ArticlePDF Available

Ablative Rayleigh-Taylor instability with strong temperature dependence of thermal conductivity

Authors:

Abstract and Figures

An asymptotic analysis of Rayleigh–Taylor unstable ablation fronts encountered in inertial confinement fusion is performed in the case of a strong temperature dependence of the thermal conductivity. At leading order the nonlinear analysis leads to a free boundary problem which is an extension of the classical Rayleigh–Taylor instability with unity Atwood number and an additional potential flow of negligible density expelled perpendicular to the front. The nonlinear evolution of the front is analysed in two-dimensional geometry by a boundary integral method. The shape of the front develops a curvature singularity within a finite time, as for the Birkhoff–Rott equation for the Kelvin–Helmholtz instability.
Content may be subject to copyright.
J. Fluid Mech. (2007), vol. 579, pp. 481–492. c
2007 Cambridge University Press
doi:10.1017/S0022112007005599 Printed in the United Kingdom
481
Ablative Rayleigh–Taylor instability with
strong temperature dependence of the
thermal conductivity
C. ALMARCHA
1,P.CLAVIN
1, L. DUCHEMIN
1AND J. SANZ
2
1Institut de Recherche sur les Ph´
enom`
enes Hors Equilibre, Universit ´
es d’Aix Marseille et CNRS, 49 rue
Joliot Curie, BP 146, 13384 Marseille cedex 13, France
2ETSI Aeronauticos, Universitad Politecnica de Madrid, Madrid 28040, Spain
(Received 12 September 2006 and in revised form 21 December 2006)
An asymptotic analysis of Rayleigh–Taylor unstable ablation fronts encountered
in inertial confinement fusion is performed in the case of a strong temperature
dependence of the thermal conductivity. At leading order the nonlinear analysis leads
to a free boundary problem which is an extension of the classical Rayleigh–Taylor
instability with unity Atwood number and an additional potential flow of negligible
density expelled perpendicular to the front. The nonlinear evolution of the front is
analysed in two-dimensional geometry by a boundary integral method. The shape of
the front develops a curvature singularity within a finite time, as for the Birkhoff–Rott
equation for the Kelvin–Helmholtz instability.
1. Introduction
In inertial confinement fusion (ICF) the high fuel density which is required for
nuclear burning can be obtained by imploding spherical shells by high-power laser
radiation (direct drive). Irradiation leads to surface ablation and drives the capsule
implosion. A layer of finite thickness (called the total thickness in the following)
is formed between the ablation front in the cold (dense) material and the critical
surface at low density and high temperature where the laser energy is deposited.
The main mechanism of energy transfer between the laser deposition region and the
cold material is thermal conduction. Owing to the temperature dependence of heat
conductivity, the typical length scale of temperature variation increases strongly across
the conduction region from dain the cold material to dcat the critical surface (hot
material), dadc. Such a scale separation is responsible for the existence of a thin
ablation front with a large density jump separating the cold and dense material from
the hot conduction region, see figure 1. Accelerating ablation fronts are Rayleigh–
Taylor (RT) unstable while the short-wavelength disturbances are stabilized by
ablation, see Bodner (1974). The self-consistent analyses of linear stability performed
by Sanz (1994, 1996), Bychkov, Goldberg & Liberman (1994), Betti et al. (1995), and
Goncharov et al. (1996) are too complicated to be extended into the nonlinear regime
without further approximations. Nonlinear analyses have been performed by Sanz
et al. (2002) and Clavin & Almarcha (2005) within the framework of a semi-empirical
discontinuous model, called the sharp boundary model (SBM), consisting of two
fluids of constant density separated by a front of zero thickness, see Betti, McCrory &
Verdon (1993), Piriz, Sanz & Iba
nez (1997), Piriz (2001) and Clavin & Masse (2004).
482 C. Almarcha, P. Clavin, L. Duchemin and J. Sanz
1.5 0.5 0
0
0.2
0.4
0.6
0.8
1.0
1.2
ρ/ρa
ν = 0.5
1
2.5
4
ν = 0.5
1
2.5
4
νx/da
0
0.2
0.4
0.6
0.8
1.0
(TTa)/(TcTa)
1.0 1.5 0.5 0
νx/da
1.0
Figure 1. Density and temperature profiles for a Spitzer-type model with increasing νand
for ΘcTc/Ta= 50.
The jump in density (across the ablation front) is a free parameter in the SBM,
and an ad hoc assumption is required to close the problem. However, in the limit
of an infinitely large density jump, the SBM leads to a simple formulation of the
nonlinear problem that can be written in conveniently reduced variables as
Ψ=0,Ψ
+=0,lim
ζ→−∞ Ψ=0,lim
ζ+Ψ+=ζ, (1.1ad)
n·∇Ψ|ζ=ζf=Uf
n
+|ζ=ζf=0,∂Ψ
∂τ +|Ψ|2
2+|Ψ+|2
2ζ=ζf
=ζf,(1.2ac)
where τand ζ=ζex+ηeyare reduced time and spatial coordinates, Uf
nis the
normal velocity of the ablation front ζf(t), and Ψ(ζ)andΨ+(ζ) are the velocity
potentials of the cold and hot flow respectively. Except for the presence of Ψ+due to
the hot flow expelled from the ablation front, the problem is like the RT problem with
unity Atwood number. The uncontrolled approximations of the semi-empirical SBM
limit the physical validity of the result, all the more so since the unknown density
jump is used in the scaling of the reduced variables in (1.1) and (1.2), see Clavin &
Almarcha (2005).
The objective of the present paper is towfold. First, by using an asymptotic analysis
extending the linear study of Sanz, Masse & Clavin (2006) to the fully nonlinear
regime, it is shown that the same equations as (1.1) and (1.2) are obtained from
the basic equations for the conservation of mass, momentum and energy in a limit
involving a strong temperature dependence of thermal conductivity, see (4.1) and (4.2).
A major difference with the SBM is that the reduced variables appearing in (1.1)
and (1.2) depend on the physical parameters defining the problem, namely the Froude
number and the power index of thermal conductivity. Secondly, we obtain solutions
in two-dimensional geometry of (1.1) and (1.2) by a boundary integral method in the
spirit of the pioneering work of Baker, Meiron & Orszag (1980), as done by Duchemin,
Josserand & Clavin (2005) in revisiting the RT instability. It is then shown that the
front develops a curvature singularity within a finite time. The singularity is similar to
the one observed in the solutions of the Birkhof–Rott equation describing the Kelvin–
Helmhotz instability. The problem is formulated in §2. The orders of magnitude are
presented in §3. The asymptotic analysis leading to (1.1) and (1.2) is developed in §4.
The nonlinear solution is presented in §5.
