You are on page 1of 42

Rev. Mod. Plasma Phys.

(2017) 1:2
https://doi.org/10.1007/s41614-017-0005-2

CHANDRASEKHAR LECTURE

Nonlinear laser–plasma interactions

P. K. Kaw1

Received: 10 November 2016 / Accepted: 22 May 2017 / Published online: 16 June 2017
Ó Division of Plasma Physics, Association of Asia Pacific Physical Societies 2017

Abstract Soon after lasers were invented, there was tremendous curiosity on the
nonlinear phenomena which would result in their interaction with a fully ionized
plasma. Apart from the basic interest, it was realized that it could be used for the
achievement of nuclear fusion in the laboratory. This led us to a paper on the
propagation of a laser beam into an inhomogeneous fusion plasma, where it was first
demonstrated that light would go up to the critical layer (where the frequency
matches the plasma frequency) and get reflected from there with a reflection
coefficient of order unity. The reflection coefficient was determined by collisional
effects. Since the wave was expected to slow down to near zero group speed at the
reflection point, the dominant collision frequency determining the reflection coef-
ficient was the collision frequency at the reflection point. It turned out that the
absorption of light was rather small for fusion temperatures. This placed a premium
on investigation of nonlinear phenomena which might contribute to the absorption
and penetration of the light into high-density plasma. An early investigation showed
that electron jitter with respect to ions would be responsible for the excitation of
decay instabilities which convert light waves into electrostatic plasma waves and
ion waves near the critical frequency. These electrostatic waves would then get
absorbed into the plasma even in the collisionless case and lead to plasma heating
which is nonlinear. Detailed estimates of this heating were made. Similar nonlinear
processes which could lead to stimulated scattering of light in the underdense region
ðx [ xp Þ were investigated together with a number of other workers. All these
nonlinear processes need a critical threshold power for excitation. Another impor-
tant process which was discovered around the same time had to do with filamen-
tation and trapping of light when certain thresholds were exceeded. All of this work
has been extensively verified in laser plasma experiments and have become the

& P. K. Kaw
predh@yahoo.com
1
Institute for Plasma Research, Gandhinagar, India

123
2 Page 2 of 42 Rev. Mod. Plasma Phys. (2017) 1:2

backbone of our present understanding of how lasers nonlinearly interact with fully
ionized fusion plasmas. A review of this early work will be presented. We shall also
present a review of our involvement in the recent work on nonlinearpenetration of
light into overdense plasmas with gradual and sharp interfaces.

Keywords Laser plasma interaction  Nonlinear phenomena  Laser


fusion  Parametric instabilities  Langmuir collapse  Relativistic nonlinearity

1 Introduction and early history

1.1 Introduction

The investigation of laser plasma interaction problem is motivated by many


applications. Some of these are: fusion of a DT pellet by irradiation with intense
laser pulses, fast ignition approach to laser fusion, propagation in overdense plasmas
by relativistic effects, particle acceleration by wakefields generated by short laser
pulses in a plasma, ion acceleration by table top lasers, intense radiation sources
generated by light plasma interaction, etc. Furthermore, research on this topic has
led to important developments in nonlinear theory of strong plasma turbulence,
theory of parametric instabilities and other areas of fundamental plasma physics.
Although the physics of what we are going to discuss is applicable to most of the
above topics, we shall motivate our discussions primarily with the laser fusion
problem. There are two major directions in which laser fusion research has
progressed. First is the original concept of using nanosecond lasers to compress DT
pellets till they reach 103–104 times original density with a series of well timed
compression shocks and then have a hot spot triggered by the convergence of these
shocks simultaneously in a given region to form a giant heating shock. This concept
has resulted in decent compressions (103) using both direct and indirect drive
schemes, but has not been able to trigger an ignited hot spot due to mixing of hot
and cold regions by Rayleigh Taylor instabilities (Lindl et al. 2014). Considerable
experimental work is still going on using this technique of laser fusion. The
alternate method being pursued is the fast ignition scheme (Tabak et al. 1994) in
which compression is done by a long-pulse nanosecond-laser and the task of
subsequent heating and triggering the ignition hot spot is given to a separate short-
pulse intense laser (tens of kilojoules in *1 ps), such as those resulting from the
chirped pulse amplification technique (Mourou 1997). Thus most of the physics
topics we discuss will be motivated by these two topics of laser fusion. The
discussion on the first topic is presented in the form of a personal quasi-historical
account of some of the early work on laser–plasma interactions, in which the author
actively participated, which illustrates how basic inroads into several areas of
fundamental plasma physics were made in a period measuring less than a decade
(late 1960s/early 1970s). This is followed by a section on relativistic nonlinear
penetration of light in overdense plasma and an implications section, which presents
a very brief account of the application of these results to modern experiments in

123
Rev. Mod. Plasma Phys. (2017) 1:2 Page 3 of 42 2

conventional laser fusion and to conceptual advances in theories of strong


turbulence of electron plasma waves. We next briefly summarize some recent
developments in the second topic, viz. physics of laser–plasma interactions in fast
ignition concept of laser fusion, again highlighting the work in which the author has
participated. Finally, we conclude by looking at the future directions in which
fundamental studies in laser plasma interaction are inevitably moving.

1.2 Early history of conceptual advances in laser–plasma interactions

So, let me start with some personal reminiscences about the conceptual advances in
laser plasma interaction physics related to the importance of parametric instabili-
ties,which took place in late 1960s/early 1970s. This will give the reader an idea of
where my work stands vis-à-vis that of others in this active field and also gives me an
opportunity of expressing my gratitude to some of those people who have helped me
in the various contributions I have made. My interest in the laser plasma interaction
problem started with my arrival as a postdoctoral associate in John Dawson’s basic
plasma theory group at Princeton in 1967. I had completed my PhD one year earlier at
IIT Delhi in India. My PhD dissertation dealt with harmonic generation and mixing by
intense ‘EM’ waves in weakly ionized plasmas such as the ionosphere. John looked at
my background and immediately suggested that I investigate the laser plasma
interaction problem with him. It is well known that when an intense pulse of light
irradiates a solid, the material immediately ablates and gets converted into a plasma
which starts expanding out into vacuum with a speed determined by the laser
intensity. The result is that subsequent interaction of the light wave with matter takes
place when the latter is in the form of an inhomogeneous plasma with a density
gradient. The scale length of the plasma is determined by the length and intensity of
the prepulse which arrives before the main light pulse (contrast ratio for the laser) and
varies between *fractions of a micron for high-contrast picosecond lasers and tens of
microns for ordinary nanosecond lasers. Maximum laser intensities have been
obtained with glass lasers with a wavelength of order 1 micron. Thus the scale length
of density variation varies between fractions of a vacuum wavelength and tens of
vacuum wavelengths. In those early days intense light pulses were nanoseconds in
length. Thus the critical thing was to study interaction of light pulses with gently
inhomogeneous plasmas. John suggested two directions. One was to investigate the
light absorption in an inhomogeneous fully ionized plasma, with a suggestion to
closely examine the collisional interaction process in a fully ionized plasma including
the effect of ion correlations on the real part of conductivity. The other problem was
to investigate the penetration of intense light pulses into overdense plasmas using
relativistic effects. This latter idea was simple. If the light pulses could be made
relativistically intense, the plasma shielding currents would saturate in each period at
 nec, where n is the electron number density and c the velocity of light in vacuum
and would be unable to shield the plasma for most of the period thus permitting
laser penetration into overdense plasmas. The former set of questions led to Dawson
et al. (1969) which described absorption in an inhomogeneous plasma and Kaw and
Dawson (1969), a paper on anomalous heating and parametric instabilities. We also
learnt about the independent and excellent contemporary work on linearized theory

123
2 Page 4 of 42 Rev. Mod. Plasma Phys. (2017) 1:2

of parametric instabilities of relevance to anomalous absorption (Nishikawa 1968a, b;


DuBois and Goldman 1965; Losada 1970; Silin 1965) and those of relevance to
stimulated scattering (Forslund et al. c1973). It became necessary to compare notes
and systemize the parametric instability theory. This was accomplished a couple of
years later (1972–1973), when I was visiting UCLA in the summers and collaborating
with George Schmidt (Kaw et al. 1973) and Drake, Lee and others (Drake et al. 1974)
paper on general theory of parametric instabilities. Meanwhile at Princeton, we had
initiated a study of the nonlinear saturated state of parametric instabilities and its
contribution to anomalous absorption. This was accomplished with the collaboration
of Bill Kruer, with whom we carried out the first PIC simulations of the parametric
instability problem (Kruer et al. 1970) which gave semi-quantitative estimates of the
magnitude of anomalous absorption due to parametric instabilities. This gave us
confidence to apply estimates based on phenomenological arguments to explain some
early experiments of Gekker and Sivukhin (Kaw et al. 1970; Gekker and Sivukhin
1969) and to explain the early results from Arecibo plasma heating experiment in the
ionosphere (Perkins and Kaw 1971). Another absorption topic given some attention in
Dawson et al. (1969) paper was the resonance absorption of obliquely incident light
waves in overdense plasmas by classical collisional/collisionless absorption mech-
anisms. This topic was subsequently discussed in detail in the Lindl and Kaw (1971)
paper published in 1971, which was directed towards calculation of ponderomotive
forces for obliquely incident waves, but used the existing theory of Denisov Ginzburg
(1970) and Denisov (1957) on resonance absorption for making the estimates.
Eventually, the quantitative application of Denisov’s collisionless absorption
mechanism by mode conversion of light waves into plasma waves at the resonance
layer was made in the 1972 paper by Freidberg et al. (1972). The light penetration
problem led to another (Kaw and Dawson 1970) paper which showed one needs
intensity  1018 W/cm2 to see relativistic effects with 1-lm lasers. This appeared to
be science fiction at that time. Little did we realize at that time that within a couple
of decades laser technology would have developed to such an extent that sub-
picosecond lasers with more powerful intensities would be available for experi-
ments. Based on our work, the light penetration physics in overdense plasmas was
explored further for astrophysical applications by Max and Perkins (1971) and for
relativistic self-focusing by Max (1976). Thus the two questions that John Dawson
posed me with on that fateful day in 1967 led to novel and interesting answers.
These were: yes, the collisional interaction-related damping for inhomogeneous
plasma layers was not very attractive for fusion. But collective interactions which
helped convert light wave energy into electrostatic plasma wave modes were
attractive candidates for anomalous absorption and absorption of light was unlikely
to be an issue. Secondly, light could indeed propagate into overdense regions by
relativistic effects. However, recent experiments have shown that for sharp
overdense plasmas (with an interface with a scale length \  light vacuum
wavelength) which naturally result when the pulses are so short that CPA (Mourou
1997) leads to relativistic intensities, standard resonance absorption was replaced by
Brunel mechanism which converted light energy into relativistic electron beams
with energy of the order of few MeV. The subsequent propagation of relativistic

123
Rev. Mod. Plasma Phys. (2017) 1:2 Page 5 of 42 2

beams into overdense plasmas has led to some very novel physics and resulted in a
very fruitful collaboration between our group at IPR (led by Amita Das) and
experimental groups at TIFR (GRK) and Osaka (KaT) (Sandhu et al. 2002; Kahaly
et al. 2008; Mondal et al. 2012; Das and Kaw 2001; Jain et al. 2003; Das et al.
2003; Sandhu et al. 2005; Saxena et al. 2006, 2007; Jain et al. 2007, 2012; Kahaly
et al. 2009; Yadav et al. 2008, 2009; Sundar and Das 2010; Yadav and Das 2010;
Gaur et al. 2009; Sundar and Das 2010; Sundar et al. 2011). I shall also briefly
summarize this work, in which I have actively participated, during the past decade
or more. Finally, I shall throughout emphasize new directions in fundamental
plasma physics, which have resulted from the laser plasma work.

2 Background physics concepts

2.1 Linear theory of light propagation and light absorption

When the light intensity is small, the interaction of light with the plasma can be
understood from linear theory. Linearly, the plasma responds to the ‘EM’ fields by
locally generating conduction currents of magnitude J ¼ n0 ev1 , where
dv1 eE
¼  eixt ð1Þ
dt m
giving J ¼ ðn0 e2 =mðixÞÞE. These 90° out-of-phase currents will attempt to shield
the plasma from the electromagnetic fields. This is clear from the Maxwell equation
4pJ 1 oE
rB¼ þ : ð2Þ
c c ot
The conduction current competes with the displacement current, the latter trying to
propel the ‘EM’ field forward as in vacuum. The result is that plasma behaves as a
dielectric with a dielectric constant given by

c2 k 2 x2p
¼  ¼ 1  ; ð3Þ
x2 x2
with x2p = 4pn0 e2 /m and the 1 coming from the displacement current and the second
term resulting from the plasma conduction current. This expression shows that as
long as x [ xp , the plasma allows the light to propagate and when x\xp , the light
is unable to propagate due to plasma shielding effects. Thus in an inhomogeneous
plasma, the light will propagate up to the critical layer where x = xp and will be
reflected from there.
There was no discussion of light absorption in the above discussion. In fact the
reflection coefficient in the above approximation can be shown to be unity with no
absorption of the light wave. This is because in the limit considered by us the
current is 90° out of phase with the electric field and so the work done by the electric
field on the plasma particles, given by time average of JE, is zero. The whole thing
can be traced to the purely inductive response of electrons to the electromagnetic

