• Keine Ergebnisse gefunden

Generation of pressure anisotropies in a weakly collisional plasma . 91

4.2 Theoretical framework

4.2.1 Generation of pressure anisotropies in a weakly collisional plasma . 91

ns the number densities,

ps =

d3wmsw2

2 fs, (4.5)

p∥s =

d3wmsw2fs (4.6)

are the perpendicular and parallel pressures, qs =

d3wmsw2

2 wfs, (4.7)

qs =

d3wmsw3fs, (4.8)

qsand qsare heat fluxes (the parallel flux of the “perpendicular internal energy” and the parallel flux of the “parallel internal energy”, respectively),w the thermal component of a particle’s velocity,fsthe distribution functions of the particles. Subtracting equation (4.4) from equation (4.3), we get an evolution equation for the pressure anisotropy:

d

dt(ps−ps) = (ps+ 2ps)1 B

dB

dt + (ps3ps) 1 ns

dns

dt ·[(qs−qs)b]3qs·b

s(ps−ps). (4.9)

Assuming that collisions are fast compared to the fluid motions, the pressure anisotropy is then small, ps−ps ps ps ps, the collisional heat fluxes are qs (1/3)qs, and the total heat flux along a field line qs = qs+qs/2 = (5/6)qs. The value of the anisotropy is set by the balance between collisional relaxation and various driving terms:

s ps−ps ps 1

νs [1

B dB

dt 2 3

1 ns

dns

dt +4·(qsb)6qs·b 15ps

]

. (4.10)

Thus, the pressure anisotropy is driven by changing magnetic-field strength, changing particle density, and by parallel heat fluxes.

It is useful to estimate the degree of anisotropy induced by different driving terms in (4.10). If we consider fluid motions with velocityuat scaleLu, variations of B at the scale of the velocity field LB = Lu, and parallel temperature gradient Ts δTs/LT at scale LT, we can evaluate the contribution ∆B,n;s of changing B and n, and the contribution

T;s of the heat fluxes, to the total anisotropy as

B,n;s u vth,s

λs

Lu, (4.11)

T;s λ2s LTLu

δTs

Ts , (4.12)

where we have used the expression for the heat flux qs = −κsTs with thermal conduc-tivity κs ∼nsvth,sλs, λs the mean free path, vth,s the thermal speed. Assume that the flow

4.2 Theoretical framework 93

velocity is nearly sonic,u∼vth,i, and that the variations of temperature are of order unity, δT /T 1. Then ∆T ∼λ2/(LuLT) for both particle species, and ∆B,n;s∼λ/Lu×vth,s/vth,i. Hence, in our ordering, for the ions, the term linked to the magnetic-field changes ∆B,n;i λ/Lu is dominant if LT λ (even in astrophysical systems with very sharp tempera-ture gradients, e.g., cold fronts or buoyant bubbles of relativistic plasma in the ICM, the magnetic-field lines are typically stretched by the fluid flow in the direction perpendicular to the gradient (e.g., Komarov et al., 2014), thus significantly increasing the scale of tem-perature variation along the field lines). For the electrons, ∆B,n;e1/40×λ/Lu, and the two contributions can be of the same order (∆B,n;e T;e), depending on the properties of the flow and the orientation of the magnetic-field lines connecting the hot and cold regions of the plasma. Note that the total anisotropy ∆tot = ∆e+ ∆i is bounded from below by the firehose instability, ∆tot > 2/β, where β is the ratio of thermal to magnetic-energy densities. If the positive ion anisotropy dominates (∆i e), then the mirror instability compels ∆i to stay below the mirror marginal level, ∆i ≲ 1/β. Therefore, in regions of high plasma β, either γ (= B1dB/dt) or νs is modified by the instabilities to keep the anisotropy between the marginal levels, 2/β < ∆tot <1/β (e.g., Melville et al., 2016). In Appendix 4.6, we calculate the ion anisotropy for the simulated cold fronts (Section 4.3) and mark the regions where the firehose and mirror instabilities could develop. Because in our work the ion anisotropy is typically dominant, the two instabilities are regulated by the ions.

In this work, we are primarily interested in electron pressure anisotropy because of its possible observational imprint in the form of polarization of thermal bremsstrahlung.

