Hostname: page-component-8448b6f56d-m8qmq Total loading time: 0 Render date: 2024-04-16T18:05:34.130Z Has data issue: false hasContentIssue false

Fusion energy using avalanche increased boron reactions for block-ignition by ultrahigh power picosecond laser pulses

Published online by Cambridge University Press:  14 July 2015

Heinrich Hora*
Affiliation:
Department of Theoretical Physics, University of New South Wales, Sydney, Australia
Georg Korn
Affiliation:
Institute of Physics ASCR, v.v.i (FZU), ELI-Beamlines project, Prague, Czech Republic
Lorenzo Giuffrida
Affiliation:
Institute of Physics ASCR, v.v.i (FZU), ELI-Beamlines project, Prague, Czech Republic
Daniele Margarone
Affiliation:
Institute of Physics ASCR, v.v.i (FZU), ELI-Beamlines project, Prague, Czech Republic
Antonino Picciotto
Affiliation:
Micro-Nano Facility, Fondazione Bruno Kessler, Trento, Italy
Josef Krasa
Affiliation:
Institute of Physics ASCR, v.v.i (FZU), ELI-Beamlines project, Prague, Czech Republic
Karel Jungwirth
Affiliation:
PALS Laboratory, Institute of Plasma Physics of the ASCR, Prague, Czech Republic
Jiri Ullschmied
Affiliation:
PALS Laboratory, Institute of Plasma Physics of the ASCR, Prague, Czech Republic
Paraskevas Lalousis
Affiliation:
Institute of Electronic Structure and Lasers FORTH, Heraklion, Greece
Shalom Eliezer
Affiliation:
Institute of Nuclear Fusion, Polytechnic University of Madrid, Madrid, Spain Soreq Research Center, Yavne, Israel
George H. Miley
Affiliation:
Department Nuclear Plasma & Radiological Engineering, University of Illinois, Urbana, Illinois
Stavros Moustaizis
Affiliation:
Laboratory of Matter Structure and Laser Physics, Technical University of Crete, Chania, Greece
Gérard Mourou
Affiliation:
DAER-IZEST, Ecole polytechnique, Route de Saclay, Palaiseau Cedex, France
*
Address correspondence and reprint requests to: Heinrich Hora, Department of Theoretical Physics, University of New South Wales, Sydney 2052, Australia. E-mail: h.hora@unsw.edu.au
Rights & Permissions [Opens in a new window]

Abstract

Exceptionally high reaction gains of hydrogen protons measured with the boron isotope 11 are compared with other fusion reactions. This is leading to the conclusion that secondary avalanche reactions are happening and confirming the results of high-gain, neutron-free, clean, safe, low-cost, and long-term available energy. The essential basis is the unusual non-thermal block-ignition scheme with picosecond laser pulses of extremely high powers above the petawatt range.

Type
Research Article
Copyright
Copyright © Cambridge University Press 2015 

1. INTRODUCTION

Fusion reactions of hydrogen protons with the boron isotope 11 (HB11) (Oliphant & Rutherford, Reference Oliphant and Rutherford1933) were from the beginning, of special interest. These occurred at collision energies of or below 100 keV, while nuclear reactions due to the Coulomb barrier needed always several MeV per proton to squeeze the nuclei to distances of their diameters of femtometers (fm). This was the first time to measure the nuclear reactions occurring at distances in the range of 100 fm and lead to the discovery of the reaction of heavy hydrogen, deuterium D, with the super heavy hydrogen tritium T (Oliphant et al., Reference Oliphant, Harteck and Rutherford1934). That these reactions occurred at 100 times higher distance showed that the DT fusion reaction was extremely exceptional occurring at collision energies of several keV.

DT is the easiest binary nuclear reaction and preferably used to study energy production in controlled nuclear fusion power generators

(1)$${\rm D} + {\rm T} = ^4 {\rm He} + {\rm n} + 17.6\,{\rm MeV}$$

resulting in clean helium (He) and in neutrons n of 14.1 MeV energy and a lifetime 12.6 min before decaying into an electron and a proton. During this lifetime the neutrons can penetrate into clean (non-radioactive, stable) nuclei and change them into dangerous radioactive nuclei. One has to realize that DT fusion primarily produces nearly four times more neutrons per gained energy than in fission reactors and another difference is that the transmuted stable nuclei in a fission reactor (referring here to the completely established fission power stations or the advancement of small modular reactors (Paterson, Reference Paterson2015) by Westinghouse or Chinese–Australian cooperation) are mostly contained in the nuclear waste and much easier kept under control than in fusion reactors where they are spreading through the blanket and escape to transmute widely dispersed clean stable nuclei into dangerous ones. This fact was described as disadvantage that DT fusion leads to the “hottest radioactive environment on earth” which was explicitly endorsed by Sir William Mitchell (Butler, Reference Butler1996). It was elaborated how this problem of DT fusion has to be overcome (Tahir & Hoffmann, Reference Tahir and Hoffmann1997). However, the following reported new aspects may lead to other options for fusion energy; it has to be acknowledged that these all would not have been reached than by standing on the shoulders of the long past and ongoing most extensive research and achievements gained from fusion work with DT.

The other important result of the HB11 fusion (Oliphant & Rutherford, Reference Oliphant and Rutherford1933) is the generation of mostly equal energetic He nuclei (α-particles)

(2)$${\rm H} + ^{11} {\rm B} = 3^4\, {\rm He} + 8.7\,{\rm MeV}{\rm.} $$

Apart from the result of the anomalous large distant reaction, the fact is important that no neutrons are being generated primarily. This was discussed (Mitchell, Reference Mitchell2000) also in contacts with J.C. Kelly (Hora et al., Reference Hora, Miley, Ghoranneviss, Malekynia and Azizi2009). It was clarified (Weaver et al., Reference Weaver, Zimmerman and Wood1973) that any side reactions produce less radioactivity per gained energy than by burning coal due to its contents of 2 ppm uranium. Therefore, any nuclear radiation problem of HB11 fusion can be ignored.

The only earlier problem before the present new developments was the fact that the HB11 fusion was extremely more difficult than DT fusion if the usual gas-dynamic compressions and thermal ignition with laser pulses of nanosecond (ns) duration were applied. Even if the advanced measurements of the fusion cross-sections at the resonance energy of 148 keV (Nevins et al., Reference Nevins and Swain2000) were used (Kouhi et al., Reference Kouhi, Ghoraneviss, Malekynia, Hora, Miley, Sari, Azizi and Razavipourm2011), the compression of the HB11 fuel had to be above 100,000 times of the solid state. The more it was surprising when instead of the ns laser pulses with their thermal ignition of the fusion, an alternative non-thermal block ignition with picosecond (ps) laser pulses of extreme high power was used in the computations (Hora et al., Reference Hora, Miley, Ghorannviss, Malekynia, Azizi and He2010), that then the difficulty for HB11 laser fusion had changed and drastically decreased to about the same limits as for DT.

This enormous improvement, however, was not sufficient for reaching conditions of economic energy production with HB11 laser fusion. Up to this stage (Hora et al., Reference Hora, Miley, Ghorannviss, Malekynia, Azizi and He2010; Lalousis et al., Reference Lalousis, Hora, Eliezer, Martnez-Val, Moustaizis, Miley and Mourou2013), the computations were still pessimistically using only binary reactions as always applied for fusion. The necessary further improvement beyond the binary reactions for HB11, however, is essential for the favorable non-thermal ps-block ignition described in the following. The experimental results for this improvement are by secondary reactions after the binary reactions which are first described in the following section before the later sections are applying this result to the details of HB11 block ignition.

2. VERY HIGH BORON FUSION EXPLAINED AS AVALANCHE REACTION

The strong interest in the HB11 fusion reaction was evident and a large number of proposals were considered including extreme non-thermal magnetic field configurations in low-density plasmas (Rostoker et al., Reference Rostoker, Binderbauer and Monkhorst1997). Substantial progress was reached when direct measurements of the reaction were performed (Belyaev et al., Reference Belyaev, Matafonov, Vinogradov, Krainov, Lisitsa, Roussetski, Ignatyev and Andrianov2005). This found attention though the laser-driven plasmas produced only 1000 reactions, just close to the level of experimental sensitivity. A convincing and very transparent result (Labaune et al., Reference Labaune, Deprierraux, Goyon, Loisel, Yahia and Rafelski2013) produced more than 106 reactions though only a complicated combination of laser-driven directed particle beams with a secondary laser–target interaction was used. After this breakthrough, the measurement of 109/sr reactions (Picciotto et al., Reference Picciotto, Margarone, Velyhan, Bellini, Krasa, Szydlowski, Bertuccio, Shi, Margarone, Prokupek, Malinowska, Krouski, Ullschmied, Laska, Kucharik and Korn2014; Margarone et al., Reference Margarone, Picciotto, Velyhan, Krasa, Kucharik, Mangione, Szydlowsky, Malinowska, Bertuccio, Shi, Crivellari, Ullschmied, Bellutti and Korn2015) was achieved. These measurements were a straightforward laser-fusion result by a unique combination of one of the largest European lasers, the PALS iodine laser at Prague, with the use of a very unusual laser-irradiated target. However, these fusion gains were too high compared with preceding knowledge.

The iodine laser matured (Hohla & Kompa, Reference Hohla and Kompa1973) to a precisely working system only after a fast inductive operation for the photo-ionizing flash lamps was possible and a high technology amplifier system was developed (Witte, 1990, private communication) for operation in an optimized environment of PALS (Jungwirth et al., Reference Jungwirth, Cejnarova, Juha, Kralikova, Krasa, Krousky, Krupickova, Laska, Masek, Mocek, Pfeifer, Präg, Renner, Rohlena, Rus, Skala, Straka and Ullschmied2001). Laser pulses in the range of 100 ps duration and up to kJ energy provided a system which was uniquely different from the broad stream of development.

