Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Page 5.

1
SECTION 5: TWO-PARTICLE SYSTEMS

1. Two-Particle Systems 2
1.1 Statement of the problem 2
1.2 The independent-particle approximation 2
2. Identical Particles and Exchange symmetry 3
2.1 Symmetry of the wave function 3
2.2 Exchange symmetry for two identical fermions with s = 1/2 5
3. Exchange forces 7
3.1 Two distinguishable particles 7
3.2 Two identical fermions or two identical bosons 8
3.3 Hydrogen molecule 9
4. The Helium Atom 10
4.1 Helium Ground State: Perturbation Theory 10
4.2 Helium Ground State: Variational Method 12
4.3 Excited states of helium 14











References
The original version of these notes was based on the textbooks by Mandl and Griffiths, although it has been
considerably modified by revision, and much detail has been added. References are given as footnotes to
sections that are closely based on other textbooks.

Original Version: July 1995; Last Major Revision: July 2003
Revision Date: 26 October 2010
Printed: 26 October 2010

Page 5.2
1. TWO-PARTICLE SYSTEMS
1.1 Statement of the problem
We wish to find the energy eigenvalues and eigenvectors of a quantum system of two particles. In principle,
this can be achieved by solving the Schrdinger equation

1 1 2 2 1 1 2 2
( , , , ) ( , , , ), H r s r s E r s r s =


where
1 1
, r s

are the position and spin coordinates of particle 1 and
2 2
, r s

are those of particle 2. The Ham-
iltonian will be of the form

2 2
2 2
1 2 1 1 2 2
1 2
( , , , ),
2 2
H V r s r s
m m
=


where the subscript on
1
2
indicates differentiation with respect to the coordinates of particle 1, and similarly
for
2
2
. The potential energy V will include terms that account for the effects of external forces on the two
particles and also terms representing the interaction between the two particles.
In general, because of the presence of the latter interaction, it is not possible to solve the Schrdinger
equation for this Hamiltonian without making some approximations.
1.2 The independent-particle approximation
A major simplification occurs if the interaction between the two particles is sufficiently weak that it can be
neglected; the potential energy then contains only terms representing the effect of external forces acting
separately on each particle. Then the Hamiltonian becomes

2 2
2 2
1 1 1 1 2 2 2 2
1 2
( , ) ( , ),
2 2
H V r s V r s
m m
=

(1)
where the potential V
1
in which particle 1 moves depends only on the coordinates of particle 1, and similarly
for V
2
. The Hamiltonian is thus the sum of two terms, each of which is the Hamiltonian of a single particle:

1 2
H H H = (2)
where

2
2
( , ).
2
i i i i i
i
H V r s
m
=

(3)
The single-particle Hamiltonians H
1
and H
2
operate on different coordinates and clearly commute with each
other and each therefore commutes with the total Hamiltonian H. Thus an eigenfunction of H can be si-
multaneously an eigenstate of both H
1
and H
2
, and it follows that we can write the two-particle wave function
as a product of single-particle wave functions of the two particles:

1 1 2 2 1 1 1 2 2 2
( , , , ) ( , ) ( , ), r s r s r s r s =

(4)
where
i
is an eigenstate of H
i
corresponding to the energy eigenvalue E
i
:
( , ) ( , ) ( 1, 2).
i i i i i i i i
H r s E r s i = =


Substitution of eqs.(2) and (3) into eq.(1) and use of eq.(4) yields
[ [ [ [
1 2 1 1 1 2 2 2 1 2 1 1 1 2 2 2
( , ) ( , ) ( , ) ( , ) H H r s r s E E r s r s =

(5)
Page 5.3
and we see that the energy of the system is

1 2
. E E E =
We conclude that, in the approximation in which the particles move completely independently of each other,
the problem reduces to solving the Schrdinger equation for a single particle, which is generally
much simpler;
the total energy is the sum of the single-particle energies and
the wave function is the product of two single-particle wave functions.
More generally, when the two particles do interact with each other, these conclusions are no longer valid.
Nonetheless, it often happens that the additional term in the Hamiltonian is sufficiently weak that it can
be treated using first-order perturbation theory.
Alternatively, the independent-particle model approximation can be used as a guide to constructing a trial
wave function in the variational method.
This will be illustrated in the discussion on helium below and in the subsequent description of many-electron
atoms.
2. IDENTICAL PARTICLES AND EXCHANGE SYMMETRY
1

