0% found this document useful (0 votes)
4K views292 pages

(Dwight R Nicholson) Introduction To Plasma Theory PDF

Uploaded by

丁安磊
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF or read online on Scribd
0% found this document useful (0 votes)
4K views292 pages

(Dwight R Nicholson) Introduction To Plasma Theory PDF

Uploaded by

丁安磊
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF or read online on Scribd
You are on page 1/ 292
Introduction to Plasma Theory Dwight AN icholson i University of lowa John Wiley & Sons New York ¢ Chichester ¢ Brisbane * Toronto * Singapore Preface The purpose of this book is to teach the basic theoretical principles of plasma physics. It is not intended to be an encyclopedia of results and techniques. Nor is it intended to be used primarily as a reference book. It is intended to develop the basic techniques of plasma physics from the beginning, namely, from Maxwell's equations and Newton’s law of motion. Absolutely no previous knowledge of plasma physics is assumed. Although the book is primarily intended for a one year course at the first or second year graduate level, it can also be used for a one or two semester Course at the junior or senior undergraduate level. Such an under- graduate course would make use of that half of the book which assumes a knowl- edge only of undergraduate electricity and magnetism, The other half of the book, suitable for the graduate level, requires familiarity with complex variables, Fourier transformation, and the Dirac delta function. The book is organized in a logical fashion. Although this is not the standard organization of an introductory course in plasma physics, ] have found that students at the graduate level respond well to this organization. After the intro- ductory material of Chapters 1 and 2 (single particle motion), the exact theories of Chapters 3 to 5 (Klimontovich and Liouville equations), which are equivalent to Maxwell’s equations plus Newton’s law of motion, are replaced via approxima- tions by the Vlasov equation of Chapter 6. Further approximations lead to the fluid theory (Chapter 7) and magnetohydrodynamic theory (Chapter 8). The book concludes with two chapters on discrete particle effects (Chapter 9) and weak turbulence theory (Chapter 10). Chapter 6, and Chapters 7 and 8, are meant to be self-contained, so that the book can easily be used by instructors who wish the standard organization. Thus, the introductory material of Chapters | and 2 can be immediately followed by Chapters 7 and 8. This would be enough material for a vii vill Preface one semester undergraduate course, while the first half of a two semester graduate course could continue with Chapter 6 on Vlasov theory, followed in the second semester by Chapters 3 to 5 on kinetic theory and then by Chapters 9 and 10. It is a pleasure to acknowledge the help of many individuals in writing this book. My views on plasma physics have been shaped over the years by dozens of plasma physicists, especially Allan N. Kaufman and Martin V. Goldman. The students in graduate plasma physics courses at the University of Colorado and the University of Iowa have contributed many useful suggestions (Sun Guo-Zheng deserves special mention). The manuscript was professionally typed and edited by Alice Conwell Shank, Gail Maxwell, Susan D. Imhoff, and Janet R. Kephart. The figures were skillfully drafted by John R. Birkbeck, Jr. and Jeana K. Wonderlich. The preparation of this book was supported by the University of Colorado, the University of Iowa, the United States Department of Energy, the United States National Aeronautics and Space Administration, and the United States National Science Foundation. Owight R. Nicholson Contents CHAPTER 1. Introduction Introduction Debye Shielding Plasma Parameter Plasma Frequency Other Parameters Collisions References Problems 2. Single Particle Motion 21 2.2 2.3 24 2.5 2.6 2.7 28 2.9 Introduction E X B Drifts Grad-B Drift Curvature Drifts Polarization Drift Magnetic Moment Adiabatic Invariants Ponderomotive Force Diffusion References Problems BAocrHe x Contents 3. Plasma Kinetic Theory I: Kilmontovich Equation 3.1 3.2 33 Introduction Klimontovich Equation Plasma Kinctic Equation References Problem 4. Plasma Kinetic Theory II: Liouville Equation and BBGKY Hierarchy 4.1 4.2 43 Introduction Liouville Equation BBGKY Hierarchy References Probiems 5. Plasma KInetic Theory III: Lenard-Balescu Equation SL 5.2 Bogoliubov's Hypothesis Lenard-Balcscu Equation References Problems 6. Vlasov Equation 6.1 Introduction Equilibrium Solutions Electrostatic Waves Landau Contour Landau Damping Wave Energy Physics of Landau Damping Nonlinear Stage of Landau Damping Stability: Nyquist Method, Penrose Criterion General Theory of Linear Vlasov Waves Linear Vlasov Waves in Unmagnetized Plasma Linear Vlasov Waves in Magnetized Plasma BGK Modes Case-Van Kampen Modes References Problems 7. 10. Contents Fluid Equations 7.1 Introduction 7.2. Derivation of the Fluid Equations from the Vlasov Equation 7.3 Langmuir Waves 7.4 Dielectric Function 7.5 Ton Plasma Waves 7.6 Electromagnetic Waves 7.7 Upper Hybrid Waves 7.8 Electrostatic lon Waves 7.9 Electromagnetic Waves in Magnetized Plasmas 7.10 Etectromagnetic Waves Along By 7.11 Alfvén Waves 7.12 Fast Magnetosonic Wave 7.13 Two-Stream Instability 7.14 Drift Waves 715 Nonlinear Ilon-Acoustic Waves—Korteweg-DeVries Equation 7.16 Nonlinear Langmuir Waves—Zakharov Equations 7.17 Parametric Instabilities References Problems 8. Magnetohydrodynamics 8.1 Introduction 8.2. MHD Equilibrium 8.3. MHD Stability 8.4 Microscopic Picture of MHD Equilibrium References Problems Discrete Particle Effects 9.1 Introduction 9.2 Debye Shielding 9.3 Fluctuations in Equilibrium References Weak Turbulence Theory 10.1 Introduction 10.2 Quasilinear Theory 10.3 Induced Scattering xi 127 127 129 132 136 138 141 144 146 150 156 161 164 166 169 171 177 181 184 185 189 189 194 200 206 208 210 21 2it 21 219 224 226 226 226 234 xii Contents 10.4. Wave-Wave Interactions References Problem APPENDIX A. Derivation of the Lenard-Balescu Equation References B. Langevin Equation, Fluctuation-Disslpation Theorem, Markov Processes, and Fokker-Planck Equation B.1 Langevin Equation and Fluctuation~Dissipation Theorem B.2_- Markov Processes and Fokker-Planck Equation References C. Pedestrian’s Guide to Complex Variables References D. Vector and Tensor Identities Reference INDEX 241 253 254 257 266 267 267 272 278 279 284 285 285 286 CHAPTER Introduction 1.1. INTRODUCTION A plasma is a gas of charged particles, in which the potential energy of a typical Particle due to its nearest neighbor is much smaller than its kinetic energy. The plasma state is the fourth state of matter: heating a solid makes a liquid, heating a liquid makes a gas, heating a gas makes a plasma. (Compare the ancient Greeks’ earth, water, air, and fire.) The word plasma comes from the Greek plasma, meaning “something formed or molded.” It was introduced to describe ionized gases by Tonks and Langmuir [1]. More than 99% of the known universe is in the plasma state. (Note that our definition excludes certain configurations such as the electron gas in a metal and so-called ‘‘strongly coupled” plasmas which are found, for example, near the surface of the sun. These need to be treated by techniques other than those found in this book.) In this book, we shall always consider plasma having roughly equal numbers of singly charged ions (+e) and electrons (~e), each with average density no (particles per cubic centimeter). In nature many plasmas have more than two species of charged particles, and many ions have more than one electron missing. It is easy to generalize the results of this book to such plasmas. EXERCISE Name a well-known proposed source of energy that involves plasma with more than one species of ion. 12 DEBYE SHIELDING Ina plasma we have many charged particles flying around at high speeds. Consider a special test particle of charge gz > 0 and infinite mass, located at the origin of a 2 Introduction three-dimensional coordinate system containing an infinite, uniform plasma. The test charge repels all other ions, and attracts al! electrons. Thus, around our test charge the electron density n, increases and the ion density decreases. The test ion gathers a shielding cloud that tends to cancel its own charge (Fig. 1.1). Consider Poisson's equation relating the electric potential to the charge densi- ty p due to electrons, ions, and test charge, Vig = —4mp = 4re(n, — n) — 41g; H(t) ab) where 8(r) = 6(x)6(y)8(z) is the product of three Dirac delta functions. After the introduction of the test charge, we wait for a long enough time that the electrons with temperature T, have come to thermai equilibrium with themselves, and the ions with temperature T; have come'to thermal equilibrium with themselves, but not so long that the electrons and ions have come to thermal equilibrium with each other at the same temperature (see Section 1.6). Then equilibrium statistical me- chanics predicts that 1, = My exp (4). mh = tt exw (52 (1.2) where cach density becomes np at large distances from the test charge where the Potential vanishes. Boltzmann’s constant is absorbed into the temperatures T, and T;, which have units of energy and are measured in units of electron-volts (eV). Assuming that eg/T, << 1 and eg/T; << 1, we expand the exponents in (1.2) and write (1.1) away from r = 0 as 1 d(,dp 1 1 Vig = =r (ASE) = anne? (= + ze (13) Fig. 1.1 A test charge in a plasma attracts particles of opposite sign and repels particles of like sign, thus forming a shielding cloud that tends to cancel its charge. Plasma Parameter 3 If we define the electron and ion Debye lengths y" (L.4) and the total Debye length =A + A? (1.5) Eg. (1.3) then becomes eg do 2 = Ze &) = Apte- (1.6) Trying a solution of the form » = @/r, we find PS 7 Se = 0% a7) The solution that falls off properly at large distances is @ & exp (-r/hp). From elementary electricity and magnetism we know that the solution to (1.1) at loca- tions very close to r = 0 is y.= q,/r; thus, the desired solution to (1.1) at all distances is (1.8) The potential due to a test charge in a plasma falls off much faster than in vacuum. This phenomenon is known as Debye shielding, and is our first example of plasma collective behavior. For distances r >> the Debye length Ap, the shielding cloud effectively cancels the test charge g,. Numerically, the Debye length of species 5 with temperature T, is roughly d, * 740[7,(eV)/n(cm™)]' in units of cm. EXERCISE Prove that the net charge in the shielding cloud exactly cancels the test charge 97. It is not necessary that g7 be a special particle. In fact, each particle in a plasma tries to gather its own shielding cloud. However, since the particles are moving, they are not completely successful. In an equal temperature plasma (ZT, = T;), a typical slowly moving ion has the full electron component of its shielding cloud and a part of the ion component, while a typical rapidly moving electron has a part of the electron component of its shielding cloud and almost none of the ion component. ~ 1.3 PLASMA PARAMETER Ina plasma where each species has density np, the distance between a particle and its nearest neighbor is roughly ny 3. The average potential energy ® of a particle due to its nearest neighbor is, in absolute value, e || ~ <— ~ yh? (1.9) 4 Introduction Our definition of a plasma requires that this potential energy be much less than the typical particle’s kinetic energy ee ry y mae’) = > T= > mys (1.10) where m, is the mass of species s,({ _) means an average over all particle velocities at a given point in space, and we have defined the thermal speed v, of species s by Ss v= 2)" (ty For electrons, ve ~ 4 X 107,12 (eV) in units of cm/s. Our definition of a plasma requires niet << T, (1.12) or an (Ts ng? (5 } >> 1 13) In € Raising each side of (1.13) to the 3/2 power, and recalling the definition (1.4) of the Debye length, we have (dropping factors of 47, etc.) Noh, >>} (1.14) where A, is called the plasma parameter of species s. (Note: Some authors call A,? the plasma parameter.) The plasma parameter is just the number of particles of “species s in a box each side of which has length the Debye length (a Debye cube). Equation (1.14) tells us that, by definition, a plasma is an ionized gas that has many particles in a Debye cube. Numerically, A, = 4 X 10°72%(eV)/no!(cm”). We will often substitute the total Debye length A pin (1.14), and define the result A = ny Xz’ to be the plasma parameter. EXERCISE Evaluate the electron thermal speed, electron Debye length, and electron plasma parameter for the following plasmas. (a) A tokamak or mirror machine with T, ~ 1 keV, my ~ 10" cm?. (b) The solar wind near the earth with T, = 10 eV, m) ~ 10 cm™. (c) The ionosphere at 300 km above the earth's surface with T, ~ 0.1 eV, : My = 108 cm”. (a) A laser fusion, electron beam fusion, or ion beam fusion plasma with T, © 1 keV, ty ~ 10 cm. (¢) The sun’s center with T, © 1 keV, my) = 10% om”. It is fairly easy to see why many ionized gases found in nature are indeed plasmas. If the potential energy of a particle due to its nearest neighbor were greater than its kinetic energy, then there would be a strong tendency for electrons and ions to bind together into atoms, thus destroying the plasma. The need to keep ions and electrons from forming bound states means that most plasmas have temperatures in excess of one electron-volt. Plasma Frequency 5 EXERCISE The temperature of intergalactic plasma is currently unknown, but it could well be much /ower than 1 eV. How could the plasma maintain itself at such a low temperature? (Hint: no ~ 10° cm”). Of course, it is possible to find situations where a plasma exists jointly with another state. For example, in the lower ionosphere there are regions where 99% of the atoms are neutral and only 1% are ionized. In this partially ionized plasma, the ionized component can be a legitimate plasma according to (1.14), where A, should be calculated using only the parameters of the ionized component. Typical- ly, there will be a continuous exchange of particles between the unionized gas and the ionized plasma, through thé processes of atomic recombination and ioniza- tion. We can now evaluate the validity of the assumption made before (1.3), that ep/T, << 1. This assumption is most severe for the nearest neighbor to the test charge (which we now take to have charge g; = +e). Using the unshielded form of the potential, we require 7, ()- bee) <<: (1.15) or noe <> w,,so w,? ~ w,?. We will see in a later chapter that the general response of an unmagnetized plasma to a perturbation in the electron density is a set of oscillations with frequencies very close to the electron plasma frequency w,. The relation among the Debye length A,, the plasma frequency w,, and the thermal speed v,, for the species s, is d, = v,/o, (1.23) EXERCISE Demonstrate (1.23). 1.5 OTHER PARAMETERS Many of the plasmas in nature and in the laboratory occur in the presence of magnetic:fields. Thus, it is important to consider the motion of an individual charged particle in a magnetic field. The Lorentz force equation for a particle of charge g, and mass m, moving in a constant magnetic field B = Bo? is me = Ee x Bye) (1.24) For initial conditions r(t = 0) = (xo,¥e,2o) and i(1 = 0) = (0, v,,v,) the solu- tion of (1.24) is x(t) = eeateea (t) =X + n (1 — cos 9,1) Lae Y(t) = yo + 9, 80 Ot 2(t) = 2 + ut (1.25) 8 Introduction where we have defined the gyrofrequency 4: Bo (1.26) EXERCISE Verify that (1.25) is the solution of (1.24) with the desired initial condi- tions. Numerically, Q, = — 2 X 10’ By (gauss, abbreviated G) in units of s', and 2, = 10* By (gauss) in units of s"' if the ions are protons. The nature of the motion (1.25) is a constant velocity in the 2-direction, and a circular gyration in the x-y plane with angular frequency || and center at the guiding center position r,, given by Tpe = (Xo + v4 /M,, Yo, Zo + v,t) (1.27) The radius of the circle in the x-y plane is the gyroradius v,/|Q,|. The mean gyroradius r, of species s is defined by setting v, equal to the thermal speed, so v/10, (1.28) EXERCISE nthe exercise below (1.14), calculate and order the frequencies w,, @;, |Q,], M5 also calculate the gyroradii r, and r; take T; = T, and use the following parameters. . (a) Protons, By = 10kG. ° (b) Protons, By = 10°5G. (c) Ot ions, By = 0.5 G. (d) Deuterons, By = 0 and B, = 10° G. {e) Protons, By = 100 G. At this point, let us briefly mention relativistic and quantum effects. For simplici- ty, we shall always treat nonrelativistic plasmas. In principle, there is no difficulty in generalizing any of the results of this course to include special relativistic effects; these are discussed at length in the book by Clemmow and Dougherty [2]. EXERCISE To what regime of electron temperature are we limited by the non- relativistic assumption? How about ion temperature if the ions are protons? ‘There are, of course, many plasmas in which special relativistic effects do be- come important. For example, cosmic rays may be thought of as a component of the interstellar and intergalactic plasma with relativistic temperature. We shall also neglect quantum mechanical effects. For most of the laboratory and astrophysical plasmas in which we might be interested, this is a good assump- tion. There are, of course, plasmas in which quantum effects are very important. An example would be solid state plasmas. As a rough criterion for the neglect of quantum effects, one might require that the typical de Broglie length #/m,v, be much less than the average distance between particles ny”. Collisions 9 EXERCISE What is the maximum density allowed by this criterion for electrons with temperature (a) 10 eV? (b) 1 keV? (c) 100 keV? In other applications, such as collisions (see next section), one might require the de Broglie length to be much smailer than the distance of closest approach of the colliding particles. In addition to these assumptions, we shall also neglect the magnetic field in many of the sections of this book. This neglect is made for simplicity, in order that the basic physical phenomena can be elucidated without the complications of a magnetic field. In practice, the magnetic field can usually be ignored when the typical frequency (inverse time scale) of a phenomenon is much larger than the gyrofrequencies of both plasma species. 1.6 COLLISIONS A typical charged particle in a plasma is at any instant interacting electrostatically (see Problem 1.5) with many other charged particles. If we did not know about Debye shielding, we might think that a typical particle is simultaneously having Coulomb collisions with all of the other particles in the plasma. However, the field of our typical particle is greatly reduced from its vacuum field at distances greater than a Debye length, so that the particle is really not colliding with particles at large distances. Thus, we may roughly think of each particle as undergoing A simultaneous Coulomb collisions. From our definition of a plasma, we know that the potential energy of interac- tion of each patticle with its nearest neighbor is small. Since the potential energy is a measure of the effect of a collision, this means that the strongest one of its A simultaneous collisions (the one with its nearest neighbor) is relatively weak. Thus, a typical charged particle in a plasma is simultaneously undergoing A weak colli- sions. We shall soon see that even though A is a large number for a plasma, the total effect of all the simultancous collisions is still weak. Of course, a weak effect can still be a very important effect. In the magnetic bottles like tokamaks and mirror machines currently being used to study controlled thermonuclear fusion plasmas, ion-ion collisions are one of the most important loss mechanisms. Mathematically, the importance of collisions is contained in an expression called the collision frequency, which is the inverse of the time it takes for a particle to suffer a collision. Exactly what is meant by a collision of a charged particle depends upon the definition, and we will consider two different definitions with different physical content. Our mathematical derivation of the collision frequency is an approximate one, intended to be simple but yet to yield the correct results within factors of two or so. A more rigorqus development can be found in the book by Spitzer [3]. (See Problem 1.6.) Consider the situation shown in Fig. 1.4. A particle of charge g, mass m is incident on another particle of charge gy and infinite mass with incident speed ug. 10 introduction Fig. 1.4 Parameters used in the discussion of the collision frequency in Section 1.6. If the incident particle were undeflected, it would have position x = vot along the upper dashed line in Fig. 1.4, being atx = 0 directly above the scattering charge gy at t = 0. The separation p of the two dashed lines is the impact parameter. If the scattering angle is small, the final parallel speed (parallel to the dashed lines) will be quite close to up. The perpendicular speed v, can be obtained by calculating the total perpendicular impulse : my, =f dt F(t) (1.29) where F, is the perpendicular force that the particle experiences in its orbit. Since the scattering angle v/v is small, we can to a good approximation use the unperturbed orbit x = vot to evaluate the right side of (1.29). This approximation is a very useful one in plasma physics. In Fig. 1.4, Newton's second law with the Coulomb force law is ¢ = $40 mt 7 ? (1.30) where # is a unit vector in the e-direction, Then 290 = Masin® _ He ss F=f sing = pan OF 7 pei 6 (1.31) where we have used p = r sin @ since the particle is assumed to be traveling along the upper dashed line. Equation (1.29) then reads = % eli de sin? 6(0) (1.32) The relation between @ and f is oa from , =p cos @ x= —reos6 = BEE = agp (1.33) so that =2 4 : a= SG : (1.34) EXERCISE Verify (1.34). Using (1.34) in (1.32), we find 2 v= tf ao sin o = Fete (1.38) Defining the quantity Collisions "1 2 Po = Gat (1.36) we have ee Po ee eee 1.37. a = 2 (1.37) which is strictly valid only when v, << vp,p >> po. In some books, the parame- ter po is called the Landau length, EXERCISE Show that if gqo > 0, then py is the distance of closest possible ap- proach for a particle of initia! speed up. Although (1.37) is not valid for large angle collisions, let us use it to get a rough idea of the impact parameter p which yields a large angle collision, we do this by setting v, equal to vg in (1,37) to obtain p = po. Thus, any impact parameter P & po will yield a targe angle collision, Suppose the incident particle is an elec- tron, and the (almost) stationary scatterer is an ion. (Although Fig. 1.4 shows a repulsive collision, our development is equally valid for attractive collisions.) The cross section for scattering through a large angle by one ion is mp,?. Consider an electron that enters a gas of ions. It will have a large angle collision after a time given roughly by setting (the total cross section of the ions in a tube of unit cross-sectional area, and length equal to the distance traveled) equal to (the unit area), or (time) X (velocity) X (number per unit volume) X (cross section) = 1. The inverse of this time gives us the collision frequency v, for /arge angle colli- sions; thus 2 Anneg?ge _ Amnce! (1.38) Py = My 2 = Mvp o! = ays mug Note that v, is proportional to the inverse third power of the particle speed. Recall that a typical charged particle in a plasma is simultaneously undergoing A collisions. Only a very few of these are of the large angle type that lead to (1.38), since a large angle collision involves a potential energy of interaction comparable to the kinetic energy of the incident particle and, by the definition of a plasma, the potential energy of a particle due to its nearest neighbor is sma]] compared to its kinetic energy. Thus, a particle undergoes many more small angle collisions than large angle collisions. It turns out that the cumulative effect of these small angle collisions is substantially larger than the effect of the large angle collisions, as we shall now show. Unlike the large angle collisions, the many small angle collisions can produce a large effect only after many of them occur. But these small angle collisions pro- duce velocity changes in random directions, some up, some down, some left, some right. We need to know how to measure the cumulative effect of many small random events. Consider a variable Ax that is the sum of many smal] random variables Ax,, €=1,2,...,N, Ax = Ax, + Ax, +t... + Ary (1.39) 12 Introduction Suppose (Ax) = 0 for each i and ((Ax)") is the same for each i, where ( ) indicates ensemble average [4]. Furthermore, suppose (Ax; Ax) = 0 if i # j, so that Ax; is uncorrelated with Ax, i # j. Then by (1.39) we have (Ax) = 0, and (dx) = (& a») ) : N = > (a) a = N (Ax, (1.40) Consider a typical particle moving in the z-direction through a gas of scattering centers. As it moves, it suffers many small angle collisions given by », which can be decomposed into random variables Av, and Av,. These latter have just the proper- ties of our random variable Ax, above. For one collision, with a given impact parameter p (Fig. 1.5), we have from (1.37) 2p 2 Uo Po (vy?) = (Av,)) + (dv,)) = Pp (4) Since Av, must have the same statistical properties as Av,, we must have 1 up? (av) = (wy?) = > “RE (142) Then by (1.40) we have, for the total x velocity Av,, N wpe? (Ave?) = N(Av,) = a (1.43) P Since we are considering a particle moving through a gas of scattering centers, it is more useful for our purposes to have the time derivative of (1.43), where on the Fig. 1.5. The incident particle is located at the origin and is traveling into the paper. It makes simultaneous small angle collisions with all of the scattering centers randomly distributed with impact parameters between p and p + dp. Collisions 13 right we shall have dN/dt = 2p dp novo as the number of scattering centers, with impact parameter between p and p + dp, which our incident particle encounters per unit time. The time derivative of (1.43) is then d q Fe (avey) = angus? FP (1.44) We have calculated (1.44) for only one set of impact parameters between p and p + dp. The same logic that led to (1.40) also allows us to sum (integrate) the right side of (1.44) over all impact parameters to obtain a total change in mean square velocity in the £-direction. Likewise, we can add the total ¥-direction and the total §-direction mean square velocities to obtain a total mean square perpendicular velocity (Av,")). With this final factor of two we have ox dp i Poin P What should we use for Piaax 20 Pin? Recall that our derivation of the scattering angle v, /vo in (1.37) uses the Coulomb force law. However, we know from Section 1.2 that the true force law is modified by Debye shielding and is essentially negligible at distances (impact parameters) much greater than a Debye length. Thus, it is consistent with the approximate nature of the present calculation to replace Pmay With A p- In the case of p,,;,, we use the fact that our scattering formula (1.37) is not valid for impact parameters p < |pol to replace p,,,, by |po|. Equation (1.45) is then d a (Ante) = 2anove’Po” (1.45) a Gp (A019?) = Qarngne'po'ln ( a ) (1.46) Since the logarithm is such a slowly varying function of its argument, it will suffice to make a very rough evaluation of A »/py. In the definition of py in (1.36) we take q = ~€,40 = +e, m = m,, and for this rough calculation replace vp by the ¢lec- tron thermal speed v, to obtain Ao Avmue _ mApia? teal liseli ze ceeleli aes where we have ignored the difference between Ap and A,. Dropping the smalt factor 2m compared to the large plasma parameter A, and using the definition (1.36) of po, we find that (1.46) becomes ~ Unnghp' = 2A (147) ‘ (Av 9?) = in (1.48) d at ( A reasonable definition for the scattering time due to small angle collisions is the time it takes (Av, '*')?) to equal uo” according to (1.48); the inverse of this time is the collision frequency v, due to small-angle collisions: _ 8rnoe" In A iy |. (149) Ye Note again the inverse cube dependence on the velocity vo. One important aspect of y, is that it is a factor 2 In A larger than the collision frequency v, for large 14 Introduction angle collisions given by (1,38). This is a substantial factor in a plasma (In A = 1. if A = 10°). Thus, the deflection of a charged particle in a plasma is predominant- ly due to the many random small angle collisions that it suffers, rather than the rare large angle collisions. Throughout one’s study of plasma physics, it is useful to identify each phenom- enon as a collective effect or as a single particle effect. The oscillation of the plasma slab in Section 1.4, characterized by the plasma frequency w,, is a collective effect involving many particles acting simultaneously to produce a large electric field. The collisional deflection of a particle, represented by the collision frequency v, in (1.49), is a single particle effect caused by many collisions with individual particles that do not act cooperatively. EXERCISE 1s the Debye shielding described in Section 1.2 a collective effect or a single particle effect? It is instructive to calculate the ratio of v, to w,, which is, taking a typical speed vp = v, in (1.49), vy Smet MA _ InA _ Ind @.. muvee, aNArA? 