Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

All-optical bright γ-ray and dense positron source by laser driven plasmas-filled cone

Open Access Open Access

Abstract

An all-optical scheme for bright γ-rays and dense ee+ pair source is proposed by irradiating a 1022 W/cm2 laser onto a near-critical-density plasmas filled Al cone. Two-dimensional (2D) QED particle-in-cell (PIC) simulations show that, a dense electron bunch is confined in the laser field due to the radiation reaction and the trapped electrons oscillate transversely, emitting bright γ-rays forward in two ways: (1) nonlinear Compton scattering due to oscillation of electrons in the laser field, and (2) Compton backwardscattering resulting from the bunch colliding with the reflected laser by the cone tip. Finally, the multi-photon Breit-Wheeler process is initiated, producing abundant ee+ pairs with a density of ∼ 1027m−3. The scheme is further demonstrated by full 3D PIC simulations, which indicates a positron number up to 2 × 109. This compact γ-rays and ee+ pair source may have many potential applications, such as the laboratory study of astrophysics and nuclear physics.

© 2016 Optical Society of America

1. Introduction

The laser technology is developing quickly in recent years. For example, by upgrading HERCULES’s laser power to 300 TW, laser intensity of ∼ 2 × 1022 W/cm2 has been firstly demonstrated in experiment in 2008 [1]. A stronger electromagnetic field with E > 1015V/m can be supplied by an intense laser such as the ELI facilities in the near future [2]. Exposed in such a strong laser field, laser-matter interactions enter into the near quantum electrodynamics (QED) regime [3, 4]. It is predicted that the laser can breakdown the pure vacuum into electron-positron pairs when the Schwinger field, Es = 1.3×1018 V/m is approached, which corresponds to a light intensity up to 1029 W/cm2 [5]. Even in the case of a single real electron in the laser focus, the required laser intensity for pairs is still of order of 1025 W/cm2 [6]. For a lower laser intensity, it is also possible to generate positrons by laser driven electron-nuclei interaction or direct photon-photon collisions [7–22]. Up to now, several schemes have been proposed to produce dense positrons under current laser conditions. For early investigations, the attention has been focused on the trident process, which relies on the colliding of laser-accelerated electrons with a high-Z nuclei. However, the process is very inefficient due to the log dependence of the pair production cross section σei on the Lorentz factor γe [9, 10]. The Bethe-Heitler (BH) process [23] also attracted great attention, which makes use of the bremsstrahlung radiation [11, 12] by decelerated electrons in a thick target. The corresponding cross section of pairs production is usually 103 times larger than that with the trident process, making it very attractive for high-number ee+ pairs production [13, 17].

With the increase of laser intensities in recent years, the multi-photon Breit-Wheeler (BW) process has been revisited, because it is capable to enhance the positron production efficiency significantly [15–22]. This process can be simply described by two steps: (1) e + le + γh, where γl refers to the laser photons and γh represents the γ photons; (2) γh + le + e+. Burke et al., have observed a signal of 106 ± 14 positrons above background in experiment and the results are the first laboratory evidence for inelastic light-by-light scatting involving only real photons [21]. Recently, Pike et al., proposed an experimental scheme for a pure photon-photon collider, suggesting that pure light can be transformed into electron-positron pairs through the BW process [22]. In theory, the positron production probabilities are determined by the parameter [16] η = γe |E + βe × cB|/Es, where E and B are the components of the laser electromagnetic field and βe is electron velocity normalized by the light speed c. It is found that, for a laser propagating parallel to the γ photons emission, η ∼ 0, which implies few positrons generation. In order to increase the cross section of γ photons colliding with the laser photons, a counterpropagating laser pulse is usually required, with the first laser generating γ photons and the second providing low energy photons. In this case, the pair production becomes more efficient because of η ≥ 1. An alternative scheme is to substitute the second laser by a reflected laser from a solid target [20]. However, the γ photons emission relies significantly on the foil surface dynamics and the required laser intensity is above 4 × 1023 W/cm2, an order of magnitude higher than avaliable in current laboratories.

