Sunteți pe pagina 1din 12

Lecture 1: Identical particles,

introduction (9/20/2005)
The main unifying theme of this course will be time-dependence (and dynamics), but we will start with identical particles.
In classical physics, particles are distinguishable. Even if two electrons
happen to have exactly the same mass, charge, and other quantities, you
may trace them and distinguish which electron is Peter and which electron
is Paul.
However, particles in quantum mechanics dont have any ID. If two electrons scatter, it is impossible (even in principle) to decide which of them is
Peter and which is Paul. Actually, both possibilities contribute to the final
wavefunction that decides about the probabilities of various results. Draw a
picture of two possible histories of electrons that scatter elastically; one cant
say which of them occured.

Recommended reading: Griffiths ch. 5.1, Gasiorowicz ch. 20.

Symmetrizing the multi-particle wavefunctions


Recall that in quantum mechanics, two particles are not generally described
by two independent waves (~x) and (~y ) but rather by a wavefunction in
six dimensions
(~r1 , ~r2 )
that determines the probability for the first one to be in the volume dV1
around ~r1 and the second to be in a small volume dV2 around ~r2 to be
dP = dV1 dV2 |(~r1 , ~r2 )|2 .
Similar statements apply for many particles, too. The wavefunction becomes
a function of 3k coordinates for k particles. Imagine a quantum mechanical
system for two electrons in an external field. The Hamiltonian is gonna be
H=

p21
p2
+ 1 + V (~r1 ) + V (~r2 ) + Vrel (~r1 ~r2 )
2m 2m
1

where p2i =
h2 2i enters the kinetic energy, V (~r) is the potential energy
between an electron and the nucleus, and Vrel (~r1 ~r2 ) is the mutual interactions of the electrons. Obviously it looks nicely symmetric in both electrons.
What do we mean mathematically? Define the exchange operator P12 that
exchanges the electrons. It means that it acts as follows:
(P12 )(~r1 , ~r2 ) = (~r2 , ~r1 ).
If the electrons wavefunction depended on other quantities, such as their
spin, the spins would have to be exchanged on the RHS, too. Note that
2
P12
= 1; it squares to the identity operator:
2
(P12
)(~r1 , ~r2 ) = (P12 (P12 ))(~r1 , ~r2 ) = (P12 )(~r2 , ~r1 ) = 1 (~r1 , ~r2 ).

Now its important to realize that the exchange operator commutes with
the Hamiltonian:
[P12 , H] P12 H H P12 = 0.

It does not matter whether you first act with the Hamiltonian and then you
switch the dependence of the resulting wavefunction on ~r1 , ~r2 , or you first
exchange the electrons and then act with the Hamiltonian. Youll get the
same results. Fine. So now, because these two operators commute, we can
simultaneously diagonalize them. Because one of the two operators is the
Hamiltonian that determines time evolution, the eigenvalues of P12 will be
conserved.
More concretely, it is useful to consider eigenstates of P12 . Because weve
proved that the exchange operator squares to 1, the same thing holds for the
eigenvalues. If we find an eigenfunction (lets start to omit some awkward
parentheses) i.e. if
P12 (~r1 , ~r2 ) = p(~r1 , ~r2 ),
then also

2
P12
(~r1 , ~r2 ) = p2 (~r1 , ~r2 ).

But we know that the latter equals (~r1 , ~r2 ) which means that p2 = 1;
the eigenvalues must satisfy the same algebraic equation as the operator.
Obviously this is solved by p = 1. How do the eigenfunctions look like?
p = +1 :

(~r1 , ~r2 ) = +(~r2 , ~r1 ),

p = 1 :

(~r1 , ~r2 ) = (~r2 , ~r1 ),

We say that the wavefunctions are symmetric in the first case, and antisymmetric in the second case.
2

