Fact-checked by Grok 2 weeks ago

Molecular physics

Molecular physics is the study of the physical properties of molecules, including the chemical bonds between atoms that form them, as well as and interactions with light and matter. This field examines multi-atomic structures through their internal motions—such as quantized rotations and vibrations—and external interactions, providing fundamental insights into molecular behavior at the quantum level. Key concepts in molecular physics revolve around the theoretical and experimental analysis of molecular structure and energy states, including electronic excitations, curves, and spectroscopic transitions. serves as a primary method, utilizing wavelengths from (for rotations, 30–150 µm) to visible/ (for electronic transitions) to determine properties like internuclear distances and bond strengths. The discipline overlaps significantly with , , and , often employing quantum mechanical models to predict molecular spectra and dynamics. Historically, molecular physics emerged in the early alongside the development of , building on foundational work in atomic structure to explain molecular bonding and spectra. Pioneering contributions from scientists like and laid the groundwork for understanding molecular quantum states, evolving into modern computational and laser-based techniques. Applications of molecular physics extend to diverse areas, including the design of quantum technologies like atomic clocks and sensors, advancements in energy conversion processes, and through precise spectroscopic analysis. It also supports by interpreting molecular signatures in spectra and aids in developing novel compounds with tailored properties.

Foundations

Definition and Scope

Molecular physics is the branch of physics that investigates the physical properties and behavior of molecules, encompassing their , states, , and interactions with , primarily through the application of and . This field employs physical principles to analyze molecules as quantum systems, deriving insights into their internal motions and responses to external perturbations. The scope of molecular physics extends to isolated molecules, ensembles in gases or liquids, and their behaviors under external fields such as electric, magnetic, or electromagnetic influences, with an emphasis on fundamental physical laws rather than chemical transformation processes. It includes the study of molecular responses in diverse environments, from ultracold gases to high-pressure conditions, providing a for understanding at the molecular without prioritizing reactivity. In distinction from , molecular physics adopts a physics-centric viewpoint that prioritizes the mathematical description of molecular wavefunctions, energy quantization, and spectroscopic signatures to elucidate the underlying principles of matter, often housed in physics departments and focusing on theoretical and experimental tools like and . This approach highlights the field's role in bridging , which examines individual atoms and electrons, with , which explores collective behaviors in solids and liquids through molecular interactions. Central to molecular physics are key concepts such as the treatment of molecules as quantum mechanical systems exhibiting discrete levels for rotational, vibrational, and electronic states, which govern their stability and transitions. These quantized levels arise from the applied to molecular Hamiltonians, enabling precise predictions of molecular properties and their evolution under physical influences.

Historical Development

The foundations of molecular physics emerged in the early 20th century through pioneering studies of molecular spectra, with Otto Oldenberg contributing key insights into the excitation of molecular vibrations by electron impact during the 1910s and 1920s. These efforts built on the emerging understanding of atomic and molecular interactions, laying groundwork for interpreting band spectra in gases. A seminal advancement came in 1925 with the formulation of the by , which explained the vertical transitions in spectra of molecules, emphasizing the rapid motion relative to electronic changes during or . The quantum era of the 1930s marked a transformative shift, as models were developed to describe molecular rotations, with John H. Van Vleck extending early quantum treatments to account for levels in diatomic molecules. Concurrently, the model was refined for molecular vibrations, approximating potential energy curves as parabolic to predict quantized vibrational states, enabling quantitative analysis of infrared spectra. These quantum mechanical frameworks, rooted in the broader foundations of wave mechanics, provided the theoretical backbone for interpreting spectroscopic data beyond classical approximations. Post-World War II advancements in the 1940s revolutionized experimental techniques, particularly through Charles H. Townes's work on , which allowed precise measurements of rotational transitions in gases at centimeter wavelengths. By the 1960s, the advent of laser spectroscopy further enhanced resolution, enabling detailed studies of molecular energy levels with tunable coherent light sources that probed fine structures unattainable by earlier methods. Key milestones in the late 20th century included the 1997 awarded to , , and William D. Phillips for developing and techniques, which extended to molecules and facilitated ultracold studies of their dynamics. In the , ultrafast emerged as a powerful tool, using femtosecond pulses to capture molecular processes like bond breaking and energy transfer. This period also witnessed a pivotal by the 1980s toward ab initio computational methods, where effective core potentials enabled accurate quantum calculations of molecular structures without empirical parameters. In the and , molecular physics advanced further with the creation of ultracold molecules for quantum and control, enabling studies of quantum many-body physics and chemistry at near-absolute zero temperatures. A major breakthrough came in 2023 with the awarded to , , and for experimental methods generating light pulses, allowing real-time observation of motion in molecules and advancing understanding of ultrafast dynamics. As of 2025, ongoing research integrates with quantum simulations to predict complex molecular behaviors, bridging theoretical models with experimental precision.

Molecular Structure

Atomic Bonds and Geometry

Atomic bonds form the foundation of molecular structure, arising from interactions between atoms that minimize the overall of the system. In molecular physics, bonds are classified into two primary types relevant to molecules: covalent and ionic. Covalent bonds result from the sharing of between atoms, typically between , leading to stable electron pairs that hold the atoms together. Ionic bonds occur through electrostatic between oppositely charged ions, often formed when are transferred from a metal to a ; while primarily characteristic of ionic solids, molecules can exhibit significant ionic character in polar bonds. These bond types are distinguished by their electron distribution and strength, with covalent bonds often exhibiting intermediate polarity based on electronegativity differences. The curve describing formation, particularly for diatomic molecules, is often modeled by the , which captures the of vibrational motion better than simpler approximations. The is given by V(r) = D_e \left(1 - e^{-a(r - r_e)}\right)^2 where D_e is the depth of the ( energy), r_e is the equilibrium , r is the internuclear distance, and a is a related to the vibrational frequency. This function approaches D_e as r increases to , with the minimum at zero representing the equilibrium bonded state; D_e is the energy required to separate the atoms to infinite distance. Derived from quantum mechanical considerations of diatomic systems, the provides a realistic description of stretching and compression, essential for understanding molecular stability. Molecular geometry, the spatial arrangement of atoms, is predicted using the Valence Shell Electron Pair Repulsion (VSEPR) theory, which posits that electron pairs around a central atom arrange themselves to minimize repulsion. In VSEPR, the number of electron domains (bonding pairs and lone pairs) determines the shape: for example, two domains yield a linear geometry as in CO₂ (O=C=O, bond angle 180°), while four domains result in tetrahedral geometry as in CH₄ (bond angles ≈109.5°). This model assumes idealized electron pair repulsions, with lone pairs exerting greater influence than bonding pairs, leading to distortions in molecules like H₂O (bent, ≈104.5°). VSEPR effectively rationalizes geometries for main-group compounds without requiring detailed quantum calculations. Complementing VSEPR, hybridization theory explains orbital overlap in these geometries: sp hybridization for linear molecules (e.g., two sp orbitals in CO₂), sp² for trigonal planar (e.g., BF₃), and sp³ for tetrahedral (e.g., CH₄), where atomic s and p orbitals mix to form equivalent hybrid orbitals with optimal directional bonding. Hybridization, introduced by Linus Pauling, aligns bond directions with observed geometries, enhancing conceptual understanding of covalent bonding. Physical properties of bonds, such as lengths, angles, and dipole moments, stem from atomic interactions and electronegativity differences. Bond lengths vary by type and atoms involved; for instance, a typical C-C single covalent bond measures 1.54 , shortening to 1.34 for double bonds due to increased . Bond angles, as predicted by VSEPR and hybridization, are influenced by and , which measures an atom's ability to attract electrons in a bond— has the highest value (4.0 on the Pauling scale), leading to polar covalent bonds. Dipole moments arise from uneven charge distribution in polar bonds, quantified as \mu = q \cdot d where q is partial charge separation and d is ; for example, HCl has a dipole moment of 1.08 D due to chlorine's higher (3.0) compared to (2.1). These factors collectively determine molecular and reactivity. At the quantum level, the Born-Oppenheimer approximation underpins the treatment of molecular bonds by separating nuclear and electronic motions, given the vast mass difference (electrons ~1/1836 of proton mass). This approximation assumes nuclei are fixed during rapid electronic motion, allowing the potential energy surface for nuclei to be computed from electronic wavefunctions. Formally, the is decoupled into electronic and nuclear parts, enabling solutions to the for bond formation without simultaneous treatment of all . This foundational assumption facilitates the study of potential energy curves and geometries in .