Here, we propose an all-optical scheme for bright γ-rays and dense positrons generation by irradiating a currently approachable laser onto near-critical-density plasmas. Our 2D PIC simulations with QED effects incorporated show that, the radiation damping force becomes large enough to compensate for the Lorentz force, causing radiation reaction trapping of a dense electron bunch in the laser field. The γ photons are emitted in two ways: (1) nonlinear Compton scattering due to the oscillation of electrons in the laser fields, and (2) Compton backwardscattering resulting from the bunch colliding with the reflected laser. As a result, abundant ee+ pairs are generated via the multi-photon BW process with a maximal positron density of 1027m−3. The scheme is further demonstrated by full 3D PIC simulations and the positron number is up to 2 × 109 with an average energy > 100 MeV. This compact γ-rays and pair source may have potential applications in many fields, e.g., laboratory studies of astrophysics and nuclear physics [24–27].

2. 2D PIC simulations and results

In the scheme, a plasma filled cone target is used because such targets have several features, such as efficient electron guiding and laser focusing. It can be attributed to the fact that the laser radiation tends to bend towards the area with a larger dielectric ε determined by [28]

ε=1ne/γenc,,
where ne is the electron density and nc is the critical density for the incident laser. On the one hand, the laser ponderomotive force pushes away the electrons from the laser axis, forming a lower density electron channel with an on-axis dielectrics larger than off-axis. It is beneficial to the laser focusing and further to decreasing the required laser intensity for radiation trapping of electrons [29,30]. On the other hand, the radial sheath electric field formed on the inner walls of the cone can effectively suppress the undesirable transverse explosion of the charged particles and improve their spatial properties [31].

We first carry out 2D simulations using QED-PIC code EPOCH [20,32]. The simulation box size is X × Y = 50λ0 × 16λ0 with λ0 = 1μm being the laser wavelength, which is sampled by 2000 × 400 cells with 20 particles in each cell. In order to reduce the required laser intensity, we use a plasma-filled cone target, as schematically shown in Fig. 1, which is initially located between 5λ0 ∼ 42λ0, with an electron density of ne=780nc and thickness of 2μm. Instead of the high-Z materials like gold, here we employ an aluminum cone so that the multiple-photon BW process is dominant over the trident process and BH process. The left cone opening radius is 6λ0, while the right tip radius is 3λ0. Meanwhile, hydrogen near-critical-density (NCD) plasma with a uniform density of 2nc is filled into the cone, which can increase the laser absorption and provide more background electrons for γ photons emission. A linearly polarized Gaussian laser pulse, with a transverse profile a=a0exp(y2/σ02), is incident from the middle of the left boundary and strikes on the right cone tip eventually, where σ0 =5λ0 is the laser focal spot radius and a0=150 is the dimensionless laser electric field amplitude, corresponding to an achievable laser intensity 3 × 1022 W/cm2 [1]. The laser pulse has a platform profile in time and the duration is τ0=10T0.

 figure: Fig. 1

Fig. 1 Schematic view of an intense linearly polarized laser striking a near-critical-density plasmas filled Al cone.

Download Full Size | PDF

As shown in Fig. 2(a), the conical target has a strong focusing effect on the laser pulse so that the laser peak intensity is enhanced from 3 × 1022 W/cm2 to 1.3 × 1023 W/cm2, corresponding to an enhanced transversely electric field ∼ 1015V/m at t=42T0. Exposed in such a strong laser field, hydrogen atoms are fully ionized, forming NCD plasmas. Simultaneously, electrons are rapidly accelerated to relativistic velocities with the Lorentz factor of the order of several hundreds. They emit abundant photons outwards and suffer from strong inward recoil forces by the photons. As a result, the radiation reaction effect becomes significant and the damping force gets comparable with the Lorentz force [33,34]. Figures 2(c) and 2(d) present the electron density and energy distribution in space at t = 42T0. We see a dense electron beam is trapped in the cone along the laser axial direction with a peak density of 7nc and energy up to 2.4 GeV at t=42T0. At t=44T0, the laser pulse penetrates through the plasma with its front arriving at the cone tip. A curved surface of the tip is thus formed by the strong laser pressure [35] and the laser wave is soon reflected, as illustrated in Fig. 2(b). Though the depletion of the laser pulse in the NCD plasmas is serious, we find the reflected laser electric field amplitude is still comparable with its original strength due to the focusing effect of the cone. This is very beneficial for the photon emission and the BW process in the following stage.