2.1 Symmetry of the wave function
If the two particles are identical, e.g. both electrons or both protons, then in quantum mechanics the particles
are indistinguishable in principle. If the two particles are interchanged, there can be no measurement that can
detect the change. Consequently, all observables of the system must be symmetric in the particle labels 1 and
2. For example, the Hamiltonian must satisfy the condition
(1, 2) (2,1) H H =
where
1 1 2 2
( , , , ) H r s r s

has been written more concisely as H(1,2). In addition, the complete two-particle wave
function
1 1 2 2
(1, 2) ( , , , ) r s r s

must obey
(1, 2) (2,1) =
since swapping the two particles cannot alter the probability distribution. It follows that
(1, 2) (2,1). = (6)
In other words, for identical particles,

The complete two-particle wave function must be
either symmetric
or anti-symmetric
under interchange of the particle coordinates.
(7)

1
This section is based in part on sections 4.4 and 5.7 of the textbook of Mandl. It has subsequently been considerably modified and
expanded. See also section 5.1.1 of the book by Griffiths.
Page 5.4
It is an empirical fact
2
that whether the wave function of a particular system of identical particles is symmetric
or antisymmetric under particle exchange depends on the intrinsic spin of the particles. Specifically,
The wave function is anti-symmetric for fermions and symmetric for bosons
bosons are particles with integral values of the intrinsic spin and which obey Bose-Einstein statistics,
and
fermions have half-integral spin
3
and obey Fermi-Dirac statistics.
We look at the consequences of exchange symmetry for two non-interacting particles (or particles that interact
only weakly). We consider two identical particles labelled 1 and 2 in states labelled A and B (representing the
quantum numbers of the states).
If the wave functions corresponding to the single-particle states A and B are denoted
A
and
B
respectively,
then possible states of the two-particle system are, in the independent-particle approximation,
(1, 2) (1) (2) and (1, 2) (1) (2)
A B B A
= =
Clearly, neither of these wave functions is symmetric or anti-symmetric under the interchange of the particle
labels. Thus they are not satisfactory wave functions for describing the state of two identical particles.
However, since both
A
(1)
B
(2) and
A
(2)
B
(1) are solutions of the two-particle Schrdinger equation with
the same energy, it follows that any linear combination of the two is also a valid solution. Therefore wave
functions that are symmetric and antisymmetric with respect to particle exchange are easily generated from
these two solutions:

(1, 2) (1) (2) (2) (1)
(1, 2) (1) (2) (2) (1)
S A B A B
A A B A B
N
N


l
=
l
l
=
l

with the exchange symmetry

(2,1) (1, 2)
(2,1) (1, 2)
S S
A A
=
=

N is the normalisation factor given by ( )
1/2
AB
2 1 N

l
=
l
.
The exchange symmetry requirement therefore implies that for two identical, indistinguishable fermions
that do not interact, the wave function must be
A
(1,2), whereas for two identical bosons it must be

S
(1,2).
Furthermore, for two identical fermions, we see that

2
(1, 2) 0 for ,
A
A B = =
i.e. two identical fermions cannot have the same set of quantum numbers
4
. This is the exclusion principle
introduced empirically for electrons by Pauli in 1924 to explain the spectra of complex atoms; it is seen

2
It was shown by Pauli in 1940 that such a correlation between spin and symmetry is necessary to construct a relativistic theory of
quantum mechanics.
3
The fact that the wave function for two electrons must be antisymmetric was first recognised by Heisenberg in 1926, and was used
in describing the spectrum of the helium atom.
4
This can be generalized to system of more than two identical particles with half-integer spin: in such a system at most one particle
can occupy any one single-particle state.
Page 5.5
to be a consequence of the more general anti-symmetrization requirement on the wave function of all
systems of identical fermions, both interacting and non-interacting.
On the other hand for bosons in the same quantum state we have
(1, 2) (1) (2) 0 for ,
S A A
A B = =
indicating that bosons do not obey the Pauli principle.
2.2 Exchange symmetry for two identical fermions with s = 1/2
Because of its importance in many branches of physics, we consider in some detail the effect of exchange
symmetry on a system of two identical fermions, such as two electrons
5
. The complete two-particle wave
function (1,2) of such a system is (in the absence of any spin-orbit coupling term in the Hamiltonian
6
) the
product of a spatial wave function
1 2
( , ) r r

and the spin function (1, 2)
S
SM
that was introduced earlier:

