Fact-checked by Grok 2 weeks ago

Bohr radius

The Bohr radius, denoted a_0, is a fundamental in , equal to $5.29177210544(82) \times 10^{-11} m, representing the most probable distance between the proton and the in the ground state of the . Introduced by in his 1913 model of the , it provides the characteristic length scale for the size of atomic orbitals in hydrogen-like systems. In Bohr's semiclassical model, the radius arises from balancing the centripetal force on the electron with the electrostatic attraction to the nucleus, under the quantization condition that the electron's angular momentum is an integer multiple of \hbar, yielding a_0 = \frac{4\pi\epsilon_0 \hbar^2}{m_e e^2} for the ground state (n=1), where m_e is the electron mass, e is the elementary charge, \epsilon_0 is the vacuum permittivity, and \hbar is the reduced Planck's constant. This expression can also be written as a_0 = \frac{\hbar}{\alpha m_e c}, with \alpha the fine-structure constant and c the speed of light. For higher energy levels, the orbital radius scales as r_n = n^2 a_0, where n is the principal quantum number. In full quantum mechanics, the Schrödinger equation for the hydrogen atom confirms the Bohr radius as the natural unit of length, appearing in the ground-state wave function \psi_{100}(r) = \frac{1}{\sqrt{\pi a_0^3}} e^{-r/a_0}, where the radial probability distribution P(r) = r^2 |\psi|^2 reaches its maximum at r = a_0. Although the expectation value of r in this state is $1.5 a_0, the Bohr radius defines the scale for atomic dimensions and serves as the base unit of length (1 bohr) in the Hartree atomic unit system, facilitating calculations in quantum chemistry and atomic physics.

Definition and Interpretation

Mathematical Definition

The Bohr radius a_0, a fundamental length scale in , is defined in the (SI) by the expression a_0 = \frac{4\pi\epsilon_0 \hbar^2}{m_e e^2}, where \epsilon_0 is the (SI unit: F/m or C²/N·m²), characterizing the strength of electric interactions in vacuum; \hbar is the reduced Planck's constant (SI unit: J·s), representing the quantum of ; m_e is the rest mass of the electron (SI unit: kg), determining the inertial response of the electron; and e is the (SI unit: C), the magnitude of the electron's charge. In Gaussian cgs units, which do not include \epsilon_0 as the constant is absorbed into the charge definition, the expression simplifies to a_0 = \frac{\hbar^2}{m_e e^2}. This radius corresponds to the most probable distance from the to the in the wavefunction of the , as determined by maximizing the radial probability distribution P(r) = 4\pi r^2 |\psi_{100}(r)|^2, where \psi_{100}(r) is the 1s orbital.

Physical Significance

The Bohr radius serves as the characteristic length scale in the where the electrostatic attraction between the proton and balances the 's , preventing collapse into the while defining the atom's spatial extent. In the , this equilibrium occurs through the equality of and Coulomb attraction, but in , it manifests via the applied to the , where the expectation value of the equals half the magnitude of the . This balance yields a stable configuration, with the total minimized at the Bohr radius, highlighting its role as the natural size parameter for atomic bound states. It approximates the "size" of the , corresponding closely to the most probable electron-nucleus distance and serving as a proxy for the radial extent in the wavefunction. Specifically, the expectation value of the electron-proton separation in the 1s orbital is \langle r \rangle = \frac{3}{2} a_0, where a_0 is the Bohr radius, indicating that the average position lies somewhat beyond the most probable one but still on the order of a_0. This makes the Bohr radius a fundamental measure of atomic dimensions, influencing phenomena from spectral lines to chemical bonding. Conceptually, the Bohr radius emerges as the unique length scale from combining —through Planck's reduced constant \hbar—with —via the electron charge e, mass m_e, and \epsilon_0. Dimensional analysis of these constants yields a_0 = \frac{4\pi\epsilon_0 \hbar^2}{m_e e^2}, the sole combination with units of length, analogous to how the c = \frac{1}{\sqrt{\mu_0 \epsilon_0}} arises as the unique velocity from purely electromagnetic constants in . This synthesis underscores the Bohr radius's foundational status in , bridging classical forces with quantum discreteness.

