+ All documents
Home > Documents > Investigation of longitudinal proton acceleration in exploded targets irradiated by intense...

Investigation of longitudinal proton acceleration in exploded targets irradiated by intense...

Date post: 15-Nov-2023
Category:
Upload: independent
View: 2 times
Download: 0 times
Share this document with a friend
12
Investigation of longitudinal proton acceleration in exploded targets irradiated by intense short-pulse laser M. Gauthier, A. Lévy, E. d'Humières, M. Glesser, B. Albertazzi, C. Beaucourt, J. Breil, S. N. Chen, V. Dervieux, J. L. Feugeas, P. Nicolaï, V. Tikhonchuk, H. Pépin, P. Antici, and J. Fuchs Citation: Physics of Plasmas (1994-present) 21, 013102 (2014); doi: 10.1063/1.4853475 View online: http://dx.doi.org/10.1063/1.4853475 View Table of Contents: http://scitation.aip.org/content/aip/journal/pop/21/1?ver=pdfcov Published by the AIP Publishing Articles you may be interested in Three dimensional effects on proton acceleration by intense laser solid target interaction Phys. Plasmas 20, 063107 (2013); 10.1063/1.4812458 Direct laser acceleration of electron by an ultra intense and short-pulsed laser in under-dense plasma Phys. Plasmas 18, 053104 (2011); 10.1063/1.3581062 Investigation of laser ion acceleration inside irradiated solid targets by neutron spectroscopy Phys. Plasmas 13, 030701 (2006); 10.1063/1.2177230 Spectral and dynamical features of the electron bunch accelerated by a short-pulse high intensity laser in an underdense plasma Phys. Plasmas 12, 073103 (2005); 10.1063/1.1948347 Proton acceleration mechanisms in high-intensity laser interaction with thin foils Phys. Plasmas 12, 062704 (2005); 10.1063/1.1927097 This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP: 207.162.24.136 On: Mon, 16 Feb 2015 16:15:03
Transcript

Investigation of longitudinal proton acceleration in exploded targets irradiated byintense short-pulse laserM. Gauthier, A. Lévy, E. d'Humières, M. Glesser, B. Albertazzi, C. Beaucourt, J. Breil, S. N. Chen, V. Dervieux,J. L. Feugeas, P. Nicolaï, V. Tikhonchuk, H. Pépin, P. Antici, and J. Fuchs Citation: Physics of Plasmas (1994-present) 21, 013102 (2014); doi: 10.1063/1.4853475 View online: http://dx.doi.org/10.1063/1.4853475 View Table of Contents: http://scitation.aip.org/content/aip/journal/pop/21/1?ver=pdfcov Published by the AIP Publishing Articles you may be interested in Three dimensional effects on proton acceleration by intense laser solid target interaction Phys. Plasmas 20, 063107 (2013); 10.1063/1.4812458 Direct laser acceleration of electron by an ultra intense and short-pulsed laser in under-dense plasma Phys. Plasmas 18, 053104 (2011); 10.1063/1.3581062 Investigation of laser ion acceleration inside irradiated solid targets by neutron spectroscopy Phys. Plasmas 13, 030701 (2006); 10.1063/1.2177230 Spectral and dynamical features of the electron bunch accelerated by a short-pulse high intensity laser in anunderdense plasma Phys. Plasmas 12, 073103 (2005); 10.1063/1.1948347 Proton acceleration mechanisms in high-intensity laser interaction with thin foils Phys. Plasmas 12, 062704 (2005); 10.1063/1.1927097

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

207.162.24.136 On: Mon, 16 Feb 2015 16:15:03

Investigation of longitudinal proton acceleration in exploded targetsirradiated by intense short-pulse laser

M. Gauthier,1,2 A. L�evy,1,3 E. d’Humieres,4 M. Glesser,1,5 B. Albertazzi,1,4 C. Beaucourt,4

J. Breil,4 S. N. Chen,1 V. Dervieux,1 J. L. Feugeas,4 P. Nicola€ı,4 V. Tikhonchuk,4 H. P�epin,5

P. Antici,1,5,6 and J. Fuchs1

1LULI, �Ecole Polytechnique, CNRS, CEA, UPMC, route de Saclay, 91128 Palaiseau, France2CEA, DAM, DIF, 91297 Arpajon, France3Sorbonne Universit�es, UPMC, Paris 06, CNRS, INSP, UMR 7588, F-75005, Paris, France4Univ. Bordeaux, CNRS, CEA, UMR 5107, F-33400 Talence, France5INRS-EMT, Varennes, PQ J3X 1S2, Canada6Dipartimento SBAI, Universita di Roma “Sapienza,” Via A. Scarpa 16, 00161 Rome, Italy

(Received 12 August 2013; accepted 6 December 2013; published online 13 January 2014)

It was recently shown that a promising way to accelerate protons in the forward direction to high

energies is to use under-dense or near-critical density targets instead of solids. Simulations have

revealed that the acceleration process depends on the density gradients of the plasma target.

Indeed, under certain conditions, the most energetic protons are predicted to be accelerated by a

collisionless shock mechanism that significantly increases their energy. We report here the results

of a recent experiment dedicated to the study of longitudinal ion acceleration in partially exploded

foils using a high intensity (�5� 1018 W/cm2) picosecond laser pulse. We show that protons

accelerated using targets having moderate front and rear plasma gradients (up to �8 lm gradient

length) exhibit similar maximum proton energy and number compared to proton beams that are

produced, in similar laser conditions, from solid targets, in the well-known target normal sheath

acceleration regime. Particle-In-Cell simulations, performed in the same conditions as the

experiment and consistent with the measurements, allow laying a path for further improvement of

this acceleration scheme. VC 2014 AIP Publishing LLC.

[http://dx.doi.org/10.1063/1.4853475]

I. INTRODUCTION

Laser-driven proton acceleration is a field of intense

research due to the interesting characteristics of this novel par-

ticle source including high brightness, high maximum energy,

high laminarity, and short duration (�ps at the source).1

Although the maximum ion energy, energy-spread and beam-

divergence still need to be significantly improved, the proton

characteristics are promising for many future applications,

such as in the medical field,2 for inertial confinement fusion,3

warm dense matter studies,4 or hybrid accelerators.5 Today,

existing multi-hundred-TW table-top laser systems generating

on-target intensities of�1019–1020 W/cm2 can routinely reach

proton energies of�15–20 MeV with a typical laser-to-proton

energy conversion efficiency of 1%–6%.6–8 The acceleration

mechanism commonly achieved at these laser intensities is

the so-called Target Normal Sheath Acceleration (TNSA)

mechanism,9,10 taking place at the target rear surface of

lm-scale solid foil.11 In this acceleration regime, high energy

electrons in the MeV regime (“hot electrons”) are generated

at the target front surface by J � B (Ref. 12) or vacuum heat-

ing13 for sharp front density gradient. These electrons, with a

mean free path (�few mm) larger than the typical target thick-

ness (�few tens of lm), cross the target and setup at the target

rear surface an intense accelerating sheath14 with a very

strong electrostatic field (�TV/m) which accelerates protons

and ions from contaminants adsorbed on the target rear sur-

face. Several alternative mechanisms for laser acceleration of

ions have also been studied. It is worth to mention the pro-

gress in the relativistic transparency regime/break-out after-

burner mechanism (BOA)15 made possible by the recent

improvements of the Trident laser system contrast.16 Another

mechanism is the collisionless shock acceleration (CSA). This

mechanism occurs when an over-critical density target is irra-

diated by a high intensity and very high energy laser.17,18 A

piston-like structure is launched by the laser at the target front,

producing a quasi-monoenergetic ion beam at the output. This

promising process has been experimentally demonstrated by

Palmer19 and Haberberger20 using a CO2 laser interacting

with near-critical gas targets (1019 cm�3).

