0% found this document useful (0 votes)
14 views28 pages

Quantum 1 Summary

The document provides an overview of key concepts in quantum physics, focusing on the wave function and the Schrödinger equation, which describes the dynamics of quantum particles. It discusses the statistical interpretation of the wave function, the concept of probability, normalization, expectation values, and the uncertainty principle. Additionally, it covers stationary states, the infinite square well, and the harmonic oscillator, detailing their mathematical formulations and physical implications.

Uploaded by

ewan leistra
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
14 views28 pages

Quantum 1 Summary

The document provides an overview of key concepts in quantum physics, focusing on the wave function and the Schrödinger equation, which describes the dynamics of quantum particles. It discusses the statistical interpretation of the wave function, the concept of probability, normalization, expectation values, and the uncertainty principle. Additionally, it covers stationary states, the infinite square well, and the harmonic oscillator, detailing their mathematical formulations and physical implications.

Uploaded by

ewan leistra
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 28

Quantum Physics 1

Ellis de Wit

’20/’21 Block Ia

1 The Wave Function


1.1 The Schrödinger Equation
The quantum description of a particle of mass m is given by the wave function, Ψ(x, t).
It’s dynamics, or time-evolution, are given by the Schrödinger equation:
∂Ψ h̄2 ∂ 2 Ψ
ih̄ =− +VΨ
∂t 2m ∂x2
Or:
∂Ψ ih̄ ∂ 2 Ψ i
= 2
− VΨ
∂t 2m ∂x h̄
The Schrödinger equation’s complex conjugate is given by:
∂Ψ∗ ih̄ ∂ 2 Ψ∗ i
=− + V Ψ∗
∂t 2m ∂x2 h̄
The wave function of a particle can be found by solving the Schrödinger equation.
Important to note here is the use of h̄ which is Planck’s constant divided by 2π:
h
h̄ = = 1.054572 × 10−34 J s

1.2 The Statistical Interpretation


The probability of finding the particle between a and b, at time t is give by:
Z b
2
P = |Ψ(x, t)| dx
a

Here we know that


2 2
|Ψ(x, t)| = Ψ∗ Ψ or |Ψ(x, t)| = Ψ2Re · Ψ2Im
2
If Ψ(x, t) isn’t a complex number then |Ψ(x, t)| = Ψ · Ψ.
2
|Ψ(x, t)| is also known as the probability density, ρ(x). This statistical interpretation introduces a
kind of indeterminacy into quantum mechanics.
There are three positions on where the particle was just before a measurement (where is was found at
point C):
1. Realist (Einstein): the particle was at point C, but for this to be true quantum mechanics is an
incomplete theory, there is some additional information (known as a hidden variable).
2. Orthodox (Bohr): the particle wasn’t really anywhere, because of the measurement the particle
was forced to be somewhere.
3. Agnostic: refuse to answer, simply because we can’t know.
Experiments have confirmed the orthodox interpretation, when a measurement is done the wave function
collapses to a spike at the point where the particle was measured to be.

1
1.3 Probability Quantum Physics 1, Summary by Ellis de Wit

1.3 Probability
1.3.1 Discrete Variables
The average value of some function of j is given by

X
hf (j)i = f (j)P (j)
j=0

The variance of a distribution is given by


2
σ 2 = x2 − hxi

Taking the square root of this gives the standard deviation


q
2
σ = hx2 i − hxi

2
And since σ 2 is clearly non negative this implies that x2 ≥ hxi

1.3.2 Continuous Variables


The probability that x lies between a and b is given by the integral of the probability density
Z b Z b
2
Pa,b = ρ(x)dx = |Ψ(x, t)| dx
a a

We also know that the probability that x lies between −∞ and +∞ is 1 (The particle has to be some-
where).
Z +∞
P−∞,+∞ = ρ(x)dx = 1
−∞

The weighted average also known as the expectation value or mean value.
Z +∞
hxi = xρ(x)dx
−∞
Z +∞
x2 = x2 ρ(x)dx
−∞
Z +∞
hf (x)i = f (x)ρ(x)dx
−∞

1.4 Normalization
We now know that: Z +∞
2
|Ψ(x, t)| dx = 1
−∞

If you get a different finite result than 1, you can normalize Ψ, which is basically re-scaling Ψ, by adding
an constant A, such that the equation above holds again. It has two constraints:
• Ψ needs to go to 0 at ∞
• Ψ must ”fall” faster than √1
x
R +∞ 2
If −∞ |Ψ(x, t)| dx gives ∞ or if Ψ has the trivial solution 0, the wave function is non-normalizable,
because it represents nonphysical particles. Physically realizable states correspond to the square-
integrable solutions to Schrödinger’s equation.
Because the total probability is time-independent a wave function will stay normalized if it has been
normalized once (Proof in the book ).

2
1.5 Momentum Quantum Physics 1, Summary by Ellis de Wit

1.5 Momentum
For a particle in state Ψ, the expectation value of x is
Z +∞ Z +∞
2
hxi = x|Ψ(x, t)| dx = Ψ∗ xΨdx
−∞ −∞

This expectation value is not the value that you are most likely to measure, nor will it be the average of
subsequent measurements. Rather if you have an ensemble of particles, each in the same state Ψ, and
measure the positions of all of them: hxi will be the average of these results.
By taking the time derivative of the expectation value of the position we find the expectation value of
the velocity:
d
hvi = hxi
dt
Z
∂ 2
= x |Ψ| dx
∂t
−ih̄
Z

= Ψ∗ Ψdx
m ∂x
It is customary to work with momentum, which is given by:
Z  
d ∂
hpi = m · hxi = −ih̄ Ψ∗ Ψ dx
dt ∂x
Therefore, given a wave function, the expectation values for location and momentum are:
Z
hxi = Ψ∗ (x)Ψdx
Z  
∗ h̄ ∂
hpi = Ψ Ψdx
i ∂x
Here the operator x ”represents” position, and the operator (h/i)(∂/∂x) ”represents” momentum.
This can be generalized by remembering that any observable is a function of x and p. Therefore the
expectation value of any such quantity, Q = Q(x, p), can be calculated by
Z  
∗ h̄ ∂
hQ(x, p)i = Ψ Q x, Ψdx
i ∂x

1.6 The Uncertainty Principle


A general property of waves is that they cannot have both a sharp position and wavelength.
The wavelength of Ψ is related to the momentum through the de Broglie formula:
h 2π h̄
p= =
λ λ
Thus a spread in wavelength corresponds to a spread in momentum. From this follows Heisenberg’s
uncertainty principle:

σx σp ≥
2

2 Time-Independent Schrödinger Equation


2.1 Stationary States
Suppose that the potential specified in the Schrödinger equation is independent of t. In that case
separation of variables can be used to solve the Schrödinger equation:

Ψ(x, t) = ψ(x)ϕ(t)

3
2.2 The Infinite Square Well Quantum Physics 1, Summary by Ellis de Wit

This gives the time-independent Schrödinger equation

h̄2 d2 ψ
− + V ψ = Eψ
2m dx2
and the general solution of the temporal part:

ϕ(t) = e−itE/h̄

Now what is so great about these separable solutions? There are three properties that are really useful,
which are
1. They are stationary states. The wave function itself of course depends on t, but the probability
density doesn’t.
2 2
|Ψ(x, t)| = ψ ∗ e+itE/h̄ ψ e−itE/h̄ = |ψ(x)|
Following from this we get that every expectation value is constant in time. So nothing ever happens
in a stationary state.

2. They are states of definite total energy (’energy eigenstates’). In classical mechanics the total
energy is called the Hamiltonian, here the corresponding Hamiltonian operator is given by:

h̄2 ∂ 2
Ĥ = − + V (x)
2m ∂x2
This gives:
Ĥψ = Eψ
And following from that we get

hHi = E and H 2 = E2

2 2
So the variance of H is σH = H 2 − hHi = 0, this gives the property that every measurement of
the total energy is certain to return the value E.

3. The general solution is a linear combination of separable solutions. For every allowed energy
there is a different wave function and combining these separable solutions can give us a more general
solution.

X ∞
X
Ψ(x, t) = cn ψn (x)e−iEn t/h̄ = cn Ψn (x, t)
n=1 n=1

In other words; the stationary states span a basis for all space.
Important to note is the fact that the separable solutions themselves are stationary states; all their
probabilities and expectation values are independent of time, but this is not the case with the general
solution.
The boundary conditions for ψ(x) are that both ψ and dψ/dx are continuous. But where the potential
goes to infinity only the first of these applies, which is important for some of the cases we will look at
next.