123
2 Page 6 of 42 Rev. Mod. Plasma Phys. (2017) 1:2

field, whereby the electron motion lags behind the electric field by 90° because of
electron inertia. The true response also has an in-phase component which is related
to instantaneous response of electrons associated with collisional effects.
The model equation for electrons in the presence of collisional effects may be
written as
ov1 eE
þ mv1 ¼  eixt ; ð4Þ
ot m
where the m term represents the friction encountered by electrons due to collisional
interaction with ions which randomizes their directed momentum and leads to a
heating of the electron fluid. The energy found in the heated electron fluid is lost
from the electromagnetic field energy which thereby suffers absorption, making the
reflection coefficient in the overdense inhomogeneous plasma less than unity. This may
be simply seen in terms of the effective dielectric constant which now becomes complex:

c2 k 2 x2p
¼  ¼ 1  : ð5Þ
x2 xðx þ imÞ
The imaginary part of  is responsible for the damping of the light wave.
In Dawson et al. (1969), we first investigated the reflection coefficient of a light
wave propagating along the density gradient in an inhomogeneous-laser produced
plasma, both by WKB methods and by solving the wave equation with linear and
parabolic density profiles. They found a very good agreement between the WKB
and exact results and found a reflection coefficient for linear profiles
  
32 L
R  exp  ; ð6Þ
15 cs0
where L is the density scale length near the reflection point and s0 is the inverse of
the collision frequency at the reflection point. The reason that s0 plays such a crucial
role in the absorption is that the group velocity of the light wave slows down to
c(1  x2p =x2 Þ1=2 tending to zero at the critical layer and the wave spends a lot of
time at the critical layer. Thus if L  cs0 the light wave is fully absorbed and if
L/cs0 \1, the light wave is partially reflected. If we calculate s0 corresponding to a
critical plasma density for 1 lm, nc  1021 /cm2, temperature *10 KeV the desired
temperature for initiating fusion reactions and L few hundred microns and use the
Rutherford formula for calculating the electron–ion collision time s0 *1/mRuther ford ,
we find L=cs0 \1. This indicates that the laser pulse will be unable to heat a few
hundred micron pellet to fusion temperatures using the conventional absorption
physics involving e–i collisions. Basically, the laser pulse will just heat the plasma
in the neighborhood of the critical layer to high temperatures and will subsequently
reflect most of the energy back.
After the Dawson et al. (1969) paper it was clear that a laser pulse incident on a
solid surface such as a DT pellet would quickly raise the temperature of plasma near
the critical surface to fusion temperatures and would more or less give a reflection
of order unity subsequently. Thus the bulk of the pellet would remain unheated to

123
Rev. Mod. Plasma Phys. (2017) 1:2 Page 7 of 42 2

fusion temperatures. The main culprit was the collisional nature of absorption. The
electron–ion collision frequency, which was responsible for the absorption, was a
strong function of electron temperature  T 3=2 due to the dependence of
Rutherford cross-section on energy. Thus it was clear that in a hot collisionless
plasma, the absorption process would have to depend on some collisionless
absorption mechanism. The only thing known was Landau damping of longitudinal
plasma waves. Since light waves do not directly Landau damp on the particles, their
phase velocity being faster than the velocity of light, we must look for mechanisms
which convert light waves into plasma waves which could then heat even a hot
collisionless plasma. The fact that such a thing was naturally happening in a uniform
plasma, when the light frequency was near the plasma frequency, was already
shown by an early work of Dawson and Oberman (1962) on the high-frequency (hf)
resistivity of a collisionless plasma.
Dawson and Oberman had been interested in the frequency dependence of the
high-frequency resistivity of a collisionless plasma and had constructed a semi-
Klimontovich type of a model to investigate this question. They saw clearly that e–i
collisions are only responsible for the loss of momentum from the electron fluid,
which meant that ions would have to be treated as discrete scattering centers. Thus
they set up a semi-Klimontovich model in which electrons were treated as a phase
space fluid using the Vlasov equation and the ions were treated as randomly
distributed discrete scattering centers, viz charged particles with delta function
positions and given Coulomb fields. To the lowest order, electron fluid simply
suffers an oscillating jitter in the light wave field E ¼ E0 cosðxtÞ. Thus we may
write zeroth-order electron distribution as
f0 ¼ feq ðv  ðeE0 =mxÞsinðxtÞÞ ð7Þ
If we go to the oscillating frame of electrons in this picture, electrons have an
equilibrium distribution (such as a Maxwellian) and the ions are discrete scattering
centers jittering about their mean positions.
The ion discreteness can be treated as a first-order effect (in the plasma parameter
expansion 1=nk3D  1) and can be described by a linearized Vlasov formalism. This
shows that the discrete ions with their shielding clouds oscillating at the light
frequency in this frame set up the characteristic linear Vlasov response at zero
frequency and x in the Vlasov electron fluid. It is as if jittering discrete Coulomb
ions are acting as antennae in this frame and radiating waves at frequency x . The
ions in turn experience a force due to the excited fields which is the negative of the
force on the electron fluid, Fx , due to the discreteness of ions. This force is
responsible for the high-frequency resistivity of the plasma and finally one arrives at
the following estimate of the plasma impedance:
 
ZðxÞ ffi 4pix=x2p ½ð1  F s Þ  iF c ; ð8Þ

where Fs and Fc are the terms in Fx =E0 which vary as sinðxtÞ, cosðxtÞ, Fx being
given by

123
2 Page 8 of 42 Rev. Mod. Plasma Phys. (2017) 1:2

  Z kmax   
eE0 2 1 1
Fx ¼ Ze dkðk4 ÞIm eixt 2  ; ð9Þ
mx 3p 0 k þ kD2 k2 ðk; xÞ
where kmax  T=e2 is the classical (or an equivalent quantum) cut-off of the
k integration.
Figure 1 illustrates the dependence of the hf resistivity on the frequency. It shows
the independence of the hf resistivity from frequency at low frequencies x\xp and
a logarithmic fall at high frequencies x [ xp due to ineffectiveness of interactions
beyond an impact parameter  vth =x. It also shows a bump in the resistivity at and
near xp which is related to the excitation of plasma waves due to discreteness of
ions. The bump is rather small for a random distribution of ions (for which the
above calculation is reported), but can be shown to assume large values when the
ions are highly correlated (Dawson and Oberman 1963). This may happen, for
example, by the excitation of ion waves in the plasma. An increase in the hf
resistivity at the critical frequency is very welcome since, as already shown earlier,
this is where the reflecting wave is spending most of its time because of the reduced
group velocity at the point of reflection. Furthermore, once the electromagnetic
wave energy is converted to plasma waves, it is likely to be absorbed by Landau
damping and other such mechanisms even in a hot collisionless plasma. We thus
turned to the general problem of conversion mechanisms of light waves into plasma
waves at the critical frequency layer.

2.2 Collective electrostatic modes in a collisionless plasma

First, we must learn something about the collective modes which a plasma can
support. An unmagnetized plasma supports two kinds of electrostatic plasma waves.
These are the high-frequency plasma oscillations involving only motion of electrons
while ions remain as a background static compensating charged fluid and the lower
frequency ion acoustic waves involving motion in both electron and ion fluids.

Fig. 1 Resistivity vs frequency

123
Rev. Mod. Plasma Phys. (2017) 1:2 Page 9 of 42 2

These can be simply described by fluid equations, since collective motion in the two
fluids dominates and discrete particle effects and/or wave–particle interaction only
lead to higher order effects like damping of the collective modes by collisions and
Landau damping, respectively. Let us start with a description of collective modes of
a collisionless plasma. For electrostatic modes in a homogeneous plasma, the basic
linearized equations are: Continuity equation for species a
ona1 on0 va1
þ ¼0 ð10Þ
ot ox
Equation of motion
ova1 ea 1 opa1
¼ E ð11Þ
ot ma na0 ma ox
Equation of state
pa1 na1
¼ ca ð12Þ
pa0 na0
Poisson equation
oE X
¼ 4pea n1a ð13Þ
ox a

When only electron motion is involved because high-frequency modes are being
considered, we take the ion number density to be a constant equal to the equilibrium
ion density. Linearising the electron equations and eliminating all variables except
the density we get

o2 ne1 2
2
2 o ne1 ð14Þ
þ xp n e1  3v the ¼ 0;
ot2 ox2
where x2p = 4pn0 e2 /m and we have chosen the ratio of specific heats c for the one-
dimensional electron fluid to be 3. This is because for these waves phase speed is
much larger than thermal speed and density compressions are accompanied by
heating, which for 1-D fluids give
p1 n1 T1 n1
¼ þ ¼3 ; ð15Þ
p0 n0 T0 n0
because T  nðc1Þ  n2 for 1-D fluids. Equation (14) shows that even in the absence
of thermal effects, plasma undergoes linear oscillations at the plasma frequency as
disturbances around the equilibrium. Physically, if the space charge balance is
disturbed somewhere, the electrons immediately react by rushing back to cancel the
space charge. But as they have inertia, they overreact by continuing to move even
after they have compensated the space charge. This causes a space charge field with
the opposite sign which pulls the electrons back towards the place of disturbance.
Plasma thus undergoes oscillations at the electron plasma frequency which is
determined by the electron number density (through the strength of space charge

123
2 Page 10 of 42 Rev. Mod. Plasma Phys. (2017) 1:2

field) and the electron mass. When finite temperature effects are taken into account,
the oscillations turn into plasma waves with a finite group velocity
 dx=dk  3vth ðkvth =xÞ, with which the wave energy propagates.
We next consider low-frequency electrostatic collective modes in a collisionless
plasma. Low-frequency modes involve motion of ions in an essential manner. When
the low frequency satisfies the condition x\\kvthe , electrons treat the perturbation
as a static one. This has two consequences. Heat produced by electron compression
cannot raise the electron temperature because the heat is rapidly conducted away by
collisionless electrons. Thus electrons behave like an isothermal fluid with p1 =p0 ¼
n1 =n0 and ratio of specific heats c = 1. Secondly, the inertia term in the equation of
motion of electrons is negligible and the force balance essentially balances the
pressure perturbation with the electric field perturbation. This gives
ne1 =ne0 ¼ e/1 =Te0 , that is electrons behave like an isothermal Boltzmann fluid. If
we linearize the ion continuity and equations of motion, treat the ions as an
adiabatic ion fluid with c = 3, i.e., Ti1 =Ti0 ¼ 2ni1 =ni0 (because ions form a cold
fluid with xkvthi ), we get
o2 ni1 o2 ni1 ð16Þ
2
þ x2pi ni1  3v2thi 2 ¼ x2pi ne1 ;
ot ox
where using Poisson equation (13) and the Boltzmann relation for ne1 we may write

ni1 ne1 k2D o2 ne1


¼  : ð17Þ
n0 n0 ne0 ox2
Taking perturbed quantities of the form expðikx  ixtÞ and combining Eqs. (16) and
(17), we get the dispersion relation for ion acoustic waves:

x2 ¼ k2 Cs 2 =ð1 þ k2 k2D Þ þ 3k2 v2thi ; ð18Þ


pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
where Cs ¼ xpi kD = ðTe0 =Mi Þ is the speed of sound in the plasma fluid with warm
electrons and cold ions. When k2 k2D  1, x2 ¼ k2 ðTe0 þ 3Ti0 Þ=Mi . In this case,
plasma is quasi-neutral with ne1  ni1 and sound wave-like disturbances propagate
in the plasma. Note that even in a cold ion plasma, acoustic disturbances like sound
waves can propagate with the speed Cs . This is because the electric fields in the
collisionless plasma communicate the existence of electron pressure to the ion fluid.
This means that ion density perturbations are surrounded by finite electron pressure
shielding clouds and when they squeeze to form an ion density compression, they
squeeze the electron fluid which has a finite temperature that resists the compression
forming the basis of an acoustic perturbation. Note also that when k2 k2D  1; i.e.,
when the wavelength is much shorter than an electron Debye length, electrons can
no longer do the shielding ne1  ni1 =k2 k2D \ni1 because the structure length resides
within an electron Debye cloud. In this case, ions support an ion plasma oscillation
very complimentary to electron plasma oscillation. In this limit ions move around
producing the space charge and electrons form the stationary compensating charge.
In the above discussion, plasma has been treated as a fluid. In reality it is made up of
discrete particles, which even if we carry out the expansion in powers of 1=ðnk3D Þ,

123
Rev. Mod. Plasma Phys. (2017) 1:2 Page 11 of 42 2

leads at the lowest order to the Vlasov formalism, where the plasma is treated as a
phase space fluid. The fluid approximation we used above, consisting of various
moments of the Vlasov equation and truncation beyond the third moment, can be
justified in a collisionless cold plasma only if x=kvthi  1 and even then misses the
characteristic damping process known as Landau damping (Landau 1946; Dawson
1961). This latter process is a wave–particle interaction process that transfers energy
from the smaller number of particles traveling faster than the wave to the wave and
takes energy from the wave to accelerate somewhat larger number of slower par-
ticles in the monotonic stable distribution, to the wave speed. Once the energy goes
into the particle distribution, it gets phase mixed away and has no chance of finding
its way into the basic wave again. In other words, one has a mechanism of con-
verting coherent wave energy into incoherent resonant particle motions and phase
mixed perturbations or effectively the temperature of these resonant electrons.

