The Kepler Problem (Part 5)

Azimuth 2025-07-23

In Part 4 we saw how the classical Kepler problem is connected to a particle moving on a sphere in 4-dimensional space, and how this illuminates the secret 4-dimensional rotation symmetry of the Kepler problem. There are various ways to quantize the Kepler problem and obtain a description of the hydrogen atom’s bound states as wavefunctions on a sphere in 4-dimensional space. Here I’ll take a less systematic but quicker approach.

I’ll start with wavefunctions on the sphere in 4-dimensional space and describe operators on these arising from rotational symmetries. Combining these with standard facts about the hydrogen atom—which I’ll take as known, rather than derive—we can set up a correspondence between such wavefunctions and bound states of a hydrogen atom.

Here’s how. Chemists use a basis of bound states of the hydrogen called |n, \ell, m \rangle where

n tells us the energy of the state, • ℓ tells us its total angular momentum, and • m tells us its angular momentum around a chosen axis, say the z axis.

For now we’re ignoring the electron’s spin: we’ll bring that in later. We’ll show that wavefunctions on the sphere in 4-dimensional space also have a nice basis of states called |n, \ell, m \rangle. This correspondence will let us show that the hydrogen atom has 4-dimensional rotation symmetry: all these rotations preserve the energy, which depends only on n.

So let’s get started!

We can identify the sphere in 4-dimensional space, S^3, with the group \text{SU}(2) of 2 \times 2 unitary matrices with determinant 1. This group acts on itself in three important ways, which combine to give the mathematics we need:

• left multiplication by g: g maps h to gh,

• right multiplication by g^{-1}: g maps h to hg^{-1},

• conjugation by g: g maps h to ghg^{-1}.

Left and right multiplication commute, so they combine to give a left action of \text{SU}(2) \times \text{SU}(2) on \text{SU}(2). We thus obtain a unitary representation R of \text{SU}(2) \times \text{SU}(2) on Hilbert space of wavefunctions on \text{SU}(2), called L^2(\text{SU}(2)), given by

(R(g_1, g_2) \psi)(g) = \psi(g_1^{-1} g g_2)

This is the key to everything in this game.

In what follows I’ll denote \text{SU}(2) as S^3 when we regard it as the unit sphere in \mathbb{R}^4, acted on by \text{SU}(2) \times \text{SU}(2) as above. So I’ll write L^2(S^3) instead of L^2(\text{SU}(2)), but they’re really the same thing.

The Peter–Weyl theorem lets us decompose L^2(S^3) into finite-dimensional irreducible unitary representations of \text{SU}(2) \times \text{SU}(2):

\displaystyle{ L^2(S^3) \cong \bigoplus_{j} V_j \otimes V_j }

Here j = 0, \frac{1}{2}, 1, \frac{3}{2}, \dots and V_j is the spin-j representation of \text{SU}(2), which is the irreducible unitary representation of dimension 2j + 1. If you don’t know the Peter–Weyl theorem, you should quit reading this article now and learn about that, because it’s a lot more important than anything I’m talking about! It’s a key tool throughout group representation theory and quantum mechanics.

The representation of \text{SU}(2) \times \text{SU}(2) on L^2(S^3) is generated by self-adjoint operators that correspond to familiar observables for bound states of the hydrogen atom! To see this, first note that the complexification of \mathfrak{su}(2) \oplus \mathfrak{su}(2) has self-adjoint elements

A_j = (\frac{1}{2} \sigma_j, 0), \qquad  B_j =  (0, \frac{1}{2} \sigma_j)

where \sigma_j are the Pauli matrices. We also use A_j and B_j to denote the corresponding self-adjoint operators on L^2(S^3). They obey the following commutation relations, which are the quantum-mechanical analogues of the Poisson brackets we saw last time:

\begin{array}{ccc}    [A_j, A_k] &=&  i\epsilon_{jk\ell} A_\ell \\ [2pt]    [B_j, B_k] &=&  i\epsilon_{jk\ell} B_\ell  \\ [2pt]   [A_j, B_k] &=& 0   \end{array}

Geometrically speaking, the skew-adjoint operators -iA_j acts on L^2(S^3) as differentiation by vector fields on S^3 \cong \text{SU}(2) that generate left translations, while the operators -iB_j act by differentiation by vector fields that generate right translations. We also use -iA_j and -iB_j to stand for these vector fields. Beware: left-invariant-vector fields generate right translations and vice versa, the vector fields -iA_j are right-invariant, while the -iB_j are left-invariant!

