You are on page 1of 31

Permutation Symmetry

and Identical Particles

Until now, our focus has largely been on the study of quantum mechanics of individual
particles.
However, most physical systems involve interaction of many (about 1023!) particles, e.g.
electrons in a solid, atoms in a gas, etc.

In classical mechanics, particles are always distinguishable


In quantum mechanics, particles can be identical and indistinguishable, e.g. electrons
in an atom or a metal.
The intrinsic uncertainty in position and momentum therefore demands separate consideration
of distinguishable and indistinguishable quantum particles.
Here we dene the quantum mechanics of many-particle systems, and address a few implications of particle indistinguishability.

Two particle systems


A general two-particle wave function with a particular energy can be written as


iEt /
(r1 , r2 , t ) (r1 , r2 )e
The Hamiltonian for whole system is therefore,


2 2 2 2
H
1
2 V (r1 , r2 , t )
2m1
2m 2


2
3 3
The normalized two-particle wavefunction
has probability | (r1 , r2 , t ) | d r1 d r2
3

of finding particle 1 in the volume d r1 and particle 2 in the volume d 3 r2


2 3 3
| (r1 , r2 , t ) | d r1d r2 1

is the normalization condition

If particle 1 is in the one particle state a (r ) and particle 2 in the state b (r )

then

(r1 , r2 ) a (r1 ) b (r2 )

However, this is possible only if we can distinguish particle 1 and 2. In quantum mechanics
two identical particles are indistinguishable so it is impossible to assign each particle to a
particular state. Instead, there are two ways of writing the wavefunction, not assigning each
particle to a particular state, namely:

(r1 , r2 ) A[ a (r1 ) b (r2 ) b (r1 ) a (r2 )]

The theory admits two kinds of identical particles: bosons, for which we use the plus sign
and fermions, for which we use the minus sign. It turns out that
All particles with integer spin are bosons, and
All particles with half-integer spin are fermions
The two-particle wavefunction makes sense only if

2
2


i
| (r1 , r2 ) | | (r2 , r1 ) | (r1 , r2 ) e (r2 , r1 )

If we introduce exchange operator


Since

P (r1 , r2 ) (r2 , r1 )

e 2i 1 0 or

It follows that the eigenvalues of P are 1. Hence

(r1 , r2 ) (r2 , r1 )

Symmetrization requirement

Particles with half-integer spin are fermions and their wavefunction must be antisymmetric
under particle exchange e.g. electron, positron, neutron, proton, quarks, muons, etc.
Particles with integer spin (including zero) are bosons and their wavefunction must be
symmetric under particle exchange e.g. pion, kaon, photon, gluon, etc.
Observe that for fermions

(r1 , r2 ) A[ a (r1 ) b (r2 ) b (r1 ) a (r2 )] 0, if a b


In other words, two identical fermions cannot occupy the same state. This is the famous
Pauli exclusion principle

Example:
Consider two non-interacting particles in an infinite potential well. As we know, the one
particle states are:
2
nx
2 2
2
n ( x)
sin(
), E n n K , where K
a
a
2ma 2
Ignoring the spins, find the ground state for (a) two distinguishable particles
(b) two identical bosons (c) two identical fermions
2 x1 x 2
2
2
sin
sin
and EG E11 (1 1 ) K 2 K
a a a
x
x
b) For two identical bosons 11 2 sin 1 sin 2 and EG E11 (12 12 ) K 2 K
a a a
c) For two identical fermions there is no state with energy E11
1
1 ( x1 ) 2 ( x2 ) 2 ( x1 ) 1 ( x2 ) 1 2 sin( ax1 ) sin( 22x2 ) sin( 2ax1 ) sin( ax2 )

2
2a
2

sin( ax1 ) sin( 22x2 ) sin( 2ax1 ) sin( ax2 )


a

a) 11

EG E12 (12 2 2 ) K 5 K

Now consider noninteracting two spin-1/2 particles. What is their ground state and energy?
For a general state, total wavefunction for two spin-1/2 particles is of the form:

( x1 , s1 ; x 2 , s 2 ) ( x1 , x 2 ) ( s1 , s 2 )
As we saw in the last lecture, the total spin of two spin-1/2 particles can be either the
singlet or triplet state, with

is antisymmetric under particle interchange.

they are symmetric under particle interchange.


For two electrons, total wavefunction, , must be antisymmetric under particle exchange.
i.e. spin singlet state must have symmetric spatial wavefunction; spin triplet states have
antisymmetric spatial wavefunction.

For the ground state, the symmetric spin triplet states are excluded because the required
antisymmetric spatial state is identically zero. The ground state of two spin-1/2 fermions
is therefore

singlet

What about the excited states?


