Fact-checked by Grok 2 weeks ago

Excited state

In , an is any of an , , or other system that possesses higher energy than the , which is the lowest possible energy configuration. These states arise when one or more electrons are promoted from occupied orbitals in the to higher-energy, unoccupied orbitals, often through absorption of photons or collisions with other particles. Excited states are inherently unstable and short-lived, typically decaying back to lower-energy states via radiative processes like or , or non-radiative pathways such as . In atomic physics, they correspond to discrete energy levels above the , as described by models like Bohr's, where transitions emit or absorb at specific wavelengths unique to each , enabling spectroscopic identification. For molecules, excited states are classified by types such as π → π** or n → π** transitions, influencing photochemical reactions, energy transfer, and phenomena like or technology. The study of these states is fundamental to understanding electronic spectra, with ultraviolet-visible absorption typically occurring in the 1-12 eV range due to the Franck-Condon principle governing vertical excitations.

Fundamentals

Definition

In , an is a of a —such as an , , or —that possesses higher than its lowest-energy configuration, typically resulting from the promotion of an to a higher orbital or the excitation of other like phonons or nuclear particles. This elevated configuration contrasts with the , which serves as the reference point of zero excess for the . The concept of excited states was introduced by in his model of the , where he proposed that electrons occupy discrete orbits with quantized energy levels, and transitions between these orbits correspond to the absorption or emission of photons. In this model, the represents the lowest orbit, while higher orbits constitute excited states, laying the groundwork for understanding atomic spectra and quantum jumps. Energy level diagrams for illustrate as discrete lines above the in bound systems, such as atoms or molecules confined by potential wells, where the yields quantized solutions. In contrast, unbound or ionized systems exhibit a of levels beyond the threshold, allowing for states without discrete quantization. Excited states can be classified by the type of motion involved, including excited states arising from promotion between molecular orbitals, vibrational excited states due to oscillations of atomic nuclei around equilibrium positions, and rotational excited states from molecular tumbling. excited states are often the primary focus in and , as they involve significantly larger gaps (typically in the visible or range) compared to the finer vibrational and rotational splittings within each level.

Ground vs. Excited States

In , the of an or represents the lowest possible energy configuration, where electrons occupy the orbitals that minimize the total energy of the system. This state is typically described by solutions to the time-independent , which yields stationary wavefunctions corresponding to definite energy eigenvalues. The is inherently stable, as any tends to return the system to this equilibrium due to the absence of lower-energy alternatives. In contrast, excited states occur when one or more electrons are promoted to higher-energy orbitals, resulting in a metastable with elevated . These states are unstable because the excess energy makes them prone to relaxation back toward the through various dissipative processes, though they can persist briefly under certain conditions. For atomic systems like , excited states correspond to higher principal quantum numbers (n > 1), such as the first excited state at n=2, which lies above the (n=1) in the energy hierarchy. The energy separation between and defines the of , with typical transitions in the visible and regions spanning 1-10 . For instance, visible light excitations range from approximately 1.6 (red) to 3.2 (violet), while UV transitions often extend to higher values up to 12 . This energy gap underscores the transient nature of excited states, as the higher-energy electrons seek to lower their potential by returning to the more stable configuration.

Excitation Mechanisms

Atomic Excitation

In isolated , excitation occurs when an is promoted from a lower to a higher , typically through interactions with external energy sources. The primary mechanisms include , collisional excitation by electrons or ions, and radiative recombination. Photon absorption is the most direct mechanism, where an in its absorbs a whose precisely matches the difference between the ground and an excited state. This process follows the E = h\nu, where [E](/page/E!) is the difference between levels, [h](/page/H+) is Planck's constant, and \nu is the of the absorbed . The resulting spectral lines are known as lines, corresponding to transitions between specific levels. Collisional excitation arises from inelastic collisions between the atom and energetic electrons or ions, where kinetic energy is transferred to an atomic , promoting it to a higher orbital without . This mechanism is prevalent in plasmas and gaseous discharges, where the collision energy exceeds the threshold but remains below the ionization potential. Radiative recombination contributes to excitation by involving the capture of a by an , forming a neutral atom while emitting a ; the excess energy often results in the atom occupying an excited state rather than the directly. In this process, the emitted 's energy is less than the , leaving the recombined in a bound excited , from which it may further . These excitation processes are governed by selection rules that determine allowed transitions, primarily for electric interactions, which dominate in atomic spectra. The key rule is \Delta l = \pm 1, where l is the orbital quantum number of the transitioning , ensuring of and parity change. Additional constraints include \Delta J = 0, \pm 1 (with J = 0 \leftrightarrow J = 0 forbidden) and \Delta S = 0 for LS coupling, where J is the total and S is the . In the hydrogen atom, these rules manifest in spectral series such as the Lyman and Balmer series, where excitations to higher n levels followed by de-excitation produce characteristic lines. The Lyman series involves transitions to or from the ground state (n=1), emitting ultraviolet photons for excitations from n=2 to higher levels dropping back to n=1. The Balmer series corresponds to transitions to or from the first excited state (n=2), producing visible light, such as the red H-alpha line at 656 nm from n=3 to n=2. Both series obey the \Delta l = \pm 1 rule, with s-to-p or p-to-s transitions being prominent. A practical example is the sodium D-line excitation, where sodium atoms absorb near 589 to promote an from the 3s to the 3p excited state, adhering to the \Delta l = +1 . The D-line doublet at 588.995 and 589.592 arises from splitting in the 3p level. This excitation is exploited in low-pressure sodium vapor lamps for streetlighting, where electrical discharge excites sodium atoms, leading to efficient yellow emission dominated by the D-lines for high-lumen output.

