• Matter and Radiation at Extremes
  • Vol. 10, Issue 2, 027403 (2025)
Zhantao Lu1,2,*, Xinglong Xie1,2, Xiao Liang1, Meizhi Sun1..., Ping Zhu1, Xuejie Zhang1, Linjun Li1,2, Hao Xue1,2, Guoli Zhang1,2, Rashid Ul Haq1,2, Dongjun Zhang1 and Jianqiang Zhu1,2|Show fewer author(s)
Author Affiliations
  • 1National Laboratory on High Power Laser and Physics, Shanghai Institute of Optics and Fine Mechanics, Chinese Academy of Sciences, Shanghai 201800, People’s Republic of China
  • 2Center of Materials Science and Optoelectronics Engineering, University of Chinese Academy of Sciences, No. 19(A), Yuquan Road, Shijingshan, Beijing 100049, People’s Republic of China
  • show less
    DOI: 10.1063/5.0235138 Cite this Article
    Zhantao Lu, Xinglong Xie, Xiao Liang, Meizhi Sun, Ping Zhu, Xuejie Zhang, Linjun Li, Hao Xue, Guoli Zhang, Rashid Ul Haq, Dongjun Zhang, Jianqiang Zhu. Effect of laser wavelength on growth of ablative Rayleigh–Taylor instability in inertial confinement fusion[J]. Matter and Radiation at Extremes, 2025, 10(2): 027403 Copy Citation Text show less

    Abstract

    The effect of drive laser wavelength on the growth of ablative Rayleigh–Taylor instability (ARTI) in inertial confinement fusion (ICF) is studied with two-dimensional numerical simulations. The results show that in the plasma acceleration phase, shorter wavelengths lead to more efficient coupling between the laser and the kinetic energy of the implosion fluid. Under the condition that the laser energy coupled to the implosion fluid is constant, the ARTI growth rate decreases as the laser wavelength moves toward the extreme ultraviolet band, reaching its minimum value near λ = 65 nm, and when the laser wavelength continuously moves toward the X-ray band, the ARTI growth rate increases rapidly. It is found that the results deviate from the theoretical ARTI growth rate. As the laser intensity benchmark increases, the position of the minimum ARTI growth rate shifts toward shorter wavelengths. As the initial sinusoidal perturbation wavenumber decreases, the position of the minimum ARTI growth rate shifts toward longer wavelengths. We believe that the conclusions drawn from the present simulations and analysis will help provide a better understanding of the ICF process and improve the theory of ARTI growth.

    I. INTRODUCTION

    Inertial confinement fusion (ICF)1–3 has attracted significant attention in recent years as a potential method for generating fusion energy. In ICF, a fuel target is compressed to extremely high density and temperature by high-energy lasers or particle beams to trigger the fusion reaction. To achieve better ICF implosion performance, it is necessary to improve the efficiency of coupling between laser energy and target and to control hydrodynamic instabilities,4–6 primarily Rayleigh–Taylor instability (RTI).7,8 RTI occurs on the outer or inner surface of the target during the acceleration or deceleration phase of the implosion in ICF. When the injected energy is deposited on the target, the surface shell absorbs this energy and converts it into a high-temperature and low-density plasma that is ejected outward at high velocity. This ablation process produces a high-pressure area, which drives the fuel to accelerate and compress inward. At the acceleration interface, RTI is triggered by the presence of different gradients of density and pressure. Meanwhile, perturbations at internal interfaces grow, coupled with ablation surface perturbations, leading to material mixing, which can severely damage the integrity of the target shell and affect the compression of the fuel and the efficiency of nuclear fusion reactions.9

    The linear theory of single-mode classical RTI shows that an initial sinusoidal perturbation with wavelength λp grows exponentially with growth rate γCRTI=ATkpg, where g is the acceleration, kp = 2π/λp is the perturbation wavenumber, and AT = (ρhρl)/(ρh + ρl) is the Atwood number. Here, ρh and ρl are the densities of the heavy and light fluids. With increasing mode amplitude, the perturbed interface evolves into an asymmetric “bubble” and “spike,” and the exponential growth of the fundamental mode saturates. According to the classical theory, when the mode amplitude reaches about 0.1λp, the phase of linear growth comes to a halt and the phase of nonlinear growth begins.10,11 It is well known that the ablation process reduces the linear growth rate of RTI, especially when the perturbation wavelength is so short that it approaches the cutoff wavelength. In addition, ablation can significantly reduce the values of the second- and third-order mode coupling coefficients, thereby slowing down the generation of second and third harmonics,12,13 which makes the saturation amplitude of the ablative RTI (ARTI) significantly greater than 0.1λp. The linear growth rate of ARTI can be approximated by the Takabe formula14,15γARTI=αkpgβkpVa, where Va=ṁ/ρa is the ablation velocity, ṁis the mass ablation rate, ρa is the density of unablated shell near the ablation front, and α and β depend on the flow parameters. It should be noted that the Takabe formula is applicable in the case of a sharp ablation front16 (kpLm << 1, where Lm is the characteristic length of the minimum density gradient near the ablation front), and there are large deviations from this formula when the perturbation wavelength is short and the characteristic width of the ablation front L0 is large.17 For this reason, Lindl proposed an improved formula18,19γ=kpg/(1+kpLm)βkpVa that includes the finite thickness of the ablation front. More complex and accurate self-consistent stability theories20–25 have been developed for the limits of large or small kpL0 and large or small Froude number Fr=Va2/gL0.