The other unusual ingredient of the experiment (Picciotto et al., Reference Picciotto, Margarone, Velyhan, Bellini, Krasa, Szydlowski, Bertuccio, Shi, Margarone, Prokupek, Malinowska, Krouski, Ullschmied, Laska, Kucharik and Korn2014; Margarone et al., Reference Margarone, Picciotto, Velyhan, Krasa, Kucharik, Mangione, Szydlowsky, Malinowska, Bertuccio, Shi, Crivellari, Ullschmied, Bellutti and Korn2015) was the use of an exotic target, consisting of silicon with an extremely high concentration of boron. Low concentration of boron produced the transistor effects from the beginning and the very high concentration is in the focus of semiconductor research for degenerate hole concentration. This is a crossing point of solid-state physics with high laser intensity interaction showing an influence by surface effects due to high trap concentrations (Boreham et al., Reference Boreham, Newman, Höpfl and Hora1995). It was the special advantage to combine the solid-state physics research of silicon with extremely high boron concentration (apart from the natural high proton concentration) for the HB11 laser fusion experiments at PALS.

The result (Picciotto et al., 2014) was most significant that a PALS laser pulse of 200 J on the special silicon target resulted in the generation of 109/sr HB11 reactions. This was considerable more than100 times higher than measured before (Labaune et al., Reference Labaune, Deprierraux, Goyon, Loisel, Yahia and Rafelski2013), with the further advantage that the experiment (Picciotto et al., 2014) was a direct and transparent laser-fusion experiment without any complex modification due to ion beams, etc.

After presenting these results with extraordinary high enhancement of the yield of induced nuclear reacting ultraclean boron fusion (Korn et al., Reference Korn, Margarone and Picciotto2014), the evaluation was followed up how these gains could have happened. The experiment was within the category of ns laser pulse-driven fusion without special conditions of pulse contrast preparation for suppression of prepulses, etc. Considering the neutron gains N from DT reactions depending on the energy E p of ns laser pulses in Figure 1 (Hora, Reference Hora2013a), an approximate relation

(3)$$N = {\rm const}\;E_{\rm p} ^2 $$

can be derived. Comparing the gain of 109 for 200 J laser pulses with the 1.9 MJ pulse of the highest measured DT reaction would correspond to 3.6 × 1017 reactions which is far above the measured value (Hurricane et al., Reference Hurricane, Callahan, Casey, Celliers, Cerjan, Dewald, Dittrich, Döppner, Hinkel, Berzak Hopkins, Kline, Le Pape, Ma, MacPhee, Milovich, Pak, Park, Patel, Remington, Salmonson, Springer and Tommasini2014). This permits the conclusion that the evaluation of the HB11 reaction (Korn et al., Reference Korn, Margarone and Picciotto2014) cannot be a binary reaction as in usual computations for fusion.

Fig. 1. Measured highest DT fusion neutron gains N per incident laser energy (full signs) of laser pulses in the range of ns duration (Hora, Reference Hora2013a) with the highest value reported (Hurricane et al., Reference Hurricane, Callahan, Casey, Celliers, Cerjan, Dewald, Dittrich, Döppner, Hinkel, Berzak Hopkins, Kline, Le Pape, Ma, MacPhee, Milovich, Pak, Park, Patel, Remington, Salmonson, Springer and Tommasini2014) given by an asterisk.

The conclusion is that the measured strongly enhanced reaction yield of HB11 is the result of secondary reactions as was discussed before (Hora et al., Reference Hora, Lalousis, Moustaizis, Földes, Miley, Yang, He, Eliezer and Martinez-Val2012a) mentioned as “secondary reactions” (Folgereaktionen) in a patent application (Hora, Reference Hora2013b) causing an avalanche increase of the reaction. Permission for no restrictions at publications was given formally by the German Patent Office (Hora, Reference Hora2014) on which documentation the open discussion is based (Hora et al., Reference Hora, Lalousis, Eliezer, Miley, Moustaizis and Mourou2014b). This can be seen from considering elastic central collisions between the mass m a of an α-particle of energy E a with an initially resting 11B nucleus with the mass m B. The energy of the boron after collision is (Hora et al., Reference Hora, Lalousis, Moustaizis, Földes, Miley, Yang, He, Eliezer and Martinez-Val2012a)

(4)$$E_{\rm B} = 4m_{\rm a} m_{\rm B} E_{\rm a} /\left( {m_{\rm a} + {\rm }m_{\rm B} } \right)^2 . $$

Using the initial energy of the α-particle from the HB11 reaction of 2.9 MeV, the first elastic hit transfers 530 keV energy from the α-particle to the boron. After this collision, the α-particle can produce a second hit to boron which is then gaining 492 keV, etc. As known from Nevins et al. (Reference Nevins and Swain2000), there is an exceptional high fusion cross-section, widely spread, with protons at 550 keV energy, about ten times higher than known from other fusion reactions. The primary produced α-particles are then causing a secondary reaction resulting in three new alphas, etc., if there is not a problem with the stopping length of the alphas.

Instead of studying several stopping power models using appropriate simplifications, the discussion of the measured strongly enhanced reaction yield of HB11 (Korn et al., Reference Korn, Margarone and Picciotto2014; Hora et al., Reference Hora, Lalousis, Guiffrida, Margarone, Korn, Eliezer, Miler, Moustaizis and Mourou2015) can be used as clarification that there is the avalanche process happening. This avalanche process for a fusion reaction is unique for HB11 in contrast to the other fusion reactions which are based only on binary reactions without the possibility of a following-up reaction.

As a further support for the used relation of Eq. (3) for DT, the reaction of DD for the ns range of laser pulses shows a similar relation (Krasa et al., Reference Krasa, Klier, Velyhan, Margarone, Krousky, Jungwirth, Skala, Pfeifer, Kravarik, Kubes, Rezak and Ullschmied2013) only with the well-known lower reaction gain by a factor of about 300. These results for the longer laser pulses are to be separated from results with ps pulses for DD fusion. In these cases, the DD gains can be higher than the normal relation of the kind of Eq. (3) with the increase of more than 104 reported by Norreys et al. (Reference Norreys, Fews, Beg, Bell, Dangor, Lee, Nelson, Schmidt, Tatarakis and Cable1998) based on extremely clean laser pulses with high contrast ratio. This case is, the first example of a fusion experiment showing the drastic difference between thermal interactions with ns pulses in contrast to the non-thermal direct ps transfer of optical laser energy into macroscopic motion of plasma blocks considered in the following sections (Hora, Reference Hora2003).

3. NON-THERMAL LASER ENERGY TRANSFER TO TARGETS

In order to explain the laser fusion mechanism on which the following results with application of the just reported avalanche reaction for HB11 is based, we summarize the main achievements of the plasma block initiation of the ignition process based on ps laser pulses. This is basically different from the mostly studied laser fusion process with ns laser pulses using gas-dynamic thermal heating, pressure-compression, and ignition processes with complex thermalizing optical absorption, electron–ion energy-delayed classical or quantum defined ion-heating mechanisms (Hora et al., Reference Hora, Sadighi-Bonabi, Yazdani, Afaideh, Nafari and Ghoranneviss2012b), including instabilities, radiation losses, thermally determined reactions, etc. The details of these processes are now mostly under control (Hurricane et al., Reference Hurricane, Callahan, Casey, Celliers, Cerjan, Dewald, Dittrich, Döppner, Hinkel, Berzak Hopkins, Kline, Le Pape, Ma, MacPhee, Milovich, Pak, Park, Patel, Remington, Salmonson, Springer and Tommasini2014), while different options of the direct or indirect drive, volume or spark ignition and self-heat by the reaction process, and re-absorption of bremsstrahlung are involved (Hora, Reference Hora2013a).

For understanding the basic difference of the non-thermal and to some extent simplified plasma block-ignition process based on non-linear (ponderomotive-dielectric determined) forces using ps laser pulses, we have to remind to the dozens of years lasting development. When the early laser pulses of frequency ω with powers below MW were focused on targets, the heating and dynamic processes followed all the preceding known physics with emission of ions of few eV according to temperatures of about 30,000 K. But at only 10 times higher laser power, ions had 1000 times higher energies and were separated by their charge numbers confirming non-thermal, electrodynamic interaction. This led to the derivation of forces by lasers in media with spatial or temporal variation of dielectric properties (Hora, Reference Hora1969; Reference Hora1985) using the refractive index n of the generated plasmas. The force density was then not only given by the gas-dynamic pressure p

(5)$${\bf f} = - \nabla p + {\bf f}_{{\rm NL}} $$

but also by Maxwell's stress tensor as Lorentz and gauge-invariant non-linear force (Hora, Reference Hora1969; Reference Hora1985; Reference Hora2000)

(6)$$\eqalign{{\bf f}_{\rm NL} & = \nabla \bullet \left[{ {\bf EE} + {\bf HH} - 0.5 ({\bf E}^2 + {\bf H}^2){\bf 1} + (1 + (\partial /\partial t)/{\rm \omega} )}\right.\cr & \quad \times \left.{( {\bf n}^2 - 1){\bf EE} }\right]/(4{\rm \pi} ) - (\partial /\partial t){\bf E} \times {\bf H}/(4{\rm \pi} c).} $$

At plane laser wave interaction with a plane plasma front, the force reduces to

(7)$$\eqalign{{\bf f}_{{\rm NL}} & = - (\partial /\partial x)\left( {{\bf E}^2 + {\bf H}^2} \right)/(8{\rm \pi} )\; \cr & = {\rm} -\! {\rm} ({\rm \omega} _{\rm p} /{\rm \omega} )^2 (\partial /\partial x)\left( {E_{\rm v} ^2 /{\bf n}} \right)/(16{\rm \pi} )}$$

showing how the force density is given by the negative gradient of the electromagnetic laser field density. Expressing the magnetic laser field from Maxwell's equations by the electric field results in the final expression which is equal to Kelvin's ponderomotive force in electrostatics of 1845. For non-linear correct results – in contrast to incorrect linear physics, see Feynman, Section 6.3 of Hora (Reference Hora2000) – all components of the stress tensor have to be used Cicchitelli (Reference Cicchitelli, Hora and Postle1990).

