Fact-checked by Grok 2 weeks ago

Potential energy surface

A potential energy surface (PES) is a geometric on which the potential energy of a molecular is plotted as a function of the coordinates representing the relative positions or geometries of its constituent atoms. This concept arises from the Born-Oppenheimer approximation, which separates the faster electronic motion from the slower motion due to the mass disparity between electrons and nuclei (typically by a factor of 1838 or more), allowing the PES to serve as an for dynamics. For a with N atoms (N ≥ 3), the PES is a multidimensional with 3N-6 dimensions, corresponding to the independent coordinates after accounting for and . In and , PESs are essential for mapping molecular energy landscapes, identifying stable structures at energy minima (where all curvatures are positive, representing equilibrium geometries like the 0.0958 nm O-H and 104.5° H-O-H in ), and locating transition states at points (with one negative direction, indicating barriers such as in the H + H₂ → H₂ + H process). These surfaces facilitate the study of mechanisms, where reactants and products occupy valleys separated by barriers, and concerted or sequential pathways can be visualized through maps in reduced dimensions (e.g., two coordinates for triatomic systems). PESs are constructed computationally by solving the electronic at various nuclear configurations, often using quantum mechanical methods, and their accurate representation is crucial for predicting rates, spectroscopic properties, and in fields ranging from gas-phase kinetics to condensed-phase simulations.

Fundamentals

Definition

In , the potential energy surface (PES) of a molecular system is defined as a multidimensional representing the E(\mathbf{R}) as a function of the nuclear coordinates \mathbf{R}, which specify the positions of the atomic nuclei. This surface encapsulates the Born-Oppenheimer potential, providing an effective description of how the energy varies with nuclear geometry, excluding the of the nuclei themselves. The PES emerges from the Born-Oppenheimer approximation, introduced in , which assumes that the much larger mass of nuclei compared to electrons allows their motions to be decoupled, enabling the electronic to be computed for fixed positions. Under this approximation, the PES is obtained by solving the electronic for a given \mathbf{R}, yielding the eigenvalues that represent the electronic energy levels, to which the classical -nuclear repulsion term is added. This distinguishes the PES from the pure electronic energy, as the former serves as the full potential governing nuclear dynamics, while the latter is an intermediate result of the electronic structure calculation. For a with N atoms, the PES constitutes a in a (3N-6)-dimensional configuration space for non-linear molecules or (3N-5)-dimensional for linear ones, after removing the three translational and (two or three) rotational of the entire system. The term "potential energy surface" originated in early quantum mechanical treatments of molecular spectra and dynamics, reflecting its geometric analogy to a that guides motions. On this surface, stable molecular conformations correspond to energy minima, small-amplitude deviations from these minima describe vibrational excitations, and chemical reactions trace paths that ascend and descend over energy barriers.

Mathematical Formulation

The Born-Oppenheimer (BO) approximation provides the foundational framework for defining a potential energy surface (PES) by separating the and in a molecular system. This separation is justified by the significant mass difference between nuclei and electrons, which leads to disparate timescales for their motions: nuclear occur on the order of femtoseconds to picoseconds, while electronic rearrangements are much faster, on scales. Under this approximation, the nuclear positions \mathbf{R} are treated as fixed parameters when solving for the electronic wavefunction, allowing the total molecular wavefunction \Psi(\mathbf{r}, \mathbf{R}) to be approximated as a product \Psi(\mathbf{r}, \mathbf{R}) \approx \psi_i(\mathbf{r}; \mathbf{R}) \chi_i(\mathbf{R}), where \mathbf{r} denotes electronic coordinates, \psi_i is the electronic wavefunction for the i-th state, and \chi_i is the nuclear wavefunction./06%3A_Adiabatic_Approximation/6.01%3A_BornOppenheimer_Approximation) The total for the is \hat{H} = \hat{T}_n + \hat{T}_e + \hat{V}_{ee} + \hat{V}_{nn} + \hat{V}_{ne}, where \hat{T}_n and \hat{T}_e are the and operators, and \hat{V}_{ee}, \hat{V}_{nn}, and \hat{V}_{ne} represent electron-electron, nucleus-nucleus, and nucleus-electron interactions, respectively. Substituting the product into the time-independent \hat{H} \Psi = E \Psi and neglecting terms involving nuclear-electronic coupling (due to the m_e / M_n \ll 1), the electronic problem decouples as an eigenvalue equation for fixed \mathbf{R}: \hat{H}_{el}(\mathbf{R}) \psi_i(\mathbf{r}; \mathbf{R}) = E_i(\mathbf{R}) \psi_i(\mathbf{r}; \mathbf{R}), where \hat{H}_{el}(\mathbf{R}) = \hat{T}_e + \hat{V}_{ee} + \hat{V}_{ne} is the electronic parametrized by coordinates, and E_i(\mathbf{R}) is the i-th electronic energy eigenvalue. The repulsion term \hat{V}_{nn}(\mathbf{R}), which depends only on \mathbf{R}, is added separately to form the for motion: V(\mathbf{R}) \approx E_i(\mathbf{R}) + V_{nn}(\mathbf{R}), with V_{nn}(\mathbf{R}) = \sum_{\alpha < \beta} \frac{Z_\alpha Z_\beta e^2}{|\mathbf{R}_\alpha - \mathbf{R}_\beta|}, where Z_\alpha are charges and the sum is over pairs. This V(\mathbf{R}) defines the PES as a in the $3N-dimensional configuration space for N atoms./06%3A_Adiabatic_Approximation/6.01%3A_BornOppenheimer_Approximation) In the adiabatic , the PES corresponds to the diagonal elements E_i(\mathbf{R}) of the in the basis of its instantaneous eigenstates, assuming no transitions between electronic states during nuclear motion. This is the standard BO framework, but it breaks down near regions of near-degeneracy (e.g., avoided crossings), where non-adiabatic couplings \langle \psi_i | \frac{\partial}{\partial \mathbf{R}} | \psi_j \rangle become significant, leading to electronic state mixing. To address such limitations, the diabatic redefines the electronic basis such that the potential matrix elements are smooth functions of \mathbf{R} with weak off-diagonal couplings, facilitating treatments of non-adiabatic dynamics; however, constructing exact diabatic states remains challenging as it requires a basis where couplings are purely nuclear kinetic in origin./06%3A_Adiabatic_Approximation/6.03%3A_Diabatic_and_Adiabatic_States) The PES is typically expressed in nuclear coordinates \mathbf{R}, which can be Cartesian coordinates \{x_\alpha, y_\alpha, z_\alpha\} for all atoms or internal coordinates such as bond lengths r_{ij} = |\mathbf{R}_i - \mathbf{R}_j|, bond angles, and dihedral angles, reducing the dimensionality from $3N to $3N-6 (or $3N-5 for linear molecules) by separating center-of-mass translation and rotation. High dimensionality precludes full visualization, so PES are often analyzed via reduced representations, such as two-dimensional contour plots along reaction coordinates or minimum energy paths./Quantum_Mechanics/11%3A_Molecules/Potential_Energy_Surface)