 figure: Fig. 2

Fig. 2 The transverse electric field evolution at t=42T0 (a) and 54T0 (b). Here, E0 = mecw0/e = 3.2 × 1012 V/m. The electron density (c) and energy distributions (d) at t=42T0.

Download Full Size | PDF

Figures 3(a)–3(d) demonstrate a bright γ-rays emission in the cone, which is produced with two independent processes: (1) nonlinear Compton scatting due to the oscillation of trapped electrons in the laser field, and (2) Compton backwardscattering resulting from the trapped electron bunch colliding with the reflected laser by the cone tip. Before the laser reflection at t=44T0, the trapped electrons have a high density and large transverse oscillation amplitude as shown in Figs. 2(c) and 2(d). This is actually the nonlinear Compton scattering, which promises high energy γ photons emission, as seen in Figs. 3(a) and 3(c). After the laser reflection, the co-moving electrons back-scatter the reflected laser photons to high energy γ photons. The spectrum of γ photons from Compton back-scatterings is broad and the maximum energy can be estimated as Eγ,max=4Ee2h¯ω0/(4Eeh¯ω0+(mec2)2), where Ee is the electron energy, and h̄ω0 is the laser photon energy. As seen in Fig. 2(d), Ee is up to 2.4 GeV at t=42T0 and will get higher later. For example, Ee ≈ 2.8 GeV at t=44T0, which leads to Eγ,max ≈ 141 MeV. This is in excellent agreement with the simulation results as shown in Fig. 3(a). It is noticed that the photon energy density gets larger, while the photon energy decreases significantly at t=48T0. This can be interpreted as follows. First, the photon energy decreasing results from the initiation of the BW process, which consumes the high energy part of γ photons to generate positrons. Figures 3(e) and 3(f) exhibit the positron density distribution at t=44T0 and 48T0. Only few positrons are produced before the laser reflection because the controlling parameter, η ∼ 0. By comparison, at t=48T0 a large number of positrons are observed, leading to the sharp decrease of the photon energy. Second, a standing wave is set up when the incident laser overlaps the reflected one. It further enhances the electric potential. On the other hand, the reflected laser by the curved surface of the tip is focused so that the subsequent Compton backscattering process gets efficient, producing high energy density photons. This interprets the increase of the photon energy density at t=48T0. In practice, we have to take into account of the pre-plasma before the cone tip, which may influence the laser reflection and alter the γ rays emission. However, the influences should be limited because of the thin thickness of the tip and sharp front of the laser pulse after its penetrating the NCD plasma.

 figure: Fig. 3

Fig. 3 The photon energy distribution [(a),(b)], the photon energy-density distribution [(c),(d)], and the positron density distribution [(e),(f)] at t=44T0 and 48T0, respectively.

Download Full Size | PDF

Figure 4(a) shows the positron energy spectra. As expected, only few positrons are generated at the very beginning. When the laser pulse is reflected, both the number and energy of positrons are dramatically increased. Here, the average Lorentz factor of the positrons generated inside the cone can be roughly determined by [36]

γ~acone=eEcone/mecwl,
where Econe is the electric field inside the cone and ωl is the laser frequency. In our case, Econe ≈ 1.42 × 1015V/m at t=46T0, leading to 〈γ〉 ≈ 441 MeV, in accordance with the simulations, ∼ 487 MeV. However, the average positron energy decreases to 187 MeV at t=48T0. It is due to the cone tip, which prevents the positrons from moving forward, as seen in Fig. 4(b).

 figure: Fig. 4

Fig. 4 The positron energy spectra (a) and the divergence angle distributions (b) at t=44T0, 46T0 and 48T0.

