Fact-checked by Grok 2 weeks ago

Fine structure

The fine structure refers to the small splitting observed in the spectral lines of atoms, particularly in , resulting from relativistic corrections to the electron's and the -orbit interaction between the electron's and orbital . This splitting, on the order of the \alpha \approx 1/137.036, breaks the degeneracy of levels that would otherwise depend only on the principal n in the non-relativistic model. In , the fine structure shift is given by \Delta E_{n j} = -\frac{\alpha^2}{n^2} E_n \left( \frac{n}{j + 1/2} - \frac{3}{4} \right), where E_n is the non-relativistic , and j is the , leading to distinct levels for different j values within the same n. The phenomenon was first experimentally resolved in 1887 by and , who observed the hydrogen Balmer-alpha line as a closely spaced doublet separated by about 0.016 nm, rather than a single line. provided the initial theoretical explanation in 1916 by extending the to include relativistic effects and elliptical orbits, introducing the \alpha = e^2 / (4\pi \epsilon_0 \hbar c) as a measure of the strength of the electromagnetic interaction. In 1928, Paul Dirac's relativistic quantum mechanical equation for the fully accounted for the fine structure in hydrogen-like atoms, deriving the energy levels exactly to order (\alpha Z)^4, where Z is the , and incorporating spin naturally without ad hoc assumptions. In , the includes three main terms: the relativistic correction H_R = -p^4 / (8 m^3 c^2), the spin-orbit coupling H_{SO} = (e^2 / (2 m^2 c^2)) (\mathbf{S} \cdot \mathbf{L} / r^3), and the term H_D = \pi (e^2 \hbar / (2 m^2 c^2)) \delta(\mathbf{r}) for s-states. These effects are most prominent in light atoms like but scale with Z^4 in heavier atoms, influencing spectra and precision measurements such as the anomalous of the . The remains a fundamental parameter in , with its value continually refined through experiments like the and comparisons of theory and measurement.

History

Early spectroscopic observations

In the mid-19th century, spectroscopic studies of vapors revealed that many emission lines appeared as closely spaced s or multiplets, deviating from the single-line predictions of early empirical models for atomic spectra. These splittings were first systematically documented using prism-based spectroscopes, which dispersed light into its component wavelengths for . For instance, the prominent sodium D-lines in the region of the spectrum were observed as a with components at 589.0 nm and 589.6 nm, corresponding to transitions in the sodium . This was initially identified as dark features in the solar spectrum by in 1814, and confirmed as emission lines from heated sodium salts by and in 1860 through laboratory flame tests. Similar patterns emerged in the spectra of other metals, such as and , where principal series lines consistently showed two closely spaced components, with separations decreasing for higher quantum numbers. These observations, made possible by improved spectrographs, suggested underlying complexities in emission beyond simple relations. The of ruled gratings in the late 19th century further enhanced ; Henry A. Rowland's rulings from onward produced gratings with thousands of lines per inch, enabling spectrographs to distinguish finer separations that prisms could not. This technological advance was crucial for quantifying multiplet structures in spectra, as gratings offered higher dispersive power and reduced chromatic aberrations compared to refractive . For , the visible spectral lines were empirically organized by in 1885 into a series following a simple reciprocal square formula, capturing the gross structure positions of lines like the Balmer-alpha (H-α) at 656.3 nm. However, high-resolution measurements soon revealed deviations, with lines appearing broader or split. In 1887, and employed —a technique using light interference patterns to measure minute wavelength differences—to resolve the fine splitting in 's Balmer lines, quantifying the separation in the Balmer-alpha line at approximately 0.18 Å. Their work, conducted with a custom echelle grating and interferometer setup, demonstrated that the apparent single lines of the gross structure concealed closely spaced components, challenging the completeness of Balmer's formula. These empirical findings, later refined by Johannes Rydberg's 1889 generalization, underscored the need to investigate substructures in atomic spectra beyond the basic Rydberg-Ritz combination principle.

