Skip to main content
Log in

Basic microscopic plasma physics from N-body mechanics

A tribute to Pierre-Simon de Laplace

  • Review Paper
  • Published:
Reviews of Modern Plasma Physics Aims and scope Submit manuscript

Abstract

Computing is not understanding. This is exemplified by the multiple and discordant interpretations of Landau damping still present after 70 years. For long deemed impossible, the mechanical N-body description of this damping, not only enables its rigorous and simple calculation, but makes unequivocal and intuitive its interpretation as the synchronization of almost resonant passing particles. This synchronization justifies mechanically why a single formula applies to both Landau growth and damping. As to the electrostatic potential, the phase mixing of many beam modes produces Landau damping, but it is unexpectedly essential for Landau growth too. Moreover, collisions play an essential role in collisionless plasmas. In particular, Debye shielding results from a cooperative dynamical self-organization process, where “collisional” deflections due to a given electron diminish the apparent number of charges about it. The finite value of exponentiation rates due to collisions is crucial for the equivalent of the van Kampen phase mixing to occur in the N-body system. The N-body approach incorporates spontaneous emission naturally, whose compound effect with Landau damping drives a thermalization of Langmuir waves. O’Neil’s damping with trapping typical of initially large enough Langmuir waves results from a phase transition. As to Coulomb scattering, there is a smooth connection between impact parameters where the two-body Rutherford picture is correct, and those where a collective description is mandatory. The N-body approach reveals two important features of the Vlasovian limit: it is singular and it corresponds to a renormalized description of the actual N-body dynamics.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5

Similar content being viewed by others

Notes

  1. The sentence defining what was called afterwards Laplace’s demon is well known : “Une intelligence qui, pour un instant donné, connaîtrait toutes les forces dont la nature est animée et la situation respective des êtres qui la composent, si d’ailleurs elle était assez vaste pour soumettre ces données à l’analyse, embrasserait dans la même formule les mouvements des plus grands corps de l’univers et ceux du plus léger atome : rien ne serait incertain pour elle, et l’avenir, comme le passé, serait présent à ses yeux.” [Essai philosophique sur les probabilités (Laplace 1840)] English translation : “An intellect, which at a certain moment would know all forces that set nature in motion, and all positions of all items of which nature is composed, if this intellect was also vast enough to submit these data to analysis, it would embrace in a single formula the movements of the greatest bodies of the universe and those of the tiniest atom; for such an intellect nothing would be uncertain and the future just like the past would be present before its eyes.” However, the genuine Laplace’s dream is reasonable, since a few sentences later he states: “Tous ses efforts dans la recherche de la vérité tendent à rapprocher [l’esprit humain] sans cesse de l’intelligence que nous venons de concevoir, mais dont il restera toujours infiniment éloigné.” English translation : “All its efforts in the quest of truth tend at moving [the human spirit] closer to the intelligence we have just conceived, but from which it will always stay infinitely distant”.

  2. This rigorous derivation is accessible to students knowing Newton’s second law of motion and the Fourier transform, but neither analytic functions, nor the Laplace transform. It also provides a correction to the lowest order expression of Landau damping. In a second step, the calculation is extended to the corresponding particle dynamics, showing that it produces an average synchronization of almost resonant passing particles with the wave.

  3. Many textbooks multiply this expression by \(\sqrt{2}\).

  4. This smoothing is similar to the one performed in the mean-field derivation of the Vlasov equation, which was recalled in the introduction.

  5. In the similar spirit of considering a wave with a slowly varying amplitude, but using smooth distribution functions, Bénisti and Gremillet (2007); Bénisti et al. (2009) derive not only Landau damping, but also its equivalent in the nonlinear regime.

  6. An arbitrary phase should be present in the sine, but it is set to 0, since it is not important for the derivation. The case with a concomitant small velocity modulation can be dealt with in a similar manner, but with longer expressions. Then a \(\varDelta \mathbf {v}_{ j} \, t \, \sin (\mathbf {k}_{\mathbf {m}} \cdot \mathbf {r}_{j0} + \psi _j)\) contribution must be added to \(\varDelta {\mathbf {r}}_j(t)\), where \(\psi _j\) and \(\varDelta \mathbf {v}_{ j}\) are, respectively, the phase and the amplitude of modulation of the velocities of the particles belonging to the same beam as particle j.

  7. The same property would hold as well for other types of interactions than Coulombian ones.

  8. Phase mixing is a classical concept in the theory of kinetic plasma waves, and especially in the van Kampen–Case approach to Landau damping (van Kampen 1955, 1957; Case 1959). Intuitively, it corresponds to the idea that the integral of a rapidly oscillating function is close to zero. Mathematically, it is grounded on the fact that \(\int {\mathrm {d}}\nu F(\nu ) \exp (- {\mathrm {i}}\nu t)\) is the Fourier transform of \(F(\nu )\), which decays for large t’s in various cases. This occurs in particular, at least on average, if \(F(\nu )\) is an \(L^2\) function; also if \(F(\nu )\) has an integrable derivative of order at least one. Having one of such properties is natural for h(v), \(g'(v)\), and \(g^{\prime \prime }(v)\), especially if g(v) is analytic, as assumed in Landau’s derivation of Landau damping (Lifshitz and Pitaevskii 1981).

  9. Phase mixing works here in the following way: because of the integration over v, the first exponential and the cosine terms in Eq. (18) produce the Fourier transforms of h(v), \(g'(v)\) and \(g^{\prime \prime }(v)\). For t large, the vanishing values of the tails of these transforms are involved, and thus neglected. For t large, in the integrals involving \(\sin (\varOmega t)/\varOmega \), this factor has non vanishing values over a vanishing domain in v. This enables extracting \(g'(\frac{\omega }{k_{\mathbf {m}}})\) out of the integral. The similar contribution involving \(g^{\prime \prime }(v)\) is neglected, since it is of higher order in \((\tau _A k_{\mathbf {m}} v_{{\mathrm {T}}})^{-1}\). We stress that we do not use Plemelj formula, in contrast with what was done in Kaufman (1972).

  10. At this point, we notice that the derivation of Landau damping in the mechanical N-body setting is more accessible to students than Landau’s derivation: the number of pages is divided by three (see for instance sections 6.3–6.5 of Nicholson (1983)), the mathematics is elementary, and there is no need to introduce Vlasov equation.

  11. Because of the Floquet exponents introduced in Appendix 1 and computed in Appendix 3, the actual condition (63) is somewhat stronger.

  12. The Bohm–Gross dispersion relation is \(\omega ^2 = \omega _{{\mathrm {p}}}^2 + 3 k^2 v_{{\mathrm {T}}}^2\).

  13. Using again phase mixing, but going to next order in \((k_{\mathbf {m}} v_{{\mathrm {T}}} \tau _A)^{-1}\) in Eq. (19), adds \( \frac{\pi \omega _{{\mathrm {p}}}^2}{k_{\mathbf {m}}^3} g^{\prime \prime } (\frac{\omega }{k_{\mathbf {m}}}) \dot{A}(t) - {\mathrm {i}}\frac{\pi \omega _{{\mathrm {p}}}^2}{2 k_{\mathbf {m}}^4} g^{\prime \prime \prime } (\frac{\omega }{k_{\mathbf {m}}}) \ddot{A}(t)\) in the right-hand side of this equation. The next orders can be obtained from Eqs. (108, 109).

  14. Actually, a possible role of trapping is a priori excluded since the bounce period is unbounded in the linear regime of Langmuir waves.

  15. This derivation is close to that in Escande et al. (2015b), but takes advantage of the simplification of Escande et al. (2015a). The derivation of Escande et al. (2016) was convenient only when taking right away the singular limit of Sect. 6.

  16. The Laplace transform in time maps a function g(t) to \(\widehat{g}(\omega ) = \int _0^{\infty } g(t) \exp ({\mathrm {i}}\omega t) {\mathrm {d}}t\) (with \(\omega \) complex).

  17. Since the arguments of functions are spelled explicitly, from now on we omit diacritics for the Laplace (or Fourier) transformed quantities.

  18. We stress that the derivation of Landau damping in this section is completely independent of that in Sect. 3 though.

  19. Actually, the analysis of Cauchy integrals, a more advanced topic in mathematics (section VIII.12 of Godement 2015), allows wilder g(v)’s. There, the analyticity of the integrals with respect to \(\omega \), not v, in the upper or lower complex half-plane mirrors the use of Laplace transform. The relevant g(v)’s are such that their absolute values are integrable, as well as that of their Fourier transforms.

  20. In the limit where \(\delta \) vanishes, the absence of these two roots does not modify Eq. (60), since they bring an infinitesimal contribution to the result.

  21. Without the phase mixing term, a wave with an initial amplitude 1 would have only half the amplitude \({\mathrm {e}}^{\gamma _\mathrm{L}(\mathbf {k}_{\mathbf {m}}) t}\) after a long enough time. This remark by Dr A. Samain was at the origin of the development of the van Kampen-like calculation in section 3.8.3 of Elskens and Escande (2002).

  22. These two figures represent actually the poles of \(1/\vert \epsilon _{{\mathrm {d}}}(\mathbf {m},\omega )\vert \), which are nothing else but the zeros of \(\epsilon _{{\mathrm {d}}}(\mathbf {m},\omega )\). Identifying the poles of a function of complex variable is easily done by identifying the closed contours in the complex plane.

  23. It was recently noted that the collisional damping rate of Langmuir waves is much smaller than the one provided by the collision frequency (see Yoon et al. 2016; Tigik et al. 2016).

  24. This property is the signature of the Floquet exponents mentioned in Appendix 1.

  25. Therefore, taking this singular limit is a way to bring to students a rapid derivation of both Landau damping and Debye shielding.