Ablative Rayleigh–Taylor instability 483
2. Formulation
In the limit of a low Mach number and for a perfect gas, ρT =ρaTa, with a constant
specific heat Cp, the fluid equations are
tρ+∇·(ρu)=0,(t+u·∇)u=ρ1p+gex,∇·[u(λ/CpρaTa)T]=0.(2.1ac)
Here, ρand Tare the density and temperature respectively, pis the pressure, u=
uex+weyis the flow velocity, xis the longitudinal coordinate and yis the transverse
coordinates, exand eyare the unit orthogonal vectors, expointing toward the hot
material, λis the heat conductivity, and gis the constant acceleration in the frame of
the capsule. The upstream boundary condition are
x→−:ρ=ρa,T=Ta,u=Vaex,p=¯
p, (2.2)
where Vais the ablation velocity and pthe unperturbed pressure, d(p+ρu2)/dx=ρg,
the overline denoting unperturbed quantities. The laser intensity I, actually an energy
flux, is absorbed at the critical density ρc, much lower than the initial density,
ρcρa.InICF,λ(T)=λaΘν,ΘT/T
adenoting the reduced temperature. The
power index is ν=5/2 for electron heat conduction and ν=13/2 for radiative heat
conduction of a fully ionized plasma. Introducing the conduction length at the cold
side, daλaaVa, three non-dimensional parameters characterize the problem: the
inverse Froude number Fr1gda/V2
a, the temperature ratio ΘcTc/Ta=ρac1
and the power index ν. The conduction length at the critical surface is dcdaΘν
cand
dcis the total thickness of the thermal wave. In conditions of ICF implosion, Θc1
and the acceleration is such that 1ν1
cFr1 < 1. Therefore, the wavelengths of
the most unstable disturbance and of the marginal mode (zero linear growth rate)
are of the same order of magnitude 2π/kmsuch that ν/dckm1/da. The thermal
perturbations cannot reach the critical surface which remains planar. The downstream
boundary conditions then reduce to
x+:TT=0,uand ppbounded.(2.3)
Equations (2.1) to (2.3) constitute the basic framework for studying the ablation
front during the implosion in ICF, see Atzeni & Meyer-Ter-Vehn (2004). The unper-
turbed temperature profile upstream of the critical surface is a solution to the first-
order equation µCp(TTa)λdT/dx=0 where µdenotes the unperturbed ablation
rate, µρaVa=I/Cp(TcTa). At the cold side, the profile of temperature (or density)
presents a sharp bend of width of order da, representative of the ablation front, see
figure 1. Choosing the origin of the x-axis at the bend, the temperature profile satisfies
a power law in the hot region where terms of order 1/Θare negligible,
ΘT(x)/Ta(νx/da)1, θ (x)Θν(x)x/da.(2.4)
This outer solution is singular at the origin where the ablation front is located.
Focusing attention on intermediate wavelengths, ν/dck1/da, equations (2.1) to
(2.3) describe a free boundary problem with two external regions: a constant-density
region of dense and cold medium (subscript ), ρ=ρa= 1, and a hot conduction
region (subscript +), ΘT/T
a=ρa+1, separated by a thin ablation front located
at x=α(y,t) (to simplify notation the Cartesian equation of the front is written for a
single-valued surface, but the method applies for a multi-valued front). Density and
temperature vary in the hot region.
Let us split the flow within the hot side into two parts, u+=u(p)
++u(r)
+,whereu(p)
+
VadaΘθis associated with the thermal flux ΘνΘ=Θθ,θΘν. According to
484 C. Almarcha, P. Clavin, L. Duchemin and J. Sanz
(2.1), ∇·u(r)
+= 0, and the thermal equation yields
t+u(r)
+·∇
Θ1+Vadaθ =0Θν, u+=u(p)
++u(r)
+,u(p)
+VadaΘθ.
(2.5)
In the limit kda0, equations (2.1) lead to jumps in density, temperature Θf+,mass
flux and pressure across the ablation front while conservation of energy gives the
mass flux in terms of the local gradient of temperature,
ρu·nuf
nx=α+
x=α=0,µ2
fρ+px=α+
x=α=0,u(r)|x=α+=u|x=α,(2.6ac)
µfaVa=(11f+)dan·∇θ|x=α+dan·∇θ|x=α+,(2.7a, b)
where x=α+and x=αdenote the hot and cold side respectively, nis the unit
vector perpendicular to the front (pointing in the direction of the hot side), uf
n
[1 + (∂α/∂y)2]1/2∂α/∂t is the normal velocity of the front in the frame of the
unperturbed front, and µfis the ablation rate (mass flux across the ablation front,
µfρ(u·nunf )|x=α). Equation (2.6c) comes from the conservation of the tangential
component of the flow velocity together with equation (2.1c) and the definition of
u(p)|x=α+.
3. Orders of magnitude
We consider intermediate accelerations, 1ν1
cFr11 in two-dimensional
geometry. An asymptotic analysis is then performed using the small parameter
ε(Fr1)1/(2(ν1)). At leading order in the distinguished limit ε0, Θc→∞,
εΘ1/2
c→∞, consistent with the real conditions, the linear growth rate σof a perturba-
tion of wavelength 2π/k is
σd
a/Va=(kda)Fr1(kda)21q
oν1,S=Kq
oK21(3.1a, b)
where S>0, K>0andq
o=d
2˜
ϕo/dξ2|ξ=0 are quantities of order unity,
Kkda/νε2ν=O(1),S(σd
a/Va)/νε2ν1=O(1),(3.2a, b)
with ˜
ϕo(ξ) the solution to the fifth-order system describing the hot conduction region
(ξ>0),
d3˜
ϕo
dξ3d˜
ϕo
dξ˜
ϕo
νξ =d˜
θo/dξ
νξ11,d2˜
θo
dξ2(˜
θo1) = ˜
ϕo
νξ1+1(3.3a, b)
and satisfying to the five boundary conditions
ξ=0:˜
ϕo=0,d˜
ϕo/dξ=0,˜
θo=0; ξ+:d
˜
ϕo/dξ=0,˜
θo=1,(3.4)
where ξk(xα), see Sanz et al. (2006). Equation (3.3b) is the heat transfer equation
(2.5), and ˜
θois related to the perturbation δθ,θ=θ(xα)+δθ, while ˜
ϕorepresents the
stream function of the solenoidal flow u(r)
+. More precisely, for a perturbed front of the
form α/da=˜
αeiky+σt,˜
θoand ˜
ϕoare defined by δθo=˜
θo˜
αeiky+σt and ϕo=˜
ϕo˜
αeiky+σt,
the subscript odenoting the leading order. Introducing δu(r)
+,u(r)
+=Va+δu(r)
+,the
components of (kda)1+1ν1δu(r)
+/Vaare (ϕo,idϕo/dξ). Pressure and vorticity
×δu(r)
+are represented by d2˜
ϕo/dξ2d(ξ1/n ˜
θo)/dξand ˜
ϕod2˜
ϕo/dξ2respectively.