    The dispersion relation for a laser in a plasma is c2k2=ω2ωpe2, where ωpe=(4πnee2/me)1/2 is the plasma frequency. At the location where the plasma frequency equals the laser frequency, the laser is reflected by the plasma, and the electron density at this point is called the critical density nc. Generally, nc = [1.115/λ(nm)2]cos θ × 1027cm−3,where θ is the incidence angle of the laser.26 The electromagnetic wave power P is depleted by the inverse bremsstrahlung process,27 and the power loss is dP/dt = −νib(t)P, where νib = (ne/ni)νei and νei is the frequency of electron–ion collisions. For direct-drive ICF, the typical laser wavelength is 351 nm, and the efficiency of coupling of the laser energy to the kinetic energy of the implosion fluid is about 6%.28–30 For indirect-drive ICF, lasers are incident into a Hohlraum composed of high-Z material, and the laser energy is converted to an X-ray radiation field with a particular energy spectral distribution.31 The energy absorbed by the target ultimately needs to be converted into kinetic energy of the implosion fluid, with the remaining energy consumed to maintain the isothermality of the corona plasma and to ablate cold material. The efficiency of conversion to kinetic energy of the implosion fluid depends on the laser wavelength.

    It is well known that in the case of direct-drive ICF, hydrodynamic instability is usually inevitable, owing to the crossing and coupling of the driving lasers. Additionally, the validity of ARTI growth theory at short laser wavelengths has not been verified. In this paper, to better understand and suppress hydrodynamic instability in ICF experiments and investigate how changes in laser wavelength affect ARTI growth, we use numerical simulations to study the effect of laser wavelength on the growth of ARTI. We adjust the intensity to keep the kinetic energy of the implosion fluid constant and thus obtain the relative coupling efficiency under an intensity benchmark. We then investigate laser action on a target for different laser wavelengths and the corresponding intensities. The simulation results show that the ARTI growth rate has its minimum in the extreme ultraviolet band, which represents a deviation from the theoretical behavior of the ARTI growth rate. Finally, the effects of the intensity benchmark and the perturbation wavenumber are investigated.

    The remainder of the paper is organized as follows. In Sec. II, the simulation methods and configurations are introduced. In Sec. III, the effects of varying laser wavelength on the growth of ARTI and laser coupling efficiency are presented and discussed. In Sec. IV, cases with different laser intensity benchmarks are presented and discussed. In Sec. V, cases with different perturbation wavenumbers are presented and discussed. Finally, the paper concludes with a summary in Sec. VI.

    II. SIMULATION SETUP

    For the simulations, we use the radiation magnetohydrodynamics code FLASH,32,33 which can handle plasmas with multiple temperatures to simulate laser-driven high-energy-density physics (HEDP) experiments. The relevant equations describing the evolution of a nonmagnetic three-temperature (3T) plasma are as follows:ρt+(ρu)=0,t(ρu)+(ρuu)+Ptot=0,t(ρeion)+(ρeionu)+Pionu=ρcv,eleτei(TeleTion),t(ρeele)+(ρeeleu)+Peleu=ρcv,eleτei(TionTele)qele+QabsQemis+Qlas,t(ρerad)+(ρeradu)+Pradu=qradQabs+Qemis.Here, ρ is the fluid density, u is the fluid velocity, Ptot= Pele+ Pion+ Prad is the total pressure, τei is the electron–ion energy relaxation time, Tele and Tion are the temperatures of the electrons and ions, eion, eele, and erad are the energies of the ions, electrons, and radiation field, cv,ele is the electron heat capacity, qele and qrad are the heat conduction fluxes of the electrons and the radiation field, Qlas is the laser energy deposition, Qabs = σ(r, λ, t)I(r, λ, t) is the absorption of radiation by electrons, where σ(r, λ, t) is the opacity of the material and I(r, λ, t) is the radiation intensity, and Qemis = 4πσ(r, λ, t)B(T, λ) is the emission of radiation by electrons, where B(T, λ) is the Planck function. The pressure, density, temperature, and internal energy are related by the equation of state.