Electronic Configuration

In molecular physics, the electronic configuration describes the distribution of electrons among the available molecular orbitals, which are delocalized wavefunctions encompassing the entire and determined by solving the for multi-electron systems. This arrangement dictates the 's chemical bonding, reactivity, and spectroscopic properties, with electrons occupying the lowest energy orbitals in accordance with the and . Unlike configurations, molecular ones account for the overlap and interaction of orbitals from constituent atoms, leading to a collective electronic structure that stabilizes the . Molecular orbitals (MOs) are constructed using the (LCAO) approximation, a foundational in where each MO is expressed as a weighted sum of basis atomic orbitals centered on the nuclei. The seminal of LCAO within the Hartree-Fock was developed by Roothaan and Hall, enabling the variational optimization of MO coefficients to minimize the total electronic energy. In this approach, bonding MOs result from constructive of atomic orbitals, lowering the energy below that of separated atoms and promoting between nuclei; antibonding MOs arise from destructive , raising the energy and creating nodal planes between atoms; non-bonding MOs, often localized on a single atom, remain largely unchanged in energy and do not contribute significantly to bonding. For instance, in the hydrogen molecule H₂, the ground-state bonding MO is \sigma_g = \frac{1}{\sqrt{2}} (1s_A + 1s_B), while the antibonding counterpart is \sigma_u^* = \frac{1}{\sqrt{2}} (1s_A - 1s_B). The electronic configuration is denoted by specifying the occupancy of each MO type, often using symmetry labels like \sigma, \pi, and \delta for diatomic molecules, with superscripts indicating the number of electrons. For the nitrogen molecule (N₂) in its ground state X^1\Sigma_g^+, the valence configuration is (\sigma_{2s})^2 (\sigma_{2s}^*)^2 (\pi_{2p})^4 (\sigma_{2p})^2, shorthand as \sigma^2 \pi^4 for the key bonding orbitals, reflecting a triple bond with two \pi and one \sigma components. Hund's rules govern the spin multiplicity in configurations with degenerate orbitals: electrons occupy orbitals singly with parallel spins to maximize total spin S (highest multiplicity $2S+1), then pair with opposite spins; for N₂, the closed-shell nature yields a singlet (S=0), but in open-shell cases like O₂, the \pi^* degeneracy follows these rules for a triplet ground state. This multiplicity influences term symbols and magnetic properties. Molecular symmetry, described by point groups, plays a crucial role in classifying orbitals and configurations, as MOs must transform according to the irreducible representations (irreps) of the group's character table. For (H₂O), which belongs to the C_{2v} with operations E, C_2 (rotation about the bisector), \sigma_v (xz plane), and \sigma_v' (yz plane), the valence atomic orbitals form symmetry-adapted linear combinations (SALCs): the oxygen 2s and 2p_z transform as A_1, the 2p_x as B_1, and 2p_y as B_2, while 1s SALCs yield one A_1 (symmetric) and one B_2 (antisymmetric). Resulting are thus labeled by these irreps (e.g., three A_1, one B_2 occupied in the ), ensuring that only orbitals of matching overlap effectively to form bonding interactions. This classification simplifies predictions of electronic transitions and selection rules. Excited electronic states arise from the promotion of an from a filled () to an empty (LUMO) orbital, altering the configuration and often leading to or triplet multiplicities depending on conservation. In a excited state, the promoted retains parallel to the vacancy (total S=0), while a triplet involves spin flip (total S=1), typically lower in energy due to reduced electron repulsion as per Hund's multiplicity rule. For example, in organic molecules, photoexcitation to S_1 () can undergo to T_1 (triplet) if the singlet-triplet energy gap \Delta E_{ST} is small (<0.37 eV), facilitated by spin-orbit coupling; this process is pivotal in applications like organic light-emitting diodes, where reverse intersystem crossing from triplet to enhances efficiency. The configuration change, such as from (\pi)^2 to (\pi)(\pi^*) in ethylene, shifts the molecule toward dissociative or reactive pathways.

Molecular Energy and Dynamics

Rotational and Vibrational Levels

The rotational energy levels of diatomic molecules are modeled using the rigid rotor approximation, which treats the molecule as two point masses separated by a fixed bond length, neglecting vibrational motion. The quantized energy levels are given by E_J = B J(J+1), where J is the rotational quantum number (J = 0, 1, 2, \dots), and B is the rotational constant defined as B = \frac{\hbar^2}{2I}, with I = \mu r_e^2 being the moment of inertia, \mu the reduced mass, and r_e the equilibrium bond length. This model arises from solving the for angular momentum in a two-body system, providing evenly spaced energy differences that scale with B, typically on the order of 1–10 cm⁻¹ for common diatomics like . Transitions between these levels obey the selection rule \Delta J = \pm 1, arising from the dipole interaction Hamiltonian in the electric dipole approximation. Vibrational energy levels in diatomic molecules are initially approximated by the quantum harmonic oscillator model, assuming a parabolic potential V(r) = \frac{1}{2} k (r - r_e)^2 near the equilibrium bond length. The resulting energy levels are E_v = \hbar \omega \left(v + \frac{1}{2}\right), where v = 0, 1, 2, \dots is the vibrational quantum number, and the angular frequency \omega = \sqrt{k / \mu} depends on the force constant k and reduced mass \mu. This approximation yields equally spaced levels, with typical spacings of 1000–4000 cm⁻¹ for molecular bonds, but it fails at higher v where the potential flattens. To incorporate anharmonicity, which accounts for bond weakening and eventual dissociation, the Morse potential is employed: V(r) = D_e \left[1 - \exp\left(-a (r - r_e)\right)\right]^2, where D_e is the well depth and a a parameter related to the curvature; the exact solutions yield vibrational levels approximated as E_v \approx \hbar \omega (v + 1/2) - \hbar \omega x_e (v + 1/2)^2, with x_e the anharmonicity constant, enabling finite v maxima near v \approx \sqrt{2 D_e / \hbar \omega}. In diatomic molecules, rotational and vibrational motions couple through the rovibrational Hamiltonian, leading to combined energy levels where the effective rotational constant varies with vibrational state: B_v = B_e - \alpha_e (v + 1/2), with \alpha_e quantifying the bond extension during vibration. Centrifugal distortion further refines the model, as rotation stretches the bond, increasing I and reducing energy; this introduces a correction term -D_e J^2 (J+1)^2 to E_J, where D_e \approx 4 B_e^3 / \omega^2, typically small but observable in high-J levels. These effects are derived from perturbation theory applied to the full vibration-rotation wavefunctions, ensuring the rigid rotor and harmonic limits as baselines. For polyatomic molecules, vibrational degrees of freedom extend to $3N - 6 normal modes for nonlinear systems or $3N - 5 for linear ones (N atoms), where each mode represents a collective oscillation treated as an independent harmonic oscillator with its own frequency \omega_i = \sqrt{k_i / \mu_i}. These modes are obtained by diagonalizing the Hessian matrix of the potential energy surface, decoupling the coordinates into orthogonal vibrations. In symmetric linear molecules like CO₂, degeneracy occurs; for example, the bending mode is doubly degenerate, corresponding to out-of-phase motions in perpendicular planes with identical \omega, while the symmetric and antisymmetric stretches are non-degenerate. Anharmonicity and mode coupling can be incorporated via multidimensional Morse-like potentials, but the harmonic normal mode basis remains foundational.

Electronic States and Transitions

In molecular physics, electronic states refer to the quantized configurations of electrons in a molecule, characterized by distinct energy levels arising from the arrangement of electrons in molecular orbitals. These states are described by potential energy surfaces (PES), which map the electronic energy as a function of nuclear coordinates under the , separating nuclear and electronic motions. The ground state PES typically features a deep minimum corresponding to the equilibrium geometry of stable molecules, while excited state PES may exhibit shallower wells, barriers, or repulsive walls that dictate photochemical reactivity. The Jablonski diagram provides a schematic representation of these electronic states, illustrating energy levels for singlet and triplet configurations along with allowed and forbidden transitions. In this framework, the ground state is the lowest singlet (S₀), with excited singlets (S₁, S₂, etc.) accessible via absorption, while triplets (T₁, T₂) arise from spin flips and are lower in energy than corresponding singlets due to exchange stabilization. Vertical transitions dominate, reflecting the rapid timescale of electronic rearrangements compared to nuclear motions. Transition probabilities between electronic states are quantified by Einstein coefficients, where B_{ij} governs absorption and stimulated emission rates, and A_{ij} describes spontaneous emission from upper (i) to lower (j) states, related through detailed balance as A_{ij}/B_{ij} = (8\pi h \nu^3)/c^3 for degenerate levels. The further modulates these probabilities, positing that electronic transitions occur vertically on the PES due to the Franck-Condon overlap integral between vibrational wavefunctions of initial and final states, favoring transitions where nuclear geometries minimally change. This results in progressions of vibrational structure within electronic bands, providing insight into PES curvatures. Spin-orbit coupling introduces relativistic interactions between electron spin and orbital angular momentum, lifting degeneracies to produce fine structure splittings, particularly pronounced in molecules with heavy atoms where the coupling constant scales with atomic number. This coupling enables intersystem crossing, a nonradiative process transferring population from singlet to triplet states via vibronic perturbations, with rates enhanced near conical intersections on intersecting PES. In heavy-element-containing molecules, such as those with iodine, intersystem crossing dominates photophysical decay pathways. Dissociation in excited electronic states often proceeds via predissociation, where a bound level couples to a dissociative continuum on the same or nearby PES, leading to broadened spectral lines and lifetimes shortened by tunneling or curve-crossing mechanisms. Photochemical pathways involve nonadiabatic transitions at avoided crossings or conical intersections between ground and excited PES, steering molecules toward specific dissociation channels, as seen in polyatomic systems like HNCO where multiple routes compete based on initial excitation energy.

