Fact-checked by Grok 2 weeks ago

Molecular orbital theory

Molecular orbital theory is a quantum mechanical method for describing the electronic structure and bonding in molecules, in which atomic orbitals combine linearly to form molecular orbitals that are delocalized over the entire molecule, allowing electrons to occupy these orbitals according to their energy levels. This approach treats the molecule as a unified system, contrasting with valence bond theory's focus on localized electron pairs between atoms. The theory originated in the late 1920s, building on the wave nature of electrons proposed by Louis de Broglie in 1923 and Schrödinger's wave equation in 1926, with Friedrich Hund introducing the core concept of molecular orbitals as solutions to the Schrödinger equation for multi-nuclei systems around 1927–1928. It was further developed in the 1930s by Robert S. Mulliken in the United States, who emphasized the molecular unit and spectroscopy applications, and by Erich Hückel in Germany, who applied it to π-electron systems in organic molecules through the Hückel molecular orbital method. These contributions established molecular orbital theory as a powerful tool for understanding delocalized electrons, surpassing limitations of earlier valence bond models like those from Heitler and London in 1927. Central to the theory is the linear combination of atomic orbitals (LCAO) approximation, where molecular orbitals are constructed as sums and differences of atomic orbitals, yielding bonding orbitals (lower energy, increased electron density between nuclei) and antibonding orbitals (higher energy, nodal planes between nuclei). Electrons fill these orbitals following the Aufbau principle, Pauli exclusion principle, and Hund's rule, with bond order calculated as half the difference between the number of bonding and antibonding electrons, providing quantitative predictions for bond strength and stability. For example, in the diatomic molecule O₂, the configuration includes two unpaired electrons in π* orbitals, explaining its paramagnetism—a feature valence bond theory initially struggled to account for. Molecular orbital theory underpins modern computational chemistry, enabling predictions of molecular geometry, spectra, and reactivity, and has been extended to transition metal complexes and solids through methods like Hartree-Fock and density functional theory. Its delocalized perspective is essential for describing conjugation, aromaticity, and photochemical processes, as seen in the Woodward-Hoffmann rules of the 1960s.

Overview and fundamentals

Definition and basic principles

Molecular orbital theory provides a quantum mechanical framework for understanding the electronic structure and bonding in molecules by describing electrons as occupying molecular orbitals (MOs), which are mathematical functions representing one-electron wavefunctions delocalized over the entire molecule. These MOs are formed as linear combinations of atomic orbitals centered on the constituent atoms, allowing electrons to be distributed across multiple nuclei rather than being localized between pairs of atoms. This delocalization captures the collective behavior of electrons in a molecule, enabling predictions of properties such as bond strengths, molecular geometries, and spectroscopic behaviors. The molecular orbitals arise as approximate solutions to the time-independent Schrödinger equation applied to the multi-electron system of the molecule, treating the nuclei and electrons as a unified entity under the influence of their mutual electrostatic interactions. Unlike atomic systems, the exact solution for molecules beyond the simplest cases like H₂⁺ is intractable due to electron correlation and nuclear repulsion terms, necessitating approximations such as the Born-Oppenheimer separation to decouple nuclear and electronic motion. The resulting MO wavefunctions thus approximate the true electronic wavefunction, providing a basis for calculating molecular energies and properties through variational principles. In this theory, electrons occupy the available MOs in a manner that minimizes the total molecular energy, adhering to three fundamental principles: the Aufbau principle, which dictates filling orbitals starting from the lowest energy level; the Pauli exclusion principle, limiting each MO to a maximum of two electrons with opposite spins; and Hund's rule, which states that for degenerate orbitals, electrons occupy separate orbitals with parallel spins to maximize total spin multiplicity before pairing. These rules ensure the ground-state electron configuration reflects the lowest possible energy while respecting quantum mechanical constraints. A simple illustration of these principles is the diatomic hydrogen molecule (H₂), where the two overlapping 1s atomic orbitals from each hydrogen atom interfere to produce a bonding molecular orbital (σ₁ₛ) of lower energy due to constructive overlap and an antibonding orbital (σ₁ₛ*) of higher energy from destructive overlap. The two valence electrons fill the bonding σ₁ₛ orbital completely, resulting in a stable single bond and an energy level diagram showing the filled bonding MO below the atomic orbital energies, with the antibonding MO unoccupied. This configuration demonstrates how orbital overlap leads to net stabilization, forming the molecular bond.

