Quantum Mechanics
Quantum Mechanics
Quantum Mechanics
Michael Li
May 5, 2016
Physical background
Photoelectric effect. Electrons in atoms and line spectra. Particle diffraction. [1]
Hydrogen atom
Spherically symmetric wave functions for spherical well and hydrogen atom.
Orbital angular momentum operators. General solution of hydrogen atom. [5]
1
Contents
Contents 2
3 Examples 4
2
1 Schrödinger Equation and Wavefunctions
Biggest difference between classical and quantum physics: You can’t be sure of any-
thing at any time. Everything is formulated in probability. The probability of a par-
Rb
ticle being in [a, b] is a |ψ(x)|2 dx, where ψ(x) is its wavefunction. Here is a non-
exhaustive list of basic principles that are good to know:
- To calculate the probability, we need to normalize the wavefunctions so that
the square of it is actually a probability distribution. But we wouldn’t do it
otherwise because it just involves fiddling with constants. However, we would
like to work for normalizable wavefunctions, or those that satisfy:
Z ∞
|ψ(x)|2 dx < ∞.
−∞
3
We have the following proposition:
Proposition. The probability density P (x, t) = |Ψ(x, t)|2 obeys
∂P ∂j
=− ,
∂t ∂x
∗
i~
where j(x, t) = − 2m Ψ∗ dΨ
dx − dΨ
dx Ψ is the probability current.
Proof.
∂P ∂Ψ ∂Ψ∗ i~ 00 i~ 00∗ ∂j
= Ψ∗ + Ψ = Ψ∗ Ψ − Ψ Ψ=− .
∂t ∂t ∂t 2m 2m ∂x
2.1 Parity
Consider the Schrödinger equation for a particle of mass m
~2 00
Hψ = − ψ + V (x)ψ = Eψ.
2m
with potential V (x) = V (−x). By changing variables x → −x, we see that ψ(x) is
an eigenfunction of H with energy E if and only if ψ(−x) is an eigenfunction of H
with energy E. There are two possibilities:
(i) If ψ(x) and ψ(−x) is the same quantum state, this can only happen if ψ(−x) =
ηψ(x) for some constant η. Then we get that η = ±1 and ψ(−x) = ±ψ(x).
We call η the parity, and say ψ has even/odd parity if η is +1/ − 1 respectively.
(ii) If ψ(x) and ψ(−x) are different quantum states, then we take linear combi-
nations ψ± (x) = α(ψ(x) ± ψ(−x)), and these are eigenstates with energy
eigenvalue E, where α is for normalization. Then by construction, ψ± (−x) =
±ψ± (x) and have parity η = ±1.
Hence, if we are given a potential with reflective symmetry V (−x) = V (x), then we
can restrict our attention and just look for solutions with ±1 parity. Now we go for
examples.
3 Examples
Infinite Well The simplest case to consider is the infinite well. Here the potential is infinite
outside the region [−a, a], so for |x| > a, we must have ψ(x) = 0, or else
V (x)ψ(x) would be infinite.
4
V
x
−a a
We require ψ = 0 for |x| > a and ψ continuous at x = ±a. Within |x| < a,
the Schrödinger equation is
~2 00
− ψ = Eψ.
2m
We simplify this to become
ψ 00 + k 2 ψ = 0,
2 2
k
where E = ~2m . Here, instead of working with the complex exponentials, we
use sin and cos since we know well when these vanish. The general solution is
thus
ψ = A cos kx + B sin kx.
Now we solve for the boundary conditions, so that ψ vanishes at ±a. This turns
out to give
~2 π 2 2
En = n ,
8ma2
where n = 1, 2, · · · , and the wavefunctions are
12 (
1 cos nπx
2a n odd
ψn (x) = .
a sin nπx
2a n even
If you stare at this long enough, you can realize they are just standing waves.
V
x
a −U
−a
The potential is given by some constant U > 0. We want to seek energy levels
for a particle of mass m, defined by the Schrödinger equation
~2 00
Hψ = − ψ + V (x)ψ = Eψ.
2m
For energies in the range −U < E < 0, we set
~2 k 2 ~2 κ2
U +E = > 0, E=− ,
2m 2m
5
where k, κ > 0 are new real constants. Note they must satisfy
2mU
k 2 + κ2 = .
