Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Plasma Physics
Plasma Physics
Plasma Physics
Ebook737 pages6 hours

Plasma Physics

Rating: 0 out of 5 stars

()

Read preview

About this ebook

A historic snapshot of the field of plasma physics, this fifty-year-old volume offers an edited collection of papers by pioneering experts in the field. In addition to assisting students in their understanding of the foundations of classical plasma physics, it provides a source of historic context for modern physicists. Highly successful upon its initial publication, this book was the standard text on plasma physics throughout the 1960s and 70s.
Hailed by Science magazine as a "well executed venture," the three-part treatment ranges from basic plasma theory to magnetohydrodynamics and microwave plasma physics. Highlights include Klimontovich's article on quantum plasmas, Buneman's writings on how to distinguish between attenuating and amplifying waves, and Yoler's clear and cogent review of magnetohydrodynamics. Professional atomic and plasma physicists and all students of plasma physics will appreciate this historic resource.
LanguageEnglish
Release dateSep 26, 2013
ISBN9780486320588
Plasma Physics

Related to Plasma Physics

Related ebooks

Physics For You

View More

Related articles

Reviews for Plasma Physics

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Plasma Physics - James E. Drummond

    Index

    CHAPTER 1

    INTRODUCTION (OSCILLATIONS IN PLASMAS)

    by J. E. Drummond

    1. History of Plasma Oscillations

    1.1. The First Problem in Plasma Physics.Plasma physics has been a recognizable field of study from about the time (1929) that Tonks and Langmuir noted [1]† the similarity of certain newly discovered oscillations in ionized gases to the oscillations of a jelly plasma. They christened the new oscillations plasma-electron oscillations and also gave the name plasma to the nearly neutral part of an ionized gas. The term is now taken in the technical sense as meaning the portion of a material body which is much larger than a Debye length and throughout which charged particles move. (See pages 2, 12, and 21 for discussion of Debye length.)

    Strong internal oscillations in an ionized gas were first proposed by Dittmer [2] to explain the anomalous high scattering observed in electric arcs. Langmuir [3] had discovered in arc discharges the presence of electrons with energies considerably greater than the total potential difference across the tube. Dittmer, using Langmuir’s probe technique [4], found the anomaly especially pronounced within a thin region a few millimeters from the filament. He hypothesized the existence of internal oscillations with periods comparable to the transit time of the primary (50-volt) electrons from the filament to the region of maximum scattering: T ∼ 10−8 sec.

    Dittmer was unable to detect these oscillations. However, before his paper was published, Penning [5] published observations of high-frequency radiations from a controlled arc and established a definite correlation between the observed high scattering and the radiations.

    In addition to the puzzling high scattering, there are two other anomalies of electrical discharges in gases that are connected with plasma oscillations. One of these is angular scattering of the primary beam a short distance from the cathode in an arc discharge. The most consistent work on this has been done by Emeleus and his students at Queen’s University, Belfast, Northern Ireland [6]. They have shown repeatedly the correlation between this scattering and the existence of the internal oscillations in discharges.

    The other phenomenon is sometimes called the Langmuir paradox. It was found that a Maxwellian distribution of electron energies exists to within a very small distance of an insulated wall in a gas discharge. This seemed quite surprising since such a wall acquires a negative charge which collects only high-energy electrons from the gas, thus distorting the Maxwellian distribution of the remaining electrons near the wall. At large distance from the wall the Maxwellian distribution would be reestablished because of energy-transfer collisions between electrons and other charged and neutral particles in the discharge. However, all the known mean free paths for energy transfer were orders of magnitude too large to account for the extremely short distance required for the reestablishment of the equilibrium distribution.