As well known, we have

\phi \in V_j \otimes V_j \implies  A^2 \phi = B^2 \phi = j(j+1) \phi

This implies that

A^2 = B^2

on all of L^2(S^3). In terms of these we can define an operator on L^2(S^3) that corresponds to the Hamiltonian for bound states of the hydrogen atom, namely

\displaystyle{ H_0 \;=\; - \frac{1}{8(A^2 + \frac{1}{4})} \;=\;- \frac{1}{8(B^2 + \frac{1}{4})} }

This is the quantum analogue of the classical Hamiltonian we saw last time! Later I’ll talk about the curious appearance of the number \frac{1}{4} here. It holds deep mysteries, but we can see why it’s needed if we know the energy levels of the hydrogen atom. On the subspace V_j \otimes V_j, the operator 4A^2 + 1 acts as multiplication by

4j(j+1) + 1 = 4j^2 + 4j + 1 = (2j + 1)^2.

In atomic physics it is traditional to work with the dimension of V_j,

n = 2j + 1

rather than j itself. Thus, we have

\displaystyle{ \phi \in V_j \otimes V_j \implies   H_0\phi = - \frac{1}{2n^2} \phi }

Great! These are precisely the usual energy eigenvalues for the hydrogen atom, expressed in units with

\hbar = e = 4 \pi \epsilon_0 = \mu = 1

where:

\hbar is Planck’s constant,

-e is the charge of the electron (yes, some idiot made the electron charge negative),

\epsilon_0 is the permittivity of the vacuum, and

\mu = m_e/(m_e + m_p) is the reduced mass of the electron.

So: we needed that \frac{1}{4} in the Hamiltonian H_0 to get the usual energy eigenvalues for hydrogen.

We have not yet brought in the action of \text{SU}(2) on S^3 by conjugation. This gives yet another important unitary representation of \text{SU}(2) on L^2(S^3). To conjugate by g we both left multiply by g and right multiply by g^{-1}. Thus, the conjugation representation of \text{SU}(2) on L^2(S^3) has self-adjoint generators

L_j = A_j + B_j

and these obey

[L_j, L_k] = i\epsilon_{jk\ell} L_\ell

As we saw last time for the classical Kepler problem, the operators L_j are the components of the angular momentum. Now we’re doing quantum mechanics, so this will be the angular momentum of a hydrogen atom—not counting the electron’s spin.

With respect to the conjugation representation, the subspace V_j \otimes V_j \subset L^2(S^3) decomposes according to the usual rules for a tensor product of two representations of \text{SU}(2):

\displaystyle{ V_j \otimes V_j \cong \bigoplus_{\ell = 0}^{2j} V_\ell }

Here \ell takes integer values going from 0 to 2j = n - 1. Thus, with respect to the conjugation representation, we can decompose L^2(S^3) into irreducible representations of \text{SU}(2), like this:

\displaystyle{ L^2(S^3) \cong \bigoplus_{n = 1}^\infty \bigoplus_{\ell = 0}^{n-1} V_\ell }

The summand V_\ell has a basis of eigenvectors for the z component of angular momentum, L_3, with eigenvalues taking integer values m ranging from -\ell to \ell. Thus L^2(S^3) has an orthonormal basis of states |n , \ell , m\rangle where:

n ranges over positive integers;

\ell ranges from 0 to n-1 in integer steps;

m ranges from -\ell to \ell in integer steps.

From the calculations thus far we have

\begin{array}{ccl}    A^2 |n, \ell, m \rangle &=& B^2 |n, \ell, m \rangle \; = \;   \frac{1}{4}(n^2-1) |n, \ell, m \rangle  \\  [8pt]  H_0 |n, \ell, m \rangle &=& \displaystyle{ - \frac{1}{2n^2} \, |n, \ell, m \rangle } \\ [8pt]      L^2 |n, \ell, m \rangle &=& \ell(\ell + 1) \, |n , \ell, m \rangle  \\ [3pt]      L_3 |n , \ell, m \rangle &=& m \, |n , \ell, m \rangle   \end{array}

The last three relations are familiar from work on the hydrogen atom. In this context

n is the ‘principal quantum number’,

\ell is the ‘azimuthal quantum number’,

m is the ‘magnetic quantum number’.

The operator H_0 corresponds to the Hamiltonian of the hydrogen atom, with the electron treated as spinless, and the states |n, \ell, m\rangle are a well-known basis of the hydrogen atom’s bound states.

So, we’ve gotten to hydrogen, starting from just the space of wavefunctions on the 3-sphere, heavily using the fact that this sphere can be seen as the group \text{SU}(2), which acts on itself in three ways. Pretty cool!

Next I’ll say more about that mysterious \frac{1}{4} in the formula for the hydrogen atom Hamiltonian. Then I’ll bring in spin.