Molecular Excitation

In molecular systems, photoexcitation primarily occurs through electronic transitions involving promotion of electrons from occupied to unoccupied molecular orbitals, with π→π* and n→π* transitions being dominant in molecules. The π→π* transitions, common in conjugated systems like alkenes and aromatics, involve excitation from π orbitals to π* antibonding orbitals and typically exhibit strong intensities due to favorable overlap. In contrast, n→π* transitions, observed in molecules with heteroatoms such as carbonyls, promote non-bonding (n) electrons to π* orbitals and are generally weaker and occur at longer wavelengths because the n orbital is higher in energy than π orbitals. These transitions lead to excited states, from which can relax the molecule vibrationally within the same electronic state, while enables spin-forbidden transitions to triplet states, facilitating longer-lived excitations./09%3A_Separation_Purification_and_Identification_of_Organic_Compounds/9.10%3A_Electronic_Spectra_of_Organic_Molecules)/Spectroscopy/Electronic_Spectroscopy/Jablonski_diagram) The Franck-Condon principle governs the nature of these excitations in polyatomic molecules, dictating that transitions are vertical on potential energy surfaces, meaning the geometry remains fixed during the ultrafast promotion due to the disparity in timescales between and motions. This results in vibronic coupling, where the excited-state potential minimum differs from the , leading to initial population of vibrationally hot states in the excited manifold. Such vertical transitions explain the broad absorption bands observed in molecular spectra, as the overlap of vibrational wavefunctions determines transition probabilities. The dynamics of molecular excitation are often illustrated using the , which depicts the ground state (S₀) and excited states such as the first (S₁) and triplet (T₁), connected by radiative and non-radiative processes. promotes the molecule from S₀ to S₁ or higher , followed by rapid vibrational relaxation within S₁; from there, can return to S₀, or to T₁ enables or other triplet-mediated pathways. This framework highlights the interplay of electronic and vibrational unique to molecules, contrasting with the simpler excitation analog./Spectroscopy/Electronic_Spectroscopy/Jablonski_diagram) Representative examples include , where UV photo at approximately 255 nm corresponds to a π→π* to S₁, initiating vibronic progression governed by Franck-Condon factors. In dyes like fluorescein, around 495 nm populates the S₁ state via a similar π→π* mechanism, leading to efficient emission at 517 nm after vibrational relaxation, underscoring applications in molecular probes./15%3A_Benzene_and_Aromaticity/15.07%3A_Spectroscopy_of_Aromatic_Compounds)