Historical Context

Origin in Bohr's Model

In 1911, proposed a model of the atom featuring a dense, positively charged surrounded by electrons in planetary-like orbits, based on his foil experiments. However, this classical model faced a critical instability: accelerating electrons in circular orbits would continuously radiate electromagnetic energy according to , causing them to spiral inward and collapse into the , contradicting the observed stability of atoms. To address these shortcomings and explain the discrete spectral lines of , such as the observed in 1885, introduced a revolutionary atomic model in a series of three papers published in 1913. In his first paper, "On the Constitution of Atoms and Molecules," Bohr postulated that electrons occupy discrete "stationary states" where they do not radiate energy despite accelerating, with transitions between states emitting or absorbing radiation at specific frequencies corresponding to the observed spectral lines. Central to this was his quantization condition for in these circular orbits: L = n \frac{h}{2\pi}, where n is a positive and h is Planck's constant, ensuring only certain orbits are allowed. For the , Bohr balanced the required for with the electrostatic attraction between the and proton, incorporating the quantization postulate. This yielded the orbital radius for the (n=1) as r_1 = \frac{4\pi\epsilon_0 [\hbar](/page/H-bar)^2}{m_e e^2}, where \hbar = h / 2\pi, m_e is the , and e is the ; this characteristic length scale is now known as the Bohr radius in honor of Bohr's foundational contribution. Subsequent papers refined the model for multi-electron systems while maintaining the core ideas for .

Evolution in Quantum Theory

The transition from Niels Bohr's semiclassical model to modern began with Werner Heisenberg's formulation of in 1925, where the Bohr radius a_0 retained its role as a fundamental scale in quantizing the atom's levels, now expressed through non-commuting dynamical matrices rather than classical orbits. This approach, developed further by and , reproduced the discrete spectrum of with a_0 emerging naturally from the correspondence principle, bridging the to a fully operational without explicit trajectories. Erwin Schrödinger's wave mechanics, introduced in 1926, provided a more intuitive reinterpretation by solving the time-independent for the , yielding radial wavefunctions scaled by a_0, the Bohr radius, which defines the extent of the electron's . In the (n=1, l=0), the exact solution gives the expectation value of the radial position as \langle r \rangle = \frac{3}{2} a_0, highlighting a_0's significance as the characteristic size beyond which the wavefunction decays exponentially. The incorporation of via Dirac's in 1928 refined this picture for high-speed electrons, introducing and yielding energy levels that include fine-structure corrections, which slightly contract the effective Bohr radius by a factor involving the \alpha \approx 1/137. In the non-relativistic limit, the Dirac hydrogen wavefunctions approach the Schrödinger solutions, but relativistic effects reduce the mean radius by about \alpha^2 / 2, or roughly 0.0027%. Throughout the , further refinements accounted for the finite nuclear mass by replacing the m_e with the \mu = m_e M / (m_e + M), where M is the proton mass, adjusting the Bohr radius to a_0' = a_0 (m_e / \mu) \approx 1.00054 a_0 for and altering expectation values accordingly. This correction, essential for precision , was integrated into quantum mechanical treatments starting from Schrödinger's framework and remains standard in atomic calculations.

