Hostname: page-component-745bb68f8f-b95js Total loading time: 0 Render date: 2025-02-10T12:19:31.876Z Has data issue: false hasContentIssue false

On the inefficiency of hole boring in fast ignition

Published online by Cambridge University Press:  01 June 2004

P. MULSER
Affiliation:
Theoretical Quantum Electronics (TQE), Darmstadt University of Technology, Darmstadt, Germany
R. SCHNEIDER
Affiliation:
Theoretical Quantum Electronics (TQE), Darmstadt University of Technology, Darmstadt, Germany
Rights & Permissions [Opens in a new window]

Abstract

Hole boring and fast ignition seem to exclude each other: When there is hole boring, no ignition occurs, and vice versa. The laser beam pressure only causes a more or less deep cone-shaped critical surface that leads to better guidance of the beam and to improved laser–plasma coupling. At laser wavelengths of the order of 1 μm, successful fast ignition requires strong anomalous laser beam–pellet coupling.

Type
Research Article
Copyright
© 2004 Cambridge University Press

1. INTRODUCTION

For inertial confinement fusion (ICF) to work successfully a sufficient amount of precompressed DT fuel must be brought up to a temperature of 6–10 keV to induce a self-sustaining nuclear burn wave. This is either accomplished by central spark ignition, where the fast heating is provided by a converging spherical shock wave, or by fast ignition (FI), with a superintense laser beam during a time interval of 10–30 ps. The extension R of the hot spot is determined by the density ρ of the compressed fuel through ρR ≥ 0.3 gcm−2 (Lindl, 1995).

The concept of FI is attractive because it offers several advantages: (1) high burn efficiency (25%), (2) no special symmetry constraints, and (3) reduced growth of hydrodynamic instabilities. However, they contrast with the fact that the laser can deposit its energy at densities not exceeding the critical density ρc, respectively particle density nc. As a consequence, electrons of relativistic energies must provide for the energy transport to the dense compressed fuel in an efficient way, that is, good collimation, and on a time scale of several tens of picoseconds (e.g., 70 ps; Hain & Mulser, 2001). Essentially three scenarios have been proposed to achieve the goal: (1) fast beam ignition (FBI) by converting the relativistic electron beam into a focused beam of ions that is absorbed in a controlled way in the compressed pellet core; (2) cone-guided fast ignition (CFI), where a hollow high-Z cone (e.g., gold) is placed in the fusion pellet with its top reaching the core to bring the absorbing critical region very close to the ignition region (Kodama et al., 2001); (3) fast coronal ignition (FCI), with the critical density brought as close as possible to the compressed core by hole boring. This appealing scheme came up nearly 10 years ago and has meanwhile apparently been accepted as a realizable FI scenario. It is therefore surprising that a detailed study of FCI combined with hole boring has not been undertaken so far. The advantages of the scheme are evident. If, for instance, a laser beam can drill a hole in the pre-compressed pellet up to the density ρ ≃ 200 gcm−3 the energy spread of the hot electron beam could be significantly reduced and perhaps better coupling of the FI laser beam, now being guided by the walls of the hole, would be an additional advantage. By deep hole boring, the use of a cone could be avoided and CFI would reduce to FCI. In this article we show that, unfortunately, hole boring in connection with FI cannot work properly: FI and hole boring nearly exclude each other.

2. HOLEBORING IN COLD MATTER

Imagine impinging a laser beam of intensity

onto the critical density of a pre-compressed pellet of local pressure p0. The laser beam acts like a piston under pressure PA, which is the sum of the light pressure pL = (1 + R)I/c, where R is the reflection coefficient and c is the light speed, and the ablation pressure pA = pc + ρc vc2 = pc(1 + Mc2); hence, with μ = I/2cpc,