3 Anomalous absorption: parametric instability

We now consider the possible excitation of electrostatic collective modes in a


plasma immersed in electric fields associated with the light wave oscillating near the
plasma frequency. Conversion of energy in the light oscillations into electrostatic
waves is like light absorption because once the energy goes into random
electrostatic waves, it cannot readily come out of the plasma and stays trapped in
the plasma until it is absorbed by collisionless wave–particle interaction processes
like Landau damping. Since wavelengths of the ‘EM’ fields are typically much
longer than those of the electrostatic fields, we can consider the limit of a dipole
pump wave, viz. one in which structure in space can be ignored, i.e.,
E ¼ E0 cosðx0 tÞ. For simplicity again we shall consider a 1-D problem in which
wave excitation along the direction of the electric field of the light wave will be
considered.
Imagine a low-frequency ion density perturbation in this plasma. Collective
modes at low frequency and long wavelength tend to be quasi-neutral. Hence the
electrons will generate a similar density perturbation ne1s  ni1s at the low
frequency. Note that the electrons are oscillating in the equilibrium field of the
light wave with a velocity v0 ¼ ðeE0 =mx0 Þsinðx0 tÞ. Nonlinear terms in the
electron equations, such as the convective nonlinearity v0 ðone1s =oxÞ in the
continuity equation will generate electron density fluctuations at high frequencies
because of the oscillation of the electron fluid along the density gradient of the low-
frequency electron density fluctuation ne1s . Thus if we imagine the low-frequency
ion acoustic perturbation has space time dependence expðixt þ ikxÞ the nonlinear
convective term generates density fluctuations at (x
x0 ) and wave vector k. Now
x0 is near the characteristic plasma frequency of the plasma. Thus the sideband
density fluctuations are being driven near the plasma frequency and with a finite k.
This source of density fluctuations thus acts as an antenna/grid trying to excite
plasma waves in the medium. This is described by the sideband generation equation,
which takes the form

123
2 Page 12 of 42 Rev. Mod. Plasma Phys. (2017) 1:2

h i
ðx
x0 Þ2 þx2p þ 3k2 v2the ve1
¼ x2p ðnel =n0 Þv0 : ð19Þ

The sideband density fluctuations will produce electric fields at the high frequencies
which will be intense if x
x0  x0  xp and the driving is of a resonant nature.
The reverse coupling of the newly excited high-frequency plasma wave fields on the
low-frequency ion acoustic dynamics is through the ponderomotive forces
v0 ove1
=ox which appear in the equation of motion of the electron fluid as con-
vective derivative terms. Physically this is the radiation pressure term rjvj2 , which
appears by combining the vrv and vB terms in the electron equation of motion.
Ponderomotive forces might also be included in the equations by the introduction of
a ponderomotive potential wp ¼ jvj2 =2, in addition to the electrostatic potential in
the low-frequency equations. We thus find the reverse coupling equation (in the
limit of quasi-neutrality ne1  ni1 )
 m v 
0
ðx2  x2s Þne1 ¼ k2 x2pi ðve1þ þ ve1 Þ; ð20Þ
4pe2 2
where xs is the characteristic ion acoustic frequency defined by the dispersion
relation in Eq. (18) (where x should be identified with xs ). Combining the two sets
of equations and eliminating all variables, we get the new dispersion relation for the
coupled ion acoustic and plasma wave system
 2 2 " #
2 2 2 k v0 2 1 1
ðx  k Cs Þ ¼ xpi þ : ð21Þ
4 ðx  x0 Þ2  x2R ðx þ x0 Þ2  x2R

Here, x2R ¼ x2p þ 3k2 v2the and defines the resonant frequency for electron plasma
waves. This dispersion relation may also be put in a succinct and general form as
" #
1 1 k2 v20 1 1
þ ¼ þ ;
ve ðk; xÞ 1 þ vi ðk; xÞ 4 ðx  x0 Þ2 ðx  x0 ; kÞ ðx þ x0 Þ2 ðx  x0 ; kÞ
ð22Þ
where ve ðk; xÞ, vi ðk; xÞ are the electron and ion susceptibility tensors at the low
frequency x; k and (x
x0 ) = 1 ? ve (x
x0 ) represent the dielectric response
function at the sideband frequencies x0 ffixp (which has no ion contribution because
ions do not respond to perturbations at the high frequencies of order x0 ffixp ). This
equation reduces to the previous one if we put

1 x2pi x2R
ve ðk; xÞ ¼ ; vi ðk; xÞ ¼ ; ve ðx
x0 ; kÞ ¼  : ð23Þ
k2 k2D x2 ðx
x0 Þ2
Equation (22) is more general because we may include kinetic effects and/or col-
lisional damping effects by using a generalized form of the susceptibility tensor. For
example, damping effects may be included by writing

123
Rev. Mod. Plasma Phys. (2017) 1:2 Page 13 of 42 2

x2pi
vi ðk; xÞ ¼ ð24Þ
xðx þ imi Þ

x2pe
ve ðk; xÞ ¼ ; ð25Þ
x
ðx
þ imÞ  3k2 v2the
where x
¼ x
x0 and mi and m are the damping rates of ion and electron modes,
respectively, and may include collisional and collisionless (Landau) damping
effects in a phenomenological manner. It also shows that the parametric process
may be considered as a four-wave interaction process where the pump ‘EM’ wave at
ðx0 ; 0Þ couples the low-frequency ion acoustic type perturbation ðx; kÞ with elec-
trostatic high-frequency sideband perturbations ðx
x0 Þ near the electron plasma
frequency. Since the shift in the frequency of the low-frequency ion acoustic type
perturbation and the growth rate can both be comparable to the low frequency itself,
it is necessary to include both upper sidebands in the analysis. This is unlike weak
turbulence theories, where the frequencies of the normal uncoupled modes are only
perturbatively modified by the coupling. Parametric instability theory is thus a step
beyond weak turbulence theories and includes a shift of the low frequency and a
growth rate comparable to the low frequency itself.
Equation (22) can be utilized to show that the coupled system exhibits
instabilities along two branches. If the natural damping of the collective modes due
to collisional effects and/or collisionless Landau damping mechanism is incorpo-
rated, the excitation of instabilities requires that the laser field exceeds a critical
value called the threshold field. We can investigate the two branches by
approximating the dispersion relation for x0  xp as:

½xðx þ imi Þ  k2 Cs2 ½ðx þ im=2Þ2  d2 þ Kd ¼ 0; ð26Þ

where

k2 v20 x0  m 
d ¼ x0  xR ; and K ¼ : ð27Þ
4 M
This equation is identical to the one derived by Nishikawa for coupled waves
(Nishikawa 1968b). The equation is known to possess two branches of instabilities
namely, the purely growing branch and the decay instabilities branch. We first
investigate the second branch, namely the so-called decay instability, which requires
that d be positive and has the minimum threshold, when certain frequency and wave
number matching conditions are satisfied:

x0 ffixR þ kCs ; 0 ¼ k þ ðkÞ ð28Þ


The dispersion relation near threshold takes the form (note d  kCs )

123
2 Page 14 of 42 Rev. Mod. Plasma Phys. (2017) 1:2

Kd
ðx  kCs þ imi =2Þðx  d þ im=2Þ þ ¼0 ð29Þ
4k2 Cs2
and the minimum threshold power, in terms of the amplitude of the laser field, is
given by
 2   
v0 m mi
¼4 : ð30Þ
vthe x0 kCs
with v0 ¼ eE0 =mx0 and depends on the damping rates of ion acoustic and plasma
waves which are excited.
This instability was first identified in DuBois and Goldman (1965) using a
different treatment. Similar work was reported in Silin (1965). In this limit, it is like
a standard three-wave interaction process consistent with pump wave interacting
with lower sideband and the low-frequency mode only. However, if the pump
amplitude is large, the growth rate can become comparable to real part of the low-
frequency mode and we must retain coupling to both upper sidebands and we get a
four-wave process with the growth rates (Nishikawa 1968a, b):
pffiffiffi !1=3
3 x2pi k2 v20
c¼ ð31Þ
2 x0 16

The second branch requires x0 \xR and leads to the result that when a certain high
threshold determined only by the damping rate of the high-frequency electron
plasma waves is exceeded, the low-frequency mode gets very strongly modified
leading to a purely growing mode. From Eq. (26), with damping rates m; mi = 0, and
putting x ¼ ic, we get the dispersion relation

ðc2 þ k2 Cs2 Þðc2 þ d2 Þ þ Kd ¼ 0 ð32Þ

with the roots


qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
c2 ¼ ðk2 Cs2 þ d2 Þ=2 þ ðk2 Cs2  d2 Þ2 =4  Kd: ð33Þ

c is real when d\0 and K exceeds a critical value. The threshold and the growth
rates are given by (threshold may be obtained by putting x ¼ 0 in Eq. (26) and
minimizing the expression with respect to d for choosing the most unstable wave
vector)
 2  
v0 m
¼4 ð34Þ
vthe x0
and the maximum growth rate well above threshold is given by (Nishikawa
1968a, b):
 1=3
m k2 v20
c¼ x0 : ð35Þ
M 8

123
Rev. Mod. Plasma Phys. (2017) 1:2 Page 15 of 42 2

Effectively, the physics of coupling is such that the frequency of the low-frequency
modes decreases, is dragged to zero, and then moves along the purely imaginary
axis in the complex plane. This is definitely a four-wave interaction process and
even at threshold requires both the sidebands at the high-frequency end. It is
therefore outside the purview of weak turbulence theory. Physically again, this
mode arises because the pressure due to ponderomotive forces is negative and
opposes the restoring force due to thermo-kinetic pressure in the ion acoustic wave.
When the threshold is exceeded, the total effective pressure for the wave becomes
negative and accumulation of density at a given location decreases the local pres-
sure, inviting more density in, thus giving a purely growing mode. This physical
interpretation for the zero-frequency purely growing mode was advanced by Kaw
and Dawson (1969), who also emphasized its importance for anomalous absorption
of laser light. This purely growing instability was first identified in a fluid treatment
in Kaw and Dawson (1969) and Nishikawa (1968a). Kaw and Dawson (1969) were
also the first to interpret this instability as the analogue of the well-known two-
stream instability for a case when the electron stream is oscillating and called it as
the OTSI. Nishikawa (1968a) also presented as a general treatment of parametric
instabilities of coupled oscillators with very different characteristic frequencies for
the first time. Kinetic theories of the instabilities were presented in Losada (1970)
and Silin (1965).
When the decay and oscillating two-stream instabilities were discovered, the next
obvious question which arose was ‘‘how to estimate the effect of these instabilities
on the absorption of the light wave’’. It is obvious that excitation of instability is a
linear effect and does not tell us how much energy is extracted from the light waves.
We need a nonlinear theory of saturation of these instabilities to figure out the rate
of energy absorption from the light wave. In order to estimate the absorption
coefficient for a light wave which excites one or both of the parametric instabilities,
it is necessary to go beyond linear theories of parametric instabilities. One must go
to a nonlinear theory of these instabilities which gives a saturation of the fluctuation
levels and thence a method of calculating the enhancement of the high-frequency
resistivity because of the enhanced ion correlations present in the saturated state.
These nonlinear theories are quite complicated. Therefore recourse was taken at this
time to carry out particle simulations of these instabilities. This work was initiated
by a collaboration with Bill Kruer at Princeton, who carried out PIC simulations of a
homogeneous plasma in the presence of a homogeneous (dipole approximation)
oscillating electric field with a frequency close to the plasma frequency. He found
that as expected parametric instabilities were excited when the threshold fields were
exceeded and eventually the coupled plasma oscillation–ion acoustic oscillation
system came to a saturation in which a turbulent state of many wave vectors was
excited. In the turbulent saturated stage, the secondary and tertiary wave modes kept
getting energy from the oscillating electric field and this kept going into the particles
by a collisionless damping mechanism. Since most collective modes are several
Debye lengths long, the excited waves have phase velocities larger than thermal.
Thus we expect to see a particle distribution function with extended tails. This is
indeed what is observed as shown in Figs. 2 and 3.

123
2 Page 16 of 42 Rev. Mod. Plasma Phys. (2017) 1:2

Fig. 2 Growth and saturation of parametrically excited electrostatic waves in the computer simulation

The saturated state is essentially obtained where the growth rate of a wave is
compensated by an enhanced damping due to coupling with many tertiary modes
which are damped by collisionless damping mechanism. Thus we may introduce a
phenomenological damping of the excited modes through a parameter meff and say
the mode saturates when the growth rate c0  meff =2. Another way of stating this for
the OTSI is that instability saturates when the applied oscillating electric field
matches the modified threshold with an increased effective damping coefficient.
This phenomenological argument gave good agreement with Kruer et al. (1970)
simulations and was soon adopted to explain certain experimental observations on
microwave–plasma interactions which were already in the literature. This defines a
large collision frequency which may then be used with the propagation physics of
the pump ‘EM’ wave in an inhomogeneous plasma to define a new reflection
coefficient
   
32 L
R  exp  meff : ð36Þ
15 c
The wave is thus fully absorbed if meff ðL=cÞ  1. This is satisfied relatively easily
and goes on to show that anomalous absorption processes can readily ensure that the
light energy is absorbed by the fusion pellet with the absorbed energy ending up in
electrostatic waves and thence into plasma particles by collisionless damping
mechanisms like Landau damping.
We now follow the phenomenological approach adopted in Kaw et al. (1970) to
explain the microwave experiments of Gekker and Sivukhin. In this experiment, a
microwave was incident on an overdense plasma and the reflection coefficient was
measured as a function of the incident microwave intensity. Gekker and Sivukhin
(1969) found that at low intensities, one observed perfect reflection. Beyond a
certain critical intensity the reflection coefficient started dropping and then fell
rather rapidly to low values. There was no indication of significant microwave
transmission (as the plasma was overdense) or scattering. Thus it seemed clear that
microwave was getting absorbed beyond a critical intensity. Kaw et al. (1970) first

123
Rev. Mod. Plasma Phys. (2017) 1:2 Page 17 of 42 2

Fig. 3 The distribution function


of x electron velocities
eE0 =mxpe vTe ¼ 0:5, x0 ¼
1:04xpe and xpe t ¼ 672;
showing extended tails on the
distribution

used a wave equation in a linear density profile to estimate the Airy amplification of
the wave field in the plasma. For low incident powers it was shown that the plasma
is stable and so the only absorption included was that due to collisional effects. This
was found to be small and led to a reflection coefficient of order unity. Next it was
shown that beyond a critical amplitude of the incident microwave power, the plasma
was unstable to decay instability and the OTSI. Phenomenologically it was assumed
that when the instability saturates, the effective collision frequency felt by the
electrons is enhanced so much that the applied wave electric field just matches the
new threshold, i.e.,
x  v 2
0 0
meff ffi : ð37Þ
4 vthe
This allows us to solve for the reflection coefficient as a function of incident power.
When this was done, the reflection coefficient curve got a reasonable interpretation
(Figs. 4, 5).
At this time, another interesting development took place. Radio astronomers at
Arecibo, in the mid-latitude region had started an interesting ionospheric heating
experiment. The idea was to heat the ionosphere with a radio wave near the critical
frequency of F2 layer  5 MHz. Since the ionosphere in this region is dominated by
electron–ion collisions, it was expected that the ionosphere would get heated up,

123
2 Page 18 of 42 Rev. Mod. Plasma Phys. (2017) 1:2

Fig. 4 Schematic of the density


profile of leading plasma edge
on which intense microwaves
are incident (Gekker and
Sivukhin 1969)

Fig. 5 Variation of the


reflection coefficient with the
power of incident electric field
strength

become less collisional and therefore lead to less damping of a probe wave which
was sent for reflection from that region. Experiments showed otherwise (Perkins and
Kaw 1971). It was found that when the incident radio wave power exceeded a
threshold, the probe signal reflection coefficient essentially went to zero. It was
shown by Perkins and Kaw (1971) that like the microwave experiment of Gekker
and Sivukhin (1969), the F2 ionosphere was also becoming unstable to decay and
oscillating two-stream instabilities beyond a critical power. This had a deleterious
effect on the probe wave leading to its anomalous absorption and scattering by
fluctuations. This experiment and its interpretation was yet another triumph for the
anomalous absorption theory which was rapidly gaining acceptance. Thus most
people agreed that the absorption of light in the fusion plasma layer is not a matter
of concern. Parametric decay and oscillating two-stream instabilities were likely to
be excited and would lead to an anomalously large absorption of the light wave by
its conversion into collective electrostatic modes which would then get absorbed by
collisionless mechanisms like Landau damping.

