This site uses cookies. By continuing to use this site you agree to our use of cookies. To find out more, see our Privacy and Cookies policy.
Paper The following article is Open access

Investigation of the Langdon effect on the nonlinear evolution of SRS from the early-stage inflation to the late-stage development of secondary instabilities

, , and

Published 15 November 2022 © 2022 The Author(s). Published by IOP Publishing Ltd
, , Citation Jie Qiu et al 2022 Nucl. Fusion 62 126072 DOI 10.1088/1741-4326/ac9b75

0029-5515/62/12/126072

Abstract

In a laser-irradiated plasma, the Langdon effect can result in a super-Gaussian electron energy distribution function (EEDF), imposing significant influences on stimulated backward Raman scattering (SRS). In this work, the influence of a super-Gaussian EEDF on the nonlinear evolution of SRS is investigated by the three wave coupling model simulation and Vlasov–Maxwell simulation for plasma parameters covering a wide range of De from 0.19 to 0.48 at both high and low intensity laser drives. In the early stage of SRS evolution, it is found that besides the kinetic effects due to electron trapping (2018 Phys. Plasmas 25 100702), the Langdon effect can also significantly widen the parameter range for the absolute growth of SRS, and the time for the absolute SRS to reach saturation is greatly shortened by the Langdon effect within certain parameter regions. In the late stage of SRS, when secondary instabilities such as decay of the electron plasma wave to beam acoustic modes, rescattering, and Langmuir decay instability become important, the Langdon effect can influence the reflectivity of SRS by affecting secondary instabilities. The comprehension of the Langdon effect on nonlinear evolution and saturation of SRS would contribute to a better understanding and prediction of SRS in inertial confinement fusion.

Export citation and abstract BibTeX RIS

Original content from this work may be used under the terms of the Creative Commons Attribution 4.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

In laser-driven inertial confinement fusion (ICF), stimulated backward Raman scattering (SRS) is one important laser plasma instability (LPI), where an incident electromagnetic wave (EMW) resonantly decays into a backscattered EMW and a forward propagating electron plasma wave (EPW) [1]. SRS needs to be suppressed in ICF, since the backscattered light can take energy away from the incident laser, and the hot electrons generated by EPW of SRS can preheat the capsule [2].

Currently, the study of SRS usually assumes a Maxwellian electron energy distribution (EEDF) [3, 4]. However, for the laser-irradiated plasma in ICF, the dominant heating mechanism is inverse bremsstrahlung (IB) heating. The Langdon effect [5, 6], i.e., the EEDF tends toward a super-Gaussian form when the IB heating rate exceeds the electron thermalization rate, can become quite important under typical hohlraum conditions. It is found that in the linear convective regime, the Langdon effect can significantly enhance the gain of SRS [7], thus providing a possible explanation for the much weaker reflectivity predicted by the ray-tracing code based on the Maxwellian EEDF gain model as compared to the experimentally measured one [8, 9]. Nevertheless, for the widely spreading range of plasma parameters and laser intensity for typical ICF, the assumption of convective SRS is invalid in some zones with the high local laser intensity which favors the Langdon effect and the reduction of Landau damping caused by the Langdon effect. The absolute growth of SRS can occur instead, wherein the amplitudes of daughter waves of SRS would grow quickly over a short distance until saturated by nonlinear saturation processes [912]. Currently, the investigation of the Langdon effect on the absolute SRS and its nonlinear behavior is still lacking. Therefore, it is necessary to examine the evolution of SRS for a wide parameter range after taking into account the Langdon effect in a self-consistent way, which is quite important for the global understanding and proper modeling of SRS in ICF.

In this work, the influence of the Langdon effect on the nonlinear evolution of SRS is investigated for a wide range of plasma parameters at both high and low-intensity laser drives. In the early growth stage of SRS, it is found that apart from the kinetic inflation due to trapped electrons [1315], the Langdon effect can broaden the parameter range for absolute SRS modes, and significantly shorten the saturation time for the absolute SRS in certain parameter regions. Over a long timescale evolution with development of the secondary instabilities such as the decay of the electron plasma wave to beam acoustic modes (BAM) [1618], rescattering of the primary scattered wave [19, 20] and Langmuir decay instability (LDI) [2123], the SRS typically demonstrates a change in dominant saturation mechanism accompanied by a change (usually drop) in the reflectivity. Under many circumstances, the Langdon effect is found to be important to these secondary instabilities and thus to the nonlinear saturation of SRS in the late evolution stage.

This paper is organized as follows: in section 2, a three wave coupling (TWC) model with consideration of the Langdon effect is given, as well as some theoretical analysis of SRS in the presence of the Langdon effect. Additionally, the physics model of a Vlasov–Maxwell code (VlaMaxW) is presented to account for both the Langdon effect and nonlinear kinetic effects. In section 3, the influence of super-Gaussian EEDFs on the nonlinear evolution of SRS is investigated for different stages. In section 4, the conclusions and some discussions are given.

2. Models with the Langdon effect considered

In SRS, the matching condition requires [1]

Equation (1)

where ωi and ki are the frequencies and wavenumbers with the subscripts i = 0, s, l for the pump wave, the scattered wave and the EPW, respectively. When the three waves of SRS are dominated by coherent single modes, the temporal-spatial evolution of SRS can be described by the coupling equations for their slowly varying envelopes. Writing the transverse electric fields of the pump wave and the scattered wave as $\mathfrak{R}[{E}_{0}{e}^{-j{\omega }_{0c}t+j{k}_{0c}x}]$, $\mathfrak{R}[{E}_{s}{e}^{-j{\omega }_{sc}t-j{k}_{sc}x}]$, and the density perturbation of the EPW as $\mathfrak{R}[\delta {n}_{l}{e}^{-j{\omega }_{lc}t+j{k}_{lc}x}]$, the envelope equations for the pump wave and the scattered wave can be written as [4, 24, 25]

Equation (2)

Equation (3)

where ωpe is the plasma frequency, ne0 is the background electron density, ωic and kic are the fundamental frequency and wavenumber of the pump wave  (i = 0) or the scattered wave  (i = s) respectively, and vi = c2 kic/ωic is the group velocity. Here, c is the light speed. The fundamental modes satisfy the dispersion relation of EMW

Equation (4)

To account for a non-Maxwellian EEDF, a kinetic description of ponderomotively driven EPW can be derived as [26],

Equation (5)

which can be rewritten as

Equation (6)

where e is the electron charge, me is the electron mass, and $\mathcal{D}$ is a function of ωl and kl

Equation (7)

describing the ponderomotive response. The electron susceptibility χe is given by

Equation (8)

where fe0 is the background EEDF. When the Langdon effect is significant, the EEDF has a super-Gaussian form [6, 7],

Equation (9)

where m is the super-Gaussian exponent, Γ is the gamma function, ${\beta }_{m}=\sqrt{3{\Gamma}(3/m)/{\Gamma}(5/m)}$, and ${v}_{\text{the}}=\sqrt{{T}_{e}/{m}_{e}}$ is the electron thermal velocity. Replacing $\mathcal{D}({\omega }_{l},{k}_{l})$ by the operator $\mathcal{D}({\omega }_{lc}+j{\partial }_{t},{k}_{lc}-j{\partial }_{x})$, and taking the slow envelope variation approximation [4]