2.2 The Infinite Square Well


Suppose we have the following potential
(
0, if 0 ≤ x ≤ a
V (x) =
∞, otherwise

A particle in this potential is completely free, except at the two ends, where an infinite force prevents it
from escaping.

4
2.2 The Infinite Square Well Quantum Physics 1, Summary by Ellis de Wit

Outside the well ψ(x) = 0, the probability of finding the particle there is zero. At the boundaries
x = 0 and x = a ψ(x) is also zero, this is a physical requirement, ψ must be continuous.
Inside the well the time-independent Schrödinger equation will become a simple harmonic oscilla-
tor. The solutions for ψ(x) inside the well are given by
r
2 nπ
ψn (x) = sin x.
a a
And the possible values for the energy are given by

h̄2 kn2 n2 π 2 h̄2


En = =
2m 2ma2
The wave function with the lowest energy, ψ1 is called the ground state, the others are called exited
states. As a collection the functions have some interesting properties:
1. They are alternately even and odd, with respect to the center of the well, starting with ψ1 , which
is even.
2. Each successive state has one more node (zero-crossing), starting with ψ1 , which has none (the
end points don’t count).
3. They are orthonormal (both orthogonal and normalized), this is defined by the following state-
ment: Z a

ψm (x) ψn (x)dx = δmn
0
Here δmn of course is the Kronecker delta;
(
0, if m 6= n
δmn =
1, if m = n

4. They are complete, any other function, f (x), can be expressed as a linear combination of them:
∞ r ∞
X 2X  nπ 
f (x) = cn ψn (x) = cn sin x .
n=1
a n=1 a

If you’ve paid attention in you mathematics courses you’ll recognize that this is just the Fourier
series for f (x).
Fourier’s trick (named by our one and only Griffiths) is a method to evaluate the coefficients cn for a
given f (x): Z
cn = ψn (x)∗ f (x) dx.

How should we interpret what cn is? cn basically tells you the ”amount of ψn that is contained in Ψ”.
2
Following from this is that |cn | tells you the probability that a measurement of the energy would yield
the value En . This of course gives that
X∞
2
|cn | = 1
n=1

Also the expectation value of the energy will be



X 2
hHi = |cn | En
n=1

The fact that the probability of getting a particular energy is independent of time, is a manifestation of
conservation of energy in quantum mechanics.

5
2.3 The Harmonic Oscillator Quantum Physics 1, Summary by Ellis de Wit

2.3 The Harmonic Oscillator


Next we want to look at the case where the potential is described by a harmonic oscillator. Clasically
this motion is governed by Hooke’s law, but the quantum problem is to solve the Schrödinger equation
for the potential
1
V (x) = mω 2 x2 .
2
This gives the following time-independent Schrödinger equation:

h̄2 ∂ 2 ψ 1
− + mω 2 x2 ψ = Eψ
2m ∂x2 2
There are two different approaches to solving this equation; a straightforward solution using the power
series method and a diabolically clever algebraic technique, using so-called ladder operators. We’ll
follow the book of course and do the ladder operators first.

2.3.1 Algebraic Method


First we rewrite our Schrödinger equation to
1  2
p + (mωx)2 ψ = Eψ

2m
In an ideal world we would be able to factor the Hamiltonian as follows:
1  2
p + (mωx)2

H=
2m
1
6 = (p + imωx) (p − imωx)
2m
Sadly we don’t live in an ideal world and this isn’t possible because p and x are operators and not
normal numbers. In general operators do not commute (x · p 6= p · x). We can however still examine
the quantities and define our ladder operators (more on the actual ladder part later):
1
a± = √ (p ± imωx)
2mh̄ω
Before we are going to continue we want to explore the wonderful world of commutators; commutators
are a measure of how badly (some) operators fail to commute. In general the commutator of operators
A and B is
[A, B] ≡ AB − BA.
So the commutator for x and p will be

[x, p] = x · p − p · x = ih̄,

this result is also known as the canonical commutation relation.


Now we can multiply a+ and a− with each other to find our Hamiltonian.
1
a− a+ = (p + imωx)(p − imωx)
2mh̄ω
1  2 i
p + (mωx)2 −

= [x, p]
2mh̄ω 2h̄
1 1
= H+
h̄ω 2
Note that the order of the ladder operators is important here, if you would reverse is this would give
1 1
a+ a− = H−
h̄ω 2
So clearly
[a+ , a− ] = −1

6
2.4 The Free Particle Quantum Physics 1, Summary by Ellis de Wit

In terms of a± we get a new form of the Schrödinger equation for the harmonic oscillator
 
1
h̄ω a± a∓ ± ψ = Eψ
2

Now we get to the ladder operator part! We call a± ladder operators, because they allow us to climb
up and down in energy, that is if ψ satisfies the Schrödinger equation with energy E, then a+ ψ satisfies
the Schrödinger equation with energy (E + h̄ω). In other words:

if Hψ = Eψ, then H(a+ ψ) = (E + h̄ω)(a+ ψ)

a+ then is called the raising operator. The same goes for a− , the lowering operator:

if Hψ = Eψ, then H(a− ψ) = (E − h̄ω)(a− ψ)

Important to note is the fact that a ”lowest rung” (ψ0 ) of our ladder exists, such that

a− ψ0 = 0

Using this we can find


 mω 1/4 2 1
ψ0 = e−mωx /2h̄
and E0 = h̄ω
π h̄ 2
And now that we know the equation of the ”lowest rung”, we can easily find all exited states
 
1
ψn (x) = An (a+ )n ψ0 (x), with En = n + h̄ω
2

Where An is the normalization constant. Some fancy algebraic footwork can also be used to find this
normalization constant. The proof in the book first finds the following relations
√ √
a+ ψn = n + 1ψn+1 and a− ψn = nψn−1 .

From these relations follow that


1
ψn = √ (a+ )n ψ0
n!
As in the case of the infinite square well, the stationary states of the harmonic oscillator are orthonormal,
so if we want we can Fourier’s trick to evaluate the coefficients, when we expand Ψ(x, 0) as a linear
combination of stationary states.

2.4 The Free Particle


Next we look at what you would think is the easiest case; the free particle. Here V (x) = 0 everywhere.
This gives that the Schrödinger equation will reduce to the same form as inside the infinite square well.
This time we will write our solutions in the exponential form:
h̄k2
Ψk (x, t) = Aei(kx− 2m t)

with √
2mE
k≡± .

Now, any function of x and t that depends on these variables in the special combination, x±vt, represents
a traveling wave of fixed profile (its shape doesn’t change as it propagates), which travels in the ∓x
direction, at speed v. The direction in which the wave travels in our case is governed by k:

k > 0 ⇒ travelling to the right,


k < 0 ⇒ travelling to the left.

So clearly the ”stationary states” of the free particle are propagating waves; their wavelength is λ =
2π/|k|, and they carry momentum
p = h̄k.

7
2.5 The Delta-Function Potential Quantum Physics 1, Summary by Ellis de Wit

Before we get to the velocities of these wave we’ll look at another important thing. This wavefunction is
not normalizable. So in the case of the free particle, the separable solutions do not represent physically
realizable states. There is no thing as a free particle with a definite energy. But we can still use these
separable solutions.
The general solution to the time-dependent Schrödinger equation is still a linear combination of separable
solutions, only this time it’s an integral over the continuous variable k:
Z +∞
1 h̄k2
Ψ(x, t) = √ φ(k)ei(kx− 2m t) dk
2π −∞

Now this wave function can be normalized (for appropriate φ(k)). But it necessarily carries a range of
k’s, and hence a range of energies and speeds. We call it a wave packet.
The question now is how to determine φ(k) so as to match the initial wave function
Z +∞
1
Ψ(x, 0) = √ φ(k)eikx dk.
2π −∞

Well the answer to this is provided by Plancherel’s theorem:


Z +∞ Z +∞
1 1
f (x) = √ F (k)eikx dk ⇐⇒ F (k) = √ f (x)e−ikx dx
2π −∞ 2π −∞

F (k) is called the Fourier transform of f (x); f (x) is the inverse Fourier transform of F (k) (the
only difference is in the sign of the exponent). There is, of course, some restriction on the allowable
functions: The integrals have to exist. This theorem gives
Z +∞ Z +∞
1 1
Ψ(x, 0) = √ φ(k)eikx dk ⇐⇒ φ(k) = √ Ψ(x, 0)e−ikx dx
2π −∞ 2π −∞
Now to the velocities. The speed of these waves is given by
r
h̄|k| E
vquantum = = .
2m 2m
But if we were to determiner the classical speed through E = (1/2)mv 2 we would get that
r
2E
vclassical = = 2vquantum
m
So apparently the quantum mechanical wave function travels at half the speed of the particle it is
supposed to represent, how does this work? The essential idea is this: A wave packet is a superposition
of sinusoidal functions whose amplitude is modulated by φ; it consists of “ripples” contained within an
“envelope.” What corresponds to the particle velocity is not the speed of the individual ripples (the
phase velocity), but rather the speed of the envelope (the group velocity) which, can be greater
than, less than, or equal to, the velocity of the ripples that go to make it up. So

vclassical = vgroup = 2vphase = 2vquantum .