An alternative scheme for laser acceleration of ions is

the acceleration from exploded targets with densities lower

than solid. Contrary to planar solid foils, exploded targets

present significant plasma density gradient on both sides. At

the rear side, density gradients decrease the electrostatic field

and degrade the ion acceleration mechanism compared to

solid targets.21–24 At the same time, the increased (compared

to a solid foil) density gradient at the front side increases the

absorption of the laser by the target, thus the hot electron

temperature, and so enhances the laser-to-ions energy con-

version efficiency.25 For a very smooth density gradient at

the rear side of the target, ions can also be accelerated in a

two-fold acceleration mechanism combining CSA and

TNSA, generating a broad energy spectrum.26,27 In contrast

to the previous CSA, the shock is launched in the back gradi-

ent by hot electrons heated by the laser: First, ions are

1070-664X/2014/21(1)/013102/11/$30.00 VC 2014 AIP Publishing LLC21, 013102-1

PHYSICS OF PLASMAS 21, 013102 (2014)

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

207.162.24.136 On: Mon, 16 Feb 2015 16:15:03

accelerated in the volume by an electric field generated by hot

electrons. This field is composed of both an inductive compo-

nent induced by the variation of a long living quasistatic mag-

netic field at the rear plasma-vacuum interface28–30 and an

electrostatic component located in the expanded sheath front

formed by the hot electrons. Due to the smooth density gradi-

ent at the rear side, the electric field monotonously decreases

with the distance from the high-density zone. Ions in the low-

density region therefore experience an electric field lower

than ions from the higher density region. As a result, ions

from the low-density region are caught by the ions coming

from the higher density region leading to the formation of an

electrostatic shock front, i.e., a peak of ion density propagat-

ing inside the decreasing (low) density ramp. The ions located

upstream are then reflected by the shock structure and acceler-

ated at velocities up to twice the shock velocity. The reflected

ions can then be accelerated again by the expansion electro-

static field which is inhomogeneous in the longitudinal and

transverse directions, leading to a broad energy spectrum.

In this paper, we investigate experimentally and numeri-

cally longitudinal ion acceleration in exploded targets irradi-

ated by an intense short (ps duration) laser pulse (SP). We use

thin solid foil targets partially exploded by a long (ns duration)

laser pulse (LP). Controlling the irradiation of the solid foil by

the LP, or varying the delay between the LP and the SP, we

are able to produce different density gradients for the plasma

targets. This way, we can study different laser ion acceleration

mechanisms and the transition between them. In particular, we

demonstrate that energetic protons can be accelerated to high

energy and high number with targets having rear gradient

characteristic length up to 8 lm. Indeed, we produce with a

good reproducibility few to several 1010 protons/sr energetic

(>1.4 MeV) protons of energies up to �8 MeV. These charac-

teristics are comparable to what we measured in the same laser

conditions and during the same experiment, in the well-known

target normal sheath acceleration regime using 10 lm gold

solid foils. We show as well that the protons accelerated using

exploded targets are emitted on a broader angle than when

using solid foils. All of these observations are found in fair

agreement with results from Particle-In-Cell (PIC) simula-

tions. The latter may allow finding the path for future optimi-

zation of ion acceleration in exploded foils as they show that,

using a longer plasma gradient and higher laser energy, CSA

would then produced higher energy protons than TNSA.

The paper is organized as follows. In Sec. II, we detail

the experimental set-up and the laser and target conditions.

Section III describes the experimental results, comparing the

proton spectra obtained when irradiating with a SP thick

solid-density planar foils (having negligible density gra-

dients) and thin targets exploded by a LP (having short to

moderate density gradients). In Sec. IV, we present simula-

tions and numerical calculations and discuss the experimen-

tal findings. Section V concludes the paper.

II. EXPERIMENTAL SET-UP

The experiment was carried out on the Etablissement

Laser de Forte Intensit�e et Energie (ELFIE) 100 TW laser fa-

cility (Laboratoire pour l’Utilisation des Lasers Intenses,

(LULI)). The experimental set-up is shown in Figure 1. A first

laser pulse, i.e., LP, s-polarized, with nominal energy E¼ 30-

40 J, Gaussian pulse duration s¼ 580 ps full width at half

maximum (FWHM), 20 lm focal spot diameter, producing

on-target intensity of �3� 1015 W/cm2 (see Figure 2) was

used to irradiate a target of variable thickness under an inci-

dent angle of 45�. As targets, we used commercially available

solid 10 lm Au and custom made plastic (Mylar) 500 nm

foils. The intensity of the LP was varied either by putting

neutral optical densities, or by defocusing it. A second laser

pulse, i.e., SP, linearly polarized, with energy of 5–8 J, 400 fs

pulse duration, 6 lm focal spot diameter (FWHM), and inten-

sity of �5� 1018 W/cm2 interacted with the exploded target,

accelerating protons in the laser-forward direction. For the

SP, the Amplified Spontaneous Emission (ASE) has been

measured to be <10�6 in intensity contrast compared to the

temporal peak of the SP and so is well below the LP intensity.

The influence of such ASE light irradiating the target prior to

the peak of the SP is therefore negligible except for the cases

where the LP intensity is strongly reduced. The SP hits the

target at normal incidence. Due to the geometry of the experi-

ment (see Figure 1), since the LP induces an asymmetric

expansion, we expect the main ion acceleration axis to be

angularly tilted with respect to the target normal axis as al-

ready observed in Ref. 26.

FIG. 1. Experimental set-up.

FIG. 2. LP temporal profile measured with an optical streak camera.