1 2
(1, 2) ( , ) (1, 2).
S
SM
r r =


As we have seen above, the complete wave function (1,2) is anti-symmetric under the exchange of the
two fermions;
As discussed earlier in this course, the spin function (1, 2)
S
SM
is either symmetric or anti-symmetric
depending on whether the value of the quantum number S is 1 or 0, respectively.
This implies that the spatial wave function must also have definite exchange symmetry:
1 2
( , ) r r

must be
symmetric if (1, 2)
S
SM
is anti-symmetric and vice versa.
Evidently, the symmetry of the spatial wave function is determined by the relative orientation of the
particle spins.
The spatial part of the wave function of a single fermion in a state with quantum numbers (nlm) is denoted
( )
nlm
r

. In the uncoupled representation (appropriate if the Hamiltonian contains no term coupling the
coordinates of the two particles so that the independent-particle picture is applicable), the spatial wave
functions of two fermions in orbitals
7
(nl) and ( ) n l are therefore of the form

1 2
( ) ( ) with , , and , ,
nlm n l m
r r m l l m l l


= =


There are (2l + 1)( ) 2 1 l such states for fixed ( ) , l l .
The total orbital angular momentum of the two-particle system is defined by:
L l l

=


leading to
. M m m =
Using the angular momentum addition theorem, in the coupled representation:
the spatial wave function of the two particles in a state of total orbital angular momentum L and
projection M is written

1 2 1 2
,
( , ) ( ; | ) ( ) ( ),
LM nlm n l m
m m
r r C lm l m LM r r


=


(8)

5
This situation is discussed in section 5.7 of the textbook by Mandl.
6
The effect of such a term will be considered later in the course.
7
An orbital is the collection of states (nlm
l
) with fixed n and l.
Page 5.6
where the sums over ( , ) m m are constrained by , m m M

= and
L takes on any one of the (2L + 1) allowed values
, 1, ,( ). L l l l l l l = (9)
In order to discuss the effects of exchange symmetry on these wave functions, we must consider separately the
two possible situations for orbitals (nl) and ( ) n l : they must be either the same or different.
Case (i) ( ) ( ) nl n l =
Particles that have the same quantum numbers nl are called equivalent particles.
From eq.(9) we have L = 0,1,2,,2l.
From symmetry properties of the vector-coupling coefficients (not discussed in this course
8
) it follows that,
for , , n l n l

=
( )
2 1 1 2
( , ) 1 ( , ).
L
LM LM
r r r r =

(10)
So under particle exchange 1 2 we find

(1, 2) is symmetric if even
anti-symmetric if odd
LM
L
L
=
=
(11)
The complete wave function of the two-particle system is the product of the spatial part
LM
(1,2), whose
symmetry depends on L, and the spin part (1, 2)
s
SM
which, as we discussed earlier, has the following
symmetry:

(1, 2) is symmetric if 1
anti-symmetric if 0
s
SM
S
S
=
=

Hence L and S can be combined only in the following combinations to produce overall anti-symmetry for the
complete two-particle wave function:

even with 0
odd with 1
L S
L S
= =
= =

Note: this applies only when the particles occupy the same orbital.
As an example, consider the ground state of the helium atom, which has electronic configuration (1s)
2
(i.e.
both electrons occupying the 1s state).
In this state, both electrons have orbital angular momentum l = 0, so that the total orbital angular
momentum is L = 0. The spatial wave function is therefore symmetric.
This means that the spin function must be antisymmetric; i.e. the spin quantum number is S = 0.
Hence the ground-state wave function is
( )
He 1 2 100 1 100 2 1 2 1 2
1
( , ) ( ) ( )
2
r r r r =


where
100
(r) represents the solution of the radial part of the single-electron Schrdinger equation, with
quantum numbers n = 1, l = 0, m = 0.


