Fact-checked by Grok 2 weeks ago

Ionization

Ionization is the process by which an atom or a molecule acquires a negative or positive charge by gaining or losing one or more electrons, thereby forming ions. This fundamental phenomenon alters the electrical balance of matter and is central to numerous physical and chemical processes. Ionization can occur through diverse mechanisms, each driven by specific energy inputs. Thermal ionization involves heating atoms or molecules to provide sufficient kinetic energy for electron ejection, commonly observed in high-temperature environments like stars or plasmas. Photoionization results from the absorption of photons with energy exceeding the ionization potential, such as in ultraviolet or X-ray interactions with matter. Impact ionization arises from collisions between charged particles, like electrons or ions, and neutral atoms, leading to electron stripping and often cascading effects in gases. Other methods include field ionization, where strong electric fields tunnel electrons out, and chemical ionization via reactive species in molecular interactions. The energy required for ionization, known as the ionization energy or potential, varies by element and electron shell, with successive removals demanding progressively more energy due to increased electrostatic attraction in the resulting ions. This process is pivotal in fields such as plasma physics, where ionized gases enable fusion research and auroral phenomena; radiation physics, for detecting and measuring ionizing particles; and atmospheric science, influencing ionospheric conductivity and radio wave propagation. In analytical chemistry, controlled ionization techniques underpin mass spectrometry for molecular identification. Understanding ionization also informs radiation safety, as it underlies biological damage from high-energy particles by creating reactive ion pairs in tissues.

Basic Concepts

Definition and Types

Ionization is the physical process by which an , , or acquires a net by either losing one or more to form positively charged cations or gaining to form negatively charged anions. This process disrupts the electrical neutrality of the involved, often requiring the input of to overcome the binding forces of the . A fundamental representation of ionization is the reaction where a or A absorbs to eject an , yielding a cation and a :
\ce{A + energy -> A^+ + e^-} In some cases, ionization can produce ion pairs, where both a cation and an anion are formed simultaneously, such as through the interaction of high-energy particles or with .
Ionization can be classified into several primary types based on the and number of electrons involved. ionization refers to the removal of one from the neutral species, which is the most common form and occurs in various -input scenarios. Multiple ionization involves the removal of two or more electrons, which can proceed sequentially—where electrons are ejected one at a time in successive steps—or non-sequentially, involving correlated electron dynamics in intense fields. Key subtypes include , where high temperatures provide the to overcome electron binding in gases or vapors; , triggered by the absorption of photons with sufficient to eject electrons; and field ionization, induced by strong that promote electron tunneling from the or . Ionization manifests in diverse contexts across physical states, influencing material properties and phenomena. In gases, it leads to the formation of plasmas, which are partially or fully ionized states consisting of free electrons, ions, and neutrals that enable electrical conductivity and collective behaviors. In aqueous solutions, ionization of electrolytes produces dissociated ions that facilitate conduction, as seen in salts like dissociating into Na⁺ and Cl⁻ ions. In solids, particularly semiconductors, processes like generate charge carriers by energetic electrons colliding with lattice atoms, enabling applications in electronics. The foundational understanding of ionization traces back to J.J. Thomson's 1897 experiments on in partially evacuated tubes, where he observed gas ionization and measured the charge-to-mass ratio of electrons, leading to their discovery.

Ionization Energy

Ionization energy is defined as the minimum energy required to remove an from a gaseous atom or in its . The first (IE₁) refers to the removal of the most loosely bound from a neutral atom, while successive ionization energies (IE₂, IE₃, etc.) correspond to removing additional electrons from the resulting cation. These successive energies increase progressively because, after the initial electron removal, the remaining electrons experience a higher (Z_eff), as there is less electronic shielding from the . In the periodic table, ionization energies exhibit clear trends: they generally decrease down a group due to increasing atomic radius and greater shielding by inner electrons, which reduces Z_eff for valence electrons, and increase across a period from left to right as Z_eff rises without a corresponding increase in shielding. Notable exceptions occur, such as the decrease between group 2 and group 13 elements (e.g., from beryllium to boron), where the electron configuration allows removal from a higher-energy p orbital rather than a stable s orbital. Ionization energies are typically measured using techniques like photoelectron spectroscopy, which determines electron binding energies from the kinetic energy of ejected photoelectrons, or atomic spectroscopy, which observes spectral lines corresponding to ionization thresholds; values are expressed in electron volts (eV) or kilojoules per mole (kJ/mol). For molecules, ionization energy relates to molecular orbital (MO) theory, where the energy required approximates the negative of the highest occupied (HOMO) energy according to , assuming no electron rearrangement. Two key types are distinguished: adiabatic ionization energy, the minimum energy for transition between relaxed ground states of the neutral and ion (accounting for geometry changes), and vertical ionization energy, which assumes instantaneous electron removal without nuclear motion, governed by the Franck-Condon principle. In photoelectron spectroscopy, the ionization energy is calculated as IE = hν - KE, where hν is the incident and KE is the measured of the ejected ; for atomic hydrogen, this yields IE₁ = 13.6 eV, corresponding to excitation from its energy of -13.6 eV to the ionized state at 0 eV. Ionization energies correlate with other atomic properties, such as , which often scales with due to the atom's tendency to attract s in bonds (e.g., Mulliken electronegativity is ( + EA)/2, where EA is ), and , as both reflect the stability of electron removal or addition.

Production Mechanisms

Thermal and Adiabatic Processes

Thermal ionization refers to the process by which atoms or molecules in a gas acquire sufficient through collisions to overcome their , resulting in the formation of ions and free electrons, typically in a state of . This mechanism dominates in high-temperature environments where the kinetic energy follows a Maxwell-Boltzmann profile, enabling electron-impact ionization. In such systems, the equilibrium ionization state is governed by the balance between ionization and recombination rates, leading to a predictable of ionization stages based on temperature and density. The Saha equation quantifies this equilibrium for a plasma, relating the densities of consecutive ionization stages to temperature. For the transition from ionization stage i to i+1, it is expressed as \frac{n_{i+1} n_e}{n_i} = \frac{2 g_{i+1}}{g_i} \left( \frac{2 \pi m_e k T}{h^2} \right)^{3/2} \exp\left( -\frac{\chi_i}{k T} \right), where n_{i+1}, n_i, and n_e are the number densities of the ions in stages i+1 and i and free electrons, respectively; g_{i+1} and g_i are the statistical weights of the respective states; m_e is the electron mass; k is Boltzmann's constant; T is the temperature; h is Planck's constant; and \chi_i is the ionization energy from stage i to i+1. This equation assumes local thermodynamic equilibrium and neglects interactions beyond binary collisions. Derived from statistical mechanics and detailed balance principles, it applies to dilute gases where pressure effects are minimal. In astrophysical contexts, the Saha equation accurately predicts ionization in stellar atmospheres, such as the high in the solar corona, where temperatures exceed 1 MK lead to nearly complete stripping of and . Similarly, in arc discharges, thermal ionization sustains conductive plasmas at temperatures around 5000–10,000 K, facilitating applications in and . However, at high densities (above ~10^{18} cm^{-3}), the assumption of equilibrium breaks down due to enhanced three-body recombination, where an ion-electron pair recombines in the presence of a third body, absorbing excess energy and suppressing net ionization. This limits the applicability of the Saha equation in dense, optically thick plasmas. Ionization in adiabatic processes occurs in gases where energy is added or removed without heat exchange with the surroundings, maintaining constant . This includes reversible or that changes and , thereby shifting the ionization balance. For example, in astrophysical outflows, adiabatic leads to cooling and recombination, while initial phases can drive ionization. This mechanism is relevant in systems without rapid non-thermal energy inputs, allowing the gas to follow paths. Practical examples include ionization in flame spectroscopy, where alkali metals like sodium ionize thermally at flame temperatures of 1500–2500 K, enhancing emission signals for trace analysis. In thermionic energy converters, used in thermoelectric generators, thermal ionization emits electrons from a (around 2000 K), generating across a gap for power conversion. Historically, Irving Langmuir's studies in the on gas discharges and thermionic phenomena at laid foundational insights into behavior, including equilibrium ionization in heated vapors, influencing modern applications. For non-equilibrium conditions, the is described by rate equations, where the net ionization for i is \frac{dn_i}{dt} = n_e n_i S_i - n_{i+1} n_e \alpha_{i+1}, with S_i the ionization coefficient and \alpha_{i+1} the recombination coefficient. The collision ionization coefficient S_i for -impact is derived from the collision cross-section \sigma_i(\epsilon), where \epsilon is the , via the Maxwellian average: S_i(T) = \int_0^\infty \sigma_i(\epsilon) v f(\epsilon, T) d\epsilon, with velocity v = \sqrt{2\epsilon / m_e} and Maxwellian distribution f(\epsilon, T) = \frac{2}{\sqrt{\pi}} (kT)^{-3/2} \sqrt{\epsilon} \exp(-\epsilon / kT). This integral is evaluated using empirical or quantum cross-sections, often approximated for thresholds near \chi_i by forms like the Seaton expression, S_i(T) \approx 10^{-7} (T/10^4 \mathrm{K})^{1/2} \exp(-\chi_i / kT) cm³/s for hydrogen-like atoms, ensuring detailed balance with recombination in equilibrium via S_i / \alpha_{i+1} = K(T) from the Saha equation. Such derivations underpin time-dependent modeling in transient plasmas.

