Lecture Notes UC4
Lecture Notes UC4
Quantum Mechanics in 3D
UC4 contents:
Or
Schrödinger equation in 3D
Thus:
where:
The potential energy V and the wave function Ψ are now functions of r = (x, y, z) and t.
The probability of finding the particle in the infinitesimal volume d 3r = dx dy dz is
| Ψ(r, t) |2 d 3r, and the normalisation condition reads:
where the spatial wave function ψn satisfies the time-independent Schrödinger equation:
With the cn constants determined by the initial wave function, Ψ(r,0), in the usual way.
Notice that if the potential admits continuum states, then the sum becomes an integral.
Schrödinger equation in spherical coordinates:
Most of the applications we will encounter involve central potentials, for which V is a
function only of the distance from the origin, V(r) → V(r).
In that case it is natural to adopt spherical coordinates, (r, θ, ϕ) (Figure 4.1).
In spherical coordinates the Laplacian takes the form:
Schrödinger equation in spherical coordinates:
In spherical coordinates, then, the time-independent Schrödinger equation reads
Variable separation:
Schrödinger equation in spherical coordinates:
Multiplying by:
Separation of variables:
Dividing by:
Solution for Φ:
m must be an integer:
Solution for ϴ:
for m ≥ 0
and Pl(x) is the lth Legendre polynomial, defined by the Rodrigues formula:
Solution for ϴ:
For negative values of m:
Pℓ(x) is a polynomial (of degree ℓ) in x, and is even or odd according to the parity of ℓ.
Pℓm(x) is not, in general, a polynomial — if m is odd it carries a factor of (1 − x 2)0.5
ℓ must be a non-negative integer.
If m > ℓ, Pℓm(x) = 0.
For any given ℓ, then, there are (2ℓ + 1) possible values of m:
Solution for ϴ:
For negative values of m:
Normalisation condition:
Spherical Harmonics:
The Radial Equation
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).
Variable change:
It tends to throw the particle outward (away from the origin), just like the centrifugal
(pseudo-)force in classical mechanics.
mp = 1.67 × 10−27 kg
e = 1.60 × 10−19 c
me = 9.11 × 10−31 kg
1 e2
From Coulomb’s law: F = the potential in SI units is:
4πϵ0 r 2
We introduce:
(for large ρ)
blows up
(for small ρ)
The hydrogen atom
The Radial Wave Function
We introduce the new function v(ρ):
(for large ρ)
(for small ρ)
Therefore:
This recursion formula determines the coefficients, and hence the function v(ρ).
For large j (this corresponds to large ρ, where the higher powers dominate):
If this were the exact result, it blows up at large ρ (so it is not normalisable):
The hydrogen atom
The Radial Wave Function: the Bohr radius
which makes v(ρ) a polynomial of order (N − 1), with (therefore) N − 1 roots, and
hence the radial wave function has N − 1 nodes.
Let’s define:
Remember:
The spatial wave functions are labeled by three quantum numbers (n, l, and m):
Bohr formula:
The ground state (that is, the state of lowest energy) is the case n = 1:
The binding energy of hydrogen (the amount of energy we would have to impart to the
electron in its ground state in order to ionise the atom) is 13.6 eV.
Remember:
Thus, l = 0 and m = 0:
The hydrogen atom
Energy levels for hydrogen
For arbitrary n, the possible
values of l are:
where:
The stationary states of the hydrogen atom are labeled by three quantum numbers:
n, l, and m.
and (for n ≠ n′) from the fact that they are eigenfunctions of Ĥ with distinct eigenvalues.
The hydrogen atom
Normalised hydrogen wave functions:
They can be visualised via density plots, in which the brightness of the cloud is
2
proportional to ψ .
| |
The hydrogen atom
The quantum numbers n, ℓ, and m can be identified from the nodes of the wave function.
m counts the number of nodes of the real (or imaginary) part of the wave function in the ϕ
direction. These nodes are planes containing the z axis on which the real or imaginary
part of ψ vanishes.
ℓ − m gives the number of nodes in the θ direction. These are cones about the z axis on
which ψ vanishes.
The hydrogen atom
Or via surfaces of constant probability density:
The hydrogen atom
The spectrum of Hydrogen:
If we put a hydrogen atom into some stationary state ψnlm, it should stay there forever.