Sommerfeld's relativistic model

In 1916, extended Bohr's planetary model of the by incorporating and permitting elliptical orbits, thereby providing the first quantitative theoretical account of the fine structure splitting observed in atomic spectral lines. Sommerfeld generalized Bohr's quantization condition using action-angle variables from , introducing two s: the principal quantum number n (related to the total action) and the k (later associated with j + 1/2, where j is the ). The quantization rules were \oint p_\phi \, d\phi = k h for the angular motion and \oint p_r \, dr = (n - k) h for the radial motion, with h as Planck's constant. To incorporate relativity, Sommerfeld accounted for the increase in with , using the relativistic p = \gamma m_0 v, where \gamma = (1 - v^2/c^2)^{-1/2}, m_0 is the rest mass, and c is the ; this modified the and orbit precession in the Keplerian problem. By solving the relativistic under these quantization conditions, Sommerfeld derived energy levels that deviated from the pure Bohr formula E_n = -13.6 \, \mathrm{eV}/n^2. The fine structure correction emerged naturally from the relativistic dynamics, manifesting as an energy shift proportional to \alpha^2, where \alpha \approx 1/137 is the dimensionless introduced by Sommerfeld and defined as \alpha = e^2 / (\hbar c) (in , with e the and \hbar = h/2\pi). This correction arises from the variation of electron velocity along the elliptical orbit, leading to a small but observable splitting of spectral lines. To lowest order in \alpha^2, the energy levels in Sommerfeld's model for hydrogen (Z=1) are given by E_{n,j} \approx -\frac{13.6 \, \mathrm{eV}}{n^2} \left[ 1 + \frac{\alpha^2}{n^2} \left( \frac{n}{j + 1/2} - \frac{3}{4} \right) \right], where n = 1, 2, \dots is the principal quantum number determining the gross energy scale, and j = 1/2, 3/2, \dots, n - 1/2 labels the fine structure sublevels (with degeneracy lifted according to the orbital angular momentum, though spin-orbit effects were not yet incorporated). This perturbative expansion of Sommerfeld's exact relativistic solution highlights how the term \alpha^2 / n^2 scales the splitting relative to the Rydberg energy. Sommerfeld's model accurately predicted the fine structure splittings in Balmer lines such as H\alpha and H\beta, aligning closely with Paschen's precise spectroscopic measurements from 1908–1914, and extended successfully to explain analogous splittings in spectra like sodium and D-lines. Despite these successes, the theory had key limitations, including its neglect of electron spin, which prevented a full of certain structures, and its restriction to one-electron systems without addressing multi-electron interactions.

Theoretical Background

Gross structure of atomic spectra

The gross structure of atomic spectra refers to the primary features observed in the emission or absorption lines of atoms, as described by non-relativistic , before accounting for finer relativistic corrections. In the case of the , the energy levels are determined by solving the time-independent for a single in the potential of the . The equation separates into radial and angular parts in spherical coordinates, with the angular solutions yielding the orbital l (ranging from 0 to n-1, where n is the principal ) and the m_l. The radial equation leads to quantized levels given by E_n = -\frac{13.6 \, \mathrm{eV}}{n^2}, where n = 1, 2, 3, \dots, independent of l and m_l, resulting in degeneracy for states with the same n but different l. This degeneracy in hydrogen means that the gross spectral structure consists of lines corresponding solely to transitions between different n levels, without splitting due to . In multi-electron atoms, however, the gross structure is modified by electron-electron interactions, which partially lift the l-degeneracy through penetration effects: inner electrons shield the imperfectly, causing energy levels with the same n but different l to separate, with s-orbitals (l=0) having lower energy than p (l=1), and so on. Despite this, the dominant energy dependence remains on n, leading to spectral series like the Lyman (ultraviolet, to n=1), Balmer (visible, to n=2), and Paschen (, to n=3) series. The frequencies of these lines are predicted by the : \nu = R \left( \frac{1}{n_1^2} - \frac{1}{n_2^2} \right), where R is the (\approx 1.097 \times 10^7 \, \mathrm{m}^{-1} for ), n_1 < n_2, and transitions obey electric dipole selection rules such as \Delta l = \pm 1. These selection rules ensure that observed gross structure lines form discrete series without the intricate substructure from spin or relativistic effects, providing a foundational framework for understanding atomic spectra as simple patterns of energy differences scaled by $1/n^2. Early spectroscopic measurements, such as those by in 1885, aligned closely with this non-relativistic model, though subtle deviations hinted at additional influences.

Relativistic effects and the fine structure constant

In the non-relativistic framework of the Schrödinger equation, atomic spectra exhibit a gross structure determined by the principal and orbital quantum numbers, but deviations arise when electron velocities become comparable to the speed of light, necessitating relativistic corrections. These effects are particularly pronounced for inner-shell electrons in atoms, where the orbital velocity v satisfies v/c \approx \alpha \approx 1/137, with c the speed of light, making special relativity essential for accurate spectral predictions. The fine structure constant \alpha, introduced by Arnold Sommerfeld in 1916 to parameterize the relativistic splitting of spectral lines, is a dimensionless quantity defined as \alpha = \frac{e^2}{4\pi \epsilon_0 \hbar c}, where e is the elementary charge, \epsilon_0 the vacuum permittivity, \hbar the reduced Planck's constant, and c the speed of light. Its measured value is \alpha \approx 7.2973525693 \times 10^{-3}, or equivalently $1/\alpha \approx 137.035999177. In quantum electrodynamics (QED), \alpha serves as the fundamental coupling constant governing the strength of electromagnetic interactions between charged particles and photons. The fine structure originates from three primary relativistic perturbations to the non-relativistic Hamiltonian: the relativistic correction to the kinetic energy, spin-orbit coupling, and the Darwin contact term. These contributions, treated perturbatively, scale as order (Z\alpha)^2 relative to the gross structure energies, where Z is the atomic number, leading to small but observable splittings in energy levels. The resulting fine structure splittings are thus proportional to \alpha^2 times the gross energy differences, typically requiring high-resolution spectroscopy to resolve, as the relative scale is on the order of $10^{-4} to $10^{-5} for light atoms.

