Tunnel ionization

Last updated

In physics, tunnel ionization is a process in which electrons in an atom (or a molecule) tunnel through the potential barrier and escape from the atom (or molecule). In an intense electric field, the potential barrier of an atom (molecule) is distorted drastically. Therefore, as the length of the barrier that electrons have to pass decreases, the electrons can escape from the atom's potential more easily. Tunneling ionization is a quantum mechanical phenomenon since in the classical picture an electron does not have sufficient energy to overcome the potential barrier of the atom.

Contents

When the atom is in a DC external field, the Coulomb potential barrier is lowered and the electron has an increased, non-zero probability of tunnelling through the potential barrier. In the case of an alternating electric field, the direction of the electric field reverses after the half period of the field. The ionized electron may come back to its parent ion. The electron may recombine with the nucleus (nuclei) and its kinetic energy is released as light (high harmonic generation). If the recombination does not occur, further ionization may proceed by collision between high-energy electrons and a parent atom (molecule). This process is known as non-sequential ionization. [1]

DC tunneling ionization

Tunneling ionization from the ground state of a hydrogen atom in an electrostatic (DC) field was solved schematically by Lev Landau, [2] using parabolic coordinates. This provides a simplified physical system that given it proper exponential dependence of the ionization rate on the applied external field. When , the ionization rate for this system is given by:

Landau expressed this in atomic units, where . In SI units the previous parameters can be expressed as:

,
.

The ionization rate is the total probability current through the outer classical turning point. This rate is found using the WKB approximation to match the ground state hydrogen wavefunction through the suppressed coulomb potential barrier.

A more physically meaningful form for the ionization rate above can be obtained by noting that the Bohr radius and hydrogen atom ionization energy are given by

,
,

where is the Rydberg energy. Then, the parameters and can be written as

, .

so that the total ionization rate can be rewritten

.

This form for the ionization rate emphasizes that the characteristic electric field needed for ionization is proportional to the ratio of the ionization energy to the characteristic size of the electron's orbital . Thus, atoms with low ionization energy (such as alkali metals) with electrons occupying orbitals with high principal quantum number (i.e. far down the periodic table) ionize most easily under a DC field. Furthermore, for a hydrogenic atom, the scaling of this characteristic ionization field goes as , where is the nuclear charge. This scaling arises because the ionization energy scales as and the orbital radius as . More accurate and general formulas for the tunneling from Hydrogen orbitals can also be obtained. [3]

As an empirical point of reference, the characteristic electric field for the ordinary hydrogen atom is about 51  V/Å (or 5.1×103 MV/cm) and the characteristic frequency is 4.1×104 THz.

AC electric field

The ionization rate of a hydrogen atom in an alternating electric field, like that of a laser, can be treated, in the appropriate limit, as the DC ionization rate averaged over a single period of the electric field's oscillation. Multiphoton and tunnel ionization of an atom or a molecule describes the same process by which a bounded electron, through the absorption of more than one photon from the laser field, is ionized. The difference between them is a matter of definition under different conditions. They can henceforth be called multiphoton ionization (MPI) whenever the distinction is not necessary. The dynamics of the MPI can be described by finding the time evolution of the state of the atom which is described by the Schrödinger equation.

Combined potential of an atom and a uniform laser field. At distances
r
<
r
0
{\displaystyle r<r_{0}}
, the potential of the laser can be neglected, while at distances with
r
>
r
0
{\displaystyle r>r_{0}}
, the Coulomb potential is negligible compared to the potential of the laser field. The electron emerges from under the barrier at
r
=
R
c
{\displaystyle r=R_{c}}
.
E
i
{\displaystyle E_{\text{i}}}
is the ionization potential of the atom. Tunnel ionization 3.png
Combined potential of an atom and a uniform laser field. At distances , the potential of the laser can be neglected, while at distances with , the Coulomb potential is negligible compared to the potential of the laser field. The electron emerges from under the barrier at . is the ionization potential of the atom.

