Sei sulla pagina 1di 28

FAST ION GENERATION BY HIGH-INTENSITY LASER IRRADIATION OF SOLID TARGETS AND APPLICATIONS

M. BORGHESI,* J. FUCHS, S. V. BULANOV,7 A. J. MACKINNON,# P. K. PATEL,# and M. ROTH**


The Queens University, School of Mathematics and Physics, Belfast BT7 1NN, United Kingdom Laboratoire pour lUtilisation des Lasers Intenses, UMR 7605 CNRS-CEA-cole Polytechnique-Universit Paris VI, 91128 Palaiseau 3, France University of Nevada, Physics Department, MS-220, Reno, Nevada 89557 7Kansai Research Establishment, APRC-JAERI, Kizu, Japan #Lawrence Livermore National Laboratory, Livermore, California **Technical University Darmstadt, Darmstadt, Germany

Received March 22, 2005 Accepted for Publication September 26, 2005

The acceleration of high-energy ion beams (up to several tens of mega-electron-volts per nucleon) following the interaction of short (t 1 ps) and intense (Il2 10 18 W{cm 2 {mm 2 ) laser pulses with solid targets has been one of the most active areas of research in the last few years. The exceptional properties of these beams (high brightness and high spectral cutoff, high directionality and laminarity, and short burst duration) distinguish them from the lower-energy ions accelerated in earlier experiments at moderate laser intensities. In view of these properties, laser-driven ion beams can be employed in a number of groundbreaking applications in the scientific, technological, and medical areas. This paper reviews the main experimental results obtained in this area in recent years, the properties of the accelerated beams, the relevant theoretical and computational models, and the main applications that have been implemented or proposed.
KEYWORDS: ion acceleration, laser-plasma interaction, inertial confinement fusion

I. INTRODUCTION One of the most important results recently obtained in laser-plasma interaction experiments is the observation of very energetic beams of ions produced from laserirradiated thin metallic foils. In a number of experiments,
*E-mail: m.borghesi@qub.ac.uk 412

performed a few years back with different laser systems and under different interaction conditions, protons with energies up to several tens of mega-electron-volts were detected behind thin foils irradiated with high-intensity pulses.13 These high-energy proton beams have fundamentally different properties from lower-energy protons observed in earlier work at lower laser intensity with laser pulses in the nanosecond and tens-of-picosecond regime,4,5 which were accelerated from the coronal plasma and emitted into a large solid angle. Beams produced during these longer interactions exhibited strong trajectory crossing and a broad energy spectrum with typical ion temperatures of ;100 keV0nucleon. These unspectacular characteristics prevented major applications. On the contrary, beams accelerated by ultraintense laser pulses exhibit a remarkable degree of beam collimation and laminarity, high cutoff energy, and emission along the normal to the unirradiated rear surface of the target. Since the first observations, an extraordinary amount of experimental and theoretical work has been devoted to the study of the characteristics and production mechanisms of these beams. Particular attention has been devoted to the exceptional accelerator-like spatial quality of the beams, and current research focuses on their optimization for use in a number of groundbreaking applications. The scope of this paper is to report on the state of the art in this area of research, with a particular view to possible application of these sources in a fast ignitor0 inertial confinement fusion ~ICF! context. The following sections will briefly outline the ion acceleration and transport mechanisms from a theoretical point of view and describe the main experimental results and the applications of possible ICF relevance. Other important applications will be briefly mentioned for the sake of completeness.
FUSION SCIENCE AND TECHNOLOGY VOL. 49 APR. 2006

Borghesi et al.

FAST ION GENERATION BY LASER IRRADIATION OF SOLID TARGETS

We will focus mainly on proton beams produced from irradiation of solid targets, as this is the most efficient and promising mechanism. Nevertheless, we note that ion acceleration can also take place in underdense plasmas, which can be obtained by ionizing gas jets targets, or can be formed by ionization and evaporation of thin solid targets due to the laser pulse pedestal and0or by the prepulse 6,7 ~even under conditions in which the target is not exploded, a preplasma is generally present in front of the target 8 !. During high-intensity propagation through an underdense plasma, Coulomb explosion processes can accelerate ions radially away from the laser axis,9,10 up to energies of the order of the lasers ponderomotive potential. This process can be enhanced by collisionless shock acceleration.11

II. THEORETICAL AND COMPUTATIONAL MODELS OF LASER-ION ACCELERATION The ion acceleration processes have been extensively investigated theoretically and numerically, mainly by means of particle-in-cell ~PIC! computer simulations. There have been several theoretical mechanisms that have been proposed to interpret the experimental observations of ion beams accelerated by ultraintense lasers from solid targets. The mechanisms that have attracted the most attention and appear to be more relevant to currently accessible experimental regimes are related to large electric fields set up by laser-accelerated electrons at target interfaces. A schematic of a typical charge and electric field configuration following high-intensity laserpulse interaction with a solid foil is shown in Fig. 1. Electrons escaping from the front target will set up an

Fig. 1. Schematic representation of the charge and electric field distribution following high-intensity laser interaction with a solid foil. The arrows below the x axis show the direction of fast ion motion.
FUSION SCIENCE AND TECHNOLOGY VOL. 49 APR. 2006

electric field that will accelerate ions in the backward direction ~i.e., toward the laser!. Electrons propagating forward into the target will set up fields in the interior of the target and at the target rear, which will accelerate ions in the forward direction. The attention of researchers has focused mainly on forward-accelerated ion beams as these exhibit higher beam quality and, typically, higher energies. Electrons can be accelerated forward inside the target by the laser through a variety of mechanisms that depend on the interaction and target conditions.12 Two cases may be distinguished depending on the plasma density gradient present at the front surface of the target. When a steep density gradient is present at the target front surface ~i.e., in high-contrast laser interactions!, electrons are forward-accelerated mainly by vacuum heating 13,14 or, for a , 1, by resonant absorption.15,16 However, these processes result in a smaller number of accelerated electrons compared with the case of electron acceleration in longer scale-length plasmas. Extended plasmas can be produced at the target front surface by the laser pulse pedestal and0or by the prepulse 17 that cause evaporation and subsequent ionization of the target material. In this case, several electron acceleration mechanisms are relevant. Despite extensive work on this topic, no clear picture has emerged yet owing to the complexity of the laser-plasma interplay. Indeed, during high-intensity propagation through an underdense plasma, the laser pulse can undergo processes such as relativistic self-focusing 1821 leading to intensity increase 22,23 or conversely detrimental processes such as filamentation,24,25 all depending on the local plasma gradient, the laser intensity, and its spatial distribution. Among the processes coupling the laser energy into fast electrons, the main ones are j B heating 26,27 and betatron acceleration.28 Once accelerated inward, the fast electrons induce a charge-separation electrostatic field at the critical surface interface. This field will in turn result in acceleration of ions swept from the target front surface.29 The typical energies of accelerated fast electrons are such that their mean free path is much larger than the thickness of the targets typically used in experiments. The target capacitance however allows only a small fraction of electrons to escape before the target is sufficiently charged so that further escape is nearly impossible. The fast electrons that are electrostatically confined on the target rear surface therefore set a charge-separation field over a Debye length.30,31 Typically, the Debye length is of the order of 1 mm, resulting in strong ~approximately teravoltper-meter! electric fields. Such fields can ionize atoms and rapidly accelerate ions normal to the initially unperturbed surface. It has been experimentally observed that this process is responsible for the acceleration of the highest-energy ions, as will be detailed later. The accelerated ions form a dense bunch of short duration that is charge neutralized by co-moving electrons. The extremely short duration of the acceleration and the fact
413

Borghesi et al.

FAST ION GENERATION BY LASER IRRADIATION OF SOLID TARGETS

that it starts from an initially cold surface are essential facts that result in the unique characteristics of the ion beam, as will be detailed in the following sections. After this initial phase ions stream into vacuum with electrons, preceded by a Debye sheath of hot electrons, as illustrated in Fig. 2. The charge-separation structure at the expansion front results in a peak of the accelerating electric field at the ion front, as shown in Fig. 3a ~Ref. 32!. The most energetic electrons always extend farther out into vacuum, maintaining the accelerating field as long as the electron temperature is high. The acceleration from the target rear has been described by several authors as an extension of the classical case of a plasma expanding into vacuum,33 driven by the ambipolar electric field generated in a narrow layer at the front of the plasma cloud. There are however fundamental differences between the shortpulse and long-pulse cases. In the latter, bulk effects and collisional ionization by thermal electrons in the coronal plasma are the dominant mechanisms. In the short-pulselaser case, however, the ion generation and acceleration mechanisms are decoupled from the stochastic laser plasma. This general scenario has been confirmed by PIC simulations, as will be seen below. Early theoretical models for the description of plasma expansion were based on the quasi-neutral behavior of the plasma and on the thermodynamical equilibrium of electrons on nanosecond timescales.3335 Under these assumptions analytical solutions of self-similar type 33 and in the framework of the renormalization group theory 36,37

Fig. 3. ~a! Structure of the electric field along the expansion direction of a plasma streaming into vacuum under the influence of a hot electron Debye sheath. ~b! Resulting spectra of the accelerated ions at different times in the expansion. ~Figures extracted from Ref. 32.!

Fig. 2. Schematic representation of laser acceleration of ions from the target rear surface. The position of the rear target surface is indicated in the drawing ~the target is located on the left!. The unneutralized Debye sheath of hot electrons at the front of the expansion induces acceleration of ions co-moving with neutralizing electrons. ~Figure extracted from Ref. 30.! 414

were found. In the subpicosecond laser irradiation case, the assumption of quasi neutrality should be abandoned since as we have seen, the process of ion acceleration finds its origin precisely in the strong charge separation, which is produced in the very early phase of the lasersolid interaction. The effect of charge separation has indeed been the object of attention of recent theoretical work,32,38,39 which has updated the freely expanding plasma model to the case of a sudden burst of energetic electrons but still in the framework of an isothermal expansion. More realistic models that take into account the finite size of the initial reservoir of particles and the adiabatic cooling and energy transfer between the ions and electrons 40 or the effect of two-electron temperatures 41 have been more recently developed. Indeed, as the beam expands, the fast electrons transfer progressively their energy to the ions, and the accelerating chargeseparation field decreases until the ion energy saturates, which is a fact that is not accounted for in the isothermal models. Charge neutralization of the ion beam is provided by the initially hot electrons that expand into vacuum with
FUSION SCIENCE AND TECHNOLOGY VOL. 49 APR. 2006

Borghesi et al.

FAST ION GENERATION BY LASER IRRADIATION OF SOLID TARGETS

the ion beam. These electrons cool down as t 2 ~Ref. 42! within a few hundreds of microns of the expansion to reach a co-moving state with the protons. Since the characteristics of this sheath of hot electrons depend on the transport properties of the fast electrons through the target, transport processes have a crucial impact on the beam properties of the ions accelerated from the rear surface. This is an area of study that is under intense scrutiny at present.43 49 Theoretical considerations based on computational and experimental results indicating that the peak proton energy increases in inverse proportion to the target thickness 29,50 have highlighted the role of electron recirculation. Electrons that have time to recirculate between the front and rear surfaces of the target within the temporal width of the pulse could give rise to an enhanced hot electron density at each surface, leading to higher accelerating electric fields and thus higher proton energies. This effect becomes less important as the target thickness increases. Theoretical considerations predict a priori that frontside acceleration yields lower energetic protons than rearside acceleration. At the laser-irradiated target front surface, the sweeping electrostatic potential that accelerates ions at the front 29 is balanced by the laser ponderomotive potential, i.e., fsweeping ; fpond . Therefore, the maximum energy that can be extracted by the ions after passing through this potential is also of the order of
fpond m e c 2 @ ~1 Il2 01.37 mm 10 18 W{cm
2

where t acceleration time, i.e., roughly the laser pulse duration 51 ion plasma frequency ~ne0 is the m i 0 initial electron sheath density, and m i is the ion mass!. ne0 Ze 2

vpi

{mm 2 !102

1# ,

A direct comparison between Emax0front and Emax0rear shows that higher ion energies can be a priori expected from rear-surface acceleration. Note however that the energy of ions accelerated from the front can be boosted by the rear-surface field if the target is thin enough that the rear-surface field is still large when these ions cross the rear surface.52 A different mechanism that has been studied theoretically and can in principle lead to ion energies much larger than the ponderomotive potential is related to the launch of an electrostatic shock from the front of the target by the ponderomotive force of the laser. Theoretical and computational work 5355 describes how the recession of the surface, driven by the laser radiation pressure, can launch an ion acoustic wave into the plasma, which in turn can evolve into an electrostatic shock. The recession of the target front surface is due to the ponderomotive pressure of the laser pulse pL 2I0c, which is of the order of gigabar at 1 10 19 W{cm 2. This pressure pushes the critical surface inward at a velocity u0c @~ncr 02ne ! ~Zme 0Mi ! ~Il201.37 10 18 !# 102 ,

where me I l mm electron mass laser power density ~intensity! ~W{cm laser wavelength ~ mm!.
2

The acceleration ~sweeping! time is however finite ~tsweeping ; 100 fs for protons and a 1-mm wavelength laser!, and if it is shorter than the laser pulse duration tlaser , i.e., the time during which the photon pressure is maintained, the maximum energy gained by the ions will be reduced. Therefore, the maximum energy of the ions ~with a charge state of Z! will be Emax0front ; Zfpond if tlaser tsweeping and Emax0front ; Zfpond ~tlaser 0tsweeping !102 if tlaser tsweeping . This has to be compared with the maximum energy that can be gained by protons accelerated at the rear. A simple estimate of the maximum energy that can be gained by the accelerated ions based on a self-similar fluid theory 32 is Emax0rear ; 2Zfpond{ ln 2{ vpi t
2

M 2e
VOL. 49

~where ncr is the critical density and ne the electron plasma density! as described by Wilks et al.27 In turn, the collisionless shock can trap and accelerate ions to high energy. As these shock-accelerated protons, after crossing the target, gain additional energy from the back-surface electrostatic field, they screen this field and reduce the acceleration for protons originating from the back. Shock-accelerated protons can have higher energies than protons accelerated from the rear target surface for very high laser intensities ~I . 10 21 W0cm 2, as inferred from the PIC simulations! and relatively thick targets. A signature of the mechanism is a plateau in the spectrum near the high-energy cutoff.55 These different mechanisms have also been studied extensively and confirmed by means of PIC simulations.29,51,56 65 A typical example of PIC simulation of the interaction between an ultraintense laser pulse and a solid target ~composed of deuterium and hydrogen! leading to the acceleration of ions is shown in Fig. 4. Figure 4a shows the density profiles of the different species at an instant in time in the expansion.65 The fast electrons precede the protons, followed by heavier deuterons for which the highest field has been screened by the protons.
415

FUSION SCIENCE AND TECHNOLOGY

APR. 2006

Borghesi et al.

FAST ION GENERATION BY LASER IRRADIATION OF SOLID TARGETS

Fig. 5. Ion energy spectra from 3-D PIC simulations.57 The lines represent ions accelerated inward at the front surface, ions accelerated at the rear surface, and the selection of ions escaping the simulating box within 10 4 -srad solid angle around the rear-surface normal.

