A linear fluid model defined by the continuity equation, and the momentum and energy balance equations, valid for describing high temperature and collisionless plasmas in the presence of an electromagnetic (EM) wave is proposed. The collisionless closure relations, the ponderomotive force, and the absorption rate of the EM wave are calculated for a relativistic temperature regime while assuming non-relativistic electron oscillation velocities in the laser electric field. The model equations coupled with the Maxwell's equations are used to study the filamentation instability (FI) in the context of inertial confinement fusion. Both the stationary and non-stationary filamentation instabilities are investigated. It is shown that the ponderomotive source of the instability is preponderant compared to the thermal source, and that the ponderomotive spatial growth rate increases significantly with the plasma temperature. The FI growth rates are compared to those of stimulated Brillouin and Raman scattering. At high temperatures, significant amplification of the filamentation instability is expected, potentially leading to non-uniform target compression, which could be detrimental for fusion.

Currently, there is growing interest to study plasmas at high temperature plasmas. Indeed, with the advancements in laser technology, powerful laser pulses are now evaluable, and their interaction with solid targets, as in inertial confinement fusion (ICF), produces high temperature plasmas up to 20keV. At such high temperatures, electron rest energy mec2 becomes non-negligible compared to the electron thermal energy Te, where me is the electron rest mass, c is the speed of light, and Te is the electron temperature in energy units. The relevant parameter that quantifies the relativistic effects is therefore z=mec2Te. Thus, as temperature increases, the relativistic parameter z decreases, and relativistic effects become more significant.

In this high plasma temperature range, the relativistic effects must be taken into account, and in such plasmas several physical phenomena should be revisited.1–6 Many works dedicated to this new relativistic regime have been reported in the literature. Using simulations, Langdon and Hinkel1 studied the nonlinear evolution of backscattering Brillouin and Raman instabilities in high temperature plasmas (320keV). In particular, the effects of the rescattering of these instabilities on the generation of hot electrons and scattering losses have been highlighted. In Ref. 2, the authors derived from the relativistic Vlasov equation the dispersion relations for electrostatic and electromagnetic (EM) waves. Using a different approach, the Vlasov equation was solved analytically in Ref. 3, and the dispersion relation of the electrostatic waves was also deduced. Bers et al. presented in Ref. 4 a new formulation for the damping rate of electron plasma waves in relativistic plasma temperatures (515keV). This result was used to investigate forward and backward Raman backscattering instabilities in the context of inertial confinement fusion. Likewise, Li et al.5 revisited the laser hosing instability in relativistic plasmas taking into account both thermal and fluid relativistic effects. The model equations were used to study short-pulse and long-pulse laser hosing instabilities using a variational method approach. Similarly, Zhao et al.6 studied Raman and Brillouin parametric instabilities induced by high intensity laser pulses propagating into an underdense relativistic plasma.

In inertial confinement fusion, it is well known that in laser-created plasmas, the EM modes couple nonlinearly with plasma eigenmodes such as electron plasma wave or ion acoustic wave, to drive parametric instabilities. These instabilities depend mainly on the nature of the interacting modes and can be broadly classified into 3-mode couplings such as Brillouin instabilities (coupling of the EM wave with an ion acoustic wave) and Raman instabilities (coupling of the EM wave with an electron plasma wave), and to 4-mode couplings such as the filamentation instability (FI) and the modulational instability. The parametric instabilities play a detrimental role in the context of inertial confinement fusion, since they mainly reduce the heating rate by the laser beam of the target and prevent an efficient compression of the target. Numerous analytical and numerical works devoted to these instabilities have been reported in the literature in order to reduce these damaging effects.7,8 The present work focuses on the FI which plays a central role in laser plasma coupling. The local increase in the intensity of the laser beam can alter transport properties and generate hot electrons. In addition, the interaction of the FI with other instabilities generated in the underdense plasma plays a detrimental role for achieving successful inertial confinement fusion.

The FI is triggered when a perturbation in the intensity of the laser beam propagating through plasma begins to grow in amplitude, forming filaments. In this work, we consider both sources of laser beam filamentation: the inverse bremsstrahlung absorption and the ponderomotive force. For both mechanisms, the ponderomotive and thermal forces push electrons from the center of the filament to the edges, creating a density depression that locally modifies the spatial refractive index profile. The refractive index, given by nref=1nencr, where ne is the electron density and ncr is the critical density, becomes higher in the filament's center than at the edges. As a result, the laser beam is refracted toward the center, reinforcing the intensity perturbation and further driving the FI.

The FI has been extensively studied in the literature,9–12 and all these studies have focused on plasmas with non-relativistic temperature. The theoretical models used are based on hydrodynamic equations coupled with Maxwell's equations. It has been shown that in highly collisional plasmas close to thermodynamic equilibrium, the spatial growth rates of the FI are relatively low typically about 2cm1 as reported in Ref. 9. Taking into account non-local effects in the fluid model, the authors have highlighted significant spatial growth rate of the FI. Similar work has been presented in Refs. 10 and 11 in weakly collisional plasmas. In addition, the interaction of the FI with other parametric instabilities was also considered. In particular, the interplay between FI and Brillouin instability was studied in Ref. 12.

To our knowledge, this instability has not been studied quantitatively in hot plasmas with temperatures typically exceeding 5 keV. At these high temperatures, relativistic effects can no longer be neglected, and the FI must be revisited in the context of inertial confinement fusion (ICF). This issue is the subject of the present work.

