You are on page 1of 7

INSTITUTE OF PHYSICS PUBLISHING Plasma Phys. Control.

Fusion 43 (2001) A31A37

PLASMA PHYSICS AND CONTROLLED FUSION PII: S0741-3335(01)28865-4

Physics of relativistic laser-plasmas


Patrick Mora
Centre de Physique Th eorique (UPR 14 du CNRS), Ecole Polytechnique, 91128 Palaiseau Cedex, France E-mail: Patrick.Mora@cpht.polytechnique.fr

Received 22 June 2001 Published 22 November 2001 Online at stacks.iop.org/PPCF/43/A31 Abstract The interaction of ultra-intense (and ultra-short) laser beams with plasmas gives rise to a variety of phenomena. The propagation is, in principle, possible in an overdense plasma if the laser intensity is above a threshold xed by the electron density. However, the solutions describing this so-called relativistically selfinduced transparency are subject to violent electron instabilities, as is already the case in an underdense plasma. The growth rates of the instabilities are so large that it appears hopeless to propagate efciently a high-intensity beam in a cold plasma. The situation is somewhat more favourable in a relativistically hot plasma, where the growth rates of the instabilities are signicantly reduced. In any case, the interaction of the laser beam with the plasma results in a strong electron acceleration and heating. The electron acceleration is both due to the plasma waves generated in the plasma by the laser beam and to the laser eld itself. In present-day experiments, the fastest electron energy can be in the range of hundreds of MeV. Correlatively, fast ions can be accelerated by the charge separation electric elds, and energetic photons due to bremsstrahlung appear, which in turn can be responsible for photonuclear reactions. The transport and interactions of all these energetic particles in and outside the plasma are interesting for various applications, such as possible laser acceleration of electrons up to the GeV range, high-energy short-duration particle sources, protron radiography, fast-ignition approach to inertial connement fusion, etc.

1. Development of ultra-short lasers During the last 1015 years, there has been a rapid development of ultra-short and ultra-intense lasers using the chirped-pulse amplication (CPA) technique [1]. In the CPA technique, a femtosecond pulse is rst stretched and chirped by a grating pair (gure 1), then sent into an amplier, keeping the pulse intensity below the damage threshold of the amplier, and then compressed back by a conjugate grating pair to its original duration.
0741-3335/01/SA0031+07$30.00 2001 IOP Publishing Ltd Printed in the UK A31

A32

P Mora

Time

Figure 1. Scheme of a chirped pulse. Here the instantaneous frequency linearly increases with time.

Whereas the long-pulse lasers work in the nanosecond range with energy in the kilojoule rangeand, in the near future, in the megajoule rangethe ultra-short-pulse lasers work in the sub-picosecond range with energy in the 11000 J range. Their comparatively low cost explains the spread in the use of these lasers in a number of laboratories. The LULI 100 TW laser at Palaiseau is one of the most powerful of these lasers. It delivers typically 30 J in 300 fs. Focussed on a 100 m2 focal spot, one obtains 1020 W cm2 , with a 1.06 m laser wavelength. The classical (non-relativistic) oscillation velocity of an electron in such a high-intensity eld is of the order of 10 times the light velocity, which means that one is by far in the relativistic interaction regime. 2. Free electron in the eld of a plane electromagnetic wave To illustrate the relativistic interaction regime, let us rst consider a free electron in a plane electromagnetic wave propagating, say, in the z direction, with electric eld in the x direction and magnetic eld in the y direction. In the low-intensity regime, one can apply the linearresponse theory for which the electron oscillates along the direction of the electric eld. In the nonlinear relativistic regime, the magnetic eld curves the electron trajectory towards the z direction. As a result, the electron has a longitudinal as well as a transverse motion, with respect to the direction of light propagation. For an electron initially at rest, one can easily obtain 2 1 p p = , (1) 2 me c where p is the electron momentum. Figure 2 shows the normalized electron trajectory in this case. In the denition of the axis, a0 = eA0 /me c, where A0 designates the peak value of the vector potential of the wave. At sufciently high intensities (a0 1), the electron motion is essentially along the z direction. There is a simple relation between the Lorentz factor of the electron, v2 , c2 and the angle between the electron trajectory and the z direction, = 1 tan =
1/2

(2)

1/2 2 . (3) 1 Note that these laws are valid only for plane waves. For a nite-size focal spot, corrections have to be considered [2].