2A, By crudely dropping the factor In A/2m and replacing A, by A, we have the easily remembered but very approximate expression (1.50) (151) Thus, the collision frequency in a plasma is very much smaller than the plasma frequency. In this respect, single particle effects are less important than collective effects. A wave with frequency near w, wil! oscillate many times before being substantially damped because of collisions. EXERCISE What is the ratio of the collisional mean free path, for a typical electron, to the electron Debye length? The collision frequency y, that we calculated in (1.49) is the one appropriate to the collisions of electrons with ions, v.,, The collision frequency v,, of electrons with electrons could be calculated in the same way, by moving to the center-of- mass frame rather than taking the scattering center to have infinite mass. This procedure would only introduce factors of two or so, so that within such factors we have y,, * vj. Next, consider ion-ion collisions between ions having the same temperature as the electrons that have collision frequency y,.. Equation (1.49) yields, with m, replaced by m,and v, = (m./m,)!? v, instead of vo, v, © (m/m,)!2 Vee. Finally, consider ions scattered by electrons (or Mack trucks scattered by pedestrians). This calculation in the center-of-mass frame would introduce another factor of (m/m,)'”, so that vj, © (N/M) V0 Suppose an electron-proton plasma is prepared in such a way that the electrons and protons have arbitrary velocity distributions, and comparable but not equal temperatures. On the time scale v,5' © v.! © A w,', the electrons will therma- Problems 15 lize via electron-electron and electron-ion collisions and obtain a Maxwellian dis- tribution. On a time scale 43 times longer, the ions will thermalize and obtain a Maxwellian at the ion temperature via ion-ion collisions. Finally, on a time scale 43 times longer stifl, the electrons and ions will come to the same temperature via ion-electron collisions. This completes our brief introduction to the important basic concepts of plasma physics. In the next chapter, we shall consider the motion of single charged parti- cles in electric and magnetic fields. REFERENCES (1) L. Tonks and I. Langmuir, Phys. Rev., 33, 195 (1929), [2] P. C. Clemmow and J. P. Dougherty, Electrodynamics of Particles and Plas- mas, Addison-Wesley, Reading, Mass., 1969. (3) L. Spitzer, Jr., Physics of Fully Ionized Gases, 2nd ed., Wiley-Interscience, New York, 1962. [4] F. Reif, Fundamentals of Statistical and Thermal Physics, McGraw-Hill, New York, 1965. PROBLEMS 1.1 Debye Shielding In the discussion of Debye shielding in Section 1.2, suppose that the ions are infinitely massive and thus cannot respond to the introduction of the test charge. How does the answer change? 1.2 Potential Energy (Birdsall’s Problem) A sphere of plasma has equal uniform densities ) of electrons and infinitely massive ions. The electrons are moved to the surface of the sphere, which they cover uniformly. What is the potential energy in the system? Sketch the electric field and electric potential as a function of radius. If the electrons initially had temperature T,, and it is found that the potential energy is equal to the total initial electron kinetic energy, what is the radius of the sphere in terms of the electron Debye length? 1.3. Total Plasma Frequency In the discussion of the plasma frequency in Section 1.4, suppose the ions are not infinitely massive but have mass m,. Modify the discussion to show that the slabs oscillate with the total plasma frequency defined in (1.22). 1.4 -Plasma in a Gravitational Field Consider an electron-proton plasma with equal temperatures T = T, = T,, no magnetic field, and a gravitational acceleration g in the —2-direction. We desire the 16 Introduction densities n,(z) and nz), where z = 0 can be thought of as the surface of a planet. If the electrons and ions were neutral, their densities would be given by the Boltzmann law n,, & exp (—m.; g2/T). Then the scale height 7/m,, g would be quite different for electrons and ions. However, this would give rise to huge electric fields that would tend to move ions up and electrons down. Taking into account the electric field, use the Boltzmann law and the initial guess thatn,(z) ~ n{2z), to be checked at the end of the calculation, to find self-consistent electron and ion density distributions. 1.5 Electrostatic Interaction Show that in nonrelativistic plasma, the Coulomb force between two typical parti- cles is much more important than the magnetic field part of the Lorentz force. 1.6 Collisions Read Sections 5.1, 5.2, and 5.3 of Spitzer [3] and compare his treatment of colli- sions to our Section 1,6. Watch out for differences in notation, and explain all apparent differences of factors of two. CHAPTER 2 Single Particle Motion 2.1. INTRODUCTION A plasma consists of many charged particles moving in self-consistent electric and magnetic fields. The fields affect the particle orbits, and the particle orbits affect the fields. The generat solution of any problem in plasma physics can be quite complicated. In this chapter, we consider the motion of a single charged particle moving in prescribed fields. After studying this part of the problem in isolation, we can proceed in following chapters to include these particle orbits in the self- consistent determination of the fields. 2.2 E x BDRIFTS Consider a particle with v, = 0 gyrating in a magnetic field By in the 2-direction, with an electric field E, in the ~f-direction perpendicular to the magnetic field as in Fig, 2.1. (The symbol 4 always means a unit vector in the a-direction.) The electric field, Ey cannot accelerate the particle indefinitely, because the magnetic field will turn the particle. (The component of electric field E,, which we ignore here, can accelerate particles indefinitely. In a plasma, the resulting current usually acts to cancel the charge that caused the electric field in the first place. There are, however, important cases where this cancellation is hindered: for example, the earth’s aurora, and tokamak runaway electrons.) What does happen? When the charge q, is positive, the ion is accelerated on the way down. This gives it a larger local gyroradius at the bottom of its orbit than at the top; recall that the gyroradius isr, = v,/Q,. Thus, the motion will be a spirat in the x-y plane as shown in Fig. 2.2, where we have used the symmetry of the situation to draw the upward part of. each orbit. We see that the orbit does not connect to itself, but has jumped a 18 ‘Single Particle Motion yy A 2B Fig. 2.1 Configuration that leads to an E x B drift certain distance to the left during one orbit. The nct result is that the particle has a drift velocity v, to the left. Let us guess how big the drift speed is. If we average over many gyroperiods, we see that the average accleration is zero. Thus. the net force must be zero. The force downward is g,E), while the force upward is (q./c)va X By. We must have mv =0=4,E, + ty, xB, (2.1) where (__) indicates an average over one gyroperiod. Taking the cross product of (2.1) with By, and assuming v, - By = 0, we find =, EX Bo vg = Be (2.2) Note that the drift velocity does not depend on the particle's charge or mass. EXERCISE What is the drift speed of an electron in the earth’s magnetosphere if |By| = 0.1 G and |E,| = 10° Y cm"? (Remember 1 sV = 300 V.) What if the particle were a uranium atom with charge g, = +57 e? Let us now make sure that our guess is correct, and that the kind of motion we have in mind, namely a gyration about the magnetic field lines, accompanied by a drift, is an exact solution to the equation of motion. This equation is mv = qBy + “ey * By (2.3) { of & Fig. 2.2. Motion of a particle of positive charge leading to the E * B drift. E X B Drifts 19 We define a new variable + by splitting v into a piece ¥ (which will turn out oscillatory) and a piece vg as given by (2.2). Taking vaaty (2.4) (2.3) becomes eee 4s 4s mv = qEy +4 9X By + “yy x By = gy +h yx By + Fy (Ey X By) x By : = oe = YX By (2.5) But the final form of (2.5) is an equation that we have already solved in (1.25) with a solution that represents gyromotion about the magnetic field. Adapting that solution we find % = vy(sin Q,1,cos 01,0) . (2.6) where v, is any constant. Thus, the total solution is x B, v = u,(sin Q,1,cos 2.1,0) + ¢ Eo °. (2.7) Note that any constant uv, is acceptable, including v, EXERCISE Whaat is thc initial velocity implied by the solution (2.7) with v, = 0? Sketch orbits with v, = 0, v, < ug v, = vy, and vy > vy, Note that this entire discussion would apply if an arbitrary (temporally and spatially constant) force F, such that F, -B, = 0 were to replace g,E, in the force equation (2.1). Thus, instead of the drift velocity (2.2), we obtain (2.8) EXERCISE What is the gravitational drift speed of an electron in a tokamak, with [By] = 10 kG? How about a proton? Does either of these drifts make it hard to confine a plasma in a volume of order (1 m)? for a time of order I s? We proceed to discuss other kinds of drifts. We have already seen that any real force gives a drift according to (2.8). We shall now see that any so-called “‘ficti- tious” force also gives a drift. For example, it is sometimes said that magnetic fields exert a “pressure.” This is , of course, not a real pressure, yet we shall find that corresponding to magnetic “pressure” is a drift related to VBy. Likewise, when an existing drift speed changes, the resulting acceleration is experienced as an “‘inertial’’ force, which then gives rise to its own drift. 20 Single Particle Motion 2.3 GRAD-B DRIFT First, let us calculate the so-called grad-B drift. To do this we need to have a feeling for expansion techniques. Suppose we have an equation for a variable x, with one small term of size e, expressed as S(x) — g(x) = 0 (2.9) Here, € is a small constant, f and g can represent a general integro-differential operator, and the solution of (2.9) when € = 0 is xo, so that f(xp) = 0. We can look for a solution x to (2.9) of the form : x=x ten tent... (2.10) Inserting (2.10) in (2.9) yields So t+ ey + ex, t+...) = lo ten tent...) (an After Taylor expanding f and g, one obtains qd 1 @& S(%0) + z (en tex te.) + tz (ete eae ete 2 dg , +... = let) +3 (tem t..dt... (2.12) 0 where df/dx, = df/dx|,,, etc. Equating the coefficients of each power of e yields f(%) = 0 (2.13) which determines Xp, and ef eget &(X0) (2.14) which determines x, = Bho) df/dxo The approximate solution x = x, + ex, is called the “solution of (2.9) to order «.” (Caution: some authors call this ‘the solution to order €”.*) We must always be careful with what we mean by small in these discussions. Something if it goes to zero as € — 0; thus, 10'%e is of order € while 10°" is of order one. Consider a particle gyrating in a magnetic field Bo2 that increases in the f- direction, as shown in Fig. 2.3. Let us guess what happens. The gyroradius will be smaller at large y than at small y, so the particle will drift as shown in Fig. 2.4 for Oy x oy Fig. 2.3 A magnetic field that increases in intensity in the J-direction. i: Grad-B Dritt 21 Fig. 2.4 Ton VB drift. ions and in Fig, 2.5 for electrons. Thus, electrons and ions drift in opposite directions and, in a plasma, a net current results. The force on a charged particle is mw = & (v x B) (2.15) Taylor expanding B about the guiding center of the particle, B =B, + (r- VB, (2.16) where B, is measured at the guiding center, and where r is measured from the guiding center (see Section 1.5), and inserting in (2.15), one obtains = (v ¥ By) + oe [v x (r+ YB] (2.17) ma = EXERCISE What assumption is being made in (2.16)? Expanding veyty, (2.18) we have mV = ome (vo * By) (2.19) which yields gyromotion, and my = E(u, x BY + Ey, * VIB) (2.20) where we are treating r-V as a small quantity, and to be consistent r must be calculated using only vo. Fig. 2.5 Electron VB drift. 22 Single Particle Motion Now we are only interested in that part of v, that represents steady drift motion; therefore, after averaging both sides of (2.20) over a gyroperiod, upon which the left side vanishes, we have o=- 2 Vea +A ko OB) (2.21) Taking v, L By, we obtain (eee v= gr eX @ WB) XB, (2.22) % Since the magnetic field B varies only in the j-direction, aB (e+ V)B) = wy 2 (2.23) Then ig & aB, ap, ve (eM =] on tue 0 |S tony Bj SB 229 By 0 oO y oy From (1.25) and (1.27), are ee wo r= { an 68.8, sin a0) (2.25) while (2.6) can be written vo = vg(sin Nytcos 9.1.0) (2.26) When we average the right side of (2.24) over a gyroperiod, the first term vanishes, and the second term yields for (2.22), using sin? 0,f = 1/2, (2.27) or (2.28) This is the grad-B drift. Recalling that Q, contains the sign of the charge, we see that the drift is in opposite directions for electrons and ions. EXERCISE Given an electron and a proton with equal energies, compare the magnitude of the grad-B drifts. 2.4 CURVATURE DRIFTS Suppose a particle is moving along a field line while gyrating about it. If the field line curves, without changing magnitude, then the particle tries to follow the field line because all motions across field lines are resisted. It therefore feels a centrifugal Curvature Drifts 23 force F, outward (Fig. 2.6), equal to F F.= 2&4, (2.29) Rs Our general drift equation (2.8) then predicts _c¢ FLX B,_ omy? wat = FEB (2.30) or : va = Gia (Ba % Bs) 231) This is the curvature drift. i Ina cylindrically symmetric vacuum field, it turns out that VB, = (—B)/R,)R» (see p. 26 of Ref. [1]); thus we may add the grad-B drift to the curvature drift to obtain we — (Ry X By) ve (ue + Fw) (2.32) Ry where we recall that vy is the perpendicular speed. A rigorous derivation of (2.31) and (2.32) can be found in Refs. (2] and (3). we Ag Fig. 2.6 Centrifuga? force felt by a particle moving along a curved field line. 24 Single Perticle Motion 2.5 POLARIZATION DRIFT We discuss next a drift that is the result of an electric field which varies with time. Since the drift is opposite for oppositely charged particles, it leads to a current called the polarization current. Consider a constant magnetic field Byz, and an electric field E(1) = — E1y, where £ is a constant (Fig. 2.7). The force equation is mi = qb +“ v xB, (2.33) We expect an E X B, drift in the (—)£-direction, which will be increasing with time. Thus, the particle is being accelerated in the (—)%-direction and, therefore, feeis an effective force in the £-direction. (The effective force is in the direction opposite to the acceleration; when one steps on the gas pedal of a car, one is forced backward into the seat.) This effective force should give rise toan F X B drift in the §-direction. Using this intuition, we consider a solution of (2.33) of the form v=o + ust + vf (2.34) where vq will contain all gyromotion, v,is the E % B drift, and v, is the polariza- tion drift, which is assumed to be constant. Substituting (2.34) into (2.33), we obtain 4 c milvy + ¥) = —qQéty + aa wo X By vBp +E vee (2.35) The assumed nature of the solution indicates the separation of this equation into pieces, with qs mVq = “po X Bo (2.36) representing gyromotion, ive = Ev, Bol (2.37) giving the polarization drift, and 0 = — g.ity — Sry 2.38) 4 oA, E() g Fig. 2.7 Configuration that leads to a polarization drift Magnetic Moment 25 giving the E x B drift, namely, cét upt = — a (2.39) Then (2.37) yields the polarization drift cm cE ec £ b(-) Se - 2.40 aby By 0, Bo ten which in vector form is (2.40) We see that (2.34) is an exact solution to (2.33). EXERCISE If E(t) = Ey cos wt, can you invent a criterion for the validity of (2.41) at each instant of time? The polarization drift leads to a polarization current J,, of electrons and pro- tons, given by nyc? dE J, = moe. — Yoo) = Bar Gp (me + md (2.42) or Puc? dE = oF a (2.43) where p,, is the mass density 2.6 MAGNETIC MOMENT The preceding sections have discussed drifts due to “forces” perpendicular to the magnetic field. There are also forces parallel to the magnetic field that are very important, leading to the concepts of magnetic moment and adiabatic invariants. Recall that the magnetic moment of a current loop with current /, area A, in c.g.s. units, is pat (2.44) Acharged particle gyrating in a magnetic field is such a current loop, with current 9,0,/2n, area tp? = mv2/02 (p, is the gyroradius), and magnetic moment tae aAbeee Ua eet Yom ,? = w = sek 2ae a? (2.45) u or (2.46) 26 ‘Single Particle Motion SE . eee z Orbit Fig. 2.8 Magnetic field with nonzero grad-B. The orbit shown is a circle in a plane perpendicular to the paper. For the discussion following (2.51), the orbit is actually a spiral with a velocity component vy in the B-direction. where W, = %m,v,? is that portion of a particle’s kinetic energy which is per- pendicular to the magnetic field. We know that magnetic moments feel a force —4VB in an inhomogeneous magnetic field. How does this work out for a charged particle? Consider a particle gyrating about the axis of a cylindrically symmetric magnetic field, whose magni- tude is changing along the axis, as shown in Fig. 2.8. EXERCISE Does the field in Fig. 2.8 satisfy Maxwell's equations? In the figure, the vertical line is a side view of the gyrating particle. Notice that at the top of the orbit (Fig. 2.9), for a positively charged particle, the v X B force has one component giving gyromotion about the field, and another component point- ing in the (—)£-direction. This latter component is constant around the gyro-orbit, and the particle is steadily accelerated away from regions of strong field. EXERCISE Show that this works the same for either sign of. the charge. The force in the (—)£-direction, evaluated anywhere on the orbit, is 4 » py, =e vB, (2.47) ¢ ¢ where r is the distance from the x-axis, and B, is the component of the magnetic field in the y-z plane in Fig. 2.8. In cylindrical coordinates, OB, 1 @ ax +r or PB) Solving this equation with B, = 0 atr = 0, and B, << B, everywhere, one ob- F= VB=0= @: > vxB Fig. 2.9 The vectors 3, #, and v x B for a particle with g, > 0 at the top of the orbit in Fig. 2'8. Magnetic Moment 27 tains r 6B, r B=-> 357 7 vel (2.48) Inserting (2.48) in (2.47), with r equal to the gyroradius p,, we obtain y= — tl & yop = — MYL yp (2.49) C2 2B or VB (2.50) as expected, Knowing the force on the particle allows the calculation of its orbit. First one needs to know how the magnetic moment m changes along the orbit. The remark- able fact is that the magnetic moment is constant along the orbit, provided the field does not change much in one gyroperiod. Let us prove this. The Lorentz force on a charged particle (with E = 0) is F = (g,/c)v X B.A small component F, of this force acts to accelerate the particle in the direction perpendicular to the Jocal magnetic field and parallel to the component of the patticle velocity vj used in the definition of y. (In Fig. 2.8, let the particle shown have a positive velocity component v) along £. Then at the top of the orbit, v x B has a component of magnitude |v, 8,| pointing into the paper.) This force is into the paper at the top of the orbit of the particle in Fig. 2.8, and is given by F=- up, (2.51) where v, is the component of velocity in the £-direction and B, is negative. The perpendicular energy of the particle then changes with time according to a ieefal 4s + (> mv?) = uF, = 2 yop, (2.52) When we use (2.48), this becomes (with r = p, at the location of the particle) @ qe ep, OB, _ 4. a Mime MD ae ee ae 258) where B, ~ B, or d 1 oB a =F oe (2.54) EXERCISE Combine (2.50) and (2.53) to show that total energy is conserved, The time rate of change of the magnetic moment is dud (Wy) _ 1 dw, aB wnals)ez ii eed ee ede 1 o_o aB Ti lReer cept tn iag =0 (2.55) 28 Single Particle Motion where the rate of change of B along the particle orbit dB/d! = v, 8B/ax has been used. Thus, (2.56) along the particle orbit, to within the accuracy of this calculation. This is an example of an adiabatic invariant, a quantity that is constant under slow changes in an externa] parameter. Although our derivation treated spatially varying mag- netic fields, the same result holds for magnetic fields with slow time variation. We are now in the position to understand the principle of mirror confinement, which is the basis for one of the two major approaches to magnetic fusion, and is also the reason for the existence of the carth’s magnetosphere. Consider a magnet- ic field created by two coils, as shown in Fig. 2.10, A particle that starts atx = 0 with energy W,, and magnetic moment y conserves both of these quantities. Sup- pose the particle initially has v, > 0; it moves to the right and feels a force to the left, F = —pVB. Does the particle get reflected by the force, or does it go past X = Xo to be lost from the machine? This will depend on its initial v,° and v,°. We have Yomu? amu w= a = RF const (2.57) and W = Yon(v? + v3) = Wy = const (2.58) As the particle moves to the right, B increases and W, = “emu? must increase to satisfy (2.57). If W, ever reaches Wy, then all the energy will be in perpendicular motion, vy will vanish, and the particle will be reflected back toward the mirror machine. This happens if hmv Wy amu + vf) eRe Bag eae cae or Pe Basin > a (2.60) uP + ug ~ Boo Defining the pitch angle 8 = tan (v{'/v)), we have from (2.60) eae mex #=0, Ss Fig. 2.10 ——~ mirror machine configuration. Adiabatic Invariants 29 (2: (2.61) Bross where we have introduced the mirror ratio R = Bw,/Bmia. Particles whose pitch angles satisfy (2.61) at the center of the machine are confined. Those that do not are lost out from the ends. While our derivation applies only to particles circling the ccntral magnetic field line, a similar statement is true for off-axis particles. 2.7 ADIABATIC INVARIANTS The magnctic moment, constant under slow spatial or temporal changes in the magnetic field, is one example of an adiabatic invariant. It often turns out that ina system with a coordinate q, and its conjugate momentum p, the action, defined by y= $e dq (2.62) is a constant under a slow change in an external parameter. Here. we have as- sumed that when there is no change in the external parameter, the motion is periodic, andf represents an intcgral over one period of the motion. In the case of a charged particle in a magnetic field, we could take, for example, measuring x from the guiding center: p = mv,,q = x, and 20/2. 1 =$ mo, de= Po mu} sin® (0.0) dt 7 a (2.63) amy ,? _ 2mmre [Way 2mre a 4 () aug: which is the magnetic moment to within a constant. One famous example of the constancy of action was derived at the 1911 Solvay conference, where Lorentz asked: “What happens if we slowly shorten the string of a swinging pendulum?” The next morning, Einstein answered: “Action [= energy/frequency] is conserved.” Let us now demonstrate the invariance of the action in the general case. We have in mind the picture of a particle bouncing in a potential well, with the shape of the well changing slowly with time, as shown in Fig, 2.1). In Hamiltonian mechanics, we have a Hamiltonian H(p,g,A) where p is the momentum (mx in the Vix) * * Fig. 2.11 Peviodic motion in a slowly changing potential well 30 Single Particle Motion figure), g is the coordinate (x in the figure), and A is the parameter that determines the shape of the well. When A is constant, the Hamiltonian is constant; when A changes, dH _ aH dx de ON dt (2.64) The Hamiltonian equations of motion are ee aH oe Dp eee a (2.65) The time derivative of the action 9; sah ody = 2)" pay (2.66) is 5 = 24" bag + read» — 9G 2.67) The last two terms vanish since the momentum is zero at the turning points; then : 2 OH b= 2 f° = ag = -20H19) — Hol =0 (2.68) a Og since 0H/dq is to be taken at fixed A, implying H(q,) and H(q,) are to be taken at fixed A, and H is a constant except for its variation with A, We have thus shown the approximate constancy of any action variable in the presence of periodic motion and slow changes of external parameters. Note that the present derivation is heuristic, since it has been assumed that the variation is so slow that the turning points g,(t) and q,(£) can be treated as continuous functions of time and can be differentiated as in (2.67) [4]. A rigorous treatment of this problem can be found in the fundamental paper of Kruskal [5]; see also Goldstein [6]. The knowledge of an adiabatic invariant can be very useful in predicting particle behavior, Let us return to the mirror machine, where the constancy of the magnet- ic moment adiabatic invariant has already enabled us to determine which particles will be confined and which will be lost. Consider a confined particle. The confined particle executes periodic motion between x, and x,. Thus, J =$p do =$ mv, de (2.69) must be a constant of the motion, even when the entire mirror field undergoes slow temporal changes, or when the mirror field is not axisymmetric. There is yet a third adiabatic invariant. As the charged particle bounces from x, to x2, it drifts perpendicular to the field lines, with speed v,. Eventually, it comes all the way around the mirror machine. Because this is like a huge gyro-orbit about the axis of the magnetic field, we define a new adiabatic invariant sap vd (2.70) where / is the distance around the mirror machine measured at some fixed x, for example x;. It turns out that J; is proportional to the total magnetic flux enclosed by the drifting motion. This invariant is useful when the mirror field is not axisym- metric, or when it undergoes slow temporal changes. Ponderomotive Force ao EXERCISE Sketch the carth’s magnetosphere, and discuss the motion leading to 4, J,, and Js. 2.8 PONDEROMOTIVE FORCE All of the examples of single particle motion considered in previous sections involve motion in a magnetic field, with or without an electric field. There is one very important single particle effect, the ponderomotive force or Miller force, that occurs in spatially varying high frequency electric fields, with or without an ac- companying magnetic field. Consider a charged particle oscillating in a high fre- quency electric field E(t) = Ey cos (wt). The motion is then a sinusoidal variation of distance with time, as shown in Fig. 2.12. Now suppose the electric field has an amplitude that varies smoothly in space, E(x,t) = E(x) cos (wt), being stronger to the right and weaker to the left. Then the first oscillation brings the particle into regions of strong field, where it can be given a strong push to the left (sce Fig. 2.13). When the field turns around, the particle is in a region of weaker field, and the push to the right is not as strong. The net result is a displacement to the left, which continues in succeeding cycles as an acceleration away from the region of strong field [7] Mathematically, the force equation is me = QE = q(x) cos wt (2.71) It is convenient to decompose x into a slowly varying component xp, called the oscillation center (compare this concept to the guiding center in a magnetic field) and a rapidly varying component x,, x = x9 + x; Here, x» = % where (__) indicates a time average over the short time 27/w. Making a Taylor expansion of t Fig. 2.12 Sinusoidal motion of a charged particle in a high-frequency electric field. 32 Single Particle Motion Weak Strong ' Fig. 2.13 Motion of a charged particle in a high-frequency electric field that is weaker to the left and stronger to the right. E(x) about the oscillation center xy, (2.71) becomes aE, m&o + #1) = 4, (F tx, S) cos wt (2.72) where dE,/dx is to be evaluated at xo. Averaging (2.72) over time, we get m&o = I a ¥ cos of (2.73) % To obtain an equation for x,, we note that ¥, >> %» since x, is high frequency, moreover, in the spirit of the Taylor expansion we have Ey >> x,(aB,/dx); there- fore (2.72) is approximately m,%, = QE Cos wt (2.74) with solution x, = —(g,E,/m,«°)(cos wf); inserting this in (2.73) and performing the time average one obtains e = — 40 aEo fo — meat dx (2.75) so that the ponderomotive force F, = m,Xp is (2.76) This formula will be easier to remember if we introduce the jitter speed § = (mae = 9,Eo/m,w; then Ra-fRZ@ (2.77) This force is very important in such applications as laser fusion, electron beam fusion, radio frequency heating of tokamaks, radio frequency plugging of mirrors, radio frequency modification of the ionosphere, and solar radio bursts. The study Oitfusion 33 of the effects of ponderomotive force is one of the areas of current basic plasma physics research. Notice that the overall mass dependence is as given in (2.76), so that the ponderomotive force acts much more strongly on electrons than on ions. A more complete derivation of the ponderomotive force, including the magnetic field in an electromagnetic wave, can be found in Schmidt [4]. 2.9 DIFFUSION We conclude this chapter with a brief discussion of the effects of collisions on the location of the guiding center of a particle in a magnetic field. The discussion of Chapter | shows that the effects of many small angle collisions in a plasma are more important than the effects of rare large angle collisions. However, it is simplest to consider here a single large angle collision between two charged parti- cles; the results can then be qualitatively applied to determine the effects of many small angle collisions. Consider first the head-on collision between two electrons at x = 0, as shown in Fig. 2.14. The last gyro-orbit, and the guiding center, of each particle before the collision are indicated in the upper half of the figure. After the collision, electron a8 Fig. 2.14 Head-on collision between two electrons in a magnetized plasma. Numbers indicate the location of the guiding centers of the two electrons. 