2.5 The Delta-Function Potential


2.5.1 Bound States and Scattering States
In classical mechanics a one-dimensional time-independent potential can give rise to two rather different
kinds of motion. If V (x) rises higher than the particle’s total energy (E) on either side, then the particle
is in a bound state. It is “stuck” in the potential well—it rocks back and forth between the turning
points, but it cannot escape. If, on the other hand, E exceeds V (x) on one side (or both), then it is in
a scattering state. The particle comes in from “infinity,” slows down or speeds up under the influence
of the potential, and returns to infinity.
The distinction is even cleaner in quantum, because the phenomenon of tunneling allows the particle

8
2.5 The Delta-Function Potential Quantum Physics 1, Summary by Ellis de Wit

to “leak” through any finite potential barrier, so the only thing that matters is the potential at infinity.
Since most potentials go to zero at infinity we get
(
E < 0 ⇒ bound state,
E > 0 ⇒ scattering state.

Because the infinite square well and harmonic oscillator potentials go to infinity as x goes there too, they
admit bound states only. Because the free particle potential is zero everywhere, it only allows scattering
states.

2.5.2 The Delta-Function Well


We all already know the Dirac delta function, but let’s give a quick overview anyway :)
The Dirac delta function is an infinitely high, infinitesimally narrow spike at the origin, whose area
is 1
  Z
0, if x 6= 0
δ(x) ≡ , with δ(x) dx = 1.
∞, if x = 0

Remember that δ(x − a) would be a spike of area 1 at the point a. And that

f (x)δ(x − a) = f (a)δ(x − a),

in particular, Z +∞ Z +∞
f (x)δ(x − a) dx = f (a) δ(x − a) dx = f (a).
−∞ −∞

That’s the most important property of the delta function: Under the integral sign it serves to “pick out”
the value of f (x) at the point a.
Let’s consider a potential of the form
V (x) = −αδ(x),
where α is some positive constant. The Schrödinger equation for the delta-function well reads

h̄2 d2 ψ
− − αδ(x)ψ = Eψ,
2m dx2
and it yields both bound and scattering states.

Bound States We’ll first look at the bound states, so E < 0. In the regions x < 0 and x > 0,
V (x) = 0, so
d2 ψ 2mE
= − 2 ψ = κ2 ψ,
dx2 h̄
where √
−2mE
κ= .

The solution for the bound states then becomes
(
Beκx , (x ≤ 0),
ψ(x) =
Be−κx , (x ≥ 0).

The fact that both sides have the same constant follows from the boundary conditions for ψ(x): ψ is
always continuous and dψ/dx is continuous except at points where the potential is infinite.
These two parts of the wave function can in the end be expressed as one function. So the delta-function
well has one bound state: √
mα −mα|x|/h̄2 mα2
ψ(x) = e ; with E = − 2 .
h̄ 2h̄

9
2.6 The Finite Square Well Quantum Physics 1, Summary by Ellis de Wit

Scattering States Now we will look at the scattering states, yay! Of course now E > 0. For x < 0
and x > 0, V (x) = 0, so
d2 ψ 2mE
= − 2 ψ = −k 2 ψ,
dx2 h̄
where √
2mE
k= .

Then we get a solution of the form
(
Aeikx + Be−ikx , (x < 0),
ψ(x) =
F eikx + Ge−ikx , (x > 0).

Sadly no terms can be ruled out, since none blow up. Imposing boundary conditions gives us two
equations:

F +G=A+B and F − G = A(1 + 2iβ) − B(1 − 2iβ), with β ≡ .
h̄2 k
Now what do all these constants represent? Well A is the amplitude of a wave coming in from the left,
B is the amplitude of a wave returning to the left, F is the amplitude of a wave traveling off to the
right, and G is the amplitude of a wave coming in from the right. For a typical scattering experiments
we know what 1 constant will be:

G = 0, (for scattering from the left).

Then we know that A is the amplitude of the incident wave, B is the amplitude of the reflected wave,
and F is the amplitude of the transmitted wave.
The relative probability that an incident particle will be reflected back is given by the reflection con-
stant:
2
|B| β2 1
R≡ 2 = 2
= 2 E/mα2 )
.
|A| 1 + β 1 + (2 h̄
The probability of transmission is given by the transmission coefficient:
2
|F | 1 1
T ≡ 2 = 2
= .
|A| 1+β 1 + (mα2 /2h̄2 E)

The higher the energy, the greater the probability of transmission.


Of course, the sum of these probabilities should be 1, and it is:

R + T = 1.

Important to notice is that these scattering wave functions are not normalizable, so they don’t actually
represent possible particle states. So we must form normalizable linear combinations of the stationary
states, just as we did for the free particle, true physical particles are represented by the resulting wave
packets.
Next we’ll quickly look at the delta-function barrier, for this we only have to flip the sign of α. This
kills the bound state, since V0 = 0, but the scattering state remains unchanged. So the particle is just
as likely to pass through the barrier as to cross over the well! The particle has some nonzero probability
of passing through the potential even if E < Vmax . We call this phenomenon tunneling.

2.6 The Finite Square Well


The finite square well has a potential of the form
(
−V0 , for − a < x < a
V (x) =
0, for |x| > a

10
2.6 The Finite Square Well Quantum Physics 1, Summary by Ellis de Wit

Bound States. First we will look at the bound states; the even solutions for this case are given by

−κx
F e
 , for x > a,
ψ(x) = D cos (lx), for 0 < x < a,

ψ(−x), for x < 0.

Imposing the boundary conditions at x = a gives us a transcendental equation for z as a function of z0


p a
tan z = (z0 /z)2 − 1 with z = la, and z0 =

This can be solved numerically or by plotting both sides of the on the same grid and looking at the
points of intersection.
the odd solutions for this case are given by

−κx
F e
 , for x > a,
ψ(x) = D sin (lx), for 0 < x < a,

−ψ(−x), for x < 0.

Here we also get a similar transcendental equation for (the same) z as a function of (the same) z0
p
− cot z = (z0 /z)2 − 1,
which can be solved in the same way.
Two limiting cases are of special interest:

1. Wide, deep well. If z0 is very large, the intersections occur just slightly below zn = nπ/2, with n
even or odd; it follows that
n2 π 2 h̄2
En + V0 ∼= .
em(2a)2

2. Shallow, narrow well. As z0 decreases, there are fewer and fewer bound states, until finally (for
z0 < π/2, where the lowest odd state disappears) only one remains.

Scattering States. For the scattering states (E > 0) we will get a solution of the form

ikx −ikx
Ae + Be
 , for x < −a,
ψ(x) = C sin (lx) + D cos (lx), for − a < x < a,

 ikx
Fe , for x > a.
Like with the delta-function well we assume that there’s no incoming wave from the right. Again A is
the incident amplitude, B is the reflected amplitude, and F is the transmitted amplitude.
There are 4 boundary conditions which can be used to eliminate C and D, and solving them gives us
two quite ugly equations:
sin (2la) 2
B=i (l − k 2 )F,
2kl
e−2ika
F = 2 +l2 ) A.
cos (2la) − i (k 2kl sin (2la)
2 2
The transmission coefficient (T = |F | /|A| ) is given by
V02
 
2a p
T −1 = 1 + sin2 2m(E + V0 ) .
4E(E + V0 ) h̄
Notice that T = 1 (the well becomes “transparent”) whenever the sine is zero, which is to say, when
2a p
2m(E + V0 ) = nπ,

where n is any integer. The energies for perfect transmission, then, are given by
n2 π 2 h̄2
En + V 0 = ,
2m(2a)2
which happen to be precisely the allowed energies for the infinite square well.