Physics of relativistic laser-plasmas

A33

0.5 k0x /a0

0.5

1 0

0.5

1.5 2 (2/) k0z/a02

2.5

Figure 2. Electron trajectory in the eld of a plane electromagnetic wave of arbitrary amplitude. a0 is the normalized amplitude of the wave, and k0 = 0 /c is the wave number.

3. Relativistic index of refraction As the linear-response theory is unable to describe the electron motion in the relativistic regime, it is not possible in general to calculate an index of refraction in this regime. However, there is a simplied case where it can be done, when the electromagnetic wave is circularly polarized, and when the longitudinal motion is supposed to be completely quenched by a charge separation eld. Then, one has N2 = 1
2 pe

(4)

where N is the index of refraction, is the light frequency, pe is the electron plasma frequency, and is the Lorentz factor in the high-intensity wave, e 2 A2 . (5) 2 m2 ec In this expression, A is the vector potential of the electromagnetic wave. Note that the relativistic effect is equivalent to a reduction of the effective electron density ne neff = . (6) As a result, the plasma can theoretically be transparent in the so-called relativistic induced transparency regime, where = 1+ nc < ne < nc ,
2 1/2

(7)

with nc = me 0 /e as the classical critical density for light at frequency [3, 4]. In fact, this relativistic induced regime is very difcult to observe [5], at least in a cold plasma, because of the electron instabilities (see section 6). 4. Relativistic ponderomotive force The nonlinear force acting on one electron can be written as fnl = e2 A2 , 2 me (8)

A34

P Mora

where means the average over the high frequency of the laser light [6]. The relativistic effect corresponds to a reduction with respect to the classical expression of the nonlinear (ponderomotive) force by the averaged Lorentz factor . On the other hand, the pressure exerted by the light on a dense plane target is given by I (9) Prad = (1 + R) , c where R is the reection rate and I the light intensity, as in the non-relativistic limit. In the relativistic regime, this radiation pressure may be very large, in the 100 Gbar range for I = 1020 W cm2 . It can push the plasma efciently and accelerate it up to a few per cent of the light velocity [7]. For a nite-size focal spot, it can drill a hole in the target (hole boring) [8]. 5. Relativistic self-focusing and guiding Let us now consider a typical laser pulse with a radial-dependent shape I (r), with dI < 0, dr and the corresponding radial-dependent index of refraction N 2 (r) = 1
2 (r) pe

(10)

(r)2

(11)

In this expression, pe is also supposed to depend on r , due to the ponderomotive force which tends to expell the electrons from the laser channel, with d pe > 0. dr As a result of the and pe dependence on r , one has (12)

dN <0 (13) dr Therefore, the plasma acts as a lens on the laser light. There exists a critical power for which this lens effect dominates the natural diffraction, Pcrit (GW) = 16.2 2 . 2 pe (14)

Above this power, the laser light self-focusses down to a focal spot that is few wavelengths in size, further increasing the laser intensity on its axis. The laser pulse can then be self-guided 1 2 w0 /c, which is the typical length on distances much longer than the Rayleigh length, zR = 2 of the focal region of a weak-amplitude laser light [9]. In this expression, w0 is the waist of the laser beam, i.e. A behaves as exp[(r/w0 )2 ] at the focal spot in vacuum. 6. Electron parametric instabilities Laser light propagation in plasma is known to be subject to parametric instabilities, where the laser energy decays into scattered light and plasma modes. For the ultra-short laser pulses, the most important instabilities are the electron plasma instabilities. The corresponding wellknown generic instability in the moderate-intensity regime is the stimulated-scattering Raman instability, where the laser light of frequency 0 and wave vector k0 decays into a scattered

