Notes 311219
Notes 311219
Notes 311219
Tuesday and Thursday, 11.05 a.m. to 11.55 a.m., Mill Lane lecture room 3
Prof. Adrian Kent (apak@damtp.cam.ac.uk)
Version dated 31.12.19. Any updated versions will be placed on the course web
page linked from www.qi.damtp.cam.ac.uk.
1
Recommended books
(*) A. Rae and J. Napolitano, Quantum Mechanics, chapters 1-5 (IOP Publish-
ing, 2002).
A very good textbook which covers a range of elementary and more advanced
topics in quantum theory, including a short discussion of the conceptual problems
of quantum mechanics. The first five chapters form a good course text for IB QM.
S. Brandt and H.D. Dahmen, The Picture Book of Quantum Mechanics (4th
edition; Springer, 2012). An excellent book with accompanying plots and simu-
lations. These give visual explanations that nicely illuminate the mathematics of
solutions of the Schrödinger equation, tunnelling, reflection from barriers, atomic
electron states, and other quantum phenomena.
2
Contents
1 Quantum Mechanics, science and technology 5
1.1 Quantum Mechanics and fundamental science . . . . . . . . . . . . . 5
1.2 Quantum Mechanics and technology . . . . . . . . . . . . . . . . . . 6
3
6.3.1 Some examples of operators . . . . . . . . . . . . . . . . . . . 53
6.4 Hermitian operators . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
6.4.1 Classical states and dynamical variables . . . . . . . . . . . . 53
6.4.2 Quantum states and observables . . . . . . . . . . . . . . . . 54
6.5 Some theorems about hermitian operators . . . . . . . . . . . . . . . 55
6.6 Quantum measurement postulates . . . . . . . . . . . . . . . . . . . 57
6.7 Expectation values . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
6.8 Commutation relations . . . . . . . . . . . . . . . . . . . . . . . . . . 59
6.8.1 The canonical commutation relations . . . . . . . . . . . . . . 60
6.9 Heisenberg’s Uncertainty Principle . . . . . . . . . . . . . . . . . . . 61
6.9.1 What does the uncertainty principle tell us? . . . . . . . . . . 63
6.10 Ehrenfest’s theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
6.10.1 Applications of Ehrenfest’s theorem . . . . . . . . . . . . . . 64
6.11 *The harmonic oscillator revisited . . . . . . . . . . . . . . . . . . . 66
4
1 Quantum Mechanics, science and technology
1.1 Quantum Mechanics and fundamental science
• the structure of atoms and molecules and their chemical interactions; i.e.
chemistry, and biochemistry and so, in principle, biology. We will begin to
discuss this towards the end of the course, when we consider quantum mechan-
ical descriptions of the hydrogen atom and, more qualitatively, other atoms.
This description is made more precise and taken further in Part II Principles
of Quantum Mechanics.
• the structure and properties of solids (and so, in principle, much of classical
mechanics). Conductivity (some basic theory of which is introduced in Part
II Applications of Quantum Mechanics) and superconductivity.
• the thermodynamics of light and other electromagnetic radiation and also how
electromagnetic radiation interacts with matter. To describe this properly
requires relativistic quantum field theory, which isn’t covered until Part III.
• the physics of subatomic particles, radioactivity, nuclear fission and fusion.
Again, we need relativistic quantum field theory to describe these phenomena
in full detail. But, as we’ll see later, even elementary quantum mechanics
gives us some useful insights into the physics of nuclear fission and fusion. For
example, we can understand the random nature of these processes, and the
way fusion and fission rates depend on relevant potentials, as a consequence
of quantum tunnelling.
Modern cosmological models are also based on quantum theory. Since we do not
have a quantum theory of gravity, and do not know for sure whether there is one,
these cosmological models are at best incomplete. Nonetheless they give a good
qualitative fit to observational data. Many physicists hope that this project can be
completed, so that we can describe the entire universe and its evolution by quantum
theory.2 These topics are covered in detail in the Part II and Part III Cosmology
courses.
1 Einstein’s general theory of relativity is the only other contender. It is an extraordinarily
beautiful theory that transformed our understanding of space and time and their relationship to
matter, and that gave us the tools to understand the cosmos. The two theories are fundamentally
incompatible, and it is uncertain precisely which parts of which theory will survive in a future uni-
fication. Still, I find quantum theory more intriguing, because of the variety of deep mathematical
concepts it combines and because it is so different from and so much stranger than the theories
that preceded it.
2 It is also true that many thoughtful physicists suspect that the project cannot be completed,
because we will need something other than a quantum theory of gravity, or because quantum
theory will turn out not to hold on large scales. If so, comparing quantum cosmological models to
observation should eventually give us insight into new physics.
5
1.2 Quantum Mechanics and technology
Many of the revolutionary technological developments of the last hundred years rely
on quantum mechanics:
3 There are nuances of definition here, and the community is still assessing the claim.
6
Figure 1: Google’s Bristlecone quantum processor. Image on left, map of qubit
connections on right.
E = hν = ~ω . (2.1)
7
2.2 Einstein’s explanation of the photoelectric effect (1905)
Experiment shows that light hitting a metal surface in a vacuum can cause
electrons to be ejected with a range of energies. To emit any electrons, the incident
light needs to have angular frequency ω satisfying ω ≥ ωmin , where ωmin is a constant
depending on the particular metal. When ω ≥ ωmin , one finds that the maximum
energy of the emitted electrons, Emax , obeys
𝑒−
𝑒−
𝑒−
𝑒− 𝑒− 𝑒− 𝑒− 𝑒− 𝑒−
8
Thus, one can explain the photoelectric effect as the result of single photons hitting
electrons near the metal surface, if one assumes that an electron needs to acquire
kinetic energy ≥ φ to overcome the binding energy of the metal. An electron which
acquires energy ~ω thus carries away energy ≤ ~ω −φ = Emax . One can also explain
the emission rate observations: the average rate of photon arrival is proportional
to the intensity of the light, and the rate of emission of electrons is proportional
to the rate of photon arrival. However, individual photons arrive, and hit electrons
so as to knock them out of the metal, at random – hence the randomly distributed
emissions of electrons.
This seems to suggest that single photons propagate through the apparatus and
nonetheless “self-interfere” in such a way that the diffraction pattern is cumulatively
built up. This is indeed how Taylor’s results were interpreted for several decades
after the development of quantum mechanics in 1926. Problems were later noticed
9
with this interpretation of this particular experiment: having the average energy flux
of a single photon does not always imply that a source is emitting single photons.
However, the experiment was subsequently repeated with genuine single photon
sources, still showing the same diffraction pattern. Qualitatively similar patterns
are seen in diffraction experiments for single electrons and other types of matter
(see below).
Third, it fails to explain why atoms belong to a finite number of chemical species,
with all members of the same species behaving identically. For instance, if a hydro-
gen atom can have an electron in any type of orbit around its nucleus, one would
expect infinitely many different types of hydrogen atom, corresponding to the in-
finitely many different possible orbits, and one would expect the atoms to have
different physical and chemical properties, depending on the details of the orbit.
10
Figure 4: Schematic illustration of the Geiger-Marsden experiment, from the web
page cited above.
Niels Bohr, in 1913, observed that these problems could be resolved in a way
consistent with Planck’s and Einstein’s earlier postulates, if we suppose that the
electron orbits of hydrogen atoms are quantised so that the orbital angular momen-
tum takes one of a discrete set of values
L = n~ , (2.6)
e2 me v 2
= . (2.7)
4π0 r2 r
If we also have
L = me vr = n~ (2.8)
then
n2 ~2 e2
= (2.9)
r3 me 4π0 r2
and hence
~2 4π0
2
r=n ( 2
) = n 2 a0 , (2.10)
me e
11
Figure 5: Spectra for the hydrogen atom. The figure shows three horizontal lines
at small distances from each other. Between the two lower lines, the Lyman series,
with four vertical red bands in compact form, is shown. This has nf = 1 and ni ≥ 2,
and wavelengths in the range 91 − 100 nanometers. The Balmer series is shown to
the right side of this series. This has nf = 2 and ni ≥ 3, and wavelengths in the
range 365 − 656 nanometers. At the right side of this, the Paschen series bands
are shown. This has nf = 3 and ni ≥ 4, and wavelengths in the range 820 − 1875
nanometers. The Rydberg formula is obtained by taking nf = n, ni = m.
Source for figure and caption: https://opentextbc.ca/physicstestbook2
Image licenced under Creative Commons.
where
4π0 ~2
a0 = ≈ 0.53 × 10−10 m (2.11)
m e e2
is the Bohr radius.
We can then obtain the energy of the n-th Bohr orbit from (2.7) and the Coulomb
potential:
1 e2 e2 e2 e4 me 1 E1
En = me v 2 − =− =− = − = ,
2 4π0 r 8π0 r 8π0 n2 a0 32π 2 20 ~2 n2 n2
(2.12)
where
e4 m e
E1 = − ≈ −13.6 eV . (2.13)
32π 2 20 ~2
Thus we have n = 1 with energy E = E1 defining the lowest possible energy
state, or ground state, of the Bohr atom. The higher energy excited states, so called
because they can be created by “exciting” the ground state atom with radiation,
correspond to n > 1. These can decay to the ground state: the ground state has no
lower energy state to decay to, and so is stable. (The Bohr model does not allow a
state with zero orbital angular momentum, which would correspond to n = 0, r = 0
and E = −∞.)
12
The energy emitted for a transition from the m-th to the n-th Bohr orbital is
Emn = Em − En . Using Emn = ~ωmn , where ωmn is the angular frequency of the
emitted photon, we have
1 1
ωmn = 2πR0 c( 2
− 2) , (2.14)
n m
where
me c e2
R0 = ( )2 , (2.15)
2~ 4π0 ~c
which agrees well with the experimentally determined value of the Rydberg con-
stant.
Figure 6: The orbits of Bohr’s planetary model of an atom; five concentric circles
are shown. The radii of the circles increase from innermost to outermost circles.
On the circles, labels E1 , E2 , up to Ei are marked. Source for figure and caption:
https://opentextbc.ca/physicstestbook2
Licenced under Creative Commons.
Bohr’s model of the atom was rather more successful than its predecessors. It
predicts the energy levels of the hydrogen atom and the spectrum of photons emitted
and absorbed. It also accounted for spectroscopic data for ionised helium (He+ ) and
for some emission and absorption spectra for other atoms. (We now understand that
these are the spectra produced by the innermost electrons, which can be modelled
in a way qualitatively similar to the electron in the hydrogen atom.)
However, as Bohr himself stressed, the model offered no explanation of atomic
physics. For example, as Rutherford commented, it’s quite mysterious that an
electron which jumps from the m-th to the n-th orbit seems to know in advance what
frequency to radiate at during the transition. Moreover, the Bohr model is quite
13
wrong about the details of electron orbits, even for the hydrogen atom. Nonetheless,
it was an important stepping stone on the path to quantum mechanics, suggesting
some link between Planck’s constant, atomic spectra and atomic structure, and the
quantisation of angular momentum and other dynamical quantities.
p = ~k , (2.16)
where k is the wave vector of the corresponding electromagnetic wave, so that
~ω
|p| = = ~|k| . (2.17)
c
14
2.7 Wave and particle models of electromagnetic radiation
We thus see the emergence of two different models of light and other electro-
magnetic radiation.