    The control equations are solved on a uniform grid, using the fifth-order weighted essentially non-oscillatory (WENO) method34 in an unsplit hydrodynamics solver.35 The laser energy deposition is calculated using the EnergyDeposition unit of FLASH, the 3T hydrodynamics update is performed using the Hydro unit, the 3T equation of state is calculated using the Eos unit, ion/electron equilibration is implemented using the Heatexchange unit, the implicit diffusion solver is implemented using the Diffuse unit, the electron thermal conductivity is calculated using the Conductivity unit, the vanLeer slope limiter36 is used to deal with shock waves, and a second-order accurate HLLC Riemann solver37 is used to calculate time-advanced fluxes. A Courant–Friedrichs–Lewy (CFL)38 number of 0.4 is adopted to ensure numerical stability.

    To obtain a steady-state flow field, we use a laser with wavelength 351 nm and intensity39,40 1 × 1014 W/cm2 to simulate the laser-driven ablation of a one-dimensional planar CH target of thickness 200 μm and density 1 g/cm3. After a few nanoseconds, the ablation surface reaches uniform acceleration motion, meaning that the state is steady.41 We then extend the steady-state flow field into two dimensions and introduce perturbations. Figure 1(a) shows the initial profiles of density, temperature, and velocity along the z axis. Figure 1(b) shows the simulation setup for the initial density profile. The upper part is the cold and dense unablated material, while the lower part is the hotter but less dense ablated plasma. The size of the simulation domain is 25.6 × 300 μm2, with a grid size of 0.1 μm,42,43 and periodic boundary conditions are set at the left and right boundaries. An initial single-mode perturbation with amplitude η = η0 cos(kpx), where η0 = 0.1 μm and kp = 2π/12.8 μm−1, is introduced at the ablation front, which is realized by shifting the one-dimensional fluid when it is extended to two dimensions. A laser with a varying wavelength but a constant intensity and a constant duration of 4 ns is vertically incident from the bottom.

    (a) Initial profiles of ρ (solid line), T (dot-dashed line), and vz (dashed line) along the z axis. (b) Simulation setup for initial density profile.

    Figure 1.(a) Initial profiles of ρ (solid line), T (dot-dashed line), and vz (dashed line) along the z axis. (b) Simulation setup for initial density profile.

    III. EFFECT OF LASER WAVELENGTH

    Studies have revealed that the coupling efficiency between laser and target during the implosion acceleration phase depends on the laser wavelength. To control variables, we take the initial laser parameters as the benchmark (with an intensity of 1 × 1014 W/cm2 and a wavelength of 351 nm) and then change the laser wavelength and adjust the intensity so that the state (implosion velocity and position of the ablation surface) of the target is consistent after 4 ns of laser action, as shown in Fig. 2. The ratio of the laser intensity used in the simulation to the benchmark is taken as the relative coupling efficiency between the input laser and the kinetic energy of the implosion fluid, as shown in Fig. 3. Here, we compare the results obtained by this method (red circles) with those from simulations and three analytic models (models A, B, and C) provided in Ref. 44. When λ ≥ 200 nm, our results match well with the simulation results from Ref. 44 (blue crosses), and when λ < 200 nm, they are in closer agreement with those of model B (green downward-pointing triangles). We believe that since the simulation results in Ref. 44 were fitted using λ = 193, 248, 351, and 527 nm, they are more accurate within this wavelength range than those of the model. For λ < 193 nm, the assumption used in model B that the laser is absorbed by inverse bremsstrahlung before the critical surface is appropriate. According to Fig. 3, the shorter the laser wavelength, the higher is the relative coupling efficiency. Here, the laser energy is acting directly on the target. When λ = 10 nm, the electromagnetic wave begins to enter the X-ray band, and the relative coupling efficiency at this wavelength is 3.5 times that at λ = 526 nm.

    Profiles of ρ along the z axis after 4 ns of laser action for different wavelengths and adjusted intensity.

    Figure 2.Profiles of ρ along the z axis after 4 ns of laser action for different wavelengths and adjusted intensity.

    Relative coupling efficiency between laser and kinetic energy of implosion fluid for different wavelengths.

    Figure 3.Relative coupling efficiency between laser and kinetic energy of implosion fluid for different wavelengths.

    On the basis of the relative coupling efficiency, we investigate laser action on a target with initial perturbation amplitude η = η0 cos(kpx), where η0 = 0.1 μm and kp = 2π/12.8 μm−1, for different laser wavelengths and corresponding intensities. The ARTI perturbations near the ablation surface after 4 ns of laser action are shown in Fig. 4, and the corresponding average ARTI growth rates γ = ln(η/η0)/t within 4 ns are shown in Fig. 5. We find that the ARTI growth rates are larger for both longer and shorter wavelengths. However, as the wavelength moves from 150 nm toward the extreme ultraviolet band, the ARTI growth rate decreases rapidly, reaching a minimum value of γ = 0.31 ns−1 near λ = 65 nm. However, as the wavelength moves continuously toward the X-ray band, the ARTI growth rate increases rapidly.

