Fact-checked by Grok 2 weeks ago

Helium atom

The helium atom is the simplest neutral multi-electron atom, consisting of a nucleus with two protons and two neutrons surrounded by two electrons in the ground-state of 1s². With an of 2, it represents the second element in the periodic table and is the first , characterized by its chemical inertness due to the fully filled first . In , the helium atom serves as a foundational example for understanding electron-electron interactions, as its includes Coulomb repulsion between the two electrons, making exact analytical solutions impossible unlike the . Approximate methods, such as the with effective nuclear charge screening (Z_eff ≈ 1.7), yield a ground-state of approximately -77.5 , close to the experimental value of -79.0 . The ground-state wavefunction is symmetric in space and antisymmetric in ( with total S=0), ensuring Pauli exclusion for the identical fermions. Helium's atomic properties highlight its stability: the first is 24.59 eV (the highest of any element), removing one to form He⁺, while the second is 54.42 eV, yielding the bare . This tight binding contributes to helium's role as the second most abundant element in the universe, formed primarily during , and its applications in , lasers, and as an inert atmosphere due to minimal reactivity.

Basic Properties

Atomic Structure

The helium atom consists of a with an of 2, comprising two protons and, in the predominant ^4\mathrm{He}, two neutrons, surrounded by two extranuclear electrons. The , often referred to as an in nuclear contexts, carries a charge of +2e, where e is the , attracting the negatively charged electrons electrostatically. This configuration makes helium the simplest stable multi-electron atom, with the electrons occupying atomic orbitals influenced by both nuclear attraction and mutual repulsion. The atomic mass of the most abundant isotope, ^4\mathrm{He}, is 4.002602 u, accounting for over 99.999% of naturally occurring helium, while the rarer ^3\mathrm{He} isotope has a mass of 3.016029 u and differs primarily in lacking one neutron, affecting its nuclear properties but not the electronic structure in basic terms. The effective atomic radius of helium, scaled from the hydrogen Bohr radius due to the higher nuclear charge, is approximately 31 pm, representing the calculated distance from the nucleus to the outermost electron density. In the ground state, the electron density distribution is spherically symmetric, with the probability density peaking near the nucleus and falling off radially, a direct consequence of the two electrons filling the 1s orbital. Electrically neutral overall, the helium atom balances the charge of +2e with two s each of charge -e, resulting in a total charge of zero. This closed-shell arrangement, where both electrons pair in the lowest-energy 1s orbital, imparts exceptional stability and chemical inertness, as the filled subshell resists further electron addition or removal under standard conditions. As a foundational model in , the helium atom exemplifies the challenges of multi-electron systems by incorporating electron-electron repulsion alongside attraction, serving as a for testing theoretical approximations beyond the hydrogen-like case. Its —yet inclusion of interparticle interactions—has driven advancements in understanding since the early 20th century.

Electron Configuration and Stability

The electron configuration of the helium atom consists of two electrons occupying the 1s orbital, denoted as $1s^2. This arrangement adheres to the , requiring the electrons to have opposite spins to occupy the same orbital. The total S=0 results in a , classified as parahelium, while the total orbital L=0 yields the ^{1}S_0. This configuration represents the lowest energy state, with both electrons paired in the innermost shell. The filled 1s subshell imparts exceptional stability to the helium atom, manifesting as high ionization energies that render it chemically inert. The first ionization energy, required to remove one electron from the neutral atom to form \ce{He+}, is 24.587 eV, while the second ionization energy, to remove the remaining electron from \ce{He+} to form the bare nucleus, is 54.418 eV. These values, the highest among all elements, reflect the strong binding due to the nuclear charge of Z=2 and the complete occupancy of the subshell, minimizing the atom's reactivity under standard conditions. In comparison to the , which has a single bound with an of 13.6 , the helium atom experiences a doubled charge that would naively suggest twice the per (approximately 27.2 each if independent). However, electron-electron repulsion significantly reduces the effective binding, resulting in the observed first of about 24.6 —tighter than but less than the independent-electron prediction. This repulsion effect underscores the challenges of multi-electron systems. Isotopic differences between ^4\ce{He} and ^3\ce{He} have negligible impact on the and quantum level occupancy at non-relativistic scales, as the variations primarily cause small shifts in energy levels rather than altering orbital assignments. calculations confirm that the remains $1s^2 for both s, with isotope shifts on the order of $10^{-3} to $10^{-5} cm^{-1} for low-lying states.