11
Quantum Physics 1, Summary by Ellis de Wit

3 Formalism
Quantum theory is based on two constructs: wave functions and operators. The state of a system
is represented by its wave function, observables are represented by operators. Mathematically, wave
functions satisfy the defining conditions for abstract vectors, and operators act on them as linear
transformations. So the natural language of quantum mechanics is linear algebra. So that’s why we
first have a recap on vector spaces and then continue on to some new concepts.

Recap on some Linear Algebra, which you should kind of know..


3.0.1 Vector Spaces
A vector space is a set of vectors |αi , |βi , ... and scalars a, b, ... which is closed under some certain
operations.
Vector addition
|αi + |βi = |γi
Vector addition is commutative: |αi + |βi = |βi + |αi and associative: |αi + (|βi + |γi). It works with
the null vector : |αi + |0i = |αi and adding a vector to it’s inverse gives the null vector: |αi + |−αi = |0i.
Scalar Multiplication
a · |αi
Scalar multiplication is distributive: a · (|αi + |βi) = a · |αi + a · |βi, and also associative a · b · |αi. The
unit scalar is given by 1 · |αi = |αi and the zero scalar is given by 0 · |αi = |0i.
A vector can be a linear combination of some vectors, if a vector isn’t a linear combination it is called
linearly independent. A collection of vectors spans the vector space if and only if every vector in the
vector space is a linear combination of this collection of vectors:
X
|αi = αi · |ei i

Here the collection of vectors |ei i is called a basis of the vector space and the number of vectors in this
basis is the dimension of the vector space.
A vector space has an orthonormal basis when

hei |ej i = δij

3.0.2 Inner Product Spaces


Given two vectors, |αi and |βi, the inner product, hα|βi, will be the complex number
X
hα|βi = αi∗ βi .
i

Two properties of the inner product are



hα|βi = hβ|αi ,

and X
hα|αi = αi∗ αi ,
i

this second property is the same as the (square of the) norm of a vector, ||α||2 , and is real and non-
negative.
Schwarz inequality says that
2
|hα|βi| ≤ hα|αi hβ|βi
Linear transformations, T , are represented by matrices, which act on vectors to produce new vectors.
  
t11 t12 · · · t1N a1
 t21 t22 · · · t2N   a2 
|βi = T |αi → b = Ta =  .
  
.. ..   .. 
 .. . .  . 
tN 1 tN 2 ··· tN N an

12
3.1 Hilbert Space Quantum Physics 1, Summary by Ellis de Wit

3.1 Hilbert Space


Now we have looked at some linear algebra basics we can apply it to quantum mechanics. In quantum
mechanics the “vectors” we encounter are functions, and they live in infinite-dimensional spaces. The
collection of all functions of x constitutes a vector space, but for our purposes it is much too large. We
will look at the set of all square-integrable functions, on a specified interval
Z b
2
f (x) such that |f (x)| dx < ∞,
a

which is called Hilbert space. So wave functions live in Hilbert space.


We define the inner product of two functions, f (x) and g(x), as follows:
Z b
hf |gi ≡ f (x)∗ g(x)dx
a

The integral Schwarz inequality says


s
Z b Z b Z b
∗ 2 2
f (x) g(x)dx ≤ |f (x)| dx |g(x)| dx
a a a

The inner product for integrals also has two notable properties which are very similar to their vector
counterparts
Z b
∗ 2
hg|f i = hf |gi and hf |f i = |f (x)| dx
a

The inner product can also give you information about functions:
• a function is said to be normalized if hf |f i = 1,
• two functions are orthogonal if hf |gi = 0,
• a set of functions are orthonormal if hfm |fn i = δmn .
A set of functions is complete if any other function (in Hilbert space) can be expressed as a linear
combination of them

X
f (x) = cn fn (x).
n=1

If the functions fn (x) are orthonormal, the coefficients are given by

cn = hfn |f i .

3.2 Observables
3.2.1 Hermitian Operators
The expectation value of an observable Q(x, p) can be expressed in inner-product notation:
D E Z
hQi = Ψ|Q̂Ψ = Ψ∗ Q̂Ψ dx.

The expectation value has to be real, thus



hQi = hQi ,

but the complex conjugate of an inner product reverses the order, so


D E D E
Ψ|Q̂Ψ = Q̂Ψ|Ψ .

This must hold true for any wave function. Observables are represented by hermitian operators.
An operator is hermitian if
D E D E
f |Q̂g = Q̂f |g for all f (x) and all g(x)

13
3.3 Eigenfunctions of a Hermitian Operator Quantum Physics 1, Summary by Ellis de Wit

3.2.2 Determinate States


Determinate states for the observable Q are states where every measurement of Q is certain to return
the same value q. We’ve seen an example of determinate states before with the stationary states, which
are determinate states of the Hamiltonian.
Determinate states are eigenfunctions of Q̂.
Q̂Ψ = qΨ
This is the eigenvalue equation for the operator Q̂; Ψ is an eigenfunction of Q̂, and q is the cor-
responding eigenvalue (a number ). The collection of all the eigenvalues of an operator is called its
spectrum. Sometimes two (or more) linearly independent eigenfunctions share the same eigenvalue; in
that case the spectrum is said to be degenerate.

3.3 Eigenfunctions of a Hermitian Operator


Our attention is thus directed to the eigenfunctions of hermitian operators. These fall into two categories:
If the spectrum is discrete (i.e., the eigenvalues are separated from one another) then the eigenfunctions
lie in Hilbert space and they constitute physically realizable states. If the spectrum is continuous (i.e.,
the eigenvalues fill out an entire range) then the eigenfunctions are not normalizable, and they do not
represent possible wave functions.

3.3.1 Discrete Spectra


Mathematically, the normalizable eigenfunctions of a hermitian operator have three important properties:
• Realness: Their eigenvalues are real.
• Orthogonality: Eigenfunctions belonging to distinct eigenvalues are orthogonal.
• Completeness: The eigenfunctions of an observable operator are complete: Any function (in
Hilbert space, a finite-dimensional vector space) can be expressed as a linear combination of them.
Unfortunately this tells us nothing about degenerate states. But, if two (or more) eigenfunctions share the
same eigenvalue, any linear combination of them is itself an eigenfunction, with the same eigenvalue, and
we can use the Gram-Schmidt orthogonalization procedure to construct orthogonal eigenfunctions
within each degenerate subspace. So even in the presence of degeneracy the eigenfunctions can be chosen
to be orthogonal.

3.3.2 Continuous Spectra


If the spectrum of a hermitian operator is continuous (so the eigenvalues are labeled by a continuous
variable, p or y, as in the examples in the book; or z, generically), the eigenfunctions are not normaliz-
able, they are not in Hilbert space and they do not represent possible physical states; nevertheless, the
eigenfunctions with real eigenvalues have Dirac orthonormality
hfz0 |fz i = δ(z − z 0 ),
which is strikingly reminiscent of true orthonormality, the indices are now continuous variables, and the
Kronecker delta has become a Dirac delta. These eigenfunctions with real eigenvalues also are complete
(with the sum now an integral).

3.4 Generalized Statistical Interpretation


Generalized statistical interpretation: If you measure an observable Q(x, p) on a particle in the
state Ψ(x, t), you are certain to get one of the eigenvalues of the hermitian operator Q̂(x, −ih̄d/dx). If
the spectrum of Q̂ is discrete, the probability of getting the particular eigenvalue qn associated with the
orthonormalized eigenfunction fn (x) is
2
|cn | , where cn = hfn |Ψi .
If the spectrum is continuous, with real eigenvalues q(z) and associated Dirac orthonormalized eigen-
functions fz (x), the probability of getting a result in the range dz is
2
|c(z)| dz, where c(z) = hfz |Ψi .