Sometimes (classical electromagnetism, diffraction experiments with a strong
light source, . . .) it is useful to model light in terms of waves:
ei(k.x−ωt) , (2.18)
ω
where k is the wave vector, ω the angular frequency, c = the speed of light in a
|k|
2πc 2π
vacuum, and the wavelength λ = ω = |k| .
Sometimes (photoelectric effect, spectroscopy in the Bohr model of the hydrogen
atom, Compton scattering, . . .) it is useful to adopt a particle model, in which light
is made up of photons with energy and momentum
E = ~ω , p = ~k . (2.19)
The word “model” is chosen deliberately here. A model can be useful (as the
wave and particle models of light are, in the appropriate contexts) without be-
ing completely correct. Indeed, G.I. Taylor’s 1909 demonstration of single photon
diffraction already gave an example of a single experiment for which neither the
wave nor the particle model of light appeared to be adequate. A simple particle
model would not predict the observed diffraction pattern, while a simple wave model
cannot explain the observation of single photons recorded on the photographic film.4
2πn~
2πr = = nλ . (2.22)
p
De Broglie hypothesised that the frequency and wavevectors of the relevant wave
were related to the energy and momentum of the particle as for photons: that is, we
have E = ~ω, p = ~k even for particles with nonzero mass. (In fact, he speculated –
incorrectly, according to our current understanding – that photons might also have
a small nonzero mass.)
Einstein wrote, on learning of de Broglie’s hypothesis, that
4 See earlier comment: to justify this fully we need to consider versions of the experiment with
15
De Broglie’s model of the atom. Electrons occupy orbits
corresponding to a integer multiple of their de Broglie wavelength.
“I believe it is the first feeble ray of light on this worst of our physics
enigmas.”
It was.
16
Figure 8: Single electron detection events building up an interference pattern in
Tonomura’s experiment. Copyright in these documents published on Hitachi World-
Wide Web Server is owned by Hitachi, Ltd.
17
Quantum weirdness: the double slit experiment
There seems, in any case, a simple way to investigate more closely. Suppose
that we have ideal detectors, which register that an electron has passed through a
region, but don’t obstruct its path. We can set up a double slit experiment with
one of these ideal detectors adjacent to each slit (for definiteness, let us say they are
on the far side, between the slit and the screen). Now we don’t need to speculate:
we can observe which slit the electron goes through. But when we do this, the
interference pattern changes: we see a superposition of two Gaussians, as predicted
by a particle model, rather than the pattern of maxima and minima predicted by
the wave model and observed in the previous experiment.
18
What is really going on in the double slit
experiment?
Why not just look and see which slit the photon goes through?
What if we include passive
detectors registering which
slit the photon goes through?
Figure 10: Observing which slit the electron goes through in the double slit exper-
iment.
1. As we already stressed, the wave model and the particle model are just that —
models. Neither of them is adequate to explain the behaviour of electrons, photons,
or other objects. Each of them can sometimes give a partial explanation of our
observations, but that explanation is not consistent with all the data.
2. In particular, the type of reasoning about electron paths that would apply
within a particle model does not generally apply in Nature. We can’t assume that
the electron follows a definite path through one slit or the other, and we can’t
assume (as we could with a classical particle) that observing which slit it goes
through makes no difference to the interference pattern.
3. Some textbooks summarize the state of affairs described in point 1 by saying
that electrons (photons, etc.) exhibit something called “wave-particle duality”.
This term can mislead, if it is interpreted as a sort of explanation of what is going
on rather than just a catchy name for it.
The fact is that our classical wave and particle models are fundamentally inad-
equate descriptions. It isn’t correct to say that an electron (or photon, etc.) is both
a wave and a particle in the classical senses of either of those words. The electron
is something different again, though it has some features in common with both. To
go further, we need a new physical model: quantum mechanics.
4. We saw that the electron interference pattern builds up over time, yet the
points at which individual electrons hit the screen do not appear to be precisely
predictable: they seem to arrive at random. It might seem natural to speculate
that this apparent randomness might be explained by the fact that we don’t have
complete data about the experiment. Perhaps, for example, the electrons leave the
19
source in slightly different directions, or perhaps they have some sort of internal
structure that we haven’t yet discussed (and perhaps hasn’t yet been discovered),
and perhaps this determines precisely where they hit the screen.
As we’ll see, according to quantum mechanics this is not the case. Quantum
mechanics, unlike classical mechanics, is an intrinsically probabilistic theory, and it
tells us that there is simply no way to predict precisely where the electron will hit
the screen, even if we have a precise and complete description of its physical state
when it leaves the source.
Now, of course, quantum mechanics might not be the final theory of nature.
It’s possible that some as yet unknown and more complete theory could underlie
quantum mechanics, and it’s logically possible that this theory (if there is one)
could be deterministic. However, there are very strong reasons to doubt that any
theory underlying quantum mechanics can be deterministic. In particular, it can be
proved (given a few very natural assumptions) that any such theory would be incon-
sistent with special relativity. This follows from Bell’s theorem and experimental
tests thereof: it is discussed further in the Part II course “Principles of Quantum
Mechanics” and in Part III courses.)
20
2.11 *Ongoing tests of quantum theory
Although later developments are beyond the scope of this course, it would be
wrong to leave the impression that the historical development of quantum mechan-
ics ended in the first part of last century. Indeed, the basic principles of quantum
mechanics are still being questioned and tested by some theorists and experimen-
talists today. Quantum theory is very well confirmed as a theory of microscopic
physics. However, the case for believing that it applies universally to physics on all
scales is much weaker.
There is a good scientific motivation for testing any scientific theory in new
domains, which is that theories developed to explain phenomena in one domain may
not necessarily apply in other domains. For example, quantum theory itself shows
us the limits of validity of Newtonian mechanics and of classical electrodynamics.
Similarly, Einstein’s general theory of relativity shows us the limits of validity of
Newton’s theory of gravity.
There is also a specific motivation for wanting to test quantum theory for macro-
scopic systems. This is that the problems in making sense of quantum theory as
an explanation of natural phenomena seem to stem from the fact that the classical
physics we see on macroscopic scales appears to emerge from quantum theory in a
way that, despite many attempts, remains hard to pin down. Many theorists be-
lieve it remains fundamentally unexplained. Many others believe it is explained or
explainable, but there is no real consensus among them about the right explanation.
Interestingly, we know there are consistent (non-relativistic) theories that agree
very precisely with quantum mechanical predictions for microscopic (small mass)
particles, but disagree for macroscopic (large mass) ones.
In the past few years, experimental technology has advanced far enough to
demonstrate diffraction of quite complex molecules. (Some descriptions of these
experiments can be found at
https://vcq.quantum.at/; see in particular the work of the Arndt and Aspelmayer
research groups.) Attempts continue to demonstrate interferometry for larger and
larger objects, motivated not only by the technological challenge but also by the
possibility of testing the validity of quantum mechanics in new domains. In Octo-
ber 2019, Fein et al. reported interference experiments for molecules of weight
larger than 25kDa (See https://www.nature.com/articles/s41567-019-0663-9 and
the rather amusing Q and A summary
https://www.quantumnano.at/detailview-news/news/facts-fiction-in-reports-on-high-
mass-interference/ ).
It should be stressed that there is to date (October 2019) no experimental evi-
dence for any deviation from quantum mechanics, which has been confirmed in an
impressive array of experiments investigating many different physical regimes. *
21
Figure 11: Interference of molecules larger than 25KDa, from the Fein et al. paper
cited above.
Part III theoretical physics courses, along with other relevant quantum field theories.
2. Although we’ve already seen the classical particle model is inadequate, we
still need some collective name for electrons, protons, neutrons and so on. Perhaps
physicists should have invented another term, but we still call these “particles”. We
will follow this tradition, so that we might say that the electron is an elementary
particle, talk about quantum mechanics applied to an abstract particle of mass m,
and so forth — always keeping in mind that the classical particle model does not
actually apply.
22
3 The one-dimensional Schrödinger equation
3.1 The 1D Schrödinger equation for a free particle
We are first going to develop quantum mechanics in one space (and one time)
dimension. We can solve the equations for simple physical models more easily in
1D than in 3D and, happily, it turns out that 1D solutions give a good qualitative
feel for a range of interesting 3D physical phenomena.
In 1924, Schrödinger developed de Broglie’s ideas further, into what became a
standard way of framing the laws of quantum mechanics.5
Recall that de Broglie postulated that matter is described by waves, and that
the energy and momentum are related to the angular frequency and wave vector by
E = ~ω and p = ~k, or in one dimension p = ~k. We can express this by associating
to a particle of energy E and momentum p a wave of the form
i p2
ψ(x, t) = exp(i(kx − ωt)) = exp( (px − t)) . (3.2)
~ 2m
Notice that we have taken ψ(x, t) to be complex. Using complex numbers to
represent waves is familiar in classical electromagnetism, where it allows us to com-
bine the electric and magnetic fields in a single equation. In that context, it’s
mathematically convenient, but the real and imaginary parts each have a simple
physical interpretation. We will see that complex-valued solutions to (generalised)
wave equations play a more essential role in quantum mechanics.
The simplest wave equation to which the de Broglie wave is the general solution
is the time-dependent 1D Schrödinger equation for a free particle:
1 ∂ ∂
(−i~ )2 ψ(x, t) = i~ ψ(x, t) . (3.3)
2m ∂x ∂t
(By a free particle we mean a particle not subject to external forces, i.e. one moving
in a potential V (x) = 0.)
5 There is an equivalent alternative way of describing quantum mechanics, developed by Heisen-
berg, Born and Jordan. But Schrödinger’s formulation is easier to work with and gives a more
intuitive physical picture – so we will follow his approach. Note that “more intuitive” here is a
relative statement. As we will see, many of the predictions of quantum mechanics are counter-
intuitive. Also, some of the intuitions Schrödinger’s picture suggests may be helpful to us in some
contexts but are not fundamentally justifiable.
23
3.2 The momentum operator
We define the momentum operator
∂
p̂ = −i~ . (3.4)
∂x
2
p
So, for the de Broglie wave ψ(x, t) = exp( ~i (px − 2m t)), we have p̂ψ = pψ. In other
words, acting with the momentum operator on the de Broglie wave is equivalent to
multiplying by the wave momentum.6 We can rewrite (3.3) as
1 2 ∂ψ
p̂ ψ = i~ . (3.5)
2m ∂t
This is our first example of a general feature of quantum mechanics. Physically
significant quantities (in this case momentum) are represented by operators. These
operators act on functions that represent physical states (in this case the idealized
state defining the de Broglie wave).