Excitation in Perturbed Systems

In perturbed systems, such as gases under or in the presence of external fields, the excitation of atoms and molecules deviates from ideal isolated conditions due to interactions that alter levels and probabilities. Collisions between excited atoms and surrounding particles lead to broadening of spectral lines, where the finite duration of radiative processes is interrupted, resulting in a line profile characterized by wings that decay as the inverse square of the frequency offset from the line center. This broadening arises from the phase shifts induced by collisions, with the (FWHM) proportional to the collision rate, typically scaling linearly with in dilute gases. Additionally, these collisions can cause a symmetric shift in the line position, though the effect is often smaller than the broadening./10%3A_Line_Profiles/10.05%3A_Pressure_Broadening) External electric fields induce the Stark effect, splitting degenerate excited levels and shifting their energies through the interaction of the atomic dipole moment with the field. For systems with permanent dipole moments, such as certain molecular excited states or Rydberg atoms, the first-order energy shift is given by \Delta E = -\vec{\mu} \cdot \vec{E}, where \vec{\mu} is the electric dipole moment and \vec{E} is the electric field vector; this linear Stark effect lifts degeneracies and can enhance or suppress excitation rates by modifying selection rules. In atomic hydrogen, for instance, the n=2 excited manifold splits into Stark sublevels, allowing tunable excitation via field-dependent transitions. Similarly, magnetic fields produce the Zeeman effect, where the magnetic dipole interaction splits excited levels according to \Delta E = g \mu_B m_j B, with g the Landé g-factor, \mu_B the Bohr magneton, m_j the magnetic quantum number, and B the field strength; this splitting is particularly pronounced in excited states with nonzero orbital angular momentum, facilitating field-controlled excitation in atomic vapors./06%3A_Perturbative_Approaches/6.02%3A_The_linear_Stark_Effect)/11%3A_Time-Independent_Perturbation_Theory/11.09%3A_Zeeman_Effect) In noble gas discharges, such as those used in helium-neon (He-Ne) lasers, collisional perturbations play a crucial role in achieving excitation suitable for lasing. Electrical discharge excites helium atoms to metastable states, which then transfer energy to neon via resonant collisions, selectively populating upper laser levels in neon while lower levels are depopulated, enabling population inversion between the $3s_2 and $2p_4 states of neon at 632.8 nm. This process relies on the perturbed environment of the discharge plasma, where pressure broadening merges lines and collisions maintain the inversion against radiative decay. However, such systems are also susceptible to quenching, where excited state populations are reduced through non-radiative collisional deactivation with ground-state atoms or molecules, often following Stern-Volmer kinetics with rate constants on the order of $10^{-10} cm³/s for typical quenchers like N₂ or O₂. Quenching competes with desired excitation pathways, limiting the efficiency in high-pressure or impure gases.

Properties and Dynamics

Lifetimes and Decay Processes

The lifetime of an excited state refers to the average time it persists before decaying back to the , governed by competing radiative and non-radiative processes. Radiative occurs through , where the excited state relaxes by emitting a , with the rate given by \frac{1}{\tau} = A, where A is the Einstein for . This underpins in excited states, typically with on the order of $10^{-9} seconds, and in triplet states, which exhibit much longer around $10^{-3} seconds due to spin-forbidden transitions. Non-radiative decay pathways include internal conversion, where excess energy is transferred to vibrational modes within the same spin multiplicity, leading to relaxation without photon emission; vibrational relaxation, which rapidly dissipates energy through molecular vibrations to the lowest vibrational level of the excited state; and intersystem crossing, a spin-forbidden transition between states of different multiplicity, such as from singlet to triplet. The overall excited-state lifetime \tau is determined by the sum of radiative (k_r) and non-radiative (k_{nr}) rates: \tau = \frac{1}{k_r + k_{nr}}. Factors such as spin-orbit coupling, enhanced by the presence of heavy atoms, significantly influence lifetimes by promoting rates, thereby shortening lifetimes and extending triplet persistence.

Excited-State Absorption

Excited-state (ESA) is a nonlinear optical process in which an atom or , already in an electronically , absorbs an additional to to a higher-lying . This typically involves a promotion from the lowest , denoted S₁, to higher states Sₙ (n > 1), occurring under conditions of high light intensity that significantly populate the initial through prior . Unlike direct multi-photon , ESA proceeds sequentially, with the first exciting the system from the S₀ to S₁, followed by a second inducing the upward , thereby enabling effective two-photon processes in systems with appropriate structures. The dynamics of ESA are governed by the absorption rate from the initial excited , which can be expressed as the excitation rate to the higher state: R = \frac{\sigma_{es} I}{h\nu} N_1, where N_1 is the in the initial excited (e.g., S₁), I is the (), \sigma_{es} is the excited-state cross-section, h is Planck's , and \nu is the . This contributes to \frac{dN_n}{dt} = R - k_{\rm [decay](/page/Decay)} N_n for the higher-state N_n (e.g., Sₙ) and depletes the initial via \frac{dN_1}{dt} = \dots - R. The equation highlights the intensity-dependent nature of the process, with the rate proportional to both the excited-state and the incident . In applications, ESA plays a critical role in dyes and saturable absorbers, where the excited-state cross-section \sigma_{es} is typically much smaller than the ground-state cross-section \sigma_{gs}, allowing efficient saturation of without significant residual losses from higher states. For instance, in lasers, low \sigma_{es} values minimize unwanted in the excited , enhancing output . Conversely, in optical limiting devices, tailored ESA properties enable reverse saturable , where increased at high intensities protects optical sensors, though optimal performance requires balancing \sigma_{es} and \sigma_{gs} for the specific . A representative example of ESA is observed in rhodamine dyes, such as , widely employed as saturable absorbers in mode-locked lasers due to their favorable nonlinear response. In these dyes, the ESA cross-section at relevant wavelengths (e.g., around 1055 nm) is approximately 2 \times 10^{-17} cm², significantly lower than the ground-state value of about 4.5 \times 10^{-17} cm² near the S₀ to S₁ , enabling effective while limiting higher-order losses. Similar is seen in , with \sigma_{es} \approx 4 \times 10^{-18} cm², underscoring the dyes' in ultrafast .

