Sei sulla pagina 1di 92

Recently Developed Quantum Kinetic

Models and Their Applications to Some


Linear Electrostatic and Electromagnetic
Plasma Modes

A thesis submitted in partial fulfillment of the


requirement for the award of degree of
Doctor of Philosophy in Physics

By

Azhar Hussain
Reg. No. 082-GCU-PHD-PHY-09

Department of Physics
GC University Lahore
God
Grant me serenity to accept the things,
I can’t change;
The courage to change the things, I can;
And the wisdom to know the difference!
Research Completion Certificate
It is certified that the research work contained in this thesis entitled “Recently
Developed Quantum Kinetic Models and Their Applications to Some Linear
Electrostatic and Electromagnetic Plasma Modes” has been carried out by Mr.
Azhar Hussain Reg. No. 082-GCU-PHD-PHY-09 under my supervision at Physics
Department, GC University Lahore, during his postgraduate studies for Doctor of
Philosophy in Physics.

Supervisor

Prof. Dr. G. Murtaza (S.I.)


Salam Chair in Physics
GC University, Lahore

Submitted Through

Prof. Dr. H. A. Shah (S. I.)


Chairperson
Department of Physics
GC University, Lahore
Declaration

I, Mr. Azhar Hussain, Reg. No. 082-GCU-PHD-PHY-09, PhD scholar at the


Department of Physics, GC University Lahore, hereby declare that the matter
printed in the thesis entitled “Recently Developed Quantum Kinetic Models and
Their Applications to Some Linear Electrostatic and Electromagnetic Plasma
Modes” is my own work and has not been printed, or published or submitted as
research work, thesis or publication in any form in any University/Research
Institution etc. in Pakistan or abroad.

Dated: ___________ ___________________

Signature of Deponent
Acknowledgements
The time I spent as a PhD student in GC University Lahore, a historic and
beautiful institute, was memorable! First of all, I express my deep gratitude to
my supervisor Prof. Dr. G. Murtaza for his guidance and care. His thoughtful
suggestions and his sympathetic attitude during the course of my work was a
great help. He was always ready to discuss the mysteries of physics not like a
teacher but like a friend. Secondly, I am extremely grateful to Prof. Dr. G.
Brodin, Umea University Sweden, who constantly guided me till the final write-
up of my thesis. During my stay at Umea, he not only introduced me to the
area of spin quantum plasma but was always there to help me, from my
residence arrangements to pay my bill with me in the post office. My time with
him at Umea, and later in the Dropbox was really fantastic!
My special thanks to Prof. Dr. M. Salimullah and Prof. Dr. N. L.
Tsintsadze who patiently made me mature in the research activity during the
early stage of my doctorate studies, when both were at Physics Department GC
University Lahore, as foreign faculty member. They not only introduced me to
quantum plasmas but also to the philosophy of physics. I have greatly profited
from their comments, hints and precious viewpoints. I am also thankful to Dr.
A. A. Malik for several fruitful discussions on the general physics.
I am gratefully obliged to Prof. Dr. Riaz Ahmad, Chairman Physics
Department (and also to Prof. Dr. H. A. Shah, ex-Chairman) for providing
sound research atmosphere. Cooperation and support of the office secretaries
and office boys from different offices, for documentation and academic purpose,
is also acknowledged.
I thank to all my research fellows and colleagues, especially to, Zafar,
Zeba, Zeshan, Tajammul, Shahid, Rasheed, Noman, Naveed, Muddasir, Jamil,
Javaid, Hafsa, Gohar, Fraz and Ayub. I will always remember them not only for
their valuable discussions and encouragement but also for lavishing dinners,
outings and movies that, I suppose, is a necessary part of the research. Part of
this research work was done during visit to Umea University Sweden (back in
2011-12). So, I would also like to acknowledge a number of colleagues at
Umea. I would especially thank to Dr. Mattias, for his time and several long
discussions on the quantum plasmas, to Jens & Martin, for their help to
understand the minor details of spin quantum kinetic theory, and to Nitin,
Chris, Amol, Mubashar and Zubair, for their nice company at snowy Umea. I
miss you all!

This work was carried out by the financial support of Higher Education
Commission (HEC) under PIN No. 063-00095-Ps3-145. This opportunity,
given by the Government of Pakistan is gratefully acknowledged.

Thanks to my family, especially to my parents for their love and prayers and to
my loving brother Mazhar, for his sincere help and immense care.

Azhar Hussain
Department of Physics
GC University Lahore
February 2014
Dedicated to all those who are toiling for
the peace, progress and prosperity of mankind!
mankind!
List of Publications

Publications included in this thesis

1. New longitudinal waves in electron-positron-ion quantum plasmas, N. L. Tsintsadze, L.


N. Tsintsadze, A. Hussain, and G. Murtaza, Euro. Phys. J. D. 64, 447 (2011).

2. On the kinetic Alfvén waves in nonrelativistic spin quantum plasmas, A. Hussain, Z.


Iqbal, G. Brodin, G. Murtaza, Physics Letters A 377, 2131 (2013).

3. Weakly relativistic quantum kinetic theory for electrostatic wave modes in magnetized
plasmas, Azhar Hussain, Gert Brodin, Martin Stefan, Phys. Plasmas 21, 032104 (2014).

Publications not included in the thesis

4. Modified Debye screening potential in a magnetized quantum plasma, M. Salimullah, A.


Hussain, I. Sara, G. Murtaza, H.A. Shah, Phys. Lett. A, 373, 2577 (2009).

5. Modified screening potential in a high density inhomogeneous quantum dusty


magnetoplasma, A. Hussain, I. Zeba, M. Salimullah, G. Murtaza, and M. Jamil, Phys.
Plasmas 17, 054504 (2010).

6. Reletivistic Bernstein waves in a degenerate plasma, Muddasir Ali, Azhar Hussain, and
G. Murtaza, Phys. Plasmas 18, 092104 (2011).

7. A comparison of parametric decay of oblique Langmuir wave in high and low density
magnetoplasmas, M. Shahid, A. Hussain, G. Murtaza, Phys. Plasmas 20, 092121 (2013).

8. On the ordinary mode and whistler mode instabilities in the degenerate anisotropic
plasmas, Z. Iqbal, A. Hussain, G. Murtaza, N. L. Tsintsadze, Phys. Plasmas 21, 032128
(2014).

9. Damping of whistler waves in non-relativistic spin quantum plasmas


Z. Iqbal, A. Hussain, G. Murtaza, Submitted to Physics of Plasmas (2014).

10. R- and L-Waves in Electron-Positron Spin Quantum Plasmas


M. Shahid, Z. Iqbal, A. Hussain, G. Murtaza, Submitted to Physicsa Scripta (2014).
Abstract
Study of some recently developed quantum kinetic models, by employing these models for linear

wave analysis, is the primary purpose of this dissertation. We have focussed here that how these

newly developed models take into account the quantum effects that arise due to electron degener-

acy, quantum Bohm-potential and spin quantum effects related to electron paramagnetism.

In the first part of the work presented in the thesis, we have studied an electron-positron-ion

degenerate Fermi gas, with or without the Madelung term. We have proposed the existence of a

new type of zero-sound wave that exhibit a damping character in contrast to the usual undamped

zero-sound waves, present in a degenerate electron-ion plasma. In an electron-hole-ion plasma,

we have found new type of longitudinal quantum sound waves that have no analogies in quantum

electron-ion plasma. The excitation of these quantum sound waves by a low-density monoener-

getic straight electron beam has also been examined. In the second part of the work, we have

focussed on the electron spin paramagnetic effects. Employing the non-relativistic spin quan-

tum kinetic theory, we have studied the effect of electron spin on the kinetic Alfvén waves in an

electron-ion plasma. We deduce that the usual kinetic Alfvén waves are modified via spin quan-

tum effects of electrons. The dimensionless parameters that determine the relative importance of

the electron spin become prominent at higher densities. It is found that the kinetic Alfvén wave

frequency decreases due to the electron spin contribution in the kinetic limit while in the inertial

limit they are almost unaffected in a hot magnetized plasma. A drawback of the non-relativistic

spin quantum kinetic theory is that it does not capture spin effects for electrostatic wave modes. To

remove this drawback, we have derived a general dispersion relation for electrostatic modes by em-

ploying recently developed weakly-relativistic quantum kinetic theory of spin quantum plasmas.

This model contains weakly-relativistic spin effects such as Thomas precession, the polarization

currents associated with the spin and also the spin-orbit coupling. It turns out that for strictly elec-

trostatic perturbations the non-relativistic spin effects vanish, and the modification of the classical

dispersion relation is solely associated with the relativistic terms. Several new wave modes appear

due the electron spin effects. As a particular example, we have studied weak-relativistic effects

for Bernstein waves in a fully degenerate plasma and proposed the existence of new waves due to

the electron spin effects, that have frequency close to the Bernstein wave frequency.

i
Contents

1 Kinetic Theory of Quantum Plasmas 1

1.1 Quantum Plasmas. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

1.2 Basic Plasma Regimes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

1.3 Kinetic Theory of Plasmas. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

1.3.1 Classical Kinetic Theory. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6

1.3.2 Quantum Kinetic Theory. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

1.4 Recently Developed Quantum Kinetic Models for Plasmas. . . . . . . . . . . . . . . 10

1.4.1 Kinetic Equation with Quantum Bohm-Potential Effect. . . . . . . . . . . 11

1.4.2 Non-relativistic Spin Quantum Kinetic Model. . . . . . . . . . . . . . . . . . 12

1.4.3 Weakly-relativistic Spin Quantum Kinetic Model. . . . . . . . . . . . . . . . 15

1.5 Summary. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

2 Motivation, Aim and Layout of the Thesis 18

2.1 Motivation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

2.2 Aim and Layout of the Thesis. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

3 New Longitudinal Waves in Electron-Positron-Ion Quantum Plasmas 22

3.1 Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

3.2 Mathematical Model. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24

3.3 Linear Analysis. . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . . . 24

3.3.1 Positron Zero Sound Waves. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

3.3.2 Positron Sound Waves. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

3.3.3 Instability Excitation by Beam-Plasma Interaction. . . . . . . . . . . . . . . 31

3.4 Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33

4 Kinetic Alfvén Waves in Non-relativistic Spin Quantum Plasmas 34

4.1 Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

4.2 Mathematical Model. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

4.3 Linear Analysis. . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . . . 36

4.3.1 General Dispersion Relation of Kinetic Alfvén Waves. . . . . . . . . . . . 39

4.3.2 Kinetic Alfvén Waves in the Kinetic Regime. . . . . . . . . . . . . . . . . . . 40

ii
4.3.3 Kinetic Alfvén Waves in the Inertial Regime. . . . . . . . . . . .. . . . . . . 41

4.4 Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

5 Electrostatic Modes in Weakly-relativistic Spin Quantum Plasmas 44

5.1 Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44

5.2 Mathematical Model. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46

5.3 Linear Analysis. . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . . . 47

5.3.1 Conductivity Components for the Free Current Density. . . . . . . . . . . . 50

5.3.2 Conductivity Components for the Polarization Current Density. . . . . . 52

5.3.3 Solutions to the Dispersion Relation. . . . . . . . . . . . . . . . . . . . . . . . . 52

5.4 Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55

6 Overall Summary and Conclusions 56

7 Appendix 59

A. Derivation of Eq. (3.6). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

B. Selection of the negative root from Eq. (4.26). . . . . . . . . . . . . . . . . . . . . . . . . 61

C. Derivation of perturbed distribution function: Eq. (5.13). . . . . . . . . . . . . . . . . 63

D. Validation of electrostatic limit for waves propagating at an angle to B0 . . . . . . 68

E. Remaining conductivity components for the polarization current density that are

needed. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71

8 Bibliography 73

iii
List of Figures

1.1 Plasma diagram in the logT − logn plane. . . . . . . . . . . . . . 2

3.1 Plot of real frequency of zero-sound waves for arbitrary values of parameter a......27

3.2 Plot of imaginary frequency of zero-sound waves for arbitrary values of parameter

a. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

3.3 Plot of Eq. (3.20) for arbitrary values of me /m p = 0.1 and n p0 /ne0 = 0.01. . . . . 30

3.4 Plot of Eq. (3.22) for arbitrary values of me /m p = 0.1 and n p0 /ne0 = 0.01 . . . . 30

4.1 Plot of the modified kinetic Alfven wave, Eq.(4.26),with spin effects for different

spin temperatures Tsp . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42

4.2 Plot shows the variation of quantum parameter Q as a function of k⊥ ρa . . . . . . 42

5.1 Left: An electron moves in a circular Bohr orbit, as seen by the nucleus. Right: The

same motion but seen from the point of view of the electron. The magnetic field B

is at electrons location and is directed out of the page . . . . . . . . . . . . . . . . . . . . . 44

5.2 Normalized frequency ωn = ω/|ωce | plotted against normalized wavenumber kn ...54

iv
Chapter 1
Kinetic Theory of Quantum Plasmas

Plasma physics has played a key role in understanding the universe and especially extraterrestrial

environments. Formation and dynamics of stars, galaxies and intergalactic medium are mainly

observed due to light signals that we receive by the virtue of charged particle dynamics there. In

laboratory, efforts to confine plasma to realize thermonuclear fusion are at full swing. Other lab-

oratory plasmas like solid state plasma, high density plasma achieved by laser-matter interaction,

have far-reaching influence on the modern technology.

1.1 Quantum Plasmas

Historically Bohm and Pines [1-4] in the decade 1950-60 studied electromagnetic interactions of

a dense electron gas (e.g., metals) treating first time the free electrons as quantum plasma. Before

this idea, researchers used to deal with a single particle (free electron) model in which the effect

of periodic positive ionic lattice on the electron motion was taken into account. Each electron was

assumed to move quite independently of all the other electrons and Coulomb interactions were

completely ignored. However, according to Bohm and Pines, a metal is described as a plasma

of very high density (some billion times larger as compared to a gas discharge) by replacing the

positive ions as a uniform distribution of positive charges. So, the electrons can no longer be

described by the laws of classical mechanics but must be treated by a theory based on quantum

mechanics. Such dense and free electron gas interacting via long range Coulomb force, with

neutralizing background of ions can be considered as a quantum plasma.

Electrodynamics of a fully ionized charged gas is frequently studied by the classical laws of

physics. The temperature of most classical plasma environments is so high (as compared to its

particle density) that usually its electrodynamic behaviour is described by classical statistical laws

of mechanics, thermodynamics and fluid mechanics. However, in some particular plasma regimes

[5-7] we have to move from the classical description of plasma to its quantum description, mostly

due to high particle number density. Owing to high particle density (as compared to classical plas-

mas), particles’ wave function overlapping exists in these environments and thus a precise study
Chapter 1 2

in such circumstances require theoretical models that take into account quantum effects also. See

an example figure (Figure 1.1) below. However, high particle density is not essentially the only

criteria to achieve a quantum plasma. Energy distribution of the particles in a semiconductor

plasma (electron-hole dynamics, with background stationary ions) is mostly based on Boltzmann

statistics. When we cool the semiconductor plasma to a very low temperature, quantum effects be-

come very conspicuous and therefore we preferably use Fermi-Dirac statistics for the the particles

instead of Boltzmann statistics. For another example, intrinsic spin associated to each plasma par-

ticle can also contribute to quantum effects, though usually this effect is ignored while describing

ordinary classical plasmas.

Figure 1.1: Plasma diagram in the logT − logn plane. IONO: ionospheric plasma; SPACE: in-
terstellar space; CORONA: solar corona; DISCHA: typical electric discharge; TOK: tokamak
experiment (magnetic confinement fusion); ICF: inertial confinement fusion; MET: metals and
metal clusters; JUP: Jupiter’s core; DWARF: white dwarf star. This figure has been taken from
Ref. [6].

Recent innovations in miniaturized electronic technology and study of high density plasma

dynamics have led us to develop new theoretical quantum plasma models. Consequently, quan-

tum fluid [8-15] and quantum kinetic models [6,16-20] have been proposed and used to check

the strength of quantum effects in plasmas. These newly developed theoretical quantum plasma

models proved to be successful in explaining the physics of astrophysical plasmas, metallic plas-

mas, semiconductor plasmas, laser produced plasmas and of Inertial Confinement Fusion (ICF),

see review articles [21-24].


Chapter 1 3

1.2 Basic Plasma Regimes

To gain the physical understanding of practical applications it is usually more suitable to use

dimensionless parameters to distinguish different plasma regimes. There exist various plasma

regimes characterized by dimensionless parameters but here we will define only degenerate/non-

degenerate regime and collisionless/collisional regime for non-relativistic plasmas. We first define

degenerate/ non-degenerate regime. It is well known that when the average de Broglie wave-

length (λB = h̄/mvtα ) of any plasma species is comparable or larger than the average inter-particle

distance1 ∼ n−1/3 , the quantum effects become prominent in plasmas. Mathematically it can be

written as [6]

n0 (< λB >3 ) ≥ 1. (1.1)

h̄2 2/3
For a plasma Eq. (1.1) can also be defined as the ratio of Fermi energy, EF = 2m (3π 2 )2/3 n0 , to

the thermal energy κB T . In the dimensionless form it can be represented as

TF 1 2/3
χ = = (3π)2/3 n0 < λB >3 . (1.2)
T 2

When χ ≥ 1, plasma lies in the degenerate (quantum) regime and degeneracy effects can no more

be neglected but for χ < 1 plasma is non-degenerate (classical). Note that for χ ≥ 1, we have

to replace the plasma statistics from Maxwellian to Fermi-Dirac. Apparently, form Eq. (1.2) one

may conclude that the quantum effects are only important for low temperature plasmas but such

conclusion is incorrect. We emphasize that it is only quantum effects associated to plasma degen-

eracy that become important in this regime. It should be realized that the important scale for the

quantum degeneracy effects is the average inter-particle distance. But there may exist other scale-

lengths where quantum effects due to wavefunction overlapping become important. For example,

the characteristic radius for the electron-ion interaction in plasmas is Zq2 /κB T (sometimes called

Landaue length). If the thermal de Broglie wavelength becomes comparable to or greater than this

scale-length, quantum interference effects become important even in non-degenerate (classical)

hot plasmas [25]. For some mathematical estimation we can write this inequality as

h̄ Zq2α
√ ≥ . (1.3)
2mα κB Tα κB Tα
1 For exact numbers use Wigner-Seitz cell n(4π/3)r3 = 1.
Chapter 1 4

For simplicity we take charge number Z = 1 and rewrite the above expression for the electrons as

2me e4
Te ≥ 2
≈ 6.3 × 105 K. (1.4)
κB h̄
So the quantum interference effects become important in the plasma environments that have tem-

perature around 6.3 × 105 K or higher. For typical fusion plasmas with densities n = 1018 cm−3

and temperature T = 108 K, the thermal de Broglie wavelength is much less than the average inter-

particle distance but it is greater than Landau-length. Similarly, numerical analysis of the plasma

parameters in the stars’ interior (n = 1027 cm−3 and T = 107 K) shows that in these environments

thermal de Broglie wavelength, Landau-length and average inter-particle distance, are all compa-

rable (∼ 10−10 cm). Another example of quantum effects in classical plasmas is the spining motion

of charged particles that can give rise to novel effects in hot Maxwellian plasmas also [24].

In order to define the collisionless/collisional regime for any plasma system, we should keep in

mind the above scaling of degenerate and non-degenerate plasmas. Collisonless plasmas are often

called ideal or weakly coulped and when collisions are unavoidable we say plasma is non-ideal

or strongly coupled. The ratio of average potential energy∼ e2 n1/3 (due to Coulomb interactions)

and average kinetic energy∼ κB T serve as a dimensionless parameter to check the ideality or

non-ideality of any plasma system. So we can define this dimensionless parameter as

e2 n1/3 n−2/3
gc = = ,
κB T 4πλD2

or equivalently it can be written as


2/3
r̄2

1 1
gc = 2
= , (1.5)
4πλD 4π nλD3

where r̄ is the average inter-particle distance and symbol gc stands for the classical coupling pa-

rameter. For a collisionless regime we must have gc << 1. Physically this means that the Debye

length should be much greater than average inter-particle distance or equivalently we can say that

the number of particles in the Debye interaction sphere should be very large. This can only be

assured if the thermal energy of particles dominates the mean interaction energy of two charged

particles. For gc ≥ 1, plasma becomes non-ideal. In terms of collision frequencies, collisionless

regime is approached when the inequality, ω p > νab , holds. Here ω p is typical plasma frequency

and νab represents collision frequency of any charged or neutral plasma species (i.e., a or b can be

replaced by e, i, or n, for electrons, ions and neutral particles, irrespective of any order).
Chapter 1 5

The above dimensionless parameter is designated to classical plasmas. When quantum degen-

eracy effects come into play, thermal velocity (energy) of particles should be replaced by Fermi

velocity (energy). However, the basic definition of dimensionless parameter in this regime will

be same as written above in Eq. (1.5). Therefore analogous to classical plasmas one can define a

dimensionless parameter for degenerate plasmas as [6]


!2/3
r̄2 1 1
gQ = 2
= , (1.6)
4πλD f 4π nλD3 f

where gQ stands for quantum coupling parameter and λD f is the Thomas-Fermi screening length.

Here again we must have gQ << 1 to be in degenerate collisionless regime. Note that the param-

eter gQ (or gc for classical case) that represents inverse of the number of particles in the Debye

interaction sphere, is of much importance to study collective effects in plasmas. In deriving the

fundamental theoretical models of plasma physics it is used as an expansion parameter and we


3
usually take the limit (nλD/D −1 << 1 to ensure collective effects. In the present work we will
f)

always deal with collisionless plasmas (with all type of collision frequencies much less than the

plasma frequency).

1.3 Kinetic Theory of Plasmas

The early credit to initiate and develop the kinetic theory of neutral gases goes to Maxwell, Boltz-

mann and Gibbs. Many physicists later worked in modeling the kinetic theory of neutral gases and

plasmas. Some well known names are Bogoliubov, Kirkwood, Klimontovich, Born and Green,

Yvon, Frieman, Prigogine and Balescu. In ordinary neutral gases short-range intermolecular in-

teractions are important but in plasmas the long-rang Coulomb interactions play the major role.

