next up previous
Next: Foreshock Physics Up: Earth's Bow Shock and Previous: Fast Mode Nature of

Kinetic Aspects of Shock Physics

Observations show that the shock structure varies primarily with the fast mode Mach number tex2html_wrap_inline868 and with the angle tex2html_wrap_inline870 between the upstream field tex2html_wrap_inline722 and the local shock normal. In particular, for quasi-perpendicular regions of the shock ( tex2html_wrap_inline874 degrees) there is a transition between ``laminar'' (smooth) and ``turbulent'' magnetic profiles in the domain tex2html_wrap_inline876 , with laminar profiles at low tex2html_wrap_inline878 (Figure 13.7).

  figure227
Figure 13.7: The top panel shows a laminar bow shock profile while the bottom panel shows a turbulent shock profile [Livesey et al., 1984; Gosling and Robson, 1985].

Laminar profiles are by definition smooth, except for upstream trains of standing whistler waves (which are related to dissipation processes at the shock). Turbulent shocks have a well-defined ``shock foot'', where the magnetic field smoothly increases over a scale length tex2html_wrap_inline880 , and an ``overshoot'' behind the ramp where the field increases above the Rankine-Hugoniot prediction on a scale tex2html_wrap_inline880 (sometimes with additional ripples) before finally reaching the average field level predicted by the Rankine-Hugoniot conditions. Particle detectors show that the foot and overshoot regions are associated with gyrating beams of solar wind protons reflected at the shock (Figure 13.8).

  figure235
Figure 13.8: Ion velocity distributions observed downstream from a marginally supercritical shock with tex2html_wrap_inline720 degrees and tex2html_wrap_inline722 almost perpendicular to the instrument's measurement plane [Schopke et al., 1983]. The dashed circle in the second frame shows the predicted locus for specularly reflected ions.

Laminar shock transitions sometimes have an upstream wave train but do not have significant levels of gyrating ions. The production of these gyrating ions is intrinsic to the both the structure and dissipation processes active at supercritical shocks, as explained more below. The definition of a ``supercritical'' shock is that gyrating ions represent the primary dissipation process active at the shock.

Quasi-parallel shocks are very turbulent with fluctuations tex2html_wrap_inline888 and an extensive foreshock region, sometimes to such a degree that it is difficult to identify where the shock transition occurs (Figure 13.9).

  figure241
Figure 13.9: Schematic illustration of the quasi-parallel and quasi-perpendicular regions of Earth's bow shock and the foreshock region upstream of the shock [Greenstadt and Fredericks, 1979]. Note that the foreshock region upstream of the quasi-parallell shock is extremely turbulent, often making it difficult to identify the shock transition itself. .

Large fluxes of gyrating ions are often observed throughout a huge volume near a quasi-parallel shock transition.

The development of reflected, gyrating beams of solar wind ions, with associated creation of foot and overshoot regions and thermalization of the downstream particles is also associated with the self-consistent development of an electrostatic potential across the shock (the ``cross-shock potential''). These phenomena can be explained as follows [Gosling et al., 1982; Schwartz et al., 1983; Goodrich, 1985; Scudder et al., 1987].

(1) During one gyroperiod solar wind protons travel about 5000 km (at tex2html_wrap_inline708 ), while the electrons travel tex2html_wrap_inline892 km, and the shock ramp is tex2html_wrap_inline894 km thick. The ions thus see the ramp as an abrupt discontinuity during their gyromotion while the electrons complete numerous gyroperiods as they cross the ramp.

(2) A self-consistent electrostatic potential is encountered at the shock ramp, which slows and resists the ion flow (but accelerates the electron flow).

(3) A fraction ( tex2html_wrap_inline896 ) of the solar wind protons have insufficient normal velocity to overcome the potential barrier and are specularly reflected at the magnetic ramp; that is, their perpendicular velocities are reversed as shown in Figure 13.10.

  figure246
Figure 13.10: Schematic from Gosling and Robson [1985]. (a) Sketch of the trajectory of an ion specularly reflecting off a shock with tex2html_wrap_inline724 degrees. (b) Idealized 2-D ion velocity distributions at several distances from the shock ramp. Specularly reflected particles move along a circle of radius tex2html_wrap_inline900 centred at the bulk flow velocity. The dashed lines are aligned parallel to the shock surface.