    ARTI perturbations near the ablation surface after 4 ns of laser action for kp = 2π/12.8 μm−1 and different wavelengths λ: (a) 351 nm; (b) 150 nm; (c) 100 nm; (d) 65 nm; (e) 40 nm; (f) 30 nm.

    Figure 4.ARTI perturbations near the ablation surface after 4 ns of laser action for kp = 2π/12.8 μm−1 and different wavelengths λ: (a) 351 nm; (b) 150 nm; (c) 100 nm; (d) 65 nm; (e) 40 nm; (f) 30 nm.

    Average ARTI growth rates γ within 4 ns for kp = 2π/12.8 μm−1 and different laser wavelengths.

    Figure 5.Average ARTI growth rates γ within 4 ns for kp = 2π/12.8 μm−1 and different laser wavelengths.

    To explore the causes of the variation of the average ARTI growth rate with laser wavelength, we analyze the variations of the main hydrodynamic parameters (g, Lm, Va, conduction layer thickness Lc, and AT) with laser action time at five representative wavelengths (351, 100, 65, 40, and 30 nm), as shown in Figs. 6(a)6(e), and then we calculate the variation of the theoretical ARTI growth rate with laser action time according to the Lindl formula,18,19 as shown in Fig. 6(f). In Fig. 7, we compare the variations with laser action time of the simulated perturbation amplitude for different laser frequencies (symbols) and the perturbation amplitude calculated from the theoretical ARTI growth rate (dashed line). We find that the shorter the laser wavelength, the smaller is the theoretical ARTI growth rate, which is mainly due to the fact that Va increases as the laser wavelength becomes shorter, while the influence of other hydrodynamic parameters is small. For λ = 351 and 100 nm, the simulated perturbation amplitude is close to the theoretical results, but for λ = 65 nm, it is smaller than these, and in particular, for λ = 40 and 30 nm, the theoretical ARTI hardly grows, which represents a large deviation from the simulated perturbation amplitude.

    Variation of main hydrodynamic parameters with laser action time at five representative wavelengths (351, 100, 65, 40, and 30 nm): (a) g; (b) Lm; (c) Va; (d) conduction layer thickness Lc; (e) AT; (f) theoretical ARTI growth rate.

    Figure 6.Variation of main hydrodynamic parameters with laser action time at five representative wavelengths (351, 100, 65, 40, and 30 nm): (a) g; (b) Lm; (c) Va; (d) conduction layer thickness Lc; (e) AT; (f) theoretical ARTI growth rate.

    Variations with laser action time of simulated perturbation amplitude for different laser frequencies (symbols) and of perturbation amplitude calculated from the theoretical ARTI growth rate (dashed line).

    Figure 7.Variations with laser action time of simulated perturbation amplitude for different laser frequencies (symbols) and of perturbation amplitude calculated from the theoretical ARTI growth rate (dashed line).

    After further analysis of the fluid state, we deem there to be two main causes of the deviation. First, when λ < 50 nm, the maximum value of the plasma density is less than the critical density, and under this condition, the laser energy can reach the ablation surface in the form of electromagnetic waves instead of being absorbed by the plasma before the critical surface, as a consequence of which the electromagnetic wave is deflected in the perturbation region of plasma density, as shown in Fig. 8. Figure 8(a) shows how the direction of the laser changes in the perturbation region of plasma density, and Fig. 8(b) shows the distribution of laser energy deposition at λ = 30 nm. For comparison, Fig. 8(c) shows the distribution of laser energy deposition at λ = 65 nm. It can be seen that owing to the deflection of the laser, the energy flow distribution for ablation on the ablation surface becomes uneven in the transverse direction, and as the laser wavelength continues to decrease, the proportion of laser energy reaching the ablation surface in the form of electromagnetic waves increases, and so the energy flow distribution is more uneven, making the ARTI perturbation amplitude larger.

    (a) Plasma density contours and direction of laser (red arrow) in the perturbation region. (b) and (c) Distributions of laser energy deposition after 3 ns of laser action at λ = 30 and 65 nm, respectively.

    Figure 8.(a) Plasma density contours and direction of laser (red arrow) in the perturbation region. (b) and (c) Distributions of laser energy deposition after 3 ns of laser action at λ = 30 and 65 nm, respectively.