Derivation and Calculation

Derivation from Bohr-Sommerfeld Model

The Bohr-Sommerfeld model applies semiclassical quantization rules to the motion of an in the Coulomb potential of the , extending the circular orbits of the original to include elliptical paths. The key quantization condition for the radial motion is given by the action integral \oint p_r \, dr = n_r h, where p_r is the radial component of the electron momentum, n_r is the radial (a non-negative ), and h is Planck's constant. This condition, combined with the angular quantization \oint p_\phi \, d\phi = k h, where k is the (a positive ), ensures levels and orbit sizes. For circular orbits, which correspond to the case of zero radial excursion (n_r = 0), the radial quantization is automatically satisfied, and the model reduces to balancing the centripetal force with the electrostatic attraction: \frac{m_e v^2}{r} = \frac{e^2}{4\pi \epsilon_0 r^2}, where m_e is the electron mass, v is the orbital speed, e is the elementary charge, \epsilon_0 is the vacuum permittivity, and r is the orbital radius. The angular momentum quantization provides m_e v r = k \hbar, with \hbar = h / 2\pi. Substituting v = k \hbar / (m_e r) into the force balance equation yields the orbital radius r_k = k^2 \frac{4\pi \epsilon_0 \hbar^2}{m_e e^2}. For the ground state (k=1), this defines the Bohr radius a_0 = \frac{4\pi \epsilon_0 \hbar^2}{m_e e^2}. In the general case of elliptical orbits, the effective radial motion in the potential leads to a semi-major axis scaling as r_n = n^2 a_0, where n = n_r + k is the principal . This preserves a_0 as the fundamental length scale for the . Arnold Sommerfeld's 1916 extension incorporated relativistic effects to account for in spectral lines, introducing a of the elliptical orbits while maintaining the non-relativistic base scale a_0 for the orbit dimensions.

Modern Quantum Mechanical Approach

In the modern quantum mechanical treatment, the Bohr radius emerges as a fundamental length scale in the exact solution to the time-independent Schrödinger equation for the hydrogen atom, which describes the two-body system of a proton and an electron in the center-of-mass frame using the reduced mass approximated as the electron mass m_e. The governing equation is -\frac{\hbar^2}{2m_e} \nabla^2 \psi(\mathbf{r}) - \frac{e^2}{4\pi \epsilon_0 r} \psi(\mathbf{r}) = E \psi(\mathbf{r}), where \psi(\mathbf{r}) is the wave function, E is the energy eigenvalue, \hbar is the reduced Planck's constant, e is the elementary charge, and \epsilon_0 is the vacuum permittivity. Due to the spherical symmetry of the Coulomb potential, the equation is solved by separation of variables in spherical coordinates: \psi_{nlm}(r, \theta, \phi) = R_{nl}(r) Y_{lm}(\theta, \phi), where Y_{lm} are spherical harmonics and R_{nl} satisfies the radial equation -\frac{\hbar^2}{2m_e} \frac{d^2 u_{nl}}{dr^2} + \left[ \frac{l(l+1) \hbar^2}{2m_e r^2} - \frac{e^2}{4\pi \epsilon_0 r} \right] u_{nl}(r) = E_{nl} u_{nl}(r), with u_{nl}(r) = r R_{nl}(r) and the effective potential comprising the centrifugal term and the attractive Coulomb potential. For the ground state (n=1, l=0, m=0), the solution yields the 1s orbital wave function \psi_{100}(r) = \frac{1}{\sqrt{\pi a_0^3}} e^{-r/a_0}, where the Bohr radius a_0 = \frac{4\pi \epsilon_0 \hbar^2}{m_e e^2} naturally sets the scale for the exponential decay, ensuring normalization and matching the boundary conditions at infinity and the origin. From this wave function, the expectation value of the radial position is \langle r \rangle = \int_0^\infty r \cdot 4\pi r^2 |\psi_{100}|^2 dr = \frac{3}{2} a_0, representing the average distance of the electron from the nucleus in the ground state. The most probable radius, found by maximizing the radial probability density $4\pi r^2 |\psi_{100}|^2, occurs at r_{\rm mp} = a_0.