Following the very first general hydrodynamic computations with the force (5), the plasma motion at sufficiently high laser intensities is dominated by the non-linear force well with some corrections by the thermodynamic processes (Hora et al., Reference Hora, Castillo, Clark, Kane, Lawrence, Miller, Nicholson-Florence, Novak, Ray, Shepanski and Tsivinsky1978). Drawn are the results for 1018 W/cm2 laser intensity from (Hora, Reference Hora1981: Fig. 10.18a & b) at plane geometry irradiation in Figure 2 incident from the right-hand side on a deuterium plasma with an initial bi-Rayleigh density profile (Fig. 10.17 of Hora, Reference Hora1981). This results in a plasma block of more than 15 vacuum wavelengths thickness moving after 1.5 ps against the laser and a deeper block inside the deuterium parallel to the laser direction into the interior of the deuterium. The acceleration of each of the plasma blocks is above 1020 cm/s2 for reaching velocities above 109 cm/s during 1.5 ps.

Fig. 2. 1018 W/cm2 neodymium glass laser incident from the right-hand side on an initially 100 eV hot deuterium plasma whose density has a very low-reflecting bi-Rayleigh profile at initial time t = 1, resulting in a laser energy density and a velocity distribution from plasma hydrodynamic computations at an interaction time t = 1.5 ps of interaction. The driving non-linear force is the negative of the varying energy density of the laser field (E 2 + H 2)/8π expressing the intensity. The dynamic development of temperature and density had accelerated the plasma block of about 20 vacuum wavelength thickness of the dielectric enlarged skin layer moving against the laser (positive velocity) and another block into the plasma (negative velocity) showing ultrahigh >1020 cm/s2 acceleration.

This ultrahigh acceleration was measured for the plane geometry by Sauerbrey (Reference Sauerbrey1996) which was 100,000 times higher than ever measured before in a laboratory. The sophisticated measurements by Doppler shift of spectral lines reflected from the outer block moving against the incident laser beam were repeated by Földes et al. (Reference Fujioka, Zhang, Ishihara, Shigemori, Hironaka, Johzaki, Sunahara, Yamamoto, Nakashima, Watanabe, Shiraga, Nishimura and Azechi2000) in agreement with the theory of non-linear force acceleration evaluated (Hora et al., Reference Hora, Badziak, Boody, Höpfl, Jungwirth, Kralikowa, Kraska, Laska, Parys, Perina, Pfeifer, Rohlena, Skala, Ullschmied, Wolowski and Woryna2002a; Reference Hora, Badziak, Read, Li, Liang, Liu, Sheng, Zhang, Osman, Miley, Zhang, He, Peng, Osman, Glowacz, Jablonski, Wolowski, Skladanowski, Jungwirth, Rohlena and Ullschmied2007). The ion density in the space charge neutral blocks was – as theoretically predicted 1978 – above 1012 Amp/cm2 (Hora et al., Reference Hora, Peng, Zhang and Osman2002b).

The generation of the ultrahigh acceleration of the non-linear force-driven plasma blocks was considered from the beginning for the explanation of the measured highly directed plasma motion with the high ion energies (Hora, Reference Hora1981) as shown in Figure 2 with the plasma block moving to the right-hand side against the laser light and the other block moving into the interior of the target. Schematically, this can be seen in Figure 3 (Hora et al., Reference Hora, Peng, Zhang and Osman2002b). Apart from the appearance of the ultrahigh velocities, the high densities of the space charge neutral plasma blocks resulted in very high ion current densities above 1012 Amp/cm2 (Hora et al., Reference Hora, Badziak, Boody, Höpfl, Jungwirth, Kralikowa, Kraska, Laska, Parys, Perina, Pfeifer, Rohlena, Skala, Ullschmied, Wolowski and Woryna2002a) which values were predicted in the computer outputs of 1978 (Hora et al., Reference Hora, Castillo, Clark, Kane, Lawrence, Miller, Nicholson-Florence, Novak, Ray, Shepanski and Tsivinsky1978; Hora, Reference Hora1981) (Fig. 2). These ion current densities are more than million times higher than in conventional accelerators. A repetition of the experiments by Sauerbrey (Reference Sauerbrey1996) was possible by Földes et al. (Reference Földes, Bakos, Gal, Juhasz, Kedves, Koscis, Szatmari and Veres2000), Figure 4, using the appropriately lower laser intensities in accurate agreement with the theory.

Fig. 3. Scheme of skin depth laser interaction, where the non-linear force accelerates a plasma block against the laser light and another block toward the target interior. In front of the blocks are electron clouds of the thickness of the effective Debye lengths [Eq. (5)] (Hora et al., Reference Hora, Peng, Zhang and Osman2002b).

Fig. 4. Intensity dependence of the velocity of the plasma front from the Doppler line shift of the reflected from the irradiation of 700 fs KrF laser pulses on an aluminum target (Földes et al., Reference Földes, Bakos, Gal, Juhasz, Kedves, Koscis, Szatmari and Veres2000).

The ultrahigh acceleration was comparably similar when treated by Target Normal Sheath Acceleration (TNSA) (Yu et al., Reference Yu, Xu, He, Yu, Ishoguro, Zhang and Wong2005), calling then the blocks “bunch” with more details of the relatively much higher Debye length than in Figure 3. The cases of Yu et al. (Reference Yu, Xu, He, Yu, Ishoguro, Zhang and Wong2005) were chosen for the lower target thickness and other plasma parameters resulting then, in a more pronounced sheath generation which could be seen also from a double-layer model (Eliezer et al., Reference Eliezer, Nissim, Martinez Val, Mima and Hora2014) arriving at corresponding results.

The most important need for the discovery to confirm the theoretical prediction of the ultrahigh acceleration (Sauerbrey, Reference Sauerbrey1996; Földes et al., Reference Földes, Bakos, Gal, Juhasz, Kedves, Koscis, Szatmari and Veres2000) and of the ultrahigh ion current densities (Hora et al., Reference Hora, Badziak, Boody, Höpfl, Jungwirth, Kralikowa, Kraska, Laska, Parys, Perina, Pfeifer, Rohlena, Skala, Ullschmied, Wolowski and Woryna2002a) was based on Mourou's et al., chirped-pulse amplification (CPA) to generate extreme powerful laser pulses of ps and shorter duration reaching now 7.12 petawatt (PW) power and 27 fs duration (Li et al., Reference Li, Liang, Chu, Gan, Lu, Yu, Yin, Leng, Lu, Wang and Xu2015) with Ti:sapphire lasers with sufficient suppression of transverse parasitic lasing have been produced with the aim that 30 PW-ps pulses may be available within few years by combining of a number of beams, if not fiber lasers with single-mode (Mourou et al., Reference Mourou, Brocklesby, Tajima and Limpert2013) and spherically focused (Hora et al., Reference Lalousis, Hora and Moustaizis2014a) output will be available. This was crucial for the measurement by Sauerbrey (Reference Sauerbrey1996) together with the second ingredient of using an extreme contrast ratio against prepulses to avoid relativistic self-focusing as clarified by Zhang et al. (Reference Zhang, He, Chen, Li, Zhang, Wang, Feng, Zhang, Tang and Zhang1998).

It should be underlined that the block ignition can be considered as a modification of the fast ignition as introduced in 1991 by Campbell (Reference Campbell2006) and elaborated by Tabak et al. (Reference Tabak, Hammer, Glinsky, Kruer, Wilks, Woodworth, Campbell, Perry and Mason1994).

4. BLOCK IGNITION FOR LASER FUSION

The non-thermal generation of plasma blocks by ultrahigh acceleration with ps laser pulses by the non-linear force can be used for the ignition of fusion following the discovery by Chu (Reference Chu1972) and Bobin (Reference Bobin, Schwarz and Hora1974) for initiation of a fusion reaction flame to ignite solid density DT fuel. The result was, however, that the ps laser pulses have at least a minimum energy flux density of few E*=108 J/cm2. Updating the theory (Hora et al., Reference Hora, Malekynia, Ghoranneviss, Miley and He2008) led to a reduction of E* by up to a factor 20.

First we report block-ignited fusion results for DT fusion before continuing about HB11. The ps pulse-generated plasma blocks with their ultrahigh acceleration initiated a fusion flame in solid density DT where plane wave laser interaction with an energy flux density E* of 108 J/cm2 was necessary (Chu, Reference Chu1972; Bobin, Reference Bobin, Schwarz and Hora1974). The updating of the computations of Chu (Reference Chu1972) arrived at similar results (Hora, Reference Hora and Kelly2009). These computations used one-fluid hydrodynamics results where the updating was necessary for the later known reduction of the thermal conduction by an inhibition factor based on the electric double layer between the hot flame and the DT fuel. Further, the stopping power had to use the Gabor collective model at this high plasma densities and not the Bethe–Bloch binary collision model. After this updating, the ignition thresholds for the fusion flame arrived at up to 20 times lower values of E* (Hora et al., Reference Hora, Malekynia, Ghoranneviss, Miley and He2008).

After initial results with one-fluid hydrodynamics (Hora, Reference Hora2009) were performed it was important to receive more information from applying the genuine two-fluid hydrodynamics (Lalousis & Hora, Reference Lalousis and Hora1983; Hora et al., Reference Hora, Lalousis and Eliezer1984) about the dynamical generation of internal electric fields and of shock processes (Betti et al., Reference Betti, Chou, Anderson, Perkins, Theobald and Sodolov2007) and diffusion. Using the same parameters for the generation of ps ultrahigh-accelerated plasma blocks of about a 5 µm depth moving into solid density DT, we are summarizing here further reported results following the initial computations (Lalousis & Hora, Reference Lalousis and Hora1983; Hora et al., Reference Hora, Lalousis and Eliezer1984). Figure 5 shows ion densities with shock profiles moving at the interaction edge into the solid density DT. A shock structure shows maximum densities four time the solid state as it is the result of the Rankine–Hugoniot simplified analytical model. The shock thickness is increasing on time what could be covered only by complete inclusion of thermal and plasma properties in our genuine two-fluid computation (Lalousis & Hora, Reference Lalousis and Hora1983; Hora et al., Reference Hora, Lalousis and Eliezer1984). It is interesting that the shock is pronounced only very late after 2 ns being only little visible at 200 ps. It should be mentioned that the same plots for three times lower E* values shows an about two times earlier building up of the shock. The shock velocity was evaluated for varying E* and resulted in Figure 6 at 2 ns. It can be seen that the fast increasing velocity at E* of 108 J/cm2 is reaching values higher than 2000 km/s with convincing ignition properties. In order to understand the reasons for the long (>200 ps) delay of the shocking process, we show in Figure 7 the longitudinal electric field which can be seen only form the genuine two-fluid hydrodynamics. It is remarkable that during the very first few hundred ps after the laser pulse in the phase of establishing the shock front, the electric field is rather high and decays only at 2 ps to comparably low values when the shock is nearly fully developed (Fig. 5).