8
See problem 5.9 on page 142 of the textbook by Mandl.
Page 5.7
Case (ii) ( ) ( ) nl n l
If , , n l n l

the argument above no longer holds, and the spatial wave function
1 2
( , )
LM
r r

as defined in
eq.(8) is neither symmetric nor anti-symmetric under particle exchange. However, since the states
1 2
( , )
LM
r r


and
2 1
( , )
LM
r r

are always orthogonal and correspond to the same energy, we can build symmetric and
anti-symmetric wave functions as follows:

[ [
[ [
1 2 2 1
1 2 2 1
1
(symm) ( , ) ( , )
2
1
(anti-symm) ( , ) ( , )
2
LM LM LM
LM LM LM
r r r r
r r r r
=
=


(12)
Possible anti-symmetric total wave functions are then the combinations

[ [
[ [
1
1 2 2 1 0
2
1
1 2 2 1 1
2
( , ) ( , ) (1, 2)
and ( , ) ( , ) (1, 2)
LM LM S
LM LM S
r r r r
r r r r

=
=





respectively.
This implies that for each value of L there must exist both singlet and triplet spin states.
This is confirmed by an examination of the spectrum of the excited states of helium, which we shall discuss
later.
3. EXCHANGE FORCES
9

Consider a one-dimensional system of two particles. We place one particle in the spatial state a with wave
function
a
(x) and the other in the state b with spatial wave function
b
(x). The two states a and b are
orthogonal and the wave functions normalized:
.
a b ab
=
We wish to compute the expectation value of the operator (x
1
x
2
)
2
in the two-particle state 1, 2 , as defined
below. This is clearly a measure of the average distance between the two particles in the two-particle state.
Note that, since x
1
and x
2
commute:

2 2 2
1 2 1 2 1 2
( ) 2 . x x x x x x =
Hence

2 2 2
1 2 1 2 1 2
1, 2 ( ) 1, 2 2 . x x x x x x = (13)
We consider two different situations.
3.1 Two distinguishable particles
We put particle 1 in state a and particle 2 in state b. The combined spatial wave function for the two-particle
state is then

1 2 1 2
( , ) ( ) ( ).
a b
x x x x =

9
This section is based on section 5.1.2 of the textbook by Griffiths.
Page 5.8
Since the particles are not identical, exchange symmetry does not impose any restriction on the wave function.
In this case, using the normalisation of the wave functions

2
2 2
1 1 1 2 1 2
2 2 2
1 1 1 2 2
2
( ) ( )
( ) ( )
,
a b
a b
a
x x x x dx dx
x x dx x dx
x


=
=
=



which represents the expectation value of x
2
in the one-particle state a. Similarly, it is easy to show that

2 2
2
b
x x =
and

1 2
. a b x x x x =
Hence from eq.(13) it follows that for two distinguishable particles

2 2 2
1 2
dist
1, 2 ( ) 1, 2 2 . a b
a b
x x x x x x = (14)
Note that we would obtain exactly the same result using the two-particle wave function
b
(x
1
)
a
(x
2
).
3.2 Two identical fermions or two identical bosons
For identical particles there are two possible combined states, which are symmetric or anti-symmetric with
respect to interchange of the two particles:

1 2 1 2 1 2
1
( , ) ( ) ( ) ( ) ( )
2
a b b a
x x x x x x

l
=
l
l

with the + sign indicating symmetric and the sign antisymmetric. The calculation of the expectation value
is now more complicated; using the orthogonality of the single-particle wave functions

2 2 2 2
1 1 1 1 2 2
2 2 2
1 1 1 2 2
2
1 1 1 1 2 2 2
2
1 1 1 1 2 2 2
2 2
2 ( ) ( )
( ) ( )
( ) ( ) ( ) ( )
( ) ( ) ( ) ( )
0 0.
a b
b a
a b b a
b a a b
a b
x x x dx x dx
x x dx x dx
x x x dx x x dx
x x x dx x x dx
x x






=

=





That is

2 2 2
1
1
.
2
a b
x x x
l
=
l

Similarly

2 2 2
2
1
2
a b
x x x
l
=
l

and

2
1 2
a b ab x x x x x =
where
( ) ( ) . ab
a b
x x x x dx =

(15)
Page 5.9
Substituting these results into eq.(13) we obtain the expectation value of the operator (x
1
x
2
)
2
for the case
of identical particles; comparing with the result for distinguishable particles eq.(14) we find