013102-2 Gauthier et al. Phys. Plasmas 21, 013102 (2014)

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

207.162.24.136 On: Mon, 16 Feb 2015 16:15:03

As diagnostics, we used two calibrated Thomson parabo-

las (TPs) located at 0� (TP A) and 15� (TP B) with respect to

the SP laser axis (see Figure 1) to measure the forward gener-

ated proton spectrum. Proton spectra measured by the TPs

were readout in an absolute manner31,32 using ImagePlates

(BAS-TR 2025 from Fuji Photo Film Co., Ltd.) that we ana-

lyzed using a FUJIFILM FLA-7000 reader. Due to technical

constraints, the detection range of TP A was limited: it could

not measure particles with energy below 1.35 MeV. Transverse

interferometry was performed using a low-energy, short (400

fs FWHM), frequency doubled (2x), optical probe beam, a

pick-off of the SP, and a Nomarski interferometer. It could

diagnose the plasma conditions and the exploded foils density

gradients. Concerning the delay between the LP and the SP, we

choose as convention that a zero delay between both pulses

corresponds to their temporal peaks hitting the target simulta-

neously whereas negative delays indicate that the peak of

the SP arrives on the target before the peak of the LP (see

Figure 3). Delays were varied from�500 to 300 ps.

III. EXPERIMENTAL RESULTS

A. Ion acceleration from planar solid foils

We first present the results of reference shots, i.e., with-

out LP (no target preformation by LP), measuring the maxi-

mum energy that can be achieved in the TNSA regime using

solid targets and these experimental conditions. We found in

the experiment that Au 10 lm foils yielded the highest pro-

ton energies for our laser conditions.33 Figure 4(a) shows

typical spectra measured at 0� (TP A) and 15� (TP B). At

normal incidence, the maximum proton energy cut-off is

around 8 MeV whereas at 15�, the cut-off is around 1.5 MeV.

Comparing the particle numbers at �1.4 MeV, we find more

than one order of magnitude difference between the 0� and

15� proton spectra confirming that the acceleration process

produces a beam that is strongly peaked in the target normal

direction.31

Thinner and lower-Z targets lead to less energetic pro-

tons.19 For example, Figure 4(b) shows spectra obtained

when irradiating 500 nm Mylar targets using the SP only.

We also observed that the acceleration process is, in this

case, less sensitive to the shot-to-shot fluctuations. However

it produces lower energy protons with cut-offs around

3-4 MeV for both TPs. Moreover, one observes that the

beam is angularly broadened compared to the thick solid tar-

get case: the cut-off energies recorded at 0� and 15� are close

to each other. This can be interpreted as follows. If one

assumes that a shock wave is launched into the plastic target

by the ablation pressure induced on the front surface by the

ASE pedestal, when it reaches the back side, it gives rise to a

dynamic expansion of the rear surface, which becomes con-

vex with a time-dependent curvature. It results in an increase

of the beam divergence. This phenomenon has already been

observed in Ref. 34 with an ASE intensity and duration simi-

lar to ours (I� 1012 W/cm2, s¼ 1 ns) using a much thicker

Al target (6 lm). We expect the effect to be even more sig-

nificant in our case since the thinner the target is, the more

sensitive it is to the prepulse.19 In addition, the time required

for the shock to pass through the target (83 ps), estimated

using the shock velocity derived in Ref. 32 (6 lm/ns), is

much shorter than our ASE duration (ns scale).

B. Ion acceleration in exploded targets

1. Plasma gradients characterization

We will first discuss the target conditions induced by

irradiating the targets by the LP. For this, we have performedFIG. 3. Intensity temporal profile of the SP and LP with a delay of Dt

between them.

FIG. 4. Proton spectra obtained at 0� (TP A) and 15� (TP B) with respect to

the target normal axis when the short pulse (I� 5� 1018 W/cm2) is interact-

ing with, respectively, a 10 lm thick gold foil (a) and a 500 nm plastic foil

(b). The spectra are all averaged over 3 different laser shots. The vertical

error bars are related to shot-to-shot variations while the horizontal ones are

both due to shot-to-shot variations and the resolution of the diagnostic (<5%

of the proton energy).

013102-3 Gauthier et al. Phys. Plasmas 21, 013102 (2014)

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

207.162.24.136 On: Mon, 16 Feb 2015 16:15:03

simulations using the two-dimensional (2D) axially symmet-

ric radiation hydrodynamic code CHIC35 reproducing our ex-

perimental conditions (e.g., using the temporal profile shown

in Figure 2 which is within the range I¼ 3� 1013–3� 1014

W/cm2, and a 45� irradiation angle). In Figure 5 are shown

2D density maps of the expanding plasma resulting from the

LP irradiation at times corresponding to a Dt of, respectively,

�350, 0, and 400 ps with respect to our convention (see Sec.

II). In Figure 5(a), one can see that for negative delays, the

target is still overdense: the laser deposits a part of its energy

in the expanding underdense front plasma and is reflected

from the target at 45� with respect to the target normal axis.

The rear density gradient starts developing and the expansion

is almost symmetric with respect to the normal target surface.

After a longer time of irradiation, one can see in Figure 5(b)

that the plasma density becomes undercritical: the LP is no

longer reflected and passes through the target. Contrary to the

front expanding plasma, the expansion of the rear plasma is

found asymmetric. The expanded plasma in the laser pulse

direction is indeed observed to be less dense and the gradient

smoother than in any other direction. At longer times (see

Figure 5(c)), the target density is well below the critical den-

sity (nc) and completely transparent for the LP that passes

though the plasma gradient.

In order to validate the density profiles obtained by the

simulations, we will now compare the front density plasma

expansions simulated with CHIC to those measured experi-

mentally by the interferometry diagnostic. In Figure 6 are

presented two different experimental electron density maps,

deduced using Abel inversion,36 at, respectively, 400 and

0 ps prior to the LP peak, with the LP intensity of, respec-

tively, 3� 1014 and 3� 1013 W/cm2. In agreement with the

Abel inversion hypothesis, the electron density map has been

symmetrized around the expansion axis. The observation

zone of our diagnostic was limited to low-density zones of

the plasma profile since at higher density the gradients are ei-

ther too strong and induce strong refraction of the probe, or

the zone is inaccessible due to the strong 2x emission. When

comparing both density maps in Figure 6, one can see that

the expanding plasma is less dense in Figure 6(b). It corre-

sponds to a longer delay and therefore to a more expanded

plasma.

Regarding the rear side plasma expansion of the target,

we were not able to detect any expanding plasma. Indeed the

non-irradiated side of the target was not sufficiently

expanded so that the plasma could be detected by our

interferometer.

In Figure 7, we compare the density profile deduced

from the interferometry images shown in Figure 6 and those

extracted from CHIC simulations for the same interaction

conditions. Although the longitudinal density range detected

by our diagnostic is very narrow, we can assess that the ex-

perimental gradient characteristic length xgl, defined by n(x)

/ exp(x/xgl), is comprised, respectively, between 15 and

45 lm in Figure 7(a) and between 30 and 48 lm in Figure

7(b). These experimental profiles are found in reasonable

agreement with the results obtained by CHIC.

