Hostname: page-component-848d4c4894-nr4z6 Total loading time: 0 Render date: 2024-05-29T02:11:01.342Z Has data issue: false hasContentIssue false

Intense ү-Ray Bursts following the Interaction of Laser Pulse with Steep Density Gradients

Published online by Cambridge University Press:  01 January 2024

S. Chintalwad
Affiliation:
Department of Physics, Indian Institute of Technology Hyderabad, Kandi, Sangareddy 502284, India
S. Krishnamurthy
Affiliation:
Department of Physics, Indian Institute of Technology Hyderabad, Kandi, Sangareddy 502284, India
S. Morris
Affiliation:
Centre for Fusion, Space and Astrophysics, Department of Physics, University of Warwick, Coventry CV4 7AL, UK
B. Ramakrishna*
Affiliation:
Department of Physics, Indian Institute of Technology Hyderabad, Kandi, Sangareddy 502284, India
*
Correspondence should be addressed to B. Ramakrishna; bhuvan@phy.iith.ac.in
Rights & Permissions [Opens in a new window]

Abstract

We investigate the production of intense ү-rays following the interaction of ultraintense laser pulse with a hybrid combination of under-dense plasma associated with a thin foil of fully ionized Al or Cu or Au at the rear side. Relativistic electrons are accelerated following the interaction of high intensity laser pulses with an under-dense plasma. These electrons are then stopped by the thin foils attached to the rear side of the under-dense plasma. This results in the production of intense-ray bursts. So, the enhancement of photon generation is due to the under-dense plasma electrons interacting with different over-dense plasma. Using open-source PIC code EPOCH, we study the effect of different electron densities in the under-plasma on photon emission. Photon emission enhancement is observed by increasing the target Z in the hybrid structure. Hybrid structure can enhance photon emission; it can increase the photon energy and yield and improve photon beam divergence. Simulations were also performed to find the optimal under-dense plasma density for ү-ray production.

Type
Research Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an open access article distributed under the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.
Copyright
Copyright © 2022 S. Chintalwad et al.

1. Introduction

Recent developments in laser intensities of the order of 1022 to 1023 Wcm−2 are possible due to the production of femto-second pulses by the introduction of the Chirped Pulse Amplification (CPA) technique [Reference Strickland and Mourou1]. The increase in laser intensity [Reference Danson, Haefner and Bromage2] has advanced the fields related to science and technology. In particular, the laser-plasma interaction has been the frontier of research with regard to accelerator physics [Reference Cowan, Fuchs and Ruhl3], radiography [Reference Perry, Sefcik and Cowan4], radiation therapy [Reference Bulanov and Khoroshkov5], and laboratory astrophysics [Reference Bulanov, Esirkepov, Kando, Koga, Kondo and Korn6]. Relativistic electrons are accelerated to MeV energies with the currently available intense laser systems. The nonlinear processes [Reference Ta Phuoc, Corde and Thaury7] associated with these relativistic electrons lead to radiation ranging from X-rays to ү-rays. High-energy X-rays are employed for very high-resolution imaging such as Thomson scattering and high-energy sources. The bright ү-ray sources are also useful in nuclear physics related to waste management and in the production of ү-ү colliders. [Reference Pike, Mackenroth, Hill and Rose8], which are of fundamental importance to the science community. Besides using proton bunches generated from the ion acceleration in the inertial confinement fusion (ICF) application [Reference Honrubia, Fernández, Temporal, Hegelich and Meyer-ter-Vehn9Reference Macchi, Antonicci and Atzeni11], the bremsstrahlung radiation generated from laser wake field acceleration (LWFA) could also be used in ICF [Reference Tikhonchuk, Gu, Klimo, Limpouch and Weber12, Reference Betti and Hurricane13]. Edward et al. [Reference Edwards, Sinclair and Goldsack14] used laser-accelerated electrons in solid targets and under-dense plasma (gas-jet) to produce a highly collimated, high-flux beam. Radiography of complex and dense materials using ү-sources was demonstrated by Glinec et al. in 2005 [Reference Glinec, Faure and Le Dain15]. However, such sources could not be employed in practical applications due to poor control of beam parameters such as divergence and energy controlling the beam quality must be improved. Such a compact ү-ray source can be used in industrial applications such as detector calibration and testing.