    The second cause of the deviation is that for longer wavelengths (>100 nm), the laser energy is absorbed before the critical surface, and the energy flow then has to pass through a thicker electron conduction layer to reach the ablation surface, as shown in Fig. 6(d). Because of the existence of ARTI, there is a perturbation of the temperature distribution in the electron conduction layer, as shown in Fig. 9, where A and B are two adjacent positions in the transverse direction, with A having a higher temperature than B. At this time, an energy flow is generated from A to B, with energy flow density q = −κ∂Te/∂x, where κTe5/2 is the coefficient of thermal conductivity. When λ = 65 nm, the critical density is close to the plasma density at the ablation front. At this time, the electron conduction layer is very thin, and so the transverse energy flow generated as the energy flow reaches the ablation surface is much smaller than that in the long-wavelength case. Therefore, for λ ≤ 65 nm, the transverse distribution of energy flow reaching the ablation surface effectively becomes uniform if the laser deflection is not considered. This effect is not taken into account in the ARTI growth rate formula, and so it is necessary to add a correction term related to Lc (or laser wavelength), which makes the formula more applicable to short-wavelength situations.

    Temperature contours of plasma after 4 ns of laser action at λ = 351 nm.

    Figure 9.Temperature contours of plasma after 4 ns of laser action at λ = 351 nm.

    IV. EFFECT OF INTENSITY BENCHMARK

    To study the effects of different values of the intensity benchmark Ib, we consider intensities of 2 × 1014 and 4 × 1014 W/cm2 with a wavelength of 351 nm and simulate laser-driven ablation of a one-dimensional target. We then extend the steady-state flow field into two dimensions and introduce perturbations again. The size of the simulation domain is now set to 25.6 × 600 μm2, because of the increased implosion velocity. We find that the relative coupling efficiency between lasers with different wavelengths and the kinetic energy of the implosion fluid remains the same for different intensity benchmarks. The ARTI average growth rate γ is shown in Fig. 10 for an initial single-mode perturbation with kp = 2π/12.8 μm−1. The growth rate reaches a minimum near λ = 55 nm for Ib = 2 × 1014 W/cm2 and near λ = 40 nm for Ib = 4 × 1014 W/cm2, which are shifts of about 10 nm and about 25 nm, respectively, toward shorter wavelengths. This is because the ablation velocity and pressure increase with increasing intensity benchmark, which results in an increase of plasma density at the ablation surface, reducing the laser wavelength needed for a noticeable deflection effect. On the other hand, the increase in ablation velocity enhances the stability of ablation, resulting in a minimum growth rate lower than that at Ib = 1 × 1014 W/cm2.

    Average ARTI growth rate γ within 4 ns at different laser wavelengths for Ib = 2 × 1014 W/cm2 and 4 × 1014 W/cm2.

    Figure 10.Average ARTI growth rate γ within 4 ns at different laser wavelengths for Ib = 2 × 1014 W/cm2 and 4 × 1014 W/cm2.

    V. EFFECT OF PERTURBATION WAVENUMBER

    In ICF, many perturbation modes grow simultaneously. During the linear growth phase, the Lindl formula can predict the growth of perturbations well under ablation conditions, and initial perturbations with mode numbers l = 140–760 eventually enter the nonlinear growth phase,45 where l is the Legendre polynomial mode of the surface perturbation amplitude. For a typical ICF spherical target with a radius of 1 mm, the corresponding perturbation wavelength range is 8–45 μm. Therefore, we initially select a wavenumber kp = 2π/12.8 μm−1, and we study cases in which the single-mode perturbation wavenumber kp = 2π/25.6 μm−1 or 2π/51.2 μm−1 for Ib = 1 × 1014 W/cm2. The results are shown in Fig. 11. The trend of variation of the ARTI growth rate with laser wavelength is basically consistent with that at kp = 2π/12.8 μm−1, but the growth rate reaches its minimum near λ = 75 nm for kp = 2π/25.6 μm−1, which is a shift of about 10 nm toward longer wavelength. For kp = 2π/51.2 μm−1, the ARTI growth rate reaches its minimum near λ = 95 nm. This is because the smaller the perturbation wavenumber, the longer is the perturbation region of plasma density, as shown in Fig. 12. This causes a laser with a given wavelength to start deflecting at a position farther away from the ablation surface as the perturbation wavenumber decreases, and so the laser wavelength needed for a noticeable deflection effect becomes longer.

    Variation of average ARTI growth rate γ within 4 ns with laser wavelength for kp = 2π/25.6 and 2π/51.2 μm−1.

    Figure 11.Variation of average ARTI growth rate γ within 4 ns with laser wavelength for kp = 2π/25.6 and 2π/51.2 μm−1.

    Density contours and perturbation region length h of plasma after 4 ns of laser action at λ = 351 nm for (a) kp = 2π/25.6 μm−1 and (b) kp = 2π/51.2 μm−1.

    Figure 12.Density contours and perturbation region length h of plasma after 4 ns of laser action at λ = 351 nm for (a) kp = 2π/25.6 μm−1 and (b) kp = 2π/51.2 μm−1.