Equation (10)

the envelope equation of EPW can be derived as

Equation (11)

Here only ${\mathcal{D}}_{r}\equiv \mathrm{R}\mathrm{e}[\mathcal{D}]$ is retained in the derivatives for $\vert \mathfrak{I}[\partial \mathcal{D}/\partial {\omega }_{l}]\vert \ll \vert \partial {\mathcal{D}}_{r}/\partial {\omega }_{l}\vert $. The group velocity of the EPW

Equation (12)

the Landau damping of EPW

Equation (13)

and the frequency mismatch

Equation (14)

Notice that in this approach, the Langdon effect can be easily incorporated just by using equation (9) when calculating χe . Furthermore, for relatively small kl λDe < 0.3 with λDe = vthe/ωpe being the Debye length, the EPW modes stimulated by SRS are close to the natural modes [26], then χe ≈ −1 and $\vert \mathfrak{I}[\partial {\epsilon}/\partial {\omega }_{l}]\vert \ll \vert \partial {{\epsilon}}_{r}/\partial {\omega }_{l}\vert $. In such cases, the dielectric operator epsilon can be adopted to describe the response of the EPW field to the ponderomotive drive from the beating of the pump wave and the scattered wave [4]. However, for large kl λDe, electron modes are strongly damped and the natural plasma modes are strongly modified [26], thus χe ≈ −1 is no longer a very good assumption for electron modes and the assumption $\vert \mathfrak{I}[\partial {\epsilon}/\partial {\omega }_{l}]\vert \ll \vert \partial {{\epsilon}}_{r}/\partial {\omega }_{l}\vert $ is broken. Then, the ponderomotive operator $\mathcal{D}=-(1+{\chi }_{e})/{\chi }_{e}$ defined here is required to describe the response of EPW field instead of epsilon. Because $\mathcal{D}$ includes both the influence of χe on the dielectric response epsilon = 1 + χe and the ponderomotive drive, which altogether determine the response of SRS [7], and the assumption $\vert \mathfrak{I}[\partial \mathcal{D}/\partial {\omega }_{l}]\vert \ll \vert \partial {\mathcal{D}}_{r}/\partial {\omega }_{l}\vert $ can be reasonably satisfied for kl λDe ranging from 0 to 1, this approach is not only valid for small kl λDe where $\mathcal{D}\approx {\epsilon}$, but also applicable to  0.3 ⩽ kl λDe ⩽ 1.

It is convenient to renormalize the wave amplitudes as

Equation (15)

where El = jeδnl /epsilon0 kl is the electrostatic field of EPW, and E0L is the pump field amplitude incident at the left boundary. Then, the TWC equations of SRS can be recast into the following simplified form

Equation (16)

Equation (17)

Equation (18)

where γ0 is the homogeneous growth rate of SRS when the Landau damping and frequency mismatch are ignored. By substituting the assumed solution forms ${E}_{s},\delta {n}_{l}/{n}_{0}\propto {e}^{{\gamma }_{0}t}$ into equations (3) and (11)

Equation (19)

can be obtained, where vos = e|E0|/me ω0c is the electron quiver velocity. Then, the instantaneous reflectivity R can be determined from the scattered wave amplitude emergent from the left boundary

Equation (20)

In this definition, R < 1 limited by pump depletion can be guaranteed if R eventually reaches a constant value [10].

SRS can be in the convective instability regime or in the absolute instability regime [10, 27, 28]. The convective instability typically occurs under strong damping condition. The convective gain coefficient κR can be derived by assuming the solution form ${a}_{s},{a}_{l}\propto {e}^{-{\kappa }_{\text{R}}x}$, where the temporal and spatial derivatives in equation (18) can be ignored compared to the Landau damping, yielding

Equation (21)

which is just the classical formula for the kinetic convective gain coefficient of SRS [7, 26]. The strong damping condition νl κR vl thus implies ${\nu }_{l}\gg {\gamma }_{0}\sqrt{{v}_{l}/{v}_{s}}$, under which the saturated reflectivity with pump depletion considered can be analytically determined from the TWC model, given by the Tang's formula [10, 29],

Equation (22)

where ɛ = |as,Right|2 vs /v0 is determined by the seed light intensity at the right boundary, and GR = 2κR L is the energy gain of SRS for an amplification length L.

The absolute instability usually occurs under strong laser drive or weak Landau damping. It is of important concern for the SRS control in ICFs [11], since SRS keeps growing until saturated by nonlinear effects, generally leading to a large reflectivity. When nonlinear effects can be ignored, by analyzing equations (17) and (18) with Laplace transform, the absolute instability condition can be derived as [10, 30]

Equation (23)

and the absolute growth rate is

Equation (24)

This TWC model can describe the evolution of SRS both in the convective and absolute regimes with the consideration of the Langdon effect and the pump depletion, but excludes all the nonlinear kinetic effects such as reduction of Landau damping and frequency shift due to trapped electrons. As simple as it is, the TWC model is helpful for understanding the initial growth of SRS and identifying the onset of nonlinear kinetic effects.

Since the super-Gaussian EEDF can significantly reduce νl [7], it is expected that the absolute instability condition is more easily met when the Langdon effect is considered. Nevertheless, in most cases the linear condition fails to be a good criterion to judge whether absolute growth can occur, since the kinetic inflation [13, 15, 31], wherein the convective SRS is transformed into absolute SRS due to electron trapping effects, is found to be quite important for SRS. To take into account the nonlinear kinetic effects, an one spatial dimensional and one velocity dimensional (1D1V) VlaMaxW has been developed, which solves the Vlasov equations

Equation (25)

Equation (26)

together with Maxwell equations

Equation (27)

Equation (28)

Equation (29)

Equation (30)

Equation (31)

Equation (32)

by numerical methods similar to reference [9]. Here the cold plasma approximation for the transverse motion p = −qα Ay is assumed, and subscript α = e, i for electron and ion, respectively. mα is the mass, qα is the charge, epsilon0 is the vacuum permittivity, p is the momentum along x direction, Ay is the vector potential, Ex is the electrostatic field, Ey and Bz are the transverse electric and magnetic fields, respectively. In the Vlasov–Maxwell simulation, the influence of super-Gaussian EEDF can be taken into account by specifying the initial one-dimensional electron distribution function ${f}_{e}^{x}$ as

Equation (33)

where fe0(v) is the isotropic super-Gaussian electron distribution given by equation (9), and the regularized incomplete gamma function $Q(s,z)\equiv {\int }_{z}^{\infty }{t}^{s-1}{e}^{-t}\mathrm{d}t/{\Gamma}(s)$. The shape of ${f}_{e}^{x}$ for different m is compared in figure 1. As seen, with increasing m, there are less electrons in both low energy range and high energy range, yet more electrons in the middle energy range of vx /vthe ∼ 1–2. For ion species, since direct laser heating is negligible, an initial one-dimensional Maxwellian distribution with ion temperature Ti is assumed.

Figure 1.

Figure 1. The one-dimensional super-Gaussian EEDF ${f}_{e}^{x}({v}_{x})$ at different m.