Fig. 4. The electron, deuteron, and proton density profiles averaged near the laser axis at 462 fs. The dot-broken line indicates the initial electrons density profile. Plot ~b! shows the protons phase plot X versus Px at the same time of plot ~a!. The dotted lines indicate the initial target surfaces. ~Figure extracted from Ref. 65.!

This compares favorably with the simplified picture shown in Fig. 1. Figure 4b shows a phase-space plot of the accelerated protons along the laser axis. Note that the protons taking off from the target front surface achieve much lower energies than the ones accelerated from the target rear surface. Being able to discriminate between ions accelerated at the target rear and at the target front, several PIC simulations confirmed that rear-surface acceleration produces higher-energy ions for laser intensities in the range of 10 17 to 10 19 W{cm 2, showing also that rear-surface acceleration induces a higher conversion efficiency of the laser energy into the ion beam energy than front-surface acceleration.29,57 Sentoku et al. obtain conversion efficiencies of up to 8% for the rearside acceleration and up to 2% for front-side acceleration ~depending on pulse length!.29 Figure 5 shows ion spectra from three-dimensional ~3-D! simulations 57 highlighting the different contributions of the front and rear mechanisms. Three-dimensional PIC simulations have also been carried out for the case in which light ions are contained
416

in a thin layer at the surface of a high-Z target.64 Such a layer can be realized in practice, for example, by coating the thin light ion layer onto a high-Z substrate. In this case a small number of light ions ~e.g., protons from hydrogen contaminants! gain energy in a time-independent electric field near the target surface, and their dynamics can be described in the test-particle approximation. This regime of ion acceleration takes on special significance for the problem of high-quality ion beam generation. The typical energy spectrum of laser-accelerated particles observed both in the experiments and in the computer simulations can be approximated by a quasi-thermal distribution with a cutoff at a maximum energy. The effective temperature that may be attributed to the fast ion beams is only within a factor of few from the maximum value of the particle energy. On the other hand, as will be discussed in Sec. IV, many applications require highquality proton beams, i.e., beams with sufficiently small energy spread, D E0E 1. Such a beam of laseraccelerated ions can be obtained using a double-layer target. As illustrated in Fig. 6, the use of a double-layer target produces fast proton beams with controlled quality and a narrow spectrum. In this scheme the target is made of two layers with ions of different electric charge and mass. The first ~front! layer consists of heavy ions with electric charge eZ i and mass m i , followed by a second ~rear! thin proton layer. Particle-in-cell simulations also indicate future perspectives for ion acceleration at the extremely high laser intensities that may be reachable in the future with pulses
FUSION SCIENCE AND TECHNOLOGY VOL. 49 APR. 2006

Borghesi et al.

FAST ION GENERATION BY LASER IRRADIATION OF SOLID TARGETS

Fig. 6. Computational results from Ref. 64. Distribution of the electric charge inside the computation region at ~a! t 40 and ~b! t 80 ~position of heavy @thick shell# and light ions @thin shell# is indicated by arrows, whereas the diffuse feature corresponds to electrons!.

in the multipetawatt power regime and beyond. An extreme ion acceleration regime, named the laser-piston mechanism, has recently been highlighted by 3-D PIC code simulations, carried out for extreme laser intensities

~up to ;10 23 W{cm 2 ! ~Refs. 39 and 63!. In this mechanism, the radiation pressure of the electromagnetic wave is directly converted into ion energy via the space-charge force related to the displacement of the electrons in a thin foil. In this regime the efficiency of the laser energy conversion into the fast ion energy can be very high. This acceleration scheme has a connection with a mechanism of ion acceleration proposed by Veksler 66,67 as well as with the snow plow acceleration mechanism suggested in Ref. 68. Formulated in the mid-1950s, Vekslers concept of the collective acceleration of ions in an electronion bunch moving in a strong electromagnetic wave had a great influence upon both particle accelerator technology and plasma physics. Up to now its direct realization was considered at moderate driving electromagnetic radiation intensities, when transverse instabilities impede the acceleration.69 In this scheme the transverse instabilities are suppressed or retarded because of the following. First, the plasma layers become quickly relativistic, during one or more laser wave periods, in the first stage of the acceleration. Because of relativistic effects the transverse instabilities in the laboratory frame grow more slowly than in the plasma reference frame. Second, the radiation pressure causes a stretching of the plasma mirror in the transverse direction, so the transverse instabilities can be retarded similarly to the slowing down of the Jeans instability in the theory of the expanding early universe.70,71 The simulations show that the foil is transformed into a cocoon where the laser pulse is almost confined ~Fig. 7!. The accelerated foil, which consists of the electron and ion layers, can be regarded as a relativistic plasma mirror copropagating with the laser pulse. The accelerated ions form a nearly flat thin plate with high density ~see Fig. 7b! and energies in the giga-electron-volt region. According to simulations for intensities ranging

Fig. 7. Computational results from Ref. 63. ~a! The ion density isosurface for n 8ncr ~a quarter removed to reveal the interior! at t 40 2p0v. ~b! The isosurface for n 2ncr, green gas for lower density at t 100 2p0v. The black curve shows the ion density along the laser pulse axis.
FUSION SCIENCE AND TECHNOLOGY VOL. 49 APR. 2006

417

Borghesi et al.

FAST ION GENERATION BY LASER IRRADIATION OF SOLID TARGETS

from 10 21 to 10 23 W{cm 2, the interaction exhibits a continuous transition from the acceleration regimes described previously to the laser-piston regime as the intensity increases. Although PIC codes have so far been the main numerical tool for ion acceleration investigations, recently published work has reported on simulations of ion acceleration from wire targets carried out using a different numerical method, namely, a gridless tree code employing a finite size particle approach.52 Although being purely electrostatic and not including electromagnetic wave propagation, the method may present advantages over PIC codes in terms of inclusion of collisional effects and the absence of geometrical boundaries.

III. EXPERIMENTAL RESULTS III.A. General Features of Laser-Driven Ion Beams The discovery of highly collimated proton beams with multi-mega-electron-volt energies was independently published by three research groups: Clarke et al.,1 Maksimchuk et al.,3 and Snavely et al.2 and Hatchett et al.30 In each of these experiments, a well-defined proton beam was observed with a roughly exponential spectrum and mean energy in the mega-electron-volt range and a high-energy cutoff in the 10- to 55-MeV range. The beam was generally emitted with a low divergence angle, with the most energetic protons having the lowest divergence angle, along the normal to the rear target surface. Following these first measurements, the process of laser-driven ion acceleration has been investigated by several experimental teams under very different physical conditions and using laser systems with different characteristics ~a review of the major laser facilities used in this area of research is provided in Ref. 72!. The laser intensities used for these studies range from a few times 10 17 W0cm 2 ~Refs. 73 and 74! up to 3 10 20 W0cm 2 ~Ref. 2!. A wide range of target thicknesses has been explored, ranging from a few wavelenghths ~Refs. 3 and 50!, up to hundreds of microns ~Refs. 2, 30, and 75 through 78! or even millimeters ~Refs. 75 and 76!; monolayer metallic or plastic targets ~Refs. 2, 30, 73, and 76! and C8H8 ~Refs. 65 and 79! as well as a range of doublelayer foils ~Refs. 3, 73, and 77 through 80! were irradiated with the laser. Ion emission has also been investigated from metallic wire targets with diameters of a few tens of microns ~Refs. 81 and 82!. Liquid droplets 83,84 have also been used. Energetic ions have been collected in the forward direction with respect to the laser and in the backward direction ~that is, on the same side of the target irradiated by the laser!, with the latter being in general less energetic than the former.
418

In these experiments, the energy spectra are generally measured by means of absolutely calibrated radiochromic film ~RCF! ~sensitive to X-rays, ions, and electrons!,85,86 nuclear activation measurements,87 CR-39 particle detectors 88 ~sensitive to ions with energies .100 keV0nucleon!, magnetic spectrometers ~from few hundred kilo-electron-volts up to 100 MeV!, and Thomson parabolas. RCF provides, with a high dynamic range, continuous spatial readout of the proton fluence, in coarsely resolved steps of proton energy by means of the range-energy relationship of the stopping power.2 It is preferentially sensitive to penetrating protons, which have a large specific energy loss and produce a high-contrast image. Electrons and X-rays generally appear as a diffuse low-intensity low-contrast background. Typical examples are shown in Fig. 8. As mentioned above, since protons have the highest charge-to-mass ratio, they are favored by the acceleration processes. The protons are present in the target as surface contaminants 4,78,89 or as compounds of the target itself or of the target coating. In a few cases,75,77,78 heating of the target was performed prior to the experiments in order to eliminate the hydrogen contaminants as much as possible and to obtain a better, controllable ion acceleration. In particular, removing the proton from the targets, or choosing H-free targets, the acceleration of heavier ions was favored. A selection of the published data on proton acceleration under relatively similar conditions of laser pulse duration, laser energy, laser intensity, and target conditions is shown in Figs. 9 and 10. Typical parameters of the proton beams obtained at some of the major laser facilities are also reported in Table I together with the parameters of the laser pulses used for these studies. Figure 9 shows the evolution of the recorded maximum proton energy as a function of either the pulse duration or laser irradiance. The data have been grouped so that similar conditions ~e.g., similar pulse durations! in experiments carried out in different laboratories could be compared. Figure 9a shows clearly that the maximum proton energy increases with the laser pulse duration, along roughly parallel lines for varying laser irradiances. This can also be seen in Fig. 9b, which shows the evolution of the maximum proton energy as a function of laser irradiance. Interestingly, we observe that for the longer pulse durations, the maximum proton energy increases as I 0.5 , whereas the increase seems to follow more closely a proportion to I for shorter laser pulses. Using PIC simulations, a change from a scaling proportional to I to a scaling proportional to I 0.5 has been observed as the laser irradiance passes from the subrelativistic domain ~i.e., I , 10 18 W{cm 2 {mm 2 ! to higher intensities.60 This can be understood if Emax @ Fpond since for I , 10 18 W{cm 2, Fpond @ I and for I . 10 18 W{cm 2, , . Fpond @ I 0.5 . In the case of Fig. 10b, the picture could be more complex because of differences in the laser contrast ratio
FUSION SCIENCE AND TECHNOLOGY VOL. 49 APR. 2006

Borghesi et al.

FAST ION GENERATION BY LASER IRRADIATION OF SOLID TARGETS

Fig. 8. ~a! Spectrum of laser-accelerated protons obtained by a magnetic spectrometer at different angles from the rear surface of a 100-mm Al foil irradiated by a petawatt laser pulse. ~b! through ~e! RCFs for the same experiment. Data from Ref. 2. The energy responses of the RCFs to the proton flux are shown in ~f !.

TABLE I Parameters of Some of the Laser Systems Used for High-Energy ~.1-MeV! Proton Beam Acceleration Experiments and Typical Parameters of the Proton Beams Produced as Reported in the Quoted References* Laser Pulse Duration ~ps! 0.04 0.06 0.06 0.1 0.15 0.32 0.4 0.45 0.5 0.7 1 Laser Energy ~J! 0.8 0.1 0.2 10 0.7 30 5 25 500 400 90 Laser Intensity ~W0cm 2 ! 6 0.7 1 7 6 5 5 3 2 10 19 10 19 to 10 19 10 18 10 20 10 19 10 19 10 19 10 18 10 20 10 20 10 20 Target Thickness ~ mm! 6 5 20 3 10 20 12.5 5 to 25 100 100 10 Maximum Proton Energy ~MeV! 8 1.2 1.5 24 2.5 20 12 10 58 44 36 Conversion Efficiency into Protons 0.2% 0.7% 1% 1% ~. 5 MeV! 12% ~.10 MeV! 7% ~.13 MeV! 5%

Laser System LOA CRIEPI, Tokyo ASTRA JanUSP MPQ LULI 100 TW CUOS GEKKO NOVA PW RAL PW RAL Vulcan

Reference 90 96, 99 97 50 95 82, 100 94 75 2 92, 101 76

T ~MeV! 0.2 3.2 3 3.4 6 4.5

*Some of the information was not provided in the references.


FUSION SCIENCE AND TECHNOLOGY VOL. 49 APR. 2006

419

Borghesi et al.

FAST ION GENERATION BY LASER IRRADIATION OF SOLID TARGETS

Fig. 9. ~a! Maximum proton energy from laser-irradiated thin ~5 to 10 mm, except for the data points marked RAL PW and Nova PW where the thickness is 100 mm! metal ~Al or Cu! targets for different experiments as a function of the laser pulse duration and for three different ranges of laser irradiances. References are as follows: LOA ~Ref. 90!, JanUSP ~Ref. 50!, LULI ~Ref. 91!, RAL PW ~Ref. 92!, Nova PW ~Ref. 2!, RAL Vulcan ~Refs. 76 and 93!, Osaka ~Ref. 65!, CUOS ~Ref. 94!, MPQ ~Ref. 95!, Tokyo ~Ref. 96!, ASTRA ~Ref. 97!, and Yokohama ~Ref. 98!. ~b! Same but as a function of the laser irradiance and for three ranges of pulse durations. The two dashed lines are trend lines proportional to I and I 0.5, respectively.

and also because of different behavior of the laser propagation in the underdense plasma in front of the targets for different laser irradiances. A dedicated experiment using exactly the same conditions would be needed to obtain a clearer picture and an understanding of the un420

derlying processes. Figure 10 shows that not only the maximum proton energy but also the efficiency of the acceleration process increases with the laser pulse duration and the laser irradiance. Here, we have chosen not to calculate a global energy conversion from laser to
FUSION SCIENCE AND TECHNOLOGY VOL. 49 APR. 2006

Borghesi et al.

