Fact-checked by Grok 2 weeks ago

Hydrogen atom

The hydrogen atom is the simplest and most fundamental unit of matter in the chemical elements, defined by an atomic number of and consisting of a single proton in the bound to a single , with no neutrons in its most common , protium (¹H). This neutral system has a of [1.00784, 1.00811] u, reflecting the dominance of ¹H, which accounts for 99.9885% of natural hydrogen abundance. In quantum mechanics, the hydrogen atom exemplifies an exactly solvable through the time-independent , yielding discrete energy levels that depend solely on the principal quantum number n and explain its characteristic atomic spectrum, including the of visible emission lines. The energy is -13.6 eV, corresponding to the electron's in the 1s orbital, while excited states lead to radiative transitions observed in astrophysical and laboratory settings. As the most abundant element, hydrogen comprises about 73.5% of the baryonic mass in the , primarily in atomic and molecular forms within , gas, and , where it drives processes that power . Recent observations as of 2025 have identified much of the previously unaccounted-for hydrogen as diffuse ionized gas surrounding galaxies. In , its small size and enable it to form the basis of covalent bonds, hydrides, and acids, while isotopes like (²H) and (³H) extend its applications in reactions and .

Basic Properties

Isotopes of Hydrogen

The hydrogen atom consists of three primary isotopes, distinguished by the number of neutrons in their nuclei, which significantly influences their and . Protium, denoted as ^[1](/page/1)\mathrm{H}, is the most abundant isotope, comprising a single proton and no neutrons in its . It accounts for approximately 99.98% of naturally occurring hydrogen atoms and is stable, with no measurable . As a single-proton , protium has a of 0 MeV, reflecting the absence of neutron-proton interactions to bind. Deuterium, or ^2\mathrm{H} (symbol D), features one proton and one neutron, making it the heaviest stable isotope of hydrogen. Its natural abundance is about 0.0156%, equivalent to one deuterium atom per roughly 6,420 hydrogen atoms in seawater. The nucleus is bound by an energy of 2.224 MeV, providing sufficient stability against dissociation. This doubled nuclear mass compared to protium alters the reduced mass in the hydrogen atom, leading to subtle shifts in atomic energy levels and spectral lines; specifically, deuterium's emission lines appear at slightly higher energies (shorter wavelengths) due to the increased reduced mass, with shifts on the order of 0.04% relative to protium. Tritium, denoted ^3\mathrm{H} (symbol T), contains one proton and two neutrons, resulting in a nuclear binding energy of 8.482 MeV, which is higher than that of deuterium but insufficient for long-term stability. It is radioactive, decaying via beta emission to with a half-life of 12.323 years. Tritium occurs only in trace amounts in nature, primarily from interactions, at abundances of approximately 10^{-18} relative to total . The additional neutron increases the nuclear mass threefold over protium, further modifying the reduced mass and causing even smaller spectral shifts compared to deuterium, though its rarity limits observational studies of neutral tritium atoms.
IsotopeSymbolNuclear CompositionNatural Abundance (atom %)StabilityBinding Energy (MeV)
Protium^1\mathrm{H}1 proton, 0 neutrons99.98Stable0
Deuterium^2\mathrm{H} (D)1 proton, 1 neutron0.0156Stable2.224
Tritium^3\mathrm{H} (T)1 proton, 2 neutrons≈ 10^{-16}Radioactive (half-life 12.323 y)8.482

Hydrogen Ion

The , denoted as H⁺, is the simplest atomic , consisting solely of a single proton with no bound electrons. This bare carries a +1 and has a mass of approximately 1.00784 atomic mass units. In , H⁺ represents the fully ionized state of the hydrogen atom, exhibiting high reactivity due to its positive charge and lack of electronic shielding. The H⁺ ion forms through the of a neutral atom, where the single is removed, requiring an of 13.59844 . This process is endothermic and occurs in high-energy environments such as plasmas or stellar atmospheres, making H⁺ prevalent in astrophysical contexts like the . Historically, the proton's identity as the was confirmed by in through experiments bombarding gas with alpha particles from , which produced hydrogen nuclei identifiable by their tracks; Rutherford named this particle the "proton" in a seminal paper published that year. In chemical contexts, particularly aqueous solutions, the bare H⁺ ion is unstable and rapidly associates with water molecules to form the hydronium ion, H₃O⁺, which serves as the solvated . This species results from of (H₂O + H⁺ → H₃O⁺) and is the active form in acid-base chemistry, influencing and facilitating reactions like . The hydronium ion's stability arises from hydrogen bonding in , preventing the bare proton's high mobility and reactivity. Isotopic variants of the hydrogen ion, such as the deuteron D⁺ (from ) or triton T⁺ (from ), exhibit behavioral differences primarily due to mass effects on reactivity. For instance, D⁺, with twice the mass of H⁺, experiences kinetic isotope effects in proton-transfer reactions, leading to slower reaction rates and altered in electrochemical processes compared to H⁺. These differences parallel those observed in neutral isotopes, where heavier nuclei influence vibrational frequencies and bond strengths.