  26. However, this has no practical consequence, since the time scale over which the system is observed, or measured, introduces a natural granularity in frequency space below which the actual continuum (or singular) limit and the discretized system should not be distinguishable.

  27. In contrast, the global exchange of energy and momentum between waves and particles is easily available in a Vlasovian setting.

  28. For completeness, most of this section follows closely section 5 of Escande et al. (2015b).

  29. Picard’s iteration technique is one of the standard methods to prove the existence and uniqueness of solutions to first order equations with given initial conditions. It uses the fact that the exact solution of equation \(\frac{{\mathrm {d}}\mathbf {X}}{{\mathrm {d}}t} = f(\mathbf {X})\) is the fixed point of the iterative process starting from \(n = 0\), and providing \(\mathbf {X}_{n+1}\) from \(\mathbf {X}_{n}\) by equation \(\frac{{\mathrm {d}}\mathbf {X}_{n+1}}{{\mathrm {d}}t} = f(\mathbf {X}_n)\) with any choice of \(\mathbf {X}_{0}(t)\). This iteration technique is very convenient, in particular to alleviate the algebra of perturbation calculations.

  30. This was the starting point of the one-dimensional N-body approach.

  31. We notice that the derivations of the self-consistent dynamics starting with a Vlasovian description by O’Neil et al. (1971), Onishchenko et al. (1970), Tennyson et al. (1994) perform kind of a zigzag with respect to the N-body description, since they go back to a finite number of degrees of freedom after going through the continuous Vlasovian description.

  32. This expression corrects Eq. (46) of Escande et al. (2015b).

  33. This further relaxation cannot be described by the self-consistent Hamiltonian, since the latter corresponds to a given bulk.

  34. Rigorously speaking, the quasilinear equations had already been mentioned in 1961 by Romanov and Filippov (1961). This reference makes the Ansatz of a Fokker–Planck equation for particle evolution and computes the corresponding diffusion coefficient; it describes the evolution of the Langmuir wave amplitude as the result of spontaneous and stimulated emission of quanta and estimates the corresponding coefficients.

  35. More information about this controversy can be found in Doxas and Cary (1997), Laval and Pesme (1999).

  36. Experimentally, studying the chaotic transport of particles in a prescribed spectrum of waves was propaedeutic to the self-consistent case too. This led to the experimental observation of resonance overlap by Doveil et al. (2005a), and of the transition from stochastic diffusion in a large set of waves to slow chaos associated to a pulsating separatrix by Doveil and Macor (2011). Nonlinear resonances excited by injected waves were both observed as a “devil’s staircase” by Macor et al. (2005) and cancelled to build a barrier to transport by Chandre et al. (2005).

  37. Equations (82, 83) are used here to avoid writing their one-dimensional analogue, which is the relevant one for this discussion.

  38. For a plateau with a finite width, the small remaining source brings a further evolution of the wave–particle system toward a state where the wave spectrum collapses toward small wavelengths together with the escape of initially resonant particles towards low bulk plasma thermal speeds (Firpo et al. 2006). This corresponds to a further step toward a new thermal equilibrium of the N-body system corresponding to the initial beam-plasma system. The description of the subsequent steps toward thermal equilibration requires to use a full N-body model.

  39. This phenomenon, also called depression of nonlinearity, was introduced in fluid mechanics (Kraichnan and Panda 1988). In Navier–Stokes turbulence, the mean-square value of the nonlinear term of the equation was found significantly depressed, i.e. smaller than the same quantity in the Gaussian field with the same energy distribution. This was identified to result from the emergence of long-lived vortices where the enstrophy cascade is inhibited. It also exists in systems with quadratic nonlinearities (Kraichnan and Panda 1988; Bos et al. 2012).

  40. This is reminiscent of the fact that the initial correlations were not disturbed in the course of the relaxation in the beam-plasma experiment (Krivoruchko et al. 1981).

  41. The envelope equation of an electron plasma wave has a sudden variation when going from the linear to the trapping regimes, in a way similar to a first order phase transition (Bénisti 2018). Furthermore, there are other aspects of non collisional damping for a wave having trapped electrons (Bénisti et al. 2012).

  42. Their derivation suggested to the first author of the present review that a direct N-body approach might be possible.

  43. The conspicuous modification of relaxation processes by magnetic fields (Ichimaru and Rosenbluth 1970; Silin 1963; Hassan and Watson 1977a; Baldwin and Watson 1977; Hassan and Watson 1977b), suggests that shielding might be modified too. They also suggest the importance of the relative ordering of spatial scales, such as Larmor radius versus the Debye length.

  44. Equation (120) corrects a typo in equation (33) of Dawson (1960).

  45. We assume \(g'(v)\) continuous, |g(v)| to be integrable, and \(g(v) \ge 0\).

  46. More precisely, the pulverization leads to infinitesimal beamlets whose summation corresponds to a coarse-graining of the previous multi-beam-multi-arrays.

References

  • R. Abe, Giant cluster expansion theory and its application to high temperature plasma. Prog. Theor. Phys. 22(2), 213–226 (1959)

    ADS  MathSciNet  MATH  Google Scholar 

  • M. Abramowitz, I.A. Stegun, Handbook of Mathematical Functions: With Formulas, Graphs, and Mathematical Tables/edited by Milton Abramowitz and Irene A. Stegun (Dover, New York, 1967)

    MATH  Google Scholar 

  • J.C. Adam, G. Laval, D. Pesme, Reconsideration of quasilinear theory. Phys. Rev. Lett. 43(22), 1671–1675 (1979)

    ADS  Google Scholar 

  • M. Antoni, Y. Elskens, D.F. Escande, Explicit reduction of N-body dynamics to self-consistent particle–wave interaction. Phys. Plasmas 5(4), 841–852 (1998)

    ADS  MathSciNet  Google Scholar 

  • D.R. Baker, N.R. Ahern, A.Y. Wong, Ion-wave echoes. Phys. Rev. Lett. 20(7), 318–321 (1968)

    ADS  Google Scholar 

  • D. Baldwin, C. Watson, Magnetized plasma kinetic theory. II. Derivation of Rosenbluth potentials for a uniform magnetized plasma. Plasma Phys. 19(6), 517–528 (1977)

    ADS  Google Scholar 

  • R. Balescu, Statistical Mechanics of Charged Particles (Interscience, London, 1963)

    MATH  Google Scholar 

  • R. Balescu, Statistical Dynamics: Matter Out of Equilibrium (Imperial Coll, London, 1997)

    MATH  Google Scholar 

  • M. Baus, J.P. Hansen, Statistical mechanics of simple Coulomb systems. Phys. Rep. 59(1), 1–94 (1980)

    ADS  MathSciNet  Google Scholar 

  • G. Bekefi, Collective emission processes in plasmas, in Plasma Physics, ed. by C. DeWitt, J. Peyraud (Gordon and Breach, Philadelphia, 1975), pp. 1–111

    Google Scholar 

  • D. Bénisti, Self-consistent theory for the linear and nonlinear propagation of a sinusoidal electron plasma wave. Application to stimulated Raman scattering in a non-uniform and non-stationary plasma. Plasma Phys. Control. Fusion 60(1), 014040 (2018)

    ADS  Google Scholar 

  • D. Bénisti, D.F. Escande, Origin of diffusion in Hamiltonian dynamics. Phys. Plasmas 4(5), 1576–1581 (1997)

    ADS  MathSciNet  Google Scholar 

  • D. Bénisti, D.F. Escande, Finite range of large perturbations in Hamiltonian dynamics. J. Stat. Phys. 92, 909–972 (1998)

    ADS  MATH  Google Scholar 

  • D. Bénisti, L. Gremillet, Nonlinear plasma response to a slowly varying electrostatic wave, and application to stimulated Raman scattering. Phys. Plasmas 14(4), 042304 (2007)

    ADS  Google Scholar 

  • D. Bénisti, D.J. Strozzi, L. Gremillet, O. Morice, Nonlinear Landau damping rate of a driven plasma wave. Phys. Rev. Lett. 103(15), 155002 (2009)

    ADS  MATH  Google Scholar 

  • D. Bénisti, O. Morice, L. Gremillet, The various manifestations of collisionless dissipation in wave propagation. Phys. Plasmas 19(6), 063110 (2012)

    ADS  Google Scholar 

  • J.T. Berndtson, J.A. Heikkinen, S.J. Karttunen, T.J.H. Pättikangas, R.R.E. Salomaa, Analysis of velocity diffusion of electrons with Vlasov-Poisson simulations. Plasma Phys. Control. Fusion 36(1), 57–71 (1994)

    ADS  Google Scholar 

  • N. Besse, Y. Elskens, D.F. Escande, P. Bertrand, On the validity of quasilinear theory. in Proceeding 38th EPS Conference on Controlled Fusion and Plasma Physics, Strasbourg, p P2.009, (2011a) http://ocs.ciemat.es/EPS2011PAP/pdf/P2.009.pdf. Accessed 3 Apr 2018

  • N. Besse, Y. Elskens, D.F. Escande, P. Bertrand, Validity of quasilinear theory: refutations and new numerical confirmation. Plasma Phys. Control. Fusion 53(2), 025012 (2011b)

    ADS  Google Scholar 

  • N. Boers, P. Pickl, On mean field limits for dynamical systems. J. Stat. Phys. 164(1), 1–16 (2016)

    ADS  MathSciNet  MATH  Google Scholar 

  • W.J. Bos, R. Rubinstein, L. Fang, Reduction of mean-square advection in turbulent passive scalar mixing. Phys. Fluids 24(7), 075104 (2012)