Fine Structure in Hydrogen

Relativistic kinetic energy correction

In the non-relativistic for the hydrogen atom, the kinetic energy of the electron is approximated as T = \frac{p^2}{2m}, where p is the electron momentum operator and m is the electron mass. However, this approximation fails at high velocities comparable to the speed of light c, which occurs for inner atomic orbitals due to the strong . The relativistic expression for the total energy of a free particle is E = \sqrt{(pc)^2 + (mc^2)^2}, and expanding this in powers of p/(mc) for p \ll mc yields the kinetic energy T \approx \frac{p^2}{2m} - \frac{p^4}{8m^3 c^2} + \cdots. The p^4 term represents the leading relativistic correction to the non-relativistic kinetic energy. To incorporate this correction into the hydrogen atom, the relativistic term is treated as a perturbation to the non-relativistic Hamiltonian H_0 = \frac{p^2}{2m} - \frac{Z e^2}{4\pi \epsilon_0 r}, where Z = 1 for hydrogen and the potential is the Coulomb interaction. The perturbation Hamiltonian is thus H' = -\frac{p^4}{8 m^3 c^2}. The first-order energy shift is given by the expectation value \Delta E = \langle \psi | H' | \psi \rangle, where \psi are the unperturbed hydrogen wavefunctions labeled by quantum numbers n, l, and m. Computing \langle p^4 \rangle directly is challenging, but it can be evaluated using the virial theorem, which relates \langle p^2 \rangle = -2 m E_n (with E_n = -\frac{m (Z \alpha c)^2}{2 n^2} the non-relativistic energy levels and \alpha the fine structure constant) and radial expectation values like \langle 1/r^3 \rangle. An alternative approach expands \langle p^4 \rangle = \langle (2m (H_0 + V))^2 \rangle, leveraging the Schrödinger equation p^2 \psi = 2m (E_n - V) \psi with V = -\frac{Z e^2}{4\pi \epsilon_0 r}, and simplifies via \langle V \rangle = 2 E_n from the virial theorem and known \langle V^2 \rangle. The resulting energy correction depends on the principal quantum number n and orbital angular momentum quantum number l, but is independent of the magnetic quantum number m due to the spherical symmetry of H'. The standard expression is \Delta E_\text{kin} = E_n \frac{(Z \alpha)^2}{n^2} \left( \frac{n}{l + \frac{1}{2}} - \frac{3}{4} \right), where the factor (Z \alpha)^2 / n^2 scales the non-relativistic energy by the square of the (for Z=1, \alpha \approx 1/137). This formula arises from the exact evaluation of \langle p^4 \rangle = \frac{(Z \alpha m c)^4}{\hbar^4 n^3 (l + 1/2)} times additional terms, but the l + 1/2 interpolation provides a compact approximation that matches the precise radial integral results closely. The correction is always negative, shifting energy levels downward more for states with smaller l (higher angular momentum barriers reduce relativistic effects), and its magnitude scales as (Z \alpha)^4 m c^2 / n^3, reflecting the v^4/c^4 relativistic order. For example, in the n=2, l=0 state of hydrogen (2s orbital), E_2 = -3.4 eV and the correction evaluates to \Delta E_\text{kin} \approx -1.5 \times 10^{-4} eV, which is on the order of the fine structure splitting compared to the gross structure spacing of about 10 eV between n=1 and n=2. This shift contributes significantly to the fine structure but is much smaller than the non-relativistic energies, justifying the perturbative approach. Higher n or l reduces the effect, with the correction vanishing in the classical limit of large orbits.