Numerical Value and Precision

Exact Expression in Terms of Fundamental Constants

The Bohr radius a_0 is expressed exactly in the (SI) as a_0 = \frac{4\pi\epsilon_0 \hbar^2}{m_e e^2}, where \epsilon_0 is the , \hbar is the reduced , m_e is the , and e is the . In the current SI framework, following the 2019 revision, e, \hbar (derived from the h), and the c (used implicitly in related definitions) are defined as exact values: e = 1.602176634 \times 10^{-19} C, h = 6.62607015 \times 10^{-34} J s, and c = 299792458 m/s, while \epsilon_0 = 8.8541878128 \times 10^{-12} F/m is exactly determined from these; m_e, however, retains a relative standard uncertainty of approximately $3.1 \times 10^{-10}. This formulation arises directly from balancing electrostatic and centripetal forces in the , scaled by quantum conditions, but remains the precise symbolic representation in for the infinite nuclear mass case. An equivalent form expresses a_0 using the fine-structure constant \alpha \approx 7.2973525643 \times 10^{-3}, which encapsulates the strength of electromagnetic interactions: a_0 = \frac{\hbar}{m_e c \alpha}. Here, \alpha = \frac{e^2}{4\pi \epsilon_0 \hbar c} is a dimensionless constant derived from the fundamental quantities above, with its 2022 CODATA value carrying the uncertainty primarily from m_e. This representation highlights the role of relativistic scales, as a_0 emerges as the Compton wavelength of the electron divided by $2\pi \alpha./19%3A_Atoms/19.03%3A_The_Hydrogen_Atom) Another alternative formulation relates a_0 to the Rydberg constant for infinite nuclear mass, R_\infty = \frac{m_e e^4}{8 \epsilon_0^2 h^3 c}, which defines the scale of hydrogen spectral lines: a_0 = \frac{\alpha}{4\pi R_\infty}. The constant R_\infty itself depends on the same fundamental parameters, with its exact relation to a_0 following from the ground-state energy E_1 = -\frac{1}{2} m_e c^2 \alpha^2 = -h c R_\infty. This form is particularly useful in spectroscopy, where R_\infty is measured precisely. For real atomic systems with finite nuclear mass, the effective Bohr radius incorporates the reduced mass \mu = \frac{m_e m_p}{m_e + m_p} of the electron-proton system (with m_p the proton mass), yielding a_H = \frac{4\pi\epsilon_0 \hbar^2}{\mu e^2} = a_0 \frac{m_e}{\mu} \approx a_0 \left(1 + \frac{m_e}{m_p}\right), where the correction factor is approximately 1.000545, enlarging the radius slightly compared to the infinite-mass a_0./07%3A_Atomic_Spectroscopy/7.04%3A_The_Bohr_Model_of_Hydrogen-like_Atoms) This variant accounts for the nucleus's motion but preserves the exact infinite-mass form as the foundational constant. Dimensional analysis confirms a_0 as the unique length scale constructible from the fundamental constants \epsilon_0 (dimensions [M^{-1} L^{-3} T^4 I^2]), [\hbar](/page/H-bar) ([M L^2 T^{-1}]), m_e ([M]), and e ([I T]), with the combination yielding dimensions of [L] exclusively through the given expression—no other independent length emerges from these quantities alone. The CODATA 2022 recommended value for the Bohr radius is a_0 = 5.29177210544(82) \times 10^{-11} m, where the uncertainty represents the standard deviation at the 1σ level and corresponds to a relative standard uncertainty of $1.55 \times 10^{-10}. This numerical value is determined through theoretical computation from fundamental physical constants, such as the reduced , the , the , and the , rather than direct measurement, with refinements arising from improved precision in these input constants (e.g., the ). The 2019 redefinition of the SI units fixed exact values for the h, e, and c, rendering derived quantities like the reduced Planck constant \hbar and \epsilon_0 exact as well; consequently, the Bohr radius's uncertainty now primarily stems from measurements of the m_e and \alpha, though the overall value has remained stable relative to the previous CODATA evaluation. In practical units, this corresponds to approximately 0.529177210544 Å or 52917.7210544 fm.