    ADS  Google Scholar 

  • T.J.M. Boyd, J.J. Sanderson, The Physics of Plasmas (Cambridge University Press, Cambridge, 2003)

    MATH  Google Scholar 

  • W. Braun, K. Hepp, The Vlasov dynamics and its fluctuations in the 1/\(N\) limit of interacting classical particles. Commun. Math. Phys. 56(2), 101–113 (1977)

    ADS  MathSciNet  MATH  Google Scholar 

  • J.D. Callen, Coulomb collision effects on linear Landau damping. Phys. Plasmas 21(5), 052106 (2014)

    ADS  Google Scholar 

  • N. Carlevaro, D. Fanelli, X. Garbet, P. Ghendrih, G. Montani, M. Pettini, Beam-plasma instability and fast particles: the Lynden-Bell approach. Plasma Phys. Control. Fusion 56(3), 035013 (2014)

    ADS  Google Scholar 

  • J.R. Cary, D.F. Escande, A.D. Verga, Non quasilinear diffusion far from the chaotic threshold. Phys. Rev. Lett. 65, 3132–3135 (1990)

    ADS  Google Scholar 

  • J.R. Cary, I. Doxas, D.F. Escande, A. Verga, Enhancement of the velocity diffusion in longitudinal plasma turbulence. Phys. Fluids B Plasma Phys. 4(7), 2062–2069 (1992)

    Google Scholar 

  • K.M. Case, Plasma oscillations. Ann. Phys. 7(3), 349–364 (1959)

    ADS  MathSciNet  MATH  Google Scholar 

  • C. Chandre, G. Ciraolo, F. Doveil, R. Lima, A. Macor, M. Vittot, Channeling chaos by building barriers. Phys. Rev. Lett. 94, 074101 (2005). (4 pp)

    ADS  Google Scholar 

  • S. Chibbaro, L. Rondoni, A. Vulpiani, Reductionism, Emergence and Levels of Reality (Springer, Heidelberg, 2014)

    MATH  Google Scholar 

  • J.M. Dawson, Plasma oscillations of a large number of electron beams. Phys. Rev. 118(2), 381–389 (1960)

    ADS  MathSciNet  MATH  Google Scholar 

  • R.L. Dobrushin, Vlasov equations. Funct. Anal. Appl. 13(2), 115–123 (1979)

    MATH  Google Scholar 

  • F. Doveil, A. Macor, Two regimes of self-consistent heating of charged particles. Phys. Rev. E 84(4), 045401 (2011)

    ADS  Google Scholar 

  • F. Doveil, M.C. Firpo, Y. Elskens, D. Guyomarc’h, M. Poleni, P. Bertrand, Trapping oscillations, discrete particle effects and kinetic theory of collisionless plasma. Phys. Lett. A 284(6), 279–285 (2001)

    ADS  Google Scholar 

  • F. Doveil, K. Auhmani, A. Macor, D. Guyomarc’h, Experimental observation of resonance overlap responsible for Hamiltonian chaos. Phys. Plasmas 12, 010702 (2005a). (4 pp)

    ADS  Google Scholar 

  • F. Doveil, D.F. Escande, A. Macor, Experimental observation of nonlinear synchronization due to a single wave. Phys. Rev. Lett. 94(8), 085003 (2005b)

    ADS  Google Scholar 

  • I. Doxas, J.R. Cary, Numerical observation of turbulence enhanced growth rates. Phys. Plasmas 4(7), 2508–2518 (1997)

    ADS  Google Scholar 

  • W.E. Drummond, D. Pines, Nonlinear stability of plasma oscillations. Nucl. Fusion Suppl. 3, 1049–1057 (1962)

    Google Scholar 

  • Y. Elskens, Finite-\(N\) dynamics admit no travelling-waves solutions for the Hamiltonian XY model and single-wave collisionless plasma model. ESAIM: Proceedings, EDP Sciences, vol 10, pp 221–215 (2001)

    MATH  Google Scholar 

  • Y. Elskens, Irreversible behaviours in Vlasov equations and many-body Hamiltonian dynamics: Landau damping, chaos and granularity. in Topics in Kinetic Theory, American Mathematical Soc., Fields Institute Communications Series, ed. by T. Passot, C. Sulem, P.L. Sulem, vol 46, pp 89–108 (2005)

  • Y. Elskens, Quasilinear limit for particle motion in a prescribed spectrum of random waves. Phys. AUC 17, 109–121 (2007)

    Google Scholar 

  • Y. Elskens, Nonquasilinear evolution of particle velocity in incoherent waves with random amplitudes. Commun. Nonlinear Sci. Numer. Simul. 15(1), 10–15 (2010)

    ADS  Google Scholar 

  • Y. Elskens, Gaussian convergence for stochastic acceleration of \(N\) particles in the dense spectrum limit. J. Stat.Phys. 148, 591–605 (2012)

    ADS  MathSciNet  MATH  Google Scholar 

  • Y. Elskens, D.F. Escande, Microscopic Dynamics of Plasmas and Chaos (Institute of Physics, Bristol, 2002)

    Google Scholar 

  • Y. Elskens, E. Pardoux, Diffusion limit for many particles in a periodic stochastic acceleration field. Ann. Appl. Probab. 20, 2022–2039 (2010)

    MathSciNet  MATH  Google Scholar 

  • Y. Elskens, D.F. Escande, F. Doveil, Vlasov equation and \(N\)-body dynamics : how central is particle dynamics to our understanding of plasmas ? Eur. Phys. J. D 68(8), 218 (2014). (7 pages)

    ADS  Google Scholar 

  • D.F. Escande, Description of Landau damping and weak Langmuir turbulence through microscopic dynamics, in Nonlinear World, vol. 2, ed. by V.G. Bar’yakhtar, et al. (World Scientific, Singapore, 1989), pp. 817–836

    Google Scholar 

  • D.F. Escande, Large scale structures in kinetic plasma turbulence, in Large Scale Structures in Nonlinear Physics, ed. by J.D. Fournier, P.L. Sulem (Springer, Berlin, 1991), pp. 73–104

    Google Scholar 

  • D.F. Escande, How to face the complexity of plasmas?, in From Hamiltonian Chaos to Complex Systems, ed. by X. Leoncini, M. Leonetti (Springer, Berlin, 2013), pp. 109–157

    Google Scholar 

  • D.F. Escande, Complexity and simplicity of plasmas, in AIP Conference Proceedings, AIP, vol. 1582, ed. by A. Das, A. Sh Sharma (Melville, 2014), pp. 22–34

  • D.F. Escande, Contributions of plasma physics to chaos and nonlinear dynamics. Plasma Phys. Control. Fusion 58(11), 113001 (2016)

    ADS  Google Scholar 

  • D.F. Escande, From thermonuclear fusion to Hamiltonian chaos. Eur. Phys. J. H (2017). https://doi.org/10.1140/epjh/e2016-70063-5

    Article  Google Scholar 

  • D.F. Escande, Y. Elskens, Proof of quasilinear equations in the chaotic regime of the weak warm beam instability. Phys. Lett. A 302, 110–118 (2002a)

    ADS  Google Scholar 

  • D.F. Escande, Y. Elskens, Proof of quasilinear equations in the chaotic regime of the weak warm beam instability. Phys. Lett. A 302(2–3), 110–118 (2002b)

    ADS  Google Scholar 

  • D.F. Escande, Y. Elskens, Quasilinear diffusion for the chaotic motion of a particle in a set of longitudinal waves. Acta Phys. Pol. B 33, 1073–1084 (2002c)

    ADS  Google Scholar 

  • D.F. Escande, Y. Elskens, Proof of quasilinear equations in the strongly nonlinear regime of the weak warm beam instability. Phys. Plasmas 10, 1588–1594 (2003)

    ADS  MathSciNet  Google Scholar 

  • D.F. Escande, F. Sattin, When can the Fokker-Planck equation describe anomalous or chaotic transport? Phys. Rev. Lett. 99, 185005 (2007). (4 pages)

    ADS  Google Scholar 

  • D.F. Escande, F. Sattin, When can the Fokker-Planck equation describe anomalous or chaotic transport? Intuitive aspects. Plasma Phys. Control. Fusion 50, 124023 (2008). (8 pp)

    ADS  Google Scholar 

  • D.F. Escande, S. Zekri, Y. Elskens, Intuitive and rigorous derivation of spontaneous emission and Landau damping of Langmuir waves through classical mechanics. Phys. Plasmas 3(10), 3534–3539 (1996)

    ADS  Google Scholar 

  • D.F. Escande, N. Besse, F. Doveil, Y. Elskens, Application of Picard iteration technique to self-consistent wave-particle interaction in plasmas. In 39th EPS Conference and 16th International Congress on Plasma Physics, Stockholm, P4.161 (2012) http://ocs.ciemat.es/EPSICPP2012ABS/pdf/P4.161.pdf. Accessed 3 Apr 2018

  • D.F. Escande, Y. Elskens, F. Doveil, Corrigendum: direct path from microscopic mechanics to Debye shielding, Landau damping, and wave-particle interaction (2015 plasma phys. control. fusion 57 025017). Plasma Phys. Control. Fusion 57(6), 069501 (2015a)

    ADS  Google Scholar 

  • D.F. Escande, Y. Elskens, F. Doveil, Direct path from microscopic mechanics to Debye shielding, Landau damping and wave-particle interaction. Plasma Phys. Control. Fusion 57(2), 025017 (2015b)

    ADS  Google Scholar 

  • D.F. Escande, Y. Elskens, F. Doveil, Uniform derivation of Coulomb collisional transport thanks to Debye shielding. J. Plasma Phys. 81(1), 305810101 (2015c)