Comparison with valence bond theory

Valence bond (VB) theory conceptualizes chemical bonding as the overlap of localized atomic orbitals from adjacent atoms, forming shared electron pairs that represent covalent bonds, often involving hybrid orbitals to achieve optimal geometry. This approach, formalized in the 1920s by Walter Heitler and Fritz London for the hydrogen molecule and extended by Linus Pauling to polyatomic systems, emphasizes the pairing of electrons with opposite spins in these localized bonds. In contrast to VB theory's focus on pairwise atomic interactions, molecular orbital (MO) theory constructs bonding descriptions using delocalized molecular orbitals that extend over the entire molecule, formed by linear combinations of atomic orbitals and filled according to the aufbau principle and Hund's rule. This delocalization in MO theory naturally accounts for resonance phenomena, such as in conjugated systems, without requiring multiple structural representations, whereas VB theory approximates such effects through resonance hybrids of localized structures. MO theory also better accommodates multicenter bonding and electron correlation in transition states, providing a more unified framework for both ground and excited states. MO theory offers significant advantages in explaining electron delocalization in extended systems, including metals where conduction bands arise from overlapping orbitals, aromatic compounds with stabilized pi systems, and odd-electron species like radicals or the triplet ground state of dioxygen, which VB theory initially struggled to describe without ad hoc adjustments like three-electron bonds. For instance, in benzene, MO theory depicts the six pi electrons occupying three delocalized molecular orbitals, yielding a single symmetric structure with equal bond lengths, while VB theory relies on two Kekulé resonance structures to approximate this uniformity, highlighting MO's efficiency in capturing conjugation. Despite these strengths, MO theory is often more computationally demanding for large molecules due to the need to solve for numerous orbitals across the system, particularly in ab initio methods, whereas VB theory's localized perspective simplifies calculations for sigma frameworks and aligns closely with intuitive chemical bonding models. In modern computational chemistry, both theories converge in their rigorous limits through methods like generalized VB or configuration interaction in MO frameworks, but VB remains preferable for qualitative interpretations of hybridization in simple molecules.

Historical development

Early contributions

The development of molecular orbital (MO) theory emerged in the mid-1920s as quantum mechanics provided a framework for understanding electronic structure beyond atomic systems. While Walter Heitler and Fritz London applied quantum mechanics to the hydrogen molecule in 1927, proposing a localized valence bond model that emphasized electron exchange between atoms, this approach highlighted the need for alternative descriptions of delocalized electrons in molecules. Friedrich Hund laid the foundational ideas for MO theory in 1927 by extending atomic orbital concepts to diatomic molecules, proposing that electrons occupy molecular orbitals formed by the interaction of atomic orbitals to explain molecular spectra. In his seminal work, Hund classified electronic states in molecules like BO, CN, and N₂ using quantum mechanical principles, introducing the notion of delocalized orbitals that span the entire molecule rather than being confined to specific bonds. Robert Mulliken advanced these concepts significantly, collaborating with Hund and formalizing MO theory through a series of papers starting in 1928, where he assigned quantum numbers to electrons in molecules and described electron configurations in terms of bonding power. By 1932, Mulliken introduced the term "orbital" to denote one-electron wavefunctions in molecules and explicitly defined bonding and antibonding molecular orbitals, with bonding orbitals resulting from constructive interference and antibonding from destructive interference of atomic orbitals. This Hund-Mulliken framework shifted emphasis from localized pairs in valence bond theory to delocalized orbitals, providing a more versatile model for molecular electronic structure. Early applications focused on the simplest molecular system, the H₂⁺ ion, where the Schrödinger equation could be solved nearly exactly. In 1927, Ø. Burrau computed the energy levels of H₂⁺ using numerical methods, yielding the first accurate molecular orbital wavefunction and demonstrating bond formation through electron delocalization between two protons. Linus Pauling further illustrated this in 1928 by applying the linear combination of atomic orbitals approximation to H₂⁺, distinguishing bonding and antibonding orbitals and showing how the bonding orbital lowers energy relative to separated atoms. These calculations validated MO theory's predictive power for dissociation energies and equilibrium bond lengths. Despite these advances, MO theory faced significant challenges in the 1920s and 1930s due to the absence of computational tools, restricting applications to simple systems like H₂⁺ and requiring semi-empirical approximations for more complex molecules. Exact solutions were infeasible for multi-electron systems, limiting widespread adoption until later methodological improvements.

Key advancements and modern refinements

In 1931, Erich Hückel introduced a semi-empirical method specifically designed for calculating the molecular orbitals of π electrons in conjugated systems, simplifying the full quantum mechanical treatment by neglecting σ bonds and using parameterized integrals to approximate electron interactions. This approach, known as Hückel molecular orbital theory, provided qualitative insights into the stability and reactivity of unsaturated hydrocarbons, such as predicting the delocalized π orbitals in benzene that contribute to its aromatic character. Hückel's method marked a significant advancement by making molecular orbital calculations computationally feasible for larger organic molecules, bridging theoretical physics and practical chemistry. During the 1950s and 1960s, John Pople advanced ab initio molecular orbital theory by developing systematic computational frameworks that integrated the Hartree-Fock self-consistent field approximation with molecular orbitals, enabling accurate predictions of molecular geometries and energies without empirical parameters. Pople's innovations included the development and standardization of Gaussian basis sets, such as the STO-nG and 6-31G families, which use contracted Gaussians to efficiently approximate Slater-type orbitals and facilitate evaluation of electron repulsion integrals essential for Hartree-Fock calculations, as detailed in his work leading to the Gaussian series of software programs. These developments, culminating in the 1970s with programs like Gaussian 70, transformed molecular orbital theory into a cornerstone of computational quantum chemistry, earning Pople the 1998 Nobel Prize in Chemistry shared with Walter Kohn. In the 1960s, the application of group theory to molecular orbital construction gained prominence through the work of F. Albert Cotton and others, who formalized the use of symmetry-adapted linear combinations of atomic orbitals to generate molecular orbitals that respect the point group symmetry of the molecule. Cotton's seminal textbook outlined how irreducible representations from group theory could classify and simplify the mixing of atomic orbitals, reducing computational complexity and providing a deeper understanding of electronic transitions and bonding symmetries in complex molecules like transition metal complexes. This refinement enhanced the predictive power of molecular orbital theory by ensuring orbitals were properly symmetrized, influencing subsequent developments in inorganic and organometallic chemistry. The 1970s and 1980s saw the rise of density functional theory (DFT) as a powerful extension of molecular orbital approaches, where Kohn-Sham orbitals serve as a basis to compute the electron density directly, incorporating exchange-correlation effects through approximate functionals. Building on the foundational Hohenberg-Kohn theorems of 1964, practical implementations in the 1970s, such as local density approximations, and 1980s advancements like generalized gradient approximations, made DFT computationally efficient for large systems while maintaining molecular orbital-like descriptions. This methodology's impact was recognized with the 1998 Nobel Prize in Chemistry awarded to Walter Kohn, revolutionizing quantum chemical calculations for molecular properties and reaction pathways. Post-2000 refinements to molecular orbital theory have focused on incorporating relativistic effects and treating excited states more accurately within quantum chemistry software suites like Gaussian and ORCA. Relativistic corrections, such as scalar relativistic pseudopotentials and Dirac-Hartree-Fock methods, address the influence of high atomic numbers on orbital energies and bonding, as exemplified in studies of heavy-element compounds where spin-orbit coupling alters molecular symmetries. For excited states, equation-of-motion coupled-cluster methods and time-dependent DFT have become standard, enabling reliable predictions of electronic spectra and photochemistry by extending single-reference molecular orbital frameworks to multireference configurations. These advancements, integrated into modern software, have expanded molecular orbital theory's applicability to relativistic systems and dynamic processes, supporting fields from materials science to biochemistry.