~2
Using these constants, the Schrödinger equation becomes
(
ψ 00 + k 2 ψ = 0 |x| < a
ψ 00 − κ2 ψ = 0 |x| > a.
We first consider the even parity solutions ψ(−x) = ψ(x). We can write our
solution as (
A cos kx |x| < a
ψ=
Be−κ|x| |x| > a
We now match ψ and ψ 0 at x = a to make it continuous. So we need
A cos ka = Be−κa
−Ak sin ka = −κBe−κa .
By parity, there is no additional information from x = −a. We can divide the
equations to obtain
k tan ka = κ.
this is still not something we can solve easily. To find when solutions exist, it
is convenient to introduce
ξ = ka, η = κa,
Hence the solution we need are solutions to
2ma2 U
η = ξ tan ξ. and ξ 2 + η 2 = .
~2
Where the second equation comes from our initial relation. We can look for
solutions by plotting these two equations.
6
V
x
Harmonic Oscillator
This is a harmonic oscillator of mass m with
1
V (x) = mω 2 x2 .
2
Note the similarity with this to the second term in a taylor expansion. And
assuming x0 is an equilibrium, the first term is zero so this is the first non-
zero approximation term! That’s why people say physics is just increasingly
complicated harmonic oscillators.
We seek all normalizable solutions to to the time-independent Schrödinger equa-
tion
~2 00 1
Hψ = − ψ + mω 2 x2 ψ = Eψ.
2m 2
12
So simplify constants, we define y = mω ~ x and E = 2E
~ω both of which is
dimensionless. Then we are left with
d2 ψ
− + y 2 ψ = Eψ.
dy 2
Now we try
1 2
ψ = f (y)e− 2 y .
Why? Well, we want to offset the y 2 factor in the differential equation here.
Then the Schrödinger equation gives
d2 f df
2
− 2y + (E − 1) = 0.
dy dy
ar y r ,
P
This is known as Hermite’s equation. We try a series solution f (y) = r≥0
and substitute in to get
X
(r + 2)(r + 1)an+2 + (E − 1 − 2r)ar y r = 0.
r≥0
2r + 1 − E
ar+2 = ar , r ≥ 0.
(r + 2)(r + 1)
7
2
So we are getting something like a series expansion of ey , which is bad, as ψ
would not be normalizable. So we need to terminate the series. This occurs iff
E = 2n + 1 for some n. Note that for each n, only one of the two independent
solutions is normalizable. So for each E, we get exactly one solution.
For n even, we have
2r − 2n
ar+2 = ar
(r + 2)(r + 1)
for r even, and ar = 0 for r odd, and the other way round when n is odd.
The solutions are thus f (y) = hn (y), where hn is a polynomial of degree n
with hn (−y) = (−1)n hn (y). For example, we have
h1 (y) = a1 y h2 (y) = a0 (1 − 2y 2 )
With constants fixed by normalization. These are known as the Hermite poly-
nomials. Thus, we have:
1
En = ~ω n + ,
2
for n = 0, 1, 2, · · · .
The wavefunctions are
mω 12
1 mω 2
ψn (x) = hn x exp − x ,
~ 2 ~
4.1 Definitions
First up, notations and some definitions:
Definition (Inner product). Let ψ(x) and φ(x) be normalizable wavefunctions at
some fixed time. We define the complex inner product by
Z ∞
(φ, ψ) = φ(x)∗ ψ(x) dx.
−∞
8
Definition (Uncertainty, or Variance). The uncertainty in position (∆x)ψ and mo-
mentum (∆p)ψ are defined by
(∆x)2ψ = h(x̂ − hx̂iψ )2 iψ = hx̂2 iψ − hx̂i2ψ ,
with exactly the same expression for momentum:
(∆p)2ψ = h(p̂ − hp̂iψ )2 iψ = hp̂2 iψ − hp̂i2ψ ,
Definition (Hermitian operator). An operator Q is Hermitian iff for all normalizable
φ, ψ, we have
Z Z
(φ, Qψ) = (Qφ, ψ) ⇔ φ∗ Qψ dx = (Qφ)∗ ψ dx.
x̂, p̂ and H are all Hermitian (From definition, just use integration by parts).
1 1 1
=− (Ψ, H(x̂Ψ)) + (Ψ, x̂(HΨ)) = (Ψ, (x̂H − H x̂)Ψ).
i~ i~ i~
But we know
~2 ~2 i~
x̂H − H x̂Ψ = − (xΨ00 − (xΨ)00 ) + (xV Ψ − V xΨ) = − Ψ0 = p̂Ψ.