Formally, we define an operator Ô to be a linear map from a space of functions7
to itself, i.e. any map such that
assume that they are suitably “well behaved”. For example, and depending on the context, we
might want to consider the space of infinitely differentiable functions, C ∞ (R), or the space of
“square integrable” functions, i.e. those satisfying Eqn. (3.11).
24
3.3 The 1D Schrödinger equation for a particle in a potential
We want to consider particles subject to a potential V (x) as well as free particles.
Examples: alpha rays scattering from a nucleus, electrons diffracting from a
crystal, buckyballs going through an interferometer, neutrons or larger massive
particles moving in a gravitational field.
p2
H= + V (x) , (3.7)
2m
namely
p̂2
Ĥ = + V (x) , (3.8)
2m
where the second operator corresponds to multiplication by V (x).
This gives the general form of the time-dependent 1D Schrödinger equation for
a single particle:
∂ψ
Ĥψ(x, t) = i~ . (3.9)
∂t
Or, more explicitly:
~2 ∂ 2 ψ ∂ψ
− + V (x)ψ = i~ . (3.10)
2m ∂x2 ∂t
Note that there is no way to prove mathematically that Eqns. (3.9, 3.10) are
physically relevant, although we have given some motivation for them in the light
of previous physical models and experimental results. As with any new physical
theory, the only real test is experiment. Since it is not yet obvious what the complex-
valued solutions to (3.9, 3.10) have to do with experimentally observable quantities,
we will first need to give rules for interpreting them physically, and then test these
predictions.
25
3.4 The wavefunction
We call a complex valued function ψ(x, t) that is a solution to (3.9) or (3.10) a
wavefunction. We say the wavefunction ψ(x, t) is normalisable (at time t) if
Z ∞
0< |ψ(x, t)|2 dx < ∞ . (3.11)
−∞
Note that for any complex valued ψ(x, t) the integral is real and non-negative.
Z ∞
|ψ(x, t)|2 dx = 1 . (3.12)
−∞
R∞
So, given a normalisable ψ(x, t) with −∞ |ψ(x, t)|2 dx = C, so that 0 < C < ∞,
we can define a normalised wavefunction C −1/2 ψ(x, t).
26
3.4.2 What the wavefunction definitely isn’t
Schrödinger initially hoped to interpret the wave function as describing a dispersed
cloud of physical material that somehow corresponds to a “smeared-out” particle.
This looks a natural interpretation at first sight, but proved untenable and was
abandoned.
One problem with this interpretation is that if a charged particle is really a
dispersed cloud of charge, we would expect to be able to detect bits of the cloud
carrying fractions of the charge of the electron. However, we always find that
charged objects carry a charge that is some integer multiple of the electron charge.
Classical electrodynamics also suggests that a dispersed charge of cloud should in-
teract repulsively with itself via the Coulomb force, and thus tend to be additionally
dispersed, in a way that we do not observe.
Even if these objections could somehow be overcome, another problem remains.
No matter how widely the electron’s wavefunction is spread out in space, when we
look for it by setting up detectors we always find an apparently pointlike particle
in a definite location. If the wavefunction really represented a dispersed cloud, this
cloud would have to suddenly coalesce into a particle at a single point when we
carry out a measurement. This would mean that parts of the cloud would have to
travel extremely fast — often much faster than light speed. This is inconsistent
with special relativity.
~2 ∂ 2 ∂
− 2
ψ(x, t) + V (x)ψ(x, t) = i~ ψ(x, t)
2m ∂x ∂t
is linear in ψ(x, t): if ψ1 and ψ2 are solutions then a1 ψ1 + a2 ψ2 is a solution too,
for any complex a1 , a2 .
27
3.6 Probabilistic interpretation of the wavefunction: the Born
rule
Max Born in 1926 explained the essential connection between the wavefunction and
experiment, via the so-called Born rule:
If we carry out an experiment to detect the position of a particle described
by a normalised wavefunction ψ(x, t), the probability of finding the particle in the
interval [x, x + dx] at time t is
Z x+dx
|ψ(y, t)|2 dy ≈ dx|ψ(x, t)|2
x
= dxρ(x, t) , (3.13)
Z b
|ψ(y, t)|2 dy . (3.14)
a
Intuitively, it may seem that (3.14) should follow from (3.13). Certainly, it would
be peculiar if the probability of finding a particle in a given interval depended on
how the interval was sub-divided (i.e. on how precise our position measurements
were). But we have already seen some apparently peculiar predictions of quantum
mechanics, which show it is not safe to rely on intuition. We should rather under-
stand (3.14) as a general postulate from which (3.13) follows as a special case. We
will see later (section 6.6) that (3.14) itself is a special case of the general quantum
measurement postulates, which apply to measurements of any physical quantity
(not only position).
28
3.6.1 Probability density and probability current
The following mathematical quantities give very useful insights into the behaviour
of solutions to the Schrödinger equation:
The probability density
We see that the Born rule justifies the interpretation of ρ(x, t) as a probability
density. If we measure the position of the particle at time t, the probability of
finding the particle in the interval [x, x + dx] is ρ(x, t)dx.
The probability current
i~ ∗ ∂
J(x, t) = − {ψ (x, t) ψ(x, t)−
2m ∂x
∂ ∗
( ψ (x, t))ψ(x, t)} . (3.16)
∂x
It is easy to verify from (3.10) that
∂J ∂ρ
+ = 0. (3.17)
∂x ∂t
Thus ρ and J do indeed satisfy a conservation equation, with ρ behaving as a
density whose total integral is conserved, and J as a current describing the density
flux.
The key point here is that ∂ρ ∂t can be written as a spatial derivative of some
quantity. This means that we can calculate the time derivative of the probability
of finding the particle in a region [a, b]:
Z b Z b
d 2 ∂
|ψ(x, t)| dx = − J(x, t)dx = J(a, t) − J(b, t) .
dt a a ∂x
(3.18)
The last term describes the probability density flux across the endpoints of the
interval – the “net flow of probability out of (or in to) the interval”.
Now if ψ is normalised, i.e. Eqn (3.12) holds, then ψ(x) → 0 as x → ±∞.
∂
Thus J(x) → 0 as x → ±∞, assuming (as we will here) that ∂x ψ(x) is bounded as
x → ∞. Thus
Z ∞
d
|ψ(x, t)|2 dx = lim J(a, t) − lim J(b, t) = 0 . (3.19)
dt −∞ a→−∞ b→∞
TheR∞total probability of finding the particle in −∞ < x < ∞ is thus constant over
time: −∞ |ψ(y, t)|2 dy = 1 for any time t. So, the Born probabilistic interpretation
is consistent: whenever we look for the particle, we will definitely find it somewhere,
and only in one place.
Notes:
29
• We will consider measurements of quantities other than position later.
• The Born rule says nothing about where the particle is if we do not measure
its position. According to the standard understanding of quantum mechanics,
this is a question with no well-defined answer: the particle’s position is not
defined unless we measure it.
As we’ll see, according to quantum mechanics, we can generally only calcu-
late the probabilities of possible measurement results: we can’t predict with
certainty which result will occur. Moreover, the theory only allows us to pre-
dict probabilities for the possible results of measurements that actually take
place in a given experiment. We cannot consistently combine these predic-
tions with those for other measurements that could have been made had we
done a different experiment instead.11
We’ll see when we discuss the general measurement postulates of quantum
mechanics in section 6.6 that measuring the position of a particle generally
changes its wavefunction. Recall the earlier discussion of the 2-slit experiment.
We found no definite answer to the question “which slit did the particle go
through?” – unless we put detectors beside the slits, and this changed the
experiment and changed the interference pattern.
11 See again the analysis of the double slit experiment above, and (for example) the relevant
30
4 Solutions of the 1D Schrödinger equation
We now look at various examples of solutions to the 1D SE for one particle:
This gives us
∂
(Ĥψ(x))T (t) = ψ(x)(i~ T (t)) , (4.3)
∂t
and so
1 1 ∂
Ĥψ(x) = (i~ T (t)) . (4.4)
ψ(x) T (t) ∂t
Since the left hand side depends only on x and the right hand side only on t, both
must equal a constant, which we call E. We thus have
−i
T (t) = exp( Et) . (4.5)
~
and the time-independent Schrödinger equation
~2 d2
− ψ(x) + V (x)ψ(x) = Eψ(x) . (4.6)
2m dx2
Solutions to the SE of the form
−i
ψ(x) exp( Et),
~
where ψ(x) is a solution to (4.6) are called stationary states.
We say ψ(x) is an eigenfunction of the hamiltonian operator Ĥ which corre-
sponds to a physical state whose energy is given by the eigenvalue E. When V = 0,
(4.6) reduces to
~2 d2
− ψ(x) = Eψ(x) . (4.7)
2m dx2
More generally, if
Aψ(x) = aψ(x) (4.8)
31
for some operator A and complex number a, then we say ψ(x) is an eigenfunc-
tion of A with eigenvalue a. The terminology is a natural extension to infinite-
dimensional spaces (of functions) of the definitions of eigenvector and eigenvalue
for finite-dimensional matrices: compare the eigenequation Av = av for matrix A,
vector v and complex number a.
As we will see, the eigenvalues of physically significant operators are observable
physical quantities. They turn out always to be real – as we would expect, since
we know of no natural way to directly observe a complex number in nature. We
will see the mathematical reason for this when we discuss the general formalism of
quantum mechanics.
Notes
2 2
k
• Equation (4.7) has solution ei(kx−ωt) , where E = ~2m = ~ω. This is the
de Broglie wave we originally considered, so our discussion is at least self-
consistent.
• This de Broglie wave solution to the 1D SE for a free particle is not normalis-
able, and so does not have a well-defined probabilistic interpretation via the
Born rule. The quantum mechanical free particle is a very useful idealisa-
tion, but not a physical solution. We never actually have a uniformly zero
potential throughout space, nor a single particle whose wavefunction is spread
uniformly over all of space. There could not be any consistent way of assign-
ing probabilities to finding such a particle in finite regions, since there is no
well-defined uniform probability distribution on the 1D real line. (The same
is true in 3D, of course.)
iEt
• Any stationary state ψ(x, t) = ψ(x)e− ~ has probability density
An important fact is that the general solution to the SE with t-independent potential
V (x) is a superposition of stationary states:
∞
iEn t
X
ψ(x, t) = an ψn (x)e− ~ . (4.9)
n=1
This follows from the fact that any wavefunction at a given time, say t = 0, can
be written as a superposition of energy eigenfunctions, so that,
∞
X
ψ(x, 0) = an ψn (x) , (4.10)
n=1
32
and from the linearity of the SE.
We will see later (from theorem 6.3) that Eqn. (4.10) is a special case of the
more general result that any wavefunction can be written as a superposition of
eigenfunctions for any operator corresponding to a physically observable quantity.