Impact Ionization

Impact ionization occurs through collisions between energetic charged particles, such as electrons or ions, and neutral atoms or molecules, where the incident particle transfers sufficient energy to eject an electron, creating an ion pair. This process is prominent in non-equilibrium conditions, such as in gas discharges, semiconductors, and radiation detectors, and can lead to avalanche effects where newly freed electrons further ionize, amplifying the initial event. The probability is governed by the impact ionization cross section \sigma(\epsilon), which is zero below the threshold energy (equal to the ionization potential) and rises sharply thereafter, often modeled by forms like the Lotz formula or quantum calculations. For electron-impact, the differential cross section depends on the kinematics, with energy and momentum conservation determining the ejected electron's energy. In gases, the Townsend avalanche coefficient \alpha quantifies the number of ionizations per unit path length: \alpha = n \int_{\chi}^\infty \sigma(\epsilon) f(\epsilon) v d\epsilon / v_d, where n is neutral density, f(\epsilon) the electron energy distribution, and v_d the drift velocity. This leads to breakdown when \exp(\alpha d) \approx 10^8 for gap distance d. Applications include gas-filled detectors like Geiger-Müller counters, where impact ionization sustains the discharge pulse, and in semiconductor avalanche photodiodes for low-light detection. In astrophysics, cosmic ray impacts ionize interstellar media.

Chemical Ionization

Chemical ionization involves the interaction of neutral molecules with pre-existing ions or reactive species, leading to charge transfer or protonation/deprotonation without direct electron removal by photons or fields. This soft ionization technique minimizes fragmentation, preserving molecular identity, and is widely used in mass spectrometry. In a typical setup, reagent gases like methane produce reactant ions (e.g., CH₅⁺ or C₂H₅⁺) via electron-impact, which then react with analytes: e.g., A + CH₅⁺ → AH⁺ + CH₄ (proton transfer if proton affinity of A > methane). The rate constants follow Langevin capture theory for ion-molecule collisions, k = 2\pi q \sqrt{\alpha / \mu}, where q is ion charge, \alpha polarizability, \mu reduced mass. Exothermic reactions proceed near the collision rate, while endothermic ones are suppressed. Developed by Field and Munson in 1965, it contrasts with hard ionization methods by producing abundant [M+H]⁺ or [M-H]⁻ ions. This method is essential for analyzing thermally labile compounds in environmental and biological samples, with minimal excess energy (~1-5 eV) compared to (~70 eV).

Photoionization Processes

is the process by which an or is ionized through the absorption of , specifically photons, leading to the ejection of an . This section focuses on single-photon and multiphoton variants, where the or combined energies exceed the (IE). These processes are fundamental in fields ranging from to , enabling selective ionization without significant thermal effects. Single-photon ionization occurs when a single photon with energy h\nu \geq IE is absorbed, promoting an electron from a bound orbital to the continuum. The probability of this process is quantified by the photoionization cross section \sigma(\omega), derived from quantum mechanical time-dependent perturbation theory, which is proportional to the square of the dipole matrix element between the initial bound state and the final continuum state: \sigma(\omega) \propto |\langle \psi_f | \mathbf{r} \cdot \mathbf{\epsilon} | \psi_i \rangle|^2, where \psi_i and \psi_f are the initial and final wave functions, \mathbf{r} is the position operator, and \mathbf{\epsilon} is the photon polarization vector. In practice, for atoms, the cross section decreases as \sigma(\omega) \sim 1/\omega^{7/2} above threshold due to the radial wave function overlap. A key application is in atmospheric science, where solar ultraviolet radiation (wavelengths ~80–100 nm) ionizes O₂ molecules in the ionosphere, producing O₂⁺ ions essential for plasma formation and radio wave propagation; measured cross sections peak near 18 eV with values around 10⁻¹⁷ cm². Multiphoton ionization (MPI) extends this to scenarios where laser intensities are moderate (10⁹–10¹² W/cm²), allowing absorption of n photons such that n h\nu \geq IE, even if individual photon energies are below threshold. In the perturbative regime, the ionization rate follows w = \sigma^{(n)} I^n, where I is the laser intensity and \sigma^{(n)} is the generalized n-photon cross section, with units adjusted by (\hbar \omega)^{n-1} for dimensionality. This process was pioneered in the 1960s using early (emitting at 694 nm), enabling the first observations of two-photon ionization in vapors like cesium. Thresholds in MPI exhibit sharp onsets, but resonances—such as intermediate excited states—enhance rates via resonance-enhanced MPI (REMPI), improving selectivity. Above-threshold ionization (ATI), a related , involves absorption of additional photons beyond the minimum n, imparting excess to the ; the shows peaks separated by h\nu, shifted by the ponderomotive energy U_p = \frac{e^2 E^2}{4 m_e \omega^2}, where E is the field amplitude, m_e the , and \omega the —representing the cycle-averaged quiver energy of a . ATI was first observed in 1979 using under intense laser fields. Experimental setups for often employ laser-based photoionization mass spectrometry (PIMS), which combines tunable lasers with time-of-flight mass analyzers to detect ionized species selectively by their ionization potentials. Development accelerated in the with ruby lasers, evolving into versatile tools for analysis by the 1970s. For molecules, photoionization dynamics are complicated by autoionizing states—superexcited levels above IE that decay spontaneously into ion + continua—and shape s, where the outgoing electron temporarily traps in a centrifugal or molecular potential barrier, manifesting as broad enhancements in cross sections (e.g., π* shape resonance in N₂ at ~15 eV). These features lead to Fano-like asymmetric profiles in spectra, influencing branching ratios and angular distributions.

Field-Induced Processes

Field-induced ionization refers to the process where strong distort the atomic or molecular potential barriers, enabling escape through quantum tunneling or barrier suppression. This mechanism is distinct from thermal or photonic excitation, as it relies primarily on the field's influence on the landscape. In static fields, typically (DC) or low-frequency (AC) fields, electrons tunnel through a triangular barrier formed by the superposition of the atomic potential and the external field. The foundational theory for this static field emission, known as cold emission due to its temperature independence, was developed in 1928 by and Lothar Nordheim, who derived an expression for the emission current density based on the for tunneling probability. The Fowler-Nordheim equation quantifies the current density J for electron emission from a metal surface in a static electric field E, approximated as: J = \frac{e^3 E^2}{8\pi h \phi} \exp\left( -\frac{8\pi \sqrt{2m \phi^3}}{3 h e E} \right), where e is the electron charge, h is Planck's constant, m is the electron mass, and \phi is the work function of the material. This exponential dependence highlights the field's critical role in reducing the barrier width, allowing observable currents at field strengths on the order of 1-10 GV/m for typical metals. Applications of static field ionization include field emission microscopy, where sharp tips emit electrons to image atomic-scale surface structures with high resolution, and ion sources for particle accelerators, such as liquid metal ion sources that generate focused beams via field evaporation of surface atoms. In dynamic fields, such as those from oscillating or pulses, the time-varying nature leads to barrier suppression rather than pure tunneling, particularly at high intensities. The Keldysh parameter \gamma = \frac{\omega \sqrt{2m I_p}}{e E}, where \omega is the field oscillation frequency, I_p is the ionization potential, and other symbols as before, delineates the regimes: perturbative multiphoton ionization dominates when \gamma \gg 1 (high frequency, low intensity), while non-perturbative tunneling or barrier suppression prevails for \gamma \ll 1 (low frequency, high intensity). This parameter, introduced in Leonid Keldysh's 1965 theory, provides a framework for understanding field-driven ionization across optical to wavelengths. In contexts, field-induced processes contribute to ionization in dielectrics, where initial electrons gain energy from the field and collide to ionize additional atoms, leading to rapid growth and . This mechanism is prominent in -matter interactions within transparent materials, where the field strength exceeds the material's bandgap, initiating a cascade that can damage or enable formation. For instance, in pulses, field-dependent rates determine the damage threshold, with critical fields around 0.1-1 MV/cm for common dielectrics like fused silica.

