You are on page 1of 75

Lecture notes for Atomic and Molecular

Physics, FYSC11, HT 2015

Joachim Schnadt

August 30, 2015


Chapter 1

Before we really start

1.1 What have you done previously?


Already in FYSA21 you have seen nearly all the quantum mechanical con-
cepts that we are going to talk about. These are:

• Particle-wave duality, de Broglie wavelengths:


λ = h/p

• Heisenberg’s uncertainty relationship for position and momentum:


∆x∆p ? h̄/2

• Energy quantisation, for example in the radiation of a black body with


radiation density
Sν∗ , Sν∗ dν dΩ dA = 2hv 3 dν dΩ dA
c2 ehν/kT −1
(cf. Fig. 1.1)

• Wave functions: Ψ(x) or Ψ(x, t)


(NB: I often write the position vector as x with components x1 = x,
x2 = y, and x3 = z, although I sometimes also might write it as r.)
Wave functions are interpreted as the probability amplitude of find-
ing the particle described by Ψ at a position x and at time t. Thus,
the

corresponding probability is |Ψ(x, t)|2 , and it is normalised to 1:
|Ψ(x, t)|2 d3 x = 1

• Orthogonality of wave functions:



ϕa and ϕb are said to be orthogonal if ϕ∗a (x)ϕb (x)d3 x = δab

2
• Quantum mechanical operators and measurements, eigenvalue equa-
tions:
⟨ ⟩ ∫
 = Ψ∗ (x)ÂΨ(x)d3 x, and
⟨ ⟩
ÂΨ(x) = aΨ(x), i.e., Â = a

• The superposition principle:


if ϕa and ϕb are wave functions, then also Ψ(x, t) = ϕa (x, t) + ϕb (x, t)
is a wave function

• Complete mutually orthogonal sets

• Particle in a potential well

• Compatible and non-compatible variables

• Commutators [A, B]
The canonical commutation relations:
[xi , xj ] = [pi , pj ] = 0 and [xi , pj ] = ih̄δij

• The parity operator P :


P Ψp (x, t) = ±Ψp (−x, t)

• The one-dimensional oscillator and ladder operators ↠,â

• Vibrational spectra

During the course we will re-encounter many of these concepts, although


they sometimes might look somewhat different from what you are used to.

1.2 The Dirac notation


Table 1.1 shows a couple of quantum mechanical concepts both in the way
you have written them previously and in the notation introduced by Paul
Dirac (1902-1984), one of the founders of quantum mechanics. The notation
is called ”Dirac notation” or ”bra-ket notation”. The main difference be-
tween wave function and Dirac notation, which indeed also is a conceptual
difference, is that the wave function notation always specifies coordinates,
while the Dirac notation is coordinate-free. If coordinates have to be speci-
fied, e.g. when you want to describe the outcome of a position measurement,
you specify that you want to use the position basis written as {⟨x|} (the

3
Figure 1.1: Spectral distribution of the radiation
density S ∗ (λ) of a blackbody (Figure taken from
http://www.ifremer.fr/droos/cours teledetection/report1/node16.html).

curly brackets indicate that the position basis is made up from a set of vec-
tors ⟨x|). Depending on your problem you also might specify other bases,
such as the momentum basis {⟨p|}.
The primary advantage of the Dirac notation is that it makes life easier:
as you see from Table 1.1 instead of having to write out integrals for e.g. the
scalar product and expectation value of an operator, one just can resort to
the time and space saving ⟨ϕa |ϕb ⟩ and ⟨Ψ|A|Ψ⟩. Beyond this, there exists
indeed a conceptual difference between the concept of a wave function and
that of a quantum mechanical state: wave functions are generally tied to
their interpretation as finding a system within a volume d3 x at position
x with probability |Ψ(x)|2 , i.e. they are tied to the position basis of the
quantum mechanical state space. The quantum mechanical state, in contrast,
is independent of the choice of basis of the quantum mechanical vector state
and thus more general (note, though, that one frequently also uses e.g. wave
functions which depend on linear momentum rather than position).
As an example we may take the hydrogen atom. In Dirac notation, the
quantum mechanical state of a hydrogen atom in the state with quantum
numbers n, l, ml – the knowledge of which fully characterises the state of the
hydrogen atom – is written |nlml ⟩. In wave function notation, the probably

4
Concept Notation that you have Dirac notation
used before

Wave function Ψ(x) = ⟨x|Ψ⟩

Quantum mechanical state |Ψ⟩



Scalar product ϕ∗a (x, t)ϕb (x, t)d3 x = ⟨ϕa |ϕb ⟩

Normalisation |ψ(x, t)|2 d3 x = 1 = ⟨ϕ|ϕ⟩

Orthogonality ϕ∗a (x, t)ϕb (x, t)d3 x = δab = ⟨ϕa |ϕb ⟩ = δab

Operator  A

Eigenvalue equation ÂΨ(x) = aΨ(x) = ⟨x|A|Ψ⟩ = a ⟨x|Ψ⟩

A |Ψ⟩ = a |Ψ⟩
⟨ ⟩ ∫
Expectation value  = Ψ∗ (x, t)ÂΨ(x, t)d3 x = ⟨A⟩ = ⟨Ψ|A|Ψ⟩

Table 1.1: The Table shows some quantum mechanical concepts in both
”conventional” and Dirac notation.

5
of finding the same hydrogen atom within a volume element d3 x at x is
written |ψn,l,ml (x)|2 d3 x. The connection between both notations is provided
by ⟨x|nlml ⟩ = ψn,l,ml (x).

6
Chapter 2

Quantum mechanics

The primary goals of this chapter are to

• introduce a particular angular momentum, the spin angular momen-


tum,

• and to introduce the concepts of quantum mechanical states and oper-


ators and the tools needed to use them.

Beyond that, we will also take an excursion in translations and time evolution
to show that it is possible to derive the canonical commutator relationships,
the replacement prescription of p by the spatial derivative ∇ in the wave
function formulation of quantum mechanics, and the Schrödinger equation.

2.1 The Stern-Gerlach experiment


The Stern-Gerlach experiment was conceived in 1921 by Otto Stern and
Walther Gerlach and carried out by Stern and Gerlach in Frankfurt in 1922.
A beam of (neutral) Ag atoms is transmitted through an inhomogenous mag-
netic field and further onto a photographic plate (see Fig. 2.1). Ag has alto-
gether 47 electrons, out of which 46 form a spherical shell with no resulting
angular momentum. The total angular momentum of the neutral Ag atom
is thus given by the spin of the 47 electron, or, in other words, the neutral
Ag atom behaves like a single electrons when it comes to its interaction with
a magnetic field1 .
1
The existence of an angular momentum intrisic to the electron has to be postulated
in non-relativistic quantum mechanics – it is an experimental fact, which finds one of its
clearest evidences in the Stern-Gerlach experiment, although the concept was unknown at
the time of the experiment and introduced first a couple of years later.

7
Figure 2.1: Stern-Gerlach experiment (Figure taken from
http://www.wikipedia.org).

8
The magnetic moment of the atom is proportional to the spin,

µ ∝ S, (2.1)
and the force onto the atom is then, entirely as in classical physics, given by
the gradient of the scalar product of the magnetic moment and the magnetic
field,
∂ ∂Bz
Fz = (µB) = µz . (2.2)
∂z ∂z
Classically, one would expect that, assuming that the spin angular momen-
tum corresponds to a spinning motion of the electron, all values of µz are
assumed between −|µ| and |µ|. Hence, the photographic plate should ex-
hibit the recording of a beam of atoms elongated along z. What instead is
observed are two distinct components that correspond to

µz ∝ Sz = ± , (2.3)
2
where h̄ is the Planck constant h̄ = 6.582210−16 eV s. The Stern-Gerlach
experiment thus shows that if the ”magnetic” isotropy of space is removed
by application of a magnetic field in a particular direction the component
of the spin in this direction is quantised, i.e., it can assume only particu-
lar values. The Stern-Gerlach experiment represents one of the most direct
proofs of quantisation and has had a profound influence on the development
of quantum mechanics.

2.1.1 Sequential Stern-Gerlach experiments


Of course it is possible to apply magnetic fields in other directions of space
than in the (arbitrarily chosen) z-direction. Then one can also let the silver
atom beam transmit sequential arrangements of Stern-Gerlach setups, such
as shown schematically in Fig. 2.2. In each of these sequential Stern-Gerlach
experiments the beam traverses first a magnetic field in the z-direction, which
leads to a split-up of the atomic beam into a beam with electrons with Sz =
+h̄/2 and one with Sz = −h̄/2. Only the former is allowed to enter the second
Stern-Gerlach setup with a magnetic field in either the z- or x-direction. As
expected, if the beam is transmitted through a second magnetic field in
the z-direction only one beam with Sz = +h̄/2 exits the setup. Also as
expected, the application of a magnetic field in the x-direction leads to a
splitting-up of the beam into two beams with Sx = +h̄/2 and Sx = −h̄/2,
respectively. If one now again stops one of the beams, namely that with Sx =
−h̄/2, and lets the other go through yet another Stern-Gerlach setup with

9
z

Figure 2.2: Sequential Stern-Gerlach experiments (Figure taken from J. J.


Sakurai, Modern Quantum Mechanics).

a magnetic field in the z-direction, something strange happens: classically,


one would expect only one beam to exit this setup with Sz = +h̄/2. The
reason is that the first beam stopper already removed all electrons with
Sz = −h̄/2. What is observed instead is two beams: one with Sz = +h̄/2
and one with Sz = −h̄/2! It is very important to see that this behaviour
is entirely unclassical. It was only the advent of the quantum mechanical
theory which could render this result plausible. Quantum mechanically, Sx
and Sz cannot be determined simultaneously, Sx and Sy are non-compatible
observables. The determination of Sx in the second Stern-Gerlach experiment
destroys all information obtained in the first Stern-Gerlach setup with a
magnetic field in the z-direction. A very peculiar behaviour indeed!
We have to leave the classical description behind, and instead go over to
an alternative formulation, which makes use of the concept of states. It is
said that the Ag atoms are in spin states, which are represented by vectors in
a two-dimensional vector space. The vector space is two-dimensional because
they are two possible outcomes of a spin direction measurement, ”spin-up”
and ”spin-down”. We write the spin states as

1 1
|Sz , +⟩ = √ |Sx , +⟩ − √ |Sx , −⟩ (2.4)
2 2

and
1 1
|Sz , −⟩ = √ |Sx , +⟩ + √ |Sx , −⟩ , (2.5)
2 2

10
or by re-arrangement
1 1
|Sx , +⟩ = √ |Sz , +⟩ + √ |Sz , −⟩ (2.6)
2 2
and
1 1
|Sz , −⟩ = − √ |Sz , +⟩ + √ |Sz , −⟩ . (2.7)
2 2
This implies that all states which can be produced in the measurement of the
z-component of the spin can be written as superpositions of states produced
in the measurement of the spin’s x-component and vice versa. The outcome
of the sequential Stern-Gerlach experiments can now be explained: after the
first experiment only the beam with atoms in the |Sz , +⟩ is allowed to enter
the second setup. This beam is a superposition of the |Sx , +⟩ and |Sx , −⟩
states, and thus after the second setup with a magnetic field in the x-direction
two beams are found. Only the |Sx , +⟩ beam is allowed to go on – but the
atoms of this beam with |Sx , +⟩ are from (2.6) seen to be either in the |Sz , +⟩
or |Sz , −⟩ state!
Now, of course there is a third component of the spin, the Sy component.
It is incorporated in the above system by complex addition:

1 i
|Sy , +⟩ = √ |Sz , +⟩ + √ |Sz , −⟩ , (2.8)
2 2
1 i
|Sy , −⟩ = √ |Sz , +⟩ − √ |Sz , −⟩ . (2.9)
2 2

2.2 The vector space of quantum mechanics


2.2.1 Requirements
In the preceding section we have seen that the quantum mechanical states can
be described as vectors in a vector space. So far we had restricted ourselves to
a two-dimensional space with basis {|Sz , +⟩, |Sz , −⟩} (alternatively: {|Sx , +⟩,
|Sx , −⟩} or {|Sy , +⟩, |Sy , −⟩}, but it is convention to use the {|Sz , +⟩, |Sz , −⟩}
basis). The concept can be generalised to include all necessary physical
quantities.
What is needed is the following: a complex vector space

• with a norm (i.e., a ”length” of the vectors, which is going to repre-


sent the probability of a measurement), which requires a scalar (inner)
product,

11
• with infinite dimensions (the number of dimensions defines the num-
ber of possible measurement outcomes, and, for example, a position
measurements thus requires an infinite number of dimensions),

• which is complete (mathematically: every Cauchy sequence converges,


which in turn implies that each superposition of vectors from that space
also converges to a vector in that space, even though the difference
between the vectors might be arbitrarily small).

2.2.2 States: kets


The requirements are exactly the requirements which are fulfilled by a Hilbert
space, named after the German mathematician David Hilbert (1862-1943).
Thus the quantum mechanical state space is a Hilbert space, with physical
states represented by ”ket” vectors |α⟩. Also a superposition of states is a
vector on the quantum mechanical state space:

|α⟩ + |β⟩ = |γ⟩ ,

and, likewise,
c |α⟩ = |α⟩ c,
where c is a complex number. Actually, c |α⟩ represents the same state as
|α⟩ alone.

2.2.3 Operators
Now that we have defined the physical states, we also have to tell how to
arrive at a measurement of a physical quantity such as Sz above. First of all, a
physical quantity which can be observed is called an observable. Observables
are represented by operators on the vector space. Operators act on states:

A |α⟩ .