    Google Scholar 

  • D.F. Escande, F. Doveil, Y. Elskens, \(N\)-body description of Debye shielding and Landau damping. Plasma Phys. Control. Fusion 58(1), 014040 (2016)

    ADS  Google Scholar 

  • M.C. Firpo, Y. Elskens, Kinetic limit of \(N\)-body description of wave-particle self-consistent interaction. J. Stat. Phys. 93(1–2), 193–209 (1998)

    ADS  MathSciNet  MATH  Google Scholar 

  • M.C. Firpo, Y. Elskens, Phase transition in the collisionless damping regime for wave-particle interaction. Phys. Rev. Lett. 84(15), 3318 (2000)

    ADS  Google Scholar 

  • M.C. Firpo, F. Doveil, Y. Elskens, P. Bertrand, M. Poleni, D. Guyomarc’h, Long-time discrete particle effects versus kinetic theory in the self-consistent single-wave model. Phys. Rev. E 64(2), 026407 (2001)

    ADS  Google Scholar 

  • M.C. Firpo, F. Leyvraz, G. Attuel, Equilibrium statistical mechanics for single waves and wave spectra in Langmuir wave-particle interaction. Phys. Plasmas 13(12), 122302 (2006)

    ADS  Google Scholar 

  • A.A. Galeev, R.Z. Sagdeev, V.D. Shapiro, V.I. Shevchenko, Is renormalization necessary in the quasi-linear theory of Langmuir oscillations? Zhurnal Eksperimentalnoi i Teoreticheskoi Fiziki 79, 2167–2174 (1980). [English trans. Sov. Phys.-JETP 52: 1095–1099 (1980)]

    ADS  Google Scholar 

  • S. Gasiorowicz, M. Neuman, R.J. Riddell, Dynamics of ionized media. Phys. Rev. 101, 922–934 (1956)

    ADS  MATH  Google Scholar 

  • R. Godement, Mathematical Analysis III, vol. 3 (Springer, Berlin, 2015)

    MATH  Google Scholar 

  • D.A. Hartmann, C.F. Driscoll, T.M. O’Neil, V.D. Shapiro, Measurements of the weak warm beam instability. Phys. Plasmas 2(3), 654–677 (1995)

    ADS  Google Scholar 

  • M. Hassan, C. Watson, Magnetized plasma kinetic theory. I. Derivation of the kinetic equation for a uniform magnetized plasma. Plasma Phys. 19(3), 237–247 (1977a)

    ADS  Google Scholar 

  • M. Hassan, C. Watson, Magnetized plasma kinetic theory. III. Fokker-Planck coefficients for a uniform magnetized plasma. Plasma Phys. 19(7), 627–649 (1977b)

    ADS  Google Scholar 

  • M. Hauray, P.E. Jabin, Particle approximation of Vlasov equations with singular forces: propagation of chaos. Ann. Sci. Ec Norm. Sup. 48, 891–940 (2015)

    MathSciNet  MATH  Google Scholar 

  • R.D. Hazeltine, The Framework of Plasma Physics (CRC Press, Boca Raton, 2018)

    Google Scholar 

  • R.D. Hazeltine, F.L. Waelbroeck, The Framework of Plasma Physics (Westview, London, 2004)

    Google Scholar 

  • J. Hubbard, The friction and diffusion coefficients of the Fokker-Planck equation in a plasma. II. Proc. R. Soc. Lond. A Math. Phys. Eng. Sci. 261(1306), 371–387 (1961)

    ADS  MathSciNet  MATH  Google Scholar 

  • S. Ichimaru, M.N. Rosenbluth, Relaxation processes in plasmas with magnetic field. Temperature relaxations. Phys. Fluids 13(11), 2778–2789 (1970)

    ADS  MATH  Google Scholar 

  • P.E. Jabin, Z. Wang, Mean field limit and propagation of chaos for Vlasov systems with bounded forces. J. Funct. Anal. 271(12), 3588–3627 (2016)

    MathSciNet  MATH  Google Scholar 

  • N.G. van Kampen, On the theory of stationary waves in plasmas. Physica 21(6–10), 949–963 (1955)

    ADS  MathSciNet  Google Scholar 

  • N.G. van Kampen, The dispersion equation for plasma waves. Physica 23(6–10), 641–650 (1957)

    ADS  MathSciNet  MATH  Google Scholar 

  • A.N. Kaufman, Reformulation of quasi-linear theory. J. Plasma Phys. 8(1), 1–5 (1972)

    ADS  Google Scholar 

  • M.K.H. Kiessling, The microscopic foundations of Vlasov theory for jellium-like Newtonian \(N\)-body systems. J. Stat. Phys. 155(6), 1299–1328 (2014)

    ADS  MathSciNet  MATH  Google Scholar 

  • R.H. Kraichnan, R. Panda, Depression of nonlinearity in decaying isotropic turbulence. Phys. Fluids 31(9), 2395–2397 (1988)

    ADS  MATH  Google Scholar 

  • S.M. Krivoruchko, V.A. Bashko, A.S. Bakai, Experimental investigations of correlation phenomena in the relaxation of velocity-spread beam in a plasma. Zhurnal Eksperimentalnoi i Teoreticheskoi Fiziki 80, 579–581 (1981). [English trans. Sov. Phys.-JETP 53: 292–298 (1981)]

    ADS  Google Scholar 

  • C. Lancellotti, From Vlasov fluctuations to the BGL kinetic equation. Il Nuovo cimento della Società italiana di fisica C 33(1), 111–119 (2010)

    Google Scholar 

  • L.D. Landau, On the vibrations of the electronic plasma. J. Phys. (USSR) 10(1), 25–34 (1946). Russian original: Zh.Eksp.Teor.Fiz. 16 (1946) 574-586

    MathSciNet  MATH  Google Scholar 

  • P.S. Laplace, Essai philosophique sur les probabilités. Bachelier, Paris (1840)

  • G. Laval, D. Pesme, Breakdown of quasilinear theory for incoherent 1-D Langmuir waves. Phys. Fluids 26(1), 52–65 (1983a)

    ADS  MATH  Google Scholar 

  • G. Laval, D. Pesme, Inconsistency of quasilinear theory. Phys. Fluids 26(1), 66–68 (1983b)

    ADS  MATH  Google Scholar 

  • G. Laval, D. Pesme, Self-consistency effects in quasilinear theory: a model for turbulent trapping. Phys. Rev. Lett. 53(3), 270–273 (1984)

    ADS  Google Scholar 

  • G. Laval, D. Pesme, Controversies about quasi-linear theory. Plasma Phys. Control. Fusion 41(3A), A239 (1999)

    ADS  Google Scholar 

  • Y.M. Liang, P. Diamond, Revisiting the validity of quasilinear theory. Phys. Fluids B Plasma Phys. 5(12), 4333–4340 (1993a)

    Google Scholar 

  • Y.M. Liang, P.H. Diamond, Weak turbulence theory of Langmuir waves: a reconsideration of the validity of quasilinear theory. Comments Plasma Phys. Control. Fusion 15, 139–149 (1993b)

    Google Scholar 

  • E.M. Lifshitz, L.P. Pitaevskii, Fizicheskaya kinetika, (Moscow, Nauka, 1979) [English trans. Landau and Lifshitz Course of Theoretical Physics 10: Physical Kinetics, Transl. Sykes JB and Franklin RN (Pergamon, Oxford, 1981)]

  • A. Macor, F. Doveil, Y. Elskens, Electrons climbing a “devils staircase” in wave-particle interaction. Phys. Rev. Lett. 95, 264102 (2005). (4pp)

    ADS  Google Scholar 

  • W.D. McComb, Renormalization Methods: A Guide for Beginners (Oxford University Press, Oxford, 2004)

    MATH  Google Scholar 

  • N. Meyer-Vernet, Aspects of Debye shielding. Am. J. Phys. 61(3), 249–257 (1993)

    ADS  Google Scholar 

  • H.E. Mynick, A.N. Kaufman, Soluble theory of nonlinear beam-plasma interaction. Phys. Fluids 21(4), 653–663 (1978)

    ADS  MATH  Google Scholar 

  • H. Neunzert, An introduction to the nonlinear Boltzmann-Vlasov equation. in Kinetic Theories and the Boltzmann Equation, ed. by C. Cercignani, Springer, Berlin, no. 1048 in Lect. Notes Math., pp 60–110 (1984)

    Google Scholar 

  • H. Neunzert, J. Wick, Die Approximation der Lösung von Integro-Differentialgleichungen durch endliche Punktmengen. in Numerische Behandlung nichtlinearer Integrodifferential- und Differentialgleichungen, ed. by R. Ansorge, W. Törnig, Springer, Berlin, no. 395 in Lect. Notes Math., pp 275–290 (1974)

    Google Scholar 

  • C.S. Ng, A. Bhattacharjee, F. Skiff, Kinetic eigenmodes and discrete spectrum of plasma oscillations in a weakly collisional plasma. Phys. Rev. Lett. 83(10), 1974–1977 (1999)

    ADS  Google Scholar 

  • C.S. Ng, A. Bhattacharjee, F. Skiff, Weakly collisional Landau damping and three-dimensional Bernstein-Greene-Kruskal modes: new results on old problems. Phys. Plasmas 13(5), 055903 (2006)

    ADS  Google Scholar 

  • D.R. Nicholson, Introduction to Plasma Theory (Wiley, New York, 1983)

    Google Scholar 

  • T.M. O’Neil, Collisionless damping of nonlinear plasma oscillations. Phys. Fluids 8(12), 2255–2262 (1965)