(Mulser & Bauer, 2004). pc and Mc are the plasma pressure and Mach number at the critical point. Equation (1) is obtained for a one-dimensional (1D) plasma flow and Te = const in the evanescent region whose extension, owing to profile steepening, is less than a quarter laser wavelength. Under the action of PA, a bow shock S traveling at speed v0 forms (Fig. 1a). In the infinitesimal time dt, the amount of matter ρ0 v0 dt undergoes the velocity change from v = 0 to v = v1, the density change from ρ0 to ρ1 and the pressure change from p0 to p1 when crossing S. With the compression ratio κ = ρ10 holds along the symmetry axis v1 = v0(1 − κ−1), and hence, the momentum balance reads

First pA = nc kTe(1 + Mc2) must be evaluated in relation to pL. The intensity interval of relevance is I = 1019–1021 W cm−2. The wavelengths considered are λ = 800 nm (Ti:Sa) and λ = 248 nm (KrF). The effective (i.e., relativistic) critical density nc = γeff nc0 is calculated from formulas given in the literature (Mulser & Bauer, 2004):

γth is defined by Eq. (4). For α/γth = 1 results γeff = (1 + 0.5)1/2γth, for α/γth ≃ 2 follows γeff = (1 + 0.8(α/γth)2)1/2γth; for (α/γth) > 5 γeff approaches the asymptotic value γeff = 0.785α. This behavior is of great help in preparing Table 1. The relativistic γ-factor used here is considerably different from simple standard expressions generally used.

Holeboring in cold matter: alternative models for matter displacement. a: Laser acts as impermeable piston; PA: piston pressure, ρ0: undisturbed pre-compressed pellet density, v0: hole boring speed. Oblique shock (b) is determined from the shock polar (c). Beyond limiting angle θ shock detaches from cone (d).

Ratio of ablation pressure pA to light pressure pL as a function of laser intensity I

A lower limit of the electron kinetic temperature Te is obtained from the requirement that under steady-state conditions the absorbed photon flux (1 − R)I must equalize the electron energy flux qe into the pellet. We set

Depending on the divergence of the electron beam qe, there is an uncertainty factor of order unity in this definition of Te. For comparison, if in the nonrelativistic expression for qe an isotropic Maxwellian distribution is used, the correct numerical factor is 2(2/π)1/2 = 1.6; instead for a collimated, that is, one-dimensional (1D), Maxwellian, it is (2π)−1/2 = 0.4. For the critical pressure we set pc = nc kTe. In the nonrelativistic case the pressure in one dimension is twice the internal energy 〈E〉, in three dimensions (isotropic) it is two thirds of it, regardless of whether the electron distribution is a Maxwellian or any other distribution. The relativistic case is more complicated. Fortunately it becomes simple again for superrelativistic electrons, that is, qe = kTe nc c if me c2 << kTe. Furthermore, in one dimension, p = nE〉 and in three dimensions, p = 1/3nE〉 holds (e.g., see the photon gas).

In Table 1 nc0, γeff, Mc, Te, and pA/pL are given for the intensities I = 1019, 2 × 1020, 2 × 1021 W cm−2 and for R = 0.5. An absorption of 50% is a realistic assumption (see, e.g., Feurer et al., 1997; Ruhl et al., 1999). The plasma pressure in the undisturbed pellet before the laser beam arrives is set equal to zero.

From Table 1 important information is extracted regarding the temperature Te, its dependence on the laser wavelength, and the pressure ratio pA/pL. Numerical FI studies do not give much chance to successful ignition for energy flux densities qe significantly less, for example, by a factor of 3, than qe = 1021 W cm−2 (Mulser & Bauer, 2004). At Ti:Sa wavelength and 1021 W cm−2 absorbed laser intensity, Te amounts to 5.3 MeV, even though the relativistic increase of nc0 is the considerable factor of 20. Because no effect is known so far which would lead to overdense light beam penetration, this temperature is a very reliable lower limit. Unless there occurs very anomalous collective coupling of such energetic electrons to the cold background, the electron beam is not stopped by the pre-compressed pellet and FI fails at λ [gsim ] 1 μm wavelength.