Fig. 5. Ion densities N i depending of the depth X after a 1 ps 3 × 108 J/cm2 laser pulse of 248 nm wavelength has irradiated solid density DT. Parameter is the time after the ps interaction.

Fig. 6. Velocity V of the fusion flame at 2 ns after the initiating ps laser pulse of 248 nm wavelength depending on the energy flux E*.

Fig. 7. Longitudinal electric field E as in case of Figure 1 for E* = 108 J/cm2 depending of the depth X at times of 40 ps; 400 ps; 1 ns; 2 ns from the highest maximum, respectively.

The studies with the genuine two-fluid hydrodynamics permits an evaluation about the decreasing velocity of the front of the fusion flame at the very long times of ns after the ps initiation of the flame. Figure 8 shows the results of fusion reaction rates with a 1 ps laser pulse of E* = 108 J/cm2. The velocity of the fusion flame is decreasing from 2130 km/s at 2 ns to 1070 km/s at 10 ns and the values of the reaction rates decay faster than in the case with only three times higher irradiation. Nevertheless, the fusion gains are not very much lower, going down at 10 ns. The gains at the time of 2 ns are nearly the same. The result of the ion density at the flame front has reached nearly the Rankine–Hugoniot value of four times compression earlier at 2 ns than in the case with three times higher irradiation of Figure 5.

Fig. 8. Reaction rates (to be multiplied by 1036 m−3/s) in solid DT at 1 ps pulse of energy flux E* = 3 × 108 J/cm2 KrF laser irradiation depending on the fuel depth × at different times up to 10 ns.

The fact of the growing thickness of the compression range on time may be understood by the thermal mechanisms around the fusion flame with a stronger effect at the first case with the higher energy flux. This can also be understood that the shock compression in the early stage appears within a too thin area such that the thermal mechanisms – completely covered by the code – are causing a dissipation of the too thin compression areas.

We may conclude that the earlier studied mechanisms of shock ignition (Betti et al., Reference Betti, Chou, Anderson, Perkins, Theobald and Sodolov2007) for fusion as evaluated for longer ns pulses is supported by our results at least for these longer times (Eliezer et al., Reference Eliezer, Nissim, Martinez Val, Mima and Hora2014). This supports the results of shock wave ignition for fusion (Betti et al., Reference Betti, Chou, Anderson, Perkins, Theobald and Sodolov2007) which requires a long (few ns) and high-energy laser pulse. The high pressure (P) shock wave thickness (d) in the ns interaction case is estimated from the equality of pressures at the interface between a flyer (or a plasma block with density ρ0 and flow velocity u 0) and the compressed target (the nuclear fuel: DT or pB11, etc. with density ρC and flow velocity u C). Since P ~ ρ0u 02 ~ ρCu C2 and the shock wave transition time in the flyer with a thickness l is t = l/u 0, one gets a shock wave thickness d of the order of d ~ (ρ0C)1/2l. In this purely shock ignition scheme – in contrast to the here studied ps plasma block ignition – the ns case values are (ρ0C) ~ 0.001 and l ~ 1 µm implying d ~ 0.03 µm. The ignition criterion is based on the requirement that the α-particles created in the DT reaction are reabsorbed in the hot spot implying a “ρR” value larger than 0.3 g/cm2 for a temperature about 10 keV and larger values for higher temperatures. In the shock wave ignition, “ρR” = ρCd ~ l0ρC)1/2 ~ 0.003 g/cm2 for our case which is two order of magnitude too low. This result fully explains why the ns interaction scheme for shock ignition (Lalousis & Hora, Reference Lalousis and Hora1983; Hora et al., Reference Hora, Lalousis and Eliezer1984; Betti et al., Reference Betti, Chou, Anderson, Perkins, Theobald and Sodolov2007) is basically different from the here discussed ps laser pulse-initiated block-ignition scheme.

It was very surprising when extending these very detailed and established results for DT fusion were applied to the fusion of HB11. Against all knowledge that fusion based on binary reactions as used for DT, the threshold E* arrived at about the same values for HB11. It was most unexpected and curious, that this case had a threshold E* of a similar level as DT (Hora et al., Reference Hora, Miley, Ghoranneviss, Malekynia and Azizi2009; Reference Hora, Miley, Ghorannviss, Malekynia, Azizi and He2010). In addition, it should be underlined that the calculations for HB11 were pessimistically performed in the same way with binary nuclear reactions only as in the case of DT because the secondary reactions of the avalanche process (see the second section of this paper) were not known at this time.

Though these calculations were based only on plane wave geometry (Hora et al., Reference Hora, Miley, Ghoranneviss, Malekynia and Azizi2009; Reference Hora, Miley, Ghorannviss, Malekynia, Azizi and He2010), the clarification of the basic change of the earlier known extreme difficulties of HB11 fusion against DT resulting now for the block ignition for HB11 as “no difficulties” was interpreted as: “This has the potential to be the best route to fusion energy” (Haan, Reference Haan and Yuandi2010).

Summarizing the block ignition, the ps ultrahigh-power laser ignition of fusion with the non-thermal direct conversion of the laser pulse energy into mechanical motion of plasma blocks is essentially different from the fusion with ns pulse ignition of fusion based on thermal interaction. In the ns case, problems are with thermalizing collision processes, delaying equipartition for ion heating, instabilities, etc. This difference is simply seen form Eq. (5), that the ns process is based on the first term of the force density and the ps process is based on the second term of non-thermal ps interaction with directly converting of the optical laser energy into macroscopic mechanical motion of the plasma blocks. This fundamental difference is including the principle of non-linearity (see reference to Feynman, Reference Feynman1987) in relation to the prediction of the block generation by the non-linear force (Fig. 3) and measured first by Sauerbrey (Reference Sauerbrey1996) as ultrahigh acceleration of the plasma blocks (Hora et al., Reference Hora, Badziak, Read, Li, Liang, Liu, Sheng, Zhang, Osman, Miley, Zhang, He, Peng, Osman, Glowacz, Jablonski, Wolowski, Skladanowski, Jungwirth, Rohlena and Ullschmied2007). The accelerations were more than 10,000 times higher than measured before in laboratories. This could then be used for the side-on ignition of Chu (Reference Chu1972) and Bobin (Reference Bobin, Schwarz and Hora1974) ps fusion-flame initiation; however, only after the CPA of shorter than ps laser pulses was available (Strickland & Mourou, Reference Strickland and Mourou1985; Mourou et al., Reference Mourou, Barty and Perry1998).

It should be underlined explicitly that the here used ps process of block ignition follows a different principle of interaction. Instead of the heating with ns pulses, the non-thermal electrodynamic non-linear force interaction is a macroscopic process known from the specific microscopic quantum properties, reducing instabilities, delays for thermal equipartion processes, and chaotic thermostatistics. This differentiation was also illustrated by Edward Teller und Lord May of Oxford; see citations in the paper of Hora (Reference Hora2011).

5. AVALANCHE BLOCK IGNITION WITH ULTRAHIGH MAGNETIC FIELDS

The just reported application of experimentally confirmed ultrahigh acceleration of plasma blocks to be used for computation of block ignition of fusion, by non-linear force-driven ps laser pulses, was based on plane geometry interaction. For thin laser beams, the energy losses in the solid targets have to be taken into account. The first steps were to study spherical generation of ps-block ignition for DTand HB11 fusion, where, however, the reached gains were not sufficiently high and a cylindrical geometry with magnetic field trapping was the next step.

Documented before the preparations to the 2012 IAEA Fusion Energy Conference (Hora et al., Reference Hora, Lalousis, Moustaizis, Földes, Miley, Yang, He, Eliezer and Martinez-Val2012a), the spherical irradiation has been studied with shock generation and increase of density (Hora et al., Reference Hora, Lalousis and Moustaizis2014a) with similar results to Malekynia and Razipur (Reference Malekynia and Razipur2013). For generating spherical converging ps laser pulses of more than EW (exawatt = 1018 W) power, this can be combined with the advantages which were opened using fiber glass generation of CPA (Mourou et al., Reference Mourou, Barty and Perry1998; Moustaizis et al., Reference Moustaizis, Lalousis, Hora, Hein, Korn and Silva2013; Hora et al., Reference Hora, Lalousis and Moustaizis2014a) amplified laser pulses as developed by the ICAN Project (Mourou et al., Reference Mourou, Brocklesby, Tajima and Limpert2013). This is also of interest for other uses than for fusion energy generation as electron or hadron accelerators with many applications including cancer therapy (Banati et al., Reference Banati, Hora, Lalousis and Moustaizis2014), as well as for generating PeV particles or for conditions of high-energy physics for exploring extremely high field processes as pair production in vacuum. This is also of interest for a fusion power station to overcome problems of lateral energy losses of the plane wave front geometry of the ICAN fiber laser (Mourou et al., Reference Mourou, Brocklesby, Tajima and Limpert2013) to be developed into spherical geometry (Hora et al., Reference Hora, Lalousis and Moustaizis2014a) as can be seen from Figure 9.

Fig. 9. Generation of a spherical shrinking laser pulse from radially directed fiber ends at a radius 1 to irradiate a spherical solid state fusion fuel 3 with the concentric grid 2 of positive electric charge for slowing down the generated α-particles from the HB11 reactions (Hora et al., Reference Hora, Lalousis and Moustaizis2014a).