If the set of energy eigenvalues is continuous, we need to replace the sum by
an integral. If there are continuous and discrete subsets, we need both sum and
integral. Thus if there is a discrete set of normalised energy eigenfunctions {ψi }N
i=1
N Z
− iE~n t iEα t
X
ψ(x, t) = an ψn (x)e + aα ψα (x)e− ~ dα , (4.11)
n=1 ∆
33
4.3 Example II: Gaussian wavepackets
This gives us
s
0 2π (k0 σ + ix)2
ψ(x, t) = C exp( ) (4.14)
σ + i~t
m 2(σ + i~t
m )
s
2 00 1 (k0 σ + ix)2 (k0 σ − ix)2
|ψ(x, t)| = C 2 2 exp( + )
σ 2 + ~mt2 2(σ + i~t m ) 2(σ − i~t
m )
s 2k σx~t
1 σ 3 k02 − σx2 + 0m
= C 00 2 2 exp( 2 2 )
σ 2 + ~mt2 σ 2 + ~mt2
s
000 1 −σ(x − km 0 ~t 2
)
=C 2 2 exp( 2 2 ) , (4.15)
σ 2 + ~mt2 σ 2 + ~mt2
where C 0 , C 00 and C 000 are constants.
Since
s s
∞ 2 2
2(σ 2 + ~mt2 )π
Z
1
|ψ(x, t)|2 dx = C 000 2 2
−∞ 2σ (σ 2 + ~mt2 )
r
π
= C 000 (4.16)
σ
the normalisation condition gives us
r
000 σ
C = . (4.17)
π
Notes:
• The Gaussian wavepacket describes an approximately localised particle, almost
all of whose probability density is within a finite region of size ≈ N (σ 2 +
34
~2 t2 12 − 21
m2 ) σ at time t, for some smallish positive number N .12 This follows since
the probability density function is a Gaussian curve with standard deviation
2 2 1 1
(σ 2 + ~mt2 ) 2 σ − 2 .
• Recall that we obtained the Gaussian wavepacket as a superposition of sta-
tionary states. Considering its evolution, we thus see that a superposition of
stationary states is not necessarily stationary, since the Gaussian wavepacket’s
probability density function varies with time.
• From Eqn. (4.15), we see the speed of the Gaussian wavepacket’s centre is
≈ ~k
m . Since its standard deviation is
0
~2 t2 1 − 1 ~t
(σ 2 + 2
)2 σ 2 ≈ 1 (4.18)
m mσ 2
for large t, the speed at which it spreads in width is ≈ ~ 1 . Note that,
mσ 2
mathematically speaking, neither is necessarily bounded by the speed of light
c. This illustrates that the Schrödinger equation is non-relativistic – i.e. in-
compatible with special relativity. If it was possible to produce Gaussian
wavepackets for any values of k0 and σ, and if the Schrödinger equation pre-
cisely described their evolution, we could use the wavepackets to send signals
faster than light.13
As this suggests, the SE is only approximately valid. 14 We cannot use Gaus-
sian wavepackets (or, unless we find experimental evidence against special
relativity, anything else) to send signals faster than light. However, the SE
is a good enough approximation to allow us to understand a great deal of
physics and chemistry.
• Pictures of an evolving Gaussian wave packet can be found at (for example)
http://demonstrations.wolfram.com/EvolutionOfAGaussianWavePacket/
0 |x| < a ,
V (x) = (4.19)
∞ |x| > a .
~2 d2
− ψ(x) + V (x)ψ(x) = Eψ(x) , (4.20)
2m dx2
12 For example, 99.7% of the probability density is within the region defined by N = 3.
13 We can also see that the SE is non-relativistic by noting that it is second order in x and first
order in t. A relativistically invariant equation must be of the same order in x and t, because
Lorentz transformations map x and t to linear combinations.
14 Quantum mechanics, which is the topic of this course, is a non-relativistic theory, which is
approximately correct when relativistic corrections are small. Physicists tend to think of quantum
theory as defining a mathematical framework, and non-relativistic quantum mechanics as a par-
ticular theory within this framework. The framework of quantum theory also includes relativistic
quantum field theories, which (as the name suggests) are consistent with special relativity. In
particular, there is a relativistically invariant equation for the evolution of an electron’s wavefunc-
tion, from which the SE emerges as a non-relativistic limit. We won’t worry about relativistic
corrections or extensions of quantum mechanics in this course; they are discussed in Part III.
35
which gives us
~2 d2
Eψ(x) |x| < a ,
− ψ(x) = (4.21)
2m dx2 (E − ∞)ψ(x) |x| > a .
Here the second alternative is a useful but unrigorous shorthand, and should per-
haps be written in quotation marks to emphasize this. The next paragraph explains
how we handle the infinity.
So, for the infinite square well we expect ψ to be continuous everywhere, and
dψ
dx to be continuous except at |x| = a. Since V = ∞ for |x| > a, we need ψ = 0 in
this region for consistency of the SE. We have
r
2mE
ψ(x) = A cos(kx)+B sin(kx) , k= in |x| < a .
~2
(4.22)
Now ψ = 0 at x = ±a gives
and hence
A cos(ka) = 0 and B sin(ka) = 0 . (4.24)
36
Hence
either A = 0 , sin(ka) = 0 ,
nπ
ψ(x) = B sin(kx) , k= for n ≥ 2 even .
2a
or B = 0 , cos(ka) = 0 ,
nπ
ψ(x) = A cos(kx) , k= for n ≥ 1 odd .
2a
(4.25)
~2 π 2 n2
E = En = for n = 1, 2, 3, . . . (4.26)
8ma2
Notes:
• We have quantised (i.e. discrete) energies for the stationary states of the infi-
nite square well. Cf. the Bohr atom – it begins to look plausible that quantum
mechanics could explain the apparent quantisation of physical states. Perhaps
the allowed electron orbits in the Bohr atom correspond to the stationary
states in a 3D Coulomb potential?15
15 We will see later that the relationship is a bit more complicated, but the intuition is along the
right lines.
37
4.5 Parity
Let us consider this last phenomenon more generally. Suppose that V (x) =
V (−x) and that ψ(x) is a solution of the time-independent SE. We have
~2 d2
− ψ(x) + V (x)ψ(x) = Eψ(x) ,
2m dx2
~2 d2
− ψ(−x) + V (x)ψ(−x) = Eψ(−x) ,
2m dx2
obtaining the second equation by substituting y = −x in the first. Hence ψ(−x) is
also a solution of the time-independent SE.
A slightly more realistic model of a confined particle is given by the finite square
well potential
−U |x| < a ,
V (x) = (4.30)
0 |x| > a .
– a crude model of the Coulomb potential of a charged nucleus. We will look
for bound state solutions – normalisable solutions – as opposed to unnormalisable
scattering solutions, whose physical interpretation we will discuss later.
~2 d2
(E + U )ψ(x) |x| < a ,
− ψ(x) = (4.31)
2m dx2 Eψ(x) |x| > a .
If E > 0, the solution for |x| > a takes the form ψ = Aeikx + Be−ikx , where
k 2 = 2mE
~ . This is unnormalisable.
16 One can also prove that the ground state (lowest energy solution), if one exists, is unique and
has even parity. In fact, one can prove the stronger result that the ground state wavefunction is
unique, real and strictly positive, which implies that it must have even parity. See for example,
volume 4 of Reed and Simon, “Methods of Modern Mathematical Physics”, Theorem XIII.4.6. A
proof is also given in Eugene Lim’s lecture notes for an earlier version of this course, available
online.
38
So we require E < 0. We will assume that E > −U , i.e. that |E| < U , since we
do not expect there to be solutions with energy lower than the minimum potential.
Let
2m|E|
k2 = , (4.32)
~2
2m(U − |E|)
l2 = , (4.33)
~2
2mU
so that k 2 + l2 = 2 . (4.34)
~
Ae−kx x > a ,
ψ(x) = (4.36)
Bekx x < −a .
For |x| < a we have
Since we know that the states of definite parity span the space of all bound
states, we can solve separately for even and odd parity states to get a complete
spanning set of solutions.
Consider the even parity solutions. If ψ(x) = ψ(−x) we have:
Ae−kx
x > a,
kx
ψ(x) = Ae x < −a , (4.38)
C cos(lx) |x| < a.
−kAe−kx
x > a,
0 kx
ψ (x) = Ake x < −a , (4.39)
−lC sin(lx) |x| < a.
As ψ and ψ 0 are continuous at x = ±a, we find
39
This gives
k = l tan(la)
or
(ka) = (la) tan(la) (4.42)
2 2mU a2
2
(ka) + (la) = . (4.43)
~2
We can solve these last two equations graphically.
The odd parity bound states (if any exist) can be similarly obtained.
The finite square well illustrates a general feature of quantum potentials V (x)
such that V (x) → 0 as x → ±∞. The time-independent SE has two linearly
independent solutions.
For E > 0, these take the form of scattering solutions
Only for special values of E < 0 can we find solutions such that B = 0 and
C(A, B) = 0, as required for normalisability. For the remaining values of E < 0,
the solutions blow up exponentially as x → −∞ or as x → ∞ (or both) and are
unphysical: neither bound states nor scattering solutions.
[These asymptotically exponential functions are not physically meaningful: they
are not normalisable and do not define eigenfunctions of Ĥ.]
40
4.7 Example V: The quantum harmonic oscillator
has potential
1
V (x) = mω 2 x2 . (4.44)
2
This is a particularly important example for several reasons.
First, it is not only solvable but (as we will see later) has a very elegant solution
that explains some important features of quantum theory.
Second, generic symmetric potentials in 1D are at least approximately described
by Eqn. (4.44). To see this, consider a potential V (x) taking a minimum value Vmin
at xmin and symmetric about xmin . Writing y = x − xmin , we have
y 2 00 y3
V (y) = V (0) + yV 0 (0) + V (0) + V 000 (0) + . . .
2! 3!
y 2 00
≈ Vmin + V (0) + O(y 4 ) .
2!
Renormalising the potential so that Vmin = 0, and taking V 00 (0) = mω 2 , we recover
(4.44). (If V 00 (0) = 0 this argument does not hold, but even in this case the harmonic
oscillator potential can be a good qualitative fit. See for example comments below
on approximating a finite square well potential by a harmonic oscillator potential.)
In fact, we can show something more general. Suppose we have a system with n
degrees of freedom that has a unique stable equilibrium. Then the small oscillations
about equilibrium can be approximated by n independent harmonic oscillators. This
means that solving the harmonic oscillator allows us to give a quantum description
of molecules excited by radiation (and hence understand the greenhouse effect and
many other phenomena) and of the behaviour of crystals and other solids.
Third, the harmonic oscillator plays an even more fundamental role in quantum
field theory, where we understand particles as essentially harmonic oscillator exci-
tations of fields. Our entire understanding of matter and radiation is thus based on
the quantum harmonic oscillator!