Construction of molecular orbitals

Linear combination of atomic orbitals method

The linear combination of atomic orbitals (LCAO) method provides a practical approximation for constructing molecular orbitals in quantum chemistry by expressing them as weighted sums of atomic orbitals centered on the atoms of the molecule. This approach, pioneered in the late 1920s and early 1930s, assumes that the molecular wavefunction can be built from the atomic orbitals of the isolated atoms, adjusted for the molecular environment. In the LCAO framework, a molecular orbital \psi_{\text{MO}} is written as \psi_{\text{MO}} = \sum_i c_i \phi_i, where \phi_i are the basis atomic orbitals (typically Slater-type or Gaussian functions) and c_i are variational coefficients that determine the contribution of each atomic orbital to the molecular orbital. These coefficients are optimized to best approximate the true molecular eigenfunctions. To determine the coefficients c_i and the corresponding orbital energies, the method employs the variational principle through the Rayleigh-Ritz procedure, which seeks to minimize the expectation value of the energy for trial wavefunctions within the chosen basis set. This minimization leads to a generalized eigenvalue problem known as the secular equation: \mathbf{H} \mathbf{c} = E \mathbf{S} \mathbf{c}, where \mathbf{H} is the Hamiltonian matrix with elements H_{ij} = \langle \phi_i | \hat{H} | \phi_j \rangle (the integrals of the Hamiltonian operator over the basis functions), \mathbf{S} is the overlap matrix with elements S_{ij} = \langle \phi_i | \phi_j \rangle, \mathbf{c} is the column vector of coefficients, and E is the energy eigenvalue. The secular equation is solved by diagonalizing the matrix \mathbf{H} in the basis defined by \mathbf{S}, yielding the lowest-energy solutions corresponding to occupied molecular orbitals. This formulation accounts for electron-nuclear attraction, electron-electron repulsion (in more advanced implementations), and kinetic energy contributions inherent in the Hamiltonian \hat{H}. The atomic orbitals forming the basis set must satisfy normalization and, ideally, orthogonality conditions to ensure a well-defined expansion. Normalization requires that each \phi_i obeys \langle \phi_i | \phi_i \rangle = 1, preventing arbitrary scaling that could distort energy estimates. Orthogonality, where \langle \phi_i | \phi_j \rangle = 0 for i \neq j, simplifies the secular equation by making \mathbf{S} the identity matrix, but in practice, basis sets are often non-orthogonal due to the spatial overlap of atomic orbitals on neighboring atoms, necessitating the inclusion of the overlap matrix \mathbf{S} to maintain accuracy. The choice of basis set size and quality directly impacts the reliability of the LCAO approximation, with larger sets providing better approximations at increased computational cost. A classic illustration of the LCAO method is its application to the hydrogen molecule (H₂), where the simplest basis consists of two 1s atomic orbitals, \phi_A and \phi_B, centered on each hydrogen nucleus. The bonding molecular orbital, which concentrates electron density between the nuclei to stabilize the bond, takes the symmetric form \psi_{\text{bonding}} = \frac{\phi_A + \phi_B}{\sqrt{2 + 2S}}, while the antibonding orbital, which places a nodal plane between the nuclei and destabilizes the system, is \psi_{\text{antibonding}} = \frac{\phi_A - \phi_B}{\sqrt{2 - 2S}}. Here, S = \langle \phi_A | \phi_B \rangle is the overlap integral, which quantifies the extent of orbital overlap and lies between 0 and 1 for H₂ at equilibrium distance; the denominators ensure normalization of the molecular orbitals. Solving the secular equation for this two-orbital basis yields the bonding energy below the atomic 1s level and the antibonding energy above it, demonstrating how LCAO captures the splitting of atomic levels into molecular orbitals.

