Helium atom
The helium atom is the simplest neutral multi-electron atom, consisting of a helium-4 nucleus with two protons and two neutrons surrounded by two electrons in the ground-state electron configuration of 1s².[1][2][3] With an atomic number of 2, it represents the second element in the periodic table and is the first noble gas, characterized by its chemical inertness due to the fully filled first electron shell.[4] In quantum mechanics, the helium atom serves as a foundational example for understanding electron-electron interactions, as its Hamiltonian includes Coulomb repulsion between the two electrons, making exact analytical solutions impossible unlike the hydrogen atom.[1][5] Approximate methods, such as the variational principle with effective nuclear charge screening (Z_eff ≈ 1.7), yield a ground-state energy of approximately -77.5 eV, close to the experimental value of -79.0 eV.[1] The ground-state wavefunction is symmetric in space and antisymmetric in spin (singlet state with total spin S=0), ensuring Pauli exclusion for the identical fermions.[1] Helium's atomic properties highlight its stability: the first ionization energy is 24.59 eV (the highest of any element), removing one electron to form He⁺, while the second is 54.42 eV, yielding the bare nucleus.[4] This tight binding contributes to helium's role as the second most abundant element in the universe, formed primarily during Big Bang nucleosynthesis, and its applications in cryogenics, lasers, and as an inert atmosphere due to minimal reactivity.[6][7]Basic Properties
Atomic Structure
The helium atom consists of a nucleus with an atomic number of 2, comprising two protons and, in the predominant isotope ^4\mathrm{He}, two neutrons, surrounded by two extranuclear electrons.[8] The nucleus, often referred to as an alpha particle in nuclear contexts, carries a charge of +2e, where e is the elementary charge, attracting the negatively charged electrons electrostatically.[9] 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.[8] 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.[10] 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.[11] Electrically neutral overall, the helium atom balances the nuclear charge of +2e with two electrons 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.[12] As a foundational model in quantum mechanics, the helium atom exemplifies the challenges of multi-electron systems by incorporating electron-electron repulsion alongside nuclear attraction, serving as a benchmark for testing theoretical approximations beyond the hydrogen-like case. Its simplicity—yet inclusion of interparticle interactions—has driven advancements in understanding atomic binding since the early 20th century.[1]Electron Configuration and Stability
The ground state electron configuration of the helium atom consists of two electrons occupying the 1s orbital, denoted as $1s^2. This arrangement adheres to the Pauli exclusion principle, requiring the electrons to have opposite spins to occupy the same orbital. The total spin quantum number S=0 results in a singlet state, classified as parahelium, while the total orbital angular momentum L=0 yields the term symbol ^{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 hydrogen atom, which has a single electron bound with an ionization energy of 13.6 eV, the helium atom experiences a doubled nuclear charge that would naively suggest twice the binding energy per electron (approximately 27.2 eV each if independent). However, electron-electron repulsion significantly reduces the effective binding, resulting in the observed first ionization energy of about 24.6 eV—tighter than hydrogen 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 electron configuration and quantum level occupancy at non-relativistic scales, as the reduced mass variations primarily cause small shifts in energy levels rather than altering orbital assignments. Ab initio calculations confirm that the ground state remains $1s^2 for both isotopes, with isotope shifts on the order of $10^{-3} to $10^{-5} cm^{-1} for low-lying states.Quantum Mechanical Framework
The Two-Electron Hamiltonian
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.[13] 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.[14] In atomic units, where \hbar = m = e = 4\pi \epsilon_0 = 1, the Hamiltonian 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 kinetic energy operators for the electrons, the next two describe the attractive potentials from the nucleus (Z = 2), and the final term captures the electron-electron repulsion, which introduces strong correlation effects.[15] The time-independent Schrödinger equation 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 wave function depending on the coordinates of both electrons, and E is the energy eigenvalue.[16] To address the multi-particle nature, the total Hamiltonian can be separated into center-of-mass and relative coordinates. The center-of-mass motion corresponds to a free particle with the total mass of the atom (nucleus plus two electrons), while the relative Hamiltonian governs the internal dynamics with a reduced mass \mu \approx m (the electron mass), given the much larger nuclear mass (approximately 7295 times the electron mass for helium-4).[5] 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 hydrogen atom.[17]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 Schrödinger equation exactly. Unlike the hydrogen atom, which admits an analytical solution due to its separable one-electron Hamiltonian, the helium Hamiltonian includes the electron-electron repulsion term $1/r_{12}, where r_{12} is the interelectronic distance. This term accounts for electron correlation, 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.[18] The non-separability arises because the $1/r_{12} interaction 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 computational complexity as the number of particles grows—a hallmark of the many-body problem. This dimensionality curse necessitates numerical methods or series expansions for any practical computation, as exact analytical progress is impossible beyond perturbative or variational frameworks.[5][19] A key quantitative measure of these challenges is the correlation energy, defined as the difference between the exact non-relativistic ground-state energy and that from the independent-particle (Hartree-Fock) approximation. For helium, this energy is approximately -0.042 hartree, representing the binding energy deficit due to unaccounted electron correlation in mean-field treatments; recovering this small but crucial fraction requires highly sophisticated methods to achieve chemical accuracy.[20] Historically, these difficulties were highlighted by Paul Dirac in 1929, who noted that while the fundamental laws of quantum mechanics 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 paradigm for quantum many-body theory.[21]Ground State Approximations
Perturbation Theory Approach