34 Single Particle Motion number 2 has the same orbit that electron number i had before the collision, and vice versa. Thus, the locations of the two guiding centers have been interchanged, and there is no net motion of the electrons. We conclude that collisions between like particles do not lead to diffusion of those particles across magnetic field lines. Next, consider the (almost) head-on collision between an electron and a slightly more energetic positron at x = 0, as shown in Fig. 2.15. The last gyro-orbit and the guiding center of each particle before the collision are indicated in the upper half of the figure. After the collision, the electron has slightly more energy than the positron, and both guiding centers have moved by two gyro-radii in the (—)2- direction. Thus, the center-of-mass of the system has moved a substantial distance in the (~)2-diréction. We conclude that collisions between unlike particles can cause significant diffusion of particles across magnetic field lines. Further discus- sion of diffusion can be found in Ref. [1] This completes our discussion of single-particle motion in prescribed electric and magnetic fields. In the next chapter, we begin a systematic treatment of plasma physics in which the electromagnetic fields and the particle orbits are determined self-consistently. BEFORE AFTER «3 Fig. 2.18 Head-on collision between an electron and a slightly more energetic positron in a magnetized plasma Problems 35 REFERENCES (1] F. F. Chen, Introduction to Plasma Physics, Plenum, New York, 1974. [2] A. Bafios, Jr., J. Plasma Phys., 1, 305 (1967). [3] T. G. Northrop, The Adiabatic Motion of Charged Particles, Wiley, New York, 1963. [4] G. Schmidt, Physics of High Temperature Plasmas, Academic, New York, 1966. [5] M. Kruskal, J. Math, Phys., 3, 806 (1962). [6] H. Goldstein, Classical Mechanics, 2nd ed., Addison-Wesley, Reading, Mass., 1980, Sect. 11-7. {7} L. D. Landau and E. M. Lifshitz, Mechanics, Addison-Wesley, Reading, Mass., 1960, Sect. 30. PROBLEMS 2.1 Example of a Drift (a) Consider a particle of charge g and mass m, initially at rest at (x,y,2) = (0,0,0), in the presence of a static magnetic field B = By2 and E = E,f. Taking Ey, By > 0, sketch the orbit of the particle when q > 0. (b) Derive an exact expression for the orbit [x(#), »(4), 2(1)] of the particle in part (a). Does this result agree with the sketch of part (a)? (c) Show that the orbit in (b) can be separated into an oscillatory term and a constant drift term. After averaging in time over the oscillatory motion, is there any net acceleration? If not, how are the forces in the problem bal- anced? (d) In what direction is the drift for g > 0? For q < 0? If there were many particles of various charges and masses present, would there be any net current? (e) Suppose the electric field were replaced by a force Fo in the f-direction. What would be the drift velocity? [Hint: Guess the answer using the last part of (c).] 2.2 Grad-B Drift Let us derive the grad-B drift in a different way. With the force equation F = (q/c)v X B, insert the zero order orbit and the Taylor expanded magnetic field. Average F over one gyroperiod to obtain an average force. Insert this average force into the general drift equation (2.8), and compare the resulting drift to the grad-B drift (2.28). 2.3 Polarization Drift Let us get the polarization drift (2.41) in a faster but less rigorous manner. With the given electric field E = —Etf, calculate an E x B drift. Relate the resulting Kl Single Perticle Motion accelerated drift to a force (being careful with signs), plug in the F X B formula (2.8), and compare the result to (2.41). 2.4 Mirror Machines (a) A mirror machine has mirror ratio 2, A Maxwellian group of electrons is released at the center of the machine. In the absence of collisions, what frac- tion of these electrons is confined? (b) Suppose the mirror machine has initially equal densities n ~ 10° cm™ of electrons and protons, each Maxwellian with a temperature I keV = 107°C. The machine is roughly one meter in size in both directions. Recalling our discussion of collisions from Chapter 1, estimate very roughly the time for (1) loss of the unconfined electrons; (2) loss of the unconfined ions; (3) Joss of many of the initially confined electrons (due primarily to which kind of collision?); why do not all of the electrons leave?, loss of the initially confined ions (due primarily to which kind of col- lision?). For fusion purposes (supposing the protons were replaced by deuterium or tritium) which of these numbers is the most relevant? 4 2.5 Drift Energy A particle of mass m and charge g in a uniform magnetic field B = By? is set into motion in the £-direction by an electric field E(1)9 that varies slowly from zero toa final value £y. Thus, at the final time the particle has an E X B drift vo. (a) Use energy arguments to show that the particle’s guiding center must have been displaced a distance vy/Q in the direction of the electric field. (b) Integrate the polarization drift velocity from time zero to time infinity to obtain a dispfacement. Does the answer agree with (a)? CHAPTER Plasma Kinetic Theory |: Klimontovich Equation 3.1. INTRODUCTION In this chapter, we begin a study of the basic equations of plasma physics. The word “kinetic” means “pertaining to motion,” so that plasma kinetic theory is the theory of plasma taking into account the motions of all of the particles. This can be done in an exact way, using the K/imontovich equation of the present chapter or the Liouville equation of the next chapter. However, we are usually not interested in the exact motion of all of the particles in a plasma, but rather in certain average or approximate characteristics. Thus, the greatest usefulness of the exact Klimon- tovich and Liouville equations is as starting points for the derivation of approxi- mate equations that describe the average properties of a plasma. In classical plasma physics, we think of the particles as point particles, each with a given charge and mass. Suppose we have a gas consisting of only one particle This particle has an orbit X,(¢) in three-dimensional configuration space x. The orbit X,(7) is the set of positions x occupied by the particle at successive times f, Likewise, the particle has an orbit V,(¢) in three-dimensional velocity space v. We combine three-dimensional configuration space x and three-dimensional velocity space v into six-dimensional phase space (xv). The density of one particle in this phase space is N(x,v,t) = 6x — X,()]6[v — V,(2)] G.1) where é[x — X,] = &(x — X,)&(y — ¥,)&(z — Z,), ete. (The properties of the Dirac delta function are reviewed in Ref. [1], p. 29, and in Ref. [2], pp. 53-54.) Note that X,, V, are the Lagrangian coordinates of the particle itself, whereas x, v are the Eulerian coordinates of the phase space. 38 Plasma Kinetic Theory | EXERCISE At any time 1, the density of particles integrated over all phase space must yield the total number of particles in the system. Verify this for the density (3.1). Next, suppose we have a system with two point particles, with respective orbits [X,(), V,(0] and (X,(1), V,(1)] in phase space (x,v). By analogy to (3.1), the particle density is 2 Naw.) = Yolk — XO — VC] (3.2) «1 : EXERCISE Repeat the previous exercise for (3.2). Now suppose that a system contains two species of particles, electrons and ions, and each species has N, particles. Then the density N, of species s is Ny N.Guvt) = 8x — X(NJolv — V1) (3.3) and the total density N is i N(xv.l) = DN,0uy,0) (3.4) EXERCISE Repeat the previous exercise for (3.4) If we know the exact positions and velocities of the particles at one time, then we know them at all Jater times. This can be seen as follows. The position X,(¢} of particle i satisfies the equation XQ = VA) (3.5) where an overdot means a time derivative. Likewise, the velocity V(‘) of particle # satisfies the Lorentz force equation 4g. 2 ¥(1) x BX(1).1) (3.6) mV (1) = QE X(1).e] + where the superscript m indicates that the electric and magnetic fields are the microscopic fields self-consistently produced by the point particles themselves, together with externally applied fields. [On the right of (3.6), the portion of E” and B” produced by particle / itself is deleted.] The microscopic fields satisfy Maxwell's equations V+ E(x) = 4p(x,1) (3.7) Vv - B%x,4) = 0 (3.8) Vx E(x) = — = eo G9) and 1 dE"(x,1) Vv xX B%x,1) = « I(x + ie oF (3.10) Klimontovich Equation 39 The microscopic charge density is ext) = > as fav N,O.0 (3.11) while the microscopic current is : mx) = Sa J awwN,(X,,1) (3.12) a EXERCISE Convince yourself that (3.11) and (3.12) yield the correct charge density and current. Equations 3.7 to 3.12 determine the exact fields in terms of the exact particle orbits, while (3,5) and (3.6) determine the exact particle orbits in terms of the exact fields. The entire set of equations is closed, so that if the positions and velocities of all particles, and the fields, are known exactly at one time, then they are known exactly at all later times. 3.2 KLIMONTOVICH EQUATION An exact equation for the evolution of a plasma is obtained by taking the time derivative of the density N,. From (3.3), this is INGeed =— 3 XV ate — X(n}élv — VA) a — EV Vale — Xuma — VOI G3) el where we have used the relations 8 ee eq (a — 4) = — Gp fla — b) and i a a TRO) = Ge 8 and where V, = (8,,4,,3,) and Vy = (0,,,,,.0,,). Using (3.5) and (3.6), we can write X; and V; in terms of V,and the fields E” and B”, whereupon (3.13) becomes ONAX V1) a No — XL V,-V,8tx — XJdlv — VJ A Exo, + V, x B'IX(),) } +V, x — XJolv — Vi) 3.14) An important property of the Dirac delta function is aia — b) = b&(a — 6) EXERCISE. How would one prove this relation? ao Plasma Kinetic Theory | This relation allows us to replace V(t) with v, and X(t) with x, on the right of (3.14) (but not in the arguments of the delta functions) so that (3.14) becomes NOV.) _ — vi Vy 2 8x — XJalv — VJ ® - [& E(x.) + me v x B%(x,0) |: vy Py dx — X,lélv — VJ (3.15) But the two summations on the right of (3.15) are just the density (3.3); therefore ING) ye, + 2 (en +4x Be )-va, = o| (3.16) ar m, c This is the exact Klimontovich equation (Klimontovich [3]; Dupree [4]). The Klimontovich equation, together with Maxwell’s equations, constitute an exact description of a plasma. Given the initial positions and velocities of the particles, the initial densities N,(x,v,t = 0) and N{x,v,f = 0) are given exactly by (3.3). The initial] fields are then chosen to be consistent with Maxwell's equations (3.7) to (3.12). With these initial conditions the problem is completely determinis- tic, and the densities and fields are exactly determined for all time. In practice, we never carry out this procedure. The Klimontovich equation contains every one of the exact single particle orbits. This is far more information than we want or need. What we really want is information about certain average properties of the plasma. We do not really care about all of the individual electro- magnetic fields contributed by the individual charges. What we do care about is the average long-range electric field, which might exist over many thousands or millions of interparticle spacings. The usefulness of the Klimontovich equation comes from its role as a starting point in the derivation of equations that describe the average properties of a plasma The Klimontovich equation can be thought of as expressing the incompressibili- ty of the “substance’’ N,(x,v,/) as it moves about in the (x,v) phase space. (Is it any wonder that a point particle is incompressible?) This can be seen as follows. Imagine a hypothetical particle with charge g,, mass m,, which at time f finds itself at the position (x,v). This hypothetical particle has an orbit in phase space deter- mined by the fields in the system. Imagine taking a time derivative of any quantity along this orbit (such.a time derivative is called a convective derivative). This derivative must include the time variation produced by the changing position in (x,v) space as well as the explicit time variation of the quantity. Thus, it must be given by D a dx iDineron tan . G17) oesit orbit where by dx/df|om we mean the change in position x of the hypothetical particle with time; tikewise for dv/dt| it. But for our hypothetical particle at position (x,v) Plasma Kinetic Equation a in phase space we know that : ax A Nowe ca and Wl = 4 ppm ea 7 fe = Fe EM + DBD (3.19) Thus, Diese reas oe * x Bp]: Diente (E+ > * BMD] V, 3.20) and the Klimontovich equation (3.16) simply says 2 Nx.) (3.21) The density of particles of species s is a constant in time, as measured along the orbit of a hypothetical particle of species s. This is true whether we are moving along the orbit of an actual particle, in which case the density is infinite, or whether we are moving along a hypothetical orbit that is not occupied by an actual particle, in which case the density is zero. Note that the density is only constant as measured along orbits of hypothetical particles; in (x,v) space at a given time it is not constant but is zero or infinite. There is yet a third way to think of the Klimontovich equation. Any fluid in which the fluid density (r,t) is neither created nor destroyed satisfies a continuity equation 8, flrt) + Ve (FV) = 0 (3.22) where V, is the divergence vector in the phase space under consideration, and V is a vector that gives the time rate of change of a fluid element at a point in phase space. (See, for example, Symon [5], p. 317.) In the present case, Ve = (Vy.Vy) and V = (dx/dt\ obi, dv/dt|ovi:). Since the particle density is neither created nor destroyed, it must satisfy a continuity equation of the form 8,N,(XV,1) + Vy (VN) + Vy: {2 [Ee +tx Be|w,} =0 (3.23) It is left as a problem to demonstrate that the continuity equation (3.23) is equiv- alent to the Klimontovich equation (3.16). 3.3. PLASMA KINETIC EQUATION Although the Klimontovich equation is exact, we are really not interested in exact solutions of it. These would contain all of the particle orbits, and would thus be far too detailed for any practical purpose. What we really would like to know are the average properties of a plasma. The Klimontovich equation tells us whether or not a particle with infinite density is to be found at a given point (x,y) in phase space. 42 Plasma Kinetic Theory | ‘What we really want to know is how many particles are likely to be found in a small volume Ax Av of phase space whose center is at (x,v). Thus, we really are not interested in the spikey function N,(x,v,4), but rather in the smooth function SAY) = (N.0GY1) (3.24) The most rigorous way to interpret (_ ) is as an ensemble average [6) over an infinite number of realizations of the plasma, prepared according to some pre- scription. For example, we could prepare an ensemble of equal temperature plas- mas, each in thermal equilibrium, and each with a test charge g, at the origin of configuration space. The resulting f, and f, would then be consistent with the discussion of Debye shielding in Section 1.2. There is another useful interpretation of the distribution function f,(x,v,t), the number of particles of species s per unit configuration space per unit velocity space, Suppose we are interested in long range electric and magnetic fields that extend over distances much larger than a Debye length. Then we can imagine a box, centered around the point x in configuration space, of a size much greater than a mean interparticle spacing, but much smaller than a Debye length (this is easy to do in a plasma; why?) We can now count the number of particles of species sin the box at time 1 with velocities m the range vto v + Av, divide by (the size of the box multiplied by Av, Av, Av,), and call the result /,(x,v,4). This number will of course fluctuate with time but, if there are very, many particles in the box, the fluctuations will be tiny and the f,(x,v,) obtained in this manner will agree very well with that obtained in the more rigorous ensemble averaging procedure. An equation for the time evolution of the distribution function f,(x,v,/) can be obtained from the Klimontovich equation (3.16) by ensemble averaging. We de- fine 6N,, 6E, and 6B by NO¥,0) = fiXG¥,t) + SNe(x,¥.0) Er(x,v,0) = E(xv,0) + 6E(x,y,2) (3.25) and Br(x,v,0) = B(xv,t) + 6B(x,v,0) where B = (B”),E = (E”), and (6N,) = (SE) = (8B) = 0. Inserting these def- initions into (3.16) and ensemble averaging, we obtain OF%Y,1) qs v > thet oe (E+ FX BY OS: (3.26) aw + : =~ Fe (GE +] * OB)-V, 6N,) Equation (3.26) is the exact form of the plasma kinetic equation. We shall meet other forms of this equation in the next chapter. The left side of (3.26) consists only of terms that vary smoothly in (x,v) space. The right side is the ensembie average of the products of very spikey quantities like éE = E” — (E”) and 5N,. Thus, the left side of (3.26) contains terms that are insensitive to the discrete-particle nature of the plasma, while the right side of (3.26) is very sensitive to the discrete-particle nature of the plasma. But the discrete-particle nature of a plasma is what gives rise to collisional effects, so that Pleama Kinetic Equation 43 the left side of (3.26) contains smoothly varying | functions representing collective effects, while the right side represents the nal effects. We have seen in Section 1.6 that the ratio of the importance ‘of collisional effects to the importance of collective effects is sometimes given by: 1/A, which is a very small number. We might guess that for many phenomena in a plasma, the right side of (3.26) has a size 1/A compared _to each of the terms on the left side; thus the right side can be neglected for the study of such phenomena. This indeed is the case, as shown in the next two chapters. This important point can be illustrated by a hypothetical exercise. Imagine that we break each electron into an infinite number of pieces, so that.ty — 3m, — 0, and e — 0, while age = constant, ¢/m, = constant, and: = constant. EXERCISE Show that in this hypothetical exercise, w, = constant, A, = constant, but T, + Oand A, > &. : Then any volume, no matter how small, would contain an infinite number of point particles, each represented by a delta function with infinitesimal charge. Statistical mechanics tells us that the relative fluctuations in such a plasma would vanish, since the fluctuations in the number of particles No in a certain volume is propor- tional to the square root of that number: Thus, on the right side of (3.26) we have 6N, ~ No ~ A}, and 6E and 6B, which are produced by 5N, behaving like (from Poisson's equation) ~ e6N, ~ Ny'No? ~ Ny! ~ A”, so that the right side becomes constant. On the left, however, each term becomes infinite asf, — =. Thus, the relative importance of the right side vanishes ~ No"! ~ A,', and we have 3.27) which is the Vlasov (7) equation (sometimes referred to as the collisionless Boltz- mann equation). This approximate equation, which neglects collisional effects, is often called the most important equation in plasma physics. Its properties will be explored in detail in Chapter 6. The fields E and B of (3.27) are the ensemble averaged fields of (3.25). They must satisfy the ensemble averaged versions of Maxwell's equations (3.7) to (3.12), which are V-E(x,t) = 4ap VBE) =0 1 oB Nee EO ie core api 4a 1 0E VXBR) = I+ os) = 0 = Za farsoun and Jao =ay=> anf avefew.t) (3.28) 44 Plasma Kinetic Theory 1 In the next two chapters we shall approach the plasma kinetic equation (3.26) from another direction, and shall use approximate methods to evaluate the colli- sional right side, In Chapter 6 we shall take up the study of the Viasov equation (3.27) REFERENCES a) @ (3) [4] 5] (6) {71 J. D. Jackson, Classical Electrodynamics, 2nd ed., Wiley, New York, 1975. K. Gottfried, Quantum Mechanics, Benjamin, New York, 1966. Yu. L. Klimontovich, The Statistical Theory of Non-equilibrium Processes in a Plasma, M.I.T. Press, Cambridge, Mass., 1967. T. H. Dupree, Phys. Fluids, 6, 1714 (1963). K. R. Symon, Mechanics, 3rd ed., Addison-Wesley, Reading, Mass., 1971. F. Reif, Fundamentals of Statistical and Thermal Physics, McGraw-Hill, New York, 1965. A.A. Vlasov, J. Phys, (U.S.S.R.), 9, 25 (1945). PROBLEM 31 Klimontovich as Continuity Prove that the continuity equation (3.23) is equivalent to the Klimontovich equa- tion (3.16). CHAPTER Plasma Kinetic Theory II: Liouville Equation 4.1 INTRODUCTION In addition to the Klimontovich equation, there is another equation, the Liouville equation, which provides an exact description of a plasma. Like the Klimontovich equation, the Liouville equation is of no direct use, but provides a starting point for the construction of approximate statistical theories. One of the most useful practical results of this approach is to provide us with an approximate form for the right side of the plasma kinetic equation (3.26), which tells us how the distribution function changes in time due io collisions. The Klimontovich equation describes the behavior of individual particles. By contrast, the Liouville equation describes the behavior of systems. Consider first a “system” consisting of one charged particJe. Suppose we measure this particle’s Position in a coordinate system,X;; then the orbit of the particle X,(/) is the set of positions x, occupied by the particle at consecutive times ft, Likewise, in velocity space we denote the orbit of the particle V,(t); this is the set of velocities taken by the particle at consecutive times f; these velocities are measured in a coordinate system v,. We thus have a phase space (X).¥;) = (%1.)1521s¥q)s¥y,s¥s,)- In this six- dimensional phase space there is one “system” consisting of one particle. The density of systems in this phase space is N(x,v.0) = 604 — X41 — Vi) (4.1) Next, consider a system of two particles. We introduce a set of coordinate axes for each particle, Particle 1 has (x;,v,) coordinate axes as before. Particle 2 has (X2,V2) coordinate axes that lay right on top of the (x,,v,) coordinate axes. The orbit X,(1), V,(1) of particle | is measured with respect to the (x,,V,) coordinate axes, while the orbit X,(f), V2(1) of particle 2 is measured with respect to the (x,,¥,) coordinate 46 Plasma Kinetic Theory II axes. We now introduce an entirely new phase space, having twelve dimensions. The phase space is (Kp s¥po¥ V2) = (XV 9Z1 5x59 Vy, 9 Vej2X2V20Z99 Vays Vyss Vea) (4.2) In this twelve-dimensional phase space, there is one system that is occupying the point [x, = X,(0), ¥, = Vi(1), x = X2(2), v2 = V2(0] at time #. The density of systems in this phase space is N(Xi5¥: Ka Vat) = Ole, — XCOlv, — Vi) 8%. — XaCH6[ve — VC] (43) EXERCISE Show that there is indeed one system in the phase space by integrat- ing the density (4.3) over all phase space. Note that the density Niin (4.3) is completely different from the density N, used in the previous chapter in the discussion of the Klimontovich equation. The density N, in Ch. 3 is the density of particles in six-dimensional phase space. The density N in (4.3) is the density of systems (each having two particles) in twelve-dimensional phase space. Finally, suppose that we have a system of N, particles. With each particle i, i = 1,2, ... No, we associate a six-dimensional coordinate system (x;,¥,). Using these 6M, coordinate axes, we construct a 6N,-dimensional phase space, analogous to the twelve-dimensional phase space in (4.2). There is one system in 6No-dimen- sional phase space; therefore the density of systems, by analogy with the density of systems (4.3), is Ny NO ViXa¥2 + XN Wy) = TD alas — XOILv, — VAT (4.4) where II, ff = fifi. - Sn EXERCISE Use (4.4) to prove that there is one system in all of phase space. 4.2 LIOUVILLE EQUATION As with the Klimontovich equation in Chapter 3, the Liouville equation is ob- tained by taking the time derivative of the appropriate density. In this case, we take the time derivative of the density of systems (4.4). Because the density of systems (4.4) is the product of 6N, terms, its time derivative involves the sum of 6N, terms. Using the relation a aX, or 8% — KM) = — G+ Va al, — X10) (4.5) and similar relations encountered in the previous chapter, the time derivative of (4.4) is Ne vg Ne Bey wl: vy, IT ay - xy - VD i 1 Meee Ny +L WV, TT x — Xo — V;) = 0 (4.6) AL pe Liouville Equation a7 Using aé(a — 5) = 86(a — 6) to replace V, by v;, and similarly for V,so that for the remainder of this chapter Vi) = ef E(x, + — x % x Brian | (4.7) and noting that the products are just the pale of systems N, (4.6) becomes (4.8) which is the Liouville equation, When combined with Maxwell’s equations and the Lorentz force equation, the Liouville equation is an exact description of a plasma For a two-component plasma with N,/2 electrons and N,/2 ions, the expression for V(t) will depend upon whether the ith particle is an electron or a proton. The Liouville equation has all of the advantages and all of the disadvantages of the Ktimontovich equation. Because.it contains all of the exact six-dimensional orbits of the individual particles in a single system orbit in 6Ny-dimensional space, it contains far more information than we want or need. Its usefulness is as a starting point in deriving a reduced statistical description, which with appropriate approx- imations can yield practical information Equation (4.8) has the form of a convective time derivative in the 6N,-dimen- sional phase space, 2 NOK Vi X20Vay > XyWnort) = (4.9) where D Pett duntE vos, (4.10) Here, V,(1) is expressed in terms of the position (X,,¥,,%2,¥3. » - - » Xw»¥w,) of the system in 6N -dimensional phase space, since that position determines the posi- tions of the particles in six-dimensional space and thus the fields at all points in six-dimensional space through Maxwell's equations. Thus, the convective time derivative, taken along the system orbit in 6No-dimensional phase space, is zero. The density of systems is incompressible. The Liouville equation (4.8) can also be put in the form of a continuity equa- tion. Recall the vector identity V+ (ab) = b+ Va + aV +b. Then Wt VN = Vas (ve) (4.11) since v, and x, are independent variables. Similarly, Vi VN = V5 (VN) (4.12) since We =: {2 [emo + ee x B"(x;,t) }} =0 (4.13) EXERCISE Prove (4.13). 48 Plasma Kinetic Theory Ii Then the Liouville equation (4.8) becomes an & & . areas TV (WN) + Y V+ (VN) =0 (4.14) it el In the form of a continuity equation, the Liouville cquation expresses the conser- vation of systems in 6N,-dimensional phase space. As we have introduced it, the Liouville equation describes the exact orbit of a single point in 6N,-dimensional phase space. An example is shown in Fig. 4.1, which is a projection of the orbit onto three of the 6N, dimensions. As the indi- vidual particles of the system move about in six-dimensional space, the system it- self moves along a continuous orbit in 6No-dimensional phase space. Suppose that we have an ensemble of such systems, prepared at time fg. At any later time = to, we define Inve LVL X2¥20 os Xn Vug Ody dxzdv, .. . diy, d¥y, to be the probability that a particular system is at the point (x,,¥,, . . . . Xx,s¥.) iD 6N,-dimensional phase space, that is, the probability that X,(4) lies between x, and x, + dx,, and V,(1) lies between v, and v, + dv, and X,(1) lies between x, and x, + dx,, and etc. Since fy, is a probability density, its integral over all 6Ny dimensions must be unity. Each system in the ensemble moves along an orbit like that shown in Fig. 4.1 We can think of this orbit as carrying its ‘piece’ of probability along with it. A large probability for point A in Fig. 4,1 at time ¢, implies a large probability for Point B at time f. In other words, we can think of the probability density as a fluid moving in the 6N,-dimensional phase space. Each element in the probability fluid moves along an exact orbit as given by the solution of the Liouville equation (4.8). Since each element of probability fluid moves along a continuous orbit, and since probability is neither created nor destroyed, the probability fluid must satisfy a ny Fig. 4.1 A projection onto three dimensions of a typical system orbit in 6 N,-dimensional phase space. BBGKY Hlerarchy a9 continuity equation in 6No-dimensional phase space of the form (4.14). Thus, fy, must satisfy = ‘ + 3 Vu OS) + > Wife) = 0 4.15) where V(t) is, as usual, calculated from the Lorentz force equation (4.7) and the fields E” and B” are the exact fields appropriate to the system that occupies this particular point in 6N,-dimensional phase space. We shall only be concerned with smooth functions fy,. Thus, we might think of a drop of ink placed in a glass of water. The initial drop contains all those systems that have a finite probability of being represented in the ensemble of systems at time fy. Ignoring diffusion, the drop may lengthen, contract, distort, squeeze, break into pieces, deform, etc., as time progresses. However, the total volume of ink is always constant; the total probability is always unity. The convection of the probability ink is expressed mathematically by reversing the steps that led from the Liouville equation (4.8) to the continuity equation (4.14). (See Problem 4.1.) Equation (4.15) becomes ths. + > wi Vite FL WV Se = 0 (4.16) which by (4.10) is Din _ > 7° (4.17) Equation (4.16) is the Liouville equation for the probability density fy,. Thus, the density of the probability ink is a constant provided that we move with the ink. The probability density /,,, is incompres 4.3 BBGKY HIERARCHY As discussed above, the density fy, represents the joint probability density that particle 1 has coordinates between (x,,¥,) and (x, + dx,,v, + dv,) and particle 2 has coordinates between (x2,¥2) and (x, + dx,,v. + dv,), and etc. We may also consider reduced probability distributions CMM MD EVEL aeadves Oendrntnt 418) which give the joint probability of particles 1 through k having the coordinates (%1,¥,) to (x, + dxisyy + dv,) and... and (Xq,¥,) to (xX, + dXi,¥y + dv,), irte- spective of the coordinates of particles k + 1,k + 2,..., No. The factor ¥* on the right of (4.18) is a normalization factor, where. is the ae spatial volume in which fy, is nonzero for all x,,X2, .. . , Xy,(Fig. 4.3). At the end of our theoretical development, we will take the limit Ng — %, V — ¢, in such a way that % = N/V is aconstant giving the average number of particles per unit real space. For the present, we assume that fy, + Oasx, — tory, — te°orz; — tee for any 50 Piasma Kinetic Theory UI 4% Fig. 4.2 Finite spatial volumg/ in which fy, is nonzero for any x,,f = I,..., No. i. Likewise, because there are no particles with infinite speed, fy, — Oas v,, ~ +e or v,, — £2 or v,, — £2 for any i. In this development, we do not care which one of the No particles is called particle number 1, etc. Thus, we always choose probability densities fy, that are completely symmetric with respect to the particle labels. For example, Fine ( z= 2cm...z,; =Scem...2) = fy 2, =5em...2; =2em...4 (4.19) provided all of the other independent variables are the same. Here, we must interchange all of the / = 7 variables with all of the / = 13 variables. This means that when we set k = 1 in (4,18), the function f,(x,.