Spectroscopy

Rotational and Vibrational Spectra

Rotational spectra of gas-phase molecules arise from transitions between quantized rotational energy levels and are typically observed in the microwave region, with frequencies ranging from a few GHz to hundreds of GHz. These transitions require the molecule to possess a permanent electric dipole moment, as homonuclear diatomic molecules like lack such a moment and are thus microwave inactive. For linear molecules, such as or , the pure rotational spectrum consists of a series of evenly spaced absorption lines corresponding to ΔJ = +1 transitions, where J is the rotational quantum number, with line spacings of approximately 2B (B being the rotational constant in cm⁻¹). For example, in , B ≈ 1.93 cm⁻¹, yielding spacings around 3.84 cm⁻¹. In polyatomic molecules classified as asymmetric tops, where the three principal moments of inertia differ significantly (I_A ≠ I_B ≠ I_C), the rotational spectra are more intricate, featuring multiple overlapping series of lines rather than simple spacing. This complexity stems from the lack of symmetry, as seen in water (H₂O), where transitions involve Wang combinations of quantum numbers and require fitting to determine the moments of inertia. These spectra provide precise structural information but demand high-resolution microwave techniques for assignment. Vibrational spectra manifest as absorption in the infrared (IR) region when molecular vibrations alter the dipole moment, with fundamental bands representing the strongest transitions from the ground vibrational state (v=0) to the first excited state (v=1). Overtone bands, such as v=0 to v=2, appear at roughly twice the fundamental frequency but are weaker (typically 1-10% intensity) due to anharmonicity in the potential energy surface, which causes deviations from exact harmonic ratios. Isotope effects significantly influence these spectra; substitution with heavier isotopes increases the reduced mass, lowering vibrational frequencies and shifting band origins, as observed in ¹³C¹⁶O versus ¹²C¹⁶O, where the fundamental frequency decreases by about 4.6%. For instance, the CO fundamental at ~2143 cm⁻¹ shifts to ~2093 cm⁻¹ for the isotopologue. Rovibrational spectra in gas-phase molecules combine vibrational and rotational transitions, observed in the mid-IR, and reveal fine structure due to simultaneous changes in both quantum numbers. For linear molecules, these spectra display distinct P (ΔJ = -1, lower frequency), Q (ΔJ = 0, near band center if perpendicular transition), and R (ΔJ = +1, higher frequency) branches, forming a banded pattern around the vibrational band origin; in CO, the fundamental band shows clear P and R branches spaced by ~2B. In polyatomic molecules, Coriolis interactions between vibrational modes introduce perturbations, splitting levels and altering branch intensities, particularly in degenerate modes like the bending vibration in CO₂ (ν₂ at ~667 cm⁻¹), where vibrational angular momentum (l-type doubling) affects the spectral resolution. These interactions are quantified by coupling constants ξ, coupling rotation to vibration. The observation of rotational and vibrational spectra is governed by the Boltzmann distribution, which determines the population of energy levels at a given temperature T. Rotational levels are more densely populated at higher temperatures, with the maximum population occurring at J_max ≈ (kT / 2B)^{1/2} - 1/2, leading to broader spectra and shifted line intensities; for example, at room temperature (~300 K), CO populates up to J ≈ 5-6, while cooling to 13 K limits it to lower J, sharpening the spectrum. Vibrational levels follow a similar distribution but with much sparser population due to higher energy spacings (θ_vib >> θ_rot), so fundamentals dominate at ambient conditions, with negligible unless T exceeds ~1000 K. This temperature dependence allows rotational temperatures to be derived from line intensity ratios, providing diagnostics for molecular environments.

Electronic and Raman Spectra

Electronic spectra of molecules arise from transitions between electronic energy levels, typically observed in the ultraviolet-visible (UV-Vis) region through absorption or emission of photons. These transitions involve promotion of an electron from the ground state to an excited electronic state, often accompanied by changes in vibrational and rotational quantum numbers, leading to structured bands rather than simple lines. In polyatomic molecules, such spectra provide insights into electronic configurations and molecular geometries, as detailed in foundational analyses of electronic structure. For example, absorption spectra in the UV-Vis range (200–800 nm) reveal π → π* or n → π* transitions in conjugated systems like , where the intense S1 ← S0 band appears around 260 nm. Vibronic progressions in spectra manifest as series of bands spaced by vibrational frequencies, arising from Franck-Condon overlaps between vibrational wavefunctions in the initial and final states. These progressions are prominent in rigid molecules, such as , where the n → π* transition shows a progression in the C=O stretching mode (ν4 ≈ 1280 cm⁻¹), reflecting the change in upon . Emission spectra, including from states and from triplet states, similarly exhibit vibronic structure; lifetimes for dyes like fluorescein are typically 4–5 ns, while lifetimes in aromatic hydrocarbons like range from milliseconds to seconds due to spin-forbidden . Selection rules govern allowed electronic transitions based on molecular symmetry and the nature of the transition dipole moment. For electric dipole transitions, the integral ∫ ψ_f* μ ψ_i dτ must be nonzero, requiring the direct product of the representations of the initial state (Γ_i), final state (Γ_f), and dipole operator (Γ_μ) to contain the totally symmetric representation (A1 in C_{2v}, for instance). Symmetry-forbidden transitions, such as those violating the (Δl = ±1 for orbitals), appear weakly unless vibronic coupling relaxes the rule, as in Herzberg-Teller interactions. Polarization effects in oriented samples or single crystals reveal transition moment directions; for instance, in anthracene crystals, the polarization of the 380 nm band aligns with the long molecular axis, aiding assignment of excited states. Raman spectroscopy probes molecular vibrations and rotations through inelastic light scattering, distinct from absorption processes. Rayleigh scattering corresponds to elastic scattering at the incident frequency (ν_0), while Raman shifts occur when the scattered photon energy is ν_0 ± ν_m, where ν_m is a molecular vibrational or rotational frequency; Stokes shifts (ν_0 - ν_m) dominate at room temperature due to population of lower vibrational levels, whereas anti-Stokes (ν_0 + ν_m) are weaker. Vibrational Raman spectra highlight polarizability changes, with strong bands for symmetric stretches in CO2 (ν_1 ≈ 1330 cm⁻¹). Rotational Raman in symmetric top molecules like CH3F exhibits O, Q, and S branches, with ΔJ = 0, ±2 selection rules, spacing determined by the moments of inertia (e.g., B ≈ 0.85 cm⁻¹ for CH3F). Nonlinear effects in include resonance enhancement, where the incident frequency approaches an electronic transition, amplifying cross-sections by factors up to 10^6 via coupling to virtual excited states. This resonance () selectively enhances vibrations coupled to the electronic state, as in proteins where the 1370 cm⁻¹ mode is intensified near the Soret band (≈400 nm), providing site-specific structural information. Electronic spectra bands often display fine rotational structure, with P, , and branches reflecting ΔJ = ±1, 0, +1 rules for linear molecules.

Intermolecular Forces

Van der Waals Interactions

Van der Waals interactions encompass the weak, non-covalent attractive forces between neutral molecules that arise primarily from and mechanisms, playing a key role in phenomena such as gas-phase clustering where these forces stabilize transient molecular aggregates. These interactions are orientation-independent and typically much weaker than covalent bonds, with energies ranging from 0.1 to 10 kJ/mol, enabling the formation of weakly bound complexes in the gas phase without dominating over thermal energies at . Dispersion forces, also known as London forces, originate from correlated fluctuations in the electron distributions of non-polar molecules, leading to instantaneous dipole moments that induce attractive interactions between them. First theoretically described by Fritz London in 1930 through a semi-classical approach involving quantum mechanical correlations in the zero-point energy of electrons, these forces are a universal component of intermolecular attractions even in systems lacking permanent dipoles. The potential energy for dispersion interactions follows an inverse sixth-power dependence on the intermolecular distance, expressed as V(r) = -\frac{C_6}{r^6}, where C_6 is a coefficient dependent on the molecular polarizabilities and ionization energies. Induction forces, or Debye interactions, occur when a permanent in one polarizes the cloud of a neighboring , creating an induced that results in attraction. This mechanism was elucidated by in the early 20th century as part of his work on molecular , highlighting how the of the induces a temporary in the polarizable species. The associated takes the form V(r) = -\frac{C_3}{r^6}, with C_3 proportional to the square of the permanent and the of the induced species, again showing the characteristic r^{-6} decay. The physical origins of both and forces are rooted in quantum fluctuations of the , contributing to the overall attractive component of intermolecular potentials while short-range repulsion arises from Pauli exclusion. A widely used model combining these effects with repulsion is the , proposed by in 1931, given by V(r) = 4\epsilon \left[ \left( \frac{\sigma}{r} \right)^{12} - \left( \frac{\sigma}{r} \right)^6 \right], where the r^{-12} term approximates the steep repulsive wall and the r^{-6} term captures the van der Waals attraction, with parameters \epsilon and \sigma fitted to experimental data for specific molecular pairs. This potential is essential for understanding the balance of forces in weakly bound systems like gas-phase clusters.