Historical Models

Classical Electromagnetic Description

Prior to the development of nuclear models, J. J. Thomson proposed in 1904 that the atom consisted of a uniform sphere of positive charge with embedded s, akin to plums in a pudding, to maintain overall neutrality. This "plum pudding" model accounted for the discovery of the but failed to explain experiments. In 1911, Ernest Rutherford's gold foil experiments revealed that alpha particles were deflected at large angles, indicating a tiny, dense, positively charged at the atom's center surrounded by orbiting s, forming a planetary model of the atom. In this classical , the hydrogen atom features a single in a around the proton , analogous to a miniature solar system governed by Newtonian mechanics and . However, the centripetal of the orbiting implies it should radiate electromagnetic energy continuously, as any accelerating charge does according to classical electrodynamics. The quantifies this radiated power for a non-relativistic point charge: P = \frac{2}{3} \frac{e^2 a^2}{c^3} in cgs units, where e is the electron charge, a is the , and c is the . This energy loss causes the orbit to spiral inward, leading to the electron's collapse onto the in approximately $10^{-11} seconds, rendering the model unstable and incompatible with the observed persistence of atoms. Classical theory also struggled to explain the spectral lines of hydrogen observed by Johann Balmer in , such as the visible series fitting an relating wavelengths to integers. In the continuous orbital model, electron transitions would produce a of frequencies rather than lines, highlighting the inadequacy of purely without additional constraints. These shortcomings prompted the search for quantized descriptions to stabilize the atom and account for spectra.

Bohr-Sommerfeld Model

In 1913, proposed a semi-classical model for the hydrogen atom that incorporated quantum ideas to address the classical electromagnetic paradoxes, such as the instability of orbits predicted by Rutherford's model. In this model, the is assumed to move in stable circular orbits around the proton, with the key postulate that the L is quantized in discrete units: L = n \hbar, where n is a positive known as the principal and \hbar = h / 2\pi is the reduced Planck's constant. The orbital radius and energy derive from balancing the classical required for against the attraction between the and proton. The centripetal force equation is \frac{m v^2}{r} = \frac{k e^2}{r^2}, where m is the , v is its speed, r is the radius, e is the , and k = 1/(4\pi\epsilon_0) is Coulomb's constant. Combining this with the quantization condition m v r = n \hbar yields the for the (n=1) as approximately 0.529 Å and discrete energy levels given by E_n = -\frac{13.6 \, \mathrm{eV}}{n^2}, which accurately match the observed lines of upon transitions between levels. In 1916, Arnold Sommerfeld extended Bohr's model to allow for elliptical orbits, generalizing the quantization by applying action-angle variables from classical mechanics. This introduced a second quantum condition on the radial action integral, leading to an additional azimuthal quantum number k (ranging from 1 to n) that determines the eccentricity of the ellipse, with circular orbits corresponding to k = n. Sommerfeld also incorporated special relativistic corrections, accounting for the variation in electron speed along the elliptical path, which causes the orbit to precess and splits the energy levels into fine structure components. This qualitative explanation of the fine structure—small deviations in spectral line positions—provided better agreement with experimental hydrogen spectra than Bohr's original circular-orbit assumption. Despite these advances, the Bohr-Sommerfeld model has significant limitations, as it relies on quantization rules without incorporating the wave nature of and fails to describe atoms with more than one due to unaccounted inter-electron interactions.

