The relativistic electron

In this chapter, we’ll be looking at the relativistic quantum mechanics for a single, free electron. We will merely introduce the important concepts and implications, instead of a fully formal relativistic treatment of the Dirac electron/field.

The Klein-Gordon equation

There is a fundamental flaw with the time-dependent Schrödinger equation \[\begin{equation} \require{physics} i \hbar \pdv{t} \Psi(\vb{r}, t) = h^c \Psi(\vb{r},t) % \thinspace , \end{equation}\] and that is that that it is not Lorentz covariant: time and space coordinates are not treated in the same way. We could however, use the squared relativistic energy for the electron: \[\begin{equation} E^2 = c^2 \norm{\vb{p}}^2 + m_e^2 c^4 \end{equation}\] and subsequently use the quantum mechanical correspondence principle \[\begin{align} & E \rightarrow i \hbar \pdv{t} \\ & \vb{p} \rightarrow -i \hbar \nabla % \label{eq:quantization_momentum} \thinspace , \end{align}\] we do end up with a Lorentz covariant equation: \[\begin{equation} \qty[ % \frac{1}{c^2} \pdv[2]{t} % - \laplacian % + \qty(\frac{m_e c}{\hbar})^2 % ] \Psi(\vb{r},t) = 0 % \thinspace , \end{equation}\] called the Klein-Gordon equation. In covariant form, it would read: \[\begin{equation} \qty[ % \partial_\mu \partial^\mu % + \qty(\frac{m_e c}{\hbar})^2 % ] \Psi(\vb{r},t) = 0 % \thinspace , \end{equation}\] from which it is evident (because only Lorentz scalars appear) that the Klein-Gordon equation is Lorentz covariant. The solutions to the Klein-Gording equation are the plane waves: \[\begin{equation} \Psi(\vb{r}, t) % = \exp( % \frac{i}{\hbar} (Et - \vb{p} \vdot \vb{r}) % ) \thinspace , \end{equation}\] characterized by the energy \[\begin{equation} E = \pm \sqrt{c^2 \norm{\vb{p}}^2 + m_e^2 c^4} \thinspace . \end{equation}\]

However, for electrons, there are a few problems with this Klein-Gordon equation. First of all, the occurrence of negative energies are hard to explain, second, it cannot explain the occurrence of the spin of an electron and third, the Klein-Gordon probability density (derived from the Klein-Gordon continuity equation) is not positive definite. This means that the Klein-Gordon was a nice first step to describe the relativistic electron, but there is better equation of motion for the free, relativistic electron.

The Dirac equation

The Dirac equation for a freely moving electron \[\begin{equation} i \hbar \pdv{t} \Psi(\vb{r},t) % = h^{c,\text{D}} \Psi(\vb{r},t) \end{equation}\] is based on the following form for the Hamiltonian: \[\begin{equation} \label{eq:Dirac_equation} h^{c,\text{D}} % = -i \hbar c \alpha^k \partial_k + \beta m_e c^2 % \thinspace , \end{equation}\] in which \(\alpha^k\) and \(\beta\) are yet unspecified objects. The components \(\alpha^i\) are not components of a contravariant vector, but they are introduced to employ Einstein’s summation convention. If we now require that \[\begin{equation} \label{eq:requirements_Dirac_matrices} \begin{cases} & \comm{\alpha^i}{\alpha^j}_+ = 2 \delta_{ij} \\ & \comm{\alpha^i}{\beta}_+ = 0 \\ & \beta^2 = 1 \end{cases} \thinspace , \end{equation}\] from which, by the way, follows that \[\begin{equation} (\alpha^i)^2 = 1 % \thinspace , \end{equation}\] we can verify that the Klein-Gordon equation is in fact embedded in the Dirac equation. This means that, given that the objects \(\alpha^k\) and \(\beta\) behave like these requirements, the Dirac equation is Lorentz covariant. Gauge invariance of the Dirac equation is discussed in subsection \(\eqref{sec:Dirac_electron_external_fields}\).