Figure 8 summarizes the characteristics front-side, back-

side plasma gradient scale length and peak density of each

simulated plasma, as obtained by CHIC for the different

delays and LP intensities explored in the experiment. First,

one can see that all gradient characteristic lengths strongly

increase with the delay and laser intensity. The target front

FIG. 5. Superposed electron density (in logarithmic scale) and LP absorption

maps obtained using CHIC 2D for LP intensity of 3� 1013 W/cm2 and

delays of, respectively, �350 (a), 0 (b), and þ400 ps (c). The LP propaga-

tion direction is indicated by the white arrows.

FIG. 6. Density maps obtained by Abel inversion from interferometry

images. These density profiles correspond to a 500 nm Mylar target irradi-

ated by a LP of, respectively, (a) I� 3� 1014 W/cm2 with Dt¼�400 ps and

(b) I� 3� 1013 W/cm2 with Dt¼ 0 ps. The maps correspond to the

low-density zone accessible to the optical probe. Absolute positions along

x-axis are consistent with the CHIC simulations shown in Figure 7.

013102-4 Gauthier et al. Phys. Plasmas 21, 013102 (2014)

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

207.162.24.136 On: Mon, 16 Feb 2015 16:15:03

surface, i.e., the surface facing the laser, is heated earlier in

time and to higher temperature compared to the rear surface.

Thus, it is obviously more expanded and exhibits a longer

gradient characteristic length. Second, we observe that the

maximum density of the exploded target drops with the delay

and intensity reaching sub-critical densities at delays from

Dt�þ100 ps (corresponding to the LP highest intensity) to

þ500 ps, (corresponding to the LP lowest intensity). It

FIG. 7. Density profiles corresponding to a 500 nm Mylar target irradiated by a LP of, respectively, (a) I� 3� 1014 W/cm2 with Dt¼�400 ps and (b)

I� 3� 1013 W/cm2 with Dt¼ 0 ps. Blue dots indicate the density profiles obtained by the interferometry diagnostic; red lines by CHIC simulations.

FIG. 8. Evolution of, respectively, the

front gradient characteristic length

(length to reach a decrease of the

density by a factor exp (1)), the rear

gradient characteristic length, and

the maximum electron density of the

exploded foil in the CHIC simulations

as a function of the delay for LP inten-

sities of I¼ 3� 1013 (blue plain curve),

I¼ 1014 (green dashed curve), and

I¼ 3� 1014 W/cm2 (red dotted curve).

013102-5 Gauthier et al. Phys. Plasmas 21, 013102 (2014)

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

207.162.24.136 On: Mon, 16 Feb 2015 16:15:03

should be noted that CHIC simulations do not take into

account the ASE of the SP. This is justified since the expan-

sion induced by the ASE is negligible compared to the

expansion caused by the LP in the delays covered by our

experiment.

2. Proton spectra

Using two different LP intensities and various Dt, we

have measured the proton spectra accelerated by the SP

(I� 5� 1018 W/cm2) irradiating exploded targets having rear

plasma gradients with characteristic length up to 8 lm

according to CHIC simulations. As can be seen in Figure 9,

an increase of the density gradient does not decrease signifi-

cantly the maximum proton energy (or energy cut-off). We

notably observed that such targets lead to similar proton

energies than the best maximum proton energy obtained

without any target preformation and this for both observation

angles (0� and 15�).In the TNSA regime, we would expect that, with such

density gradients, the energy cut-off decreases notably.37,38

In addition, we observed that for certain gradient characteris-

tic lengths such as 0.4 or 2.4 lm, the most energetic protons

are found at 15� from the target normal, rather than at 0�.For example, as shown in Figure 10, we measured at 0� and

15� for a LP intensity I¼ 3� 1013 W/cm2 and Dt¼�400 ps

(corresponding to a gradient characteristics length of 0.4 lm)

a higher proton energy cutoff on the TP positioned at 15�

(�7–8 MeV); however, on the TP located at 0� we only

measured 2.8–3.7 MeV. This indicates that the acceleration

mechanism is affected by the asymmetric expansion caused

by the LP. While the maximum proton energy is dependent

on the electrostatic field, the angle with which the protons

stem out of the target strongly depends on the geometry of

the sheath field since ions are accelerated normally to the

isopotential.1 We also observed that the most energetic pro-

tons generated using exploded targets have energies that are

similar—if not better—to those obtained in the best TNSA

conditions using solid targets. In addition, when comparing

the different spectra, we observe that the number of high

energy protons (>1.4 MeV) is also comparable to the one

measured in the case of a 10 lm solid-density target. For

example, we calculated from the spectra that we produced

�6.2� 1010 protons/sr above 1.4 MeV at 15� with a 500 nm

thick plastic foil exploded by a ns laser pulse. These numbers

are close to the proton beams we produced with a 10 lm

gold foil where we recorded �2� 1011 protons/sr above

1.4 MeV at 0� (see Figure 4(a)).

In order to further investigate the beam angular charac-

teristics, we tilted the target at 7�, 15�, 22�, and 30� with

respect to the initial target surface as shown in Figure 11 in

which the rotation angle is noted h. We kept the other param-

eters fixed (LP intensity I¼ 3� 1013 W/cm2 and

Dt¼�400 ps), since in this configuration we had experimen-

tally obtained the highest proton energies. Note that due to

geometrical effects, the rotation of the target changes the LP

intensity by a factor cos(45�-h), inducing an intensity varia-

tion �26%, and thus an electron temperature variation �9%

according to the scaling in Ref. 39. We assume this effect to

be negligible with respect to the shot-to-shot fluctuations of

the laser beam.

A comparison of the proton energy cut-offs measured for

different target angles is presented in Figure 12. It should be

noted that no protons were detected at 0� (TP A) when the

target was tilted at 15�, 22�, and 30� with respect to the initial

target surface. It means that for these shots, the maximum

FIG. 10. Proton spectra measured at 0� and 15� when irradiating 500 nm

plastic foils by the LP of intensity I¼ 3� 1013 W/cm2 with Dt¼�400 ps.

FIG. 11. Angle of the target compared to the SP axis.

FIG. 9. Proton energy cut-off obtained using 500 nm Mylar exploded foil as

a function of the characteristic length of the target rear density gradient. The

proton energy cut-offs measured without LP, i.e., with negligible target pre-

formation either with a 500 nm Mylar foils or a 10 lm Au (hashed zones

delimited by, respectively, plain and dashed lines) are added for

comparison.

013102-6 Gauthier et al. Phys. Plasmas 21, 013102 (2014)

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

207.162.24.136 On: Mon, 16 Feb 2015 16:15:03

proton energy did not reach more than 1.35 MeV. The most

energetic protons are clearly produced close to the target nor-

mal axis, but energetic protons are still found with angles up

to 15�. The accelerated ion beam is therefore not isotropic,

but less collimated than the one produced from solid targets.