Quantum Mechanical Treatment

Schrödinger Equation Formulation

The quantum mechanical description of the hydrogen atom begins with the non-relativistic of a proton and an interacting via the potential. To simplify this, the system is transformed into an equivalent one-body problem using the center-of-mass frame, where the relative motion is governed by the \mu = \frac{m_e m_p}{m_e + m_p}, with m_e the and m_p the proton mass; since m_p \gg m_e, \mu \approx m_e to a high degree of accuracy. The Hamiltonian for this effective one-body system is H = \frac{\mathbf{p}^2}{2\mu} - \frac{e^2}{4\pi\epsilon_0 r}, where \mathbf{p} is the , r = |\mathbf{r}| is the radial distance between the particles, e is the , and \epsilon_0 is the ; this form captures the of the reduced particle and the attractive potential. The time-independent is then H \psi(\mathbf{r}) = E \psi(\mathbf{r}), where \psi(\mathbf{r}) is the wave function and E is the eigenvalue, providing the foundation for finding stationary states of the atom. Given the spherical symmetry of the Coulomb potential, the equation is solved in spherical coordinates (r, \theta, \phi), where the wave function separates as \psi(r, \theta, \phi) = R(r) Y(\theta, \phi), with R(r) the radial part and Y(\theta, \phi) the angular part. The Laplacian operator in spherical coordinates is \nabla^2 = \frac{1}{r^2} \frac{\partial}{\partial r} \left( r^2 \frac{\partial}{\partial r} \right) + \frac{1}{r^2 \sin \theta} \frac{\partial}{\partial \theta} \left( \sin \theta \frac{\partial}{\partial \theta} \right) + \frac{1}{r^2 \sin^2 \theta} \frac{\partial^2}{\partial \phi^2}, leading to the separated radial equation involving \frac{d^2 R}{dr^2} + \frac{2}{r} \frac{dR}{dr} and the angular equation. The angular part corresponds to the operators for total L^2 Y = \ell(\ell+1) \hbar^2 Y and z-component L_z Y = m \hbar Y, where \ell and m are quantum numbers, with Y(\theta, \phi) expressed as ; this separation exploits the commutativity of L^2 and L_z with the . For bound states, the energy E < 0 ensures square-integrable wave functions that vanish as r \to \infty, imposing quantization through boundary conditions on the radial function, such as R(r) \to 0 as r \to 0 and exponential decay at large r.

Wavefunctions and Quantum Numbers

The solutions to the time-independent Schrödinger equation for the hydrogen atom are stationary states described by wavefunctions ψ_{n l m_l}(r, θ, φ) that depend on three spatial quantum numbers: the principal quantum number n = 1, 2, 3, ..., which determines the size and energy scale of the orbital; the orbital angular momentum quantum number l = 0, 1, ..., n-1, which specifies the orbital's angular momentum magnitude; and the magnetic quantum number m_l = -l, -l+1, ..., l, which describes the projection of the angular momentum along a chosen axis. Additionally, the electron's intrinsic spin introduces a fourth quantum number m_s = ±1/2, though it does not appear in the non-relativistic Schrödinger equation for the orbital wavefunction. These quantum numbers arise naturally from the separation of variables in spherical coordinates, ensuring the wavefunctions form a complete, orthonormal basis for the Hilbert space of the system. The angular dependence of the wavefunction is given by the spherical harmonics Y_{l}^{m_l}(θ, φ), which are eigenfunctions of the angular momentum operators and satisfy the associated differential equations on the unit sphere. These functions are complex-valued, with |Y_{l}^{m_l}|^2 providing the angular probability distribution, and they are normalized such that ∫ |Y_{l}^{m_l}|^2 dΩ = 1 over the solid angle dΩ = sinθ dθ dφ. The spherical harmonics encode the quantum mechanical analogs of classical orbital shapes, with l=0 corresponding to s-orbitals (spherically symmetric) and higher l to p, d, etc., orbitals. The radial part R_{n l}(r) of the wavefunction solves the radial Schrödinger equation and takes the form R_{n l}(r) = N e^{-ρ/2} ρ^l L_{n-l-1}^{2l+1}(ρ), where ρ = 2r/(n a_0), N is a normalization constant, a_0 is the (approximately 0.529 Å), and L_k^α(ρ) are associated , which are polynomial solutions to the ensuring finite behavior at the origin and exponential decay at infinity. For the ground state (n=1, l=0, m_l=0), the full wavefunction simplifies to ψ_{100}(r) = \frac{1}{\sqrt{\pi a_0^3}} e^{-r/a_0}, which is spherically symmetric and peaks at the . This explicit form was derived by solving the separated equations, confirming the wavefunction's normalization ∫ |ψ|^2 dV = 1 over all space. The hydrogen wavefunctions satisfy orthogonality relations: ∫ ψ_{n l m_l}^* ψ_{n' l' m_l'} dV = δ_{n n'} δ_{l l'} δ_{m_l m_l'}, where δ are , allowing them to serve as basis functions for expanding arbitrary states. This orthogonality stems from the self-adjoint nature of the Hamiltonian and the boundary conditions of the problem. According to the , the square modulus |ψ_{n l m_l}|^2 represents the probability density for finding the electron in a volume element dV, providing a statistical description of the electron's position rather than a classical trajectory.