The output of the fibers at a sphere 1 of Figure 9 of the amplified beam of the fibers can be directed radially without needing any further optics. The wave front is of such a high-quality single-mode focusing which has (Mourou et al., Reference Mourou, Brocklesby, Tajima and Limpert2013) to go to a diameter of 10 µm such that intensities at about 1021 W/cm2 are reached when the spherically converging laser pulse the surface of a target 3. The fibers automatically permit an exceptionally high quality of single-mode beam uniformity where the usually undesired maxima in the beam profile with glass lasers are automatic eliminated due to the fiber optics quality. The resulting gains for irradiation of uncompressed solid density fuel of DT or of HB11 including avalanche reactions with subsequent dynamic compression was evaluated needing ps laser pulses of EW for fusion gains of few hundred where limitations are given by the limited spherical fuel geometry (Hora et al., Reference Hora, Lalousis and Moustaizis2014a). This case is not of interest for the following consideration.

Another option instead of spherical geometry is to consider cylindrical geometry for the reaction by application of an external pulsed ultrahigh magnetic field. This can be produced by the interaction between of an ns laser beam with a capacitor coil target (Fig. 10). This mechanism has been described and experimentally demonstrated for the first time in 1984 (Daido et al., Reference Daido, Miki, Mima, Fujita, Sawaki, Fujita, Kitagawa, Nakai and Yamanaka1986) and recently improved by Fujioka et al. (Reference Fujioka, Zhang, Ishihara, Shigemori, Hironaka, Johzaki, Sunahara, Yamamoto, Nakashima, Watanabe, Shiraga, Nishimura and Azechi2013) at Gekko XII facility. More details about measurements of the ultrahigh magnetic fields were reported by Santos et al. (Reference Santos, Bailly-Gandvoux, Giuffrida, Forester-Colleoni, Fuijoka, Zang, Korneev, Boullaud, Dorand, Batani, Chervrrot, Cross, Crowston, Dubois, Gazave, Greofri, d'Humieres, Hulin, Ishihara, Kojima, Loyez, Marques, Nicolai, Peyrusse, Poye, Raffestin, Ribolzi, Roth, Schaumann, Serred, Tikhonchuk, Vacar and Woolsley2015). These magnetic fields with intensities of few kilotesla (kT), contained in few cubic millimeters and with time duration of few ns, generated with laser pulses above kJ, can be employed to trap the cylindrical plasma generated in the HB11 target for a fusion reaction in the coil of Figure 10.

Fig. 10. Generation of a 5 kT magnetic field in the coils (Fujioka et al., Reference Fujioka, Zhang, Ishihara, Shigemori, Hironaka, Johzaki, Sunahara, Yamamoto, Nakashima, Watanabe, Shiraga, Nishimura and Azechi2013) by firing a >kJ ns laser pulse 1 into the hole between the plates. The HB11 fusion fuel is coaxially located in the coils and the block ignition of the fusion flame is produced by the ps laser pulse for block ignition from laser 2.

The computation without involving secondary avalanche reactions – only with binary HB11 reactions – shows sufficient trapping by a cylindrical 10 kT magnetic field from fluid hydrodynamics (Lalousis et al., Reference Lalousis, Hora and Moustaizis2014; Reference Lalousis, Moustaizis, Hora and Miley2015).

The following two-fluid computations use a solid cylinder of 1 cm in length and 1 mm radius as a coaxial HB11 fuel target in the coil of Figure 10 where the generated α-particles are included as a third fluid using a magnetic field of 10 kT. The mean gyroradius of the α-particles from the HB11 reaction in a 10 kT field is about 43 µm. This is sufficiently small compared with the 1 mm radius of the cylinder to achieve strong magnetic field confinement of the alphas. The radial dependence at the time of 100 ps is shown in Figures 11–13 and the propagation of the fusion along the cylindrical axis results in a length x ~ 0.83 cm within 1 ns (Fig. 14). Owing to the low electron temperatures (Fig. 12) the cyclotron radiation generated in the cylindrical reaction volume (Lalousis et al., Reference Lalousis, Moustaizis, Hora and Miley2015) is <3 kJ/ns (Gulkis, Reference Gulkis and Meyers1987). Thus cyclotron radiation losses can be neglected.

Fig. 11. Solid HB11 cylinder of 1 mm radius irradiated by a ps-1020 W/cm2 laser pulse of 248 nm wavelength at concentric radius of 1 mm. After 100 ps, the dependence on the radius r is shown for the density of electrons N e, boron N b and hydrogen N h (sequence at r = 0 from above) using a 10 kT magnetic field (Lalousis et al., Reference Lalousis, Moustaizis, Hora and Miley2015).

Fig. 12. Same as Figure 11 with electron and ion temperatures.

Fig. 13. Same case as Figure 11: Solid HB11 cylinder of 1 mm radius irradiated by ps-1020 W/cm2 laser pulse of 248 nm wavelength. Radial density at 100 ps of primary produced α-particles N_a ending at 1 mm laser irradiation radius and the magnetic field strength depending on the radius r (in untouched fuel 10 kT at radius above 1 mm).

Fig. 14. HB11 fusion reaction rate at different depths x in 1D computation parallel to the magnetic field, at times after the ps generation of the fusion flame initiated by 1020 W/cm2, ps, 248 nm wavelength laser pulse.

Figure 13 shows the radial dependence of the α-particle density and of the magnetic field parallel to the cylindrical axis at the time of 100 ps. While the alphas are well confined to the initial radius of 1 mm and the magnetic field is modified at this periphery only due to E × B rotation, etc. For the propagation of the fusion flame along the axis of the cylinder, the one-dimensional (1D) computation of the reaction rate is shown in Figure 14. The total number of α-particles depending on the square of r has been increased considerably showing the successful ignition though only binary reactions were included in the computations. If one included the preliminary approximation of avalanche reactions the reaction result is rather modest taking into account that a 3 EW-ps laser was needed to reach a reaction gain in the range above 500.

If one could assume the additional avalanche process, a solid fuel cylinder of 1 cm length and 2 mm diameter block-ignited by a 3 MJ-ps laser pulse of 3 EW power will produce more than 1 gigawatt (GW) electrical energy (Lalousis et al., Reference Lalousis, Hora and Moustaizis2014) again with a gain only of about 300. The produced α-particles are gyrating with a radius of 43 µm which type of particle motion is not included in the hydrodynamics though most of the relevant plasma mechanisms were sufficient to such extend that the essential dynamic and confining processes are covered.

The same computations with a B11 fuel cylinder of 1 cm length and 0.2 mm diameter (Lalousis et al., Reference Lalousis, Hora and Moustaizis2014) resulted in some radial expansion of the plasma with binary reactions as shown in Figure 15 with the radial distribution of the α-particles during times up to 1000 ps. Though this was only for binary reactions, the ignition process is evident. For the case that avalanche reactions are included for irradiation with 1020 W/cm2 at the axis of 0.2 mm diameter on the HB11 solid fuel cylinder of 1 cm length and 1 mm radius, it is estimated that the laser pulse 2 (Fig. 10) of 30 kJ energy and 1 ps duration (30 PW power) is producing more than 1 GJ energy. The estimation is based on excluding the α gyration for the hydrodynamics. Though the genuine many-fluid codes takes care of highly non-equilibrium plasma conditions (Fortov & Lakubov, Reference Fortov and Lakubov1998; Davidson, Reference Davidson2001), it may be preferred to base further steps on specific experiments of HB11 reactions following the preceding measurements (Korn et al., Reference Korn, Margarone and Picciotto2014; Picciotto et al., Reference Picciotto, Margarone, Velyhan, Bellini, Krasa, Szydlowski, Bertuccio, Shi, Margarone, Prokupek, Malinowska, Krouski, Ullschmied, Laska, Kucharik and Korn2014; Margarone et al., Reference Margarone, Picciotto, Velyhan, Krasa, Kucharik, Mangione, Szydlowsky, Malinowska, Bertuccio, Shi, Crivellari, Ullschmied, Bellutti and Korn2015).

Fig. 15. α density N a depending on the radius r at different times (from lowest to highest curves for 100, 500 and 1000 ps, respectively) showing ignition from the increase of the curves on time (Lalousis et al., Reference Lalousis, Hora and Moustaizis2014).

6. POSSIBLE ECONOMIC CLEAN AND SAFE FUSION REACTOR

The experimental proof of the avalanche reactions for HB11 (Hora et al., Reference Hora, Lalousis and Moustaizis2014a; Korn et al., Reference Korn, Margarone and Picciotto2014) results in the scheme for combination of the non-linear force-driven plasma block ignition with ultrahigh accelerations together with the ultrahigh magnetic fields for cylindrical trapping of the fusion reaction. In the case for igniting 1 cm long cylinder of HB11 of 1 mm radius can then be based on of the irradiation by the laser pulse 2, Figure 10, with the laser beam diameter of 200 µm and an intensity of 1020 W/cm2 of a 30 PW pulse of ps duration. The computation of the α-particle generation – even very pessimistically using only binary reactions as in the DT case – as shown in Figure 16 with a 10 kT magnetic field results in the progress of the reaction gain.

Fig. 16. The HB11 fusion without a radioactive radiation problem is based on a block ignition Hora et al. (Reference Hora, Badziak, Read, Li, Liang, Liu, Sheng, Zhang, Osman, Miley, Zhang, He, Peng, Osman, Glowacz, Jablonski, Wolowski, Skladanowski, Jungwirth, Rohlena and Ullschmied2007) with a 30 kJ-ps laser pulse 2 (Fig. 10) where the solid hydrogen–boron fuel in the cylindrical axis of the magnetic coil is trapped by a 10 kT field sustained for about 1 ns after being generated by a laser pulse 1 of comparable duration. The central reaction unit (Fig. 10) is electric charged to the level of −1.4 million volts against the wall of a sphere producing α-particles (He nuclei) of more than a GJ energy, of which a small part is needed for the operation of the laser pulses. One part of the gained costs of electricity is needed for the apparatus of the central reaction and for the boron metal of the fuel being destroyed at each reaction (Hora et al., Reference Hora, Lalousis and Moustaizis2014a).

There is a comparably slow expansion of the growing α-particle density generated by the fusion reaction against the magnetic field. Including the avalanche reactions into the 1 ns reaction results in the production of considerably higher gain of 1 Gigajoule energy in the generated of initially mono-energetic generated α-particles generated by the laser pulse 2 (Fig. 10) needing only 30 kJ energy (Hora et al., Reference Hora, Lalousis, Eliezer, Miley, Moustaizis and Mourou2014b). On the basis of an avalanche process during up to 1 ns duration, the energy in the α-particles is more than 1 GJ (277 kWh). For conversion of most of the 2.99 MeV α-particle energy into electrostatic energy, the particles have to be slowed down by an electric field of −1.4 MV. This field has to be generated at each reaction between the central reaction unit against the spherical mantle of the reactor of Figure 16.

For a power station the main part of the generated energy of the α-particles from the level of a −1.4 MV voltage into polyphase alternating current can be performed as known from the megavolt-direct-current transmission line techniques Kanngiesser (Reference Kanngiesser, Huang and Lips1994). If the reactor works with a frequency of 1 Hz, the electric current for conversion is 714 Amp averaged between each fusion reaction.

7. SUMMARY

Combining the experimental result of most exceptionally high gains of proton–boron (HB11) fusion (Korn et al., Reference Korn, Margarone and Picciotto2014; Picciotto et al., Reference Picciotto, Margarone, Velyhan, Bellini, Krasa, Szydlowski, Bertuccio, Shi, Margarone, Prokupek, Malinowska, Krouski, Ullschmied, Laska, Kucharik and Korn2014; Margarone et al., Reference Margarone, Picciotto, Velyhan, Krasa, Kucharik, Mangione, Szydlowsky, Malinowska, Bertuccio, Shi, Crivellari, Ullschmied, Bellutti and Korn2015) based on the concluded avalanche mechanisms (Hora et al,. Reference Hora, Lalousis, Eliezer, Miley, Moustaizis and Mourou2014b; Reference Lalousis, Moustaizis, Hora and Miley2015) and evaluated now in more details, with the results of >PW-ps laser-driven block ignition (Hora et al., Reference Hora, Miley, Ghorannviss, Malekynia, Azizi and He2010; Lalousis et al., Reference Lalousis, Hora and Moustaizis2014) was leading to a first step of producing ultraclean GJ fusion energy possibly by 30 kJ laser irradiation. Cylindrical magnetic trapping with ultrahigh magnetic fields in the 10 kT range is necessary. The open treatment of these developments was approved (Hora, Reference Hora2014) to follow the route for designing of an economic long-lasting energy source. This furthermore has to be based on appropriate scientific and technological explorations.

The presented results are the work of more than 40 years research (Hora, Reference Hora1969; Reference Hora2000; Reference Hora2009; Hora et al., Reference Hora, Lalousis, Eliezer, Miley, Moustaizis and Mourou2014b; Reference Lalousis, Moustaizis, Hora and Miley2015) possible only after providing extremely high-power ps laser pulses (Strickland & Mourou, Reference Strickland and Mourou1985; Mourou et al., Reference Mourou, Barty and Perry1998) for demonstration of the predicted picoseocond block acceleration (Sauerbrey, Reference Sauerbrey1996) and ignition with ultrahigh magnetic fields (Fujioka et al., Reference Fujioka, Zhang, Ishihara, Shigemori, Hironaka, Johzaki, Sunahara, Yamamoto, Nakashima, Watanabe, Shiraga, Nishimura and Azechi2013). The key result is the here detailed evaluation of the highly increased reaction rate for HB11 fusion (Korn et al., Reference Korn, Margarone and Picciotto2014) as an avalanche process (Hora et al., Reference Hora, Lalousis, Eliezer, Miley, Moustaizis and Mourou2014b; Reference Lalousis, Moustaizis, Hora and Miley2015). This is not the end but may be the beginning of large-scale research to one of possible options for environmentally acceptable ultraclean and economic energy production.