This decrease of the beam collimation could be explained by

the geometry of the SP. Indeed, according to Ref. 27, the pro-

tons are expected to be preferentially emitted at an angle

lying in between the direction of the plasma expansion gener-

ated by the LP and the direction of the SP propagation: the

long living magnetic field induced by the hot electron current

generated by the SP slightly deflects the beam from the direc-

tion of the plasma expansion. In our experiment, the direction

of the plasma expansion is given by the axis normal to the tar-

get, w direction in Figure 11 while the SP remains collinear

to x. The proton beams direction is therefore tilted from w to

x. This effect is not visible when both the SP and the plasma

expansion are pointing in the same direction (w and x are col-

linear, h¼ 0), but becomes perceptible when the angle

between the two axis increases. Nevertheless, the fact that we

clearly observe an energy cut-off decrease when increasing

the angle h means that the effect of the long living magnetic

fields is not strong enough to overcome the direction induced

by the plasma expansion generated by the LP. It could be the

case with a higher intensity or a more energetic SP producing

a higher number of hot electrons and thus generating a stron-

ger magnetic field.

It should also be noted that the highest energy is found

in TP B at 15� in the case where the target is positioned at

0�, i.e., when the target surface is normal to the SP. This cor-

responds to an asymmetric expansion of the rear plasma: the

plasma expansion direction is slightly modified in the direc-

tion of the LP propagation axis, tilting the proton beam in

the direction given by TP B.

IV. SIMULATIONS AND DISCUSSION

To study the transition between the sharp and the long

gradient regime and to understand the possibility of

obtaining high energy protons for various gradient condi-

tions, we have performed 2D PIC simulations with the

PICLS code.40 Depending on the gradient characteristic

length, several regimes, already studied theoretically, can be

obtained. (1) For sharp density gradients, classical TNSA is

dominant even when irradiating lower-than-solid density tar-

gets.41 (2) For longer gradients, on both sides of the target, a

trade-off takes place between the laser absorption increase

on the interaction side and the decrease of the TNSA electric

field on the rear side due to the lengthening of the density

gradient. The acceleration regime is similar to the gradient

regime studied by Ref. 19 in which the maximum proton

energy decreases with increasing plasma gradient length. (3)

When the target starts to become transparent during the inter-

action, the laser-to-target coupling is significantly improved.

The increase of the hot electron temperature can counterbal-

ance the effect of the characteristic scale length in the peak

electric field formula E ¼ kbTh

eklss

ffiffiffiffiffiffiffiffiffiffi2

expð1Þ

q(where Th is the hot

electrons temperature, kb the Boltzmann constant, e the

charge of an electron, and klss the density gradient scale

length) and lead to higher energy protons.23 In this regime,

for long rear side density gradients, a secondary accelerating

mechanism can occur through a strong electrostatic

shock15,19 and enhance the maximum proton energy. This

rear-side shock is different from front-side collisionless

shocks already observed when the laser is reflected at the

front of a dense target15 and from the BOA, observed when

using a higher contrast SP laser, coupled to very thin tar-

gets.13 (4) For too long gradients, obtained with a strongly

expanded target, the laser-to-target coupling decreases since

the target becomes too low density and thus too transparent

to the laser light for producing energetic electrons.

The PIC simulations illustrating these regimes were per-

formed using a 1.057 lm wavelength, 350 fs (FWHM) dura-

tion, p-polarized laser pulse focused on a 10 lm diameter

spot. The spatial and temporal profiles are truncated

Gaussians. The laser intensity of 8� 1018 W/cm2 is similar

to the intensity of the SP in our experiment. The pulse is

injected from the left side of the simulation box.

The density profile in the first simulation corresponds to

a long negative delay between the two pulses reproducing

the case of the SP interacting almost directly with the thin

foil used in the experiments. We used the truncated (as can

be seen in Figure 13(a)) CHIC density profile obtained for a

delay of �500 ps and a LP intensity of 3� 1013 W/cm2: the

gradient characteristic length of the rear plasma profile is

below 1 lm. Since the laser energy and the small target

thickness lead to a quick heating of the target, collisions

inside the plasma can be neglected. The target is composed

of C6þ ions, protons, and electrons with a 300 nc maximum

electron density. The initial electron, proton, and carbon ion

temperatures are set to zero. The plasma is located 70 lm

away from the left side in the 176 by 160 lm simulations

box. The spatial and temporal steps are respectively

Dx¼Dy¼ 15 nm and Dt¼ 0.05 fs. The boundary conditions

we used are absorbing in x and y. The density profile we

used and the obtained proton phase space on axis are shown

respectively in Figures 13(a) and 13(b).

FIG. 12. Proton energy cut-off recorded at 0� (TP A) and 15� (TP B) to the

SP axis as a function of the angle between the normal of the irradiated target

and the SP propagation axis.

013102-7 Gauthier et al. Phys. Plasmas 21, 013102 (2014)

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

207.162.24.136 On: Mon, 16 Feb 2015 16:15:03

This simulation exhibits a standard TNSA acceleration

process with the maximum energy protons coming from the

back surface and a strong laser absorption leading to high

energy protons. The maximum proton energy is here

15.6 MeV. This is higher than our best measurement per-

formed in the TNSA regime, i.e., obtained using 500 nm

Mylar target and the sole SP (see Sec. III A). However, this

was expected since 2D simulations are known to overesti-

mate the final proton energy. These observations indicate

that we are in the regime (1) described before.

In a second simulation, we used (as initial density pro-

file) for the PIC simulation the CHIC density profile obtained

with a LP intensity of 1014 W/cm2 and Dt¼þ100 ps. In this

case, the density profile corresponds to a lower density, with

a 2.53 nc maximum electron density, and larger plasma with

smoother gradients (see Figure 14(a)): the gradient charac-

teristic length of the rear plasma profile is around 28 lm. We

are dealing here with a plasma gradient that is much longer

that those we investigated experimentally. The simulation

parameters are the same as above except for the following

parameters. The target has a maximum electron density

slightly higher than the critical density. The plasma is

located 100 lm away from the left side in the 800 by 128 lm

simulation box. The spatial and temporal steps are, respec-

tively, Dx¼Dy¼ 100 nm and Dt¼ 0.33 fs.

The obtained proton phase space along the x-axis is

shown in Figure 14(b). In this simulation, the maximum

energy protons are also accelerated by a strong electrostatic

field in the decreasing density gradient, but a secondary

acceleration process occurs. The first step is not as efficient

as in the case described in Figure 13. The long density gradi-

ent at the back side of the target will lead to a lower acceler-

ating electric field, but the strong laser-to-target coupling can

compensate this decrease and still leads to high energy pro-

tons. In our case, this first step leads to a maximum proton

velocity of around 0.1 c. In a second step, protons are accel-

erated further in the gradient by an electrostatic shock over a

short distance which leads to a distinct feature in the proton

phase space that is clearly visible at x¼ 660 lm in Figure

14(b). In this case the maximum proton energy at saturation

is 13.2 MeV, which is very close to the maximum energy

obtained in the case described above in Figure 13. This case

corresponds to the regime (3) described before.