This basic difference has led the scientists to modify the Boltzmann equation for plasmas to in-

corporate long-range interactions. Usually two approaches are followed. Either we start with

Klimontovich equation and after taking some ensemble averaging recover the Boltzman equation

or we start with Liouville equation, derive the BBGKY-hierarchy and truncate it under certain

approximations, to get the Boltzmann equation. Below we will follow the first approach to derive

classical collisionless Boltzmann equation. For understanding quantum kinetic theory, we will

have a look at the celebrated Wigner distribution function and Wigner evolution equation also.
Chapter 1 6

1.3.1 Classical Kinetic Theory

In classical description of plasmas, charged particles are considered as point particles. We usually

have huge number of particles in any plasma system. But let us first focus on only one particle. We

assume that this single particle moves randomly in six-dimensional phase space (x, v) and as time

passes it forms an orbit having coordinates X1 (t), V1 (t). The density (location) of this particle in

phase space can be achieved by the following spiky-type of distribution [26]

N(x, v,t) = δ [x − X1 (t)]δ [v − V1 (t)], (1.7)

where δ [x − X1 (t)] ≡ δ [x − X1 (t)]δ [y −Y1 (t)]δ [z − Z1 (t)] and same holds for velocity coordinates.

Note that x, v are the Eulerian coordinates in phase space while X1 (t), V1 (t) are the Lagrangian

coordinates of the particle itself. Next, if we are observing two particles instead of one then the

above definition can be written as


2
N(x, v,t) = ∑ δ [x − Xn (t)]δ [v − Vn (t)], (1.8)
n=1

which can be generalized for N-particles as


N
N(x, v,t) = ∑ δ [x − Xn (t)]δ [v − Vn (t)]. (1.9)
n=1

Time evolution of the above Eq. (1.9) gives us an exact evolution equation for classical plasmas.

So, time derivative gives us


N
∂ N(x, v,t)
= − ∑ Ẋn (t).∇x δ [x − Xn (t)]δ [v − Vn (t)]
∂t n=1
N
− ∑ V̇n (t).∇v δ [x − Xn (t)]δ [v − Vn (t)], (1.10)
n=1

where dot over Xn (t) and Vn (t) in the above equation shows the time derivative. For plasmas we

can replace V̇n (t) by the Lorentz force and also Ẋn (t) = Vn (t). After this substitution we use the

property of Dirac delta function aδ (a − b) = bδ (a − b) on the right hand side of the equation to

replace Vn (t) and Xn (t) by v and x respectively. The outcome is


N
∂ N(x, v,t)
= −v.∇x ∑ δ [x − Xn (t)]δ [v − Vn (t)]
∂t n=1
  N
qα m qα
− E [x,t] + v × Bm [x,t] .∇v ∑ δ [x − Xn (t)]δ [v − Vn (t)], (1.11)
mα mα c n=1
Chapter 1 7

where Em [x,t] and Bm [x,t] are the total fields that contain self-consistent fields and also the exter-

nal fields. Here we can see that the two summations in Eq. (1.11) is exactly the definition (1.9),

therefore we finally can write


 
∂ N(x, v,t) qα m qα m
+ v.∇x N(x, v,t) + E [x,t] + v × B [x,t] .∇v N(x, v,t) = 0, (1.12)
∂t mα mα c

which is the exact Klimontovich equation. This equation contains the detail of the orbits of all the

particles which is a much more information than needed and can not be used for practical purposes.

As we are usually interested to find the number of particles likely to be found (not exactly found)

in a small volume element ∆x∆v centered at (x, v). Therefore we take the ensemble average of the

spiky-distribution N(x,v,t), which can be written as

hN(x, v,t)i ≡ f (x, v,t), (1.13)

where the symbol hi, shows the ensemble average and f (x, v,t) is a smoothly varying distribution

function. For practical purposes, like in plasmas, we do not care about the trajectory of each

particle rather we are interested in some average fields produced by the long-range electromagnetic

forces and average number of particles to observe. Therefore using the following definitions

N(x, v,t) = f (x, v,t) + δ N(x, v,t), (1.14)

Em (x,t) = E(x,t) + δ Em (x,t), (1.15)

Bm (x,t) = B(x,t) + δ Bm (x,t), (1.16)

in Eq. (1.12) and after taking the ensemble averaging, we finally get
    
∂f qα qα qα m qα m
+ v.∇x f + E+ v × B .∇v f = − δE + v × δ B .∇v δ N . (1.17)
∂t mα mα c mα mα c

In obtaining Eq. (1.17), we have used hN(x, v,t)i = f (x, v,t), hEm (x,t)i = E(x,t), hBm (x,t)i =

B(x,t) and hδ N(x, v,t)i = hδ Em (x,t)i = hδ Bm (x,t)i = 0. Quantities on left hand side (LHS) of

Eq.(1.17) vary smoothly in phase space and contribute to collective effects while all collisional

effects are hidden on right hand side (RHS). For collisionless plasmas the RHS of above equation

can be set equal to zero and consequently we get the collisionless Boltzmann equation which is

often misnomered as Vlasov equation [27]. For brevity, we have not presented all mathematical

steps so one should see Ref. [26] for more details.


Chapter 1 8

1.3.2 Quantum Kinetic Theory

In quantum kinetic description of plasmas, particles should be treated as waves and one should

solve the Schrodinger equation for N-particles which is hard task. However this hard task can be

made simple if we are not interested in dynamics of each particle rather we want to know some

average (collective) behaviour. In quantum kinetic theory, many particle density-matrix approach

can serve this purpose [28]. In simple words, density-matrix is a quantum mechanical ensemble

averaging used for many particle dynamics. Once we specify the state of a many particle system

by the density matrix, the time evolution of this matrix gives us the dynamical behaviour of the

system. Such a time evolution equation for many particles is called quantum Liouville equation

or von Neumann equation. Density matrix in quantum kinetic theory plays the same role as the

particle probability distribution (e.g., Maxwellian) plays in classical kinetic theory.

There exist a parallel approach in which quantum phenomena can be described in classical

phase space that was first put forward by Wigner [29]. The beauty of this method is, one works

in simple phase space variables instead of quantum operator algebra but gets the same results

as expected from quantum mechanics. Extensive work has been done to study Wigner-type of

distribution functions and to extend the Wigner evolution equation for practical applications see

e.g., work by Irving [30], Klimontovich [31], Tatarskii [32], and Arnold [33]. So it may be of

interest to know something about Wigner distribution function and evolution equation for this

function.

Wigner distribution gives us a joint probability at any phase space pint (r, p), which is perhaps

its most interesting feature. This feature enables us to find the expectation value (or ensemble

average) of a quantum operator in the classical phase space. Any phase space distribution can be

uniquely defined only when a proper rule of association is specified for it. For example, when

using celebrated Wigner distribution function, Weyl rule of association is used for calculating

the expectation values of physical observables [34]. Weyl transform simply converts a quantum

mechanical operator, say Â(x̂, p̂), into a function with phase space variables A(x, p). Note that joint

probability can not be found in quantum mechanics due to the restriction of Heisenberg uncertainty

principle. In its simplest form, Wigner distribution function can be written as

1
Z
fw (x, p,t) = ψ ∗ (x + λ /2,t)ψ(x − λ /2,t)eipλ /h̄ dλ , (1.18)
2π h̄
Chapter 1 9

where ψ ∗ (x + λ /2,t)ψ(x − λ /2,t) = O(x, λ ) can be considered as overlap function [35] that gives

spatial correlation. It means that we are considering overlapping of the wave functions at some

”advanced” position x + λ /2 with its conjugate at position x − λ /2. Some general properties of

Wigner distribution function are:

(1) It is real-valued and normalized to unity.

(2) x-space or p-space probabilities can be computed by integrating it over p or x variables respec-

-tively.

(3) Magnitude of Wigner function should satisfy | fw (x, p,t)| ≤ 1 ≈ 0.318 (in a.u.).

(4) It gets negative values so it is not a true probability distribution. Therefore it is often called as

the quasi-probability distribution.

(5) It is not unique and other quasi-probability distributions can also be constructed satisfying

conditions (1)-(4).

The equation of motion for Wigner function can be derived by taking the time evolution of Eq.

(1.18), with the aid of Schrodinger equation. However here we will present the approach of Refs.

[36,37]. We start with Schrodinger equation in its non-relativistic form

∂ ψ(r,t) h̄2
ih̄ = − ∇2 ψ(r,t) +V (r,t)ψ(r,t). (1.19)
∂t 2m

Writing the above equation for two different points r1 , r2 we have

∂ ψ1 (r1 ,t) h̄2 2


ih̄ + ∇ ψ1 (r1 ,t) −V (r1 ,t)ψ1 (r1 ,t) = 0, (1.20)
∂t 2m 1

∂ ψ2 (r2 ,t) h̄2 2


ih̄ + ∇ ψ2 (r2 ,t) −V (r2 ,t)ψ2 (r2 ,t) = 0. (1.21)
∂t 2m 2
Multiply Eq. (1.20) by ψ2∗ (r2 ,t) and after taking complex conjugate of Eq. (1.21), multiply it with

ψ1 (r1 ,t) and subtract the resulting equations to get

∂ h̄2
ψ1 (r1 ,t)ψ2∗ (r2 ,t)+ ∇21 − ∇22 ψ1 (r1 ,t)ψ2∗ (r2 ,t)−[V (r1 ,t) −V (r2 ,t)] ψ1 (r1 ,t)ψ2∗ (r2 ,t) = 0.

ih̄
∂t 2m
(1.22)

Here we introduce new variables λ = r1 − r2 and r1 − r = 21 (r1 + r2 ) which give us the following

transformations

λ λ ∂2
ψ1 (r1 ,t) → ψ(r + ,t); ψ2 (r2 ,t) → ψ(r − ,t); ∇21 − ∇22 → 2 . (1.23)
2 2 ∂ r∂ λ
Chapter 1 10

Substituting these transformations in Eq. (1.22) we obtain

h̄2 ∂2
 
∂ λ ∗ λ λ λ
0 = ih̄ ψ(r + ,t)ψ (r − ,t) + 2 ψ(r + ,t)ψ ∗ (r − ,t)
∂t 2 2 2m ∂ r∂ λ 2 2
 
λ λ λ λ
− V (r + ,t) −V (r − ,t) ψ(r + ,t)ψ ∗ (r − ,t). (1.24)
2 2 2 2
1
Multiplying Eq. (1.24) by (2π h̄)3
exp(ip.λ /h̄) and integrating over λ we finally get
Z  
∂ fw p · ∇ 1 λ λ λ λ
+ fw − V (r + ,t) −V (r − ,t) ψ(r+ ,t)ψ ∗ (r− ,t)eip.λ /h̄ dλ = 0, (1.25)
∂t m ih̄ 2 2 2 2
1
ψ(r + λ2 ,t)ψ ∗ (r − λ2 ,t)eip.λ /h̄ dλ . This equation is
R
where we have used the definition fw = (2π h̄)3

formally known as Wigner evolution equation. Sometimes this equation is equivalently written in

the Wigner-Moyal [38] form as

h̄ ←
− − →
 
∂ fw p · ∇ 2
+ fw − V (r) sin ∇r · ∇ p fw = 0. (1.26)
∂t m h̄ 2

The arrows on ∇r and ∇ p in the last term show that ∇r acts only on V (r) while ∇ p only on fw .

We have briefly presented some basics of Wigner’s quantum kinetic theory, for details see review

articles [34,39].

1.4 Recent Quantum Kinetic Models for Plasmas

The seminal paper of Manfredi titled, ”How to model quantum plasmas” [6] has offered plasma

physicists to develop new, simple and clearly-defined quantum plasma models that could be treated

mathematically and numerically. In this pursuit, Tsintsadze [18] proposed a new quantum kinetic

model for Fermi particles inspired from the old work of Kuzelev [17]. This model captures particle

dispersion effects in a very simple and elegant way. To study spin-paramagnetic effects in plasmas,

a novel kinetic model (non-relativistic) has been developed on rigorous grounds by Zamanian [19]

that was earlier put forward by Brodin [40]. Asenjo [41] further extended the spin kinetic theory

to study semi-relativistic effects in plasmas. These theoretical developments in Spin Quantum

Plasmas (SQPs) are an extension of the the old work of Cowley [42] and Kulsurd [43]. However,

the present reformulation of spin kinetic theories captures many new effects. Especially, these

newly proposed kinetic models for SQPs [19,41] deal with some new type of resonances due

to electron spin motion that never appeared in the literature before. We stress here that other

parallel theoretical approaches are also well established and can be used to study quantum kinetic
Chapter 1 11

effects. In this context see Refs. [4, 44] for Random Phase Approximation (RPA) approach, Refs.

[5, 45] for Quantum Electrodynamics (QED) approach and also see the work by Jungel [46] for

semiconductor devices. However, we will prefer here to introduce only the above mentioned three

models [18,40,41] that are the essential groundwork of this dissertation and are briefly described

in the subsections below.

1.4.1 Kinetic Equation with Quantum Bohm-Potential Effect

The starting point of the mathematical formulation of quantum kinetic equation proposed in Ref.

[18], is the non-relativistic evolution equation for spin- 21 particles that contains Pauli’s term also.

However if we study plasma waves without external magnetic field (as we do in Chapter 3) spin

contributions vanish and the Pauli’s term can be neglected. Therefore, in expressing the mathemat-

ical exposition of the quantum kinetic equations [18], we will use a Hamiltonian that contains only

kinetic and potential energies of a charged spin- 12 particle, moving in an electromagnetic field.

We start with single particle Schrodinger wave equation

h̄2 2 e2 A2
 
∂ Ψα ieh̄
ih̄ =− ∇ Ψα − (A · ∇ + ∇A) + + q α φ Ψα , (1.27)
∂t 2mα 2mα c 2mα c2

where ψα is single particle wavefunction, α stands for particle species (electron/photon), A(r,t)

and φ (r,t) represent vector and scalar potentials respectively. Using the following Madelung-

representaion of the wavefunction

Ψα (r,t) = Rα (r,t) e(iSα (r,t)/h̄) , (1.28)

in equation (1.27) and separating real and imaginary parts we obtain the following two equations

∂ R2α
 
2 pα
+ ∇ · Rα = 0, (1.29)
∂t mα

h̄2
   2 
∂ pα uα × B ∇ Rα
+ (u · ∇) uα = qα E + + ∇ , (1.30)
∂t c 2mα Rα
where the amplitude Rα (r,t), and the action function Sα (r,t) are real-valued functions. An im-

portant restriction on Rα (r,t) in Madelung-representation is that it must hold Rα (r,t) ≥ 0 at all

point [35]. R2α (r,t) = |ψ(r,t)|2 ≡ ns (r,t) represents the probability density associated to single

particle and pα = mα uα = ∇Sα (r,t) − qcα A(r,t) is the particle’s momentum. In quantum me-

chanics, Eq. (1.29) shows the conservation of probability flux, and Eq. (1.30) is called quantum
Chapter 1 12

Hamilton-Jacobi equation which represents force acting on a single particle. Now we connect the

probability density of single particle ns (r,t) with the plasma particle density per unit volume as
Z
2
s
n(r,t) = Nn (r,t) = N|ψ(r,t)| = d 3 p f (r, p,t), (1.31)

where N is the total number of particles in a unit volume and f (r, p,t) represents one particle

distribution function. Above expression shows that, in a unit volume, the single particle probability

density can be converted into many particle number density when multiplied by the total number

of particles present in that unit volume. Next, we take the well known Landau’s equation for a

Fermi gas2 that is [47]

∂ fα (r, p,t) dp ∂ fα (r, p,t)


+ (v · ∇) fα (r, p,t) + = C( fα ), (1.32)
∂t dt ∂p

where C( fα ) contributes collisional effects. If we neglect C( fα ) (i.e., collisions) and make use of

Eqs. (1.30) and (1.31) into the above Eq. (1.32), we obtain the following result
 2 √ 
h̄2
  
∂ fα uα × B ∇ nα ∂ fα
+ (u · ∇) fα + qα E + + ∇ √ · = 0, (1.33)
∂t c 2mα nα ∂p

which is the required quantum kinetic equation for a Fermionic plasma with particle dispersion
dp
effects. Though the substitution of dt from Eq. (1.30) into Eq. (1.32) is crucial, but this mathe-

matical trick is plausible and elegantly captures particle dispersion effects. Moreover this equation

is much simpler than Wigner’s equation that also accounts for the same type of effects. However,

Wigner equation is more general but if one wants to concentrate just on particle dispersion effects

the above Eq. (1.33) is also a good choice.

1.4.2 Non-relativistic Spin Quantum Kinetic Model

The quantum kinetic equation that we are going to present in this section is for classical hot

plasmas but has the potential to study spin-paramagnetic effects. In this sense, this approach

should be called as semiclassical approach because it captures both classical effects and quantum

spin-effects. The quantum kinetic equation proposed in Ref. [40] has been derived heuristically

and provides an easy understanding of the model.

We are interested here to incorporate spin effects in ordinary collisionless Boltzmann equation.
2 Landau originally derived this equation to study Fermi liquid theory.
Chapter 1 13

We know that the classical collisionless Boltzmann equation in (x, v) phase space has the form

[26]
 
Df ∂ dx dv
≡ + |orbit · ∇x + |orbit · ∇v f = 0, (1.34)
Dt ∂t dt dt
dx dv
where we can guess that dt |orbit is just the velocity of single particle and dt |orbit is its acceleration

which, for plasmas, is usually replaced by the acceleration produced by the Lorentz force. Eq.

(1.34) simply shows that the density of particles remains constant along the phase space orbit.

We now extend the usual phase space (x, v) concept to incorporate spin effects. We can construct

intuitively an extended phase space that gives the particle conservation in the form
 
Df ∂ dx dv ds
≡ + |orbit · ∇x + |orbit · ∇v + |spin · ∇s f = 0, (1.35)
Dt ∂t dt dt dt
where we have introduced extra degrees of freedom for spin variable. To find the time evolution of

such a system we should preferably use quantum mechanics, but here we are doing plasma physics
dv
not quantum mechanics! Therefore we just take some aid of quantum mechanics to find dt and
ds
dt , and then we will come back to the above Eq. (1.35).

The Hamiltonian of a non-relativistic spin- 12 particle (say an electron), that reflects interactions

with electric and magnetic fields, can be written as


h̄2
H=− [∇ − (iqA)/h̄]2 + µe σ · B + qφ , (1.36)
2m2
where q = −|e| is the electronic charge, µe = −(g/4)eh̄/m is the electron magnetic moment, and

the vector quantity σ contains Pauli spin matrices. g in the definition of µe is the electron g-factor

and has the value g ≈ 2.00232. In Heisenberg picture, we write time evolution of any operator as
dF ∂F 1
= + [F, H]. (1.37)
dt ∂t ih̄
dv
When we use Eq. (1.36) in this equation, after some basic quantum algebra, the derivatives dt and
ds
dt can be calculated. This results into
 
dv 1 2µe
= q(E + v × B) + ∇(s · B) , (1.38)
dt m h̄
ds 2µe
= (s × B), (1.39)
dt h̄
where, s = h̄/2. We now substitute Eqs. (1.38) and (1.39) back in Eq. (1.35) to get the following

semiclassical collisionless Boltzmann equation,


   
∂f 1 2µe ∂ f 2µe ∂f
+ (v · ∇) f + q E+v×B + ∇(s · B) · + (s × B) · = 0. (1.40)
∂t m h̄ ∂v h̄ ∂s
Chapter 1 14

This equation contains two new terms as compared to the classical version of the theory. Last

term appears due to spin-precessional effects while second last term represents magnetic dipole

force. Due to additional spin effects in Eq. (1.40), an obvious modification is necessary in the

current density because when we consider electrons as small magnetic moments, they collectively

contribute to a magnetization current. Therefore the total current density for this case can be

written as [40]
2µe
Z Z
J = J f ree + ∇ × M = q v f dΩ + ∇× s f dΩ, (1.41)

where we have replaced usual phase space volume element with dΩ = d 3 vd 2 s. Here we have

suppressed the third degree of freedom for spin variable by assuming that we are working with a

sphere that has unit length. We also note that f = f (r, v, s,t), which shows that we should now

modify the equilibrium distribution function also to study plasma waves. The equilibrium state of a

plasma is most commonly characterized by Maxwellian distribution function. The simple remedy

to modify this distribution for spin effects is just to add an extra energy term µe · B (associated to

magnetic dipole effects), together with kinetic energy. So the modified equilibrium Maxwellian

distribution [40] for plasmas can be written as


" #
n0 m 3/2 mv2 /2 + 2µe s · B0 /h̄
f0 = ( ) exp − . (1.42)
4π 2πκB T κB T

The above presented heuristic model developed in Ref. [40] was later extended on more rigorous

grounds by Zamanian [19]. So following Ref. [19] we can rewrite above equations (1.40)-(1.42)

with some modifications as

∂f hq µe i 2µe
0= + v · ∇x f + (E + v × B) + ∇x (ŝ · B + B · ∇ŝ ) · ∇v f + (ŝ × B) · ∇ŝ f , (1.43)
∂t m m h̄
Z Z 
2 3 2 3
J = J f ree + ∇x × M = q v f d sd v + µe ∇x × 3ŝ f d sd v , (1.44)

and

f0 = f0+ + f0− = (n0 /4π)F0+ (v2 )[(1 + cos θs )] + F0− (v2 )[(1 − cos θs )]. (1.45)

Comparison of Eqs. (1.40) and (1.43) shows that the vector s = h̄/2 now has been replaced by ŝ,

that is a spin vector with unit magnitude. Note that in Eq. (1.43), we have an additional quantum

term proportional to → B · ∇ŝ . This term represents modification in the magnetic dipole force due

to the spread out of the spin probability distribution. Smearing out of the spin probability over
Chapter 1 15

the whole unit sphere also contributes to an extra constant factor of 3 in the magnetization current

density, as written in Eq. (1.44). The modified distribution function in Eq. (1.45) represents a

limiting case of the full distribution (that takes into account Fermi-Dirac statistic, spread of spin

probability and Landau quantization also) proposed by Zamanian in Ref. [19]. In Eq. (1.45) for

two distinct spin-up and spin-down populations, F0± (v2 ) represents the Maxwellian distributions,

while 1 ± cos θs contributes to the spin probability distributions of the electrons.