Table 1 underlines the importance of transport and collective deposition research at relativistic intensities. Switching to KrF lasers is a remedy; however it brings Te down only by a factor of 4 instead of the wavelength ratio squared, that is, 10.4, because with shorter wavelength the relativistic α-factor reduces also and hence γeff lowers from γeff = 20 to γeff = 9. One has to bear in mind that Te = 1.29 MeV is a lower limit. Future experiments have to show whether absorption takes place at nc, increasing according to Eq. (3), or whether, due to enormous profile steepening at the critical point, laser beam penetration follows a more modest density increase.

Another important aspect is that the hole boring bow shock always runs into a preheated region of pressure p0pc because the electron energy transport is faster than v0. Therefore the maximum effective pressure determining the hole boring speed v0 is PAp0 = pL + pc(1 + Mc2). For all intensities considered in Table 1 PApc is very close to pL.

The efficiency of hole boring in a cold background plasma (almost no energy exchange of the electrons from the critical region) may be studied by assuming a pellet density profile of the form ρ0 = h(x/L)s. From the pressure balance (2), PA = fρ0 v02 + pc, f = (1 − κ−1), the depth x/L as a function of t is given by

The pressure balance (2) also holds in the absence of a shock, that is, f = 0. Choosing I = 2 × 1021, R = 0.5, h = 400 gcm−3, t = 20 ps, f = 0.5 (corresponding to κ = 2, typical compression in numerical simulations), L = 0.01 cm and s = 2, one finds with PApc = 1.5 × 1.05pL (see Table 1) x/L = 0.45, ρ/h = 0.21, ρ = 83 gcm−3. With the exponent s = 3, x/L = 0.53, ρ/h = 0.15, ρ = 59 gcm−3. Now a typical “hole boring prepulse,” frequently quoted in the literature, of I = 1019 W cm−2, PApc = 1.5 × 1.2 (Table 1), s = 2, and t = 50 fs is considered: x/L = 0.23, ρ/h = 0.055, ρ = 22 gcm−3. After 20 ps, the laser has penetrated up to ρ = 8 gcm−3 only. Thus, at such an intensity hole boring is rather modest even in the most favorable case of cold plasma.

There remain some uncertainties originating from the laser pulse transverse intensity distribution and beam width, in the bow-shock model all summarized by the factor f. It is therefore illuminating to use an alternative model that is also accessible to an analytical treatment. As soon as the bow shock invades a density region ρ0 [gsim ] 1 gcm−3 (i.e., 5 times solid DT) the mass flows pc Mc and ρ0 v0 obey the inequality pc Mc << ρ0 v0. As a consequence, there is nearly stagnation on the axis in front of the piston P. This means that the laser acts nearly like an impermeable piston, deviating almost all plasma laterally. Therefore a solid cone-shaped piston of narrow aperture angle θ propagating at speed v0 is considered. From the vertex of the piston P a coaxial conical shock of aperture angle α originates (Fig. 1b). Under the assumption of negligible pressure p0, the possible states behind the shock cone (region 1) can be determined from the shock polar, which in Fig. 1c is given for κ = 4. When normalized to v0 it reads Vy2 = (1 − Vx)(Vx − κ−1). From the continuous transition of the tangential component through the shock front, that is, v0∥ = v1∥ (see, e.g., any textbook of gas dynamics), one derives