Ablative Rayleigh–Taylor instability 485
Let us now consider the orders of magnitude. According to (3.3) and (3.4), ˜
θois of
order unity, δθo(ξ)=O(α/da). The perturbed ablation rate δµfis evaluated from its
definition (2.7) with δθ =O(kδθ), δµf/¯
µ=O(). The order of magnitude δu(r)
+/Va=
O(ε2˜
ϕo) is obtained from the expression of δu(r)
+/Vain terms of ϕoby using
(kda)1=O(ε2), see (3.2). Downstream of the ablation front, in the region where
ξ=O(1), the unperturbed temperature Θ(ξ) is, according to (kda)1=O(ε2),
of order 12, see (2.4). The order of magnitude δu(p)
+(ξ)/Va=O(ε2)isobtained
from the definition u(p)
+VadaΘθwith δθ =O(kδθ). In the hot conduction region
a quasi-steady approximation applies because ρ+t=O(ερaVax), as a consequence of
(3.2), σ/Vak=O(ε1), together with ρ+a=1+=O(ε2). The pressure is evaluated
from ρaVaxδu+≈−xδp+, and the orders of magnitude in the hot side are
δθo(ξ)=O(α/da)µ
f/¯
µ=O()u(p)
+(ξ)/Va=O(ε2),(3.5)
δp+(ξ)aV2
a=O(ε2)u(r)
+(ξ)/Va=O(ε2˜
ϕo).(3.6)
In the cold region upstream of the ablation front where the density is constant, the
pressure is evaluated from (2.1) by using σ/Vak=O(ε1), xδukδu,xδpkδp,
so that ρaσδu
∼−p,andδu/Va=O(δpaV2
a). The order of magnitude of
the pressure being the same on both sides of the ablation front, the perturbed flow
velocity in the cold side is smaller than in the hot conduction region by a factor ε,
but larger than the ablation velocity by a factor 1:
δpaV2
a=O(ε2)u/Va=O(ε1)u
nf −|δu|ξ=0 =O(ε|δu|ξ=0).(3.7)
4. Nonlinear analysis
In the limit we consider from now on, ν1, the last term on the left-hand side of
equation (3.3a) together with the right-hand side of (3.3b) become negligible so that
θis a solution of Laplace equation. An explicit solution of (3.4) is
d˜
ϕo/dξ=eξ
0
x1e2xdx+e
ξ
ξ
x1e2xdx,
showing that ˜
ϕoand d˜
ϕo/dξare of order 1,δu(r)
+=O(δu(p)
+) see (3.5), (3.6). This
also indicates the existence of a boundary layer at ξ= 0 where the vorticity ×δu(r)is
localized, ξ1: d˜
ϕo/dξξ(1ξ1), d2˜
ϕo/dξ21ξ1;ξ=O(1): d2˜
ϕo/dξ2=O(1).
This suggests performing the nonlinear analysis in the distinguished limits
ν→∞
2(Fr1)1/(ν1) 0
cρac→∞
2Θc→∞,(4.1)
for amplitudes of the wrinkled front of same order as the wavelength,
αk =O(1)/d
a=O(ε2νν1),(4.2)
see (3.2). The parameter Θcwill no longer appear in the rest of the analysis, since the
last two conditions in (4.1) are there only to ensure the validity of (2.3), kmdc1,
while ε0 ensures kmda1, see Sanz et al. (2006).
Considering length and time scales of the order of the most linearly amplified
wavelength and its linear growth, the useful non-dimensional variables are, according
to (3.2),
ζ=(r/da)ε2Fr1=(tVa/da)εFr1with ε2Fr1=νε2ν,(4.3)
see (4.1). In the system of coordinates, ζ=(ζ,η), ζand ηdenote the longitudinal
and transverse coordinates respectively, and the coordinates of the ablation front
486 C. Almarcha, P. Clavin, L. Duchemin and J. Sanz
will be denoted by ζf. The orders of magnitude of the fully nonlinear regime are
obtained by introducing (4.2) into (3.5), (3.6) and (3.7) after using (3.2). Hence the non-
dimensional quantities of order unity in the limit (4.1) for characterizing the ablation
rate µ µµ, the normal velocity of the front uf
nand the flow u±u±u±,
π±π±π±with π±p±ρ±gx are M(ζf), Uf
n(ζf), U±(ζ), Π±(ζ),
defined as follows:
µ/µM, uf
nVaUf
nε, (4.4a, b)
u/VaU/ε, πaV2
aΠ2,u+/VaU+2,π+aV2
aΠ+2.
(4.5)
Equation (4.4b) is obtained from (3.2), σ/Vak=O(1), ∂α/∂t =O(σ/k).
Let us now show that a constant-density approximation holds at leading order
downstream of the front. According to (2.4) and (3.2a), one has Θ=(ξ/K)1ε2,
with, according to (3.5) and (4.2), δθ =O(ε2ν). The perturbations of θ,pand up
+
decreasing to zero exponentially fast with increasing ξ, the perturbation of ρvanishes
outside the region ξ=O(1) where ρis constant and δρ is smaller than ρat leading
order in the limit (4.1),
ξ=O(1): ρo+(ξ)a=1o=ε2ρ
o+(ξ)a=δθoΘν+1
o=O(ε2).(4.6)
In addition, according to the comment above (4.1), the flow is potential in this region,
uo+=u(p)
o+=Vadaθo2.
The vicinity of the origin ξ= 0 is more tricky, and requires special attention. The
boundary conditions (3.4) at ξ= 0 together with the expression q
o=d
2˜
ϕo/dξ2|ξ=0 are
obtained from (2.6) and (2.7) for a matching region on the hot side of the ablation
front corresponding to Θ=O(1). There, the variation of the dynamical pressure
δ(µ2
f)issmallerthanδp by a factor ε, so that relation (2.6b)is[δpo]+
=0. Not
only is the vorticity localized inside the boundary layer 16Θ<12, but also the
variations of dynamical pressure are of order 1in this layer, while the variation of
pressure is of order unity. However the thickness (measured with the ξ-variable) of
the layer shrinks to zero in the limit ν→∞. Therefore the analysis may be carried
out in the limit (4.1) by incorporating the layer ε>ρ+a
2into the ablation front
where µ2
fmust be retained in the jump in pressure. The ablation front becomes a
vortex sheet, separating two regions of constant density ρ=ρa,ρ+=ε2ρawhere the
flow is in irrotational motion.
The space variables ξand ζbeing obtained from the original space variable x/da
with the same scaling, see (3.2) and (4.3), the leading-order Euler equations can then
be written in non-dimensional form
U/∂τ +(U·˜
)U=˜
Π,∂U+/∂ζ +(U+·˜
)U+=˜
Π+,(4.7)
with ˜
∇·U=0 and ˜
∇·U+=0,˜
denoting the gradient with respect to ζ. Here, the
approximation of a negligible unperturbed velocity in the cold flow, Va/u=O(ε)see
(4.5), and a steady-state approximation in the hot flow, t=O(εu+x), has been used.