Standard image High-resolution image

In the following, the Vlasov–Maxwell simulation is primarily used to investigate the influence of super-Gaussian EEDF on the nonlinear evolution (e.g., inflation and saturation) of SRS, while the TWC simulation is conducted to provide a valuable reference to help comprehend the simulation results, where the fundamental mode of the EPW is chosen at the peak value of κR. For small kl λDe, it is quite close to the natural mode of EPW which satisfies  1 + χe (ωl , kl ) = 0, and also δωl ≈ 0. While for kl λDe > 0.3, this can deviate significantly from the natural mode, and also δωl can be nonzero.

3. The Langdon effect on the nonlinear evolution of SRS

In the Vlasov–Maxwell and TWC model simulations, a homogeneous He plasma with length L0 = 200λ0 is assumed, where the laser vacuum length λ0 = 0.351 μm, L0 is on the speckle length scale $\sim 8{f}^{2}{\lambda }_{0}$ of a realistic laser beam with f being the f-number [2], and the low-Z He plasma is the typical environment for significant SRS generation in ICFs [2, 32]. In the Vlasov–Maxwell simulation, also two additional collision layers with lengths  2 × 20λ0 and ramp electron density profiles are appended to both sides of the plasma to eliminate effects of the sheath field. One linearly polarized laser beam with intensity I0 is incident from the left boundary at t = 0 with a rise time  20T0, while the seed light with frequency ωsc and intensity Is = 10−6 I0 is incident from the right boundary at t = 100T0 with a rise time  20T0, where T0 = 2π/ω0 is the laser period. Since the typical range of super-Gaussian exponent m for low-Z He plasma is between 2 and 3 as analyzed in reference [7], simulation cases with m = 2 and m = 2.9 are compared to demonstrate impacts of the Langdon effect in this work. As known, ${k}_{l}{\lambda }_{\text{De}}\propto \sqrt{{T}_{e}/{n}_{e}}$, is a key parameter to determine the saturation mechanism of SRS [1, 33], so we widely scan kl λDe from 0.48 to 0.19 in simulations, by changing Te from 5 keV to 0.8 keV with ne = 0.1nc (nc is the critical density). Laser intensity is typically chosen as I15 = 2.82  (I15 = I0 [W cm−2]/1015), which is achievable in small laser speckles in ICF. Besides, a low laser intensity I15 = 0.11 is also used for the low kl λDe cases.

In the Vlasov–Maxwell simulations, the temporal evolution of the electrostatic field in the kl -space together with the instantaneous reflectivity are shown with different Te and m in figures 2 and 3 for I15 = 2.82 and I15 = 0.11, respectively. Except the cases shown in figures 2(a) and 3(a), where the SRS reflectivity is saturated at a low level due to convective saturation [10, 30]; the nonlinear effects are obvious in other cases, where the onset of early-stage saturation of the reflectivity generally occurs before significant broadening of the kl -spectrum induced by the secondary instabilities and even the cascaded instabilities. Therefore, the SRS evolution can be approximately divided into the early stage where the secondary instabilities play negligible roles, and the late stage where the secondary instabilities dominate and result in non-stationary variation of the reflectivity. In the following subsection 3.1, the influences of the Langdon effect on the early growth and saturation of SRS are mainly studied. Then, the differences in the dominant saturation mechanism and reflectivity of SRS for different m in the late stage are discussed in subsection 3.2.

Figure 2.

Figure 2. The evolution of kl -spectrum of the EPW field  [log10|El (kl )|2] for several cases with Te = 4.5, 3.2, 2.8 keV and m = 2, 2.9 when I15 = 2.82. The corresponding reflectivity versus time is displayed in the bottom panels, where dividing-line between the early stage and the later stage is marked by the red dotted vertical lines. The condition ne = 0.1 nc , λ0 = 351 nm for a homogeneous He plasma with Ti = Te /5 and length of 200λ0 is taken.

Standard image High-resolution image
Figure 3.

Figure 3. The evolution of kl -spectrum of the electrostatic field  [log10|Ez (kl )|2] for several cases with Te = 2, 1.5, 1.2 keV and m = 2, 2.9 when I15 = 0.11. The corresponding reflectivity versus time is displayed in the bottom panels, where dividing-line between the early stage and the later stage is marked by the red dotted vertical lines. The condition ne = 0.1 nc , λ0 = 351 nm for a homogeneous He plasma with Ti = Te /5 and length of 200λ0 is taken.

Standard image High-resolution image

3.1. The early-stage growth and saturation of SRS

As shown in figures 2 and 3, SRS grows convectively with a low saturation level at the larger kl λDe, but grows quickly to a high early-stage saturation level at the smaller kl λDe. This issue can be predicted by the linear criterion ${\nu }_{l0}\leqslant 2{\gamma }_{0}\sqrt{{v}_{l}/{v}_{s}}$ for absolute SRS, where νl0 is the initial Landau damping. Since νl0 decreases rapidly with decreasing kl λDe, there exists one critical ${[{k}_{l}{\lambda }_{\text{De}}]}_{c}$ at which the equality of the criterion can be satisfied. For ${k}_{l}{\lambda }_{\text{De}} > {[{k}_{l}{\lambda }_{\text{De}}]}_{\text{c}}$, the SRS growth is convective. The reflectivity can saturate at a lower level and nonlinear effects are quite weak, as presented in figures 4(c) and (d). In such cases, the reflectivity calculated by the Tang's formula (22) agrees well with the simulation results of the TWC model, as shown in figures 4(a) and (b). For ${k}_{l}{\lambda }_{\text{De}}< {[{k}_{l}{\lambda }_{\text{De}}]}_{\text{c}}$, the SRS growth is absolute, implying that SRS can keep growing until saturated by nonlinear effects, as shown by the time evolution of reflectivity in figures 4(c) and (d), which is calculated by the TWC model and hence the only nonlinear effect is the pump depletion. Since nonlinear effects only become important at large reflectivity, the saturated reflectivity for absolute SRS growth can not be too small. Consequently, when ${[{G}_{\text{R}}]}_{\text{c}}$ at ${[{k}_{l}{\lambda }_{\text{De}}]}_{\text{c}}$ is small and hence the convectively saturated level is low, there would be a sharp increase of reflectivity (especially, much sharper than the prediction by the Tang's model) at the transition from convective to absolute SRS, as exampled in figure 4(b). In comparison, the change of R near ${[{k}_{l}{\lambda }_{\text{De}}]}_{\text{c}}$ is much more gradual in figure 4(a), since for ${[{G}_{\text{R}}]}_{\text{c}} > 15$ (cf table 1), the reflectivity due to convective amplification is already sufficiently large to incur nonlinear saturation effects.

Figure 4.

Figure 4. The saturated reflectivity versus kl λDe from the TWC model (solid lines) for (a) I15 = 2.82 and (b) I15 = 0.11 at m = 2 (in red) and m = 2.9 (in blue). Correspondingly, the temporal evolution of the reflectivity for cases indicated by asterisks, crosses and triangles in (a) and (b) is presented in (c) and (d), respectively. In (a) and (b), the reflectivity versus kl λDe calculated by the Tang's formula is also plotted as dotted lines, while ${[{k}_{l}{\lambda }_{\text{De}}]}_{c}$ for absolute growth calculated from the linear criterion equation (23) is indicated for both m = 2 and m = 2.9.