¥,,0 is (to within a normaliza- tion constant) the number of particles per unit real space per unit velocity space. Thus, this function f,(x,,¥,,1) has the same meaning (to within a normalization constant) as the function f,(x,v,¢) introduced in the previous chapter in connection with the plasma kinetic equation. To keep the theory as simple as possible, we shall ignore any external electric and magnetic fields. We shall deal with only one species of No particles; it is easy enough to generalize the results to a plasma with two species of N,/2 particles each at the end of the development. For some purposes, such as calculating electron- electron collisional effects, the second species can be introduced as a smeared-out ion background of density mo, which simply neutralizes the total electron charge. Finally, we adopt the Coulomb model, which ignores the magnetic fields produced by the charged particle motion. In this model, the acceleration Xo Va) = 3 ay (4.20) i where 2 = —4__ (x, - x) (4.21) = BBGKY Hlerarchy 51 is the acceleration of particle i due to the Coulomb electric field of particle j. Since a particle exerts no force on itself, we use (4.21) only if / # J; if i = j, we use = 0. Equation (4.21) replaces Maxwell's equations and the Lorentz force law. a The Liouville equation . 16) becomes 4 oe + Be Vx, Int 3 a,- Wf. = 0 (4.22) pl Equations for the ae distributions f, are obtained by integrating the Liou- ville equation (4.22) over all Xq1,ViersXiea+Viv2s - - » XnVnge FOr example, to obtain the equation for fy,, we integrate (4.22) over all xy, and vy,, obtaining Q ® Yn dry dng Gt + f dns a 3 8 Vin % o + faeadrr, > 3 ay Va Sue = (4.23) a A Term © is easy, since we can move the time derivative outside the integral to obtain a a O = Sf dxcdvante = VI Fo Su (424) where the definition (4.18) has been used. Term @ is also easy. In the first Ny ~ 1 terms in the sum, the integration variables are independent of the operator v,- V,; this operator can then be moved outside the integration and we again obtain a term proportional to fy,.. The last term in the sum, with / = No, is fii. dU, Ces, Ban, tas Arne Yo, Bea Min = f 40x Aa, A205, % Ful To, + 2 similar terms =0 (4.25) since fy, vanishes at the boundaries of the system that have been placed at xy, = te, etc. Thus, ia @="v 3s V5 Va Snot (4.26) 1 Term @ is not much harder. Splitting the double sum. mM Net Met Not Ne DUe= dh wat 2 Bras + : Bins + Beare Am whe we get 52 Plasma Kinetic Theory i! ne ' ¥ Day Ve Foe a pl 3 1 Ng + faruden Sans + Voy, Srs fl i + fdxndvne 3 anys Vo So (427) =" where thei = Ny,j = Ny term has been discarded because ay,x, = 0. The second term on the right vanishes after direct integration with respect to dv,,, and evalua- tion at v,,, = +, ctc. The remaining terms in ®, ®, and @, after multiplication by Vs, are ® ® ® 7 Net Mgt Sp for tL wey fae FD ay V fives ar el el fl ® + PY fden.dvny ain,’ Ve fun = 0 (428) a This is the desired equation for fy.-1. Notice that it does not depend only on fx.-1; the last term @ depends on fy,. We have made no approximations in deriving (4.28); within the Coulomb model, it is exact, Having succeeded in deriving the equation for fy,.;, Jet us proceed to derive the equation for fy,,. To do this, we integrate (4.28) over all xy,-, and over all Vy... AS in (4.24), term @ yields V 8, fyyer- EXERCISE Use the definition (4.18) to explain the difference between the power of V encountered here and that encountered in (4.24). As in (4.26), term ® yields one term that vanishes upon integration, leaving a sum from LtoNo ~ 2. In term @, we do as in (4.27); we split the double (Ny — 1)sum into a double (No — 2) sum plus two singte (N, — 2) sums, the i = No — 1, J = No — 1 term vanishing since ay.-1,x,1 = 0. Term @ becomes Ned Neo Dav YY ay Va Sues mF Not + farwrdvncr S ayner* VoSoer Ne + f dvr ver 3 anes? Vag Suet (4.29) a The last term on the right vanishes upon direct integration with respect to Vj,-1- BBGKY Hierarchy 53 For term @ we have ® Ml wo Jf Pe: Avner My, By, Big” Va Sn = HOS fan arn ain V, fein dnrtn (430) a where the N, — 1 term in the sum vanishes upon doing the dvy,.; integration. The variables (Xy,.Vy,) and (X,,-1-Vvy-s) are simply dummy variables of integration on the far right of (4.30). Therefore, we can switch the labels Ny and Ny — 1, so that ay, becomes a,,). The density fy, can stay the same, however, because it has the symmetry property (4.19). Equation (4.30) becomes @ = VS febiyes dng tines Vy, f dn defi a ; VM yor Y faders dns diacr *VoSues @3n > u which is identical with the middle term on the right of term @ in (4.29). Collecting all of the terms in @, ®, @, and @ and dividing by V, we obtain Need Ned Stat Bw Vyfuer + Say Vases el fl + 25 fan V1 Be ot a This equation for fy,-2 is quite similar in structure to (4.28) for fy,-:. Notice again that this equation does not involve only fy,-2, but also involves fy,., in the last term on the left. By comparing (4.28) and (4.32), we see a pattern emerging. Using the same manipulations that we have been using (see Problem 4.2), we can generate an equation similar to (4.28) and (4.32) for arbitrary k. This equation is vIn = 0 (4.32) 2K da wht S Sa: Wife fl (4.33) (No — & + mT 3 Jari dren 810s Ve fin = 0 fork = 1,2,...,No — 2. This is the BBGKY hierarchy (Bogoliubov [1]; Born and Green (2); Kirkwood (3, 4]; and Yvon (5]). Each equation for f, is coupled to the next higher equation through the /,,, term. EXERCISE Verify that (4.22) for fy, and (4.32) for fyy-z are in agreement with (4.33), Verify that (4.28) for fy,.. is in agreement with (4.33), provided that fy, is replaced by V% fy, in (4.33) [see (4.18)]. 54 Plasma Kinetic Theory Il As it stands, the BBGKY hierarchy (4.33) is still exact (within the Coulomb model) and is just as hard to solve as the original Liouville equation (4.22). It consists of Ny coupled integro-differential equations. Progress will come only when we take just the first few equations, fork = 1,k = 2,etc.,and then use an approximation to close the set and cut off the dependence on higher equations. From (4.33) the kK = 1 equation is ACM Fv Va fi NoW~1 a 7 fdr dra aia Vy, AM okaeto) =0 (4.34) This is coupled to the k = 2 equation through /,. One way to proceed is to find some approximation for fy in terms of f,. If we can do this, then (4.34) will be written entirely in terms of f,, and we will have a complete description of the time evolution of f,(x1,¥;,f) given the initial value f,(x,,v,,f = 0). This is a good point at which to repeat our interpretation of the functions Ail%,¥;,0) and fy(%,,¥15Xp,¥250). We have said before that f, is equivalent to f, in the plasma kinetic equation; when multiplied by y= No/V, it is the ensemble averaged number of particles per unit real space per unit velocity space at the point (x,,¥,) in six-dimensional phase space. + EXERCISE Use the definition to show that f dv, fi(x;,¥1,1) = 1 provided that none of the functions f,, k = 1,2,..., Ng depend upon the positions x;,X2, +) XN We may also say that /;(x,,v,,4)d@x, dv, is the probability that a given particle finds itself in the region of phase space between (x,,¥;) and (x, + dx;,v, + dv,). The interpretation of f; js similar to the interpretation of f,. The function f, is the ensemble averaged number of particles per unit x, real space per unit x, real space per unit v, velocity space per unit v, velocity space. We may also say that FA(%,¥1,%2,¥o,t) is proportional to the joint probability that particle | finds itself at (%4,¥;) and particle 2 finds itself at (x;,V,). Since in this discussion all particles are of the same species, we know that an exact expression for f, would include the fact that no two particles (electrons, for example) can occupy the same Spatial location. Thus, an exact expression for f, must have the property that f; ~ 0 asx, + %, regardless of the values of v, and v2. In developing an approximate expression for Fz, we could of course lose this property. Another property that f, should have is symmetry with respect to the particle labels: f3(x;,¥1,%2.Vast) = SaO%2,VoXuVist)- This symmetry occurs because the original fy, has such symmetry, by assumption. It turns out that f; has an intimate relation to f,, which can be seen by an elementary example from probability theory. Suppose we have two loaded dice, each of which always rolls a five. Then the probability distribution for the value of the throws of either die is P(x) = &x — 5) (4.35) The joint probability that the value of the first die will be x and the value of the second die will be y is P,(x,y) = (x — 5)6(y — 5) (4.36) BBGKY Hierarchy 55 But by (4.35) this is just PAxy) = P(x)Pi(y) (4.37) This separation always occurs when two quantities are statistically independent, that is, the value of one quantity does not depend on the value of the other quantity. Thus, it is always useful in considering joint probability distributions to factor out the piece that would be there if the two quantities were uncorrelated. Thus, for the dice we have Pi(xy) = Py(x)P((y) + SP(x,y) (4.38) where P(x,y) = 0 by (4.37). For a plasma, we define the correlation function B(X1,VLX2,¥250) BY LGV) = ACV DAO V290) + g(¥1,¥1,%2,¥2t) (4.39) This is the first step in the Mayer [6] cluster expansion. EXERCISE From the definitions of fy, and f,, convince yourself that f; has the same units as f; f,. We are ready to insert the.form (4.39) into the equation (4.34) for f,, which becomes ASX Viet) +H VS + ng f drs diy ay + VAG ns DAG!) + 8(X,,¥1.%2,¥2,0)] = 0 (4.40) where we have replaced (Ny — 1)/V by no because we are interested only in systems with N, >> 1. Suppose one assumes that the correlation function vanishes. That is, we assume that the particles in the plasma behave as if they were completely independent of the particular positions and velocities of the other particles. This assumption would be exactly valid if we performed the pulverization procedure discussed in the previous chapter, in which my + %,e — 0,m, —- 0,A — ©, H@ = constant, e/m, = constant, v, = constant, w, = constant, and A, = constant. Then each particle would have zero charge, and its presence would not affect any other particle. Collective effects could of course still happen, as these involve only f, and not g. When we set g equal to zero, (4.40) becomes af TW Whi + (to f dy dvs aye Fira Has)] “Vo, Kivut) = 0 (aly But the quantity in brackets is just the acceleration a,, produced on particle | by particle 2, integrated over the probability distribution f,(x2,V2,f) of particle 2. This is the ensemble averaged acceleration experienced by particle 1 due to all other particles, (4,4) = mg f dy dy 913 f0.¥2) (4.42) EXERCISE Convince yourself that a is normalized correctly. 56 Plasme Kinetic Theory 1 Then (4.41) becomes ahtue Veh ta A which we recognize as our old friend the Vlasov equation. The Vlasov equation is probably the most useful equation in plasma physics, and a large portion of this book is devoted to its study. For our present purposes, however, it is not enough, It does not include the collisional effects that are represented by the two-particle correlation function g. We would like to have at least an approxjmate equation that does include collisional effects and that, there- fore, predicts the temporal evolution of f, due to collisions. We must therefore return to the exact k = 1 equation (4.40) and find some method to evaluate g. Since g is defined through (4.39) as g = f, ~ fi fi, we must go back to the k& = 2 equation in the BBGKY hierarchy in order to obtain an equation for fy and, hence, for g. Setting k = 2 in (4.33) and using (Ny — 2)/V = mo, one has ® ® @ Of + (Wi Va, + a VA + (an Ve, + ar Wh (4.43) @ + no} dxsdvj(aj, °° Vy, + an WA = 0 (4.44) We have seen that it is useful to factor out the part, f, offi = ff, + g, which exists when the particles are uncorrelated. Likewise, it is useful to factor from f, the part that would exist when the particles are uncorrelated, plus those parts that result from tWo-particle correlations. This leads to the next step in the Mayer cluster expansion, which is A123) = ADAADLE) + fi(g23) + F(2)g3) + A3)g(12) + 4023) (4.45) where we have introduced a simplified notation: (1) = (X1,¥;), (2) = (X),¥2), and (3) = (x,,¥,). Equation (4.45) will be explored further in Problems 4.4 and 4.5. Our procedure is to insert (4.45) into (4.44) and neglect (123). This means that we neglect three-particle correlations, or three-body collisions. It turns out that these correlations are of higher order in the plasma parameter A; therefore their neglect is quite well justified for many purposes. The resulting set of equations constitute two equations in two unknowns f, and g. Thus, we have truncated the BBGKY hierarchy while retaining the effects of collisions to a good approxima- ton. Inserting (4.45) for f; and fr = f, f; + g into the k = 2 BBGKY equation (4.44), we find for the numbered terms: Oo. © ; @ = ADAG + /(DAW + 812) ® @ Fw Vs AML + vy Vs. g2) + (1 ~ 2) BBGRY Hlerarchy 57 12 VACA) + aa? Vy, g2) + 1 = 3 © @= no f a3 as Vy, LADAQAAG) + A()g23) @ + f,(2)g(13) + fiB)g(12)} + (1 = 2} (4.46) where d3 = dx,dy, and {1 ~ 2} means that all of the preceding terms on the right side are repeated with the symbols | and 2 interchanged. Recall that g(12) = g(21) by the symmetry of f,. Many of the terms in (4.46) can be eliminated using the k = 1 BBGKY equation (4.40). For example, O+O+O+O= fA + 41° % AO) + np f a3 a5 + %, LADS + UMA = [left side of (4.40)]f,(2) = 0 (4.47) Term © likewise combines with three of the {I -+ 2] terms to vanish, leaving 2012) + (y+ Va, + va > Vy, e(12) = ~ nV, + ans WADAQ) + 20201 ~ {gf d3 y+ Vs, L239) + HU + {1 Bl 48) Together with (4.40) which in the condensed notation reads AQ) +u- yA) + ry f a2 a “Vy, UDA) + g(12) = 0 (4.49) we have two equations in the two unknowns f, and g. We have truncated the BBGKY hierarchy by ignoring three-particle correlations. In practice, (4.48) and (4.49) are impossibly difficult to solve, either analytically or numerically. They are two coupled nonlinear integro-differentia] equations ina twelve-dimensional phase space. The present thrust of plasma kinetic theory con- sists in finding certain approximations to g(12) that are then inserted in (4.49). Using the definition of the acceleration a in (4.42), we rewrite (4.49) as A) + VS + LW Si = A of @2 ain + Vy, #2) (4.50) which is in exactly the same form as the plasma kinetic equation (3.26). Most of the discussion in this chapter has been exact, in particular, the deriva- tion of the Liouville equation and the BBGKY hierarchy. Even the approxima- tions that lead to (4.48) and (4.49) are extremely good ones, for example, ! << No and the neglect of three-particle collisions. By contrast, the approximations needed to convert (4.48) and (4.49) into manageable form are sometimes quite drastic and 58 Plasma Kinetic Theory iI less justifiable, as will be seen in the next chapter. Further discussion of the Liouville equation and the BBGKY hierarchy can be found in the books of Mont- gomery and Tidman [7], Montgomery [8], Clemmow and Dougherty (9), Krall and Trivelpiece [10], and Klimontovich [11]. REFERENCES [1] N,N. Bogoliubov, Problems of a Dynamical Theary in Statistical Physics, State Technical Press, Moscow, 1946. [2] M. Born and H. S. Green, A General Kinetic Theory of Liquids, Cambridge University Press, Cambridge, England, 1949. [3) J. G. Kirkwood, J. Chem. Phys., 14, 180 (1946). [4] J. G. Kirkwood, J. Chem. Phys. 15, 72 (1947) (5) J. Yvon, La Theorie des Fluides et I'Equation d'Etat, Hermann et Cie, Paris, 1935, {6] J. E. Mayer and M. G. Mayer, Statistical Mechanics, Wiley, New York, 1940. [7] D. C. Montgomery and D. A. Tidman, Plasma Kinetic Theory, McGraw- Hill, New York, 1964. {8} D. C. Montgomery, Theory of the Unmagnetized Plasma, Gordon and Breach, New York, 1971. (91 P.C. Clemmow and J. P. Dougherty, Electrodynamics of Particles and Plas- mas, Addison-Wesley, Reading, Mass., 1969. [10] N. A. Krall and A, W. Trivelpiece, Principles of Plasma Physics, McGraw- Hill, New York, 1973. ® (11) Yu. L. Klimontovich, The Statistical Theory of Non-equilibrium Processes in a Plasma, M.1.T. Press, Cambridge, Mass., 1967. PROBLEMS 4.1 Continuity vs. Convective Demonstrate the equivalence between the convective derivative form of the Liou- ville equation (4.16) and the continuity equation (4.15). 4.2 BBGKY Hierarchy Integrate (4.32) over all xy,-7 and ¥y,-2 to obtain the k = Ny — 3 equation of the BBGKY hierarchy, and compare your result to (4.33). 4.3, Normalization Explain in detail the normalization of (4.42). Problems. 59 4.4 Three-Point Correlations (Coins) In (4.45) we define a three-point joint probability function f, in terms of the one-point probability f;, the two-point correlation function g, and the three-point correlation function A. Suppose we apply this kind of thinking to the case of three coins, each of which can come up heads (+) or tails (—}. What is the meaning of f, in this case? Write out f, in the form (4.45), and evaluate f;,/;, g, and hin each of the following cases, (a) All three coins are “honest,” that is, each coin is equally likely to come up heads or tails, and each coin is unaffected by any other coin. (b) Because the coins are mysteriously locked together, in any one throw all three are heads or tails, the result changing randomly from throw to throw. (c) All three coins afways come up tails (dy The first two coins always come up heads, while the third is honest. Note that here the probability functions are not symmetric, so that, for example, J,(1) is not the same function as /,(3). 4.5 Three-Point Correlations (Dice) In (4.45) we define a three-point joint probability function f; in terms of the one-point probability f;, the two-point correlation function g, and the three-point correlation functidWA. Suppose we apply this kind of thinking to the case of three dice, each of which can take on integer values from one through six. What is the meaning of f, in this case? Write out f; in the form (4.45), and evaluate f;, f,, g, and A in each of the following cases. (a) All three dice are “honest,” that is, the value of each die is equally likely one through six and is independent of the value of any other die. (b) Because the dice are mysteriously locked together, in one throw all three always show the same value, the value changing randomly from throw to throw with all six values equally likely. (c) All of the dice always come up “five.” (d) The first two dice always come up “two”; the other one is “honest.” 4.6 BBGKY Hierarchy In this chapter, we derive the BBGK Y hierarchy from the Liouville equation. This can be done in a completely different way [10], starting with the Klimontovich equation. Explain, by using words and writing equations only for illustration, how the k = 1 and& = 2 equations of the BBGKY hierarchy can be obtained from the Klimontovich equation CHAPTER Plasma Kinetic Theory III: Lenard-Balescu Equation 5.1 BOGOLIUBOV'S HYPOTHESIS In the preceding chapter, the BBGKY hierarchy is truncated by neglecting three- particle correlations (three-body collisions). For a good plasma, this is probably a very good approximation, although no rigorous proof exists. The spirit of the approximation is the same as that of Section 1.6, where the collision frequency is calculated as a series of two-body collisions, even though the particle is interacting with A particles simultaneously. Since the collision of particle A with particle B is usually a small angle collision, its effect on the orbit of particle A is small, thus making a negligible effect on the simultaneous collision of particle 4 with particle C. The result of our truncation of the BBGKY hierarchy is the set of coupled equations (4.48) and (4.50) in the two unknowns f,(x,,¥).2) and g(X;,V,,%2s¥25f). These equations are quite intractable in general. However, there is one set of simplifying assumptions that is both physically very important and allows the exact (almost) solution of (4.48) and (4.50). Consider a spatially homogeneous ensemble of plasmas. This means that any function of one spatial variable must be independent of that variable; so Sk%¥.0) = fi(v.0) and a(x,,t) = a() = 0 by (4.21) and (4.42), Any ensemble averaged function of two spatial variables can only be a function of the difference between those variables; therefore we write g = g(x, — X2,¥1,¥,f). With these assumptions, (4.50) simplifies considerably and becomes 8, A(V11) = — rm f dx, dv, ayy > Vy, B(K) — X2.¥i,Vat) (5.1) Bogoliubov’s Hypothes!s 61 Equation 4.48 simplifies since two terms are of the form Uma f a3 ais fi) Vs, (12) = a Vy, (12) = 0 (5.2) leaving 8) (Xt — Xa VEVat) + V+ Vx, B(12) + vo + Vy, (12) + (a+ Vy, + a+ Ve@C12) + maf d3 ars + Vy, F(DE23) + 9 f d3 any Vs, FDU) = — (a2 Vy, + an WAM A(2) (5.3) ‘We now wish to argue that the fourth term on the left is smaller than all the other terms and can be discarded. Recall the pulverization procedure of the previous chapter. By that argument, as well as the discussion of collisions in Section 1.6, we argue that the two-point correlation function g is higher order in the plasma parameter A than /,; thus g/f, ~ A”!. The acceleration aj. ~ e?/m, ~ A™'since e/m, is constant and e ~ np! ~ A‘; here, we phrase our discussion in terms of electrons. Thus, all terms in (5.3) are ~ A” except for the fourth term on the left, which is ~ A’?, We discard this term, leaving $e) +g tig =S (5.4) where V, and V, are operators defined by Vig(12) = v, + Vx, g(12) + bre f dda 8291 VLD (6) Vig(12) = + Vs, $12) + (mf Ban, 09] 42) 67) and the source function S is SQ > VV.) = — (an Vo, + an’ Vy.) AC) A(2) (5.8) In this chapter we alternate between the notations (1) and (x,,v,) depending on convenience. For simplicity, we suppose that we are dealing with an electron plasma. A neutralizing ion background can be thought to be present; it is consid- ered to be smoothed out so that it does not contribute explicitly to the acceleration a,,, which by (4.21) is 2 ay = ——— 7 (% - , 5.9) = Ga, = xP (x, — x) 6.9) The important physical situation to which this discussion applies is as follows. Imagine a beam of electrons incident on a Maxwellian electron plasma in the 62 Plasma Kinetic Theory II! &-direction. Then the function Flo) = f dv, do, flv) (5.10) has the form shown in Fig. 5.1. Ignoring questions of stability (see Chapter 6), we recognize that the beam of electrons represented by the bump at large positive v, will experience collisions that will eventually (t — °) produce a new Maxwellian at a higher temperature. By the discussion of Section 1.6 we can predict the time scale for this process to be ~ v,,~ w,/A. The solution of (5.1) and (5.4) which we are about to obtain should yield a very good theoretical description for this important process. This evolution is encountered in such applications as electron beam-pellet fusion and (when generalized to ions) ohmic heating of tokamaks. The further assumption that allows us to solve the (still very complicated) set of equations (5.1) and (5.4) is Bogoliubov’s hypothesis. The assumption is that the two-point correlation function g relaxes on a time scale very short compared to the time scale on which f; relaxes [1]. Imagine introducing a test electron into a plasma. The other electrons will adjust to the presence of the test electron in roughly the time it takes for them to have a collision with it. With a typical speed v, and a typical length A,, the time for a collision is ~ \,/v, ~ w,'. By contrast, the time for f, to change because of collisions is ~ Aw, 7; thus it is indeed quite reasonable to assume that g relaxes quickly compared to f,. Mathematically, we incorporate this assumption by ignoring the time dependence of f,(v,,t) and F,(¥z,t) in the source function S on the right of (5.4). Equation (5.4) is then a Jinear equation for g with a known, constant (in time) source function on the right. We can solve such a linear equation for g(x, — X),¥,,¥2,¢ — °°) where f + 0° is understood to refer to the short time scale on which g relaxes. The solution for g will then depend on the factors f,(v,,1) and f,(v,,/) in the source function (5.8). When this solution for g is substituted into the right side of (5.1), there results a single nonlinear integro-differential equation in the one unknown function f,. We Fig. 5.1. Distribution F(v,) defined in (5.10) for an electron beam incident on a plasma. Lenard-Balescu Equation 63 have finally achieved our goal of truncating the BBGKY hierarchy and have expressed the entire plasma kinetic equation (5.1) in terms of the one unknown function f,(¥,,0). The implementation of this procedure is straightforward but complicated. In order to understand it, it is useful to have first studied the material in Chapter 6 on the Vlasov equation. Thus, we will not perform the derivation here; it is included in Appendix A, The reader who is studying plasma physics for the first time may wish to accept the results as given here, and proceed to read Appendix A after a thorough study of Chapter 6. The solution of (5.1) and (5.4) uses the techniques of Fourier transformation in space, Laplace transformation in time, and their inverses. The conventions used in this book are as follows: a Sf (k) =| ane ees f (x) oD f(x) = f ae e™ f(k) (5.12) S(w) =f dt el" f(0) (5.13) 0 Ko =f ese) (5.14) 1, Ore where the integrals over x, k, and f are usually along the real axes while the integral over w is along the Laplace inversion contour to be discussed later. Expressed in terms of the difference variable x = x, — x, the acceleration a,, in (5.9) is 7 5.19 =—; a a(x) mix * (5.15) with Fourier transform ~ik ay(k) = ok) (5.16) where ce 5.17) AK) = Tage 6. is the Fourier transform of the Coulomb potential e Herts (5.18) ell (See Problem 5.1.) Then, as shown in Appendix A, the solution of (5.1) and (5.4), under the Bogoliubov hypothesis, is (hk) af(vst) _ _ Bim ance oe m2 w Jax av % le(k,k + v)I? or £ > X 8k + (¥ — VIL) Vy SY) — SOV LM] (5.19) 64 Plasma Kinetic Theory tI! which is the Lenard-Balescu equation (Rets. [2] to [6]). In this equation, we have dropped the subscript 1 from v,, and the subscript | from f,, and have used the dielectric function k-Y, Ret fore “2 which will be studied in detail] in the next chapter. The velocity integral must be performed along the Landau contour, as discussed in the next chapter. The inter- pretation of the Lenard-Balescu equation (5.19), and several alternate forms, will be discussed in the next section. (ko) = 1 + 5.2 LENARD-BALESCU EQUATION The Lenard-Balescu equation (5.19) is obtained from the BBGKY hierarchy after several assumptions: three-particle correlations are negligible, the ensemble of plasmas is spatially homogeneous, and the two-particle correlation function g relaxes much faster than the one-particle distribution function /;. Thus, the Len- ard-Balescu equation is applicable to situations such as the collisional relaxation of a beam in a plasma, but is not applicable in general to the collisional damping of spatially inhomogeneous wave motion or any phenomena that involve high frequencies like w,. The right side of (5.19) represents the physics of two-particle collisions, since the right side of (5.1) is proportional to the two-particle correlation function g. This is indicated by the factor ¢(k)/e(k,k * v), which appears squared, It will be shown in the next chapter that the dielectric function e(k,w) represents the plasma shielding of the field of a test charge. Thus, this term in (5,19) represents the interaction of one particle (together with its shielding cloud) with the potential field of another particle (together with its shielding cloud), that is, the collision of two shielded particles. There is a problem with the Lenard-Balescu equation (5.19) as it stands. If one converts the & integration into spherical coordinates, and takes into account the forms (5.17) of g(k) and (5.20) of e(k,w), one finds that at large k the integral diverges like f dk/k ~ In k. Thus, just as in the derivation of the collision frequen- cy in Section 1.6, we find a logarithmic divergence at large k, or small distances. In Section 1.6 we cut off the spatial integral at the lower limit po, where py is the impact-parameter for large angle collisions, It is argued in Section 1.6 that the physicai formulation is not valid for large angle collisions, thus producing an unphysical divergence at short distances. The same thing is going on here. The derivation of the Lenard—Balescu equation is based on the assumption that in the expression S12) = AG) fi) + e(12) (5.21) we have |g| << |f, fil. This assumption led us to discard a term in (5.3) to obtain (5.4). However, this assumption is not always valid. It is not possible for two electrons to get very close to each other; therefore, we must have f, — 0 as X, —* xz, which implies g = — f; f,. Thus, for small values of |x, — x,| (large k), it is not correct to assume |g] << | f; f;|. In practice, since the divergence is Bogollubov's Hypothesis 65 logarithmic, we can simply cut off the integral in (5.19) at some upper limit wave number corresponding to some lower limit spatial scale. For this purpose, the impact parameter (Landau length) po for large angle collisions (see Section 1.6) would be a reasonable choice. The Lenard-Balescu equation (5.19) has several desirable features [4-5]. These are: (a) Iff2O0att =0,f2 Oatallr, (b) ‘Particles are conserved: d/dt f dv f(v,t) (c) Momentum is conserved: d/dt f dvv f(v,) = (a) Kinetic energy is conserved: d/dt f dv v? f(v,t) = 0. (&) Any Maxwellian is a time-independent solution. (f) Ast — &, any f satisfying (a) approaches a Maxwellian. A simplified but fairly accurate form of the Lenard-Balescu equation (5.19) can be obtained as follows. We rewrite (5.19) in the form POD = — v fav B ww — WSOFW) (5.22) with the tensor aw) __ Brn f kkg*(k) Tack we OE — VI il m2 (5.23) where the definition (5.17) has been used, and where the dimensionless function y is found from (5.20) to be Wkk+v) = vefav ees (5.24) Again, the velocity integral must be performed along the Landau contour, as discussed in the next chapter. The wave number integral in (5.23) is performed as follows. When we orient the &, axis in the v — v’ direction, the Q,, component of the tensor Q is Qilvv') ce ae 1 6k) a ae fat, ak, dk, “EE Wow) TF WRAP The factor 6(k,) implies Q;; = 0 if either i = 1 or j = 1, The k, integration is trivially performed using this factor. In cylindrical coordinates with k, = k cos 6, k, = k sin 6, and cutting off the integration at an upper wave number ky = po, we find, using Q,, as an example, eee 2nge* 1 Ostee) = — aaa Leslee 9 Since wis a function of @ but not of k [sec (5.24)], the wave number integration can be performed (Problem 5.3). The result is 66 Plasma Kinetic Theory II! nge* ‘Qn Im[yin (1 + k@A2/ Onlv.v') = — aoa f d ae ee sin?