Let us now go back to the jump conditions across the front. According to (4.4) and
(4.5), uf
n=O(|u|), |u|=O(Va), the normal velocity of the ablation front is larger
than the ablation velocity Va, but smaller than the perturbed velocity of the expelled
hot fluid, |u|=O(ε|u+|). Mass conservation, equation (2.6a), then leads to
n·U|ζ=ζf=Uf
n,n·(ex+U+)|ζ=ζf=1+M, (4.8a, b)
Ablative Rayleigh–Taylor instability 487
where Uf
nis the velocity normal to the front ξf(t) in the system of coordinates
(4.3). Equation (4.8a) expresses that the front moves with the cold fluid, and (4.8b)
that the flow velocity of the hot flow is controlled by the ablation rate only. In
other words, the ablation rate is negligible in the cold side, |u|/Va=O(1)and
µ/¯
µ=O(1), while the front velocity is negligible in front of the flow velocity in the
hot fluid, uf
n=O(ε|u+|). By using (2.4), dan·∇θ|ζ=ζf=n·ex, equation (4.8b)together
with equation (2.7a) leads to n·U+|ζ=ζf=dan·∇(θθ)|ζ=ζf. This can be written in
terms of a velocity potential Φ+of order unity, n·U+|ζ=ζf=n·˜
Φ+|ζ=ζf, where by
definition Φ+(θθ)Fr1=O(1), see (4.3) and (4.6). According to the leading order
of (2.4) and (2.5), θ=0,θ = 0, and the velocity potential Φ+is a solution to Laplace
equation. On the other hand, the hot fluid velocity being larger than the cold fluid
velocity, |u+|=O(|u|), conservation of transverse momentum implies that the hot
fluid is expelled quasi-perpendicularly to the front, t·(ex+U+)|ζ=ζf=0, where tis the
unitary vector tangent to the front. Viewed from the hot side, the front is an isotherm,
t·∇θ|ζ=ζf= 0 equivalent to t·(ex+˜
Φ+)|ζ=ζf=0, so that t·U+|ζ=ζf=t·˜
Φ+|ζ=ζf.
Hence the boundary condition of the hot flow is U+|ζ=ζf=˜
Φ+|ζ=ζf.Thisisinfull
agreement with an incompressible and a potential flow and slaved to the temperature,
as in the linear approximation, U+=˜
Φ+.
The flows Uand U+being irrotational, (4.7) can be written as U/∂τ =
˜
{Π+|U|2
/2}and U+/∂ζ =˜
{Π++|U+|2
/2}, giving
∂Ψ/∂ τ =Π−|˜
Ψ|2
/2,∂Φ
+/∂ζ =Π+−|˜
Φ+|2
/2,(4.9)
where Ψis the velocity potential of the cold flow, U=˜
Ψ, with limζ→−∞ Ψ=0
and n·˜
Ψ|ζ=ζf=Uf
n, see (4.8). According to (2.3) and (2.4), the boundary conditions
for Φ+are limζ+Φ+=0 and Φ+|ζ=ζf=ζf. The dynamics of the ablation front
ζf(τ) is then obtained by solving two Laplace equations, ˜
Ψ=0 and ˜
Φ+= 0, with
the four boundary conditions just mentioned above, plus a fifth one at the front for
conservation of longitudinal momentum, obtained from equation (2.6b),
[ππ+]x=α=(1+M)22(α/da)Fr1.(4.10)
The first term on the right-hand side is the dynamical pressure and the second
is the acceleration. According to (4.1), (4.2) and (4.3), (α/da)Fr1=ζf2,sothat
the three terms in (4.10) are of same order, see (4.5) and the discussion after (4.6)
concerning the vicinity of ξ= 0. Therefore, according to equation (4.8b) together with
(4.9), the conservation of longitudinal momentum across the front takes the following
non-dimensional form:
∂Ψ
∂τ ∂Φ+
∂ζ +|˜
Ψ|2
2|˜
Φ+|2
2ζ=ζf
=1[n·(ex+˜
Φ+)|ζ=ζf]2+ζf.(4.11)
This equation is similar to that obtained in the semi-empirical SBM of Clavin &
Almarcha (2005) for Fr1=O(1) in the limit of a large jump in density across the abla-
tion front. According to the present analysis, equation (4.11) is obtained in a systematic
way in the limit (4.1). Moreover, according to (4.6), the unknown jump in density
appearing in the scalings of the SBM is now determined: ρ+a=(Fr1)1/(ν1) .
Recalling that constant terms are irrelevant and that the hot flow is orthogonal to
the ablation front, the right hand side of (4.11) may be written in a simpler form
|n·(ex+˜
Φ+)|ζ=ζf=|ex+˜
Φ+|ζ=ζf.
488 C. Almarcha, P. Clavin, L. Duchemin and J. Sanz
Z = f (z)
ζ
ξ
f (z) = e–(i)z
M
Figure 2. Conformal map used to transform the periodic surface into a bounded domain.
Considering infinitesimal disturbances in the form e±i, the linear analysis leads
to the linear growth rate S=KK2, in full agreement with (3.1), since the solution
to (3.3) and (3.4) shows that limν+[d2˜
ϕo/dξ2|ξ=0] = 1, see Sanz et al. (2006). When
the velocity potential of the total flow Ψ+Φ++ζis introduced into (4.11), the
free boundary problem takes the simpler form (1.1), (1.2) describing an extension of
the RT problem for an inviscid fluid above a vacuum without surface tension. The
numerical solution of this problem is discussed in the next section.
5. Results and discussion
Accurate numerical solutions to (1.1) and (1.2) are obtained in periodic two-
dimensional geometry with an extension of the boundary integral method of
Duchemin et al. (2005). We first briefly explain the method. Thanks to the potential
flow approximation, all the useful information is concentrated on the interface. The
power of boundary integral methods consists of reducing the dimension of the problem
by one: from two dimensions to one dimension in our case. The method utilizes a
highly non-uniform set of collocation points placed on the front and redistributed at
each time-step in order to ensure a good accuracy in regions where the curvature of the
front is large. Particular attention is paid to the redistribution in order to be sure that
it does not introduce spurious effects. At each time-step, knowing Ψand Φ+=Ψ+ζ
on the front (i.e. on the collocation points), the Laplace equation is solved to find the
velocities of the fluid particles on each side of the front. This computation is performed
using the conformal map Z=e
iz,z=ξ+iζ, to transform the periodic domain into a
bounded domain where the Cauchy theorem applies, see figure 2. The discrete version
of the Cauchy theorem gives us the streamfunctions corresponding respectively to
Ψon the cold side and Φ+on the hot side. Knowing the streamfunctions and the
velocity potentials on each side, we deduce the normal and tangential gradients of Ψ
and Φ+. The points on the front are advected using (1.2a) while the velocity potential
Ψis updated using the boundary condition (1.2c), after introducing the material
derivative ∂Ψ/∂ τ +|Ψ|2/2. Finally the time-stepping method used to obtain the
evolution of the front pattern is a fourth-order Runge–Kutta scheme with a time-step
decreasing according to the minimum acrlength between two successive points on the
surface.
Ablative Rayleigh–Taylor instability 489
π
π
ππ
–2π
0
0
τ = 10.4
8.29
6
Figu re 3. Numerical solution of (1.1), (1.2) (solid line) at various times for a sinusoidal
disturbance with the wavelength of the most linearly amplified mode, initial amplitude: 0.1.
The curvature singularity is spontaneously formed at τ=8.29. A smooth solution (dashed
line) is obtained when a small surface-tension-like term, 0.004 curvature, is added from the
beginning in (1.2c).