123
Rev. Mod. Plasma Phys. (2017) 1:2 Page 19 of 42 2

4 Filamentation and trapping of light

The phenomenon of filamentation and trapping of light due to ponderomotive forces


in a plasma was first reported by Kaw, Schmidt and Wilcox in 1973 (Kaw et al.
1973). This collaboration occurred in the summer of 1973 at UCLA when I was
visiting from my base position in India. It was an exciting period in the development
of theory of parametric instabilities. Drake was working for his PhD degree and was
helping develop key theoretical concepts on how to incorporate ponderomotive
forces in the Vlasov treatment with Y.C. Lee and the rest of us (this led to the
famous six authors’ paper on parametric instabilities due to ‘EM’ waves in a
plasma). All of us were also involved in analyzing the dispersion relation in various
limits and recovering known instabilities from it and see the various connections. It
is in these fertile conditions that George, Tom and I started analyzing the problem of
filamentation and trapping. It was clear that such an effect must arise. For consider a
plasma with light and plasma uniformly distributed. Now consider a wave front with
a fluctuation in the intensity such that the intensity is more at the center than at the
edges. It is clear that ponderomotive forces will try to push matter out radially and
this will make the density in the center to be less than at the edges. Thus the
dielectric constant in the center increases, viz.  ¼ 1  ðx2p =x2 Þ increases. This has
the effect of slowing down the wave front in the center than at the edges. Another
way of stating this is that the wave front acquires a curvature which focuses the laser
and increases the intensity further. There is thus a positive feed back and the
channels should keep getting evacuated of the plasma and filled with intense light
till channels of order skin depth with very limited plasma in them will form in the
midst of the original plasma. We shall show that the filamentation occurs when a
threshold is exceeded. It thus appears that the uniformly distributed state of light and
plasma above a threshold power is unstable to the precipitation of light where the
overlap region between high-density plasma and intense light is reduced just as in a
chemical solution, the solute separates from the solvent and precipitates out.
Thermodynamically, the overlap region has electrons moving with a jitter velocity
with respect to ions, which is clearly like a free energy source which is being
minimized subject to the relevant constraints. A variational problem which
demonstrates this physics remains to be formulated as yet.
The filamentation problem can be carried out as a spatial instability problem with
electromagnetic fields having the expðixtÞ dependence. The simplest version is
one where E due to ‘EM’ field is in X-direction, propagation is in Z-direction and
the light intensity variation is in Y-direction. The plasma can be treated as a
dielectric medium with the cold plasma dielectric constant  ¼ 1  ðx2p =x2 Þ. The
plasma density can be rearranged by the structure of the electromagnetic fields. This
is because electromagnetic fields apply forces vrv and vB which may be
combined to write a ponderomotive force term rjvj2 /2, where v ¼ eE=mx is the
jitter velocity of electrons with respect to ions. This force is from strong field to
weak field direction in a intensity modulated signal. The force appears only on the
electron fluid, but cannot drive electrons away because space charge fields are
produced which hold the electrons in place. The result is that electrons have a

123
2 Page 20 of 42 Rev. Mod. Plasma Phys. (2017) 1:2

Boltzmann-like distribution of density in which both electrostatic and ponderomo-


tive potentials play a role. Physically, this is a balance between pressure gradient
term with isothermal electrons and the electrostatic and ponderomotive forces, that
is
ne ¼ n0 expðeð/ þ /p Þ=Te Þ ð38Þ

Fp ¼ er/p ¼ re2 jEj2 =2mx2 : ð39Þ

Ions do not see the ponderomotive potential and only feel the electrostatic fields and
thus give

ni ¼ n0 expðe/=Ti Þ: ð40Þ
Invoking quasi-neutrality of the zero frequency density perturbations, we find an
expression for / in terms of the ponderomotive potential, which when substituted in
ne give the following expression:

ne =n0 ¼ exp½e2 jEj2 =2mx2 ðTe þ Ti Þ : ð41Þ

This leads to an expression for the nonlinear dielectric constant as


 
 ¼  jEj2 ¼ 1  ðx2p0 =x2 ÞexpðbjEj2 Þ: ð42Þ

Noting that electrostatic component of the field satisfies the equation: rðEÞ = 0,
the nonlinear wave equation now takes the form

r2 E þ rðEr=Þ ¼ ðx2 =c2 ÞE: ð43Þ

We now consider an EM plane wave E ¼ E0 cosðxt  k0 zÞ with a dispersion rela-


1=2
tion x=k0 ¼ c=0 , where 0 ¼ ðE20 =2Þ . Introducing a perturbation dE ¼
e1 ðxÞcosðxt  k0 zÞ þ e2 ðxÞsinðxt  k0 zÞ; we may write linearized coupled equa-
tions as

oe2 0 x2
2k0 þ r2 e1 þ rðE0 rE0 :e1 Þ ¼ 2 0 E0 E0 :e1 ; ð44Þ
oz 0 c

oe1 0
2k0 þ r2 e2 þ k0 ez ðE0 rE0 :e1 Þ ¼ 0: ð45Þ
oz 0
These equations may be used to derive dispersion relation for modes growing in the
Z-direction as  expðkjj z þ iðkx x þ ky yÞÞ, namely
h i
ðkx2 þ ky2 Þ ðkx2 þ ky2 Þ þ E02 0 ½ðkx2 =0 Þ  ðx2 =c2 Þ þ 4k02 kjj2 ¼ 0: ð46Þ

This gives the condition for instability as

123
Rev. Mod. Plasma Phys. (2017) 1:2 Page 21 of 42 2

x2 k2 kx2 þ ky2
[ xþ : ð47Þ
c2 0 E02 0

For kx ¼ 0 , the fastest growth rate has the value kjjmax ¼ k40 E02 0 =0 and for ky ¼ 0,
this growth rate is
k0 2 0

1=2
kjjmax ¼ E0  =0 1 þ E02 0 =0 : ð48Þ
4
Thus one expects the light wave to breakup first into slabs in the direction along
wave electric field vector and almost simultaneously into filaments (one can see this
effect clearly in the first principal simulations of Nishihara et. al.[??] specially near
the critical plasma frequency). Substituting for 0 shows that the typical scale length
of the fastest growing mode is of the order some what longer than the light
wavelength. We can also show that a beam cross-sectional area L2 will self-focus
provided the power in the beam exceeds a critical power given by
c
Pc ¼ ðTe þ Ti Þ½1 þ x2 =x2p ð1  x2p =x2 Þ1=2 ; ð49Þ
4pr0
where r0 is the classical electron radius.
What will be the ultimate fate of the filamentation instabilities? Presumably this
system wants to go towards a nonlinear equilibrium in which the time dependence is
absent and the dimensions are fixed by the magnitude of nonlinearity. One such
solution for the nonlinear equilibrium (for the case when the electric filed is pointing
in the X-direction and all quantities are varying Y-direction only) may be obtained
by solving the nonlinear equation:
 
o2 E x2 2 k 2 c2
þ 1  aexpð0:5bE Þ  E ¼ 0; ð50Þ
oy2 c2 x2
in which a ¼ x2p0 =x2 . The solution may be found by the method of quadratures and
which produces the first integral of the form

ðg0 Þ2  g2 þ k½g2 þ expðg2 Þ  1 ¼ 0; ð51Þ

where g2 ¼ 0:5bE2 , k ¼ x2p0 =ðx2p0 þ k2 c2  x2 Þ and the derivate is with respect to


pffiffiffi
n ¼ xp0 y=c k. A slab solution corresponds to the case when g, g0 both vanish as
n ! 0. The maximum value of the g0 determines the wave number k by the eigen
value condition k ¼ g20 =½g20 þ expðg20 Þ  1 . For cubic nonlinearity with g0 1 the
scale length of EM field is given by the inverse of ðxp0 =cÞE0 and is relatively long.
For g0 1 on the other hand, scale length is of the order of c=xp0 , the skin depth.
This result is reasonable because the channel is completely evacuated and the light
can propagate into the overdense regions on the sides only up to a skin depth. The
paper by Kaw et al. (1973) includes cavity resonator-type solutions in addition to
the wave guide solutions described above.

123
2 Page 22 of 42 Rev. Mod. Plasma Phys. (2017) 1:2

5 Stimulated scattering: general theory of parametric instabilities

Thus far we have predominantly discussed only the conversion of incident EM


fields into collective electrostatic modes. Such processes are responsible for
anomalous absorption of the light pulse with a frequency close to the plasma
frequency. But once the genie of three-wave interactions is out we must consider
other nonlinear processes which may also be excited. In fact, filamentation is indeed
such a process with a light wave going to a transverse electrostatic perturbation at
rest and primarily forward scattered waves in the interior. Furthermore, news came
to Princeton that while we were busy with parametric excitation of electrostatic
waves, it was recognized by the Los Alamos group (Forslund et al. 1973) that
equally important processes are conversion of incident light wave into a plasma
wave and a scattered light wave at the shifted sideband frequency (viz. SRS or so-
called stimulated Raman scattering) or a three-wave process converting incident
light wave into an ion acoustic wave and a scattered wave at shifted sideband
frequency (SBS or stimulated Brillouin scattering). Furthermore, since the low-
frequency mode dispersion relation is strongly modified by the pump wave in
parametric instability, it need not even be a low-frequency normal mode of the
plasma. Thus if x0  x1  kvthe ; kvthi , where x1 refers to the scattered electromag-
netic wave and k = k0 - k1 is the wave number of the scattering electrostatic field,
which may be a perturbation suffering intense Landau damping in the absence of the
pump wave but may be driven unstable by an intense pump, the process is labeled as
stimulated Compton scattering or SCS. These processes are deleterious and convert
ingoing light wave into a scattered wave which may be backward directed or
directed sideways and thus play a harmful role in the light plasma interaction for
fusion. They are called stimulated scattering processes and are important in the
parameter space where xp  x0 up to xp  x0 (SBS, SCS) and xp  x0 to
2xp  x0 (SRS) that is, in most of the underdense region which is encountered by
the light wave before it reaches the critical frequency. Thus if most of the light gets
scattered before it reaches the critical layer, the good processes never come into play
and the problem becomes very complex. In addition to these essentially three-wave
processes, there are four-wave processes like modulational instabilities and
filamentation which essentially favor the process of forward scattering of light
which are useful for transporting light to the critical frequency point.
A universal treatment for all parametric instabilities driven by an electromagnetic
wave in a plasma was presented by Drake et al. (1974) in 1974 and is still
considered a gospel paper on the subject. It introduced the general concept of a
ponderomotive potential coming from the ponderomotive force related to rjvj2 .
This together with the electrostatic potential described the selfconsistent electro-
static modes and their excitation by electromagnetic perturbations. The pump and
scattered electromagnetic waves were described by the Maxwell equations in which
additional currents due to the movement of electrostatic density fluctuations by the
velocity of the pump wave generated the source sideband currents. It has been
shown in references Forslund et al. (1973) and Drake et al. (1974) that the general

123
Rev. Mod. Plasma Phys. (2017) 1:2 Page 23 of 42 2

dispersion relation for parametric instabilities in the presence of pump wave E ¼


E0 cosðk0 x  x0 tÞ takes the form
!
1 1 k2 jk v0 j2 jk v0 j2 jkþ v0 j2 jkþ v0 j2
þ ¼ 2D
 2 2 þ 2  2 2 ;
ve ðk; xÞ 1 þ vi ðk; xÞ 4 k  x  k  kþ Dþ x þ kþ þ
ð52Þ

k ¼ k0  k1 ; ð53Þ

v0 ¼ eE0 =mx0 ; 2
x
; k
are x
x0 and jk
k0 j2 ; ð54Þ

2 2
D
¼ k
c  x2

¼ k2 c2
2kk0 c2 2xx0  x2 : ð55Þ

This equation is the most general dispersion relation for parametric instabilities
driven by a plane EM wave. When 
[defined below Eq. (22) as a dispersion
function for electrostatic sideband waves] tends to zero, k  v0 terms dominate and
we have coupling to electrostatic waves as already described in Sect. 3. On the other
hand, if D
tends to zero, 
is non-resonant and does not vanish. In this case, EM
waves are excited and we get the new stimulated scattering instabilities. We should
first consider the case when only one of D
is zero, i.e., only one upper sideband is
excited. Thus, we consider a case D tending to zero and Dþ not equal to zero, i.e.,
anti-Stokes component is non-resonant.
For xx0 , we may approximate D as
 
c2 k:k0 c2 k2
D Dðk ; x Þ ¼ c2 k2
 x2 þ x2p ¼ 2x0 x  þ : ð56Þ
x0 2x0
It is clear that D tends to zero if k ¼ 2k0 cosh; where h is the angle between k and
k0 ; i.e., between the excited electrostatic mode and the direction of propagation of
the pump wave.
Figure 6 illustrates this result in a geometrical form and shows that the h ¼ 0
corresponds to backscattering (i.e., the scattered Stokes components ðx ; k Þ
propagates opposite to the incident EM wave) and h ¼ p=2 corresponds to almost
forward scattering or filamentation. Note also that k ¼ 2k0 cosðhÞ ensures the