Energy Levels and Eigenstates

The energy levels of the hydrogen atom, derived from the time-independent Schrödinger equation, are discrete and labeled by the principal quantum number n = 1, 2, 3, \dots. These bound states have energies given by E_n = -\frac{\mu e^4}{(4\pi \epsilon_0)^2 2 \hbar^2 n^2}, where \mu is the reduced mass of the electron-proton system, e is the elementary charge, \epsilon_0 is the vacuum permittivity, and \hbar is the reduced Planck's constant. This formula arises from separating the Schrödinger equation in spherical coordinates and solving the radial equation, yielding eigenvalues independent of the orbital angular momentum quantum number l and the magnetic quantum number m_l. For hydrogen-1 (protium), the proton mass greatly exceeds the electron mass, so \mu \approx m_e, the electron mass. In this infinite nuclear mass approximation, the energies simplify to E_n \approx -13.598 \, \text{eV} / n^2, with the ground state (n=1) at approximately -13.6 \, \text{eV}. The negative sign indicates bound states below the ionization threshold at E=0. This level structure explains the stability of the atom and the quantized nature of its excitation energies. The independence of E_n from l (where $0 \leq l < n) and m_l (-l \leq m_l \leq l) results in accidental degeneracy: each energy level n accommodates n^2 distinct eigenstates, corresponding to all possible combinations of l and m_l. These eigenstates are the familiar hydrogenic wavefunctions \psi_{n l m_l}(r, \theta, \phi), which define the spatial probability distributions for the electron. The spectroscopic implications of these levels are profound, as transitions between them produce the characteristic emission and absorption lines in the hydrogen spectrum. In the electric dipole approximation, the dominant transitions obey selection rules \Delta l = \pm 1 and \Delta m_l = 0, \pm 1, derived from the nonzero matrix elements of the dipole operator \mathbf{r} between states of opposite parity. These rules dictate allowed radiative decays, with forbidden transitions (\Delta l = 0, \pm 2) occurring at much weaker rates via higher-order multipoles. The wavelengths of allowed transitions follow the Rydberg formula: \frac{1}{\lambda} = R_\infty \left( \frac{1}{n_1^2} - \frac{1}{n_2^2} \right), where n_2 > n_1 are the upper and lower principal quantum numbers, and R_\infty = 10\,973\,731.568\,160(21) \, \text{m}^{-1} is the Rydberg constant for infinite nuclear mass (CODATA 2018). This empirical formula, originally proposed for series in alkali spectra and confirmed for hydrogen, groups transitions into named spectral series based on n_1: the Lyman series (n_1=1, ultraviolet, discovered 1906), Balmer series (n_1=2, visible, discovered 1885), and Paschen series (n_1=3, infrared, discovered 1908). For example, the Balmer-alpha line (n_2=3 \to n_1=2) appears at 656.3 nm, a prominent red line in stellar spectra. Isotope effects introduce subtle shifts in these levels and spectra due to variations in the \mu = m_e M / (m_e + M), where M is the nuclear mass. For (^2H, M \approx 2 m_p), \mu is slightly larger than for protium (^1H), leading to energy levels deeper by about 0.05% and corresponding blue-shifts in spectral lines (e.g., ~2.19 cm⁻¹ for Balmer-alpha). These shifts, first theoretically anticipated in the reduced-mass correction and experimentally verified in following deuterium's discovery, enable precise measurements and distinguish isotopic signatures in astrophysical observations.