Depending on the density gradient at the back side of

the target, it is therefore possible to accelerate ions to similar

energies than when using sharp rear gradients target (com-

pare Figures 13 and 14) but through different processes.

Figures 15(a) and 15(b), respectively, show the forward

proton energy spectrum obtained in the second simulation

(regime (3)) and the associated divergence distribution for

protons with energy higher than 1 MeV. Overlaid are the

same plots obtained with a 500 nm thick (non-exploded)

solid foil (i.e., in the TNSA regime (1)). The divergence dis-

tribution in regime (3) is more complex than in regime (1).

The divergence is broader, and additional features are

observed at �35� and for angles higher than 60�. This latest

feature is even observed for protons with energy higher than

1 MeV corresponding to protons that are accelerated almost

FIG. 14. (a) Density profile obtained

with the CHIC code for a 500 nm

Mylar foil exploded with a 1014 W/cm2

intensity and aþ 100 ps delay. (b)

Normalized proton phase space lineout

in the longitudinal direction 7.42 ps af-

ter the interaction of the maximum of

the SP. The laser comes from the left.

FIG. 13. (a) Density profile obtained

with the CHIC code for an exploding

foil with a 3� 1013 W/cm2 intensity

(LP) and Dt¼�500 ps. (b) Normalized

proton phase space lineout in the longi-

tudinal direction (along the x-axis) 0.6

and 1 ps after the interaction of the

maximum of the SP (m is the proton

mass and c the speed of light). The

laser comes from the left.

013102-8 Gauthier et al. Phys. Plasmas 21, 013102 (2014)

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

207.162.24.136 On: Mon, 16 Feb 2015 16:15:03

perpendicular to the laser axis in the channel created by the

laser through the Coulomb explosion mechanism.42

To study the transition between these two efficient laser

proton acceleration regimes (regimes (1) and (3)) and to

compare with our experimental results, we have simulated

intermediate density gradients. Here, we decreased the

plasma length used for the case illustrated in Figure 14 by

changing the rear surface profile characteristics, starting

from the CHIC profile obtained for a LP intensity of

1014 W/cm2 and a delay of þ100 ps. In Figure 16, we show

the proton phase spaces obtained at two simulation times for

characteristic gradient lengths of 5 and 10 lm, i.e., at the

frontier between regimes (2) and (3). When decreasing the

gradient scale length to 10 lm, the laser coupling with the

target is still high, but the electrostatic field in the density

gradient is decreased. This leads to moderately high energy

protons (5.2 MeV) as evidenced by the lower maximum

velocities reached and shown in Figure 16(a). In this case,

the coupling efficiency is not enough to compare to either

TNSA with thin targets (regime (1), Figure 13) or shock

acceleration with exploded foils (regime (3), Figure 14).

Figure 16(b) corresponds to the case of a 5 lm scale

length for which the laser coupling is even more decreased

and will lead to lower energy protons (2.8 MeV). The higher

coupling efficiency in the 10 lm scale length case is also evi-

denced by the enhanced backward (towards the laser) accel-

eration visible in Figure 16(a) compared to Figure 16(b).

When comparing the evolution from a steep back sur-

face density gradient in Figure 13 to a longer and smoother

one in Figures 14 and 16, it is clear that it is possible to

accelerate high energy protons on a wide range of target pa-

rameters in the near-critical laser ion acceleration regime.

This is very similar to our experimental observation that

MeV protons can be generated with various delays and vari-

ous LP intensities.

In addition, these simulations indicate that there are two

optimum regimes for laser proton acceleration: the thin tar-

get TNSA and the exploded foil CSA, generating similar

maximum proton energies. For thin target TNSA, high laser

contrast is necessary. One drawback of this regime is that the

various techniques to improve the laser contrast also

decrease the total laser energy in the focal spot (it is the case

using plasma mirrors43 or non-linear systems laser frequency

doubling). For exploded foil shock acceleration, a small

range of density gradient scale length can lead to this regime.

According to simulations, in between, a large range of target

parameters will lead to moderately high energy protons

through degraded TNSA.

V. CONCLUSION/PERSPECTIVES

We have studied the longitudinal acceleration of protons

using a high intensity picosecond laser pulse irradiating a

500 nm thick plastic foil exploded by a nanosecond pulse.

We varied the nanosecond pulse intensity and the delay

between the picosecond and the nanosecond pulses to study

the acceleration mechanism in various target conditions hav-

ing short to moderate plasma density gradients, i.e., with

FIG. 15. Comparison of (a) the proton

energy spectra in the forward direction

for all angles, and (b) the divergence

distribution for protons with energy

>1 MeV in the forward direction in the

case of the SP interaction with a

500 nm plastic foil non-exploded and

one exploded with a 1014 W/cm2 inten-

sity and aþ 100 ps delay.

FIG. 16. Proton phase space on axis at

3.96 ps (blue) and at 5.28 ps (red) after

the beginning of the PIC simulation in

the case of the SP interacting with an

exploded foil with a rear surface gradi-

ent with characteristic lengths of

10 lm (a) and 5 lm (b).

013102-9 Gauthier et al. Phys. Plasmas 21, 013102 (2014)

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

207.162.24.136 On: Mon, 16 Feb 2015 16:15:03

gradient characteristic lengths from 0 to 8 lm. We have

shown, in agreement with PIC simulations, that, in these gra-

dient density ranges, protons can be accelerated in number

and energy similar to the one recorded when using thin solid

planar foils in the TNSA regime. In addition, we highlighted

that the use of exploded targets generates a broader angular

proton distribution compared to solid targets and accelera-

tion anisotropy. The requirements for such acceleration

mechanism to occur are (1) high laser intensity to initiate a

strong expansion in the decreasing density gradient, as well

as (2) a density profile tailored so the plasma at the back is

long enough for ions to be reflected, and (3) a high laser

energy to be able to propagate in a near-critical target so that

energetic electrons reach the back gradient to initiate a

strong expansion that is also necessary to reach ion reflec-

tion. We expect that at higher SP energies we would be able

to achieve higher electric fields in longer gradients. This

should improve underdense laser ion acceleration and enable

to observe clearer the CSA acceleration process.27,44 As a

consequence, this should allow obtaining higher energy pro-

tons than with the conventional TNSA regime.

ACKNOWLEDGMENTS

We acknowledge support from the LULI technical teams.