Dipole and Hydrogen Bonding

Dipole-dipole interactions, also known as Keesom interactions, arise from the electrostatic attraction between permanent electric in polar molecules. These forces are orientation-dependent, with the interaction energy for two fixed given by U = \frac{1}{4\pi\epsilon_0} \frac{\mu_1 \mu_2}{r^3} \left[ \hat{\mu}_1 \cdot \hat{\mu}_2 - 3 (\hat{\mu}_1 \cdot \hat{r}) (\hat{\mu}_2 \cdot \hat{r}) \right], where \mu_1 and \mu_2 are the dipole moments, r is the intermolecular distance, and the terms involving unit vectors \hat{\mu}_1, \hat{\mu}_2, and \hat{r} account for the angular dependence. In thermal environments, molecules rotate freely, leading to an averaged potential that scales as $1/r^6 and is weaker than the fixed-orientation case. Hydrogen bonding represents a particularly strong and directional form of dipole-dipole interaction, occurring when a covalently bonded to an electronegative atom X (such as , , or F) interacts with another electronegative atom Y, denoted as X–H···Y. These bonds typically exhibit energies ranging from 10 to 40 kJ/mol, significantly stronger than typical dipole-dipole forces due to the partial positive charge on and the electrons on Y. In extended structures like chains, hydrogen bonds display , where the formation of additional bonds enhances the strength of existing ones through mutual , often nearly doubling the terminal bond energy in short chains. Higher-order electrostatic effects, such as , play a key role in nonpolar linear molecules like CO₂, which possesses no net but a substantial moment of approximately 4.1 × 10⁻²⁶ esu cm², arising from the charge distribution with negative oxygen atoms relative to the central carbon. This moment induces anisotropic attractions between molecules, influencing their alignment and contributing to intermolecular forces beyond . For systems involving charged , ion- interactions dominate, where the is U = -\frac{q \mu \cos\theta}{4\pi\epsilon_0 r^2}, with q as the ion charge, \mu the , and \theta the angle between the and the ion-molecule line; these forces are crucial in polar solvents solvating s. Spectroscopically, these electrostatic interactions manifest in (IR) spectra through shifts in vibrational frequencies. Hydrogen bonding characteristically causes red-shifts in X–H stretching modes, as the weakened bond lengthens and its force constant decreases, with the shift magnitude increasing with bond strength. For instance, in dimers or chains, O–H stretches shift by tens to hundreds of cm⁻¹ toward lower wavenumbers, providing a direct probe of bond formation and .

Theoretical Approaches

Quantum Mechanical Models

The quantum mechanical description of molecules begins with the time-independent , \hat{H} \psi = E \psi, where \hat{H} is the , \psi is the , and E is the eigenvalue. The full for a consisting of nuclei and electrons is given by \hat{H} = \hat{T}_n + \hat{T}_e + \hat{V}_{ee} + \hat{V}_{nn} + \hat{V}_{en}, where \hat{T}_n and \hat{T}_e are the operators for the nuclei and electrons, respectively, \hat{V}_{ee} is the electron-electron repulsion, \hat{V}_{nn} is the nucleus-nucleus repulsion, and \hat{V}_{en} is the electron-nucleus attraction. This many-body captures the non-separable interactions in molecular systems, making exact solutions intractable for systems beyond the . To make the problem tractable, the exploits the large mass difference between electrons and nuclei (electrons are ~1836 times lighter than protons), assuming nuclear positions change negligibly during electronic motion. This separates the into electronic and nuclear parts: the electronic is solved first for fixed nuclear coordinates \mathbf{R}, yielding potential energy surfaces E(\mathbf{R}) that serve as effective potentials for the nuclear wave function. The approximation is valid when the energy spacing of electronic states exceeds vibrational frequencies, with corrections (non-adiabatic effects) small for most ground-state properties but relevant in conical intersections or photochemical processes. With this separation, the focus shifts to the electronic Hamiltonian. The Hartree-Fock method employs a mean-field approximation, where each electron moves in an average potential created by the other electrons and nuclei, leading to a set of single-particle equations solved self-consistently via a wave function. This approach neglects electron correlation but provides a foundational description of molecular electronic structure. For improved accuracy, configuration interaction methods incorporate electron correlation by expanding the wave function as a of multiple s, often built from Hartree-Fock orbitals as a basis, to account for instantaneous electron-electron interactions beyond the mean field. The underpins many of these approximations, stating that for any normalized \phi, the expectation value of the \langle \phi | \hat{H} | \phi \rangle provides an upper bound to the true ground-state , with equality only for the exact . are constructed to minimize this , as in Hartree-Fock or configuration interaction. complements this by treating small interactions, such as electron correlation or external fields, as perturbations to a solvable zeroth-order , yielding corrections to energies and wave functions via series expansions. For and time evolution, the time-dependent i \hbar \frac{\partial \psi}{\partial t} = \hat{H} \psi governs the of the wave function under a time-independent or time-varying , enabling the study of transient processes like excitations or reactive .

Computational Methods

Computational methods in molecular physics provide numerical frameworks for approximating solutions to the molecular , enabling the prediction of structures, energies, and properties for systems where analytic solutions are intractable. These approaches span a hierarchy of approximations, from highly accurate but computationally demanding techniques to more efficient semi-empirical and classical methods, with trade-offs primarily determined by the treatment of correlation and the size of the system. The choice of method depends on the desired balance between precision—often benchmarked against experimental data or higher-level calculations—and scalability to larger molecules or ensembles. Ab initio methods solve the electronic without empirical parameters, relying on systematic approximations to the many-body . The Hartree-Fock () method serves as the foundational approach, approximating the as a single of molecular orbitals and minimizing the energy via the . The HF equations are solved self-consistently: \left[ -\frac{1}{2}\nabla^2 - \sum_I \frac{Z_I}{r_I} + \sum_{j=1}^N \left( J_j(\mathbf{r}) - K_j(\mathbf{r}) \right) \right] \phi_i(\mathbf{r}) = \epsilon_i \phi_i(\mathbf{r}) where J_j and K_j are the and operators, respectively. HF captures exchange effects but neglects electron correlation, typically underestimating bond dissociation energies by 10–50 kcal/mol (10–30%) for small molecules, depending on the system, at a computational cost scaling as O(N^4) for N basis functions. To include correlation, post-HF methods such as second-order Møller-Plesset (MP2) add pairwise electron interactions perturbatively, improving accuracy for and reaction energies, though at O(N^5) cost. More robust coupled-cluster methods like CCSD(T)—coupled cluster with singles, doubles, and perturbative triples—achieve near-quantitative accuracy (errors <1 kcal/mol for many thermochemical properties) and is considered the "gold standard" for calculations, but scales as O(N^7), limiting it to systems with ~50 atoms. These methods require a finite basis set to expand the molecular orbitals, introducing basis set superposition error (BSSE) that diminishes with larger sets. Minimal basis sets like STO-3G approximate Slater-type orbitals (STOs) using three Gaussian functions each for efficiency in geometry optimizations, suitable for qualitative studies but yielding energies ~10–20 kcal/mol too high. Correlated consistent basis sets, such as Dunning's cc-pVQZ (with quadruple-zeta valence functions and ), provide near-complete basis set limits for post-HF calculations, reducing BSSE to <1 kcal/mol for bond dissociation, though increasing cost cubically with . Complete basis set extrapolations are often employed to estimate the infinite-basis limit. Density functional theory (DFT) offers a computationally efficient by reformulating the problem in terms of the n(\mathbf{r}), grounded in the Hohenberg-Kohn theorems, which establish that the ground-state energy is a unique functional of n(\mathbf{r}). The Kohn-Sham framework maps the interacting system to a non-interacting reference of orbitals \psi_i, with the including the exact exchange-correlation functional E_{xc}: \left[ -\frac{1}{2}\nabla^2 + v_{\text{eff}}(\mathbf{r}) \right] \psi_i(\mathbf{r}) = \epsilon_i \psi_i(\mathbf{r}), \quad v_{\text{eff}}(\mathbf{r}) = v_{\text{ext}}(\mathbf{r}) + \int \frac{n(\mathbf{r}')}{|\mathbf{r}-\mathbf{r}'|} d\mathbf{r}' + v_{xc}(\mathbf{r}), solved iteratively like HF but at O(N^3) cost. The local density approximation (LDA) approximates E_{xc} using the uniform electron gas, providing reasonable geometries but overbinding energies by 5–10 kcal/mol. Hybrid functionals like B3LYP incorporate ~20% Hartree-Fock exchange with Becke's gradient-corrected exchange and Lee-Yang-Parr correlation, achieving chemical accuracy (~3 kcal/mol) for diverse properties like barrier heights and noncovalent interactions, widely adopted for medium-sized molecules up to hundreds of atoms. DFT's efficiency stems from avoiding explicit wave function correlation, though it struggles with strong correlation or dispersion without corrections. For larger systems where and DFT become prohibitive, semi-empirical methods approximate the framework by parameterizing two-electron integrals and core interactions from experimental data, reducing cost to O(N^2) or linear. The Austin Model 1 (AM1) extends modified neglect of diatomic overlap (MNDO) with Gaussian core corrections to improve bonding and lone-pair interactions, achieving errors of ~5–10 kcal/ in heats of formation for molecules. The parameterized model 3 (PM3) refines AM1 parameterization for a broader periodic table, enhancing accuracy for transition metals and while maintaining speed for simulations of thousands of atoms, though it underestimates dispersion. These methods enable rapid screening in or materials modeling, trading some rigor for scalability. Molecular dynamics (MD) simulations extend static quantum calculations to dynamic processes, evolving atomic trajectories under forces derived from potential energy surfaces. Classical MD employs empirical force fields—such as or CHARMM, parameterized from quantum data and experiments—to compute bonded (bonds, angles) and non-bonded (van der Waals, electrostatic) interactions, enabling microsecond-scale simulations of biomolecules at low cost (~nanoseconds per day on GPUs). Quantum effects like zero-point motion are neglected, limiting accuracy for light atoms or low temperatures. Path integral MD (PIMD) incorporates quantum statistics by representing each particle as a ring polymer of P beads (typically P=32, scaling as O(N P^2)), exactly treating nuclear quantum effects like tunneling and delocalization in vibrational spectra or proton transfer, as validated in simulations where classical MD overestimates diffusion by 20–30%. Hybrid quantum-classical schemes, like , couple high-level quantum regions to classical surroundings for reactions.