Reactivity and Applications

Photochemical Reactions

Excited states in molecules often exhibit enhanced reactivity compared to their ground states due to altered potential energy surfaces that lower activation barriers for chemical transformations. In photochemical reactions, the promotion of an electron to an excited state can lead to bond weakening or distortion, facilitating processes that are forbidden or slow in the ground state. A key feature enabling this reactivity is the presence of conical intersections, where the potential energy surfaces of the ground and excited states touch, allowing ultrafast nonradiative transitions and efficient channeling of energy into chemical pathways. This mechanism is central to many photochemical processes, as it permits rapid relaxation while promoting selective bond breaking or forming. Common types of photochemical reactions driven by excited states include and . In , absorption of light induces a geometric rearrangement, such as the cis-trans isomerization of in the visual protein , where the 11-cis- converts to all-trans- upon excitation, triggering the vision signaling cascade with near-unity . involves the cleavage of bonds to produce fragments, exemplified by the atmospheric photodissociation of O₂ in the Schumann-Runge bands (175-205 nm), where O₂ + hν → 2O generates oxygen atoms essential for formation. These reactions highlight how excited states enable specific transformations by accessing repulsive or distorted geometries unavailable in the . The efficiency of photochemical reactions is quantified by the , Φ, defined as the number of product molecules formed per absorbed: \Phi = \frac{\text{number of product molecules}}{\text{number of photons absorbed}} Values of Φ are often less than 1 due to competing decay processes, such as , , or , which divert the excited-state energy away from productive chemical channels. For instance, in solution-phase reactions, vibrational relaxation or interactions can reduce Φ, emphasizing the role of environmental factors in determining reaction outcomes. In biological systems, excited-state photochemistry underpins critical processes like photosynthesis and DNA damage. In photosynthesis, excitation of chlorophyll a in photosystem II leads to rapid charge separation, where the excited singlet state transfers an electron to a pheophytin acceptor within picoseconds, initiating the electron transport chain that drives ATP synthesis and carbon fixation. Conversely, ultraviolet excitation of DNA bases, particularly thymine, promotes [2+2] cycloaddition to form cyclobutane pyrimidine dimers, a primary lesion in UV-induced mutagenesis and skin cancer. These examples illustrate the dual role of excited states in enabling life-sustaining reactions while posing risks from uncontrolled photochemistry.

Spectroscopic Techniques

Absorption spectroscopy in the ultraviolet-visible (UV-Vis) range is a fundamental technique for characterizing excited states by measuring the wavelengths at which molecules absorb light to promote electrons from ground to excited electronic states. Typically spanning 200-800 nm, UV-Vis absorption reveals excitation spectra that correspond to electronic transitions such as π-π* in conjugated systems or n-π* in carbonyl compounds, with absorption maxima shifting to longer wavelengths as conjugation increases, as seen in polyenes like absorbing in the visible region. The Beer-Lambert law quantifies these absorptions, where absorbance A = \epsilon l c (with \epsilon as absorptivity, l as path length, and c as concentration), enabling determination of excited-state energies and oscillator strengths for transitions like those in at around 165 nm. Emission spectroscopy, particularly , complements by monitoring the decay of excited states back to the , providing insights into radiative lifetimes and quantum yields. In , a excited to a (S₁) emits light upon returning to the (S₀), with the process following an described by I(t) = I_0 e^{-t/[\tau](/page/Tau)}, where \tau is the fluorescence lifetime, typically on the order of nanoseconds for dyes. This technique is widely used to track excited-state dynamics, such as in biomolecules where Stokes shifts arise from vibrational relaxation in S₁ before . Time-resolved methods like pump-probe spectroscopy enable the observation of ultrafast excited-state dynamics on timescales, crucial for capturing processes such as and evolution. In this approach, a pump excites the sample to an electronic state, followed by a delayed probe that monitors changes in absorption or photoelectron emission, achieving resolutions down to 50 fs for nuclear motions in molecules like . Broadband implementations, often using transient absorption, reveal -scale relaxations, as in the 20 fs dynamics of tetramers in photosynthetic complexes. Phosphorescence spectroscopy probes longer-lived triplet states (T₁), where emission occurs via spin-forbidden transitions from T₁ to S₀, resulting in lifetimes from milliseconds to seconds and red-shifted spectra relative to . This requires from S₁ to T₁, typically enhanced in rigid or low-temperature environments to minimize non-radiative decay, as observed in aromatic hydrocarbons like . Delayed , arising from mechanisms like or thermal activation from T₁ back to S₁, provides additional characterization of triplet populations, with emission delays on the scale in systems such as organic phosphors. Laser-induced fluorescence (LIF) exemplifies these techniques in applied contexts, such as , where a excites trace like NO₂ to fluorescent excited states, and the ensuing is detected to quantify concentrations with high . In field campaigns like PARADE 2011, LIF measured NO₂ at parts-per-billion levels by tuning to specific vibronic transitions around 450 nm, enabling real-time monitoring of radical dynamics without interference from other .