When the intensity of the laser is strong, the lowest-order perturbation theory is not sufficient to describe the MPI process. In this case, the laser field on larger distances from the nucleus is more important than the Coulomb potential and the dynamic of the electron in the field should be properly taken into account. The first work in this category was published by Leonid Keldysh. [4] He modeled the MPI process as a transition of the electron from the ground state of the atom to the Volkov states (the state of a free electron in the electromagnetic field [5] ). In this model, the perturbation of the ground state by the laser field is neglected and the details of atomic structure in determining the ionization probability are not taken into account. The major difficulty with Keldysh's model was its neglect of the effects of Coulomb interaction on the final state of the electron. As is observed from the figure, the Coulomb field is not very small in magnitude compared to the potential of the laser at larger distances from the nucleus. This is in contrast to the approximation made by neglecting the potential of the laser at regions near the nucleus. A. M. Perelomov, V. S. Popov and M. V. Terent'ev [6] [7] included the Coulomb interaction at larger internuclear distances. Their model (which is called the PPT model after their initials) was derived for short-range potential and includes the effect of the long-range Coulomb interaction through the first-order correction in the quasi-classical action. In the quasi-static limit, the PPT model approaches the ADK model by M. V. Ammosov, N. B. Delone, and V. P. Krainov. [8]

Many experiments have been carried out on the MPI of rare gas atoms using strong laser pulses, through measuring both the total ion yield and the kinetic energy of the electrons. Here, one only considers the experiments designed to measure the total ion yield. Among these experiments are those by S. L. Chin et al., [9] S. Augst et al. [10] and T. Auguste et al. [11] Chin et al. used a 10.6 μm CO2 laser in their experiment. Due to the very small frequency of the laser, the tunneling is strictly quasi-static, a characteristic that is not easily attainable using pulses in the near infrared or visible region of frequencies. These findings weakened the suspicion on the applicability of models basically founded on the assumption of a structureless atom. S. Larochelle et al. [12] have compared the theoretically predicted ion versus intensity curves of rare gas atoms interacting with a Ti:sapphire laser with experimental measurement. They have shown that the total ionization rate predicted by the PPT model fits very well the experimental ion yields for all rare gases in the intermediate regime of Keldysh parameter.

Analytical formula for the rate of MPI

The dynamics of the MPI can be described by finding the time evolution of the state of the atom which is described by the Schrödinger equation. The form of this equation in the electric field gauge, assuming the single active electron (SAE) approximation and using dipole approximation, is the following

where is the electric field of the laser and is the static Coulomb potential of the atomic core at the position of the active electron. By finding the exact solution of equation (1) for a potential ( the magnitude of the ionization potential of the atom), the probability current is calculated. Then, the total MPI rate from short-range potential for linear polarization, , is found from

where is the frequency of the laser, which is assumed to be polarized in the direction of the axis. The effect of the ionic potential, which behaves like ( is the charge of atomic or ionic core) at a long distance from the nucleus, is calculated through first order correction on the semi-classical action. The result is that the effect of ionic potential is to increase the rate of MPI by a factor of

Where and is the peak electric field of laser. Thus, the total rate of MPI from a state with quantum numbers and in a laser field for linear polarization is calculated to be

where is the Keldysh's adiabaticity parameter and . The coefficients , and are given by

The coefficient is given by

,

where

The ADK model is the limit of the PPT model when approaches zero (quasi-static limit). In this case, which is known as quasi-static tunnelling (QST), the ionization rate is given by

.

In practice, the limit for the QST regime is . This is justified by the following consideration. [13] Referring to the figure, the ease or difficulty of tunneling can be expressed as the ratio between the equivalent classical time it takes for the electron to tunnel out the potential barrier while the potential is bent down. This ratio is indeed , since the potential is bent down during half a cycle of the field oscillation and the ratio can be expressed as

,

where is the tunneling time (classical time of flight of an electron through a potential barrier, and is the period of laser field oscillation.

MPI of molecules

Contrary to the abundance of theoretical and experimental work on the MPI of rare gas atoms, the amount of research on the prediction of the rate of MPI of neutral molecules was scarce until recently. Walsh et al. [14] have measured the MPI rate of some diatomic molecules interacting with a 10.6 μm CO2 laser. They found that these molecules are tunnel-ionized as if they were structureless atoms with an ionization potential equivalent to that of the molecular ground state. A. Talebpour et al. [15] [16] were able to quantitatively fit the ionization yield of diatomic molecules interacting with a Ti:sapphire laser pulse. The conclusion of the work was that the MPI rate of a diatomic molecule can be predicted from the PPT model by assuming that the electron tunnels through a barrier given by instead of barrier which is used in the calculation of the MPI rate of atoms. The importance of this finding is in its practicality; the only parameter needed for predicting the MPI rate of a diatomic molecule is a single parameter, . Using the semi-empirical model for the MPI rate of unsaturated hydrocarbons is feasible. [17] This simplistic view ignores the ionization dependence on orientation of molecular axis with respect to polarization of the electric field of the laser, which is determined by the symmetries of the molecular orbitals. This dependence can be used to follow molecular dynamics using strong field multiphoton ionization. [18]