1.4.3 Weakly-relativistic Spin Quantum Kinetic Model

The evolution equation for weakly-relativistic spin quantum plasmas has been derived recently

by Asenjo in Ref. [41]. To introduce this model here in detail, the complete exposition of the

mathematical steps involved, is a hopeless job because lengthy and cumbersome quantum algebra

is involved. Therefore, we will very briefly overview the necessary assumptions and mathematical

steps involved to obtain the desired evolution equation. For details one can see Refs. [41] and

[19].

We start with the weakly-relativistic (having terms of the order 1/c2 ) Hamiltonain [41,48]

1  q 2 qh̄ h̄2 q h̄q h  q i


H = mc2 + qφ + p̂ − A − σ · B + 2 2 ∇ · E − 2 2 σ · E × p̂ − A
2m c 2mc 8m c 4m c c
ih̄2 q 1  q 4
− 2 2 σ · (∇ × E) + 3 2 p̂ − A . (1.46)
8m c 8m c c

Here the first four terms give usual Pauli’s Hamiltonian with the addition of rest mass energy term,

mc2 . Note that all remaining terms are of order 1/c2 , which are higher corrections to Pauli’s equa-

tion. Fifth and the last term are the Darwin term and mass-velocity correction term respectively.

Sixth and seventh spin-dependent correction terms show Thomas precession and spin-orbit cou-

pling respectively.

We now define a density matrix for spinors after choosing the proper basis that is

ραβ (x, y,t) = hx, α|ρ̂|y, β i, (1.47)

where |x, αi serve as basis states. Here z-axis is taken as the axis of quantization and along this

axis, α = 1 shows a spin-up state and α = 2 a spin-down state. Taking into account Eq. (1.46),

we are now interested to find the evolution equation for this density matrix that can be written as

∂ ρ̂
ih̄ = [Ĥ, ρ̂]. (1.48)
∂t
Chapter 1 16

In the next step we need a spin-modified Wigner transform fw (x, p, s,t), that can be used to convert

the above density matrix evolution Eq. (1.48), into the evolution equation of fw (x, p, s,t) in an

extended phase-space (x, p, s,t).

If we go back to Eq. (1.18), we see that this definition holds for pure quantum states. When we

have mixed states 3 , we can rewrite Eq. (1.18) as

1
Z
fw (x, p,t) = d 3 λ eip·λ /h̄ ρ (x + λ /2; x − λ /2,t) , (1.49)
(2π h̄)3

where ρ stands for density matrix. As pointed out in Refs [41,19], one can include spin degree

of freedom in above Eq. (1.49) and the modified Winger distribution (joint phase-space and spin-

space distribution), can be written as

1
f (x, p, s,t) = Tr(1 + s.σ )W (x, p,t), (1.50)

where symbol Tr is the trace that can be calculated writing 2 × 2-matrix for spin variables and

1
Z R 1/2
−iλ /h̄·[p− qc −1/2 dηA(x+ηλ ,t)]
W (x, p,t) = d3λ e ρ (x + λ /2; x − λ /2,t) , (1.51)
(2π h̄)3

is the Wigner-Stratonovich transform [41,19]. Integral over vector potential A in this transform

ensures the gauge invariance. Distribution function in Eq. (1.50), satisfies the following properties
Z
f (x, s,t) = d 3 p f (x, p, s,t), (1.52)
Z
f (p, s,t) = d 3 x f (x, p, s,t). (1.53)

Combining Eqs. (1.50), (1.51) and (1.48), one can finally get the following evolution equation

[41]
     
∂f p µ 1 p µ
0= + + E × (s + ∇s ) · ∇x f + q E + + E × (s + ∇s ) × B · ∇ p f
∂t me 2me c c me 2me c
h̄2 q
    
2µ p×E p×E
+ s× B− · ∇s f + µ (s + ∇s ) · ∂xi B − ∂ pi f − 2 2 ∂xi (∇ · E)∂ pi f . (1.54)
h̄ 2me c 2me c 8m c

Note that, in the above evolution equation (1.54), the second and third terms contain a velocity

which is now modified by the spin effects as

p µ
v = v(x, p, s,t) = + E × (s + ∇s ). (1.55)
me 2me c
3 Forexample., if we want to describe a plasma with large collection of particles and each particle is in its own state
that is different from others.
Chapter 1 17

In the fourth and fifth terms of Eq. (1.54), the usual magnetic field B is now modified as B →

(B − p × E/2mc). Here we can see that these terms are directly coupled with spin effects so

we can say that these terms appear due to spin-orbit coupling. The last term appears due to the

Darwin-term in the Hamiltonian, Eq. (1.46), and contributes to electron Zitterbewegung-effect. In

simple words this effect shows that electron in relativistic theory is not a point particle rather it

spreads out over a volume of order4 ∼ (h̄/mc)3 [48]. Further consequences of the above mentioned

effects on plasmas will be discussed in Chapter 5.

1.5 Summary

In this chapter we have briefly introduced the basic characteristics of quantum plasmas and men-

tioned some areas where such plasmas can be realized. We have also presented here the difference

between classical and quantum kinetic theories and showed the basic mathematical formulation of

both the theories. In the last section, we have discussed three particular quantum kinetic models

that provide necessary groundwork of this dissertation.

4 We recall that (h̄/mc) is the Compton wavelength.


Chapter 2

Motivation, Aim and Layout of the Thesis

2.1 Motivation

Originating from a diverse theoretical background, physics of quantum plasmas is now a subject

of great interest. Study of quantum effects related to particle dispersion was investigated in the

remote past [4,31] but it has revived in recent years using fluid or kinetic approach [6,15-20].

Perhaps it is Manfredi [16] who first rediscovered (and redefined) quantum plasmas but this area of

research got overwhelming support after the novel works of Haas [11,49-52], Manfredi [6,10,16]

and Anderson [53]. Fluid set of equations for quantum plasmas can be derived either by taking the

moments of Wigner equation or by Madelung transforming the Schrodinger equation followed by

proper ensemble averaging. Though the quantum corrections that appear in various plasma modes

due to particle dispersion are usually small but recently Crouseilles et al. [14] has shown that it

is not the case always. Crouseilles et al. proposed that for the high frequency plasma waves, in

nanometric objects, with length-scale comparable to the Thomas-Fermi screening length, particle

dispersion effects due to Madelung-term (also called Bohm-potential term) become comparable to

the electron degeneracy effects (h̄ω pe /EF )2 → 1 . Therefore the basic motivation is to study the


quantum effects in plasmas and to explore how such effects are examined using quantum kinetic

models.

The kinetic picture of plasma is usually studied by Wigner equation coupled with the set of

Maxwell’s equations. In Ref. [22] it has been shown that under certain approximations Wigner

equation can be converted to a collisionless Boltzmann-type of equation with quantum correc-

tion term. This equation is then applied to study the strength of quantum corrections in electron

plasma waves and ion-acoutic waves. The results obtained look similar to those that are achieved

by quantum hydrodynamic model but are more general. Recently, Tsintsadze [18] has proposed

a new quantum kinetic model for fully degenerate plasmas that successfully captures the particle

dispersion effects via Madelung-term in a different way (as compared to Wigner equation). If we

compare the results of this model [18] with those obtained from other quantum kinetic models
Chapter 2 19

[22,31], it is clear that this approach is not only mathematically simpler but retrieves old results

also. This model, that describes the dense Fermi gases, can make some predictions of the forma-

tion of neutral Bose atoms and of the existence of new bound states (electron-ion) [37]. However

the success story of this model mostly revolves around the longitudinal waves. Employing this

model, various aspects of longitudinal waves have been studied in Refs. [18,37,54,55] that in-

corporate particle dispersion effects. The novelty of this model became one of the motivations to

study particle dispersion effects.

Spin Quantum Plasma (SQP) is another growing area of research in the kinetic theory of

plasma physics5 . New kinetic models [19,40,41] have been proposed to study the collective elec-

tron paramagnetic effects in the plasma waves and instabilities. In Ref [40] it has been explored

for the first time that the spin-magnetization effects of electrons can give rise to new type of wave-

particle resonances. The theoretical model proposed in Ref. [40] was extended in Ref. [19] on

more rigorous grounds based on density-matix approach, starting from basic quantum mechan-

ics. It it pointed out that this newly developed model can be successfully applied to high density,

high magnetic field environments. It is important to note that in the long wavelength limit (waves

longer than the de Broglie wavelength), particle dispersion effects are usually negligible [19,24].

However, even in the long wavelength limit there do exist some quantum effects and that are ef-

fects associated to the intrinsic spin of particle species. Naturally these effects can be observed

only in magnetized plasmas. Magnetic dipole force and spin precession are the major effects that

have been explored extensively [19,40,41,56-60]. In Refs. [19,40,56,57] effect of electron spin

on some linear plasma waves has been investigated while in Refs. [58-60] nonlinear phenomena

in SQPs have also been addressed. Apart from several successful applications of non-relativistic

theory of SQPs, it has a drawback also. Spin effects in electrostatic waves can not be captured

using non-relativistic kinetic theory of SQPs. To overcome this drawback non-relativistic theory

of SQPs has been further extended to include semi-relativistic effects [41]. In past, except for few

attempts [42,43], electron-spin effects have not been considered in the kinetic theory of plasmas.

The above mentioned recent SQP models can have some possible applications in ICF, spintronics

as well as in astrophysics, though not yet tested empirically.


5 Fluid description of spin quantum plasmas is also well established but we will not discuss that here.
Chapter 2 20

2.2 Aim and Layout of the Thesis

In the last two decades, quantum kinetic theory for plasmas has attracted many physicists and

mathematicians who have successfully developed and applied the theory to certain physical phe-

nomena of their interest. The quest of quantum kinetic theory was to explore the underlying

physics of miniaturized electronic devices and other high density environments where the par-

ticles’ wavefunctions overlap. In finding the physics of various phenomena, different quantum

kinetic models have been proposed. These efforts divided the research into two sub-departments:

(i) development of new theoretical models and (ii) finding applications of the newly developed

models. The aim of the present work is to apply some recently developed quantum kinetic models

(more specifically, the three models that have been discussed in the preceding chapter in Section

1.4) to study some basic plasma modes. Due to large variety of quantum kinetic models that

explain different aspects of plasma physics, it is natural to explore the importance of compet-

ing kinetic models . Therefore after making sound background in the kinetic theory we select

three particular quantum kinetic models (see sections 1.4.1, 1.4.2 and 1.4.3) for studying various

plasma modes. We focused on how a certain model works and what are its applicability limits.

The idea was to investigate the same aspect of any plasma phenomenon using these three different

models and to compare the results obtained. However, as work progressed we discovered that

each model has some limitations, and no single aspect of plasma phenomenon (e.g., electrostatic

waves only) can be studied by these three different models simultaneously. For example, non-

relativistic kinetic model for SQPs [19] can not capture spin effects in electrostatic waves. For

another example, model proposed by Tsintsadze [18] is good in studying electrostatic waves but

due to Bohm-potential term in the kinetic equation, its application to the electromagnetic modes

becomes problematic and consequently no work in this direction has yet appeared. This demands

to reconsider the basic theoretical foundations of this model. Such restrictions have led us to work

in the specific problem area and this is the main reason to select different problems for each model.

Moreover, quantum effects have been studied extensively by fluid description of plasmas, there-

fore, it is also important to explore how such effects can be captured by quantum kinetic models.

In the present work, phenomenological model used in Chapter 3 is of interest since its relative

simplicity makes it easier to apply on the more complicated problems while more accurate models
Chapter 2 21

have been employed in Chapter 4 and 5 because these models have been derived from the first

principles.

Chapter 3 presents some novel aspects of longitudinal waves in fully degenerate unmagnetized

electron-positron-ion (EPI) plasmas. Theoretical model that covers the mathematical analysis of

Chapter 3 has been developed by N. L. Tsintsadze and L. N. Tsintsadze[18] (see Section 1.4.1).

In this chapter we have made linear analysis of Positron Zero-Sound waves and Positron Sound

waves. Significance of positron species in Zeor-Sound waves has been explained. Quantum cor-

rections to these plasma modes via Madelung quantum term has also been investigated. We have

further examined the excitation of Positron sound waves by a low-density monoenergetic straight

electron beam.

In Chapter 4 we have studied the effect of electron spin on the kinetic Alfven waves in clas-

sical hot magnetized electron-ion SQPs. We have used the quantum kinetic model that has been

developed by Zamanian [19]. Here we have made linear analysis of the Kinetic Alfven waves. It

will be made clear in this chapter how the quantum effects via electron spin modify the kinetic

Alfven waves in both kinetic and inertial regimes. The benefits and drawbacks of the analysis has

also been discussed at the end of chapter.

In Chapter 5 we have extended the analysis of magnetized electron-ion SQPs in the semi-

relativistic regime. Here we have employed semi-relativistic model for SQPs that has been re-

cently proposed by Asenjo[41]. We have derived the general conductivity tensor components to

study various linear electrostatic plasma modes in a magnetized spin quantum plasma. We have

particularly discussed in detail the semi-relativistic effects in the Bernstein mode.


Chapter 3

New Longitudinal Waves in Electron-Positron-Ion


Quantum Plasmas

3.1 Introduction

Advent of electron-positron (e-p) plasmas, which gained momentum in 1990s, was the major

factor which motivated researchers to work in electron-positron-ion (e-p-i) plasma. In laboratory,

Surko and Murphy [61], successfully accumulated large number of positrons and used them as

plasma particles. They proposed that the positrons in the form of positronium atom can be injected

in the hot plasma and thus can be used to probe anomalous transport phenomena in the tokamak.

Shortly after the proposal of Surko and Murphy, work in e-p plasmas gained momentum. In Refs.

[62,63], linear and non-linear collective plasma modes have been studied in the non-relativistic e-p

plasmas. It turns out that some basic plasma modes (like whistler mode) are totally absent in e-p

plasmas while such modes do exist in the electron-ion (e-i) plasmas. Such peculiarities motivated

plasma community to look into the properties of e-p-i plasmas. In this pursuit, large-amplitude

non-linear structures were studied in Refs. [64-67], which proved to be an initial step towards

understanding e-p-i plasmas. Recent interest in laser-plasma interaction [68-70] experiments has

given new impetus to this area of research. Apart from such laboratory e-p-i plasmas, it is believed

that the existence of electron-ion or e-p-i plasmas is common in planetary interiors, in compact

astrophysical objects e.g., the interior of white dwarf stars, magnetospheres of neutron stars and

magnetars. Moreover, electron-hole-ion plasma in semiconductors can also be categorized as an

e-p-i plasma, though the mass of hole may be smaller or larger than positron.

The properties of linear electron oscillations in a dense Fermi plasma have been studied in the

past [2,71-74]. Recently, a new type of quantum kinetic equations of the Fermi particles of various

species were derived, and a general set of fluid equations describing the quantum plasma was ob-

tained [18, 75]. This kinetic equation for the Fermi quantum plasma was used in Refs. [18, 75] to

study the propagation of small longitudinal perturbations in an electron-ion collisionless plasmas,

deriving a quantum dispersion equation. The dispersion properties of electrostatic oscillations in


Chapter 3 23

quantum plasmas have been discussed later in Refs. [76, 77]. The quantum linear and nonlinear

ion acoustic waves were investigated in Ref. [52]. The effects of the quantization of the orbital

motion of electrons and the spin of electrons on the propagation of longitudinal waves in the quan-

tum plasma have been reported recently [55, 78], as well as the instabilities of these waves due

to an electron beam [79]. The dust acoustic solitary waves in a quantum plasma have been also

considered [80], assuming that the dust grains are also quantum particles. A great deal of interest

was recently drawn to the study of linear and nonlinear waves in the hydrodynamic approximation,

in particular, in ion acoustic solitary waves [81-85], in the quantum electron-positron-ion plasma,

assuming electrons and positrons to be inertialess. It should be noted that the addition inertialess

positrons to the quantum electron-ion plasma merely results in a slight frequency shift in the dis-

persion equation of the ion acoustic waves, but doesn’t produce new waves. Most importantly the

damping rate of the quantum ion acoustic waves has not been considered in the above publications

and the criteria for the existence of such waves has not been established, which will be the focus

of the present work.

The work presented in this chapter contains two different quantum plasmas: one is the Fermi

gas composed of electrons, positrons and ions (in this case, the masses of the electrons and

positrons are equal, me = m p ), and the other is of electrons, holes and ions. Note that the electrons

and holes have different masses due to interaction between particles. For instance, in a solid-

state plasma or a quantum liquid the effective mass of charge carriers (electrons and holes) differs

from that of free electrons. Semiconductors, containing light negative (electrons) and heavy pos-

itive (holes) charge carriers, can be degenerate (ne ≥ 1016 − 1018 cm−3 ) with the effective mass

of electrons m∗e ≈ (0.01 − 0.1)me , and the degeneracy occurs at temperatures T < 102 K. In the

degenerate solid-state plasma or the Fermi liquid, although the particles are crowded together only

a few angström apart, the mean free path is longer than a few centimeters at low temperature. Two

factors are responsible for such long mean free paths. One is the Pauli exclusion principle and

the other is the screening of the Coulomb interaction between the particles [86-88]. Hence, in our

investigation we can suppose the plasma to be collisionless at low temperatures, and we take into

account the linear Landau damping. Unless otherwise stated, the ions are assumed to be immobile.
Chapter 3 24

3.2 Mathematical Model

We consider the propagation of small longitudinal perturbations (H = 0 , E = −∇φ ) in an electron-

positron(hole)-ion plasma. To study this problem we use mathematical model presented in Section

(1.4.1) and rewrite Eq. (1.33) as follows [75]


∂ fα ∂ fα h̄2 1 √ ∂ fα
+ (v · ∇) fα − eα ∇φ + ∇ √ 4 nα = 0, (3.1)
∂t ∂p 2mα nα ∂p
where the last term is the Madelung term, the suffix α stands for the particle species, h̄ is the Planck

constant divided by 2π and nα is the density for the different species of particles and defined as
dp
Z
nα = 2 fα . (3.2)
(2π h̄)3
The factor 2 is on account of the particle spin. We express a distribution function as fα = fα0 +

δ fα , where fα0 is the unperturbed, stationary isotropic homogeneous distribution function and

δ fα is the small perturbed part.

3.3 Linear Analysis

After linearizing equation (3.1) and the Poisson equation with respect to perturbation we obtain
h̄2
 
∂ δ nα ∂ foα
δ fα + (v · ∇)δ fα + −eα ∇δ φ + ∇4 · = 0, (3.3)
∂t 4mα noα ∂p
and
dp
Z
4δ φ = −4π ∑ eα 2 δ fα , (3.4)
α (2π h̄)3
where noα in the equation (3.3) is the equilibrium density of the plasma particles.

In our consideration we suppose that the electrons and positrons are the completely degenerate.

Namely the equilibrium distribution functions for the electrons and positrons are assumed to be
p2F
the step functions foα = Θ(µα − εα ), which is one for µα = εFα = 2mα ≥ εα and zero for εF α < ε,

and its derivative is


∂ foα
= −vδ (εα − εF α ) (3.5)
∂p
Then supposing that δ fα and δ φ vary like exp [i(k · r − ωt)] and taking into account definition

(3.5), from Eq. (3.3) and (3.4) follows the general quantum dispersion relation 6
2
3ω pα
 
1 ω ω + kvFα
1 + Σα 2 2 1− ln = 0, (3.6)
k vFα Γα 2kvFα ω − kvFα
6 See Appendix A for details.
Chapter 3 25

where
3h̄2 k2
 
ω ω + kvFα
Γα = 1 + 2 2 1− ln . (3.7)
4mα vFα 2kvFα ω − kvFα
In the following we examine the dispersion Eq.(3.6), which describes oscillation properties both

of electrons and positrons, for some interesting cases.

3.3.1 Positron Zero Sound Waves

A weak electrostatic interactions of electrons with each other in an electron-ion plasma were dis-

cussed in the remote past [71]-[74], and has been demonstrated that the undamped zero sound

can exist in an electron-ion almost Fermi gas. Landau [89], studying a weak interactions between

He3 atoms at sufficiently low temperatures, has shown that the undamped zero sound (ω ' kvFe )

can propagate in an almost ideal Fermi gas, i.e., in the degenerate electron-ion plasma the elec-

trons oscillations happen not to damp until ω/k → vFe , and the necessary condition is that the

length of the waves must be larger than the Thomas-Fermi screening length. However, in the

electron-positron-ion degenerate Fermi gas the situation drastically changes. The point is that for

the electron-ion plasma the Eqs. (3.6) and (3.7) are purely real in the frequency range ω  kvFe ,

ω  kvFi , and in the collisionless approximation the dissipation is totally absent in the degener-

ate electron-ion plasma. Hereafter, we investigate electron-positron oscillations in the degenerate

electron-positron-ion plasma, with the ions being treated as a uniform background of positive

charge.

We first consider the zero sound waves in an electron-positron-ion gas without the Madelung

term. From Eq.(3.6), choosing the frequency range kvFe > ω ∼ kvF p , taking me = m p and noting

ln(ω − kvFe ) = ln | ω − kvFe | −iπδ (ω − kvFe ), we get the following dispersion relation
" " # !#
2 2
λ p2 iπω λ p2
ω = kvF p 1 + 2 exp −2 k λ p + 2 − , (3.8)
ρ kvFe λe2
√ √
where λ p = vF p / 3ω pp and λe = vFe / 3ω pe are the positron and electron Thomas-Fermi screen-

ing lengths respectively, and ρ 2 = (λ p2 λe2 )/(λ p2 + λe2 ). From the above dispersion relation we can

separate real and imaginary frequencies as


" "  1/3 #!#
n 0e
ωr = kvF p 1 − 2 exp −2 1 + k2 λ p2 + , (3.9)
n0p
"  1/3 #!
n0e
ωi = −4πkvF p exp −4 1 + k2 λ p2 + . (3.10)
n0p
Chapter 3 26

We notice that the exponential damping in the imaginary part of frequency is twice stronger than in

the real one. It should be noted that in an electron-ion plasma the zero sound waves do not have the

imaginary part of frequency. However, in the case considered here, it does exist owing to positrons.