Orbital Visualization

The shapes of hydrogen atom orbitals are determined by the angular part of the wavefunction and reflect the quantum mechanical of the . The s orbitals (l = 0) are spherically symmetric, with distributed evenly around the in a ball-like form that decreases radially outward. The p orbitals (l = 1) exhibit a , consisting of two lobes separated by a nodal plane through the , oriented along the x, y, or z axes depending on the m_l. Higher orbitals, such as d (l = 2), display more intricate cloverleaf or double- configurations with additional nodal planes, illustrating the increasing complexity as l increases. Nodal structures further characterize these orbitals, where nodes are regions of zero electron probability density. The number of radial nodes, which are spherical surfaces where the radial wavefunction vanishes, is given by n - l - 1, with n as the quantum number and l the . The number of nodes, corresponding to conical or planar surfaces due to the , equals l. For example, the 1s orbital (n=1, l=0) has no nodes, resulting in a smooth spherical ; the 2p orbital (n=2, l=1) has one node (a plane) and no radial nodes; while the 3d orbital (n=3, l=2) has two nodes and no radial nodes. These nodes arise from the oscillatory nature of the solutions to the radial and angular Schrödinger equations. Boundary surface plots provide a common visualization tool, typically depicting isosurfaces that enclose a specified probability, such as 90% of the total , to represent the orbital's extent. For the 1s orbital, cross-sections reveal an of probability from the , with no nodes interrupting the smooth distribution. In contrast, a cross-section of the 2p_z orbital along the z-axis shows two lobes separated by a nodal at the , with peaking away from the origin due to the angular dependence. These plots emphasize the volume where the is most likely found, rather than precise paths. Orbital visualizations represent time-averaged probability densities, |\psi|^2, integrated over all time, rather than instantaneous positions or classical trajectories, as the electron does not follow definite paths in . This averaging highlights the stationary nature of the states, where the remains constant despite the underlying wavefunction's evolution. For orbitals with l > 0, the non-zero probability density near the , despite the classical centrifugal barrier, is interpreted as quantum tunneling, allowing the electron to penetrate regions forbidden in a semiclassical picture.

Extensions and Alternatives

Relativistic Effects and Fine Structure

The non-relativistic for the hydrogen atom predicts energy levels that depend solely on the principal n, resulting in degeneracies for states with the same n but different orbital quantum numbers \ell. Observations of the hydrogen spectrum, however, reveal small splittings within these levels, collectively known as the , which arise from relativistic effects and electron spin. These corrections are of \alpha^2 times the gross energies, where \alpha \approx 1/137 is the , defined as \alpha = e^2 / (4\pi \epsilon_0 \hbar c). The provides the fundamental relativistic framework for describing the in the hydrogen atom's potential V(r) = -e^2 / (4\pi \epsilon_0 r). This first-order relativistic , i \hbar \partial \psi / \partial t = [c \vec{\alpha} \cdot \vec{p} + \beta m_e c^2 + V(r)] \psi, inherently includes the electron's spin as a 4-component and satisfies the Dirac equation for both positive and negative energy states. The exact bound-state solutions for the hydrogen atom were obtained shortly after Dirac's formulation, yielding energy eigenvalues that depend on n and the j, but not on \ell. These solutions introduce a k = \pm (j + 1/2), which distinguishes states with j = \ell \pm 1/2 and ensures the correct relativistic kinematics. The splitting is captured in the approximate derived from the Dirac eigenvalues, expanded to order \alpha^2: E_{n j} \approx E_n \left[ 1 + \frac{\alpha^2}{n^2} \left( \frac{n}{j + 1/2} - \frac{3}{4} \right) \right], where E_n = -\frac{13.6 \, \mathrm{[eV](/page/EV)}}{n^2} is the non-relativistic Bohr . This shows that levels with the same n but different j are split, with the shift scaling as \alpha^2 E_n / n. For example, in the n=2 manifold, the j=1/2 states lie below the j=3/2 states by an amount on the order of $10^{-4} \, \mathrm{[eV](/page/EV)}. A key component of the is the spin-orbit coupling, which emerges naturally in the as the interaction between the electron's spin magnetic moment \vec{\mu}_s = -g_e \mu_B \vec{S} / \hbar (with g_e \approx 2) and the effective \vec{B} produced by the electron's orbital motion in the nuclear \vec{E} = (Ze/r^2) \hat{r}. In the electron's , this field appears as \vec{B} = -(\vec{v} \times \vec{E}) / c^2, leading to an interaction H_{SO} = -\vec{\mu}_s \cdot \vec{B} / 2 (the factor of 1/2 accounts for ). The resulting energy shift is \Delta E_{SO} \propto \langle \vec{L} \cdot \vec{S} \rangle / r^3, which splits \ell levels according to j. This coupling contributes dominantly to the fine structure splitting observed in alkali-like atoms but is unified with other relativistic terms in hydrogen via Dirac theory. The relativistic reduced mass correction refines the treatment by accounting for the proton's finite mass m_p in a fully relativistic manner, beyond the non-relativistic \mu \approx m_e (1 - m_e / m_p). In the Dirac framework, this introduces recoil effects of order (m_e / m_p) \alpha^2 E_n, shifting all levels downward by approximately - (m_e / m_p) E_n \alpha^2 / n^2 and slightly modifying the splittings. For , this correction is small, about 0.05% of the gross energy, but essential for precision spectroscopy. Although the Dirac equation successfully explains most fine structure, it predicts degeneracy between the $2S_{1/2} and $2P_{1/2} states for n=2. In 1947, Lamb and Retherford measured an anomalous splitting of 1057.8 MHz (about $4.37 \times 10^{-6} \, \mathrm{eV}) between these levels using microwave excitation in a beam of excited hydrogen atoms, revealing a discrepancy that served as a precursor to quantum electrodynamic refinements.