    ADS  MathSciNet  Google Scholar 

  • T.M. O’Neil, J.H. Malmberg, Transition of the dispersion roots from beam-type to Landau-type solutions. Phys. Fluids 11(8), 1754–1760 (1968)

    ADS  Google Scholar 

  • T.M. O’Neil, J.H. Winfrey, J.H. Malmberg, Nonlinear interaction of a small cold beam and a plasma. Phys. Fluids 14(6), 1204–1212 (1971)

    ADS  Google Scholar 

  • I.N. Onishchenko, A.R. Linetskii, N.G. Matsiborko, V.D. Shapiro, V.I. Shevchenko, Contribution to the nonlinear theory of excitation of a monochromatic plasma wave by an electron beam. ZhETF Pis Red 12, 281–285 (1970). [Engl. Transl. 1960, JETP Lett.12 281–285 12: 281–285]

    Google Scholar 

  • A. Piel, Plasma Physics: An Introduction to Laboratory, Space, and Fusion Plasmas (Springer, Heidelberg, 2017)

    MATH  Google Scholar 

  • C. Roberson, K.W. Gentle, Experimental test of the quasilinear theory of the gentle bump instability. Phys. Fluids 14(11), 2462–2469 (1971)

    ADS  Google Scholar 

  • Y.A. Romanov, G.F. Filippov, Interaction of fast electron beams with longitudinal plasma waves. Zh Eksp Theor. Phys. 40, 123–132 (1961). [English trans. Sov. Phys.-JETP 13: 87–92 (1961)]

    Google Scholar 

  • M.N. Rosenbluth, W.M. MacDonald, D.L. Judd, Fokker-Planck equation for an inverse-square force. Phys. Rev. 107, 1–6 (1957)

    ADS  MathSciNet  MATH  Google Scholar 

  • N. Rostoker, Superposition of dressed test particles. Phys. Fluids 7(4), 479–490 (1964)

    ADS  MathSciNet  Google Scholar 

  • D.D. Ryutov, Landau damping: half a century with the great discovery. Plasma Phys. Control. Fusion 41(3A), A1–A12 (1999)

    ADS  Google Scholar 

  • E.E. Salpeter, On Mayer’s theory of cluster expansions. Ann. Phys. 5(3), 183–223 (1958)

    ADS  MathSciNet  MATH  Google Scholar 

  • V. Silin, On relaxation of electron and ion temperatures of fully ionized plasma in a strong magnetic field. SOVIET PHYSICS JETP 16(4), 1281–1285 (1963)

    ADS  MathSciNet  Google Scholar 

  • H. Spohn, Large Scale Dynamics of Interacting Particles (Springer Science & Business Media, Heidelberg, 2012)

    MATH  Google Scholar 

  • J.L. Tennyson, J.D. Meiss, P.J. Morrison, Self-consistent chaos in the beam-plasma instability. Phys. D 71, 1–17 (1994)

    MathSciNet  MATH  Google Scholar 

  • K. Theilhaber, G. Laval, D. Pesme, Numerical simulations of turbulent trapping in the weak beam-plasma instability. Phys. Fluids 30(10), 3129–3149 (1987)

    ADS  Google Scholar 

  • S.F. Tigik, L.F. Ziebell, P.H. Yoon, Collisional damping rates for plasma waves. Phys. Plasmas 23(6), 064504 (2016)

    ADS  Google Scholar 

  • S.I. Tsunoda, F. Doveil, J.H. Malmberg, Experimental test of the quasilinear theory of the interaction between a weak warm electron beam and a spectrum of waves. Phys. Rev. Lett. 58, 1112–1115 (1987a)

    ADS  Google Scholar 

  • S.I. Tsunoda, F. Doveil, J.H. Malmberg, An experimental test of quasilinear theory. Phys. Scr. 40, 204–205 (1987b)

    ADS  Google Scholar 

  • S.I. Tsunoda, F. Doveil, J.H. Malmberg, Experimental test of quasilinear theory. Phys. Fluids B 3, 2747–2757 (1991)

    ADS  Google Scholar 

  • A.A. Vedenov, E.P. Velikhov, R.Z. Sagdeev, Quasilinear theory of plasma oscillations. Nuclear Fusion Suppl. 2, 465–475 (1962)

    Google Scholar 

  • A. Volokitin, C. Krafft, Velocity diffusion in plasma waves excited by electron beams. Plasma Phys. Control. Fusion 54(8), 085002 (2012)

    ADS  Google Scholar 

  • P.H. Yoon, L.F. Ziebell, E.P. Kontar, R. Schlickeiser, Weak turbulence theory for collisional plasmas. Phys. Rev. E 93(3), 033203 (2016)

    ADS  MathSciNet  Google Scholar 

Download references

Acknowledgements

D. F. E. is grateful to the members of Equipe Turbulence Plasma in Marseilles, since the theory reviewed in this paper is the result of three decades of collaboration with them. He thanks Professor M. Kikuchi for suggesting him to write this review. He also thanks Professor A. Sen for many useful suggestions, and Professor P. Huneman for pointing out to him the book “Reductionism, emergence and levels of reality” by Chibbaro et al. He thanks Drs F. Bonneau, M.-C. Firpo, and F. Sattin for helpful comments on the manuscript. Also D. F. G. Minenna who brought the precious views of a newcomer in the field. One of the authors (D. Z.) has been supported by the A*MIDEX project (no. ANR-11-IDEX-0001-02) funded by the “Investissements d’Avenir” French Government program, managed by the French National Research Agency (ANR).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to D. F. Escande.

Ethics declarations

Conflict of interest statement

On behalf of all authors, the corresponding author states that there is no conflict of interest.

Appendices

Appendix 1: Rigorous derivation à la Kaufman

A(t) is entire

The dynamics considered in Sect. 3 corresponds to the linearized motion of the N electrons with respect to a given multi-beam-multi-array, when a single wave with wavevector \(\mathbf {k}_{\mathbf {m}}\) is excited at \(t=0\). Let \(\mathbf {r}_{j0}\) be the initial position of the unperturbed beam particle with index j, and \(\mathbf {v}_{j}\) be its velocity, and let \(\varDelta {\mathbf {r}}_j (t)= \mathbf {r}_j (t) - \mathbf {r}_{j0} - \mathbf {v}_j t\) be the mismatch of the actual position of particle j with respect to the unperturbed beam particle with the same index. Setting \(\mathbf {r}_j = \mathbf {r}_{j0} + \mathbf {v}_j t + \varDelta {\mathbf {r}}_j (t)\) in Eq. (8), we replace \(\tilde{\varphi }\) with its expansion to first order in the \(\varDelta {\mathbf {r}}_j (t)\)’s

$$\begin{aligned} \tilde{\varphi } (\mathbf {m},t) = - \sum _{l = 1}^N \frac{ e}{\varepsilon _0 k_{\mathbf {m}}^2} \exp [- {\mathrm {i}}\mathbf {k}_{\mathbf {m}} \cdot (\mathbf {r}_{l0} + \mathbf {v}_{l} t)] \ [1 - {\mathrm {i}}\mathbf {k}_{\mathbf {m}} \cdot \varDelta {\mathbf {r}}_l(t)]. \end{aligned}$$
(101)

Using Eqs. (8 and 101), the linearized particle dynamics defined by Eq. (9) is then given by

$$\begin{aligned} \varDelta \ddot{\mathbf {r}}_j = \frac{{\mathrm {i}}e}{L^3 m} \mathbf {k}_{\mathbf {m}} \ \tilde{\varphi }(\mathbf {m},t) \exp [{\mathrm {i}}\mathbf {k}_{\mathbf {m}} \cdot (\mathbf {r}_{j0} + \mathbf {v}_{j} t)] + {\mathrm {c.c.}}\,. \end{aligned}$$
(102)

Because of ordering (10), for each beam the corresponding values of \(\exp [{\mathrm {i}}\mathbf {k}_{\mathbf {m}} \cdot \mathbf {r}_{l0}]\) are uniformly distributed on the unit circle, and their global contribution to the 1 factor in the last bracket of Eq. (101) vanishes. Therefore Eq. (102) yields the compact expression

$$\begin{aligned} \varDelta \ddot{\mathbf {r}}_j = - \frac{2 \omega _{{\mathrm {p}}}^2}{N k_{\mathbf {m}}^2} \mathbf {k}_{\mathbf {m}} \ \sum _{l = 1}^N \cos [\mathbf {k}_{\mathbf {m}} \cdot [\mathbf {r}_{j0} - \mathbf {r}_{l0} + (\mathbf {v}_{j} - \mathbf {v}_{l}) \, t \, ] \, \mathbf {k}_{\mathbf {m}} \cdot \varDelta {\mathbf {r}}_l(t). \end{aligned}$$
(103)

This defines a system of N linear differential equations whose coefficients are entire functions of t. Therefore, the \(\varDelta {\mathbf {r}}_l(t)\)’s are entire functions. Through Eq. (101), this property is transferred to \(\tilde{\varphi } (\mathbf {m},t)\) and to the amplitude \(A(t) = \tilde{\varphi } (\mathbf {m},t) \exp ({\mathrm {i}}\omega t)\) for \(\omega \) real: A(t) is an entire function.

If the \(\mathbf {k}_{\mathbf {m}} \cdot \mathbf {v}_{j}\)’s are multiples of a given number, Eq. (103) has coefficients with some period T, and belongs to the Floquet class of differential equations. Then its solutions are of the type

$$\begin{aligned} U(t) = V(t) {\mathrm {e}}^{\alpha t} , \end{aligned}$$
(104)

where V(t) is a vector of period T, and \(\alpha \) a complex number. The corresponding Floquet exponents are the \(\beta _\sigma \)’s introduced at the end of Sect. 5.3. A similar equation was met in the self-consistent wave–particle approach introduced in Sect. 9.1 (section 3 of Elskens and Escande 2002). Its full solution turned out to be a superposition of wave-like and ballistic solutions. This is remarkable, since such equations are generally not explicitly solvable with elementary functions, even for the simplest one, the Mathieu equation.