For angles θ such that the straight line starting from the origin does not have any point in common with the shock polar, the shock S forms at a finite distance from the cone with finite curvature on the axis (Fig. 1c), and Eq. (2) applies again. For a strong shock, equivalent to p0 = 0, the limiting angle is given by tan θ = (κ − 1)/2κ1/2. For κ = 4 results, tan θ = ¾ and θ = 37°. The corresponding angle α follows from Figure 1, that is, θ = 64°, and the angular factor in Eq. (6), corresponding to the f-factor of Eq. (5), assumes the value 0.32. For the cone and shock angles of θ = 28° and α = 40° indicated in Figure 1, the angular factor is 0.29. The factor decreases monotonically with decreasing aperture angle of the cone and, at given PAp0, v0 increases. For a strong shock of Figure 1a,c the factor becomes f = 1 − ¼ = 0.75. The comparison of the two models shows that a laser beam with a narrow cone-shaped intensity profile accelerates hole boring owing to an f-factor less than 0.5 in the denominator of Eq. (5): with f = 0.29, x/L increases by a factor of 1.15 (s = 2) and 1.11 (s = 3). The conclusion, also to the authors' surprise, is that even in the most favorable situation of cold matter, that is, collimated hot electrons interacting with the dense pellet core only, hole boring at all relevant laser beam intensities is not very efficient.

3. HOLEBORING IN HOT MATTER

3.1. Numerical results

First of all, there is the experience with computer simulations of FI on the absence of any hole boring. Specifically, a 5-mg DT pellet, pre-compressed to 350 gcm−3, was successfully ignited by a 75 kJ laser beam of absorbed intensity of I = 1021 W cm−2, deposited at the density ρ = 4.6 gcm−3 during 20 ps (Mulser & Bauer, 2004). The energy transport was accomplished by flux-limited Spitzer-diffusive heat conduction, and R = 0, that is, no reflection, was set. During the entire time of irradiation no indication of hole boring or concave deposition surface deformation could be observed. It can be concluded with certainty that in a diffusive model for qe any hole boring under the above conditions stops at a density ρ not higher than 5 gcm−3. The reason is that, due to pellet heating, the pressure p0 balances the light pressure pL. From other computer runs (Hain & Mulser, 2001) with 3-mg DT pellets, successfully ignited by I = 2 × 1020 W cm−2, 12–16 kJ “free ignition energy,” deposited at 3–5 gcm−3 over 20 ps and diffusively transported to the dense core without a flux limiter (therefore the low ignition energy), it can be safely concluded that hole boring or crater formation stopped at densities ρ < 3 gcm−3. It was a common experience of the authors that, rather than a tendency to crater formation, there was a tendency to push the laser piston back, away from the deposition region, owing to a fast increase of the counterpressure p0.

3.2. Analytical considerations

From the ignition condition ρR = 0.3, at ρ = 400 gcm−3 a radius R = 7.5 μm results. The energy needed for ignition is typically 15 kJ (Atzeni, 1999; Mulser & Bauer, 2004). Let us now assume that the compressed core is heated by an electron beam of energy flux density qe through one spot hemisphere during the time τ = 20 ps. The energy balance 2πR2qeτ = 15 kJ leads to qe = 2.1 × 1020 W cm−2. Surprising enough, this value agrees with the hydrodynamic beam calculations without flux limit (Hain & Mulser, 2001; see Fig. 2 therein). The comparison with the ignition energy of 75 kJ when the heat flux limit is introduced shows that in a diffusive model, most of the FI energy is not supplied to the compressed core during the phase of high compression. The location where the flux limit plays its decisive role is the high density jump behind the laser deposition region. To transport the necessary FI energy flux qe = 2.1 × 1020 W cm−2 to the compressed core of radius R = 7.5 μm with

for DT (Z = 1; Braginskii 1966), Te must be as high as Te = 34 keV, in good agreement with the numerical simulations.