Symmetry considerations and approximations

In molecular orbital theory, the symmetry properties of a molecule, characterized by its point group, play a crucial role in simplifying the construction of molecular orbitals via the linear combination of atomic orbitals (LCAO) method. Atomic orbitals are classified according to the irreducible representations (irreps) of the point group, ensuring that only orbitals transforming under the same irrep can mix to form molecular orbitals. This symmetry classification reduces the size of the secular determinant, transforming the full matrix into smaller block-diagonal submatrices, one for each irrep, which drastically lowers the computational complexity for larger systems. To exploit this symmetry efficiently, basis atomic orbitals are combined into symmetry-adapted linear combinations (SALCs), which are linear combinations that individually transform as a single irrep of the point group. SALCs are generated by projecting the basis orbitals onto the irreps using the group's character table and projection operators, grouping equivalent orbitals (e.g., those on identical atoms) into sets that match the molecule's symmetry. This approach not only confirms which combinations are allowed but also provides the initial guess for molecular orbitals within each symmetry block, enhancing both accuracy and efficiency in solving the secular equations. Further approximations are essential for practical calculations, particularly regarding overlap integrals in the LCAO framework. The zero-overlap approximation sets the overlap integral S_{ij} = \langle \phi_i | \phi_j \rangle = 0 for i \neq j, assuming atomic orbitals on different centers are orthogonal or their overlap is negligible due to distance. This simplifies the overlap matrix to the identity, decoupling the secular equation and avoiding the need to compute or invert non-unitary matrices, though it introduces some error in bond length dependencies. Semi-empirical methods build on these symmetries by parameterizing the Hamiltonian integrals to incorporate experimental data. In Hückel molecular orbital theory, a foundational semi-empirical approach for π systems, the diagonal elements of the Hamiltonian are set to the coulomb integral \alpha, representing the energy of an isolated atomic orbital, while off-diagonal elements between adjacent orbitals are the resonance integral \beta, capturing bonding interactions; all other integrals are zero. These parameters are not computed ab initio but adjusted empirically (\alpha often around -11 eV for carbon, \beta about -3 to -4 eV), allowing rapid predictions of orbital energies while relying on symmetry to limit the basis. An illustrative example is the application of these techniques to the oxygen molecule (O₂), which belongs to the D_{\infty h} point group. Only valence 2s and 2p orbitals are included in the LCAO basis to focus on bonding, reducing the full set from core-involved calculations. Symmetry classifies these into σ_g, σ_u, π_u, and π_g irreps, separating the secular determinant into independent 1×1 or 2×2 blocks (e.g., the π_u block mixes the two p_y orbitals into degenerate combinations). With the zero-overlap approximation, the resulting molecular orbitals explain O₂'s double bond and paramagnetism without solving a large matrix.

Classification and types of molecular orbitals

Bonding, antibonding, and non-bonding orbitals

In molecular orbital theory, molecular orbitals are categorized as bonding, antibonding, or non-bonding based on their energies relative to the parent atomic orbitals and their influence on electron density distribution, particularly in the internuclear region. Bonding molecular orbitals form through constructive overlap of atomic orbitals, resulting in lower energy levels than the isolated atomic orbitals and an accumulation of electron density between the atomic nuclei, which promotes attraction between the positively charged nuclei and stabilizes the molecule. This increased density enhances the electrostatic shielding and lowers the overall potential energy of the system. Antibonding molecular orbitals, in contrast, arise from destructive interference of atomic orbitals, leading to higher energy levels than the atomic orbitals and a nodal plane perpendicular to the internuclear axis where electron density is minimized or absent. This configuration places electron density primarily outside the bonding region, effectively repelling the nuclei and destabilizing the molecule when occupied. The presence of a node between the nuclei distinguishes these orbitals and reduces overlap, contributing to their elevated energy. Non-bonding molecular orbitals exhibit energies comparable to those of the contributing atomic orbitals and do not produce a net increase or decrease in electron density between the nuclei, as their wavefunctions are localized primarily on one atom with minimal overlap. These orbitals typically represent lone pairs or electrons in atomic orbitals that do not participate effectively in bonding, maintaining an atomic-like character without significantly affecting molecular stability. The general energy ordering places bonding orbitals below the atomic orbital energies, non-bonding orbitals at approximately the same level, and antibonding orbitals above, with the degree of orbital occupancy dictating bond strength—electrons in bonding orbitals strengthen the bond, while those in antibonding orbitals weaken it, and non-bonding electrons contribute neither effect. For instance, in the homonuclear diatomic molecule N₂, a qualitative molecular orbital diagram illustrates several filled bonding orbitals and empty antibonding orbitals, yielding a stable configuration with electrons occupying lower-energy bonding states that promote a strong triple bond.