Physics of relativistic laser-plasmas

A35

1.0 0.8 0.6 0.4 0.2 0.0 50 100 150 X / 0 200

Figure 3. Snapshots of the prole of the pulse intensity as it propagates through the plasma slab. 1 Snapshots are plotted at time intervals of 4000 . The intensity is normalized to the initial peak intensity. This gure corresponds to a one-dimensional simulation and is taken from [12].

down-shifted electromagnetic wave (1 , k1 ) and an electron plasma wave (p , kp ), with the conservation laws 0 = 1 + p , k0 = k1 + kp . (15) (16)

In the relativistic regime, some of the properties of the Raman instability are modied [10]. In particular, it can occur at any density, provided that the laser can propagate, i.e. ne < 1 nc at moderate intensities. Moreover, instabilities nc , while it was limited to ne < 4 previously known in the moderate-intensity regime as two-plasmon instability and relativistic lamentation are in fact unied with the Raman instability in the relativistic regime. The growth 1 rate of the instability can be extremely large, up to 2 0 , which corresponds to a timescale of a femtosecond for a 1.06 m laser light. It results in a rapid destruction of the wave front of the light, a strong scattering of the light, and a strong electron heating. Once the plasma is heated to a temperature in the MeV range, the growth rate and the domain of wave vectors kp for which the instability occurs are reduced [11]. As a typical scenario, let us consider a 300 fs laser pulse propagating through a 200 m width plasma slab of density 0.2nc , with a peak intensity of 1020 W cm2 (see gure 3). Numerous one- and two-dimensional particle-in-cell simulations were performed in this context by Adam et al [12]. As a result of the instability, the front part of the pulse is strongly depleted, leading to a strong increase of the initially cold electron temperature. Once the electrons have reached MeV temperature, the instability is less efcient and the back part of the pulse is able to propagate accross the plasma slab. 7. Electron acceleration As a general rule, laserplasma interaction in the relativistic regime is very efcient in producing very energetic particles, primarily electrons [13]. The resultant electron energy spectrum strongly depends on the parameters of the experiment (pulse intensity, duration and focal spot size, plasma density and geometry, etc.). Electrons in the 100 MeV range have been detected in plasma of electron density in the 1019 cm3 range obtained by instantaneous ionization of gas jets [1416]. Numerous energetic electrons are obtained similarly (though with comparatively lower energies) in laser-solid experiments. The acceleration mechanisms

A36

P Mora

are the subject of numerous studies. The mechanisms differ, depending on the parameters of the experiment. Typically, in low-density experiments, the plasma wave resulting from the electron Raman instability [17] can reach such an amplitude that it breaks generating electrons in the tens of MeV range. At higher plasma density, including the interaction with initially solid targets, the acceleration mechanisms are more complex. It involves plasma waves, as in low-density plasmas, and, more generally, the electric and magnetic low-frequency elds selfconsistently generated in the plasma, and the high-frequency electromagnetic waves (laser + scattered waves) [18]. The propagation of these electrons in solid targets is a major subject of present experimental and theoretical studies [1922]. 8. Ions and other high-energy particle sources The energetic electrons may in turn transfer part of their energy to ions or generate x- or -rays. The energy transfer to ions occurs mainly via charge separation elds either at the front or at the rear side of thin foils [23]. Protons in the tens of MeV range were observed [2426]. Heavy ions up to 450 MeV have also been obtained with high-Z material targets [27]. Protons can be used for proton radiography with a high resolution. The x- and -rays result due to bremsstrahlung of electrons on ions of the target. Finally, one can have nuclear reactions due to fast ions or photo-induced [2830]. 9. Fast ignition of thermonuclear targets One of the most exciting applications of ultra-fast and ultra-intense laser pulses is the fast ignition of thermonuclear targets [31], either via the energetic electrons created by the laserplasma interaction or via the fast ions accelerated by the electrons in thin foils conguration [32]. In both cases, the idea is to bring a signicant part of the energy of the ultra-short pulse close to the core of a thermonuclear target previously compressed by a long laser pulse (in the 10 ns range). The heated part is supposed to burn its nuclear fuel and ignite the whole target. The scheme allows a reduction of the long-pulse energy normally necessary to lead to ignition and to the same burn efciency. A complete discussion of the fast-ignition concept and studies is given in the paper by J Meyer-ter-Vehn in these proceedings. 10. Conclusion The physics of relativistic laserplasmas has renewed the interest in laserplasma interaction in the past 10 years [33]. Promising applications such as fast-ignition of thermonuclear pellets, proton radiography, high-energy particles and x-ray sources, are foreseen. Acknowledgments Numerous discussions with J C Adam, A H eron and G Laval are acknowledged. References
[1] Mourou G and Umstadter D 1992 Phys. Fluids B 4 2315 Perry M D and Mourou G 1994 Science 264 917 [2] Quesnel B and Mora P 1998 Phys. Rev. E 58 3719 [3] Lefebvre E and Bonnaud G 1995 Phys. Rev. Lett. 74 2002 [4] Gu erin S, Laval G, Mora P, Adam J C and H eron A 1996 Phys. Plasmas 3 2693