If we tickle it slightly (by collision with another atom, say, or by shining light on it), the
atom may undergo a transition to some other stationary state:
- 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.
Rydberg formula
where:
In the classical theory of central forces, energy and angular momentum are the fundamental
conserved quantities. The angular momentum of a particle (with respect to the origin) is
given by:
We need to obtain the eigenvalues and the eigenfunctions of the angular momentum
operators.
The hydrogen atom
Angular momentum: Eigenvalues
The operators Lx and Ly do not commute:
Remember:
Now,
We have:
Thus:
The hydrogen atom
Angular momentum: Ladder operator technique
There must also exist a “bottom rung”, fb , such that:
Remember:
Comparing with:
So the eigenvalues of Lz are ℏm, where m goes from −l to +l, in N integer steps.
where:
The uncertainty principle implies that we cannot know all three components of L.
Actually, there aren’t three components — a particle simply cannot have a determinate
angular momentum vector.
If Lz has a well-defined value, then Lx and Ly do not.
The hydrogen atom
Angular momentum: Eigenfunctions
We will see that f ℓm = Yℓm, i.e. the eigenfunctions of L 2 and Lz are the spherical harmonics.
Since:
Here:
The unit vectors θ ̂ and ϕ̂ can be resolved into their cartesian components:
The hydrogen atom
Angular momentum: Eigenfunctions
In components:
And:
The hydrogen atom
Angular momentum: Eigenfunctions
In particular:
Conclusion:
Spherical harmonics are the eigenfunctions of L 2 and Lz. When we solved the Schrödinger
equation by separation of variables, we were inadvertently constructing simultaneous
eigenfunctions of the three commuting operators H, L 2 and Lz:
The hydrogen atom
Angular momentum: Eigenfunctions
We can rewrite Schrödinger’s equation as follows:
The algebraic theory of angular momentum permits ℓ (and hence also m) to take on
half -integer values:
The electron (as far as we know) is a structureless point, and its spin angular momentum
cannot be decomposed into orbital angular momenta of constituent parts.
Suffice it to say that elementary particles carry intrinsic angular momentum (S) in
addition to their “extrinsic” angular momentum (L).
The hydrogen atom
Spin:
where S± ≡ Sx ± iSy. The eigenvectors are not spherical harmonics (they’re not
functions of θ and ϕ at all), and there is no reason to exclude the half-integer
values of s and m:
The hydrogen atom
Spin:
Every elementary particle has a specific and immutable value of s, which we call
the spin of that particular species:
By contrast, the orbital angular momentum quantum number ℓ (e.g. for an electron
in a hydrogen atom) can take on any (integer) value, and will change from one to
another when the system is perturbed.
s is fixed for any given particle, so the theory of spin is comparatively simple.
The hydrogen atom
Spin 1/2:
s = 1/2 is the spin of the particles that make up ordinary matter (protons, neutrons,
and electrons), as well as all quarks and all leptons.
Using these as basis vectors, the general state of a spin-1/2 particle can be represented by
a two-element column matrix (or spinor):
where:
The hydrogen atom
Spin 1/2:
The spin operators become matrices:
Conclusion:
The hydrogen atom
Spin 1/2:
Similarly:
For which:
Remember:
The hydrogen atom
Spin 1/2 (Pauli spin matrices).
The possible values for Sx are the same as those for Sz. The eigenspinors are obtained via:
For a particle in the state χ+, what is the z-component of a particle’s spin angular
momentum?
We can answer unambiguously: +ℏ/2.
What is the net spin, s, of the combination, and what is the z component, m?
Thus:
The net spin, s, is much more subtle. If you combine spin s1 with spin s2, what total
spins s can we get?
The answer is that you get every spin from (s1 + s2) down to (s1 − s2)
Example 1:
If you package together a particle of spin 3/2 with a particle of spin 2,
you could get a total spin of 7/2, 5/2, 3/2, or 1/2, depending on the configuration.
Example 2:
If a hydrogen atom is in the state ψnlm , the net angular momentum of the electron (spin
1 1
plus orbital) is ℓ + or ℓ −
2 2
If you now throw in spin of the proton, the atom’s total angular momentum quantum
number is ℓ + 1 or ℓ − 1.