Quantum Electrodynamics Approach

Quantum electrodynamics (QED), the relativistic of , describes the hydrogen atom as a of an and proton interacting through the exchange of virtual photons, incorporating all-order radiative corrections to achieve unprecedented precision in atomic spectra. This framework extends the non-relativistic Schrödinger and relativistic Dirac treatments by including quantum fluctuations of the , such as virtual electron-positron pairs and photon loops, which modify the Coulomb potential and electron self-interactions. A hallmark of the QED approach is the , the energy splitting between the otherwise degenerate 2S_{1/2} and 2P_{1/2} states, calculated to be 1057.845(9) MHz through contributions from (negative shift of about -27 MHz) and electron self-energy (positive dominant term of about 1085 MHz). This effect, first theoretically approximated by Bethe using a non-relativistic and later fully computed in , arises from the electron's interaction with its own electromagnetic field and the screened nuclear charge due to virtual pairs. Modern bound-state calculations refine this value, confirming the experimental measurement from . The of the hydrogen , manifesting as the 21 cm radio line at , results from the magnetic dipole-dipole spin-spin between the and proton spins, with the dominant Fermi term proportional to the squared wavefunction density at the . This s-state , derived from the non-relativistic limit of the and corrected by radiative effects (about 0.1% of the total), splits the F=1 and F=0 hyperfine levels by the hyperfine constant A ≈ . The line's observation in provided early confirmation of predictions for weak in atomic systems. QED also predicts the anomalous magnetic moment of the bound , characterized by the deviation a_e = (g-2)/2 from the Dirac value g=2, with leading-order contribution α/(2π) ≈ 0.001159652 from one-loop and vertex corrections. Higher-order QED loops up to five orders contribute additional terms, yielding a_e(theory) = 0.00115965218073(28) for the , while binding effects in introduce small corrections of order (α Z)^4 m/M_p. measurements of the bound-electron g-factor in hydrogen-like ions test these QED contributions. Theoretical QED predictions for hydrogen energy levels, including all known radiative, relativistic, and recoil corrections, agree with experimental to relative precisions exceeding 10^{-12}, as verified in measurements of transitions like 1S-2S and 2S-Rydberg. For instance, the determination of 2S-nP transitions achieves uncertainties below 1 kHz, matching QED calculations that incorporate up to three-loop and effects. This concordance, spanning over 12 decimal places in reduced units for key intervals like the , underscores QED's validity while probing beyond-standard-model physics through discrepancies in proton radius extractions.