Bosons and fermions


Now imagine 3, 4, or more electrons, and prepare them in the eigenstate of
P12 , P13 , and other exchange operators. If the electrons are indistinguishable,
all these eigenstates must be equal. So for any kind of particles, the eigenvalues of all exchange operators must be universally either +1 this is when
we call the particles bosons - or they must equal 1 which is the case of
fermions. Now it is natural to assume that the wavefunction of bosons must
be symmetric and the wavefunction of identical fermions must be antisymmetric - with respect to any exchange of a pair of particles. More general
wavefunctions, especially antisymmetric wavefunctions for bosons and symmetric for fermions, are simply forbidden by mother Nature.
This assertion actually follows from quantum field theory that describes
multiparticle states as action of creation operators for single-particle states
acting on the vacuum state called |0i. For example, in quantum field
theory, a state where one electron is at place X and another one is at Y is
written as
b (X)b (Y )|0i = b (Y )b (X)|0i
which follows from b (X)b (Y ) = b (Y )b (X). (In the real life, the dagger
usually kills. But in quantum field theory, it creates things.) Translated to
the language of ordinary wavefunctions this means our usual
(~r2 , ~r1 ) = (~r1 , ~r2 ).
For bosons there would be a plus sign.
Because these creation operators commute for bosons and anticommute
for fermions the required (anti)symmetry of the wavefunction follows. Also,
one can follow Wolfgang Pauli and prove the so-called spin-statistics relation
using quantum field theory:
the bosons, i.e. the particles with symmetric wavefunctions, p = +1,
must have integral spin j Z, such as 0, 1, 2, . . .; examples include
photons, W,Z-bosons, Higgs bosons, and bound states containing an
even number of fermions
the fermions, i.e. the particles with antisymmetric wavefunctions, p =
1, must have half-integral spin j Z + 1/2 such as 0.5, 1.5, 2.5 and so
on; this includes electrons (and neutrinos), quarks, protons, neutrons,
and bound states containing an odd number of other electrons
3

These relations between the spin and the symmetry of the wavefunction
follow from the assumption that the energy is bounded from below and the
norm of the states (and thus the probabilities) must be non-negative.
The wavefunction of two Helium atoms is easily seen to be symmetric
(p = +1) because the exchange of two Helium atoms (He42 ) is nothing else
than the exchange of two pairs of electrons, two pairs of neutrons, and two
pairs of protons. And p2e p2n p2p = +1 regardless of the individual signs. But
more generally, one needs quantum field theory to prove the spin-statistics
relation above.

Statistics, the Pauli exclusion principle, and the distribution


You know from high school chemistry that two electrons cant occupy exactly
the same state of an atom. This Pauli exclusion principle can easily proved
from antisymmetry of the fermionic wavefunctions:
(MA, MA) = (MA, MA)

(MA, MA) = 0.

We chose Massachusetts but it holds for any states; the wavefunction is


always antisymmetric whatever basis we choose for the Hilbert space. It is
antisymmetric in the coordinate representation; it is also antisymmetric in
the momentum representation (which can be trivially seen from the Fourier
transform);
p1 , p~2 ) = (~
p2 , p~1 )
(~

it is also antisymmetric if the states such as 1s(up), 1s(down), 2s(up), 2s(down)


etc. are chosen as the basis. The antisymmetry of the fermions wavefunctions is crucial for the diversity of atoms and the existence of things such as
life.
Bosons behave very differently than the fermions. They not only have no
difficulties to live in the same state, but they actually like it which is the
driving force behind the so-called Bose-Einstein condensation and behind the
coherence of LASERs (photons are bosons).
In classical physics, the thermal distribution of particles treats all of them
independently. The probability that a given particle appears in the state of
energy E decreases as the Maxwell-Boltzmann factor
p=

1
1
1
exp(E/kT ) =
Z
Z exp(E/kT ) + 0.
4

where Z is a normalization factor that must be chosen in such a way that


the probabilities add up to one. In quantum statistical physics, the particles
are indistinguishable and we must look at particular states, not particles. In
a given state, the number Nf of fermions of the same kind may be either 0
or 1. In another state, the number of bosons Nb of the same kind may be 0,
1, 2, 3, etc.
If the energy of the state is Ef , the energy of Nf fermions in this state is
Nf Ef ; similarly if you replace f by b. Different configurations described by
the number of bosons and fermions in various states are weighted again by
the Maxwell-Boltzmann factor.
Fermi-Dirac distribution
Consider the fermions. The number Nf for a given state is either 0 or 1. The
probability that there is exactly one fermion there must be exp(Ef /kT )
times the probability that there is no fermion. Because the probabilities add
up to one, it implies that
p(Nf = 0) =