The physical reason for such damping is obvious from Eq.(3.10). Under the same assumptions as

above, the frequencies of the zero sound waves with the Madelung term are modified as
    1/3   1/3 
k 2 λ 2 + n0e iπω n0e
p n0p  kvFe n0p
ω = k vF p 1 + 2 exp −2 1 + − (3.11)
   
1+a 1+a


1/3
3 ωq2

n0e h̄k2
where a = k2 v2F p n0p , which plays an important role and can take any value and ωq = 2me

is the quantum oscillation frequency. Real and imaginary parts from the above equation can be

separated as
    1/3  
n0e
k2 λ p2 + n0p π
 
ωr = k vF p 1 + 2 exp −2 1 +  cos (3.12)
    
1+a 1+a

   1/3 
n0e
k2 λ p2 + n0p
ωi = −2k vF p exp −2 1 +
  
1+a


    1/3  
n0e

π

2π k2 λ p2 + n0p  2

π

sin + exp −2 1 +  cos , (3.13)
   
1+a 1+a 1+a 1+a

"  1/3 #!
n
k2 λ p2 + n 0e
0p
In deriving above Eq. (3.13) we have assumed exp −2 1 + 1+a  1. As an exam-

ple if we take a=1, that can easily be satisfied for semiconductor plasmas, the result is a purely

quantum effect
   1/3 
n0e
k2 λ p2 + n0p
ωr ' kvF p and ωi = −2kvFe exp −2 1 +  . (3.14)
  
2

It should be noted that in the usual zero sound waves one has k2 λ p2 >> 1, but here the zero sound

waves exist even in the limit k2 λ p2 ≈ 1 due to the factor (n0e /n0p )1/3 in the exponential part of

Eqs.(3.8)-(3.13). We also note that the damping of the positron zero sound waves, as follows from

Eqs.(3.10) and (3.13), is due to the electrons alone.


Chapter 3 27

Figure 3.1: Plot of real frequency of zero-sound waves for arbitrary values of a = 0 (black solid
line corresponds to plot of Eq. (3.9)), a = 3 (blue dashed line) and a = 6 (red dotted line). Both
dashed and dotted lines correspond to Eq. (3.12). Here we have fixed the ratio (ne0 /n p0 )1/3 = 5.

Figure 3.2: Plot of imaginary frequency of zero-sound waves for arbitrary values of a = 0 (black
solid line corresponds to plot of Eq. (3.10)), a = 0.001 (blue dashed line) and a = 0.005 (red
dotted line). Both dashed and dotted lines correspond to Eq. (3.13). Here we have again taken the
ratio (ne0 /n p0 )1/3 = 5 as fixed.
Chapter 3 28

It should be emphasized that Eqs.(3.8)-(3.13) are new spectrum and damping rates for positron

oscillations. In the case, when a ' 1, i.e., k2 λ p2 ∼ (n0p /n0e )1/3 , the real part of the frequency

(3.12) becomes ω ' kvF p , which means that in such plasma the phonon waves, which are purely

quantum, can propagate with slow damping. Figures (3.1) and (3.2), express the effects of quan-

tum parameter a on the real as well as imaginary frequency of zero-sound waves. In Fig. (3.1)

we noticed that the parameter a significantly affects the real frequency of zero-sound waves at

wavenumbers comparable to the inverse of λF p . However, for all a ≤ 1, it does not have any

significant affect on the real frequency. In Fig. (3.2) we observed that the parameter a actively

participates in the damping process of zero sound waves even at very small values. As this param-

eter approaches more and more closer to unity, magnitude of the damping rate increases, as shown

in Fig. (3.2). It is also important to mention that the damping of zero-sound waves is significant

only for wavenumbers comparable to the inverse of λF p .

3.3.2 Positron Sound Waves

Recently in Ref.[75], Tsintsadze et al. obtained the nonlinear kinetic equation of quantum electron

and ion plasma in the one-fluid approximation and derived the dispersion relation for the quantum

ion sound waves with damping rate in the frequency range kvFe  ω  kvFi . In this section,

we will show that a similar dispersion equation exists in the electron-positron(hole)-ion quantum

plasmas in the frequency range kvFe  ω  kvF p . The effect of ions is entirely ignored except

for the compensation of the electron-positron charges in the equilibrium state. In the following,

we study special cases in which the effective mass of electron is less than the mass of positron. In

semiconductors, for example, we often have situations when the mass of the hole becomes much

greater than the effective mass of the electron.

In the frequency range ω  kvFα ( α = e, p), i.e., when the phase velocity is larger than

the Fermi velocities of both species of charge carriers, expressions (3.6) and (3.7) are purely real,

and in the collisionless limit the dissipation is absent in the degenerate plasma. One can imme-

diately see from Eqs.(3.6) and (3.7) that in this case the dispersion relation coincides with the

Klimontovich and Silin’s [72] spectrum, but with small corrections due to the positrons oscilla-

tions. Namely, it reads

3
ω 2 = ω pe
2 2
+ ω pp + k2 (v2Fe + v2F p ) + ωqe
2 2
+ ωqp . (3.15)
5
Chapter 3 29

We note here that the contribution of the positrons in this equation is insignificant because n0e >

n0p and m p > me . We also note that this spectrum is valid only for the long wavelength limit,

k2 λα2  1. We now first consider the case without the Madelung term. For the frequency range

kvFe > ω > kvF p , we can write Eq.(3.6) as


2 
3 ω pe
 2
iπ ω ω pp
1+ 2 2 1+ − 2 = 0. (3.16)
k vFe 2kvFe ω

From Eq.(3.16), separating the real ωr and the imaginary ωi parts of the frequency, it is straight

forward to obtain the following relations


!
2 k2 λ 2
ω pp e 3 k2 v2F p
ωr2 = 1+ , (3.17)
1 + k2 λe2 5 ωr2
!
π ωr4 3 k2 v2F p 1
ωi = − 2
1+ , (3.18)
4 ω pp 5 ωr2 k2 λe2 kvFe

where λe = √1 vFe . Here k2 v2F p  ωr2 . From this inequality follows a very important relation.
3 ω pe

Namely, for the existence of waves, described by Eqs. (3.17) and (3.18), the inequality m p /me >

(n0e /n0p )1/3 must be satisfied. Expressions (3.17) and (3.18) resemble those for the ion sound

waves, but now ωr  ω pi . This spectrum (3.17) with the damping rate (3.18) are novel, and we

therefor call them the positron sound waves. It should be noted that in the Fermi electron-positron-

ion gas, when m p = me , such waves are absent. For sufficiently long waves in a plasma, as in any

quasineutral medium, the longitudinal sound waves should propagate. In these new waves (3.17)

the electrons and positrons move in unison to a first approximation, so there are no uncompensated
q
me
charges, i.e., k2 λe2  1. Introducing the positron sound velocity vs = 3m p
vFe , we have
 1/2
n0p
ωr = kvs , (3.19)
n0e
 1/2
π me n0p
ωi = − kvs . (3.20)
2 mp n0e
We should emphasize that the real part of the frequency is larger than the imaginary one, since
 1/2
me n0p
m p n0e << 1. In the opposite case k2 λe2  1, the equations (3.17) and (3.18) reduce to
 
9 2 2
ωr2 = 2
ω pp 1 + k λp , (3.21)
5
2
π ω pp 1
ωi = − . (3.22)
2 kvFe k λe2
2
Chapter 3 30

Figure 3.3: Plot of Eq. (3.20) for arbitrary values of (me /m p ) = 0.1 and (n p0 /ne0 ) = 0.01. These
numbers have been chosen to ensure the condition (m p /me ) > (n0e /n0p )1/3 .

Figure 3.4: Plot of Eq. (3.22) for arbitrary values of (me /m p ) = 0.1 and (n p0 /ne0 ) = 0.01. These
numbers have been chosen to ensure the condition (m p /me ) > (n0e /n0p )1/3 .

The frequency (3.21) evidently is the plasma frequency for the positron oscillations and Eq.

(3.22) is the corresponding damping for the short wavelength waves. Figs. (3.3) and (3.4), show

that the damping rate for long wavelength waves is much weaker than short wavelength waves.

Note that the mass ratio of electron to positron (hole) depends on the type of semiconductor ma-

terial e.g., in Silicon and GaAs, me /m p < 1 holds, while in Germanium the opposite is true, i.e.,

me /m p > 1. Next, in the case when the Madelung term is incorporated, the equation (3.6) takes
Chapter 3 31

the form  
2 2
2 3 k vF p
2 ω pp 1 + 5 ω2
3ω pe
 
1 iπω
1+ 2 2 1+ − =0, (3.23)
k vFe Γe 2kvFe ω 2Γp
where
2
3ωqe

iπω
Γe = 1 + 2 2 1 + (3.24)
k vFe 2kvFe
and
2
ωqp
Γp = 1 − . (3.25)
ω2
From Eq.(3.23), the real frequency can be expressed as
 
2 2
2 3 k vF p
ω pp 1 + 5 ω 2
r
ωr2 = ωqp
2
+ 2
3 ω pe
(3.26)
1 + k 2 v2
Fe

which for k2 v2Fe >> 2


ω pe and k2 v2Fe << 2 ,
reads respectively
ω pe
 
2 2 2 9 2 2
ωr = ωqp + ω pp 1 + k λ p (3.27)
5

and
k2 v2Fe
 
me n0p
ωr2 = 2
ωqp + . (3.28)
m p n0e 3
The corresponding damping rates are given by
 2
3π ω pp ω pe
ωi = − ω pp (3.29)
4 kvFe kvFe

and
 
π me n0p
ωi = − kvFe . (3.30)
12 m p n0e
Here we noticed that the quantum corrections due to quantum (Madelung) term appear in the real

frequencies for both long and short wavelength limits but such contributions does not participate

in the damping process, i.e., in Eqs. (3.29) and (3.30). Therefore, we can equivalently write the

results (3.29) and (3.30) into the form of Eqs. (3.20) and (3.22), with some extra constant factors

that appear due to expansions of log-terms.

3.3.3 Instability Excitation by Beam-Plasma Interaction

We now propose the excitation of the new type of waves (3.27) and (3.28) by a straight electron

beam, with the density nb much less than the plasma density, which is injected into a degener-

ate electron-hole plasma. We assume that the electron beam obeys the Maxwellian distribution
Chapter 3 32

function (since the density is low). Starting from Eq.(3.1), and following the usual procedure, we

obtain the dispersion relation including the electron beam contribution

1 + δ εe + δ ε p + δ εb = 0 , (3.31)

where
2
3ω pα
 
1 ω ω + kvFα
δ εα = 2 2 1− ln , α = e, p (3.32)
k vFα Γα 2kvFα ω − kvFα
and
 0 
ωb2 1

ω
δ εb = 2 2 1 − I+ . (3.33)
k vtb Γb k vtb
2 /2 Rx 2 /2
Here ω 0 = ω − k · u , u is the velocity of electron beam, I+ (x) = xe−x i∞ dτ eτ , and
2
ωqb
  0 
ω
Γb = 1 + 2
1 − I+ . (3.34)
k2 vtb k vtb
0
In the frequency range kvFe > ω > kvF p using the asymptotic value [90] of the function I+ ( kωvtb )

for ω 0 >> k vtb , from Eq.(3.31) we get


2 2 2
3 ω pe
 
iπ ω ω pp ω pb
1+ 1 + − − =0. (3.35)
k2 v2Fe 2kvFe ω 2 − ωqp
2 2
(ω − k · u)2 − ωqb

Examining the dispersion relation (3.35), we consider the hydrodynamic instability in the long
2 >> k2 v2 . For ω = ω + γ (where ω is given by Eq.(3.28) and ω =
wavelength limit, 3ω pe Fe r r

k · u − ωqb + γ, with |γ| << ω , Eq.(3.35) yields


1/2
ωr2 − ωqp
2

1 m p n0b
Im γ = √ (3.36)
2 mb n0p ωr ωqb

or

1 n0b n0p me k2 v2
Im γ = √ √ Fe (3.37)
6 n0e mb m p ωr ωqb
It should be noted that this growth rate is also purely quantum.

For the same frequency range kvFe > ω > kvF p , but with ω 0 << k vtb , Eq.(3.31) casts in the

form
2 2 2 
π ω0
r
3 ω pe
  
iπ ω ω pp ω pb
1+ 2 2 1+ − 2 2
+ 2 2 1+i =0 (3.38)
k vFe 2kvFe ω − ωqp k vtb 2 kvtb
which leads to the following result

(ωr2 − ωqp
2 )2
  r  
3π m p n0e 1 π m p n0b ωr − k · u 2 2 2
ωi = − − (ωr − ωqp ) . (3.39)
4 me n0p k3 v3Fe 2 3
2 mb n0p ωr k3 vtb
Chapter 3 33

So that for the kinetic instability the following inequality


3 
 r
π mb n0e vtb
k·u > 1+3 ωr (3.40)
2 me n0b v3Fe

should be satisfied.

3.5 Conclusions

We have investigated the propagation of small longitudinal perturbations in an electron-positron-

ion and electron-hole-ion plasmas, deriving a general quantum dispersion equation. Studying this

dispersion relation, in the electron-positron degenerate Fermi gas with or without the Madelung

term, we have revealed a new type of zero sound waves, which slowly damp for frequencies

ω ≤ kvF p . Whereas, in an electron-hole plasmas we found new longitudinal quantum waves,

which have no analogies in quantum electron-ion plasmas. We have proposed the generation of

these new waves by a straight electron beam, with the density much less than the plasma density.

These investigations may play an essential role for the description of complex phenomena that

appear in dense astrophysical objects, and in the next generation intense laser-solid density plasma

experiments, as well as may have a potential application in modern technology.


Chapter 4
Kinetic Alfvén Waves in Non-relativistic Spin Quantum
Plasmas

4.1 Introduction

The idea of Alfvén waves was first put forward by Alfvén, giving the concept of ”magnetohydro-

dynamic waves in a conducting fluid”, in his pioneering work [91] . Alfvén speculated that a low

frequency electromagnetic wave (later named Alfvén waves) takes part in the dynamic processes

that always occur inside the sun. Later extensive research on the Alfvén waves in space and fu-

sion plasmas [92, 93, 94] demonstrated that these are among the most important electromagnetic

plasma waves. For large perpendicular wave number the Alfvén waves are generalized to kinetic

Alfvén Waves (KAWs) by incorporating ions gyro radii effects or to inertial Alfvén waves (IAWs)

by including electrons inertial effects, which makes the waves dispersive in nature [95, 96]. The

fundamental dispersion relations of these waves separates in the KAW dispersion relation
  
2 2 2 2 2 3 Te
ω = kk vA 1 + k⊥ ρi +
4 Ti

for vte  vA and the IAW dispersion relation


!−1
c2 k2
2
ω = kk2 v2A 1 + 2⊥
ω pe

for vte  vA [97], where vte is the electron thermal velocity vA is the Alfvén velocity, ρi is the

ion gyro radius, ω pe is the electron plasma frequency, c is the speed of light in vacuum, Te (Ti )

is the electron (ion) temperature and kk (k⊥ ) is the wave number parallel (perpendicular) to the

external magnetic field. These waves have been experimentally observed both in space [98, 99] and

laboratory [100, 101] plasmas. Study of the KAWs has revealed their importance in the particle

energization in magnetized plasmas and can be applied to the tokamak plasma heating [102, 103,

104], auroral electron acceleration [105, 106], sunspot and solar coronal plasma heating [107,

108]. Nonlinear aspects of KAWs have also been studied in literature [109]. Recent advancement

in the field of quantum plasmas demands the study of KAWs for the high density plasma regimes

also, and this is what is the purpose of the present chapter.


Chapter 4 35

Much of the focus on quantum plasmas has been to study quantum dispersion effects using

the Bohm potential or quantum statistical effects incorporating the Fermi pressure [11, 14, 49,

52, 110, 111, 112, 113]. However a few works [12, 19, 40, 114] studied spin quantum effects

in plasmas including the contributions from the magnetic diploe force, spin precession and spin

magnetization. Within the framework of fluid model, Ref. [12] derived spin multifluid equations

by taking the macroscopic spin density as an additional variable. Their analysis suggested that spin

effects are important for strong magnetic fields when the Zeeman energy becomes comparable to

the characteristic kinetic energy (which could be thermal energy or the Fermi energy, see also

Ref. [114]), which applies to astrophysical environments such as pulsars and magnetars. In Ref.

[19], using quantum kinetic theory it was found that spin-induced damping of Alfvén waves could

occur for plasma parameters that lies in a supposedly classical regime. In Ref. [40] a new type

of mode in magnetized plasmas has been investigated which emerge due to the combined effect

of electron spin precession and the magnetic dipole force. Based on the theory developed in Ref.

[19], a general conductivity tensor was derived in Ref. [57]. The conductivity tensor was used to

discuss Weibel-like instability associated with the electron spin properties. A review of the recent

work on spin quantum plasmas has been presented in Ref. [24].

Usually it becomes of interest to examine how the quantum effects come into play when the

plasma density is of the order n ∼ 1026 cm−3 or higher. The electron dynamics may be quantum

mechanical at the same time as the ion motion is classical, due to the mass difference. However

certain low frequency (occurring on ion time scale) phenomena, can still be significantly modified

due to quantum nature of electrons. Some simple examples of such modes are ion-acoustic waves,

ion-cyclotron waves, magnetosonic waves and KAWs [12, 115, 116, 117]. Motivated by recent

studies of spin quantum plasmas, we have here made an effort to investigate spin quantum effects

in the low frequency Alfvén waves in a high density hot magnetized plasma, using nonrelativistic

linear spin-kinetic theory. We have studied collective electron spin effects on the KAWs in the

kinetic (vte  vA ) as well as in the inertial limit (vte  vA ) regime.


Chapter 4 36

4.2 Mathematical Model

The mathematical structure needed for the proposed problem is based on the recently developed

nonrelativistic spin quantum plasma model [19, 57] presented in Section (1.4.2). For clarity we

rewrite Eqs. (1.43) and (1.44) below that will serve the purpose,
 
∂f qα µe 2µe
0= + v · ∇x f + (E + v × B) + ∇x (ŝ · B + B · ∇ŝ ) · ∇v f + (ŝ × B) · ∇ŝ f , (4.1)
∂t mα me h̄

and
Z Z 
2 3 2 3
J = J f + JM = J f + ∇x × M = qα v f d sd v + µα ∇x × 3ŝ f d sd v , (4.2)

where α stands for electrons(e) or ions(i), µe = gqe h̄/4me , is the electron magnetic moment,

g = 2.00232 is the electron spin g-factor and qe is the electron charge, me is the electron mass,

h = 2π h̄ is Planck’s constant and ŝ is the unit spin vector. Note that, in writing Eq. (4.1), we

have taken spin-effects for electrons only and neglected such effects for ions due to the relative

mass difference. We use spherical coordinates in spin space such that d 2 s = sin θs dθs dϕs , and

∇ŝ = θ̂ ∂ /∂ θs + ϕ̂(1/ sin θs )∂ /∂ ϕs . The system is closed by Maxwell equations.

4.3 Linear Analysis

Next we focus on linear wave propagation in the presence of an unperturbed magnetic field B0 =

B0 ẑ. The wave properties are fully described by the conductivity tensor defined by ji = σi j E j where

we use the Einstein summation convention. Here σi j are defined to include all plasma currents,

i.e. free currents J f as well as magnetization currents ∇x × M . After linearizing Eq. (4.1) and

Eq. (4.2), one can find the following components of the electron contribution to the conductivity

tensor (as derived by Ref. [57]).


−2iq2α ωcα
2
2 2 ∂ f0ν
∞ 2 n Jn ∂ v2
Z
3 mα k⊥
σxx = ∑ ∑ d v
ν=+,− n=−∞ ω − kk vk − nωcα
" h̄
 ∂ f0ν
∞ Z
−iµ 2
e 2 2
ν f 0ν + me nω ce + kk vk ∂ v2
+ ∑ ∑ d3v kk Jn
ν=+,− n=−∞ h̄ω ω − kk vk − nωce + ωcg
−ν f0ν + mh̄e nωce + kk vk ∂∂ fv0ν2
 #
+ (4.3)
ω − kk vk − nωce − ωcg
Chapter 4 37

and
−2iq2α 2 2 ∂ f0ν
mα Jn vk ∂ v2
∞ Z
σzz = ∑ ∑ d3v
ν=+,− n=−∞ ω − kk vk − nωcα
ν f0ν + mh̄e nωce + kk vk ∂∂ fv0ν2
" 
∞ Z
−iµ 2
3 e 2 2
+ ∑ ∑ d v k J
ν=+,− n=−∞ h̄ω ⊥ n ω − kk vk − nωce + ωcg
−ν f0ν + mh̄e nωce + kk vk ∂∂ fv0ν2
 #
+ . (4.4)
ω − kk vk − nωce − ωcg

where the spin precession frequency is ωcg = (g/2)ωce , ωce denote the electron cyclotron fre-

quency, and ν = +, − shows the summation over spin-up(down) electrons. The ion contribution

is classical and can be found by replacing index e against index i in the first (classical) terms of

Eqs. (4.3) and (4.4). To proceed further, we first define the spin-dependent distribution function to

analyze Eqs. (4.3) and (4.4) that is, f0± = (1/4π)[(1 ± cos θs )]F0± (v2 ). Note that f0 = f0+ + f0− ,

and F0± (v2 ) is the usual distribution function for the up and down populations of electrons, which

could be e.g. a Maxwellian or a Fermi-Dirac distribution. The number density of particles with

F0± (v2 ) d 3 v. For a plasma in thermodynamic equi-


R
spin up (down) n0+ (n0− ) is given by n0± =

librium that satisfy T  TF , where T is the thermal temperature and TF is the Fermi temperature,

the relation between the spin-up and spin-down distributions is given by

n0+ − n0− = tanh[µe B0 /kB T ](n0+ + n0− ), (4.5)

where we define n0+ + n0− ≡ n0 . In a more general case, the relation between spin-up and spin-

down states may not correspond to thermodynamic equilibrium. For example when the external

magnetic field is changed on a relatively fast scale (faster than the spin relaxation time), we do not

expect the distribution of spin-up and spin-down states to be given by Eq. (4.5). More generally

we can account for a deviation from thermodynamic equilibrium by introducing a spin temperature

Tsp defined by

n0+ − n0− = tanh[µe B0 /kB Tsp ](n0+ + n0− ), (4.6)

where Tsp in general differ from T . The relation (4.6) will be used from now on to carry out the

summation over the up- and down spin -distributions. Furthermore we will consider the case T

 TF and take F0± (v2 ) to be Maxwellian.