References

REFERENCES

Banati, R.Hora, H., Lalousis, P. & Moustaizis, S. (2014). Ultrahigh laser acceleration of plasma blocks with ultrahigh ion density for fusion and hadron therapy. J. Intense Pulsed Laser Appl. Adv. Phys. 4, 1116.Google Scholar
Belyaev, V.S., Matafonov, A.P., Vinogradov, V.I., Krainov, V.P., Lisitsa, V.S., Roussetski, A.S., Ignatyev, G.N. & Andrianov, V.P. (2005). Observation of neutronless fusion reactions in picosecond laser plasmas. Phys. Rev. E 72, 026406.CrossRefGoogle ScholarPubMed
Betti, R., Chou, C.D., Anderson, K.S., Perkins, L.J., Theobald, W. & Sodolov, A.A. (2007). Shock ignition of thermonuclear fuel with high areal density. Phys. Rev. Lett. 98, 155001.CrossRefGoogle ScholarPubMed
Bobin, J.-L. (1974). Nuclear fusion reactions in fronts propagating in solid DT. (Schwarz, H. and Hora, H., Eds.) Laser Interaction and Related Plasma Phenomena, Vol. 4B, p. 465. New York: Plenum Press.Google Scholar
Boreham, B.W., Newman, D.S., Höpfl, R. & Hora, H. (1995). Depressed photoemission from Görlich cathodes at high laser light intensities. J. Appl. Phys. 78, 58485850.CrossRefGoogle Scholar
Butler, D. (1996). Now Europe's physicists seek shift in strategic fusion research. Nature 380, 655.CrossRefGoogle Scholar
Campbell, E.M. (2006). High intensity laser plasma interaction and application to inertial fusion and high energy density physics D.Sc. (Doctor of Science) Thesis. Sydney, Australia: University of Western Sydney, 723 p.Google Scholar
Chu, M.S. (1972). Thermonuclear reactions waves at high densities. Phys. Fluids 15, 412.CrossRefGoogle Scholar
Cicchitelli, L., Hora, H. & Postle, R. (1990). Longitudinal field components of laser beams in vacuum. Phys. Rev. A 41, 37273732.CrossRefGoogle ScholarPubMed
Daido, H., Miki, F., Mima, K., Fujita, M., Sawaki, K., Fujita, H., Kitagawa, Y., Nakai, S. & Yamanaka, C. (1986). Generation of a strong magnetic field by an intense CO2-laser pulse. Phys. Rev. Lett. 56, 846.CrossRefGoogle Scholar
Davidson, R. (2001). Nonneutral Plasmas. Singapore: World Scientific.CrossRefGoogle Scholar
Eliezer, S., Nissim, N., Martinez Val, J.M., Mima, K. & Hora, H. (2014). Double layer acceleration by laser radiation. Laser Part. Beams 32, 211217.CrossRefGoogle Scholar
Feynman, R. (1987). see H. Hora, Laser Plasma Physics, SPIE Books, Belllingham WA 2000, Section 6.3; Hora. H. Laser and Particle Beams 29, 275 (2011).Google Scholar
Földes, I.B., Bakos, J.S., Gal, K., Juhasz, Y., Kedves, M.A., Koscis, G., Szatmari, S. & Veres, G. (2000). Properties of high Harmonics generation by UV laser pulses on solid surfaces. Laser Phys. 10, 264.Google Scholar
Fortov, V.E. & Lakubov, (1998). Physics of Nonideal Plasmas. Singapore: World Scientific.Google Scholar
Fujioka, S., Zhang, Z.Ishihara, K., Shigemori, K., Hironaka, Y., Johzaki, T., Sunahara, A., Yamamoto, N., Nakashima, H., Watanabe, T., Shiraga, H., Nishimura, H. & Azechi, H. (2013). Kilotesla magnetic field due to a capacitor-coil target driven by high power laser. Sci. Rep. 3, 11701176.CrossRefGoogle ScholarPubMed
Gulkis, S. (1987). In Encyclopaedia of Physical Science and Technology (Meyers, R.A., Ed.), Vol. 11, p. 644. San Diego: Academic Press, Eq. (33).Google Scholar
Haan, S. (2010). see in Yuandi, Li, Editor, Nuclear Power without radioactivity Highlights in Chemical Technology, 24 March 2010 Royal Soc. Chemistry, London; or A. Calilasseria. Physics. Rational Skepticism Forum, see google “boron fusion”, (22 April 2010) http://www.physorg.com/news190295239.html.Google Scholar
Hohla, K. & Kompa, K.L. (1973). Gigawatt photochemical iodine laser. Appl. Phys. Lett. 22, 7780.CrossRefGoogle Scholar
Hora, H. (1969). Nonlinear confining and deconfining forces associated with the interaction of laser radiation with plasma. Phys. Fluids 12, 182191.CrossRefGoogle Scholar
Hora, H. (1981). Physics of Laser Driven Plasma. New York: Wiley, figures 10.18 a & b.Google Scholar
Hora, H. (1985). The transient electrodynamic forces at laser–plasma Interaction. Phys. Fluids 28, 37063707.CrossRefGoogle Scholar
Hora, H. (2000). Laser Plasma Physics. Belllingham, WA: SPIE Books.Google Scholar
Hora, H. (2003). Skin-depth theory explaining anomalous picosecond–terawatt laser plasma interaction II. Czech. J. Phys. 53, 199217.CrossRefGoogle Scholar
Hora, H. (2009). Laser fusion with nonlinear force driven plasma blocks: thresholds and dielectric effects. Laser Part. Beams 29, 207222.CrossRefGoogle Scholar
Hora, H. (2011). Distingished celebration for Professor George H. Miley by the University of Illinois, Urbana Illinois, USA. Laser Part. Beams 29, 275278.CrossRefGoogle Scholar
Hora, H. (2013 a). Extraodrinary jump of increasing laser fusion gains experienced at volume ignition for combination with NIF experiments. Laser Part. Beams 31, 229232.CrossRefGoogle Scholar
Hora, H. (2013 b). Neutron-free production of nuclear fusion reactions without additional fields. German Patent Application 10 2013 016 306.6Google Scholar
Hora, H. (2014). Permission for open use of German Patent application 10 2013 016 305.6. 15 Jan.Google Scholar
Hora, H., Badziak, J., Boody, F.P., Höpfl, R., Jungwirth, K., Kralikowa, B., Kraska, J., Laska, L., Parys, P., Perina, V., Pfeifer, M., Rohlena, K., Skala, J., Ullschmied, J., Wolowski, J. & Woryna, E. (2002 a). Effects of ps and ns. laser pulses for giant ion source. Opt. Commun. 207, 333337.CrossRefGoogle Scholar
Hora, H., Badziak, J., Read, M.N., Li, Y.-T., Liang, T.-J., Liu, H., Sheng, Z.-M., Zhang, J., Osman, F., Miley, G.H., Zhang, W., He, X., Peng, H., Osman, F., Glowacz, S., Jablonski, S., Wolowski, S., Skladanowski, Z., Jungwirth, K., Rohlena, K. & Ullschmied, J. (2007). Fast ignition by laser driven beams of very high intensity. Phys. Plasmas 14, 072701/1–7.CrossRefGoogle Scholar
Hora, H., Castillo, R., Clark, R.G., Kane, E.L., Lawrence, V.E., Miller, R.D.C., Nicholson-Florence, M.F., Novak, M.M., Ray, P.S., Shepanski, J.R. & Tsivinsky, I. (1978). Calculation of intertal confienement fusion gains using a collective model for reahat, bremsstrahlung and fuel depletion for higly efficient elecgtrodynamic laser compressions. Plasma Phys. Control. Nucl. Fusion Res. Innsbruck 1978 Nucl. Fusion (Suppl. 1979) 3, 237246.Google Scholar
Hora, H. & Kelly, J.C. (2009). Nuclear energy without radioactivity. Austr. Phys. 46, 111113.Google Scholar
Hora, H., Lalousis, P. & Eliezer, S. (1984). Analysis of the inverted double layers in nonlinear force produced cavitons at laser-plasma interaction. Phys. Rev. Lett. 53, 16501652.CrossRefGoogle Scholar
Hora, H., Lalousis, P., Eliezer, S., Miley, G.H., Moustaizis, S. & Mourou, G. (2014 b). New option for solving the climatic problems with non-thermal laser driven boron fusion and ultrahigh magnetic fields. arXiv 1412.4190, see also N. Byrne, E. Hunt and M. Beilharz (2014). News Release Austr. Physics Congress December 2014, .Google Scholar
Hora, H., Lalousis, P., Guiffrida, L., Margarone, D., Korn, G., Eliezer, S., Miler, G.H., Moustaizis, S. & Mourou, G. (2015) Petawatt laser pulses for proton-boron high gain fusion with avalanche reactions excluding problems of nuclear radiation. SPIE Proceedings 9515, 951518/1–15.Google Scholar
Hora, H., Lalousis, P. & Moustaizis, S. (2014 a). Fiber ICAN laser with exawatt–picosecond pulses for fusion without nuclear radiation problems. 32, 6368.CrossRefGoogle Scholar
Hora, H., Lalousis, P., Moustaizis, S., Földes, I., Miley, G.H., Yang, X., He, X.T., Eliezer, S. & Martinez-Val, J.-M. (2012 a). Shock Studies in Nonlinear Force Driven Laser Fusion with Ultrahigh Plasma Block Acceleration. IAEA Proc. Fusion Energy, San Diego October 2012. Paper IFE/P6–03, 8 pages (IAEA Vienna 2013) http://www-naweb.iaea.org/napc/physics/FEC/FEC2012/papers/27_IFEP603.pdf.Google Scholar
Hora, H., Malekynia, B., Ghoranneviss, M., Miley, G.H. & He, X. (2008). Twenty times lower ignition thresholds for laser driven fusion using collective effects and the inhibition factor. Appl. Phys. Lett. 93, 011101/1–3.CrossRefGoogle Scholar
Hora, H., Miley, G.H., Ghoranneviss, M., Malekynia, B. & Azizi, N. (2009) Laser-optical path to nuclear energy without radioactivity: fusion of hydrogen-boron by nonlinear force driven plasma blocks. Opt. Commun. 283, 41244126.CrossRefGoogle Scholar
Hora, H., Miley, G.H., Ghorannviss, M., Malekynia, H., Azizi, N. & He, X-T. (2010). Fusion energy without radioactivity: laser ignition of solid hydrogen-boron(11) fuel. Energy Environ. Sci. 3, 479486.CrossRefGoogle Scholar
Hora, H., Peng, H.-S., Zhang, W. & Osman, F. (2002 b). New skin depth interaction by ps-TW laser pulses and consequences for fusion energy. SPIE Proc. 4914, 4248.CrossRefGoogle Scholar
Hora, H., Sadighi-Bonabi, R., Yazdani, R., Afaideh, A., Nafari, F. & Ghoranneviss, M. (2012 b). Effect of quantum correction of the acceleration and delayed heating of plasma blocks. Phys. Rev. E 85, 036404/1–6.Google ScholarPubMed
Hurricane, O.A., Callahan, D.A., Casey, D.T., Celliers, P.M., Cerjan, C., Dewald, E.L., Dittrich, T.R., Döppner, T., Hinkel, D.E., Berzak Hopkins, L.F., Kline, J.L., Le Pape, S., Ma, T., MacPhee, A.G., Milovich, J.L., Pak, A., Park, H.-S., Patel, P.K., Remington, B.A., Salmonson, J.D., Springer, P.T. & Tommasini, R. (2014). Fuel gain exceeding unity in an inertially confined fusion implosion. Nature 506, 343348.CrossRefGoogle Scholar
Jungwirth, K., Cejnarova, A., Juha, L., Kralikova, B., Krasa, J., Krousky, E., Krupickova, P., Laska, L., Masek, K., Mocek, T., Pfeifer, M., Präg, A., Renner, O., Rohlena, K., Rus, B., Skala, J., Straka, P. & Ullschmied, J. (2001). The prague asterix laser system phys. Plasmas 8, 24952501.CrossRefGoogle Scholar
Kanngiesser, K.W., Huang, D.H. & Lips, H. (1994). Highvoltage direct current transmission – Systems and Planning. Siemens monographs Munich, EV HA 7.Google Scholar
Korn, G., Margarone, D. & Picciotto, A. (2014). Boron-Proton Nuclear Fusion Enhancement Induced in Boron-doped Silicon Targets by Low-contrast Pulsed Lasers. IZEST ELI-NP Conf. Paris, 17–18 September 2014.Google Scholar
Kouhi, M., Ghoraneviss, M., Malekynia, B., Hora, H., Miley, G.H., Sari, A.H., Azizi, N. & Razavipourm, S.S. (2011). Resonance effect for strong increase of fusion gains at thermal compression for volume ignition of Hydrogen Boron-11. Laser Part. Beams 29, 125.CrossRefGoogle Scholar
Krasa, J., Klier, D., Velyhan, A., Margarone, D., Krousky, E., Jungwirth, K., Skala, J., Pfeifer, M., Kravarik, J., Kubes, P., Rezak, K. & Ullschmied, J. (2013). Observation of repeitive bursts in emission of fast ions and and neutrons in sub-nanosecond laser-solid experiemnts. Laser Part. Beams 31, 395401.CrossRefGoogle Scholar
Labaune, C., Deprierraux, S., Goyon, S., Loisel, C., Yahia, G. & Rafelski, J. (2013). Fusion reactions initiated by laser acceleratedparticle beams in laser produced plasmas. Nat. Commun. 4, 2506.Google Scholar
Lalousis, P. & Hora, H. (1983). First direct electron and ion fluid computation of high electrostatic fields in dense inhomogeneous plasmas with subsequent nonlinear laser interaction. Laser Part. Beams 1, 283304.CrossRefGoogle Scholar
Lalousis, P., Hora, H. & Moustaizis, S. (2014). Optimized boron fusion with magnetic trapping by laser driven shock initiation at nonlinear force driven ultrahigh acceleration. Laser Part. Beams 32, 409411.CrossRefGoogle Scholar
Lalousis, P., Hora, H., Eliezer, S., Martnez-Val, J.-M., Moustaizis, S., Miley, G.H. & Mourou, G. (2013). Shock mechanisms by ultrahigh laser accelerated plasma blocks in solid density targets for fusion. Phys. Lett. A 377, 885888.CrossRefGoogle Scholar
Lalousis, P., Moustaizis, S., Hora, H. & Miley, G.H. (2015). Kilotesla magnetic assisted fast laser ignited Boron-11 hydrogen fusion with nonlinear force driven ultrahigh accelerated plasma blocks. J. Fusion Energy 34, 6267.CrossRefGoogle Scholar
Li, R.X., Liang, X.Y., Chu, Y.X., Gan, Z.B., Lu, H.H., Yu, L.H., Yin, D.J., Leng, Y.X., Lu, X.M., Wang, C. & Xu, Z.Z. (2015). Progress towards a 10 PW ultraintense laser facility: Demonstration of 5 PW high gain large aperture Ti:saphhire amplifier. 4th Advanced Lasers and Photon Sources (ALPS′15), Yokohama, Japan April 22–24.Google Scholar
Malekynia, B. & Razipur, S.S. (2013). Internal propagation of fusion flame with shock strong of laser driven plasma block for advanced nuclear fuels ignition. Chin. Phys. B22, 055202.CrossRefGoogle Scholar
Margarone, D., Picciotto, A., Velyhan, A., Krasa, J., Kucharik, M., Mangione, A., Szydlowsky, A., Malinowska, A., Bertuccio, G., Shi, Y., Crivellari, M., Ullschmied, J., Bellutti, P. & Korn, G. (2015). Advanced scheme for high-yield laser driven nuclear reactions. Plasma Phys. Control. Fusion 57, 014030.CrossRefGoogle Scholar
Mitchell, W. (2000). Keynote lecture at the ECLIM conference Schliersee/Germany.Google Scholar
Mourou, G., Barty, C.P.L. & Perry, M.D. (1998). Ultrahigh-intensity lasers: physics of the extreme on a tabletop. Phys. Today 51, 2230.Google Scholar
Mourou, G., Brocklesby, B., Tajima, T. & Limpert, L. (2013). The future is fibre accelerators. Nat. Photonics 7, 258261.CrossRefGoogle Scholar
Moustaizis, S., Lalousis, P. & Hora, H. (2013). A LIF scheme for HiPER application based on the combination of ultrahigh laser nonlinear force driven plasma blocks and the relativistic acceleration of ion blocks. In High Power, High Energy and Hig-intensity Laser Technology and Research using Extreme Light: Entering New Frontiers with Petawatt-class Lasers. (Hein, J., Korn, G. and Silva, L.O., Eds.). Proceedings of SPIE Vol. 8780, paper 878029/1–10.Google Scholar
Nevins, W.M. & Swain, R. (2000). The thermonuclear fusion rate coefficients for p-11B reactions. Nucl. Fusion 40, 865.CrossRefGoogle Scholar
Norreys, P.A., Fews, A.P.Beg, F.N., Bell, A.R., Dangor, A.E., Lee, P., Nelson, M.B., Schmidt, H., Tatarakis, M. & Cable, M.D. (1998). Neutron production from picosecond laser irradiation of deuterated targets at intensities of 1019 W/cm2 Plasma Phys. Control. Fusion 40, 175.CrossRefGoogle Scholar
Oliphant, M.L.E., Harteck, P. & Rutherford, L. (1934). 100 kilovolt discharges in deuterium plasmas. Proc. R. Soc. Lond. A 144, 692714.Google Scholar
Oliphant, M.L.E. & Rutherford, L. (1933). Experiments on the transmutations of elements by protons. Proc. R. Soc. Lond. A 141, 259281.Google Scholar
Paterson, A. (2015). Latest developments on small modular reactors. Bull. R. Soc. NSW No. 385, 4.Google Scholar
Picciotto, A., Margarone, D., Velyhan, A., Bellini, P., Krasa, J., Szydlowski, A., Bertuccio, G., Shi, Y., Margarone, A., Prokupek, J., Malinowska, A., Krouski, E., Ullschmied, J., Laska, L., Kucharik, M. & Korn, G. (2014). Boron-proton nuclear-fusion enhancement induced in boron-doped silicon targets by low-contrast pulsed laser. Phys. Rev. X 4, 031030.Google Scholar
Rostoker, N., Binderbauer, M.W. & Monkhorst, H.J. (1997). Colliding beam fusion reactor. Science 278, 14191422.CrossRefGoogle ScholarPubMed
Santos, J.J., Bailly-Gandvoux, M., Giuffrida, L., Forester-Colleoni, P., Fuijoka, S., Zang, Z., Korneev, Ph, Boullaud, R., Dorand, S., Batani, D., Chervrrot, M., Cross, J., Crowston, R., Dubois, J.-L., Gazave, J., Greofri, G., d'Humieres, E., Hulin, S., Ishihara, K., Kojima, S., Loyez, E., Marques, J.-R., Nicolai, Ph., Peyrusse, O., Poye, A., Raffestin, D., Ribolzi, J., Roth, M., Schaumann, G., Serred, F., Tikhonchuk, V.T., Vacar, Ph. & Woolsley, N. (2015). Laser-driven platform for generation and characterization of strong quasi-staticmagnetic fields. Phys. Plasmas arXiv 1503.00247/1–16.Google Scholar
Sauerbrey, R. (1996). Acceleration of femtosecond laser produced plasmas. Phys. Plasmas 3, 47124716.Google Scholar
Strickland, D. & Mourou, G. (1985). Compression of amplified chirped optical pulses. Opt. Commun. 56, 219222.CrossRefGoogle Scholar
Tabak, M., Hammer, J., Glinsky, M.N., Kruer, W.L., Wilks, S.C., Woodworth, J., Campbell, E.M., Perry, M.D. & Mason, R.J. (1994). Ignition and high gain with ultrapowerful lasers. Phys. Plasmas 1, 16261634.Google Scholar
Tahir, N.A. & Hoffmann, D.H.H. (1997). Development of advanced inertial fusion targets. Laser Part. Beams 15, 575587.CrossRefGoogle Scholar
Weaver, T., Zimmerman, G. & Wood, L. (1973). Exotic CTR Fuels: Non-thermal Effects and Laser Fusion Applications. Lawrence Livermore National Laboratory Report UCRL – 74938.Google Scholar
Yu, W., Xu, H., He, F., Yu, M.Y., Ishoguro, S., Zhang, J. & Wong, A.Y. (2005). Direct acceleration of solid-density plasma bunch by ultraintense laser. Phys. Rev. E 72, 046401.CrossRefGoogle ScholarPubMed
Zhang, M., He, J.T., Chen, D.B., Li, Z.H., Zhang, Y., Wang, L., Feng, B.L., Zhang, D.F., Tang, X.W. & Zhang, J. (1998). Effects of a prepulse on γ-ray radiation produced by a femtosecond laser with only 5-mJ energy. Phys. Rev. E 57, 37463748.CrossRefGoogle Scholar
Figure 0