Notice that ℓ can be achieved in two distinct ways, depending on whether the electron
1 1
alone is in the ℓ + configuration or the ℓ − configuration.
2 2
The hydrogen atom
Addition of Angular Momenta
The combined state | s m ⟩ with total spin s and z-component m will be some linear
combination of the composite states | s1 s2 m1 m2 ⟩:
Because the z components add, the only composite states that contribute are those for
which:
If two particles (of spin 2 and spin 1) are at rest in a box, and the total spin is 3, and its z
component is 0, then a measurement of Sz could return the value ℏ (with probability 1/5),
(1)
or 0 (with probability 3/5), or −ℏ (with probability 1/5). Notice that the probabilities add up
to 1 (the sum of the squares of any column on the Clebsch–Gordan table is 1).
The hydrogen atom
Clebsch-Gordan coefficients
Electron in a Magnetic Field
A spinning charged particle constitutes a magnetic dipole. Its magnetic dipole moment,
μ, is proportional to its spin angular momentum, S.
The gyromagnetic ratio of an object whose charge and mass are identically distributed is
gs q
, where gs is the spin g-factor, q is the charge and m is the mass.
2m
For reasons that are fully explained only in relativistic quantum theory, the gyromagnetic
e
ratio of the electron is (almost) exactly twice the classical value: γ = −
m
so the Hamiltonian matrix for a spinning charged particle, at rest in a magnetic field B, is:
Larmor precession
Imagine a particle of spin 1/2 at rest in a uniform magnetic field, which points in the z-
direction:
Since the Hamiltonian is time independent, the general solution to the time-dependent
Schrödinger equation,
where | a |2 + | b |2 = 1.
Larmor precession
With no essential loss of generality we can write a = cos(α/2) and b = sin(α/2),
where α is a fixed angle whose physical significance will appear in a moment.
Thus :
Similarly:
This force can be used to separate out particles with a particular spin orientation.
Imagine a beam of heavy neutral atoms, traveling in the y direction, which passes
through a region of static but inhomogeneous magnetic field (Figure 4.15)—say
where B0 is a strong uniform field and the constant α describes a small deviation from
homogeneity.
Stern-Gerlach experiment
The force on these atoms is:
But because of the Larmor precession about B0, Sx oscillates rapidly, and
averages to zero; the net force is in the z direction:
and the beam is deflected up or down, in proportion to the z component of the spin
angular momentum.
Classically we’d expect a smear (because Sz would not be quantised), but in fact the
beam splits into 2s + 1 separate streams, demonstrating the quantisation of
angular momentum.
In silver atoms all the inner electrons are paired, in such a way that their angular
momenta cancel. The net spin is simply that of the outermost— unpaired—electron, so
1
in this case s = , and the beam splits in two.)
2
Stern-Gerlach experiment
We tend casually to assume that the initial state of a system is known (the
Schrödinger equation tells us how it subsequently evolves)—but it is natural to
wonder how we get a system into a particular state in the first place.
Experiment 1:
When a second, identical, S-G apparatus is placed at the exit of the first apparatus, only
z+ is seen in the output of the second apparatus.
This result is expected since all neutrons at this point are expected to have z+ spin, as
only the z+ beam from the first apparatus entered the second apparatus
Source:
https://en.wikipedia.org/wiki/Stern%E2%80%93Gerlach_experiment
Stern-Gerlach experiments ?
Experiment 2:
The middle system shows what happens when a different S-G apparatus is placed at the
exit of the z+ beam resulting of the first apparatus, the second apparatus measuring the
deflection of the beams on the x axis instead of the z axis.
The second apparatus produces x+ and x- outputs. Now classically we would expect
to have one beam with the x characteristic oriented + and the z characteristic oriented +,
and another with the x characteristic oriented - and the z characteristic oriented +.
Stern-Gerlach experiments ?
Experiment 3:
The output of the third apparatus which measures the deflection on the z axis again
shows an output of z- as well as z+.
Given that the input to the second S-G apparatus consisted only of z+, it can be inferred
that a S-G apparatus must be altering the states of the particles that pass through it.
Source: https://www.youtube.com/watch?v=PH1FbkLVJU4