Applications in Atomic Systems

Role in Hydrogen Atom Ground State

In the ground state of the hydrogen atom, the Bohr radius a_0 serves as the fundamental length scale governing the spatial distribution of the electron's probability density. The ground-state wave function, obtained from the Schrödinger equation, is \psi_{100}(r) = \frac{1}{\sqrt{\pi a_0^3}} e^{-r/a_0}, where the exponential decay is parameterized by a_0 \approx 5.29 \times 10^{-11} m. This results in a spherically symmetric electron cloud, with the probability density |\psi|^2 highest near the nucleus but extending indefinitely, reflecting the quantum delocalization of the electron. The root-mean-square radius, defined as \sqrt{\langle r^2 \rangle}, quantifies the typical extent of this distribution and equals \sqrt{3} a_0 \approx 9.17 \times 10^{-11} m. Similarly, the expectation value of the radius is \langle r \rangle = \frac{3}{2} a_0 \approx 7.94 \times 10^{-11} m, both exceeding the most probable radius of a_0 due to the asymmetric tail of the radial probability distribution P(r) = 4\pi r^2 |\psi|^2. The cumulative probability of locating the within a sphere of radius a_0 is given by integrating the radial probability up to r = a_0, yielding P(r < a_0) = 1 - (1 + 2 \frac{r}{a_0}) e^{-2 r / a_0} \big|_{r=a_0} = 1 - 3 e^{-2} \approx 0.594, or 59.4%. This means nearly 41% of the probability lies beyond a_0, underscoring the delocalized character of the ground-state electron wave function over scales comparable to and larger than the Bohr radius. The radius enclosing 50% probability is approximately $0.85 a_0, further illustrating that the electron is not confined to a classical orbit but spreads out probabilistically. The ground-state energy is E_1 = -13.6 eV, corresponding to the ionization energy required to remove the electron from this delocalized state at the a_0 scale. This energy arises from the balance of kinetic and potential terms in the Hamiltonian, scaled by a_0 via E_n = -\frac{1}{2} m_e (\alpha c / n)^2 for principal quantum number n=1, where \alpha is the fine-structure constant. Experimentally, this value and the associated size scale are verified through high-precision spectroscopy of hydrogen's emission lines, which match the theoretical predictions derived from the , as well as X-ray scattering experiments on hydrogen-like ions that probe the electron density distribution.

Extension to Hydrogen-like Ions

The Bohr model readily extends to hydrogen-like ions, which consist of a nucleus with atomic number Z > 1 and a single orbiting , such as singly ionized (He^+, Z=2) and doubly ionized (Li^{2+}, Z=3). In these systems, the electron experiences a attraction proportional to Z, leading to a contraction of the orbital radius compared to neutral . The ground-state radius, known as the generalized Bohr radius a_Z, is given by a_Z = \frac{a_0}{Z}, where a_0 is the Bohr radius for . This inverse scaling reflects the increased nuclear charge pulling the closer to the nucleus. The energy levels in hydrogen-like ions follow a similar scaling, with the principal quantum number n determining the states. The energy is E_n = -\frac{Z^2 E_1}{n^2}, where E_1 is the ground-state energy for (approximately -13.6 ). This quadratic dependence on Z arises from the stronger potential, which deepens the and increases the by Z^2. Consequently, the orbital size remains inversely proportional to Z, as the balance between and electrostatic attraction in the dictates a radius that shrinks linearly with the effective charge. In deriving these expressions, the electron mass is replaced by the reduced mass \mu to account for the nuclear motion. For hydrogen-like ions with heavy nuclei (M \gg m_e), \mu \approx m_e, providing an excellent approximation since the nucleus is nearly stationary. The precise reduced mass is \mu = \frac{m_e M}{m_e + M} \approx m_e \left(1 - \frac{m_e}{M}\right), and the corresponding Bohr radius becomes a_Z^\mu \approx \frac{a_0}{Z} \left(1 + \frac{m_e}{M}\right), where the correction term arises because the effective mass \mu < m_e leads to a slightly larger radius than the infinite-mass approximation. This adjustment is minimal for high-Z ions due to their larger nuclear masses. Hydrogen-like ions with high Z enable precision measurements that test (QED) in strong fields, where the small a_Z brings relativistic and QED corrections to the forefront. For instance, g-factor measurements of hydrogen-like tin (^{49+}) have achieved 0.5 parts-per-billion precision, allowing stringent tests of QED contributions beyond the Dirac-Coulomb framework. These experiments, often using electron beam ion traps, validate theoretical predictions for QED effects in highly charged systems.