Sigma, pi, and other orbital symmetries

Molecular orbitals in diatomic molecules are classified by their angular momentum quantum number and symmetry relative to the internuclear axis, using Greek letters to indicate the number of vertical nodal planes: σ for zero, π for one, δ for two, and φ for three. This notation, introduced by Robert S. Mulliken in the late 1920s, draws an analogy to atomic orbital labels (s, p, d, f) and reflects the projection of orbital angular momentum along the bond axis. Sigma (σ) molecular orbitals form through end-on (head-on) overlap of atomic orbitals aligned along the internuclear axis, producing no angular nodes in planes containing that axis and resulting in cylindrical symmetry. Common examples include the overlap of two s orbitals or two p_z orbitals (where z is the bond axis), concentrating electron density directly between the nuclei. In homonuclear diatomics like F₂, the valence σ orbitals derive primarily from 2s and 2p_z atomic orbitals, forming the backbone of the bonding framework. Pi (π) molecular orbitals arise from parallel, sideways overlap of atomic p orbitals, such as p_x-p_x or p_y-p_y, introducing one nodal plane containing the internuclear axis and localizing electron density in lobes above and below the bond. This overlap is less efficient than σ due to poorer directional alignment, leading to higher energy π orbitals compared to σ in many cases. In F₂, the degenerate π orbitals stem from the 2p_x and 2p_y sets, contributing to the overall molecular stability. For heavier elements, particularly transition metals, delta (δ) molecular orbitals emerge from the side-on overlap of d orbitals, such as d_{xy} with d_{xy} or d_{x^2-y^2} with d_{x^2-y^2}, featuring two nodal planes containing the bond axis. These δ interactions are observed in metal-metal bonds, as in the Re-Re δ bond of [Re₂Cl₈]²⁻. Phi (φ) molecular orbitals, analogous but rarer, involve f orbital overlaps with three such nodal planes and appear in f-block actinide complexes where oxidation state variations enable selective activation of higher-order bonding. In polyatomic molecules, σ symmetries often incorporate hybrid atomic orbitals for directional bonding. For instance, sp hybrid orbitals on a central atom, formed by mixing one s and one p orbital, overlap end-on with ligand orbitals to create linear σ frameworks, as in BeH₂ where two sp hybrids on Be form σ bonds with H 1s orbitals. These hybrid contributions extend σ bonding beyond simple atomic overlaps while maintaining the characteristic head-on symmetry. Heteronuclear diatomics like CO illustrate how σ and π symmetries adapt to differing atomic energies. In CO's molecular orbital diagram, σ orbitals from C 2p_z and O 2p_z show polarization with density shifted toward the more electronegative oxygen, while π orbitals from 2p_x/y sets exhibit similar asymmetry, enhancing the molecule's reactivity at carbon. This contrasts with F₂'s symmetric homonuclear arrangement, where equal atomic contributions yield balanced σ and π densities.

Applications and properties

Bond order and molecular stability

In molecular orbital theory, the bond order (BO) provides a quantitative measure of the strength and stability of a chemical bond between two atoms, calculated as half the difference between the number of electrons occupying bonding molecular orbitals (n_b) and those in antibonding molecular orbitals (n_a): \text{BO} = \frac{1}{2} (n_b - n_a) This formula reflects the net bonding contribution from electron occupancy in the molecular orbital diagram. A bond order of 1 corresponds to a single bond, as seen in H_2, where two electrons fill a bonding \sigma_{1s} orbital and none occupy the antibonding counterpart. Fractional values, such as 0.5 in the H_2^+ ion, indicate a weaker, half-filled bond with one electron in the bonding orbital, predicting lower stability compared to neutral H_2. Integer values greater than 1 denote multiple bonds, while a bond order of 0 signals instability and the absence of a bound molecule. For homonuclear diatomic molecules, bond order occupancy directly predicts molecular existence and bonding type. Nitrogen (N_2) has a bond order of 3, with 8 bonding electrons (from \sigma_{2s}, \sigma_{2p_z}, and two \pi_{2p} orbitals) and 2 antibonding electrons (in \sigma_{2s}^* orbital), corresponding to a strong triple bond. Oxygen (O_2) exhibits a bond order of 2, with 8 bonding and 4 antibonding electrons, forming a double bond that is paramagnetic due to two unpaired electrons in degenerate \pi^* orbitals. In contrast, helium (He_2) has a bond order of 0, as its four electrons occupy the bonding \sigma_{1s} and antibonding \sigma_{1s}^* orbitals with two electrons each, rendering the molecule unstable and unbound under standard conditions. Higher bond orders correlate with increased bond strength, manifested in greater dissociation energies and shorter bond lengths, as the net electron density between nuclei enhances orbital overlap and lowers potential energy. For instance, N_2 has a dissociation energy of approximately 941 kJ/mol and a bond length of 110 pm, compared to O_2's 498 kJ/mol and 121 pm, reflecting the trend where each incremental bond order unit strengthens the bond. In heteronuclear diatomics, fractional bond orders arise due to unequal atomic orbital contributions, as in nitric oxide (NO) with a bond order of 2.5 (11 valence electrons: 8 bonding, 3 antibonding), resulting in an intermediate bond strength between double and triple bonds. This fractional value underscores how molecular orbital theory accommodates varying electronegativities without integer constraints, aiding predictions of reactivity in such species.

