Fact-checked by Grok 2 weeks ago

Thomson scattering

Thomson scattering is the of by a free , such as an , in the classical low-energy limit where the is much less than the rest mass energy of the (approximately 0.511 MeV). In this process, the of the incident wave exerts a on the particle, causing it to oscillate and re-emit radiation at the same frequency but in a different direction, with no net energy transfer to the particle. The scattering is coherent and independent of the incident wavelength under these conditions, distinguishing it from inelastic processes like that occur at higher energies. The theoretical foundation of Thomson scattering was developed by J. J. Thomson in 1906, who calculated the scattering cross-section using classical electrodynamics shortly after his discovery of the electron. The total Thomson cross-section for an electron is a constant value of \sigma_T = \frac{8\pi}{3} r_e^2 \approx 6.65 \times 10^{-25} cm², where r_e is the classical electron radius (approximately 2.82 × 10^{-13} cm). The differential cross-section depends on the scattering angle \theta and polarization, given by \frac{d\sigma}{d\Omega} = \frac{r_e^2}{2} (1 + \cos^2 \theta) for unpolarized light, leading to forward-backward symmetry and partial polarization of the scattered radiation perpendicular to the incident direction. This classical description holds for non-relativistic particles and when the particle's motion during one wave cycle is much smaller than the wavelength, but it breaks down at higher energies where quantum effects dominate, transitioning to the Klein-Nishina regime. Thomson scattering plays a crucial role in plasma physics as a non-invasive diagnostic tool for measuring electron density and temperature, with laser-based implementations providing precise, first-principles data in laboratory experiments. In astrophysics, it governs the propagation of radiation through ionized media, contributing to phenomena such as the polarization of light in reflection nebulae, light echoes around supernovae, and the damping of anisotropies in the cosmic microwave background. For heavier particles like protons, the cross-section is significantly smaller (by a factor of about (m_e / m_p)^2 \approx (1/1836)^2 \approx 3 \times 10^{-7}), making electron scattering the dominant process in most astrophysical and laboratory plasmas.

Fundamentals

Definition and Scope

Thomson scattering is the of by a free , such as an , where the scattered radiation has the same as the incident radiation. This process occurs when the h\nu is much smaller than the rest mass energy of the , m_e c^2 \approx 511 keV, treating the electron as quasi-free and non-relativistic. In the basic mechanism, the electric field of the incident electromagnetic wave exerts a Lorentz force on the charged particle, accelerating it and inducing oscillatory motion. The oscillating particle then acts as a dipole radiator, re-emitting electromagnetic radiation at the same frequency as the incident wave, with no net energy transfer to the particle. This classical description relies on electrodynamics and assumes coherent scattering without significant quantum recoil effects. The regime of applicability requires that the scattering be describable by classical electrodynamics, which holds when the wavelength is much larger than the Compton wavelength \lambda_C = h / (m_e c) \approx 2.426 \times 10^{-12} m, ensuring the interaction is non-relativistic and the electron's velocity remains much less than the . Deviations occur at higher energies, where quantum effects become prominent. The phenomenon is named after J. J. Thomson, who first derived its classical form in 1906 while investigating the scattering of Röntgen radiation by gases, providing an early theoretical framework for radiation interactions with free charges. This work was pivotal in the development of scattering theories preceding .