Connection to Atomic Units

In the atomic unit system, also known as Hartree atomic units, the Bohr radius a_0 serves as the fundamental unit of length, set equal to 1, alongside the electron mass m_e = 1, elementary charge e = 1, and reduced Planck's constant \hbar = 1. This choice of units eliminates the need for explicit fundamental constants in many equations of atomic physics, simplifying theoretical and numerical work. With these conventions, the time-independent for the reduces to a dimensionless form: -\frac{1}{2} \nabla^2 [\psi](/page/Psi) - \frac{1}{[r](/page/R)} [\psi](/page/Psi) = E [\psi](/page/Psi), where distances are measured in units of a_0, energies in s (E_h), and the potential represents the attraction between the proton and . The (n=1) in these units is \psi_{100}(r) = \pi^{-1/2} e^{-[r](/page/R)}, with the corresponding energy eigenvalue E = -1/2 ./01%3A_Chapters/1.07%3A_Hydrogen_Atom) The primary advantages of atomic units lie in their facilitation of dimensionless formulations, which streamline computations in and by reducing numerical errors and avoiding conversions between disparate scales. This system is particularly beneficial in methods like Hartree-Fock self-consistent field calculations, where integrals over atomic orbitals are evaluated more efficiently without scaling factors. An extension of this framework appears in Rydberg atomic units, where a_0 remains the base length unit, but the energy scale is halved to the Rydberg (Ry = E_h / 2), aligning the hydrogen ground state energy at -1 Ry for spectroscopic applications.

Comparisons with Other Length Scales

The Bohr radius a_0, with a value of approximately $5.291772 \times 10^{-11} m, serves as a scale for orbitals, particularly in the . In comparison, the r_e = \frac{e^2}{4\pi\epsilon_0 m_e c^2} \approx 2.818 \times 10^{-15} m represents a much smaller scale, equivalent to a_0 \alpha^2, where \alpha \approx 1/137 is the . This relation highlights the point-like nature of the in classical electrodynamics versus the extended orbital size in quantum models, with r_e being about 18,800 times smaller than a_0. Another key relativistic length scale is the Compton wavelength of the electron, \lambda_C = \frac{h}{m_e c} \approx 2.426 \times 10^{-12} m, which is the wavelength associated with the electron's rest mass energy and marks the onset of quantum field effects. This is approximately $0.046 a_0, or equivalently $2\pi a_0 \alpha, making it smaller than the Bohr radius but larger than the classical electron radius by a factor of about 860. The reduced Compton wavelength \bar{\lambda}_C = \frac{\hbar}{m_e c} \approx 3.862 \times 10^{-13} m is even smaller, at roughly a_0 \alpha \approx a_0 / 137, emphasizing how non-relativistic quantum mechanics applies at the atomic scale while relativistic corrections become significant near the Compton scale. On the nuclear scale, typical radii follow the semi-empirical formula R \approx 1.2 A^{1/3} fm, where A is the and 1 fm = $10^{-15} m, yielding sizes of 1 to 10 fm for light to heavy nuclei. This places nuclear dimensions about $10^4 to $10^5 times smaller than a_0, underscoring the vast separation between strong interactions and electromagnetic atomic binding, which prevents electron wavefunctions from probing directly. In , lattice constants of common solids range from 2 to 6 (e.g., 3.61 for , 5.43 for ), corresponding to roughly 4 to 11 times a_0 since a_0 \approx 0.529 . This comparability illustrates how orbitals overlap in crystals to form energy bands, a foundational concept in solid-state theory where the Bohr radius provides the natural unit for interatomic distances and delocalization.