@ (5.27) It turns out (as can be seen more clearly after a study of the following chapter) that the dimensionless function y is of order unity. In addition, we recognize the factor kod, = 2,/po from Section 1.6 to be (within factors of order unity) the plasma parameter A. Thus, we neglect unity compared to ky?A,?/y, and In (ys) compared to In (k,?A2) © In (A?) = 2 In A, to obtain On(vv') = Onlvwv) = — Similar arguments yield Q.; = Qu = 0. A tensor with only the Q2. and Qs, components nonzero can be a expressed in terms of the unit tensor T= Kk, + kk, + kyks; with g = v — v' and recalling that k, = 8, we have 4 fence A g T- ge (5.29) This expression is known as the Landau form tor 3 With @ in the form (5.29), it is possible to put the Lenard—Balescu equation (5.22) in the form of a Fokker-Planck equation. The general Fokker-Planck equa- tion is a very important equation in all aspects of statistical physics, and is derived from first principles in Appendix B. Following Montgomery and Tidman [5], we notice that 2arnye! mily — vl Ina (5.28) Bev) = — VY Vg = fla (5.30) so that with an integration by parts (5.22) becomes afte) = EBEIK 9, 0, f+ Wal df) = 0) f av V0, Woe FO) __ 2rme* in 1: LOW dv eI me = 2-6 f dv WO WE SN} s 2rnge! Ind A — 4v,- [rweefav SW fe) m, + WY: F099, 9, f av gf} (8.31) where in the first step we have used Vig = — V,g, and in the third step we have used (V,- V,)¢ = 2/g. This is in the standard form of a Fokker-Planck equation (see Appendix B), V,Vy: (BSC) (5.32) Vy TA S(Y)) + Lenard-Balescu Equation 67 where the coefficient of dynamic friction _ Bane! In A » £1) Aca) = vif av eae (5.33) and the diffusion coefficient 4rnget In A creep ata Biv.9 = vVfav lv — visW0 (5.34) With the coefficients (5.33) and (5.34), Eq. (5.32) is known as the Landau form of the Fokker-Planck equation. The meaning of the terms in the Fokker-Planck equation is discussed in Ap- pendix B. The coefficient of dynamic friction A represents the slowing down of a typical particle because of many small angle collisions. The diffusion coefficient represents the increase of a typical particle's velocity (in the direction perpendicu- lar to its instantaneous velocity) because of many small angle collisions. Thus, the two terms on the right side of the Fokker—Planck equation (5.32) tend to balance each other. They are in perfect balance when f is a Maxwellian, as shown in Problem 5.5. The Landau form of the Fokker-Planck equation (5.32) has been solved nu- merically by MacDonald et al. [7] (Fig. 5.2). The initial distribution function f(v,t = 0) = f(Iv|,f = 0) is spherically symmetric in velocity space. Figure $.2 shows the steady progression of the distribution, as time increases, toward a Max- wellian. At late times, there is an overshoot at low speeds, which indicates that it Dimensionless velocity distribution Dimensionless velocity Fig. 5.2 Time evolution of a spherically symmetric électron distribution function as ob- tained from a numerical solution of the Landau form of the Fokker-Planck equation (5.32) by MacDonald et al. [7]. 68 Plasma Kinetic Theory II : takes a long time to populate the high speed tail of the Maxwellian. (Remember that Coulomb collisions become quite weak for fast particles.) There exist even simpler forms of the Fokker-Planck equation [4] but these are not too accurate and are used only to get a rough idea of collisional effects. One is Fave — wr t 2a) (535) where vy is a collision frequency, and vy is a constant velocity. An even cruder model, which is not related to the development of the present chapter, is the Krook model, La-- wf) (5.36) where f, is the appropriate Maxwellian distribution. Equation (5.36) is also called the BGK equation, after Bhatnagar, Gross, and Krook [8]. This brings us to the end of our study of plasma kinetic theory including the effects of two-body collisions. The materia] in this chapter can be truly appreciated only after a careful study of Appendices A and B. However, Appendix A itself can best be understood after one has mastered the treatment of the Vlasov equation, to which we turn our attention in the next chapter. REFERENCES [1] N.N, Bogoliubov, Problems of a Dynamical Theory in Statistical Physics, State Technical Press, Moscow, 1946. [2] A. Lenard, Ann. Phys. (New York), 10, 390 (1960). [3] R. Balescu, Phys. Fluids, 3, 52 (1960). {4} P. C. Clemmow and J. P. Dougherty, Electrodynamics of Particles and Plas- mas, Addison-Wesley, Reading, Mass., 1969. [5] D.C. Montgomery and D. A. Tidman, Plasma Kinetic Theory, McGraw-Hill, New York, 1964. [6] D.C. Montgomery, Theory of the Unmagnetized Plasma, Gordon and Breach, New York, 1971. [7] W. M. MacDonald, M. N. Rosenbluth, and W. Chuck, Phys. Rev., 107, 350 (1957). {8} P. L. Bhatnagar, E. P. Gross, and M. Krook, Phys. Rev., 94, 511 (1954). [9] M. Rosenbluth, W. M. MacDonald, and D. L. Judd, Phys. Rev., 107, 1 (1957). PROBLEMS $.1 ) Fourier Transforms Find the Fourier transforms (5.16) and (5.17). (Hint: Use spherical polar coordi- nates with k «x = kr cos 6.) Problems 69 ox $2 L) Lenard-Balescu Equation CS After referring to Clemmow and Dougherty [4], and Montgomery and Tidman (5], sketch the proofs of properties (a) to (f) of the Lenard-Balescu equation as listed below (5.21). 5.3 An Integral With the help of a table of integrals, perform the integration in (5.26). 5.4 Simpler Derivation of the Landau Form The development of the Landau form for G, from (5.23) to (5.28), is the standard one. However, a simpler one exists. In (5.23), replace ¢ by unity, and cut off the wave number integration at a lower wave number A,' as well as at the upper wave number py"!. Show that (5.28) resuits, The replacement of ¢ by unity is equivalent to ignoring the shielding, as can be seen in (5.20). aa :5.5 /Maxwellian Show that a Maxwellian is an exact time-independent solution of both the Len- ard-Balescu equation (5.19) and the Landau form of the Fokker-Pianck equation (5.32). 5.6 Two-Point Correlation Function Discuss the meaning of f; = f,f, + g. Why should g depend on f,? In particular, how would g change as we turn up the temperature of a Maxwellian? 5.7 Plasmas and Brownian Motion Discuss the analogy between collisional effects on a particle in a plasma and Brownian motion. Explain why the collisional effects can be described by a Fok- ker-Planck equation. Thus, using only words, explain how we could use the results of Section 1.6 on collisions to obtain the Fokker-Planck equation directly, without starting from Liouville + BBGKY — Lenard-Balescu — Fokker-Planck. This is actually the technique used by Rosenbluth et al. (9] 5.8 Units Check alt of the units in (5.19) to (5.36). Using crude dimensional arguments, derive the model (5.35) from the Fokker-Planck equation (5.32) and the coeffi- cients (5.33) and (5.34). CHAPTER Viasov Equation 6.1. INTRODUCTION Possibly the single most important equation in plasma physics is the Viasov equa- tion, This equation describes the evolution of the distribution function f,(x,v,1) in six-dimensional phase space. As discussed in Chapter 3, the distribution function f,(x,v,) can be thought of as the ensemble averaged number of point particles per unit six-dimensional phase space. It can also be thought of as the number of particles at any given time ¢, in a small region of the six-dimensional phase space of a single plasma, divided by the volume of the small region of six-dimensional phase space. As discussed in Chapter 3, the Viasov equation becomes exact in the limit that the number of particles A in a Debye cube becomes infinite. The Vlasov equation arises naturally from the Ktimontovich equation (Chapter 3) or from the BBGKY hierarchy (Chapter 4) when the effects of collisions are ignored. For this reason, the Vlasov equation is also called the collisionless Boltz- manu equation. By ignoring the effects of collisions from the start, we can derive the Vlasov equation as follows. Consider f,(x,v,/) as a probability density associated with an ensemble of sys- tems. This probability density can be thought of as a fluid in six-dimensional phase space. Since particles are neither created nor destroyed, this fluid must satisfy a continuity equation with the form OSV) + Vee (& orbit vy. a tf nit ar (6.1) Equilibrium Solutions 71 where d/dt|oi refers to the orbit of the fluid element at the position (x,v) in phase space. But the fluid represents the probability density of particles; therefore the orbit of the fluid element must be the same as the orbit of a particle of species s at position: x with velocity v. With this identification, we have immediately & hoe = (6.2) and ee = £ [Ban + tx Bx,0 | (63) dt lon mM, i c - where because the effects of collisions are being ignored the fields E and B are the smooth, ensemble averaged fields satisfying Maxwell’s equations (3.28), Equation (6.1) becomes 8, FOGWD + Vx (Vf) +Zy,- @+txpys]=o (6.4) ms u c ff Fh With the vector identity V (ab) = bV-a + a: Vb, we find 8, fv + ¥ (6.5) which is the Viasov equation [1). EXERCISE Verify that the two terms dropped in going from (6.4) to (6.5) in- deed vanish. When this equation, one for each species, is combined with Maxwell’s equations (3.28), we have a complete description of the behavior of a plasma. Although in principle the Vlasov equation only:applies to an ensemble of plasmas, in practice Wwe assume that, because of the large number of particles in a single plasma, the fluctuations are so small that the Vlasov equation yields good predictions for a single plasma. Since collisions have been ignored, the Vlasov equation applies only when collisional effects are unimportant. Often, this means that we are limited to phenomena with a characteristic frequency w >> v,; * w,/A. 6.2 EQUILIBRIUM SOLUTIONS For time scales short compared to a collision time »,;! ~ A w,', we are interested in finding steady-state solutions to the Vlasov equation (6.5), that is, those for which 4,/, = 0. (In this chapter, the words “equilibrium” and “steady-state” are used synonymously.) Of course, there is no guarantee that such steady-state solu- tions are stable to small perturbations. (A pencil standing on its tip is a steady- state solutign, but not a stable one.) As we look for solutions to the Vlasov equation, it is useful to interpret the left side of (6.5) as the total time derivative of f, along a particle orbit. Consider a 72 Vlasov Equation particle of species s whose orbit in six-dimensional phase space is X(t), V(1), where X(4) is the function that gives the position x in real space of the particle at time 1, and V(r) is the function that gives the position vin velocity space of the particle at time ¢. Then the total time derivative of any quantity, measured along the test particle’s orbit in phase space, is ‘Do aX(s) avi) Dect at ar a dx dv =F ae toon 98 at lowe” a : 4s fal . Ha tveV + 7 (E+ % B):Vs (6.6) where we have inserted (6.2) and (6.3). Thus, the Vlasov equation (6.5) simply says D Dr fown = 0 (6.7) Knowledge of the form (6.7) gives us one way to solve the Vlasov equation (6.5). Suppose we construct: fj out of functions C{x,¥,1) that are constants of the motion along the orbit of.a particle. Then by (6.7), a ICO) = & ob c=0 (6.8) so that the Vlasov equation (6.5) is satisfied. Thus, any distribution that is a function only ofthe constants of the motion of the individual! particle orbits is a solution of the Vlasov equation. In the present section we are interested only in equilibrium solutions that do not depend explicitly on time. Noting that the fields E and B in (6.5) can be combina- tions of externally imposed fields and self-consistent fields, we consider the follow- ing cases. CASE A: E = B = In the absence of external fields, the energy 4m,v* and momentum m,v = mM,(v,,Vy,v;) of a particle are constants of the motion. Thus, any function Ss = lve Vie) (6.9) is a solution of the time-independent Vlasov equation. This can also be seen by writing the time-independent Vlasov equation with no external fields, ve Us =0 (6.10) to which (6.9) is seen to be a solution. CASE B: E = 0, B = CONSTANT In the presence of a uniform background magnetic field, the total particle momen- tum is no longer a constant. If we choose the magnetic field in the 2-direction, then the constants of the motion are the momentum m,v, in the Z-direction and the energy 4im,u,? = Yam,(v, + v,2) in the plane perpendicular to the magnetic Electrostatic Waves 73 field. Thus, any function SI, = Flv, v,) (6.11) is an equilibrium solution to the Viasov equation in the presence of a uniform magnetic field. EXERCISE Show (Chapter 1) that v, is a constant of the motion in a uniform magnetic field, Verify by direct calculation that (6.11) is a solution of the Vlasov equation (6.5) in this case. CASE C: B = 0, E = E(x) #0 In the presence of an arbitrary electric field E(x) = — < d/dx (x) in the < direction, the particle constants of the motion are the momenta m,v, and m,v,, and the energy 4m,u. + g, v(x) associated with motion in the £-direction. Thus, Se = foe + 245 AV My yd.) (6.12) is an equilibrium distribution function: (Note that f, can also depend upon the sign of v,.) EXERCISE Verify by direct substitution that (6.12) is a solution of the Vlasov equation (6.5) in this case. In addition to these three simple cases, there are other important examples that are used. For example, in Chapter 2 we discussed the adiabatic invariants that are approximate constants of motion. Using these adiabatic invariants, one can con- struct approximate equilibrium distribution functions. Such solutions find wide applications in the study of magnetic confinement devices such as the tokamak and mirror machine. 6.3 ELECTROSTATIC WAVES One of the simplest and most instructive predictions of Vlasov theory is the existence of electrostatic waves, waves that have only an electric field with no magnetic field, and in the small amplitude limit have a time and spatial depen- dence ~ exp (ik +x — iwt) + c.c. (complex conjugate), where ki|E EXERCISE Show that for waves witb k||E, Maxwell’s equations predict no magnetic field. We begin with the simple situation of a plasma with no applied electric or magnet- ic fields. Each species has a distribution function Si = fo tha (6.13) where fo = fio(¥) is one of the equilibrium solutions discussed in the previous section, and f,,(x,v,¢) is a smal! perturbation associated with the small-amplitude waye. For each species, [ar fal) = (6.14) 74 Viesov Equation where 7ipis the average number of particles per unit configuration space. Choosing the electric field in the £-direction, and treating waves with a spatial variation in the £-direction only, the Vlasov equation is af + wah +H Ea,f, = 0 (6.15) With fia zero order quantity, and f,, and E small quantities of first order, we look for linearized solutions of (6.15). The zero order terms in (6.15) yield 8, fio + ve Ox fo = 0 (6.16) which is trivially satisfied by our equilibrium solutions fi) = fio(v). The first order terms in (6.15) are Ifa + U9, fa + i Eafe = 0 (6.17) Looking for plane wave solutions ~ exp(ikx — iwt) this is ~ tw fy + tkucfy =~ BO, fa (6.18) or : Salesest) = IL Ba, fav) (6.19) o — ky, The only one of Maxwell's equations (3.28) needed for electrostatic waves is Poisson’s equation, which in the present case is ikE = Ame(n — n,) ane f dv Gf) = ane f avy ~ fa) "8, ec) ~itnete fay [ fu, Me fe) wo ~ ky. w — key i (6.20) i Eliminating £ from both sides we obtain the dispersion relation for electrostatic waves in an unmagnetized plasma, (6.21) : 5G) = ag au, fale + x fa, dv. fo) (6.22) Notice that the ion component of g is reduced by the factor m,/m,. For example, if the electrons and ions are Maxwellian, we have fo = tay exp [“(u2 + v? + u2)/2u2] (6.23) whereupon Electrostatic Waves 75 elo) = Bayne, exp(-ve/202) Baty, exP(-v2/2u2) (6.24) where as usual v,? =/T,/m,. Kd EXERCISE Verify that the Maxwellian (6.23) satisfies the normalization (6.14). For equal temperatures Tj = T,,we have ¥) << y,, and g(u) is as shown in Fig. 6.1. Notice that g(u) has the units (velocity) '. The ion contribution appears tiny when compared to the electron contribution. However. for low frequency motions the ion contribution can be very important, as inthe jon-acor ve Let us use the dispersion relation (6.21) to find the relation between { frequency w and wave number k,Yor high frequency electron plasma waves called Langmuir waves. The high frequency of these waves implies that the massive ions do not have time to respond to them, so we ignore the ion contribution to g(u) in (6.22), that is, we let m, — ©, This is equivalent to ignoring the ion motion in our derivation of the plasma frequency in Section 1.4. The dispersion relation (6.21) includes an integration over an integrand with a pole at {)=_w/k. This pole must be handied with care. For the present, suppose we restrict ourselves to waves such that w/k >> wforallu for which g(u) is appreciable, so that d,g(u) = Oatu = w/k (Fig. 6.2). With this assumption, (6.21) can be integrated by parts to obtain oe 1-7 % where the boundary terms vanish h beca se g(u — +0) = 0. Expanding the de- nominator up to and including second ofder terms in uk/a, we find 1-2 favo (1 + 2Hk = EXERCISE Verify the expansion. (6.25) (6.26) tu) Jon contribution u Fig. 6.1. The function g(u) as predicted by (6.22) for an equal temperature Maxwellian plasma, 76 Vlasov Equation aw) oy wk Fig. 6.2 The Langmuir wave calculation of Section 6.3 is appropriate only when the phase speed «/K of the wave is much larger than the thermal speed », With g(u) given by the first term on the right of (6.24), we have [ du g(u) = 1, f du g(u)u = 0,f du g(u)u? = v2. EXERCISE Verify these statements. Equation 6.26 then predicts (6.27) (6.28) which js the famous Langmuir wave dispersion relation; it can easily be committed to memory. EXERCISE {a) Obtain (6.28) from (6.27) with the given assumptions. (b) Verify that (6.28) is consistent with the given assumptions. (c) Show that (6.28) is equivalent to w = w,(1 + 3k?A2/2). (d) Use the result of Problem J .3 to modify (6.28) to include the effects of ion motion. Jn the next section we shall return to the dispersion relation (6.21) and show how to properly treat the pole when we do not have @, g(u) = O atu = w/k. 6.4 LANDAU CONTOUR in this section, we present a more complete treatment of electrostatic waves, which includes a careful evaluation of the pole in the dispersion relation (6.21). As shown by Landau [2], the best way to proceed is by solving the Vlasov equation (6.17) and Poisson’s equation (6.20) in the context of an initial value problem Lendau Contour 7 To simplify the discussion, we treat only high frequency Langmuir waves and ignore ion motion. Looking for waves with the spatial dependence ~ exp (ikx), and denoting the first order electron distribution by f,(k,v,t), Eq. (6.17) becomes af; + iku, fi — (e/m)E Auf = 0 (6.29) When we use the Laplace transform convention (5.13) to (5.14), and the fact that the Laplace transform of d,g(s)}is —iwg(w) — g(¢ = 0), the Laplace transform of (6.29) is tw f,(k,v,w) + iky, fy > (e/m,)E(w)9y, fo = flk.v,t = 0) (6.30) ve : EXERCISE Demonstrate that the Laplace transform of d,g(t) is ~iw g(w) — g(t = 0). Poisson's equation (6.20) is in this case ikE(@) = — are f dv flkiy,o) e (6.31) Solving (6.30) for f,(k,¥,a), we obtain (e/m E(w) 4, foo + AilKev,t = 0) ike) = Oe eee (6.32) which when substituted in (6.31) yields > ‘ne?/ aural ik [ ay 5 | E(a) = atte filleyt = 0) = fav (Ob (6.33) The factor in brackets on the left is the djelectric function as d, : kya) = 1 — Bef au 5 es (6.34) where the definition (6.22) of g(u) has been used. Note that (kw) = e*(k,a) Equation (6.33) then becomes “ane flkiwt = 0) EO) = Biieay faa ca [As with all Laplace transforms, E(a) is defined only for 4, sufficiently large, and the inverse Laplace transform F=f $% Eee" (6.36) is carried out along the Laplace contour as shown in Fig. 6.3. The Laplace contour must pass above all poles of E(w), which by (6.35) includes all zeros of e(k,w). With the Laplace contour as shown in Fig, 6.3, we have w, > 0 everywhere on the contour and thus @, > 0 in the evaluation of e(k,w) in (6.34). By analytically continuing the function g() to the entire complex u = u, + iu,-plane, we can think of the integration in (6.34) as occurring along the real y-axis, with a pole at u = w/k in the upper half w-plane, as shown in Fig. 6.4. 78 Viasov Equation Fig, 6.3. Laplace contour The inverse Laplace transform (6.36) is accomplished by analytically continuing E(w) to the entire complex w-plane. By (6.35), when we analytically continue E(w), we must analytically continue e(k,). This means that when w crosses the real o-axis from above to below, we cannot allow. the pole in Fig. 6.4 to cross the integration contour; if it did, the value of e(k,@) would jump by —27i times the residue atu = w/k. Thus, we must deform the integration contour in the complex u-plane as shown in Fig. 6.5 when w, < 0. The two sets of contours shown in Figs. 6.4 and 6.5 are collectively known as the Landau contour. A similar contour must be used to evaluate the other integral in (6.35). With E(w) in (6.35) analytically continued to the entire complex w-plane, the Z contour in (6.36) can be deformed as shown in Fig. 6.6. We do not attempt to close the contour with a semicircle in the lower half w-plane, since it is not clear from Fig. 6.4 Contour of integration used in evaluating the dielectric function (6.34) when w, > 0. Landau Contour 79 Fig. 6.5 Contour of integration used in evaluating the dielectric function (6.34) when a, <0. (6.35) that E(w) falis off fast enough at large negative w; for this to be done. Rather, we treat each part of the-contour in Fig. 6.6 separately. There are four types of contributions. First, there are the poles of E(w) such as the one at point 4 in Fig. 6.6. We assume that this pole comes from a zero of e(k,w) rather than from the other factor on the right of (6.35). Denoting the frequency at point A by wy, this pole contributes a term to E(t) with the time dependence exp(—iw4!). We call these contributions the normal modes, and we note that since the frequencies of the ,are given by e(k,«w) = 0, they correspond to the waves found by solving the dispersion relation (6.21) in the previous section. After some time has Fig. 6.6 Deformed Laptace contour used in taking the inverse Laplace transform (6.36). 80 Vlasov Equation elapsed, the dominant contribution from the normal modes comes from the one with the largest imaginary part of the frequency, as in point A of Fig. 6.6. If this is positive, the normal mode is unstable and grows with time; if it is negative, the normal mode is damped and decays with time. Second, there are contributions from segments like the one from point B to point C, With the contour a distance y > 0.below the real w-axis, this contribu- tion is of the form : © E(t) ~ {2 Blw)et = eH! f s E(wyeniet (6.37) which decays very.rapidly with time. Thus, after an initial sransient.period. ot time, these contributions can be ignored. Third. there are contributions from the two segments like the one from point D to point E, of the form Ed BW) ~ f Fo Elayeis! Fd = eins [S Elay + fader (6.38) L These contributions can be ignored since Efé).is.small for.ay — 9°; by (6.35) it varies as w' for large w. Fourth, the segment from point E to infinity gives a contribution E(t) ~ f “go emt E(an) (6.39) which vanishes since for wy) —_° the integrand scillates infinitely fast. Thus, the response to an initial perturbation consists of norma! modes that oscillate with the normal mode frequencies given by the dispersion relation (6.21), and transients. After some time, the normal mode with the largest imaginary part dominates. This discussion refers to a single wave number k. A spatially localized perturbation in a plasma will have a spectrum of wave numbers, and the response after an initial transient period will consist of a spectrum of normal modes at different wave numbers. If any one of the norma! mode frequencies has a positive imaginary part, the plasma is unstable and the perturbation grows with time. If all of the normal mode frequencies have negative imaginary parts, the perturbation eventually damps away. In the next section, we quantitatively evaluate the real and imaginary parts of the normal mode frequencies in terms of the zero-order distri- bution function fio(v). 65 LANDAU DAMPING In this section, we return to the Langmuir wave dispersion relation (6.21) and use our knowledge of the integration contours to carefully evaluate the contribution of the pole atu = w/k. This calculation is especially elegant when we assume that |w,| << |w,{, which can be checked after w is calculated. Writing the dispersion relation (6.21) in the form Landau Damping 81 e(k,w) = 6, + ig = 0 (6.40) and Taylor expanding about w = w, yields elkyw,) + te(kw,) + fa; dette) it 0 (6.41) Here the term ~ w, d¢,/dw is ignored because it is the Ponte of w;, Which is smaii, and d¢,/aw, which can be assumed to be small because w, ~“@, which can be seen by equating’ the real and imaginary parts of (6.41) este to zero; this yields (kyu) (6.42) a aaaaHe ei as kia) : a ON ela V/dohemm, 0/80] y=, (6.43) When «, = 0, the Landau contour is as shown in Fig. 6.7. The Plemelj formula (see Appendix C), as applied to the contour in Fig. 6.7, is and : =P (43 ) + rid(u — a) (6.44) u-@ where P means principal value; this allows (6.34) to be written 1 ~ Seef" dy 8) _ (kre, u— (@/) ni we 4, 0 ae (6.45) where for any function f(u), Pf” du LO = him [ tu FO + f a 0] (6.46) = ua eo Le u-@ late u—a The second term on the right of (6.44) comes from integrating around the semicir- cle in Fig. 6.7, which yields one-half of 2ni times the residue, The integrand of (6.46) is shown as Fig. 6.8. The two large contributions of opposite sign cancel each other, so that the principal value is not very sensitive to the properties of the function f(y) near the pole at u = a. From (6.42) and (6.45) we determine the real part of the frequency ‘a, from ‘ mere we 4, Bu) «(k,o,) = 0 = “pf au Tab (6.47) atu = w,/k, integrate For purposes of this integration, we can take d,g(u) by parts as in (6.25) to (6.27), and obtain fe gin ak eyes! efk,w,) = 1 Pes =0 (6.48) which by (6.28) yields, with the assumption ku? << w?, o? = a2 + 3k7u2 (6.49) 82 Vlasov Equation Fig. 6.7 Landau contour when w, = 0. The straight line portions of the contour are exactly on the real axis. [Note that the expressions below (6.26) are valid for an arbitrary g(u), provided that f du g(w)u = 0 and provided that one defines v, by f du g(w)u? = v2] In order to calculate w, with (6.43) we need ek, de(kw,) _ 2w2 2 See) = ee) 0 Tee (6.50) where terms ©(k?A,2) have been ignored. Then by (6.43) and (6.45), ‘Tre Pa 0 = RF delet (6.51) The total normal mode frequency is finally ) + This equation is valid for all Langmuir waves such that KA, << 1. The generali- zation of (6.52) to encompass all wave numbers has been presented by Jackson [3]. (6.52) 3 (145k Fig. 68 Principal value integrand. Wave Energy 83 When the slope of the distribution function 4,g(u)|,20/« is negative, as with the Maxwellian distribution in Fig. 6.2, Langmuir waves haveg, <“Qand are Landau damped. When the slope is positive at u = w,/k, as it is for.atange of wave numbers in the “‘bump-on-tail” distribution in Fig. 5.1 att = 0, then these wave numbers grow exponentiall With a Maxwellian g(u) given by the first term on the right of (6.24) one can explicitly evaluate w,, which is mw? u w= ae exp (0/202 )Ijeu se o. (F Vg Tex OP (BAenp (2K) (6.53) This vanishes for k — 0 and increases rapidly with increasing wave number, such that for KA, > 0.3 the damping w, es @, is so large that such waves are never observed. In Section 6.7 we shall consider the ‘physical mechanism of Landau damping. Heuristically, this can be considered as follows. Consider a wave with a phase speed V, = w,/k ina Maxwellian plasma (Fig. 6.2). Those particles with speeds u very close to ve interact strongly with the wave. Particles with speeds slightly faster than V,, are grabbed by the wave and slowed down, giving up energy to the wave, while particles with speeds slightly slower than the wave are sped up, taking energy from the wave. Since in a Maxwellian plasma there are more particles with speeds slightly less than V, than with speeds slightly greater than V,,, the net result is an energy gain by the particles and an energy loss by the wave; this is Landau damping. On the other hand, consider a wave with phase speed V,, = w,/k = usuch that d,(g) > Oas on the left half of the “bump-on-tail” in Fig. 5.1. Now there are more particles slightly faster than V, than slightly slower, the particles lose net energy to the wave, and the wave grows. These ideas will be made quantitative in Section 6.7. 