The standard representation

There are different ways of choosing the objects \(\boldsymbol{\alpha} = (\alpha_x, \alpha_y, \alpha_z)\) and \(\beta\) such that the requirements in equation \(\eqref{eq:requirements_Dirac_matrices}\) are fulfilled. One particular choice of representation is called the standard representation, in which the \(\alpha\)- and \(\beta\)-matrices are both \((4 \times 4)\)-matrices. In this representation, the \(\alpha\)-matrices are given by \[\begin{equation} \label{eq:Dirac_alpha_matrices} \alpha^k = % \begin{pmatrix} % 0 & \sigma_k \\ % \sigma_k & 0 % \end{pmatrix} \thinspace , \end{equation}\] in which \(\sigma_k\) are the Pauli spin matrices (cfr. equation \(\eqref{eq:Pauli_spin_matrices}\)). In the standard representation, the \((4 \times 4)\)-\(\beta\) matrix is given by: \[\begin{equation} \label{eq:Dirac_beta_matrix} \beta = % \begin{pmatrix} % I_2 & 0 \\ % 0 & -I_2 % \end{pmatrix} \thinspace . \end{equation}\] From the previous definitions, it automatically follows that all \(\alpha^k\) and \(\beta\)-matrices are Hermitian, which should be the case since the Dirac Hamiltonian must be Hermitian.

Since, in the standard representation, the objects \(\alpha^k\) and \(\beta\) are \(4 \times 4\)-matrices, the state vector \(\Psi(\vb{r}, t)\) must also be \(4\)-dimensional and is called a \(4\)-spinor or bispinor. Because of the block structure of the \(\alpha\)- and \(\beta\)-matrices, this 4-spinor is often split up into two 2-spinors: \[\begin{equation} \Psi(\vb{r}, t) = % \begin{pmatrix} % \Psi^{(1)}(\vb{r}, t) \\ \Psi^{(2)}(\vb{r}, t) \\ \Psi^{(3)}(\vb{r}, t) \\ \Psi^{(4)}(\vb{r}, t) % \end{pmatrix} % = % \begin{pmatrix} % \Psi^\text{L}(\vb{r}, t) \\ \Psi^\text{S}(\vb{r}, t) % \end{pmatrix} \thinspace , \end{equation}\] which are called the large and the small components, but more on that later.

If we introduce the \(k\)-th component of the momentum operator in coordinate representation as (cfr. the quantization of the momentum in equation \(\eqref{eq:quantization_momentum}\)): \[\begin{equation} p^c_k = -i \hbar \partial_k \end{equation}\] then we can write the Dirac equation in a compact form as \[\begin{equation} i \hbar \pdv{t} \Psi(\vb{r}, t) % = \qty( % c \boldsymbol{\alpha} \vdot \vb{p}^c + \beta m_e c^2 % ) \Psi(\vb{r}, t) % \thinspace . \end{equation}\] By subsequently introducing the small and large components, we can see that the Dirac equation admits a \((2 \times 2)\)-superstructure: \[\begin{equation} i \hbar \pdv{t} % \begin{pmatrix} % \Psi^\text{L}(\vb{r}, t) \\ \Psi^\text{S}(\vb{r}, t) % \end{pmatrix} = \begin{pmatrix} m_e c^2 & c \boldsymbol{\sigma} \vdot \vb{p}^c \\ c \boldsymbol{\sigma} \vdot \vb{p}^c & - m_e c^2 \end{pmatrix} % \begin{pmatrix} % \Psi^\text{L}(\vb{r}, t) \\ \Psi^\text{S}(\vb{r}, t) % \end{pmatrix} \thinspace , \end{equation}\] or in full \(4\)-component form: \[\begin{equation} i \hbar \pdv{t} % \begin{pmatrix} % \Psi^{(1)}(\vb{r}, t) \\ \Psi^{(2)}(\vb{r}, t) \\ \Psi^{(3)}(\vb{r}, t) \\ \Psi^{(4)}(\vb{r}, t) % \end{pmatrix} = \begin{pmatrix} m_e c^2 % & 0 % & p^c_z % & p^c_x - i p^c_y \\ 0 % & m_e c^2 % & p^c_x + i p^c_y % & -p_z \\ p^c_z % & p^c_x - i p^c_y % & -m_e c^2 % & 0 \\ p^c_x + i p^c_y % & -p_z % & 0 % & - m_e c^2 \end{pmatrix} \begin{pmatrix} % \Psi^{(1)}(\vb{r}, t) \\ \Psi^{(2)}(\vb{r}, t) \\ \Psi^{(3)}(\vb{r}, t) \\ \Psi^{(4)}(\vb{r}, t) % \end{pmatrix} \thinspace . \end{equation}\] in which we have used the at first sight strange ‘inner product’ of the Pauli sigma matrices with the momentum operator: \[\begin{equation} \boldsymbol{\sigma} \vdot \vb{p}^c % = \begin{pmatrix} % p^c_z & p^c_x -i p^c_y \\ p^c_x + i p^c_y & -p_z % \end{pmatrix} \thinspace . \end{equation}\] By the way, this operator can be equivalently written as: \[\begin{equation} \boldsymbol{\sigma} \vdot \vb{p}^c % = \frac{1}{r^2} ( % \boldsymbol{\sigma} \vdot \vb{r} % ) [ % \vb{r} \vdot \vb{p}^c % + i (\boldsymbol{\sigma} \vdot \vb{l}) % ] % \thinspace , \end{equation}\] in which the so-called spin-orbit coupling term is made explicit in the last term.