Download Full Size | PDF

3. The robustness of the scheme

In the process of the 2D simulations, we also consider the influences of the cone tip thicknesses and density of NCD plasmas on the pair production. As shown in Fig. 5(a), a 2μm thick tip is optimal for the positron generation under our laser conditions. This is easy to understand: a 1μm thick cone tip is so thin as to cause relativistic transparency of the tip for the laser pulse, making the laser reflection weaken and the BW process inefficient. By comparison, a 3μm tip is a litter thick, which absorbs considerable laser energy so that the efficiency of the BW process becomes low.

 figure: Fig. 5

Fig. 5 The positron number evolutions with different cone tip thicknesses (a) and NCD plasma densities (b).

Download Full Size | PDF

Meanwhile, the NCD plasma density also plays an important role for the positron generation. As seen in Fig. 5(b), we consider three different NCD plasmas: 0.8nc, 2nc, and 4nc. It is found that 2nc plasma is optimal for the positron generation in our scheme. When 0.8nc plasmas is filled in the cone, only few electrons are trapped so that the photon emission becomes trivial. On the contrary, using a 4nc plasma is also disadvantageous, because most laser energy is consumed by the NCD plasmas and only a small part laser wave is reflected by the cone tip. Therefor, an optimal value of the NCD plasma density exsits, e.g., 2nc in our case.

Finally, we conduct full 3D simulations to demonstrate the robustness of the scheme, as shown in Fig. 6. Here, the box size is X × Y × Z = 50λ0 × 16λ0 × 16λ0, sampled by 600 × 128 × 128 cells and 8 pseudoions per cell. All other parameters are the same as those in the 2D simulations. It is obvious to see that a dense relativistic electron bunch is trapped along the laser axis with a density of tens nc. These electrons oscillate in the lase fields and emit γ-rays forward. Finally, we obtain a positron beam with a peak density of 1nc and a number > 2 × 109. Meanwhile, the average positron energy is up to 107 MeV when laser front has just been reflected at t=48T0, which is smaller than that in the 2D case. As we know, in a more realistic 3D case, the focusing effect of the cone on the laser pulse becomes weakened due to the another dimension along the Z direction as compared with the 2D case. This results in a smaller laser electric field in the cone (not shown here). Finally, the positron energy gets smaller according to the Eq. (2).

 figure: Fig. 6

Fig. 6 3D simulation results: (a) the electron density distribution at t=57T0, (b) the positron density distribution at t=57T0, and the positron energy spectrum evolution.

Download Full Size | PDF

4. Conclusion

In conclusion, we propose an all-optical scheme for bright γ-rays emission and dense ee+ pair production by using the QED-PIC simulations. It is shown that the radiation reaction leads to the electron trapping in the laser field. These electrons oscillate transversely and emit high-energy γ photons through the nonlinear Compton scattering and Compton backwardscattering processes. Finally, the multi-photon BW process is initiated and dense ee+ pairs are produced with a density of ∼ 1027m−3. Full 3D PIC simulations indicate the final positron number is up to 109, which may have many potential applications in fundamental science and astrophysics.

Acknowledgments

This work was financially supported by the National Natural Science Foundation of China (Project Nos.11474360, 11205243, 11175255, 91230205, and 11375265), the National Basic Research Program of China (Grant No. 2013CBA01502), and the Research Project of NUDT (JQ14-02-02). The authors wish to acknowledge CFSA at University of Warwick for allowing usage of EPOCH.