To perform the d 3 v-integrations in Eqs. (4.3) and (4.4), we have use the following standard inte-

grals:
Chapter 4 38

Z ∞ v2

− 2
Jn2 v⊥ dv⊥ e vtα 2
= vtα Γn (χ⊥α ) (4.7)
0

v2
k
− 2
dvk e 1√ vtα
Z ∞
=− πZ(ζnα ) (4.8)
−∞ ω − kk vk − nωcα kk

v2
k
− 2
vk dvk e vtα √
vtα
Z ∞
=− π (1 + ζnα Z(ζnα )) (4.9)
−∞ ω − kk vk − nωcα kk
and similarly
v2
k
− 2
dvk e 1√vte
Z ∞
=− πZ(ζnsp ) (4.10)
−∞ ω − kk vk − nωce ± ωcg kk

v2
k
− 2
vk dvk e vte √
vte
Z ∞
=− π (1 + ζnsp Z(ζnsp )) (4.11)
−∞ ω − kk vk − nωce ± ωcg kk
Using the above mentioned integrals and the definition εi j = δi j + iσi j /ε0 ω, we can define the

required components of dielectric tensor as


2 2
g2 h̄kk
∞   ∞
ω pα Γn (χ⊥α )
2 ω ω pe
εxx = 1 + ∑ ∑ n χ⊥α Z(ζnα ) − ∑ Γn (χ⊥e ) 2
α ω2 n=−∞ kk vtα n=−∞ 8 mvte ω

µe B0 µe B 0
× tanh[ ]Z(ζnsp+ ) − tanh[ ]Z(ζnsp− )
kB Tsp kB Tsp

− 2 {nωce (Z(ζnsp+ ) + Z(ζnsp− ))
mvte

+kk vte (2 + ζnsp+ Z(ζnsp+ ) + ζnsp− Z(ζnsp− ))} , (4.12)

and
2
ω pα ∞  
ω
εzz = 1+2∑ ∑ Γn (χ⊥α )ζnα (1 + ζnα Z(ζnα )) kk vtα
α ω2
n=−∞
2 2 

g2 h̄ k⊥ ω pe µe B0 µe B0
− ∑ Γn (χ⊥e ) 2 tanh[ ]Z(ζnsp+ ) − tanh[ ]Z(ζnsp− )
n=−∞ 8 mv k
te k ω k T
B sp kB Tsp


− 2
{nω ce (Z(ζnsp+ ) + Z(ζnsp− )) + k v
k te (2 + ζnsp+ Z(ζnsp+ ) + ζnsp− Z(ζnsp− ))} , (4.13)
mvte

where Z(ζnα ) is the usual plasma dispersion function with argument ζnα = (ω − nωcα )/kk vtα ,
2 v2 /2ω 2 = k2 ρ 2 , and v2 = 2T /m . It is important to mention here that in obtaining
χ⊥α = k⊥ tα cα ⊥ α tα α α

above results we have performed summation over ν also which gives us Z(ζnsp+ ) and Z(ζnsp− ).
Chapter 4 39

These Z-functions can be considered as spin modified plasma dispersion functions having argu-

ments ζnsp+ = (ω − nωce + ωcg )/kk vte and ζnsp− = (ω − nωce − ωcg )/kk vte .

4.3.1 General Dispersion Relation of Kinetic Alfven Waves

The dispersion relation for the KAWs in the low frequency and long parallel wavelength limit is

given by the following determinant


s − n2
c c + εs + n n

εxx + εxx k εxz xz k ⊥
= 0· (4.14)
εc + εs + n n c s 2
k ⊥ εzz + εzz − n⊥

zx zx

where, nk = ckk /ω and n⊥ = ck⊥ /ω. Here εicj and εisj represent the classical and quantum (due to

electron spin) contributions to dielectric response function. Expanding the determinant, we obtain

the general dispersion relation for KAWs as


 
c s
− n2k εzzc = n2⊥ − εzzs εxx
 c
εxx + εxx (4.15)

c ε c >> (ε c )2 but ε sp ε sp = (ε sp )2 which


In obtaining Eq. (4.15) we have used the fact that εxx zz xz xx zz xz

shows that the relative importance of the xz-components are much more important for the spin

contributions. Now, we first simplify εxx = ε cxx +ε sxx and εzz = ε czz +ε szz . Separating n = 0 and n 6= 0

terms, expanding plasma dispersion functions for ζnα >> 1, ζnsp± >> 1, ζ0sp± >> 1, combining

terms for n = ±1, making use of the identity Γ−n (χ⊥α ) = Γn (χ⊥α ) we have
2
tanh[ µkeBBT0 ] tanh[ µkeBBT0 ]
" #
2 2
ω pα ∞
2 Γn (χ⊥α ) g2 h̄kk ω pe
εxx = 1 + ∑ 2 ∑ + Γ0 (µ⊥e ) 2 − (4.16)
α ωcα n=1 χ⊥α 8 m ω ω + ωcg ω − ωcg
2 2 
g2 h̄kk ∞
   
ω pe µe B0 1 1 h̄ωce 1 1
+
8 m n=1 ∑ Γn (χ⊥e ) ω 2 tanh[ kB T ] ω + ∆ωce − ω − ∆ωce − mv2 ω + ∆ωce − ω − ∆ωce .
te

After taking into account ω  ∆ωce = ωcg − ωce and ω  ωcg we can write above equation as

2 2 µB
2 2 tanh[ e 0 ]
ω pα 1 − Γ0 (χ⊥α ) g2 h̄kk ω pe kB T
εxx = 1+∑ 2 + Γ0 (χ⊥e ) 2
α ω cα χ ⊥α 8 m ω ω cg
2 h̄k 2 2     
g k ω pe µe B0 1 h̄ωce 1
+ (1 − Γ0 (χ⊥e )) 2 tanh[ ] − 2
(4.17)
8 m ω kB T ∆ωce mvte ∆ωce

Some further simplification can be achieved by taking the following assumptions (1 − Γ0 (χ⊥α )) /χ⊥α ≈

1 − 43 k⊥
2 ρ 2 , Γ (χ
α
2 2
0 ⊥α ) ≈ 1 − k⊥ ρα , and me  mi . Under these assumptions we obtain

c2
 
3 2 2
εxx = 1 + 2 1 − k⊥ ρi + n2k Q ≡ εxx
c
+ n2k Q, (4.18)
vA 4
Chapter 4 40

and similar simplifications for εzz lead us to


2
ω pα
 
ω
εzz = 1 + 2 ∑ Γ0 (χ⊥α )ζ0α (1 + ζ0α Z(ζ0α )) + n2⊥ Q ≡ εzzc + n2⊥ Q (4.19)
α ω2 kk vtα

where n2k Q = εxx


s , n2 Q = ε s and Q is defined as
⊥ zz

µe B0
2
g2 h̄ω pe  tanh[ kB Tsp ] g2 2 2  h̄ω pe
2   
2 2 µe B0 h̄ωce 1
Q= 1 − k ρ
⊥ e + k ρ tanh[ ]− (4.20)
4 mc2 ωcg 8 ⊥ e mc2 kB Tsp 2
mvte ∆ωce

Using Eqs. (4.18) and (4.19) back in Eq. (4.15) we can write
 
c
εxx − n2k (1 − Q) εzzc = n2⊥ (1 − Q)εxx
c
. (4.21)

Eq. (4.21) is the general dispersion relation for KAWs that contains the electron spin contributions

in the form of dimensionless quantum parameter Q.

4.3.2 Kinetic Alfven Waves in the Kinetic Regime

In the kinetic regime, we assume vti  ω/kk  vte therefore 1+ζ0e Z(ζoe ) = 1, and 1+ζ0i Z(ζoi ) '

−1/2ζ0i2 = −kk2 vti2 /2ω 2 . Thus under these conditions equation (4.21) takes the form
"  c2 k2 #"
kk2 vti2
!#
c2 1 − k2 ρ 2 1 − k2 ρ 2

3 2 2 k
1+ 2 1 − k⊥ ρi − 2 (1 − Q) 1 + 2 ⊥2 e + 2 ⊥2 i −
vA 4 ω kk λDe kk λDi 2ω 2
c2 k 2 c2
  
3 2 2
= 2⊥ (1 − Q) 1 + 2 1 − k⊥ ρi (4.22)
ω vA 4

For c2 >> v2A and kk2 λDe


2  1 we can rewrite above equation as

ω 4 − Bω 2 +C = 0, (4.23)

where
 
3 2 2
B= kk2 v2A 1 + k⊥ ρi + k⊥ ρa (1 + k⊥ ρe ) (1 − Q) + kk2 c2s (1 − k⊥
2 2 2 2 2 2 2 2
ρi )(1 + k⊥ ρe ), (4.24)
4
3 2 2
C = kk2 v2A (1 + k⊥ ρi )(1 − Q)kk2 c2s (1 − k⊥
2 2 2 2
ρi )(1 + k⊥ ρe ). (4.25)
4
The solution of equation (4.23) can be written as
r !
B 4C
ω2 = 1± 1− 2 (4.26)
2 B

Eq. (4.26) is the modified dispersion relation of KAWs in the kinetic regime with electron spin

effects, represented by the expression Q . If we set quantum contributions zero (i.e., h̄ → 0)


Chapter 4 41

Q vanishes and we safely retrieve the classical results. A plot of Eq. (4.26) is shown in Fig.

(4.1) below for a plasma density of n0 = 2 × 1026 cm−3 a magnetic field B0 = 108 G an electron

temperature Te = 2 × 107 K and cold ions (Ti = 104 K) for various spin temperatures Tsp . This

example is not strictly within the limit of the applicability as we have assumed TF  Te and our

example have TF ≤ Te . However, we still expect the plot to be qualitatively correct in this regime.

As our plotting variable on x-axis is k⊥ ρa , it might be of interest that how spin-contributing factor

Q varies with k⊥ ρa . Therefore we have shown below a plot of k⊥ ρa versus Q in Fig. (4.2). Here

ρa = cs /ωci is the acoustic gyro radius and cs is the usual ion-acoustic velocity.

In plotting Eq. (4.26), we have taken the negative root only, in order to get the kinetic Alfvén

wave rather than the ion-acoustic mode. Note however that this ordering holds for a high density

plasma with β larger than unity 7 . For lower β it is the positive root that becomes the kinetic

Alfvén wave. For parallel propagation we take k⊥ = 0 in Eq. (4.26) and obtain modified dispersion

relation of pure Alfvén waves with spin effects,


!
2
g2 h̄ ω pe µB0
ω 2 = kk2 v2A 1− tanh[ ] (4.27)
4 mc2 ωcg kB Tsp

4.3.3 Kinetic Alfven Waves in the Inertial Regime

The KAWs in inertial regime are named as inertial Alfvén waves. In inertial regime, we assume

that both ions and electrons are cold in this case, i.e., ω/kk  vt i,e therefore in Eq. (4.21) we take
2 = −k2 v2 /2ω 2 , which leads to
1 + ζ0i,e Z(ζoi,e ) ' −1/2ζ0i,e k t i,e

"  c2 k2 #"
kk2 vte
2
!
kk2 vti2
!#
c2 1 − k2 ρ 2 1 − k2 ρ 2

3 2 2 k
1+ 2 1 − k⊥ ρi − 2 (1 − Q) 1 + 2 ⊥2 e − + 2 ⊥2 i −
vA 4 ω kk λDe 2ω 2 kk λDi 2ω 2
c2 k 2 c2
  
3 2 2
= 2⊥ (1 − Q) 2 1 − k⊥ ρi (4.28)
ω vA 4

For c2 >> v2A , kk2 λDe


2  1 and ω/k  v
k ti,e we can write above equation as

ω2 (1 + 43 k⊥ 2 ρ 2 )(1 − Q)
i
= (4.29)
kk2 v2A 1 + c2 k⊥2 (1 − Q)(1 + k2 ρ 2 )
ω2 pe ⊥ e

7 See Appendix B
Chapter 4 42

Figure 4.1: Plot of the modified kinetic Alfven wave, Eq.(4.26),with spin effects for different spin
temperatures Tsp . We have taken n = 2 × 1026 cm3 = ne0 = ni0 , B0 = 108 G, Te = 2 × 107 K and
Ti = 104 K. Black line shows kinetic Alfvn wave without spin quantum effects(h̄ → 0). Red, blue
and brown lines show spin quantum effects for spin temperatures Tsp = 0.5Te , Tsp = 0.2Te , and
Tsp = 0.1Te respectively.

Figure 4.2: Plot shows the variation of quantum parameter Q as a function of k⊥ ρa .


Chapter 4 43

where we have already defined Q in Eq. (4.20). This is modified dispersion relation of KAWs in

inertial regime with spin effects. Here again if we set h̄ → 0, Q vanishes and we obtain the usual

classical results.

4.4 Conclusions

2 > v2 ) as
We have here evaluated the spin contribution for shear Alfvén waves in the kinetic (vte A
2 < v2 ) regime based on the conductivity tensor derived by Lundin and
well as in the inertial (vte A

Brodin [57]. We note that to some extent the spin effects are suppressed for low frequencies

considered here (i.e. much lower than ∆ωce ). From a mathematical point of view this happens

because several terms with denominators ω −kk vk ±∆ωce in the conductivity tensor approximately

cancels for low frequencies. Physically this corresponds to (approximately) averaging out the

effects of spin polarization on these slow scales. Nevertheless in case the plasma density is high

certain effects due to the electron spin can be seen. In Fig. (4.1) we have picked values ne

= 2 × 1026 cm−3 , B0 = 1 × 108 G, and Te ≈ 2 × 107 K , which is of relevance for e.g. inertial

confinement fusion (ICF) plasmas, where the quasi-static self-generated magnetic field can be

large [118]. We have graphically shown the effect of electron spin on the kinetic Alfvén waves

by focusing on variations in the spin temperature. The effects of spin is significant when there

is a deviation from thermodynamic equilibrium, such that the lower energy spin state is more

populated than in the equilibrium state. Such a situation may occur e.g. if B0 is decreased faster

than the spin relaxation time. Analysis of Eq. (4.29) for the above mentioned ICF parameters

shows that electron spin quantum effects in IAWs behave differently as compared KAWs Eq.

(4.26). Spin quantum effects appear both in connection to gyro radii (electron or ion) effects and

electron inertial effects. We observe a competition between spin effects that appear due to these

terms which mars the overall spin quantum effects in IAWs, as can be seen from Eq. (4.29). So,

we deduce that IAWs are less affected by the spin quantum effects in a hot magnetized plasma.
Chapter 5

Electrostatic Modes in Weakly Relativistic Spin Quantum


Plasmas

5.1 Introduction

The very origin of the weakly relativistic effects is the relativistic Dirac Hamiltonian (see Ref.

[48]). Spin-orbit coupling effects are the most frequently encountered weakly relativistic effects.

Spin-orbit coupling for single-electron atoms appear to be weak while for multielectron atoms it

becomes reasonably strong. To get the basic understanding of this effect we consider a single-

electron atom case (e.g., Hydrogen atom). If the electron orbits the nucleus with velocity v, then

in the electron rest frame the nucleus will orbit the electron with velocity −v (see Fig. below).

In electron’s rest frame it constitutes a current J = −Zev [119]. This current loop produces a

Figure 5.1: Left: An electron moves in a circular Bohr orbit, as seen by the nucleus. Right: The
same motion but seen from the point of view of the electron. The magnetic field B is at electron’s
location and is directed out of the page. This figure has been taken from Ref. [119].

magnetic field B, according to Ampere’s law as

µ0 J × r Zeµ0 v × r
B= =−
4π r3 4π r3

To express B in terms of electric field E, we use Coulomb’s law

Ze r
E=
4πε0 r3
Chapter 5 45

Combining the last two equations we can write

1
B=− v × E,
c2

which can give rise to a precession frequency

eB e
=ω =− v × E.
me me c2

When we derive the above expression from relativistic considerations using transformation of

velocities we get a correction factor of 1/2 in the above expression, sometimes called Thomas

factor. For details of Thomas precession see Appendix O of Ref. [119].

Spin-orbit coupling is mostly addressed from the point of view of atomic physics, however

the concept is well-defined from a general point of view. Basically, the spin-orbit interaction is

the first order relativistic correction to the magnetic dipole energy in the Pauli Hamiltonian. To

see the general physical significance of this effect we should not visualize a specific orbit at all.

In a fully ionized gas like plasmas, we may think of the velocity v as coming from some sort of

orbit, but apparently we only need a nonzero velocity to get an effect of this type. It is fortunate

that we don’t need to care about detailed orbits to get a proper expression, since the orbits can be

tremendously complicated, as is the case in plasmas.

In the past, only a limited work has appeared in the literature [42, 43], to study spin-orbit

interaction effects in plasmas. However, recent interest to explore such effects in plasmas [120-

125] has motivated researchers to develop new type of quantum hydrodynamic and kinetic models.

In Ref. [120], taking into account Rashba spin-orbit interaction, collective plasma excitations have

been studied in a quantum wire. In Ref. [121], new quantum hydrodynamic equations have been

derived that incorporate spin-current and spin-orbit interaction effects. It has been proved that

new wave branches exist in plasmas, if we excite a plasma from an external source and consider

the magnetic moments of plasma particles. In a recent investigation of electrostatic plasma waves

[122], it has been found that the the group velocity of plasma modes as well as the growth rate

of the plasma instability are enhanced by electron spin interactions. In Ref. [123], it has been

shown that new magnetization and polarization waves can propagate in plasmas if spin-orbital

interaction is taken into account. Study of Whistler waves in Ref [124] reveals that the spin-

orbital corrections may become important for strong magnetic fields and high particle density
Chapter 5 46

concentrations. Outcome of spin-orbit coupling effects in quantum plasmas, based on quantum

hydrodynamic model of Ref. [125], proves that the spin-quantum corrections that are produced by

the magnetic moments of plasmas particles are more important than the quantum Bohm-potentail

corrections, at least for strong-radiation case. Recently, following Ref. [41], a study of linear and

non-linear plasma waves [59] shows that semi-relativistic effects may become prominent for high

density and high magnetic field environments.

In the present chapter we will start with the quantum kinetic model [41] that we have intro-

duced in Section (1.4.3), to include weakly relativistic effects such as spin-orbit interaction and

Thomas precession [119]. The model is used to analyze linear electrostatic waves propagating

at an arbitrary angle with respect to an external magnetic field. A general dispersion relation is

derived which reduces to the well-known classical result when the limit h̄ → 0 is taken. One of the

main effects of the full model is to allow for new electrostatic wave modes that is absent unless

the spin-orbit interaction is present. The dispersion relation is solved numerically for an example

of this kind.

5.2 Mathematical Model

In Ref. [41], Asenjo et al. derived a weakly relativistic model based on the Dirac equation. By

making successive Foldy–Wouthuysen transformations [126] of the Dirac Hamiltonian, a weakly

relativistic Hamiltonian can be calculated that only contains the positive energy states (i.e. the elec-

trons). As exploited mathematically in Section (1.4.3), a density matrix can be constructed, whose

evolution is described by the weakly relativistic Hamiltonian. By making a Wigner-Stratonovich

transformation [127] of the density matrix, together with a Q-transformation (see Eq. (12) in [41])

a scalar quasi-distribution function can be defined. As a result, we get the kinetic evolution equa-

tion for a weakly relativistic spin- 21 collisionless plasma, that we have already exposed in Section

(1.4.3). After omitting one term of Eq. (1.51), we rewrite it here again for the sake of convenience

of the reader and rename it as Eq. (5.1).


     
∂f p µ 1 p µ
0= + + E × (s + ∇s ) · ∇x f + q E + + E × (s + ∇s ) × B · ∇p f
∂t me 2me c c me 2me c
    
2µ p×E p×E
+ s× B− · ∇s f + µ (s + ∇s ) · ∂xi B − ∂pi f . (5.1)
h̄ 2me c 2me c
Chapter 5 47

Here f = f (x, p, s,t) is the quasi-distribution function defined on a phase space extended by two

spin dimensions (denoted s) on the unit sphere, in addition to the traditional space and momentum

coordinates x and p. We use the notation m for the mass and µ is the magnetic moment of the

particle (= geh̄/4me for electrons, where g = 2.002319 is the spin g-factor), q is its charge. The

index x, p or s on the nabla operator indicates that it acts on the respective coordinates. The

omitted term is associated with the Darwin term in the Hamiltonian, and was dropped since in

a long scale length expansion it is smaller than the other terms 8 . This Vlasov-like equation is

coupled to Maxwells equations

∇ · E = 4πρT , (5.2)
1 ∂ E 4π
∇×B = + JT , (5.3)
c ∂t c
where the total charge and current density are given by

ρT = ρF + ∇ · P, (5.4)


JT = JF + ∇ × M + P. (5.5)
∂t
R
Here ρF = q dΩ f is the free charge density and the free current density, the polarization and

magnetization are given by


 
p 3µ
Z
Jf = q dΩ + E×s f, (5.6)
me 2me c
s×p
Z
P = −3µ dΩ f, (5.7)
2me c
Z
M = 3µ dΩ s f , (5.8)

respectively, where dΩ = d3 p sin θs dθs dφs using spherical coordinates θs and φs in spin space. For

a more complete discussion of the model presented here, containing e.g. conservation laws and a

physical interpretations of the momentum variable, see Ref. [41].


8 Eq. (5.1) is valid for long wavelengths, which means scale lengths much longer than the thermal de Broglie length
and also much longer than the Compton wavelength. The last condition is needed to drop the contribution to Eq.
(5.1) that would have originated from the Darwin term in the Hamiltonian. At the same time we will consider the
electrostatic approximation, which applies for sufficiently short wavelengths. As seen from the derivation in Appendix
D, the conditions of sufficiently short wavelengths and sufficiently long wavelengths can be fulfilled simultaneously.
Chapter 5 48

5.3 Linear Analysis

Here we assume that the magnetic field can be written as B = B0 + B1 , where subscripts 0 and 1

are used for equilibrium and perturbed quantities. (The electric field is simply written as E1 , since

the equilibrium part is taken to be zero, i.e. E0 = 0.) We also write the total distribution function

as f = f0 + f1 . Firstly, we discuss the thermodynamic expressions for the background spin and

momentum distribution given by f0 . In Eq. (60) by Ref. [19] a general expression for f0 is

given for a magnetized plasma, where the z-axis is taken along the unperturbed magnetic field, i.e.