FAST ION GENERATION BY LASER IRRADIATION OF SOLID TARGETS

laser facility, i.e., using the highest achieved peak power and laser energy. A well-collimated, forward-directed picosecond bunch of 3 10 13 protons with energies up to ;60 MeV, and an energy content of up to 48 J ~12% of the laser energy! was generated as a result of the interaction of 400 J, in 1-mm, 0.5-ps laser pulses, focused in an 8- to 9-mm focal spot, producing an intensity up to 3 10 20 W0cm 2, with a 100-mm-thick CH target.2,30 Similar energies, but five times fewer protons, were observed using Au targets.2,30 III.B. Experiments on Acceleration Mechanisms There have been conflicting interpretations of the available experimental evidence for what concerns the location of the accelerating mechanisms, and two different models of the origin of these protons have emerged: one suggesting proton acceleration from the front 1,3,8 and the other the rear 2,30,80 of the target. The annular angular ring pattern of proton emission, observed in a number of experiments, has been variously interpreted as supporting acceleration from the front of the target in Clark et al.1 or from the rear in Murakami et al.75 and in 3-D PIC code simulations by Pukhov.57 Experiments on the Nova Petawatt laser 2,30 were explained in terms of electrostatic sheath at the rear surface of the target. In particular, targets that had a wedged rear surface were shown to produce two proton beams, one from each surface. This was interpreted as consistent with the rear acceleration mechanism occurring at the rear surfaces of the wedge. A number of experiments aimed at isolating the generating mechanism have been carried out. For instance, ion acceleration from laser irradiation of water droplets has been described by Karsch et al.83 From time-of-flight neutron measurements, these authors concluded that sheath acceleration occurring at the rear surface of each droplet was the dominant mechanism of ion acceleration in this experiment. Mackinnon et al.80 observed eradication of the high-energy component of the proton beam when introducing a small preformed plasma scale length on the rear of the target, as is consistent with a rear-surface acceleration mechanism.31 Conversely, Nemoto et al.8 clearly observed front-surface accelerated energetic deuterons using a nuclear activation technique; the same experimental data have also been reported in Ref. 3 along with different results obtained with a frequency-doubled laser. Experiments using heated targets to drive off surface contaminants have also investigated the source proton production. Zepf et al.77 measured proton flux reduction by factors of 2 to 3 when targets with plastic layers were heated above their melt temperature. The conclusion inferred from this experiment was that the most energetic protons originated from the front of the target. Hegelich et al.78 also used resistive heating to drive off contaminant hydrocarbon layers on targets coated
421

Fig. 10. ~a! Number of accelerated protons in a 1-MeV energy bin at 10 MeV of energy as a function of the laser pulse duration and for two different ranges of laser irradiances. The references the data are extracted from are the same as in Fig. 8. ~b! Same as ~a! but as a function of the laser irradiance and for two ranges of pulse durations.

protons since the data reported in the various experiments differ in terms of spectral energy range so that a consistent number cannot be tabulated and compared for the various experiments. The point of comparison has therefore been chosen to be the number of protons with 10 MeV of energy accelerated in the forward direction. We can assume that when this number increases, the total number of protons, and hence the conversion efficiency, will increase. The present published record in terms of highest proton energy has been achieved using the Nova Petawatt
FUSION SCIENCE AND TECHNOLOGY VOL. 49 APR. 2006

Borghesi et al.

FAST ION GENERATION BY LASER IRRADIATION OF SOLID TARGETS

Fig. 11. ~a! Schematic of a Gaussian sheath formation on a solid target rear surface with the image of the laser focal spot on target at full power; protons are accelerated normal to iso-density contours in the sheath, defining a maximum emission angle that corresponds to the inflexion point in the sheath. ~b! A sharp boundary is consistently observed, at each proton energy, in the angular distribution of the protons observed on an RCF: The experiment uses a 48-mm Au target irradiated by a 350-fs, I ; 2 10 19 W0cm 2 laser pulse at normal incidence.49 ~c! An electrostatic model 49 for the acceleration of protons that uses a Gaussian sheath reproduces well the experimental observation of ~b! as shown in ~d!.

with calcium flouride ~CaF2 ! layers. In this case the targets were heated enough to drive off all detectable sources of contaminants but remained below the melting temperature of the CaF2 layer. The light ion spectrum produced by these heated targets indicated that a rear-surface mechanism was responsible for accelerating ions to multimega-electron-volt energies in this experiment. Most recently, another method of removing contaminants has used an ion gun to sputter off oxide layers present on both target surfaces continuously up until the laser shot is fired, as described by Allen et al.89 The advantage of this technique is that the ion gun action can be localized so that either the front or rear of the target are cleaned, thus allowing the experimenter to isolate the location of the proton beam. When the front surface of the target was cleaned, no discernable effect was observed on the proton beam; however, when the rear surface was cleaned, the proton beam was essentially eradicated. This experiment showed conclusively that the rear target sheath acceleration was the dominant mechanism. The same conclusion was reached in another experiment 91 that used nuclear activation measurements of laser-accelerated deuterons ~deposited on a single target surface! and protons. This allowed direct and quantitative comparison of front- and rear-side acceleration under identical laser conditions. It was shown that for protons .3.5 MeV, front-side acceleration accounts for ,3% of the total energy of the accelerated protons, the rest being provided by rear-side acceleration. III.C. Emission Characteristics of Proton Beams As mentioned in Sec. III.B, most of the available experimental evidence indicates that the main source of high-energy protons originates from rear-surface accel422

eration. In this section, we therefore concentrate on the spatial and angular characteristics of the rear-surface accelerated protons, the front-surface accelerated ones being, as mentioned before, widely divergent and poorly laminar since they originate from the laser holebored critical interface. Some of the unique and most interesting characteristics of the laser-accelerated proton beams are discussed in this section. Since the ions are accelerated by the electron sheath on the target rear surface, their spatial and angular characteristics will be determined by the electron sheath spatial distribution. Such distribution depends not only on the target rear-surface condition but also on the distribution of the electrons generated at the laser-target frontsurface interface and on their subsequent evolution in the target. Therefore, the protons spatial and angular characteristics will depend on the target material, the target surface roughness, the target shape, and the laser focal distribution on the target. In the case of controlled, ideal conditions for the laser and the target ~i.e., a conductor target, a mirrorlike smooth target surface, and a smooth laser focal distribution!, it is observed, as shown in Fig. 11, that the proton beam exhibits a smooth angular distribution 49 with a sharp boundary.2 The proton angular distribution is here observed on RCF dosimetry media.85 The spatial distribution of the protons in a given RCF layer gives the angular emission pattern within a specific proton energy range. If one imagines that the electron sheath follows a generic bell-shaped spatial distribution, the existence of a sharp angular boundary in the proton angular distribution is easily understood. Indeed, protons will be accelerated everywhere in the sheath, normal to the local iso-density contour. As sketched in Fig. 11a, the central region, as well as the wings of the sheath, will produce
FUSION SCIENCE AND TECHNOLOGY VOL. 49 APR. 2006

Borghesi et al.

FAST ION GENERATION BY LASER IRRADIATION OF SOLID TARGETS

protons normal to the surface while intermediate zones will accelerate protons at an angle. The presence of an inflexion point in the sheath will therefore produce a maximum angle of acceleration. A simple electrostatic model of acceleration assuming a bell-shaped sheath 49 reproduces well this behavior, as illustrated in Figs. 11c and 11d. The precise angle of the boundary depends on the particular parameters of the sheath. Deviations from the ideal conditions mentioned above will perturb the proton angular distribution. A roughened target surface will induce spatial modulations in the electron sheath and therefore local random orientation of the protons during the initial acceleration that will result in local angular modulations in the proton dose, i.e., filamentlike structures.82 Similarly, an insulator target or a modulated laser focal spot will produce the same result. Indeed, even starting from a smooth distribution of fast electrons at the laser-target interface, it has been experimentally shown that the transport is disrupted in insulators 49 : Modulations appear in the electron beam, which forms the sheath when reaching the target rear surface, inducing filaments in the proton angular distribution as in the case of a roughened target surface. Fast electron generation and transport in solid matter are a complex topic that is still not well understood ~see, e.g., Ref. 102!. However, it appears that at least for conductor targets, the electron distribution at the source can be imprinted by the laser focal distribution.49 A modulated laser irradiating a conductor target will produce a modulated electron beam. If this beam is transported smoothly through the conductor target, it will produce a modulated sheath resulting in a filamented proton angular distribution. Small modulations in the laser focus are difficult to diagnose and can vary shot to shot unless there is a dynamic correction for the laser wave front. They are therefore a likely explanation for the transverse profile dishomogeneities frequently observed in proton beams.1,2

A modulation of the proton beam angular distribution can be obtained purposefully by micromachining on the surface shallow grooves like on an optical grating. This produces a periodic modulation of the beam angular envelope, as observed in Fig. 12 ~Refs. 49 and 51!. As protons are first accelerated normal to the surface, the grooves on the surface induce a modulation of the takeoff angle. Because of the global sheath expansion, an overall near-linear divergence is added to this initially imprinted angular modulation of the beam. Projected on a film stack far away, this results in a modulation of the proton dose, as observed. Using such modulations of the beam intensity, it is possible to image the proton-emitting surface and thus to measure directly the source size. In the example of Fig. 12, the measured diameter of the emission zone is of the order of 50 mm @full-width at half-maximum ~FWHM!# at 4.5 MeV and is ;30 mm ~FWHM! for .9-MeV protons. Alternate ways to determine the source size that rely on beam trajectory reconstruction by projection of patterned objects 103 or knife edges 104 are possible. Other interesting information that can be retrieved readily from Fig. 12 is the dependence of the divergence angle versus the source position of the accelerated protons. One observes easily that within the central portion of the beam, the phase-space correlation is almost exactly linear. This remarkable result suggests that the relativistic electron sheath has a nearly Gaussian radial distribution in its density profile. Indeed, as the accelerating field is E F ~kThot 0e!~nhot 0nhot !, where nhot ; exp@eF~r!0kThot # is the Boltzmann distribution of the hot electrons in the sheath, a bell-shaped sheath will result in a nearly linear relation between the radial position and the radial electric field in the sheath. This confirms the assumption of a bell-shaped sheath used to explain the result of Fig. 11. A further confirmation has been recently provided by the detection of the field

Fig. 12. ~a! Experimental angular distribution of 10-MeV protons accelerated from a thin Al solid density foil of 60-mm thickness irradiated by a laser with a Gaussian intensity distribution. The rear surface of the foil has sinusoidal microgrooves in one direction. ~b! Normalized lateral momentum distribution obtained with an effective simulation.107 ~Figure extracted from Ref. 51.!
FUSION SCIENCE AND TECHNOLOGY VOL. 49 APR. 2006

423

Borghesi et al.

FAST ION GENERATION BY LASER IRRADIATION OF SOLID TARGETS

associated with a bell-shaped sheath driving the expansion of a laser-driven proton beam.105 These measurements, obtained with high spatial and temporal resolution, have been performed by using the proton probing techniques that will be described in Sec. IV.A. The proton beam angular distribution can also be modified by curving the target surface in order to focus down the protons to a tight spot. This has been theoretically predicted 31,58 and experimentally demonstrated 106 via enhanced heating of a secondary target, as will be discussed in Sec. IV.B. Figure 13 shows computational results from 3-D PIC simulations of the acceleration of a beam of protons from a spherical shell target,58 clearly indicating the focusing effect due to the spherically deformed Debye sheath. Smooth or filamented, the proton beam has an angular envelope that decreases with energy.2,107 Using the aforementioned technique of directly imaging the protonemitting surface by imprinting grooves on the target rear surface, it has been shown 107 that for protons .4.5 MeV, the decrease in the angular envelope with energy is due to a decrease of the emitting zone. Such a decrease of the emission zone is expected for a transversally bell-shaped electron density distribution. In such a sheath, the highest-

energy protons are accelerated in the central, highdensity portion of the sheath, whereas lower energies come from the wings of the sheath distribution and thus are emitted at larger angle.49 Ring structures have also been observed in the proton angular distributions. These are of two distinct types: a ring that appears inside the proton beam for low proton energies 108 and a ring that appears on the edge of the proton beam at all energies.1 The latter ring is observed directly on the CR-39 nuclear track detector, as can be seen in Fig. 10 of Ref. 77. This feature is not visible in raw RCF data but can be revealed by a subtractive analysis method 109 that allows the retrieval of the angularly resolved spectrum of the beam.110 The ring at the proton beam edge appears to decrease with the proton energy, and although it has been attributed to the effect of magnetic fields inside the target,1 it may be also due to the fact, mentioned above, that the proton angular envelope decreases with energy.57 Regarding the other ring that appears inside the proton beam for low proton energies, it is due to proton acceleration in the wings of the sheath: The lower the proton energy, the larger will be the acceleration zone and the larger the emission angle0divergence, as seen above. However, when the proton energy is low

Fig. 13. Three-dimensional PIC simulation of proton acceleration from a thin, dense, spherical plasma shell ~22-mm radius and 5-mm thickness! by a 70-fs petawatt laser pulse: ~a! 3-D proton distribution ;400 fs after the interaction ~subbox boundaries: 5 , x , 25 mm, 5 , y , 25 mm, and 4 , z , 32 mm!; ~b! and ~c! cross section of the fast proton beam density at ~a! z 19 mm and ~b! z 26 mm. The geometric center of the spherical target is located at z 30 mm. ~Adapted from Ref. 58.! 424
FUSION SCIENCE AND TECHNOLOGY VOL. 49 APR. 2006

Borghesi et al.

FAST ION GENERATION BY LASER IRRADIATION OF SOLID TARGETS

enough, the emission zone reaches the wings of the sheath where the emission angle reduces again. This produces an accumulation of dose, in the shape of a ring. Laminarity is another remarkable characteristic of these proton beams. The degree of laminarity of chargedparticle beams can be expressed in terms of their emittance, which is proportional to the area of the bounding ellipsoid of the distribution of particles in phase-space. The highest-quality ion beams have the lowest values of transverse and longitudinal emittance, indicating a low effective transverse ion temperature and a high degree of angle-space and time-energy correlation. Using the groove imaging technique discussed previously, it is possible to fully reconstruct the transverse phase-space and thus determine the transverse emittance.107 In this way, it has been experimentally shown that for protons of up to 10 MeV, the transverse emittance is as low as 0.004 mm{mrad, i.e., 100-fold better than typical radiofrequency accelerators and at a substantially higher ion current ~kiloampere range!. Indeed, for typical proton accelerators ~e.g., the CERN SPS!, the emittance from the proton injector linac is ;1 mm{mrad ~normalized root-mean-square! and up to 3.5 mm{mrad within the synchrotron, with 10 11 protons per bunch. It has also been shown 107 that the removal of the co-moving electrons after 1 cm of the quasi-neutral ~protons and electrons! beam expansion did not significantly increase the measured proton transverse emittance. This last observation is important since in order to take advantage of the exceptionally small proton beam emittance in future applications, e.g., to capture them into a postaccelerator, removal of the co-moving electrons without significantly perturbing the protons is crucial. The ultralow measured emittance stems from the extremely strong, transient acceleration that takes place from a cold, initially unperturbed surface and from the fact that during much of the acceleration the proton space charge is neutralized by the co-moving hot electrons. Such an acceleration process represents a potentially near-ideal kind of ion diode as compared either to the ion beams generated from plasma plumes,4 i.e., from the laser-heated turbulent plasma on the front side of the foils, or to the conventional plasma discharge ion sources used in accelerators. It should be noted that transverse emittance can also be determined by alternate means, like using projection, on a far distant film, of objects such as knife edges 111 or meshes 103 placed in the ion beam path, although these methods are less precise. In terms of applications using the ion beam as a projection source, having a low-emittance beam is equivalent to projecting from a pointlike source located in front of the target, with much smaller transverse extent than the ion-emitting region on the target surface.103 This allows high resolution ~point projection! in chargedparticle radiography or ion patterned lithography. Regarding the longitudinal emittance, the energy spread of the laser-accelerated proton beam is large, from zero to
FUSION SCIENCE AND TECHNOLOGY VOL. 49 APR. 2006

tens of mega-electron-volts; however, because of the extremely short duration of the accelerating field ~,10 ps!, the longitudinal phase-space energy-time product is probably ,10 4 eV{s. PIC simulations show that longitudinally the acceleration is extremely laminar 31 in the sense that the spread of proton energies in a given longitudinal slice is very small. A good longitudinal velocity chirp of the beam is important since it could allow producing monochromatic beams by coupling the beam to the field gradient of a postaccelerator. III.D. Heavy Ion Beams As mentioned previously, regardless of the material used in laser-ion acceleration experiments, protons have always been observed as the dominant ion species.2 The origin of the protons is surface contaminants, e.g., water vapor or hydrocarbons, providing a layer containing protons at the target surface. Because of the low ionization potential and high charge-to-mass ratio, hydrogen is among the first ion species produced and most efficiently accelerated. The cloud of accelerated protons then screens the electric field generated by the electrons for all the other ion species. The key for the efficient acceleration of heavy ions is the removal of any proton or light ion contaminants. Heavy ions have also been observed backward to the laser originating from the front side.108 But, these ions, even though high in energy ~4 MeV0u!, do not have the high-quality beam characteristic as those in the laser direction from the rear surface. For the latter, recent experiments have demonstrated heavy ion acceleration up to .5 MeV0u, which corresponds to ion energies usually available at the end of conventional accelerators of hundreds of meters in length.78 First attempts to remove the hydrogen contaminants used resistively heated Al targets up to temperatures of a few hundred degrees. Even a partial removal of the hydrogenous contaminants strongly enhanced the acceleration of heavier ions ~i.e., carbon!.82 A reduction by a factor of 10 of the hydrogenous contaminants increased the energy of the carbon ions by a factor of 2.5 and their number by two orders of magnitude. Using tungsten as a thermally stable target resist and coating the rear surface with the material of interest, the target could be heated to .12008C. Spectra from such targets show no accelerated protons at all and a strongly increased heavy ion component. The maximum energy could be enhanced by a factor of 5 compared to Al targets and the conversion efficiency by a factor of 10 ~Ref. 78!. In cases where ohmic heating of target materials of interest is not possible because of a low evaporation point, laser heating has been demonstrated to be an appropriate option to remove the contamination layers. In this case the intensity of the laser heating the rear surface and evaporating the proton layers and its timing, with respect to the short pulse, have to be matched carefully.
425

Borghesi et al.