PET (positron emission tomography) is a radioisotope-based imaging technology used in cancer diagnosis [Reference Halperin, Perez and Brady16]. Proton beams from conventional accelerators, such as LINAC, are generally used to produce radio isotopes. However, radioisotopes can also be produced by photo-nuclear processes arising from ү-ray irradiation. LWFA is a promising path for ү-ray-based radioisotope production which has emerged from current technology. This method uses compact tabletop lasers over high-end particle accelerators which are inexpensive and accessible. Radioisotope activation with the ү-rays produced by the bremsstrahlung of LWFA electrons have been demonstrated by various experiments [Reference Cowan, Hunt and Phillips17Reference Giulietti, Bourgeois and Ceccotti19]. For practical applications, it is crucial to achieve the desired beam features such as the source size, divergence, and electron-photon conversion efficiency. Using conventional accelerators in the low-energy range, analysis of bremsstrahlung sources for target material and thickness was performed by Marzini et al. [Reference Marziani, Taibi, Di Domenico and Gambaccini20], and LWFA-based bremsstrahlung sources are a key advancement for laser plasma-based particle sources which can be used for various applications. On the other hand, the Direct Laser Acceleration (DLA) mechanism accelerates electrons by laser electromagnetic waves in the ion channel or the plasma bubble through betatron resonance. In (DLA), the electrons in the ion channel or the plasma bubble are imparted with much higher energy than in vacuum [Reference Stupakov and Zolotorev21]. The (DLA) regime is also very attractive due to a wide variety of potential practical applications such as the production of the MeV range of ultrarelativistic electrons that can serve as compact X-ray radiation sources [Reference Stark, Toncian and Arefiev22]. Apart from this, the (DLA) regime is also useful in novel sources of other energetic particles such as ions [Reference Schollmeier, Sefkow and Geissel23], neutrons [Reference Pomerantz, McCary and Meadows24], and positrons [Reference Gahn, Tsakiris and Pretzler25].

Recent studies have shown that the ү-rays produced by laser-driven bremsstrahlung [Reference Kmetec, Gordon, Macklin, Lemoff, Brown and Harris26Reference Hatchett, Brown and Cowan28] and Compton scattering or Thomson scattering [Reference Ben-Ismal, Lundh and Rechatin29Reference Nerush, Kostyukov, Ji and Pukhov32] which are based on laser-wake field accelerators [Reference Cipiccia, Islam and Ersfeld33Reference Di Piazza, Müller, Hatsagortsyan and Keitel36] can produce highly collimated and quasimonochromatic ү-rays. Bremsstrahlung is produced when a charged particle is stopped by the strong electric fields inside the atomic nuclei. When a laser interacts with a plasma, the emitted photon energy is determined by the laser’s intensity, as well as target parameters such as thickness, atomic number, and the cross section of electron interaction with an atomic nucleus. The interaction of a high-intensity laser pulse with a target is divided into two steps: first, the laser-plasma interaction produces MeV electrons [Reference Forslund, Kindel and Lee37Reference Brunel39], and then, in a high Z solid target, the process of the production of bremsstrahlung when a high intensity laser pulse interacts with matter is basically due to the production of MeV electrons by laser and subsequent quenching of these electrons in the strong field of the atomic nuclei. The interaction cross section to produce bremsstrahlung radiation is increased with increase in Z of the target [Reference Koch and Motz40]. The processes involving the atomic nucleus, such as bremsstrahlung e + Z e + γ + Z , play an important role in the production of plasmas with a high charge state or/and large atomic number Z. The interaction of a high-energy electron bunch with a high-Z target such as copper or gold is where bremsstrahlung radiation would be most important. With the help of high laser intensities, the bremsstrahlung may produce comparable amounts of ү-ray radiation. Laser Wake-Field Acceleration (LWFA) [Reference Tajima and Dawson41] or (DLA) mechanisms can accelerate electrons in under-dense plasma. In a double-layer interaction setup [Reference Jirka, Klimo, Gu and Weber42, Reference Wang, Yan and Chen43], firstly, a high-intense laser pulse interacts with the under-dense plasma and accelerates hot electrons in the under-dense region by LWFA or DLA [Reference Pukhov, Sheng and Meyer-ter-Vehn44, Reference Gahn, Tsakiris and Pukhov45], and then these accelerated electrons are stopped by the nuclear potential of the foil target placed at the rear side of the under-dense region. Higher electron energy leads to more efficient ү-ray emission as the bremsstrahlung cross section increases with electron energy [Reference Morris, Robinson and Ridgers46, Reference Chintalwad, Krishnamurthy, Ramakrishna and Ridgers47]. In plasma, the bremsstrahlung emission is strongly dependent on the interaction cross section between the incoming electron and the nuclear potential of the target. Furthermore, the cross section depends on the Z of the target. The cross section (σb) is proportional to α r e 2 Z 2 where α = e 2 / c = 1 / 137 is the fine structure constant and r e = e 2 / m c 2 is the classical electron radius, respectively. The interaction cross section increases with increasing the Z of the target. For the Al (Z = 13) target, interaction cross section has σ b 169 α r e 2 , for Cu target (Z = 29), it is around σ b 841 α r e 2 , and for the Au target (Z = 79), cross section will be σ b 6241 α r e 2 [Reference and Motz48, Reference Tsai49].