Spin-orbit coupling

The spin-orbit coupling originates from the relativistic interaction between the electron's spin magnetic moment and the effective magnetic field experienced by the electron due to its orbital motion around the proton in the Coulomb electric field of the nucleus. In the rest frame of the electron, the proton appears to move, generating a magnetic field that couples to the electron's spin, with the factor of 1/2 arising from the Thomas precession correction to avoid double-counting the relativistic effects. This phenomenon was first quantitatively derived by Llewellyn Thomas in 1926 using a classical model of the spinning electron in the hydrogen atom. The spin-orbit interaction is described by the perturbative Hamiltonian term
H_{\rm SO} = \frac{1}{2 m_e^2 c^2} \frac{1}{r} \frac{dV}{dr} \, \mathbf{S} \cdot \mathbf{L},
where m_e is the electron mass, c is the speed of light, r is the radial distance, V(r) is the Coulomb potential V(r) = -\frac{Z e^2}{4\pi \epsilon_0 r} (with Z=1 for hydrogen), \mathbf{S} is the spin angular momentum operator, and \mathbf{L} is the orbital angular momentum operator. For the hydrogen Coulomb potential, \frac{dV}{dr} = \frac{e^2}{4\pi \epsilon_0 r^2}, so the Hamiltonian simplifies to
H_{\rm SO} = \frac{e^2}{8\pi \epsilon_0 m_e^2 c^2} \frac{\mathbf{S} \cdot \mathbf{L}}{r^3}.
This form incorporates the g-factor of approximately 2 for the electron spin and the Thomas precession factor of 1/2.
To compute the first-order energy correction due to spin-orbit coupling, perturbation theory is applied in the basis of states that diagonalize the total angular momentum \mathbf{J} = \mathbf{L} + \mathbf{S}, as H_{\rm SO} commutes with \mathbf{J}^2 and J_z. The expectation value of \mathbf{S} \cdot \mathbf{L} is obtained from the identity
\mathbf{S} \cdot \mathbf{L} = \frac{1}{2} \left( J^2 - L^2 - S^2 \right),
yielding
\langle \mathbf{S} \cdot \mathbf{L} \rangle = \frac{\hbar^2}{2} \left[ j(j+1) - l(l+1) - s(s+1) \right],
where j is the total angular momentum quantum number, l is the orbital quantum number, and s = 1/2 is the spin quantum number. The radial expectation value \langle 1/r^3 \rangle for hydrogenic wave functions in the state |n, l\rangle is \langle 1/r^3 \rangle = \frac{Z^3}{a_0^3 n^3 l (l + 1/2) (l + 1)}, where a_0 is the and n is the principal quantum number. Substituting these into the perturbation theory expression gives the spin-orbit energy shift.
The resulting spin-orbit contribution to the energy levels of hydrogen is
\Delta E_{\rm SO} = - E_n^{(0)} \frac{(Z \alpha)^2}{n^2} \frac{j(j+1) - l(l+1) - s(s+1)}{ l (l + 1/2) (l + 1)},
where E_n^{(0)} = -\frac{13.6 \, \rm eV}{n^2} is the non-relativistic ground-state energy scaled to principal quantum number n, \alpha \approx 1/137 is the , and Z = 1 for hydrogen. With s(s+1) = 3/4, this formula depends on j but not on the magnetic quantum numbers m_j or m_l.
This spin-orbit term lifts the degeneracy between states with j = l + 1/2 and j = l - 1/2 (for l \geq 1), while leaving l = 0 states unaffected since \mathbf{L} = 0. For the n=2, l=1 (2p) subshell of hydrogen, the states split into $2p_{3/2} (j = 3/2) and $2p_{1/2} (j = 1/2), with an energy separation of approximately $4.5 \times 10^{-5} \, \rm eV (or 10.97 GHz, corresponding to 0.365 cm^{-1} in wavenumber). This splitting contributes to the observed fine structure in the hydrogen spectral lines, such as the .

Darwin term

The Darwin term is a relativistic correction arising in the perturbative expansion of the for an electron in an external potential, first derived by in his 1928 calculation of the fine structure of hydrogen-like atoms using the . It appears explicitly in the , which decouples the positive and negative energy components of the Dirac wave function to obtain a non-relativistic Hamiltonian with relativistic corrections of order (v/c)^2. The term takes the form H_D = \frac{\hbar^2}{8 m^2 c^2} \nabla^2 V, where m is the electron mass, c is the speed of light, and V is the external potential. For the Coulomb potential V(r) = -\frac{Z e^2}{4\pi \epsilon_0 r} in a hydrogen-like atom, the Laplacian yields \nabla^2 V = \frac{Z e^2}{\epsilon_0} \delta^3(\mathbf{r}), transforming H_D into a contact interaction proportional to the Dirac delta function at the nucleus. This delta function ensures the expectation value \langle H_D \rangle vanishes for states with angular momentum l > 0, as the corresponding wave functions have zero probability density at the origin, \psi_{nlm}(0) = 0. For s-states (l = 0), the shift is nonzero and given by \Delta E_D = \frac{(Z \alpha)^4 m c^2}{2 n^3}, where \alpha is the , Z is the , and n is the principal ; this follows from evaluating \langle \delta^3(\mathbf{r}) \rangle = |\psi_{n00}(0)|^2 = \frac{Z^3}{\pi n^3 a_0^3}, with a_0 the . Physically, the Darwin term accounts for the electron's —a trembling motion inherent to the , with amplitude on the of the reduced \hbar / (m c)—which effectively smears the electron's position and allows s-wave functions, which would otherwise be forbidden from penetrating the classically singular potential at the , to interact with a folded potential there. For the ground state (n=1, l=0), this shift is of \alpha^2 times the relativistic kinetic energy correction, providing a partial positive offset that ensures consistency with the exact Dirac energy levels when combined in the perturbative series.

