Part IB Quantum Mechanics
Based on lectures by J. M. Evans
Notes taken by Dexter Chua
Michaelmas 2015
These notes are not endorsed by the lecturers, and I have modified them (often
significantly) after lectures. They are nowhere near accurate representations of what
was actually lectured, and in particular, all errors are almost surely mine.
Physical background
Photo electric effect. Electrons in atoms and line spectra. Particle diffraction. [1]
Schr¨odinger equation and solutions
De Broglie waves. Schr¨odinger equation. Superposition principle. Probability interpre-
tation, density and current. [2]
Stationary states. Free particle, Gaussian wave packet. Motion in 1-dimensional
potentials, parity. Potential step, square well and barrier. Harmonic oscillator. [4]
Observables and expectation values
Position and momentum op erators and expectation values. Canonical commutation
relations. Uncertainty principle. [2]
Observables and Hermitian operators. Eigenvalues and eigenfunctions. Formula for
expectation value. [2]
Hydrogen atom
Spherically sy mmetric wave functions for spherical well and hydrogen atom.
Orbital angular momentum operators. General solution of hydrogen atom. [5]
Contents
0 Introduction
0.1 Light quanta
0.2 Bohr model of the atom
0.3 Matter waves
1 Wavefunctions and the Schr¨odinger equation
1.1 Particle state and probability
1.2 Operators
1.3 Time evolution of wavefunctions
2 Some examples in one dimension
2.1 Introduction
2.2 Infinite well particle in a box
2.3 Parity
2.4 Potential well
2.5 The harmonic oscillator
3 Expectation and uncertainty
3.1 Inner products and expectation values
3.2 Ehrenfest’s theorem
3.3 Heisenberg’s uncertainty principle
4 More results in one dimensions
4.1 Gaussian wavepackets
4.2 Scattering
4.3 Potential step
4.4 Potential barrier
4.5 General features of stationary states
5 Axioms for quantum mechanics
5.1 States and observables
5.2 Measurements
5.3 Evolution in time
5.4 Discrete and continuous spectra
5.5 Degeneracy and simultaneous measurements
6 Quantum mechanics in three dimensions
6.1 Introduction
6.2 Separable eigenstate solutions
6.3 Angular momentum
6.4 Joint eigenstates for a spherically symmetric potential
7 The hydrogen atom
7.1 Introduction
7.2 General solution
7.3 Comments
0 Introduction
Quantum mechanics (QM) is a radical generalization of classical physics. Pro-
found new features of quantum mechanics include
(i)
Quantisation — Quantities such as energy are often restricted to a discrete
set of values, or appear in definite amounts , called quanta.
(ii)
Wave-particle duality Classical concepts of particles and waves become
merged in quantum mechanics. They are different aspects of a single entity.
So an electron will no longer be thought of a “particle” but an entity that
has properties of both particles and waves.
(iii)
Probability and uncertainty Predictions in quantum mechanics involve
probability in a fundamental way. This probability does not arise from our
lack of knowledge of the system, but is a genuine uncertainty in reality. In
particular, there are limits to what we can ask about a physical system,
even in principle. For example, the Heisenberg uncertainty principle entails
that we cannot accurately know both the position and momentum of a
particle.
Quantum mechanics also involves a new fundamental constant
h
or
~
=
h
2π
. The
dimension of this is
[h] = ML
2
T
1
= [energy] × [time] = [position] ×[momentum].
We can think of this constant as representing the “strength” of quantum effects.
Despite having these new profound features, we expect to recover classical physics
when we take the limit ~ 0.
Historically, there are a few experiments that led to the development of
quantum mechanics.
0.1 Light quanta
In quantum mechanics, light (or electromagnetic waves) consists of quanta called
photons. We can think of them as waves that come in discrete “packets” that
behave like particles.
In particular, photons behave like particles with energy
E
=
=
~ω
, where
ν
is the frequency and
ω
= 2
πν
is the angular frequency. However, we usually
don’t care about ν and just call ω the frequency.
Similarly, the momentum is given by
p
=
h/λ
=
~k
, where
λ
is the wavelength
and k = 2π is the wave number.
For electromagnetic waves, the speed is
c
=
ω/k
=
νλ
. This is consistent
with the fact that photons are massless particles, since we have
E = cp,
as entailed by special relativity.
Historically, the concept of quanta was introduced by Planck. At that time,
the classical laws of physics did not accurately depict how the spectrum of black-
body radiation will behave. In particular, it predicted that a black body will emit
an infinite amount of energy through radiation, which is clearly nonsense. Using
the concept of quanta and the energy-frequency relation, Planck was to derive the
spectrum of “black-body” radiation that is consistent with experimental results.
However, they were not yet sure that light indeed come in quanta physically. It
could have just been a mathematical trick to derive the desired result.
The physical reality of photons was clarified by Einstein in explaining the
photo-electric effect.
When we shine some light (or electromagnetic radiation
γ
) of frequency
ω
onto certain metals, this can cause an emission of electrons (
e
). We can measure
the maximum kinetic energy K of these electrons.
γ
e
Experiments show that
(i)
The number of electrons emitted depends on the intensity (brightness) of
the light, but not the frequency.
(ii)
The kinetic energy
K
depends only (linearly) on the frequency but not the
intensity.
(iii) For ω < ω
0
(for some critical value ω
0
), no electrons are emitted at all.
This is hard to understand classically, but is exactly as expected if each electron
emitted is due to the impact with a single photon. If
W
is the energy required
to liberate an electron, then we would expect
K
=
~ω W
by the conservation
of energy. We will have no emission if ω < ω
0
= W/~.
0.2 Bohr model of the atom
When we heat atoms up to make them emit light; or shine light at atoms so that
they absorb light, we will find that light is emitted and absorbed at very specific
frequencies, known as the emission and absorption spectra. This suggests that
the inner structure of atoms is discrete.
However, this is not the case in the classical model. In the classical model,
the simplest atom, the hydrogen, consists of an electron with charge
e
and
mass m, orbiting a proton of charge +e and mass m
p
m fixed at the origin.
The potential energy is
V (r) =
e
2
4πε
0
1
r
,
and the dynamics of the electron is governed by Newton’s laws of motions, just as
we derived the orbits of planets under the gravitational potential in IA Dynamics
and Relativity. This model implies that the angular momentum
L
is constant,
and so is the energy E =
1
2
mv
2
+ V (r).
This is not a very satisfactory model for the atom. First of all, it cannot
explain the discrete emission and absorption spectra. More importantly, while
this model seems like a mini solar system, electromagnetism behaves differently
from gravitation. To maintain a circular orbit, an acceleration has to be applied
onto the electron. Indeed, the force is given by
F =
mv
2
r
=
e
2
4πε
0
1
r
2
.
Accelerating particles emit radiation and lose energy. So according to classical
electrodynamics, the electron will just decay into the proton and atoms will
implode.
The solution to this problem is to simply declare that this cannot happen.
Bohr proposed the Bohr quantization conditions that restricts the classical orbits
by saying that the angular momentum can only take values
L = mrv = n~
for
n
= 1
,
2
, ···
. Using these, together with the force equation, we can solve
r
and v completely for each n and obtain
r
n
=
4πε
0
me
2
~
2
n
2
v
n
=
e
2
4πε
0
1
~n
E
n
=
1
2
m
e
2
4πε
0
~
2
1
n
2
.
Now we assume that the electron can make transitions between different energy
levels
n
and
m > n
, accompanied by emission or absorption of a photon of
frequency ω given by
E = ~ω = E
n
E
m
=
1
2
m
e
2
4πε
0
~
2
1
n
2
1
m
2
.
0
E
m
E
n
γ
This model explains a vast amount of experimental data. This also gives an
estimate of the size of a hydrogen atom:
r
1
=
4πε
0
me
2
~
2
5.29 × 10
11
m,
known as the Bohr radius.
While the model fits experiments very well, it does not provide a good
explanation for why the radius/angular momentum should be quantized. It
simply asserts this fact and then magically produces the desired results. Thus,
we would like a better understanding of why angular momentum should be
quantized.
0.3 Matter waves
The relations
E = = ~ω
p =
h
λ
= ~k
are used to associate particle properties (energy and momentum) to waves. They
can also be used the other way round to associate wave properties (frequency
and wave number) to particles. Moreover, these apply to non-relativistic particles
such as electrons (as well as relativistic photons). This
λ
is known as the de
Broglie wavelength.
Of course, nothing prevents us from assigning arbitrary numbers to our
particles. So an immediate question to ask is is there any physical significance
to the “frequency” and “wavenumber” of particles? Or maybe particles in fact
are waves?
Recall that the quantization of the Bohr model requires that
L = rp = n~.
Using the relations above, this is equivalent to requiring that
= 2πr.
This is exactly the statement that the circumference of the orbit is an integer
multiple of the wavelength. This is the condition we need for a standing wave
to form on the circumference. This looks promising as an explanation for the
quantization relation.
But in reality, do electrons actually behave like waves? If electrons really are
waves, then they should exhibit the usual behaviour of waves, such as diffraction
and interference.
We can repeat our favorite double-slit experiment on electrons. We have a
sinusoidal wave incident on some barrier with narrow openings as shown:
λ
wave
density of
electrons
δ
At different points, depending on the difference
δ
in path length, we may
have constructive interference (large amplitude) or destructive interference (no
amplitude). In particular, constructive interference occurs if
δ
=
, and
destructive if δ = (n +
1
2
)λ.
Not only does this experiment allow us to verify if something is a wave. We
can also figure out its wavelength λ by experiment.
Practically, the actual experiment for electrons is slightly more complicated.
Since the wavelength of an electron is rather small, to obtain the diffraction
pattern, we cannot just poke holes in sheets. Instead, we need to use crystals as
our diffraction grating. Nevertheless, this shows that electrons do diffract, and
the wavelength is the de Broglie wavelength.
This also has a conceptual importance. For regular waves, diffraction is
something we can make sense of. However, here we are talking about electrons.
We know that if we fire many many electrons, the distribution will follow the
pattern described above. But what if we just fire a single electron? On average,
it should still follow the distribution. However, for this individual electron, we
cannot know where it will actually land. We can only provide a probability
distribution of where it will end up. In quantum mechanics, everything is
inherently probabilistic.
As we have seen, quantum mechanics is vastly different from classical mechan-
ics. This is unlike special relativity, where we are just making adjustments to
Newtonian mechanics. In fact, in IA Dynamics and Relativity, we just “derived”
special relativity by assuming the principle of relativity and that the speed
of light is independent of the observer. This is not something we can do for
quantum mechanics — what we are going to do is just come up with some theory
and then show (or claim) that they agree with experiment.
1 Wavefunctions and the Schr¨odinger equation
The time evolution of particles in quantum mechanics is governed by the
Schr¨odinger equation, which we will come to shortly. In general, this is a
difficult equation to solve, and there aren’t many interesting cases where we are
able to provide a full solution.
Hence, to begin with, we will concentrate on quantum mechanics in one
(spatial) dimension only, as the maths is much simpler and diagrams are easier
to draw, and we can focus more on the physical content.
1.1 Particle state and probability
Classically, a point particle in 1 dimension has a definitive position
x
(and
momentum
p
) at each time. To completely specify a particle, it suffices to write
down these two numbers. In quantum mechanics, this is much more complicated.
Instead, a particle has a state at each time, specified by a complex-valued
wavefunction ψ(x).
The physical content of the wavefunction is as follows: if
ψ
is appropriately
normalized, then when we measure the position of a particle, we get a result
x
with probability density function
|ψ
(
x
)
|
2
, i.e. the probability that the position
is found in [
x, x
+
δx
] (for small
δx
) is given by
|ψ
(
x
)
|
2
δx
. Alternatively, the
probability of finding it in an interval [a, b] is given by
P(particle position in [a, b]) =
Z
b
a
|ψ(x)|
2
dx.
What do we mean by “appropriately normalized”? From our equation above, we
see that we require
Z
−∞
|ψ(x)|
2
dx = 1,
since this is the total probability of finding the particle anywhere at all. This
the normalization condition required.
Example (Gaussian wavefunction). We define
ψ(x) = Ce
(xc)
2
2α
,
where c is real and C could be complex.
c
We have
Z
−∞
|ψ(x)|
2
dx = |C|
2
Z
−∞
e
(xc)
2
α
dx = |C|
2
(απ)
1
2
= 1.
So for normalization, we need to pick C = (1/απ)
1/4
(up to a multiple of e
).
If
α
is small, then we have a sharp peak around
x
=
c
. If
α
is large, it is
more spread out.
While it is nice to have a normalized wavefunction, it is often inconvenient
to deal exclusively with normalized wavefunctions, or else we will have a lot of
ugly-looking constants floating all around. As a matter of fact, we can always
restore normalization at the end of the calculation. So we often don’t bother.
If we do not care about normalization, then for any (non-zero)
λ
,
ψ
(
x
) and
λψ
(
x
) represent the same quantum state (since they give the same probabilities).
In practice, we usually refer to either of these as “the state”. If we like fancy
words, we can thus think of the states as equivalence classes of wavefunctions
under the equivalence relation ψ φ if φ = λψ for some non-zero λ.
What we do require, then, is not that the wavefunction is normalized, but
normalizable, i.e.
Z
−∞
|ψ(x)|
2
dx < .
We will very soon encounter wavefunctions that are not normalizable. Mathe-
matically, these are useful things to have, but we have to be more careful when
interpreting these things physically.
A characteristic property of quantum mechanics is that if
ψ
1
(
x
) and
ψ
2
(
x
) are
wavefunctions for a particle, then
ψ
1
(
x
) +
ψ
2
(
x
) is also a possible particle state
(ignoring normalization), provided the result is non-zero. This is the principle of
superposition, and arises from the fact that the equations of quantum mechanics
are linear.
