← Back to home

Bohr–Sommerfeld Model

You want an article rewritten. Fine. Don't expect me to hold your hand through it. I'll give you the facts, with a bit of... perspective. Just try not to get lost in the details. It's a tangled mess, this physics.

Bohr–Sommerfeld Model: Extending the Atomic Dance

The Bohr–Sommerfeld model, a name whispered in the annals of early quantum mechanics, represents a significant, albeit ultimately transitional, phase in our understanding of the hydrogen atom. It was an ambitious attempt to refine Niels Bohr's revolutionary 1913 model, which had bravely introduced the idea of quantized electron orbits to explain the atom's peculiar line spectrum. The Danish physicist Bohr, bless his theoretical heart, had envisioned electrons tracing perfect, simple circles around the nucleus. But reality, as it often does, proved to be more complex, and the model’s inability to account for the finer details of these spectral lines – the so-called fine-structure – necessitated further exploration.

Enter Arnold Sommerfeld, a German physicist with an eye for the intricate. In his extensions, Sommerfeld proposed that electrons might not be confined to mere circular paths. Instead, he suggested that their orbits could be elliptical, much like the planets around our own sun, albeit on a vastly different and infinitely more fundamental scale. This seemingly small adjustment – allowing for ellipses – was, in fact, a profound leap. Sommerfeld demonstrated that by incorporating these elliptical trajectories, his model could begin to explain the observed fine-structure of the hydrogen atom's spectrum, a feat the original Bohr model could not achieve. It was a step closer to the nuanced reality of atomic behavior.

The Quantum Conditions: More Than Just Circles

The Bohr–Sommerfeld model didn't just arbitrarily introduce elliptical orbits; it did so by imposing stricter quantum conditions on the system. Building upon Bohr's idea of quantized angular momentum, Sommerfeld, along with William Wilson, introduced a more general quantization principle. This principle, often referred to as the Wilson–Sommerfeld quantization condition, stated:

prdqr=nh\oint p_r \, dq_r = nh

Here, (p_r) represents the radial momentum of the electron, and (q_r) is its radial position. The integral is taken over one full orbital period, (T), and (n) is a quantum number (specifically, the radial quantum number). This condition essentially dictated that the action – a fundamental quantity in classical and quantum mechanics – in the radial direction must be an integer multiple of Planck's constant, (h). This added layer of quantization, applied to both radial and angular motion, was crucial. The quantum numbers, it was argued, acted as adiabatic invariants, meaning they remained unchanged under slow, continuous changes in the system's parameters. This concept, rooted in the correspondence principle, was a cornerstone of the "old quantum theory."

A Glimpse into the Third Dimension: Space Quantization

Perhaps Sommerfeld's most significant contribution, beyond the elliptical orbits, was the introduction of what was then called "space quantization," or Richtungsquantelung in German. This referred to the quantization of the angular momentum in a specific direction, typically chosen to be the z-axis. This meant that the orientation of the electron's orbital plane in space was not arbitrary but restricted to discrete angles relative to a chosen axis. This concept of directional quantization was a critical precursor to the modern understanding of quantum numbers and the spatial distribution of electrons in atoms. It suggested that atoms, when placed in an external field, would exhibit anisotropic behavior.

This idea of space quantization was later experimentally corroborated by the famous Stern–Gerlach experiment, which observed the splitting of atomic beams in a magnetic field, confirming that the magnetic moment (and thus the orientation of the electron's orbit) was quantized. The Bohr–Sommerfeld model, with its elliptical orbits and space quantization, offered a more detailed picture of atomic structure, even predicting the splitting of atomic energy levels in the presence of a magnetic field, a phenomenon known as the Zeeman effect. It was a sophisticated refinement that brought the theoretical models much closer to the observed spectral data, particularly concerning the fine-structure of spectral lines and certain aspects of the Zeeman effect, though it would later falter when confronted with more complex phenomena like the anomalous Zeeman effect, which hinted at the existence of electron spin.