    In 1929, Tonks and Langmuir [1] presented their now famous theory of plasma oscillations which provided a single mechanism to explain each of these anomalies. Their theory was based upon a completely uniform macroscopically neutral, zero-temperature model of the plasma. A small displacement of a slab of electrons from their equilibrium position gave rise through coulomb interaction to a restoring force which in first order was linear. Simple harmonic motion resulted with a frequency which has become known as the characteristic plasma-electron frequency, or the Langmuir frequency.

    There has been some objection to this latter name since the characteristic-frequency parameter appeared much earlier in the classical work of Lorentz [7]. It should be pointed out, however, that the waves treated by Lorentz were transverse electromagnetic waves, while the oscillations deduced by Tonks and Langmuir were longitudinal waves of entirely different character. In order to emphasize the difference, waves of essentially the Langmuir type have been called electric sound waves, although, as we shall see, this is not an entirely adequate term.

    The theory and experiment on plasma oscillations have been significantly extended since the early pioneer work. The restriction to zero temperature was removed by the theoretical work of Landau [8] in 1946. The Langmuir frequency now appeared to be not a unique frequency of compressional vibration but the lower limit of a spectrum of such vibration frequencies which had minimum diffusion-type damping associated with them. This spectrum was governed by a dispersion law relating frequency to the wave number of the disturbance.

    In addition to the lower limit provided by the Langmuir frequency, Landau’s work showed the existence of an upper limit to the frequency spectrum obtained by setting the wavelength of the oscillation equal to the Debye length [9] (the distance required for the screening of a charge by a hot electrolyte). This is the wavelength at which the minimum diffusion damping rate becomes comparable with the oscillation frequency.

    In 1952, Gabor published [10] an article in which he calculated the rate of energy exchange between an individual electron and the whole spectrum of plasma oscillations in a discharge, under the assumption of equipartition of energy between the various modes of plasma oscillation. This yielded an energy-transfer rate orders of magnitude too small to account for the phenomena that plasma oscillations were originally invented to explain. Thus the refined theory seemed to explain much less than the original plasma-oscillation theory. Gabor was thus driven back to the original assumption of the existence of a strong, coherent, nearly single frequency oscillation within the discharge.

    In order to test this conclusion more severely, Gabor, Ash, and Dracott [11] in 1955 sent a special beam of electrons through a discharge and allowed the emerging beam to strike a fluorescent screen. An externally supplied time base deflection allowed the display of a resulting 100- to 120-Mc/sec wave which was randomly modulated at a much lower frequency. They found that surprisingly large amplitude oscillations (comparable to the static field strength in a sheath) existed in the positive ion sheath separating the plasma from an insulated wall, but not in the plasma itself.

    This required new extensions of the theory, namely, to include the effects of boundaries and beams on plasma oscillations. Both of these extensions were available in the theoretical work of Bohm and Gross and others [12]. One prediction of the Bohm and Gross theory was that an electron beam sent through a homogeneous plasma should excite plasma oscillations. An experimental test of this theory made by Merrill and Webb [13] already existed as of 1939, as pointed out and analyzed by Twiss [14] in 1951. However, the agreement between this experiment and theory was not quite certain because the electron beam in the Merrill and Webb experiment entered the plasma through an ion sheath. In order to eliminate this objection, Looney and Brown [15] arranged (1952–1954) for the sheathless injection of an electron beam into a plasma and found in this case no oscillations. This seemed to contradict the Bohm and Gross theory. What was needed, it appeared, was a systematic means of interpreting the dispersion laws such as that due to Bohm and Gross. Such a systematic procedure is given in this book by Sturrock [16]. Another approach emphasizing the importance of the coupling elements for detecting oscillations is provided by 0. Buneman [17].