Computation

Methods

The construction of potential energy surfaces (PESs) often begins with methods, which compute electronic energies at specific nuclear configurations to provide the data points necessary for surface representation. The Hartree-Fock (HF) method serves as a foundational approach, solving the by assuming a single to approximate the wavefunction, yielding single-point energies that form the basis for PES grids in small molecular systems. Post-HF techniques enhance accuracy by accounting for electron correlation; for instance, second-order Møller-Plesset (MP2) corrects HF limitations through perturbative inclusion of double excitations, while coupled-cluster methods like CCSD(T), which incorporate singles, doubles, and perturbative triples excitations, achieve near-quantitative accuracy for benchmark PESs, with typical errors below 1 kcal/mol for small molecules. (DFT) offers a computationally efficient alternative, approximating the exchange-correlation functional to compute energies and forces, widely used for single-point evaluations in larger systems due to its balance of speed and reliability, often with hybrid functionals like B3LYP for improved PES fidelity. Once data points are obtained, fitting and techniques construct analytical representations of the PES to enable efficient evaluation across configuration space. Global fitting methods, such as permutation- polynomial neural networks (PIP-NN), combine invariant to permutations with neural networks to fit high-dimensional PESs, achieving high accuracy for reactive systems like H + H₂ by training on sparse datasets. Local approaches, exemplified by , build the PES as a weighted sum of expansions around selected grid points, suitable for regions near minima or transition states, with modifications like modified enabling smoother surfaces for gas-surface interactions. Machine learning (ML) approaches have revolutionized PES construction in the by learning complex mappings from atomic coordinates to energies and forces, often surpassing traditional fits in scalability and accuracy for large systems. potentials like (Accurate Neural network Interaction) employ hierarchical architectures to predict DFT-level energies for small to medium-sized molecules, transferable across chemical space with relative energy errors around 0.6 kcal/ after training on extensive configurations. Recent extensions, such as ANI-2x and Schrödinger-ANI (as of 2024), support larger systems up to hundreds of atoms. SchNet, a continuous-filter , models quantum interactions via message-passing on atomic graphs, reproducing global PESs for small molecules with mean absolute errors below 0.1 kcal/ while ensuring rotational invariance. (GPR) provides in PES fitting, constructing smooth surfaces from sparse data via functions, as demonstrated in high-dimensional models for large molecules where it achieves chemical accuracy with thousands of training points. Recent advances (as of 2025) include automated frameworks for exploring and fitting PESs, such as those using for efficient data generation, and Kolmogorov-Arnold Networks (KANs) substituting multi-layer perceptrons in ML frameworks for improved prediction of PES and properties. DeepMD uses deep s to learn many-body potentials, enabling accuracy in simulations of solids and liquids, with applications in showing forces predicted to within 0.1 eV/Å. Additionally, physics-informed methods improve PES quality by incorporating experimental data like Feshbach resonances (2024). For simpler systems, semiclassical and empirical potentials offer analytically tractable approximations without full quantum calculations. The , V(r) = D_e (1 - e^{-a(r - r_e)})^2, captures anharmonic diatomic vibrations with dissociation energy D_e and equilibrium distance r_e, widely used for gas-phase diatomics in early PES studies. The , V(r) = 4ε [(σ/r)^{12} - (σ/r)^6], models van der Waals interactions in non-bonded atoms, essential for representing repulsive and attractive walls in rare-gas clusters and surface adsorption PESs. These empirical forms, parameterized from experimental data, facilitate rapid dynamics but lack transferability to reactive regimes. Specialized software facilitates the implementation of these methods. For calculations, Gaussian provides robust tools for , post-HF, and DFT single-point energies, supporting PES grid generation via and optimization modules. MOLPRO excels in multireference correlated methods like CASSCF and MRCI for accurate PESs in excited states, with built-in surface-fitting utilities. The program system specializes in high-level multiconfiguration interaction for non-adiabatic PESs, enabling state-specific energy computations. For ML potentials, LAMMPS integrates interfaces to models like DeepMD and SchNet, allowing on-the-fly evaluation during large-scale . PESs can be represented either through pre-computed grid-based methods, where energies are tabulated on a discrete mesh for in , or via on-the-fly computation, where electronic structure calculations are performed dynamically at each step to avoid fitting errors, though at higher cost; the depends on system dimensionality and required accuracy.

Challenges

Computing accurate potential energy surfaces (PESs) faces significant scaling challenges due to the exponential increase in computational cost with system size, particularly for high-level ab initio methods like coupled-cluster singles, doubles, and perturbative triples, CCSD(T). This method scales as approximately O(N^7), where N is the number of basis functions, limiting its application to systems of up to about 50 atoms even with modern hardware and local correlation approximations. For larger molecules, such as biomolecules or clusters, the prohibitive cost necessitates approximations or lower-level theories, compromising accuracy in regions like transition states or dissociation limits. Multireference cases, often arising from near-degeneracies in electronic states (e.g., in bond-breaking processes or transition metals), further complicate PES . Single-reference methods like CCSD(T) fail in these regions, leading to qualitative errors such as incorrect behaviors. Multireference approaches, such as complete active space self-consistent field (CASSCF) or multireference configuration interaction (MRCI), are required to properly describe static correlation, but they introduce additional challenges like selecting an appropriate active space and handling intruder states, increasing both setup complexity and computational expense. High-dimensionality and exacerbate these issues, as PESs for polyatomic molecules span $3N-6 internal coordinates (N being the number of atoms), invoking of dimensionality in global fitting or gridding. Fitting accurate representations, such as permutation-invariant polynomials or neural networks, requires exponentially more data points as dimensionality grows, making exhaustive sampling infeasible for systems beyond triatomics without techniques like decoupling or . , prominent in floppy modes or large-amplitude vibrations, further demands non-perturbative treatments to capture coupling effects accurately. Key error sources in PES computations include basis set incompleteness, which underestimates correlation energies by 1-5 kcal/mol depending on the basis quality, and truncation of electron correlation, as in approximate methods that neglect higher excitations. These systematic errors propagate into derived properties like reaction barriers or vibrational frequencies. To mitigate them, strategies intelligently sample key regions—such as minima, saddles, and paths—using from surrogate models (e.g., Gaussian processes) to prioritize evaluations, reducing the total number of expensive calculations by orders of magnitude while focusing on chemically relevant configurations. Validation of PESs relies on comparison to experimental data, particularly from (e.g., or Raman spectra) to vibrational levels and anharmonicities, or scattering experiments for . Discrepancies, such as shifts in zero-point energies exceeding 10 cm⁻¹, highlight inaccuracies. For machine learning-based PESs (ML-PESs), uncertainty quantification via ensemble methods or provides on predictions, enabling reliable use in simulations by flagging extrapolation risks. Strategies to address these challenges include multi-level methods like ONIOM (Our own N-layered Integrated and ), which partitions the system into high-accuracy regions treated with methods and low-accuracy surroundings approximated at semi-empirical or levels, achieving CCSD(T)-like accuracy for systems up to hundreds of atoms at reduced cost. Fragment-based approaches, such as energy-based fragmentation, decompose the PES into additive contributions from molecular fragments, extrapolated to the full system, effectively handling large-scale computations while controlling truncation errors through systematic convergence tests. Recent tools like , a robust adaptive nature-inspired global explorer (2025), enhance efficiency in optimizing high-dimensional PES landscapes.