14
3.5 The Uncertainty Principle Quantum Physics 1, Summary by Ellis de Wit

Upon measurement, the wave function “collapses” to the corresponding eigenstate, for continuous spectra
the collapse is to a narrow range about the measured value.
The momentum space wave function, Φ(p, t), is given by
Z +∞
1
Φ(p, t) = √ e−ipx/h̄ Ψ(x, t) dx.
2π h̄ −∞

It is essentially the Fourier transform of the (position space) wave function Ψ(x, t), which, by Plancherel’s
theorem, is its inverse Fourier transform:
Z +∞
1
Ψ(x, t) = √ eipx/h̄ Φ(p, t) dp.
2π h̄ −∞
According to the generalized statistical interpretation, the probability that a measurement of momentum
would yield a result in the range dp is
2
|Φ(p, t)| dp

3.5 The Uncertainty Principle


3.5.1 Proof of the Generalized Uncertainty Principle
The (generalized) uncertainty principle is given by
 D E
2 2 1
σA σB ≥ [Â, B̂]
2i

There is an “uncertainty principle” for every pair of observables whose operators do not commute, these
observables are also known as incompatible observables. Incompatible observables do not have shared
eigenfunctions, at least, they cannot have a complete set of common eigenfunctions. By contrast, com-
patible (commuting) observables can have complete sets of simultaneous eigenfunctions.
Note that the uncertainty principle is not an extra assumption in quantum theory, but rather a conse-
quence of the statistical interpretation.

3.5.2 The Minimum-Uncertainty Wave Packet


We have twice encountered wave functions that hit the position-momentum uncertainty limit: the ground
state of the harmonic oscillator and the Gaussian wave packet for the free particle. This raises an
interesting question: What is the most general minimum-uncertainty wave packet?
Tweaking the proof of the generalized uncertainty principle a little bit gives rise to the following condition
for minimum uncertainty:
g(x) = iaf (x), where a is real.
(Also f = (Â − hAi)Ψ and g = (B̂ − hBi)Ψ, this is shown in the proof of the generalized uncertainty
principle in the book.)
For the position-momentum uncertainty principle this criterion becomes:
!
ĥ d
− hpi Ψ = ia(x − hxi)Ψ,
i dx

for which the general solution is


2
Ψ(x) = Ae−a(x−hxi) /2h̄ ihpix/h̄.
e

Evidently the minimum-uncertainty wave packet is a gaussian—and the two examples we encountered
earlier were gaussians.

15
3.6 Dirac Notation Quantum Physics 1, Summary by Ellis de Wit

3.5.3 The Energy-Time Uncertainty Principle


The position-momentum uncertainty principle is often paired with the energy-time uncertainty prin-
ciple,

∆t ∆E ≥ .
2
In the context of special relativity the energy-time form might be thought of as a consequence of the
position-momentum version, because x and t go together in the position-time four-vector, while p and
E go together in the energy-momentum four-vector.
In the energy-time uncertainty principle ∆t doesn’t represent the standard deviation of a collection of
time measurements, but it represents the amount of time it takes the expectation value of Q to change
by one standard deviation. This is shown by the formula

d hQi
σQ = ∆t.
dt

And the time derivative of the expectation value of some operator, Q(x, p, t), is given by
* +
d i D E ∂ Q̂
hQi = [Ĥ, Q̂] + .
dt h̄ ∂t

3.6 Dirac Notation


The notation of the inner product, hα|βi, is called the bra-ket, or Dirac, notation and can be split into
two pieces, bra, hα|, and ket, |βi. The ket is a ”normal” column vector
 
a1
 a2 
|αi =  .  ,
 
 .. 
an

and the corresponding bra is its Hermitian conjugate († ), which is the combination of the transpose
(> ) and the complex conjugate (∗ ):

hα| = a∗1 a∗2 · · · a∗n




4 Quantum Mechanics in Three Dimensions


4.1 Schrödinger Equation in Spherical Coordinates
When generalizing to 3 dimensions, momentum becomes:


p→ ∇
i
Thus:
∂Ψ h̄2 2
ih̄ =− ∇ Ψ + V Ψ.
∂t 2m
Here
∂2 ∂2 ∂2
∇2 ≡ + +
∂x2 ∂y 2 ∂z 2
of course is the Laplacian. The potential energy and the wave function now are functions of r = (x, y, z)
and t. The normalization condition now reads
Z
2
|Ψ| d3 r =, 1

16
4.1 Schrödinger Equation in Spherical Coordinates Quantum Physics 1, Summary by Ellis de Wit

with the integral taken over all space and the infinitesimal volume d3 r = dx dy dz. If the potential
is independent of time we can still use the separation of variables and there will be a complete set of
stationary states:
Ψn (r, t) = ψn (r) e−iEn t/h̄ .
And the the time-independent Schrödinger equation becomes:

h̄2 2
− ∇ ψ + V ψ = Eψ
2m

4.1.1 Seperation of Variables


To make our lives easier in the long run we want to adopt spherical coordinates, (r, θ, φ),

x = r sin θ cos φ,
y = r sin θ sin φ,
z = r cos θ.

In spherical coordinates the Laplacian takes the form


∂2
   
1 ∂ ∂ 1 ∂ ∂ 1
∇2 = 2 r2 + 2 sin θ + 2 2 .
r ∂r ∂r r sin θ ∂θ ∂θ r sin θ ∂φ2
In spherical coordinates, then, the time-independent Schrödinger equation reads

h̄2 1 ∂ ∂2ψ
     
2 ∂ψ 1 ∂ ∂ψ 1
− r + 2 sin θ + 2 2 + V ψ = Eψ
2m r2 ∂r ∂r r sin θ ∂θ ∂θ r sin θ ∂φ2
We begin by looking for solutions that are separable into products:

ψ(r, θ, φ) = R(r)Y (θ, φ)

Some derivation later gives us two equations, one that only depends on r and one that depends on θ and
φ. Both must be a constant, we will write this “separation constant” in the form l(l + 1):

2mr2
 
1 d dR
r2 − [V (r) − E] = l(l + 1);
R dr dr h̄2
1 ∂2Y
   
1 1 ∂ ∂Y
sin θ + = −l(l + 1).
Y sin θ ∂θ ∂θ sin2 θ ∂φ2

4.1.2 The Angular Equation


First we will look at the angular equation, that depends on θ and φ. For this equation we can once again
use separation of variables:
Y (θ, φ) = Θ(θ)Φ(φ).
This again gives us two equations, one that only depends on θ and one that only depends on φ. Again
both must be a constant, this time we will call the separation constant m2 :
  
1 1 d dΘ
sin θ + l(l + 1) sin2 θ = m2 ;
Θ sin θ dθ dθ
1 d2 Φ
= −m2 .
Φ dφ2
The general solution of Φ(φ) then is
Φ(φ) = eimφ ,
here m is an integer that can be positive and negative. Also it is natural to require that

Φ(φ + 2π) = Φ(φ)

The general solution of Θ(θ) isn’t that easy. The solution is

Θ(θ) = APlm (cos θ),

17
4.2 The Hydrogen Atom Quantum Physics 1, Summary by Ellis de Wit

here Plm is the associated Legendre function, defined by


 |m|
2 |m|/2 d
Plm (x) ≡ (1 − x ) Pl (x),
dx

and Pl (x) is the lth Legendre polynomial, defined by the Rodrigues formula:
 l
1 d
Pl (x) ≡ (x2 − 1)l .
2l l! dx

Important to note is that l must be a nonnegative integer and that the values of m range from −l to l, with
integer intervals, because of this for any given l there are (2l + 1) possible values of m. Mathematically
speaking there exist solutions for any old values of l and m, but physically speaking some solutions are
unacceptable.
The normalized angular wave functions are called spherical harmonics:
s
m (2l + 1) (l − |m|)! imφ m
Yl (θ, φ) =  e Pl (cos θ),
4π (l + |m|)!

here  = (−1)m for m ≥ 0 and  = 1 for m ≤ 0. These are automatically orthogonal, so


Z 2π Z π h i
∗ 0
[Ylm (θ, φ)] Ylm
0 (θ, φ) sin θ dθ dφ = δll0 δmm0 .
0 0

For historical reasons, l is called the azimuthal quantum number, and m the magnetic quantum
number.

4.1.3 The Radial Equation


Notice that the angular part of the wave function, Y (θ, φ), is the same for all spherically symmetric
potentials; the actual shape of the potential, V (r), affects only the radial part of the wave function,
R(r).
The radial equation is given by

h̄2 d2 u h̄2 l(l + 1)


 
− + V + u = Eu,
2m dx2 2m r2

here the m’s are masses and


u(r) ≡ rR(r).
As you can see the radial equation is identical in form to the one-dimensional Schrödinger equation,
except that the effective potential contains an extra term called the centrifugal term. It tends to
throw the particle outward (away from the origin).
For the radial equation the normalization condition becomes
Z ∞
2
|u| dr = 1.
0

This is all we can do with the radial part of the wave function until a specific potential V (r) is provided.