In the perturbation theory approach to the helium atom, the Hamiltonian is separated into an unperturbed part and a perturbation to approximate the ground state energy. The unperturbed Hamiltonian H_0 consists of the kinetic energy operators and the attractive Coulomb interactions between the nucleus and each electron, 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.[22] 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 series expansion in powers of V.[22] The unperturbed ground state wavefunction \psi_0 is the antisymmetrized product of hydrogenic 1s orbitals for Z = 2, but for the spatial part in the singlet spin state (relevant for the ground state), 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.[22] 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.[22][4] This first-order approximation underestimates the binding energy (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 perturbation 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.[22]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 wave function. The variational theorem states that for any trial wave function \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 Hamiltonian and E_\mathrm{ground} is the exact ground state energy. A simple trial wave function for the helium ground state 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 parameter representing the effective nuclear charge and a_0 is the Bohr radius. This form implicitly accounts for screening but neglects explicit electron correlation. 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}.[23] 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 nucleus from the electric field experienced by the other electron, thereby reducing the effective nuclear charge 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 nucleus simultaneously. As a result, each electron perceives a diminished nuclear attraction, leading to weaker binding compared to a non-interacting model.[24] A simple empirical approach to estimate this screening is provided by Slater's rules, which assign a shielding constant \sigma of 0.30 to the contribution from the other electron in the same 1s shell. Thus, for helium, 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 electron. This approximation captures the essence of intra-shell screening without solving the full quantum mechanical problem.[25] Within the variational framework, a more refined estimate emerges from optimizing the trial wavefunction parameter, yielding Z_{\text{eff}} = Z \left(1 - \frac{5}{16Z}\right) \approx 1.6875 for Z = 2. This value indicates that the electron-electron interaction effectively reduces the nuclear charge by about 16% , consistent with the physical picture of correlated electron densities where the probability of finding both electrons close to the nucleus is lowered due to repulsion.[1] The screening effect explains the deviation of helium's ground state energy from the independent-electron approximation, where neglecting repulsion would predict each electron bound by -13.6 Z^2 = -54.4 eV, for a total of -108.8 eV; in reality, the exact non-relativistic ground state energy is -79.0 eV, with the reduced magnitude attributable to screening that weakens the overall binding.[26]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.[14] 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.[14][27] The Pauli exclusion principle requires the total wavefunction of the two identical fermions to be antisymmetric under particle exchange, which is achieved by combining the spin and spatial parts with opposite symmetries: the antisymmetric spin function for singlets pairs with a symmetric spatial function, and the symmetric spin function for triplets pairs with an antisymmetric spatial function.[14] 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 Coulomb repulsion compared to the symmetric spatial wavefunction of singlets.[14] Consequently, for configurations with the same n and l, triplet states lie lower in energy than their corresponding singlet states, a result first quantitatively analyzed using perturbation theory in the early quantum mechanical treatments of helium.[14][27] Electric dipole transitions in helium obey the selection rule ΔS = 0, restricting optical transitions to occur within the same spin manifold—either between singlet states (parahelium series) or triplet states (orthohelium series)—due to the conservation of total spin in the weak spin-orbit coupling regime.[14] This separation underlies the historical observation of distinct spectral series for parahelium and orthohelium.[27]Helium Atomic Spectrum
The atomic spectrum of neutral helium was first identified in the Sun's chromosphere during the total solar eclipse 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.[28] Independently, British astronomer Joseph Norman Lockyer detected the same line on October 20, 1868, from ground-based observations in England using a spectroscope on the sun in daylight and proposed it originated from a new element in the Sun, naming it helium after the Greek word helios for "sun," as no matching terrestrial element was known at the time.[29] This discovery marked the first identification of an element via spectroscopy before its isolation on Earth in 1895. The emission spectrum of the helium atom features discrete lines from electron transitions between excited states, classified into series based on quantum number changes and spin multiplicity (singlet or triplet configurations). These series arise due to the Pauli exclusion principle and electron exchange effects, with orthohelium (triplet states) producing more visible lines than parahelium (singlet states). The principal series involves transitions from 1s np states to the 1s 2s state, occurring primarily in the ultraviolet 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.[30] Complementary series include the sharp series (involving s → p transitions), diffuse series (d → p), and fundamental series (f → d), each manifesting in both singlet 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 helium (He II) but was historically linked to neutral helium interpretations before Bohr's model clarified its origin.[31] The Rydberg formula, 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 effective nuclear charge 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 nuclear 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.[32] Such experiments confirm energy level assignments and support QED validations without accessing ionization thresholds.[33]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.[34] 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.[35][23] 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 hartree, corresponding to an ionization energy of approximately 23.07 eV. 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 hartree or better, yielding ionization energies approaching 24.59 eV.[34][23] The electron correlation energy, representing the adjustment to the electron-electron repulsion beyond the first-order mean-field contribution, amounts to approximately -0.042 hartree (\approx -1.14 eV) in helium, which increases the ionization energy by this amount relative to Hartree-Fock approximations (where E_\ce{HF} \approx -2.86168 hartree and I \approx 23.45 eV).[34]| Method | Ground-State Energy (hartree) | Ionization Energy (eV) |
|---|---|---|
| Perturbation Theory (1st order) | -2.75 | 20.4 |
| Perturbation Theory (high order) | ≈ -2.9037 | ≈ 24.59 |
| Variational (simple screened) | -2.8477 | 23.07 |
| Variational (improved Hylleraas) | -2.903724 | 24.59 |
| Exact non-relativistic | -2.903724377 | 24.587 |