References and links

1. V. Yanovsky, V. Chvykov, G. Kalinchenko, P. Rousseau, T. Planchon, T. Matsuoka, A. Maksimchuk, J. Nees, G. Cheriaux, G. Mourou, and K. Krushelnick, “Ultra-high intensity- 300-TW laser at 0.1 Hz repetition rate,” Opt. Express 16, 2109–2114 (2008). [CrossRef]   [PubMed]  

2. See www.eli-np.ro for Extreme Light Infrastructure-Nuclear Physics.

3. V. I. Ritus, “Quantum effects of the interaction of elementary particles with an intense electromagnetic field,” J. Russ. Laser Res. 6, 497–617 (1985). [CrossRef]  

4. A. Di Piazza, C. Müller, K. Z. Hatsagortsyan, and C. H. Keitel, “Extremely high-intensity laser interactions with fundamental quantum systems,” Rev. Mod. Phys. 84, 1177–1228 (2012). [CrossRef]  

5. J. Schwinger, “On gauge invariance and vacuum polarization,” Phys. Rev. 82, 664–679 (1951). [CrossRef]  

6. A. M. Fedotov, N. B. Narozhny, G. Mourou, and G. Korn, “Limitation on the attainable intensity of high power lasers,” Phys. Rev. Lett. 105, 080402 (2010). [CrossRef]  

7. G. Sarri, K. Poder, J. M. Cole, W. Schumaker, A. Di Piazza, B. Reville, T. Dzelzainis, T. Dzelzainis, D. Doria, L. A. Gizzi, G. Grittani, S. Kar, C. H. Keitel, K. Krushelnick, S. Kuschel, S. P. D. Mangles, Z. Najmudin, N. Shukla, L. O. Silva, D. Symes, A. G. R. Thomas, M. Vargas, J. Vieira, and M. Zepf, “Generation of neutral and high-density electron-positron pair plasmas in the laboratory,” Nat. Commun. 6, 6747 (2015). [CrossRef]   [PubMed]  

8. G. Sarri, W. Schumaker, A. Di Piazza, M. Vargas, B. Dromey, M. E. Dieckmann, V. Chvykov, A. Maksimchuk, V. Yanovsky, Z. H. He, B. X. Hou, J. A. Nees, A. G. R. Thomas, C. H. Keitel, and M. Zepf, “Table-top laser-based source of femtosecond, collimated, ultrarelativistic positron beams,” Phys. Rev. Lett. 110, 255002 (2013). [CrossRef]   [PubMed]  

9. E. P. Liang, S. C. Wilks, and M. Tabak, “Pair production by ultraintense lasers,” Phys. Rev. Lett. 81, 4887–4890 (1998). [CrossRef]  

10. K. Nakashima and H. Takabe, “Numerical study of pair creation by ultraintense lasers,” Phys. Plasmas 9, 1505–1512 (2002). [CrossRef]  

11. D. A. Gryaznykh, Y. Z. Kandiev, and V. A. Lykov, “Estimates of electron-positron pair production in the interaction of high-power laser radiation with high-Z targets,” JETP Lett. 67, 257–262 (1998). [CrossRef]  

12. H. Chen, S. C. Wilks, J. D. Bonlie, E. P. Liang, J. Myatt, D. F. Price, D. D. Meyerhofer, and P. Beiersdorfer, “Relativistic positron creation using ultraintense short pulse lasers,” Phys. Rev. Lett. 102, 105001 (2009). [CrossRef]   [PubMed]  

13. D. Li, K. Imasaki, S. Miyamoto, K. Horikawa, K. Ishihara, S. Amano, and T. Mochizuki, “Positron generation through laser Compton scattering gamma ray,” Appl. Phys. Lett. 94, 091112 (2009). [CrossRef]  

14. Y. H. Yan, Y. C. Wu, Z. Q. Zhang, J. Teng, J. Q. Yu, D. X. Liu, K. G. Dong, L. Wei, W. Fan, L. F. Cao, Z. E. Yao, and Y. Q. Gu, “Monte Carlo simulation study of positron generation in ultra-intense laser-solid interactions,” Phys. Plasmas 19, 023114 (2012). [CrossRef]  

