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

Ultra-bright γ-ray flashes and dense attosecond positron bunches from two counter-propagating laser pulses irradiating a micro-wire target

Open Access Open Access

Abstract

We propose a novel scheme to generate ultra-bright ultra-short γ-ray flashes and high-energy-density attosecond positron bunches by using multi-dimensional particle-in-cell simulations with quantum electrodynamics effects incorporated. By irradiating a 10 PW laser pulse with an intensity of 1023 W/cm2 onto a micro-wire target, surface electrons are dragged-out of the micro-wire and are effectively accelerated to several GeV energies by the laser ponderomotive force, forming relativistic attosecond electron bunches. When these electrons interact with the probe pulse from the other side, ultra-short γ-ray flashes are emitted with an ultra-high peak brightness of 1.8 × 1024 photons s−1mm−2mrad−2 per 0.1%BW at 24 MeV. These photons propagate with a low divergence and collide with the probe pulse, triggering the Breit-Wheeler process. Dense attosecond ee+ pair bunches are produced with the positron energy density as high as 1017 J/m3 and number of 109. Such ultra-bright ultra-short γ-ray flashes and secondary positron beams may have potential applications in fundamental physics, high-energy-density physics, applied science and laboratory astrophysics.

© 2017 Optical Society of America

Corrections

29 August 2017: A typographical correction was made to the title.

1. Introduction

With the development of multi-PW laser facilities such as Extreme Light Infrastructure (ELI) [1] and SG-II laser facilities [2], people are expected to generate femtosecond optical pulses of intensity up to 1023−24 W/cm2, which opens up wide possibilities for theoretical and experimental investigation of strong-field quantum electrodynamics (QED) effects. It is predicted that the QED vacuum is unstable [3] when exposed to ultra-intense laser light with electric field approaching the Schwinger limit [4] (ES = 1.3 × 1018 V/m), which corresponds to a light intensity up to 1029 W/cm2. Finally, ee+ pairs can be spontaneously created from vacuum. However, the required laser intensity is at least 6 orders of magnitude higher than the current achievable, making it unlikely to demonstrate the pairs creation directly under terrestrial laboratory conditions. Fortunately, recent studies showed that the laser-plasma interaction in the near-QED regime can significantly reduce the threshold of the laser intensity for dense pair creation. With the projected PW laser facilities, this offers a very promising approach to creating dense ee+ pairs by ultra-intense laser-plasma interactions.

At the intensity above 1023 W/cm2, the laser-matter interaction becomes strongly dissipative [5] and enters into the near QED regime. Efficient laser energy transform into high energy γ-rays occurs, which enables copious ee+ pair production via various mechanisms, such as trident process, Bethe-Heitler (B-H) process [6] and Breit-Wheeler (B-W) process [7]. For example, irradiating an ultra-thin foil by an ultra-intense laser pulse with intensity ~ 1023 W/cm2, the multi-photon B-W process typically has a much larger cross-section of the pair creation and therefore dominates the pair production over the trident process and B-H process [8, 9].

High energy γ-ray flashes and dense pair plasmas are ubiquitous in the high energy astrophysical environment, such as γ-ray bursts, black hole and pulsar wind nebulae [5, 10]. The possibility of probing the process of laboratory astrophysics and strong-field QED has motivated the pursuit of the generation of bright γ-ray bursts and high-flux relativistic pair jets [11–17]. Recently, a large amount of effort has been dedicated to acquiring bright γ-rays and dense positrons in the near-QED regime. For example, Ridgers et al. [12] numerically demonstrated the production of pure ee+ pair plasmas by an ultra-intense laser striking solid foils. During the hole-boring process, energetic electrons radiate high energy γ-photons via the so-called skin-depth emission, which is able to trigger the B-W process so that dense relativistic ee+ pairs are produced in the laser fields. Luo et al. [13] and Chang et al. [14] proposed an alternative method by irradiating the target with two ultra-intense colliding laser pulses. Due to the symmetric compression of the target, the positron production is significantly enhanced. Recently, Liu et al. [15] and Zhu et al. [16, 17] reported, for the first time, the bright γ-ray source and dense positron beam generation in the near-critical-density (NCD) plasmas. By using double cone targets, the required laser intensity for the radiation trapping effect is reduced by two orders of magnitude, so that the B-W process is triggered at current laser intensities. It is also shown that a dense positron bunch could be generated in gas plasmas driven by two counter-propagating laser pulses [8]. In this scenario, the first laser is responsible for accelerating the electron bunches to GeV energies in the laser wakefield regime [18, 19], while the second is used for triggering the nonlinear Compton scattering and B-W process. In spite of these different concepts, the proposed schemes seem to be complicated, which depend on either an engineering-shaped cone structure or staged acceleration. Meanwhile, the obtained γ-photons are highly localized and the resulting positrons merger into a single bunch. This poses a great challenge in the beam manipulating and processing for potential applications where special structure of photon beam and positron bunch are required.