Experimental Techniques

Spectroscopic Measurements

Spectroscopic measurements in molecular physics rely on precise to observe , , or of by molecules, revealing details about their energy levels and interactions. These techniques primarily involve photon-molecule interactions in controlled settings, such as gas cells or chambers, to capture spectra with minimal environmental . Fourier transform spectrometers and laser-based systems are central to achieving the high needed for resolving fine rotational-vibrational structures in molecular spectra. Microwave and () spectroscopy employs () spectrometers to measure rotational and vibrational transitions in gas-phase molecules. In , chirped-pulse techniques generate broadband excitation, allowing rapid acquisition of spectra across several GHz with resolutions better than 10 kHz, equivalent to approximately 3 × 10^{-7} cm^{-1}, enabling the detection of subtle hyperfine splittings. For spectroscopy, instruments use to achieve resolutions as high as 0.005 cm^{-1} or better, limited by the maximum difference in the interferometer, which is crucial for resolving closely spaced vibrational bands in polyatomic molecules. These resolutions allow observation of spectral patterns arising from quantized energy levels without significant broadening from Doppler effects in low-pressure samples. Laser-based methods enhance precision and sensitivity for molecular spectroscopy. Tunable diode lasers, such as distributed (DFB) or quantum cascade lasers (QCLs), provide narrow linewidths (typically <1 MHz) and rapid tuning over 1-10 cm^{-1} ranges, enabling high-resolution of trace species with detection limits down to (ppb). For even greater sensitivity, (CRDS) confines laser light in high-finesse optical cavities, measuring times to detect coefficients as low as 10^{-10} cm^{-1}, ideal for detection in molecular physics, such as NO_2 at concentrations below 1 ppb. These techniques minimize path-length uncertainties through multi-pass cells or cavity enhancement, achieving precisions suitable for studying weak molecular transitions. Time-resolved spectroscopic measurements capture ultrafast using pump-probe configurations. In these setups, a "" pulse excites the to a higher energy state, followed by a delayed "" pulse that monitors transient or , resolving femtosecond-scale processes like vibrational relaxation or electronic state evolution. effects, such as quantum beats between vibrational modes, manifest as oscillatory signals in the differential , providing insights into wavepacket in systems like polyatomic . Resolutions down to 10 are routine with titanium-sapphire lasers, allowing observation of intramolecular couplings without significant . Gas-phase measurements offer unperturbed spectra, reflecting intrinsic molecular properties, but require low pressures to reduce collisional broadening. In contrast, matrix isolation traps molecules in inert cryogenic matrices (e.g., or at 10-20 K) to stabilize reactive species and suppress rotational motion, yielding sharper vibrational lines; however, host-guest interactions induce matrix shifts of 1-13 cm^{-1} and site-specific splittings, necessitating corrections for accurate comparison to gas-phase data. This approach avoids perturbations from intermolecular collisions prevalent in denser gas samples, though it introduces subtle environmental effects that must be quantified.

Scattering and Diffraction

Scattering and diffraction techniques in molecular physics employ beams of electrons, neutrons, X-rays, or other particles to interrogate molecular structures and interactions at the scale, providing spatial information complementary to spectroscopic methods. These approaches exploit interference patterns arising from wave-like particle off molecular potentials, yielding insights into bond lengths, positions, and pathways. In gas-phase or condensed environments, such experiments reveal details obscured by motion or intermolecular forces, with resolutions down to angstroms for structural parameters. Gas-phase electron diffraction (GED) is a for elucidating the structures of small, volatile molecules, where a beam of high-energy electrons scatters from a gaseous sample, producing patterns that encode internuclear distances. This technique achieves determinations with typical accuracies of 0.01 or better for rigid molecules, enabling precise characterization of geometries free from crystal-packing distortions. However, due to rapid rotational averaging in the isotropic gas phase, the observed patterns reflect thermal ensembles rather than instantaneous snapshots, necessitating corrections for vibrational and conformational averaging during . Neutron scattering excels in probing light atoms, particularly , owing to the element's large coherent length, which provides strong signals even in hydrogen-rich molecular systems where X-rays falter due to weak contrasts. In molecular physics, maps static structures in powders or liquids, while inelastic captures dynamic processes such as molecular by measuring transfers corresponding to or librational modes. This sensitivity to hydrogen positions has been pivotal in studies of hydrogen-bonded networks and catalytic surfaces, offering vibrational spectra with resolutions tied to the 's de Broglie . X-ray diffraction remains indispensable for resolving structures of larger molecules, such as compounds or biomolecules, crystallized into ordered lattices that amplify signals through constructive . For these systems, coordinates are refined from data, with typical resolutions of 0.5–1 for small molecules, though higher precision is attainable under cryogenic conditions. Thermal displacements, which broaden peaks and reduce intensities, are quantified via Debye-Waller factors—exponential terms that model the mean-square vibrational amplitudes and ensure accurate refinement of bond lengths and angles in the presence of disorder. Crossed-beam scattering experiments collide controlled molecular beams to dissect gas-phase , isolating elementary steps and measuring product distributions through cross-sections that quantify probabilities as functions of and velocity. Pioneered in the , these studies reveal stereodynamics, such as direct rebound versus stripping mechanisms, with collision energies tunable from thermal to hyperthermal regimes. Such data, often visualized in polar plots, directly inform surfaces and geometries for bimolecular reactions.

Applications and Research

Industrial and Technological Uses

Molecular physics plays a pivotal role in through the application of spectroscopic techniques for . (IR) and , which probe vibrational modes of molecular bonds, enable the identification of composition, crystallinity, and defects in industrial settings. These non-destructive methods facilitate real-time during and , ensuring material consistency for applications like and composites. For instance, distinguishes between polymer phases and detects additives at low concentrations, supporting efficient manufacturing workflows. In optoelectronic devices, molecular physics informs the rational design of for light-emitting diodes (LEDs) and . By engineering molecular structures to optimize electronic coupling and energy transfer, researchers achieve high-efficiency organic LEDs with external quantum efficiencies over 30% and photovoltaic cells with power conversion efficiencies surpassing 18%. These designs leverage intermolecular forces to control charge separation and recombination, enabling flexible, low-cost devices for displays and harvesting. Theoretical predictions from quantum mechanical models enable such tailored molecular architectures. Pharmaceutical development benefits from vibrational spectroscopy in analyzing drug-protein binding interactions. Techniques like 2D-IR spectroscopy detect spectral shifts in amide vibrations upon ligand binding, providing insights into binding affinities and conformational changes without labels. This approach supports , as seen in studies classifying binding to with over 90% accuracy. Additionally, aids in chiral molecule separation by distinguishing enantiomers through microwave three-wave mixing, which measures chirality-dependent signals in gas-phase samples. This enables quantitative enantiomeric excess determination in mixtures, crucial for synthesizing pure enantiomers in production where one form may be therapeutic and the other inactive or toxic. Sensor technologies draw on molecular physics for via IR detectors that exploit molecular absorption fingerprints. Fourier-transform IR spectrometers identify trace pollutants, such as volatile organic compounds or , at parts-per-billion levels, enabling for air quality and emissions control in industrial sites. In precision timekeeping, of molecules reduces thermal motion, achieving temperatures below 1 to minimize decoherence in optical clocks. This enhances clock stability to 10^{-18} fractional uncertainty, supporting applications in GPS and . Energy conversion technologies rely on molecular physics to elucidate interactions in fuel cells and . In polymer electrolyte fuel cells, ionomer-catalyst interfaces govern proton conductivity and oxygen reduction kinetics, with molecular simulations revealing optimal binding energies for Pt nanoparticles that boost durability to over 10,000 hours. For , understanding van der Waals and electrostatic interactions at active sites improves selectivity in processes like ammonia synthesis, where tailored metal-organic frameworks enhance reaction rates by factors of 10.00015-4)