Solution to all orders

By expanding \(A(t-\tau )\) in Taylor series, the second term of Eq. (16) becomes (with \(k = k_{\mathbf {m}}\) and \(v_\phi = \omega /k)\))

$$\begin{aligned} S= & {} - {\mathrm {i}}\omega _{\mathrm {p}}^2 \sum _{n=0}^{+\infty } \frac{{\mathrm {d}}^n A(t)}{{\mathrm {d}}t^n} \int \frac{\partial }{\partial \varOmega } \int _0^t (-1)^n\frac{\tau ^n}{n!} \exp ( - {\mathrm {i}}\varOmega \tau ) \, {\mathrm {d}}\tau \, g(v) {\mathrm {d}}v \nonumber \\= & {} - {\mathrm {i}}\omega _{\mathrm {p}}^2 \sum _{n=0}^{+\infty } \frac{{\mathrm {d}}^n A(t)}{{\mathrm {d}}t^n} \int \frac{\partial }{\partial \varOmega } \int _0^t \frac{{\mathrm {i}}^{3n}}{n!} \frac{\partial ^n}{\partial \varOmega ^n} \exp ( - {\mathrm {i}}\varOmega \tau ) \, {\mathrm {d}}\tau \, g(v) {\mathrm {d}}v \nonumber \\= & {} \omega _{\mathrm {p}}^2 \sum _{n=0}^{+\infty } \frac{{\mathrm {i}}^{3n}}{n!} \frac{{\mathrm {d}}^n A(t)}{{\mathrm {d}}t^n} \int \frac{\partial ^{n+1}}{\partial \varOmega ^{n+1}} \frac{\exp ( - {\mathrm {i}}\varOmega t)-1}{\varOmega } \, g(v) {\mathrm {d}}v \nonumber \\= & {} \omega _{\mathrm {p}}^2 \sum _{n=0}^{+\infty } \frac{{\mathrm {i}}^{3n}}{n! k^{n+1}} \frac{{\mathrm {d}}^n A(t)}{{\mathrm {d}}t^n} \int (-1)^{n+1} g^{(n+1)}(v) \, \frac{\exp ( - {\mathrm {i}}\varOmega t)-1}{\varOmega } {\mathrm {d}}v \nonumber \\= & {} - \omega _{\mathrm {p}}^2 \sum _{n=0}^{+\infty } \frac{{\mathrm {i}}^{n}}{n! k^{n+1}} \frac{{\mathrm {d}}^n A(t)}{{\mathrm {d}}t^n} \int g^{(n+1)}(v) \frac{\exp ( - {\mathrm {i}}\varOmega t)-1}{\varOmega } {\mathrm {d}}v \nonumber \\&\quad \sim _{t\rightarrow +\infty } \omega _{\mathrm {p}}^2 \sum _{n=0}^{+\infty } \frac{{\mathrm {i}}^{n}}{n! k^{n+1}} \frac{{\mathrm {d}}^n A(t)}{{\mathrm {d}}t^n} \left[ {\mathrm{P}} \! \! \! \int \frac{g^{(n+1)}(v)}{\varOmega } {\mathrm {d}}v + {\mathrm {i}}\frac{\pi }{k} g^{(n+1)}(v_\phi ) \right] . \end{aligned}$$
(105)

The time evolution of A(t), and the frequency \(\omega \), are then derived from

$$\begin{aligned} 1 = \frac{\omega _{\mathrm {p}}^2}{A} \sum _{n=0}^{+\infty } \frac{{\mathrm {i}}^{n}}{n! k^{n+1}} \frac{{\mathrm {d}}^n A(t)}{{\mathrm {d}}t^n} \left[ {\mathrm{P}} \! \! \! \int \frac{g^{(n+1)}(v)}{\varOmega } {\mathrm {d}}v + {\mathrm {i}}\frac{\pi }{k} g^{(n+1)}(v_\phi ) \right] . \end{aligned}$$
(106)

Looking for a solution \(A(t)=A(0)e^{\nu t}\) with \(\nu \) real, Eq. (106) reads

$$\begin{aligned} 1 = \omega _{\mathrm {p}}^2 \sum _{n=0}^{+\infty }\frac{{\mathrm {i}}^{n}}{n! k^{n+1}} \nu ^n \left[ {\mathrm{P}} \! \! \! \int \frac{g^{(n+1)}(v)}{\varOmega } {\mathrm {d}}v + {\mathrm {i}}\frac{\pi }{k} g^{(n+1)}(v_\phi ) \right] , \end{aligned}$$
(107)

whose real part is

$$\begin{aligned} 1 = \omega _{\mathrm {p}}^2 \sum _{p=0}^{+\infty } \frac{(-1)^p}{(2p)! k^{2p+1}} \nu ^{2p} \left[ {\mathrm{P}} \! \! \! \int \frac{g^{(2p+1)}(v)}{\varOmega } {\mathrm {d}}v - \frac{\pi \nu }{(2p+1)k^2} g^{(2p+2)}(v_\phi )\right] , \end{aligned}$$
(108)

and imaginary part is

$$\begin{aligned} 0 = \sum _{p=0}^{+\infty } \frac{(-1)^p\nu ^{2p}}{(2p+1)!k^{2p}} \left[ \nu {\mathrm{P}} \! \! \! \int \frac{g^{(2p+2)}(v)}{\varOmega } {\mathrm {d}}v +\pi (2p+1)g^{(2p+1)}(v_\phi ) \right] . \end{aligned}$$
(109)

Since \(g^{(n+1)}(v) \sim v_{{\mathrm {T}}}^{-n}g'(v)\), Eqs. (108) and (109) may be seen as expansions in the small parameter \(\varepsilon \equiv \nu / (k \lambda _{\mathrm {D}}\omega _{{\mathrm {p}}})\). Then, these equations may be solved order by order. The first two orders were used in Sect. 3.1.

It is also noteworthy that, when g(v) is analytic, Eq. (107) reads

$$\begin{aligned} 1 = \frac{\omega _{\mathrm {p}}^2}{k} \left[ {\mathrm{P}} \! \! \! \int \frac{g'(v + {\mathrm {i}}\nu / k)}{\varOmega } {\mathrm {d}}v + {\mathrm {i}}\frac{\pi }{k} g'(v_\phi +i\nu /k) \right] , \end{aligned}$$
(110)

which is exactly the formula obtained by Landau.

Appendix 2: Relevance of transients before Landau damping

In this Appendix, we discuss the relevance of transients that would occur before the self-consistent electrostatic field may experience Landau damping. To do so, we have to specify how the self-consistent electrostatic field has actually been generated, an issue that is usually eluded. This implies that we do account for the external drive (e.g. a laser, a polarized grid, electrodes...) used to induce the self-consistent field in the plasma, when calculating the electron motion. The calculation is, therefore, slightly more general than that leading to Eqs. (1418). Such a generalization has already been performed in Bénisti and Gremillet (2007) when the self-consistent field was slowly driven. An important point of Bénisti and Gremillet (2007) was to prove that a wave may be considered as slowly varying, provided that its complex amplitude did not change much during a time interval of the order of \(\tau _{\mathrm{mix}}=(k v_{\mathrm {T}})^{-1}\). More precisely, the typical wave growth rate, \(\gamma \), had to be such that \(\vert \gamma \vert \tau _{\mathrm{mix}}<0.1\).

We now further discuss the importance of the product \(\vert \gamma \vert \tau _{\mathrm{mix}}\) in terms of the transients. For the sake of simplicity, the discussion is restricted to the situation when the electrons only feel the effect of the drive during a finite time interval, namely when \(0<t<\tau _{\mathrm {d}}\) (see Bénisti and Gremillet 2007 for a calculation that does not make use of this hypothesis). We also assume that the total force, including the effect of the self-consistent field and the drive, derives from an effective potential which reads \(\varphi =A(t) \exp [{\mathrm {i}}(\mathbf {k_m}.\mathbf {r}-\omega t)] + {\mathrm {c.c.}}\) (which has been proved to be correct in Bénisti and Gremillet 2007 when the plasma wave was laser driven). Because we include the effect of the drive, we can now integrate the electrons motion from the time when they are at equilibrium, \(\mathbf {r}_j=\mathbf {r}_{j0}\) when \(t=0\). Then, Eq. (14) is changed into

$$\begin{aligned} \varDelta \mathbf {r}_{j1}(t) = \alpha \mathbf {k}_{\mathbf {m}} \int _{0}^{t} \tau A(t-\tau ) \exp [{\mathrm {i}}(\varOmega _j (t-\tau ) + \mathbf {k}_{\mathbf {m}} \cdot \mathbf {r}_{j0})] {\mathrm {d}}\tau + {\mathrm {c.c.}} \end{aligned}$$
(111)

If the wave is slowly driven, A(t) is a slowly varying function, and one may stop the Taylor expansion of \(A(t-\tau )\) at first order, which yields

$$\begin{aligned} A(t)= & {} \int \frac{\omega _{{\mathrm {p}}}^2}{k_{\mathbf {m}}} A(t) g'(v) \frac{1-\cos (\varOmega t)+ {\mathrm {i}}\sin (\varOmega t)}{\varOmega } {\mathrm {d}}v \nonumber \\&- \, {\mathrm {i}}\frac{\omega _{{\mathrm {p}}}^2}{k_{\mathbf {m}}^2} \dot{A}(t) \int g^{\prime \prime }(v) \frac{\cos (\varOmega t)- {\mathrm {i}}\sin (\varOmega t) -1}{\varOmega } {\mathrm {d}}v, \end{aligned}$$
(112)

which is the same as Eq. (18) except that the term proportional to h(v) no longer appears, since the calculation has been performed with \(\delta \mathbf {r}_{ j}=\mathbf {0}\). Hence, unlike in Eq. (18), no transient is expected from this term. Now, when \(t>\tau _{\mathrm {d}}\) it is clear that A(t) is nothing but the amplitude of the self-consistent potential, since the electrons no longer feel the effect of the drive. Moreover, because it is slowly driven, only when \(\tau _{\mathrm {d}}\gg \tau _{\mathrm{mix}}\) may the self-consistent field reach a significant amplitude, and may effectively be Landau damped. Hence, when \(t>\tau _{\mathrm {d}}\), Eq. (112) leads to Eq. (19). This means that, if the wave is slowly driven, it is Landau damped just after the drive has been turned off, and there is no transient.