2 2
2
1 2 1 2
dist
1, 2 ( ) 1, 2 1, 2 ( ) 1, 2 2 . ab x x x x x

= (16)
If the two spatial wave functions do not overlap significantly, then from eq.(15)
( ) ( ) 0 ab
a b
x x x x dx = =


and hence from eq.(16)

2 2
1 2 1 2
dist
1, 2 ( ) 1, 2 1, 2 ( ) 1, 2 . x x x x

= (17)
We conclude that:
The exchange symmetry property of the wave function for identical particles is not important if they are
spatially far apart.
Otherwise, from eq.(16) we see that:
Two identical particles in a symmetric spatial state tend, on average, to be somewhat closer together than
two distinguishable particles in the same two single-particle states.
Two identical particles in an anti-symmetric spatial state tend to be somewhat further apart than two
distinguishable particles in the same two states.
In other words, the two-particle system behaves as though there is a force of repulsion or attraction between
the two identical particles; this is the so-called exchange force. It is a purely geometrical consequence of the
symmetrization requirement (or, equivalently, the Pauli principle).
3.3 Hydrogen molecule
The measured spin of the ground state of the hydrogen molecule provides evidence for the existence of the
exchange force. In the simplest model, the molecular ground state comprises one electron in the atomic ground
state centred on nucleus 1 and one electron in the atomic ground state centred on nucleus 2.
If the two electrons are in the symmetric spatial state (hence in the anti-symmetric spin state with
S = 0), the exchange force is attractive and tends to concentrate the electrons towards the middle of the
molecule, resulting in an accumulation of negative charge that would attract the two protons inwards.
This is the covalent bond that holds the molecule together.
If the two electrons are in the anti-symmetric spatial state (hence in the symmetric spin state with
S = 1), the exchange force is repulsive and results in an accumulation of negative charge towards the
outside of the molecule, that would attract the protons outwards. The molecule is therefore not held
together, but pulled apart.
We conclude that the stable molecule must be in the S = 0 electronic spin state. This is confirmed by ex-
periment.
Page 5.10
4. THE HELIUM ATOM
Consider a helium atom or helium-like ion containing a nucleus of charge Ze and two orbiting electrons. We
assume that the nucleus is infinitely heavy, so that centre of mass motion can be neglected. Ignoring
fine-structure effects (which will be included later), the Hamiltonian is given by
( )
2 2 2
2 2
1 2
0 1 2 0 1 2
1 1 1
2 4 4
Ze e
H
m r r r r
1

=


( )

(18)
in an obvious notation. For the moment we retain Z in the expression for the Hamiltonian, so that it is also
valid for two-electron ions such as H

, Li
+
, Be
++
and so on.
The first term represents the kinetic energy of the two electrons,
the second represents the interaction of each electron with the nucleus of charge +Ze, and
the third term is the electron-electron interaction energy.
No exact method exists to solve the Schrdinger equation with this Hamiltonian, because of the form of the
last term (which couples the motion of the two electrons and is non-central in nature), so we must resort to
approximation techniques; we shall apply both perturbation theory and the variational method.
4.1 Helium Ground State: Perturbation Theory
10

We can write the Hamiltonian as

0
H H V =
where
H
0
is the Hamiltonian of two non-interacting electrons each moving independently in a central Coulomb
potential, which is treated as the unperturbed Hamiltonian:
( )
2 2
2 2
0 1 2
0 1 2
1 1
2 4
Ze
H
m r r
1
=


( )


V is the electron-electron interaction, which is treated as a perturbation
11
:

2
0 12
1
.
4
e
V
r
=
The unperturbed Hamiltonian splits into separate terms for each of the particles; this is exactly the situation
discussed earlier in connection with the independent particle approximation, so that we can invoke all the
results obtained there:
the unperturbed wave function is a product of single-particle wave functions and
the unperturbed energy is a sum of single-particle energies.
We therefore have to solve the single-particle Schrdinger equation with Hamiltonian

2 2
2
0
1
2 4
i i
i
Ze
H
m r
=

;

10
The original version of this section was based on section 7.2 of the textbook by Mandl.
11
This is clearly not justified for small Z since V cannot be substantially smaller than the interaction energy of an electron with the
nucleus.
Page 5.11
this is identical to that of the hydrogen atom except for the presence of Z. So in the helium ground state, in
zeroth order, each electron is in a hydrogen (1s) state with wave function