Physics of relativistic laser-plasmas [5] [6] [7] [8] [9] [10] [11] [12] [13] [14] [15] [16] [17] [18] [19] [20] [21] [22] [23] [24] [25] [26] [27] [28] [29] [30] [31] [32] [33] Fuchs J et al 1998 Phys. Rev. Lett. 80 2326 Mora P and Antonsen T M Jr 1997 Phys. Plasmas 4 217 Denavit J 1992 Phys. Rev. Lett. 69 3052 Takahashi K et al 2000 Phys. Rev. Lett. 84 2405 Chessa P Mora P and Antonsen T M Jr 1998 Phys. Plasmas 5 3451 Quesnel B, Mora P, Adam J C, Gu erin S, H eron A and Laval G 1997 Phys. Rev. Lett. 78 2132; Quesnel B, Mora P, Adam J C, H eron A and Laval G 1997 Phys. Plasmas 4 3358 Adam J C, H eron A, Laval G and Mora P 2001 Phys. Plasmas 8 1664 Adam J C, H eron A, Laval G and Mora P 2000 Phys. Rev. Lett. 84 3598 Amiranoff F et al 1998 Phys. Rev. Lett. 81 995 Modena A et al 1995 Nature 377 606 Gahn C et al 1999 Phys. Rev. Lett. 83 4772 Malka V et al 2001 Phys. Plasmas 8 2605 Antonsen T M Jr and Mora P 1992 Phys. Rev. Lett. 69 2204 Antonsen T M Jr and Mora P 1993 Phys. Fluids B 5 1440 Pukhov A et al 1999 Phys. Plasmas 6 2847 Pukhov A and Meyer-ter-Vehn J 1997 Phys. Rev. Lett. 79 2686 Gremillet L et al 1999 Phys. Rev. Lett. 83 5015 Borghesi M et al 1999 Phys. Rev. Lett. 83 4309 Honda M, Meyer-ter-Vehn J and Pukhov A 2000 Phys. Rev. Lett. 85 2128 Pukhov A 2001 Phys. Rev. Lett. 86 3562 Clark E L et al 2000 Phys. Rev. Lett. 84 670 Snavely R A et al 2000 Phys. Rev. Lett. 85 2945 Mackinnon A J et al 2001 Phys. Rev. Lett. 86 1769 Clark E L et al 2000 Phys. Rev. Lett. 85 1654 Disdier L et al 2000 Phys. Rev. Lett. 82 1454 Ledingham K W D et al 2000 Phys. Rev. Lett. 84, 899 Cowan T E et al 2000 Phys. Rev. Lett. 84 903 Tabak M et al 1994 Phys. Plasmas 1 1626 Roth M et al 2001 Phys. Rev. Lett. 86 436 Umstadter D 2001 Phys. Plasmas 8 1774

A37

You might also like