<<

Tailoring -generated plasmas for efficient nuclear excitation by electron capture

Yuanbin Wu,1, ∗ Jonas Gunst,1, † Christoph H. Keitel,1 and Adriana P´alffy1, ‡ 1Max-Planck-Institut f¨urKernphysik, Saupfercheckweg 1, D-69117 Heidelberg, Germany (Dated: February 7, 2018) The optimal parameters for nuclear excitation by electron capture in environments gener- ated by the interaction of ultra-strong optical with solid matter are investigated theoretically. As a case study we consider a 4.85 keV nuclear transition starting from the long-lived 93mMo iso- mer that can lead to the release of the stored 2.4 MeV excitation energy. We find that due to the complex plasma dynamics, the nuclear excitation rate and the actual number of excited nuclei do not reach their maximum at the same laser parameters. The nuclear excitation achievable with a high-power optical laser is up to twelve and up to six orders of magnitude larger than the values predicted for direct resonant and secondary plasma-mediated excitation at the x-ray free electron laser, respectively. Our results show that the experimental observation of the nuclear excitation of 93mMo and the subsequent release of stored energy should be possible at laser facilities available today.

Novel coherent light sources open unprecedented pos- that the maximal number of depleted isomers for realis- sibilities for the field of laser-matter interactions [1]. tic laser setup parameters may reach for the first time The X-ray Free Electron Laser (XFEL) [2, 3] for in- measurable values. Although still far from the final goal, stance can drive low-energy electromagnetic transitions this is a further milestone on the way to the realization of in nuclei. Ultra-strong optical laser systems with up to controlled energy storage and release via nuclear isomers. few petawatt power [4–8] are very efficient in generat- We consider a strong optical laser that interacts with ing plasma environments [9], which host complex inter- a solid-state target containing a fraction of nuclei in the actions between photons, electrons, ions and the atomic isomeric state. NEEC and photoexcitation may occur in nucleus. Nuclear excitation in laser-generated hot plas- the generated plasma. In the resonant process of NEEC, mas involving optical lasers [10–26], or cold high-density a free electron recombines into a vacant bound atomic plasmas [27] at the XFEL [28, 29] have been under inves- state with the simultaneous excitation of the nucleus. tigation. Special attention has been attracted by nuclear The isomers can then be excited to a trigger state which transitions starting from long-lived excited states. Such rapidly decays to the nuclear ground state and releases states are also known as nuclear isomers and are par- the stored energy. We consider in the following the case ticularly interesting due to their potential to store large of 93mMo for which recent claims have been made [40] on amounts of energy over long periods of time [30–37]. A the first observation of NEEC following the proposals in typical example is 93mMo at 2.4 MeV, for which an ad- Refs. [41, 42]. ditional excitation of only 4.85 keV could lead to the de- Free electrons in the plasma cover a broad energy pletion of the isomer and release on demand of the stored range such that many NEEC resonance channels may energy. contribute to the net NEEC rate λneec. This can be ex- For both optical- and x-ray laser-generated plasmas, pressed as the convolution over the electron energy E of the process of nuclear excitation by electron capture the almost Dirac-delta-like NEEC single-resonance cross (NEEC) [38, 39] into the atomic shell has proven to have section σneec and the free-electron flux φe, summed over a significant contribution. As secondary process in the all charge states q and capture channels αd, cold plasma environment generated by the interaction X Z of the XFEL with solid-state targets, NEEC can exceed λneec(Te, ne) = Pq(Te, ne) dE σneec(E)φe(E,Te, ne).