Current Frontiers

Recent advances in molecular physics have pushed the boundaries of ultracold molecule research, particularly through the manipulation of Feshbach resonances to form Bose-Einstein condensates (BECs) of molecules. Post-2010 experiments have demonstrated the creation of stable ultracold molecular BECs by associating fermionic atoms into bosonic dimers near Feshbach resonances, enabling the study of superfluidity and quantum degeneracy in composite particles. For instance, in 2013, researchers produced ultracold fermionic Feshbach molecules of ^{23}Na^{40}K with binding energies tunable via magnetic fields, achieving near-unitary interactions that bridge the BEC-BCS crossover regime. Further progress in 2023 involved preparing bosonic Feshbach molecules of ^{40}K atoms in optical lattices, where Feshbach tuning allowed for high-fidelity conversion (up to 100%) into the second Bloch band, revealing long lifetimes (~100 ms) and relaxation channels dominated by dimer-dimer collisions. These experiments highlight the potential for engineering exotic many-body phases, such as paired superfluids, inaccessible in atomic systems alone. Attosecond physics has revolutionized the real-time observation of electron dynamics within molecules, earning the 2023 Nobel Prize in Physics for the development of attosecond pulse generation techniques. Using high-harmonic generation (HHG) from laser-irradiated gases, isolated attosecond pulses (down to 43 as) enable probing of ultrafast electron motion on its natural timescale, capturing processes like photoemission delays in molecular orbitals. A key 2017 experiment measured attosecond delays (21 as difference) between 2s and 2p electron emissions in neon, providing direct insight into electron correlation effects that extend to molecular systems like water, where 2020 interferometry revealed 50–70 as delays in photoemission from liquid versus gas phases, underscoring solvent influences on electron dynamics. These methods have opened attosecond streaking and pump-probe spectroscopy for molecules, allowing visualization of charge migration and Auger decay in real time, with applications to photochemical reactions. Quantum simulation using molecules in optical s has emerged as a powerful tool to mimic complex condensed matter phenomena, leveraging the rich internal of polar molecules. Ultracold dipolar molecules, such as KRb, trapped in optical lattices simulate extended Hubbard models with long-range interactions, studies of quantum magnetism and fractionalization not feasible in atomic systems. In 2023, experiments with Feshbach molecules in higher orbital bands of square lattices demonstrated precise control over binding energies (e.g., 22.7 kHz) and interaction strengths, facilitating simulations of phases and topological insulators by tuning lattice depth and resonance positions. This approach offers superior accuracy for emulating electron-lattice couplings in solids, with molecule lifetimes extended to milliseconds, paving the way for probing quantum phase transitions in synthetic dimensions. In , single-molecule has advanced the understanding of dynamics since 2000, revealing heterogeneous pathways and transient intermediates through techniques like (). (smFRET) tracks conformational changes at the nanometer scale, distinguishing unfolded subpopulations and their interconversions in real time, as demonstrated in post-2000 studies of the cold shock protein (Csp), where relaxation times of ~250 μs were mapped during folding. For larger proteins, 2011 high-throughput smFRET experiments on unfolded the folding landscape, identifying parallel pathways and barriers with millisecond resolution, challenging ensemble-averaged models. Additionally, 2010 observations inside chaperonin cages (/GroES) showed substrate proteins folding in ~1 second, with smFRET quantifying compaction and avoidance of aggregation, informing mechanisms of assisted folding in cellular environments. These techniques continue to elucidate misfolding in diseases like by capturing rare events in α-synuclein dynamics.