As mentioned above, the Raman instability (see Refs. 1, 4, and 6) has been studied in very hot plasmas. The present work therefore extends the study of the parametric instabilities in such plasmas.

We used new relativistic and collisionless hydrodynamic model equations that incorporate purely kinetic effects. This model serves as the counterpart to a kinetic model based on the Vlasov equation. The dissipation of the hydrodynamic modes in plasmas is included through the thermal conductivity and the viscosity coefficient, which account for the well-known Landau damping.13,14 Additionally, we emphasize that new transport coefficients, namely, the convective heat flux and temperature anisotropy coefficients, are taken into account in this collisionless hydrodynamic model. These coefficients become negligible in the collisional approximation, and therefore they are not accounted for in the standard hydrodynamic equations.

Using these hydrodynamic model equations and Maxwell's equations, we investigate the FI in a new regime that is not only relativistic but also collisionless. Most previous models, whether relativistic or non-relativistic, are based on the continuity equation and the momentum and energy balance equations in the collisional regime. Additional assumptions on the pressure gradient term (choice of the adiabatic parameter, isothermal or adiabatic approximation, etc.) are used to make the system of equations self-consistent. These assumptions are valid only in the collisional regime. The present collisionless model which describes perturbed plasmas with respect to equilibrium is derived without any complementary assumptions.

This work is organized as follows. In Sec. II, we present the hydrodynamic model which includes both collisionless and relativistic effects. These equations are split into low-frequency equations and high-frequency equations including the wave equation of EM modes. Then, in Sec. III, we calculate the thermal and ponderomotive drives. Section IV is devoted to the dispersion relation of the FI and to the numerical analysis of the growth rate. Finally, we give in Sec. V a discussion and a summary of this work.

The FI of an EM wave arises from nonlinear coupling between the EM wave and the plasma. The main assumptions used in this hydrodynamic model are the consideration of the collisionless transport and relativistic temperature effects on plasma electrons. Ions, because of their strong inertia, are considered as non-relativistic.

The model used to describe the plasma is based on ion and electron hydrodynamic equations, while the EM wave as usual is described by the Maxwell's equations. It is assumed that the EM wave consists of a high-frequency (HF) electric field of constant amplitude of zero order, and a small perturbed electric field of first order. In complex notations, they are written, respectively, as EHF0=E0expiω0t+ik0·rex and δEHF=δE+expiω+t+ ik+·rex+δEexpiωt+ik·rex. The parameters defining these fields are the amplitudes E0, δE+, and δE, the frequencies ω0, ω+, and ω, and the wave vectors k0, k+, and k. The wave vector directions are specified as follows: k0=k0ez, k=kzez+  kyey. The unperturbed plasma is assumed to be homogeneous, defined by the electron density ne0, electron temperature Te0, ion density ni0, and ion temperature Ti0. In the presence of the EM wave, a coupling occurs between the EM wave and the plasma's eigenmodes. At lowest order, the perturbed electron and ion hydrodynamic quantities are δne,δTe,δVe and δni,δTi,δVi, corresponding, respectively, to the density, temperature, and fluid velocity for each species. All hydrodynamic quantities are assumed to be described by Fourier modes of the form, δT,δn,δVexpiωt+ik·r, where ω is the frequency and k is the wave vector of the hydrodynamic modes. The modes considered in this work are resonant modes which ensure efficient wave–plasma coupling. These modes satisfy the resonance conditions ω=ω ∓ ω0 and k=k ∓ k0.

The system of hydrodynamic equations describing the perturbed state of the plasma is divided into two groups: low-frequency equations and high-frequency equations. These two timescales are defined by ω1 for the low-frequency hydrodynamic modes and by ω01 for the high-frequency modes, with ω0ω. We now present each set of equations.

  • The relativistic and collisionless electron hydrodynamic equations that describe the perturbed state are the continuity equations and the momentum and energy balance equations,15 
    (1)
    (2)
    (3)

    where me is the electron mass, qe is the electron charge, δΠ¯¯e is the stress tensor, δFpe is the ponderomotive force, δqe is the heat flux, and δjHF·EHF is the inverse bremsstrahlung absorption rate averaged on the time scale T=2πω0, where jHF is the high-frequency current density. Moreover, we have used the parameters G=K3z0K2z0, Kn(z0) being the nth-kind modified Bessel function and H=231+z021G2+5z0G where here z0=mec2Te0. The nonlinear ponderomotive and inverse bremsstrahlung absorption terms are coupling terms between the EM wave and the plasma, and they are the driven terms for the FI.

    For the sake of clarity, we present the explicit expressions of the heat flux and the stress tensor in  Appendix A. The expressions of these transport quantities for electrons are
    (4)
    (5)

    The transport coefficients are the thermal conductivity KTe, the viscosity coefficient ηe, which both represent the dissipation terms, and αVe and αTe, respectively, the convective and temperature anisotropy coefficients. We found that the Onsager symmetry is fulfilled, i.e., αTe=αVe. For electrons, the explicit expressions of these four coefficients are given by Eqs. (A3)–(A6). We represent in Fig. 1 these transport coefficients as functions of the relativistic parameter z0, where we can see that these coefficients increase with increasing relativistic effects (or decreasing z0).