In this paper, we propose a novel scheme for ultra-bright γ-ray flashes and high-energy-density positron bunches generation. By irradiating an ultra-intense laser pulse with intensity of 1023 W/cm2 onto a micro-wire target, the electrons near the wire surface are quickly dragged-out by the laser transverse fields, forming attosecond bunches in space, and are effectively accelerated to several GeV energies by the laser ponderomotive force. These attosecond electron bunches behave like “relativistic flying mirrors (RFMs)” [5], which subsequently backscatter the probe laser photons to high energy γ-photons along the electron motion direction. The emitted γ-photons are bunched in space, forming ultra-high brightness γ-ray flashes characterized by a spatial gap of about one laser wavelength. As soon as they collide with the probe pulse, the multi-photon B-W process is triggered and dense attosecond ee+ pair bunches are produced. Our multi-dimensional simulations indicate that the final positron energy density is as high as 1 × 1017 J/m3 and the positron number is up to 109 per shot. The produced ultra-bright ultra-short γ-ray flashes and attosecond positron bunches with high energy and density may have diverse applications in fundamental physics, applied science and laboratory astrophysics.

2. Two-dimensional simulations and results

We first carry out two-dimensional (2D) particle-in-cell (PIC) simulations using the open source code EPOCH [20] to demonstrate feasibility of the scheme. The simulation box size is x × y = 14λL × 12λL, which is sampled by 5600 × 3600 cells with 64 pseudo-electrons and 36 pseudo-carbons in each cell. Here λL = cT0 = 1 μm is the laser wavelength and T0 is the laser period. The left driving pulse is linearly polarized along the y direction, which is grazing incident on the wire target from the left to the right, as schematically shown in Fig. 1. The laser profile can be expressed as a=a0e(y/σL)2e(t/τL)2, where a0 = eE0/me0 is the dimensionless amplitude of the laser electric field with e and me the charge and mass of electron, c the speed of light in vacuum, and E0 the laser peak electric field strength, σL = 4.5λL is the focus spot radius and τL = 4T0 is the pulse duration. The probe pulse is polarized along the same direction and propagates from the right to the left. Its profile is given by a=a0e(y/σL)2 sin4(πt/τL′), with τL = 3T0. Both laser pulses have the same peak intensity of 1023 W/cm2. The wire target is located between x0 = 1 μm and x1 = 7 μm, with radius of 0.2 μm. The wire material is graphite with realistic density of 1.2 g/cm3, corresponding to a number density of ne = 360nc, where nc = 1.1 × 1021 cm−3 is the critical density for λL = 1 μm laser pulse. Since the laser intensity under consideration is extremely high, we assume the target to be completely ionized (C6+) and its initial temperature is 1 keV. Here the skin-depth of the plasma δ=(λL/2π)nc/ne can be resolved, while the initial Debye length is extremely small. However, due to the abrupt increase of the plasma temperature, the Debye length becomes much larger soon so that it can be resolved in only 1T0 of the laser-wire target interaction.

 figure: Fig. 1

Fig. 1 Schematic view of ultra-bright γ-ray flashes emission and attosecond ee+ bunches generation by two ultra-intense colliding lasers irradiating a micro-wire target.

Download Full Size | PDF

2.1. Attosecond electron bunches generation

High-energy-density electrons are vital to stimulate the QED effects in ultra-intense laser fields because they determine the cross-section of the electron-photon collision. During the laser-wire interaction under discussion, the amplitude of the driving laser’s electric field is sufficiently strong to immediately break the Langmuir oscillation of electrons on the wire surface [21], so that some electrons in the skin layer could be spontaneously extracted into vacuum. For a linearly-polarized laser pulse, the surface electrons are mainly dragged out by negative transverse laser electric field (Ey < 0), forming a series of sawtooth-like spikes as shown in Fig. 2(a). Due to the non-uniform distribution of the laser electric field in the transverse direction, the laser ponderomotive force acting on the attosecond electrons closer to the wire surface is much stronger than that in the margin region. Therefore, these electrons experience both stronger transverse laser electric field and transverse ponderomotive force, making them flee the wire surface and resulting in an arrow structure in space. From the simulation results above, one sees that the longitudinal thickness of each bunch is merely 0.2 μm, which is much less than the wavelength of the laser pulse. This indicates that the duration of each bunch in time is exactly in the attosecond domain, i.e., 660 as at t = 8T0, which is at the same level as in [22, 23]. On the other hand, we find the electron number density in each spike is up to 100nc, with charge of about 5 nC. The obtained peak electron density in our scheme is several orders of magnitude higher than in the gas scenario [19] and at least 3 times larger than those obtained in the NCD plasmas [17].

 figure: Fig. 2

Fig. 2 Density distribution of attosecond electron bunches at t = 8T0 (a). Phase-space distribution of electrons (x, px) at t = 8T0 (b). Evolution of the electron energy spectrum (c). Distribution of the electron divergence angle at t = 8T0 and 12T0 (d).