Quantum Mechanical Framework

The Two-Electron

The quantum mechanical description of the helium atom begins with the non-relativistic Hamiltonian for a two-electron system in the presence of a nucleus of charge Z = 2. In SI units, the Hamiltonian operator is given by H = -\frac{\hbar^2}{2m} (\nabla_1^2 + \nabla_2^2) - \frac{Z e^2}{4\pi \epsilon_0 r_1} - \frac{Z e^2}{4\pi \epsilon_0 r_2} + \frac{e^2}{4\pi \epsilon_0 r_{12}}, where m is the electron mass, \hbar is the reduced Planck's constant, e is the elementary charge, \epsilon_0 is the vacuum permittivity, \mathbf{r}_1 and \mathbf{r}_2 are the position vectors of the two electrons relative to the nucleus, and r_{12} = |\mathbf{r}_1 - \mathbf{r}_2| is the inter-electron distance. This expression accounts for the kinetic energy of each electron, the Coulomb attraction between each electron and the nucleus, and the Coulomb repulsion between the two electrons. In , where \hbar = m = e = 4\pi \epsilon_0 = 1, the simplifies to H = -\frac{1}{2} (\nabla_1^2 + \nabla_2^2) - \frac{2}{r_1} - \frac{2}{r_2} + \frac{1}{r_{12}}. The first two terms represent the operators for the electrons, the next two describe the attractive potentials from the (Z = 2), and the final term captures the electron-electron repulsion, which introduces strong correlation effects. The time-independent for the system is H \psi(\mathbf{r}_1, \mathbf{r}_2) = E \psi(\mathbf{r}_1, \mathbf{r}_2), where \psi(\mathbf{r}_1, \mathbf{r}_2) is the six-dimensional depending on the coordinates of both electrons, and E is the eigenvalue. To address the multi-particle nature, the total can be separated into center-of-mass and relative coordinates. The center-of-mass motion corresponds to a with the total mass of the atom ( plus two electrons), while the relative governs the internal dynamics with a \mu \approx m (the ), given the much larger nuclear mass (approximately 7295 times the for ). However, the $1/r_{12} repulsion term couples the electron coordinates inseparably, preventing an exact analytical separation into independent single-particle problems as achieved for the .

Challenges in Exact Solution

The helium atom represents the simplest nontrivial example of a quantum many-body system, where the presence of two electrons introduces significant challenges in solving the exactly. Unlike the , which admits an analytical solution due to its separable one-electron , the helium includes the electron-electron repulsion term $1/r_{12}, where r_{12} is the interelectronic distance. This term accounts for , capturing the instantaneous avoidance of electrons due to their mutual repulsion, which renders the wavefunction non-separable in the electron coordinates and prevents a closed-form solution. The non-separability arises because the $1/r_{12} couples the motions of the two electrons, making the problem inherently multidimensional. For two electrons, the spatial wavefunction depends on six coordinates (three for each electron), leading to an exponential increase in as the number of particles grows—a hallmark of the . This dimensionality curse necessitates numerical methods or series expansions for any practical computation, as exact analytical progress is impossible beyond perturbative or variational frameworks. A key quantitative measure of these challenges is the , defined as the difference between the exact non-relativistic ground-state energy and that from the independent-particle () approximation. For , this energy is approximately -0.042 , representing the deficit due to unaccounted in mean-field treatments; recovering this small but crucial fraction requires highly sophisticated methods to achieve chemical accuracy. Historically, these difficulties were highlighted by in 1929, who noted that while the fundamental laws of are known, their exact application to many-body systems like the helium atom leads to equations too complicated to solve analytically, establishing helium as a for quantum many-body .