In this paper, we investigate the photon emission by the bremsstrahlung mechanism; we use the open-source particle-in-cell (PIC) code EPOCH [Reference Arber, Bennett and Brady50] for this study. The bremsstrahlung mechanism is studied using the Monte Carlo (MC) algorithm, which is already included in the EPOCH code [Reference Morris, Robinson and Ridgers46]. We study the relative strength of the bremsstrahlung radiation for various densities of the under-dense plasma. When an ultra-intense laser with an intensity I = 4 × 10 21 W c m 2 is used to irradiate the under-dense plasma with the density of 0.02n c to 2n c, by attaching fully ionized targets such as Al or Au or Cu to the rear side of the under-dense plasma, the effect of the rear side target electron density on the emission of ү-rays is discussed in detail.

2. Simulation Set-Up

The simulations were performed with the 2D PIC code EPOCH. The simulation box covers (0 μm to 50 μm) along the x-direction and (−10 μm to 10 μm) along the y-direction with 5000 × 2000 cells with a cell size of 10 nm along the x-direction and y-direction. The under-dense hydrogen plasma with an electron density of (0.02n c to 2n c) is placed at 1 μm to 41 μm with 40 μm thickness. Electrons and ions have 32 macroparticles per cell. A P-polarized laser pulse of wavelength 0.8 μm with a focal spot radius of 1 μm is used. The peak intensity is 4 × 10 21 W c m 2 with a pulse duration of FWHM 30 fs. The laser pulse has a Gaussian spatial and temporal profile. The laser pulse enters the simulation box from x = 0 μm left boundary at the start of the simulation (t = 0), Al, or Cu, or Au with a thickness of 1 μm attached to the rear side of the under-dense plasma has been considered as a fully ionized plasma with the electron density n e = 455n c for Al, n e = 2238n c for Cu, and n e = 4234.4n c for Au where n c is electron critical density n c = ε 0 m e ω 0 2 / 4 π e 2 with laser frequency ω 0 = 2 π c / λ 0 where ε0, me, e, and λ are permittivity of free space, the mass of an electron, electron charge, and laser wavelength, respectively. The critical density corresponding to the laser wavelength λ = 0.8 μm is n c = 1.71 × 10 21 c m 3 . Only photons with energies greater than 50 keV are considered in all simulations. For computational convenience and in order to look into only the high-energy photons, we have considered only those photons with energy above 50 keV. The interaction set-up for double layer simulation is shown in Figure 1.

Figure 1: Simulation interaction set-up.

3. Results and Discussions

To study the photon emission in this double layer interaction setup, firstly, the interaction of a laser pulse with a 40 μm thick under-dense hydrogen plasma containing electrons and protons with a density of 0.2nc is considered. The interaction of ultraintense laser pulses with various under-dense plasma densities is then compared. The simulation results showing the electron density in under-dense plasma and corresponding electric field are presented in Figures 2(a) and 2(b). When an ultraintense laser pulse interacts with under-dense plasma, the laser pulse propagates through the plasma with minimal energy loss. As shown in Figure 2(a), the bubble regime is not formed sufficiently for the LWFA of the electron acceleration scheme. Figure 2(b) shows the corresponding transverse electric field in the under-dense plasma. When the laser pulse penetrates the plasma, the ponderomotive force causes electrons to be expelled sideways. Ions form a positively charged sheath behind the laser pulse because they are not expelled as quickly as electrons. The electrons are then pulled back towards the laser axis by the electrostatic field created.

Figure 2: (a) Electron density distribution in under-dense plasma for 0.2n c (b) corresponding transverse electric field, (c) electron energy, and (d) photon energy spectrum for 0.2n c.

The electron and photon energy spectra are shown in Figures 2(c) and 2(d). In the electron energy spectrum, the purple line indicates the electron energy in the under-dense plasma when the laser pulse approaches the end of the under-dense plasma, and the electron energy is higher in under-dense plasma than after interacting with the rear foil. The electron energy in the under-dense target is 189 MeV. The corresponding electron energy values for Al, Cu, and Au are 85 Mev, 75 MeV, and 80 MeV, respectively. Only when these high-energy electrons from the under-dense plasma interact with the rear foil attached to the under-dense target can photon emission be enhanced. Figure 2(d) shows the photon energy for all three foils; the obtained maximum photon energy values for Al, Cu, and Au are 61.3 MeV, 116 MeV, and 149.6 MeV, respectively. The highest photon energy is found in the case of the Au target. As increasing the Z of the target, the bremsstrahlung emission increases due to the increase in the interaction cross section between the electron and the ion. The photon number from the bremsstrahlung mechanism is shown in Figure 3.

Figure 3: Photon number (above) and angular distribution of photons (below) for 0.2n c.

The comparison of photon numbers generated from Al, Cu, or Au indicate that the number generated by Al is smaller than that generated by Cu and Au. The emitted photon number is found to be around 4 × 109, 4.2 × 1012, and 4.5 × 1013 for Al, Cu, and Au, respectively. The results indicate a maximum bremsstrahlung photon energy and the photon numbers for higher Z (Au). The simulation time is restricted to 300 fs, but the photon emission is expected to happen at a longer time scale. In addition, the angular distribution of photons is depicted in Figure 3. In Al target, the focused photon number along the θ = 0° is higher compared to Cu and Au. In case of Cu and Au, the emitted photons are mostly perpendicular to the laser direction, at about θ = ±90°. This transverse emission has also been observed by Morris, S. et al. [Reference Morris, Robinson and Ridgers46]. But in the case of the Au target, the emitted photons along the transverse directions are one order of magnitude higher than the Cu target. The photon emission direction is longitudinal for Al, and in the case of Cu or Au, the photons are emitted along the transverse direction.

We have performed additional simulations to understand the role of a higher under-dense plasma density and its corresponding effect on the emission of photons. We performed simulations by setting the density of the first layer ranging from 0.02n c to 2n c keeping all simulation parameters similar except for the under-dense plasma density. Figures 4(a) and 4(b) show the electron density and electric field for the n e = 2n c. Due to the self-focusing effect, the laser pulse gains a smaller transverse profile and a higher peak intensity; the electrons in the under-dense plasma are highly focused compared to the n e = 0.2n c. The laser pulse loses its significant energy within the under-dense region for higher plasma density and less amount of the high energy electron bunch gets reflected by the foil leading to reduced photon emission. The corresponding electron and photon energy spectra are shown in Figures 4(c) and 4(d). In the electron energy spectrum, the electron energy in the under-dense plasma is taken just before interacting with the foil in both the cases 0.2n c and 2n c. The electron energy in the under-dense target is 135 MeV, and electron energy for Al, Cu, and Au is 58 MeV, 65 MeV, and 68 MeV, respectively. The photon energy is higher in the Au case than in the Al and Cu cases. The corresponding photon energy for Al, Cu, and Au is 29.1 MeV, 55.3 MeV, and 81.3 MeV, respectively.

Figure 4: (a) Electron density distribution in under-dense plasma for 2n c (b) corresponding transverse electric field, (c) electron energy, and (d) photon energy spectrum for 2n c.