Standard image High-resolution image

Table 1. Summary of critical parameters for the transition from convective to absolute SRS growth from the linear TWC model and from the Vlasov–Maxwell simulation. The homogeneous He plasma with ne = 0.1 nc , Ti = Te /5 and length of  200λ0  (λ0 = 351 nm) is taken.

I15 m Linear TWCVlasov–Maxwell
${[{k}_{l}{\lambda }_{\text{De}}]}_{\text{c}}$ $\frac{{[{\nu }_{l0}]}_{\text{c}}}{0.01{\omega }_{0}}$ ${[{G}_{\text{R}}]}_{\text{c}}$ ${[{k}_{l}{\lambda }_{\text{De}}]}_{\text{c}}$ $\frac{{[{\nu }_{l0}]}_{\text{c}}}{0.01{\omega }_{0}}$ ${[{G}_{\text{R}}]}_{\text{c}}$
2.8220.2780.20820.50.422.581.68
 2.90.3380.24116.40.4652.151.84
0.1120.2380.03584.990.2690.1541.12
 2.90.2950.04293.920.3160.1121.45

In addition to decreasing kl λDe, increasing m also leads to a decrease of νl0, and thus can result in the transition from convective to absolute SRS in certain regions of kl λDe. Comparing the lines with asterisks in figures 4(c) and (d), for the same laser and plasma parameters, SRS is in the convective regime when m = 2 but can grow absolutely and eventually saturated by nonlinear effects when m = 2.9. Consequently, ${[{k}_{l}{\lambda }_{\text{De}}]}_{\text{c}}$ becomes larger for a greater m, as listed in table 1, indicating that the Langdon effect can broaden the parameter range for absolute SRS instability as shown in figure 4. In table 1, it is found that the difference in ${[{k}_{l}{\lambda }_{\text{De}}]}_{\text{c}}$ for different m results in close (slightly higher for increasing m) values of ${[{\nu }_{l0}]}_{\text{c}}$ in the linear prediction. This is because in equation (23), γ0 and vs are almost independent of kl λDe, while according to the matching condition of SRS, it can be proved that ${v}_{l}\approx 3{v}_{\text{the}}^{2}{k}_{l}/{\omega }_{l}$ increases moderately with larger ${[{k}_{l}{\lambda }_{\text{De}}]}_{c}$ at greater m, in contrast to the strong rise of νl0 with increasing kl λDe.

Considering that nonlinear kinetic effects can not be included by the TWC model, results from Vlasov–Maxwell simulations are shown in figure 5 for some cases. Due to the nonlinear kinetic effects, the convective SRS in the linear prediction can become absolute. One example is presented in figure 5(c), where the initial growth of SRS agrees well with the TWC model before t < 500T0, after which the TWC model predicts convective saturation while the Vlasov–Maxwell simulation exhibits the continual growth of SRS toward much higher reflectivity. A detailed investigation shows that with the growth of the EPW field, resonant electrons with vx vphl are trapped, resulting in flattening of the EEDF around vphl, as shown in figure 5(d). Correspondingly, the Landau damping is reduced [34, 35] while the EPW frequency is downshifted causing a frequency mismatch [3638]. This in turn induces a frequency upshift of the scattered wave that tends to restore the frequency matching resonance, as shown in the bottom panel of figure 5(c) for the adjustment period  500T0 < t < 2700T0. The frequency mismatch induces a phase mismatch δmis and thus a reduction of growth rate by  cos δmis, impairing the SRS growth, while the nonlinear reduction of νl favors the SRS growth. In the case shown in figure 5(c), the competition between these two factors results in oscillation of the reflectivity during  1500T0 < t < 2500T0, wherein significant frequency mismatch exists in the plasma region further away from the left boundary (e.g. x/λ0 = 140 in figure 5(c)). After a period of adjustment, ultimately at t > 2700T0 over a large plasma region the frequency upshift of the scattered wave becomes sufficiently large to compensate the frequency downshift of the EPW, reducing the frequency mismatch to a low level. Consequently, the nonlinear absolute growth rate ${\gamma }_{\text{abs,NL}}\approx 2{\gamma }_{0}\sqrt{{v}_{l}/{v}_{s}}\,\cos \,{\delta }_{\text{mis}}-{\nu }_{\text{NL}}$ exceeds zero over a large plasma region, and SRS enters into the absolute growth period. Until when t > 3100T0 the EPW amplitude in the plasma region near the left boundary (e.g. x/λ0 = 60 in figure 5(c)) is so large that the frequency downshift of the EPW can not be completely compensated to maintain the SRS resonance, resulting in saturation of the absolute SRS growth.

Figure 5.

Figure 5. The early-stage evolution of the instantaneous reflectivity for (a) I15 = 2.82 and (b) I15 = 0.11. The circles denote tearly defined as the time until when 90% of the EPW energy is contained within  [0.98klc , 1.02klc ]. The diamonds denote tsat corresponding to the turning point of the reflectivity, which is determined by fitting  log10R versus t with a piecewise linear function. The case m = 2 and kl λDe = 0.415 in (a) is replotted in (c), where the reflectivity versus time is compared to the prediction of TWC model in the upper panel, while the frequency shifts of the EPW and the scattered wave are displayed in the bottom panel for distances x = 60λ0 and x = 140λ0 from the left boundary. In (d), the flatted EEDF is shown for two cases in (a) at m = 2 and m = 2.9, where vphl is indicated by vertical dotted lines. The condition ne = 0.1  nc , λ0 = 351 nm for a homogeneous He plasma with Ti = Te /5 and length of 200λ0 is taken.

Standard image High-resolution image

To illustrate the Langdon effect on the early-stage saturation of SRS in Vlasov–Maxwell simulation, we need to measure the early-stage saturated reflectivity Rsat,e. However, as shown in figures 2(b)–(f) and 3(b)–(f), for the absolute SRS, the reflectivity versus time after the saturation is irregular and nonstationary, making it necessary to clearly define Rsat,e in a reasonable way. Here we define Rsat,e as the averaged reflectivity over the time period from the onset of early-stage saturation  (tsat) until the onset of significant secondary instabilities  (tearly). In practice, tsat can be identified as the time when the growth of the reflectivity becomes flattened, and tearly can be defined as the time until when  90% of the EPW field energy is contained within  [0.98klc , 1.02klc ], as indicated in figures 5(a) and (b). For the convective SRS growth, on the other hand, the reflectivity would eventually become constant with time or in some cases oscillate in a regular way and hence have a constant mean value; correspondingly, it is natural to identify the saturated reflectivity Rsat as the steady (mean) value of the reflectivity.