Properties

Key Features

The potential energy surface (PES) of a molecular system is characterized by stationary points where the of the with respect to nuclear coordinates vanishes. These points include minima, which correspond to stable molecular structures such as equilibrium geometries, where the (second derivatives of the ) is positive definite, indicating positive curvatures in all directions. Maxima represent unstable configurations with negative curvatures in all directions, though they are rare and less physically relevant in typical PES landscapes. Saddle points, particularly first-order saddles, feature one negative curvature (along the ) and positive curvatures in the remaining directions; these define transition states that separate reactant and product minima and govern the of chemical . Reaction paths on the PES connect these stationary points and delineate the progression of chemical transformations. The minimum energy path () is the lowest-energy trajectory linking reactants to products via the , defined as the path where the energy is minimized perpendicular to the path direction at every point. The intrinsic reaction coordinate (IRC), introduced by Fukui, provides a mass-weighted variant of the , following the steepest descent in mass-scaled coordinates from the , ensuring it captures the dynamical flow of the reaction without external forces. Energy barriers along these paths represent the height difference between minima and adjacent saddle points, dictating reaction rates in . In quantum treatments, tunneling permits reactions at energies below the classical barrier by allowing wavefunctions to penetrate the forbidden region, with the tunneling probability increasing for thinner or lower barriers. (ZPE) corrections, arising from vibrational ground-state energies, raise the effective barrier height since ZPE is typically lower at the than at minima, reducing the classical barrier but often not eliminating the need for tunneling in light-atom transfers. The of the PES also includes valleys, which are broad low-energy regions funneling trajectories toward minima and representing conformational basins; , which are high-energy crests separating valleys and often associated with points; and seams, such as seams in multi-state systems, where energy levels cross and enable non-adiabatic transitions. These features influence branching in reactive , where trajectories may diverge at or seam points, leading to multiple product channels from a single collision. Locating these features involves optimization techniques tailored to the PES curvature. The steepest descent method follows the negative direction from an initial guess, converging to the nearest minimum or along reaction paths by iteratively reducing energy in the direction of maximum slope. The Newton-Raphson method, a second-order approach, uses both the and to predict steps toward stationary points, offering quadratic convergence near minima or saddles when the Hessian is appropriately projected to enforce the desired index (number of imaginary frequencies). Visualization of PES features aids in interpreting multidimensional landscapes. For two-dimensional slices, such as collinear triatomic systems, contour plots display energy levels as nested curves, with minima as enclosed lows and saddles as col-like passes. In three or more dimensions, isosurfaces render constant-energy volumes, revealing valleys as extended troughs and ridges as elevated divides, though projection artifacts can obscure details in higher dimensions. In excited electronic states, distinctions between attractive and repulsive PES branches can alter these features, such as by introducing seams that facilitate rapid radiationless decay.

Attractive and Repulsive Surfaces

In the context of excited electronic states, potential energy surfaces (PES) are often categorized as attractive or repulsive based on their shape along internuclear coordinates in diatomic molecules. Attractive PES feature a potential well with a minimum energy at an equilibrium separation, supporting bound vibrational and rotational states, as seen in the ground-state ^1Σ_g^+ configuration of homonuclear diatomics like , where electron sharing stabilizes the bond. Conversely, repulsive PES exhibit no such minimum, resulting in monotonically increasing energy with decreasing separation and leading to spontaneous dissociation; the lowest excited ^1Σ_u^+ state of exemplifies this, correlating asymptotically to two ground-state hydrogen atoms without barrierless separation from the united atom limit. These distinctions arise from symmetry and electron configuration: gerade (g) states like ^1Σ_g^+ promote bonding due to constructive orbital overlap, while ungerade (u) states like ^1Σ_u^+ enforce antibonding character from destructive interference. For diatomic systems, attractive PES are commonly modeled using the Morse potential, which captures the anharmonic well depth and equilibrium distance: V(R) = D_e \left[1 - e^{-a(R - R_e)}\right]^2, where D_e is the dissociation energy from the bottom of the well, R_e is the equilibrium bond length, and a governs the well's width; this form approximates quantum mechanical solutions for bound states near the equilibrium. Repulsive PES, in contrast, are dominated by short-range exponential terms reflecting core-core and Pauli repulsion, often expressed as V(R) \propto e^{-bR} for the inner wall, transitioning to weaker long-range attractions like van der Waals forces at large separations; such forms ensure rapid energy rise at small R, precluding stable molecules. These analytical models facilitate understanding of state-specific behaviors without full ab initio computation. Interactions between multiple PES introduce conical intersections and avoided crossings, critical for excited-state dynamics. Conical intersections arise when two PES of the same spin and symmetry degeneracy occur, forming a double cone topology in the energy landscape; encircling such a point induces a Berry phase shift of π in the electronic wave function, as theoretically established for polyatomic systems. Avoided crossings, meanwhile, occur between PES of identical symmetry in the adiabatic basis, where non-adiabatic coupling prevents degeneracy, resulting in a gap and oscillatory energy profiles. These features enable efficient radiationless transitions, with conical intersections providing funnel-like pathways for ultrafast decay from excited to lower states. Non-adiabatic effects at these junctions govern trajectory branching in , often simulated via surface hopping algorithms that probabilistically switch between PES based on coupling strengths. In fewest-switches surface hopping (FSSH), motion follows classical trajectories on the adiabatic surface, with hops to another surface determined by the time-dependent Schrödinger equation's non-adiabatic probability flux, conserving energy through velocity rescaling; this approach captures decoherence and near intersections. Such dynamics are pivotal in , where vertical excitation to a repulsive PES drives impulsive bond breaking, yielding fragment kinetic energies reflective of the surface's slope, and in charge transfer reactions, where repulsive ionic surfaces facilitate electron hopping between donor-acceptor pairs at close approach. Multi-state PES are represented in either adiabatic or diabatic bases to handle s. The adiabatic diagonalizes the electronic for fixed nuclei, yielding PES as eigenvalues with off-diagonal non-adiabatic couplings that vary strongly with , complicating near intersections. The diabatic , by contrast, rotates the basis to make electronic wave functions nearly nuclear-coordinate-independent, transferring coupling to constant off-diagonal matrix elements while producing crossing (rather than avoided) surfaces; this is advantageous for quasiclassical treatments, as it simplifies hopping probabilities and preserves in multi-state manifolds. The transformation between representations relies on unitary matrices satisfying the non-adiabatic coupling conditions, enabling accurate modeling of state mixing in excited systems.

Applications

Reaction Dynamics

Potential energy surfaces (PES) serve as the foundational framework for elucidating reaction mechanisms and in chemical , mapping the variations as a of nuclear coordinates to predict how reactants evolve into products. By integrating PES with dynamical methods, researchers can simulate trajectories, compute constants, and analyze outcomes, bridging theoretical models with experimental observations. These surfaces, often constructed from calculations, allow for the exploration of both classical and quantum effects in reactive processes. Classical trajectory methods, particularly quasiclassical trajectory (QCT) calculations, propagate ensembles of classical on a PES to study and reaction probabilities, incorporating initial quantum via sampling techniques. QCT is widely used for gas-phase reactions, providing insights into disposal, product distributions, and cross-sections by averaging over thousands of initialized with realistic distributions. This approach excels in multidimensional systems where full quantum is computationally prohibitive, revealing details like recrossing at transition states that refine rate predictions. Transition state theory (TST) derives thermal rate constants from the barrier heights and geometries on the PES, assuming a quasi-equilibrium between reactants and the while neglecting post-barrier recrossing. variational TST optimizes the dividing surface along the reaction path to minimize recrossing, improving accuracy for barrierless or submerged barrier reactions by evaluating the at the variational . These methods rely on the PES topology near the to compute functions and coefficients, offering a statistical foundation for in complex environments. Quantum dynamics extends TST through variational formulations and instanton theory to account for tunneling, where the path represents the most probable tunneling trajectory on the PES, enabling computation of quantum rate constants via approximations. Variational quantum TST incorporates multidimensional tunneling corrections, such as small-curvature or semiclassical methods, to capture below-barrier influenced by the PES . Instanton theory, particularly in its ring-polymer form, efficiently handles low-temperature regimes where tunneling dominates, providing benchmark rates for validating PES accuracy. In reactive scattering, PES features dictate direct mechanisms, involving prompt abstraction or stripping with minimal complex formation, versus indirect mechanisms that proceed through long-lived intermediates, as probed by crossed molecular beam experiments and dynamics simulations. Opacity functions, which quantify reaction probability as a function of impact parameter, reveal these distinctions; low impact parameters favor indirect paths with isotropic , while higher ones promote direct rebound or stripping with forward/backward biases. High-fidelity PES enable quantum calculations to resolve interferences and resonances shaping differential cross-sections. Isotope effects in arise from PES mass dependence, altering vibrational frequencies and zero-point energies that influence barrier and product . Heavier reduce tunneling and shift stereodynamics, such as product rotational alignment, toward classical limits, as seen in differential cross-sections and polarization-dependent distributions computed on isotopic PES variants. Stereodynamic probes, including vector correlations, trace how PES controls approach geometries and product orientations, highlighting orientation-dependent reactivity. PES play a pivotal role in by enabling rate constant calculations for ion-molecule reactions in the , where low temperatures amplify tunneling and barrierless paths on multiwell surfaces. In modeling, detailed PES explorations of multiwell systems, such as alkyl + O₂ reactions, inform simulations of pressure-dependent , capturing isomerizations, stabilizations, and chain-branching essential for ignition and formation predictions. These applications underscore the need for global, accurate PES to integrate dynamics into large-scale kinetic models.