Example (Superposition of Gaussian wavefunctions). Take
ψ(x) = B
exp
(x c)
2
2α
+ exp
x
2
2β

.
Then the resultant distribution would be something like
We choose
B
so that
ψ
in a normalized wavefunction for a single particle. Note
that this is not two particles at two different positions. It is one particle that is
“spread out” at two different positions.
It is possible that in some cases, the particles in the configuration space
may be restricted. For example, we might require
`
2
x
`
2
with some
boundary conditions at the edges. Then the normalization condition would not
be integrating over (−∞, ), but [
`
2
,
`
2
].
1.2 Operators
We know that the square of the wavefunction gives the probability distribution of
the position of the particle. How about other information such as the momentum
and energy? It turns out that all the information about the particle is contained
in the wavefunction (which is why we call it the “state” of the particle).
We call each property of the particle which we can measure an observable.
Each observable is represented by an operator acting on
ψ
(
x
). For example, the
position is represented by the operator
ˆx
=
x
. This means that (
ˆ
)(
x
) =
(
x
).
We can list a few other operators:
position ˆx = x ˆ = (x)
momentum ˆp = i~
x
ˆ = i~ψ
0
(x)
energy H =
ˆp
2
2m
+ V (ˆx) Hψ =
~
2
2m
ψ
00
(x) + V (x)ψ(x)
The final
H
is called the Hamiltonian, where
m
is the mass and
V
is the potential.
We see that the Hamiltonian is just the kinetic energy
p
2
2m
and the potential
energy
V
. There will be more insight into why the operators are defined like
this in IIC Classical Dynamics and IID Principles of Quantum Mechanics.
Note that we put hats on
ˆx
and
ˆp
to make it explicit that these are operators,
as opposed to the classical quantities position and momentum. Otherwise, the
definition ˆx = x would look silly.
How do these operators relate to the actual physical properties? In general,
when we measure an observable, the result is not certain. They are randomly
distributed according to some probability distribution, which we will go into full
details later.
However, a definite result is obtained if and only if
ψ
is an eigenstate, or
eigenfunction, of the operator. In this case, results of the measurements are the
eigenvalue associated. For example, we have
ˆ =
if and only if ψ is a state with definite momentum p. Similarly,
Hψ = Eψ
if and only if ψ has definite energy E.
Here we are starting to see why quantization occurs in quantum mechanics.
Since the only possible values of
E
and
p
are the eigenvalues, if the operators
have a discrete set of eigenvalues, then we can only have discrete values of
p
and
E.
Example. Let
ψ(x) = Ce
ikx
.
This has a wavelength of
λ
= 2
π/k
. This is a momentum eigenstate, since we
have
ˆ = ~ψ
0
= (~k)ψ.
So we know that the momentum eigenvalue is p = ~k. This looks encouraging!
Note that if there is no potential, i.e. V = 0, then
Hψ =
ˆp
2
2m
ψ =
~
2
2m
ψ
00
=
~
2
k
2
2m
ψ.
So the energy eigenvalue is
E =
~
2
k
2
2m
.
Note, however, that our wavefunction has
|ψ
(
x
)
|
2
=
|C|
2
, which is a constant.
So this wavefunction is not normalizable on the whole line. However, if we
restrict ourselves to some finite domain
`
2
x
`
2
, then we can normalize by
picking C =
1
`
.
Example. Consider the Gaussian distribution
ψ(x) = C exp
x
2
2α
.
We get
ˆ(x) = i~ψ
0
(x) 6= (x)
for any number p. So this is not an eigenfunction of the momentum.
However, if we consider the harmonic oscillator with potential
V (x) =
1
2
Kx
2
,
then this
ψ
(
x
) is an eigenfunction of the Hamiltonian operator, provided we
picked the right α. We have
Hψ =
~
2
2m
ψ
00
+
1
2
Kx
2
ψ = Eψ
when
α
2
=
~
2
Km
. Then the energy is
E
=
~
2
q
K
m
. This is to be verified on the
example sheet.
Despite being a simple system, the harmonic oscillator is incredibly useful in
theoretical physics. We will hence solve this completely later.
Definition
(Time-independent Schr¨odinger equation)
.
The time-independent
Schr¨odinger equation is the energy eigenvalue equation
Hψ = Eψ,
or
~
2
2m
ψ
00
+ V (x)ψ = Eψ.
This is in general what determines what the system behaves. In particular,
the eigenvalues E are precisely the allowed energy values.
1.3 Time evolution of wavefunctions
So far, everything is instantaneous. The wavefunction specifies the state at a
particular time, and the eigenvalues are the properties of the system at that
particular time. However, this is quantum mechanics, or quantum dynamics. We
should be looking at how things change. We want to know how the wavefunction
changes with time. This is what we will get to now.
Time-dependent Schr¨odinger equation
We will write Ψ instead of
ψ
to indicate that we are looking at the time-dependent
wavefunction. The evolution of this Ψ(
x, t
) is described by the time-dependent
Schr¨odinger equation.
Definition
(Time-dependent Schr¨odinger equation)
.
For a time-dependent
wavefunction Ψ(x, t), the time-dependent Schr¨odinger equation is
i~
Ψ
t
= HΨ. ()
For a particle in a potential V (x), this can reads
i~
Ψ
t
=
~
2
2m
2
Ψ
x
2
+ V (x.
While this looks rather scary, it isn’t really that bad. First of all, it is linear. So
the sums and multiples of solutions are also solutions. It is also first-order in
time. So if we know the wavefunction Ψ(
x, t
0
) at a particular time
t
0
, then this
determines the whole function Ψ(x, t).
This is similar to classical dynamics, where knowing the potential
V
(and
hence the Hamiltonian
H
) completely specifies how the system evolves with
time. However, this is in some ways different from classical dynamics. Newton’s
second law is second-order in time, while this is first-order in time. This is
significant since when our equation is first-order in time, then the current state
of the wavefunction completely specifies the evolution of the wavefunction in
time.
Yet, this difference is just an illusion. The wavefunction is the state of the
particle, and not just the “position”. Instead, we can think of it as capturing the
position and momentum. Indeed, if we write the equations of classical dynamics
in terms of position and momentum, it will be first order in time.
Stationary states
It is not the coincidence that the time-independent Schr¨odinger equation and
the time-dependent Schr¨odinger equation are named so similarly (and it is also
not an attempt to confuse students).
We perform separation of variables, and consider a special class of solutions
Ψ(
x, t
) =
T
(
t
)
ψ
(
x
), where Ψ(
x,
0) =
ψ
(
x
) (i.e.
T
(0) = 1). If
ψ
satisfies the
time-independent Schr¨odinger equation
Hψ = Eψ,
then since
H
does not involve time derivatives, we know Ψ is an energy eigenstate
at each fixed t, i.e.
HΨ = EΨ.
So if we want this Ψ to satisfy the Schr¨odinger equation, we must have
i~
˙
T = ET.
The solution is obvious:
T (t) = exp
iEt
~
.
We can write our full solution as
Ψ(x, t) = ψ(x) exp
iEt
~
.
Note that the frequency is
ω
=
E
~
. So we recover the Energy-frequency relation
we’ve previously had.
Definition (Stationary state). A stationary state is a state of the form
Ψ(x, t) = ψ(x) exp
iEt
~
.
where
ψ
(
x
) is an eigenfunction of the Hamiltonian with eigenvalue
E
. This term
is also sometimes applied to ψ instead.
While the stationary states seem to be a rather peculiar class of solutions
that would rarely correspond to an actual physical state in reality, they are in
fact very important in quantum mechanics. The reason is that the stationary
states form a basis of the state space. In other words, every possible state can be
written as a (possibly infinite) linear combination of stationary states. Hence, by
understanding the stationary states, we can understand a lot about a quantum
system.
Conservation of probability
Note that for a stationary state, we have
|Ψ(x, t)|
2
= |ψ(x)|
2
,
which is independent of time. In general, this is true in most cases.
Consider a general Ψ(
x, t
) obeying the time-dependent Schr¨odinger equation.
Proposition. The probability density
P (x, t) = |Ψ(x, t)|
2
obeys a conservation equation
P
t
=
j
x
,
where
j(x, t) =
i~
2m
Ψ
dx
dx
Ψ
is the probability current.
Since Ψ
Ψ
0
is the complex conjugate of Ψ
0∗
Ψ, we know that Ψ
Ψ
0
Ψ
0∗
Ψ is
imaginary. So multiplying by
i
ensures that
j
(
x, t
) is real, which is a good thing
since P is also real.
Proof.
This is straightforward from the Schr¨odinger equation and its complex
conjugate. We have
P
t
= Ψ
Ψ
t
+
Ψ
t
Ψ
= Ψ
i~
2m
Ψ
00
i~
2m
Ψ
00∗
Ψ
where the two V terms cancel each other out, assuming V is real
=
j
x
.
The important thing here is not the specific form of
j
, but that
P
t
can be
written as the space derivative of some quantity. This implies that the probability
that we find the particle in [a, b] at fixed time t changes as
d
dt
Z
b
a
|Ψ(x, t)|
2
dx =
Z
b
a
j
x
(x, t) dx = j(a, t) j(b, t).
We can think of the final term as the probability current getting in and out of
the interval at the boundary.
In particular, consider a normalizable state with Ψ
,
Ψ
0
, j
0 as
x ±∞
for fixed t. Taking a −∞ and b +, we have
d
dt
Z
−∞
|Ψ(x, t)|
2
dx = 0.
What does this tell us? This tells us that if Ψ(
x,
0) is normalized, Ψ(
x, t
) is
normalized for all
t
. Hence we know that for each fixed
t
,
|
Ψ(
x, t
)
|
2
is a probability
distribution. So what this really says is that the probability interpretation is
consistent with the time evolution.
2 Some examples in one dimension
2.1 Introduction
In general, we are going to consider the energy eigenvalue problem for a particle
in 1 dimension in a potential V (x), i.e.
Hψ =
~
2
2m
ψ
00
+ V (x)ψ = Eψ.
In other words, we want to find the allowed energy eigenvalues.
This is a hard problem in general. However, we can consider the really easy
case where
V
(
X
) =
U
, where
U
is a constant. Then we can easily write down
solutions.
If U > E, then the Schr¨odinger equation is equivalent to
ψ
00
κ
2
ψ = 0,
where κ is such that U E =
~
2
κ
2
2m
. We take wlog κ > 0. The solution is then
ψ = Ae
κx
+ Be
κx
.
On the other hand, if U < E, then the Schr¨odinger equation says
ψ + k
2
ψ = 0,
where k is picked such that E U =
~
2
k
2
2m
. The solutions are
ψ = Ae
ikx
+ Be
ikx
.
Note that these new constants are merely there to simplify our equations. They
generally need not have physical meanings.
Now why are we interested in cases where the potential is constant? Wouldn’t
that be just equivalent to a free particle? This is indeed true. However, knowing
these solutions allow us to to study piecewise flat potentials such as steps, wells
and barriers.
x
V
x
V
x
V
Here a finite discontinuity in
V
is allowed. In this case, we can have
ψ, ψ
0
continuous and
ψ
00
discontinuous. Then the discontinuity of
ψ
00
cancels that of
V , and the Schr¨odinger equation holds everywhere.
In this chapter, we will seek normalizable solutions with
Z
−∞
|ψ(x)|
2
dx.
This requires that
ψ
(
x
)
0 as
x ±∞
. We see that for the segments and
the end, we want to have decaying exponentials
e
κx
instead of oscillating
exponentials e
ikx
.
2.2 Infinite well particle in a box
The simplest case to consider is the infinite well. Here the potential is infinite
outside the region [
a, a
], and we have much less to think about. For
|x| > a
,
we must have ψ(x) = 0, or else V (x)ψ(x) would be infinite.
x
V
a
a
V (x) =
(
0 |x| a
∞ |x| > a.
We require
ψ
= 0 for
|x| > a
and
ψ
continuous at
x
=
±a
. Within
|x| < a
, the
Schr¨odinger equation is
~
2
2m
ψ
00
= Eψ.
We simplify this to become
ψ
00
+ k
2
ψ = 0,
where
E =
~
2
k
2
2m
.
Here, instead of working with the complex exponentials, we use
sin
and
cos
since
we know well when these vanish. The general solution is thus
ψ = A cos kx + B sin kx.
Our boundary conditions require that ψ vanishes at x = ±a. So we need
A cos ka ±B sin ka = 0.
In other words, we require
A cos ka = B sin ka = 0.
Since
sin ka
and
cos ka
cannot be simultaneously 0, either
A
= 0 or
B
= 0. So
the two possibilities are
(i) B = 0 and ka = /2 with n = 1, 3, ···
(ii) A = 0 and ka = /2 with n = 2, 4, ···
Hence the allowed energy levels are
E
n
=
~
2
π
2
8ma
2
n
2
,
where n = 1, 2, ···, and the wavefunctions are
ψ
n
(x) =
1
a
1
2
(
cos
x
2a
n odd
sin
x
2a
n even
.
These are normalized with
R
a
a
|ψ
n
(x)|
2
dx.
ψ
1
:
x
V
a a
ψ
2
:
x
V
a a
ψ
3
:
x
V
a a
ψ
4
:
x
V
a a
This was a rather simple and nice example. We have an infinite well, and the
particle is well-contained inside the box. The solutions just look like standing
waves on a string with two fixed end points something we (hopefully) are
familiar with.
Note that
ψ
n
(
x
) = (
1)
n+1
ψ
n
(
x
). We will see that this is a general feature
of energy eigenfunctions of a symmetric potential. This is known as parity.
2.3 Parity
Consider the Schr¨odinger equation for a particle of mass m
Hψ =
~
2
2m
ψ
00
+ V (x)ψ = Eψ.
with potential
V (x) = V (x).