B0 = B0 ẑ. The general formula is very complicated due to the effect of Landau quantization, which

is significant whenever the Zeeman energy µB0 is comparable to the kinetic energy Ek (which can

be either kB T or kB TF , depending on whether the temperature T or the Fermi temperature TF is

larger). In order to simplify the expressions we will consider the limit when µB0 < Ek , such that

the effect of Landau quantization is reasonably small. For this case the distribution function in

thermodynamic equilibrium can be written as

f0 = f0 (p2 , θs ) = fˆ0 (p2 )[c+ (1 + cos θs ) + c− (1 − cos θs )] (5.9)

independent of the ordering between T and TF . Here fˆ0 (p2 ) is normalized such that fˆ0 (p2 )d 3 p =
R

1, and the constants c+ and c− should produce the number density of the spin-up and spin-down

fˆ0 (p2 )c± (1 ± cos θs )dΩ = n0± , where the total number density is n0 =
R
states, respectively, i.e.

n0+ +n0− . Firstly picking the case TF  T (corresponding to a large and negative chemical poten-

tial in Eq. (60) of Ref. [19]) we obtain c± = n0± /4π = n0 exp(±µB0 /kB T )/[cosh(µB0 /kB T )4π].

As a result the degree of spin polarization in the unperturbed state is found to be n0+ − n0−

= n0 tanh(µB0 /kB T ) for TF  T . Note that since we have assumed the Zeeman energy to

be small, we should not apply this expression for a too large argument of the tanh-function,

such that a first or second order Taylor-expansion will suffice in most cases. Next we con-

sider the opposite regime TF  T , in which case the chemical potential should be equated with

the Fermi energy kB TF , and as a consequence fˆ0 (p2 ) changes from a Maxwellian to a Fermi-

Dirac distribution (i.e. fˆ0 (p2 ) ∝ exp(−p2 /me vt2 ) changes to fˆ0 (p2 ) being constant for p2 <

p2F = me kB TF /2). Also using µB0  kB TF (to make Landau quantization negligible) we then

get c± = n0± /4π ' (n0 /2)[1 ± 3µB0 /2kB TF ]/4π. Now the spin polarization in the unperturbed

state becomes n0+ − n0− = 3n0 µB0 /2kB TF . As pointed out above the expressions presented here
Chapter 5 49

is the thermodynamic equilibrium expressions. If the unperturbed magnetic field changes faster

than the spin relaxation time, it may be adequate to apply another degree of spin polarization in the

unperturbed state (see e.g. Ref. [57]). In the present work we will limit ourselves to a background

state in thermodynamic equilibrium, however.

The linearization of equation (5.1) in the electrostatic limit gives us

∂ f1 p 2µ
0 = + · ∇x f1 + s × B0 · ∇s f1
∂t me h̄
+µ∂xi [B0 · (s + ∇s )]∂pi f1 − qE1 · ∇p f0
µ i
+ ∂ [(p × E1 ) · (s + ∇s )] ∂pi f0
2me c x

− [E1 × (s + ∇s )] × B0 · ∇p f0
2me c2
µ
+ s × (p × E1 ) · ∇s f0 . (5.10)
h̄me c

In the electrostatic limit terms proportional to B1 are dropped by definition. However, generally

for waves propagating at an angle to B0 , the perturbed magnetic field B1 is nonzero. Nevertheless

for sufficiently short wavelengths it is possible to neglect these terms. See the Appendix D for a

formal validation of this approximation.

To solve Eq. (5.10) for the perturbed distribution function f1 we make the ansatz

1
f1 = ∑ gab (p⊥ , pz , θs )ψa (φ p , p⊥ ) √ exp(−ibφs ), (5.11)
a,b=−∞ 2π
where

1
ψa (φ p , p⊥ ) = √ exp[−i(aφ p − k⊥ p⊥ sin φ p /ωce )]

∞  
1 k⊥ p⊥
= √ ∑ l me ωce exp[i(l − a)φ p ],
J (5.12)
2π l=−∞
ωce is the electron cyclotron frequency and Jl is the Bessel function of order l. Using cylindrical

coordinates for the momentum variable (p⊥ , φ p , pz ) and spherical coordinates for the spin vari-

able (θs , φs ), and following the standard techniques [57] based on the orthogonality of the ψa

eigenfunctions, we have9

Jl (x)ei(l−a)φ p Aeiφs
∞ 
f1 = ∑
a,l=−∞ 2i ω − kz pz /me − aωce + ωcg
Ce−iφs

B
+ + f0 , (5.13)
ω − kz pz /me − aωce ω − kz pz /me − aωce − ωcg
9 See Appendix C for details.
Chapter 5 50

where x = k⊥ p⊥ /me ωce and ωcg = 2µB0 /h̄ = (g/2)ωce is the spin precession factor. Note that the

spin precession frequency is very close to the electron cyclotron frequency. The expressions for

A, B and C are
  
µ ∂
A = − pz Ex sin θs + cos θs (ame ωce + kz pz ) Ja
me c ∂ θs
   
∂ 0 me ωce 2 me ωce 0
+p⊥ Ez sin θs + cos θs me ωce 2aJa − 2a Ja + a Ja − Ja
∂ θs k⊥ p⊥ k⊥ p⊥
 
me ωce ∂
+kz pz Ja0 + a Ja
k⊥ p⊥ ∂ p2
   
qµB0 p⊥ ∂ 0 me ωce ∂
+ 2
Ez sin θs + cos θs Ja + a Ja
me c ∂ θs k⊥ p⊥ ∂ p2
   
µ me ωce ∂
− Ez p⊥ Ja0 + a Ja − Ex pz Ja , (5.14)
h̄me c k⊥ p⊥ ∂ θs

 
amωce ∂
B = 4qJa Ex + Ez pz
k⊥ ∂ p2
     
2µ ∂ 0 me ωce 0 ∂
− p⊥ Ex cos θs − sin θs 2ame ωce Ja − Ja + kz pz Ja
me c ∂ θs k⊥ p⊥ ∂ p2
 
2qµB0 ame ωce ∂ ∂
− Ja Ex cos θs − sin θs , (5.15)
me c2 k⊥ ∂ θs ∂ p2

  
µ ∂
C = pz Ex sin θs + cos θs (ame ωce + kz pz ) Ja
me c ∂ θs
   
∂ me ωce me ωce
+p⊥ Ez sin θs + cos θs me ωce 2aJa0 − 2a Ja − a2 Ja + Ja0
∂ θs k⊥ p⊥ k⊥ p⊥
 
me ωce ∂
+kz pz Ja0 − a Ja
k⊥ p⊥ ∂ p2
  
qµB0 p⊥ ∂ 0 me ωce ∂
− E z sin θ s + cos θs J a − a J a
me c2 ∂ θs k⊥ p⊥ ∂ p2
   
µ 0 me ωce ∂
− Ez p⊥ −Ja + a Ja − Ex pz Ja . (5.16)
h̄me c k⊥ p⊥ ∂ θs

Note that in the above expressions for A, B and C and throughout the text below, it is understood

that the argument of the Bessel function is (k⊥ p⊥ /me ωce ), unless stated otherwise.

5.3.1 Conductivity Components for the Free Current Density

According to our notation we let all currents within linear theory be computed from a conductivity

tensor, ji = σi j E j , where ji contains the free current, the magnetization current as well as the po-

larization current. Each of these three contributions are well defined, and we may thus separate the
Chapter 5 51

magn pol
different contributions into σi j = σifree
j + σi j + σi j . Furthermore, some of the contributions to

σi j have the electric field as a direct source, whereas (if we avoid the electrostatic approximation)

some have the perturbed magnetic field as a source. Since we can always rewrite the magnetic

field in terms of the electric field using Faraday’s law, the magnetic field contributions can still be

included in σi j , which is the convention we adopt here. Nevertheless, it is of interest to separate

the terms depending on the underlaying source term, and thus all different contributions (σifree
j ,
magn pol
σi j , σi j ) can be divided into those originating from the electric field and those originating
free−E
from the (perturbed) magnetic field. Hence we make the division σifree
j = σi j + σifree−B
j where

σifree−E
j originates from the electric field terms in Eq. (5.1) whereas σifree−B
j originates from the
pol
magnetic field terms. A similar division is made of σi j . In Appendix D we show that for short
magn
wavelengths electrostatic approximations can be made, which means that the terms σi j σifree−B
j
pol−B
and σi j do not contribute to the dispersion relation, and furthermore only four out of the nine

Cartesian tensor components are needed. Given this we first calculate the four relevant compo-

nents of σifree−E
j . Substituting Eq. (5.13) into (5.6), the required components of the free part of the

conductivity tensor are found to be


16π 2 qωce

aJa me ωce
Z
free−E
σxx = ∑ dp⊥ dpz p⊥ 2qa p⊥ Ja
k⊥ a i(ω − kz pz /me − aωce ) k⊥ p⊥
(me ωce )2

µ
− p⊥ 2ame ωce Ja0 − a Ja
me c k⊥ p⊥
ˆ 2
 
0 (n0+ − n0− ) ∂ f 0 (p )
+kz pz Ja , (5.17)
n0 ∂ p2
32π 2 aωce Ja2 ∂ fˆ0 (p2 )
Z
free−E
= dp⊥ dpz p⊥ pz , (5.18)
k⊥ ∑
σxz
a i(ω − kz pz /me − aωce ) ∂ p2

16π 2 q pz Ja
Z
free−E
σzx = ∑ dp⊥ dpz p⊥ ×
me a i(ω − kz pz /me − aωce )
(me ωce )2
 
me ωce µ
2qa Ja − p⊥ 2ame ωce Ja0 − a Ja
k⊥ me c k⊥ p⊥
ˆ 2
 
0 (n0+ − n0− ) ∂ f 0 (p )
+kz pz Ja , (5.19)
n0 ∂ p2
and
32π 2 q2 Ja2 ∂ fˆ0 (p2 )
Z
σzzfree−E = dp⊥ dpz p⊥ p2z . (5.20)
me ∑ a i(ω − kz pz /me − aωce ) ∂ p2
For notational brevity the above expressions contain only the electron contribution. The ion part

is obtained trivially by dropping the spin terms proportional to µ(since the ion spin corrections
Chapter 5 52

are much smaller than the electron ones) and by replacing the electron charge and mass by the ion

values. In obtaining Eqs. (5.17-5.20), we have performed all angular integrations.

5.3.2 Conductivity Components for the Polarization Current Density

Substituting Eq. (5.13) into Eq. (5.7) we calculate the required components of the polarization
pol−E
part of the conductivity tensor - σi j . For the xx-component we have
(
4µ iµ
Z
pol−E 2
σxx = ωπ ∑ dp⊥ dpz p⊥ Ja2 ×
me c a me c(ω − k p
z z /m e − aω ce + ω cg )
 2
p (n0+ − n0− )


− p2z (ame ωce + kz pz ) 2 − z
∂p h̄ n0
2  
iµ pz 2 ∂ (n0+ − n0− )
− J (ame ωce + kz pz ) 2 −
me c(ω − kz pz /me − aωce − ωcg ) a ∂p h̄n0

2i me ωce 2 0 (n0+ − n0− )
+ 2qa p J Ja
ω − kz pz /me − aωce k⊥ p⊥ ⊥ a n0
)
µ p2⊥ 2
 
(m ω
e ce ) ∂
− 2ame ωce Ja02 − a J 0 Ja + kz pz Ja02 fˆ0 (p2 ). (5.21)
me c k⊥ p⊥ a ∂ p2
pol−E pol−E pol−E
Similarly we have also evaluated σxz , σzx and σzz , that are shown in Appendix E. In

obtaining Eq. (5.21), we have again performed all angular integrations.

In addition to polarization currents, we also have magnetization currents. However, as pointed

out above and demonstrated in Appendix D these terms do not contribute to the dispersion relation

in the electrostatic limit. Interestingly, this is not because they are necessarily small compared to

e.g. the polarization currents, but because the symmetry properties of various parts of σi j means

that cancellations occur in the dispersion relation for short wavelengths. In the further text we will

analyze the conductivity tensor in detail. If we compare the components with previous results [57],

we observe that it contains several terms that were missing in the nonrelativistic version [57] of the

theory. These terms will be crucial for the detailed dispersion relation presented in the example

below.

5.3.3 Solutions to the Dispersion Relation

There are a number of parameters that determine the relative magnitude of the quantum contri-

butions to the electrostatic dispersion relation Eq. (D.9) derived in Appendix D. These include

h̄ω p /me c2 , h̄ω p /kB T and µB B0 /kB T , where ω p is the plasma frequency. The different quantum

terms are generally proportional to some combination of these factors. If the Fermi temperature
Chapter 5 53

TF is higher than the electron temperature, T is typically replaced by TF when the order of mag-

nitude of the quantum terms are estimated. It should be noted that only the quantum relativistic

effects contribute in the electrostatic case. For the electromagnetic case when the non-relativistic

spin contributions survive, the magnitude of these terms can be significantly larger (cf. Ref. [57]).

Excluding the fully quantum relativistic regime (where h̄ω p /me c2 ∼ 1), the quantum terms of Eq.

(D.9) may still be significant if we are close to any of the resonances of the new terms. As an

illustrative example, let us consider waves with a frequency close to 2ωce propagating perpendic-

ular to the magnetic field, i.e. we let kz = 0 in which case the general dispersion relation (D.9)
free−E + σ pol−E
immediately reduces to σxx xx = iε0 ω. Furthermore, since by assumption ω ≈ 2ωce , in

the sums over a we can focus on the terms with small denominators. Due to the fact ωce ' ωcg

these are ω − (3ωce − ωcg ), ω − (ωce + ωcg ) and ω − 2ωce , and as a further consequence we are

also dropping the ion contribution. Moreover, we pick the case with T  TF , in which case we

have a Fermi-Dirac momentum distribution, and a thermodynamic equilibrium background spin

polarization given by n0+ − n0− = 3n0 µB0 /2kB TF . Under these conditions the dispersion relation

can be written
2 Z 1  (h̄ω )2 Z 1  
ω pe ωce 2
 p
2 + pe ωce 4 9 2 2 p 2

1 = 2 v2
duJ2 ζ 1 − u du u + u J1 ζ 1 − u
(ω − 2ωce ) k⊥ F −1 m2e c4 ω − (ωcg + ωce ) −1 10
(h̄ω pe )2
Z 1  
ωce 4 9 2 2 p 2

+ 2 4 du 3u + u J3 ζ 1 − u , (5.22)
me c ω − (3ωce − ωcg ) −1 10
where vF = (2kB TF /me )1/2 is the Fermi velocity, ζ = kvF /ωce (with k = k⊥ ) and we have used

the properties of the Heaviside and delta functions to carry out one of the momentum integrations,

and the remaining variable of integration is changed to u = pz /me vF . The first term of the right

hand side gives the classical-type of Bernstein resonance in a degenerate plasma, which is now

complemented by the nearby spin resonances. Solving the dispersion relation numerically for a

magnetic field of 109 T and a number density 7 × 1034 m−3 results in three wave modes that are

shown in Fig. (5.2). The frequencies approach the precise resonances for very large wavenumber,

but deviates somewhat from the resonance frequencies for smaller k. The middle curve is the

degenerate Bernstein mode modified by the coupling to the nearby spin induced modes, which

have slightly higher and lower frequencies as seen in Fig (5.2). For these modes to be coupled

as in Fig. (5.2), very strong magnetic fields are needed, as can be found e.g. in an astrophysical

context [128, 129, 130]. It should be noted that the mode structure close to other resonances are
Chapter 5 54

similar to that shown in Fig. (5.2). For each degenerate cyclotron or Bernstein resonance there

Figure 5.2: Normalized frequency ωn = ω/|ωce | plotted against normalized wavenumber kn for
Eq. (5.22). The parameter values are B0 = 109 T and n0 = 7 × 1034 m−3 .

is one spin resonance with slightly higher frequency, and one spin resonance with slightly lower

frequency.
Chapter 5 55

5.4 Conclusions

In the present chapter we have calculated the electrostatic dispersion relation in a plasma taking

into account the weakly relativistic effects associated with the electron spin using a kinetic model.

The weakly relativistic terms are typically smaller than the non-relativistic terms. However, for

electrostatic perturbations the non-relativistic spin terms vanish but the (weakly) relativistic re-

mains. These terms are typically proportional to the parameters h̄ω p /me c2 and/or µB B0 /me c2

which are very small (when compared with the terms that appear due to electron degeneracy ef-

fects) unless the electron number density and/or magnetic field is large. To some extent the effect

of the spin terms can be enhanced due to the presence of resonances, which partially reduces the

need for extremely large densities and magnetic field strengths. As investigated in section (5.3.3)

these spin resonances give rise to several new wave modes. Due to the close vicinity of some

of the spin resonances to the classical-type of degenerate Bernstein modes, we also find that the

degenerate Bernstein modes can be influenced by the spin resonances to a certain extent.

Our results suggest that the applications of the general theory should primarily be sought in

astrophysical environments such as pulsar and magnetar environments, where the magnetic field

may have values in the range B0 ∼ 106 − 1010 T [128, 129, 130] and the density can be of the order

n0 ∼ 1033 − 1036 m−3 (i.e. a thick accretion disc or a neutron star crust). To some extent inertial

confinement fusion plasmas might be of interest, although the non-relativistic spin terms is likely

to be of more interest in that regime. It should also be stressed that for the stronger magnetic

fields (i.e. magnetars with B0 ∼ 1010 T), the effects such as Landau quantization [19] should be

included, and there is also a need for a fully relativistic treatment of the spin, rather than the weakly

relativistic description used here.


Chapter 6

Overall Summary and Conclusions

Here we briefly present the overall summary of the work presented in this thesis.

In chapter 3 we have investigated the longitudinal waves in an unmagenitized electron-positron-

ion quantum (fully degenerate) plasma using kinetic theory. We have particularly studied positron

zero-sound waves and positron sound waves. Usual zero-sound waves in a degenerate electron-ion

plasma exist in the short wavelength domain k2 λFe


2  1 and remain undamped due to the fact that

particles having velocities greater than the Fermi velocity are not available which can contribute to

absorption. However, when we add a third specie (positron) in a degenerate electron-ion plasma,

scenario becomes very different. In the present study we found that positron zero-sound waves

damp provided ne0 > n p0 holds in a degenerate electron-positron-ion plasma. This is because the

Fermi velocity of particles is directly proportional to the number density in the degenerate plasmas

and ne0 > n p0 ensures that vFe > vF p . Therefore, for positron particles, electrons act as a source

of damping. Moreover, we also examined that positron zero-sound waves exist for k2 λFe
2 ' 1 and

act as a collective electrostatic mode, with or without Madelung term contribution. Therefore

we speculate that such positron zero-sound modes can be used to probe shielding phenomena of

positron sound waves. We have also discussed the damping rates of positron sound waves with

and without Madelung term and proposed the generation of such modes by a straight (classical)

electron beam. We discussed the excitation of a hydrodynamic instability in a degenerate electron-

positron-ion plasma for the beam frequency ω 0  kvtb (ω 0 is the Doppler shifted frequency of the

beam) and a kinetic instability for ω 0  kvtb . We expect that our findings may prove to be useful

in studying the complex phenomena of next generation laser-solid density plasma experiments, in

modern technology (e.g., miniaturized semiconductor devices) as well as in dense astrophysical

environments.

In chapter 4 we have presented the detailed analysis of kinetic Alfvén waves (KAWs) in a clas-

sical hot magnetized spin quantum plasma using kinetic theory. The dispersion relation for KAWs

has been obtained by employing spin-modified Vlasov-Maxwell set of equations. The general-

ized dispersion relation for KAWs captures the magnetic dipole force, spin precession dynamics
Chapter 6 57

as well as spin magnetization effects due to electrons. The two limiting cases of KAWs in the

kinetic and inertial limit have been discussed. It is found that the spin quantum effects suppress
2 > v2 ), in a hot magnetized plasma. However,
the Alfvén wave frequency in the kinetic limit (vte A
2 < v2 ), when the inertial contribution in the dispersion relation is dominat-
in the inertial limit (vte A

ing, the spin effects increasing and decreasing the frequency are equally large and cancels. The

dispersive nature of the KAWs with spin effects depends considerably on the ratios h̄ω pe /mc2 and

µB0 /kB T . This suggests that a careful analysis of Landau quantization effects as well as relativis-

tic effects can provide new insights. For a limiting case, we have also derived the spin-modified

dispersion relation of pure Alfven waves which becomes unstable for very high density plasma

(when TF  T ). However, very high density plasmas inherently contain the relativistic effects

and in the non-relativistic theory of SQPs such high densities are not allowed. This shows that

some new Alfvenic wave instabilities can be found in the dense plasma regimes which demands

to investigate the nature of KAWs in relativistic plasma environments. The analysis made here

also indicate that spin effects can be more important when the wave frequencies are increased,

approaching ∆ωce . The present results may be applicable to the astrophysical environments like

White Dwarfs or stars and Magnetars as well as to the ICF plasmas.

In chapter 5 we have extended the analysis of non-relativistic SQPs to the weakly-relativistic

regime. The linearized theory has been solved in a homogeneous magnetized plasma. Here we fo-

cused on the regime where TF  T , and used the Fermi-Dirac distribution. The basic motivation to

work in weakly-relativistic regime is to explore electron spin effects attributed to the electrostatic

plasma modes which can not be captured using non-relativistic version of SQPs. In contrast to

the non-relativistic case a prominent contribution comes from the polarization currents associated

with the electron spin in the weakly-relativistic case. We derived a general dispersion relation for

electrostatic modes for oblique propagation (kx , kz ), and calculated all the required components

of free and polarization currents. As an example, we applied the results for studying electron

Bernstein mode in a degenerate plasma. We noticed that the electron spin contributions to the

Bernstein mode solely appear due to the weakly-relativistic effects. If we drop weak-relativistic

contributions in our final result (for Bernstein wave), all the spin effects vanish and we are left

with the usual classical-type of electron Bernstein waves in a degenerate plasma. Moreover, we

found that in fully degenerate SQPs, Bernstein mode is modified by nearby resonances and some
Chapter 6 58

new modes that have frequencies very close to the Bernstein mode (but not exactly equal) can

be excited by incorporating collective electron spin effects. It turns out that the new wave modes

are localized to rather narrow frequency bands, unless the plasma density is very high and/or the

external magnetic field is very strong. An example from an astrophysical regime is presented and

the implications of our results are discussed.