Spectroscopy and Photochemistry

In vibrational spectroscopy, the curvature of the potential energy surface (PES) at equilibrium geometries provides essential information about molecular vibrations, particularly through the Hessian matrix, which yields the second derivatives of the potential with respect to nuclear coordinates for normal mode analysis. This harmonic approximation underpins the prediction of fundamental frequencies observed in infrared (IR) and Raman spectra, where normal modes represent collective atomic displacements that diagonalize the Hessian, facilitating the assignment of spectral bands to specific molecular motions. Beyond the harmonic limit, anharmonic effects arising from higher-order terms in the PES expansion enable the interpretation of overtone and combination bands, which appear at frequencies that deviate from simple integer multiples of fundamentals due to coupling and distortion along the surface. For instance, in polyatomic molecules, these anharmonicities, captured via perturbation theory or variational methods on the PES, explain the richness of vibrational spectra in systems like water clusters. Electronic spectroscopy relies on PESs to quantify the intensities of vibronic transitions via Franck-Condon factors, which measure the overlap of vibrational wavefunctions between ground and excited states during vertical transitions where positions remain fixed on the ultrafast timescale. These factors, computed from the and differences between the two PESs, determine the relative strengths of or lines, as seen in the UV-visible spectra of diatomic molecules like , where the progression of vibrational structure reflects the offset between potential minima. In polyatomic systems, Duschinsky rotations—mixing of normal modes between states—further modulate these factors, providing insights into changes upon excitation. Photochemical reactions often proceed on excited-state PESs, where conical intersections or avoided crossings facilitate rapid or by allowing nonadiabatic transitions back to the . For example, in the of in , the excited-state PES features a barrierless path leading to a , enabling ultrafast torsion and recovery of the ground-state within picoseconds. Similarly, in molecules like OCS involves repulsive excited-state surfaces that drive bond breaking, with the PES topology dictating product quantum yields and branching ratios. These surfaces, often mapped using multireference , reveal how initial excitation populates reactive pathways distinct from thermal ground-state mechanisms. Time-resolved spectroscopic methods, such as pump-probe techniques, directly probe dynamics along PESs, capturing wavepacket evolution and surface crossings in real time. In the of , ultrafast transient absorption reveals motion from the Franck-Condon region toward a on the excited-state PES, completing trans-to-cis in about 100 femtoseconds before . These experiments validate theoretical PES constructions by matching observed timescales to simulated trajectories, highlighting the role of vibrational in steering photochemical outcomes. In applications to biomolecules, protein folding landscapes are often approximated as high-dimensional PESs, where minima correspond to folded states and barriers represent kinetic traps, guiding the interpretation of spectroscopic probes like fluorescence resonance energy transfer (FRET) to map energy funnels. Seminal energy landscape theory posits that evolution has shaped these surfaces to be minimally frustrated, funneling denatured ensembles toward the native fold, as evidenced in simulations of proteins like CI2 where rugged terrains near the unfolded basin smooth out approaching the minimum. Such approximations extend PES concepts from small molecules to macromolecular dynamics, aiding the design of folding modulators via spectroscopic validation.

Examples

H + H₂ System

The H + H₂ exchange reaction, H + H₂ → H₂ + H, represents a foundational benchmark in the study of bimolecular reactions, serving as the simplest prototypical system for atom-diatom exchange dynamics. In its collinear configuration, the potential energy surface (PES) is effectively two-dimensional, parameterized using Jacobi coordinates: the radial distance R between the incoming H atom and the center of mass of the H₂ molecule, and the intramolecular H-H bond length r of the diatom. This reduction highlights the essential features of reactive scattering without the complications of full three-dimensional geometry, while the full 3D PES incorporates an additional polar angle \theta to describe bending modes. The symmetric nature of the H₃ system features identical asymptotic reactant and product channels connected via a central saddle-point barrier near an equilateral triangular geometry, with shallow van der Waals wells (~75 μE_h or ~0.002 eV deep) in the entrance and exit channels. Key characteristics of the H + H₂ PES include a central saddle-point barrier at approximately 0.42 eV above the reactant , located near the geometry with bond lengths of about 1.8 a₀. This barrier supports two primary reactive s observable in plots of the PES: the direct (or stripping) pathway, akin to an Eley-Rideal-like process where the incoming H atom abstracts one H from H₂ at larger separations with minimal deflection, and the , involving a leading to backscattered products after interaction with the repulsive wall. The barrier's position shifts effectively from early (favoring translational for reaction) in the vibrational of H₂ to late (favoring vibrational ) when H₂ is vibrationally excited, as vibrational motion in the reactant aligns the system more favorably with the per Polanyi's rules, enhancing reactivity and forward scattering. plots of early PES representations, such as those derived from valence-bond formulations, vividly illustrate these paths: the curves gently toward the product at extended R, while trajectories bounce off high- contours near the origin. Early representations of the H + H₂ PES include the analytical Porter-Karplus surface from 1964, a semiempirical valence-bond model incorporating overlap and three-center exchange terms that provided the first realistic depiction for collinear quantum calculations, despite some inaccuracies in barrier height and exothermicity. Subsequent improvements came with fitted surfaces, such as the Muckerman surface from 1979, which refined the PES using configuration interaction methods to better match spectroscopic data and enable accurate quasiclassical trajectory simulations of cross sections. These historical PES facilitated exact quantum reactive computations, revealing Feshbach resonances—temporary in quasibound states above the barrier—that modulate probabilities at low energies, with on the order of picoseconds. Moreover, stereodynamic studies on isotopic variants like H + highlight orientation-dependent effects, where reactant alignment influences product rotational distributions and differential cross sections, as seen in crossed-beam experiments resolving forward versus backward . As a prototype, the H + H₂ PES has profoundly influenced extensions to polyatomic systems, providing a testing ground for scaling methods like diatomics-in-molecules and permutation-invariant polynomials to higher dimensions, where multidimensional barriers and mode-specific dynamics emerge in reactions like H + CH₄. Its simplicity allows exact quantum benchmarks that validate approximations for complex PES, such as those involving multiple wells or conical intersections, paving the way for accurate modeling of and .