Ground State Approximations

Perturbation Theory Approach

In the approach to the helium atom, the is separated into an unperturbed part and a perturbation to approximate the energy. The unperturbed H_0 consists of the kinetic energy operators and the attractive interactions between the and each , treating the electrons as independent hydrogen-like particles with nuclear charge Z = 2: H_0 = h_1 + h_2, \quad h_i = \frac{\mathbf{p}_i^2}{2m} - \frac{Z e^2}{r_i}, where i = 1, 2 labels the electrons, \mathbf{p}_i is the momentum operator, m is the electron mass, e is the elementary charge, and r_i is the distance from the nucleus to electron i. The perturbation V accounts for the electron-electron repulsion: V = \frac{e^2}{r_{12}}, where r_{12} = |\mathbf{r}_1 - \mathbf{r}_2| is the inter-electron distance. This separation is justified because the repulsion term is smaller than the attractive terms for tightly bound inner electrons, allowing a in powers of V. The unperturbed wavefunction \psi_0 is the antisymmetrized product of hydrogenic 1s orbitals for Z = 2, but for the spatial part in the spin state (relevant for the ), it simplifies to the symmetric product: \psi_0(\mathbf{r}_1, \mathbf{r}_2) = \frac{Z^3}{\pi a_0^3} \exp\left( -\frac{Z (r_1 + r_2)}{a_0} \right), where a_0 is the Bohr radius; the full wavefunction includes the spin part but does not affect the first-order energy calculation due to spatial symmetry. The unperturbed energy is E^{(0)} = -Z^2 hartrees = -4 hartrees ≈ -108.8 eV. The first-order energy correction is the expectation value of the perturbation in the unperturbed state: E^{(1)} = \int \psi_0^* V \psi_0 \, d\tau = \left\langle \frac{e^2}{r_{12}} \right\rangle_0. Evaluating this integral yields E^{(1)} = \frac{5}{8} Z hartrees. For Z = 2, E^{(1)} = \frac{5}{4} hartrees ≈ 34.0 eV, so the approximate ground state energy is E \approx E^{(0)} + E^{(1)} = -2.75 hartrees ≈ -74.8 eV, compared to the experimental value of -79.0 eV. This first-order approximation underestimates the (overestimates the energy by about 5%) because it neglects electron correlation effects beyond the mean field, where the electrons avoid each other more effectively than in the independent-particle model. Higher-order corrections improve the result but face challenges: the V is singular at r_{12} = 0, leading to divergent contributions in second and higher orders unless regularization techniques, such as those involving exponential expansions, are employed.

Variational Method

The variational method provides an upper bound to the ground state energy of the helium atom through the use of trial wave functions that approximate the true . The variational theorem states that for any trial \psi_t, the expectation value \langle \psi_t | \hat{H} | \psi_t \rangle / \langle \psi_t | \psi_t \rangle \geq E_\mathrm{ground}, where \hat{H} is the and E_\mathrm{ground} is the exact energy. A simple trial for the helium assumes independent electrons in screened hydrogen-like 1s orbitals, given by \psi_t(\mathbf{r}_1, \mathbf{r}_2) = \left( \frac{\alpha^3}{\pi a_0^3} \right) e^{-\alpha (r_1 + r_2)/a_0}, where \alpha is a variational representing the and a_0 is the . This form implicitly accounts for screening but neglects explicit electron . The energy expectation value E(\alpha) is minimized by solving \frac{dE}{d\alpha} = 0, which yields the optimal \alpha = Z - \frac{5}{16} = \frac{27}{16} for nuclear charge Z = 2. Substituting this value gives the variational energy E_\mathrm{var} = -\left( \frac{27}{16} \right)^2 \times 2 \, \mathrm{Ry} \approx -77.5 \, \mathrm{eV}. This variational result is superior to the first-order perturbation theory estimate, as the flexible parameter \alpha implicitly incorporates some electron correlation effects beyond the independent-particle approximation. To achieve higher precision, trial functions are extended to include explicit dependence on the interelectronic distance r_{12}, such as in Hylleraas-type expansions of the form \psi_t = \sum c_{klm} r_1^k r_2^l r_{12}^m e^{-\alpha (r_1 + r_2)}. Egil Hylleraas's 1929 calculation using a 10-term expansion of this type yielded a ground state energy of approximately -79.0 \, \mathrm{eV}, remarkably close to the experimental value.