Download Full Size | PDF

In such a geometry, the electrons on the wire surface will be dragged out with a large transverse momentum. Meanwhile, the J × B force turns the transverse motion of electrons into the direction of laser’s propagation so that the electrons can be further accelerated by the longitudinal ponderomotive force [24]. The laser ponderomotive acceleration or so-called direct laser acceleration is quite efficient for electron acceleration, which is usually dominant in the context of ultra-intense laser-plasma interaction. In this regime, the ponderomotive force is proportional to the laser intensity and reciprocal to the laser duration, i.e., ∼ πa02/2τ0, so that an ultra-short ultra-intense laser pulse is capable of accelerating electrons to energies up to 0.255a02MeV [25]. Thus, in a few laser cycles the electrons can be immediately accelerated to relativistic velocities and therefore remain in the acceleration phase for a long time. However, the compact electron structure will be deteriorated due to the space charge effect as well as the ponderomotive force scattering. Both effects result in an increasing beam dispersion at a later stage. In order to maintain the compact structure meanwhile accelerate the electrons to sufficiently high energy for the γ-ray emission, the electron bunches are expected to collide with the probe pulse as soon as they flee the right wire tip. Finally, the attosecond electron bunches can acquire sufficient energy in a short time with the maximum energy up to 10 GeV, as shown in Fig. 2(c). At t = 9T0, their collisions with the probe pulse take place and the radiation loss of electrons becomes significant after the collisions, e.g., at t = 13T0. Figure 2(d) shows the electron divergence before and after the collisions. We see the attosecond electron bunches have a high collimation before the collisions (∼ 10°), which is very favorable for the γ-ray emission at the following stage.

2.2. Ultra-bright ultra-short γ-ray flashes emission

In the nonlinear QED regime, two significant quantum effects dominate the laser-plasma interaction at ultra-high laser intensities. Firstly, ultra-relativistic electrons quiver violently in the laser fields and concomitantly radiate a large amount of photons in the synchrotronlike way. Alternatively, as the electrons counter-propagate with the laser pulse, they are able to scatter off the incident laser photons to high energy γ-photons in the strong laser fields. When the photon energy in the electron rest frame is comparable to the electron rest mass, multi-photon scattering occurs, which is called multi-photon nonlinear Compton scattering (NCS) [26, 27], i.e., (e + le + ωγ), where ωl and ωγ denote the laser photon and γ-photon, respectively. Secondly, the emitted high energy γ-photons can further decay into ee+ pairs via the B-W process, and therefore generate copious ee+ pairs. Under the quasi-stationarity and weak field approximation [28], the cross-section of these processes can be assumed to be only dependent on the following two crucial gauge-invariant parameters:

η=eme3c4|FμvPv|γeES|E+β×cB|,
χ=e22me3c4|Fμvkv|ωγ2me3c4ES|E+k×cB|,
where ħ is the reduced plank constant, Fμν is the electromagnetic (EM) field tensor, Pν and kν are the four-momentum of electrons and photons, respectively, E is the component of the electric field perpendicular to the electron’s motion and B is the magnetic field, β is the electron’s velocity normalized by the speed of light in vacuum c, and ħk is the γ-photon momentum. In theory, η characterizes the probability of γ-photon emission by an electron with Lorentz factor γe in the EM field, while χ determines the probability of the ee+ pairs production through the interaction of high energy γ-photons with the EM field. The three-vector form of η indicates that the terms E and β × cB in Eq. (1) above are additive once the energetic electrons propagate along the optical axis anti-parallel to the incident laser pulse. This results in η ≈ 2γEL/ES, which promotes the probability of the γ-photon emission. Furthermore, in the context of such a head-on collision, the laser frequency is Doppler up-shifted in the frame of reference of counter-propagating electrons and the transformed field experienced by electron is amplified, which is preferable for the γ-photon emission with a low laser intensity [29].