Classical Electron Model

In the classical electron model of Thomson scattering, the is treated as a point charge of magnitude e and mass m_e, undergoing non-relativistic motion without or quantum mechanical effects. The incident , modeled as a with \mathbf{E}, drives the electron's motion through the \mathbf{F} = -e \mathbf{E}, assuming the contribution is negligible for low velocities. This accelerates the , causing it to oscillate harmonically at the of the incident wave, with the displacement proportional to E / (m_e \omega^2), where \omega is the . The resulting generates dipole radiation, as the oscillating charge acts as an \mathbf{p} = -e \mathbf{r}, re-radiating energy in all directions except along the axis. For unpolarized incident , the treatment averages over the two orthogonal polarizations, yielding an effective driving . The scattered power exhibits an angular dependence proportional to \sin^2 \theta, where \theta is between the incident propagation direction and the observation point relative to the acceleration vector. Polarization plays a key role: linearly polarized incident produces scattered with polarization aligned to the incident case, while unpolarized input results in partially polarized scattered , preferentially perpendicular to the scattering plane, with the degree of polarization given by (1 - \cos^2 \phi)/(1 + \cos^2 \phi), where \phi is the scattering . As an elastic process, Thomson scattering conserves in the electron's , such that the scattered photon's frequency matches the incident frequency, with no net transfer to the . This model yields the Thomson cross-section \sigma_T \approx 6.65 \times 10^{-25} cm², characterizing the effective scattering area per .

Derivation

Differential Cross-Section

The differential cross-section in Thomson scattering quantifies the probability of scattering into a particular direction, expressed as the scattered power per unit divided by the incident intensity. In the classical treatment, an incident electromagnetic with \mathbf{E} = \hat{\epsilon} E_0 \cos(\omega t) drives a of charge -e and mass m_e according to the equation of motion m_e \ddot{\mathbf{r}} = -e \mathbf{E}, assuming non-relativistic conditions and neglecting magnetic forces and radiation damping. The resulting oscillatory motion has acceleration \mathbf{a} = (e E_0 / m_e) \hat{\epsilon} \cos(\omega t), with time-averaged squared amplitude \langle a^2 \rangle = (e^2 E_0^2)/(2 m_e^2). The time-averaged power radiated per unit solid angle by this oscillating dipole follows from the classical formula for non-relativistic dipole radiation: \frac{dP}{d\Omega} = \frac{e^2 \langle a^2 \rangle}{16 \pi^2 \epsilon_0 c^3} \sin^2 \theta, where \theta is the angle between the acceleration direction (aligned with the incident polarization \hat{\epsilon}) and the observation direction \mathbf{n}. The incident wave has time-averaged intensity (Poynting flux) I = \frac{1}{2} c \epsilon_0 E_0^2. The differential cross-section is then \frac{d\sigma}{d\Omega} = \frac{dP/d\Omega}{I}, yielding \frac{d\sigma}{d\Omega} = r_e^2 \sin^2 \theta for linearly polarized incident light, where r_e = \frac{e^2}{4 \pi \epsilon_0 m_e c^2} \approx 2.82 \times 10^{-15} m is the classical electron radius. This expression was first derived by J. J. Thomson in his classical analysis of light scattering by charged particles. For unpolarized incident light, the cross-section requires averaging over the two orthogonal states. Assuming the scattering plane is defined by the incident direction \mathbf{k}_i and \mathbf{n}, with scattering \phi between \mathbf{k}_i and \mathbf{n}, the average yields \frac{d\sigma}{d\Omega} = r_e^2 \frac{1 + \cos^2 \phi}{2}. This form is of the azimuthal around \mathbf{k}_i and exhibits between forward (\phi = 0^\circ) and backward (\phi = 180^\circ) directions, where the value reaches r_e^2, with a minimum of \frac{1}{2} r_e^2 at \phi = 90^\circ. The differential cross-section has units of area per (m²/sr) and normalizes such that its integral over all solid angles gives the total Thomson cross-section.