Screening Effect and Effective Charge

In the helium atom, the screening effect refers to the phenomenon where each of the two 1s electrons partially shields the from the experienced by the other , thereby reducing the Z_{\text{eff}} below the bare nuclear charge Z = 2. This shielding arises primarily from the Coulomb repulsion between the electrons, which causes their spatial distributions to correlate such that they tend to avoid occupying the same region near the simultaneously. As a result, each perceives a diminished nuclear attraction, leading to weaker binding compared to a non-interacting model. A simple empirical approach to estimate this screening is provided by , which assign a shielding constant \sigma of 0.30 to the contribution from the other in the same 1s . Thus, for , Z_{\text{eff}} \approx Z - \sigma = 2 - 0.30 = 1.70, reflecting the partial neutralization of the nuclear charge by the electron cloud of the companion . This approximation captures the essence of intra-shell screening without solving the full quantum mechanical problem. Within the variational framework, a more refined estimate emerges from optimizing the wavefunction , yielding Z_{\text{eff}} = Z \left(1 - \frac{5}{16Z}\right) \approx 1.6875 for Z = 2. This value indicates that the - interaction effectively reduces the nuclear by about 16% , consistent with the physical picture of correlated electron densities where the probability of finding both electrons close to the is lowered due to repulsion. The screening effect explains the deviation of helium's energy from the independent-electron approximation, where neglecting repulsion would predict each bound by -13.6 Z^2 = -54.4 , for a total of -108.8 ; in reality, the exact non-relativistic energy is -79.0 , with the reduced magnitude attributable to screening that weakens the overall binding.

Excited States and Energy Levels

Singlet and Triplet Configurations

In the excited states of the helium atom, one electron remains in the 1s orbital while the other is promoted to a higher orbital such as 2s or 2p, resulting in configurations denoted as 1s nl where n ≥ 2 and l is the orbital angular momentum quantum number of the excited electron. These configurations give rise to distinct spin states due to the two electrons' spins coupling to total spin quantum numbers S = 0 or S = 1. The S = 0 states, known as singlets or parahelium, feature antiparallel electron spins and a symmetric spatial wavefunction, while the S = 1 states, known as triplets or orthohelium, have parallel spins and an antisymmetric spatial wavefunction. The requires the total wavefunction of the two identical fermions to be antisymmetric under particle exchange, which is achieved by combining the and spatial parts with opposite symmetries: the antisymmetric function for s pairs with a symmetric spatial function, and the symmetric function for triplets pairs with an antisymmetric spatial function. In triplet states, the antisymmetric spatial wavefunction introduces an exchange node where the electrons' positions coincide (r_1 = r_2), effectively increasing the average separation and reducing repulsion compared to the symmetric spatial wavefunction of s. Consequently, for configurations with the same n and l, triplet states lie lower in energy than their corresponding states, a result first quantitatively analyzed using in the early quantum mechanical treatments of . Electric dipole transitions in helium obey the selection rule ΔS = 0, restricting optical transitions to occur within the same manifold—either between states (parahelium series) or triplet states (orthohelium series)—due to the conservation of total spin in the weak spin-orbit coupling regime. This separation underlies the historical observation of distinct spectral series for parahelium and orthohelium.