Figure 3(a) and 3(c) show the photon number density distribution during the γ-ray emission process, which can be divided into two independent stages (I) and (II). In the initial stage (I), the collision between the attosecond electron bunches and the probe pulse has not yet occurred and the emission results from the electron acceleration in the strong laser fields. This is actually the synchrotronlike radiation. It is interesting to see that several ultra-short γ-ray flashes are emitted along the motion direction of the attosecond electron bunches. These denser γ-photons are generated and mainly concentrated at the tail of each electron bunch. By contrast, there are only a few photons emitted on the laser axis. We can interpret it from the electron energy and dynamical motion at this stage. Here, the radiation power is proportional to the fourth power of electron energy, i.e., Pγ=e2cγe4/6πε0ρ2, where ρ is the radius of the trajectory and ε0 is the permittivity of vacuum [30]. For each attosecond electron bunch, the electrons located farther away from the micro-wire surface are extracted much earlier and therefore obtain more energies from the laser pulses as compared to those near the wire surface. Meanwhile, the radius of the trajectory is also much smaller for these high energy electrons. Thus these electrons with higher energy and smaller curvature radius emit more photons. However, near the wire surface the electrons are just extracted by the laser fields and have a small energy so that we do not see any dense photon emission here. On the other hand, these photons have a large divergence angle and fly away from the micro-wire surface as soon as they are emitted by the electrons. Finally we see higher power and brighter γ-ray emission at the tail of each electron bunch, as is seen in Fig. 3(a). Figure 3(b) presents the angular distribution of electron bunches at 9T0. As expected, the photon emission is concentrated in a cone along the direction of ±45° with respect to the wire axis. The corresponding cut-off energy is up to 4 GeV, as shown in Fig. 3(e).

 figure: Fig. 3

Fig. 3 Density distribution of high energy γ-photons at t = 9T0 (a) and 11T0 (c). The red dashed lines mark the micro-wire locations. Angle-resolved photon energy spectrum at t = 9T0 (b) and 11T0 (d). Here the cone angle φ = 12° and θ = 20°. Evolution of the photon energy spectra (e), total photon number, and photon average energy (f). Here (I) and (II) mark different stages of the photon emission during the laser-wire interaction.

Download Full Size | PDF

In the second stage (II), the attosecond electron bunches start leaving the wire tip and subsequently collide with the probe pulse. As a result, a large amount of γ-photons are emitted via the multi-photon NCS process. It is clear to see that the resulting γ-ray flashes have a similar structure to the attosecond electron bunches. The duration of each pulse can be approximately estimated as τγ=τe+τL/4γe2 with τe the electron bunch duration. For ultra-relativistic electrons with γe ≫ 1, the second term above in r.h.s. is negligible and the γ-ray source has almost the same duration of the attosecond bunch. On the other hand, these photons have a high collimation and are located almost at the same positions of the attosecond electron bunches. Note that these electrons are ultra-relativistic here, which can follow the photons at this stage. Figure 3(d) presents the photon angular distribution after the collision. We see these photons propagate along the laser axial with a smaller cone angle of 20° and source size of 0.2 × 2 × 0.8 μm3. Compared to the emission at the stage (I), the photon number increases significantly and the average photon energy is also boosted sharply as seen in Fig. 3(f). For example, at t = 11T0, the photon cut-off energy is as high as 8 GeV and the average photon energy is about 24 MeV with an equivalent total photon number up to 1.7 × 1013. This indicates that nearly 10% of the total laser energy is transferred to the γ-photons.

Since ultra-intense laser pulses are employed in our scheme, the laser-wire target interaction is highly nonlinear and high-order multi-photon scattering occurs. The number of the absorbed laser photons is a random value with its own distribution law, but the first order of scattering has the largest cross-section which occupies the major portion of the scattering spectrum [31]. Finally, we may estimate the energy of the scattered photons on assumption of a head-on collision of electrons and photons by virtue of the formula below [32].

εγ=4εe2ω04εeω0+(mec2)2,
where εe and ħω0 are the electron energy and the laser photon energy, respectively. Thus, GeV γ-photons could be obtained though their number is relatively small. At t = 11T0, the equivalent photon number in each flash is estimated to be 3.5 × 1011 and the peak photon brightness is of the order of 4 × 1024 photons s−1mm−2mrad−2 per 0.1%BW at 24 MeV. This γ-ray flash source with such ultra-short durations is one of the brightest, as compared to ever reported in ELI-NP, SXFEL, etc [33–36].