By changing variables
x x
, we see that
ψ
(
x
) is an eigenfunction of
H
with
energy
E
if and only if
ψ
(
x
) is an eigenfunction of
H
with energy
E
. There
are two possibilities:
(i)
If
ψ
(
x
) and
ψ
(
x
) represent the same quantum state, this can only happen
if
ψ
(
x
) =
ηψ
(
x
) for some constant
η
. Since this is true for all
x
, we can
do this twice and get
ψ(x) = ηψ(x) = η
2
ψ(x).
So we get that
η
=
±
1 and
ψ
(
x
) =
±ψ
(
x
). We call
η
the parity, and say
ψ has even/odd parity if η is +1/ 1 respectively.
For example, in our particle in a box, our states ψ
n
have parity (1)
n+1
.
(ii)
If
ψ
(
x
) and
ψ
(
x
) represent different quantum states, then we can still
take linear combinations
ψ
±
(x) = α(ψ(x) ±ψ(x)),
and these are also eigenstates with energy eigenvalue
E
, where
α
is for
normalization. Then by construction,
ψ
±
(
x
) =
±ψ
±
(
x
) and have parity
η = ±1.
Hence, if we are given a potential with reflective symmetry
V
(
x
) =
V
(
x
), then
we can restrict our attention and just look for solutions with definite parity.
2.4 Potential well
We will consider a potential that looks like this:
x
V
a
a
U
The potential is given by
V (x) =
(
U |x| < a
0 |x| a
for some constant
U >
0. Classically, this is not very interesting. If the energy
E <
0, then the particle is contained in the well. Otherwise it is free to move
around. However, in quantum mechanics, this is much more interesting.
We want to seek energy levels for a particle of mass
m
, defined by the
Schr¨odinger equation
Hψ =
~
2
2m
ψ
00
+ V (x)ψ = Eψ.
For energies in the range
U < E < 0,
we set
U + E =
~
2
k
2
2m
> 0, E =
~
2
κ
2
2m
,
where
k, κ >
0 are new real constants. Note that these coefficients are not
independent, since U is given and fixed. So they must satisfy
k
2
+ κ
2
=
2mU
~
2
.
Using these constants, the Schr¨odinger equation becomes
(
ψ
00
+ k
2
ψ = 0 |x| < a
ψ
00
κ
2
ψ = 0 |x| > a.
As we previously said, we want the Schr¨odinger equation to hold even at the
discontinuities. So we need ψ and ψ
0
to be continuous at x = ±a.
We first consider the even parity solutions
ψ
(
x
) =
ψ
(
x
). We can write our
solution as
ψ =
(
A cos kx |x| < a
Be
κ|x|
|x| > a
We match ψ and ψ
0
at x = a. So we need
A cos ka = Be
κa
Ak sin ka = κBe
κa
.
By parity, there is no additional information from x = a.
We can divide the equations to obtain
k tan ka = κ.
this is still not something we can solve easily. To find when solutions exist, it is
convenient to introduce
ξ = ak, η = aκ,
where these two constants are dimensionless and positive. Note that this
η
has
nothing to do with parity. It’s just that we have run out of letters to use. Hence
the solution we need are solutions to
η = ξ tan ξ.
Also, our initial conditions on k and κ require
ξ
2
+ η
2
=
2ma
2
U
~
2
.
We can look for solutions by plotting these two equations. We first plot the
curve η = ξ tan ξ:
ξ
η
1π
2
3π
2
5π
2
7π
2
9π
2
11π
2
The other equation is the equation of a circle. Depending on the size of the
constant 2ma
2
U/~
2
, there will be a different number of points of intersections.
ξ
η
So there will be a different number of solutions depending on the value of
2ma
2
U/~
2
. In particular, if
(n 1)π <
2mUa
2
~
2
1/2
< ,
then we have exactly n even parity solutions (for n 1).
We can do exactly the same thing for odd parity eigenstates. . . on example
sheet 1.
For
E >
0 or
E < U
, we will end up finding non-normalizable solutions.
What is more interesting, though is to look at the solutions we have now. We
can compare what we’ve got with what we would expect classically.
Classically, any value of
E
in the range
U < E <
0 is allowed, and the
motion is deeply uninteresting. The particle just goes back and forth inside the
well, and is strictly confined in a x a.
Quantum mechanically, there is just a discrete, finite set of allowed energies.
What is more surprising is that while
ψ
decays exponentially outside the well, it
is non-zero! This means there is in theory a non-zero probability of finding the
particle outside the well! We call these particles bound in the potential, but in
fact there is a non-zero probability of finding the particle outside the well.
2.5 The harmonic oscillator
So far in our examples, the quantization (mathematically) comes from us requiring
continuity at the boundaries. In the harmonic oscillator, it arises in a different
way.
x
V
This is a harmonic oscillator of mass m with
V (x) =
1
2
2
x
2
.
Classically, this has a motion of
x
=
A cos ω
(
t t
0
), which is something we
(hopefully) know well too.
This is a really important example. First of all, we can solve it, which is a
good thing. More importantly, any smooth potential can be approximated by a
harmonic oscillator near an equilibrium x
0
, since
V (x) = V (x
0
) +
1
2
V
00
(x
0
)(x x
0
)
2
+ ··· .
Systems with many degrees like crystals can also be treated as collections of
independent oscillators by considering the normal modes. If we apply this to
the electromagnetic field, we get photons! So it is very important to understand
the quantum mechanical oscillator.
We are going to seek all normalizable solutions to the time-independent
Schr¨odinger equation
Hψ =
~
2
2m
ψ
00
+
1
2
2
x
2
ψ = Eψ.
So simplify constants, we define
y =
~
1
2
x, E =
2E
~ω
,
both of which is dimensionless. Then we are left with
d
2
ψ
dy
2
+ y
2
ψ = Eψ.
We can consider the behaviour for
y
2
E
. For large
y
, the
y
2
ψ
term will
be large, and so we want the
ψ
00
term to offset it. We might want to try the
Gaussian
e
y
2
/2
, and when we differentiate it twice, we would have brought
down a factor of y
2
. So we can wlog set
ψ = f(y)e
1
2
y
2
.
Then the Schr¨odinger equation gives
d
2
f
dy
2
2y
df
dy
+ (E 1) = 0.
This is known as Hermite’s equation. We try a series solution
f(y) =
X
r0
a
r
y
r
,
and substitute in to get
X
r0
(r + 2)(r + 1)a
n+2
+ (E 1 2r)a
r
y
r
= 0.
This holds if and only if
a
r+2
=
2r + 1 E
(r + 2)(r + 1)
a
r
, r 0.
We can choose
a
0
and
a
1
independently, and can get two linearly independent
solutions. Each solution involves either all even or all odd powers.
However, we have a problem. We want normalizable solutions. So we want
to make sure our function does not explode at large
y
. Note that it is okay if
f
(
y
) is quite large, since our
ψ
is suppressed by the
e
1
2
y
2
terms, but we cannot
grow too big.
We look at these two solutions individually. To examine the behaviour of
f(y) when y is large, observe that unless the coefficients vanish, we get
a
p
/a
p2
1
p
.
This matches the coefficients of
y
α
e
y
2
for some power
α
(e.g.
P
p0
y
2p
p!
). This
is bad, since our ψ will then grow as e
1
2
y
2
, and cannot be normalized.
Hence, we get normalizable
ψ
if and only if the series for
f
terminates to
give a polynomial. This occurs iff
E
= 2
n
+ 1 for some
n
. Note that for each
n
,
only one of the two independent solutions is normalizable. So for each
E
, we get
exactly one solution.
So for n even, we have
a
r+2
=
2r 2n
(r + 2)(r + 1)
a
r
for r even, and a
r
= 0 for r odd, and the other way round when n is odd.
The solutions are thus
f
(
y
) =
h
n
(
y
), where
h
n
is a polynomial of degree
n
with h
n
(y) = (1)
n
h
n
(y).
For example, we have
h
0
(y) = a
0
h
1
(y) = a
1
y
h
2
(y) = a
0
(1 2y
2
)
h
3
(y) = a
1
y
2
3
y
3
.
These are known as the Hermite polynomials. We have now solved our harmonic
oscillator. With the constant restored, the possible energy eigenvalues are
E
n
= ~ω
n +
1
2
,
for n = 0, 1, 2, ···.
The wavefunctions are
ψ
n
(x) = h
n
~
1
2
x
exp
1
2
~
x
2
,
where normalization fixes a
0
and a
1
.
As we said in the beginning, harmonic oscillators are everywhere. It turns
out quantised electromagnetic fields correspond to sums of quantised harmonic
oscillators, with
E
n
E
0
= n~ω
This is equivalent to saying the
n
th state contains
n
photons, each of energy
~ω
.
3 Expectation and uncertainty
So far, all we’ve been doing is solving Schr¨odinger’s equation, which is arguably
not very fun. In this chapter, we will look at some theoretical foundations of
quantum mechanics. Most importantly, we will learn how to extract information
from the wavefunction
ψ
. Two important concepts will be the expectation
and uncertainty. We will then prove two important theorems about these
Ehrenfest’s theorem and the Heisenberg uncertainty principle.
3.1 Inner products and expectation values
Definitions
Definition
(Inner product)
.
Let
ψ
(
x
) and
φ
(
x
) be normalizable wavefunctions
at some fixed time (not necessarily stationary states). We define the complex
inner product by
(φ, ψ) =
Z
−∞
φ(x)
ψ(x) dx.
Note that for any complex number α, we have
(φ, αψ) = α(φ, ψ) = (α
φ, ψ).
Also, we have
(φ, ψ) = (ψ, φ)
.
These are just the usual properties of an inner product.
Definition
(Norm)
.
The norm of a wavefunction
ψ
, written,
kψk
is defined by
kψk
2
= (ψ, ψ) =
Z
−∞
|ψ(x)|
2
dx.
This ensures the norm is real and positive.
Suppose we have a normalized state
ψ
, i.e.
kψk
= 1, we define the expectation
values of observables as
Definition
(Expectation value)
.
The expectation value of any observable
H
on
the state ψ is
hHi
ψ
= (ψ, Hψ).
For example, for the position, we have
hˆxi
ψ
= (ψ, ˆ) =
Z
−∞
x|ψ(x)|
2
dx.
Similarly, for the momentum, we have
hˆpi
ψ
= (ψ, ˆ) =
Z
−∞
ψ
(i~ψ
0
) dx.
How are we supposed to interpret this thing? So far, all we have said about
operators is that if you are an eigenstate, then measuring that property will give
a definite value. However, the point of quantum mechanics is that things are
waves. We can add them together to get superpositions. Then the sum of two
eigenstates will not be an eigenstate, and does not have definite, say, momentum.
This formula tells us what the average value of any state is.
This is our new assumption of quantum mechanics the expectation value
is the mean or average of results obtained by measuring the observable many
times, with the system prepared in state ψ before each measurement.
Note that this is valid for any operator. In particular, we can take any
function of our existing operators. One important example is the uncertainty:
Definition
(Uncertainty)
.
The uncertainty in position (∆
x
)
ψ
and momentum
(∆p)
ψ
are defined by
(∆x)
2
ψ
= h(ˆx hˆxi
ψ
)
2
i
ψ
= hˆx
2
i
ψ
hˆxi
2
ψ
,
with exactly the same expression for momentum:
(∆p)
2
ψ
= h(ˆp hˆpi
ψ
)
2
i
ψ
= hˆp
2
i
ψ
hˆpi
2
ψ
,
We will later show that these quantities (∆
x
)
2
ψ
and (∆
y
)
2
ψ
are indeed real
and positive, so that this actually makes sense.
Hermitian operators
The expectation values defined can be shown to be real for
ˆx
and
ˆp
specifically,
by manually fiddling with stuff. We can generalize this result to a large class of
operators known as Hermitian operators.
Definition
(Hermitian operator)
.
An operator
Q
is Hermitian iff for all nor-
malizable φ, ψ, we have
(φ, ) = (Qφ, ψ).
In other words, we have
Z
φ
dx =
Z
()
ψ dx.
In particular, this implies that
(ψ, ) = (, ψ) = (ψ, )
.
So (ψ, ) is real, i.e. hQi
ψ
is real.
Proposition.
The operators
ˆx
,
ˆp
and
H
are all Hermitian (for real potentials).
Proof.
We do
ˆx
first: we want to show (
φ, ˆ
) = (
ˆxφ, ψ
). This statement is
equivalent to
Z
−∞
φ(x)
(x) dx =
Z
−∞
((x))
ψ(x) dx.
Since position is real, this is true.
To show that
ˆp
is Hermitian, we want to show (
φ, ˆ
) = (
ˆpφ, ψ
). This is
equivalent to saying
Z
−∞
φ
(i~ψ
0
) dx =
Z
−∞
(i~φ
0
)
ψ dx.
This works by integrating by parts: the difference of the two terms is
i~[φ
ψ]
−∞
= 0
since φ, ψ are normalizable.
To show that H is Hermitian, we want to show (φ, Hψ) = (Hφ, ψ), where
H =
h
2
2m
d
2
dx
2
+ V (x).
To show this, it suffices to consider the kinetic and potential terms separately.
For the kinetic energy, we just need to show that (
φ, ψ
00
) = (
φ
00
, ψ
). This is true
since we can integrate by parts twice to obtain
(φ, ψ
00
) = (φ
0
, ψ
0
) = (φ
00
, ψ).
For the potential term, we have
(φ, V (ˆx)ψ) = (φ, V (x)ψ) = (V (x)φ, ψ) = (V (ˆx)φ, ψ).
So H is Hermitian, as claimed.
Thus we know that hxi
ψ
, hˆpi
ψ
, hHi
ψ
are all real.