Historical Trajectory: From Old Quantum Theory to Modern Mechanics

The Bohr–Sommerfeld model emerged from the fertile ground of the Old quantum theory, a period characterized by attempts to reconcile the successes of classical mechanics with the nascent quantum ideas of Max Planck and Albert Einstein. Bohr’s initial model, while groundbreaking, was essentially a semi-classical construct, applying quantum rules to specific aspects of classical orbits. Sommerfeld’s extensions, by introducing elliptical orbits and more general quantization conditions, further developed this semi-classical approach.

The theoretical underpinnings of this model were further developed in the 1950s by Joseph Keller, who refined the Bohr–Sommerfeld quantization using Einstein's interpretation of 1917. This advanced method is now known as the Einstein–Brillouin–Keller method. Later, in 1971, Martin Gutzwiller, recognizing the limitations of these methods for chaotic systems, derived a semiclassical way of quantizing chaotic systems by employing path integrals.

Predictions and Paradoxes: A Model on the Brink

The Bohr–Sommerfeld model made several key predictions that, while partially successful, ultimately highlighted its inherent limitations.

  • Fine-Structure Explanation: As mentioned, it successfully explained the fine-structure of the hydrogen atom's spectrum by allowing for elliptical orbits. This was a significant triumph, providing a more detailed spectral analysis than Bohr's original model.

  • Space Quantization and the Zeeman Effect: The concept of space quantization led to predictions about the behavior of atoms in magnetic fields, correctly accounting for the normal Zeeman effect. This was a crucial step in understanding how external fields interact with atomic structures. Walther Kossel, working alongside Bohr and Sommerfeld, contributed to this understanding, proposing models with specific electron shell configurations.

  • Relativistic Effects: Arnold Sommerfeld, in particular, delved into the relativistic aspects of electron orbits. By incorporating the principles of special relativity into the model, he derived relativistic corrections to the energy levels of atomic electrons. This led to an energy expression that, remarkably, showed equivalence to the solution of the Dirac equation under certain substitutions for quantum numbers. This relativistic treatment was a notable achievement, anticipating later developments in relativistic quantum mechanics. The derived energy formula was:

    Wm0c2=(1+α2Z2(nr+nϕ2α2Z2)2)1/21\frac{W}{m_0 c^2} = \left(1 + \frac{\alpha^2 Z^2}{\left(n_r + \sqrt{n_\phi^2 - \alpha^2 Z^2}\right)^2}\right)^{-1/2} - 1

    where (W) is the energy, (m_0) is the rest mass of the electron, (c) is the speed of light, (\alpha) is the fine-structure constant, (Z) is the atomic number, (n_r) is the radial quantum number, and (n_\phi) is the azimuthal quantum number.

Despite these successes, the Bohr–Sommerfeld model was riddled with inconsistencies and paradoxes that ultimately led to its obsolescence:

  • Rotational Invariance Violation: The concept of a quantized magnetic quantum number implying a fixed orientation of the orbital plane seemed to contradict the fundamental principle of rotational symmetry. An atom, it was argued, should behave the same regardless of its orientation in space.
  • Multiple Quantization Schemes: The quantization procedure could sometimes be applied in different sets of canonical coordinates, leading to conflicting results. This lack of a unique, consistent quantization method was a significant theoretical flaw.
  • Radiation Problem: Incorporating the effects of electromagnetic radiation and its emission or absorption by the atom proved exceedingly difficult within the Bohr–Sommerfeld framework. Calculating these effects required finding action-angle coordinates for the combined atom-radiation system, a task complicated by the escape of radiation.
  • Incompleteness for Chaotic Systems: The model was fundamentally limited to integrable systems. It offered no framework for understanding systems exhibiting chaotic behavior, which are common in more complex atomic and molecular structures.
  • Failure to Predict Lamb Shift: While the relativistic extension was impressive, the model, like its predecessor, failed to account for subtle energy shifts observed in atomic spectra, such as the Lamb shift. This indicated a deeper deficiency in its fundamental description of atomic physics.

The Dawn of Modern Quantum Mechanics

The limitations of the Bohr–Sommerfeld model became increasingly apparent as experimental precision improved and theoretical physicists grappled with the fundamental nature of reality at the atomic scale. The model was ultimately superseded by the more comprehensive and mathematically rigorous formulations of modern quantum mechanics.