Helium Atomic Spectrum

The atomic spectrum of neutral helium was first identified in the Sun's during the total on August 18, 1868, when French astronomer Pierre Jules César Janssen observed an unknown bright yellow emission line at 587.6 nm during spectroscopic analysis of solar prominences. Independently, astronomer Joseph Norman Lockyer detected the same line on October 20, 1868, from ground-based observations in using a spectroscope on in daylight and proposed it originated from a new element in the Sun, naming it helium after the Greek word for "sun," as no matching terrestrial element was known at the time. This discovery marked the first identification of an element via before its isolation on Earth in 1895. The of the helium atom features discrete lines from transitions between excited states, classified into series based on changes and multiplicity ( or triplet configurations). These series arise due to the and exchange effects, with orthohelium (triplet states) producing more visible lines than parahelium ( states). The principal series involves transitions from 1s np states to the 1s 2s state, occurring primarily in the to visible range; a representative example is the intense yellow line at 587.6 nm, assigned to the triplet transition 3³D → 2³P in the diffuse subclass. Complementary series include the sharp series (involving sp transitions), diffuse series (dp), and fundamental series (fd), each manifesting in both and triplet systems with converging lines toward series limits around 50–60 eV. The Pickering series, observed in hot stellar atmospheres, corresponds to transitions in singly ionized (He ) but was historically linked to neutral interpretations before Bohr's model clarified its origin. The , adapted for helium, describes these series wavenumbers as \bar{\nu} = R_\text{He} \left( \frac{1}{n_1^2} - \frac{1}{n_2^2} \right), where R_\text{He} is the helium Rydberg constant, but requires an Z_\text{eff} = Z - \sigma (with Z = 2 and screening constant \sigma \approx 0.3–1.7 depending on the state) to account for electron-electron repulsion and screening, deviating from the hydrogenic case. This modification enables accurate prediction of line positions, particularly for Rydberg states where outer electrons experience reduced attraction. Advances in laser spectroscopy have enabled precise resolution of fine and hyperfine structure in helium lines, providing benchmarks for atomic theory. For ^3He, Doppler-free laser excitation measures hyperfine splittings in the 2³P state to sub-MHz accuracy, revealing isotope shifts and testing relativistic corrections. Such experiments confirm energy level assignments and support QED validations without accessing ionization thresholds.

Ionization and Precision Calculations

Theoretical Ionization Energy

The ionization energy of the helium atom is defined as the difference between the ground-state energy of the He⁺ ion and that of the neutral helium atom, I = E(\ce{He+}) - E(\ce{He}). The He⁺ ion is a one-electron hydrogen-like system with nuclear charge Z = 2, yielding a ground-state energy of E(\ce{He+}) = -2 hartree = -54.417760 eV. High-precision variational calculations provide the non-relativistic ground-state energy of neutral helium as E(\ce{He}) = -2.903724377 hartree = -79.005147 eV, resulting in a theoretical non-relativistic ionization energy of I = 24.587387 eV. In the Rayleigh-Schrödinger perturbation theory approach, the electron-electron repulsion term $1/r_{12} is treated as a perturbation to the unperturbed two-electron system of independent hydrogenic orbitals with Z = 2, which has zeroth-order energy E^{(0)} = -4 hartree. The first-order correction is E^{(1)} = 5/8 Z = 1.25 hartree, giving a total energy of -2.75 hartree and an ionization energy of approximately 20.4 eV, which underestimates the true value by treating correlation inadequately. Including higher-order terms significantly improves the result; for instance, calculations through 13th order in perturbation theory converge to within 0.001 eV of the non-relativistic limit of 24.587 eV. The variational method yields upper bounds to the ground-state energy by optimizing trial wavefunctions. A simple screened hydrogenic trial function \psi = (Z'^3/\pi) e^{-Z'(r_1 + r_2)} with optimized effective charge Z' = 27/16 = 1.6875 gives E = -(Z')^2 = -2.84765625 , corresponding to an ionization energy of approximately 23.07 . More advanced variational approaches, such as those incorporating Hylleraas coordinates to explicitly account for electron correlation via terms like r_{12}, achieve energies of -2.9037 or better, yielding ionization energies approaching 24.59 . The electron correlation energy, representing the adjustment to the electron-electron repulsion beyond the mean-field contribution, amounts to approximately -0.042 (\approx -1.14 ) in helium, which increases the by this amount relative to Hartree-Fock approximations (where E_\ce{HF} \approx -2.86168 and I \approx 23.45 ).
MethodGround-State Energy ()Ionization Energy ()
(1st order)-2.7520.4
(high order)≈ -2.9037≈ 24.59
Variational (simple screened)-2.847723.07
Variational (improved Hylleraas)-2.90372424.59
Exact non-relativistic-2.90372437724.587