An example of solution is shown in figure 3. Unlike the RT instability with unity
Atwood number (A= 1) and zero surface tension, the problem does not possess a
smooth solution for more than a finite time, as for the solution of the Birkhoff–Rott
equation studied by Moore (1979) and also for the RT instability with A<1, see
Baker, Caflisch & Siegel (1993) and Matsuoka & Nishihara (2006). The shape of
the front develops a singularity within a finite time, the inclination of the tangent
to the front stays finite and a continuous but infinite curvature develops. The points
of maximum and minimum curvature (of opposite sign on both sides of the zero
curvature point) collapse in finite time, while the absolute value of the curvature blows
up. Using the boundary integral method, we were able to approach the singularity
with great accuracy. Figure 4 shows the time evolution of the following quantities: the
derivative of the curvature with respect to the arclength sat the inflection point κ(s),
the maximum and minimum curvatures around the inflection point κmax and κmin and
the distance in arclength between these last two points smax smin. The critical time
τ0at which the singularity develops is found by requiring that the behaviour of these
quantities in powers of τ0τis observed for a large number of decades. A remarkable
behaviour in powers of τ0τis observed independently of the wavenumber. The
scaling laws derived from numerical fitting in figure 4 are self-consistent since a
variation of the curvature (τ0τ)1on an arclength distance (τ0τ)4/3gives a
curvature derivative in (τ0τ)7/3. The exponents are accurate within a few percent.
Figure 5 shows the rescaled curvature as a function of the rescaled arclength for
490 C. Almarcha, P. Clavin, L. Duchemin and J. Sanz
10–4 10–3 10–2 10–1 100
10–6
10–4
10–2
100
102
104
106
108
1010
(τ0τ)4/3
(τ0τ)–7/3
(τ0τ)–1
κ(s)
κmax
κmin
smaxsmin
τ0τ
Figure 4. κ(s), κmax ,κmin and smax smin as a function of τ0τin a log-log plot.
–6 –4 –2 0 2 4 6
–0.8
–0.6
–0.4
–0.2
0
0.2
0.4
Figure 5. Rescaled curvature according to (τ0τ)1as a function of the arclength rescaled
according to (τ0τ)4/3.
various time-steps. The very good collapse of these curves exhibits a self-similarity
solution of the problem close to the critical time τ0.
The nonlinearity introduced by ablation |˜
Φ+|2plays a key role in the singularity
formation, since such a singularity is not observed in the RT problem with A=1 for
which Φ+does not exist, see Duchemin et al. (2005). There is however an indication
that the singularity is smoothed out at following orders of the asymptotic expansion
when the finite thickness of the ablation front is taken into account. If a small
Ablative Rayleigh–Taylor instability 491
surface-tension-like term is introduced into the right-hand side of (1.2c) for modelling
the finite thickness effect of the curved front, as for wrinkled flames, see Pelc´
e & Clavin
(1982), the arclength between the extrema of curvature still decreases to zero but in
infinite time, exponentially, while the curvature modulus increases exponentially. The
rest of the front shape is not influenced sensitively by the small surface-tension-like
effect. For periodic conditions and infinitesimal initial disturbances, the tip of the
bubble reaches the same constant curvature and velocity in the long time limit as
in the RT problem investigated by Taylor (1950), but with an overshoot. This is
confirmed by a local analysis with a single-mode approximation of the Layzer-type,
see Layzer (1955), but with a different wavenumber for Ψand Ψ+, showing that
the contribution of |˜
Ψ+|2in (1.2) gives a constant term in the long time limit. The
spike continues to accelerate while its radius increases with time, unlike the radius of
curvature of the RT spike which decreases in time like 13, see Clavin & Williams
(2005) and Duchemin et al. (2005).
To conclude, the free boundary problem (1.1), (1.2) is mathematically ill-posed for
general classes of initial conditions, since a curvature singularity appears within a
finite time, but it provides a simple and useful framework for studying the nonlinear
growth of the hydrodynamical instability of ablation fronts in ICF, at least before
the sudden formation of the singularity. With our regularizing scheme, no roll-up
is observed after the singularity, see figure 3, unlike in Kelvin–Helmholtz instability.
A very accurate numerical method is required to observe the singularity. An open
question is whether the smooth solution of equations (1.1), (1.2) modified by a surface,
tension-like term is physically relevant for the long-term evolution, beyond the time at
which the singularity spontaneously appears in the absence of the surface-tension-like
term. The vortex produced by front wrinkling could become no longer negligible
as soon as a strong increase of curvature develops locally, and a vortex sheet could
appear in the hot fluid. If this were the case, the singularity would indicate a transition
in the dynamics of the pattern, beyond which the present analysis would fail. This
seems to be the case in the direct numerical simulation presented by Betti & Sanz
(2006) who investigated the role of the vorticity for ν=5/2, Fr1=0.2 and a real
density profile. In this respect the influence of higher-order effects and the physical
relevance of both the singularity and the limit of strong temperature dependence of
thermal conductivity remain to be investigated.
This research was supported by the French program “Instabilit ´
es Hydrodynamiques
intervenant en FCI en lien avec le LMJ” sponsored by the CEA-DIF and the CNRS.
We would like to thank Professor Yves Pomeau for fruitful discussions and careful
reading of the manuscript.
REFERENCES
Atzeni, S. & Meyer-Ter-Vehn, J. 2004 The Physics of Inertial Fusion, 1st edn. Oxford Science
Publications.
Baker, G., Caflisch, R. E. & Siegel, M. 1993 Singularity formation during Rayleigh–Taylor
instability. J. Fluid Mech. 252, 51–78.
Baker,G.R.,Meiron,D.I.&Orszag,S.A.1980 Vortex simulation of the Rayleigh–Taylor
instability. Phys. Fluids 23, 1485–1490.
Betti, R., Goncharov, V., McCrory, R. L. & Verdon, C. P. 1995 Self-consistent cutoff wave
numbers of the Rayleigh–Taylor instability. Phys. Plasmas 2, 3844–3851.
Betti,R.,McCrory,R.L.&Verdon,C.P.1993 Stability analysis of unsteady ablation fronts.
Phys. Rev. Lett. 71, 3131–3134.
492 C. Almarcha, P. Clavin, L. Duchemin and J. Sanz
Betti, R. & Sanz, J. 2006 Bubble acceleration in the ablative Rayleigh–Taylor instability. Phys.
Rev. Lett. 97, 205002.
Bodner, S. 1974 Rayleigh–Taylor instability and laser-pellet fusion. Phys. Rev. Lett. 33, 761–764.
Bychkov,V.,Golberg,S.M.&Liberman,M.A.1994 Self-consistent model of the Rayleigh–Taylor
instability in ablatively accelerated laser plasma. Phys. Plasmas 1, 2976–2986.
Clavin, P. & Almarcha, C. 2005 Ablative Rayleigh–Taylor instability in the limit of an infinitely
large density ratio. C. R. M´
ec. 333, 379–388.
Clavin,P.&Masse,L.2004 Instabilities of ablations fronts in inertial fusion: a comparison with
flames. Phys. Plasmas 11, 690–705.
Clavin,P.&Williams,F.2005 Asymptotic spike evolution in Rayleigh–Taylor instability. J. Fluid
Mech. 525, 105–113.
Duch em in, L., Jos serand, C. & Clavin, P. 2005 Asymptotic behavior of the Rayleigh–Taylor
instability. Phys. Rev. Lett. 94, 224501.
