Sunteți pe pagina 1din 25

Plasma Phys. Control. Fusion 38 (1996) 769–793.

Printed in the UK

REVIEW ARTICLE

Short-pulse laser–plasma interactions

P Gibbon and E Förster


Max Planck Society, Research Unit ‘X-Ray Optics’ at the Friedrich Schiller University Jena,
Max-Wien-Platz 1, D-07743 Jena, Germany

Received 14 March 1996

Abstract. Recent theoretical and experimental research with short-pulse, high-intensity lasers
is surveyed with particular emphasis on new physical processes that occur in interactions
with low- and high-density plasmas. Basic models of femtosecond laser–solid interaction are
described including collisional absorption, transport, hydrodynamics, fast electron and hard x-ray
generation, together with recently predicted phenomena at extreme intensities, such as gigagauss
magnetic fields and induced transparency. New developments in the complementary field of
nonlinear propagation in ionized gases are reviewed, including field ionization, relativistic self-
focusing, wakefield generation and scattering instabilities. Applications in the areas of x-ray
generation for medical and biological imaging, new coherent light sources, nonlinear wave
guiding and particle acceleration are also examined.

1. Introduction

There are few technological advances which can lay claim to the number and variety of
new research fields as the arrival of the femtosecond laser. In less than a decade, ‘short-
pulse’ lasers have found applications in the physical sciences, medicine and engineering,
and dozens of new research groups have been created under the rubric of ultrafast science.
So what is so special about short-pulse lasers? First, their duration, reduced from a few
tens of picoseconds in the mid-1980s to a state-of-the-art 10 fs, allows phenomena to be
investigated which were simply too ‘fast’ for the old generation of lasers. Second, the ability
to generate coherent light pulses 1000 times shorter means that for the same energy and cost,
a laser beam can be focused to 1000 times greater intensity than was previously possible.
Thus, whereas one used to speak of ‘high intensities’ of 1014 –1015 W cm−2 , nowadays
fluxes of 1018 W cm−2 are routinely achieved with benchtop lasers, and systems capable of
reaching a dizzy 1021 W cm−2 are already under construction in several laboratories around
the world.
The electric field strengths of such lasers are orders of magnitude higher than those
binding electrons to atoms, which means that a gaseous or solid target placed at the laser
focus will undergo rapid ionization. The plasma formed in this manner will comprise
the usual fluid-like mixture of electrons and ions, but many of its basic properties will
be essentially controlled by the laser field, rather than by its own density and temperature.
Under these conditions, many of the old rules of laser–plasma interaction must be rewritten; a
fact which has prompted a number of popular articles heralding new physics and applications
(Burgess and Hutchinson 1993, Perry and Mourou 1994, Joshi and Corkum 1995). Of
course, it is not essential to focus to these extremes to do science with such lasers: there are
plenty of applications, such as high-harmonic generation in gases (L’Huillier and Balcou

0741-3335/96/060769+25$19.50
c 1996 IOP Publishing Ltd 769
770 P Gibbon and E Förster

1993), which need a large number of pump photons delivered in a short time, but not
necessarily squeezed into a small focal spot. A high-power laser system therefore has the
versatility to access completely different physics depending upon the optical arrangement
and the nature of the target. Consider, for example, the standard T3 (‘Table-Top-Terawatt’)
system now becoming standard equipment in laser and physics laboratories alike. With
a few changes to the focusing optics and target area, such a system can be used for
harmonic generation at 1015 W cm−2 , while at the same time providing a source for studying
instabilities (Darrow et al 1992) and relativistic propagation (Borisov et al l992, Monot et
al l995b) of intense beams in tenuous (underdense) plasmas at 1018 W cm−2 , and for
generating hard x-rays from solid targets (Murnane et al 1991).
This new regime of laser–matter interaction has been made possible thanks to a technique
known as chirped-pulse-amplification (CPA) (Strickland and Mourou 1985, Maine et al
1988). In CPA, a short pulse is first stretched to a nanosecond duration, amplified by a
factor > 108 , and then recompressed to its original duration. While it is not our intention
to provide a comprehensive review of state-of-the-art technology here, we list some of the
larger multi-terawatt systems in table 1.

Table 1. Multi-terawatt laser systems worldwide

Laboratory Name Type λ Power Pulse length Rep. rate


(nm) (TW) (fs) (Hz)

LLNL, USA Petawatta Ti-Sa 1053 > 1000 500–104 –


UCSD, USA –a Ti-Sa 800 10–100 10 10
LULI, FR –a Nd:glass 1053 > 100 400–104 –
Limeil, FR P102 Nd:glass 1053 55 400 10−3
RAL, UK VULCAN Nd:glass 1053 35 400–2500 –
RAL, UK TITANIAa KrF 248 > 15 150 –
ILE, JP – Nd:glass 1053 30 1000
LOA, FR LIF Ti-Sa 800 10 50 10
Saclay, FR a Ti-Sa 800 10 30 10
a under construction/upgrade

In view of the impending escalation in technology towards a new class of petawatt


lasers with pulse length of only a few optical cycles, it is perhaps an appropriate time to
take stock of the progress which has been made so far in understanding how high-intensity
short-pulse lasers interact with ionized matter, and to review the areas in which this new
interaction physics is already being put to practical use in other fields. In so doing, we shall
restrict ourselves to laser–plasma interaction and we shall not attempt to cover the related
field of atomic physics and nonlinear optics of bound electrons, on which the reader can
find excellent reviews elsewhere (L’Huillier et al 1995).
The article is organized as follows. In section 2 we first present a brief overview of the
physical processes that can play a role in short-pulse laser interactions with both high- and
low-density plasmas. The underlying theory behind the ‘laser–solid’ phenomena, including
absorption, thermal transport, hydrodynamics, generation of fast particles, hard x-rays and
magnetic fields, is then described in section 3 together with a broadly chronological review
of related experiments. In section 4 we consider interactions with underdense plasmas,
including propagation effects, relativistic focusing, parametric instabilities, novel x-ray laser
schemes and particle acceleration. In section 5, we describe a number of applications
based on the physical ideas introduced in the two previous sections, and indicate the likely
directions of future research. Our philosophy here is to try and provide a logical link between
Short-pulse laser–plasma interactions 771

the basic physics of laser–plasma interaction and the end-user who is mainly interested in
the characteristics of the photons or particles emitted and, in particular, in whether the
femtosecond laser–plasma source has advantages over conventional sources.

2. Basics of short-pulse plasma physics

Whether the target medium placed at the focus is gaseous or solid, short-pulse, high-intensity
interactions with matter generally involve a number of physical processes: ionization,
propagation and refraction, generation of plasma waves, and the subsequent thermal and
hydrodynamic evolution of the target material. The importance of any one of these
processes depends heavily upon the laser parameters, and we shall shortly see how the
evolution in laser technology towards shorter pulses and higher intensities has shifted the
research emphasis from atomic physics and linear laser–plasma wave coupling, to extremely
nonlinear collective phenomena. Likewise, on the application side, we have seen a shift
towards harder, brighter x-ray sources and production of more energetic particles. This has
presented a challenge both to theorists, who have to solve more complex equations to model
the physics, and to experimentalists, who have to devise more sophisticated diagnostics in
order to isolate or exploit these effects.

Figure 1. Physics of fs laser interaction with (a) solids and (b) gases.
772 P Gibbon and E Förster

We have attempted to summarize the main interaction physics in figure 1, which indicates
the intensity range in which various phenomena predominantly occur. Above intensities of,
say, 1016 W cm−2 , many of the effects actually depend upon the laser irradiance I λ2 , which
means that the ‘threshold’ intensity for a given phenomenon can vary depending upon the
laser wavelength. This applies, for example, to any effect which involves the quiver velocity
of an electron in the electric field of the laser,
 1/2
vos I λ2
= 0.84 .
c 1018 W cm−2 µm2
Thus, to give an electron an oscillation energy Uos = 1/2me vos
2
of 1 keV, we need a 1 µm
laser with I = 10 W cm , or a 0.248 µm KrF laser with I = 1.6 × 1017 W cm−2 .
16 −2

As with laser–plasma interactions generally, this simple scaling applies to many of


the phenomena to be considered here. Indeed, there are many instances in which the
physics overlaps with earlier work using nanosecond glass and CO2 lasers. The pulse
duration, though not necessarily a prerequisite for the study or existence of the effects
depicted in figure 1, is of central importance for applications that exploit the ‘ultrafast’
aspect of the physics. For example, an early motivation for studying femtosecond laser–
solid interactions was the widespread interest in hard x-ray sources for ‘real-time’ probing
of chemical reactions that take place on a sub-picosecond timescale (Murnane et al 1991).

3. Femtosecond laser–solid interactions

The question of what happens to a solid target when it is subjected to irradiation by a


short-pulse laser is a simple one to pose, but actually quite complicated to answer. More
than one physical picture is possible depending on whether the material is treated as a dense
conductor, or a ‘sandwich’ of cold solid plus a hot, thin layer of plasma in the region of
the laser’s focal spot. This division still persists in the modelling and interpretation of
contemporary experiments at high intensities. To date, there is no single model which can
adequately describe all the main pieces of interaction physics.
In order to see why, it is initially helpful to look back at the early ideas put forward
at the end of the 1980s in anticipation of the outcome of such experiments. Although
the subject of laser–plasma interaction has been established in the context of laser fusion
for at least 25 years prior to the arrival of short pulses, a number of authors (Gamaliy and
Tikhonchuk 1988, More et al 1988a, Mulser et al 1989, Milchberg and Freeman 1989) were
quick to point out that much of the traditional physics would not apply to sub-picosecond
interactions for essentially two reasons. First, the short pulse duration means that there is
not enough time for a substantial region of ‘coronal’ plasma to form in front of the target
during the interaction. Second, owing to the steep density gradient, laser energy can be
deposited at much higher densities than in nanosecond interactions, where it is absorbed at
or below the critical density Nc . This is the density at which the plasma becomes overdense
for an electromagnetic wave with frequency ω0 , and is defined by ω02 = 4π eNc /me , where
e and me are the electron charge and mass, respectively.