In addition, we also looked into the photon number and distribution of the photons in different targets in Figure 5. The emitted photon number is found to be around 1.4 × 108, 4 × 1012, and 4.3 × 1013 for Al, Cu, and Au, respectively. The generated photon number by the Al target is smaller than the Cu and Au targets. The maximum photon energy and the corresponding photons emitted are relatively higher for Au. In addition, we also depict the angular distribution of photons. In the Al target, the focused photons are relatively high compared to Cu and Au. But in the case of the Au target, the photons are comparatively higher than the Cu target. For a given laser pulse, the photon yield and energy can be affected by the parameters of under-dense plasma density. We performed a series of PIC simulations by varying the under-dense plasma density to obtain the optimal density for the highest photon yield and corresponding energy. Here, we changed the density of the under-dense plasma, but the length of the under-dense plasma is fixed to 40 μm. The results shown in Figure 6 are the photon energy dependence on different densities in all three targets. The photon number and energy decrease with increasing under-dense plasma density. The optimal density is found to be n e = 0.2n c. At this density, the number and energy of the photons are high compared to other densities. If the length and density of an under-dense plasma are too small, the resulting electron energy and number are limited. When under-dense plasma is too long and too dense, filamentation may occur, and thus, the laser absorption efficiency will be reduced in hydrogen layers with a higher density. Both are disadvantageous to the generation of high-energy photons.

Figure 5: Photon number (above) and angular distribution of photons (below) for 2n c.

Figure 6: The photon energy dependency on different densities in under-dense plasma.

4. Conclusion

A hybrid double-layer setup has been studied via 2D EPOCH simulations for the gamma ray production following the interaction of high intensity (I = 4 × 1021 Wcm−2) short-pulse laser. The results confirm photon energy enhancement for Au (high Z) target with the interaction of MeV electrons driven by the laser pulse interacting with the under-dense plasma in the hybrid structure. MeV electrons are accelerated in the under-dense region through well-known laser-plasma mechanisms such as LWFA and DLA. In conclusion, our simulations suggest an improved photon beam divergence and conversion efficiency of laser energy to photons if a high Z layer is present in the hybrid target. This study will be of fundamental importance for various applications in high-intensity laser-plasma interactions.

Data Availability

Data will be made available from the corresponding author upon reasonable request.

Conflicts of Interest

The authors declare that they have no conflicts of interest.

Acknowledgments

The authors would like to acknowledge the EPOCH code used in this research which the University of Warwick developed as part of UK EPSRC Project No. EP/G054940/1, EP/G054950/1, EP/G056803/1, EP/G055165/1, and EP/M022463/1. The authors also acknowledge the Council of Scientific and Industrial Research (CSIR) and imprint projects (IMP/2019/000275).