  • Similarly, the non-relativistic and collisionless ion hydrodynamic equations read
    (6)
    (7)
    (8)
    where mi and qi are, respectively, the ion masse and the ion charge. Due to the ion high inertia, the relativistic effects mic2Ti01, the ion ponderomotive force, and the absorption of EM energy by ions are negligible. The non-relativistic transport coefficients are calculated in Ref. 14, and they read
    (9)
    (10)

    where KTi=952π, ηi=252π, and αTi=αVi=25.

  • Poisson's equation relates the electrostatic field to the plasma density perturbation
    (11)

    where ε0 is the vacuum permittivity.

FIG. 1.

Thermal conductivity KTe (solid line), viscosity coefficient ηe (dashed line), and anisotropic temperature coefficient αTe (dotted line) as functions of the relativistic parameter z0=mec2Te0.

FIG. 1.

Thermal conductivity KTe (solid line), viscosity coefficient ηe (dashed line), and anisotropic temperature coefficient αTe (dotted line) as functions of the relativistic parameter z0=mec2Te0.

Close modal

  • The high-frequency hydrodynamic equations only concern electrons (low inertia), and they can be written as
    (12)
    (13)
    (14)

    where νei is the electron–ion collision frequency.15 

  • The last set of equation is the wave equation of EM modes. To determine this equation, we use the Maxwell–Faraday equation which connects the high-frequency electric δEHF to the high-frequency magnetic fields δBHF,
    (15)
    and the Maxwell–Ampère equation
    (16)
    where μ0=1/ε0c2 is the magnetic permeability of the vacuum, and
    (17)

    is the high-frequency current density.

First, we calculate the velocities δVeHF+, δVeHF, and VeHF0. Using Eqs. (12)–(14), we obtain
(18)
(19)
Now, we take the rotational of Eq. (15) and using Eq. (16), we get
(20)
Replacing relations (18) and (19) into Eq. (17) and rearranging terms, we deduce the wave equation for the perturbed EM waves,
(21)
where ωpe=ne0qe2ε0me is the electron plasma frequency. We assumed that νeiω0 and used the dispersion relation of free EM modes, ω02=k02c2+ωpe2G.

To proceed further, we need to calculate the driving terms for the FI in the relativistic approximation, specifically the absorption rate and the ponderomotive force.

To calculate the absorption rate, we use Eqs. (13) and (14), retaining only the relevant first-order terms,
(22)
which becomes
(23)
where
(24)
is the absorption rate with the electron–ion collision frequency15 
(25)

Zi is the ion charge number, and lnΛ is the Coulomb logarithm. We note that in the non-relativistic limit, z01, K3z011z0expz0, and collision frequency scales as νei1Te03/2.

The ponderomotive force is a low-frequency force that accounts for the elastic transfer of momentum from photons to electrons. It is derived from the product of two high-frequency terms averaged on the high-frequency scale. The first term is the convection term
(26)
and the second term is the magnetic force
(27)
where the magnetic fields B0HF and δBHF++δBHF can be calculated from the Maxwell–Faraday equation,
(28)
(29)
and
(30)

The ponderomotive force considered in this work is transverse, i.e., δFpe=δFpeey, meaning we are not concerned here with the longitudinal component, δFpe=δFpeez.

The two terms of Eq. (26) cancel each other because the high-frequency velocities are parallel to the Ox-axis, and as transverse modes, they do not depend on the spatial variable x. The first term in Eq. (27) is longitudinal (along ez) and thus does not contribute to the desired transverse ponderomotive force. Only the second term in Eq. (27) remains, which can be rewritten by retaining only the transverse components as follows:
(31)
After simplification, we obtain
(32)

The hydrodynamic equations (1)–(3) and (6)–(8) and the wave equation (21) together with the expressions of the absorption rate (23) and of the ponderomotive force (32) are the self-consistent basic equations of this work. In the following, it will be used to study the FI in relativistic plasma temperature.

The equations of the model established above are linear with respect to the perturbed hydrodynamic variables and the perturbed electrostatic and EM fields. As a first step, we focus on the ions. Using the continuity [Eq. (6)], the momentum [Eq. (7)], and Poisson's equations [Eq. (11)], and the closure relations (A1) and (A2), we deduce
(33)
where ωpi=ni0qi2ε0mi is the ion plasma frequency. Likewise, the ion energy equation can be recast as
(34)
From Eqs. (33) and (34), we deduce the following relationship between the perturbed ion and electron densities:
(35)
Similarly, by following the same stages, from the electron equations (1)–(3), we obtain a second relationship between the ion and electron densities
(36)
We have also only kept the dominant terms of these two source terms. Equations (35) and (36) are two linear and homogeneous equations with respect to the variables δne and δni. They admit a non-trivial solution if the determinant of their four coefficients is zero. The resulting equation is the desired dispersion relation of the FI. For clarity, its explicit expression is given in  Appendix B, where it appears as an eighth-degree polynomial in the frequency ω of the hydrodynamic modes. It corresponds to Eq. (B1) in  Appendix B that we rewrite here for convenience,

The coefficients Ci=08k,ne0,Te0 are expressed as functions of the plasma and EM wave parameters.

  • Steady-state filamentation instability.