Goncharov, V., Betti, R., McCrory, R. L., Sorotokin, P. & Verdon, C. P. 1996 Self-consistent
stability analysis of ablation fronts with large Froude numbers. Phys. Plasmas 3, 1402–1414.
Layzer, D. 1955 On the instability of superimposed fluids in a gravitational field. Astrophys. J. 122,
1–12.
Matsuoka, C. & Nishihara, K. 2006 Erratum: Vortex core dynamics and singularity formations
in incompressible Richtmyer-Meshkov instability. Phys. Rev. E73, 026304 (2006). Phys. Rev.
E74, 049902.
Moore, D. W. 1979 The spontaneous appearance of a singularity in the shape of evolving vortex
sheet. Proc. R. Soc. Lond. A365, 105–119.
Pelc´
e, P. & Cl av i n, P. 1982 Influence of hydrodynamics and diffusion upon the stability limits of
laminar premixed flames. J. Fluid Mech. 124, 219–237.
Piriz, A. 2001 Hydrodynamic instability of ablation fronts in inertial confinement fusion. Phys.
Plasmas 8, 997–1002.
Piriz, A. R., Sanz, J. & Iba
nez, L. F. 1997 Rayleigh–Taylor instability of the steady ablation fronts:
The discontinuity model revisited. Phys. Plasmas 4, 1117–1126.
Sanz , J. 1994 Self-consistent analytical model of the Rayleigh–Taylor instability in inertial
confinement fusion. Phys. Rev. Lett. 73, 2700–2703.
Sanz , J. 1996 Self-consistent analytical model of the Rayleigh–Taylor instability in inertial
confinement fusion. Phys. Rev. E53, 4026–4045.
Sanz,J.,Masse,L.&Clavin,P.2006 The linear Darrieus-Landau and Rayleigh–Taylor instabilities
in inertial confinement fusion revisited. Phys. Plasmas 13, 102702.
Sanz , J., Ram´
ırez,J.,Ramis,R.,Betti,R.&Town,R.P.J.2002 Nonlinear theory of the ablative
Rayleigh–Taylor instability. Phys. Rev. Lett. 89, 195002.
Taylor, G. I. 1950 The instability of liquid surfaces when accelerated in a direction perpendicular
to their plan. Proc. R. Soc. Lond. 201, 192–196.
... Cette dernière raison rend le modèle complet difficile à étudier, que ce soit analytiquement ou numériquement. Au cours des dernières décennies, de nombreux modèles simplifiés [CADS07,MC04,SMC06] ont vu le jour: frontières raides, faible vorticité, analyses auto-consistantes etc... Ces modèles analytiques sont complexes et reposent sur des heuristiques, rendant parfois difficile leur interprétation physique. Notons également que la plupart de ces modèles sont étudiés dans la limite de très grands exposants de conductivité m 1, ce qui ne correspond pas forcément à la réalité physique m = 7/2 (bien que ces modèles semblent quand même fournir des résultats en accord avec les simulations et expériences). ...
... This study is motivated by the context of Inertial Confinement Fusion, in which this model arises. Numerical computations and heuristics on the full hydrodynamical model [CADS07,MC04] yield dispersion relations very close to the one we obtain here. Our results are also very close to the ones established in [CMR11], where the authors consider a slightly different model which is not Galilean invariant. ...
... This nonlinear diffusion arises in very high temperature hydrodynamics and Physics of Plasmas [ZR66]. For example, in the context of Inertial Confinement Fusion [CADS07,Alm07,MC04], the dominant mechanism of heat transfer is the so called Spitzer electronic diffusion, which corresponds to m = 7/2. We will consider only diffusion exponents m > 3. Equations (1.2.1a)-(1.2.1b) correspond to conservation of energy and mass, where the transversal part of the flow is neglected. ...
... Another issue of importance for the ablative RT instability in ICF is the dynamics of spikes studied recently in Refs. [158,161,162]. Based on the assumption that the asymptotic flow field near the spike tip is close to that imposed by free fall, the approach to free- fall acceleration is shown in Ref. [161] to follow an inverse power law in time. ...
... Long time numerical simulations of the RT instability at At ¼ 1 demonstrated that the spike's curvature evolves as ft 3 , while the overshoot in acceleration shows good agreement with the t À5 law [162]. In addition, an asymptotic analysis of the spike shape at the RT unstable ablation fronts has been performed in Ref. [158] for a strong temperature dependence of the thermal conductivity. It has been demonstrated that the spike shape develops a curvature singularity within a finite time, similar in nature to the KH instability. ...
Article
The overarching objective of the present endeavor is to demonstrate the universal character of combustion phenomena for various areas of modern physics, focusing on inertial confinement fusion (ICF) in this review. We present the key features of laser deflagration, and consider the similarities and differences between the laser plasma flow and the slow combustion front. We discuss the linear stage of the Rayleigh–Taylor instability in laser ablation, short-wavelength stabilization of the instability due to the mass flow, and demonstrate the importance of the concepts and methods of combustion science for an understanding of the corresponding ICF processes. We show the possibility of the Darrieus–Landau instability in the laser ablation flow and discuss the specific features of the instability at the linear and nonlinear stages as compared to the combustion counterpart of this phenomenon. We consider the nonlinear stage of the Rayleigh–Taylor instability in the ICF and generation of ultra-high magnetic field by the instability, and show that proper understanding of vorticity production in the laser plasma and, hence, of the magnetic field generation requires concepts from combustion science.KeywordsCombustionInertial confinement fusionDeflagrationRayleigh–Taylor instabilityDarrieus–Landau instability
... For example in physics of plasmas, and particularly in the context of Inertial Confinement Fusion, the dominant mechanism of heat transfer is the so-called electronic Spitzer heat diffusion and corresponds to m = 5/2 (see e.g. [13,23]). Suitably rescaling time and space yields the nonlinear parabolic problem ...
Article
We study the free boundary regularity of the traveling wave solutions to a degenerate advection-diffusion problem of Porous Medium type, whose existence was proved in \cite{MonsaingonNovikovRoquejoffre}. We set up a finite difference scheme allowing to compute approximate solutions and capture the free boundaries, and we carry out a numerical investigation of their regularity. Based on some nondegeneracy assumptions supported by solid numerical evidence, we prove the Lipschitz regularity of the free boundaries. Our simulations indicate that this regularity is optimal, and the free boundaries seem to develop Lipschitz corners at least for some values of the nonlinear diffusion exponent. We discuss analytically the existence of corners in the framework of viscosity solutions to certain periodic Hamilton-Jacobi equations, whose validity is again supported by numerical evidence.
... This is relevant for the study of asymptotic spike evolution at ablation front. [50][51][52] Also of note is the coalescence of k ¼ 50 lm bubbles into larger features between 17.7 ns and 18.7 ns. This indicates that the ablative RTI has reached a high level of non-linearity. ...