Other representations

We can write Dirac’s equation in many different forms. Let us start by introducing the Dirac matrices \(\gamma^\mu = (\gamma^0, \gamma^i)\): \[\begin{equation} \gamma^0 = \beta = % \begin{pmatrix} I_2 & 0 \\ 0 & -I_2 % \end{pmatrix} \thinspace , \end{equation}\] which is a Hermitian matrix, and \[\begin{equation} \gamma^i = \beta \alpha^i % = \begin{pmatrix} 0 & \sigma_k \\ -\sigma_k & 0 \end{pmatrix} \thinspace , \end{equation}\] which are anti-Hermitian matrices. They furthermore fulfill the anticommutation relations \[\begin{equation} \comm{\gamma^\mu}{\gamma^\nu}_+ = 2 \eta^{\mu \nu} % \label{eq:anticommutator_gamma_matrices} % \thinspace , \end{equation}\] which is said to define a Clifford algebra. It is then possible to obtain different representations of the Dirac equations by doing similarity transformations of the form \[\begin{equation} \tilde{\gamma}^\mu = M \gamma^\mu M^{-1} \end{equation}\] with \(M\) an arbitrary invertible matrix. The Dirac matrices \(\gamma^\mu\) are used in the Einstein summation convention, but are not strictly Lorentz 4-vectors. In fact, they are Lorentz scalars, as they have the same values in every frame of reference. Just by an abuse of notation, we can calculate the seemingly covariant components as \[\begin{equation} \gamma_\nu = \eta_{\nu \mu} \gamma^\mu % \thinspace . \end{equation}\] Using the Dirac matrices \(\gamma^\mu\), the Dirac equation can be written in covariant form: \[\begin{equation} \label{eq:Dirac_equation_gamma_matrices} \qty( % -i \hbar \gamma^\mu \partial_\mu % + m_e c % ) \Psi(x) = 0 % \thinspace . \end{equation}\] It can be shown that this equation is Lorentz covariant (i.e. it has the same form in two interial frames related through Lorentz transformations), which means that even though \(\gamma^\mu\) are not the components of a contravariant vector, the combination \(\gamma^\mu \partial_\mu \Psi\) is Lorentz covariant, so we can treat \(\gamma^\mu \partial_\mu\) as a Lorentz scalar.

If we take the Hermitian adjoint of this equation, and subsequently multiply by \(\gamma^0\) on the right, we obtain the Dirac adjoint equation: \[\begin{equation} i \partial_{\mu} \bar{\Psi}(x) \gamma^{\mu} % + \frac{m_e c}{\hbar} \bar{\Psi}(x) = 0 % \thinspace , \end{equation}\] in which we have defined the Dirac adjoint (Bransden and Joachain 2000) as \[\begin{equation} \bar{\Psi}(x) = \Psi^\dagger(x) \gamma^0 % \thinspace . \end{equation}\] By multiplying the Dirac equation on the left with \(\bar{\Psi}\) and the Dirac adjoint equation on the right with \(\Psi\), we find a continuity equation: \[\begin{equation} \partial_\mu j^\mu(x) = 0 % \thinspace , \end{equation}\] if we define the \(4\)-current as: \[\begin{equation} j^\mu(x) = \bar{\Psi}(x) \gamma^\mu \Psi(x) % \thinspace . \end{equation}\] In terms of the Dirac \(\alpha\) and \(\beta\)-matrices, the time component gives the Dirac distribution: \[\begin{equation} \frac{j^0(x)}{c} = \rho(\vb{r}, t) % = \Psi^\dagger(\vb{r}, t) \Psi(\vb{r}, t) \end{equation}\] and the spatial components give the Dirac 3-current density: \[\begin{equation} \vb{j}(\vb{r}, t) = % c \Psi^\dagger(\vb{r}, t) % \boldsymbol{\alpha} % \Psi(\vb{r}, t) % \thinspace , \end{equation}\] which are related through the continuity equation: \[\begin{equation} \pdv{ % \rho(\vb{r}, t) % }{t} % + \boldsymbol{\nabla} \vdot \vb{j}(\vb{r}, t) = 0 % \thinspace . \end{equation}\]