2m m m
So done.The second part is similar. So we have:
d 1
hp̂iΨ = (Ψ̇, p̂Ψ) + (Ψ, p̂Ψ̇) = (Ψ, (p̂H − H p̂)Ψ)
dt i~
Then of course, we have
−~2
(p̂H − H p̂)Ψ = −i~ ((Ψ00 )0 − (Ψ0 )00 ) − i~((V (x)Ψ)0 − V (x)Ψ0 )
2m
= −i~V 0 (x)Ψ.
So done.
9
Theorem (Heisenberg’s uncertainty principle). If ψ is any normalized state (at any
fixed time), then
~
(∆x)ψ (∆p)ψ ≥ .
2
Yes, this is the famed theorem that we cannot know anything exactly. To prove
this we need the following:
Definition (Commutator). Let Q and S be operators. Then the commutator is de-
noted and defined by
[Q, S] = QS − SQ.
This is a measure of the lack of commutativity of the two operators.
In particular, the commutator of position and momentum is
[x̂, p̂] = x̂p̂ − p̂x̂ = i~.
Now we prove the Uncertainty principle:
Proof. Choose α = hx̂iψ and β = hp̂iψ , and define
X = x̂ − α, P = p̂ − β.
Then we have
(∆x)2ψ = (ψ, X 2 ψ) = (Xψ, Xψ) = kXψk2
(∆p)2ψ = (ψ, P 2 ψ) = (P ψ, P ψ) = kP ψk2
Then we have
(∆x)ψ (∆p)ψ = kXψkkP ψk
≥ |(Xψ , PΨ )|
≥ | Im(Xψ , PΨ )|
h i
1
≥ (Xψ, P ψ) − (P ψ, Xψ)
2i
h i 1
1 ~ ~
= (ψ, XP ψ) − (ψ, P Xψ) = (ψ, [X, P ]ψ) = (ψ, ψ) = .
2i 2i 2 2
10
5.2 Scattering
Consider the time-dependent Schrödinger equation with a potential barrier. We send
a wavepacket to a barrier, and in quantum mechanics, we would expect some part to
pass it, and some part to be reflected:
AΨref BΨtr
11
Potential Step We consider a wavepacket going into this:
V (x)
E < U We apply the standard method, introducing constants k, κ > 0 such that
~2 k 2 ~2 κ2
E= , U −E = .
2m 2m
Then the Schrödinger equation becomes
(
ψ 00 + k 2 ψ = 0 x < 0
ψ 00 − κ2 ψ = 0 x > 0
The solutions are ψ = Ieikx + Re−ikx for x < 0, and ψ = Ce−κx for
x > 0 (since ψ has to be bounded).
Since ψ and ψ 0 are continuous at x = 0, we have the equations
(
I +R=C
.
ikI − ikR = −κC
So we have
k − iκ 2k
R= I, C= I.
k + iκ k + iκ
If x < 0, ψ(x) is a superposition of beams (momentum eigenstates) with
|I|2 particles per unit length in the incident part, and |R|2 particles per
unit length in the reflected part, with p = ±~k. The current is
~k ~k
j = jinc + jref = |I|2 − |R|2 ,
m m
The probability of reflection is
|jref | |R|2
Pref = = = 1,
|jinc | |I|2
~2 k 2 ~2 κ2
E= , E−U = ,
2m 2m
We then match the two solutions again (skipping some steps to show you
the important part): (
I +R=T
ikI − ikR = ikT.
12
We can solve these to obtain
k−κ 2k
R= I, T = I.
k+κ k+κ
Our flux on the left is now
~k ~k
j = jinc + jref = |I|2 − |R|2 ,
m m
while the flux on the right is
~κ
j = jtr |T |2 .
m
The probability of reflection and transmission is
2
|R|2 |T |2 κ
|jref | k−κ |jtr | 4kκ
Pref = = 2
= , Ptr = = 2
= .
|jinc | |I| k+κ |jinc | |I| k (k + κ)2
x
0 a
Now consider a stationary state with energy E with 0 < E < U . We set the
constants
~2 k 2 ~2 κ2
E= , U −E = .