Article
Full-text available
Academic tests in physical regimes not encountered in Inertial Confinement Fusion will help to build a better understanding of hydrodynamic instabilities and constitute the scientifically grounded validation complementary to fully integrated experiments. Under the National Ignition Facility (NIF) Discovery Science program, recent indirect drive experiments have been carried out to study the ablative Rayleigh-Taylor Instability (RTI) in transition from weakly nonlinear to highly nonlinear regime [A. Casner et al., Phys. Plasmas 19, 082708 (2012)]. In these experiments, a modulated package is accelerated by a 175 eV radiative temperature plateau created by a room temperature gas-filled platform irradiated by 60 NIF laser beams. The unique capabilities of the NIF are harnessed to accelerate this planar sample over much larger distances (’1.4 mm) and longer time periods (’12 ns) than previously achieved. This extended acceleration could eventually allow entering into a turbulent-like regime not precluded by the theory for the RTI at the ablation front. Simultaneous measurements of the foil trajectory and the subsequent RTI growth are performed and compared with radiative hydrodynamics simulations. We present RTI growth measurements for two-dimensional single-mode and broadband multimode modulations. The dependence of RTI growth on initial conditions and ablative stabilization is emphasized, and we demonstrate for the first time in indirect-drive a bubblecompetition, bubble-merger regime for the RTI at ablation front
... Rayleigh-Taylor instability is a fundamental phenomenon in fluid dynamics and it occurs when a heavier fluid lies above a lighter fluid in the presence of a constant gravitational field [30,35]. Interest in this problem stems from its importance to a wide variety of applications such as oceanography, mixing ( [2], and references therein), thermal convection [10], finger selection in splashes [17], inertial confinement fusion [39,31,3] and other natural phenomena and technical applications [32]. When fluids are immiscible, a sharp interface exists between them and any perturbation of the interface causes formation of a pattern of rising bubbles of lighter fluid and falling spikes of heavier fluid. ...
Article
Full-text available
This study considers the nonlinear dynamics of stratified immiscible fluids when an electric field acts perpendicular to the direction of gravity. A particular setup is investigated in detail, namely two stratified fluids inside a horizontal channel of infinite extent. The fluids are taken to be perfect dielectrics and a constant horizontal field is imposed along the channel. The sharp interface separating the two fluids may or may not support surface tension and the Rayleigh-Taylor instability is typically present when the heavier fluid is on top. A novel system of partial differential equations that describe the interfacial position and the leading order horizontal velocity in the fluid layers, is studied analytically and computationally. The system is valid in the asymptotic limit of one layer being asymptotically thin compared to the second fluid layer, and as a result nonlocal electrostatic terms arise due to the multiscale nature of the physical setup. The initial value problem on spatially periodic domains is solved numerically and it is shown that a sufficiently strong electric field can linearly stabilize the Rayleigh-Taylor instability to produce nonlinear quasi-periodic oscillations in time that are quite close to standing waves. In situations when the instability is present, the system is shown to generically evolve to touch-up singularities with the interface touching the upper wall in finite time while the leading order horizontal velocity blows up. Accurate numerical solutions allied with asymptotic analysis show that the terminal states follow self-similar structures that are different if surface tension is present or absent, but with the electric field present. In the presence of surface tension the touch-up is found to take place with bounded interfacial gradients but unbounded curvature, with electrostatic effects relegated to higher order. If surface tension is absent, however, the electric field supports touch-up with a local cusp structure so that the interfacial gradients themselves are unbounded. The self-similar solutions are of the second kind and extensive simulations are used to extract the scaling exponents. Distinct and independent methods are described and implemented and agreement between them is excellent.
Article
Academic tests in physical regimes not encountered in Inertial Confinement Fusion will help to build a better understanding of hydrodynamic instabilities and constitute the scientifically grounded validation complementary to fully integrated experiments. Under the National Ignition Facility (NIF) Discovery Science program, recent indirect drive experiments have been carried out to study the ablative Rayleigh-Taylor Instability (RTI) in transition from weakly nonlinear to highly nonlinear regime [A. Casner et al., Phys. Plasmas 19, 082708 (2012)]. In these experiments, a modulated package is accelerated by a 175 eV radiative temperature plateau created by a room temperature gas-filled platform irradiated by 60 NIF laser beams. The unique capabilities of the NIF are harnessed to accelerate this planar sample over much larger distances ( ≃ 1.4 mm) and longer time periods ( ≃ 12 ns) than previously achieved. This extended acceleration could eventually allow entering into a turbulent-like regime not precluded by the theory for the RTI at the ablation front. Simultaneous measurements of the foil trajectory and the subsequent RTI growth are performed and compared with radiative hydrodynamics simulations. We present RTI growth measurements for two-dimensional single-mode and broadband multimode modulations. The dependence of RTI growth on initial conditions and ablative stabilization is emphasized, and we demonstrate for the first time in indirect-drive a bubble-competition, bubble-merger regime for the RTI at ablation front.
Article
We propose a model for the study of ablation fronts encountered in inertial confinement fusion flows. Following the ideas of physicists specially Sanz's ones, we try to justify the modeling of this problem by a Hele–Shaw equation posed in the complement of a bounded domain the boundary of which is evolving with time. This equation is coupled with an advection equation for the vorticity of the fluid. We propose a numerical strategy based on the fat boundary method to discretize the system and we present some numerical results which show that the numerical method is feasible to simulate ablative Raleigh–Taylor instabilities in the context of these plasma flows.
Article
We study the 2D system of incompressible gravity driven Euler equations in the neighborhood of a particular smooth density profile ρ0(x) such that ρ0(x)=ρaξ(xL0), where ξ is a nonconstant solution of ξ̇=ξν+1(1−ξ), L0>0 is the width of the ablation region, ν>1 is the thermal conductivity exponent, and ρa>0 is the maximum density of the fluid. The linearization of the equations around the stationary solution (ρ0,0→,p0), ∇p0=ρ0g→ leads to the study of the Rayleigh equation for the perturbation of the velocity at the wavenumber k: −ddx(ρ0(x)du¯dx)+k2(ρ0(x)−gγ2ρ0′(x))u¯=0. We denote by the terms ‘eigenmode and growth rate’ an L2(R) solution of the Rayleigh equation associated with a value of γ. The purpose of this paper is twofold: •derive the following expansion in kL0, for small kL0, of the unique reduced linear growth rate γgk∈[14,1]gkγ2=1+2Γ(1+1ν)(2kL0ν)1ν+a2(kL0)2ν+O(kL0) where a2 is explicitly known, provided ν>2,•prove the nonlinear instability result for small times in the neighborhood of a general profile ρ0(x) such that k0(x)=ρ0′(x)ρ0(x) is regular enough, bounded, and k0(x)(ρ0(x))−12 bounded (which is the case for ρaξ(xL0)), thanks to the existence of Λ such that γ≤Λ for all possible growth rates and at least one growth rate γ belongs to (Λ2,Λ). This generalizes the result of Guo and Hwang [Y. Guo, H.J. Hwang, On the dynamical Rayleigh–Taylor instability, Arch. Ration. Mech. Anal. 167 (3) (2003) 235–253], which was obtained in the case ρ0(x)≥ρl>0.
Article
Full-text available
The linear and non-linear sensitivity of the 180 kJ baseline HiPER target to high-mode perturbations, i.e. surface roughness, is addressed using two-dimensional simulations and a complementary analysis by linear and non-linear ablative Rayleigh–Taylor models. Simulations provide an assessment of an early non-linear stage leading to a significant deformation of the ablation surface for modes of maximum linear growth factor. A design using a picket prepulse evidences an improvement in the target stability inducing a delay of the non-linear behavior. Perturbation evolution and shape, evidenced by simulations of the non-linear stage, are analyzed with existing self-consistent non-linear theory.