    The Sturrock approach was utilized in 1958 by Drummond and Chang [18] in an attempt to resolve the paradox between the Looney and Brown experiment and the Bohm and Gross theory. This work showed that the Bohm and Gross dispersion law yields a prediction of amplification rather than oscillation. A similar conclusion had been reached by Pierce earlier by different methods [12]. This result seems to be contradicted by the failure of Looney and Brown to observe amplification [19]. An analysis of the parameters used by Looney and Brown shows that the gain of their system should be less than 20 db and hence probably less than coupling losses. In a similar arrangement, Demirkhanov [20] found strong interaction between a modulated beam and a plasma. As pointed out by Sturrock, the electron-beam diameter used by Looney and Brown was much smaller than the plasma wavelengths would have been and hence did not approximate the requirements of an infinitely broad beam postulated by Bohm and Gross. A theory of the dispersion of waves in and about a finite beam in a plasma has been given by Budker [21] in 1956.

    Looney and Brown went on in their experiment to establish an ion sheath over the port of entry of the electrons into the plasma. For such cases, they found standing (electric sound) waves of plasma oscillations and concluded that some special modulation mechanism was active in the sheath.

    Gordon [15], in a later thesis, pointed out that whenever a sheath was formed over the entry port, one was also formed over the beam-collection electrode, because the latter was electrically connected to the port electrode. Thus, secondary electrons originating on the collector electrode would be accelerated toward the entry electrode where they would be reflected back into the plasma.

    Furthermore, the multiple reflections which such electrons could experience might be phase selective. Therefore, secondary electrons, produced at such a time that in their transit through the oscillating plasma they would give up some of their kinetic energy of the oscillations, would not be quite able to reach the negative port electrode, whereas the secondary electrons produced during the opposite half cycle would acquire energy from the oscillations and thus be able to reach and be collected by the port electrode. By operating the port and collector electrodes at different potentials, Gordon was able to verify some of the detailed predictions of his theory. Since he was never able to obtain the oscillations without some kind of reflected beam, however, a number of possible mechanisms might have been causing the oscillations: bunching by sheath oscillations, selective reflection, and the presence of multiple beams.

    In order to separate out the various possible causes, Kofoid [22] arranged to have two independently generated electron beams interpenetrate in a plasma. He was able to eliminate some of the possible extraneous causes of oscillations and yet demonstrate coherent standing waves.

    We have thus seen that the first problem in plasma physics was that of the interaction of individual electrons with the whole body of electrons in an ionized gas. The body (i.e., plasma) property is represented by plasma oscillations, as has been emphasized by Bohm and Pines [23]. This is still a central problem in the field, as is witnessed by the existence of runaway electrons in thermonuclear machines [24]. If the highspeed electrons which arise in these machines interact strongly with the plasma, they will produce high-frequency instabilities—amplification or oscillations.

    1.2. Coherent Radiation from Plasmas.If the first problem in plasma physics can be said to be the interaction of plasma oscillations with individual electrons, the second problem is the interaction of plasma oscillations with the electromagnetic radiation field. It is almost as old a problem as the first one. The existence of plasma oscillations was hypothesized to explain the existence of the anomalies connected with the behavior of electrons in electric arcs. The case for the existence of these oscillations was considerably strengthened when radio waves of about the right frequency were found to be emitted from arcs. However, the theory of the oscillations showed them to be longitudinal, or nonradiating. It is true that Tonks and Langmuir also deduced the existence of transverse waves in plasmas, but these propagation modes were uncoupled to the plasma oscillation modes. This made necessary the assumption that edge effects or inhomogeneities would couple the two types of waves. Such has now been derived by the theoretical work of Field [25], Denisov [26], and Zhelezniakov [27]. Field showed that, for a plasma wave incident obliquely on a plane discontinuity, the conduction and displacement currents which usually cancel no longer do so. This produces a magnetic field and hence a diverging Poynting vector. This is of importance in explaining the radio noise emitted by the sun [28] and stars as well as by radio stars [29]. It is also of importance in understanding the radio noise emitted by fluorescent lights and, much more importantly, from thermonuclear reactors [30]. Of course it would also be important in designing a plasma oscillator or noise power source [31].