We thank Professor Yasuhiko Sentoku for usage of the code

PICLS and Dr. T. Vinci for usage of the code neutrino to ana-

lyze the interferometry measurements. This work was sup-

ported by the European Commission (CRISP FP-7 Contract

No. 283745) (P. Antici), FRQNT/FRQSC (Grant No. 174726)

(P. Antici), CRSNG decouverte (Grant No. 435416) and FP-7

Laserlab-Europe (Grant Agreement No. 602 284464 and

Grant No. 001528) (M. Gauthier, A. L�evy, M. Glesser, B.

Albertazzi, S. N. Chen, V. Dervieux, P. Antici, J. Fuchs),

National Science Foundation (Grant No. 600 1064468) (S. N.

Chen), HPC resources of CINES under allocations 2011-604

056129, 2012-056129, and 2013-056129 made by GENCI

(Grand Equipement National de Calcul Intensif) (E.

d’Humieres, C. Beaucourt, J. Breil, J. L. Feugeas, P. Nicolai,

V. Tikhonchuk). Support from the Aquitaine Regional

Council is also acknowledged.

1T. E. Cowan, J. Fuchs, H. Ruhl, A. Kemp, P. Audebert, M. Roth, R.

Stephens, I. Barton, A. Blazevic, E. Brambrink, J. Cobble, J. Fernandez, J.

C. Gauthier, M. Geissel, M. Hegelich, J. Kaae, S. Karsch, G. P. Le Sage,

S. Letzring, M. Manclossi, S. Meyroneinc, A. Newkirk, H. Pepin, and N.

Renard-LeGalloudec, Phys. Rev. Lett. 92, 204801 (2004).2S. V. Bulanov and V. S. Khoroshkov, Plasma Phys. Rep. 28, 453 (2002);

V. Malka, S. Fritzler, E. Lefebvre, E. d’Humieres, R. Ferrand, G. Grillon,

C. Albaret, S. Meyroneinc, J.-P. Chambaret, A. Antonetti, and D. Hulin,

Med. Phys. 31, 1587 (2004).3M. Roth, T. E. Cowan, M. H. Key, S. P. Hatchett, C. Brown, W. Fountain,

J. Johnson, D. M. Pennington, R. A. Snavely, S. C. Wilks, K. Yasuike, H.

Ruhl, F. Pegoraro, S. V. Bulanov, E. M. Campbell, M. D. Perry, and H.

Powell, Phys. Rev. Lett. 86, 436 (2001).4A. Mancic, J. Robiche, P. Antici, P. Audebert, C. Blancard, P. Combis, F.

Dorchies, G. Faussurier, S. Fourmaux, M. Harmand, R. Kodama, L.

Lancia, S. Mazevet, M. Nakatsutsumi, O. Peyrusse, V. Recoules, P.

Renaudin, R. Shepherd, and J. Fuchs, High Energy Dens. Phys. 6, 21

(2010).5P. Antici, M. Migliorati, A. Mostacci, L. Picardi, L. Palumbo, and C.

Ronsivalle, Phys. Plasmas 18, 073103 (2011); P. Antici, M. Fazi, A.

Lombardi, M. Migliorati, L. Palumbo, P. Audebert, and J. Fuchs, J. Appl.

Phys. 104, 124901 (2008).6S. Buffechoux, J. Psikal, M. Nakatsutsumi, L. Romagnani, A. Andreev, K.

Zeil, M. Amin, P. Antici, T. Burris-Mog, A. Compant-La-Fontaine, E.

D’Humieres, S. Fourmaux, S. Gaillard, F. Gobet, F. Hannachi, S. Kraft, A.

Mancic, C. Plaisir, G. Sarri, M. Tarisien, T. Toncian, U. Schramm, M.

Tampo, P. Audebert, O. Willi, T. E. Cowan, H. Pepin, V. Tikhonchuk, M.

Borghesi, and J. Fuchs, Phys. Rev. Lett. 105, 015005 (2010).7S. Fourmaux, S. Buffechoux, B. Albertazzi, D. Capelli, A. L�evy, S.

Gnedyuk, L. Lecherbourg, P. Lassonde, S. Payeur, P. Antici, H. P�epin, R.

S. Marjoribanks, J. Fuchs, and J. C. Kieffer, Phys. Plasmas 20, 013110

(2013).8K. Zeil, S. D. Kraft, S. Bock, M. Bussmann, T. E. Cowan, T. Kluge, J.

Metzkes, T. Richter, R. Sauerbrey, and U. Schramm, New J. Phys. 12,

045015 (2010).9S. P. Hatchett, C. G. Brown, T. E. Cowan, E. A. Henry, J. S. Johnson, M.

H. Key, J. A. Koch, A. B. Langdon, B. F. Lasinski, R. W. Lee, A. J.

Mackinnon, D. M. Pennington, M. D. Perry, T. W. Phillips, M. Roth, T. C.

Sangster, M. S. Singh, R. A. Snavely, M. A. Stoyer, S. C. Wilks, and K.

Yasuike, Phys. Plasmas 7, 2076 (2000).10S. C. Wilks, A. B. Langdon, T. E. Cowan, M. Roth, M. Singh, S. Hatchett,

M. H. Key, D. Pennington, A. MacKinnon, and R. A. Snavely, Phys.

Plasmas 8, 542 (2001).11J. Fuchs, Y. Sentoku, E. d’Humieres, T. E. Cowan, J. Cobble, P. Audebert,

A. Kemp, A. Nikroo, P. Antici, E. Brambrink, A. Blazevic, E. M.

Campbell, J. C. Fern�andez, J.-C. Gauthier, M. Geissel, M. Hegelich, S.

Karsch, H. Popescu, N. Renard-LeGalloudec, M. Roth, J. Schreiber, R.

Stephens, and H. P�epin, Phys. Plasmas 14, 053105 (2007).12S. C. Wilks, W. L. Kruer, M. Tabak, and A. B. Langdon, Phys. Rev. Lett.

69, 1383 (1992).13F. Brunel, Phys. Rev. Lett. 59, 52 (1987).14P. Antici, J. Fuchs, M. Borghesi, L. Gremillet, T. Grismayer, Y.

Sentoku, E. d’Humieres, C. A. Cecchetti, A. Mancic, A. C. Pipahl, T.

Toncian, O. Willi, P. Mora, and P. Audebert, Phys. Rev. Lett. 101,

105004 (2008).15L. Yin, B. J. Albright, B. M. Hegelich, and J. C. Fernandez, Laser Part.

Beams 24, 291 (2006); L. Yin, B. J. Albright, B. M. Hegelich, K. J.

Bowers, K. A. Flippo, T. J. T. Kwan, and J. C. Fern�andez, Phys. Plasmas

14, 056706 (2007).16B. M. Hegelich, D. Jung, B. J. Albright, M. Cheung, B. Dromey, D. C.

Gautier, C. Hamilton, S. Letzring, R. Munchhausen, S. Palaniyappan, R.