1/2
3
100
3
0 0
exp
nlm
Z Zr
a a

1
1

= =


( )
( )

where a
0
is the Bohr radius. Each electron has energy in Rydbergs (defined in the section on the Stark effect):

2
1
Ry
n
E E Z = =
Hence in the independent-particle approximation, the spatial wave function of the helium ground state is the
product of two single-electron wave functions:

( )
3
1 2
100 1 100 2
3
0 0
( ) ( ) exp
Z Z r r
r r
a a

1
1

=



( )
( )

Since the spatial wave function is symmetric under particle exchange, the electrons must be in the anti-
symmetric S = 0 state. The complete wave function is therefore

( )
( )
100 1 100 2
3
1 2
1 2 1 2
3
0 0
(1, 2) ( ) ( ) (1, 2)
1
exp
2
r r
Z Z r r
a a

=
1
1
=


( )
( )
(19)
The energy of this unperturbed state is the sum of the two single-electron energies:

(0) 2
1
2 Ry. E Z =
The first-order correction to the energy is, after considerable calculation
12
:

( )
2
2 2 3
(1) 1 2 3 3
1 2 1 3
0 12 0 12 0 0
2
0 0
1 1 2
(1, 2) (1, 2) exp
4 4
5 5
Ry
4 8 4
e e Z Z r r
E d r d r
r r a a
e Z
Z
a

1
1

= =



( )
( )
= =


So the energy of the ground state is, in this approximation,

( )
2
1
5
2 1 Ry.
8
E Z
Z
=
The table compares calculated and measured energies (in eV) for the eight lightest helium-like atoms and ions.
Z
(0)
1
E
(1)
1
E

1
E Expt.
H

1 -27.2 17.0 -10.2 -14.3


He 2 -108.8 34.0 -74.8 -79.0
Li
+
3 -244.8 51.0 -193.8 -198.1
Be
++
4 -435.2 68.0 -367.2 -371.6
B
+++
5 -679.9 85.0 -594.9 -599.6
C
4+
6 -979.1 102.0 -877.1 -882.1
N
5+
7 -1332.6 119.0 -1213.7 -1219.1
O
6+
8 -1740.6 136.0 -1604.6 -1610.7

12
See pp 286-8 of Quantum Physics by S Gasiorowicz, 2
nd
edition, published by Wiley (1974), or pp 263-4 (1
st
edition) or pp 300-1
(2
nd
edition) of the book by Griffiths.
Page 5.12
Note that:
The magnitude of the perturbation relative to the unperturbed value is 5/8Z , which is over 60% for the
hydrogen ion, 30% for helium, 20% for lithium, falling to 8% for oxygen.
The approximation is consistently too high by about 4-6 eV, which represents an error of 5% for helium,
2% for lithium, 1% for beryllium, decreasing to only 0.4% for oxygen.
The approximation is surprisingly accurate, even for helium, given the magnitude of the perturbation for small
Z.
4.2 Helium Ground State: Variational Method
13

In the simplest approximation, as described above, each electron in the ground state is represented by a
hydrogenic (1s) wave function, suggesting that a suitable trial function for the helium ground state is
( )
( )
3
1 2
1 2
3
0 0
, exp
r r
r r
a a

l
= l
l
l

(20)
where is the variational parameter. Note that the wave function is correctly normalized.
The Hamiltonian is given by eq.(18):
( )
2 2 2
2 2
1 2
0 1 2 0 1 2
1 1 1
.
2 4 4
Ze e
H
m r r r r
1

=


( )


With this trial function the expectation value
14
of the kinetic energy operator is

( )
2 2 2
2 2
2 2 3 3
1 2 1 1 2 2
2 2
2
0
2 2 2
2 2 Ry
2
d r d r
m m m
m a

=
1

= =


( )



and the expectation value of the electron-nucleus potential energy is

2 2
0 1 2 0 0
1 1
2 4 Ry.
4 4
Ze Ze
Z
r r a


1 1

= =



( ) ( )

The electron-electron interaction energy is

2 2
0 12 0 0
1 5 5
Ry
4 8 4 4
e e
r a



1

= =


( )

Note that the integral appearing in the last computation is identical to that encountered in the perturbation
approach but with Z replaced by . In each of these calculations we have converted the energy into Rydbergs.
Collecting these results, we find that the variational energy is given by