Multidimensional Systems

In polyatomic molecules, potential energy surfaces (PES) become high-dimensional, with the number of internal degrees of freedom given by $3N-6 for nonlinear systems, where N is the number of atoms, leading to significant computational complexity. For instance, the dissociation of CH₄ on a Ni(111) surface requires a 15-dimensional PES to capture the full configuration space, including molecular orientation and surface interactions, as developed using neural network fitting to ab initio data. In larger systems like proteins, the PES dimensionality reaches thousands due to the vast conformational space, where folding pathways are explored through coarse-grained models that approximate the landscape as a funnel-shaped surface supporting numerous local minima. To manage this complexity, reduced-dimensionality approaches project the full PES onto key coordinates, such as the Hamiltonian, which separates motion along a minimum from transverse vibrations and rotations. Hamiltonians facilitate calculations by treating the explicitly while coupling it to a subset of orthogonal modes, enabling rate constant evaluations via flux correlation functions. Dividing surfaces, often perpendicular to the at the , further simplify barrier crossing analyses in for multidimensional systems. Global PES aim to represent the entire configuration space analytically, often via permutationally invariant polynomials or neural networks fitted to points, ensuring accuracy across all regions including asymptotes. In contrast, local PES focus on specific features like seams or states, using methods for points without requiring full-space coverage, which is practical for polyatomic dynamics where global fits demand extensive data. Representative examples illustrate these concepts in polyatomic reactions. The SN₂ reaction Cl⁻ + CH₃Cl → ClCH₃ + Cl⁻ features a double-well PES with a central barrier, where front-side and back-side attacks lead to inversion or retention mechanisms, as captured in a full-dimensional surface revealing ion-pair complexes. Pericyclic reactions like the Diels-Alder between and ethene exhibit a concerted PES with a single , where the surface's is modulated by substituents, as explored in grid-based RHF calculations showing asynchronous distortions in electron-deficient cases. Sampling high-dimensional PES poses challenges due to rare events and ergodicity breaking, addressed by enhanced methods like simulations for equilibrium distributions and , which add history-dependent Gaussians to bias collective variables and reconstruct free energies. efficiently explores barriers in polyatomic systems by flattening the landscape along reaction coordinates, though convergence requires careful variable selection to avoid incomplete sampling of orthogonal modes. Recent applications include QM/MM-derived PES for , where treats the and the surrounding protein, enabling profiles for reactions like Kemp elimination in ketosteroid isomerase, revealing electrostatic stabilization of transition states. These hybrid surfaces capture multidimensional effects, such as proton transfers coupled to conformational changes, with barriers reduced by 10-20 kcal/mol compared to solution phase.

History and Advances

Historical Development

The concept of the potential energy surface (PES) originated with the , proposed by and in 1927. This seminal work separated the fast electronic motion from the slower nuclear motion in molecules, treating the electronic energy as a function of fixed nuclear positions and thereby laying the theoretical foundation for representing molecular energies as hypersurfaces in nuclear coordinate space. Building on this, extended in 1928 to encompass chemical reactions, providing early insights into how interactions could generate barriers and valleys for reactive pathways. This extension highlighted the role of PES in describing activation processes and reaction mechanisms, bridging with chemical reactivity. A practical milestone came in 1931 when Henry Eyring and constructed the first semi-empirical PES for the H + H₂ reaction using valence bond principles combined with empirical estimates of integrals. Their London-Eyring-Polanyi (LEP) surface approximated the energy landscape for collinear triatomic configurations, enabling qualitative predictions of reaction rates and introducing the idea of a for the . From the 1940s through the 1960s, researchers applied valence bond and emerging theories to refine PES for more complex systems, focusing on accurate representations of minima, barriers, and intersections. Notable efforts included Joseph O. Hirschfelder's calculations of PES for the H₃ system, which explored multidimensional features and dynamics using variational methods. These developments emphasized semi-empirical adjustments to capture bonding and repulsive interactions, advancing applications in and simple modeling. The 1970s marked a pivotal shift toward computations of PES, facilitated by the availability of supercomputers that handled the intensive calculations required for non-empirical methods. This era enabled higher-fidelity surfaces for triatomic and larger systems, reducing reliance on approximations and improving predictive power for reaction dynamics. Prominent contributors included Karplus, whose work in the 1960s and 1970s integrated empirical and early PES into simulations, elucidating conformational changes and energy transfer. Similarly, Donald G. Truhlar advanced quasiclassical trajectory methods on PES during this period, quantifying rate constants and stereodynamics for polyatomic reactions. Their efforts established PES as central to computational studies of chemical processes.

Modern Developments

In the 2000s, coupled-cluster (CC) methods achieved greater accuracy for potential energy surfaces (PESs) of larger molecular systems through the development of local correlation approaches, enabling near-spectroscopic precision for molecules with up to several dozen atoms while reducing computational cost. Concurrently, double-hybrid density functional theory (DFT) emerged as a cost-effective alternative, incorporating second-order perturbation theory with hybrid exchange-correlation functionals to improve PES descriptions for thermochemistry and reaction barriers, as exemplified by the B2PLYP functional. The marked a revolution in PES construction driven by (ML), with kernel-based methods like scalable Gaussian process regression (sGDML) enabling data-efficient modeling of global force fields at coupled-cluster accuracy for molecular dynamics simulations of organic molecules. Deep neural networks further advanced transferable PESs, as in PhysNet, which predicts energies, forces, and dipole moments with chemical accuracy across diverse systems by embedding physical symmetries and equivariances into the architecture. From 2020 to 2025, contributed to PES exploration via the (VQE) on noisy intermediate-scale quantum (NISQ) devices, demonstrating feasibility for mapping diatomic curves and small-molecule ground states with hybrid quantum-classical workflows. techniques, incorporating uncertainty-driven sampling, enhanced PES efficiency by adaptively selecting points during , reducing training data needs by orders of magnitude while maintaining fidelity in complex landscapes. For multi-state PESs, non-adiabatic advanced in the early 2020s, with methods like those using to compute derivative couplings from energy and gradient data alone, enabling accurate simulations of photochemical processes without explicit wavefunction calculations, as implemented in frameworks for polyatomic systems. Recent applications of PESs to have focused on catalytic surfaces and defects, where ML-interpolated potentials model adsorption energies and reaction pathways on metal oxides and perovskites, accelerating discovery of efficient electrocatalysts for CO₂ reduction. Looking ahead, integrating PES modeling with experiments through inverse design—leveraging generative to optimize molecular structures from desired properties—promises accelerated materials , with ongoing efforts emphasizing closed-loop workflows combining simulations and high-throughput by 2025.