Fig. 1. Measured highest DT fusion neutron gains N per incident laser energy (full signs) of laser pulses in the range of ns duration (Hora, 2013a) with the highest value reported (Hurricane et al., 2014) given by an asterisk.

Figure 1

Fig. 2. 1018 W/cm2 neodymium glass laser incident from the right-hand side on an initially 100 eV hot deuterium plasma whose density has a very low-reflecting bi-Rayleigh profile at initial time t = 1, resulting in a laser energy density and a velocity distribution from plasma hydrodynamic computations at an interaction time t = 1.5 ps of interaction. The driving non-linear force is the negative of the varying energy density of the laser field (E2 + H2)/8π expressing the intensity. The dynamic development of temperature and density had accelerated the plasma block of about 20 vacuum wavelength thickness of the dielectric enlarged skin layer moving against the laser (positive velocity) and another block into the plasma (negative velocity) showing ultrahigh >1020 cm/s2 acceleration.

Figure 2

Fig. 3. Scheme of skin depth laser interaction, where the non-linear force accelerates a plasma block against the laser light and another block toward the target interior. In front of the blocks are electron clouds of the thickness of the effective Debye lengths [Eq. (5)] (Hora et al., 2002b).

Figure 3

Fig. 4. Intensity dependence of the velocity of the plasma front from the Doppler line shift of the reflected from the irradiation of 700 fs KrF laser pulses on an aluminum target (Földes et al., 2000).

Figure 4

Fig. 5. Ion densities Ni depending of the depth X after a 1 ps 3 × 108 J/cm2 laser pulse of 248 nm wavelength has irradiated solid density DT. Parameter is the time after the ps interaction.

Figure 5

Fig. 6. Velocity V of the fusion flame at 2 ns after the initiating ps laser pulse of 248 nm wavelength depending on the energy flux E*.

Figure 6

Fig. 7. Longitudinal electric field E as in case of Figure 1 for E* = 108 J/cm2 depending of the depth X at times of 40 ps; 400 ps; 1 ns; 2 ns from the highest maximum, respectively.

Figure 7

Fig. 8. Reaction rates (to be multiplied by 1036 m−3/s) in solid DT at 1 ps pulse of energy flux E* = 3 × 108 J/cm2 KrF laser irradiation depending on the fuel depth × at different times up to 10 ns.

Figure 8

Fig. 9. Generation of a spherical shrinking laser pulse from radially directed fiber ends at a radius 1 to irradiate a spherical solid state fusion fuel 3 with the concentric grid 2 of positive electric charge for slowing down the generated α-particles from the HB11 reactions (Hora et al., 2014a).

Figure 9

Fig. 10. Generation of a 5 kT magnetic field in the coils (Fujioka et al., 2013) by firing a >kJ ns laser pulse 1 into the hole between the plates. The HB11 fusion fuel is coaxially located in the coils and the block ignition of the fusion flame is produced by the ps laser pulse for block ignition from laser 2.

Figure 10

Fig. 11. Solid HB11 cylinder of 1 mm radius irradiated by a ps-1020 W/cm2 laser pulse of 248 nm wavelength at concentric radius of 1 mm. After 100 ps, the dependence on the radius r is shown for the density of electrons Ne, boron Nb and hydrogen Nh (sequence at r = 0 from above) using a 10 kT magnetic field (Lalousis et al., 2015).

Figure 11

Fig. 12. Same as Figure 11 with electron and ion temperatures.

Figure 12

Fig. 13. Same case as Figure 11: Solid HB11 cylinder of 1 mm radius irradiated by ps-1020 W/cm2 laser pulse of 248 nm wavelength. Radial density at 100 ps of primary produced α-particles N_a ending at 1 mm laser irradiation radius and the magnetic field strength depending on the radius r (in untouched fuel 10 kT at radius above 1 mm).

Figure 13

Fig. 14. HB11 fusion reaction rate at different depths x in 1D computation parallel to the magnetic field, at times after the ps generation of the fusion flame initiated by 1020 W/cm2, ps, 248 nm wavelength laser pulse.

Figure 14

Fig. 15. α density Na depending on the radius r at different times (from lowest to highest curves for 100, 500 and 1000 ps, respectively) showing ignition from the increase of the curves on time (Lalousis et al., 2014).

Figure 15

Fig. 16. The HB11 fusion without a radioactive radiation problem is based on a block ignition Hora et al. (2007) with a 30 kJ-ps laser pulse 2 (Fig. 10) where the solid hydrogen–boron fuel in the cylindrical axis of the magnetic coil is trapped by a 10 kT field sustained for about 1 ns after being generated by a laser pulse 1 of comparable duration. The central reaction unit (Fig. 10) is electric charged to the level of −1.4 million volts against the wall of a sphere producing α-particles (He nuclei) of more than a GJ energy, of which a small part is needed for the operation of the laser pulses. One part of the gained costs of electricity is needed for the apparatus of the central reaction and for the boron metal of the fuel being destroyed at each reaction (Hora et al., 2014a).