3.1. Collisional heating


Not surprisingly, the first theoretical works appearing at the end of the 1980s concentrated
on the issue of laser energy absorption in step-like density profiles. More et al (1988a)
proposed a simple model for normal incidence interactions, in which the laser energy is
absorbed in a small skin layer of depth δ at the target surface, see figure 2.
Short-pulse laser–plasma interactions 773

Figure 2. Schematic of short-pulse heating of solid-density plasma.

In this picture, the plasma is assumed to be ionized to some degree and can be treated
as a conductor with a permittivity:
ωp2
ε =1− (1)
ω0 (ω0 + iνei )
where νei is the electron–ion collision frequency, given by:
νei ' 3 × 10−6 ZNe Te−3/2 log 3s −1
where Z is the number of free electrons, Ne is the electron density in cm−3 , Te is the
temperature in eV and log 3 is the Coulomb logarithm. By matching the electric and
magnetic fields at the vacuum–plasma boundary, one can obtain the absorption coefficient
for highly overdense plasmas (ωp /ω0  1) in two limits (More et al 1988a, Gamaliy 1994):
 2ν


ei
νei  ω0
 ω
p
A= (2)

 2ω
 0 (νei /ωp )1/2 νei > ω0 .
ωp
This result can be generalized to arbitrary angles of incidence and polarization (s or p)
using the Fresnel equations (Born and Wolf 1980). The absorbed energy is then used to
determine the temperature from the equation of state for the plasma. This can be calculated
using the ideal gas law at low densities, or the Thomas–Fermi statistical model at high
densities (Pfalzner 1991). Further refinements can be made by including non-equilibrium
and low-temperature effects (More et al 1988b). The Thomas–Fermi model is also often
used to determine the number of free electrons, or ionization degree, but for very short
pulses it is more appropriate to solve the atomic populations explicitly (Edwards and Rose
1993). Although one can include most collisional and radiative effects this way, a fully self-
consistent ionization model, including field-ionization (Pfalzner 1992) and non-Maxwellian
effects still remains a challenge.
Given the temperature and ionization state, the next step is to determine how the energy
is transported into the target. More et al (1988a) did this by solving the diffusion equation
with conductivities based on the usual Spitzer–Härm heat flow (Zel’dovich and Raizer 1966)
but modified for high densities (Lee and More 1984). This can be made more sophisticated
by using a Fokker–Planck treatment of collisions which explicitly takes into account non-
local transport (Rozmus and Tikhonchuk 1990). With either approach, a self-similar solution
can be found which describes the heat-front penetration into the target and sets an upper
limit on the pulse duration above which hydrodynamics becomes significant.
774 P Gibbon and E Förster

3.2. Hydrodynamic models


Until now we have only considered the idealized step-profile target. In reality, the plasma
will have a finite gradient, either because the pulse length is longer than the thermal
expansion time (which decreases with increasing intensity), or because a prepulse is present
which is sufficiently long and intense to ionize and ablate material from the target surface
prior to the arrival of the main pulse. At low intensities (e.g. I < 1014 W cm−2 ), this is
not such a problem even with contrast ratios of 104 –105 , because the prepulse will remain
below the threshold for plasma formation. On the other hand, typical pulse lengths in
early experiments were around 1 ps, so some thermal expansion was inevitable (Milchberg
and Freeman 1989, Fedosejevs et al 1990a). In this case, the Fresnel equations are no
longer valid and, to make quantitative comparison with experiments, it becomes necessary
to calculate the absorption in a finite density profile.
This can be done by solving the Helmholtz equations for the electromagnetic wave
numerically (Milchberg and Freeman 1989, Kieffer et al 1989a, Fedosejevs 1990b). As
before, the absorption model can be coupled to a set of coupled equations for the heat flow
and ionization, but for self-consistency it is better to solve the hydrodynamics as well.

Figure 3. Modelling of laser–solid interaction.

These ingredients, ionization, collisions, wave propagation, thermal transport and


hydrodynamics, form the basis for many standard ‘short-pulse’ simulation codes now in
common use (Ng et al 1994, Davis et al 1995, Teubner et al 1996a). More sophisticated
codes also exist which solve the Fokker–Planck (FP) equation to include non-local heat
flow (Town et al 1994, 1995, Matte et al 1994, Limpouch et al 1994). The self-consistent
treatment of ionization dynamics and FP heat flow leads to a strongly non-Maxwellian
electron distribution function, which modifies the heat-flow penetration into the target in
Short-pulse laser–plasma interactions 775

comparison with the classical Spitzer–Härm theory of thermal transport. These workhorse
models are depicted schematically in figure 3.

3.3. Short-pulse x-rays


One of the main motivations behind the development of femtosecond laser–plasma models
is to understand and optimize the generation of short x-ray pulses (Kühlke et al 1987,
Stearns et al 1988, Murnane et al 1989, Teubner et al 1995). In particular, one would like
to be able to predict the yield and duration of x-rays emitted from the plasma, which depend
upon the interplay of a number of effects: absorption, heating, ionization, recombination
and transport. By treating the plasma as a high-density, blackbody radiator, Rosen (1990)
and Milchberg et al (1991) demonstrated the possibility of generating sub-picosecond x-rays
via rapid cooling of the target. More sophisticated treatments including hydrodynamics and
detailed atomic modelling have been able to determine the pulse duration of particular lines
(Audebert et al 1994). Experimental measurements of x-ray pulse duration are currently
instrument-limited to around 1 ps (Kieffer et al 1993, Shepherd et al 1994) and new
techniques will have to be developed, for example, using atomic excitation and ionization
(Barty et al 1995), to measure pulses in the sub-100 fs regime.

3.4. Collisionless absorption


Early experiments performed at modest intensities (Milchberg et al 1988, Kieffer et al
1989b, Murnane et al 1989, Landen et al 1989, Fedosejevs et al 1990a) were in good
agreement with hydrodynamic models, in terms of both the measured reflectivity, and the
typical plasma parameters inferred from atomic x-ray spectra. However, a problem soon
became apparent as lasers were upgraded and intensities increased. First, for intensities
above 1015 W cm−2 or so, the plasma temperature rises so quickly that collisions become
ineffective during the interaction: Te /eV ∼ I 1/2 ω0 t; νei ∼ I −3/4 t −3/2 (Rozmus and
Tikhonchuk 1990). Second, the electron quiver velocity is comparable to the thermal
velocity, thus reducing the effective collision frequency further (Pert 1995):
vte3
νeff ' νei 2 + v 2 )3/2
. (3)
(vos te

In other words, collisional absorption starts to turn off for irradiances I λ2 >
1015 W cm−2 µm2 , and therefore could not account for the high absorption observed,
for example, by Chaker et al (1991) and Meyerhofer et al (1993).
Given this discrepancy, alternative absorption mechanisms were sought which did not
rely on collisions between electrons and ions. In fact, there are a number of collisionless
processes which can couple laser energy to the plasma. The best known and most studied
of these is resonance absorption, although it is not immediately clear how effective this is in
steep density gradients. In the standard picture (Ginzburg 1964), a p-polarized light wave
tunnels through to the critical surface (Ne = Nc ), where it excites a plasma wave. This
wave grows over a number of laser periods and is eventually damped either by collisions at
low intensities or by particle trapping and wave breaking at high intensities (Kruer 1988).
For long density scale lengths, defined by k0 L  1, where k0 = 2π/λ is the laser wave
vector and L−1 ≡ |d(log Ne )/dx|x=xc , the absorption rate has a self-similar dependence upon
the parameter (k0 L)2/3 sin2 θ (Ginzburg 1964). This behaviour is more-or-less independent
of the damping mechanism provided the pump amplitude is small, a condition which we
shall quantify shortly. ‘Collisionless’ resonance absorption can therefore be modelled in a
776 P Gibbon and E Förster

hydro-code even if the collision rate is too small to give significant inverse-bremsstrahlung
(Rae and Burnett 1991). This is usually done by introducing a phenomenological collision
frequency in the vicinity of the critical surface (Forslund et al 1975a) such that one recovers
the ∼ 50% optimum absorption rate in the long scale-length limit.
Although the overall energy balance will be taken care of in this manner, the way in
which it is divided into thermal and suprathermal electron heating can only be determined
self-consistently using a kinetic approach such as particle-in-cell (PIC) (Birdsall and
Langdon 1985) or Vlasov simulation.
Resonance absorption was studied extensively in the 1970s and 1980s with two-
dimensional PIC codes in order to understand the origin of fast electrons generated in
nanosecond laser–plasma interactions (Estabrook et al 1975, Forslund et al 1977, Estabrook
and Kruer 1978, Adam and Héron 1988, Kruer 1988). Ironically, fast electron generation
is highly undesirable in the ICF context because it leads to preheating of the fuel, thus
preventing targets from being compressed to the necessary densities for high thermonuclear
gain. With the advent of short pulses, however, fast electrons are very much back in fashion
because they generate hard x-rays as they travel through the cold part of the target behind
the hot plasma where they are generated.
As hinted at earlier, resonance absorption ceases to work in its usual form in very
steep density gradients. To see this, consider a resonantly driven plasma wave at the critical
density with a field amplitude Ep . In a sharp-edged profile, there will be little field swelling,
and Ep will be roughly the same as the incident laser field E0 . Electrons will therefore
undergo oscillations along the density gradient with an amplitude
Xp ' eE0 /me ω02 = vos /ω0 .
The resonance breaks down if this amplitude exceeds the density scale length L, i.e. if
vos /ω0 L > 1.
Under these conditions, it is no longer useful to speak of electrons being heated by
a plasma wave, since this wave is destroyed and rebuilt afresh each cycle. This simple
fact was pointed out by Brunel (1987), who proposed an alternative mechanism in which
electrons are directly heated by the p-polarized component of the laser field. According
to Brunel’s model, electrons are dragged away from the target surface, turned around and
accelerated back into the solid all within half a laser cycle. These electrons are simply
absorbed because the field only penetrates to a skin depth or so. Assuming electrons gain a
velocity vos during their vacuum orbit, the absorption fraction can be estimated as (Bonnaud
et al 1991)
 vos