The Dirac distribution and current density can be written in spinor (two-component) and bispinor (four-component) form as follows (G. Dyall and Faegri 2007): \[\begin{align} \rho(\vb{r}, t) &= \Psi^\dagger(\vb{r}, t) \Psi(\vb{r}, t) % \label{eq:Dirac_distribution} \\ &= \Psi^{\text{L} \dagger}(\vb{r}, t) \Psi^{\text{L}}(\vb{r}, t) % + \Psi^{\text{S} \dagger}(\vb{r}, t) \Psi^{\text{S}}(\vb{r}, t) \\ &= \sum_u^4 \Psi^{(u) \dagger}(\vb{r}, t) \Psi^{(u)}(\vb{r}, t) \end{align}\] and \[\begin{align} \vb{j}(\vb{r}, t) % &= c \Psi^\dagger(\vb{r}, t) % \boldsymbol{\alpha} % \Psi(\vb{r}, t) \\ &= c % \qty( % \Psi^{\text{L} \dagger}(\vb{r}, t) % \boldsymbol{\sigma} % \Psi^{\text{S}}(\vb{r}, t) % + \Psi^{\text{S} \dagger}(\vb{r}, t) % \boldsymbol{\sigma} % \Psi^{\text{L}}(\vb{r}, t) % ) \\ &= c \begin{pmatrix} \Psi^{(1)*} \Psi^{(4)} % + \Psi^{(2)*} \Psi^{(3)} % + \Psi^{(3)*} \Psi^{(2)} % + \Psi^{(4)*} \Psi^{(1)} \\ % -i \Psi^{(1)*} \Psi^{(4)} % + i \Psi^{(2)*} \Psi^{(3)} % - i \Psi^{(3)*} \Psi^{(2)} % + i \Psi^{(4)*} \Psi^{(1)} \\ % \Psi^{(1)*} \Psi^{(3)} % - \Psi^{(2)*} \Psi^{(4)} % + \Psi^{(3)*} \Psi^{(1)} % - \Psi^{(4)*} \Psi^{(2)} \end{pmatrix} \thinspace , \end{align}\] in which we have not written the explicit time- and position-dependence of the components of the bispinor in the last equation.

Negative-energy states

Let us now turn to the solution of the Dirac equation. For a particle at rest, the kinetic contributions in the Dirac equation vanish, so we are left with \[\begin{equation} i \hbar \pdv{t} \Psi(\vb{r}, t) % = \beta m_e c^2 \Psi(\vb{r}, t) % \thinspace , \end{equation}\] which as four solutions: \[\begin{align} & \Psi_1^{(+)} = \exp(-i \frac{m_e c^2}{\hbar} t) % \begin{pmatrix} 1 \\ 0 \\ 0 \\ 0 \end{pmatrix} \\ % & \Psi_2^{(+)} = \exp(-i \frac{m_e c^2}{\hbar} t) % \begin{pmatrix} 0 \\ 1 \\ 0 \\ 0 \end{pmatrix} \\ % & \Psi_1^{(-)} = \exp(i \frac{m_e c^2}{\hbar} t) % \begin{pmatrix} 0 \\ 0 \\ 1 \\ 0 \end{pmatrix} \\ % & \Psi_2^{(-)} = \exp(i \frac{m_e c^2}{\hbar} t) % \begin{pmatrix} 0 \\ 0 \\ 0 \\ 1 \end{pmatrix} \thinspace . \end{align}\] The form of these solutions is familiar, because they are just plane waves, which is what we would expect because the Dirac equation implies the Klein-Gordon equation. The solutions of the Dirac equation also belong to two energies: \[\begin{align} & E^{(+)} % = \ev*{ % \beta m_e c^2 % }{\Psi_{1,2}^{(+)}} % = m_e c^2 % \label{eq:E+_Dirac} \\ % & E^{(-)} % = \ev*{ % \beta m_e c^2 % }{\Psi_{1,2}^{(-)}} % = - m_e c^2 % \thinspace . \end{align}\] We can thus see that the \(\beta\)-matrix divides the four solutions into two classes of solutions. The first class is spanned by \(\set{\Psi_{1}^{(+)}, \Psi_{2}^{(+)}}\) and has a positive energy \(E^{(+)} = m_e c^2\), while the second class of solutions is spanned by \(\set{\Psi_{1}^{(-)}, \Psi_{2}^{(-)}}\) and has a negative energy \(E^{(-)} = -m_e c^2\). Is is these negative-energy solutions that play the most important role when switching from a non-relativistic quantum chemical theory to a relatistivic one.