Computational Methods

Approaches to Excited-State Calculations

Time-dependent methods provide a framework for calculating vertical excitation energies, which correspond to electronic transitions without nuclear rearrangement, by solving the time-dependent in a linear response approximation. Time-dependent Hartree-Fock (TD-HF) theory, introduced as an extension of the Hartree-Fock method to time-dependent perturbations, computes excited states through (RPA)-like equations that account for -hole excitations. This approach yields excitation energies and transition properties but often overestimates them due to neglect of electron correlation. Time-dependent density functional theory (TD-DFT), building on the TD-HF formalism but using Kohn-Sham orbitals, offers a more efficient alternative by incorporating exchange-correlation effects via approximate functionals; the linear response formulation, as developed by Casida, diagonalizes a to obtain excitation energies ω and oscillator strengths. The f, which quantifies the intensity of a , is given by f = \frac{2 m_e \omega}{3 \hbar e^2} |\boldsymbol{\mu}|^2 where m_e is the , ω is the excitation frequency, ħ is the reduced Planck's , e is the , and μ is the between ground and excited states. In TD-DFT, these quantities are derived from response functions, enabling predictions of spectra for medium-sized molecules. Configuration interaction methods construct excited-state wavefunctions as linear combinations of determinants from the ground-state reference. The configuration interaction singles (CIS) approach, a simple yet size-consistent method, targets singly excited configurations relative to the Hartree-Fock ground state, providing a balanced treatment of electron correlation for vertical excitations and allowing geometry optimization of excited states. For higher accuracy, equation-of-motion coupled cluster (EOM-CC) methods, particularly EOM-CCSD, extend coupled cluster theory by applying excitation operators to the correlated ground state, yielding well-behaved wavefunctions and excitation energies that approach chemical accuracy for single-reference systems. Multireference approaches are essential for excited states involving significant static correlation or open-shell character, where single-reference methods fail. The complete active space self-consistent field (CASSCF) method optimizes a multiconfigurational wavefunction by distributing active electrons fully among active orbitals, capturing near-degeneracies and providing qualitatively correct descriptions of surfaces for such states. As an illustrative application, TD-DFT has been employed to model the absorption spectrum of the green fluorescent protein (GFP) chromophore, where calculations on the anionic form predict a bright π→π* transition at 2.59 eV, aligning with the observed ~480 nm (2.58 eV) absorption maximum and highlighting the method's utility in biological systems.