Appendix A

Derivation of Eq. (3.6)

We start with the following set of linearized equations

h̄2
 
∂ δ nα ∂ foα
δ fα + (v · ∇)δ fα + −eα ∇δ φ + ∇4 = 0, (A.1)
∂t 4mα noα ∂p

4δ φ = −4π ∑ eα δ nα , (A.2)
α
dp
Z
δ nα = 2 δ fα . (A.3)
(2π h̄)3
Assuming the harmonic solutions for which all the variables vary like exp[i(k · v − ωt)], (A.1) and

(A.2) can be written as

∂ foα h̄2 k2 ∂ foα


−iωδ fα + i(v · k)δ fα − ieα δ φ k · −i δ nα k · = 0, (A.4)
∂p 4mα noα ∂p

k2 δ φ = −4π ∑ eα δ nα , (A.5)
α

From (A.4) we can write

k · ∂∂fpoα
h̄2 k2 k · ∂∂fpoα
δ f α = eα δ φ + δ nα , (A.6)
k · v − ω 4mα noα k·v−ω

Integrating both side of above equation over volume element dp we can write
Z
2
Z
2 k · ∂∂fpoα h̄2 k2
Z
2 k · ∂∂fpoα
δ fα dp = eα δ φ dp + δ n α dp, (A.7)
(2π h̄)3 (2π h̄)3 k · v − ω 4mα noα (2π h̄)3 k · v − ω

Note that the variables δ φ and δ nα can be taken out of the integration because both are function

of space and time only. LHS of (A.7) is by definition δ nα , so we can combine this term to the last

term on RHS to get


∂ foα
R 2 k· ∂ p
eα δ φ (2π h̄)3 k·v−ω
dp
δ nα = ∂ foα
, (A.8)
2 2
2 k· ∂ p
1 − 4mh̄αknoα
R
(2π h̄)3 k·v−ω
dp
which simply can be written as
eα δ φ I
δ nα = 2 2 , (A.9)
1 − 4mh̄αknoα I
where
Z
2 k · ∂∂fpoα
I= dp. (A.10)
(2π h̄)3 k · v − ω
Appendix A 60

Now one can evaluate the integral I in (A.10) for background equilibrium distribution function. In

our case, we have taken equilibrium Fermi distribution function as a step function for TF >> T .

Performing all the integrations in I we obtain


  
−3noα ω ω + kvFα
I= 1− ln . (A.11)
mv2Fα akvFα ω − kvFα

Note that while performing integrations we have introduced the standard definition noα = p3Fα /3π 2 h̄3

for Fermions . Using (A.11) back in (A.9) and inserting the resulting δ nα in (A.5) one can get
2
3ω pα
 
1 ω ω + kvFα
1 + Σα 2 2 1− ln = 0, (A.12)
k vFα Γα 2kvFα ω − kvFα

which is the required Eq. (3.6).


Appendix B

Selection of the negative root from Eq. (4.26)

We have bi-quadratic equation:

ω 4 − Bω 2 +C = 0, (B.1)

with the solution r !


2 B 4C
ω = 1± 1− 2 . (B.2)
2 B
To get the idea for the selection of proper root that gives us KAWs we make our analysis simple.

We drop all the quantum effects, and let k⊥ → 0 in (B.2) and rewrite the factors 1 − 4C
B2
and B. This

gives us
2 2 2 2 2 4 2 2 2 2 2 2 2
4C (kk vA + kk cs ) − 4kk cs vA (kk vA − kk cs )
1− 2 = = 2 2 , (B.3)
B (kk2 v2A + kk2 c2s )2 (kk vA + kk2 c2s )2
and

B = (kk2 v2A + kk2 c2s ). (B.4)

Case-I

We first take the root of (B.3) for v2A > c2s (low-beta plasma), that is
2 2 2 2
4C (kk vA − kk cs )
r
1− = . (B.5)
B2 (kk2 v2A + kk2 c2s )

As the root is always defined as a positive value, (B.5) holds for v2A > c2s . Using (B.5) and (B.4)

back in (B.2) we can write

(kk2 v2A − kk2 c2s )


!
2ω 2 = (kk2 v2A + kk2 c2s ) 1 ± . (B.6)
(kk2 v2A + kk2 c2s )

It is clear from (B.6) that positive-root gives us ω 2 = kk2 v2A and negative-root gives us ω 2 = kk2 c2s .

So for v2A > c2s we need positive-root from the solution (B.1) to get the Alfven waves.

Case-II

Here we consider (B.3) again for the case c2s > v2A (high-beta plasma). In this limit the root should

be defined as
2 2 2 2
4C (kk cs − kk vA )
r
1− 2 = 2 2 . (B.7)
B (kk vA + kk2 c2s )
Appendix B 62

Using (B.7) and (B.4) back in (B.2) we obtain

(kk2 c2s − kk2 v2A )


!
2ω 2 = (kk2 v2A + kk2 c2s ) 1 ± . (B.8)
(kk2 v2A + kk2 c2s )

It is clear from (B.8) that negative-root gives us ω 2 = kk2 v2A while positive-root gives us

ω 2 = kk2 c2s . So for c2s > v2A (high-beta plasma) we need negative-root from the solution (B.1) to get

the Alfven waves. In principle the analysis above says which sign to take. However, there might be

difference between picking signs in (B.1) without quantum and k⊥ −contributing terms and with

such contributions. Therefore, having more of an experimental approach, we can just change the

sign in front of the root sign and look for the root for which the ratio ω/kk vA goes approximately

unity because the quantum and k⊥ −contributions are small (but important) corrections to the pure

Alfven waves.
Appendix C

Derivation of perturbed distribution function: Eq. (5.13)

After taking the electrostatic approximation we start with the following linearized equation

∂ f1 p 2µ q
+ · ∇x f1 + s × B0 .∇s f1 + p × B0 .∇ p f1 =
∂t m h̄ mc
µ i
− qE1 .∇ p f0 + ∂ [(p × E1 ).(s + ∇s )] ∂ pi f0
2mc x
qµ µ
− 2
[E1 × (s + ∇s )] × B0 .∇ p f0 + s × (p × E1 ).∇s f0 , (C.1)
2mc h̄mc

We will use the perturbed distribution function f1 in the eigenfunctions as


∞ ∞
1
f1 = ∑ ∑ gab (p⊥ , pz , θs )ψ(φ p , p⊥ ) √ exp(−ibφs ), (C.2)
a=−∞ b=−∞ 2π

where

1
ψ(φ p , p⊥ ) = √ exp[−i(aφ p − k⊥ p⊥ sin φ p /ωce )]

∞  
1 k⊥ p⊥
= √ ∑ Jl ωce exp[i(l − a)φ p ]. (C.3)
2π l=−∞

Multiplying Eq. (C.1) with ψa∗ eibφs / 2π after using f1 from Eq. (C.2) and integrating over φ p

and φs we get

−i(ω − kz pz /m − aωc − bωcg )gab = Iab , (C.4)

where
ψa∗ eibφs
Z 2π Z 2π
Iab = dφ p dφs (RHS) √ . (C.5)
0 0 2π
Note that if we successfully evaluate the integrals Iab , we can use this in Eq. (C.4) to find gab

which in turn is used to find the perturbed distribution function f1 . Using cylindrical coordinates

for momentum variable and spherical polar coordinates for spin variable we can write

p = x̂p⊥ cos φ p + ŷp⊥ sin φ p + ẑpz

s = x̂ sin θs cos φs + ŷ sin θs sin φs + ẑ cos θs


∂ 1 ∂
∇s = (x̂ cos θs cos φs + ŷ cos θs sin φs − ẑ sin θs ) + (−x̂ sin φs + ŷ cos φs )
∂ θs sin θs ∂ φs
Appendix C 64

Taking into account (C.1) and using the above coordinate transformations we obtain term by term

as
q ∂
(p × B0 ).∇ p = −ωc (C.6)
mc ∂ φp
2µ ∂
(s × B0 ).∇s = −ωcg (C.7)
h̄ ∂ φs
 

−qE1 .∇ p f0 = −q Ex p⊥ cos φ p + Ez pz 2 2 f0 (C.8)
∂p

µ i
+ ∂ [(p × E1 ).(s + ∇s )]∂ pi f0 =
2mc x 
µ ∂ ∂
i pz Ex sin φs (sin θs + cos θs ) − p⊥ Ex sin φ p (cos θs − sin θs )
2mc ∂ θs ∂ θs

∂ ∂
+p⊥ Ez (cos φs sin φ p − sin φs cos φ p )(sin θs + cos θs ) (kx p⊥ cos φ p + kz pz )2 2 f0 (C.9)
∂ θs ∂p


− [E1 × (s + ∇s )] × B0 .∇ p f0 =
2mc2 
qµ ∂
2
B0 p⊥ Ex cos φ p (cos θs − sin θs )
2mc ∂ θs

∂ ∂
−Ez (sin φ p sin φs + cosφ p cos φs )(sin θs + cos θs ) 2 2 f0 (C.10)
∂ θs ∂p
 
µ µ ∂
+ s × (p × E1 ).∇s f0 = Ez P⊥ (sin φ p sin φs + cosφ p cos φs ) − cos φs pz Ex f0 (C.11)
h̄mc h̄mc ∂ θs
Due to harmonic exponentials on the RHS of Eq. (C.5), we can perform integrations easily in

Iab involving sine and cosine. We write down results for each term separately, after performing

integrations over φ p and φs on the RHS of Eq. (C.5), that are


 

term1 = −q Ex p⊥ cos φ p + Ez pz 2 2 f0 →
∂p
 
−q ∂
Ex p⊥ (πδl=a∓1 )(2πδb=0 ) + Ez pz (2πδl=a )(2πδb=0 ) 2 2 f0 , (C.12)
2π ∂p
Appendix C 65


µ ∂
term2 = i pz Ex sin φs (sin θs + cos θs )(kx p⊥ cos φ p + kz pz )
2mc ∂ θs

−p⊥ Ex sin φ p (cos θs − sin θs )(kx p⊥ cos φ p + kz pz )
∂ θs

∂ ∂
+p⊥ Ez (cos φs sin φ p − sin φs cos φ p )(sin θs + cos θs )(kx p⊥ cos φ p + kz pz ) 2 2 f0 →
∂ θs ∂p

iµ ∂
pz Ex (sin θs + cos θs )(ibπδb=±1 ){kx p⊥ πδl=a∓1 + kz pz 2πδl=a }
2π(2mc) ∂ θs
∂ π
−p⊥ Ex (cos θs − sin θs ){kx p⊥ i(a − l) δl=a∓2 + kz pz i(a − l)πδl=a∓1 }2πδb=0
∂ θs 2

∂ π π
+p⊥ Ez (sin θs + cos θs ) kx p⊥ {πδb=±1 i(a − l) δl=a∓2 − ibπδb=±1 (πδl=a + δl=a∓2 )}
∂ θs 2 2


+kz pz {πδb=±1 i(a − l)πδl=a∓1 − ibπδb=±1 πδl=a∓1 } 2 2 f0 , (C.13)
∂p


qµ ∂
term3 = 2
B0 p⊥ Ex cos φ p (cos θs − sin θs )
2mc ∂ θs

∂ ∂
−Ez (sin θs + cos θs )(sin φ p sin φs + cosφ p cos φs ) 2 2 f0 →
∂ θs ∂p

qµ ∂
2
B0 p⊥ Ex (cos θs − sin θs )(πδl=a∓1 )(2πδb=0 )
2π(2mc ) ∂ θs

∂ ∂
−Ez (sin θs + cos θs ){i(a − l)πδl=a∓1 ibπδb=±1 + πδl=a∓1 πδb=±1 } 2 2 f0 , (C.14)
∂ θs ∂p

 
µ ∂
term4 = Ez P⊥ (sin φ p sin φs + cosφ p cos φs ) − cos φs pz Ex f0 →
h̄mc ∂ θs

µ
Ez P⊥ {ibπδb=±1 i(a − l)πδl=a∓1 + πδb=±1 πδl=a∓1 } (C.15)
2π(h̄mc)


−Ex pz (πδb=±1 2πδl−a=0 ) f0 ,
∂ θs

where symbol δ in all results written above shows a Kronecker-Delta (It is not a Dirac-Delta!).

Note that summation over l in Eqs. (C.12)-(C.15) has not been performed yet. For performing

summation over l we use the following properties of Bessel functions

Ja (x)
Ja+1 (x) + Ja−1 (x) = 2a (C.16)
x

−Ja+1 (x) + Ja−1 (x) = 2Ja0 (x) (C.17)

which in our case leads to

δl=a Jl → Ja (x) (C.18)


2a
δl=a∓1 Jl → Ja (x) (C.19)
x
Appendix C 66

(a − l)δl=a∓1 Jl → 2Ja0 (x) (C.20)


1 2a2 2
δl=a + δl=a∓2 Jl → 2 Ja (x) − Ja0 (x) (C.21)
2 x x
 
2a 0 2a
(a − l)δl=a∓2 Jl → 4 J (x) − 2 Ja (x) (C.22)
x a x
k⊥ p⊥
where x stands for mωce .
After using the above definitions, Eqs. (C.12)-(C.15) can be rewritten as
 
mωce ∂
term1 = −qπ 2a Ex p⊥ Ja (x) + 2Ez pz Ja (x) δb=0 2 2 f0 (C.23)
k⊥ p⊥ ∂p
  
−π µ ∂ mωce
term2 = pz Ex (sin θs + cos θs )(bδb=±1 ) kx p⊥ 2a Ja + 2kz pz Ja (C.24)
4mc ∂ θs k⊥ p⊥
   
∂ mωce 0 mωce 2 0
−p⊥ Ex (cos θs − sin θs )2δb=0 kx p⊥ 2 2a J − 2a( ) Ja + kz pz 2Ja
∂ θs k⊥ p⊥ a k⊥ p⊥
     
∂ mωce 0 mωce 2 mωce 2 mωce 0
+p⊥ Ez (sin θs + cos θs )δb=±1 kx p⊥ 2 2a Ja − 2a( ) Ja − b 2a2 ( ) Ja − 2 Ja
∂ θs k⊥ p⊥ k⊥ p⊥ k⊥ p⊥ k⊥ p⊥
 
mωce ∂
+kz pz 2Ja0 − b 2a Ja 2 2 f0
k⊥ p⊥ ∂p

πqµB0 p⊥ ∂ mωce
term3 = 2
2Ex (cos θs − sin θs ){2a Ja (x)}δb=0
4mc ∂ θs k⊥ p⊥

∂ mωce ∂
+Ez (sin θs + cos θs ){b2Ja (x)0 − 2a Ja (x)}δb=±1 2 2 f0 (C.25)
∂ θs k⊥ p⊥ ∂p
 
πµ 0 mωce ∂
term4 = Ez P⊥ {−b2Ja (x) + 2a Ja }δb=±1 − 2Ex pz Ja (x)δb=±1 f0 (C.26)
2h̄mc k⊥ p⊥ ∂ θs
Using Eqs. (C.23)-(C.26) in Eq. (C.4) we obtain
"  
π mωce ∂
gab = −q 2a Ex p⊥ Ja (x) + 2Ez pz Ja (x) δb=0 2 2
−i(ω − kz pz /m − aωce − bωcg ) k⊥ p⊥ ∂p
  
µ ∂ mωce
− pz Ex (sin θs + cos θs )(bδb=±1 ) kx p⊥ 2a Ja + 2kz pz Ja
4mc ∂ θs k⊥ p⊥
   
∂ mωce 0 mωce 2 0
−p⊥ Ex (cos θs − sin θs )2δb=0 kx p⊥ 2 2a J − 2a( ) Ja + kz pz 2Ja
∂ θs k⊥ p⊥ a k⊥ p⊥
     
∂ mωce 0 mωce 2 2 mωce 2 mωce 0
+p⊥ Ez (sin θs + cos θs )δb=±1 kx p⊥ 2 2a J − 2a( ) Ja − b 2a ( ) Ja − 2 J
∂ θs k⊥ p⊥ a k⊥ p⊥ k⊥ p⊥ k⊥ p⊥ a
 
0 mωce ∂
+kz pz 2Ja − b 2a Ja 2 2
k⊥ p⊥ ∂p

qµB0 p⊥ ∂ mωce
+ 2
2Ex (cos θs − sin θs ){2a Ja (x)}δb=0
4mc ∂ θs k⊥ p⊥

∂ 0 mωce ∂
+Ez (sin θs + cos θs ){b2Ja (x) − 2a Ja (x)}δb=±1 2 2
∂ θs k⊥ p⊥ ∂p
  #
µ mω ce ∂
+ Ez P⊥ {−b2Ja0 (x) + 2a Ja }δb=±1 − 2Ex pz Ja (x)δb=±1 f0 (C.27)
2h̄mc k⊥ p⊥ ∂ θs
Appendix C 67

Using Eq. (C.27) in Eq. (C.2) we obtain the following perturbed distribution function
" 
Jl ei(l−a)φ p e−ibφs
∞ 
mωce ∂
f1 = ∑ −2i(ω − kz pz /m − aωce − bωcg ) −q 2a k⊥ p⊥ Ex p⊥ Ja (x) + 2Ez pz Ja (x) δb=0 2 ∂ p2
a,b,l=−∞
  
µ ∂ mωce
− pz Ex (sin θs + cos θs )(bδb=±1 ) kx p⊥ 2a Ja + 2kz pz Ja
4mc ∂ θs k⊥ p⊥
   
∂ mωce 0 mωce 2 0
−p⊥ Ex (cos θs − sin θs )2δb=0 kx p⊥ 2 2a J − 2a( ) Ja + kz pz 2Ja
∂ θs k⊥ p⊥ a k⊥ p⊥
     
∂ mωce 0 mωce 2 2 mωce 2 mωce 0
+p⊥ Ez (sin θs + cos θs )δb=±1 kx p⊥ 2 2a J − 2a( ) Ja − b 2a ( ) Ja − 2 J
∂ θs k⊥ p⊥ a k⊥ p⊥ k⊥ p⊥ k⊥ p⊥ a
 
mωce ∂
+kz pz 2Ja0 − b 2a Ja 2 2
k⊥ p⊥ ∂p

qµB0 p⊥ ∂ mωce
+ 2
2Ex (cos θs − sin θs ){2a Ja (x)}δb=0
4mc ∂ θs k⊥ p⊥

∂ 0 mωce ∂
+Ez (sin θs + cos θs ){b2Ja (x) − 2a Ja (x)}δb=±1 2 2
∂ θs k⊥ p⊥ ∂p
  #
µ 0 mωce ∂
+ Ez p⊥ {−b2Ja (x) + 2a Ja }δb=±1 − 2Ex pz Ja (x)δb=±1 f0 (p2 , θs ) (C.28)
2h̄mc k⊥ p⊥ ∂ θs

After performing the sum over b we get



Jl ei(l−a)φ p
Ãeiφs + B̃ + C̃e−iφs f0

f1 = ∑ (C.29)
a,l=−∞ 2i

where

A B C
à = , B̃ = , C̃ =
(ω − kz pz /m − aωcα + ωcg ) (ω − kz pz /m − aωcα ) (ω − kz pz /m − aωcα − ωcg )
(C.30)

where the expressions for A, B and C are already exposed in the text of chapter 5.
Appendix D

Validation of the electrostatic limit for waves propagating


at an angle to B0

In general for linear waves in homogeneous medium we can write the current density as

ji = σi j E j , (D.1)

where the conductivity tensor σi j = σi j (ω, k) and we have used the summation convention. The

conductivity tensor as defined here includes all currents, i.e. free currents, magnetization currents

and polarization currents. Eliminating the magnetic field in Ampere’s law using Faraday’s law we

obtain
k 2 c2 ki k j c2 iσi j
  
1− 2 δi j − + E j = 0. (D.2)
ω ω2 ε0 ω
The non-trivial solutions is obtained by putting the determinant to zero, which give the dispersion

relation as

k2 c2 ki k j c2 iσi j
  
det 1− 2 δi j − + = 0. (D.3)
ω ω2 ε0 ω
While for an oblique direction of propagation, the waves are typically not exactly electrostatic, we

can let k2 c2 /ω 2 → ∞ to get the electrostatic limit, assuming the conductivity tensor to be of order

unity in this expansion parameter. To order k6 c6 /ω 6 we get an identity that is always fulfilled and

to order k4 c4 /ω 4 we get

k2 c4 2
[k (σxx − iε0 ω) + k⊥ kz (σxz + σzx ) + kz2 (σzz − iε0 ω)] = 0, (D.4)
ω4 ⊥

where the wave-vector is assumed to lie in the xz-plane for convenience. Eq. (D.4) shows that

only four conductivity tensor components are needed, and agrees with the dispersion relation one

gets by assuming E = −∇Φ from the start, and using Poisson’s equation rather than the full set

of Maxwell’s equations. While the above derivation of the electrostatic limit did not rely on any

specific model for the particles such as e.g. the Vlasov equation, it still does not apply directly to

the spin-kinetic model we are interested in. The reason is that the assumption that σi j is of order

unity in our expansion parameter k2 c2 /ω 2 does not apply. In particular, certain terms in σi j → ∞
Appendix D 69

as k2 c2 /ω 2 → ∞. To address our situation, it is therefore convenient to divide the current density

into different contributions as


magn pol
σi j = σifree
j + σi j + σi j , (D.5)

where the three conductivity tensor components are the ones resulting from the free current den-

sity, the magnetization current density and the polarization current density, respectively. The terms

in the conductivity tensor diverging when k2 c2 /ω 2 → ∞ is the ones proportional to the magnetiza-
magn
tion current density, i.e. certain terms in σi j . Specifically the diverging terms σidj can be written

[57]
−kz2 ikz2
 
2 k⊥ kz
c
σidj = C 2  −ikz2 −kz2 ik⊥ kz  , (D.6)
ω 2
k⊥ kz −ik⊥ kz k⊥
where C is a non-diverging scalar. However, although the individual terms become large, the

structure of the matrix σidj implies that its contribution to the total dispersion relation nevertheless

vanish to the order k4 c4 /ω 4 .