As in mathematics, operators may have eigenvectors or eigenstates, as they


are called in quantum mechanics. If the eigenstates of an observable A are
denoted by |a′ ⟩, |a′′ ⟩, |a′′ ⟩, . . ., then

A |a′ ⟩ = a′ |a′ ⟩ ,

A |a′′ ⟩ = a′′ |a′′ ⟩ ,


...,

12
where the a′ , a′′ , . . . are numbers.
The introduction of operators in this way makes it possible to properly
describe what happens in the sequential Stern-Gerlach experiments above.
The z-component of the spin angular momentum is an observable and thus
represented by an operator Sz . It has eigenstates |Sz , +⟩ and |Sz , −⟩ with
eigenvalues h̄/2 and −h̄/2 (you already see that the eigenvalues will represent
the outcomes of a measurement!):

Sz |Sz , +⟩ = |Sz , +⟩ , (2.10)
2

Sz |Sz , −⟩ = − |Sz , −⟩ . (2.11)
2
Similar definitions apply to the operators Sx and Sy .
In the sequential Stern-Gerlach experiments the first Stern-Gerlach setup
is a measurement of the spin in the z-direction, which corresponds to an
application of Sz to the states of the silver atom beam, which can be in
either state |Sz , +⟩ or |Sz , −⟩; the outcome of the measurement is thus either
h̄/2 or −h̄/2 and leaves the atoms in the corresponding states |Sz , ±⟩:

Sz |Sz , ±⟩ = ± |Sz , ±⟩ . (2.12)
2
For the second and third experiment in Fig. 2.2 the second setup traversed
by the |Sz , +⟩ beam corresponds to a measurement of the spin in the x-
direction. According to (2.4), an application of Sx to |Sz , +⟩ yields ±h̄/2
and leaves the atoms in states |Sx , ±⟩. In the third experiment of Fig. 2.2
only |Sx , +⟩ is allowed to traverse the final magnetic field in z-direction. This
corresponds again to a measurement of Sz . An application of the operator Sz
to (2.6) yields immediately the outcome of the experiment: the z-component
of the spin can either be h̄/2 or −h̄/2, and after measurement the atoms are
either in state |Sz , +⟩ or |Sz , −⟩.
Other examples for observables are the position and momentum operators
x and p:

x |x′ ⟩ = x′ |x′ ⟩ , (2.13)


′ ′ ′
p |p ⟩ = p |p ⟩ . (2.14)
Observe the difference between operators x and p and their eigenvalues x′
and p′ , which are numbers!
Finally, two operators A and B are equal if their application to an arbi-
trary state yields the same results:

A = B if (and only if) A |α⟩ = B |α⟩ ∀ |α⟩ . (2.15)

13
2.2.4 Dimensionality of the vector space
As already indicated above, the dimensionality of the quantum mechanical
vector space of interest i determined by the number of possible outcomes
of a measurement (we will typically restrict ourselves to considering only a
relevant subspace of the total Hilbert space of quantum mechanics, which
comprises dimensions for all possible outcomes of measurements of all phys-
ical quantities). In the Stern-Gerlach experiment one really has only two
different outcomes of the measurement, which in a bit more colloquial lan-
guage are called ”spin-up” and ”spin-down”. You might argue that you could
measure the z-component of the spin in the x-, y-, and z-directions, with a
total of six dimensions. However, equations (2.4) to (2.9) show that the out-
come of a measurement of Sx and Sy always can be described in terms of the
z-component of the spin (of course the use of the z-component is a conven-
tion); hence two dimensions are enough. Any spin state can be written as
superposition of the two possible eigenstates of Sz :

|Ψ⟩ = c+ |Sz , +⟩ + c− |Sz , −⟩ .

More generally, if an observables has eigenstates {|a′ ⟩} (observe the use


of curly brackets, which again indicates a set) then a general state can be
written as

|Ψ⟩ = ca′ |a′ ⟩ . (2.16)
a′

2.2.5 States: bras


Now from Mathematics it is known that each vector space has a dual space,
which also is a vector space – a mirror image, so to say. Of course also the
quantum mechanical Hilbert space has a dual space. The vectors of this dual
space are written as

⟨a| ,

and they are called bra’s. The bra which corresponds to a ket c |a⟩ is c∗ ⟨a|,
where c∗ is the complex conjugate of c.
The scalar product (inner product) of a bra ⟨a| and a ket |b⟩ is written as

⟨a|b⟩ . (2.17)
It has the following property:

14
⟨a|b⟩ = ⟨b|a⟩∗ . (2.18)
This implies that ⟨a|a⟩ is a real number (how do you see that?). The scalar
product of a bra with its corresponding ket is positive or zero:

⟨a|a⟩ ≥ 0. (2.19)
This property is important since the scalar product defines the norm and thus
the length of a vector. Since the square of the length of a quantum mechanical
vector ⟨a|a⟩ is coupled to its probability interpretation (cf. Table 1.1) this
value has to be zero or larger and cannot be negative. It is also practical
(but not necessary) to require that

⟨a|a⟩ = 1. (2.20)
For an non-normalised state |a′ ⟩ this length normalisation is achieved by
1
|a⟩ = √ |a′ ⟩ . (2.21)
′ ′
⟨a |a ⟩
Finally, two states |a⟩ and |b⟩ are said to be orthogonal if

⟨a|b⟩ = ⟨b|a⟩ = 0.

2.2.6 More on operators


Operators in bra space and Hermitian operators
Since we have operators on ket space, we also need operators on bra space.
In the Dirac notation operators are defined on kets from the left and on
bras from the right side (the arrows merely indicate, into which direction the
operators act in the equations):


|b⟩ = X |a⟩ ,


⟨b| = ⟨a| Y .
The operator in dual (bra) space that corresponds to the operator X in
ket space is X † , the Hermitian adjoint of X. Thus

X |a⟩ does not correspond to ⟨a| X,

but
X |a⟩ corresponds to ⟨a| X † .

15
Operators have matrix representations (we will come back to this point
soon); the Hermitian adjoint of an operator is then obtained by taking the
complex conjugate of the transpose of the operator’s matrix in a matrix
representation:

Hermitian adjoint: aij −→ a∗ji . (2.22)


The operator is then said to be Hermitian (or self-adjoint) if the Hermitian
adjoint is the same as the matrix, which corresponds to the operator. An
example of Hermitian operators is given by the Pauli matrices:

( ) ( ) ( )
0 1 0 −i 1 0
aij
1 0 i 0 0 −1
( ) ( ) ( )
0 1 0 −i 1 0
a∗ji
1 0 i 0 0 −1

The eigenvalues of a Hermitian operator A are real, and, furthermore the


eigenkets of A corresponding to different eigenvalues are orthogonal. This is
easily shown: Assume that A is Hermitian. Then

A |a′ ⟩ = a′ |a′ ⟩
(2.23)
⇒ ⟨a′′ | A† = a′′∗ ⟨a′′ | ,
but also

⟨a′′ | A† = ⟨a′′ | A,

since A is Hermitian.
Multiply now the first equation in (2.23) by ⟨a′′ | from the left and the
second by |a′ ⟩ from the right, taking into account that A is Hermitian. This
gives

⟨a′′ |A|a′ ⟩ = a′ ⟨a′′ |a′ ⟩ ,


⟨a′′ |A|a′ ⟩ = a′′∗ ⟨a′′ |a′ ⟩ .

Taking the difference yields

0 = (a′ − a′′∗ ) ⟨a′′ |a′ ⟩ .

If a′′ = a′ , then ⟨a′′ |a′ ⟩ = 1, and thus a′ = a′′∗ = a′∗ and thus a′ is real. Hence
the eigenvalues of a Hermitian operator A are real.

16
If, in contrast, a′′ ̸= a′ , then a′ − a′′∗ = a′ − a′′ ̸= 0 and hence ⟨a′′ |a′ ⟩ = 0
or, in other words, |a′′ ⟩ and |a′ ⟩ are orthogonal. The eigenvectors, which
corresponds to different eigenvalues of the same Hermitian operator A are
orthogonal!

Operator multiplication and outer product


Two operators might act onto a state one after the other (this corresponds
to two subsequent measurements). In general, operators do not commute,
i.e., if X and Y are two operators, then in general

XY |a⟩ ̸= Y X |a⟩ .
You can rationalise this fact by considering a real problem. You are familiar
with the Heisenberg position-momentum uncertainty principle, which states
that ∆x∆p ? h̄/2. This implies that first measuring the momentum and
then the position of a quantum mechanical particle does not necessarily give
the same measurement result as first measuring the position and then the
momentum - x and p do not commute. You should also realise that the order
of operators (in ket space), which corresponds to the order of measurements,
is from right to left. The rightmost operator acts first onto the state.
We have already considered operators in bra space and stated that

X |a⟩ corresponds to ⟨a| X † .


From mathematics we know that the Hermitian adjoint of an operator prod-
uct is given by

(XY )† = Y † X † .
This also makes sense physically, since operators in bra space act to the left.
One can build operators from a product of states. The outer product of
two states |a⟩ and |b⟩ is defined as

X = |b⟩⟨a|;
forming such an outer product results in the formation of an operator, which
here is called X. The Hermitian adjoint of this operator is

X † = |a⟩⟨b|.
Finally, we have already encountered matrix elements (cf. Table 1.1) of
the form ⟨b|X|a⟩ (you will soon see why these constructs are called matrix
elements). For these the following relationship is valid:

17
⟨ ⟩∗
⟨b|X|a⟩ = a X † b .

Of course, if X is Hermitian, i.e., if X is an observable, then ⟨b|X|a⟩ =


⟨a|X|b⟩∗ .

2.2.7 Bases of state space


In section 2.2.4 we already discussed briefly that the state space of interest
is spanned by the eigenvectors of the observable of interest. For example,
in the Stern-Gerlach experiment the spin state space is two-dimensional and
spanned by |Sz , +⟩ and |Sz , −⟩, which are the eigenvectors of the observable
Sz . After an Sz measurement the silver atom is either in state |Sz , +⟩ or
in state |Sz , −⟩. |Sz , +⟩ cannot be described in terms of |Sz , −⟩ and vice
versa; |Sz , +⟩ and |Sz , −⟩ are linearly independent, which implies that they
are orthogonal: ⟨Sz , +|Sz , −⟩ = 0. |Sz , +⟩ and |Sz , −⟩ form a basis of the
spin state space.
More generally, if we call the basis vector of the state space of interest
|a′ ⟩, |a′′ ⟩, . . ., and if we normalise them,
′ ′′
|a′N ⟩ = √|a ′⟩ |a′′N ⟩ = √ |a′′ ⟩ ′′ ...,
⟨a |a′ ⟩ ⟨a |a ⟩

then |a′ ⟩, |a′′ ⟩, . . . form a complete orthonormal basis of state space.


Thus an arbitrary normalised state |α⟩ in state space can be written as a
sum of basis vectors, which here are assumed to be already normalised (i.e.,
we skip the index N ):

|α⟩ = ca′ |a′ ⟩. (2.24)
a′

The coefficients ca′ are retrieved by multiplying equation (2.24) with ⟨a′ | from
the left:

⟨a′ |α⟩ = ca′ .

|α⟩ can thus also be written as


∑ ∑
|α⟩ = ⟨a′ |α⟩ |a′ ⟩ = |a′ ⟩ ⟨a′ |α⟩.
a′ a′

This in turn implies the very important relationship that

18

|a′ ⟩⟨a′ | = 1 (2.25)
a′

(2.25) is called the closure or completeness relation (SV: fullständighets-


relationen). The relation expresses the completeness of the basis {|a′ ⟩} (what
does this imply for an arbitrary vector on the quantum mechanical state
space?). The meaning of the coefficients ca′ = ⟨a′ |α⟩ is that they give the
probability amplitude for finding the system described by |α⟩ to be in state
|a′ ⟩, i.e., |ca′ |2 is the corresponding probability.
We also can assign a meaning to the outer product |a′ ⟩⟨a′ |: it projects
the state it acts on (it is an operator!) onto the state |a′ ⟩. |a′ ⟩⟨a′ | is said to
be a projector :
∑ ∑
|a′ ⟩ ⟨a′ |α⟩ = |a′ ⟩ ⟨a′ |a′′ ⟩ ⟨a′′ |α⟩ = |a′ ⟩ δa′ a′′ ca′′ = ca′ |a′ ⟩ .
a′′ a′′

2.2.8 Matrix representations


Of course one always can choose different bases for the state space of interest.
For example, one basis for the spin space of the Stern-Gerlach experiment is
the {|Sz , +⟩ , |Sz , −⟩} basis. Other choices would
⟩be, e.g., {|Sx , +⟩ , |Sx , −⟩}
(n)
or {|Sy , +⟩ , |Sz , −⟩}. For a particular basis { a }, with n = 1 . . . N , oper-
ators and vectors can be written in matrix notation2 :

 ⟨ ⟩ ⟨ ⟩ ⟨ ⟩ 

a(1) X a(1) a(1) X a(2) . . . a(1) X a(N )
 ⟨ ⟩ ⟨ ⟩ ⟨ ⟩ 

 a(2) X a(1) a(2) X a(2) . . . a(2) X a(N ) 

ˆ
X= 
,

(2.26)
 ⟨ . . . ⟩ ... ... ⟨ . . . ⟩ 
(N ) (1) (N ) (N )
a X a ... . . . a X a
 ⟨ ⟩ 

a(1) α
 ⟨ ⟩ 

 a(2) α 

|α⟩ =
ˆ
,

(2.27)
 ⟨ . . . ⟩ 

a(N ) α
(⟨ ⟩∗ ⟨ ⟩∗ ⟨ ⟩∗ )
⟨α| =
ˆ a(1) α , a(2) α , . . . , a(N ) α . (2.28)
Equation (2.26) shows clearly why the construct ⟨b|X|a⟩ is called a matrix
element.
2
You should read the sign =
ˆ as ”corresponds to”.

19
Observe that in equation (2.26) to (2.28) above there is no equality be-
tween what is to the left and the right of the ”=”
ˆ sign – this is a correspon-
dence. If you want to write an equation that corresponds to equation (2.26)
then you should apply the closure relation two times:
∑∑
X= |a′′ ⟩ ⟨a′′ |X|a′ ⟩ ⟨a′ |. (2.29)
a′′ a′

Exercise: Matrix notation vs Dirac notation




1. Write the operator |β⟩⟨α| in matrix notation using the basis { a(n) }.

2. Consider, again, the electronic spin. If we define |+⟩ ≡ |Sz , +⟩ and


|−⟩ ≡ |Sz , −⟩ and knowing that Sz |±⟩ = ± h̄2 |±⟩ write the matrix/vector
notations for |+⟩, |−⟩, Sz , S+ ≡ h̄|+⟩⟨−|, and S− ≡ h̄|−⟩⟨+|.