Limitations and Advances

Time-dependent density functional theory (TD-DFT), while widely used for excited-state calculations, systematically underestimates excitation energies for charge-transfer (CT) states due to the self-interaction error and the lack of long-range correlation in standard approximate functionals. This limitation arises particularly in systems where the electron density shifts significantly between donor and acceptor sites, leading to errors exceeding 1 eV in many organic molecules and dye-sensitized materials. Similarly, single-reference methods such as coupled-cluster singles and doubles (CCSD) or standard TD-DFT struggle with near-degeneracies, where multiple electronic configurations contribute comparably, resulting in unphysical orbital occupations or convergence failures in strongly correlated systems like transition metal complexes. These issues often necessitate multireference approaches, such as complete active space self-consistent field (CASSCF), to capture static correlation effects accurately. Incorporating solvent effects in excited-state computations presents further challenges, as implicit models like polarizable continuum model (PCM) treat the solvent as a dielectric continuum, which oversimplifies local interactions and hydrogen bonding in polar environments, leading to typical errors of 0.2–0.3 eV in excitation energies. Explicit solvation models, by contrast, include discrete solvent molecules via or hybrids, better capturing dynamical solvent reorganization during electronic transitions but at a high computational cost that limits system sizes to hundreds of atoms. Hybrid approaches combining PCM with explicit solvation layers have shown improved accuracy for excitation energies in aqueous solutions compared to pure implicit treatments, though they require careful parameterization for perturbed systems like biomolecules. Recent advances have addressed these gaps through many-body methods like combined with the Bethe-Salpeter equation (GW-BSE), which excels in by accounting for screened interactions and excitonic effects, yielding energies and optical spectra with errors below 0.2 for semiconductors such as and transition metal dichalcogenides. In molecular dynamics, (ML) potentials trained on data have enabled efficient simulations of non-adiabatic processes post-2020, with models like diabatic artificial neural networks (DANN) accelerating excited-state trajectories for photoswitches by orders of magnitude while maintaining chemical accuracy in and coupling predictions. Enhancements to the ΔSCF method, which relaxes orbitals for individual excited states, have improved stability and reliability through theoretical foundations linking it to ensemble , allowing robust calculations of core-excited states in large systems with reduced basis set sensitivity. Looking ahead, real-time TD-DFT (RT-TDDFT) emerges as a promising avenue for simulating non-adiabatic dynamics, propagating the time-dependent Kohn-Sham equations to capture ultrafast processes like conical intersections in photochemical reactions, with recent ML integrations reducing computational overhead for attosecond-scale events in polyatomic molecules.