Magnetism and unpaired electrons

Molecular orbital theory accounts for the magnetic properties of molecules by examining the occupancy of molecular orbitals, particularly whether electrons are paired or unpaired. Paramagnetic molecules possess one or more unpaired electrons, leading to a net magnetic moment that causes attraction to an external magnetic field, while diamagnetic molecules have all electrons in paired states within filled orbitals, resulting in no net magnetic moment and weak repulsion from the field. This distinction arises from the electron configuration determined by the aufbau principle combined with Hund's rule of maximum multiplicity, which dictates that, in a set of degenerate molecular orbitals, electrons will occupy separate orbitals with parallel spins before pairing to achieve the lowest energy state with the highest total spin. A classic example is the dioxygen molecule (O₂), where molecular orbital theory predicts a ground state with two unpaired electrons in the degenerate π{2p} antibonding orbitals. According to the MO configuration (σ{1s}^2 σ^{1s}^2 σ{2s}^2 σ^{2s}^2 σ{2p}^2 π_{2p}^4 π^{2p}^2), the two electrons in the π^*{2p} orbitals occupy each degenerate orbital singly with parallel spins, forming a triplet state (³Σ_g^-) that explains the observed paramagnetism of O₂. This prediction was a key early success of MO theory, first proposed by Friedrich Hund in his analysis of molecular spectra, resolving a limitation of valence bond theory which incorrectly suggested a singlet ground state. In contrast, the nitrogen molecule (N₂) has the MO configuration (σ_{1s}^2 σ^{1s}^2 σ{2s}^2 σ^{2s}^2 π{2p}^4 σ_{2p}^2) with all 10 valence electrons paired in filled orbitals, rendering it diamagnetic as confirmed experimentally. The influence of atomic number on orbital ordering further illustrates these magnetic behaviors in lighter diatomic molecules. For diboron (B₂), strong s-p mixing raises the energy of the σ_{2p} orbital above the degenerate π_{2p} orbitals, leading to the valence configuration (σ_{2s}^2 σ^{2s}^2 π{2p}^2) where the two electrons in the π_{2p} set are unpaired with parallel spins per Hund's rule, making B₂ paramagnetic. In dicarbon (C₂), the additional two electrons fully pair within the same π_{2p} orbitals (σ_{2s}^2 σ^{2s}^2 π{2p}^4), resulting in a diamagnetic ground state. These predictions align with experimental observations of B₂'s paramagnetism in vapor phase studies and C₂'s diamagnetism. Unpaired electrons predicted by MO theory also enable spectroscopic detection via electron spin resonance (ESR), a technique that measures the absorption of microwave radiation by the magnetic moments of these electrons in a magnetic field, providing insights into spin states and molecular environments in paramagnetic species.

Extensions to complex systems

Polyatomic molecules and delocalization

In molecular orbital theory, the construction of molecular orbitals for polyatomic molecules employs the linear combination of atomic orbitals (LCAO) method, where valence atomic orbitals from all atoms in the molecule are combined to form a set of molecular orbitals equal in number to the contributing atomic orbitals. This results in multiple bonding and antibonding pairs that accommodate the total valence electrons, providing a delocalized description of the electronic structure across the entire system. Seminal semiempirical approaches, such as the complete neglect of differential overlap (CNDO) method developed by Pople, Santry, and Segal, parameterized the integrals in these calculations using atomic data for elements like carbon, nitrogen, oxygen, and fluorine, enabling practical computations for general polyatomics. A central concept in applying molecular orbital theory to polyatomic molecules is electron delocalization, in which electrons occupy orbitals extending over multiple atoms rather than being confined to specific bonds. This delocalization arises naturally from the LCAO framework, as the molecular orbitals are linear superpositions of atomic contributions, leading to electron density distributed throughout the molecule and enhancing stability, particularly in systems with conjugated or extended frameworks. Early foundational work by Mulliken emphasized how this delocalized nature distinguishes molecular orbital theory from localized valence bond models, allowing for better prediction of properties in complex molecules. An illustrative example of delocalization in polyatomic molecules is the allyl cation, a three-carbon system (C₃H₅⁺) with a conjugated pi framework. Here, the three p orbitals from the carbon atoms combine to form three pi molecular orbitals: a fully bonding orbital (lowest energy, occupied by two electrons), a non-bonding orbital, and an antibonding orbital. The delocalized nature of the bonding pi molecular orbital spreads the electron density over all three carbons, lowering the overall energy compared to a localized double bond plus isolated p orbital and stabilizing the positive charge. In linear triatomic molecules such as CO₂, molecular orbital theory reveals delocalization effects that result in equivalent C-O bonds due to the symmetric spreading of sigma and pi electron density across the O-C-O framework. The valence molecular orbitals, formed from carbon's 2s, 2p and oxygen's 2s, 2p orbitals, include delocalized bonding MOs such as the sigma 3σ_g and pi 1π_u, along with non-bonding 4σ_g (O 2pσ lone pairs), yielding equivalent bond lengths of approximately 1.16 Å and an effective bond order of 2. Molecular orbital theory further ties into the prediction of aromaticity in polyatomic cyclic systems, where delocalized pi orbitals filled with 4n+2 electrons (n = 0,1,2,...) achieve closed-shell configurations that confer exceptional stability, as originally formulated in Hückel's framework for conjugated pi systems. MO theory extends further to periodic solids via crystal orbital methods and to coordination compounds through ligand field molecular orbital approaches, incorporating d-orbitals and metal-ligand interactions.