    Another important means of coupling energy from plasma oscillations to the radiation field is provided by the establishment of a static magnetic field within the plasma [25–28, 32]. In this case, the conduction electrons can no longer move in the direction of the electric field strength, so again conduction and displacement currents do not cancel and a magnetic field is produced. There are certain resonant frequencies, approximately integral multiples of the electron cyclotron frequency, for which there is a phase resonance between the electric field strength and the conduction current density. For these frequencies the orbits of individual electrons may expand or contract, exchanging energy with the radiation field. At these frequencies the wavelength of a right-hand circularly polarized oscillation propagating along the static magnetic field becomes very short, and electrons thermally drifting along the magnetic field lines may travel faster than the phase velocity of the wave. This gives rise to a form of Cherenkov radiation [33] and its inverse process. Electrons moving toward the source of a right-hand circularly polarized wave see a Doppler-shifted frequency. Thus, power begins to be absorbed at a lower frequency of the source than the electron cyclotron frequency [32]. Mahaffey [94] has observed this shift.

    A general formalism for treating small oscillations in heterogeneous plasmas including quasi-static electric and magnetic fields has been given by Drummond, Gerwin, and Springer [27]. They have derived the necessary and sufficient conditions for the total power transfer between a plasma and the radiation field to be identically zero. Application of this to a special case seems to disagree with a calculation of Tidman’s [27]. The difference may be due to a transient energy storage. Gerwin, Springer, and Drummond have derived the conditions for maximum steady-state power transfer [27].

    Sturrock [27] has recently shown that nonlinearities should give rise to second-harmonic radiation from plasma oscillations of homogeneous plasmas and has applied his theory to explain the observations of type 2 solar bursts (the small amount of fundamental signal observed being presumed due to radiation coupling by the small magnetic field).

    Given a coupling mechanism, the question still remains as to how, in certain situations, the requisite beams or turbulences, etc., arise and propagate. This is the important subject of magnetohydrodynamics (MHD) treated in the second part of the present volume and outlined in the next subsection of this chapter.

    1.3. Magnetohydrodynamics.Magnetohydrodynamics (MHD), or, better, magnetofluidynamics (MFD), as a subject of astrophysical interest began at about the time that radio-frequency plasma physics started. MFD dealt with problems of solar hydrodynamics [34] and the origin of the earth’s [35] and the sun’s [36] magnetic fields. Jeffreys [37] in 1926 was able to conclude from seismic observations (which showed that transverse waves are never transmitted through the earth’s core while longitudinal waves are transmitted with very little damping, etc.) that there seems to be no reason to deny that the earth’s metallic core is truly fluid. It was proved by Angenheister and Bartels [38] in 1928 that most of the magnetic field of the earth must arise from electric currents flowing within the core of the earth rather than around the earth, † This established one important case of an incompressible conducting fluid moving in a magnetic field (the conditions usually assumed for MHD analysis). Thus was born the MHD problem. It has been extensively studied since, with much interesting work remaining to be done. The present status of the astrophysical and geophysical aspects of MHD have been recently reviewed in book form [39] and thus are not dealt with further in the present volume, except to point out that the MHD motions may be responsible for the conditions favorable for generation and radiation of radio oscillations by the sun, as mentioned in the preceding subsection.

    It is interesting to note the early interaction between pure (impractical) astrophysical MHD and the very practical (mundane) problems such as mercury jet commutation [40] by Hartmann, one of the important early workers in the field [41]. Another practical application of MFD, magnetically self-focusing streams, was investigated in relation to the breakdown of high-voltage tubes by Bennett [42] in 1934. Magnetic self-focusing of streams has since been renamed the pinch effect by Tonks [43] in 1939. Today this effect is the basis of operation of a considerable number of the fusion research machines [44]. However, there does not seem to be any didactic review on the Bennett pinch or on the stability of relativistic self-focusing streams. Thus, two chapters (also containing new results) on these subjects are included in the present volume.