In this subsection, we study the FI in the stationary limit: ω0. In this case, the hydrodynamic modes scale as δXexpik·r. The instability is analyzed as an amplification of the laser intensity in exp(Γz), where Γ=ikz is the real spatial growth rate. The dispersion relation for the FI can be easily deduced from the general dispersion relation (B1) by setting ω=0, yielding the following dispersion relation: C0=0, which explicitly gives from Eqs. (B10) and (B11) the following spatial growth rate:
(37)
where v0=eE0meω0 is the quiver velocity. Within the condition ZiTe/Ti1 used in this work, the condition k2λDi2k2λDe21 is thus fulfilled, where λDi,e are the ion and electron Debye lengths, and the expression of the spatial growth rate becomes
(38)

We should note that Eq. (37) can be useful for further applications, particularly in plasmas where the parameter ZiTeTi has low values close to 1. We have checked numerically that the two expressions (37) and (38) give very close numerical results. In both equations, the negative term k2 stands for the diffraction (a stabilizing term), while the two positive terms represent the source terms of the instability. They correspond, respectively, to the thermal and ponderomotive drive terms. They depend, respectively, on the relevant parameter νav02 and v02, which stand, respectively, for the inverse bremsstrahlung absorption and the ponderomotive force.

It is important to note that in the thermal drive, dissipation is accounted for only through the thermal conductivity KTe; the viscosity, which represents the second dissipative term, does not appear in Eq. (38). Moreover, additional transport coefficients, such as the electron convective coefficient αVe as well as all ion transport coefficients, do not contribute to the spatial growth rate of the stationary FI.

We have numerically calculated in Figs. 2 and 3, the spatial growth rate as a function of the wavenumber k for, respectively, the ponderomotive and thermal sources. The electron temperature and the corresponding laser intensity are

  • Te0=1keV and I=4.2×1013W/cm2 as in Ref. 9;

  • Te0=3keV and I=5×1014W/cm2 as in Ref. 12;

  • Te0=20keV and I=1016W/cm2 as in Ref. 1.

FIG. 2.

Spatial growth rate of the FI driven by the ponderomotive force as a function of the wavenumber. The plasma parameters used are the electron temperatures Te0=1keV (dashed line), Te0=3keV (dotted line), and Te0=20keV (solid line), the underdense electron density ne0=0.1nc, where nc is the electron critical density, the laser wavelength λL=1.06μm, and the ion charge number Zi=6.

FIG. 2.

Spatial growth rate of the FI driven by the ponderomotive force as a function of the wavenumber. The plasma parameters used are the electron temperatures Te0=1keV (dashed line), Te0=3keV (dotted line), and Te0=20keV (solid line), the underdense electron density ne0=0.1nc, where nc is the electron critical density, the laser wavelength λL=1.06μm, and the ion charge number Zi=6.

Close modal
FIG. 3.

Spatial growth rate of the FI driven by the inverse bremsstrahlung absorption as a function of the wavenumber. The other plasma parameters used are the same as in Fig. 2.

FIG. 3.

Spatial growth rate of the FI driven by the inverse bremsstrahlung absorption as a function of the wavenumber. The other plasma parameters used are the same as in Fig. 2.

Close modal

A relationship between temperature and laser intensity would allow expressing the growth rate only as a function of temperature Teo, i.e., the relativistic parameter zo. This would be particularly useful for interpreting the role of relativistic effects on the FI. However, to our knowledge, this relationship I0=fI0 requires developing a complete analytical hydrodynamic model that describes the laser–plasma interaction, which is beyond the scope of this work. We just note that a hydrodynamic model was derived by Fabbro et al.16 for non-relativistic temperature plasmas, leading to the scaling law I0Te03/2. In addition, we assume also as in Ref. 10 that Ti0Te04.

Figures 2 and 3 show that the growth rate exhibits a maximum for an optimal wavenumber kopt. For the thermal and ponderomotive FI, the spatial growth rate scales, respectively, as ΓthSthkk4 and ΓpSpk2k4. For both sources, the diffraction term k4 becomes dominant over the source terms in the high wavenumber range, whereas in the small wavenumber range, the two source terms dominate the diffraction term.

The spatial growth rate in Fig. 2 driven by the ponderomotive force is independent of the relativistic transport coefficients KTe, ηe, and αTe. These modes are therefore not subject to damping due to thermal conduction or viscosity. However, the inertia caused by the relativistic effects, which is contained into the parameter G2, will reduce the growth rate. This occurs as if the effective mass increases, thereby decreasing the oscillation velocity of electrons in the high-frequency EM field. We note that the parameter G is the counterpart of the Lorentz factor 1v2/c21/2 for fluid relativistic effects. Figure 4 illustrates its variation with respect to the relativistic parameter z0. In the non-relativistic limit, we find G=1, and within the temperature range considered in this work, G increases slowly as z0 decreases. Finally, the relativistic quiver velocity v0/G increases with increasing laser intensity, and it is reduced by relativistic inertial effects.

FIG. 4.

Relativistic inertia factor G due to random motion as a function of the relativistic factor z0.

FIG. 4.

Relativistic inertia factor G due to random motion as a function of the relativistic factor z0.

Close modal

In Fig. 3, where only the thermal drive is considered, the spatial growth rate decreases as the temperature increases. This behavior can be attributed to the increase in the dissipative transport coefficients KTe and αTe (see Fig. 1), the rise in the inertia parameter G, and the decrease in the absorption rate νa [Eq. (24)].