Experimental Determination

The ionization energy of the helium atom, defined as the energy required to remove one from the neutral to form He⁺ in its , has been determined experimentally through a progression of techniques, starting from early electron impact methods and evolving to sophisticated spectroscopic approaches. One of the earliest measurements was conducted via electron impact in the Franck-Hertz experiment, which established the threshold at approximately 24.5 , providing initial confirmation of discrete energy levels in atoms. This historical benchmark laid the foundation for subsequent refinements, highlighting the quantized nature of and processes. Modern determinations rely on high-precision photoionization spectroscopy, particularly using to probe thresholds where the photoionization cross-section rises sharply. These experiments measure the onset of ion production as a function of , yielding accurate values by analyzing the edges near 24.59 . Complementary methods involve to high-lying Rydberg states followed by series , where transition frequencies to states with principal quantum numbers n approaching converge to the . Such techniques, employing tunable lasers and radiofrequency fields for state-selective detection, have achieved relative precisions of ~10^{-9} or better as of 2025. The accepted experimental value for the ground-state ionization energy of ⁴He is 24.587389011(25) × 10^{-3} eV, obtained from detailed spectroscopic compilations and confirmed through precision laser-based measurements. This uncertainty of 2.5 × 10^{-8} eV reflects the high accuracy of contemporary experiments, now approaching limits set by quantum electrodynamic corrections when compared to theory. For the isotope ³He, the ionization energy is slightly lower at approximately 24.58609 eV, with a difference of about 1.3 × 10^{-3} eV arising primarily from the effect due to the lighter nuclear mass.

Advanced Theoretical Methods

Advanced theoretical methods for the helium atom extend beyond basic approximations to achieve unprecedented precision, incorporating explicit correlation effects, relativistic dynamics, and corrections. The Hylleraas-configuration (Hy-CI) approach uses basis functions that explicitly include the interelectron r_{12}, enabling variational solutions to the non-relativistic with exceptional accuracy. By the 1990s, Hy-CI calculations had reached precisions of $10^{-10} for the ground-state energy, as demonstrated in large-scale expansions with thousands of terms. Relativistic effects are accounted for using the Dirac-Coulomb-Breit , which includes and other fine-structure terms beyond the non-relativistic limit. These corrections contribute approximately 0.0001 eV to the ground-state , arising primarily from mass-velocity and terms in the Breit interaction. Full incorporation of these effects shifts the non-relativistic benchmark energy of -2.903724377 by small but measurable amounts. QED contributions, analogous to the Lamb shift in hydrogen, introduce radiative corrections due to virtual photon exchanges, with leading terms on the order of $10^{-6} eV for the helium ground state. Higher-order QED effects, including electron-electron radiative interactions, are evaluated ab initio, achieving overall precisions of $10^{-12} hartree in 2020s calculations that combine Hy-CI bases with perturbative QED insertions. The complete relativistic and QED-corrected ground-state energy is -2.903724375(4) hartree, matching experimental ionization energies to high fidelity. These methods have profound applications in testing QED validity and refining fundamental constants, such as the , through comparisons with precision . Recent ab initio calculations using correlated basis spline functions confirm theoretical predictions to 12 decimal digits for key energy levels, highlighting helium's role as a benchmark for beyond-standard-model physics.