    The one area of MFD which has developed most rapidly in the last few years, and which has not yet received review, pertains to magnetoaerodynamics research. Therefore, an important chapter by Yoler has been included in the present volume.

    1.4. Quantum Plasma Physics.Just as the first problem of classical plasma physics dealt with the exchange of energy between an individual electron and a collective motion of many particles, so also did the first problem of quantum plasma physics. In 1934, Bloch [45] extended his famous work of the preceding year [46] on the stopping power of charged particles to include the quantization of the excited states of the Fermi gas. He treated these not as states with holes and excited particles but as states of soundlike oscillation of the gas. He showed that agreement with results based on a superposition of states determined by the hole theory was obtained if the equations of motion of the sound waves were properly chosen and if the zero-point amplitudes of the waves had values which would be expected for a sound field obeying Bose statistics.

    In 1936 Rudberg [47] observed that electrons of several hundred volts reflected from solid samples of the noble metals preferentially lost amounts of energy characteristic (eigenlosses) of the target metal. Rudberg and Slater correctly [48] interpreted each of these eigenlosses not in terms of excitation of a quantum in the Bloch sound field (representing a superposition of many individual particle excitations) but rather as a single excitation of an individual particle transition from the filled d band to the top of the s-p band. Five years elapsed before the first electron-beam-transmission measurement through aluminum and beryllium foils was reported by Ruthemann [49]. His measurements showed eigenlosses that could not be interpreted as single electron transitions, but the relation of his observed eigenlosses to the quanta of Bloch’s sound field in the Fermi gas was to lie unsuspected for more than a decade.

    In the meantime, other attempts to explain the magnitude of the eigenlosses failed [50]. In 1948 Lang [51], on the basis of experiments on 7.6-kev electrons penetrating aluminum, estimated the mean free path for the scattering process and showed that the samples of metal which had been used for determining the value of the anomalous eigenlosses had been thin enough to rule out multiple scattering.

    Also in 1948 Ruthemann [52], and in the following year Mollenstedt [53], extended his research to other metals where much less sharply defined eigenlosses were found. In 1947 Bogolyubov [54] obtained, in regard to a problem of superfluity, a dispersion law for plasma oscillations by using a theory of soundlike excitations and the method of second quantization.

    In 1949 Bohm [55] produced a theory of superconductivity based upon his anticipation that plasma oscillations in metals would be very difficult to disturb by external perturbing forces, a fact later proved by Tomonaga [56].

    In 1950 Tomonaga [56] extended Bloch’s work by proving, for a one-dimensional system of interacting fermions, that the excited states of Bloch sound waves obey Bose statistics. "This fact provides us with a new possibility of treating an assembly of Fermi particles in terms of the equivalent assembly of Bose particles, namely, the assembly of sound quanta. The field equation for the sound wave is found to be linear irrespective of the absence or presence of mutual interaction between particles [unlike the equation for the wave function ψ in the usual method] so that this method is a very useful means of dealing with many Fermion problems. It is also applicable to the case where the interparticle force is not weak. In the case of force of too short a range this method fails [56]. The possibility of this new method was found independently by Bohm [and Pines] [57] who discuss a very interesting phenomenon of plasmalike oscillations in a degenerate electron gas from a very similar point of view [56]. They utilize the linearity of the field equations to study such a pure correlation phenomenon as plasma oscillations of an electron gas" [56] and are the first to point out the connection of this theory with the eigenloss experiments of Ruthemann [52] and Lang [51].

    In 1951 Bohm and Staver [58] and in 1952 Staver [59] used the collective approach to describe the ionic vibrations (plasma ion oscillations) and the coupling between these and the plasma electron oscillations.

    Also in 1952 Klimontovich and Silin [60] generalized Tomonaga’s quantum plasma theory. A review of this and their later work which has not been widely enough known among scientists not reading Russian is contained in Chap. 2 of the present volume.