( )
2
5
( ) 2 2 Ry.
16
E Z
l
=
l
l

Putting the partial derivative of the energy with respect to equal to zero gives for the value of the pa-
rameter that minimizes the energy

13
The original version of this section was based on section 8.2.2 of the textbook by Mandl. See also section 7.2 of the book by
Griffiths.
14
Note that the integrals needed for the calculation of K and V for helium are identically double those determined in the
corresponding variational calculation for the ground state of hydrogen.
Page 5.13

0
5
16
Z =
and substitution back into the expression for the variational energy provides the minimum energy

( )
2
0
5
( ) 2 Ry.
16
E Z =
The table compares energies (in eV) calculated using the variational method with experimental values and
energies from first-order perturbation theory.

Z
0
( ) E (expt) E E(pert)
H

1 -12.9 -14.36 -10.2


He 2 -77.4 -79.0 -74.8
Li
+
3 -196.4 -198.1 -193.8
Be
++
4 -369.8 -371.6 -367.2
B
+++
5 -597.6 -599.6 -594.9
C
4+
6 -879.8 -882.1 -877.1
N
5+
7 -1216.3 -1219.1 -1213.7
O
6+
8 -1607.3 -1610.7 -1604.6
Note:
The variational and perturbation methods differ in their predictions by
25
128
Ry = 2.65 eV.
The variational method produces the more accurate energies in all cases, with errors increasing with Z
from 1.4 eV to 3.4 eV, and the relative error decreasing from 10% for H

, 3.3% for He, to 0.2% for O


6+
.
The energy measured for the H

ion, -14.36 eV, is 0.76 eV below the ground state energy of hydrogen, in-
dicating that the ion is actually bound by 0.76 eV, decaying into a neutral hydrogen atom plus a free electron
if this amount of energy is supplied to the ion. However, both calculations fail to predict the existence of a
bound state, since the calculated energies are both above -13.60 eV.
With our choice of trial wave function, the variational parameter has a physical interpretation. Comparing
(20) with eqs.(19) we see that the trial function

represents two electrons moving independently in a


Coulomb field with charge e . The variational wave function
0

therefore represents two independent


electrons moving in a Coulomb field whose effective charge is

( ) eff 0
5
.
16
Z e e Z e = =
It appears that the nuclear charge seen by an electron is partially screened by the other electron, being reduced
on average from its full value Ze to Z
eff
e. The calculation takes into account, approximately, the screening
effect that will be discussed later in connection with the central-field approximation for multi-electron atoms.
A better trial function for the ground state of the two-electron system is
( )
1 2 2 1
, 1 2
0 0
, exp exp
r r r r
r r
a a

l l
= l l
l l
l l


which has two variational parameters, allowing different screening for the two electrons. Note that the second
term is required to make the spatial wave function symmetric under particle exchange. This leads to an upper
Page 5.14
bound of -78.20 eV for the helium ground state, compared with the measured value -79.0 eV. It also correctly
predicts that the H

ion is bound.
4.3 Excited states of helium
15

Again, as a first approximation, we neglect the electron-electron interaction (and any relativistic effects) so
that each electron moves independently in the Coulomb field created by the nucleus. We place one electron
in the orbit = (n,l) and the other in the orbit ( , ) n l = . The energy of the state is then the sum of the two
electron energies:

( )
(0)
2 2
1 1
4 Ry. E
n n

=
However, it is easily seen that one electron must be in the n = 1 orbit for all bound excited states:
The ionization energy of helium, which is the energy needed to completely remove one electron from the
atom, leaving the He
+
ion in the ground state plus a free electron, is found by substituting
n = 1, n = , resulting in
(0)
ion
E = 54.4 eV (in this approximation).
Placing both electrons in the excited electronic state 2 n n = = gives
(0)
27.2 E = eV, which is above
the ionization energy. States with , 2 n n will obviously lie above this. All such states will therefore
decay by emitting an electron, leaving a positively-charged helium ion.
Hence for all bound helium states we must have one electron in the lowest energy state, i.e.
= (n = 1,l = 0). This is illustrated in the figure below (note that the zero of the energy scale is the
ionization energy).