To better understand the behavior of the growth rates with respect to temperature, we derive their scaling laws. Using Eq. (38) and the scaling law of Ref. 15, I0v02Te3/2, we can determine the maximum growth rate of the ponderomotive and thermal source terms with respect to temperature, i.e., with respect to relativistic effects. Although this scaling law is valid for non-relativistic plasmas, it is assumed that it remains valid for moderately relativistic plasmas since the electron mass inertia factor G does not vary significantly in the temperature range used in this work. The scaling law of the growth rates are ΓthSthkk4 and ΓpSpk2k4, where Sth=12Gvte2c2νav02ω021αTevteKTe and Sp=12vte2c2ωpe2v02G2. We deduce readily kopt,th=Sth41/3 and kopt,p=Sp21/2 which give, respectively, the maximum growth rates ΓthSth2/3 and ΓpSp. A simple analysis of these growth rates shows that Γth1Te3/2 and ΓpTe1/2, where we have used νa1/T0e3/2. Thus, the behavior of the spatial growth rate with respect to the temperature is opposite for the thermal and ponderomotive drives. As a conclusion, from Figs. 2 and 3, the ponderomotive FI dominates the thermal FI and even more significantly when the plasma temperature increases.

Moreover, in the ultrarelativistic limit, z00, G, αT1, and vteKTeπ6c; thus, the thermal and ponderomotive sources should decrease drastically, and we expect that in strong relativistic regime the FI should be stabilized.

We give in Fig. 5 the maximum spatial growth rate and the corresponding optimum wavenumber which maximizes this growth rate, as a function of the position defined by the electron density in the underdense plasma. We present the results for the temperatures Te0=1keV and Te0=20keV and using only the dominant ponderomotive source. The plasma parameters are given in Fig. 2.

FIG. 5.

Maximum spatial growth rate (solid line) and optimum wavenumber (dashed line) as functions of the electron density normalized by the critical density, for non-relativistic [Fig. 4(a)] and relativistic [Fig. 4(b)] electron temperatures. The other plasma parameters are the same as in Fig. 2.

FIG. 5.

Maximum spatial growth rate (solid line) and optimum wavenumber (dashed line) as functions of the electron density normalized by the critical density, for non-relativistic [Fig. 4(a)] and relativistic [Fig. 4(b)] electron temperatures. The other plasma parameters are the same as in Fig. 2.

Close modal

The results in Fig. 5 illustrate the dependence of the maximum growth rate and the corresponding optimum wavenumber on electron density in homogeneous plasmas. This analysis allows for a qualitative study of what occurs in an inhomogeneous plasma. If we assume that the plasma is inhomogeneous with a characteristic spatial scale length L, the local approximation L1 is well verified, where k is the FI wavenumber. Typically, L is of the order of a hundred micrometers, and k is greater than 105m1. Therefore, the dispersion relation of FI, derived for a homogeneous plasma, remains valid in inhomogeneous plasmas where the local approximation is verified. Furthermore, as seen in Fig. 5, the growth rate increases significantly with electron density. For instance, for Te0=20keV, the growth rate is approximately Γ3.8×105m1 at the layer nenc=0.1 and about Γ9.4×105m1 at the layer nenc=0.6. This is due to the decrease in the wavenumber k0 of the EM wave via the dispersion relation of EM waves in plasmas. Additionally, we observe that the optimum wavenumber increases with electron density. We have also checked that these wavenumbers are consistent with the collisionless approximation kλei>1 used in the fluid model, since for nenc=0.6 we found λei1.5×104m and about k105m1.

  • Non-stationary filamentation instability.

In this section, we focus on the non-stationary modes of the FI. To do these, we must numerically solve Eq. (B1) in  Appendix B. This equation is an eighth-order polynomial with real coefficients for the even powers of ω and purely imaginary coefficients for the odd powers. By changing the variable ω by Ω=iω, the coefficients of the polynomial are all real. We used standard numerical methods to calculate the roots of this polynomial,17 assuming a real wavenumber k and a complex frequency ω=ωr+iωi, where ωr is the frequency and ωi, if positive, would be the temporal growth rate. The results obtained for the complex frequency ω are represented in Figs. 6 and 7 for, respectively, the frequency ωrk and the growth rate ωik for two plasma temperatures Te0=3keV and Te0=20keV. These results account for both the thermal and ponderomotive sources of instability.

FIG. 6.

Frequency as a function of the wavenumber of the FI driven by the thermal and the ponderomotive sources for electron temperatures Te0=3keV (dashed line) and Te0=20keV (solid line). The other plasma parameters used are the same as in Fig. 2.

FIG. 6.

Frequency as a function of the wavenumber of the FI driven by the thermal and the ponderomotive sources for electron temperatures Te0=3keV (dashed line) and Te0=20keV (solid line). The other plasma parameters used are the same as in Fig. 2.

Close modal
FIG. 7.

Temporal growth rate as a function of the wavenumber of the FI driven by the thermal and the ponderomotive sources for electron temperatures Te0=3keV (dashed line) and Te0=20keV (solid line). The other plasma parameters used are the same as in Fig. 2.

FIG. 7.

Temporal growth rate as a function of the wavenumber of the FI driven by the thermal and the ponderomotive sources for electron temperatures Te0=3keV (dashed line) and Te0=20keV (solid line). The other plasma parameters used are the same as in Fig. 2.

Close modal

For the real dispersion relation ωk (see Fig. 6), the low-frequency modes exhibit dispersive properties, and they move at different velocities leading to the deformation of the wave envelope. The frequency magnitude is around a few 1012s1 which is significantly lower than the laser frequency ω01.8×1015s1, thus justifying the low-frequency approximation used in the hydrodynamic model equations. On the other hand, for wavenumbers typically greater than 7×105m1, we can see that the frequency is about ωrk>7×105m1=9.75×1012s1. This frequency corresponds to the ion plasma frequency ωpi. We have checked numerically by varying the ion mass that this property is always fulfilled. In Fig. 7, examining the temporal growth rate within this frequency range ωik>7×105m1, we note strong damping of this mode. Under these conditions, electrons and ions remain largely unaffected by the EM wave, and they oscillate at the low plasma frequency ωpi.