Furthermore, observe that
X = ˆx α, P = ˆp β
are (similarly) Hermitian for any real α, β. Hence
(ψ, X
2
ψ) = (ψ, X(Xψ)) = (Xψ, Xψ) = kXψk
2
0.
Similarly, we have
(ψ, P
2
ψ) = (ψ, P (P ψ)) = (P ψ, P ψ) = kP ψk
2
0.
If we choose
α
=
hˆxi
ψ
and
β
=
hˆpi
ψ
, the expressions above say that (∆
x
)
2
ψ
and
(∆p)
2
ψ
are indeed real and positive.
Cauchy-Schwarz inequality
We are going to end with a small section on a technical result that will come
handy later on.
Proposition
(Cauchy-Schwarz inequality)
.
If
ψ
and
φ
are any normalizable
states, then
kψkkφk |(ψ, φ)|.
Proof. Consider
kψ + λφk
2
= (ψ + λφ, ψ + λφ)
= (ψ, ψ) + λ(ψ, φ) + λ
(φ, ψ) + |λ|
2
(φ, φ) 0.
This is true for any complex λ. Set
λ =
(φ, ψ)
kφk
2
which is always well-defined since
φ
is normalizable, and then the above equation
becomes
kψk
2
|(ψ, φ)|
2
kφk
2
0.
So done.
3.2 Ehrenfest’s theorem
We will show that, in fact, quantum mechanics is like classical mechanics.
Again, consider a normalizable state Ψ(
x, t
) satisfying the time-dependent
Schr¨odinger equation, i.e.
i~
˙
Ψ = HΨ =
ˆp
2
2m
+ V (ˆx)
Ψ.
Classically, we are used to
x
and
p
changing in time. However, here
ˆx
and
ˆp
are
fixed in time, while the states change with time. However, what does change
with time is the expectations. The expectation values
hˆxi
Ψ
= (Ψ, ˆxΨ), hˆpi
Ψ
= (Ψ, ˆpΨ)
depend on t through Ψ. Ehrenfest’s theorem states the following:
Theorem (Ehrenfest’s theorem).
d
dt
hˆxi
Ψ
=
1
m
hˆpi
Ψ
d
dt
hˆpi
Ψ
= −hV
0
(ˆx)i
Ψ
.
These are the quantum counterparts to the classical equations of motion.
Proof. We have
d
dt
hˆxi
Ψ
= (
˙
Ψ, ˆxΨ) + , ˆx
˙
Ψ)
=
1
i~
HΨ, ˆxΨ
+
Ψ, ˆx
1
i~
H
Ψ
Since H is Hermitian, we can move it around and get
=
1
i~
, H(ˆxΨ)) +
1
i~
, ˆx(HΨ))
=
1
i~
, (ˆxH H ˆx)Ψ).
But we know
(ˆxH H ˆx)Ψ =
~
2
2m
(xΨ
00
(xΨ)
00
) + (xV Ψ V xΨ) =
~
2
m
Ψ
0
=
i~
m
ˆpΨ.
So done.
The second part is similar. We have
d
dt
hˆpi
Ψ
= (
˙
Ψ, ˆpΨ) + , ˆp
˙
Ψ)
=
1
i~
HΨ, ˆpΨ
+
Ψ, ˆp
1
i~
H
Ψ
Since H is Hermitian, we can move it around and get
=
1
i~
, H(ˆpΨ)) +
1
i~
, ˆp(HΨ))
=
1
i~
, (ˆpH H ˆp)Ψ).
Again, we can compute
(ˆpH H ˆp)Ψ = i~
~
2
2m
((Ψ
00
)
0
0
)
00
) i~((V (x)Ψ)
0
V (x
0
)
= i~V
0
(x.
So done.
Note that in general, quantum mechanics can be portrayed in different
“pictures”. In this course, we will be using the Schr¨odinger picture all the time,
in which the operators are time-independent, and the states evolve in time. An
alternative picture is the Heisenberg picture, in which states are fixed in time,
and all the time dependence lie in the operators. When written in this way,
quantum mechanics is even more like classical mechanics. This will be explored
more in depth in IID Principles of Quantum Mechanics.
3.3 Heisenberg’s uncertainty principle
We will show that, in fact, quantum mechanics is not like classical mechanics.
Statement
The statement of the uncertainty principle (or relation) is
Theorem
(Heisenberg’s uncertainty principle)
.
If
ψ
is any normalized state (at
any fixed time), then
(∆x)
ψ
(∆p)
ψ
~
2
.
Example. Consider the normalized Gaussian
ψ(x) =
1
απ
1
4
e
x
2
.
We find that
hˆxi
ψ
= hˆpi
ψ
= 0,
and also
(∆x)
2
ψ
=
α
2
, (∆p)
2
ψ
=
~
2
2α
.
So we get
(∆x)
ψ
(∆p)
ψ
=
~
2
.
We see that a small
α
corresponds to
ψ
sharply peaked around
x
= 0, i.e. it has
a rather definite position, but has a large uncertainty in momentum. On the
other hand, if
α
is large,
ψ
is spread out in position but has a small uncertainty
in momentum.
Recall that for
α
=
~
, this is the lowest energy eigenstate for harmonic
oscillator with
H =
1
2m
ˆp +
1
2
2
ˆx
2
,
with eigenvalue
1
2
~ω
. We can use the uncertainty principle to understand why
we have a minimum energy of
1
2
~ω
instead of 0. If we had a really small energy,
then we would just sit at the bottom of the potential well and do nothing, with
both a small (and definite) momentum and position. Hence for uncertainty to
occur, we need a non-zero ground state energy.
To understand where the uncertainty principle come from, we have to under-
stand commutation relations.
Commutation relations
We are going to define a weird thing called the commutator. At first sight, this
seems like a weird definition, but this turns out to be a really important concept
in quantum mechanics.
Definition
(Commutator)
.
Let
Q
and
S
be operators. Then the commutator
is denoted and defined by
[Q, S] = QS SQ.
This is a measure of the lack of commutativity of the two operators.
In particular, the commutator of position and momentum is
[ˆx, ˆp] = ˆxˆp ˆpˆx = i~.
This relation results from a simple application of the product rule:
(ˆxˆp ˆpˆx)ψ = i~ψ
0
(i~()
0
= i~ψ.
Note that if α and β are any real constants, then the operators
X = ˆx α, P = ˆp β
also obey
[X, P ] = i~.
This is something we will use when proving the uncertainty principle.
Recall that when we proved Ehrenfest’s theorem, the last step was to calculate
the commutator:
[ˆx, H] = ˆxH H ˆx =
i~
m
ˆp
[ˆp, H] = ˆpH H ˆp = i~V
0
(ˆx).
Since
H
is defined in terms of
ˆp
and
ˆx
, we can indeed find these relations just
using the commutator relations between
ˆx
and
ˆp
(plus a few more basic facts
about commutators).
Commutator relations are important in quantum mechanics. When we first
defined the momentum operator
ˆp
as
i~
x
, you might have wondered where
this weird definition came from. This definition naturally comes up if we require
that
ˆx
is “multiply by
x
(so that the delta function
δ
(
x x
0
) is the eigenstate
with definition position
x
0
), and that
ˆx
and
ˆp
satisfies this commutator relation.
With this requirements, ˆp must be defined as that derivative.
Then one might ask, why would we want
ˆx
and
ˆp
to satisfy this commutator
relation? It turns out that in classical dynamics, there is something similar
known as the Poisson bracket
{·, ·}
, where we have
{x, p}
= 1. To get from
classical dynamics to quantum mechanics, we just have to promote our Poisson
brackets to commutator brackets, and multiply by i~.
Proof of uncertainty principle
Proof of uncertainty principle. Choose α = hˆxi
ψ
and β = hˆpi
ψ
, and define
X = ˆx α, P = ˆp β.
Then we have
(∆x)
2
ψ
= (ψ, X
2
ψ) = (Xψ, Xψ) = kXψk
2
(∆p)
2
ψ
= (ψ, P
2
ψ) = (P ψ, P ψ) = kP ψk
2
Then we have
(∆x)
ψ
(∆p)
ψ
= kXψkkPψk
|(Xψ, P ψ)|
|im(Xψ, P ψ)|
1
2i
h
(Xψ, P ψ) (P ψ, Xψ)
i
=
1
2i
h
(ψ, XP ψ) (ψ, P Xψ)
i
=
1
2i
(ψ, [X, P ]ψ)
=
~
2
(ψ, ψ)
=
~
2
.
So done.
4 More results in one dimensions
4.1 Gaussian wavepackets
When we solve Schr¨odinger’s equation, what we get is a “wave” that represents
the probability of finding our thing at each position. However, in real life,
we don’t think of particles as being randomly distributed over different places.
Instead, particles are localized to some small regions of space.
These would be represented by wavefunctions in which most of the distribution
is concentrated in some small region:
These are known as wavepackets.
Definition
(Wavepacket)
.
A wavefunction localised in space (about some point,
on some scale) is usually called a wavepacket.
This is a rather loose definition, and can refer to anything that is localized in
space. Often, we would like to consider a particular kind of wavepacket, known
as a Gaussian wavepacket.
Definition (Gaussian wavepacket). A Gaussian wavepacket is given by
Ψ
0
(x, t) =
α
π
1/4
1
γ(t)
1/2
e
x
2
/2γ(t)
,
for some γ(t).
These are particularly nice wavefunctions. For example, we can show that for
a Gaussian wavepacket, (∆
x
)
ψ
(∆
p
)
ψ
=
~
2
exactly, and uncertainty is minimized.
The Gaussian wavepacket is a solution of the time-dependent Schr¨odinger
equation (with V = 0) for
γ(t) = α +
i~
m
t.
Substituting this
γ
(
t
) into our equation, we find that the probability density is
P
0
(x, t) = |Ψ
0
(x, t)|
2
=
α
π
1/2
1
|γ(t)|
e
αx
2
/|γ(t)|
2
,
which is peaked at
x
= 0. This corresponds to a particle at rest at the origin,
spreading out with time.
A related solution to the time-dependent Schr¨odinger equation with
V
= 0 is
a moving particle:
Ψ
u
(x, t) = Ψ
0
(x ut) exp
i
mu
~
x
exp
i
mu
2
2~
t
.
The probability density resulting from this solution is
P
u
(x, t) = |Ψ
u
(x, t)|
2
= P
0
(x ut, t).
So this corresponds to a particle moving with velocity u. Furthermore, we get
hˆpi
Ψ
u
= mu.
This corresponds with the classical momentum, mass × velocity.
We see that wavepackets do indeed behave like particles, in the sense that we
can set them moving and the quantum momentum of these objects do indeed
behave like the classical momentum. In fact, we will soon attempt to send them
to a brick wall and see what happens.
In the limit
α
, our particle becomes more and more spread out in
space. The uncertainty in the position becomes larger and larger, while the
momentum becomes more and more definite. Then the wavefunction above
resembles something like
Ψ(x, t) = Ce
ikx
e
iEt/~
,
which is a momentum eigenstate with
~k
=
mu
and energy
E
=
1
2
mu
2
=
~
2
k
2
2m
.
Note, however, that this is not normalizable.
4.2 Scattering
Consider the time-dependent Schr¨odinger equation with a potential barrier. We
would like to send a wavepacket towards the barrier and see what happens.
u
Ψ
Classically, we would expect the particle to either pass through the barrier or get
reflected. However, in quantum mechanics, we would expect it to “partly” pass
through and “partly” get reflected. So the resultant wave is something like this:
AΨ
ref
BΨ
tr
Here Ψ, Ψ
ref
and Ψ
tr
are normalized wavefunctions, and
P
ref
= |A|
2
, P
tr
= |B|
2
.
are the probabilities of reflection and transmission respectively.
This is generally hard to solve. Scattering problems are much simpler to
solve for momentum eigenstates of the form
e
ikx
. However, these are not
normalizable wavefunctions, and despite being mathematically convenient, we
are not allowed to use them directly, since they do not make sense physically.
These, in some sense, represent particles that are “infinitely spread out” and
can appear anywhere in the universe with equal probability, which doesn’t really
make sense.
There are two ways we can get around this problem. We know that we can
construct normalized momentum eigenstates for a single particle confined in a
box
`
2
x
`
2
, namely
ψ(x) =
1
`
e
ikx
,
where the periodic boundary conditions require
ψ
(
x
+
`
) =
ψ
(
x
), i.e.
k
=
2πn
`
for some integer
n
. After calculations have been done, the box can be removed
by taking the limit ` .
Identical results are obtained more conveniently by allowing Ψ(
x, t
) to repre-
sent beams of infinitely many particles, with
|
Ψ(
x, t
)
|
2
being the density of the
number of particles (per unit length) at
x, t
. When we do this, instead of having
one particle and watching it evolve, we constantly send in particles so that the
system does not appear to change with time. This allows us to find steady
states. Mathematically, this corresponds to finding solutions to the Schr¨odinger
equation that do not change with time. To determine, say, the probability of
reflection, roughly speaking, we look at the proportion of particles moving left
compared to the proportion of particles moving right in this steady state.
In principle, this interpretation is obtained by considering a constant stream
of wavepackets and using some limiting/averaging procedure, but we usually
don’t care about these formalities.
For these particle beams, Ψ(
x, t
) is bounded, but no longer normalizable.
Recall that for a single particle, the probability current was defined as
j(x, t) =
i~
2m
Ψ
0
ΨΨ
0∗
).
If we have a particle beam instead of a particle, and Ψ is the particle density
instead of the probability distribution, j now represents the flux of particles at
x, t, i.e. the number of particles passing the point x in unit time.
Recall that a stationary state of energy
E
is of the form Ψ(
x, t
) =
ψ
(
x
)
e
iEt/~
.
We have
|Ψ(x, t)|
2
= |ψ(x)|
2
,
and
j(x, t) =
i~
2m
(ψ
ψ
0
ψψ
0∗
).