To obtain the stationary states of the quantum harmonic oscillator we need to
solve
~2 d2 1
− 2
ψ(x) + mω 2 x2 ψ(x) = Eψ(x) . (4.45)
2m dx 2
We define r
mω 2E
y= x and α = . (4.46)
~ ~ω
We have
d2 ψ
2
− y 2 ψ = −αψ . (4.47)
dy
41
2
− y2
ψ(y) = H(y)e , (4.48)
X∞
H(y) = am y m , (4.49)
m=0
and trying to match the coefficients am to produce an exact solution.
d2 y2 2 y2
(H exp(− )) + (α − y )H exp(− ) = 0 (4.51)
dy 2 2 2
00 0 2 y2 2
(H − 2yH + Hy − H + (α − y )H) exp(− ) = 0 (4.52)
2
00 0 y2
(H − 2yH + (α − 1)H) exp(− ) = 0 (4.53)
2
We have
∞
X
H 0 (y) = mam y m−1 , (4.54)
m=0
X∞
H 00 (y) = m(m − 1)am y m−2
m=0
X∞
= am+2 (m + 2)(m + 1)y m . (4.55)
m=0
(2m + 1 − α)
am+2 = am for m = 0, 1, 2, . . . . (4.57)
(m + 2)(m + 1)
which determines the am for m ≥ 2 in terms of a0 and a1 .
42
2
For large m, (4.57) gives am+2 ≈ m am , which asymptotically describes the
2
coefficients of ey :
∞ ∞
y2
X 1 2n X
e = y = bn y n (4.58)
n=0
n! n=0
where for n odd we have bn = 0 and for n even we have
1 2 2
bn+2 = n bn = bn ≈ bn . (4.59)
2 +1 n+2 n
So, if the power series is infinite, we expect
2
H(y) ≈ Cey
H(y)
for some nonzero constant C: i.e. that Cey2
→ 1 as y → ∞. This implies that
y2 1 2
ψ(y) ≈ H(y)e− 2 ≈ Ce 2 y , (4.60)
2E
α= = 2m + 1 (m = 0, 1, 2, . . .) (4.65)
~ω
43
i.e.
1
E = (m + )~ω (m = 0, 1, 2, . . .) . (4.66)
2
The harmonic oscillator has energy levels equally spaced, separated by ~ω, with
minimum (or zero-point) energy 21 ~ω.
H0 = 1 , H1 = 2y , H2 = 4y 2 − 2 , H3 = 8y 3 − 12y , . . . . (4.67)
17 In fact one can derive a general expression (see e.g. Schiff, “Quantum Mechanics”, 3rd edition):
2 dn −y2
Hn (y) = (−1)n ey e .
dy n
44
5 Tunnelling and Scattering
Let us reconsider the bound state solutions for the finite square well potential:
−kx
e x > a,
ψ(x) ∼ ekx x < −a , (5.1)
cos(lx) or sin(lx) |x| < a.
The Born rule (3.13) thus tells us that the probability of finding the particle
outside the well is non-zero, even though the particle’s energy is less than the
potential height in this region (V = 0 in |x| > a by our convention).
Clearly, a classical particle with E < 0 would never make its way outside the
well. We have here a suggestive argument for the existence of an intrinsically
quantum phenomenon – the possibility of tunnelling through a potential barrier
into a classically forbidden state.
However, the physical interpretation of this calculation is complicated by the fact
that we can only assign meaning in quantum mechanics to things we can detect,
and we do not have any way to detect negative energy particles. So, we cannot
actually observe a bound state particle outside the region |x| < a unless we alter
the potential.
What we can and do observe are quantum particles tunnelling through finite
width potential barriers through which a classical particle of the same energy could
not pass.
ψ ∼ ψr + ψt . (5.3)
45
Some pictures of this process can be found in Schiff (3rd edition, pp 106-9), and
in Brandt and Dahmen, “Picture Book of Quantum Mechanics”.
ψ = exp(ikx) + R exp(−ikx) as x → −∞
ψ = T exp(ik 0 x) as x → ∞ .
R exp(−ikx) as x → −∞ ,
, and a transmitted plane wave of form
T exp(ik 0 x) as x → ∞ .
The beam picture is perhaps the most intuitive, and we will use it, considering
beam scattering from various potentials. We can interpret |R|2 as the density of
particles in the reflected beam, |T |2 as the density in the transmitted beam. Recall
p
that we have p = ~k and so the speed v = m = ~k
m . The particles in the incoming,
reflected and transmitted beams thus have speeds
~k ~k ~k 0
,− , (5.6)
m m m
46
respectively.
We define the particle flux in the beams to be the number of particles per second
in the beam passing a fixed point. We have
~k
m incoming ,
2
flux = velocity × density = m 0|R|
~k
reflected ,
2
m |T |
~k
transmitted.
(5.7)
Particle conservation – no particles are destroyed or created in the scattering process
– thus implies that
~k ~k 2 ~k 0 2
= |R| + |T | . (5.8)
m m m
We will verify this in particular examples.
~2 d2
(E − U )ψ(x) x > 0 ,
− ψ(x) = (5.10)
2m dx2 Eψ(x) x < 0.
For the moment we consider U > 0 (a step rather than a drop).
We want solutions of the form exp(ikx) + R exp(−ikx) for x < 0; these have
2 2
k
E = ~2m > 0.
Case 1: E < U Define
√ p
2mE 2m(U − E)
k= , l= . (5.11)
~ ~
We have
d2 l2 ψ(x)
x > 0,
ψ(x) = 2 (5.12)
dx2 −k ψ(x) x < 0 .
So
A exp(−lx) x > 0,
ψ(x) = (5.13)
exp(ikx) + R exp(−ikx) x < 0 .
We see there is no transmitted plane wave in this case.
47
Continuity of ψ and ψ 0 at x = 0 gives
1+R=A (5.14)
Hence
2k
A= (5.16)
k + il
k − il
R= . (5.17)
k + il
In particular |R|2 = 1: the reflected flux equals the incoming flux, and thus the
reflection probability for any given incoming particle is one.
and
T exp(ilx) x > 0,
ψ(x) = (5.19)
exp(ikx) + R exp(−ikx) x < 0 .
Continuity of ψ, ψ 0 at x = 0 gives
T =1+R (5.20)
l
T = 1−R. (5.21)
k
Hence
2k
T = , (5.22)
k+l
k−l
R= . (5.23)
k+l
~k
FI = (5.24)
m
~k 2 ~k k − l 2
FR = |R| = ( ) (5.25)
m m k+l
~l ~l 4k 2
FT = |T |2 = ( ). (5.26)
m m (k + l)2
48
We see that
~ k(k − l)2 + 4k 2 l
F R + FT =
m (k + l)2
~ k(k + l)2
=
m (k + l)2
~k
=
m
= FI (5.27)
k ≈ l , FR ≈ 0 , FT ≈ 1 .
This implies near-perfect transmission, again as classical intuition would suggest.
3. But the results in general do not accord with classical intuition. Consider for
example the case U < 0 and E |U |. Here we have
l k , |R|2 ≈ 1 , |T |2 ≈ 0 .
In other words, we find near perfect reflection from a downward step, precisely
the opposite result to that indicated by classical intuition.
We have
T exp(ikx) for x > a ,
ψ(x) = A exp(−lx) + B exp(lx) for 0 < x < a ,
exp(ikx) + R exp(−ikx) for x < 0 ,
(5.29)
49
where
~2 k 2 ~2 l2
E= , U −E = . (5.30)
2m 2m
4ikle−ika
T = . (5.35)
(2ikl + l2 − k 2 )e−la + (2ikl − l2 + k 2 )ela
This gives
16k 2 l2
|T |2 = . (5.36)
4k 2 l2 (ela + e−la )2 + (l2 − k 2 )2 (ela − e−la )2
16k 2 l2
|T |2 ≈ . (5.37)
4k 2 l2 exp(2la) + (l2 − k 2 )2 exp(2la)
16k 2 l2
≈ exp(−2la) . (5.38)
(k 2 + l2 )2
√
2m(U −E)
Since l = ~ , we find
2a p
|T |2 ≈ exp(− 2m(U − E)) . (5.39)
~
50
5.4.1 *Important examples
• Nuclear fission.
• Nuclear fusion.
• Muon-catalysed nuclear fusion.
51
6 The basic formalism of quantum mechanics
2. The inner product is anti-linear in the first entry and linear in the second:
52
6.3 Operators
Recall that we defined an operator Ô to be any linear map from a space of functions18
(for example C ∞ (R) or L2 (R)) to itself, i.e. any map such that
53
positions and pi its momentum. Classical dynamical variables – for instance the
energy
n
X p2i
E= + V (x1 , . . . , xn ) (6.4)
i=1
2mi
– are defined by functions on phase space. Note that the classical state of a finite
number of particles can be described by a finite number of parameters.
In principle, the classical state and (hence) the value of all classical dynamical
variables can be measured with arbitrary precision.
∂
p̂ = −i~ (6.5)
∂x
x̂ = x (i.e. multiplication by x) (6.6)
~2 ∂ 2
Ĥ = − + V (x) (6.7)
2m ∂x2
Another example is the parity operator
54
6.5 Some theorems about hermitian operators
Theorem 6.1. The eigenvalues of a hermitian operator are real.
one can find a more general notion of normalisability which covers both sets of eigenfunctions,
and more general versions of the theorems can be precisely framed in terms of this condition.
This definition includes bound states and scattering solutions to the time-independent SE, but
not solutions which blow up exponentially. A discussion can be found in, for example, Messiah,
“Quantum Mechanics”, vol. 1, chap V.9.*
55
Theorem 6.3. The discrete and continuous sets of eigenfunctions of any her-
mitian operator together form a complete orthogonal basis of the physical wavefunc-
tions, i.e. of the normalisable complex-valued functions ψ(x) of one real variable
x.
Proof. This is quite hard to prove in complete generality. We will assume it without
proof in this course.
uous set of eigenfunctions {ψα }α∈∆ , where the indexing set ∆ is some sub-interval
of the real numbers. Then any wavefunction ψ can be written as
N
X Z
ψ= ai ψi + aα ψα dα , (6.9)
i=1 ∆
as
∞
X
ψ= (ψi , ψ)ψi . (6.10)
i=1
56
6.6 Quantum measurement postulates
• Postulate 1 Every quantum observable O is represented by a hermitian op-
erator Ô.
• Postulate 2 The possible outcomes of a measurement of O are the eigenvalues
of Ô.
Notes
• It follows from postulates 2 and 3 that the total probability of all possible
outcomes is
X X X
|ai |2 = (ai ψi , ai ψi ) = (ai ψi , aj ψj ) = (ψ, ψ) = 1 .
i i ij
(6.12)
So the postulates are consistent: the sum of the probabilities of all possible
outcomes is 1, and so if you carry out a measurement you will certainly get
some outcome and you will only get one outcome. (We already verified this
in the case of the Born rule for position measurements.)
57
• If the wavefunction ψ is an eigenfunction ψi of Ô, the measurement outcome
will be λi with probability one. For example, a stationary state obeying
Ĥψ = Eψ will always give outcome E if the energy is measured.
• But unless the wavefunction ψ is an eigenfunction of the measured observable,
the measurement outcome is not definitely predictable. In contrast to classical
mechanics, a quantum observable does not generally have a definite value on
a quantum state.