4.2 The Hydrogen Atom


The hydrogen atom consists of a heavy, essentially motionless proton, of charge e, together with a much
lighter electron (charge −e) that orbits around it, bound by the mutual attraction of opposite charges.
From Coulomb’s law, the potential energy is

e2 1
V (r) = −
4π0 r

18
4.2 The Hydrogen Atom Quantum Physics 1, Summary by Ellis de Wit

4.2.1 The Radial Wave Function


The radial equation now is

h̄2 d2 u e2 1 h̄2 l(l + 1)


 
− + − + u = Eu,
2m dx2 4π0 r 2m r2
and we will be looking at the discrete bound states, representing the hydrogen atom.
The allowed energies of this equation are given by the Bohr formula:
"  2 2 #
m e 1 E1
En = − 2 2
= 2 , with n = 1, 2, 3, ...
2h̄ 4π0 n n

The energy of the ground state is given by:


"  2 2 #
m e
E1 = − = −13.6 eV
2h̄2 4π0

Evidently the binding energy of hydrogen (the amount of energy you would have to impart to the
electron in the ground state in order to ionize the atom) is 13.6 eV.
The most probable distance between the nucleus and the electron in a hydrogen atom in its ground state
is the so-called Bohr radius and is given by

4π0 h̄2
a≡ = 0.529 × 10−10 m
me2
The normalized hydrogen wave functions are

ψnlm (r, θ, φ) = Rnl (r)Ylm (θ, φ)


s 
3  l
2 (n − l − 1)! −r/na 2r  2l+1
Ln−l−1 (2r/na) Ylm (θ, φ).

= 3
e
na 2n[(n + l)!] na

Here  p
d
Lpq−p (x) ≡ (−1) p
Lq (x)
dx
is an associated Laguerre polynomial, and
 q
d
Lq (x) ≡ ex (e−x xq )
dx
is the qth Laguerre polynomial.
The ground state of hydrogen is given by
1
ψ100 (r, θ, φ) = √ e−r/a
πa3

4.2.2 The Spectrum of Hydrogen


If you perturb an hydrogen atom (by collision with another atom, say, or by shining light on it), the
electron may undergo a transition to some other stationary state—either by absorbing energy, and
moving up to a higher-energy state, or by giving off energy (typically in the form of electromagnetic
radiation), and moving down. In practice transitions are constantly occurring, the energy of the given
off light (photons) corresponds to the difference in energy between the initial and final states:
!
1 1
Eγ = Ei − Ef = −13.6 eV − 2 .
n2i nf

The Planck formula says that the energy of a photon is proportional to its frequency:

Eγ = hν.

19
4.3 Angular Momentum Quantum Physics 1, Summary by Ellis de Wit

And the Rydberg formula connects the wavelength to the principal quantum number
!
1 1 1
=R − 2 ,
λ n2f ni

where the Rydberg constant is given by:


2
e2

m
R≡ = 1.097 × 107 m−1 .
4πch̄3 4π0

The different series of hydrogen transitions also have names:


• The Lyman series consists of transitions to the ground state (nf = 1).
• The Balmer series consists of transitions to the first exited state (nf = 2).
• The Paschen series consists of transitions to the second exited state (nf = 3).
And the list goes on like that.

4.3 Angular Momentum


Classically, the angular momentum of a particle (with respect to the origin) is given by the formula

L = r × p.

This gives us in quantum mechanics three angular momentum operators:

Lx = ypz − zpy , Ly = zpx − xpz , Lz = xpy − ypx ,

with of course

px → −ih̄ ∂/∂x, py → −ih̄ ∂/∂y, pz → −ih̄ ∂/∂z.

4.3.1 Eigenvalues
Our three angular momentum operators sadly do not commute:

[Lx , Ly ] = ih̄Lz , [Ly , Lz ] = ih̄Lx , [Lz , Lx ] = ih̄Ly .

This makes them incompatible observables. According to the generalized uncertainty principle,


σLx σLx ≥ |hLz i|.
2
But introducing the square of the total angular momentum,

L2 ≡ L2x + L2y + L2z ,

helps us immensely, since this does commute with all three angular momentum operators. With this we
can find simultaneous eigenstates of L2 and (let’s say) Lz :

L2 f = λf and Lz f = µf.

We’ll use a “ladder operator” technique like we’ve seen before

L± ≡ Lx ± iLy .

This operator of is of course compatible with L2 and its commutator with Lz is [Lz , l± ] = ±h̄L± . If f
is an eigenfunction of L2 and Lz , so also is L± f :

L2 (L± f ) = λ(L± f ), Lz (L± f ) = (µ ± h̄)(L± f ).

20
4.4 Spin Quantum Physics 1, Summary by Ellis de Wit

We call L+ the “raising” operator, because it increases the eigenvalue of Lz by h̄, and L− the “lowering”
operator, because it lowers the eigenvalue by h̄. In contrary to the harmonic oscillator ladder this ladder
has both a top and a bottom rung, such that

L+ ftop = 0 L− fbottom = 0

The eigenvalues of L2 and Lz are given by

L2 flm = h̄2 l(l + 1)flm and Lz flm = h̄mflm ,

where

l = 0, 1/2, 1, 3/2, ...; m = −l, −l + 1, ..., l − 1, l,

and the eigenfunctions are characterized by the numbers l and m. For a given value of l, there are 2l + 1
different values of m (i.e., 2l + 1 “rungs” on the “ladder”).

4.3.2 Eigenfunctions
The operator Lz is given by
h̄ ∂
Lz =
i ∂φ
and the operator L2 is given by

1 ∂2
   
1 ∂ ∂
L2 = −h̄2 sin θ + .
sin θ ∂θ ∂θ sin2 θ ∂φ2
But we’ve solved this before! The result (appropriately normalized) is the spherical harmonic, Ylm (θ, φ).
Conclusion: Spherical harmonics are eigenfunctions of L2 and Lz . There is a curious final twist to this
story, for the algebraic theory of angular momentum permits l (and hence also m) to take on half -integer
values, whereas separation of variables yielded eigenfunctions only for integer values.

4.4 Spin
In classical mechanics, a rigid object admits two kinds of angular momentum: orbital (L = r × p),
associated with the motion of the center of mass, and spin (S = Iω), associated with motion about
the center of mass. Elementary particles carry intrinsic angular momentum (S) in addition to their
“extrinsic” angular momentum (L). The algebraic theory of spin is a carbon copy of the theory of orbital
angular momentum, beginning with the fundamental commutation relations:

[Sx , Sy ] = ih̄Sz , [Sy , Sz ] = ih̄Sx , [Sz , Sx ] = ih̄Sy .

It follows (as before) that the eigenvectors of S 2 and Sz satisfy

S 2 |s mi = h̄2 s(s + 1) |s mi ; Sz |s mi = h̄m |s mi ;

and p
S± |s mi = h̄ s(s + 1) − m(m ± 1) |s (m ± 1)i ,
with S± ≡ Sx ± iSy and

s = 0, 1/2, 1, 3/2, ...; m = −s, −s + 1, ..., s − 1, s.

It so happens that every elementary particle has a specific and immutable value of s, which we call the
spin of that particular species, some particles include:
• spin 0: pi mesons,
• spin 1/2: proton, neutron, electron, neutrino, and quarks,
• spin 1: photon, Z and W bosons, gluons,
• spin 3/2: deltas,
• spin 2: gravitons.

21
4.4 Spin Quantum Physics 1, Summary by Ellis de Wit

4.4.1 Spin 1/2


When you have a particle with spin 1/2 then there are just two eigenstates:

1 1
Spin up : and
2 2

1 1
Spin down : (− ) .
2 2

Using these as basis vectors, the general state of a spin-1/2 particle can be expressed as a two-element
column matrix (or spinor):  
a
χ= = aχ+ + bχ− ,
b
with
 
1
χ+ = representing spin up, and
0
 
0
χ− = representing spin down.
1

Then using the eigenvalues of the operators S 2 , S± and Sz we can find the matrix representation for the
spin operators. For this we get:
     
3 1 0 0 1 0 0
S2 = h̄2 S+ = h̄ S− = h̄
4 0 1 0 0 1 0

 

For Sx , Sy and Sz it is neater to write S = 2 σ, with
     
0 1 0 −i 1 0
σx ≡ σy ≡ σz ≡
1 0 i 0 0 −1

These last three matrices are the Pauli spin matrices.