The triplet state has the form

1
2

antisymmetric

symmetric

n1 ( x1 ) n2 ( x2 ) n2 ( x1 ) n1 ( x2 )1 , where 1 is defined as before

So T could be any of the three configurations in 1 with energy E12 = E1 + E2 = 5K


The singlet state is

symmetric

antisymmetric

S n 1 ( x1 ) n 2 ( x 2 ) n 2 ( x1 ) n 1 ( x 2 ) 0 , where 0 is defined as before


with energy E12 = 5K

To summarize the previous examples; complete quantum state vector then has the form

For bosons
The two parts must have the same exchange symmetry,
or else the full eigenstate would not be symmetric.

For fermions
The two parts cannot have the same exchange symmetry,
or else the full eigenstate would not be antisymmetric.

To construct wavefunctions for three or more fermions, let us suppose that they do not interact,
and are conned by a spin-independent potential,

H H S ( pi , ri ), where H S is the single particle Hamiltonian


i

p2
HS
V (r )
2m

Eigenfunctions of Schrodinger equation involve products of states of single-particle


Hamiltonian, HS
However, simple products
do not have required
antisymmetry under exchange of any two particles.
We could achieve antisymmetrization for particles 1 and 2 by subtracting the same product
with 1 and 2 interchanged,

However, wavefunction must be antisymmetrized under all possible exchanges. So, for 3
particles, we must add together all 3! permutations of 1, 2, 3 in the state a, b, c with factor
-1 for each particle exchange.

Such a sum is known as a Slater determinant:

and can be generalized to N particles as:

Antisymmetry of wavefunction under particle exchange follows from antisymmetry of Slater


determinant,
The determinant is non-vanishing only if all three states a, b,c are different manifestation
of Paulis exclusion principle: two identical fermions can not occupy the same state.
The conditions on wavefunction antisymmetry imply spin-dependent correlations even where
the Hamiltonian is spin-independent, and leads to numerous physical manifestations...

Example : Hydrogen H2 gas


With two spin-1/2 proton degrees of freedom, H2 can adopt a spin singlet (parahydrogen)
or spin triplet (orthohydrogen) wavefunction
The space part is described by rotational modes with orbital quantum numbers l = 0, 1, 2, . . .
Orthohydrogen: Nuclear spin part is symmetric (spin triplet) and
nuclear space part is antisymmetric (odd l).
Parahydrogen: Nuclear spin part is antisymmetric (spin singlet)
and nuclear space part is symmetric (even l).
Moment of inertia of molecule

These two forms have different sets of rotational energies and hence different specific heats
H2 gas at equilibrium contains 75% orthohydrogen and 25%
parahydrogen. The 3:1 ratio is a reection of the nuclear spin
degeneracy.

Another implication of the symmetrization of the wavefunctions is the separation between


the particles.
If we ignore spin, and consider just the spatial wavefunction
with
for bosons, and
for fermions, we can calculate the expectation value of the square of the separation between
two particles.

Case I: Distinguishable Particles

Similarly

and

where we write
to indicate that we just have the x2 expectation value w.r.t. the state .
. Our distinguishable particles are then, on average,
apart

Case II: Indistinguishable particles


Here, we have

The relevant expectation values are

and provided

then orthonormality gives

and the above reduces to


since there is no distinguishing between the particles.
But

where the last term is just the obvious

If we compare with the distinguishable case we see that the difference resides in the
final term

So for the bosonic (upper signs) case, the particles are slightly closer together than distin
guishable particles, and for fermions (lower signs) case, they are slightly further apart.

Again, we see that the cross term vanishes if the wave functions do not significantly overlap,
another indication that localization (i.e. a classical interpretation) works at some level.
The force is fictitious in a classical sense, but we can think of a bosonic attractive force
pulling two particles closer together, and a fermionic repulsive force tending to push two
particles apart. This is the so-called exchange force.

Total wave function has to be symmetric or antisymmetric, we have to put together complete
two-electron state including the spin:

Thus, a symmetric spin state (e.g., a triplet) must be associated with an antisymmetric
spatial wave function (an antibonding combination), while an antisymmetric spin state
(e.g., a singlet) must be associated with a symmetric spatial w.f. (a bonding combination).
Sure enough, the chemists tell us that the covalent bonding requires the two electrons to
occupy the singlet state, with total spin zero.

Schematic picture of the covalent bond: (a) symmetric combination produces attractive
force; (b) antisymmetric combination produces repulsive force.

Atoms

Electron-electron interaction

For a neutral atom,

For fermions, the spatial part of the wave function must satisfy
complete wave function

and the

must be antisymmetric with respect to interchange of any two particles.


For hydrogen, Z = 1, so there is no contribution from the electron-electron interaction term.
is then simply a one-electron wave function.