Combined perturbative corrections

The total fine structure Hamiltonian for the hydrogen atom in the non-relativistic limit is given by the sum of the relativistic corrections to the H_{\text{kin}}, the spin-orbit interaction H_{\text{SO}}, and the term H_{\text{Darwin}}, such that H_{\text{fs}} = H_{\text{kin}} + H_{\text{SO}} + H_{\text{Darwin}}. These terms arise as first-order perturbations to the and collectively account for relativistic effects up to order (Z\alpha)^4, where Z is the and \alpha is the . Heisenberg and Jordan first demonstrated in the framework of that incorporating the spin-orbit coupling alongside relativistic corrections reproduces the observed fine structure splittings originally predicted by Sommerfeld's semi-classical model. To first order in perturbation theory, the combined energy shift for a state with principal quantum number n and total angular momentum quantum number j is \Delta E_{\text{fs}} = E_n \frac{(Z\alpha)^2}{n^2} \left( \frac{n}{j + 1/2} - \frac{3}{4} \right), where E_n = -\frac{13.6 \, \text{eV}}{n^2} is the unperturbed Bohr energy. This formula, derived by summing the individual perturbative contributions, shows that the fine structure shift depends only on n and j, independent of the orbital angular momentum quantum number l. The relativistic kinetic energy and Darwin term together yield an l-dependent correction proportional to \langle p^4 \rangle / (8 m_e^3) and a contact term for l=0 states, respectively; adding the spin-orbit term H_{\text{SO}} \propto \mathbf{L} \cdot \mathbf{S} causes the l dependence to cancel, resulting in a pure j-dependence that aligns with the Landé interval rule for level spacings. This combined perturbative treatment has key spectral implications for . For the n=2 level, it predicts degeneracy between the $2s_{1/2} and $2p_{1/2} states (both with j=1/2), while the $2p_{3/2} state ( j=3/2 ) lies higher by approximately $0.365 \, \text{cm}^{-1}. Consequently, transitions from these levels to the exhibit fine structure doublets, such as in the Balmer \alpha line at $656.3 \, \text{nm}, with the $2p_{3/2} \to 1s_{1/2} component at slightly higher frequency than $2p_{1/2} \to 1s_{1/2}. These predictions agree with high-resolution spectroscopic measurements to within about 0.1%, confirming the validity of the perturbative approach before corrections like the are considered.