References

  1. [1]
    Helium Atom - Richard Fitzpatrick
    A helium atom consists of a nucleus of charge $+2 e$ surrounded by two electrons. Let us attempt to calculate its ground-state energy.
  2. [2]
    The Structure of the Atom – MCC AST - Maricopa Open Digital Press
    Helium has two protons in its nucleus instead of the single proton that characterizes hydrogen. In addition, the helium nucleus contains two neutrons, particles ...
  3. [3]
    The Electron Configurations of Atoms
    Therefore, the electron configuration of hydrogen is written: For helium (atomic number 2), which has two electrons, the electron configuration is: He: 1s2.
  4. [4]
    Atomic Data for Helium (He) - Physical Measurement Laboratory
    Ionization energy 198310.669 cm-1 (24.587387 eV) Ref. M02 He II Ground State 1s 2S1/2. Ionization energy 438908.8789 cm-1 (54.417760 eV) Ref. MK00b.
  5. [5]
    [PDF] Lecture 21-22, Helium Atom - DSpace@MIT
    We refer to this picture as an independent electron approximation. Within the wavefunction the electrons behave as if they do not interact, because we have ...
  6. [6]
    The Helium Atom - NASA Scientific Visualization Studio
    Jul 3, 2007 · Helium nuclei were created in the Big Bang and contain two protons and two neutrons each. Helium is the second most abundant element, ...
  7. [7]
    1 Properties and History | The Impact of Selling the Federal Helium ...
    It has a particularly stable and symmetrical structure. The nucleus of the atom consists of two protons and either one or two neutrons, depending on the isotope ...
  8. [8]
    Atomic Data for Helium (He) - Physical Measurement Laboratory
    Atomic Data for Helium (He). Atomic Number = 2. Atomic Weight = 4.002602. Reference E95. Isotope Mass Abundance Spin Mag Moment 3He 3.01603 0.000138% 1/2 ...Missing: radius | Show results with:radius
  9. [9]
    physics.nist.gov Elemental Data Index: 2 Helium
    2 - Helium (He) information about the ground-state, ionization energy, and atomic weight data. Atomic Weight: 4.002 602(2) Ionization Energy: eVMissing: properties radius<|control11|><|separator|>
  10. [10]
    WebElements Periodic Table » Helium » radii of atoms and ions
    Helium's calculated atomic radius is 31 pm, covalent radius is 28 pm, and the Pauling radius of He(I) is 93 pm. No ionic radii are listed for He.Missing: effective | Show results with:effective
  11. [11]
    Averaged electron densities of the helium-like atoms - arXiv
    Aug 2, 2019 · Abstract:Different kinds of averaging of the wavefunctions/densities of the two-electron atomic systems are investigated.
  12. [12]
    Helium Atom - an overview | ScienceDirect Topics
    Helium atoms in their ground electronic state are chemically inert. A helium atom's two electrons form a closed shell. It cannot enter in a chemical ...Missing: inertness | Show results with:inertness
  13. [13]
    [PDF] HELIUM ATOM R r2 r1 r2 r1
    atomic units. In this system of units we choose our unit of mass to be equal to the electron mass, me, our unit of charge to be equal to the electron charge ...
  14. [14]
    [PDF] helium.pdf
    Bound states of helium with quantum numbers. Energies are measured in atomic units. Notice the gap in the energy scale between the ground state and the excited ...
  15. [15]
    [PDF] Helium Atom, Approximate Methods
    Apr 22, 2008 · The Helium atom has 2 electrons with coordinates r1 and r2 as well as a single nucleus with coordinate R. The nuclues carries a Z = +2e charge.
  16. [16]
    [PDF] 5.61 F17 Lecture 22: Helium Atom - MIT OpenCourseWare
    quantum mechanics says we have to make the wavefunction antisymmetric, antisymmetry by itself does not affect our energy for Helium. The simplest way for us to ...
  17. [17]
    [PDF] The helium atom
    Center of mass separation. Focus on the relative problem. Center of mass is assumed to be at the nucleus. ▫ Good approximation as nuclei get heavier.
  18. [18]
    [PDF] Comparing many-body approaches against the helium atom exact ...
    Apr 1, 2019 · The Schrödinger equation cannot be solved in a closed form. The calculation of many-body correlation energies and correlation effects presents ...
  19. [19]
    [PDF] Solving the Helium Atom
    May 4, 2007 · This method can, in principle, be used to calculate very complicated quantum systems like molecules with good accuracy and speed.<|separator|>
  20. [20]
    Correlation Effects in Atoms. I. Helium | Phys. Rev.
    Starting from the Hartree-Fock Hamiltonian as zeroth-order approximation, we have evaluated the correlation energy through fifth order in perturbation theory.
  21. [21]
    Quantum Mechanics of One- and Two-Electron Atoms - Google Books
    Quantum Mechanics of One- and Two-Electron Atoms. Front Cover. Hans A. Bethe, Edwin E. Salpeter. Springer Science & Business Media, Dec 6, 2012 - Science - 370 ...