Note that Sx , Sy , Sz and S 2 are hermitian, since they represent observables and S± aren’t hermitian,
thus not observable.
2
If you measure Sz on a particle in the general state χ, you could get +h̄/2, with probability |a| , or
2
−h̄/2, with probability |b| . Since these are the only possibilities we get,
2 2
|a| + |b| = 1

4.4.2 Electron in a Magnetic Field


A spinning charged particle constitutes a magnetic dipole. Its magnetic dipole moment, µ, is pro-
portional to its spin angular momentum, S:

µ = γS,

here γ is called the gyromagnetic ratio.


When a magnetic dipole is placed in a magnetic field B, it experiences a torque, µ × B, which tends to
line it up parallel to the field (just like a compass needle). The energy associated with this torque is

H = −mu · B,

so the Hamiltonian of a spinning charged particle, at rest in a magnetic field B, is

H = −γB · S.

22
Quantum Physics 1, Summary by Ellis de Wit

4.4.3 Addition of Angular Momenta


Suppose now that we have two spin-1/2 particles—for example, the electron and the proton in the ground
state of hydrogen. Each can have spin up or spin down, so there are four possibilities in all:

↑↑, ↑↓, ↓↑, ↓↓,

where the first arrow refers to the electron and the second to the proton. This combination of two
spin-1/2 particles can carry a total spin of 1 or 0. The three states with s = 1 are
 
 |1 1i = ↑↑ 
|1 0i = √12 (↑↓ + ↓↑) s = 1 (the triplet combination).
|1 − 1i = ↓↓
 

The orthogonal state with m = 0 carries s = 0:


 
1
|0 0i = √ (↑↓ − ↓↑) s = 0 (the singlet).
2
So in a more general form: If you combine spin s1 with spin s2 , the total spins you can get are

s = (s1 + s2 ), (s1 + s2 − 1), (s1 + s2 − 2), . . . , |s1 − s2 |

(Roughly speaking, the highest total spin occurs when the individual spins are aligned parallel to one
another, and the lowest occurs when they are antiparallel.)

5 Identical Particles
5.1 Two-Particle Systems
The state of a two-particle system is a function of the coordinates of particle one (r1 ), the coordinates
of particle two (r2 ), and the time:
Ψ(r1 , r2 , t).
Its time evolution is determined (as always) by the Schrödinger equation:
∂Ψ
ih̄ = HΨ
∂t
where H is the Hamiltonian for the whole system:

h̄2 2 h̄2 2
H=− ∇1 − ∇ + V (r1 , r2 , t)
2m1 2m2 2
The statistical interpretation carries over in the obvious way:
2
|Ψ(r1 , r2 , t)| d3 r1 d3 r2

is the probability of finding particle 1 in the volume d3 r1 and particle 2 in the volume d3 r2 ; evidently Ψ
must be normalized in such a way that
Z
2
|Ψ(r1 , r2 , t)| d3 r1 d3 r2 = 1

5.1.1 Bosons and Fermions


In quantum mechanics particles are indistinguishable in principle. Because of this the wave function of
particle 1 that is in the (one-particle) state ψa (r), and particle 2 that is in the state ψb (r) isn’t simply a
product:
ψ(r1 , r2 ) 6= ψa (r1 )ψb (r2 ) (if particles are indistinguishable).
Because we can’t do this we need to construct a wave function that is noncommittal as to which particle
is in which state. There are actually two ways to do it:

ψ± (r1 , r2 ) = A[ψa (r1 )ψb (r2 ) ± ψb (r1 )ψa (r2 )]

23
5.1 Two-Particle Systems Quantum Physics 1, Summary by Ellis de Wit

Thus the theory admits two kinds of identical particles: bosons, for which we use the plus sign, and
fermions, for which we use the minus sign. It so happens that
(
all particles with integer spin are bosons, and
all particles with half integer spin are fermions.

This connection between spin and statistics can be proved in relativistic quantum mechanics; in the
nonrelativistic theory it is taken as an axiom.
The famous Pauli exclusion principle states that two or more identical fermions cannot occupy the
same state;
ψ− (r1 , r2 ) = A[ψa (r1 )ψa (r2 ) − ψa (r2 )ψa (r1 )] = 0.
This doesn’t go for bosons, because there we add the products of the states together.
There is a more general (and more sophisticated) way to formulate this problem. Let us define the
exchange operator, P , which interchanges the two particles:

P f (r1 , r2 ) = f (r2 , r1 )

Clearly, P 2 = 1, and it follows that the eigenvalues of P are ±1. Now, if the two particles are identical,
the Hamiltonian must treat them the same: m1 = m2 and V (r1 , r2 ) = V (r2 , r1 ). It follows that P and
H are compatible observables and hence we can find a complete set of functions that are simultaneous
eigenstates of both. That is to say, we can find solutions to the Schrödinger equation that are either
symmetric (eigenvalue +1) or antisymmetric (eigenvalue -1) under exchange:

ψ(r1 , r2 ) = ±ψ(r2 , r1 ).

For identical particles the wave function is required to satisfy this symmetrization requirement.
Bosons will get the plus sign and fermions the minus sign.

5.1.2 Exchange Forces


Next we’re gonna look at a simple one-dimensional example, to get some sense of what the symmetrization
requirement actually does.
Suppose one particle is in state ψa (x), and the other is in state ψb (x), and these two states are orthogonal
and normalized. If the two particles are distinguishable, then the combined wave function is

ψ(x1 , x2 ) = ψa (x1 )ψb (x2 ),

for identical bosons we get

ψ+ (x1 , x2 ) = A[ψa (x1 )ψb (x2 ) + ψb (x2 )ψa (x1 )],

and for identical fermions we get

ψ− (x1 , x2 ) = A[ψa (x1 )ψb (x2 ) − ψb (x2 )ψa (x1 )].

Now let’s calculate the expectation value of the square of the separation distance between the two
particles,
(x1 − x2 )2 = x21 + x22 − 2 hx1 x2 i

Case 1: Distinguishable particles. For distinguishable particles the expectation value of the square
of the separation distance between the two particles becomes

(x1 − x2 )2 d
= x2 a
+ x2 b
− 2 hxia hxib .

Case 2: Identical particles. For identical particles the expectation value of the square of the sepa-
ration distance between the two particles becomes
2
(x1 − x2 )2 ±
= x2 a
+ x2 b
− 2 hxia hxib ∓ 2|hxiab | .

24
5.2 Atoms Quantum Physics 1, Summary by Ellis de Wit

It is clear that the only difference resides in the final term:


2
(x1 − x2 )2 ±
= (x1 − x2 )2 d
∓ 2|hxiab | .

Identical bosons (the upper signs) tend to be somewhat closer together, and identical fermions (the lower
signs) somewhat farther apart, than distinguishable particles in the same two states. Notice that this
only makes a difference when the two wave functions actually overlap, if the wavefunctions don’t overlap
hxiab vanishes. As a practical matter, therefore, it’s okay to pretend that electrons with nonoverlapping
wave functions are distinguishable.
The interesting case is when there is some overlap of the wave functions. The system behaves as though
there were a “force of attraction” between identical bosons, pulling them closer together, and a “force of
repulsion” between identical fermions, pushing them apart. We call it an exchange force.
So far we’ve been ignoring spin. The complete state of the electron includes not only its position wave
function, but also a spinor, describing the orientation of its spin:

ψ(r)χ(s).

When we put together the two-electron state, it is the whole works, not just the spatial part, that has
to be antisymmetric with respect to exchange. Now, a glance at the composite spin states reveals that
the singlet combination is antisymmetric (and hence would have to be joined with a symmetric spatial
function), whereas the three triplet states are all symmetric (and would require an antisymmetric spatial
function).

5.2 Atoms
A neutral atom, of atomic number Z, consists of a heavy nucleus, with electric charge Ze, surrounded
by Z electrons (mass m and charge −e). The Hamiltonian for this system is
Z  Z
h̄2 2 Ze2 e2
    X
X 1 1 1
H= − ∇ − + .
j=1
2m j 4π0 rj 2 4π0 |rj − rk |
j6=k

The term in curly brackets represents the kinetic plus potential energy of the jth electron, in the electric
field of the nucleus; the second sum (which runs over all values of j and k except j = k) is the potential
energy associated with the mutual repulsion of the electrons (the factor of 1/2 in front corrects for the
fact that the summation counts each pair twice). The problem is to solve Schrödinger’s equation for the
wave function ψ(r1 , r2 , . . . , rZ ). Because electrons are identical fermions, however, not all solutions are
acceptable: only those for which the complete state (position and spin),

ψ(r1 , r2 , . . . , rZ )χ(s1 , s2 , . . . , sZ ),

is antisymmetric with respect to interchange of any two electrons. In particular, no two electrons can
occupy the same state.