FAST ION GENERATION BY LASER IRRADIATION OF SOLID TARGETS

The energy spectrum of the heavy ions and their charge state distribution also provide detailed information about the accelerating electric field at the rear surface. It was shown that in a typical experiment collisional ionization and recombination in flight are negligible, and so the detected charge states directly image the electric field strength due to the field ionization process. The results that have been obtained match very well the estimated field strength, also predicted by theory,30 and range from 10 11 V0m up to a few 10 12 V0m. The accelerating field deduced from the ion acceleration is consistent with the observed proton energies with nonheated targets. For example, in typical experiments fluorine ions were accelerated up to 100 MeV, i.e., .5 MeV0nucleon at a maximum charge state of 7 ~see Fig. 14!. This corresponds to an electric field of 2 TV0m, which would have accelerated protons, if present, to energies up to 25 MeV, i.e., the maximum energies found in experiments with nonheated targets under similar experimental conditions. The conversion efficiency in heavy ions is very high. Similar to the results obtained for proton beams, conversion efficiencies of up to 4% from laser to ion beam energy have been measured. Since the accelerating mechanism for protons and heavy ions is the same, excellent beam quality can be expected also for the heavy ions. Experiments using structured targets have recently provided measurements of the emittance of heavy ion beams, indicating a beam quality comparable to that of the laser-accelerated proton beams and superior to conventionally accelerated ion beams. The future prospects for laser-accelerated heavy ion beam applications are as promising as those for protons. Recently, the initiation of fusion reactions of laser-accelerated heavy ions with secondary target material has been demonstrated result-

ing in a variety of short-lived radioisotopes, some of them relevant for medical applications.112 Furthermore, because of the high intensity, short pulse length, excellent beam quality, and directionality, the idea of light ion fast ignition has also been developed.113 One of the most promising applications of laser-driven ion beams is their use as next-generation ion sources for conventional accelerators. Thanks to the high particle number ~more than 10 12 ions per pulse! and the excellent beam quality, this might greatly enhance the luminosity of particle accelerators. Laser-accelerated heavy ions exhibit, as the protons, a broad energy distribution up to a characteristic cutoff energy. If matched appropriately, the ion content in a single charge state can be as high as 80% of the total number of ions. The challenge for applications as nextgeneration ion sources lies mainly in designing a matching section into a conventional accelerator that can accept high beam currents and simultaneously rotate the longitudinal phase-space in order to achieve a monochromatic beam.

IV. APPLICATIONS In this section, the main applications of laser-driven ion beams ~either achieved or proposed! will be reviewed, together with a description of the requirements of each application in terms of beam parameters. A summary of the beam properties required by some of these applications is given in Table II. IV.A. Radiography and Imaging with Laser-Accelerated Protons The use of ion beams, and particularly proton beams, for radiographic applications was first proposed in the 1960s. Quasi-monochromatic beams of ions from conventional accelerators have been used for detecting areal density variations in samples, with spatial resolution. This was achieved by exploiting the energy deposition properties of the particles in matter through a number of different methods: differential stopping radiography; marginal range radiography, which is based on the enhanced sensitivity of ions to areal density variations toward the end of their range; and scattering radiography, which exploits the intensity pattern created via scattering in the ion beam intensity cross section for samples with thickness smaller than the stopping range. Radiography with very high energy protons ~;1 to 10 GeV! is being developed as a tool for weapon testing.114 Ion beams from accelerators have also been employed on some occasions for electric field measurements in plasmas via the detection of the proton deflection, e.g., in Ref. 115. In practice, the difficulties and high cost involved in coupling externally produced particle beams of sufficiently high energy to laser-plasma experiments ~or indeed magnetic confinement experiments! and the relatively long duration of ion
FUSION SCIENCE AND TECHNOLOGY VOL. 49 APR. 2006

Fig. 14. F 7 spectra from heated and unheated targets ~from Ref. 78!. More than 5 MeV0nucleon are achieved for F 7 ions when the targets are heated. 426

Borghesi et al.

FAST ION GENERATION BY LASER IRRADIATION OF SOLID TARGETS

TABLE II Parametric Requirements of Some of the Applications of Laser-Driven Proton Beams Application Radiography ~density detection! Beam Requirements Low emittance Short duration High energy ~for dense matter probing! Low emittance Short duration Broadband for multiframing Low emittance0focusability High flux Spectral tailoring Short duration Low emittance0focusability High flux Small energy spread High energy ~50 to 250 MeV! High repetition ~duty cycle! Removal of co-moving electrons Monochromatic beam High repetition High current Removal of co-moving electrons Ultralow emittance High flux Small energy spread High repetition High energy High flux High repetition

Imaging0deflectometry ~electromagnetic field mapping! Isochoric heating ~warm dense matter!

Fast ignition Protontherapy

Industrial application ~implantation, lithography!

Injection into accelerators

PET

pulses produced from conventional accelerators has limited the application of such diagnostic techniques. The unique properties of protons from high-intensity laser-matter interactions, particularly in terms of spatial quality and temporal duration, have opened up a totally new area of application of proton probing0proton radiography. Several experiments have been carried out in which laser-driven proton beams have been employed as a backlighter for static and dynamic target assemblies, in some cases a secondary target irradiated by a separate laser pulse. As seen in Sec. III.C, the protons emitted from a laser-irradiated foil can be described as emitted from a virtual, pointlike source located in front of the target.103 A point-projection imaging scheme is therefore automatically achieved. The magnification of the system is determined by M 1 L0h, with L and h, respectively,
FUSION SCIENCE AND TECHNOLOGY VOL. 49 APR. 2006

the object-to-detector and source-to-object distances. Density variations in the target probed can be detected via modifications of the proton beam density cross section, caused by differential stopping of the ions in the case of thick targets or by scattering in the case of thin targets. Similarly, electric or magnetic fields in the sample region can be revealed by the proton deflection and the associated modifications in the proton density pattern. The detector employed in proton radiography0 probing experiments consists mainly of RCF films, often arranged in a multilayer package. The multilayered arrangement offers the possibility of energy-resolved measurements despite the beams broad energy spectrum as in each layer the dose deposited is mainly due to the protons within a narrow range dE, which reach the Bragg peak in the layer. Typical values for RCF packages and protons with 5- to 10-MeV energies are dE ; 0.5 MeV. In experiments with low proton fluxes, track detectors ~CR39! are a viable alternative to RCF, and the proton density distribution is obtained directly from the density of tracks after etching. As mentioned above, backlighting with laser-driven protons has intrinsically high spatial resolution, which is determined by the size d of the virtual proton source and the width ds of the point spread function of the detector ~mainly due to scattering near the end of the proton range!. For large magnification, the spatial resolution is a few microns. However, resolution degradation due to multiple scattering affects this technique when the object probed is not very thin, basically increasing the size of the effective proton source. Energy dispersion provides the technique with an intrinsic multiframe capability. In fact, since the sample to be probed is situated at a finite distance from the source, protons with different energies reach it at different times. As the detector performs spectral selection, each RCF layer contains, in first approximation, information pertaining to a particular time. Depending on the experimental conditions, two-dimensional proton deflection map frames spanning up to 100 ps can be obtained in a single shot. The ultimate limit of the temporal resolution is given by the duration of the proton burst t at the source, which is of the order of the laser pulse duration. However, other effects are also important: The finite energy resolution of the detector layers and the finite transit time of the protons through the region where the fields are present normally limit the resolution to a few picoseconds. Several radiographic applications of laser-produced protons have been reported to date. Roth et al.82 have demonstrated proton radiography of macroscopic, thick compound targets using laser-produced protons with energy of 5 to 10 MeV, in a projection arrangement with M ; 1. Because of proton stopping, the RCF exposure provided a negative image of the areal density of the target. Radiography of thick masks with large magnification and approximately micron resolution has been shown at Los Alamos National Laboratory employing
427

Borghesi et al.

FAST ION GENERATION BY LASER IRRADIATION OF SOLID TARGETS

sub-mega-electron-volt protons produced at moderate irradiance.116 Radiography of thin objects, i.e., with areal density much smaller than the proton range, has also been demonstrated.103 Because of multiple small angle scattering by the ions inside the object under test, protons that propagate through the object suffer a larger lateral spread compared to protons that propagate in vacuum. At the edges of the object, the difference in the lateral spreads produces a perturbation in the beam intensity profile. If the transverse size of the object is larger than the perturbation width, then the perturbations due to different edges do not superimpose, and only the contour of the object is visible in the image. If instead the transverse size of the object is smaller than the perturbations width, then the perturbations due to two adjacent edges can superimpose producing deep minima in the intensity profile. Radio-

graphs of thin objects ~e.g., 5-mm Cu meshes irradiated by 15-MeV protons, or 7-mm CH spherical shells; see Fig. 15a! have been obtained by this technique.103,117 Density diagnosis via proton radiography has potential application in ICF. Protons of 50 to 100 MeV produced by a petawatt laser would be sufficiently energetic to propagate through cold, compressed National Ignition Facility cores. A preliminary test studying the compression of empty CH shells under multibeam isotropic irradiation at the moderate irradiance of 10 13 W0cm 2 has been carried out in an experiment at the Rutherford Appleton Laboratory.117 Radiographs of the target at various stages of compression were obtained ~see Fig. 15!. Modeling of proton propagation through target and detector using Monte Carlo codes permits the retrieval of density and core size at maximum compression ~3 g0cm 3 and 80 mm! in good agreement with hydrodynamic

Fig. 15. ~a! A 7-MeV proton radiograph of undriven 500-mm-diam, 7-mm-wall-thickness shell. ~b! Proton radiograph taken of highly asymmetric implosion, caused by mistimed heater beams. Beams from lower left side of figure are 1 to 2 ns earlier than beams from upper right. ~c! Proton radiograph, in 7-MeV protons, of a 500-mm-diam microballoon with a 3-mm wall at a time close to stagnation. The core is slightly elliptical and has assembled below the center point of the shell. ~d! Radial lineout taken through center of proton radiograph data and Monte Carlo simulation output for fixed peak density and varying core size ~FWHM!. 428
FUSION SCIENCE AND TECHNOLOGY VOL. 49 APR. 2006

Borghesi et al.

FAST ION GENERATION BY LASER IRRADIATION OF SOLID TARGETS

simulations. Preliminary work on experiments aiming to use protons as a shock diagnostic in laser-irradiated solids has also been carried out.117119 Probably the most important applications to date of proton probing are related to the unique capability of this technique to detect electrostatic fields in plasmas.120,121 This has made possible obtaining for the first time direct information on electric fields arising through a number of laser-plasma interaction processes.117,122,123 The high temporal resolution is here fundamental in allowing the detection of highly transient fields following short-pulse interaction. When the protons cross a region with a nonzero electric field, they are deflected by the transverse component E4 of the field. The proton transverse deflection at the proton detector plane is equal to
b

Dr4 where 2 m p vp 02 e b L

eL
0

2 ~E4 0m p vp ! dl ,

proton kinetic energy its charge distance over which the field is present distance from the object to the detector.

As a consequence of the deflections, the proton beam cross-section profile undergoes variations showing local modulations in the proton density. Assuming the proton density modulation to be small, dn0n 0 1, where n 0 and dn are, respectively, the unperturbed proton density and proton density modulation at the detector plane, we obtain dn0n 0 6div~Dr4 !60M, where M is the geometrical magnification. The value of the electric field amplitude and spatial scale can then be determined if a given functional dependence of E4 can be inferred a priori, e.g., from theoretical or geometrical considerations. Thin meshes inserted in the beam ~e.g., between the proton source and the object! are sometimes used as markers of the different parts of the proton beam cross sections, in a proper proton deflectometry arrangement particularly suited to revealing relatively large-scale

fields.117 As mentioned earlier, the meshes impress a modulation pattern in the beam before propagating through the electric field configuration to be probed. The beam is effectively divided in a series of beamlets, and their deflection Dr4 can be obtained directly from the deflection of the impressed pattern. A proton moir technique employing two grids to generate a set of moir fringes has also been proposed as a way to increase the sensitivity to small electric fields.124,125 A general analysis method, applicable to both proton imaging and proton deflectometry data, consists of using particle-tracing codes to follow the propagation of the protons through a given 3-D field structure, which can be modified iteratively until the computational proton profile reproduces the experimental ones. State-of-the-art tracers allow realistic simulations including experimental proton spectrum and emission geometry, as well as detector response. Typical data obtained with this technique are shown in Fig. 16. The data show, with picosecond resolution, the evolution of the electric fields surrounding a 50-mm Ta wire irradiated by the Vulcan, 1-ps, 25-TW pulse.122 The RCF layers shown ~second layers of the radiochromic stacks, corresponding to proton energies of 6 to 7 MeV! have been obtained in separate shots for different delays between the two chirped pulse amplification pulses ~one interacts with the wire; the other generates the protons used to radiograph!. When the proton probe arrived on target before the interaction pulse ~Fig. 16a!, only the shadow of the Ta wire, thick enough to slow down and scatter the protons, is visible, with some small effect visible in the interaction region due to preplasma present ahead of the interaction. However, when the probe is roughly coincident with the interaction ~Fig. 16b!, a dramatic effect is observed, with the protons being deflected away from the surface of the wire, which charges up because of hot electron expulsion. At the same time an ion front appears to be driven from the target, as also observed on irradiated foils.105 The charge is seen to decay in a few tens of picoseconds, as filamentary structures ~horizontal striations! appear, likely to be associated with the development of an electromagnetic

Fig. 16. Proton projection images of a 50-mm Ta wire for different interaction-probing delays: ~a! 12 ps, ~b! 2 ps, ~c! 8 ps, ~d! 18 ps, and ~e! 30 ps ~the absolute timing precision was ;5 ps!. The energy of the protons employed was 6 to 7 MeV.
FUSION SCIENCE AND TECHNOLOGY VOL. 49 APR. 2006

429

Borghesi et al.