Tunneling time

The question of how long a tunneling particle spends inside the barrier region has remained unresolved since the early days of quantum mechanics. It is sometimes suggested that the tunneling time is instantaneous because both the Keldysh and the closely related Buttiker-Landauer [19] times are imaginary (corresponding to the decay of the wavefunction under the barrier). In a recent publication [20] the main competing theories of tunneling time are compared against experimental measurements using the attoclock in strong laser field ionization of helium atoms. Refined attoclock measurements reveal a real and not instantaneous tunneling delay time over a large intensity regime. It is found that the experimental results are compatible with the probability distribution of tunneling times constructed using a Feynman path integral (FPI) formulation. [21] [22] However, later work in atomic hydrogen has demonstrated that most of the tunneling time measured in the experiment is purely from the long-range Coulomb force exerted by the ion core on the outgoing electron. [23]

Reference s

  1. Corkum, P. B. (1993-09-27). "Plasma perspective on strong field multiphoton ionization". Physical Review Letters. 71 (13). American Physical Society (APS): 1994–1997. doi:10.1103/physrevlett.71.1994. ISSN   0031-9007. PMID   10054556. S2CID   29947935.
  2. L.D. Landau and E.M. Lifshitz, Quantum Mechanics (Pergamon, New York, 1965), 2nd ed., pg 276.
  3. Yamabe, Tokio; Tachibana, Akitomo; Silverstone, Harris J. (1977-09-01). "Theory of the ionization of the hydrogen atom by an external electrostatic field". Physical Review A. 16 (3): 877–890. doi:10.1103/PhysRevA.16.877.
  4. Keldysh L V 1965 Soviet Phys. JETP 2354
  5. Wolkow, D. M. (1935). "Über eine Klasse von Lösungen der Diracschen Gleichung". Zeitschrift für Physik (in German). 94 (3–4). Springer Science and Business Media LLC: 250–260. doi:10.1007/bf01331022. ISSN   1434-6001. S2CID   123046147.
  6. Perelemov A M, Popov V S and Terent'ev M V 1966 SovietPhys. JETP, 23 924
  7. Perelemov A M and Popov V S 1967 Soviet Phys.JETP, 25 336
  8. Ammosov M V, Delone N B and Krainov V P 1986 SovietPhys. JETP, 64 1191
  9. Chin, S L; Yergeau, F; Lavigne, P (1985-04-28). "Tunnel ionisation of Xe in an ultra-intense CO2 laser field (1014 W cm−2) with multiple charge creation". Journal of Physics B: Atomic and Molecular Physics. 18 (8). IOP Publishing: L213–L215. doi:10.1088/0022-3700/18/8/001. ISSN   0022-3700.
  10. Augst, S.; Meyerhofer, D. D.; Strickland, D.; Chint, S. L. (1991-04-01). "Laser ionization of noble gases by Coulomb-barrier suppression". Journal of the Optical Society of America B. 8 (4). The Optical Society: 858. doi:10.1364/josab.8.000858. ISSN   0740-3224.
  11. Auguste, T; Monot, P; Lompre, L A; Mainfray, G; Manus, C (1992-10-28). "Multiply charged ions produced in noble gases by a 1 ps laser pulse at lambda = 1053 nm". Journal of Physics B: Atomic, Molecular and Optical Physics. 25 (20). IOP Publishing: 4181–4194. doi:10.1088/0953-4075/25/20/015. ISSN   0953-4075. S2CID   250751215.
  12. Larochelle, S; Talebpour, A; Chin, S L (1998-03-28). "Non-sequential multiple ionization of rare gas atoms in a Ti:Sapphire laser field". Journal of Physics B: Atomic, Molecular and Optical Physics. 31 (6). IOP Publishing: 1201–1214. Bibcode:1998JPhB...31.1201L. doi:10.1088/0953-4075/31/6/008. ISSN   0953-4075. S2CID   250747225.
  13. CHIN, S. L. (2004). "From multiphoton to tunnel ionization". Advances in Multi-Photon Processes and Spectroscopy. Vol. 16. WORLD SCIENTIFIC. pp. 249–271. doi:10.1142/9789812796585_0003. ISBN   978-981-256-031-5. ISSN   0218-0227.