• We can extendP postulates 3 and 4 to the case when Ô has degenerate eigen-
values. If ψ = i ai ψi is measured, where {ψi }∞i=1 are orthonormalised eigen-
functions of Ô and {ψi }i∈I are a complete set of orthonormalised P
eigenfunc-
tions with the same eigenvalue λ, the probability of outcome λ is i∈I |ai |2 ,
and the state
P resulting after a measurement with outcome λ is (up to normal-
isation) i∈I ai ψi .
• The projection postulate is so called because it implies that the the post-
measurement wavefunction ψi (x, t) is obtained from the pre-measurement
wavefunction ψ(x, t) by the action of the projection operator Pi defined by
Pi : ψ → (ψi , ψ)ψi , up to normalisation. We call Pi a projection since it maps
any state onto its component in a particular linear subspace, namely the sub-
space spanned by ψi – an action analogous to, for instance, the projection of
a 3D vector (x, y, z) onto its x-component (x, 0, 0).
We can similarly justify this definition of expectation value for the position op-
erator x̂ from the Born rule. Recall that the probability of obtaining a position
measurement outcome in the interval [x, x + dx] is given by |ψ(x)|2 dx. The expec-
tation value of a position measurement is thus
58
Z ∞ Z ∞
2
x|ψ(x)| dx = ψ ∗ (x)xψ(x)dx = (ψ, x̂ψ) .
−∞ ∞
Note that the expectation value is linear with respect to real scalars: i.e.
for any hermitian operators A, B and any real numbers a, b. We restrict to a, b real
here because the interpretation of hÂiψ as an expectation value of an observable
requires that A is hermitian, since observables are always represented by hermitian
operators. A complex multiple of a hermitian operator is not generally hermitian:
if A is hermitian then (aA)† = a∗ A† .
[A, B] = AB − BA . (6.16)
[A, A] = 0 ,
[A, B] = −[B, A] ,
[A, BC] = [A, B]C + B[A, C]
[AB, C] = A[B, C] + [A, C]B . (6.17)
Note that the commutator [A, B] is a sum of products of operators, and thus
itself an operator. Note also that it depends linearly on both entries.
59
The commutator plays a crucial role in describing symmetries of quantum me-
chanical systems, as we will see when we consider angular momentum. It also
gives a way of calibrating how close two operators are to having simultaneously
determinable eigenvalues: see the following note and the later discussion of the
uncertainty principle.
∂ψ
x̂p̂ψ = −ix~
∂x
∂
p̂x̂ψ = −i~ (xψ)
∂x
∂ψ
= −i~ψ − ix~
∂x
so [x̂, p̂]ψ = x̂p̂ψ − p̂x̂ψ = i~ψ
60
and as this is true for all ψ we have
It can be shown (though not in this course: see Part II Principles of Quantum
Mechanics) that these commutation relations essentially characterise the position
and momentum operators. That is, any pair of operators satisfying these relations
is equivalent (in a sense that can be made precise) to position and momentum.
Note that Theorem 6.1 implies that the expectation value and the uncertainty
are always real, as we would expect if they are physically meaningful.
Exercise Verify that (∆ψ A)2 is the statistical variance of the probability
distribution for the possible outcomes of the measurement of A on ψ, and ∆ψ A is
the distribution’s standard deviation.
Lemma 6.5. The uncertainty ∆ψ A ≥ 0 and ∆ψ A = 0 if and only if ψ is an
eigenfunction of A.
Aψ = hAiψ ψ , (6.22)
61
Theorem 6.7 (the generalised uncertainty relations). If A and B are any two
observables, and ψ is any state, then
1
∆ψ A∆ψ B ≥ |(ψ, [A, B]ψ)| . (6.23)
2
Proof. We have
(ψ, {A0 , B 0 }ψ) = (({A0 , B 0 })† ψ, ψ) = (({A0 , B 0 })ψ, ψ) = (ψ, {A0 , B 0 }ψ)∗ .
(6.25)
A similar argument shows that (ψ, [A0 , B 0 ]ψ) is imaginary.
So
1
|(ψ, A0 B 0 ψ)|2 = (|(ψ, {A0 , B 0 }ψ)|2 + |(ψ, [A0 , B 0 ]ψ)|2 ) . (6.26)
4
Combining (6.24) and (6.26), we have that
1
(∆ψ A)2 (∆ψ B)2 ≥ |(ψ, [A, B]ψ)|2 . (6.27)
4
Taking the square root gives (6.23).
Corollary 6.7.1. (the Heisenberg uncertainty principle for position and momen-
tum).
1
(∆ψ x)(∆ψ p) ≥ ~ (6.28)
2
Proof. Taking A = x̂ and B = p̂, we have [A, B] = i~, and the result follows from
Thm. 6.23.
Thus, the smaller the uncertainty in position, ∆ψ x, the greater the minimum
possible uncertainty in momentum, ∆ψ p, and vice versa.
62
Lemma 6.8. If
x̂ψ = iap̂ψ (6.29)
for some real parameter a, then (∆ψ x)(∆ψ p) = 12 ~.
which is the condition for the first term on the RHS of (6.26) to vanish. We also
have that hx̂iψ = iahp̂iψ and, since both expectations are real, this implies that
hx̂iψ = hp̂iψ = 0. Hence
which means we have equality in the Schwarz’s inequality (6.24) used to derive
(6.28).
Lemma 6.9. The condition (6.29) holds if and only if ψ(x) = C exp(−bx2 ) for
some constants b, C.
∂
Proof. If x̂ψ = iap̂ψ for some real a, we have that xψ = a~ ∂x ψ and so ψ(x) =
2 1
C exp(−bx ) for some real b = − 2a~ , and because we have equality in (6.28) we know
the uncertainty is minimised. Conversely, any wavefunction of the form ψ(x) =
C exp(−bx2 ) satisfies x̂ψ = iap̂ψ for some real a.
63
and B. (∆ψ A) and (∆ψ B) are the standard deviations for the outcomes of mea-
surements of A and B, but these hypothetical measurements are alternative possible
measurements on the same state ψ, not actual measurements carried out one after
the other. If we measured, say, first A and then B, the first measurement would
collapse the wavefunction onto an eigenfunction of A, and the second measurement
would hence not generally be a measurement on the original state ψ.
Proof. We have
d ∞ ∗
Z
dhAiψ
= ψ Aψdx
dt dt −∞
Z ∞
∂ψ ∗ ∂A ∂ψ
= ( Aψ + ψ ∗ ψ + ψ ∗ A )dx
−∞ ∂t ∂t ∂t
Z ∞
∂A i
= h iψ + ((Ĥψ)∗ Aψ − ψ ∗ A(Ĥψ))dx
∂t ~ −∞
i ∞ ∗
Z
∂A
= h iψ + (ψ ĤAψ − ψ ∗ A(Ĥψ))dx
∂t ~ −∞
i ∂A
= h[Ĥ, A]iψ + h iψ . (6.31)
~ ∂t
64
∂ Ĥ
Since none of these operators is explicitly time-dependent, we have that ∂t =
∂ x̂ ∂ p̂
∂t = ∂t = 0 and so the h ∂A
∂t iψ term on the RHS of (6.30) vanishes in each case,
giving
d dV
hp̂iψ = −h iψ ,
dt dx
d 1
hx̂iψ = hp̂iψ ,
dt m
d
hĤiψ = 0 . (6.35)
dt
d 1
These are quantum versions of the classical laws dt x= m p (which follows from
d dV d
p = mv), dt p = − dx (which follows from F = ma), and dt E = 0 (conservation of
total energy).
65
6.11 *The harmonic oscillator revisited
By considering commutation relations, we can give a much nicer and more illumi-
nating derivation of the energy spectrum of the harmonic oscillator. This derivation
forms part of the material for the Part II Principles of Quantum Mechanics course.
It is non-examinable material, in the sense that if you are asked to derive the energy
spectrum (without any method being stipulated) then the derivation given earlier is
a perfectly adequate answer. However, the derivation below is simpler and slicker,
and of course it also may be used in this context.
Recall that the harmonic oscillator hamiltonian is
p̂2 1
Ĥ = + mω 2 x̂2
2m 2
1 iω
= (p̂ + imωx̂)(p̂ − imωx̂) + [p̂, x̂]
2m 2
1 ~ω
= (p̂ + imωx̂)(p̂ − imωx̂) + (6.36)
2m 2
Define the operator a = √ 1 (p̂ − imωx̂). Since p̂ and x̂ are hermitian, we have
2m
a† = √ 1 (p̂
2m
+ imωx̂), and
1
Ĥ = a† a + ~ω . (6.37)
2
Ĥψ = Eψ .
We then have
66
In particular, if it were true that an ψ 6= 0 for all n, there would be eigenfunctions
of arbitrarily low energy, and so there would be no ground state.
However, given any physical wavefunction ψ, we have that
Z ∞
−~2 d2 ψ 1
hĤiψ = ψ∗ ( + mω 2 x2 ψ)dx
−∞ 2m dx2 2
≥ 0,
since both terms are non-negative. (Important note: this argument can obviously
be generalised to show that, if we have any potential V such that V (x) ≥ 0 for all
x, then hĤiψ ≥ 0 for all states ψ.)
Since Ĥ = a† a+ ~ω
2 and aψ0 = 0, we have Ĥψ0 =
~ω
2 ψ0 , giving us the previously
~ω
obtained value of 2 for the ground state energy. We have also obtained a closed
form expression (6.46) for the ground state and hence for the excited states,
1 mωx2
(a† )n ψ0 = C( √ (p̂ + imωx̂))n exp(− ), (6.47)
2m 2~
and we see immediately that their energies are (n + 21 )~ω, as previously obtained
by a less direct argument.
We can similarly show that there cannot be eigenfunctions ψ 0 with energies
taking values other than (n + 12 )~ω). If there were, then am ψ 0 cannot vanish for
any m, since ψ0 is the unique wavefunction annihilated by a. So there would be
negative energy eigenfunctions, which contradicts the result shown above.
With a little more thought, we can similarly also show that the eigenspaces
must all be non-degenerate: i.e. there is (up to scalar multiplication) only one
eigenfunction of each energy.
The derivation of the harmonic oscillator spectrum in this subsection illustrates
an important general feature: symmetries or regularities in a quantum mechanical
spectrum (such as the regular spacing of the harmonic oscillator energy levels)
suggest the existence of a set of operators whose commutation relations define the
symmetry or explain the regularity (in this case, the operators Ĥ, a and a† ). *
67
7 The 3D Schrödinger equation
7.1 Quantum mechanics in three dimensions
We can develop quantum mechanics in three dimensions following the analogy with
classical mechanics that we used to obtain the 1D Schrödinger equation. The clas-
sical state of a single particle is described by six dynamical variables: its position
P3 P3
x = (x1 , x2 , x3 ) = i=1 xi ei and momentum p = (p1 , p2 , p3 ) = i=1 pi ei , where ei
are the standard orthonormal basis vectors. The particle’s energy is
p.p
H= + V (x) . (7.1)
2M
(Throughout this section, we use M to denote mass, to avoid confusion with the
L3 angular momentum eigenvalue denoted by m which we introduce below.)