Fig. 6 Wave vector diagram for


non-forward scattering

123
2 Page 24 of 42 Rev. Mod. Plasma Phys. (2017) 1:2

jk  k0 j ¼ jk0 j so the D 0 is trivially satisfied for xx0 ; because it reduces to


c2 k02  x20 þ x2p ¼ 0 which is the dispersion relation for pump wave. In other words,
the Manley–Rowe relations for the pump and scattered waves require that the
triangle of wave vector produce a wave vector for the scattered wave which lies on
the semi-circle of radius jk0 j.
The dispersion relation for three-wave interaction now takes the form
1 1 2k02 v20
þ ¼ w2 ðh; /Þ; ð57Þ
ve 1 þ vi x0 ðx  MxÞ
where Mx ¼ k:Vg  d, Vg ¼ c2 k0 =x0 , d ¼ c2 k2 =2x0 and wðh; /Þis multiplicative
angular factor determined by the polarization of the light wave. We now discuss the
above equation for scattering off electron modes. In this case, vi is zero and ve is
determined by the susceptibility at electron plasma wave frequency. The dispersion
relation then takes the form:

ðx  MxÞðx2  x2R þ ixCp Þ ¼ 2x0 x2p w2 ðv20 =c2 Þ; ð58Þ

where Cp is the damping rate of the free plasma waves. At x xR , the excited
electrostatic wave satisfies the natural dispersion relation for electron plasma waves
and the equation near threshold reduces to

ðx  xR þ iCp Þðx  Mx þ iC Þ ¼ ðv20 =c2 Þw2 x0 xp ; ð59Þ

where C is the collisional damping rate of the EM wave. This equation demon-
strates decay instabilities of the pump wave into a Raman scattered wave and
plasma waves when threshold given by
  
v20 1 Cp C
2
¼ 2 ð60Þ
c w xp x0
is exceeded. The growth rate a little above threshold is given by

c ¼ ðv0 =cÞwðx0 xp Þ1=2 : ð61Þ

Well above threshold dispersion relation is altered and yields approximate


expression
pffiffiffiffi pffiffiffiffi
x ¼ w2=3 ðv0 =cÞ2=3 ðx0 x2p =4Þ1=3 ½1 þ 3i; 1  3i; 2 : ð62Þ

Clearly when the pump is intense low-frequency mode can be driven far away from
its natural frequency xp . This also suggests even if low-frequency dynamics is
dominated by wave–particle interactions such that x kvthe ; one can still excite
unstable modes if the pump is intense. This has been investigated in Drake et al.
(1974) and forms basis of the stimulated Compton scattering by electrons. Similar
consideration can be given to modes supported by the ion fluids and leads to
stimulated Brillouin scattering. Table 1 gives the summary of the threshold and
growth rates for the electron modes of stimulated scattering instabilities.

123
Rev. Mod. Plasma Phys. (2017) 1:2 Page 25 of 42 2

Table 1 Dependence of growth m0 eE0


¼ mx Growth rate
rate on pump strength for c 0c

electron wave scattering m0


2 x0 xp
 
Cp CCp 1=2
 mc0  x0 xp c Cp
ðxp x0 Þ1=2
 1=2 m0

1=2
xp m0 Cp xp x0
x0 [ c [ 1=2 c
ðxp x0 Þ
m0
 1=2 m0
 1=2
x
c [ xp0 c x2p x0
 1=2 m0
2
m0
[ Cx0 k0 kD c x0 Im v1
c e

Induced Compton Scattering

We now turn to modulational instabilities which are long wavelength instabilities


with k\2k0 cosh these require that x  k:Vg , Vg ¼ c2 k0 =x0 and both Dþ ; D be
retained in the dispersion relation as they are both equally resonant the resulting
four-wave interaction dispersion relation now takes the form
"  #1
1 1 2v20 2 iC 2 2
þ ¼ 2d x  kVg þ d : ð63Þ
1 þ vi ve c 2
2
When jxj  kvi , vDi;e ¼ ke;i =k2 ; Eq. (63) can readily be solved to give
 1=2
iC 2v2 k2 1
x ¼ kVg 
d 1  20 De : ð64Þ
2 c k2 ð1 þ Ti =Te Þ
This solution is unstable if the minimum threshold given by

v20 C
¼ ð1 þ Ti =Te Þk02 k2De ð65Þ
c2 x 0
is exceeded. This happens when C 2d c2 k2 =x0 . For k  vg ¼ 0, this a purely
growing mode and in this limit it is just the filamentation instability of Sect. 4,
producing standing density striation across the incident ‘EM’ wave. From Eq. (63),
we can derive an expression for growth rate of filamentation instability for which a
typical value is cffiðv0 =cÞxpi .
Once the threshold is exceeded, the modes grow with typical growth rates as
shown in the table. One issue which was particularly important for the stimulated
scattering processes was the modification of threshold electric field in inhomoge-
neous plasmas. This theory was extended and put into a firm form by Rosenbluth
(1972). He pointed out that in an inhomogeneous medium, WKB theory demands
that for each wave, k in the direction of plasma inhomogeneity change, keeping
omega constant as per the dispersion relation. This results in a finite region along the
direction of inhomogeneity where the three waves can be said to have their ks and
xs matched. The region of amplification is thus defined by the matching condition
and is of a finite size. To see quantitatively how a finite size of the interaction region
affects the threshold, one might consider a homogeneous plasma problem in which
the pump wave is finite and uniform only over a region of size L and resonant

123
2 Page 26 of 42 Rev. Mod. Plasma Phys. (2017) 1:2

matching condition is satisfied. The backscattered ‘EM’ wave will stay in the region
of finite pump only for a time of order L/c. The inverse of this time is like a damping
time of the backscattered wave and is typically much shorter than the damping time
due to collisions. It therefore replaces the collisional damping rate C ðx2p =2x0 Þ by
convective damping rate c/L which typically makes the threshold higher. The
general condition for absolute instability is that V1 V2 \0, where V1 ; V2 are the group
velocities of the two decay waves and c20 L2 / jV1 jjV2 j [ 1; where c0 is the nonlinear
matrix element responsible for instability. When we consider the case of an
inhomogeneous medium in which k along the density gradient for all waves
changes, the phase mismatch factor is zero at one x and changes linearly in the
neighborhood of zero. This means that delta d ¼ k0x þ k1x þ k2x ¼ 0 at one x and
varies as k0 x around zero, where k0 ¼ dðk0x þ k1x þ k2x Þ=dx. This phase mismatch
gives a phase factor expðik0 x2 =2Þ which essentially means there is no resonant
coupling beyond k0 x2 =2  1. Thus the inhomogeneous medium behaves as if the
medium has a size of order L2eff  2=k0 and the amplification factor in inhomoge-
neous medium is of order or greater than unity if c20 /k0 jV1 V2 j [ 1.
Rosenbluth further investigated the nature of parametric instabilities in an
inhomogeneous medium, whether they are convective or absolute in nature. He
found that for V1 V2 [ 0, the instabilities are convective and propagate away from
the region where they are initiated. For V1 V2 \0 on the other hand, the wave
instability is absolute with a difference. Unlike homogeneous media where absolute
instability grows indefinitely at the point of initiation by a wave packet and can
saturate only by nonlinear effects, in this case the instability saturates, without
nonlinear effects, with an amplification factor c20 / k0 jV1 V2 j at the point of initiation.
Thus the instability may never go into an extremely nonlinear state if the
amplification factor is not too large.
We may calculate k0 from the dispersion relation for pump and decay waves. For
example, for backscattered SRS waves with density gradient providing the
inhomogeneity where

k0x ¼ ðx20  x2p Þ1=2 =c ¼ k0 ð66Þ

k1x ¼ ðx21  x2p Þ1=2 =c ’ k0 ð67Þ

k2x ¼ ðx2  x2p Þ1=2 =vthe ¼ 2k0 ð68Þ

k0 ¼ dðk0x þ k1x þ k2x Þ=dx ð69Þ

k0 is dominated by dk2x =dx ¼ ð1=4k0 v2the Þðdx2p =dxÞ. Similarly the dominant contri-
bution to k0 may come from gradients in temperature, plasma flow, etc., depending
on the wave mode being considered and will determine amplification factor for the
inhomogeneous problem. In this manner, the threshold of stimulated scattering for
inhomogeneous media may be calculated (Nishikawa and Liu 1976).

123
Rev. Mod. Plasma Phys. (2017) 1:2 Page 27 of 42 2

In order to assess the importance a given stimulated scattering process, one


essentially calculates the amplification factor for a given instability. If it is not too
large, the scattering is not too important and may essentially be calculated from the
above linear calculations. If the exponent becomes a factor much larger than unity,
the scattering process becomes nonlinear and we must retain nonlinear effects such
as pump depletion and other quasi-linear processes, wave–wave interactions
including cross-coupling of the electrostatic waves of SRS and SBS, etc., and
calculate the amplification in the presence of nonlinear processes. One also couples
the equations with those evolving the plasma profiles by an appropriate hydrody-
namic code which includes selfconsistent effects like energy and momentum into
the plasma through the damping of electrostatic waves involved in the scattering
process. One such outstanding combination of codes is LASNEX and PF3D codes
(Lindl et al. 2014; Still et al. 2000).

6 Relativistic nonlinear propagation of light in plasmas

For relativistically intense laser light pulses, the parameter ðeE=mxcÞ [ 1. This
corresponds to a threshold incident power density of order
Ith  m2 x2 c3 =ð4pe2 Þ  1018 W=cm2 for a 1-lm laser. This much power is routinely
available these days in picosecond lasers using the chirped pulse amplification
technique. Impressive studies using such lasers are being carried out on fast ignition
concept of laser fusion, basic studies of warm dense matter, table top accelerators
and radiation sources, laboratory simulation of astrophysical phenomena, etc.
As already discussed, small-amplitude electromagnetic waves propagate linearly
through a plasma, treating it as a dispersive dielectric. The cold plasma dielectric
constant differs from unity and was shown in Sect. 2.1 to be given by
 ¼ ðc2 k2 =x2 Þ ¼ 1  x2p =x2 : ð70Þ

The departure from unity arises because the free electrons in the plasma generate a
conduction current density in response to the electric fields created by the changing
magnetic fields. This opposes the displacement current and attempts to shield the
plasma from the magnetic field of the incoming wave. Note that
xp ¼ ð4pn0 e2 =mÞ1=2 . For x\xp , \0, the propagation k is imaginary (evanescent
wave) and the shielding is complete. Thus we say that ‘EM’ waves do not propagate
in overdense plasmas (x\xp ) and are reflected from them. Let us see how the
concept of overdense plasmas gets modified for relativistically intense waves with
the parameter a = ðeE=mxcÞ [ 1. We expect the electrons will become heavier
due to relativistic effects and reduce the effective plasma frequency, so that waves
in the range xpeff \x should be able to propagate. This effect was found by
Akhiezer and Polovin (1956) for circularly polarized waves. For linearly polarized
waves one discovers high phase velocity almost transverse waves in which the
conduction current saturates at n0 ec for most of the wave period leading to nonlinear
sawtooth ‘EM’ waves propagating in overdense plasmas. Such solutions were first
found numerically by Kaw and Dawson (1970) and soon thereafter given an

123
2 Page 28 of 42 Rev. Mod. Plasma Phys. (2017) 1:2

analytical description by Max and Perkins (1971). The basic equations describing
these phenomena are the Maxwell equations and the relativistic equations for the
cold electron fluid, the latter being written as (Akhiezer and Polovin 1956):
op vB
þ vrp ¼ eðE þ Þ; ð71Þ
ot c

p ¼ mvð1  v2 =c2 Þ1=2 : ð72Þ

We look for one-dimensional nonlinear solutions by making the traveling wave


ansatz and introducing the phase velocity u with the variables
u p xpe
b¼ ; q¼ ; n¼ ðz  utÞ ð73Þ
c mc c
to get the following nonlinear ordinary differential equations:

o2 qx 1 bqx
2
þ 2 ¼ 0; ð74Þ
on ðb  1Þ bð1 þ q2 Þ1=2  qz

o2 qy 1 bqy
2
þ 2
¼ 0; ð75Þ
on ðb  1Þ bð1 þ q2 Þ1=2  qz

o2 ½bqz  ð1 þ q2 Þ1=2 qz
þ ¼ 0: ð76Þ
on2 bð1 þ q2 Þ1=2  qz
One class of exact nonlinear solutions is that of circularly polarized ‘EM’ waves
with qz ¼ 0 and q=constant, viz.

e2 E02
q2 ¼ ; E02 ¼ Ex2 þ Ey2 ; ð77Þ
m2 c2 x2
 
mcxq x
Ex ¼ sin n ; ð78Þ
e bxpe
 
mcxq x
Ey ¼ cos n ; ð79Þ
e bxpe
 1=2
x2pe e2 E02
¼1 1 þ ; ð80Þ
x2 m2 c2 x2
Propagation into overdense plasmas will take place provided
 1=4
e2 E 2
xpe 1 þ 2 2 0 2 \x: ð81Þ
m c x
Circularly polarized transverse waves arise in the special case where the vB forces
along the longitudinal direction exactly balance out. In a more general situation such

123
Rev. Mod. Plasma Phys. (2017) 1:2 Page 29 of 42 2

Fig. 7 Strong linearly polarized ‘EM’ waves. The curve shows the transverse and longitudinal
component of electron momentum as well as the transverse electric field. All quantities have been
normalized to their maximum value. The period is P

as a linearly or elliptically polarized case, we get mixed transverse longitudinal


modes because the vB forces do not exactly balance out. Such cases were
numerically investigated in Kaw and Dawson (1970) and analytically described in
Kaw and Dawson (1970) and Max and Perkins (1971) (Fig. 7).

