Hostname: page-component-8448b6f56d-42gr6 Total loading time: 0 Render date: 2024-04-20T12:33:13.232Z Has data issue: false hasContentIssue false

Effective laser driven proton acceleration from near critical density hydrogen plasma

Published online by Cambridge University Press:  15 February 2016

Ashutosh Sharma*
Affiliation:
ELI-ALPS, Szeged, Hungary
Alexander Andreev
Affiliation:
ELI-ALPS, Szeged, Hungary Max-Born Institute, Berlin, Germany
*
Address correspondence and reprint requests to: Ashutosh Sharma, ELI-ALPS, Szeged, Hungary. E-mail: ashutosh.sharma@eli-alps.hu
Rights & Permissions [Opens in a new window]

Abstract

Recent advances in the production of high repetition, high power, and short laser pulse have enabled the generation of high-energy proton beam, required for technology and other medical applications. Here we demonstrate the effective laser driven proton acceleration from near-critical density hydrogen plasma by employing the short and intense laser pulse through three-dimensional (3D) particle-in-cell (PIC) simulation. The generation of strong magnetic field is demonstrated by numerical results and scaled with the plasma density and the electric field of laser. 3D PIC simulation results show the ring shaped proton density distribution where the protons are accelerated along the laser axis with fairly low divergence accompanied by off-axis beam of ring-like shape.

Type
Research Article
Copyright
Copyright © Cambridge University Press 2016 

1. INTRODUCTION

Beams of high-energy ions and protons could be produced by directing a high power ultra-short laser pulse onto a thin target, as investigated in 2000 by Snavely et al. (Reference Snavely, Key, Hatchett, Cowan, Roth, Phillips, Stoyer, Henry, Sangster, Singh, Wilks, MacKinnon, Offenberger, Pennington, Yasuike, Langdon, Lasinski, Johnson, Perry and Campbell2000). Acceleration takes place at the target–vacuum interfaces where laser-accelerated relativistic electrons form a dense electron plasma sheath (field ~TV/m) that ionizes surface atoms and accelerates ions in the target normal. These beams have extreme laminarity, ultra-short duration, and high particle number per bunch that distinguish them from beams produced by conventional sources (e.g., accelerators). In particular the medical applications of such beams give rise to the field of radiation therapy (Loeffler& Durante, Reference Loeffler and Durante2013). Recent progress in generating the high-energy (>50 MeV) ions from intense laser–matter interactions (1018–1021 Wcm−2; (Hatchett et al., Reference Hatchett, Brown, Cowan, Henry, Johnson, Key, Koch, Langdon, Lasinski, Lee, Mackinnon, Pennington, Perry, Phillips, Roth, Sangster, Singh, Snavely, Stoyer, Wilks and Yasuike2000; Snavely et al., Reference Snavely, Key, Hatchett, Cowan, Roth, Phillips, Stoyer, Henry, Sangster, Singh, Wilks, MacKinnon, Offenberger, Pennington, Yasuike, Langdon, Lasinski, Johnson, Perry and Campbell2000; Wilks et al., Reference Wilks, Langdon, Cowan, Roth, Singh, Hatchett, Key, Pennington, MacKinnon and Snavely2001; Borghesi et al., Reference Borghesi, Mackinnon, Campbell, Hicks, Kar, Patel, Price, Romagnani, Schiavi and Willi2004; Fuchs et al., Reference Fuchs, Antici, d'Humières, Lefebvre, Borghesi, Brambrink, Cecchetti, Kaluza, Malka, Manclossi, Meyroneinc, Mora, Schreiber, Toncian, Pépin and Audebert2006; Hegelich et al., Reference Hegelich, Albright, Cobble, Flippo, Letzring, Paffett, Ruhl, Schreiber, Schulze and Fernández2006; Robson et al., Reference Robson, Simpson, Clarke, Ledingham, Lindau, Lundh, McCanny, Mora, Neely, Wahlström, Zepf and McKenna2007) has opened up new areas of research, with applications in radiography (Mackinnon et al., Reference Mackinnon, Patel, Borghesi, Clarke, Freeman, Habara, Hatchett, Hey, Hicks, Kar, Key, King, Lancaster, Neely, Nikkro, Norreys, Notley, Phillips, Romagnani, Snavely, Stephens and Town2006), oncology (Bulanov & Khoroshkov, Reference Bulanov and Khoroshkov2002), astrophysics (Baraffe, Reference Baraffe2005), imaging (Fritzler et al., Reference Fritzler, Malka, Grillon, Rousseau, Burgy, Lefebvre, D'Humieres, McKenna and Ledingham2003), high-energy-density physics (Dyer et al., Reference Dyer, Bernstein, Cho, Osterholz, Grigsby, Dalton, Shepherd, Ping, Chen, Widmann and Ditmire2008), and ion-proton beam fast ignition (Roth et al., Reference Roth, Cowan, Key, Hatchett, Brown, Fountain, Johnson, Pennington, Snavely, Wilks, Yasuike, Ruhl, Pegoraro, Bulanov, Campbell, Perry and Powell2001; Key et al., Reference Key, Akli, Beg, Chen, Chung, Freeman, Foord, Green, Gu, Gregori, Habara, Hatchett, Hey, Hill, King, Kodama, Koch, Lancaster, Lasinski, Langdon, MacKinnon, Murphy, Norreys, Patel, Patel, Pasley, Snavely, Stephens, Stoeckl, Tabak, Theobald, Tanaka, Town, Wilks, Yabuuchi and Zhang2006a, Reference Key, Freeman, Hatchett, MacKinnon, Patel, Snavely and Stephensb). The fast recent progress has hinted that the extreme parameters of extreme light infrastructure will allow the production of ultra-high-energy ions (GeV and beyond) which will open the door to future unique applications like time and space resolved radiography of dense matter (Borghesi et al., Reference Borghesi, Sarri, Cecchetti, Kourakis, Hoarty, Stevenson, James, Brown, Hobbs, Lockyear, Morton, Willi, Jung and Dieckmann2010), injectors study for medical applications (Muramatsu and Kitagawa, Reference Muramatsu and Kitagawa2012) for ion beam physics (Hoffmann et al., Reference Hoffmann, Blazevic, Ni, Rosmej, Roth, Tahir, Tauschwitz, Udrea, Varentsov, Weyricj and Maron2005).

Recent development (Nickles et al., Reference Nickles, Ter-Avetisyan, Schnurer, Sokollik, Sandner, Schreiber, Hilscher, Jahnke, Andreev and Tikhonchuk2007; Daido et al., Reference Daido, Nishiuchi and Pirozhkov2012; Jung et al., Reference Jung, Albright, Yin, Gautier, Dromey, Shah, Palaniyappan, Letzring, Wu, Shimada, Johnson, Habs, Roth, Fernandez and Heglich2015) in laser driven ion acceleration from overdense plasmas demonstrated stimulated mechanisms. Target-normal sheath acceleration (TNSA) (Wilks et al., Reference Wilks, Kruer, Tabak and Langdon1992; Cowan et al., Reference Cowan, Fuchs, Ruhl, Kemp, Audebert, Roth, Stephens, Barton, Blazevic, Brambrink, Cobble, Fernández, Gauthier, Geissel, Hegelich, Kaae, Karsch, Le Sage, Letzring, Manclossi, Meyroneinc, Newkirk, Pépin and Renard- LeGalloudec2004), radiation pressure acceleration (RPA) (Esirkepov et al., Reference Esirkepov, Borghesi, Bulanov, Mourou and Tajima2004; Macchi et al., Reference Macchi, Cattani, Liseykina and Cornolti2005; Robinson et al., Reference Robinson, Gibbon, Zepf, Kar, Evans and Bellei2009; Palmer et al., Reference Palmer, Dover, Pogorelsky, Babzien, Dudnikova, Ispiriyan, Polyanskiy, Schreiber, Shkolnikov, Yakimenko and Najmudin2011; Yao et al., Reference Yao, Li, Cao, Zheng, Huang, Xiao and Skoric2014), shock wave acceleration (Silva et al., Reference Silva, Marti, Davies, Fonseca, Ren, Tsung and Mori2004; Haberberger et al., Reference Haberberger, Tochitsky, Fiuza, Gong, Fonseca, Silva, Mori and Joshi2012; Fiuza et al., Reference Fiuza, Stockem, Boella, Fonseca, Silva, Haberberger, Tochitsky, Mori and Joshi2013) and relativistic transparency regime (Henig et al., Reference Henig, Kiefer, Markey, Gautier, Flippo, Letzring, Johnson, Shimada, Yin, Albright, Bowers, Fernández, Rykovanov, Wu, Zepf, Jung, Liechtenstein, Schreiber, Habs and Hegelich2009; Yin et al., Reference Yin, Albright, Bowers, Jung, Fernández and Hegelich2011; Jung et al., Reference Jung, Yin, Albright, Gautier, Letzring, Dromey, Yeung, Hörlein, Shah and Palaniyappan2013; Roth et al., Reference Roth, Jung, Falk, Guler, Deppert, Devlin, Favalli, Fernandez, Gautier, Geissel, Haight, Hamilton, Hegelich, Johnson, Merrill, Schaumann, Schoenberg, Schollmeier, Shimada, Taddeucci, Tybo, Wagner, Wender, Wilde and Wurden2013) are the novel mechanisms which are the center of experiments and theoretical investigations.

The high-energy ion acceleration from near critical density target is also of high relevance because recent analytical and simulation study reported high-energy protons. Matsukado et al. (Reference Matsukado, Esirkepov, Kinoshita, Daido, Utsumi, Li, Fukumi, Hayashi, Orimo, Nishiuchi, Bulanov, Tajima, Noda, Iwashita, Shirai, Takeuchi, Nakamura, Yamazaki, Ikegami, Mihara, Morita, Uesaka, Yoshii, Watanabe, Hosokai, Zhidkov, Ogata, Wada and Kubota2003) investigated the role of prepulse in acceleration of protons from thin foil. In their research it was shown that initially prepulse evaporate the irradiated region of thin foil, thereafter the main pulse interacting with the underdense plasma generates the electrostatic field due to magnetic field expansion and hence accelerate the ions. Bulanov et al. (Reference Bulanov, Bychenkov, Chvykov, Kalinchenko, Litzenberg, Matsuoka, Thomas, Willingale, Yanovsky, Krushelnick and Maksimchuk2010a) investigated the scaling laws and optimal conditions for proton acceleration by the magnetic vortex mechanism (MVA) in near critical density targets. Nakamura et al. (Reference Nakamura, Bulanov, Esirkepov and Kando2010) also derived the energy scaling of ions by MVA using the particle-in-cell (PIC) simulations.

Recently Gu et al. (Reference Gu, Zhu, Li, Yu, Huang, Zhang, Kong and Kawata2014) investigated the bunch of protons of maximum energy 1 GeV using the 2.5D PIC simulation where protons experience multi acceleration mechanism [at the rear surface of target the protons are accelerated by TNSA and MVA, and later on by the long range breakout afterburner acceleration (Yin et al., Reference Yin, Albright, Hegelich and Fernández2006, Reference Yin, Albright, Bowers, Jung, Fernández and Hegelich2011) during the whole acceleration process]. Kawata et al. (Reference Kawata, Nagashima, Takano, Izumiyama, Kamiyama, Barada, Kong, Gu, Wang, Ma, Wang, Zhang, Xie, Zhang and Dai2014) discussed the compact and controllable laser baser ion accelerator by using a solid target with a fine sub-wavelength structure or a near critical density gas plasma. The energy efficiency from laser to ion is improved by employing such hybrid target. The ion acceleration is also demonstrated using the cluster-gas target with an ultra-short laser pulse. The PIC simulation study by Fukuda et al. (Reference Fukuda, Faenov, Tampo, Pikuz, Nakamura, Kando, Hayashi, Yogo, Sakaki, Kameshima, Pirozhkov, Ogura, Mori, Esirkepov, Koga, Boldarev, Gasilov, Magunov, Yamauchi, Kodama, Bolton, Kato, Tajima, Daido and Bulanov2009) revealed the production of high-energy ions at the rear side of near-critical plasma target, due to the formation of dipole vortex structure at rear edge. Using an intense circularly polarized laser, Gao et al. (Reference Gao, Wang, Lin, Zou and Yan2012) demonstrated efficiently the acceleration of protons from a foam-Carbon foil target. Recently Gauthier et al. (Reference Gauthier, Levy, D'Humieres, Glesser, Albertazzi, Beaucourt, Brelli, Chen, Dervieux, Feugeas, Nicolai, Tikhonchuk, Pepin, Antici and Fuchs2014) have shown a promising approach to accelerate protons to high-energies using the near-critical target. The simulation study explores the relevance of density gradient of plasma target on acceleration process. Another interesting mechanism is reported by Bake et al. (Reference Bake, Zhang, Xie, Hong and Wang2012) where the proton acceleration is demonstrated via combined mechanism of RPA and plasma bubble field. By 2D PIC simulations the energy enhancement of protons bunch is shown by using the background plasma with negative density gradient. 3D PIC simulation results reported by Lemos et al. (Reference Lemos, Martins, Dias, Marsh, Pak and Joshi2012) were confirmed in an experiment conducted at University of California, Los Angeles (UCLA). In their experimental study few MeV energetic ions were observed in an underdense plasma by employing a 50 fs/5 TWs Ti:Sapphire laser.

The most stable and well understood mechanism so far is the TNSA, which usually requires long pulse duration in order to reach high cut-off energy. Since at a few laboratories the ultra-short laser pulses of high repetition rate are in the center of interest, we have to consider different mechanisms, especially to avoid debris problem. The scheme of interest may be MVA, which is more efficient in near-critical density plasma.

We investigated in this 3D PIC simulation study the acceleration of protons from the near critical density hydrogen target using the ultra-short (17 fs) and high power [in petawatt (PW) regime] laser, which is not explored in past (Amitani et al., Reference Amitani, Esirkepov, Bulanov, Nishihara, Kuznetsov and Kamenets2002) through the 3D simulations. We further explored the spatial distribution of high-energy protons which are generated via a composite acceleration mechanism of MVA and post-acceleration by the rest-over laser field itself. The 3D simulations are performed in this research to suit the experimental conditions for ion acceleration application, to employ the high repetition PW laser facility of ELI-ALPS (http://www.eli-hu.hu/). 3D simulation is also of relevance for MVA mechanism of ion acceleration because the magnetic vortex structure is a 3D entity, where fast electrons propagate inside the channel and the return current flows along the channel wall, which accompanies the radial electric field and azimuthal magnetic field. Thus it is desired to model the 3D simulation to investigate the effective proton acceleration utilizing the MVA mechanism, which can be advantageous over 1.5D or 2D simulation. We reported the simulation results in this study which delineate the high-energy protons with ring shaped density distribution at the plasma–vacuum interface.