FAST ION GENERATION BY LASER IRRADIATION OF SOLID TARGETS

instability. From the deflection surface, fields of ;10 10 V0m are deduced. Evidence of an ultrafast current flowing through the wire and progressively neutralizing the charge lost is also provided by the data.122 The wire evolution under these interaction conditions has been described as a laser-driven Z-pinch. 126 Beside fields associated with fast electron dynamics, the technique has provided novel information ~via their associated electrostatic field! on late-time ionic structures developing from nonlinear phenomena after highintensity interactions, including ~a! the first observations of postsolitons, arising from merging of solitons created in the wake of the laser pulse in near-critical plasmas,123 and ~b! the first observation of a quasi-stationary ionic modulation arising from a two-stream instability in the wake of a 30-fs pulse propagating in a tenuous plasma.127 Application of the technique to nanosecond-produced plasmas of ICF interest has also yielded interesting results. Filamentary structures tentatively associated with electron transport instabilities have been seen in plasmas produced by moderate irradiation ~10 13 to 10 14 W0cm 2 ! at the Rutherford Appleton Laboratory.120 In experiments carried out at Laboratoire pour lUtilisation des Lasers Intenses ~LULI!, electric fields associated with the pressure gradients have been detected for the first time in plasmas produced by 10 15 W0cm 2 irradiation of metal wires and foils by using the deflectometry technique.117 By particle-tracing analysis of the pattern, fields of the order of 10 9 V0m have been obtained. LASNEX simulations, combined with particle tracing through the field structure predicted, reproduce in magnitude and geometry the experimental observations. The magnetic field can also deflect protons, and there is some scope for the application of proton probing techniques to this purpose. IV.B. Proton Heating The high-energy flux and short temporal duration of laser-generated proton beams has recently led researchers to investigate their capability for heating solid matter to high-temperature plasma states. The heating of materials to a plasma state can be achieved by many means; however, for studies of fundamental properties such as the equation of state or opacity, it is often desirable to create large-area, or large-volume, plasmas in a uniform, single temperature and density state. Heating of materials with ions has already been demonstrated with accelerator-based and electrical-pulsed ion sources. The heavy ion synchroton at GSI Darmstadt produces more than 10 10 ions ~Ar 18 ! in a 250-ns pulse capable of heating several millimeters of lead to almost 1-eV temperatures ~1 eV 11 604 K! ~Ref. 128!. Pulsed electrical sources such as the PBFA-II facility in Sandia National Laboratories have produced 10 17 protons in 20-ns pulses, focused to 1-cm diameter, with subsequent heat430

ing of initially solid aluminum to 35 eV ~Ref. 129!. One facet of these sources is that because of the relatively long pulse durations, the materials undergo significant hydrodynamic expansion during the heating period. Lasergenerated proton beams with temporal durations of just a few picoseconds may provide a means of very rapid heating, on a timescale shorter than the hydrodynamic timescale. Experimental measurements of the conversion efficiency of laser energy into protons have ranged between 2 to 7% ~Refs. 1, 2, and 50!, dependent on the precise laser parameters, interaction conditions, and target material and thickness. The temporal duration of the proton pulse near the source has been shown to be of picosecond order both experimentally 105,122 and in extensive modeling.29 The proton energy spectrum can often be approximated as a single temperature exponential distribution with kT of a few mega-electron-volts. Most measurements have been made above a minimum threshold proton energy of 3 to 4 MeV; thus, the spectrum and conversion efficiency into protons at the lowest energies is rather uncertain. If one assumes that the spectrum continues in an exponential form, one can estimate the energy deposition of the beam in a secondary target. The first demonstration of laser-generated proton heating was obtained by Patel et al.106 In this experiment a 10-J pulse from the 100-fs JanUSP laser at Lawrence Livermore National Laboratory was focused onto an Al foil producing a 100- to 200-mJ proton beam. A second 10-mm-thick Al foil was placed in the path of the proton beam a distance of 250 mm from the first. A timeresolved single-wavelength measurement was made of the rear-surface emission in a narrow band at ;570 nm. If the target is heated volumetrically, one would expect to see a thermal emission trace with a sharp, almost instantaneous, rise followed by a gradual fall as the surface expands outward and cools. This pattern is indeed observed in the data, as shown in Fig. 17. The absolute measured intensity at this wavelength provided an estimate of the initial temperature of the rear surface, which was in this case 4 6 1 eV. A focused proton beam, produced from a spherically shaped target, was seen to heat a smaller region to a significantly higher temperature, ;23 eV. In these experiments the isochoric heating by the protons is volumetric but not uniform. The exponential form of the proton energy spectrum will result in a significant temperature gradient along the direction of the beam, with the highest energy deposition, and thus temperature, at the front surface of the heated foil. Further work will be required, optimizing the proton beam spectral characteristics, or the target design, to minimize these gradients in energy deposition. For some applications, such as proton fast ignition ~PFI!, this feature of proton heating may not be of primary concern; however, it is of some consequence for the achievement of the idealized single temperature and density states so desired for
FUSION SCIENCE AND TECHNOLOGY VOL. 49 APR. 2006

Borghesi et al.

FAST ION GENERATION BY LASER IRRADIATION OF SOLID TARGETS

Fig. 17. ~a! Experimental setup for flat and focusing target geometries. Each target consisted of a flat or hemispherical 10-mmthick Al target irradiated by the laser and a flat 10-mm-thick Al foil to be heated by the protons. ~b! Corresponding streak camera images showing space- and time-resolved thermal emission at 570 nm from the rear side of the proton-heated foil. The streak camera images an 800-mm spatial region with a 1-ns temporal window.

precise quantitative measurements of material properties in extreme, strongly coupled plasma regimes. IV.C. Proton Fast Ignition Since the discovery of intense, short, energetic bursts of ions with excellent beam quality generated by the short-pulse-lasersolid-matter interaction, the idea of using those beams for fast ignition was introduced.113,130,131 Protons have several advantages compared to other ion species and electrons. They can penetrate more deeply into a target to reach the high-density region, where the hot spot is to be formed, because of the quadratic dependence of the stopping power on the charge state. Also, as with all ions, they exhibit a characteristic maximum of the energy deposition at the end of their range ~in the Bragg peak!, which is desirable in order to heat a localized volume efficiently. As sketched in Fig. 18,
FUSION SCIENCE AND TECHNOLOGY VOL. 49 APR. 2006

the basic idea is to use multiple, short-pulse lasers irradiating a thin curved foil. The protons are accelerated from the rear surface of the foil and, because of the parabolic geometry, are focused into the compressed fuel. With respect to electrons, the higher mass and the neutralized space charge make them less likely to be subject to instabilities. As has been shown in Sec. III.C, laser-driven ion beams can have an excellent beam quality in terms of emittance ~permitting to focus them down into a small volume!, the pulse duration is short, and the particle numbers are high. According to current experimental and theoretical understanding, the prospects for PFI are quite promising, although a number of issues still need to be addressed quantitatively in the parameter range that is relevant to fast ignition. One of the requirements for PFI is the possibility of focusing the proton beams into a small volume. As
431

Borghesi et al.

FAST ION GENERATION BY LASER IRRADIATION OF SOLID TARGETS

Fig. 18. Schematic view of PFI in indirectly driven ICF ~not to scale!. The rear surface of the laser target is shaped to focus the proton beam into the spark volume.

discussed in Sec. IV.B, it has recently been demonstrated that proton beam focusing is indeed possible and spot sizes of ;50 mm have been achieved.106 This is still larger than required by PFI; however, in principle the ultralow transverse emittance reported for these proton beams should allow focusing down the beam to spot sizes orders of magnitude smaller than what has been achieved so far. In practice, there are several issues that need to be considered. The first issue is the degree of neutralization of the beam, which is composed of protons but also of comoving and neutralizing electrons. If the beam is not perfectly neutralized or if there are instabilities that develop in the flow and modulate the electron and ion density along the propagation axis, an emittance-limited focus may not be achieved because of Coulomb forces increasing near the focus. The second issue concerns the geometry of the focusing device and of the accelerating electron sheath. The aforementioned 50-mm spot size was achieved using protons accelerated from a hemispherical target. The concave rear surface of the hemisphere, where the proton acceleration takes place, induced a curvature of the sheath. However, this may have only partially compensated the natural Gaussian-like curvature of the sheath, which is responsible, for example, for the variation in divergence angles for different proton energies observed in the experiments. The final divergence of the accelerated beam results from the combined and competing effect of the target curvature and of the sheath curvature. This may explain why a relatively large focal spot was achieved compared to what could be in principle assumed from the determination of the beam emittance. However, one has to take into account the fact that larger irradiated areas on the target front surface as required for PFI ~see below! would flatten the electron distribution at the rear surface.
432

This would result not only in a single divergence angle for different energies but also in a much smaller initial divergence angle, for mostly all of the energies, which could be more easily compensated in order to reach the required focal spot diameters. The proton pulse length is of the right order of magnitude for PFI. The protons are not monochromatic but rather have an exponential energy distribution. This seemed to be a concern at the beginning for two reasons. First, because of the different energies, the dispersion of the proton pulse from the source to the target lengthens the pulse. Therefore, a maximum source-pellet distance of the order of a few millimeters is necessary for the energy deposition of the PFI beam to take place within the disassembly time for the hot spot region. On the other hand, this relatively short distance raises the concern if the thin metallic foil, which is to be the source of the protons, can be kept cold enough not to develop a density gradient at the rear surface, which would diminish the accelerating field. A second concern was related to the stopping power. Because of the difference in initial velocity, the energy deposition of protons with different kinetic energy is spread over a larger volume. Slower protons are stopped earlier and do not contribute to the creation of the ignitor spark. Recent numerical simulations have partly relieved those concerns. Simulations by Basko et al. ~presented at the Workshop on Fast Ignition of Fusion Targets, 2002, Tampa, Florida! have shown that the protective shield placed in front of the source can withstand the X-ray flux of the pellet compression and keep the rear surface of the source foil cold enough for efficient acceleration from the rear surface. In the simulation the drive pulse caused the pellet stagnation at 48 ns after the beginning of the irradiation. The thickness of the protective foil was found to be crucial not only with respect to thermal shielding of the proton source but also because of the closure of the acceleration gap. For a foil too thin, the displacement of the rear side ~the one facing the proton source! due to thermal expansion and ablative foil acceleration can become quite substantial. A thickness of a few tens of microns on the other hand provides thermal shielding as well as sufficient mechanical stability. A monochromatic proton beam is actually not the optimum to heat a hot spot in a fusion target. Numerical simulations have shown that one has to take into account the decrease of the stopping power of the nuclear fuel with increasing plasma temperature.113 Therefore, an exponential energy spectrum, like the one that is generated by target normal sheath acceleration, is the most favorable one. The protons with the highest energies arrive first and penetrate deeply into the fuel. By the time the proton number increases and the target temperature rises, the stopping power is reduced, thereby compensating for the lower initial energy of the incoming protons. As a result, the majority of the protons deposit their energy within the same volume.
FUSION SCIENCE AND TECHNOLOGY VOL. 49 APR. 2006

Borghesi et al.

FAST ION GENERATION BY LASER IRRADIATION OF SOLID TARGETS

In contrast to conventional fast ignition, the PFI can be easily implemented into indirectly driven scenarios, necessary for example in heavy ion ICF. For practical reasons, there is another advantage of PFI. Since focusing of the ion beam is provided by the geometry of the source foil, the pointing and focusing requirements for the high-energy short-pulse-laser beams are significantly relaxed. In fact, the laser beam foci can be as large as a few hundred microns, which as mentioned previously will flatten the shape of the electron sheath driving the protons and enhance their focusability. As a consequence, the final optics of the short-pulse lasers can be mounted relatively far from the reactor center, which may increase their lifetime. Existing short-pulse lasers have already demonstrated intensities that are sufficient for generating proton energy spectra required for PFI. Regardless of the nature of the ignitor beam, calculations show a minimum deposited energy requirement for fast ignition of the order of 10 kJ. Measured and published efficiencies for converting the primary laser energy into ion beams appear favorable for PFI. While experiments employing lowenergy short-pulse lasers have conversion efficiencies ,1%, the efficiency changes to a few percent for systems of a few joules of energy and increases up to .10% for systems having hundreds of joules at comparable focused intensities. Extrapolating the conversion efficiencies to multikilojoule laser systems conversion efficiencies of .10% can be expected, which would result in the need for a few hundred kilojoules of short-pulse-laser energy for PFI. The most detailed theoretical analysis of PFI requirements so far has been published by Temporal et al.113 for proton beams with an exponential energy spectrum. Following their assumptions a total proton energy of ;26 kJ at an effective temperature of 3 MeV is required. These values have been found to be consistent with a sourceto-target distance of 4 mm, which would allow for application in indirectly driven ICF. It is interesting to note that the protons, which effectively heat the hot spot, contain only 10 kJ of the total energy and range from 3 to 10 MeV. Techniques for shaping the energy spectrum of the laser-accelerated protons could lead to a reduction of the required laser beam energy. The simulations by Temporal et al. not only refer to an effective ion spectrum rather than to a monoenergetic ion beam but also take into account pulse lengthening due to velocity dispersion as well as range lengthening due to temperature effects. The total number of protons needed for ignition is close to 10 16, and a typical proton beam temperature of 3 MeV is commonly obtained in experiments at 5 10 19 W0cm 2. Assuming a pulse length of 4 ps ~which would increase the damage threshold of modern dielectric compressor gratings! and a conversion efficiency of 10%, a total laser energy of 260 kJ would be needed. To match the temperature conditions, the lasers would have to irradiate a proton source surface of 13 mm 2, correspondFUSION SCIENCE AND TECHNOLOGY VOL. 49 APR. 2006

ing to a diameter of ;4 mm. The question arises if there are enough protons available at the nonirradiated rear surface of the source. Early experiments have shown an increase in proton numbers of about an order of magnitude when irradiating bulk plastic 2 ; however, this was at the cost of beam quality due to electron beam breakup. Later experiments have demonstrated undisturbed electron transport ~and consequently the production of a highly laminar ion beam! through hydrogen-containing layers of up to 1000- thickness, coated at the rear of metallic targets. Assuming a CH layer at the rear surface with 400- thickness, an area of little less than 13 mm 2 would contain the 10 16 protons required for igniting the hot spot. A similar area would need to be irradiated at the front surface. Both areas would be tolerable in a geometry of 4-mm source to hot spot distance, as proposed in the simulation by Temporal et al., and still allow for efficient focusing. Currently, there are plans to build multikilojoule shortpulse laser systems in the United States, Japan, United Kingdom, and France. These facilities will allow experiments in a more relevant regime for fast ignition than the existing ones and provide the opportunity to better gauge the feasibility of PFI as an alternative way to achieve ignition. However, several issues need to be studied beforehand, even with the existing, smaller-scale facilities. There is a need to determine the exact degree of neutralization of the accelerated proton beam; the stopping power of protons in dense, hot plasmas ~this could be addressed using lasers coupled to ion accelerators!; techniques to achieve smaller proton focal spots; and ways of achieving energy selection within the proton beam. It will also be necessary to enhance simulation capabilities so that realistic models of proton acceleration and focusing can be used for predictive capability. Simulations of implosion and fuel assembly need also to progress so that more realistic geometry of compressed fuels can be used for calculations of proton stopping and beam requirements in order to refine the parameters required for PFI. IV.D. Nuclear Reactions Initiated by Laser-Driven Ions and Applications The interaction of laser-driven high-energy ions with secondary targets can initiate nuclear reactions of various types, which, as mentioned before, can been used as a tool to diagnose the beam properties. For example, 63 Cu~ p, n! 63 Zn reactions in copper stacks are used to quantify the proton numbers through measurement of b decay of 63 Zn nuclei using a NaI detector-based coincidence counting system.93,132 More recently, techniques employing a single Cu layer have been used in which a range of isotopes resulting from a protoninduced nuclear reaction is analyzed in order to reconstruct the proton spectrum.133 This also presents the opportunity to carry out nuclear physics experiments in laser laboratories rather than
433

Borghesi et al.