ðn  ð1=4ÞPÞ2
qx ¼ q0  ; 0  n  ð1=2ÞP; ð82Þ
2ðb2  1Þ

ðn  ð3=4ÞPÞ2
qx ¼ q0  ; 1=2P  n  P; ð83Þ
2ðb2  1Þ
where P is the period of the periodic wave. For much of the period vx tends to c and
the solution for transverse momentum is a series of linked parabolas. The corre-
sponding electric field has a sawtooth shape. The longitudinal variable and its
description both in the main solutions and the boundary layer are described in Max
and Perkins (1971). This solution also shows that light will penetrate into an
overdense plasma.

7 Implications and recent advances

7.1 Application to laser fusion

The most important application of the above-mentioned physical concepts has been
in the problem of direct and indirect drive for laser fusion. A direct example of this
is the preparation and post-experiment analysis for the NIF campaign (Lindl et al.

123
2 Page 30 of 42 Rev. Mod. Plasma Phys. (2017) 1:2

2014; Still et al. 2000) on utilization of indirect drive hohlraum capsules for laser
fusion. In these experiments, light is allowed to enter a hohlraum capsule in the
form of a large number of oriented and focused laser beams falling on the high
Z-coated inner wall of the hohlraum. The high Z plasma is supposed to absorb the
incoming laser light and then emit X-ray radiation, which is trapped inside the
hohlraum cavity giving a thermal spectrum of X-ray radiation with a characteristic
temperature  300eV. However, because of the high intensities involved, SRS, SBS
and filamentation have a great deal of impact on the above process. The effects are
largely deleterious since they spread the directed laser energy in directions away
from where it is aimed and also convert part of the energy into electrostatic waves
which produce energetic electrons and ions by Landau damping. To make an
estimate of the magnitude of stimulated scattering extensive numerical codes have
been developed and benchmarked (e.g., PF3D) with experiments done in the
laboratory. These usually are in the form of beam propagation codes, which trace
the propagation of incident laser beams into an expanding plasma with its
inhomogeneities which are being selfconsistently generated by sophisticated
multispecies hydrodynamic (including multigroup electrons to take account of the
energetic electron population) codes like LASNEX. The stimulated scattering
calculations are done with various degrees of sophistication. The simplest is to
calculate the amplification factor of unstable Brillouin or Raman scattered wave in
the presence of plasma inhomogeneities like, density, temperature, flow velocity
inhomogeneity, etc. The more advanced calculations retain nonlinear effects like
pump depletion, interference of electrostatic wave components of SRS and SBS
such as collapse of SRS plasma waves by SBS ion acoustic waves and growth of
SBS waves by the ion acoustic cavities created by SRS electron plasma waves. One
can also include the focusing of laser beams with RPP (random phase plates) which
is used sometimes to weaken the impact of parametric instabilities (Still et al. 2000;
Berger et al. 1998). Local absorption of the electrostatic component of the waves
can lead to production of energetic electrons and/or ions which then deposit their
energy into the plasma. Here also several degrees of sophistication are possible. The
ions tend to deposit their energy locally through collisional effects. The electrons,
on the other hand, have a tendency to go to longer distances and do nonlocal heating
by excitation of return currents, excitation of turbulence leading to turbulent
heating, etc. The deposition of energy and momentum into the plasma leads to an
impact on the overall expansion that the plasma is undergoing. Nonlocal electron
thermal conductivity effects can also be incorporated.
This is also the fate of the energy which reaches the critical surface for reflection
which also promotes the energy deposition in electrostatic waves/energetic particles
by phenomena such as resonance absorption, decay and OTSI, etc. The magnitude
of the energetic electron problem can only be ascertained by carrying out detailed
theory of the parametric instabilities in the presence of an evolving plasma density
profile. Energetic electrons are a problem because they have the feature of entering
the high-density region of the plasma and preheating it before the compressional
shock being sent to compress the plasma reaches there. Preheating makes the
compression difficult and is hence an undesirable feature of these instabilities. The
upshot of these investigations is that although the entire range of physical

123
Rev. Mod. Plasma Phys. (2017) 1:2 Page 31 of 42 2

phenomena may not have been incorporated in the calculations yet, one has been
able to understand the observed phenomena and gainfully optimize them such that
more than 80% absorption of the incident laser energy is assured and even preheat
can be controlled to values less than a few percent of the thermal energy ultimately
trapped in the capsule. Furthermore, issues on the creation of asymmetries in the
thermal radiation have also been understood and techniques to ameliorate their
impact discovered (Murphy et al. 1998; Turner et al. 2000, 2003). Thus one may
justifiably argue that the theory of parametric instabilities as applied to laser fusion
has reached a high level of maturity since the early days of discovery of these
effects. It is another matter that our overall capability to accurately predict the laser
energies required to ignite a hohlraum fusion capsule has not reached a satisfactory
level of sophistication yet (Lindl et al. 2014).

7.2 Collisionless linear light absorption: some novel empirical results

For a period of about one and a half decades, our laser plasma group at IPR led by
Amita Das has been actively interacting with the experimental laser plasma group at
TIFR , headed by G. Ravindra Kumar. Our first collaboration was an empirical
investigation of waves breaking in the resonance absorption of a light wave
obliquely incident on an inhomogeneous plasma. This is described below (Sandhu
et al. 2005). Besides this we have mostly looked at the physics of interaction of
short laser pulses with overdense plasmas and have provided the modeling expertise
to explain the experimental results. We have also been collaborating with
experimental and simulation groups in Osaka. I have been actively participating
in all of this activity. The important contributions of this joint activity are described
in Sect. 7.4.
Resonance absorptions is a linear mechanism which can directly couple a light
wave to electrostatic waves and thus lead to collisionless absorption. This
mechanism was discussed by Denisov in the ionospheric context and is known as
resonance absorption. In resonance absorption, one considers a p polarized wave
obliquely incident on a plasma layer. It goes to the point xcosh ¼ xp and reflects
from there. This is the cut-off point for the wave. But beyond the cut-off is the
resonance x ¼ xp . If this point is not too far off from the cut-off, the ‘EM’ field
tunnels to the resonance and excites longitudinal waves there. The basic reason is
that for an obliquely incident p polarized wave, there is a component of the electric
field along the density gradient. Oscillations induced by the light wave create
density fluctuations oscillating at the light frequency and this can resonantly pump
the plasma waves at the resonance point which may then get damped, by collisions,
Landau damping in warm plasmas, wave breaking in cold plasmas or simply by
propagating away from the region of resonance and eventually giving their energy
to the particles. Denisov demonstrated that the absorption of the light wave by
coupling to plasma waves is independent of the detailed mechanism of absorption of
the secondary waves and instead depends critically on the angle of incidence for
which an optimum exists. If the incident angle is too small, the electric field has too
small an component along the density gradient and the coupling is weak. If the angle

123
2 Page 32 of 42 Rev. Mod. Plasma Phys. (2017) 1:2

is too large (closer 0 to 90 ), the coupling is good but the cut-off is too far from the
resonance and there is very little tunneling to the resonance point. Thus there is an
optimum at an intermediate angle, which for linear density profiles is given by
 ð1=3Þ
xL
U¼ sinh ffi 0:8: ð84Þ
c
And this is where the absorption and coupling to plasma waves by resonance
absorption is maximum. Resonance absorption is applicable when the inhomoge-
neous plasma layer produced by expansion is rather long (  or greater than frac-
tions of a vacuum light wavelength). When the layer is of a size much smaller than a
wavelength which happens when the light pulse has a short duration (sub-pi-
cosecond) and a high contrast (  1010 has been achieved), then the light pulse
typically faces an overdense plasma with a sharp interface. As we shall see later the
physics of absorption in this case is much different.
Around the year 2000, Ravindra Kumar and his associates at TIFR (Sandhu et al.
2005) initiated experimental work on resonance absorption in cold plasmas, where
wave breaking is the mode of conversion of plasma wave energy to energetic
particle energies. Wave breaking is a nonlinear wave–particle interaction in a cold
plasma. It occurs typically when the kinetic energy picked up by the particle during
interaction with the wave is enough to make it resonantly interact with the wave
even though its original velocity is negligible (cold plasma), i.e., when
e/ [ ðm=2Þðx=kÞ2 or Vtr  ð2e/=mÞ1=2 [ ðx=kÞ this condition is satisfied, one
expects the wave to interact with all the cold particles in the plasma and to deliver
its energy to the particles. In the beautiful experiment done at TIFR (Sandhu et al.
2005), this phenomenon has been observed. What the TIFR group did was to
provide a very good evidence of this optimization process using a prepulse and a
delayed laser main pulse to prepare a plasma with varying scale length of density
variation to study the optimization with respect to the key parameter. They
measured the level of plasma oscillations by the level of second-harmonic
generation by such waves in an inhomogeneous plasma and the process of wave
breaking induced collisionless heating by diagnosis of the soft X-ray emission due
to bremsstrahlung from such hot electrons. As shown in the Fig. 8a, b:
The experiments show that for cases where the level of plasma waves excited is
not large enough for wave breaking, both the variations are similar, i.e., the second-
harmonic efficiency and the X-ray generation both go through a single peak and
then decay as expected for a process which has an optimal angle. On the other hand,
when the intensity of the plasma waves excited is large enough for wave breaking
the second-harmonic efficiency, showed a characteristic bite out in the second-
harmonic emission as the key optimization index U (as defined in Eq. 84) was
varied by changing the scale length of density variation (in turn modified by
changing the delay between the prepulse and the main laser pulse) whereas the soft
X-ray emission was going through a peak. We know by Denisov’s theory that if one
varies the scale length of density variation, the resonance absorption process will
pass through an optimum where the generation of plasma waves will be most
intense. When the plasma wave produced by resonance absorption is below the

123
Rev. Mod. Plasma Phys. (2017) 1:2 Page 33 of 42 2

Fig. 8 a Integrated yield of X-rays from 30 to 300 keV, b second-harmonic efficiency /I2x =Ix2 , as a
function of time delay between the prepulse and the main pulse. Each is normalized with respect to yield
with main pulse alone

critical amplitude, it does not break and one sees a strong signal in the second-
harmonic radiation detector as expected. For parameters where, the generated
plasma wave is intense enough to cause wave breaking, the second-harmonic signal
suffers a bite out and the soft X-ray signal which gives an idea of the number of
energetic particles generated through the bremsstrahlung that they produce (in the
form of soft X-rays) passes through a peak. At higher density scale lengths, the
plasma wave generation becomes less optimal again and the second-harmonic signal
picks up again with a concomitant reduction in the bremsstrahlung radiation signal.
We have provided a quantitative theory of this observed variation in reference
Sandhu et al. (2005), thus providing a key evidence for the importance of resonance
absorption in plasmas with gentle scale lengths.
Another mechanism that may be used to cause collisionless damping of light
waves in sharp interface overdense plasmas is by the excitation of surface waves by
laser pulses. I shall briefly describe this mechanism because this has also been

123
2 Page 34 of 42 Rev. Mod. Plasma Phys. (2017) 1:2

verified experimentally in recent years by Kahaly et al. (2008). It is well known


(Kaw and McBride 1970) that a sharp interface plasma supports surface waves with
the dispersion relation
c2 kjj2 x2p  x2
¼ ð85Þ
x2 x2p  2x2

This gives us mixed electrostatic–electromagnetic waves with a surface phase


pffiffiffi
velocity below the velocity of light and a maximum frequency xp = 2 in the
electrostatic limit. Since these waves are traveling with a velocity less than c, they
can interact with the particles and are therefore good candidates for transferring
energy from a laser beam to particles provided they can be excited by light waves.
Furthermore, if the sharp interface opens up, their frequency matches the natural
frequency of bulk plasma waves at some local point in the interface. This is the
point where mode conversion occurs and the surface wave loses most of its energy
to bulk waves. Surface waves are therefore excellent candidates for acting as
intermediaries for transferring laser energy to particles. Unfortunately, there is a
difficulty, viz. surface waves have a surface phase velocity less than the velocity of
light and a light wave obliquely incident on a surface will have a phase velocity
greater than the velocity of light. Thus a light wave does not couple to surface
plasma waves directly. However, if we slow down the light wave by coupling it to
microscopic and nano structures on the surface we can hope to get a good coupling.
In Kahaly et al. (2008) it has been shown how nano structures in the form of a sub-
lambda grating can be used to resonantly excite surface waves and essentially arrive
at a near 100% energy absorption from the laser, the energy then immediately
appearing in the particle distribution in collisionless plasmas.