References

  1. [1]
    Hydrogen | H (Element) - PubChem - NIH
    Hydrogen is a chemical element with symbol H and atomic number 1. Classified as a nonmetal, Hydrogen is a gas at 25°C (room temperature).
  2. [2]
    [PDF] Notes on Atomic Structure 1. Introduction 2. Hydrogen Atoms and ...
    Jan 7, 2011 · A.​​ Hydrogen is the simplest atom, consisting of a single electron and a nucleus. We will treat the hydrogenlike atoms generally, allowing for a ...
  3. [3]
    Atomic Weights and Isotopic Compositions for Hydrogen
    Isotope · Relative Atomic Mass · Isotopic · Composition · Standard Atomic Weight · Notes. 1, H, 1, 1.007 825 032 23(9), 0.999 885(70), [1.007 84, 1.008 11] ...
  4. [4]
    [PDF] hydrogen.pdf - University of California, Berkeley
    Hydrogen played a central role in the history of quantum mechanics. By the ... matrix mechanics, one of the first problems solved was the hydrogen atom.
  5. [5]
    [PDF] Quantum Theory of the Hydrogen Atom
    1913: Niels Bohr uses quantum theory to explain the origin of the line spectrum of hydrogen. 1. The electron in a hydrogen atom can exist only in discrete ...
  6. [6]
    4.3 The Hydrogen Atom - FAMU-FSU College of Engineering
    The hydrogen atom consists of a nucleus which is just a single proton, and an electron encircling that nucleus. The nucleus, being much heavier than the ...
  7. [7]
    Hydrogen atom spectrum - Oregon State University
    A hydrogen atom is the simplest atom. Its nucleus consists of one proton, and it has one electron bound to the nucleus. The electron normally exists in its ...Missing: structure | Show results with:structure
  8. [8]
    Origin of the Elements
    Aug 9, 2000 · Approximately 73% of the mass of the visible universe is in the form of hydrogen. Helium makes up about 25% of the mass, and everything else ...
  9. [9]
    Hydrogen explained - U.S. Energy Information Administration (EIA)
    Jun 23, 2023 · Hydrogen is also the most abundant element in the universe. The sun, and other stars, are essentially giant balls of hydrogen and helium gases.
  10. [10]
    Isotope Basics | NIDC
    Hydrogen-1, or protium, is the most prevalent hydrogen isotope, accounting for 99.98% hydrogen atoms, and has no neutrons. ... number, Z is the atomic number, and ...
  11. [11]
  12. [12]
    [PDF] Review of the isotope effect in the hydrogen spectrum
    As the electron emitted radiation and lost energy, the orbit would shrink and eventually (in a predicted 10−12 seconds!) the electron would spiral into the.
  13. [13]
    H-3 - Nuclear Data Center at KAERI
    Binding Energy: 8481.821 +- 0.004 keV; Beta Decay Energy: B- 18.591 +- 0.001 ... Decay energy: 0.019 MeV. Possible parent nuclides: Neutron emission from ...
  14. [14]
    DOE Explains...Deuterium-Tritium Fusion Fuel - Department of Energy
    Deuterium and tritium are isotopes of hydrogen, the most abundant element in the universe. While all isotopes of hydrogen have one proton, deuterium also has ...
  15. [15]
    NIST: Atomic Spectra Database - Ionization Energies Form
    This form provides access to NIST critically evaluated data on ground states and ionization energies of atoms and atomic ions.
  16. [16]
    Hydrogen atom - the NIST WebBook
    View reactions leading to H + (ion structure unspecified), Electron affinity determinations, Ionization energy determinations, De-protonation reactions.Gas phase ion energetics data · References
  17. [17]
    NIST Atomic Spectra Database Ionization Energies Data
    Ionization Energy (eV), Uncertaintyb (eV), References. 1, Hydrogen, 1s, 2S1/2, (13.598434599702), 0.000000000012, HDEL. 2, Helium, 1s2, 1S0, 24.587389011 ...
  18. [18]
    Rutherford, transmutation and the proton - CERN Courier
    May 8, 2019 · Rutherford, transmutation and the proton. 8 May 2019. The events leading to Ernest Rutherford's discovery of the proton, published in 1919.
  19. [19]
    Hydronium ion - American Chemical Society
    Sep 28, 2020 · All acidic aqueous solutions contain protonated water, known commonly as the hydronium ion (H 3 O + ). Brønsted acids release one or more of their protons ( ...
  20. [20]
    (PDF) Atomic Models, J.J. Thomson's “Plum Pudding” Model
    PDF | In 1897, Joseph John Thomson (1856–1940) had announced the discovery of a corpuscle. Others soon called it ▻ electron, despite Thomson's stubborn.
  21. [21]
    [PDF] LXXIX. The scattering of α and β particles by matter and the structure ...
    The atom is supposed to consist of a number N of negatively charged corpuscles, accompanied by .m equal quantity of positive electricity uniformly dis- trihuted ...
  22. [22]
    [physics/0205046] A Hundred Years of Larmor Formula - arXiv
    May 16, 2002 · Abstract: Sir Joseph LARMOR showed in 1897 that an oscillating electric charge emits radiation energy proportional to (acceleration)^2.Missing: power | Show results with:power