  14. Walsh, T D G; Decker, J E; Chin, S L (1993-02-28). "Tunnel ionization of simple molecules by an intense CO2 laser". Journal of Physics B: Atomic, Molecular and Optical Physics. 26 (4). IOP Publishing: L85–L90. doi:10.1088/0953-4075/26/4/002. ISSN   0953-4075. S2CID   250888196.
  15. Talebpour, A; Larochelle, S; Chin, S L (1998-01-28). "Suppressed tunnelling ionization of the molecule in an intense Ti:sapphire laser pulse". Journal of Physics B: Atomic, Molecular and Optical Physics. 31 (2). IOP Publishing: L49–L58. doi:10.1088/0953-4075/31/2/003. ISSN   0953-4075. S2CID   250791262.
  16. Talebpour, A.; Yang, J.; Chin, S.L. (1999). "Semi-empirical model for the rate of tunnel ionization of N2 and O2 molecule in an intense Ti:sapphire laser pulse". Optics Communications. 163 (1–3). Elsevier BV: 29–32. doi:10.1016/s0030-4018(99)00113-3. ISSN   0030-4018.
  17. Talebpour, A; Larochelle, S; Chin, S L (1998-06-28). "Multiphoton ionization of unsaturated hydrocarbons". Journal of Physics B: Atomic, Molecular and Optical Physics. 31 (12). IOP Publishing: 2769–2776. doi:10.1088/0953-4075/31/12/012. ISSN   0953-4075. S2CID   250797867.
  18. Jaron-Becker, A. (2012). "Molecular Dynamics in Strong Laser Fields". IEEE Journal of Selected Topics in Quantum Electronics. 18 (1). Institute of Electrical and Electronics Engineers (IEEE): 105–112. doi:10.1109/jstqe.2011.2108271. ISSN   1077-260X. S2CID   16703524.
  19. Büttiker, M.; Landauer, R. (1982-12-06). "Traversal Time for Tunneling". Physical Review Letters. 49 (23). American Physical Society (APS): 1739–1742. doi:10.1103/physrevlett.49.1739. ISSN   0031-9007.
  20. Landsman, Alexandra S.; Weger, Matthias; Maurer, Jochen; Boge, Robert; Ludwig, André; et al. (2014-11-14). "Ultrafast resolution of tunneling delay time". Optica. 1 (5). The Optical Society: 343. arXiv: 1301.2766 . doi: 10.1364/optica.1.000343 . ISSN   2334-2536.
  21. Fertig, H. A. (1990-11-05). "Traversal-Time Distribution and the Uncertainty Principle in Quantum Tunneling". Physical Review Letters. 65 (19). American Physical Society (APS): 2321–2324. doi:10.1103/physrevlett.65.2321. ISSN   0031-9007. PMID   10042518.
  22. Yamada, Norifumi (2004-10-18). "Unified Derivation of Tunneling Times from Decoherence Functionals". Physical Review Letters. 93 (17). American Physical Society (APS): 170401. doi:10.1103/physrevlett.93.170401. ISSN   0031-9007. PMID   15525052.
  23. Sainadh, U. Satya; Xu, Han; Wang, Xiaoshan; Atia-Tul-Noor, A.; Wallace, William C.; Douguet, Nicolas; Bray, Alexander; Ivanov, Igor; Bartschat, Klaus; Kheifets, Anatoli; Sang, R. T.; Litvinyuk, I. V. (April 2019). "Attosecond angular streaking and tunnelling time in atomic hydrogen". Nature. 568 (7750): 75–77. doi:10.1038/s41586-019-1028-3. hdl: 10072/387846 . ISSN   1476-4687. PMID   30886392. S2CID   81977455.

Further reading

Related Research Articles

Spontaneous emission is the process in which a quantum mechanical system transits from an excited energy state to a lower energy state and emits a quantized amount of energy in the form of a photon. Spontaneous emission is ultimately responsible for most of the light we see all around us; it is so ubiquitous that there are many names given to what is essentially the same process. If atoms are excited by some means other than heating, the spontaneous emission is called luminescence. For example, fireflies are luminescent. And there are different forms of luminescence depending on how excited atoms are produced. If the excitation is effected by the absorption of radiation the spontaneous emission is called fluorescence. Sometimes molecules have a metastable level and continue to fluoresce long after the exciting radiation is turned off; this is called phosphorescence. Figurines that glow in the dark are phosphorescent. Lasers start via spontaneous emission, then during continuous operation work by stimulated emission.