We now analyze the behavior of the damping rate with respect to the wavenumbers, as shown in Fig. 7. The most unstable modes are located in the wavenumber spectrum about k6.2×105m1, and the growth rate of these unstable modes increases with temperature. For Te0=20keV, the maximum growth rate is around ωi2.9×1011s1, and it is about ωi1.3×1011s1 for Te0=3keV. Thus, the increase in the temperature leads to moderate increase in the growth rates. On the other hand, the relativistic effects significantly increase the spectral width of unstable modes. For Te0=20keV, it can be deduced from Fig. 7 that in the wavenumber range (5.3×105 and 6.6×105m1), the growth rate is greater than ωi1011s1. Under typical ICF conditions where the pulse duration is on the order of a nanosecond, we should obtain a high amplification factor of the filament intensity, although it will be limited by nonlinear effects.18–20 Thus, as for the stationary modes, the low-frequency modes are significantly amplified, and this leads to a strong filamentation of the laser pulse via mainly the ponderomotive drive. This should play a detrimental role in the context of the inertial confinement fusion.

  • The FI is driven throughout the underdense plasma and coexists with other parametric instabilities, including stimulated Brillouin scattering (SBS) and stimulated Raman scattering (SRS). We propose to compare the magnitude of the growth rates of these two instabilities with that of the FI. For this purpose, we derive the dispersion relations of SBS and SRS instabilities in plasmas at relativistic temperatures. For clarity, these derivations are presented in  Appendix C [see Eqs. (C3) and (C4)]. Comparison of these three instabilities shows the following:

  • In inertial confined fusion, both the FI and SBS instabilities can occur throughout the entire underdense plasma ne<nc, where nc is the critical density, whereas the SRS is driven in region under the quarter of the critical density, i.e., ne<nc4.

  • The frequency spectra of the hydrodynamic modes driven by the FI and the SBS lie in the low-frequency range. In contrast, the SRS spectrum is located in the high-frequency range, and it therefore differs from that of the FI.

  • The FI wave vectors of the unstable hydrodynamic modes differ in both magnitude and direction from those of the SBS and SRS. Indeed, the direction of the FI wave vector k is transverse to the EM pump wave vector k0, whereas k takes arbitrary directions for the SBS and SRS instabilities. The magnitude of the SBS and SRS wave numbers k could be of the order of k0, while that of the FI is clearly less important, i.e., kk0.

  • As shown in Fig. 6, for plasma temperatures of 3 and 20keV, the maximum FI growth rates are, respectively, about 1011s1 and 3×1011s1. Under the same temperature and density conditions, from Eqs. (C3) and (C4), the SBS and SRS growth rates take the values ωiB3keV=3.3×1011s1,ωiB20keV=4.1×1011s1, ωiR3keV=4.8×1011s1, and ωiR20keV=5.2×1011s1. We used the maximum growth rates corresponding to the backward-SBS with k=2k0 and to the backward-SRS with k=k012ωpeω0.21 We can note that in the underdense plasma, the maximum SBS and SRS growth rates are higher than that of the FI.

  • It is well known that spatial gradients in density and temperature can significantly affect the behavior of plasma instabilities. In this work, however, our analysis is restricted to homogeneous plasmas, as our primary objective is to investigate FI in relativistic plasmas. Thus, we ignore this effect on the growth rate of the SBS and SRS and consider that it is beyond the scope of this work. The FI is not sensitive to plasma inhomogeneity, and thus our results remain valid in inhomogeneous plasmas.

For the steady-state case, we consider zero-frequency hydrodynamic quantities, which are the density, temperature, velocity perturbations, and also the corresponding modulations in intensity. Consequently, the density fluctuations are purely growing modes meaning that ion acoustic waves are not involved in this coupling. Instead, these fluctuations correspond to local variations in plasma density driven by thermal and ponderomotive sources. The increase in density enhances the filament intensity via the refractive index, which in turn, by feedback, further increases the density. This amplification of the filament occurs more and more as the EM pump wave propagates in the plasma.

For the non-stationary case, the results are exact since no assumption is made about the real part of the frequency, which is calculated on the same footing as its imaginary part. We then describe the coupling between the EM wave with the ion wave by numerically solving the dispersion relation for the parameters fixed in Figs. 6 and 7. This is a spatiotemporal problem that requires an analysis on the ω,k Fourier-space to determine the nature of the amplification (absolute, convective instability, etc.).

In this work, a model of linear hydrodynamic equations to describe plasmas for temperature relativistic regime and non-relativistic electron oscillation velocities in the laser electric field is proposed. Electrons are treated within the relativistic approximation, allowing for high temperatures above 10keV, while ions, due to their greater inertia, are described in the non-relativistic approximation. The low-frequency closure relations used are exact and are calculated for collisionless plasmas. In general, collisions contribute significantly to the damping of plasma hydrodynamic fluctuations which could suppress the development of plasma instabilities. Consequently, instabilities are more readily driven in the collisionless wavenumber range, defined by the Knudsen number kλei>1, where λei is the electron mean-free-path. This condition is consistently satisfied in the numerical applications of this study. Moreover, this hydrodynamic model can be used to investigate the dispersion relations of plasma modes as well as other instabilities.