  22. [22]
    Atomic Shielding Constants | Phys. Rev.
    Atomic Shielding Constants. J. C. Slater. Jefferson Physical Laboratory ... J. C. Slater, Phys. Rev. 34, 1293 (1929); J. C. Slater, Phys. Rev. 35, 509 (1930).
  23. [23]
    None
    Summary of each segment:
  24. [24]
    How Scientists Discovered Helium, the First Alien Element, 150 ...
    Aug 17, 2018 · On August 18, 1868, Janssen managed to do just that. He became the first person to observe helium, an element never before seen on Earth, in the ...
  25. [25]
    This Month in Physics History | American Physical Society
    Some 5,000 miles away, on October 20, 1868, the English astronomer Joseph Norman Lockyer also succeeded in observing the solar prominences in broad daylight.
  26. [26]
    Strong Lines of Helium ( He ) - Physical Measurement Laboratory
    Intensity, Vacuum Wavelength (Å), Spectrum, Reference. 15 c, 231.4541, He II, GM65. 20 c, 232.5842, He II, GM65. 30 c, 234.3472, He II, GM65.Missing: principal 587.6
  27. [27]
    The Pickering Series in O Type Stars - Nature
    The Pickering series, consisting of the lines 5411, 4542, 4200, etc., in stellar spectra, is due to ionised helium.Missing: sharp fundamental
  28. [28]
    [PDF] precision measurements of the hyperfine structure in the 23p state
    In this experiment, the hyperfine structure splittings in the 23P state of the 3He atom are measured to new levels of precision. The basic experimental setup ...
  29. [29]
    Precision spectroscopy of atomic helium | National Science Review
    The fine-structure splitting in the 23P state of helium is considered to be the most suitable for determining the fine-structure constant α in atoms. After more ...
  30. [30]
    Ground-state energies for helium, and | Phys. Rev. A
    Apr 15, 2002 · Access by Google Scholar Library Links. Ground-state energies for helium, H − , and P s −. G. W. F. Drake, Mark M. Cassar, and Razvan A. Nistor.
  31. [31]
    Perturbation calculations of the helium ground state energy
    A comparison is presented for the ground state of the helium atom of high order Rayleigh-Schrödinger perturbation theories based upon a hydrogenic, ...
  32. [32]
    8.2: Perturbation Theory and the Variational Method for Helium
    ### Summary of Calculated Ground State Energies and Ionization Energies for Helium
  33. [33]
    An experimental determination of the ionisation potential for ...
    The experimental determination of the ionization potential for helium was 20.5 volts, using a method of accelerating electrons until positive ions were ...
  34. [34]
    The absolute photoionization cross sections of helium, neon, argon ...
    Agreement with theoretical calculations on helium is demonstrated to be within ± 2-3% from threshold down to the double ionization threshold at 79 eV.
  35. [35]
    Ionization Energy of the Metastable State of from Rydberg-Series ...
    Aug 27, 2021 · We report on a new determination of the ionization energy of the 2 S 0 1 metastable state of He at a precision of 32 kHz (or 3 × 10 - 11 ) by ...<|control11|><|separator|>
  36. [36]
    Energy levels for the stable isotopes of atomic helium(4He I and 3He I)
    We calculate very accurate ab initio ionization energies for both 4He I and 3He I as well as the isotope shifts for n = 1 to 10, L = 0 to 7 and combined ...
  37. [37]
    Obtaining microhartree accuracy for two-electron systems with ...
    This accuracy is comparable to that of the Hylleraas-CI method and for diatomics approaches that obtained with explicitly- correlated basis sets in elliptical ...
  38. [38]
    [PDF] Relativistic corrections for the ground electronic state of molecular ...
    This assertion is supported by the example of the helium atom. The nonrecoil relativistic correc- tion to the 4He ionization energy is 16 904.024 MHz, while.Missing: 0.0001 | Show results with:0.0001<|separator|>
  39. [39]
    High Precision Theory of Atomic Helium - IOPscience
    Essentially exact calculations of the nonrelativistic wave functions and energies of helium are reviewed, together with the lowest-order relativistic ...Missing: ground | Show results with:ground
  40. [40]
    Radiative QED contribution to the helium Lamb shift | Phys. Rev. A
    Jan 6, 2021 · We present a derivation of the last unknown part of the α 7 ⁢ m contribution to the Lamb shift of a two-electron atom, induced by the radiative QED effects.
  41. [41]
    spline basis functions to the leading relativistic and QED corrections ...
    Dec 21, 2023 · In this paper, we detail the extension of C-BSBFs towards calculating leading relativistic and quantum electrodynamics (QED) corrections for energy levels of ...<|control11|><|separator|>