References

  1. [1]
    potential-energy surface (P04780) - IUPAC Gold Book
    potential-energy surface ... A geometric hypersurface on which the potential energy of a set of reactants is plotted as a function of the coordinates representing ...
  2. [2]
    [PDF] Potential Energy Surfaces
    A potential energy surface is an effective potential en- ergy function for the relevant nuclear degrees of freedom. The latter are usually defined as all ...
  3. [3]
    Potential Energy Surface - Chemistry LibreTexts
    Oct 19, 2023 · Potential energy surface (PES) is a mathematical or a graphical function that shows a relationship between the energy of a molecule and its ...
  4. [4]
    Potential Energy Surface - an overview | ScienceDirect Topics
    Potential energy surface (PES) is defined as a representation of the energy of a molecular system as a function of the positions of its nuclei, arising from ...
  5. [5]
    6.1: Born–Oppenheimer Approximation - Chemistry LibreTexts
    Dec 12, 2020 · These effective potentials are known as Born–Oppenheimer or adiabatic potential energy surfaces (PES). For the nuclear degrees of freedom ...
  6. [6]
    Zur Quantentheorie der Molekeln - Born - 1927 - Wiley Online Library
    Click on the article title to read more ... Zur Quantentheorie der Molekeln. M. Born,. M. Born. Search for more papers by this author · R. Oppenheimer,.
  7. [7]
    Potential‐Energy Surfaces, the Born–Oppenheimer Approximations ...
    Jun 27, 2016 · The concept of a potential-energy surface (PES) is central to our understanding of spectroscopy, photochemistry, and chemical kinetics.
  8. [8]
    The Potential-Energy Surface Can Be Calculated Using Quantum ...
    Mar 8, 2025 · The Dimensionality of a Potential Energy Surface​​ a system that is described by a function of the three distances. For more complicated systems, ...Mathematical Definition and... · The Dimensionality of a...
  9. [9]
    Potential Energy Surfaces — CH40208 - Python in Chemistry
    For a system containing N atoms the relative positions of these atoms is described by ( 3 N − 6 ) degrees of freedom (this becomes ( 3 N − 5 ) in the case of a ...Missing: dimensionality | Show results with:dimensionality
  10. [10]
    463 Potential Energy Surfaces - Oxford Academic
    A unimolecular reactant molecule consisting of N atoms has a multidimensional potential energy surface which depends on 3N-6 independent coordinates. For the ...
  11. [11]
    2.6: Potential Energy Surfaces - Chemistry LibreTexts
    Apr 3, 2016 · A potential energy surface (PES) describes the potential energy of a system, especially a collection of atoms, in terms of certain parameters.
  12. [12]
    [PDF] XII. The Born-Oppenheimer Approximation ∑ ∑ ∑ ∑ ∑ - MIT
    The Born-. Oppenheimer. (BO) approximation is probably the most fundamental approximation in chemistry. From a practical point of view, it will allow.
  13. [13]
    On the Choice of the Ab Initio Level of Theory for Potential Energy ...
    The average accuracies of the Hartree–Fock and MP2 methods are 1500–8000 and 400–1000 cm–1, respectively. The standard CCSD(T) method provides errors of 900 ...
  14. [14]
    Potential energy surfaces and dynamic properties via ab initio ...
    Feb 20, 2024 · In hybrid functionals, varying the percentage of exact exchange derived from the Hartree-Fock method using Kohn-Sham orbitals have been ...
  15. [15]
    Permutation invariant polynomial neural network approach to fitting ...
    Aug 6, 2013 · This so-called PIP-NN approach is applied to the H + H2 and Cl + H2 systems and the analytical potential energy surfaces for these two systems ...Missing: SHEP | Show results with:SHEP
  16. [16]
    Modified Shepard interpolation of gas-surface potential energy ...
    The interpolation developed here is suitable for constructing potential energy surfaces by sampling classical trajectories using the Grow procedure and ...
  17. [17]
    ANI-1: an extensible neural network potential with DFT accuracy at ...
    ANI is a new method designed with the intent of developing transferable neural network potentials that utilize a highly-modified version of the Behler and ...
  18. [18]
    SchNet: A continuous-filter convolutional neural network for ... - arXiv
    Jun 26, 2017 · This includes rotationally invariant energy predictions and a smooth, differentiable potential energy surface. Our architecture achieves ...
  19. [19]
    Gaussian Process Regression for Materials and Molecules
    Aug 16, 2021 · Here, we present a detailed study on constructing potential energy surfaces using a machine learning method, namely, Gaussian process regression ...
  20. [20]
    Deep Potential Molecular Dynamics: A Scalable Model with the ...
    Apr 4, 2018 · We introduce a scheme for molecular simulations, the deep potential molecular dynamics (DPMD) method, based on a many-body potential and interatomic forces.
  21. [21]
    [PDF] Interpolation and fitting of potential energy surfaces - Uni Siegen
    Mar 17, 2002 · plotting of the energy points as a function of one or two coordinates might help to find erroneous input data. – Ralph Jaquet, University Siegen ...
  22. [22]
    Lennard-Jones potential - Wikipedia
    The Lennard-Jones potential (also termed the LJ potential or 12-6 potential; named for John Lennard-Jones) is an intermolecular pair potential.
  23. [23]
    Molpro quantum chemistry package
    Molpro is a comprehensive system of ab initio programs for advanced molecular electronic structure calculations, designed and maintained by H.-J. Werner and PJ ...Ordering information · Molpro manual · Support · Gmolpro graphical user interface
  24. [24]
    The COLUMBUS Quantum Chemistry Programs Home Page - GitLab
    Aug 21, 2023 · Introduction. COLUMBUS is a collection of programs for high-level ab initio molecular electronic structure calculations.
  25. [25]
    [PDF] Using LAMMPS to map potential energy surfaces for rapid ionic ...
    LAMMPS is used to map potential energy surfaces for ionic liquids, enabling rapid force field development and simulation-assisted materials development.
  26. [26]
    MCTDH on-the-fly: Efficient grid-based quantum dynamics without ...
    Apr 4, 2018 · One does not need to pre-fit or pre-compute a global PES; instead, the potential energy operator is learnt on-the-fly using PES evaluations ( ...
  27. [27]
    State-of-the-art local correlation methods enable affordable gold ...
    Aug 28, 2024 · The largest conventional CCSD(T) computations with a reliable basis set convergence can reach 20–30 atoms, and this limit can be roughly doubled ...
  28. [28]
    Potential energy surface studies via a single root multireference ...
    Aug 18, 2010 · The systems considered here exhibit varying degrees of degeneracy at different regions of PES. The treatment of these systems via traditional ...
  29. [29]
    Fitting high-dimensional potential energy surface using active ...
    Nov 4, 2015 · With the growth of the dimensionality, approximation becomes even more difficult, resulting in the curse of dimensionality and subsequent ...
  30. [30]
    [PDF] POTENTIAL ENERGY SURFACES
    expansion technique for obtaining a wave function may suffer from two interrelated sources of error: the incompleteness of the one- and n-particle basis sets.
  31. [31]
    Benchmarking State-of-the-Art Ab Initio Potential Energy Surfaces
    Jul 27, 2025 · High-quality potential energy surfaces (PESs) are prerequisites for quantitative atomistic simulations with both quantum and classical ...
  32. [32]
    Fast uncertainty estimates in deep learning interatomic potentials
    Apr 27, 2023 · We propose a method to estimate the predictive uncertainty based on a single neural network without the need for an ensemble.Missing: spectroscopy | Show results with:spectroscopy
  33. [33]
    [PDF] The ONIOM Method and Its Applications - UC Berkeley math
    Apr 8, 2015 · ONIOM(B3LYP/6-311++G(d,p):UFF) reactive potential energy surfaces for decomposition of methylnitramide in the gas phase (black, solid lines) ...