1
,
exp(Ef /kT ) + 1

p(Nf = 1) =

exp(Ef /kT )
.
exp(Ef /kT ) + 1

What is the average number of particles in that state? It is obviously equal


to the probability that there is one particle:
hNf i =

1
exp(Ef /kT ) + 1

In Maxwell-Boltzmann statistics, the +1 term in the denominator was missing; the number of particles in a different state was simply proportional to
the probability distribution of each individual distinguishable particle.
Bose-Einstein distribution
Consider the bosons. The number Nb for a given state is either 0 or 1 or 2
or any non-negative integer. The probability that there are Nb bosons must
depend on Nb like exp(Nb Eb /kT ). But the sum of these numbers must add
up to one. You can see that the unique solution is
p(Nb = nb ) = exp(nb Eb /kT ) exp((nb + 1)Eb /kT )
5

What is the average number of bosons in the state?


hNb i =

p(Nb = nb )nb =

nb =0

1
exp(Eb /kT ) 1

The series can be calculated as a derivative of a geometric series with respect


to an extra parameter. The result may be more important: note that it has
a similar form like the Maxwell-Boltzmann or the Fermi-Dirac distribution,
but the relative sign in the denominator differs from the fermions. When
we add some volumes of the phase space (the density of states which gives
a power law factor 3 ) to the Bose-Einstein formula above, we obtain the
famous blackbody formula due to Planck (1900).
Note that for Ef , Eb  kT , the 1 term in the denominator may be
neglected (the exponential wins) and the quantum distributions reduce to the
classical distribution. However, for states whose energy Ef , Eb is comparable
to kT (or smaller), quantum mechanics and indistinguishability is always
important.

The Slater determinant


Let us return to the wavefunctions. Imagine that you want to construct a
multi-particle wavefunction for n identical particles that are as independent
as possible and that are described by the wavefunctions
1 (r),

2 (r),

3 (r)

...

n (r).

Our first guess is the product


(r1 , r2 , . . . rn ) = 1 (r1 )2 (r2 ) . . . n (rn ).
But because i are different functions, this does not have the right (anti)symmetry.
We must (anti)symmetrize the wavefunction to obtain
(r1 , r2 , . . . rn ) = A

permutations
X

p1 (r1 )p2 (r2 ) . . . pn (rn )

for the bosons or


(r1 , r2 , . . . rn ) = A

permutations
X

(1)p p1 (r1 )p2 (r2 ) . . . pn (rn )

where p is the sign of the permutation for fermions. The latter wavefunction
can also be written as the Slater determinant:

A det i (rj ) =
i,j

1 (r1 ) 2 (r1 )
1 (r2 ) 2 (r2 )
..
..
.
.
1 (rn ) 2 (rn )

n (r1 )
n (r2 )
..
..
.
.
n (rn )

The normalization constant A may be determined easily if all one-particle


R
wavefunctions i (r) are orthogonal, i.e. j (r)i (r) = 0 for i 6= j. Then the
norm of the multiparticle wavefunction is simply a sum of n! identical terms
(the mixed terms cancel by orthogonality) which tells us that
1
A= .
n!
Note that for two-particle states, the (anti)symmetrized wavefunction is as
simple as
1 (r1 )2 (r2 ) 2 (r1 )1 (r2 )

(r1 , r2 ) =
.
2

Exchange forces
Because two fermions cant occupy the same state, you may think that
they dont even want to be too close. And you would be right. Even
non-interacting fermions tend to be further apart than if they were distinguishable. And bosons tend to be closer together than non-interacting
distinguishable particles would be.
To see it, start with distinguishable particles whose wavefunction is
(x1 , x2 ) = a (x1 )a (x2 ).
Calculate
h(x1 x2 )2 i = hx21 i 2hx1 x2 i + hx22 i

It is easy to see that the result is simply

. . . = hx2 ia + hx2 ib 2hxia hxib .