Theoretical Descriptions

Semi-Classical Approaches

Semi-classical approaches to ionization integrate classical descriptions of electron trajectories with quantum mechanical elements to model the escape of electrons from atomic or molecular potentials under external fields. In these models, the electron's motion is treated classically once it is sufficiently far from the or ion core, influenced by the combined and external field potentials, while quantum transitions or initial conditions account for the departure from bound states. This hybrid framework is especially effective for Rydberg states, where the large orbital radii allow classical-like behavior, enabling the study of field-induced escape dynamics without full quantum wavefunction computations. A central model within this paradigm is the classical over-the-barrier ionization (OBI) framework, where ionization occurs when the external field suppresses the Coulomb barrier below the electron's energy level, permitting classical escape over the saddle point of the potential. This model incorporates Coulomb focusing, in which the attractive field of the residual ion core bends and converges the trajectories of outgoing electrons, modifying their asymptotic momentum distributions and enhancing recollision probabilities in intense laser fields. Developed as an extension of earlier classical trajectory methods, the OBI approach provides intuitive insights into barrier penetration without relying on tunneling probabilities. These semi-classical techniques find key applications in interpreting high-order above-threshold ionization (ATI) spectra, where classical electron trajectories driven by laser fields explain the characteristic energy plateaus, cutoffs, and angular distributions observed in photoelectron experiments. By simulating ensembles of trajectories launched from quantum-initialized states, the models capture rescattering effects that contribute to the multi-photon absorption features beyond the ponderomotive limit. Additionally, they elucidate the role of classical in atomic systems, particularly for Rydberg atoms in or static fields, where irregular, diffusive trajectories lead to ionization thresholds and sensitivity to field perturbations. Despite their strengths, semi-classical approaches have limitations, particularly at low energies where quantum tunneling through the barrier dominates, rendering classical escape improbable and necessitating full quantum treatments. Originating in the 1970s through advancements in , these methods evolved from semiclassical impact-parameter approximations to address electron promotion and capture, providing a foundation for later strong-field extensions.

Quantum Mechanical Models

Quantum mechanical models of ionization fundamentally rely on solving the time-dependent (TDSE) for an system interacting with an external field, treating the field as a to the unperturbed Hamiltonian. These approaches capture the quantum nature of , including wavefunction and probabilities, contrasting with semi-classical methods by emphasizing full operator-based descriptions rather than hybrid trajectories. For weak fields, where the perturbation does not significantly distort the bound states, time-independent provides corrections to energy levels, while time-dependent formulations yield rates to continuum states representing ionization. The foundational framework for time-dependent perturbation theory was established by Dirac in 1926, enabling the calculation of transition amplitudes between initial bound states and final continuum states under weak, time-varying perturbations such as electromagnetic fields. In the limit of a continuum of final states, this leads to Fermi's golden rule, which gives the transition rate \Gamma from an initial state |i\rangle to final states |f\rangle as \Gamma = \frac{2\pi}{\hbar} \left| \langle f | H' | i \rangle \right|^2 \rho(E_f), where H' is the perturbation Hamiltonian (e.g., the dipole interaction - \mathbf{d} \cdot \mathbf{E}(t) for an electric field \mathbf{E}(t)) and \rho(E_f) is the density of final states at energy E_f matching the initial energy plus absorbed photon energy. This rate quantifies photoionization probabilities in weak laser fields, assuming first-order processes where the electron is promoted directly to the continuum. For static weak fields, time-independent perturbation theory similarly shifts bound-state energies, revealing avoided crossings that signal potential ionization pathways, though exact continuum transitions require the time-dependent extension. For the simplest atomic system, the , exact solutions to the exist in specific configurations, providing benchmarks for ionization models. In a uniform static electric , the separates exactly in (\xi, \eta, \phi), where \xi = r + z and \eta = r - z, yielding separable wavefunctions and energy levels labeled by quantum numbers n, n_1, m (with n = n_1 + n_2 + |m| + 1). This solution shows linear energy shifts for low-lying states and quadratic for higher ones, with the field-induced mixing of states lowering the ionization threshold and enabling tunneling-like escape in stronger fields, though perturbative limits suffice for weak perturbations. These exact hydrogenic results, first derived using in the early quantum era, underpin approximations for more complex atoms by highlighting field-induced asymmetry in probability distributions. In time-dependent scenarios, such as pulses, advanced basis states account for field-dressing of electrons. Volkov states describe free electrons in a plane-wave field, providing exact solutions to the TDSE for a particle in a classical electromagnetic wave, with the wavefunction incorporating oscillatory shifts due to the vector potential \mathbf{A}(t). These states are used to construct "dressed" continuum wavefunctions in ionization amplitudes, capturing the quiver motion of ionized electrons without atomic binding. For monochromatic or periodic fields, transforms the TDSE into a time-independent eigenvalue problem via quasienergy states, where solutions take the form \psi(\mathbf{r}, t) = e^{-i \epsilon t / \hbar} \phi(\mathbf{r}, t) with \phi periodic in the field cycle T, \phi(\mathbf{r}, t + T) = \phi(\mathbf{r}, t). This approach reveals Floquet sidebands in the energy spectrum, corresponding to multi-photon ionization channels, and is particularly useful for high-frequency fields where breaks down but periodicity persists. In the perturbative regime, Floquet methods recover Dirac's results, while non-perturbatively they predict dynamical stabilization against ionization. For multi-electron atoms, where electron correlation complicates single-particle pictures, configuration interaction (CI) approximations expand the many-body wavefunction as a linear combination of Slater determinants from a basis of orbitals, solving the TDSE or TISE variationally to include interactions beyond mean-field Hartree-Fock. Seminal applications, such as Hylleraas' early CI for helium ground-state energy, demonstrate how mixing configurations (e.g., promoting an to a continuum orbital) captures effects in ionization potentials and rates. Truncated CI, like singles-doubles (CISD), balances accuracy and computation for systems like or atoms, yielding ionization energies within 0.1 of experiment by accounting for double excitations that mimic shake-up during ejection. These methods extend hydrogenic models by incorporating effective potentials, providing a quantum foundation for understanding correlated multi-electron escape in weak fields. Such full-wavefunction approaches serve as limits for semi-classical approximations in more intense regimes.