the direct nuclear photoexcitation by six orders of mag- q,αd nitude [28, 29] for the 4.85 keV excitation starting from (1) 93m the Mo isomeric state. In this Letter, we show that by Here, Pq is the probability to find ions charge state arXiv:1708.04826v2 [physics.plasm-ph] 6 Feb 2018 tailoring optical-laser-generated plasmas to harness max- q in the plasma as a function of electron temperature imum nuclear excitation via NEEC, a further six orders Te and density ne. The dependence φe(Te) determines of magnitude increase in the nuclear excitation and sub- the quantitative contribution of the NEEC resonances. sequent isomer depletion compared to the case of cold The theoretical formalism for the calculation of the XFEL-generated plasmas can be reached. As an inter- NEEC cross section σneec has been presented elsewhere esting point, we find that due to the complexity of the [28, 29, 43, 44]. The total NEEC excitation number Nexc processes involved, the plasma and correspondingly laser is connected to the rate λneec via parameters for reaching the maximal NEEC rate are not Z Z identical to the ones that provide the maximal number 3 Nexc = d r dt niso(r, t) λneec(Te, ne; r, t), (2) of nuclei actually excited. Our calculations demonstrate Vp 2 where niso denotes the number density of isomers and 23 3 24 neec 5.0 ne = 10 cm 10 6.0 L Vp is the plasma volume. Let us assume in a first ap- neec M neec

proximation homogeneous plasma conditions over the ] 23 Nexc 2.5 3.0 N 10 3 ] E plasma lifetime τ . Then the total number of excited 1

p E m s 1e 2 1e 4 C c [ 22 0.0 0.0 nuclei is N = N λ (T , n )τ , with N the num- [ 10

exc iso neec e e p iso c 2 21 3 4 e

e n = 10 cm ×10 e 10 × y e

2.4 1.6 x t i n ber of isomers in the plasma. Assuming a spherical c s i

t

n 21 a

e 10 plasma the plasma lifetime is approximatively given by e t

1.2 0.8 t d i a

p o

¯ r n

τp = Rp mi/(TeZ) [29, 45] with the ion mass mi, the n

o 20

C 1e 4 1e 8

N 10 ¯ 0.0 0.0 r t average charge state Z and the plasma radius R . N E p iso 4 19 3 8 e ne = 10 cm c x E ×10 10 × c 1.6 e l