Article
Full-text available
This paper is devoted to the study of the deceleration phase of inertial confinement capsules. The purpose is to obtain a zero-dimensional model that has the form of a closed system of ordinary differential equations for the main hydrodynamic quantities. The model takes into account the energy released by nuclear reactions, a nonlocal model for the -particle energy deposition process, and radiation loss by electron bremsstrahlung. The asymptotic analysis is performed in the case of a strong temperature dependence of the thermal conductivity. We finally study the beginning of the expansion phase after stagnation to derive an ignition criterion. © 2008 American Institute of Physics.
Article
Full-text available
A vortex technique capable of calculating the Rayleigh–Taylor instability to large amplitudes in inviscid, incompressible, layered flows is introduced. The results show the formation of a steady-state bubble at large times, whose velocity is in agreement with the theory of Birkhoff and Carter. It is shown that the spike acceleration can exceed free fall, as suggested recently by Menikoff and Zemach. Results are also presented for instability at various Atwood ratios and for fluids having several layers.
Article
Full-text available
Within the framework of the quasi-isobaric approximation (low Mach number), a self-consistent stability analysis of ablation fronts in inertial confinement fusion is performed for all the modes with a wavelength larger than the conduction length in the cold material. The validity domain is ranging from short to long wavelength modes, shorter and larger than the total thickness of the thermal wave, respectively. The analysis leads to a single analytical expression for the dispersion relation valid in the whole range of modes, including the transition regime (wavelength of the same order of magnitude as the total thickness). The hydrodynamic instabilities of ablation fronts are thus described by a unified theory in a large domain of conditions, ranging from weak acceleration in the early stage of irradiation to strong acceleration during the main implosion phase. In the weak acceleration regime, the transition between Darrieus-Landau unstable modes and damped oscillatory modes is described. Comparison with numerical results shows a good agreement. Comparison with the previous analyses sheds new light on the stabilization mechanism. The result of the sharp boundary model is recovered in the limit of large power index for thermal conduction.
Book
The book is devoted to targets for nuclear fusion by inertial confinement and to the various branches of physics involved. It first discusses fusion reactions and general requirements for fusion energy production. It then introduces and illustrates the concept of inertial confinement fusion by spherical implosion, followed by detailed treatments of the physics of fusion ignition and burn, and of energy gain. The next part of the book is mostly devoted to the underlying physics involved in inertial fusion, and covers hydrodynamics, hydrodynamic stability, radiative transport and equations-of-state of hot dense matter, laser and ion beam interaction with plasma. It discusses different approaches to inertial fusion (direct-drive by laser, indirect-drive by laser or ion beams), including recent developments in fast ignition. The goal of the book is to give an introduction to this subject, and also to provide practical results even when derived on the basis of simplified models.
Article
The evolution of a small amplitude initial disturbance to a straight uniform vortex sheet is described by the Fourier coefficients of the disturbance. An approximation to the exact evolution equation for these coefficients is proposed and it is shown, by an asymptotic analysis valid at large times, that the solution of the approximate equations develops a singularity at a critical time. The critical time is proportional to In(epsilon exp -1), where epsilon is the initial amplitude of the disturbance and the singularity itself is such that the nth Fourier coefficient decays like n exp -2.5 instead of exponentially. Evidence (not conclusive, however) is present to show that the approximation used is adequate. It is concluded that the class of vortex layer motions correctly modelled by replacing the vortex layer by a vortex sheet is very restricted; the vortex sheet is an inadequate approximation unless it is everywhere undergoing rapid stretching.
Article
It is shown that, when two superposed fluids of different densities are accelerated in a direction perpendicular to their interface, this surface is stable or unstable according to whether the acceleration is directed from the heavier to the lighter fluid or vice versa. The relationship between the rate of development of the instability and the length of wave-like disturbances, the acceleration and the densities is found, and similar calculations are made for the case when a sheet of liquid of uniform depth is accelerated.
Article
The Rayleigh-Taylor instability in laser-driven spherical implosions can be stabilized by convective flow and by the "fire-polishing" effect, but the size of the stabilization effect depends on details of the thermal conductivity near the ablation surface.
Article
The instability of ablation fronts strongly accelerated toward the dense medium under the conditions of inertial confinement fusion (ICF) is addressed in the limit of an infinitely large density ratio. The analysis serves to demonstrate that the flow is irrotational to first order, reducing the nonlinear analysis to solve a two-potential flows problem. Vorticity appears at the following orders in the perturbation analysis. This result simplifies greatly the analysis. The possibility for using boundary integral methods opens new perspectives in the nonlinear theory of the ablative RT instability in ICF. A few examples are given at the end of the Note. To cite this article: P. Clavin, C. Almarcha, C. R. Mecanique 333 (2005).
Article
We study temporal evolution of an interface in the Richtmyer- Meshkov instability numerically. The interface is treated as a vortex sheet and the Birkhoff-Rott equation is used in order to describe motion of a vortex sheet. We show that redistribution of grid points to equal arclength and the application of the Fourier series expansion for numerical differentiations and integrations make it possible to perform long-time caluculations. Successive profiles of a vortex sheet and the temporal evolution of the sheet strength are presented, and especially the evolution of the sheet strength of a vortex core, defined as a point at which the absolute values of the curvature and strength of a sheet become maximum, is discussed. It is found that the sheet strength of a vortex core takes a maximal value at a finite time and turns to gradually decreasing when the Atwood number in the system is non-zero.
Article
A comparison with flames sheds new light on the dynamics of ablation fronts in inertial confinement fusion (ICF). The mathematical formulation of the problem in ICF is the same as for flames propagating upwards. The difference concerns the Froude number Fr, yielding a different order of magnitude for the nondimensional wave number of the marginally stable disturbances. When the thermal conductivity varies strongly, as is the case in ICF, a wide range of characteristic (diffusive) lengths is involved across the wave structure. For disturbances with intermediate wavelengths, a “universal” diffusive relaxation rate of thermal waves is exhibited with no dependence on the heat conductivity. This is a key point for describing the dynamics of strongly accelerated ablation fronts whose marginally stable wavelength is much shorter than the total wave thickness. The coupling of hydrodynamics and heat conduction is analyzed in a way similar to flame theory, through the derivation of a kinematic relation for the ablation front including its thermal relaxation. A transition between the regimes of flames and ablation fronts in ICF is exhibited with decreasing Fr. For a moderate acceleration, Fr≫1, the result for flames is recovered. For a large acceleration, Fr of order unity, the thermal relaxation, when coupled with hydrodynamics, is shown to damp out the Darrieus–Landau instability, yielding the known result in ICF for strongly accelerated ablation fronts. For a wide class of models, including the simple two-length-scale model, the description is shown to be independent of the model. A weakly nonlinear analysis, valid irrespective of the number of unstable modes, is carried out for describing the early development of nonlinear structures of the ablation front in ICF. The role of the Darrieus–Landau instability at the early stage of irradiation is pointed out. © 2004 American Institute of Physics.