Let us now investigate the situation when the wave may no longer be considered as slowly varying when \(t < \tau _{\mathrm {d}}\). Then, Eq. (112) is no longer valid, and we have to pay a specific attention to the electron motion during the driving phase (\(t\le \tau _{\mathrm {d}})\). To do so, we still assume that the electrons are at equilibrium when \(t=0\). Moreover, we use Eq. (111), which is exact, to calculate the shift in their positions at \(t=\tau _{\mathrm {d}}\), induced by the drive and the self-consistent electric fields. This yields

$$\begin{aligned} \varDelta \mathbf {r}_{j1}(t)= & {} \alpha \mathbf {k}_{\mathbf {m}} \int _{0}^{\tau _{\mathrm {d}}} \tau A(\tau _{\mathrm {d}}-\tau ) \exp [{\mathrm {i}}(\varOmega _j (\tau _{\mathrm {d}}-\tau ) + \mathbf {k}_{\mathbf {m}} \cdot \mathbf {r}_{j0})] {\mathrm {d}}\tau + {\mathrm {c.c.}}\nonumber \\\equiv & {} \delta \mathbf {r}_{ j} \sin (\mathbf {k}_{\mathbf {m}} \cdot \mathbf {r}_{j0}+\psi _0). \end{aligned}$$
(113)

As noted in Sect. 3.1, the phase \(\psi _0\) has no importance in the derivation, and we henceforth drop it. Using the result of Eq. (113), we now calculate the shift in position when \(t>\tau _{\mathrm {d}}\), which will let us conclude about the evolution of the wave amplitude. Using again Eq. (111), we find

$$\begin{aligned} \varDelta \mathbf {r}_{j_1}(t)= & {} \delta \mathbf {r}_{ j} \sin (\mathbf {k}_{\mathbf {m}} \cdot \mathbf {r}_{j0}) + \left\{ \alpha \mathbf {k_m}\int _{\tau _{\mathrm {d}}}^t \tau \mathcal {A}(t-\tau ) {\mathrm {d}}\tau \right. \nonumber \\&+ \left. \alpha \mathbf {k_m} \int _0^{\tau _{\mathrm {d}}} \tau \left[ \mathcal {A}(t-\tau )-\mathcal {A}(\tau _{\mathrm {d}}-\tau ) \right] {\mathrm {d}}\tau + {\mathrm {c.c.}}\right\} \ , \end{aligned}$$
(114)

where we introduced

$$\begin{aligned} \mathcal {A}(t) \equiv A(t)\exp [{\mathrm {i}}(\varOmega _j t + \mathbf {k}_{\mathbf {m}} \cdot \mathbf {r}_{j0})]. \end{aligned}$$
(115)

In Sect. 3.1, the limit \(\tau _{\mathrm {d}}\rightarrow 0\) was considered, which yields Eq. (14). In this Appendix, we specify how small \(\tau _{\mathrm {d}}\) has to be for the results of Sect. 3.1 to be valid. First, we want the third term in the right-hand side of Eq. (114) to be negligible, which is only true if \(t\gg \tau _{\mathrm {d}}\) and if A(t) does not abruptly vanish within a time interval smaller than \(\tau _{\mathrm {d}}\). The first condition implies that the results of Sect. 3.1 are only valid when \(t\gg \tau _{\mathrm {d}}\). The second condition is only true provided that \(\gamma _\mathrm{L} \tau _{\mathrm {d}}\ll 1\).

Neglecting the third term of Eq. (114) and following the same steps as in Sect. 3.1, one finds that Eq. (18) is changed into

$$\begin{aligned} A(t) =&\int \left[ \frac{Ne}{2\varepsilon _0 k_{\mathbf {m}}^2 }\, h(v) \exp (- {\mathrm {i}}\varOmega t) \right. \nonumber \\&\left. + \frac{\omega _{{\mathrm {p}}}^2}{k_{\mathbf {m}}} A(t) g'(v) \frac{{\mathrm {e}}^{-{\mathrm {i}}\varOmega \tau _{\mathrm {d}}}-\cos (\varOmega t)+ {\mathrm {i}}\sin (\varOmega t)}{\varOmega } \right] {\mathrm {d}}v \nonumber \\&- \, {\mathrm {i}}\frac{\omega _{{\mathrm {p}}}^2}{k_{\mathbf {m}}^2} \dot{A}(t) \int g^{\prime \prime }(v) \frac{\cos (\varOmega t)- {\mathrm {i}}\sin (\varOmega t) - {\mathrm {e}}^{-{\mathrm {i}}\varOmega \tau _{\mathrm {d}}}}{\varOmega } {\mathrm {d}}v. \end{aligned}$$
(116)

Then, Eq. (18) is recovered only in the limit \(\varOmega \tau _{\mathrm {d}}\rightarrow 0\). Hence, this equation is only valid provided that the wave could reach a significant amplitude, due to the external drive, during a time much smaller than \(\tau _{\mathrm{mix}}\) and the plasma period. When the latter condition is fulfilled, Eq. (18) implies that, once the drive is turned off, one has to wait for a time of the order of \(\tau _{\mathrm{mix}}\) before Landau damping is effective.

Therefore, we recover here the usual difference in a system’s response to a sudden excitation compared to an adiabatic one. When a system is subjected to a sudden change, it usually rapidly oscillates before entering a stationary, or slowly varying, regime. These transient oscillations do not exist under an adiabatic-like external force, and the system keeps on varying in a smooth way.

Note that we only derived the wave evolution in two opposite limits, either when the driving time, \(\tau _{\mathrm {d}}\), was much smaller than \(\tau _{\mathrm{mix}}\) and the plasma period, or when it was much smaller than \(\tau _{\mathrm{mix}}\). The situation when \(\tau _{\mathrm {d}}\) is either of the order of the plasma period or of \(\tau _{\mathrm{mix}}\) would deserve further investigation.

Appendix 3: Infinite number of beams

This appendix provides calculations similar to those in Dawson (1960), but reformulated in a form suitable for the derivation of Sect. 5, and with the correction of several errors. We first focus on \( \epsilon _{{\mathrm {d}}1}(\mathbf {m},\omega )\) defined by Eq. (51). To separate the regular and the singular parts of this quantity, we add and subtract to the right-hand side of Eq. (51) the quantity

$$\begin{aligned} \omega _{{\mathrm {p}}}^2\left[ \frac{\pi ^2 g(\omega /k)}{k^2 \delta \sin ^2(\pi \omega /k \delta )}-\frac{2 \pi g'(\omega /k)}{k^2}\cot (\pi \omega /k \delta )\right] , \end{aligned}$$
(117)

where \(k=k_{\mathbf {m}}\). This yields

$$\begin{aligned} \epsilon _{{\mathrm {d}}1}(\mathbf {m},\omega )=&1 - \omega _{{\mathrm {p}}}^2 \left[ \frac{\pi ^2 g(\omega /k)}{k^2 \delta \sin ^2(\pi \omega /k \delta )}-\frac{2 \pi g'(\omega /k)}{k^2} \cot (\pi \omega /k \delta ) \right. \nonumber \\&\left. + \varSigma _{\sigma = - \infty }^{\infty }\left( \frac{[g(\sigma \delta ) -g(\omega /k)]\delta }{ (\omega - \sigma k \delta )^2}+\frac{2 g'(\omega /k) \omega \delta }{k (\omega ^2 - (\sigma k \delta )^2)} \right) \right] , \end{aligned}$$
(118)

where use has been made of the relations (see Abramowitz and Stegun 1967)Footnote 44

$$\begin{aligned} \frac{\pi ^2 }{\sin ^2(\pi x)}= & {} \varSigma _{\sigma = - \infty }^{\infty }\frac{1}{ (x - \sigma )^2} \end{aligned}$$
(119)
$$\begin{aligned} \pi \cot (\pi x)= & {} \varSigma _{\sigma = - \infty }^{\infty }\frac{x}{ (x^2 - \sigma ^2)}. \end{aligned}$$
(120)

We notice that the function of \(\sigma \delta \) inside the summation in Eq. (118) has no real poles. Therefore, the sum passes smoothly to an integral as \(\delta \) goes to zeroFootnote 45. In this limit, Eq. (118) becomes

$$\begin{aligned} \epsilon _{{\mathrm {d}}1}(\mathbf {m},\omega )=&1 - \omega _{{\mathrm {p}}}^2 \left[ \frac{\pi ^2 g(\omega /k)}{k^2 \delta \sin ^2(\pi \omega /k \delta )}-\frac{2 \pi g'(\omega /k)}{k^2} \cot (\pi \omega /k \delta ) \right. \nonumber \\&\qquad \left. + \int _{- \infty }^{\infty }\left( \frac{g'(v)}{ k(\omega - k v)}+\frac{2 g'(\omega /k) \omega }{k (\omega ^2 - k^2 v^2)} \right) {\mathrm {d}}v \right] . \end{aligned}$$
(121)