    VI. CONCLUSION

    The effect of laser wavelength on the growth of ARTI has been studied through numerical simulations. In the plasma acceleration phase, shorter wavelengths lead to more efficient coupling between the laser and the kinetic energy of the implosion fluid. Under the condition that the laser energy coupled to the implosion fluid is constant, the ARTI growth rate decreases as the wavelength moves toward the extreme ultraviolet band, reaching a minimum value of γ = 0.31 ns−1 near λ = 65 nm. As the wavelength continuously moves toward the X-ray band, the ARTI growth rate increases rapidly. This is due to three causes. First, the ablation velocity Va becomes larger as the laser wavelength decreases, which enhances the stability of ablation. Second, when the maximum value of the plasma density is less than the critical density, laser energy can reach the ablation surface in the form of electromagnetic waves, which are deflected in the perturbation region of plasma density, resulting in an energy flow distribution for ablation on the ablation surface that is uneven in the transverse direction. Under this condition, the theoretical ARTI growth rate deviates significantly from the simulated results. Third, for longer wavelengths, the energy flow has to pass through a thicker electron conduction layer to reach the ablation surface, and the temperature distribution in the electron conduction layer is perturbed by the presence of ARTI, which leads to the generation of a transverse energy flow and thus to an uneven transverse distribution of energy flow reaching the ablation surface. It is therefore necessary to modify the ARTI growth formula.

    As the laser intensity benchmark increases, the position of the minimum ARTI growth rate shifts toward shorter wavelengths. This is because the ablation velocity and pressure increase with increasing intensity benchmark, which results in an increase in plasma density at the ablation surface, as a consequence of which the laser wavelength needed for a noticeable deflection effect becomes shorter.

    As the initial perturbation wavenumber decreases, the position of the minimum ARTI growth rate shifts toward longer wavelengths. This is because the smaller the perturbation wavenumber, the longer is the perturbation region of plasma density, and a laser of given wavelength starts to be deflected at a position farther away from the ablation surface.

    Thus, numerical simulation and theoretical analysis have indicated that extreme ultraviolet lasers of specific wavelengths have higher coupling efficiency and can better control the growth of instability, thereby significantly reducing the laser energy required for fusion ignition and achieving higher fusion energy gain. At present, the main methods to generate high-power extreme ultraviolet light sources are by multiplying the frequency of high-power lasers46 or by generating high-order harmonics through laser interaction with solid materials.47,48 Therefore, the use of extreme ultraviolet lasers to achieve ICF ignition is an important topic worthy of further investigation in the future.

    ACKNOWLEDGMENTS

    Acknowledgment. This work is supported by the National Natural Science Foundation of China (NSFC) (Grant Nos. 12074399, 12204500, and 12004403), the Key Projects of the Intergovernmental International Scientific and Technological Innovation Cooperation (Grant No. 2021YFE0116700), the Shanghai Natural Science Foundation (Grant No. 20ZR1464400), and the Shanghai Sailing Program (Grant No. 22YF1455300).

    References

    [1] J.Nuckolls, A.Thiessen, L.Wood, G.Zimmerman. Laser compression of matter to super-high densities: Thermonuclear (CTR) applications. Nature, 239, 139-142(1972).

    [2] K. S.Anderson, T. R.Boehly, R. S.Craxton, V. N.Goncharov, D. R.Harding, J. P.Knauer, R. L.McCrory, P. W.McKenty, D. D.Meyerhofer, J. F.Myattet?al.. Direct-drive inertial confinement fusion: A review. Phys. Plasmas, 22, 110501(2015).

    [3] R.Betti, O. A.Hurricane. Inertial-confinement fusion with lasers. Nat. Phys., 12, 435-448(2016).

    [4] E. E.Meshkov. Instability of the interface of two gases accelerated by a shock wave. Fluid Dyn., 4, 101-104(1969).

    [5] Helmholtz. XLIII. On discontinuous movements of fluids. London, Edinburgh Dublin Philos. Mag. J. Sci., 36, 337-346(1868).

    [6] M. S.Plesset. On the stability of fluid flows with spherical symmetry. J. Appl. Phys., 25, 96-98(1954).

    [7] L.Rayleigh. Scientific Papers II, 200(1900).

    [8] G. I.Taylor. The instability of liquid surfaces when accelerated in a direction perpendicular to their planes. I. Proc. R. Soc. London, Ser. A, 201, 192-196(1950).

    [9] K. S.Anderson, R.Betti, P. Y.Chang, J.Edwards, M.Fatenejad, J. D.Lindl, R. L.McCrory, R.Nora, D.Shvarts, B. K.Spears. Thermonuclear ignition in inertial confinement fusion and comparison with magnetic confinement. Phys. Plasmas, 17, 058102(2010).

    [10] D.Layzer. On the instability of superposed fluids in a gravitational field. Astrophys. J., 122, 1(1955).