6.6 WAVE ENERGY In Section 6.5 the real and imaginary parts of the normal mode frequency for longitudinal waves are determined from a knowledge of the dielectric function e(k,w). The importance of the dielectric function for longitudinal waves is due to its equivalence to Poisson’s equation ikE = 4np (6.54) which is effectively replaced by tke(k,wE = 0 (6.55) for purposes of calculating normal modes [see (6.33)}. Thus, all of the physics contained in Poisson’s equation is also contained in the dielectric function. It is important to note that the “|” in the dielectric function fia Fea _ age) (kya) = 1 $5 du V- W/k (6.34) 84 Vlasov Equation comes from the left side of Poisson's equation (the “vacuum” contribution) where- as the other term comes from the right side of Poisson’s equation and represents the contribution of the plasma (plasma = medium = dielectric). The dielectric function ¢(k,w) provides a very useful approach to wave energy. In this section, we present a somewhat heuristic demonstration of the relation between the dielectric function and wave energy. A more rigorous development can be found in Landau and Lifshitz (4). When we deal with energy, we deal with squared quantities, such as electric fields; therefore we must be certain to have only real quantities before squaring. Consider a real oscillatory electric field EW) = + E(1) exp (~iw,t) + + EX*(1) exp (i,t) (6.56) at a fixed spatial point. The time-averaged electric field energy density at this spatial point is ~-FP_1 = z > We = He = yy Bex lion FE? + BF exp Qia,t)] _ IEP a (6.57) where the terms at frequencies + 2iw,t vanish on averaging over the fast time scale 2n/e,. With the real electric field £ one can associate a real current J, such that J(1) = YaJ(2) exp (—iw,t) + ¥ J*(t) exp (iw,1). Since we are dealing with linear waves, there is a linear relation between current and electric field, M(t) = o(wE(t) (6.58) where o(w) is the conductivity. There is a simple relationship between the conduc- tivity o(w) and the dielectric function e(w), Ampere’s law for longitudinal waves yields 4 any OE Geet eee c =o (+ i)e c tw oe ee at = = ew)E (6.59) with which we identify 4: do) = 1+ Srigte) (6.60) where in accordance with electromagnetic theory we have introduced the dis- placement D = cE. Wave Energy 85 Let us now develop an expression for the time rate of change of wave energy. From the first form of Ampere’s law in (6.59), we obtain for real fields o£ ‘i = and (6.61) Multiplying each side by ee oe 7 7 Gy = — £9 =- i [E exp (—iw,t) + E* exp (ia,t)\[oE exp (—ia,t) + o*E* exp (iw,t)] (6.62) where we have used (6.56) and (6.58). Taking the average of both sides over the short time 27/a,, the second harmonic terms disappear and we have, using (6.57), alo 4d pp 1 én at EP = t (Eo*E* + E*oE) (6.63) At this point we expand o(w) exactly as we expanded e(w) in (6.41). From (6.60), _ 4rodw,) aaa (6.64) while 4 ea) = Aree) (6.65) Consistent with the assumptions used in (6.41), we assume |o,(w,)| << |o,(w,)| and expand oa) = ofw,) + gal to. do; = ofw,) + ia(w,) ~ 0% To (6.66) where we have ignored the term w(do,/dw)|, as being second order in small quantities. By (6,56), the normal mode electric field E(1) has the time dependence exp(w,t); therefore dE/dt = wE, and we can write do, do,| o te | ar (667) whereupon (6.63) a 1 pda) Fae 2p dnl a or i lel aa de | a + das) a) adh ae 4 2 a Ap 2 = o,(w,)|E| wae — IE| (6.68) Moving a term to the left side, we find 86 Vlasov Equation o,(w,)|E|? (6.69) field energy particle energy dissipation where we recognize the first term on the left as the time derivative of the electric field energy. Because the second term on the left involves o;, we identify it as the time derivative of the particle energy contained in the current. The right side represents dissipation due to o,(w,), which is proportional 10 €(w,) and thus to «; by (6.43). Since 0,(w,) represents dissipation, it is known as the resistive part of the conductivity, while o(w,) represents the particle energy in the wave and is thus called the reactive part of the conductivity. The two terms on the left of (6.69) represent the time derivative of the total wave energy Win. Thus, as a function of the electric field energy density (6.57), do, Wo = (: ~ dn aa ) W, (6.70) From (6.64), the constant can be written do, a 1-42 Go] = Gy loco), (6.71) ® so the total wave energy is d do [elo)],, Me (6.72) For example, for Langmuir waves with kA, — 0, we have [see (6.48)] 2 We ta) = 1 - (6.73) Therefore =2 (6.74) d i @, do Mle, = 1+ TF so there are equal amounts of energy in particles and electric field in a Langmuir wave EXERCISE Compare this result to your result in Problem 6.4. Since the wave energy is ~ |E|*, and E ~ exp (w,t), we must have aw, a = tw Mo (6.75) Let us verify that (6.75) is indeed satisfied by (6.69), which says qh. 1 1 Ge Ty CMWEP = — F ow) 6rW, Woo (6.76) t] ~ Brole) (a7da\loodg Physics of Landau Damping 87 But (d/da)lwe)|,, = ode/da)|y, since e(w,) = 0 by (6.42), and oa,) = ce@,)/4r from (6.65); thus (6.76) is dW _ _ _2efa,) a ae de,/da|,, Woo = 2a Keo (6.77) where (6.43) has been used to insert w;. Thus, (6.75) is indeed satisfied. The convenient formulas (6.43) for growth rate and (6.72) for wave energy in terms of the dielectric function e(X,w) are applicable to all electrostatic waves in a plasma, and are very useful in practice. Because of the form (6.34), one can plaus- ibly state that a full knowledge of e(k,w) for all values of k and w implies a full knowledge of the distribution function g(u). In the next section, we consider in de- tail the effect of an electrostatic wave on the distribution function; this leads to a microscopic understanding of Landau damping. 6.7. PHYSICS OF LANDAU DAMPING In Section 6.5 the phenomenon of Landau damping (6.52) is introduced as a mathematical consequence of the solution of the dispersion relation. In the present section we discuss the detailed physics of Landau damping [5]. This is done by considering the effect of the small. wave, associated with the perturbed distribution function f,(x,v,f = 0), on the background plasma represented by fo(v). For con- venience in this section, we denote the x-component of velocity by v, and suppress the “electron” subscript on f, and fy. Consider a linear Langmuir wave of the form E\(x',t) = Ey sin (kx — @,f) where x’ denotes the laboratory frame of reference, Ey is a small constant, and for the moment we ignore the imaginary part of the frequency. In the frame of reference x moving with the phase speed t,/k > 0 with respect to the laboratory frame, the wave field is independent of time and is given by E,(x) = Ey sin (kx), as shown in Fig. 6.9. All of the particles in the background distribution function fo(v) are affected by this electric field, and some are speeded up while others are slowed down. We focus our attention on only those particles in fy(v) that have speed vy in the lab frame at t = 0 and, thus, have speed # = vp — vin the frame moving with the wave phase speed vu, = ,/k (Fig. 6.10). This “beam” of particles with speed vp in the lab frame and speed @, in the wave frame will see the energy of some £, (x) Fig. 6.9 Stationary electric field as seen in the frame x = x’ — (w,/k)t moving to the right at speed w,/k with respect to the laboratory frame x’. 88 Viesov Equation feo) Fig. 6.10 Beam of electrons, ail with speed y, with respect to the laboratory frame and speed io = v — vy with respect to the moving frame. of its members increase with time and the energy of some of its members decrease with time, depending on the initial particle position x9. After a time f, a particle will have experienced a change in speed Av (independent of frame) so that the particte’s energy, as measured in the lab frame, suffers a change 1 1 AE = > mAuy + Avy? — CE mevy = mudv + Of(Av)’] (6.78) where we shall ignore the term in (Av)? in what follows; it can be shown that in the present derivation this term gives a negligible contribution to quantities of inter- est. We are interested in the average change in energy over a wavelength, (AE) x = rte Vol AY) xy (6.79) Since Av can be calculated in any frame, we work in the wave frame. Then fae se fi aun =f arya = =H f dt’ sin k(t") (6.80) 0 m, 0 where ¥(1) is the orbit of the particle. If we insert in (6.80) the unperturbed particle orbit (7) = xy + ot, without the effects of the wave field, we would find (Av),, = 0. Thus, we must include the lowest order correction to the particle orbit due to the effect of the wave. This is done as follows: ee f ao Au(t) = Xo +f a” i Ey iy — sf dt’” sin k&(1"") ae Xp + yt” (6.81) where we have gone far enough to pick up the lowest order correction in E,. Performing the last integration we find Physics of Landau Damping 69 eee ie . = om mG By + mie [cos (kxp + kit”) ~— cos kxo] (6.82) The next to last integration yields ae eee ek, RU!) = x9 + Gel’ — Tatas cos (kx) Ey. a + Fetag sin (kx + kipt’) — sin (kx,)] (6.83) The first integrand is of the form sin (kx) + kUot’ + A) where & is proportional to Ey, teEy & mi, cos (kx) 2. [sin (kxy + Kit’) — sin(kxo)) (6.84) mM, Kuig Since we are looking for the lowest order correction in Ey, we can Taylor expand sin(a + A) = sina + Acosa (6.85) to lowest order in A. Then Ey fi . du(t) =~ =f dt'[& cos (kxp + Kigt') + sin (kx9 + kipt’)] (6.86) e Jo We next wish to average Av over one wavelength, upon which the sin term disap- pears. The other terms are evaluated using the identities (sin (uw — a)cos(u ~ 5)), = — + sin (a — b) (6.87) (sin (w — a) sin (w ~ 6)), = (cos (w — a) cos (uw — b)), 7 + cos(a — b) (6.88) where { ), means an average over one period of the variable u. We find (Av), = (2) x f : ar{- wa sin (kit’) t ee +& cos (Kiet’)} (6.89) ‘The integration can be performed, and yields Ey 1 (Av(2))y, = (2) RGF Mleos (kit) = + kiot sin (kigt)} (6.90) Our aim is to form the change of energy, (AE),, = m-vy(Av),, and integrate over all velocities vy in the lab frame to find the total change in energy of the particles. Before doing so, let us evaluate (6.90) at early time, such that kit << 1. Note that early time for one “beam” of velocity tj may not be early time for another 90 Viasov Equation “beam” of velocity >> tp, Using the formulas 5 sinx = x — 7 +... (6.91) and cox 1-25 + + . (6.92) we find Acosx — 1) + xsinx ~ (6:93) Therefore (6.90) becomes = [eumee— (2) pats Bh kit << 1 (6.94) Thus, we see that “beams” with} > 0, that is, those moving faster than the wave, are indeed slowed down at early times, in the sense of an average over x). “Beams” with @ < 0, those moving slower than the wave, are sped up in an average sense. We are now ready to integrate the spatially averaged energy change over all velocities, in the lab frame. We cannot use (6.94) for (Av),,, because kigt << 1is not true for all “beams” at a given time t. Rather, we use (6.90) to find the total energy change W(7), Wi = f° dy KKBBg = J dv fludmvldrd., (695) where we have used (6.78). Since (6.90) involves velocities in the wave frame, it is convenient to make the change of variable % = vu» — v, in (6.95), which be- comes WO) = mf” dvs FoNbs + v-MAr). (6.96) where Ftp) = fo(dy + v,). We expect the major contribution to (6.96) to come from particles with velocities close to i) ~ 0; this can be seen from the fact that the expression (6.90) for (Av),, varies as (i). We therefore expand f (iy) ~ 7) + % 7(0). The product Toi + v,) = Fie + FO)u, + 9° F'O) + BF) (6.97) has four terms, Since (Av),, in (6.90) is odd in Bp, the two terms in (6.97) that are even in Gy will give zero when the integration in (6.96) is performed. Only the two odd terms in (6.97), 7(O)t, and tv, 7’(0) contribute. Usually, the latter term makes a much larger contribution. EXERCISE Verify this statement for a Maxwellian, with vu, >> v,. Physics of Landau Damping a1 If we keep only the latter term, (6.96) becomes Ey) pa wn) = mee) (= sy #0 ta[c0s (kit) — 1) + kigt sin ean) (6.98) With the change of variable x = kgf, the integral 7 in (6.98) becomes * : = x f 7 [2(cosx — 1) + xsinx] (6.99) The second term is found from an integral table to yield fae = (6.100) while the other term ylelds —27. EXERCISE (a) Verify the last statement by changing the limits of integration to (0,°), care~ fully integrating by parts, and using (6.100). (b) Evaluate (6,100) using contour integration. First, move the contour off of the (nonexistent) pole at z = 0 and then expand the sine in terms of expo- nentials. Thus, we find My =- 5 7 FO (2). (6.101) or E, W(t) = — 7 ne Kod (5) t (6.102) or identifying fy with nog, and taking w, ~ ,, 3 wn) Bu Eot (6.103) Equation 6.103 shows that the total particle energy is changing as the first power of time t, and is positive when g’(v,) < 0 and negative when g’(v,) > 0. The energy gained or lost by the particles must come from the wave. The rate of change of wave energy Wave must be equal and opposite to the rate of change of particle energy. From (6.103), 3 4 Pome = 4 WO) = 5 8 UEP (6.104) The total wave energy, averaged over a wavelength, is Wrae = 2Ey? (sin? (kx)),/8% = Ey’/87, where the factor of 2 is introduced because a Langmuir wave has equal amounts of energy in electric field energy and in particle kinetic energy. Thus, (6.104) is SF Wome = BE gu) Wowe (6.105) 92 Vlasov Equation If the electric field amplitude is varying with time as E,(t) ~ exp (yf), then the wave energy Wraye ~ exp (2-1); thus WWrave dt Comparing (6.106) and (6.105), we find + Fr ae) (6.107) which is in exact agreement with the formula (6.52) obtained by contour integra- tion of the linearized Vlasov equation. We see therefore that Landau damping is related to the initia] behavior of the background particles, with particles moving slightly faster than the wave being slowed at early times and particles moving slightly more slowly than the wave being sped up at early times; this is true only in a spatially averaged sense. The net Landau damping (or Landau growth) comes from contributions from all particles, averaged over space; however, because of the to? dependence in (6.90), particles close to the wave phase speed give the biggest contribution, which is why g‘(v,) is so important. The theory just developed is a linear one and thus is exact only for waves of infinitesimal amplitude. In the next section we discuss heuristically the modifica~ tion of these ideas for waves of finite amplitude. = 2y Mowe (6.106) 6.8 NONLINEAR STAGE OF LANDAU DAMPING In previous sections we have treated linear Landau damping, first by integrating the linearized Vlasov equation, and then by considering the detailed orbits of the background distribution function. Let us next look at the consequences of finite wave amplitude. We again think in terms of the initial value problem. We consider the back- ground distribution function, in the presence of the wave £(Z,t) = Eo(f) sin (kX — a,t). In the wave frame, moving at velocity v, w,/k with respect to the lab frame, the electric field is E(x) =:E, sin kx, with Corresponding electrostatic potential (x) = (£)/k} cos kx (Fig. 6.11). In this wave frame, ignoring the slow time dependence of E,(f), the electrons see a time-independent-electric field, so their total energy H = — ey(x) + mv? is aconstant, when v is measured in the wave frame. For each particle, the constant H is determined by the initial position and initial velocity; therefore, H = — eglx()] + Ym) =~ ey[x(t = 0)) + mvt = 0) (6.108) The corresponding equation of motion is eae eapen obo n@ny cat mt = — eysinkx = ~~ F— (—coskx) (6.109) Thus, the particles are moving in a potential well — eg(x) = — (eE,/k) cos kx Nonlinear Stage of Landau Damping 93 E(x) (x) Fig. 6.11 Electric field and electrostatic potential in the Langmuir wave frame. Consider all particles with vy = 0 at time ¢ = 0. These particles have different energies depending on their position; from (6.108), H = — eg[x(t = 0)] (6.110) Att = 0, these particles find themselves in a potential field as shown in Fig. 6.12. Particles at A do not moye. Particles at begin to move in.the well, withconstant energy. Those at C do the same. Those at D.are marginally stable; a slight pertur- bation will allow them to begin moving in the well. Each of these particles oscil- lates in the wel | frequeney_of oscillation, which decreases as we bottom of the well, at B'for instance, we can find the frequency of oscillation by expanding the force about x, = 0; thus from (6.109), mk = — eBy sin kx ~ — cE kx (6.111) from which we identify the characteristic frequency of oscillation, eEok wy = aa (6.112) or (6.113) where a, is known as the bounce frequency. eb (x) Fig. 6.12 Potential well for electrons in a Langmuir wave. 94 Viasov Equation Linear Landau damping was derived on the basis of only small perturbations in particle orbits. However, after one-half of a bounce period, the particle at B has moved a substantial fraction of a wavelength, and linear theory is invalidated. Thus, we expect linear Landau damping to hold only for short times, such that v2 < O and to the left if v < 0, as shown in Fig. 6.14. During the time each particle represented bya dat moves from jg det to the tip of its arrow, we have the eriod of linear Iandau damping, or growt| . When the slope of the distribution function i is negative, this early time behavior results in Landau | damping, with the trapped particles contribute. - However, after a substantial fraction of a bounce period, the trapped particles are smeared out around. their phase space orbits and the stage of linear Landau damping is over. The smearing process is facilitated by Fig. 6.13 Particle orbits in the v-x plane, neglecting the self-consistent change in the wave electric field due to the motion of the particles. Nonlinear Stage of Landau Damping 95 Fig. 6.14 Initial particle orbits in x-v phase space, the fact that. neighboring trapped.particleorhits have slightly different bounce perjods. Atx = Oin the wave frame, the initial distribution might be a Maxwelli- an, as in Fig. 6.15. At a much later time after many bounce periods, the particles with velocity.» ~ uy are smeares out, and the distribution looks as in Fig. 6.16. The flat region is uy — v, < vu < v, + u,, where the trapping speed. ois defined by Yam? = 2lewlan (6.115) 7 i =2 (<) elt? (6.116) < 2 7 _ , [tEo\! » =2 (Fe) (6.117) Note the relation between the bounce frequency @; from (6.113), and the trapping speed vu, from (6.117), oy = ku, (6.118) how Fig. 6.15 Maxwellian ats = 0, x = 0. 96 Vlasov Equation fl wpm [Saye % Fig. 6.16 Distribution at x = 0,2 >> 0. ‘We can now attempt to construct an overall scenario for the initial Jalue prob- lem of a finite amplitude Langmuir wave. At early time \t << wy 'y we have propriate damping.zate. At ¢ ae ‘wy, the trapped particles have gone, through a half a bounce and as they start to come through the second half of a bounce, they can put back into the wave some of the energy that they initially took out of it. This reversal of energy is by no means complete, however, because by this time the trapped particles are out of phase with each other. We can then construct a picture of the behavior of wave amplitude versus time as shown in Fig. 6.17. Here, y;.is the Landau damping rate, and we have assumed fy, >> w». For very large time, far off the right side of the figure, the curve will approach a straight line, all the phases will be completely mixed, and the wave will become a BGK mode [6] (see Section 6.13). Much of the present discus- sion has been heuristic. A completely self-consistent and nonlinear treatment of Langmuir waves is a very interesting current topic of research; see, for example, References [7] to [9]. log 1E (2) | a no oa Fig. 6.17 Langmuir wave amplitude versus time, ‘Stability: Nyquist Method, Penrose Criterion 97 6.9 STABILITY: NYQUIST METHOD, PENROSE CRITERION In Section 6.2 we discussed methods of constructing Vlasov equilibria. Once we have found these equilibria, we must ask the question: Are they stable or unstable? For example, we know that when the ions and electrons are both Maxwellian with no relative drift, we expect the system to be stable. On the other hand, when the electrons form a cold beam moving through cold ions, elementary fluid theory (see Chapter 7) predicts instability. The question of whether or not a spatially uniform equilibrium is stable can be answered by the Nyquist method. [Note that the equilibrium must be uniform; this means that the Nyquist method unfortunately cannot determine the stability of BGK modes (see Section 6.13)]. We know that all information concerning the linear stability of an equilibrium to electrostatic perturbations is contained in the dielectric function ¢e(k,w), which is obtained by linearizing the basic physical equations about an equilibrium. Knowledge of e(k,w) everywhere in the complex w-plane determines all of the electrostatic stability properties of a system. Consider a general function e(k,w). Regarding k as a fixed real positive con- stant, we can then consider ¢ to be a function of w only. Form a new function (1/e) de/dw. Then it turns out that ot f see ee =n aieees of (k,w) inside the contour —_(6,119) dat J € dw where ¢, is any closed contour in the complex w-plane, the integration is in the counterclockwise direction, and we assume de/@w has no poles in the enclosed region, and € has only simple zeros. The derivation of (6.119) is as follows: Near any simple zero ay of e, we have robe ko) = 0. +e | (w —@) +... (6.120) ‘ae bey 29 while : de _ a ae a tw Bw |, al. Comes tices (6.121) Thus, near wy, we have La 1 see (6.122) € dw w —~ Ws A trivial application of the residue theorem to the left side of (6.119) then yields the right side of (6,119). Thus, (6.119) tells us the number of roots of ew,k) = Oin a certain region of w-space. In order to locate all the unstable roots, we simply need to evaluate (6.119) along a contour that includes all of the upper half w- plane, since having e(k,w) = 0 when w, > 0 means instability. In Fig. 6.18, (6.119) would yield N = 2, while in Fig. 6.19, (6.119) would yield N = 0 (left), N = | (middle), and N = 3 (right). As one integrates around a contour in the w-plane, it is possible to draw a corresponding contour in the complex e-plane. In that plane we have from (6.119) 1 1 de IT I fo (6.123) 98 Vlasov Equation oy Fig. 6,18 Contour c,, encircling the upper half of the w-plane. where c, is the contour in the e-plane obtained by evaluating «k,w) at each point on the contour c,, in the w-plane. Thus, the middle term in (6.123) says that the contour c, must pick up N zeros of ¢, so that in the eplane it must circle the origin N times. Three examples are shown in Fig. 6.19. We thus have a powerful technique, the Nyquist method, for determining wheth- er a physical system described by a dielectric function e(k,w) is stable or not. We simply draw the curve ¢, in the e-plane, found by mapping the curve c,, which encircles the upper half w-plane. If c, does not encircle the origin « = 0, then the system is stable. If c, does encircle the origin one or more times, the system is unstable. The Nyquist method by itself does not tell us. the growth rate of the instability. . €=0 <=0 €=0 Sp Be ey as ‘ § 6 4 * 5 a “ Fig. 6.19 Three examples of contours c, and the corresponding contours ¢,. Stability: Nyquist Method, Penrose Criterion 99 Let us test these ideas using the Vlasov~Poisson system, as represented by e(k,w) in (6.34) for any w, and in (6.45) for w, = 0. We are trying to map the c, contour onto the e-plane. First, consider the semicircle at ¢%. Then (6.34) yields (kw) = 1 (6.124) everywhere on the semicircle, since for |w| — ° the second term in (6.34) van- ishes. The remainder of the c,, contour is the path from w = —% to w = + along the real w-axis. But this is precisely the situation when we can use the form (6.45). By looking at the sign of the imaginary term in (6.45), we can see that as w — +20, ¢ > 0, while as w — —%, ¢; <0. Also, for large Jw], ¢, = 1 — w2/o2. We thus have the beginning of our path c, in the e-plane as shown in Fig. 6.20. If the remaining part of the path c, looks as it does in Fig. 6.21, we shal] have no instability because the origin is not encitcled. However, if the remainder of c, looks as in Fig. 6.22 we have one unstable mode, because the origin is encircled once in the counterclockwise direction. Note that it is impossible to obtain the contour shown in Fig. 6.23 because this encircles the origin in the clockwise sense, predicting N = —1 by (6.123), which is nonsense. Because of the handy formula (6.45), which describes the entire path c, except for the point e = 1, we know immediately all the places where c, crosses the real axis. These are just the places where ¢; = 0, or by (6.45), where d,g(u)|,,, = 0. Thus, a single humped g(u) has only one position u) where d,g(u) = 0. In this case, ¢, can only cross the real ¢-axis in one place, and we immediately know that this is a stable system. This is because it is not possible, with what we already know about the contour ¢,, to encircle the origin in a counterclockwise sense and only cross the real ¢-axis once. Fig. 6.20 Portion of the contour ¢, that comes from all portions of the ¢, contour in Fig. 6.18 with |e] = &. 100 Vlasov Equation & Fig. 6.21 A contour c, that would yield no instability. We can verify this conclusion by evaluating ¢, at the position w vanishes. From (6.45), 2 Be Oe Je du _4g(4) e u— w/k =1-%p le dy 8H) _ (6.125) eR u— Uo i If in the numerator of the integrand we subtract zero, in the form of 0 = d,g(o), we can write ce Al g(u) = B(uo)) =l GP fa ee (6.126) 6 ‘ Fig. 6.22 A contour ¢, that indicates one unstable mode. Stability: Nyquist Method, Penrose Criterion 101 Fig. 6.23. A contour ¢, that can never occur because it indicates N = —1 We can integrate (6.126) by parts, because all quantities are well defined at the singularity, to obtain @(@ = ku) = 1+ Ze vf du (su) _ (sha) ee (6.127) where the principal value symbol is ae integrand in (6.127) is Positive definite, since g(uo) is the maximum value of g. Thus, (6.127) yields ~ elo = kup) > 1! (6.128) so that the picture of c, is as shown in Fig. 6.24, confirming our prediction that the origin is not encircled and that there is no instability for a single-humped distribu- In interpreting this result we must remember that ‘e(u) includes an ion contribu- tion as well as an electron contribution. Thus, one situation in which the above result holds is when a single humpe ribution_is moving through a background of infinitely massive io ce, (6.22) shows that g(u) de- pends only on the electron distribution. Next, consider the case where the distribution has a double hump, with a relative minimum between them (Fig. 6.25). Then (6.45) predicts three values of «; (ea kup) Fig. 6.24 Contour ¢, obtained from Vlasov-Poisson theory for a single-humped distribu- tion g(u). 102 Viasov Equation gi w Hy My he Fig. 6.25 A double-humped distribution, with zero slope at uw), uo, and u,. w@ = ku;, @ = kuo, and @ = ku; where c, crosses the seal axis. If the absolute maximum of g(u) occurs at_uy = uj, then it is straightforward to show that ed = ku,) > 1s EXERCISE Verity this, using the same argument as in (6.127) and (6.128). It furthermore must be the case that as we move along the w contour from w = —& tow = +22, we encounter the crossings of the real _e-axis in the order w= kun), ew = kifg), and them efw = kuz). Thus, the first part of ¢,looks as shown in Fig. 6.26. We are now v allowed two more crossings of the real e-axis. One possibil- ity is as shown in Fig. 6.27, which does not give in instability, Another possibility is shown in Fig. 6.28, which indicates one unstable + root. We see that a necessary condition for instability is oo : (wo = ku) < OY (6.129) which from (6.45) is a ae = we 48 U4). ew = ku) = 1 ~ Fy P feu 280 (6.130) 4 Su € (a= kuy) * Fig. 6.26 Portions of ¢, for a double-humped distribution. Stabillty: Nyquist Method, Penrose Criterion 103 Fig. 6.27 Contour c, for a stable double-humped distribution. By once again subtracting 0 = d,g(u) in the numerator, we can integrate by parts to obtain 2 i eo = kg) = 1 SE fF au et = se a (6.131) where the justification is the same as in (6.127). Now (6.129) says we need ew = kup) < 0 for instability, But this will be assured if (6.132) Fig. 6.28 Contour ¢, for an unstable double-humped distribution. 104 Viasov Equation tw) Fig. 6.29 Three regipns A, B, C that contribute to the integral in (6,132). for when (6.132) is true, (6.131) ranges from em = kuy) = for k ~ © to @ = kup) = -% fork — 0. Thus, for some value of k, we must have e(@ = kg) <0 while (@ = ku,) >Q)which is the necessary condition for in- stability. Equation (6.132) is called the Penrose criterion [11, 12]; it is a necessary and sufficient condition for the linear instability of a Viasov-Poisson equilibrium. Consider the integration in (6.132) as applied to the three regions shown in Fig. 6.29. In all of region B the integrand is negative, while in regions A and C the integrand is positive. Thus, the negative contribution in B -thust exceed the positive contributions TréixAjand C in order that the Penrose Criterion be satisfied. Notice that the negative contributic {Bis enhanced by aeep hols} EXERCISE Show that if g(1m) = 0, the Penrose criterion is always satisfied. The Nyquist method also tells us the range of unstable wave numbers. By requiring! e(w = kuo) <0 and ew = kt) > 0, we find that Eq. (6.131) for g(a) Yeett Fig. 6.30 Two drifting Maxwellians with v, < vyqn that are unstable Stability: General Theory of Linear Viesov Waves 105 et) Fig. 6.31 Two drifting Maxwellians with vain < v, that are stable. (w= kup) and the equivalent expression for ew = ku) yield a? f ay EO. 8(uo)] (u [a — g(42)) af Prvfalesyd) (6.138) where we have taken the zero order charge density and current to vanish. Combin- ing (6.135) to (6.138) with Maxwell’s equations VE, = 470 (6.139) --1#® VxE,= Purr) (6.140) v-B, =0 (6.141) and = ny, 1 vx B= aE J+ CoO (6.142) we have 4 complete set of linear equations with which to find the dispersion relation for an arbitrary linear wave. Let us now solve the first order Vlasov equation (6.136). Recall that in Section 6.2 we introduced the-concept of unperturbed orbits of hypothetical particles, so that the zero order Vlasov equation (6.135) could be written 2 tain =0 (6.143) We then proceeded to find equilibrium distribution functions that were functions of the constants of motion of the hypothetical particles. Consider a hypothetical particle moving in a given force field, consisting of zero order electric and magnet- ic fields. For the present, we allow the zero order electric and. magnetic fields to be functions of both space and time; later we will make them constant. From New- ton’s laws of motion we can follow the orbit of a particle to find X(1), its orbit in real space, and V(t), its orbit in velocity space, The equations for these varjgbles are X(2) = Va) (6.144) and MiG vo = 2 ferxon +2 x moon} 145 In particular, we associate one of these orbits with every point (x,¥,2) in seven- dimensional phase space, by choosing the appropriate constant of integration. That is, we choose Stability: General Theory of Linear Vlasov Waves 107 X(r) =x - f Xr) ar" (6.146) i or RC eee f vr") dt” (6.147) |. which satisfies (6.144) and also has the property that X(t) = x (6.148) Similarly, for velocity, we choose the constant of integration such that vir) =v — f Vora” (6.149) R Thus, the orbit [X(7’),V(1’)] is the orbit of that particle which reaches the position (x,¥) at time f. : Consider any function Mont ah ae [Hr | =F evan v=V(r) +X) WhO WO voxcey v=vir) + V(r) > Wt Ww) (6.150) vaver'y Along the unperturbed oe we have X(1') = V(r’) = vy, and va’) = = & feo ae wo x BX) (6.151) Thus, the right side of ary is just the left side of (6.136) when h = f,,, and we can write (6.136) as : d rs Fe FAVED ware) =~ *& Exe.) vevir') vr + 2 x Bx} SW Lo VE) eoxir) (6.152) vevry Both sides of (6.152) can be integrated from t’ = —% to #’ = f along the unper- turbed orbit that ends up at X(¢) = x and V(t) = v. The result is Su Oovit) = FalX(0),V0)0 = 2] af fw ferry + 20 x BW} = Wy FolX V1 [yreniey (6.153) vsviry) 108 Viasov Equation Equation (6.153) is a forma! solution for f;, where in the integrand we must evaluate E,, B,, and fj at the correct point of the unperturbed orbit (X(t), V(1')] of the hypothetical particle at time ¢’. From now on, we consider only uniform, stationary zero order fields Ep = 0, Bp = constant, and fo = f,o(v) a function only of velocity, although (6.153) can be used in more general cases. Let us look for plane wave solutions to (6,153), E,(x,0) = E exp (—iat + ik+ x) (6.154) B(x, = B exp (—iot + ik x) (6.155) and SalX,v,0) = f(y) exp (tat + ik + x) (6.156) where E and B are constant vectors. For the moment, we take Im(w) > 0, so the wave is exponentially growing with respect to time. Then a finite amplitude at time 1 implies f,,() = —2e) = 0, and we can ignore the contribution of f,,(¢" = —2) in (6.153). The philosophy here is to find a dispersion relation valid for Im(w) > 0, and then to analytically continue the dispersion relation for arbitrary w. This is the same technique used earlier for Langmuir waves, which led in that case to the Landau contour for evaluating the electrostatic dispersion relation. Equation (6.153) now reads vu" Flvyexp(—iwt + ik+x) = — Le. ar[ E+ oO. xB | + Vy fol¥ yaviery XP [Tiwt” + ik - XC) (6.157) Moving exp (—iwt + ik + x) to the right we find se aaseane Que pate: very 0) = > fiale+“e xB Ve Lol yay EXP (riw(’ — D+ ik + (XC) — xd] (6.158) Making the change of variable r = 1" — ¢ we have jy =-# Pale+*O xa] ms Wy Fool¥ Vv exp {ivr + ik [X(r) — XJ} (6.159) where we realize that in this notation, [X(r),V(7)] is the orbit that yields [X(r = 0), Vir = 0)) = (xv). 