    In 1953 Wolff [60] showed that because of the presence of the lattice an effective mass correction should be applied to the electrons that undergo plasma oscillations in metals. He also showed that "in metals with occupied d bands, such as Cu, Ni, and Ag, there is a strong coupling between the plasma wave and the d electrons which gives rise to frequency broadening. This explains why the plasma lines observed by Ruthemann [52] and Lang [51] are so much wider in these elements than in Be or A1 [49]" [60].

    In 1953 Pines applied the theory of quantum plasma physics to the calculation of the binding energy, specific heat, ferromagnetism, and stopping power of metals [61].

    In 1954 Marton and Leder [62], Watanabe [63], and Kleinn [64] continued the extension of transmission measurements to a wide range of solids, finding generally complex eigenloss spectra. It has been suggested by Pines [65], Gabor [66], and Ferrell [67] that at least one line in the energy loss spectrum of each metal may be due to collective excitation [68].

    From 1954 on, experiments have been conducted to distinguish between the collective and the individual particle excitations produced in metals by fast electrons. These involve measurements of the eigenloss energy and intensity as a function of the angle of scattering of the incident electron [69], the state of division of the target [68, 70], the surface impurity content [71], and the temperature [72]. The results for many eigenlosses have agreed with predictions based upon the excitation of plasma oscillation quanta (plasmons) [65–67, 73].

    Just as for classical, so with quantum plasma physics the second problem is the interaction of plasma oscillations with the electromagnetic radiation field. The theory has been developed by Ferrell [74], and the extremely attractive experiments that he has proposed are under way [74].

    In the following section of this chapter, the basic statistical mechanics upon which classical plasma theory is based will be reviewed.

    2. Basic Statistical Mechanics of Plasmas

    The purpose of this section is to review the basic principles of the classical statistical mechanics that apply to plasmas. The first subsection presents a derivation of the Boltzmann equation from first principles. This equation, which is sometimes thought to be beyond derivation, involves a number of assumptions which are not always valid for plasmas. These are pointed out in the first subsection, and the improved BFP (Boltzmann-Fokker-Planck) equation is considered in the second subsection.

    2.1. Derivation of Boltzmann Equation.The general method for treating an N-body problem is statistical mechanics. The fundamental result of classical statistical mechanics is Liouville’s theorem [75], which can be put in the form

    where fN is the ensemble density function in the 6N is the complete Hamiltonian for the system. The symbol fN is interpreted as the probability density of finding a system in a state specified by the positions and momenta of all N of the particles. The brackets used here are the usual Poisson brackets:

    where the qk and pk are the coordinates and conj ugate momenta of the system and u and v are any functions of qk, pk, and time t. To find the probability density f1 for a single particle being in a given region of its phase space, one integrates over the coordinates and momenta of all the other N 1 particles [76].

    A typical term in the Poisson brackets of (2.1) is

    If this is integrated over qj, and pjmust be evaluated at the extremities of the coordinates which may be taken outside the plasma, where fN where fN will be zero. Thus the only terms in the Poisson bracket that remain after integration are the terms that contain derivatives with respect to the coordinates or momenta of the first particle. Thus

    must be specified.

    for 1 ≤ i ≤ 3. The quantity h is the interaction potential between the first and kth particles (only two body potentials allowed). Thus (2.5) becomes

    where the last integration has been performed over the coordinates and momenta except those of the first and kth particles, and the two-particle distribution function is defined as

    where the coordinates and momenta of only two particles are omitted from the integration.

    is further assumed to depend only upon the coordinates and possibly momenta of the first and kth particles, and not explicitly upon k, then the sum over k in by N 1, which is then called ϕ. Thus (2.7) becomes

    Thus the integration of (2.1) has given

    The term on the right represents two-body interactions.