15. G. Breit and J. A. Wheeler, “Collision of two light quanta,” Phys. Rev. 46, 1087–1091 (1934). [CrossRef]  

16. A. R. Bell and J. G. Kirk, “Possibility of prolific pair production with high-power lasers,” Phys. Rev. Lett. 101, 200403 (2008). [CrossRef]   [PubMed]  

17. J. G. Kirk, A. R. Bell, and I. Arka, “Pair production in counter-propagating laser beams,” Plasma Phys. Controlled Fusion 51, 085008 (2009). [CrossRef]  

18. J. X. Liu, Y. Y. Ma, J. Zhao, T. P. Yu, X. H. Yang, L. F. Gan, G. B. Zhang, J. F. Yan, H. B. Zhuo, J. J. Liu, Y. Zhao, and S. Kawata, “High-flux low-divergence positron beam generation from ultra-intense laser irradiated a tapered hollow target,” Phys. Plasmas 22, 103102 (2015). [CrossRef]  

19. W. Luo, Y. B. Zhu, H. B. Zhuo, Y. Y. Ma, Y. M. Song, Z. C. Zhu, X. D. Wang, X. H. Li, I. C. E. Turcu, and M. Chen, “Dense electron-positron plasmas and gamma-ray bursts generation by counterpropagating quantum electrodynamics-strong laser interaction with solid targets,” Phys. Plasmas 22, 063112 (2015). [CrossRef]  

20. C. P. Ridgers, C. S. Brady, R. Duclous, J. G. Kirk, K. Bennett, T. D. Arber, A. P. L. Robinson, and A. R. Bell, “Dense electron-positron plasmas and ultraintense γ rays from laser-irradiated solids,” Phys. Rev. Lett. 108, 165006 (2012). [CrossRef]  

21. D. L. Burke, R. C. Field, G. Horton-Smith, J. E. Spencer, D. Walz, S. C. Berridge, W. M. Bugg, K. Shmakov, A. W. Weidemann, C. Bula, K. T. McDonald, E. J. Prebys, C. Bamber, S. J. Boege, T. Koffas, T. Kotseroglou, A. C. Melissinos, D. D. Meyerhofer, D. A. Reis, and W. Ragg, “Positron production in multiphoton light-by-light scattering,” Phys. Rev. Lett. 79, 1626–1629 (1997). [CrossRef]  

22. O. J. Pike, F. Mackenroth, E. G. Hill, and S. J. Rose, “A photon-photon collider in a vacuum hohlraum,” Nat. Photonics 8, 434–436 (2014).

23. W. Heitler, The Quantum Theory of Radiation (Clarendon, 1954).

24. B. A. Remington, “High energy density laboratory astrophysics,” Plasma Phys. Controlled Fusion 47, A191–A203 (2005). [CrossRef]  

25. R. Ruffini, G. Vereshchagin, and S. S. Xue, “Electron-positron pairs in physics and astrophysics: from heavy nuclei to black holes,” Phys. Rep. 487, 1–140 (2010). [CrossRef]  

26. D. W. Gidley, H. G. Peng, and R. S. Vallery, “Positron annihilation as a method to characterize porous materials,” Annu. Rev. Mater. Res. 36, 49–79 (2006). [CrossRef]  

27. Y. C. Wu, J. Jiang, S. J. Wang, A. Kallis, and P. G. Coleman, “Porosity and crystallization of water ice films studied by positron and positronium annihilation,” Phys. Rev. B 84, 064123 (2011). [CrossRef]  

28. T. P. Yu, W. Yu, F. Q. Shao, S. X. Luan, D. B. Zou, Z. Y. Ge, G. B. Zhang, J. W. Wang, W. Q. Wang, X. H. Li, J. X. Liu, J. M. Ouyang, and A. Y. Wong, “High-energy-density electron jet generation from an opening gold cone filled with near critical-density plasma,” J. Appl. Phys. 117, 023105 (2015). [CrossRef]  