Shah, H.-C. Wu, L. Yin, and J. C. Fern�andez, “160 MeV laser-accelerated

protons from CH2 nano-targets for proton cancer therapy,” e-print

arXiv:1310.8650.17L. O. Silva, M. Marti, J. R. Davies, and R. A. Fonseca, Phys. Rev. Lett.

92, 015002 (2004).18F. Fiuza, A. Stockem, E. Boella, R. A. Fonseca, L. O. Silva, D.

Haberberger, S. Tochitsky, C. Gong, W. B. Mori, and C. Joshi, Phys. Rev.

Lett. 109, 215001 (2012).19C. A. Palmer, N. P. Dover, I. Pogorelsky, M. Babzien, G. I. Dudnikova,

M. Ispiriyan, M. N. Polyanskiy, J. Schreiber, P. Shkolnikov, V.

Yakimenko, and Z. Najmudin, Phys. Rev. Lett. 106, 014801 (2011).20D. Haberberger, S. Tochitsky, F. Fiuza, C. Gong, R. A. Fonseca, L. O.

Silva, W. B. Mori, and C. Joshi, Nature Phys. 8, 95 (2012).21M. C. Kaluza, J. Schreiber, M. I. K. Santala, G. D. Tsakiris, K. Eidmann,

J. Meyer-ter-Vehn, and K. J. Witte, Phys. Rev. Lett. 93, 045003 (2004).22T. Grismayer and P. Mora, Phys. Plasmas 13, 032103 (2006).23J. Fuchs, C. A. Cecchetti, M. Borghesi, T. Grismayer, E. d’Humieres, P.

Antici, S. Atzeni, P. Mora, A. Pipahl, L. Romagnani, A. Schiavi, Y. Sentoku,

T. Toncian, P. Audebert, and O. Willi, Phys. Rev. Lett. 99, 015002 (2007).24A. L�evy, R. Nuter, T. Ceccotti, P. Combis, M. Drouin, L. Gremillet, P.

Monot, H. Popescu, F. R�eau, E. Lefebvre, and P. Martin, New J. Phys. 11,

093036 (2009).25P. Antici, J. Fuchs, E. d’Humieres, J. Robiche, E. Brambrink, S. Atzeni, A.

Schiavi, Y. Sentoku, P. Audebert, and H. P�epin, New J. Phys. 11, 023038

(2009).26E. d’Humi�eres, J. L. Feugeas, P. Nicola€ı, S. Gaillard, T. Cowan, Y.

Sentoku, and V. Tikhonchuk, J. Phys.: Conf. Ser. 244, 042023 (2010).27E. d’Humieres, P. Antici, M. Glesser, J. Boeker, F. Cardelli, S. Chen, J. L.

Feugeas, F. Filippi, M. Gauthier, A. Levy, P. Nicola€ı, H. P�epin, L.

Romagnani, M. Scisci�o, V. T. Tikhonchuk, O. Willi, J. C. Kieffer, and J.

Fuchs, Plasma Phys. Controlled Fusion 55, 124025 (2013).28S. V. Bulanov, D. V. Dylov, T. Zh. Esirkepov, F. F. Kamenets, and D. V.

Sokolov, Plasma Phys. Rep. 31, 369 (2005).

013102-10 Gauthier et al. Phys. Plasmas 21, 013102 (2014)

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

207.162.24.136 On: Mon, 16 Feb 2015 16:15:03

29A. Yogo, H. Daido, S. V. Bulanov, K. Nemoto, Y. Oishi, T. Nayuki, T.

Fujii, K. Ogura, S. Orimo, A. Sagisaka, J.-L. Ma, T. Zh. Esirkepov, M.

Mori, M. Nishiuchi, A. S. Pirozhkov, S. Nakamura, A. Noda, H.

Nagatomo, T. Kimura, and T. Tajima, Phys. Rev. E 77, 016401 (2008).30T. Nakamura, J. K. Koga, T. Z. Esirkepov, M. Kando, G. Korn, and S. V.

Bulanov, Phys. Rev. Lett. 108, 195001 (2012).31A. Mancic, J. Fuchs, P. Antici, S. A. Gaillard, and P. Audebert, Rev. Sci.

Intrum. 79, 073301 (2008).32T. Bonnet, M. Comet, D. Denis-Petit, F. Gobet, F. Hannachi, M. Tarisien, M.

Versteegen, and M. M. Al�eonard, Rev. Sci. Instrum. 84, 013508 (2013).33A. Mancic, Ph.D. thesis, Ecole Polytechnique, Palaiseau, 2010.34F. Lindau, O. Lundh, A. Persson, P. McKenna, K. Osvay, D. Batani, and

C.-G. Wahlstr€om, Phys. Rev. Lett. 95, 175002 (2005).35J. Breil, P. Maire, Ph. Nicolai, and G. Schurtz, J. Phys.: Conf. Ser. 112,

022035 (2008).36K. Bockasten, J. Opt. Soc. Am. 51, 943 (1961).37J. Fuchs, P. Antici, E. d’Humieres, E. Lefebvre, M. Borghesi, E.

Brambrink, C. Cecchetti, T. Toncian, H. P�epin, and P. Audebert, J. Phys.

IV (France) 133, 1151 (2006).

38P. Antici, J. Fuchs, E. d’Humieres, E. Lefebvre, M. Borghesi, E.

Brambrink, C. A. Cecchetti, S. Gaillard, L. Romagnani, Y. Sentoku, T.

Toncian, O. Willi, P. Audebert, and H. P�epin, Phys. Plasmas 14, 030701

(2007).39C. Fauquignon and F. Floux, Phys. Fluids 13, 386 (1970); J. L. Bobin,

ibid. 14, 2341 (1971).40Y. Sentoku and A. J. Kemp, J. Comp. Phys. 227, 6846 (2008).41L. Willingale, S. P. D. Mangles, P. M. Nilson, R. J. Clarke, A. E. Dangor,

M. C. Kaluza, S. Karsch, K. L. Lancaster, W. B. Mori, Z. Najmudin, J.

Schreiber, A. G. R. Thomas, M. S. Wei, and K. Krushelnick, Phys. Rev.

Lett. 96, 245002 (2006).42K. Krushelnick, E. L. Clark, Z. Najmudin, M. Salvati, M. I. K.

Santala, M. Tatarakis, and A. E. Dangor, Phys. Rev. Lett. 83, 737

(1999).43A. L�evy, T. Ceccotti, P. D’Oliveira, F. R�eau, M. Perdrix, F. Qu�er�e, P.

Monot, M. Bougeard, H. Lagadec, P. Martin, J.-P. Geindre, and P.

Audebert, Opt. Lett. 32, 310 (2007).44E. d’Humieres and V. Tikhonchuk, AIP Conf. Proc. 1299, 704

(2010).

013102-11 Gauthier et al. Phys. Plasmas 21, 013102 (2014)

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

207.162.24.136 On: Mon, 16 Feb 2015 16:15:03


Recommended