The quantum-mechanical calculation of the hydrogen atom is,
in my opinion, the single most important model in all of physics:
miraculously, it is possible to find closed-form solutions
for the wave function of an electron in a proton’s potential well.
The results are highly educational, and also qualitatively
tell us a lot about all other chemical elements.
We start from the time-independent Schrödinger equation,
where μ is the reduced mass
of the electron-proton system,
and V is the proton’s Coulomb potential:
Now we have two simpler equations than the one we started with.
We multiply them by R and Y respectively,
and define C≡ℓ(ℓ+1) to help us later
(ℓ is unknown for now).
The results are the radial equation
and the angular equation:
Note that this calculation has not really been specific to hydrogen so far:
it is applicable to all spherically symmetric quantum systems.
Angular equation
Let us keep this generality, by keeping V unspecified for now,
In that case, the radial equation cannot be solved yet,
but the angular one can. We separate the variables again:
Y(θ,φ)=Θ(θ)Φ(φ)
Insert this into the equation and multiply by
sin2θ/(ΘΦ) to get a clean separation:
sin2θΘΘ′′+sinθcosθΘΘ′+ℓ(ℓ+1)sin2θ=−ΦΦ′′
Each term depends on θ or φ but not both,
so there exists a constant m2 such that:
m2=sin2θΘΘ′′+sinθcosθΘΘ′+ℓ(ℓ+1)sin2θ=−ΦΦ′′
These are two distinct equations;
multiplying by Θ and Φ respectively yields:
m2Θ−m2Φ=sin2θΘ′′+sinθcosθΘ′+ℓ(ℓ+1)sin2θΘ=Φ′′
Clearly the latter is the simplest, so we start there.
It is an eigenvalue problem for m2,
but it looks like a harmonic oscillator equation,
so the solutions are easily found to be:
Φ(φ)=eimφ
Because the coordinate φ is only defined in the interval [0,2π],
we demand periodic boundary conditions Φ(0)=Φ(2mπ),
which tells us that m is an integer.
The other equation, for Θ, needs a bit more work.
We write it out like so:
0=dθ2d2Θ+sinθcosθdθdΘ+(ℓ(ℓ+1)−sin2θm2)Θ
And then perform a change of variables θ→ξ
where ξ≡cosθ, leading to:
This result can be recognized as
Legendre’s generalized equation,
a known eigenvalue problem for ℓ(ℓ+1),
which has solutions when ℓ is a non-negative integer.
Those solutions are called the associated Legendre polynomialsPℓm(x) of degree ℓ and order m.
For a given ℓ, there exist 2ℓ+1
such “polynomials” (they actually contain square roots too)
indexed by the integer m in the range [−ℓ,ℓ],
so e.g. for ℓ=2 there is m=−2,−1,0,1,2.
We now have:
Yℓm(θ,φ)∝Pℓm(cosθ)eimφ
We are still missing a constant factor,
found by imposing the normalization condition:
∫02π∫0π∣Yℓm∣2sinθdθdφ=1
Calculating the normalization constant (not shown here) leads to the
following definition of the so-called spherical harmonicsYℓm of degree ℓ and order m:
These are important functions:
the wave function of any spherically symmetric quantum system
is a superposition of Yℓm with r-dependent coefficients.
And, as befits a (component of a) wave function,
they form an orthonormal basis, specifically:
∫02π∫0πYℓmYℓ′m′sinθdθdφ=δℓℓ′δmm′
Radial equation
With the angular part solved, we now turn to the radial part.
Introducing u(r)=rR(r), such that the derivatives of R(r) become:
R′=r2ru′−uR′′=r3r2u′′−2ru′+2u
Inserting this into the radial equation
and cancelling some of the terms:
After multiplying by ℏ2/(2μr) and rearranging, this turns into:
Eu=−2μℏ2u′′+(V+2μℏ2r2ℓ(ℓ+1))u
Here it is useful to define an effective potentialVeff(r) as below.
Keep in mind that ℓ is known after solving the angular equation:
Veff≡V+2μℏ2r2ℓ(ℓ+1)
This yields a relation of the same form as the
time-independent Schrödinger equation, just with V replaced by
Veff. This is the “true” radial equation,
an eigenvalue problem for E:
Eu=−2μℏ2u′′+Veffu
Now, finally, we specialize for the hydrogen atom.
Coulomb’s law tells us the attractive force F(r)
between the electron and the proton,
which we integrate to find the potential energy V(r):
F(r)=4πε0r2q2⟹V(r)=−4πε0rq2
Where q<0 is the electron’s charge,
and ε0 is the permittivity of free space.
Note that V<0, so there is a natural distinction
between bound statesE<0
(where the electron is trapped in the proton’s well),
and scattering statesE>0
(where the electron is free).