Helium atom
For helium, Z = 2, leading to a single electron-electron interaction term. This is enough to turn
the proper solution of the Schrodinger equation into a true many-body wave function.
The Hamiltonian now reads

If we ignore the e-e interaction for the moment, then the spatial
part of the wavefunction can be approximated as the product of
two one electron hydrogen wave functions:

where the Bohr radius is half as large as for hydrogen and

Specifically, the ground state is

E0 = 8(-13.6) = -109 eV

Because 0 is a symmetric function, the spin wave function must be antisymmetric. Thus, the
ground state of He is a singlet configuration, wherein the spins are aligned oppositely.
The actual ground state is indeed a singlet, but the energy is only -79 eV. Such poor agreement
is expected since we neglected the positive (repulsive) contribution of e-e interaction
The excited states of He consist of one electron in the hydrogenic ground state and the other in
an excited state:
The spatial portion of this wave function can be constructed either symmetrically or
antisymmetrically
leading to the possibility of either antisymmetric or symmetric spin portions, respectively,
known as parahelium and orthohelium.
If you try to put both electrons in excited states, one immediately drops into the ground state,
releasing enough energy to knock the other electron into the continuum and yielding a He ion
(He+). This process is known as an Auger transition.

Energy level diagram for He (relative to He+, -54.4 eV).


Note that parahelium (antisymmetric spin) energies are
uniformly higher than orthohelium counterparts.

Quantum Mechanical Statistics (brief)

So far, we have been dealing with the ground state of systems. Statistical mechanics deals
with the occupation of states when a system is excited. The fundamental assumption of
statistical mechanics is that, in thermal equilibrium, every distinct state with the same
total energy is occupied with equal probability. Temperature is simply a measure of the total
energy of a system in thermal equilibrium.
The only change from classical statistical mechanics occasioned by quantum mechanics
has to do with how we count distinct states, which depends on whether the particles
involved are distinguishable, identical fermions, or identical bosons.

Simple example: Consider a system of 2 identical particles for which 3 single-particle states are
available. If, as in classical mechanics, the particles are considered distinguishable, they may be
labeled as A and B. On the other hand, both particles must have the same label, say A, if they
are considered indistinguishable. The tables below enumerate all distinct states for distinguisha
ble classical particles or for indistinguishable fermions and bosons.

Classically, 32 = 9 distinct microstates are


available, but only 3 are available to fermions
or 6 to bosons.

So, the results vary dramatically depending on the fundamental nature of the particles.

Quantum statistical mechanics

One particle

( x)

Two particles

N particles

( x1 , x2 )

( x1 , x2 , x3.....xN )

Quantum statistical mechanics


N particles

Thermal equilibrium, T

Quantization of the energy


for individual particles

( E1 , E2 , E3 ,...)
Total energy:

E N1E1 N2 E2 N3 E3 ...

N N1 N2 N3 ...

Quantum statistical mechanics


3 kind of particles

Distinguishable particle

Identical fermions
Antisymmetric state
Identical bosons
Symmetric state

For each type of particles:


List all the possible configurations
Determine the number of combinations of each configuration
Determine the probability of each energy for a given particle
Generally for an arbitrary potential, for which the one-particle energies are E1, E2, E3with
degeneracies d1, d2, d3 If we put N particles into this potential, the number of distinct states
corresponding to the configuration (N1, N2, N3,) is Q, which depends strongly on whether
the particles are distinguishable, identical fermions, or identical bosons.

Distinguishable particle

Identical fermions

d nNn
Qdis N1 , N 2 , N3 ,.. N !
n 1 N n !

Q fermions N1 , N 2 , N3 ,..
n 1

dn !
N n ! d n N n !

Identical bosons

Qbosons N1 , N 2 , N3 ,..
n 1

dn 1 !

N n ! d n 1!

In thermal equilibrium, the most probable configuration is that one which can be attained
in the largest number of different ways. We want to find this subject to two constraints:

This is best handled using Lagrange multipliers. It is also useful to maximize ln (Q) rather
than Q, which turns the products of factorials into sums. Let

where and b are the Lagrange multipliers. We now seek the values of Nn which satisfy
The most probable occupation numbers are:
Distinguishable particles:
Identical fermions:
Identical bosons:

Physical significance of and b


Insight into the physical meaning of and b can be gained by substituting one of the
above equations for Nn into the sums which yield N and E, but that requires assuming
a specific model for V. Let's use a simple V, a three-dimensional infinite square well:

n x n y n z
2 2
Ek
k , where k
,
,

2m
lx l y lz

For distinguishable particles,

The second equation leads us to define T by


is customarily expressed in term of a chemical potential:
We can now write expressions for the most probable number of particles in a particular
one-particle state with energy e as:

Distinguishable particles (Maxwell-Boltzmann):


Identical fermions (Fermi-Dirac):

Identical bosons (Bose-Einstein):

The Fermi-Dirac distribution for T = 0 and for T somewhat above zero.

You might also like