Next we note that the magnetic field gives a contribution to the free current density and also to

the polarization current density. These contributions come partly from the magnetic dipole force

term and partly from the spin-precession term in Eq. (5.1) of the main text. Thus we can divide

σifree free−E
j = σi j + σifree−B
j , (D.7)

where the second term comes from the source term that explicitly depend on the perturbed mag-

netic field in (5.1). A similar division can be made of the polarization current density. However,

since Bi = εi jk k j Ek /ω, the contribution σifree−B


j will be of the form

jifree−B = χifree−B
j B j = χifree−B
j ε jkl k j Ek /ω, (D.8)

such that σifree−B


j = χifree−B
j ε jkl k j where χifree−B
j is a magnetic susceptibility tensor whose detailed

properties are irrelevant for our discussion. The important point is that the antisymmetry property

of the Levi-Chivita tensor εi jk means that the contribution from σifree−B


j cancel when the electro-

static limit is taken (i.e. the contributions to the general dispersion relation in Eq. (D.3) cancel to
pol−B
order k4 c4 /ω 4 ). An identical argument holds for σi j . Thus in order to calculate the general

dispersion relation we need to find four tensor components (σxx , σxz , σzx and σzz as in Eq. (D.4)),
pol−E
and only the direct contribution from the electric field, σifree−E
j rather than σifree
j and σi j rather
Appendix D 70

pol
than σi j . Taking this into account keeping terms up to order k4 c4 /ω 4 we find that Eq. (D.3)

reduces to

iε0 ω = sin2 θ σxx


free−E pol−E

+ σxx
free−E pol−E free−E pol−E
+ cos2 θ σzzfree−E + σzzpol−E ,
 
+ sin θ cos θ σxz + σxz + σzx + σzx (D.9)

where the angle θ is defined by k = k⊥b


x + kzb
z = k(sin θb
x + cos θb
z). Eq. (D.9) can be derived more

directly (but less rigorously) from Poisson’s equation. It should be noted that while the full tensor

σi j is Hermitian when the pole contributions are neglected, the same is not true for the sub-parts,

σifree−E
j etc. As a result we cannot invoke symmetry arguments for the tensor and must compute
free−E and σ free−E , and similarly for the polarization current contribution.
both σxz zx
Appendix E

Remaining conductivity components for the polarization


current density that are needed

The remaining components of the conductivity tensor emanating from the polarization currents

that are needed are:


(
4µωπ 2 i µ p⊥ pz
Z
pol−E
σxz = ∑ dp⊥ dpz × (E.1)
me c a (ω − kz pz /me − aωce + ωcg ) me c
(me ωce )2
 
0 2 0 me ωce kz pz 2 ∂
2ame ωce Ja Ja − a (1 − a)Ja + kz pz Ja Ja + a Ja
k⊥ p⊥ k⊥ p⊥ ∂ p2
  
me ωce 2 (n0+ − n0− )
+ Ja Ja0 + a J
k⊥ p⊥ a h̄n0
i µ p⊥ pz
− ×
(ω − kz pz /me − aωce − ωcg ) me c
(me ωce )2
 
me ωce kz pz 2 ∂
2ame ωce Ja Ja0 − a )(1 + a)Ja2 + kz pz Ja Ja0 − a Ja
k⊥ p⊥ k⊥ p⊥ ∂ p2
  
me ωce (n0+ − n0− )
− Ja0 − a Ja
k⊥ p⊥ h̄n0
)
4i 0 (n0+ − n0− ) ∂
+ qp⊥ pz Ja Ja fˆ0 (p2 )p⊥ ,
(ω − kz pz /me − aωce ) n0 ∂ p2

(
4µ 2 ωπ 2 i
Z
pol−E 2
σzx = 2 2 ∑ dp⊥ dpz p⊥ pz × (E.2)
3me c a (ω − kz pz /me − aωce + ωcg )
2
 
2 (me ωce ) 2 me ωc 2 0 0 ∂
a Ja + kz pz a Ja + ame ωce Ja Ja + kz pz Ja Ja
k⊥ p⊥ k⊥ p⊥ ∂ p2
  
me ωc 2 (n0+ − n0− )
+ Ja Ja0 + a J
k⊥ p⊥ a h̄n0
i
− ×
(ω − kz pz /m − aωce − ωcg )
(me ωce )2 2
 
me ωc 2 ∂
ame ωce Ja Ja0 + kz pz Ja Ja0 − a2 ( Ja − kz pz a Ja
k⊥ p⊥ k⊥ p⊥ ∂ p2
  )
me ωc 2 (n0+ − n0− )
− Ja Ja0 − a J fˆ0 (p2 )
k⊥ p⊥ a h̄n0
Appendix E 72

and
(
4µ 2 ωπ 2 −i
Z
σzzpol−E = 2 2 ∑ dp⊥ dpz p3⊥ × (E.3)
3me c a (ω − kz pz /me − aωce + ωcg )
(me ωce )2 02 (me ωce )3 2

2ame ωce Ja02 − a(1 − 3a) Ja Ja (1 − a) J
k⊥ p⊥ (k⊥ p⊥ )2 a
me ωce kz pz 02 me ωce kz pz 2 2 ∂
  
02
+kz pz Ja + 2a Ja Ja Ja
k⊥ p⊥ k⊥ p⊥ ∂ p2
  2  
02 me ωce 02 me ωce 2 (n0+ − n0− )
+ Ja + 2a Ja Ja Ja
k⊥ p⊥ k⊥ p⊥ h̄n0
i
− ×
(ω − kz pz /me − aωce − ωcg )
(me ωce )2 02 (me ωce )3 2

2ame ωce Ja02 − a(1 + 3a) Ja Ja (1 + a) J
k⊥ p⊥ (k⊥ p⊥ )2 a
me ωce kz pz 02 me ωce kz pz 2 2 ∂
  
02
+kz pz Ja − 2a Ja Ja Ja
k⊥ p⊥ k⊥ p⊥ ∂ p2
)
me ωce 02 me ωce 2 2 (n0+ − n0− )
   
02
− Ja − 2a Ja Ja Ja fˆ0 (p2 ).
k⊥ p⊥ k⊥ p⊥ h̄n0
Bibliography

[1] D. Pines and D. Bohm, Phys. Rev. 85, 338 (1952).

[2] D. Bohm and D. Pines, Phys. Rev. 92, 609 (1953).

[3] D. Pines, Phys. Rev. 92, 626 (1953).

[4] D. Pines, J. Nucl. Energy, Part C: Plasma Physics, Vol 2, pp. 5 to 17 (1961).

[5] E. A. Rauscher, J. Plasma Physics 2, 517 (1968).

[6] G. Manfredi, How to model Quantum Plasmas, Fields Inst. Commun. 46, 263 (2005).

[7] M. Bonitz, A. Filinov, J. Boning, J. W. Dufty, in Introduction to Complex Plasmas

edited by M. Bonitz, N. Horing, P. Ludwig, Springer-Verlag Berlin Heidelberg (2010) p. 41.

[8] C. L. Gardner, J. Appl. Math. 54, 409 (1994).

[9] L. S. Kuz’menkov and S. G. Maksimov, Theo. Math. Phys. 118, 227 (1999).

[10] G. Manfredi and F. Haas, Phys. Rev. B 64, 075316 (2001).

[11] F. Hass, Phys. Plasmas 12, 062117 (2005).

[12] M. Marklund and G. Bordin, Phys. Rev. Lett. 98, 025001 (2007).

[13] B. Eliasson and P. K. Shukla, Phys. Scr. 78, 025503 (2008).

[14] N. Crouseilles, P.-A. Hervieux, G. Manfredi, Phys. Rev. B 78, 155412 (2008).

[15] F. Haas, Quantum Plasmas, An Hydrodynamic Approach, Springer Science+Business Media

(2011).
Bibliography 74

[16] G. Manfredi and M. R. Feix, in Advances in Kinetic Theory and Computing: Selected Papers

edited by B. Perthame, World Scientific Publishing Co. Pte. Ltd. (1994) p. 109.

[17] M. V. Kuzelev and A. A. Rukhadze, Physics-Uspekhi 46, 603 (1999).

[18] N. L. Tsintsadze and L. N. Tsintsadze, Euro. Phys. Lett. 88, 35001 (2009).

[19] J. Zamanian, M. Marklund, G. Brodin, New J. Phys. 12, 043019 (2010)

[20] M. Bonitz, AIP Conf. Proc. 1421, 135 (2012).

[21] P. K. Shukla and B. Eliasson, Plasma Phys. Control. Fusion 52, 124040 (2010).

[22] P. K. Shukla and B. Eliasson, Physics-Uspekhi 53, 51 (2010).

[23] P. K. Shukla and B. Eliasson, Rev. Mod. Phys. 83, 885 (2011).

[24] G. Brodin, M. Marklund, J. Zamanian, M. Stefan, Plasma Phys. Control. Fusion, 53, 074013

(2011).

[25] V.E . Fortov, I. T. Iakubov, A. G. Khrapak, Physics of Strongly Coupled Plasma, Oxford

University Press Inc., New York (2006), p. 2.

[26] D. R. Nicholson, Introduction to Plasma Theory, John Wiley and Sons, Inc. (1983) p. 37.

[27] M. Henon, Astron. Astrophys. 114, 211 (1982).

[28] R. L. Liboff, Kinetic Theory: Classical, Quantum, and Relativistic Descriptions, Springer-

Verlag New York, Inc. (2003) p. 341.

[29] E. Wigner, Phys. Rev. 40, 749 (1932).

[30] J. H. Irving and R. W. Zwanzig, J. Chem. Phys. 19, 1173 (1951).

[31] Y. L. Klimontovich and V. P. Silin, Sov. Phys. Usp. 3, 84 (1960).

[32] V. I. Tatarskii, Sov. Phys. Usp. 26, 311 (1983).

[33] A. Arnold and H. Steinrck, Z. Angew. Math. Phys. 40, 793 (1989).

[34] Hai-Woong Lee, Phys. Rep. 259 147 (1995).


Bibliography 75

[35] R. E. Wyatt, Quantum Dynamics with Trajectories,Springer Science+Business Media, Inc.,

New York (2005), p. 68.

[36] T. B. Materdey and C. E. Seyler, Int. J. Mod. Phys. B 17, 4555 (2005).

[37] N. L. Tsintsadze, AIP Conf. Proc. 1306, 75 (2010).

[38] J. E. Moyal, in Quantum Mechanics in Phase Space: An Overview with Selected Papers,

edited by C. K. Zachos, D. B. Fairlie, T. L. Curtright, World Scientific Publishing Co. Pte.

Ltd. (2005) p. 167.

[39] H. Hillery, R. F. O’Connell, M. O. Scully, E. P. Wigner, Phys. Rep. 106, 121 (1984).

[40] G. Brodin, M. Marklund, J. Zamanian, A. Ericsson, P. L. Mana, Phys. Rev. Lett. 101, 245002

(2008).

[41] F. A Asenjo, J. Zamanian, M. Marklund, G. Brodin, P. Johansson, New J. Phys. 14, 073042

(2012).

[42] S. C. Cowley, R .M. Kulsrud, E. Valeo, Phys. Fluids 29, 430 (1986).

[43] R. M. Kulsrud, E. J. Valeo, S. C. Cowley, Nucl. Fusion 26, 1443 (1986).

[44] W. R. Chappell, Phys. Rev. 152, 113 (1966).

[45] D. B. Melrose, Plasma Phys. 16, 845 (1974).

[46] A. Jungel, Transport Equations for Semiconductors, Lect. Notes Phys. 773, Springer, Berlin

Heidelberg (2009).

[47] L. D. Landau and E. M. Lifshitz, Statistical Physics Part 2, Pergamon Press Ltd. (1980) p.

14.

[48] P. Strange, Relativistic Quantum Mechanics with Applications in Condensed Matter and

Atomic Physics, Cambridge University Press, (1998) p. 117.

[49] F. Haas, G. Manfredi, M. Feix, Phys. Rev. E 62, 2763 (2000).

[50] F. Haas, G. Manfredi, J. Goedert, Phys. Rev. E 64, 026413 (2001).


Bibliography 76

[51] F. Haas, G. Manfredi, J. Goedert, Braz. J. Phys. 33, 128 (2003).

[52] F. Haas, L. G. Garcia, J. Goedert, G. Manfredi, Phys. Plasmas 10, 3858 (2003).

[53] D. Anderson, B. Hall, M. Lisak, M. Marklund, Phys. Rev. E 65, 046417 (2002).

[54] L. N. Tsintsadze and N. L. Tsintsadze, J. Plasma Phys. 76, 403 (2010); e-print arXiv:

physics/0911.4788v1.

[55] L. N. Tsintsadze, AIP Conf. Proc. 1306, 89 (2010).

[56] J. Zamanian, M. Stefan, M. Marklund, G. Brodin, Phys. Plasmas 17, 102109 (2010).

[57] J. Lundin and G. Brodin, Phys. Rev. Lett. 82, 056407 (2010).

[58] M. Stefan, J. Zamanian, G. Brodin, A. P. Misra, M. Marklund, Phys. Rev. E 83, 036410

(2011).

[59] M. Stefan and G. Brodin, Phys. Plasmas 20, 012114 (2013).

[60] G. Brodin and M. Stefan, Phys. Rev. E 88, 023107 (2013).

[61] C. M. Surko and T. J. Murphy, Phys. Fluids B, 2, 1372 (1990).

[62] N. Iwamoto, Phys. Rev. E 47, 604 (1993).

[63] G. P. Zank and R. G. Greaves, Phys. Rev. E 51, 6079 (1995).

[64] V. I. Berezhiani and S. M. Mahajan, Phys. Rev. Lett. 73, 1110 (1994).

[65] V. I. Berezhiani, M. Y. El-Ashry, U. A. Mofiz, Phys. Rev. E 50, 448 (1994).

[66] V. I. Berezhiani and S. M. Mahajan, Phys. Rev. E 52, 1968 (1995).

[67] S. I. Popel, S. V. Vladimirov, P. K. Shukla, Phys. Plasmas 2, 716 (1995).

[68] W. P. Leemans, R. Duarte, E. Esarey, S. Fournier, C. G. R. Geddes, D. Lockhart, C. B.

Schroeder, C. Toth, J. L. Vay, S. Zimmermann, AIP Conf. Proc. 1299, 3 (2010).

[69] C. P. Ridgers, C. S. Brady, R. Duclous, J. G. Kirk, K. Bennett, T. D. Arber, A. R. Bell, Phys.

Plasmas 20, 056701 (2013).


Bibliography 77

[70] A. J. Kemp and L. Divol, Phys. Rev. Lett. 109, 195005 (2012).

[71] I. I. Goldman, Zh. Eksp. Teor. Fiz. 17, 681 (1947).

[72] Y. L. Klimontovich and V. P. Silin, Zh. Eksp. Teor. Fiz. 23, 151 (1952).

[73] D. Bohm, Phys. Rev. 85, 166 (1952).

[74] D. Bohm and D. Pines, in Plasma Physics, edited by J.E.Drummond (New York: McGraw-

Hill, 1961).

[75] N. L. Tsintsadze and L. N. Tsintsadze, AIP Conf. Proc. 1177, 18 (2009).

[76] B. Eliasson and P. K. Shukla, J. Plasma Phys. 76, 7 (2010).

[77] L. N. Tsintsadze and N. L. Tsintsadze, J. Plasma Phys. 76, 403 (2010).

[78] L. N. Tsintsadze, e-print arXiv: http://arxiv.org/abs/0911.0133v1.

[79] L. N. Tsintsadze, e-print arXiv: http://arxiv.org/abs/1005.3408v1.

[80] S. Ali and P. K. Shukla, Phys. Plasmas 13, 022313 (2006); ibid, 13, 129902 (2006).

[81] S. Ali, W. M. Moslem, P. K. Shukla, R. Schlickeiser, Phys. Plasmas 14, 082307 (2007).

[82] A. Mushtaq and S. A. Khan, Phys. Plasmas 14, 052307 (2007).

[83] P. Chatterjee, K. Roy, G. Mondal, S. V. Muniandy, S. L. Yap, C. S. Wong, Phys. Plasmas 16,

122112 (2009).

[84] E. F. El-Shamy, R. Sabry, W. M. Moslem, P. K. Shukla, Phys. Plasmas 17, 082311 (2010).

[85] A. S. Bains, A. P. Misra, N. S. Saini, T. S. Gill, Phys. Plasmas 17, 012103 (2010).

[86] L. D. Landau and E. M. Lifshitz, Statistical Physics, 3rd Edition, Part 1 (Butterworth-

Heinemann, Oxford, 1998).

[87] D. Kremp, M. Schlanges, W.D.Kraeft, Quantum Statistics of Nonideal Plasmas (Springer-

Verlag, Berlin Heidelberg, 2005).

[88] N. L. Tsintsadze, A. Rasheed, H. A. Shah, G. Murtaza, Phys. Plasmas 16, 112307 (2009).
Bibliography 78

[89] L. D. Landau, Zh. Eksp. Teor. Fiz. 30, 1058 (1956); ibid, 35, 97 (1958).

[90] A. F. Alexandrov, L. S. Bogdankevich, A. A. Rukhadze, Principals of Plasma Electrodynam-

ics (Springer, Heidelberg, 1984).

[91] H. Alfvén, Nature (London) 150, 405 (1942).

[92] L. Chen, Plasma Phys. Control. Fusion 50, 124001 (2008).

[93] Neil F. Cramer, The Physics of Alfvén Waves, WILEY-VCH Verlag Berlin GmbH, Berlin

(2001).

[94] W. Gekelman, S. Vincena, B. Van Compernolle, G. J. Morales, J. E. Maggs, P. Pribyl, T. A.

Carter, Phys. Plasmas 18 , 055501 (2011).

[95] R. L. Lysak and W. Lotko, J. Geophys. Res. 101, 5085 (1996).

[96] M. F. Bashir, Z. Iqbal, I. Aslam, G. Murtaza, Phys. Plasmas 17, 102112 (2010).

[97] A. Hasegawa and C. Uberoi, The Alfvén Wave, DOE Critical Review Series, Advances in

Fusion Science and Engineering, Technical Information Service. US Department of Energy,

Washington, DC (1982).

[98] V. M. Chmyrev, S. V. Bilichenko, O. A. Pokhotelov, V. A. Marchenko, V. I. Lazarev, A. V.

Streltsov, L. Stenflo, Phys. Scripta 38, 841 (1988).

[99] C. C. Chaston, C.W. Carlson, W. J.Peria, R. E. Ergun, J. P. McFadden, Geophys. Res. Lett.

26, 647 (1999).

[100] D. Leneman, W. Gekelman, J. Maggs, Phys. Rev. Lett. 82, 2673(1999).

[101] C. A. Kletzing, S. R. Bounds, J. M-Hiner, W. Gekelman, C. Mitchell, Phys. Rev. Lett. 90,

035004 (2003).

[102] A. Hasegawa and L. Chen, Phys. Fluids 19, 1924 (1976).

[103] D. W. Ross, G. L. Chen, S. M. Mahajan, Phys. Fluids 25, 652 (1982) .

[104] L. Huang, X. M. Qin, N. Ding, Y. X. Long, Chin. Phys. Lett. 8, 232 (1991).
Bibliography 79

[105] D. J. Wu and J. K. Chao, Phys. Plasmas 10, 3787 (2003).

[106] D. J. Wu and J. K. Chao, J. Geophys. Res. 109, A06211 (2004).

[107] D. J. Wu and C. Fang, Astrophys. J. 659, L181 (2007).

[108] D. J. Wu and C. Fang, Astrophys. J. 596, 656 (2003).

[109] P.K. Shukla and L. Stenflo, Phys. Scripta T60, 32 (1995).

[110] L. G. Garcia, F. Haas, L. P. L. de Oliveira, J. Goedert, Phys. Plasmas 12, 012302 (2005).

[111] H. Ren, Z. Wu, P. K. Chu, Phys. Plasmas 14, 062102 (2007).

[112] F. Haas, G. Manfredi, P. K. Shukla, P.-A. Hervieux, Phys. Rev. B 80, 073301 (2009).

[113] N. Shukla, P. K. Shukla, B. Eliasson, L. Stenflo, Phys. Lett. A 374, 1749 (2010).

[114] Mubashar Iqbal, J. Plasma Phys. 79, 19 (2012).

[115] A. Mushtaq and D. B. Melrose, Phys. Plasmas 16, 102110 (2009).

[116] A. Mushtaq and S. V. Vladimirov, Phys. Plasmas 17, 102310 (2010).

[117] A. J. Keane, A. Mushtaq, M. S. Wheatland, Phys. Rev. E 83, 066407 (2011).

[118] U. Wagner, M. Tatarakis, A. Gopal, F. N. Beg, E. L. Clark, A. E. Dangor, R. G. Evans, M.

G. Haines, S. P. D. Mangles, P. A. Norreys, M.-S.Wei, M. Zepf, K. Krushelnick, Phys. Rev.

E 70, 026401 (2004).

[119] Robert Eisberg, Quantum Physics of Atoms, Molecules, Solids, Nuclei, and Particles, Sec-

ond Edition, John Wiley and Sons (1985) p.278. This material is reproduced with permission

of John Wiley and Sons, Inc.

[120] G. Gumbs, Phys. Rev. B 70, 235314 (2004).

[121] P. A. Andreev and L. S. Kuz’menkov, in Progress In Electromagnetics Research Symposium

Proceedings, Marrakesh, Morocco, 1047 (2011).

[122] Dae-Han Ki and Young-Dae Jung, Appl. Phys. Lett. 99, 121506 (2011).
Bibliography 80

[123] M. I. Trukhanova, Intl. J. Mod. Phys. B 26, 1250004 (2012).

[124] M. I. Trukhanova, Eur. Phys. J. D 67:32 (2013).

[125] Ze Cheng and Shan Wu, Eur. Phys. Lett. 100 45004 (2012).

[126] L. L. Foldy and S. A. Wouthuysen, Phys. Rev. 78, 29 (1950).

[127] R. L. Stratonovich, Dok. Akad. Nauk. SSSR 1, 72 (1976); R. L. Stratonovich, Sov. Phys.–

Dokl. 1, 414 (1976) (Engl. Transl.).

[128] C. Kouveliotou, S. Dieters, T. Strohmayer, J. van Paradijs, G. J. Fishman, C. A. Meegan, K.

Hurley, J. Kommers, I. Smith, D. Frail, T. Murakami, Nature 393, 235 (1998).

[129] D. M. Palmer, S. Barthelmy, N. Gehrels, Nature 434, 1107 (2005).

[130] A. K. Harding and D. Lai, Rep. Prog. Phys. 69, 2631 (2006).

Potrebbero piacerti anche