Exact Relativistic Treatment

Dirac equation solutions for hydrogen

The Dirac equation provides the relativistic wave equation for a single electron of mass m and charge -e in an external electrostatic potential \phi, given by (c \vec{\alpha} \cdot \vec{p} + \beta m c^2 - e \phi) \psi = E \psi, where \vec{\alpha} and \beta are $4 \times 4 Dirac matrices, \vec{p} = -i \hbar \nabla is the momentum operator, c is the speed of light, E is the total energy (including rest mass), and \psi is a four-component spinor wave function. For the hydrogen atom (Z=1), the scalar potential is the Coulomb potential \phi = e / (4 \pi \epsilon_0 r), so the interaction term is V = -e \phi = -e^2 / (4 \pi \epsilon_0 r). To solve this equation exactly for bound states, the time-independent is separated in spherical coordinates, exploiting the rotational invariance of the potential. The four-component \psi(\vec{r}) is expressed as \psi(\vec{r}) = \begin{pmatrix} \frac{F(r)}{r} \Omega_{\kappa m_j}(\theta, \phi) \\ i \frac{G(r)}{r} \Omega_{-\kappa m_j}(\theta, \phi) \end{pmatrix}, where F(r) and G(r) are the large and small radial components, respectively, and \Omega_{\kappa m_j} are two-component spherical spinors characterized by the total quantum numbers j (with m_j its projection) and the relativistic quantum number \kappa = \pm (j + 1/2), which relates the orbital angular momentum l to j via l = j \pm 1/2 (specifically, \kappa = -(l+1) for j = l - 1/2 and \kappa = l for j = l + 1/2). An additional n = 1, 2, \dots arises from the radial equation, satisfying n \geq j + 1/2. This separation yields two coupled first-order differential equations for F(r) and G(r), which can be solved analytically. The exact energy eigenvalues depend only on n and j (independent of l and m_j), given by E_{n j} = m c^2 \left[ 1 + \left( \frac{Z \alpha}{n - (j + 1/2) + \sqrt{(j + 1/2)^2 - (Z \alpha)^2}} \right)^2 \right]^{-1/2}, where \alpha = e^2 / (4 \pi \epsilon_0 \hbar c) is the fine structure constant and Z is the nuclear charge ( Z=1 for hydrogen). This formula reproduces the exact relativistic spectrum, including fine structure splittings. In the non-relativistic limit (Z \alpha \ll 1), it expands to the Schrödinger binding energy plus relativistic corrections: E_{n j} \approx m c^2 - \frac{m c^2 (Z \alpha)^2}{2 n^2} + \mathcal{O}((Z \alpha)^4). The radial wave functions F(r) and G(r) are expressed in terms of confluent hypergeometric functions (or equivalently, associated ) multiplied by exponential factors, with involving gamma functions to account for the relativistic parameters. For (Z=1), explicit forms are available for low-lying states, such as the (n=1, j=1/2, \kappa=-1), where F(r) dominates for large r while G(r) is small but non-negligible near the origin. The probability density |\psi|^2 reveals relativistic effects, notably a contraction of the radial distribution for inner (s-like) orbitals: the expectation value \langle r \rangle for the 1s decreases by approximately \frac{3}{2} \alpha^2 (or about 0.004%) compared to the non-relativistic case due to increased velocity near the , pulling the wave function inward. This contraction is more pronounced for higher Z. For low Z, these exact solutions reduce to perturbative approximations of the non-relativistic plus fine structure corrections.

Fine structure formula from Dirac theory

The exact energy levels in the Dirac theory for the are given by E_{nj} = mc^2 \left[ 1 + \left( \frac{Z\alpha}{n - \left(j + \frac{1}{2}\right) + \sqrt{\left(j + \frac{1}{2}\right)^2 - (Z\alpha)^2}} \right)^2 \right]^{-\frac{1}{2}}, where m is the rest , c is the , \alpha is the , Z is the , n is the principal quantum number, and j is the . To obtain the fine structure correction, this expression is expanded in powers of the small parameter Z\alpha using the . The leading term is the rest energy mc^2, followed by the non-relativistic Bohr energy E_n = -\frac{1}{2} m (Z\alpha)^2 c^2 / n^2, and the fine structure shift \Delta E_\mathrm{fs} at order (Z\alpha)^4: E_{nj} \approx mc^2 + E_n + \Delta E_\mathrm{fs}, with \Delta E_\mathrm{fs} = E_n \frac{(Z\alpha)^2}{n^2} \left( \frac{n}{j + \frac{1}{2}} - \frac{3}{4} \right). This expansion confirms the relativistic corrections derived perturbatively from the non-relativistic Schrödinger equation. In the Dirac framework, the fine structure arises naturally from the relativistic wave equation, unifying the relativistic kinetic energy correction, spin-orbit coupling, and Darwin term without requiring separate perturbative additions; these effects emerge from the structure of the Dirac Hamiltonian and its coupling to the Coulomb potential. Higher-order terms in the expansion, such as those at order (Z\alpha)^6, account for effects like nuclear recoil, but the fine structure is conventionally defined up to the (Z\alpha)^4 contribution. For (Z=1), the Dirac formula exactly reproduces the perturbative fine structure result at this order, as higher terms are negligible given \alpha \approx 1/137. For higher Z, deviations appear because Z\alpha is no longer small, indicating the need for quantum electrodynamic corrections beyond the Dirac approximation.