Direct heating of the pellet core by electrons produced in the critical region requires a mean free path λe = (2nDTσ)−1, where σ is a Coulomb cross section, of the order of 2R. Because the moderately relativistic differential cross section σΩ is σΩ = γ−2σR (Sakurai, 1967), where σR is the Rutherford cross section, the total cross section σ for momentum transfer is

where λD is the Debye length, λB is the reduced de Broglie wavelength, and

For the ratios u = 1.0, 1.5, and 2.0, me c2 = 511 keV, the Coulomb logarithms for ρDT = 400 gcm−3 (corresponding to nDT = 1026 cm−3) and Te = 40 keV are ln Λ = 5.6, 5.2, and 4.9, and the mean free paths λe amount to

Hence, it can be concluded that electrons of kTe > 1.5 me c2 = 770 keV are not stopped by an ICF pellet of standard dimensions (here: 5 mg mass and ρmax = 400 gcm−3) and direct FI with fast electrons produced by a laser of λ [gsim ] λTi:Sa fails (see Table 1) if binary collisions with cold electrons and ions are the only interaction mechanism. At λ = λKrF there seems to exist a possibility for FI on the basis of binary transport mechanisms as considered until now. However, one must be aware of the fact that the estimates for kTe in Table 1 are minimum values. In addition, numerical FI studies with collimated electron beams show that the required beam energies are closer to 100 than to 10 kJ (Hain, 1999).

Owing to the lack of knowledge on energy transfer from the critical region to the target interior, hole boring can be analyzed only on the basis of more or less reasonable assumptions which, at the same time, are basic enough to allow us to come to valid conclusions.

3.2.1. Flux-limited Spitzer model

It is assumed that the energy transport from the absorption region occurs diffusively. Owing to the low collisionality qe is nonlocal and strongly inhibited. The evolution of p0 = n0 kTe0 is evaluated for qe = I(1 − R) = 1021 W cm−2 = const over τ = 20 ps at λ = 248 nm (KrF). In a plasma of constant density, the heat diffusion equation

can be integrated for qe = κ0Tmax5/2Tmax/∂x = const under the approximation 〈Te〉 = Tmax. By observing that qe = cVTexF, a straightforward calculation leads to the evolution of the heat front position xF(t) as follows:

Interestingly enough, if ∂Te /∂x is approximated by Te /xF(t), an expression for xF(t) is obtained in which only the numerical factor (9/28)2/9 = 0.78 is replaced by 1.0 (Zel'dovich & Raizer, 1967). Owing to the strong temperature dependence of qe, the rectangular temperature profile T(x,t) is approximately preserved (or, at least, remains self-similar) when the plasma is inhomogeneous. By setting, for convenience,

, with

(h, L from (5)),

, and integrating xF = 7xF/9t, xF(t) is deduced for variable specific heat:

Its derivative

is consistent with

if the second term in the bracket is much larger than 1. This is generally fulfilled except for irrelevant, very short times (t < 2 × 10−13 s in the following). xF(t), 〈Te〉(t), and p0(t) are evaluated now for qe = 1021 W cm−2 at KrF wavelength. From Table 1 and Eq. (8), the ratio κλe = λe /Lc ≃ 103 is found (λe is the electron mean free path, Lc = 1/κ is the scale length at critical density, κ = 640 cm−1) for which, according to Tahraoui and Bendib (2002) a heat flux inhibition factor fq = 1.2 × 10−4 results, that is, in (12) κ0 = 2 × 10−6 [cgsK] has to be replaced by κ0 = 2 × 10−6fq = 2.4 × 10−10. With these parameters, Table 2 is obtained. In the neighborhood of the critical density the shock wave propagates at a speed v0c/20 if f = 0.5 is set. Thus, the heat wave preheats the undisturbed pellet density creating in this way the counterpressure p0. From Table 2 it follows that hole boring stops at a time below 0.5 ps at the negligible depth of approximately 4–5 μm.

Thermal wave front xF, electron temperature 〈Te〉, and electron pressure p0 near xF for absorbed laser flux qe = 1021 W cm−2 (KrF)

The hole boring pressure here is PApL = (1 + R)I/c = 1 Tbar. PA scales like I, p0 roughly like αqe /xFI1/2I4/9 < I. Owing to the monotonic decrease of fq with decreasing laser flux density, p0Iδ, with

, hole boring reduces with decreasing intensity of irradiation. Hence, in agreement with the numerical simulation for I [lsim ] 1021 W cm−2, hole boring is suppressed when flux-limited Spitzer–Braginskii-type energy transport takes place. It should be noted that fq used here (and all similar fq from the published literature) is strictly valid for electron distribution functions close to equilibrium, a condition which may not be fulfilled at superhigh laser intensities.