References

  1. [1]
    The Bohr Model
    The states with successively more energy than the ground state are called the first excited state, the second excited state, and so on. Beyond an energy ...
  2. [2]
    Understanding the Atom - Imagine the Universe! - NASA
    When an electron temporarily occupies an energy state greater than its ground state, it is in an excited state. An electron can become excited if it is given ...
  3. [3]
    Spectroscopy of Electronically Excited States
    Most molecules have bound higher energy excited electronic states in addition to the ground electronic state E 0.
  4. [4]
    [PDF] Chapter 7. Quantum Theory and Atomic Structure
    (first) energy level, called the ground state. • The second energy level (second stationary state) and all higher levels are called excited states. •A ...
  5. [5]
    I. On the constitution of atoms and molecules - Taylor & Francis Online
    On the constitution of atoms and molecules. N. Bohr Dr. phil. Copenhagen. Pages 1-25 | Published online: 08 Apr 2009.
  6. [6]
    [PDF] Quantum Mechanics - Northern Illinois University
    Apr 2, 2025 · More generally, in quantum mechanics the energies of bound states turn out to be quantized (discrete). There are also unbound (ionized) states ...Missing: diagram | Show results with:diagram
  7. [7]
    Photochemistry - MSU chemistry
    A photochemical reaction occurs when internal conversion and relaxation of an excited state leads to a ground state isomer of the initial substrate molecule, or ...
  8. [8]
    [PDF] 1 The Schrödinger equation - MIT OpenCourseWare
    Sep 13, 2013 · This is the equation for ψ that makes Ψ(x, t) = e−iEt/ ψ(x) a stationary state of energy E. Any of the three boxed equations above is referred ...
  9. [9]
    Quantum Theory - FSU Chemistry & Biochemistry
    The lowest energy state (n=1) is called the ground state of the atom ... or higher) the atom is said to be in an excited state. •As n becomes larger ...
  10. [10]
    [PDF] Introduction to Excited Electronic States
    Ultraviolet / visible (UV/vis) spectra are dominated by electronic transitions. Electronic transitions typically occur in the 1-12 eV range.
  11. [11]
    6.27 Optical Applications - FAMU-FSU College of Engineering
    The visible range of light corresponds to photons with energies from about 1.6 eV (red) to 3.2 eV (violet). In terms of the wave length of the light, the range ...
  12. [12]
    Atomic Spectroscopy
    May 7, 2012 · The figure on the right shows a high energy photon with Ephoton = hν being absorbed, resulting in a 2s→3s electron excitation; similarly, a ...
  13. [13]
    Atomic Excitation - an overview | ScienceDirect Topics
    Collisions between charged particles and either free or bound electrons result in ionization or excitation of absorber atoms, whereas interactions with the ...
  14. [14]
    Atomic Spectros. - Spectral Lines - NIST
    Selection rules for discrete transitions. Electric dipole (E1) ("allowed"), Magnetic dipole (M1) ("forbidden"), Electric quadrupole (E2) ("forbidden") ...
  15. [15]
    Radiative and Dielectronic Recombination | NIST
    Feb 25, 2010 · Radiative recombination involves an ion capturing an electron and emitting a photon. Dielectronic recombination involves a bound electron being ...
  16. [16]
  17. [17]
    1.4: The Hydrogen Atomic Spectrum
    ### Summary of Lyman and Balmer Series for Hydrogen Atom
  18. [18]
    Hydrogen-Like Atoms:Sodium - HyperPhysics
    The sodium spectrum is dominated by the bright doublet known as the Sodium D-lines at 588.9950 and 589.5924 nanometers.Missing: excitation | Show results with:excitation
  19. [19]
    Sodium Vapor Lamps - RP Photonics
    They largely exploit the characteristic orange 589-nm D-line of sodium atoms, which has also been extensively studied in physics.
  20. [20]
    Franck-Condon Principle - an overview | ScienceDirect Topics
    According to the Franck–Condon principle, all electronic transitions are vertical (Figure 2) that is, they occur without change in the position of the nuclei.
  21. [21]
    Fluorescein - OMLC
    This page summarizes the optical absorption and emission data of Fluorescein that is available in the PhotochemCAD package, version 2.1a (Du 1998, Dixon 2005).
  22. [22]
    A Qualitative Excited State Dynamics Model from First-Principles
    Apr 2, 2025 · Same-spin transitions correspond to the internal conversion mechanism, while electron spin-flipping transitions correspond to the inter-system ...3. Theory · 3.1. Radiative Decay · 4.3. Radiative Decay
  23. [23]
    Interplay of Fluorescence and Phosphorescence in Organic ...
    Jun 21, 2017 · Essentially, the typical lifetimes of fluorescence (nanoseconds) and phosphorescence (milliseconds to seconds) differ greatly with direct ...
  24. [24]
    [PDF] Internal Conversion and Intersystem Crossing Dynamics ... - OSTI.GOV
    May 26, 2022 · Internal conversion (1a) and intersystem crossing (1b) are the two major mechanisms of nonradiative decay, with the former involving states with ...
  25. [25]
    [PDF] Fluorescent Signaling Based on Control of Excited State Dynamics ...
    (16) As τ ) 1/(kr + knr) and φ ) kr/(kr + knr), kr ) φ/τ. (17) Quantum yields were determined by standard methods with tryptophan as a reference. See ...
  26. [26]
    State-specific heavy-atom effect on intersystem crossing processes ...
    Jan 30, 2013 · On the basis of the computed potential energy profiles and spin-orbit couplings, we proposed three competitive, efficient nonadiabatic pathways ...
  27. [27]
    Excited-state Absorption – ESA - RP Photonics
    Here, the transition rate of an ion subject e.g. to some optical pump intensity is R E S A = σ I p / h ν p with the ESA transition cross-section , which ...What is Excited-state... · Strength of Excited-state... · ESA in Upconversion Lasers
  28. [28]
    Excited State Absorption - an overview | ScienceDirect Topics
    ESA is the mechanism in which two photons are successively absorbed by an electron affiliated with the multiple energy levels.
  29. [29]
    Optimization of optical limiting devices based on excited-state ...