References

  1. [1]
    Atomic Molecular and Optical Sciences - AMOS Experiment at LBNL
    Molecular physics is the study of the physical properties of molecules and of the chemical bonds between atoms that bind them into molecules. Its most ...
  2. [2]
  3. [3]
    Atomic, Molecular and Optical Physics
    Research in AMO physics has a long history--from building the foundations of quantum mechanics to continuing today at the cutting edge of science.
  4. [4]
    Chapter: 8 Applications of Atomic, Molecular, and Optical Physics
    AMO physics provides theoretical and experimental methods and essential data to neighboring areas of science such as chemistry, astrophysics, condensed-matter ...
  5. [5]
  6. [6]
    Learn about Molecular Physics - Taylor & Francis Online
    Aims and scope​​ Molecular Physics is a well-established international journal publishing original high quality papers in chemical physics and physical chemistry.
  7. [7]
    Guide to the James Franck Papers 1882-1966 - UChicago Library
    Otto Oldenberg, "Excitation of Molecular Vibration by Electron Impact." Typescript with corrections in the author's hand, 2 leaves. Box 24 Folder 18. Robert W ...
  8. [8]
    Quantum physics in America, 1920-1935 9780405125850 ...
    ... 1920's was that of molecular structure of gases as evidenced by their band spectra. This is not surprising since there were several American laboratories ...
  9. [9]
    Kuhn Losses Regained: Van Vleck from Spectra to Susceptibilities
    These 1926 papers only dealt with the special case in which the rigid rotator was used to model the gas molecules. While crossing the Atlantic in June 1926, Van ...
  10. [10]
    [PDF] MOLECULAR VIBRATIONS: FROM HARMONIC OSCILLATORS TO ...
    In particular it is useful to think of the molecule as a set of harmonic oscillators at low energies whereas at higher energies it is more appropriate to think ...
  11. [11]
    [PDF] Charles H. Townes - Nobel Lecture
    late 1940's was microwave spectroscopy, the study of interactions between microwaves and molecules. From this research, considerable information could be ...
  12. [12]
    The early days of precision laser spectroscopy - Physics Today
    Jan 1, 2007 · In the 1960s and 1970s, spectroscopists developed a host of nonlinear techniques to measure the interaction of light and matter with a ...
  13. [13]
    The Nobel Prize in Physics 1997 - NobelPrize.org
    The Nobel Prize in Physics 1997 was awarded jointly to Steven Chu, Claude Cohen-Tannoudji and William D. Phillips for development of methods to cool and trap ...
  14. [14]
    Roadmap of ultrafast x-ray atomic and molecular physics - IOPscience
    Jan 9, 2018 · Here we capture the perspectives of 17 leading groups and organize the contributions into four categories: ultrafast molecular dynamics, multidimensional x-ray ...<|separator|>
  15. [15]
    Ab initio effective core potentials for molecular calculations ...
    Jan 1, 1985 · A consistent set of ab initio effective core potentials (ECP) has been generated for the main group elements from Na to Bi using the procedure originally ...
  16. [16]
    Metallic Bonding - an overview | ScienceDirect Topics
    A metallic bond is a bond resulting from attractions between positive ions and surrounding mobile electrons. It is a form of chemical bonding that arises from ...
  17. [17]
    Diatomic Molecules According to the Wave Mechanics. II. Vibrational ...
    Diatomic Molecules According to the Wave Mechanics. II. Vibrational Levels. Philip M. Morse.
  18. [18]
    Inorganic stereochemistry - Quarterly Reviews, Chemical Society ...
    Inorganic stereochemistry. R. J. Gillespie and R. S. Nyholm, Q. Rev. Chem. Soc., 1957, 11, 339 DOI: 10.1039/QR9571100339. To request permission to reproduce ...Missing: PDF | Show results with:PDF
  19. [19]
    THE NATURE OF THE CHEMICAL BOND. II. THE ONE-ELECTRON ...
    Riddhimoy Pathak, Mridul Krishna Sharma, Kanishka Biswas. Pauling's Third Rule as a Guide for Designing Low Thermal Conducting Chalcogenides. Chemistry of ...
  20. [20]
    Article Longest C–C Single Bond among Neutral Hydrocarbons with ...
    Apr 12, 2018 · The Csp3–Csp3 bond length is generally 1.54 Å, as in the case of ethane 1. Although this standard bond length is observed in almost all ...
  21. [21]
    [PDF] On the Quantum Theory of Molecules
    In order to determine the eigenfunctions and thereby the transition probabilities only to the zeroth approximation, the energy calculation must be carried out ...
  22. [22]
    [PDF] arXiv:1912.12029v2 [physics.comp-ph] 11 Feb 2020
    Feb 11, 2020 · 1. Electronic structure calculations on molecular systems most often employ the linear combination of atomic orbitals. (LCAO) approach, where ...<|separator|>
  23. [23]
  24. [24]
    On the complexity of the absorption spectrum of molecular nitrogen
    Jun 30, 2008 · The electronic configuration of homonuclear N2 in its X1Σg+ ground state is: ( 1 s σ g ) 2 ( 1 s σ u ) 2 ( 2 s σ g ) 2 ( 2 s σ u ) 2 ( 2 p π u ) ...
  25. [25]
    Term rules for simple metal clusters | Scientific Reports - Nature
    Oct 26, 2015 · Hund's term rules are only valid for isolated atoms, but have no generalization for molecules or clusters of several atoms.
  26. [26]
    [PDF] MOLECULAR SYMMETRY, GROUP THEORY, & APPLICATIONS
    As another example, we will use group theory to construct the molecular orbitals of H2O (point group C2v) using a basis set consisting of all the valence ...
  27. [27]
    Promoting Singlet/triplet Exciton Transformation in Organic ... - Nature
    Jul 24, 2017 · Our discovery highlights systematically the critical importance of vertical transition configuration of excited states in promoting the singlet/triplet exciton ...Introduction · Molecular Selection And... · Excited State Similarity (s)
  28. [28]
    Molecular Spectra Vol I : Herzberg,Gerhard. - Internet Archive
    Nov 15, 2006 · Publication date: 1950/00/00. Topics: NATURAL SCIENCES, Physics, Mechanics of gases. Aeromechanics. Plasma physics.
  29. [29]
    Theoretical analysis of the centrifugal distortion contributions to the ...
    The centrifugal distortion contributions to the rotational energies of diatomic molecules are derived from the resolution of the vibration-rotation wave ...
  30. [30]
    [PDF] Structure and spectra of polyatomic molecules
    In CO2 the bending mode is doubly degenerate (see Figure 5.4). The excitation of a degenerate vibrational mode leads to a vibrational angular momentum l. In ...
  31. [31]
    Potential Energy Surface - an overview | ScienceDirect Topics
    The potential energy surface, or hypersurface, describes the energy of a molecular assembly and its value depends on the coordinates of all the atoms in the ...
  32. [32]
    Neural Network Potential Energy Surfaces for Small Molecules and ...
    We review progress in neural network (NN)-based methods for the construction of interatomic potentials from discrete samples (such as ab initio energies)
  33. [33]
    [PDF] Aleksander Jabłoński - Chemistry
    The original diagram of the energy levels was published in a short paper in Nature in 1933 (copy of the title and the original diagram below, reprinted by ...Missing: citation | Show results with:citation
  34. [34]
    Generalized Einstein relations between absorption and emission ...
    Einstein's relationships between absorption and emission conflict with the time-energy uncertainty principle by ascribing a finite radiative lifetime to the ...
  35. [35]
    Nuclear Motions Associated with Electron Transitions in Diatomic Molecules
    ### Summary of Franck-Condon Principle for Vertical Transitions
  36. [36]
    Spin–orbit coupling and intersystem crossing in molecules - 2012
    Jul 22, 2011 · This article focusses on approximate SOC operators for practical use in molecular applications and reviews state-of-the-art theoretical methods for evaluating ...INTRODUCTION · SPIN–ORBIT COUPLING · INTERSYSTEM CROSSING...
  37. [37]
    Understanding and Controlling Intersystem Crossing in Molecules
    Among the spin-dependent interaction terms that enable a crossover between states of different electron spin multiplicities, spin–orbit coupling (SOC) is by far ...<|separator|>
  38. [38]
    Dynamics of Excited Molecules: Predissociation | Chemical Reviews
    The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a ...
  39. [39]
    Multiple Dissociation Pathways in HNCO Decomposition Governed ...
    Sep 23, 2023 · It is well established that multiple dissociation pathways relate to nonadiabatic transitions between multiple coupled PESs.
  40. [40]
    None
    ### Summary of Rotational and Vibrational Spectra from Lecture 2
  41. [41]
    None
    ### Summary of Rotational Spectra of Polyatomic Molecules, Asymmetric Tops, and Microwave Region
  42. [42]
    None
    Summary of each segment:
  43. [43]
    [PDF] Molecular energy levels and spectroscopy
    A transition between two vibrational states gives rise to a vibrational band, made up of P, Q and R branches, corresponding to transitions between rotational ...Missing: PQR | Show results with:PQR
  44. [44]
    [PDF] Intermolecular Forces
    The larger the surface area of the molecules the greater the dispersion forces. • Typical energy 0.05 – 40 kJ mol–1. • Dispersion forces exist between all ...
  45. [45]
    [PDF] Van Der Waals Interactions
    a single van der Waals interaction typically ranges from 0.4 to 4 kJ/mol, compared to covalent bonds which often exceed 200 kJ/mol. Importantly, van der Waals ...<|separator|>
  46. [46]
    [PDF] Lecture 10. The Dispersion Force and van der Waals Interaction
    In this lecture, we will discuss the last contribution to the vdW interactions, the London dispersion energy. 10.1. Classical view of dispersion interactions.
  47. [47]
    [PDF] Viewpoint - American Physical Society
    Apr 8, 2013 · First invoked in 1930 by Fritz London, dispersion forces are a necessary consequence of the quantum electron cor- relations that govern the ...
  48. [48]
    [PDF] Van der Waals Forces (2)
    This is often referred to as the Debye interaction or the induction interaction. It constitutes the second of three inverse sixth power contributions to the ...
  49. [49]
    [PDF] Molecule Matters van der Waals Molecules
    1210. In this article, Debye's contributions to understanding van der Waals attractive forces, specifically, dipole–induced di- pole interaction, is discussed.
  50. [50]
    A double exponential potential for van der Waals interaction - NIH
    Jun 7, 2019 · The Lennard-Jones (L-J) potential is a mathematically simple model that approximates vdw interaction. It was first proposed in 1924 by John ...
  51. [51]
    Keesom Interaction - an overview | ScienceDirect Topics
    Keesom interaction is defined as the interaction between two molecules ... interaction potential w(r) can now be calculated by Eqs. 77, 74, and 73. For ...
  52. [52]
    Hydrogen Bond - an overview | ScienceDirect Topics
    Typically, they have bond strengths between about 10 and 40 kJ mol−1 and lengths of about 0.18 nm[9]. The actual strength of a particular hydrogen bond depends ...
  53. [53]
    Strong Short‐Range Cooperativity in Hydrogen‐Bond Chains - PMC
    Chains of hydrogen bonds are prevalent structural motifs in supramolecular and biological systems. H bonds are widely proposed to exhibit positive cooperativity ...
  54. [54]
    The quadrupole moment of the carbon dioxide molecule - Journals
    A direct measurement of the magnitude and sign of the quadrupole moment of the carbon dioxide molecule has been made by determining the birefringence.
  55. [55]
    Sorption, Structure and Dynamics of CO2 and Ethane in Silicalite at ...
    Dec 28, 2018 · CO2 is a linear molecule with no dipole moment, but with a finite quadrupole moment. This quadrupole moment of CO2 gives rise to interesting ...
  56. [56]
    NMR and IR Investigations of Strong Intramolecular Hydrogen Bonds
    The length of the X–H bond usually increases on hydrogen bond formation leading to a red shift of the infrared X–H stretching frequency and an increase in the ...
  57. [57]
    Intermolecular Interactions and Spectroscopic Signatures of ... - NIH
    The red shift, characteristic to the hydrogen bond formation, is slightly bigger in trimer than in dimer, probably due to the additional spatial restriction ...
  58. [58]
    [PDF] The Molecular Schrödinger Equation - Hunt Research Group