(4) The specularly reflected protons move so as to gain energy from the convection electric field and develop a large gyrospeed tex2html_wrap_inline902 given by

equation252

as well as a different gyrocenter velocity. Here tex2html_wrap_inline904 is the initial speed in the local shock frame.

(5) The gyrocenter velocity is directed downstream for quasi-perpendicular shocks ( tex2html_wrap_inline906 degrees) but upstream for quasi-parallel shocks [Gosling et al., 1982].

(6) For quasi-perpendicular shocks, then, the reflected protons gyrate upstream for a partial gyro-orbit before encountering the shock again, now with sufficient normal speed to overcome the potential barrier and pass downstream as a component with large apparent thermal energy. The current associated with this motion causes the magnetic field to increase in the downstream region. The spatial extent of the foot and overshoot are then of order the gyroradius of the specularly reflected ions.

(7) The potential develops as a result of the different motions of electrons and ions. Starting from the electron momentum equation, for a perpendicular shock with the normal in the tex2html_wrap_inline908 direction and the magnetic field in the tex2html_wrap_inline910 direction, the momentum equation for the electron fluid is

equation263

whence after neglecting the electron inertia term one may finally write

equation274

The cross-shock potential tex2html_wrap_inline912 thus results from the gradients in pressure and magnetic field energy across the shock (an ambipolar potential) as well as the effects of the net ion drifts associated with the gyrating ions (second term). The cross-shock potential is primarily oriented in the shock's normal direction.

Order of magnitude estimates are that

equation288

These estimates explain the characteristic size of the cross-shock potential and the amount of ion heating downstream from the shock. Moreover, the above balance between the upstream flow energy and the downstream temperature enhancement also explains qualitatively how the overall Rankine-Hugoniot conditions for the shock transition can be obeyed on the large scale, as they must be despite very different microphysics.

Figure 13.11 shows the development of gyrating ion beams in hybrid simulations (fluid electrons and particle ions), while Figure 13.12 shows the corresponding spatial profiles of the magnetic field strength and the cross-shock potential [Leroy et al., 1983].

  figure297
Figure 13.11: Ion phase space plots tex2html_wrap_inline728 and tex2html_wrap_inline730 from hybrid shock simulations [Leroy et al., 1982; Goodrich, 1985]. Here x is the coordinate along the shock normal and the magnetic field is in the z direction, tex2html_wrap_inline736 , and tex2html_wrap_inline738 . The times for each panel are (a) t = 0, (b) tex2html_wrap_inline742 , (c) tex2html_wrap_inline744 , (d) tex2html_wrap_inline746 , and (e) tex2html_wrap_inline748 . Note the development of the gyrating ion beams.

  figure313
Figure 13.12: Plots of the magnetic field and cross-shock potentials as a function of x and time for the simulations in Figure 13.11 [Leroy et al., 1982; Goodrich, 1985]. Note the development of the foot and overshoot in the potential and magnetic field.

Qualitatively the gyrating ions comprise a second ion population moving with significant relative speed to the core solar wind ions, especially perpendicular to the magnetic field. The associated gradients in the particle distribution function imply the possibility of instabilities redistributing this free energy and ending up with a single, broadened and thermalized ion distribution in the magnetosheath. Figures 13.4 and 13.8 show this process observationally [Schopke et al., 1983].

A qualitatively important point is that the cross-shock potential and formation of gyrating ion beams involve reversible physics in the sense that no dissipation, energy loss, or entropy change is involved. How can this lead to an irreversible shock transition with a change in entropy between the upstream and downstream states? The answer is that wave-particle interactions associated with unstable particle distributions lead to dissipation and an increase in entropy. Nevertheless, much of the apparent ``heating'' of the electrons and ions across the shock can be understood in terms of ``reversible'' physics and the operation of the cross-shock potential.

A final remark is that for quasi-parallel shocks the gyrating ions have gyrocenters directed upstream from the shock, leading to an extended upstream region filled with gyrating ions and the large-amplitude MHD waves they drive. These upstream ``foreshock'' regions are then very turbulent with much evidence of Fermi acceleration and wave-particle interactions.


next up previous
Next: Foreshock Physics Up: Earth's Bow Shock and Previous: Fast Mode Nature of

Iver Cairns
Thu Sep 9 09:54:58 EST 1999