Strong Field Approximations

The strong-field approximation (SFA) provides a framework for describing ionization processes in intense fields, where the is dominated by the external rather than atomic potentials. In this approach, the from a bound initial state to a final state is computed using the element, which neglects the after ionization and treats the as propagating in the alone. The ionization rate w is then obtained as w = \frac{1}{T} \sum_f |c_f(T)|^2, where T is the pulse duration, the sum is over final states, and c_f(T) arises from the operator in the . The SFA amplitudes are evaluated using the saddle-point method applied to the classical action S = \int^t \frac{ [ \mathbf{p} + \mathbf{A}(t') ]^2 }{2m} dt', where \mathbf{p} is the canonical momentum, \mathbf{A}(t) is the of the field, and m is the . This integral yields saddle points that determine the dominant contributions to the transition amplitude, resulting in an ionization rate approximated by w \approx \exp(-2 \operatorname{Im} S), capturing the exponential suppression due to tunneling or multiphoton processes. The approximation is valid in the regime of high intensities where the ponderomotive energy U_p = \frac{e^2 E^2}{4 m \omega^2} (with E the field and \omega the ) greatly exceeds the I_p, ensuring that the field dressing of the continuum states dominates over atomic binding. Extensions of the SFA distinguish between velocity-gauge and length-gauge formulations; the velocity gauge, which uses the directly, is typically preferred for its and accuracy in describing trajectories, while the length gauge, based on the , can introduce gauge-dependent artifacts at low frequencies. Historically, the SFA traces to the 1965 work of Keldysh, who introduced the foundational non-perturbative treatment for multiphoton and tunneling ionization, with key developments in the 1980s by Reiss formalizing the approach for intense fields.

Advanced Phenomena

Tunnel Ionization Dynamics

Tunnel ionization occurs when an in an or escapes through an exponentially suppressed potential barrier induced by a strong , rather than surmounting it classically. This quantum mechanical process was first theoretically described for static fields in the context of field emission from atoms. In such scenarios, the barrier is distorted by the field, allowing the electron wavefunction to penetrate and emerge on the other side with a probability that decays exponentially with the barrier width. For alternating electromagnetic fields, particularly those from intense lasers, the Ammosov-Delone-Krainov (ADK) model provides a widely used expression for the tunneling ionization rate of complex atoms and ions from arbitrary initial states. The ADK rate is given by w = |C_{n^* l^*}|^2 \frac{Z^3}{F} \left( \frac{2 Z}{n^* F} \right)^{2 n^* - |m| - 1} \exp\left[ -\frac{2 Z^3}{3 n^{*3} F} \right], where F is the field strength, Z is the charge of the residual ion, n^* = Z / \sqrt{2 I_p} is the effective with I_p the ionization potential, |m| is the absolute value of the , and C_{n^* l^*} is a related to the initial orbital. This model builds on the strong field approximation, adapting static tunneling concepts to time-varying fields. In the quasi-static approximation, valid for low-frequency fields where \omega \ll I_p / \hbar, the ionization rate is treated as instantaneous and depends on the field's phase at the moment of tunneling, effectively averaging the static rate over the field's cycle. For higher frequencies, dynamic tunneling effects become prominent, involving imaginary-time trajectories in the electron's path under the oscillating field; here, the Keldysh parameter \gamma = \omega \sqrt{2 I_p} / F quantifies the transition from tunneling (\gamma \ll 1) to multiphoton regimes, with modifications to the barrier penetration accounting for the field's temporal variation. Key observables in tunnel ionization dynamics include ionization delay times, which measure the temporal offset between peak field strength and electron release—often on the scale—and transverse distributions of ionized electrons, revealing the tunneling exit geometry and field-induced acceleration. These have been probed experimentally using attoclock techniques and streaking methods. Historically, while Oppenheimer's 1928 work laid the static foundation, adaptations for fields emerged in the with the advent of tabletop intense systems, enabling studies of AC tunneling in gases and solids.

Multiple Ionization Effects

Multiple ionization refers to the removal of more than one from an or under intense fields, often revealing correlated dynamics beyond independent single- processes. In strong-field regimes, such as those accessed by Ti:sapphire lasers in the 1990s, experiments on like and demonstrated enhanced double ionization yields at specific intensities, signaling non-sequential mechanisms. Non-sequential double ionization (NSDI) occurs when an initially tunnel-ionized recollides with the parent , ejecting a second through electron-impact ionization. This recollision model, proposed in the early 1990s, explains the characteristic "knee" in the double-ionization yield versus intensity curve, where the yield rises more steeply than predicted by sequential ionization at intensities around 10^14–10^15 W/cm² for . Early observations in the mid-1990s using Ti:sapphire lasers confirmed this knee structure in , attributing it to correlated ejection driven by the recolliding 's peaking at about 3.17 U_p, where U_p is the ponderomotive . Population trapping arises in () fields when electrons are localized in stable Rydberg-like states, librating without ionizing due to AC Stark shifts that detune multiphoton resonances. This effect stabilizes population in high-lying states by shifting their energies such that the field-dressed levels avoid further ionization pathways, observed in atoms under femtosecond Ti:sapphire pulses where trapping efficiency depends on the frequency and intensity matching the Stark-shifted resonance conditions. In , trapping manifests as reduced ionization rates at specific field parameters, contrasting with over-the-barrier in unbound trajectories. For molecules, inner-valence multiphoton ionization (MPI) excites electrons from deeper orbitals, often leading to unstable dications and rapid . In (N₂), inner-valence ionization populates repulsive states of the N₂²⁺ dication, resulting in N⁺ + N⁺ fragmentation with releases around 5–10 , as probed by wavelength-selected XUV pulses. This process highlights site-specific , where inner-shell destabilizes the molecular more effectively than outer-valence removal, contributing to observed times on the scale. Correlated effects in molecules are exemplified by charge-resonance enhanced ionization (CREI), where the potential energy curves of ionic states cross at a critical internuclear (typically 2–3 a₀ for diatomic systems), enabling resonant coupling that boosts the second ionization rate. In and H₂⁺, CREI amplifies multiple ionization as the stretches during the laser pulse, with enhancement factors up to orders of magnitude over atomic rates at internuclear separations matching the . These correlations were first theoretically detailed in the mid-1990s and later observed in experiments with intense pulses, underscoring the role of motion in multi-electron ejection.

Kramers-Henneberger Transformations

The Kramers-Henneberger transformation is a unitary transformation used to analyze the dynamics of atoms exposed to intense, oscillating fields by shifting to a that tracks the classical quiver motion of the induced by the field. This approach, originally developed by W. Henneberger in 1968 as a method for atoms in intense , builds on earlier ideas from H. A. Kramers in the 1950s regarding accelerated frames in . In the , extensions of this framework were applied to , facilitating the study of field-dressed states in periodic potentials. The transformation effectively dresses the atomic potential with the laser field, providing insight into how intense fields modify electronic structure without direct inclusion of the in the . In the Kramers-Henneberger , the coordinate transformation follows the quiver displacement \alpha(t), defined as \alpha(t) = -\int^t A(t') \, dt' / c, where A(t) is the of the field and c is the (in , this simplifies accordingly). This shift to an accelerated incorporates the 's oscillatory drive into the potential, yielding a time-dependent V(r - \alpha(t)). For high-frequency fields where the laser period is much shorter than atomic response times, a cycle-averaged, quasi-static emerges: V_{\text{eff}}(r) = \frac{1}{T} \int_0^T V(r - \alpha(t)) \, dt, with T = 2\pi / \omega the optical period. This averaged potential captures the ponderomotive effects and field-induced modifications to binding, such as stabilization or destabilization of states. The transformation finds key applications in describing quiver-averaged potentials for atoms and molecules in high-frequency regimes, where the electron's oscillatory motion smears the Coulomb interaction, leading to altered energy levels and ionization thresholds. In molecular systems, it predicts phenomena like bond softening, where the effective potential reduces internuclear barriers and lowers dissociation energies, or bond hardening under certain field polarizations that enhance stability. These predictions arise from the displacement of charge distributions in the laser-dressed frame, offering a length-gauge perspective complementary to velocity-gauge formulations in strong-field approximations (SFA), which simplify multiphoton and tunneling ionization rates. Despite its utility, the Kramers-Henneberger transformation has limitations, particularly breaking down for low-frequency fields where the quiver period approaches atomic timescales, preventing valid cycle averaging and leading to non-adiabatic effects. It also falters under strong interactions, as the assumption of nearly free-electron quiver motion fails when binding energies compete with the field-driven oscillations. These constraints highlight its suitability primarily for the high-frequency approximation in intense-field physics.