From the above estimates, it is clear that in the case of astrophysical systems with large temperature gradients, the driving term linked to heat fluxes must be taken into account along with the driving by the magnetic-field changes. We do this in detail for cold fronts in Section 4.3.

4.2.2 Polarization of bremsstrahlung by electron anisotropy

Consider first the polarization of bremsstrahlung emission from an electron beam deflected by a single ion. At low energies (compared to the kinetic energy of an electron), photons produced by small-angle scattering of the electrons off the ion are polarized in the plane perpendicular to the electron beam due to the mainly perpendicular acceleration that slightly changes only the direction of the electron velocity. At higher energies, when both the direction and the magnitude of the electron velocity change significantly, polarization becomes dominated by the acceleration the electrons experience parallel to the beam.

Below, we demonstrate that the latter regime is of first importance for our problem, because the degree of polarization is considerably larger at high energies in the case of thermal bremsstrahlung from a cloud of anisotropic electrons.

Bremsstrahlung emission from a beam of electrons of energy εis fully described by dif-ferential cross sections per unit solid angle and photon energy,d2σ(ε, ϵ, θ) andd2σ(ε, ϵ, θ), for the components perpendicular and parallel to the radiation plane (spanned by an emit-ted photon’s and an initial electron’s momenta). Here, d2 = d2/(dϵdΩ), ϵ stands for the

Figure 4.1: The degree of bremsstrahlung polarizationP(ϵ, θ) = (d2σ−d2σ)/(d2σ+d2σ) from a beam of electrons of energyε = 8 KeV as a function of the emitted photon’s energy ϵ and the angle θ between the beam axis and the photon’s momentum.

emitted photon’s energy, and θ the angle between the emitted photon’s momentum and the beam axis. We use the fully relativistic cross sections calculated by Gluckstern & Hull (1953) in the first Born approximation, which are appropriate for the problem at hand1. Because the formulae in the original paper by Gluckstern & Hull (1953), as well as those given later by Bai & Ramaty (1978), are both subjected to typos, we provide the correct explicit expressions for the cross sections in Appendix 4.7. The degree of polarization is P(ϵ, θ) = (d2σ −d2σ)/(d2σ+d2σ) = d2σ1/d2σ0, where d2σ1 is the differential cross section of the polarized emission, d2σ0 of the total emission. Its dependence on the photon energy and direction with respect to the beam axis is illustrated in Fig. 4.1. The transition between the perpendicular and parallel polarization occurs at photon energy ϵ ∼ε/8. As noted before, the perpendicular polarization at low energies is produced by small-angle scattering of the electrons, while the parallel is the result of collisions that significantly change the electron energy.

Since the differential cross sections presented above are essentially the ’Green’s func-tions’ of the bremsstrahlung emission, the total and polarized emission from a cloud of elec-trons can be found by integrating over the electron distribution function. Let us introduce

1The limit of validity of this approximation is given by condition ε/mec2 (Z/137)2 , where Z is the charge of the scattering ion in atomic units,ε the energy of an outgoing electron (Gluckstern & Hull, 1953). ForZ = 1, the condition is satisfied for outgoing electrons at energiesε 30 eV.

4.2 Theoretical framework 95

a spherical coordinate system and assume that the electron distribution is axisymmetric with respect to the magnetic-field direction, taken to be the z axis. We denote the unit vector in the direction of the incoming electron ˆp= (sinθ0cosϕ0,sinθ0sinϕ0,cosθ0), and the direction of the line of sight ˆk= (sinθ,0,cosθ) (choose ϕ = 0 without loss of general-ity because the resulting polarization pattern is also axisymmetric). The geometry of the vectors is illustrated in Fig. 4.2. The polarization directions perpendicular and parallel to the plane spanned by the vectors ˆpand ˆk (the radiation plane) are, respectively,

ˆe = pˆ×kˆ

|pˆ×kˆ|, (4.13)

eˆ =

kˆ×p×k)ˆ

|pˆ×kˆ| . (4.14)