Hückel molecular orbital theory for conjugated systems

Hückel molecular orbital (HMO) theory is a semi-empirical quantum mechanical method developed to approximate the π molecular orbitals in planar conjugated hydrocarbon systems, such as alternant hydrocarbons with sp²-hybridized carbon atoms. It simplifies the full molecular orbital approach by focusing exclusively on the π electrons contributed by p_z atomic orbitals perpendicular to the molecular plane, while treating the σ framework as a rigid, non-interacting backbone that provides the geometric scaffold. This approximation neglects σ-π interactions and electron-electron repulsion, assuming overlap only between adjacent p orbitals in the conjugated chain or ring. The method solves the Schrödinger equation variationally using a linear combination of atomic orbitals (LCAO) basis, yielding qualitative insights into electronic structure, stability, and reactivity in organic molecules. Central to HMO theory are two empirical parameters: α, the Coulomb integral representing the energy of an electron in an isolated 2p_z orbital (typically set as the zero-energy reference), and β, the resonance (or overlap) integral quantifying the interaction between adjacent p orbitals, which is negative (β < 0) to reflect bonding stabilization. All non-adjacent interactions are ignored (set to zero), and the overlap integral S is approximated as zero for simplicity. For linear conjugated systems like polyenes, the Hamiltonian matrix is tridiagonal, leading to a secular determinant whose eigenvalues give the π orbital energies. In cyclic conjugated systems, such as annulenes with n atoms, the secular equation incorporates periodic boundary conditions, resulting in closed-form solutions for the energy levels: E_k = \alpha + 2\beta \cos\left(\frac{2\pi k}{n}\right) where k = 0, 1, ..., n-1 labels the molecular orbitals, ordered from lowest to highest energy. Each non-degenerate orbital (except degeneracies at k and n-k) is doubly occupied in the ground state up to the total number of π electrons. A graphical mnemonic for these cyclic energy levels, known as the Frost circle method, aids visualization: a regular n-sided polygon is inscribed in a circle with one vertex at the bottom (touching the energy reference α), and the vertical positions of the vertices correspond to the orbital energies scaled by |β| (with bonding orbitals below α and antibonding above). Degenerate pairs occur at equal heights, and the lowest orbital is always non-degenerate and fully bonding. For neutral hydrocarbons, the lowest n/2 orbitals (or appropriate filled set) are doubly occupied, providing a quick assessment of stability; for example, systems with 4n+2 π electrons fill only bonding orbitals, enhancing aromatic character. This method efficiently predicts the pattern without solving the determinant explicitly. Illustrative examples highlight HMO applications. In ethene (n=2), the secular determinant yields two π orbitals: a bonding MO at energy α + β (doubly occupied by two electrons) and an antibonding MO at α - β, establishing the basic C=C double bond with delocalization over the two carbons. For 1,3-butadiene (n=4, linear), the four π MOs have energies α + 1.618β (bonding, filled), α + 0.618β (bonding, filled), α - 0.618β (antibonding, empty), and α - 1.618β (antibonding, empty), demonstrating partial delocalization across the chain and a total π stabilization energy of approximately 2.236|β| relative to two isolated double bonds. Benzene (n=6, cyclic), a cornerstone case, features six π MOs: the lowest at α + 2β (doubly occupied), a degenerate pair at α + β (each doubly occupied), a degenerate pair at α - β (empty), and the highest at α - 2β (empty). With six π electrons, all bonding orbitals are filled, yielding a delocalization energy of 2|β| and uniform π electron distribution, underpinning its aromatic stability. HMO theory extends to quantitative predictions of molecular properties. Bond orders are calculated as weighted sums of orbital coefficients between adjacent atoms, p_{ij} = ∑{occupied} c{i,m} c_{j,m}, where values between 1 and 2 indicate partial double-bond character; in benzene, all C-C bonds have order 1.5, reflecting equalization. Charge densities at atom i, q_i = 1 + ∑{occupied} |c{i,m}|^2 (for neutral hydrocarbons with one π electron per carbon), are unity in symmetric cases like benzene but vary in alternants or substituted systems, guiding reactivity—for instance, electrophilic attack prefers sites of highest π density in alternant hydrocarbons. These predictions align with observed patterns in conjugated systems, such as ortho/para directing effects in electrophilic aromatic substitution.