4x1 LINEAR VLASOV WAVES IN UNMAGNETIZED PLASMA Let us evaluate (6.159) for the case of an equilibrium plasma with no zero order fields. Then the unperturbed orbits are from (6.147) and (6.149): vi‘) =v (6.160) and . Xi’) =x —We- 2) =xtvr (6.161) Linear Viasov Waves In Unmagnetized Plasma 109 Since V(r) does not depend on r, (6.159) becomes ’ joy =~ (e+ x6). vs 6 xf dr exp (—iwr + ik- vr) (6.162) Since Im(w) > 0, the integral is well behaved and we find : : a 1 £ drexp (iar + ik-v7) = ay (6.163) so that (6.164) Taking B = Oand looking for electrostatic waves, we could combine (6.164) for each species with Poisson's equation to obtain our old Viasov-Poisson dispersion relation, leading to Langmuir waves, ion-acoustic waves, and Landau damping. Taking B,(x,1) = B exp (—iws + ik « x) # 0, we find from Ampere’s law (6.142) and Faraday’s law (6.140) that ‘ ik XE, = a B, (6.165) and an iw ik x B= hI --> BE (6.166) or Clk % (kX E,)) = —i4rad — aE; (6.167) For isotropic zero order distribution functions fo(v) = fio(v), the term(v x B) - Vy fw in (6.164) vanishes. EXERCISE Verify this. Then Jat) = Ta fdvvfulxvs) “9, = ¥ aexp(-ior + ike») fav Ere v (6.168) Let us look for transverse waves, such thatE 1 k.Thenk x (k X £) = — KE, and (6.167) becomes (? — Re)E = —4ntw 2 fe ees where the harmonic dependence has been factored out. Suppose the vectors are arranged as shown in Fig. 6.32. Then the numerator in (6.169) has a termE - V, = ee. k-y* (6.169) 110 Viasov Equation 2, By yey ak Fig. 6.32 Vector components for an clectromagnetic wave in an unmagnetized plasma E 4,,, while the denominator is of the form « ~ kv,. Thus, we can integrate by parts in the numerator, picking out only the term E 4, v = Ey = E£. We find (vx) a (6.170) 2 (@? — RetyE = Arak Fm f dvs Where g(v,) is as usual defined by (6.22) and where we have ignored the ion dependence by allowing m, + ©. Dividing out the constant vector E, we finally obtain the dispersion relation for electromagnetic waves in unmagnetized plasma, Be + oe f dv, we) (6.171) Whenever there might be a problem with the pole in (6.171), the Landau contour must be used, for the same reason as in the discussion of Langmuir waves. The integral in (6.171) can easily be performed with the assumption w/k >> v, forall vu, Of interest. Then @ — kv, ~ «, the integral {© du, g(v,) = 1, and we find ow = wd + Ke% (6.172) for linear electromagnetic waves in unmagnetized plasma. EXERCISE Show from (6.172) that even in a more accurate treatment of the integral in (6.171), there is no need to consider a contribution involving BUD v=ask In this section we have seen how electrostatic and electromagnetic waves in a uniform plasma with no external fields are treated via the Vlasov equation. In the next section we shall set up the equivalent procedure for a uniformly magnetized plasma 6.12 LINEAR VLASOV WAVES IN MAGNETIZED PLASMA In the previous section we have seen that the linearized Vlasov equation can be solved by integrating along the orbits of hypothetical particles moving in the zero Linear Viasov Waves in Megnetized Plasme WwW order fields; these orbits are called unperturbed because they do not fee] the effect of the wave motion for which one is looking. In an unmagnetized plasma, the unperturbed orbits are simple, and it is straightforward to evaluate the perturbed distribution function. Consider the evaluation of (6.159) in the presence of a uniform background magnetic field. Now the unperturbed orbits are spirals around the magnetic field lines, satisfying Newton’s law mN(r) = oa V(r) X By (6.173) Choosing the magnetic field in the £-direction, the gyro-orbits that satisfy X(r = 0) = x, V(r = 0) = v, are Vr) = v, (6.174) Zr) = 2+ ur (6.175) Vr) = v, cos ( — Q,7) (6.176) ; wo a i Xx) = x — GF sin(e — O47) + Fe sine (6.177) Vir) = v, sin (ge — O47) (6.178) and Yr) = yp t+ jr 008 (2 = Or) — Toe (6.179) where the gyrofrequency is 2, = a and @ is a constant 0 S y S 2n EXERCISE Verify that (6.174) to (6.179) satisfy (6.173) with the appropriate boundary conditions. Inserting the orbit (6.174) to (6.179) into (6.159), we can carry out the integration over 7 in (6.159). Then using Maxwell's equations to eliminate B in terms of E, we could obtain a general dispersion relation for waves in a uniformly magnetized plasma. This dispersion relation would contain all of the waves to be encountered in the next chapter on fluid theory, for example, Alfvén waves, upper-hybrid waves, and extraordinary waves. In addition, entirely new wave modes appear in the Vlasov formulation, which are impossible to obtain from a fluid formulation. Known as Bernstein modes, these waves depend on the detailed interaction of the wave motion with the gyro-orbits of the particles. Because the details of the evaluation of (6.159) are quite tedious (see (13], p. 405 and [14] and [15]), we shall simply sketch the derivation and point to the physical- ly interesting terms. For the zero order distribution, we choose the natural func- tion of the constants of the motion fig = fio(v,,u,). Then with vy, = (v,? + v,2)? we have Vy Fool ¥1¥)I every = (89, fo + 99, fa + 24, fo—vey = (Vk + ¥,9) * au, fa + £ 4, fu (6.180) EXERCISE Verify this equation. 112 Vlasov Equation Every term in (6.180), except V, and V,, is a constant of motion of a particle orbit and can be taken outside the integration in (6.159). In general, the perturbed magnetic field B can have three components. However, the combination WV XB) + Vifo = (-VivB, + v.¥,B,) x By, So + (V,B, — BLV,) a, fw (6.181) has only single terms, V,, V, that depend on 7 and must be kept inside the integral. After taking the constants E and 6 (expressed in terms of E through Maxwell's equations if desired) as well as v and fo(v,,v,) outside the integral, all remaining terms are of the form Vr) T =fa V,(7)| exp {—fwr + ik + (X(7) — x} (6.182) 1 The integrals (6.182) can be evaluated in terms of the identities etsin? = S' J,(a) exp (ind) (6.183) and eviasind =! J,(a) exp (—ind) (6.184) where J, is the Besse! function of order m. Without loss of generality, we choose the wave number to be k = k,& + k,% then (6.175) and (6.177) yield k + [X(r) — x] = k,v,7 hy ae 4 Bete i a, Sin(e ~ O47) + sing (6.185) Thus, choosing the factor unity in (6.182) as an example, (6.182) becomes 1 f dr ew iurtik.v.r e-iasin (0,7) giasing (6.186) where a = k,v,/Q,. With (6.183) and (6.184) we have 1 = fi drermrrton S 54a Ee X ertorins SF (ayeine (6.187) The integration can now be performed because only three exponential factors depend on 7. We find =¥ S| a 1= Ida) SY J,(aeeh “He = Tn, = ke) (6.188) a a A glance back at (6.159) shows that we have just calculated the term involving £, Oy, fo(V,.v,); this term is therefore Lineer Viasov Waves In Magnetized Plasma 113 Heavies Ne 10, = kv) Fay) = 24, a, folv.w) S z- (6.189) Fm ee {+ other terms involving E and B} where a = k,v,/Q,. The other terms in (6.189) gre no harder to calculate than the one shown; for example, the integral (6.182) involving V,(r) = v, cos(g — 0,7) is easily calculated using V,(r) = v,[exp (ig — #M,7) + exp(-ig + 19,7))/2, which fits naturally into the form (6.187) and has the effect of shifting the indices of the Bessel functions up and down by one. Eliminating the perturbed magnetic field using Maxwell’s equations, we find as in the previous section [Eqs. (6.167), (6.169)} of + fk x (k x BY = —4riw S gf Povf (vy (6.190) ana ‘ where 7,(v) from (6.189) is linearly proportional to the various components of E. Thus, (6.190) is three equations in three unknowns; setting the determinant of the coefficients equal to zero yields the horrendous dispersion relation found in Ref. 13, Eqs. (8.10.10) and (8.10.11). In (6.190), the velocity integral must, as usual, be performed along the Landau contour. By taking appropriate limits, we could obtain from (6.190) all of the waves of fluid theory. In addition, there occur waves called Bernstein modes {16]. These modes propagate across the magnetic field, so that k, — 0 in the denominator of (6.189), and are predominantty electrostatic. Then the denominator can be taken outside the integral of (6.189), and the dispersion relation will contain sums of terms like (a — /0,)71. This term goes from —° to +e as w changes by a small amount 2,. The result is a set of modes, one for each harmonic of the cyclotron frequency of each species. One of these modes corresponds to the upper-hybrid mode (see Chapter 7) in the limit of small wave number. The dispersion diagram is qualitatively as shown in Fig. 6.33. This figure is for the case wry = (02 + w2)'? between 2|,| and 3|2,|. Notice that there are waves for any frequency such that 19,| ayy there are stop’ bands where no waves exist. There are other interesting features of (6.189) and (6.190). Recall that in the unmagnetized case, Landau damping came from a resonant denominator of the form (w — kv)", representing strong interaction with particles with speeds equal to the wave phase speed, v = w/k. From (6.189), we get the same kind of reso- nant denominator, but involving only the component of velocity v, along the magnetic field. The same procedure as in the unmagnetized case will yield damp- ing involving those paticles with parallel speeds 7 = 0,41,42,.... (6.191) If k, + 0, and w # JO, for any /, there is no damping 114 Vlasov Equation hte 121 Fig. 6.33 Sketch of the dispersion curves for electron Bernstein modes. When / = 0 in (6.191), we have our old friend Landau damping. When / # 0, we have cyclotron damping. Physically, cyclotron damping occurs when the parti- cle sees a wave whose Doppler shifted frequency is the gyrofrequency or some harmonic thereof: w — k,v, = 19, = +1,42,.. (6.192) Suppose the wave is circularly polarized, or has at least one component that is circularly polarized, For example, consider / = 1, Then the particle might see the field shown in Fig. 6.34 as it goes around its gyro-orbit. Evidently, the particle can be continuously accelerated and thus the wave is damped. The concept of cyclotron damping has an interesting extension to relat ——E Fig. 6.34 Orientation of the electric field of a wave, and the position of a particle, at four points around the particle’s gyro-orbit. t BGK Modes 115 mass is a function of particle speed. The resonance condition (6.192) then depends on v, as well as v,. This completes our discussion of linear waves in uniform Vlasov plasma. In the next section, we encounter our first example of a nonlinear wave. 6.13 BGK MODES In preceding sections we have studied the linear waves that can exist in a Viasov plasma. As the intensity of such waves is increased, nonlinear effects that are ignored in the linear derivation become important. There are many different non- linear effects, and much current research in plasma physics is devoted to the theoretical, experimental, and numerical study of nonlinear waves. In this section, we introduce an important class of nonlinear waves, the BGK modes, named after Bernstein, Greene, and Kruskal [6]. In Case C of Section 6.2 we encountered an equilibrium distribution function in the presence of a spatially varying electrostatic potential. A BGK mode involves just such a distribution function, where the electrostatic potential is produced self-consistently by the distribution function through Poisson’s equation. For simplicity, let us consider the time independent situation with spatial varia- tion only in the %-direction. Then - cach species the Vlasov equation is (va co nm, # a 2) Les v) =0 (6.193) =_v,. The potential g(x) must be determined self-consistently through Poisson's Teg oe = dre if dfx) — L dv fix, »| (6.194) where it is understood that the v, and v, dependencies of f,(x,v) have been integrat- ed over. We already know the solutions of (6.193); these are just the equilibrium distribution functions (6.12). Thus, we can pick two arbitrary functions f,[v2 + 2q,0(x)/m,], one for each species, insert these in (6.194), and solve the resulting equation, which is ye _ ax ane dv {z [v*— 2eotvm. | ane [¥ + reotxr/m] } (6.195) This equation must be solved for ¢(x) subject to appropriate boundary conditions. For example, we may wish to look for periodic wavelike solutions, or for localized soliton solutions. It turns out that there exists a huge number of solutions to the nonlinear integro-ditferential equation (6.195). Let us begin to study (6.195) by looking at a very simple case, where each species is a cold beam of-pazticles, each particle of species s having the same speed at a given position, Thus, we choose SAX V) = Aryye LP — 2eplxV/ mg — ve} (6.196) 116 Viasov Equation where for definiteness we choose only the positive root inside the delta function; that is, we recall the relation — Y= yo) a0) = | “E (6.197) YY \yny, where y, is the solution of f(y) = 0. Then (6.196) becomes FAR0) = mo ~ CY ~ 5) (6.198) where B(x) = Lud + 2eg(x)/m]!? (6.199) Similarly, for the ions we take Sie,2) = my % &v ~ 4) (6.200) with 5Ax) = [v7 — 2eg(x)/m]'? (6.201) Here, v; and v, are arbitrary constants that we choose large enough that (6.199) and (6.201) always yield real positive values for 3, and i, We have chosen the normalization constants 1, the same for ions and electrons; we must check at the end of the calculation that this gives an overall neutral plasma. We now look for spatially periodic solutions to (6.195). Integrating f,, f, over all velocity space, we find ndx) = No a (6.202) and nfo) = no (6.203) a so that (6.195) becomes Pe _ ee Ge = bance (# #) (6.204) or a 2eelx)\'? Zegix) \'"? Ga = Anne {(1 + ) -( - ae) } (6.205) We notice the fortunate circumstance that (6.205) is in the form of a pseudopoten- tial equation; that is, it is in the form ig ed me o Vie) (6.206) with Vy) = — dang {mad ( + 2)" m2 (1 eee y} (6.207) Mv, my; BGK Modes 117 ve Fig. 6.35 Pscudopotential well used in finding BGK modes for an electron beam traveling through an ion beam. As a specific example, let us choose mv? = my»; Me) =~ 4rnar {(1 + 2ee\" 4 (1 - 20" 7 (6.208) The sketch of V(¢) is as shown in Fig. 6.35. Equations of the form (6.206) are called pseudopotential equations because of théir resemblance to Newton's law of motion mt = F(x) = — dV(x)/dx. With an initial choice of the ‘‘pseudoenergy” somewhere between —(32)"? noT and —8nmoP, the “pseudoparticle” oscillates forever in the pseudopotential well, producing a spatially periodic potential that oscillates between —gpand gq, as shown in Fig. 6.36. The function g(x) is a periodic function but is not a sine function; it becomes a sine function in the limit of very small @3; In the limit of small yo, we can make analytic progress by expanding the square roots in (6,205), assuming ey(x)/T << 1 forall x. We obtain = T. Then 2 2 oy Srna (6.209) with solution A(x) = Go sin (2'? X/ Ren) (6.210) where we have defined an effective Debye length dat = v/e, (6.211) Recall that v, here is a constant and not a thermal speed. ox) % “bo Fig. 6.36 Periodic BGK madgs for an electron beam moving through an ion beara. 118 Viasov Equation Our physical picture of this BGK mode, both the nonlinear version (6.206) and the linear limit (6.210), is as follows. A spatially periodic potential exists. The ion beam is accelerated through regions of large negative potential and thus has a lower density there, while the electron beam is decelerated in regions of large negative potential and thus has a higher density there. The net result is a negative net charge in regions of negative potential, of exactly the right amount to produce the negative potential./The opposite argument produces the regions of positive potential. The potential and densities thus have the phase relationships shown in Fig. 6.37. The important point is that this physical process works not only in the linear regime of (6.210), but also the nonlinear regime of (6.206). In the preceding discussion we assumed that the ion and electron velocities were large enough so that none of them were trapped in the electrostatic potential wells. In other words, we had ion energy > eg(x) for all x, and electron energy > — e¢(x) for all x. We can also consider the case where some of the electrons or ions are trapped in the potential wells. Amazingly, it turns out that almost any potential 4(2) can be constructed by. choosing appropriate distributions of trapped electrons, untrapped electrons, trapped ions, and untrapped ions. ot) 4 Fig. 6.37 Phase relationships among electrostatic potential, electron density, ion density, and net charge density in the BGK mode of Fig. 6.36. BGK Modes 119 For this discussion it is convenient to think in terms of distribution functions that depend on energy\@Zxather than distribution functions that depend on velocity v. With the substitution H = Yem,v? + g,y, we have dH = m,u dv = m,{(2/m,) (H — 4,9)]'?, dv = (2m(H — q,)]' du and = —__ SAH) dH Sv) dv = flvH)] dv = m(H — 4.0)" (6.212) Equation 6.212 applies to particles with positive speeds. If we assume for conve- nience (we do not need to do this in general) that there are equal numbers of left- -going and right-going particles, then we have mo) = [odoinn = 2 [ doftea) a LAx,H) =2 [at i aa where the lower limit. on the energy integration must be taken to be-9,;'no particle can have energy less than g,¢(x) for then its velocity would be imaginary since ‘hm,v? = H — gy. Consider a periodic potential, as shown in Fig. 6,38. Then any ion with total energy # less than.¢gqyz Will be trapped between the potential hills; the ion with energy egy oscillates forever on the solid line ghown, An ion with total energy H greater than egms will travel forever to the left or to the tight. The electrons, however, see the potentiafupside down, because of their negative charge. Thus, any electron with energy tess than —e¢mio will be trapped between the potential minima, while electrons with energies greater than'-egnistravel forever to the left or to the right. Suppose we are given a completely arbitrary periodic potential g(x), a given distribution fH) of ions (both trapped and untrapped), and a given distribution F(H), H > — egain. Then it turns out that the distribution of trapped electrons SAH), —eg < H < —emin, can always be chosen so that. Poisson's equation is satisfied and the given potential g(x) is indeed produced. Detaits of this calcula- tion can be found on p. 436 of Ref. [13]. Note that f, can be different for v <0 and v > 0. There are many practical applications of BGK modes, including the nonlinear stage of a Landau damped Langmuir wave (Section 6.8), and the theories of shock (6.213) O(n) SAN Vg Fig. 6.38 Periodic potential of a BGK mode that contains trapped and untrapped par- ticles. rin 120 Vlasov Equation waves and double layers. By contrast, the theory of Case-Van Kampen modes, to be presented in the next section, has very little practicat application; nevertheless, this theory teaches us much about the analytic structure of the Vlasov equation. 6.14 CASE-VAN KAMPEN MODES In Sections 6.1 to 6.5 we studied Langmuir waves and Landau damping by linear- izing the Vlasov and Poisson equations, eliminating the perturbed distribution function f,, and solving the initial value problem for the electric field £. This'led us to look for normial modes of E, which were found by setting the dielectric function ¢(w,k) equal to zero. In this way we found one normal mode for every value of k, with w(k) ~ w(1 + %kA2) + iy where y is the Landau damping rate. There is another way to approach this problem, and that is to eliminate E and look for normal modes of f,. In this way we find the Case-Van Kampen modes (17, 18). The Viasov-Poisson system is ast vas — = as=o (6.214) and 4,E = 4ne [% te aso] (6.215) where one-dimensional variations are considered, v JF. v,, and /(v) is the one- dimensional electron distribution. Equation (6.2/4) refers to electrons only, the ions being fixed (m, — ©). Linearizing (6.214) and (6.215) exactly as in Section 6.3 we find - E 8 I fit vd, f = BE (6.216) and 8,E(x,1) = — 4re if dv f,(%,v,1) (6.217) where fo(v) = @g(u), [= dv g(v) = 1. We can now look for norma! modes off, that have the spatial and time dependence exp(—iwt + ikx). Note that this is not the same procedure as used previously in Section 6.4 for the electric field E. There, we assumed only the spatial dependence exp(ikx), and used Laplace transform techniques to consider the complete time evolution. At late times, we found that only the normal modes given by the zeros of e(w,k) were important. Here we are looking immediately for normal modes in space and time. We are not considering an initial value problem; the connection between the normal modes found here and an initial value problem must be established separately. Looking for sotutions f,(x,»,) = 7,(v) exp(—iwt + ikx),E = Ey exp(~iwt + tkx), (6.216) and (6.217) become myeEy Og are ade (6.218) (iw + iku) fF, = (Case-Van Kampen Modes 121 and Ey = ee aL dvj(v) (6.219) from which we can eliminate £, to obtain (» - ar = 25 3 € griwrtike (6.229) is the electric field associated with the normal mode (6.222) These normal modes are peculiar, both mathematically and physically. Math- ematically, we have f\(v = @/k) — © because of the 6-function in (6.222). But it is not consistent to linearize the Vlasov equation with f = f, + f, and then find f, infinite! Physically, (6.222) says that we must have a finite number of particles per unit spatial volume with velocity exactly equal to w/k, which is impossible to do. We conclude that the individual Case-Van Kampen modes as given by (6.222) are not physically relevant. What then is the importance of the modes in (6.222)? The importance of the modes in (6.222) lies in the possibility of creating a physically and mathematically acceptable disturbance by adding up many such modes. Consider a fixed wave number k. Since the basic linearized Vlasov-Pois- son system is indeed linear, we may construct a solution by taking any linear combination of the solutions in (6.222). The general solution is Airut) = et [deve F(yae(w) (6.230) where we label the normal modes of (6.222) by their frequency w, and ¢(w) is an arbitrary weighting function. For sufficiently well behaved c(w), the singularities in F(v,w) will be smoothed out, and f,(x,,f) will be a mathematically and physi- cally nice function. For a given initial condition f,(x,v,? = 0), the function c(w) must be chosen such that (6.230) yields the correct solution at 1 = 0. Inserting (6.222) for f,(v) into (6.230), we find fileut) = ay Pf dw + ellxk ete" e(w = ku) ew) (@/k) a Ketter lw = ku) P ede (6.231) Case-Van Kampen Modes 123 where we have used &(v — w/k) = ké(w — kv). Consider the middle term in (6.231). This term does not damp away at jate times, but rather oscillates in velocity faster and faster with increasing time, as shown in Fig, 6.39. This behavior is due to the free streaming of particles, and would occur for f,(x,v.f) even if the charge of the electrons were zero. For example, consider the linearized Vlasov equation (6.216) when the charge is zero; a, + ikuf, = 0 (6.232) where spatial dependence e“* is assumed. The solution of (6.232) is SQ) = fiw = djeweer (6.233) which becomes more and more pathological with increasing time. Physically, a small number of collisions would wipe out this behavior at late time. Returning 1o the general case with a nonzero charge, we can ask; Does the electric field behave more reasonably? Yes it does. From Poisson's equation, “4 are Bx) = f£ dv f,(v,x.1) (6.234) E(x,t) = 4mei e** f dv e~*¢(w = ku) (6.235) where the first and last terms on the right of (6.231) cancel upon integration. Ale=0, », £=0) f(x=0, v, £>0) Fig. 6.39 A portion of the distribution function of a Case-Van Kampen mode at two different times. 124 Vlasov Equation EXERCISE Verify this. The right side of (6.235) is in the form of a Fourier transform, It turns out that when c(w) is correctly chosen to represent the initial f,(x,v,f) through (6.230), then (6.235) is exactly equivalent to the expression (6.35) together with (6.34). Thus, the right side of (6.235) would produce transients, as before, together with the correct Landau damped normal modes. Thus, there is complete agreement between the norma) mode picture for E(x,t) of Sections 6.4 and 6.5, and the normal mode picture for f\(x,v,F) of the present section. The latter is somewhat more complete, since f,(x,v,t) determines E(x,t) and therefore determines its own future at all times, while E(x,1) could be produced by many different functions f,(x,v,1) and therefore does not determine its own future at all times. In practice, either ap- proach may be used because E(x,’) does pick out the slowly Landau damped normal modes of the system at late times, after the transients have died away. This brings us to the end of our discussion of the Vlasov equation. We recall that the Vlasov equation is obtained as an approximate theory from the Klimon- tovich and Liouville approaches of Chapters 3 to 5 by neglecting the physics of collisions. In the next two chapters, we take moments of the Vlasov equation to obtain even simpler and more approximate theories of a plasma; these are the fluid theory and magnetohydrodynamics. REFERENCES [1] A.A. Vlasov, J. Phys. (U.S.S.R.), 9, 25 (1945). (2) L. D. Landau, J. Pays. (U.S.S.R), 10, 25 (1946). [3] J.D. Jackson, J. Nucl, Energy, Part C, 1, 171 (1960). (4) L. D. Landau and E. M. Lifshitz, Electrodynamics of Continuous Media, Pergamon, Oxford, 1960. [5] 5. Dawson, Phys. Fluids, 4, 869 (1961) {6] 1. B. Bernstein, J. M. Greene, and M. D. Kruskal, Phys. Rev., 108, 546 (1957). (7) T.M. O'Neil, J. H. Winfrey, and J. H. Malmberg, Phys. Fluids, 14, 1204 1974), [8] G. J. Morales and J. H. Malmberg, Pays. Fluids, 17, 609 (1974). [9] T. O'Neil, Phys. Fluids, 8, 2255 (1965). [10] C. S. Gardner, Phys. Fluids, 6, 839 (1963). [11] R. Penrose, Phys. Fluids, 3, 258 (1960) [12] P. Noerdlinger, Phys. Rev., 118, 879 (1960). {13]_N. A. Krall and A. W. Trivelpiece, Principles of Plasma Physics, McGraw- Hill, New York, 1973. [14] T. H. Stix, The Theory of Plasma Waves, McGraw-Hill, New York, 1962, Chs. 8 and 9. [15] G. Bekefi, Radiation Processes in Plasmas, Wiley, New York, 1966. Problems 125 (16) I. B. Bernstein, Phys. Rev., 109, 10 (1958). [17] K. M. Case, Ann. Phys. (N.Y.), 7, 349 (1959), [18] _N. G. Van Kampen, Physica, 21, 949 (1955). - PROBLEMS 6.1 Resistive vs. Reactive Instabilities The type of instability found in (6.52) when d,g(¥)ly-u/e > 0 is often called a “resistive” instability; these are instabilities to which (6.42) and (6.43) apply. Show why (6.42) and (6.43) do not apply to the dielectric function @) = 1 + w,?/w*, which yields a “reactive” instability. 6.2 Ton-Acoustic Waves The dispersion relation (6.21) has a branch describing ion-acoustic waves as well as one describing Langmuir waves. We consider g(1) given by two Maxwellians as in (6.24), with T; << T,, and look for a wave with phase speed such that u << w/k << v, For the ion contribution, expand the denominator in (6.21) as in (6.26). For the electron contribution, approximate the integral by ignoring w/k << uy in the denominator. Then solve the dispersion relation e(k,w) = Oto find » = w(k) for ion-acoustic waves. . 6.3 Electrostatic Waves The dispersion relation (6.21) for electrostatic waves must be solved using the Landau contour of Section 6.4; alternatively, the integral can be evaluated for w; > 0 and the result analytically continued to w,; < 0. Evaluate (6.21) and find the norma! modes w(k) for the following distribution functions g(u). (a) Cold plasma, g(u) = d(u). (b) Cold beam, g(u) = &(u — 4). (©) Square distribution, g(u) = (2c) for |u| c; cis a real positive constant. (@) Cauchy distribution, g(u) = (¢/m)(u? + c)"!. (Why can there never be a true Cauchy distribution?) 