  34. [34]
    [2411.12467] On Multilevel Energy-Based Fragmentation Methods
    Nov 19, 2024 · Energy-based fragmentation methods approximate the potential energy of a molecular system as a sum of contribution terms built from the energies of particular ...
  35. [35]
    Exploring potential energy surfaces for chemical reactions: An ...
    Jul 11, 2003 · Potential energy surfaces form a central concept in the application of electronic structure methods to the study of molecular structures, ...
  36. [36]
    The reaction path intrinsic reaction coordinate method and the ...
    This lowest energy path or minimum energy path (MEP) is often referred as reaction path (RP). From a mathematical point of view, the RP is defined as a curved ...
  37. [37]
    Nature of the chemical reaction path - Fukui - Wiley Online Library
    Mathematical formulation of the “intrinsic reaction coordinate” is decisively helpful in a quantitative, nonempirical understanding of the characteristic ...
  38. [38]
    The role of potential energy surface in quantum mechanical tunneling
    Oct 1, 2020 · Tunneling is more feasible in thermo-neutral symmetric potential energy surface. •. Narrow PES is more effective for tunneling than broad ...
  39. [39]
    An open quantum systems approach to proton tunnelling in DNA
    May 5, 2022 · We determine that the quantum tunnelling contribution to the proton transfer rate is several orders of magnitude larger than the classical over-the-barrier ...
  40. [40]
    (PDF) The set of valley ridge inflection points on the potential energy ...
    Aug 10, 2025 · The MP2/6-31G** potential energy surface (PES) of the water molecule is used as a model to locate curves of valley ridge inflection (VRI) ...Missing: seams scattering
  41. [41]
    Bifurcations on Potential Energy Surfaces of Organic Reactions - PMC
    The intrinsic reaction coordinate (IRC) is the most common quantum mechanical reaction pathway ... minimum energy pathway (MEP). However, when the gradient ...
  42. [42]
    Quadratic steepest descent on potential energy surfaces. I. Basic ...
    Oct 1, 1993 · A novel second‐order algorithm is formulated for determining steepest‐descent lines on potential energy surfaces.
  43. [43]
    [PDF] Introduction to potential energy surfaces
    Sep 27, 2021 · Computing the PES for a collinear reaction (continued). The PES is often more easily visualized as a contour plot: 0.6 0.8. 1. 1.2 1.4 1.6 1.8.
  44. [44]
    [PDF] Profile Contour Plots: Alternative Projections of 3D Free Energy ...
    We present profile contour plots as a technique for representing the. 3D energy surfaces encountered in computational chemistry. Pro- file contour plots are ...
  45. [45]
    Electron pair density in the lowest 1Σu+ and 1Σg+ states of H2
    Aug 19, 2011 · The prototype cases of the lowest 1 and 1 states of H2, which exhibit large variation in types of bonding, ranging from strongly ionic to ...
  46. [46]
    [PDF] PotentialEnergy Curves for the X 1Σg+, b3Σu+, and C 1Πu States of ...
    In the present work we have extended to large inter- nuclear distances the previous calculation of the ground- state energy of H2 using a generalized James-and-.
  47. [47]
    The intersection of potential energy surfaces in polyatomic molecules
    (1999) Location of conical intersections by the Pancharatnam connection and the sign-change theorem of Longuet-Higgins: a model calculation with the H4 ...
  48. [48]
    On the Role of Conical Intersections of Two Potential Energy ...
    On the Role of Conical Intersections of Two Potential Energy Surfaces of the Same Symmetry in Photodissociation. ... The Journal of Chemical Physics 2012, 137 (9) ...Missing: seminal | Show results with:seminal
  49. [49]
    High-dimensional ab initio potential energy surfaces for reaction ...
    This article discusses potential energy surfaces (PESs) for reaction dynamics, fits to ab initio energies, and examples of bimolecular and unimolecular ...Missing: astrochemistry | Show results with:astrochemistry
  50. [50]
    [PDF] coupling of classical and quasiclassical trajectories with electronic ...
    In classical and quasiclassical trajectory chemical dynamics simulations, the atom- istic dynamics of collisions, chemical reactions, and energy transfer ...
  51. [51]
    Potential-energy surfaces and transition-state theory
    Studies include: using zero-point energies to correct the classical activation energy; comparing the calculated quantities with experimental values; determining ...
  52. [52]
    Quantum instanton evaluation of the thermal rate constants and ...
    Mar 20, 2007 · In the present paper, we will demonstrate that the QI theory, in the classical limit, actually tends to the classical variational TST (VTST).
  53. [53]
    Spiers Memorial Lecture: New directions in molecular scattering
    May 1, 2024 · Ever-improving scattering experiments and exact quantum-dynamics calculations based on highly accurate potential energy surfaces have driven our ...
  54. [54]
    Theoretical study of stereodynamics for the reactions Cl+H2/HD/D2
    Mar 8, 2003 · Studies on the dynamical stereochemistry of the Cl + H 2 reaction and its isotopic variants, especially the isotope effect on the product ...
  55. [55]
    [2409.07211] Ab initio Calculations for Astrochemistry - arXiv
    Sep 11, 2024 · Then, the role of quantum chemistry in the investigation of interstellar-relevant potential energy surfaces will be described, focusing on ...
  56. [56]
    [PDF] well Potential Energy Surfaces in Combustion (and Atmospheric ...
    This review describes recent experimental and theoretical advances that deliver increasingly complete views of their reaction mechanisms. New methods for ...Missing: astrochemistry | Show results with:astrochemistry
  57. [57]
    Vibrational spectroscopy by means of first‐principles molecular ...
    Mar 1, 2022 · ... potential energy surface (PES) at finite ... spectrum of bulk liquid water revealed by intermolecular instantaneous normal mode analysis.
  58. [58]
    Modeling Anharmonic Effects in the Vibrational Spectra of High ...
    Apr 24, 2023 · ... potential energy surface, the selection ... Efficient anharmonic vibrational spectroscopy for large molecules using local-mode coordinates.
  59. [59]
    Franck-Condon factors by counting perfect matchings of graphs with ...
    ... Franck-Condon factors. A short comment on notation: we use boldface letters ... potential energy surface and their corresponding eigenfunction for the ...
  60. [60]
    Franck–Condon Factors in Electronic Excitations from the Ground ...
    May 11, 2023 · FCFs are used to explain vibronic transitions of molecules observed in the UV–vis and photoelectron spectra of diatomic and polyatomic molecules.
  61. [61]
    Potential energy surface crossings in organic photochemistry
    Potential energy surface crossings in organic photochemistry. F. Bernardi, M. Olivucci and M. A. Robb, Chem. Soc. Rev., 1996, 25, 321 DOI: 10.1039/CS9962500321.Missing: femtosecond dynamics<|control11|><|separator|>
  62. [62]
    [PDF] Review Article Exploring Multiple Potential Energy Surfaces
    On excited state PESs, molecules frequently take unexpected geometries that are very different from those of stable ground state molecules. Hence, the advantage ...
  63. [63]
    Excited-state methods for molecular systems: Performance, pitfalls ...
    May 1, 2025 · Proper theoretical descriptions of ground and excited states are critical for understanding molecular photophysics and photochemistry.
  64. [64]
    [PDF] Understanding protein folding with energy landscape theory Part II
    ... energy landscape theory of protein folding is a statistical description of a protein's complex potential energy surface, where individual folding events are ...
  65. [65]
    (PDF) Global potential energy surfaces for the H 3 system. Analytical ...