2.3. High-energy-density attosecond positron bunches production

It is well-known that a plane EM wave of arbitrary intensity and frequency is impossible to create ee+ pairs from vacuum [37]. In our configuration, copious ee+ pairs are generated via the multi-photon B-W process in the interaction zone of the two colliding focused laser pulses. The collision between ultra-relativistic electron bunches and the probe pulse begins at t = 9T0, while the pair creation is gradually enhanced from the collision onward. Here, the electron energy and density determine the final γ-photon emission via the NCS process, while the pair creation is highly dependent on the γ-photon’s distribution. Finally, the electron bunches are able to transfer the feature of the ultra-short duration to first the photons and then positrons, so that we can obtain bright γ-ray flashes and compact attosecond positron bunches, as is seen in Fig. 4(a). Once generated, these positrons are also accelerated by the laser pressure of the driving pulse, like the attosecond electrons, and propagate forward along the laser axial direction. Figures 4(b) and 4(c) present the positron energy spectrum and divergence angle distribution. We see the positron emission angle is relatively small, with a cone angle of 10° at t = 11T0, which is beneficial to the further applications in practice. Meanwhile, the positron bunches have a broad energy spectrum ranging from several MeV to GeV. The corresponding mean energy is up to 338 MeV at t = 11T0, which is even higher than the parent γ-photon’s energy (several tens of MeV) due to the laser acceleration of positrons. The equivalent final positron number in 2D is about 1×1010 and the corresponding laser energy conversion efficiency to the positrons is about 0.07%, which is at least 3 times larger than in the B-H experiments [11] and is much higher than that in the laser wakefield scenario [8, 18]. In theory, we can estimate the final positron number by [17]

Ne+=εe+mec2(αESE)3/4ω0mec2.
where εe+ is the positron energy and α = e2/4πϵ0ħc is the fine-structure constant. The predicted final positron number Ne+=2.7×1010, which is of the same order of magnitude of the simulation result above. Note that the equation is obtained under assumption that the laser pulse is focused to the diffraction limit. Though our employed laser intensity is only one quarter of [13]. we obtained almost the same positron energy and density. Here, each positron bunch is overdense and the maximal number density is up to 1022 cm−3. In the central region of each positron bunch, the energy density is approximately ∼ 1017 J/m3, which is 6 orders of magnitude higher than the threshold for studying high-energy-density physics (1011 J/m3). To the best of our knowledge, it is the first time to report the generation of such attosecond positron bunches with high energy and density, which may have various potential applications in future.

 figure: Fig. 4

Fig. 4 Energy distribution (a) and energy spectra (b) of positrons. Positron divergence angle distribution at t = 9T0, 10T0, and 11T0 (c).

Download Full Size | PDF

3. Robustness of the scheme

At ultra-high laser intensities, nonlinear effects during the QED processes become remarkable and the electron dynamics is strongly modified by the radiation reaction [30, 38]. With the increase of the laser intensity, the cross-section of collisions between electron-photon and between photon-photon increases so that both the number of pairs and the laser energy conversion efficiency increase significantly. Figure 5(a) shows the effective temperature of the attosecond electron bunches against the laser dimensionless parameter a0. In these simulations, we keep all other parameters unchanged but only vary a0. We see the electron temperature tends to be saturated after a0 = 300. This nonlinear phenomenon results from the modification of radiation reaction [20, 39]. At a relatively low laser intensity, the electron acceleration is more important and the radiation loss of electrons is not serious. Once the radiation damping force becomes comparable with the Lorenz force [30], FR/FL=(2/3)k0reγe2βea0~1, the electrons dynamics changes and the electrons act as an energy converter, so that more laser energy is transferred to the γ-photons. Here, k0 = 2π/λ0 and re is the classical electron radius. This corresponds to a threshold a0 ∼ 320, assuming an average relativistic factor γe¯=500, which is in excellent agreement with the simulation result above. Figure 5(b) presents the final yield of secondary particles and the corresponding laser energy conversion efficiency against the parameter a0. As the laser intensity increases, both the production of γ-photons and positrons and the corresponding laser energy conversion increase. Especially, once a0 > 450 the positron production becomes highly efficient.

 figure: Fig. 5

Fig. 5 Effective temperature of attosecond electron bunches against the laser dimensionless parameter a0 (a). Final yield of secondary particles and the corresponding laser energy conversion efficiency against the parameter a0 (b). Here, the total number of photons and positrons, and the conversion efficiencies are all normalized by the corresponding maximal value in the case with a0 = 750. Number density distribution of γ-photons in full 3D simulations at t = 11T0 (c). The xz panel shows the projection of the photon density. Energy distribution of positron bunches in space at t = 11T0 (d). The insets in (c) and (d) show the corresponding energy spectra.