Often, when solving a scattering problem, the solution will involve sums of
momentum eigenstates. So it helps to understand these better.
Our momentum eigenstates are
ψ(x) = Ce
ikx
,
which are solutions to the time-independent Schr¨odinger equation with
V
= 0
with E =
~
2
k
2
2m
.
Applying the momentum operator, we find that
p
=
~k
is the momentum of
each particle in the beam, and
|ψ
(
x
)
|
2
=
|C|
2
is the density of particles in the
beam. We can also evaluate the current to be
j =
~k
m
|C|
2
.
This makes sense.
~k
m
=
p
m
is the velocity of the particles, and
|C|
2
is how many
particles we have. So this still roughly corresponds to what we used to have
classically.
In scattering problems, we will seek the transmitted and reflected flux
j
tr
,
j
ref
in terms of the incident flux
j
inc
, and the probabilities for transmission and
reflection are then given by
P
tr
=
|j
tr
|
|j
inc
|
, P
ref
=
|j
ref
|
|j
inc
|
.
4.3 Potential step
Consider the time-independent Schr¨odinger equation for a step potential
V (x) =
(
0 x 0
U x > 0
,
where
U >
0 is a constant. The Schr¨odinger equation is, in case you’ve forgotten,
~
2
2m
ψ
00
+ V (x)ψ = Eψ.
x
V (x)
We require ψ and ψ
0
to be continuous at x = 0.
We can consider two different cases:
(i)
0
< E < U
: We apply the standard method, introducing constants
k, κ >
0
such that
E =
~
2
k
2
2m
, U E =
~
2
κ
2
2m
.
Then the Schr¨odinger equation becomes
(
ψ
00
+ k
2
ψ = 0 x < 0
ψ
00
κ
2
ψ = 0 x > 0
The solutions are
ψ
=
Ie
ikx
+
Re
ikx
for
x <
0, and
ψ
=
Ce
κx
for
x >
0
(since ψ has to be bounded).
Since ψ and ψ
0
are continuous at x = 0, we have the equations
(
I + R = C
ikI ikR = κC
.
So we have
R =
k
k +
I, C =
2k
k +
I.
If
x <
0,
ψ
(
x
) is a superposition of beams (momentum eigenstates) with
|I|
2
particles per unit length in the incident part, and
|R|
2
particles per
unit length in the reflected part, with p = ±~k. The current is
j = j
inc
+ j
ref
= |I|
2
~k
m
|R|
2
~k
m
,
The probability of reflection is
P
ref
=
|j
ref
|
|j
inc
|
=
|R|
2
|I|
2
= 1,
which makes sense.
On the right hand side, we have
j
= 0. So
P
tr
= 0. However,
|ψ
(
x
)
|
2
6
= 0
in this classically forbidden region.
(ii) E > U: This time, we set
E =
~
2
k
2
2m
, E U =
~
2
κ
2
2m
,
with k, κ > 0. Then the Schr¨odinger equation becomes
(
ψ
00
+ k
2
ψ = 0 x < 0
ψ
00
+ κ
2
ψ = 0 x > 0
Then we find
ψ
=
Ie
ikx
+
R
ikx
on
x <
0, with
ψ
=
T e
ikx
on
x >
0. Note
that it is in principle possible to get an
e
ikx
term on
x >
0, but this
would correspond to sending in a particle from the right. We, by choice,
assume there is no such term.
We now match ψ and ψ
0
at x = 0. Then we get the equations
(
I + R = T
ikI ikR = ikT.
We can solve these to obtain
R =
k κ
k + κ
I, T =
2k
k + κ
I.
Our flux on the left is now
j = j
inc
+ j
ref
= |I|
2
~k
m
|R|
2
~k
m
,
while the flux on the right is
j = j
tr
|T |
2
~κ
m
.
The probability of reflection is
P
ref
=
|j
ref
|
|j
inc
|
=
|R|
2
|I|
2
=
k κ
k + κ
2
,
while the probability of transmission is
P
tr
=
|j
tr
|
|j
inc
|
=
|T |
2
κ
|I|
2
k
=
4kκ
(k + κ)
2
.
Note that P
ref
+ P
tr
= 1.
Classically, we would expect all particles to be transmitted, since they all
have sufficient energy. However, quantum mechanically, there is still a
probability of being reflected.
4.4 Potential barrier
So far, the results were slightly weird. We just showed that however high energy
we have, there is still a non-zero probability of being reflected by the potential
step. However, things are really weird when we have a potential barrier.
Consider the following potential:
x
V
U
0
a
We can write this as
V (x) =
0 x 0
U 0 < x < a
0 x a
We consider a stationary state with energy
E
with 0
< E < U
. We set the
constants
E =
~
2
k
2
2m
, U E =
~
2
κ
2
2m
.
Then the Schr¨odinger equations become
ψ
00
+ k
2
ψ = 0 x < 0
ψ
00
κ
2
ψ = 0 0 < x < a
ψ
00
+ k
2
ψ = 0 x > a
So we get
ψ = Ie
ikx
+ Re
ikx
x < 0
ψ = Ae
κx
+ Be
κx
0 < x < a
ψ = T e
ikx
x > a
Matching ψ and ψ
0
at x = 0 and a gives the equations
I + R = A + B
ik(I R) = κ(A B)
Ae
κa
+ Be
κa
= T e
ika
κ(Ae
κa
Be
κa
) = ikTe
ika
.
We can solve these to obtain
I +
κ ik
κ + ik
R = T e
ika
e
κa
I +
κ + ik
κ ik
R = T e
ika
e
κa
.
After lots of some algebra, we obtain
T = Ie
ika
cosh κa i
k
2
κ
2
2kκ
sinh κa
1
To interpret this, we use the currents
j = j
inc
+ j
ref
= (|I|
2
|R|
2
)
~k
m
for x < 0. On the other hand, we have
j = j
tr
= |T |
2
~k
m
for
x > a
. We can use these to find the transmission probability, and it turns
out to be
P
tr
=
|j
trj
|
|j
inc
|
=
|T |
2
|I|
2
=
1 +
U
2
4E(U E)
sinh
2
κa
1
.
This demonstrates quantum tunneling. There is a non-zero probability that the
particles can pass through the potential barrier even though it classically does
not have enough energy. In particular, for
κa
1, the probability of tunneling
decays as
e
2κa
. This is important, since it allows certain reactions with high
potential barrier to occur in practice even if the reactants do not classically have
enough energy to overcome it.
4.5 General features of stationary states
We are going to end the chapter by looking at the difference between bound
states and scattering states in general.
Consider the time-independent Schr¨odinger equation for a particle of mass
m
Hψ =
~
2
2m
ψ
00
+ V (x)ψ = Eψ,
with the potential
V
(
x
)
0 as
x ±∞
. This is a second order ordinary
differential equation for a complex function
ψ
, and hence there are two complex
constants in general solution. However, since this is a linear equation, this
implies that 1 complex constant corresponds to changing
ψ 7→ λψ
, which gives
no change in physical state. So we just have one constant to mess with.
As |x| , our equation simply becomes
~
2
2m
ψ
00
= Eψ.
So we get
ψ
(
Ae
ikx
+ Be
ikx
E =
~
2
k
2
2m
> 0
Ae
κx
+ Be
κx
E =
~
2
κ
2
2m
< 0.
So we get two kinds of stationary states depending on the sign of
E
. These
correspond to bound states and scattering states.
Bound state solutions, E < 0
If we want ψ to be normalizable, then there are 2 boundary conditions for ψ:
ψ
(
Ae
κx
x −∞
Be
κx
x +
This is an overdetermined system, since we have too many boundary conditions.
Solutions exist only when we are lucky, and only for certain values of
E
. So
bound state energy levels are quantized. We may find several bound states or
none.
Definition
(Ground and excited states)
.
The lowest energy eigenstate is called
the ground state. Eigenstates with higher energies are called excited states.
Scattering state solutions, E > 0
Now
ψ
is not normalized but bounded. We can view this as particle beams, and
the boundary conditions determines the direction of the incoming beam. So we
have
ψ
(
Ie
ikx
+ Re
ikx
x −∞
T e
ikx
x +
This is no longer overdetermined since we have more free constants. The solution
for any E > 0 (imposing condition on one complex constant) gives
j
(
j
inc
+ j
ref
j
tr
=
(
|I|
2
~k
m
|R|
2
~k
m
x −∞
|T |
2
~k
m
x +
We also get the reflection and transmission probabilities
P
ref
= |A
ref
|
2
=
|j
ref
|
|j
inc
|
P
tr
= |A
tr
|
2
=
|j
tr
|
|j
inc
|
,
where
A
ref
(k) =
R
I
A
tr
(k) =
T
I
are the reflection and transmission amplitudes. In quantum mechanics, “ampli-
tude” general refers to things that give probabilities when squared.
5 Axioms for quantum mechanics
We are going to sum up everything we’ve had so far into a big picture. To do
so, we will state a few axioms of quantum mechanics and see how they relate to
what we have done so far.
Our axioms (or postulates) will be marked by bullet points in this chapter,
as we go through them one by one.
5.1 States and observables
States of a quantum system correspond to non-zero elements of a complex
vector space
V
(which has nothing to do with the potential), with
ψ
and
αψ physically equivalent for all α C \ {0}.
Furthermore, there is a complex inner product (
φ, ψ
) defined on
V
satisfying
(φ, α
1
ψ
1
+ α
2
ψ
2
) = α
1
(φ, ψ
1
) + α
2
(φ, ψ
2
)
(β
1
φ
1
+ β
2
φ
2
, ψ) = β
1
(φ
1
, ψ) + β
2
(φ
2
, ψ)
(φ, ψ) = (ψ, φ)
kψk
2
= (ψ, ψ) 0
kψk = 0 if and only if ψ = 0.
Note that this is just the linear algebra definition of what a complex inner
product should satisfy.
An operator A is a linear map V V satisfying
A(αψ + βφ) = αAψ + βAφ.
For any operator
A
, the Hermitian conjugate or adjoint, denoted
A
is
defined to be the unique operator satisfying
(φ, A
ψ) = (Aφ, ψ).
An operator Q is called Hermitian or self-adjoint if
Q
= Q.
A state χ 6= 0 is an eigenstate of Q with eigenvalue λ if
= λχ.
The set of all eigenvalues of Q is called the spectrum of Q.
A measurable quantity, or observable, in a quantum system corresponds to
a Hermitian operator.
So far, we have worked with the vector space of functions (that are sufficiently
smooth), and used the integral as the inner product. However, in general, we can
work with arbitrary complex vector spaces and arbitrary inner products. This
will become necessary when we, say, study the spin of electrons in IID Principles
of Quantum Mechanics.
Even in the case of the vector space of (sufficiently smooth) functions, we will
only work with this formalism informally. When we try to make things precise,
we will encounter a lot of subtleties such as operators not being well-defined
everywhere.
5.2 Measurements
Key results for Hermitian operators
Our next set of axioms relate Hermitian operators with physical measurements.
Before we state them, we first look at some purely mathematical properties of
Hermitian operators. We first prove some purely mathematical results about
Hermitian operators.
Proposition. Let Q be Hermitian (an observable), i.e. Q
= Q. Then
(i) Eigenvalues of Q are real.
(ii)
Eigenstates of
Q
with different eigenvalues are orthogonal (with respect to
the complex inner product).
(iii)
Any state can be written as a (possibly infinite) linear combination of
eigenstates of
Q
, i.e. eigenstates of
Q
provide a basis for
V
. Alternatively,
the set of eigenstates is complete.
Note that the last property is not actually necessarily true. For example,
the position and momentum operators do not have eigenstates at all, since the
“eigenstates” are not normalizable. However, for many of the operators we care
about, this is indeed true, and everything we say in this section only applies to
operators for which this holds.
Proof.
(i) Since Q is Hermitian, we have, by definition,
(χ, ) = (Qχ, χ).
Let χ be an eigenvector with eigenvalue λ, i.e. = λχ. Then we have
(χ, λχ) = (λχ, χ).
So we get
λ(χ, χ) = λ
(χ, χ).
Since (χ, χ) 6= 0, we must have λ = λ
. So λ is real.
(ii) Let = λχ and = µφ. Then we have
(φ, ) = (Qφ, χ).
So we have
(φ, λχ) = (µφ, χ).
In other words,
λ(φ, χ) = µ
(φ, χ) = µ(φ, χ).
Since λ 6= µ by assumption, we know that (φ, χ) = 0.
(iii) We will not attempt to justify this, or discuss issues of convergence.
Measurement axioms
Consider an observable
Q
with discrete spectrum (i.e. eigenvalues) and normalized
eigenstates. From the results we just had, any given state can be written
ψ =
X
n
α
n
χ
n
,
with
n
= λ
n
χ
n
, and (χ
m
, χ
n
) = δ
mn
. This means we can calculate α
n
by
α
n
= (χ
n
, ψ).
We can assume also that
α
n
6
= 0 since we can just pretend the terms with
coefficient 0 do not exist, and that the corresponding
λ
n
are distinct by choosing
an appropriate basis. Then
The outcome of a measurement is some eigenvalue of Q.
The probability of obtaining λ
n
is
P
n
= |α
n
|
2
,
where α
n
= (χ
n
, ψ) is the amplitude.
The measurement is instantaneous and forces the system into the state
χ
n
.
This is the new state immediately after the measurement is made.
The last statement is rather weird. However, if you think hard, this must be
the case. If we measure the state of the system, and get, say, 3, then if we
measure it immediately afterwards, we would expect to get the result 3 again with
certainty, instead of being randomly distributed like the original state. So after
a measurement, the system must be forced into the corresponding eigenstate.
