Quantum 1 Summary
Quantum 1 Summary
Ellis de Wit
’20/’21 Block Ia
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
2
And since σ 2 is clearly non negative this implies that x2 ≥ hxi
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
Ψ(x, t) = ψ(x)ϕ(t)
3
2.2 The Infinite Square Well Quantum Physics 1, Summary by Ellis de Wit
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.
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
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
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
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.
[x, p] = x · p − p · x = ih̄,
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:
a+ then is called the raising operator. The same goes for a− , the lowering operator:
Important to note is the fact that a ”lowest rung” (ψ0 ) of our ladder exists, such that
a− ψ0 = 0
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 .
with √
2mE
k≡± .
h̄
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:
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 −∞
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
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.
Remember that δ(x − a) would be a spike of area 1 at the point a. And that
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= .
h̄
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= .
h̄
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:
mα
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:
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)
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.
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.
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,
h̄
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.
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
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
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
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.
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
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
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.
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
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
and the corresponding bra is its Hermitian conjugate († ), which is the combination of the transpose
(> ) and the complex conjugate (∗ ):
h̄
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
x = r sin θ cos φ,
y = r sin θ sin φ,
z = r cos θ.
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:
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
17
4.2 The Hydrogen Atom Quantum Physics 1, Summary by Ellis de Wit
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|)!
For historical reasons, l is called the azimuthal quantum number, and m the magnetic quantum
number.
This is all we can do with the radial part of the wave function until a specific potential V (r) is provided.
e2 1
V (r) = −
4π0 r
18
4.2 The Hydrogen Atom Quantum Physics 1, Summary by Ellis de Wit
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
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
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
L = r × p.
with of course
4.3.1 Eigenvalues
Our three angular momentum operators sadly do not commute:
This makes them incompatible observables. According to the generalized uncertainty principle,
h̄
σLx σLx ≥ |hLz i|.
2
But introducing the square of the total angular momentum,
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.
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 :
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
where
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:
and p
S± |s mi = h̄ s(s + 1) − m(m ± 1) |s (m ± 1)i ,
with S± ≡ Sx ± iSy and
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
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
h̄
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
µ = γS,
H = −mu · B,
H = −γB · S.
22
Quantum Physics 1, Summary by Ellis de Wit
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 = ↓↓
(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
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.
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
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,
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
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
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.
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.
27
12.4 Schrodinger’s cat Quantum Physics 1, Summary by Ellis de Wit
1
ψ = √ (ψalive + ψdead )
2
28