Multi-Electron Atoms

Scaling and qualitative features

In multi-electron atoms, the fine structure splitting for inner electron shells scales proportionally to Z^4 \alpha^2, where Z is the and \alpha is the , in contrast to the gross structure energy levels that scale as Z^2. This Z^4 dependence stems from the relativistic corrections—such as spin-orbit coupling and relativistic —being of relative order (Z \alpha)^2 to the non-relativistic energies, which themselves scale as Z^2. For outer shells, this scaling is moderated by electron screening effects, where inner electrons reduce the Z_\mathrm{eff} experienced by electrons, thereby weakening the fine structure relative to inner shells. As Z increases, the parameter Z \alpha grows, and based on non-relativistic Schrödinger solutions breaks down when Z \alpha \approx 1, typically for inner shells in atoms with Z > 30, due to the comparable magnitude of relativistic and non-relativistic contributions. In such cases, more sophisticated approaches are required, including variational methods to incorporate variationally or the Dirac-Fock method, which solves the many-electron self-consistently to account for both relativistic kinematics and electron-electron interactions. Qualitatively, the fine structure in multi-electron atoms intertwines with electrostatic electron-electron interactions, leading to distinct schemes that depend on Z. For lighter atoms (low Z), where spin-orbit effects are relatively weak, the (Russell-Saunders) coupling scheme dominates, in which individual orbital angular momenta \mathbf{l}_i couple to total \mathbf{L} = \sum \mathbf{l}_i and spins \mathbf{s}_i to total \mathbf{S} = \sum \mathbf{s}_i, followed by coupling of \mathbf{L} and \mathbf{S} to total \mathbf{J}. In heavier atoms (high Z), stronger relativistic effects favor the jj coupling scheme, where each electron's total \mathbf{j}_i = \mathbf{l}_i + \mathbf{s}_i couples directly to the overall \mathbf{J} = \sum \mathbf{j}_i, better capturing the dominance of individual spin-orbit interactions over electrostatic correlations. This transition from to jj coupling reflects the increasing relative importance of fine structure over gross structure splittings with rising Z.

Examples in alkali and heavier atoms

In alkali atoms, the fine structure arises primarily from the spin-orbit interaction of the and is well-described by LS coupling, where the total orbital and angular momenta couple to form terms split by total J. A representative example is sodium, where the 3p ^2P term splits into ^2P_{1/2} and ^2P_{3/2} levels separated by 17.22 cm^{-1}, manifesting in the optical 3p → 3s transition as the yellow D-line doublet at wavelengths of 589.158 (^2P_{3/2} → ^2S_{1/2}) and 589.757 (^2P_{1/2} → ^2S_{1/2}). This splitting reflects the perturbative nature of the interaction in light alkali atoms, with the higher-J level lying above the lower-J one in . In heavier alkali atoms like cesium, the fine structure splitting increases due to higher , reaching 554 cm^{-1} for the 6p ^2P term and producing the prominent D lines at 852.347 nm (^2P_{3/2} → ^2S_{1/2}) and 894.593 nm (^2P_{1/2} → ^2S_{1/2}). These lines, separated by about 42 nm in wavelength, were resolved in atomic spectra using diffraction gratings as early as the late 19th and early 20th centuries, confirming the doublet nature predicted by spin-orbit theory. For heavier atoms beyond the series, such as mercury (Z=80), relativistic effects and strong spin-orbit coupling shift the dominant scheme to jj-coupling, where individual momenta couple separately before total J formation. The 6s6p ^3P splits into J=0, 1, and 2 levels with energies at 37,645 cm^{-1} (^3P_0), 39,412 cm^{-1} (^3P_1), and 44,043 cm^{-1} (^3P_2), yielding splittings of 1,767 cm^{-1} (J=1 to J=0) and 4,631 cm^{-1} (J=2 to J=1), for a total span of about 6,398 cm^{-1} or 0.79 . This large splitting, scaling roughly as Z^4 for , highlights the enhanced relativistic influence in high-Z atoms and contributes to the complex triplet structure observed in mercury's UV-visible spectrum. Occasional anomalies, such as inversion of the fine structure ordering, arise from interaction perturbing the levels; for instance, in the p^2 of oxygen, the ^3P exhibits inverted splitting with J=2 as the lowest level (unlike the normal J=0 lowest in carbon's analogous ), due to mixing with higher-lying configurations that alters the effective spin-orbit matrix elements.