    [11] V. N.Goncharov. Analytical model of nonlinear, single-mode, classical Rayleigh–Taylor instability at arbitrary Atwood numbers. Phys. Rev. Lett., 88, 134502(2002).

    [12] T.Ikegawa, K.Nishihara. Ablation effects on weakly nonlinear Rayleigh–Taylor instability with a finite bandwidth. Phys. Rev. Lett., 89, 115001(2002).

    [13] R.Betti, J.Ramírez, R.Ramis, J.Sanz, R. P. J.Town. Nonlinear theory of the ablative Rayleigh–Taylor instability. Phys. Rev. Lett., 89, 195002(2002).

    [14] K.Mima, L.Montierth, R. L.Morse, H.Takabe. Self-consistent growth rate of the Rayleigh–Taylor instability in an ablatively accelerating plasma. Phys. Fluids, 28, 3676(1985).

    [15] S. E.Bodner. Rayleigh–Taylor instability and laser-pellet fusion. Phys. Rev. Lett., 33, 761(1974).

    [16] J. D.Lindl, D. H.Munro, M.Tabak. Hydrodynamic stability and the direct drive approach to laser fusion. Phys. Fluid. Plasma Phys., 2, 1007-1014(1990).

    [17] R.Betti, V. N.Goncharov, R. L.McCrory, C. P.Verdon. Growth rates of the ablative Rayleigh–Taylor instability in inertial confinement fusion. Phys. Plasmas, 5, 1446(1998).

    [18] S. G.Glendinning, S. W.Haan, B. A.Hammel, J. D.Kilkenny, J. P.Knauer, J. D.Lindl, D.Munro, B. A.Remington, C. P.Verdon, S. V.Weber. A review of the ablative stabilization of the Rayleigh–Taylor instability in regimes relevant to inertial confinement fusion. Phys. Plasmas, 1, 1379-1389(1994).

    [19] J.Lindl. Development of the indirect-drive approach to inertial confinement fusion and the target physics basis for ignition and gain. Phys. Plasmas, 2, 3933-4024(1995).

    [20] J.Sanz. Self-consistent analytical model of the Rayleigh–Taylor instability in inertial confinement fusion. Phys. Rev. Lett., 73, 2700(1994).

    [21] R.Betti, V. N.Goncharov, R. L.McCrory, C. P.Verdon. Self-consistent cutoff wave number of the ablative Rayleigh–Taylor instability. Phys. Plasmas, 2, 3844-3851(1995).

    [22] R.Betti, V. N.Goncharov, R. L.McCrory, P.Sorotokin, C. P.Verdon. Self-consistent stability analysis of ablation fronts with large Froude numbers. Phys. Plasmas, 3, 1402-1414(1996).

    [23] R.Betti, V. N.Goncharov, R. L.McCrory, C. P.Verdon. Self-consistent stability analysis of ablation fronts with small Froude numbers. Phys. Plasmas, 3, 4665-4676(1996).

    [24] R.Betti, V. N.Goncharov, R. L.McCrory, P.Sorotokin, C. P.Verdon. Self-consistent stability analysis of ablation fronts in inertial confinement fusion. Phys. Plasmas, 3, 2122-2128(1996).

    [25] L. F.Ibanez, A. R.Piriz, J.Sanz. Rayleigh–Taylor instability of steady ablation fronts: The discontinuity model revisited. Phys. Plasmas, 4, 1117-1126(1997).

    [26] F. F.Chen. Introduction to Plasma Physics(1974).

    [27] E. G.Harris, J. F.Seely. Heating of a plasma by multiphoton inverse bremsstrahlung. Phys. Rev. A, 7, 1064(1973).

    [28] R.Boni, R. S.Craxton, J.Delettrez, L.Goldman, R.Keck, R. L.McCrory, W.Seka, D.Shvarts, J. M.Soures. Measurements and interpretation of the absorption of 0.35 μm laser radiation on planar targets. Opt. Commun., 40, 437-440(1982).

    [29] F.Amiranoff, R.Fabbro, E.Fabre, C.Garban-Labaune, C. E.Max, A.Michard, J.Virmont, M.Weinfeld. Effect of laser wavelength and pulse duration on laser-light absorption and back reflection. Phys. Rev. Lett., 48, 1018(1982).

    [30] R. S.Craxton, J.Delettrez, R. L.Keck, R. L.McCrory, M. C.Richardson, W.Seka, J. M.Soures. Absorption physics at 351 nm in spherical geometry. Phys. Rev. Lett., 54, 1656(1985).

    [31] P.Amendt, R. L.Berger, S. G.Glendinning, S. H.Glenzer, S. W.Haan, R. L.Kauffman, O. L.Landen, J. D.Lindl, L. J.Suter. The physics basis for ignition using indirect-drive targets on the national ignition facility. Phys. Plasmas, 11, 339-491(2004).