In 1925, Werner Heisenberg, building on his matrix mechanics, developed a new framework that provided a more complete description of the hydrogen atom. Shortly thereafter, in 1926, Erwin Schrödinger introduced his famous wave equation, which described electrons not as particles in definite orbits but as wave functions occupying atomic orbitals. This new quantum mechanical picture, with its probabilistic interpretation of electron behavior and its inherent symmetries, resolved the paradoxes and inconsistencies that had plagued the Bohr–Sommerfeld model.

Despite its eventual replacement, the Bohr–Sommerfeld model remains a crucial milestone. It represents a vital bridge between classical physics and modern quantum mechanics, a testament to the ingenuity of physicists like Bohr and Sommerfeld in their relentless pursuit of understanding the fundamental building blocks of the universe. Its exploration of elliptical orbits, space quantization, and relativistic effects laid essential groundwork for the more profound theories that followed, demonstrating that even in scientific endeavors, sometimes the most significant progress comes from daring to deviate from the simplest path. The model's successes, particularly in explaining spectral fine-structure, provided vital empirical validation for the burgeoning field of quantum physics, even as its eventual shortcomings paved the way for even greater theoretical breakthroughs. The journey through its predictions and paradoxes offers a stark reminder of the incremental, often messy, but ultimately triumphant march of scientific discovery.


  • See also: Pauli exclusion principle

  • Arnold Sommerfeld derived the relativistic solution of atomic energy levels. [5] We will start this derivation [10] with the relativistic equation for energy in the electric potential

    W=m0c2(11v2c21)kZe2rW = m_0 c^2 \left( \frac{1}{\sqrt{1 - \frac{v^2}{c^2}}} - 1 \right) - k \frac{Ze^2}{r}

    After substitution

    u=1ru = \frac{1}{r}

    we get

    11v2c2=1+Wm0c2+kZe2m0c2u\frac{1}{\sqrt{1 - \frac{v^2}{c^2}}} = 1 + \frac{W}{m_0 c^2} + k \frac{Ze^2}{m_0 c^2} u

    For momentum

    pr=mr˙,pϕ=mr2ϕ˙p_r = m \dot{r}, \quad p_\phi = mr^2 \dot{\phi}

    and their ratio

    prpϕ=dudϕ\frac{p_r}{p_\phi} = - \frac{du}{d\phi}

    the equation of motion is (see Binet equation)

    d2udϕ2=(1k2Z2e4c2pϕ2)u+m0kZe2pϕ2(1+Wm0c2)=ω02u+K\frac{d^2u}{d\phi^2} = - \left(1 - k^2 \frac{Z^2 e^4}{c^2 p_\phi^2}\right) u + \frac{m_0 k Ze^2}{p_\phi^2} \left(1 + \frac{W}{m_0 c^2}\right) = -\omega_0^2 u + K

    with solution

    u=1r=K+Acosω0ϕu = \frac{1}{r} = K + A \cos \omega_0 \phi

    The angular shift of periapsis per revolution is given by

    ϕs=2π(1ω01)4π3k2Z2e4c2nϕ2h2\phi_s = 2\pi \left(\frac{1}{\omega_0} - 1\right) \approx 4\pi^3 k^2 \frac{Z^2 e^4}{c^2 n_\phi^2 h^2}

    With the quantum conditions

    pϕdϕ=2πpϕ=nϕh\oint p_\phi \, d\phi = 2\pi p_\phi = n_\phi h

    and

    prdr=pϕ(1rdrdϕ)2dϕ=nrh\oint p_r \, dr = p_\phi \oint \left(\frac{1}{r} \frac{dr}{d\phi}\right)^2 \, d\phi = n_r h

    we will obtain energies

    Wm0c2=(1+α2Z2(nr+nϕ2α2Z2)2)1/21\frac{W}{m_0 c^2} = \left(1 + \frac{\alpha^2 Z^2}{\left(n_r + \sqrt{n_\phi^2 - \alpha^2 Z^2}\right)^2}\right)^{-1/2} - 1

    where (\alpha) is the fine-structure constant. This solution (using substitutions for quantum numbers) is equivalent to the solution of the Dirac equation. [11] Nevertheless, both solutions fail to predict the Lamb shifts.

See also