We now compute the zeros of \(\epsilon _{{\mathrm {d}}1}(\mathbf {m},\omega )\) and write \(\omega = \alpha + {\mathrm {i}}\beta \). We first consider those with \(\beta \) vanishing when \(\delta \) goes to zero. If \(\beta \) vanished like or faster than \(\delta \), the first term in the bracket of Eq. (121) would diverge, while the second and third one would remain finite, which is impossible. Therefore, \(\beta \) vanishes slower than \(\delta \), which forces the cotangent to converge toward \(- \mu \, {\mathrm {i}}\), where \(\mu = \pm 1\) is the sign of \(\beta \). Then, to stay finite, the first term requires \(\beta \) to scale like \(\delta \, |\ln (\delta /v_{{\mathrm {T}}})\,|\). With this in mind, and looking for solutions in the vicinity of \(n k \delta \), Eq. (121) requires

$$\begin{aligned}&\omega _{{\mathrm {p}}}^2 \left[ - \frac{4 \pi ^2 g(n \delta ) \exp [2 \pi ({\mathrm {i}}\mu \alpha _1 - |\beta |)/k \delta ]}{k^2 \delta }+\frac{{\mathrm {i}}\mu \pi g'(n \delta )}{k^2}\right] \nonumber \\&\quad = 1 + {\mathrm{P}} \! \! \! \int _{- \infty }^{\infty } \frac{\omega _{{\mathrm {p}}}^2 g'(v)}{ k^2(n \delta - v)} \, {\mathrm {d}}v, \end{aligned}$$
(122)

for \(n \delta \) in the support of g, with \(\alpha _1 = \alpha - n k \delta \). Equation (121) provides two contributions to the term in \(g'(n \delta )\): one from the term in cotangent, and one from the pole of the term in \(g'(v)\) in the integral, while the two poles of the term in \(g'(\omega /k)\) bring contributions cancelling each other. Solving for \(\alpha _1\) and \(\beta \) yields

$$\begin{aligned}&\tan \frac{2 \pi \alpha _1}{k \delta } = - \frac{\pi \omega _{{\mathrm {p}}}^2}{k^2} g'(n \delta )/\left[ 1 + {\mathrm{P}} \! \! \! \int _{- \infty }^{\infty } {\mathrm {d}}v \frac{\omega _{{\mathrm {p}}}^2 g'(v)}{ k^2(n \delta - v)}\right] \, , \end{aligned}$$
(123)
$$\begin{aligned}& \beta = \mu \frac{k \delta }{2 \pi } \ln \left\{ \left[ \frac{k^2 \delta }{4 \pi ^2 \omega _{{\mathrm {p}}}^2 g(n \delta )}\right] \right. \times \nonumber \\& \qquad \qquad \qquad \left. \left[ \left( 1 \! \! + {\mathrm {P}}\! \! \! \int _{- \infty }^{\infty } \! \! \! {\mathrm {d}}v \frac{\omega _{{\mathrm {p}}}^2 g'(v)}{ k^2(n \delta - v)} \right) ^2 \! \! + \! \left( \frac{\pi \omega _{{\mathrm {p}}}^2 g'(n \delta )}{k^2} \right) ^2 \right] \right\}. \end{aligned}$$
(124)

Equation (123) yields \(2 \pi \alpha _1/k \delta \) modulo \(\pi \), and the right solution is obtained by requiring \(\cos (2 \pi \mu \alpha _1/k \delta )\) to have the opposite sign to the denominator of this equation.

Like the natural frequencies of the beams, the roots are spaced \(k \delta \) apart in \(\alpha \). Therefore, the above zeros have real parts between these natural frequencies. There are two roots for each beam, since \(\mu \) can be either positive or negative, and no root away form the support of g. Thus we obtain two modes for each beam, as required.

We now compute \( \frac{\partial \epsilon }{\partial \omega } (\mathbf {m}, \omega )\) defined in Eq. (54) for the case of a vanishing imaginary part of \(\omega \) when \(\delta \) goes to zero. Here again, we handle the singularity by adding and subtracting to the right-hand side of Eq. (54) the quantity

$$\begin{aligned} 2 \omega _{{\mathrm {p}}}^2\left[ - \frac{\pi ^3 \cos (\pi \omega /k \delta ) g(\omega /k)}{k^3 \delta ^2 \sin ^3(\pi \omega /k \delta )} - \frac{\pi ^2 g'(\omega /k)}{k^2 \delta \sin ^2(\pi \omega /k \delta )} + \frac{\pi g^{\prime \prime }(\omega /k)}{2 k^3}\cot (\pi \omega /k \delta )\right] . \end{aligned}$$
(125)

These terms can also be written in the form of sums using again Eqs. (119 and 120), and (see Abramowitz and Stegun 1967)

$$\begin{aligned} \pi ^3 \frac{\cos (\pi x) }{\sin ^3(\pi x)}= & {} \varSigma _{\sigma = - \infty }^{\infty }\frac{1}{ (x - \sigma )^3}. \end{aligned}$$
(126)

Using the latter expression, for \(\delta \) small we find that \( \epsilon '_{\sigma ,\mu }\) is given by Eq. (56).

Appendix 4: Approaching the singular limit by coarse-graining

In the introduction, we recalled the pulverization procedure for deriving Vlasov equation from the BBGKY hierarchy. The singular limit can be obtained by a coarse-graining procedure, which is germane to the pulverization procedure: each particle is substituted by a continuum of particles with velocities close to its velocity, with a mismatch in velocity \(\varDelta \mathbf {v}\) distributed with the continuous distribution \(P(\varDelta \mathbf {v})\), instead of a discrete distribution in the case of the pulverizationFootnote 46. This procedure may be viewed as the counterpart in velocity of the quantum regularization of small spatial scales for collisions recalled in the introduction. Indeed, the coarse-graining in velocity may appear as a way to account for the quantum uncertainty on the particle velocities.

The calculation leading to Eq. (47) can be performed again, but the summation over particles now involves an integral over the nearby velocities of the coarse-grained system. Then \(\varphi ^{({\mathrm{bal}})}(\mathbf {m},\omega )\) is substituted with

$$\begin{aligned} \varphi _{{\mathrm{cg}},j}^{({\mathrm{bal}})}(\mathbf {m},\omega ) = - \frac{{\mathrm {i}}e}{\varepsilon _0 k_{\mathbf {m}}^2} \int \frac{\exp [- {\mathrm {i}}\mathbf {k}_{\mathbf {m}} \cdot \mathbf {r}_j(0)]}{\omega -\mathbf {k}_{\mathbf {m}} \cdot ({\dot{\mathbf {r}}}_j(0) + \mathbf {u})} P(\mathbf {u}) {\mathrm {d}}^3 \mathbf {u}, \end{aligned}$$
(127)

and

$$\begin{aligned} f_0(\mathbf {v}) = \sum _{\sigma = 1}^{n_{\mathrm {b}}} N_\sigma P(\mathbf {v}-\mathbf {w}_\sigma ) , \end{aligned}$$
(128)

where \(n_{\mathrm {b}}\) is the number of beams. If the width of P is large with respect to the edge of an elementary cube of the velocity grid, \(f_0\) is a smooth function, the grid may be taken as very tight, and for practical purposes \(P(\mathbf {u})\) may be taken as a Dirac distribution in Eq. (127), which then becomes Eq. (46).

Appendix 5: Shielded Coulomb potential by a singular limit of the many-beam description

If in Eq. (65) we take first the limit \(\delta \rightarrow 0\) and then the limit \(\varepsilon \rightarrow 0^{+}\), according to Eq. (121), \(\epsilon _{\mathrm {d}}(\mathbf {m},\mathbf {k}_{\mathbf {m}} \cdot \mathbf {v}+ {\mathrm {i}}\varepsilon )\) converges toward

$$\begin{aligned} \epsilon _{\mathrm{lim}} (\mathbf {m},\mathbf {k}_{\mathbf {m}} \cdot \mathbf {u}) = 1 + \omega _{{\mathrm {p}}}^2 \, {\mathrm{P}} \! \! \! \int \frac{\frac{\partial g}{\partial v}(v)}{k_{\mathbf {m}}(\mathbf {k}_{\mathbf {m}} \cdot \mathbf {u}- k_{\mathbf {m}} v)} {\mathrm {d}}v - {\mathrm {i}}\mu \frac{\pi \omega _{{\mathrm {p}}}^2}{k_{\mathbf {m}}^2} \frac{\partial g}{\partial v}\left( \frac{\mathbf {k}_{\mathbf {m}} \cdot \mathbf {u}}{k_{\mathbf {m}}}\right) . \end{aligned}$$
(129)

This is nothing but the contribution of \(\epsilon (\mathbf {m},\mathbf {k}_{\mathbf {m}} \cdot \mathbf {u}+ {\mathrm {i}}\varepsilon )\) in the same limit (see for instance equation (9.12) of Nicholson 1983). Therefore, for \(\delta \) small enough, Eq. (64) becomes Eq. (68), and Eq. (65) becomes Eq. (69).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Escande, D.F., Bénisti, D., Elskens, Y. et al. Basic microscopic plasma physics from N-body mechanics. Rev. Mod. Plasma Phys. 2, 9 (2018). https://doi.org/10.1007/s41614-018-0021-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s41614-018-0021-x

Keywords

Navigation