7.3 Strong turbulence of plasma waves: Langmuir collapse

The parametric instability theory, in particular, the OTSI has had a profound
influence on the development of theory of strong turbulence of electron plasma
waves. Weak turbulence theory of nonlinear plasma waves is dominated by three-
wave interaction processes, of which the resonant decay instability is the most
important. This is because the basic ansatz of weak turbulence theory that the
nonlinear interaction parameter cnl  xL (the linear eigen frequencies , of which xs
is lower) rules out other robust instabilities like OTSI in which nonlinear interaction
can shift real part of xL to zero. Now it may be noted that the only resonant decay
instability permitted by the dispersion relations is a plasma wave going into another
plasma wave and an ion acoustic wave. This may be seen from the Manley–Rowe
relations (energy, momentum conservation in the interaction) as applied to decay
instabilities, viz.:
x1 ¼ x2 þ xs ; k1 ¼ k2 þ k3 ð86Þ

where

123
Rev. Mod. Plasma Phys. (2017) 1:2 Page 35 of 42 2

x1 ¼ xk1 ¼ ðx2p þ 3k12 v2the Þ1=2 ; ð87Þ

x2 ¼ xk2 ¼ ðx2p þ 3k22 v2the Þ1=2 ; ð88Þ

k3 kD xpi
xs ¼ : ð89Þ
ð1 þ k32 k2D Þ1=2
This has the obvious implication that decay instability shifts the energy to plasma
waves with lower frequency and hence longer wavelengths. Multiple sequential
decay instabilities then shift the energy to the longest wavelength of the system
leading to a condensation of plasma wave energy in the longest wavelength mode
and the so-called ‘‘infra-red‘‘ catastrophe. Zakharov (1972) saw clearly that OTSI
was a way out of the catastrophe because it shifted the energy to shorter wave-
lengths. We may recall the basic feature of OTSI, viz. pump frequency x0 \xR the
frequency of the excited plasma wave , which here translates to x1 \x2 or shifting
of energy to sharper scales. Thus, if our description includes both decay processes
and OTSI, we have the possibility of describing strong turbulence of plasma waves
where xs  cnl  dxR (linear dispersive width of the plasma line)  xp . Zakharov
then showed that the dominant nonlinear physics of electron plasma waves is
contained in the nonlinear equations

o2 Ns
 lr2 Ns ¼ lr2 jEj2 ; ð90Þ
ot2
 
oE
r i þ r2 E ¼ r  ðNs EÞ; ð91Þ
ot
which are written in normalized variables
pffiffiffiffiffiffiffiffi
xp t ¼ t; x= 3=2kD ¼ x; l ¼ ð2=3Þðm=MÞ; jEj2 =8pnT ¼ jEj2 ; dn=n0 ¼ Ns :
ð92Þ
These equations are known as the Zakharov equations and contain the dominant
dynamics of nonlinear plasma waves and their interaction with ion acoustic waves.
Note that the fundamental physics of parametric interactions, namely driving of
low-frequency acoustic waves by the radiation pressure of plasma waves (Eq. 90)
and the coupling of ion waves with plasma waves to regenerate plasma waves, Ns E
term in second equation, form the heart of these equations. Zakharov noticed that if
the o2 Ns =ot2 term could be neglected the combined equation becomes the nonlinear
Schrödinger equation:
oE
i þ r2 E þ jEj2 E ¼ 0 ð93Þ
ot
Zakharov and Shabat (1972) discovered an inverse scattering formalism for solving
the 1-D version of Eq. (93) exactly. This showed that starting from arbitrary initial

123
2 Page 36 of 42 Rev. Mod. Plasma Phys. (2017) 1:2

conditions the solutions go into formation of envelope solitons such as given by the
expression.

E ¼ Em sinðkx  xtÞ=coshðk0 nÞ where n ¼ x  Vg t ð94Þ

Vg ¼ 3kk2D xp ; k0 kD ¼ Em =ð48pnTÞ1=2 ; x ¼ xp ½1 þ ð3=2Þk2 k2D  Em2 =ð32pnTÞ


ð95Þ
Envelope solitons are coherent nonlinear entities where the plasma wave is trapped
in a density cavity that it digs itself by ponderomotive forces and plasma nonlin-
earity and diffraction/dispersion effects balance. Thus in one dimension, one can
define the strongly turbulent state of electron plasma waves, to be formed by a
collection of envelope solitons. Each envelope soliton is a collection of plasma
waves which have been woven into a coherent stationary state by the large nonlinear
width cnl which overpowers the randomness due to xs and dxR where latter is the
frequency spread of the plasma waves due to thermal dispersion. Thus it needs a
minimum threshold intensity of plasma waves, which is given by
Em2 =4pnT [ ðMxR Þ2 . The fact that one will get a stationary state in 1-D is clearly
indicated by a conservation law of the nonlinear schrodinger Eq. (93), viz.
Z
jEj2 d 3 r ¼ constant: ð96Þ

This conservation law is a conservation law for action density, jEj2 =xp , which is
directly proportional to the number of plasmons involved in the interactions. This
number is conserved because only low-frequency interactions with dxp  xp are
involved in the interactions. Equation (96) clearly shows that as the interaction
proceeds jEj2 Ld  constant, where d is the number of dimension. The nonlinear
schrodinger equation has 3 terms. The second term o2 E=ox2 is a dispersive/
diffractive term and essentially spreads the perturbation. The third term or the
nonlinear term jEj2 E on the other hand, produces a steepening of the structures and
generation of sharper structures. Let us suppose we start with a large perturbation so
nonlinear term is large and the structure starts steepening. It must do so keeping
jEj2 L constant. This means that L  1=jEj2 and that the diffraction term
E=x2  E=L2  jEj4 E and therefore increases faster than jEj2 E. Thus the system will
be driven to a stationary state where the two terms balance and we get envelope
soliton solutions.
On the other hand, if the plasma was described by three-dimensional equations
with o2 E=ox2 replaced by r2 E, the equations are found to have no stationary
solutions but only lead to collapsing cavity solutions. This also follows from the
above argument. In this case, jEj2 L3  constant and as the nonlinear entity steepens,
the nonlinear term dominates over the dispersive term because of a faster inverse L
dependence and so the collapse continues. Thus a three-dimensional cavity keeps
shrinking in size till it becomes of order the Debye length and then the fields have
very rapid wave–particle interaction with the plasma particles and the entire plasma

123
Rev. Mod. Plasma Phys. (2017) 1:2 Page 37 of 42 2

wave energy is converted into extended tails of energetic particles adorning the
electron velocity distributions. This has resulted in a totally new picture of strong
plasma turbulence in 3-D. The basic conclusion is that no stationary states are
possible and the energy goes into collapsing cavities of plasma waves trapped in ion
wave depressions they create themselves and these cavities eventually reach the size
of Debye length at which point intense wave–particle interaction occurs and the
energy is soaked up by the particles. A large amount of simulation data exists which
agrees with this remarkable theory of strong plasma wave turbulence Goldman
(1984).

7.4 Recent studies of light interaction with overdense plasmas

When the plasma vacuum interface is sharp, the mechanism of resonance absorption
described above does not work. The oscillating electric field from an obliquely
incident p polarized laser pulls electrons out of a nonlinear evanescent layer and
speedily accelerates them inwards to high energies (called Brunel mechanism or
vacuum heating), generating bursts of relativistic electron currents which propagate
towards the plasma interior. If the incident light pulse is relativistically intense, then
a strong mechanism to send bursty relativistic electron currents inwards exists even
for normal incidence through the JB nonlinear term. These high-intensity
relativistic currents injected into the high-density plasma lead to a host of other
effects, which we shall discuss now. These studies are of interest to fast ignition
concept of laser fusion. They also provide fundamentally new information about the
formation and evolution of turbulent magnetic fields in the laboratory.
The electron current densities are huge, typically of order nc ec  1012 A=cm2
leading to a total current  10 MA in a focal spot about 30 microns in diameter.
This is greater than the Alfvén limit of electron current which can propagate in
vacuum

IA ffi 17000bcAmps; c ¼ ð1  b2 Þ1=2 ð97Þ

At I [ IA in vacuum, the electrons are turned around by the magnetic fields they
themselves create and so propagation is impossible. Propagation in a plasma is,
however, possible. This is because the plasma shields itself from the magnetic fields
associated with the incoming relativistic current. This is done by creating return
currents in the background plasma in such a manner that the net magnetic field
vanishes. This system of oppositely directed electron currents is, however, sus-
ceptible to plasma instabilities with the result that the transport and decay of for-
ward relativistic currents is different from that of the background plasma return
currents. This leads to the generation of turbulent magnetic fields which can tran-
siently reach peak values up to 100 megagauss. A diagnosis of the nature and
evolution of these turbulent magnetic fields can thus become a very good probe of
the energy transport of the incident laser energy by the relativistic electrons they
produce. The second series of experiments at TIFR which I wish to describe here,
has been devoted to understanding the nature of generation mechanisms of turbulent
magnetic fields in the interaction of short picosecond laser pulses with overdense

123
2 Page 38 of 42 Rev. Mod. Plasma Phys. (2017) 1:2

plasmas. Here, by a clever use of the adjustable time delay between the pump and
the probe for the magnetic field measurement, the evolution of such magnetic fields
has been studied on a picosecond time scale and has led to several novel results.
Magnetic fields, being mostly oriented perpendicular to the direction of laser
propagation, are measured by the Cotton-Mouton effect which measures change in
the polarization parameters of the laser beam as it propagates in the magnetized
plasma. I shall now provide a brief account of this work together with the inter-
pretation we have provided for the observations. A PIC simulation of interaction of
an intense laser pulse with an overdense plasma (Sentoku et al. 2003) shows that a
bunch of relativistic electrons is produced at the surface and penetrate towards the
interior with current densities of order 1012 A=cm2 . The plasma attempts to shield
itself from the transient magnetic fields by the excitation of return currents in the
background cold plasma. This system of oppositely directed electron currents is
susceptible to current filamentation and beam plasma instabilities which break up
the bulk volume currents into separated sheet currents. These sheet currents then
tear, merge and coalesce and are broken up along the direction of current flow by
Kelvin–Helmholtz-type current shear instabilities (Das and Kaw 2001) leading to
the formation of EMHD current dipoles with forward relativistic currents along the
central axis and surrounded by a return current in a cylindrical shell. Such EMHD
dipoles with dimensions of order the skin depth are interesting nonlinear solutions
of the EMHD fluid equations (Das 1999). Kinetic effects on such EMHD dipoles
have also been analyzed by PIC simulations and the parameter space where they
have a long lifetime has been established (Hata et al. 2013). Interesting interaction
dynamic of dipoles with each other has been studied by Das and her associates as
has been their propagation into inhomogeneous plasmas (Das 1999; Yadav et al.
2008, 2009; Yadav and Das 2010; Gaur et al. 2009).
As these structures propagate into higher density regions they become sharper
leading to shock formation and shock heating of the electron fluid by a turbulence
driven anomalous resistivity/anomalous electron viscosity mechanism (Das 1999;
Yabuuchi et al. 2009). In Yabuuchi et al. (2009), certain features of the Osaka
heating experiment have been explained by the shock heating mechanism. These
results should be of considerable interest to the fast ignition fusion enthusiasts and
should therefore be also verified by PIC simulations which can resolve skin depth
size structures at even the highest densities involved. Such simulations remain to be
done. In Sandhu et al. (2002), we observed and modeled the residual field left in the
plasma for about  10 picoseconds after the 150 femto-second laser pulse and
associated REB pulse had left the scene. The decay of plasma currents was faster
than that expected of a plasma with the observed plasma parameters and was
ascribed to an anomalous resistivity associated with turbulent motions of the
electron fluid (Jain et al. 2007). From the fast ignition community the consensus
seems to be that the electron fluid-induced magnetic turbulence is detrimental and is
responsible for giving a large divergence to the incoming REB preventing it from
efficiently hitting the compressed core and thereby heating it. Currently, suggestions
of providing transverse nanostructures of size sharper than skin depth on the target,
which stabilizes some of the current-driven electron fluid instabilities (Mishra et al.

123
Rev. Mod. Plasma Phys. (2017) 1:2 Page 39 of 42 2

2014) and that of strongly magnetizing the electron fluid by externally produced
axial magnetic fields  –10 kT, are being investigated for the fast ignition problem
(Hohenberger et al. 2012).
The study of turbulent magnetic fields, their origins and evolution, is a topic of
great interest to astrophysicists. It has been found that the later stages of the above
experiments can reproduce certain universal features of decaying solar astrophysical
turbulence. Spectral features of the generated turbulent magnetic fields for time
scales up to  1 picosecond were studied in Mondal et al. (2012) and were found to
be consistent with features of turbulence in electron fluids. On the other hand, when
multi-picosecond evolution of the decaying turbulence was carefully investigated, it
was fond that after a few picoseconds the spectral power density develops a knee at
kqi  1 signaling the participation of ion motion in the turbulent plasma. It is also
found that the slope of the spectral power density has different values on the two
sides of the knee, as is expected in a plasma with turbulent motions of magnetized
ions. Many new features of the evolving spectral power density have been noted and
given an explanation on the basis of the properties of the underlying turbulence
(G Ravindra et al. Spectra of magnetic turbulence in multi picosecond laser
experiment. Private communication).

8 Conclusions

We have presented a review of laser interaction with plasmas of relevance to


nuclear fusion. The discussion was divided into two main parts based on interaction
of lasers with weakly inhomogeneous plasma layers and those of interaction with
very sharp layers. In the former, we considered parametric instabilities as the main
nonlinear effect and showed that they can assist in the absorption of light by
conversion into collective electrostatic modes near the critical layer of reflection
ðx  xp Þ but lead to deleterious scattering processes like SBS, SCS and SRS and
filamentation in the underdense regions. In plasmas with sharp interfaces, we briefly
showed how the relativistic electron beams generated at the interface lead to
selfconsistent generation of turbulent magnetic fields which create a divergence of
the beams, thereby reducing their contribution to core plasma heating in the fast
ignition concept. It was also projected that nonlinear entities like EMHD current
dipoles may change this picture if they play a prominent role in the interactions.
What can we say of the future? It is obvious that work on fusion by lasers will
continue and methods will keep getting invented which help us avoid the deleterious
effects and continue to utilize the good effects. Other many applications of lasers
will also progress on many fronts. New things will open up on the horizons of laser–
plasma interactions as the intensity of laser goes up. Currently, intensities up to
1021 W=cm2 have been reached. Soon one will get into the regime of 1023 
1024 W=cm2 range and such effects as effect of radiation reaction forces on the
motion of single electrons and collective electron fluid will begin to play a role. One
expects effects due to deviations from the 90 degree phase difference between
velocity and electric fields, in addition to other novel effects. From 1026 

123
2 Page 40 of 42 Rev. Mod. Plasma Phys. (2017) 1:2

1029 W=cm2 pair production effects will come in (the former power is sufficient to
produce pairs from gamma rays generated by laser light scattered by beams of
electrons) and change the qualitative picture of interaction of light with plasmas.
Similarly, X-ray and gamma ray lasers may also become available with their own
opportunities. One cannot see clearly further into the future, but it promises to be an
exciting time for the study of laser–plasma interactions.