References

  1. [1]
    Bohr radius - CODATA Value
    Click symbol for equation. Bohr radius $a_0$. Numerical value, 5.291 772 105 44 x 10-11 m. Standard uncertainty, 0.000 000 000 82 x 10-11 m.
  2. [2]
    2.5: The Shape of Atomic Orbitals - Chemistry LibreTexts
    Aug 30, 2022 · For the hydrogen atom, the peak in the radial probability plot occurs at r = 0.529 Å (52.9 pm), which is exactly the radius calculated by Bohr ...
  3. [3]
    Niels Bohr's First 1913 Paper: Still Relevant, Still ... - AIP Publishing
    Nov 1, 2018 · The theory of the hydrogen atom is essentially already complete at this point in the paper. The quantization of orbital angular momentum is ...
  4. [4]
    6.4 Bohr's Model of the Hydrogen Atom - University Physics Volume 3
    Sep 29, 2016 · Bohr's model of the hydrogen atom, proposed by Niels Bohr in 1913, was the first quantum model that correctly explained the hydrogen emission ...
  5. [5]
    Bohr radius - CODATA Value
    Bohr radius $ a_0 = \hbar/\alpha m_{\rm e}c = ; Numerical value, 5.291 772 105 44 x 10-11 m ; Standard uncertainty, 0.000 000 000 82 x 10-11 m.
  6. [6]
    Fine Structure of Hydrogen Energy Levels - Richard Fitzpatrick
    with $ {\cal E} = E/(m_e\,c^2 . Here, $ a_0=4\pi\,\epsilon_0\,\hbar^2 is the Bohr radius, and $ \alpha=e^2/(4\pi\,\epsilon_0 the fine structure constant.
  7. [7]
  8. [8]
    The Most Probable Radius Hydrogen Ground State - HyperPhysics
    It may seem a bit surprising that the average value of r is 1.5 x the first Bohr radius, which is the most probable value. The extended tail of the probability ...
  9. [9]
    [PDF] 6.4 Hydrogen atom - MIT
    May 4, 2009 · The Bohr radius is the radius of the orbit with the lowest energy (the ground state). The energy is a sum of kinetic and potential energy. This.Missing: significance | Show results with:significance
  10. [10]
    [PDF] Why atoms don't collapse Notes on General Chemistry
    Nov 19, 2004 · The balance of kinetic and potential energy in an atom is what keeps its electrons from collapsing into the nucleus. We can see how this ...
  11. [11]
    [PDF] Dimensional Analysis - UF Physics Department
    This is called the Bohr radius, or simply the bohr, because in the Bohr model it is the radius of the smallest orbit for an electron circling a fixed proton.
  12. [12]
    I. On the constitution of atoms and molecules - Taylor & Francis Online
    On the constitution of atoms and molecules. N. Bohr Dr. phil. Copenhagen. Pages 1-25 | Published online: 08 Apr 2009.
  13. [13]
    Quantum Mechanics (1925-1927)
    Between 1925 and 1927, Heisenberg participated prominently in the dis- covery, formulation and application of quantum mechanics. In this period he.
  14. [14]
    [PDF] 1926-Schrodinger.pdf
    The wave-equation and its application to the hydrogen atom. §5. The intrinsic reason for the appearance of discrete characteristic frequencies. §6. Other ...
  15. [15]
    [PDF] The Quantum Theory of the Electron - UCSD Math
    Oct 26, 2013 · By P. A. M. DIRAC, St. John's College, Cambridge. (Communicated by R. H. Fowler, F.R.S.-Received January 2, 1928.) The new quantum mechanics ...
  16. [16]
    [PDF] XXXVII. On the constitution of atoms and molecules - Zenodo
    Apr 8, 2009 · 1913. 2 K. Downloaded by [Duke University Libraries] at 11:53 30 July 2012. Page 5. 478. Dr. ~q. Bohr on the Constitution electrons in the rings ...
  17. [17]
    [PDF] {How Sommerfeld extended Bohr's model of the atom (1913–1916)}
    Jan 30, 2014 · With α → 0 the fine-structure formula became the simple formula of Bohr's model. In his Academy treatise Sommerfeld had cho- sen α = π2e4.