2.2.9 The position operator and the position basis of


state space
The action of the position operator is defined by

x |x′ ⟩ = x′ |x′ ⟩ , (2.30)


where x is an operator (actually, it is an observable) and x′ a scalar – the
outcome of the position measurement represented by x. Typically, the possi-
ble values of x′ are infinitesimally closely spaced. Just think of a free particle
anywhere in space.
Since x is an observable its eigenvectors must span a subspace of the
quantum mechanical state space. This is the position subspace. Since the
eigenvalues are infinitesimally closely spaced, there must exist an indefinite
number of them, and hence the vectors |x′ ⟩ cannot form a discrete basis, but
the basis must be continuous and the number of dimensions of the position
subspace is infinite. Therefore, the sum in the closure relation must be
replaced by an integral:
∑ ∫ ∞
′ ′
|a ⟩⟨a | = 1 −→ dx′ |x′ ⟩⟨x′ | = 1. (2.31)
a′ −∞

An arbitrary state |α⟩ can then be expanded as


∫ ∞
|α⟩ = dx′ |x′ ⟩ ⟨x′ |α⟩, (2.32)
−∞

or, in three dimensions:

20

|α⟩ = d3 x′ |x′ ⟩ ⟨x′ |α⟩. (2.33)
If you compare this to the discrete expansion
∑ ∑
|α⟩ = |a′ ⟩ ⟨a′ |α⟩ = ca′ |a′ ⟩,
a′ a′

then you see that ⟨x′ |α⟩ must be a probability amplitude, as are the coef-
ficients ca′ . From before you know that probability amplitudes related to a
position measurement are given by the wave functions, and we can identify

ψ(x′ ) = ⟨x′ |α⟩ . (2.34)

Scalar products of position states


For discrete bases we found for the basis states that

⟨a′ |a′′ ⟩ = δa′ a′′ ,

where δa′ a′′ is the Kronecker delta. For a continuous basis such as the position
basis things are bit more complicated:

⟨x′ |x′′ ⟩ = δ(x′ − x′′ ), (2.35)


which looks very similar to the expression for the discrete basis. However,
δ(x′ − x′′ ) is the delta function, a generalised function with the following two
properties:
∫ ∞
dx δ(x − x0 )f (x) = f (x0 ), (2.36)
−∞
∫ ∫ ∞
∞ dδ(x) df (x) df (x)
dx f (x) = − dx δ(x) =− . (2.37)
−∞ dx −∞ dx dx x=0
The latter property further implies that
dδ(−x) dδ(x)
=− ,
dx dx
dδ(−x)
x = −δ(x).
dx
If the delta-function has a vector argument, this implies that it represents
the product of the delta functions of the vector components:

δ(x − x0 ) = δ(x − x0 )δ(y − y0 )δ(z − z0 ). (2.38)

21
2.2.10 Comparison of relations for a discrete basis and
for a continuous basis
The following list compares the expressions for a discrete basis and a con-
tinuous basis (here the position basis has been chosen, but it could be any
continuous basis, e.g., the momentum basis):

• Orthonormality: ⟨a′ |a′′ ⟩ = δa′ a′′ ←→ ⟨x′ |x′′ ⟩ = δ(x′ − x′′ ),


∑ ∫
• Closure relation: a′ |a′ ⟩⟨a′ | = 1 ←→ dx′ |x′ ⟩⟨x′ | = 1,

• Expansion of an arbitrary state: ∫



|α⟩ = a′ |a′ ⟩ ⟨a′ |α⟩ ←→ |α⟩ = dx′ |x′ ⟩ ⟨x′ |α⟩,
∑ ∫
• Normalisation: a′ |⟨a′ |α⟩|2 = 1 ←→ dx′ |⟨x′ |α⟩|2 = 1,

• Scalar product of arbitrary states: ∫



⟨β|α⟩ = a′ ⟨β|a′ ⟩ ⟨a′ |α⟩ ←→ ⟨β|α⟩ = dx′ ⟨β|x′ ⟩ ⟨x′ |α⟩,
{ ⟩}

• Matrix elements of a matrix A with eigenstates a(i) :
⟨ ⟩
(j)
a A a(i) = a(i) δij ←→ ⟨x′′ |x|x′ ⟩ = x′ δ(x′ − x′′ ).

Question: Can you think of any observables (other than the position), which
typically would have a continuous spectrum of eigenvalues?

Task : Using equation (2.34), rewrite the closure relation for a discrete basis
in wave function notation. Give a physical interpretation of the outcome.

2.3 The spin vectors and spin operators


2.3.1 Explicit construction
Already in section 2.1.1 we wrote down expressions for the states |Sz , ±⟩,
|Sx , ±⟩, and |Sy , ±⟩. While we at that time postulated what they have to
look like, we will now construct them explicitly. The same we will do for the
spin operators Sx , Sy , and Sz .
We start by again considering the third of the sequential Stern-Gerlach
experiments in Figure 2.2. Since the outcome of this experiment clearly
shows that the silver atoms coming out of the Stern-Gerlach analyser with
the magnetic field in the x-direction can be in both the |Sz , +⟩ and |Sz , −⟩

22
states, and since the probability for both of these possibilities is 1/2 we can
write
1
|⟨Sz , +|Sx , +⟩|2 = |⟨Sz , −|Sx , +⟩|2 = ,
2
or, equivalently,
1
|⟨Sz , +|Sx , +⟩| = |⟨Sz , −|Sx , +⟩| = √ . (2.39)
2
From equation (2.39) we get the normalisation factors for expressing the
states |Sx , +⟩ and |Sx , −⟩ in terms of the states |Sz , ±⟩. By convention we
furthermore set the complex phase factor of the |Sz , +⟩ state in the expres-
sions for |Sx , +⟩ to 1, while the complex phase factor of the |Sz , −⟩ is yet to
be determined:
1 1
|Sx , +⟩ = √ |Sz , +⟩ + √ eiδ1 |Sz , −⟩ . (2.40)
2 2
Since we know that, necessarily, ⟨Sx , −|Sx , +⟩=0 (why?) we can easily obtain
an expression for |Sx , −⟩. Suppose that

A B
|Sx , −⟩ = √ |Sz , +⟩ + √ |Sz , −⟩ ,
2 2

where A and B are complex constants (how do we know that this a valid
statement?). Again, by convention, A is set to 1, and only B has to be
determined from equation (2.40). Since

1 B∗
⟨Sx , −| = √ ⟨Sz , +| + √ ⟨Sz , −| ,
2 2
and thus
⟨Sx , −|Sx , +⟩ = 21 ⟨Sz , +|Sz , +⟩ + 1
2
eiδ1 ⟨Sz , +|Sz , −⟩
∗ B∗
+ B2 ⟨Sz , −|Sz , +⟩ + 2
eiδ1 ⟨Sz , −|Sz , −⟩ ,

and since furthermore ⟨Sz , +|Sz , +⟩ = 1, ⟨Sz , −|Sz , −⟩ = 1, and ⟨Sz , −|Sz , +⟩ =
⟨Sz , +|Sz , −⟩ = 0, we find B ∗ = −e−iδ1 and thus
1 1
|Sx , −⟩ = √ |Sz , +⟩ − √ eiδ1 |Sz , −⟩ . (2.41)
2 2
In the same way expressions are obtained for |Sy , ±⟩:

23
1 1
|Sx , ±⟩ = √ |Sz , +⟩ ± √ eiδ2 |Sz , −⟩ . (2.42)
2 2
δ1 and δ2 can now be determined from the scalar product of |Sy , ±⟩ with
|Sx , ±⟩, which by symmetry can be determined from equation (2.39):
1
|⟨Sy , ±|Sx , ±⟩| = √ .
2
Putting in the expressions for |Sx , ±⟩ and |Sy , ±⟩, equations (2.41) and (2.42),
one obtains
1
1
1 ± ei(δ1 −δ2 ) = √ ,
2 2
i.e., δ2 − δ1 = ±π/2. We choose δ1 = 0 and δ2 = π/2 and thus arrive at
1 1
|Sx , ±⟩ = √ |Sz , +⟩ ± √ |Sz , −⟩ , (2.43)
2 2
1 i
|Sy , ±⟩ = √ |Sz , +⟩ ± √ |Sz , −⟩ . (2.44)
2 2
This means that the expressions for the components of the spin in the x and
y directions now have been fully expressed in terms of the z-component of
the spin.
We still have to write the observables Sx , Sy , and Sz in terms of the z-
component of the spin, since such a formulation will tell us exactly how these
observables act on the different spin states. This is straightforwardly done
by using the fact that any observable can be written in terms of a sum of
projectors onto the eigenstates of the observable. This is seen by applying
the closure relation (2.25) two times to an observable A:
∑∑ ∑
A= |a′′ ⟩ ⟨a′′ |A|a′ ⟩ ⟨a′ | = a′ |a′ ⟩⟨a′ |, (2.45)
a′ a′′ a′

since ⟨a′′ |A|a′ ⟩ = a′ δa′ a′′ . The eigenvalues of Sz (and likewise of Sx and
Sy ) are known from experiment to be ±h̄/2. Since the eigenstates of Sz are
|Sz , +⟩ and |Sz , −⟩ one readily obtains

Sz = {|Sz , +⟩⟨Sz , +| − |Sz , −⟩⟨Sz , −|} . (2.46)
2
The expressions for Sx and Sy are obtained in the same way starting from

Sx,y = {|Sx,y , +⟩⟨Sx,y , +| − |Sx,y , −⟩⟨Sx,y , −|} ,
2
24
and then putting in the expressions for |Sx , ±⟩ and |Sy , ±⟩, equations (2.43)
and (2.44), respectively. This gives


Sx = {|Sz , +⟩⟨Sz , −| + |Sz , −⟩⟨Sz , +|} , (2.47)
2


Sy = {−i|Sz , +⟩⟨Sz , −| + i|Sz , −⟩⟨Sz , +|} . (2.48)
2

2.3.2 The commutators and anticommutators of the


spin operators
Definition of commutators and anticommutators
We will see shortly that the commutators of observables play an immensely
important role in quantum mechanics. The commutator [A, B] of two oper-
ators A and B is defined as

[A, B] ≡ AB − BA. (2.49)


Likewise, the anticommutator {A, B} is defined as

{A, B} ≡ AB + BA. (2.50)

Exercise
It is now your task to show that

[Si , Sj ] = iεijk h̄Sk , (2.51)


where


 +1 if the ijk are an even permutation
εijk ≡ −1 if the ijk are an odd permutation (2.52)


0 otherwise
is the so-called Levi-Civita symbol. Observe that the ijk are ordered and
that they can assume the values 1,2, and 3, which stand for x, y, and z.
Show also that

1 2
{Si , Sj } = h̄ δij . (2.53)
2
Useful relations are

25
{A, B} = [A, B] + 2BA,

{A, B} = {B, A} .

Relation (2.52) is fundamental in the theory of angular momentum: it is


the defining relation for an angular momentum. Moreover, it expresses that
the components of the angular momentum fulfil an uncertainty relationship.

The square of the spin angular momentum


Another very important spin operator is the square of the spin S 2 . It is the
sum of the squares of the spin components:

S 2 ≡ Sx2 + Sy2 + Sz2 .

The squaring of the operators Sx , Sy , and Sz expresses that they are ap-
plied twice: Sx2 |α⟩ = Sx Sx |α⟩. For example, with |α⟩ = |Sz , +⟩ one gets
Sx2 |Sz , +⟩ = Sx Sx |Sz , +⟩ = Sx h̄/2 |Sz , +⟩ = h̄2 /4 |Sz , +⟩.
It is important that you realise that S 2 = S · S is a scalar product on
cartesian space. It is not a scalar product on the quantum mechanical state
space!
Since {Si , Sj } = 21 h̄2 δij we find that

3
S 2 = h̄2 1, (2.54)
4
where 1 is the unit operator. Often the unit operator is not written out so
that it is written S 2 = 34 h̄2 , but of course since the left-hand side of the
equation is an operator, also the right-hand side must be an operator.
Another important property of S 2 is that it commutes with every com-
ponent of the spin:
[ ]
S 2 , Si = 0. (2.55)
This is easily shown when recognising that

[A, B + C] = [A, B] + [A, C]

for any operators A, B, and C, and then using the definition of the commu-
tator.

26
2.4 Commuting (Compatible) observables
2.4.1 The case of non-degenerate eigenstates
We have just encountered an example of two observables which commute,
namely S 2 and, e.g., Sz . The compatability of the two observables has a very
important implication for their eigenstates, and that is what we will show
now.
Let us assume that we have two observables A and B which commute,
i.e. [A, B] = 0. Furthermore, let us assume that the eigenstates of A are
non-degenerate, i.e., to each eigenstate of A there exists a unique eigenvalue
a and vice versa. If we now label the eigenvalues of A as previously by a′ ,
a′′ , . . . then

⟨a′′ |[A, B]|a′ ⟩ = (a′′ − a′ ) ⟨a′′ |B|a′ ⟩ = 0.

If a′′ ̸= a′ this must mean that ⟨a′′ |B|a′ ⟩ = 0, or, in other words, the matrix
elements of B in the basis {|a′ ⟩} are diagonal. This implies that the states
{|a′ ⟩} are eigenstates of B, as well, as is shown by applying the closure
relation to B two times:
∑∑ ∑
B= |a′′ ⟩ ⟨a′′ |B|a′ ⟩ ⟨a′ | = |a′ ⟩ ⟨a′ |B|a′ ⟩ ⟨a′ | ,
a′ a′′ a′

which is multiplied by |a′ ⟩ from the right to give

B |a′ ⟩ = ⟨a′ |B|a′ ⟩ |a′ ⟩ . (2.56)


This is a very important result that shows that if [A, B] = 0 then any eigen-
state |a′ ⟩ of A is also an eigenstate of B with eigenvalue ⟨a′ |B|a′ ⟩.

2.4.2 Degeneracy
Two (or more) eigenstates of an observable A are said to be degenerate if the
corresponding eigenvalues are the same:
⟩ ⟩

A a′(i) = a′ a′(i) ,

with i larger than 1. For example the eigenstates of the energy operator,
the Hamiltonian, for the hydrogen atom are degenerate, as is illustrated
in Figure 2.3. The ”problem” is now that just specifying the energy of a
particular state is not enough for exactly specifying which state the hydrogen
atom is in. Further labels are needed (and shown in the Figure).