 3f (θ) A1
c
A= (4)

 3 f (θ) vos A∼1
8 c
where f (θ ) = sin3 θ/ cos θ.
The hot electron temperature inferred from this model and from electrostatic PIC
simulations is just
Thot ∝ vos
2
' 3.6I16 λ2µ (5)

where I16 is the intensity in 1016 W cm−2 and λµ the wavelength in microns.
While this capacitor model predicts both high absorption and a strong hot temperature
scaling, subsequent studies with electromagnetic codes showed that the mechanism saturates
at high intensity due to deflection of the electron orbits by the v ∧ B force (Estabrook and
Short-pulse laser–plasma interactions 777

Kruer 1986, Brunel 1988). On the other hand, the saturation effect can be partially overcome
by using two incident beams at ±45◦ .
Just as for collisional absorption, the situation becomes more complicated for realistic
profiles with finite density gradients. This case was considered by Gibbon and Bell
(1992), who found a highly complex transition between resonance absorption and vacuum
heating depending upon the irradiance and scale length. For high irradiances and short
scale lengths, the absorption saturates at around 10–15%, but for intermediate values (e.g.
I λ2 = 1016 W cm−2 µm2 , L/λ ∼ 0.1), the absorption can be as high as 70%. These
results are in good agreement with a more recent analytical study of absorption in short
scale-length profiles (Andreev et al 1994). The hot electron temperature scales according
to
1/3
Thot ' 8 I16 λ2N keV (6)
which is somewhat lower than the scaling from simulations of resonance absorption in
steepened density profiles (Forslund et al 1977, Estabrook and Kruer 1978, Kruer 1988).
In simulations with mobile ions, however, a rather different picture emerges. The strong
space charge created by electrons circulating outside the surface pulls out an underdense
ion shelf which can drastically√alter the absorption (Brunel 1988, Gibbon 1994). After
a characteristic time ts ' 100 A/Zλµ fs, the absorption and hot electron distribution
resemble those seen in early simulations; the main difference being that the pressure
balance assumed and imposed in long-pulse (ns) interactions (Forslund et al 1977) is
unlikely to be achieved for femtosecond pulses. This lack of hydrodynamic equilibrium can
strongly influence the hot electron fraction and temperature scaling. For extreme intensities
(I > 1020 W cm−2 µm2 ) at normal incidence, energy is transferred directly to the ions
through the formation of a collisionless shock (Denavit 1992).
A third collisionless mechanism which is closely related to vacuum heating is the
anomalous skin effect. This is actually a well known effect in solid-state physics (Ziman
1969), and was originally studied for step-like vacuum–plasma interfaces by Weibel (1967).
Its potential significance for short-pulse interaction was therefore recognized quite early
by a number of workers (Gamaliy and Tikhonchuk 1988, Mulser et al 1989, Rozmus and
Tikhonchuk 1990, Andreev et al 1992).
Physically, the anomalous skin effect is not as mysterious as it sounds. Consider first
the situation for the normal skin effect. Electrons within the skin layer δ = c/ωp oscillate
in the laser field and dissipate energy through collisions with ions. The oscillation energy
is thus locally thermalized provided that the electron mean-free path la = vte /νei is smaller
than the skin depth. Now imagine that the temperature is increased so that la > δ, and that
the mean thermal excursion length vte /ω0 > δ. Under these conditions, the laser field is
carried further into the plasma and the effective collision frequency is given by the excursion
time in the anomalous skin layer δa , i.e. νeff = vte /δa , where δa = (c2 vte /ω0 ωp2 )1/3 (Weibel
1967). For normal incidence in the overdense limit, one finds (Rozmus and Tikhonchuk
1990, Andreev et al 1992):
 1/3  1/6  1/3
ω0 vte ω02 Te Nc
A' δa = ' . (7)
c cωp2 511 keV Ne
Self-consistent theories of the anomalous skin effect solve the full Vlasov equation in
order to correctly describe the non-local relationship between the current and the electric
field (Weibel 1967, Mulser et al 1989, Gamaliy 1994, Matte and Aguenaou 1992, Andreev
et al 1992). None the less, just as with resonance absorption, the effect can also be included
in a hydrodynamic model by replacing νei with an effective collision frequency.
778 P Gibbon and E Förster

The Fresnel equations can again be used to obtain the angular absorption dependence in
the step-profile limit (Andreev et al 1992). The maximum absorption for p-polarized light
is nominally A ∼ 2/3 at grazing incidence angles, independent of density and temperature,
but can be enhanced if the distribution function is anisotropic. A more complete study of
anomalous skin absorption including relativistic effects has been made by Ruhl and Mulser
(1995).

3.5. Ultrahigh intensities

As we saw earlier, ion motion can alter the electron dynamics by changing the density
profile near the critical surface. As long as this motion remains normal to the gradient, the
absorption and hydrodynamics can still be modelled in one dimension, even for obliquely
incident light. This picture becomes inadequate if a hole is formed, or if the surface develops
ripples. Both of these situations can occur for finite focal spot sizes (which are typically
diffraction-limited to 2–10 µm) and at extreme irradiances (I λ2 > 1018 W cm−2 µm2 ).
This regime was studied by Wilks et al (l992) using two-dimensional PIC simulation. They
found that tightly focused, normally incident light can bore a hole several wavelengths deep
through moderately overdense plasma on the sub-picosecond timescale.
‘Hole boring’ results from a combination of three effects. First, the light pressure, PL =
2I /c ' 600I18 Mbar, vastly exceeds the thermal plasma pressure, Pe = 160N23 Te Mbar
(where N23 is the density in 1023 cm−3 and I18 is the intensity in 1018 W cm−2 ). This will
cause the plasma to be pushed inwards preferentially at the centre of the focal spot. Second,
a radial ponderomotive force due to the transverse intensity gradient ∇r I pushes electrons
away from the centre of the beam, creating a charge separation which pulls the ions out.
Third, the skin depth is enhanced where the laser intensity is greatest due to relativistic
decrease in the effective plasma frequency: ωp0 = ωp /γ , where γ = (1 + vos 2
/2c2 )1/2 .
At sufficiently extreme intensities this could lead to ‘induced transparency’, where the
laser beam is transmitted through a nominally overdense plasma instead of being reflected
(Lefebvre and Bonnaud 1995). As a hole is formed, the absorption and hot electron
temperature both increase because density gradients are formed parallel to the laser electric
field (Wilks 1993).
Another two-dimensional effect is magnetic field generation, which has long been a
subject of fascination in the field of laser–plasma interactions. Of particular interest are the
large DC fields which can arise from electron transport around the focal spot. In short-pulse
interactions, there are at least three mechanisms which can generate B-fields:

(i) radial transport where the electron temperature and density gradients are not parallel
(Stamper et al 1971), giving a source term ∂B/∂t ∝ ∇Ne ∧ ∇Te ;
(ii) DC currents in steep density gradients driven by temporal variations in the
ponderomotive force (Sudan 1993, Wilks et al 1992);
(iii) hot electron surface currents (Brunel 1988, Gibbon 1994, Ruhl and Mulser 1995).

The first of these mechanisms, which occurs on the hydrodynamic timescale, persists long
after the laser pulse and can cause strong pinching of the ablated plasma (Bell et al 1993). In
contrast, the other two mechanisms will occur predominantly at early times (10 fs < t 6 τp )
in interactions at normal and oblique incidence, respectively. For intensities of 1019 W cm−2 ,
the magnitude of the B-field can be 109 G and above.
Short-pulse laser–plasma interactions 779

3.6. Hot electrons and hard x-ray generation


There are essentially three signatures of high-intensity collective effects in laser–plasma
interactions: high angular-dependent absorption, hard x-rays and fast ions. All three of
these have since been verified experimentally for sub-picosecond pulses. Absorption of
50–60% for p-polarized light has been found for intensities of 1016 and above (Kieffer et
al 1989b, Klem et al 1993, Meyerhofer et al 1993, Teubner et al 1993, Sauerbrey et al
1994), x-rays in the keV–MeV range have been measured (Audebert et al 1992, Kmetec
et al 1992, Klem et al 1993, Chen et al 1993, Rousse et al 1994, Schnürer et al 1995,
Teubner et al 1996b), and fast ion blow-off has been seen (Meyerhofer et al 1993, Fews et
al 1994).
Unlike their softer cousins (see section 3.3), hard x-rays are a direct result of hot
electrons produced in the vicinity of the focal spot. By virtue its long mean-free path, a
fast electron can penetrate into the cold region of the target beyond the heat front, where it
either emits bremsstrahlung via collisions with ions, or produces line radiation by knocking
out a bound K-shell electron.

Figure 4. Hot electron temperature measurements in short-pulse experiments (squares) compared


with PIC simulations (filled circles). Experimental data are taken from: Meyerhofer et al (1993)–
LLE; Rousse et al (1994)–LULI; Teubner et al (1996b)–Jena/Sprite; Jiang et al (1995)–INRS;
Schnürer et al (1995)–MBI; Kmetec et al (1992)–Stanford; Fews et al (1994)–LLNL/IC/RAL.