For example,
hx21 i =

dx1 x21 |a (x1 )|2


7

dx2 |b (x2 )|2 = hx2 ia

because the second integral in the product equals one (normalization). The
same thing applies to hx2 ib . Also, the integral calculating 2hx1 x2 i factorizes
as indicated.
For identical particles we must change a couple of things, starting from
the wavefunction. The sign below refers to bosons or fermions, respectively.
1
(x1 , x2 ) = [a (x1 )b (x2 ) b (x1 )a (x2 )].
2
In this case,
hx21 i =

dx1 dx2 x21 | (x1 , x2 )|2

1
2Z
1
+
2
1Z

2Z
1

2
=

(1)

dx1 x21 |a (x1 )|2

dx2 |b (x2 )|2

(2)

dx1 x21 |b (x1 )|2

dx2 |a (x2 )|2

(3)

dx1 x21 a (x1 )b (x1 )

dx2 b (x2 )a (x2 )

(4)

dx1 x21 b (x1 )a (x1 )

dx2 a (x2 )b (x2 )

(5)

The last two terms (among the four) vanish because their latter integral
factor equals zero by orthogonality of a and b . The first two terms give
you
1
. . . = (hx2 ia + hx2 ib )
2
2
Obviously, the result for hx2 i is the same thing because of an a-b symmetry.
The last term in the expectation value of the distance, namely hx1 x2 i, requires
an extra calculation:
hx1 x2 i =
+

Z
1Z
2
dx1 x1 |a (x1 )|
dx2 x2 |b (x2 )|2
2Z
Z
1
2
dx1 x1 |b (x1 )|
dx2 x2 |a (x2 )|2
2Z
Z
1

dx1 x1 a (x1 )b (x1 ) dx2 x2 b (x2 )a (x1 )


2Z
Z
1

dx1 x1 b (x1 )a (x1 ) dx2 x2 a (x2 )b (x1 )


2
hxia hxib |hxiab |2

(6)
(7)
(8)
(9)
(10)

where we defined the so-called overlap integral


hxiab =

dx x a (x)b (x)

as the first factor in the third term (with x1 x, and it also appears in
the conjugated form, both in the third and the fourth term). This overlap
integral is only nonzero if the particles overlap in space; only in this case,
their indistinguishability affects the expectation value (and thus the force).
Otherwise we get back the previous result. The qualitative conclusion of
the calculated expectation values is, of course, that bosons tend to be closer
together while the fermions tend to be further apart.

Spins and binding energies of some nuclei and atoms


Hydrogen molecules and helium
Even a qualitative analysis of this result shows that the indistinguishability of
the particles matters and it can be shown to be responsible for the differences
between various superficially similar systems in atomic (electrons moving
around nuclei), molecular (electrons moving among many nuclei) and nuclear
(protons and neutrons moving inside nuclei) physics.
For example: Hydrogen is a gas that likes to form the H2 molecules. But
the helium and the hydrogen bind neither to He2 nor HeH. What causes
the difference? The molecule H2 is a bound state with a large (negative)
binding energy for the electrons. This bound is expected to be strong if the
electrons are concentrated in between the two nuclei.
How does it work? Imagine two electrons with the same spins (up-up).
Then the dependence of the wavefunction on the spins is symmetric, and the
dependence of the wavefunction on the position must be antisymmetric. The
electrons tend to be further apart, and they dont want to be concentrated
in the middle simultaneously: that causes anti-binding. Weve shown that
its the case when the total spin is S = 1.
But what about the total spin of the electrons equal to S = 0? In that
case, the dependence of the wavefunction on the spins is described by the
singlet state (singlet means that there exists a single polarization only; the
projection of the spin can only be zero because the spin is zero):
1
= [ | ia | ib | ia | ib ]
2
9

The Fermi statistics requires the total wavefunction that can be usefully
written in a factorized form as long as the Hamiltonian does not directly
depend on the spins (or if it at least splits to two Hamiltonians that depend
on the positions and spins separately)
(x1 , x2 ; s1 , x2 ) = (x1 , x2 )(s1 , s2 )
to be antisymmetric
(x1 , x2 ; s1 , x2 ) = (x2 , x1 ; s2 , x1 ).
But because the spin wavefunction is already antisymmetric in our singlet
state
(s1 , s2 ) = (s2 , s1 ),
the spatial wavefunction for the electrons must actually be symmetric
(x1 , x2 ) = +(x2 , x1 ).
This effectively causes the electrons to behave as bosons. Both of them
occupy the space in between the nuclei, and the bound of H2 becomes strong.
If at least one of the atoms is helium, it contains electrons both with spin
up and spin down (the helium is a singlet itself). Therefore the other atom
either H or He must have at least one electron that shares the same
spin state with an electron of the first helium atom. Thats enough of an
antibinding to destroy chances for a He2 or HeH bound state.
We need to perform a more careful analysis for the more complex atoms;
see Gasiorowicz.
Nuclei and isospin
Spin 1/2 particles can be spinning up or down. Later we will see that
they are spinors a representation of the SU(2) group. The same mathematics may be applied in other contexts with different interpretations than
the actual spinning in space.
Consider protons and neutrons. They have a different charge. But the
electric charge is not so important in the nuclei where the strong nuclear force
is much more significant. From this perspective, a proton and a neutron look
like twins; a doublet. They may be described as nucleons that only differ
by the isospin (isotopical spin): the proton is a nucleon with spin up while
10