5.2.1 Helium
After hydrogen, the simplest atom is helium (Z = 2). The Hamiltonian,

h̄2 2 1 2e2 h̄2 2 1 2e2 e2


   
1
H= − ∇1 − + − ∇2 − +
2m 4π0 r1 2m 4π0 r2 4π0 |r1 − r2 |

consists of two hydrogenic Hamiltonians (with nuclear charge 2e), one for electron 1 and one for electron
2, together with a final term describing the repulsion of the two electrons. If we ignore this last term we
can separate the Schrödinger equation, and the solutions can be written as products of hydrogen wave
functions:
ψ(r1 , r2 ) = ψlnm (r1 )ψl0 n0 m0 (r2 ).
This will have only half the Bohr radius, and four times the Bohr energies. The total energy would be
−13.6
E = 4(En En0 ), with En = eV.
n2

25
Quantum Physics 1, Summary by Ellis de Wit

In particular, the ground state would be


8 −2(r1 +r2 )/a)
ψ0 (r1 , r2 ) = ψ100 (r1 )ψ100 (r2 ) = e , with E0 = −109 eV.
πa3
ψ0 is a symmetric function, so the spin state has to be antisymmetric, so it should be a singlet configu-
ration. The actual ground state of helium is indeed a singlet, but the experimentally determined energy
is -78.975 eV, so ignoring the repulsion of the two electrons didn’t give a good value in that regard.
The excited states of helium consist of one electron in the hydrogenic ground state, and the other in an
excited state:
ψnlm ψ100
We can construct both symmetric and antisymmetric combinations, in the usual way; the former go with
the antisymmetric spin configuration (the singlet), and they are called parahelium, while the latter
require a symmetric spin configuration (the triplet), and they are known as orthohelium. The ground
state is necessarily parahelium; the excited states come in both forms.

5.2.2 The Periodic Table


To first approximation (ignoring their mutual repulsion altogether), individual electrons occupy one-
particle hydrogenic states (n, l, m), called orbitals. Only two electrons can occupy any given orbital
(one with spin up, and one with spin down—or). There are n2 hydrogenic wave functions (all with the
same energy En ) for a given value of n, so the n = 1 shell holds 2 electrons, the n = 2 shell holds 8, n = 3
takes 18, and in general the nth shell can accommodate 2n2 electrons. Qualitatively, the horizontal rows
on the Periodic Table correspond to filling out each shell, but the electron-electron repulsion throws
this counting off.
The state of a particular electron is represented by the pair nl, with n (the number) giving the shell, and
l (the letter) specifying the orbital angular momentum; the magnetic quantum number m is not listed,
but an exponent is used to indicate the number of electrons that occupy the state in question. This is
shown in the following table.

l
n s p d f g

1 2 - - - -
2 2 6 - - -
3 2 6 10 - -
4 2 6 10 14 -

The numbers in the table indicate the total number of atoms that each shell can hold and the arrows
indicate the order in which the shells get filled, so the 1s shell gets filled first, then 2s, 2p, 3s and so on.
Another way to express the state of an atom is using the following notation:
2S+1
LJ ,

with S as the total spin, L as the total orbital angular momentum and J as the grand total (orbital plus
spin).

12 Afterword
In this section we look at some paradoxes that will conclude whether the realist viewpoint (did the
physical system ”actually have” the attribute in question prior to the measurement) or the orthodox
position (did the act of measurement itself ”create” the property, limited only by the statistical constraint
imposed by the wave function) is the right one.

26
12.1 EPR paradox Quantum Physics 1, Summary by Ellis de Wit

12.1 EPR paradox


The EPR paradox (or the Einstein-Podolsky-Rosen Paradox) is a thought experiment intended to
demonstrate an inherent paradox in the early formulations of quantum theory. It is among the best-
known examples of quantum entanglement.
Consider the decay of a neutral pi meson into an electron and a positron:

π 0 → e− + e+ .

Both sides will have spin 0, so the electron and positron are in the singlet configuration:
1
√ (↑− ↓+ − ↓− ↑+ )
2
The two particles are entangled. If the electron is found to have spin up, the positron must have spin
down and vice versa, no matter the distance if you measure spin up for the electron, you will immediately
know that the positron will have spin down. For the realists this is logical, the particles will have had
their spins from the moment they were created, quantum mechanics just didn’t know it. But for the
”orthodox” view this is a problem; your measurement of the electron collapsed the wave function, and
instantaneously ”produced” the spin of the positron 20 meters (or 20 light years) away. The fundamental
assumption on which the EPR argument rests is that no influence can propagate faster than the speed of
light. We call this the principle of locality, but the collapse of the wave function-whatever its ontological
status-is instantaneous.

12.2 Bell’s theorem


Bell suggested a generalization of the EPR/Bohm experiment: Instead of orienting the electron and
positron detectors along the same direction, he allowed them to be rotated independently. The first
measures the component of the electron spin in the direction of a unit vector a, and the second measures
the spin of the positron along the direction b. Bell proposed to calculate the average value of the product
of the spins, for a given set of detector orientations. Call this average P (a, b). For arbitrary orientations,
quantum mechanics predicts
P (a, b) = −a · b
What Bell discovered is that this result is incompatible with any local hidden variable theory.
The famous Bell inequality is given by

|P (a, b) − P (a, c)| ≤ 1 + P (b, c).

This holds for any local hidden variable theory, but it is easy to show that the quantum mechanical
prediction is incompatible with Bell’s inequality.
With Bell’s modification, then, the EPR paradox proves something far more radical than its authors
imagined: If they are right, then not only is quantum mechanics incomplete, it is downright wrong. On
the other hand, if quantum mechanics is right, then no hidden variable theory is going to rescue us
from the nonlocality. The results of the experiment were in excellent agreement with the predictions of
quantum mechanics, and clearly incompatible with Bell’s inequality. So nature itself is fundamentally
nonlocal. Causality is of course still a thing, causal influences cannot propagate faster than light, but
”ethereal” influences (like the entanglement) seemingly can.

12.3 No-clone theorem


You don’t need to know this section for the exam
Quantum measurements are typically destructive, in the sense that they alter the state of the system
measured. This is how the uncertainty principle is enforced in the laboratory. You might wonder why
we don’t just make a bunch of identical copies (clones) of the original state, and measure them, leaving
the system itself unscathed. It can’t be done. Indeed, if you could build a cloning device, quantum
mechanics would be out the window. For example you would be able to send messages using the EPRB
experiment and we just figured out that is a no-no. You can make a machine to clone spin-up electrons
and spin-down electrons, but it’s going to fail for any nontrivial linear combinations.

27
12.4 Schrodinger’s cat Quantum Physics 1, Summary by Ellis de Wit

12.4 Schrodinger’s cat


You don’t need to know this section for the exam
A cat is penned up in a steel chamber, along with the following device (which must be secured against
direct interference by the cat): in a Geiger counter, there is a tiny bit of radioactive substance, so small,
that perhaps in the course of the hour one of the atoms decays, but also, with equal probability, perhaps
none; if it happens, the counter tube discharges and through a relay releases a hammer that shatters a
small flask of hydrocyanic acid. If one has left this entire system to itself for an hour, one would say
that the cat still lives if meanwhile no atom has decayed. The first atomic decay would have poisoned it.
The psi-function of the entire system would express this by having in it the living and dead cat mixed
or smeared out in equal parts.

1
ψ = √ (ψalive + ψdead )
2

12.5 Quantum Zeno paradox


You don’t need to know this section for the exam
In 1977 Misra and Sudarshan proposed what they called the quantum Zeno effect as a dramatic
experimental demonstration of the collapse of the wave function. Their idea was to take an unstable
system (an atom in an excited state, say), and subject it to repeated measurements. Each observation
collapses the wave function, resetting the clock, and it is possible by this means to delay indefinitely the
expected transition to the lower state. Proof shows that a continuously observed unstable system never
decays at all, but for this to work the measurements must be made extremely rapidly.
As it turns out, the experiment is impractical for spontaneous transitions, but it can be done using
induced transitions, and the results are in excellent agreement with the theoretical predictions. Unfor-
tunately, this experiment is not as compelling a confirmation of the collapse of the wave function as its
designers hoped; the observed effect can be accounted for in other ways.

28

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy