Hostname: page-component-8448b6f56d-jr42d Total loading time: 0 Render date: 2024-04-23T06:23:37.479Z Has data issue: false hasContentIssue false

Observation of resistive Weibel instability in intense laser plasma

Published online by Cambridge University Press:  29 April 2020

S. Krishnamurthy
Affiliation:
Department of Physics, Indian Institute of Technology Hyderabad, Sangareddy, Telangana502285, India
K. Makur
Affiliation:
Department of Physics, Indian Institute of Technology Hyderabad, Sangareddy, Telangana502285, India
B. Ramakrishna*
Affiliation:
Department of Physics, Indian Institute of Technology Hyderabad, Sangareddy, Telangana502285, India
*
Author for correspondence: B. Ramakrishna, Department of Physics, Indian Institute of Technology Hyderabad, Sangareddy, Telangana502285, India. E-mail: bhuvan@phy.iith.ac.in
Rights & Permissions [Opens in a new window]

Abstract

We observe experimentally periodic proton beam filamentation in laser-produced dense plasma using multilayered (CH–Al–CH) sandwich targets. The accelerated MeV proton beams from these targets exhibit periodic frozen filaments up to 5–10 µm as a result of resistive Weibel instabilities in the expanding plasma. The evolution of strong self-generated resistive magnetic fields at the targets interface is attributed to such plasma effects, which are supported, by our theory and simulations. We suggest that the resistive Weibel instability could be effectively employed to understand the evolution of magnetic fields in laser-generated plasma in the astrophysics scenario or the advanced fast igniter approach of the inertial confinement fusion.

Type
Letter to the Editor
Copyright
Copyright © The Author(s), 2020. Published by Cambridge University Press

Introduction

Remarkable progress in the field of intense laser-matter interaction has been witnessed with the advent of a chirped pulse amplification (CPA) technique (Strickland and Mourou, Reference Strickland and Mourou1985). Laser-driven ion acceleration is immensely important as it has the potential to overcome the barrier faced by conventional accelerators. Amongst target normal sheath cceleration (TNSA), radiation pressure acceleration (RPA), collisionless shock acceleration (CSA), and other mechanisms of laser-driven ion acceleration, TNSA is the most widely accepted mechanism (Mora, Reference Mora2003), in which the laser pulse ionizes the front surface of the target and the electrons acquire energy from the peak of the pulse, these electrons are termed as fast electrons as they have higher kinetic energy. Fast electrons advance through the material and extend into the vacuum thereby creating a charge separation leading to a sheath electric field, which accelerates the ions in the vicinity. The accelerated ion energies are therefore dependent on the fast electrons. In addition to ion acceleration, the laser-generated fast electrons and their propagation in the target medium are important in various other applications, such as fast ignition in inertial confinement fusion (ICF) (Craxton et al., Reference Craxton, Anderson, Boehly, Goncharov, Harding, Knauer, McCrory, McKenty, Meyerhofer and Myatt2015; Betti and Hurricane, Reference Betti and Hurricane2016), generation of X rays (Gahn et al., Reference Gahn, Pretzler, Saemann, Tsakiris, Witte, Gassmann, Schätz, Schramm, Thirolf and Habs1998; Norreys et al., Reference Norreys, Santala, Clark, Zepf, Watts, Beg, Krushelnick, Tatarakis, Dangor and Fang1999; Yu et al., Reference Yu, Jiang, Kieffer and Krol1999; Ledingham et al., Reference Ledingham, Spencer, McCanny, Singhal, Santala, Clark, Watts, Beg, Zepf and Krushelnick2000), γ rays (Cowan et al., Reference Cowan, Perry, Key, Ditmire, Hatchett, Henry, Moody, Moran, Pennington and Phillips1999; Gahn et al., Reference Gahn, Tsakiris, Pretzler, Witte, Delfin, Wahlström and Habs2000; Chen et al., Reference Chen, Wilks, Bonlie, Liang, Myatt, Price, Meyerhofer and Beiersdorfer2009), and positrons (Clark et al., Reference Clark, Krushelnick, Davies, Zepf, Tatarakis, Beg, Machacek, Norreys, Santala, Watts and Dangor2000; Maksimchuk et al., Reference Maksimchuk, Gu, Flippo, Umstadter and Yu Bychenkov2000; Snavely et al., Reference Snavely, Key, Hatchett, Cowan, Roth, Phillips, Stoyer, Henry, Sangster and Singh2000). When fast electrons propagate in a dense solid target, the return cold electron currents are setup so that the requirement of charge neutrality is fulfilled in the plasma. Equilibrium between fast electrons and return electrons are unstable due to electromagnetic perturbation leading to Weibel instability, which produces strong filamentation of the fast electron beam. The efficiency of energy transport becomes very poor and intense fast electrons cannot transport beyond the filament length (Bell and Kingham, Reference Bell and Kingham2003). Instabilities have been modeled by connecting collisional and collisionless plasmas (Bret, Reference Bret2010) and also by considering quantum effects in the pre-compressed target (Bret et al., Reference Bret, Fernandez and Anfray2009) in order to understand the effects on the energy transfer from beam to pellet in the fast ignition scenario.

TNSA yields polyenergetic ions because the ions are accelerated by the sheath electric field, which is Gaussian in shape. The divergence of laser-accelerated protons (10–30°) is also set depending on the energy of the accelerated beam. Efforts have been made to gain control over the ion beam profile by employing a photoresist (PR) (Metzkes et al., Reference Metzkes, Kluge, Zeil, Bussmann, Kraft, Cowan and Schramm2014), hydrocarbon (CH) layer (Ramakrishna et al., Reference Ramakrishna, Tayyab, Bagchi, Mandal, Upadhyay, Weng, Murakami, Cowan, Chakera, Naik and Gupta2015) on the rear surface of the target in order to reduce the transverse modulations in the ion beam. In a multilayered target, the material boundary will cause a change in resistivity, which affects the magnetic field generation, and growth thereby affecting the flow of electrons as described in the theoretical work (Robinson and Sherlock, Reference Robinson and Sherlock2007) and a proof-of-principle experiment (Ramakrishna et al., Reference Ramakrishna, Kar, Robinson, Adams, Markey, Quinn, Yuan, McKenna, Lancaster, Green, Scott, Norreys, Schreiber and Zepf2010). Other methods like using a shaped target, quadrupole magnets, and compact ion lens driven by a secondary laser beam have been attempted to achieve geometrical focusing of the laser-driven protons (Toncian et al., Reference Toncian, Borghesi, Fuchs, d'Humieries, Antici, Audebert, Brambrink, Cecchetti, Pipahl, Romagnani and Willi2006; Kar et al., Reference Kar, Markey, Simpson, Bellei, Green, Nagel, Kneip, Carroll, Dromey, Willingale, Clark, McKenna, Najmudin, Krushelnick, Norreys, Clarke, Neely, Borghesi and Zepf2008, Reference Kar, Markey, Borghesi, Carroll, McKenna, Neely, Quinn and Zepf2011; Nishiuchi et al., Reference Nishiuchi, Daito, Ikegami, Daido, Mori, Orimo, Ogura, Sagisaka, Yogo, Pirozhkov, Sugiyama, Kiriyama, Okada, Kanazawa, Kondo, Shimomura, Tanoue, Nakai, Sasao, Wakai, Sakaki, Bolton, Choi, Sung, Lee, Oishi, Fujii, Nemoto, Souda, Noda, Iseki and Yoshiyuki2009). Since the typical laser to proton conversion efficiency previously reported from many experiments is less than 10% (Fuchs et al., Reference Fuchs, Antici, d'Humières, Lefebvre, Borghesi, Brambrink, Cecchetti, Kaluza, Malka, Manclossi, Meyroneinc, Mora, Schreiber, Toncian, Pépin and Audebert2006; Robson et al., Reference Robson, Simpson, McKenna, Ledingham and Clarke2006), it is crucial to maintain maximum proton flux for various applications, including hadron therapy and fast ignition. In the fast ignition approach, some proposals are currently being investigated to use proton beams in addition to the fast electrons to heat the core of the imploding capsule using CH layered targets. This design can reduce the core arrival time lags due to different ion energies, and ion beams with wide energy spectra can be accepted (Sakagami et al., Reference Sakagami, Johzaki, Sunahara and Nagatomo2016). However, the plasma instabilities affect the energy and beam profile of the ions and understanding such filamentary effects in proton beam is important that is otherwise proven to be detrimental for ICF.

Experimental details

The experiment was carried out at HZDR, Dresden, using the 150 TW Ti:sapphire laser system delivering 30 fs (FWHM) at 10 Hz repetition rate, at a central wavelength of 800 nm. The p-polarized laser beam was focused to a focal spot of 8 µm diameter (FWHM) with an off-axis parabolic mirror with a focal length of 250 mm, reaching a peak intensity of the order of ~3 × 1020 W/cm2 (Schramm et al., Reference Schramm, Bussmann, Irman, Siebold, Zeil, Albach, Bernert, Bock, Brack and Branco2017). The dispersed ion beam was detected by a micro-channel plate (MCP) detector coupled to a 16-bit camera. A stack of radiochromic films (RCFs) was placed at 5 cm behind the target to detect the ion beam profile. RCF was used because it is an absolutely calibrated dosimetry medium, which changes its color on exposure to ionizing radiation. Protons of a given energy will traverse only through a well-defined thickness of the film before depositing the energy at the Bragg peak (Xu et al., Reference Xu, Liao, Wu, Geng, Li, Zhu, Li, Hu, Shou, Chen, Lu, Ma, Zhao, Zhu, Lin and Yan2019). Therefore, each layer of RCF stack could be associated with a definite minimum proton energy (Fig. 1).

Fig. 1. (Color online): A schematic diagram of the experimental setup showing the RCF position behind the target.

Results and discussion

In this paper, we report the observation of proton beam filamentation at relativistic laser intensities. The filamentary structures observed in the proton beam are observed to uniformly evolve as the ultrafast (ps) plasma channel expands, which is clearly evidenced by the proton beam distribution recorded on each layer (6–10 MeV) of RCF stack as shown in Figure 2. The fast electron beam generated at the critical density surface of the plasma is observed to be divergent in the laser-plasma interaction (Santala et al., Reference Santala, Zepf, Watts, Beg, Clark, Tatarakis, Krushelnick, Dangor, McCanny, Spencer, Singhal, Ledingham, Wilks, Machacek, Wark, Allott, Clarke and Norreys2000). Very large instantaneous currents (106–109 A) produced will lead to strong collective effects (Bell et al., Reference Bell, Davies, Guerin and Ruhl1997; Davies et al., Reference Davies, Bell, Haines and Guerin1997), which results in fine structure (i.e., filaments) (Bell and Kingham, Reference Bell and Kingham2003) and hence divergence of the beam. The target chosen for the present experiment is CH–Al–CH and Al–CH–Al multilayered targets. First, given the large resistivity expected in dielectric materials, plastic targets are ideal candidates to investigate the generation of resistive fields (Jung et al., Reference Jung, Osterholz, Wenbrück, Kiselev, Pretzler, Pukhov, Willi, Kar, Borghesi, Nazarov, Karsch, Clarke and Neely2005). Second, using multilayers, it is possible to change the resistivity gradient, which can have a direct impact on the magnetic fields generated.

Fig. 2. (Color online): Proton beam profiles (a) 10 MeV, (b) 8 MeV, and (c) 6 MeV recorded on RCF from the target rear surface CH–Al–CH (2.5 µm); (d) shows a smooth beam recorded on RCF from the target rear surface of Al–CH–Al (2.5 µm).

The theory behind collimation of fast electrons in a structured target could be understood by combining ohm's law with Faraday's law for the growth rate of magnetic field resulting in the following equation:

(1)$$\displaystyle{{{\rm \partial} B} \over {{\rm \partial} t}} = {\rm \eta} \nabla \times j + \lpar {\nabla {\rm\eta}} \rpar \times j, $$

where η is the resistivity and ј is the fast electron current density. The first term corresponds to the magnetic fields that act to push fast electrons toward regions of higher fast electron density, while the second term corresponds to the magnetic fields that push the fast electrons toward regions of higher resistivity. The first term of Eq. (1) is the origin of electromagnetic instability. A fast electron current density perturbation normal to the streaming direction (i.e., normal to the direction of j) will generate the magnetic field that would push more fast electrons into the high density region from the low density region (the higher the density of fast electrons is, the higher the current density is: j). More and more fast electrons are constricted together, which further strengthens the density perturbation. Consequently, the generated field gets stronger further according to Eq. (1), and the density ripple magnitude gets larger and larger. As a result, the Weibel instability is created and filaments (a sequence of high and low density structures) are formed.

Figure 2a2c shows filamentation in the proton beam imprint from the rear side of the CH–Al–CH target, as recorded on an RCF arising due to self-generated resistive instabilities (Sentoku et al., Reference Sentoku, Mima, Kojima and Ruhi2000, Reference Sentoku, d'Humieres, Romagnani, Audebert and Fuchs2011). Figure 2d represents the proton beam profile of a 6 MeV RCF observed from the Al–CH–Al target, which represents a smooth collimated beam with no net-like pattern. Robinson and Sherlock (Reference Robinson and Sherlock2007) have highlighted the potential for harnessing these self-generated fields for guided electron beam transport in targets specifically designed to give rapid magnetic field growth at a well-defined boundary.

The divergence of the beam calculated for the CH–Al–CH foil is around 30° calculated from the recorded RCF data. The common inference from these figures is that whenever a low-Z thin transport layer is present on the rear side, we can observe a filamented proton beam imprint on the RCF. Similar experiments were conducted on the HZDR DRACO Laser facility using thin foil titanium (Ti) targets (2 and 3 µm), wherein they have reported transverse modulations in the proton beam profile (Metzkes et al., Reference Metzkes, Kluge, Zeil, Bussmann, Kraft, Cowan and Schramm2014). They have used a very thick hydrocarbon transport layer (3.5 µm) at the rear side of 2-µm Ti foil to control the spatial modulations in the beam. In this experiment, we have used the advantage of inherent resistivity gradient of the targets to control the filamentation without increasing the thickness of the rear side transport layer.

Figure 3a represents the magnified proton beam profile, which depicts an increasingly strong modulation for thinner targets. Particularly, striking is the profile which exhibits a high-contrast modulation with an elaborate net-like periodic uniform structure. Inspection of consecutive films illustrates that the structure persists deep into the stack with almost constant formation, indicating that it is formed well in advance of the detector surface. Also, the filamentary structures are constantly present in all ranges of energy as shown in Figure 2a2c, indicating that the filamentation has occurred well inside the target with the fast electrons and the corresponding structure is mapped onto the protons through the sheath electric field.

Fig. 3. (a) A zoom in the profile depicting a net-like pattern in the proton beam imprint. (b) A line out of the pattern shows periodicity in the net pattern.

The typical structure size at the stack is 0.2 cm over a beam that extends 5 cm on the stack. Assuming a beam originating from a radial size of 100–200 µm (Borghesi et al., Reference Borghesi, Mackinnon, Campbell, Hicks, Kar, Patel, Price, Romagnani, Schiavi and Willi2004) and that the expansion of the structure is ballistic from the target surface where it was generated to the detector, implies an instability wavelength, of 6–8 µm or less. This is greater than the typical wavelength of surface modulations on the foils, which is about 0.1 µm. We have further investigated the mechanism for the filamentation of the beam observed in the experiment. Resistivity (Gremillet et al., Reference Gremillet, Bonnaud and Amiranoff2002; Debayle and Tikhonchuk, Reference Debayle and Tikhonchuk2008; Robinson et al., Reference Robinson, Kingham, Ridgers and Sherlock2008), ionization (Krasheninnikov et al., Reference Krasheninnikov, Kim, Frolov and Stephens2005), and Weibel (Fried, Reference Fried1959; Weibel, Reference Weibel1959; Grassi et al., Reference Grassi, Grech, Amiranoff, Pegoraro, Macchi and Riconda2017) instabilities are typically invoked to explain the filamentation of fast currents observed experimentally (Borghesi et al., Reference Borghesi, Mackinnon, Bell, Malka, Vickers, Willi, Davies, Pukhov and Meyerter-Vehn1999; Gremillet et al., Reference Gremillet, Amiranoff, Baton, Gauthier, Koenig, Martinolli, Pisani, Bonnaud, Lebourg, Rousseaux, Toupin, Antonicci, Batani, Bernardinello, Hall, Scott, Norreys, Bandulet and Pepin1999; Tatarakis et al., Reference Tatarakis, Beg, Clark, Dangor, Edwards, Evans, Goldsack, Ledingham, Norreys, Sinclair, Wei, Zepf and Krushelnick2003; Wei et al., Reference Wei, Beg, Clark, Dangor, Evans, Gopal, Ledingham, McKenna, Norreys, Tatarakis, Zepf and Krushelnick2004). In physical terms, resistive filamentation occurs because any transverse perturbation in the current density will resistively grow a magnetic field that drives fast electrons into the perturbation. As this increases the local current density, the magnetic field gets further enhanced and a positive feedback is created. Our experimental observation that the filamentation wavelength λ = 2π/k = 6–8 µm agrees well with the reported wavelength for Weibel instability to be about 5–15 µm (Romagnani et al., Reference Romagnani, Robinson, Clarke, Doria, Lancia, Nazarov, Notley, Pipahl, Quinn, Ramakrishna, Wilson, Fuchs, Willi and Borghesi2019) using the proton probe technique. From our observations, we conclude that Weibel instabilities at faster time scales may be the root cause to seed resistive instabilities in the target leading to the filamentation.

PIC simulations

To investigate the filamentation and corresponding correlation to the divergence of the proton beam from sandwich targets, 2D particle-in-cell (PIC) simulation was carried out using the EPOCH (Arber et al., Reference Arber, Bennett, Brady, Lawrence-Douglas, Ramsay, Sircombe, Gillies, Evans, Schmitz, Bell and Ridgers2015). The simulation domain was chosen to be 40 µm × 40 µm along x- and y-directions with a grid resolution of 20 nm along each direction accommodating 32 macroparticles per cell. The laser pulse of duration 30 fs, wavelength 0.8 µm propagates along the positive x-direction with a transverse Gaussian profile. The laser focal spot and intensity were taken to be 8 µm and 3 × 1020 W/cm2 in accordance with the parameters used during the experiment. The simulations were carried out for sandwich targets Al–CH–Al (1–0.5–1 µm) and CH–Al–CH (1–0.5–1 µm) with n Al = 10n c, and for the plastic CH layer, we assumed n carbon = 10n c, n H = 20n c, where n c = 1.74 × 1021 cm−3. Both the targets were coated with a 50 nm proton layer on the rear side in order to account for the effect of contaminant layer in the experiment, structures of fast electrons are mapped on to the sheath electric field, which is directly correlated to the acceleration of protons. Therefore, protons accelerated from the rear side of the target would also exhibit filamentation as observed from the experiment.

Figure 4a represents that the magnetic field inside Al–CH–Al cease to grow once the modulations reach the boundary of the interface between CH–Al layers on the rear side, and therefore, the modulations are not propagated through to the rear surface of the target. Whereas in case of CH–Al–CH (Fig. 4b), it can be clearly seen that the filaments do not get suppressed near the boundary of the rear side layers and propagate through to the rear surface of the target. Similarly, the number density of electrons in Al–CH–Al (Fig. 4c) do not show filamentary structures in the layer corresponding to the rear side of the target, but CH–Al–CH (Fig. 4d) clearly has the filamentary structure in the layer at the rear side of the target. These filamentary structures of fast electrons are mapped on to the sheath electric field, which is directly correlated to the acceleration of protons. Therefore, protons accelerated from the rear side of the target would also exhibit filamentation as observed from the experiment.

Fig. 4. Magnetic field within the target for (a) Al–CH–Al and (b) CH–Al–CH. The electron number density within the target for (c) Al–CH–Al and (d) CH–Al–CH.

The parameter, which indicates the divergence of the proton beam, is the spread in the transverse momentum Py of the protons. The spread in the transverse momentum Py is measured at Px = 20 × 10−21 kg m/s for both the targets. Al–CH–Al (Fig. 5a) has a spread of 2.98 × 10−21 kg m/s, whereas CH–Al–CH (Fig. 5b) has a spread of 4.11 × 10−21 kg m/s, indicating that the protons from CH–Al–CH are more divergent than that from Al–CH–Al as observed from the experiment. The histogram of the electric field over the entire simulation domain is shown in Figure 6. The maximum value of the sheath electric field reached for Al–CH–Al is about 10 TV/m, whereas for CH–Al–CH, the maximum is about 6 TV/m. This implies that the formation of filamentary structures inside the target rear surface may have caused perturbation in the sheath field and, therefore, may have direct correlation in the observed reduction in the sheath electric field for the target CH–Al–CH.

Fig. 5. The phase space distribution of protons at 1ps for (a) Al–CH–Al and (b) CH–Al–CH.

Fig. 6. A histogram of the electric field for both the targets.

In conclusion, we have shown that self-resistive fields are very important to understand the transport of fast electron/ion beams, which can control the flow of mega-ampere currents in the intense laser-plasma interaction. This provides a direction to understand the instabilities arising at the interface of resistivity gradient targets that will otherwise lead to filamentation/pinching of the electron beams. Our experimental findings are in good agreement with the qualitative analysis performed using 2D PIC simulations. This offers a route to control and investigate these instabilities, which may otherwise prove detrimental for fast ignition experiments. Also, the magnetic fields of about 104 T are observed in the simulation, which may have implications in the astrophysical scenario (Cassam-Chenai et al., Reference Cassam-Chenai, Hughes, Reynoso, Badenes and Moffett2008; Suzuki-Vidal, Reference Suzuki-Vidal2015), that is, when a supernova collapses to form a neutron star a similar counter streaming charged particles may give rise to instability leading to generation of strong magnetic field. Charged particles streaming in such high magnetic fields may also contribute to gamma ray bursts.

Acknowledgments

B. Ramakrishna acknowledges financial support from the DST-YSS/2014/000622 and DST-DAAD (INT/FRG/DAAD/p-237/2014) fellowships from the Government of India. We thank the laser particle acceleration division of HZDR for their kind support and fruitful discussions. We thank Prof. Masakatsu Murakami and S. M. Weng for the discussions on PIC simulations. The EPOCH code used in this paper was, in part, funded by the UK EPSRC grants EP/G054950/1, EP/G056803/1, EP/G055165/1, and EP/ M022463/1.

References

Arber, TD, Bennett, K, Brady, CS, Lawrence-Douglas, A, Ramsay, MG, Sircombe, NJ, Gillies, P, Evans, RG, Schmitz, H, Bell, AR and Ridgers, CP (2015) Contemporary particle-in-cell approach to laser-plasma modelling. Plasma Physics and Controlled Fusion 57, 113001.CrossRefGoogle Scholar
Bell, AR and Kingham, RJ (2003) Resistive collimation of electron beams in laser-produced plasmas. Physical Review Letters 91, 035003.CrossRefGoogle ScholarPubMed
Bell, AR, Davies, JR, Guerin, S and Ruhl, H (1997) Fast-electron transport in high-intensity short-pulse laser-solid experiments. Plasma Physics and Controlled Fusion 39, 653.CrossRefGoogle Scholar
Betti, R and Hurricane, OA (2016) Inertial-confinement fusion with lasers. Nature Physics 12, 435.CrossRefGoogle Scholar
Borghesi, M, Mackinnon, AJ, Bell, AR, Malka, G, Vickers, C, Willi, O, Davies, JR, Pukhov, A and Meyerter-Vehn, J (1999) Observation of collimated ionization channels in aluminum coated glass targets irradiated by ultraintense laser pulses. Physical Review Letters 83, 4309.CrossRefGoogle Scholar
Borghesi, M, Mackinnon, AJ, Campbell, DH, Hicks, DG, Kar, S, Patel, PK, Price, D, Romagnani, L, Schiavi, A and Willi, O (2004) Multi-MeV proton source investigations in ultraintense laser-foil interactions. Physical Review Letters 92, 055003.CrossRefGoogle ScholarPubMed
Bret, A (2010) Collisional and collisionless beam plasma instabilities. Laser and Particle Beams 28, 491495.CrossRefGoogle Scholar
Bret, A, Fernandez, MFJ and Anfray, JM (2009) Unstable spectrum of a relativistic electron beam interacting with a quantum collisional plasma: application to the Fast Ignition Scenario. Plasma Physics and Controlled Fusion 51, 075011.CrossRefGoogle Scholar
Cassam-Chenai, G, Hughes, JP, Reynoso, EM, Badenes, C and Moffett, D (2008) Morphological evidence for azimuthal variations of the cosmic-ray ion acceleration at the blast wave of SN 1006. The Astrophysical Journal 680, 11801197.CrossRefGoogle Scholar
Chen, H, Wilks, SC, Bonlie, JD, Liang, EP, Myatt, J, Price, DF, Meyerhofer, DD and Beiersdorfer, P (2009) Relativistic positron creation using ultraintense short pulse lasers. Physical Review Letters 102, 105001.CrossRefGoogle ScholarPubMed
Clark, EL, Krushelnick, K, Davies, JR, Zepf, M, Tatarakis, M, Beg, FN, Machacek, A, Norreys, PA, Santala, MIK, Watts, I and Dangor, AE (2000) Measurement of energetic transport through magnetized plasma from intense laser interactions with solids. Physical Review Letters 84, 670.CrossRefGoogle Scholar
Cowan, TE, Perry, MD, Key, MH, Ditmire, TR, Hatchett, SP, Henry, EA, Moody, JD, Moran, MJ, Pennington, DM and Phillips, TW (1999) High energy electrons, nuclear phenomena and heating in petawatt laser-solid experiments. Laser and Particle Beams 17, 773.CrossRefGoogle Scholar
Craxton, RS, Anderson, KS, Boehly, TR, Goncharov, VN, Harding, DR, Knauer, JP, McCrory, RL, McKenty, PW, Meyerhofer, DD and Myatt, JF (2015) Direct-drive inertial confinement fusion: a review. Physics of Plasmas 22, 110501.CrossRefGoogle Scholar
Davies, JR, Bell, AR, Haines, MG and Guerin, SM (1997) Short-pulse high-intensity laser-generated fast electron transport into thick solid targets. Physical Review E 56, 7193.CrossRefGoogle Scholar
Debayle, A and Tikhonchuk, VT (2008) Filamentation instability of a fast electron beam in a dielectric target. Physical Review E 78, 066404.CrossRefGoogle Scholar
Fried, BD (1959) Mechanism for instability of transverse plasms waves. Physics of Fluids 2, 337.CrossRefGoogle Scholar
Fuchs, J, Antici, P, d'Humières, E, Lefebvre, E, Borghesi, M, Brambrink, E, Cecchetti, CA, Kaluza, M, Malka, V, Manclossi, M, Meyroneinc, S, Mora, P, Schreiber, J, Toncian, T, Pépin, H and Audebert, P (2006) Laser-driven proton scaling laws and new paths towards energy increase. Nature Physics 2, 48.CrossRefGoogle Scholar
Gahn, C, Pretzler, G, Saemann, A, Tsakiris, GD, Witte, KJ, Gassmann, D, Schätz, T, Schramm, U, Thirolf, P and Habs, D (1998) MeV γ-ray yield from solid targets irradiated with fs-laser pulses. Applied Physics Letters 73, 3662.CrossRefGoogle Scholar
Gahn, C, Tsakiris, GD, Pretzler, G, Witte, KJ, Delfin, C, Wahlström, CG and Habs, D (2000) Generating positrons with fs-laser pulses. Applied Physics Letters 77, 2662.CrossRefGoogle Scholar
Grassi, A, Grech, M, Amiranoff, F, Pegoraro, F, Macchi, A and Riconda, C (2017) Electron Weibel instability in relativistic counter streaming plasmas with flow-aligned external magnetic fields. Physical Review E 95, 023203.CrossRefGoogle Scholar
Gremillet, L, Amiranoff, F, Baton, SD, Gauthier, JC, Koenig, M, Martinolli, E, Pisani, F, Bonnaud, G, Lebourg, C, Rousseaux, C, Toupin, C, Antonicci, A, Batani, D, Bernardinello, A, Hall, T, Scott, D, Norreys, P, Bandulet, H and Pepin, H (1999) Time resolved observation of ultrahigh intensity laser-produced electron jets propagating through transparent solid targets. Physical Review Letters 83, 5015.CrossRefGoogle Scholar
Gremillet, L, Bonnaud, G and Amiranoff, F (2002) Filalmented transport of laser-generated relativistic electrons penetrating a solid target. Physics of Plasmas 9, 941.CrossRefGoogle Scholar
Jung, R, Osterholz, J, Wenbrück, KL, Kiselev, S, Pretzler, G, Pukhov, A, Willi, O, Kar, S, Borghesi, M, Nazarov, W, Karsch, S, Clarke, R and Neely, D (2005) Study of electron-beam propagation through preionized dense foam plasma. Physical Review Letters 94, 195001.CrossRefGoogle Scholar
Kar, S, Markey, K, Simpson, PT, Bellei, C, Green, JS, Nagel, SR, Kneip, S, Carroll, DC, Dromey, B, Willingale, L, Clark, EL, McKenna, P, Najmudin, Z, Krushelnick, K, Norreys, P, Clarke, RJ, Neely, D, Borghesi, M and Zepf, M (2008) Dynamic control of laser-produced proton beams. Physical Review Letters 100, 105004.CrossRefGoogle ScholarPubMed
Kar, S, Markey, K, Borghesi, M, Carroll, DC, McKenna, P, Neely, D, Quinn, MN and Zepf, M (2011) Ballistic focussing of polyenergetic protons driven by petawatt laser pulses. Physical Review Letters 106, 225003.CrossRefGoogle Scholar
Krasheninnikov, SI, Kim, AV, Frolov, BK and Stephens, R (2005) Intense electron beam propagation through insulators: ionization front structure and stability. Physics of Plasmas 12, 073105.CrossRefGoogle Scholar
Ledingham, KWD, Spencer, I, McCanny, T, Singhal, RP, Santala, MIK, Clark, E, Watts, I, Beg, FN, Zepf, M and Krushelnick, K (2000) Photonuclear physics when a multiterrawatt laser pulse interacts with solid targets. Physical Review Letters 84, 899.CrossRefGoogle Scholar
Maksimchuk, A, Gu, S, Flippo, K, Umstadter, D and Yu Bychenkov, V (2000) Forward ion acceleration in thin films driven by a high-intensity laser. Physical Review Letters 84, 4108.CrossRefGoogle ScholarPubMed
Metzkes, J, Kluge, T, Zeil, K, Bussmann, M, Kraft, SD, Cowan, TE and Schramm, U (2014) Experimental observation of transverse modulations in laser-driven proton beams. New Journal of Physics 16, 023008.CrossRefGoogle Scholar
Mora, P (2003) Plasma expansion into a vacuum. Physical Review Letters 90, 185002.CrossRefGoogle ScholarPubMed
Nishiuchi, M, Daito, I, Ikegami, M, Daido, H, Mori, M, Orimo, S, Ogura, K, Sagisaka, A, Yogo, A, Pirozhkov, AS, Sugiyama, H, Kiriyama, H, Okada, H, Kanazawa, S, Kondo, S, Shimomura, T, Tanoue, M, Nakai, Y, Sasao, H, Wakai, D, Sakaki, H, Bolton, P, Choi, IW, Sung, JH, Lee, J, Oishi, Y, Fujii, T, Nemoto, K, Souda, H, Noda, A, Iseki, Y and Yoshiyuki, T (2009) Focussing and spectral enhancement of a repetition-rated, laser-driven, divergent multi-MeV proton beam using permanent quadrupole magnets. Applied Physics Letters 94, 061107.CrossRefGoogle Scholar
Norreys, PA, Santala, M, Clark, E, Zepf, M, Watts, I, Beg, FN, Krushelnick, K, Tatarakis, M, Dangor, AE and Fang, X (1999) Observation of highly directional γ-ray beam from ultrashort, ultraintense laser pulse interactions with solids. Physics of Plasmas 6, 2150.CrossRefGoogle Scholar
Ramakrishna, B, Kar, S, Robinson, APL, Adams, DJ, Markey, K, Quinn, MN, Yuan, XH, McKenna, P, Lancaster, KL, Green, JS, Scott, RHH, Norreys, PA, Schreiber, J and Zepf, M (2010) Laser-driven fast electron collimation in targets with resistivity boundary. Physical Review Letters 105, 135001.CrossRefGoogle ScholarPubMed
Ramakrishna, B, Tayyab, M, Bagchi, S, Mandal, T, Upadhyay, A, Weng, SM, Murakami, M, Cowan, TE, Chakera, JA, Naik, PA and Gupta, PD (2015) Filamentation control and collimation of laser accelerated MeV protons. Plasma Physics and Controlled Fusion 12, 125013.CrossRefGoogle Scholar
Robinson, APL and Sherlock, M (2007) Magnetic collimation of fast electrons produced by ultraintense laser irradiation by structuring the target composition. Physics of Plasmas 14, 083105.CrossRefGoogle Scholar
Robinson, APL, Kingham, RJ, Ridgers, CP and Sherlock, M (2008) Effect of transverse density modulations on fast electron transport in dense plasmas. Plasma Physics and Controlled Fusion 50, 065019.CrossRefGoogle Scholar
Robson, L, Simpson, PT, McKenna, P, Ledingham, KWD and Clarke, RJ (2006) Scaling of proton acceleration driven by petawatt laser-plasma interactions. Nature Physics 3, 58.CrossRefGoogle Scholar
Romagnani, L, Robinson, APL, Clarke, RJ, Doria, D, Lancia, L, Nazarov, W, Notley, MM, Pipahl, A, Quinn, K, Ramakrishna, B, Wilson, PA, Fuchs, J, Willi, O and Borghesi, M (2019) Dynamics of the electromagnetic fields induced by fast electron propagation in near-solid-density media. Physical Review Letters 122, 025001.CrossRefGoogle ScholarPubMed
Sakagami, H, Johzaki, T, Sunahara, A and Nagatomo, H (2016) Integrated simulations for ion beam assisted fast ignition. Journal of Physics: Conference Series 1688, 012096.Google Scholar
Santala, MIK, Zepf, M, Watts, I, Beg, FN, Clark, E, Tatarakis, M, Krushelnick, K, Dangor, AE, McCanny, T, Spencer, I, Singhal, RP, Ledingham, KWD, Wilks, SC, Machacek, AC, Wark, JS, Allott, R, Clarke, RJ and Norreys, PA (2000) Effect of the plasma density scale length on the direction of fast electrons in relativistic laser-solid interactions. Physical Review Letters 84, 1459.CrossRefGoogle ScholarPubMed
Schramm, U, Bussmann, M, Irman, A, Siebold, M, Zeil, K, Albach, D, Bernert, C, Bock, S, Brack, F and Branco, J (2017) First results with the novel petawatt laser acceleration facility in Dresden. Journal of Physics: Conference Series 874, 012028.Google Scholar
Sentoku, Y, Mima, K, Kojima, S and Ruhi, H (2000) Magnetic instability by the relativistic laser pulses in overdense plasmas. Physics of Plasmas 7, 689.CrossRefGoogle Scholar
Sentoku, Y, d'Humieres, E, Romagnani, L, Audebert, P and Fuchs, J (2011) Dynamic control over mega-ampere electron currents in metals using ionization-driven resistive magnetic fields. Physical Review Letters 107, 135005.CrossRefGoogle ScholarPubMed
Snavely, RA, Key, MH, Hatchett, SP, Cowan, TE, Roth, M, Phillips, TW, Stoyer, MA, Henry, EA, Sangster, TC and Singh, MC (2000) Intense high energy proton beams from petawatt laser-irradiation of solids. Physical Review Letters 85, 2945.CrossRefGoogle ScholarPubMed
Strickland, D and Mourou, G (1985) Compression of amplified chirped pulses. Optics Communications 55, 447.CrossRefGoogle Scholar
Suzuki-Vidal, F (2015) How to spark a field. Nature Physics 11, 9899.CrossRefGoogle Scholar
Tatarakis, M, Beg, FN, Clark, EL, Dangor, AE, Edwards, RD, Evans, RG, Goldsack, TJ, Ledingham, KWD, Norreys, PA, Sinclair, MA, Wei, MS, Zepf, M and Krushelnick, K (2003) Propagation instabilities of high-intensity laser-produced electron beams. Physical Review Letters 90, 175001.CrossRefGoogle ScholarPubMed
Toncian, T, Borghesi, M, Fuchs, J, d'Humieries, E, Antici, P, Audebert, P, Brambrink, E, Cecchetti, CA, Pipahl, A, Romagnani, L and Willi, O (2006) Ultrafast laser-driven microlens to focus and energy select Mega-Electron Volt protons. Science 312, 410.CrossRefGoogle ScholarPubMed
Wei, MS, Beg, FN, Clark, EL, Dangor, AE, Evans, RG, Gopal, A, Ledingham, KWD, McKenna, P, Norreys, PA, Tatarakis, M, Zepf, M and Krushelnick, K (2004) Observations of the filamentation of high-intensity laser-produced electron beams. Physical Review E 70, 056412.CrossRefGoogle ScholarPubMed
Weibel, ES (1959) Spontaneously growing transverse waves in a plasma due to an anisotropic velocity distribution. Physical Review Letters 2, 83.CrossRefGoogle Scholar
Xu, XH, Liao, Q, Wu, MJ, Geng, YX, Li, DY, Zhu, JG, Li, CC, Hu, RH, Shou, YR, Chen, YH, Lu, HY, Ma, WJ, Zhao, YY, Zhu, K, Lin, C and Yan, XQ (2019) Detection and analysis of laser driven proton beams by calibrated Gafchromic HD-V2 and MD-V3 radiochromic films. Review of Scientific Instruments 90, 033306.CrossRefGoogle ScholarPubMed
Yu, J, Jiang, Z, Kieffer, JC and Krol, A (1999) Hard x-ray emission in high intensity femtosecond laser-target interaction. Physics of Plasmas 6, 1318.CrossRefGoogle Scholar
Figure 0

Fig. 1. (Color online): A schematic diagram of the experimental setup showing the RCF position behind the target.

Figure 1

Fig. 2. (Color online): Proton beam profiles (a) 10 MeV, (b) 8 MeV, and (c) 6 MeV recorded on RCF from the target rear surface CH–Al–CH (2.5 µm); (d) shows a smooth beam recorded on RCF from the target rear surface of Al–CH–Al (2.5 µm).

Figure 2

Fig. 3. (a) A zoom in the profile depicting a net-like pattern in the proton beam imprint. (b) A line out of the pattern shows periodicity in the net pattern.

Figure 3

Fig. 4. Magnetic field within the target for (a) Al–CH–Al and (b) CH–Al–CH. The electron number density within the target for (c) Al–CH–Al and (d) CH–Al–CH.

Figure 4

Fig. 5. The phase space distribution of protons at 1ps for (a) Al–CH–Al and (b) CH–Al–CH.

Figure 5

Fig. 6. A histogram of the electric field for both the targets.