Applications and Distinctions

Practical Uses

Ionization processes underpin a wide array of practical applications in science, technology, and industry, enabling the manipulation of charged particles for precise analysis, energy generation, and environmental control. In analytical techniques, electron impact ionization remains a cornerstone of , where a beam of 70 electrons bombards gaseous molecules, fragmenting them into characteristic ions that are separated and detected to identify chemical compositions with high sensitivity. Similarly, photoelectron spectroscopy utilizes to examine surface properties, ejecting electrons from atoms in the top few nanometers of a material to reveal elemental composition, chemical states, and bonding configurations essential for and research. In energy and propulsion systems, field emission ionization facilitates the operation of ion thrusters, particularly (FEEP) devices, where intense electric fields extract and accelerate s from liquid metal propellants like cesium or , providing micro-Newton thrust levels with specific impulses exceeding 8000 seconds for station-keeping and deep-space missions. In fabrication, relies on ionization within reactive gas plasmas to generate charged species and radicals that anisotropically remove material layers, enabling the creation of intricate nanostructures in microchips with feature sizes below 5 nm. Medical applications harness ionizing radiation for therapy, directing high-energy beams such as X-rays or protons to induce ionization in tumor cells, causing DNA strand breaks that preferentially kill malignant tissue while minimizing damage to surrounding healthy structures through techniques like intensity-modulated radiation therapy. Environmentally, corona discharge ionization is employed in air purification systems, where a high-voltage electrode creates a plasma that produces reactive ions and ozone to neutralize airborne pathogens, volatile organic compounds, and particulates, though careful design is needed to limit byproduct emissions. In , ionization forms critical layers in planetary ionospheres, such as Earth's F-layer where solar radiation ionizes oxygen and to create a that reflects high-frequency radio waves and influences signals and auroral displays. Stellar spectra analysis exploits ionization equilibria to infer astrophysical conditions, as the ratios of ionized to neutral lines (e.g., He II versus He I) reveal surface temperatures ranging from 10,000 K in O-type stars to 3,000 K in M-types, aiding in the determination of elemental abundances and evolutionary stages. Post-2020 advancements have leveraged pulse technology for real-time imaging of ionization dynamics, using isolated pulses shorter than 100 to capture wave packets during tunnel ionization in atoms, enabling visualization of charge migration on timescales for potential applications in quantum control. This progress was recognized by the 2023 , awarded to , , and for pioneering methods to generate light pulses used in probing motion during ionization. Safety considerations distinguish , which possesses sufficient energy (typically >12.4 eV for photons, corresponding to ultraviolet wavelengths shorter than 100 nm) to eject electrons from atoms and potentially cause cellular damage leading to cancer, from like visible light or microwaves that lacks this capability and poses lower biological risks at equivalent exposures. Regulatory thresholds, such as the International Commission on Radiological Protection's limits of 20 mSv per year averaged over five years for occupational exposure to , ensure protection by maintaining doses below levels associated with stochastic health effects.

Distinction from Dissociation

Ionization refers to the process by which an atom or loses one or more , resulting in the formation of positively charged ions, without necessarily involving the breaking of chemical . In contrast, is the separation of a molecular into two or more smaller molecular , such as atoms or fragments, through the cleavage of chemical , and it does not inherently involve a net loss of from the . These processes can occur independently in molecular , but they are distinct in their primary mechanisms and outcomes: ionization produces charged that retain their molecular structure, whereas dissociation yields neutral or charged fragments depending on prior ionization states, but emphasizes rupture over electron removal. In molecular contexts, single ionization typically creates a stable charged molecular ion without immediate breaking, as the positive charge is delocalized across the . For instance, the ionization of the hydrogen molecule (H₂) produces the stable H₂⁺ ion, which maintains a with a energy of approximately 2.8 , contrasting with the neutral H₂ into two hydrogen atoms (H + H) that requires only about 4.5 . The for H₂ to form H₂⁺ is significantly higher, around 15.4 , highlighting how ionization thresholds exceed typical energies by an . Dissociation can proceed via direct pathways, where places the on a repulsive leading to immediate bond cleavage, or through , involving coupling between a bound vibrational state and a dissociative continuum, resulting in delayed fragmentation without net loss. Overlaps arise when ionization induces , such as through the population of repulsive excited states in the ion, or via charge transfer processes that redistribute charge without pure bond breaking; however, these differ from pure , which avoids ejection. Multiple ionization can escalate this to , where sequential removal creates highly charged that repel and fragment due to electrostatic forces. To distinguish these processes experimentally, is commonly employed, as it separates ions by , allowing identification of intact molecular ions (indicative of ionization without ) versus fragment ions (signaling or combined processes) based on their yield and distributions.

Ionization Energy Tables

Ionization energies provide quantitative measures of the energy required to remove s from atoms and molecules, serving as empirical data that embody discussed in basic concepts. The first ionization energy (IE₁) represents the energy to remove the most loosely bound from a neutral atom in its . Data compiled here are drawn from the NIST Atomic Spectra Database, which offers critically evaluated values based on experimental and theoretical assessments, with updates as of 2024 incorporating relativistic corrections for heavier elements (Z > 80) where quantum electrodynamic effects influence binding. Uncertainties in these measurements typically range from 0.0001 eV for light elements to 0.1 eV or more for superheavy elements like (Og, Z=118), where direct experimental data are limited and relativistic effects significantly lower the IE₁ compared to non-relativistic predictions. For instance, the IE₁ of (Au, Z=79) is 9.2257 eV, but relativistic stabilization of the 6s orbital reduces it relative to lighter analogs.

Table 1: First Ionization Energies of Elements (Z=1–118)

The table below lists IE₁ in electronvolts (eV) and kilojoules per mole (kJ/mol), converted using the factor 1 eV ≈ 96.485 kJ/mol. Values reflect ground-state transitions and highlight : IE₁ generally increases across a due to rising but decreases down a group due to increasing . Notable anomalies include (Be, 9.323 eV), where the filled 2s² subshell confers exceptional stability, resulting in a higher IE₁ than expected compared to (8.298 eV); similar stability appears in (14.534 eV) over oxygen (13.618 eV) due to half-filled p subshell. For visualization, plotting IE₁ across periods reveals sawtooth patterns peaking at (e.g., Ne: 21.565 eV) and minima at alkali metals (e.g., Na: 5.139 eV). Data sourced from NIST Atomic Spectra Database (2024).
ZElementIE₁ (eV)IE₁ (kJ/mol)
1H13.598441312.0
2He24.587412372.3
3Li5.39172520.2
4Be9.32270899.5
5B8.29699800.6
6C11.260301086.5
7N14.534141402.3
8O13.617661313.9
9F17.422821681.0
10Ne21.564602080.7
11Na5.13908495.8
12Mg7.64624737.7
13Al5.98577577.5
14Si8.15168786.5
15P10.486691011.8
16S10.36001999.6
17Cl12.967641251.2
18Ar15.759621520.5
19K4.34066418.8
20Ca6.11316589.8
21Sc6.56149633.1
22Ti6.82812658.8
23V6.74619650.9
24Cr6.76651652.9
25Mn7.43404717.3
26Fe7.90247762.5
27Co7.88101760.4
28Ni7.63988737.1
29Cu7.72638745.5
30Zn9.39420906.4
31Ga5.99930578.8
32Ge7.89943762.2
33As9.78860944.5
34Se9.75238941.0
35Br11.813811139.9
36Kr13.999601350.8
37Rb4.17713403.0
38Sr5.69487549.5
39Y6.21730599.8
40Zr6.63390640.1
41Nb6.75885652.1
42Mo7.09243684.3
43Tc7.11938686.8
44Ru7.36050710.2
45Rh7.45890719.7
46Pd8.33686804.4
47Ag7.57623731.0
48Cd8.99382867.8
49In5.78635558.3
50Sn7.34392708.6
51Sb8.60839830.6
52Te9.00962869.3
53I10.451261008.4
54Xe12.129831170.4
55Cs3.89390375.7
56Ba5.21169502.9
57La5.57690538.1
58Ce5.53860534.4
59Pr5.64737545.6
60Nd5.52514533.1
61Pm5.58172538.8
62Sm5.64370544.9
63Eu5.67037547.1
64Gd6.14980593.4
65Tb5.86380565.8
66Dy5.93905573.0
67Ho6.02151581.0
68Er6.10770589.3
69Tm6.18430596.7
70Yb6.25416603.4
71Lu5.42587523.5
72Hf6.82507658.5
73Ta7.54957728.6
74W7.86403758.6
75Re7.83352755.8
76Os8.43820814.3
77Ir8.96702865.2
78Pt8.95883864.4
79Au9.22553890.1
80Hg10.437501007.1
81Tl6.10840589.4
82Pb7.41668715.6
83Bi7.28550703.3
84Po8.41400812.1
85At9.31750899.0
86Rn10.748501036.7
87Fr4.07270393.0
88Ra5.27840509.3
89Ac5.38020519.4
90Th6.30670608.6
91Pa5.89060568.7
92U6.19400597.6
93Np6.26570604.6
94Pu6.02580581.6
95Am5.97380576.6
96Cm5.99140578.2
97Bk6.19770598.0
98Cf6.28170606.3
99Es6.36770614.4
100Fm6.41900619.4
101Md6.49626
102No6.58635
103Lr4.9473
104Rf6.0579
105Db6.2598
106Sg6.4617
107Bh6.6637
108Hs6.8656
109Mt7.0675
110Ds7.2695
111Rg7.4714
112Cn7.6734
113Nh7.6733
114Fl7.7743
115Mc8.0772
116Lv8.4810
117Ts8.9859
118Og8.6829
Note: Values for Z=101–118 are theoretical or extrapolated due to experimental challenges, with uncertainties up to ±1 eV; relativistic effects contract s-orbitals, lowering IE₁ for superheavies like Og.