Total Cross-Section

The cross-section for Thomson scattering is obtained by integrating the differential cross-section over all , yielding a frequency-independent result in the classical low-energy limit. For unpolarized incident , this integration gives the Thomson cross-section \sigma_T = \frac{8\pi}{3} r_e^2, where r_e = \frac{e^2}{4\pi\epsilon_0 m_e c^2} \approx 2.818 \times 10^{-15} m is the . Numerically, \sigma_T \approx 6.65 \times 10^{-29} m^2, or equivalently 0.665 s (with 1 = 10^{-28} m^2). To derive this, begin with the differential cross-section for : \frac{d\sigma}{d\Omega} = r_e^2 \frac{1 + \cos^2 \theta}{2}, where \theta is the scattering angle. The total cross-section is then \sigma_T = \int \frac{d\sigma}{d\Omega} \, d\Omega = r_e^2 \int_0^{2\pi} d\phi \int_0^\pi \frac{1 + \cos^2 \theta}{2} \sin \theta \, d\theta, exploiting azimuthal . The \phi integral yields $2\pi, and the factor of $1/2 simplifies it to \pi r_e^2 \int_0^\pi (1 + \cos^2 \theta) \sin \theta \, d\theta. Substituting u = \cos \theta (du = -\sin \theta \, d\theta), the limits change from u=1 to u=-1, giving \pi r_e^2 \int_{-1}^1 (1 + u^2) \, du = \pi r_e^2 \left[ u + \frac{u^3}{3} \right]_{-1}^1 = \pi r_e^2 \cdot \frac{8}{3} = \frac{8\pi}{3} r_e^2. Physically, \sigma_T represents the effective scattering probability per free electron for low-energy photons, serving as a fundamental constant in radiative transfer calculations where electron-photon interactions dominate. This classical result is independent of photon frequency, reflecting the non-relativistic acceleration of the electron by the electromagnetic wave. Compared to the geometric cross-section inferred from the classical electron radius (\pi r_e^2 \approx 2.5 \times 10^{-29} m^2), \sigma_T is of comparable scale but larger by a factor of about 2.7, underscoring the wave-like nature of the interaction rather than a purely geometric one. The Thomson cross-section is valid only when the photon energy h\nu \ll m_e c^2 \approx 511 keV, ensuring negligible and energy loss during scattering; at higher energies, quantum effects lead to the Klein-Nishina formula, reducing the cross-section.

Applications

Astrophysical Contexts

In the solar corona and atmospheres, free s scatter photospheric light through Thomson scattering, which dominates the K-corona's continuum emission and leads to a reversal of the photosphere's due to the favorable scattering geometry at larger solar radii. This effect is particularly evident in white-light observations, where the scattered intensity increases toward the limb, providing a diagnostic for coronal electron densities and s. Measurements of the scattered light's further enable inferences about the coronal and temperature structure, as the degree of polarization depends on the viewing angle relative to the electron's acceleration. Thomson scattering played a pivotal role during the recombination at z \approx 1100, when the transitioned from ionized to neutral gas, establishing the surface of last for the (). The high at this time resulted in an enormous \tau_e = n_e \sigma_T L, where n_e is the , \sigma_T is the Thomson cross-section, and L is the relevant path length, causing photons to diffuse randomly and imprint the plasma's velocity and density perturbations onto the anisotropies. This process sets the characteristic angular scale of the power spectrum, with the acoustic peaks arising from baryon-photon oscillations damped by the , and subsequent at lower redshifts (z \sim 6-10) adds a small secondary of \tau \approx 0.058 (as of Planck PR4, 2025), generating large-scale E-mode . In the , particularly within H II regions ionized by massive stars, Thomson scattering by free s contributes to the polarization of free-free emission at the region's edges and produces extended halos around embedded compact sources. These halos arise as photospheric or nebular light is scattered by the diffuse electron gas, broadening the apparent size of sources and altering their observed profiles, which is observable in radio and optical wavelengths where free-free processes also dominate the thermal emission. Such effects are crucial for interpreting the and energetics of H II regions, as they modulate the escape of and influence the region's ionization balance. In accretion disks surrounding black holes, such as in binaries like , the Thomson optical depth \tau_T often exceeds unity due to dense populations in the inner disk and , leading to multiple scatterings that Comptonize soft disk photons into the observed hard spectrum. This Comptonization process, operating in the Thomson limit for photon energies below ~100 keV, produces a power-law tail in the spectrum, with the depending on the temperature and , as seen in the hard state of where the 's hot s (kT_e \sim 100 keV) upscatter thermal disk emission. Observations of such systems reveal variability tied to changes in \tau_T, highlighting Thomson scattering's role in regulating radiation transport and efficiency in these extreme environments. In supernova remnants, relativistic electrons accelerated at the shock fronts scatter synchrotron radiation produced by the same population, operating in the semi-Thomson regime where photon energies in the remain below the energy. This synchrotron self-Compton () process generates gamma-ray emission complementary to the observed radio and synchrotron, with the applying to lower-energy photons and transitioning to at higher energies, constraining the maximum energies in young remnants like . Such contributes to the non-thermal broadband spectrum, providing insights into particle acceleration efficiency and strengths within these cosmic accelerators.