  23. [23]
    [PDF] Classical Lifetime of a Bohr Atom 1 Problem - Kirk T. McDonald
    Since the electron is accelerating, a classical analysis suggests that it will continuously radiate energy, and therefore the radius of the orbit would shrink ...
  24. [24]
    [PDF] Philosophical Magazine Series 6 I. On the constitution of atoms and ...
    On the theory of this paper the only neutral atom which contains a single electron is the hydrogen atom. The per- manent state of this atom should ...
  25. [25]
    [PDF] Zur Quantentheorie der Spektrallinien
    1916. Page 93. Zur Quantentheorie der Spektrallinien. 93. (56). C = 1,430 + 7 * 0,010. Berechnen wir C andererseits rnit unseren Werten (54) und (55), so ...Missing: Arnold | Show results with:Arnold
  26. [26]
    Sommerfeld's elliptical atomic orbits revisited—A useful preliminary ...
    Sep 1, 2014 · While the Schrödinger method overcomes the limitations in the Bohr-Sommerfeld description, many of the general features associated with the ...Missing: source | Show results with:source
  27. [27]
    Hydrogen Atom - Galileo
    The hydrogen atom consists of two particles, the proton and the electron, interacting via the Coulomb potential V(→r1−→r2)=e2/r, where as usual r=|→r1−→r2|.
  28. [28]
    [PDF] 4.5 The Hydrogen Atom
    The simplest of all atoms is the Hydrogen atom, which is made up of a positively charged proton with rest mass mp = 1.6726231 × 10-27 kg, and.
  29. [29]
  30. [30]
    [PDF] An undulatory theory of the mechanics of atoms and molecules - ISY
    The paper gives an account of the author's work on a new form of quantum theory. §1. The Hamiltonian analogy between mechanics and optics. §2. The.
  31. [31]
    [PDF] Schrödinger's original quantum–mechanical solution for hydrogen
    Jul 4, 2021 · In this work, we show how the Laplace method can be used to solve for the quantum–mechanical energy eigenfunc- tions of the hydrogen atom, ...
  32. [32]
    [PDF] Chapter 10 The Hydrogen Atom The Schrodinger Equation in ...
    It is a particle in a box with spherical, soft walls. Finally, the hydrogen atom is one of the precious few realistic systems which can actually be solved ...
  33. [33]
  34. [34]
  35. [35]
    The Hydrogen atom - Chemistry 301
    Orbital Shapes. The hydrogen atoms orbitals are the "wavefunction" portion of the quantum mechanical solution to the hydrogen atom. The wavefunctions tell us ...Missing: density | Show results with:density
  36. [36]
    Visualization of Atomic Orbitals
    For the hydrogen atom, Z = 1. The electron possesses angular momentum by virtual of its motion around the atom (sometimes called orbital angular momentum). It ...Missing: nodes boundary
  37. [37]
    [PDF] Ψnlm(r,θ,φ) - MIT OpenCourseWare
    There are three common plots used to help us visualize an s orbital: (1) Probability density Ψ2 plot of s orbitals in which density of dots represents ...
  38. [38]
    19 The Hydrogen Atom and The Periodic Table - Feynman Lectures
    The hydrogen atom in any particular state is a particle with a certain “spin” j—the quantum number of the total angular momentum. Part of this spin comes from ...
  39. [39]
    Current advances: The fine-structure constant
    The fine-structure constant α is of dimension 1 (i.e., it is simply a number) and very nearly equal to 1/137. It is the "coupling constant" or measure of the ...<|separator|>
  40. [40]
    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, 10.1038 ...
  41. [41]
    [PDF] Lectures on Quantum Mechanics - arXiv
    HYDROGEN ATOM VIA DIRAC EQUATION. 149. 34 Hydrogen Atom via Dirac Equation. We determine the quantum stationary states of the hydrogen atom from the Dirac ...
  42. [42]
    [PDF] The hydrogen atom as relativistic bound system - arXiv
    Mar 18, 2019 · There are three types of relativistic corrections in the H atom energy levels which can be listed in order of decreasing magnitude: α) fine ...
  43. [43]
    [1506.07239] Origin of the Spin-Orbit Interaction - arXiv
    Jun 24, 2015 · The interaction energy U is due to the coupling of the induced electric dipole p=(v/c)x mu with the electric field En of the nucleus. Assuming ...
  44. [44]
    [1107.1737] Relativistic Reduced-Mass and Recoil Corrections to ...
    Jul 8, 2011 · Relativistic corrections to vacuum polarization of order alpha (Zalpha)^4 mu are calculated, with a full account of the reduced-mass dependence.
  45. [45]
    Fine Structure of the Hydrogen Atom by a Microwave Method
    Quantum Milestones, 1947: Lamb Shift Verifies New Quantum Concept. Published 24 February, 2025. The discovery of a small discrepancy in hydrogen's atomic ...
  46. [46]
    Stringent test of QED with hydrogen-like tin | Nature
    Oct 4, 2023 · The so far heaviest measured g factor of hydrogen-like ions is 28Si13+, which allowed for a stringent test of QED in low-to-medium-Z ions. Here ...