References

  1. [1]
    Molecular Orbital theory (MO) is the most important ... - Chemistry 301
    Molecular Orbital theory (MO) is the most important quantum mechanical theory for describing bonding in molecules. It is an approximate theory (as any theory ...
  2. [2]
    7.7 Molecular Orbital Theory – Chemistry Fundamentals
    Molecular orbital (MO) theory describes electron behavior in a molecule using combinations of atomic wave functions. It forms bonding and antibonding orbitals.
  3. [3]
    Physics in Chemistry from the 1920s to the 1960s | Isis
    Sep 19, 2018 · The Hund-Mulliken-Hückel approach became the “molecular orbital theory.” The “valence bond theory” was more congenial to chemists because it ...
  4. [4]
    [PDF] An Introduction to Hartree-Fock Molecular Orbital Theory
    Hartree-Fock theory is fundamental to electronic structure, describing each electron's motion by a single-particle function, and is the basis of molecular ...<|control11|><|separator|>
  5. [5]
    [PDF] Robert S. Mulliken - Nobel Lecture
    1966 ROBERT S.MULLIKEN. 6. See the introduction to *Report on Molecular Orbital Theory*, J. Chim. Phys., 46. (1949) 497, and references given there. 7. F. Hund, ...
  6. [6]
  7. [7]
  8. [8]
  9. [9]
  10. [10]
  11. [11]
  12. [12]
    Burrau, . (1927) Berechnung des Energiewertes des ... - Scirp.org.
    Jul 31, 2024 · Burrau, Ø. (1927) Berechnung des Energiewertes des Wasserstoffmolekül-Ions H+2 im Normalzustand. Det Kgl Danske Videnskabernes Selskab, 7, 2-18.Missing: O. | Show results with:O.
  13. [13]
    The Application of the Quantum Mechanics to the Structure of the ...
    The application of the quantum mechanics to the structure of the hydrogen molecule and hydrogen molecule-ion and to related problems.
  14. [14]
    [PDF] THE PATH TO MOLECULAR ORBITAL THEORY - iupac
    ABSTRACT. The early history of the development of molecular orbital (MO) theory is reviewed. Aided by analogies to atomic spectra and based on evidence from.Missing: Friedrich Robert
  15. [15]
    [PDF] John A. Pople - Nobel Lecture
    the orbital theory of many-electron systems was laid by Hartree, Fock and ... Schleyer and J. A. Pople, "Ab Initio Molecular Orbital. Theory", John Wiley ...
  16. [16]
    Perspective: Relativistic effects | The Journal of Chemical Physics
    This perspective article discusses some broadly-known and some less broadly-known consequences of Einstein's special relativity in quantum chemistry.
  17. [17]
    Multireference Approaches for Excited States of Molecules
    This review is devoted to a specific class of approaches, usually denoted as multireference (MR) methods, the generality of which is needed for solving many ...
  18. [18]
    [PDF] Lecture 13: Molecular Orbital Theory - MIT OpenCourseWare
    Oct 6, 2014 · Readings for today: Section 3.8 – 3.11 Molecular Orbital Theory (Same in 5th and 4th ed.) ... generate 2 molecular orbitals (one bonding and one ...
  19. [19]
    9.7: Molecular Orbitals - Chemistry LibreTexts
    Jul 7, 2023 · In molecular orbital theory, electrons in antibonding orbitals effectively cancel the stabilization resulting from electrons in bonding orbitals ...Missing: seminal papers
  20. [20]
    5.1.3: Molecular orbitals from d orbitals - Chemistry LibreTexts
    Sep 11, 2022 · Generally, there are three types of bonding and antibonding interactions that may occur with d orbitals: sigma ( σ ), pi ( π ), and delta ( δ ) ...Missing: phi | Show results with:phi
  21. [21]
  22. [22]
    Unlocking Novel δ and φ Bonding Modes in Actinides via Oxidation ...
    Apr 14, 2025 · In this work, we demonstrate that variations in actinide oxidation state across the early actinide series can be used as a lever to selectively activate or ...Missing: phi | Show results with:phi
  23. [23]
    10: Bonding in Polyatomic Molecules - Chemistry LibreTexts
    Nov 15, 2022 · This page discusses molecular orbital theory's application to polyatomic molecules, highlighting hybrid orbitals in molecular shapes and ...
  24. [24]
    5.3.2: Polar bonds - Chemistry LibreTexts
    May 3, 2023 · The valence molecular orbitals in both atoms are the 2 ⁢ s and 2 ⁢ p orbitals. The molecular orbital diagram for carbon monoxide (Figure 5 .
  25. [25]
  26. [26]
    [PDF] Molecular Orbital Theory - MIT OpenCourseWare
    (c) The N-N bond is stronger in N2 since an electron is removed from an anti- bonding orbital, increasing the bond order from 2.5 to 3. Page 2. LECTURE 13. 3. ( ...
  27. [27]
    [PDF] Molecular Orbital (MO) Theory
    The higher energy anti-bonding orbital (2) destabilizes the molecule. Consider H2. When the 1s orbitals of each H- atom subtract, an anti-bonding orbital forms.
  28. [28]
    Valence Bond Theory—Its Birth, Struggles with Molecular Orbital ...
    Mar 15, 2021 · This essay describes the successive births of valence bond (VB) theory during 1916–1931. The alternative molecular orbital (MO) theory was born in the late ...
  29. [29]
    Zur Deutung einiger Erscheinungen in den Molekelspektren
    Cite this article. Hund, F. Zur Deutung einiger Erscheinungen in den Molekelspektren. Z. Physik 36, 657–674 (1926). https://doi.org/10.1007/BF01400155.
  30. [30]
    [PDF] Bond-order and Magnetic Behavior of Diatomic Species without ...
    (1-20)e-s using molecular orbital theory is time consuming. Keeping ... -,B2. +. 9. 1. 1.73. Para magnetic. B2, Be2. 2-, HF. 10. 2. 2.82. Para magnetic. B2. -,C2.
  31. [31]
    Semiempirical Molecular Orbitals for General Polyatomic Molecules ...
    A consistent one‐electron semiempirical molecular orbital method based on the approximation Hij=Sij(2−| Sij |)(Hii+Hjj)/2has been used with atomic data to ...
  32. [32]
    The origins and later developments of molecular orbital theory
    Jan 23, 2012 · Hund in 1926 speculated on the assignment of quantum numbers for the electronic states of diatomic molecules.
  33. [33]
    Quantentheoretische Beiträge zum Benzolproblem
    Quantentheoretische Beiträge zum Benzolproblem. I. Die Elektronenkonfiguration des Benzols und verwandter Verbindungen. Published: March 1931. Volume 70, pages ...
  34. [34]