References

Strickland, D. and Mourou, G., “Compression of amplified chirped optical pulses,Optics Communications, vol. 56, no. 3, pp. 219221, 1985.10.1016/0030-4018(85)90120-8CrossRefGoogle Scholar
Danson, C. N., Haefner, C., Bromage, J. et al., “Petawatt and exawatt class lasers worldwide,High Power Laser Science and Engineering, vol. 7, p. e54, 2019.10.1017/hpl.2019.36CrossRefGoogle Scholar
Cowan, T. E., Fuchs, J., Ruhl, H. et al., “Ultralow emittance, multi-MeV proton beams from a laser virtual-cathode plasma accelerator,Physical Review Letters, vol. 92, no. 20, Article ID 204801, 2004.10.1103/PhysRevLett.92.204801CrossRefGoogle ScholarPubMed
Perry, M. D., Sefcik, J. A., Cowan, T. et al., “Hard x-ray production from high intensity laser solid interactions (invited),Review of Scientific Instruments, vol. 70, no. 1, pp. 265269, 1999.10.1063/1.1149442CrossRefGoogle Scholar
Bulanov, S. V. and Khoroshkov, V. S., “Feasibility of using laser ion accelerators in proton therapy,Plasma Physics Reports, vol. 28, no. 5, pp. 453456, 2002.10.1134/1.1478534CrossRefGoogle Scholar
Bulanov, S. V., Esirkepov, T. Z., Kando, M., Koga, J., Kondo, K., and Korn, G., “On the problems of relativistic laboratory astrophysics and fundamental physics with super powerful lasers,Plasma Physics Reports, vol. 41, no. 1, pp. 151, 2015.10.1134/S1063780X15010018CrossRefGoogle Scholar
Ta Phuoc, K., Corde, S., Thaury, C. et al., “All-optical Compton gamma-ray source,Nature Photonics, vol. 6, no. 5, pp. 308311, 2012.10.1038/nphoton.2012.82CrossRefGoogle Scholar
Pike, O. J., Mackenroth, F., Hill, E. G., and Rose, S. J., “A photon-photon collider in a vacuum hohlraum,Nature Photonics, vol. 8, no. 6, pp. 434436, 2014.10.1038/nphoton.2014.95CrossRefGoogle Scholar
Honrubia, J. J., Fernández, J. C., Temporal, M., Hegelich, B. M., and Meyer-ter-Vehn, J., “Fast ignition of inertial fusion targets by laser-driven carbon beams,Physics of Plasmas, vol. 16, no. 10, Article ID 102701, 2009.10.1063/1.3234248CrossRefGoogle Scholar
Roth, M., Cowan, T. E., Key, M. H. et al., “Fast ignition by intense laser-accelerated proton beams,Physical Review Letters, vol. 86, no. 3, pp. 436439, 2001.10.1103/PhysRevLett.86.436CrossRefGoogle ScholarPubMed
Macchi, A., Antonicci, A., Atzeni, S. et al., “Fundamental issues in fast ignition physics: from relativistic electron generation to proton driven ignition,Nuclear Fusion, vol. 43, no. 5, pp. 362368, 2003.10.1088/0029-5515/43/5/309CrossRefGoogle Scholar
Tikhonchuk, V., Gu, Y. J., Klimo, O., Limpouch, J., and Weber, S., “Studies of laser-plasma interaction physics with low-density targets for direct-drive inertial confinement schemes,Matter and Radiation at Extremes, vol. 4, no. 4, Article ID 045402, 2019.10.1063/1.5090965CrossRefGoogle Scholar
Betti, R. and Hurricane, O. A., “Inertial-confinement fusion with lasers,Nature Physics, vol. 12, Article ID 435448, 2016.Google Scholar
Edwards, R. D., Sinclair, M. A., Goldsack, T. J. et al., “Characterization of a gamma-ray source based on a laser-plasma accelerator with applications to radiography,Applied Physics Letters, vol. 80, no. 12, pp. 21292131, 2002.10.1063/1.1464221CrossRefGoogle Scholar
Glinec, Y., Faure, J., Le Dain, L. et al., “High-resolution-ray radiography produced by a laser-plasma driven electron source,Physical Review Letters, vol. 94, no. 2, Article ID 025003, 2005.10.1103/PhysRevLett.94.025003CrossRefGoogle ScholarPubMed
Halperin, E. C., Perez, C. A., and Brady, L. W., Principles and Practice of Radiation Oncology, Wolters Kluwer Health, Philadelphia, PA, USA, 2008.Google Scholar
Cowan, T. E., Hunt, A. W., Phillips, T. W. et al., “Photonuclear fission from high energy electrons from ultraintense laser-solid interactions,Physical Review Letters, vol. 84, no. 5, pp. 903906, 2000.10.1103/PhysRevLett.84.903CrossRefGoogle ScholarPubMed
Reed, S. A., Chvykov, V., Kalintchenko, G., Matsuoka, T., Rousseau, P., and Yanovsky, V., “Photonuclear fission with quasi-monoenergeticn electron beams from laser wake fields,Applied Physics Letters, vol. 89, no. 23, Article ID 231107, 2006.10.1063/1.2400400CrossRefGoogle Scholar