2. 3D PIC SIMULATION RESULTS

The regime of laser ion acceleration from near critical density gas targets can be realized when the laser pulse propagates through a near critical density target that is much longer than the pulse itself and the pulse forms a density channel. In MVA regime, when a tightly focused laser pulse interacts with the near-critical density plasma, the ponderomotive force of the laser expels electrons and ions in the transverse direction, forming the electron and ion density channel. A portion of the electrons are accelerated in the direction of laser pulse propagation by the longitudinal electric field. The motion of these electrons generates a magnetic field, which circulates in the channel around the propagation axis. Upon exiting the channel, the magnetic field expands into vacuum and the electron current is dissipated. Some of the electrons leave the plasma channel while few of them return to sustain the magnetic field on the rear side of plasma–vacuum interface. The magnetic field displaces the electron component of plasma with regard to the ion component and a strong quasi-static electric field is generated that can both accelerate and collimate the ions. For optimum acceleration [Bulanov et al. (Reference Bulanov, Bychenkov, Chvykov, Kalinchenko, Litzenberg, Matsuoka, Thomas, Willingale, Yanovsky, Krushelnick and Maksimchuk2010a, Reference Bulanov, Litzenberg, Pirozhkov, Thomas, Willingale, Krushelnick and Maksimchukb)] in case of MVA, the laser spot size should match the size of the self-focusing channel in order to avoid filamentation. The laser pulse energy should be depleted near the target rear to create the dipole vortex at the exit side of the channel and it is not wasted to transmission.

The effectiveness of the MVA mechanism requires the efficient transfer of laser energy to the fast electrons in the plasma which are accelerated in the plasma channel along the laser propagation direction. Thus the optimal condition for efficient proton acceleration can be estimated by equating laser pulse energy in plasma waveguide to the energy of electrons. The optimum plasma length (l) can be estimated from the assumption that all laser energy is transferred to the electrons in the plasma channel. If each electron has an average energy of a Lm ec 2, the optimum length can be written as (Bulanov et al., Reference Bulanov, Litzenberg, Pirozhkov, Thomas, Willingale, Krushelnick and Maksimchuk2010b) l = a Lcτ(n c/n e) K, where K is the geometry constant (K is 0.1 in 2D case and 0.074 in 3D case) and the laser pulse amplitude (a L), can be determined by the laser power (P) and plasma density (n e) as a L = (8Π (P 0/P c)(n e/n c))1/3.

Based on the theoretical approach (as outlined previously), we demonstrate here through the 3D PIC simulations the acceleration of protons by employing the ELI-ALPS high-field (HF) laser interaction with the near critical density hydrogen plasma. The incident laser parameters are chosen comfortably within the capabilities of ELI-ALPS facility. The expected parameter regime for ELI-ALPS HF laser is as follows: Laser wavelength λL = 800 nm, laser energy εL = 34J, and pulse duration t L = 17 fs. We consider here the interaction of a linearly polarized laser pulse with the plasma target. The laser pulse considered here is Gaussian in space and time where the beam radius is r b = 1.042 µm obtained by focusing lens of F = 2. The tightly focused beam will initially diverge in the plasma and will expel electrons to form a plasma channel, then after the divergence of laser will stop and the most of the laser energy in plasma channel will be transferred to plasma electrons to accelerate it. The plasma target proposed here is hydrogen of density n i = n e = 5.22 × 1021 cm−3 (three times of the critical density n c) and the optimum thickness of plasma target is l = 25 µm, to utilize the maximum laser energy transfer to plasma electrons [see (1)]. 3D PIC simulations were carried out using the fully relativistic electromagnetic code PIConGPU (Burau, Reference Burau2010; Bussmann et al., Reference Bussmann, Burau, Cowan, Debus, Huebl, Juckeland, Kluge, Nagel, Pausch, Schmitt, Schramm, Schuchart and Widera2013). The boundary conditions are periodic throughout the simulation. We considered here the simulation box of dimension 10 × 200 × 20 µm3 corresponding to the grid size 256 × 5120 × 512 with cell size of 40 nm. The time step is 66.7 as. The laser pulse (polarization of laser field is along x-axis) incidents on plasma along the y-axis in XZ plane. The plasma target starts at z = 0.0 µm and terminates at 25.0 µm. The 3D simulation performed in this study limits to the simulation size corresponding to the available graphical processing units (GPUs) computing facility. We considered here n i = 3n c to minimize the optimal plasma channel length and limit 3D simulations. In our simulations we considered intensities of the order of 1023 W/cm2 or smaller so we did not incorporate the radiation reaction force (see Gao et al., Reference Gao, Wang, Lin, Zou and Yan2012) along with the Lorentz force for proton acceleration.

A liquid hydrogen jet [Kühnel et al. (Reference Kühnel, Fernández, Tejeda, Kalinin, Montero and Grisenti2011)] might be an interesting option for MVA to obtain the near critical density plasma in laboratory with ultra-intense femtosecond pulses because, besides allowing the acceleration of protons, being a “continuously flowing” target it would allow for high repetition rate operation, which is of high relevance for applications with high repetition PW laser.

The PIC simulation results (as shown by Fig. 1), summarizes the laser pulse penetration through the near critical density hydrogen plasma and consequently the generation of magnetic field at the plasma–vacuum interface.

Fig. 1. (a) The distribution of the magnetic field of vortex structure followed by the laser magnetic field, as the laser pulse channels inside the hydrogen plasma, (b) the variation of magnetic field at plasma–vacuum interface in XZ plane, (c) scaling of magnetic field with the incident laser power, analytical result (black solid line), and simulation results (red dot); at time t = 0.2 psec.

The generation of magnetic field can be illustrated in the following manner. The relativistic electrons accelerated within the plasma channel, follow the laser pulse and exit from the rear side of the plasma channel as the pulse exist at the plasma–vacuum interface. The fast electrons leave the plasma and then return back into it under the influence of an unneutralized electric charge. The electrons form a toroidal vortex in 3D geometry (dipole vortex in 2D case) due to such motion and the electric current of vortex generates a quasi-static magnetic field. Thus in case of homogeneous plasma with sharp plasma vacuum boundary at both ends (front and rear) of plasma channel, the magnetic field produced by fast electron beam expands along the plasma slab boundary. The rapid variation in magnetic field due to the vortex motion produces a strong quasi-static electric field. The ions are accelerated to high-energies due to the quasi-static electric field at plasma–vacuum interface. The magnitude of the magnetic field can be estimated by employing the Ampere's law, B ≈ 4Πηn eer ch where r ch is the plasma channel radius which is equal to the radius of focused laser beam. We further simplified the magnetic field to scale it with initial laser-plasma parameter and obtained as,

(1)$$B = 32 \times {10^{{\rm -} 8}}{\rm \eta} \sqrt {{a_{\rm L}}{n_{\rm e}}\left\lceil {{\rm c}{{\rm m}^{{\rm -} 3}}} \right\rceil} T,$$

(in terms of Tesla). The magnitude of magnetic field (using η = 0.6 taken from the simulation) is ~20.0 × 104T, which is close to the simulation results as shown by Figure 1c. Thus the magnetic field (1) increases with the focused laser field and the plasma density, and consequently the strong magnetic field can be generated in laboratory by utilizing the high power laser interaction with the near-critical plasma. The evolution of the magnetic field depends dominantly on the scale on which the plasma density varies. In case of an inhomogeneous plasma with a density gradient at plasma–vacuum interface, the magnetic vortex moves down the density gradient and expands in forward direction due to the decrease in plasma density and in lateral direction due to force acting on the vortex in the $\nabla nX\Omega $ direction (Nycander and Isichenko, Reference Nycander and Isichenko1990). Due to the forward and lateral expansion of magnetic vortex structure the electron density in the current filament decreases to a value where the condition of charge quasi-neutrality fails to hold. Consequently the current filament undergoes a Coulomb expansion and leads the vanishing of magnetic field. The abrupt decrease in magnetic field induces a strong electric field which accelerates the ion beam to high-energies. PIC simulation studies (Bulanov et al., Reference Bulanov, Dylov, Esirkepov, Kamenets and Sokolov2005; Nakamura et al., Reference Nakamura, Bulanov, Esirkepov and Kando2010) of high power laser interaction with gas target have shown the relevance of plasma density gradient for ion acceleration, however in this study we focus the ion acceleration utilizing the ultra-short–ultra-intense laser pulse interacting the near critical density plasma with sharp plasma–vacuum boundary.

We show the evolution of electron and ion density in Figure 2 at 200 fs, when the laser pulse exits from the plasma channel. A strong wakefield of the order of several GV/cm is generated in the plasma, as the ultra-intense laser propagates in the near critical density plasma. The bunch of trapped electrons is then accelerated to 100 s of MeV along the channel axis. As fast electron beam ejects out of the plasma into vacuum, the most energetic electrons escapes, forming a plasma potential barrier, which prevents further acceleration of low energy electrons.

Fig. 2. Evolution of (a) electron density (n e) and (b) ion density (n i) at time 200 fs, after pulse exits the plasma channel. The electron and ion densities are normalized to relativistic modified critical plasma density ${n_{\rm cr}} = \sqrt {{\rm \gamma}}{n_{\rm e}} $ where γ is the relativistic factor (here γ ≈ 12) as its variation is shown by the color bar. The X, Y and Z-axis are shown in units of laser wavelength.

As these electrons are pulled back into plasma (return current), it generates a magnetic vortex field of magnitude several megagauss (as shown in Figure 3a). The expansion of the magnetic field produces a longitudinal inductive electric field (by Faraday's Law), which accelerates the protons further. The longitudinal electric field (as shown in Fig. 3b) attains the magnitude of the order of TV/m.

Fig. 3. Evolution of (a) magnetic vortex field (Bz) and (b) longitudinal electric field (Ey) at (200 fs) just after pulse exits the plasma channel. The color bar shown for Bz is expressed in units of 8.52 × 104 T and for Ey in units of 25.5 TV/m (E L~500 TV/m). The Y and Z-axis are shown in units of laser wavelength. The dark red and blue lines ahead of Bz and Ey correspond to the laser field.

The ion filament maintains over long distance because of dominant nature of focusing longitudinal field over the diverging magnetic vortex field. This electric field accelerates and collimates protons from the thin proton filament which is formed along the propagation channel. Figure 4 shows the longitudinal and transverse proton momentum along the propagation direction (Y-axis) at different time instant to demonstrate the acceleration of protons at the rear side of plasma target in vacuum.

Fig. 4. Longitudinal momentum (p y) (a,b) and transverse momentum (p z) (c,d) of accelerated protons along the propagation direction (Y-axis) at time instant (a,c) 0.2 psec (b,d) 0.53 psec. The Y-axis is shown in units of laser wavelength (0.8 µm) and longitudinal and transverse proton momentum is expressed in units of m ec.

Figure 4a, c shows the acceleration of protons at time instant 0.2 psec where protons are accelerated to high energy from longitudinal electric field generated due to the expansion of magnetic field. Figure 4c, d further explain the extended acceleration of protons at time instant 0.53 psec, since as time advances the magnetic field dissipates and the acceleration of proton is due to the energy transfer from leading electrons through the rest-over laser field. The electrons and protons accelerated along the Y-axis are expelled by the laser field in transverse direction and the transverse expansion is favored because of coulomb explosion of particles at axis. From the phase space data (as shown in Fig. 4) one can obtain the proton angular distributions. For a specific energy range, the distribution can be calculated by calculating the number of particles going in a given angle found from

(2)$${\rm \theta} = {\cos ^{{\rm -} 1}}\left( {{\,p_y}/\sqrt {{\,p_x^2} + p_y^2 + p_z^2}}\, \right)$$

The energy spectrum of the accelerated proton bunch which is propagating close to the axis, shown in Figure 5 at time t = 0.2 psec (a) and 0.67 psec (b). The maximum cut-off energy, we obtain in this case is around 1.1 GeV at 0.67 psec. In the inset we show the energy spectrum of all protons, which are propagating along laser direction at different angle. The longitudinal momentum of protons is maintained from t = 0.2 ps to 0.67 psec since the responsible factor for their acceleration is the longitudinal electric field.

Fig. 5. Energy distribution of protons propagating close (−5° to + 5°) to axis at 0.2 psec (a) and 0.67 psec (b). Inset plot: Proton energy distribution while considering the all accelerated protons.

We studied further the proton density (energy) at different time instant during the acceleration process (as shown in Fig. 6), after the laser field exits from the plasma channel. At time 167 fs when the ions form thin filament, they are accelerated to energy of 0.16 GeV and at time 200 fs the ions gain energy 0.35 GeV due to the strong longitudinal electric field produced due to time varying magnetic field. The ring shaped proton distribution (as shown in Fig. 6b) can be explained as an expansion of proton filament, which is formed due to ponderomotive expulsion of electrons from the region which is having a dimension of the order of the laser spot diameter. In Figure 6c we show the angular distribution of protons, which indicates that, the high energy protons are well collimated with the small divergence angle. It is worth to note that the results shown in Figure 5 correspond to the protons propagating close to the ring and the results shown in Fig. 5a is corresponding to Fig. 6b at 0.2 psec.

Fig. 6. The proton density distribution corresponding to different maximum proton energy (at different time instant t = (a) 0.17 psec, (b) 0.2 psec corresponding to propagation direction at y = (a) 26 µm and (b) 48 µm. (c) The angular distribution of protons at 0.2 psec, where inset shows 3D view of proton energy distribution. The color bar corresponds to ion density normalized with critical plasma density. In (a) the pink background corresponds to unperturbed proton density since plot (a) corresponds to plasma–vacuum interface at the rear side of the plasma channel.

The generation of magnetic field at plasma–vacuum interface is shown in Figure 3a and subsequently the quasi-static electric field accelerating protons in Figure 3b. However the simulation results (see Fig. 3a, b) also show the electric and magnetic field of transmitted laser field, which is not completely depleted in the plasma channel. Thus the protons accelerated by the MVA mechanism at plasma–vacuum interface may also get influenced by the transmitted laser fields. As the pulse exits the plasma channel (as shown in Fig. 2a), the leading hot electrons propagate in vacuum and form a bow in front of the protons. The magnetic field upon exiting the channel pushes further protons in respect of electrons and collimated proton bunch follows the leading electrons (co-moving with the laser field). Thus leading electrons transfer their energy to proton through the charge separation field and electrons are continuously getting energy from the laser pulse, which extends the acceleration length to obtain the high proton energy. We explore further the mechanism of acceleration of proton by investigating the evolution of magnetic field (Fig. 7a) and longitudinal electric field (Fig. 7b) which is responsible for the acceleration of electrons and protons. At initial acceleration stage the ions gain energy from the longitudinal electric field which is produced by the expansion of magnetic field. The magnetic field is maintained for 106 fs after the laser pulse exits the plasma channel (as shown in Fig. 7a, since the magnetic field of vortex structure exits along the Y-axis for distance 32 µm starting from 30 to 70 laser wavelength, which corresponds to 106 fs time period). In this time period the protons gain energy 670 MeV. After this time the acceleration continues at slower rate due to the expansion of magnetic field. The post acceleration mechanism which is responsible after this time is due to longitudinal electric field between electrons, which is co-moving with laser field (as shown in Fig. 7c, d) and the accelerated protons. The proton energy density (Fig. 7c) and electron energy density (Fig. 7d) at time 0.53 psec is shown to reveal the post acceleration process along the propagation axis. Thus as time advances the magnetic field dissipates and the dominant acceleration of proton is due to the energy transfer from leading electrons.

Fig. 7. The evolution of (a) magnetic vortex field (Bz) (b) longitudinal electric field (Ey), (c) electron energy density, and (d) ion energy density; along the propagation direction (Y-axis) followed by the laser field at 530 fsec (just before where we get the maximum proton energy). The Y-axis are shown in units of laser wavelength. Bz and for Ey are expressed in the same units as in Figure 3 while the electron and ion energy densities are expressed in units of n cm ec 2.

Following the model of Bulanov et al. (Reference Bulanov, Esarey, Schroeder, Leemans, Bulanov, Margarone, Korn and Haberer2015) the maximum proton energy from MVA can be written as ${E_{\rm P}} = {\rm \eta \gamma} _{\rm e}^2 {a_{\rm L}}{m_{\rm e}}{c^2}$ where the proton energy (E P) scales with the laser power as P 0.67 and ${{\rm \gamma} _{\rm e}} = \left( {\sqrt {2/1.84}} \right){(2P/K{P_{\rm c}})^{1/6}}{({n_{{\rm cr}}}/{n_{\rm e}})^{1/3}}$ is the Lorentz factor of accelerated bulk electrons (which can be obtained from the condition that electron velocity is equal to the group velocity of an electromagnetic pulse propagating in a waveguide).

In Figure 8, we show the dependence of maximum proton energy on the laser power for the optimized value of laser-target parameter. We obtained the scaling of proton energy with the laser power as E P ~ P 0.67, which is consistent with the model result.

Fig. 8. The dependence of maximal proton energy on the laser power. The black line curve shows the maximum proton energy from simulation results, while the blue dot corresponds to the laser power where the simulation is performed.

The proton energy scaling demonstrates that few hundred MeV–GeV protons can be obtained by employing the high power laser (from 100 TW–2 PW) interaction with the hydrogen near critical density plasma. The 3D simulation results in the considered regime show the laser-to-proton conversion efficiency ~1% (for protons with energy >100 MeV and cut-off energy ~1 GeV). The highest TNSA proton energy (~70 MeV) has been obtained by using single shot 80 J laser pulse and specially designed targets, where the conversion efficiency (for proton >4 MeV energy) was also about 1% [Gaillard et al. (Reference Gaillard, Kluge, Flippo, Bussmann, Gall, Lockard, Geissel, Offermann, Schollmeier, Sentoku and Cowan2011)]. In our case even PW range laser has repetition rate about 10 Hz, it means that our laser will have at least ten times higher efficiency in the production of fast protons.

3. CONCLUSION

In conclusion, we demonstrated the acceleration of ions from near-critical density hydrogen plasma employing the ultra-short laser pulse by 3D PIC simulation. We obtained the maximum proton energy ~GeV by employing the 2 PW/17 fs laser (http://www.eli-hu.hu/). The simulation results showed that an ion filament is formed on the axis of the plasma due to space charge attraction of high charge of the accelerated electron bunch. Significant part of the energetic ions forms a ring like spatial distribution. These protons are initially accelerated to sub GeV energy by a longitudinal electric field created by the time-varying magnetic field due to the return current of the lower energy electrons. We explored further that the MVA is effective for about 100 fs (in case of our parameters) after the laser pulse exits from plasma–vacuum interface, further acceleration of protons occurred from the electrons co-moving with the laser pulse. The final proton energy was thus determined by MVA and the post acceleration. The magnetic field generated at the rear side of plasma–vacuum interface increases with the incident laser power for a given plasma density. We scaled the dependence of magnetic field on initial laser-plasma parameter. The analytical formulation shows the dependence of magnetic field on the laser power and the plasma density. 3D PIC simulation results show the ring shaped proton density distribution where the protons are accelerated along the laser axis with fairly low divergence accompanied by off-axis beam of ring-like shape. The target consideration of near critical gas plasma can be a better alternative to a thin solid foil because of its low cost, debris -free and can be operated at high repetition rate (>10 Hz) for efficient potential applications.

ACKNOWLEDGEMENTS

We performed 3D PIC simulation utilising the code PIConGPU [Bussmann et al. (Reference Cowan, Fuchs, Ruhl, Kemp, Audebert, Roth, Stephens, Barton, Blazevic, Brambrink, Cobble, Fernández, Gauthier, Geissel, Hegelich, Kaae, Karsch, Le Sage, Letzring, Manclossi, Meyroneinc, Newkirk, Pépin and Renard- LeGalloudec2013)] in this research work. We acknowledge support of the Department of Information Services and Computing, Helmholtz-Zentrum Dresden-Rossendorf (HZDR), Germany; for providing access to the GPU Compute Cluster Hypnos. The authors also thank M. Bussmann, A. Huebl and the PIConGPU developer team for fruitful discussions regarding the simulation work; PIConGPU is developed and maintained by the Computational Radiation Physics Group at the Institute of Radiation Physics, HZDR.

References

REFERENCES

Amitani, H., Esirkepov, T., Bulanov, S., Nishihara, K., Kuznetsov, A. & Kamenets, F. (2002). Accelerated dense ion filament formed by ultra-intense laser in plasma slab. AIP Conf. Proc. 611, 340345.CrossRefGoogle Scholar
Bake, M.A., Zhang, S., Xie, B.S., Hong, X.R. & Wang, H.Y. (2012). Energy enhancement of proton acceleration in combinational radiation pressure and bubble by optimizing plasma density. Phys. Plasmas 19, 083103083109.Google Scholar
Baraffe, I. (2005). Structure and evolution of giant planets. Space Sci. Rev. 116, 6776.Google Scholar
Borghesi, M., Mackinnon, A.J., Campbell, D.H., Hicks, D.G., Kar, S., Patel, P.K., Price, D., Romagnani, L., Schiavi, A. & Willi, O. (2004). Multi-MeV proton source investigations in ultraintense laser-foil interactions. Phys. Rev. Lett. 92, 055003055006.CrossRefGoogle ScholarPubMed
Borghesi, M., Sarri, G., Cecchetti, C.A., Kourakis, I., Hoarty, D., Stevenson, R.M., James, S., Brown, C.D., Hobbs, P., Lockyear, J., Morton, J., Willi, O., Jung, R. & Dieckmann, M. (2010). Progress in proton radiography for diagnosis of ICF-relevant plasmas. Laser Part. Beams 28, 277284.Google Scholar
Bulanov, S.V. & Khoroshkov, V.S. (2002). Feasibility of using laser ion accelerators in proton therapy. Plasma Phys. Rep. 28, 453456.Google Scholar
Bulanov, S.S., Bychenkov, V.Y., Chvykov, V., Kalinchenko, G., Litzenberg, D.W., Matsuoka, T., Thomas, A.G.R., Willingale, L., Yanovsky, V., Krushelnick, K. & Maksimchuk, A. (2010 a). Generation of GeV protons from 1 PW laser interaction with near critical density targets. Phys. Plasmas 17, 043105043112.CrossRefGoogle ScholarPubMed
Bulanov, S.S., Esarey, E., Schroeder, C.B., Leemans, W.P., Bulanov, S.V., Margarone, D., Korn, G. & Haberer, T. (2015). Helium-3 and helium-4 acceleration by high power laser pulses for hadron therapy. Phys. Rev. ST Accel. Beams 18, 061302061307.CrossRefGoogle Scholar
Bulanov, S.S., Litzenberg, D.W., Pirozhkov, A.S., Thomas, A.G.R., Willingale, L., Krushelnick, K. & Maksimchuk, A. (2010 b). Laser acceleration of protons from near critical density targets for application to radiation therapy, http://arxiv.org/abs/1007.3971Google Scholar
Bulanov, S.V., Dylov, D.V., Esirkepov, T.Zh., Kamenets, F.F. & Sokolov, D.V. (2005). Ion acceleration in a dipole vortex in a laser plasma corona. Plasma Phys. Rep. 31, 369381.CrossRefGoogle Scholar
Bussmann, M., Burau, H., Cowan, T.E., Debus, A., Huebl, A., Juckeland, G., Kluge, T., Nagel, W.E., Pausch, R., Schmitt, F., Schramm, U., Schuchart, J. & Widera, R. (2013). Radiative Signatures of the Relativistic Kelvin-Helmholtz Instability, Proc. of the Int. Conf. on High Performance Comp., Networking, Storage and Anal. 5:1–5:12.Google Scholar
Burau, H. (2010). PIConGPU: A fully relativistic particle-in-cell code for a GPU cluster. IEEE Trans. Plasma Sci. 38, 28312839.CrossRefGoogle Scholar
Cowan, T.E., Fuchs, J., Ruhl, H., Kemp, A., Audebert, P., Roth, M., Stephens, R., Barton, I., Blazevic, A., Brambrink, E., Cobble, J., Fernández, J., Gauthier, J.C., Geissel, M., Hegelich, M., Kaae, J., Karsch, S., Le Sage, G.P., Letzring, S., Manclossi, M., Meyroneinc, S., Newkirk, A., Pépin, H. & Renard- LeGalloudec, N. (2004). Ultralow emittance, multi-MeV proton beams from a laser virtual-cathode plasma accelerator. Phys. Rev. Lett. 92, 204801204804.CrossRefGoogle ScholarPubMed
Daido, H., Nishiuchi, M. & Pirozhkov, A.S. (2012). Review of laser-driven ion sources and their applications. Rep. Prog. Phys. 75, 056401056471.Google Scholar
Dyer, G.M., Bernstein, A.C., Cho, B.I., Osterholz, J., Grigsby, W., Dalton, A., Shepherd, R., Ping, Y., Chen, H., Widmann, K. & Ditmire, T. (2008). Equation-of-state measurement of dense plasmas heated with fast protons. Phys. Rev. Lett. 101, 1500215005.CrossRefGoogle ScholarPubMed
Esirkepov, T., Borghesi, M., Bulanov, S.V., Mourou, G. & Tajima, T. (2004). Highly efficient relativistic-ion generation in the laser-piston regime. Phys. Rev. Lett. 92, 175003175006.Google Scholar
Fiuza, F., Stockem, A., Boella, E., Fonseca, R.A., Silva, L.O., Haberberger, D., Tochitsky, S., Mori, W.B. & Joshi, C. (2013). Ion acceleration from laser-driven electrostatic shocks. Phys. Plasmas 20, 056304056311.Google Scholar
Fritzler, S., Malka, V., Grillon, G., Rousseau, J.P., and Burgy, F., Lefebvre, E., D'Humieres, E., McKenna, P. & Ledingham, K.W.D. (2003). Proton beams generated with high-intensity lasers: Applications to medical isotope production. Appl. Phys. Lett. 83, 30393041.Google Scholar
Fuchs, J., Antici, P., d'Humières, E., Lefebvre, E., Borghesi, M., Brambrink, E., Cecchetti, C.A., Kaluza, M., Malka, V., Manclossi, M., Meyroneinc, S., Mora, P., Schreiber, J., Toncian, T., Pépin, H. & Audebert, P. (2006). Laser-driven proton scaling laws and new paths towards energy increase. Nature Phys. 2, 4854.Google Scholar
Fukuda, Y., Faenov, A.Ya., Tampo, M., Pikuz, T.A., Nakamura, T., Kando, M., Hayashi, Y., Yogo, A., Sakaki, H., Kameshima, T., Pirozhkov, A.S., Ogura, K., Mori, M., Esirkepov, T.Z., Koga, J., Boldarev, A.S., Gasilov, V.A., Magunov, A.I., Yamauchi, T., Kodama, R., Bolton, P.R., Kato, Y., Tajima, T., Daido, H. & Bulanov, S.V. (2009). Energy increase in multi-MeV ion acceleration in the interaction of a short pulse laser with a cluster-gas target. Phys. Rev. Lett. 103, 165002165005.CrossRefGoogle Scholar
Gaillard, S.A., Kluge, T., Flippo, K.A., Bussmann, M., Gall, B., Lockard, T., Geissel, M., Offermann, D.T., Schollmeier, M., Sentoku, Y. & Cowan, T.E. (2011). Increased laser-accelerated proton energies via direct laser-light-pressure acceleration of electrons in microcone targets. Phys. Plasmas 18, 056710056720.Google Scholar
Gao, L., Wang, H., Lin, C., Zou, Y. & Yan, X. (2012). Efficient proton beam generation from a foam-carbon foil target using an intense circularly polarized laser. Phys. Plasmas 19, 083107083110.Google Scholar
Gauthier, M., Levy, A., D'Humieres, E., Glesser, M., Albertazzi, B., Beaucourt, C., Brelli, J., Chen, S.N., Dervieux, V., Feugeas, J.L., Nicolai, P., Tikhonchuk, V., Pepin, H., Antici, P. & Fuchs, J. (2014). Investigation of longitudinal proton acceleration in exploded targets irradiated by intense short-pulse laser. Phys. Plasmas 21, 013102013112.Google Scholar
Gu, Y.J., Zhu, Z., Li, X.F., Yu, Q., Huang, S., Zhang, F., Kong, Q. & Kawata, S. (2014). Stable long range proton acceleration driven by intense laser pulse with underdense plasmas. Phys. Plasmas 21, 063104063109.Google Scholar
Haberberger, D., Tochitsky, S., Fiuza, F., Gong, C., Fonseca, R.A., Silva, L.O., Mori, W.B. & Joshi, C. (2012). Collisionless shocks in laser-produced plasma generate monoenergetic high-energy proton beams. Nat. Phys. 8, 9599.Google Scholar
Hatchett, S.P., Brown, C.G., Cowan, T.E., Henry, E.A., Johnson, J.S., Key, M.H., Koch, J.A., Langdon, A.B., Lasinski, B.F., Lee, R.W., Mackinnon, A.J., Pennington, D.M., Perry, M.D., Phillips, T.W., Roth, M., Sangster, T.C., Singh, M.C., Snavely, R.A., Stoyer, M.A., Wilks, S.C. & Yasuike, K. (2000). Electron, photon, and ion beams from the relativistic interaction of Petawatt laser pulses with solid targets. Phys. Plasmas 7, 20762082.CrossRefGoogle Scholar
Hegelich, B.M., Albright, B.J., Cobble, J., Flippo, K., Letzring, S., Paffett, M., Ruhl, H., Schreiber, J., Schulze, R.K. & Fernández, J.C. (2006). Laser acceleration of quasi-monoenergetic MeV ion beams. Nature 439, 441444.Google Scholar
Henig, A., Kiefer, D., Markey, K., Gautier, D.C., Flippo, K.A., Letzring, S., Johnson, R.P., Shimada, T., Yin, L., Albright, B.J., Bowers, K.J., Fernández, J.C., Rykovanov, S.G., Wu, H.C., Zepf, M., Jung, D., Liechtenstein, V.Kh., Schreiber, J., Habs, D. & Hegelich, B.M. (2009). Enhanced laser-driven ion acceleration in the relativistic transparency regime. Phys. Rev. Lett. 103, 045002045005.Google Scholar
Hoffmann, D.H.H., Blazevic, A., Ni, P., Rosmej, O., Roth, M., Tahir, N.A., Tauschwitz, A., Udrea, S., Varentsov, D., Weyricj, K., and Maron, Y. (2005). Present and future perspectives for high energy density physics with intense heavy ion and laser beams. Laser Part. Beams 23, 4753.CrossRefGoogle Scholar
Jung, D., Albright, B.J., Yin, L., Gautier, D.C., Dromey, B., Shah, R., Palaniyappan, S., Letzring, S., Wu, H.-C., Shimada, T., Johnson, R.P., Habs, D., Roth, M., Fernandez, J.C. & Heglich, M. (2015). Scaling of ion energies in the relativistic-induced transparency regime. Laser Part. Beams 33, 695703.Google Scholar
Jung, D., Yin, L., Albright, B.J., Gautier, D.C., Letzring, S., Dromey, B., Yeung, M., Hörlein, R., Shah, R. & Palaniyappan, S. (2013). Efficient carbon ion beam generation from laser-driven volume acceleration. New J. Phys. 15, 023007023016.Google Scholar
Kawata, S., Nagashima, T., Takano, M., Izumiyama, T., Kamiyama, D., Barada, D., Kong, Q., Gu, Y.J., Wang, P.X., Ma, Y.Y., Wang, W.M., Zhang, W., Xie, J., Zhang, H. & Dai, D. (2014). Controllability of intense-laser ion acceleration. High Power Laser Sc. Eng. 2, 112.Google Scholar
Key, M.H., Akli, K., Beg, F., Chen, M.H., Chung, H.K., Freeman, R.R., Foord, M.E., Green, J.S., Gu, P., Gregori, G., Habara, H., Hatchett, S.P., Hey, D., Hill, J.M., King, J.A., Kodama, R., Koch, J.A., Lancaster, K., Lasinski, B.F., Langdon, B., MacKinnon, A.J., Murphy, C.D., Norreys, P.A., Patel, N., Patel, P., Pasley, J., Snavely, R.A., Stephens, R.B., Stoeckl, C., Tabak, M., Theobald, W., Tanaka, K., Town, R., Wilks, S.C., Yabuuchi, Y.B. & Zhang, B. (2006 a). Study of electron and proton isochoric heating for fast ignition. J. Phys. IV France 133, 371378.Google Scholar
Key, M.H., Freeman, R.R., Hatchett, S.P., MacKinnon, A.J., Patel, P.K., Snavely, R.A. & Stephens, R.B. (2006 b). Proton fast ignition. Fusion Sci. Technol. 49, 440452.Google Scholar
Kühnel, M., Fernández, J.M., Tejeda, G., Kalinin, A., Montero, S. & Grisenti, R.E. (2011). Time-resolved study of crystallization in deeply cooled liquid parahydrogen. Phys. Rev. Lett. 106, 245301245304.CrossRefGoogle ScholarPubMed
Lemos, N., Martins, J.L., Dias, J.M., Marsh, K.A., Pak, A. & Joshi, C. (2012). Forward directed ion acceleration in a LWFA with ionization-induced injection. J. Plasma Phys. 78, 327331.Google Scholar
Loeffler, J.S. & Durante, M. (2013). Charged particle therapy – optimization, challenges and future directions. Nat. Rev. Clin. Oncol. 10, 411424.CrossRefGoogle ScholarPubMed
Macchi, A., Cattani, F., Liseykina, T.V. & Cornolti, F. (2005). Laser acceleration of ion bunches at the front surface of overdense plasmas. Phys. Rev. Lett. 94, 165003165006.Google Scholar
Mackinnon, A.J., Patel, P.K., Borghesi, M., Clarke, R.C., Freeman, R.R., Habara, H., Hatchett, S.P., Hey, D., Hicks, D.G., Kar, S., Key, M.H., King, J.A., Lancaster, K., Neely, D., Nikkro, A., Norreys, P.A., Notley, M.M., Phillips, T.W., Romagnani, L., Snavely, R.A., Stephens, R.B. & Town, R.P.J. (2006). Proton radiography of a laser-driven implosion. Phys. Rev. Lett. 97, 045001045004.Google Scholar
Matsukado, K., Esirkepov, T., Kinoshita, K., Daido, H., Utsumi, T., Li, Z., Fukumi, A., Hayashi, Y., Orimo, S., Nishiuchi, M., Bulanov, S.V., Tajima, T., Noda, A., Iwashita, Y., Shirai, T., Takeuchi, T., Nakamura, S., Yamazaki, A., Ikegami, M., Mihara, T., Morita, A., Uesaka, M., Yoshii, K., Watanabe, T., Hosokai, T., Zhidkov, A., Ogata, A., Wada, Y. & Kubota, T. (2003). Energetic protons from a few-micron metallic foil evaporated by an intense laser pulse. Phys. Rev. Lett. 91, 215001215004.CrossRefGoogle ScholarPubMed
Muramatsu, M. & Kitagawa, A. (2012). A review of ion sources for medical accelerators. Rev. Sci. Instrum. 83, 02B90902B915.Google Scholar
Nakamura, T., Bulanov, S.V., Esirkepov, T.Zh. & Kando, M. (2010). High-energy ions from near-critical density plasmas via magnetic vortex acceleration. Phys. Rev. Lett. 105, 135002135005.CrossRefGoogle ScholarPubMed
Nickles, P.V., Ter-Avetisyan, S., Schnurer, M., Sokollik, T., Sandner, W., Schreiber, J., Hilscher, D., Jahnke, U., Andreev, A. & Tikhonchuk, V. (2007). Review of ultrafast ion acceleration experiments in laser plasma at Max Born Institute. Laser Part. Beams 25, 347363.Google Scholar
Nycander, J. & Isichenko, M.B. (1990). Motion of dipole vortices in a weakly inhomogeneous medium and related convective transport. Phys. Fluids B2, 20422047.CrossRefGoogle Scholar
Palmer, C.A.J., Dover, N.P., Pogorelsky, I., Babzien, M., Dudnikova, G.I., Ispiriyan, M., Polyanskiy, M.N., Schreiber, J., Shkolnikov, P., Yakimenko, V. & Najmudin, Z. (2011). Monoenergetic proton beams accelerated by a radiation pressure driven shock. Phys. Rev. Lett. 106, 014801014804.Google Scholar
Robinson, A.P.L., Gibbon, P., Zepf, M., Kar, S., Evans, R.G. & Bellei, C. (2009). Relativistically correct hole-boring and ion acceleration by circularly polarized laser pulses. Plasma Phys. Contr. Fus. 51, 024004024017.CrossRefGoogle Scholar
Robson, L., Simpson, P.T., Clarke, R.J., Ledingham, K.W.D., Lindau, F., Lundh, O., McCanny, T., Mora, P., Neely, D., Wahlström, C.G., Zepf, M. & McKenna, P. (2007). Scaling of proton acceleration driven by petawatt-laser–plasma interactions. Nature Phys. 3, 5862.Google Scholar
Roth, M., Cowan, T.E., Key, M.H., Hatchett, S.P., Brown, C., Fountain, W., Johnson, J., Pennington, D.M., Snavely, R.A., Wilks, S.C., Yasuike, K., Ruhl, H., Pegoraro, F., Bulanov, S.V., Campbell, E.M., Perry, M.D. & Powell, H. (2001). Fast ignition by intense laser-accelerated proton beams. Phys. Rev. Lett. 86, 436439.Google Scholar
Roth, M., Jung, D., Falk, K., Guler, N., Deppert, O., Devlin, M., Favalli, A., Fernandez, J., Gautier, D., Geissel, M., Haight, R., Hamilton, C.E., Hegelich, B.M., Johnson, R.P., Merrill, F., Schaumann, G., Schoenberg, K., Schollmeier, M., Shimada, T., Taddeucci, T., Tybo, J.L., Wagner, F., Wender, S.A., Wilde, C.H. & Wurden, G.A. (2013). Bright laser-driven neutron source based on the relativistic transparency of solids. Phys. Rev. Lett. 110, 044802044806.Google Scholar
Silva, L.O., Marti, M., Davies, J.R., Fonseca, R.A., Ren, C., Tsung, F.S. & Mori, W.B. (2004). Proton shock acceleration in laser-plasma interactions. Phys. Rev. Lett. 9, 015002015005.Google Scholar
Snavely, R.A., Key, M.H., Hatchett, S.P., Cowan, T.E., Roth, M., Phillips, T.W., Stoyer, M.A., Henry, E.A., Sangster, T.C., Singh, M.S., Wilks, S.C., MacKinnon, A., Offenberger, A., Pennington, D.M., Yasuike, K., Langdon, A.B., Lasinski, B.F., Johnson, J., Perry, M.D. & Campbell, E.M. (2000). Intense high-energy proton beams from petawatt-laser irradiation of solids. Phys. Rev. Lett. 85, 29452948.CrossRefGoogle ScholarPubMed
Wilks, S.C., Kruer, W.L., Tabak, M. & Langdon, A.B. (1992). Absorption of ultra-intense laser pulses. Phys. Rev. Lett. 69, 13831386.Google Scholar
Wilks, S.C., Langdon, A.B., Cowan, T.E., Roth, M., Singh, M., Hatchett, S., Key, M.H., Pennington, D., MacKinnon, A. & Snavely, R.A. (2001). Energetic proton generation in ultra-intense laser–solid interactions. Phys. Plasmas 8, 542549.CrossRefGoogle Scholar
Yao, W., Li, B., Cao, L., Zheng, F., Huang, T., Xiao, H. & Skoric, M.M. (2014). Generation of monoenergetic proton beams by a combined scheme with an overdense hydrocarbon target and an underdense plasma gas irradiated by ultra-intense laser pulse. Laser Part. Beams 32, 583589.CrossRefGoogle Scholar
Yin, L., Albright, B.J., Bowers, K.J., Jung, D., Fernández, J.C. & Hegelich, B.M. (2011). Three-dimensional dynamics of breakout afterburner ion acceleration using high-contrast short-pulse laser and nanoscale targets. Phys. Rev. Lett. 107, 045003045006.CrossRefGoogle ScholarPubMed
Yin, L., Albright, B.J., Hegelich, B.M. & Fernández, J.C. (2006). GeV laser ion acceleration from ultrathin targets: The laser break-out afterburner. Laser Part. Beams 24, 291298.Google Scholar
Figure 0

Fig. 1. (a) The distribution of the magnetic field of vortex structure followed by the laser magnetic field, as the laser pulse channels inside the hydrogen plasma, (b) the variation of magnetic field at plasma–vacuum interface in XZ plane, (c) scaling of magnetic field with the incident laser power, analytical result (black solid line), and simulation results (red dot); at time t = 0.2 psec.

Figure 1

Fig. 2. Evolution of (a) electron density (ne) and (b) ion density (ni) at time 200 fs, after pulse exits the plasma channel. The electron and ion densities are normalized to relativistic modified critical plasma density ${n_{\rm cr}} = \sqrt {{\rm \gamma}}{n_{\rm e}} $ where γ is the relativistic factor (here γ ≈ 12) as its variation is shown by the color bar. The X, Y and Z-axis are shown in units of laser wavelength.

Figure 2

Fig. 3. Evolution of (a) magnetic vortex field (Bz) and (b) longitudinal electric field (Ey) at (200 fs) just after pulse exits the plasma channel. The color bar shown for Bz is expressed in units of 8.52 × 104 T and for Ey in units of 25.5 TV/m (EL~500 TV/m). The Y and Z-axis are shown in units of laser wavelength. The dark red and blue lines ahead of Bz and Ey correspond to the laser field.

Figure 3

Fig. 4. Longitudinal momentum (py) (a,b) and transverse momentum (pz) (c,d) of accelerated protons along the propagation direction (Y-axis) at time instant (a,c) 0.2 psec (b,d) 0.53 psec. The Y-axis is shown in units of laser wavelength (0.8 µm) and longitudinal and transverse proton momentum is expressed in units of mec.

Figure 4

Fig. 5. Energy distribution of protons propagating close (−5° to + 5°) to axis at 0.2 psec (a) and 0.67 psec (b). Inset plot: Proton energy distribution while considering the all accelerated protons.

Figure 5

Fig. 6. The proton density distribution corresponding to different maximum proton energy (at different time instant t = (a) 0.17 psec, (b) 0.2 psec corresponding to propagation direction at y = (a) 26 µm and (b) 48 µm. (c) The angular distribution of protons at 0.2 psec, where inset shows 3D view of proton energy distribution. The color bar corresponds to ion density normalized with critical plasma density. In (a) the pink background corresponds to unperturbed proton density since plot (a) corresponds to plasma–vacuum interface at the rear side of the plasma channel.

Figure 6

Fig. 7. The evolution of (a) magnetic vortex field (Bz) (b) longitudinal electric field (Ey), (c) electron energy density, and (d) ion energy density; along the propagation direction (Y-axis) followed by the laser field at 530 fsec (just before where we get the maximum proton energy). The Y-axis are shown in units of laser wavelength. Bz and for Ey are expressed in the same units as in Figure 3 while the electron and ion energy densities are expressed in units of ncmec2.

Figure 7

Fig. 8. The dependence of maximal proton energy on the laser power. The black line curve shows the maximum proton energy from simulation results, while the blue dot corresponds to the laser power where the simulation is performed.