The true radial equation, after dividing by E, is now given by:
Inserting this into the radial equation and dividing out all common factors gives:
0=4ρw′′+4(ℓ+1−ρ)w′+(ρ0−2(ℓ+1))w
Let us rearrange this to put it in a more suggestive form.
Keep in mind that w=w(2ρ):
0=(2ρ)w′′+((2ℓ+1)+1−(2ρ))w′+(2ρ0−ℓ−1)w
This can be recognized as Laguerre’s generalized equation,
a well-known eigenvalue problem for λ≡(ρ0/2−ℓ−1).
It has solutions when λ is a non-negative integer,
in other words for ρ0=2n with n=1,2,3,...,
which also tells us that ℓ cannot be larger than n−1.
Then the solutions are the so-called associated Laguerre polynomialsLn−ℓ−12ℓ+1(2ρ), therefore:
u(ρ)∝ρℓ+1e−ρLn−ℓ−12ℓ+1(2ρ)
We are still missing a constant factor,
found by imposing the normalization condition:
∫0∞R2r2dr=1
Calculating the normalization constant (not shown here)
leads to this radial solution Rnℓ(r):
Meanwhile, by isolating the definitions of κ and ρ0 for E,
we find the eigenenergies to be:
E=−2μℏ2κ2=−2μℏ2(2πε0ℏ2ρ0μq2)2
Since ρ0=2n, these allowed Bohr energiesEn
of the electron are as follows:
En=−n212ℏ2μ(4πε0q2)2
At this point, it is customary to also define
the reduced Bohr radiusa0∗, given by:
a0∗≡nκ1=μq24πε0ℏ2≈5.295×10−11m=0.5295A˚
The non-reduced Bohr radiusa0 simply uses
the electron’s raw mass me instead of μ.
Roughly speaking, a0∗ is the most probable electron-proton distance
after a measurement of the electron’s position while it is in its ground state.
This is often to used to write Rnℓ(r) as:
Multiplying the angular and radial parts together,
we thus arrive at the following expression
for the full wave function ψnℓm:
ψnℓm(r,θ,φ)=Rnℓ(r)Yℓm(θ,φ)
The indices n, ℓ, and m are the quantum numbers,
which describe the state of the electron.
There is also a fourth not shown here, the spin quantum number,
which is +1/2 or −1/2 for spin-up or spin-down electrons respectively.
The principal quantum numbern, often called the shell number,
gives the energy level (shell) of the electron,
because the other numbers do not appear in En’s formula.
Since En=E1/n2,
the energy differences decrease with increasing n,
so electrons in higher shells can be excited more easily
(i.e. they need less energy to get excited).
The azimuthal quantum numberℓ gives the subshell
of shell n in which the electron is located.
It takes integer values from 0 to n−1 inclusive,
with 0, 1, 2, and 3 respectively
also called the s, p, d, and f subshells.
The electron’s total angular momentum is given by ℏℓ(ℓ+1).
The magnetic quantum numberm splits the electrons in each subshell
into orbitals, and takes integer values from −ℓ to ℓ.
The z-component of the electron’s angular momentum is ℏm.
The total degeneracy of each energy level n
can be calculated as the sum of an arithmetic series,
and is found to be n2 excluding spin (or 2n2 with spin).
Unsurprisingly, all these wave functions form an orthonormal basis
(although not a complete one unless the scattering states with E>0 are included):
When an excited electron drops from a state with energy Ei
to a lower level Ef, it emits a photon with energy ℏω,
where ω is the angular frequency of the resulting
electromagnetic wave:
ℏω=Ei−Ef=E1(ni21−nf21)
The corresponding vacuum wavelength is λ0=2πc/ω,
leading to the Rydberg formula,
which was discovered empirically before the hydrogen atom had been solved:
λ01=2πcω=RH(ni21−nf21)
Quantum mechanics then successfully gave a theoretical value
to the experimentally determined Rydberg constantRH (or R∞ if the raw electron mass me is used):
RH=2πℏc∣E1∣=4πℏ3cμ(4πε0q2)2≈1.097×107m−1
The transitions from excited states to the ground state nf=1
correspond to ultraviolet spectral lines known as the Lyman series.
Similarly, transitions to nf=2 give visible lines known as the Balmer series,
and transitions to nf=3 explain the Paschen series of infrared lines.
The Rydberg constant is not to be confused
with the Rydberg energyRy,
which is the ionization energy of ground-state hydrogen,
and is sometimes used as a unit in calculations:
Ry≡2πℏcRH=∣E1∣=2ℏ2μ(4πε0q2)2≈13.61eV
The point is that the hydrogen atom’s solution gave clear
explanations for known experimental data,
and settled the mystery of what an atom actually looks like.
While other elements’ atoms generally do not have such closed-form solutions
(because they have more than one electron),
their orbitals are qualitatively very similar.
In short, this model is the foundation of our modern understanding of atoms.