    Aug 6, 2025 · Figure 1 shows the HD-H + VLABP PES as a function of the Jacobi coordinates R and θ for a fixed internuclear r HD distance of 1.442 bohr, ...
  66. [66]
    [PDF] A new tunneling path for reactions such as H+H2→H2+H - SciSpace
    FIG. 1, Plot of potential energy contours for the H + H2 - H2. + H reaction using the Porter-Karplus surface. ... reaction, for the Porter-Karplus potential ...<|separator|>
  67. [67]
    High vibrational excitation of the reagent transforms the late-barrier ...
    Jan 24, 2024 · It is found that with the HOD vibration excitation increasing from vOH = 1 to 4, the product H2 becomes increasingly vibrationally excited and ...
  68. [68]
    A new H 3 potential energy surface and its implications for chemical ...
    A new H3 potential energy surface and its implications for chemical reaction ... Muckerman, Quantum Chemistry Program Exchange, #229, Indiana University ...
  69. [69]
    [PDF] SCATTERING RESONANCES IN THE SIMPLEST CHEMICAL ...
    the first exact quantum-mechanical H + H2 collinear computations. They observed that the total reactive and nonreactive cross sections displayed pronounced ...<|separator|>
  70. [70]
    Accurate Global Potential Energy Surfaces for the H + CH 3 OH ...
    Jun 12, 2020 · Here we develop global analytical PESs for this system using the permutation-invariant polynomial neural network (PIP-NN) and the high-dimensional neural ...Missing: astrochemistry | Show results with:astrochemistry
  71. [71]
    The H+H2 reactive system. Progress in the study of the dynamics of ...
    Aug 6, 2025 · The reactions studied are H + H2, F + H2, Cl + H2 and O(1D) + H2, for which reliable potential-energy surfaces (PES) are available. The QCT ...
  72. [72]
    Potential energy surfaces for polyatomic reaction dynamics
    Improvement of Fourteen Coupled Global Potential Energy Surfaces of 3A′ States of O + O2. The Journal of Physical Chemistry A 2025, 129 (13) , 3166-3175 ...
  73. [73]
    Methane dissociation on Ni(111): A fifteen-dimensional potential ...
    Oct 8, 2015 · In the present work, we develop a highly accurate, fifteen-dimensional potential energy surface (PES) of CH4 interacting on a rigid flat ...
  74. [74]
    Energy landscapes for proteins described by the UNRES coarse ...
    The potential energy landscape encodes the information required to properly fold from the sequence and contains all possible conformations, which can be ranked ...
  75. [75]
    [PDF] Reaction path Hamiltonian and the unified reaction valley approach
    This reduction of dimensionality permits the solution of the classical Liouville equation for the evaluation of the flux required in a rate constant eval ...
  76. [76]
    Generalized approximation to the reaction path: The formic acid ...
    Feb 27, 2008 · THE REACTION SPACE HAMILTONIAN IN REDUCED DIMENSIONALITY. Having established that { s 1 , s 2 , s 3 } = { d 1 , d 2 , d 3 } we proceed with the ...
  77. [77]
    Constructing Potential Energy Surfaces for Polyatomic Systems ...
    Oct 9, 2011 · Different methods of constructing potential energy surfaces in polyatomic systems are reviewed, with the emphasis put on fitting, interpolation, and analytical
  78. [78]
    Revealing a double-inversion mechanism for the F - Nature
    Jan 19, 2015 · Here, we report an accurate global analytic potential energy surface (PES) for the F−+CH3Cl SN2 reaction, which describes both the back-side and ...
  79. [79]
    and UHF–AM1 potential-energy surfaces for the diels–alder reaction ...
    The potential-energy surfaces for the Diels–Alder reactions between substituted butadienes and ethenes have been investigated by intensive RHF- and UHF-grid ...
  80. [80]
    Improving the Efficiency of Variationally Enhanced Sampling with ...
    Metadynamics is an enhanced sampling method designed to flatten free energy surfaces uniformly. However, the highest-energy regions are often irrelevant to ...
  81. [81]
    Lifting the Curse of Dimensionality on Enhanced Sampling of ...
    Apr 2, 2018 · A common challenge to applying metadynamics to the study of complex systems is selecting the proper collective variables to bias.<|separator|>
  82. [82]
    Large-scale QM/MM free energy simulations of enzyme catalysis ...
    Hybrid quantum mechanical–molecular mechanical (QM/MM) simulations provide key insights into enzyme structure–function relationships.
  83. [83]
    Unraveling the energetic significance of chemical events in enzyme ...
    Oct 8, 2020 · Hybrid quantum mechanics/molecular mechanics (QM/MM) and molecular dynamics (MD) studies have validated the acylation mechanism and provided ...
  84. [84]
    London, Fritz | Encyclopedia.com
    Fritz London and his younger brother, Heinz, were the only children of Franz ... In 1928 he formulated a description of chemical reactions as ...
  85. [85]
    Theoretical Chemistry in the 20th Century | Royal Society
    Feb 24, 2023 · A potential energy surface (PES) is the electronic energy expressed as a function of the nuclear coordinates in a molecule. It arises ...
  86. [86]
    [PDF] 1 Comparison of quantum mechanical and empirical potential ...
    Possibly the earliest example of a PES dates to 1931 with the empirical potential of Eyring and Polanyi[11] for the interaction of three hydrogen atoms, which ...
  87. [87]
    [PDF] CLASSICS
    Nov 11, 2013 · Polanyi's classic paper, co-authored by Henry Eyring, reproduced in this issue, was truly a classic, mapping out the potential energy ...
  88. [88]
    Three Millennia of Atoms and Molecules - ACS Publications
    Trajectories on the potential energy surface of H3 were calculated by the American chemical physicists Joseph Hirschfelder (1911-1990), Henry Eyring and Bryan.
  89. [89]
    [PDF] Potential Energy Surfaces and Elastic and Inelastic Scattering," DG ...
    They used valence-bond theory with the two configurations mentioned above; for linear symmetric H, one may consider instead one symmetry-adapted configuration, ...
  90. [90]
    Dynamics Calculations Based on Ab Initio Potential Energy Surfaces
    One of the most exciting developments in theoretical chemistry in the last few years has been the production of ab initio potential energy surfaces by ...
  91. [91]
    Publications of Donald G. Truhlar | The Journal of Physical ...
    “The Representation and Use of Potential Energy Surfaces in the Wide Vicinity of a Reaction Path for Dynamics Calculations on Polyatomic Reactions,” D. G. ...
  92. [92]
    Coupled-cluster theory in quantum chemistry | Rev. Mod. Phys.
    Feb 22, 2007 · Today, coupled-cluster theory offers the most accurate results among the practical ab initio electronic-structure theories applicable to ...
  93. [93]
    Analytic derivatives for perturbatively corrected “double hybrid ...
    Mar 30, 2007 · In this work, the applicability of the new functionals is extended to the exploration of potential energy surfaces with analytic gradients. The ...
  94. [94]
    Benchmarking Adaptive Variational Quantum Eigensolvers - Frontiers
    Here we benchmark the accuracy of VQE and ADAPT-VQE to calculate the electronic ground states and potential energy curves for a few selected diatomic molecules.<|separator|>
  95. [95]
    Nonadiabatic Derivative Couplings Calculated Using Information of ...
    May 21, 2023 · Utilizing this algorithm, one can calculate NACMEs using only the information of potential energy surfaces (PESs), i.e., energies, and gradients ...Missing: PyUNIx | Show results with:PyUNIx
  96. [96]
    Modeling Potential Energy Surface by Force Fields for ...
    The concept of the potential energy surface (PES) in computational simulations is essential for studying material properties and heterogeneous catalytic ...
  97. [97]
    Cross-disciplinary perspectives on the potential for artificial ...
    Apr 25, 2025 · We present ten diverse perspectives on the impact of AI coming from those with a range of backgrounds from experimental chemistry, computational chemistry, ...