Giulietti, A., Bourgeois, N., Ceccotti, T. et al., “Intenseγ-ray source in the giant-dipole-resonance range driven by 10-TW laser pulses,Physical Review Letters, vol. 101, no. 10, Article ID 105002, 2008.10.1103/PhysRevLett.101.105002CrossRefGoogle ScholarPubMed
Marziani, M., Taibi, A., Di Domenico, G., and Gambaccini, M., “Optimization of radiography applications using x-ray beams emitted by compact accelerators. part I. Monte Carlo study of the hard x-ray spectrum,Medical Physics, vol. 36, no. 10, pp. 46834701, 2009.10.1118/1.3223357CrossRefGoogle ScholarPubMed
Stupakov, G. V. and Zolotorev, M. S., “Ponderomotive laser acceleration and focusing in vacuum for generation of attosecond electron bunches,Physical Review Letters, vol. 86, no. 23, pp. 52745277, 2001.10.1103/PhysRevLett.86.5274CrossRefGoogle ScholarPubMed
Stark, D. J., Toncian, T., and Arefiev, A. V., “Enhanced multi-MeV photon emission by a laser-driven electron beam in a self-generated magnetic field,Physical Review Letters, vol. 116, no. 18, Article ID 185003, 2016.10.1103/PhysRevLett.116.185003CrossRefGoogle Scholar
Schollmeier, M., Sefkow, A. B., Geissel, M. et al., “Laser-to-hot-electron conversion limitations in relativistic laser matter interactions due to multi-picosecond dynamics,Physics of Plasmas, vol. 22, no. 4, Article ID 043116, 2015.10.1063/1.4918332CrossRefGoogle Scholar
Pomerantz, I., McCary, E., Meadows, A. R. et al., “Ultrashort pulsed neutron source,Physical Review Letters, vol. 113, no. 18, Article ID 184801, 2014.10.1103/PhysRevLett.113.184801CrossRefGoogle ScholarPubMed
Gahn, C., Tsakiris, G. D., Pretzler, G. et al., “Generating positrons with femtosecond-laser pulses,Applied Physics Letters, vol. 77, Article ID 26622664, 2000.10.1063/1.1319526CrossRefGoogle Scholar
Kmetec, J. D., Gordon, C. L., Macklin, J. J., Lemoff, B. E., Brown, G. S., and Harris, S. E., “MeV x-ray generation with a femtosecond laser,Physical Review Letters, vol. 68, no. 10, pp. 15271530, 1992.10.1103/PhysRevLett.68.1527CrossRefGoogle ScholarPubMed
Norreys, P. A., Santala, M., Clark, E. et al., “Observation of a highly directional γ-ray beam from ultrashort, ultraintense laser pulse interactions with solids,Physics of Plasmas, vol. 6, no. 5, pp. 21502156, 1999.10.1063/1.873466CrossRefGoogle Scholar
Hatchett, S. P., Brown, C. G., Cowan, T. E. et al., “Electron, photon, and ion beams from the relativistic interaction of petawatt laser pulses with solid targets,Physics of Plasmas, vol. 7, no. 5, pp. 20762082, 2000.10.1063/1.874030CrossRefGoogle Scholar
Ben-Ismal, A., Lundh, O., Rechatin, C. et al., “Compact and high-quality-ray source applied to 10m-range resolution radiography,Applied Physics Letters, vol. 98, no. 26, Article ID 264101, 2011.Google Scholar
Brady, C. S., Ridgers, C. P., Arber, T. D., Bell, A. R., and Kirk, J. G., “Laser absorption in relativistically underdense plasmas by synchrotron radiation,Physical Review Letters, vol. 109, no. 24, Article ID 245006, 2012.10.1103/PhysRevLett.109.245006CrossRefGoogle ScholarPubMed
Liu, J., Yu, J., Shou, Y. et al., “Generation of bright γ-ray/hard x-ray flash with intense femtosecond pulses and double-layer targets,Physics of Plasmas, vol. 26, no. 3, Article ID 033109, 2019.10.1063/1.5085306CrossRefGoogle Scholar
Nerush, E. N., Kostyukov, I. Y., Ji, L., and Pukhov, A., “Gamma-ray generation in ultrahigh-intensity laser-foil interactions,Physics of Plasmas, vol. 21, no. 1, Article ID 013109, 2014.10.1063/1.4863423CrossRefGoogle Scholar
Cipiccia, S., Islam, M. R., Ersfeld, B. et al., “Gamma-rays from harmonically resonant betatron oscillations in a plasma wake,Nature Physics, vol. 7, no. 11, pp. 867871, 2011.10.1038/nphys2090CrossRefGoogle Scholar
Esarey, E., Schroeder, C. B., and Leemans, W. P., “Physics of laser-driven plasma-based electron accelerators,Reviews of Modern Physics, vol. 81, no. 3, Article ID 12291285, 2009.10.1103/RevModPhys.81.1229CrossRefGoogle Scholar