3.2.2. Anomalous electron beam stopping

At laser wavelengths of the order of 1 μm (Nd, Ti:Sa, iodine) and beam energies in the multikilojoule range the bulk of the electrons acquires energies too high to be slowed down by a standard-size ICF pellet. The situation may change only if strongly enhanced collective stopping occurs somewhere in the pellet. Fortunately there are indications of such an anomalously intensified energy exchange, from numerical simulations (Ruhl, 2002; Sentoku et al., 2002), as well as from analytical and semianalytical modeling (Das & Kaw, 2001; Jain et al., 2003). Anomalous effects are expected to be bounded to the neighborhood of the critical region and to decay in the more remote dense zones of the pellet owing to collisional damping and the absence of forces driving instabilities to a high saturation level.

The alternative model in this section assumes that the more or less collimated electron jets from Table 1 are subject to collective stopping over a length of perhaps a few tens of microns. According to Das and Kaw (2001) and Jain et al. (2003) a quasilinear friction term νeff of the order of

acts on the fast electrons. At intensities I ≃ 1021 W cm−2 a self-generated magnetic field of 1 GGauss, corresponding to an electron cyclotron frequency ωc = 2 × 1016 s−1, may be assumed. Hence νeff [gsim ] 1014 s−1 will result and a relativistic electron moving at speed c thermalizes over a few λe ≃ 3 μm, typically 3λe = 10 μm. On the basis of present understanding this means that Eqs. (12) are applicable with a flux inhibition factor fq ≃ 10−4, after thermalization has been achieved, to fluxes qe = 1021 at Ti:Sa laser wavelength too. Owing to the short thermalization time τ < 10−13 s−1 Table 2 does not change for KrF. For Ti:Sa beams, slightly higher values for xF, 〈Te〉, and p0 are calculated, for example, xF = 43.6 μm, 〈Te〉 = 1.7 MeV, p0 = 2.7 Tbar at t = 5 × 10−13 s−1.

4. CONCLUSION

The foregoing analysis on fast ignition of a fusion pellet with super-intense laser beams has revealed two important aspects:

  1. The energy requirements are such that the average kinetic energy of the electrons, created at laser wavelengths λ ≃ 1 μm (Nd, Ti:Sa, iodine), is too high to allow stopping in the pre-compressed pellet by binary collisions. In this situation of weak electron–pellet interaction, moderate hole boring, induced mainly by light pressure, takes place but pellet ignition fails. At laser beam wavelengths λ [lsim ] λNd/4 (KrF) pellet ignition cannot be categorically excluded on the basis of collisional interaction only. However, then very quickly a thermal pressure exceeding the light pressure is generated and hole boring is stopped. Numerical studies confirm the analytical findings.
  2. Fast ignition with acceptable laser energies in the infrared (and perhaps in the near UV too) requires anomalous, that is, collective, electron beam–pellet interaction somewhere in the moderately compressed pellet region, with the consequence that hole boring is suppressed again.

Hence, on the basis of present knowledge on electron beam–plasma coupling, we conclude that with fast ignition with energies not exceeding 100 kJ considerably, fast ignition and hole boring exclude each other.

ACKNOWLEDGMENT

Discussion of some special gas dynamic aspects with M. Kanapathipillai is gratefully acknowledged.