We then apply this model to study the FI of a laser beam under typical conditions of laser–plasma interaction in the context of ICF. The model takes into account EM waves propagating through the plasma, with amplitude modulations transverse to the wave's propagation direction. These EM waves are described by Maxwell's equations which are coupled with the hydrodynamic equations to make a self-consistent system. The absorption of the wave energy via inverse bremsstrahlung and the ponderomotive force are both considered, and they represent the two driven terms of the FI.

In a first step, the FI is studied in the steady-state approximation, and the dispersion relation is derived by considering both sources of the instability. The analysis reveals that highly unstable stationary modes are excited in the plasma. It also emphasized the dominance of the ponderomotive source over the thermal source to drive the FI. The calculated growth rates show that strongly unstable modes can be driven near the critical density.

In a second step, the FI is studied within the framework of the non-stationary approximation. The dispersion relation is in the form of an eighth-degree polynomial with respect to the complex frequency ω=ωr+iωi. The numerical solution of the dispersion relation revealed highly unstable modes. It was shown that the growth rates depend moderately on the electron temperature. For the two temperatures used in this work, Te0=3keV and Te0=20keV, the growth rate is greater than 1011s1.

In conclusion, for hot plasmas with a temperature exceeding 10keV, the FI can play a detrimental role in the context of ICF. Indeed, the strong modulation of the laser intensity should increase the lack of symmetry of the compression wave propagating toward the center of the target. It results thus a non-efficient coupling between the laser energy deposition and the compression wave.

This work has been carried out within the framework of the PRFU 2023–2026 (Projet de Recherche et de Formation Universitaire) and under Grant Agreement No. B00L02UN160420230002.

The authors have no conflicts to disclose.

H. Benmakrelouf: Conceptualization (equal); Investigation (equal); Methodology (equal); Writing – original draft (equal). K. Bendib-Kalache: Conceptualization (equal); Investigation (equal); Methodology (equal); Writing – original draft (equal). A. Bendib: Conceptualization (equal); Investigation (equal); Methodology (equal); Writing – original draft (equal).

The data that support the findings of this study are available within the article.

To calculate the collisionless closure relations, we must solve the electron Vlasov equation which describes the electron distribution function in the phase space r,p, where r and p are, respectively, the electron vector position and the momentum. This was done by the authors first, in non-relativistic plasmas,14 then in relativistic plasmas in temperature.22 We are interested in this work by the closure relations in relativistic plasmas. In Ref. 22, they were presented in the form of integrals not convenient to use. We propose in this  Appendix A simpler and more explicit writing of these integrals by explicitly calculating some of them and expressing the others with respect to modified Bessel functions. We give below the new expressions of the closure relations as a function of electron relativistic parameter z0=mec2Te0. The collisionless transport quantities are written
(A1)
(A2)
where δqxe is the heat flux, and δΠxxe is the stress tensor. These quantities are defined by the thermal conductivity coefficient
(A3)
the convective heat flux coefficient
(A4)
the viscosity coefficient
(A5)
and the temperature anisotropy coefficient
(A6)
where
with =K3z0K2z0, where K2z0 and K3z0 are the modified Bessel function of kind 2 and 3, respectively. The coefficients (A3)–(A6) constitute reliable closure relations for low-frequency hydrodynamic equations.
Solving the determinant of Eqs. (35) and (36), we obtain the following expression of the dispersion relation:
(B1)
where
(B2)
(B3)
(B4)
(B5)
(B6)
(B7)
(B8)
(B9)
(B10)
where
(B11)
is the inverse bremsstrahlung source term.

Both SBS and SRS instabilities are three-waves resonant processes. They correspond to the decay of the EM pump wave ω0,k0 into a scattered EM wave ω1,k1 and an ion acoustic wave ω,k in the case of SBS and an electron plasma wave ω,k in the case of SRS. The associated parametric resonance conditions are given by ω0=ω1+ω and k0=k1+k.

1. SBS instability
From Eqs. (35) and (36), keeping only the Stokes resonant mode, the dispersion relation of the SBS reads
(C1)
where νtz0=νeiz0ωpe22Gω02. The Landau damping rate of the ion acoustic wave is23 
where μ=miZme, τ=ZTeTi, and the coefficients A and B are numerical fits given by

The relativistic effects are included in the parameter G, νtz0, and νiaz0. The associated parametric resonance conditions for the wavenumbers are presented in the following scheme:

Since ωω0,ω1, it results that ω0ω1, and thus k0k1. In particular, for the pump EM mode and the scattered EM mode, the following relations ω0ω1 and k0k1 hold, and thus k=2k0sinθ2. For typical cases (a) a forward SBS with θ=0, (b) a backscattering SBS for θ=π, and (c) a side scattering SBS for θ=π2, we obtain, respectively, k0, k=2k0, and k=2k0.

