Radiation trapping

Last updated

Radiation trapping, imprisonment of resonance radiation, radiative transfer of spectral lines, line transfer or radiation diffusion is a phenomenon in physics whereby radiation may be "trapped" in a system as it is emitted by one atom and absorbed by another. [1] [2]

Contents

Classical description

Classically, one can think of radiation trapping as a multiple-scattering phenomena, where a photon is scattered by multiple atoms in a cloud. This motivates treatment as a diffusion problem. As such, one can primarily consider the mean free path of light, defined as the reciprocal of the density of scatterers and the scattering cross section:

One can assume for simplicity that the scattering diagram is isotropic, which ends up being a good approximation for atoms with equally populated sublevels of total angular momentum. In the classical limit, we can think of the electromagnetic energy density as what is being diffused. So, we consider the diffusion constant in three dimensions,

where is the transport time. [3] The transport time accounts for both the group delay between scattering events and Wigner's delay time, which is associated with an elastic scattering process. [4] It is written as

where is the group velocity. When the photons are near resonance, the lifetime of an excited state in the atomic vapor is equal to the transport time, , independent of the detuning. [5] This comes in handy, since the average number of scattering events is the ratio of the time spent in the system to the lifetime of the excited state (or equivalently, the scattering time). Since in a 3D diffusion process the electromagnetic energy density spreads as , we can find the average number of scattering events for a photon before it escapes:

Finally, the number of scattering events can be related to the optical depth as follows. Since , the number of scattering events scales with the square of the optical depth. [6]

Derivation of the Holstein equation

In 1947, Theodore Holstein attacked the problem of imprisonment of resonance radiation in a novel way. Foregoing the classical method presented in the prior section, Holstein asserted that there could not exist a mean free path for the photons. His treatment begins with the introduction of a probability function , which describes the probability that a photon emitted at is absorbed within the volume element about the point . Additionally, one can enforce atom number conservation to write

where represent the number increase and decrease in population of excited atoms, and is the number density of excited atoms. If the reciprocal lifetime of an excited atom is given by , then is given by

Then is obtained by considering all other volume elements, which is where the introduction of becomes useful. The contribution of an outside volume to the number of excited atoms is given by the number of photons emitted by that outside volume multiplied by the probability that those photons are absorbed within the volume . Integration over all outside volume elements yields

Substituting and into the particle conservation law, we arrive at an integral equation for the density of excited atoms  the Holstein equation [7]

Finding the escape probability of photons from the Holstein equation

Now to find the escape probability of the photons, we consider solutions by ansatz of the form

Observing the Holstein equation, one can note that these solutions are subject to the constraint

Aided by the exchange symmetry of , namely that , one can use variational methods to assert that leads to

Completing the square and introducing the escape probability , whose definition follows from that all particles must either be absorbed or escape with a summed probability of 1, an equation in terms of the escape probability is derived:

Numerical methods for solving the Holstein equation

Many contemporary studies in atomic physics utilize numerical solutions to Holstein's equation to both show the presence of radiation trapping in their experimental system and to discuss its effects on the atomic spectra. Radiation trapping has been observed in a variety of experiments, including in the trapping of cesium atoms in a magneto-optical trap (MOT), in the spectroscopic characterization of dense Rydberg gases of strontium atoms, and in lifetime analyses of doped ytterbium(III) oxide for laser improvement. [8] [9] [10]

To solve or simulate the Holstein equation, the Monte Carlo method is commonly employed. An absorption coefficient is calculated for an experiment with a certain opacity, atomic species, Doppler-broadened lineshape, etc., and then a test is made to see whether the photon escapes after flights through the atomic vapor (see Figure 1 in the reference). [11]

Other methods include transforming the Holstein equation into a linear generalized eigenvalue problem, which is more computationally expensive and requires the usage of several simplifying assumptions, including but not limited to that the lowest eigenmode of the Holstein equation is parabolic in shape, the atomic vapor is spherical, the atomic vapor has reached a steady state after the near-resonant laser has been shut off, etc. [8]

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.

<span class="mw-page-title-main">Stimulated emission</span> Release of a photon triggered by another

Stimulated emission is the process by which an incoming photon of a specific frequency can interact with an excited atomic electron, causing it to drop to a lower energy level. The liberated energy transfers to the electromagnetic field, creating a new photon with a frequency, polarization, and direction of travel that are all identical to the photons of the incident wave. This is in contrast to spontaneous emission, which occurs at a characteristic rate for each of the atoms/oscillators in the upper energy state regardless of the external electromagnetic field.

<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 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.

<span class="mw-page-title-main">Fokker–Planck equation</span> Partial differential equation

In statistical mechanics and information theory, the Fokker–Planck equation is a partial differential equation that describes the time evolution of the probability density function of the velocity of a particle under the influence of drag forces and random forces, as in Brownian motion. The equation can be generalized to other observables as well. The Fokker-Planck equation has multiple applications in information theory, graph theory, data science, finance, economics etc.

In physics, mean free path is the average distance over which a moving particle travels before substantially changing its direction or energy, typically as a result of one or more successive collisions with other particles.