Table 2: Successive Ionization Energies for the First 10 Elements

Successive ionization energies (IEₙ) increase nonlinearly as each removal occurs from a more positively charged ion, with sharp jumps after removing core electrons (e.g., after valence shell depletion). The table shows IE₁ to IE₅ (or available) in eV for elements Z=1–10, illustrating trends like helium's high IE₂ (54.417 eV) due to removing from a 1s¹ ion. Data from NIST Atomic Spectra Database (2024).
ElementIE₁ (eV)IE₂ (eV)IE₃ (eV)IE₄ (eV)IE₅ (eV)
13.598
He24.58754.417
5.39275.640122.45
Be9.32318.211153.89217.72
B8.29825.15437.93259.37340.22
C11.26024.38347.88864.494392.09
N14.53429.60147.44577.47597.869
O13.61835.12154.93577.411113.90
F17.42334.97162.70687.139114.24
Ne21.56541.063.097.0126.0
Uncertainties are <0.01 eV for these light elements. These values underscore shell effects, such as the large IE₃ for (122.45 ) marking 1s removal. For common molecules, ionization energies differ from due to bonding. (H₂O) has an adiabatic IE of 12.621 ± 0.002 (0 K ground-state to ground-state ion transition) and a vertical IE of approximately 12.62 (Franck-Condon maximum), reflecting minimal geometry change upon ionization from the highest occupied . (CO₂) exhibits an adiabatic IE of 13.777 ± 0.001 and vertical IE of 13.78 , with the linear structure preserved in the ion, leading to close values. These molecular data, from NIST Chemistry WebBook (2024), highlight how delocalized orbitals in CO₂ stabilize the neutral, raising its IE relative to oxygen.