6.4 Ton-Acoustic Wave Energy Apply the wave energy formula (6.72) to the ion-acoustic dielectric function de- rived in Problem 6.2. Is most of the energy in electric field energy or in particle energy? Explain physically and in detail why this is so (see discussion in Chapter 7). 65 Two Drifting Cauchy Distributions Consider a Vlasov equilibrium consisting of infinitely massive ions and two coun- terstreaming Cauchy distributions (see Problem 6.3) such that. 126 (a) (b) ©) 6.6 (a) (b) Vlasov Equation Dee! 1 il . sO = 7 Lapa te t ware Sketch g(u). Apply Gardner's theorem to show that stability is guaranteed fora < A/V/3. Apply the Penrose criterion to show that this equilibrium is stable fora < A, and unstable for a > A. Isotropic Stability Consider a plasma with infinitely massive ions and an electron distribution function that is the surface of a sphere in three-dimensional velocity space, FAV) = C WU — vg) = C Oflu? + vy? + v7)? — vo] Calculate g(u), sketch it, and use Gardner's theorem to show that this distri- bution is stable to electrostatic perturbations. Consider any isotropic distribution function LAY) = F.%) Use Gardner's theorem to show that such a distribution is stable to electro- static perturbations. CHAPTER Fluid Equations 7.1 INTRODUCTION ‘There are many phenomena in plasma physics that can be studied by thinking of the plasma as two interpenetrating fluids, an electron fluid and an ion fluid. In this approach, it is not necessary to consider the fact that each species consists of particles with different velocities. The advantage of this approach is its simplicity; it leads to equations in three spatial dimensions and time rather than in the seven-dimensional phase space of Vlasov theory (Chapter 6). The disadvantage of this approach is that it misses velocity-dependent effects such as Landau damping. In this section, we introduce the fluid equations heuristically, for the benefit of those readers who have not yet studied Chapter 6 on Viasov theory. In the next section, we present a more rigorous derivation of the fluid equations from the Vlasov equation. The first equation of fluid theory is the continuity equation, which expresses the fact that the fluid is not being created or destroyed, so that the only way that the fluid density n,(x,1) of fluid species s can change at a point x = (x,y,z) is by having a net amount of fluid enter or Jeave a small spatial volume including that point. The density n, is the number of particles of species s per unit volume. To every element of fluid there corresponds a velocity vector V,(x,1) that gives the velocity of the fluid element at the point x at time ¢. Mathematically, the continuity equation for fluid species s is 4,n,(X,1) + V+ (AV) (7.1) where V = (8,,0,,0,) is the usual gradient operator in three-dimensional configu- 128 Fiuld Equations ration space. A derivation of the fluid continuity equation can be found in most undergraduate mechanics books (see, for example, Ref. [1]). The second equation of fluid theory is the force equation, which is simply Newton’s second law of motion for a fluid. This can be written for fluid species s as nym,V (x,t) = F,(x,0) : (7.2) where F,(x,1) is the force per unit volume acting on the fluid element at position x at time f. The time derivative on the teft of Newton’s law refers to the fluid element as an entity and therefore must be taken along the orbit of the fluid element. Thus, V1) = a,V, + (#) orbit av, + (V, + UV, (7.3) On the right side of (7.2) are all of the forces that act on a fluid element. One such force is the pressure gradient force. A fluid of charged particles has a pressure P.(x,1) = n,(x,t)T,(x,1) and an associated force per unit volume —VP,. Another force is the Lorentz force per unit volume, q,7,(x,t)E(x,1) + (9,/c)n,(x,1)V.(x,t) B(x,t). With these forces, Eq. (7.2) becomes il nm aN, + nm, VV, = ~ UP, + a.nB + nv, xB (7.4) or a1 aV. + ¥,-° VV, = ve, +H e+ vee | (75) ™, me which can be thought of as the force equation per particle, The fields E(x,4) and B(x,t) are the macroscopic fields (those which would be measured by a probe), as discussed in Chapter 3. With the given fluid quantities, the total charge density p is defined by pt) = YL gnOut) (76) while the total current density J is defined by ID) = Y anv.) 7) . When combined with Maxwell’s equations V + E(x,t) = 4ap (7.8) V+ Bx = 0 (79) Vx Ea.) =- + 3B (7.10) vx Ba = “Za + Log a.) the fluid equations provide a complete, but approximate, description of plasma physics. A more careful development of the fluid equations from the Vlasov Derivation of the Fluid Equations from the Vlasov Equation 129 equation is provided in the next section. The reader who has not yet studied the Vlasov equation (Chapter 6) can proceed directly to Section 7.3. + 7.2 DERIVATION OF THE FLUID EQUATIONS FROM THE VLASOV EQUATION Except for the neglect of collisions, the Vlasov equation (Chapter 6) is an exact description of a plasma. By taking velocity moments of the Vlasov equation in seven-dimensional (x,v,£) space, an infinite hierarchy of equations in four-dimen- sional (x,/) space can be derived. When an appropriate truncation of this infinite hierarchy is carried out, the standard fwo-fluid theory of plasma physics is ob- tained. This procedure is reminiscent of the truncation of the BBGKY hierarchy in Chapter 4 that led to the plasma kinetic equation and thence to the Vlasov equation. The Vlasov equation (6.5) is ® ® ® ® hom) tv Wht EFT XB DL=0 63) We use the normalization nxt) = f dv £0%.0) (7.12) and note that the fluid velocity V, is V(x.) = ase vhley,0) (7.13) The first fluid equation (the continuity equation) is obtained by integrating (6.5) over all velocity space (i.e., we first multiply by “unity”). The first term yields 8,n,(x,t). The second term is O=fav-y The third and fourth terms vanish upon performing the velocity integration. =vVy “fav vf, = Vy + (nV,) (7.14) EXERCISE Show the above. The result is the exact continuity equation On (x,t) + Vy (nV) = 0 (7.15) which agrees with (7.1). (Except in this section, V, is denoted by V in this chapter.) The force equation is obtained by multiplying (6.5) by v and integrating over all velocity space. This yields 0} @® ® ® F fart far v4+& farv[(e+ty xa). ws, (7.16) 130 Fluid Equations In term Q, we have a,(1,V,) by (7.13). In @, we perform the manipulation w - V, L=v- Wwf) = Vx + (w/,). Since f,(x,v,1) is a probability distribution, the ensemble average of any quantity is fav g, 7 a.) (g) favs, Thus, term @ is Va of dvwi, = Ve + Culv ») (7.18) Term @ is easily evaluated by an integration by parts, yielding —(9,/m,)En,, EXERCISE Verify this result for at least one component of the combination vE- V,. In term @, it is useful to move Vy, to the left, obtaining Vy - [(v * B) f,J; an integration by parts then yields eae f ade x @® HE dv(v X B)f, me Ns B) (7.19) where we have evaluated each component in (7.19) using (7.13). Combining all terms, (7.16) becomes nV.) + V > (n,Q) n, (e (7.20) which is the fluid force equqtion for species s. Multiplying through by the mass m,, we see that each term has units of (force/volume). Equation (7.20) is also called the momentum equation, since it determines the time rate of change of momentum per unit volume. Note that the continuity equation (7.15) for a, involves the function V,, and the force equation (7.20) for V, involves the function (v y). It is clear that every equation for m factors of y will involve a term with # + | factors of v. Thus, to obtain a complete description of a plasma, we need an infinite number of moment equations as derived from the Vlasov equation. This is equivalent to replacing the seven-dimensional Vlasov equation by an infinite number of four-dimensional fluid equations. In practice, we seek to truncate this series of equations by using a physical argument to evaluate the term with n + 1 factors of y, rather than using the fluid equation for that term. For example, we shall use physical arguments to evaluate the (v v) term in (7.20), so that the force equation (7.20), the continuity equation (7.15), and Maxwell’s equations become a complete description of the plasma. Jn detailed descriptions of plasmas found, for example, in magnetic con- finement devices and in the solar wind, the fluid equation for (v v) is used and physica] arguments are used to evaluate terms with three components of velocity [2]. The equation for the time derivative of (v v) is known as the energy equation There are various circumstances where it is easy to evaluate (v v), For example, suppose the species is cold, so that all particles have the same macroscopic velocity V,. Then f,(x,v,t) = n,(x,)a(v — V,); thus Derivation of the Fluid Equations trom the Viasov Equation 131 Wy) = at dvnw8v ~ V.) = V,V, (7.21) EXERCISE Verify (7.21), recalling that 6(v — V,) = &(v, — V,,) &v, — Vy) 8u. — Vee). Another case is where the distribution function /,(x,v,/) is isotropic at each point in space. Then with UpUy aoe UsU, wy= (fun yy, oy (7.22) Vive UB Udy we have V, = 0, and upon taking the average, all of the off-diagonal terms in (7.22) vanish. EXERCISE Prove this case. The diagonal terms are (v,2) = (v2) = (v,2) = v,? where u,(x) is the therma! speed. Equation (7.22) becomes wy) = v(x) T (7.23) where T is the unit tensor, and we take into account the possibility that the temperature (TJ, = m,v,?) is spatially dependent. The second term in (7.20) than takes the form Vs (nv) = V+ (nuel) = V(rQv2) = VP,/m, (7.24) where the pressure P, = nymu2 = n,T, More generally, we might have a distribution that has a net velocity V, in a certain direction, and has an isotropic velocity distribution in the frame moving with velocity V,; therefore (vy, — Vy.)°) = (vy — Vo?) = (ve — Va) = ve = P,/m,n,. Then Pg = VY. + (7.25) and Vi (advy)) = V+ (VV) + 4 VP, = (V+ VV.) + (Vs + V)(aV.) + z VP, (7.26) EXERCISE By writing out components, or by any other method, justify the manipulations in (7.26) The force equation (7.20) becomes, multiplying by m,, A,(mn,V.) + (V > V)nnV) + (V. + VXmn,V,) = -VP, + 4m, (E tivex 8} (7.27) 132 Fluld Equations Equation (7.27) can be simplified by subtracting the continuity equation (7.15), multiplied by V,, from the left side. We find msn, aN, + mandVs Os = —VP, + an, (E+ 2 v, x BY (7.28) in agreement with (7.5). When combined with Maxwell’s equations, and when some means is found for describing the pressure P,, Eqs. (7.15) and (7.28) provide a complete description of fluid plasma ‘behavior. We shall find several different means for describing the pressure. For variations in one direction only, the pressure is P, = n,T,=1,m,(v”) where we evaluate (v*) along the direction of variation. If we are dealing with a motion, a wave for example, which is slowly varying compared to the equilibra- tion time of species s, we might have isothermal behavior so VP, = V(a.T,) = T,Va, (7.29) On the other hand, a rapidly varying compression may involve adiabatic motion, so UP, = V(n,T,) = n,0T, + Tn, = y.T.Vn, (7.30) where y, = (2 + D)/D is the so-called “ratio of specific heats.” Here, D is the number of dimensions that share in the increased temperature, and it is assumed that the motion in (7.30) involves only small departures of the density and temper- ature from their unperturbed values n,, T,. In succeeding sections on wave mo- tions in fluid plasma, we shall apply these ideas. EXERCISE Verify (7.30) using the basic ideas of adiabatic compressions. 7.3. LANGMUIR WAVES Now that we have developed a complete description of a plasma, in the form of the two-fluid equations, what can we do with it? The first thing we can do is to study the various kinds of waves that can propagate through a plasma. Waves are very important. They propagate energy from one part of a plasma to another. They send information out of the plasma that enables an external observer to know what is occurring inside. They can become unstable, growing as they propagate, to such large amplitudes that they disrupt the confinement of a plasma. Our first example of a wave is a very simple one, the electron plasma wave, or Langmuir wave (also called a space-charge wave). Suppose the ions are infinitely massive, so that they do not contribute to the motion, but have a fixed particle density vo and a fixed charge density eny. Then we need only three equations to describe the electron motion: the electron continuity equation, the electron force equation, and Poisson’s equation. These are (in one dimension, with B, = 0) an. + (mV) = 0 (7.31) mudaV. + V.aV.) = bP. — enE (7.32) Langmiur Waves 133 and 0,E = 4me(m — 1.) (7.33) We seek solutions to (7,31) to (7.33) in the form of small amplitude waves, where the electric field has a sinusoidal spatial variation. In order to look for such small amplitude waves, we first linearize (7.31) to (7.33). With n. = ty +m, (7.34) E=& : (7.35) and veo= vy (7.36) we first neglect the pressure P,, assuming that the electrons are cold. Then the only zeroth order contribution from (7.31) to (7.33) is amy = 0 (7.37) which is trivially satisfied. The first order terms are Gn, + no Av, = 0 (7.38) MINo dv, = —eNgk, (7.39) and a,E, = —4aen, (7.40) ‘These equations are now linear, and we may look for wave solutions in which each variable has the form cos (kx — wt + 6), E,(x,1) = EB, cos (kx — wt) (7.41) ni(x,t) = A, cos (kx — wt + @,) (7.42) and v(x) = 8, cos (kx — wt + 6,) (7.43) where £,,, #,, and %, are real constants and 6,, 6, are possible phase shifts. It turns out that it is very awkward to use sin’s and cos’s, and it is very convenient to use solutions that vary as exp.(—iwt + ikx). We may do this by noting that if (7.41) to (7.43) is a solution of the linearized equations (7.38) to (7.40), then the expression obtained by giving each of (7.41) to (7.43) a phase shift of 1/2 will also be a solution, where sin replaces cos in (7.41) to (7.43). Any linear combination of these two sets of solutions will also be a solution; in particular, [cos solution] + i[sin solution] is a solution, of the form E\(x,t) = E, exp (~iwt + ikx) (7.44) n,(x,1) = A, exp (i8,) exp (iat + ikx) (7.45) and u\(x,2) = 5, exp (i6,) exp (—iwl + kx) (7.46) If we next absorb the phase factors exp (i8,) and exp (i8,) into new complex con- stants 7, = A, exp (i6,), 0; = 0, exp (i6,), we have 134 Fluid Equations E,(x,t) = E, exp (—iwr + ikx) (7.47) n,(x,1) = it, exp (~iwt + ikx) (7.48) and vi(%t) = d, exp (iat +.1kx) (7.49) After we have obtained the solutions (7.47) to (7.49), we can add them to their complex conjugate to obtain the physically relevant real solutions, if desired. EXERCISE Why is the complex conjugate of (7.47) to (7.49) also a solution? Inserting the assumed wave solution (7.47) to (7.49) into the linearized equa- tions (7.38) to (7.40), one obtains ian, + iknd, = 0 (7.50) —iomangy, = eng, (7.51) and . ikE, = —4rren, (7.52) where we have divided each side by exp (fot + ikx). Solving (7.50) for ¥, in terms of f;, and solving (7.52) for 7, in terms of £,, and inserting in (7.51), we find wm, £, Tame = Ey (7.53) Upon dividing by the constant E,, we see that rene = ee = a (7.54) or £ w, (7.55) for our cold plasma oscillations; thus the wave frequency is just our old friend the electron plasma frequency. EXERCISE If we had kept the ion component with mass m;, can you guess what the wave frequency would be? Thus, we have shown that any expression of the form (7.44) to (7.46) is a solution of (7.38) to (7.40), provided the frequency is given by (7.55). Note that this is true for arbitrary wave number k. EXERCISE Using (7.50) to (7.52), determine the phase shifts between E,, ,, and v,. Choosing a value of k, add (7.47) to (7.49) to its complex conjugate to obtain real solutions, and sketch these solutions with their appropriate phase shifts. The expression (7.55) for wis called a dispersion relation, because it is supposed to represent the relation between frequency w and wave number k, In this case the dispersion relation is trivial, because it does not depend on k. The group velocity Langmiur Waves 135 do _ Y= a =0 (7.56) does not depend on wave number k, so these waves are dispersionless. An initial wave packet will not propagate, but merely oscillates forever at frequency w = w,. Thus, the physics of these waves is as simple as the physics of the oscillating slabs used to derive the plasma frequency in Chapter 1 It is always useful to sketch dispersion relations. In this case, the sketch of (7.55) is simple, consisting of two straight lines at w = + w,, as shown in Fig. 7.1. The most serious assumption in our derivation of the cold plasma waves is the neglect of the pressure term in (7.32). Let us repeat the derivation, including the pressure P,. We have VP. = Vin.) = V(t + mMXTo + T,)) = VT, + T,Vn,, keeping only first order terms. Now in order to relate the first order temperature change T, in the wave to the first order density change ,, we must go outside the fluid theory. We consider long wavelength waves, such that a typical electron travels only a fraction of a wavelength A in one wave period; then the compression of the wave will be an adiabatic one. Thus, the assumption vu? << A, or w/k >> v, (7.57) leads us to consider adiabatic compressions. If we further assume that the collision frequency is small, ye << wo (7.58) then the change in temperature during the compression along the direction of wave propagation will not be transmitted to the other two directions. We conclude that our compression is a one-dimensional adiabatic compression; which means that VP, = nVT, + TyVn, = 3T)Vn, (7.59) by the expression below (7.30). With the expression (7.59) for VP,, our previous derivation goes through with the addition of one term; thus (7.50) to (7.52) are replaced by Fig. 7.1. Dispersion relation for electron plasma waves when the pressure is ignored. 136 Fluld Equations Fig. 7.2. Langmuir wave dispersion diagram. ih, + ikngs, = 0 (7.60) siwmNyd, = —eng£, — ik3T yt, (7.61) ikE, = —4men, (7.62) Solving (7.60) for 3;, solving (7.62) for 7i,, and inserting in (7.61), one finds (w? — w2 — 3k°T)/m)E, = 0 (7.63) or wo = w2 + 3K? To/m, = w2 +.3k v2 (7.64) which is the famous Langmuir wave dispersion relation (see Eq. 6.28). Since we have used the assumption (7.57), the dispersion relation (7.64) has a limited range of validity, restricted to Om Me ue 0 the pressure acts as an additional restor- ing force which gives the wave a higher frequency 7.4 DIELECTRIC FUNCTION Itis often useful to draw an analogy between a plasma and a dielectric medium. Recall that in ordinary dielectric theory, we are able to replace Poisson's equation for charges in a vacuum, Dielectric Function 137 vV-E=4np (7.69) by v-D=0 (7.70) where D=€-E a7 Here © is the dielectric tensor, the properties of the medium have been incorporated into the displacement D, and we assume there are no additional free charges. This same operation can of course be performed for a plasma as well as for a dielectric medium. In one dimension, and assuming plane wave fields, Eq. (7.69) is i ikE = Aap (7.72) so that if we can write p in terms of E it will be easy to identify the dielectric function: 4 ik (2 ~“P) = kez =0 (7.73) For cold plasma waves this has been done in (7.50) to (7.52), which are easily written in the form 2 ik(1 -2r)E=0 (7.74) so that the dielectric function is 2 da) = 1— _ (7.75) and we notice that the dispersion relation (7.55) is precisely equivalent to equating €() to zero, dw =-0 >w=te, (7.76) Thus, we see that the normal modes of a plasma are obtained from the zeros of the dielectric function (see Chapter 6). EXERCISE Verify (7.74). Ina similar fashion, because (7.60) to (7.62) for Langmuir waves can be written in the form 2 ik [} ~ ates | E=0 (1.77) we identify the dielectric function wk) = 1 ~ oa (7.78) the zeros of which yield the normal modes (7.67): wk) = 0 = afk) (7.79) ak) = (1 + 372)? (7.80) The dielectric function has been studied in more detail in Chapter 6. 138 Fluid Equations 7.5 ION PLASMA WAVES In Section 7.3 we studied high frequency electron plasma oscillations, with fre- quency w near the electron plasma frequency w,. For these waves, the ion motion is negligible and irrelevant. Here, we consider low frequency waves, w S w;, where the ion motion dominates the wave physics. (In an unmagnetized plasma, the words “low frequency” refer to w < uw, while “high frequency” refers to w = w,. In a magnetized plasma, we often have the ordering 2; << w, << a < |O,|. (Why?) Then “low frequency” means @ < 0,, and other frequencies are called low or high depending on what they are being compared to.) To include both electron and ion physics, we need five fluid equations: electron and ion continuity, electron and ion force, and Poisson's equation. These are, from (7.1) and (7.5), an, + A,(nV,) = 0 (7.81) mn, BV, + mnVe IV, = —d,P, — en£ (7.82) aa, + (nV) = 0 (7.83) mn, AV, + mnV, dV, = ~8,P, + enE (7.84) and 6,E = 4me(n, — m,) (7.85) This set of equations is not as bad as it looks. First, because we intend to linearize, all V a,V terms disappear. Second, all pressure terms can be written 0, P., = Tes 4,n,,;, Where temporarily we do not specify the coefficients y,,, and T,; are the unperturbed electron and ion temperatures. Then linearizing (7.81) to-(7.85) with n, = no + n,,, and all other quantities first order, we obtain Gyig + no AY, = 0 (7.86) mjtiy 8,V, = — eT eM — eMoE (7.87) ty + ny ,V; = 0 (7.88) ming AV; = — iF; Aut + eMgE (7.89) and aE = 4re(n, — 1) (7.90) Before solving (7.86) to (7.90), let us guess the properties of the wave we are looking for. The ions will have a sinusoidal density perturbation as shown in Fig, 7.3. Since the frequency is very low, the electrons see an almost static ion density perturbation, and they will try to obtain the same density in order to prevent huge electric fields. However, since the electrons are flying about very fast, the attempt Fig. 7.3. Electron and ion density perturbations in an ion plasma wave. lon Plasme Waves 139 to exactly cancel the ion charge distribution will not totally succeed; rather, the electrons try to smear themselves out more smoothly. Thus, the electron density perturbation is slightly smaller than the ion density perturbation, and the resulting density difference produces the electric field of the waves. EXERCISE Draw the electric field produced by the densities in Fig. 7.3. We have seen the tendency for this cancellation of electron and ion charge before, in Problem 1.4, “Plasma in a gravitational field.” This important property is called quasineuirality, and it is found in almost all low frequency plasma behavior This discussion encourages us to took for a solution of (7,86) to (7.90) with my ~ ny. We therefore ignore Poisson's equation; and we find from (7.86) and (7.88) that we must have V, = V;. We eliminate the electric field by adding (7.87) and (7,89) to obtain ane Ox Next, we eliminate the velocity V, by taking the time derivative of (7.86) and inserting the result in the spatial derivative of (7.91), to obtain = —(yT. + yi) (7.91) ang ang (m. + my) Be = T+ eT) St (7.92) or, neglecting m, compared to m,, & an, Gas Se (7.93) where we have defined the sound speed nya os cao ¥ : (7.94) Assuming a plane wave solution of the form exp (—iwt + ikx), (7.93) yields the ion-acoustic -'ispersion relation Re? (7.95) The name ion-acoustic arises from the similarity between the dispersion relation (7.95) and the equivalent relation for sound waves traveling through a gas. It is difficult to determine the regime of validity of (7.95) from the foregoing discussion. We do know, however, that we have neglected the difference (;, — ”.,), Which by Poisson’s equation is proportional to 0,E ~ KE. We thus expect that (7.95) is limited to small k; we shall see in a moment that this is so. What shall we take for +, y;? In practice, this depends on the region of density, temperature, and wave number in which we are working. It may be the case that the ion motions are adiabatic in one dimension, soy; = 3. It may also be that collisions are important enough to redistribute the wave compressional energy in three dimensions, so that y, = (D + 2)/D = 5/3 for adiabatic compressions in three dimensions. As for the electrons, it is the case that a typical electron travels 140 Fluid Equations many wavelengths in one wave period; that is, the distance traveled in one period ufo ~ v/ke, >> ke", since > Ge sateen) >? Thus, the electrons are communicating over many wavelengths during one wave period so that they remain isothermal; we therefore choose the iosthermal y, = 1. When T, >> T,, we then have the very simple and easy to remember formula (T./m)? | (7.96) In other words, the sound speed is the thermal speed that the ions would have if they had the electron temperature. Let us now return to a more exact solution of (7.86) to (7.90) that does not assume quasineutrality. Because the electron mass is very small, we ignore the electron “inertia” term on the left side of (7.87), upon which (7.87) yields On, _ —eMy i ae Tr E (7.97) For the ions, we eliminate V, from the spatial derivative of (7.89) by using the time derivative of (7.88), to obtain en, _ any OE — my SP =~ vs Gt t em Fy (7.98) Looking for plane wave solutions to (7.90), (7.97), and (7.98) (or Fourier trans- forming, if you like), we find tkE = 4re(n, — 1.) (7,99) = Ze thang = Sao E (7.100) and (ot ny = 28 ike (7.101) mm Inserting (7.100) and (7,101) into (7.99) we find je ae TE eI 7 i [ ~ em, * mim | aed (7.102) from which we identify the dielectric function wo? oe aa eytym, + Pym, = ° (7.103) the zeros of which yield the dispersion relation «(k). Solving (7.103), we find wk) = (7.104) which is the general dispersion relation for ion plasma waves. EXERCISE Solve (7.103) to obtain (7.104). Electromagnetic Waves 141 Langmuir +/g an J ‘Slope /3 Vy Be Fig. 7.4 Dispersion diagrams for electron plasma waves and ion plasma waves (7, = 0). In the small kA, limit, we regain the ion-acoustic dispersion relation (7.95). We further expect that (7.104) is only valid when the second term on the right is larger than the first. If this were not so, we would have w/k ~ (T/m)\”? ~ v, which would mean that many ions would have speeds of the same order as the wave phase speed. When this is the case, we do not expect fluid theory to be valid; rather, we must use the Vlasov equation to properly treat those particles which can interact resonantly with the wave, Another interesting limit of (7.104) is reached when 7; ~ Oand kA, >> 1; we then find ow? ~ w? (7.105) which are ion plasma waves oscillating at the ion plasma frequency. Because the wavelength is short compared to the electron Debye length, the electrons are incapable of shielding, and we have ions oscillating in a uniform background of negative charge. This is quite analogous to our cold electron plasma oscillations at @ = w,, which are electrons oscillating in a uniform background of positive charge. We can now draw the dispersion diagrams of electron plasma waves and ion plasma waves on the same diagram; we do this schematically for the case 7; ~ 0 in Fig. 7.4. Note that the dispersive (k2A,2) term in the denominator of (7.104) becomes more important at larger k, leading to a transition from the acoustic behavior at small k (7.95) to the oscillations at w = w, (7.105) at large k. 7.6 ELECTROMAGNETIC WAVES The only other waves in unmagnetized homogeneous plasma are electromagnetic waves. We shall find that these waves are high frequency, w = w,, so that we can ignore ion motion. We shall further find that these waves are transverse, k * E = 0 and k - B = 0, so that we can ignore Poisson's equation and the V - B = 0 equation. The fluid equations that we then need are 142 Fluid Equations vxe=—Lop (7.106) vxp= y4s+ on (7.107) c c J=-—en,V, (7.108) mit.dN. + mndV, UV, = — UP, — en, - a Vv. x B (7.109) an, + V- (nV) = 0 (7.110) ‘We have assumed that the wave is transverse, k - E = 0. Ifk is in the £-direction, we may choose E in the -direction; then by Faraday's law (7.106), we have B along 2 (Fig. 7.5). We next assume that there is no zero order component of V_, and that we can neglect the V, X B force; this assumption must be checked at the end of the calculation. We look for solutions that have k - V, = 0; then (7.110) predicts 4n./ar = 0, so n, = mo everywhere and we can ignore VP, in (7.109). With no further assumption, the (V, + V)V, term in (7.109) also vanishes. We have left the equations vxe=-Lap ay i vx B=— ate nV. + 4 ak 7.112) mgt 3, = — enoE (7.113) Taking the time derivative of (7.112) and the curl of (7.111), we eliminate 3,V,and B in (7.112) to obtain ae 2 ~ eo KV x Ey = MH pL SE (7.114) or taking a plane wave solution, V * V x — k, we get xk 28 Fig. 7.5 Vector relationships in an electromagnetic wave in unmagnetized plasma. Electromagnetic Waves 143 (w? — ke? — w2Z)E = 0 (7.115) which yields the electromagnetic dispersion relation Letting the plasma density approach zero we fegain the free space light waves with w = ke. Note the similarity between the electromagnetic dispersion relation and the Langmuir dispersion relation, where c? is replaced by 3v,’. On the same dispersion diagram, the two branches look as shown in Fig. 7.6. Recall that in the theory of optical media (air, water, etc.) it is useful to define an index of refraction pak (7.117) @ for light traveling through the medium, From (7.116) we see that in a plasma the index of refraction is (7.118) According to (7.118), the index of refraction becomes imaginary when w < w, Thus, for real w one obtains imaginary k, corresponding to evanescence. The result of this evanescence is that when an electromagnetic wave impinges on an inhomogeneous plasma, it reflects at the point where w = w,, called the position of critical density. This effect is important in laser fusion and in the interaction of radio waves with the ionosphere (Fig. 7.7). We can now reconsider our neglect of the (g/c)V, * B force in this derivation. By Faraday’s law, B ~ (ck/w)E ~ nE, but 2 <1 always by (7.118), so B

You might also like