    [32] B.Fryxell, D. Q.Lamb, P.MacNeice, K.Olson, P.Ricker, R.Rosner, F. X.Timmes, J. W.Truran, H.Tufo, M.Zingale. FLASH: An adaptive mesh hydrodynamics code for modeling astrophysical thermonuclear flashes. Astrophys. J. Suppl. Ser., 131, 273(2000).

    [33] M.Fatenejad, N.Flocke, C.Graziani, G.Gregori, D. Q.Lamb, D.Lee, J.Meinecke, A.Scopatz, P.Tzeferacos, K.Weide. FLASH MHD simulations of experiments that study shock-generated magnetic fields. High Energy Density Phys., 17, 24-31(2015).

    [34] C. W.Shu. Essentially Non-Oscillatory and Weighted Essentially Non-Oscillatory Schemes for Hyperbolic Conservation Laws, 325-432(1998).

    [35] A. E.Deane, C.Federrath, D.Lee. A new multidimensional unsplit MHD solver in FLASH3. Numerical Numbering of Space Plasma Flows: ASTRONUM-2008, 406, 243(2009).

    [36] B.Van Leer. Towards the ultimate conservative difference scheme. J. Comput. Phys., 135, 229-248(1997).

    [37] W.Speares, M.Spruce, E. F.Toro. Restoration of the contact surface in the HLL–Riemann solver. Shock Waves, 4, 25-34(1994).

    [38] S. K.Doss, R. J.Gelinas, K.Miller. The moving finite element method: Applications to general partial differential equations with multiple large gradients. J. Comput. Phys., 40, 202-249(1981).

    [39] M.Desselberger, M. J.Lamb, M.Savage, O.Willi. Measurement of the Rayleigh–Taylor instability in targets driven by optically smoothed laser beams. Phys. Rev. Lett., 65, 2997(1990).

    [40] R.Betti, T. R.Boehly, D. K.Bradley, T. J. B.Collins, V. N.Goncharov, J. P.Knauer, P. W.Mckenty, D. D.Meyerhofer, V. A.Smalyuk, C. P.Verdonet?al.. Single-mode, Rayleigh–Taylor growth-rate measurements on the OMEGA laser system. Phys. Plasmas, 7, 338-345(2000).

    [41] H. J.Kull. Incompressible description of Rayleigh–Taylor instabilities in laser-ablated plasmas. Phys. Fluid. Plasma Phys., 1, 170-182(1989).

    [42] B. J.Albright, E.Vold, L.Yin. Plasma transport simulations of Rayleigh–Taylor instability in near-ICF deceleration regimes. Phys. Plasmas, 28, 092709(2021).

    [43] Y. J.Li, Z. Y.Li, Y. X.Liu, L. F.Wang, J. F.Wu, W. H.Ye, K. G.Zhao. Thin-shell effects on nonlinear bubble evolution in the ablative Rayleigh–Taylor instability. Phys. Plasmas, 29, 082102(2022).

    [44] S. P.Obenschain, A. J.Schmitt. The importance of laser wavelength for driving inertial confinement fusion targets. I. Basic physics. Phys. Plasmas, 30, 012701(2023).

    [45] K.Fang, Y.-T.Li, J.Zhang, Z.Zhang. Analytical studies of Rayleigh–Taylor instability growth of double-cone ignition scheme in 2020 winter experimental campaign. Acta Phys. Sin., 71, 035204(2022).

    [46] M.Chen, S. B.Dai, Z. M.Wang, F.-F.Zhang, S. J.Zhanget?al.. 2.14 mW deep-ultraviolet laser at 165 nm by eighth-harmonic generation of a 1319 nm Nd:YAG laser in KBBF. Laser Phys. Lett., 13, 035401(2016).

    [47] D. A.Browne, M. B.Gaarde, S.Ghimire, G.Ndabashimiye, D. A.Reis, K. J.Schafer, M.Wu. Solid-state harmonics beyond the atomic limit. Nature, 534, 520(2016).

    [48] S.Han, H.Kim, S.Kim, S. W.Kim, Y. J.Kim, Y. W.Kim, I. Y.Park. High-harmonic generation by field enhanced femtosecond pulses in metal-sapphire nanostructure. Nat. Commun., 7, 13105(2016).

    Zhantao Lu, Xinglong Xie, Xiao Liang, Meizhi Sun, Ping Zhu, Xuejie Zhang, Linjun Li, Hao Xue, Guoli Zhang, Rashid Ul Haq, Dongjun Zhang, Jianqiang Zhu. Effect of laser wavelength on growth of ablative Rayleigh–Taylor instability in inertial confinement fusion[J]. Matter and Radiation at Extremes, 2025, 10(2): 027403
    Download Citation