~2 2 ∂
− ∇ ψ(x, t) + V (x)ψ(x, t) = i~ ψ(x, t) . (7.7)
2M ∂t
Using the method of separation of variables, as before, we can derive the 3D time-
independent Schrödinger equation
~2 2
− ∇ ψ(x) + V (x)ψ(x) = Eψ(x) . (7.8)
2M
68
We can define the probability density and current
and as in the 1D case (cf (3.17)) we can show that they obey a conservation equation
∂ρ
+ ∇.J = 0 . (7.11)
∂t
Notice that the 3D Schrödinger equation, like the 1D SE, is linear and the
superposition principle thus applies to its solutions: there is a physical solution
corresponding to any linear combination of two (or more) physical solutions.
The Born rule naturally extends to the 3D case: the probability of finding a
particle in a small 3D volume V which contains a point x0 is
Z
|ψ(x, t)|2 d3 x ≈ V |ψ(x0 , t)|2 (7.12)
V
= V ρ(x0 , t) .
We can thus define the uncertainty ∆ψ A as in (6.21), using the definition (7.13) for
expectation values.
The proofs of theorems 6.1 and 6.2 apply to hermitian operators on 3D wave-
functions just as to hermitian operators on 1D wavefunctions.
Theorem 6.3 also extends to hermitian operators on 3D wavefunctions: the dis-
crete and continuous sets of eigenfunctions of a hermitian operator form a complete
orthogonal basis of th normalisable complex-valued functions ψ(x) of 3D vectors x.
69
It is convenient to use spherical polar coordinates
1 d 2d
∇2 ψ = (r )ψ
r2 dr dr
d2 ψ 2 dψ
= 2 +
dr r dr
2
1 d
= (rψ) . (7.16)
r dr2
so we have
~2 1 d2
− (rψ(r)) + V (r)ψ(r) = Eψ(r) , (7.17)
2M r dr2
which we can rewrite as
~2 d 2
− (rψ(r)) + V (r)(rψ(r)) = E(rψ(r)) . (7.18)
2M dr2
Conversely, any odd parity solution of the 1D SE for −∞ < r < ∞ defines a
solution to (7.18) with φ(r) → 0 as r → 0 and dφdr finite at r = 0. Provided that
21 See IA Vector Calculus, or e.g. Collinson “Introductory Vector Analysis”, Chap. 12.
22 * See Dirac, “The Principles of Quantum Mechanics” (4th edition), Chap. VI for a full
discussion. *
70
V (r) is finite and continuous at r = 0, these are the correct continuity conditions
for solutions of the 3D SE: they imply that ψ and ψ 0 are continuous at the origin.
Solving (7.18) thus becomes equivalent to finding odd parity solutions to the 1D
SE for −∞ < r < ∞.
Comments We will show later (see Thm. 7.2) that the ground state (the
lowest energy bound state, if there is one) of a 3D quantum system with spherically
symmetric potential is itself spherically symmetric. (Cf. the 1D result that the
ground state of a symmetric potential always has even parity.) Hence we can always
use the method above to obtain the ground state.
One might wonder whether there might not exist even parity solutions φ+ (r) of
the 1D SE with the property that φ+ (0) = dφ +
dr (0) = 0, which would also define
solutions to (7.18) for 0 ≤ r < ∞ with the appropriate properties. The following
lemma rules this out.
Lemma 7.1. There are no even parity solutions φ+ (r) of the 1D SE with the
property that φ+ (0) = dφ +
dr (0) = 0.
Proof. If such a solution φ+ were to exist, we could define a continuous odd parity
solution φ− (r) by
φ+ (r) r ≥ 0,
φ− (r) = (7.20)
−φ+ (−r) r < 0 .
Then, by the superposition principle, φ(r) = φ+ (r) − φ− (r) would also be a
solution. But we have φ(r) = 0 for r > 0, so that all derivatives of φ vanish for
r ≥ 0. The Schrödinger equation has no non-trivial solutions with this property:
hence φ(r) = 0 for all r. Hence φ+ (r) = φ− (r) = 0 for all r, so in particular
the hypothesised even parity solution φ+ is not a physical solution, as it vanishes
everywhere.
has potential
1
V (r) = M ω 2 r2 .
2
The general method we have just given for constructing spherically symmetric
stationary states shows that its spherically symmetric stationary states are related
by (7.19) to the odd parity bound states of the 1D harmonic oscillator, and have
the same energies. Thus the lowest energy spherically symmetric stationary state –
i.e. the ground state – has energy 32 ~ω, and the higher energy (excited) spherically
symmetric states have energies (2n + 32 )~ω for positive integer n.
has potential
−U r < a ,
V (r) = (7.21)
0 r > a.
By the above argument, spherically symmetric stationary states correspond to odd
parity bound states of the 1D square well potential
−U |x| < a ,
V (x) = (7.22)
0 |x| > a .
71
These, if they exist, can be obtained by the graphical method used earlier to
obtain 1D square well potential bound states. In particular, one can show (cf.
section 4 above and example sheet I, question 10) that there exists an odd parity
bound state if and only if r
2M U π
2
≥ . (7.23)
~ 2a
So, if this condition is not satisfied, the 3D spherical square well does not have a
spherically symmetric stationary state: i.e. it does not have a ground state, and
thus does not have any bound states.
As this illustrates, 3D potential wells (continuous potentials with V (x) ≤ 0 for
all x, V (x) < 0 for some x, and V (x) = 0 for |x| > a, for some finite a) do not
necessarily have bound states. This is in contrast to the 1D case:
Exercise (important!): Show that all 1D potential wells have at least
one bound state.
~2 d2 ψ 2 dψ e2
− ( 2 + )− ψ(r) = Eψ(r) , (7.24)
2M dr r dr 4π0 r
√
e2 M −2M E
for some E < 0. Writing a = 2π0 ~2 , b= ~ , we have
d2 ψ 2 dψ a
+ + ψ − b2 ψ = 0 . (7.25)
dr2 r dr r
If we try the ansatz
we see the first and fourth terms dominate the other two for large r, and cancel one
another precisely.
P∞ This suggests trying an ansatz of the form ψ(r) = f (r) exp(−br),
with f (r) = n=0 an rn , in the hope of finding values of the coefficients an such
that the four terms cancel precisely to all orders. (Cf. our first solution to the
harmonic oscillator.)
This implies that ψ(r) = O(r−1 ) as r → 0, i.e. that ψ can at worst have a singularity
of order r−1 at zero. We also require that ψ should correspond to a continuous
wavefunction. This excludes a singularity of order r−1 , so we require that ψ is
regular — i.e. has a finite limit — as r → 0.
23 *This can be shown to be an excellent approximation (see Part II Principles of Quantum
Mechanics): it gives the correct energy levels up to an overall constant factor (the same for each
energy level) of order 1 + (me /mp ) ≈ 1.0005.*
72
We have
d2 f 2 df 1
2
+ ( − 2b) + (a − 2b)f (r) = 0 . (7.27)
dr r dr r
Hence
∞
X
(an n(n − 1)rn−2 + 2an nrn−2 − 2ban nrn−1 + (a − 2b)an rn−1 ) = 0 (7.28)
n=0
This gives
(2b(n − 1) − (a − 2b))
an = an−1
n(n − 1) + 2n
2bn − a
= an−1 . (7.30)
n(n + 1)
We thus have that an → 2b
n an−1 for large n. If the coefficients do not vanish for large
n, this means they have the asymptotic behaviour of the coefficients of exp(2br),
i.e. f (r) ≈ C exp(2br). This would give ψ(r) ≈ C exp(2br) exp(−br) = exp(br),
leading to an unnormalisable and thus unphysical wavefunction. So there must be
some integer N ≥ 1 for which aN = 0, and we can take N to be the smallest such
integer.
~2 a2
E=−
8M N 2
M e4
=− , (7.31)
32π 2 20 ~2 N 2
which is precisely the energy spectrum of the Bohr orbits, but now derived from
quantum mechanics (though still with an assumption of spherical symmetry, which
we will need to relax to obtain the general orbital wavefunction).
From
2bn − a
a = 2bN and an = an−1 (7.32)
n(n + 1)
we obtain
n−N
an = an−1 2b . (7.33)
n(n + 1)
This gives solutions of the form
1 N = 1,
f (r) = (1 − br) N = 2, (7.34)
(1 − 2br + 23 (br)2 ) N = 3 ,
73
and generally f (r) = L1N −1 (2br) where L1N −1 is one of the associated Laguerre
polynomials.24
The corresponding wavefunctions are ψ(r) = CL1N −1 (2br) exp(−br), where the
constant C is determined by normalisation.
L = x ∧ p, Li = ijk xj pk , (7.39)
∂
L̂ = −i~x̂ ∧ ∇ , L̂i = −i~ijk xˆj , (7.40)
∂xk
and the total angular momentum
74
7.6.1 Angular momentum commutation relations
∂ ∂
[L̂i , L̂j ] = −~2 ilm jnp [x̂l , xˆn ]
∂xm ∂xp
∂ ∂ ∂ ∂
= −~2 ilm jnp ([x̂l , xˆn ] + xˆn [x̂l , ])
∂xm ∂xp ∂xm ∂xp
∂ ∂ ∂ ∂
= −~2 ilm jnp (x̂l [ , xˆn ] + xˆn [x̂l , ] )
∂xm ∂xp ∂xp ∂xm
∂ ∂
= −~2 ilm jnp (x̂l δmn − xˆn δlp )
∂xp ∂xm
∂ ∂
= −~2 mil mpj x̂l + ~2 pjn pmi xˆn
∂xp ∂xm
∂ ∂
= −~2 (δip δlj − δij δlp )x̂l + ~2 (δjm δni − δji δnm )(xˆn )
∂xp ∂xm
∂ ∂ ∂ ∂
= −~2 (xˆj − δij x̂l − x̂i + δij x̂l )
∂xi ∂xl ∂xj ∂xl
= i~ijk L̂k .
=
Since the L̂i do not commute, they are not simultaneously diagonalisable. However,
L̂2 and any one of the L̂i can be simultaneously diagonalised, since [L̂2 , L̂i ] = 0.
We also have
X
[L̂i , p̂2j ] = 2i~ijk p̂j p̂k = 0 . (7.46)
j
qP
2 P 2
Now we have that r̂ = j xˆj . We also have that [L̂i , j xˆj ] = 0. One can
show directly from this (see part II Principles of Quantum Mechanics), or check
by calculation (Exercise), that [L̂i , r̂] = 0. More generally, one can show that
[L̂i , V (r)] = 0 for any spherically symmetric potential V (r). We also have that
p̂.p̂ 1 X 2
[L̂i , ] = [L̂i , p ] = 0. (7.47)
2m 2m j j
75
So, for any spherically symmetric potential V (r), we have that
~2 2
[L̂i , Ĥ] = [L̂i , − ∇ + V (r)]
2M
= 0, (7.48)
In other words, Ĥ, L̂i and L̂2 all commute with one another.