References

A.I. Akhiezer, R.V. Polovin. Theory of wave motion of an electron plasma. Soviet Phys. JETP, 3 (1956)
R.L. Berger, C.H. Still, E.A. Williams, A.B. Langdon, On the dominant and subdominant behavior of
stimulated raman and Brillouin scattering driven by nonuniform laser beams. Phys. Plasmas 5(12),
4337–4356 (1998)
A. Das, Nonlinear aspects of two-dimensional electron magnetohydrodynamics. Plasma Phys. Control.
Fusion 41(3A), A531 (1999)
A. Das, P. Kaw, Nonlocal sausage-like instability of current channels in electron magnetohydrodynamics.
Phys Plasmas 8(10), 4518–4523 (2001)
A. Das, P. Kaw, Propagation of electron magnetohydrodynamic structures in a two-dimensional
inhomogeneous plasma. Phys. Plasmas 15(6), 062308 (2008)
A. Das, P. Kaw, S. Sengupta, Anomalous energy dissipation of electron current pulses propagating
through an inhomogeneous collisionless plasma medium. Phys. Plasmas 16(4), 040701 (2009)
J. Dawson, On landau damping. Phys. Fluids (1958–1988) 4(7), 869–874 (1961)
J. Dawson, P. Kaw, B. Green, Optical absorption and expansion of laser-produced plasmas. Physics of
Fluids (1958–1988) 12(4), 875–882 (1969)
J. Dawson, C. Oberman, High-frequency conductivity and the emission and absorption coefficients of a
fully ionized plasma. Phys. Fluids (1958–1988) 5(5), 517–524 (1962)
J. Dawson, C. Oberman, Effect of ion correlations on high-frequency plasma conductivity. Phys. Fluids
(1958–1988) 6(3), 394–397 (1963)
N.G. Denisov, On a singularity of the field of an electromagnetic wave propagated in an inhomogeneous
plasma. Soviet Phys. JETP-USSR 4(4), 544–553 (1957)
J.F. Drake, P.K. Kaw, Y.-C. Lee, G. Schmid, C.S. Liu, M.N. Rosenbluth, Parametric instabilities of
electromagnetic waves in plasmas. Phys. Fluids (1958–1988) 17(4), 778–785 (1974)
D.F. DuBois, M.V. Goldman, Radiation-induced instability of electron plasma oscillations. Phys. Rev.
Lett. 14, 544–546 (1965)
D.W. Forslund, J.M. Kindel, E.L. Lindman, Nonlinear behavior of stimulated Brillouin and raman
scattering in laser-irradiated plasmas. Phys. Rev. Lett. 30(16), 739 (1973)
J.P. Freidberg, R.W. Mitchell, R.L. Morse, L.I. Rudsinski, Resonant absorption of laser light by plasma
targets. Phys. Rev. Lett. 28(13), 795 (1972)
G. Gaur, S. Sundar, S.K. Yadav, A. Das, P. Kaw, S. Sharma, Role of natural length and time scales on
shear driven two-dimensional electron magnetohydrodynamic instability. Phys. Plasmas 16(7),
072310 (2009)
I.R. Gekker, O.V. Sivukhin, Soviet J. Exp. Theor. Phys. 9, 243 (1969)
V.M. Goldman, Strong turbulence of plasma waves. Rev. Modern Phys. 56(4), 709 (1984)
M. Hata, H. Sakagami, A. Das, Kinetic effects on robustness of electron magnetohydrodynamic
structures. Phys. Plasmas 20(4), 042303 (2013)
M. Hohenberger, P.-Y. Chang, G. Fiksel, J.P. Knauer, R. Betti, F.J. Marshall, D.D. Meyerhofer, F.H.
Séguin, R.D. Petrasso, Inertial confinement fusion implosions with imposed magnetic field
compression using the omega laser). Phys. Plasmas 19(5), 056306 (2012)
N. Jain, A. Das, P. Kaw, S. Sengupta, Nonlinear electron magnetohydrodynamic simulations of sausage-
like instability of current channels. Phys. Plasmas 10(1), 29–36 (2003)
N. Jain, A. Das, P. Kaw, S. Sengupta, Sausage instabilities in the electron current layer and its role in the
concept of fast ignition. Nuclear Fusion 44(1), 98 (2003)

123
Rev. Mod. Plasma Phys. (2017) 1:2 Page 41 of 42 2

N. Jain, A. Das, P. Kaw, Sudip Sengupta, Role of current shear driven turbulence in anomalous stopping
of energetic electrons. Phys. Lett. A 363(1), 125–129 (2007)
N. Jain, A. Das, S. Sengupta, P. Kaw, Nonlinear electron-magnetohydrodynamic simulations of three
dimensional current shear instability. Phys. Plasmas 19(9), 092305 (2012)
P.K. Kaw, J.M. Dawson, Laser-induced anomalous heating of a plasma. Phys. Fluids (1958–1988) 12(12),
2586–2591 (1969)
P. Kaw, J. Dawson, Relativistic nonlinear propagation of laser beams in cold overdense plasmas. Phys.
Fluids (1958–1988) 13(2), 472–481 (1970)
P.K. Kaw, J.B. McBride, Surface waves on a plasma half-space. Phys. Fluids (1958–1988) 13(7),
1784–1790 (1970)
P. Kaw, G. Schmidt, T. Wilcox, Filamentation and trapping of electromagnetic radiation in plasmas.
Phys. Fluids (1958–1988) 16(9), 1522–1525 (1973)
P. Kaw, E. Valeo, J.M. Dawson, Interpretation of an experiment on the anomalous absorption of an
electromagnetic wave in a plasma. Phys. Rev. Lett. 25, 430–433 (1970)
G.R. Kumar, S. Sengupta, P.K. Kaw, Laser-pulse-induced second-harmonic and hard x-ray emission: role
of plasma-wave breaking. Phys. Rev. Lett. 95(2), 025005 (2005)
Lev Davidovich Landau, On the vibrations of the electronic plasma. Zh. Eksp. Teor. Fiz. 10, 25 (1946)
J.D. Lindl, P.K. Kaw, Ponderomotive force on laser-produced plasmas. Phys. Fluids (1958–1988) 14(2),
371–377 (1971)
J. Lindl, O. Landen, J. Edwards, E. Moses et al., Review of the national ignition campaign 2009–2012.
Phys. Plasmas 21(2), 020501 (2014)
J.R.S. Losada, Electrostatic plasma instabilities excited by a high-frequency electric field. Phys. Fluids
13(6), 1533–1542 (1970)
C.E. Max, Strong self-focusing due to the ponderomotive force in plasmas. Phys. Fluids (1958–1988)
19(1), 74–77 (1976)
C. Max, F. Perkins, Strong electromagnetic waves in overdense plasmas. Phys. Rev. Lett. 27, 1342–1345
(1971)
S.K. Mishra, P. Kaw, A. Das, S. Sengupta, G.R. Kumar, Stabilization of beam-weibel instability by
equilibrium density ripples. Phys. Plasmas 21(1), 012108 (2014)
G. Mourou, The ultrahigh-peak-power laser: present and future. Appl. Phys. B 65(2), 205–211 (1997)
K. Nishikawa, Parametric excitation of coupled waves i. general formulation. J. Phys. Soc. Japan 24(4),
916–922 (1968)
K. Nishikawa, Parametric excitation of coupled waves. ii. parametric plasmon-photon interaction. J. Phys.
Soc. Japan 24(5), 1152–1158 (1968)
K. Nishikawa, C.S. Liu, General formalism of parametric excitation. Adv. Plasma Phys. 1, 3–81 (1976)
F.W. Perkins, P.K. Kaw, On the role of plasma instabilities in ionospheric heating by radio waves.
J. Geophys. Res. 76(1), 282–284 (1971)
R.E. Turner, P.A. Amendt, O.L. Landen, L.J. Suter, R.J. Wallace, and B.A. Hammel. Role of laser beam
geometry in improving implosion symmetry and performance for indirect-drive inertial confinement
fusion. Phys. Plasmas, 10(6) (2003)
N.M. Rosenbluth, Parametric instabilities in inhomogeneous media. Phys. Rev. Lett. 29(9), 565 (1972)
A.S. Sandhu, A.K. Dharmadhikari, P.P. Rajeev, G.R. Kumar, S. Sengupta, A. Das, P.K. Kaw, Laser-
generated ultrashort multimegagauss magnetic pulses in plasmas. Phys. Rev. Lett. 89, 225002
(2002)
V. Saxena, A. Das, A. Sen, P. Kaw, Fluid simulation studies of the dynamical behavior of one-
dimensional relativistic electromagnetic solitons. Phys. Plasmas 13(3), 032309 (2006)
V. Saxena, A. Das, S. Sengupta, P. Kaw, A. Sen, Stability of nonlinear one-dimensional laser pulse
solitons in a plasma. Phys. Plasmas 14(7), 072307 (2007)
Y. Sentoku, K. Mima, P. Kaw, K. Nishikawa, Anomalous resistivity resulting from mev-electron
transport in overdense plasma. Phys. Rev. Lett. 90(15), 155001 (2003)
S. Subhendu Kahaly, G.Ravindra Mondal, S. Kumar, Sengupta, A. Das, P.K. Kaw, Polarimetric detection
of laser induced ultrashort magnetic pulses in overdense plasma. Phys. Plasmas 16(4), 043114
(2009)
S.K. Subhendu Kahaly, W.M. Yadav, S. Wang, Z.M. Sengupta, A. Sheng, P.K. Das, Kaw, and G
Ravindra Kumar. Near-complete absorption of intense, ultrashort laser light by sub-k gratings. Phys.
Rev. Lett. 101(14), 145001 (2008)

123
2 Page 42 of 42 Rev. Mod. Plasma Phys. (2017) 1:2

V.S. Mondal, W.J. Narayanan, A.D. Ding, B.H. Lad, S. Ahmad, W.M. Wang, Z.M. Sheng, S. Sengupta,
P. Kaw et al., Direct observation of turbulent magnetic fields in hot, dense laser produced plasmas.
Proceed. Nat. Acad. Sci. 109(21), 8011–8015 (2012)
S. Sundar, A. Das, Electron velocity shear driven instability in relativistic regime. Phys. Plasmas 17(2),
022101 (2010)
S. Sundar, A. Das, Free energy source for flow shear driven instabilities in electron-magnetohydrody-
namics. Phys. Plasmas 17(4), 042106 (2010)
S. Sundar, A. Das, V. Saxena, P. Kaw, A. Sen, Relativistic electromagnetic flat top solitons and their
stability. Phys. Plasmas 18(11), 112112 (2011)
C.H. Still, R.L. Berger, A.B. Langdon, D.E. Hinkel, L.J. Suter, E.A. Williams. Filamentation and forward
Brillouin scatter of entire smoothed and aberrated laser beams. Phys. Plasmas 7(5), 2023–2032
(2000)
T. J. Murphy, J. M. Wallace, N. D. Delamater, Cris W. Barnes, P. Gobby, A. A. Hauer, E. Lindman, G.
Magelssen, J. B. Moore, J. A. Oertel, R. Watt, O. L. Landen, P. Amendt, M. Cable, C. Decker, B.
A. Hammel, J. A. Koch, L. J. Suter, R. E. Turner, R. J. Wallace, F. J. Marshall, D. Bradley, R.
S. Craxton, R. Keck, J. P. Knauer, R. Kremens and J. D. Schnittman. and J. D. Schnittman.,
Hohlraum symmetry experiments with multiple beam cones on the omega laser facility. Phys. Rev.
Lett. 81, 108–111 (1998)
M. Tabak, J. Hammer, M.E. Glinsky, W.L. Kruer, S.C. Wilks, J. Woodworth, E.M. Campbell, M.D.
Perry, R.J. Mason. Ignition and high gain with ultrapowerful lasers. Phys. Plasmas 1(5), 1626–1634
(1994)
R.E. Turner, P. Amendt, O.L. Landen, S.G. Glendinning, P. Bell, C. Decker, B.A. Hammel, D. Kalantar,
D. Lee, R. Wallace et al., Demonstration of time-dependent symmetry control in hohlraums by
drive-beam staggering. Phys. Plasmas 7(1), 333–337 (2000)
V.L. Ginzburg, The propagation of electromagnetic waves in plasmas. International Series of
Monographs in Electromagnetic Waves, 2nd rev. and enl. edn. (Pergamon, Oxford, 1970),
pp. 260–280
VP Silin. Parametric resonance in a plasma. Zh. Eksperim. i Teor. Fiz., 48, 1965
V Ev Zakharov and AB v Shabat. Exact theory of two-dimensional self-focussing and one-dimensional
self-modulating waves in nonlinear media. Sov. Phys.-JETP (Engl. Transl.), 34, 1972
L. William, P.K. Kruer, J.M. Kaw, Dawson, C. Oberman, Anomalous high-frequency resistivity and
heating of a plasma. Phys. Rev. Lett. 24(18), 987 (1970)
T. Yabuuchi, A. Das, G.R. Kumar, H. Habara, P.K. Kaw, R. Kodama, K. Mima, P.A. Norreys, S.
Sengupta, K.A. Tanaka, Evidence of anomalous resistivity for hot electron propagation through a
dense fusion core in fast ignition experiments. New J. Phys. 11(9), 093031 (2009)
S.K. Yadav, A. Das, Nonlinear studies of fast electron current pulse propagation in a two dimensional
inhomogeneous plasma. Phys. Plasmas 17(5), 052306 (2010)
V.E. Zakharov, Collapse of Langmuir Waves. Soviet J. Exp. Theor. Phys. 35, 908 (1972)

123

You might also like