Bremsstrahlung appears as a continuum anywhere in the 0.1 keV–MeV range depending


on the laser intensity and plasma parameters (Kmetec et al 1992, Schnürer et al 1995),
whereas inner-shell line emission can be 1–100 keV depending upon the atomic number
of the target material (Soom et al 1993, Rousse et al 1994, Jiang et al 1995). ‘Cold’
line radiation is presently the more interesting of these for applications because of its
typically narrow bandwidth and high potential brightness. Kα radiation is also an important
diagnostic tool. The spectral intensity of photons emitted is directly dependent upon the
electron energy. This can be exploited by observing that electrons slow down and eventually
780 P Gibbon and E Förster

stop in cold material, so that the total line emission will in general depend upon the target
thickness. The hot electron energy can therefore be inferred from a ‘sandwich’ experiment
(e.g. Al on Si) in which the thickness of the front layer is varied (Hares et al 1979). The
change in the Kα line ratios with thickness can then be fitted to a characteristic temperature
or distribution function.
Although experimental hot temperature measurements in the sub-picosecond regime are
still scarce, a picture is gradually emerging which suggests that
(i) temperatures are lower than for long-pulse experiments at the same intensities,,
(ii) the scaling law is Thot ' (I λ2 )1/3−1/2 .
A summary of these experiments is shown in figure 4 together with results from recent
(one-dimensional) PIC simulations.

3.7. High-harmonic generation


Thanks to short-pulse technology, the subject of harmonic generation by ultra-intense laser–
solid interaction has also enjoyed a resurgence of interest as a means of producing short-
wavelength, coherent light (Kohlweyer et al 1995, Von der Linde et al 1995). Physically,
harmonics are generated by the highly nonlinear plasma oscillations at the surface of the
target as described in section 3.4. This effect was first demonstrated with long-pulse CO2
lasers by Carman and co-workers at Los Alamos in the early 1980s (Carman et al 1981a,
b) and was suggested as a means of inferring the plasma density from the highest harmonic
observed.
More recent work by Gibbon (1996) has demonstrated that for intensities I >
1018 W cm−2 , the harmonic ‘cutoff’
√ predicted by earlier theories (Bezzerides et al 1982,
Grebogi et al 1983), ωn = Ne /Nc ω0 no longer appears in the reflected spectrum.
Theoretically, over 60 harmonics can be generated with efficiencies > 10−6 for modest
plasma densities Ne /Nc ∼ 10 − 30. This prediction has been verified in an experiment with
the Vulcan CPA system at RAL, UK, in which over 70 harmonics were observed using
2.5 ps pulses at intensities up to 1019 W cm−2 (Norreys et al 1996). In terms of raw
power and scalability with wavelength and intensity, this mechanism is hard to beat: the
efficiency essentially depends upon I λ2 , whereas the minimum wavelength is just λ/nmax .
For example, Norreys et al estimate that the power converted into the 38th harmonic at
28 nm was 24 MW. On the other hand, it was also found that the harmonics were emitted
more-or-less isotropically, an effect which was attributed to surface dimpling. This is in
contrast to the near-specular emission in the preceding experiments using 100 fs pulses
(Kohlweyer et al 1995, Von der Linde et al 1995), and it will be a challenge to optimize
the coherence of harmonic high-order emission for these ultrashort-pulse durations.

4. Laser interaction with low-density (tenuous) plasmas

Focusing a high-intensity laser pulse into a gas elicits a completely different character
of interaction from that considered in section 3. With solids, the pulse interacts with a
few microns of essentially mirror-like material; with gas targets, the pulse propagates over
millimetres, during which time it can ionize, distort, refract and accelerate particles, and so
on. The interaction physics is therefore determined not only by the spot size σ0 and intensity
I at the focus, but also by the focusing geometry or the Rayleigh diffraction length:
πσ02
ZR = (8)
λ
Short-pulse laser–plasma interactions 781

where σ0 is the 1/e2 spot size, defined for a Gaussian beam by: I (r) = I0 exp(−r 2 /σ02 ). In
this section we describe some of these new effects in nonlinear optics and how they might
be usefully exploited.

4.1. Multiphoton ionization and heating


One of the first obstacles to studying laser–plasma interactions at high intensities is the
production of a plasma with accurately known properties. The simplest way to do this is to
fill a target chamber with a gas at atmospheric pressure and ionize it by focusing the laser
inside the chamber. Rapid ionization occurs by virtue of the fact that the laser field perturbs
the Coulomb barrier of the atom, allowing electrons to tunnel free. For hydrogen-like ions,
this process can be modelled according to a theory by Keldysh (1965), who derived a
DC ionization probability:
 5/2 "   #
Ei Ea 2 Ei 3/2 Ea
W = 4ωa exp − (9)
Eh EL 3 Eh EL
where Ei and Eh , are the ionization potentials of the atom and hydrogen respectively,
Ea = m2 e5 / h4 is the atomic electric field, EL is the laser field, and ωa is the atomic
frequency (4 × 1016 s−1 ). This formula is particularly useful in the regime e2 EL2 /4Mω02 
Ei  hν, which is precisely the operating regime of femtosecond laser systems, where
the photoelectric effect (that is, single-photon ionization) cannot account for the observed
ionization rates. In fact (9) and its generalizations for oscillating fields and more complex
atoms (Ammosov et al 1986) have proved remarkably effective in describing the ionization
physics of femtosecond interactions. This theory of so-called ‘above-threshold-ionization’
(ATI) has been verified experimentally by measuring the energies of emitted electrons
(Freeman et al 1987, Augst et al 1989) and multiply-charged ions (Auguste et al 1992b).
An important implication of ATI is that the kinetic energy of electrons pulled out from
atoms by the laser field is typically smaller than both the ionization potential and the quiver
energy (Burnett and Corkum 1987). This low ‘residual’ energy arises because within an
optical cycle, ionization occurs at the peak of the electric field where the quiver velocity is
zero. On the other hand, any departure from linear polarization will increase the random
energy acquired by ‘newly-born’ electrons, and ultimately increase the final temperature of
the plasma created.
This issue is of central importance to novel x-ray laser schemes using short pulses
(Amendt et al 1991), which rely on the rapid creation of a highly ionized, cold plasma
which subsequently recombines and lases back to the ground state. If the plasma is too hot,
the scheme will saturate and the x-ray efficiency will be low. A number of theoretical
(Penetrante and Bardsley 1991, Pert 1995) and experimental (Offenberger et al 1993,
Dunne et al 1994, Blyth et al 1995) efforts have therefore concentrated on determining
and minimizing the plasma temperature. In particular, inverse-bremsstrahlung heating
and collective heating due to parametric instabilities (see section 4.3) could impair the
effectiveness of such schemes. Consequently, short wavelengths (λ = 1/4 µm) and pulse
lengths shorter than 100 fs should improve the chances of success for this type of scheme.

4.2. Nonlinear refraction


The subject of nonlinear propagation of electromagnetic (EM) waves in plasmas is too vast
to do justice to in this review. Interest in this field predates the use of short pulses by several
decades, in the context of ICF, ionospheric physics and astrophysics. Nevertheless, there
782 P Gibbon and E Förster

were again a number of works appearing at the end of the 1980s pointing out the ways in
which femtosecond pulses should behave differently from longer (pico–nanosecond) pulses
in underdense plasmas.
At first sight, one might think that for laser intensities such that EL  Ei , a plasma
would be instantly created by the leading foot of the pulse, leaving the main part propagating
through a fully ionized, uniform plasma. Unfortunately, the situation is complicated by a
phenomenon known as ionization-induced defocusing (Auguste et al 1992a, Leemans et
al 1992, Rae 1993). Near the front of an intense pulse, where the field is close to the
ionization threshold, the gas at the centre of the beam will be ionized more, giving rise to
a steep radial density gradient. The refractive index of the plasma, given by:
 
Ne (r, z, t) 1/2
η(r, z, t) = 1 − (10)
Nc
will therefore have a minimum on axis and act as a defocusing lens for the rear portion of
the beam. The result is that for high gas pressures, the laser beam is deflected well before
it can reach its nominal focus (Auguste et al 1992a).
To circumvent this problem, experiments requiring high intensities are usually performed
either with a preformed plasma (Durfee III and Milchberg 1993, Mackinnon et al 1995), or
using a gas-jet configuration in which the beam is focused in vacuum before it actually enters
the gas (Auguste et al 1994). Interaction at the maximum intensity is then guaranteed and
one can essentially neglect the ionization physics. It is then possible to study the interaction
of the laser fields with free electrons at intensities of 1018 W cm−2 and above. To see what
new effects can be expected in this regime, it is helpful to examine the wave equation for
an electromagnetic wave in a plasma:
 2 
∂ Ne A
− c ∇ A = JNL =
2 2
. (11)
∂t 2 γ
The nonlinear current JNL on the right-hand side of (11) contains both the coupling of
the laser field to the plasma and high-intensity effects such as relativistic self-focusing. The
latter effect arises due to a change in the refractive index via electrons quivering in the laser
field at velocities close to the speed of light. This phenomenon has been known for some
time (Litvak 1968, Max et al 1974), but it is only through short-pulse technology that it
has become possible to study it experimentally. The reason for this is that there is a power
threshold (Sprangle et al 1988):
Ne
Pc = 17 GW (12)
Nc
at which beam diffraction is balanced by self-focusing. For typical electron densities
available from a gas jet, namely 1018 –1020 cm−3 , one needs a multi-terawatt laser to have
a reasonable chance of seeing the effect. On the other hand, theoretical and computational
studies have demonstrated that self-focusing should be accompanied by partial or complete
expulsion of electrons from the beam centre (Sun et al 1987, Mori et al 1988, Borisov et
al 1990, Chen and Sudan l993, Pukhov and Meyer-ter-Vehn 1996), forming a kind of self-
sustained optical fibre. Recent experiments have reported evidence of extended propagation
over several Rayleigh lengths (Borisov et al 1992, Sullivan et al 1994, Monot et al 1995b,
Mackinnon et al 1995). However, the interpretation of these results has been complicated
by the fact that the diagnostics used to image the focused beam rely on scattering of the
laser light from plasma electrons (Gibbon et al 1995).
Short-pulse laser–plasma interactions 783