2m 2m
Then the Schrödinger equations become
ψ 00 + k 2 ψ = 0 x<0
00 2
ψ −κ ψ =0 0<x<a
00 2
ψ +k ψ =0 x>a
So we get
I +R=A+B
ik(I − R) = κ(A − B)
Ae κa
+ Be−κa = T eika
κ(Aeκa − Be−κa ) = ikT eika .
13
We can work out the boring algebra to obtain:
−1
k 2 − κ2
−ika
T = Ie cosh κa − i sinh κa
2kκ
Therefore we can use these to find the transmission probability, and it turns out
to be −1
|T |2 U2
|jtrj | 2
Ptr = = = 1 + sinh κa .
|jinc | |I|2 4E(U − E)
This demonstrates quantum tunneling. There is a non-zero probability that the
particles can pass through the potential barrier even though it classically does
not have enough energy.
~2 00
− ψ = Eψ.
2m
Which give solutions like:
( 2 2
Aeikx + Be−ikx k
E = ~2m >0
ψ∼ ~2 κ 2
Aeκx + Be−κx E = − 2m < 0.
Which is overdetermined. That’s why bound state energy levels are quantized.
If it is not bounded, then we have:
(
Ieikx + Re−ikx x → −∞
ψ∼
T eikx x → +∞
So we have:
|jref |
Pref = |Aref |2 =
|jinc |
|jtr |
Ptr = |Atr |2 = ,
|jinc |
R T
where Aref (k) = I and Atr (k) = I .
Note. In quantum mechanics, Amplitude squared generally gives probability.
14
6 Axioms for Quantum Mechanics
Note. This section is in the notes but honestly I fail to see how they can test this. But
still read through it. It will help a lot when (if) you do PQM next year.
Now we would list the axioms, and some useful results:
i~Ψ̇ = HΨ,
where H is a Hermitian operator, the Hamiltonian; this holds at all times except
at the instant a measurement is made. This is called the wavefunction collapse.
Why? Well………You just asked a million dollar question.
- For any observable Q, the number of linearly independent eigenstates with
eigenvalue λ is the degeneracy of the eigenvalue. In other words, the degen-
eracy is the dimension of the eigenspace
Vλ = {ψ : Qψ = λψ}.
15
Not an Axiom What if we wanted to distinguish degenerate states? Then we need two observ-
ables A and B, and the state needs to be an eigenstate for both of them. From
linear algebra, we can know that this is possible if A and B commutes.
Not an Axiom
Theorem (Ehrenfest’s theorem in General). If Q is any operator with no ex-
plicit time dependence, then
d
i~ hQiΨ = h[Q, H]iΨ ,
dt
where [Q, H] = QH − HQ is the commutator. The proof is done by expanding
out the definition and replacing −i~Ψ̇ with HΨ.
P
Not an Axiom For a state at time 0 written as Ψ(0) = n αn ψn , then using stationary states,
we can write the general form as:
X
Ψ(t) = αn e−iEn t/~ ψn .
n
Definition (Structureless particle). A structureless particle is one for which all ob-
servables can be written in terms of position and momentum.
In reality, this is mostly not true but we don’t care. The method of solving this is
simply by separation of variables where we look for solutions in one dimension and
multiply them together. I know. It is not exciting. Let’s get to the more exciting stuff:
L = x̂ ∧ p̂ = −i~x ∧ ∇.
∂
Li = εijk x̂j p̂k = −i~εijk xj .
∂xk
For example, we have
∂ ∂
L3 = x̂1 p̂2 − x̂2 p̂1 = −i~ x1 − x2 .
∂x2 ∂x1
16
Definition. The total angular momentum operator is
[Li , Lj ] = i~εijk Lk [Li , x̂j ] = i~εijk x̂k , [Li , p̂j ] = i~εijk p̂k
∂
L3 = −i~
∂φ
∂±iϕ ∂
L± = L1 ± iL2 = ±~e ± i cot θ
∂θ ∂φ
1 ∂2
2 2 1 ∂ ∂
L = −~ sin θ + .
sin θ ∂θ ∂θ sin2 θ ∂φ2
From above, we can see that we can measure L3 and L2 simultaneously, so there
are simultaneous eigenfunctions of these operators, which we call Y`m (θ, ϕ), with
` = 0, 1, 2, · · · and m = 0, ±1, ±2, · · · , ±`. These have eigenvalues ~m for L3 and
~2 `(` + 1) for L2 .
In spherical polars, we have:
~2 1 ∂ 2 1 1 2
H=− 2
r+ L + V (r).