  18. [18]
  19. [19]
    [PDF] P. LeClair
    To find hri, we have to integrate rP(r) from 0 to ∞. Use the same substitution ρ = r/ao, dρ = dr/ao, the resulting integral is known.
  20. [20]
    [PDF] 2018 codata recommended values of the fundamental constants of ...
    2018 CODATA RECOMMENDED VALUES OF THE FUNDAMENTAL. CONSTANTS OF PHYSICS ... Bohr radius ¯h/(αmec) a0. 5.291 772 109 03(80) × 10−11 m. Bohr magneton e¯h ...
  21. [21]
    [PDF] 2022 codata recommended values of the fundamental constants of ...
    2022 CODATA RECOMMENDED VALUES OF THE FUNDAMENTAL. CONSTANTS OF PHYSICS ... Bohr radius ¯h/(αmec) a0. 5.291 772 105 44(82) × 10−11 m. Bohr magneton e¯h ...
  22. [22]
    7.4: The Bohr Model of Hydrogen-like Atoms - Physics LibreTexts
    Mar 5, 2022 · The quantity represented by the symbol μ is called the reduced mass of the electron. It is slightly less than the actual mass of the electron.Missing: correction | Show results with:correction
  23. [23]
    Energy Levels of Hydrogen Atom - Richard Fitzpatrick
    Here, $ \mu = m_e \,m_p/(m_e+ m_p)\simeq m_e is the reduced mass, which takes into account the fact that the electron (of mass $ m_e$ ) ...<|control11|><|separator|>
  24. [24]
    Stringent test of QED with hydrogen-like tin - PMC - PubMed Central
    Here we report on our high-precision, high-field test of QED in hydrogen-like 118Sn49+. The highly charged ions were produced with the Heidelberg electron beam ...
  25. [25]
    Towards Precision Spectroscopy of Antiprotonic Atoms for Probing ...
    Jan 15, 2025 · PAX (antiProtonic Atom X-ray spectroscopy) is a new experiment with the aim to test strong-field quantum electrodynamics (QED) effects by ...
  26. [26]
    atomic unit of length - CODATA Value
    Click symbol for equation. atomic unit of length $a_0$. Numerical value, 5.291 772 105 44 x 10-11 m. Standard uncertainty, 0.000 000 000 82 x 10-11 m.
  27. [27]
    atomic units (A00504) - IUPAC Gold Book
    The advantage of atomic units is that if all calculations are directly expressed in such units, the results do not vary with any revision of the numerical ...
  28. [28]
    [PDF] 23. The Hydrogen Atom - Physics
    With the understanding that all distances are measured in units of a0 and all energies are measured in units of Eh, the radial Schrödinger equation becomes. −.
  29. [29]
    Atomic Units (a.u.) and Conversion Factors - Wiley Online Library
    These units have many advantages, not least that they bring the electronic. Schrödinger equation to its intrinsically simplest form, expressed in pure numbers.
  30. [30]
    classical electron radius - CODATA Value
    classical electron radius $r_{\rm e}$ ; Numerical value, 2.817 940 3205 x 10-15 m ; Standard uncertainty, 0.000 000 0013 x 10-15 m.
  31. [31]
    Compton wavelength† - CODATA Value
    Compton wavelength† $\lambda_{\rm C}$. Numerical value, 2.426 310 235 38 x 10-12 m. Standard uncertainty, 0.000 000 000 76 x 10-12 m.
  32. [32]
  33. [33]
    Nuclear Units - HyperPhysics
    Krane comments that the evidence points to a mass radius and a charge radius which agree with each other within about 0.1 fermi. Since heavy nuclei have ...
  34. [34]
    Lattice Constants for all the elements in the Periodic Table
    Sodium. 429.06, 429.06, 429.06 pm ; Magnesium. 320.94, 320.94, 521.08 pm ; Aluminum. 404.95, 404.95, 404.95 pm ; Silicon. 543.09, 543.09, 543.09 pm ; Phosphorus ...
  35. [35]
    Lattice Constant - an overview | ScienceDirect Topics
    The maximum number of lattice constants for a unit cell can be three. These are, in general, on the order of a few angstroms and are experimentally determined ...