27
E

n=3 3s 3p 3d

n=2 2s 2p

1s
n=1

Figure 2.3: Energy levels of the non-relativistic hydrogen atom. n is the


principal quantum number which ”labels” the energy; s, p, and d are the
labels for the angular momentum quantum number. The degeneracy of the
levels of a particular energy is 2n2 -fold, i.e., there exist 2n2 states of the same
energy.

2.4.3 Commuting observables with degenerate eigen-


states
These additional labels are provided by the eigenvalues of further observ-
ables. It can be shown that if A is an observable with degenerate eigenstates
and if it commutes with an observable B, i.e., [A, B] = 0, then there exist
simultaneous eigenstates of A and B, very much as in the non-degenerate
case. Then

A |a′ , b′ ⟩ = a′ |a′ , b′ ⟩ ,
(2.57)
B |a′ , b′ ⟩ = b′ |a′ , b′ ⟩ .
Note that the common eigenstates of A and B now were labelled by the
corresponding eigenvalues for both A and B.

Example
As we will see later during the course, the orbital angular momentum squared
L2 is an observables with eigenvalues h̄2 l(l + 1), where l = 1 . . . n is the
angular momentum quantum number. Each eigenstate of L2 is (2l + 1)-fold

28
degenerate. The necessary additional label is provided by the eigenvalues of
the z-component of the angular momentum Lz , which are ml h̄, with ml =
−l, −l + 1, . . . , +l.

Complete sets of compatible observables


In the general case one needs to find a maximum set of commuting observ-
ables

[A, B] = [B, C] = [A, C] = . . . = 0

to fully describe the system. The eigenstates are then labelled by the eigen-
values of the different observables:
A |a′ , b′ , c′ , . . .⟩ = a′ |a′ , b′ , c′ , . . .⟩ ,

B |a′ , b′ , c′ , . . .⟩ = b′ |a′ , b′ , c′ , . . .⟩ ,

C |a′ , b′ , c′ , . . .⟩ = c′ |a′ , b′ , c′ , . . .⟩ .

...

The scalar product of two such states is then ⟨a′ , b′ , c′ , . . . |a′′ , b′′ , c′′ , . . . ⟩ =
δa′ a′′ δb′ b′′ δc′ c′′ . . ., and the closure relation is written
∑∑∑
. . . |a′ , b′ , c′ , . . . ⟩⟨a′ , b′ , c′ , . . . | = 1. (2.58)
a′ b′ c′
What happens now if one measures the properties corresponding to two
commuting observables A and B? If all eigenstates of A and B are non-
degenerate, then the measurement of an eigenvalue of A determines the cor-
responding eigenvalue of B, as well (−→A depicts a measurement of A):

|α⟩ −→A |a′ , b′ ⟩ −→B |a′ , b′ ⟩ −→A |a′ , b′ ⟩ .

This is different if A and B are degenerate. Then the determination of


A does not specify B, i.e., the state after measurement of A is still in a
superposition of all possible eigenstates of B:

n ⟩ ⟩ ⟩

′(j) ′(i)
|α⟩ −→A c(i)
a a ,b −→B a′(j) , b′(k) −→A a′(j) , b′(k) ,
i=1

where j and k are fix labels, but i is a running label. From the last measure-
ment you also see that A and B can be determined simultaneously if (and
only if ) A and B commute.

29
2.5 Translation, momentum, and time evolu-
tion
In principle, we now have the main tools for being able to proceed study-
ing orbital angular momentum, with the hydrogen atom as the primary ex-
ample. It is quite interesting, though, that the fundamental equation, the
Schrödinger equation, as well as important relationships such as the canon-
ical commutator relations can be derived from quite simple principles. Also
the standard prescription of replacing momentum p by the ∇ operation when
acting on a (position-dependent) wave function can be derived quite easily.
Therefore, we here make a little detour to explore these things.
What we will need to do to derive the relationships for the components of
x and p is to consider translations. We will start with infinitesimal transla-
tions and then consider finite ones. Since the order of position measurement
and translation matters (i.e., whether the system first is translated and the
position measured then, or vice versa), we will see that the corresponding
operators do not commute. Translation and momentum are connected, in
that momentum generates translation.
To derive the Schrödinger equation we will have to consider the time
evolution of quantum systems. This will leads us to the Schrödinger equation
in a straightforward manner, but we will still have to postulate that it is the
Hamiltonian which governs time evolution.

Question: What is a translation?

2.5.1 Infinitesimal translations


We will start out with infinitesimal translations, i.e., by translations by an
infinitesimal distance dx′ . If the initial state of the system before the trans-
lation is characterised by a position state |x′ ⟩, then after the infinitesimal
translation it is
T (dx′ ) |x′ ⟩ = |x′ + dx′ ⟩ .
The operator T (dx′ ) |x′ ⟩ is called the infinitesimal translation operator. If
T (dx′ ) |x′ ⟩ is applied to a general state |α⟩, then one finds the following:

T (dx′ ) |α⟩ = T (dx′ ) d3 x′ |x′ ⟩ ⟨x′ |α⟩ =
∫ ∫
= d3 x′ |x′ + dx′ ⟩ ⟨x′ |α⟩ = d3 x′ |x′ ⟩ ⟨x′ − dx′ |α⟩ = (2.59)

= d3 x′ |x′ ⟩ ψα (x′ − dx′ ).

30
In the first step we used the closure relation. The last step is due to a change
of variable from x′ to x′ − dx′ . This does not affect the limits of integration,
which are ±∞, and neither does it affect the volume element. You see here,
that the consideration of the translation implies that we have to consider the
wave function of the translated state ψα (x′ − dx′ ).
It seems that we do not know very much about the translation operator
T (dx′ ). However, we can require a number of properties:

• We require that if |α⟩ is normalised, then
⟨ also T (dx ) |α⟩ needs
⟩ to
† ′ ′
be normalised (why?), i.e. 1 = ⟨α|α⟩ = α T (dx ) T (dx ) α . This
implies that T (dx′ ) is unitary:

T † (dx′ )T (dx′ ) = 1 ⇔ T † (dx′ ) = T −1 (dx′ ).

• It should be possible to combine subsequent translations to a single


translation:

T (dx′′ ) T (dx′ ) = T (dx′ + dx′′ ) (observe that the first translation


carried out is that to the right by dx′ )

• Reverse translations are the same as translations in the negative direc-


tion:

T (−dx′ ) = T −1 (dx′ )

• Zero translations amount to the unity operator (no change):

limdx′ →0 T (dx′ ) = 1.
These four requirements are fulfilled by

T (dx′ ) = 1 − i K·dx′ , (2.60)

where K stands for the triple of Hermitian operators (Kx , Ky , Kz ). Note


that the · marks the scalar product in 3d space (not in quantum mechanical
state space!).

Task : Show that the infinitesimal translation operator (2.60) fulfills all re-
quirements!

Question: What do you think is the physical meaning of K?

31
Relationship between x and K
We now consider the order of position measurement and translation. If the
translation is carried out first, one finds

x T (dx′ ) |x′ ⟩ = (x′ + dx′ ) |x′ + dx′ ⟩ . (2.61)

If one, instead, first carries out the position measurement then one arrives at

T (dx′ )x |x′ ⟩ = x′ |x′ + dx′ ⟩ . (2.62)

Obviously, and not very surprising, the operators x and T (dx′ ) do not com-
mute – the order matters. For the commutator we find

[x, T (dx′ )] |x′ ⟩ = dx′ |x′ + dx′ ⟩ ≃ dx′ |x′ ⟩ . (2.63)

Since the relation is valid for an arbitrary state |x′ ⟩, we can identify

[x, T (dx′ )] = dx′ . (2.64)

Putting in the definition of T (dx′ ) (equation (2.60)) leads to

−ix(K·dx′ ) + i(K·dx′ )x = dx′ .

You can now choose dx′ to be oriented along a vector xj , i.e., dx′ ≡ dx′j =
dx′ x̂j , where x̂j is the unit vector along that direction. Multiplying the
equation by a vector dx′i = dx′i x̂i , which is chosen to be normal to x̂j (i.e.,
dx′j ·dx′i = δij ), and then dividing by dx′i dx′j leads to

[xi , Kj ] = iδij . (2.65)

The canonical commutator relations


We have already said that momentum generates translation. Since we have
seen that K generates an infinitesimal translation, there must exist a linear
relationship between both quantities:

K = const · p. (2.66)

From a comparison to de Broglie’s relation


2π p
=
λ h
and by identifying k with the wave vector,

|k| = ,
λ
32
we see that the constant in (2.66) must be given by 1/h̄. For the infinitesimal
translation operator (2.60) this results in

p·dx′
T (dx′ ) = 1 − i , (2.67)

and relation (2.65) turns into

[xi , pj ] = ih̄δij . (2.68)

This is one of the canonical commutator relations for position and momen-
tum, that you have talked about in FYSA21! Consider once more how we
have arrived at it: it follows from the fact that it matters in which order
one executes an infinitesimal translation and a position measurement! Since
(2.68) is another formulation of the uncertainty relation for position and
momentum
h̄2
⟨(∆x)2 ⟩⟨(∆p)2 ⟩ ? ,
4
this uncertainty relation is related to the fact that the order of translation
and position measurement matters!!
As you know there are two other canonical commutator relations, namely
that for the components of the position and that for the components of the
momentum:
[xi , xj ] = 0,
(2.69)
[pi , pj ] = 0.
The former of the these two relations derives from the fact that all three
components of the position vector can be determined simultaneously to an
arbitrary degree of accuracy; the latter from the fact that the order of subse-
quent translations does not matter for the result of the total translation. We
do not show this here formally, but you should remember all three canonical
commutator relations (2.68) and (2.69).

2.5.2 Finite translations


In the preceding section we considered infinitesimal translations. We now di-
rect our interest at finite translations of length ∆x′ . Such a finite translations
is achieved by combining N infinitesimal translation of length dx′ = ∆x′ /N
and letting N go to infinity. Without any loss of generality, we also as-
sume that ∆x′ is parallel to x̂′ . The action of the finite translation operator

33
T (∆x′ ) is then given by
( ) ( ) ( )
∆x′ ∆ x′ ∆x′
T (∆x′ ) |x′ ⟩ = limN →∞ T N
T N
... T N
|x′ ⟩ =
( ) ( )
(2.70)
ipx ∆x′ N ′
= limN →∞ 1 − N h̄
= exp −i px ∆x

.

In this equation px is an operator! To understand what the meaning of a


function of an operator is, consider the Taylor expansion of the operator. In
our case, if A is an operator, then exp A corresponds to
A2
exp A = 1 + A + + ...,
4
i.e., it is the identity operator plus A applied once plus A applied twice, etc.

2.5.3 Momentum operator in the position basis


The aim of this section is to derive the standard prescription of replacing
the momentum operator p by the operator −ih̄∇ when dealing with wave
functions. We will do this here in one dimension, but the result is readily
generalised to three dimensions. Consider the action of the finite translation
operator onto a general state |α⟩:

′ px ∆x′
T (∆x ) |α⟩ = (1 − i ) |α⟩ . (2.71)

The left-hand side can be rewritten using the closure relation for the position
basis:
∫ ∫
T (∆x′ ) |α⟩ = dx′ T (∆x′ ) |x′ ⟩ ⟨x′ |α⟩ = dx′ |x′ + ∆x′ ⟩ ⟨x′ |α⟩ =
∫ ∫ ( )
= dx′ |x′ ⟩ ⟨x′ − ∆x′ |α⟩ = dx′ |x′ ⟩ ⟨x′ |α⟩ − ∆x′ ∂x∂ ′ ⟨x′ |α⟩ (2.72)

= |α⟩ − dx′ |x′ ⟩ ∆x′ ∂x∂ ′ ⟨x′ |α⟩.
In this derivation use has been made of
∂ ′ ∆ψα ψα (x′ − ∆x′ ) − ψα (x′ )
− ψα (x ) = − = .
∂x′ ∆x′ ∆x′
Combining equations (2.71) and (2.72) provides an expression for the action
of the operator px onto a state |α⟩:
∫ ( )

px |α⟩ = dx |x ⟩ −ih̄ ′ ⟨x′ |α⟩ .
′ ′
(2.73)
∂x

34
Now we consider the scalar product of another position state |x′′ ⟩ with
px |x′ ⟩ and the scalar product of an arbitrary state |β⟩ with px |x′ ⟩: Using
⟨x′′ |x′ ⟩ = δ(x′′ − x′ ) and ⟨x′ |α⟩ = ψα (x′ ) we find


⟨x′′ |px |α⟩ = −ih̄ ⟨x′′ |α⟩ (2.74)
∂x′′
and ∫

⟨β|px |α⟩ = )ψα (x′ ).
dx′ ψβ∗ (x′ )(−ih̄ (2.75)
∂x′
These two equations together correspond to the prescription

px ←→ − ih̄ , (2.76)
∂x′
for the action of the momentum operator onto wave functions. You have
encountered this prescription already in FYSA21. Here this prescription
has been derived from the translation properties of momentum. In three
dimensions, the prescription is

p ←→ − ih̄∇′ . (2.77)

2.5.4 Dynamics in quantum systems


The term dynamics always refers to a time evolution. The time evolution
of a quantum mechanical state can be put into an operator, very much in
the same way as we put the translation of a position state in a translation
operator. So we ask ourselves how a state |α⟩ ≡ |α t0 ⟩ at time t0 will evolve
in time. At later times that same state is denoted by |α t t0 ⟩, where the
specification of t0 represents the boundary condition that |α t0 ⟩ and |α t t0 ⟩
are identical at t0 . We introduce the time evolution operator U (t, t0 ) which
takes the state |α t0 ⟩ to |α t t0 ⟩:

|α t t0 ⟩ = U (t, t0 ) |α t0 ⟩ . (2.78)

As for the infinitesimal translation operator we require the following:

• the conversation of probability, expressed by that both the state at t0


and the time-evolved state have to be normalised:
⟨ ⟩

⟨α t t0 |α t t0 ⟩ = α t0 U † (t, t0 )U (t, t0 ) α t0 = 1, from which follows

U † (t, t0 )U (t, t0 ) = 1, i.e., U (t, t0 ) is unitary,

35
• that subsequent time evolutions can be combined to a single time evo-
lution (t2 > t1 > t0 ) (observe that U (t1 , t0 ) acts prior to U (t2 , t1 )):

U (t2 , t0 ) = U (t2 , t1 )U (t1 , t0 ),


• that infinitesimal time evolutions tend toward the unity operator:

limdt→0 U (t0 + dt, t0 ) = 1.