    In this same fashion the Born-Green hierarchy of differential equations for the n-particle distributions can be obtained, the equation for each fn containing the next higher distribution in the term on the right. One can obtain the usual collision term of Boltzmann’s equation by neglecting three-particle interactions in the equation for f2 and integrating over the duration of one collision. Then

    where f1 and h are assumed nearly constant during a collision and where Boltzmann’s distribution function is

    Under the assumption neglecting three-body interactions, the equation for f2 becomes

    Thus (2.11) becomes

    But one should now investigate whether or not it is valid to assume that f1 and h remain nearly constant during the period required for a collision. This time is given roughly by

    sec.

    However, for an individual electron-electron or electron-ion encounter the period is always comparable with the period of plasma oscillations. This can be shown as follows: The diameter of interaction for a charge is that given by the Debye-Huckel theory of electrolytes.

    Thus

    Furthermore, the whole Born-Green hierarchy can be expected to be important since one may anticipate with Langmuir that the many-body interaction is the crux of the organized behavior of electrons in a plasma. One is thus left with two alternatives: to return to basic statistical mechanics, working directly with Liouville’s equation or some other approximation of this,† or to adopt the self-consistent-field point of view. From this latter point of view, the electrons do not interact directly with one another. Rather, they produce an electromagnetic field which is expressed as a function of position and time. The electromagnetic field then acts on each electron through its position and momentum and hence enters into h on the exact left-hand side of Eq. (2.14). The further requirement is then made that the resulting distribution of electrons be just that which produced the electromagnetic field [78]. In most theories, it is this last step that yields the dispersion law. In the self-consistent-field approach the electromagnetic field is given by

    The distribution function f is given by (2.14), where h is determined by the electromagnetic field interaction using the field satisfying (2.18). The term on the right-hand side of (2.14) now contains only electron-atom interactions and close-encounter electron-electron and electron-ion collisions. In more familiar form, (2.14) can be written as

    A word of warning might be given about the use of this equation. It should not be used to compute the reaction of a plasma to an applied electromagnetic field represented by h in (2.14) or by E and H in (2.19). It is not valid in this case because the total electromagnetic field which contains the reaction itself must be used in order for this equation to be correct. Furthermore, Eq. (2.18) does not give the total electromagnetic field; it gives only the expectation of the current density was used in it. Thus Eqs. (2.18) and (2.19) do not form a closed set. The essence of the self-consistent-field approximation is the assumption that they do.

    Equations (2.18) and (2.19) form a nonlinear set, for the coefficients E and H depend in turn upon moments of f. Frequently these equations are linearized and solved for small perturbations away from an equilibrium distribution. Attempts have been made to predict the effects of nonlinearities arising from strong coherent interactions among plasma electrons. This has been done both for direct solutions of the Boltzmann equation [79] [Eq. (2.19)] and for solutions of the magneto-hydrodynamic equations [39, 80] which are derived from the Boltzmann equation. However, an important point is usually overlooked in these endeavors. This is that a linearization has already been imposed in deriving the Boltzmann equation. When we used Boltzmann’s assumption of molecular chaos we evidently assumed that correlation effects could be neglected. Yet for long-range coulomb forces we would expect these correlations to be important for all but very small deviations from equilibrium. In writing the two-particle distribution function fof two one-particle distribution functions we not only omit nonlinear terms but linear terms as well. Thus the whole character of the distributions will be altered in the time required for the correlations to become effective. The correlations we speak of will give rise to effects such as dynamical friction. In order to include them we give up trying to follow rapid changes in the one-particle distribution function, and, following Boltzmann, Born and Green, Grad, Chandrasekhar, and others, we average f1 over the time interval required for a rapid fluctuation of force or a collision.

    2.2. The BFP Equation.We can represent the force acting on an electron as the sum of a coherent long-range force due to the organized behavior of the plasma and a stochastic force due to statistical fluctuations in the distribution of neighboring electrons. The average period of the fluctuating force can be estimated to be of the order of the time τ between electrons.

    while the plasma period is

    Thus the statistical fluctuations can be expected to be much more rapid than the frequencies of coherent oscillation for plasmas of smaller density than this or greater temperature. For these plasmas it is possible to define a time interval ΔT and much greater than the average period of stochastic fluctuations τ (see Fig. 1, page 22).