In Vlasov–Maxwell simulation, Rsat,e versus kl λDe is shown in figures 6(a) and (b) for different m, whereas tsat versus kl λDe is displayed in figures 6(c) and (d) for cases with absolute SRS growth. As a comparison, Rsat,e with an alternative definition of tearly, i.e. the time until when 90% of the EPW energy is contained within  [0.95klc , 1.05klc ], as well as the time-averaged reflectivity over the entire simulation time with ttsat, is also shown. As seen, Rsat,e is insensitive to the individual choice in the definition of tearly. In fact, for relatively large kl λDe, kinetic electron trapping plays a key role. In such cases, the nonlinear evolution in the early stage is predominately determined by (i) nonlinear reduction of Landau damping, (ii) nonlinear frequency downshift of EPW and the accompanied frequency upshift of the scattered wave, and (iii) pump depletion that becomes important when the reflectivity rises to a high level; while secondary instabilities that significantly broaden the kl -spectrum of EPW and denote the end of the early stage, mainly consist of trapped particle instability and generation of beam acoustic modes (cf section 3.2). As shown in figures 2(a)–(f) and 3(a)–(e), broadening of the kl -spectrum of EPW is sudden and abrupt. Consequently, tearly can be delineated with a great accuracy, yielding a robust Rsat,e. Things are different for the three low- kl λDe cases indicated by open squares in figures 6(b) and (d), where the early-stage saturation is predominately caused by pump depletion, after which broadly featured LDI begins to become important, resulting in broadening of the kl -spectrum. As shown in figure 3(f), in these cases broadening of the kl spectrum is much more gradual, making tearly prone to individual choices. However, Rsat,e remains insensitive to individual choices of tearly since Rsat,e is mainly contributed by the first reflectivity peak with large width and high amplitude, which is mainly saturated by the pump depletion and hence always contained within t < tearly.

Figure 6.

Figure 6. (a) and (b) Rsat,e versus kl λDe at different m in Vlasov–Maxwell simulations for (a) I15 = 2.82 and (b) I15 = 0.11. Three types of cases are distinguished: (I) absolute growth when the kl -spectrum broadening of EPW due to secondary instabilities is abrupt (open circles). (II) Absolute growth when the kl -spectrum broadening of EPW is gradual (open squares). (III) Convective growth with rather weak nonlinear effects (solid circles). For comparison, Rsat,e calculated using an alternative definition of tearly (90% of the EPW field energy is contained within  [0.95klc , 1.05klc ]) is displayed by the plus symbols, while the time-averaged reflectivity over the entire simulation time with t > tsat is shown as open triangles. (c) and (d) tsat versus kl λDe at different m for cases with absolute growth. The condition ne = 0.1 nc , λ0 = 351 nm for a homogeneous He plasma with Ti = Te /5 and length of 200λ0 is taken.

Standard image High-resolution image

In figures 6(a) and (b), with decreasing kl λDe, an abrupt rise in Rsat,e appears at one critical ${[{k}_{l}{\lambda }_{\text{De}}]}_{\text{c}}$, where the transition from convective to absolute SRS growth occurs. ${[{k}_{l}{\lambda }_{\text{De}}]}_{\text{c}}$ as obtained from the Vlasov–Maxwell simulations, as well as the initial Landau damping νl0 and the convective gain GR corresponding to ${[{k}_{l}{\lambda }_{\text{De}}]}_{\text{c}}$, is listed in table 1 for both m = 2 and m = 2.9 at I15 = 2.82 and I15 = 0.11. It can be seen that due to nonlinear kinetic effects, ${[{k}_{l}{\lambda }_{\text{De}}]}_{\text{c}}$ can far exceed that from the TWC model, leading to much smaller ${[{G}_{\text{R}}]}_{\text{c}}$ and hence a sharp rise in the reflectivity at the transition. Nevertheless, even in the presence of kinetic effects, ${[{k}_{l}{\lambda }_{\text{De}}]}_{\text{c}}$ is larger for a greater m. Also the values of ${[{\nu }_{l0}]}_{\text{c}}$ corresponding to ${[{k}_{l}{\lambda }_{\text{De}}]}_{\text{c}}$ are quite close for different m, as in the prediction of the TWC model. However, contrary to the TWC model, in the Vlasov–Maxwell simulation ${[{\nu }_{l0}]}_{\text{c}}$ is lower for larger m. To comprehend this point, notice that after a proper account of the nonlinear modification to νl and δωl by kinetic effects, equations (16)–(18) can still be used to describe the early-stage evolution of SRS. Various approximations for the nonlinear evolution of νl and δωl have been proposed in the literature, of which one relatively simple bounce-averaged model is given in references [35, 39] as

Equation (34)

Equation (35)

where the bouncing frequency ${\omega }_{B}=\sqrt{e{E}_{l}{k}_{l}/{m}_{e}}$. In this modified TWC model, when other parameters such as γ0, vl , vs and v0 are kept nearly the same, the initial Landau damping νl0 and δωl /ωB that depicts the strength of nonlinear frequency shift, determine the nonlinear evolution of SRS, and hence whether the absolute SRS growth can occur or not. Since the increase of either the Landau damping or the frequency shift would impair the SRS growth, it can be expected that with increasing δωl /ωB , a lower νl0 is required for the onset of absolute SRS growth. This is indeed the effect of increasing m, which leads to greater δωl /ωB at the same νl0, as elucidated in figure 7. Consequently, ${[{\nu }_{l0}]}_{c}$ must be smaller for increasing m to overcome the effect of stronger nonlinear frequency shift at greater m. It can be further understood that the different variation of δωl and νl0 with m ultimately results from δωl ∝ ∂2 fe0/∂2 vx , in contrast to νl0 ∝ ∂fe0/∂vx [7]. This general understanding should hold even though the simplified model specified by equations (34) and (35) is not precise, indicating that apart from the Landau damping, the influence of super-Gaussian EEDFs on the nonlinear frequency shift is also an important factor to affect the early-stage development of SRS.

Figure 7.

Figure 7. Left axis. kl λDe versus νl0 for m = 2 (in red) and m = 2.9 (in blue), where ${[{k}_{l}{\lambda }_{\text{De}}]}_{\text{c}}$ from the Vlasov–Maxwell simulation is indicated by the plus and asterisk symbols. Right axis. δωl ω0/νl ωB versus νl for m = 2 and m = 2.9, where the nonlinear frequency shift δωl is estimated using equation (35).

Standard image High-resolution image

Below ${[{k}_{l}{\lambda }_{\text{De}}]}_{\text{c}}$, Rsat,e is quite insensitive to kl λDe, except for an intermediate range of kl λDe ∼ 0.33–0.4 for m = 2.9 and kl λDe ∼ 0.28–0.35 for m = 2, where Rsat,e increases with decreasing kl λDe as shown in figure 6(a). For the high kl λDe range ( kl λDe > 0.4 for m = 2.9 and kl λDe > 0.35 for m = 2), the phase velocity of the EPW is located in the bulk region of the EEDF, leading to strong Landau damping and also strong electron-trapping induced nonlinearity (e.g., νl0/γ0 ≫ 1 and δωl /γ0 ≫ 1). Despite the difference in the initial Landau damping, after the adjustment period, the absolute SRS growth become similar for different kl λDe and m, until when the reflectivity reaches the level $\sim 0.02$, where secondary instabilities become important and the early stage ends, as shown in figure 5(a). As a result, the dependence of Rsat,e on both kl λDe and m is much weaken. In the intermediate range of kl λDe with weaker Landau damping and kinetic nonlinearity, the Landau damping and the kinetic nonlinear shift are comparable to γ0 (e.g., νl0 ∼ 0.5–2.9 and δωl /γ0 ∼ 0.3–1.3 for kl λDe ∼ 0.28–0.35 at m = 2, I15 = 2.82 and δne /ne0 = 0.02). Thus, the nonlinear adjustment is insufficient to smear the effects of decreasing initial Landau damping and nonlinear frequency shift when kl λDe decreases or m increases. Consequently, Rsat,e increases with decreasing kl λDe or increasing m. For the low kl λDe range (kl λDe < 0.33 for m = 2.9 and kl λDe < 0.28 for m = 2) with even smaller initial Landau damping, after the adjustment period, the Landau damping is nearly negligible. The absolute growth rate ${\gamma }_{\text{abs,NL}}\lesssim \mathrm{max}[{\gamma }_{\text{abs}}]\equiv 2{\gamma }_{0}\sqrt{{v}_{l}/{v}_{s}}$, and also the evolution of SRS towards the early-stage saturation, becomes quite similar for different kl λDe and m, as shown in figure 5(b). As a result, Rsat,e again becomes nearly independent of kl λDe and m.

In figures 6(c) and (d), it can be seen that the saturation time tsat is quite large near ${[{k}_{l}{\lambda }_{\text{De}}]}_{\text{c}}$. As shown in figure 5(c), here an oscillating plateau of the reflectivity is formed before the absolute growth period, significantly lengthening the adjustment period. This is because the counteracting effects of nonlinear Landau damping reduction and the frequency shift induced phase mismatch, nearly balance during the adjustment period. A slight decrease of kl λDe or increase of m, weakens both the Landau damping and the frequency shift, thus breaks the balance and significantly reduces the adjustment time, leading to a sharp drop of tsat with decreasing kl λDe or increasing m near ${[{k}_{l}{\lambda }_{\text{De}}]}_{\text{c}}$, as shown in figures 6(c) and (d). This decreasing trend of tsat holds until for kl λDe far below ${[{k}_{l}{\lambda }_{\text{De}}]}_{\text{c}}$, where tsat tends to a value nearly independent of kl λDe and m. Here, the SRS growth is absolute even in the absence of nonlinear kinetic effects, the adjustment period is negligible, and tsat is primarily contributed by the absolute growth period. The absolute growth rate ${\gamma }_{\text{abs}}\approx 2{\gamma }_{0}\sqrt{{v}_{l}/{v}_{s}}$ is nearly independent of kl λDe and m, and hence so is tsat.

3.2. The late-stage saturation of SRS

Now we examine the impacts of the Langdon effect on the nonlinear saturation of SRS in the late stage when secondary instabilities are important. Since the nonlinear behavior and dominant saturation mechanism in the late stage are quite different between the high kl λDe regime  (0.25 ≲ kl λDe ≲ 0.45) which is investigated under a high intensity drive  (I15 = 2.82), and the low kl λDe regime  (0.18 ≲ kl λDe ≲ 0.3) that is studied under a low intensity drive  (I15 = 0.11), in the following we discuss them separately.

For I15 = 2.82 and Te = 3.2 keV as shown in figures 2(c) and (d) for m = 2 and m = 2.9 respectively, a burst behavior of the reflectivity is exhibited. In the active phase, the trapped particle induced nonlinearity ultimately results in a chaotic state, wherein a nearly continuous kl -spectrum of the EPW field spreading from klc ≈ 1.5ω0/c to about ω0/c is generated. As can be seen, the downward broadening of the wavenumber is quite abrupt. It is caused by vortex-merging processes in the phase space, which begins with pairwise merging of phase-space 'holes' or vortices near the phase velocity of EPW [40, 41], as shown in figure 8 near t ≈ 1500T0 for m = 2 and t ≈ 920T0 for m = 2.9, respectively. This in turn gives rise to broadband incoherent EPW field consisting of beam acoustic modes [1618]. The decay of the resonant (and usually downshifted) EPW into BAMs breaks the three wave resonance condition of SRS, and thus serves as an efficient saturation mechanism for SRS. The corresponding typical kl ωl spectrum is shown in figure 9(b), where the BAM feature is obvious. Note that the adjusted resonant point at  (ωl , kl ) ≈ (0.37ω0/c, 1.50ω0/c), which corresponds to the intersection of BAMs with the Stokes curve, is downshifted relative to the linear resonance mode (ωlc , klc ) = (0.39ω0, 1.47ω0/c). Correspondingly, the backscattered feature in the transverse electric field at  (ωs , ks ) ≈ (0.63ω0, − 0.55ω0/c), as shown in figure 9(a), is upshifted relative to the linear resonance mode  (ωsc , ksc ) ≈ (0.61ω0, −0.52ω0/c). For m = 2 where the bursts are well separated, the temporal-spatial evolution of the EPW, the scattered wave and the pump wave is shown in figures 10(a)–(c). It can be seen that BAMs are developed between  1400T0 and  1800T0, while their convection and damping lead to oscillation of the reflectivity during this period. Then at 2000T0, a strong peak occurs. Accordingly, the laser pump is depleted, while the decay to BAMs is significantly enhanced by the strong EPW field. As a result, the reflectivity quickly drops to near zero at t ∼ 2200T0. Then, the pump intensity is restored. However, the BAM packet needs a much longer time  (∼L/vl ) to damp or convect through the plasma [42]. Inside the packet, the rapid decay to BAMs caused by the large EPW field keeps the (upshifted) backscattered light at a low level. As this scattered light moves outside the packet into the unperturbed plasma region behind the packet, it is off-resonance with  Δωδωl . This limits its further amplification, and also produces a beat pattern separated at τ = 2πω [43]; correspondingly, many high frequency minor bursts appears during the quiescent period between 2500T0 and  3000T0. When the packet has almost convected out of the plasma  (t ∼ 3200T0), a new major burst coming from new SRS growth in the plasma, now almost clear of the incoherent BAMs, occurs. Again, the pump depletion that takes effect instantaneously when the reflectivity is large, and the generation of BAMs whose effects last a long time, begin to suppress the reflectivity. The continual suppression and recovery of SRS lead to a sequence of major bursts with period about 1300T0, intervened by many minor bursts. For m = 2.9, the decay to BAMs and the pump depletion play similar roles and act as the predominant saturation mechanism for tearly < t < 4000T0. Nevertheless, as shown in figures 10(d) and (f), due to the smaller Landau damping and hence the greater growth rate and gain of SRS at larger m, significant SRS re-growth can occur behind the packet even when less than half the plasma is clear of the incoherent BAM field, permitting several packets to coexist and interact. For example, at t ∼ 2000T0, new packet II is formed near the left boundary when the previous packet I has just convected half through the plasma. The bursts of the reflectivity at t ∼ 2200T0 are generated deep inside packet I at x ≈ 150λ0, and further amplified across packet II on its path to the left boundary, similar to the high gain case in [43]. So, the bursts become overlapped, while the significant interaction between bursts leads to a less regular burst behavior. Consequently, compared to m = 2, the quiescent period is significantly reduced, and the average reflectivity is enhanced. Besides, over a long timescale t > 4000T0, in addition to decay to BAMs, rescattering also become important, featured by the remarkable kl -features near −0.54ω0/c [53]. The typical ωs ks spectrum of the scattered wave, and the corresponding ωl kl spectrum of the EPW, are shown in figures 9(c) and (d), respectively, where the feature with  (ωl , kl ) ≈ (0.32ω0, −0.54ω0/c) and  (ωs , ks ) ≈ (0.3ω0, 0) is due to rescattering of the primary upshifted backward scattered wave with  (ωs , ks ) ≈ (0.64ω0, −0.54ω0/c). Consequently, the SRS becomes more chaotic, leading to more irregular variation of the reflectivity.