References

REFERENCES

Atzeni, S. (1999). Inertial fusion fast ignitor: Igniting pulse parameter window vs the penetration depth of the heating particles and the density of the precompressed fuel. Phys. Plasmas 6, 33163326.Google Scholar
Braginskii, S.I. (1966). Transport processes in plasmas. In Review of Plasma Physics (Leontovich, M.A., Ed.) Vol. 1. New York: Consultants Bureau, p. 209.
Das, A. & Kaw, P. (2001). Non-local sausage-like instability of current channels in electron magnetohydrodynamics. Phys. Plasmas 8, 45184523.Google Scholar
Feurer, T., Theobald, W., Sauerbrey, R., Uschmann, I., Altenbernd, D., Teubner, U., Gibbon, P., Forster, E., Malka, G. & Miquel, J.L. (1997). Onset of diffuse reflectivity and fast electron flux inhibition in 528-nm-laser solid interactions at ultrahigh intensity. Phys. Rev. E 56, 46084614.Google Scholar
Hain, S. (1999). Propagation of Intense Laser Radiation through Matter. Herdecke, Germany: GCA-Verlag (in German); p. 128.
Hain, S. & Mulser, P. (2001). Fast ignition without hole boring. Phys. Rev. Lett. 86, 10151018.Google Scholar
Jain, N., Das, A., Kaw, P. & Sengupta, S. (2003). Nonlinear electron magnetohydrodynamic simulations of sausage-like instability in magnetic channels. Phys. Plasmas 10, 2936.Google Scholar
Kodama, R., Norreys, P.A., Mima, K., Dangor, A.E., Evans, R.G., Fujita, H., Kitagawa, Y., Krushelnick, K., Miyakoshi, T., Miyanaga, N., Norimatsu, T., Rose, S.J., Shozaki, T., Shigemori, K., Sunahara, A., Tampo, M., Tanaka, K.A., Toyama, Y., Yamanaka, Y. & Zepf, M. (2001). Fast heating of ultrahigh-density plasma as a step towards laser fusion ignition. Nature 412, 798802.Google Scholar
Lindl, J. (1995). Development of the indirect-drive approach to inertial confinement fusion and the target physics basis for ignition and gain. Phys. Plasmas 2, 39334024.Google Scholar
Mulser, P. & Bauer, D. (2004). Fast ignition of fusion pellets with superintense lasers: Concepts, problems, and prospectives. Laser and Part. Beams 22, 512.Google Scholar
Ruhl, H. (2002). 3D kinetic simulation of super-intense laser-induced anomalous transport. Plasma Sources Sci. Technol. 11, A1A5.Google Scholar
Ruhl, H., Macchi, A., Mulser, P., Cornolti, F. & Hain, S. (1999). Collective dynamics and enhancement of absorption in deformed targets. Phys. Rev. Lett. 82, 20952098.Google Scholar
Sakurai, J.J. (1967). Advanced Quantum Mechanics. Reading, MA: Addison-Wesley Publishing Co. p. 193.
Sentoku, Y., Mima, K., Sheng, Z.M., Kaw, P., Nishihara, K & Nishikawa, K. (2002). Three-dimensional particle-in-cell simulations of energetic electron generation and transport with relativistic laser pulses in overdense plasmas. Phys. Rev. E 65, 046408.Google Scholar
Tahraoui, A. & Bendib, A. (2002). Collisional heat flux in laser heated plasmas. Phys. Plasmas 9, 30893097.Google Scholar
Zel'dovich, Ya.B. & Raizer, Yu.P. (1967). Physics of Shock Waves and High-Temperature Hydrodynamic Phenomena. New York: Academic Press, Chap. X.
Figure 0

Holeboring in cold matter: alternative models for matter displacement. a: Laser acts as impermeable piston; PA: piston pressure, ρ0: undisturbed pre-compressed pellet density, v0: hole boring speed. Oblique shock (b) is determined from the shock polar (c). Beyond limiting angle θ shock detaches from cone (d).

Figure 1

Ratio of ablation pressure pA to light pressure pL as a function of laser intensity I

Figure 2

Thermal wave front xF, electron temperature 〈Te〉, and electron pressure p0 near xF for absorbed laser flux qe = 1021 W cm−2 (KrF)