Plasma Diagnostics

In laser Thomson scattering diagnostics, a high-power beam is directed into the , where it scatters off free , providing a non-perturbing of key parameters. The scattered is analyzed to determine electron T_e from the caused by the thermal motion of electrons, with the width of the scaling as \Delta \omega / \omega_0 \propto (T_e / m_e [c](/page/Speed_of_light)^2)^{1/2}, where \omega_0 is the incident , m_e is the electron , and [c](/page/Speed_of_light) is the . Electron n_e is inferred from the total scattered power, which is proportional to n_e in the non-collective regime. The regime depends on the parameter k \lambda_D, where k is the wave number of the scattered light and \lambda_D is the ; for k \lambda_D \gg 1 (non-collective regime), the is incoherent and dominated by individual motions, allowing straightforward extraction of n_e and T_e from the Rayleigh-like spectrum. In contrast, the collective regime (k \lambda_D \lesssim 1) involves coherent from , complicating the analysis but enabling measurements of additional parameters like wave spectra. Noise from collective fluctuations is typically mitigated through signal averaging over multiple shots or spatial points to improve . Ion Thomson scattering extends these diagnostics to heavier ions by probing ion-acoustic waves in the collective regime, often using longer-wavelength probes such as CO_2 lasers to match the lower ion velocities. The spectral shift and broadening reveal the ion velocity distribution and temperature T_i, with the ion-acoustic feature appearing at frequencies \omega \approx k c_s, where c_s is the ion sound speed depending on T_e and T_i. This technique has been used to measure ion species temperature ratios in fusion-relevant plasmas. These methods find critical applications in laboratory plasma experiments, including at the (NIF), where Thomson scattering characterizes underdense plasmas in hohlraums to validate ignition conditions. In magnetic confinement devices like tokamaks, edge diagnostics monitor n_e and T_e profiles to assess divertor performance and stability. Thomson scattering also supports space plasma simulations on facilities such as the Large Plasma Device, replicating or magnetospheric conditions for validation against satellite data. Instrumentation typically includes collection optics, such as off-axis Schwarzschild telescopes, to relay the scattering volume to high-resolution spectrometers like triple-grating or fiber-based designs, which resolve spectral features down to 0.1 nm while rejecting stray light via filters. Detectors, often intensified CCDs or streak cameras, capture time-resolved spectra, enabling measurements in dynamic plasmas with repetition rates up to 20 kHz.