A solution for the infinitesimal time evolution operator is readily found (re-
member the solution for the infinitesimal translation operator):
H dt
U (t0 + dt, t0 ) = 1 − i
, (2.79)

where H is an Hermitian operator (which will turn out to be the Hamilto-
nian).
Consider now two subsequent evolutions from t0 to t and then further to
t + dt:
U (t + dt, t0 ) = U (t + dt, t)U (t, t0 ) = (1 − i Hh̄dt )U (t, t0 )

U (t + dt, t0 ) − U (t, t0 ) = −i Hh̄ dtU (t, t0 ),


or in differential form:

ih̄U (t, t0 ) = HU (t, t0 ). (2.80)
∂t
This is the Schrödinger equation, although it might look oddly unfamiliar to
you. The normal formula is obtained by multiplying by |α t0 ⟩:

ih̄ U (t, t0 ) |α t0 ⟩ = HU (t, t0 ) |α t0 ⟩ ⇔
∂t

ih̄ |α t t0 ⟩ = H |α t t0 ⟩ . (2.81)
∂t
We have derived the Schrödinger equation by just considering the time
evolution of a physical state. Unfortunately, we a priori do not know very
much about the operator H. Actually, that H is the classical Hamiltonian
needs to be postulated. But it makes sense: the Hamiltonian has the right
dimension, namely [energy], which is required by the form of the infinitesimal
time evolution operator (2.79), since the dimension of h̄ is [energy × time].
Furthermore, also in classical mechanics the Hamiltonian is the generator
of time evolution. So it makes sense to postulate that the same is true in
quantum mechanics.

36
2.6 Where do we stand now?
At this point it seems appropriate to take a step back and look at what has
been achieved. We have introduced the quantum mechanical state space and
vectors and operators on that space. We have seen the connection to wave
functions, and the connection between observables, which are Hermitian op-
erators, and measurements has been discussed. Spin has been introduced,
together with the spin states and spin operators. We have discussed compat-
ible (commuting) observables and seen that the eigenvalues to the common
eigenvectors of observables with degenerate spectra can be used to label the
eigenvectors. Finally, we have considered translation, which led to the canon-
ical commutator relations for position and momentum, and time evolution,
which resulted in the time-dependent Schrödinger equation.
In the following the aim is to discuss another angular momentum, namely
orbital angular momentum. Orbital angular momentum can be cast into the
same picture as spin angular momentum and follows the same set of rules
as spin. Its handling in wave function notation is, however, a bit more
complicated than spin angular momentum. Since it is advantageous to work
with a real problem, we will start out with the hydrogen atom, which is a two-
body system and therefore can be described in terms of a central potential.
Orbital angular momentum plays a central role in the discussion.

37
Chapter 3

The hydrogen atom

3.1 The hydrogen atom in classical and semi-


classical physics

The complete description of the hydrogen atom is the most basic problem
of Atomic Physics. Here, the starting point is the classical description of
the hydrogen atom. Already in the 19th century it became clear that the
experimental observations did not agree with any classical solution, however,
and therefore Niels Bohr introduced a new, phenomenological model, now
known as the Bohr Model, which could explain some of the observations.
The Bohr Model was one of the highlights of early quantum mechanics, but
was soon replaced by a better, truly quantum mechanical model, which we
will treat in detail.

3.1.1 Statement of the problem

The hydrogen atom is the smallest of all atoms. It contains one electron
and a proton, which are held together by Coulomb attraction. Hence, we
are dealing with a two-body problem characterised by a force acting along a
line between the proton and electron and a potential, which depends on the
distance |re − rp | only: V = V (|re − rp |). Classical physics show that such
a two-body problem can be reduced to a one-body central field problem.

38
e-

rep
re

R
p+
rp

Figure 3.1: Illustration of the vectors involved in the description of the hy-
drogen atom. R is the vector to the centre-of-mass of the hydrogen atom.

3.1.2 The classical approach


Figure 3.1 shows the notation that we are going to use in the following
treatment. Newton’s second law gives us the following:

dV r e − r p
me r¨e = −∇e V (|r e − r p |) = − ,
drep rep

where rep = |r e − r p |. Likewise,

dV r p − r e
mp r̈ p = −∇e V (rep ) = − .
drep rep

The vectors r e and r p can be exchanged against the vectors R and r ep to


the centre-of-mass and between the proton and electron, respectively:

re = me +mp
M
re = me
M e
r + mp
M e
r + mp
M p
r − mp
M p
r =
(3.1)
me r e +mp r p mp mp
= M
+ r
M ep
=R+ r ,
M ep

and, in the same way,


me
rp = R − r ep . (3.2)
M
39
M is the mass of the total system of the proton and electron. Adding the
two equations gives
me r̈ e + mp r̈ p = me R̈ + me mp
M
r̈ ep + mp R̈ − me mp
M
r̈ ep = M R̈ =

dV r ep dV r ep
= − drep rep
+ drep rep
= 0,

that is
M R̈ = 0, (3.3)
which means that the centre-of-mass moves without acceleration, quite as
expected for a hydrogen atom in no outer field. Taking the difference of
equations (3.1) and (3.2) one sees that
{ }
1
2
(me r̈ e − mp r̈ p ) = 1
2
me R̈ + me mp
M
r̈ ep − mp R̈ + me mp
M
r̈ ep =
( )
dV r e −r p
me mp
M
r̈ ep ≡ µr̈ ep = 1
2
2 drep rep
≡ − dr
dV
ep
r̂ ep ,

where r̂ ep denotes the unit vector along r ep and µ is the reduced mass of the
system. In other words, we have shown that the separation of proton and
electron, r ep , obeys a Newtonian law
dV
µr̈ ep = − r̂ ep , (3.4)
drep
and that the two-body problem can be reduced to a one-body problem. V is
the central potential, which in the case of the hydrogen atom is the Coulomb
potential
e2
V =− . (3.5)
4πε0 rep
Since the force is purely radial, i.e. F (r ep ) = f (rep )r̂ ep = − dr
dV
ep
r̂ ep , we see
that
dV
f (rep ) = − . (3.6)
drep

Orbital angular momentum


The torque r ep × F is zero. This means that the angular momentum l =
r ep × p is a constant of the motion. It follows that the motion is strictly
planar: r ep · l = 0, and we may choose a coordinate system such that
x = rep cos ϕ,
(3.7)
y = rep sin ϕ.

40
Using these cylindrical coordinates one finds for the length of the angular
momentum
l = µrep2 ϕ̇, (3.8)
which, like l, is a constant of the motion. It turns out that also the z-
component of the angular momentum
lz = µ (xẏ − y ẋ) (3.9)
is a constant of the motion.

Remark : Here we find both parallels and differences to the quantum mechan-
ical treatment, as we will see later during the course. Classically, l, l, and
lz are constants of the motion. Quantum mechanically, in contrast, l is no
constant of the motion, while both l and lz are. The reason is that quantum
mechanics does not allow the full determination of the angular momentum.

Task: Show that the torque r ep × F vanishes. Show also that the angular
momentum l = r ep × p is a constant of the motion. Show that the motion
of the electron around the nucleus (or rather the motion of the nucleus and
electron around the centre-of-mass) is planar. Finally, show also that l and
lz are constants of the motion. Phyiscally, why must lz be a constant of the
motion?

Centrifugal barrier
The goal of a classical treatment would be to obtain the (classical) trajectory
of the electron around the nucleus rep (t). The starting point for finding rep (t)
would be the energy
( )
E = Ekin + V = 21 µ (ẍ2 + ÿ 2 ) + V (rep ) = 12 µ r̈ep
2
+ r2ep ϕ̈2 + V (rep ) =

l2
= 21 µr̈ep
2
+ 2µr2ep
+ V (rep ) = 12 µr̈ep
2
+ Veff (rep ),
(3.10)

41
l2
2 µ rep2

Energy

Veff

- A / rep

rep

Figure 3.2: Illustration of the shape of the effective potential.

l2
where Veff (rep ) contains the centrifugal barrier 2µr2ep
. For a −1/r-potential the
potential curves are drawn in Figure 3.2.

Remark : We will also encounter the centrifugal barrier in the full quantum
mechanical treatment of the hydrogen atom.

Remark : The energy, which we consider above, is not a constant of the


motion, i.e., we have neglected an important contribution. Which one?

If we had the full energy, we could, in principle, determine the classical


trajectory by calculating the time dependence t(rep ) and then inverting it to
give rep (t). Even in the classical theory this becomes very demanding, due
to the occurring integrals. One gets quite a bit further by instead explicitly
considering the forces which act on the electron and neglecting radiation.
With the potential given by Eq. (3.5), the relevant forces are the Coulomb
force and the balancing centrifugal force:

e2 µv 2
2
= = µω 2 rep ,
4πε0 rep rep

42
where ω is the angular frequency ω = v/rep . Solving for ω 2 gives

e2 /4πε0 1
ω2 = 3
= 2, (3.11)
µrep T

with T the time period for one revolution. This corresponds exactly to
Kepler’s third law, with the gravitational potential replaced by a Coulomb
potential:

Kepler’s third law states that the orbital period T 2 is proportional


to the cube a3 (= r3ep ) of the semi-major axis of its orbit.

From the balance of forces we also find an expression for the energy:

1 1 e2 /4πε0 e2 /4πε0
E = µv 2 + V (rep ) = µv 2 − =− . (3.12)
2 2 rep 2rep
This is this far classical physics takes us under the neglect of radiation. In
order to get a step further we now need to introduce the quantum hypothesis
of Planck. It was Niels Bohr in Copenhagen (1885 - 1962), who did this first
in his semi-classical model of the atom. If you are interested you can find his
original work on the internet (N. Bohr, Philosoph. Mag. Ser. 6 (1913) 26, 1,
accessible from http://www.informaworld.com/smpp/philmagarchive db=all).

3.1.3 The Bohr Model (1913)


In contradiction to classical theory, Bohr assumed that there are only certain
allowed orbits for the electron, and that radiation due to electron acceleration
is not allowed. The only way radiation can be emitted or absorbed is by
electrons passing from one orbit to another. The lost or gained energy is
assumed to be a multiple of Planck’s energy quantum h̄ω = hν:

∆E = k h̄ω, i.e.,
(3.13)
∆E
ω= kh̄
.

Without any loss of generality we here assume that ∆E > 0. From Eq. (3.12)
an expression for the difference of the energies of an electron in different orbits
is found ( )
e2 /4πε0
∆E = 2 1
rep
− r′ , i.e.,
1
ep

( )
1 e2 /4πε0 e2 /4πε0 ∆rep
ω= kh̄ 2
1
rep
− 1
r′ep
≃ 2kh̄ r2ep
.

43
Here it has been assumed that rep ≃ r′ep , which is true only for two orbits
with a large distance of the electron from the nucleus. This is no further
limitation, since the physics for large orbits carry over to smaller orbits.
We can now equate the classical result in Eq. (3.11) with the present
value:
2
2 /4πε (e2 /4πε0 ) (∆rep )2
ω 2 = e µr 3
0
= 4k2 h̄2 r2
, i.e.,
ep ep

4 h̄2 me 2
(∆rep )2 = e2 /4πε0 me µ
k rep = 4a0 mµe k 2 rep ,

√ √
=⇒ ∆rep = 2 a0 rep mµe k.

a0 is the Bohr radius (a0 = 0.529 Å= 0.529 × 10−10 m). Let k = 1 and
assume that there exists a quantum number n (a quantum number is always
an integer number) so that

me x
rep = anx µ
,

me x
r′ep = a(n + 1)x µ
.

This ansatz comes about from requiring that the orbit radius should scale
with a label (the quantum number n) and that it needs to have the dimension
of a length – therefore the length parameter a. For n ≫ 1 and using a Taylor
expansion of (n + 1)x one finds

me x
∆rep = (a(n + 1)x − anx ) µ

{( )x ( ) x−1 ( )x }
≃a me
µ
2 x
n +x me
µ
2
(n + 1) x−1
−n x me
µ
2
=

( ) x−1 ( ) x−1
= ax(n + 1)x−1 me
µ
2
≃ axnx−1 me
µ
2
=

√ √ √ √ x ( )x
= 2 a0 rep mµe = 2 a0 an 2 mµe 4

√ x√ ( )1− x
2n1− 2 a0 me 2
=⇒ a= x µ
.

By construction, a should be independent of n, and this implies x = 2 and


a = a0 , which gives
me
rep = a0 n2 , (3.14)
µ

44
where n, as said before, is an integer number. Approximating me ≃ µ (or
defining a0 using the reduced mass of the hydrogen atom rather than the
electron rest mass) one gets the more commonly seen

rep = a0 n2 . (3.15)

For the energy one finds


e2 /4πε0 1 µ
E= 2a0 n2 me
(3.16)
e2 /4πε0 1
me ≃ µ : E = 2a0 n2
.

This result for the energy of the hydrogen atom turns out to be correct even
in a proper quantum mechanical treatment!
The result applies, approximately, also for atoms composed of a heavier
nuclei and an electron.

The Rydberg constant and Rydberg formula


One can express the energy in wave numbers instead of the here used energy.
Wave numbers (ν̄ = 1/λ = ν/c = E/hc) are a commonly used quantity
in some kinds of spectroscopy. In wave numbers the difference of energies
becomes: 2
( ) ( )
ν̄ = e2a/4πε 0 µ
0 hc me
1
n 2 − n
1
′ 2 ≡ R ∞
µ
me n
1
2 − n
1
′ 2 ≃
( ) (3.17)
≃ R∞ n2 − n′ 2 .
1 1

The Rydberg constant of indefinite mass is R∞ = 1.0973 × 107 m−1 . The


formula is called Rydberg formula.

Good to remember : With n = 1 and n′ = ∞ you find hcR∞ = 13.6 eV,


which is the ground state ionisation energy for an atom with one electron
and an indefinitely heavy nucleus, but of course the value also applies to the
hydrogen atom due to the small difference between me and µ.

For real nuclei with atomic number Z and mass mz the Rydberg constant
becomes ( )
µ mZ me
RZ = R∞ = R∞ ≃ R∞ 1 − . (3.18)
me mZ + me mZ
In addition, we need to take into account that real nuclei have a charge
+Ze, so that the Coulomb interaction with the electron becomes Ze2 /4πε0 r2ep .