FAST ION GENERATION BY LASER IRRADIATION OF SOLID TARGETS

in accelerator facilities and to apply the products of the reaction processes in several areas. Reactions initiated by laser-accelerated high-Z ions have also been studied. McKenna et al.92,112,134 have shown that fusion reactions between fast heavy ions from a laser-produced plasma and stationary atoms in an adjacent activation sample create compound nuclei in excited states, which deexcite through the evaporation of protons, neutrons, and alpha particles. The emission of precise energy gamma rays from the residual nuclei in the activated samples was measured and used with calculated reaction cross sections to make quantitative measurements of heavy ion acceleration from a laser-produced plasma. Nuclear reactions of interest for spallation physics have also been investigated by employing the multi-mega-electron-volt proton beams.101 The broad energy distribution of the beams is in this case advantageous for the determination of residual nuclide generation arising from specific spallation processes such as evaporation. Multi-mega-electron-volt proton beams, generated from intense laser-plasma interactions, have also been used to induce nuclear reactions in low-Z materials such as 11 B and H2 18 O, in order to produce short-lived positronemitting isotopes of medical interest,90,135,136 i.e., for positron emission tomography ~PET!. In PET, the patient is injected with a pharmaceutical labeled with a short-lived positron-emitting isotope. This radiopharmaceutical is metabolized at specific sites in the body. Positrons annihilate with electrons to produce two counterpropagating gamma rays. After their detection, specific sites of high uptake of the pharmaceutical may be imaged. PET has proven to be extremely useful in medical imaging of blood flow and amino acid transport and in the detection of tumors. The radioisotopes used in PET are typically short-lived positron emitters such as 11 C, 13 N, 15 O, and 18 F. These reactions are carried out by using up to 20MeV protons or similar energy deuterons from cyclotrons with the concomitant problems of large size and cost and extensive radiation shielding. Production of shortlived isotopes via laser-driven proton beams may be economical in the near future with the possibility of employing moderate-energy, ultrashort, high-repetition tabletop lasers. Extrapolations based on present results point to the possibility of reaching the gigabecquerel activities required for PET therapy when laser systems capable of delivering 1-J, 30-fs pulses, focused at 10 20 W0cm 2, with kilohertz repetition will become available.90 Neutrons are an important product of the nuclear reactions produced above, with potential applications in cancer therapy ~boron neutron capture therapy!, neutron radiography, and transmutation of nuclear waste. The potential for laser-driven neutron sources is considerable and offers advantages over accelerator- and reactordriven sources in terms of cost, compactness, brightness, and short duration, particularly advantageous for fast neutron radiography 137 and studies of impulsive damage of matter.138 In view of these potential applications, several
434

experiments have studied the production of neutrons initiated by laser-driven proton beams on secondary targets. Experiments 137,139 carried out at the Vulcan laser facility have studied neutron generation via 11 B~ p, n! 11 C reactions and via the 7 Li~ p, n! 7 Be reaction revealing neutron yields up to 4 10 9 sr 10laser pulse ~result obtained for a laser intensity of 3 10 20 W0cm 2 ! ~Ref. 139!. Neutron production has also been observed during interactions with targets containing deuterium, either solid 140 145 or gaseous,9,146,147 directly irradiated by high-intensity laser pulses. In this case the neutrons are produced in the course of fusion reactions of the type D~d, n! 3 He involving laser-accelerated deuterium ions. In recent solid-target experiments, the neutron spectra produced through this process have been used as a diagnostic of the laser-driven deuterium ions inside the target 142144 in order to provide information on acceleration taking place at the front surface or in the bulk of the target. IV.E. Hadrontherapy Hadrontherapy 148152 is a form of radiotherapy that uses protons, neutrons, or carbon ions to irradiate cancer tumors. The use of protons and carbon ions in radiotherapy has several advantages to the more widely used X-ray radiotherapy. First, the proton beam scattering on the atomic electrons is weak, and thus, there is less irradiation of healthy tissues in the vicinity of the tumor. Second, the slowing-down length for a proton with given energy is fixed, which avoids undesirable irradiation of healthy tissues at the rear side of the tumor. Third, the well-localized maximum of the proton energy losses in matter ~the Bragg peak! leads to a substantial increase of the irradiation dose in the vicinity of the proton stopping point. After .40 yr of experimental investigations, several dedicated hadrontherapy centers are operating, and others are at present under construction ~see Ref. 153!. These centers have accelerators specialized for medical applications from which the proton beams are delivered into three to five procedure rooms equipped with treatment units. A necessary and costly treatment unit of the hospitalbased proton therapy centers is the gantry, which is a device that provides multidirectional irradiation of a lying patient. The proton energy window of therapeutical interest ranges between 60 and 250 MeV, depending on the location of the tumor. Proton beams with the required parameters are currently obtained using conventional charged-particle accelerators such as synchrotrons, cyclotrons, and linacs.154 The use of laser-based accelerators has been proposed as an alternative,155159 which could lead to advantages in terms of device compactness and costs. A laser accelerator could be used simply as a highefficiency ion injector for the proton accelerator or could replace altogether a conventional proton accelerator.
FUSION SCIENCE AND TECHNOLOGY VOL. 49 APR. 2006

Borghesi et al.

FAST ION GENERATION BY LASER IRRADIATION OF SOLID TARGETS

Because of the broad energy and angular spectra of the protons, an energy selection and beam collimation system will be needed.160,161 Typically, DE0E ; 10 2 is required for optimal dose delivery over the tumor region. All-optical systems have also been proposed in which the ion beam acceleration takes place in the treatment room itself and ion beam transport and delivery issues are so minimized.155 In this case the beam energy spread and divergence would have to be minimized by controlling the beam and target parameters.58,64 The required energies for deep-seated tumors ~.200 MeV! are still in the future but appear within reach considering the ongoing developments in the field. A demanding requirement to be satisfied is also the system duty factor, i.e., the fraction of time during which the proton beam is available for use that must not be ,0.3. IV.F. Technological Applications There is certainly scope for application of laserdriven ion sources in areas of industrial0manufacturing interest, particularly if high-repetition tabletop, moderatecost laser systems with the required parameters for efficient ion generation will be developed. Areas of obvious interest are ion implantation in substrates and manufacturing techniques such as deep proton projection lithography and micromachining.162164 While surface implantation of moderate-energy ions accelerated by subnanosecond pulses has been already studied,165,166 the higher ion energies attainable with high-intensity pulses could allow deep implantation via ion projection approaches,167,168 which is of interest for semiconductor device manufacturing. The beams exceptional quality in terms of emittance could play a crucial role in this type of application. In addition, the short duration of laser-driven ion bursts opens up to investigation a whole new regime of physical conditions relating to the interaction of ionizing radiation with matter and the induced material modifications. This is a highly topical area for many technological applications, including ion implantation or control of material damage in fusion reactors. Since the response of radiation-damaged materials takes place typically over timescales in the picosecond range, impulsive ~picosecond! particle irradiation of matter provides now favorable conditions to carry out pump-probe experiments with unprecedented temporal resolution.

tional spatial and temporal qualities of the beams, the high-energy and high particle numbers achievable, and the application of these properties to a number of applications, both in plasma physics and in interdisciplinary areas. Work is undergoing worldwide aimed at the optimization and control of the acceleration mechanism and of the beam properties. Results in this area, coupled with the continuous development in laser technology, are bound to further expand the field of applicability of these sources to include medical or technological applications.

ACKNOWLEDGMENTS
The authors acknowledge discussions with F. Pegoraro and M. Lontano and support from the QUB-IRCEP scheme.

REFERENCES
1. E. L. CLARK et al., Measurements of Energetic Proton Transport Through Magnetized Plasma from Intense Laser Interactions with Solids, Phys. Rev. Lett., 84, 670 ~2000!. 2. R. A. SNAVELY et al., Intense High-Energy Proton Beams from Petawatt-Laser Irradiation of Solids, Phys. Rev. Lett., 85, 2945 ~2000!. 3. A. MAKSIMCHUK, S. GU, K. FLIPPO, D. UMSTADTER, and V. Y. BYCHENKOV, Forward Ion Acceleration in Thin Films Driven by a High-Intensity Laser, Phys. Rev. Lett., 84, 4108 ~2000!. 4. S. J. GITOMER, R. D. JONES, F. BEGAY, A. W. EHLER, J. F. KEPHART, and R. KRISTAL, Fast Ions and Hot-Electrons in the Laser-Plasma Interaction, Phys. Fluids, 29, 2679 ~1986!. 5. A. P. FEWS et al., Plasma Ion Emission from High-Intensity Picosecond Laser-Pulse Interactions with Solid Targets, Phys. Rev. Lett., 73, 1801 ~1994!. 6. D. GIULIETTI et al., Production of Ultracollimated Bunches of Multi-MeV Electrons by 35 fs Laser Pulses Propagating in ExplodingFoil Plasmas, Phys. Plasmas, 9, 3655 ~2002!. 7. K. MATSUKADO et al., Energetic Protons from a Few-Micron Metallic Foil Evaporated by an Intense Laser Pulse, Phys. Rev. Lett., 91, 215001 ~2003!. 8. K. NEMOTO, A. MAKSIMCHUK, S. BANERJEE, K. FLIPPO, G. MOUROU, D. UMSTADTER, and V. Y. BYCHENKOV, LaserTriggered Ion Acceleration and Table Top Isotope Production, Appl. Phys. Lett., 78, 595 ~2001!. 9. S. FRITZLER et al., Ion Heating and Thermonuclear Neutron Production from High-Intensity Subpicosecond Laser Pulses Interacting with Underdense Plasmas, Phys. Rev. Lett., 89, 165004 ~2002!. 10. K. KRUSHELNICK et al., Multi-MeV Ion Production from HighIntensity Laser Interactions with Underdense Plasmas, Phys. Rev. Lett., 83, 737 ~1999!. 11. M. S. WEI et al., Ion Acceleration by Collisionless Shocks in High-Intensity-Laser-Underdense-Plasma Interaction, Phys. Rev. Lett., 93, 155003 ~2004!. 12. F. AMIRANOFF, Fast Electron Production in Ultra-Short HighIntensity Laser-Plasma Interaction and Its Consequences, Meas. Sci. Technol., 12, 1795 ~2001!. 13. P. GIBBON and A. R. BELL, Collisionless Absorption in SharpEdged Plasmas, Phys. Rev. Lett., 68, 1535 ~1992!.

V. CONCLUSIONS Research in the area of laser-driven ion sources has grown at a phenomenal pace since the first experiments reporting multi-mega-electron-volt proton acceleration from laser-irradiated foils, as testified by the high number of publications appearing in the scientific literature. Most of the present research is motivated by the excepFUSION SCIENCE AND TECHNOLOGY VOL. 49 APR. 2006

435

Borghesi et al.

FAST ION GENERATION BY LASER IRRADIATION OF SOLID TARGETS


35. A. V. GUREVICH, L. V. PARIISKAYA, and L. P. PITAEVSKII, Self-Similar Motion of a Rarefied Plasma, Sov. Phys. JETP, 22, 449 ~1966!. 36. V. F. KOVALEV and V. Y. BYCHENKOV, Analytic Solutions to the Vlasov Equations for Expanding Plasmas, Phys. Rev. Lett., 90, 185004 ~2003!. 37. D. S. DOROZHKINA and V. E. SEMENOV, Exact Solution of Vlasov Equations for Quasineutral Expansion of Plasma Bunch into Vacuum, Phys. Rev. Lett., 81, 2691 ~1998!. 38. M. PASSONI and M. LONTANO, One-Dimensional Model of the Electrostatic Ion Acceleration in the Ultraintense Laser-Solid Interaction, Laser Part. Beams, 22, 163 ~2004!. 39. S. V. BULANOV, T. Z. ESIRKEPOV, J. KOGA, T. TAJIMA, and D. FARINA, Concerning the Maximum Energy of Ions Accelerated at the Front of a Relativistic Electron Cloud Expanding into Vacuum, Plasma Phys. Rep., 30, 18 ~2004!. 40. S. BETTI, F. CECCHERINI, F. CORNOLTI, and F. PEGORARO, Expansion of a Finite Size Plasma in Vacuum, Plasma Phys. Control. Fusion, 47, 521 ~2005!. 41. M. PASSONI, V. T. TIKHONCHUK, M. LONTANO, and V. Y. BYCHENKOV, Charge Separation Effects in Solid Targets and Ion Acceleration with a Two-Temperature Electron Distribution, Phys. Rev. E, 69, 026411 ~2004!. 42. V. F. KOVALEV, V. Y. BYCHENKOV, and V. T. TIKHONCHUK, Particle Dynamics During Adiabatic Expansion of a Plasma Bunch, J. Exp. Theor. Phys., 95, 226 ~2002!. 43. L. GREMILLET, G. BONNAUD, and F. AMIRANOFF, Filamented Transport of Laser-Generated Relativistic Electrons Penetrating a Solid Target, Phys. Plasmas, 9, 941 ~2002!. 44. J. J. SANTOS et al., Fast Electron Transport in Ultraintense Laser Pulse Interaction with Solid Targets by Rear-Side Self-Radiation Diagnostics, Phys. Rev. Lett., 89, 025001 ~2002!. 45. J. R. DAVIES, Electric and Magnetic Field Generation and Target Heating by Laser-Generated Fast Electrons, Phys. Rev. E, 68, 056404 ~2003!. 46. F. A. BIBI, J. P. MATTE, and J. C. KIEFFER, Fokker-Planck Simulations of Hot Electron Transport in Solid Density Plasma, Laser Part. Beams, 22, 97 ~2004!. 47. J. J. HONRUBIA, A. ANTONICCI, and D. MORENO, Hybrid Simulations of Fast Electron Transport in Conducting Media, Laser Part. Beams, 22, 129 ~2004!. 48. R. J. KINGHAM and A. R. BELL, An Implicit Vlasov-FokkerPlanck Code to Model Non-Local Electron Transport in 2-D with Magnetic Fields, J. Comput. Phys., 194, 1 ~2004!. 49. J. FUCHS et al., Spatial Uniformity of Laser-Accelerated Ultrahigh-Current MeV Electron Propagation in Metals and Insulators, Phys. Rev. Lett., 91, 255002 ~2003!. 50. A. J. MACKINNON et al., Enhancement of Proton Acceleration by Hot-Electron Recirculation in Thin Foils Irradiated by Ultraintense Laser Pulses, Phys. Rev. Lett., 88, 215006 ~2002!. 51. H. RUHL, T. COWAN, and J. FUCHS, The Generation of MicroFiducials in Laser-Accelerated Proton Flows, Their Imaging Property of Surface Structures and Application for the Characterization of the Flow, Phys. Plasmas, 11, L17 ~2004!. 52. P. GIBBON, F. N. BEG, E. L. CLARK, R. G. EVANS, and M. ZEPF, Tree-Code Simulations of Proton Acceleration from LaserIrradiated Wire Targets, Phys. Plasmas, 11, 4032 ~2004!. 53. J. DENAVIT, Absorption of High-Intensity Subpicosecond Lasers on Solid Density Targets, Phys. Rev. Lett., 69, 3052 ~1992!.
FUSION SCIENCE AND TECHNOLOGY VOL. 49 APR. 2006