Note that these axioms are consistent in the following sense. If
ψ
is normalized,
then
(ψ, ψ) =
X
α
m
χ
m
,
X
α
n
χ
n
=
X
m,n
α
m
α
n
(χ
m
, χ
n
)
=
X
m,n
α
m
α
n
δ
mn
=
X
n
|α
n
|
2
=
X
n
P
n
= 1
So if the state is normalized, then the sum of probabilities is indeed 1.
Example.
Consider the harmonic oscillator, and consider the operator
Q
=
H
with eigenfunctions χ
n
= ψ
n
and eigenvalues
λ
n
= E
n
= ~ω
n +
1
2
.
Suppose we have prepared our system in the state
ψ =
1
6
(ψ
0
+ 2ψ
1
4
).
Then the coefficients are
α
0
=
1
6
, α
1
=
2
6
, α
4
=
i
6
.
This is normalized since
kψk
2
= |α
0
|
2
+ |α
1
|
2
+ |α
4
|
2
= 1.
Measuring the energy gives the following probabilities:
Energy Probability
E
0
=
1
2
~ω P
0
=
1
6
E
1
=
3
2
~ω P
1
=
2
3
E
4
=
9
2
~ω P
4
=
1
6
If a measurement gives
E
1
, then
ψ
1
is the new state immediately after measure-
ment.
Expressions for expectation and uncertainty
We can use these probability amplitudes to express the expectation and uncer-
tainty of a state in a more familiar way.
Proposition.
(i) The expectation value of Q in state ψ is
hQi
ψ
= (ψ, ) =
X
λ
n
P
n
,
with notation as in the previous part.
(ii) The uncertainty (δQ)
ψ
is given by
(∆Q)
2
ψ
= h(Q hQi
ψ
)
2
i
ψ
= hQ
2
i
ψ
hQi
2
ψ
=
X
n
(λ
n
hQi
ψ
)
2
P
n
.
From this, we see that
ψ
is an eigenstate of
Q
with eigenvalue
λ
if and only
if hQi
ψ
= λ and (∆Q)
ψ
= 0.
Proof.
(i) Let ψ =
P
α
n
χ
n
. Then
=
X
α
n
λ
n
χ
n
.
So we have
(ψ, ) =
X
n,m
(α
m
, χ
n
, α
n
λ
n
χ
n
) =
X
n
α
n
α
n
λ
n
=
X
λ
n
P
n
.
(ii) This is a direct check using the first expression.
Example.
Consider the harmonic oscillator as we had in the previous example.
Then the expectation value is
hHi
ψ
=
X
n
E
n
P
n
=
1
2
·
1
6
+
3
2
·
2
3
+
9
2
·
1
6
~ω =
11
6
~ω.
Note that the measurement axiom tells us that after measurement, the
system is then forced into the eigenstate. So when we said that we interpret the
expectation value as the “average result for many measurements”, we do not
mean measuring a single system many many times. Instead, we prepare a lot of
copies of the system in state ψ, and measure each of them once.
5.3 Evolution in time
The state of a quantum system Ψ(t) obeys the Schr¨odinger equation
i~
˙
Ψ = HΨ,
where
H
is a Hermitian operator, the Hamiltonian; this holds at all times
except at the instant a measurement is made.
In principle, this is all there is in quantum mechanics. However, for practical
purposes, it is helpful to note the following:
Stationary states
Consider the energy eigenstates with
Hψ
n
= E
n
ψ
n
, (ψ
m
, ψ
n
) = δ
mn
.
Then we have certain simple solutions of the Schr¨odinger equation of the form
Ψ
n
= ψ
n
e
iE
n
t/~
.
In general, given an initial state
Ψ(0) =
X
n
α
n
ψ
n
,
since the Schr¨odinger equation is linear, we can get the following solution for all
time:
Ψ(t) =
X
n
α
n
e
iE
n
t/~
ψ
n
.
Example. Consider again the harmonic oscillator with initial state
Ψ(0) =
1
6
(ψ
0
+ 2ψ
1
4
).
Then Ψ(t) is given by
Ψ(t) =
1
6
(ψ
0
e
t/2
+ 2ψ
1
e
3t/2
4
e
9t/2
).
Ehrenfest’s theorem (general form)
Theorem
(Ehrenfest’s theorem)
.
If
Q
is any operator with no explicit time
dependence, then
i~
d
dt
hQi
Ψ
= h[Q, H]i
Ψ
,
where
[Q, H] = QH HQ
is the commutator.
Proof. If Q does not have time dependence, then
i~
d
dt
, QΨ) = (i~
˙
Ψ, QΨ) + , Qi~
˙
Ψ)
= (HΨ, QΨ) + , QHΨ)
= (Ψ, (QH HQ)Ψ)
= (Ψ, [Q, H]Ψ).
If
Q
has explicit time dependence, then we have an extra term on the right,
and have
i~
d
dt
hQi
Ψ
= h[Q, H]i
Ψ
+ i~h
˙
Qi
Ψ
.
These general versions correspond to classical equations of motion in Hamiltonian
formalism you will (hopefully) study in IIC Classical Dynamics.
5.4 Discrete and continuous spectra
In stating the measurement axioms, we have assumed that our spectrum of
eigenvalues of
Q
was discrete, and we got nice results about measurements.
However, for certain systems, the spectra of
ˆp
and
H
may be continuous. This
is the same problem when we are faced with non-normalizable momentum
wavefunctions.
To solve this problem, we can make the spectra discrete by a technical device:
we put the system in a “box” of length
`
with suitable boundary conditions of
ψ
(
x
). We can then take
`
at the end of the calculation. We’ve discussed
this in some extent for momentum eigenstates. We shall revisit that scenario in
the more formal framework we’ve developed.
Example.
Consider
ψ
(
x
) with periodic boundary conditions
ψ
(
x
+
`
) =
ψ
(
x
).
So we can restrict to
`
2
x
`
2
.
We compare with the general axioms, where
Q = ˆp = i~
d
dx
.
The eigenstates are
χ
n
(x) =
1
`
e
ik
n
x
, k
n
=
2πn
`
.
Now we have discrete eigenvalues as before given by
λ
n
= ~k
n
.
We know that the states are orthonormal on
`
2
x
`
2
, i.e.
(χ
n
, χ
m
) =
Z
`
2
`
2
χ
m
(x)
χ
n
(x) dx = δ
mn
.
We can thus expand our solution in terms of the eigenstates to get a complex
Fourier series
ψ(x) =
X
n
α
n
χ
n
(x),
where the amplitudes are given by
α
n
= (χ
n
, ψ).
When we take the limit as
n
, the Fourier series becomes a Fourier integral.
There is another approach to this problem (which is non-examinable). We
can also extend from discrete to continuous spectra as follows. We replace the
discrete label n with some continuous label ξ. Then we have the equation
ξ
= λ
ξ
χ
ξ
.
These are eigenstates with an orthonormality conditions
(χ
ξ
, χ
η
) = δ(ξ η),
where we replaced our old
δ
mn
with
δ
(
ξ η
), the Dirac delta function. To
perform the expansion in eigenstates, the discrete sum becomes an integral. We
have
ψ =
Z
α
ξ
χ
ξ
dξ,
where the coefficients are
α
ξ
= (χ
ξ
, ψ).
In the discrete case,
|α
n
|
2
is the probability mass function. The obvious gener-
alization here would be to let
|α
ξ
|
2
be our probability density function. More
precisely,
Z
b
a
|α
ξ
|
2
dξ = probability that the result corresponds to a ξ b.
Example.
Consider the particle in one dimension with position as our operator.
We will see that this is just our previous interpretation of the wavefunction.
Let our operator be Q = ˆx. Then our eigenstates are
χ
ξ
(x) = δ(x ξ).
The corresponding eigenvalue is
λ
ξ
= ξ.
This is true since
ˆ
ξ
(x) = (x ξ) = ξδ(x ξ) = ξχ
ξ
(x),
since δ(x ξ) is non-zero only when x = ξ.
With this eigenstates, we can expand this is
ψ(x) =
Z
α
ξ
χ
ξ
(x) dξ =
Z
α
ξ
δ(x ξ) dξ = α
x
.
So our coefficients are given by α
ξ
= ψ(ξ). So
Z
b
a
|ψ(ξ)|
2
dξ
is indeed the probability of measuring the particle to be in
a ξ b
. So we
recover our original interpretation of the wavefunction.
We see that as long as we are happy to work with generalized functions like
delta functions, things become much nicer and clearer. Of course, we have to be
more careful and study distributions properly if we want to do this.
5.5 Degeneracy and simultaneous measurements
Degeneracy
Definition
(Degeneracy)
.
For any observable
Q
, the number of linearly inde-
pendent eigenstates with eigenvalue
λ
is the degeneracy of the eigenvalue. In
other words, the degeneracy is the dimension of the eigenspace
V
λ
= {ψ : = λψ}.
An eigenvalue is non-degenerate if the degeneracy is exactly 1, and is degenerate
if the degeneracy is more than 1.
We say two states are degenerate if they have the same eigenvalue.
In one dimension, the energy bound states are always non-degenerate, as we
have seen in example sheet 2. However, in three dimensions, energy bound states
may be degenerate. If
λ
is degenerate, then there is a large freedom in choosing
an orthonormal basis for
V
λ
. Physically, we cannot distinguish degenerate states
by measuring
Q
alone. So we would like to measure something else as well to
distinguish these eigenstates. When can we do this? We have previously seen
that we cannot simultaneously measure position and momentum. It turns out
the criterion for whether we can do this is simple.
Commuting observables
Recall that after performing a measurement, the state is forced into the corre-
sponding eigenstate. Hence, to simultaneously measure two observables
A
and
B
, we need the state to be a simultaneously an eigenstate of
A
and
B
. In other
words, simultaneous measurement is possible if and only if there is a basis for
V
consisting of simultaneous or joint eigenstates χ
n
with
n
= λ
n
χ
n
, Bχ
n
= µ
n
χ
n
.
Our measurement axioms imply that if the state is in
χ
n
, then measuring
A
,
B
in rapid succession, in any order, will give definite results
λ
n
and
µ
n
respectively
(assuming the time interval between each pair of measurements is short enough
that we can neglect the evolution of state in time).
From IB Linear Algebra, a necessary and sufficient condition for
A
and
B
to
be simultaneously measurable (i.e. simultaneously diagonalizable) is for
A
and
B to commute, i.e.
[A, B] = AB BA = 0.
This is consistent with a “generalized uncertainty relation”
(∆A)
ψ
(∆B)
ψ
1
2
|h[A, B]i
ψ
|,
since if we if have a state that is simultaneously an eigenstate for
A
and
B
, then
the uncertainties on the left would vanish. So
h
[
A, B
]
i
ψ
= 0. The proof of this
relation is on example sheet 3.
This will be a technique we will use to tackle degeneracy in general. If we have
a system where, say, the energy is degenerate, we will attempt to find another
operator that commutes with
H
, and try to further classify the underlying states.
When dealing with the hydrogen atom later, we will use the angular momentum
to separate the degenerate energy eigenstates.
6 Quantum mechanics in three dimensions
Finally, we will move on and discuss quantum systems in three dimensions, since
we (probably) live in a three-dimensional world. These problems are usually
substantially more difficult, since we have to deal with partial derivatives. Also,
unlike the case of one dimension, we will run into degenerate states, where
multiple eigenstates have the same energy. We thus have to find some other
observables (that commute with the Hamiltonian) in order to fully classify the
eigenstates.
6.1 Introduction
To begin with, we translate everything we’ve had for the one-dimensional world
into the three-dimensional setting.
A quantum state of a particle in three dimensions is given by a wavefunction
ψ
(
x
) at fixed time, or Ψ(
x, t
) for a state evolving in time. The inner product is
defined as
(ϕ, ψ) =
Z
ϕ(x)
ψ(x) d
3
x.
We adopt the convention that no limits on the integral means integrating over
all space. If ψ is normalized, i.e.
kψk
2
= (ψ, ψ) =
Z
|ψ(x)|
2
d
3
x = 1,
then the probability of measuring the particle to be inside a small volume
δV
(containing x) is
|ψ(x)|
2
δV.
The position and momentum are Hermitian operators
ˆ
x = (ˆx
1
, ˆx
2
, ˆx
3
), ˆx
i
ψ = x
i
ψ
and
ˆ
p = (ˆp
1
, ˆp
2
, ˆp
3
) = i~ = i~
x
1
,
x
2
,
x
3
We have the canonical commutation relations
[ˆx
i
, ˆp
j
] = i~δ
ij
, [ˆx
i
, ˆx
j
] = [ˆp
i
, ˆp
j
] = 0.
We see that position and momentum in different directions don’t come into
conflict. We can have a definite position in one direction and a definite momentum
in another direction. The uncertainty principle only kicks in when we are in the
same direction. In particular, we have
(∆x
i
)(∆p
j
)
~
2
δ
ij
.
Similar to what we did in classical mechanics, we assume our particles are
structureless.
Definition
(Structureless particle)
.
A structureless particle is one for which all
observables can be written in terms of position and momentum.
In reality, many particles are not structureless, and (at least) possess an
additional quantity known as “spin”. We will not study these in this course, but
only mention it briefly near the end. Instead, we will just pretend all particles
are structureless for simplicity.
The Hamiltonian for a structureless particle in a potential V is
H =
ˆ
p
2
2m
+ V (
ˆ
x) =
~
2
2m
2
+ V (x).
The time-dependent Schr¨odinger equation is
i~
Ψ
t
= HΨ =
~
2
2m
2
Ψ + V (x.
The probability current is defined as
j =
i~
2m
Ψ ΨΨ
).
This probability current obeys the conservation equation
t
|Ψ(x, t)|
2
= −∇ · j.