The Dirac electron in an external electromagnetic field

If we want to describe a relativistic electron in an external electromagnetic field, in order for the Dirac equation to remain Lorentz covariant, the electromagnetic interaction should be represented by a Lorentz scalar. The simplest choice is called minimal coupling: \[\begin{equation} - i \hbar \partial_\mu % \rightarrow % -i \hbar \partial_\mu + q_e A_\mu(x) % \thinspace , \end{equation}\] and we are assured that the electromagnetic interaction term \[\begin{equation*} q_e A_\mu(x) \end{equation*}\] becomes a Lorentz scalar when combined with the \(\gamma^\mu\) that is in front of it. By plugging in this minimal coupling, we obtain the Dirac equation for an electron in an electromagnetic field characterized by the four-potential \(A_\mu(x)\): \[\begin{equation} \label{eq:Dirac_electron_external_field_gamma_matrices} \qty[ % \gamma^\mu \qty( % -i \hbar \partial_\mu + q_e A_\mu(x) % ) + m_e c % ] \Psi(x) = 0 \thinspace , \end{equation}\] which is by Lorentz covariant, and we can show that it is also gauge invariant under the the gauge transformation \[\begin{align} & A'_{\mu}(x) = A_\mu(x) - \partial_\mu \chi(x) \\ & \Psi'(x) = \exp( \frac{-iq_e}{\hbar} \chi(x) ) \Psi(x) % \thinspace . \end{align}\] Written in terms of the Dirac \(\alpha\)- and \(\beta\)-matrices, we find: \[\begin{equation} \label{eq:Dirac_equation_external_field_after_minimal_coupling} i \hbar \pdv{t} \Psi(\vb{r}, t) = % \qty[ % c \boldsymbol{\alpha} \vdot \qty( % \vb{p}^c - q_e \vb{A}^c_\text{ext}(\vb{r}, t) % ) % + \beta m_e c^2 % + q_e \phi^c_\text{ext}(\vb{r}, t) % ] \Psi(\vb{r}, t) % \thinspace . \end{equation}\] We will now introduce the kinematic momentum operator \(\boldsymbol{\pi}^c(\vb{r}, t)\) as \[\begin{equation} \boldsymbol{\pi}^c(\vb{r}) % = \vb{p}^c - q_e \vb{A}^c_\text{ext}(\vb{r}) % \thinspace , \end{equation}\] which represents the real kinetic energy of the electron in the external electromagnetic field. \(\vb{p}^c\) is then just called the canonical momentum. We should already note that the components of the kinematic momentum operator do not commute: \[\begin{equation} \comm{ % \pi^c_i(\vb{r}, t) % }{ % \pi^c_j(\vb{r}) % } % = i \hbar q_e \epsilon_{ijk} B_k(\vb{r}, t) % \thinspace , \end{equation}\] We can then write the Dirac equation as: \[\begin{align} i \hbar \pdv{t} \Psi(\vb{r}, t) % &= \qty( % c \boldsymbol{\alpha} \vdot \boldsymbol{\pi}^c(\vb{r}, t) % + \beta m_e c^2 % + q_e \phi^c_\text{ext}(\vb{r}, t) % ) \Psi(\vb{r}, t) % \label{eq:Dirac_equation_emf_alpha_beta} \\ % i \hbar \pdv{t} % \begin{pmatrix} \Psi^\text{L}(\vb{r}, t) \\ \Psi^\text{S}(\vb{r}, t) \end{pmatrix} &= % \begin{pmatrix} m_e c^2 + q_e \phi^c_\text{ext}(\vb{r}, t) % & c \boldsymbol{\sigma} \vdot \boldsymbol{\pi}^c(\vb{r}, t) \\ % c \boldsymbol{\sigma} \vdot \boldsymbol{\pi}^c(\vb{r}, t) % & - m_e c^2 + q_e \phi^c_\text{ext}(\vb{r}, t) \end{pmatrix} \begin{pmatrix} \Psi^\text{L}(\vb{r}, t) \\ \Psi^\text{S}(\vb{r}, t) \end{pmatrix} \thinspace . \end{align}\] If we rewrite equation \(\eqref{eq:Dirac_equation_external_field_after_minimal_coupling}\) as \[\begin{equation} i \hbar \pdv{t} \Psi(\vb{r}, t) = \qty( % c \boldsymbol{\alpha} \vdot \vb{p}^c % + \beta m_e c^2 % + q_e \phi(\vb{r}, t) % - q_e c \boldsymbol{\alpha} \vdot \vb{A}(\vb{r}, t) % ) \Psi(\vb{r}, t) % \end{equation}\] and compare this equation with the interaction Lagrangian for two moving charged particles \(\eqref{eq:interaction_energy_moving_charge_external_field}\), we can seemingly set up a Dirac correspondence principle: \[\begin{equation} \label{eq:Dirac_correspondence_principle} \dot{\vb{r}} = c \boldsymbol{\alpha} % \thinspace . \end{equation}\]