Download Full Size | PDF

Finally, we carry out a 3D simulation to demonstrate the feasibility of the scheme. Here, the box size is x × y × z = 14λL × 12λL × 12λL, sampled by 1400 × 480 × 480 cells and 12 pseudo-electrons and 6 pseudo-carbons in each cell. All parameters are the same as in the 2D cases except that the z direction is considered. Figures 5(c) and 5(d) present the density and energy distribution of γ-photons and positrons, respectively. It is shown that the peak brightness of the γ-ray flashes is up to 1.8 × 1024 photons s−1mm−2mrad−2 per 0.1%BW at 24 MeV. These γ-photons can subsequently decay into high-energy-density positron bunches via the B-W process. The total positron number is about 1 × 109 per shot with a clear spatial attosecond structure. As compared to the 2D simulations above, both the yield and the number density of γ-photons and positrons decrease, since the expansion of the produced attosecond electron bunches along the z direction should be taken into account in the 3D configuration. However, the resulting γ-ray flashes and the attosecond structure of positrons are obvious. At t = 11T0, the positron energy achieved is up to 600 MeV and the photon density is as high as 300nc. This demonstrates that the underlying physics during the laser-wire interaction in the full 3D configuration does not vary. With the progress of projected PW laser facilities in the near future, the resulting ultra-bright ultra-short γ-ray flashes and dense attosecond positron bunches may be tested soon, which might have diverse applications in various fields.

4. Conclusion

In summary, we propose a novel scheme to generate ultra-bright ultra-short γ-ray flashes and high-energy-density attosecond positron bunches by using multi-dimensional QED-PIC simulations. It is shown that the attosecond electron bunches are extracted from the micro-wire by the laser transverse fields and are accelerated forward to ultra-relativistic energies by the laser ponderomotive force. As soon as the attosecond electron bunches flee the wire tip and collide with the probe pulse, ultra-brilliant γ-ray flashes are emitted with ultra-high brightness of 1.8 × 1024 photons s−1mm−2mrad−2 per 0.1%BW at 24 MeV. These γ-photons can subsequently decay into high-energy-density positron bunches via the B-W process. It is shown that the total positron number in full 3D simulations is about 1 × 109 per shot with a clear spatial attosecond structure and over critical density. These unique properties are beneficial for further applications in laboratory astrophysics and high energy physics in the future.

Funding

National Natural Science Foundation of China (NSFC) (11622547, 11675264, 11474360, 11375265); National Basic Research Program of China (2013CBA01504); Science Challenge Project (TZ2016005); Hunan Provincial Natural Science Foundation of China (2017JJ1003); Research Project of NUDT (JQ14-02-02).

Acknowledgments

All simulations were performed on TianHe High performance Computers at NUDT. The EPOCH code was developed as part of the UK EPSRC grants EP/G054940/1. T.P.Y acknowledges the scholarship from China Scholarship Council (201506115018).

References and links

1. See more information on the website, http://www.eli-np.ro/.

2. X. Xie, J. Zhu, Q. Yang, J. Kang, H. Zhu, M. Sun, and A. Guo, “Introduction to SG-II 5 PW Laser Facility,“ in Conference on Lasers and Electro-Optics, OSA Technical Digest (2016) (Optical Society of America, 2016), paper SM1M.7.

3. E. S. Fradkin and S. M. Shvartsman, Quantum Electrodynamics: With Unstable Vacuum (Springer, 2011).

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

5. S. V. Bulanov, T. Zh. Esirkepov, M. Kando, J. Koga, K. Kondo, and G. Korn, “On the problems of relativistic laboratory astrophysics and fundamental physics with super powerful lasers,” Plasma Phys. Rep. 41(1), 1–51 (2015). [CrossRef]  

6. H. Bethe and W. Heitler, “On the stopping of Fast Particles and on the Creation of Positive Elctrons,” Pro. R. Soc. Lond. A 146, 83112 (1934).

7. G. Breit and J. A. Wheeler, “Collision of two light quanta,” Phys. Rev. 46(12), 1087–1091 (1934). [CrossRef]  

8. T. G. Blackburn, C. P. Ridgers, J. G. Kirk, and A. R. Bell, “Quantum radiation reaction in laser-electron-beam collisions,” Phys. Rev. Lett. 112, 015001 (2014). [CrossRef]   [PubMed]  

9. K. Nakashima and H. Takabe, “Numerical study of pair creation by ulraintense lasers,” Phys. Plasmas 9(5), 1505–1512 (2002). [CrossRef]  

10. R. Ruffini, G. Vereshchagin, and S. S. Xue, “Electron-positron pairs in physics and astrophysics: from heavy nuclei to black holes,” Phys. Rep. 487(1–4), 1–140 (2010). [CrossRef]  

11. H. Chen, S. C. Wilks, D. D. Meyerhofer, J. Bonlie, C. D. Chen, S. N. Chen, C. Courtois, L. Elberson, G. Gregori, and W. Kruer, “Relativistic quasimonoenergetic positron jets from intense laser-solid interactions,” Phys. Rev. Lett. 105(1), 015003 (2010). [CrossRef]   [PubMed]  