2. SRS instability
The dispersion relation of SRS is similar to that of SBS. We change in Eq. (C1) the ion acoustic wave by the electron plasma wave, and we obtain
(C2)
where
and ω±=ω±ω0 and k±=k ± k0.
The damping rate of electron plasma waves is3 
where
and K=z0kvteωpe. The SRS wavenumbers resonance scheme is the same as the SBS represented above.
Solving Eq. (C1), we obtain, for the SBS, the growth rate
(C3)
Similarly for the SRS, the growth rate is easily deduced from Eq. (C2) (Ref. 21),
(C4)
1.
A. B.
Langdon
and
D. E.
Hinkel
,
Phys. Rev. Lett.
89
,
015003
(
2002
).
2.
J.
Bergman
and
B.
Eliasson
,
Phys. Plasmas
8
,
1482
(
2001
).
3.
B.
Touil
,
A.
Bendib
, and
K.
Bendib-Kalache
,
Phys. Plasmas
24
,
022111
(
2017
).
4.
A.
Bers
,
I. P.
Shkarosfsky
, and
M.
Shoucri
,
Phys. Plasmas
16
,
022104
(
2009
).
5.
G.
Li
,
W.
Mori
, and
C.
Ren
,
Phys. Rev. Lett.
110
,
155002
(
2013
).
6.
Y.
Zhao
,
J.
Zheng
,
M.
Chen
,
L.-L.
Yu
,
S.-M.
Weng
,
C.
Ren
,
C.-S.
Liu
, and
Z.-M.
Sheng
,
Phys. Plasmas
21
,
112114
(
2014
).
7.
R. K.
Kirkwood
,
J. D.
Moody
,
J.
Kline
,
E.
Dewald
,
S.
Glenzer
,
L.
Divol
,
P.
Michel
,
D.
Hinkel
,
R.
Berger
,
E.
Williams
,
J.
Milovich
,
L.
Yin
,
H.
Rose
,
B.
MacGowan
,
O.
Landen
,
M.
Rosen
, and
J.
Lindl
,
Plasma Phys. Controlled Fusion
55
,
103001
(
2013
).
8.
R. S.
Craxton
,
K. S.
Anderson
,
T. R.
Boehly
,
V. N.
Goncharov
,
D. R.
Harding
,
J. P.
Knauer
,
R. L.
McCrory
,
P. W.
McKenty
,
D. D.
Meyerhofer
,
J. F.
Myatt
,
A. J.
Schmitt
,
J. D.
Sethian
,
R. W.
Short
,
S.
Skupsky
,
W.
Theobald
,
W. L.
Kruer
,
K.
Tanaka
,
R.
Betti
,
T. J. B.
Collins
,
J. A.
Delettrez
,
S. X.
Hu
,
J. A.
Marozas
,
A. V.
Maximov
,
D. T.
Michel
,
P. B.
Radha
,
S. P.
Regan
,
T. C.
Sangster
,
W.
Seka
,
A. A.
Solodov
,
J. M.
Soures
,
C.
Stoeckl
, and
J. D.
Zuegel
,
Phys. Plasmas
22
,
110501
(
2015
).
9.
10.
T. B.
Kaiser
,
B. I.
Cohen
,
R. L.
Berger
,
B. F.
Lasinski
,
A. B.
Langdon
, and
E. A.
Williams
,
Phys. Plasmas
1
,
1287
(
1994
).
11.
A.
Bendib
,
F.
Bouzid
,
K.
Bendib
, and
G.
Matthieussent
,
Phys. Plasmas
6
,
4008
(
1999
).
12.
V.
Yu. Bychenkov
,
W.
Rozmus
,
A. V.
Brantov
, and
V. T.
Tikhonchuk
,
Phys. Plasmas
7
,
1511
(
2000
).
13.
G. W.
Hammett
and
F. W.
Perkins
,
Phys. Rev. Lett.
64
,
3019
(
1990
).
14.
K.
Bendib
and
A.
Bendib
,
Phys. Plasmas
6
,
1500
(
1999
).
15.
D. I.
Dzhavakhishvili
and
N. L.
Tsintsadze
,
Zh. Eksp. Teor. Fiz.
64
,
1314
(
1973
)
D. I.
Dzhavakhishvili
and
N. L.
Tsintsadze
[
Sov. Phys. JETP
37
,
666
(
1973
)];
O. J.
Pike
and
S. J.
Rose
,
Phys. Rev. E
93
,
053208
(
2016
).
[PubMed]
16.
R.
Fabbro
,
C. E.
Max
, and
E.
Fabre
,
Phys. Fluids
28
,
1463
(
1985
).
17.
H.
William
,
S. A.
Teukolsky
,
W. T.
Vetterling
, and
B. P.
Flannery
,
Numerical Recipes
(
Cambridge University Press
,
Cambridge, UK
,
2007
), Chap. 9.
18.
F.
Califano
,
R.
Prandi
,
F.
Pegoraro
, and
S. V.
Bulanov
,
Phys. Rev. E
58
,
7837
(
1998
).
19.
E.
Sobacchi
,
Y.
Lyubarsky
,
A. M.
Beloborodov
,
L.
Sironi
, and
M.
Iwamoto
,
Astrophys. J. Lett.
943
,
L21
(
2023
).
20.
S.
Sharma
,
N.
Kumar
,
S.
Hussain
, and
R. P.
Sharma
,
Laser Part. Beams
35
,
10
(
2016
).
21.
W. L.
Kruer
,
The Physics of Laser Plasma Interaction
(
Addison-Wesley
,
New York
,
1988
).
22.
K.
Bendib
,
A.
Bendib
, and
G.
Matthieussent
,
Phys. Plasmas
11
,
4254
(
2004
).
23.
B.
Touil
,
A.
Bendib
,
K.
Bendib-Kalache
, and
C.
Deutsch
,
Laser Part. Beams
34
,
178
(
2016
).