    Introduction. • the basic Schrödinger equation o H is the molecular Hamiltonian, and is an operator, think of it as a special.
  59. [59]
    [PDF] An Introduction to Configuration Interaction Theory - - Sherrill Group
    In the first paper on quantum mechanics, Heisenberg used matrix mechanics to cal- culate the frequencies and intensities of spectral lines [2]. Later, when ...Missing: original | Show results with:original
  60. [60]
    [PDF] The Variation Principle
    This principle allows us to calculate an upper bound for the ground state energy by finding the trial wavefunction ϕ for which the integral is minimised (hence ...
  61. [61]
    [PDF] Perturbation theory
    Feb 24, 2006 · All the material required is covered in “Molecular Quantum Mechanics” fourth edition by Peter Atkins and Ronald Friedman (OUP 2005) ...
  62. [62]
    [PDF] DAMTP - 9 Perturbation Theory II: Time Dependent Case
    In this chapter we'll study perturbation theory in the case that the perturbation varies in time. Unlike the time–independent case, where we mostly wanted ...
  63. [63]
    Semiempirical Quantum Mechanical Methods for Noncovalent ...
    Apr 13, 2016 · Semiempirical (SE) methods can be derived from either Hartree–Fock or density functional theory by applying systematic approximations, ...
  64. [64]
    A mathematical and computational review of Hartree-Fock SCF ...
    May 2, 2007 · Abstract: We present here a review of the fundamental topics of Hartree-Fock theory in Quantum Chemistry. From the molecular Hamiltonian, ...
  65. [65]
    [PDF] Introduction to Quantum Chemistry - Zenodo
    Apr 24, 2025 · • CC methods such as CCSD and CCSD(T) are the most robust methods. ... • CCSD(T) is considered as the current gold standard of post-HF methods.
  66. [66]
    Basis Sets | Gaussian.com
    May 17, 2021 · Gaussian 16 uses basis sets like STO-3G, 3-21G, 6-21G, 4-31G, 6-31G, 6-311G, and cc-pVDZ, cc-pVTZ, cc-pVQZ, cc-pV5Z, cc-pV6Z. STO-3G is the ...
  67. [67]
    Enabling large-scale quantum path integral molecular dynamics ...
    Apr 24, 2023 · Molecular dynamics (MD) simulations have evolved to establish the relationship between atomic structure behavior and the functions or properties ...<|separator|>
  68. [68]
    Molecular Structure and Chirality Detection by Fourier Transform ...
    Dec 19, 2014 · We describe a three-wave mixing experiment using time-separated microwave pulses to detect the enantiomer-specific emission signal of the chiral molecule.
  69. [69]
    Perspective: The first ten years of broadband chirped pulse Fourier ...
    Gas-phase microwave experiments routinely achieve frequency resolution better than 10 kHz, so that a typical full-band spectrum covering 8–18 GHz contains (10 ...II. BACKGROUND AND... · Sensitivity · Resolution · Implementation of CP-FTMW...
  70. [70]
    How an FTIR Spectrometer Operates - Chemistry LibreTexts
    Apr 9, 2023 · (4) The resolution is extremely high (0.1 ~ 0.005 cm-1). (5) The scan range is wide (1000 ~ 10 cm-1). (6) The interference from stray light is ...
  71. [71]
    Improvement of the Detection Sensitivity for Tunable Diode Laser ...
    In this review paper, the important advances in TDLAS detection sensitivity are discussed, including the selection of absorption lines, the improvement of diode ...
  72. [72]
    Trace gas detection with cavity ring down spectroscopy
    Apr 1, 1995 · Trace gas detection of small molecules has been performed with cavity ring down (CRD) absorption spectroscopy in the near UV part of the ...
  73. [73]
    Advances in cavity-enhanced methods for high precision molecular ...
    Jun 7, 2024 · In this review article, we discuss the current status of highly precise and highly sensitive laser spectroscopy for fundamental tests and measurements.
  74. [74]
    Ultrafast pump-probe spectroscopy: femtosecond dynamics in ...
    Ultrafast internal conversion dynamics through the on-the-fly simulation of transient absorption pump–probe spectra with different electronic structure methods.
  75. [75]
    Hetero-site-specific X-ray pump-probe spectroscopy for ... - Nature
    May 23, 2016 · Here we show experimental evidence of a hetero-site pump-probe signal. By using two-colour 10-fs X-ray pulses, we are able to observe the femtosecond time ...
  76. [76]
    On the synergy of matrix-isolation infrared spectroscopy and ...
    Nov 9, 2020 · In the spectrum, these interactions cause a shift in frequencies compared to the gas phase, the so-called matrix shift (typically below 4 cm−1 ...
  77. [77]
    Matrix isolation infrared spectroscopy | NIST
    Sep 4, 2009 · Studies in the gas phase offer the potential for the most precise, detailed infrared spectroscopic measurements.Missing: perturbations | Show results with:perturbations<|control11|><|separator|>
  78. [78]
    [PDF] Improved Methods and Study of 1,1,3,3-tetramethylcyclobutane and ...
    The gas phase electron-diffraction (GED) technique has been the ... Typical bond length accuracies are 0.01-0.004 Å ... In contrast, for small molecules, ...
  79. [79]
    A Critical Assessment of Modern Gas Electron Diffraction Data from ...
    Jun 25, 2020 · In detailed examination and comparison of the obtained results we have determined the average experimental precision of 0.004 Å for bond lengths ...
  80. [80]
    Requiem for gas-phase electron diffraction | Structural Chemistry
    Apr 11, 2023 · Electron diffraction yielded thermal average lengths between the nuclear positions and the rotational transitions yielded distances between ...
  81. [81]
    Neutron scattering and hydrogen storage - ScienceDirect.com
    Hydrogen has the largest scattering interaction with neutrons of all the elements in the periodic table making neutron scattering ideal for studying hydrogen ...
  82. [82]
    [PDF] An Introduction to Neutron Scattering
    Why do Neutron Scattering? • To determine the positions and motions of atoms in condensed matter. – 1994 Nobel Prize to Shull and Brockhouse cited these ...
  83. [83]
    Neutrons “101” – A Primer for Earth Scientists - GeoScienceWorld
    Sep 1, 2021 · Neutrons can lose or gain energy to or from the target atom during inelastic scattering, and this information can be used to describe how atoms ...
  84. [84]
    (IUCr) Models of thermal motion in small-molecule crystallography
    The Debye–Waller factor, introduced a century ago, remains a fundamental component in the refinement of crystal structures against X-ray, ...
  85. [85]
    Everything you always wanted to know about the Debye–Waller ...
    Jun 6, 2025 · The Debye–Waller factor that causes the observed diffraction intensities to decrease at finite temperature due to the atom oscillating back and forth.
  86. [86]
    [PDF] Yuan Tseh Lee - Nobel Lecture
    The idea of crossed molecular beams experiments is in a sense to “visualize” the details of a chemical reaction by tracing the trajectories of the reaction ...
  87. [87]
    Crossed-beam studies of reaction dynamics - PubMed
    This article reviews recent progress in our understanding of gas-phase neutral reaction dynamics as made possible by improvements in the crossed molecular beam ...Missing: seminal | Show results with:seminal
  88. [88]
    Effectively Analyzing Polymers Using IR, Raman Spectroscopy and ...
    Vibrational spectroscopy (infrared (IR) and Raman) is especially useful for the analysis of polymers in an industrial setting. IR and Raman are highly specific ...
  89. [89]
    Spectroscopic Polymer Characterization - ScienceDirect.com
    Spectroscopic techniques for the study of polymers must yield high-resolution, narrow linewidth spectra that provide selectivity and structural information.
  90. [90]
    Application of spectroscopic methods in the polymer industry ...
    A special way of characterization of polymers by means of spectroscopic methods (IR, NMR, UV) is described. The top method is the IR spectroscopy.
  91. [91]
    Molecular Design of Photovoltaic Materials for Polymer Solar Cells
    Jan 30, 2012 · In this Account, I discuss the basic requirements and scientific issues in the molecular design of high efficiency photovoltaic molecules.
  92. [92]
    Rational molecular and device design enables organic solar cells ...
    Feb 28, 2024 · The impact of our molecular design on photovoltaic performance is investigated by devices with conventional structure of indium tin oxide (ITO) ...
  93. [93]
    Inverse molecular design from first principles: Tailoring organic ...
    May 10, 2022 · For light-emitting diodes (LEDs), further development of molecular materials will enable realization of efficient and stable blue LEDs that meet ...
  94. [94]
    Optical Screening and Classification of Drug Binding to Proteins in ...
    We demonstrate that ultrafast 2D-IR spectroscopy can overcome these issues by providing a direct, label-free optical measurement of protein–drug binding in ...
  95. [95]
    Vibrational Spectroscopy as a Tool for Studying Drug-Cell Interaction
    A particularly promising application is the use of vibrational spectroscopic techniques to study the interaction of drugs with cells. Many studies have ...Missing: pharmaceuticals | Show results with:pharmaceuticals<|separator|>
  96. [96]
    Quantitative Chiral Analysis by Molecular Rotational Spectroscopy
    This method offers the possibility for direct chiral analysis of molecules in complex chemical mixtures without the need for chemical separation protocol ...
  97. [97]
    Chiral rotational spectroscopy | Phys. Rev. A
    Sep 8, 2016 · We introduce chiral rotational spectroscopy, a technique that enables the determination of the orientated optical activity pseudotensor components.Abstract · Article Text · INTRODUCTION · SUMMARY AND OUTLOOK
  98. [98]
    Applications of infrared spectroscopy in environmental contamination
    IR spectroscopy is used for environmental contamination analysis of pesticides, cyanide, metalloids, heavy metals, and microplastics.
  99. [99]
  100. [100]
    Direct laser cooling of a symmetric top molecule - Science
    Sep 11, 2020 · Laser cooling of atomic systems has enabled substantial advances in quantum simulation, precision clocks, and quantum many-body physics (1–4).
  101. [101]
  102. [102]
    Refining Fuel Cells: The Critical Role of Ionomer-Catalyst Molecular ...
    Nov 6, 2024 · The interaction between the carbon support, active sites, and ionomer is crucial for oxygen reduction reaction catalysts.
  103. [103]
    Designing fuel cell catalyst support for superior catalytic activity and ...
    Oct 18, 2022 · In this study, we introduce a bottom-up designed spherical carbon support with intrinsic Nitrogen-doping that permits uniform dispersion of Pt catalyst.
  104. [104]
    Ultracold Fermionic Feshbach Molecules of | Phys. Rev. Lett.
    Here we report on the production of ultracold weakly bound fermionic molecules of Na 23 K 40 near a broad Feshbach resonance [13] . The NaK molecule is ...
  105. [105]
    Ultracold Feshbach molecules in an orbital optical lattice - Nature
    Mar 20, 2023 · Here we demonstrate the preparation of ultracold Feshbach molecules of fermionic atoms in the second Bloch band of an optical square lattice.<|separator|>
  106. [106]
    [PDF] Scientific Background to the Nobel Prize in Physics 2023
    Oct 3, 2023 · Studies of light-matter interaction in the attosecond time domain have recently expanded to new vistas, examining molecules, liquids and solids.
  107. [107]
  108. [108]
  109. [109]
    Accuracy of quantum simulators with ultracold dipolar molecules
    This optical lattice can be implemented straightforwardly in experiments by counterpropagating laser beams, whereas the hard-wall boundaries can be engineered ...
  110. [110]
    Toward dynamic structural biology: Two decades of single-molecule ...
    Jan 19, 2018 · Since its first implementation in 1996, smFRET experiments both confirmed previous hypotheses and discovered new fundamental biological ...
  111. [111]
    Single-molecule fluorescence spectroscopy maps the folding ...
    Oct 11, 2011 · Here we introduce a novel method, based on high-throughput single-molecule fluorescence experiments, which is specifically geared towards tracing the dynamics ...
  112. [112]
    Single-molecule spectroscopy of protein folding in a chaperonin cage
    Jun 14, 2010 · We use single-molecule Förster resonance energy transfer to follow the folding of a protein inside the GroEL/GroES chaperonin cavity over a time range from ...<|control11|><|separator|>