14. F. BRUNEL, Not-So-Resonant, Resonant Absorption, Phys. Rev. Lett., 59, 52 ~1987!. 15. V. L. GINZBURG, The Propagation of Electromagnetic Waves in Plasmas, Pergamon Press, New York ~1964!. 16. W. L. KRUER, The Physics of Laser Plasma Interaction, AddisonWesley, New York ~1988!. 17. M. I. K. SANTALA et al., Effect of the Plasma Density Scale Length on the Direction of Fast Electrons in Relativistic Laser-Solid Interactions, Phys. Rev. Lett., 84, 1459 ~2000!. 18. A. G. LITVAK, Finite-Amplitude Wave Beams in a MagnetoActive Plasma, Sov. Phys. JETP, 30, 344 ~1969!. 19. P. SPRANGLE, C. M. TANG, and E. ESAREY, Relativistic SelfFocusing of Short-Pulse Radiation Beams in Plasmas, IEEE Trans. Plasma Sci., 15, 145 ~1987!. 20. A. B. BORISOV, A. V. BOROVSKIY, V. V. KOROBKIN, A. M. PROKHOROV, C. K. RHODES, and O. B. SHIRYAEV, Stabilization of Relativistic Self-Focusing of Intense Subpicosecond Ultraviolet Pulses in Plasmas, Phys. Rev. Lett., 65, 1753 ~1990!. 21. P. MONOT, T. AUGUSTE, P. GIBBON, F. JAKOBER, and G. MAINFRAY, Experimental Demonstration of Relativistic SelfChanneling of a Multiterawatt Laser-Pulse in an Underdense Plasma, Phys. Rev. Lett., 74, 2953 ~1995!. 22. M. BORGHESI et al., Relativistic Channeling of a Picosecond Laser Pulse in a Near-Critical Preformed Plasma, Phys. Rev. Lett., 78, 879 ~1997!. 23. A. PUKHOV and J. MEYER-TER-VEHN, Relativistic Magnetic Self-Channeling of Light in Near-Critical Plasma: Three-Dimensional Particle-in-Cell Simulation, Phys. Rev. Lett., 76, 3975 ~1996!. 24. P. E. YOUNG and P. R. BOLTON, Propagation of Subpicosecond Laser Pulses Through a Fully Ionized Plasma, Phys. Rev. Lett., 77, 4556 ~1996!. 25. Z. M. SHENG, K. NISHIHARA, T. HONDA, Y. SENTOKU, K. MIMA, and S. V. BULANOV, Anisotropic Filamentation Instability of Intense Laser Beams in Plasmas Near the Critical Density, Phys. Rev. E, 6406, 066409 ~2001!. 26. W. L. KRUER and K. ESTABROOK, JXB Heating by Very Intense Laser-Light, Phys. Fluids, 28, 430 ~1985!. 27. S. C. WILKS, W. L. KRUER, M. TABAK, and A. B. LANGDON, Absorption of Ultra-Intense Laser-Pulses, Phys. Rev. Lett., 69, 1383 ~1992!. 28. A. PUKHOV, Z. M. SHENG, and J. MEYER-TER-VEHN, Particle Acceleration in Relativistic Laser Channels, Phys. Plasmas, 6, 2847 ~1999!. 29. Y. SENTOKU, T. E. COWAN, A. KEMP, and H. RUHL, High Energy Proton Acceleration in Interaction of Short Laser Pulse with Dense Plasma Targets, Phys. Plasmas, 10, 2009 ~2003!. 30. S. P. HATCHETT et al., Electron, Photon, and Ion Beams from the Relativistic Interaction of Petawatt Laser Pulses with Solid Targets, Phys. Plasmas, 7, 2076 ~2000!. 31. S. C. WILKS et al., Energetic Proton Generation in Ultra-Intense Laser-Solid Interactions, Phys. Plasmas, 8, 542 ~2001!. 32. P. MORA, Plasma Expansion into a Vacuum, Phys. Rev. Lett., 90, 185002 ~2003!. 33. A. V. GUREVICH and L. P. PITAEVSKII, Sov. Phys. JETP, 36, 274 ~1972!. 34. J. E. CROW, P. L. AUER, and J. E. ALLEN, Expansion of a Plasma into a Vacuum, J. Plasma Phys., 14, 65 ~1975!.

436

Borghesi et al.

FAST ION GENERATION BY LASER IRRADIATION OF SOLID TARGETS


77. M. ZEPF et al., Proton Acceleration from High-Intensity Laser Interactions with Thin Foil Targets, Phys. Rev. Lett., 90, 064801 ~2003!. 78. M. HEGELICH et al., MeV Ion Jets from Short-Pulse-Laser Interaction with Thin Foils, Phys. Rev. Lett., 89, 085002 ~2002!. 79. J. BADZIAK et al., Fast Proton Generation from Ultrashort Laser Pulse Interaction with Double-Layer Foil Targets, Phys. Rev. Lett., 87, 215001 ~2001!. 80. A. J. MACKINNON et al., Effect of Plasma Scale Length on Multi-MeV Proton Production by Intense Laser Pulses, Phys. Rev. Lett., 86, 1769 ~2001!. 81. F. N. BEG et al., Target Charging Effects on Proton Acceleration During High-Intensity Short-Pulse Laser-Solid Interactions, Appl. Phys. Lett., 84, 2766 ~2004!. 82. M. ROTH et al., Energetic Ions Generated by Laser Pulses: A Detailed Study on Target Properties, Phys. Rev. Spec. Top. Accel. Beams, 5, 061301 ~2002!. 83. S. KARSCH et al., High-Intensity Laser Induced Ion Acceleration from Heavy-Water Droplets, Phys. Rev. Lett., 91, 015001 ~2003!. 84. M. SCHNURER, S. TER-AVETISYAN, S. BUSCH, M. P. KALACHNIKOV, E. RISSE, W. SANDNER, and P. V. NICKLES, MeV-Proton Emission from Ultrafast Laser-Driven Microparticles, Appl. Phys. B Lasers Opt., 78, 895 ~2004!. 85. N. V. KLASSEN, L. VAN DER ZWAN, and J. CYGLER, GafChromic MD-55: Investigated as a Precision Dosimeter, Med. Phys., 24, 1924 ~1997!. 86. W. L. McLAUGHLIN et al., Novel Radiochromic Films for Clinical Dosimetry, Radiat. Prot. Dosimet., 66, 263 ~1996!. 87. M. A. STOYER et al., Nuclear Diagnostics for Petawatt Experiments, Rev. Sci. Instrum., 72, 767 ~2001!. 88. F. H. SEGUIN et al., Spectrometry of Charged Particles from Inertial-Confinement-Fusion Plasmas, Rev. Sci. Instrum., 74, 975 ~2003!. 89. M. ALLEN, P. K. PATEL, A. MACKINNON, D. PRICE, S. WILKS, and E. MORSE, Direct Experimental Evidence of Back-Surface Ion Acceleration from Laser-Irradiated Gold Foils, Phys. Rev. Lett., 93, 265004 ~2004!. 90. S. FRITZLER et al., Proton Beams Generated with HighIntensity Lasers: Applications to Medical Isotope Production, Appl. Phys. Lett., 83, 3039 ~2003!. 91. J. FUCHS et al., Comparison of Laser Ion Acceleration from the Front and Rear Surfaces of Thin Foils, Phys. Rev. Lett., 94, 045004 ~2005!. 92. P. McKENNA et al., Characterization of Proton and Heavier Ion Acceleration in Ultrahigh-Intensity Laser Interactions with Heated Target Foils, Phys. Rev. E, 70, 036405 ~2004!. 93. I. SPENCER et al., Laser Generation of Proton Beams for the Production of Short-Lived Positron Emitting Radioisotopes, Nucl. Instrum. Methods Phys. Res. B Beam Interact. Mater. Atoms, 183, 449 ~2001!. 94. A. MAKSIMCHUK et al., High-Energy Ion Generation by Short Laser Pulses, Plasma Phys. Rep., 30, 473 ~2004!. 95. M. KALUZA, J. SCHREIBER, M. I. K. SANTALA, G. D. TSAKIRIS, K. EIDMANN, J. MEYER-TER-VEHN, and K. J. WITTE, Influence of the Laser Prepulse on Proton Acceleration in Thin-Foil Experiments, Phys. Rev. Lett., 93, 045003 ~2004!. 96. T. FUJII et al., MeV-Order Proton and Carbon Ion Acceleration by Irradiation of 60 fs TW Laser Pulses on Thin Copper Tape, Appl. Phys. Lett., 83, 1524 ~2003!.

54. A. ZHIDKOV, M. UESAKA, A. SASAKI, and H. DAIDO, Ion Acceleration in a Solitary Wave by an Intense Picosecond Laser Pulse, Phys. Rev. Lett., 89, 215002 ~2002!. 55. L. O. SILVA, M. MARTI, J. R. DAVIES, R. A. FONSECA, C. REN, F. S. TSUNG, and W. B. MORI, Proton Shock Acceleration in Laser-Plasma Interactions, Phys. Rev. Lett., 92, 015002 ~2004!. 56. E. DHUMIERES, E. LEFEBYRE, L. GREMILLET, and V. MALKA, Proton Acceleration Mechanisms in High-Intensity Laser Interaction with Thin Foils, Phys. Plasmas, 12, 062704 ~2005!. 57. A. PUKHOV, Three-Dimensional Simulations of Ion Acceleration from a Foil Irradiated by a Short-Pulse Laser, Phys. Rev. Lett., 86, 3562 ~2001!. 58. H. RUHL et al., Computer Simulation of the Three-Dimensional Regime of Proton Acceleration in the Interaction of Laser Radiation with a Thin Spherical Target, Plasma Phys. Rep., 27, 363 ~2001!. 59. F. PEGORARO et al., Ion Acceleration Regimes in Underdense Plasmas, IEEE Trans. Plasma Sci., 28, 1177 ~2000!. 60. Y. SENTOKU et al., High-Energy Ion Generation in Interaction of Short Laser Pulse with High-Density Plasma, Appl. Phys. B. Lasers Opt., 74, 207 ~2002!. 61. Y. SENTOKU et al., High Density Collimated Beams of Relativistic Ions Produced by Petawatt Laser Pulses in Plasmas, Phys. Rev. E, 62, 7271 ~2000!. 62. T. Z. ESIRKEPOV et al., Ion Acceleration by Superintense Laser Pulses in Plasmas, JETP Lett., 70, 82 ~1999!. 63. T. ESIRKEPOV, M. BORGHESI, S. V. BULANOV, G. MOUROU, and T. TAJIMA, Highly Efficient Relativistic-Ion Generation in the Laser-Piston Regime, Phys. Rev. Lett., 92, 175003 ~2004!. 64. T. Z. ESIRKEPOV et al., Proposed Double-Layer Target for the Generation of High-Quality Laser-Accelerated Ion Beams, Phys. Rev. Lett., 89, 175003 ~2002!. 65. Y. MURAKAMI et al., Observation of Proton Rear Emission and Possible Gigagauss Scale Magnetic Fields from Ultra-Intense Laser Illuminated Plastic Target, Phys. Plasmas, 8, 4138 ~2001!. 66. V. I. VEKSLER, in Proc. CERN Symp. High Energy Accelerators and Pion Physics, Geneva, Switzerland, 1956, Vol. 1, p. 80. 67. V. I. VEKSLER, At. Energy, 2, 427 ~1957!. 68. M. ASHOUR-ABDALLA, J. N. LEBOEUF, T. TAJIMA, J. M. DAWSON, and C. F. KENNEL, Ultrarelativistic Electromagnetic Pulses in Plasmas, Phys. Rev. A, 23, 1906 ~1981!. 69. G. A. ASKARYAN, At. Energy, 4, 71 ~1958!. 70. S. WEINBERG, Gravitation and Cosmology, John Wiley & Sons, New York ~1972!. 71. E. M. LIFSHITZ, J. Phys. USSR, 10, 116 ~1946!. 72. J. D. ZUEGEL et al., Laser Challenges for Fast Ignition, Fusion Sci. Technol., 49, 453 ~2006!. 73. J. BADZIAK et al., Generation of Energetic Protons from Thin Foil Targets Irradiated by a High-Intensity Ultrashort Laser Pulse, Nucl. Instrum. Methods Phys. Res. A Accel. Spectrom. Dect. Assoc. Equip., 498, 503 ~2003!. 74. F. N. BEG et al., A Study of Picosecond Laser-Solid Interactions up to 10~19! W cm~-2!, Phys. Plasmas, 4, 447 ~1997!. 75. K. KRUSHELNICK et al., Energetic Proton Production from Relativistic Laser Interaction with High Density Plasmas, Phys. Plasmas, 7, 2055 ~2000!. 76. M. ZEPF et al., Fast Particle Generation and Energy Transport in Laser-Solid Interactions, Phys. Plasmas, 8, 2323 ~2001!.
FUSION SCIENCE AND TECHNOLOGY VOL. 49 APR. 2006

437

Borghesi et al.