Figure 1: Calculated spectrum of the helium atom in the independent-particle approximation

15
A discussion of the first few excited states in terms of the variational method can be found on page 193 of the textbook of Mandl.
We do not follow his approach in these notes. See for example pp 289-291 of the book by Gasiorowicz, pp 323-4 of Quantum
Mechanics by Sara M McMurry, published by Addison-Wesley (1993), or pp 445-6 of Quantum Mechanics, 2
nd
edition, by P
Merzbacher, published by Wiley (1970)
Page 5.15
Still neglecting the electron-electron interaction, possible antisymmetric two-electron wave functions for an
excited state have the form

1 2 0 1 2
1 2 1 1 2
(1, 2) ( , ) ( , )
(1, 2) ( , ) ( , )
S
S
r r s s
r r s s

=
=
=
=



where the labels on the and indicate the exchange symmetry of the spatial wave function.
Since the spin wave function is antisymmetric for S = 0, it must be multiplied by a symmetric spatial
wave function
+
.
The symmetric spin wave function for S = 1 must be multiplied by an antisymmetric spatial wave
function

.
In terms of the single-electron spatial wave functions ( ) r


, the two-electron spatial wave functions are
[ [
1 2 1 2 2 1
1
( , ) ( ) ( ) ( ) ( ) .
2
r r r r r r



=


We now include the effects of the electron-electron repulsion. In first-order perturbation theory, the energy
correction due to this interaction is
( ) ( )
(1)
(1, 2) (1, 2)
(1, 2) (1, 2) 1, 2 1, 2
(1, 2) (1, 2) ,
S S
E V
V
V




=
=
=

where we have used the fact that the perturbation

2
0 1 2
1
4
e
V
r r
=


is independent of the spins. Hence

[ [
[ [
(1)
1 2 1 2 2 1 2 1
1 2 2 1 2 1 1 2
1
( ) ( ) ( ) ( ) ( ) ( ) ( ) ( )
2
1
( ) ( ) ( ) ( ) ( ) ( ) ( ) ( ) .
2
E r r V r r r r V r r
r r V r r r r V r r






=




The first and second matrix elements are equal, as are the third and fourth, so we can write the energy
correction as

(1)
E I J

=
where the direct (Coulomb) interaction is

1 2 1 2
2
2 3 3
1 2 1 2
0 1 2
( ) ( ) ( ) ( )
1
( ) ( )
4
I r r V r r
e
d r d r r r
r r



=
=






and the exchange integral is

1 2 2 1
2
3 3
1 2 1 2 2 1
0 1 2
( ) ( ) ( ) ( )
1
( ) ( ) ( ) ( ).
4
J r r V r r
e
d r d r r r r r
r r





=
=






The first term is equivalent to the classical result, and is what would be obtained if the electrons were
distinguishable.
The second term has no classical counterpart, and results from the indistinguishability of the particles.
Page 5.16
The energy I is manifestly positive, and the energy J is found by calculation
16
to be positive but smaller (as
expected intuitively by inspection of the integral). Hence

(1)
is positive E

and
(1) (1)
. E E


We see that:
The general effect of the electron-electron repulsion is to increase the energy of an excited state, as would
be expected of a repulsive force.
The effect of the exchange symmetry is to split each level into two levels, separated by the exchange
energy 2J and characterized by S = 0 (upper level) and S = 1 (lower level).
The energy of the perturbed states therefore depends on their spin, even though the perturbation is the
Coulomb interaction between the electrons, which is clearly independent of spin. (See also the earlier
discussion on exchange forces.)
In summary, there are two sets of excited states for helium:
S = 0 states, called parahelium,
S = 1 states, called orthohelium.
These were originally thought to be from two different types of helium, hence the different names.
The observed spectrum is illustrated in the figure below
17
, which shows S = 0 states as solid lines and S = 1
states as dashed lines. The energies are indicated in eV relative to the ionization energy. The quantum
numbers (n,l) refer to one electron, the other being in the n = 1 orbital for all states.
Comparison can be made with the previous figure, which shows the spectrum calculated by ignoring the
electron-electron interaction and the exchange symmetry requirements. The effect of the perturbation is large,
especially for the ground state (as noted earlier). The splitting of all excited states according to S confirms the
qualitative predictions made above.

16
See Gasiorowicz, pp 289-290
17
Adapted from the figure on page 325 of the textbook of McMurry.
Page 5.17

Figure 2: Experimental energy levels of atomic helium

You might also like