Figure 8.

Figure 8. The EEDF averaged over one EPW wavelength  ([200π/klc , 202π/klc ]) exhibits the formation of holes near the phase velocity for (a) m = 2  (t ⩽ 1400T0) and (d) m = 2.9  (t ⩽ 930T0), which is smoothed out by vertex merging processes at t ⩾ 1600T0 for m = 2 and t = 960T0 for m = 2.9, respectively. And the phase space vortex-merging of EEDF begins with pairwise coalescence in phase space near (b) and (c)  t ≈ 1500T0 for m = 2 and (e) and (f) t ≈ 920T0 for m = 2.9. The condition ne = 0.1 nc , λ0 = 351 nm, Te = 3.2 keV, and I15 = 2.82 for a homogeneous He plasma with Ti = Te /5 and length of 200λ0 is taken.

Standard image High-resolution image
Figure 9.

Figure 9. Some ωk spectra of (a), (c) and (e) the transverse electric field  [log10|Ey (ω, k)|2] and (b), (d) and (f) the electrostatic field  [log10|El (ωl , kl )|2] for cases in figure 2. The dispersion relations for the EPW and the scattered wave are shown as white dashed lines. The Stokes curves as the locus of the EPW modes that is phase matched for the electromagnetic decay of the pump, i.e.,  (ω0ωl , kl k0) satisfies the dispersion relation of the EMW, is plotted as yellow lines. The condition ne = 0.1 nc , λ0 = 351 nm for a homogeneous He plasma with Ti = Te /5 and length of 200λ0 is taken.

Standard image High-resolution image
Figure 10.

Figure 10. The temporal and spatial evolution of (a) and (d) the electrostatic fields and the transverse electric fields of (b) and (e) the scattered wave and (c) and (f) the pump wave for m = 2 and m = 2.9. The root mean square over one wavelength is taken for each field. The reflectivity versus time is shown for comparison as black lines in the right region of panels (a) and (d). The condition ne = 0.1nc , λ0 = 351 nm for a homogeneous He plasma with Ti = Te /5 and length of 200λ0 is taken.

Standard image High-resolution image

For Te = 2.8 keV and I15 = 2.82 as shown in figures 2(e) and (f), the evolution of SRS after t > tearly can be further divided into three periods: I. In the initial period with t < 4000T0, the generation of BAMs plus the pump depletion remains the dominant saturation mechanism. The SRS reflectivity consists of overlapped bursts for both m = 2 and m = 2.9, and though less apparent, still the quiescent period is shorter for m = 2.9. II. In the intermediate period with  4000T0 < t < 6000T0 for m = 2 and  4000T0 < t < 7000T0 for m = 2.9, rescattering aids in limiting the SRS level. III. In the final period with t > 6500T0 for m = 2 and t > 8500T0 for m = 2.9, the decay to low- k Langmuir branch (including modes due to rescattering and forward SRS) is significantly enhanced, leading to the kl -spectrum continuously ranging from klc down to zero. As shown in figure 9(f) for the typical ωl kl spectrum of the EPW, both BAMs and the low- k Langmuir branch with features due to rescattering and forward SRS contained are nearly fully occupied. Correspondingly, the ωs ks spectrum of the scattered wave shown in figure 9(e) exhibits features of rescattering with (ωs , ks ) around (0.3ω0, 0) and forward SRS with (ωs , ks ) around (0.68ω0/c, 0.61ω0/c). Such fully developed incoherence results in the drop of SRS reflectivity in the final period, as demonstrated in figures 2(e) and (f).

The variation of the late-stage saturation mechanism with Te , together with the corresponding average reflectivity, is summarized in figure 11(a) for m = 2 and m = 2.9 at I15 = 2.82. For all cases, initially the dominant saturation mechanism is decay to BAMs plus the pump depletion. In this period, the lower Landau damping and hence greater gain for increasing m results in shorter quiescent period and hence greater average reflectivity. With decreasing Te or increasing m, rescattering and decay to low-kl Langmuir branch can become important in the later period, usually leading to a more chaotic state with reduced average reflectivity. Consequently, the dependence of the averaged reflectivity on m becomes more uncertain in the later period, and the average reflectivity can be smaller for increasing m in some cases.

Figure 11.

Figure 11. The average reflectivity and dominant secondary processes during the late stage of SRS. In (a) with higher kl λDe and I15 = 2.82, three saturation mechanisms are distinguished: the decay to BAMs (open squares), decay to BAMs plus rescattering (open diamonds), and decay to BAMs and low-kl Langmuir branch (open circles). In (b) with lower kl λDe and I15 = 0.11, four saturation mechanisms are distinguished: TPI (open squares), TPI plus LDI (open triangles), LDI (open diamonds) and LDI cascade (open circles). When there is a distinct change in the dominant saturation mechanism, the average reflectivities are calculated separately for the initial period and the final period. Notice that in all cases the pump depletion can also play some role in the saturation of SRS, especially at the time with large instantaneous reflectivity. The condition ne = 0.1nc , λ0 = 351 nm for a homogeneous He plasma with Ti = Te /5 and length of 200λ0 is taken.

Standard image High-resolution image

For Te = 1.5 keV and I15 = 0.11 as shown in figures 3(c) and (d), in the initial period after t > tearly (t < 12000T0 for m = 2 and t < 7000T0 for m = 2.9), the dominant secondary process is trapped particle instability (TPI) [4446], featured by the appearance of two sidelobes with kl = klc ± ΔkTPI around the primary component klc . The sidelobe frequency is resonant with the electron bouncing frequency in the wave frame, thus satisfying ΔωTPI − ΔkTPI vphl = ±ωB [45], where ΔωTPIωTPIωlc ,  ΔkTPI = kTPIklc , and vphl = ωlc /klc is the EPW phase velocity. As shown in figure 12(a) for the typical ωl kl spectrum of the EPW when TPI dominates, the most significant mode occurs along the dispersion relation of EPW, thus ΔωTPIkTPI = vl , giving ΔkTPI = ±ωB /(vphlvl ). For Te = 1.5 keV, using  2πeEl /me ω0 c ≈ 0.01 estimated from the simulation data, it can be obtained  ΔkTPI ≈ ±0.26ω0/c, consistent with the sidelobe locations on the kl -spectrum as shown in figures 3(c) and (d).

Figure 12.