2µ r ∂r 2µ r2
We can check that
[Li , H] = [L2 , H] = 0.
This implies we can use the eigenvalues of H, L2 and L3 to label our solutions to the
equation. Therefore, we can simultaneously measure them, so the joint eigenstates
must be:
ψ(x) = R(r)Y`m (θ, ϕ),
17
Now we then try to solve the Schrodinger equation, which gives:
~2 1 d2 ~2
− 2
(rR) + `(` + 1)R + V R = ER.
2µ r dr 2µr2
Now R(r) is called the radial part of the wavefunction, and we will work with χ(r) =
rR(r), often called the radial wavefunction. Then the equation turns into:
~2 00 ~2 `(` + 1)
− χ + χ + V χ = Eχ.
2µ 2µr2
The radial Schrödinger Equation. NOw since we would like R to be finite as r → 0, so
χ = 0 at r = 0. The normalization conditions turns into:
Z ∞
|R(r)|2 r2 dr < ∞.
0
V
a
r
−U
We now look for bound state solutions to the Schrödinger equation with −U < E <
0, with total angular momentum quantum number `.
For r < a, our radial wavefunction χ obeys
`(` + 1)
χ00 − χ + k 2 χ = 0,
r2
where k is a new constant obeying
~2 k 2
U +E = .
2µ
For r ≥ a, we have
`(` + 1)
χ00 − χ − κ2 χ = 0,
r2
with κ obeying
~2 κ2
E=− .
2µ
We can solve in each region and match χ, χ0 at r = a, with boundary condition
χ(0) = 0. Note that given this boundary condition, solving this is equivalent to
solving it for the whole R but requiring the solution to be odd.
18
Solving this for general ` is slightly complicated (would be taught in Applied
Quantum Mechanics course). So we shall look at some particular examples.
For ` = 0, we have no angular term, and we have done this before. The general
solution is (
A sin kr r < a
χ(r) =
Be−κr r>a
Matching the values at x = a determines the values of k, κ and hence E.
For ` = 1, it turns out the solution is just
(
1
A cos kr − kr sin kr r < a
χ(r) = 1
−κr .
B 1 + kr e r>a
8.1 Introduction
Consider an electron moving in a Coulomb potential
e2 1
V (r) = − .
4πε0 r
This potential is due to a proton stationary at r = 0. We follow results from the last
section of the last chapter, and set the mass µ = me , the electron mass. The joint
energy eigenstates of H, L2 and L3 are of the form
19
The goal of this chapter is to understand all the (normalizable) solutions to this
equation (∗).
Now we start by guessing. For large r, we get
R00 ∼ κ2 R.
When we substitute this in, we will get three kinds of terms. The r` e−κr terms match,
and so do the terms of the form r`−2 e−κr . Finally, we see the r`−1 e−κe terms match
if and only if
(` + 1)κ = λ.
Hence, for any integer n = ` + 1 = 1, 2, 3, · · · , there are bound states with energies
2
~2 λ2 e2
1 1
En = − = − me .
2me n2 2 4πε0 ~ n2
But this is only one solution as for each energy level, there are many possible
angular momentums.
20
Now each side of the equality is equidimensional, which is great for series solutions.
This equation is regular singular at r = 0, so we guess a solution of the form
∞
X
f (r) = ap rp+σ , a0 6= 0.
p=0
κn = λ
for some n ≥ ` + 1. So the resulting energy levels are exactly those we found before:
2
~2 2 ~2 λ2 e2
1 1
En = − κ =− = − me .
2me 2me n2 2 4πε0 ~ n2
for n = 1, 2, 3, · · · . This n is called the principle quantum number.
For any given n, the possible angular momentum quantum numbers are
` = 0, 1, 2, 3, · · · , n − 1
m = 0, ±1, ±2, · · · , ±`.
with
Rn`r = r` gn` (r)e−λr/n ,
where gn` (r) are (proportional to) the associated Lagnerre polynomials.
In general, the “shape” of probability distribution for any electron state depends on
r and θ, ϕ mostly through Y`m . For ` = 0, we have a spherically symmetric solutions
21
The degeneracy of each energy level En is
n−1
X `
X n−1
X
1= (2` + 1) = n2 .
`=0 m=−` `=0
So. We solved the hydrogen atom. Kind of. By guessing. But hey, we solved it.
22