Thus
∂ X ∂xi ∂
=
∂θ i
∂θ ∂xi
∂ ∂ ∂
= r cos θ cos φ + r cos θ sin φ − r sin θ , (7.51)
∂x1 ∂x2 ∂x3
∂ X ∂xi ∂
=
∂φ i
∂φ ∂xi
∂ ∂
= −r sin θ sin φ + r sin θ cos φ . (7.52)
∂x1 ∂x2
We thus obtain
∂ ∂ ∂ ∂
i~(cos φ cot θ + sin φ ) = −i~(x2 − x3 )
∂φ ∂θ ∂x3 ∂x2
= L̂1 , (7.53)
∂ ∂ ∂ ∂
i~(sin φ cot θ + cos φ ) = −i~(x3 − x1 )
∂φ ∂θ ∂x1 ∂x3
= L̂2 , (7.54)
∂ ∂ ∂
−i~ = −i~(x1 − x2 )
∂φ ∂x2 ∂x1
= L̂3 . (7.55)
76
We can also obtain
X
L̂2 = L̂2i = (L̂1 + iL̂2 )(L̂1 − iL̂2 ) + i[L̂1 , L̂2 ] + L̂23
i
2 1 d d m2
−~ ( (sin θ ) − )Y (θ) = λY (θ) . (7.58)
sin θ dθ dθ sin2 θ
From a physics perspective, the key fact about this equation is that we can show it
has non-singular solutions if and only if λ = ~2 l(l + 1) for some integers l ≥ 0 and
m such that m is in the range −l ≤ m ≤ l.
Less crucial for now are the details, although it is interesting to see them. The
solutions are called the associated Legendre functions Pl,m (θ). They can be obtained
by reducing the equation to a standard form, using the substitution w = cos θ. Since
θ is in the range 0 ≤ θ ≤ π, we have −1 ≤ w ≤ 1. We obtain the equation
d dY m2
−~2 ((1 − w2 ) ) − (λ − )Y = 0 . (7.59)
dw dw 1 − w2
For m = 0 and λ = ~2 l(l + 1) this is Legendre’s differential equation for functions
of degree l, which has solution Pl (w). For general m it’s an associated Legendre
differential equation.25
The associated Legendre functions can be obtained from the Legendre polyno-
mials Pl by
d|m|
Pl,m (θ) = (sin θ)|m| Pl (cos θ) . (7.60)
d(cos θ)|m|
We thus have the overall solution given by the spherical harmonic with total
angular momentum quantum number l and L̂3 quantum number m:
(C.U.P., 1996).
77
an eigenfunction of L̂2 and L̂3 with eigenvalues ~2 l(l + 1) and ~m respectively.
(For plots of some spherical harmonics see e.g.
mathworld.wolfram.com/SphericalHarmonic.html.)
2 2 ∂2 2 ∂ 2 1
−~ ∇ = −~ ( 2 + ) + 2 L̂2 . (7.63)
∂r r ∂r r
We can thus rewrite the SE as
~2 ∂ 2 2 ∂ 1
− ( + )+ L̂2 )ψ(r, θ, φ) + V (r)ψ(r, θ, φ) = Eψ(r, θ, φ) . (7.64)
2M ∂r2 r ∂r 2M r2
~2 d2 2 d ~2
− ( + )ψ(r)+( l(l+1))+V (r))ψ(r) = Eψ(r) .
2M dr2 r dr 2M r2
(7.65)
So, we have a standard 1D radial Schrödinger equation for ψ(r), with the modified
potential
~2 l(l + 1)
V (r) + .
2M r2
Comments:
• We have seen that we can find a basis of simultaneous eigenfunctions of Ĥ, L̂2
and L̂3 , with eigenvalues E, l and m respectively. Since the modified potential
depends on l but not m, if an energy eigenspace with eigenvalue E contains
any state with L̂2 eigenvalue l, it must contain states with all the associated
L̂3 eigenvalues: m = −l, −l + 1, . . . , l. This greatly simplifies the analysis of
orbital angular momentum eigenstates associated with a given energy level.
• If the angular momentum l = 0 then also m = 0, and the function Y00 (θ, φ) is
constant. Thus all zero angular momentum states are spherically symmetric.
Conversely, since the Ylm for l 6= 0 are orthogonal to Y00 , all spherically
symmetric states have zero angular momentum. This makes sense physically,
since a state ψ with hLiψ 6= 0 by definition has a nonzero vector associated
with it, which breaks spherical symmetry.
78
• *The fact that, in quantum mechanics, we can express −~2 ∇2 in terms of a
differential operator involving only r together with a term proportional to the
operator L2 is a consequence of the fact that, in classical mechanics, kinetic
energy can be expressed as a sum of terms proportional to the radial momen-
tum squared and the angular momentum squared. However, to derive the
first result from the second requires addressing one or two subtleties beyond
our scope. (A discussion can be found in Dirac’s book “The Principles of
Quantum Mechanics”, 4th edition, Section 38.)*
• *We can understand qualitatively why we should expect the radial SE to de-
pend on the angular momentum l, by noting that the “extra potential energy”
2
term l(l+1)~
2M r 2 corresponds to the potential needed to produce the centripetal
L2
force mr 3 which would keep a classical particle of angular momentum L in a
p
circular orbit, if we set L = ~ l(l + 1).*
7.7.1 Degeneracies
As noted above, the values of E for which this equation is solvable clearly may
depend on l but not on m. As there are (2l + 1) possible values of m = −l, −l +
1, . . . , l, each energy level would have degeneracy (2l + 1), assuming there are no
further degeneracies.
Proof. The proof is by contradiction. Suppose that ψ(r, θ, φ) = ψ(r)Ylm (θ, φ), for
some l > 0, is the lowest energy solution and has energy E. We have that
~2 d2 2 d ~2
− ( + )ψ(r)+( l(l+1))+V (r))ψ(r) = Eψ(r) .
2M dr2 r dr 2M r2
(7.66)
Now as Ĥ, L̂2 and L̂3 are commuting hermitian operators, the space of wave-
functions is spanned by their simultaneous eigenstates. In particular, the space
of zero angular momentum wavefunctions is spanned by orthonormal eigenstates
ψi (r, θ, φ) of Ĥ with E = Ei and l = m = 0, which have the form
79
∞
~2 d2
Z
2 d
E= ψ ∗ (r)(− ( 2+ )ψ(r)
r=0 2M dr r dr
~2
+( l(l + 1)) + V (r))ψ(r)) (7.68)
2M r2
Z ∞
~2 d2 2 d
= ψ ∗ (r)(− ( 2+ + V (r)))ψ(r)
r=0 2M dr r dr
Z ∞
~2
+ ψ ∗ (r)( l(l + 1))ψ(r) (7.69)
r=0 2M r2
Z ∞
~2 d2 2 d
> ψ ∗ (r)(− ( 2+ + V (r)))ψ(r)
r=0 2M dr r dr
Z ∞X
~2 d2 2 d X
= ci ψi∗ (r)(− ( 2+ + V (r))) cj ψj (r) . (7.70)
r=0 i 2M dr r dr j
Since we have that E > i |ci |2 Ei and that i |ci |2 = 1, we must have that
P P
E > Ei for at least one value of i. Hence E is not the lowest energy eigenvalue, in
contradiction to our original assumption.
d2 2 d 1 a
( 2
+ )ψ(r) + (− 2 l(l + 1)) + )ψ(r) = b2 ψ(r) . (7.72)
dr r dr r r
As we saw in discussing Eqn. (7.26), we see that the ansatz ψ(r) ≈ exp(−br) means
that the two terms which are largest asymptotically (the first term on the LHS and
the term on the RHS) cancel. This again suggests trying an ansatz of the form
ψ(r) = f (r) exp(−br), for a power series f (r).
However, the new singular term ( r12 l(l + 1)) means that the previously obtained
solutions are not generally valid.
P∞It turns out to be convenient to write the power series in the form f (r) =
n+σ
a
n=0 n r , where σ, a constant to be determined, is chosen so that a0 6= 0: i.e.
the power series begins with a term proportional to rσ .
Considering the coefficient of rσ−2 , we have −(σ(σ − 1) + 2σ) + l(l + 1) = 0 or
σ(σ + 1) = l(l + 1), a quadratic equation with roots σ = l and σ = −(l + 1). As
l ≥ 0, we choose σ = l to avoid a divergence at r = 0.
We now have
(n + l)2b − a
an = an−1 for n ≥ 1 . (7.73)
n(n + 2l + 1)
80
2b
As before, if the power series does not terminate this reduces to an ≈ n an−1 for
large n, which would give us f (r) ≈ exp(2br) and
ψ(r) ≈ exp(2br) exp(−br) ≈ exp(br) ,
a divergent and unnormalisable wavefunction, which is physically unacceptable.
The power series must thus terminate, so we have a = 2b(n + l), for some n ≥ 1.
a
Rewriting, we have b = 2N for some N ≥ l + 1, giving the same overall set of
solutions for b, and thus the same energy levels (i.e. the Bohr energy levels), as the
spherically symmetric case with l = 0 we considered earlier:
M e4 1
E=− 2 .
32π 0 ~ N 2
2 2
(The first of these degeneracies occurs only for a Coulomb force law; the second, as
we have seen, holds for any central potential.) The total number of values of (m, l)
consistent with N is thus
N −1 l N −1
X X X 1
1= (2l + 1) = 2( N (N − 1)) + N = N 2 . (7.76)
2
l=0 m=−l l=0
In fact, the true degeneracy of the N th energy level of the hydrogen atom in
a full non-relativistic quantum mechanical treatment is 2N 2 : the extra factor of 2
arises from an intrinsically quantum mechanical degree of freedom, the electron spin,
which has no direct classical analogue. (This is covered in the Part II Principles of
Quantum Mechanics course.)
81
It turns out that for relatively small atoms this gives qualitatively the right
form, with corrections arising from the electron-electron interactions that can be
calculated using perturbation theory.26 However, we also need to allow for the
Pauli exclusion principle, which implies that no two electrons in the same atom can
be in the same state. So the lowest overall energy state is given by filling up the
energy levels in order of increasing energy, starting with the lowest. Allowing for
the twofold degeneracy arising from spin, as above, we have 2N 2 states in the N th
energy level. This gives us an atom with a full energy level with Z = 2, 10 = 8+2, . . .
for N = 1, 2, . . .; these are the elements helium, neon, . . .. The elements with outer
electrons in the 1st and 2nd energy levels fill out the corresponding first two rows of
the periodic table. The analysis gets more complicated as atoms get larger, because
electron-electron interactions become more important, and this qualitative picture
is not adequate for the third and higher rows of the periodic table.
We can understand that helium and neon are chemically inert (unreactive) as
a consequence of the fact that they have full energy levels, which turns out to be
a very stable state that does not easily undergo transitions by capturing or losing
electrons.
82