N 1.2 can be estimated introducing the isomer fraction embed- 19 E 10 ded in the original solid-state target f , N = f n V , iso iso iso i p 0.8 0.6 where ni stands for the ion number density in the plasma. 1018 93m −5 0.0 0.0 A Mo isomer fraction of fiso ≈ 10 embedded in 0 1 2 3 4 5 6 7 0 1 2 3 4 5 6 7 solid-state Niobium foils can be generated by intense Temperature [keV] Tmax [keV] 14 93 93m (≥ 10 protons/s) beams [28] via the 41Nb(p,n) 42Mo reaction [46]. FIG. 1. (color online). Left graph: NEEC rate λneec (blue, Numerical results for λneec and the corresponding total solid curve) and the total number of excited isomers Nexc (red, number of excited isomers Nexc for an arbitrary plasma dash-dotted curve), as well as the individual contributions L M radius of 40 µm are presented in Fig. 1. For the cal- λneec (orange, dashed curve) and λneec (green, dash-dotted culation of σneec we use a theoretical prediction for the curve) from the L and M shell, respectively, as a function of reduced nuclear transition probability [47]. We model the electron temperature Te for selected electron densities ne. the plasma conditions by a relativistic distribution for A plasma radius of 40 µm has been assumed in the calcula- tions of Nexc. Right graph: Temperatures Tmax as functions the free electrons and a charge state distribution com- L M of density, for maximizing Nexc, λneec, λneec and λneec, respec- puted with the radiative-collisional code FLYCHK [48]. tively, at each particular ne. The relativistic electronic wave functions [49] and bind- ing energies are in first approximation calculated inde- pendently of Te and ne, which are accounted for only behavior of λneec and Nexc becomes more involved as the indirectly via the charge state distribution. For a spe- charge distribution Pq shows a complex dependence on 21 cific charge state, we further assume that the ion is in its the plasma conditions ne and Te. Between ne = 10 −3 23 −3 ground state and recombination of the NEEC electron cm and 10 cm we see that with increasing ne the occurs in a free orbital. Among these assumptions, ne- atomic shell contributions change significantly and λneec glecting the plasma-induced ionization potential depres- is much enhanced. sion [50] is the most severe appoximation, as binding en- The temperature Tmax at which Nexc or the total or ergies may vary by few eV to hundreds of eV depending partial shell contributions λneec reach a maximum for on the plasma density. Using the Steward-Pyatt model each density value ne is depicted in the right graph of [50] which also has reasonable agreement with more re- Fig. 1. Naively, one would expect that Tmax is approxi- cently developed methods [51, 52] for our purpose, we mately the same for Nexc and for λneec. This is however 21 −3 estimate that the consequences for Nexc are even for the only true at high densities starting from 10 cm . Ac- case of high-density plasmas with large ionization poten- cording to our approximation for τp, the chosen plasma tial depression only on the level of 10%. lifetime is Te-dependent. In particular at low electron NEEC into the K shell is energetically forbidden for densities, τp acts as a weighting function proportional −1/2 the 4.85 keV transition in Mo. The results for the dom- to (Te) shifting the maximum of Nexc to lower tem- inant recombination channels into the L and M atomic peratures. The optimal plasma conditions for the total shells are presented individually in Fig. 1. For the total excitation number can thus drastically differ from the NEEC rate λneec, further smaller contributions from the optimal conditions for λneec in this model. We note that recombination into the N and O shells were also taken the arbitrary choice of Rp only influences the absolute into account. Both λneec and Nexc increase with increas- scale of the NEEC excitation number, not the position 19 −3 ing electron density ne. In the range ne = 10 cm to of Tmax. 1020 cm−3, our calculations show that the charge state A comparison with nuclear photoexcitation assuming a distribution Pq is nearly unaffected for a fixed tempera- black-body radiation spectrum at the given plasma tem- 21 −3 ture Te, while λneec is enhanced by a factor of 10 main- perature Te shows that at ne = 10 cm NEEC domi- 22 taining the same functional dependence on Te. This indi- nates for Te < 1.6 keV and for higher densities ne = 10 −3 cates that at low densities the boost in λneec is (almost) cm up to a temperature of 5 keV. The actual photoex- a pure density effect coming from the increasing number citation in the plasma should be even lower in particular of free electrons present in the plasma (φe ∝ ne). In- at low densities because photons may easier escape the 23 creasing the electron density to even higher values, the finite plasma volume. For the high density ne ≥ 10 3