4.3. Wakefield excitation and instabilities


Another important effect which modifies beam propagation is the excitation of plasma
waves. These can be generated either by Raman-type instabilities (Kruer 1988) or by an
appropriate choice of parameters so that a ‘wake’ is produced behind the pulse. In both
cases, large-amplitude electrostatic fields can be generated which are able to accelerate
electrons to very high energies over short distances (< 1 m), potentially to GeV levels.
Not surprisingly, therefore, much of the motivation for studying short-pulse interactions has
come from the ‘plasma accelerator’ community, and we review some of the issues involved
in this application later in section 5.5.
On the physics side, a topic which has attracted some controversy is the interplay of
relativistic focusing and plasma wave generation. Several groups have tackled this problem
by solving the two-dimensional envelope equations for the laser beam together with the
nonlinear fluid response of the plasma (Sprangle et al 1990, 1992, Andreev et al 1992,
Abramyan et al 1992, Antonsen and Mora 1992, Krall et al 1994). Simulations using this
approach typically predict that for pulse lengths a few times longer than the plasma period,
the envelope breaks up into beamlets of length λp (Andreev et al 1992, Sprangle et al 1992).
This effect, known as ‘self-modulation’ was interpreted by these authors as follows: owing
to the finite pulse shape, a small wake plasma wave is excited non-resonantly, which results
in an oscillating density perturbation within the pulse envelope. The EM waves therefore
see a refractive index which is alternately peaked and dented at intervals of λp /2. These
portions of the pulse will therefore focus and diffract, respectively, leading to modulations
in the envelope with period λp . These modulations subsequently enhance the plasma wake.
The net result is that a large amount of energy can be scattered outside the original
focal cone (Antonsen and Mora 1992), leaving a large-amplitude plasma wave close to the
beam axis. On the other hand, one is tempted to conclude that the nominal requirement
for wakefield excitation can be relaxed, and one can take pulse lengths τp  ωp−1 .
Unfortunately, this picture is not complete: envelope models neglect a number of important
physical processes which turn out to be just as important. First, parametric instabilities
(Drake et al 1974, Forslund et al 1975b) such as Raman backscatter (RBS) and Raman
forward scatter (RFS) are explicitly excluded by the paraxial-ray approximation used in
these models. The growth rate for RBS with a relativistic pump is (Sakharov and Kirsanov
1994, Guérin et al 1995):
√  1/3  2/3
3 ω0 ωp2 a0
0B = (13)
2 2 γ0
whereas for RFS, we have (Mori et al 1994):
ωp2 a0
0F = √ . (14)
8ω0 (1 + a02 )1/2
In RBS, the pump wave decays into a plasma wave plus an EM wave travelling back
towards the focal lens. The instability therefore grows from the foot of the pulse towards
the rear, and the number of e-foldings depends upon the pulse length. For RFS, however,
the instability growth depends upon the propagation length as well, and is thus potentially
more damaging for applications using short, high-intensity pulses.
The second important effect excluded from all fluid models, whether paraxial or not, is
wave breaking (Koch and Albritton 1974), which causes the plasma wave to lose coherence
and heat electrons. As in laser–solid interactions, a kinetic model is essential to treat this
process self-consistently. Several groups have already presented PIC simulations which
784 P Gibbon and E Förster

largely confirm the initial growth scaling of the RFS and RBS instabilities but which also
follow them to saturation (Bulanov et al 1992, Decker et al 1994). The state-of-the-
art in this area is currently claimed by the UCLA/LLNL groups, who have been able to
model actual experimental parameters with two-dimensional PIC simulations comprising
over 107 particles (Tzeng et al 1996). These studies have shown that RFS can also
excite large-amplitude plasma waves and induce modulations in the pulse in a manner
indistinguishable from the self-modulational instability observed with fluid models.
Experiments at Livermore and RAL, UK largely corroborate these findings, but the
overall understanding of propagation effects is far from complete. In an experiment using
a 600 fs pulse and a 1 mm helium gas jet, Coverdale et al (1995) demonstrated that up to
50% of the light is scattered out of the focal cone, a result that is apparently at odds with an
experiment performed under very similar conditions at Saclay (France), where collimation
of the exiting beam was observed (Monot et al 1995a).

5. Applications of short-pulse laser-plasma sources

So far in this review, we have concentrated mainly on the basic physical issues of
femtosecond laser–plasma interactions. While it is true that much current research is
curiosity driven, an equally important motivating factor is the extent to which laser-plasmas
can be used as primary sources of photons and electrons for other purposes. From the
preceding sections, it should by now be clear that there are basically three main areas of
application:

(i) generation of hard and soft incoherent x-rays,


(ii) coherent short-wavelength light sources,
(iii) particle acceleration.

In this section we consider some specific applications in a little more detail, where possible
comparing laser-plasma sources with more traditional ones. As far as the x-ray sources
are concerned, a vital component of any successful application will be the development of
suitable optics for the x-ray photons, which in many cases is a technological challenge in
itself (Förster et al 1992, 1994, Attwood 1992).

5.1. Medical imaging


The possibility of creating ultrafast x-ray flashlamps has been one of the major motivations
behind short-pulse technological developments (Murnane et al 1991). As we saw in
sections 3.3 and 3.6, a laser-produced plasma can be crudely regarded as a polychromatic
continuum source with peaks of line radiation characteristic of the target material. The
latter component is generally regarded as the more interesting because of its well defined
emission wavelength, narrow bandwidth and high intensity.
An application which is presently under serious evaluation by several groups is medical
and biological imaging. Since their discovery a century ago (Röntgen 1895), x-rays have
been exploited for this purpose with increasing sophistication, so it is natural to ask what
improvements can be offered by laser-plasma sources. The requirements of medical imaging
are essentially threefold: first, x-ray photon energies need to be 20–100 keV to allow
transmission through the body; second, the bandwidth should be narrow to minimize the
dose from unwanted soft x-rays; third, a high degree of tunability is needed to distinguish
between different types of tissue. Although the traditional x-ray tube meets these general
Short-pulse laser–plasma interactions 785

specifications, a laser-plasma source may offer some additional advantages (Herrlin et al


1993):
• First, a short-pulse x-ray source delivers a high yield in a picosecond duration, as
opposed to micro–milliseconds for conventional sources. This offers the possibility of
substantial dose reduction by using a time-gating technique to eliminate scattered x-rays
which degrade the image contrast (Gordon III et al 1995).
• Second, differential imaging with rapid, simultaneous exposure is possible (Tillman et
al 1996). This technique, traditionally implemented using large synchrotron sources,
requires rapid exposure by two x-ray lines with photon energies above and below
the K-absorption edge of a ‘contrast agent’. Subtraction of the two resulting images
enhances the parts of the sample containing the contrast agent and suppresses unwanted
information.
• Third, the small target size makes it possible to conceive novel image projections where
the x-ray source is placed inside the object of investigation (Tillman et al 1995).
While preliminary experiments by the Lund group are very encouraging, a number of
issues remain to be clarified, particularly concerning dosage. For instance, although the
overall x-ray dose may be reduced by using laser-plasma sources, it is not yet clear what
kind of damage the higher intensities may cause to the molecular structure of living tissue.
Furthermore, the optimization of these sources, in brightness, pulse length and size, will
depend upon the ability to control the interaction physics.

5.2. Microscopy, holography and interferometry


The realization of a tunable, coherent light source in the 1–100 nm wavelength range
promises to open up as many new possibilities as the development of the laser in the
1960s. In a tutorial review of XUV sources, Attwood (1992) gives three basic definitions
which characterize coherent radiation: brightness, coherence length and transverse resolving
power. The brightness B is defined as:
B = 8/(1A 1 BW) (15)
where 8 is the number of photons per second, 1A is the source area, 1 is the solid angle
of the emitted radiation, and BW is the bandwidth δλ/λ. The temporal coherence length,
lcoh = λ2 /21λ (16)
determines the detail in which information can be resolved along the propagation path. The
resolving power is set by the diffraction limit:
dθ > λ/2π (17)
where d is the minimum resolvable distance on the object and θ is the observation half-angle.
Currently, there are two major high-brightness light sources with this wavelength range:
synchrotron undulators (Winick 1994) and kilojoule-class x-ray lasers (XRL) (Elton 1990).
Short-pulse lasers offer two alternative routes: optical-field- or inner-shell photoionized XRL
schemes and harmonic generation in gases or plasmas. At the time of writing, none of these
novel schemes offers a water-window source, but their compactness, scaling properties, high
efficiency and superior time resolution compared with long-pulse XRL schemes make them
well worth pursuing.
For applications where coherence is not essential, such as standard microscopy (see
for example, (Rochow and Tucker 1994)), electron beams can deliver both resolution and
contrast right down to atomic (sub-Angstrom) levels. The main advantage of photons over
786 P Gibbon and E Förster

electrons is their ability to pass less destructively through aqueous solutions. Moreover,
scanning electron microscopy normally requires carefully prepared biological samples, either
freeze-dried or treated with hydrophobic agents, a process which can alter the cellular
structure. With soft x-ray pulses from laser-produced plasmas (section 3.3), on the other
hand, one can imagine studying living cells with a time resolution sufficient to capture
dynamical processes on a sub-nanosecond timescale.
Apart from good temporal coherence, a prerequisite for biological holography is a
wavelength within the so-called ‘water-window’ between the absorption K-edges of oxygen
(23.2 Å) and carbon (43.7 Å) (Solem and Chapline 1984). This choice allows transmission
of the probe beam through the sample while providing natural contrast between proteins
(i.e. carbon) and water (oxygen). For example, this should yield information on protein
structures in their natural (aqueous) environment.
An important application of coherent XUV sources which is quite widespread is plasma
density diagnosis. In ICF and astrophysical plasmas, densities can be well above the critical
density Nc ' 1021 λ−2 −2
µ cm , which make them difficult or impossible to probe with visible
or UV lasers. A soft x-ray laser with λ < 20 nm, on the other hand, has a critical density of
1024 cm−3 or above, and can be used to obtain the plasma density using interferometry (Da
Silva et al 1995). An added advantage of ultrafast XUV schemes would be an improvement
in spatial resolution to sub-micron levels, by freezing hydrodynamic motion.