12. 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(16), 165006 (2012). [CrossRef]   [PubMed]  

13. 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 counter-propagating quantum electrodynamics-strong laser interaction with solid targets,” Phys. Plasmas 22(6), 063112 (2015). [CrossRef]  

14. H. X. Chang, B. Qiao, Z. Xu, X. R. Xu, C. T. Zhou, X. Q. Yan, S. Z. Wu, M. Borghesi, M. Zepf, and X. T. He, “Generation of overdense and high-energy electron-positron-pair plasmas by irradiation of a thin foil with two ultraintense lasers,” Phys. Rev. E 92(5), 053107 (2015). [CrossRef]  

15. J. J. Liu, T. P. Yu, Y. Yin, X. L. Zhu, and F. Q. Shao, “All-optical bright γ-rays and dense positron source by laser driven plasmas-filled cone,” Opt. Express 24(14), 15978–15986 (2016). [CrossRef]   [PubMed]  

16. X. L. Zhu, Y. Yin, T. P. Yu, F. Q. Shao, Z. Y. Ge, W. Q. Wang, and J. J. Liu, “Enhanced electron trapping and γ ray emission by ultra-intense laser irradiating a near-critical-density plasma filled gold cone,” New J. Phys. 17(5), 053039 (2015). [CrossRef]  

17. X. L. Zhu, T. P. Yu, Z. M. Sheng, Y. Yin, I. C. E. Turcu, and A. Pukhov, “Dense GeV electron-positron pairs generated by lasers in near-critical-density plasmas,” Nat. Commun. 7, 13686 (2016). [CrossRef]   [PubMed]  

18. M. Lobet, X. Mathieu, E. D’Humières, and L. Gremillet, “Generation of high-energy electron-positron beams in the collision of a laser-accelerated electron beam and a multi-petawatt laser,” Phys. Rev. ST Accel. Beams 20, 043401 (2017). [CrossRef]  

19. J. X. Liu, Y. Y. Ma, T. P. Yu, J. Zhao, X. H. Yang, L. F. Gan, G. B. Zhang, Y. Zhao, S. J. Zhang, J. J. Liu, H. B. Zhuo, F. Q. Shao, and S. Kawata, “Enhanced electron-positron pair production by ultra intense laser irradiating a compound target,” Plasma Phys. Control. Fusion 58, 125007 (2016). [CrossRef]  

20. C. P. Ridgers, J. G. Kirk, R. Duclous, T. G. Blackburn, C. S. Brady, K. Bennett, T. D. Arber, and A. R. Bell, “Modelling gamma-ray photon emission and pair production in high-intensity laser-matter interactions,” J. Comput. Phys. 260, 273–285 (2014). [CrossRef]  

21. Y. Y. Ma, Z. M. Sheng, Y. T. Li, W. W. Chang, X. H. Yuan, M. Chen, H. C. Wu, J. Zheng, and J. Zhang, “Dense quasi-monoenergetic attosecond electron bunches from laser interaction with wire and slice targets,” Phys. Plasmas 13, 110702 (2006). [CrossRef]  

22. N. Naumova, I. Sokolov, J. Nees, A. Maksimchuk, V. Yanovsky, and G. Mourou, “Attosecond electron bunches,” Phys. Rev. Lett. 93(19), 195003 (2004). [CrossRef]   [PubMed]  

23. L. X. Hu, T. P. Yu, F. Q. Shao, W. Luo, and Y. Yin, “A bright attosecond x-ray pulse train generation in a double-laser-driven cone target,” J. Appl. Phys. 119, 509–513 (2016). [CrossRef]  

24. Q. Kong, S. Kawata, S. Miyauchi, K. Sakai, Y. K. Ho, and A. A. Andereev, “Electron bunch trapping and compression by an intense focused pulse laser,” Phys. Rev. E 69(5), 056502 (2004). [CrossRef]  

25. W. M. Wang, Z. M. Sheng, Y. T. Li, L. M. Chen, S. Kawata, and J. Zhang, “Monoenergetic electron bunches generated from thin solid foils irradiated by ultrashort, ultraintense circularly polarized lasers,” Phys. Rev. ST Accel. Beams 13, 071301 (2010). [CrossRef]  

26. F. Ehlotzky, K. Krajewska, and J. Z. Kamiski, “Fundamental processes of quantum electrodynamics in laser fields of relativistic power,” Rep. Prog. Phys. 72, 046401 (2009). [CrossRef]  

27. K. Krajewska and J. Z. Kamiski, “Compton process in intense short laser pulses,” Phys. Rev. A 85, 13582–13591 (2012). [CrossRef]  