40 ELI-beamlines PETAL LULI VULCAN 1

Z 30 10 3 N Epulse [J] 1500 3500 100 500

] 10 1 E

20 E τ [fs] 150 5000 1000 500

s pulse C [ 15 16

10 10 2 c 2

I [W/cm ] 10 e λ [nm] 1053 1053 1053 1053 e e 4 x n 10 c −2 −2 −2 i

N 2.4 × 10 1.9 1.1 × 10 2.7 × 10

t exc a e t 3 t i a 10 o r

n

C 5 TABLE I. Laser parameters and maximal Nexc achieved at 10 N E 15 2 e x E Exc. Nexc 4 the optimal laser intensity Iopt = 3.5 × 10 W/cm for ELI- 10 c N = 100 fs beamlines L4 [4], PETAL [5, 6], LULI [7] and VULCAN [8] NEEC rate = 500 fs lasers. 10 6 neec = 1000 fs 10 5 1015 1016 Laser intensity [W/cm2] the spherical plasma model with the lifetime τp. We use the smallest length scale out of Rfocal and dp to calcu- FIG. 2. (color online). The NEEC rate λneec and the total late τp for a lower-limit estimate of the NEEC excitation. excitation number Nexc per laser pulse as functions of laser in- Numerical results for λ and for the total excitation tensity. The inset shows the average charge state Z¯ calculated neec with the help of FLYCHK. See text for further explanations. number Nexc per laser pulse are presented in Fig. 2 as a function of the laser intensity. We consider a pulse energy of 100 J, wavelength of 1053 nm, and laser pulse duration values of 100, 500 and 1000 fs, respectively. Also here the −3 cm parameter regime, NEEC is the dominant nuclear optimal laser intensities Iopt at which λneec and respec- excitation mechanism. tively Nexc are maximal do not coincide. For the assumed 16 In the following we proceed to determine how the opti- laser parameters, λneec is maximized by Iopt = 1.3 × 10 mal NEEC parameter region in the temperature-density W/cm2 at a temperature of 5.1 keV and a density of landscape may be accessed by a short laser pulse. We 5.9 × 1019 cm−3. In contrast, the optimal intensity for 15 2 discern in our treatment two cases, namely the low- and Nexc per laser pulse is 3.5 × 10 W/cm independent of high-density plasmas, and refine accordingly our plasma the laser pulse duration in the range of the considered model. Firstly, we consider the case of a low density values. The electron temperature and density achieved (underdense) plasma, which can be generated via the in- at this intensity are 1.4 keV and 6.6×1019 cm−3, respec- teraction of a strong optical laser with a thin target. The tively, leading to a charge state distribution with Z¯ ∼ 30 plasma generation process typically evolves in two steps (see inset of Fig. 2) where capture channels into the M [53]: (i) a preplasma is formed by the prepulse of the shell still exceed the L-shell contribution. For dp < Rfocal laser; (ii) this preplasma is subsequently heated by the (the case for the parameters of Fig. 2) the plasma life- main laser pulse potentially up to keV electron energies. time is determined by dp and in turn by τpulse. The We model the plasma following the approach in NEEC excitation becomes stronger with increasing laser Refs. [53, 54]. With the help of the so-called scaling law, pulse duration τpulse reaching its maximum at the value 2 the electron temperature is given as Te ≈ 3.6I16λµ keV, where dp = Rfocal. For even longer pulse durations we 16 2 where I16 is the laser intensity in units of 10 W/cm need to use Rfocal in our model to determine the plasma and λµ the wavelength in microns [55–57]. The electron lifetime and this leads to a decrease of λneec. density can be estimated as ne = Ne/Vp where Ne is the In Table I we evaluate the optimal laser intensity Iopt total number of electrons and the plasma volume is given and the expected maximal NEEC excitation Nexc for re- 2 by Vp = πRfocaldp, with Rfocal the focal radius of the alistic parameters of high-power optical lasers which are laser and the plasma thickness dp = cτpulse determined currently available or under construction. The excita- by the speed of light c and the laser pulse duration τpulse. tion Nexc per laser pulse is up to six orders of magnitude The electron number can be related to the absorbed laser larger than the one [∼ 10−6, recalculated for the parame- energy fEpulse via Ne = fEpulse/Te. Since experimen- ters considered here] in the XFEL-generated cold (T =350 tal results in Refs. [58, 59] show that the laser absorp- eV) plasma [28, 29]. The largest value of 1.9 excitations tion is almost independent of the target material and per pulse should be reached with the PETAL laser which thickness, we adopt an universal absorption coefficient provides both high laser power and long pulse duration. f = f(I, λ) which is a cubic interpolation to theoretical We now turn to the case of high electron densities, results based on a Vlasov-Fokker-Planck code presented which promises the strongest nuclear excitation accord- in Ref. [58]. For the considered intensity range between ing to Fig. 1. Experiments and simulations have shown 1015 and 2 × 1016 W/cm2, the absorption fraction f lies that it is possible to isochorically heat targets at solid- between 0.1 and 0.2. state density to temperatures of a few hundred eV or For the case of focal radius, plasma thickness and even a few keV [60–62]. Since in this regime the heating plasma radius of similar scale, we may again consider of the target is mainly conducted by secondary parti- 4

2.0 ps 3.0 ps 4.0 ps direction over the region of Afocal. We consider a laser 103 pulse energy of 100 J, which leads for the pulse duration ] 2 1 10

s and laser intensity adopted in the PIC simulation to a

[ 1 10 −4 2 c

e focal spot area of approximatively 2 × 10 cm . Re- e Regression n 100 PIC sults for λneec are presented in the upper panel of Fig. 3. 10 1 The rate is maximized at depths x with optimal plasma