References

  1. [1]
    Hydrogen Fine Structure - HyperPhysics
    It is found to be a closely-spaced doublet. This splitting is called fine structure and was one of the first experimental evidences for electron spin.Missing: definition | Show results with:definition
  2. [2]
    [PDF] Quantum Physics III Chapter 2: Hydrogen Fine Structure
    Apr 2, 2022 · HYDROGEN ATOM FINE STRUCTURE bound state spectrum can be better appreciated using the speed of light to consider both the fine structure ...
  3. [3]
    Current advances: The fine-structure constant
    The usual fine-structure constant α is defined as the square of the completely screened charge, that is, the value observed at infinite distance or in the ...
  4. [4]
    [PDF] Fine Structure in Hydrogen and Alkali Atoms
    In these notes we consider the fine structure of hydrogen-like and alkali atoms, which concerns the effects of relativity and spin on the dynamics of the ...
  5. [5]
    [PDF] Fine Structure of the Hydrogen Atom Part I - MIT
    This was discovered by Michelson and Morley* in 1887 to be actually a doublet. An explanation of this fine structure was given by Sommerfeld in 1916 in.
  6. [6]
    This Month in Physics History | American Physical Society
    The man who did the most to improve the precision of diffraction gratings, thereby revolutionizing the field of spectroscopy, was the 19th century American ...Missing: alkali metals multiplets
  7. [7]
    Zur Quantentheorie der Spektrallinien - Sommerfeld - 1916
    Zur Quantentheorie der Spektrallinien. A. Sommerfeld,. A. Sommerfeld ... Download PDF. back. Additional links. About Wiley Online Library. Privacy Policy ...
  8. [8]
    [PDF] {How Sommerfeld extended Bohr's model of the atom (1913–1916)}
    Jan 30, 2014 · Abstract. Sommerfeld's extension of Bohr's atomic model was moti- vated by the quest for a theory of the Zeeman and Stark effects. The.
  9. [9]
  10. [10]
  11. [11]
    [PDF] Sommerfeld fine structure constant α and its physical interpretation
    The fine structure constant α, introduced by Sommerfeld, is 2πe²/hc, a dimensionless number. Its inverse is approximately 137.
  12. [12]
    Fine Structure of Hydrogen - Richard Fitzpatrick
    This modification of the energy levels of a hydrogen atom due to a combination of relativity and spin-orbit coupling is known as fine structure. Now, it is ...Missing: (Zα) | Show results with:(Zα)
  13. [13]
    The Motion of the Spinning Electron - Nature
    Download PDF. Letter; Published: 10 April 1926. The Motion of the Spinning Electron. L. H. THOMAS. Nature volume 117, page 514 (1926)Cite this article. 12k ...
  14. [14]
  15. [15]
    [PDF] Real hydrogen atom - bingweb
    E = -3.01884 x 10-5 eV for the 2 2P1/2 state The energy difference is 3∆ = 4.52826 x 10-5 eV. Fig. Splitting of 2 P level into two states (2 2P3/2 and 2 2P1/2) ...
  16. [16]
    The wave equations of the electron - Journals
    The wave equations of the electron. Charles Galton Darwin. Google Scholar · Find this author on PubMed · Search for more papers by this author. Charles Galton ...
  17. [17]
    On the Dirac Theory of Spin 1/2 Particles and Its Non-Relativistic Limit
    On the Dirac Theory of Spin 1/2 Particles and Its Non-Relativistic Limit. Leslie L. Foldy · Siegfried A. Wouthuysen* ... Phys. Rev. 78, 29– Published 1 April ...
  18. [18]
    The quantum theory of the electron - Journals
    Wu S (2025) Exact solutions of Dirac equation for hydrogen atom using the linear combination of orthogonal Laguerre basis functions, Scientific Reports ...
  19. [19]
    Die Energieniveaus des Wasserstoffatoms nach der Diracschen ...
    Cite this article. Gordon, W. Die Energieniveaus des Wasserstoffatoms nach der Diracschen Quantentheorie des Elektrons. Z. Physik 48, 11–14 (1928). https://doi.
  20. [20]
    Relativistic Effects in the Electronic Structure of Atoms | ACS Omega
    Sep 22, 2017 · Periodic trends in relativistic effects are investigated from 1 H through 103 Lr using Dirac–Hartree–Fock and nonrelativistic Hartree–Fock calculations.
  21. [21]
    [PDF] Notes on Atomic Structure 1. Introduction 2. Hydrogen Atoms and ...
    Jan 7, 2011 · For heavier atoms, the transition frequencies scale as ω∝Z2, the radii scale as 1/Z, and so the transition rates scale as Z4, e.g. they are 16× ...
  22. [22]
    Relativistic Perturbation Theory Formalism to Computing Spectra ...
    The theoretical results are obtained within the Hartree-Fock, Dirac-Fock methods by Ostrovsky-Sheynerman, relativistic PT theory with the empirical model ...
  23. [23]
    [PDF] LS-coupling and jj-coupling
    Spin-orbit interaction in multi-electron atoms. • We now have two effects to ... Fine structure in LS-coupling. • Now, we add the spin-orbit term of the ...
  24. [24]
    [PDF] Sodium D Line Data - Daniel A. Steck
    May 27, 2000 · This document presents physical and optical properties of sodium, including the D line components, D1 and D2, with D2 being more relevant for ...
  25. [25]
    [PDF] Cesium D Line Data - Daniel A. Steck
    Jan 23, 1998 · In this reference we present many of the physical and optical properties of cesium that are relevant to various quantum optics experiments.
  26. [26]
    [PDF] Atomic Physics - Doyle Group | Harvard
    5.2 The jj-coupling scheme. To calculate the fine structure in the trS-coupling scheme v/e treated the spin-orbit interaction as a perturbation on a term). 2s ...
  27. [27]
    [PDF] Lecture 12 Atomic structure - TCM
    Therefore, the energy levels for the oxygen ground state configuration are the same as for carbon, except that the fine structure multiplet is inverted, in ...