45
The energy of a hydrogen-like atom (i.e., nucleus plus one electron) is then
(remember that a0 contains the factor e2 /4πϵ0 !)

e2 /4πε0 Z 2 Z2
E=− = −hcR Z 2. (3.19)
2a0 n2 n

Standing waves
The derivation of the distance between electron and nucleus in the Bohr
model is more straightforward if one applies the wave/matter duality (which
was not known to Bohr at the time he developed his model). Any stationary
state of the electron must be described by a standing wave. This condition
can be written as
2πrep = nλelectron , (3.20)
with n an integer number. From de Broglies wave/matter duality we find
the wavelength associated with the electron as
h h
λelectron = = , (3.21)
p µv
which gives for the speed of the electron
h
v=n .
2πµrep
From the balance of centrifugal and central forces we have
e2
e2 /4πε0 4πε0
(2πµrep )2
rep = = ,
µv 2 µn2 h2
and hence
n 2 h2 1 1 me
rep = 2 2
= a0 n2 ,
e /4πε0 µ 4π µ
which is the same result as we found before in Equation (3.14).

3.1.4 Experimental observations


One of the primary driving forces for the introduction of new atomic models
in the 19th and 20th century – and among these models in particular the
Bohr Model – was the observation that atoms emit and absorb light only
at certain, discrete energies. This was first shown by Gustav Kirchhoff and
Robert Bunsen in 1859, and they probably used setups similar to those shown
in Figure 3.3 and 3.4 (well, without the computer and diode). Also the

46
Figure 3.3: Typical optical emission setup. The source is hot, so that the
source atoms are heated up and emit light (Figure from Demtröder, Atoms,
Molecules and Photons, Springer, Berlin, Heidelberg, 2006).

Figure 3.4: Typical optical absorption setup. The light source needs to
be continuous (Figure from Demtröder, Atoms, Molecules and Photons,
Springer, Berlin, Heidelberg, 2006).

emission series of the hydrogen atom, for which an empirical formula was
developed by Johann Jakob Balmer in the 1880s, will have been measured
by similar setups. Both the emission and absorption series of the hydrogen
atom are shown in Figure 3.5. Balmer’s formula was later generalised by
Janne Rydberg to formula (3.17), which also describes the other emission
and absorption series of the hydrogen atom. The Balmer series corresponds
to transitions to/from the Bohr orbit with label n = 2. Its main part is found
in the visible regime. Other series are the Lyman (from/to n = 1), Paschen
(from/to n = 3), Brackett (from/to n = 4), and Pfund series (from/to n = 5).
The wavelength, energy, and wavenumber regimes of these series are specified
in Table 3.1.

47
Figure 3.5: The Balmer series in emission and absorption (Figure from
http://www.astronomyknowhow.com/hydrogen-alpha.htm).

48
Energy (eV) Wavelength (nm) Wavenumber (cm−1 )

Lyman (n=1) n’=2 10.20 122 8.2×106

n’=∞ 13.60 91 1.1×107

Balmer (n=2) n’=3 1.89 656 1.5×106

n’=∞ 3.40 365 2.7×106

Paschen (n=3) n’=4 0.66 1875 5.3×105

n’=∞ 1.51 828 1.2×106

Table 3.1: Energy, wavelength, and wavenumber regimes of the Lyman,


Balmer, and Paschen series.

3.1.5 Moseley’s Law


In the 1910s Henry Gwyn Jeffreys Moseley measured the x-ray spectra of
many elements by bombarding the materials by high-voltage electrons. This
leads to the excitation of inner electrons. The excited atoms will eventually
decay under the emission of x-rays with characteristic energies. Moseley
found a relationship between the frequency of the emitted x-rays and the
atomic number: √
ν ∝ Z. (3.22)
This relationship is today called ”Moseley’s Law” and is in qualitative agree-
ment with Rydbergs formula
{ }
Z2 Z2
ν = cR∞ − ′2 .
n2 n
The formula and measured energies do not comply too well, since the
formula is formulated for atoms with one electron. The agreement can be
improved by the introduction of shell-dependent screening factors σn :
{ }
(Z − σn )2 (Z − σn′ )2
ν = cR∞ − .
n2 n′2

49
Figure 3.6: Moseley’s recording of the frequencies of the characteristic x-ray
emitted by atoms with atomic number Z (from Phil. Mag. 27, 703 (1914)).

50
The term ”screening” describes the effect of the other electrons, which do
not participate in the optical transition, onto the nuclear charge seen by
the electron, which ”jumps” from one orbit to another. As is seen from the
formula, the charge is reduced to an effective charge (Z − σn )e.
The term ”shell” describes the orbits in the Bohr Model and are labelled
as K-shell for n = 1, L-shell for n = 2, M -shell for n = 3, and so on.

3.1.6 Shortcomings of the Bohr model


The Bohr model correctly predicts the energies observed in the emission
and absorption series of atomic hydrogen. Nevertheless, it has a number of
severe shortcomings, which made it necessary to introduce a fully quantum
mechanical model. Some of the shortcomings are:

• The Bohr model does not comply with wave/matter duality: it does not
obey Heisenberg’s uncertainty relation ∆x∆p ? h̄. Both position and
momentum are fully classical and can be determined to any precision
simultaneously.

• In the Bohr model the orbital angular momentum for hydrogen in its
ground state is |l| = nh̄ (show this by yourself!). From experiment it
is know that it actually is zero.

• The Bohr model fails in explaining the intensities of the spectral lines.

• The spectra of heavier atoms can only be explained if ad-hoc assump-


tions are made (see Moseley’s law).

• The fine and hyperfine structure in the spectra of hydrogen and other
atoms cannot be explained. An explanation of this structure actually
goes beyond non-relativistic quantum mechanics, as well, but can be
treated when making some assumptions.

In this course we will treat the non-relativistic quantum mechanical the-


ory of the hydrogen atom and other atoms. Before we do that, we introduce
the concept of the spin by means of a consideration of the Stern-Gerlach ex-
periment. This experiment provides a very nice and pretty straightforward
illustration of fundamental quantum effects, and therefore we already intro-
duce spin now, although we will not need it in the first quantum mechanical
treatment of the hydrogen atom. We will also use it to repeat and introduce
some quantum mechanical concepts and tools.

51
3.2 Solution of the time-dependent Schrödinger
equation
In the preceeding chapter we have arrived at the time-dependent Schrödinger
equation:

|α t t0 ⟩ = H |α t t0 ⟩ .
ih̄
∂t
If we now consider a hydrogen atom without external field, it is clear that
it constitutes a stationary system, which implies that the Hamiltonian does
not explicitly depend on time or, in other words, that it is conserved (the
hydrogen atom is said to be a ”conservative system”):
∂H
= 0.
∂t
If we assume that H has eigenstates, these must be time-indepdent, too. We
label the states by a number of quantum numbers, here represented by what is
to become the principal quantum number n and a label for all other quantum
numbers, τ . Since H is the energy observable, we label the eigenvalues by E:

H |n τ ⟩ = En,τ |n τ ⟩ . (3.23)
This is the stationary Schrödinger equation, and it actually just expresses
that the hydrogen atom is a system with stationary energy eigenstates. The
derivation of the solution of the time-independent Schrödinger equation is
the main theme of this chapter – but we also would like to know the corre-
sponding solution to the time-dependent Schrödinger equation. A reasonable
ansatz is to say that this solution must build on the solution of the time-
independent Schrödinger equation, and we therefore postulate that time-
dependent solution is a superposition of the time-independent states with all
time dependence in the coefficients of the sum:

|α t t0 ⟩ = cn τ (t) |n τ ⟩ ⇔ (3.24)
n τ
cn,τ (t) = ⟨n τ |α t t0 ⟩ . (3.25)
If we use 3.24 in the time-dependent Schrödinger equation and remember
that H is Hermitian and we therefore know how it acts to the left, then we
find a differential equation:
⟨ ⟩ ⟨ ⟩
∂ ∑
n τ ih̄ ∂t

α t t0 = n τ ih̄ ∂t ′
n ,τ ′ cn ′ ,τ ′ (t) ′ ′
n τ =


= ih̄ ∂t cn,τ (t) = ⟨n τ |H|α t t0 ⟩ = En,τ ⟨n τ |α t t0 ⟩ = En,τ cn,τ (t),

52
i.e.

cnτ (t) = En,τ cn,τ (t).
ih̄ (3.26)
∂t
We can write down the solutions to this equation right away:
t−t0
cn,τ (t) = cn,τ (t0 )e−iEn,τ h̄ , (3.27)

with
cn,τ (t0 ) = ⟨n τ |α t0 t0 ⟩ = ⟨n τ |α t0 ⟩ . (3.28)
Thus, spelled out, the solution to the time-dependent Schrödinger equation
is
∑ t−t0
|α t t0 ⟩ = ⟨n τ |α t0 ⟩ e−iEn,τ h̄ |n τ ⟩ . (3.29)
n,τ

All time dependence is packed in the phase factor.


While we now have the general form of the solutions to the time-dependent
Schrödinger equation, we are missing the solutions |n τ ⟩ and corresponding
eigenvalues En,τ of the time-independent Schrödinger equation – and finding
these is the programme for now.

Question: What does it imply for the probability of finding the hydrogen
atom in a certain state and for superpositions of different states that the
time dependence is found in the phase factor?

Remark/Question: Solutions (3.29) are, in this form, valid only for the dis-
crete part of the hydrogen energy spectrum. Part of the energy spectrum is
continuous. For this part (3.29) has to be rewritten as
∑∫ t−t0
|α t t0 ⟩ = dE ⟨Eτ |α t0 ⟩ e−iE h̄ |Eτ ⟩ . (3.30)
τ

Comparison with (3.29) shows that the energy E here takes the same labelling
role for the states as the principal quantum number n does for the discrete
states. But what does it actually mean when I say that a part of the energy
spectrum is continuous? And which part of the hydrogen energy spectrum

53
is continuous and why? Using the Bohr model, what are the theoretical
boundaries of the continuous part of the hydrogen spectrum in absorption
and emission (given as energies and wavelengths)?

3.3 Solution of the time-independent Schrödinger


equation
We will now set out to find the analytical form of the solutions to the time-
independent Schrödinger equation (3.23). From classical mechanics we know
that the total energy is given by the kinetic energy plus the potential energy,
which in the present case stems from the Coulomb attraction between proton
and electron. Hence we readily have

p2 p2 e2 p2 e2
H = Ek + V = +V = − ≡ − , (3.31)
2µ 2µ 4πε0 rep 2µ 4πε0 r
where µ = mmee+m
mp
p
is the reduced mass of the hydrogen atom and where I
for convenience have skipped the label on the distance between electron and
proton. The time-independent Schrödinger equation is thus
{ }
p2 e2/4πε0
− |n τ ⟩ = En,τ |n τ ⟩ . (3.32)
2µ r
Much of the solution of this equation will now be done in the position basis,
i.e. using wave functions. At suitable points we will return to Dirac notation,
though.
In order to rewrite (3.32) with wave functions, we multiply from the left
by a position bra:
⟨ 2 ⟩
p e2/4πε0
x 2µ − r
n τ = En,τ ⟨x|n τ ⟩ ⇔

⟨ 2 ⟩
p e2/4πε0
x 2µ n τ − x r
n τ = En,τ ψn,τ (x).

The matrix element ⟨x|p2 |n τ ⟩ can be evaluated using equation (2.74) and
for a moment setting |β⟩ ≡ p |n τ ⟩:

⟨x|p2 |n τ ⟩ = ⟨x|p|β⟩ = −ih̄∇ ⟨x|β⟩ = −ih̄∇ ⟨x|p|n τ ⟩ =

= −h̄2 ∇2 ⟨x|n τ ⟩ = −h̄2 ∇2 ψn,τ (x).

54
The Coulomb potential contains only the position operator, and in the po-
sition basis we simply can replace it by the position eigenvalue (cf. equation
(2.13) and note that I for simplicity using the same symbol r for the position
operator and the position eigenvalue), i.e. ⟨x|V (r)|n τ ⟩ = V (r) ⟨x|n τ ⟩ =
V (r)ψn,τ (x). The Schrödinger equation thus becomes, as expected,
{ }
h̄2 ∇2 e2/4πε0
− − ψn,τ (x) = En,τ ψn,τ (x). (3.33)
2µ r
In spherical coordinates, i.e. with x = r sin θ cos ϕ, y = r sin θ sin ϕ, z =
r cos ϕ (cf. Figure 3.7), the Laplacian operator ∇2 is rewritten as
( )
1 ∂2 1 ∂2 1 ∂ 1 ∂2
∇2 = r+ 2 + + , (3.34)
r ∂r 2 r ∂θ2 tan θ ∂θ sin2 θ ∂ϕ2
and with the abbreviation
( )
∂2 1 ∂ 1 ∂2
l ≡ −h̄
2 2
+ + (3.35)
∂θ2 tan θ ∂θ sin2 θ ∂ϕ2
we find the Schrödinger equation in spherical coordinates
{ }
h̄2 1 ∂ 2 l2 e2/4πε0
− r + − ψn,τ (x) = En,τ ψn,τ (x) (3.36)
2µ r ∂r2 2µr2 r
The second in the curly brackets looks suspiciously like the centrifugal barrier,
which we encountered in equation (3.10) and figure 3.2, and, indeed, l2 is not
a mere abbreviation, but it is the square of the orbital angular momentum.
Note the l2 does not depend on r, while the remaining part of the Hamiltonian
does not depend on θ and ϕ. All angular dependence of the Schrödinger
equation is thus confined to the centrifugal barrier term!

Remark: From (3.36) you see that the Hamiltonian H and l2 commute:
[H, l2 ] = 0 (how do you see that?). From section 2.4.3 we know that this
implies that H and l2 have common eigenstates, and this we will also see
explicitly in the following. We have also already discussed that one needs to
find a maximum set of commuting observables to label all the states. For the
hydrogen atom H and l2 obviously form part of such a set.

55
z
r
θ
y
r

ϕ
x

Figure 3.7: Spherical coordinates.

Remark: Please note that equation (3.34) is really a short-hand notation for
( )
1 ∂2 1 ∂2 1 ∂ 1 ∂2
∇ f (x) =
2
r+ 2 + + f (x),
r ∂r 2 r ∂θ2 tan θ ∂θ sin2 θ ∂ϕ2

where f is a function of x. This implies that derivative in the first term also
acts on the r written behind! Forgetting about this is a common source of
mistakes.