In particle physics, the Dirac equation is a relativistic wave equation derived by British physicist Paul Dirac in 1928. In its free form, or including electromagnetic interactions, it describes all spin-1/2 massive particles, called "Dirac particles", such as electrons and quarks for which parity is a symmetry. It is consistent with both the principles of quantum mechanics and the theory of special relativity, and was the first theory to account fully for special relativity in the context of quantum mechanics. It was validated by accounting for the fine structure of the hydrogen spectrum in a completely rigorous way.

<span class="mw-page-title-main">Bremsstrahlung</span> Electromagnetic radiation due to deceleration of charged particles

In particle physics, bremsstrahlung is electromagnetic radiation produced by the deceleration of a charged particle when deflected by another charged particle, typically an electron by an atomic nucleus. The moving particle loses kinetic energy, which is converted into radiation, thus satisfying the law of conservation of energy. The term is also used to refer to the process of producing the radiation. Bremsstrahlung has a continuous spectrum, which becomes more intense and whose peak intensity shifts toward higher frequencies as the change of the energy of the decelerated particles increases.

<span class="mw-page-title-main">Ionization</span> Process by which atoms or molecules acquire charge by gaining or losing electrons

Ionization is the process by which an atom or a molecule acquires a negative or positive charge by gaining or losing electrons, often in conjunction with other chemical changes. The resulting electrically charged atom or molecule is called an ion. Ionization can result from the loss of an electron after collisions with subatomic particles, collisions with other atoms, molecules, electrons, positrons, protons, antiprotons and ions, or through the interaction with electromagnetic radiation. Heterolytic bond cleavage and heterolytic substitution reactions can result in the formation of ion pairs. Ionization can occur through radioactive decay by the internal conversion process, in which an excited nucleus transfers its energy to one of the inner-shell electrons causing it to be ejected.

Matter waves are a central part of the theory of quantum mechanics, being half of wave–particle duality. At all scales where measurements have been practical, matter exhibits wave-like behavior. For example, a beam of electrons can be diffracted just like a beam of light or a water wave.

In physics, the Hamilton–Jacobi equation, named after William Rowan Hamilton and Carl Gustav Jacob Jacobi, is an alternative formulation of classical mechanics, equivalent to other formulations such as Newton's laws of motion, Lagrangian mechanics and Hamiltonian mechanics.

In quantum physics, Fermi's golden rule is a formula that describes the transition rate from one energy eigenstate of a quantum system to a group of energy eigenstates in a continuum, as a result of a weak perturbation. This transition rate is effectively independent of time and is proportional to the strength of the coupling between the initial and final states of the system as well as the density of states. It is also applicable when the final state is discrete, i.e. it is not part of a continuum, if there is some decoherence in the process, like relaxation or collision of the atoms, or like noise in the perturbation, in which case the density of states is replaced by the reciprocal of the decoherence bandwidth.

In physics, the gyromagnetic ratio of a particle or system is the ratio of its magnetic moment to its angular momentum, and it is often denoted by the symbol γ, gamma. Its SI unit is the radian per second per tesla (rad⋅s−1⋅T−1) or, equivalently, the coulomb per kilogram (C⋅kg−1).

In quantum physics, the spin–orbit interaction is a relativistic interaction of a particle's spin with its motion inside a potential. A key example of this phenomenon is the spin–orbit interaction leading to shifts in an electron's atomic energy levels, due to electromagnetic interaction between the electron's magnetic dipole, its orbital motion, and the electrostatic field of the positively charged nucleus. This phenomenon is detectable as a splitting of spectral lines, which can be thought of as a Zeeman effect product of two relativistic effects: the apparent magnetic field seen from the electron perspective and the magnetic moment of the electron associated with its intrinsic spin. A similar effect, due to the relationship between angular momentum and the strong nuclear force, occurs for protons and neutrons moving inside the nucleus, leading to a shift in their energy levels in the nucleus shell model. In the field of spintronics, spin–orbit effects for electrons in semiconductors and other materials are explored for technological applications. The spin–orbit interaction is at the origin of magnetocrystalline anisotropy and the spin Hall effect.

<span class="mw-page-title-main">Larmor formula</span> Gives the total power radiated by an accelerating, nonrelativistic point charge

In electrodynamics, the Larmor formula is used to calculate the total power radiated by a nonrelativistic point charge as it accelerates. It was first derived by J. J. Larmor in 1897, in the context of the wave theory of light.

The theoretical and experimental justification for the Schrödinger equation motivates the discovery of the Schrödinger equation, the equation that describes the dynamics of nonrelativistic particles. The motivation uses photons, which are relativistic particles with dynamics described by Maxwell's equations, as an analogue for all types of particles.