29. L. L. Ji, A. Pukhov, I. Yu. Kostyukov, B. F. Shen, and K. Akli, “Radiation-reaction trapping of electrons in extreme laser fields,” Phys. Rev. Lett. 112, 145003 (2014). [CrossRef]   [PubMed]  

30. X. L. Zhu, Y. Yan, T. P. Yu, F. Q. Shao, Z. Y. Ge, W. Q. Wang, and J. J. Liu, “Enhanced electron trapping and gamma-ray emission by ultra-intense laser irradiating a near-critical density plasma filled gold cone,” New J. Phys. 17, 053039 (2015). [CrossRef]  

31. D. B. Zou, H. B. Zhou, T. P. Yu, H. C. Wu, X. H. Yang, F. Q. Shao, Y. Y. Ma, Y. Yin, and Z. Y. Ge, “Enhanced laser-radiation-pressure-driven proton acceleration by moving focusing electric-fields in a foil-in-cone target,” Phys. Plasmas 22, 023109 (2015). [CrossRef]  

32. C. P. Ridgers, C. S. Brady, R. Duclous, J. G. Kirk, K. Bennett, T. D. Arber, and A. R. Bell, “Dense electron-positron plasmas and bursts of gamma-rays from laser-generated quantum electrodynamic plasmas,” Phys. Plasmas 20, 056701 (2013). [CrossRef]  

33. M. Chen, A. Pukhov, T. P. Yu, and Z. M. Sheng, “Radiation reaction effects on ion acceleration in laser foil interaction,” Plasma Phys. Control. Fusion 53, 86–89 (2011). [CrossRef]  

34. T. P. Yu, A. Pukhov, Z. M. Sheng, F. Liu, and G. Shvets, “Bright betatronlike X rays from radiation pressure acceleration of a mass-limited foil target,” Phys. Rev. Lett. 110, 045001 (2013). [CrossRef]   [PubMed]  

35. T. P. Yu, A. Pukhov, G. Shvets, M. Chen, T. H. Ratliff, S. A. Yi, and V. Khudik, “Simulations of stable compact proton beam acceleration from a two-ion-species ultrathin foil,” Phys. Plasmas 18, 043110 (2011). [CrossRef]  

36. J. Koga, T. Z. Esirkepov, and S. V. Bulanov, “Nonlinear Thomson scattering in the strong radiation damping regime,” Phys. Plasmas 12, 093106 (2005). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (6)

Fig. 1
Fig. 1 Schematic view of an intense linearly polarized laser striking a near-critical-density plasmas filled Al cone.
Fig. 2
Fig. 2 The transverse electric field evolution at t=42T0 (a) and 54T0 (b). Here, E0 = mecw0/e = 3.2 × 1012 V/m. The electron density (c) and energy distributions (d) at t=42T0.
Fig. 3
Fig. 3 The photon energy distribution [(a),(b)], the photon energy-density distribution [(c),(d)], and the positron density distribution [(e),(f)] at t=44T0 and 48T0, respectively.
Fig. 4
Fig. 4 The positron energy spectra (a) and the divergence angle distributions (b) at t=44T0, 46T0 and 48T0.
Fig. 5
Fig. 5 The positron number evolutions with different cone tip thicknesses (a) and NCD plasma densities (b).
Fig. 6
Fig. 6 3D simulation results: (a) the electron density distribution at t=57T0, (b) the positron density distribution at t=57T0, and the positron energy spectrum evolution.

Equations (2)

Equations on this page are rendered with MathJax. Learn more.

ε = 1 n e / γ e n c , ,
γ ~ a cone = e E cone / m e cw l ,
Select as filters


Select Topics Cancel
© Copyright 2023 | Optica Publishing Group. All Rights Reserved