Task: Show the validity of equation (3.35) by using the definition of the
orbital angular momentum l = x × p and the prescription (2.77) for p.

Task: Show the validity of equation (3.34). One way of doing this is to use
the expressions for x, y, z in spherical coordinates to derive expressions for
the differentials dxi = ∂x∂r
i
dr + ∂x
∂ϕ
i
dϕ + ∂x
∂θ
i
dθ. Solve for dr, dθ, and dϕ and
∂ ∂r ∂ ∂ϕ ∂ ∂θ ∂
use these to express the ∂xi = ∂xi ∂r + ∂xi ∂ϕ + ∂x i ∂θ
.

To get further we multiply by 2µr2 (observe that you have to do this


”from the left”):

56
{ 2
}
−h̄2 r ∂r

2r + l −
2 e2/4πε0 2µr ψ
n,τ (x) = 2µr En,τ ψn,τ (x) ⇔
2

{ 2
}
l2 ψn,τ (x) = h̄2 r ∂r 2 r + 2µr /4πε0 − 2µr En,τ
∂ e2 2
ψn,τ (x).

Now the operator on the left-hand side only depends on the angular variables
θ and ϕ, while the operator on the right-hand side only depends on the radial
variable r, i.e. we have separated the variables. This suggests that we also
should separate the variables in the wave function using the ansatz

ψn,τ = Rn,τ (r)Yn,τ (θ, ϕ), (3.37)


because we then complete the separation of variables:

{ 2
}
l2 Rn,τ (r)Yn,τ (θ, ϕ) = h̄2 r ∂r
∂ e2 2
2 r + 2µr /4πε0 + 2µr En,τ Rn,τ (r)Yn,τ (θ, ϕ) =
{ 2
}
= R(r)l2 Yn,τ (θ, ϕ) = Yn,τ h̄2 r ∂r
∂ e2 2
2 r + 2µr /4πε0 + 2µr En,τ Rn,τ (r) ⇔
{ }
∂2
2 h̄2 r r+2µre2/4πε0 +2µr2 En,τ Rn,τ (r)
l Yn,τ (θ,ϕ)
=
∂r 2
Yn,τ (θ,ϕ) Rn,τ (r)
(3.38)
Since the left-hand side only depends on θ and ϕ, while the right-hand side
only depends on r, they must equal a (dimensionless) constant, which we
call ωl h̄2 . Actually, we will show below that ωl has the form l(l + 1), with
l an integer number. l is the orbital angular momentum quantum number,
which we will use alongside the principal quantum n and further quantum
numbers to be determined for labelling the electronic states of the hydrogen
atom. Now using the constant, for the left-hand side we find
( )
l2 Yn,τ (θ, ϕ) = ωl h̄2 Yn,τ (θ, ϕ) = l(l + 1)h̄2 Yn,τ (θ, ϕ) . (3.39)

Obviously this is an eigenvalue equation for l2 , and, at the same time, it


is the angular Schrödinger equation of the hydrogen atom, which will now
proceed to solve. Before we do that, for book-keeping reasons we write down
the remaining part of (3.38) (this is the radial Schrödinger equation of the
hydrogen atom, which we will have to solve a bit later):
{ }
h̄2 ∂ 2 ωl h̄2 e2/4πε0
− + − rRn,τ (r) = En,τ rRn,τ (r). (3.40)
2µ ∂r2 2µr2 r

57
Remark: Please note that l ̸= |l| in contradiction with conventional mathe-
matical notation. While slightly confusing, this is the accepted and conven-
tional physics notation, and you should be aware of it.

3.3.1 Solution of the angular Schrödinger equation of


the hydrogen atom
Using equation (3.35) we can rewrite equation (3.39) as

( )
∂2 ∂ ∂2
− sin θ 2 + sin θ cos θ
2
− ωl sin2 θ Yn,τ (θ, ϕ) = Yn,τ (θ, ϕ). (3.41)
∂θ ∂θ ∂ϕ2

As before for r on the on side and θ and ϕ on the other, we have here
separated the operators according to variables θ and ϕ. This leads us to the
ansatz

Yn,τ (θ, ϕ) = Tn,τ (θ)Fn,τ (ϕ), (3.42)


and we can finalise the separation of variables:
( )
∂2
sin2 ∂θ2

+ sin θ cos θ ∂θ − ωl sin2 θ Tn,τ (θ) ∂2
F (ϕ)
∂ϕ2 n,τ
=− ≡ m2l
Tn,τ (θ) Fn,τ (ϕ)

Again, both sides of the equation need to equal a constant, which we have
named m2l . ml will be the magnetic quantum number and constitute yet
another (necessary) label for the electronic states of the hydrogen atom.

Labels of the electronic states - the quantum numbers


If we neglect spin (and this we do for right now), we now have in total three
equations to solve,

{ }
h̄2 ∂ 2 l(l+1)h̄2 e2/4πε0
2µ ∂r2
− 2µr2
− r
rRn,τ (r) = En,τ rRn,τ (r),
( )
∂2
sin2 ∂θ 2

+ sin θ cos θ ∂θ − ωl sin2 θ Tn,τ (θ) = m2l Tn,τ (θ), (3.43)

∂2
F (ϕ)
∂ϕ2 n,τ
= m2l Fn,τ (ϕ),

58
which is fully sufficient to determine the three unknown functions R(r), T (θ),
and F (ϕ). We also have three labels or quantum numbers, n, ωl , ml (or rather
n, l, ml ), and these will be sufficient to fully label the solutions. Hence, we
can replace τ by the set l, ml . We note that there occurs no n or l in the
function for F (ϕ) and no n in the function for T (θ), and hence we can skip
these labels:

Fml (ϕ) ≡ Fn,τ ,

Tl,ml (θ) ≡ Tn,τ .

Thus, the angular part of the solution of the Schrödinger equation for the
hydrogen atom is labelled by l and ml only, but not by n, which will be
needed for labelling the radial parts:

Yl,ml (θ, ϕ) ≡ Yn,τ .

In the following we will find rules for which values l, ml , and n may take.

Solution for F
The solution to the last of the three equations in (3.43) is

1
Fn,τ (ϕ) ≡ Fml (ϕ) = √ eiml ϕ , (3.44)

where the prefactor is found from the normalisation requirement
∫ 2

N eiml ϕ dϕ = 1. (3.45)

N is the normalisation constant, which from this turns out to be N = 1/ 2π .

Question: Why is (3.45) required?

Task: From the its symmetry, show that (3.44) implies that ml only can take
integer values:
ml = 0, ±1, ±2, . . . (3.46)

59
∂2
2
Instead of the equation ∂ϕ 2 Fml (ϕ) = ml Fml (ϕ) we also can simply write

F (ϕ) = iml Fml (ϕ). If we then identify
∂ϕ ml

h̄ ∂
lz = (3.47)
i ∂ϕ
we can also write

lz Fml = ml h̄Fml . (3.48)


This is the eigenvalue equation for the z-component of the orbital angular
momentum.

Task: Show the validity of (3.47) using the definition of the orbital angular
momentum.

Task: Show that l2 and lz commute: [l2 , lz ] = 0.

[l2 , lz ] = 0 implies that the observables l2 and lz have common eigen-


states. From equation (3.48) it is seen that the Fml , which we already know,
are eigenfunctions of lz . Since lz does not depend on θ also the product
Yl,ml (θ, ϕ) = Tl,ml (θ)Fml (ϕ) must be an eigenfunction of lz . But we already
now that the Yl,ml (θ, ϕ) are eigenfunctions of l2 (cf. equation (3.39))!

Ladder operators
In order to find the full form of the Yl,ml we would like to solve the second
equation in (3.43). This is done iteratively, as you will see in the following.
We start by defining two ladder operators:

L± = Lx ± iLy , (3.49)
where the Lx and L − y are the x- and y-components of the orbital angular
momentum. You should now solve the following tasks concerned with the
action of ladder operators on the functions Yl,ml . You will also derive rules
for the quantum numbers l and ml .

60
Task:
Using the definition of l and the canonical commutator relations show that
the components of l obey the following:

[lx , ly ] = ih̄lz

[ly , lz ] = ih̄lx (3.50)

[lz , lx ] = ih̄ly
What do these relations imply for [ly , lx ], [lz , ly ], and [lx , lz ]?
Show now that l2 and l± commute:
[ ]
l2 , l± = 0.

Furthermore show that


[lz , l± ] = ±h̄l± .
Using these relations show that

ωl − m2l ≥ 0.

Show that l+ Yl,ml (θ, ϕ) is an eigenfunction of lz with eigenvalue (ml + 1)h̄.


Likewise, show that l− Yl,ml (θ, ϕ) is an eigenfunction of lz with eigenvalue
(ml − 1)h̄.
Convince yourself that this implies that
Yl,ml +1 (θ, ϕ) = N l+ Yl,ml (θ, ϕ),

Yl,ml −1 (θ, ϕ) = M l− Yl,ml (θ, ϕ),


where N and M are normalisation constants.
Show for the products of l+ and − the following:
l+ l− = l2 − lz2 + h̄lz ,

l− l+ = l2 − lz2 − h̄lz .
Using these relationships you can perform the normalisation and obtain N
and M above. Show that
N= √ 1
,
h̄ ωl −ml (ml +1)

M= √ 1
.
h̄ ωl −ml (ml −1)

61
Also convince yourself that the preceding statements imply that for a
given ωl there will a maximum and minimum value for ml , which we call
ml,max and ml,min .
Since there are maximum values of ml , it is clear that

l+ Ylml,max = 0,

l− Ylml,min = 0.

Show that
ml,max (ml,max + 1) = ml,min (ml,min − 1) = ωl ,

that
ml,max = −ml,min ,

and that
positive integer
ml,max = .
2
If we define l = ml,max with l an integer (since we actually have shown previ-
ously that ml must be an integer number), show that

ωl = l(l + 1).

What are the values that ml can assume?

So, what have we learnt? A couple of different things, actually. First,


we see that the ladder operators can switch between the Yl,ml with the same
l but different ml . Ladder operators are very important tools, especially in
quantum field theories, where they are used to generate different states in
the same way as they are used here to generate the different states of the
hydrogen atom. Second, for a given l this means that we have a tool at hand
to generate the different Yl,ml iteratively. Third, we have derived rules for
two of the quantum numbers used to label the eigenstates of the hydrogen
atom: l is an integer and ml can run from ... (well, this you should have
answered by yourself).
As a comment, all what we have formulated above we could have formu-
lated in Dirac notation as well:

62

l± |l, ml ⟩ = h̄ l(l + 1) − ml (ml ± 1) |l, ml ± 1⟩ ,

l+ |l, ml,max ⟩ = 0,
(3.51)
l− |l, ml,min ⟩ = 0,

lz l± |l, ml ⟩ = (ml ± 1)h̄l± |l, ml ⟩ .

Full solution of the angular Schrödinger equation of the hydrogen


atom
We are now in the position to find the solutions to the middle equation in
(3.43), which we reformulate, using ωl = l(l + 1), as

∂2 1 ∂ m2l
− 2
+ − 2
2 Tl,ml = l(l + 1)h̄ T (θ). (3.52)
∂θ tan θ ∂θ sin θ
Actually, we will not directly solve this equation, but it is important, because
it, after multiplication with F (ϕ) and R(r) represents the eigenvalue equation
for l2 :

l2 ψn,l,ml = l(l + 1)h̄2 ψn,l,ml . (3.53)


To find the T (θ) we instead write l+ in spherical coordinates and let it act
on Yl,ml,max :
( )
iϕ ∂ ∂
l+ Yl,ml,max = h̄e + i cot θ Yl,ml,max .
∂θ ∂ϕ

Since we know that ml,max = l and since we know the solutions F (ϕ) this
gives
( )
h̄eiϕ ∂
∂θ

+ i cot θ ∂ϕ Tll (θ) √12π eilϕ = 0 ⇔


T (θ)
∂θ ll
= l cot θTll (θ).

As you can confirm easily this equation has the solution

Tll (θ) = cl (sin θ)l . (3.54)


The cl s are normalisation constants for which we require (why?)

|Yl,ml (θ, ϕ)|2 sin θ dθ = 1.

63

For c0 , for example, one finds c0 = 1/ 2.
Hence, for maximum ml = l the overall solution for the angular Schrödinger
equation of the hydrogen atom is
1
Yll (θ, ϕ) = √ cl sinl θeilϕ . (3.55)

All other Yl,ml (θ, ϕ) can be generated from Yll (θ, ϕ) by application of l− :
Yl,l−1 (θ, ϕ) = l− Yll (θ, ϕ),

2
Yl,l−2 (θ, ϕ) = l− Yll (θ, ϕ),

..
.

(2l+1)
Yl,−l (θ, ϕ) = l− Yll (θ, ϕ).

The spherical harmonics


The Yl,ml are the so-called spherical harmonics. The first couple of them are
given in Table 3.2. Since they are complex functions it is not entirely trivial
to plot them. However, physically we are interested in the probability rather
than the probability amplitude of the wave functions, i.e. we are interested
in |Yl,ml |2 . The squares |Yl,ml |2 do not depend on ϕ, and hence one can plot
them in so-called polar diagrams as in Fig. 3.8 for a number of selected
spherical harmonics. In these polar diagrams one plots the endpoint of a
vector with length |Yl,ml |2 for each possible value of θ.

Task: Plot all polar diagrams for l = 2 in Matlab or any other suitable
software.