    Molecules exhibiting excited-state absorption have received considerable attention as promising candidates for use in passive optical limiters. [7],[12],[13] ...
  30. [30]
    Excited state absorption: a key phenomenon for the improvement of ...
    Excited state absorption: a key phenomenon for the improvement of biphotonic based optical limiting at telecommunication wavelengths. Q. Bellier, N. S. Makarov, ...
  31. [31]
    [PDF] Excited-State Absorption Cross-Sections in Rhodamine Dyes ...
    Excited-state absorption spectroscopy with picosecond light pulses provides new information on energy levels and on the dynamics of transitions i1 *3 ). Intense ...Missing: mechanism | Show results with:mechanism
  32. [32]
    Conical intersections in molecular photochemistry – the role of ...
    In this paper, we present a simple method for determining the existence of a conical intersection for a given chemical reaction.
  33. [33]
    Photoisomerization Mechanism of Rhodopsin and 9-cis ... - NIH
    Ultrafast, highly efficient photoisomerization of the chromophore is an outstanding feature of retinal proteins such as bacteriorhodopsin and rhodopsin.Photoisomerization Mechanism... · Figure 3 · Figure 2Missing: seminal paper
  34. [34]
    A Correlated‐K Parameterization for O2 Photolysis in the Schumann ...
    May 9, 2024 · In today's atmosphere, the absorption of incoming ultraviolet radiation by O2 plays a decisive role in creating O atoms that can react to form ...
  35. [35]
    Quantum Yield - an overview | ScienceDirect Topics
    Quantum yield (ϕ) is defined as the number of moles of product formed per Einstein of light absorbed by a photosensitive agent in a photochemical reaction.
  36. [36]
    The initial charge separation step in oxygenic photosynthesis - Nature
    Apr 27, 2022 · In this study, we investigated the excited state dynamics of the PSII-RC via 2DEV spectroscopy. Both highly excitation frequency-dependent ...
  37. [37]
    UV‐induced DNA Damage: The Role of Electronic Excited States
    Oct 5, 2015 · Electronic interactions among DNA bases give rise to excited states delocalized over two or more bases. As a result, the excited state ...
  38. [38]
    20.3: Excited Electronic States: Electronic Spectroscopy of Molecules
    Feb 2, 2016 · When the energy from UV or visible light is absorbed by a molecule, one of its electrons jumps from a lower energy to a higher energy molecular orbital.
  39. [39]
    Principles and Theory of Fluorescence Spectroscopy - HORIBA
    Typically, the excited state decays in an exponential manner, as indicated in the equation below. The use of fluorescence lifetime has its advantages over that ...
  40. [40]
    Spectroscopic and Structural Probing of Excited-State Molecular ...
    Apr 22, 2020 · We report on a combination of structural (relativistic ultrafast electron diffraction, or UED) and spectroscopic (time-resolved photoelectron spectroscopy, or ...
  41. [41]
    Femtosecond Dynamics of Excited States of Chlorophyll Tetramer in ...
    It was measured by a broadband femtosecond laser pump-probe spectroscopy within the range from 400 to 750 nm and with a time resolution of 20 fs-20 …
  42. [42]
    Laser-induced fluorescence-based detection of atmospheric ... - AMT
    Mar 7, 2019 · Laser-induced fluorescence-based detection of atmospheric nitrogen dioxide and comparison of different techniques during the PARADE 2011 field ...
  43. [43]
    Time-Dependent Hartree---Fock Theory for Molecules
    Time-Dependent Hartree—Fock Theory for Molecules. A. D. McLACHLAN and M. A. ... M. A. Ball and A. D. McLachlan, Mol. Phys. (to be published). OutlineMissing: URL | Show results with:URL
  44. [44]
  45. [45]
    Toward a systematic molecular orbital theory for excited states
    Foresman · Martin Head-Gordon · John A. Pople · Michael J. Frisch. ACS Legacy Archive. Open PDF. The Journal of Physical Chemistry. Cite this: J. Phys. Chem.
  46. [46]
    The equation of motion coupled‐cluster method. A systematic ...
    A comprehensive overview of the equation of motion coupled‐cluster (EOM‐CC) method and its application to molecular systems is presented.
  47. [47]
    The complete active space SCF (CASSCF) method in a Newton ...
    Feb 15, 1981 · The CASSCF wave function is formed from a complete distribution of a number of active electrons in a set of active orbitals.Missing: URL | Show results with:URL
  48. [48]
    Quantum Chemical Benchmark Studies of the Electronic Properties ...
    This series of two papers focuses on accurate calculations of the properties of biological chromophores with ab initio methods using the model GFP chromophore, ...Figure 1 · 2 Computational Methods · Figure 3
  49. [49]
    Neutral excitation density-functional theory: an efficient and ... - Nature
    Jun 2, 2020 · However, LR-TDDFT has two significant limitations: its computational cost, which severely limits the size of the systems that it can be used to ...
  50. [50]
    Surface Hopping Nested Instances Training Set for Excited-state ...
    Jul 26, 2025 · These methods are essential for accurately capturing electronic state mixing and near-degeneracies, where single-reference approaches often ...Methods · Quantum Chemistry Reference... · Technical Validation
  51. [51]
    Improved prediction of solvation free energies by machine-learning ...
    Jun 18, 2021 · We develop and introduce the Machine-Learning Polarizable Continuum solvation Model (ML-PCM) for a substantial improvement of the predictability of solvation ...
  52. [52]
    pyGWBSE: a high throughput workflow package for GW-BSE ...
    Feb 13, 2023 · GW-BSE is a many body perturbation theory based approach to explore the quasiparticle (QP) and excitonic properties of materials. GW ...
  53. [53]
    Excited state non-adiabatic dynamics of large photoswitchable ...
    Jun 15, 2022 · Here we introduce a diabatic artificial neural network (DANN), based on diabatic states, to accelerate such simulations for azobenzene derivatives.Results · Ml Architecture And Training · Virtual ScreeningMissing: post- | Show results with:post-