Schumaker, W., Sarri, G., Vargas, M. et al., “Measurements of high-energy radiation generation from laser-Wakefield accelerated electron beams,Physics of Plasmas, vol. 21, no. 5, Article ID 056704, 2014.10.1063/1.4875336CrossRefGoogle Scholar
Di Piazza, A., Müller, C., Hatsagortsyan, K. Z., and Keitel, C. H., “Extremely high-intensity laser interactions with fundamental quantum systems,Reviews of Modern Physics, vol. 84, no. 3, pp. 11771228, 2012.10.1103/RevModPhys.84.1177CrossRefGoogle Scholar
Forslund, D. W., Kindel, J. M., and Lee, K., “Theory of hot-electron spectra at high laser intensity,Physical Review Letters, vol. 39, no. 5, pp. 284288, 1977.10.1103/PhysRevLett.39.284CrossRefGoogle Scholar
Kruer, W. L. and Estabrook, K., “J×B heating by very intense laser light,Physics of Fluids, vol. 28, no. 1, pp. 430432, 1985.10.1063/1.865171CrossRefGoogle Scholar
Brunel, F., “Not-so-resonant, resonant absorption,Physical Review Letters, vol. 59, no. 1, pp. 5255, 1987.10.1103/PhysRevLett.59.52CrossRefGoogle ScholarPubMed
Koch, H. W. and Motz, J. W., “Bremsstrahlung cross-section formulas and related data,Reviews of Modern Physics, vol. 31, no. 4, pp. 920955, 1959.10.1103/RevModPhys.31.920CrossRefGoogle Scholar
Tajima, T. and Dawson, J. M., “Laser electron accelerator,Physical Review Letters, vol. 43, no. 4, pp. 267270, 1979.10.1103/PhysRevLett.43.267CrossRefGoogle Scholar
Jirka, M., Klimo, O., Gu, Y. J., and Weber, S, “Enhanced photon emission from a double-layer target at moderate laser intensities,Scientific Reports, vol. 10, p. 8887, 2020.10.1038/s41598-020-65778-4CrossRefGoogle ScholarPubMed
Wang, H. Y., Yan, X. Q., Chen, J. E. et al., “Efficient and stable proton acceleration by irradiating a two-layer target with a linearly polarized laser pulse,Physics of Plasmas, vol. 20, no. 1, Article ID 013101, 2013.Google Scholar
Pukhov, A., Sheng, Z. M., and Meyer-ter-Vehn, J., “Particle acceleration in relativistic laser channels,Physics of Plasmas, vol. 06, p. 2847, 1999.10.1063/1.873242CrossRefGoogle Scholar
Gahn, C., Tsakiris, G. D., Pukhov, A. et al., “Multi-mev electron beam generation by direct laser acceleration in high-density plasma channels,Physical Review Letters, vol. 83, p. 4772, 1999.10.1103/PhysRevLett.83.4772CrossRefGoogle Scholar
Morris, S., Robinson, A., and Ridgers, C., “Highly efficient conversion of laser energy to hard x-rays in high-intensity laser-solid simulations,Physics of Plasmas, vol. 28, no. 10, Article ID 103304, 2021.10.1063/5.0055398CrossRefGoogle Scholar
Chintalwad, S., Krishnamurthy, S., Ramakrishna, B., and Ridgers, C. P., “Photon emission enhancement studies from the interaction of ultra-intense laser pulses with shaped targets,Physical Review E—Statistical Physics, Plasmas, Fluids, and Related Interdisciplinary Topics, vol. 105, Article ID 025205, 2022.Google ScholarPubMed
and Motz, K., “Bremsstrahlung cross-section formulas and related data,Reviews of Modern Physics, vol. 31, 1959.Google Scholar
Tsai, Y.-S., “Pair production and bremsstrahlung of charged leptons,Reviews of Modern Physics, vol. 46, no. 4, pp. 815851, 1974.10.1103/RevModPhys.46.815CrossRefGoogle Scholar
Arber, T. D., Bennett, K., Brady, C. S. et al., “Contemporary particle-in-cell approach to laser-plasma modelling,Plasma Physics and Controlled Fusion, vol. 57, no. 11, Article ID 113001, 2015.10.1088/0741-3335/57/11/113001CrossRefGoogle Scholar
Figure 0

Figure 1: Simulation interaction set-up.

Figure 1

Figure 2: (a) Electron density distribution in under-dense plasma for 0.2nc (b) corresponding transverse electric field, (c) electron energy, and (d) photon energy spectrum for 0.2nc.

Figure 2

Figure 3: Photon number (above) and angular distribution of photons (below) for 0.2nc.

Figure 3

Figure 4: (a) Electron density distribution in under-dense plasma for 2nc (b) corresponding transverse electric field, (c) electron energy, and (d) photon energy spectrum for 2nc.

Figure 4

Figure 5: Photon number (above) and angular distribution of photons (below) for 2nc.

Figure 5

Figure 6: The photon energy dependency on different densities in under-dense plasma.