3.3.2 Solution of the radial Schrödinger equation of


the hydrogen atom
Now that we have found the solutions Yl,ml (θ, ϕ) to the angular Schrödinger
of the hydrogen atom we need to also find the solutions R(r) to the radial
Schrödinger equation of the hydrogen atom (3.40) or, with u(r) ≡ rR(r) and
noting that the equation does not contain ml and the solutions (neither wave
functions nor energies) thus will not depend on ml ,

64
Table 3.2: Spherical harmonics.

l ml Yl,ml

1

0 (s) 0 2 π


1 (p) -1 1
2
3

sin θe−iϕ

1 3
1 (p) 0 2 2π
cos θ

1 (p) +1 - 12 3

sin θeiϕ


2 (d) -2 1
4
15

sin2 θe−2iϕ

2 (d) -1 1
2
15

cos θ sin θe−iϕ
√ ( )
2 (d) 0 1
4
5
π
2 cos2 θ − sin2 θ

2 (d) +1 − 12 15

cos θ sin θeiϕ

2 (d) +2 − 41 15

sin2 θe2iϕ


3 (f) -3 1
8
35
π
sin3 θe−3iϕ

3 (f) -2 1
4
105

cos θ sin2 θe−2iϕ

3 (f) -1 1
8
21π sin θ (5 cos2 θ − 1) e−iϕ

3 (f) 0 1
4
7π (5 cos3 θ − 3 cos θ)

3 (f) +1 − 18 21π sin θ (5 cos2 θ − 1) eiϕ

3 (f) +2 − 41 105

cos θ sin2 θe2iϕ

3 (f) +3 −6518 35
π
sin3 θe3iϕ
Figure 3.8: Polar diagrams of selected spherical harmonics. Please note the
different scaling of the diagrams. Please also note that the diagrams are
rotationally symmetric about the z-axis.
l=0, m=0 80
z
60

40 θ
|Ylm|2
20

-80 -60 -40 -20 20 40 60 80


x 10-3 -20

-40

-60
x 10-3

φ
l=1, m=0 l=1, m=+-1
0.2 z z
0.10

0.1 0.05

-0.2 -0.1 0.1 0.2 -0.10 -0.05 0.05 0.10

-0.1 -0.05

-0.2 -0.10

φ φ
l=3, m=0 l=3, m=+-1
z
15
z 4
10
2
5

-15 -10 -5 5 10 15 -4 -2 2 4
-5
-2
-10
-4
-15

φ φ
l=3, m=+-2 l=3, m=+-3
z z
0.15
0.10
0.10
0.05
0.05

-0.10 -0.05 0.05 0.10 -0.15 -0.10 -0.05 0.05 0.10 0.15
-0.05
-0.05
-0.10
-0.10
-0.15

φ φ

66
( )
h̄2 ∂ 2 l(l + 1)h̄2 e2/4πε0
− + − u(r) = En,l u(r). (3.56)
2µ ∂r2 2µr2 r

Asymptotic behaviour for r → ∞


For r → ∞ the second and third term in equation (3.56) go to zero, leaving
∂2 2µ
u(r) = − Eu(r),
∂r2 h̄2
with solution
√ √
2µE 2µE
u(r → ∞) = Aei h̄
r
+ Be−i h̄
r
.

Note that the energy E can be both positive and negative, with a positive
energy corresponding to an unbound state and a negative to a bound state.
For positive energies the solution is simply a travelling wave, with the first
part corresponding to an outbound √ and the second to an inbound wave. For
negative energies we rename κ ≡ −2µE h̄
and find
√ √
−i2 2µE −i2 2µE
−i
uE<0 (r → ∞) = Ae i h̄ + Be h̄ = Ae−κr + Beκr .

R(r) needs to be finite for all r, because it cannot be normalised otherwise.


The second term in uE<0 (r) = rR(r) and thus also in R(r), however, diverges
and thus B = 0. Hence
A −κr
uE<0 (r → ∞) = Ae−κr ⇔ RE<0 (r → ∞) = e .
r
The general solution to equation (3.56) will have to exhibit exactly this
asymptotic behaviour.

Behaviour for r → 0
In order to single out the term that will determine the asymptotic behavior
we replace

u(r) = re−κr y(r). (3.57)


In the ansatz we have also put in an r to account for that u(r) = rR(r). Now
we make a power-law ansatz for y(r):


y(r) = bj r j . (3.58)
j=s

67
For r → 0 the lowest power is most important, and hence

y(r → 0) = bs rs .

This we put into equation (3.56) and, again, just retain the lowest-power
terms (i.e. the rs−1 terms) and find

−s(s + 1) + l(l + 1) = 0. (3.59)


Therefore s=l or s = −(l + 1). However, the solution ψs=−(l+1) (r, θ, ϕ) =
1
Y (θ, ϕ) to the Schrödinger equation diverges at the origin and thus is
rl+1 l,ml
not normalisable. We therefore choose

s = l. (3.60)
The remaining possible solution is thus u(r → 0) = re−κr bs rl , which goes to
zero at the origin:

u(0) = 0. (3.61)

General solution - energies and rules for quantum numbers


If we now put the ansatz (3.57) for u(r) into (3.56) we find the following
equation for y(r) (it’s a little work, and one has to use the relationship
between κ and E):

∂ 2 y(r) 2 ∂y(r) ∂y(r) l(l + 1) µ 2κ


+ − 2κ − y(r) + 2 e2/4πε0
2 y(r) − y(r) = 0.
∂r2 r ∂r ∂r r2 h̄ r r
(3.62)
Using the power series ansatz (3.58), remembering that s = l (equation
(3.60)) and requiring that the coefficients of equals powers of r need to be
zero for the left-hand side of (3.62) being zero one arrives at a recursion
formula for the bj (also this is a little work):
2
κ(j + l) − 4πε
e µ
0 h̄
2
bj = 2bj−1 2
. (3.63)
j + 2jl + j
The power series can only have a finite number of summands, otherwise y(r)
– and with it R(r) – would diverge. Thus, there is a bm−1 , which is the last
non-vanishing coefficient of the power series (note that m is an integer number
≥ 1). bm is zero, and hence the nominator in (3.63) needs to disappar:

µ e2
κ(m + l) = . (3.64)
h̄2 4πε0

68

−2µE
If we replace κ = h̄
we find the energies of bound states of the hydrogen
atom:

µe4 1
E=− 2 2
, (3.65)
8ε0 h (m + l)2
where we have replaced h̄ = h/2π. Further, we rename

n=m+l (3.66)
and, impressively, we find exactly the same energy as Bohr found in his
model:

µe4 1
En = − , (3.67)
8ε20 h2 n2
(cf. equation (3.16))!
From (3.66) we also note that

l ≤ n − 1, (3.68)
or, acutally, since l must be positive or zero

0 ≤ l ≤ n − 1. (3.69)
This also implies that

n ≥ 1. (3.70)

General solution - radial wave functions


We can rewrite the recursion formula (3.63) by using equation (3.64) for κ
h̄2
and the definition of the Bohr radius a0 = e2/4πε 0µ
to yield
m−j 1
bj = −2bj−1 . (3.71)
j(j + 2l + 1)(m + l) a0
The solutions are, as you can check easily by yourself,
( )j ( )j
bj = − a10 2
m+l
(m−1)! (2l+1)!
(m−j−1)! j!(j+2l+1)!
b0 =
(3.72)
( )j ( )j
bj = − a10 2
n
(n−l−1)! (2l+1)!
(n−l−j−1)! j!(j+2l+1)!
b0 .
We can now use this recursion formula to determine the solutions Rnl (r) to
the radial Schrödinger equation of the hydrogen atom (remember Rnl (r) =

69
− 1 ∑
= e−κr y(r) = e a0 n n−l−1 r
u(r)/r
j=0 bj rj+l ). Before we can do that, however, we
have to determine the zeroth coefficient b0 . This is done by requiring that
R10 (r) is normalised:

∫∞ ∫∞ a30
0
2
R10 (r) r2 dr = 0 b20 e−2r/a0 r2 dr = 4
b20 ≡ 1 ⇔
(3.73)
2
b0 = 3/2 .
a0

Question: Where does the r2 in the integration come from?

Using equations (3.73) and (3.72) and requiring that the wave functions be
normalised (bit of work, again) one finds:

2 −r/a0
R10 (r) = 3/2 e ,
a0

( )
R20 (r) = 2
(2a0 )3/2
1− r
2a0
e−r/2a0 ,

1 √1 r e− /2a0 ,
r
R21 (r) = (2a0 )3/2 3 a0

( )
R30 (r) = 1
√ 3/2 6 − 4 ar0 + 4r2
9a20
e−r/3a0 ,
9 3a0

( ) (3.74)
2r −r/3a0
R31 (r) = 1
√ 3/2 4− 2r
3a0 3a0
e ,
9 6a0

1 4r2 −r/3a0
R32 (r) = √ 3/2
9a 2e ,
9 30a0 0

···,
{( )3 }1/2 ( )l
Rnl (r) = − 2
a0 n
(n−l−1)!
2n((n+l)!)3
2 a0rn e−r/a0 n Ln+l,2l+1 (2r/a0 n) ,

where the Ln+l,2l+1 are the so-called associated Laguerre polynomials, which
you can find tabulated, and where the term in curly brackets comes from the
normalisation of the wave function.

70
3.3.3 The energies of the hydrogen atom
We found that the energies of the hydrogen atom are (equation (3.67))
µe4 1 e2/4πε0 µ 1 e2/4πε0 1 1
En = − 2 2 2
= − 2
≈ − 2
= −13.6eV 2 .
8ε0 h n 2a0 me n 2a0 n n
The energy only depends on the principal quantum number n (= 1, 2, 3, ...),
but not the orbital angular momentum quantum number l (0 ≤ l < n).
This means that states with the same n, but different l are degenerate (i.e.
they have the same energy, although they are different states). The degree
of degeneracy of the states with quantum number n is


n−1
dn = (2l + 1) = n2 , (3.75)
l=0

i.e. for a given n there are n2 states with the same energy. This is also seen
in the energy diagram of the hydrogen atom in Fig. 3.9. In this diagram
the energy E = 0 is the vacuum level or ionisation energy – in a hydrogen
atom with an energy above this limit the electron is not bound to the proton
anymore.

3.3.4 Wave functions of the hydrogen atom


Altogether we have found recursion formulas for both the angular and radial
parts of the wave functions of the hydrogen atom,

ψnlml (r, θ, ϕ) = Rnl (r)Ylml (θ, ϕ). (3.76)


This puts us into the position to be able to investigate and discuss the solu-
tions quantitatively.

Notation and nomenclature


For historic reasons a particular nomenclature is used for the wave functions
with a particular n and l. Thus, instead of using l = 0, 1, 2, 3, 4, 5, ... one says
s, p, d, f, g, h, ..., where s stands for ”sharp”, p for ”principal”, d for ”diffuse”,
f for fundamental. The other letters just follow the alphabet. The reason
for these names will become clearer when we look at optical transitions in
the hydrogen atom. Wave functions with n = 1, 2, 3, 4, ... are said to belong
to the K−, L−, M −, N −, ... shells. The term ”shell” will become clearer
when we discuss the energy diagram of the hydrogen atom.
Very often the term orbital is as a synonym for the term wave function.
The most correct definition of ”orbital” is that it is the part of space in

71
Figure 3.9: Energy diagram for the hydrogen atom without consideration of
spin.
E (eV)

Shell
0 ...
...

...

...

...

...

...

...
N n=4 5g
M n=3 4f
3s 3p 3d
L n=2
2s 2p

1s
K n=1 -13.6
l=0 l=1 l=2 l=3 l=4 ...

72
which the electron is found with a certain, pre-defined probability. Thus, in
principal ”orbital” refers to the square of the wave function rather than the
wave function itself, but, nevertheless, the term is often used more sloppily
and interchangeably with ”wave function”, ”quantum mechanical state”, or
”square of the wave function”. The state with n = 1 and l = 0 is then called
the 1s−orbital, that with n = 2 and l = 0 the 2s−orbital, that with n = 2
and l = 1 the 2p−orbital etc.

Discussion of the radial solutions

Task: Plot the radial wave functions of the hydrogen atom, using e.g. MAT-
LAB, for n = 1, n = 2, and n = 3. What do you learn from the curves about
the hydrogen atom?

What is physically more interesting than the radial wave functions them-
selves are their squares (why?). Using the squares one derives the probability
of finding the electron between r and r + dr as follows:

∫ π ∫ 2π
dP = |R(r)|2 |Ylml (θ, ϕ)|2 r2 dr sin θ dθ dϕ = R2 (r)r2 dr, (3.77)
θ,0 ϕ,0

where we have used the fact that the spherical harmonics are normalised.
r2 R2 (r) is the radial probability density.

Task: Plot both R2 (r) and the radial probability densities for n = 1, n = 2,
and n = 3 and compare them. What is the main difference and why? Con-
sider the probability density at r = 0. Why is it zero? What do the densities
tell you about the size of the hydrogen atom in the different states? Without
considering the formula for the energy, if you compare wave functions with
different principal quantum numbers n - how do you think their energies
differ qualitatively and why?

73
.1 Liten atomfysikalisk och kvantmekanisk or-
dbok
Som man kan se från tabellen nedan är det oftast möjligt att direkt översatta
den engelska facktermen till svenska. För fullständighetens skull har jag ändå
tagit med sådana triviala översättningar och liksaså har jag tagit med sådana
begrepp som bör vara bekanta.

74
Engelska Svenska
adjoint, hermitian adjoint konjugat, hermiteskt konjugat
angular momentum rörelsemängdsmoment
aufbau principle uppbyggnadsprincipen
generaliserad rörelsemängd,
canonical momentum,
kanonisk rörelsemängd,
conjugate momentum,
kanoniskt konjugerad rörelsemängd,
canonically conjugate momentum
konjugerad rörelsemängd
central field approximation centralfältsapproximationen
closure relation fullständighetsrelation
completeness relation fullständighetsrelation
cross section tvärsnitt
decay sönderfall
degenerated degenererat
electron density elektrontäthet
fluorescence fluorescens
inner product, scalar product inre produkt, skalärprodukt
Lamb shift Lambskiftet
lifetime livstid
LS coupling LS-koppling
magnetic quantum number magnetiskt kvanttal
momentum rörelsemängd
orbital orbital
orbital angular momentum banrörelsemängdsmoment
orbital angular momentum quantum number bankvanttalet
outer product yttre produkt
parity paritet
perturbation theory störningsteori
photoionisation fotojonisation
principal quantum number huvudkvanttal
quantum number kvanttal
scalar product, inner product skalärprodukt, inre produkt
scatter, scattering sprida, spridning
selection rule urvalsregel
shell skal
spectral line spektrallinje
spherical harmonics klotytefunktioner
spin spinn
spin-orbit coupling spinn-ban-koppling
state tillstånd
transition övergång
translation 75 translation
unitary unitär
wave function vågfunktion
work function utträdesarbete
x-ray photoelectron spectroscopy röntgenfotoelektronspektroskopi

You might also like