Comparisons and Extensions

Relation to Compton Scattering

Compton scattering describes the inelastic interaction between a photon and a free electron, where energy and momentum conservation lead to a shift in the photon's wavelength given by \Delta \lambda = \frac{h}{m_e c} (1 - \cos \theta), with h as Planck's constant, m_e the electron mass, c the speed of light, and \theta the scattering angle. In this process, the electron recoils, gaining kinetic energy, and the scattering is inherently quantum mechanical. In the low-energy limit where the incident h\nu \ll m_e c^2 \approx 511 keV, the becomes negligible, the scattering turns , and the Compton cross-section reduces to the classical Thomson cross-section \sigma_T = \frac{8\pi}{3} r_e^2, with r_e the . Here, the electron can be treated as essentially at rest post-scattering, aligning with the non-relativistic approximation. The relativistic quantum mechanical treatment of is provided by the Klein-Nishina formula for the differential cross-section: \frac{d\sigma}{d\Omega} = \frac{r_e^2}{2} \left( \frac{E'}{E} \right)^2 \left( \frac{E'}{E} + \frac{E}{E'} - \sin^2 \theta \right), where E is the incident , E' = \frac{E}{1 + \frac{E}{m_e c^2}(1 - \cos \theta)} is the scattered , and \theta is the scattering angle. For E \ll m_e c^2, this formula approximates the Thomson differential cross-section \frac{d\sigma}{d\Omega} = \frac{r_e^2}{2} (1 + \cos^2 \theta), recovering the classical result without quantum recoil effects. The Thomson approximation holds well for photon energies below approximately 10 keV in X-ray regimes, where the relative deviation from Klein-Nishina is small, but transitions to full Compton behavior around 10–100 keV as relativistic effects become significant. In quantum electrodynamics, Thomson scattering corresponds to the tree-level Feynman diagrams (s- and u-channel) for photon-electron scattering in the low-energy regime, providing the leading-order description before higher-order corrections.

Distinction from Rayleigh Scattering

Thomson scattering and Rayleigh scattering are both elastic processes in which electromagnetic radiation is scattered without a change in photon energy, but they differ fundamentally in the nature of the scatterer and the applicable physical regime. Thomson scattering arises from the interaction of light with free, unbound electrons, where the electron oscillates under the influence of the incident electric field and re-radiates as a dipole, with the scattering cross-section independent of the radiation frequency. In contrast, Rayleigh scattering involves coherent scattering by bound electrons in neutral atoms or molecules, treated as induced dipoles in particles much smaller than the wavelength (a ≪ λ), leading to a total cross-section σ_R that scales strongly with frequency as σ_R ∝ ω⁴ (or equivalently ∝ λ⁻⁴), which explains phenomena like the blue color of the sky due to preferential scattering of shorter wavelengths. This frequency dependence in Rayleigh scattering stems from the harmonic oscillator model of the bound electron, where the response is resonant near the atom's natural frequency ω₀, and for ω ≪ ω₀, the approximation σ_R ≈ σ_T (ω/ω₀)⁴ holds, with σ_T being the Thomson cross-section. A key distinction lies in the scatterer type and resulting charge dynamics: Thomson scattering accelerates free charges with no net restoring force, producing scattering independent of the medium's atomic structure, whereas Rayleigh scattering induces oscillations in neutral systems without net charge acceleration, as the atom as a whole responds to the field. Both processes exhibit similar angular dependence in their differential cross-sections, with dσ/dΩ ∝ sin²θ due to the dipole radiation pattern, where θ is the angle between the incident and the scattering direction; however, Rayleigh scattering applies specifically to small, polarizable particles, while Thomson treats the electron as a point charge. In partially ionized plasmas, both mechanisms can coexist, with Thomson scattering dominating interactions with free electrons and Rayleigh scattering contributing from neutral atoms, though Rayleigh becomes prominent at optical wavelengths where neutrals are prevalent. In dense media, such as optically thick atmospheres or gases, multiple scattering events complicate both processes, but Thomson scattering remains particularly relevant in highly ionized environments like stellar interiors or fusion plasmas, where free s abound, whereas is more characteristic of neutral or lowly ionized gases. This distinction underscores their complementary roles in diagnostics: Thomson for electron properties in plasmas, and for neutral and measurements.