References

  1. [1]
    Ionization - Energy Education
    Sep 3, 2018 · Ionization is the process by which ions are formed by gain or loss of an electron from an atom or molecule.
  2. [2]
    Ionization - an overview | ScienceDirect Topics
    Ionization is defined as the process in which high-energy radiation removes an electron from an atom, resulting in the formation of ion pairs and potentially ...
  3. [3]
    Ionization Processes - Computational Electromagneticaerodynamics
    Apr 1, 2016 · The main mechanisms for ionization consist of thermal, electron impact, chemical reactions, photoionization, and microwave or electron cyclotron ...
  4. [4]
    NIST: Electron-Impact Cross Section Database - Intro.
    This ionization cross section database is specifically designed for electron-impact ionization. It is versatile and can provide cross sections for atoms as ...
  5. [5]
    Ionization - Nondestructive Evaluation Physics : X-Ray - NDE-Ed.org
    The term "ionization" refers to the complete removal of an electron from an atom following the transfer of energy from a passing charged particle.
  6. [6]
    Ionization Energies - Chemistry LibreTexts
    Jan 29, 2023 · ionization energy is a measure of the energy needed to pull a particular electron away from the attraction of the nucleus. A high value of ...Definition: First Ionization Energy · Factors affecting the size of...
  7. [7]
    [PDF] Chapter 2: Overview of Atmospheric Ionizing Radiation (AIR)
    Hess's studies found the ionization rates to decrease with altitude up to 500 meters followed by a steady increase at higher altitudes to where the ground level.
  8. [8]
    Mass Spectrometry Ionization Methods - Chemistry at Emory
    Electron Impact ionization (EI) · Fast Atom Bombardment (FAB) · Electrospray ionization (ESI) · Atmospheric Pressure Chemical Ionization (APCI) · Matrix Assisted ...
  9. [9]
    PRINCIPLES OF IONIZING RADIATION - Toxicological ... - NCBI - NIH
    INTRODUCTION. This chapter provides an overview of the principles of ionizing radiation before a discussion of the health effects in Chapter 3.
  10. [10]
    Ionization - Nondestructive Evaluation Physics : Atomic Elements
    Ionization is any process that changes the electrical balance within an atom. If we remove an electron from a stable atom, the atom becomes electrically ...
  11. [11]
  12. [12]
    Non-sequential double ionization with near-single cycle laser pulses
    Aug 8, 2017 · Non-sequential double ionization (NSDI) in intense near-infrared laser fields is a fundamental process with electron-electron correlation ...
  13. [13]
    PHOTOIONIZATION Definition & Meaning - Dictionary.com
    the phenomenon in which the absorption of electromagnetic radiation by an atom in a gas induces the atom to emit a bound electron and thereby become ionized.<|control11|><|separator|>
  14. [14]
    Field ionization | physics - Britannica
    Oct 7, 2025 · Intense fields, of the order of 10 8 volts per centimetre, can be generated in the neighbourhood of sharp points and edges of electrodes.
  15. [15]
    About Plasmas and Fusion - Princeton Plasma Physics Laboratory
    The free negative electrons and positive ions in a plasma allow electric current to flow through it. In a plasma, electrons are freed from their atoms, allowing ...
  16. [16]
    Distribution Functions and Ionization Rates for Hot Electrons in ...
    The distribution of electrons in a semiconductor at high electric field is governed by a Boltzmann equation which describes the effects of the field.
  17. [17]
    Joseph John “J. J.” Thomson | Science History Institute
    In 1897 Thomson discovered the electron and then went on to propose a model for the structure of the atom. His work also led to the invention of the mass ...
  18. [18]
    Ionization Energy and Electron Affinity
    The first ionization energy of an element is the energy needed to remove the outermost, or highest energy, electron from a neutral atom in the gas phase.
  19. [19]
    Ionization Energy - Chemistry 301
    Ionization energy (I.E.) is the energy required to remove an electron from an atom in the gas phase. For example the ionization energy for a sodium atoms is ...Missing: definition | Show results with:definition
  20. [20]
    7.3 Energetics of Ion Formation
    Ionization energy is the energy needed to remove an electron from a gaseous atom. It's always positive, and successive ionization energies increase.<|control11|><|separator|>
  21. [21]
    [PDF] Periodic Trends Review Answer Key
    Ionization Energy Exceptions. - Group 2 to Group 13: The ionization energy decreases from beryllium (Be) to boron (B) because the added electron in boron ...
  22. [22]
    [PDF] Photoelectron Spectroscopy - Sandiego
    The kinetic energy (KE) of the ejected electron is measured, and the energy required to remove the electron from the atom is calculated from the difference.
  23. [23]
    [PDF] Vertical vs. Adiabatic Ionization Energies in Solution and Gas-Phase
    Sep 1, 2018 · 1, following Koopman's theorem and the Franck– Condon principle, initial vertical ionization from the HOMO occurs rapidly on the timescale of ...
  24. [24]
    4.3 The Hydrogen Atom - FAMU-FSU College of Engineering
    Such a loss of the electron is called “ionization” of the atom. The ionization energy of the hydrogen atom is 13.6 eV; this is the minimum amount of energy ...
  25. [25]
    [PDF] MITOCW | ocw-5-111-f08-lec10_300k
    So, in other words, we can just think of electronegativity as being the average of that ionization energy and the electron affinity. This should make sense ...
  26. [26]
    [PDF] Plasma ionization. Saha equation.
    The Saha equation describes ionization of plasma in thermal equilibrium. When plasma is not in thermodynamic equilibrium it is important to study elementary ...
  27. [27]
    [PDF] IONIZATION, SAHA EQUATION
    It is customary to express the ionization energy in electron volts, 1eV = 1.602 × 10−12 erg. Ionization of the most abundant elements, hydrogen and helium, is ...
  28. [28]
    [PDF] Ionization Equilibrium - Saha's Equation - MIT OpenCourseWare
    In equilibrium, an equal flux must be leaving the metal. The (bi-directional) electron current density is then given by,. 4πeme je = eΓe = h3.
  29. [29]
    [PDF] How the Saha Ionization Equation Was Discovered - arXiv
    On applying the ionization equation, Saha found that these alkali atoms would be almost fully ionized under the conditions of the solar atmosphere. A singly ...
  30. [30]
    [PDF] Photoionization Models for High Density Gas
    Three body recombination will affect the ionization balance of the plasma and the emitted recombination spectrum. As three body recombination occurs ...
  31. [31]
    Ion Chemistry in Atmospheric and Astrophysical Plasmas
    As the Universe cooled by adiabatic expansion, recombination occurred and molecular formation was driven by catalytic reactions involving the relict electrons ...
  32. [32]
    Thermal and chemical ionization in flames - SpringerLink
    In our experiments the rate of ion formation in the front of a hydrogen flame seeded with potassium exceeded the purely thermal ionization rate by 0.5–2 orders.
  33. [33]
    [PDF] Thermionic Energy Conversion in the Twenty-first Century
    Nov 8, 2017 · Thermionic energy conversion (TEC) is the direct conversion of heat into electricity by the mechanism of thermionic emission, ...
  34. [34]
    [PDF] IRVING LANGMUIR - National Academy of Sciences
    These studies included the discovery and detailed investigation of the formation of atomic hydrogen by contact of molecular hydrogen with a hot tungsten ...
  35. [35]
    [PDF] Chapter 3 Collisions in Plasmas
    Each electron in this distribution is losing momentum to the ions at a rate given by the collision frequency νp = ni q2 e q2 i. (4π 0). 2. 4π (me + mi) mim2.
  36. [36]
    [PDF] Photoionization of Atoms - UNL Digital Commons
    According to time-dependent perturbation theory, the photoionization cross section is proportional to the absolute square of the matrix element of (24.4).
  37. [37]
    Multiphoton Ionization (MPI) - an overview | ScienceDirect Topics
    Multiphoton ionization is defined as the process in which a molecule is ionized through the simultaneous absorption of two or more photons, ...Missing: formula | Show results with:formula
  38. [38]
    Molecular multiphoton ionization spectroscopy
    MPI of benzene in a supersonic jet demonstrates dramatic resolution improvement and shows that natural linewidth information can be gained for large molecules.
  39. [39]
    Above-threshold ionization - ScienceDirect.com
    ATI was discovered in 1979 and is now expected to be a universal phenomenon observable in all atomic species. The effect occurs readily in light pulses short ...
  40. [40]
    Thermal Desorption/Tunable Vacuum–Ultraviolet Time-of-Flight ...
    This paper describes thermal desorption/tunable vacuum–ultraviolet photoionization time-of-flight aerosol mass spectrometry (TD-VUV-TOF-PIAMS)
  41. [41]
    Resonance phenomena in molecular photoionization
    The effects of autoionization and shape resonances on the branching ratios and asymmetry parameters for several systems are discussed. The potential and current ...
  42. [42]
    Electron emission in intense electric fields - Journals
    At ordinary temperatures these currents are completely independent of the temperature. The formula for these current is I = Ce─a/F, (1) Which is, of course, ...
  43. [43]
    Scanning Field Emission Microscopy with Polarization Analysis ...
    Here we provide a comprehensive review of all strategies that have been adopted to perform the spin polarized experiments in the Fowler-Nordheim regime of STM.
  44. [44]
    Field emission ion source - US2809314A - Google Patents
    Fortunately, the gas which is most frequently used in ion sources for particle accelerators is also a Vgas which is capable of permeating most metals. This ...<|control11|><|separator|>
  45. [45]
    Field Dependent Avalanche Ionization Rates in Dielectrics
    Feb 24, 2009 · Collision-driven avalanche ionization is accepted to dominate laser dielectric interactions at pulse lengths > 100 fs [4, 5] . On this time ...Missing: induced | Show results with:induced
  46. [46]
    [PDF] Theory of avalanche ionization induced in transparent dielectrics by ...
    The main regularities of development of the electron avalanche are investigated on basis of the solution of the derived equation, and critical field values are ...
  47. [47]
    A classical model for multiple-electron capture in slow collisions of ...
    A new version of the classical over-barrier model for multiple-electron capture in slow collisions between ions and atoms is described. This version ...
  48. [48]
    Classical analysis of Coulomb effects in strong-field ionization of H - 2
    Sep 10, 2013 · We analyze the distortion of the molecular frame photoelectron angular distributions of H 2 + ionized by a strong, circularly polarized infrared laser field