Then ˆe is rotated by angle χ (see Fig. 4.2) with respect to the y direction, which is the perpendicular polarization direction in the referencexz plane that contains the line of sight ˆk. The angleχ is expressed as

cosχ= ˆey = (sinθcosθ0cosθsinθ0cosϕ0)/sinθ, (4.15) where θ is the angle between ˆp and ˆk:

cosθ = cosθcosθ0+ sinθsinθ0cosθ0. (4.16) Linear polarization (for unpolarized electrons, bremsstrahlung photons are never circu-larly polarized) is described by the two independent Stokes parameters P1 and P2: P1 corresponds to the degree of polarization with respect to a given reference plane (xz in our case); P2 to the degree of polarization with respect to a plane rotated around the line of sight by π/4 from the reference plane. For a given momentum of the initial electron, P1 and P2, normalized by the total intensity, are transformed by rotation of the radiation plane relative to the reference plane as

P1,ˆp = cos 2χd2σ1 d2σ0, P2,ˆp = sin 2χd2σ1

d2σ0. (4.17)

Thus, knowing the expression for the angle χ [equation (4.15)] between the radiation and reference planes, we can calculate the degree of polarization of bremsstrahlung emission from a cloud of electronsP1,2 =I1,2/I0, whereI1,2is the intensity of the polarized emission and I0 the total intensity, both integrated over the electron distribution F(ε, θ0):

I0(ϵ, θ) = ni

ϵ

1

1

d(cosθ0)

0

0 v(ε)F(ε, θ0)d2σ0(ε, ϵ, θ), (4.18) I1(ϵ, θ) = ni

ϵ

+1

1

d(cosθ0)

0

0 v(ε)F(ε, θ0) cos 2χ d2σ1(ε, ϵ, θ), (4.19) I2(ϵ, θ) = ni

ϵ

+1

1

d(cosθ0)

0

0 v(ε)F(ε, θ0) sin 2χ d2σ1(ε, ϵ, θ). (4.20)

Figure 4.2: Geometry for the problem of the polarization of bremsstrahlung emission from a cloud of electrons.

Due to the axisymmetry of the electron distribution function, I2 integrates to zero [see, e.g., the appendix of Haug (1972) for a mathematical proof of this], and the total degree of linear polarization isP =I1/I0.

The distribution functionF(ε, θ0) is related to the velocity distribution functionf(v, θ0) as

F(ε, θ0) = v2 f(v, θ0)dv

dε. (4.21)

For the velocity distribution function, we employ a bi-Maxwellian:

f(v, θ0) =ne ( me

2πT

) ( me 2πT

)1/2

exp [

−mev2 2T0

(T0

T sin2θ0+T0

T cos2θ0 )]

. (4.22) where T0 = (1/3)T+ (2/3)T is the total temperature. If the anisotropy

∆ = T−T

T0 (4.23)

is small, and ∆mev2/(2T0)1, one can expand the distribution function to the first order in ∆:

f(v, θ0) = f0(v) +δf(v, θ0), (4.24) where f0(v) is an isotropic Maxwell distribution at temperature T0:

f0(v) = ne ( me

2πT0 )3/2

exp (

−mev2 2T0

)

, (4.25)

4.2 Theoretical framework 97

Figure 4.3: The degree of bremsstrahlung polarization from a cloud of electrons with a bi-Maxwellian distribution at temperatureT0 = 8 keV and anisotropy level ∆ = 0.25 [(4.23)]

as a function of the emitted photon energyϵ and the angle between the axis of anisotropy and the line of sight. Results in the linear approximation [equations (4.24-(4.25)] are shown in red for comparison. The polarization degree is plotted with the minus sign to facilitate comparison with Fig. 4.1. The opposite sign comes from the fact that the electron pressure anisotropy ∆ is defined to be positive for T > T [equation (4.23)].

and the anisotropic perturbation is

δf(v, θ0) = ∆ mev2 2T0

(1

3cos2θ0 )

f0(v). (4.26)

Using equations (4.20), (4.21), and (4.26), we obtain the degree of polarization of thermal bremsstrahlung for a small anisotropic perturbation of the electron distribution, when the linear approximation (expansion in ∆) [equation (4.24)] is applicable:

P(ϵ, T0, θ) = ∆ sin2θ G(ϵ, T0), (4.27) where G(ϵ, T0) becomes a function of ϵ/T0 at temperaturesT0 ≲10 keV. At ϵ∼ a few T0, G(ϵ, T0)1. The degree of polarization from a cloud of anisotropic electrons with ∆ = 0.25 at T0 = 8 KeV is shown in Fig. 4.3 in black for a general bi-Maxwellian distribution, and in red in the linear approximation [equation (4.27)]. We see that the linear approximation holds at least up to ∆0.25.