Figure 12. Some ωl kl spectra of the electrostatic field [log10|El (ωl , kl )|2] for cases in figure 3. In each panel, the upper part shows features due to the EPW, while the bottom part demonstrates features due to the LDI generated IAW. The dispersion relations for the EPW are shown as white dashed lines. The Stokes curves as the locus of the EPW modes that is phase matched for the electromagnetic decay of the pump, i.e.,  (ω0ωl , kl k0) satisfies the dispersion relation of the EMW, is plotted as yellow lines. The condition ne = 0.1nc , λ0 = 351 nm for a homogeneous He plasma with Ti = Te /5 and length of 200λ0 is taken.

Standard image High-resolution image

As time evolves, Langmuir decay instability [21, 23], where a primary Langmuir wave (LW) decays into a secondary Langmuir wave and an ion acoustic wave (IAW), begins to become important. LDI satisfies the matching condition

Equation (36)

where i denotes the stage number of the Langmuir cascade with i = 0 corresponding to the primary LW, and ωa and ka are the frequency and wave number of the IAW, respectively. The dispersion relation for the LW can be approximated by ${\omega }_{l,i}^{2}={\omega }_{\text{pe}}^{2}+3{k}_{l,i}^{2}{v}_{\text{the}}^{2}$, while the dispersion relation for the IAW is approximately ωa = |ka |cs , where ${c}_{s}\approx \sqrt{Z{T}_{e}/M}$ is the acoustic velocity with Z and M being the charge and mass of the ion species. Substituting these dispersion relations into equation (36) yields kl,i+1 ≈ −kl,i + ΔkLDI and ka ≈ 2kl,i , where the wavenumber difference between two successive cascade step is

Equation (37)

The LDI threshold can be estimated by [47]

Equation (38)

where νa /ωa ≈ 0.099 for Te /Ti = 5 in He plasma considered here. For m = 2 and Te = 1.5 keV, when t ≈ 10000T0, El /El,LDI ≈ 1.14, LDI with one cascade step is excited and results in the reduction of reflectivity in the later period, as seen from figure 3(c). The corresponding ωl kl spectrum of the electrostatic field is shown in figure 12(b), where the features at kl ≈ −1.51ω0/c + 0.07ω0/c due to the secondary LW and at kl ≈ 3ω0/c due to IAW are obvious. For m = 2.9, due to decreasing νl with increasing m, El,LDI is reduced while El is generally greater, leading to El /El,LDI ≈ 15 at t ≈ 7000T0. Consequently, at least five LDI cascade steps are apparent from figures 3(d) and 12(c) for t > 10000T0, while the sidelobes of TPI become insignificant. This indicates that LDI cascade becomes the dominant saturation mechanism, limiting the reflectivity to a very low level about  5% for t > 15000T0.

For Te = 1.2 keV and m = 2 as shown in figure 3(e), in the initial period t < 8000T0, TPI plus the pump depletion is still the dominant saturation mechanism. However, when t ∼ 8000T0, El/El,LDI ≈ 4, so LDI cascade can be excited and the reflectivity in the later period is significantly reduced, as shown in figures 3(e) and 12(d). For Te = 1.2 keV and m = 2.9 as shown in figure 3(f), νl /ωl ∼ 10−6 is quite small. As a result, SRS is strongly driven and grows rapidly at the early time, leading to El/El,thr ≈ 100 and δne /ne ≈ 0.06 at t = 5000T0. Broad-featured LDI is developed, as shown in figure 12(e) for  6000T0 < t < 8000T0. As more cascade steps are excited, the kl -spectrum is gradually broadened (in contrast to the sudden broadening of the kl -spectrum in other cases). When t ∼ 10000T0, multiple cascade steps with broad spectral features have been excited, as shown in figure 12(f). In this strongly-driven regime, SRS is quite turbulent, and the instantaneous reflectivity varies over a wide range. This leads to a greater average reflectivity in the later period compared to m = 2.

The variation of the late-stage saturation mechanism with Te , together with the corresponding average reflectivity, is summarized in figure 11(b) for m = 2 and m = 2.9 at I15 = 0.11. Except for the three strongly driven cases with low Te (Te = 0.8 keV and m = 2, Te = 0.8 keV and m = 2.9, Te = 1.2 keV and m = 2.9), initially the dominant mechanism is TPI plus the pump depletion, while LDI and LDI cascade can develop over time for some Te and m, significantly reducing the reflectivity. Increasing m is favorable for excitation of LDI or LDI cascade since the LDI threshold is reduced while the EPW amplitude is typically greater before the onset of LDI due to the lower Landau damping. This can in turn result in a much stronger drop of the reflectivity in the later period than m = 2, thus the reflectivity at m = 2.9 in the later period can be smaller than m = 2.

4. Discussion and summary

In summary, the influence of the Langdon effect on the nonlinear evolution of SRS over a long timescale is investigated for a wide range of plasma parameters. For the early stage of SRS, it is found that the Langdon effect can significantly widen the parameter range for absolute SRS growth, and the kinetic nonlinear effect can widen this parameter range further. The time for SRS to reach the early-stage saturation is significantly reduced by the Langdon effect except when kl λDe is far below ${[{k}_{l}{\lambda }_{\text{De}}]}_{\mathrm{c}}$. For the late stage of SRS, at high kl λDe, initially the dominant saturation mechanism is decay to BAMs plus the pump depletion, wherein the time-varying reflectivity is composed of a series of (possibly overlapped and irregular) bursts. The Langdon effect can shorten the quiescent period and hence increase the average reflectivity. Additional secondary instabilities such as rescattering of the primary scattered wave, and the generation of low- k Langmuir branch can also develop over time, typically further reducing the reflectivity. The Langdon effect favors the development of additional secondary processes, though the effect is generally too weak to make a great difference in the reflectivity. At low kl λDe, the saturation mechanisms include TPI plus the pump depletion, which dominates in the initial period of the late stage except for very low kl λDe, and LDI with single or multiple cascade steps, which typically becomes important in the later time period. The Langdon effect decreases the threshold for LDI, thus LDI or LDI cascade with more cascade step is easier to excite. This can significantly suppress the reflectivity at the later time, even leading to smaller reflectivity than m = 2.

This work, in combination with the study of the Langdon effect on the convective SRS in reference [7], indicates the importance of the Langdon effect to both the linear and nonlinear development of SRS. The reduction of Landau damping due to the Langdon effect and nonlinear kinetic trapping, as well as the resultant change from convective to absolute SRS growth, can serve as possible reasons for the underestimated SRS reflectivity obtained by the convective gain model in numerical analysis of the experiments under the assumption of Maxwellian EEDFs [8, 9, 11], which is helpful for further improvement to the physical modeling and simulation of SRS. However, for SRS in the complicated realistic plasma of ICF experiments [9, 32, 48, 49], it should be noted that nonuniform plasma condition [10] and the high dimensional effects such as wavefront bowing, self-focusing, sideloss, etc [5052], which are not present in the current model, also play important roles in the evolution of SRS. So, further work with multi-dimensional effects considered should be done under more realistic laser and plasma conditions to assess the importance of the Langdon effect to SRS in ICF experiments.

Acknowledgments

This work was supported by the National Natural Science Foundation of China (Grant Nos. 11875093, 12275032, 12205021 and 11875091), and the Project supported by CAEP Foundation (Grant No. CX20210040).

Please wait… references are loading.