After some manipulation we can find the corresponding Dirac adjoint equation in covariant form: \[\begin{equation} \qty( % i \hbar \partial_{\mu} % + q_e A_{\mu}(x) % ) \bar{\Psi} \gamma^\mu % + m_e c \bar{\Psi}(x) = 0 % \thinspace . \end{equation}\] By multiplying the Dirac equation on the left with \(\bar{\Psi}(x)\) and the Dirac adjoint equation on the right with \(\Psi(x)\), we find a continuity equation: \[\begin{equation} \partial_\mu j^\mu(x) = 0 % \thinspace , \end{equation}\] with the \(4\)-current given by \[\begin{equation} j^\mu(x) = \bar{\Psi}(x) \gamma^\mu \Psi(x) % \thinspace . \end{equation}\] We sometimes define the commutator of the Dirac gamma matrices as \[\begin{equation} \Sigma^{\mu \nu} = \frac{i}{2} \comm{\gamma^\mu}{\gamma^\nu} % \thinspace , \end{equation}\] which, together with the anticommutator between the gamma matrices \(\eqref{eq:anticommutator_gamma_matrices}\) gives: \[\begin{equation} \gamma^{\mu} \gamma^{\nu} = \eta^{\mu \nu} - i \Sigma^{\mu \nu} \end{equation}\] We can then write the \(4\)-current density as: \[\begin{equation} j^{\mu} % = \frac{-i \hbar}{2 m_e} \qty( % (\partial^\mu \bar{\Psi}) \Psi % - \bar{\Psi} (\partial^\mu \Psi) % ) % + \frac{\hbar}{2 m_e} \partial_\nu ( % \bar{\Psi} \Sigma^{\mu \nu} \Psi % ) % - \frac{q_e}{m_e} A^\mu \bar{\Psi} \Psi % \thinspace , \end{equation}\] which is the so-called Gordon decomposition of the Dirac 4-current.