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

29. I. V. Sokolov, J. A. Nees, V. P. Yanovsky, N. M. Naumova, and G. A. Mourou, “Emission and its back-reaction accompanying electron motion in relativistically string and QED-strong pulsed laser field,” Phys. Rev. E. Stat. Nonlin. Soft. Matter. Phys. 81(2), 036412 (2010). [CrossRef]  

30. J. D. Jackson, Classical Electrodynamics, 3rd ed. (John Wiley & Sons, 2007), Chap. 14.

31. C. Bula, K. T. Mcdonald, E. J. Prebys, C. Bamber, S. Boege, T. Kotseroglou, and C. R. Field, “Observation of nonlinear effects in Compton scattering,” Phys. Rev. Lett. 76(17), 3116 (1996). [CrossRef]   [PubMed]  

32. C. Sun and Y. K. Wu, “Theoretical and simulation studies of characteristics of a Compton light source,” Phys. Rev. ST Accel. Beams 14(4), 044701 (2011). [CrossRef]  

33. K. Ta Phuoc, S. Corde, C. Thaury, V. Malka, A. Tafzi, J. P. Goddet, R. C. Shah, S. Sebban, and A. Rousse, “All-optical compton gamma-ray source,” Nat. Photonics 6, 308–311 (2013). [CrossRef]  

34. G. Sarri, D. J. Corvan, W. Schumaker, J. M. Cole, A. Di Piazza, and M. Zepf, “Ultrahigh brilliance multi-MeV γ-ray beams from nonlinear relativistic Thomson scattering,” Phys. Rev. Lett. 113, 224801 (2014). [CrossRef]  

35. C. A. Ur, “Gamma beam system at ELI-NP,” AIP Conference Proceedings 1645, 237 (2015). [CrossRef]  

36. H. L. Wu, J. H. Chen, B. Liu, D. Wang, and Z. T. Zhao, “Proposal for a high brightness γ-ray source at SXFEL,” Nucl. Sci. Tech. 26, 050103 (2015).

37. V. B. Berestekii, E. M. Lifshitz, and L. P. Pitaevskii, Quantum Electrodynamics, (Pergamon Press, 1982).

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

39. T. P. Yu, Z. M. Sheng, Y. Yin, H. B. Zhuo, Y. Y. Ma, F. Q. Shao, and A. Pukhov, “Dynamics of laser mass-limited foil interaction at ultra-high laser intensities,” Phys. Plasmas 21, 053105 (2014). [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 (5)

Fig. 1
Fig. 1 Schematic view of ultra-bright γ-ray flashes emission and attosecond ee+ bunches generation by two ultra-intense colliding lasers irradiating a micro-wire target.
Fig. 2
Fig. 2 Density distribution of attosecond electron bunches at t = 8T0 (a). Phase-space distribution of electrons (x, px) at t = 8T0 (b). Evolution of the electron energy spectrum (c). Distribution of the electron divergence angle at t = 8T0 and 12T0 (d).
Fig. 3
Fig. 3 Density distribution of high energy γ-photons at t = 9T0 (a) and 11T0 (c). The red dashed lines mark the micro-wire locations. Angle-resolved photon energy spectrum at t = 9T0 (b) and 11T0 (d). Here the cone angle φ = 12° and θ = 20°. Evolution of the photon energy spectra (e), total photon number, and photon average energy (f). Here (I) and (II) mark different stages of the photon emission during the laser-wire interaction.
Fig. 4
Fig. 4 Energy distribution (a) and energy spectra (b) of positrons. Positron divergence angle distribution at t = 9T0, 10T0, and 11T0 (c).
Fig. 5
Fig. 5 Effective temperature of attosecond electron bunches against the laser dimensionless parameter a0 (a). Final yield of secondary particles and the corresponding laser energy conversion efficiency against the parameter a0 (b). Here, the total number of photons and positrons, and the conversion efficiencies are all normalized by the corresponding maximal value in the case with a0 = 750. Number density distribution of γ-photons in full 3D simulations at t = 11T0 (c). The xz panel shows the projection of the photon density. Energy distribution of positron bunches in space at t = 11T0 (d). The insets in (c) and (d) show the corresponding energy spectra.

Equations (4)

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

η = e m e 3 c 4 | F μ v P v | γ e E S | E + β × c B | ,
χ = e 2 2 m e 3 c 4 | F μ v k v | ω γ 2 m e 3 c 4 E S | E + k × c B | ,
ε γ = 4 ε e 2 ω 0 4 ε e ω 0 + ( m e c 2 ) 2 ,
N e + = ε e + m e c 2 ( α E S E ) 3 / 4 ω 0 m e c 2 .
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.