5.3. Ultrafast probing of atomic structure


Two techniques which have been used for some time to probe the structure of matter at
the atomic and molecular level are in situ x-ray diffraction and x-ray spectroscopy. These
methods have quite different source requirements and optical arrangements, see figure 5,
but share the exciting new possibilities offered by ultrafast time resolution. Consider, for
example, the typical timescales and length scales of protein motions shown in table 2 (Petsko
and Ringe 1984).

Table 2. Timescales and length scales of protein motions.

Time (s) Deflection (Å)

Atomic vibration 10−15 –10−11 0.01–1


Collective 10−12 –10−3 0.01–5
Bond breaking/joining 10−9 –10−3 0.5–10

Diffractometry exploits the interference effect created by adjacent atomic planes (Bragg
scattering) to obtain global structural information about fluid or crystal samples. Since x-
ray diffraction measurements can be directly inverted to atomic positions or bond lengths,
it is conceivable that ultrafast exposures on the 100 fs timescale would ultimately allow
‘filming’ of dynamic processes such as phase changes or chemical reactions (Barty et al
1995). Progress towards this goal has been recently achieved by Tomov et al (1995) using a
scheme similar to that shown in figure 5(b). They demonstrated a pump–probe experiment
to observe changes in the lattice temperature of a gold crystal on a 10 ps timescale.
Spectroscopy can also reveal information on atomic structure, but its interpretation
is generally complicated by uncertainties in bulk properties. An exception to this is
extended x-ray absorption fine structure (EXAFS), which yields direct information on the
near neighbours of a given atom. Soft x-rays from laser-plasmas have been successfully
Short-pulse laser–plasma interactions 787

(a)

(b)

Figure 5. X-ray optical arrangements for (a) pump–probe diffractometry and (b) ultrafast
absorption spectroscopy.

used as EXAFS sources for some time owing to their high brightness and sub-nanosecond
recording capability (Eason et al 1984). Again, short-pulse sources have been proposed as
a means of extending the time resolution down to the sub-picosecond regime (Tallents et al
1990). An advantage of spectroscopic techniques over diffractometry is that the required
x-ray photon flux is several orders of magnitude lower. Preliminary proof-of-principle
experiments have none the less concentrated on the near-edge spectrum (XANES), where
the source and detection requirements can be relaxed even further (Ráksi et al 1995).
788 P Gibbon and E Förster

5.4. Lithography
While lithography is often cited as a potential application for laser-plasma x-ray sources,
it is not obvious that ultrashort-pulse lengths bring any real advantage. To the authors’
knowledge, short-pulse systems have not yet been seriously evaluated in the context. In
order to use lasers for x-ray exposure of resists, one needs a short, tunable wavelength
(around 10 Å) for high resolution, combined with high average power to meet throughput
requirements. This does not rule out short-pulse x-ray sources in special cases, but
nanosecond lasers currently appear to represent the most promising option in this field
(Chaker et al 1990, Maldonado 1995).

5.5. Bench-top particle accelerators


The demise of the superconducting super collider (SSC) has remotivated the search for
alternatives to conventional particle accelerator technology. It has been realized for some
time that plasma could form the basis for a new generation of compact accelerators thanks
to their ability to support much larger electric fields. Conventional synchrotrons and linacs
operate with field gradients limited to around 100 MV m−1 . A plasma, on the other hand,
which is already ionized, can theoretically sustain a field 104 times larger, given by:
 1/2
me cωp Ne
Ep = ' GV cm−1 . (18)
e 1018 cm−3
To accelerate particles, these fields must propagate with velocities approaching the speed
of light. In a seminal paper, Tajima and Dawson (1979) proposed two methods of coupling
the transverse electromagnetic energy of a high-power laser into longitudinal plasma waves
with high phase velocity. The first requires a pulse length matched to the plasma period
such that τp ' π/ωp (Gorbunov and Kirsanov 1987, Sprangle et al 1988), which translates
into a technical requirement:
−1/2
tfwhm > 50N18 fs. (19)
This condition could not be met with the technology available at that time, so they
proposed an alternative ‘beat-wave’ scheme, in which two lasers are used with frequencies
chosen so that ω1 − ω0 = ωp . In contrast to the wakefield scheme, where a plasma wave
is forcibly driven up by the pulse, the beat-wave method relies on a more gentle build-up
over tens or even hundreds of picoseconds. In both cases, the plasma wave has a phase
velocity:
 
ωp2 1/2
Vph = c 1 − 2 . (20)
ω0
An electron trapped in such a wave will be accelerated over at most half a wavelength (after
which it starts to be decelerated), giving an effective acceleration length
 −3/2  
λp Ne λ −2
La = ' 3.2 cm. (21)
2(c − vph ) 1018 cm−3 µm
Combining (18) and (21), we obtain the maximum energy gain of an electron accelerated
by the plasma wave:
−1 −2
1U = Ep La ' 3.2N18 λµ GeV. (22)
Thus, a multi-terawatt Ti-Sa laser is in principle capable of accelerating an electron to 5 GeV
in a distance of 5 cm through a plasma with density 1018 cm−3 . In practice, acceleration will
Short-pulse laser–plasma interactions 789

be limited by other factors, such as laser diffraction or instabilities. For example, comparing
the Rayleigh length (8) with (21), we typically have ZR  La , so some means of guiding
the laser beam over the dephasing length must be found to optimize the energy coupling.
Whether relativistic or channel guiding (section 4.2) can be combined with large-amplitude
plasma wave generation has yet to be proven experimentally, but this will be one of the
goals of ‘second generation’ plasma-based accelerators in the near future, see Katsouleas and
Bingham (1996). To date, there have been some notable experiments demonstrating particle
acceleration with a beat-wave scheme (Kitagawa et al 1992, Clayton et al 1993, Everett
et al 1994, Amiranoff et al 1995). First experiments with short-pulse lasers (Nakajima et
al 1995, Modena et al 1995) have also achieved acceleration of thermal electrons to over
40 MeV, but the underlying mechanism has been attributed to Raman instabilities rather
than to ‘clean’ wakefield excitation. In an important step towards the latter, radial plasma
waves generated by ‘cigar-shaped’ pulses have been observed with both temporal and spatial
resolution by the LULI group (Marquès et al 1996).

5.6. Novel fusion concepts


One of the most exciting alternative schemes to conventional ICF to emerge recently is
the so-called ‘fast igniter’ concept (Tabak et al 1994). In the standard inertial confinement
scenario, ignition is achieved by compressing a deuterium–tritium pellet to high density in
such a way that a hot spot is created at the centre. This hot spot ignites first, providing
a spark which then propagates outwards to burn the surrounding higher density fuel. To
do this, laser energies of more than 1 MJ are needed for significant gain and implosion
symmetry has to be controlled to better than 1%.
In the fast igniter scheme, the hot spot is replaced by an external energy source, thus
theoretically relaxing the driver, compression and uniformity requirements by up to an order
of magnitude. Max et al (1974) proposed using a short-pulse, high-intensity laser to deliver
energy to the compressed fuel via fast electrons. The scheme works in three stages. First,
the fuel is compressed to a radius of around 10 µm, the equivalent of an α-particle range
at an areal density of 0.4 g cm−2 . Next, a ‘prepulse’ several hundred picoseconds long
with a tailored intensity profile is used to create an optical channel through the underdense
corona and push the critical surface closer to the centre of the target. Finally, a short pulse
with intensity of around 1020 W cm−2 is sent through the channel to the high-density core,
where it heats electrons to energies up to 1 MeV. These fast electrons penetrate the fuel,
where they thermalize and, ultimately, heat the ions to fusion temperatures (5–10 keV).
Although the scheme is highly speculative, it has none the less sparked a revival
of interest in the energy coupling and transport physics of high-intensity laser–plasma
interactions. A large number of experiments are therefore being planned and carried out
to investigate the issues outlined in sections 3.4–3.6. As our understanding of short-pulse
physics matures, we shall no doubt see improvements on the original Livermore scheme
and perhaps even more radical ideas for laser fusion.

6. Summary

The first decade of research on femtosecond laser–plasma interactions has proved


enormously rich both in variety of physics and in the potential for new and sometimes
unexpected applications. Each improvement in technology, whether in power, shorter
pulse duration, sharper focusing optics, or higher contrast ratios, sparks a fresh round of
experiments with a ‘look-and-see’ spirit. In this review we have tried to document the
790 P Gibbon and E Förster

changing emphasis in physics, from weakly nonlinear optics to extreme relativistic, kinetic
processes, which has accompanied these developments. In addition, we have explored some
of the areas in which femtosecond laser-produced plasmas show great promise as sources
of fast particles and short-wavelength radiation. With a new generation of high repetition-
rate, multi-terawatt lasers coming online this year, we can expect many of these ultrafast
applications to be realized in the near future.

Acknowledgments

Thanks are due to a number of colleagues who have provided valuable advice, comments
and, in some cases, unpublished material, notably: A Andreev, A R Bell, C B Darrow,
G Hölzer, J C Gauthier, J C Kieffer, T Missalla, S Svanberg, U Teubner, R P J Town and
I Uschmann.