The adiabatic theorem is a concept in quantum mechanics. Its original form, due to Max Born and Vladimir Fock (1928), was stated as follows:

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.

Tunnel ionization is a process in which electrons in an atom pass through the potential barrier and escape from the atom. 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.

<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.

<span class="mw-page-title-main">Electron scattering</span> Deviation of electrons from their original trajectories

Electron scattering occurs when electrons are displaced from their original trajectory. This is due to the electrostatic forces within matter interaction or, if an external magnetic field is present, the electron may be deflected by the Lorentz force. This scattering typically happens with solids such as metals, semiconductors and insulators; and is a limiting factor in integrated circuits and transistors.

In the physics of electromagnetism, the Abraham–Lorentz force is the recoil force on an accelerating charged particle caused by the particle emitting electromagnetic radiation by self-interaction. It is also called the radiation reaction force, the radiation damping force, or the self-force. It is named after the physicists Max Abraham and Hendrik Lorentz.

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.

<span class="mw-page-title-main">Struve function</span>

In mathematics, the Struve functionsHα(x), are solutions y(x) of the non-homogeneous Bessel's differential equation:

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.

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 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">Phase contrast magnetic resonance imaging</span>

Phase contrast magnetic resonance imaging (PC-MRI) is a specific type of magnetic resonance imaging used primarily to determine flow velocities. PC-MRI can be considered a method of Magnetic Resonance Velocimetry. It also provides a method of magnetic resonance angiography. Since modern PC-MRI is typically time-resolved, it provides a means of 4D imaging.

References

  1. Primer on Collisions
    • Molisch, Andreas F.; Oehry, Bernard P. (1998), Radiation Trapping in Atomic Vapours, Oxford: Oxford University Press, ISBN   0-19-853866-9 , retrieved June 18, 2006.
  2. van Rossum, M. C. W.; Nieuwenhuizen, Th. M. (1999-01-01). "Multiple scattering of classical waves: microscopy, mesoscopy, and diffusion". Reviews of Modern Physics. 71 (1): 313–371. arXiv: cond-mat/9804141 . Bibcode:1999RvMP...71..313V. doi:10.1103/RevModPhys.71.313. S2CID   119044791.
  3. Wigner, E. P. (1954-04-01). "The Problem of Multiple Scattering". Physical Review. 94 (1): 17–25. Bibcode:1954PhRv...94...17W. doi:10.1103/PhysRev.94.17.
  4. Labeyrie, G.; Vaujour, E.; Müller, C. A.; Delande, D.; Miniatura, C.; Wilkowski, D.; Kaiser, R. (2003-11-26). "Slow Diffusion of Light in a Cold Atomic Cloud". Physical Review Letters. 91 (22): 223904. Bibcode:2003PhRvL..91v3904L. doi:10.1103/PhysRevLett.91.223904. PMID   14683240.
  5. Weiss, Patrizia; Araújo, Michelle O.; Kaiser, Robin; Guerin, William (2018-06-15). "Subradiance and radiation trapping in cold atoms". New Journal of Physics. 20 (6): 063024. arXiv: 1803.01646 . Bibcode:2018NJPh...20f3024W. doi: 10.1088/1367-2630/aac5d0 . ISSN   1367-2630.
  6. Holstein, T. (1947-12-15). "Imprisonment of Resonance Radiation in Gases". Physical Review. 72 (12): 1212–1233. Bibcode:1947PhRv...72.1212H. doi:10.1103/PhysRev.72.1212.
  7. 1 2 Fioretti, A; Molisch, A. F; Müller, J. H; Verkerk, P; Allegrini, M (1998-04-15). "Observation of radiation trapping in a dense Cs magneto-optical trap". Optics Communications. 149 (4): 415–422. Bibcode:1998OptCo.149..415F. doi:10.1016/S0030-4018(97)00704-9. ISSN   0030-4018.
  8. Sadler, D. P.; Bridge, E. M.; Boddy, D.; Bounds, A. D.; Keegan, N. C.; Lochead, G.; Jones, M. P. A.; Olmos, B. (2017-01-24). "Radiation trapping in a dense cold Rydberg gas". Physical Review A. 95 (1): 013839. arXiv: 1607.07767 . Bibcode:2017PhRvA..95a3839S. doi:10.1103/PhysRevA.95.013839. hdl: 10072/393488 . S2CID   56448828.
  9. Auzel, F.; Baldacchini, G.; Laversenne, L.; Boulon, G. (2003-10-01). "Radiation trapping and self-quenching analysis in Yb3+, Er3+, and Ho3+ doped Y2O3". Optical Materials. Proceedings of the Fifth French-Israeli Workshop on Optical Properties of Inorganic Materials. 24 (1): 103–109. Bibcode:2003OptMa..24..103A. doi:10.1016/S0925-3467(03)00112-5. ISSN   0925-3467.
  10. Wiorkowski, P.; Hartmann, W. (1985-03-15). "Investigation of radiation imprisonment: Application to time resolved fluorescence spectroscopy". Optics Communications. 53 (4): 217–220. Bibcode:1985OptCo..53..217W. doi:10.1016/0030-4018(85)90158-0. ISSN   0030-4018.