] 12 conditions for NEEC. The peak propagates through the V

e 8 k target and disappears at around 4 ps as target heating [

e 4

T leads afterwards to temperatures exceeding the optimal 0 25 ] 10 value. The analysis of the data sampled from 1 to 4 ps in 3 100-fs steps shows that the integrated NEEC rate reaches m 1024 c [

its maximum at 3.1 ps and drops roughly to half its value e 23 n 10 at 4 ps. 0 500 1000 0 500 1000 0 500 1000 x [nm] x [nm] x [nm] Using the regression curves for λneec calculated with the fitted ne and Te functions, we solve Eq. (2) in a two- FIG. 3. (color online). Electron density, temperature and step procedure to obtain Nexc. First, for each time in- the NEEC rate based on the PIC simulation as functions of stant t the product of NEEC rate and isomer density is target depth x. The laser has peak intensity I = 1018 W/cm2 integrated with respect to x over the whole target thick- and λ = 800 nm wavelength. The raw data averaged over 10 ness dt and multiplied by the focal spot area Afocal to nm intervals is presented together with a linear polynomial account for the perpendicular directions. Second, the and a third order exponential fit for ne and Te, respectively. outcomes of the spatial integration are interpolated as a Regression curves for λneec calculated with the fitted ne and function of time leading to Nexc(t) which is then inserted Te functions are shown in the upper graphs. into the time integral in Eq. (2). For t > 4 ps, we extrapo- late Nexc(t) assuming an exponential functional behavior initially following the slope at 4 ps. The time integration cles, i.e., hot electrons generated in the laser-target in- converges approximatively after 10 ps, leading to an exci- teraction, a more sophisticated model is necessary com- tation number of 1.8 isomers per pulse via NEEC which pared to the low-density case. We have performed a one- is almost identical with the best value at low densities dimensional (1D) particle-in-cell (PIC) simulation of a obtained with the PETAL parameters. With laser repe- Nb solid target with 1 µm thickness and Nb density of tition rates of few Hz for 100 J pulses, the threshold of 22 −3 nnb = 5.5×10 cm interacting with a high-power laser one isomer depletion per second should be reached pro- −5 using the EPOCH code [63]. The isomer fraction of 10 viding a detectable signal. The experimental signature is small enough to be neglected here in the determination of the nuclear excitation would be a gamma-ray photon of the plasma conditions. The laser has a Gaussian pro- of approx. 1 MeV released in the decay cascade of the 18 2 file in time with peak intensity I = 10 W/cm , laser triggering level in 93Mo. An evaluation of the plasma duration τpulse = 500 fs, and laser wavelength λ = 800 black-body and bremsstrahlung radiation spectra at this nm, respectively. A linear preplasma with the thickness photon energy shows that the signal-to-background ratio of 0.5 µm is considered in front of the solid target. Ion- is very high. Notable here is that in the high-density case ization is not included explicitly in the simulation; as a a 100 J laser available at many facilities around the world representative order of the electron density, we fix the is competitive with a kJ-laser facility. charge state to 10. Tailoring the plasma conditions for NEEC promises To include the effect of atomic ionization and recom- a 12 orders of magnitude increase of the 93mMo deple- bination events, we averaged the raw data for electron tion compared to the direct driving of the nuclear tran- temperature Te and ion density ni from the PIC sim- sition with an XFEL laser. An experimental proof of ulation over 10 nm intervals, and used these values as this scenario appears to be possible with present high- input for the radiative-collisional model implemented in power optical lasers. The PIC simulation has been car- FLYCHK [48] to obtain charge state distributions and ried out in the direction with the smallest length scale (corrected) electron densities. The electron density and of the plasma. Modeling the expansion in the perpen- temperature values are shown in the lower and middle dicular direction of the laser incidence, a roughly 10 to panels of Fig. 3 for the time instants 2, 3 and 4 ps as a 100-times longer plasma lifetime can be expected to boost function of the target penetration depth x. Nexc. A further enhancement can be achieved by employ- For the high-density region, we evaluate the NEEC ing a combination of optical and x-ray lasers as envisaged rate as a function of target depth x and time t by insert- for instance at HIBEF [64] at the European XFEL [65]. ing the PIC-simulation results for Te and the corrected X-rays-generated inner shell holes could then provide the ne values into Eq. (1). The plasma is assumed to be optimal capture state independently from the hot plasma homogeneous only in the plane perpendicular to the x conditions. We note however that further substantial im- 5 provements are required for practical energy storage ap- (2008). plications. In our calculation, only a 10−10 fraction of [21] G. Mourou and T. Tajima, Science 331, 41 (2011). the isomers in the plasma volume are depleted. In ad- [22] A. V. Andreev, R. V. Volkov, V. M. Gordienko, A. M. dition, the total isomer energy stored in the microscopic Dykhne, M. P. Kalashnikov, P. M. Mikheev, P. V. Nikles, A. B. Savel’ev, E. V. Tkalya, R. A. Chalykh, et al., Jour- plasma volume is still far from typical requirements of nal of Experimental and Theoretical Physics 91, 1163 macroscopic every-day life applications. (2000). [23] A. V. Andreev, V. M. Gordienko, and A. B. Savel’ev, Quantum Electronics 31, 941 (2001). [24] C. Granja, J. Kuba, A. Haiduk, and O. Renner, Nuclear Physics A 784, 1 (2007). ∗ [email protected] [25] O. Renner, L. Juha, J. Krasa, E. Krousky, M. Pfeifer, † [email protected] A. Velyhan, C. Granja, J. Jakubek, V. Linhart, T. Slav- ‡ Palff[email protected] icek, et al., Laser and Particle Beams 26, 249 (2008). [1] A. Di Piazza, C. M¨uller,K. Z. Hatsagortsyan, and C. H. [26] F. Gobet, C. Plaisir, F. Hannachi, M. Tarisien, T. Bon- Keitel, Rev. Mod. Phys. 84, 1177 (2012). net, M. Versteegen, M. Al´eonard,G. Gosselin, V. M´eot, [2] Linac Coherent Light Source - LCLS (2017). and P. Morel, Nuclear Instruments and Methods in [3] XFEL @ SACLA, Official Website (2017), Physics Research Section A: Accelerators, Spectrometers, http://xfel.riken.jp/eng/sacla/. Detectors and Associated Equipment 653, 80 (2011). [4] ELI-beamlines L4 beam line webpage (2017), [27] S. M. Vinko, O. Ciricosta, B. I. Cho, K. Engelhorn, H.