References

Abramyan L A, Litvak A G, Mironov V A and Sergeev A M 1992 Sov. Phys.–JETP 75 978–82


Adam J C and Héron A 1988 Phys. Fluids 31 2602–14
Amendt P, Eder D C and Wilks S C 1991 Phys. Rev. Lett. 66 2589–92
Amiranoff F et al 1995 Phys. Rev. Lett. 74 5220–3
Ammosov M V, Delone N B and Krainov V P 1986 Sov. Phys.–JETP 64 1191
Andreev A A, Gamaliy E G, Novikov V N, Semakhin A N and Tikhonchuk V T 1992 Sov. Phys.–JETP 74 963–73
Andreev A A, Limpouch J and Semakhin A N 1994 Bull. Russ. Acad. Sci. 58 1056–63
Antonsen Jr T M and Mora P 1992 Phys. Rev. Lett. 69 2204–7
Attwood D 1992 Phys. Today 45 24–31
Audebert P, Geindre J P, Gauthier J C, Mysyrowicz A, Chamberet J P and Antonetti A 1992 Europhys. Lett. 19
189–94
Audebert P, Geindre J P, Rousse A, Gauthier J C, Mysyrowicz A, Grillon G and Antonetti A 1994 J. Phys. B: At.
Mol. Opt. Phys. 27 3303–14
Augst S, Strickland D, Meyerhofer D, Cin S L and Eberly J 1989 Phys. Rev. Lett. 63 2212–5
Auguste T, Monot P, Lompré L-A, Mainfray G and Manus C 1992a Opt. Commun. 89 145–8
——1992b J. Phys. B: At. Mol. Phys. 25 4181–94
Auguste T, Monot P, Mainfray G, Manus C, Gary S and Louis-Jacquet M 1994 Opt. Commun. 105 292–6
Barty C P J et al 1995 Time Resolved Electron and X-ray Diffraction vol 2521 (Bellingham, USA: SPIE) pp 246–57
Bell A R et al 1993 Phys. Rev. E 48 2087–93
Bezzerides B, Jones R D and Forslund D W 1982 Phys. Rev. Lett. 49 202–5
Birdsall C K and Langdon A B 1985 Plasma Physics via Computer Simulation (New York: McGraw-Hill)
Blyth W J, Preston S G, Offenberger A A, Key M H, Wark J S, Najmudin Z, Modena A, Djaoui A and Dangor A E
1995 Phys. Rev. Lett. 74 554–7
Bonnaud G, Gibbon P, Kindel J and Williams E 1991 Laser Part. beams 9 339–54
Borisov A B, Borovskiy A V, Korobkin V V, Prokhorov A M, Rhodes C K and Shiryaev O B 1990 Phys. Rev.
Lett. 65 1753–6
Borisov A B et al 1992 Phys. Rev. Lett. 68 2309–12
Born M and Wolf E 1980 Principles of Optics 6th edn (Oxford: Pergamon)
Brunel F 1987 Phys. Rev. Lett. 59 52–5
——1988 Phys. Fluids 31 2714–9
Bulanov S V, Inovenkov I N, Kirsanov V I, Naumova N M and Sakharov A S 1992 Phys. Fluids 4 1935–42
Burgess D and Hutchinson H 1993 New Scientist 140 28–33
Burnett N H and Corkum P B 1987 J. Opt. Soc. Am. B 6 1195–9
Carman R L, Rhodes C K and Benjamin R F 1981a Phys. Rev. A 24 2649–63
Carman R L, Forslund D W and Kindel J M 1981b Phys. Rev. Lett. 46 29–32
Chaker M, Boily S, Lafontaine B, Keffer J C, Pépin H, Toubhans I and Fabbro R 1990 Microelectron. Eng. 10
91–105
Chaker M, Kieffer J C, Matte J P, Pépin H, Audebert P, Maine P, Strickland D, Bado P and Mourou G 1991 Phys.
Fluids B 3 167–75
Short-pulse laser–plasma interactions 791

Chen X L and Sudan R N 1993 Phys. Rev. Lett. 70 2082–5


Chen H, Soom B, Yaakobi B, Uchida S and Meyerhofer D D 1993 Phys. Rev. Lett. 70 3431–4
Clayton C E et al 1993 Phys. Rev. Lett. 70 37–40
Coverdale C A, Darrow C B, Decker C D, Mori W B, Tzeng K-C, Marsh K A, Clayton C E and Joshi C 1995
Phys. Rev. Lett. 74 4659–62
Da Silva L B et al 1995 Rev. Sci. Instrum. 66 574–8
Darrow C B, Coverdale C, Perry M D, Mori W B, Clayton C, Marsh K and Joshi C 1992 Phys. Rev. Lett. 69
442–5
Davis J, Clark R and Giuliani J 1995 Laser Part. Beams 13 3–18
Decker C D, Mori W B and Katsouleas T 1994 Phys. Rev. E 50 R3338–41
Denavit J 1992 Phys. Rev. Lett. 69 3052–5
Drake J F, Kaw P K, Lee Y C, Schmidt G, Liu C S and Rosenbluth M N 1974 Phys. Fluids 17 778–85
Dunne M, Afshar-Rad T, Edwards J, MacKinnon A J, Viana S M, Willi O and Pert G 1994 Phys. Rev. Lett. 72
1024–7
Durfee III C G and Milchberg H M 1993 Phys. Rev. Lett. 71 2409–12
Eason R W, Bradley D K, Kilkenny J D and Greaves G N 1984 J. Phys. C 17 5067–74
Edwards J and Rose S J 1993 J. Phys. B: At. Mol. Opt. Phys. 26 L523–7
Elton R C 1990 X-Ray Lasers (Boston, MA: Academic)
Estabrook K G and Kruer W L 1978 Phys. Rev. Lett. 40 42–5
——1986 Resonant absorption in very steep density gradients Technical report: Lawrence Livermore National
Laboratory
Estabrook K G, Valeo E J and Kruer W L 1975 Phys. Fluids 18 1151–9
Everett M, Lal A, Gordon D, Clayton C E, Marsh K A and Joshi C 1994 Nature 368 527–9
Fedosejevs R, Ottman R, Sigel R, Kühnle G, Szatmari S and Schäfer F P 1990a Phys. Rev. Lett. 64 1250–3
——1990b Appl. Phys. B 50 79–99
Fews A P, Norreys P A, Beg F N, Bell A R, Dangor A E, Danson C N, Lee P and Rose S J 1994 Phys. Rev. Lett.
73 1801–4
Forslund D W, Kindel J M, Lee K, Lindman E L and Morse R L 1975a Phys. Rev. A 11 679–83
Forslund D W, Kindel J M and Lindman E L 1975b Phys. Fluids 18 1002–16
Forslund D W, Kindel J M and Lee K 1977 Phys. Rev. Lett. 39 284–8
Förster E, Gäbel K and Uschmann I 1992 Rev. Sci. Instrum. 63 5012–6
Förster E, Fill E E, Gäbel K, He H, Missalla T, Renner O, Uschmann I and Wark J 1994 J. Quant. Spectrosc.
Radiat. Transfer 51 101–11
Freeman R R, Bucksbaum P H, Milchberg H, Darack S, Schumacher D and Geusic M E 1987 Phys. Rev. Lett. 59
1092
Gamaliy E G 1994 Laser Part. beams 12 185–208
Gamaliy E G and Tikhonchuk V T 1988 JETP Lett. 48 453–5
Gibbon P 1994 Phys. Rev. Lett. 73 664–7
——1996 Phys. Rev. Lett. 76 50–3
Gibbon P and Bell A R 1992 Phys. Rev. Lett. 68 1535–8
Gibbon P, Monot P, Auguste T and Mainfray G 1995 Phys. Plasmas 2 1305–10
Ginzburg V L 1964 The Propagation of Electromagnetic Waves in Plasmas (New York: Pergamon)
Gorbunov L M and Kirsanov V I 1987 Sov. Phys.–JETP 66 290–4
Gordon III C L, Yin G Y, Lemoff B E, Bell P M and Barty C P J 1995 Opt. Lett. 20 1056–8
Grebogi C, Tripathi V K and Chen H-H 1983 Phys. Fluids 26 1904–8
Guérin S, Laval G, Mora P, Adam J C, Héron A and Bendib A 1995 Phys. Plasmas 2 2807–14
Hares J D, Kilkenny J D, Key M H and Lunney J G 1979 Phys. Rev. Lett. 42 1216–9
Herrlin K, Svahn G, Olsson C, Pettersson H, Tillman C, Persson A, Wahlström C-G and Svanberg S 1993
Radiology. 189 65–8
Jiang Z, Kieffer J C, Matte J P, Chaker M, Peyrusse O, Gilles D, Korn G, Maksimchuk A, Coe S and Mourou G
1995 Phys. Plasmas 2 1702–11
Joshi C J and Corkum P B 1995 Phys. Today 48 36–43
Katsouleas T and Bingham R (eds) 1996 IEEE Trans. Plas. Sci. in press
Keldysh L V 1965 Sov. Phys.–JETP 20 1307
Kieffer J C, Matte J P, Bélair S, Chaker M, Audebert P, Pépin H, Maine P, Strickland D, Bado P and Mourou G
1989a IEEE J. Quantum Electron. 25 2640–7
Kieffer J C, Audebert P, Chaker M, Matte J P, Johnston T W, Maine P, Delettrez J, Strickland D, Bado P and
Mourou G 1989b Phys. Rev. Lett. 62 760–3
792 P Gibbon and E Förster