References

  1. [1]
    [PDF] CHAPTER 23 The Interaction of Light with Matter: I - Scattering
    Thomson scattering: is the elastic (coherent) scattering of electromag- netic radiation by a free charged particle, as described by classical electro- magnetism ...Missing: definition key
  2. [2]
    Thomson Scattering - Richard Fitzpatrick
    Thomson Scattering. When an electromagnetic wave is incident on a charged particle, the electric and magnetic components of the wave exert a Lorentz force on ...Missing: facts | Show results with:facts
  3. [3]
    Thomson Scattering of Radiation - Duke Physics
    This is the Thomson formula for scattering of radiation by free charge. It works for X-rays for electrons or $\gamma$ -rays for protons.Missing: definition key facts
  4. [4]
    [PDF] Thomson and Rayleigh Scattering Initial questions
    In the limit we're considering, the new frequency is exactly the same as the old frequency. That's Thomson scattering.Missing: key facts
  5. [5]
    None
    ### Historical Context of Thomson Scattering Discovery and Key Developments
  6. [6]
    [PDF] Thomson scattering on the Large Plasma Device - BaPSF
    Aug 1, 2025 · Thomson scattering (TS) is a powerful, first-principles, and non-invasive plasma diagnostic that derives electron density and temperature from ...Missing: key | Show results with:key
  7. [7]
    [PDF] LXX. On the number of corpuscles in an atom - Gilles Montambaux
    Published online: 16 Apr 2009. To cite this article: Prof. J.J. Thomson M.A. F.R.S. (1906) LXX. On the number ... Scattering of R6ntgen Radiation by Gases.-.
  8. [8]
    [PDF] LECTURE 9 Thomson Scattering
    For scattering problems, this holds when the distance electrons wander in an atom (typically of order 1 Angstrom) is << than the wavelength of the radiation— ...Missing: condition m_e
  9. [9]
    [PDF] Classical, Non-Relativistic Theory of Scattering of Electromagnetic ...
    the bound atomic/molecular electrons behave as if they are free – i.e. as in Thomson scattering of free electrons! When. 0 ω ω. the behavior of the factor ...
  10. [10]
    Thomson cross section - CODATA Value
    Numerical value, 6.652 458 7051 x 10-29 m ; Standard uncertainty, 0.000 000 0062 x 10-29 m ; Relative standard uncertainty, 9.3 x 10 ...
  11. [11]
    [PDF] 3. Interaction of Neutral Particles with Matter
    Compared to the ionisation cross sections discussed in Section 2, these cross sections ll ! = 6.65 10-25 cm2 = 0.665 b (barn). [cm2 / Atom]. K.
  12. [12]
    [PDF] Nonlinear Thomson scattering: A tutorial
    In Thomson scattering, an electron that is initially at rest may acquire relativistic velocities in the fields of high- intensity light and through this ...Missing: key | Show results with:key
  13. [13]
    [PDF] A Technique to Measure Coronal Electron Density, Temperature ...
    Thomson scattering is elastic, which means that the energy and wavelength of the photon in- cident on the electron are preserved upon scattering. Thomson ...
  14. [14]
    [PDF] 29. Cosmic Microwave Background - Particle Data Group
    Aug 11, 2022 · Since this corresponds to a scattering optical depth τ ≃ 0.06, then roughly 6% of CMB photons were re-scattered at the reionization epoch, with ...
  15. [15]
    Anisotropies in the CMB - M. White et al.
    tau / dz, for Thomson scattering optical depth tau . This function measures the probability that the radiation was last scattered in a redshift interval dz ...
  16. [16]
    CBI limits on 31 GHz excess emission in southern H ii regions
    ... for pure free–free emission. Free–free emission is intrinsically unpolarized, but can be polarized at the edges of H ii regions by Thomson scattering. The ...
  17. [17]
    Radiation mechanisms and geometry of Cygnus X-1 in the soft state
    A likely model for the hard state of Cyg X-1 embodies Comptonization of blackbody photons from an optically thick accretion disc in a hot, thermal, optically ...
  18. [18]
    What is Thomson Scattering for Plasma Diagnostics
    Thompson Scattering can be used to non-invasively determine key plasma characteristics (taking a Surfatron plasma as an example) such as electron density ...