This implies that for any fixed volume V ,
d
dt
Z
V
|Ψ(x, t)|
2
d
3
x =
Z
V
· j d
3
x =
Z
V
j · dS,
and this is true for any fixed volume V with boundary V . So if
|Ψ(x, t)| 0
sufficiently rapidly as |x| , then the boundary term disappears and
d
dt
Z
|Ψ(x, t)|
2
d
3
x = 0.
This is the conservation of probability (or normalization).
6.2 Separable eigenstate solutions
How do we solve the Schr¨odinger equation in three dimensions? Things are
more complicated, since we have partial derivatives. Often, those we can solve
are those we can reduce to one-dimensional problems, using the symmetry of
the system. For example, we will solve the hydrogen atom exploiting the fact
that the potential is spherically symmetric. To do so, we will use the method of
separation of variables, as we’ve seen in the IB Methods course.
Consider the simpler case where we only have two dimensions. The time-
independent Schr¨odinger equation then gives
Hψ =
~
2
2m
2
x
2
1
+
2
x
2
2
ψ + V (x
1
, x
2
)ψ = Eψ.
We are going to consider potentials of the form
V (x
1
, x
2
) = U
1
(x
1
) + U
2
(x
2
).
The Hamiltonian then splits into
H = H
1
+ H
2
,
where
H
i
=
~
2
2m
2
x
2
i
+ U
i
(x
i
).
We look for separable solutions of the form
ψ = χ
1
(x
1
)χ
2
(x
2
).
The Schr¨odinger equation then gives (upon division by ψ = χ
1
χ
2
) gives
~
2
2m
χ
00
1
χ
1
+ U
1
+
~
2
2m
χ
00
2
χ
2
+ U
2
= E.
Since each term is independent of x
2
and x
1
respectively, we have
H
1
χ
1
= E
1
χ
1
, H
2
χ
2
= E
2
χ
2
,
with
E
1
+ E
2
= E.
This is the usual separation of variables, but here we can interpret this physically
in this scenario, the two dimensions are de-coupled, and we can treat them
separately. The individual
E
1
and
E
2
are just the contributions from each
component to the energy. Thus the process of separation variables can be seen
as looking for joint or simultaneous eigenstates for
H
1
and
H
2
, noting the fact
that [H
1
, H
2
] = 0.
Example. Consider the harmonic oscillator with
U
i
(x
i
) =
1
2
ω
2
x
2
i
.
This corresponds to saying
V =
1
2
2
kxk
2
=
1
2
2
(x
2
1
+ x
2
2
).
Now we have
H
i
= H
0
(ˆx
i
, ˆp
i
),
with H
0
the usual harmonic oscillator Hamiltonian.
Using the previous results for the one-dimensional harmonic oscillator, we
have
χ
1
= ψ
n
1
(x
1
), χ
2
= ψ
n
2
(x
2
),
where
ψ
i
is the
i
th eigenstate of the one-dimensional harmonic oscillator, and
n
1
, n
2
= 0, 1, 2, ···.
The corresponding energies are
E
i
= ~ω
n
i
+
1
2
.
The energy eigenvalues of the two-dimensional oscillator are thus
E = ~ω (n
i
+ n
2
+ 1)
for
ψ(x
1
, x
2
) = ψ
n
1
(x
1
)ψ
n
2
(x
2
).
We have the following energies and states:
State Energy Possible states
Ground state E = ~ω ψ = ψ
0
(x
1
)ψ
0
(x
2
)
1st excited state E = 2~ω
ψ = ψ
1
(x
1
)ψ
0
(x
2
)
ψ = ψ
0
(x
1
)ψ
1
(x
2
)
We see there is a degeneracy of 2 for the first excited state.
Separable solutions also arise naturally when
H
(i.e. the potential
V
) has
some symmetry. For example, if the potential is spherically symmetric, we can
find solutions
ψ(x) = R(r)Y (θ, φ),
where r, θ, φ are the spherical polars.
6.3 Angular momentum
Recall that in IA Dynamics and Relativity, we also solved a spherically symmetric
system in classical dynamics. We obtained beautiful, exact solutions for the
orbits of particle in a gravitational potential in terms of conic sections.
To do so, we came up with a conserved quantity known as the angular
momentum. Using the fact that this is conserved, we were able to make a lot of
calculations and derive many results. We can also use this conserved quantity to
understand, say, why the Earth doesn’t just fall into the sun directly.
To understand spherically symmetric potentials in quantum mechanics, it is
also helpful to understand the angular momentum.
Definitions
Definition
(Angular momentum)
.
The angular momentum is a vector of oper-
ators
L =
ˆ
x
ˆ
p = i~x .
In components, this is given by
L
i
= ε
ijk
ˆx
j
ˆp
k
= i~ε
ijk
x
j
x
k
.
For example, we have
L
3
= ˆx
1
ˆp
2
ˆx
2
ˆp
1
= i~
x
1
x
2
x
2
x
1
.
Note that this is just the same definition as in classical dynamics, but with
everything promoted to operators.
These operators are Hermitian, i.e.
L
= L,
since
ˆx
i
and
ˆp
j
are themselves Hermitian, and noting the fact that
ˆx
i
and
ˆp
j
commute whenever i 6= j.
Each component of
L
is the angular momentum in one direction. We can
also consider the length of L, which is the total angular momentum.
Definition
(Total angular momentum)
.
The total angular momentum operator
is
L
2
= L
i
L
i
= L
2
1
+ L
2
2
+ L
3
3
.
Again, this is Hermitian, and hence an observable.
Commutation relations
We have the following important commutation relations for angular momentum:
[L
i
, L
j
] = i~ε
ijk
L
k
. (i)
For example,
[L
1
, L
2
] = i~L
3
.
Recall that in classical dynamics, an important result is that the angular mo-
mentum is conserved in all directions. However, we know we can’t do this in
quantum mechanics, since the angular momentum operators do not commute,
and we cannot measure all of them. This is why we have this
L
2
. It captures the
total angular momentum, and commutes with the angular momentum operators:
[L
2
, L
i
] = 0 (ii)
for all i.
Finally, we also have the following commutation relations.
[L
i
, ˆx
j
] = i~ε
ijk
ˆx
k
, [L
i
, ˆp
j
] = i~ε
ijk
ˆp
k
. (iii)
These are rather important results.
L
i
and
L
2
are observables, but (i) implies we
cannot simultaneously measure, say
L
1
and
L
2
. The best that can be done is to
measure
L
2
and
L
3
, say. This is possible by (ii). Finally, (iii) allows computation
of the commutator of L
i
with any function of ˆx
i
and ˆp
i
.
First, we want to prove these commutation relations.
Proposition.
(i) [L
i
, L
j
] = i~ε
ijk
L
k
.
(ii) [L
2
, L
i
] = 0.
(iii) [L
i
, ˆx
j
] = i~ε
ijk
ˆx
k
and [L
i
, ˆp
j
] = i~ε
ijk
ˆp
k
Proof.
(i)
We first do a direct approach, and look at specific indices instead of general
i and j. We have
L
1
L
2
= (i~)
2
x
2
x
3
x
3
x
2
x
3
x
1
x
1
x
3
= ~
2
x
2
x
3
x
3
x
1
x
1
x
2
2
x
2
3
x
2
3
2
x
2
x
1
+ x
3
x
1
2
x
2
x
3
Now note that
x
2
x
3
x
3
x
1
= x
2
x
3
2
x
3
x
1
+ x
2
x
1
.
Similarly, we can compute L
2
L
1
and find
[L
1
, L
2
] = ~
2
x
2
x
1
x
1
x
2
= i~L
3
,
where all the double derivatives cancel. So done.
Alternatively, we can do this in an abstract way. We have
L
i
L
j
= ε
iar
ˆx
a
ˆp
r
ε
jbs
ˆx
b
ˆp
s
= ε
iar
ε
jbs
(ˆx
a
ˆp
r
ˆx
b
ˆp
s
)
= ε
iar
ε
jbs
(ˆx
a
(ˆx
b
ˆp
r
[ˆp
r
, ˆx
b
])ˆp
s
)
= ε
iar
ε
jbs
(ˆx
a
ˆx
b
ˆp
r
ˆp
s
i~δ
br
ˆx
a
ˆp
s
)
Similarly, we have
L
j
L
i
= ε
iar
ε
jbs
(ˆx
b
ˆx
a
ˆp
s
ˆp
r
i~δ
as
ˆx
b
ˆp
r
)
Then the commutator is
L
i
L
j
L
j
L
i
= i~ε
iar
ε
jbs
(δ
br
ˆx
a
ˆp
s
δ
as
ˆx
b
ˆp
r
)
= i~(ε
iab
ε
jbs
ˆx
a
ˆp
s
ε
iar
ε
jba
ˆx
b
ˆp
r
)
= i~((δ
is
δ
ja
δ
ij
δ
as
)ˆx
a
ˆp
s
(δ
ib
δ
rj
δ
ij
δ
rb
)ˆx
b
ˆp
r
)
= i~(ˆx
i
ˆp
j
ˆx
j
ˆp
i
)
= i~ε
ijk
L
k
.
So done.
(ii) This follows from (i) using the Leibnitz property:
[A, BC] = [A, B]C + B[A, C].
This property can be proved by directly expanding both sides, and the
proof is uninteresting.
Using this, we get
[L
i
, L
2
] = [L
i
, L
j
L
j
]
= [L
i
, L
j
]L
j
+ L
j
[L
i
, L
j
]
= i~ε
ijk
(L
k
L
j
+ L
j
L
k
)
= 0
where we get 0 since we are contracting the antisymmetric tensor
ε
ijk
with
the symmetric tensor L
k
L
j
+ L
j
L
k
.
(iii)
We will use the Leibnitz property again, but this time we have it the other
way round:
[AB, C] = [A, C]B + A[B, C].
This follows immediately from the previous version since [
A, B
] =
[
B, A
].
Then we can compute
[L
i
, ˆx
j
] = ε
iab
[ˆx
a
ˆp
b
, ˆx
j
]
= ε
iab
([ˆx
a
, ˆx
j
]ˆp
b
+ ˆx
a
[ˆp
b
, ˆx
j
])
= ε
iab
ˆx
a
(i~δ
bj
)
= i~ε
ija
ˆx
a
as claimed.
We also have
[L
i
, ˆp
j
] = ε
iab
[ˆx
a
ˆp
b
, ˆp
j
]
= ε
iab
([ˆx
a
, ˆp
j
]ˆp
b
+ ˆx
a
[ˆp
b
, ˆp
j
])
= ε
iab
(i~δ
aj
ˆp
b
)
= i~ε
ijb
ˆp
b
.
Spherical polars and spherical harmonics
Recall that angular momentum is something coming from rotation. So let’s
work with something with spherical symmetry. We will exploit the symmetry
and work with spherical polar coordinates. To do so, we first need to convert
everything we had so far from Cartesian coordinates to polar coordinates.
We define our spherical polar coordinates (r, θ, ϕ) by the usual relations
x
1
= r sin θ cos ϕ
x
2
= r sin θ sin ϕ
x
3
= r cos θ.
If we express our angular momentum operators as differential operators, then we
can write them entirely in terms of
r, θ
and
ϕ
using the chain rule. The formula
for
L
3
will be rather simple, since
x
3
is our axis of rotation. However, those for
L
1
and
L
2
will be much more complicated. Instead of writing them out directly,
we instead write down the formula for
L
±
=
L
1
± iL
2
. A routine application of
the chain rule gives
L
3
= i~
ϕ
L
±
= L
1
± iL
2
= ±~e
±
θ
± i cot θ
ϕ
L
2
= ~
2
1
sin θ
θ
sin θ
θ
+
1
sin
2
θ
2
ϕ
2
.
Note these operators involve only
θ
and
ϕ
. Furthermore, the expression for
L
2
is something we have all seen before we have
2
=
1
r
2
r
2
r
1
~
2
r
2
L
2
.
Since we know
[L
3
, L
2
] = 0,
there are simultaneous eigenfunctions of these operators, which we shall call
Y
`m
(
θ, ϕ
), with
`
= 0
,
1
,
2
, ···
and
m
= 0
, ±
1
, ±
2
, ··· , ±`
. These have eigenval-
ues ~m for L
3
and ~
2
`(` + 1) for L
2
.
In general, we have
Y
`m
= const e
imϕ
P
m
`
(cos θ),
where
P
m
`
is the associated Legendre function. For the simplest case
m
= 0, we
have
Y
`0
= const P
`
(cos θ),
where P
`
is the Legendre polynomial.
Note that these details are not important. The important thing to take away
is that there are solutions
Y
`m
(
θ, ϕ
), which we can happily use and be confident
that someone out there understands these well.
6.4 Joint eigenstates for a spherically symmetric potential
Unfortunately, it is completely standard to use
m
to indicate the eigenvalue of
L
3
, as we did above. Hence, here we shall consider a particle of mass
µ
in a
potential V (r) with spherical symmetry. The Hamiltonian is
H =
1
2µ
ˆp
2
+ V =
~
2
2µ
2
+ V (r).
We have seen that we can write H as
H =
~
2
2µ
1
r
2
r
2
r +
1
2µ
1
r
2
L
2
+ V (r).
The first thing we want to check is that
[L
i
, H] = [L
2
, H] = 0.
This implies we can use the eigenvalues of
H
,
L
2
and
L
3
to label our solutions
to the equation.
We check this using Cartesian coordinates. The kinetic term is
[L
i
,
ˆ
p
2
] = [L
i
, ˆp
j
ˆp
j
]
= [L
i
, ˆp
j
]ˆp
j
+ ˆp
j
[L
i
, ˆp
j
]
= i~ε
ijk
(ˆp
k
ˆp
j
+ ˆp
j
ˆp
k
)
= 0
since we are contracting an antisymmetric tensor with a symmetric term. We
can also compute the commutator with the potential term
[L
i
, V (r)] = i~ε
ijk
x
j
V
x
k
= i~ε
ijk
x
j
x
k
r
V
0
(r)
= 0,
using the fact that
r
x
i
=
x
i
r
.