    Thus the rapidly fluctuating forces may be treated statistically as done by Chandrasekhar for Brownian motion [81]. His result is

    where E(r,t) and H(r,t) are, respectively, the self-consistent electric and magnetic fields, γ(f;v) is the expectation value of the rapidly fluctuating statistical acceleration, and q(f;v) is one-half the expectation value of τ times the square of this acceleration. They are evaluated by Gasioro-wicz, Neuman, and Riddell [85] and by Rostoker [85]. Rostoker [85] has recently pointed out that in some systems, such as a plasma penetrated by an electron beam, the self-consistent Boltzmann and Maxwell equations amplify any inhomogeneous fluctuations introduced. Thus, for such systems, the FP terms γ and q are greatly amplified, and dynamical friction and diffusion-in-velocity are greatly enhanced.

    3. Theory of Plasma Oscillations

    3.1. Zero -temperature Finite Amplitude Theory.With Tonks and Langmuir [1] consider a uniform, zero-temperature plasma. Let there be a small x-axis displacement ξ(x,t) of the electrons originally at x, but let the distribution of the massive + charges remain uniform. As pointed out by Dawson [82], if the displacements are such that the ordering of the electrons remains constant, i.e., such that if electron A at x1 starts out to the left of electron B at x2, then after the displacements ξ(x1,t) and ξ(x2,t) A is still to the left of B, then the total charge density which has crossed the plane x + ξ(x,t) is neξ(x,t), where n is the initially constant density of electrons. Thus the electric field at this plane is given by Gauss’s theorem.

    This field produces an acceleration of the displaced electrons [82]:

    Curves of the electric field produced by these oscillations are given in Dawson’s article [which from Eq. (3.1) gives

    Thus the electrons in a homogeneous plasma are seen to have a natural resonant frequency called the plasma frequency or Langmuir frequency. If ion motion were included, the frequency, as shown by Allis [92], would be somewhat higher because the electron mass m in the equation for ωp [above Eq. (2.17)] would have to be replaced by the reduced mass

    where M is the ion mass.

    3.2. Electromagnetic Waves.How are plasma oscillations related to electromagnetic waves? Plasma oscillations exist in the presence of charge bunching, whereas electromagnetic waves exist in the presence of currents: displacement currents in free space and total currents elsewhere.

    Consider Maxwell’s expression of Faraday’s induction law:

    together with Maxwell’s equation

    Taking the curl of the former and using the latter in it, one obtains

    and ξ compared with 1],

    Combining these into Eq. (3.5) gives

    Taking the divergence of this yields Thus for a general time dependence, Eq. (3.7) can be written

    This is the vector Klein-Gordon equation. It describes the transverse vibrations of a string imbedded in an elastic medium [83] such as a jelly plasma.

    Eq. (3.7) becomes

    where

    either an electromagnetic or an electric sound oscillation can occur (both have infinite wave length).

    Lorentz had earlier given the result [84]

    for electromagnetic waves in a medium where individual electrons have a natural resonant frequency ω0.

    3.3. Plasma Ion Oscillations.We omitted ion motion in calculating electron plasma oscillation. One feels justified in this because the ions are so much more massive than the electrons that they could not have responded much at the high frequency of the electron oscillations. One wonders, however, if there might not have been similar oscillations at a low frequency. For a low frequency, the electrons should be able to respond so completely that they will come into equilibrium with the potential in a small fraction of an oscillation period. On the other hand, the ions will respond to the field by acceleration as the electrons did for the higher-frequency oscillations. Following Tonks and Langmuir [1], we have (considering now only very small ion

    Enjoying the preview?
    Page 1 of 1