  19. [19]
    Electron temperature and density measurement by Thomson ...
    Sep 6, 2022 · Electron temperature and density measurement by Thomson scattering with a high repetition rate laser of 20 kHz on LHD.
  20. [20]
    Thomson Scattering: Spectral Density - PlasmaPy
    α > 1 corresponds to collective scattering, while α < 1 corresponds to non-collective scattering. Depending on which of these regimes applies, fitting the ...
  21. [21]
    [PDF] Thomson-Scattering Measurements in the Collective and Non ...
    May 14, 2010 · These measurements allow an accurate local measurement of fundamental plasma parameters: electron temperature, density and ion temperature.
  22. [22]
    [PDF] Thomson scattering diagnostics of nonthermal plasma from particle ...
    Abstract—Optical Thomson scattering is now a mature di- agnostic tool for precisely measuring local plasma density and temperature.
  23. [23]
    Thomson-scattering techniques to diagnose local electron and ion ...
    This new technique allows a simultaneous time resolved local measurement of electron density and temperature.
  24. [24]
    Thomson scattering diagnostic for the measurement of ion species ...
    Simultaneous Thomson scattering measurements of collective electron-plasma and ion-acoustic fluctuations have been utilized to determine ion species ...
  25. [25]
    Core ion measurements with collective Thomson scattering for ...
    Feb 20, 2024 · Here we assess the accuracy with which a collective Thomson scattering (CTS) diagnostic can provide key measurements for burn control in the planned European ...
  26. [26]
    [PDF] UC San Diego - eScholarship.org
    The first application of Thomson scattering on plasma devices used in fusion ... Thomson scattering from inertial confinement fusion hohlraum plasmas was ... NIF ...<|separator|>
  27. [27]
    [PDF] Simulated performance of the optical Thomson scattering diagnostic ...
    An optical Thomson scattering diagnostic has been designed for the National Ignition Facility (NIF) to characterize under-dense plasmas.<|separator|>
  28. [28]
    Physics feasibility study of a collective Thomson scattering ...
    Here we explore conceptual design options and potential measurement capabilities of a collective Thomson scattering diagnostic at SPARC.<|separator|>
  29. [29]
    [PDF] The preliminary design of the optical Thomson scattering diagnostic ...
    May 8, 2016 · The collection telescope is an off-axis Schwarzschild design that relays a 50 µm spot (Thomson volume) to the entrance of the spectrometers. One ...
  30. [30]
    High-speed fiber-based spectrometer for plasma Thomson scattering
    We present a novel concept for a Thomson scattering diagnostic, based on a high-speed fiber optic spectrometer. The high-speed fiber optic spectrometer ...
  31. [31]
    [PDF] graphs of the Compton energy-angle relationship and the Klein ...
    Nishina Formula for Free Electrons. The Compton scattering by free electrons is described to a very good approximation by the. Klein-Nishina formula.
  32. [32]
    [PDF] Thomson and Rayleigh Scattering
    ... set ω to ω0. Then we get σ(ω) = πσT. 2τ ω2. 0τ/2π. (ω−ω0)2+(ω2. 0τ/2)2. = πσT. 2τ. Γ/2π. (ω−ω0)2+(Γ/2)2. = 2π2e2 mc. Γ/2π. (ω−ω0)2+(Γ/2)2. (15). Page 5. where Γ ...
  33. [33]
    [PDF] Thomson scattering, Rayleigh scattering and all that
    (1.4). This is known as the classical Thomson cross section. When the incident energy of the photon ¯hω becomes comparable to the rest mass of the electron ...
  34. [34]
    Influence of Rayleigh scattering on the ion feature of Thomson ...
    Jun 1, 2001 · The result shows that the profile of the scattering spectrum is nearly the same as that of Thomson scattering with inclusion of Rayleigh ...Missing: ionized | Show results with:ionized
  35. [35]
    Two-color scattering for the measurement of neutrals at the edge of ...
    Jun 15, 2021 · The scattering signal is the summation of the two scattering intensities, with the Thomson signal typically dominating the Rayleigh signal under ...