FAST ION GENERATION BY LASER IRRADIATION OF SOLID TARGETS


117. A. J. MACKINNON et al., Proton Radiography as an Electromagnetic Field and Density Perturbation Diagnostic, Rev. Sci. Instrum., 75, 3531 ~2004!. 118. M. KOENIG et al., High Pressures Generated by Laser Driven Shocks: Applications to Planetary Physics, Nucl. Fusion, 44, S208 ~2004!. 119. S. LE PAPE et al., Novel Diagnostic of Shock Fronts in Low-Z Dense Plasmas, Astrophys. Space Sci., 298, 313 ~2005!. 120. M. BORGHESI et al., Electric Field Detection in Laser-Plasma Interaction Experiments via the Proton Imaging Technique, Phys. Plasmas, 9, 2214 ~2002!. 121. M. BORGHESI et al., Proton Imaging Detection of Transient Electromagnetic Fields in Laser-Plasma Interactions, Rev. Sci. Instrum., 74, 1688 ~2003!. 122. M. BORGHESI et al., Measurement of Highly Transient Electrical Charging Following High-Intensity Laser-Solid Interaction, Appl. Phys. Lett., 82, 1529 ~2003!. 123. M. BORGHESI et al., Macroscopic Evidence of Soliton Formation in Multiterawatt Laser-Plasma Interaction, Phys. Rev. Lett., 88, 135002 ~2002!. 124. A. J. MACKINNON, P. K. PATEL, D. W. PRICE, D. HICKS, M. BORGHESI, and L. ROMAGNANI, First Observation of Moire Fringes in a Proton Beam Generated by a 100 fs Laser Pulse, Rev. Sci. Instrum., 74, 1917 ~2003!. 125. A. J. MACKINNON, P. K. PATEL, D. W. PRICE, D. HICKS, L. ROMAGNANI, and M. BORGHESI, Proton Moire Fringes for Diagnosing Electromagnetic Fields in Opaque Materials and Plasmas, Appl. Phys. Lett., 82, 3188 ~2003!. 126. F. N. BEG et al., High-Intensity-Laser-Driven Z Pinches, Phys. Rev. Lett., 92, 095001 ~2004!. 127. M. BORGHESI et al., Plasma Ion Evolution in the Wake of a High-Intensity Ultrashort Laser Pulse, Phys. Rev. Lett., 94, 145003 ~2005!. 128. D. H. H. HOFFMANN et al., Plasma Physics with Intense Laser and Ion Beams, Nucl. Instrum. Methods Phys. Res. B Beam Interact. Mater. Atoms, 161, 9 ~2000!. 129. J. BAILEY et al., Observation of K-Alpha X-Ray-Satellites from a Target Heated by an Intense Ion-Beam, Laser Part. Beams, 8, 555 ~1990!. 130. M. ROTH et al., Fast Ignition by Intense Laser-Accelerated Proton Beams, Phys. Rev. Lett., 86, 436 ~2001!. 131. S. ATZENI, M. TEMPORAL, and J. J. HONRUBIA, A First Analysis of Fast Ignition of Precompressed ICF Fuel by LaserAccelerated Protons, Nucl. Fusion, 42, L1 ~2002!. 132. M. I. K. SANTALA et al., Production of Radioactive Nuclides by Energetic Protons Generated from Intense Laser-Plasma Interactions, Appl. Phys. Lett., 78, 19 ~2001!. 133. J. M. YANG et al., Nuclear Reactions in Copper Induced by Protons from a Petawatt Laser-Foil Interaction, Appl. Phys. Lett., 84, 675 ~2004!. 134. P. McKENNA et al., Effect of Target Heating on Ion-Induced Reactions in High-Intensity Laser-Plasma Interactions, Appl. Phys. Lett., 83, 2763 ~2003!. 135. K. W. D. LEDINGHAM, P. McKENNA, and R. P. SINGHAL, Applications for Nuclear Phenomena Generated by Ultra-Intense Lasers, Science, 300, 1107 ~2003!. 136. K. W. D. LEDINGHAM et al., High Power Laser Production of Short-Lived Isotopes for Positron Emission Tomography, J. Phys. D Appl. Phys., 37, 2341 ~2004!.
FUSION SCIENCE AND TECHNOLOGY VOL. 49 APR. 2006

97. I. SPENCER et al., Experimental Study of Proton Emission from 60-fs, 200-mJ High-Repetition-Rate Tabletop-Laser Pulses Interacting with Solid Targets, Phys. Rev. E, 67, 046402 ~2003!. 98. H. KISHIMURA et al., Enhanced Generation of Fast Protons from a Polymer-Coated Metal Foil by a Femtosecond Intense Laser Field, Appl. Phys. Lett., 85, 2736 ~2004!. 99. Y. OISHI et al., Dependence on Laser Intensity and Pulse Duration in Proton Acceleration by Irradiation of Ultrashort Laser Pulses on a Cu Foil Target, Phys. Plasmas, 12, 073102 ~2005!. 100. M. ALLEN et al., Proton Spectra from Ultraintense LaserPlasma Interaction with Thin Foils: Experiments, Theory, and Simulations, Phys. Plasmas, 10, 3283 ~2003!. 101. P. McKENNA et al., Broad Energy Spectrum of LaserAccelerated Protons for Spallation-Related Physics, Phys. Rev. Lett., 94, 084801 ~2005!. 102. H. RUHL and P. MULSER, Relativistic Vlasov Simulation of Intense Fs Laser Pulse-Matter Interaction, Phys. Lett. A, 205, 388 ~1995!. 103. M. BORGHESI et al., Multi-MeV Proton Source Investigations in Ultraintense Laser-Foil Interactions, Phys. Rev. Lett., 92, 055003 ~2004!. 104. J. SCHREIBER et al., Source-Size Measurements and Charge Distributions of Ions Accelerated from Thin Foils Irradiated by HighIntensity Laser Pulses, Appl. Phys. B Lasers Opt., 79, 1041 ~2004!. 105. L. ROMAGNANI et al., Dynamics of Electric Fields Driving Laser-Acceleration of Multi-MeV Protons, Phys. Rev. Lett., 95, 195001 ~2005!. 106. P. K. PATEL et al., Isochoric Heating of Solid-Density Matter with an Ultrafast Proton Beam, Phys. Rev. Lett., 91, 125004 ~2003!. 107. T. E. COWAN et al., Ultralow Emittance, Multi-MeV Proton Beams from a Laser Virtual-Cathode Plasma Accelerator, Phys. Rev. Lett., 92, 204801 ~2004!. 108. E. L. CLARK et al., Energetic Heavy-Ion and Proton Generation from Ultraintense Laser-Plasma Interactions with Solids, Phys. Rev. Lett., 85, 1654 ~2000!. 109. E. BRESCHI, M. BORGHESI, M. GALIMBERTI, D. GLULLETTI, L. A. GIZZI, and L. ROMAGNANI, A New Algorithm for Spectral and Spatial Reconstruction of Proton Beams from Dosimetric Measurements, Nucl. Instrum. Methods Phys. Res. A Accel. Spectrom. Dect. Assoc. Equip., 522, 190 ~2004!. 110. E. BRESCHI et al., Spectral and Angular Characterization of Laser-Produced Proton Beams from Dosimetric Measurements, Laser Part. Beams, 22, 393 ~2004!. 111. T. E. COWAN, Laser Generation of Accelerator Quality Proton Beams, High Brightness Beams, World Scientific, Singapore ~2000!. 112. P. McKENNA et al., Demonstration of Fusion-Evaporation and Direct-Interaction Nuclear Reactions Using High-Intensity LaserPlasma-Accelerated Ion Beams, Phys. Rev. Lett., 91, 075006 ~2003!. 113. M. TEMPORAL, J. J. HONRUBIA, and S. ATZENI, Numerical Study of Fast Ignition of Ablatively Imploded Deuterium-Tritium Fusion Capsules by Ultra-Intense Proton Beams, Phys. Plasmas, 9, 3098 ~2002!. 114. N. S. P. KING et al., An 800-MeV Proton Radiography Facility for Dynamic Experiments, Nucl. Instrum. Methods Phys. Res. A Accel. Spectrom. Dect. Assoc. Equip., 424, 84 ~1999!. 115. C. W. MENDEL, Jr., and J. N. OLSEN, Charge-Separation Electric Fields in Laser Plasmas, Phys. Rev. Lett., 34, 859 ~1975!. 116. J. A. COBBLE, R. P. JOHNSON, T. E. COWAN, N. RENARD-LE GALLOUDEC, and M. ALLEN, High Resolution Laser-Driven Proton Radiography, J. Appl. Phys., 92, 1775 ~2002!.

438

Borghesi et al.

FAST ION GENERATION BY LASER IRRADIATION OF SOLID TARGETS


154. W. WIESZCZYCKA and W. H. SCHARF, Proton Radiotherapy Accelerators, World Scientific, River Edge, New Jersey ~2001!. 155. S. V. BULANOV, T. Z. ESIRKEPOV, V. S. KHOROSHKOV, A. V. KUNETSOV, and F. PEGORARO, Oncological Hadrontherapy with Laser Ion Accelerators, Phys. Lett. A, 299, 240 ~2002!. 156. S. V. BULANOV and V. S. KHOROSHKOV, Feasibility of Using Laser Ion Accelerators in Proton Therapy, Plasma Phys. Rep., 28, 453 ~2002!. 157. E. FOURKAL and C. MA, Laser-Accelerated Carbon Ion Beams for Radiation Therapy, Med. Phys., 30, 1448 ~2003!. 158. E. FOURKAL, R. PRICE, C. MA, and A. POLLACK, Energy and Intensity Modulated Radiation Therapy Using Laser Accelerated Proton Beams, Med. Phys., 31, 1884 ~2004!. 159. V. MALKA et al., Practicability of Protontherapy Using Compact Laser Systems, Med. Phys., 31, 1587 ~2004!. 160. E. FOURKAL, J. LI, M. DING, T. TAJIMA, and C. MA, Particle Selection System for Laser-Accelerated Proton Therapy, Med. Phys., 29, 1326 ~2002!. 161. E. FOURKAL, J. S. LI, M. DING, T. TAJIMA, and C. M. MA, Particle Selection for Laser-Accelerated Proton Therapy Feasibility Study, Med. Phys., 30, 1660 ~2003!. 162. H. OTTEVAERE, B. VOLCKAERTS, J. LAMPRECHT, J. SCHWIDER, A. HERMANNE, I. VERETENNICOFF, and H. THIENPONT, Two-Dimensional Plastic Microlens Arrays by Deep Lithography with Protons: Fabrication and Characterization, J. Opt. A Pure Appl. Opt., 4, S22 ~2002!. 163. J. MELNGAILIS, A. A. MONDELLI, I. L. BERRY, and R. MOHONDRO, A Review of Ion Projection Lithography, J. Vac. Sci. Technol. B, 16, 927 ~1998!. 164. I. RAJTA, I. GOMEZ-MORILLA, M. H. ABRAHAM, and A. Z. KISS, Proton Beam Micromachining on PMMA, Foturan and CR-39 Materials, Nucl. Instrum. Methods Phys. Res. B Beam Interact. Mater. Atoms, 210, 260 ~2003!. 165. L. TORRISI et al., Implantation of Ions Produced by the Use of High Power Iodine Laser, Appl. Surf. Sci., 217, 319 ~2003!. 166. J. KRASA et al., Highly Charged Ions Generated with Intense Laser Beams, Nucl. Instrum. Methods Phys. Res. B Beam Interact. Mater. Atoms, 205, 355 ~2003!. 167. J. MEIJER, B. BURCHARD, K. IVANOVA, B. E. VOLLAND, I. W. RANGELOW, M. RUB, and G. DEBOY, High-Energy Ion Projection for Deep Ion Implantation as a Low Cost High Throughput Alternative for Subsequent Epitaxy Processes, J. Vac. Sci. Technol. B, 22, 152 ~2004!. 168. J. M. MIAO and H. L. HARTNAGEL, High-Energy Ion Implantation: An Alternative Technology for Micromachining ThreeDimensional GaAs Structures, Sens. Actuator A Phys., 114, 505 ~2004!.

137. K. L. LANCASTER et al., Characterization of Li-7~p, n!Be-7 Neutron Yields from Laser Produced Ion Beams for Fast Neutron Radiography, Phys. Plasmas, 11, 3404 ~2004!. 138. L. J. PERKINS et al., The Investigation of High Intensity Laser Driven Micro Neutron Sources for Fusion Materials Research at High Fluence, Nucl. Fusion, 40, 1 ~2000!. 139. J. M. YANG et al., Neutron Production by Fast Protons from Ultraintense Laser-Plasma Interactions, J. Appl. Phys., 96, 6912 ~2004!. 140. P. A. NORREYS et al., Neutron Production from Picosecond Laser Irradiation of Deuterated Targets at Intensities of 10~19! W cm~-2!, Plasma Phys. Control. Fusion, 40, 175 ~1998!. 141. L. DISDIER, J. P. GARCONNET, G. MALKA, and J. L. MIQUEL, Fast Neutron Emission from a High-Energy Ion Beam Produced by a High-Intensity Subpicosecond Laser Pulse, Phys. Rev. Lett., 82, 1454 ~1999!. 142. H. HABARA et al., Fast Ion Acceleration in Ultraintense Laser Interactions with an Overdense Plasma, Phys. Rev. E, 69, 036407 ~2004!. 143. H. HABARA et al., Momentum Distribution of Accelerated Ions in Ultra-Intense Laser-Plasma Interactions via Neutron Spectroscopy, Phys. Plasmas, 10, 3712 ~2003!. 144. H. HABARA et al., Ion Acceleration from the Shock Front Induced by Hole Boring in Ultraintense Laser-Plasma Interactions, Phys. Rev. E, 70, 046414 ~2004!. 145. C. TOUPIN, E. LEFEBVRE, and G. BONNAUD, Neutron Emission from a Deuterated Solid Target Irradiated by an Ultraintense Laser Pulse, Phys. Plasmas, 8, 1011 ~2001!. 146. T. DITMIRE et al., High-Energy Ions Produced in Explosions of Superheated Atomic Clusters, Nature, 386, 54 ~1997!. 147. G. GRILLON et al., Deuterium-Deuterium Fusion Dynamics in Low-Density Molecular-Cluster Jets Irradiated by Intense Ultrafast Laser Pulses, Phys. Rev. Lett., 89, 065005 ~2002!. 148. W. T. CHU, B. A. LUDEWIGT, and T. R. RENNER, Instrumentation for Treatment of Cancer Using Proton and Light-Ion Beams, Rev. Sci. Instrum., 64, 2055 ~1993!. 149. V. S. KHOROSHKOV and K. K. ONOSOVSKY, State-of-theArt in Proton Therapy Techniques, Instrum. Exp. Tech., 38, 149 ~1995!. 150. G. KRAFT, Tumortherapy with Ion Beams, Nucl. Instrum. Methods Phys. Res. A Accel. Spectrom. Dect. Assoc. Equip., 454, 1 ~2000!. 151. R. R. WILSON, Radiological Use of Fast Protons, Radiology, 47, 487 ~1946!. 152. U. AMALDI, Cancer Therapy with Particle Accelerators, Nucl. Phys. A, 654, 375C ~1999!. 153. U. AMALDI et al., LIBOA Linac-Booster for Protontherapy: Construction and Tests of a Prototype, Nucl. Instrum. Methods Phys. Res. A Accel. Spectrom. Dect. Assoc. Equip., 521, 512 ~2004!.

FUSION SCIENCE AND TECHNOLOGY

VOL. 49

APR. 2006

439

Potrebbero piacerti anche