Kieffer J C et al 1993 Phys. Fluids B 5 2676–81


Kitagawa Y et al 1992 Phys. Rev. Lett. 68 48–51
Klem D E, Darrow C, Lane S and Perry M D 1993 Short-pulse High-intensity Lasers and Applications II vol 1860
ed H A Baldis (Bellingham, USA: SPIE) pp 98–102
Kmetec J D, Gordon C L, Macklin J J, Lemoff B E, Brown G S and Harris S E 1992 Phys. Rev. Lett. 68 1527–30
Koch P and Albritton J 1974 Phys. Rev. Lett. 32 1420–3
Kohlweyer S, Tsakiris G D, Wahlström C-G, Tillman C and Mercer I 1995 Opt. Commun. 117 431–8
Krall J, Ting A, Esarey E and Sprangle P 1994 Phys. Rev. E 48 2157–60
Kruer W L 1988 The Physics of Laser Plasma Interactions (New York: Addison-Wesley)
Kühlke D, Herpes U and Von der Linde D 1987 Appl. Phys. Lett. 50 1785
Landen O L, Stearns D G and Campbell E M 1989 Phys. Rev. Lett. 63 1475
Lee Y T and More R M 1984 Phys. Fluids 27 1273–86
Leemans W, Clayton C E, Mori W B, Marsh K A, Kaw P K, Dyson A, Joshi C and Wallace J M 1992 Phys. Rev.
A 46 1091–105
Lefebvre E and Bonnaud G 1995 Phys. Rev. Lett. 74 2002–5
L’Huillier A and Balcou P 1993 Phys. Rev. Lett. 70 774–7
L’Huillier A et al 1995 J. Nonlin. Opt. Phys. Mat. 4 647–65
Limpouch J, Drska L and Liska R 1994 Laser Part. beams 12 101–10
Litvak A G 1968 Zn. Eksp. Teor. Fiz. 57 344–7
Mackinnon A J, Allfrey S, Borghesi M, Iwase A, Willi O and Walsh F 1995 Laser Interaction with Matter Inst.
Phys. Conf. Ser. vol 140 ed S J Rose pp 337–40
Maine P, Strickland D, Bado P, Pessot M and Mourou G 1988 IEEE J. Quantum Electron. 24 398–403
Maldonado J R 1995 Applications of Laser Plasma Radiation II vol 2523 ed M C Richardson and G A Kyrala
(Bellingham, USA: SPIE) pp 2–22
Marquès J R, Geindre J P, Amiranoff F, Audebert P, Gauthier J C, Antonetti A and Grillon G 1996 Phys. Rev.
Lett. in press
Matte J P and Aguenaou K 1992 Phys. Rev. A 45 2558–66
Matte J P, Kieffer J C, Ethier S, Chaker M and Peyrusse O 1994 Phys. Rev. Lett. 72 1208–11
Max C E, Arons J and Langdon A B 1974 Phys. Rev. Lett. 33 209–12
Meyerhofer D D, Chen H, Delettrez J A, Soom B, Uchida S and Yaakobi B 1993 Phys. Fluids B 5 2584–8
Milchberg H M and Freeman R R 1989 J. Opt. Soc. Am. B 6 1351–5
Milchberg H M, Freeman R R, Davey S C and More R M 1988 Phys. Rev. Lett. 61 2364–7
Milchberg H M, Lyubomirsky I and Durfee III C G 1991 Phys. Rev. Lett. 67 2654–7
Modena A et al 1995 Nature 377 606–8
Monot P, Auguste T, Gibbon P, Jackober F and Mainfray G 1995a Phys. Rev. E 52 R5780–3
Monot P, Auguste T, Gibbon P, Jackober F, Mainfray G, Dulieu A, Louis-Jacquet M, Malka G and Miquel J L
1995b Phys. Rev. Lett. 74 2953–6
More R M, Zinamon Z, Warren K H, Falcone R and Murnane M 1988a J. Physique C7 49 43–51
More R M, Warren K H, Young D A and Zimmerman G B 1988b Phys. Fluids 31 3059–78
Mori W B, Joshi C, Dawson J M, Forslund D W and Kindel J M 1988 Phys. Rev. Lett. 60 1298–301
Mori W B, Decker C D, Hinkel D E and Katsouleas T 1994 Phys. Rev. Lett. 72 1482–5
Mulser P, Pfalzner S and Cornolti F 1989 Laser Interaction with Matter ed G Velarde et al (Singapore: World
Scientific) pp 142–5
Murnane M M, Kapteyn H C and Falcone R W 1989 Phys. Rev. Lett. 62 155–8
Murnane M M, Kapteyn H C, Rosen M D and Falcone R W 1991 Science 251 531–6
Nakajima K et al 1995 Phys. Rev. Lett. 74 4428–31
Ng A, Celliers P, Forsman A, More R M, Lee Y T, Perrot F, Dharma-wardana M W C and Rinker G A 1994
Phys. Rev. Lett. 72 3351–4
Norreys P A et al 1996 Phys. Rev. Lett. 76 1832–5
Offenberger A A, Blyth W, Dangor A E, Djaoui A, Key M H, Najmudin Z and Wark J S 1993 Phys. Rev. Lett.
71 3983–6
Penetrante B M and Bardsley J N 1991 Phys. Rev. A 43 3100
Perry M D and Mourou G 1994 Science 264 917–24
Pert G J 1995 Phys. Rev. E 51 4778–89
Petsko G A and Ringe D 1984 Ann. Rev. Biophys. Bioeng. 13 331–71
Pfalzner S 1991 Appl. Phys. B 53 203–6
——1992 J. Phys. B: At. Mol. Opt. Phys. 25 L545–9
Pukhov A and Meyer-ter-Vehn J 1996 Phys. Rev. Lett. submitted
Short-pulse laser–plasma interactions 793

Rae S C 1993 Opt. Commun. 97 25–8


Rae S C and Burnett K 1991 Phys. Rev. A 44 3835–40
Ráksi F, Wilson K R, Jiang Z, Ikhlef A, Côté C Y and Kieffer J-C 1995 Applications of Laser Plasma Radiation
II vol 2523 ed M C Richardson and G A Kyrala (Bellingham, USA: SPIE) pp 306–15
Rochow T G and Tucker P A 1994 Introduction to Microscopy by Means of Light, Electrons, X Rays or Acoustics
2nd edn (New York: Plenum)
Röntgen W C 1895 Nature 53 274
Rosen M D 1990 Femtosecond to Nanosecond High-Intensity Lasers and Applications vol 1229 ed E M Campbell
pp 160–70
Rousse A, Audebert P, Geindre J P, Falliès F, Gauthier J C, Mysyrowicz A, Grillon G and Antonetti A 1994 Phys.
Rev. E 50 2200–7
Rozmus W and Tikhonchuk V T 1990 Phys. Rev. A 42 7401
Ruhl H and Mulser P 1995 Phys. Lett. 205A 388–92
Sakharov A S and Kirsanov V I 1994 Phys. Rev. E 49 3274–82
Sauerbrey R, Fure J, LeBlanc S P, Van Wonterghem B, Teubner U and Schäfer F P 1994 Phys. Plasmas 1 1635–42
Schnürer M, Kalashnikov M P, Nickles P V, Schlegel T, Sandner W, Demchenko N, Nolte R and Ambrosi P 1995
Phys. Plasmas 2 3106–10
Shepherd R, Booth R, Price D, Bowers M, Swan D, Bonlie J, Young B, Dunn J, White B and Stewart R 1994
Rev. Sci. Instrum. 66 719–21
Solem J C and Chapline G F 1984 Opt. Eng. 23 193–203
Soom B, Chen H, Fisher Y and Meyerhofer D D 1993 J. Appl. Phys. 74 5372–7
Sprangle P, Esarey E, Ting A and Joyce G 1988 Appl. Phys. Lett. 53 2146–8
Sprangle P, Esarey E and Ting A 1990 Phys. Rev. Lett. 64 2011–4
Sprangle P, Esarey E, Krall J and Joyce G 1992 Phys. Rev. Lett. 69 2200–3
Stamper J, Papadapoulos K, Sudan R N, McLean E, Dean S and Dawson J 1971 Phys. Rev. Lett. 26 1012–5
Stearns D G, Landen O L, Campbell E M and Scofield J H 1988 Phys. Rev. A 37 1684–90
Strickland D and Mourou G 1985 Opt. Commun. 56 219–21
Sudan R 1993 Phys. Rev. Lett. 70 3075–8
Sullivan A, Hamster H, Gordon S P, Nathel H and Falcone R W 1994 Opt. Lett. 19 1544–6
Sun G-Z, Ott E, Lee Y C and Guzdar P 1987 Phys. Fluids 30 526–32
Tabak M, Hammer J, Glinsky M E, Kruer W L, Wilks S C, Woodworth J, Campbell E M, Perry M D and Mason R J
1994 Phys. Plasmas 1 1626–34
Tajima T and Dawson J M 1979 Phys. Rev. Lett. 43 267–70
Tallents G J, Key M H, Ridgeley A, Shaikh W, Lewis C L S, O’Neill D, Davidson S J, Freeman N J and Perkins D
1990 J. Quant. Spectrosc. Radiat. Transfer 43 53–60
Teubner U, Bergmann J, Van Wonterghem B, Schäfer F P and Sauerbrey R 1993 Phys. Rev. Lett. 70 794–7
Teubner U, Wülker C, Förster E and Theobald W 1995 Phys. Plasmas 2 972–81
Teubner U, Gibbon P, Förster E, Falliès F, Audebert P, Geindre J P and Gauthier J C 1996a Phys. Plasmas in
press
Teubner U et al 1996b submitted for publication
Tillman C, Persson A, Wahlström C-G, Svanberg S and Herrlin K 1995 Appl. Phys. B 61 333–8
Tillman C, Mercer I, Svanberg S and Herrlin K 1996 J. Opt. Soc. Am. B 13 1–7
Tomov I V, Chen P and Rentzepis P M 1995 J. Appl. Cryst. 28 358–62
Town R P J, Bell A R and Rose S 1994 Phys. Rev. E 50 1413–21
——1995 Phys. Rev. Lett. 74 924–7
Tzeng K-C, Mori W B and Decker C D 1996 Phys. Rev. Lett. in press
Von der Linde D, Engers T, Jenke G, Agostini P, Grillon G, Nibbering E, Mysyrowicz A and Antonetti A 1995
Phys. Rev. A 52 R25–7
Weibel E S 1967 Phys. Fluids 10 741–8
Wilks S C 1993 Phys. Fluids B 5 2603–8
Wilks S C, Kruer W L, Tabak M and Langdon A B 1992 Phys. Rev. Lett. 69 1383–6
Winick H 1994 Synchrotron Radiation Sources: A Primer (Singapore: World Scientific)
Zel’dovich Ya B and Raizer Yu P 1966 Physics of Shock Waves and High-Temperature Phenomena (New York:
Academic)
Ziman J M 1969 Principle of the Theory of Solids (Cambridge: Cambridge University Press)

S-ar putea să vă placă și