the neutrons isospin points down (these directions have nothing to do with
the real space).
Why is there a deuteron - a bound state of one proton and one neutron
np (heavy Hydrogen nucleus) but no pp state (neutron-free helium) or nn
(double-neutron)? And why does the deuteron have spin one? The total
wavefunction of two nucleons may be written as
= space spin isospin
and it must be antisymmetric for nucleons that are fermions because of their
spin 1/2. The ground state of the deuteron tries to minimize the angular momentum, much like the ground state of the Hydrogen atom. This implies that
space is symmetric. This means that spin isospin must be antisymmetric.
Were left with two possibilities:
1. spin antisymmetric, isospin symmetric:
J = 0 : spin

| i| i | i| i

=
,
2

I = 1 : iso

|pi|pi

= (|pi|ni + |ni|pi)/ 2

|ni|ni

2. spin symmetric, isospin antisymmetric


J = 1 : spin =

| i| i

2 ,
(|
i|
i
+
|
i|
i)/

| i| i

I = 0 : iso =

|pi|ni |ni|pi

Because of more quantitative and difficult calculations that we cant do


here, nuclear physics prefers the second alternative as the only allowed bound
state. This implies that the particle must have one proton and one neutron
(deuteron) and its spin must be 1 (a triplet).

Atoms and Hunds rules


You may want to read Griffiths, ch. 5.2. We will only sketch the Hunds
rules here.
The atoms are described by the Hamiltonian for N electrons (kinetic plus
potential energy for the nucleus) plus their pairwise interactions.
11

We can solve a simpler system in which the electrons interactions with


each others are neglected. In that case, one obtains many copies of the
electrons in the Hydrogen eigenstates, described by n, l, m, sz where n =
1, 2, 3, . . . is the principal quantum number, l = 0, 1, 2, . . . n 1 is the orbital
quantum number, m = l, l + 1, . . . , l is its z-projection, and sz = 1/2 is
the spin. The Pauli principle says that we can only put one electron to each
state. The electrons energy would only depend on their principal quantum
numbers. In reality, the self-interactions rearrange the shells a bit, so that
the most bound ones are (in this order)
1s, 2s, 2p, 3s, 3p, 4s, 3d, . . .
where the letters s, p, d, f, g, h, i . . . stand for sharp, principal, diffuse, fundamental, g-shaped, h-shaped, i-shaped, and so forth, and denote
l = 0, 1, 2, 3, 4, 5, 6 . . .
Moreover, the lowest energy eigenstates of the atoms may usually be determined by phenomenological Hunds rules.
couple the valence electrons (those in the highest partially occupied
shell) to give maximum total spin; for chemists: make as many spins
as possible parallel, compute the total sz = ms , and call it s (thats the
highest component of a multiplet)
given this choice, construct the maximum l compatible with the Pauli
principle; for chemists: try to obey the previous rule but fill in the
shells to obtain the maximum lz = ml and call this maximum value l
(again the highest component of the multiplet)
the total angular momentum j is j = l + s if the shell is more than
half-full, and j = l s otherwise

The explanation why two of these rules work is the following: the first rule
tries to make the spin wavefunction as symmetric as possible, which means
that the spatial wavefunction is antisymmetric as possible. Such an antisymmetry reduces the probability that the electrons are too close to each other,
and it reduces the positive energy contribution from their self-interactions.
The third rules is implied by a spin-orbital interaction proportional to l s
which is a linear function of j 2 for fixed l, s. The rule tries to minimize this
term if the shell is at most half-occupied; if it is more than half-full, it is
better to think in terms of holes for which the conclusion reverses.
12

S-ar putea să vă placă și