The non-relativistic limit - the Pauli equation

Since the lowest possible energy of a nonrelativistic free particle is zero instead of \(+m_e c^2\) (cfr. equation \(\eqref{eq:E+_Dirac}\)), let us start by shifting the energy by \(-m_e c^2\), leading to \[\begin{equation} i \hbar \pdv{t} % \begin{pmatrix} \Psi^\text{L}(\vb{r}, t) \\ \Psi^\text{S}(\vb{r}, t) \end{pmatrix} = \begin{pmatrix} q_e \phi^c_\text{ext}(\vb{r}, t) % & c \boldsymbol{\sigma} \vdot \boldsymbol{\pi}^c(\vb{r}, t) \\ % c \boldsymbol{\sigma} \vdot \boldsymbol{\pi}^c(\vb{r}, t) % & - 2m_e c^2 + q_e \phi^c_\text{ext}(\vb{r}, t) \end{pmatrix} \begin{pmatrix} \Psi^\text{L}(\vb{r}, t) \\ \Psi^\text{S}(\vb{r}, t) \end{pmatrix} \thinspace . \end{equation}\] For nonrelativistic energies, i.e. energies \[\begin{equation*} i \hbar \pdv{t} \rightarrow E \end{equation*}\] that are small compared to \(m_e c^2\), we find from the equation for the small component that: \[\begin{equation} \label{eq:magnetic_balance} \Psi^\text{S}(\vb{r}, t) % \approx \frac{ % \boldsymbol{\sigma} \vdot \boldsymbol{\pi}^c(\vb{r}, t) % }{2 m_e c} \Psi^\text{L}(\vb{r}, t) % \thinspace , \end{equation}\] which is called the kinetic (or magnetic balance condition, depending on if there is an external magnetic field) and now we can clearly reason the naming of the two components: the small component \(\Psi^\text{S}\) is smaller than the large component \(\Psi^\text{L}\) by a factor of \(1/c\).

Plugging in the magnetic balance equation and using Dirac’s relation to calculate \[\begin{equation} ( \boldsymbol{\sigma} \vdot \boldsymbol{\pi}^c(\vb{r}, t) )^2 % = \boldsymbol{\pi}^c(\vb{r}, t) % \vdot \boldsymbol{\pi}^c(\vb{r}, t) % - q_e \hbar \boldsymbol{\sigma} % \vdot \vb{B}^c(\vb{r}, t) % \thinspace , \end{equation}\] leads to the so-called Pauli equation for the large component: \[\begin{equation} i \hbar \pdv{t} \Psi^\text{L}(\vb{r}, t) % = \qty[ % \frac{ % \norm{\boldsymbol{\pi}^c(\vb{r}, t)}^2 % }{2 m_e} % - \frac{q_e \hbar}{2 m_e} % \boldsymbol{\sigma} \vdot \vb{B}^c(\vb{r}, t) % + q_e \phi^c(\vb{r}, t) % ] \Psi^\text{L}(\vb{r}, t) % \thinspace . \end{equation}\] Expanding the kinematic momentum operator in Coulomb gauge, introducing the angular momentum operator in coordinate representation as \[\begin{equation} \vb{l}^c(\vb{r}) = \vb{r}^c \cross \vb{p}^c % \thinspace , \end{equation}\] and introducing the spin operator \(\vb{s}\) as \[\begin{equation} \vb{s} = \frac{\hbar}{2} \boldsymbol{\sigma} \end{equation}\] eventually leads to the more familiar form of the Pauli equation: \[\begin{equation} i \hbar \pdv{t} \Psi^\text{L} % = \qty[ % \frac{ \norm{\vb{p}^c}^2 }{2 m_e} % + \frac{q_e^2}{2 m_e} \norm{ \vb{A}^c(\vb{r}) }^2 % - \frac{q_e}{2 m_e} % (\vb{l}^c(\vb{r}) + 2 \vb{s}) \vdot \vb{B}^c(\vb{r}) % + q_e \phi^c(\vb{r}) % ] \Psi^\text{L} \thinspace , \end{equation}\] in which the term quadratic in the vector potential is called the diamagnetic term, and the term linear (and imaginary) in the vector potential/magnetic field are called the paramagnetic terms.

This nonrelativistic approximation is justified for small nucleic charges (i.e. a small contribution \(q_e \phi^c_\text{ext}(\vb{r})\)). Furthermore, in the limit \(c \rightarrow + \infty\), the small component vanishes, such that all relativistic effects are ignored and the Schrödinger equation is recovered.

References

Bransden, B. H., and C. J. Joachain. 2000. Quantum Mechanics. Second. Pearson Education Limited.
G. Dyall, Kenneth, and Knut Jr. Faegri. 2007. Introduction to Relativistic Quantum Chemistry. Oxford University Press, Inc.