-K. https://www.eli-beams.eu/en/facility/lasers/l4-10- Chung, C. R. D. Brown, T. Burian, J. Chalupsky, R. W. pw-kj-beamline/. Falcone, C. Graves, et al., Nature 482, 59 (2012). [5] Petawatt Aquitaine Laser – PETAL webpage (2017), [28] J. Gunst, Y. A. Litvinov, C. H. Keitel, and A. P´alffy, http://petal.aquitaine.fr/spip.php?lang=en. Phys. Rev. Lett. 112, 082501 (2014). [6] A. Casner, T. Caillaud, S. Darbon, A. Duval, I. Thfouin, [29] J. Gunst, Y. Wu, N. Kumar, C. H. Keitel, and A. P´alffy, J. P. Jadaud, J. P. LeBreton, C. Reverdin, B. Rosse, Physics of Plasmas 22, 112706 (2015). R. Rosch, et al., High Energy Density Phys. 17, Part [30] P. Walker and G. Dracoulis, Nature 399, 35 (1999). A, 2 (2015). [31] A. Aprahamian and Y. Sun, Nat. Phys. 1, 81 (2005). [7] LULI2000 laser system webpage (2017), [32] D. Belic, C. Arlandini, J. Besserer, J. de Boer, J. J. Car- https://portail.polytechnique.edu/luli/en/facilities roll, J. Enders, T. Hartmann, F. K¨appeler, H. Kaiser, //luli2000-laser-system. U. Kneissl, et al., Phys. Rev. Lett. 83, 5242 (1999). [8] Vulcan laser webpage (2017), [33] C. B. Collins, F. Davanloo, M. C. Iosif, R. Dussart, J. M. https://www.clf.stfc.ac.uk/Pages/Vulcan-laser.aspx. Hicks, S. A. Karamian, C. A. Ur, I. I. Popescu, V. I. [9] P. Mulser and D. Bauer, High power laser-matter inter- Kirischuk, J. J. Carroll, et al., Phys. Rev. Lett. 82, 695 action, vol. 238 (Springer, Berlin, Heidelberg, 2010). (1999). [10] M. R. Harston and J. F. Chemin, Phys. Rev. C 59, 2462 [34] D. Belic, C. Arlandini, J. Besserer, J. de Boer, J. J. Car- (1999). roll, J. Enders, T. Hartmann, F. K¨appeler, H. Kaiser, [11] G. Gosselin and P. Morel, Phys. Rev. C 70, 064603 U. Kneissl, et al., Phys. Rev. C 65, 035801 (2002). (2004). [35] J. J. Carroll, Laser Phys. Lett. p. 275 (2004). [12] G. Gosselin, V. M´eot,and P. Morel, Phys. Rev. C 76, [36] A. P´alffy, J. Evers, and C. H. Keitel, Phys. Rev. Lett. 044611 (2007). 99, 172502 (2007). [13] P. Morel, V. M´eot, G. Gosselin, D. Gogny, and [37] G. D. Dracoulis, P. M. Walker, and F. G. Kondev, Re- W. Younes, Phys. Rev. A 69, 063414 (2004). ports on Progress in Physics 79, 076301 (2016). [14] V. M´eot,J. Aupiais, P. Morel, G. Gosselin, F. Gobet, [38] V. I. Goldanskii and V. A. Namiot, Phys. Lett. B 62, 393 J. N. Scheurer, and M. Tarisien, Phys. Rev. C 75, 064306 (1976). (2007). [39] A. P´alffy, Contemporary Physics 51, 471 (2010). [15] P. Morel, V. M´eot, G. Gosselin, G. Faussurier, and [40] First experimental evidence of NEEC, talk presented by C. Blancard, Phys. Rev. C 81, 034609 (2010). C. Chiara at the Topical Meeting on Isotope-Based En- [16] M. Comet, G. Gosselin, V. M´eot,P. Morel, J.-C. Pain, ergy Sources (The George Washington University, Wash- D. Denis-Petit, F. Gobet, F. Hannachi, M. Tarisien, and ington, DC, May 15-16, 2017). M. Versteegen, Phys. Rev. C 92, 054609 (2015). [41] S. A. Karamian and J. J. Carroll, Physics of Atomic Nu- [17] K. W. D. Ledingham, I. Spencer, T. McCanny, R. P. clei 75, 1362 (2012). Singhal, M. I. K. Santala, E. Clark, I. Watts, F. N. Beg, [42] M. Polasik, K. S labkowska, J. J. Carroll, C. J. Chiara, M. Zepf, K. Krushelnick, et al., Phys. Rev. Lett. 84, 899 L. Syrocki, E. W¸eder,and J. Rzadkiewicz, Phys. Rev. C (2000). 95, 034312 (2017). [18] T. E. Cowan, A. W. Hunt, T. W. Phillips, S. C. Wilks, [43] A. P´alffy, W. Scheid, and Z. Harman, Phys. Rev. A 73, M. D. Perry, C. Brown, W. Fountain, S. Hatchett, 012715 (2006). J. Johnson, M. H. Key, et al., Phys. Rev. Lett. 84, 903 [44] A. P´alffy, Z. Harman, and W. Scheid, Phys. Rev. A 75, (2000). 012709 (2007). [19] P. Gibbon, Short Pulse Laser Interactions with Matter. [45] V. P. Krainov and M. B. Smirnov, Phys. Rep. 370, 237 An Introduction (Imperial College Press, London, 2005). (2002). [20] K. M. Spohr, M. Shaw, W. Galster, K. W. D. Ledingham, [46] Experimental Nuclear Reaction Data - EXFOR (2017), L. Robson, J. M. Yang, P. McKenna, T. McCanny, J. J. http://www-nds.iaea.org/exfor/exfor.htm. Melone, K.-U. Amthor, et al., New J. Phys. 10, 043037 [47] M. Hasegawa, Y. Sun, S. Tazaki, K. Kaneko, and 6

