particles 22 11/10/2019


we have

x' , x"; A  A  x' , x"  A  x", x '  A

REFERENCES
R. Shankar, Principles of Quantum Mechanics (Kluwer Academic, 1980).
J.S.. Townsend, A Modern Approach to Quantum Mechanics, second edition (University Science Book, 2012).

Identical particles 23 11/10/2019

You might also like