In spectroscopy, the Autler–Townes effect, is a dynamical Stark effect corresponding to the case when an oscillating electric field is tuned in resonance to the transition frequency of a given spectral line, and resulting in a change of the shape of the absorption/emission spectra of that spectral line. The AC Stark effect was discovered in 1955 by American physicists Stanley Autler and Charles Townes.

Resonance fluorescence is the process in which a two-level atom system interacts with the quantum electromagnetic field if the field is driven at a frequency near to the natural frequency of the atom.

A hydrogen-like atom (or hydrogenic atom) is any atom or ion with a single valence electron. These atoms are isoelectronic with hydrogen. Examples of hydrogen-like atoms include, but are not limited to, hydrogen itself, all alkali metals such as Rb and Cs, singly ionized alkaline earth metals such as Ca+ and Sr+ and other ions such as He+, Li2+, and Be3+ and isotopes of any of the above. A hydrogen-like atom includes a positively charged core consisting of the atomic nucleus and any core electrons as well as a single valence electron. Because helium is common in the universe, the spectroscopy of singly ionized helium is important in EUV astronomy, for example, of DO white dwarf stars.

The Kapitza–Dirac effect is a quantum mechanical effect consisting of the diffraction of matter by a standing wave of light. The effect was first predicted as the diffraction of electrons from a standing wave of light by Paul Dirac and Pyotr Kapitsa in 1933. The effect relies on the wave–particle duality of matter as stated by the de Broglie hypothesis in 1924.

Surface-extended X-ray absorption fine structure (SEXAFS) is the surface-sensitive equivalent of the EXAFS technique. This technique involves the illumination of the sample by high-intensity X-ray beams from a synchrotron and monitoring their photoabsorption by detecting in the intensity of Auger electrons as a function of the incident photon energy. Surface sensitivity is achieved by the interpretation of data depending on the intensity of the Auger electrons instead of looking at the relative absorption of the X-rays as in the parent method, EXAFS.

An electric dipole transition is the dominant effect of an interaction of an electron in an atom with the electromagnetic field.

Heat transfer physics describes the kinetics of energy storage, transport, and energy transformation by principal energy carriers: phonons, electrons, fluid particles, and photons. Heat is thermal energy stored in temperature-dependent motion of particles including electrons, atomic nuclei, individual atoms, and molecules. Heat is transferred to and from matter by the principal energy carriers. The state of energy stored within matter, or transported by the carriers, is described by a combination of classical and quantum statistical mechanics. The energy is different made (converted) among various carriers. The heat transfer processes are governed by the rates at which various related physical phenomena occur, such as the rate of particle collisions in classical mechanics. These various states and kinetics determine the heat transfer, i.e., the net rate of energy storage or transport. Governing these process from the atomic level to macroscale are the laws of thermodynamics, including conservation of energy.

In quantum mechanics, magnetic resonance is a resonant effect that can appear when a magnetic dipole is exposed to a static magnetic field and perturbed with another, oscillating electromagnetic field. Due to the static field, the dipole can assume a number of discrete energy eigenstates, depending on the value of its angular momentum (azimuthal) quantum number. The oscillating field can then make the dipole transit between its energy states with a certain probability and at a certain rate. The overall transition probability will depend on the field's frequency and the rate will depend on its amplitude. When the frequency of that field leads to the maximum possible transition probability between two states, a magnetic resonance has been achieved. In that case, the energy of the photons composing the oscillating field matches the energy difference between said states. If the dipole is tickled with a field oscillating far from resonance, it is unlikely to transition. That is analogous to other resonant effects, such as with the forced harmonic oscillator. The periodic transition between the different states is called Rabi cycle and the rate at which that happens is called Rabi frequency. The Rabi frequency should not be confused with the field's own frequency. Since many atomic nuclei species can behave as a magnetic dipole, this resonance technique is the basis of nuclear magnetic resonance, including nuclear magnetic resonance imaging and nuclear magnetic resonance spectroscopy.

<span class="mw-page-title-main">Lorentz oscillator model</span> Theoretical model describing the optical response of bound charges

The Lorentz oscillator model describes the optical response of bound charges. The model is named after the Dutch physicist Hendrik Antoon Lorentz. It is a classical, phenomenological model for materials with characteristic resonance frequencies for optical absorption, e.g. ionic and molecular vibrations, interband transitions (semiconductors), phonons, and collective excitations.