T. Mizusaki, Phys. Lett. B 696, 197 (2011). D. A. Liedahl, et al., Phys. Rev. Lett. 100, 085004 [48] H.-K. Chung, M. H. Chen, W. L. Morgan, Y. Ralchenko, (2008). and R. W. Lee, High Energy Density Phys. 1, 3 (2005). [59] D. F. Price, R. M. More, R. S. Walling, G. Guethlein, [49] K. G. Dyall, I. P. Grant, C. T. Johnson, F. A. Parpia, R. L. Shepherd, R. E. Stewart, and W. E. White, Phys. and E. P. Plummer, Comput. Phys. Commun. 55, 425 Rev. Lett. 75, 252 (1995). (1989). [60] A. Saemann, K. Eidmann, I. E. Golovkin, R. C. Mancini, [50] J. C. Steward and K. D. Pyatt, Jr., Astrophys. J. 144, E. Andersson, E. F¨orster,and K. Witte, Phys. Rev. Lett. 1203 (1966). 82, 4843 (1999). [51] C. Lin, G. R¨opke, W.-D. Kraeft, and H. Reinholz, Phys. [61] P. Audebert, R. Shepherd, K. B. Fournier, O. Peyrusse, Rev. E 96, 013202 (2017). D. Price, R. Lee, P. Springer, J.-C. Gauthier, and [52] S. X. Hu, Phys. Rev. Lett. 119, 065001 (2017). L. Klein, Phys. Rev. Lett. 89, 265001 (2002). [53] J. Fuchs, P. Antici, E. d’Humi`eres, E. Lefebvre, [62] Y. Sentoku, A. J. Kemp, R. Presura, M. S. Bakeman, M. Borghesi, E. Brambrink, C. A. Cecchetti, M. Kaluza, and T. E. Cowan, Physics of Plasmas 14, 122701 (2007). V. Malka, M. Manclossi, et al., Nature Physics 2, 48 [63] T. D. Arber, K. Bennett, C. S. Brady, A. Lawrence- (2006). Douglas, M. G. Ramsay, N. J. Sircombe, P. Gillies, [54] Y. Wu and A. P´alffy, The Astrophysical Journal 838, 55 R. G. Evans, H. Schmitz, A. R. Bell, et al., Plasma (2017). Physics and Controlled Fusion 57, 113001 (2015), the [55] F. Brunel, Phys. Rev. Lett. 59, 52 (1987). EPOCH code was developed under the project that was [56] G. Bonnaud, P. Gibbon, J. Kindel, and E. Williams, in part funded by the UK EPSRC grants EP/G054950/1, Laser and Particle Beams 9, 339 (1991). EP/G056803/1, EP/G055165/1 and EP/ M022463/1. [57] P. Gibbon and E. F¨orster,Plasma Physics and Controlled [64] Helmholtz International Beamline for Extreme Fields Fusion 38, 769 (1996). at the European XFEL, Official Website (2017), [58] Y. Ping, R. Shepherd, B. F. Lasinski, M. Tabak, H. Chen, https://www.hzdr.de/db/Cms?pNid=427&pOid=35325. H. K. Chung, K. B. Fournier, S. B. Hansen, A. Kemp, [65] European XFEL, Official Website (2017), http://www.xfel.eu/.