  49. [49]
    Classical-Quantum Correspondence for Above-Threshold Ionization
    Mar 17, 2014 · We measure high resolution photoelectron angular distributions (PADs) for above-threshold ionization of xenon atoms in infrared laser fields.
  50. [50]
    Chaotic ionization of highly excited hydrogen atoms
    The experimental and theoretical studies of the microwave ionization of highly excited hydrogen atoms provide a unique opportunity to explore two new ...
  51. [51]
    Theory of Penning Ionization. I. Atoms - AIP Publishing
    Apr 1, 1970 · The theory of Penning ionization (PI) and related associative ionization (AI) is developed and examined in a classical, semiclassical, and
  52. [52]
    A semiclassical approach to collisional ionization with application to ...
    A semiclassical approach based on the propagation of classical trajectories on potential surfaces analytically continued into the complex plane, ...
  53. [53]
    On the theory of quantum mechanics | Proceedings of the Royal ...
    Bhattacharya K and Dalgarno A (1997) A modification of the Dirac time-dependent perturbation theory, Proceedings of the Royal Society of London. A ...
  54. [54]
    [PDF] Symphony on Strong Field Approximation
    Keldysh theory was further developed to calculate total ionization rates for various atomic species and states by M.V. Ammosov, N.B. Delone and V.P. Krainov ...
  55. [55]
    Strong-field approximation for intense-laser--atom processes
    The strong-field approximation (SFA) can be and has been applied in both length gauge and velocity gauge with quantitatively conflicting answers.
  56. [56]
    [PDF] soviet physics jetp - MIT
    Expressions are obtained for the probability of ionization of atoms and solid bodies in the field of a strong electromagnetic wave whose frequency is lower than ...
  57. [57]
    [PDF] Tunnel ionization of complex atoms and of atomic ions in an ...
    The paper derives an expression for the probability of tunnel ionization of complex atoms and atomic ions in an alternating field, using quasiclassical  ...
  58. [58]
    Theory of molecular tunneling ionization | Phys. Rev. A
    Abstract. We have extended the tunneling ionization model of Ammosov-Delone-Krainov (ADK) for atoms to diatomic molecules by considering the symmetry property ...Missing: dynamics seminal
  59. [59]
    First-principles calculations for the tunnel ionization rate of atoms ...
    May 7, 2004 · The Ammosov-Delone-Krainov (ADK) model [2] is based on the quasistatic approximation and has been applied for ionization of atoms with success.
  60. [60]
    Keldysh parameter, photoionization adiabaticity, and the tunneling ...
    Oct 17, 2016 · We also show that the ratio of the Keldysh γ to the frequency of the driver field indeed defines an important time scale of photoionization.Missing: dynamic | Show results with:dynamic
  61. [61]
    Reconciling Conflicting Approaches for the Tunneling Time Delay in ...
    Several recent attoclock experiments have investigated the fundamental question of a quantum mechanically induced time delay in tunneling ionization.
  62. [62]
    Full experimental determination of tunneling time with attosecond ...
    Jul 7, 2022 · We propose and demonstrate a scheme to experimentally determine the tunneling ionization time in an attoclock without any theoretical calculation.
  63. [63]
    Tunneling ionization of atoms and atomic ions in an intense laser ...
    It is shown that the space–time distribution of the laser field affects the yield of ions produced with tunneling. We derive an expression for the threshold ...Missing: adaptations | Show results with:adaptations
  64. [64]
    [PDF] Strong field double ionization: What is under the “knee”? - arXiv
    May 2, 2009 · Both uncorrelated (“sequential”) and correlated (“nonsequential”) processes contribute to the double ionization of the helium atom in strong ...
  65. [65]
    Excitation of Rydberg wave packets in the tunneling regime
    Oct 3, 2017 · To explain the origin of such trapping of population into Rydberg states, two mechanisms have been proposed: the first involves ac-Stark-shifted ...<|separator|>
  66. [66]
    Dynamics of N2 Dissociation upon Inner-Valence Ionization by ...
    Jan 13, 2015 · Here, we apply femtosecond time-resolved photoelectron and photoion spectroscopy to study dissociative ionization of nitrogen, the most abundant ...Missing: dication | Show results with:dication
  67. [67]
    Inner-valence states of N2+ and the dissociation dynamics studied ...
    Jun 16, 2006 · Dissociation products formed from the inner-valence N 2 + states are determined by threshold photoelectron-photoion coincidence spectroscopy.Missing: multiphoton dication seminal
  68. [68]
    Charge-resonance-enhanced ionization of diatomic molecular ions ...
    Oct 1, 1995 · We study the ionization of the H+2 molecular ion in intense, short-pulse laser fields by numerically solving the three-dimensional time-dependent Schrödinger ...Missing: original | Show results with:original
  69. [69]
    Charge Resonance Enhanced Ionization of C O 2 Probed by Laser ...
    Aug 2, 2011 · This process, called charge resonance enhanced ionization (CREI), is well understood for diatomic molecules, such as H 2 + / D 2 + , in terms of ...Missing: original paper
  70. [70]
    A remark on the Kramers-Henneberger transformation - ScienceDirect
    It is demonstrated that the Kramers-Henneberger transformation is the first step in a systematic construction of Floquet states. For a harmonic oscillator ...
  71. [71]
    Ionization of atoms in intense laser pulses using the Kramers ...
    We present a general technique for the numerical calculation of the behavior of an electron in an intense laser field.
  72. [72]
    Interference patterns in ionization of Kramers–Henneberger atom
    Oct 11, 2022 · We demonstrate that ionization of such an atom exhibits some molecular-like features such as low order interference maxima in photoelectron momentum spectra.
  73. [73]
    Electron Impact or Chemical Ionization for Mass Spectrometry - AZoM
    Aug 22, 2024 · Electron impact ionization delivers the highest energy to the analytes leading to 'hard' ionization and causes a large amount of molecular fragmentation.
  74. [74]
    Surface Analysis: X-ray Photoelectron Spectroscopy and Auger ...
    This fundamental review is on the subject of X-ray photoelectron spectroscopy (XPS), also called electron spectroscopy for chemical analysis (ESCA), and Auger ...
  75. [75]
    [PDF] ENDURANCE TEST OF THE MICRONEWTON FEEP THRUSTER
    The FEEP Thruster​​ FEEP (Field Emission Electric Propulsion) is an advanced electrostatic thruster capable of delivering very low thrust with very high accuracy ...<|separator|>
  76. [76]
    Future of plasma etching for microelectronics: Challenges and ...
    Jun 7, 2024 · Plasma etching is an essential semiconductor manufacturing technology required to enable the current microelectronics industry.
  77. [77]
    Radiation Therapy for Cancer - NCI
    May 15, 2025 · Radiation therapy is a type of cancer treatment that uses high doses of radiation to kill cancer cells and shrink tumors.Side Effects · External Beam Radiation · Brachytherapy
  78. [78]
    Residential Air Cleaners (Second Edition): A Summary of Available ...
    Sep 7, 2016 · Ozone-generating devices sold as air cleaners use UV light or corona discharge and are meant to control indoor air pollutants. Table 1 provides ...
  79. [79]
    10 Things to Know About the Ionosphere - NASA Science
    Dec 10, 2019 · Because it's formed when particles are ionized by the Sun's energy, the ionosphere changes from Earth's day side to night side.
  80. [80]
    Spectral Analysis - Imagine the Universe! - NASA
    Spectral analysis uses a spectrum, a chart of light intensity vs. energy, to determine elements, temperature, density, and magnetic fields in stars.<|separator|>
  81. [81]
    Coherent imaging of an attosecond electron wave packet - Science
    Jun 16, 2017 · Here we show, by using photoionization of neon, that a train of attosecond pulses synchronized with an infrared laser field can be used to ...
  82. [82]
    Press release: The Nobel Prize in Physics 2023 - NobelPrize.org
    Oct 3, 2023 · Attosecond physics gives us the opportunity to understand mechanisms that are governed by electrons. The next step will be utilising them,” says ...
  83. [83]
  84. [84]
    dissociation (D01801) - IUPAC Gold Book
    The separation of a molecular entity into two or more molecular entities (or any similar separation within a polyatomic molecular entity).
  85. [85]
    Hydrogen - the NIST WebBook
    Ionization energy determinations ; 15.431 ± 0.022, TE, Villarejo, 1968, RDSH ; 15.439 ± 0.015, PE, Collin and Natalis, 1968, RDSH.Gas phase ion energetics data · References
  86. [86]
    [PDF] Bond dissociation energies in simple molecules
    ... dissociation energies of the H2 , HD, and D2 Molecules, J. Mol. Spect., 5, 482 (1960). [32] Wieland, K.,. Bandensysteme B( 22+)—» X( 22+. ) und Dis ...
  87. [87]
    Predissociation - an overview | ScienceDirect Topics
    Predissociation is a dissociation mechanism where bound-state vibrational levels acquire some dissociative character due to interactions with a repulsive ...
  88. [88]
    Multiple ionization and Coulomb explosion of molecules, molecular ...
    Coulomb explosion, caused by high energy species, is used to determine molecular structures and trace ultrafast chemical reactions. It is also used to study ...
  89. [89]
  90. [90]
    NIST: Atomic Spectra Database - Ionization Energies Form
    This form provides access to NIST critically evaluated data on ground states and ionization energies of atoms and atomic ions.Missing: 1-118 | Show results with:1-118
  91. [91]
    Ground Levels and Ionization Energies for the Neutral Atoms | NIST
    This NIST database provides ground levels and ionization energies for neutral atoms, based on a literature survey, with some uncertainties. Ground levels are ...Missing: successive first
  92. [92]
    Calculated Ionization Energy for H 2 O (Water) - CCCBDB
    Experimental Ionization Energy is 12.621 ± 0.002 eV. Please note! These calculated ionizataion energies have the vibrational zero-point energy ( zpe ) included.
  93. [93]
    Water - the NIST WebBook
    Ionization energy determinations. IE (eV), Method, Reference, Comment. 12.65 ± 0.05, EI, Snow and Thomas, 1990, LL. 12.6188 ± 0.0009, PI, Page, Larkin, et al., ...Missing: adiabatic | Show results with:adiabatic
  94. [94]
    Calculated Ionization Energy for CO 2 (Carbon dioxide)
    Experimental Ionization Energy is 13.777 ± 0.001 eV. Please note! These calculated ionizataion energies have the vibrational zero-point energy ( zpe ) included.Missing: H2O | Show results with:H2O
  95. [95]
    Carbon dioxide - the NIST WebBook
    Ionization energy determinations ; 13.774 ± 0.003, PI, Parr and Taylor, 1974 ; 13.9 ± 0.2, EI, Semenov, Volkov, et al., 1973 ; 13.776 ± 0.008, PI, Parr and Taylor, ...Missing: H2O adiabatic