Now that
H
,
L
2
and
L
3
are a commuting set of observables, we have the joint
eigenstates
ψ(x) = R(r)Y
`m
(θ, ϕ),
and we have
L
2
Y
`m
= ~
2
`(` + 1)Y
`m
` = 0, 1, 2, ···
L
3
Y
`m
= ~mY
`m
m = 0, ±1, ±2, ··· , ±`.
The numbers
`
and
m
are usually known as the angular momentum quantum
numbers. Note that
`
= 0 is the special case where we have a spherically
symmetric solution.
Finally, we solve the Schr¨odinger equation
Hψ = Eψ
to obtain
~
2
2µ
1
r
d
2
dr
2
(rR) +
~
2
2µr
2
`(` + 1)R + V R = ER.
This is now an ordinary differential equation in
R
. We can interpret the terms
as the radial kinetic energy, angular kinetic energy, the potential energy and
the energy eigenvalue respectively. Note that similar to what we did in classical
dynamics, under a spherically symmetric potential, we can replace the angular
part of the motion with an “effective potential”
~
2
2µr
2
`(` + 1).
We often call
R
(
r
) the radial part of the wavefunction, defined on
r
0.
Often, it is convenient to work with
χ
(
r
) =
rR
(
r
), which is sometimes called
the radial wavefunction. Multiplying the original Schr¨odinger equation by
r
, we
obtain
~
2
2µ
χ
00
+
~
2
`(` + 1)
2µr
2
χ + V χ = Eχ.
This is known as the radial Schr¨odinger equation.
This has to obey some boundary conditions. Since we want
R
to be finite as
r
0, we must have
χ
= 0 at
r
= 0. Moreover, the normalization condition is
now
1 =
Z
|ψ(x)|
2
d
3
x =
Z
|R(r)|
2
r
2
dr
Z
|Y
`m
(θ, ϕ)|
2
sin θ dθ dϕ.
Hence, ψ is normalizable if and only if
Z
0
|R(r)|
2
r
2
dr < .
Alternatively, this requires
Z
0
|χ(r)|
2
dr < .
Example
(Three-dimensional well)
.
We now plot our potential as a function of
r:
r
V
U
a
This is described by
V (r) =
(
0 r a
U r < a
,
where U > 0 is a constant.
We now look for bound state solutions to the Schr¨odinger equation with
U < E < 0, with total angular momentum quantum number `.
For r < a, our radial wavefunction χ obeys
χ
00
`(` + 1)
r
2
χ + k
2
χ = 0,
where k is a new constant obeying
U + E =
~
2
k
2
2µ
.
For r a, we have
χ
00
`(` + 1)
r
2
χ κ
2
χ = 0,
with κ obeying
E =
~
2
κ
2
2µ
.
We can solve in each region and match
χ, χ
0
at
r
=
a
, with boundary condition
χ
(0) = 0. Note that given this boundary condition, solving this is equivalent to
solving it for the whole R but requiring the solution to be odd.
Solving this for general
`
is slightly complicated. So we shall look at some
particular examples.
For
`
= 0, we have no angular term, and we have done this before. The
general solution is
χ(r) =
(
A sin kr r < a
Be
κr
r > a
Matching the values at x = a determines the values of k, κ and hence E.
For ` = 1, it turns out the solution is just
χ(r) =
(
A
cos kr
1
kr
sin kr
r < a
B
1 +
1
kr
e
κr
r > a
.
After matching, the solution is
ψ(r) = R(r)Y
1m
(θ, ϕ) =
χ(r)
r
Y
1m
(θ, ϕ),
where m can take values m = 0, ±1.
Solution for general
`
involves spherical Bessel functions, and are studied
more in-depth in the IID Applications of Quantum Mechanics course.
7 The hydrogen atom
Recall that at the beginning of the course, we said one of the main motivations
for quantum mechanics was to study how atoms worked and where the spectrum
of the hydrogen atom came from. This is just an application of what we have
done for general spherically symmetric potentials, but this is so important that
we give it a separate chapter.
7.1 Introduction
Consider an electron moving in a Coulomb potential
V (r) =
e
2
4πε
0
1
r
.
This potential is due to a proton stationary at
r
= 0. We follow results from
the last section of the last chapter, and set the mass
µ
=
m
e
, the electron mass.
The joint energy eigenstates of H, L
2
and L
3
are of the form
φ(x) = R(r)Y
`m
(θ, ϕ)
for ` = 0, 1, ··· and m = 0, ±1, ··· , ±m.
The radial part of the Schr¨odinger equation can be written
R
00
+
2
r
R
0
`(` + 1)
r
2
R +
2λ
r
R = κ
2
R, ()
with
λ =
m
e
e
2
4πε
0
~
2
, E =
~
2
κ
2
2m
e
.
Note that here we work directly with
R
instead of
χ
, as this turns out to be
easier later on.
The goal of this chapter is to understand all the (normalizable) solutions to
this equation ().
As in the case of the harmonic oscillator, the trick to solve this is see what
happens for large
r
, and “guess” a common factor of the solutions. In the case of
the harmonic oscillator, we guessed the solution should have
e
y
2
/2
as a factor.
Here, for large r, we get
R
00
κ
2
R.
This implies R e
κr
for large R.
For small
r
, we by assumption know that
R
is finite, while
R
0
and
R
00
could
potentially go crazy. So we multiply by
r
2
and discard the
rR
and
r
2
R
terms to
get
r
2
R
00
+ 2rR
0
`(` + 1)R 0.
This gives the solution R r
`
.
We’ll be bold and try a solution of the form
R(r) = Cr
`
e
κr
.
When we substitute this in, we will get three kinds of terms. The
r
`
e
κr
terms
match, and so do the terms of the form r
`2
e
κr
. Finally, we see the r
`1
e
κe
terms match if and only if
2(`r
`1
)(κe
κr
) + 2(r
`1
)(κe
κr
) + 2λr
`1
e
κe
= 0.
When we simplify this mess, we see this holds if and only if
(` + 1)κ = λ.
Hence, for any integer
n
=
`
+ 1 = 1
,
2
,
3
, ···
, there are bound states with
energies
E
n
=
~
2
2m
e
λ
2
n
2
=
1
2
m
e
e
2
4πε
0
~
2
1
n
2
.
These are the energy levels of the Bohr model, derived within the framework of
the Bohr model. However, there is a slight difference. In our model, the total
angular momentum eigenvalue is
~
2
`(` + 1) = ~
2
n(n 1),
which is not what the Bohr model predicted.
Nevertheless, this is not the full solution. For each energy level, this only
gives one possible angular momentum, but we know that there can be many
possible angular momentums for each energy level. So there is more work to be
done.
7.2 General solution
We guessed our solution
r
`
e
κr
above by looking at the asymptotic behaviour
at large and small
r
. We then managed to show that this is one solution of the
hydrogen atom. We are greedy, and we want all solutions.
Similar to what we did for the harmonic oscillator, we guess that our general
solution is of the form
R(r) = e
κr
f(r).
Putting it in, we obtain
f
00
+
2
r
f
0
`(` + 1)
r
2
f = 2
κf
0
+ (κ λ)
f
r
.
We immediately see an advantage of this substitution now each side of the
equality is equidimensional, and equidimensionality makes our life much easier
when seeking series solution. This equation is regular singular at
r
= 0, and
hence we guess a solution of the form
f(r) =
X
p=0
a
p
r
p+σ
, a
0
6= 0.
Then substitution gives
X
p0
((p + σ)(p + σ 1) `(` + 1))a
p
r
p+σ2
=
X
p0
2(κ(p + σ + 1) λ)a
p
r
p+σ1
.
The lowest term gives us the indicial equation
σ(σ + 1) `(` + 1) = (σ `)(σ + ` + 1) = 0.
So either
σ
=
`
or
σ
=
(
`
+ 1). We discard the
σ
=
(
`
+ 1) solution since this
would make f and hence R singular at r = 0. So we have σ = `.
Given this, the coefficients are then determined by
a
p
=
2(κ(p + `) λ)
p(p + 2` + 1)
a
p1
, p 1.
Similar to the harmonic oscillator, we now observe that, unless the series termi-
nates, we have
a
p
a
p1
2κ
p
as
p
, which matches the behaviour of
r
α
e
2κr
(for some
α
). So
R
(
r
) is
normalizable only if the series terminates. Hence the possible values of λ are
κn = λ
for some
n `
+ 1. So the resulting energy levels are exactly those we found
before:
E
n
=
~
2
2m
e
κ
2
=
~
2
2m
e
λ
2
n
2
=
1
2
m
e
e
2
4πε
0
~
2
1
n
2
.
for n = 1, 2, 3, ···. This n is called the principle quantum number.
For any given n, the possible angular momentum quantum numbers are
` = 0, 1, 2, 3, ··· , n 1
m = 0, ±1, ±2, ··· , ±`.
The simultaneous eigenstates are then
ψ
n`m
(x) = R
n`
(r)Y
`m
(θ, ϕ),
with
R
n`
(r) = r
`
g
n`
(r)e
λr/n
,
where g
n`
(r) are (proportional to) the associated Laguerre polynomials.
In general, the “shape” of probability distribution for any electron state
depends on
r
and
θ, ϕ
mostly through
Y
`m
. For
`
= 0, we have a spherically
symmetric solutions
ψ
n00
(x) = g
n0
(r)e
λr/n
.
This is very different from the Bohr model, which says the energy levels depend
only on the angular momentum and nothing else. Here we can have many
different angular momentums for each energy level, and can even have no angular
momentum at all.
The degeneracy of each energy level E
n
is
n1
X
`=0
`
X
m=`
1 =
n1
X
`=0
(2` + 1) = n
2
.
If you study quantum mechanics more in-depth, you will find that the degeneracy
of energy eigenstates reflects the symmetries in the Coulomb potential. Moreover,
the fact that we have
n
2
degenerate states implies that there is a hidden symmetry,
in addition to the obvious
SO
(3) rotational symmetry, since just
SO
(3) itself
should give rise to much fewer degenerate states.
So. We have solved the hydrogen atom.
7.3 Comments
Is this it? Not really. When we solved the hydrogen atom, we made a lot of
simplifying assumptions. It is worth revisiting these assumptions and see if they
are actually significant.
Assumptions in the treatment of the hydrogen atom
One thing we assumed was that the proton is stationary at the origin and the
electron moves around it. We also took the mass to be
µ
=
m
e
. More accurately,
we can consider the motion relative to the center of mass of the system, and we
should take the mass as the reduced mass
µ =
m
e
m
p
m
e
+ m
p
,
just as in classical mechanics. However, the proton mass is so much larger and
heavier, and the reduced mass is very close to the electron mass. Hence, what
we’ve got is actually a good approximation. In principle, we can take this into
account and this will change the energy levels very slightly.
What else? The entire treatment of quantum mechanics is non-relativistic.
We can work a bit harder and solve the hydrogen atom relativistically, but the
corrections are also small. These are rather moot problems. There are larger
problems.
Spin
We have always assumed that particles are structureless, namely that we can
completely specify the properties of a particle by its position and momentum.
However, it turns out electrons (and protons and neutrons) have an additional
internal degree of freedom called spin. This is a form of angular momentum, but
with
`
=
1
2
and
m
=
±
1
2
. This cannot be due to orbital motion, since orbital
motion has integer values of
`
for well-behaved wavefunctions. However, we still
call it angular momentum, since angular momentum is conserved only if we take
these into account as well.
The result is that for each each quantum number
n, `, m
, there are two
possible spin states, and the total degeneracy of level
E
n
is then 2
n
2
. This agrees
with what we know from chemistry.
Many electron atoms
So far, we have been looking at a hydrogen atom, with just one proton and one
electron. What if we had more electrons? Consider a nucleus at the origin with
charge +
Ze
, where
Z
is the atomic number. This has
Z
independent electrons
orbiting it with positions x
a
for a = 1, ··· , Z.
We can write down the Schr¨odinger equation for these particles, and it looks
rather complicated, since electrons not only interact with the nucleus, but with
other electrons as well.
So, to begin, we first ignore electron-electron interactions. Then the solutions
can be written down immediately:
ψ(x
1
, x
2
, ··· , x
Z
) = ψ
1
(x
1
)ψ
2
(x
2
) ···ψ
Z
(x
Z
),
where
ψ
i
is any solution for the hydrogen atom, scaled appropriately by
e
2
7→ Ze
2
to accommodate for the larger charge of the nucleus. The energy is then
E = E
1
+ E
2
+ ··· + E
Z
.
We can next add in the electron-electron interactions terms, and find a more
accurate equation for
ψ
using perturbation theory, which you will come across
in IID Principles of Quantum Mechanics.
However, there is an additional constraint on this. The Fermi-Dirac statistics
or Pauli exclusion principle states that no two particles can have the same state.
In other words, if we attempt to construct a multi-electron atom, we cannot
put everything into the ground state. We are forced to put some electrons in
higher energy states. This is how chemical reactivity arises, which depends on
occupancy of energy levels.
For
n
= 1, we have 2
n
2
= 2 electron states. This is full for
Z
= 2, and this
is helium.
For
n
= 2, we have 2
n
2
= 8 electron states. Hence the first two energy
levels are full for Z = 10, and this is neon.
These are rather stable elements, since to give them an additional electron, we
must put it in a higher energy level, which costs a lot of energy.
We also expect reactive atoms when the number of electrons is one more
or less than the full energy levels. These include hydrogen (
Z
= 1), lithium
(Z = 3), fluorine (Z = 9) and sodium (Z = 11).
This is a recognizable sketch of the periodic table. However, for
n
= 3 and
above, this model does not hold well. At these energy levels, electron-electron
interactions become important, and the world is not so simple.